Keywords

1 Introduction

Modulation spaces were introduced in Feichtinger’s seminal technical report [17], and prove themselves as useful family of Banach spaces of tempered distributions in time-frequency analysis, [4, 10, 28]. The main purpose of this survey article is to enlighten some properties of modulation spaces in a rather self-contained manner. In contrast to the most common situation, our analysis includes both quasi-Banach and Banach modulation spaces within the framework of ultradifferentiable functions and ultradistributions of Gelfand–Shilov type. For that reason we collect necessary background material in a rather detailed preliminary section.

Motivated by recent applications of modulation spaces in the context of nonlinear harmonic analysis and its applications, cf. [4,5,6, 14, 22, 38, 39, 47, 54] we focus our attention to boundedness for multiplications and convolutions for elements in such spaces. The basic results in that direction go back to the original contribution [17], and were thereafter reconsidered by many authors in different contexts. Let us give a brief, and unavoidably incomplete account on the related results.

In Sect. 3 we formulate in Theorems 3.5 and 3.7 bilinear versions of more general multiplication and convolution results in [54, Section 3]. The contents of Theorems 3.5 and 3.7 in the unweighted case for modulation spaces M p, q can be summarized as follows.

Proposition 1.1

Let p j, q j ∈ (0, ], j = 0, 1, 2,

$$\displaystyle \begin{aligned} \theta _1=\max \left (1 ,\frac 1{p_0},\frac 1{q_1},\frac 1{q_2} \right ) \quad \mathit{\text{and}}\quad \theta _2=\max \left (1 ,\frac 1{p_1},\frac 1{p_2} \right ). \end{aligned}$$

Then

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} M^{p_1,q_1}\cdot M^{p_2,q_2} &\subseteq M^{p_0,q_0},& \qquad \frac 1{p_1}+\frac 1{p_2} &=\frac 1{p_0}, & \quad \frac 1{q_1}+\frac 1{q_2} &= \theta _1 + \frac 1{q_0}, \\ {} M^{p_1,q_1}* M^{p_2,q_2} &\subseteq M^{p_0,q_0}, & \qquad \frac 1{p_1}+\frac 1{p_2} &= \theta _2+\frac 1{p_0}, & \quad \frac 1{q_1}+\frac 1{q_2} &= \frac 1{q_0}. \end{array}\end{aligned} $$

The general multiplication and convolution properties in Sect. 3 also overlap with results by Bastianoni, Cordero and Nicola in [2], by Bastianoni and Teofanov in [1], and by Guo et al. in [32].

The multiplication relation in Proposition 1.1 for p j, q j ≥ 1 was obtained already in [17] by Feichtinger. It is also obvious that the convolution relation was well-known since then (though a first formal proof of this relation seems to be given first in [48]). In general, these convolution and multiplication properties follow the rules

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \ell ^{p_1}&*\ell ^{p_2}\subseteq \ell ^{p_0}, & \quad \ell ^{q_1} &\cdot \ell ^{q_2}\subseteq \ell ^{q_0} & \quad &\Rightarrow & \quad M^{p_1,q_1}&*M^{p_2,q_2}\subseteq M^{p_0,q_0} \end{array}\end{aligned} $$

and

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \ell ^{p_1} &\cdot \ell ^{p_2}\subseteq \ell ^{p_0}, & \quad \ell ^{q_1} &* \ell ^{q_2}\subseteq \ell ^{q_0} & \quad &\Rightarrow & \quad M^{p_1,q_1} &\cdot M^{p_2,q_2}\subseteq M^{p_0,q_0}, \end{array}\end{aligned} $$

which goes back to [17] in the Banach space case and to [25] in the quasi-Banach case. See also [19] and [42] for extensions of these relations to more general Banach function spaces and quasi-Banach function spaces, respectively.

In Sect. 3 we basically review some results from [54]. To make this survey self-contained we give the proof of Theorem 3.7 in unweighted case. In contrast to [32], we do not deduce any sharpness for our results.

To show Proposition 1.1 in the quasi-Banach setting, apart from the usual use of Hölder’s and Young’s inequalities, additional arguments are needed. In our situation we discretize the situations in similar ways as in [2] by using Gabor analysis for modulation spaces, and then apply some further arguments, valid in non-convex analysis. This approach is slightly different compared to what is used in [32] which follows the discretization technique introduced in [55], and which has some traces of Gabor analysis.

We refer to [54] for a detailed discussion on the uniqueness of multiplications and convolutions in Proposition 1.1.

