1 Introduction

The Bargmann-Fock space \({{\mathcal {F}}}^2({\mathbb {C}})\) is the Hilbert space consisting of those analytic functions \(f\in H({\mathbb {C}})\) such that

$$\begin{aligned} \Vert f\Vert _{{{\mathcal {F}}}}^2 = \int _{{\mathbb {C}}}|f(z)|^2 e^{-\pi |z|^2}\ dA(z) < +\infty , \end{aligned}$$

where dA(z) denotes the Lebesgue measure. \({{\mathcal {F}}}^2({\mathbb {C}})\) admits a reproducing kernel \(K_w(z) = e^{\pi {\overline{w}}z},\) which means that

$$\begin{aligned}f(w) = \langle f, K_w\rangle ,\ \ f\in {{\mathcal {F}}}^2({\mathbb {C}}). \end{aligned}$$

The normalized monomials

$$\begin{aligned} e_n(z) = \left( \frac{\pi ^n}{n!}\right) ^{\frac{1}{2}} z^n,\ \ n\ge 0, \end{aligned}$$

form an orthonormal basis. For a fixed \(a\in {\mathbb {C}}\) the translation operator

$$\begin{aligned} W_a:{{\mathcal {F}}}^2({\mathbb {C}})\rightarrow {{\mathcal {F}}}^2({\mathbb {C}}),\ \left( W_a f\right) (z) = f(z-a)e^{-\frac{\pi }{2} |a|^2 + \pi z{\overline{a}}}, \end{aligned}$$

is an isometry (see [14, Proposition 2.38]). We denote \(d\lambda (z) = e^{-\pi |z|^2} dA(z),\) so \({{\mathcal {F}}}^2({\mathbb {C}})\) is a closed subspace of \(L^2({\mathbb {C}}, d\lambda ).\) The orthogonal projection

$$\begin{aligned} P:L^2({\mathbb {C}}, d\lambda )\rightarrow {{\mathcal {F}}}^2({\mathbb {C}}) \end{aligned}$$

is the integral operator

$$\begin{aligned} \left( Pf\right) (z) = \int _{{\mathbb {C}}} f(w)K_w(z)\ d\lambda (w). \end{aligned}$$

For a measurable and bounded function F on \({\mathbb {C}}\) the Toeplitz operator with symbol F is defined as

$$\begin{aligned} T_F(f)(z) = P(Ff)(z) = \int _{{\mathbb {C}}} F(w)f(w)K_w(z)\ d\lambda (w). \end{aligned}$$

The systematic study of Toeplitz operators on the Fock space started in [3, 4]. Since then it has been a very active research area. We refer to [14, Chapter 6], where boundedness and membership in the Schatten classes is discussed.

It is obvious that

$$\begin{aligned}T_F:{{\mathcal {F}}}^2({\mathbb {C}})\rightarrow {{\mathcal {F}}}^2({\mathbb {C}}) \end{aligned}$$

is a bounded operator and

$$\begin{aligned} \Vert T_F(f)\Vert \le \Vert Ff\Vert _{L^2({\mathbb {C}}, d\lambda )} \le \Vert F\Vert _\infty \cdot \Vert f\Vert . \end{aligned}$$

In particular, \(\Vert T_F\Vert \le 1\) whenever \(\Vert F\Vert _\infty \le 1.\) If moreover \(T_F\) is compact, which happens for instance when \(F\in L^1({\mathbb {C}}),\) then \(\Vert T_F\Vert \) is strictly less than 1 but, as far as we know, no precise estimate for the norm is known. The main result of the paper gives a bound for \(\Vert T_F\Vert \) in the case that the symbol F is radial, real-valued, and satisfies some integrability condition. For Toeplitz operators with radial symbols we refer to [11]. Besides Toeplitz operators on the Fock space we consider time-frequency localization operators with Gaussian window, also known as anti-Wick operators. They were introduced by Daubechies [7] as filters in signal analysis and can be obtained from Toeplitz operators on the Fock space after applying Bargmann transform.

2 Toeplitz Operators on the Fock Space

The Toeplitz operator defined by a real valued symbol F is self-adjoint. This is immediate from the identity

$$\begin{aligned} \langle T_F(f), g\rangle = \int _{{\mathbb {C}}} F(z) f(z)\overline{g(z)}d\lambda (z) \end{aligned}$$

for all \(f,g\in {{\mathcal {F}}}^2({\mathbb {C}}).\) In this case we have

$$\begin{aligned} \Vert T_F\Vert = \sup _{\Vert f\Vert =1}\left| \langle T_F(f), f\rangle \right| \le \sup _{\Vert f\Vert =1}\int _{{\mathbb {C}}} |F(z)|\cdot |f(z)|^2d\lambda (z). \end{aligned}$$

A symbol F is said to be radial with respect to \(a\in {\mathbb {C}}\) if \(F(z) = g(|z-a|)\) for some bounded and measurable function g on \([0, +\infty ).\) The main result of the paper is as follows.

