Home Lp-Fourier analysis for certain differential-reflection operators
Article Publicly Available

Lp-Fourier analysis for certain differential-reflection operators

  • Salem Ben Said EMAIL logo , Asma Boussen and Mohamed Sifi
Published/Copyright: November 6, 2016

Abstract

In a previous paper we introduced a family of differential-reflection operators ΛA,ε acting on smooth functions defined on , where the spectral problem for the operators ΛA,ε has been studied. Here A is a Sturm–Liouville function with additional hypotheses and -1ε1. Via the eigenfunctions of ΛA,ε, we introduce in this paper a generalized Fourier transform A,ε. An Lp-harmonic analysis for A,ε is developed when 0<p21+1-ε2. In particular, an Lp-Schwartz space isomorphism theorem for A,ε is proved.

MSC 2010: 43A32; 43A15

1 Introduction

In [2] (see also [3]) we considered some aspects of harmonic analysis associated to the following family of one-dimensional (A,ε)-operators:

(1.1)ΛA,εf(x)=f(x)+A(x)A(x)(f(x)-f(-x)2)-εϱf(-x),

where A is a so-called Chébli function on (i.e. A is a continuous +-valued function on satisfying certain regularity and convexity hypotheses), ϱ is the index of A and -1ε1. We note that ϱ is non-negative. The function A and the real number ε are the deformation parameters giving three well-known cases (as special examples) when one of the following conditions holds:

  1. A(x)=Aα(x)=|x|2α+1 and ε arbitrary (Dunkl’s operators [15]),

  2. A(x)=Aα,β(x)=|sinhx|2α+1(coshx)2β+1 and ε=0 (Heckman’s operators [19]),

  3. A(x)=Aα,β(x)=|sinhx|2α+1(coshx)2β+1 and ε=1 (Cherednik’s operators [12]).

We mention that a differential-reflection operator built in terms of a Chébli function was built for the first time in [21].

In [2] we proved that for λ the equation

ΛA,εf(x)=iλf(x),

where f:, admits a unique solution ΨA,ε(λ,) satisfying ΨA,ε(λ,0)=1. Moreover, we established in [2, Theorems 3.4 and 3.5] suitable estimates for the growth of the eigenfunction ΨA,ε(λ,x) and of its partial derivatives.

In this paper we are concerned with a development of an Lp-harmonic analysis for a generalized Fourier transform A,ε when 0<p21+1-ε2. Here

A,εf(λ)=f(x)ΨA,ε(λ,-x)A(x)𝑑x

for fL1(,A(x)dx).

Using the estimates for the growth of ΨA,ε(λ,x), we get holomorphic properties of A,ε on the space Lp(,A(x)dx) with

(1.2){p=1for ϱ=0 or [ϱ>0 and ε=0],1p<21+1-ε2for ϱ>0 and ε[-1,1]{0}.

A Riemann–Lebesgue lemma is also obtained for p is as in (1.2).

We then turn our attention to an Lp-Schwartz space isomorphism theorem for A,ε. In [18] Harish-Chandra proved an L2-Schwartz space isomorphism for the spherical Fourier transform on non-compact Riemannian symmetric spaces. This result was extended to Lp-Schwartz spaces with 0<p<2 by Trombi and Varadarajan [26] (see also [16, 17, 13]). In the early nineties, Anker gave a new and simple proof of their result, based on the Paley–Wiener theorem for the spherical Fourier transform on Riemannian symmetric spaces [1]. Recently, Anker’s method was used in [22] to prove an Lp-Schwartz space isomorphism theorem for the Heckman–Opdam hypergeometric functions. Our approach is inspired by Anker’s paper [1]. More precisely, for -1ε1 and 0<p21+1-ε2, we define the tube domain

p,ε:={λ|Imλ|ϱ((2/p)-1-1-ε2)}.

Denote by 𝒮p() the Lp-Schwartz space on , and by 𝒮(p,ε) the Schwartz space on the tube domain p,ε. We prove that A,ε is a topological isomorphism between 𝒮p() and 𝒮(p,ε) with

{0<p1for ϱ=0 or [ϱ>0 and ε=0],0<p<21+1-ε2for ϱ>0 and ε[-1,1]{0}

(see Theorem 3.12).

We close this paper by establishing a result in connection with pointwise multipliers of 𝒮(p,ε). More precisely, for arbitrary γ0, a function ψ defined on the tube domain γ:={λ|Imλ|γ} is called a pointwise multiplier of 𝒮(γ) if the mapping ϕψϕ is continuous from 𝒮(γ) into itself. In [4] Betancor et al. characterized the set of pointwise multipliers of the Schwartz spaces 𝒮(γ) for arbitrary γ0. Under the assumptions

{0<p1for ϱ=0 or [ϱ>0 and ε=0],22+1-ε2p<21+1-ε2for ϱ>0 and ε[-1,1]{0},

we proved that if T is in the dual space 𝒮p() of 𝒮p() such that ψ:=A,ε(T) is a pointwise multiplier of 𝒮(p,ε), then for any s there exist and continuous functions fm defined on , m=0,1,,, such that

T=m=0ΛA,εmfm

and, for every such m,

supx(|x|+1)se(2p-1-ε2)ϱ|x||fm(x)|<.

The organization of this paper is as follows: In Section 2 we recapitulate some definitions and basic notations, as well as some results from literature. Further, we recall from [2] some results on ΨA,ε(λ,x). In Section 3 we develop an Lp-harmonic analysis for the Fourier transform A,ε, where we mainly prove an Lp-Schwartz space isomorphism theorem for A,ε. Finally, in Section 4 we characterize the distributions T𝒮p() so that A,ε(T) are pointwise multipliers of the Schwartz space 𝒮(p,ε).

2 Background

2.1 The Chébli transform

In this subsection we present results from [10, 11, 24, 25].

Throughout this paper we will denote by A a function on satisfying the following hypotheses:

  1. A(x)=|x|2α+1B(x), where α>-12 and B is any even, positive and smooth function on with B(0)=1.

  2. A is increasing and unbounded on +.

  3. A/A is a decreasing and smooth function on +*, and hence the limit 2ϱ:=limx+A(x)/A(x) exists and is non-negative.

Such a function A is called a Chébli function (it is also called a Chébli–Trimèche function). From (H1) it follows that

(2.1)A(x)A(x)=2α+1x+C(x),x0,

where C:=B/B is an odd and smooth function on .

Let ΔA, or simply Δ, be the following second order differential operator:

Δ=d2dx2+A(x)A(x)ddx.

For μ we consider the Cauchy problem

(2.2)Δf(x)=-(μ2+ϱ2)f(x)with f(0)=1 and f(0)=0.

In [11] the author proved that system (2.2) admits a unique solution φμ. For every μ the solution φμ is an even smooth function on and the map μφμ(x) is analytic. The following Laplace-type representation of φμ can be found in [11] (see also [24]).

Lemma 2.1

For every xR* there is a non-negative even continuous function K(|x|,) supported in [-|x|,|x|] such that for all μC we have

φμ(x)=0|x|K(|x|,t)cos(μt)𝑑t.

The following estimates of the eigenfunctions φμ can be found in [10, 11, 25].

Lemma 2.2

Let μC such that |Imμ|ϱ. Then we have the following results:

  1. φ±iϱ(x)=1, φ-μ(x)=φμ(x) and |φμ(x)|1.

  2. e-ϱ|x|φ0(x)c(|x|+1)e-ϱ|x|.

  3. |φμ(x)|φiImμ(x)e|Imμ||x|φ0(x).

  4. |φμ(x)|c(ϱ2+|μ|2)e|Imμ||x|φ0(x).

The Chébli transform of fL1(+,A(x)dx) is defined by

(2.3)Δ(f)(μ):=+f(x)φμ(x)A(x)𝑑x.

