1. Introduction and Preliminaries
In ([
1], Section 11, Section 12), J. Prüss investigated the solvability of the following problem on the line:
where
X and
Y are complex Banach spaces such that
Y is densely embedded into
X,
and
The author introduced the notion of a vector-valued homogeneous space
and analyzed the well-posedness of the equation
where
and
The existence of strong solutions to (
2) has been proved in the case that the forcing term
belongs to a subspace
of
In particular, for such inhomogeneities
, the Fourier–Carleman spectrum
is compact (cf. ([
1], Subsection 0.5, Subsection 0.6) for more details on the subject), where the real spectrum of (
2) is defined through
and
is a proper subspace of the space consisting of all functions
, which admits an extension to an entire function of exponential growth (the strong solution
of (
2) also enjoys this feature; see ([
1], Theorem 11.1). For some other references concerning the abstract Volterra integro-differential equations on the line, one may refer, e.g., to the paper [
2] by V. E. Fedorov and N. M. Skripka and [
3] by R. Ponce. The abstract Volterra integro-differential inclusions on the real line with generalized Weyl fractional derivatives have recently been investigated in [
4]; cf. also [
5,
6] for some other important references worth mentioning.
Here, we briefly analyze the following extension of problem (
1):
where
X and
Y are complex Banach spaces,
Y is embedded into
B is a closed linear operator with a domain and range contained in
,
,
and the operator
is injective. For simplicity, we will not consider the perturbations of term
which are bounded in variation, and we will apply the usually considered Fourier transform in place of the Fourier–Carleman transform considered in [
1].
The structure of this paper is as follows: We first explain the basic notation and terminology used throughout the paper. After that, we recall the basic definitions and results on
-regularized
C-pseudoresolvent families (
Section 1.1) and generalized Weyl integro-differential operators (
Section 1.2).
In
Section 2, we investigate the well-posedness of abstract degenerate non-scalar Volterra integral Equation (
3) on the line. The first structural result of
Section 2 is Proposition 2, where we consider the situation in which
is a global
-regularized
C-pseudoresolvent family. Here, we prove the existence of solutions to the Cauchy problem
In Proposition 3, we clarify a uniqueness result for the above problem, where we assume the existence of a global
-regularized
C-uniqueness family
for (
3); cf. [
7] for the notion and more details in this direction. In Theorem 1, we apply the Fourier transform in the space of tempered vector-valued distributions in order to analyze the existence and uniqueness of solutions to Problem (
3); cf. also Example 1, where we provide some applications to the kernels
where
,
,
is a closed linear operator on
X (
) and the vector space
is equipped with the graph norm
In Theorem 2, we consider the differential and analytical properties of mapping
which is incredibly important in our analysis. After that, we are ready to apply the operator-valued version of Mikhlin’s theorem ([
8], Proposition 8.2.3) in the analysis of the existence of a solution
of (
3) which satisfies
and
; cf. Theorem 3 for more details. The existence and uniqueness of almost-periodic-type solutions to (
3) are briefly considered in Section Almost-Periodic-Type Solutions to (
3).
For convenience, we will work in the setting of complex Banach spaces. We will reconsider the results established by J. Prüss in ([
1], Section 11, Section 12) for a degenerate abstract Cauchy problem on the line (
3) somewhere else. We will not consider here the abstract semilinear Volterra integral equations of non-scalar type on the line as well.
Before explaining the notation and terminology used throughout the paper, we feel it is our duty to emphasize that we have not been able to find any practical applications of the established results in physical or engineering models and also that our theoretical findings are not accompanied here by any numerical illustrations or concrete examples. Although some notations probably could be refined, we have done our best to increase the clarity and readability of this paper as well as the exposition of the material.
Notation and preliminaries. If , then we set where denotes the Euler Gamma function and the Dirac delta distribution. If and then we set
Unless specified otherwise, we assume henceforward that
and
are complex Banach spaces such that
Y is embedded into
X as well as that
where
denotes the space consisting of all continuous linear mappings from
Y into
X and
We assume that the operator
is injective and
B is a closed linear operator in
X such that
by
We denote the Banach space
where the graph norm
is defined by
The symbol
stands for the identity operator on
For more details about the integration of functions with values in Banach spaces, we refer the reader to [
7,
8]. We will use the same notion and notation as in [
7]. Finally, let us recall that the Fourier transform on the real line and its inverse transform are defined by
respectively; see [
8,
9,
10,
11,
12] and the references quoted therein.
1.1. -Regularized C-Pseudoresolvent Families
We need the following notion (cf. ([
7], Definition 2.9.2)):
Definition 1. Let , , and . A family in is called an -regularized C-pseudoresolvent family if the following holds:
- (S1)
The mappings , and , are continuous in X for every fixed , and , .
