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

Two results on the Dunkl maximal operator

Luc DELEAVAL Institut de Mathématiques de Jussieu, Université Pierre et Marie Curie, 175, rue du Chevaleret, 75013 Paris, France deleaval@math.jussieu.fr
(Date: July, 2010)
Abstract.

In this article, we first improve the scalar maximal theorem for the Dunkl maximal operator by giving some precisions on the behavior of the constants of this theorem for a general reflection group. Next we complete the vector-valued theorem for the Dunkl-type Fefferman-Stein operator in the case 2d\mathbb{Z}_{2}^{d} by establishing a result of exponential integrability corresponding to the case p=+p=+\infty.

Key words and phrases:
Dunkl operators, Maximal operator, Fefferman-Stein-type operator, Real analysis.
2000 Mathematics Subject Classification:
42B10, 42B25
The author is pleased to express his thanks to Banafsheh for her advises in english.

1. Introduction

Dunkl operators provide an essential tool to extend Fourier analysis on Euclidean spaces and analysis on Riemannian symmetric spaces of Euclidean type. Since their invention in 1989, these operators have largely contributed, in the setting of root systems and associated reflection groups, to the development of harmonic analysis and to the theory of multivariable hypergeometric functions.
In this paper, we focus on the Dunkl maximal operator MκWM_{\kappa}^{W} which is defined by

MκWf(x)=supr>01μκW(Br)|df(y)τxW(χBr)(y)dμκW(y)|,xd,M_{\kappa}^{W}f(x)=\sup_{r>0}\frac{1}{\mu_{\kappa}^{W}(B_{r})}\biggl{|}\int_{\mathbb{R}^{d}}f(y)\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\,\mathrm{d}\mu_{\kappa}^{W}(y)\biggr{|},\quad x\in\mathbb{R}^{d},

where χBr\chi_{{}_{B_{r}}} is the characteristic function of the Euclidean ball of radius rr centered at the origin, τxW\tau_{x}^{W} is the Dunkl translation and μκW\mu_{\kappa}^{W} is a weighted Lebesgue measure invariant under the action of the reflection group WW (see Section 2 for more details). This operator, which reduces to the well-known Hardy-Littlewood maximal operator in the case where the multiplicity function κ\kappa is equal to 0 (see Section 2 for details), is of particular interest for harmonic analysis associated with root systems. Nevertheless, the structure of the Dunkl translation prevents us from using the tools of real analysis (covering lemma, weighted inequality, Calderón-Zygmund decomposition…) and makes the study of MκWM_{\kappa}^{W} difficult.
However, Thangavelu and Xu [21] succeeded in proving the following scalar maximal theorem, where we denote by Lp(μκW)L^{p}(\mu_{\kappa}^{W}) the space Lp(d;μκW)L^{p}(\mathbb{R}^{d};\mu_{\kappa}^{W}) (for 1p+1\leqslant p\leqslant+\infty) and we use the shorter notation W,p\mathopen{\|}\cdot\mathclose{\|}_{W,p} instead of Lp(μκW)\mathopen{\|}\cdot\|_{L^{p}(\mu_{\kappa}^{W})}.

Theorem 1.1 (Scalar maximal theorem).

Let ff be a measurable function defined on d\mathbb{R}^{d}.

  1. (1)

    If fL1(μκW)f\in L^{1}(\mu_{\kappa}^{W}), then for every λ>0\lambda>0,

    μκW({xd:MκWf(x)>λ})CλfW,1,\mu_{\kappa}^{W}\Bigl{(}\Bigl{\{}x\in\mathbb{R}^{d}:M_{\kappa}^{W}f(x)>\lambda\Bigr{\}}\Bigr{)}\leqslant\frac{C}{\lambda}\|f\|_{W,1},

    where C=C(d,κ)C=C(d,\kappa) is a constant independent of ff and λ\lambda.

  2. (2)

    If fLp(μκW)f\in L^{p}(\mu_{\kappa}^{W}) with 1<p+1<p\leqslant+\infty, then MκWfLp(μκW)M_{\kappa}^{W}f\in L^{p}(\mu_{\kappa}^{W}) and

    MκWfW,pCfW,p,\bigl{\|}M_{\kappa}^{W}f\bigr{\|}_{W,p}\leqslant C\|f\|_{W,p},

    where C=C(d,κ,p)C=C(d,\kappa,p) is a constant independent of ff.

In order to prove this theorem, Thangavelu and Xu have used the following Hopf-Dunford-Schwartz ergodic theorem (see [3]).

Theorem 1.2 (Hopf-Dunford-Schwartz ergodic theorem).

Let XX be a measurable space and let mm be a positive measure on XX. Let {Tt}t0\{T_{t}\}_{t\geqslant 0} be a contraction semigroup of operators on Lp(X;m)L^{p}(X;m), that is, a semigroup which satisfies, for every p[1,+]p\in[1,+\infty] and every fLp(X;m)f\in L^{p}(X;m),

TtfLp(X;m)fLp(X;m).\|T_{t}f\|_{L^{p}(X;m)}\leqslant\|f\|_{L^{p}(X;m)}.

Define

Mf(x)=supt>0|1t0tTsf(x)ds|.Mf(x)=\sup_{t>0}\biggl{|}\frac{1}{t}\int_{0}^{t}T_{s}f(x)\,\mathrm{d}s\biggr{|}.
  1. (1)

    If fL1(X;m)f\in L^{1}(X;m), then for every λ>0\lambda>0,

    m({xX:Mf(x)>λ})2λfL1(X;m).m\Bigl{(}\Bigl{\{}x\in X:Mf(x)>\lambda\Bigr{\}}\Bigr{)}\leqslant\frac{2}{\lambda}\|f\|_{L^{1}(X;m)}.
  2. (2)

    If fLp(X;m)f\in L^{p}(X;m), with 1<p+1<p\leqslant+\infty, then MfLp(X;m)Mf\in L^{p}(X;m) and

    MfLp(X;m)CfLp(X;m),\bigl{\|}Mf\bigr{\|}_{L^{p}(X;m)}\leqslant C\|f\|_{L^{p}(X;m)},

    where C=C(p)C=C(p) is a constant independent of ff.

We will see that we can use the previous theorem in order to refine the scalar maximal theorem. More precisely, our first result is the following, where we denote by 2γ2\gamma the degree of homogeneity of the measure μκW\mu_{\kappa}^{W}.

Theorem 1.3.

Let ff be a measurable function defined on d\mathbb{R}^{d}.

  1. (1)

    There exists a numerical constant CC such that if fL1(μκW)f\in L^{1}(\mu_{\kappa}^{W}), then for every λ>0\lambda>0,

    μκW({xd:MκWf(x)>λ})Cd+2γλfW,1.\mu_{\kappa}^{W}\Bigl{(}\Bigl{\{}x\in\mathbb{R}^{d}:M_{\kappa}^{W}f(x)>\lambda\Bigr{\}}\Bigr{)}\leqslant C\frac{d+2\gamma}{\lambda}\|f\|_{W,1}.
  2. (2)

    There exists a numerical constant CC such that if fLp(μκW)f\in L^{p}(\mu_{\kappa}^{W}) with 1<p<+1<p<+\infty, then

    MκWfW,pCpp1d+2γfW,p.\bigl{\|}M_{\kappa}^{W}f\bigr{\|}_{W,p}\leqslant C\frac{p}{p-1}\sqrt{d+2\gamma}\,\|f\|_{W,p}.

In the particular case where γ=0\gamma=0, the previous theorem coincides with a theorem due to Stein and Strömberg for the Hardy-Littlewood maximal operator ([18]).

Our second result deals with the vector-valued extension of the scalar maximal theorem which has been proved in [2] in the case where the reflection group is 2d\mathbb{Z}_{2}^{d}. Let us recall this theorem. Denote by κ2d\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}} the Dunkl-type Fefferman-Stein operator given for a sequence f=(fn)n1f=(f_{n})_{n\geqslant 1} of measurable functions by

κ2df=(Mκ2dfn)n1.\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f=\bigl{(}M_{\kappa}^{\mathbb{Z}_{2}^{d}}f_{n}\bigr{)}_{n\geqslant 1}.
Theorem 1.4 (Vector-valued maximal theorem).

Let W=2dW=\mathbb{Z}_{2}^{d} and let f=(fn)n1f=(f_{n})_{n\geqslant 1} be a sequence of measurable functions defined on d\mathbb{R}^{d}.

  1. (1)

    Let 1<r<+1<r<+\infty. If frL1(μκ2d)\|f\|_{\ell^{r}}\in L^{1}(\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}), then for every λ>0\lambda>0,

    μκ2d({xd:κ2d(x)r>λ})Cλfr2d,1,\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in\mathbb{R}^{d}:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}(x)\|_{\ell^{r}}>\lambda\Bigr{\}}\Bigr{)}\leqslant\frac{C}{\lambda}\bigl{\|}\,\|f\|_{\ell^{r}}\bigr{\|}_{\mathbb{Z}^{d}_{2},1},

    where C=C(d,κ,r)C=C(d,\kappa,r) is a constant independent of (fn)n1(f_{n})_{n\geqslant 1} and λ\lambda.

  2. (2)

    Let 1<r,p<+1<r,p<+\infty. If frLp(μκ2d)\|f\|_{\ell^{r}}\in L^{p}(\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}), then

    κ2dfr2d,pCfr2d,p,\bigl{\|}\,\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f\|_{\ell^{r}}\bigr{\|}_{\mathbb{Z}_{2}^{d},p}\leqslant C\bigl{\|}\,\|f\|_{\ell^{r}}\bigr{\|}_{\mathbb{Z}_{2}^{d},p},

    where C=C(d,κ,p,r)C=C(d,\kappa,p,r) is a constant independent of (fn)n1(f_{n})_{n\geqslant 1}.

We will see that no analogue of (2)(2) holds when p=+p=+\infty. However, in this case we will give the following result of exponential integrability on every compact set, which generalizes the classical one due to Fefferman and Stein (see [10] or [20, page 75]).

Theorem 1.5.