The following Plancherel and inversion formulas for Δ are proved in [11].

Theorem 2.3

There exists a unique positive measure π with support R+ such that FΔ induces an isometric isomorphism from L2(R+,A(x)dx) onto L2(R+,π(dμ)), and for any fL1L2(R+,A(x)dx) we have

+|f(x)|2A(x)𝑑x=+|Δ(f)(μ)|2π(dμ).

The inverse transform is given by

(2.4)Δ-1g(x)=+g(μ)φμ(x)π(dμ).

To have a nice behavior of the Plancherel measure π we must add a further (growth) restriction on the function A. Following [24], we will assume that A/A satisfies the following additional hypothesis:

  1. There exists a constant δ>0 such that for all x[x0,) (for some x0>0), we have

    A(x)A(x)={2ϱ+e-δxD(x)if ϱ>0,2α+1x+e-δxD(x)if ϱ=0,

    with D being a smooth function bounded together with its derivatives.

In these circumstances the Plancherel measure π is absolutely continuous with respect to the Lebesgue measure and has density |c(μ)|-2, where c is a continuous function on + and zero free on +* (see [5]). Moreover, by [25, Proposition 6.1.12 and Corollary 6.1.5] (see also [8]), for μ we have the following:

  1. If ϱ0 and α>-1/2, then |c(μ)|-2|μ|2α+1 whenever |μ|1.

  2. If ϱ>0 and α>-1/2, then |c(μ)|-2|μ|2 whenever |μ|1.

  3. If ϱ=0 and α>0, then |c(μ)|-2|μ|2α+1 whenever |μ|1.

In the literature, the function c is called Harish-Chandra’s function of the operator Δ. We refer to [7] for more details on the c-function.

Henceforth we will assume that Chébli’s function A satisfies the additional hypothesis (H4). It follows that for |x| large enough we have the following results:

  1. A(x)=O(|x|2α+1) for ϱ=0.

  2. A(x)=O(e2ϱ|x|) for ϱ>0.

We close this section by giving some basic results of (the analogue of) the Abel transform associated to the second order differential operator Δ.

Denote by 𝒟e() the space of even and compactly supported functions in C(). In [24] the author has proved that the Abel transform defined on 𝒟e() by

(2.5)𝒜f(y)=12|x|>|y|K(|x|,y)f(x)A(x)𝑑x

is an automorphism of 𝒟e() satisfying

Δ=euc𝒜,

where euc is the Euclidean Fourier transform.

2.2 A family of differential-reflection operators

Recall from (1.1) the differential-reflection operator ΛA,ε. In view of (2.1) and the hypothesis (H4) on A/A, the space 𝒟() (of smooth functions with compact support on ) and the space 𝒮() (of Schwartz functions on ) are invariant under the action of ΛA,ε.

Let S denote the symmetry (Sf)(x):=f(-x). For fC() such that

supx(1+|x|)res|x||f(t)(x)|<

for every r,t and for some 2ϱs<, and for gC() such that g and all its derivatives are at most of polynomial growth we have

(2.6)ΛA,εf(x)g(x)A(x)𝑑x=-f(x)(ΛA,ε+2εϱS)g(x)A(x)𝑑x.

Let λ and consider the initial data problem

(2.7)ΛA,εf(x)=iλf(x),f(0)=1,

where f: is a C1-function.

Theorem 2.4

Theorem 2.4 (see [2, Theorem 3.2])

For all λC, there exists a unique solution ΨA,ε(λ,) to problem (2.7). Further, for every xR, the function λΨA,ε(λ,x) is analytic on C. More explicitly:

  1. For iλεϱ, we have

    (2.8)ΨA,ε(λ,x)=φμε(x)+1iλ-εϱφμε(x),where με2:=λ2+(ε2-1)ϱ2.
  2. For iλ=εϱ, we have

    ΨA,ε(λ,x)=1+2εϱsg(x)A(x)0|x|A(t)𝑑t.

The following theorem contains estimates for the growth of the eigenfunction ΨA,ε.

Theorem 2.5

Theorem 2.5 (see [2, Theorem 3.4])

Suppose that -1ε1 and xR. Then we have the following results:

  1. For λ we have |ΨA,ε(λ,x)|2.

  2. For λ=a+ib we have |ΨA,ε(λ,x)|ΨA,ε(0,x)e|b||x|.

  3. For λ=0 we distinguish the following two cases:

    1. For ε=0 we have ΨA,0(0,x)=1.

    2. For ε0 there is a constant cε>0 such that

      ΨA,ε(0,x)cε(|x|+1)e-ϱ(1-1-ε2)|x|.

Remark 2.6

By [2, Theorem 4.10] we may replace the first statement above by |ΨA,ε(λ,x)|1 for all λ.

We remind the reader that φiϱ(x)=1 and

φi1-ε2ϱ(x)c(|x|+1)e-ϱ(1-1-ε2)|x|;

see Lemma 2.2.

Theorem 2.7

Theorem 2.7 (see [2, Theorem 3.5])

We have the following results:

  1. Assume that λ and |x|x0 for some x0>0. Given N, there is a positive constant c such that

    |NxNΨA,ε(λ,x)|c(|λ|+1)Ne|Imλ||x|φi1-ε2ϱ(x).
  2. Assume that λ and x. Given M, there is a positive constant c such that

    |MλMΨA,ε(λ,x)|c|x|Me|Imλ||x|φi1-ε2ϱ(x).

2.3 Intertwining operators

First, let us recall the following Laplace-type representation of the eigenfunction ΨA,ε(λ,) which can be found in [2, Corollary 4.4, Theorem 4.10].

Theorem 2.8

For every xR*, there is a non-negative continuous function Kε(x,) supported in [-|x|,|x|] such that for all λC, we have

(2.9)ΨA,ε(λ,x)=|y|<|x|𝕂ε(x,y)eiλy𝑑y.

For fC() we define VA,εf by

VA,εf(x)=|y|<|x|𝕂ε(x,y)f(y)𝑑yfor x0 and VA,εf(0)=f(0).

Observe that

ΨA,ε(λ,x)=VA,ε(eiλ)(x).
Theorem 2.9

Theorem 2.9 (see [2, Theorem 4.7])

The operator VA,ε is the unique automorphism of C(R) such that

ΛA,εVA,ε=VA,εddx,

where ΛA,ε is the differential-reflection operator (1.1).

Below we will deal with the dual operator VA,εt of VA,ε in the sense that

VA,εf(x)g(x)A(x)𝑑x=f(y)VA,εtg(y)𝑑y.

This can be written as

(2.10)VA,εtg(y)=|x|>|y|𝕂ε(x,y)g(x)A(x)𝑑x.

Denote by Ce() the space of even functions in C(). For fCe() we set

εf(x):=f(x)-ϱε|x|2|y|<|x|f(y)J1(ϱεx2-y2)x2-y2𝑑y,

where J1 is the Bessel function of the first kind and ϱε:=1-ε2ϱ. If ε=±1, then ϱ±1=0, and therefore ±1=id. By [23, Proposition 2.1] (see also [20, Theorem 5.1]) the integral transform ε is an automorphism of Ce(), and its inverse transform is given by

ε-1f(x)=f(x)+ϱε|x|2|y|<|x|f(y)I1(ϱεx2-y2)x2-y2𝑑y,

where I1 is the modified Bessel function of the first kind.

Let 𝒟e() be the space of even functions in 𝒟(). For g𝒟e() put

(2.11)εtg(y):=g(y)-ϱε2|x|>|y||x|g(x)J1(ϱεx2-y2)x2-y2𝑑x.

Theorem 2.10

Theorem 2.10 (see [2, Theorem 4.2])

The transform integral Eεt is an automorphism of De(R), and its inverse transform is given by

ε-1tg(y)=g(y)+ϱε2|x|>|y||x|g(x)I1(ϱεx2-y2)x2-y2𝑑x.

