2. Bilateral Fractional Integral and Derivative
In this paper
is a nonempty (possibly unbounded) open interval,
u is a real function of one variable and
. The support of a function
u is denoted by
. The notation
stands for the classical pointwise derivative;
, or shortly
D, denotes the distributional derivative with respect to the variable
x. For every open interval
, we denote by
the set of absolutely continuous functions with the domain in the interval
A, which coincides ([
20]) with the space the Gagliardo–Sobolev space
when they are both endowed with the standard norm
. Moreover, we set
as the set of measurable functions which are Lebesgue integrable on every compact subset of
A, and
and
, where
denotes the measures whose total variation on
A is bounded. We denote by
and
respectively the space of distributions and the space of tempered distributions on the open set
A. We denote by
the space of of Hölder continuous functions on the set
K.
For the reader’s convenience, we recall the definition of Gagliardo’s fractional Sobolev spaces
([
21,
22]). For any
, we set
which is a Banach space endowed with the norm
and we recall also the definition of the Riemann–Liouville fractional integral and derivative of order
s for
-functions, whose standard references can be found in the book by Samko et al. [
9].
In the sequel, H denotes the Heaviside function if , if , while denotes the sign function if , if , .
Definition 1. (Riemann–Liouville fractional integral)
Assume and .
The left-side and right-side Riemann–Liouville fractional integrals and are defined by setting, respectively, Here,
denotes the Euler gamma function [
23].
Notice that and in general, for every strictly positive integer value, , coincides with the n-th order primitive, vanishing at together with all derivatives up to order .
Both
and
are absolutely continuous functions if
since they are primitives of
functions, whereas we can only say that they are
functions if
,
(see [
9]): indeed, jump discontinuities are allowed if
, as shown by the next example.
Example 1. Set . Then for every there is , , s.t. is discontinuous. For instance, consider , thus, exploiting the Euler beta function , one gets Thus, is a piecewise constant function on with a jump at .
Next, we recall the classical definition of left and right Riemann–Liouville fractional derivatives as in [
9,
24,
25,
26,
27].
Definition 2. (Classical Riemann–Liouville fractional derivative)
Assume and .
The left Riemann–Liouville derivative of u at is defined byfor every value of x such that this derivative exists. Similarly, we may define the right Riemann–Liouville derivative of u at asfor every value of x such that this derivative exists. Then we introduce the distributional Riemann–Liouville fractional derivative as in [
25]: a refinement of the Riemann–Liouville fractional derivative, obtained by the plain substitution of the pointwise classical derivative with the distributional derivative
Definition 3. (Distributional Riemann–Liouville fractional derivative)
Assume and . The distributional left Riemann–Liouville derivative of u, , is defined by Similarly, we may define the distributional right Riemann–Liouville derivative of u, , as Remark 1. The distributional Riemann–Liouville fractional derivatives provide a suitable refinement of the classical ones for the purposes of the present paper. However, we emphasize that they coincide on every function u such that is absolutely continuous, as it was always the case in the classical applications of fractional derivatives ([28,29,30]). In Lemma 5 below, we examine the case when the above pointwise-defined derivative exists a.e. and defines an function coincident with the distributional derivative D, respectively of and .
In the sequel, we omit the suffix of the interval without loss of information, since in this paper, we do not consider any other fractional derivative than the Riemann–Liouville one; we omit also the endpoints and suffix whenever they are clearly established.
Therefore, we will write shortly , , , , and , respectively, in place of , , , , and .
One of the disadvantages of the one-side Riemann–Liouville derivative and integral, as defined above, is the fact that only one endpoint of the interval plays a role (see (
56) and (
57) ) since they are “anisotropic” definitions (see [
31] and Lemma 6). On the other hand, if we aim to exploit such definitions in a variational context, we have to deal with boundary conditions so that both interval endpoints must play a role ([
32]). Therefore, we introduced the bilateral fractional integral and derivative, by keeping separate their “even” and “odd” parts:
Definition 4. For every we set the even and odd versions of bilateral fractional integrals and derivatives: Whenever
, the convolution in (
8)–(
11) has to be understood, without relabeling, as the convolution of the trivial extension of
u (still an
function with support on
) with either
or
(both belonging to
). Also
have to be understood, without relabeling, as the natural extension for
, provided by the convolution of the trivial extension of
u with the corresponding kernels (here,
H denotes the Heaviside function):
namely
Remark 2. In (9) and (11), denotes the distributional derivative in , but obviously its restriction as a distribution on the open set is understood whenever one works in the bounded interval . Up to a normalization constant (see (
25)),
is called the Riesz potential of
u ([
1,
9]). These fractional integrals
turn out to be in
(thus
on every bounded interval
I) for every
, since they are convolutions of
with an
kernel. Moreover, we have the next result.