Let f=(fn)n1f=(f_{n})_{n\geqslant 1} be a sequence of measurable functions defined on d\mathbb{R}^{d} and let 1<r<+1<r<+\infty. If frL(μκ2d)\|f\|_{\ell^{r}}\in L^{\infty}(\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}) is such that

μκ2d(suppfrr)<+,\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)}<+\infty,

then the function κ2dfrr\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f\|_{\ell^{r}}^{r} is exponentially integrable on every compact set. More precisely there exists a constant Cd,κ,rC_{d,\kappa,r}, which depends only on dd, κ\kappa and rr, such that for every compact subset KK of d\mathbb{R}^{d} and for every ε\varepsilon satisfying

0ε<log(2)2Cd,κ,rfrr2d,,0\leqslant\varepsilon<\frac{\log(2)}{2C_{d,\kappa,r}\bigl{\|}\,\|f\|_{\ell^{r}}^{r}\bigr{\|}_{\mathbb{Z}_{2}^{d},\infty}},

we have the inequality

Keεκ2df(x)rrdμκ2d(x)μκ2d(K)+2εCd,κ,rfrr2d,max{2μκ2d(K);μκ2d(suppfrr)}log(2)2εCd,κ,rfrr2d,.\int_{K}\mathrm{e}^{\varepsilon\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}}\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(x)\\ \leqslant\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(K)+\frac{2\varepsilon C_{d,\kappa,r}\bigl{\|}\,\|f\|_{\ell^{r}}^{r}\bigr{\|}_{\mathbb{Z}_{2}^{d},\infty}\max\Bigl{\{}2\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(K);\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)}\Bigr{\}}}{\log(2)-2\varepsilon C_{d,\kappa,r}\bigl{\|}\,\|f\|_{\ell^{r}}^{r}\bigr{\|}_{\mathbb{Z}_{2}^{d},\infty}}.

The paper is organized as follows. In the next section, we collect some definitions and results related to Dunkl’s analysis. We then give in Section 3 the proof of Theorem 1.3. A counterexample in the case p=+p=+\infty is given for the vector-valued maximal theorem in Section 4 and the substitute result contained in Theorem 1.5 is established.

2. Preliminaries

This section is devoted to the preliminaries and background. We only focus on the aspects of the Dunkl theory which will be relevant in what follows. For a large survey about this theory, the reader may especially consult [9, 16] and the references therein.
Let W𝒪(d)W\subset\mathcal{O}(\mathbb{R}^{d}) be a finite reflection group associated with a reduced root system \mathcal{R} (not necessarily crystallographic) and let κ:\kappa:\mathcal{R}\to\mathbb{C} be a multiplicity function, that is, a WW-invariant function. We assume in this article that κ\kappa takes value in [0,+[[0,+\infty[.
The (rational) Dunkl operators TξT_{\xi}^{\mathcal{R}} on d\mathbb{R}^{d}, introduced in [4], are the following κ\kappa–deformations of directional derivatives ξ\partial_{\xi} by reflections:

Tξf(x)=ξf(x)+α+κ(α)f(x)f(σα(x))x,αξ,α,xd,T_{\xi}^{\mathcal{R}}f(x)=\partial_{\xi}f(x)+\sum_{\alpha\in\mathcal{R}_{+}}\kappa(\alpha)\frac{f(x)-f(\sigma_{\alpha}(x))}{\langle x,\alpha\rangle}\langle\xi,\alpha\rangle,\quad x\in\mathbb{R}^{d},

where σα\sigma_{\alpha} denotes the reflection with respect to the hyperplane orthogonal to α\alpha, and +\mathcal{R}_{+} denotes a positive subsystem of \mathcal{R}. The definition is of course independent of the choice of a positive subsystem since κ\kappa is WW-invariant. The most important property of these operators is their commutativity, that is, TξTξ=TξTξT^{\mathcal{R}}_{\xi}T^{\mathcal{R}}_{\xi^{\prime}}=T^{\mathcal{R}}_{\xi^{\prime}}T^{\mathcal{R}}_{\xi} ([4]). Therefore, we are naturally led to consider the eigenfunction problem

(2.1) Tξf=y,ξfξdT^{\mathcal{R}}_{\xi}f=\langle y,\xi\rangle f\quad\forall\xi\in\mathbb{R}^{d}

with ydy\in\mathbb{C}^{d} a fixed parameter. This problem has been completely solved by Opdam ([11]).

Theorem 2.1.

Let ydy\in\mathbb{C}^{d}. There exists a unique solution f=EκW(,y)f=E_{\kappa}^{W}(\cdot,y) of

Tξf=y,ξfξdT^{\mathcal{R}}_{\xi}f=\langle y,\xi\rangle f\quad\forall\xi\in\mathbb{R}^{d}

which is real-analytic on d\mathbb{R}^{d} and satisfies f(0)=1f(0)=1. Moreover the Dunkl kernel EκWE_{\kappa}^{W} extends to a holomorphic function on d×d\mathbb{C}^{d}\times\mathbb{C}^{d}.

In fact, the existence of a solution has already been proved by Dunkl ([5]). Indeed, he noticed the existence of an intertwining operator VκWV_{\kappa}^{W} which satisfies

VκW(𝒫nd)=𝒫nd,VκW|𝒫0d=Id|𝒫0d,TξVκW=VκWξξd,V^{W}_{\kappa}(\mathcal{P}^{d}_{n})=\mathcal{P}^{d}_{n},\quad\ {V^{W}_{\kappa}}_{|_{\mathcal{P}^{d}_{0}}}=\mathrm{Id}_{|_{\mathcal{P}^{d}_{0}}},\quad\ T^{\mathcal{R}}_{\xi}V^{W}_{\kappa}=V^{W}_{\kappa}\partial_{\xi}\ \,\,\forall\xi\in\mathbb{R}^{d},

where 𝒫n\mathcal{P}_{n} denotes the space of homogeneous polynomials of degree nn in dd variables. Since the exponential function is a solution of (2.1) when κ=0\kappa=0 (that is, when Tξ=ξT^{\mathcal{R}}_{\xi}=\partial_{\xi}), he naturally set Eκ(,y)=VκW(e,y)E_{\kappa}(\cdot,y)=V_{\kappa}^{W}(\mathrm{e}^{\langle\cdot,y\rangle}). Unfortunately, the Dunkl kernel is explicitly known only in some special cases; when the root system is of A2A_{2}-type ([7]), of B2B_{2}-type ([8]) and when the reflection group is 2d\mathbb{Z}_{2}^{d} ([6, 22]). Nevertheless we know that this kernel has many properties in common with the classical exponential to which it reduces when κ=0\kappa=0. For significant results on this kernel and the intertwining operator, the reader may especially consult [1, 5, 11, 12, 14, 17]. The Dunkl kernel is of particular interest as it gives rise to an integral transform which is taken with respect to a weighted Lebesgue measure invariant under the action of WW and which generalizes the Euclidean Fourier transform.
More precisely, let us introduce the measure dμκW(x)=hκ2(x)dx\mathrm{d}\mu_{\kappa}^{W}(x)=h_{\kappa}^{2}(x)\,\mathrm{d}x where the weight given by

hκ2(x)=α+|x,α|2κ(α)h_{\kappa}^{2}(x)=\prod_{\alpha\in\mathcal{R}_{+}}|\langle x,\alpha\rangle|^{2\kappa(\alpha)}

is homogeneous of degree 2γ2\gamma with

γ=α+κ(α).\gamma=\sum_{\alpha\in\mathcal{R}_{+}}\kappa(\alpha).

Then for every fL1(μκW)f\in L^{1}(\mu_{\kappa}^{W}), the Dunkl transform of ff, denoted by κW(f)\mathcal{F}_{\kappa}^{W}(f), is defined by

κW(f)(x)=cκWdEκW(ix,y)f(y)dμκW(y),xd,\mathcal{F}_{\kappa}^{W}(f)(x)=c_{\kappa}^{W}\int_{\mathbb{R}^{d}}E^{W}_{\kappa}(-ix,y)f(y)\,\mathrm{d}\mu_{\kappa}^{W}(y),\quad x\in\mathbb{R}^{d},

where cκWc_{\kappa}^{W} is the Mehta-type constant

cκW=(dex22dμκW(x))1.c_{\kappa}^{W}=\biggl{(}\int_{\mathbb{R}^{d}}\mathrm{e}^{-\frac{\|x\|^{2}}{2}}\,\mathrm{d}\mu^{W}_{\kappa}(x)\biggr{)}^{-1}.

Let us point out that the Dunkl transform coincides with the Euclidean Fourier transform when κ=0\kappa=0 and that it is more or less a Hankel transform when d=1d=1. The two main properties of the Dunkl transform are given in the following theorem ([1, 6]).

Theorem 2.2.

  1. (1)

    Inversion formula   Let fL1(μκW)f\in L^{1}(\mu_{\kappa}^{W}). If κW(f)\mathcal{F}_{\kappa}^{W}(f) is in L1(μκW)L^{1}(\mu_{\kappa}^{W}), then we have the following inversion formula:

    f(x)=cκWdEκW(ix,y)κW(f)(y)dμκW(y).f(x)=c_{\kappa}^{W}\int_{\mathbb{R}^{d}}E^{W}_{\kappa}(ix,y)\mathcal{F}_{\kappa}^{W}(f)(y)\,\mathrm{d}\mu_{\kappa}^{W}(y).
  2. (2)

    Plancherel theorem   The Dunkl transform has a unique extension to an isometric isomorphism of L2(μκW)L^{2}(\mu_{\kappa}^{W}).

The Dunkl transform shares many other properties with the Fourier transform. Therefore, it is natural to associate a generalized translation operator with this transform.
There are many ways to define the Dunkl translation but we use the definition which most underlines the analogy with the Fourier transform. It is the definition given in [21] with a different convention. Let xdx\in\mathbb{R}^{d}. The Dunkl translation fτxWff\mapsto\tau_{x}^{W}f is defined on L2(μκW)L^{2}(\mu_{\kappa}^{W}) by the equation

κW(τxWf)(y)=EκW(ix,y)κW(f)(y),yd.\mathcal{F}_{\kappa}^{W}(\tau_{x}^{W}f)(y)=E_{\kappa}^{W}(ix,y)\mathcal{F}_{\kappa}^{W}(f)(y),\quad y\in\mathbb{R}^{d}.

It is useful to have a set of functions for which the above equality holds pointwise. One can take the set

𝒜κW(d)={fL1(μκW):κW(f)L1(μκW)},\mathcal{A}_{\kappa}^{W}(\mathbb{R}^{d})=\bigl{\{}f\in L^{1}(\mu_{\kappa}^{W}):\mathcal{F}_{\kappa}^{W}(f)\in L^{1}(\mu_{\kappa}^{W})\bigr{\}},

which is a subset of L2(μκW)L^{2}(\mu_{\kappa}^{W}) (since it is contained in the intersection of L1(μκW)L^{1}(\mu_{\kappa}^{W}) and LL^{\infty}). For f𝒜κW(d)f\in\mathcal{A}_{\kappa}^{W}(\mathbb{R}^{d}), the inversion formula allows us to write

τxWf(y)=cκWdEκW(ix,z)EκW(iy,z)κW(f)(z)dμκW(z).\tau_{x}^{W}f(y)=c_{\kappa}^{W}\int_{\mathbb{R}^{d}}E_{\kappa}^{W}(ix,z)E_{\kappa}^{W}(iy,z)\mathcal{F}_{\kappa}^{W}(f)(z)\,\mathrm{d}\mu^{W}_{\kappa}(z).