Theorem 2.11

Theorem 2.11 (see [2, Lemma 4.8, Theorem 4.9])

The dual operator VA,εt from (2.10) satisfies the following properties:

  1. The operator VA,εt is the unique automorphism of 𝒟() satisfying the intertwining property

    ddyVA,εt=VA,εt(ΛA,ε+2εϱS),

    where S denotes the symmetry (Sf)(x):=f(-x).

  2. The operator VA,εt can be expressed as

    VA,εtg(y)=ε-1t𝒜(ge)(y)-(εϱ-ddy)ε-1t𝒜(Jgo)(y),

    where Jh(x):=-xh(t)𝑑t and 𝒜 is the Abel transform (2.5).

3 Fourier transform of Lp-Schwartz spaces

Assume that -1ε1. For fL1(,A(x)dx) put

A,εf(λ)=f(x)ΨA,ε(λ,-x)A(x)𝑑x.

To state its alleged inverse transform, let us introduce the following Plancherel measure:

(3.1)πε(dλ)=|λ|λ2-(1-ε2)ϱ2|c(λ2-(1-ε2)ϱ2)|2𝟙]-1-ε2ϱ,1-ε2ϱ[(λ)dλ,

where c is Harish-Chandra’s function associated to the second order differential operator Δ (see Section 2 for more details on the c-function). In the sequel we set fˇ(x):=f(-x).

Theorem 3.1

Let f be a smooth function with compact support on R. Then we have the following formulas:

  1. (Inversion formula)

    (3.2)f(x)=14A,ε(f)(λ)ΨA,ε(λ,x)(1-εϱiλ)πε(dλ)almost everywhere.
  2. (Plancherel formula)

    (3.3)|f(x)|2A(x)𝑑x=14A,ε(f)(λ)A,ε(fˇ)(-λ)¯(1-εϱiλ)πε(dλ).

    We may rewrite (3.3) for two smooth and compactly supported functions f and g as

    f(x)g(-x)A(x)𝑑x=14A,ε(f)(λ)A,ε(g)(λ)(1-εϱiλ)πε(dλ).

Proof.

For the sake of completeness, we provide a detailed proof. (i) We set Jh(x):=-xh(t)𝑑t. Using the superposition (2.8) of the eigenfunction Ψε(λ,x), we obtain

A,εf(λ)=2Δ(fe)(με)+2(iλ+εϱ)Δ(Jfo)(με),

where FΔ is the Chébli transform (2.3) and με is so that με2=λ2+(ε2-1)ϱ2. By the inversion formula (2.4) for Δ we deduce that

(3.4)f(x)=+{Δ(fe)(με)φμε(x)+Δ(Jfo)(με)φμε(x)}π(dμε).

Let us write φμε and φμε in terms of ΨA,ε as follows:

φμε(x)=12(ΨA,ε(-λ,-x)+ΨA,ε(-λ,x)),φμε(x)=iλ+εϱ2(ΨA,ε(-λ,-x)-ΨA,ε(-λ,x)).

Consequently, formula (3.4) becomes

f(x)=12+ΨA,ε(-λ,-x){Δ(fe)(με)+(iλ+εϱ)Δ(Jfo)(με)}π(dμε)
+12+ΨA,ε(-λ,x){Δ(fe)(με)-(iλ+εϱ)Δ(Jfo)(με)}π(dμε)
(3.5)=14+{ΨA,ε(-λ,-x)A,ε(f)(λ)+ΨA,ε(-λ,x)A,ε(fˇ)(λ)}πε(dλ).

Further, it is easy to check that

(3.6)ΨA,ε(λ,x)=(1+εϱiλ)ΨA,ε(-λ,-x)-εϱiλΨA,ε(λ,-x),

and therefore

(3.7)A,ε(fˇ)(λ)=(1+εϱiλ)A,ε(f)(-λ)-εϱiλA,ε(f)(λ).

In view of (3.6) and (3.7) we obtain

+A,ε(f)(λ)ΨA,ε(-λ,-x)πε(dλ)=+ΨA,ε(λ,x)A,ε(f)(λ)(1-εϱiλ)πε(dλ)
(3.8)++ΨA,ε(-λ,x)A,ε(f)(λ)(εϱiλ)πε(dλ),

and

+A,ε(fˇ)(λ)ΨA,ε(-λ,x)πε(dλ)=+ΨA,ε(-λ,x)A,ε(f)(-λ)(1+εϱiλ)πε(dλ)
(3.9)++ΨA,ε(-λ,x)A,ε(f)(λ)(-εϱiλ)πε(dλ).

Substituting (3.8) and (3.9) into (3.5), we get the inversion formula (3.2). (ii) Using the fact that ΨA,ε(λ,x)¯=ΨA,ε(-λ,x) for λ, we have

A,ε(gˇ)(λ)¯=g(x)¯Ψε(-λ,x)A(x)𝑑x.

Application of the identity (3.5) for f gives

(3.10)f(x)g(x)¯A(x)𝑑x=14+{A,ε(f)(λ)A,ε(g)(λ)¯+A,ε(fˇ)(λ)A,ε(gˇ)(λ)¯}πε(dλ).

Further, from (3.6) it follows that

(3.11)A,ε(fˇ)(-λ)¯=(1+εϱiλ)A,ε(f)(λ)¯-εϱiλA,ε(f)(-λ)¯.

Therefore,

(3.12)(1-εϱiλ)A,ε(f)(λ)A,ε(fˇ)(-λ)¯=(1+ε2ϱ2λ2)|A,ε(f)(λ)|2-εϱiλ(1-εϱiλ)A,ε(f)(λ)A,ε(f)(-λ)¯.

Now let us rewrite (3.11) as

A,ε(f)(-λ)¯=iλiλ-εϱA,ε(fˇ)(λ)¯+εϱ-iλ+εϱA,ε(f)(λ)¯.

Hence

A,ε(f)(λ)A,ε(f)(-λ)¯=iλiλ-εϱA,ε(f)(λ)A,ε(fˇ)(λ)¯+εϱ-iλ+εϱ|A,ε(f)(λ)|2,

which implies

(-εϱiλ)(iλ-εϱiλ)A,ε(f)(λ)A,ε(f)(-λ)¯=-(εϱiλ)A,ε(f)(λ)A,ε(fˇ)(λ)¯-(ε2ϱ2λ2)|A,ε(f)(λ)|2.

Thus (3.12) becomes

(3.13)(1-εϱiλ)A,ε(f)(λ)A,ε(fˇ)(-λ)¯=|A,ε(f)(λ)|2-εϱiλA,ε(f)(λ)A,ε(fˇ)(λ)¯,

which is the key identity towards the Plancherel formula (3.3). Moreover, from (3.13) we also have

(3.14)(1-εϱiλ)A,ε(fˇ)(-λ)¯A,ε(f)(λ)=|A,ε(fˇ)(-λ)|2-εϱiλA,ε(fˇ)(-λ)¯A,ε(f)(-λ).