Lemma 1. If , , and then belong to .
Proof. See Lemmas 2.5 and 3.6 (iii) in [
1]. □
The behavior of all the above operators, as
or
, is clarified by subsequent Lemmas of the present section, whose proof can be found in [
1].
Notice that both and belong to , for ; hence, the convolution with any function is well defined and belongs to ; moreover in as , while has no limit in as , where denotes the space of tempered distributions.
Fractional derivatives degenerate developing singularities as ; nevertheless, they can be made convergent to meaningful limits by suitable normalization.
Lemma 2. Assume , and choose the constants in the Fourier transform such that . Then Remark 3. Notice that relations (21), (23) and (24) tell that, as , both (left Riemann–Liouville fractional derivative of order s of u) and (even Riemann–Liouville fractional derivative of order s of u) converge in to the distributional derivative , while converges in to . On the other hand relationship (22) means that (odd Riemann–Liouville fractional derivative of order s of u) fades as but, when suitably normalized as , it converges in to the Gagliardo fractional derivative of order 1 of u, say . Fractional integrals degenerate developing singularities as ; indeed the convolution term fulfills in as ; nevertheless, fractional integrals are convergent to meaningful limits by suitable normalization.
Lemma 3. Assume , with and set the constants in the Fourier transform such that . Then Lemma 4. Assume , .
If , then If , then If , then If , then Lemma 5. Assume , . Then If in addition , then If in addition , then Remark 4. Every distributional fractional derivative (left, right, even, and odd) appearing in the statements of Lemmas 3–5, which are proved in [1] with fractional classical derivatives , still hold true in the present formulation with corresponding distributional derivatives by exactly the same proof, since the assumptions ensure that all derivatives are evaluated on local absolute continuous functions. Remark 5. Notice that, when is replaced by a bounded interval, the identities (27), (28), (33) and (34) require an additional correction term, taking into account of boundary values (see (52) and (53) in Theorem 1), whereas (31) and (32) remain true (see (136), (137)). Symmetries of even or odd functions are inherited neither by fractional integrals, nor by fractional derivatives. Nevertheless, the next lemma holds true.
Lemma 6. For every , and , by setting For every , and every even function , we get For every and every odd function , we obtain Proof.
By inserting
in place of
s in (
36), if
v is even we obtain (
37) via
Even
v entails
,
,
and
; hence, (
36) and (
37) entail, respectively, (
38) and (
39).
Odd
v entails
,
,
and
; hence, (
36), (
37) entail, respectively, (
40), (
41). □
Results listed above (mainly Lemmas 3 and 4 proved in [
1]) lead to the natural definition of the operators representing the bilateral version of Riemann–Liouville fractional derivatives and integrals, as stated below. Results similar to the ones in Lemma 6 can be found also in [
33].
Definition 5. (Bilateral Riemann–Liouville fractional integral of order s) Definition 6. (Bilateral Riemann–Liouville fractional derivative of order s) 3. The Bilateral Fractional Sobolev Space
From now on, we consider only functions defined on a bounded interval .
As already mentioned in [
25], possible naïve definitions of bilateral fractional Sobolev spaces could be set by
, where
and
for example, a definition which refers to
-functions whose classical Riemann–Liouville fractional derivative of prescribed order
exists finitely almost everywhere and belongs to
.
