1 Introduction and Result

The Hermite polynomials \(H_n(x)\) (\(n\ge 0\)) are defined by the orthogonal relation (cf. [15, 16])

$$\begin{aligned} \int _{-\infty }^{\infty }e^{-x^2}H_n(x)H_m(x)dx=\pi ^{\frac{1}{2}}2^n n!\delta _{nm}, \end{aligned}$$

and the Hermite functions are given by

$$\begin{aligned} {\mathcal {H}}_n(x)=\left( \pi ^{\frac{1}{2}}2^n n!\right) ^{-\frac{1}{2}}e^{-\frac{x^2}{2}}H_n(x), \quad n=0,1,2,\ldots , \end{aligned}$$

which are orthonormal over \((-\infty ,\infty )\) associated with the Lebesgue measure.

For \(f\in L({\mathbb {R}})\), its Hermite expansion is

$$\begin{aligned} f(x)\sim \sum _{n=0}^{\infty }a_n(f){\mathcal {H}}_n(x),\quad a_n(f)=\int _{-\infty }^{\infty }f(x){\mathcal {H}}_n(x)dx. \end{aligned}$$

The Hardy type inequality for Hermite expansions of functions in the Hardy space \(H^1({\mathbb {R}})\) has been studied in several works. The first step was done by Kanjin [5] who proved the inequality

$$\begin{aligned} \sum _{n=1}^{\infty }n^{-\frac{29}{36}}|a_n(f)|\le A\Vert f\Vert _{H^1({\mathbb {R}})}, \quad f\in H^1({\mathbb {R}}), \end{aligned}$$
(1)

where \(A>0\) is a constant independent of \(f\). Balasubramanian and Radha [2] extended Kanjin’s result to \(H^p({\mathbb {R}})\), \(0<p\le 1\). Radha and Thangavelu [10] (cf [17] also) obtained inequalities of Hardy type for \(d\)-dimensional Hermite and special Hermite expansions for \(d\ge 2\), where the constant they determined, in place of \(\frac{29}{36}\), is

$$\begin{aligned} \sigma =\left( \frac{d}{2}+1\right) \left( \frac{2-p}{2}\right) \end{aligned}$$

for \(H^p({\mathbb {R}}^d)\), \(0<p\le 1\). In comparison with the case of \(d\)-dimension (\(d\ge 2\)), the Hardy inequality for one-dimensional Hermite expansions should be the one as (1) but with \(\frac{3}{4}\) instead of \(\frac{29}{36}\). However the method in [10] does not work for \(d=1\). An improved form of (1) with \(\frac{3}{4}+\epsilon \) for \(\epsilon >0\) in place of \(\frac{29}{36}\) was obtained by Kanjin [6]. Moreover Kanjin [6] proved his inequality for all \(f\in L^1({\mathbb {R}})\), that is,

$$\begin{aligned} \sum _{n=1}^{\infty }n^{-\frac{3}{4}-\epsilon }|a_n(f)|\le A\Vert f\Vert _{L^1({\mathbb {R}})}. \end{aligned}$$
(2)

This again leads Kanjin to conjecture that the possible form of the Hardy inequality for Hermite expansions would be

$$\begin{aligned} \sum _{n=1}^{\infty }n^{-\frac{3}{4}}|a_n(f)|\le A\Vert f\Vert _{H^1({\mathbb {R}})}. \end{aligned}$$
(3)

We shall give a positive answer to this conjecture in the present paper. Kanjin [6] also showed that there exists a function \(f_0\in L^1({\mathbb {R}})\) such that

$$\begin{aligned} \sum _{n=1}^{\infty }n^{-\frac{3}{4}}|a_n(f_0)|=\infty , \end{aligned}$$

so the Hardy space norm is required in (3).

The proofs of (1) and (2) in [5, 6] were based on the pointwise estimate of the Hermite functions as follows: for given \(\tau >0\), there exist positive constants \(A\), \(\eta \) and \(\xi \) such that

$$\begin{aligned} |{\mathcal {H}}_n(x)|\le A\left( |x|+\sqrt{n}\right) ^{-\frac{1}{4}} \left( ||x|-\sqrt{2n}|+n^{-\frac{1}{6}}\right) ^{-\frac{1}{4}} \Psi _n(x) \end{aligned}$$
(4)

holds for all \(x\in {\mathbb {R}}\) and \(n\ge 1\), where

$$\begin{aligned} \Psi _n(x)=\left\{ \begin{array}{ll} 1,&{}\quad \hbox {for}\quad 0\le |x|\le \sqrt{2n};\\ \exp \left( -\eta n^{\frac{1}{4}}||x|-\sqrt{2n}|^{3/2}\right) ,&{}\quad \hbox {for}\quad \sqrt{2n}\le |x|\le (1+\tau )\sqrt{2n};\\ e^{-\xi x^2},&{}\quad \hbox {for}\quad (1+\tau )\sqrt{2n}\le |x|. \end{array}\right. \end{aligned}$$
(5)

A separate description of (4) is given in [1] due to Skovgaard. The unified and simplified form as (4) is stated in [7], in virtue of the relation [16], (1.1.52), (1.1.53)] of Hermite polynomials and Laguerre polynomials and a unified description [9], (2.2)] of Laguerre polynomials based on [8] and the table in [1, p. 699].