In Sect. 4 we apply the results from previous parts in the framework of the so called Gabor product. It is introduced in [14] in order to derive a phase space analogue to the usual convolution identity for the Fourier transform. The main motivation is to use such kind of products in a phase-space formulation of certain nonlinear equations. As noticed in [14], among other interesting characteristics of phase-space representations, the initial value problem in phase-space may be well-posed for more general initial distributions. This means that the phase-space formulation could contain solutions other than the standard ones. We refer to [11,12,13], where the phase-space extensions are explored in different contexts. Here we illustrate this approach by considering the nonlinear cubic Schrödinger equation, which appear for example in Bose-Einstein condensate theory [35]. We also refer to [4, Chapter 7] for an overview of results related to well-posedness of the nonlinear Schrödinger equations in the framework of modulation spaces, see also [3, 38, 39].

2 Preliminaries

In this section we give an exposition of background material related to the definition and basic properties of modulation spaces. Thus we recall some facts on the short-time Fourier transform and related projections, the (Fourier invariant) Gelfand-Shilov spaces, weight functions, and mixed-norm spaces of Lebesgue type. We also recall convolution and multiplication in weighted Lebesgue sequence spaces.

2.1 The Short-Time Fourier Transform

In what follows we let \(\mathcal F\) be the Fourier transform which takes the form

$$\displaystyle \begin{aligned} (\mathcal Ff)(\xi )= \widehat f(\xi ) \equiv (2\pi )^{-\frac d2}\int _{\mathbb R^{d}} f(x)e^{-i\langle x,\xi\rangle }\, dx \end{aligned}$$

when \(f\in L^1(\mathbb R^{d})\). Here 〈 ⋅ ,  ⋅ 〉 denotes the usual scalar product on \(\mathbb R^{d}\). The same notation is used for the usual dual form between test functions and corresponding (ultra-)distributions. We recall that map \(\mathcal F\) extends uniquely to a homeomorphism on the space of tempered distributions \(\mathcal S^{\prime }(\mathbb R^{d})\), to a unitary operator on \(L^2(\mathbb R^{d})\) and restricts to a homeomorphism on the Schwartz space of smooth rapidly decreasing functions \(\mathcal S(\mathbb R^{d})\), cf. (29). We also observe with our choice of the Fourier transform, the usual convolution identity for the Fourier transform takes the forms

$$\displaystyle \begin{aligned} \mathcal F (f\cdot g) = (2\pi )^{-\frac d2}\widehat f *\widehat g \quad \text{and}\quad \mathcal F (f* g) = (2\pi )^{\frac d2}\widehat f \cdot \widehat g \end{aligned} $$
(1)

when \(f,g\in \mathcal S(\mathbb R^{d})\).

In several situations it is convenient to use a localized version of the Fourier transform, called the short-time Fourier transform, STFT for short. The short-time Fourier transform of \(f\in \mathcal S^{\prime }(\mathbb R^{d})\) with respect to the fixed window function \(\phi \in \mathcal S(\mathbb R^{d})\) is defined by

$$\displaystyle \begin{aligned} (V_\phi f)(x,\xi ) &\equiv (2\pi )^{-\frac d2} (f,\phi ( \cdot \, -x)e^{i\langle \, \cdot \,,\xi\rangle })_{L^2}. {} \end{aligned} $$
(2)

Here \(( \cdot \, ,\, \cdot \, )_{L^2}\) denotes the unique continuous extension of the inner product on \(L^2(\mathbb R^{d})\) restricted to \(\mathcal S(\mathbb R^{d})\) into a continuous map from \(\mathcal S^{\prime }(\mathbb R^{d}) \times \mathcal S(\mathbb R^{d})\) to \(\mathbb C\).

We observe that using certain properties for tensor products of distributions,

$$\displaystyle \begin{aligned} (V_\phi f)(x,\xi ) &= \mathcal F (f\cdot \overline {\phi ( \cdot \, -x)})(\xi ). \end{aligned} $$
((2)′)

(cf. [33, 52]). If in addition \(f\in L^p(\mathbb R^{d})\) for some p ∈ [1, ], then

$$\displaystyle \begin{aligned} (V_\phi f)(x,\xi ) &= (2\pi )^{-\frac d2}\int _{\mathbb R^{d}}f(y) \overline {\phi (y-x)}e^{-i\langle y,\xi\rangle }\, dy. \end{aligned} $$
((2)″)

We observe that the domain of V ϕ is \(\mathcal S^{\prime }(\mathbb R^{d})\). The images are contained in \(C^\infty (\mathbb R^{2d})\), the set of smooth functions defined on the phase space \(\mathbb R^{d}\times \mathbb R^{d}\simeq \mathbb R^{2d}\).

