Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-29T00:37:30.037Z Has data issue: false hasContentIssue false

On the kernel of the $(\kappa ,a)$-Generalized fourier transform

Published online by Cambridge University Press:  14 August 2023

Dmitry Gorbachev
Affiliation:
Department of Applied Mathematics and Computer Science, Tula State University, 300012 Tula, Russia; E-mail: dvgmail@mail.ru
Valerii Ivanov
Affiliation:
Department of Applied Mathematics and Computer Science, Tula State University, 300012 Tula, Russia; E-mail: ivaleryi@mail.ru
Sergey Tikhonov
Affiliation:
Centre de Recerca Matemàtica, Campus de Bellaterra, Edifici C, 08193 Bellaterra, Barcelona, Spain; E-mail: stikhonov@crm.cat Catalan Institution for Research and Advanced Studies, Passeig de Lluís Companys 23, 08010 Barcelona, Spain; E-mail: stikhonov@crm.cat Department of Mathematics, Universitat Autònoma de Barcelona, Building C Science Faculty, 08193 Bellaterra, Barcelona, Spain; E-mail: stikhonov@crm.cat

Abstract

For the kernel $B_{\kappa ,a}(x,y)$ of the $(\kappa ,a)$-generalized Fourier transform $\mathcal {F}_{\kappa ,a}$, acting in $L^{2}(\mathbb {R}^{d})$ with the weight $|x|^{a-2}v_{\kappa }(x)$, where $v_{\kappa }$ is the Dunkl weight, we study the important question of when $\|B_{\kappa ,a}\|_{\infty }=B_{\kappa ,a}(0,0)=1$. The positive answer was known for $d\ge 2$ and $\frac {2}{a}\in \mathbb {N}$. We investigate the case $d=1$ and $\frac {2}{a}\in \mathbb {N}$. Moreover, we give sufficient conditions on parameters for $\|B_{\kappa ,a}\|_{\infty }>1$ to hold with $d\ge 1$ and any a.

We also study the image of the Schwartz space under the $\mathcal {F}_{\kappa ,a}$ transform. In particular, we obtain that $\mathcal {F}_{\kappa ,a}(\mathcal {S}(\mathbb {R}^d))=\mathcal {S}(\mathbb {R}^d)$ only if $a=2$. Finally, extending the Dunkl transform, we introduce nondeformed transforms generated by $\mathcal {F}_{\kappa ,a}$ and study their main properties.

Type
Analysis
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press

1 Introduction

Let as usual $\Delta $ be the Laplacian operator in $\mathbb {R}^{d}$ . For the Fourier transform

$$\begin{align*}\mathcal{F}(f)(y)=(2\pi)^{-d/2}\int_{\mathbb{R}^{d}}f(x)e^{-i\langle x,y\rangle}\,dx ,\end{align*}$$

Howe [Reference Howe16] obtained the following spectral decomposition of $\mathcal {F}$ using the harmonic oscillator $-({\Delta -|x|^2})/2$ and its eigenfunctions forming the basis in $L^{2}(\mathbb {R}^{d})$ :

$$\begin{align*}\mathcal{F}=\exp\Bigl(\frac{i\pi d}{4}\Bigr)\exp\Bigl(\frac{i\pi}{4}\,\bigl(\Delta-|x|^{2}\bigr)\Bigr). \end{align*}$$

Among other applications, this decomposition is useful to define the fractional power of the Fourier transform (see [Reference Ben Saïd, Kobayashi and Ørsted4, Reference Kobayashi and Mano19]).

During the last 30 years, a lot of attention has been given to various generalizations of the Fourier transform. As an important example, to develop harmonic analysis on weighted spaces, the Dunkl transform was introduced in [Reference Dunkl12]. The Dunkl transform $\mathcal {F}_\kappa $ is defined with the help of a root system $\Omega \subset \mathbb {R}^{d}$ , a reflection group $G\subset O(d)$ , and multiplicity function $\kappa \colon \Omega \to \mathbb {R}_+$ , such that $\kappa $ is G-invariant. Here, G is generated by reflections $\{\sigma _{\alpha }\colon \alpha \in \Omega \}$ , where $\sigma _{\alpha }$ is a reflection with respect to hyperplane $(\alpha ,x)=0$ .

The differential-difference Dunkl Laplacian operator $\Delta _{\kappa }$ plays the role of the classical Laplacian [Reference Rösler20]. If $\kappa \equiv 0$ , we have $\Delta _{\kappa }=\Delta $ . Dunkl Laplacian allows us to define the Dunkl harmonic oscillator $\Delta _{\kappa }-|x|^{2}$ and the Dunkl transform

$$\begin{align*}\mathcal{F}_{\kappa}=\exp\Bigl(\frac{i\pi}{2}\,\Bigl(\frac{d}{2}+\langle \kappa\rangle\Bigr)\Bigr) \exp\Bigl(\frac{i\pi}{4}\bigl(\Delta_{\kappa}-|x|^{2}\bigr)\Bigr), \end{align*}$$

where

$$\begin{align*}\langle \kappa\rangle=\frac{1}{2}\sum_{\alpha\in \Omega}\kappa(\alpha). \end{align*}$$

Further extensions of Fourier and Dunkl transforms were obtained by Ben Saïd, Kobayashi, and Ørsted in [Reference Ben Saïd, Kobayashi and Ørsted4]. They defined the a-deformed Dunkl harmonic oscillator

$$\begin{align*}\Delta_{\kappa,a}=|x|^{2-a}\Delta_{\kappa}-|x|^{a},\quad a>0, \end{align*}$$

and the $(\kappa ,a)$ -generalized Fourier transform

(1.1) $$ \begin{align} \mathcal{F}_{\kappa,a}=\exp\Bigl(\frac{i\pi}{2}\,(\lambda_{\kappa,a}+1)\Bigr) \exp\Bigl(\frac{i\pi}{2a}\,\Delta_{\kappa,a}\Bigr), \end{align} $$

which is a two-parameter family of unitary operators in $L^{2}(\mathbb {R}^{d},d\mu _{\kappa ,a})$ equipped with the norm

$$\begin{align*}\|f\|_{2,d\mu_{\kappa,a}}=\Bigl(\int_{\mathbb{R}^{d}}|f(x)|^{2}\,d\mu_{\kappa,a}(x)\Bigr)^{1/2}. \end{align*}$$

Here

$$\begin{align*}\lambda_{\kappa, a}=\frac{2\lambda_{\kappa}}{a},\quad \lambda_{\kappa}=\langle \kappa\rangle+\frac{d-2}{2},\quad d\mu_{\kappa,a}(x)=c_{\kappa,a}v_{\kappa,a}(x)\,dx,\quad v_{\kappa,a}(x)=|x|^{a-2}v_{\kappa}(x), \end{align*}$$
$$\begin{align*}v_{\kappa}(x)=\prod_{\alpha\in \Omega}|\langle\alpha,x\rangle|^{\kappa(\alpha)},\quad c^{-1}_{\kappa,a}=\int_{\mathbb{R}^{d}}e^{-|x|^{a}/a}v_{\kappa,a}(x)\,dx. \end{align*}$$

Throughout the paper, we assume that $ d+2\langle \kappa \rangle +a-2=2\lambda _{\kappa }+a>0$ or, equivalently, $\lambda _{\kappa , a}>-1$ . Note that under this condition, the weight function $v_{\kappa ,a}$ is locally integrable.

For $a=2$ , (1.1) reduces to the Dunkl transform, while if $a=2$ and $\kappa \equiv 0$ , then (1.1) is the classical Fourier transform. For $a\ne 2$ , we arrive at deformed Dunkl and Fourier transforms, which have various applications. In particular, for $a=1$ and $\kappa \equiv 0$ , the deformed Dunkl transform is the unitary inversion operator of the Schrödinger model of minimal representation of the group $O(N+1,2)$ [Reference Kobayashi and Mano19].

The unitary operator $\mathcal {F}_{\kappa ,a}$ on $L^{2}(\mathbb {R}^{d},d\mu _{\kappa ,a})$ can be written as the integral transform [Reference Ben Saïd, Kobayashi and Ørsted4, (5.8)]

$$ \begin{align*} \mathcal{F}_{\kappa,a}(f)(y)=\int_{\mathbb{R}^{d}}B_{\kappa,a}(x,y)f(x)\,d\mu_{\kappa,a}(x) \end{align*} $$

with the continuous symmetric kernel $B_{\kappa ,a}(x,y)$ satisfying $B_{\kappa ,a}(0,y)=1$ . In particular, $B_{0,2}(x,y)=e^{-i\langle x,y\rangle }$ . One of the fundamental questions in the theory of deformed transforms is to investigate basic properties of the kernel $B_{\kappa ,a}(x,y)$ , in particular, to know when it is uniformly bounded. To illustrate the importance of this property, note that the condition $|B_{\kappa ,a}(x,y)|\le M$ implies the Hausdorff-Young inequality

$$\begin{align*}\bigl\|\mathcal{F}_{\kappa,a}(f)\bigr\|_{p',d\mu_{k,a}}\le M^{2/p-1}\bigl\|f\bigr\|_{p,d\mu_{\kappa,a}},\quad 1\le p\le 2,\quad \frac{1}{p}+\frac{1}{p'}=1. \end{align*}$$

A more important problem is to describe parameters so that there holds

(1.2) $$ \begin{align} \|B_{\kappa,a}\|_{\infty}=\sup_{x,y\in\mathbb{R}^{d}}|B_{\kappa,a}(x,y)|=B_{\kappa,a}(0,0)=1. \end{align} $$

In this case, the Hausdorff-Young inequality holds with the constant $1$ and one can define the generalized translation operator $\tau ^yf(x)$ in $L^{2}(\mathbb {R}^{d},d\mu _{k,a})$ by

$$\begin{align*}\mathcal{F}_{\kappa,a}(\tau^yf)(z)=B_{\kappa,a}(y,z)\mathcal{F}_{\kappa,a}(f)(z) \end{align*}$$

(see [Reference Gorbachev, Ivanov and Tikhonov15]), and moreover, its norm equals 1.

Let us list the known cases when (1.2) holds:

In this paper, we continue to study the case

$$\begin{align*}\frac{2}{a}\in\mathbb{N}. \end{align*}$$

Its importance was discussed in [Reference Ben Saïd, Kobayashi and Ørsted4], [Reference Boubatra, Negzaoui and Sifi7], and [Reference Constales, De Bie and Lian10]. For example, in this case, one has a simple inversion formula for the $(\kappa ,a)$ -generalized Fourier transform [Reference Ben Saïd, Kobayashi and Ørsted4, Theorem 5.3] and an explicit expression for the generalized translation operator $\tau ^y$ [Reference Boubatra, Negzaoui and Sifi7]. Moreover, the kernel $B_{\kappa ,2/n}$ with $n\in \mathbb {N}$ appears for dihedral groups G [Reference Constales, De Bie and Lian10].

Our first goal in this paper is, on the one hand, to extend the list of parameters for which (1.2) holds for $\frac {2}{a}\in \mathbb {N}$ ; on the other hand, to point out the cases when (1.2) does not hold. The following theorem describes positive results, where, for completeness, we include all known cases.

Theorem 1.1 (see [Reference Constales, De Bie and Lian10, $d\ge 2$ ]).

Let $0<a\le 1$ , $\frac {2}{a}\in \mathbb {N}$ . If $d=1$ , $\langle \kappa \rangle \ge \frac {1}{2}$ or $d\ge 2$ , $\langle \kappa \rangle \ge 0$ , then equality (1.2) is true.

Our proof of Theorem 1.1 for $d=1$ is based on an integral representation of $B_{\kappa ,a}$ with the special kernel and a study of positiveness of this kernel. This approach is closely related to the theory of positive definite functions.

In the general case $d\ge 1$ , we give the proof based on the approach developed in the papers [Reference De Bie and Lian6, Reference Constales, De Bie and Lian10, Reference Deleaval and Demni11], see Section 4.1, and the alternative proof based on representation with positive kernels, see Subsection 4.3.

With regard to negative results, we obtain the following theorem, where we specify parameters when Theorem 1.1 does not hold.

Theorem 1.2. In either of the following cases:

  • $d=1$ , $0<a\le 1$ , and $\langle \kappa \rangle =\frac {1}{2}-\frac {a}{4}$ , or

  • $d\ge 1$ , $a\in (1,2)\cup (2,\infty )$ and $\langle \kappa \rangle \ge 0$ ,

we have

(1.3) $$ \begin{align} \|B_{\kappa,a}\|_{\infty}>1. \end{align} $$

The rest of the paper is organized as follows. Section 2 is devoted to the proof of Theorem 1.1 in the case $d=1$ . In Section 3, we study the properties of the one-dimensional kernel $B_{\kappa ,a}$ for $\lambda _{\kappa ,a}<0$ . In particular, in Section 3.1, we investigate positive definiteness of kernels of the integral transforms generated by ${\mathcal {F}}_{\kappa ,a}$ . In Section 4, we prove Theorem 1.1 in full generality as well as Theorem 1.2 (Section 4.3).

In Section 5, we study the question of how the $\mathcal {F}_{\kappa ,a}$ transform acts on Schwartz functions. The Schwartz space $\mathcal {S}(\mathbb {R}^{d})$ is invariant under the classical Fourier transform $\mathcal {F}_{0,2}$ and the Dunkl $\mathcal {F}_{\kappa ,2}$ (see [Reference de Jeu8]), but the case of deformed transforms is more complicated. In fact, we show that $\mathcal {S}(\mathbb {R}^d)$ is not invariant under $\mathcal {F}_{\kappa ,a}$ for $a\neq 2$ , which contradicts a widely used statement in [Reference Johansen18] (see Remark 5.3). If $\frac {a}{2}\notin \mathbb {N}$ , then the generalized Fourier transform may not be infinitely differentiable, and if $\frac {2}{a}\notin \mathbb {N}$ , then it may not be rapidly decreasing at infinity. For $d=1$ and $\frac {2}{a}\in \mathbb {N}$ , the generalized Fourier transform of $f\in \mathcal {S}(\mathbb {R})$ is rapidly decreasing due to the representation $\mathcal {F}_{\kappa ,a}(f)(y)=F_1\bigl (|y|^{a/2}\bigr )+yF_2\bigl (|y|^{a/2}\bigr )$ , where the even functions $F_1,F_2\in \mathcal {S}(\mathbb {R})$ (see Proposition 5.4).

Finally, in Section 6, we study one-dimensional nondeformed unitary transforms generated by $\mathcal {F}_{\kappa ,a}$ :

$$ \begin{align*} \mathcal{F}_{r}^{\lambda}(g)(v)=\int_{-\infty}^{\infty}e_{2r+1}(uv,\lambda)g(u)\,\frac{|u|^{2\lambda+1} \,du}{2^{\lambda+1}\Gamma(\lambda+1)}, \end{align*} $$

where $r\in \mathbb {Z}_+$ , $\lambda \ge -1/2$ , and the kernel

$$\begin{align*}e_{2r+1}(uv,\lambda)=j_{\lambda}(uv)+i(-1)^{r+1}\,\frac{(uv)^{2r+1}}{2^{2r+1}(\lambda+1)_{2r+1}}\, j_{\lambda+2r+1}(uv) \end{align*}$$

is an eigenfunction of the differential-difference operator

$$\begin{align*}\delta_{\lambda}g(u)=\Delta_{\lambda+1/2}g(u)-2r(\lambda+r+1)\,\frac{g(u)-g(-u)}{u^2}. \end{align*}$$

Here, $\Delta _{\lambda +1/2}$ is the one-dimensional Dunkl Laplacian for $\langle \kappa \rangle =\lambda +\frac {1}{2}$ and

$$\begin{align*}(a)_0=1,\quad (a)_n=\frac{\Gamma(a+n)}{\Gamma(a)}=a(a+1)\cdots(a+n-1),\quad n\ge 1, \end{align*}$$

is the Pochhammer symbol. Note that such unitary transforms give new examples of an important class of Bessel-Hankel type transforms with the kernel $k(uv)$ (see, e.g. [Reference Titchmarsh24, Chapter VIII]). In particular, they generalize the one-dimensional Dunkl transform ( $r=0$ ).