A direct consequence of (4) and (5) is

$$\begin{aligned} |{\mathcal {H}}_n(x)|\le An^{-\frac{1}{12}}, \end{aligned}$$

and \({\mathcal {H}}_n(x)\) attains this bound near the point \(x=\sqrt{2n}\). But for most \(x\) it has a much smaller bound as a multiple of \(n^{-\frac{1}{4}}\). It is a very hard work to apply such a non-proportional property of the Hermite functions as Kanjin did in [5, 6], and certainly, it is also difficult to achieve the best result for related problems. However, if for \(d\ge 2\), we denote by \(\Phi _{\alpha }\), \(\alpha \in {\mathbb {N}}^d\), the \(d\)-dimensional Hermite functions, namely,

$$\begin{aligned} \Phi _{\alpha }(x_1,\ldots ,x_d)={\mathcal {H}}_{\alpha _1}(x_1)\cdots {\mathcal {H}}_{\alpha _d}(x_d),\quad \alpha =(\alpha _1,\ldots ,\alpha _d), \end{aligned}$$

then there exists a constant \(A>0\) independent of \(n\) and \((x_1,\ldots ,x_d)\) such that (see [16, Lemma 3.2.2])

$$\begin{aligned} \sum _{|\alpha |=n}|\Phi _{\alpha }(x_1,\ldots ,x_d)|^2\le A(n+1)^{\frac{d}{2}-1}. \end{aligned}$$
(6)

Obviously this is not true for \(d=1\). The bound in (6) has been used in research of various problems for \(d\ge 2\), as in [16] for example; it is also the key in the proof of the inequalities of Hardy type in [10] for \(d\)-dimensional Hermite expansions for \(d\ge 2\).

In order to prove the Hardy inequality (3), we shall follow a different approach, by evaluating the square integration of the Poisson integral associated to Hermite expansions of functions in \(H^1({\mathbb {R}})\).

Indeed, we shall work with the generalized Hermite expansions of functions in \(H^1({\mathbb {R}})\). If \(\lambda >-1/2\), the generalized Hermite polynomials \(H_n^{(\lambda )}(x)\) (\(n\ge 0\)) are defined by (see [3])

$$\begin{aligned} H_{2k}^{(\lambda )}(x)=&\left( \frac{k!}{\Gamma (k+\lambda +1/2)}\right) ^{1/2}L_k^{(\lambda -1/2)}(x^2),\end{aligned}$$
(7)
$$\begin{aligned} H_{2k-1}^{(\lambda )}(x)=&\left( \frac{(k-1)!}{\Gamma (k+\lambda +1/2)}\right) ^{1/2} xL_{k-1}^{(\lambda +1/2)}(x^2), \end{aligned}$$
(8)

where \(L_n^{(\alpha )}(x)\) (\(\alpha >-1\), \(n\ge 0\)) are the Laguerre polynomials determined by the orthogonal relation (see [15, 16])

$$\begin{aligned} \int _0^{\infty }e^{-x}x^{\alpha }L_n^{(\alpha )}(x)L_m^{(\alpha )}(x)dx=\frac{\Gamma (n+\alpha +1)}{n!}\delta _{mn}. \end{aligned}$$

The system \(\left\{ H_n^{(\lambda )}(x)\right\} \) is orthonormal over \((-\infty ,+\infty )\) with respect to the weight \(|x|^{2\lambda }e^{-x^2}\), and \(H_n(x)=H_n^{(0)}(x)\) (\(n\ge 0\)) are the usual Hermite polynomials (up to constants).

The generalized Hermite functions \({\mathcal {H}}_n^{(\lambda )}(x)\) (\(n\ge 0\)) are given by

$$\begin{aligned} {\mathcal {H}}_n^{(\lambda )}(x)=e^{-\frac{x^2}{2}}|x|^{\lambda }H_n^{(\lambda )}(x), \end{aligned}$$

which are orthonormal over \((-\infty ,\infty )\) associated with Lebesgue measure. For a function \(f\in L({\mathbb {R}})\), its generalized Hermite expansion is

$$\begin{aligned} f\sim \sum _{n=0}^{\infty }a_n^{(\lambda )}(f){\mathcal {H}}_n^{(\lambda )}(x),\qquad \qquad a_n^{(\lambda )}(f)=\int _{-\infty }^{\infty }f(t){\mathcal {H}}_n^{(\lambda )}(t)dt. \end{aligned}$$
(9)

In what follows we assume that \(\lambda \ge 0\). Our main result is stated as follows.

Theorem 1.1

Let \(\lambda \ge 0\). Then there exists a constant \(A>0\) such that for all \(f\in H^1({\mathbb {R}})\),

$$\begin{aligned} \sum _{n=1}^{\infty }n^{-\frac{3}{4}}|a_n^{(\lambda )}(f)|\le A\Vert f\Vert _{H^1({\mathbb {R}})}. \end{aligned}$$
(10)

The generalized Hermite polynomials \(H_n^{(\lambda )}(x)\) (\(n\ge 0\)) were used in a Bose-like oscillator calculus in [11]; their further generalizations to the Dunkl setting in several variables can be found in [13] and those in the \(A_{d-1}\) case are eigenfunctions of the Hamiltonian of the linear quantum Calogero-Moser-Sutherland model (with some modification) in \({\mathbb {R}}^d\) (see [13] and references therein). It would be interesting to extend various results on the usual Hermite expansions such as in [10] and others to orthogonal expansions associated to these generalizations. We remark that in [4], the Hermite functions were used in a description of Feichtinger’s space \(S_0\).

Throughout the paper, \(U\lesssim V\) means that \(U\le cV\) for some positive constant \(c\) independent of variables, functions, \(n\), \(k\), etc., but possibly dependent of the parameter \(\lambda \).

