1 Introduction

A classical version of the uncertainty principle states that an integrable function f defined on the real line and its Fourier transform \({\hat{f}} \) cannot be simultaneously and sharply localized unless \(f=0\) almost everywhere. An important result making this precise, is the Hardy Theorem (see [13]):

Theorem 1

Let \(\alpha , \beta , c\) be positive real numbers and f a measurable function on \({\mathbb {R}}\) such that:

  1. (i)

    \( \vert f(t)\vert \le c \text {e}^{-\alpha \pi t^2}, \ \ t\in {\mathbb {R}}\),

  2. (ii)

    \( \vert {\hat{f}}(k)\vert \le c \text {e}^{-\beta \pi k^2}, \ \ k\in {\mathbb {R}}\).

If \(\alpha \beta > 1\), then \(f= 0\) a.e. If \(\alpha \beta = 1\) then \(f(t)=b\text {e}^{-\alpha \pi t^2}\), for some constant b. If \(\alpha \beta < 1\), then any finite linear combination of Hermite functions satisfies (i) and (ii).

For the Fourier transformation we use the normalization

$$\begin{aligned} {{\hat{f}}}(k)=\int _{{\mathbb {R}}}f(t)\hbox {e}^{-2i\pi t k}dt,\ k\in {\mathbb {R}}. \end{aligned}$$

Note that Hardy’s theorem is also valid on \({\mathbb {R}}^n\) (see [19,  Theorem 4]). Much efforts have been deployed to prove Hardy-like theorems for various classes of non-Abelian connected Lie groups. Specifically, analogues and variants of Hardy’s theorem have been shown for nilpotent Lie groups [1, 2, 14, 19, 21], some classes of solvable Lie groups [3], non-compact connected semisimple Lie groups G with finite center [9, 15, 17, 18] and motion groups [16, 20]. For more information and further references concerning the entire subject, we refer the readers to the excellent monograph by Thangavelu [22].

The continuous Gabor transform (also known as windowed Fourier transform) is a classical tool in mathematical signal processing. Roughly speaking, it is the Fourier transform of a signal f seen through a sliding window \(\psi \). For \(f\in L^2({\mathbb {R}}^n)\) and a nonzero function \(\psi \in L^2({\mathbb {R}}^n)\) called a window function, the continuous Gabor transform with respect to \(\psi \) is defined on \({\mathbb {R}}^n\times \widehat{{\mathbb {R}}}^n\) by

$$\begin{aligned} {\mathcal {G}}_\psi f(x,w): =\int _{{\mathbb {R}}^n} f(y){\overline{\psi }}(y-x)e^{-2i\pi \langle y,w \rangle }dy. \end{aligned}$$

According to [11], we have for all \(f_1\), \(f_2\), \(\psi _1\), \(\psi _2\in L^2({\mathbb {R}}^n)\) the functions \({\mathcal {G}}_{\psi _1}f_1\) and \({\mathcal {G}}_{\psi _2}f_2\) belong to \(L^{2}({\mathbb {R}}^n\times \hat{{\mathbb {R}}}^n)\) and

$$\begin{aligned} \langle {\mathcal {G}}_{\psi _1}f_1,{\mathcal {G}}_{\psi _2}f_2\rangle _{L^2({\mathbb {R}}^n\times \hat{{\mathbb {R}}}^n)}=\langle f_1 , f_2\rangle _{L^2({\mathbb {R}}^n)}\overline{\langle \psi _1,\psi _2\rangle }_{L^2({\mathbb {R}}^n)}. \end{aligned}$$
(1)

This transform plays an important role in time-frequency analysis namely by providing an interesting way to study the local frequency spectrum of signals. For a detailed discussion of time-frequency analysis, we refer the readers to [11]. It has been shown in the early 2000s that many uncertainty principles for the Fourier transform have a counterpart for the continuous Gabor transform (see e.g. [7, 12]). We specify that a Hardy-type theorem has been established in [12,  Theorem 2.6.2].

Theorem 2

Let \(f,\psi \in L^2({\mathbb {R}}^n)\). Assume that

$$\begin{aligned} \Big \vert {\mathcal {G}}_\psi f(x,w)\Big \vert \le C \text {e}^{-\pi (\alpha x^2+\beta w^2)/2}, \end{aligned}$$

for some constants \(\alpha , \beta , C>0\). Then three cases can occur.

  1. (i)

    If \(\alpha \beta > 1\), then either \(f=0\) a.e. or \(\psi = 0\) a.e.

  2. (ii)

    If \(\alpha \beta = 1\) and \({\mathcal {G}}_\psi f\) is not zero almost everywhere, then both f and \(\psi \) are multiples of some time-frequency shift of the Gaussian \(\text {e}^{-\alpha \pi x^2}\).

  3. (iii)

    If \(\alpha \beta < 1\), then the decay condition is satisfied whenever f and \(\psi \) are finite linear combinations of Hermite functions.

On the other hand, the continuous Gabor transform has also been extended to separable locally compact unimodular group of type I (see [10]). One should notice that, in the Euclidean setting, the continuous Gabor transform has many symmetries which are lost in the Lie group setting (the dual of G does not identify with G) and this is then a serious obstacle for stating uncertainty principles for the continuous Gabor transform. However, some attempts to extend Theorem 2 on special classes of non-Abelian Lie groups have already been made. In particular, we cite here the work of Bansal, Kumar and Sharma [6], where the authors generalized Hardy’s theorem for the Gabor transform on locally compact abelian groups having noncompact identity component and groups of the form \({\mathbb {R}}^n\times K\), where K is a compact group having irreducible representations of bounded dimension. When it comes to connected nilpotent Lie groups, only a conjecture are now available. In the same reference, the previous authors conjecture that if \(\alpha ,\beta \) and C are positive real numbers and \(f,\psi \) are square integrable functions on connected nilpotent Lie group \(G=\exp \mathfrak {g} \) such that \( \Vert {\mathcal {G}}_\psi f (g,\pi _l)\Vert _{HS}\le C \text {e}^{-\pi (\alpha \Vert g\Vert ^2+\beta \Vert l\Vert ^2)/2}\) for all \((g,l)\in G\times {\mathcal {W}}\), then \(f= 0\) a.e. or \(\psi = 0\) a.e. provided that \(\alpha \beta >1\). Here \({\mathcal {W}}\) is a suitable cross-section for the generic coadjoint orbits in \(\mathfrak {g}^*\), the vector space dual of \(\mathfrak {g}\), and \(\Vert g\Vert \) and \(\Vert l\Vert \) are substitutes (in terms of bases of \(\mathfrak {g}\) and \(\mathfrak {g}^*\)) for the Euclidean norms on \({\mathbb {R}}^n\) and \(\widehat{{\mathbb {R}}}^n={\mathbb {R}}^n\). For details and unexplained notation see Sect. 2. They also proved that this conjecture fails for a connected nilpotent Lie group G having a square integrable irreducible representation. This paper is the first attempt to establish analog of Hardy’s theorem for Gabor transform on nilpotent Lie groups. By exploiting Hardy’s Theorem for \({\mathbb {R}}\) and a localized version of the Plancherel formula, we show in Sect. 4 that the above-mentioned conjecture holds. Our main result is the following:

Theorem 3

Let G be connected, simply connected nilpotent Lie group. Let \(f,\psi \in L^2(G)\) be such that

$$\begin{aligned} \Vert {\mathcal {G}}_\psi f (g,\pi _l)\Vert _{HS}^2\le C e^{-{\pi \over 2} (\alpha \Vert g\Vert ^2+\beta \Vert l\Vert ^2)}, \end{aligned}$$
(2)

for all \((g,l)\in G\times {{\mathcal {W}}}\), where \(\alpha ,\beta \) and C are positive real numbers. If \(\alpha \beta >1\), then either \(f=0\) a.e. or \(\psi =0\) a.e.