2 Proof of Theorem 1.1 in the one-dimensional case

In what follows, we assume that

$$\begin{align*}d=1,\quad a>0,\quad \kappa=\langle\kappa\rangle\ge 0,\quad \lambda_{\kappa}=\kappa-\frac{1}{2},\quad 2\lambda_{\kappa}+a>0,\quad\lambda=\lambda_{\kappa,a}=\frac{2\lambda_{\kappa}}{a}, \end{align*}$$
$$\begin{align*}v_{\kappa,a}(x)=|x|^{2\kappa+a-2},\quad d\mu_{\kappa,a}(x)=c_{\kappa,a}v_{\kappa,a}(x)\,dx,\quad c_{\kappa,a}=\frac{1}{2a^{\lambda}\Gamma(\lambda+1)}, \end{align*}$$

and $\mathcal {F}_{\kappa ,a}$ is the $(\kappa ,a)$ -generalized Fourier transform (1.1) on the real line. Firstly, let us investigate when the kernel of $\mathcal {F}_{\kappa ,a}$ is uniformly bounded. Using [Reference Ben Saïd, Kobayashi and Ørsted4, Section 5], we can write the kernel as

(2.1) $$ \begin{align} B_{\kappa,a}(x,y)=j_{\lambda}\Bigl(\frac{2}{a}\,|xy|^{a/2}\Bigr)+ \frac{\Gamma(\lambda+1)}{\Gamma(\lambda+1+2/a)}\,\frac{xy}{(ai)^{2/a}}\, j_{\lambda+\frac{2}{a}}\Bigl(\frac{2}{a}\,|xy|^{a/2}\Bigr), \end{align} $$

where $j_{\lambda }(x)=2^{\lambda }\Gamma (\lambda +1)x^{-\lambda }J_{\lambda }(x)$ is the normalized Bessel function and $J_{\lambda }(x)$ is the classical Bessel function. Then the asymptotic behavior of $J_{\lambda }(x)$ (see [Reference Watson25, Chapter VII, 7.1]) immediately allows us to derive the following

Proposition 2.1. Let $d=1$ . The conditions

(2.2) $$ \begin{align} 0<a\le 2,\quad \kappa\ge\frac{1}{2}-\frac{a}{4},\quad\text{or}\quad a \ge 2,\quad \kappa\ge 0, \end{align} $$

are necessary and sufficient for boundedness of the kernel $B_{\kappa ,a}(x,y)$ .

The main goal of this section is to prove Theorem 1.1 for $d=1$ .

Proof. Note that $B_{\kappa ,a}(x,y)=b_{\kappa ,a}(xy)$ , where

(2.3) $$ \begin{align} b_{\kappa,a}(x)=j_{\lambda}\Bigl(\frac{2}{a}\,|x|^{a/2}\Bigr)+ \frac{\Gamma(\lambda+1)}{\Gamma(\lambda+1+2/a)}\,\frac{x}{(ai)^{2/a}}\, j_{\lambda+\frac{2}{a}}\Bigl(\frac{2}{a}\,|x|^{a/2}\Bigr). \end{align} $$

Therefore, under the conditions of Theorem 1.1, it suffices to establish the inequality $|b_{\kappa ,a}(x)|\le 1$ for $x\in \mathbb {R}$ .

Let $a=\frac {2}{R}$ , $R\in \mathbb {N}$ . Equality (2.3) can be written as

$$\begin{align*}b_{\kappa,a}(x)=j_{\lambda}\Bigl(R|x|^{1/R}\Bigr)+ \frac{\Gamma(\lambda+1)}{\Gamma(\lambda+1+R)}\Bigl(\frac{R}{2}\Bigr)^{R}(-i)^{R}\,x j_{\lambda+R}\Bigl(R|x|^{1/R}\Bigr). \end{align*}$$

Let $x\in \mathbb {R}$ . In the case $R=2r+1$ , $a=\frac {2}{2r+1}$ , $r\in \mathbb {Z}_{+}$ , and Footnote 3 $v=(2r+1)x^{\frac {1}{2r+1}}$ , $x=\bigl (\frac {v}{2r+1}\bigr )^{2r+1}$ , there holds

(2.4) $$ \begin{align} e_{2r+1}(v,\lambda)=b_{\kappa,a}\Bigl(\Bigl(\frac{v}{2r+1}\Bigr)^{2r+1}\Bigr)=j_{\lambda}(v)+ i(-1)^{r+1}\,\frac{v^{2r+1}}{2^{2r+1}(\lambda+1)_{2r+1}}\, j_{\lambda+2r+1}(v). \end{align} $$

In the case $R=2r$ , $a=\frac {1}{r}$ , $r\in \mathbb {N}$ , and $v=2r|x|^{\frac {1}{2r}}{\operatorname{sign}} {}\ x$ , $x=\big (\frac {v}{2r}\big )^{2r}{\operatorname{sign}} {}\ v$ , we have

(2.5) $$ \begin{align} e_{2r}(v,\lambda)=b_{\kappa,a}\Bigl(\Bigl(\frac{v}{2r}\Bigr)^{2r}{\operatorname{sign}}{}\ v\Bigr)=j_{\lambda}(v)+ (-1)^{r}\,\frac{v^{2r}}{2^{2r}(\lambda+1)_{2r}}\,j_{\lambda+2r}(v)\,{\operatorname{sign}}{}\ v. \end{align} $$

In order to see that $|e_{2r+1}(v,\lambda )|, |e_{2r}(v,\lambda )|\le 1$ , we will need several auxiliary results. We start with the following identity

(2.6) $$ \begin{align} \frac{v^2}{4(\lambda+1)(\lambda+2)}\,j_{\lambda+2}(v)=j_{\lambda+1}(v)-j_{\lambda}(v), \end{align} $$

which follows from the recurrence relation for the Bessel function $J_{\lambda }(v)$ (see [Reference Watson25, Chapter III, 3.2]). Then, by induction, we establish

Lemma 2.2. If $r\in \mathbb {N}$ , then

(2.7) $$ \begin{align} \frac{v^{2r}}{2^{2r}(\lambda+1)_{2r}}\,j_{\lambda+2r}(v)= (-1)^{r}j_{\lambda}(v)+\sum_{s=1}^{r-1}(-1)^{s+r} \binom{r}{s}\frac{(\lambda+r)_{s}}{(\lambda+1)_{s}}\,j_{\lambda+s}(v) +\frac{(\lambda+r+1)_{r-1}}{(\lambda+1)_{r-1}}\,j_{\lambda+r}(v). \end{align} $$

Proof. For $r=1$ , the needed formula coincides with (2.6). Assume that (2.7) is valid for every $k\le r-1$ and $\lambda $ . Denote by $a_{s}^r(\lambda )$ , $s=0,1,\dots ,r$ , the coefficients by $j_{\lambda +s}(v)$ in the decomposition (2.7). Taking into account (2.6) and the inductive assumption, we derive that

$$ \begin{align*} \frac{v^{2r}}{2^{2r}(\lambda+1)_{2r}}\,j_{\lambda+2r}(v)&= \frac{v^{2r-2}}{2^{2r-2}(\lambda+1)_{2r-2}}\,\frac{v^2}{4(\lambda+2r-1)(\lambda+2r)}\,j_{(\lambda+2r-2)+2}(v)\\ &=\frac{v^{2r-2}}{2^{2r-2}(\lambda+1)_{2r-2}}\,\{j_{(\lambda+1)+2r-2}(v)-j_{\lambda+2r-2}(v)\}\\ &=\frac{\lambda+2r-1}{\lambda+1}\sum_{s=1}^{r}a_{s-1}^{r-1}(\lambda+1)j_{\lambda+s}(v)- \sum_{s=0}^{r-1}a_{s}^{r-1}(\lambda)j_{\lambda+s}(v). \end{align*} $$

It is enough to show that

$$\begin{align*}a_{0}^{r}(\lambda)=-a_{0}^{r-1}(\lambda),\quad a_{r}^{r}(\lambda)=\frac{\lambda+2r-1}{\lambda+1}\,a_{r-1}^{r-1}(\lambda+1), \end{align*}$$
$$\begin{align*}a_{s}^{r}(\lambda)=\frac{\lambda+2r-1}{\lambda+1}\,a_{s-1}^{r-1}(\lambda+1)-a_{s}^{r-1}(\lambda),\quad s=1,\dots,r-1. \end{align*}$$

Indeed, using the induction step, we have that

$$\begin{align*}a_{0}^{r}(\lambda)= (-1)^r,\quad a_{r}^{r}(\lambda)=\frac{\lambda+2r-1}{\lambda+1}\,\frac{(\lambda+r+1)_{r-2}}{(\lambda+2)_{r-2}}= \frac{(\lambda+r+1)_{r-1}}{(\lambda+1)_{r-1}}, \end{align*}$$

and, for $s=1,\dots ,r-1$ ,

$$ \begin{align*} a_s^r(\lambda)&=(-1)^{s+r}\Bigl\{\frac{\lambda+2r-1}{\lambda+1} \binom{r-1}{s-1}\frac{(\lambda+r)_{s-1}}{(\lambda+2)_{s-1}}+\binom{r-1}{s}\frac{(\lambda+r-1)_{s}}{(\lambda+1)_{s}}\Bigr\} \\[4pt] &=(-1)^{s+r}\binom{r}{s}\frac{(\lambda+r)_{s}}{(\lambda+1)_{s}}\Bigl\{\frac{\lambda+2r-1}{\lambda+r+s-1}\,\frac{s}{r}+ \frac{\lambda+r-1}{\lambda+r+s-1}\,\frac{r-s}{r}\Bigr\}\\[4pt] &=(-1)^{s+r}\binom{r}{s}\frac{(\lambda+r)_{s}}{(\lambda+1)_{s}}, \end{align*} $$

which completes the proof.

Lemma 2.3. For $r\in \mathbb {Z}_{+}$ , we have

$$\begin{align*}\frac{v^{2r+1}}{2^{2r+1}(\lambda+1)_{2r+1}}\,j_{\lambda+2r+1}(v)=(-1)^{r+1}\sum_{s=0}^{r}(-1)^{s} \binom{r}{s}\frac{(\lambda+r+1)_{s}}{(\lambda+1)_{s}}\,j_{\lambda+s}'(v). \end{align*}$$

Proof. Using Lemma 2.2 and the equality

$$\begin{align*}j_{\lambda}'(v)=-\frac{v}{2(\lambda+1)}\,j_{\lambda+1}(v), \end{align*}$$

we derive that

$$ \begin{align*} &\frac{v^{2r+1}}{2^{2r+1}(\lambda+1)_{2r+1}}\,j_{\lambda+2r+1}(v)=\frac{v}{2(\lambda+1)}\, \frac{v^{2r}}{2^{2r}(\lambda+2)_{2r}}\,j_{(\lambda+1)+2r}(v)\\ &\quad =\frac{v}{2(\lambda+1)}\Bigl\{(-1)^{r}j_{\lambda+1}(v)+\sum_{s=1}^{r-1}(-1)^{s+r} \binom{r}{s}\frac{(\lambda+r+1)_{s}}{(\lambda +2)_{s}}\,j_{\lambda+1+s}(v)+ \frac{(\lambda+r+2)_{r-1}}{(\lambda+2)_{r-1}}\,j_{\lambda+1+r}(v)\Bigr\}\\ &\qquad\qquad\qquad\qquad = (-1)^{r+1}j_{\lambda}'(v)+\sum_{s=1}^{r-1}(-1)^{s+r+1} \binom{r}{s}\frac{(\lambda+r+1)_{s}}{(\lambda+1)_s}\,j_{\lambda+s}'(v)- \frac{(\lambda+r+1)_{r}}{(\lambda+1)_r}\,j_{\lambda+r}'(v). \end{align*} $$

Taking into account (2.4) and (2.5), Lemmas 2.2 and 2.3, and

(2.8) $$ \begin{align} j_{\lambda}(v)=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}e^{-ivt}\,dt,\quad j_{\lambda}'(v)=-ic_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}te^{-ivt}\,dt \end{align} $$

with $c_{\lambda }=\frac {\Gamma (\lambda +1)}{\sqrt {\pi }\Gamma (\lambda +1/2)}$ , $\lambda>-1/2$ (see [Reference Watson25, Chapter III, 3.3], we arrive at the following integral representations of the functions $e_{2r+1}(v,\lambda )$ , and $e_{2r}(v,\lambda )$ .

Lemma 2.4. If $r\in \mathbb {Z}_{+}$ , $\lambda>-1/2$ , then

(2.9) $$ \begin{align} e_{2r+1}(v,\lambda)=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}q_{2r+1}(t,\lambda)e^{-ivt}\,dt, \end{align} $$

where $q_{2r+1}(t,\lambda )$ is a polynomial of degree $2r+1$ with respect to t given by

$$\begin{align*}q_{2r+1}(t,\lambda)=1+t\sum_{s=0}^{r}(-1)^{s} \binom{r}{s}\frac{(\lambda+r+1)_{s}}{(\lambda+1/2)_{s}}\,(1-t^2)^s. \end{align*}$$

Lemma 2.5. If $r\in \mathbb {N}$ , $\lambda>-1/2$ , then

$$ \begin{align*} e_{2r}(v,\lambda)=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}q_{2r}(t,\lambda)e^{-ivt}\,dt, \end{align*} $$

where $q_{2r}(t,\lambda )$ is a polynomial of degree $2r$ with respect to t given by

$$\begin{align*}q_{2r}(t,\lambda)=q_{2r}(t,v,\lambda) =1+{\operatorname{sign}}{}\ v\,\Bigl\{\sum_{s=0}^{r}(-1)^{s} \binom{r}{s}\frac{(\lambda+r)_{s}}{(\lambda+1/2)_{s}}\,(1-t^2)^s\Bigr\}. \end{align*}$$

For our further analysis, it is important to know for which $\lambda $ the polynomials $q_{2r}(t,\lambda )$ , $q_{2r+1}(t,\lambda )$ are nonnegative on $[-1,1]$ . If $r=0$ , $q_{1}(t,\lambda )=1-t\ge 0$ on $[-1,1]$ , that is, $q_{1}$ does not depend on $\lambda $ . This special case corresponds to parameters $a=2,\, \kappa \ge 0$ , and hence (2.9) is a well-known integral representation of the kernel of the one-dimensional Dunkl transform. In other cases, positivity conditions for $q_{2r}(t,\lambda )$ and $q_{2r+1}(t,\lambda )$ depend on $\lambda $ . We will see that there holds

(2.10) $$ \begin{align} q_{2r}(t,\lambda)=1+{\operatorname{sign}} v\,P_{2r}^{(\lambda-1/2)}(t),\quad q_{2r+1}(t,\lambda)=1+P_{2r+1}^{(\lambda-1/2)}(t), \end{align} $$

where $\{P_n^{(\alpha )}(t)\}_{n=0}^{\infty }$ is the system of Gegenbauer (ultraspherical) polynomials, that is, the family of polynomials orthogonal on $[-1,1]$ with respect to the weight function $(1-t^2)^{\alpha }$ , $\alpha>-1$ , normalized by $P_n^{(\alpha )}(1)=1$ . Note that, $P_n^{(-1/2)}(t)=\cos {}(n\arccos t)$ are the Chebyshev polynomials. For $\lambda>-1/2$ , we have

(2.11) $$ \begin{align} C_n^{\lambda}(t)=\frac{\Gamma(2\lambda+n)}{n!\,\Gamma(2\lambda)}\,P_n^{(\lambda-1/2)}(t),\quad \lambda\neq 0,\quad\mbox{and}\quad \lim_{\lambda\to 0}\frac{1}{\lambda}\,C_n^{\lambda}(t)=\frac{2}{n}\,P_n^{(-1/2)}(t), \end{align} $$

where $C_n^{\lambda }(t)$ are the Gegenbauer polynomials given in [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, Chapter X, 10.9].