In Fourier analysis, the translation operator ff(+x)f\mapsto f(\cdot+x) (to which the Dunkl translation reduces when κ=0\kappa=0) is positive and LpL^{p}-bounded. In the Dunkl setting, τxW\tau_{x}^{W} is not a positive operator ([12, 21]) and the Lp(μκW)L^{p}(\mu_{\kappa}^{W})-boundedness is still a challenging problem, apart from the trivial case where p=2p=2 (thanks to the Plancherel theorem and the fact that |EκW(ix,y)|1|E_{\kappa}^{W}(ix,y)|\leqslant 1). The most general result we have is given in the following theorem ([17, 21]), where we denote by Lradp(μκW)L^{p}_{\mathrm{rad}}(\mu_{\kappa}^{W}) the space of radial functions in Lp(μκW)L^{p}(\mu_{\kappa}^{W}).

Theorem 2.3.

  1. (1)

    For every pp satisfying 1p21\leqslant p\leqslant 2 and for every xdx\in\mathbb{R}^{d}, the Dunkl translation τxW:Lradp(μκW)Lp(μκW)\tau_{x}^{W}:L^{p}_{\mathrm{rad}}(\mu_{\kappa}^{W})\to L^{p}(\mu_{\kappa}^{W}) is a bounded operator.

  2. (2)

    Let fL1(μκW)f\in L^{1}(\mu_{\kappa}^{W}) be a bounded, radial and positive function. Then τxWf0\tau_{x}^{W}f\geqslant 0 for every xdx\in\mathbb{R}^{d}.

The last result we mention about the Dunkl translation is the following.

Theorem 2.4.

Let fLrad1(μκW)f\in L^{1}_{\mathrm{rad}}(\mu_{\kappa}^{W}). Then, for every xdx\in\mathbb{R}^{d},

dτxWf(y)dμκW(y)=df(y)dμκW(y).\int_{\mathbb{R}^{d}}\tau_{x}^{W}f(y)\,\mathrm{d}\mu_{\kappa}^{W}(y)=\int_{\mathbb{R}^{d}}f(y)\,\mathrm{d}\mu_{\kappa}^{W}(y).

Another important tool in the Dunkl analysis is the Dunkl-type heat semi-group which has been mainly studied by Rösler ([13, 15]). We are searching for solutions u𝒞2(d×]0,+[)𝒞b(d×[0,+[)u\in\mathcal{C}^{2}(\mathbb{R}^{d}\times]0,+\infty[)\cap\mathcal{C}_{b}(\mathbb{R}^{d}\times[0,+\infty[) of the following Cauchy problem for the generalized heat equation:

(CH)κ:{ΔκWu(x,t)=tu(x,t)(x,t)d×]0,+[,u(,0)=f(CH)_{\kappa}:\begin{cases}\Delta_{\kappa}^{W}u(x,t)&=\partial_{t}u(x,t)\quad\forall(x,t)\in\mathbb{R}^{d}\times]0,+\infty[,\\ u(\cdot,0)&=f\end{cases}

with initial data ff in the Schwartz space 𝒮(d)\mathcal{S}(\mathbb{R}^{d}) and where ΔκW=j=1d(Tej)2\Delta_{\kappa}^{W}=\sum_{j=1}^{d}(T_{e_{j}}^{\mathcal{R}})^{2} is the Dunkl Laplacian. It is easily noticed that a solution of ΔκWu(x,t)=tu(x,t)\Delta_{\kappa}^{W}u(x,t)=\partial_{t}u(x,t) is given on d×]0,+[\mathbb{R}^{d}\times]0,+\infty[ by the generalized Gaussian which is defined for every xdx\in\mathbb{R}^{d} by

qtW(x)=cκW(2t)d2+γex24tq_{t}^{W}(x)=\frac{c_{\kappa}^{W}}{(2t)^{\frac{d}{2}+\gamma}}\mathrm{e}^{-\frac{\|x\|^{2}}{4t}}

and which has the following two properties.

Proposition 2.1.

  1. (1)

    For every t>0t>0,

    dqtW(x)dμκW(x)=1.\int_{\mathbb{R}^{d}}q_{t}^{W}(x)\,\mathrm{d}\mu_{\kappa}^{W}(x)=1.
  2. (2)

    For every t>0t>0 and every xdx\in\mathbb{R}^{d},

    κW(qtW)(x)=cκWetx2.\mathcal{F}_{\kappa}^{W}(q_{t}^{W})(x)=c_{\kappa}^{W}\mathrm{e}^{-t\|x\|^{2}}.

The Dunkl-type heat kernel QκWQ_{\kappa}^{W} is defined by taking the Dunkl translation of qtWq_{t}^{W}, that is, according to [13],

QκW(x,y,t)\displaystyle Q_{\kappa}^{W}(x,y,t) =τxWqtW(y)\displaystyle=\tau_{x}^{W}q_{t}^{W}(-y)
=cκW(2t)d2+γe(x2+y2)4tEκW(x2t,y2t),x,yd,t>0.\displaystyle=\frac{c_{\kappa}^{W}}{(2t)^{\frac{d}{2}+\gamma}}\mathrm{e}^{-\frac{(\|x\|^{2}+\|y\|^{2})}{4t}}E_{\kappa}^{W}\Bigl{(}\frac{x}{\sqrt{2t}},\frac{y}{\sqrt{2t}}\Bigr{)},\quad x,y\in\mathbb{R}^{d},t>0.

This positive kernel ([13]) allows us to define a generalized heat operator (or Dunkl-type heat operator). More precisely for every fLp(μκW)f\in L^{p}(\mu_{\kappa}^{W}), with 1p+1\leqslant p\leqslant+\infty, and for every t0t\geqslant 0, we set

HtWf={df(y)QκW(,y,t)dμκW(y)ift>0,fift=0.H_{t}^{W}f=\begin{cases}\int_{\mathbb{R}^{d}}f(y)Q_{\kappa}^{W}(\cdot,y,t)\,\mathrm{d}\mu_{\kappa}^{W}(y)&\text{if}\ t>0,\\ f&\text{if}\ t=0.\end{cases}

The following fundamental result about this operator is due to Rösler (see [13] and [15]).

Theorem 2.5.

For every pp satisfying 1p+1\leqslant p\leqslant+\infty, the family {HtW}t0\{H_{t}^{W}\}_{t\geqslant 0} is a positive and contraction semigroup on Lp(μκW)L^{p}(\mu_{\kappa}^{W}). Moreover, for every f𝒮(d)f\in\mathcal{S}(\mathbb{R}^{d}), the function uu given for every (x,t)d×[0,+[(x,t)\in\mathbb{R}^{d}\times[0,+\infty[ by

u(x,t)=HtWf(x)u(x,t)=H_{t}^{W}f(x)

belongs to 𝒞2(d×]0,+[)𝒞b(d×[0,+[)\mathcal{C}^{2}(\mathbb{R}^{d}\times]0,+\infty[)\cap\mathcal{C}_{b}(\mathbb{R}^{d}\times[0,+\infty[) and is a solution of the Cauchy problem (CH)κ(CH)_{\kappa}.

We can easily improve the previous theorem.

Theorem 2.6.

For every pp satisfying 1p+1\leqslant p\leqslant+\infty, the family {HtW}t0\{H_{t}^{W}\}_{t\geqslant 0} is a symmetric diffusion semigroup on Lp(μκW)L^{p}(\mu_{\kappa}^{W}), that is, a semigroup which satisfies:

  1. (1)

    HtWH_{t}^{W} is a contraction on Lp(μκW)L^{p}(\mu_{\kappa}^{W});

  2. (2)

    HtWH_{t}^{W} is symmetric, that is, self-adjoint on L2(μκW)L^{2}(\mu_{\kappa}^{W});

  3. (3)

    HtWH_{t}^{W} is positive;

  4. (4)

    HtW(1)=1H_{t}^{W}(1)=1.

The reader is referred to the book of Stein [19] for a detailed study of this kind of semigroup.

3. Behavior of the Dunkl maximal operator in the scalar case

In this section we give the proof of Theorem 1.3 for a general reflection group. For convenience, we prove each point of the theorem separately. Thus, we first establish the following result.

Theorem 3.1.

There exists a numerical constant CC such that for every fL1(μκW)f\in L^{1}(\mu_{\kappa}^{W}) and every λ>0\lambda>0,

μκW({xd:MκWf(x)>λ})Cd+2γλfW,1.\mu_{\kappa}^{W}\Bigl{(}\Bigl{\{}x\in\mathbb{R}^{d}:M_{\kappa}^{W}f(x)>\lambda\Bigr{\}}\Bigr{)}\leqslant C\frac{d+2\gamma}{\lambda}\|f\|_{W,1}.

To prove this theorem, we need two lemmas. The first one is basic calculus. Before stating it, we introduce some notation.

Notation.

We define

a(Sd1)=Sd1hκ2(x)dω(x)a(S^{d-1})=\int_{S^{d-1}}h^{2}_{\kappa}(x)\,\mathrm{d}\omega(x)

where ω\omega is the usual Lebesgue measure on Sd1S^{d-1}. We also write

qtW|Sd1=cκW(2t)d2+γe14t.{q_{t}^{W}}_{|_{S^{d-1}}}=\frac{c_{\kappa}^{W}}{(2t)^{\frac{d}{2}+\gamma}}\mathrm{e}^{-\frac{1}{4t}}.

With this notation, we can formulate the lemma.

Lemma 3.1.

We have the following equalities:

μκW(B1)=a(Sd1)d+2γ,\displaystyle\mu_{\kappa}^{W}(B_{1})=\frac{a(S^{d-1})}{d+2\gamma},
(cκW)1=2d2+γ1Γ(d2+γ)a(Sd1),\displaystyle(c_{\kappa}^{W})^{-1}=2^{\frac{d}{2}+\gamma-1}\Gamma\Bigl{(}\frac{d}{2}+\gamma\Bigr{)}a(S^{d-1}),
0+qtW|Sd1dt=cκW2d2+γ4Γ(d2+γ1).\displaystyle\int_{0}^{+\infty}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t=\frac{c_{\kappa}^{W}2^{\frac{d}{2}+\gamma}}{4}\Gamma\Bigl{(}\frac{d}{2}+\gamma-1\Bigr{)}.

Moreover if d+2γ8d+2\gamma\geqslant 8, then

1d+2γ+qtW|Sd1dtcκW2d2+γ4(d+2γ4)d2+γ1ed+2γ4.\int_{\frac{1}{d+2\gamma}}^{+\infty}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t\leqslant\frac{c_{\kappa}^{W}2^{\frac{d}{2}+\gamma}}{4}\Bigl{(}\frac{d+2\gamma}{4}\Bigr{)}^{\frac{d}{2}+\gamma-1}\mathrm{e}^{-\frac{d+2\gamma}{4}}.
Proof.