- (S2)
Put , , . Then (S2) means , , and is locally Lipschitz continuous in .
- (S3)
The resolvent equationshold, and (4), resp. (5), is called the first resolvent equation, resp. the second resolvent equation.
Let us recall that (S3) can be equivalently written as
where
cf. also ([
1], p. 153).
By continuity, we mean continuity in
We need to recall the notion from ([
7], Definition 2.9.3):
Definition 2. Let , , and . A strongly continuous operator family is said to be an -regularized C-uniqueness family if For further information, the reader may consult [
7] and the references quoted therein.
1.2. Generalized Weyl Fractional Derivatives
For Weyl fractional calculus, we refer to the research monographs [
13] by K. S. Miller and B. Ross and [
14] by S. G. Samko, A. A. Kilbas and O. I. Marichev; cf. also the research articles quoted in [
4]. In ([
15], Definition 3), we recently introduced the notion of a generalized Weyl
-fractional derivative
of a locally integrable function
where
X is a complex Banach space. A special case of
is the usually considered Weyl fractional derivative
Definition 3. Suppose that , is a locally integrable function, and The generalized Weyl fractional derivative of function is well defined if the mapping is well defined and m-times continuously differentiable by The function
is said to be a generalized Weyl
a-integral of function
If
for some
then the class of functions for which the above integral absolutely converges was first considered by M. J. Lighthill in [
16].
We need the following auxiliary results from [
4]:
Lemma 1. (i) Suppose that , and If -
then the term is well defined, the term is well defined and we have
- (ii)
Suppose that , Z is a complex Banach space, the operator family is strongly continuous, for each we have and . If then the term is well defined, the term is well defined and we have
We can similarly prove the next proposition:
Proposition 1. Suppose that , Z is a complex Banach space, the operator family is strongly continuous, and . Ifthen the term is well defined, the term is well defined and we have 2. Well-Posedness of Abstract Degenerate Non-Scalar Volterra Integral Equations on the Line
In the following proposition, all considered integrals are taken with respect to the topology of space
X (cf. also ([
7], Proposition 11.4)):
Proposition 2. Assume that , , and is a global -regularized C-pseudoresolvent family. Suppose that the functionsare well defined and Lebesgue measurable on the real line, the integrals and are convergent andThen, we have Furthermore, if the function is continuous, the functions and are well defined and continuous on the real line and the integrals defining the functions and are absolutely convergent, then the function is also continuous on the real line and we have Proof. It is clear that
The prescribed assumptions imply, together with ([
7], Theorem 1.2.3), Proposition 1 and (S3)’, that
as claimed. Suppose now that the requirements in the second part of the proposition hold. To show that
is continuous on the real line and (
8) holds, observe first that the integral which defines the function
is absolutely continuous and the function
is continuous. It can be simply proved by means of Lemma 1(ii) that
so that
,
Since the functions
and
are continuous on the real line and the integral which defines the function
is absolutely convergent, we can similarly prove with the help of Lemma 1(ii) that
and
This simply implies the desired assertion by differentiation of (
7). □
In the commonly considered case the integral must be defined for all so that the boundedness of function is not sufficient enough for applications of Proposition 2.
The uniqueness of solutions to Problem (
8) with
is examined in the following result:
Proposition 3. Assume that and is an -regularized C-uniqueness family for (3). Suppose, further, that is a continuous function, for a.e. , there exist and such that the integral is absolutely convergent in X for each the integral is absolutely convergent in X for each and Then, for all
Proof. Since we have assumed that the integral
is absolutely convergent in
X, Proposition 1 and (
9) together imply that the integral
is convergent in
X for each
and
Using the functional equality of
the above yields
so that
Since
and the integral
is absolutely convergent in
X for each
Lemma 1(i) implies
Keeping in mind that the function
is continuous, we can differentiate
m-times the last equality to show that
,
By the injectiveness of regularizing operator
C, we finally obtain
□
Concerning possible applications of Proposition 2 and Proposition 3 with
we would like to say that these results can be successfully applied to the
-regularized
C-pseudoresolvent families constructed in ([
17], Theorem A.12, Corollary A.13) and
-regularized
C-pseudoresolvent families constructed in ([
7], Theorem 2.9.7). In a degenerate setting, possible applications can be given to the abstract degenerate fractional integro-differential equations considered on p. 221 in [
7].
Suppose now that, for every
and
there is a unique continuous function
such that (
3) holds with
the function
is differentiable and
,
Then, the closedness of
B implies
, and we have
Multiplying (
3) with
and using the uniqueness of the solutions, it follows that
so that there is a unique element
s.t.