Lemma 2.6. For $\lambda>-1/2$ and $r\ge 0$ , there hold

$$ \begin{align*} P_{2r}^{(\lambda-1/2)}(t)=\sum_{s=0}^{r}(-1)^{s} \binom{r}{s}\frac{(\lambda+r)_{s}}{(\lambda+1/2)_{s}}\,(1-t^2)^s \end{align*} $$

and

$$ \begin{align*} P_{2r+1}^{(\lambda-1/2)}(t)=t\sum_{s=0}^{r}(-1)^{s} \binom{r}{s}\frac{(\lambda+r+1)_{s}}{(\lambda+1/2)_{s}}\,(1-t^2)^s. \end{align*} $$

Therefore, (2.10) holds.

Proof. Since $ \frac {(-r)_s}{s!}=(-1)^{s}\binom {r}{s}, $ taking into account (2.11), [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, Chapter X, 10.9(21)], we get

$$ \begin{align*} \sum_{s=0}^{r}(-1)^{s}\binom{r}{s}&\frac{(\lambda+r)_{s}}{(\lambda+1/2)_{s}}\,(1-t^2)^s =\sum_{s=0}^{r}\frac{(-r)_s(\lambda+r)_{s}}{s!\,(\lambda+1/2)_{s}}\,(1-t^2)^s\\ &\qquad\quad ={}_{2}F_1(-r,\lambda+r;\lambda+1/2;1-t^2)=\frac{(2r)!\,\Gamma(2\lambda+1)} {2\Gamma(2\lambda+2r)}\,\frac{1}{\lambda}\,C_{2r}^{\lambda}(t)=P_{2r}^{(\lambda-1/2)}(t). \end{align*} $$

Similarly, using (2.11), [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, Chapter X, 10.9(22)], we arrive at

$$ \begin{align*} t\sum_{s=0}^{r}(-1)^{s}&\binom{r}{s}\frac{(\lambda+r+1)_{s}}{(\lambda+1/2)_{s}}\,(1-t^2)^s =t\sum_{s=0}^{r}\frac{(-r)_s(\lambda+r+1)_{s}}{s!\,(\lambda+1/2)_{s}}\,(1-t^2)^s\\ &=t\,{}_{2}F_1(-r,\lambda+r+1;\lambda+1/2;1-t^2)=\frac{(2r+1)!\,\Gamma(2\lambda+1)} {2\Gamma(2\lambda+2r+1)}\,\frac{1}{\lambda}\,C_{2r+1}^{\lambda}(t)=P_{2r+1}^{(\lambda-1/2)}(t). \end{align*} $$

We complete the proof of Theorem 1.1 noting that $|P_{n}^{(\lambda -1/2)}(t)|\le 1$ for $t\in [-1,1]$ under the condition $\lambda \ge 0$ or, equivalently, $\kappa \ge 1/2$ (see [Reference Szegö23, Chapter VII, 7.32.2]). Then in light of Remark 3.2 below and Lemma 2.6, the polynomials $q_{2r}(t,\lambda )$ and $q_{2r+1}(t,\lambda )$ are nonnegative on $[-1,1]$ and therefore Lemmas 2.4 and 2.5, together with (2.4) and (2.5), yield the statement of Theorem 1.1. To illustrate this, for $a=\frac {2}{2r+1}$ , $v=(2r+1)x^{\frac {1}{2r+1}}$ , we have

$$\begin{align*}|b_{\kappa,a}(x)|=|e_{2r+1}(v,\lambda)|\le e_{2r+1}(0,\lambda)=1.\\[-34pt] \end{align*}$$

The following integral representations of the kernel $B_{\kappa ,a}(x,y)$ follow from the lemmas above.

Corollary 2.7. If $\lambda =(2\kappa -1)/a$ and $\kappa>\frac {1}{2}-\frac {a}{4}$ , then for $x\in \mathbb {R}$

$$\begin{align*}B_{\kappa,a}(x,y)=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}(1+P_{2r+1}^{(\lambda-1/2)}(t))e^{-i(2r+1)(xy)^{\frac{1}{2r+1}}{}t}\,dt,\quad a=\frac{2}{2r+1},\quad r\in\mathbb{Z}_+, \end{align*}$$

and

$$\begin{align*}B_{\kappa,a}(x,y)=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}(1+{\operatorname{sign}}{}\ (xy)\,P_{2r}^{(\lambda-1/2)}(t))e^{-i(2r|xy|^{\frac{1}{2r}}{\operatorname{sign}}{}\ (xy))t}\,dt,\quad a=\frac{1}{r},\quad r\in\mathbb{N}. \end{align*}$$

Remark 2.8. The representations of $B_{\kappa ,a}(x,y)$ given in Corollary 2.7 can be also obtained from

$$\begin{align*}n!\int_{0}^{\pi}e^{iz\cos\theta}C_{n}^{\lambda}(\cos\theta)(\sin\theta)^{2\lambda}\,d\theta= 2^{\lambda}\sqrt{\pi}\,\Gamma(\lambda+1/2)(2\lambda)_n\,i^nz^{-\lambda}J_{\lambda+n}(z) \end{align*}$$

(see [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, Chapter X, 10.9(38)]) without applying Lemmas 2.2 and 2.3. These lemmas are of independent interest.

Remark 2.9. The positiveness of the polynomials $q_{2r}(t,\lambda )$ , $q_{2r+1}(t,\lambda )$ is sufficient, but not necessary, for the estimate $|B_{\kappa ,a}(x,y)|\le 1$ to hold. In [Reference Gorbachev, Ivanov and Tikhonov15], it was mentioned that for $a=1$ , this estimate holds also for $1/4<\kappa _0\le \kappa <1/2$ , $\kappa _{0}\approx 0.44$ .

Note that the case $\kappa <1/2$ corresponds to $\lambda <0$ , which we study in detail in the next section.

3 The case $d=1$ and $\lambda _{\kappa ,a}<0$

Let us investigate whether the polynomials $q_{2r}(t,\lambda )$ and $q_{2r+1}(t,\lambda )$ are nonnegative for $\lambda =\lambda _{\kappa ,a}=(2\kappa -1)/a<0$ . In order to do this, we decompose them by the polynomials $q_{2r}(t,0)$ and $q_{2r+1}(t,0)$ . First, we decompose the Gegenbauer polynomials in terms of the Chebyshev polynomials $P_n^{(-1/2)}(\cos t)=\cos nt$ using the well-known result (see [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, 10.9(17)])

(3.1) $$ \begin{align} C_{n}^{\lambda}(\cos\theta)=\sum_{s=0}^{n}\frac{(\lambda)_s(\lambda)_{n-s}}{s!\,(n-s)!}\cos{}(n-2s)\theta. \end{align} $$

Let us start with the case $R=2r+1$ .

Lemma 3.1. For any $r\in \mathbb {Z}_+$ and $\lambda>-1/2$ , there holds

(3.2) $$ \begin{align} P_{2r+1}^{(\lambda-1/2)}(t)=\sum_{s=0}^{r}b_s^r(\lambda)P_{2r+1-2s}^{(-1/2)}(t), \end{align} $$

where

(3.3) $$ \begin{align} b_s^r(\lambda)=\frac{(2r+2-s)_{s}(\lambda)_{s}(\lambda+r+1)_{r-s}}{4^r(\lambda+1/2)_{r}s!},\quad s=0,1,\dots,r. \end{align} $$

Proof. Taking into account (2.11) and (3.1), we obtain

$$\begin{align*}\frac{\Gamma(2\lambda+2r+1)}{(2r+1)!\,\Gamma(2\lambda)}\,P_{2r+1}^{(\lambda-1/2)}(t)= C_{2r+1}^{\lambda}(t)=2\sum_{s=0}^{r}\frac{(\lambda)_s(\lambda)_{2r+1-s}}{s!\,(2r+1-s)!}\,P_{2r+1-2s}^{(-1/2)}(t). \end{align*}$$

Hence, the duplication formula for the gamma function implies (3.3).

Remark 3.2. We see that in the decomposition (3.2), the zero coefficient is positive, and all other coefficients are also positive for $\lambda>0$ and negative for $\lambda <0$ . Note that the normalization of the Gegenbauer polynomials $P_{n}^{(\alpha )}(1) =1$ implies the equality

$$\begin{align*}\sum_{s=0}^{r}b_s^r(\lambda)=1 ,\end{align*}$$

and for $\lambda>0$ , the Gegenbauer polynomials $P_{2r+1}^{(\lambda -1/2)}(t)$ are the convex hull of the Chebyshev polynomials $P_{2r+1-2s}^{(-1/2)}(t)$ , $s=0,1,\dots ,r$ . In particular, one has $|P_{2r+1}^{(\lambda -1/2)}(t)|\le 1$ , $t\in [-1,1]$ , $\lambda \ge 0$ .

Thus, we are in a position to state the required result on the decomposition for the polynomial $q_{2r+1}(t,\lambda )$ .

Corollary 3.3. For any $r\in \mathbb {Z}_+$ and $\lambda>-1/2$ , there holds

$$\begin{align*}q_{2r+1}(t,\lambda)=\sum_{s=0}^{r}b_s^r(\lambda)q_{2r-2s+1}(t,0). \end{align*}$$

Using (2.10) and the properties of the coefficients $b_s^r(\lambda )$ , $s=0,1,\dots ,r$ , from Remark 3.2, we establish the following result.

Corollary 3.4. For any $r\in \mathbb {N}$ and $\lambda \in (-1/2,0)$ , the polynomial $q_{2r+1}(t,\lambda )$ is negative at the points of local minimum of the Chebyshev polynomial $P_{2r+1}^{(-1/2)}(t)= \cos {}(({2r+1})\arccos t)$ .

In the case $R=2r$ , similarly, one can obtain the decomposition of the polynomial $q_{2r}(t,\lambda )$ .

Lemma 3.5. For any $r\in \mathbb {N}$ and $\lambda>-1/2$ , there holds

$$\begin{align*}P_{2r}^{(\lambda-1/2)}(t)=\sum_{s=0}^{r}d_s^r(\lambda)P_{2r-2s}^{(-1/2)}(t), \end{align*}$$

where

$$\begin{align*}d_s^r(\lambda)=\frac{2(2r+1-s)_{s}(\lambda)_{s}(\lambda+r)_{r-s}}{4^r(\lambda+1/2)_{r}s!},\quad s=0,1,\dots,r-1,\quad d_r^r(\lambda)=\frac{(r+1)_{r}(\lambda)_{r}}{4^r(\lambda+1/2)_{r}r!}. \end{align*}$$

Proof. In light of (2.11) and (3.1), we have

$$\begin{align*}\frac{\Gamma(2\lambda+2r)}{(2r)!\,\Gamma(2\lambda)}\,P_{2r}^{(\lambda-1/2)}(t)= C_ {2r}^{\lambda}(t)=2\sum_{s=0}^{{r-1}}\frac{(\lambda)_s(\lambda)_{2r-s}}{s!\,(2r-s)!}\,P_{2r-2s}^{(-1/2)}(t)+\frac{((\lambda)_r)^2}{(r!)^2}\,P_{0}^{(-1/2)}(t). \end{align*}$$

Then, using the duplication formula for gamma function, we obtain for $s=0,1,\dots ,r-1$

$$\begin{align*}d_s^r(\lambda)= \frac{2(2r)!\,(\lambda)_s\Gamma(2\lambda)\Gamma(\lambda+2r-s)}{(2r-s)!\,s!\,\Gamma(\lambda)\Gamma(2\lambda+2r)} =\frac{2(2r+1-s)_{s}(\lambda)_{s}(\lambda+r)_{r-s}}{4^r(\lambda+1/2)_{r}s!}, \end{align*}$$

while for $s=r$ ,

$$\begin{align*}d_r^r(\lambda)=\frac{(2r)!\,(\lambda)_r\Gamma(2\lambda)\Gamma(\lambda+r)}{(r!)^{2}\Gamma(\lambda)\Gamma(2\lambda+2r)} =\frac{(r+1)_{r}(\lambda)_{r}}{4^r(\lambda+1/2)_{r}r!}.\\[-34pt] \end{align*}$$

Corollary 3.6. For any $r\in \mathbb {N}$ and $\lambda>-1/2$ , the following decomposition holds

$$\begin{align*}q_{2r}(t,\lambda)=\sum_{s=0}^{r}d_s^r(\lambda)q_{2r-2s}(t,0). \end{align*}$$

Since $ d_0^r(\lambda )>0$ and ${\operatorname{sign}} d_s^r(\lambda )={\operatorname{sign}} \lambda $ , $s=1,\dots ,r$ , for $\lambda>-1/2$ , taking into account (2.10), we arrive at the following result.

Corollary 3.7. If $r\in \mathbb {N}$ and $\lambda \in (-1/2,0)$ , then the polynomial $q_{2r}(t,\lambda )$ is negative at the points of local extremum of the Chebyshev polynomial $P_{2r}^{(-1/2)}(t)=\cos {}(2r\arccos t)$ .

Corollaries 3.4 and 3.7 show that the polynomials $q_{2r}(t,\lambda )$ and $q_{2r+1}(t,\lambda )$ for $r\ge 1$ change sign and, therefore, the method of the proof of the estimate $|B_{\kappa ,a}(x,y)|\le 1$ used in Theorem 1.1 cannot be applied when $\lambda <0$ . In this case, the problem remains open.

3.1 On positive definiteness

Recall that a continuous function f on $\mathbb {R}$ is positive definite if for any $\{v_s\}_{s=1}^{n}\subset \mathbb {R}$ and $\{z_s\}_{s=1}^{n}\subset \mathbb {C}$ there holds

$$\begin{align*}\sum_{s,l=1}^{n}z_s\overline{z_{l}}f(v_s-v_l)\ge 0. \end{align*}$$

Bochner’s theorem states that any continuous positive definite function $f(v)$ , $f(0)=1$ , is the Fourier transform of a probability measure.

Recall that the function $e_{2r+1}(v,\lambda )$ is given in (2.4). By (2.9) and (2.10),

$$\begin{align*}e_{2r+1}(v,\lambda)=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}(1+P_{2r+1}^{(\lambda-1/2)}(t))e^{-ivt}\,dt. \end{align*}$$

Then, by Remark 3.2 and Corollaries 3.3 and 3.4, we deduce the following

Corollary 3.8. The function $e_{2r+1}(v,\lambda )$ is positive definite if and only if $\lambda \ge 0$ .

4 The kernel of the generalized Fourier transform in the multivariate case

4.1 When the kernel is bounded by one

In this section, for the sake of completeness, we give the proof of Theorem 1.1 in the general case, following ideas from the paper [Reference Constales, De Bie and Lian10]. We stress that the proof below also allows one to deal with the case $d=1$ .