Actually, if the classical Riemann–Liouville fractional derivative
of
u exists a.e. for
x for some
, then
is differentiable almost everywhere, referring to the same
s; nevertheless, such an a.e. derivative does not provide complete information about the distributional derivative of the fractional integral
, when
is not an absolutely continuous function. Thus, the differential properties are not completely described by the pointwise fractional derivative, though existing almost everywhere in
. This shows that the previous definitions
and
are not suitable to obtain an integration by parts formula, whereas the appropriate ones refer to distributional Riemann–Liouville fractional derivative
in Definition 3, namely, they are given by
Therefore, to develop a satisfactory theory of fractional Sobolev spaces, we introduced a more effective function space in [
25], by defining the fractional Sobolev spaces related to one-sided fractional derivatives, which are recalled in subsequent Definition 7, where we confine to the case
.
Definition 7. We recall the definitions of Riemann–Liouville fractional Sobolev spaces related to one-sided fractional derivatives, as introduced in [25]: Explicitly, the properties entail, respectively, that the distributional derivatives belong to , thus and .
Here, we introduce also the “even” and “odd” fractional Sobolev spaces.
Definition 8. The even/odd Riemann–Liouville fractional Sobolev spaces are Eventually, we define the bilateral Riemann–Liouville fractional Sobolev spaces, with the aim to achieve a symmetric framework.
Definition 9. The (Bilateral) Riemann–Liouville Fractional Sobolev spaces.
For every , we set , that is, Notice that, concerning Definition 9, by exploiting (12) and (13), we get also Theorem 1. Assume and is bounded. Then, the (bilateral) Riemann–Liouville fractional Sobolev space (Definition 9) is a normed space when endowed with the natural norm The set is a Banach space and, for every there is such that Every can be represented by both Proof. The map is a norm on , indeed,
is equivalent to the norm since belongs to , and ; analogously is a norm for , due to , and .
The completeness of with respect to such a norm when is bounded, follows by the completeness of and together with the fact that is a Cauchy sequence in the norm if and only if is a Cauchy sequence in and are Cauchy sequences in .
Estimate (
51) and representations (
52) and (
53) follow by (
129), (
130) and (
135) of Proposition 12, which is shown in
Section 5. □
Remark 6. Thanks to , we have replaced the terms in the norm (50) by , where D denotes the distributional derivative. Obviously, the terms and in the norm (50) could be alternatively replaced by and , referring to (8), (16) and (17), still achieving an equivalent norm on . Example 2. For every , the constant functions and both belong to the space . Spaces , test functions on , and , continuously differentiable functions, are contained in .
Example 3. For every , the discontinuous piecewise constant belongs to .
Indeed, both and belong to .
Example 4. For every and , the function belongs to . This claim is straightforward for and ; we refer to (97) and (98) for Example 5. Function with belongs to if and only if .
Indeed, if , while , summarizing belongs to for ; on the other hand, belongs to for and is bounded on if and only if , due to Summarizing, and taking into account Example 4 for , In the particular case , we recover with and unbounded in a right neighborhood of .
Theorem 2. (Integration by parts in)
Next, identities hold true for , : Proof. Identity (
56) follows by (
2), (
5) and
Identity (
57) follows by (
3), (
4) and similar computations.
Identities (
58) and (
59) follow by the subtraction and sum of (
56) and (
57). □
Remark 7. Notice that when u is representable, e.g., under a slightly stronger condition, then we find a more symmetric formulation. For instance, (59) translates into Lemma 7. holds true with the related uniform estimate: there is a constant such that Proof. By computations in Example 3, we know that the Heaviside function belongs to ; thus if the embedding holds true, then it is strict.
Recalling the definition ([
34]) of right Caputo fractional derivative
and its relationship with the right Riemann–Liouville fractional derivative
and taking into account
we get (7) and (
62). □
Theorem 3. [Compactness in]
Assume that the interval is bounded, the parameter s fulfills , and Then there exist , , and a subsequence such that, without relabeling, Proof. Claim (i) follows by (
51) and (
63) and reflexivity of
for any fixed
; thus, by choosing a sequence
and extracting a diagonal sequence, we get the claim for a unique subsequence and unique
u valid for every
q fulfilling
. Moreover, such
u belongs to
. Eventually, for
there is a measure
such that
in
, but such
must be equal to
u, then
in
.