2 Backgrounds

2.1 Continuous Gabor Transform

Let G be a separable locally compact unimodular group of type I, and let dg be its Haar measure. We endow the unitary dual of G with the Mackey Borel structure. We denote by \(L^p(G)\) the space of \(L^p\)-functions on G for \(p\ge 1\), and we define

$$\begin{aligned} \pi (f)=\int _Gf(g)\pi (g)dg,\quad \pi \in {{\hat{G}}}, \, f\in L^1(G). \end{aligned}$$

Then by the abstract Plancherel theorem, there exists a unique Borel measure \(\rho \) on \({{\hat{G}}}\) such that for any function \(f\in L^1(G)\cap L^2(G)\),

$$\begin{aligned} \int _G \vert f(g)\vert ^2dg=\int _{{{\hat{G}}}}\Vert \pi (f) \Vert _{HS}^2d\rho (\pi ), \end{aligned}$$

where \(\Vert \pi (f) \Vert _{HS}=\big (\text {tr}\big (\pi (f)^*\pi (f)\big )\big )^{1/2}\) denotes the Hilbert-Schmidt norm of \(\pi (f)\).

Let \(f \in C_c(G)\), the set of all continuous complex-valued functions on G with compact supports, and \(\psi \) a fixed nonzero function in \(L^2(G)\), usually called window function. For \((x,\pi )\in G\times {{\hat{G}}}\), the continuous Gabor transform of f with respect to the window function \(\psi \) is defined as a measurable field of operators on \(G\times {{\hat{G}}}\) by

$$\begin{aligned} {\mathcal {G}}_\psi f(x,\pi ) :=\int _Gf(g){\overline{\psi }}(x^{-1}g)\pi (g)dg. \end{aligned}$$

Let \(f_\psi ^x\) be the function defined on G by

$$\begin{aligned} f_\psi ^x(g)=f(g){\overline{\psi }} (x^{-1}g),\qquad \forall g\in G. \end{aligned}$$

Then, \( f_\psi ^x\in L^1(G)\cap L^2(G)\) and

$$\begin{aligned} \pi (f_\psi ^x)=\int _Gf_\psi ^x(g)\pi (g)dg=\int _Gf(g){\overline{\psi }}(x^{-1}g)\pi (g)dg={\mathcal {G}}_\psi f(x,\pi ). \end{aligned}$$
(3)

By the Plancherel theorem, \({\mathcal {G}}_\psi f(x,\pi )\) is a Hilbert-Schmidt operator for all \(x \in G\) and for almost all \(\pi \in {{\hat{G}}}\). Furthermore,

$$\begin{aligned} \int _G \int _{{{\hat{G}}}} \Vert {\mathcal {G}}_\psi f(x,\pi )\Vert _{HS}^2d\rho (\pi )\, dx=\Vert \psi \Vert _2^2\Vert f \Vert _2^2. \end{aligned}$$
(4)

Thus, the continuous Gabor transform \( {\mathcal {G}}_\psi : f\mapsto {\mathcal {G}}_\psi f\) (\(f\in C_c(G)\)) is a multiple of an isometry. So, we can extend \( {\mathcal {G}}_\psi \) uniquely to a bounded linear operator on \(L^2(G)\) which we still denote by \( {\mathcal {G}}_\psi \) and this extension satisfies (4) for each \( f \in L^2(G)\).

2.2 Nilpotent Lie Groups

We begin this subsection by reviewing some useful facts and notations for nilpotent Lie group. This material is quite standard, we refer the reader to [8] for details. We assume henceforth that \(G=\exp \mathfrak {g}\) is a connected, simply connected nilpotent Lie group.

Let \({\mathcal {B}}=\{X_1,...,X_n\}\) be a strong Malcev basis of \(\mathfrak {g}\) passing through the center of \(\mathfrak {g}\). We introduce a norm function on G by setting for \( x = \exp (x_ 1X_1+ \cdots +x_ nX_n )\in G, \ {x_ j} \in {\mathbb {R}}\),

$$\begin{aligned} \Vert x\Vert = \sqrt{{({x_ 1^2}+ \cdots +{x_n^2})}}. \end{aligned}$$

The map:

$$\begin{aligned} {{{\mathbb {R}}}^n} \rightarrow G , \ ({x_ 1},...,{x_ n}) \mapsto \displaystyle \exp \big (\sum _ {j = 1}^{n }{x_ j} {X_ j}\big ) \end{aligned}$$

is a diffeomorphism and maps the Lebesgue measure on \( {{\mathbb {R}}}^n \) to the Haar measure on G. In this setup, we shall identify G as set with \( {{\mathbb {R}}}^{n}\). We consider the Euclidean norm of \(\mathfrak {g}^*\) with respect to the basis \( {\mathcal {B}}^*=\{X_1^*,...,X_n^*\}\), that is,

$$\begin{aligned} \Big \Vert { \displaystyle \sum _ {j = 1}^{n }{l_ j} {X_ j^*}}\Big \Vert =\sqrt{( l_1^2 + \cdots + l_n^2 )}= \Vert l \Vert , \ \ l_j\in {\mathbb {R}}. \end{aligned}$$

Let \({\mathcal {U}} \) denote the Zariski open subset of \(\mathfrak {g} ^*\) consisting of all elements in generic orbits with respect to the basis \({\mathcal {B}}^*\). Let S be the set of jump indices, and set \(T = \{1,...,n \}\backslash S\) and \( V_T = {\mathbb {R}}\text {-span}\{X_i^*\ : \ i \in T \}\). Then, \({\mathcal {W}} ={{\mathcal {U}} \cap V_T}\) is a cross section of the generic orbits and \({\mathcal {W}}\) supports the Plancherel measure on \( {\hat{G}}\). Let Pf(l) denote the Pfaffian of the skew-symmetric matrix \( M_S(l)= ( l([X_i,X_j]))_{ i,j \in S} .\) Then, one has that:

$$\begin{aligned} {{\vert Pf(l) \vert }^2}= \det {M_S (l)}. \end{aligned}$$

If dl is the Lebesgue measure on \( {\mathcal {W}},\) then \( d \tau = \vert Pf(l) \vert dl\) is a Plancherel measure for \( {\hat{G}}\). Let dg be the Haar measure on G. For \(\varphi \in L^1 (G) \cap L^2 (G)\), the Plancherel formula reads:

$$\begin{aligned} {{\Vert \varphi \Vert }^2 _ 2} = \displaystyle \int _G {\vert \varphi (g) \vert }^2 dg = \displaystyle \int _ {{\mathcal {W}} } { \Vert \pi _l( \varphi ) \Vert }^2_{HS} d \tau (l). \end{aligned}$$
(5)

3 Some Lemmas

In this section we prove three results, Lemmas 1, 2 and 3 which are required to prove Theorem 3.