Theorem 1

Let \(F\in L^1({\mathbb {C}})\cap L^\infty ({\mathbb {C}})\) be a real-valued and radial symbol with respect to \(a\in {\mathbb {C}}.\) Then

$$\begin{aligned} \Vert T_F\Vert \le \Vert F\Vert _\infty \Bigg (1 - \exp \Bigg (-\frac{\Vert F\Vert _1}{\Vert F\Vert _\infty }\Bigg )\Bigg ).\end{aligned}$$

An expression for the norm of Toeplitz operators with radial symbols can be found in [11] but it is unclear how the estimate provided by Theorem 1 can be obtained from it.

For the proof we will need some auxiliary results. First we observe that for \(|F(z)| = g(|z|)\) and \(f = \sum _{n=0}^\infty b_n e_n\) we have, after changing to polar coordinates,

$$\begin{aligned} \int _{{\mathbb {C}}} |F(z)|\cdot |f(z)|^2d\lambda (z)= & {} \sum _{n=0}^\infty |b_n|^2\int _{{\mathbb {C}}}g(|z|)|e_n(z)|^2\ d\lambda (z) \\= & {} \sum _{n=0}^\infty |b_n|^2 2\pi \int _0^\infty g(r)\pi ^n\frac{r^{2n+1}}{n!}e^{-\pi r^2}\ dr \\= & {} \sum _{n=0}^\infty |b_n|^2 \int _0^\infty g\Big (\sqrt{\frac{t}{\pi }}\Big )\frac{t^n}{n!} e^{-t}\ dt. \end{aligned}$$

The d-dimensional Lebesgue measure of a set \(\varOmega \subset {\mathbb {R}}^d\) is denoted \(|\varOmega |\) both for \(d=1\) and \(d=2.\)

Lemma 1

Let \(I\subset [0, +\infty )\) be a measurable set with finite Lebesgue measure. Then

$$\begin{aligned} \frac{1}{n!}\int _I s^ne^{-s}\ ds \le 1-e^{-|I|}. \end{aligned}$$

Proof

  1. (a)

    We first assume that I is a finite union of bounded intervals. The function \(h(s) = \frac{s^n}{n!}e^{-s}\) attains its absolute maximum at \(s = n.\) Then h increases on [0, n] and decreases on \([n, +\infty ).\) We consider \(a\le n\le b\) such that

$$\begin{aligned} n - a = |I\cap [0, n]|\ ,\ b-n = |I\cap [n, +\infty )|.\end{aligned}$$

Then

$$\begin{aligned} \frac{1}{n!}\int _I s^n e^{-s}\ ds\le & {} \int _a^b h(s)\ ds = \frac{e^{-a}}{n!}\int _0^{b-a}(t+a)^n e^{-t}\ dt \\= & {} \sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \frac{a^{n-k}}{n!}e^{-a}\int _0^{|I|} t^k e^{-t}\ dt \\= & {} \sum _{k=0}^n \frac{a^{n -k}}{(n -k)!}e^{-a}\frac{1}{k!}\int _0^{|I|} t^k e^{-t}\ dt\\\le & {} \sup _{0\le k\le n}\frac{1}{k!}\int _0^{|I|} t^k e^{-t}\ dt = \int _0^{|I|}e^{-t}\ dt. \end{aligned}$$

For the last identity observe that

$$\begin{aligned} \frac{1}{k!}\int _0^{s} t^k e^{-t}\ dt = 1 - e^{-s}\sum _{j=0}^k \frac{s^j}{j!}. \end{aligned}$$
  1. (b)

    For a general measurable set I with finite measure the conclusion follows from part (a) and the fact that for every \(\varepsilon > 0\) there is a set J,  finite union of bounded intervals, with the property that

$$\begin{aligned} |J\setminus I| + |I\setminus J| \le \varepsilon . \end{aligned}$$

\(\square \)

Lemma 2

Let \((I_k)_{k=1}^N\) be disjoint sets with finite measure and \(0\le \varepsilon _k\le 1\) for every \(1\le k\le N.\) Then, for every \(p\in {\mathbb {N}}_0\) we have

$$\begin{aligned} \sum _{k=1}^N \varepsilon _k\int _{I_k}\frac{t^p}{p!}e^{-t}\ dt \le 1 - \exp \left( -\sum _{k=1}^N \varepsilon _k |I_k|\right) . \end{aligned}$$