The compact embedding
valid for any
(Rellich Theorem) entails the existence of
and
in
fulfilling, up to subsequences,
By
and the Mazur Theorem, there is a sequence of convex combinations
, which is strongly converging: precisely,
strongly in
for every
with
,
,
. Hence, by (
63),
is a continuous map from
to
,
and
, hence, we obtain
and hence, in
. Moreover, by (
69),
is bounded in
; then, there exists
such that, possibly up to subsequences,
Taking into account (
70), (
71) and the uniqueness of limit in
, we obtain
Taking into account , we set , so solves Abel integral equation . By the semigroup property, ; hence, . Therefore (by Proposition 2), the Abel integral equations have a unique solution in , given by , .
Set . . So u solves the Abel equation . Moreover, by the semigroup property, ; hence, . Therefore (Proposition 3), the Abel equation has a unique solution in , given by .
By
,
strongly in
, hence,
Then, by (
65),
. Hence we have shown claims (ii) and (iii).
Moreover, the convergence is also in the sense of distributions and the sequence is bounded in
; therefore,
belongs to
and, again up to subsequences,
We can deal with by the same argument, exploiting Corollary 1 for the backward Abel integral equation , leading to . □
Remark 8. The boundedness of is an essential assumption in the previous compactness theorem, not only to exploit the Rellich theorem, but also to avoid slow non-integrable decay at infinity of the fractional integral: indeed, even for an integrable compactly supported u, we may have at , e.g., if .
Remark 9. We emphasize that in Theorem 3, we cannot improve (64), since may belong to . Indeed, we can choose if , if , if and if . Thus, f belongs to and is uniformly bounded. Solving the Abel equations and with Propositions 2 and 3 provides , whereas is uniformly bounded in ; hence, is uniformly bounded in , due to Lemma 7.
We recall a well-known result [
9] (Theorem 2.1) concerning the
-representability of functions.
Theorem 4. [-representability] Given , then
for some if and only if for some if and only if Moreover, in the affirmative case, say, when there exists such that (resp. ), we obtain In
Section 5, we provide a self-contained proof of the above result together with a discussion of the related forward and backward Abel equation in the distributional framework, even in the cases when
or
(see Propositions 2 and 3 and Corollaries 1 and 2).
Here, we show that the representability result has a natural extension to the bilateral case.
Proof. Since
the claim follows by Definition 9, Theorem 4, Proposition 1 (semigroup property of fractional integral), Propositions 2 and 3 and Corollaries 1 and 2. □
Next, we make explicit some embedding relationship between and .
Theorem 6. The following strict embeddings hold true: where we refer to Definitions 7–9 about , shortly denoted here, versus the naïve definition at the beginning of the present Section 3. Proof. Without loss of generality, we assume .
Strict embeddings of
in
and of
in
are shown respectively by
and
: see (
55) in Example 5.
Therefore, in order to show (
74), it is sufficient to show an example for the strict embedding
: indeed, the proof of
is achieved by replacing the variable
t with
in the counterexample showing the other strict embedding by exploiting the symmetry with respect to
, analogous to the one with respect to
in (
36) and (
37).
We first note that
follows by definition (
4): the existence of a weak derivative in
of
entails the existence of the fractional derivative
, coincident with the almost everywhere defined fractional derivative
(x).
The strict embeddings , and follows by the subsequent argument, which, for any fixed , provides the existence of a function in and a function in .
Given , we show a function z in such that .
Precisely, by denoting
V the Cantor–Vitali function on
([
21]), we claim that
Indeed
V is
-Hölder continuous with
. So
belongs to
for every
by Theorem 3.1 in [
9] and the fact that
. Therefore
(hence
) for
. Moreover,
for
: indeed, due to continuity of
V in
,
is continuous in
and we obtain
By Hölder continuity
, we obtain
Then
Therefore, the limit above is equal to 0, as , thus proving the claim .
Summarizing, , and , for .
Therefore we can consider the Abel integral equation in the distributional setting
and solve it; by Proposition 3, the unique solution is given by
, and fulfils
. Moreover
a.e. on
, whereas
, which is a nontrivial bounded measure. Explicitly
fulfills
. So far, we have proved the first embedding chain in (
74) for
.