2 Some Facts on the Poisson kernel

The Poisson kernel of the generalized Hermite polynomials \(H_n^{(\lambda )}(x)\) (\(n\ge 0\)) is, for \(0\le r<1\),

$$\begin{aligned} P^{(\lambda )}(r;x,y)=\sum _{n=0}^{\infty }r^nH_n^{(\lambda )}(x)H_n^{(\lambda )}(y). \end{aligned}$$

If we write \(P^{(\lambda )}(r;x,y)\) into two parts, one being the summation for even \(n\) and the other for odd \(n\), then from (7), (8), and by [15] or [16], (1.1.47)], we have

$$\begin{aligned} P^{(\lambda )}(r;x,y)=\frac{(1-r^2)^{-\lambda -1/2}}{\Gamma (\lambda +1/2)} \exp \left( -\frac{r^2}{1-r^2}(x^2+y^2)\right) E_{\lambda }\left( \frac{2rxy}{1-r^2}\right) , \end{aligned}$$

where \(E_{\lambda }(z)\) is the one-dimensional Dunkl kernel

$$\begin{aligned} E_\lambda (z)=j_{\lambda -1/2}(iz)+\frac{z}{2\lambda +1}j_{\lambda +1/2}(iz), \end{aligned}$$

and \(j_{\alpha }(z)\) is the normalized Bessel function

$$\begin{aligned} j_{\alpha }(z)=2^{\alpha }\Gamma (\alpha +1)\frac{J_{\alpha }(z)}{z^{\alpha }} =\Gamma (\alpha +1)\sum _{n=0}^{\infty }\frac{(-1)^n(z/2)^{2n}}{n!\Gamma (n+\alpha +1)}. \end{aligned}$$

If \(\lambda =0\), \(E_0(z)=e^z\), and for \(\lambda >0\), a Laplace-type representation of \(E_\lambda (z)\) is (cf. [12, Lemma 2.1])

$$\begin{aligned} E_\lambda (z)=c_{\lambda }'\int _{-1}^1e^{zt}(1+t)(1-t^2)^{\lambda -1}dt, \quad c'_{\lambda }=\frac{\Gamma (\lambda +1/2)}{\Gamma (\lambda )\Gamma (1/2)}. \end{aligned}$$
(11)

By (), the Poisson kernel \({\mathcal {P}}^{(\lambda )}(r;x,y)=\sum _{n=0}^{\infty }r^n{\mathcal {H}}_n^{(\lambda )}(x){\mathcal {H}}_n^{(\lambda )}(y)\) of the generalized Hermite functions \({\mathcal {H}}_n^{(\lambda )}\) (\(n\ge 0\)) can be written as

$$\begin{aligned} {\mathcal {P}}^{(\lambda )}(r;x,y)=|xy|^{\lambda }\frac{(1-r^2)^{-\lambda -1/2}}{\Gamma (\lambda +1/2)} \exp \left( -\frac{1+r^2}{1-r^2}\cdot \frac{x^2+y^2}{2}\right) E_{\lambda }\left( \frac{2rxy}{1-r^2}\right) . \end{aligned}$$

The Poisson integral of the generalized Hermite expansion of a function \(f\in L^1({\mathbb {R}})\) is defined by

$$\begin{aligned} f_{\lambda }(r;x)=\int _{-\infty }^{\infty }{\mathcal {P}}^{(\lambda )}(r;x,y)f(y)dy. \end{aligned}$$
(12)

For \(\lambda =0\), \({\mathcal {P}}^{(0)}(r;x,y)\) is consistent with the Poisson kernel of the usual Hermite functions \({\mathcal {H}}_n\) (\(n\ge 0\)) (see [16])

$$\begin{aligned} {\mathcal {P}}(r;x,y)=\frac{(1-r^2)^{-\frac{1}{2}}}{\sqrt{\pi }} \exp \left( -\frac{2r(x-y)^2+(1-r)^2(x^2+y^2)}{2(1-r^2)}\right) . \end{aligned}$$

In order to give some necessary estimates of the Poisson kernel \({\mathcal {P}}^{(\lambda )}(r;x,y)\) we rewrite it into

$$\begin{aligned} {\mathcal {P}}^{(\lambda )}(r;x,y)=\frac{(1-r^2)^{-\lambda -1/2}}{\Gamma (\lambda +1/2)}|xy|^{\lambda }\psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y),\quad \xi =\frac{2rx}{1-r^2}, \end{aligned}$$
(13)

where

$$\begin{aligned} \psi _{r,x}(y)=\exp \left( -\frac{1+r^2}{1-r^2}\frac{x^2+y^2}{2}+\frac{2r|xy|}{1-r^2}\right) . \end{aligned}$$

We shall need several preliminary lemmas.

Lemma 2.1

For \(y\ne 0\),

  1. (i)

    \(\left| e^{-|\xi y|}E_{\lambda }(\xi y)\right| \lesssim (1+|\xi y|)^{-\lambda }\);

  2. (ii)

    \(\left| \frac{\partial }{\partial y}\left( e^{-|\xi y|}E_{\lambda }(\xi y)\right) \right| \lesssim |\xi |(1+|\xi y|)^{-\lambda -1}\).

Proof

If \(\lambda =0\), both (i) and (ii) are trivial. In what follows we assume that \(\lambda >0\). From (11) we have

$$\begin{aligned} \left| e^{-|\xi y|}E_{\lambda }(\xi y)\right|&\le 2c_{\lambda }'\int _{-1}^1e^{-|\xi y|(1-|t|)}(1-t^2)^{\lambda -1}dt \lesssim \int _{0}^1e^{-|\xi y|(1-t)}(1-t)^{\lambda -1}dt\\&=|\xi y|^{-\lambda }\int _0^{|\xi y|}e^{-s}s^{\lambda -1}ds, \end{aligned}$$

which is bounded by a multiple of \((1+|\xi y|)^{-\lambda }\), since \(\int _0^{A}e^{-s}s^{\lambda -1}ds\asymp \left( A/(A+1)\right) ^{\lambda }\).

