This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Inverses of integral transforms of RKHSs

Akira Yamada Tokyo Gakugei University [email protected]
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, RKHS
2020 Mathematics Subject Classification:
Primary 46E22, Secondary 42A38

1. Introduction

The Fourier transform \mathcal{F} and its inverse transform \mathcal{F}^{*} for fL1()f\in L^{1}(\mathbb{R}) are given by

(f)(x)\displaystyle\mathcal{F}(f)(x) =12πf(t)eitx𝑑t,\displaystyle=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}f(t)e^{-itx}\,dt, (1)
(f)(t)\displaystyle\mathcal{F}^{*}(f)(t) =12πf(x)eitx𝑑x,\displaystyle=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}f(x)e^{itx}\,dx, (2)

Note that the integral kernel of the Fourier transform, eitxe^{-itx}, and the integral kernel of the inverse Fourier transform, eitxe^{itx}, are complex conjugates of each other. Integral transforms in LpL^{p} 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 \mathcal{H} of functions on a set EE is called an RKHS if the operator

ff(x)f\in\mathcal{H}\mapsto f(x)\in\mathbb{C}

is bounded for any ff\in\mathcal{H} and point xEx\in E. Then, by Riesz’s representation theorem, there exists a function kxk_{x}\in\mathcal{H} such that f(x)=f,kxf(x)=\langle f,k_{x}\rangle. This function kxk_{x} is called the reproducing kernel of \mathcal{H} for the point xx, and the two-variable function k(x,y)=ky,kxk(x,y)=\langle k_{y},k_{x}\rangle is called the reproducing kernel of \mathcal{H}. 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 \mathcal{H} be a Hilbert space and let VV be a vector space. For a linear map A:VA\colon\mathcal{H}\to V with closed kernel, there exists a unique Hilbert space structure on ranA=A()\operatorname{ran}A=A(\mathcal{H}) of VV so that the linear map A:ranAA\colon\mathcal{H}\to\operatorname{ran}A is a coisometry. In other words, we define an inner product on ranA\operatorname{ran}A such that, for f,g(kerA)f,g\in(\ker A)^{\perp},

Af,AgranA=f,g.\langle Af,Ag\rangle_{\operatorname{ran}A}=\langle f,g\rangle_{\mathcal{H}}.

This Hilbert space is called the operator range of the linear map AA (cf. [Sarason94])

Let ϕ:E\phi\colon E\to\mathcal{H} be a map from a set EE to a Hilbert space \mathcal{H}. Define an operator ϕ^:E\hat{\phi}\colon\mathcal{H}\to\mathbb{C}^{E} by, for ff\in\mathcal{H},

(ϕ^f)(x)=f,ϕ(x),xE.(\hat{\phi}f)(x)=\langle f,\phi(x)\rangle_{\mathcal{H}},\quad x\in E.

Note that kerϕ^=ϕ(E)\ker\hat{\phi}=\phi(E)^{\perp} is closed in \mathcal{H}. The operator range ϕ^()\hat{\phi}(\mathcal{H}) of the map ϕ^\hat{\phi} is a RKHS on EE, and the operator ϕ^:ϕ^()\hat{\phi}\colon\mathcal{H}\to\hat{\phi}(\mathcal{H}) is called the integral transform of \mathcal{H} induced by the map ϕ\phi. If the map ϕ\phi is clear from the context, we write with f^=ϕ^f\hat{f}=\hat{\phi}f and ^=ϕ^()\hat{\mathcal{H}}=\hat{\phi}(\mathcal{H}).

Theorem 2.1 ([Saitoh97]*pp. 20–23).

Under the above settings, the following hold:

  1. (i)

    For any ff\in\mathcal{H}, we have f^^f\|\hat{f}\|_{\hat{\mathcal{H}}}\leq\|f\|_{\mathcal{H}}. Equality holds if and only if f(kerϕ^)=span¯ϕ(E)f\in(\ker\hat{\phi})^{\perp}=\operatorname{\overline{span}}\phi(E), where span¯ϕ(E)\operatorname{\overline{span}}\phi(E) denotes the closed linear span of the set ϕ(E)\phi(E). In particular, ϕ^:^\hat{\phi}\colon\mathcal{H}\to\hat{\mathcal{H}} is an isometry if and only if ϕ(E)\phi(E) is complete in \mathcal{H}, i.e., the linear span of ϕ(E)\phi(E) is dense in \mathcal{H}.

  2. (ii)

    The integral transform ^\hat{\mathcal{H}} of \mathcal{H} is a RKHS on EE, and its reproducing kernel for the point yEy\in E is the function ϕ(y)^\widehat{\phi(y)}. Moreover, the reproducing kernel of ^\hat{\mathcal{H}} is given by

    ϕ(y),ϕ(x).\langle\phi(y),\phi(x)\rangle_{\mathcal{H}}. (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 ϕ:E\phi\colon E\to\mathcal{H} be a map from a set EE to a Hilbert space \mathcal{H} of functions on a set TT, and let ^\hat{\mathcal{H}} be the integral transform of \mathcal{H} induced by ϕ\phi. Then, if all of the conditions

  1. (i)

    ϕ(E)\phi(E)is complete in \mathcal{H},

  2. (ii)

    ϕ(t,)¯^\overline{\phi(t,\cdot)}\in\hat{\mathcal{H}}, for any tTt\in T,

  3. (iii)

    f^,ϕ(t,)¯^\langle\hat{f},\overline{\phi(t,\cdot)}\rangle_{\hat{\mathcal{H}}}\in\mathcal{H}, for any ff\in\mathcal{H} and tTt\in T,

  4. (iv)

    f^(x),ϕ(y),ϕ(x)^=f^(x),ϕ(t,x)¯^,ϕ(t,y)\langle\hat{f}(x),\langle\phi(y),\phi(x)\rangle_{\mathcal{H}}\rangle_{\hat{\mathcal{H}}}=\langle\langle\hat{f}(x),\overline{\phi(t,x)}\rangle_{\hat{\mathcal{H}}},\phi(t,y)\rangle_{\mathcal{H}}, for any ff\in\mathcal{H} and x,yEx,y\in E,

are satisfied, the inverse of the integral transform ϕ^\hat{\phi} is given by

f(t)=f^,ϕ(t,)¯^,tT.f(t)=\langle\hat{f},\overline{\phi(t,\cdot)}\rangle_{\hat{\mathcal{H}}},\quad t\in T. (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 EE be a set, and let \mathcal{H} be a Hilbert space of functions defined on a set TT. Denote by ^\hat{\mathcal{H}} the integral transform of \mathcal{H} induced by a map ϕ:E\phi\colon E\to\mathcal{H}, and let ϕ(t,x)=ϕ(x)(t)\phi(t,x)=\phi(x)(t) for (t,x)T×E(t,x)\in T\times E. Then, the following are equivalent:

  1. (i)

    For every tTt\in T we have ϕ(t,)¯^\overline{\phi(t,\cdot)}\in\hat{\mathcal{H}}, and the inverse transform formula (4) holds for all fspan¯ϕ(E)f\in\operatorname{\overline{span}}\phi(E).