In the sequel, we show that, given any , we can adapt the Cantor–Vitali function in such a way that it is s-Hölder continuous for any ; hence, we recover the strict embedding for any s in , and hence, for any s in , due to the generic choice of .
Indeed, given
, we can replace the construction of Cantor
-middle set
(say, a set whose Hausdorff dimension is
, which leads to the
Hölder continuous Cantor–Vitali function
) with the Cantor-like
-middle set
, with Hausdorff dimension
, which leads to the
Hölder continuous Cantor–Vitali generalized function
, where
Notice that as and as , so that spans the interval as runs over . Moreover, for .
Again by Proposition 3, we get that is representable, say there exists (unique) s.t. s.t. for , and we claim that , , for : indeed these claims about the generalized Cantor–Vitali function can be proved by the same procedure dealing with the definition of , as it is sketched below.
The function is of bounded variations since is monotone, as it is the uniform limit of a sequence of monotone nondecreasing functions.
Continuity of follows from uniform convergence of standard iterative approximations by piecewise linear functions. The absolutely continuous part of the distributional derivative is identically 0 since is locally constant on an open set of Lebesgue measure 1: indeed, it is a union of open intervals, which is iteratively obtained by approximation with finite unions whose measure fulfills the recursive scheme: , , so that as .
The worst case for differential quotients of
n-th approximations of
is provided by
, so that
is the biggest real
s.t.
is uniformly bounded for
, say
So
belongs to
for every
by Theorem 3.1 of [
9] and taking into account that
. Therefore,
that is
for
.
Moreover,
. Indeed, by continuity of
in
, we obtain
Since
, we get
Thus
Summarizing, if
is the generalized Cantor–Vitali function and
,
since
is a nontrivial Cantor measure with no atomic part, whereas
a.e.; moreover,
where, to achieve (
77), we exploit Proposition 2 to solve the backward Abel integral equation in the distributional framework
; indeed,
,
, then the unique solution
of
is
, which fulfills
. Hence, by evaluating the distributional derivative
D, we get
which is a nontrivial Cantor measure with no atomic part, whereas
a.e. □
We list some properties concerning the comparison of bilateral Riemann–Liouville fractional Sobolev spaces
with classical spaces: Gagliardo fractional Sobolev spaces
, functions of bounded variation
and
, De Giorgi’s space of special bounded variation functions, whose derivatives have no Cantor part ([
21,
35] for example).
Theorem 7. Let be such that . Then with continuous injection, say Proof. Straightforward consequence of Theorem 3.2 of [
25] and Definition 9. □
In [
25], we have compared
and
with
, and proved
This inclusion was refined by a recent result (Theorem 3.4 in [
2]) showing
On the other hand, for every
,
is contained neither in
nor in
, due to remarkable examples of Weierstrass-type functions. Indeed a Weierstrass function
w can be defined ([
36]) so that
w belongs to
, but
w does not belong to
since it is nowhere differentiable. Fix
and set
Notice that the the constant subtraction entails , thus preventing a singularity of at .
Theorem 8. Let be such that . Then with continuous injection. Precisely, .
We emphasize that in the case of an unbounded interval , there is no chance for a compactness statement analogous to Theorem 3 in (a, b), since the Rellich theorem cannot be applied.
On the other hand, the property entails a stronger qualitative condition on u than in the case of with a boundedness of , as clarified by the next remark.
Remark 10. If , , then and is bounded in a neighborhood of . Property may fail for if .
Indeed, entails , hence , , hence then, exploiting the Fourier transform , , hence is bounded and .
If , then, referring to (8) and (10), neither nor belong to (or even to with , when ), moreover the summability of may fail at infinity due to a decay of order , therefore may be unbounded around . Remark 11. Notwithstanding Remark 10 (excluding nontrivial constant functions from the space ), if we restrict to bounded intervals, a constant function belongs to , for every bounded interval and every value of K. Indeed, 4. Bilateral Fractional Bounded Variation Space
Possible naïve definitions could be provided, for
, by
which refer to
-functions whose classical pointwise-defined Riemann–Liouville fractional derivative of prescribed order
is a bounded measure.