The equalities are easy to prove by passing to polar coordinates and by substitution. Therefore we only prove the inequality. By definition we have

1d+2γ+qtW|Sd1dt=cκW1d+2γ+1(2t)d2+γe14tdt,\int_{\frac{1}{d+2\gamma}}^{+\infty}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t=c_{\kappa}^{W}\int_{\frac{1}{d+2\gamma}}^{+\infty}\frac{1}{(2t)^{\frac{d}{2}+\gamma}}\mathrm{e}^{-\frac{1}{4t}}\,\mathrm{d}t,

which leads after a change of variables to

1d+2γ+qtW|Sd1dt=cκW2d2+γ40d+2γ4td2+γ2etdt.\int_{\frac{1}{d+2\gamma}}^{+\infty}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t=\frac{c_{\kappa}^{W}2^{\frac{d}{2}+\gamma}}{4}\int_{0}^{\frac{d+2\gamma}{4}}t^{\frac{d}{2}+\gamma-2}\mathrm{e}^{-t}\,\mathrm{d}t.

Since for d+2γ8d+2\gamma\geqslant 8 and t[0,d+2γ4]t\in[0,\frac{d+2\gamma}{4}] we have

td2+γ2et(d+2γ4)d2+γ2ed+2γ4,t^{\frac{d}{2}+\gamma-2}\mathrm{e}^{-t}\leqslant\Bigl{(}\frac{d+2\gamma}{4}\Bigr{)}^{\frac{d}{2}+\gamma-2}\mathrm{e}^{-\frac{d+2\gamma}{4}},

we obtain

0d+2γ4td2+γ2etdt(d+2γ4)d2+γ1ed+2γ4,\int_{0}^{\frac{d+2\gamma}{4}}t^{\frac{d}{2}+\gamma-2}\mathrm{e}^{-t}\,\mathrm{d}t\leqslant\Bigl{(}\frac{d+2\gamma}{4}\Bigr{)}^{\frac{d}{2}+\gamma-1}\mathrm{e}^{-\frac{d+2\gamma}{4}},

and the inequality is proved. ∎

The second lemma allows us to reduce the inequality of Theorem 3.1 to a more convenient one.

Lemma 3.2.

If there exists t0>0t_{0}>0 such that

1μκW(B1)C(d,κ)t00t0qtW|Sd1dt,\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant\frac{C(d,\kappa)}{t_{0}}\int_{0}^{t_{0}}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t,

then for every fL1(μκW)f\in L^{1}(\mu_{\kappa}^{W}) and every λ>0\lambda>0,

μκW({xd:MκWf(x)>λ})4C(d,κ)λfW,1,\mu_{\kappa}^{W}\Bigl{(}\Bigl{\{}x\in\mathbb{R}^{d}:M_{\kappa}^{W}f(x)>\lambda\Bigr{\}}\Bigr{)}\leqslant 4\frac{C(d,\kappa)}{\lambda}\|f\|_{W,1},

where C(d,κ)C(d,\kappa) is the same positive constant in both hypothesis and conclusion of the lemma.

Proof.

We can assume that ff is nonnegative. If there exists t0>0t_{0}>0 such that

1μκW(B1)C(d,κ)t00t0qtW|Sd1dt,\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant\frac{C(d,\kappa)}{t_{0}}\int_{0}^{t_{0}}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t,

we can deduce that for nn large enough,

(3.1) 1μκW(B1)2C(d,κ)t01nt0qtW|Sd1dt.\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant\frac{2C(d,\kappa)}{t_{0}}\int_{\frac{1}{n}}^{t_{0}}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t.

Then we claim that for every ydy\in\mathbb{R}^{d},

(3.2) 1μκW(B1)χB1(y)2C(d,κ)t01nt0qtW(y)dt.\frac{1}{\mu_{\kappa}^{W}(B_{1})}\chi_{{}_{B_{1}}}(y)\leqslant\frac{2C(d,\kappa)}{t_{0}}\int_{\frac{1}{n}}^{t_{0}}q_{t}^{W}(y)\,\mathrm{d}t.

Indeed if y>1\|y\|>1, there is nothing to do since χB1(y)=0\chi_{{}_{B_{1}}}(y)=0. If y1\|y\|\leqslant 1, it is enough to use (3.1) and the fact that qtW|Sd1qtW(y){q_{t}^{W}}_{|_{S^{d-1}}}\leqslant q_{t}^{W}(y).
As a result, we can write, for every r>0r>0 and every ydy\in\mathbb{R}^{d},

1μκW(Br)χBr(y)=1rd+2γμκW(B1)χB1(yr)2C(d,κ)rd+2γt01nt0qtW(yr)dt,\frac{1}{\mu_{\kappa}^{W}(B_{r})}\chi_{{}_{B_{r}}}(y)=\frac{1}{r^{d+2\gamma}\mu_{\kappa}^{W}(B_{1})}\chi_{{}_{B_{1}}}\Bigl{(}\frac{y}{r}\Bigr{)}\leqslant\frac{2C(d,\kappa)}{r^{d+2\gamma}t_{0}}\int_{\frac{1}{n}}^{t_{0}}q_{t}^{W}\Bigl{(}\frac{y}{r}\Bigr{)}\,\mathrm{d}t,

which leads after a change of variables to

1μκW(Br)χBr(y)2C(d,κ)r2t0r2nr2t0qtW(y)dt.\frac{1}{\mu_{\kappa}^{W}(B_{r})}\chi_{{}_{B_{r}}}(y)\leqslant\frac{2C(d,\kappa)}{r^{2}t_{0}}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(y)\,\mathrm{d}t.

Let xdx\in\mathbb{R}^{d}. By Theorem 2.3(2)(2),

1μκW(Br)τxW(χBr)(y)2C(d,κ)r2t0τxW(r2nr2t0qtW()dt)(y).\frac{1}{\mu_{\kappa}^{W}(B_{r})}\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\leqslant\frac{2C(d,\kappa)}{r^{2}t_{0}}\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}(-y).

Let us temporarily assume that

(3.3) τxW(r2nr2t0qtW()dt)(y)=r2nr2t0τxW(qtW)(y)dt.\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}(-y)=\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(-y)\,\mathrm{d}t.

This implies

1μκW(Br)τxW(χBr)(y)2C(d,κ)r2t0r2nr2t0τxW(qtW)(y)dt.\frac{1}{\mu_{\kappa}^{W}(B_{r})}\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\leqslant\frac{2C(d,\kappa)}{r^{2}t_{0}}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(-y)\,\mathrm{d}t.

Multiplying both sides by f(y)f(y) and integrating over d\mathbb{R}^{d} we obtain

1μκW(Br)df(y)τxW(χBr)(y)dμκW(y)2C(d,κ)r2t0r2nr2t0HtWf(x)dt,\frac{1}{\mu_{\kappa}^{W}(B_{r})}\int_{\mathbb{R}^{d}}f(y)\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\,\mathrm{d}\mu_{\kappa}^{W}(y)\leqslant\frac{2C(d,\kappa)}{r^{2}t_{0}}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}H_{t}^{W}f(x)\,\mathrm{d}t,

and so

1μκW(Br)df(y)τxW(χBr)(y)dμκW(y)2C(d,κ)r2t00r2t0HtWf(x)dt,\frac{1}{\mu_{\kappa}^{W}(B_{r})}\int_{\mathbb{R}^{d}}f(y)\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\,\mathrm{d}\mu_{\kappa}^{W}(y)\leqslant\frac{2C(d,\kappa)}{r^{2}t_{0}}\int_{0}^{r^{2}t_{0}}H_{t}^{W}f(x)\,\mathrm{d}t,

This easily yields

1μκW(Br)df(y)τxW(χBr)(y)dμκW(y)2C(d,κ)sups>0(1s0sHtWf(x)dt),\frac{1}{\mu_{\kappa}^{W}(B_{r})}\int_{\mathbb{R}^{d}}f(y)\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\,\mathrm{d}\mu_{\kappa}^{W}(y)\leqslant 2C(d,\kappa)\sup_{s>0}\biggl{(}\frac{1}{s}\int_{0}^{s}H_{t}^{W}f(x)\,\mathrm{d}t\biggr{)},

from which we can deduce that

MκWf(x)2C(d,κ)sups>0(1s0sHtWf(x)dt).M_{\kappa}^{W}f(x)\leqslant 2C(d,\kappa)\sup_{s>0}\biggl{(}\frac{1}{s}\int_{0}^{s}H_{t}^{W}f(x)\,\mathrm{d}t\biggr{)}.

Since {HtW}t0\{H_{t}^{W}\}_{t\geqslant 0} is a semigroup which satisfies the contraction property on Lp(μκW)L^{p}(\mu_{\kappa}^{W}) (Theorem 2.5), the first point of the Hopf-Dunford-Schwartz ergodic theorem implies the desired conclusion and we are left with the task of establishing (3.3).

Let nn be an integer. Due to Proposition 2.1, it is easily seen that the radial function r2nr2t0qtW()dt\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t is in L1(μκW)L^{1}(\mu_{\kappa}^{W}) so Theorem 2.3 leads to

τxW(r2nr2t0qtW()dt)L1(μκW).\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}\in L^{1}(\mu_{\kappa}^{W}).

Since on one hand

κW(r2nr2t0qtW()dt)=r2nr2t0cκWet2dtL1(μκW),\mathcal{F}_{\kappa}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}=\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}c_{\kappa}^{W}\mathrm{e}^{-t\|\cdot\|^{2}}\,\mathrm{d}t\in L^{1}(\mu_{\kappa}^{W}),

and on the other hand

κW(τxW(r2nr2t0qtW()dt))=EκW(ix,)κW(r2nr2t0qtW()dt),\mathcal{F}_{\kappa}^{W}\biggl{(}\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}\biggr{)}=E_{\kappa}^{W}(ix,\cdot)\,\mathcal{F}_{\kappa}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)},

we find

κW(τxW(r2nr2t0qtW()dt))L1(μκW).\mathcal{F}_{\kappa}^{W}\biggl{(}\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}\biggr{)}\in L^{1}(\mu_{\kappa}^{W}).