Proof

We denote by n the number of indexes k such that \(0< \varepsilon _k < 1\) and we proceed by induction on n. For \(n = 0\) this is the content of Lemma 1. Let us now assume \(n = 1.\) Let \(1\le j\le N\) be the coordinate with the property that \(0< \varepsilon _j < \) and check that

$$\begin{aligned} \psi (\varepsilon ):= \sum _{k\ne j}\int _{I_k}\frac{t^p}{p!}e^{-t}\ dt + \varepsilon \int _{I_j}\frac{t^p}{p!}e^{-t}\ dt + \exp \left( -\sum _{k\ne j}|I_k| - \varepsilon |I_j|\right) \le 1\end{aligned}$$

for every \(0 \le \varepsilon \le 1.\) In fact, \(\psi (0) \le 1\) and \(\psi (1)\le 1\) follow from Lemma 1. Moreover, the critical point \(\varepsilon _0\) of \(\psi \) satisfies

$$\begin{aligned} \int _{I_j}\frac{t^p}{p!}e^{-t}\ dt = |I_j|\exp \left( -\sum _{k\ne j}|I_k| - \varepsilon _0 |I_j|\right) . \end{aligned}$$

Hence

$$\begin{aligned} \psi (\varepsilon _0)= & {} \sum _{k\ne j}\int _{I_k}\frac{t^p}{p!}e^{-t}\ dt + \varepsilon _0|I_j|\exp \left( -\sum _{k\ne j}|I_k| - \varepsilon _0 |I_j|\right) \\&+ \exp \left( -\sum _{k\ne j}|I_k| - \varepsilon |I_j|\right) \\= & {} \sum _{k\ne j}\int _{I_k}\frac{t^p}{p!}e^{-t}\ dt + \left( 1 + \varepsilon _0|I_j|\right) \exp \left( -\sum _{k\ne j}|I_k| - \varepsilon |I_j|\right) . \end{aligned}$$

Since

$$\begin{aligned} 1 + \varepsilon _0|I_j| \le \exp \big (\varepsilon _0 |I_j|\big ) \end{aligned}$$

we conclude

$$\begin{aligned} \psi (\varepsilon _0) \le \sum _{k\ne j}\int _{I_k}\frac{t^p}{p!}e^{-t}\ dt + \exp \left( -\sum _{k\ne j}|I_k|\right) \le 1. \end{aligned}$$

Let us assume that the Lemma holds for \(n = \ell \) (\(0\le \ell < N\)) and let \(n = \ell +1.\) We consider the function \(\psi :[0,1]^{\ell +1}\rightarrow {\mathbb {R}}\) defined by

$$\begin{aligned} \psi ({\varvec{\varepsilon }}):= \sum _{k=1}^{\ell +1}\varepsilon _k \int _{I_k}\frac{t^p}{p!}e^{-t}\ dt + \sum _j\int _{J_j}\frac{t^p}{p!}e^{-t}\ dt + \exp \left( -\sum _k\varepsilon _k |I_k| - \sum _j |J_j|\right) \end{aligned}$$

for \({\varvec{\varepsilon }} = (\varepsilon _1, \ldots , \varepsilon _{\ell +1}).\) The induction hypothesis means that \(\psi ({\varvec{\varepsilon }}) \le 1\) whenever \({\varvec{\varepsilon }}\) is in the boundary of \([0,1]^{\ell +1}.\) The lemma is proved after checking that \(\psi ({\varvec{\varepsilon }}_0) \le 1,\) where \({\varvec{\varepsilon }_0}\) is a critical point of \(\psi .\) Proceeding as before,

$$\begin{aligned} \psi ({\varvec{\varepsilon }}_0)= & {} \left( \sum _{k=1}^{\ell +1}\varepsilon _k |I_k| + 1\right) e^{-\sum _k \varepsilon _k |I_k|} e^{-\sum _j |J_j|} + \sum _j \int _{J_j}\frac{t^p}{p!}e^{-t}\ dt \\\le & {} \exp \left( -\sum _j |J_j|\right) + \sum _j \int _{J_j}\frac{t^p}{p!}e^{-t}\ dt \le 1. \end{aligned}$$

\(\square \)

