Abstract

For the evolution equation with a scalar type spectral operator in a Banach space, conditions on are found that are necessary and sufficient for all weak solutions of the equation on to be strongly infinite differentiable on or . Certain effects of smoothness improvement of the weak solutions are analyzed.

1. Introduction

Consider the evolution equation with a scalar type spectral operator in a complex Banach space .

Following [1], we understand by a weak solution of equation (1.1) on an interval () such a vector function that (1) is strongly continuous on ; (2)for any , ( is the domain of an operator, is the operator adjoint to , and is the pairing between the space and its dual ).

As was shown [2], Theorem 4.2, the general weak solution of (1.1) on () is as follows: the operator exponentials , , defined in the sense of the operational calculus for scalar type spectral operators (see Section 2).

Here, we are to find conditions on that are necessary and sufficient for all weak solutions of (1.1) on to be strongly infinite differentiable on or .

This goal attained; all the principal results of paper [3] and the corresponding ones of [4] obtain their natural generalizations.

2. Preliminaries

Henceforth, unless specified otherwise, let be a scalar type spectral operator in a complex Banach space with a norm , its spectral measure (the resolution of the identity), and the operator's spectrum, with the latter being the support for the former.

Observe that, in a Hilbert space, the scalar type spectral operators are those similar to the normal ones [5].

For scalar type spectral operators, there has been developed an operational calculus for Borel measurable functions defined on [6, 7]. With being such a function, a new scalar type spectral operator is defined as follows: where ( is the characteristic function of a set ) and are bounded scalar type spectral operators on defined in the same manner as for normal operators (see, e.g., [8, 9]).

In particular, The properties of the spectral measure, , and the operational calculus underlying our discourse are exhaustively delineated in [6, 7]. We will outline here a few noteworthy facts.

Due to its strong countable additivity, the spectral measure is bounded [10], that is, there is an such that ( is the -algebra of Borel sets in the complex plane ).

Observe that is used in (2.6) to designate the norm in the space of bounded linear operators on . We will adhere to this rather common economy of symbols adopting the same notation for the norm in the dual space as well.

For any and , let be the total variation of the complex-valued measure on .

As we discussed in [4, 11], with being an arbitrary Borel measurable function on , for any , , and , where is the constant from (2.6).

In particular, for , , and (or any Borel set containing ), we have Further, for a Borel measurable nonnegative function on , a , and a sequence of pairwise disjoint Borel sets in , Indeed, since for the spectral measure [6, 7] for the total variation, we have Whence, due to nonnegativity of [12], We shall need the following regions in the complex plane: where and are positive constants.

The terms spectral measure and operational calculus frequently referred to will be abbreviated to s.m. and o.c., respectively.

3. Differentiability of a Particular Weak Solution

Proposition 3.1. Let and be a subinterval of an interval . A weak solution of (1.1) on is times strongly differentiable on if and only if in which case,

Proof. β€œOnly if” part.
Let and suppose that a weak solution of (1.1) on is times strongly differentiable on a subinterval .
Then for any , Hence, by the closedness of the operator (cf. [1]), This concludes proving the β€œonly if” part for , in which case, as is to be noted, the subinterval can shrink to a single point .
Let and let the interval be not a singleton. Then differentiating (3.4) at an arbitrary , we obtain with the increments being such that .
Since by the closedness of , we infer that Considering (3.4) and (3.7), Continuing inductively in this manner, we arrive at β€œIf” part.
Let be a weak solution of (1.1) on an interval () such that for an and a subinterval , As was discussed (see Section 1), with some
The fact that , , implies by the properties of the o.c. and [2], Proposition 3.1, that, for any , Given a and an arbitrary , let us choose a segment () so that is its midpoint if or if . For increments such that and any , we have Indeed, for any and an arbitrary ,
Therefore, for any , For , , let us fix an arbitrary and consider the sequence of vector functions Since , , by the properties of the o.c., Due to the boundedness of , , by the properties of the o.c., , , is a bounded linear operator on (, , [7]). Hence, the vector functions are strongly continuous on .
For an arbitrary segment , we have Hence, for any , the vector function , , is strongly continuous on , being the uniform limit of the sequence of strongly continuous on vector functions on any segment .
Let us fix an arbitrary and integrate (3.15) for between and an arbitrary . Considering the strong continuity of , , we have Whence, as follows from the Hahn-Banach Theorem, By the strong continuity of , , Consequently, by (3.15), for , Whence, analogously, Continuing inductively in this manner, we infer that, for any ,

Corollary 3.2. Let be a subinterval of an interval . A weak solution of (1.1) on is strongly infinite differentiable on if and only if in which case,

Thus, we have obtained generalizations of Proposition 4.1 and Corollary 4.1 of [3], respectively.