For every \(x, w \in {\mathbb {R}}^n\), we denote by \({\mathcal {M}}_w\) and \({\mathcal {T}}_x\) the modulation and the translation operators defined respectively on \(L^2({\mathbb {R}}^n)\) by

$$\begin{aligned}&\forall z\in {\mathbb {R}}^n,\quad {\mathcal {M}}_wf(z)=e^{2i\pi \langle z,w\rangle }f(z), \\&\forall z\in {\mathbb {R}}^n,\quad {\mathcal {T}}_xf(z)=f(z-x). \end{aligned}$$

Then we deduce that,

$$\begin{aligned} \forall z\in {\mathbb {R}}^n,\quad {\mathcal {M}}_w({\mathcal {T}}_xf)(z)=e^{2i\pi \langle z,w\rangle }f(z-x), \end{aligned}$$
(6)

and

$$\begin{aligned} \forall z\in {\mathbb {R}}^n,\quad {\mathcal {T}}_x({\mathcal {M}}_wf)(z)=e^{-2i\pi \langle x,w\rangle }e^{2i\pi \langle z,w\rangle }f(z-x). \end{aligned}$$
(7)

The results in the following lemma are quite standard.

Lemma 1

Let \(f,\psi \in L^2({\mathbb {R}}^n)\) and \(\xi , \lambda , y, z \in {\mathbb {R}}^n\). Then,

  1. (i)

    \({\mathcal {G}}_{({\mathcal {M}}_\xi {\mathcal {T}}_z\psi )}({\mathcal {M}}_\lambda {\mathcal {T}}_yf)(x,w) = e^{2i\pi \langle x,\xi \rangle } e^{-2i\pi \langle y,w-\lambda +\xi \rangle } {\mathcal {G}}_\psi f(x-y+z,w-\lambda +\xi ). \)

    In particular, \({\mathcal {G}}_\psi ({\mathcal {M}}_\lambda {\mathcal {T}}_yf)(x,w)=e^{-2i\pi \langle y,w\rangle } e^{2i\pi \langle y,\lambda \rangle } {\mathcal {G}}_\psi f(x-y,w-\lambda )\).

  2. (ii)

    \({\mathcal {G}}_\psi f(-x,-w)=e^{-2i\pi \langle x, w\rangle }\overline{{\mathcal {G}}_f\psi (x,w)}.\)

  3. (iii)

    Let \(F(x,w)= {\mathcal {G}}_\psi f(x,w) {\mathcal {G}}_\psi f(-x,-w)e^{2i\pi \langle x,w\rangle }.\) Then,

    $$\begin{aligned} {\hat{F}}(\nu ,\theta )=F(-\theta ,\nu ),\quad \nu , \theta \in {\mathbb {R}}^n. \end{aligned}$$

Let’s fix as above a strong Malcev basis \(\{X_1,...,X_n\}\) of \(\mathfrak {g}\) such that \(X_1\) is in the center of \(\mathfrak {g}\). For \(a= (a_2,...,a_n)\in {\mathbb {R}}^{n-1}\), let \((f^g_\psi )_a\) be the complex valued function defined on \({\mathbb {R}}\) by

$$\begin{aligned} (f^g_\psi )_a(t)=f^g_\psi (t,a) =f^g_\psi \Big (\exp \big (tX_1+\sum _{j=2}^na_jX_j\big )\Big ). \end{aligned}$$

For \(k\in {\mathbb {R}}\) and \(s=(s_2,...,s_n)\in {\mathbb {R}}^{n-1}\), let \(g=\exp \big (kX_1+\sum _{j=2}^ns_jX_j\big )\) and \(f^{k,s}_\psi =f^g_\psi \). It is easy to see that \(f^g_\psi \in L^1(G)\) for all \(g\in G\), it sufficient to use Cauchy–Schwarz inequality. Moreover by [5,  Lemma 3.1], we have

$$\begin{aligned} {\mathcal {G}}_\psi f(g,\pi _l)=\pi _l(f^g_\psi ), \end{aligned}$$
(8)

for all \(g\in G\). We should also mention that \(f^g_\psi \in L^2(G)\), for almost all \(g\in G\). In fact,

$$\begin{aligned} \int _G\int _G\vert f^g_\psi (x)\vert ^2dx\, dg =\int _G\int _G\vert f(x)\vert ^2\vert \psi (g^{-1}x)\vert ^2dx\, dg =\Vert f \Vert _2^2\Vert \psi \Vert _2^2<\infty . \end{aligned}$$

Then obviously \(\int _G\vert f^g_\psi (x)\vert ^2dx<\infty \), for almost all \(g\in G\).

Lemma 2

Let \(f,\psi \in L^2(G)\) meet the condition (2) of Theorem 3. Then

$$\begin{aligned} I(f^{k,s}_\psi ):= \int _{{\mathbb {R}}^{n-1}} \Bigg ( \int _{{\mathbb {R}}} \big \vert (f^{k,s}_\psi )_a(t)\big \vert dt\Bigg )^2 da <\infty , \end{aligned}$$

for all \(k\in {\mathbb {R}}\) and almost all \(s=(s_2,...,s_n)\in {\mathbb {R}}^{n-1}\).

Proof

By using (2), we have

$$\begin{aligned}&\int _G\int _{\mathcal {W}} (1+\Vert g \Vert ^2) \Vert {\mathcal {G}}_\psi f(g,\pi _l)\Vert _{HS}^2 \vert Pf(l)\vert dl\,dg\nonumber \\&\quad \le \int _G\int _{\mathcal {W}} (1+\Vert g \Vert ^2) e^{-\pi (\alpha \Vert g\Vert ^2+\beta \Vert l\Vert ^2)} \vert Pf(l)\vert dl\,dg. \end{aligned}$$
(9)

Assume that the degree of the polynomial function Pf(l) is equal to \(\delta \). Then,

$$\begin{aligned} \vert Pf(l)\vert\le & {} \text {cst} \ (1+\Vert l\Vert ^2)^\frac{\delta }{2}.\\\le & {} \text {cst} \ (1+\Vert l\Vert ^2)^\delta \end{aligned}$$

Therefore, the integral on the right hand side of (9) converges. Hence,

$$\begin{aligned} \infty> & {} \int _G(1+\Vert g \Vert ^2)\left( \int _{\mathcal {W}} \Vert \pi (f_\psi ^g) \Vert ^2_{HS}\vert Pf(l)\vert dl\right) dg \\= & {} \int _G\int _G (1+\Vert g \Vert ^2)\vert f(x)\vert ^2 \vert \psi (g^{-1}x)\vert ^2dx\,dg \end{aligned}$$

(using Plancherel formula of G)

$$\begin{aligned}&\ge \int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}} \int _{{\mathbb {R}}}(1+\vert k\vert ^{2}) \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2\\&\quad \times \Big \vert \psi \Big ( \exp \Big (kX_1+\sum _{j=2}^{n}s_jX_j\Big )^{-1}\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2 dt\, da\, dk\, ds. \end{aligned}$$

