Inverses of integral transforms of RKHSs
Abstract.
The Fourier transform and its inverse are well-known to have complex conjugate integral kernels. S. Saitoh demonstrated that this relationship extends to the theory of integral transforms of Hilbert spaces of functions under certain conditions. In this paper, we derive a necessary and sufficient condition for the inverse of an integral transform of a Hilbert space of functions to be represented by a complex conjugate integral kernel. As an application, we present an alternative proof of Plancherel’s theorem using the theory of reproducing kernels.
Key words and phrases:
integral transform, inverse transform, RKHS2020 Mathematics Subject Classification:
Primary 46E22, Secondary 42A381. Introduction
The Fourier transform and its inverse transform for are given by
(1) | ||||
(2) |
Note that the integral kernel of the Fourier transform, , and the integral kernel of the inverse Fourier transform, , are complex conjugates of each other. Integral transforms in spaces, when abstracted within the framework of Hilbert spaces, become the integral transforms of Hilbert spaces, which will be discussed later. In this case, the image space of the integral transform becomes a reproducing kernel Hilbert space (RKHS), and it is well known that its reproducing kernel is expressed by the inner product (Theorem 2.1). In 1982, S. Saitoh showed in [Saitoh82] that in the integral transform of Hilbert spaces, under certain conditions, the integral kernel of the inverse transform becomes the complex conjugate of the integral kernel of the original transform, similar to the case of the Fourier transform described above (Theorem 2.2, Equation (4)).
This paper establishes a close relationship between RKHSs and the inverse transform formula (4) (Theorem 3.1). We derive a generalized inverse formula (6) as an extension of (4) and apply it to provide an alternative proof of Plancherel’s theorem. Throughout this paper, all operators are linear unless otherwise stated.
2. Integral transforms of Hilbert spaces
A Hilbert space of functions on a set is called an RKHS if the operator
is bounded for any and point . Then, by Riesz’s representation theorem, there exists a function such that . This function is called the reproducing kernel of for the point , and the two-variable function is called the reproducing kernel of . For more details on RKHS, refer to [Aronszajn50, PaulsenRaghupathi16]. For the theory and applications of integral transforms of Hilbert spaces, see [Saitoh97, SaitohSawano16] and the references therein.
Definition 2.1.
Let be a Hilbert space and let be a vector space. For a linear map with closed kernel, there exists a unique Hilbert space structure on of so that the linear map is a coisometry. In other words, we define an inner product on such that, for ,
This Hilbert space is called the operator range of the linear map (cf. [Sarason94])
Let be a map from a set to a Hilbert space . Define an operator by, for ,
Note that is closed in . The operator range of the map is a RKHS on , and the operator is called the integral transform of induced by the map . If the map is clear from the context, we write with and .
Theorem 2.1 ([Saitoh97]*pp. 20–23).
Under the above settings, the following hold:
-
(i)
For any , we have . Equality holds if and only if , where denotes the closed linear span of the set . In particular, is an isometry if and only if is complete in , i.e., the linear span of is dense in .
-
(ii)
The integral transform of is a RKHS on , and its reproducing kernel for the point is the function . Moreover, the reproducing kernel of is given by
(3)
The following is a prototype of the inversion theorem [Saitoh82]*Theorem 3.2 for the integral transforms of Hilbert spaces.
Theorem 2.2 ([Saitoh82]*Theorem 3.1).
Let be a map from a set to a Hilbert space of functions on a set , and let be the integral transform of induced by . Then, if all of the conditions
-
(i)
is complete in ,
-
(ii)
, for any ,
-
(iii)
, for any and ,
-
(iv)
, for any and ,
are satisfied, the inverse of the integral transform is given by
(4) |
Although the hypotheses are somewhat restrictive, Theorem2.2 can be regarded as a generalization of the inverse Fourier transform outlined in Introduction.
3. Main Results
The following is our main result concerning the inverse transform.
Theorem 3.1 (Main Theorem).
Let be a set, and let be a Hilbert space of functions defined on a set . Denote by the integral transform of induced by a map , and let for . Then, the following are equivalent:
-
(i)
For every we have , and the inverse transform formula (4) holds for all .
-
(ii)
The closed subspace of is a RKHS on .
In particular, if is an RKHS on and is complete in , then the inverse transform formula (4) holds for any .
Proof.
(i)(ii) By Theorem 2.1, the integral transform is a contraction. Thus, from the identity (4) and Schwarz’s inequality, for and , we have
Therefore, is a RKHS on .
(ii)(i) Let denote the reproducing kernel of the RKHS for the point . Since , for any , we have
Thus, . Moreover, since the integral transform is an isometric isomorphism on , for any , we have
Therefore, (4) holds. The final assertion follows because the completeness of in is equivalent to . ∎
To apply Theorem 3.1, it is necessary to know explicitly the inner product of the image of the integral transform. To this end, we prepare the following lemma, which provides a sufficient condition for the norm of an RKHS to be expressed in terms of the norm of another space.
Lemma 3.1.