  2. (ii)

    The closed subspace span¯ϕ(E)\operatorname{\overline{span}}\phi(E) of \mathcal{H} is a RKHS on TT.

In particular, if \mathcal{H} is an RKHS on TT and ϕ(E)\phi(E) is complete in \mathcal{H}, then the inverse transform formula (4) holds for any ff\in\mathcal{H}.

Proof.

(i)\implies(ii) By Theorem 2.1, the integral transform is a contraction. Thus, from the identity (4) and Schwarz’s inequality, for fspan¯ϕ(E)f\in\operatorname{\overline{span}}\phi(E) and tTt\in T, we have

|f(t)|ϕ(t,)¯^f^^ϕ(t,)¯^f.|f(t)|\leq\|\overline{\phi(t,\cdot)}\|_{\hat{\mathcal{H}}}\|\hat{f}\|_{\hat{\mathcal{H}}}\leq\|\overline{\phi(t,\cdot)}\|_{\hat{\mathcal{H}}}\|f\|_{\mathcal{H}}.

Therefore, span¯ϕ(E)\operatorname{\overline{span}}\phi(E) is a RKHS on TT.

(ii)\implies(i) Let ktk_{t} denote the reproducing kernel of the RKHS span¯ϕ(E)\operatorname{\overline{span}}\phi(E) for the point tTt\in T. Since ktk_{t}\in\mathcal{H}, for any xEx\in E, we have

ϕ(t,x)¯=ϕ(x),kt¯=kt,ϕ(x)=k^t(x).\overline{\phi(t,x)}=\overline{\langle\phi(x),k_{t}\rangle}_{\mathcal{H}}=\langle k_{t},\phi(x)\rangle_{\mathcal{H}}=\hat{k}_{t}(x).

Thus, ϕ(t,)¯=k^t^\overline{\phi(t,\cdot)}=\hat{k}_{t}\in\hat{\mathcal{H}}. Moreover, since the integral transform ϕ^\hat{\phi} is an isometric isomorphism on span¯ϕ(E)=(kerϕ^)\operatorname{\overline{span}}\phi(E)=(\ker\hat{\phi})^{\perp}, for any fspan¯ϕ(E)f\in\operatorname{\overline{span}}\phi(E), we have

f(t)=f,kt=f^,k^t^=f^,ϕ(t,)¯^.f(t)=\langle f,k_{t}\rangle_{\mathcal{H}}=\langle\hat{f},\hat{k}_{t}\rangle_{\hat{\mathcal{H}}}=\langle\hat{f},\overline{\phi(t,\cdot)}\rangle_{\hat{\mathcal{H}}}.

Therefore, (4) holds. The final assertion follows because the completeness of ϕ(E)\phi(E) in \mathcal{H} is equivalent to span¯ϕ(E)=\operatorname{\overline{span}}\phi(E)=\mathcal{H}. ∎

To apply Theorem 3.1, it is necessary to know explicitly the inner product of the image ^\hat{\mathcal{H}} 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 HH be an RKHS on a set EE with reproducing kernel kk, and let FF be a subset of EE. Denote by ky=k(,y)k_{y}=k(\cdot,y) the reproducing kernel for a point yEy\in E. Assume the following conditions:

  1. (i)

    If fHf\in H and f|F=0f|_{F}=0, then f=0f=0, i.e., FF is a uniqueness set for HH.

  2. (ii)

    There exists a map T:{kx}xFKT\colon\{k_{x}\}_{x\in F}\to K to a Hilbert space KK such that, for all x,yFx,y\in F,

    k(x,y)=Tky,TkxK.k(x,y)=\langle Tk_{y},Tk_{x}\rangle_{K}. (5)

Then, TT extends uniquely to an isometry T~\tilde{T} from HH to KK, and HH is isometrically isomorphic to a closed subspace of KK. In particular, if K=L2(F,dμ)K=L^{2}(F,d\mu) and TT is defined by the restriction to FF, then the extended isometry T~\tilde{T} coincides with the restriction of functions in HH to FF.

Proof.

The space H|FH|_{F} of restrictions of functions in HH to FF is an RKHS on FF, with the restriction of kk to F×FF\times F as its reproducing kernel. This RKHS structure is induced by the operator range of the restriction operator ρ:HH|F\rho\colon H\to H|_{F}. Since ρ\rho is injective by (i), ρ:HH|F\rho\colon H\to H|_{F} is an isometry by the definition of the operator range. Moreover, from (ii), for cic_{i}\in\mathbb{C} and xiFx_{i}\in F, we have

icikxi|FH|F2=i,jcic¯jk(xj,xi)=i,jcic¯jTkxi,TkxjK=iciTkxiK2.\Bigl{\|}\sum_{i}c_{i}k_{x_{i}}|_{F}\Bigr{\|}_{H|_{F}}^{2}=\sum_{i,j}c_{i}\overline{c}_{j}k(x_{j},x_{i})=\sum_{i,j}c_{i}\overline{c}_{j}\langle Tk_{x_{i}},Tk_{x_{j}}\rangle_{K}=\Bigl{\|}\sum_{i}c_{i}Tk_{x_{i}}\Bigr{\|}_{K}^{2}.

This identity implies that if icikxi|F=0\sum_{i}c_{i}k_{x_{i}}|F=0, then iciTkxi=0\sum_{i}c_{i}Tk{x_{i}}=0. Thus, the mapping

T~:icikxi|FH|FiciTkxiK\tilde{T}\colon\sum_{i}c_{i}k_{x_{i}}|_{F}\in H|_{F}\mapsto\sum_{i}c_{i}Tk_{x_{i}}\in K

is a well-defined linear isometry extending the map TT. By (i), the set {kx}xF\{k_{x}\}_{x\in F} is complete in HH, which means that the set span{kx}xF\operatorname{span\,}\{k_{x}\}_{x\in F} is dense in HH. Consequently, the composition T~ρ:span{kx}xFK\tilde{T}\rho\colon\operatorname{span\,}\{k_{x}\}_{x\in F}\to K extends uniquely to an isometry from HH to KK. Thus, HH is isometrically isomorphic to the closed subspace of KK. The last assertion can be seen from the fact that, since HH is an RKHS, strong convergence implies pointwise convergence, and in LpL^{p} 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 \mathcal{H} be a Hilbert space and let ϕ\phi be a map ϕ:E\phi\colon E\to\mathcal{H}. If 𝒲\mathcal{W} is an RKHS on a set TT, and if S:𝒲S\colon\mathcal{H}\to\mathcal{W} is an isometry, then the composite mapping Eϕ𝑆𝒲E\overset{\phi}{\to}\mathcal{H}\overset{S}{\to}\mathcal{W} is called a transformation sequence on TT.