Let $d\in \mathbb {N}$ , $a>0$ , $\langle \kappa \rangle \ge 0$ , and $\eta =\lambda _{\kappa }=\langle \kappa \rangle +\frac {d-2}{2}>0$ . For $w\ge 0$ and $\tau \in [-1,1]$ , define

(4.1) $$ \begin{align} \Psi_{a}^d(w,\tau)=2^{2\eta/a}\Gamma\Bigl(\frac{2\eta+a}{a}\Bigr)\sum_{j=0}^{\infty}e^{-\frac{i\pi j}{a}}\,\frac{\eta+j}{\eta}\,w^{-2\eta/a}J_{\frac{2(\eta+j)}{a}}(w)C_{j}^{\eta}(\tau). \end{align} $$

Putting in (4.1)

(4.2) $$ \begin{align} x,y\in\mathbb{R}^d,\quad x=|x|x',\quad y=|y|y',\quad x',y'\in\mathbb{S}^{d-1},\quad w=\frac{2}{a}\,(|x||y|)^{a/2},\quad \tau=\langle x',\,y'\rangle, \end{align} $$

we obtain the function

$$ \begin{align*} K_{a}^d(x,y)=a^{2\eta/a}\Gamma\Bigl(\frac{2\eta+a}{a}\Bigr)(|x||y|)^{-\eta}\sum_{j=0}^{\infty}e^{-\frac{i\pi j}{a}}\,\frac{\eta+j}{\eta}\,J_{\frac{2(\eta+j)}{a}}\Bigl(\frac{2}{a}(|x||y|)^{a/2}\Bigr)C_{j}^{\eta}(\langle x',\,y'\rangle). \end{align*} $$

Note that $K_{a}^d(x,y)=K_{a}^d(y,x)$ and $\Psi _{a}^d(0,\tau )=K_{a}^d(0,y)=1$ .

Let $V_{\kappa }f(x)$ be the intertwining operator in the Dunkl harmonic analysis [Reference Rösler20], which is a positive operator satisfying

$$\begin{align*}V_{\kappa}f(x)=\int_{\mathbb{R}^d}f(\xi)\,d\mu_{x}^{\kappa}(\xi). \end{align*}$$

The representing measures $\mu _{x}^{\kappa }(\xi )$ are compactly supported probability measures with $\text {supp}\,\mu _{x}^{\kappa }(\xi )\subset \text {co}{}\{gx\colon g\in G\}$ [Reference Rösler22].

The kernel of the generalized Fourier transform $\mathcal {F}_{\kappa ,a}$ is given by [Reference Ben Saïd, Kobayashi and Ørsted4, Chapters 4, 5]

(4.3) $$ \begin{align} &B_{\kappa,a}(x,y)=V_{\kappa}K_{a}^d(x,|y|\cdot)(y') \notag\\ &\qquad\quad=a^{2\eta/a}\Gamma\Bigl(\frac{2\eta+a}{a}\Bigr)(|x||y|)^{-\eta}\sum_{j=0}^{\infty}e^{-\frac{i\pi j}{a}}\,\frac{\eta+j}{\eta}\,J_{\frac{2(\eta+j)}{a}}\Bigl(\frac{2}{a}(|x||y|)^{a/2}\Bigr)V_{\kappa}C_{j}^{\eta}(\langle x',{\cdot}\,\rangle)(y'). \end{align} $$

If $\langle \kappa \rangle =0$ , that is, $\eta =(d-2)/2$ , then $V_{\kappa }=\text {I}$ is the identity operator and $K_{a}^d(x,y)=B_{0,a}(x,y)$ .

Since the operator $V_{\kappa }$ is positive and $V_{\kappa }1=1$ , the proof of equality (1.2) can proceed as follows: If $|\Psi _{a}^d(w,\tau )|\le 1$ , then $|B_{\kappa ,a}(x,y)|\le 1$ .

Let further $\frac {2}{a}\in \mathbb {N}$ , $a=\frac {2}{R}$ . The inequality $|\Psi _{a}^d(w,\tau )|\le 1$ was proved in [Reference Constales, De Bie and Lian10] under the conditions $\eta>0$ and $\eta =(d-2)/2$ . We note that the proof of this estimate in [Reference Constales, De Bie and Lian10] also holds for $\eta =\langle \kappa \rangle +(d-2)/2>0$ . Hence, (1.2) is valid for $\langle \kappa \rangle +(d-2)/2>0$ . If $\eta =0$ , then $d=1$ , $\kappa =1/2$ or $d=2$ , $\kappa \equiv 0$ , and in these cases, (1.2) holds as well (see Theorem 1.1 and [Reference De Bie5]). This completes the proof of Theorem 1.1.

4.2 Positive definiteness of $\Psi _{a}^d$ . Another proof of Theorem 1.1

In [Reference Constales, De Bie and Lian10], to evaluate the kernel $B_{0,a}(x,y)$ , the authors used the Laplace transform of the function $\Psi _{a}^d$ given by (4.1). In the general case $d\ge 1$ , we give another proof of the estimate $|\Psi _{a}^d|\le 1$ based on the approach from the papers [Reference De Bie and Lian6, Reference Deleaval and Demni11], which allows one to show positive definiteness of the function $\Psi _{a}^d(w,\tau )$ with respect to w. Note also that for $a=1,2$ , the functions

$$ \begin{align*} \Psi_{1}^d(w,\tau)=j_{\lambda_{\kappa}-1/2}(w\sqrt{(1+\tau)/2}\,), \quad \Psi_{2}^d(w,\tau)=e^{-iw\tau}, \end{align*} $$

are positive definite (see [Reference Ben Saïd, Kobayashi and Ørsted4, Example 4.18], [Reference Constales, De Bie and Lian10]).

Let

$$\begin{align*}I_{\eta}(b)=\Bigl(\frac{b}{2}\Bigr)^{\eta}\,\sum_{j=0}^{\infty}\frac{1}{j!\,\Gamma(j+\eta+1)}\Bigl(\frac{b^2}{4}\Bigr)^j \end{align*}$$

be the modified Bessel function of the first kind (see [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, Chapter 7, 7.2.2]) and

$$\begin{align*}\Phi_2^{(m)}(\beta_1,\dots,\beta_m;\gamma;x_1,\dots,x_m)= \sum_{j_1,\dots,j_m\ge 0}\frac{(\beta_1)_{j_1}\cdots (\beta_m)_{j_m}}{(\gamma)_{j_1+\cdots+j_m}}\,\frac{x_1^{j_1}}{j_{1}!}\cdots\frac{x_m^{j_m}}{j_{m}!} \end{align*}$$

be the second Humbert function of m variables (see [Reference Exton14, Chapter 2, 2.1.1.2]). In the case when $\gamma -\sum _{j=1}^m\beta _j>0$ and $\beta _j>0$ , $j=1,\dots ,m$ , the hypergeometric function $\Phi _2^{(m)}$ admits the following integral representation

(4.4) $$ \begin{align} \Phi_2^{(m)}(\beta_1,\dots,\beta_m;\gamma;x_1,\dots,x_m) =C_{\beta}^{(\gamma)}\int_{T^m}e^{\sum_{j=1}^mx_jt_j}\Bigl(1-\sum_{j=1}^{m}t_j\Bigr)^{\gamma-\sum_{j=1}^m\beta_j-1} \prod_{j=1}^{m}t_j^{\beta_j-1}\,dt_1\cdots dt_m, \end{align} $$

where

$$\begin{align*}C_{\beta}^{(\gamma)}=\frac{\Gamma(\gamma)}{\Gamma(\gamma-\sum_{j=1}^{m}\beta_j)\prod_{j=1}^{m}\Gamma(\beta_j)} \end{align*}$$

and

$$\begin{align*}T^m=\Bigl\{(t_1,\dots,t_m)\colon t_j\ge 0,\ j=1,\dots,m,\ \sum_{j=1}^{m}t_j\le 1\Bigr\} \end{align*}$$

is the unit simplex in $\mathbb {R}^m$ ([Reference Chamayou and Wesolowski9, Reference Humbert17]).

Taking into account the results from [Reference De Bie and Lian6, Reference Deleaval and Demni11], we obtain the following proposition, which gives an alternative proof of Theorem 1.1.

Proposition 4.1. Let $d\in \mathbb {N}$ , $a=\frac {2}{R}$ , $R\in \mathbb {N}$ , $R\ge 2$ , $\eta =\langle \kappa \rangle +\frac {d-2}{2}>0$ , $\tau \in [-1,1]$ , and $q=\arccos \tau $ . The function $w \mapsto \Psi _{a}^d(w,\tau )$ is a positive definite entire function of exponential type

(4.5) $$ \begin{align} \theta(a,\tau)= \begin{cases} \cos\frac{q}{R}, & R=2r\ \text{and}\ 0\le q\le\pi \ \text{or}\ R=2r+1\ \text{and}\ 0\le q\le\pi/2,\\ \cos\frac{\pi-q}{R}, & R=2r+1\ \text{and}\ \pi/2\le q\le\pi. \end{cases} \end{align} $$

Proof. Let us consider the function

$$\begin{align*}f_{R,\eta}(b,\tau)=\Gamma(R\eta+1)\Bigl(\frac{b}{2}\Bigr)^{\eta}\sum_{j=0}^{\infty} \frac{j+\eta}{\eta}\,I_{R(j+\eta)}(b)C_{j}^{\eta}(\tau), \quad R\in\mathbb{N}. \end{align*}$$

Let $\tau =\langle x',\,y'\rangle $ , $b=-iw$ , $R=\frac {2}{a}$ , $R\ge 2$ . Since $I_{\eta }(-iw)=e^{-i\frac {\pi \eta }{2}}J_{\eta }(w)$ [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, Chapter 7, 7.2.2], then $f_{R,\eta }(b,\tau )=\Psi _{a}^d(w,\tau )$ . It is known (see [Reference De Bie and Lian6, Reference Deleaval and Demni11]) that

$$\begin{align*}f_{R,\eta}(b,\tau)=b_0\Phi_{2}^{(R-1)}(\eta,\dots,\eta;R\eta;b_1-b_0,\dots,b_{R-1}-b_0), \end{align*}$$

where $b_j=b\cos {}(\frac {q-2\pi j}{R})$ , $j=0,\dots ,R-1$ . In light of (4.4) and using $\eta>0$ , we get

(4.6) $$ \begin{align} &\Psi_{a}^d(w,\tau)=f_{R,\eta}(-iw,\tau)\notag\\ &\qquad=C_{\eta}\int_{T^{R-1}}e^{-iw\{\cos{}(\frac{q}{R})+\sum_{j=1}^{R-1}(\cos{}(\frac{q-2\pi j}{R})-\cos{}(\frac{q}{R}))t_j\}}\Bigl(1-\sum_{j=1}^{R-1}t_j\Bigr)^{\eta-1}\prod_{j=1}^{R-1}t_j^{\eta-1}\, \,dt_1\cdots dt_{R-1}, \end{align} $$

where $C_{\eta }=\Gamma (R\eta )/(\Gamma (\eta ))^R$ . For $\alpha>0$ and $\beta \in \mathbb {R}$ , the function $\alpha e^{i\beta w}$ is positive definite, hence the function $\Psi _{a}^d(w,\tau )$ is also positive definite with respect to w.

The representation (4.6) implies that $\Psi _a^d(w,\tau )$ is an entire function of exponential type with respect to w. Let us calculate its type. Since for $j=1,\dots ,R-1$ , $q=\arccos \tau $ , $\tau \in [-1,1]$ , $ \cos \frac {q}{R}-\cos \frac {q-2\pi j}{R}=2\sin \frac {\pi i}{R}\,\sin \frac {\pi j-q}{R}\ge 0, $ then for any $(t_1,\dots ,t_{R-1})\in T^{R-1}$ ,

$$\begin{align*}\min_{1\le j\le R-1}\cos\frac{q-2\pi j}{R}\le \cos\frac{q}{R}+\sum_{j=1}^{R-1}\Bigl(\cos\frac{q-2\pi j}{R}-\cos\frac{q}{R}\Bigr)t_j\le \cos\frac{q}{R}. \end{align*}$$

Since

$$ \begin{align*} \min_{1\le j\le R-1}\cos\frac{q-2\pi j}{R}= \begin{cases}-\cos\frac{q}{R}, & R=2r\ \text{or}\ R=2r+1\ \text{and}\ 0\le q\le\pi/2,\\ -\cos\frac{\pi-q}{R}, & R=2r+1\ \text{and}\ \pi/2\le q\le\pi, \end{cases} \end{align*} $$

the type of the function $\Psi _{a}^d(w,\tau )$ is given by (4.5).

Remark 4.2. Let $t,\tau \in (-1,1)$ , $\lambda =2\eta /a$ , $a=1/r$ , $r\in \mathbb {N}$ . Then the function $\Psi _{a}^{d}(w,\tau )$ has the following integral representation

(4.7) $$ \begin{align} \Psi_{a}^{d}(w,\tau)=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}\sum_{j=0}^{\infty}\frac{\eta+j}{\eta}\,P_{2j/a}^{(\lambda-1/2)}(t) C_{j}^{\eta}(\tau)e^{-iwt}\,dt. \end{align} $$

Indeed, using (2.8) and (4.1) and Lemmas 2.2 and 2.6, we get

$$ \begin{align*} &\Psi_{a}^d(w,\tau)=\sum_{j=0}^{\infty}(-1)^{rj}\,\frac{\eta+j}{\eta}\,\frac{w^{2rj}}{2^{2rj}(\lambda+1)_{2rj}}\,j_{\lambda+2rj}(w) C_{j}^{\eta}(\tau)\\ &=\sum_{j=0}^{\infty}\frac{\eta+j}{\eta}\Bigl\{j_{\lambda}(w)+\sum_{s=1}^{rj-1}(-1)^s\binom{rj}{s}\frac{(\lambda+rj)_s}{(\lambda+1)_s}\,j_{\lambda+s}(w) {+}(-1)^{rj}\,\frac{(\lambda+rj+1)_{rj-1}}{(\lambda+1)_{rj-1}}\,j_{\lambda+rj}(w)\Bigr\}C_{j}^{\eta}(\tau)\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad=c_{\lambda}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}\sum_{j=0}^{\infty}\frac{\eta+j}{\eta}\,P_{2rj}^{(\lambda-1/2)}(t) C_{j}^{\eta}(\tau)e^{-iwt}\,dt. \end{align*} $$

The integral representation (4.7) implies that

$$ \begin{align*} \psi_a^d(t,\tau)=\sum_{j=0}^{\infty}\frac{\eta+j}{\eta}\,P_{2rj}^{(\lambda-1/2)}(t) C_{j}^{\eta}(\tau)\ge 0,\quad t,\,\tau\in (-1,1), \end{align*} $$

is equivalent to the positive definiteness of the function $\Psi _{a}^d(w,\tau )$ . Proposition 4.1 shows that the function $\psi _{a}^d(t,\tau )$ is positive and its support as a function of t lies on the interval $I_a=[-\theta (a,\tau ),\theta (a,\tau )]$ .

In the case $a=1$ , one can easily write an explicit formula for $\psi _{1}^d(t,\tau )$ . In more detail, we have $r=1$ , $\lambda =2\eta $ , and $\theta (a,\tau )=\cos \frac {q}{2}=\sqrt {(1+\tau )/2}$ . Then using [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi2, Chapter VIII, 8.7(31)], we obtain

$$ \begin{align*} \psi_{1}^d(t,\tau)&=(2\pi c_{\lambda})^{-1}(1-t^2)^{1/2-2\eta}\int_{-\infty}^{\infty}j_{\eta-1/2}(w\sqrt{(1+\tau)/2})e^{iwt}\,dw\\ &=c(\eta)(1-t^2)^{1/2-2\eta}(1+\tau)^{1/2-\eta}(1+\tau-2t^2)^{\eta-1}\chi_{I_1}(t)\\ &=c(\eta)(1-t^2)^{1/2-2\eta}(1+\tau)^{1/2-\eta}(\tau-P_2^{(-1/2)}(t))^{\eta-1}\chi_{I_1}(t), \end{align*} $$

where $\chi _{I_1}(t)$ is the characteristic function of the interval $I_1$ and $P_2^{(-1/2)}(t)= \cos {}(2\arccos t)$ .