Actually, if the Riemann–Liouville fractional derivative of u exists for a.e. x for some , then is differentiable almost everywhere, referring to the same s; nevertheless, we have no information on the distributional derivative of the fractional integral .
These differential properties are not completely described by the point-wise derivative, though it exists almost everywhere. This shows that the previous definitions of
,
and
are not suitable to obtain an integration by the parts formula. Therefore, to develop a satisfactory theory of fractional bounded variation spaces, as we did for fractional Sobolev spaces in [
25], we introduce a more suitable function space: the
bilateral fractional bounded variation space , as defined in the sequel.
Remark 12. We recall that, as long as these classical fractional derivatives are evaluated on absolutely continuous functions, as it was done in all previous section, using the operators of the classical Definition 2 provides the same results as the distributional Definition 3: for this reason, we keep the usual classical notations (, and the corresponding short forms , ). However, in the present section, we evaluate fractional derivatives on functions of bounded variations, a setting where the two definitions provide different evaluations.
Next, inspired by [
2], where the nonsymmetric spaces are studied also in the case of higher order derivatives, we introduce the bilateral Riemann–Liouville bounded variation space, with the aim to achieve a symmetric framework.
Definition 10. The (bilateral) Riemann–Liouville fractional bounded variation spaces. For every , we set where, referring to Definition 3, Theorem 9. Assume that the interval is bounded and the parameter s fulfills .
Then, the space is a normed space endowed with the norm Contribution in the norm (81) can be replaced by . Moreover, is a Banach space and for every , there is , such that Every can be represented by both Proof. We emphasize that here
and
replace, respectively,
and
which were in representations (
52) and (
53) of
functions, since in the present
setting, there are not pointwise defined values, though there are well-defined finite right and left limits at every point in
.
The map is a norm on , indeed,
is equivalent to the norm
, since
belongs to
,
and
; analogously
is a norm for
, due to
,
and
. Therefore, terms
can be replaced, respectively, by
in the natural norm
The other claims follow by the same proof of Theorem 1 for the fractional Sobolev setting, where actually only the Proposition 2 and Corollary 1 about Abel forward and backward integral equation must be suitably tuned as stated in Remark 17. □
Example 6. The constant functions and belong to the space . In general, the space of test function on is contained in .
Example 7. Heaviside function H belongs to , thanks to Example 3.
Example 8. Function belongs to if , since due to Example 1.
Due to the unboundedness of in a right neighborhood of (due to Example 5), we obtain that does not belong to .
In general, for , belongs to if .
Theorem 10. (integration by parts in )
Next, identities hold true for , : Proof. Exactly the same proof of Theorem 2, but the facts that, here, the distributional derivatives in replaces the almost everywhere pointwise derivative in and the integrals at the left-hand side are evaluated with respect to the measures , , and , in place of Lebesgue measure. □
Theorem 11. [Compactness in]
Assume that , the interval is bounded and Then, there exist and a subsequence such that, without relabeling, Proof. The proof can be achieved by exactly the same argument used in the proof of compactness in (Theorem 3). □
Remark 13. since and .
Indeed, the first embedding follows by (78) and is strict due to (79); the second embedding is obviously strict, about the third embedding notice that (see Example 8). In addition, we can rewrite (74) as follows since, referring to notations (76) and (77) in the proof of Theorem 6, 5. Abel Equation in and Some Useful Relationships
Here, for the reader’s convenience, first, we recall some basic algebra of fractional differential calculus, then we extend to the distributional setting some classical results about Abel integral equations: these suitably tuned claims are exploited in
Section 3 and
Section 4 to prove the main properties of
and
, with
: Theorems 1, 3, 5, 6, 9 and 11.
All the results stated in this section are independent of the ones of previous sections.
To avoid confusion with the standard notation of variable s in the Laplace transform, here, we label by , instead of s, the index of fractional integral, fractional derivative and fractional Sobolev space.
All along this section: the Laplace transformable function refers to a measurable function v on with support contained in such that there exists for which is a Lebesgue-integrable function; the Laplace transformable distribution refers to a distribution v on with support contained in such that there exists for which is a tempered distribution; and in all cases, denotes the Laplace transform of v.