Noting that,

$$\begin{aligned}&\exp \Big (kX_1+\sum _{j=2}^{n}s_jX_j\Big )^{-1}\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\nonumber \\&\quad =\exp \Big ( \big (t-k+Q_1(a,s)\big )X_1+ \sum _{j=2}^{n} Q_j(a,s)X_j \Big ), \end{aligned}$$
(10)

where, for \(1\le j\le n\), \(Q_j\) is a polynomial function depending on \(a=(a_2,...,a_n)\) and \(s=(s_2,...,s_n)\). Furthermore, one can write

$$\begin{aligned} Q_j(a,s)= a_j-s_j+Q_j'(a_{j+1},...,a_n,s_{j+1},...,s_n), \quad j=2,...,n. \end{aligned}$$
(11)

It follows that,

$$\begin{aligned} \infty> & {} \int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}} (1+\vert k\vert ^{2}) \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2\\&\times \Big \vert \psi \Big ( \exp \Big ( \big (t-k+Q_1(a,s)\big )X_1+ \sum _{j=2}^{n} Q_j(a,s)X_j \Big )\Big )\Big \vert ^2 dt\,da\,dk\, ds \\= & {} \int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}}\Big (1+ \big \vert t- r+Q_1(a,s)\big \vert ^{2}\Big ) \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2\\&\times \Big \vert \psi \Big ( \exp \Big ( rX_1+ \sum _{j=2}^{n} Q_j(a,s)X_j \Big )\Big )\Big \vert ^2 dt\,da\,dr\, ds \end{aligned}$$

(by substituting \(r=t-k+Q_1(a,s)\) for k). Now let’s use the change of variable \(\sigma _j=Q_j(a,s)\), \(j=2,...,n\), for fixed value of a. Note that, from (11) this change of variable has Jacobian 1. We then obtain,

$$\begin{aligned} \infty> & {} \int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}}\Big (1+\big \vert t- r+R(a,\sigma )\big \vert ^{2}\Big )\Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2\\&\times \Big \vert \psi \Big ( \exp \Big ( rX_1+\sum _{j=2}^{n} \sigma _jX_j \Big )\Big )\Big \vert ^2 dt\,da\,dr\, d\sigma , \end{aligned}$$

where \(R(a,\sigma )\) is a polynomial function depending on \(a=(a_2,...,a_n)\) and \(\sigma =(\sigma _2,...,\sigma _n)\). Therefore,

$$\begin{aligned}&\Big \vert \psi \Big ( \exp \Big ( rX_1+\sum _{j=2}^{n} \sigma _jX_j \Big )\Big )\Big \vert ^2 \int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}}\Big (1+ \big \vert t- r+R(a,\sigma )\big \vert ^{2}\Big ) \\&\quad \times \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2 dt\, da <\infty , \end{aligned}$$

for almost all \(r,\sigma _2,..., \sigma _n\in {\mathbb {R}}\). As \(\psi \) is non identically zero, there exists \(r_0,\sigma _0=(\sigma _2^0, ..., \sigma ^0_n) \) such that

$$\begin{aligned} \psi \Big ( \exp \Big (r_0X_1+ \sum _{j=2}^{n} \sigma ^0_jX_j \Big )\Big ) \ne 0, \end{aligned}$$

and

$$\begin{aligned} \int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}} \Big (1+\big \vert t- r_0+R(a,\sigma _0)\big \vert ^{2}\Big ) \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2 dt\, da <\infty .\nonumber \\ \end{aligned}$$
(12)

On the other hand, we have

$$\begin{aligned} I(f^{k,s}_\psi )= & {} \int _{{\mathbb {R}}^{n-1}}\Bigg (\int _{{\mathbb {R}}}\Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big ){\overline{\psi }}\\&\Big (g^{-1}\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert dt\Bigg )^2 da \\= & {} \int _{{\mathbb {R}}^{n-1}}\Bigg (\int _{{\mathbb {R}}}\Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert \\&\times \Big \vert \psi \Big (\exp \Big ( \big (t-k+Q_1(a,s)\big )X_1+ \sum _{j=2}^{n} Q_j(a,s)X_j\Big )\Big )\Big \vert dt\Bigg )^2 da \end{aligned}$$

(using (10))

$$\begin{aligned}&\le \int _{{\mathbb {R}}^{n-1}}\Bigg (\int _{{\mathbb {R}}}{\Big \vert \psi \Big (\exp \Big ( \big (t-k+Q_1(a,s)\big )X_1+ \sum _{j=2}^{n} Q_j(a,s)X_j\Big )\Big )\Big \vert ^2\over 1+\big \vert t- r_0+R(a,\sigma _0)\big \vert ^2}dt\Bigg ) \\&\quad \times \Bigg ( \int _{{\mathbb {R}}} \Big (1+\big \vert t- r_0+R(a,\sigma _0)\big \vert ^2\Big ) \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2 dt\Bigg )da \end{aligned}$$

(using Cauchy–Shwartz inequality),

$$\begin{aligned}&\le \int _{{\mathbb {R}}^{n-1}}\Bigg (\int _{{\mathbb {R}}}\Big \vert \psi \Big (\exp \Big ( \big (t-k+Q_1(a,s)\big )X_1+ \sum _{j=2}^{n} Q_j(a,s)X_j\Big )\Big )\Big \vert ^2dt\Bigg ) \\&\qquad \times \Bigg ( \int _{{\mathbb {R}}} \Big (1+\big \vert t- r_0+R(a,\sigma _0)\big \vert ^2\Big ) \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2 dt\Bigg )da \\&\quad = \int _{{\mathbb {R}}^{n-1}}\Bigg (\int _{{\mathbb {R}}}\Big \vert \psi \Big (\exp \Big (zX_1+ \sum _{j=2}^{n} Q_j(a,s)X_j\Big )\Big )\Big \vert ^2dz\Bigg ) \\&\qquad \times \Bigg ( \int _{{\mathbb {R}}} \Big (1+\big \vert t- r_0+R(a,\sigma _0)\big \vert ^2\Big ) \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2 dt\Bigg )da \end{aligned}$$

(using the change of variable \(z=t-k+Q_1(a,s)\)). By substituting \(Q_j(a,s)\) for \(s_j\), \(j=2,...,n\), using Eq. (11), we have

$$\begin{aligned}&\int _{{\mathbb {R}}^{n-1}} I(f^{k,s}_\psi )\, ds\le \Vert \psi \Vert _2^2\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}} \Big (1+\big \vert t- r_0+R(a,\sigma _0)\big \vert ^2 \Big )\\&\quad \Big \vert f\Big (\exp \Big (t X_1+\sum _{j=2}^{n}a_jX_j\Big )\Big )\Big \vert ^2 dt\, da, \end{aligned}$$

which is finite by (12). This implies that, \(I(f^{k,s}_\psi )\) is finite for all \(k\in {\mathbb {R}}\) and almost all \(s\in {\mathbb {R}}^{n-1}\). \(\square \)