Theorem 3.2 (Inverse Transform Theorem).

Let Eϕ𝑆𝒲E\overset{\phi}{\to}\mathcal{H}\overset{S}{\to}\mathcal{W} be a transformation sequence on TT. Let ^\hat{\mathcal{H}} denote the integral transform of \mathcal{H} induced by ϕ\phi. Then, for all fspan¯ϕ(E)f\in\operatorname{\overline{span}}\phi(E) and tTt\in T, the following identity holds:

(Sf)(t)=f^,(Sϕ)(t,)¯^.(Sf)(t)=\langle\hat{f},\overline{(S\phi)(t,\cdot)}\rangle_{\hat{\mathcal{H}}}. (6)
Proof.

Both ^=ϕ^()\hat{\mathcal{H}}=\hat{\phi}(\mathcal{H}) and Sϕ^(𝒲)\widehat{S\phi}(\mathcal{W}) are RKHSs on EE, and their reproducing kernels coincide since SS is an isometry. Thus, ^=Sϕ^(𝒲)\hat{\mathcal{H}}=\widehat{S\phi}(\mathcal{W}). Since an isometry is a closed map, so S(span¯ϕ(E))=span¯S(ϕ(E))S(\operatorname{\overline{span}}\phi(E))=\operatorname{\overline{span}}S(\phi(E)). By applying Theorem 3.1 to the integral transform induced by Sϕ:E𝒲S\phi\colon E\to\mathcal{W}, for all fspan¯ϕ(E)f\in\operatorname{\overline{span}}\phi(E), we have

(Sf)(t)=Sϕ^(Sf),(Sϕ)(t,)¯^.(Sf)(t)=\langle\widehat{S\phi}(Sf),\overline{(S\phi)(t,\cdot)}\rangle_{\hat{\mathcal{H}}}.

Since SS is an isometry, Sϕ^(Sf)=f^\widehat{S\phi}(Sf)=\hat{f}. This proves the identity (6). ∎

Next we give a simple application of the inverse transform theorem.

Example 1 (cf. [SaitohSawano16]*pp. 116, 153).

Let \mathcal{H} be a separable Hilbert space, and let {gn}n=1\{g_{n}\}_{n=1}^{\infty} be a complete orthonormal system for \mathcal{H}. The Fourier coefficients of ff\in\mathcal{H} with respect to {gn}\{g_{n}\} are defined as the sequence (f^n)(\hat{f}_{n})\in\mathbb{C}^{\mathbb{N}}, where

f^n=f,gn.\hat{f}_{n}=\langle f,g_{n}\rangle_{\mathcal{H}}.

Consider the integral transform ^\hat{\mathcal{H}} of \mathcal{H} induced by the map ψ:E\psi\colon E\to\mathcal{H}. By Parseval’s identity, the map

S:f(f^n)2S\colon f\in\mathcal{H}\mapsto(\hat{f}_{n})\in\ell^{2}

is an isometry. Since 2\ell^{2} is an RKHS on \mathbb{N} with the reproducing kernel δij\delta_{ij} (Kronecker delta), the sequence 𝜓𝑆2\mathbb{N}\overset{\psi}{\to}\mathcal{H}\overset{S}{\to}\ell^{2} is a transformation sequence on \mathbb{N}. From the inverse transform formula (6), for all fspan¯ψ(E)f\in\operatorname{\overline{span}}\psi(E) and nn\in\mathbb{N}, we have

f^n\displaystyle\hat{f}_{n} =f^,ψ()^n¯^=f^(x),ψ(x),gn¯^.\displaystyle=\langle\hat{f},\overline{\widehat{\psi(\cdot)}n}\rangle_{\hat{\mathcal{H}}}=\langle\hat{f}(x),\overline{\langle\psi(x),g_{n}\rangle}\mathcal{H}\rangle_{\hat{\mathcal{H}}}. (7)

Expressing the inverse transform as a Fourier series, we have that for fspan¯ψ(E)f\in\operatorname{\overline{span}}\psi(E), the following holds in the sense of strong convergence in \mathcal{H}:

f=n=0f^,ψ()^n¯^gn.f=\sum_{n=0}^{\infty}\langle\hat{f},\overline{\widehat{\psi(\cdot)}_{n}}\rangle_{\hat{\mathcal{H}}}g_{n}.

3.1. RKHSs isometric to One-Dimensional L2L^{2} 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 L2L^{2} spaces on intervals of \mathbb{R}, as these are among the most important Hilbert spaces.

Definition 3.2.

The notation (a,b)(a,b) represents the set of real numbers strictly between the two extended real numbers aa and bb in ¯={±}\overline{\mathbb{R}}=\mathbb{R}\cup\{\pm\infty\}, i.e.,

(a,b)={{x:a<x<b},(a<b),,(a=b),{x:b<x<a},(b<a).(a,b)=\begin{cases}\{x\in\mathbb{R}\colon a<x<b\},&(a<b),\\ \emptyset,&(a=b),\\ \{x\in\mathbb{R}\colon b<x<a\},&(b<a).\end{cases}

The notations for half-open intervals [a,b)[a,b), closed intervals [a,b][a,b], and so forth, are defined analogously.

Under this definition, note that for c,x,y¯c,x,y\in\overline{\mathbb{R}}, the following identity holds:

(c,x)(c,y)=(c,med{x,y,c}),(c,x)\cap(c,y)=(c,\operatorname{med}\{x,y,c\}),

where med{x,y,z}\operatorname{med}\{x,y,z\} denotes the median of x,y,z¯x,y,z\in\overline{\mathbb{R}}.

Definition 3.3.

Let I=(a,b)I=(a,b), with a<b-\infty\leq a<b\leq\infty, be an open interval in ¯\overline{\mathbb{R}}, and let c[a,b]c\in[a,b]. Consider the space of absolutely continuous complex-valued functions fAC(I)f\in AC(I) on the interval II such that f(c)=0f(c)=0 and fLρ2(I)f^{\prime}\in L_{\rho}^{2}(I), equipped with the inner product

f,g=abf(t)g(t)¯ρ(t)𝑑t,\langle f,g\rangle=\int_{a}^{b}f^{\prime}(t)\overline{g^{\prime}(t)}\,\rho(t)\,dt,

which is called the first-order Sobolev-type space on II and is denoted by Hc,ρ(I)H_{c,\rho}(I). Here, the weight function ρ\rho is assumed to be positive and measurable on II and satisfies the following condition:

χ(c,x)ρL1((c,x)),xI.\frac{\chi_{(c,x)}}{\rho}\in L^{1}((c,x)),\quad\forall x\in I. (8)
Remark 3.1.