Consequently, τxW(r2nr2t0qtW()dt)𝒜κW(d)\tau_{x}^{W}\bigl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\bigr{)}\in\mathcal{A}_{\kappa}^{W}(\mathbb{R}^{d}) and we can use the inversion formula to obtain

τxW(r2nr2t0qtW()dt)(y)=cκWdEκW(ix,z)EκW(iy,z)r2nr2t0cκWetz2dtdμκW(z),\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}(y)=c_{\kappa}^{W}\int_{\mathbb{R}^{d}}E_{\kappa}^{W}(ix,z)E_{\kappa}^{W}(iy,z)\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}c_{\kappa}^{W}\mathrm{e}^{-t\|z\|^{2}}\,\mathrm{d}t\,\mathrm{d}\mu_{\kappa}^{W}(z),

that is,

τxW(r2nr2t0qtW()dt)(y)=r2nr2t0(cκWdEκW(ix,z)EκW(iy,z)cκWetz2dμκW(z))dt.\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}(y)\\ =\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}\biggl{(}c_{\kappa}^{W}\int_{\mathbb{R}^{d}}E_{\kappa}^{W}(ix,z)E_{\kappa}^{W}(iy,z)c_{\kappa}^{W}\mathrm{e}^{-t\|z\|^{2}}\,\mathrm{d}\mu_{\kappa}^{W}(z)\biggr{)}\,\mathrm{d}t.

The inversion formula leads to

τxW(r2nr2t0qtW()dt)(y)=r2nr2t0τxW(qtW)(y)dt,\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}(y)=\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t,

and (3.3) is established. The lemma is proved.∎

Remark.

Of course, we can take the limit of (3.3) to prove

(3.4) τxW(0r2t0qtW()dt)(y)=0r2t0τxW(qtW)(y)dt.\tau_{x}^{W}\biggl{(}\int_{0}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}(y)=\int_{0}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t.

On one hand, r2nr2t0qtW()dt0r2t0qtW()dt\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\to\int_{0}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t in L1(μκW)L^{1}(\mu_{\kappa}^{W}) as nn goes to infinity. Therefore, by Theorem 2.3(1)(1),

τxW(r2nr2t0qtW()dt)τxW(0r2t0qtW()dt)\tau_{x}^{W}\biggl{(}\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}\to\tau_{x}^{W}\biggl{(}\int_{0}^{r^{2}t_{0}}q_{t}^{W}(\cdot)\,\mathrm{d}t\biggr{)}

in L1(μκW)L^{1}(\mu_{\kappa}^{W}) as nn goes to infinity.

On the other hand

00r2t0τxW(qtW)(y)dtr2nr2t0τxW(qtW)(y)dt=0r2nτxW(qtW)(y)dt.0\leqslant\int_{0}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t-\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t=\int^{\frac{r^{2}}{n}}_{0}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t.

But

d0r2nτxW(qtW)(y)dtdμκW(y)=0r2ndτxW(qtW)(y)dμκW(y)dt=r2n\int_{\mathbb{R}^{d}}\int^{\frac{r^{2}}{n}}_{0}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t\,\mathrm{d}\mu_{\kappa}^{W}(y)=\int^{\frac{r^{2}}{n}}_{0}\int_{\mathbb{R}^{d}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}\mu_{\kappa}^{W}(y)\,\mathrm{d}t=\frac{r^{2}}{n}

since thanks to Theorem 2.4,

dτxW(qtW)(y)dμκW(y)=dqtW(y)dμκW(y)=1.\int_{\mathbb{R}^{d}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}\mu_{\kappa}^{W}(y)=\int_{\mathbb{R}^{d}}q_{t}^{W}(y)\,\mathrm{d}\mu_{\kappa}^{W}(y)=1.

Consequently

r2nr2t0τxW(qtW)(y)dt0r2t0τxW(qtW)(y)dt\int_{\frac{r^{2}}{n}}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t\to\int_{0}^{r^{2}t_{0}}\tau_{x}^{W}(q_{t}^{W})(y)\,\mathrm{d}t

in L1(μκW)L^{1}(\mu_{\kappa}^{W}) as nn goes to infinity. The equality (3.4) is thus true.

We are now in a position to prove Theorem 3.1.

Proof.

Due to Lemma 3.2, it is enough to find t0>0t_{0}>0 such that

(3.5) 1μκW(B1)Cd+2γt00t0qtW|Sd1dt,\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant C\frac{d+2\gamma}{t_{0}}\int_{0}^{t_{0}}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t,

where CC is a numerical constant. Set t0=1d+2γt_{0}=\tfrac{1}{d+2\gamma}. The inequality of Lemma 3.1 asserts that for d+2γ8d+2\gamma\geqslant 8,

1d+2γ+qtW|Sd1dtcκW2d2+γ4(d+2γ4)d2+γ1ed+2γ4.\int_{\frac{1}{d+2\gamma}}^{+\infty}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t\leqslant\frac{c_{\kappa}^{W}2^{\frac{d}{2}+\gamma}}{4}\Bigl{(}\frac{d+2\gamma}{4}\Bigr{)}^{\frac{d}{2}+\gamma-1}\mathrm{e}^{-\frac{d+2\gamma}{4}}.

On one hand, Stirling’s formula implies that

(n4)n21en4=on+(Γ(n21)),\Bigl{(}\frac{n}{4}\Bigr{)}^{\frac{n}{2}-1}\mathrm{e}^{-\frac{n}{4}}=\underset{n\to+\infty}{\mathrm{o}}\Bigl{(}\Gamma\Bigl{(}\frac{n}{2}-1\Bigr{)}\Bigr{)},

and, on the other hand, the third equality of Lemma 3.1 yields

0+qtW|Sd1dt=cκW2d2+γ4Γ(d2+γ1).\int_{0}^{+\infty}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t=\frac{c_{\kappa}^{W}2^{\frac{d}{2}+\gamma}}{4}\Gamma\Bigl{(}\frac{d}{2}+\gamma-1\Bigr{)}.

Therefore, we conclude that there exists a numerical constant CC such that

(3.6) cκW2d2+γΓ(d2+γ1)C01d+2γqtW|Sd1dt.c_{\kappa}^{W}2^{\frac{d}{2}+\gamma}\Gamma\Bigl{(}\frac{d}{2}+\gamma-1\Bigr{)}\leqslant C\int_{0}^{\frac{1}{d+2\gamma}}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t.

By the first two equalities of Lemma 3.1 we can write

cκW2d2+γΓ(d2+γ1)=2Γ(d2+γ1)μκW(B1)(d+2γ)Γ(d2+γ),c_{\kappa}^{W}2^{\frac{d}{2}+\gamma}\Gamma\Bigl{(}\frac{d}{2}+\gamma-1\Bigr{)}=\frac{2\,\Gamma\bigl{(}\frac{d}{2}+\gamma-1\bigr{)}}{\mu_{\kappa}^{W}(B_{1})(d+2\gamma)\Gamma\bigl{(}\frac{d}{2}+\gamma\bigr{)}},

and by inserting this in (3.6) we are led to

1μκW(B1)C(d+2γ)Γ(d2+γ)Γ(d2+γ1)01d+2γqtW|Sd1dt,\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant C\frac{(d+2\gamma)\,\Gamma\bigl{(}\frac{d}{2}+\gamma\bigr{)}}{\Gamma\bigl{(}\frac{d}{2}+\gamma-1\bigr{)}}\int_{0}^{\frac{1}{d+2\gamma}}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t,

which finally implies that

1μκW(B1)C(d+2γ)201d+2γqtW|Sd1dt.\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant C(d+2\gamma)^{2}\int_{0}^{\frac{1}{d+2\gamma}}{q_{t}^{W}}_{|_{S^{d-1}}}\,\mathrm{d}t.

This proves (3.5) and establishes the theorem. ∎

We now turn to Theorem 1.3(2)(2) that we recall below.

Theorem 3.2.

There exists a numerical constant CC such that for every fLp(μκW)f\in L^{p}(\mu_{\kappa}^{W}) with 1<p<+1<p<+\infty,

MκWfW,pCpp1d+2γfW,p.\bigl{\|}M_{\kappa}^{W}f\bigr{\|}_{W,p}\leqslant C\frac{p}{p-1}\sqrt{d+2\gamma}\,\|f\|_{W,p}.
Remark.

This result is better that what one would obtain by using Theorem 3.1, the LL^{\infty} case and the Marcinkiewicz interpolation theorem.

In order to prove Theorem 3.2, we need the following lemma which reduces the inequality of that theorem to a more convenient one.

Lemma 3.3.

If there exists t0>0t_{0}>0 such that

1μκW(B1)C(d,κ)qt0W|Sd1,\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant C(d,\kappa){q_{t_{0}}^{W}}_{|_{S^{d-1}}},

then there exists a numerical constant CC such that for every pp satisfying 1<p<+1<p<+\infty and every fLp(μκW)f\in L^{p}(\mu_{\kappa}^{W}),

MκWfW,pCpp1C(d,κ)fW,p,\bigl{\|}M_{\kappa}^{W}f\bigr{\|}_{W,p}\leqslant C\frac{p}{p-1}C(d,\kappa)\,\|f\|_{W,p},

where C(d,κ)C(d,\kappa) is the same positive constant in both hypothesis and conclusion of the lemma.

Proof.

The proof is quite similar to the one of Lemma 3.2. One can assume that ff is nonnegative. If there exists t0>0t_{0}>0 such that

(3.7) 1μκW(B1)C(d,κ)qt0W|Sd1,\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant C(d,\kappa){q_{t_{0}}^{W}}_{|_{S^{d-1}}},

then we can deduce that for every ydy\in\mathbb{R}^{d},

(3.8) 1μκW(B1)χB1(y)C(d,κ)qt0W(y).\frac{1}{\mu_{\kappa}^{W}(B_{1})}\chi_{{}_{B_{1}}}(y)\leqslant C(d,\kappa)q_{t_{0}}^{W}(y).

Indeed, if y>1\|y\|>1, it is obvious since χB1(y)=0\chi_{{}_{B_{1}}}(y)=0. If y1\|y\|\leqslant 1, it is enough to use (3.7) and the fact that qt0W|Sd1qt0W(y){q_{t_{0}}^{W}}_{|_{S^{d-1}}}\leqslant q_{t_{0}}^{W}(y).

Consequently, for every r>0r>0, we can write

1μκW(Br)χBr(y)\displaystyle\frac{1}{\mu_{\kappa}^{W}(B_{r})}\chi_{{}_{B_{r}}}(y) =1rd+2γμκW(B1)χB1(yr)\displaystyle=\frac{1}{r^{d+2\gamma}\mu_{\kappa}^{W}(B_{1})}\chi_{{}_{B_{1}}}\Bigl{(}\frac{y}{r}\Bigr{)}
C(d,κ)rd+2γqt0W(yr)=C(d,κ)qr2t0W(y).\displaystyle\leqslant\frac{C(d,\kappa)}{r^{d+2\gamma}}q_{t_{0}}^{W}\Bigl{(}\frac{y}{r}\Bigr{)}=C(d,\kappa)q_{r^{2}t_{0}}^{W}(y).