Furthermore, from (11) we have

$$\begin{aligned} \frac{\partial }{\partial y}\left( e^{-|\xi y|}E_{\lambda }(\xi y)\right)&=-c_{\lambda }'\xi \int _{-1}^1e^{-\xi y(1-t)}(1+t)^{\lambda }(1-t)^{\lambda }dt\quad \hbox {for}\quad \xi y>0;\\&=c_{\lambda }'\xi \int _{-1}^1e^{\xi y(1-t)}(1+t)^{\lambda -1}(1-t)^{\lambda +1}dt\quad \,\, \hbox {for}\quad \xi y<0. \end{aligned}$$

In the both cases we split the integral into two parts as \(\int _{-1}^1=\int _{-1}^0+\int _0^1\), to get

$$\begin{aligned} \left| \frac{\partial }{\partial y}\left( e^{-|\xi y|}E_{\lambda }(\xi y)\right) \right|&\lesssim |\xi |\left[ e^{-|\xi y|}+\int _{0}^1e^{-|\xi y|(1-t)}(1-t)^{\lambda }dt\right] \\&=|\xi |\left[ e^{-|\xi y|}+|\xi y|^{-\lambda -1}\int _0^{|\xi y|}e^{-s}s^{\lambda }ds\right] \end{aligned}$$

which implies the desired estimate in part (ii), since \(\int _0^{A}e^{-s}s^{\lambda }ds\asymp \left( A/(A+1)\right) ^{\lambda +1}\). \(\square \)

Lemma 2.2

  1. (i)

    For \(0\le r<1\),

    $$\begin{aligned} \psi _{r,x}(y)=\exp \left( -\frac{r(|x|-|y|)^2}{1-r^2}\right) \exp \left( -\frac{1-r}{1+r}\frac{x^2+y^2}{2}\right) ; \end{aligned}$$
    (14)
  2. (ii)

    for \(0\le r\le \frac{1}{2}\), \(\psi _{r,x}(y)\lesssim \exp \left( -c_1(x^2+y^2)\right) \), and for \(\frac{1}{2}\le r<1\), \(\psi _{r,x}(y)\lesssim \exp \left( -c_2\frac{(|x|-|y|)^2}{1-r}\right) \), where \(c_1,c_2>0\) are independent of \(x,y\) and \(r\).

Part (i) is obvious and implies part (ii) immediately.

Lemma 2.3

For \(x,y\in (-\infty ,\infty )\) and \(r\in [0,1)\),

$$\begin{aligned} \frac{(1-r)^{1/2}|x|}{1-r+|xy|}\lesssim 1+\frac{\left| |x|-|y|\right| }{(1-r)^{1/2}}. \end{aligned}$$

Indeed, for \(|y|\le |x|/2\) the left hand side is bounded by \(|x|/(1-r)^{1/2}\le 2\left| |x|-|y|\right| /(1-r)^{1/2}\), and for \(|y|\ge |x|/2\), by \(|x|/\left( 2|xy|^{1/2}\right) \le 1\).

Proposition 2.4

  1. (i)

    If \(0\le r\le \frac{1}{2}\), then for \(y\ne 0\),

    $$\begin{aligned} \left| \frac{\partial }{\partial y}{\mathcal {P}}^{(\lambda )}(r;x,y)\right| \lesssim |x|^{\lambda }|y|^{\lambda -1}\left( \lambda +|xy|+|y|^2\right) \exp \left( -c_1(x^2+y^2)\right) ;\nonumber \\ \end{aligned}$$
    (15)
  2. (ii)

    if \(\frac{1}{2}\le r<1\), then for \(y\ne 0\),

    $$\begin{aligned} \left| \frac{\partial }{\partial y}{\mathcal {P}}^{(\lambda )}(r;x,y)\right| \lesssim \lambda \frac{(1-r)^{-1/2}|x|^{\lambda }|y|^{\lambda -1}}{(1-r+|xy|)^{\lambda }}\psi _{r,x}(y) +(1-r)^{-1}\psi _{r,x}(y)^{\frac{1}{2}}.\nonumber \\ \end{aligned}$$
    (16)

Proof

For \(y\ne 0\), from (13) we have

$$\begin{aligned} \frac{\partial }{\partial y}{\mathcal {P}}^{(\lambda )}(r;x,y)=&\frac{(1-r^2)^{-\lambda -1/2}}{\Gamma (\lambda +1/2)}\left( U_1(x,y)+U_2(x,y)\right) , \end{aligned}$$

where

$$\begin{aligned}&U_1(x,y)=\lambda |x|^{\lambda }|y|^{\lambda -2}y\left( \psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right) ,\\ {}&U_2(x,y)=|xy|^{\lambda }\left( \psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right) '_y. \end{aligned}$$