In the above definition, when cc is an endpoint of II, i.e., when c=ac=a or c=bc=b, special attention is needed regarding the value of f(c)f(c). For any x,yIx,y\in I with x<yx<y, by the Cauchy-Schwarz inequality, we have

|f(x)f(y)|2=|xyf(t)𝑑t|2xy|f|2ρ𝑑txydtρf2xydtρ.|f(x)-f(y)|^{2}=\Bigl{|}\int_{x}^{y}f^{\prime}(t)\,dt\Bigr{|}^{2}\leq\int_{x}^{y}|f^{\prime}|^{2}\rho\,dt\int_{x}^{y}\frac{dt}{\rho}\leq\|f\|^{2}\int_{x}^{y}\frac{dt}{\rho}.

By assumption (8), the right-hand side becomes arbitrarily small as x,ycx,y\to c. Thus, by Cauchy’s convergence criterion, for any fHc,ρ(I)f\in H_{c,\rho}(I), the limit limxcf(x)\lim\limits_{x\to c}f(x) exists. This limit value is defined as f(c)f(c). Furthermore, from (8), if ρ1\rho\equiv 1, then cc must be in \mathbb{R}.

The following theorem is a slight generalization of the RKHSs discussed in [Saitoh84a] and [Saitoh88]*p. 75. We will show that the space Hc,ρ(I)H_{c,\rho}(I) is an RKHS isometric to Lρ2(I)L_{\rho}^{2}(I).

Theorem 3.3.

Hc,ρ(I)H_{c,\rho}(I) is an RKHS on I{c}I\cup\{c\}, and the following hold:

  1. (i)

    The reproducing kernel of Hc,ρ(I)H_{c,\rho}(I) is given by

    k(x,y)=(c,med{x,y,c})dtρ(t)=|cmed{x,y,c}dtρ(t)|.k(x,y)=\int_{(c,\operatorname{med}\{x,y,c\})}\frac{dt}{\rho(t)}=\Bigl{|}\int_{c}^{\operatorname{med}\{x,y,c\}}\frac{dt}{\rho(t)}\Bigr{|}.
  2. (ii)

    Let χE\chi_{E} denote the characteristic function of a set EE. Then, Hc,ρ(I)H_{c,\rho}(I) is the integral transform of the Hilbert space Lρ2(I)L_{\rho}^{2}(I) via the mapping

    ϕ:xIχ(c,x)ρLρ2(I),\phi\colon x\in I\mapsto\frac{\chi_{(c,x)}}{\rho}\in L_{\rho}^{2}(I),

    that is, Hc,ρ(I)=ϕ^(Lρ2(I))H_{c,\rho}(I)=\hat{\phi}(L_{\rho}^{2}(I)).

  3. (iii)

    The integral transform ϕ^:Lρ2(I)Hc,ρ(I)\hat{\phi}\colon L_{\rho}^{2}(I)\to H_{c,\rho}(I) is an isometry. In particular, the family of functions ϕ(I)\phi(I) is complete in Lρ2(I)L_{\rho}^{2}(I).

Proof.

(i) Let ky=k(,y)k_{y}=k(\cdot,y). From (8), it is clear that kyHc,ρ(I)k_{y}\in H_{c,\rho}(I). We verify the reproducing property of kyk_{y}. We consider cases based on the relationship between yy and cc. If ycy\leq c,

k(x,y)\displaystyle k(x,y) ={xycρ1𝑑t,(xc),0,(cx).\displaystyle=\begin{cases}\int_{x\vee y}^{c}\rho^{-1}dt,&(x\leq c),\\ 0,&(c\leq x).\end{cases} (9)

Hence, for fHc,ρ(I)f\in H_{c,\rho}(I),

f,ky=Ifky¯ρ𝑑t=ycf(ρ1)ρ𝑑t=cyf(t)𝑑t=f(y).\langle f,k_{y}\rangle=\int_{I}f^{\prime}\overline{k_{y}^{\prime}}\rho\,dt=\int_{y}^{c}f^{\prime}\cdot(-\rho^{-1})\rho\,dt=\int_{c}^{y}f^{\prime}(t)\,dt=f(y).

The case cyc\leq y is handled similarly.

(ii) From assumption (8), the mapping ϕ\phi is well-defined. The reproducing kernel of the integral transform ϕ^(Lρ2(I))\hat{\phi}(L_{\rho}^{2}(I)) is given by (3):

ϕ(y),ϕ(x)Lρ2(I)\displaystyle\langle\phi(y),\phi(x)\rangle_{L_{\rho}^{2}(I)} =Iχ(c,x)(c,y)ρ𝑑t=(c,med{x,y,c})dtρ.\displaystyle=\int_{I}\frac{\chi_{(c,x)\cap(c,y)}}{\rho}\,dt=\int_{(c,\operatorname{med}\{x,y,c\})}\frac{dt}{\rho}.

This coincides with the reproducing kernel of Hc,ρ(I)H_{c,\rho}(I). Hence, ϕ^(Lρ2(I))=Hc,ρ(I)\hat{\phi}(L_{\rho}^{2}(I))=H_{c,\rho}(I).

(iii) For fLρ2(I)f\in L_{\rho}^{2}(I), the integral transform f^=ϕ^(f)\hat{f}=\hat{\phi}(f) is, by definition,

f^(x)=Ifχ(c,x)ρρ𝑑t={cxf𝑑t,(xc),cxf𝑑t,(xc).\hat{f}(x)=\int_{I}f\frac{\chi_{(c,x)}}{\rho}\rho\,dt=\begin{cases}-\int_{c}^{x}f\,dt,&(x\leq c),\\ \int_{c}^{x}f\,dt,&(x\geq c).\end{cases}

Hence, |f^(x)|=|f(x)||\hat{f}^{\prime}(x)|=|f(x)| a.e. xIx\in I, and

f^Hc,ρ(I)2=I|f^|2ρ𝑑t=I|f|2ρ𝑑t=fLρ2(I)2.\|\hat{f}\|_{H_{c,\rho}(I)}^{2}=\int_{I}|\hat{f}^{\prime}|^{2}\rho\,dt=\int_{I}|f|^{2}\rho\,dt=\|f\|_{L_{\rho}^{2}(I)}^{2}.

Therefore, the integral transform ϕ^\hat{\phi} is an isometry, and by Theorem 2.1, the set ϕ(I)\phi(I) is complete in Lρ2(I)L_{\rho}^{2}(I). ∎

3.2. Transformation Sequences and Tensor Products

To handle the Fourier transform on N\mathbb{R}^{N} using the inversion theorem, it is necessary to identify a manageable RKHS isometric to the NN-dimensional L2L^{2} 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 i=1ni\otimes_{i=1}^{n}\mathcal{H}_{i} of Hilbert spaces i\mathcal{H}_{i} (i=1,,ni=1,\dots,n) is the completion of the algebraic tensor product i=1ni\odot_{i=1}^{n}\mathcal{H}_{i}, equipped with an inner product satisfying

i=1nai,i=1nbii=1ni=i=1nai,bii\langle\otimes_{i=1}^{n}a_{i},\otimes_{i=1}^{n}b_{i}\rangle_{\otimes_{i=1}^{n}\mathcal{H}_{i}}=\prod_{i=1}^{n}\langle a_{i},b_{i}\rangle_{\mathcal{H}_{i}} (10)

for any ai,biia_{i},b_{i}\in\mathcal{H}_{i} (i=1,,ni=1,\dots,n).