Coming back to (
3), we obtain
Therefore, the integral
is convergent for every
and
Consequently, we have
If
is a vector-valued function and the integral
is convergent for some
, then we tacitly assume henceforth that
is locally integrable. By a Fourier-transformable function, we mean any locally integrable function
such that the integral
is convergent for all
The condition (
10) has an important role in the formulation of the subsequent theorem, which seems to be new even for the equations of scalar type, with
:
Theorem 1. (i) Suppose that the operator is well defined and injective for all . Then, there exists a unique function such that the integral is convergent in X for every , (3) holds with and - (ii)
Suppose that (10) holds, , the function is locally integrable in Y and the regular distribution determined by belongs to the space of tempered Y-valued distributions. If , then is a solution of (3) for a.e. provided that and
Proof. Suppose that the function
is Fourier transformable, (
3) holds with
and
Since
B is closed, it follows that
Furthermore, the Fubini theorem yields that the function
,
is Fourier transformable and
so that
The injectivity of operator
implies
Since
, the Fourier inversion formula ([
8], Theorem 1.8.1 d) yields that
for a.e.
which completes the proof of (i). The proof of (ii) is quite similar. First of all, it is clear that
Taking into account the prescribed assumptions, the above simply implies by the reverse procedure that
for any
Since
,
and
the Fourier inversion formula ([
8], Theorem 1.8.1 d) yields that (
3) holds for a.e.
. □
Unfortunately, we cannot expect that there exists a function
such that
so the solution
of (
3), where
cannot be found in the form
with
and
More precisely, the following holds:
Proposition 4. Suppose that , and (12) holds. Then, Proof. Suppose the contrary. Take any
such that
Then,
and
Define
by
and
By ([
8], Theorem 1.8.1 a), the equality (
13) implies
It is clear that
so that the Riemann–Lebesgue lemma ([
8], Theorem 1.8.1 c) yields that
and
. Keeping in mind these facts, the closedness of
B and (
14) together imply
which is a contradiction. □
Now we will illustrate Theorem 1(ii) with the following example:
Example 1. The most symptomatic case for applications to multi-term problems is the case in which we havewhere , and is a closed linear operator on X (). We endow the vector space with the graph norm under which Y is complete. In our concrete situation, we have Assume now that there exist closed linear operator A on X and complex polynomials where such that and for cf. [7] for the notion and more details. Despite the negative result established in Proposition 4, the requirements necessary for applications of Theorem 1 are satisfied in many concrete situations, especially in those situations where the forcing term belongs to the Schwartz space of rapidly decreasing functions , and the functions , and , admit polynomially bounded holomorphic extensions to the strip , where a sufficiently small number Take, for instance, where and for Then, the functions , and , are infinitely differentiable and all their derivatives are polynomially bounded on the real line, which follows by means of the product rule, Cauchy integral theorem and equalityso that the solution of (3) and the function belong to the space For some concrete examples, we may refer to ([7], Theorem 1.2.7, Example 2.5.8). We can also provide many illustrative applications of Theorem 1(ii) to the abstract (degenerate) Volterra equations of the scalar type on the line. Let us note that Theorem 1(ii) can be also applied to the abstract degenerate non-scalar Volterra equations on the line, provided that the forcing term belongs to the space of rapidly decreasing ultra-differentiable functions of Beurling type as well as that the functions , and , admit ultra-polynomially bounded holomorphic extensions to the strip see ([18], Proposition 4.1) and ([7], Example 2.2.18) for more details in this direction. For the sequel, we need the next result:
Theorem 2. Suppose that the mapping , is continuous and Then, the following holds:
- (i)
If the assumption implies and and the assumptions and imply and for then the following holds: Furthermore, if for each there exists such that there exist analytic mappings and such that , and then the mapping is analytic, the mapping is analytic, , , , - (ii)
If , , then the following holds:
Proof. Denote
Then we have
Let us fix now a number
Since the operator
is injective, the above equality implies after a simple computation that
Since
, the dominated convergence theorem implies
Moreover, the mapping
,
is continuous so that
in
Taken together with the above equality, we obtain
This simply yields (
15) and (
18). The second statement in (i) remains to be proved. Towards this end, observe first that the mapping
is analytic. Let
be fixed. From the first part of (i) and the identity theorem for holomorphic functions ([
8], Proposition A.2), it follows that
Using the last equality, prescribed commutativity assumptions and ([
8], Proposition A.3), we find that the mapping
is analytic and
. This simply yields that
,
as well as that (
16) and (
17) hold true, finishing the proof of the theorem. □