Let xdx\in\mathbb{R}^{d}. Due to Theorem 2.3(2)(2) we can assert that

1μκW(Br)τxW(χBr)(y)C(d,κ)τxW(qr2t0W,κ)(y).\frac{1}{\mu_{\kappa}^{W}(B_{r})}\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\leqslant C(d,\kappa)\tau_{x}^{W}(q^{W,\kappa}_{r^{2}t_{0}})(-y).

Multiplying both sides by f(y)f(y) and integrating over d\mathbb{R}^{d} we obtain

1μκW(Br)df(y)τxW(χBr)(y)dμκW(y)C(d,κ)Hr2t0W,κf(x),\frac{1}{\mu_{\kappa}^{W}(B_{r})}\int_{\mathbb{R}^{d}}f(y)\tau_{x}^{W}(\chi_{{}_{B_{r}}})(-y)\,\mathrm{d}\mu_{\kappa}^{W}(y)\leqslant C(d,\kappa)H^{W,\kappa}_{r^{2}t_{0}}f(x),

from which we deduce

(3.9) MκWf(x)C(d,κ)supt>0HtWf(x).M_{\kappa}^{W}f(x)\leqslant C(d,\kappa)\sup_{t>0}H_{t}^{W}f(x).

But {HtW}t0\{H^{W}_{t}\}_{t\geqslant 0} is a symmetric diffusion semigroup on Lp(μκW)L^{p}(\mu_{\kappa}^{W}) (Theorem 2.6). Therefore, by a result due to Stein (see [19, chapter 4]), for every pp satisfying 1<p<+1<p<+\infty,

supt>0HtWfW,pCpp1fW,p,\bigl{\|}\sup_{t>0}H_{t}^{W}f\bigr{\|}_{W,p}\leqslant C\frac{p}{p-1}\|f\|_{W,p},

where CC is a numerical constant. By using this inequality in (3.9) we obtain the desired result. ∎

We can now turn to the proof of Theorem 3.2.

Proof.

Thanks to the previous lemma, it is enough to find t0>0t_{0}>0 such that

1μκW(B1)Cd+2γqt0W|Sd1\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant C\sqrt{d+2\gamma}\,{q_{t_{0}}^{W}}_{|_{S^{d-1}}}

or, equivalently, such that

1μκW(B1)CcκWd+2γ(12t0)d2+γe14t0.\frac{1}{\mu_{\kappa}^{W}(B_{1})}\leqslant Cc_{\kappa}^{W}\sqrt{d+2\gamma}\Bigl{(}\frac{1}{2t_{0}}\Bigr{)}^{\frac{d}{2}+\gamma}\mathrm{e}^{-\frac{1}{4t_{0}}}.

On one hand, the first two equalities of Lemma 3.1 allow us to write

1μκW(B1)=cκW(d+2γ)2d2+γ1Γ(d2+γ),\frac{1}{\mu_{\kappa}^{W}(B_{1})}=c_{\kappa}^{W}(d+2\gamma)2^{\frac{d}{2}+\gamma-1}\Gamma\Bigl{(}\frac{d}{2}+\gamma\Bigr{)},

and on the other hand, Stirling’s formula gives us

2n2Γ(n2)=On+(nn12en2).2^{\frac{n}{2}}\Gamma\Bigl{(}\frac{n}{2}\Bigr{)}=\underset{n\to+\infty}{\mathrm{O}}\bigl{(}n^{\frac{n-1}{2}}\mathrm{e}^{-\frac{n}{2}}\bigr{)}.

We finally obtain the desired result by choosing t0=12d+4γt_{0}=\tfrac{1}{2d+4\gamma}. ∎

4. Exponential integrability in the vector-valued case

This section is devoted to the proof of Theorem 1.5 in the case where the reflection group is 2d\mathbb{Z}_{2}^{d} (that is, the root system is ={±ej:1jd}\mathcal{R}=\{\pm e_{j}:1\leqslant j\leqslant d\}). Let us recall some facts related to Dunkl’s analysis associated with this particular reflection group.

In this case, an explicit formula of the intertwining operator Vκ2dV_{\kappa}^{\mathbb{Z}_{2}^{d}} is known (see [22]) and there is an explicit formula for the Dunkl translation. In the one-dimensional case, the following formula has been proved by Rösler in the setting of signed hypergroups (see [12]):

(4.1) τx2f(y)=1211f(x2+y2+2xyt)(1+x+yx2+y2+2xyt)ψκ(t)dt+1211f(x2+y2+2xyt)(1x+yx2+y2+2xyt)ψκ(t)dt\tau^{\mathbb{Z}_{2}}_{x}f(y)=\frac{1}{2}\int_{-1}^{1}f\Bigl{(}\sqrt{x^{2}+y^{2}+2xyt}\Bigr{)}\biggl{(}1+\frac{x+y}{\sqrt{x^{2}+y^{2}+2xyt}}\biggr{)}\psi_{\kappa}(t)\,\mathrm{d}t\\ +\frac{1}{2}\int_{-1}^{1}f\Bigl{(}-\sqrt{x^{2}+y^{2}+2xyt}\Bigr{)}\biggl{(}1-\frac{x+y}{\sqrt{x^{2}+y^{2}+2xyt}}\biggr{)}\psi_{\kappa}(t)\,\mathrm{d}t

where ψ\psi is given by ψκ(t)=(B(κ,12))1(1+t)(1t2)κ1\psi_{\kappa}(t)=\bigl{(}B\bigl{(}\kappa,\frac{1}{2}\bigr{)}\bigr{)}^{-1}(1+t)(1-t^{2})^{\kappa-1} (with BB the beta function) and where κ\kappa is the only value taken by the multiplicity function κ\kappa. This formula implies an explicit one in the case 2d\mathbb{Z}_{2}^{d} which gives us the boundedness of the Dunkl translation. In order to give an equivalent formula to (4.1), we need to introduce some notation.

Notation.

  1. (1)

    For x,y,zx,y,z\in\mathbb{R}, we put

    σx,y,z={12xy(x2+y2z2)if x,y0,0if x=0 or y=0,\sigma_{x,y,z}=\begin{cases}\frac{1}{2xy}(x^{2}+y^{2}-z^{2})&\text{if }x,y\neq 0,\\ 0&\text{if }x=0\text{ or }y=0,\end{cases}

    as well as

    ϱ(x,y,z)=12(1σx,y,z+σz,x,y+σz,y,x).\varrho(x,y,z)=\frac{1}{2}(1-\sigma_{x,y,z}+\sigma_{z,x,y}+\sigma_{z,y,x}).
  2. (2)

    For x,y,z>0x,y,z>0, we put

    Kκ(x,y,z)=22κ2(B(κ,12))1Δ(x,y,z)2κ2(xyz)2κ1χ[|xy|,x+y](z),K_{\kappa}(x,y,z)=2^{2\kappa-2}\Bigl{(}B\Bigl{(}\kappa,\frac{1}{2}\Bigr{)}\Bigr{)}^{-1}\frac{\Delta(x,y,z)^{2\kappa-2}}{(xyz)^{2\kappa-1}}\chi_{{}_{[|x-y|,x+y]}}(z),

    where Δ(x,y,z)\Delta(x,y,z) denotes the area of the triangle (perhaps degenerated) with sides x,y,zx,y,z.

With this notation, (4.1) can be reformulated (using a change of variables) as follows:

(4.2) τx2f(y)=f(z)dνx,y2(z)\tau^{\mathbb{Z}_{2}}_{x}f(y)=\int_{\mathbb{R}}f(z)\,\mathrm{d}\nu^{\mathbb{Z}_{2}}_{x,y}(z)

where the measure νx,y2\nu^{\mathbb{Z}_{2}}_{x,y} is given by

dνx,y2(z)={𝒦κ(x,y,z)dμκ2(z)if x,y0,dδx(z)if y=0,dδy(z)if x=0,\mathrm{d}\nu^{\mathbb{Z}_{2}}_{x,y}(z)=\begin{cases}\mathcal{K}_{\kappa}(x,y,z)\,\mathrm{d}\mu^{\mathbb{Z}_{2}}_{\kappa}(z)&\text{if }x,y\neq 0,\\ \mathrm{d}\delta_{x}(z)&\text{if }y=0,\\ \mathrm{d}\delta_{y}(z)&\text{if }x=0,\end{cases}

with

𝒦κ(x,y,z)=Kκ(|x|,|y|,|z|)ϱ(x,y,z).\mathcal{K}_{\kappa}(x,y,z)=K_{\kappa}\bigl{(}|x|,|y|,|z|\bigr{)}\varrho(x,y,z).

We have the following one-dimensional result due to Rösler, .

Theorem 4.1.

The measure νx,y2\nu^{\mathbb{Z}_{2}}_{x,y} satisfies

  1. (1)

    suppνx,y2=[|x||y|,||x||y||][||x||y||,|x|+|y|]\mathrm{supp}\,\nu^{\mathbb{Z}_{2}}_{x,y}=\Bigl{[}-|x|-|y|,-\bigl{|}|x|-|y|\bigr{|}\Bigr{]}\bigcup\Bigl{[}\bigl{|}|x|-|y|\bigr{|},|x|+|y|\Bigr{]} for x,y0x,y\neq 0.

  2. (2)

    νx,y2()=1\nu^{\mathbb{Z}_{2}}_{x,y}(\mathbb{R})=1 and νx,y2=|dνx,y2|4\|\nu^{\mathbb{Z}_{2}}_{x,y}\|=\int_{\mathbb{R}}|\mathrm{d}\nu^{\mathbb{Z}_{2}}_{x,y}|\leqslant 4, for x,yx,y\in\mathbb{R}.

We now prove that no analogue of Theorem 1.4(2)(2) holds when p=+p=+\infty.

Proposition 4.1.

The conclusion of the second point of Theorem 1.4 does not hold when p=+p=+\infty.

Proof.

Let d=1d=1 and let f=(fn)n1f=(f_{n})_{n\geqslant 1} where fn=χ[2n1,2n[f_{n}=\chi_{{}_{[2^{n-1},2^{n}[}} for every n1n\geqslant 1. We easily see that

fr=χ[1,+[L(μκ2)\|f\|_{\ell^{r}}=\chi_{{}_{[1,+\infty[}}\in L^{\infty}(\mu_{\kappa}^{\mathbb{Z}_{2}})

while we will prove that

(4.3) κ2frL(μκ2).\|\mathcal{M}^{\mathbb{Z}_{2}}_{\kappa}f\|_{\ell^{r}}\notin L^{\infty}(\mu_{\kappa}^{\mathbb{Z}_{2}}).