Before stating the next lemma, we need a localized version of the Plancherel measure (see [4]). Let \(Z=\exp {\mathfrak {z}}\) be the center of G and fix a nonzero vector \(X_1\) of \(\mathfrak {z}\). Let \(A=\exp \mathfrak {a}= \exp {\mathbb {R}}X_1\) be the closed connected subgroup of Z and \(\chi =\chi _\zeta \), \(\zeta =l_1X_1^* \in \mathfrak {a}^{ * }\), be the unitary character of A, defined by

$$\begin{aligned} \chi _\zeta (\exp tX_1)=\text {e}^{-2i\pi tl _1}. \end{aligned}$$

Let \({{\hat{G}}}_\chi =\{ \pi \in {{\hat{G}}} \ : \ \pi _{ \vert A}=\chi \cdot I \}\). For \(1\le p<+\infty \), let \(L^p(G/A, \zeta )\) be the set of all measurable functions \(\varphi :G\rightarrow {\mathbb {C}}\) such that \(\varphi (xa)={\overline{\chi }}(a)\varphi (x)\) for almost all \(x\in G\) and \(a\in A\) and such that

$$\begin{aligned} \Vert \varphi \Vert _{L^p(G/A)}^p=\int _{G/A}\vert \varphi (x)\vert ^pd{\dot{x}}<+\infty . \end{aligned}$$

Moreover, let \({\mathfrak {g}^*_\zeta }=\zeta + \mathfrak {a}^{\perp }\) and \({\mathcal {W}}_{ \zeta } ={\mathcal {W}}\cap {\mathfrak {g}_\zeta ^*}\). In this case, the Plancherel formula reads: if

$$\begin{aligned} \pi (\varphi )=\int _{G/A}\varphi (x)\pi (x)d{\dot{x}},\ \pi \in {{\hat{G}}}_\chi , \end{aligned}$$

then, for \(\varphi \in L^1( G/A, \zeta )\cap L^2( G/A, \zeta )\) we have:

$$\begin{aligned} \Vert \varphi \Vert _2=\Big (\int _{{{\mathcal {W}}}_{\zeta }}\Vert \pi _{ l}(\varphi ) \Vert ^2_{HS}\vert Pf(l)\vert dl\Big )^{1\over 2}. \end{aligned}$$
(13)

If d is the maximal dimension of coadjoint orbits in \(\mathfrak {g}^*\), then T has \(n-d\) elements and thus \(V_T\) can be identified with \({\mathbb {R}}^{n-d}\). We can identify \(V_T\) with \({\mathbb {R}}X_1^*\oplus {\mathbb {R}}^{n-d-1}\). We denote by

$$\begin{aligned} \ p_*:V_T \rightarrow {\mathbb {R}}X_1^*, \ \ \ l\mapsto l_1 X_1^* \end{aligned}$$

the canonical projection. As \({\mathcal {W}}\) is a Zariski open set of \(V_T\), \(p_*({\mathcal {W}})\) is also a nonempty Zariski open set of \({\mathbb {R}}\). Then it will be convenient to write elements \(l\in {\mathcal {W}}\), as \((l_1,l')\) where \(l_1\in \ p_*({\mathcal {W}})\) and \(l'\in {\mathcal {W}}_{{l_1}}=\{ l' \in {\mathbb {R}}^{n-d-1} \ : \ (l_1,l') \in {\mathcal {W}} \}\). It turns out that \({\mathcal {W}}_{{l_1}}\) is also a Zariski open set of \({\mathbb {R}}^{n-d-1}\) for each \(l_1\in p_*({\mathcal {W}})\). The set \({\mathcal {W}}_{{l_1}}\) corresponds obviously to the cross-section \({{\mathcal {W}}}_{\zeta }\) used in the localized version of the Plancherel formula in (13). On the other hand, we obtain a decomposition of the Plancherel measure: for a function \(F\in C_c({\mathcal {W}})\), we have

$$\begin{aligned} \int _{\mathcal {W}}F(l)dl=\int _{{\mathbb {R}}}\int _{{\mathcal {W}}_{l_1}}F(l)dl'\, dl_1, \end{aligned}$$
(14)

where the measure \(dl'\) is induced on \({\mathcal {W}}_{l_1}\) by the Lebesque measure on \({\mathcal {W}}\).

Lemma 3

Let \(f,\psi \in L^2(G)\) satisfying condition (2) of Theorem 3 and \(\gamma \in ]0,\beta [\). Then there exists \(c>0\), such that

$$\begin{aligned} \int _{{\mathbb {R}}^{n-1}}\left| \widehat{(f_\psi ^{k,s})_a}(l_1) \right| ^2da \le c\exp (-\pi (\alpha k^2+\alpha \Vert s\Vert ^2+\gamma l_1^2)), \end{aligned}$$

for all \(k, l_1\in {\mathbb {R}}\) and almost all \(s\in {\mathbb {R}}^{n-1}\).

Proof

Let h(ks) be the function defined on \({\mathbb {R}}\) by

$$\begin{aligned} h(k,s)(\lambda )=\int _{{\mathbb {R}}^{n-1}} ((f_\psi ^{k,s})_a*(f_\psi ^{k,s})_a^\star )(\lambda )da, \end{aligned}$$

where \(\lambda \in {\mathbb {R}}\) and \((f_\psi ^{k,s})_a^\star (\lambda )=\overline{(f_\psi ^{k,s})_a(-\lambda )}\). Then, \(h(k,s)\in L^1({\mathbb {R}})\), for all \(k\in {\mathbb {R}}\) and almost all \(s\in {\mathbb {R}}^{n-1}\). In fact, from Lemma 2

$$\begin{aligned} \int _{\mathbb {R}}\vert h(k,s)(\lambda )\vert d\lambda\le & {} \int _{\mathbb {R}}\int _{{\mathbb {R}}^{n-1}}\int _{\mathbb {R}}\big \vert (f_\psi ^{k,s})_a(t)\big \vert \big \vert (f_\psi ^{k,s})_a(t-\lambda )\big \vert dt\, da\, d\lambda \nonumber \\= & {} \int _{{\mathbb {R}}^{n-1}}\Bigg (\int _{\mathbb {R}}\big \vert (f_\psi ^{k,s})_a(t)\big \vert dt\Bigg )^2da<\infty , \end{aligned}$$
(15)

for all \(k\in {\mathbb {R}}\) and almost all \(s\in {\mathbb {R}}^{n-1}\). Thus,

$$\begin{aligned} \widehat{h(k,s)}(l_1)=\int _{{\mathbb {R}}^{n-1}}\big \vert \widehat{(f_\psi ^{k,s})_a}(l_1) \big \vert ^2da. \end{aligned}$$
(16)

Identifying \(A=\exp {\mathbb {R}}X_1\) with \({\mathbb {R}}\). Following the idea of Kaniuth and Kumar [14], for \(u\in L^1(A)\cap L^2(A)\) define \(u*f_\psi ^{k,s}\) on G by

$$\begin{aligned} u*f_\psi ^{k,s}(x)=\int _{A}u(z)f_\psi ^{k,s}(z^{-1}x)\,dz \end{aligned}$$

and then \(h(k,s)_u: {\mathbb {R}}\rightarrow {\mathbb {C}}\) by

$$\begin{aligned} h(k,s)_u(t)=\int _{{\mathbb {R}}^{n-1}}\Big ((u*f_\psi ^{k,s})_a* \big ((u*f_\psi ^{k,s})_a\big )^\star \Big )(t)\,da. \end{aligned}$$

