1. Introduction
The uniform well-posedness of the Cauchy problem for first-order differential equations is traditionally studied in terms of semigroups of operators. Briefly, a resolving
-continuous semigroup of operators
of the equation
exists if and only if the respective Cauchy problem
is uniformly well posed [
1,
2,
3,
4]. In this case, every operator
maps the initial data
to a value
of a solution of the problem at time
. Such operators for a first-order equation form a semigroup, but for other equations, it is not so. But even in the absence of the semigroup property, the study of resolving families of operators of various differential equations allows us to obtain important qualitative results on their solutions. Therefore, the issue of the existence of a resolving family for a differential equation is very important. This issue is equivalent to the fulfillment of the criterion in the form of Hille–Yosida conditions for the operator
A from a first-order equation or their generalizations for other equations. Strongly continuous resolving families and criteria of their existence have been studied for second-order equations [
5,
6,
7], for some integro-differential equations [
8], for evolution integral equations [
9], for equations with the Gerasimov–Caputo derivative [
10,
11,
12], with the Riemann–Liouville derivative [
13,
14], and with a distributed Gerasimov–Caputo derivative [
15]. Various applications in physics of integro-differential equations and differential equations with fractional derivatives, including the Hilfer derivative, which is the focus of this work, can be found in [
16,
17,
18].
Here, we study strongly continuous resolving families for a linear equation in a Banach space
:
where
is the Hilfer derivative [
16] of an order
,
, of a kind
, and
A is a linear closed operator in
. The Cauchy-type problem
is considered for Equation (
1). Hereafter,
is the Riemann–Liouville fractional derivative for
, and the Riemann–Liouville fractional integral for
(see below for details) is
,
.
In the second section, the main definitions are given, and some basic properties of operators from a resolving family for Equation (
1) are proved. In the third section, it is shown that for
, Equation (
1) has a resolving family of operators in the case of a bounded operator
A only. Hence, the only case of interest is
. The main result is proved here. It is a theorem on necessary and sufficient conditions of the existence of a resolving family of operators for Equation (
1). These conditions generalize the Hille–Yosida conditions for
-continuous semigroups of operators. The most difficult issue is the sufficiency of these conditions; it was solved using Phillips-type approximations [
1,
19] and by applying the properties of operator
A’s resolvents without assuming the presence of the Radon–Nikodym property in a considered Banach space. Due to the main theorem, problem (
1) and (
2) have a solution, and their uniqueness is proved. In the fourth section, we show the unique solution of existence for Cauchy-type problem (
2) for the inhomogeneous equation
under some assumptions on
f. The fifth section contains the construction of a family of operators
,
such that for
, where
and
, there exists a resolving family for Equation (
1), and for
, it is true only for
or
. The last section contains an application of this result for the investigation of an initial boundary value problem with a time fractional-order one-dimensional Schrödinger equation.
In the case of
, the obtained abstract results coincide with the corresponding results of [
10,
11,
12]. Note the works [
20,
21] on equations with the Hilfer derivative and their resolving families of various classes.
2. Primary Results
Let
be a Banach space. For
and
, let
be the Riemann–Liouville fractional integral,
,
,
be the usual derivative of the
m-th order, and
be the Riemann–Liouville derivative of
. For
, we will also denote
.
Denote
,
The Hilfer fractional derivative of an order
,
, and of a kind
is defined as
For sufficiently smooth h, we have . For , due to our designations, we obtain
Remark 1. For , the Hilfer fractional derivative coincides with the Riemann–Liouville derivative, and for , the Hilfer derivative is the Gerasimov–Caputo fractional derivative (see [22,23,24]). Denote the Laplace transform for as or . Denote also . Hereafter, the principal branch of the power function will be used.
Lemma 1. [
20].
Let , , have the Laplace transform, and . Then, It is known (see, e.g., [
25,
26]) that
Let be the Banach space of all linear continuous operators from to , and let denote the set of all linear closed operators, densely defined in , acting on space . We supply the domain of an operator through the norm of its graph, and thus, we obtain the Banach space .
We will use the notation for the set of all functions h that belong to and are absolutely continuous on every segment -th derivative.
Consider the Cauchy-type problem
for an equation
with
,
,
. A solution for problem (
4) and (
5) is a function
such that
,
, and condition (
4) and equality (
5) for
are valid.