Additionally, the tensor product i=1nSi:i=1nii=1n𝒦i\otimes_{i=1}^{n}S_{i}\colon\otimes_{i=1}^{n}\mathcal{H}_{i}\to\otimes_{i=1}^{n}\mathcal{K}_{i} of bounded operators Si:i𝒦iS_{i}\colon\mathcal{H}_{i}\to\mathcal{K}_{i} (i=1,,ni=1,\dots,n) between Hilbert spaces is the unique bounded operator satisfying

(i=1nSi)(i=1nai)=i=1nSiai,(\otimes_{i=1}^{n}S_{i})(\otimes_{i=1}^{n}a_{i})=\otimes_{i=1}^{n}S_{i}a_{i}, (11)

for aiia_{i}\in\mathcal{H}_{i} (i=1,,ni=1,\dots,n) (see, e.g., [ReedSimon72, PaulsenRaghupathi16]).

The following result regarding the tensor product of RKHSs is well known.

Proposition 3.1 ([PaulsenRaghupathi16]*p. 73).

Let XX be an RKHS on EE and YY an RKHS on FF. The tensor product Hilbert space XYX\otimes Y is an RKHS on E×FE\times F. If kx1k_{x}^{1} is the reproducing kernel of XX for xEx\in E and ky2k_{y}^{2} is the reproducing kernel of YY for yFy\in F, then the reproducing kernel of XYX\otimes Y for (x,y)E×F(x,y)\in E\times F is given by kx1ky2k_{x}^{1}\otimes k_{y}^{2}.

Next, we describe the relationship between tensor product Hilbert spaces and integral transforms. In what follows, the reproducing kernel of the integral transform ϕ^()\hat{\phi}(\mathcal{H}) for a point xEx\in E, defined by a map ϕ:E\phi\colon E\to\mathcal{H}, is denoted by k[ϕ]xk[\phi]_{x}.

Proposition 3.2.

Let i\mathcal{H}_{i} be Hilbert spaces, EiE_{i} be sets, and ϕi:Eii\phi_{i}\colon E_{i}\to\mathcal{H}_{i} be maps (i=1,,ni=1,\dots,n). Then, the integral transform defined by the map i=1nϕi:i=1nEii=1ni\otimes_{i=1}^{n}\phi_{i}\colon\prod_{i=1}^{n}E_{i}\to\otimes_{i=1}^{n}\mathcal{H}_{i} such that (i=1nϕi)(x1,,xn)=i=1nϕi(xi)(\otimes_{i=1}^{n}\phi_{i})(x_{1},\dots,x_{n})=\otimes_{i=1}^{n}\phi_{i}(x_{i}) satisfies:

  1. (i)

    For any fiif_{i}\in\mathcal{H}_{i} (i=1,,ni=1,\dots,n), i=1nϕi^(i=1nfi)=i=1nϕ^ifi\widehat{\otimes_{i=1}^{n}\phi_{i}}(\otimes_{i=1}^{n}f_{i})=\otimes_{i=1}^{n}\hat{\phi}_{i}f_{i}.

  2. (ii)

    For x=(xi)i=1nEix=(x_{i})\in\prod_{i=1}^{n}E_{i}, k[i=1nϕi]x=i=1nk[ϕi]xik[\otimes_{i=1}^{n}\phi_{i}]_{x}=\otimes_{i=1}^{n}k[\phi_{i}]_{x_{i}}. In particular, as RKHSs on i=1nEi\prod_{i=1}^{n}E_{i},

    i=1nϕi^(i=1ni)=i=1nϕ^i(i).\widehat{\otimes_{i=1}^{n}\phi_{i}}\Bigl{(}\otimes_{i=1}^{n}\mathcal{H}_{i}\Bigr{)}=\otimes_{i=1}^{n}\hat{\phi}_{i}(\mathcal{H}_{i}).
  3. (iii)

    If ϕi(Ei)\phi_{i}(E_{i}) is complete in i\mathcal{H}_{i} (i=1,,ni=1,\dots,n), then (i=1nϕi)(i=1nEi)(\otimes_{i=1}^{n}\phi_{i})(\prod_{i=1}^{n}E_{i}) is also complete in i=1ni\otimes_{i=1}^{n}\mathcal{H}_{i}.

Proof.

First, note that both the integral transform i=1nϕi^(i=1ni)\widehat{\otimes_{i=1}^{n}\phi_{i}}(\otimes_{i=1}^{n}\mathcal{H}_{i}) and the tensor product Hilbert space i=1nϕ^ii\otimes_{i=1}^{n}\hat{\phi}_{i}\mathcal{H}_{i} are RKHSs on i=1nEi\prod_{i=1}^{n}E_{i}. For xiEix_{i}\in E_{i} and fiif_{i}\in\mathcal{H}_{i} (i=1,,ni=1,\dots,n), (i) follows from a straightforward calculation based on the definition. To prove (ii), using (3), the reproducing kernel of i=1ni\otimes_{i=1}^{n}\mathcal{H}_{i} for y=(yi)y=(y_{i}) is

k[i=1nϕi]x(y)\displaystyle k[\otimes_{i=1}^{n}\phi_{i}]_{x}(y) =i=1nϕi(y),i=1nϕi(x)i=1ni=i=1nϕi(yi),ϕi(xi)i\displaystyle=\langle\otimes_{i=1}^{n}\phi_{i}(y),\otimes_{i=1}^{n}\phi_{i}(x)\rangle_{\otimes_{i=1}^{n}\mathcal{H}_{i}}=\prod_{i=1}^{n}\langle\phi_{i}(y_{i}),\phi_{i}(x_{i})\rangle_{\mathcal{H}_{i}}
=i=1nk[ϕi]xi(yi)=(i=1nk[ϕi]xi)(y).\displaystyle=\prod_{i=1}^{n}k[\phi_{i}]_{x_{i}}(y_{i})=(\otimes_{i=1}^{n}k[\phi_{i}]_{x_{i}})(y).

