In what follows, we recall some definitions, notations, and results that we need in the sequel. Throughout this paper,
is a Banach space, and
,
, are closed linear operators defined on a common domain
which is dense in
; the notations
and
represent the resolvent set of the operator
and the domain of
endowed with the graph norm, respectively. For
, we fix
, and we represent by
the norm of
in
. Let
and
be Banach spaces. In this paper, the notation
stands for the Banach space of bounded linear operators from
into
endowed with the uniform operator topology, and we abbreviate this notation to
when
. Furthermore, for appropriate functions
, the notation
denotes the Laplace transform of
. The notation
stands for the closed ball with center at
and radius
in
. On the other hand, for a bounded function
and
, the notation
is given by
and we simplify this notation to
when no confusion about the space
arises.
We will define the phase space
axiomatically, using ideas and notations developed in [19]. More precisely,
will denote the vector space of functions defined from
into
endowed with a seminorm denoted
, and such that the following axioms hold.
-
(A)
If
,
,
is continuous on
and
, then for every
the following conditions hold:
-
(i)
is in
,
-
(ii)
,
-
(iii)
,
where
is a constant;
,
is continuous,
is locally bounded, and
are independent of
.
-
(A1)
For the function
in (A), the function
is continuous from
into
.
-
(B)
The space
is complete.
Example 2.1 (the phase space
).
Let
, and let
be a nonnegative measurable function which satisfies the conditions (g-5), (g-6) in the terminology of [19]. Briefly, this means that
is locally integrable and there exists a nonnegative, locally bounded function
on
, such that
, for all
and
, where
is a set with Lebesgue measure zero. The space
consists of all classes of functions
, such that
is continuous on
, Lebesgue-measurable, and
is Lebesgue integrable on
. The seminorm in
is defined by
The space
satisfies axioms (A), (A1), (B). Moreover, when
and
, we can take
,
, and
, for
(see [19, Theorem
] for details).
For additional details concerning phase space we refer the reader to [19].
To obtain our results, we assume that the integrodifferential abstract Cauchy problem
has an associated
-resolvent operator of bounded linear operators
on
.
Definition 2.2.
A one parameter family of bounded linear operators
on
is called a
-resolvent operator of (2.3)-(2.4) if the following conditions are verified.
-
(a)
The function
is strongly continuous and
for all
and
.
-
(b)
For
,
, and
for every
.
The existence of a resolvent operator for problem (2.3)-(2.4) was studied in [20]. In this paper, we have considered the following conditions.
-
(P1)
The operator
is a closed linear operator with
dense in
. There is positive constants
, such that
where
,
,
for some
, and
for all
.
-
(P2)
For all
,
is a closed linear operator,
, and
is strongly measurable on
for each
. There exists
(the notation
stands for the set of all locally integrable functions from
into
) such that
exists for
and
for all
and
. Moreover, the operator-valued function
has an analytical extension (still denoted by
) to
such that
for all
, and
, as
.
-
(P3)
There exists a subspace
dense in
and positive constant
, such that
,
,
for every
and all
.
In the sequel, for
and
,
for
,
, are the paths
and
oriented counterclockwise. In addition,
are the sets
We now define the operator family
by
The following result has been established in [20, Theorem 2.1].
Theorem 2.3.
Assume that conditions (P1)–(P3) are fulfilled, then there is a unique
-resolvent operator for problem (2.3)-(2.4).
In what follows, we always assume that the conditions (P1)–(P3) are verified.
We consider now the nonhomogeneous problem.
In the rest of this section, we discuss existence and regularity of solutions of
where
and
. In the sequel,
is the operator function defined by (2.11). We begin by introducing the following concept of classical solution.
Definition 2.4.
A function
,
is called a classical solution of (2.12)-(2.13) on
if
; the condition (2.13) holds and (2.12) is verified on
.
Definition 2.5.
Let
, we define the family
by
for each
.
The proof of the next result is in [20]. For reader's convenience, we will give the proof.
Lemma 2.6.
If the function
is exponentially bounded in
, then
is exponentially bounded in
.
Proof.
If there are constants,
such that
, we obtain
where
.
The next result follows from Lemma 2.6. We will omit the proof.
Lemma 2.7.
If the function
is exponentially bounded in
, then
is exponentially bounded in
.
We now establish a variation of constants formula for the solutions of (2.12)-(2.13). The proof of the next result is in [20]. For reader's convenience, we will give the proof.
Theorem 2.8.
Let
. Assume that
and
is a classical solution of (2.12)-(2.13) on
, then
Proof.
The Cauchy problem (2.12)-(2.13) is equivalent to the Volterra equation
and the
-resolvent equation (2.6) is equivalent to
To prove (2.16), we notice that
Therefore,
We obtain
It is clear from the preceding definition that
is a solution of problem (2.3)-(2.4) on
for
.
Definition 2.9.
Let
. A function
is called a mild solution of (2.12)-(2.13) if
The proof of the next result is in [20]. For reader's convenience, we will give the proof.
Theorem 2.10.
Let
and
. If
, then the mild solution of (2.12)-(2.13) is a classical solution.
Proof.
To begin with, we study the case in which
. Let
be the mild solution of (2.12)-(2.13) and assume that
. It is easy to see that
and
where
is given by
. From [7, Lemma 3.12], we obtain that
is a classical solution and satisfies
Moreover, from (2.24) and taking into account that
for all
and
, we deduce the existence of constants
,
(which are independent from
) such that
Now, we assume that
. Let
be a sequence in
such that
in
. From [7, Lemma 3.12], we know that
,
, is a classical solution of (2.12)-(2.13) with
in the place of
. By using the estimate (2.25), we deduce the existence of functions
, such that
in
and
in
. These facts, jointly with our assumptions on
, permit to conclude that
in
. On the other hand,
we obtain
Now, by making
on
we conclude that
is a classical solution of (2.12)-(2.13). The proof is finished.
The proof of the next result is in [20]. For reader's convenience, we will give the proof.
Theorem 2.11.
Let
and
. If
, then the mild solution of (2.12)-(2.13) is a classical solution.
Proof.
Let
, there is
on
such that
on
and
on
. Put
proceeding as in the proof of Theorem 2.10. It follows that
is a classical solution of (2.12)-(2.13). Moreover, from [7, Lemma 3.13], we obtain
from which we deduce the existence of positive constants
(independent from
) such that
With similar arguments as in the proof of Theorem 2.10, we conclude that
is a classical solution of (2.12)-(2.13). We omit additional details. The proof is completed.
To establish our existence results, we need the following Lemma.
Lemma 2.12.
Let
. If
is compact for some
, then
and
are compact for all
.
Proof.
From the resolvent identity, it follows that
is compact for every
. We have from [20, Lemma 2.2] that
is a compact operator for
; therefore,
is a compact operator for
.
From [20, Lemma 2.5], we have,
is uniformly continuous for
, for any
fixed, we can select points
, such that if
, we obtain
, for all
and
.
Therefore, for all
, we have that
Noting now that
from (2.31), we find that
Thus,
where
is compact and
, then we observe that
as
. This permits us to conclude that
is relatively compact in
. This proves that
is a compact operator for all
.
For completeness, we include the following well-known result.
Theorem 2.13 (Leray-Schauder alternative, [21, Theorem
]).
Let
be a closed convex subset of a Banach space
with
. Let
be a completely continuous map. Then,
has a fixed point in
or the set
is unbounded.