Let be an RKHS on a set with reproducing kernel , and let be a subset of . Denote by the reproducing kernel for a point . Assume the following conditions:
-
(i)
If and , then , i.e., is a uniqueness set for .
-
(ii)
There exists a map to a Hilbert space such that, for all ,
(5)
Then, extends uniquely to an isometry from to , and is isometrically isomorphic to a closed subspace of . In particular, if and is defined by the restriction to , then the extended isometry coincides with the restriction of functions in to .
Proof.
The space of restrictions of functions in to is an RKHS on , with the restriction of to as its reproducing kernel. This RKHS structure is induced by the operator range of the restriction operator . Since is injective by (i), is an isometry by the definition of the operator range. Moreover, from (ii), for and , we have
This identity implies that if , then . Thus, the mapping
is a well-defined linear isometry extending the map . By (i), the set is complete in , which means that the set is dense in . Consequently, the composition extends uniquely to an isometry from to . Thus, is isometrically isomorphic to the closed subspace of . The last assertion can be seen from the fact that, since is an RKHS, strong convergence implies pointwise convergence, and in spaces mean convergence implies the existence of a subsequence that converges almost everywhere. ∎
The inverse transform formula (4) holds for the integral transform of an RKHS. Even for a general Hilbert space, a similar formula can be obtained if there exists an RKHS which is isometric to it. For convenience we make the following definition.
Definition 3.1.
Let be a Hilbert space and let be a map . If is an RKHS on a set , and if is an isometry, then the composite mapping is called a transformation sequence on .
Theorem 3.2 (Inverse Transform Theorem).
Let be a transformation sequence on . Let denote the integral transform of induced by . Then, for all and , the following identity holds:
(6) |
Proof.
Next we give a simple application of the inverse transform theorem.
Example 1 (cf. [SaitohSawano16]*pp. 116, 153).
Let be a separable Hilbert space, and let be a complete orthonormal system for . The Fourier coefficients of with respect to are defined as the sequence , where
Consider the integral transform of induced by the map . By Parseval’s identity, the map
is an isometry. Since is an RKHS on with the reproducing kernel (Kronecker delta), the sequence is a transformation sequence on . From the inverse transform formula (6), for all and , we have
(7) |
Expressing the inverse transform as a Fourier series, we have that for , the following holds in the sense of strong convergence in :
3.1. RKHSs isometric to One-Dimensional spaces
To apply the inverse transform theorem, we need to identify suitable RKHSs that are isometric to Hilbert spaces. Here, we focus on RKHSs isometric to spaces on intervals of , as these are among the most important Hilbert spaces.
Definition 3.2.
The notation represents the set of real numbers strictly between the two extended real numbers and in , i.e.,
The notations for half-open intervals , closed intervals , and so forth, are defined analogously.
Under this definition, note that for , the following identity holds:
where denotes the median of .
Definition 3.3.
Let , with , be an open interval in , and let . Consider the space of absolutely continuous complex-valued functions on the interval such that and , equipped with the inner product
which is called the first-order Sobolev-type space on and is denoted by . Here, the weight function is assumed to be positive and measurable on and satisfies the following condition:
(8) |
Remark 3.1.
In the above definition, when is an endpoint of , i.e., when or , special attention is needed regarding the value of . For any with , by the Cauchy-Schwarz inequality, we have
By assumption (8), the right-hand side becomes arbitrarily small as . Thus, by Cauchy’s convergence criterion, for any , the limit exists. This limit value is defined as . Furthermore, from (8), if , then must be in .
The following theorem is a slight generalization of the RKHSs discussed in [Saitoh84a] and [Saitoh88]*p. 75. We will show that the space is an RKHS isometric to .
Theorem 3.3.
is an RKHS on , and the following hold:
-
(i)
The reproducing kernel of is given by
-
(ii)
Let denote the characteristic function of a set . Then, is the integral transform of the Hilbert space via the mapping
that is, .
-
(iii)
The integral transform is an isometry. In particular, the family of functions is complete in .
Proof.
(i) Let . From (8), it is clear that . We verify the reproducing property of . We consider cases based on the relationship between and . If ,
(9) |
Hence, for ,
The case is handled similarly.
(ii) From assumption (8), the mapping is well-defined. The reproducing kernel of the integral transform is given by (3):
This coincides with the reproducing kernel of . Hence, .
(iii) For , the integral transform is, by definition,
Hence, a.e. , and
Therefore, the integral transform is an isometry, and by Theorem 2.1, the set is complete in . ∎
3.2. Transformation Sequences and Tensor Products
To handle the Fourier transform on using the inversion theorem, it is necessary to identify a manageable RKHS isometric to the -dimensional space. Here, we aim to construct the desired higher-dimensional RKHS through the tensor product of the one-dimensional RKHS introduced in the previous section.
The tensor product Hilbert space of Hilbert spaces () is the completion of the algebraic tensor product , equipped with an inner product satisfying
(10) |
for any ().
Additionally, the tensor product of bounded operators () between Hilbert spaces is the unique bounded operator satisfying
(11) |
for () (see, e.g., [ReedSimon72, PaulsenRaghupathi16]).