Proof of Theorem 1

We first assume \(a = 0,\) that is, F is radial. After replacing F by \(G = \frac{F}{\Vert F\Vert _\infty }\) if necessary we can assume that \(\Vert F\Vert _\infty = 1.\) Since F is radial we have \(F(z) = g\big (|z|\big ).\) We aim to prove that

$$\begin{aligned} \int _{{\mathbb {C}}}\left| g\big (|z|\big )\right| \cdot \left| f(z)\right| ^2 e^{-\pi |z|^2}\ dA(z) \le 1 - \exp \left( -2\pi \int _0^\infty r\left| g(r)\right| \ dr\right) \end{aligned}$$

for every entire function \(f(z) = \sum _{p=0}^\infty b_p e_p\) such that \(\sum _{p=0}^\infty |b_p|^2 = 1.\) We have

$$\begin{aligned} \int _{{\mathbb {C}}}\left| g\big (|z|\big )\right| \cdot \left| F(z)\right| ^2 e^{-\pi |z|^2}\ dA(z) = \sum _{p=0}^\infty |b_p|^2\int _0^\infty \left| g\left( \sqrt{\frac{t}{\pi }}\right) \right| \cdot \frac{t^p}{p!}e^{-t}\ dt. \end{aligned}$$

Let us first assume

$$\begin{aligned} g = \sum _{k=1}^N \varepsilon _k \chi _{I_k},\ \left| \varepsilon _k\right| \le 1, \end{aligned}$$
(1)

where \((I_k)_{k=1}^N\) are disjoint intervals. Then, Lemma 1 gives

$$\begin{aligned} \sum _{p=0}^\infty |b_p|^2\int _0^\infty \left| g\left( \sqrt{\frac{t}{\pi }}\right) \right| \cdot \frac{t^p}{p!}e^{-t}\ dt\le & {} 1 - \exp \left( -\sum _{k=1}^N |\varepsilon _k| |J_k|\right) \\= & {} 1 - \exp \left( -2\pi \int _0^\infty r|g(r)|\ dr\right) \\= & {} 1 - \exp \left( -\Vert F\Vert _1\right) . \end{aligned}$$

We used \(J_k = \left\{ t:\ \sqrt{\frac{t}{\pi }}\in I_k\right\} \) and \(|J_k| = 2\pi \int _{I_k}r dr.\) Theorem 1 is proved for g as in (1). Let us now assume that \(\Vert g\Vert _\infty \le 1\) and \(g\in L^1({\mathbb {R}}^+,rdr)\cap L^\infty ({\mathbb {R}}^+).\) Then there is a sequence \((g_n)_n\) of step functions as in (1) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\int _0^\infty |g_n(r)-g(r)|\ r dr = 0. \end{aligned}$$

We put \(F_n(z):= g_n(|z|).\) According to [12, Theorem 3.5] there is a constant \(K > 0\) such that

$$\begin{aligned} \Vert T_G\Vert \le K\sup _{z\in {{\mathbb {C}}}}\int _{D(z,1)}|G|\ dA \end{aligned}$$
(2)

for every bounded symbol G,  which implies

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert T_F - T_{F_n}\Vert \le K\lim _{n\rightarrow \infty }\Vert F_n - F\Vert _1 = 0. \end{aligned}$$

We finally conclude

$$\begin{aligned} \Vert T_F\Vert \le 1 - \exp \big (-\Vert F\Vert _1\big ). \end{aligned}$$

In the case \(a\ne 0,\) the identity

$$\begin{aligned} \int _{{\mathbb {C}}} g\big (|z-a|\big )|f(z)|^2 d\lambda (z) = \int _{{\mathbb {C}}} g\big (|u|\big )\left| (W_{-a}f)(u)\right| ^2 d\lambda (u) \end{aligned}$$

and the fact that \(W_{-a}\) is an isometry gives the conclusion. We can also argue from the fact that \(W_{-a}\circ T_F = T_G\circ W_{-a},\) where \(G(z) = g\big (|z|\big ).\) \(\square \)

In particular, if \(\varOmega \subset {{\mathbb {C}}}\) presents radial symmetry with respect to some point then

$$\begin{aligned} \int _\varOmega \left| f(z)\right| ^2 d\lambda (z) \le \Big (1 - e^{-|\varOmega |}\Big )\cdot \int _{{{\mathbb {C}}}}\left| f(z)\right| ^2 d\lambda (z) \end{aligned}$$
(3)

for every \(f\in {{\mathcal {F}}}^2({{\mathbb {C}}}).\)

The question arises whether inequality (3) holds for every subset \(\varOmega .\) This is related to a conjecture by Abreu and Speckbacher in [1] (see the next section). We do not have an answer to this question except for monomials or its translates.