Hence, k[i=1nϕi]x=i=1nk[ϕi]xik[\otimes_{i=1}^{n}\phi_{i}]_{x}=\otimes_{i=1}^{n}k[\phi_{i}]_{x_{i}}. By Proposition 3.1, the left-hand side and right-hand side represent the reproducing kernels of the RKHS i=1nϕi^(i=1ni)\widehat{\otimes_{i=1}^{n}\phi_{i}}(\otimes_{i=1}^{n}\mathcal{H}_{i}) and the RKHS i=1nϕ^i(i)\otimes_{i=1}^{n}\hat{\phi}_{i}(\mathcal{H}_{i}) for xx, respectively. The equality of reproducing kernels implies that these RKHSs are identical. (iii) follows from the fact that the span of (i=1nϕi)(i=1nEi)=i=1n(ϕi(Ei))(\otimes_{i=1}^{n}\phi_{i})(\prod_{i=1}^{n}E_{i})=\otimes_{i=1}^{n}(\phi_{i}(E_{i})) is dense in the algebraic tensor product i=1ni\odot_{i=1}^{n}\mathcal{H}_{i}, which is straightforward to verify. ∎

Transformation sequences are closed under tensor products.

Proposition 3.3.

If EϕiiSi𝒲iE\overset{\phi_{i}}{\to}\mathcal{H}_{i}\overset{S_{i}}{\to}\mathcal{W}_{i} are transformation sequences on TiT_{i} (i=1,,Ni=1,\dots,N), then

i=1NEi{\prod_{i=1}^{N}E_{i}}i=1Ni{\otimes_{i=1}^{N}\mathcal{H}_{i}}i=1N𝒲i{\otimes_{i=1}^{N}\mathcal{W}_{i}}i=1Nϕi\scriptstyle{\otimes_{i=1}^{N}\phi_{i}}i=1NSi\scriptstyle{\otimes_{i=1}^{N}S_{i}}

is a transformation sequence on i=1NTi\prod_{i=1}^{N}T_{i}.

Proof.

It suffices to show that the operator i=1NSi\otimes_{i=1}^{N}S_{i} is an isometry. This can be demonstrated by showing isometry for simple tensors, which follows directly from (10) and the definition (11). ∎

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 a>0a>0, let Ia=(a,a)I_{a}=(-a,a). We define a mapping ϕ:L2(Ia)\phi\colon\mathbb{C}\to L^{2}(I_{a}) by

ϕ(t,x)=ϕ(x)(t)=12πeitx¯,(t,x)Ia×.\phi(t,x)=\phi(x)(t)=\frac{1}{\sqrt{2\pi}}e^{it\overline{x}},\quad(t,x)\in I_{a}\times\mathbb{C}. (12)

The set ϕ()\phi(\mathbb{C}) is complete in L2(Ia)L^{2}(I_{a}), because the family {eiπnt/a}n\{e^{i\pi nt/a}\}_{n\in\mathbb{Z}} is a complete orthogonal system in IaI_{a}. Therefore, by Theorem 2.1, the integral transform ϕ^:L2(Ia)ϕ^(L2(Ia))\hat{\phi}\colon L^{2}(I_{a})\to\hat{\phi}(L^{2}(I_{a})) is an isometry. The integral transform induced by ϕ\phi is related to the Fourier transform \mathcal{F} on L2()L^{2}(\mathbb{R}) as follows: for fL2()f\in L^{2}(\mathbb{R}),

ϕ^(f|Ia)=12πIaf(t)eit𝑑t=(fχIa).\hat{\phi}(f|_{I_{a}})=\frac{1}{\sqrt{2\pi}}\int_{I_{a}}f(t)e^{-it\cdot}\,dt=\mathcal{F}(f\chi_{I_{a}}).

The image ϕ^(L2(Ia))\hat{\phi}(L^{2}(I_{a})) of the integral transform, denoted by PW(a)PW(a), is an RKHS on \mathbb{C}, called the Paley-Wiener space of order aa. The reproducing kernel Ka(x,y)K_{a}(x,y) of PW(a)PW(a) is given by, for (x,y)2(x,y)\in\mathbb{C}^{2},

Ka(x,y)\displaystyle K_{a}(x,y) =ϕ(y),ϕ(x)L2(Ia)=12πIaeity¯eitx¯¯𝑑t=sin(a(xy¯))π(xy¯).\displaystyle=\langle\phi(y),\phi(x)\rangle_{L^{2}(I_{a})}=\frac{1}{2\pi}\int_{I_{a}}e^{it\overline{y}}\overline{e^{it\overline{x}}}\,dt=\frac{\sin(a(x-\overline{y}))}{\pi(x-\overline{y})}. (13)

Since the integral kernel ϕ(x)¯\overline{\phi(x)} of PW(a)PW(a) is an entire function of xx, it is easy to see that differentiation and integration can be interchanged. This implies that elements of PW(a)PW(a) are entire functions. The uniqueness theorem for analytic functions implies that if an entire function vanishes on \mathbb{R}, then it is identically zero on \mathbb{C}. Thus, the real line \mathbb{R} is a uniqueness set for PW(a)PW(a). The following lemma provides the equality (5) for the norm in PW(a)PW(a), which is essential for our purposes.

Lemma 4.1.

For all x,yx,y\in\mathbb{R},

sin(a(tx))sin(a(ty))(tx)(ty)𝑑t=πsin(a(yx))yx.\int_{\mathbb{R}}\frac{\sin(a(t-x))\sin(a(t-y))}{(t-x)(t-y)}\,dt=\pi\frac{\sin(a(y-x))}{y-x}.
Proof.

Using partial fraction decomposition, we have

sin(a(tx))sin(a(ty))(tx)(ty)𝑑t\displaystyle\int_{\mathbb{R}}\frac{\sin(a(t-x))\sin(a(t-y))}{(t-x)(t-y)}\,dt =1yx{sin(a(tx))sin(a(ty))tydt\displaystyle=\frac{1}{y-x}\Bigl{\{}\int_{-\infty}^{\infty}\frac{\sin(a(t-x))\sin(a(t-y))}{t-y}\,dt (14)
sin(a(tx))sin(a(ty))txdt}\displaystyle\qquad\qquad-\int_{-\infty}^{\infty}\frac{\sin(a(t-x))\sin(a(t-y))}{t-x}\,dt\Bigr{\}} (15)
=F(x,y)F(y,x)yx.\displaystyle=\frac{F(x,y)-F(y,x)}{y-x}. (16)

Here, the two integrals on the right-hand side are improper integrals, and the first one is denoted by F(x,y)F(x,y). To compute F(x,y)F(x,y), we make the substitution u=tyu=t-y:

F(x,y)\displaystyle F(x,y) =sin(a(tx))sin(a(ty))ty𝑑t\displaystyle=\int_{-\infty}^{\infty}\frac{\sin(a(t-x))\sin(a(t-y))}{t-y}\,dt (17)
=sin(a(t+yx))sin(at)t𝑑t\displaystyle=\int_{-\infty}^{\infty}\frac{\sin(a(t+y-x))\sin(at)}{t}\,dt (18)
=limnnncos(at)sin(a(yx))sin(at)t𝑑t\displaystyle=\lim_{n\to\infty}\int_{-n}^{n}\frac{\cos(at)\sin(a(y-x))\sin(at)}{t}\,dt (19)
=sin(a(yx))2sin(2at)t𝑑t=sin(a(yx))0sintt𝑑t\displaystyle=\frac{\sin(a(y-x))}{2}\int_{-\infty}^{\infty}\frac{\sin(2at)}{t}\,dt=\sin(a(y-x))\int_{0}^{\infty}\frac{\sin t}{t}\,dt (20)
=π2sin(a(yx)).\displaystyle=\frac{\pi}{2}\sin(a(y-x)). (21)