For \(0\le r\le \frac{1}{2}\), it follows from Lemmas 2.1(i) and 2.2(ii) that \((1-r^2)^{-\lambda -1/2}U_1(x,y)\) is bounded by a multiple of \(\lambda |x|^{\lambda }|y|^{\lambda -1}\exp \left( -c_1(x^2+y^2)\right) \), and for \(\frac{1}{2}\le r<1\), by a multiple of

$$\begin{aligned} \lambda \frac{(1-r)^{-\lambda -1/2}|x|^{\lambda }|y|^{\lambda -1}}{(1+|\xi y|)^{\lambda }}\psi _{r,x}(y) \lesssim \lambda \frac{(1-r)^{-1/2}|x|^{\lambda }|y|^{\lambda -1}}{(1-r+|xy|)^{\lambda }}\psi _{r,x}(y). \end{aligned}$$

In order to verify that \((1-r^2)^{-\lambda -1/2}U_2(x,y)\) has the desired estimates as in (15) and (16), we shall show that for \(0\le r\le \frac{1}{2}\),

$$\begin{aligned} \left| \left( \psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right) '_y\right| \lesssim (|x|+|y|)\exp \left( -c_1(x^2+y^2)\right) ; \end{aligned}$$
(17)

and for \(\frac{1}{2}\le r<1\),

$$\begin{aligned}&\left| \left( \psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right) '_y\right| \nonumber \\ {}&\quad \lesssim \frac{\psi _{r,x}(y)}{(1-r)^{1-\lambda }} \left[ \frac{\left| |x|-|y|\right| +(1-r)^2|y|}{(1-r+|xy|)^{\lambda }}+\frac{(1-r)|x|}{(1-r+|xy|)^{\lambda +1}}\right] . \end{aligned}$$
(18)

Indeed, since from (14),

$$\begin{aligned}&\left( \psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right) '_y\\&\quad =\left[ \frac{2r(|x|-|y|)(\hbox {sgn}\,y){-}(1{-}r)^2y}{1{-}r^2}\left( e^{-|\xi y|}E_{\lambda }(\xi y)\right) {+}\left( e^{-|\xi y|}E_{\lambda }(\xi y)\right) '_y\right] \psi _{r,x}(y), \end{aligned}$$

by Lemma 2.1 we have

$$\begin{aligned} \left| \left( \psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right) '_y\right| \lesssim \left[ \frac{\left| |x|-|y|\right| +(1-r)^2|y|}{(1-r)(1+|\xi y|)^{\lambda }} +\frac{|\xi |}{(1+|\xi y|)^{\lambda +1}}\right] \psi _{r,x}(y). \end{aligned}$$

Thus (17) and (18) follow immediately.

For \(0\le r\le \frac{1}{2}\), from (17) it is obvious that

$$\begin{aligned} \left| (1-r^2)^{-\lambda -1/2}U_2(x,y)\right| \lesssim |xy|^{\lambda }(|x|+|y|)\exp \left( -c_1(x^2+y^2)\right) ; \end{aligned}$$

and for \(\frac{1}{2}\le r<1\), from (18) we have

$$\begin{aligned} \left| (1-r^2)^{-\lambda -1/2}U_2(x,y)\right| \lesssim \frac{\psi _{r,x}(y)}{1-r} \left[ \frac{\left| |x|-|y|\right| }{(1{-}r)^{1/2}}{+}(1-r)^{3/2}|y|+\frac{(1-r)^{1/2}|x|}{1-r+|xy|}\right] . \end{aligned}$$

By Lemma 2.3, it is easy to see that the expression in the brackets above is bounded by a multiple of \(\psi _{r,x}(y)^{-1/2}\), and so \(\left| (1-r^2)^{-\lambda -1/2}U_2(x,y)\right| \lesssim \psi _{r,x}(y)^{1/2}/(1-r)\). The proof of the proposition is finished. \(\square \)

3 Proof of the Main Result

The following theorem is crucial in the proof of Theorem 1.1.

Theorem 3.1

Assume that \(\lambda \ge 0\). There is a positive constant \(A\) such that for all \(f\in H^1({\mathbb {R}})\),

$$\begin{aligned} \int _0^1(1-r)^{-\frac{3}{4}}\left( \int _{-\infty }^{\infty }|f_{\lambda }(r;y)|^2dy\right) ^{\frac{1}{2}}dr\le A\Vert f\Vert _{H^1({\mathbb {R}})}, \end{aligned}$$
(19)

where \(f_{\lambda }(r;y)\) is the Poisson integral (12) associated to the generalized Hermite expansion of \(f\).

The proof of Theorem 3.1 is based upon some norm estimates stated below.

Lemma 3.2

  1. (i)

    If \(\lambda =0\) or \(\lambda >1\), then for \(y,\bar{y}\in (-\infty ,\infty )\),

    $$\begin{aligned} \left( \int _{-\infty }^{\infty } \left| {\mathcal {P}}^{(\lambda )}(r;x,y)-{\mathcal {P}}^{(\lambda )}(r;x,\bar{y})\right| ^2dx\right) ^{\frac{1}{2}} \lesssim \frac{|y-\bar{y}|}{(1-r)^{3/4}}; \end{aligned}$$
    (20)
  2. (ii)

    If \(0<\lambda \le 1\), then for \(y,\bar{y}\in (-\infty ,\infty )\),

    $$\begin{aligned} \left( \int _{-\infty }^{\infty } \left| {\mathcal {P}}^{(\lambda )}(r;x,y)-{\mathcal {P}}^{(\lambda )}(r;x,\bar{y})\right| ^2dx\right) ^{\frac{1}{2}} \lesssim \frac{|y{-}\bar{y}|^{\lambda }}{(1{-}r)^{(2\lambda {+}1)/4}} \left( 1{+}\frac{|y{-}\bar{y}|}{(1{-}r)^{1/2}}\right) . \end{aligned}$$