The short-time Fourier transform appears in different contexts and under different names. In quantum mechanics it is rather common to call it the coherent state transform (see e.g. [37]). It is also closely related to the so-called Wigner distribution or radar ambiguity function (see e.g. [36]). In time-frequency analysis, it is also sometimes called the Voice transform.

The main idea with the design of short-time Fourier transform is to get the Fourier content, or the frequency resolution of localized functions and distributions. Roughly speaking, short-time Fourier transforms give a simultaneous information both on functions or distributions themselves as well as their Fourier transforms in the sense that the map

$$\displaystyle \begin{aligned} x\mapsto V_\phi f(x,\xi ) \end{aligned}$$

resembles on f(x), while the map

$$\displaystyle \begin{aligned} \xi \mapsto V_\phi f(x,\xi ) \end{aligned}$$

resembles on \(\widehat f(\xi )\).

As for the ordinary Fourier transform, there are several mapping properties which hold true for the short-time Fourier transform. As an elegant way to approach such properties in the framework of distributions, we may follow ideas given in [24] by Folland.

In fact, let T be the semi-conjugated tensor map

$$\displaystyle \begin{aligned} T(f,\phi ) &= f\otimes \overline \phi , \end{aligned} $$
(3)

U be the linear pullback

$$\displaystyle \begin{aligned} (UF)(x,y) &= U(y,y-x) \end{aligned} $$
(4)

and \(\mathcal F _2\) be the partial Fourier transform given by

$$\displaystyle \begin{aligned} (\mathcal F _2F)(x,\xi ) &= (2\pi )^{-\frac d2}\int _{\mathbb R^{d}}F(x,y)e^{-i\langle y,\xi\rangle }\, dy . \end{aligned} $$
(5)

Then

$$\displaystyle \begin{aligned} V_\phi f = (\mathcal F _2 \circ U\circ T)(f,\phi ) , \end{aligned} $$
(6)

when \(f,\phi \in \mathcal S(\mathbb R^{d})\).

We observe that the mappings

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} T : \mathcal S(\mathbb R^{d}) \times \mathcal S(\mathbb R^{d}) &\to \mathcal S(\mathbb R^{2d}), & \quad U,\mathcal F _2 : \mathcal S(\mathbb R^{2d}) &\to \mathcal S(\mathbb R^{2d}) \end{array}\end{aligned} $$
(7)

are continuous and uniquely extendable to continuous mappings

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} T : \mathcal S^{\prime}(\mathbb R^{d}) \times \mathcal S^{\prime}(\mathbb R^{d}) &\to \mathcal S^{\prime}(\mathbb R^{2d}), & \quad U,\mathcal F _2 : \mathcal S^{\prime}(\mathbb R^{2d}) &\to \mathcal S^{\prime}(\mathbb R^{2d}), \end{array}\end{aligned} $$
(8)

which in turn restricts to isometric mappings

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} T : L^2(\mathbb R^{d}) \times L^2(\mathbb R^{d}) &\to L^2(\mathbb R^{2d}), & \quad U,\mathcal F _2 : L^2(\mathbb R^{2d}) &\to L^2(\mathbb R^{2d}). \end{array}\end{aligned} $$
(9)

Here that T is isometric means that

$$\displaystyle \begin{aligned} \Vert T(f,\phi )\Vert _{L^2(\mathbb R^{2d})} = \Vert f\Vert _{L^2(\mathbb R^{d})}\Vert \phi\Vert _{L^2(\mathbb R^{d})}. \end{aligned}$$

It is now natural to define V ϕ f as the right-hand side of (6) when \(f,\phi \in \mathcal S^{\prime }(\mathbb R^{d})\), in which V ϕ f is well-defined as an element in \(\mathcal S^{\prime }(\mathbb R^{2d})\).