Substituting this into (16), we obtain the desired result. ∎

In Lemma 3.1 let TT be the restriction map to \mathbb{R}, and set H=PW(a)H=PW(a), E=E=\mathbb{C}, F=F=\mathbb{R}, and K=L2()K=L^{2}(\mathbb{R}). Lemma 4.1 implies that the norm of PW(a)PW(a) is given by the L2()L^{2}(\mathbb{R}) norm (cf. [Saitoh97]*p. 62): for fPW(a)f\in PW(a),

fPW(a)2=|f(x)|2𝑑x.\|f\|_{PW(a)}^{2}=\int_{\mathbb{R}}|f(x)|^{2}\,dx. (22)

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 t=(ti),x=(xi)Nt=(t_{i}),x=(x_{i})\in\mathbb{C}^{N}, let tx=i=1Ntix¯it\cdot x=\sum_{i=1}^{N}t_{i}\overline{x}_{i}.

Theorem 4.1 (Plancherel’s Theorem).

For fL1(N)f\in L^{1}(\mathbb{R}^{N}) (NN\in\mathbb{N}), the Fourier transform f\mathcal{F}f and the conjugate Fourier transform f\mathcal{F}^{*}f are defined for tNt\in\mathbb{R}^{N} by

f(t)\displaystyle\mathcal{F}f(t) =1(2π)N/2Nf(x)eitx𝑑x,\displaystyle=\frac{1}{(2\pi)^{N/2}}\int_{\mathbb{R}^{N}}f(x)e^{-it\cdot x}\,dx, (23)
f(t)\displaystyle\mathcal{F}^{*}f(t) =1(2π)N/2Nf(x)eitx𝑑x.\displaystyle=\frac{1}{(2\pi)^{N/2}}\int_{\mathbb{R}^{N}}f(x)e^{it\cdot x}\,dx. (24)

Then, the restrictions of \mathcal{F} and its conjugate \mathcal{F}^{*} to L1(N)L2(N)L^{1}(\mathbb{R}^{N})\cap L^{2}(\mathbb{R}^{N}) extend to unitary operators on L2(N)L^{2}(\mathbb{R}^{N}), and they are inverses of each other.

Proof.

The integral transform ϕ^:L2(Ia)PW(a)\hat{\phi}\colon L^{2}(I_{a})\to PW(a) is a Fourier transform. Since ϕ^\hat{\phi} is an isometry, its extension ϕ^:L2(Ia)L2()\hat{\phi}\colon L^{2}(I_{a})\to L^{2}(\mathbb{R}) is also an isometry. As L2(Ia)NL2(IaN)L^{2}(I_{a})^{\otimes N}\cong L^{2}(I_{a}^{N}), the operator ϕ^N:L2(Ia)NPW(a)N\hat{\phi}^{\otimes N}:L^{2}(I_{a})^{\otimes N}\to PW(a)^{\otimes N} can be identified with the Fourier transform on L2(IaN)L^{2}(I_{a}^{N}). By Proposition 3.3, the tensor product of isometries is an isometry. Therefore, ϕ^N\hat{\phi}^{\otimes N} is an isometry. The Fourier transform on L2(N)L^{2}(\mathbb{R}^{N}) is defined as the mean convergence of the Fourier transform on the Cartesian products IaNI_{a}^{N} as aa\to\infty. From the above, it follows that the Fourier transform :L2(N)L2(N)\mathcal{F}:L^{2}(\mathbb{R}^{N})\to L^{2}(\mathbb{R}^{N}) is an isometry. Similarly, the conjugate Fourier transform :L2(N)L2(N)\mathcal{F}^{*}:L^{2}(\mathbb{R}^{N})\to L^{2}(\mathbb{R}^{N}) is also an isometry.

Next, we show that \mathcal{F} and \mathcal{F}^{*} are mutual inverses, i.e., ==I\mathcal{F}\mathcal{F}^{*}=\mathcal{F}^{*}\mathcal{F}=I, where II denotes the identity. Let 𝒲(Ia)=H0,1(Ia)\mathcal{W}(I_{a})=H_{0,1}(I_{a}). Then 𝒲(Ia)\mathcal{W}(I_{a}) is an RKHS on IaI_{a}, and the indefinite integral operator S:fL2(Ia)0f(x)𝑑x𝒲(Ia)S\colon f\in L^{2}(I_{a})\mapsto\int_{0}^{\bullet}f(x)\,dx\in\mathcal{W}(I_{a}) is an isometry. Considering the transformation sequence on IaI_{a},

Ia{I_{a}}L2(Ia){L^{2}(I_{a})}𝒲(Ia){\mathcal{W}(I_{a})}ϕ\scriptstyle{\phi}S\scriptstyle{S}

and taking the NN-fold tensor product and applying Proposition 3.3, we obtain the following transformation sequence on IaNI_{a}^{N}:

IaNL2(Ia)N𝒲(Ia)NϕNSN.\leavevmode\hbox to152.63pt{\vbox to19.33pt{\pgfpicture\makeatletter\hbox{\hskip 76.31258pt\lower-9.71622pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-76.31258pt}{-9.61638pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 8.37585pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.07031pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${I_{a}^{N}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 8.37585pt\hfil&\hfil\hskip 45.8003pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-17.49478pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${L^{2}(I_{a})^{\otimes N}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 21.80032pt\hfil&\hfil\hskip 46.1364pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-17.83089pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\mathcal{W}(I_{a})^{\otimes N}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 22.13643pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-59.36087pt}{-7.11638pt}\pgfsys@lineto{-36.16086pt}{-7.11638pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-35.96088pt}{-7.11638pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-53.02684pt}{-3.40251pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\phi^{\otimes N}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{8.23975pt}{-7.11638pt}\pgfsys@lineto{31.43976pt}{-7.11638pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{31.63974pt}{-7.11638pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{14.31128pt}{-4.76361pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{S^{\otimes N}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}.

Since the tensor product Hilbert space L2(Ia)NL^{2}(I_{a})^{\otimes N} is isomorphic to L2(IaN)L^{2}(I_{a}^{N}), we have, for t=(tj),x=(xj)Nt=(t_{j}),\ x=(x_{j})\in\mathbb{R}^{N},

ϕN(t,x)=1(2π)N/2eitx.\phi^{\otimes N}(t,x)=\frac{1}{(2\pi)^{N/2}}e^{it\cdot x}.