For all xx\in\mathbb{R} and n1n\geqslant 1, by definition of Mκ2χ[2n1,2n[M^{\mathbb{Z}_{2}}_{\kappa}\chi_{{}_{[2^{n-1},2^{n}[}},

Mκ2χ[2n1,2n[(x)1μκ2(]|x|2n,|x|+2n[)2n12nτx2(χ]|x|2n,|x|+2n[)(y)dμκ2(y).M^{\mathbb{Z}_{2}}_{\kappa}\chi_{{}_{[2^{n-1},2^{n}[}}(x)\\ \geqslant\frac{1}{\mu_{\kappa}^{\mathbb{Z}_{2}}(]-|x|-2^{n},|x|+2^{n}[)}\int_{2^{n-1}}^{2^{n}}\tau^{\mathbb{Z}_{2}}_{x}(\chi_{{}_{]-|x|-2^{n},|x|+2^{n}[}})(-y)\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}}(y).

We claim that for every y[2n1,2n[y\in[2^{n-1},2^{n}[,

(4.4) τx2(χ]|x|2n,|x|+2n[)(y)=1.\tau^{\mathbb{Z}_{2}}_{x}(\chi_{{}_{]-|x|-2^{n},|x|+2^{n}[}})(-y)=1.

Indeed, by using (4.2) we can write

222κB(κ,12)τx2(χ]|x|2n,|x|+2n[)(y)=χ]|x|2n,|x|+2n[(z)χ]||x||y||,|x|+|y|[(|z|)Δ(|x|,|y|,|z|)2κ2(|xyz|)2κ1ϱ(x,y,z)dμκ2(z),2^{2-2\kappa}B\Bigl{(}\kappa,\frac{1}{2}\Bigr{)}\tau^{\mathbb{Z}_{2}}_{x}(\chi_{{}_{]-|x|-2^{n},|x|+2^{n}[}})(-y)\\ =\int_{\mathbb{R}}\chi_{{}_{]-|x|-2^{n},|x|+2^{n}[}}(z)\chi_{{}_{]||x|-|y||,|x|+|y|[}}(|z|)\frac{\Delta(|x|,|y|,|z|)^{2\kappa-2}}{(|xyz|)^{2\kappa-1}}\varrho{(x,-y,z)}\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}}(z),

and, since |x|+|y|<|x|+2n|x|+|y|<|x|+2^{n}, we deduce that

222κB(κ,12)τx2(χ]|x|2n,|x|+2n[)(y)=χ]||x||y||,|x|+|y|[(|z|)Δ(|x|,|y|,|z|)2κ2(|xyz|)2κ1ϱ(x,y,z)dμκ2(z),2^{2-2\kappa}B\Bigl{(}\kappa,\frac{1}{2}\Bigr{)}\tau^{\mathbb{Z}_{2}}_{x}(\chi_{{}_{]-|x|-2^{n},|x|+2^{n}[}})(-y)\\ =\int_{\mathbb{R}}\chi_{{}_{]||x|-|y||,|x|+|y|[}}(|z|)\frac{\Delta(|x|,|y|,|z|)^{2\kappa-2}}{(|xyz|)^{2\kappa-1}}\varrho{(x,-y,z)}\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}}(z),

that is,

τx2(χ]|x|2n,|x|+2n[)(y)=νx,y2().\tau^{\mathbb{Z}_{2}}_{x}(\chi_{{}_{]-|x|-2^{n},|x|+2^{n}[}})(-y)=\nu^{\mathbb{Z}_{2}}_{x,-y}(\mathbb{R}).

We obtain (4.4) by using Theorem 4.1. As a result we have the inequality

Mκ2χ[2n1,2n[(x)1μκ2(]|x|2n,|x|+2n[)2n12ndμκ2(y),M^{\mathbb{Z}_{2}}_{\kappa}\chi_{{}_{[2^{n-1},2^{n}[}}(x)\geqslant\frac{1}{\mu_{\kappa}^{\mathbb{Z}_{2}}(]-|x|-2^{n},|x|+2^{n}[)}\int_{2^{n-1}}^{2^{n}}\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}}(y),

and since dμκ2(y)=|y|2κdy\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}}(y)=|y|^{2\kappa}\mathrm{d}y, we are led to

Mκ2χ[2n1,2n[(x)2n(2κ+1)2(n1)(2κ+1)2(|x|+2n)2κ+1.M^{\mathbb{Z}_{2}}_{\kappa}\chi_{{}_{[2^{n-1},2^{n}[}}(x)\geqslant\frac{2^{n(2\kappa+1)}-2^{(n-1)(2\kappa+1)}}{2(|x|+2^{n})^{2\kappa+1}}.

For every xx satisfying |x|2n|x|\leqslant 2^{n} we thus obtain

Mκ2χ[2n1,2n[(x)2n(2κ+1)2(n1)(2κ+1)2(n+1)(2κ+1)+1,M^{\mathbb{Z}_{2}}_{\kappa}\chi_{{}_{[2^{n-1},2^{n}[}}(x)\geqslant\frac{2^{n(2\kappa+1)}-2^{(n-1)(2\kappa+1)}}{2^{(n+1)(2\kappa+1)+1}},

and after simplifications we get

Mκ2χ[2n1,2n[(x)12(2κ+1)22κ+2.M^{\mathbb{Z}_{2}}_{\kappa}\chi_{{}_{[2^{n-1},2^{n}[}}(x)\geqslant\frac{1-2^{-(2\kappa+1)}}{2^{2\kappa+2}}.

Thus, for every xx\in\mathbb{R},

κ2frr{n:2n|x|}(12(2κ+1)22κ+2)r=+,\|\mathcal{M}^{\mathbb{Z}_{2}}_{\kappa}f\|_{\ell^{r}}^{r}\geqslant\sum_{\{n:2^{n}\geqslant|x|\}}\Bigl{(}\frac{1-2^{-(2\kappa+1)}}{2^{2\kappa+2}}\Bigr{)}^{r}=+\infty,

and the proof of (4.3) is finished. ∎

A substitute result in the case p=+p=+\infty is given in Theorem 1.5 that we now prove. In order to do that, we need three lemmas. The first one is just a trivial functional equality.

Lemma 4.1.

Let XX be a measure space and let mm be a positive measure on XX. Let φ\varphi be an increasing continuously differentiable function on [0,+[[0,+\infty[ which satisfies φ(0)=0\varphi(0)=0. Then for every measurable function ff on XX,

Xφ(|f(x)|)dm(x)=0+φ(λ)m({xX:|f(x)|>λ})dλ.\int_{X}\varphi\bigl{(}|f(x)|\bigr{)}\,\mathrm{d}m(x)=\int_{0}^{+\infty}\varphi^{\prime}(\lambda)m\Bigl{(}\Bigl{\{}x\in X:|f(x)|>\lambda\Bigr{\}}\Bigr{)}\,\mathrm{d}\lambda.
Proof.

Since φ\varphi^{\prime} is nonnegative, the Fubini theorem yields

0+φ(λ)m({xX:|f(x)|>λ})dλ=X(0|f(x)|φ(λ)dλ)dm(x).\int_{0}^{+\infty}\varphi^{\prime}(\lambda)m\Bigl{(}\Bigl{\{}x\in X:|f(x)|>\lambda\Bigr{\}}\Bigr{)}\,\mathrm{d}\lambda=\int_{X}\biggl{(}\int_{0}^{|f(x)|}\varphi^{\prime}(\lambda)\,\mathrm{d}\lambda\biggr{)}\,\mathrm{d}m(x).

The desired result follows by integrating and using the fact that φ(0)=0\varphi(0)=0. ∎

The second lemma gives us the behavior of the constant of Theorem 1.4(2)(2) when pp grows and rr is fixed. The result of exponential integrability is closely related to this behavior.

Lemma 4.2.

The constant C(d,κ,p,r)C(d,\kappa,p,r) of the vector-valued maximal theorem is such that

C(d,κ,p,r)=Op+(p1r).C(d,\kappa,p,r)=\underset{p\to+\infty}{O}\bigl{(}p^{\frac{1}{r}}\bigr{)}.
Proof.

Since the parameter rr is fixed, it is enough to consider the proof of Theorem 1.4 when p>rp>r. As explained in [2], once we have constructed the operator Mκ2d,RM_{\kappa}^{\mathbb{Z}_{2}^{d},R} (see [2] for the definition) and shown that it satisfies a scalar maximal theorem and a weighted inequality, we can follow almost verbatim the proof of the vector-valued maximal theorem for the Hardy-Littlewood maximal operator (see [10] or [20]). Thus, it is easily seen that the dependence on pp is given by the constant (with exponent 1r\tfrac{1}{r}) of the maximal theorem for Mκ2d,RM_{\kappa}^{\mathbb{Z}_{2}^{d},R} and for the space Lppr(μκ2d)L^{\frac{p}{p-r}}(\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}). Since this constant is obtained by interpolation, we can write

C(d,κ,p,r)=C(d,κ,r)(C(d,κ)pprppr1)prpr.C(d,\kappa,p,r)=C(d,\kappa,r)\biggl{(}C(d,\kappa)\frac{\frac{p}{p-r}}{\frac{p}{p-r}-1}\biggr{)}^{\frac{p-r}{pr}}.

But it is obvious that

(C(d,κ)pprppr1)prp=Op+(p),\biggl{(}C(d,\kappa)\frac{\frac{p}{p-r}}{\frac{p}{p-r}-1}\biggr{)}^{\frac{p-r}{p}}=\underset{p\to+\infty}{O}\bigl{(}p\bigr{)},

and the lemma is proved. ∎

The last lemma gives us (under the hypothesis of Theorem 1.5) a sharp estimate of the measure of the set {xK:κ2df(x)rr>λ}\{x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\} for every λ>0\lambda>0, where KK denotes a compact subset of d\mathbb{R}^{d}. More precisely we have the following inequality.

Lemma 4.3.

Let f=(fn)n1f=(f_{n})_{n\geqslant 1} be a sequence of measurable functions defined on d\mathbb{R}^{d} and let rr satisfy 1<r<+1<r<+\infty. If frL(μκ2d)\|f\|_{\ell^{r}}\in L^{\infty}(\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}) is such that

μκ2d(suppfrr)<+,\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)}<+\infty,

then for every compact subset KK of d\mathbb{R}^{d} and every λ>0\lambda>0,

μκ2d({xK:κ2df(x)rr>λ})max{2μκ2d(K);μκ2d(suppfrr)}e(log(2)2Cfrr2d,)λ,\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\\ \leqslant\max\Bigl{\{}2\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(K);\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)}\Bigr{\}}\mathrm{e}^{-\Bigl{(}\frac{\log(2)}{2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}}\Bigr{)}\lambda},

where C=C(d,κ,r)C=C(d,\kappa,r) is independent of (fn)n1(f_{n})_{n\geqslant 1}, KK and λ\lambda.