First we recall some relationships concerning fractional integral of powers of
x in
:
where
is the incomplete Beta function:
.
Hence, since both conditions
and
hold true when
, one obtains the fractional derivative of power functions of
x in
:
In the particular case
we obtain
Thus has a nontrivial kernel, as it is the case of the linear operators .
More in general, by (
100), we know that
The converse holds too (see Proposition 8).
Proposition 1. (Semigroup property of fractional integral )
For every , , with , we have In general, if , , , , then if , , , , then Proof. Consider the trivial extension of
v and the standard extension of related subsequent fractional integrals as defined by
We assume first
. By denoting
V, the Laplace transform of
v and taking into account of
and (
97), we obtain, for
,
hence, claim (
103) follows by the injectivity of the Laplace transform.
hence, claim (
104) follows by the injectivity of the Laplace transform.
In general, we obtain, for
,
which proves (
105). Identity (
106) is achieved in the same way. The case of general
is achieved by translation. Moreover, given
with
, by considering the trivial extension of
v on
, we obtain
hence, (
107) is proved. Identity (
108) is achieved in the same way. □
Proposition 2. Assume , , and f belongs to .
Then the Abel integral equation admits the solution u given by which is unique among Laplace-transformable functions evaluated with translated variable .
Proof. Whenever necessary, we consider the trivial extension (namely, 0 valued) on of every function and if necessary on , without relabeling the function name. Thus, every related fractional integral set as a function defined over has a trivial extension, which coincides on with the same fractional integral of the trivial extension, namely, it has support contained on .
First, we assume
. In such a case,
f is a Laplace-transformable function: we denote by
and
their Laplace transform evaluated at the variable
s. If a Laplace transformable solution
u exists, then its Laplace transform
must fulfill the transformed equation. We have
We evaluate
: reminding that
, where
w is any
-transformable distribution, and here,
and
denote respectively the distributional derivative on the open set
and on
. By taking into account that
belongs to
, we know that
is a well-defined real value. Thus, by formula
applied to
under the assumption
, we obtain
where the four last equalities are understood in the sense a.e. on
, coherently with the fact that
because it coincides with the derivative of the function
and vanishes on
. Moreover
u is unique due to the injectivity of the Laplace transform. Then (
110) is proved when
.
If
, we can exploit the solution Formula (
111) proved in case
: assume
on
and set
; then
and
have support on
and, hence, are Laplace transformable functions.
Thus we have the Abel equation
, that is
By (
111), we get
that is
□
Remark 14. At a first glance, both technical assumptions in Proposition 2, namely and , may look strange or unnatural.
However, they cannot be circumvented: actually, they are both necessary conditions for the existence of a solution of Equation (109). Let us check this claim: if such a solution as exists, then belongs to ; moreover, due to the semigroup property of fractional integrals (see Proposition 1), hence, is the primitive of an function; thus belongs to and .
Remark 15. Condition may be not easy to check. However, it can be replaced by stronger conditions, which are much easier to check. Indeed, if either there exists a finite value or f is bounded in a neighborhood of 0, then .
Remark 16. For the unnormalized Abel equation, , namely as a straightforward consequence of Proposition 2 and Euler reflection formula, , under the assumption , we recover the next formula for the unique solution u in :still under the requirement that necessary conditions and hold true. Now, we remove the assumption and look for solutions in .
Proposition 3. Assume that , and f belongs to the space .
Then, the Abel integral equation in the distributional framework admits a unique solution u among Laplace transformable distributions evaluated at (variable translation), which is the bounded measure on with support contained in given by In (116), actually u denotes the trivial extension outside , andrepresents the distributional convolution whose evaluation, namely f, is identically 0 on and possibly non-zero on . Proof. Same proof of Proposition 2. Only the step in (
111) with
has to be slightly modified: denoting by
and
the distributional derivative respectively in
and
, setting
,
,
,
and
we exploit the fact that
is a finite well-defined value (since
entails
), and we replace (
111) by
□
Corollary 1. Assume , , the value exists and is finite (possibly substituted by weaker condition ) and f belongs to .