It is not hard to see that

$$\begin{aligned} h(k,s)_u(t)=\int _{{\mathbb {R}}^{n-1}} \big ((u*(f_\psi ^{k,s})_a)*(u*(f_\psi ^{k,s})_a)^\star \big )(t)\,da. \end{aligned}$$

Therefore, for every \(\eta _1\in {\mathbb {R}}\)

$$\begin{aligned} \widehat{h(k,s)_u}(\eta _1)&=\int _{{\mathbb {R}}^{n-1}} \Big \vert \big (u*(f_\psi ^{k,s})_a\big )^{\widehat{}}(\eta _1)\Big \vert ^2\,da \nonumber \\&=\vert {\hat{u}}(\eta _1)\vert ^2\int _{{\mathbb {R}}^{n-1}} \Big \vert \widehat{(f_\psi ^{k,s})_a}(\eta _1)\Big \vert ^2\,da =\vert {\hat{u}}(\eta _1)\vert ^2\widehat{h(k,s)}(\eta _1) \end{aligned}$$
(17)

(using Eq. (16)). Hence,

$$\begin{aligned} \int _{\mathbb {R}}\big \vert \widehat{h(k,s)_u}(\eta _1)\big \vert d\eta _1&=\int _{{\mathbb {R}}}\vert {\hat{u}}(\eta _1)\vert ^2\big \vert \widehat{h(k,s)}(\eta _1)\big \vert d\eta _1 \\&\le \int _{\mathbb {R}} \vert {\hat{u}}(\eta _1)\vert ^2 \Vert h(k,s) \Vert _1d\eta _1=\Vert u\Vert _2^2\Vert h(k,s) \Vert _1, \end{aligned}$$

which is finite by (15). By the inversion formula for \({\mathbb {R}}\), we have

$$\begin{aligned}&\int _{{\mathbb {R}}}\widehat{h(k,s)_u}(\eta _1)\,d\eta _1 =h(k,s)_u(0)\nonumber \\&\qquad =\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}} \big \vert u*(f_\psi ^{k,s})_a (t)\big \vert ^2\,dt\,da =\Vert u*f_\psi ^{k,s}\Vert _2^2. \end{aligned}$$
(18)

Now fix \(l_1\in {\mathbb {R}}\) and let \(u_m\in L^1(A) \), \(m \in {\mathbb {N}}^*\) such that \(\widehat{u_m}(\eta _1)=1\) for \(\eta _1\in V_m(l_1) =[l_1-(1/ 2m),l_1+(1/ 2m)]\) and \(\widehat{u_m}(\eta _1)=0\) on the complement of \(V_m(l_1)\). Noting that, \(u_m\) is also in \(L^2(A)\). Indeed, since \(\widehat{u_m}\in L^1(A)\cap L^2(A)\),

$$\begin{aligned} \Vert u_m\Vert _2=\Vert {\check{u}}_m\Vert _2=\big \Vert \widehat{ \widehat{u_m}}\big \Vert _2=\Vert \widehat{u_m}\Vert _2<\infty , \end{aligned}$$

where \({\check{u}}_m(z)=u_m(-z)\).

As \(\widehat{h(k,s)}\) is continuous and \(V_m(l_1)\) has length 1/m, we have: for all \(k\in {\mathbb {R}}\) and almost all \(s\in {\mathbb {R}}^{n-1}\),

$$\begin{aligned} \widehat{h(k,s)}(l_1)&=\lim _{m\rightarrow \infty }m \int _{V_m(l_1)}\widehat{h(k,s)}(\eta _1)\,d\eta _1 =\lim _{m\rightarrow \infty }m \int _{{\mathbb {R}}}\widehat{h(k,s)}(\eta _1) {{\textbf {1}}}_{V_m(l_1)}\,d\eta _1 \\&= \lim _{m\rightarrow \infty }m \int _{{\mathbb {R}}}\widehat{h(k,s)_{u_m}}(\eta _1) \,d\eta _1\qquad (\text {using Eq.} (17))\\&=\lim _{m\rightarrow \infty }m \Vert u_m*f_\psi ^{k,s} \Vert _2^2 \qquad (\text {using Eq.} (18)) \\ {}&=\lim _{m\rightarrow \infty }m \int _{{\mathbb {R}}} \int _{{\mathcal {W}}_{\eta _1}}\vert Pf(\eta )\vert \Vert \pi _\eta (u_m*f_\psi ^{k,s}) \Vert ^2_{HS} \,d\eta '\, d\eta _1 \\&\qquad \qquad \qquad \qquad \qquad \qquad \quad (\text {using Eqs.} (5)\, \hbox {and}\, (14)) \\&=\lim _{m\rightarrow \infty }m \int _{{\mathbb {R}}}\vert \widehat{u_m}(\eta _1) \vert ^2\biggl (\int _{{\mathcal {W}}_{\eta _1}}\vert Pf(\eta )\vert \Vert \pi _\eta (f_\psi ^{k,s}) \Vert ^2_{HS} \,d\eta '\biggl )\,d\eta _1\\&=\lim _{m\rightarrow \infty }m \int _{V_m(l_1)}I\eta _1\,d\eta _1, \end{aligned}$$

where \(\eta =(\eta _1,\eta ')\) and \(\displaystyle I\eta _1=\int _{{\mathcal {W}}_{\eta _1}}\vert Pf(\eta )\vert \Vert \pi _\eta (f_\psi ^{k,s}) \Vert ^2_{HS} \,d\eta '\).

Now by (2) and (8), we obtain

$$\begin{aligned} I\eta _1 \le C \int _{{\mathcal {W}}_{\eta _1}} \vert Pf(\eta )\vert \exp ({-\pi (\alpha k^2+\alpha \Vert s\Vert ^2+\beta \Vert \eta \Vert ^2) }) d\eta '. \end{aligned}$$

Since \(Pf(\eta )\) is a polynominal function of \(\eta \), there exists \(R>0\) such that

$$\begin{aligned} \vert {Pf(\eta )}\vert \exp ({-\pi (\beta -\gamma ) \Vert \eta \Vert ^2 })\le 1 \end{aligned}$$

for all \(\eta \in \mathfrak {g} ^*\) with \(\Vert {\eta }\Vert \ge R\). Let \(K\ge 1\) such that \(\vert Pf(\eta )\vert \le K\) for all \(\eta \in \mathfrak {g}^*\) with \(\Vert {\eta }\Vert \le R\). It follows that,

$$\begin{aligned} I\eta _1&\le C K \exp (-\pi (\alpha k^2+\alpha \Vert s\Vert ^2))\int _{{\mathcal {W}}_{\eta _1}} \exp \big ({-\pi \gamma \Vert \eta \Vert ^2 }\big ) d\eta ' \\&= C K \exp ({-\pi (\alpha k^2+\alpha \Vert s\Vert ^2)})\int _{{\mathbb {R}}^{n-d-1}} \exp \big ({-\pi \gamma \Vert (\eta _1,\eta ')\Vert ^2}\big ) d\eta ' \\&= C K \exp (-\pi (\alpha k^2+\alpha \Vert s\Vert ^2+\gamma \eta _1^2))\int _{{\mathbb {R}}^{n-d-1}} \exp \big ({-\pi \gamma \Vert \eta '\Vert ^2}\big ) d\eta '\\&= c \exp (-\pi (\alpha k^2+\alpha \Vert s\Vert ^2+\gamma \eta _1^2)), \end{aligned}$$

for some \(c>0\). Therefore,

$$\begin{aligned} \widehat{h(k,s)}(l_1)&\le c\lim _{m\rightarrow \infty }m\int _{V_m(l_1)} \exp ({-\pi (\alpha k^2+\alpha \Vert s\Vert ^2+\gamma \eta _1^2)})d\eta _1 \\ {}&=c \exp ({-\pi ( \alpha k^2+\alpha \Vert s\Vert ^2+\gamma l_1^2)}). \end{aligned}$$