Proof.

For every pp satisfying 1p<+1\leqslant p<+\infty, the Chebyshev inequality yields

μκ2d({xK:κ2df(x)rr>λ})1λp{xK:κ2df(x)rr>λ}κ2df(x)rprdμκ2d(x),\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\\ \leqslant\frac{1}{\lambda^{p}}\int_{\bigl{\{}x\in K:\,\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\bigr{\}}}\|\mathcal{M}^{\mathbb{Z}^{d}_{2}}_{\kappa}f(x)\|_{\ell^{r}}^{pr}\,\mathrm{d}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(x),

which implies, by enlarging the domain of integration,

μκ2d({xK:κ2df(x)rr>λ})1λpdκ2df(x)rprdμκ2d(x).\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\leqslant\frac{1}{\lambda^{p}}\int_{\mathbb{R}^{d}}\|\mathcal{M}^{\mathbb{Z}^{d}_{2}}_{\kappa}f(x)\|_{\ell^{r}}^{pr}\,\mathrm{d}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(x).

By applying the vector-valued maximal theorem for κ2d\mathcal{M}^{\mathbb{Z}^{d}_{2}}_{\kappa} we get

μκ2d({xK:κ2df(x)rr>λ})(C(d,κ,pr,r))prλpdf(x)rprdμκ2d(x),\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\leqslant\frac{\bigl{(}C(d,\kappa,pr,r)\bigr{)}^{pr}}{\lambda^{p}}\int_{\mathbb{R}^{d}}\|f(x)\|_{\ell^{r}}^{pr}\,\mathrm{d}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(x),

where C=C(d,κ,pr,r)C=C(d,\kappa,pr,r) is the constant of Theorem 1.4 (and which is independent of (fn)n1(f_{n})_{n\geqslant 1}, KK and λ\lambda). Thanks to Lemma 4.2 we are lead to

μκ2d({xK:κ2df(x)rr>λ})Cppλpdf(x)rprdμκ2d(x),\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\leqslant\frac{Cp^{p}}{\lambda^{p}}\int_{\mathbb{R}^{d}}\|f(x)\|_{\ell^{r}}^{pr}\,\mathrm{d}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(x),

where C=C(d,κ,r)C=C(d,\kappa,r) is independent of (fn)n1(f_{n})_{n\geqslant 1}, KK, λ\lambda and pp. The hypothesis of the lemma allows us to write

(4.5) μκ2d({xK:κ2df(x)rr>λ})(Cpfrr2d,λ)pμκ2d(suppfrr).\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\\ \leqslant\biggl{(}\frac{Cp\,\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}}{\lambda}\biggr{)}^{p}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)}.

We now exploit (4.5) by choosing pp in terms of λ\lambda. If λ2Cfrr2d,\lambda\geqslant 2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}, we put

p=λ2Cfrr2d,1.p=\frac{\lambda}{2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}}\geqslant 1.

Then (4.5) can be reformulated as follows:

μκ2d({xK:κ2df(x)rr>λ})(12)pμκ2d(suppfrr),\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\leqslant\Bigl{(}\frac{1}{2}\Bigr{)}^{p}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)},

that is,

(4.6) μκ2d({xK:κ2df(x)rr>λ})e(log(2)2Cfrr2d,)λμκ2d(suppfrr).\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\leqslant\mathrm{e}^{-\Bigl{(}\frac{\log(2)}{2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}}\Bigr{)}\lambda}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)}.

If λ2Cfrr2d,\lambda\leqslant 2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}, we immediately obtain

μκ2d({xK:κ2df(x)rr>λ})μκ2d(K).\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\leqslant\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(K).

Since

e(log(2)2Cfrr2d,)λ12,\mathrm{e}^{-\Bigl{(}\frac{\log(2)}{2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}}\Bigr{)}\lambda}\geqslant\frac{1}{2},

we find that

(4.7) μκ2d({xK:κ2df(x)rr>λ})2μκ2d(K)e(log(2)2Cfrr2d,)λ.\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\leqslant 2\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(K)\mathrm{e}^{-\Bigl{(}\frac{\log(2)}{2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}}\Bigr{)}\lambda}.

The desired result is then a trivial consequence of (4.6) and (4.7). ∎

We are now in a position to prove Theorem 1.5.

Proof.

Let KK be a compact subset of d\mathbb{R}^{d} and let ε\varepsilon be a real number which satisfies

0ε<log(2)2Cd,κ,rfrr2d,,0\leqslant\varepsilon<\frac{\log(2)}{2C_{d,\kappa,r}\bigl{\|}\,\|f\|_{\ell^{r}}^{r}\bigr{\|}_{\mathbb{Z}_{2}^{d},\infty}},

where Cd,κ,rC_{d,\kappa,r} is the constant of Lemma 4.3. We first write

Keεκ2df(x)rrdμκ2d(x)=μκ2d(K)+K(eεκ2df(x)rr1)dμκ2d(x).\int_{K}\mathrm{e}^{\varepsilon\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}}\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(x)=\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(K)+\int_{K}\Bigl{(}\mathrm{e}^{\varepsilon\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}}-1\Bigr{)}\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(x).

We then apply the equality of Lemma 4.1 to the function φ:teεt1\varphi:t\mapsto\mathrm{e}^{\varepsilon t}-1 to obtain

Keεκ2df(x)rrdμκ2d(x)=μκ2d(K)+ε0+eελμκ2d({xK:κ2df(x)rr>λ})dλ.\int_{K}\mathrm{e}^{\varepsilon\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}}\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(x)\\ =\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(K)+\varepsilon\int_{0}^{+\infty}\mathrm{e}^{\varepsilon\lambda}\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\Bigl{\{}x\in K:\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}>\lambda\Bigr{\}}\Bigr{)}\,\mathrm{d}\lambda.

Thanks to Lemma 4.3, we are led to

Keεκ2df(x)rrdμκ2d(x)μκ2d(K)+εmax{2μκ2d(K);μκ2d(suppfrr)}0+eλ(εlog(2)2Cfrr2d,)dλ.\int_{K}\mathrm{e}^{\varepsilon\|\mathcal{M}_{\kappa}^{\mathbb{Z}_{2}^{d}}f(x)\|_{\ell^{r}}^{r}}\,\mathrm{d}\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(x)\\ \leqslant\mu_{\kappa}^{\mathbb{Z}_{2}^{d}}(K)+\varepsilon\max\Bigl{\{}2\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}(K);\mu^{\mathbb{Z}^{d}_{2}}_{\kappa}\Bigl{(}\mathrm{supp}\|f\|_{\ell^{r}}^{r}\Bigr{)}\Bigr{\}}\int_{0}^{+\infty}\mathrm{e}^{\lambda\Bigl{(}\varepsilon-\frac{\log(2)}{2C\|\|f\|_{\ell^{r}}^{r}\|_{\mathbb{Z}_{2}^{d},\infty}}\Bigr{)}}\,\mathrm{d}\lambda.

The condition on ε\varepsilon and an integration allow us to conclude. ∎

References

  • [1] Marcel de Jeu. The Dunkl transform. Invent. Math., 113(1):147–162, 1993.
  • [2] Luc Deleaval. Fefferman-Stein inequalities for the Z2dZ^{d}_{2} Dunkl maximal operator. J. Math. Anal. Appl., 360(2):711–726, 2009.
  • [3] Nelson Dunford and Jacob T. Schwartz. Linear Operators. I. General Theory. With the assistance of W. G. Bade and R. G. Bartle. Pure and Applied Mathematics, Vol. 7. Interscience Publishers, Inc., New York, 1958.
  • [4] Charles F. Dunkl. Differential-difference operators associated to reflection groups. Trans. Amer. Math. Soc., 311(1):167–183, 1989.
  • [5] Charles F. Dunkl. Integral kernels with reflection group invariance. Canad. J. Math., 43(6):1213–1227, 1991.
  • [6] Charles F. Dunkl. Hankel transforms associated to finite reflection groups. In Hypergeometric functions on domains of positivity, Jack polynomials, and applications (Tampa, FL, 1991), volume 138 of Contemp. Math., pages 123–138. Amer. Math. Soc., Providence, RI, 1992.
  • [7] Charles F. Dunkl. Intertwining operators associated to the group S3S_{3}. Trans. Amer. Math. Soc., 347(9):3347–3374, 1995.
  • [8] Charles F. Dunkl. An intertwining operator for the group B2B_{2}. Glasg. Math. J., 49(2):291–319, 2007.
  • [9] Charles F. Dunkl and Yuan Xu. Orthogonal polynomials of several variables, volume 81 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 2001.
  • [10] C. Fefferman and E. M. Stein. Some maximal inequalities. Amer. J. Math., 93:107–115, 1971.
  • [11] E. M. Opdam. Dunkl operators, Bessel functions and the discriminant of a finite Coxeter group. Compositio Math., 85(3):333–373, 1993.
  • [12] Margit Rösler. Bessel-type signed hypergroups on 𝐑{\bf R}. In Probability measures on groups and related structures, XI (Oberwolfach, 1994), pages 292–304. World Sci. Publ., River Edge, NJ, 1995.
  • [13] Margit Rösler. Generalized Hermite polynomials and the heat equation for Dunkl operators. Comm. Math. Phys., 192(3):519–542, 1998.
  • [14] Margit Rösler. Positivity of Dunkl’s intertwining operator. Duke Math. J., 98(3):445–463, 1999.
  • [15] Margit Rösler. One-parameter semigroups related to abstract quantum models of Calogero type. In Infinite dimensional harmonic analysis (Kyoto, 1999), pages 290–305. Gräbner, Altendorf, 2000.
  • [16] Margit Rösler. Dunkl operators: theory and applications. In Orthogonal polynomials and special functions (Leuven, 2002), volume 1817 of Lecture Notes in Math., pages 93–135. Springer, Berlin, 2003.
  • [17] Margit Rösler. A positive radial product formula for the Dunkl kernel. Trans. Amer. Math. Soc., 355(6):2413–2438 (electronic), 2003.
  • [18] E. M. Stein and J.-O. Strömberg. Behavior of maximal functions in 𝐑n{\bf R}^{n} for large nn. Ark. Mat., 21(2):259–269, 1983.
  • [19] Elias M. Stein. Topics in harmonic analysis related to the Littlewood-Paley theory. Annals of Mathematics Studies, No. 63. Princeton University Press, Princeton, N.J., 1970.
  • [20] Elias M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993.
  • [21] Sundaram Thangavelu and Yuan Xu. Convolution operator and maximal function for the Dunkl transform. J. Anal. Math., 97:25–55, 2005.
  • [22] Yuan Xu. Orthogonal polynomials for a family of product weight functions on the spheres. Canad. J. Math., 49(1):175–192, 1997.