Indeed, we obtain (3.14) in three steps:

  1. Replace f by fˇ in (3.13).

  2. Substitute λ by -λ in the resulting identity from step 1.

  3. Take the complex conjugates in the resulting identity from step 2.

Putting the pieces together, we arrive at

A,ε(f)(λ)A,ε(fˇ)(-λ)¯(1-εϱiλ)πε(dλ)
=+|A,ε(f)(λ)|2πε(dλ)-+A,ε(f)(λ)A,ε(fˇ)(λ)¯(εϱiλ)πε(dλ)
   +-|A,ε(fˇ)(-λ)|2πε(dλ)--A,ε(f)(-λ)A,ε(fˇ)(-λ)¯(εϱiλ)πε(dλ)
=+{|A,ε(f)(λ)|2+|A,ε(fˇ)(λ)|2}πε(dλ),

which compares very well with 4fL2(,A(x)dx)2 (see (3.10)). ∎

Remarks 3.2

Notice that:

  1. For ε=1, the Plancherel formula (3.3) corrects [9, Theorem 5.13] (stated without a proof).

  2. For ε=0 we can prove the following (stronger) versions of the inversion and the Plancherel formulas:

    1. If fL1(,A(x)dx) and A,0(f)L1(,π0(dλ)), then

      f(x)=14A,0(f)(λ)ΨA,0(λ,x)π0(dλ)almost everywhere.
    2. If fL1L2(,A(x)dx), then

      A,0fL2(,π0(dλ))andA,0fL2(,π0(dλ))=2fL2(,A(x)dx).
    3. There exists a unique isometry on L2(,A(x)dx) that coincides with (1/2)A,0 on L1L2(,A(x)dx).

The following lemma will be needed in the proof of a Paley–Wiener theorem for A,ε.

Lemma 3.3

For R>0 denote by DR(R) the space of smooth functions with support inside [-R,R]. Then fDR(R) if and only if VA,εtfDR(R).

Proof.

The direct statement follows from (2.10). The converse direction is more involved. On the one hand, one can prove that

VA,ε-1tg(y)=𝒜-1εt(ge)(y)+(εϱ+ddy)𝒜-1εt(Jgo)(y),

where g=ge+go and Jh(x):=-xh(t)𝑑t. On the other hand, from (2.11) and [6, Lemma 4.10] it follows that if ge𝒟R(), then 𝒜-1εt(ge)𝒟R(). Further, one may check that go𝒟R() if and only if Jgo𝒟R(). As Jgo is an even function, it follows from above that

𝒜-1εt(Jgo)𝒟R().

For R>0, let PWR() be the space of entire functions h on which are of exponential type and rapidly decreasing, i.e. for all t we have

supλ(|λ|+1)te-R|Imλ||h(λ)|<.

Denote by PW() the union of the spaces PWR(), R>0.

Theorem 3.4

Theorem 3.4 (Paley–Wiener Theorem)

Assume that -1ε1. The Fourier transform FA,ε is a topological linear isomorphism between D(R) and PW(C). More precisely, it is a topological linear isomorphism between DR(R) and PWR(C) for all R>0.

Proof.

The proof is standard. We shall only indicate how to proceed towards the statement. On the one hand, the Fourier transform A,ε can be written as A,ε(f)=eucVA,εt(fˇ), where euc is the Euclidean Fourier transform and VA,εt is the intertwining operator (2.10). This is a direct consequence of (2.9). Now, in the light of Lemma 3.3, appealing to the Paley–Wiener theorem for the Euclidean Fourier transform euc we get the desired statement. ∎

For -1ε1 and 0<p21+1-ε2 set ϑp,ε:=2p-1-1-ε2. Observe that 121+1-ε22. We introduce the tube domain

p,ε:={λ|Imλ|ϱϑp,ε}.
Proposition 3.5

For all λC1,ε and for all xR we have |ΨA,ε(λ,x)|2.

Proof.

Let R>0 be arbitrary but fixed and let

1,ε:={ν|ν|ϱ(1-1-ε2)}.

Application of the maximum modulus principle together with the fact that |ΨA,ε(λ,x)|ΨA,ε(iImλ,x) in the domain [-R,R]+i1,ε implies that the maximum of |ΨA,ε(λ,x)| is obtained when λ belongs to the boundary of i1,ε, that is λ=iη with |η|=ϱ(1-1-ε2). Now recall that ΨA,ε(iη,x)+ΨA,ε(iη,-x)=2φμε(x) when ε0,±1, and ΨA,ε(iη,x)+ΨA,ε(iη,-x)=2 when ε=0,±1. Further, the parameter με satisfies

με2=λ2-(1-ε2)ϱ2=-ϱ2(1-21-ε2{1-1-ε2})0,

and therefore μεi with |με|ϱ. Using the fact that ΨA,ε(iη,x)>0 for all x, together with the fact that φμε(x)1 for με as above (see Lemma 2.2), we obtain that ΨA,ε(iη,x)2 for all x and -1ε1. ∎

Corollary 3.6

Assume that ϱ0 and ε[-1,1]. Let fL1(R,A(x)dx). Then the following properties hold:

  1. The Fourier transform A,ε(f)(λ) is well defined for all λ1,ε. Moreover,

    |A,ε(f)(λ)|2fL1(,A(x)dx)for all λ1,ε.
  2. The function A,ε(f) is holomorphic on ̊1,ε, the interior of 1,ε.

  3. (Riemann–Lebesgue lemma.) We have

    (3.15)limλ1,ε,|λ||A,ε(f)(λ)|=0.

Proof.

The first two statements are direct consequences of Proposition 3.5, the fact that ΨA,ε(λ,) is holomorphic in λ, and Morera’s theorem. For the Riemann–Lebesgue lemma, a classical proof for the Euclidean Fourier transform carries over. More precisely, assume that f𝒟() (the space of smooth functions with compact support on ). Now use the Paley–Wiener Theorem 3.4 to conclude that the limit (3.15) holds for test functions; the general case then follows from the fact that 𝒟() is dense in L1(,A(x)dx). ∎

Next we discuss some properties of the Fourier transform A,ε on Lp(,A(x)dx) with p>1. Observe that the set ̊p,ε, the interior of p,ε in , is empty whenever ϱ=0 or p=21+1-ε2. Observe also that when ε=0, we have 21+1-ε2=1.

Lemma 3.7

Assume that ϱ>0, and fLp(R,A(x)dx) with 1<p<21+1-ε2 and ε[-1,1]{0}. Then the following properties hold:

  1. The Fourier transform A,ε(f)(λ) is well defined for all λ in ̊p,ε. Moreover,

    |A,ε(f)(λ)|cfLp(,A(x)dx)for all λ̊p,ε.
  2. The function A,ε(f) is holomorphic on ̊p,ε.

  3. (Riemann–Lebesgue lemma.) We have

    limλ̊p,ε,|λ||A,ε(f)(λ)|=0.

Proof.

The first two statements follow easily from the estimate

|ΨA,ε(λ,x)|c(|x|+1)e|Imλ||x|e-ϱ|x|(1-1-ε2)for ϱ>0,

the fact that A(x)c|x|βe2ϱ|x| for some β>0 (a consequence of hypothesis (H4) on the function A), the fact that ΨA,ε(λ,) is holomorphic in λ, and Morera’s theorem. The Riemann–Lebesgue lemma is established exactly as for (3.15) by approximating any function in the space Lp(,A(x)dx) by compactly supported smooth functions. ∎

Henceforth, when ϱ=0, we will assume that the parameter α appearing in the hypothesis (H1) is strictly positive. This is due to the fact that when ϱ=0, estimates for the c-function |c(μ)| for |μ|1 are evaluable only when α>0 (see Section 2).

Theorem 3.8

Under the same assumptions as in Corollary 3.6 and Lemma 3.7, the Fourier transform FA,ε is injective on Lp(R,A(x)dx).

Proof.