4. Differentiability of Weak Solutions

Theorem 4.1. Every weak solution of (1.1) on is strongly infinite differentiable on if and only if there is a such that the set is bounded.

Proof. β€œIf” part.
Let be such that the set is bounded and a weak solution of (1.1) on . Then (see Section 1) with some
For any , and an arbitrary , Indeed, the former integral is finite due to the boundedness of the set , the finiteness of the measure and the continuity of the integrated function on .
For the latter one, we have Further, for any and , Indeed, since, due to the boundedness of the set , the set is void for all sufficiently large 's, In addition to this
By the properties of the o.c. and [2], Proposition 3.1, (4.2) and (4.4) imply that
Then, by Corollary 3.2, is strongly infinite differentiable on .
β€œOnly if” part.
We will prove this part by contrapositive.
Assume that, for any , the set is unbounded.
In particular, for any , unbounded is the set .
Hence, we can choose a sequence of points in the complex plane as follows:
The latter implies in particular that the points are distinct (, ).
Since, for any , the set is open in , there exists an such that this set along with the point contains the open disk of radius centered at Then, for any , , Further, since the points , , are distinct, we can regard the radii of the disks, , , to be small enough so that Whence, by the properties of the s.m., Observe also, that the subspaces , , are nontrivial since , with being an open set.
By choosing a unit vector (), , we obtain a vector sequence such that ( is the Kronecker delta symbol).
As is readily verified, (4.14) implies that the vectors are linearly independent.
Moreover, there is an such that Indeed, the opposite implies the existence of a subsequence such that Then, for any , by selecting a vector such that , considering (4.14) and (2.6), we arrive at the contradiction: As follows from the Hahn-Banach Theorem, for any , there is an such that Concerning the sequence of the real parts, , there are two possibilities: it is either bounded or unbounded above.
Suppose that the sequence is bounded above, that is, there is such an that Let By (4.14), For any and an arbitrary , Similarly, for any , By [2], Proposition 3.1, (4.22) and (4.23) imply that
Therefore (see Section 1), , , is a weak solution of (1.1) on .
Let (cf. (4.18)).
Considering (4.18) and (4.15), we have Similarly to (4.22), Whence, by [2], Proposition 3.1, .
Consequently, by Proposition 3.1, the weak solution , , of (1.1) on is not strongly differentiable at 0.
Now, assume that the sequence is unbounded above.
Therefore, there is a subsequence such that Then the vector is well defined.
By (4.14), For any and an arbitrary , similar to (4.22), Indeed, for all sufficiently large 's, and by (4.27), Analogously, for an arbitrary ,
From (4.30) and (4.32), by [2], Proposition 3.1, we infer that
Therefore (see Section 1), , , is a weak solution of (1.1) on .
For any , , by (4.11), (4.12), and (4.27), Hence, for , , Using this estimate, we obtain Whence, by [2], Proposition 3.1, .
Therefore, by Proposition 3.1, the weak solution , , of (1.1) on is not strongly differentiable at 0.
With every possibility concerning considered, we infer that the opposite to the assumption that, for a certain , the set is bounded, allows to single out a weak solution of (1.1) on that is not strongly differentiable at 0, much less strongly infinite differentiable on .
Thus, the β€œonly if” part has been proved by contrapositive.

Theorem 5.1 of [3] has been generalized.

Theorem 4.2. Every weak solution of (1.1) on is strongly infinite differentiable on if and only if there is a such that, for any , the set is bounded.