Finally, Eq. (16) allows us to conclude. \(\square \)

4 Proof of Theorem 3

For \(a=(a_2,...,a_n), s=(s_2,...,s_n)\in {\mathbb {R}}^{n-1}\), let \(f_{a,s}\),\(\psi _{a,s}\) be the complex-valued functions defined on \({\mathbb {R}}\) by

$$\begin{aligned} f_{a,s}(t)= & {} f\Big ( \exp \Big ( \big (t-Q_1(a,s)\big )X_1+ \sum _{j=2}^{n} a_jX_j \Big )\Big ),\\&\text {and} \ \ \psi _{a,s}(t)=\psi \Big (\exp \Big ( tX_1+ \sum _{j=2}^{n} Q_j(a,s)X_j \Big ) \Big ), \end{aligned}$$

where the polynomial functions \(Q_j\) are defined as in (10). Then obviously \(f_{a,s},\psi _{a,s}\in L^2({\mathbb {R}})\), for almost all \(a\in {\mathbb {R}}^{n-1}\) and all \(s\in {\mathbb {R}}^{n-1}\). For fixed \(\lambda \), \(y\in {\mathbb {R}}\), let \(F_{\lambda ,y}\) and \(K_{\lambda ,y}\) be the functions defined on \({\mathbb {R}}\times {\mathbb {R}}\) by

$$\begin{aligned} F_{\lambda ,y}(k,l_1)={\mathcal {G}}_{\psi _{a,s}}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(k,l_1) {\mathcal {G}}_{\psi _{a,s}}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(-k,-l_1)e^{2i\pi k l_1}, \end{aligned}$$

and

$$\begin{aligned} K_{\lambda ,y}(k,l_1)=\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}^{n-1}}F_{\lambda ,y}(k,l_1) \phi (a,s) da\, ds, \end{aligned}$$

where \(\phi \in {\mathcal {S}}({\mathbb {R}}^{n-1}\times {\mathbb {R}}^{n-1})\), the Schwartz space of \({\mathbb {R}}^{n-1}\times {\mathbb {R}}^{n-1}\). Now for fixed \(\mu \in {\mathbb {R}}\), let \(R_{\lambda ,y,\mu }\) be the function defined on \({\mathbb {R}}\) by

$$\begin{aligned} R_{\lambda ,y,\mu }(k)=K_{\lambda ,y}(k,.)^{\widehat{}}(\mu ), \end{aligned}$$
(19)

where \(K_{\lambda ,y}(k,.)^{\widehat{}}\) is the partial Fourier transform of \(K_{\lambda ,y}\) with respect the second variable \(l_1\). It follows, using Lemma 1, that

$$\begin{aligned} {\widehat{R}}_{\lambda ,y,\mu }(w)={\widehat{K}}_{\lambda ,y}(w,\mu )= & {} \displaystyle \int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}^{n-1}}{\hat{F}}_{\lambda ,y}(w,\mu )\phi (a,s)da\, ds\nonumber \\= & {} \int _{{\mathbb {R}}^{n-1}}\displaystyle \int _{{\mathbb {R}}^{n-1}}F_{\lambda ,y}(-\mu ,w)\phi (a,s)da\, ds=K_{\lambda ,y}(-\mu ,w).\nonumber \\ \end{aligned}$$
(20)

Lemma 4

There exists a positive constant \(C_1\) such that

$$\begin{aligned} \Big \vert R_{\lambda ,y,\mu }(k)\Big \vert \le C_1 \, e^{-\alpha \pi k^2}. \end{aligned}$$

Moreover, the constant \(C_1\) does not depend on \(\lambda \), \(\mu \) and y.

Proof

From Eq. (19) we have,

$$\begin{aligned}&\vert R_{\lambda ,y,\mu }(k)\vert \\&\quad =\vert K_{\lambda ,y} (k,.)^{\widehat{}}(\mu )\vert \\&\quad \le \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}^{n-1}}\Big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(k,l_1)\Big \vert \\&\qquad \Big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(-k,-l_1)\Big \vert \Big \vert \phi (a,s)\Big \vert da\, ds\, dl_1\\&\quad \le \text {cst} \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}^{n-1}}\Big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(k,l_1)\Big \vert \\&\qquad \Big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(-k,-l_1)\Big \vert da\, ds\, dl_1 . \end{aligned}$$

By using Cauchy-Schwartz inequality, we obtain

$$\begin{aligned} \vert R_{\lambda ,y,\mu }(k)\vert\le & {} \text {cst}\left( \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}}\left( \int _{{\mathbb {R}}^{n-1}}\big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(k,l_1)\big \vert ^2da\right) ds\,dl_1\right) ^{\frac{1}{2}}\\&\times \left( \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}}\left( \int _{{\mathbb {R}}^{n-1}}\big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(-k,-l_1)\big \vert ^2da\right) ds\,dl_1\right) ^{\frac{1}{2}}. \end{aligned}$$

Remark that,

$$\begin{aligned} \Big \vert {\mathcal {G}}_{\psi _{a,s}}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(k,l_1)\Big \vert =\Big \vert {\mathcal {G}}_{\psi _{a,s}}f_{a,s}(k-y,l_1-\lambda )\Big \vert \end{aligned}$$

(using i) in Lemma 1)

$$\begin{aligned}&=\Bigg \vert \displaystyle \int _{{\mathbb {R}}}f_{a,s}(t)\overline{\psi _{a,s}}(t-k+y)\text {e}^{-2i\pi t(l_1-\lambda )}dt\Bigg \vert \\&\qquad =\Bigg \vert \int _{{\mathbb {R}}}f\left( \exp \left( (t-Q_1(a,s))\right) X_1+ \sum _{j=2}^{n} a_jX_j \right) \\&\qquad \times \ {\overline{\psi }}\left( \exp \left( ( t-k+y)X_1+ \sum _{j=2}^{n} Q_j(a,s)X_j \right) \right) \text {e}^{-2i\pi t(l_1-\lambda )}dt\Bigg \vert \\&\quad =\Bigg \vert \int _{{\mathbb {R}}}f\left( \exp \left( rX_1+ \sum _{j=2}^{n} a_jX_j \right) \right) \\&\qquad \times {\overline{\psi }}\left( \exp \left( ( r-(k-y)+Q_1(a,s))X_1+ \sum _{j=2}^{n} Q_j(a,s)X_j \right) \right) \text {e}^{-2i\pi r(l_1-\lambda )}dr\Bigg \vert \end{aligned}$$

(by substituting \(r=t-Q_1(a,s)\) for t)

$$\begin{aligned}&=\Big \vert \int _{{\mathbb {R}}}f\left( \exp \left( rX_1+ \sum _{j=2}^{n} a_jX_j \right) \right) \\&\quad \times {\overline{\psi }}\left( \exp \left( (k-y) X_1+ \sum _{j=1}^{n} s_jX_j \right) ^{-1}\exp \left( r X_1+\sum _{j=2}^{n} a_jX_j\right) \right) \text {e}^{-2i\pi r(l_1-\lambda )}dr\Big \vert \end{aligned}$$

(using Eq. (10))

\(=\displaystyle \left| \int _{{\mathbb {R}}}f_\psi ^{k-y,s}\left( \exp \left( rX_1+ \sum _{j=2}^{n} a_jX_j \right) \right) \text {e}^{-2i\pi r(l_1-\lambda )}dr\right| =\left| \widehat{(f_\psi ^{k-y,s})_a}(l_1-\lambda )\right| \).