Definition 1. Let , . A family of operators is called resolving for Equation (5) if the following conditions are satisfied: - (i)
,
- (ii)
For every ,
- (iii)
, for all ,
- (iv)
For every is a solution of the Cauchy-type problem , , , for Equation (5).
Remark 2. It is easy to show that for any , problem (4) and (5) has a solution . We write
if a resolving family of operators for Equation (
5) with a constant
in condition (i) exists.
Remark 3. In the case of , Equation (5) has a resolving family of operators that can be calculated (see, e.g., [27]):where is the Mittag–Leffler function. Thus, with some ω depending on . Definition 2. Let , . An operator is called the operator of the class for a some constant if the following two conditions are fulfilled:
(i) If , then
(ii) Remark 4. Conditions (i) and (ii) generalize the Hille–Yosida conditions and agree with them if . Indeed, in this case, we can use the formula for the n-th-order derivative of the resolvent and obtain Remark 5. In Theorem 2.8 [11], it was proved that the equation has a resolving family of operators, whereis the Gerasimov–Caputo derivative if and only if . Lemma 2. Let , . Then, .
Proof. Denote
for
,
. For
,
, we have
due to equality (10) [
13] with
instead of
. The proof is completed. □
Remark 6. The assertion of Lemma 2 for , was proved in [13]. Lemma 3. Let , , , and let be a resolving family of operators. Then, and Proof. For
,
, it follows from Lemma 1 that
due to conditions (iii) and (iv) of Definition 1 and the closedness of
A. This implies the invertibility of the operator
and the fulfilment of (
6). The boundedness of
follows from the obtained inclusion of
.
If we differentiate the equality
with respect to
, we obtain, for
,
Consequently, □
Corollary 1. Let , . Then, .
Lemma 4. Let , , . Then, for every , , Proof. Due to the condition
, for
, the inequalities
and
are valid. For
,
The density in
of the domain
implies that
for every
.
Analogously, we can obtain the required statement for every . □
Corollary 2. Let , , . Then, for arbitrary , the domain is dense in the space .
Proof. Consider . Then, and □
Corollary 3. Let , , . Then, for arbitrary , the domain is dense in the space with the graph norm of operator A.
Proof. Due to Lemma 4, for
, we obtain
and
Thus,
□
Theorem 1. Let , . There exist a resolving family of operators for Equation (5). Then, the family of operators is continuous in the norm of at the point if and only if . Proof. For
, by Lemma 3,
; hence,
By the assumption of the theorem, the function
is continuous on
; in addition,
. For any
, take a number
such that
for every
; therefore,
Here, we take into account that
for all
. Hence, for sufficiently large
, we have
, and there exists a continuous inverse operator for
. Thus,
.
Now, suppose that
. Hence, the inverse Laplace transform of
is defined for
. Take
for
. Then, for
, we have equalities
which imply that
as
□
3. Existence Criterion of a Resolving Family of Operators
Theorem 2. Let and with . Then, .
Proof. If
, then Corollary 1 implies that
. Therefore, the complex plane without some bounded set belongs to the resolvent set
. Hence, for sufficiently large
, it is necessary that
. Then, by Corollary 1,
Due to Lemma 5.2 [
7], operator
A is bounded. □
Let
, for
. For
and
, define using the Phillips approximations [
1] (Theorem 6.3.3) for the inverse Laplace transform as follows:
This series is uniformly convergent on all segments
, with any
; moreover, for
,
Due to the asymptotic expansion of the Euler gamma function [
28] (§1.18)
we have
The asymptotics of the Mittag–Leffler function [
11] (p. 12)
imply that
and
Consequently, for all
,
It is not difficult to prove the infinite differentiability of at every .
Lemma 5. Let , , . Then, there exists the strong limit , which is uniform with respect to t on all segments of the form , .
Proof. Take in further considerations
, for which
; hence,
If
, then we have
Therefore,
and the integral
converges.
Hence, , since is analytic on .
If
,
,
, then for sufficiently large
,
If
,
, we pass limit
in the equality
The domain is dense in the space due to Corollary 2. In addition, the family is uniformly bounded on segments , . Hence, the convergence of is strong and uniform with respect to t on segments . □
Denote
,
. Due to (
7),
Lemma 5 implies that
is a strongly continuous family of operators. Due to the proof of Lemma 5, we have for
,
Theorem 3. Let , . Then, .