Example 1

Let \(k_w = e^{-\frac{\pi }{2}|w|^2}K_w\) be the normalized reproducing kernel of \({{\mathcal {F}}}^2({{\mathbb {C}}}).\) Then, for every set \(\varOmega \subset {{\mathbb {C}}}\) with finite measure we have

$$\begin{aligned} \int _\varOmega \left| k_w(z)\right| ^2 d\lambda (z) \le 1 - e^{-|\varOmega |}. \end{aligned}$$

Proof

In fact, \(k_w = W_w\left( e_0\right) .\) Hence

$$\begin{aligned} \int _\varOmega |k_w(z)|^2 d\lambda (z) = \int _{\varOmega -w} d\lambda (z) \end{aligned}$$

and the conclusion follows from the fact that the last integral attains its maximum when \(\varOmega \) is a disc centered at w (see the comment after [1, Conjecture 1]). \(\square \)

It is easy to check that when \(\varOmega \) is a disc centered at point \(\omega \) the inequality in Example 1 is an identity.

Proposition 1

Let \(\varOmega \subset {\mathbb {R}}^2\) be a set with finite measure. Then, for every \(n\in {\mathbb {N}}\) and \(a\in {{\mathbb {C}}},\)

$$\begin{aligned} \int _\varOmega \left| W_a(e_n)(z)\right| ^2 d\lambda (z)\le 1-e^{-|\varOmega |}. \end{aligned}$$

Proof

Since

$$\begin{aligned} \int _\varOmega \left| W_a(e_n)(z)\right| ^2 d\lambda (z) = \int _{\varOmega -a} \left| e_n(z)\right| ^2 d\lambda (z) \end{aligned}$$

we can assume that \(a = 0.\) For every \(\theta \in [0,2\pi ]\) we denote

$$\begin{aligned} \varOmega _\theta = \left\{ r \ge 0:\ re^{i\theta }\in \varOmega \right\} . \end{aligned}$$

Then

$$\begin{aligned} \int _\varOmega \left| e_n(z)\right| ^2 d\lambda (z)= & {} \frac{\pi ^n}{n!}\int _\varOmega \left| z^n\right| ^2 e^{-\pi |z|^2}\ dA(z)\\= & {} \frac{\pi ^n}{n!}\int _0^{2\pi }\Bigg (\int _{\varOmega _\theta }r^{2n}e^{-\pi r^2}2\pi r\ dr\Bigg ) \frac{d\theta }{2\pi }\\= & {} \int _0^{2\pi }\Bigg (\int _{I_\theta }\frac{t^n}{n!} e^{-t}\ dt\Bigg ) \frac{d\theta }{2\pi }, \end{aligned}$$

where

$$\begin{aligned} I_\theta = \left\{ t = \pi r^2:\ r\in \varOmega _\theta \right\} . \end{aligned}$$

Since \(|\varOmega | < \infty \) then a.e. \(\theta \in [0,2\pi ]\) we have

$$\begin{aligned} |I_\theta | = 2\pi \int _{\varOmega _\theta }r dr < +\infty . \end{aligned}$$

Moreover, by Lemma 1,

$$\begin{aligned} \int _0^{2\pi }\left( \int _{I_\theta }\frac{t^n}{n!} e^{-t}\ dt\right) \frac{d\theta }{2\pi } \le \int _0^{2\pi } \left( 1-e^{-|I_\theta |}\right) \frac{d\theta }{2\pi }. \end{aligned}$$

Finally we consider the convex function \(f(t) = e^{-t}-1\) and the probability measure \(\frac{d\theta }{2\pi }\) and put \(h(\theta ) = |I_\theta |.\) Jensen’s inequality gives

$$\begin{aligned} f\left( \int _0^{2\pi }h(\theta )\frac{d\theta }{2\pi }\right) \le \int _0^{2\pi }f\left( h(\theta )\right) \frac{d\theta }{2\pi }, \end{aligned}$$

which means

$$\begin{aligned} \int _0^{2\pi } \left( 1-e^{-|I_\theta |}\right) \frac{d\theta }{2\pi }\le & {} 1 - \exp \left( -\int _0^{2\pi }|I_\theta |\ \frac{d\theta }{2\pi }\right) \\= & {} 1 - \exp \left( -\int _0^{2\pi }\left( \int _{\varOmega _\theta } r\ dr\right) \ d\theta \right) \\= & {} 1 - e^{-|\varOmega |}. \end{aligned}$$

\(\square \)

We finish the section with some examples of sets \(\varOmega \) with infinite Lebesgue measure for which the Toeplitz operator with symbol \(F=\chi _\varOmega \) has norm as small as we want.