It results that,

$$\begin{aligned} \big \vert R_{\lambda ,y,\mu }(k) \big \vert\le & {} \text {cst}\left( \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}} \left( \int _{{\mathbb {R}}^{n-1}}\big \vert \widehat{(f_\psi ^{k-y,s})_a}(l_1 - \lambda )\big \vert ^2da\right) ds\,dl_1\right) ^{\frac{1}{2}}\\&\times \left( \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}} \int _{{\mathbb {R}}^{n-1}}\big \vert \widehat{(f_\psi ^{-k-y,s})_a}(-l_1 - \lambda )\big \vert ^2da\,ds\,dl_1\right) ^{\frac{1}{2}} \\\le & {} \text {cst} \left( \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}} e^{-\pi (\alpha \vert k-y\vert ^2+\gamma (l_1-\lambda )^2+\alpha \Vert s\Vert ^2)} ds\,dl_1\right) ^{\frac{1}{2}}\\&\times \left( \displaystyle \int _{{\mathbb {R}}}\int _{{\mathbb {R}}^{n-1}} e^{-\pi (\alpha \vert k+y\vert ^2+\gamma (l_1+\lambda )^2+\alpha \Vert s \Vert ^2)} ds\,dl_1\right) ^{\frac{1}{2}} \quad (\text {using Lemma}\, 3) \\\le & {} \text {cst}\, e^{-\frac{\pi }{2}\alpha (\vert k-y\vert ^2+\vert k+y\vert ^2)}\left( \int _{{\mathbb {R}}^{n-1}} e^{-\pi \alpha \Vert s\Vert ^2} ds\right) \left( \int _{{\mathbb {R}}}e^{-\pi \gamma l_1^2} dl_1\right) \\\le & {} \text {cst}\, e^{-\pi \alpha k^2}, \end{aligned}$$

which is the desired result. \(\square \)

Lemma 5

There exists a positive constant \(C_2\) such that

$$\begin{aligned} \big \vert {\hat{R}}_{\lambda ,y,\mu }(w)\big \vert \le C_2\, e^{-\pi \gamma w^2}. \end{aligned}$$

Moreover, the constant \(C_2\) does not depend on \(\lambda \), y and \(\mu \).

Proof

By (20), we have

$$\begin{aligned}&\big \vert {\hat{R}}_{\lambda ,y,\mu }(w)\big \vert =\big \vert K_{\lambda ,y}(-\mu ,w)\big \vert \\&\quad \le \text {cst}\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}^{n-1}}\Big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(-\mu ,w)\Big \vert \Big \vert {\mathcal {G}}_{\psi _a,s}({\mathcal {M}}_\lambda {\mathcal {T}}_yf_{a,s})(\mu ,-w)\Big \vert dads. \end{aligned}$$

As in the proof of the Lemma 4 we can show that,

$$\begin{aligned}&\big \vert {\hat{R}}_{\lambda ,y,\mu }(w)\big \vert \le \text {cst}\, e^{-\frac{\pi \gamma }{2}(\vert w-\lambda \vert ^2+\vert w+\lambda \vert ^2)}e^{\frac{-\pi a}{2}(\vert \mu +y\vert ^2+\vert -\mu + y\vert ^2) } \times \left( \int _{{\mathbb {R}}^{n-1}} e^{-\pi \alpha \Vert s \Vert ^2 } ds\right) \\&\quad \le \text {cst}\, e^{-\pi \gamma w^2}, \end{aligned}$$

which is the desired result. \(\square \)

We have shown finally that \(R_{\lambda ,y,\mu }\) verifies the decay conditions of Hardy theorem on \({\mathbb {R}}\). Since \(\alpha \beta >1\), we can choose \(0<\gamma < \beta \) such that \(\alpha \gamma >1\). We conclude that \(R_{\lambda ,y,\mu }=0\) a.e. and \({\hat{R}}_{\lambda ,y,\mu }=0\) for all \(\lambda , y, \mu \in {\mathbb {R}}\). In (20), allowing \(\phi \) to vary through the space of Schwartz functions on \({\mathbb {R}}^{n-1}\times {\mathbb {R}}^{n-1}\), we obtain \(F_{\lambda ,y}(-\mu ,w)=0\) for all \( \lambda , y, \mu \) in \({\mathbb {R}}\) and almost all \(w \in {\mathbb {R}}\). As \(F_{-\lambda ,-y}\) is continuous on \({\mathbb {R}}\times {\mathbb {R}}\),

$$\begin{aligned} \vert F_{-\lambda ,-y}(0,0)\vert =\vert G_{\psi _{a,s}f_{a,s}(y,\lambda )}\vert ^2=0 \end{aligned}$$

(using i) in Lemma 1). Hence, \({\mathcal {G}}_{\psi _{a,s}}f_{a,s}=0\) a.e. By using Eq. (4), we have

$$\begin{aligned} \Vert \psi _{a,s}\Vert _2^2 \Vert f_{a,s}\Vert _2^2 =0, \end{aligned}$$

which implies either \(\psi _{a,s}=0\) a.e. or \(f_{a,s}=0\) a.e. Observe that,

$$\begin{aligned}&\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}^{n-1}} \Vert f_{a,s} \Vert _2^2\Vert \psi _{a,s} \Vert _2^2da\, ds \\&\quad =\int _{{\mathbb {R}}^{n-1}}\int _{{\mathbb {R}}^{n-1}}\left( \int _ {\mathbb {R}}\Big \vert f\left( \exp \left( \left( t-Q_1(a,s)\right) X_1+ \sum _{j=2}^{n} a_jX_j \right) \right) \Big \vert ^2 dt\right) \\&\qquad \times \left( \int _ {\mathbb {R}} \Big \vert \psi \left( \exp \left( tX_1+ \sum _{j=2}^{n} Q_j(a,s)X_j \right) \right) \Big \vert ^2 dt\right) da\, ds =\Vert f \Vert _2^2\Vert \psi \Vert _2^2 \end{aligned}$$

(by substituting \(t-Q_1(a,s)\) for t and \(Q_j(a,s)\) for \(s_j\), \(j=2,...,n\), using Eq. (11)). This allow us to achieve the proof.