The following result regarding the tensor product of RKHSs is well known.
Proposition 3.1 ([PaulsenRaghupathi16]*p. 73).
Let be an RKHS on and an RKHS on . The tensor product Hilbert space is an RKHS on . If is the reproducing kernel of for and is the reproducing kernel of for , then the reproducing kernel of for is given by .
Next, we describe the relationship between tensor product Hilbert spaces and integral transforms. In what follows, the reproducing kernel of the integral transform for a point , defined by a map , is denoted by .
Proposition 3.2.
Let be Hilbert spaces, be sets, and be maps (). Then, the integral transform defined by the map such that satisfies:
-
(i)
For any (), .
-
(ii)
For , . In particular, as RKHSs on ,
-
(iii)
If is complete in (), then is also complete in .
Proof.
First, note that both the integral transform and the tensor product Hilbert space are RKHSs on . For and (), (i) follows from a straightforward calculation based on the definition. To prove (ii), using (3), the reproducing kernel of for is
Hence, . By Proposition 3.1, the left-hand side and right-hand side represent the reproducing kernels of the RKHS and the RKHS for , respectively. The equality of reproducing kernels implies that these RKHSs are identical. (iii) follows from the fact that the span of is dense in the algebraic tensor product , which is straightforward to verify. ∎
Transformation sequences are closed under tensor products.
Proposition 3.3.
If are transformation sequences on (), then
is a transformation sequence on .
4. Application
4.1. Paley-Wiener Space
As preparation for treating Fourier transforms using the theory of reproducing kernels, we consider the Paley-Wiener space. For , let . We define a mapping by
(12) |
The set is complete in , because the family is a complete orthogonal system in . Therefore, by Theorem 2.1, the integral transform is an isometry. The integral transform induced by is related to the Fourier transform on as follows: for ,
The image of the integral transform, denoted by , is an RKHS on , called the Paley-Wiener space of order . The reproducing kernel of is given by, for ,
(13) |
Since the integral kernel of is an entire function of , it is easy to see that differentiation and integration can be interchanged. This implies that elements of are entire functions. The uniqueness theorem for analytic functions implies that if an entire function vanishes on , then it is identically zero on . Thus, the real line is a uniqueness set for . The following lemma provides the equality (5) for the norm in , which is essential for our purposes.
Lemma 4.1.
For all ,
Proof.
Using partial fraction decomposition, we have
(14) | ||||
(15) | ||||
(16) |
Here, the two integrals on the right-hand side are improper integrals, and the first one is denoted by . To compute , we make the substitution :
(17) | ||||
(18) | ||||
(19) | ||||
(20) | ||||
(21) |
Substituting this into (16), we obtain the desired result. ∎
4.2. Plancherel’s Theorem
In this section, we provide an alternative proof of Plancherel’s theorem using integral transforms and the inversion theorem. This approach offers a different perspective compared to standard proofs that often rely on approximation by smooth functions. We use the notation from the previous section. For , let .
Theorem 4.1 (Plancherel’s Theorem).
For (), the Fourier transform and the conjugate Fourier transform are defined for by
(23) | ||||
(24) |
Then, the restrictions of and its conjugate to extend to unitary operators on , and they are inverses of each other.
Proof.
The integral transform is a Fourier transform. Since is an isometry, its extension is also an isometry. As , the operator can be identified with the Fourier transform on . By Proposition 3.3, the tensor product of isometries is an isometry. Therefore, is an isometry. The Fourier transform on is defined as the mean convergence of the Fourier transform on the Cartesian products as . From the above, it follows that the Fourier transform is an isometry. Similarly, the conjugate Fourier transform is also an isometry.
Next, we show that and are mutual inverses, i.e., , where denotes the identity. Let . Then is an RKHS on , and the indefinite integral operator is an isometry. Considering the transformation sequence on ,
and taking the -fold tensor product and applying Proposition 3.3, we obtain the following transformation sequence on :
Since the tensor product Hilbert space is isomorphic to , we have, for ,
On the other hand, by the definition of the tensor product of maps and Fubini’s theorem, for and , we have
Since the span of simple tensors is dense in , it follows that for all ,
i.e., is the -dimensional indefinite integral operator. Similarly, the norm of which is an RKHS on satisfies
for . Therefore, when and , applying the inverse transformation formula (6), we obtain, for , ,
(25) |
By Theorem 3.1, for fixed , we have . Since the Fourier transform is an isometry, () in . Thus, letting in (25), we obtain, for all and ,
(26) |
If, in addition, , then, since , we can apply Fubini’s theorem to interchange the order of integration in (26), obtaining
Furthermore, if we write and , then by Fubini’s theorem we have,
Differentiating both sides with respect to at , we obtain, for almost every ,
Repeating this differentiation times with respect to at , we conclude that for almost every . Now, if , then by integration by parts we obtain the identity
Since the left-hand side is bounded and since , it is easy to see that . Since is dense in , we conclude that the set of with is dense in . Therefore, by the boundedness of and , the desired identity holds by continuity. Similarly, it can be shown that . Thus, Plancherel’s theorem is demonstrated. ∎