4.3 Parameters when the kernel is not bounded by one

We will show that in some cases the uniform norm of the kernel $B_{\kappa ,a}(x,y)$ is not bounded by $1$ and can be either finite or infinite. We mention that the conditions when the norm is not finite are known only in the one-dimensional case. In particular, if $d=1$ , $0<a<2$ , and $\kappa <\frac {1}{2}-\frac {a}{4}$ , then $\|B_{\kappa ,a}\|_{\infty }=\infty $ (see Proposition 2.1).

Proof of Theorem 1.2.

Let first $d=1$ and $\lambda =\lambda _{\kappa ,a}=\frac {2\kappa -1}{a}$ . Changing variables $x=(\frac {a}{2}|v|)^{2/a}{\operatorname{sign}}\ {}v$ in (2.3), we get

(4.8) $$ \begin{align} e_{\kappa,a}(v,\lambda)=b_{\kappa,a}\Bigl(\Bigl(\frac{a}{2}\,|v|\Bigr)^{2/a}{\operatorname{sign}}{}\ v\Bigr) =j_{\lambda}(v)&+\frac{\Gamma(\lambda+1)\cos\frac{\pi}{a}}{2^{2/a}\Gamma(\lambda+1+2/a)}\, |v|^{2/a}j_{\lambda+\frac{2}{a}}(v){\operatorname{sign}}{}\ v\notag\\ &{}-i\,\frac{\Gamma(\lambda+1)\sin\frac{\pi}{a}}{2^{2/a}\Gamma(\lambda+1+2/a)}\, |v|^{2/a}j_{\lambda+\frac{2}{a}}(v){\operatorname{sign}}{}\ v. \end{align} $$

If $0<a\le 1$ and $\kappa =\frac {1}{2}-\frac {a}{4}$ , then $\lambda =-1/2$ , $j_{-1/2}(v)=\cos v$ , $j_{1/2}(v)=\frac {\sin v}{v}$ .

Assume first that $\cos \frac {\pi }{a}=0$ or $a=\frac {2}{2r+1}$ , $r\in \mathbb {N}$ . Since $\sin \frac {\pi }{a}\neq 0$ , $j_{-1/2}(2\pi )=1$ , ${j_{1/2}(2\pi )=0}$ , and $j_{2r+1/2}(2\pi )\neq 0$ (see [Reference Watson25, Chapter 15, 15.28]), then from (4.8), we get $|e_{\kappa ,a}(2\pi ,-1/2)|=|e_{2r+1}(2\pi ,-1/2)|>1$ .

Let now $\cos {}(\pi /a)\neq 0$ . In light of (4.8), there holds

$$\begin{align*}|e_{\kappa,a}(v,\lambda)|\ge \Bigl|\cos v+\frac{\Gamma(\frac{1}{2})\cos\frac{\pi}{a}}{2^{2/a}\Gamma(\frac{2}{a}+\frac{1}{2})}\, |v|^{2/a}j_{\frac{2}{a}-\frac{1}{2}}(v){\operatorname{sign}}{}\ v\Bigr|. \end{align*}$$

Since

$$\begin{align*}(2\pi s)^{2/a}j_{\frac{2}{a}-\frac{1}{2}}(2\pi s)=\frac{2^{2/a}\Gamma(\frac{2}{a}+\frac{1}{2})}{\Gamma(\frac{1}{2})}\, \Bigl\{\cos\frac{\pi}{a}+O\Bigl(\frac{1}{s}\Bigr)\Bigr\} \end{align*}$$

as $s\to +\infty $ (see [Reference Watson25, Chapter VII, 7.1]), we deduce that for sufficiently large s

$$\begin{align*}|e_{\kappa,a}(2\pi s,\lambda)|\ge 1+\cos^2\frac{\pi}{a}+O\Bigl(\frac{1}{s}\Bigr)>1, \end{align*}$$

which completes the proof of (1.3). Note that similar arguments also imply the proof of (1.3) for $1<a<2$ and $\kappa =\frac {1}{2}-\frac {a}{4}$ .

Now, we consider the multivariate case. Suppose that $d\ge 1$ , $a\in (1,2)\cup (2,\infty )$ , $\langle \kappa \rangle \ge 0$ , and $\eta =\lambda _{\kappa }$ , $\lambda =\frac {2\eta }{a}=\frac {2\langle \kappa \rangle +d-2}{a}$ . Recall that $x'$ , $y'$ , w are defined in (4.2). In view of (2.2) and the proof for the case $d=1$ , we may assume that $\eta ,\lambda>-1/2$ and $\cos \frac {\pi }{a}\neq 0$ .

Note that (4.3) can be equivalently written as

$$\begin{align*}B_{\kappa,a}(x,y) =\sum_{j=0}^{\infty}e^{-\frac{i\pi j}{a}}\,\frac{\eta+j}{\eta}\,\frac{\Gamma(\lambda+1)}{2^{2j/a}\,\Gamma(\lambda+1+2j/a)}\,w^{2j/a} j_{\lambda+\frac{2j}{a}}(w)V_{\kappa}C_{j}^{\eta}(\langle x',{\cdot}\,\rangle)(y'). \end{align*}$$

For Gegenbauer polynomials (2.11), the following estimate holds

$$\begin{align*}\sup_{-1\le t\le 1}\Bigl|\frac{1}{\eta}\,C_j^{\eta}(t)\Bigr|\le \widetilde{c}(\eta)j^{2\eta-1},\quad j\ge 1, \end{align*}$$

[Reference Ben Saïd, Kobayashi and Ørsted4, Lemma 4.9]. Using (2.11), we get

$$\begin{align*}\Bigl|\frac{\eta+j}{\eta}\,V_{\kappa}C_j^{\eta}(\langle x',{\cdot}\,\rangle)(y')\Bigr|\le c(\eta)(j^{2\eta}+1). \end{align*}$$

Since $|j_{\lambda }(w)|\le 1$ , $\frac {1}{\eta }\,C_1^{\eta }(t)=2t$ , then for $0\le w\le 1$ , we have

$$ \begin{align*} \Bigl|\sum_{j=2}^{\infty}e^{-\frac{i\pi j}{a}}\,\frac{\eta+j}{\eta}\,\frac{\Gamma(\lambda+1)}{2^{2j/a}\Gamma(\lambda+1+2j/a)}\,w^{2j/a} &j_{\lambda+\frac{2j}{a}}(w)V_{\kappa}C_{j}^{\eta}(\langle x',{\cdot}\,\rangle)(y')\Bigr|\\ &{}\le c(\eta)w^{4/a}\sum_{j=2}^{\infty}\frac{\Gamma(\lambda+1)(j^{2\eta}+1)}{2^{2j/a}\Gamma(\lambda+1+2j/a)} \le c(a,\kappa)w^{4/a} \end{align*} $$

and

$$ \begin{align*} B_{\kappa,a}(x,y)=j_{\lambda}(w)&+\frac{2(\eta+1)\Gamma(\lambda+1)\cos\frac{\pi}{a}}{2^{2/a}\Gamma(\lambda+1+2/a)}\,w^{2/a} j_{\lambda+\frac{2}{a}}(w)V_{\kappa}(\langle x',{\cdot}\,\rangle)(y')\\ &{}-i\,\frac{2(\eta+1)\Gamma(\lambda+1)\sin\frac{\pi}{a}}{2^{2/a}\Gamma(\lambda+1+2/a)}\,w^{2/a} j_{\lambda+\frac{2}{a}}(w)V_{\kappa}(\langle x',{\cdot}\,\rangle)(y')+O(w^{4/a}),\quad w\to 0. \end{align*} $$

Therefore,

(4.9) $$ \begin{align} |B_{\kappa,a}(x,y)|\ge \Bigl|j_{\lambda}(w)+\frac{2(\eta+1)\Gamma(\lambda+1)\cos\frac{\pi}{a}}{2^{2/a}\Gamma(\lambda+1+2/a)}\,w^{2/a} j_{\lambda+\frac{2}{a}}(w)V_{\kappa}(\langle x',{\cdot}\,\rangle)(y')+O(w^{4/a})\Bigr|. \end{align} $$

Since the operator $V_{\kappa }$ is an isomorphism on the space of homogeneous polynomials of degree $1$ , then for suitable $x',y'\in \mathbb {S}^{d-1}$ , we deduce

(4.10) $$ \begin{align} V_{\kappa}(\langle x',{\cdot}\,\rangle)(y')=\int_{\mathbb{R}^d}\langle x',\xi\rangle\,d\mu_{y'}^{\kappa}(\xi)\neq 0,\quad {\operatorname{sign}} (V_{\kappa}(\langle x',{\cdot}\,\rangle)(y'))={\operatorname{sign}}{}\ (\cos{}(\pi/a)). \end{align} $$

We have

$$\begin{align*}j_{\lambda}(w)=1+O(w^2),\quad j_{\lambda+\frac{2}{a}}(w)=1+O(w^2) \end{align*}$$

as $w\to 0$ . This, (4.9) and (4.10) imply that, for given $x',y'\in \mathbb {S}^{d-1}$ as above and sufficiently small positive w,

$$\begin{align*}|B_{\kappa,a}(x,y)|\ge 1+\frac{2(\eta+1)\Gamma(\lambda+1)|\!\cos\frac{\pi}{a}|}{2^{2/a}\Gamma(\lambda+1+2/a)}\,w^{2/a} |V_{\kappa}(\langle x',{\cdot}\,\rangle)(y')|+O(w^{4/a})+O(w^2)>1.\\[-42pt] \end{align*}$$

Remark 4.3.

  1. (i) For $0<a\le 1$ , $\frac {2}{a}\in \mathbb {N}$ , equality (1.2) holds provided $\kappa \ge \kappa _0(a)$ , where $\frac {1}{2}-\frac {a}{4}<\kappa _0(a)\le 1/2$ . The problem of determining $\kappa _0(a)$ is open.

  2. (ii) Theorem 1.2 shows that our conjecture in [Reference Gorbachev, Ivanov and Tikhonov15] asserting that (1.3) holds under the condition $2\langle \kappa \rangle +d+a\ge 3$ is not valid for $d\ge 1$ and $a\in (1,2)\cup (2,\infty )$ .

5 Image of the Schwartz space

In this section, for the $(\kappa ,a)$ -generalized Fourier transform, we study the image of the Schwartz space $\mathcal {S}(\mathbb {R}^d)$ . Denote by $\mathcal {S}(\mathbb {R}_+)$ the subspace of even functions from $\mathcal {S}(\mathbb {R})$ .

5.1 The case of $a>0$

Let $d\nu _{\eta }(u)=b_{\eta }u^{2\eta +1}\,du$ , $b_{\eta }^{-1}=2^{\eta }\Gamma (\eta +1)$ , $d\nu _{\eta ,a}(u)=b_{\eta ,a}u^{2\eta +a-1}du$ , $b_{\eta ,a}^{-1}=a^{2\eta /a}\Gamma (2\eta /a+1)$ ,

$$\begin{align*}H_{\eta}(f_0)(v)=\int_{0}^{\infty}f_0(u)j_{\eta}(uv)\,d\nu_{\eta}(u),\quad f_{0}\in L^1(\mathbb{R}_+,d\nu_{\eta}),\quad \eta\ge-1/2, \end{align*}$$

be the Hankel transform and

$$\begin{align*}H_{\eta,a}(f_0)(v)=\int_{0}^{\infty}f_0(u)j_{\frac{2\eta}{a}}\Bigl(\frac{2}{a}\,(vu)^{a/2} \Bigr)\,d\nu_{\eta,a}(u),\quad 2\eta+a\ge 1, \end{align*}$$

the a-deformed Hankel transform (see [Reference Ben Saïd, Kobayashi and Ørsted4, Reference Gorbachev, Ivanov and Tikhonov15]).

Recall that $\lambda _{\kappa }=\langle \kappa \rangle +(d-2)/2$ , $\lambda =\lambda _{\kappa ,a}=2\lambda _{\kappa }/a$ .

Example 5.1. Consider $f(x)=e^{-|x|^2}\in \mathcal {S}(\mathbb {R}^d)$ , $f(x)=f_0(\rho )$ , $\rho =|x|$ . If $|y|=v$ , $\rho =(a/2)^{1/a}u^{2/a}$ , then (see [Reference Ben Saïd, Kobayashi and Ørsted4, Reference Gorbachev, Ivanov and Tikhonov15])

$$ \begin{align*} \mathcal{F}_{\kappa,a}(f)(y)=H_{\lambda_{\kappa},a}(f_0)(v)&= \int_{0}^{\infty}f_0(\rho)j_{\frac{2\lambda_{\kappa}}{a}}\Bigl(\frac{2}{a}\,(v\rho)^{a/2} \Bigr)\,d\nu_{\lambda_{\kappa},a}(\rho)\\ &= \int_{0}^{\infty}g_a(u)j_{\lambda}\Bigl(\Bigl(\frac{2}{a}\Bigr)^{1/2}\,v^{a/2}u\Bigr)\,d\nu_{\lambda}(u), \end{align*} $$

where

$$\begin{align*}g_a(u)=\exp \Bigl(-\Bigl(\frac{a}{2}\Bigr)^{2/a}u^{4/a}\Bigr),\quad u\ge 0. \end{align*}$$

Assuming that $\mathcal {F}_{\kappa ,a}(f)(y)$ is rapidly decreasing at infinity, the same is true for the function

$$\begin{align*}G_a(v)=H_{\lambda}(g_a)(v)=\int_{0}^{\infty}g_a(u) j_{\lambda}(uv)\,d\nu_{\lambda}(u). \end{align*}$$

Considering even extensions of the functions $g_a$ and $G_a$ , we get the following representation via the Dunkl transform:

$$\begin{align*}G_a(v)=\int_{-\infty}^{\infty}g_a(u)j_{\lambda}(uv)\,d\mu_{\kappa,2}(u)=\mathcal{F}_{\kappa,2}(g_a)(v),\quad v\in\mathbb{R}. \end{align*}$$

If $2/a$ is a noninteger, then either $4/a\not \in \mathbb {N}$ or $4/a=2s+1$ , $s\in \mathbb {Z}_+$ , and the function

$$\begin{align*}g_a(u)=\exp \Bigl(-\Bigl(\frac{a}{2}\Bigr)^{2/a}|u|^{4/a}\Bigr),\quad u\in \mathbb{R}, \end{align*}$$

has finite smoothness at the origin. On the other hand, since $g_a,G_a\in L^1(\mathbb {R},d\mu _{\kappa ,2})$ , we obtain

$$\begin{align*}g_a(u)=\mathcal{F}_{\kappa,2}(G_a)(u)=\int_{-\infty}^{\infty}G_a(v)j_{\lambda}(uv)\,d\mu_{\kappa,2}(v) ,\end{align*}$$

and the right-hand side is infinitely differentiable at the origin if $G_a(v)$ is fast decreasing. This contradiction shows that, for $2/a\not \in \mathbb {N}$ , $\mathcal {F}_{\kappa ,a}(f)(y)$ cannot rapidly decay at infinity and so it is not a Schwartz function.

Moreover, let us denote for convenience the derivative of f by $\partial f$ . Since

$$\begin{align*}(\partial_v^2j_{\lambda}(uv))\bigr|_{v=0}=-\frac{u^2}{2(\lambda+1)},\quad (\partial_v^4j_{\lambda}(uv))\bigr|_{v=0}=\frac{3u^4}{4(\lambda+1)(\lambda+2)}, \end{align*}$$

then

$$\begin{align*}G_a(v)=\xi_0+\xi_1v^2+O(v^4),\qquad v\to 0,\qquad \xi_1\neq 0, \end{align*}$$

and

$$\begin{align*}\mathcal{F}_{\kappa,a}(f)(y)=\xi_0+\widetilde{\xi}_1|y|^a+O(|y|^{2a}),\qquad y\to 0,\qquad \widetilde{\xi}_1\neq 0. \end{align*}$$

If a is not even, then $\mathcal {F}_{\kappa ,a}(f)(y)$ has finite smoothness at the origin.