Proposition 2

For every \(\varepsilon > 0\) there exists \(\varOmega \) with infinite Lebesgue measure such that

$$\begin{aligned} \int _\varOmega |f(z)|^2 d\lambda (z) \le \varepsilon \int _{{{\mathbb {C}}}} |f(z)|^2 d\lambda (z) \end{aligned}$$

for every \(f\in {{\mathcal {F}}}^2.\)

Proof

Let \(K > 0\) as in (2) and let \((\varOmega _n)_n\) be a sequence of bounded sets with Lebesgue measure \(|\varOmega _n| = \frac{\varepsilon }{K}\) and such that \(\text{ dist } (\varOmega _n, \varOmega _m) > 2\) whenever \(n\ne m,\) and take \(\varOmega = \cup _{n\in {{\mathbb {N}}}} \varOmega _n.\) Since each disc D(z, 1) meets at most one set \(\varOmega _n\) we have \(|\varOmega \cap D(z,1)|\le \frac{\varepsilon }{K}.\) The estimate (2) turns out \(\Vert T_{\chi _\varOmega }\Vert \le \varepsilon ,\) which gives the conclusion. \(\square \)

3 Time-Frequency Localization Operators

For \(F\in L^1({{\mathbb {C}}})\) we denote by \(H_F:L^2({\mathbb {R}})\rightarrow L^2({\mathbb {R}})\) the localization operator

$$\begin{aligned} H_F f = \int _{{\mathbb {C}}} F(z)\ \langle f, \pi (z)h_0\rangle \ \pi (z)h_0\ dA(z). \end{aligned}$$

Here \(h_0(t) = 2^{1/4} e^{-\pi t^2}\) is the Gaussian and \(\pi (z)\) is the time-frequency shift, defined for \(z = x+i\omega \) as

$$\begin{aligned} \big (\pi (z)f\big )(t) = e^{2\pi i \omega t}f(t-x),\ f\in L^2({\mathbb {R}}). \end{aligned}$$

In case F is the characteristic function of a set \(\varOmega \) we write \(H_\varOmega \) instead of \(H_{\chi _\varOmega }.\) We refer to [5] or [6, Chapter 4] for general facts concerning localization operators.

For \(f,g\in L^2({\mathbb {R}}),\) the expression

$$\begin{aligned} \big (V_g f\big )(z) := \langle f, \pi (z)g\rangle \end{aligned}$$

is the short time Fourier transform of f with window g,  known as Gabor transform in the case where the window \(g = h_0\) is the Gaussian.

If F is real-valued then \(H_F\) is a selfadjoint operator on \(L^2({\mathbb {R}}),\) hence

$$\begin{aligned} \Vert H_F\Vert = \sup _{\Vert f\Vert _2 = 1}\left| \langle H_F f,\ f\rangle \right| \le \sup _{\Vert f\Vert _2 = 1}\int _{{{\mathbb {C}}}} |F(z)|\cdot \left| \left( V_{h_0} f\right) (z)\right| ^2 dA(z). \end{aligned}$$

There is a connection between localization operators and Toeplitz operators on the Fock space via the Bargmann transform.

The Bargmann transform is the surjective and unitary operator

$$\begin{aligned} {{\mathcal {B}}}:L^2({\mathbb {R}}) \rightarrow {{\mathcal {F}}}^2({{\mathbb {C}}}) \end{aligned}$$

defined as

$$\begin{aligned} \big ({{\mathcal {B}}}f\big )(z) = 2^{1/4}\int _{{\mathbb {R}}}f(t) e^{2\pi t z -\pi t^2-\frac{\pi }{2}z^2}\ dt. \end{aligned}$$

It was introduced in [2] and has the important property that the Hermite functions are mapped into normalized analytic monomials. More precisely, \({{\mathcal {B}}}(h_n) = e_n,\) where \(h_n\) is defined via the so called Rodrigues formula as

$$\begin{aligned} h_n(t) = \frac{2^{1/4}}{\sqrt{n!}}\left( \frac{-1}{2\sqrt{\pi }}\right) ^n e^{\pi t^2}\frac{d^n}{dt^n}\left( e^{-2\pi t^2}\right) ,\ \ n\ge 0. \end{aligned}$$

Then \(\left( h_n\right) _{n\ge 0}\) forms an orthonormal basis for \(L^2({\mathbb {R}}).\) The Gabor transform of Hermite functions is well-known (see for instance [9, Chapter 1.9]). In fact, for \(z = x + i\xi ,\)

$$\begin{aligned} \langle h_n, \pi (z)h_0\rangle = e^{-i\pi x\xi - \frac{\pi }{2}|z|^2}\sqrt{\frac{\pi ^n}{n!}}{\overline{z}}^n. \end{aligned}$$
(4)