Take q such that p+q=pq. For fLp(,A(x)dx) and g𝒟() we have

|(f,g)A|:=|f(x)g(-x)A(x)𝑑x|fLp(,A(x)dx)gLq(,A(x)dx)

and

|(A,ε(f),A,ε(g))πε|:=|A,ε(f)(λ)A,ε(g)(λ)(1-εϱiλ)πε(dλ)|
supλ|A,ε(f)(λ)|A,ε(g)L1(,π~ε(dλ))
(3.16)cfLp(,A(x)dx)A,ε(g)L1(,π~ε(dλ)),

where π~ε(dλ):=|1-εϱiλ|πε(dλ). Above we have used Corollary 3.6 and Lemma 3.7 to get (3.16). Therefore, the mappings f(f,g)A and f(A,ε(f),A,ε(g))πε are continuous functionals on Lp(,A(x)dx). Now,

(f,g)A=(1/4)(A,ε(f),A,ε(g))πε

for all f𝒟() (see Theorem 3.1) and by continuity for all fLp(,A(x)dx). Assume that fLp(,A(x)dx) and that A,ε(f)=0. Then for all g𝒟() we have (f,g)A=(1/4)(A,ε(f),A,ε(g))πε=0 and therefore f=0. ∎

For -1ε1 and 0<p21+1-ε2 let 𝒮p() be the space consisting of all functions fC() such that

supx(|x|+1)sφ0(x)-2/p|f(k)(x)|<

for any s and any k. The topology of 𝒮p() is defined by the seminorms

σs,k(p)(f)=supx(|x|+1)sφ0(x)-2/p|f(k)(x)|.

We pin down that 𝒮p() is a dense subspace of Lq(,A(x)dx) for all qp.

The following facts are standard; see for instance [14, Appendix A].

Lemma 3.9

  1. 𝒮p() is a Fréchet space with respect to the seminorms σs,k(p).

  2. 𝒟() is a dense subspace of 𝒮p().

Recall from above the tube domain

p,ε:={λ|Imλ|ϱϑp,ε},

where ϑp,ε=2p-1-1-ε2.

The Schwartz space 𝒮(p,ε) consists of all complex valued functions h that are analytic in the interior of p,ε and such that h and all its derivatives extend continuously to p,ε and satisfy

supλp,ε(|λ|+1)t|h()(λ)|<

for any t and any . The topology of 𝒮(p,ε) is defined by the seminorms

(3.17)τt,(ϑp,ε)(h):=supλp,ε(|λ|+1)t|h()(λ)|.

For ϑp,ε=0 or ϱ=0, the space 𝒮(p,ε) is the classical Schwartz space on . By [6, Lemma 4.17] the Paley–Wiener space PW() is dense in the Schwartz space 𝒮(p,ε).

Lemma 3.10

The Fourier transform FA,ε maps Sp(R) continuously into S(Cp,ε) and is injective with

{0<p1for ϱ=0 or [ϱ>0 and ε=0],0<p<21+1-ε2for ϱ>0 and ε[-1,1]{0}.

Proof.

Let f𝒮p(). For λp,ε we have

|A,ε(f)(λ)||f(x)||ΨA,ε(λ,-x)|A(x)𝑑x
|f(x)|φ0(x)-2/pφ0(x)2/pΨA,ε(0,-x)e|Imλ||x|A(x)𝑑x
c1|f(x)|φ0(x)-2/p(|x|+1)2/p+1e-2ϱ|x|A(x)𝑑x.

Under hypothesis (H4) on Chébli’s function A, there exists a β>0 such that A(x)c|x|βe2ϱ|x|. Hence,

|A,ε(f)(λ)|c2|f(x)|φ0(x)-2/p(|x|+1)2/p+1|x|β𝑑x<.

This proves that A,ε(f) is well defined for all f𝒮p() when -1ε1 and 0<p21+1-ε2. Moreover, since the map λΨA,ε(λ,x) is holomorphic on , it follows that for all f𝒮p(), the function A,ε(f) is analytic in the interior of p,ε and continuous on p,ε. Furthermore, by Theorem 2.7 (ii), we have

|A,ε(f)(λ)(k)|c3|f(x)|φ0(x)-2/p(|x|+1)2/p+k+1|x|β𝑑x<.

Thus, all derivatives of A,ε(f) extend continuously to p,ε. Next, we will prove that given a continuous seminorm τ on 𝒮(p,ε), there exists a continuous seminorm σ on 𝒮p() such that

τ(A,ε(f))c4σ(f)for all f𝒮p().

Note that the space 𝒮(p,ε) and its topology are also determined by the seminorms

(3.18)hτ~t,(ϑp,ε)(h):=supλp,ε|{(λ+1)th(λ)}()|,

where t and are two arbitrary positive integers. Invoking the identity (2.6), we have

(iλ)rA,ε(f)(λ)=(iλ)rfˇ(x)ΨA,ε(λ,x)A(x)𝑑x=fˇ(x)ΛA,εrΨA,ε(λ,x)A(x)𝑑x
=(-1)r(ΛA,ε+2εϱS)rfˇ(x)ΨA,ε(λ,x)A(x)𝑑x=ΛA,εrf(-x)ΨA,ε(λ,x)A(x)𝑑x
=A,ε(ΛA,εrf)(λ)

for r, where S denotes the symmetry Sf(x)=f(-x). Above we have used (ΛA,ε+2εϱS)rS=(-1)rSΛA,εr. Thus,

{(iλ)rA,ε(f)(λ)}()=ΛA,εrf(x)λΨA,ε(λ,-x)A(x)𝑑x.

On the one hand, using Theorem 2.7 (ii), we obtain

|{(iλ)rA,ε(f)(λ)}()|c5|ΛA,εrf(x)|(|x|+1)φi1-ε2ϱ(x)e|Imλ||x|A(x)𝑑x
=c5|x|a|ΛA,εrf(x)|(|x|+1)φi1-ε2ϱ(x)e|Imλ||x|A(x)𝑑x
+c5|x|>a|ΛA,εrf(x)|(|x|+1)φi1-ε2ϱ(x)e|Imλ||x|A(x)𝑑x
c6|x|a|ΛA,εrf(x)|φ0(x)-2/p(|x|+1)2/p++1e-2ϱ|x|A(x)𝑑x
+c6|x|>a|ΛA,εrf(x)|φ0(x)-2/p(|x|+1)2/p++1e-2ϱ|x|A(x)𝑑x.

On the other hand, by mimicking the proof of [6, Lemma 4.18], the following holds:

  1. For |x|a we have

    |ΛA,εrf(x)|c7(i=0r|f(i)(x)|+i=0r-1|f(i)(-x)|+i=0rm=0nr|f(i)(ξm)|),

    where ξm=ξm(x,r)]-|x|,|x|[.

  2. For |x|>a we have

    |ΛA,εrf(x)|c7(i=0r|f(i)(x)|+i=0r-1|f(i)(-x)|).

The estimate

τ(A,ε(f))c8finiteσ(f)for all f𝒮p()

is now a matter of putting the pieces together.

We obtain the injectivity of the transform A,ε on 𝒮p() from the facts that A,ε is injective on Lq(,A(x)dx) for a certain range of q (see Theorem 3.8) and that 𝒮p() is a dense subspace of Lq(,A(x)dx) for qp.

This concludes the proof of Lemma 3.10. ∎

Lemma 3.11

Let -1ε1 and 0<p21+1-ε2. The inverse Fourier transform FA,ε-1:PW(C)D(R) given by

A,ε-1h(x)=14h(λ)ΨA,ε(λ,x)(1-εϱiλ)πε(dλ)

is continuous for the topologies induced by S(Cp,ε) and Sp(R).

Proof.

Let f𝒟() and hPW() so that f=A,ε-1(h). Given a seminorm σ on 𝒮p(), we should find a continuous seminorm τ on 𝒮(p,ε) such that σ(f)cτ(h).

Denote by g the image of h by the inverse Euclidean Fourier transform euc-1. Making use of the Paley–Wiener Theorem 3.4 for A,ε and the classical Paley–Wiener theorem for euc, we have the following support conservation property:

supp(f)IR:=[-R,R]supp(g)IR.