The following statement follows from Example 5.1.

Proposition 5.2. Let $d\in \mathbb {N}$ .

  1. (i) The condition $\frac {a}{2}\in \mathbb {N}$ is necessary for the embedding $\mathcal {F}_{\kappa ,a}(\mathcal {S}(\mathbb {R}^d))\subset C^{\infty }(\mathbb {R}^d)$ to hold.

  2. (ii) The condition $\frac {2}{a}\in \mathbb {N}$ is necessary for the set $\mathcal {F}_{\kappa ,a}(\mathcal {S}(\mathbb {R}^d))$ to consist of rapidly decreasing functions at infinity.

Remark 5.3. We see that $\mathcal {F}_{\kappa ,a}(\mathcal {S}(\mathbb {R}^d))=\mathcal {S}(\mathbb {R}^d)$ only for $a=2$ . Hence, the claim in [Reference Johansen18, Lemma 2.12] that the Schwartz space is invariant under the generalized Fourier transform is false for $a\neq 2$ .

The conditions in Proposition 5.2 are also sufficient, at least in the one-dimensional case. Indeed, suppose $\lambda =(2\kappa -1)/a\ge -1/2$ and denote the even and odd parts of f, as usual, by

$$\begin{align*}f_{e}(x)=\frac{f(x)+f(-x)}{2},\quad f_{o}(x)=\frac{f(x)-f(-x)}{2}. \end{align*}$$

Then the one-dimensional generalized Fourier transform can be written as

(5.1) $$ \begin{align} \mathcal{F}_{\kappa,a}(f)(y)=2\int_{0}^{\infty}(B_{\kappa,a}(\,{\cdot}\,,y))_{e}(x) f_e(x)\,d\mu_{\kappa,a}(x)+ 2\int_{0}^{\infty}(B_{\kappa,a}(\,{\cdot}\,,y))_{o}(x) f_o(x)\,d\mu_{\kappa,a}(x). \end{align} $$

Putting in (5.1)

(5.2) $$ \begin{align} \begin{gathered} x=x(u)=\Bigl(\frac{a}{2}\Bigr)^{\frac{1}{a}}\,u^{\frac{2}{a}},\quad g(u)=Af(u)=f(x(u))=f\Bigl(\Bigl(\frac{a}{2}\Bigr)^{\frac{1}{a}}\,u^{\frac{2}{a}}\Bigr),\\ g_e(u)=Af_e(u),\quad g_o(u)=Af_o(u),\quad x,u\in\mathbb{R}_+, \end{gathered} \end{align} $$

and taking into account (2.1), we deduce

(5.3) $$ \begin{align} \mathcal{F}_{\kappa,a}(f)(y)&=\int_{0}^{\infty}Af_e(u)j_{\lambda} \Bigl(\Bigl(\frac{2}{a}\Bigr)^{1/2}|y|^{a/2}u\Bigr)\,d\nu_{\lambda}(u)\notag \\ &\quad \ {}+e^{-\frac{\pi i}{a}}\Bigl(\frac{2}{a}\Bigr)^{1/a}y\int_{0}^{\infty}u^{-2/a}Af_o(u) j_{\lambda+\frac{2}{a}}\Bigl(\Bigl(\frac{2}{a}\Bigr)^{1/2} |y|^{a/2}u\Bigr)\,d\nu_{\lambda+\frac{2}{a}}(u)\notag \\ &\qquad\qquad =H_{\lambda}(g_e)\Bigl(\Bigl(\frac{2}{a}\Bigr)^{1/2}|y|^{a/2}\Bigr)+ e^{-\frac{\pi i}{a}} \Bigl(\frac{2}{a}\Bigr)^{1/a} yH_{\lambda+\frac{2}{a}}(u^{-2/a}g_o)\Bigl(\Bigl(\frac{2}{a}\Bigr)^{1/2}|y|^{a/2}\Bigr). \end{align} $$

If $f\in \mathcal {S}(\mathbb {R})$ , then $g_e(u)$ , $u^{-2/a}g_o(u)$ decrease rapidly at infinity and there exist

$$\begin{align*}\lim_{u\to 0+0}u^{-2/a}g_o(u)=\Bigl(\frac{a}{2}\Bigr)^{\frac{1}{a}}f'(0). \end{align*}$$

Therefore, the representation (5.3) shows that, for even a, we have the embedding $\mathcal {F}_{\kappa ,a}(\mathcal {S}(\mathbb {R}))\subset C^{\infty }(\mathbb {R})$ .

Further, if $\frac {2}{a}\in \mathbb {N}$ , $f\in \mathcal {S}(\mathbb {R})$ , then the functions $g_e(u),u^{-2/a}g_o(u) \in \mathcal {S}(\mathbb {R}_+)$ as well as (see [Reference Rösler20]) the even functions

$$\begin{align*}H_{\lambda}(g_e)(v),\ H_{\lambda+\frac{2}{a}}(u^{-2/a}g_o)(v)\in \mathcal{S}(\mathbb{R}). \end{align*}$$

Hence, we prove the following

Proposition 5.4. Suppose $\frac {2}{a}\in \mathbb {N}$ , $\lambda \ge -1/2$ , $f\in \mathcal {S}(\mathbb {R})$ , then

(5.4) $$ \begin{align} \mathcal{F}_{\kappa,a}(f)(y)=F_1\bigl(|y|^{a/2}\bigr)+yF_2\bigl(|y|^{a/2}\bigr), \end{align} $$

where the even functions $F_1,F_2\in \mathcal {S}(\mathbb {R})$ .

Thus, for $2/a\in \mathbb {N}$ , the generalized Fourier transform decreases rapidly at infinity and we arrive at the following result.

Proposition 5.5.

  1. (i) The embedding $\mathcal {F}_{\kappa ,a}(\mathcal {S}(\mathbb {R}))\subset C^{\infty }(\mathbb {R})$ is valid if and only if $\frac {a}{2}\in \mathbb {N}$ .

  2. (ii) The set $\mathcal {F}_{\kappa ,a}(\mathcal {S}(\mathbb {R}))$ consists of rapidly decreasing functions at infinity if and only if $\frac {2}{a}\in \mathbb {N}$ .

Let

$$\begin{align*}\mathcal{S}_{org}(\mathbb{R})=\{f\in\mathcal{S}(\mathbb{R})\colon \partial^{k}f(0)=0,\ k\in\mathbb{N}\}. \end{align*}$$

The class $\mathcal {S}_{org}(\mathbb {R})$ is dense in $L^{2}(\mathbb {R},d\mu _{\kappa ,a})$ . Suppose $a>0$ , $f(x)\in \mathcal {S}_{org}(\mathbb {R})$ , then the functions $g_e(u),u^{-2/a}g_o(u)$ belong to $\mathcal {S}(\mathbb {R}_+)$ , cf. (5.3). Therefore, the even functions $H_{\lambda }(g_e)(v)$ , $H_{\lambda +\frac {2}{a}}(u^{-2/a}g_o)(v)$ also belong to $\mathcal {S}(\mathbb {R})$ .

Thus, we obtain the following

Proposition 5.6. Suppose $a>0$ , $\lambda \ge -1/2$ , and $f\in \mathcal {S}_{org}(\mathbb {R})$ ; then the generalized Fourier transform $\mathcal {F}_{\kappa ,a}(f)$ enjoys the representation (5.4).

5.2 The case of irrational a

We will show that if a is irrational, any nontrivial Schwartz function possesses similar properties to the Gaussian function in Example 5.1. To see this, we need auxiliary properties of the kernel of the generalized Fourier transform.

Let $\mathbb {S}^{d-1}$ be the unit sphere in $\mathbb {R}^{d}$ , $x=\rho x'$ , $\rho =|x|\in \mathbb {R}_+$ , $x'\in \mathbb {S}^{d-1}$ , and $dx'$ be the Lebesgue measure on the sphere. If $a_{\kappa }^{-1}=\int _{\mathbb {S}^{d-1}}v_{\kappa }(x')\,dx'$ , $d\sigma _{\kappa }(x')=a_{\kappa }v_{\kappa }(x')\,dx'$ , then $d\mu _{\kappa ,a}(x)=d\nu _{\lambda _{\kappa },a}(\rho )\,d\sigma _{\kappa }(x')$ and $c_{\kappa ,a}=b_{\kappa ,a}a_{\kappa }$ .

Denote by $\mathcal {H}_{n}^{d}(v_{\kappa })$ the subspace of $\kappa $ -spherical harmonics of degree $n\in \mathbb {Z}_{+}$ in $L^{2} (\mathbb {S}^{d-1},d\sigma _{\kappa })$ (see [Reference Dunkl and Xu13, Chapter 5]). Let $\mathcal {P}_{n}^{d}$ be the space of homogeneous polynomials of degree n in $\mathbb {R}^{d}$ . Then $\mathcal {H}_{n}^{d}(v_{\kappa })$ is the restriction of $\ker \Delta _{\kappa }\cap \mathcal {P}_{n}^{d}$ to the sphere $\mathbb {S}^{d-1}$ .

If $l_{n}$ is the dimension of $\mathcal {H}_{n}^{d}(v_{\kappa })$ , we denote by $\{Y_{n}^{j}\colon j=1,\ldots ,l_{n}\}$ the real-valued orthonormal basis of $\mathcal {H}_{n}^{d}(v_{\kappa })$ in $L^{2}(\mathbb {S}^{d-1},d\sigma _{\kappa })$ . A union of these bases forms an orthonormal basis in $L^{2}(\mathbb {S}^{d-1},d\sigma _{\kappa })$ consisting of k-spherical harmonics.

Let us rewrite (4.3) as follows

$$\begin{align*}B_{\kappa,a}(x,y) =\sum_{j=0}^{\infty}e^{-\frac{i\pi j}{a}}\,\frac{\lambda_{\kappa}+j}{\lambda_{\kappa}}\,\frac{\Gamma(2\lambda_{\kappa}/a+1)(|x||y|)^{j}}{a^{2j/a}\,\Gamma(2(\lambda_{\kappa}+j)/a+1)}\, j_{\frac{2(\lambda_{\kappa}+j)}{a}}\Bigl(\frac{2}{a}\,(|x||y|)^{a/2}\Bigr)V_{\kappa}C_{j}^{\lambda_{\kappa}}(\langle x',{\cdot}\,\rangle)(y'). \end{align*}$$

Integrating this and using orthogonality of $\kappa $ -spherical harmonics, as in the case of the Dunkl kernel (see [Reference Dunkl and Xu13, Theorem 5.3.4], [Reference Rösler21, Corollary 2.5]), we obtain the following crucial property of the kernel of the generalized Fourier transform.

Proposition 5.7. If $x,y\in \mathbb {R}^d$ , $x=\rho x'$ , $y=vy'$ , then

$$\begin{align*}\int_{\mathbb{S}^{d-1}}B_{\kappa,a}(x,vy')Y_{n}^{j}(y')\,d\sigma_{\kappa}(y')=\frac{e^{-\frac{i\pi n}{a}}\Gamma(2\lambda_{\kappa}/a+1)}{a^{2n/a} \,\Gamma(2(\lambda_{\kappa}+n)/a+1)}\,v^n j_{\frac{2(\lambda_{\kappa}+n)}{a}}\Bigl(\frac{2}{a}\,(\rho v)^{a/2}\Bigr)Y_n^j(x). \end{align*}$$

Proposition 5.8. For irrational a and a nontrivial function $f\in \mathcal {S}(\mathbb {R}^d)$ , we have $\mathcal {F}_{\kappa ,a}(f)\notin \mathcal {S}(\mathbb {R}^d)$ .

Proof. 1. First, assume that $f(x)=\rho ^n\psi (\rho )Y_n^j(x')$ , where $n\in \mathbb {Z}_+$ , $x=\rho x'$ , $|x'|=1$ , and $\psi \in \mathcal {S}(\mathbb {R}_+)$ . Since $\rho ^nY_n^j(x')=Y_n^j(x)$ is a homogeneous polynomial of degree n, then $f(x)=\psi (|x|)Y_n^j(x)\in \mathcal {S}(\mathbb {R}^d)$ . If $y=vy'$ , $\rho =(a/2)^{1/a}u^{2/a}$ , then by [Reference Gorbachev, Ivanov and Tikhonov15]

$$ \begin{align*} \mathcal{F}_{\kappa,a}(f)(y)=e^{-\frac{i\pi n}{a}}Y_n^j(y)&H_{\lambda_{\kappa}+n,a}(\psi)(v) =e^{-\frac{i\pi n}{a}}Y_n^j(y)\int_{0}^{\infty}\psi(\rho) j_{\frac{2(\lambda_{\kappa}+n)}{a}} \Bigl(\frac{2}{a}\,(v\rho)^{a/2} \Bigr)\,d\nu_{\lambda_{\kappa},a}(\rho)\\ & =e^{-\frac{i\pi n}{a}}Y_n^j(y)\int_{0}^{\infty}\psi\Bigl(\Bigl(\frac{a}{2}\Bigr)^{1/a}u^{2/a}\Bigr) j_{\frac{2(\lambda_{\kappa}+n)}{a}}\Bigl(\Bigl(\frac{2}{a}\Bigr)^{1/2}\,v^{a/2}u \Bigr)\,d\nu_{\frac{2(\lambda_{\kappa}+n)}{a}}(u). \end{align*} $$

Moreover, following the ideas used in Example 5.1, we obtain that the function

$$\begin{align*}g_a(v)=\int_{0}^{\infty}\psi\Bigl(\Bigl(\frac{a}{2}\Bigr)^{1/a}u^{2/a}\Bigr)j_{\frac{2(\lambda_{\kappa}+n)}{a}}\Bigl(\Bigl(\frac{2}{a}\Bigr)^{1/2}\,v^{a/2}u \Bigr)\,d\nu_{\frac{2(\lambda_{\kappa}+n)}{a}}(u) \end{align*}$$

with irrational a cannot decrease rapidly as $v\to \infty $ , and it has finite smoothness at the origin. Therefore, it follows that $\mathcal {F}_{\kappa ,a}(f)\notin \mathcal {S}(\mathbb {R}^d)$ .

2. Let now $f\in \mathcal {S}(\mathbb {R}^d)$ be any nonzero function. Then its spherical $\kappa $ -harmonic expansion is given by

$$\begin{align*}f(\rho x')=\sum_{n=0}^{\infty}\sum_{j=1}^{l_{n}}f_{nj}(\rho)Y_{n}^{j}(x'),\quad f_{nj}(\rho)=\int_{\mathbb{S}^{d-1}}f(\rho x')Y_{n}^{j}(x')\,d\sigma_{\kappa}(x') \end{align*}$$

(see [Reference Gorbachev, Ivanov and Tikhonov15]). Since the subspaces $\mathcal {H}_n^d(v_{\kappa })$ are orthogonal, then $f_{nj}^{(j)}(0)=0$ , $j=0,1,\dots ,n-1$ . Changing variables $x'\to -x'$ implies $f_{nj}(-\rho ) =(-1)^nf_{nj}(\rho ). $ Hence, setting a nonzero function $g_{nj}(f)(x)=f_{nj}(\rho )Y_{n}^{j}(x')\in \mathcal {S}(\mathbb {R}^d)$ , we have $g_{nj}(x)=\rho ^n\psi (\rho )Y_{n}^{j}(x')$ with some $\psi \in \mathcal {S}(\mathbb {R}_+)$ .