Proof

For \(\lambda >1\) and \(y>\bar{y}\) we have

$$\begin{aligned} \left\| {\mathcal {P}}^{(\lambda )}(r;\cdot ,y)-{\mathcal {P}}^{(\lambda )}(r;\cdot ,\bar{y})\right\| _{L^2({\mathbb {R}})}&=\left\| \int _{\bar{y}}^y\frac{\partial }{\partial z}{\mathcal {P}}^{(\lambda )}(r;\cdot ,z)dz\right\| _{L^2({\mathbb {R}})}\\&\le \int _{\bar{y}}^y\left\| \frac{\partial }{\partial z}{\mathcal {P}}^{(\lambda )}(r;\cdot ,z)\right\| _{L^2({\mathbb {R}})}dz. \end{aligned}$$

If \(0\le r\le \frac{1}{2}\), then by Proposition 2.4(i), \(\left\| \frac{\partial }{\partial z}{\mathcal {P}}^{(\lambda )}(r;\cdot ,z)\right\| _{L^2({\mathbb {R}})}\lesssim 1\), so that

$$\begin{aligned} \left\| {\mathcal {P}}^{(\lambda )}(r;\cdot ,y)-{\mathcal {P}}^{(\lambda )}(r;\cdot ,\bar{y})\right\| _{L^2({\mathbb {R}})} \lesssim |y-\bar{y}|\lesssim \frac{|y-\bar{y}|}{(1-r)^{3/4}}. \end{aligned}$$

Since for \(\lambda >1\) and \(\frac{1}{2}\le r<1\), by Lemma 2.3 we have

$$\begin{aligned} \frac{(1-r)^{-1/2}|x|^{\lambda }|y|^{\lambda -1}}{(1-r+|xy|)^{\lambda }} \le \frac{(1-r)^{-1/2}|x|}{1-r+|xy|}\lesssim \frac{1}{1-r}\left( 1+\frac{\left| |x|-|y|\right| }{(1-r)^{1/2}}\right) \lesssim \frac{\psi _{r,x}(y)^{-\frac{1}{2}}}{1-r}, \end{aligned}$$

inserting this into (16) yields that \(\left| \frac{\partial }{\partial y}{\mathcal {P}}^{(\lambda )}(r;x,y)\right| \lesssim (1-r)^{-1}\psi _{r,x}(y)^{\frac{1}{2}}\) and hence

$$\begin{aligned} \left\| {\mathcal {P}}^{(\lambda )}(r;\cdot ,y)-{\mathcal {P}}^{(\lambda )}(r;\cdot ,\bar{y})\right\| _{L^2({\mathbb {R}})} \le \frac{1}{1-r}\int _{\bar{y}}^y\left\| \psi _{r,x}(z)^{\frac{1}{2}}\right\| _{L^2({\mathbb {R}})}dz \lesssim \frac{|y-\bar{y}|}{(1-r)^{3/4}}. \end{aligned}$$

This proves part (i) for \(\lambda >1\).

If \(\lambda =0\), by Proposition 2.4 \(\left| \frac{\partial }{\partial y}{\mathcal {P}}^{(0)}(r;x,y)\right| \) is bounded by a multiple of \((|x|+|y|)e^{-c_1(x^2+y^2)}\) for \(0\le r\le \frac{1}{2}\), and \((1-r)^{-1}\psi _{r,x}(y)^{\frac{1}{2}}\) for \(\frac{1}{2}\le r<1\). Hence the above process still works for \(\lambda =0\). This completes the proof of part (i).

Now we turn to the proof of part (ii). For \(0<\lambda \le 1\) and \(y>\bar{y}\), we write

$$\begin{aligned}&{\mathcal {P}}^{(\lambda )}(r;x,y)-{\mathcal {P}}^{(\lambda )}(r;x,\bar{y})\\&\quad = \frac{(1-r^2)^{-\lambda -1/2}}{\Gamma (\lambda +1/2)}\left[ \int _{\bar{y}}^y U_2(x,z)dz +|x|^{\lambda }(|y|^{\lambda }-|\bar{y}|^{\lambda })\psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right. \\&\quad \quad +\left. \int _{\bar{y}}^y|x|^{\lambda }(|\bar{y}|^{\lambda }-|z|^{\lambda }) \left( \psi _{r,x}(z)e^{-|\xi z|}E_{\lambda }(\xi z)\right) '_zdz\right] , \end{aligned}$$

where \(U_2(x,z)\) is defined as the same as in the proof of Proposition 2.4. There we have shown that (for all \(\lambda \ge 0\)) \((1-r^2)^{-\lambda -1/2}U_2(x,z)\) is bounded by a multiple of \(|xz|^{\lambda }(|x|+|z|)\exp \left( -c_1(x^2+z^2)\right) \) for \(0\le r\le 1/2\), and of \(\psi _{r,x}(z)^{1/2}/(1-r)\) for \(1/2\le r<1\). From the proof of part (i) above it follows that

$$\begin{aligned} \left\| {\mathcal {P}}^{(\lambda )}(r;\cdot ,y)-{\mathcal {P}}^{(\lambda )}(r;\cdot ,\bar{y})\right\| _{L^2({\mathbb {R}})} \lesssim&\frac{|y-\bar{y}|}{(1-r)^{3/4}}+V_1+V_2, \end{aligned}$$
(21)

where

$$\begin{aligned} V_1&=\frac{|y-\bar{y}|^{\lambda }}{(1-r)^{\lambda +1/2}} \left( \int _{-\infty }^{\infty }\left| |x|^{\lambda }\psi _{r,x}(y)e^{-|\xi y|}E_{\lambda }(\xi y)\right| ^2dx\right) ^{1/2},\\ V_2&=\frac{|y-\bar{y}|^{\lambda }}{(1-r)^{\lambda +1/2}}\int _{\bar{y}}^y\left( \int _{-\infty }^{\infty }\left| |x|^{\lambda } \left( \psi _{r,x}(z)e^{-|\xi z|}E_{\lambda }(\xi z)\right) '_z\right| ^2dx\right) ^{1/2}dz. \end{aligned}$$