For j1 let ωjC() with ωj=0 on Ij-1 and ωj=1 outside of Ij. Assume that ωj and all its derivatives are bounded uniformly in j. We will write gj=ωjg, and define hj:=euc(gj) and fj:=A,ε-1(hj). Note that gj=g outside Ij. Hence, by the above support property, fj=f outside Ij. We shall estimate the function

x(|x|+1)sφ0(x)-2/p|fj(k)(x)|

on Ij+1Ij with j1. Recall that fj=f on Ij+1Ij. In view of Theorem 2.7 (i) we have

|fj(k)(x)|14|hj(λ)||xkΨA,ε(λ,x)||1-εϱiλ|πε(dλ)
c1φi1-ε2ϱ(x)|hj(λ)|(|λ|+1)k|1-εϱiλ|πε(dλ),

where

|1-εϱiλ|πε(dλ)=λ2+ε2ϱ2λ2-(1-ε2)ϱ21|c(λ2-(1-ε2)ϱ2)|2𝟙]-1-ε2ϱ,1-ε2ϱ[(λ)dλ.

By knowing about the asymptotic behavior of the c-function (see Section 2), one comes to

|fj(k)(x)|c2φi1-ε2ϱ(x)τt1,0(0)(hj)

for some integer t1>0. It follows that

supxIj+1Ij(|x|+1)sφ0(x)-2/p|fj(k)(x)|c3js+1eϱj(2p-1+1-ε2)τt1,0(0)(hj).

Recall that the two seminorms τt,(ϑp,ε) (see (3.17)) and τ~t,(ϑp,ε) (see (3.18)) are equivalent on 𝒮(p,ε). Since hj=euc(gj), it follows that

(1+λ)t1hj(λ)==0t1(t1)λeuc(gj)(λ).

Denote by C and by Cj the compact supports of g and gj, respectively. Thus,

τ~t1,0(0)(hj)=0t1(t1)Cj|gj()(y)|𝑑y
c4=0t1supyCj(|y|+1)2|gj()(y)|
=c4=0t1supw{±1}supyCj,+(y+1)2|gj()(wy)|,

where Cj,+:=Cj+. Now one uses the Leibniz rule to compute the derivatives of gj=ωjg. Since ωj=0 on Ij-1 and since ωj and all its derivatives are bounded uniformly in j, we then have

τ~t1,0(0)(hj)c5=0t1supw{±1}supyC+Ij-1(y+1)2|g()(wy)|,

where C+:=C+. Hence,

js+1eϱj(2p-1+1-ε2)τ~t1,0(0)(hj)c6=0t1supw{±1}supyC+Ij-1(y+1)s+3eϱy(2p-1+1-ε2)|g()(wy)|
c7=0t1supw{±1}supyC+Ij-1(y+1)s+3eϱy(2p-1-1-ε2)|g()(wy)|.

Recall that g(x)=euc-1(h)(x), where euc is the Euclidean Fourier transform and hPW(). By Cauchy’s integral theorem, it is known that

p(u)eαug()(u)=cp(iλ){(iλ-α)h(λ+iα)}eiλu𝑑λ

for any polynomial p[u]. Hence,

=0t1supw{±1}supyC+Ij-1(y+1)s+3eϱy(2p-1-1-ε2)|g()(wy)|c8r=0s+3sup|Imλ|ϱϑp,ε(|λ|+1)t2|h(r)(λ)|
=c8r=0s+3τt2,r(ϑp,ε)(h)

for some integer t2>0.

It remains for us to estimate the function

x(|x|+1)sφ0(x)-2/p|f(k)(x)|

on I1=[-1,1]. First, it is not hard to prove that for |x|1 there is a positive constant c and an integer mk1 such that

|kxkΨA,ε(λ,x)|c(|λ|+1)mk|iλ-εϱ|φ0(x)

for λ such that |λ|1-ε2ϱ. Now, arguing as above, we have

|f(k)(x)|c1φ0(x)|h(λ)|(|λ|+1)mk|iλ-εϱ||1-εϱiλ|πε(dλ).

Since I1 is compact, it follows that

supxI1(|x|+1)sφ0(x)-2/p|f(k)(x)|c2|h(λ)|(|λ|+1)mk|iλ-εϱ||1-εϱiλ|πε(dλ)c3τt,0(0)(h)

for some integer t>0.

This finishes the proof of Lemma 3.11. ∎

In summary, we have proved the following theorem.

Theorem 3.12

The Fourier transform FA,ε is a topological isomorphism between Sp(R) and S(Cp,ε) with

(3.19){0<p1for ϱ=0 or [ϱ>0 and ε=0],0<p<21+1-ε2for ϱ>0 and ε[-1,1]{0}.

4 Pointwise multipliers

For -1ε1 and p as in (3.19) denote by 𝒮p() and by 𝒮(p,ε) the topological dual spaces of 𝒮p() and 𝒮(p,ε), respectively.

Let f be a Lebesgue measurable function on such that

|f(x)|φ0(x)2/p(|x|+1)-A(x)𝑑x<

for some . Then the functional Tf defined on 𝒮p() by

Tf,ϕ=4f(x)ϕ(-x)A(x)𝑑x,ϕ𝒮p(),

is in 𝒮p(). Indeed,

|Tf,ϕ|cσ,0(p)(ϕ)|f(x)|φ0(x)2/p(|x|+1)-A(x)𝑑x<.

Further, since p21+1-ε22, the Schwartz space 𝒮p() can be seen as a subspace of 𝒮p() by identifying f𝒮p() with Tf𝒮p().

Now let h be a measurable function on such that

|h(λ)|(|λ|+1)-|1-εϱiλ|πε(dλ)<

for some . Here πε(dλ) denotes the Plancherel measure (3.1),

πε(dλ)=|λ|λ2-(1-ε2)ϱ2|c(λ2-(1-ε2)ϱ2)|2𝟙]-1-ε2ϱ,1-ε2ϱ[(λ)dλ,

where c is Harish-Chandra’s function associated to the operator Δ (see Section 2). Then the functional 𝒯h defined on 𝒮(p,ε) by

𝒯h,ψ=h(λ)ψ(λ)(1-εϱiλ)πε(dλ),ψ𝒮(p,ε),

is in the dual space 𝒮(p,ε). In fact,

|𝒯h,ψ|cτt,0(0)(ψ)|h(λ)|(|λ|+1)-t|1-εϱiλ|πε(dλ)<.

Moreover, since |c(μ)|-2|μ|2α+1 for |μ| large (with α>-1/2) and

|c(μ)|-2{|μ|2for |μ|1 and ϱ>0,|μ|2α+1for |μ|1 and ϱ=0 (with α>0),

it follows that the Schwartz space 𝒮(p,ε) can be identified with a subspace of 𝒮(p,ε).

For T in 𝒮p() we define the distributional Fourier transform A,ε(T) of T on 𝒮(p,ε)=A,ε(𝒮p()) by

A,ε(T),A,ε(ϕ)=T,ϕ,ϕ𝒮p().

This can be written as

A,ε(T),ψ=T,A,ε-1(ψ),ψ𝒮(p,ε).

This definition is an extension of the Fourier transform on 𝒮p(). Indeed, let f𝒮p(). Applying Fubini’s theorem, for every ϕ𝒮p(), we have

𝒯A,ε(f),A,ε(ϕ)=A,ε(f)(λ)A,ε(ϕ)(λ)(1-εϱiλ)πε(dλ)
=f(x){A,ε(ϕ)(λ)ΨA,ε(λ,-x)(1-εϱiλ)πε(dλ)}A(x)𝑑x
=4f(x)ϕ(-x)A(x)𝑑x
=Tf,ϕ.

Hence, A,ε(Tf)=𝒯A,ε(f).

A function ψ defined on the tube domain p,ε is called a pointwise multiplier of 𝒮(p,ε) if the mapping ϕψϕ is continuous from 𝒮(p,ε) into itself. The following statement comes from [4, Proposition 3.2], with changes appropriate to our setting.

Lemma 4.1

Let ψ be a function defined on Cp,ε. Then ψ is a pointwise multiplier of S(Cp,ε) if and only if ψ satisfies the following three conditions:

  1. ψ is holomorphic in the interior of p,ε.