In order to apply the results obtained in case 1, we need to show that

$$ \begin{align*}\mathcal{F}_{\kappa,a}(g_{nj}(f))(y)=g_{nj}(\mathcal{F}_{\kappa,a}(f))(y).\end{align*} $$

Indeed, we note that

$$\begin{align*}\mathcal{F}_{\kappa,a}(g_{nj}(f))(y)=e^{-i\pi n/a}Y_n^j(y)H_{\lambda_{\kappa}+n,a}(\psi)(v). \end{align*}$$

On the other hand, in view of Proposition 5.7, we deduce that

$$ \begin{align*} g_{nj}(\mathcal{F}_{\kappa,a}(f))(y)&=Y_{n}^{j}(y')\int_{\mathbb{S}^{d-1}}\mathcal{F}_{\kappa,a}(f)(vy')Y_{n}^{j}(y')\,d\sigma_{\kappa,a}(y')\\ &=e^{-\frac{i\pi n}{a}}Y_{n}^{j}(y) \int_{\mathbb{R}^d}\frac{f(x)\Gamma(2\lambda_{\kappa}/a+1)}{a^{2n/a} \,\Gamma(2(\lambda_{\kappa}+n)/a+1)} j_{\frac{2(\lambda_{\kappa}+n)}{a}}\Bigl(\frac{2}{a}\,(\rho v)^{a/2}\Bigr)Y_n^j(x)\,d\mu_{\kappa,a}(x) \\ &=e^{-\frac{i\pi n}{a}}Y_{n}^{j}(y)\int_{0}^{\infty}\psi(\rho) j_{\frac{2(\lambda_{\kappa}+n)}{a}}\Bigl(\frac{2}{a}\,(\rho v)^{a/2}\Bigr)\,d\nu_{\lambda_{\kappa}+n,a}(\rho)\\ &=e^{-\frac{i\pi n}{a}}Y_n^j(y)H_{\lambda_{\kappa}+n,a}(\psi)(v). \end{align*} $$

Assuming here that $f,\mathcal {F}_{\kappa ,a}(f)\in \mathcal {S}(\mathbb {R}^d)$ yields $g_{nj}(f),\mathcal {F}_{\kappa ,a}(g_{nj}(f))\in \mathcal {S}(\mathbb {R}^d)$ , which contradicts the first case.

Summarizing, the generalized Fourier transform for irrational a drastically deforms even very smooth functions. It was mentioned in [Reference Ben Saïd, Kobayashi and Ørsted4, Chapter 5] that the generalized Fourier transform has a finite order only for rational a. Therefore, the case of irrational a is of little interest in harmonic analysis.

6 Nondeformed unitary transforms generated by $\mathcal {F}_{\kappa ,a}$

Let $a=\frac {2}{2r+1}$ and $\lambda =\lambda _{\kappa ,a}=(2\kappa -1)/a\ge -1/2$ . Recalling that A is given by (5.2), we may assume that $x, u\in \mathbb {R}$ . Since

$$\begin{align*}\int_{-\infty}^{\infty}|f(x)|^2\,d\mu_{\kappa,a}(x)=\int_{-\infty}^{\infty}|Af(u)|^2\,d\widetilde{\nu}_{\lambda}(u),\qquad d\widetilde{\nu}_{\lambda}(u) =\frac{|u|^{2\lambda+1}\,du}{2^{\lambda+1} \Gamma(\lambda+1)}, \end{align*}$$

the linear operator $A\colon L^{2}(\mathbb {R},d\mu _{\kappa ,a})\to L^{2}(\mathbb {R},d\widetilde {\nu }_{\lambda })$ is an isometric isomorphism and the inverse operator is given by $A^{-1}g(x)=g((2r+1)^{1/2}\,x^{1/(2r+1)})$ .

In view of (2.4) and (5.2), $B_{\kappa ,a}(x,y)=e_{2r+1}(uv,\lambda )$ and

$$ \begin{align*} A\mathcal{F}_{\kappa,a}(f)(v)=\int_{\mathbb{R}}e_{2r+1}(uv,\lambda)Af(u) \,d\widetilde{\nu}_{\lambda}(u). \end{align*} $$

This formula defines the nondeformed transform $\mathcal {F}_{r}^{\lambda }$ , for $\lambda> -1/2$ and $r\in \mathbb {Z}_+$ ,

$$ \begin{align*} \mathcal{F}_{r}^{\lambda}(g)(v)&=\int_{-\infty}^{\infty}e_{2r+1}(uv,\lambda)g(u) \,d\widetilde{\nu}_{\lambda}(u)\\ &=\int_{-\infty}^{\infty}\Bigl(j_{\lambda}(uv)+i(-1)^{r+1}\,\frac{(uv)^{2r+1}}{2^{2r+1}(\lambda+1)_{2r+1}}\, j_{\lambda+2r+1}(uv)\Bigr)g(u) \,d\widetilde{\nu}_{\lambda}(u)\\ &=c_{\lambda}\int_{-\infty}^{\infty}\int_{-1}^{1}(1-t^2)^{\lambda-1/2}(1+P_{2r+1}^{(\lambda-1/2)}(t))\,e^{-iuvt}\,dt\,g(u) \,d\widetilde{\nu}_{\lambda}(u). \end{align*} $$

Moreover, its kernel satisfies the estimate $|e_{2r+1}(uv,\lambda )|\le M_{\lambda }<\infty $ and, importantly, $M_{\lambda }=1$ for $\lambda \ge 0$ . If $r=0$ , we recover the one-dimensional Dunkl transform. Below, we study invariant subspaces ( $\subset C^{\infty }$ ) of the $\mathcal {F}_{r}^{\lambda }$ transform. The Plancherel theorem for $\mathcal {F}_{\kappa ,a}$ given by

$$\begin{align*}\int_{-\infty}^{\infty}|\mathcal{F}_{\kappa,a}(f)(y)|^2\,d\mu_{\kappa,a}(y)=\int_{-\infty}^{\infty}|f(x)|^2\,d\mu_{\kappa,a}(x), \quad f\in L^{2}(\mathbb{R},d\mu_{\kappa,a}), \end{align*}$$

implies that $\mathcal {F}_{r}^{\lambda }$ is a unitary operator in $L^{2}(\mathbb {R},d\widetilde {\nu }_{\lambda })$ , that is

$$\begin{align*}\int_{-\infty}^{\infty}|\mathcal{F}_{r}^{\lambda}(g)(v)|^2\,d\widetilde{\nu}_{\lambda}(v) =\int_{-\infty}^{\infty}|g(u)|^2\,d\widetilde{\nu}_{\lambda}(u). \end{align*}$$

Since the reverse operator satisfies $(\mathcal {F}_{\kappa ,a})^{-1}(f)(x)=\mathcal {F}_{\kappa ,a}(f)(-x)$ [Reference Ben Saïd, Kobayashi and Ørsted4, Theorem 5.3], we have

$$ \begin{align*} (\mathcal{F}_{r}^{\lambda})^{-1}(f)(u)=\int_{-\infty}^{\infty}\overline{e_{2r+1}(uv,\lambda)}f(v) \,d\widetilde{\nu}_{\lambda}(v). \end{align*} $$

If $g, \mathcal {F}_{r}^{\lambda }(g)\in L^{1}(\mathbb {R},d\widetilde {\nu }_{\lambda })$ , then one may assume that $g, \mathcal {F}_{r}^{\lambda }(g)\in C_b(\mathbb {R})$ . Moreover, the inversion formula

(6.1) $$ \begin{align} g(u)=\int_{-\infty}^{\infty}\overline{e_{2r+1}(uv,\lambda)}\mathcal{F}_{r}^{\lambda}(g)(v)\,d\widetilde{\nu}_{\lambda}(v) \end{align} $$

holds not only in an $L_2$ sense but also pointwise.

Considering the derivatives of the kernel $e_{2r+1}(uv,\lambda )$ , we note that

$$\begin{align*}\partial_{v}^{n}e_{2r+1}(uv,\lambda)=(iu)^{n}c_{\lambda}\int_{-1}^{1}t^{n} (1-t^2)^{\lambda-1/2}(1+P_{2r+1}^{(\lambda-1/2)}(t))\,e^{-iuvt}\,dt, \end{align*}$$

and so,

$$\begin{align*}|\partial_{v}^{n}e_{2r+1}(uv,\lambda)|\le M_{\lambda}|u|^{n}. \end{align*}$$

Then, for $g\in \mathcal {S}(\mathbb {R})$ , we have

(6.2) $$ \begin{align} |\partial^{n}\mathcal{F}_{r}^{\lambda}(g)(v)|=\Bigl|\int_{-\infty}^{\infty}g(u)\partial_{v}^{n}e_{2r+1}(uv,\lambda) \,d\widetilde{\nu}_{\lambda}(u)\Bigr|\le M_{\lambda}\int_{-\infty}^{\infty}|u|^{n}|g(u)|\,d\widetilde{\nu}_{\lambda}(u)<\infty. \end{align} $$

Therefore, $\mathcal {F}_{r}^{\lambda }(\mathcal {S}(\mathbb {R}))\subset C^{\infty }(\mathbb {R})$ . However, $\mathcal {F}_{r}^{\lambda }(\mathcal {S}(\mathbb {R}))\not \subset \mathcal {S}(\mathbb {R})$ . Indeed, assuming $g,\mathcal {F}_{r}^{\lambda }(g)\in \mathcal {S}(\mathbb {R})$ , by orthogonality of the Gegenbauer polynomials for $s=0,1,\dots ,r-1$ ,

$$\begin{align*}\int_{-1}^{1}t^{2s+1} (1-t^2)^{\lambda-1/2}(1\pm P_{2r+1}^{(\lambda-1/2)}(t,\lambda))\,dt=0 \end{align*}$$

and

(6.3) $$ \begin{align} \partial^{2s+1}\mathcal{F}_{r}^{\lambda}(g)(0)=0,\quad \partial^{2s+1}g(0)=0, \end{align} $$

which is not true for arbitrary g.

Put for $n\in \mathbb {Z}_+$

$$\begin{align*}\mathcal{S}_{n}(\mathbb{R})=\{g\in\mathcal{S}(\mathbb{R})\colon \partial^{2s+1}g(0)=0,\quad s=0,1,\dots,n-1\},\quad \mathcal{S}_{0}(\mathbb{R})=\mathcal{S}(\mathbb{R}). \end{align*}$$

The set $\mathcal {S}_{n}(\mathbb {R})$ is dense in $L^{2}(\mathbb {R},d\mu _{\kappa ,a})$ and in $L^{2}(\mathbb {R},d\widetilde {\nu }_{\lambda })$ .

Example 6.1. Consider the function $g_{2s+1}(u)=u^{2s+1}e^{-u^2}\in \mathcal {S}_{s}(\mathbb {R})$ , $s\in \mathbb {Z}_+$ . By means of (2.1), (2.4), [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi1, Chapter VI, 6.1], and [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi2, Chapter VIII, 8.6(14)], we get

$$ \begin{align*} \mathcal{F}_{r}^{\lambda}(g_{2s+1})(v)&=i(-1)^{r+1}v^{-\lambda}\int_{0}^{\infty}u^{\lambda+2s+2}e^{-u^2}J_{\lambda+2r+1}(uv)\,du\\ &=ic_{r,\lambda,s}\,v^{2r+1}\sum_{l=0}^{\infty}\frac{(\lambda+s+r+2)_l}{(\lambda+2r+2)_l}\Bigl(-\frac{v^2}{4}\Bigr)^{l}\\ &=ic_{r,\lambda,s}\,v^{2r+1}\Phi\Bigl(\lambda+s+r+2,\lambda+2r+2,-\frac{v^2}{4}\Bigr),\quad c_{r,\lambda,s}>0. \end{align*} $$

We consider two cases. If $s=0,1,\dots ,r-1$ , then asymptotics as $v\to \infty $ [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi3, Chapter VI,6.13.1] implies

$$\begin{align*}\Phi\Bigl(\lambda+s+r+2,\lambda+2r+2,-\frac{v^2}{4}\Bigr)=\frac{\Gamma(\lambda+2r+2)} {\Gamma(r-s)}\Bigl(\frac{v^2}{4}\Bigr)^{-(\lambda+s+r+2)}\Bigl(1+O\Bigl(\frac{1}{v^2}\Bigr)\Bigr), \end{align*}$$

that is, $\mathcal {F}_{r}^{\lambda }(g_{2s+1})\notin \mathcal {S}(\mathbb {R})$ . If $s\ge r$ , then applying the Kummer transform [Reference Bateman, Erdélyi, Magnus, Oberhettinger and Tricomi3, Chapter VI,6.3.7)] gives us

(6.4) $$ \begin{align} \mathcal{F}_{r}^{\lambda}(g_{2s+1})(v)&=ic_{r,\lambda,s}\,v^{2r+1}e^{-v^2/4} \Phi\Bigl(r-s,\lambda+2r+2,\frac{v^2}{4}\Bigr)\nonumber \\ &=ic_{r,\lambda,s}\,v^{2r+1}e^{-v^2/4} \sum_{l=0}^{s-r}\frac{(r-s)_l}{(\lambda+2r+2)_l}\Bigl(\frac{v^2}{4}\Bigr)^{l}, \end{align} $$

that is, $\mathcal {F}_{r}^{\lambda }(g_{2s+1})(v)\in \mathcal {S}_{r}(\mathbb {R})$ .

Since $g_{2r+1}\in \mathcal {S}_{r}(\mathbb {R})$ and $\mathcal {F}_{r}^{\lambda }(g_{2r+1})(v)\in \mathcal {S}_{r}(\mathbb {R})$ , one can conjecture that $\mathcal {F}_{r}^{\lambda }(\mathcal {S}_{r}(\mathbb {R}))= \mathcal {S}_{r}(\mathbb {R})$ . In order to show this (see Proposition 6.3), we will need some auxiliary results.

Recall that for the weight function $|x|^{2\lambda +1}$ , the differential-difference Dunkl operator of the first and second order are given by

$$\begin{align*}T_{\lambda+1/2}g(u)=\partial g(u)+(\lambda+1/2)\,\frac{g(u)-g(-u)}{u}, \end{align*}$$
$$\begin{align*}\Delta_{\lambda+1/2}g(u)=T_{\lambda+1/2}^2g(u)=\partial^{2} g(u)+\frac{2\lambda+1}{u}\,\partial g(u)-(\lambda+1/2)\,\frac{g(u)-g(-u)}{u^2} \end{align*}$$

(see [Reference Ben Saïd, Kobayashi and Ørsted4, Reference Rösler20]). Let us define the operator

$$\begin{align*}\delta_{\lambda}g(u)=T_{\lambda+1/2}^2g(u)-2r(\lambda+r+1)\,\frac{g(u)-g(-u)}{u^2}, \end{align*}$$

which is obtained by changing variables $x=x(u)$ as in (5.2) in the Dunkl Laplacian

$$\begin{align*}\Delta_{\kappa}f(x)=\frac{(2r+1)^{2r-1}}{u^{4r}}\,\delta_{\lambda}g(u). \end{align*}$$

By direct calculations, we verify that the kernel $e_{2r+1}(uv,\lambda )$ is the eigenfunction of $\delta _{\lambda }$ :

(6.5) $$ \begin{align} (\delta_{\lambda})_ue_{2r+1}(uv,\lambda)=-|v|^2e_{2r+1}(uv,\lambda). \end{align} $$

Using [Reference Rösler20, Proposition 2.18] for $g\in \mathcal {S}(\mathbb {R})$ , we have

(6.6) $$ \begin{align} \int_{-\infty}^{\infty}T_{\lambda+1/2}^2g(u)e_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u)= \int_{-\infty}^{\infty}g(u)(T_{\lambda+1/2}^2)_ue_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u). \end{align} $$

Suppose $g\in \mathcal {S}_1(\mathbb {R})$ , then

$$\begin{align*}\frac{g(u)-g(-u)}{u^2}=\frac{2g_{o}(u)}{u^2}\in \mathcal{S}(\mathbb{R}) \end{align*}$$

and

$$\begin{align*}\int_{-\infty}^{\infty}\frac{g(u)-g(-u)}{u^2}\,e_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u)= \int_{-\infty}^{\infty}g(u)\,\frac{e_{2r+1}(uv,\lambda)-e_{2r+1}(-uv,\lambda)}{u^2}\,d\widetilde{\nu}_{\lambda}(u). \end{align*}$$

This, (6.5) and (6.6) for any $g\in \mathcal {S}_1(\mathbb {R})$ yield

(6.7) $$ \begin{align} \int_{-\infty}^{\infty}\delta_{\lambda}g(u)e_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u)&= \int_{-\infty}^{\infty}g(u)(\delta_{\lambda})_ue_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u) \nonumber \\ &=-|v|^2\int_{-\infty}^{\infty}g(u)e_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u). \end{align} $$

Applying (6.7) for $g\in \mathcal {S}_n(\mathbb {R})$ , we get

(6.8) $$ \begin{align} \int_{-\infty}^{\infty}\delta_{\lambda}^ng(u)e_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u)= (-1)^n|v|^{2n}\int_{-\infty}^{\infty}g(u)e_{2r+1}(uv,\lambda)\,d\widetilde{\nu}_{\lambda}(u). \end{align} $$

Lemma 6.2. Suppose $g\in \mathcal {S}(\mathbb {R})$ , and

$$\begin{align*}a_n(g)(v)=\sum_{k=0}^n\sum_{l=0}^k\frac{1}{(k-l)!}\,\frac{\partial^{2l+1}g(0)}{(2l+1)!}\,v^{2k+1}e^{-v^2}, \end{align*}$$

then $g-a_n(g)\in \mathcal {S}_{n+1}(\mathbb {R})$ .