Proposition 2.1

The map

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} (f,\phi ) &\mapsto V_\phi f & &:\, & &\mathcal S(\mathbb R^{d}) \times \mathcal S(\mathbb R^{d}) & &\to \mathcal S(\mathbb R^{2d}) {} \end{array}\end{aligned} $$
(10)

is continuous, which extends uniquely to a continuous map

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} (f,\phi ) &\mapsto V_\phi f & &:\, & &\mathcal S^{\prime}(\mathbb R^{d}) \times \mathcal S^{\prime}(\mathbb R^{d}) & &\to \mathcal S^{\prime}(\mathbb R^{2d}), {} \end{array}\end{aligned} $$
(11)

which in turn restricts to an isometric map

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} (f,\phi ) &\mapsto V_\phi f & &:\, & &L^2(\mathbb R^{d}) \times L^2(\mathbb R^{d}) & &\to L^2(\mathbb R^{2d}). {} \end{array}\end{aligned} $$
(12)

If \(\phi \in \mathcal S(\mathbb R^{d})\) and \(f\in \mathcal S^{\prime }(\mathbb R^{d})\), then (11) shows that \(V_\phi f\in \mathcal S^{\prime }(\mathbb R^{2d})\). On the other hand, it is easy to see that the right-hand side of (2) defines a smooth function. Consequently beside (11) and (10), we also have the continuous map

$$\displaystyle \begin{aligned} (f,\phi ) \mapsto V_\phi f :\, \mathcal S^{\prime}(\mathbb R^{d}) \times \mathcal S(\mathbb R^{d}) \to \mathcal S^{\prime}(\mathbb R^{2d}) \cap C^\infty (\mathbb R^{2d}). \end{aligned} $$
(13)

For short-time Fourier transform, the Parseval identity is replaced by the so-called Moyal identity, also known as the orthogonality relation given by

$$\displaystyle \begin{aligned} (V_{\phi}f,V_{\psi}g)_{L^2(\mathbb R^{2d})} = (\psi ,\phi )_{L^2(\mathbb R^{d})}(f,g)_{L^2(\mathbb R^{d})}, \end{aligned} $$
(14)

when \(f,g,\phi ,\psi \in \mathcal S(\mathbb R^{d})\). The identity (14) is obtained by rewriting the short-time Fourier transforms by (2) and then applying the Parseval identity in suitable ways. We observe that the right-hand side makes sense also when f, g, ϕ and ψ belong to other spaces than \(\mathcal S(\mathbb R^{d})\). For example we may let

$$\displaystyle \begin{aligned} (f,g,\phi ,\psi ) &\in \mathcal S^{\prime}(\mathbb R^{d})\times \mathcal S(\mathbb R^{d})\times \mathcal S(\mathbb R^{d}) \times \mathcal S^{\prime}(\mathbb R^{d}), \\ {} (f,g,\phi ,\psi ) &\in \mathcal S(\mathbb R^{d})\times \mathcal S^{\prime}(\mathbb R^{d})\times \mathcal S^{\prime}(\mathbb R^{d}) \times \mathcal S(\mathbb R^{d}), \\ {} (f,g,\phi ,\psi ) &\in \mathcal S^{\prime}(\mathbb R^{d})\times \mathcal S(\mathbb R^{d})\times L^q(\mathbb R^{d}) \times L^{q^{\prime}}(\mathbb R^{d}) \\ {} \text{or}\qquad (f,g,\phi ,\psi ) &\in L^p(\mathbb R^{d})\times L^{p^{\prime}}(\mathbb R^{d})\times L^q(\mathbb R^{d}) \times L^{q^{\prime}}(\mathbb R^{d}), \end{aligned} $$
(15)

when p, p , q, q ∈ [1, ] satisfy

$$\displaystyle \begin{aligned} \frac 1p+\frac 1{p^{\prime}}=\frac 1q+\frac 1{q^{\prime}}=1. \end{aligned}$$

By Moyal’s identity (14) it follows that if \(\phi \in \mathcal S(\mathbb R^{d})\setminus \{ 0 \}\), then the identity operator on \(\mathcal S^{\prime }(\mathbb R^{d})\) is given by

$$\displaystyle \begin{aligned} \operatorname{Id} = \left (\Vert \phi\Vert _{L^2}^{-2}\right ) \cdot V_\phi ^*\circ V_\phi , \end{aligned} $$
(16)

provided suitable mapping properties of the (L 2-)adjoint \(V_\phi ^*\) of V ϕ can be established. Obviously, \(V_\phi ^*\) fullfils

$$\displaystyle \begin{aligned} (V_\phi ^*F,g) _{L^2(\mathbb R^{d})} = (F,V_\phi g) _{L^2(\mathbb R^{2d})} \end{aligned} $$
(17)

when \(F\in \mathcal S(\mathbb R^{2d})\) and \(g\in \mathcal S(\mathbb R^{d})\).