On the other hand, by the definition of the tensor product of maps and Fubini’s theorem, for I(t)=j=1N(0,tj)(IaN)I(t)=\prod_{j=1}^{N}(0,t_{j})\ (\subset I_{a}^{N}) and f1,,fNL2(Ia)f_{1},\dots,f_{N}\in L^{2}(I_{a}), we have

SN(j=1Nfj)(t)=j=1NSfj(tj)=j=1N0tjfj(sj)dsj=I(t)(j=1Nfj)(s)ds.S^{\otimes N}(\otimes_{j=1}^{N}f_{j})(t)=\prod_{j=1}^{N}Sf_{j}(t_{j})=\prod_{j=1}^{N}\int_{0}^{t_{j}}f_{j}(s_{j})\,ds_{j}=\int_{I(t)}(\otimes_{j=1}^{N}f_{j})(s)\,ds.

Since the span of simple tensors j=1Nfj\otimes_{j=1}^{N}f_{j} is dense in L2(IaN)L^{2}(I_{a}^{N}), it follows that for all fL2(IaN)f\in L^{2}(I_{a}^{N}),

(SNf)(t)=I(t)f(s)𝑑s,(S^{\otimes N}f)(t)=\int_{I(t)}f(s)\,ds,

i.e., SNS^{\otimes N} is the NN-dimensional indefinite integral operator. Similarly, the norm of PW(a)NPW(a)^{\otimes N} which is an RKHS on N\mathbb{C}^{N} satisfies

fPW(a)N2=N|f(x)|2𝑑x\|f\|_{PW(a)^{\otimes N}}^{2}=\int_{\mathbb{R}^{N}}|f(x)|^{2}dx

for fPW(a)Nf\in PW(a)^{\otimes N}. Therefore, when fL2(N)f\in L^{2}(\mathbb{R}^{N}) and f|IaNL2(IaN)f|_{I_{a}^{N}}\in L^{2}(I_{a}^{N}), applying the inverse transformation formula (6), we obtain, for t=(tj)IaNt=(t_{j})\in I_{a}^{N}, x=(xj)Nx=(x_{j})\in\mathbb{R}^{N},

I(t)f(s)𝑑s=1(2π)N/2N(fχIaN)(x)I(t)eisx𝑑s𝑑x.\int_{I(t)}f(s)\,ds=\frac{1}{(2\pi)^{N/2}}\int_{\mathbb{R}^{N}}\mathcal{F}(f\chi_{I_{a}^{N}})(x)\int_{I(t)}e^{is\cdot x}\,ds\,dx. (25)

By Theorem 3.1, for fixed tIaNt\in I_{a}^{N}, we have I(t)eisx𝑑sL2(N)\int_{I(t)}e^{is\cdot x}\,ds\in L^{2}(\mathbb{R}^{N}). Since the Fourier transform is an isometry, (fχIaN)(f)\mathcal{F}(f\chi_{I_{a}^{N}})\to\mathcal{F}(f) (aa\to\infty) in L2(N)L^{2}(\mathbb{R}^{N}). Thus, letting aa\to\infty in (25), we obtain, for all fL2(N)f\in L^{2}(\mathbb{R}^{N}) and tNt\in\mathbb{R}^{N},

I(t)f(x)𝑑x\displaystyle\int_{I(t)}f(x)\,dx =1(2π)N/2N(f)(x)I(t)eisx𝑑s𝑑x.\displaystyle=\frac{1}{(2\pi)^{N/2}}\int_{\mathbb{R}^{N}}\mathcal{F}(f)(x)\int_{I(t)}e^{is\cdot x}\,ds\,dx. (26)

If, in addition, (f)L1(N)\mathcal{F}(f)\in L^{1}(\mathbb{R}^{N}), then, since |eisx|=1|e^{is\cdot x}|=1, we can apply Fubini’s theorem to interchange the order of integration in (26), obtaining

I(t)f(x)𝑑x=I(t)(f)(s)𝑑s.\int_{I(t)}f(x)\,dx=\int_{I(t)}\mathcal{F}^{*}\mathcal{F}(f)(s)\,ds.

Furthermore, if we write x=(x1,x)x=(x_{1},x^{\prime}) and t=(t1,t)Nt=(t_{1},t^{\prime})\in\mathbb{R}^{N}, then by Fubini’s theorem we have,

0t1I(t)f(x1,x)𝑑x𝑑x1=0t1I(t)(f)(x)𝑑x𝑑x1.\int_{0}^{t_{1}}\int_{I(t^{\prime})}f(x_{1},x^{\prime})\,dx^{\prime}\,dx_{1}=\int_{0}^{t_{1}}\int_{I(t^{\prime})}\mathcal{F}^{*}\mathcal{F}(f)(x^{\prime})\,dx^{\prime}\,dx_{1}.

Differentiating both sides with respect to t1t_{1} at t1=x1t_{1}=x_{1}, we obtain, for almost every xN1x^{\prime}\in\mathbb{R}^{N-1},

I(t)f(x1,x)𝑑x=I(t)(f)(x)𝑑x.\int_{I(t^{\prime})}f(x_{1},x^{\prime})\,dx^{\prime}=\int_{I(t^{\prime})}\mathcal{F}^{*}\mathcal{F}(f)(x^{\prime})\,dx^{\prime}.

Repeating this differentiation N1N-1 times with respect to t2,,tNt_{2},\dots,t_{N} at t2=x2,,tN=xNt_{2}=x_{2},\dots,t_{N}=x_{N}, we conclude that f(x)=(f)(x)f(x)=\mathcal{F}^{*}\mathcal{F}(f)(x) for almost every xNx\in\mathbb{R}^{N}. Now, if fCc(N)f\in C_{c}^{\infty}(\mathbb{R}^{N}), then by integration by parts we obtain the identity

(2Nfx12xN2)=(1)Ni=1Nxi2(f).\mathcal{F}\Bigl{(}\frac{\partial^{2N}f}{\partial x_{1}^{2}\cdots\partial x_{N}^{2}}\Bigr{)}=(-1)^{N}\prod_{i=1}^{N}x_{i}^{2}\cdot\mathcal{F}(f).

Since the left-hand side is bounded and since (f)L2(N)\mathcal{F}(f)\in L^{2}(\mathbb{R}^{N}), it is easy to see that (f)L1(N)\mathcal{F}(f)\in L^{1}(\mathbb{R}^{N}). Since Cc(N)C_{c}^{\infty}(\mathbb{R}^{N}) is dense in L2(N)L^{2}(\mathbb{R}^{N}), we conclude that the set of ff with (f)L1(N)\mathcal{F}(f)\in L^{1}(\mathbb{R}^{N}) is dense in L2(N)L^{2}(\mathbb{R}^{N}). Therefore, by the boundedness of \mathcal{F} and \mathcal{F}^{*}, the desired identity =I\mathcal{F}^{*}\mathcal{F}=I holds by continuity. Similarly, it can be shown that =I\mathcal{F}\mathcal{F}^{*}=I. Thus, Plancherel’s theorem is demonstrated. ∎

References