Since for \(z = x + i\xi \) we have ( [10, 3.4.1])

$$\begin{aligned} \big (V_{h_0} f\big )(x,-\xi ) = e^{i\pi x\xi }\cdot \big ({{\mathcal {B}}}f\big )(z)\cdot e^{-\frac{\pi |z|^2}{2}} \end{aligned}$$

then, for every \(f\in L^2({\mathbb {R}})\) and \(F\in L^1({{\mathbb {C}}})\cap L^\infty ({{\mathbb {C}}})\) we obtain

$$\begin{aligned} \int _{{{\mathbb {C}}}} |F(z)|\cdot \left| \left( V_{h_0} f\right) (z)\right| ^2 dA(z) = \int _{{{\mathbb {C}}}} |F(z)|\cdot \left| \big (Bf\big )(z)\right| ^2\ d\lambda (z). \end{aligned}$$

Consequently, all the estimates in the previous section can be translated into estimates concerning localization operators.

Abreu, Speckbacher conjecture in [1] that, among all the sets with a given measure, \(\Vert H_\varOmega \Vert \) attains its maximum when \(\varOmega \) is a disc, up to perturbations of Lebesgue measure zero. This turns out to be equivalent to the validity of inequality (3) for every function in the Fock space or, equivalently, to the fact that

$$\begin{aligned} \Vert f\Vert _2^2 \le e^{|\varOmega |}\int _{{{\mathbb {C}}}\setminus \varOmega } \left| \big (V_{h_0}f\big )(z)\right| ^2 dA(z)\ \ \forall f\in L^2({\mathbb {R}}). \end{aligned}$$

In this regard it is worth noting that Nazarov [13] proved the existence of two absolute constants AB such that

$$\begin{aligned} \Vert f\Vert _2^2 \le Ae^{B\cdot |S|\cdot |\Sigma |}\left( \int _{{\mathbb {R}}\setminus S} |f|^2 + \int _{{\mathbb {R}}\setminus \Sigma }|{\widehat{f}}|^2\right) \end{aligned}$$

for every \(f\in L^2({\mathbb {R}})\) and for any pair \((S,\Sigma )\) of sets with finite measure. Also, it follows from [8, Theorem 4.1] that for every set \(\varOmega \subset {{\mathbb {R}}}^{2}\) thin at infinity and for every \(g\in L^2({{\mathbb {R}}})\) there exist a constant \(C > 0\) such that

$$\begin{aligned} \Vert f\Vert _2^2 \le C\int _{{{\mathbb {C}}}\setminus \varOmega } \left| \big (V_{g}f\big )(z)\right| ^2 dA(z)\ \ \forall f\in L^2({\mathbb {R}}). \end{aligned}$$

From Theorem 1 and Proposition 1 we get the following.

Corollary 1

Let \(F\in L^1({{\mathbb {C}}})\cap L^\infty ({{\mathbb {C}}})\) be a real-valued and radial symbol with respect to \(a\in {{\mathbb {C}}}.\) Then

$$\begin{aligned} \Vert H_F\Vert \le \Vert F\Vert _\infty \Bigg (1 - \exp \Bigg (-\frac{\Vert F\Vert _1}{\Vert F\Vert _\infty }\Bigg )\Bigg ). \end{aligned}$$

Corollary 2

Let \(\varOmega \subset {\mathbb {R}}^2\) be a set with finite measure. Then, for every \(n\in {\mathbb {N}},\)

$$\begin{aligned} \left| \langle H_\varOmega h_n,\ h_n\rangle \right| \le 1-e^{-|\varOmega |}. \end{aligned}$$

We fix a non-zero window \(g\in L^2({\mathbb {R}}).\) The modulation space \(M^1({\mathbb {R}}),\) also known as Feichtinger algebra, is the set of tempered distributions \(f\in {{\mathcal {S}}}^\prime ({\mathbb {R}})\) such that

$$\begin{aligned} \Vert f\Vert _{M^1}:= \int _{{{\mathbb {C}}}}\left| \langle f, \pi (z)g\rangle \right| dA(z) < +\infty . \end{aligned}$$

The use of different windows g in the definition of \(M^1({\mathbb {R}})\) yields the same spaces with equivalent norms. It is well known that \(M^1({\mathbb {R}})\) is continuously included in \(L^2({\mathbb {R}})\) and

$$\begin{aligned} \Vert f\Vert _2 = \Vert V_g f\Vert _ 2 \le \Vert V_g f\Vert _1 \end{aligned}$$

whenever \(f\in M^1({\mathbb {R}})\) and \(\Vert g\Vert _2 = 1.\) See for instance [10, 3.2.1] for the first identity.