Proof. The inclusion is proved in Corollary 1. Let . In this case, as it is proved above, there exists the strongly continuous family .
If
, then the derivatives
are sums of natural powers of
, which are multiplied by scalar functions. Consequently,
Due to these relations, and for . If we pass the limit as , then and for due to the closedness of A.
If
, we have
From (
9), it follows that
Hence,
.
If
, then
and
, and for
, we have
, since
.
For any
, choose
, where
and the denotation
is used. Since the operator
commutes with
by the construction, we have
If
, then we choose
and have
After using
on the left and right sides of equality (
10) and
on both sides of (
11), we obtain
, and for
,
, we have
.
If
, then
In this case, for
,
,
Hence,
. Due to Corollary 3, the domain
is dense in
; therefore, equality (
12) can be extended on
. Indeed,
for
,
. Therefore, for
,
and
, and for
,
.
Thus,
is a resolving family of operators for Equation (
5). □
Remark 7. Due to Lemma 3, the resolving family of operators for Equation (5) is unique. Remark 8. We can affirm that for , . Indeed, due to Lemma 2 and Theorem 2, , and by means of Remark 3, .
Corollary 4. Let , , , and . Then, the Cauchy-type problem for Equation (5) has a unique solution. It has a form . Proof. Due to Theorem 3, it remains to prove that the problem solution is unique.
Use
in Equation (
5) and obtain
Here,
, and the Laplace convolution * has a form
Equality (
13) is true for every solution of the Cauchy-type problem
for Equation (
5). In particular, for
, the equality
is valid; therefore, for another solution
z,
Differentiating by
t, we obtain
,
. □
Corollary 5. Let , , , and . Then, the Cauchy-type problemfor Equation (5) has a unique solution. It has a form . Proof. As in the previous proof, we can prove that for a solution of problem (
5) and (
14) with
,
,
Here,
by the definition of a solution. Taking the solution
, we have
therefore, for another solution
z,
and
,
.
For a solution
z for problem (
5) and (
14) with
, the function
solves this problem with
. Hence,
. □
5. Simple Examples of Operators from
Take the Banach space , which consists of sequences , where for , such that the series is convergent.
Consider the Cauchy-type problem
where
,
,
, and
. Set
and
for
such that
, which make up the domain
. It is not difficult to prove that for all
,
and
. Let
Then, for every
and
, we have
and
. So, we will consider
only.
Solving problem (
20) and (
21) for every
individually, we obtain that the resolving family of equation has a form
where
(see Remark 3). Since
, we have for a fixed
, due to the asymptotic expansion of the Mittag–Leffler function ([
11], p. 13) for
,
Therefore,
. Since for
the equality
holds, and for every
, we have
Thus,
for every
.
Now, take
. Then,
, and for a fixed
using the asymptotic expansion of the Mittag–Leffler function ([
11], p. 12), we obtain
as
. So,
for
,
. If
or
, (
20) is the Cauchy problem for the equation
,
, or
,
,
and
or
.
Analogously, we can consider the case .
6. Application to Initial Boundary Value Problem
For
, consider in a bounded domain
with a smooth boundary
, the initial boundary value problem
for the linear-time fractional Schrödinger equation
where
,
,
is the Riemann–Liouville fractional integral with respect to the time variable
t,
is the Hilfer fractional derivative with respect to
t,
i is the imaginary unit,
is the Laplace operator, and
. In setting
, linear operator
is defined as
,
,
. Thus, problem (
22)–(
24) is a partial case of problem (
4) and (
5).
It is known that
, where
are real negative eigenvalues of the Laplace operator, which are numbered in non-decreasing order, taking into account their multiplicities. Let
be the corresponding eigenfunctions of the operator
A, which form an orthonormal basis in
. Decomposing all the vectors in
according to this basis, we present problem (
22)–(
24) as follows:
Here,
and
,
. Thus, we obtained problem (
20) and (
21) in the
space, since the Fourier coefficient sequences
for every
and
belong to
.
For simplicity, let and ; then, , , . Arguing in a similar manner as in the previous section, we obtain that this operator is analogous to operator () and that there exists a family of resolving operators for every and .