Then, the backward Abel integral equation admits a solution u, unique among Laplace transformable functions evaluated at (sign change and translation), which is given by Proof. Taking into account that
, set
, and hence,
,
, and choose
. Then
Therefore, we can apply Proposition 2 to an Abel equation on
:
□
Corollary 2. Assume that , and f belongs to the space .
Then the backward Abel integral equation in the distributional framework admits a unique solution u among Laplace transformable distributions evaluated at (say with sign change and translation), which is the bounded measure with support contained in given by In (120), actually u denotes the trivial extension outside , andrepresents the distributional convolution whose evaluation, namely f, is identically 0 on and possibly non-zero on . Proof. Same proof of Corollary 1, but exploiting Proposition 3 instead of Proposition 2. Notice that the trivial extension of a function in has compact support and can be dealt with as a Laplace transformable distribution evaluated at the variable . □
Example 9. We mention some basic examples of solution u for Abel integral equationonwithand, more in general for distributional Abel integral equationandwith support condition.
If , then , for , due to Proposition 2.
If , then , for , due to Proposition 2.
If , then , for , , due to Proposition 2.
These relationships are deduced by Proposition 2: in the first and second item, notice that entails (see Remark 15), while in third item entails both and . Thus, we get the three claims above by applying the relationships If , , , then the solution u with support on to distributional backward Abel equation is given by .
Indeed , thus, by Proposition 3,
Then, u solves the Abel equation since, by representation (14), we obtain If then the solution u to backward Abel equation is given by , due to Corollary 1 since .
If , , , then the solution u with support on to backward distributional Abel equation is given by . Indeed so , so, by Corollary 2,
Then, such u solves the Abel equation since, by representation (14),
Lemma 8. Fix a value .
If a Laplace transformable function u fulfils on the half-line , then , for a suitable constant C.
If a function , with , fulfils on , then , for a suitable constant K.
If a function , with , fulfils on , then , for a suitable constant C.
Proof. The property
entails
is constant. Thus, for a suitable constant function
K, we have that
u fulfills the Abel integral equation:
, moreover
since
, and due to (
96) and the boundedness of
Then, by Proposition 2, the solution
u of the Abel equation
is
This proves the first and second claim, since an -transformable function is an function on every bounded interval. The third one follows in the same way, by applying Corollary 1 to the backward Abel equation . □
Lemma 8 provides the inverse of (
100). Hence, summarizing
Lemma 9. Assume that the interval is bounded, , , belongs to and .
and for every ; moreover, there is such that The same claims hold true when , and are replaced, respectively, by , and in the assumptions and the claims.
Proof. By considering
as the unknown in the Abel integral equation
we know by Proposition 2 that there is a solution
fulfilling the integral equation: such
is the unique solution in
and fulfills
Thus
. Moreover, by (
127),
and the semigroup property of
(Proposition 1),
Summing up
,
; then
,
and (
125), (
126) follow by standard embedding of fractional integrals. □
If we remove the assumption in Lemma 9, then we must add suitable corrections to both v and , as stated by the next theorem.
Theorem 12. Assume that bounded, , , belongs to .
and for every ; moreover, there is such that Explicitly, for every given , we have The same claims hold true when , and are replaced respectively by , and in the assumptions and the claims.
Proof. Since
belongs to
, it has a finite right value
at
, labeled by
, say
. By (
97),
,
. We set
then
,
and
. We know by Proposition 3 that there is a solution
with
fulfilling the integral equation
such
is the unique solution with support on
and fulfills
Thus
. By (
132),
and the semigroup property of
Hence, by setting
and taking into account (
99), we obtain
Thus , ; then , . The function belongs to for every , due to the boundedness of the interval. By standard embedding of fractional integrals, the function belongs to for every .
Summarizing,
for every
and (
129) and (
130) hold true. □
Remark 17. We emphasize that in Theorem 12 the fractional integrals and derivatives , , and are understood in the distributional sense provided by Definitions 1 and 3. Referring to Definition 10, with bounded, (131) reads as follows Moreover, in a bounded interval we have since ; whereas where , indeed by Lemma 8 the kernel of is made by functions of the kind , which all belong to and fulfill on