  2. For every t the derivatives ψ(t) extend continuously to p,ε.

  3. For every t there exists nt such that

    supλp,ε(|λ|+1)-nt|ψ(t)(λ)|<.

Theorem 4.2

Suppose that

{0<p1for ϱ=0 or [ϱ>0 and ε=0],22+1-ε2p<21+1-ε2for ϱ>0 and ε[-1,1]{0}.

If TSp(R) such that ψ:=FA,ε(T) is a pointwise multiplier of S(Cp,ε), then for any sN there exist N and continuous functions fm defined on R, m=0,1,,, such that

T=m=0ΛA,εmfm

and, for every such m, we have

(4.1)supx(|x|+1)sφ0(x)-2p+1-ε2|fm(x)|<.

Here ΛA,ε is the differential-reflection operator (1.1).

Proof.

It is assumed that ψ=A,ε(T) is a pointwise multiplier of 𝒮(p,ε). Then by Lemma 4.1, for all t there is an integer nt such that

(4.2)supλp,ε(|λ|+1)-nt|ψ(t)(λ)|<.

Fix s and consider an integer that will be specified later. Define the function κ on p,ε by

κ(λ)=(iλ+ϱ+1)-ψ(λ).

In view of our assumption on p, the function κ satisfies the first and the second conditions in the definition of the space 𝒮(p,ε). Further, since |ΨA,ε(λ,x)|2 for all λ, we have

|A,ε-1(κ)(x)|:=|cκ(λ)ΨA,ε(λ,x)(1-εϱiλ)πε(dλ)|c1|κ(λ)||1-εϱiλ|πε(dλ),

where

|1-εϱiλ|πε(dλ)=λ2+ε2ϱ2λ2-(1-ε2)ϱ21|c(λ2-(1-ε2)ϱ2)|2𝟙]-1-ε2ϱ,1-ε2ϱ[(λ)dλ.

Thus, in view of the estimate (4.2) and the behavior of |c(μ)|-2 for small and large |μ|, it follows that A,ε-1(κ)(x) exists for all x provided that >n0+2α+2. Here the parameter n0 comes from (4.2). Moreover, for all ϕ𝒮p(), Fubini’s theorem leads to

ϕ(-x)A,ε-1(κ)(x)A(x)𝑑x=c1ϕ(-x)(κ(λ)ΨA,ε(λ,x)(1-εϱiλ)πε(dλ))A(x)𝑑x
=c1κ(λ)(ϕ(-x)ΨA,ε(λ,x)A(x)𝑑x)(1-εϱiλ)πε(dλ)
=c1κ(λ)A,ε(ϕ)(λ)(1-εϱiλ)πε(dλ).

It follows that the inverse Fourier transform A,ε-1(κ) of κ as an element of 𝒮(p,ε) concurs with the classical Fourier transform of κ. Further,

T=A,ε-1((iλ+ϱ+1)κ)=m=0(m)(ϱ+1)-mΛA,εmA,ε-1(κ):=m=0ΛA,εmfm.

It remains for us to show that, given s, the functions fm satisfy (4.1), provided that is large enough. To do so, we will use a similar approach to that in the proof of Lemma 3.11.

Set ξ:=A,ε-1(κ) and g:=euc-1(κ), where euc denotes the Euclidean Fourier transform. Observe that if is large enough, then g is well defined. For j1 let ωjC() such that ωj=0 on Ij-1:=[-(j-1),j-1] and ωj=1 outside of Ij. We shall assume that ωj and all its derivatives are bounded uniformly in j.

We set gj:=ωjg, and define κj:=euc(gj) and ξj=A,ε-1(κj). Since ωj=1 outside of Ij, it follows that gj-g=0 outside of Ij, that is supp(gj-g)Ij. Using the support conservation property from the proof of Lemma 3.11, we deduce that ξ may differ from ξj only inside Ij. Now we will estimate the function

(4.3)x(|x|+1)sφ0(x)-2p+1-ε2ξ(x),

first on I1 and next on Ij+1Ij for j1.

We claim that

|ΨA,ε(λ,x)|c2(|λ|+1)φ0(x)

for λ such that |λ|1-ε2ϱ. Indeed, as λ is such that |λ|1-ε2ϱ, it follows that με. Thus, the claim follows from the superposition (2.8) of ΨA,ε(λ,x) and the facts that |φμε(x)|φ0(x) and |φμε(x)|c(με2+ϱ2)φ0(x) (see Lemma 2.2).

From the claim above we have

|ξ(x)|c3|κ(λ)||ΨA,ε(λ,x)||1-εϱiλ|πε(dλ)
c4φ0(x)|κ(λ)|(|λ|+1)|1-εϱiλ|πε(dλ).

Since I1 is compact, we deduce that for every s we have

supxI1(|x|+1)sφ0(x)-2p+1-ε2|ξ(x)|<

whenever >n0+2α+3. Here the parameter n0 comes from (4.2). Now we consider the estimate of function (4.3) on Ij+1Ij for j1. Recall that ξ=ξj outside of Ij.

Arguing as above, we obtain

|ξj(x)|c5φ0(x)supλ]-1-ε2ϱ,1-ε2ϱ[|(λ+1)t1κj(λ)|

for some integer t1>2α+3. It follows that

supxIj+1Ij(|x|+1)sφ0(x)-2p+1-ε2|ξj(x)|c6jse(2p-1-1-ε2)ϱjsupλ]-1-ε2ϱ,1-ε2ϱ[|(λ+1)t1κj(λ)|.

Since κj=euc(gj) with gj=ωjg, we claim that

(4.4)|(λ+1)t1κj(λ)|c7q=0t1supw{±1}supx+Ij-1(x+1)2|g(q)(wx)|.

Indeed, on the one hand, we have

(4.5)(λ+1)t1κj(λ)=r=0t1crgj(x)xreiλxdx=r=0t1crωj(x)g(x)xreiλxdx.

On the other hand, we have

(4.6)(ωjg)(r)(x)=q=0rcqg(q)(x)ωj(r-q)(x)0as |x|+.

In fact, starting from g=euc-1(κ), we obtain

(4.7)g(q)(x)=cκ(λ)(iλ)qeiλx𝑑λ.

Thus, if >n0+t1+1 then by the Riemann–Lebesgue lemma for the Euclidean Fourier transform, we obtain g(q)(x)0 as |x|. Thus (4.6) holds true. Now in view of (4.6) we may rewrite (4.5) as

(λ+1)t1κj(λ)=r=0t1q=0rcq,rg(q)(x)ωj(r-q)(x)eiλx𝑑x.

Recall that the function ωj vanishes on Ij-1 and is bounded, together with all its derivatives, uniformly in j. Therefore,

|(λ+1)t1κj(λ)|cq=0t1Ij-1|g(q)(x)|𝑑x
cq=0t1supxIj-1(|x|+1)2|g(q)(x)|.

This finishes the proof of claim (4.4).

It follows that

jse(2p-1-1-ε2)ϱjsupλ]-1-ε2ϱ,1-ε2ϱ[|(λ+1)t1κj(λ)|
c7q=0t1supw{±1}supx+Ij-1(x+1)s+2e(2p-1-1-ε2)ϱx|g(q)(wx)|.

Next we shall prove that the right-hand side is finite. Assume first that ϱ=0. By (4.7) we have

(4.8)(x+1)s+2g(q)(wx)=r=0s+2cq,rκ(λ)λqλreiλwxdλ.