Proof. β€œIf” part.
Let be such that, for an arbitrary , the set is bounded.
Let be a weak solution of (1.1) on . Then (see Section 1) with some
For any , , and an arbitrary ,
The first integral in this sum is finite due the boundedness of the set , the finiteness of the measure , and the continuity of the integrated function on .
The finiteness of the second integral is proved in absolutely the same manner as for the corresponding integral in the proof of the β€œif” part of Theorem 4.1.
Finally for the third one, considering that is arbitrary and setting , we have Further, for any and , Indeed, since, due to the boundedness of the set , the set is void for all sufficiently large 's, Similarly to the β€œif” part of Theorem 4.1 and (4.38), we have By the properties of the o.c. and [2], Proposition 3.1, (4.37) and (4.39) imply that Whence, by Corollary 3.2, is strongly infinite differentiable on .
β€œOnly if” part.
As well as in Theorem 4.1, we will prove this part by contrapositive.
Thus, we assume that, for any , there is such a that the set is unbounded.
Let us show that this assumption can even be strengthened. To wit: there is such a that, for any , the set is unbounded.
Indeed, there are two possibilities: (1)for a certain , the set is unbounded; (2)for any , the set is bounded.
In the first case, the set is unbounded with the very , for which the set is unbounded, and an arbitrary .
In the second case, based on the premise we infer that, for any , the set is unbounded. Then so is the set for any and .
Thus, let us fix a such that the set is unbounded for an arbitrary .
In particular, for any , the set is unbounded.
Hence, we can select a sequence of points in the complex plane in the following manner: The latter, in particular, implies that the points are distinct.
Since, for any , the set is open in , there exists such an that this set, along with the point , contains the open disk of radius centered at Hence, for any , , Since the points are distinct, we can regard the radii of the disks, , to be small enough so that
By the properties of the s.m., the latter implies Observe that the subspaces , , are nontrivial since , being an open set.
By choosing a unit vector (), , we obtain a vector sequence such that In the same manner as in the proof of Theorem 4.1, one can show that there is an such that As follows from the Hahn-Banach Theorem, for any , there is an such that Concerning the sequence of the real parts, , there are two possibilities: it is either bounded or unbounded.
First, assume that the sequence is bounded, that is, there is an such that Let By (4.49), In absolutely the same fashion as it was done in the case of bounded above sequence in the proof of the β€œonly if” part of Theorem 4.1, it is shown that
Therefore (see Section 1), , , is a weak solution of (1.1) on .
Let (cf. (4.51)).
Taking (4.50) and (4.51) into account, we have Thus, Whence, by the properties of the o.c. and [2], Proposition 3.1, .
Consequently, by Proposition 3.1, the weak solution , of (1.1) on is not once strongly differentiable on .
Now, assume that the sequence is unbounded. Therefore, there is a subsequence such that Suppose that as .
Without restricting generality, we can regard that Considering this case just like the analogous one in the proof of the β€œonly if” part of Theorem 4.1, one can show that the vector belongs to and, hence, (see Section 1) , , is a weak solution of (1.1) on .
For any , , by (4.46), (4.47), and (4.59), Hence, for , , Using this estimate, we have Whence, by the properties of the o.c. and [2], Proposition 3.1, .
Therefore, by Proposition 3.1, the weak solution , , of (1.1) on is not once strongly differentiable on .
Now, suppose that as .
Without restricting generality, we can regard that Let For any and an arbitrary , Analogously, for an arbitrary , From (4.66) and (4.67), by [2], Proposition 3.1, we infer that
Therefore (see Section 1), , , is a weak solution of (1.1) on .
Let (cf. (4.51)).
Taking into account (4.50) and (4.51), we have For any , , by (4.47) and (4.64), Hence, for , , Using these estimates, we have Whence, by the properties of the o.c. and [2], Proposition 3.1, .
Therefore, by Proposition 3.1, the weak solution , , of (1.1) on is not once strongly differentiable on .
With every possibility concerning considered, we infer that the opposite to the assumption that there is a such that, for any , the set is bounded, allows to single out a weak solution of (1.1) on that is not once strongly differentiable on , much less strongly infinite differentiable on .
Thus, the β€œonly if” part has been proved by contrapositive.

Theorem 5.2 of [3] and Theorem 4.1 of [4] have been generalized.

5. Certain Effects of Smoothness Improvement

As we observed in the proofs of the β€œonly if” parts of Theorems 4.1 and 4.2, the opposits to the β€œif” parts’ premises imply that there is a weak solution of (1.1) on , which is not strongly differentiable at 0 or, respectively, once strongly differentiable on .

Therefore, the case of finite strong differentiability of the weak solutions is superfluous and we obtain the following effects of smoothness improvement.

Proposition 5.1. If every weak solution of (1.1) on is strongly differentiable at 0, then all of them are strongly infinite differentiable on .

Proposition 5.2. If every weak solution of (1.1) on is once strongly differentiable on , then all of them are strongly infinite differentiable on .

These statements generalize Propositions 6.1 and 6.2 of [3], respectively, the latter agreeing with the case when is a linear operator (not necessarily spectral of scalar type) generating a -semigroup (cf. [1, 13]).

6. Final Remarks

Due to the scalar type spectrality of the operator , all the above criteria are formulated exclusively in terms of its spectrum, no restrictions on the operator's resolvent behavior necessary, which makes them inherently qualitative and more transparent than similar results for -semigroups (cf. [14]).

If a scalar type spectral generates a -semigroup (cf. [15]), we immediately obtain the results of paper [4] regarding infinite differentiability.

Acknowledgment

The author would like to express his gratitude to Mr. Michael and Jonell McCullough.