If \(0\le r\le \frac{1}{2}\), then by Lemmas 2.1(i) and 2.2(ii),

$$\begin{aligned} V_1&\lesssim |y-\bar{y}|^{\lambda } \left( \int _{-\infty }^{\infty }|x|^{2\lambda }\exp \left( -2c_1(x^2+y^2)\right) dx\right) ^{1/2}\nonumber \\&\lesssim |y-\bar{y}|^{\lambda }\lesssim \frac{|y-\bar{y}|^{\lambda }}{(1-r)^{(2\lambda +1)/4}}; \end{aligned}$$
(22)

and from (17),

$$\begin{aligned} V_2&\lesssim |y-\bar{y}|^{\lambda } \int _{\bar{y}}^y\left( \int _{-\infty }^{\infty }|x|^{2\lambda }(|x|+|z|)^2\exp \left( -2c_1(x^2+z^2)\right) dx\right) ^{1/2}dz\nonumber \\&\lesssim |y-\bar{y}|^{\lambda +1}\lesssim \frac{|y-\bar{y}|^{\lambda +1}}{(1-r)^{(2\lambda +3)/4}}. \end{aligned}$$
(23)

If \(\frac{1}{2}\le r<1\), then by Lemmas 2.1(i), 2.2(ii) and 2.3,

$$\begin{aligned} V_1&\lesssim \frac{|y-\bar{y}|^{\lambda }}{(1-r)^{1/2}} \left( \int _{-\infty }^{\infty }\left| \psi _{r,x}(y)\frac{|x|^{\lambda }}{(1-r+|xy|)^{\lambda }}\right| ^2dx\right) ^{1/2} \nonumber \\&\lesssim \frac{|y-\bar{y}|^{\lambda }}{(1-r)^{(\lambda +1)/2}} \left( \int _{-\infty }^{\infty }\left( 1+\frac{\left| |x|-|y|\right| }{(1-r)^{1/2}}\right) ^{2\lambda } \exp \left( -2c_2\frac{(|x|-|y|)^2}{1-r}\right) dx\right) ^{1/2}\nonumber \\&\lesssim \frac{|y-\bar{y}|^{\lambda }}{(1-r)^{(2\lambda +1)/4}}; \end{aligned}$$
(24)

and by (18), Lemmas 2.2(ii) and 2.3,

$$\begin{aligned} V_2&\lesssim \frac{|y-\bar{y}|^{\lambda }}{(1-r)^{(\lambda +2)/2}} \int _{\bar{y}}^y\left( \int _{-\infty }^{\infty }\left[ \left( \frac{(1-r)^{1/2}|x|}{1-r+|xz|}\right) ^{\lambda } \left( \frac{\left| |x|-|z|\right| }{(1-r)^{1/2}}+(1-r)^{\frac{3}{2}}|z|\right) \right. \right. \nonumber \\&\left. \left. \qquad +\left( \frac{(1-r)^{1/2}|x|}{1-r+|xz|}\right) ^{\lambda +1}\right] ^2\psi _{r,x}(z)^2dx\right) ^{1/2}dz\nonumber \\&\lesssim \frac{|y-\bar{y}|^{\lambda }}{(1-r)^{(\lambda +2)/2}} \int _{\bar{y}}^y\left( \int _{-\infty }^{\infty }\left[ \left( 1+\frac{\left| |x|-|z|\right| }{(1-r)^{1/2}}\right) ^{\lambda } \left( \frac{\left| |x|-|z|\right| }{(1-r)^{1/2}}+(1-r)^{\frac{3}{2}}|z|\right) \right. \right. \nonumber \\&\left. \left. \qquad +\left( 1+\frac{\left| |x|-|z|\right| }{(1-r)^{1/2}}\right) ^{\lambda +1}\right] ^2\psi _{r,x}(z)^2dx\right) ^{1/2}dz\nonumber \\&\lesssim \frac{|y-\bar{y}|^{\lambda }}{(1-r)^{(\lambda +2)/2}} \int _{\bar{y}}^y\left( \int _{-\infty }^{\infty }\psi _{r,x}(z)dx\right) ^{1/2}dz\nonumber \\&\lesssim \frac{|y-\bar{y}|^{\lambda +1}}{(1-r)^{(2\lambda +3)/4}}. \end{aligned}$$
(25)

Collecting the estimates of (22), (23), (24) and (25) into (21), we finish the proof of part (ii) of the lemma. \(\square \)

Now we come to the proof of Theorem 3.1.

Let us first recall the atom characterization of the Hardy space \(H^1({\mathbb {R}})\) (cf. [14, 18]). An \(H^1\) atom is a measurable function \(a\) on \({\mathbb {R}}\) satisfying

  1. (i)

    \(\hbox {supp}\, a\subseteq I\) for some interval \(I\subset {\mathbb {R}}\);

  2. (ii)

    \(\Vert a\Vert _{L^{\infty }}\le |I|^{-1}\), \(|I|\) denoting the length of \(I\);

  3. (iii)

    \(\int _{-\infty }^{\infty }a(x)dx=0\).