By expressing the scalar product and the short-time Fourier transform in terms of integrals in (17), it follows by straight-forward manipulations that the adjoint in (17) is given by

$$\displaystyle \begin{aligned} (V_\phi ^*F)(x) = (2\pi )^{-\frac d2} \iint _{\mathbb R^{2d}} F(y,\eta )\phi (x-y)e^{i\langle x,\eta\rangle }\, dyd\eta , \end{aligned} $$
(18)

when \(F\in \mathcal S(\mathbb R^{2d})\). We may now use mapping properties like (11)–(12) to extend the definition of \(V_\phi ^*F\) when F and ϕ belong to various classes of function and distribution spaces. For example, by (11), (10) and (12), it follows that the map

$$\displaystyle \begin{aligned} (F,g)\mapsto (F,V_\phi g) _{L^2(\mathbb R^{2d})} \end{aligned}$$

defines a sesqui-linear form on \(\mathcal S(\mathbb R^{2d})\times \mathcal S^{\prime }(\mathbb R^{d})\), \(\mathcal S^{\prime }(\mathbb R^{2d})\times \mathcal S(\mathbb R^{d})\) and on \(L^2(\mathbb R^{2d})\times L^2(\mathbb R^{d})\). This implies that if \(\phi \in \mathcal S(\mathbb R^{d})\), then \(V_\phi ^*\) in (17) is continuous from \(\mathcal S(\mathbb R^{2d})\) to \(\mathcal S(\mathbb R^{d})\) which is uniquely extendable to a continuous map \(\mathcal S^{\prime }(\mathbb R^{2d})\) to \(\mathcal S^{\prime }(\mathbb R^{d})\), and to \(L^2(\mathbb R^{2d})\) to \(L^2(\mathbb R^{d})\). That is, the mappings

$$\displaystyle \begin{aligned} \begin{alignedat}{3} V_\phi ^* : \mathcal S(\mathbb R^{2d}) \to \mathcal S(\mathbb R^{d}), \qquad &V_\phi ^* & &: & \mathcal S^{\prime}(\mathbb R^{2d}) &\to \mathcal S^{\prime}(\mathbb R^{d}) \\ {} \text{and}\qquad &V_\phi ^* & &: & L^2(\mathbb R^{2d}) &\to L^2(\mathbb R^{d}) \end{alignedat} \end{aligned} $$
(19)

are continuous.

2.2 STFT Projections and a Suitable Twisted Convolution

If \(\phi \in \mathcal S(\mathbb R^{d})\) satisfies \(\Vert \phi \Vert _{L^2}=1\), then (16) shows that \(V_\phi ^*\circ V_\phi \) is the identity operator on \(\mathcal S^{\prime }(\mathbb R^{d})\). If we swap the order of this composition we get certain types of projections. In fact, for any \(\phi \in \mathcal S(\mathbb R^{d})\setminus \{ 0 \}\), let P ϕ be the operator given by

$$\displaystyle \begin{aligned} P_\phi \equiv \Vert \phi\Vert _{L^2}^{-2}\cdot V_\phi \circ V_\phi ^*. \end{aligned} $$
(20)

We observe that P ϕ is continuous on \(\mathcal S(\mathbb R^{2d})\), \(L^2(\mathbb R^{2d})\) and \(\mathcal S^{\prime }(\mathbb R^{2d})\) due to the mapping properties for V ϕ and \(V_\phi ^*\) above.

It is clear that \(P_\phi ^*=P_\phi \), i.e. P ϕ is self-adjoint. Furthermore, P ϕ is an projection:

Hence,

$$\displaystyle \begin{aligned} P_\phi ^* = P_\phi \quad \text{and}\quad P_\phi ^2=P_\phi , \end{aligned} $$
(21)

which shows that P ϕ is an orthonormal projection.

The ranks of P ϕ are given by

$$\displaystyle \begin{aligned} P_\phi (\mathcal S(\mathbb R^{2d})) = V_\phi (\mathcal S(\mathbb R^{d})), \qquad P_\phi (L^2 (\mathbb R^{2d})) &= V_\phi (L^2 (\mathbb R^{d})), \\ {} \text{and}\qquad P_\phi (\mathcal S^{\prime}(\mathbb R^{2d})) &= V_\phi (\mathcal S^{\prime}(\mathbb R^{d})). \end{aligned} $$
(22)