We claim that

(4.9)(κ(λ)λq)(r)0as |λ|+,

provided that is large enough. Indeed, this claim follows immediately from the fact that

(κ(λ)λq)(r)=a=0rcaλq-r+aκ(a)(λ)(with r-aq)
(4.10)=a=0rb=0aca,bλq-r+a(iλ+1)--a+bψ(b)(λ),

together with the fact that ψ satisfies (4.2). Thus, by (4.9) we may rewrite (4.8) as

(4.11)(x+1)s+2g(q)(wx)=r=0s+2cq,r(κ(λ)λq)(r)eiλwx𝑑λ.

Using again the fact that ψ satisfies (4.2) together with the double sum (4.10), we obtain from (4.11) that

supw{±1}supx+Ij-1(x+1)s+2|g(q)(wx)|<

for ϱ=0, provided that is large enough.

Now assume that ϱ>0. Since g=euc-1(κ) and κ is holomorphic in the interior of p,ε, Cauchy’s integral theorem gives

p(u)eαug(q)(u)=cstp(iλ){(iλ-α)qκ(λ+iα)}eiλu𝑑λ,

with p(x)=(x+1)s+2 and α=(2p-1-1-ε2)ϱ. The same argument as above implies that

supw{±1}supx+Ij-1(x+1)s+2e(2p-1-1-ε2)ϱx|g(q)(wx)|<,

provided that is large enough.

Putting the pieces together, we conclude that

supxIj+1Ij(|x|+1)sφ0(x)-2p+1-ε2|ξj(x)|<

for large enough. ∎

References

[1] Anker J.-P., The spherical Fourier transform of rapidly decreasing functions. A simple proof of a characterization due to Harish-Chandra, Helgason, Trombi, and Varadarajan, J. Funct. Anal. 96 (1991), 331–349. 10.1016/0022-1236(91)90065-DSearch in Google Scholar

[2] Ben Saïd S., Boussen A. and Sifi M., Intertwining operators associated to a family of differential-reflection operators, Mediterr. J. Math. 13 (2016), 4129–4151. 10.1007/s00009-016-0736-2Search in Google Scholar

[3] Ben Saïd S., Boussen A. and Sifi M., Lp harmonic analysis for differential-reflection operators, C. R. Math. Acad. Sci. Paris 354 (2016), no. 5, 510–516. 10.1016/j.crma.2016.01.020Search in Google Scholar

[4] Betancor J. D., Betancor J. J. and Méndez J. M. R., Convolution operators on Schwartz spaces for Chébli–Trimèche hypergroups, Rocky Mountain J. Math. 37 (2007), no. 3, 723–761. 10.1216/rmjm/1182536162Search in Google Scholar

[5] Bloom W. R. and Xu Z., The Hardy–Littlewood maximal function for Chébli–Trimèche hypergroups, Applications of Hypergroups and Related Measure Algebras (Seattle 1993), Contemp. Math. 183, American Mathematical Society, Providence (1995), 45–70. 10.1090/conm/183/02054Search in Google Scholar

[6] Bloom W. R. and Xu Z., Fourier transforms of Schwartz functions on Chébli–Trimèche hypergroups, Monatsh. Math. 125 (1998), no. 2, 89–109. 10.1007/BF01332821Search in Google Scholar

[7] Bracco O., Fonction maximale associée à des opérateurs de Sturm–Liouville singuliers, Thèse, Université Louis Pasteur (Strasbourg I), Strasbourg, 1999. Search in Google Scholar

[8] Bracco O., Propriétés de la mesure spectrale pour une classe d’opérateurs différentiels singuliers sur (0,), C. R. Acad. Sci. Paris Ser. I Math. 329 (1999), no. 4, 299–302. 10.1016/S0764-4442(00)88570-5Search in Google Scholar

[9] Bouzeffour F., Fitouhi A., Ghazouani S. and Nemri A., On harmonic analysis related with the generalized Dunkl operator, Integral Transforms Spec. Funct. 23 (2012), no. 8, 609–625, 783. 10.1080/10652469.2011.618807Search in Google Scholar

[10] Chébli H., Sur un théorème de Paley–Wiener associé à la décomposition spectrale d’un opérateur de Sturm–Liouville sur (0,), J. Funct. Anal. 17 (1974), 447–461. 10.1016/0022-1236(74)90052-4Search in Google Scholar

[11] Chébli H., Théorème de Paley–Wiener associé à un opérateur différentiel singulier sur (0,),, J. Math. Pures Appl. (9) 58 (1979), no. 1, 1–19. 10.5802/jedp.178Search in Google Scholar

[12] Cherednik I., A unification of Knizhnik–Zamolodchnikov equations and Dunkl operators via affine Hecke algebras, Invent. Math. 106 (1991), 411–432. 10.1007/BF01243918Search in Google Scholar

[13] Clerc J.-L., Transformation de Fourier sphérique des espaces de Schwartz, J. Funct. Anal. 37 (1980), 182–202. 10.1016/0022-1236(80)90040-3Search in Google Scholar

[14] Delorme P., Espace de Schwartz pour la transformation de Fourier hypergéométrique, J. Funct. Anal. 168 (1999), no. 1, 239–312. 10.1006/jfan.1999.3449Search in Google Scholar

[15] Dunkl C., Differential-difference operators associated to reflection groups, Trans. Amer. Math. Soc. 311 (1989), 167–183. 10.1090/S0002-9947-1989-0951883-8Search in Google Scholar

[16] Ehrenpreis L. and Mautner F. I., Some properties of the Fourier-transform on semisimple Lie Groups. III, Trans. Amer. Math. Soc. 90 (1959), 431–484. 10.1090/S0002-9947-1959-0102755-3Search in Google Scholar

[17] Flensted-Jensen M., Spherical functions of a real semisimple Lie group. A method of reduction to the complex case, J. Funct. Anal. 30 (1978), 106–146. 10.1016/0022-1236(78)90058-7Search in Google Scholar

[18] Harish-Chandra , Spherical functions on a semisimple Lie group. I, Amer. J. Math. 80 (1958), 241–310. 10.2307/2372786Search in Google Scholar

[19] Heckman G. J., An elementary approach to the hypergeometric shift operators of Opdam, Invent. Math. 103 (1991), no. 2, 341–350. 10.1007/BF01239517Search in Google Scholar

[20] Lions J. L., Opérateurs de Delsarte et problèmes mixtes, Bull. Soc. Math. France 84 (1956), 9–95. 10.24033/bsmf.1467Search in Google Scholar

[21] Mourou M. A. and Trimèche K., Transmutation operator and Paley–Wiener theorem associated with a singular differential-difference operator on the real line, Anal. Appl. (Singap.) 1 (2003), 43–70. 10.1142/S0219530503000090Search in Google Scholar

[22] Narayanan E. K., Pasquale A. and Pusti S., Asymptotics of Harish-Chandra expansions, bounded hypergeometric functions associated with root systems, and applications, Adv. Math. 252 (2014), 227–259. 10.1016/j.aim.2013.10.027Search in Google Scholar

[23] Thyssen M., Sur certains opérateurs de transmutation particuliers, Mem. Soc. R. Sci. Liége Coll. 8 V. Sér. 6 (1961), no. 3, 1–32. Search in Google Scholar

[24] Trimèche K., Transformation intégrale de Weyl et théorème de Paley–Wiener associés à un opérateur différentiel singulier sur (0,), J. Math. Pures Appl. (9) 60 (1981), no. 1, 51–98. Search in Google Scholar

[25] Trimèche K., Generalized wavelets and hypergroups, Gordon and Breach Science, Amsterdam, 1997. Search in Google Scholar

[26] Trombi P. C. and Varadarajan V. S., Spherical transforms of semisimple Lie groups, Ann. of Math. (2) 94 (1971), 246–303. 10.2307/1970861Search in Google Scholar

Received: 2016-1-15
Revised: 2016-9-29
Accepted: 2016-9-29
Published Online: 2016-11-6
Published in Print: 2017-1-1

© 2017 by De Gruyter

Downloaded on 4.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/apam-2016-0009/html?lang=en
Scroll to top button