Proof. We have $g-a_n(g)\in \mathcal {S}(\mathbb {R})$ . Applying the Leibniz rule for $s=0,1,\dots ,n$ ,

$$ \begin{align*} \frac{\partial^{2s+1}a_n(g)(0)}{(2s+1)!}&=\frac{1}{(2s+1)!}\sum_{k=0}^s\sum_{l=0}^k\frac{\partial^{2l+1}g(0)}{(k-l)!\,(2l+1)!} \binom{2s+1}{2k+1}(2k+1)!\partial^{2s-2k}(e^{-u^2})(0)\\ &=\sum_{k=0}^s\sum_{l=0}^k\frac{\partial^{2l+1}g(0)(-1)^{s+k}}{(k-l)!\,(2l+1)!\,(s-k)!} = \sum_{l=0}^s\frac{\partial^{2l+1}g(0)}{(2l+1)!}\sum_{m=0}^{s-l}\frac{(-1)^{s+l+m}}{m!\,(s-l-m)!}\\ &=\frac{1}{(s-l)!}\sum_{l=0}^s\frac{(-1)^{s+l}\partial^{2l+1}g(0)}{(s-l)!\,(2l+1)!}(1-1)^{s-l}=\frac{\partial^{2s+1}g(0)}{(2s+1)!}.\\[-40pt] \end{align*} $$

Proposition 6.3. Let $\lambda>-1/2$ and $r\in \mathbb {Z}_+$ . We have $\mathcal {F}_{r}^{\lambda }(\mathcal {S}_{r}(\mathbb {R}))=\mathcal {S}_{r}(\mathbb {R})$ .

If $r=0$ , we recover the result by de Jeu [Reference de Jeu8] for the Dunkl transform.

Proof. Let $r\in \mathbb {N}$ and $g\in \mathcal {S}_{r}(\mathbb {R})$ . It is enough to show that g satisfies the condition:

(6.9) $$ \begin{align} \partial^{m}(v^{2n}\mathcal{F}_{r}^{\lambda}(g)(v))\ \text{is bounded for any}\ m\in\mathbb{Z}_+,\ n\ge r+1. \end{align} $$

We have

$$\begin{align*}\partial^{m}(v^{2n}\mathcal{F}_{r}^{\lambda}(g)(v))=\partial^{m}(v^{2n}\mathcal{F}_{r}^{\lambda}(g-a_{n-1}(g))(v)) +\partial^{m}(v^{2n}\mathcal{F}_{r}^{\lambda}(a_{n-1}(g))(v)). \end{align*}$$

Since $\partial ^{2l+1}g(0)=0$ , $l=0,1,\dots ,r-1$ , then

$$\begin{align*}a_{n-1}(g)(v)=\sum_{k=r}^{n-1}\sum_{l=r}^k\frac{1}{(k-l)!}\, \frac{\partial^{2l+1}g(0)}{(2l+1)!}\,v^{2k+1}e^{-v^2}\in\mathcal{S}_{r}(\mathbb{R}). \end{align*}$$

By (6.4), condition (6.9) is valid for $a_{n-1}(g)$ . By Lemma 6.2, $g-a_{n-1}(g)\in \mathcal {S}_n(\mathbb {R})$ . In light of (6.8), $v^{2n}\mathcal {F}_{r}^{\lambda }(g-a_{n-1}(g))(v)$ is the $\mathcal {F}_{r}^{\lambda }$ -transform of the function $(-1)^n\delta _{\lambda }^n(g-a_{n-1}(g))\in \mathcal {S}(\mathbb {R})$ . Applying for this transform inequality (6.2), we get the property (6.9) for $g-a_{n-1}(g)$ . Therefore, $\mathcal {F}_{r}^{\lambda }(g)\in \mathcal {S}(\mathbb {R})$ . By virtue of (6.1) and (6.3), $\mathcal {F}_{r}^{\lambda }(g)\in \mathcal {S}_{r}(\mathbb {R})$ .

Remark 6.4. An alternative proof of Proposition 6.3 reads as follows. Let $g\in \mathcal {S}_{r}(\mathbb {R})$ . In light of (5.1)–(5.3), we have the following representation

$$ \begin{align*} \mathcal{F}_{r}^{\lambda}(g)(v)&=\int_{-\infty}^{\infty}j_{\lambda}(uv)g(u)\,d\widetilde{\nu}_{\lambda}(u)+ \frac{i(-1)^{r+1}v^{2r+1}}{2^{2r+1}(\lambda+1)_{2r+1}}\int_{-\infty}^{\infty}u^{2r+1}j_{\lambda+2r+1}(uv)g(u)\,d\widetilde{\nu}_{\lambda}(u)\\ &=\int_{0}^{\infty}j_{\lambda}(uv)g_{e}(u)\,d\nu_{\lambda}(u)+i(-1)^{r+1}v^{2r+1}\int_{0}^{\infty}j_{\lambda+2r+1}(uv)u^{-(2r+1)}g_{o}(u) \,d\nu_{\lambda+2r+1}(u)\\ &=H_{\lambda}(g_e)(v)++i(-1)^{r+1}v^{2r+1}H_{\lambda+2r+1}(u^{-(2r+1)}g_o)(v) =F_1(v)+v^{2r+1}F_2(v), \end{align*} $$

where even functions $F_1,F_2\in \mathcal {S}(\mathbb {R})$ . Therefore, $\mathcal {F}_{r}^{\lambda }(g)\in \mathcal {S}_{r}(\mathbb {R})$ . The first proof was given to underline the important properties of $\mathcal {F}_{r}^{\lambda }$ -transform and its kernel.

Since $f\in \mathcal {S}(\mathbb {R})$ implies $Af\in \mathcal {S}_r(\mathbb {R})$ , Proposition 6.3 yields the following result.

Corollary 6.5. Let $a=\frac {2}{2r+1}$ and $\kappa \ge \frac {r}{2r+1}$ . If $f\in \mathcal {S}(\mathbb {R})$ , then $\mathcal {F}_{\kappa ,a}(f)((2r+1)^{-(r+1/2)} v^{2r+1})\in \mathcal {S}_r(\mathbb {R})$ or, equivalently, $\mathcal {F}_{\kappa ,a}(f)(v)=g((2r+1)^{1/2}\,v^{1/(2r+1)})$ , $g\in \mathcal {S}_r(\mathbb {R})$ ; cf. (5.4).

Now, we discuss the case $\lambda =-1/2$ .

Remark 6.6. If $\lambda =-1/2$ , $r\in \mathbb {Z}_+$ , then

$$\begin{align*}\delta_{-1/2}g(u)=\partial^{2}g(u)-r(2r+1)\,\frac{g(u)-g(-u)}{u^2},\quad e_{1}(uv,-1/2)=e^{-iuv}\quad (r=0). \end{align*}$$

Taking into account (2.4), and passing to the limit in (2.4) as $\lambda \to -1/2$ , we deduce that for $r\ge 1$

$$ \begin{align*} e_{2r+1}(uv,-1/2)&=\cos{}(uv)+ i(-1)^{r+1}\,\frac{(uv)^{2r+1}}{2^{2r+1}(1/2)_{2r+1}} j_{2r+1/2}(uv)\\ &=e^{-iuv}-(r+1/2)\int_{-1}^{1}\sum_{s=0}^{r-1}(-1)^s\binom{r}{s+1}\frac{(r+3/2)_s}{s!}\,(1-t^2)^ste^{-iuvt}\,dt. \end{align*} $$

Taking this into account and analyzing the proofs above, we note that all mentioned results in this section for the transform $\mathcal {F}_{r}^{\lambda }$ in the case $\lambda>-1/2$ are also valid for $\lambda =-1/2$ . In particular, $\mathcal {F}_{r}^{-1/2}$ , $r\in \mathbb {Z}_+$ , are the unitary transforms in the nonweighted $L^2(\mathbb {R},dx)$ , where $\mathcal {F}_{0}^{-1/2}$ corresponds to the classical Fourier transform.

Acknowledgements

We would like to thank the unknown referees for their valuable comments and suggestions which helped to improve the paper. The research of D. Gorbachev and V. Ivanov was performed by a grant of RSF (project 18-11-00199), https://rscf.ru/project/18-11-00199. S. Tikhonov was partially supported by PID2020-114948GB-I00, 2021 SGR 00087, the CERCA Programme of the Generalitat de Catalunya, Severo Ochoa and María de Maeztu Program for Centers and Units of Excellence in R&D (CEX2020-001084-M), and the Ministry of Education and Science of the Republic of Kazakhstan (AP 14870758).

Competing interest

The authors have no competing interest to declare.

Footnotes

1 The case $d=1$ was first considered in [Reference Ben Saïd, Kobayashi and Ørsted4]. The assertion [Reference Ben Saïd, Kobayashi and Ørsted4, Theorem 5.11] was corrected in [Reference Gorbachev, Ivanov and Tikhonov15].

2 The case $ \langle \kappa \rangle> 0$ was announced in [Reference Constales, De Bie and Lian10, Remark 3]. The proof is similar to the one of [Reference Constales, De Bie and Lian10, Theorem 9].

3 As usual, $x^{\frac {1}{2r+1}}=-|x|^{\frac {1}{2r+1}}$ for $x<0$ .

References

Bateman, H., Erdélyi, A., Magnus, W., Oberhettinger, F. and Tricomi, F. G., Higher Transcendental Functions, II (McGraw Hill Book Company, New York, 1953).Google Scholar
Bateman, H., Erdélyi, A., Magnus, W., Oberhettinger, F. and Tricomi, F. G., Tables of Integral Transforms, II (McGraw-Hill Book Company, New York, 1954).Google Scholar
Bateman, H., Erdélyi, A., Magnus, W., Oberhettinger, F., Tricomi, F. G., et al., Higher Transcendental Functions, I (McGraw Hill Book Company, New York, 1953).Google Scholar
Ben Saïd, S., Kobayashi, T. and Ørsted, B., ‘Laguerre semigroup and Dunkl operators’, Compos. Math. 148(4) (2012), 12651336.CrossRefGoogle Scholar
De Bie, H., ‘The kernel of the radially deformed Fourier transform’, Integral Transforms Spec. Funct. 24 (2013), 10001008.CrossRefGoogle Scholar
De Bie, H. and Lian, P., ‘The Dunkl kernel and intertwining operator for dihedral groups’, J. Funct. Anal. 280(7) (2021), 108932.CrossRefGoogle Scholar
Boubatra, M., Negzaoui, S. and Sifi, M., ‘A new product formula involving Bessel functions’, Integral Transforms Spec. Funct. 33(3) (2022), 247263.CrossRefGoogle Scholar
de Jeu, M. F. E., ‘The Dunkl transform’, Invent. Math. 113 (1993), 147162.CrossRefGoogle Scholar
Chamayou, J. F. and Wesolowski, J., ‘Lauricella and Humbert functions through probabilistic tools’, Integral Transf. Spec. Funct. 20 (2009), 529538.CrossRefGoogle Scholar
Constales, D., De Bie, H. and Lian, P., ‘Explicit formulas for the Dunkl dihedral kernel and the $\left(\kappa, a\right)$ -generalized Fourier kernel’, J. Math. Anal. Appl. 460(2) (2018), 900926.CrossRefGoogle Scholar
Deleaval, L. and Demni, N., ‘Generalized Bessel functions of dihedral-type: Expression as a series of confluent Horn functions and Laplace-type integral representation’, Ramanujan J. 54 (2021), 197217.CrossRefGoogle Scholar
Dunkl, C. F., ‘Hankel transforms associated to finite reflections groups’, Contemp. Math. 138 (1992), 123138.CrossRefGoogle Scholar
Dunkl, C. F. and Xu, Y., Orthogonal Polynomials of Several Variables (Cambridge University Press, Cambridge, 2001).CrossRefGoogle Scholar
Exton, H., Multiple Hypergeometric Functions and Applications (Ellis Horwood, Chichester, NY, 1983), 61132.Google Scholar
Gorbachev, D. V., Ivanov, V. I. and Tikhonov, S. Yu., ‘Pitt’s inequalities and uncertainty principle for generalized Fourier transform’, Int. Math. Res. Not. 2016(23) (2016), 71797200.CrossRefGoogle Scholar
Howe, R., ‘The oscillator semigroup’, in The Mathematical Heritage of Hermann Weyl, Proceedings of Symposia in Pure Mathematics vol. 48 (American Mathematical Society, Providence, 1988). R. O. Wells, editor.CrossRefGoogle Scholar
Humbert, P., ‘The confluent hypergeometric functions of two variables’, Proc. Roy. Soc. Edinburgh 41 (1920), 7382.CrossRefGoogle Scholar
Johansen, T. R., ‘Weighted inequalities and uncertainty principles for the $\left(k;a\right)$ -generalized Fourier transform’, Int. J. Math. 27(3) (2016), 1650019.CrossRefGoogle Scholar
Kobayashi, T. and Mano, G., ‘The Schrödinger model for the minimal representation of the indefinite orthogonal group $O\left(p;q\right)$ ’, in Memoirs of the American Mathematical Societies vol. 213(1000) (American Mathematical Society, Providence, 2011), 132 pp.Google Scholar
Rösler, M., ‘Dunkl operators. Theory and applications’, in Orthogonal Polynomials and Special Functions, Lecture Notes in Mathematics vol. 1817 (Springer-Verlag, Berlin, Heidelberg, 2002), 93135.Google Scholar
Rösler, M., ‘A positive radial product formula for the Dunkl kernel’, Trans. Amer. Math. Soc. 355(6) (2003), 24132438.CrossRefGoogle Scholar
Rösler, M., ‘Positivity of Dunkl’s intertwining operator’, Duke Math. J. 98 (1999), 445463.CrossRefGoogle Scholar
Szegö, G., Orthogonal Polynomials, third edn (American Mathematical Society, Providence, RI, 1974).Google Scholar
Titchmarsh, E. C., Introduction to the Theory of Fourier Integrals, second edn (Clarendon Press, Oxford, 1948).Google Scholar
Watson, G. N., A Treatise on the Theory of Bessel Functions, second edn (Cambridge University Press, Cambridge, 1966).Google Scholar