Proposition 3

Let \(\varOmega \subset {\mathbb {R}}^2\) be a set with finite measure. Then, for every \(f\in M^1({\mathbb {R}})\) and \(n\in {\mathbb {N}}_0\) we have

$$\begin{aligned} \int _\varOmega \left| \left( V_{h_0} f\right) (z)\right| ^2 dA(z) \le \Vert V_{h_n} f\Vert _1^2\cdot \big (1-e^{-|\varOmega |}\big ). \end{aligned}$$

Proof

It suffices to prove the proposition under the additional assumption that \(\Vert V_{h_n}f\Vert _1 = 1.\) Fixed \(n\in {\mathbb {N}}_0\) we consider the set

$$\begin{aligned} B:=\left\{ \pi (z)h_n:\ z\in {{\mathbb {C}}}\right\} \subset L^2({\mathbb {R}}). \end{aligned}$$

Then

$$\begin{aligned} B^\circ := \left\{ g\in L^2({\mathbb {R}}):\ \left| \langle g, \pi (z)h_n\rangle \right| \le 1\right\} = \left\{ g\in L^2({\mathbb {R}}):\ \Vert V_{h_n}g\Vert _\infty \le 1\right\} . \end{aligned}$$

We have

$$\begin{aligned} \left| \langle f, g\rangle \right| = \left| \langle V_{h_n}f, V_{h_n}g\rangle \right| \le \Vert V_{h_n}f\Vert _1\cdot \Vert V_{h_n}g\Vert _\infty \le 1 \end{aligned}$$

for every \(g\in B^\circ ,\) which means that \(f\in B^{\circ \circ }.\) According to the bipolar theorem,

$$\begin{aligned} f = L^2-\lim _{k\rightarrow \infty } f_k \end{aligned}$$

where each \(f_k\) is in the absolutely convex hull of B. For each \(k\in {\mathbb {N}}\) we can find scalars \((\alpha _j)_{j= 1}^N\) and points \((z_j)_{j=1}^N\) such that \(f_k = \sum _{j=1}^N\alpha _j \pi (z_j)h_n\) and \(\sum _{j=1}^N |\alpha _j| \le 1.\) Then

$$\begin{aligned} \Bigg (\int _\varOmega \left| \left( V_{h_0} f_k\right) (z)\right| ^2 dA(z)\Bigg )^{\frac{1}{2}}= & {} \Bigg (\int _\varOmega \left| \langle f_k, \pi (z)\varphi \rangle \right| ^2 dA(z)\Bigg )^{\frac{1}{2}} \\\le & {} \sum _{j=1}^N |\alpha _j|\Bigg (\int _\varOmega \left| \langle \pi (z_j)h_n, \pi (z)\varphi \rangle \right| ^2 dA(z)\Bigg )^{\frac{1}{2}}\\= & {} \sum _{j=1}^N |\alpha _j|\Bigg (\int _\varOmega \left| \langle h_n, \pi (z-z_j)\varphi \rangle \right| ^2 dA(z)\Bigg )^{\frac{1}{2}}\\= & {} \sum _{j=1}^N |\alpha _j|\Bigg (\int _{\varOmega -z_j} \left| \langle h_n, \pi (z)\varphi \rangle \right| ^2 dA(z)\Bigg )^{\frac{1}{2}}\\= & {} \sum _{j=1}^N |\alpha _j|\left| \langle H_{\varOmega -z_j}h_n, h_n\rangle \right| ^{\frac{1}{2}} \le \big (1-e^{-|\varOmega |}\big )^{\frac{1}{2}}. \end{aligned}$$

Finally,

$$\begin{aligned} \int _\varOmega \left| \left( V_{h_0} f\right) (z)\right| ^2 dA(z) = \lim _{k\rightarrow \infty }\int _\varOmega \left| \left( V_{h_0} f_k\right) (z)\right| ^2 dA(z) \le 1-e^{-|\varOmega |}. \end{aligned}$$

\(\square \)

The next result is a direct consequence of Proposition 2.

Corollary 3

For every \(\varepsilon > 0\) there exists \(\varOmega \) with infinite Lebesgue measure such that

$$\begin{aligned} \Vert H_\varOmega \Vert \le \varepsilon . \end{aligned}$$