Observe that the statements of Bernstein’s lemma ([
8], Lemma 8.2.1) and Mikhlin’s theorem ([
8], Proposition 8.2.3) continue to hold for the vector-valued functions. Using this fact, it readily follows that there exists a function
such that
,
where
is given by (
11), provided that
and there exists a sufficiently small real number
such that
Now we would like to state the following result:
Theorem 3. Suppose that (10) holds, , and there exists such that (19) holds and Then, there exists a function such that , , and (3) holds for a.e. Proof. Clearly, there exists a function
such that
,
Since
and (
20) holds, we similarly obtain
Furthermore,
Now the statement follows Theorem 1(ii). □
Under certain reasonable assumptions, we have the following identity:
so that Theorem 2 can be used to state a straightforward corollary of Theorem 3. To illustrate this result, let us assume that
and
, where
and
A is a closed linear operator on
Then, we have
so that Theorem 2 can be applied in many concrete situations in which the terms
and
decay sufficiently enough as
Finally, let us note that it would be very tempting to state some notable results on the existence of
-solutions of (
3) in the case of
Almost-Periodic-Type Solutions to (3)
The notion of almost periodicity was introduced by H. Bohr around 1924–1926 ([
19,
20,
21]); cf. also [
22,
23] and the references quoted therein. Suppose that
is continuous. If
then a number
is said to be an
-period for
if
By
we denote the set consisting of all
-periods for
. Let us recall that
is almost periodic if, for each
, the set
is relatively dense in
i.e., there exists
such that any subinterval of
of length
l meets
. Any almost-periodic function
is bounded and uniformly continuous, and the set of all almost-periodic functions
, denoted by
is a vector space with the usual operations. Equipped with the sup-norm,
is a Banach space.
Let
Then, we say that a function
is Stepanov
p-almost periodic if there is
such that any subinterval of
of length
l contains a point
such that
Further on, a function is said to be
- (i)
equi-Weyl-
p-almost periodic, if, for each
, we can find
and
such that any interval
of length
L contains a point
such that
- (ii)
Weyl-
p-almost periodic, if, for each
, we can find
such that any interval
of length
L contains a point
such that
Any almost-periodic function is Stepanov p-almost periodic, any Stepanov-p-almost-periodic function is equi-Weyl-p-almost periodic and any equi-Weyl-p-almost-periodic function is Weyl-p-almost periodic. All these inclusions are strict. Let us also note that the set of all Stepanov-p-almost-periodic functions , resp. equi-Weyl-p-almost-periodic functions , is a vector space with the usual operations, which is no longer true for Weyl-p-almost-periodic functions.
We will first observe that the Banach contraction principle can be successfully applied in the analysis of the existence and uniqueness of almost periodic solutions to the Problem (
1), provided that
,
and
is almost periodic. Towards this end, it suffices to observe that the mapping
given by
is a well-defined contraction.
Concerning Proposition 2, we will only observe here that the solution
inherits certain Weyl almost-periodic behavior from the forcing term
cf. [
22]. It is also worth noting that, for every
the abstract Cauchy problem (
3) is equivalent to the problem
where
, as well as that the problem (
8) is equivalent with the problem
This observation allows us to consider the situations in which the kernel
is not uniformly integrable in
but only exponentially bounded. Here, it is worth noting that, if the operator family
in
is a global
-regularized
C-pseudoresolvent family (the operator family
in
is a global
-regularized
C-uniqueness family), then
is a global
-regularized
C-pseudoresolvent family (
is a global
-regularized
C-uniqueness family). For example, if we assume that the forcing term
is almost periodic and the operator family
is exponentially decaying, then the solution
of (
21) will be almost periodic as well ([
22]). For many important examples and applications of
-regularized
-pseudoresolvent families, we refer the reader to [
1].
We close this subsection by stating the next proposition:
Proposition 5. Let Then, the following holds:
- (i)
Suppose that If is almost periodic and (3) holds, then the function is almost periodic. - (ii)
Suppose that and If is Stepanov p-almost periodic and (3) holds, then the function is Stepanov p-almost periodic. - (iii)
Let and let Suppose that is equi-Weyl-p-almost periodic, resp. Weyl-p-almost periodic and Weyl p-bounded, provided that , resp. provided that , and (3) holds. Then, the function is equi-Weyl-p-almost periodic, resp. can be represented as the difference of two Weyl-p-almost-periodic functions.
Proof. We will prove only (iii), because the proofs of (i) and (ii) are quite similar to the proof of (iii); cf. also ([
22], Proposition 2.6.11). Using the argumentation given in the proof of ([
22], Theorem 2.11.4), we can show that the mapping
is equi-Weyl-
p-almost periodic, resp. Weyl-
p-almost periodic. Since
the function
is also equi-Weyl-
p-almost periodic, resp. Weyl-
p-almost periodic, which simply implies the required conclusion. □