For a function \(f\in H^1({\mathbb {R}})\), there are a sequence \(\{a_k\}\) of \(H^1\) atoms and a sequence \(\{\lambda _k\}\) of complex numbers satisfying \(\sum |\lambda _k|<\infty \), such that \(f=\sum _{k=1}^{\infty }\lambda _k a_k\) and

$$\begin{aligned} A_1\Vert f\Vert _{H^1({\mathbb {R}})}\le \sum |\lambda _k|\le A_2\Vert f\Vert _{H^1({\mathbb {R}})}, \end{aligned}$$

where \(A_1,A_2\) are positive constants independent of \(f\).

In order to prove Theorem 3.1, it suffices to verify the inequality (19) for each \(H^1\) atom \(f=a\), namely, there exists a absolute constant \(A>0\), such that

$$\begin{aligned} \int _0^1(1-r)^{-\frac{3}{4}}\Vert a_{\lambda }(r;\cdot )\Vert _{L^2({\mathbb {R}})}dr\le A \end{aligned}$$
(26)

holds for all \(H^1\) atoms \(a\) satisfying (i), (ii) and (iii) above, where \(a_{\lambda }(r;y)\) is the Poisson integral (12) of \(a\) associated to its generalized Hermite expansion.

By Parseval’s inequality, from (i) and (ii) we have

$$\begin{aligned} \Vert a_{\lambda }(r;\cdot )\Vert _{L^2({\mathbb {R}})}\le \Vert a\Vert _{L^2({\mathbb {R}})}\le |I|^{-\frac{1}{2}}. \end{aligned}$$
(27)

(27) is useful for atoms with larger supports, but we also need a more accurate estimate for atoms with smaller supports.

For an atom \(a\) satisfying (i), (ii) and (iii), let \(\bar{y}\) be the left endpoint of the interval \(I\). Applying the cancelation property (iii) and Minkowski’s inequality, we have

$$\begin{aligned} \Vert a_{\lambda }(r;\cdot )\Vert _{L^2({\mathbb {R}})}&=\left( \int _{-\infty }^{\infty } \left[ \int _I\left( {\mathcal {P}}^{(\lambda )}(r;x,y)-{\mathcal {P}}^{(\lambda )}(r;x,\bar{y})\right) a(y)dy\right] ^{2}dx \right) ^{\frac{1}{2}}\nonumber \\&\le \int _I|a(y)|\left( \int _{-\infty }^{\infty } \left| {\mathcal {P}}^{(\lambda )}(r;x,y)-{\mathcal {P}}^{(\lambda )}(r;x,\bar{y})\right| ^2dx\right) ^{1/2}dy. \end{aligned}$$

By means of Lemma 3.2 we obtain

$$\begin{aligned} \Vert a_{\lambda }(r;\cdot )\Vert _{L^2({\mathbb {R}})}\lesssim \frac{|I|}{(1-r)^{3/4}} +\frac{|I|^{\lambda }}{(1-r)^{(2\lambda +1)/4}}+\frac{|I|^{\lambda +1}}{(1-r)^{(2\lambda +3)/4}}. \end{aligned}$$
(28)

If \(\lambda =0\) or \(\lambda >1\), the first term on the right hand side appears only.

If \(|I|\ge 1\), from (27) it is trivial that

$$\begin{aligned} \int _0^1(1-r)^{-\frac{3}{4}}\Vert a_{\lambda }(r;\cdot )\Vert _{L^2({\mathbb {R}})}dr\le 4|I|^{-\frac{1}{2}}\le 4; \end{aligned}$$

and if \(|I|<1\), we split the integral into two parts, over \([1-|I|^2,1)\) and \([0,1-|I|^2]\) respectively, and apply (27) and (28) to obtain

$$\begin{aligned}&\int _0^1(1-r)^{-\frac{3}{4}}\Vert a_{\lambda }(r;\cdot )\Vert _{L^2({\mathbb {R}})}dr\lesssim \int _{1-|I|^2}^1|I|^{-1/2}(1-r)^{-3/4}dr\\&\quad +\int _0^{1-|I|^2}\left( \frac{|I|}{(1-r)^{3/2}} +\frac{|I|^{\lambda }}{(1-r)^{(\lambda +2)/2}}+\frac{|I|^{\lambda +1}}{(1-r)^{(\lambda +3)/2}}\right) dr. \end{aligned}$$

It is easy to see that the values of the two integrals above are independent of \(|I|<1\). This proves (26), and hence Theorem 3.1.

Proof of Theorem 1.1

By Hölder’s inequality one has

$$\begin{aligned} \sum _{n=0}^{\infty }r^{2n}|a_n^{(\lambda )}(f)|&\le \left( \sum _{n=0}^{\infty }r^{2n}\right) ^{1/2}\left( \sum _{n=0}^{\infty }r^{2n}|a_n^{(\lambda )}(f)|^{2}\right) ^{1/2}\\&\le (1-r)^{-1/2}\Vert f_{\lambda }(r;\cdot )\Vert _{L^2}. \end{aligned}$$

Multiplying \((1-r)^{-1/4}\) on both sides and taking integration over \([0,1)\) respectively, then by Theorem 3.1 we get

$$\begin{aligned} \sum _{n=0}^{\infty }B(2n+1,3/4)|a_n^{(\lambda )}(f)|\le A\Vert f\Vert _{H^1}, \end{aligned}$$

where \(B(\alpha ,\beta )\) denotes the beta function. Finally, the Hardy inequality (10) follows from the fact that \(B(2n+1,3/4)\sim (n+1)^{-3/4}\). The proof of Theorem 1.1 is completed.\(\square \)