In fact, if \(F\in \mathcal S^{\prime }(\mathbb R^{2d})\), then

$$\displaystyle \begin{aligned} P_\phi F = V_\phi f, \end{aligned}$$

where \(f=\Vert \phi \Vert _{L^2}^{-2}V_\phi ^*F\in \mathcal S^{\prime }(\mathbb R^{d})\). This shows that \(P_\phi (\mathcal S^{\prime }(\mathbb R^{2d})) \subseteq V_\phi (\mathcal S^{\prime }(\mathbb R^{d}))\). On the other hand, if \(f\in \mathcal S^{\prime }(\mathbb R^{d})\) and F = V ϕ f, then

$$\displaystyle \begin{aligned} P_\phi F = \Big (V_\phi \circ \Big ( \Vert \phi\Vert _{L^2}^{-2} \cdot V_\phi ^* \circ V_\phi \Big ) \Big )f = V_\phi f, \end{aligned}$$

which shows that any element in \(V_\phi (\mathcal S^{\prime }(\mathbb R^{d}))\) equals an element in \(P_\phi (\mathcal S^{\prime }(\mathbb R^{2d}))\), i.e. \(P_\phi (\mathcal S^{\prime }(\mathbb R^{2d})) = V_\phi (\mathcal S^{\prime }(\mathbb R^{d}))\). This gives the last identity in (22). In the same way, the first two identities are obtained.

Remark 2.2

Let \(F\in \mathcal S^{\prime }(\mathbb R^{2d})\). Then it follows from the last identity in (22) that F = V ϕ f for some \(f\in \mathcal S^{\prime }(\mathbb R^{d})\), if and only if

$$\displaystyle \begin{aligned} F= P_\phi F. \end{aligned} $$
(23)

Furthermore, if (23) holds, then F = V ϕ f with

$$\displaystyle \begin{aligned} f=(\Vert \phi\Vert _{L^2}^{-2})\cdot V_\phi ^*F . \end{aligned} $$
(24)

There is a twisted convolution which is linked to the projection in (20). In fact, if \(F\in \mathcal S(\mathbb R^{2d})\) and \(\phi \in \mathcal S(\mathbb R^{d})\setminus \{ 0 \}\), then it follows by expanding the integrals for V ϕ and \(V_\phi ^*\) in (20), and performing some straight-forward manipulations that

$$\displaystyle \begin{aligned} P_\phi F = \Vert \phi\Vert _{L^2}^{-2}\cdot V_\phi \phi *_V F , \qquad F\in \mathcal S^{\prime}(\mathbb R^{2d}), \end{aligned} $$
(25)

where the twisted convolutionV is defined by

$$\displaystyle \begin{aligned} (F*_VG)(x,\xi ) &= (2\pi )^{-\frac d2} \iint _{\mathbb R^{2d}} \! \! F(x-y,\xi -\eta )G(y,\eta )e^{-i\langle y,\xi -\eta\rangle }\, dyd\eta . \notag \\ {} &= (2\pi )^{-\frac d2} \iint _{\mathbb R^{2d}} \! \! F(y,\eta )G(x-y,\xi - \eta )e^{-i\langle x-y,\eta\rangle }\, dyd\eta , {} \end{aligned} $$
(26)

when \(F,G\in \mathcal S(\mathbb R^{2d})\). We observe that the definition of ∗V is uniquely extendable in different ways. For example, Young’s inequality for ordinary convolution also holds for the twisted convolution. Moreover, the map (F, G)↦F ∗V G extends uniquely to continuous mappings from \(\mathcal S(\mathbb R^{2d}) \times \mathcal S^{\prime }(\mathbb R^{2d})\) or \(\mathcal S^{\prime }(\mathbb R^{2d})\times \mathcal S(\mathbb R^{2d})\) to \(\mathcal S^{\prime }(\mathbb R^{2d})\). By straight-forward computations it follows that

$$\displaystyle \begin{aligned} (F*_VG)*_VH = F*_V(G*_VH), \end{aligned} $$
(27)

when \(F,H\in \mathcal S(\mathbb R^{2d})\) and \(G\in \mathcal S^{\prime }(\mathbb R^{2d})\), or \(F,H\in \mathcal S^{\prime }(\mathbb R^{2d})\) and \(G\in \mathcal S(\mathbb R^{2d})\).

Let \(f\in \mathcal S^{\prime }(\mathbb R^{d})\) and \(\phi _j\in \mathcal S(\mathbb R^{d})\), j = 1, 2, 3. By straight-forward applications of Parseval’s formula it follows that

$$\displaystyle \begin{aligned} \big ( (V_{\phi _2}\phi _3) *_V(V_{\phi _1}f) \big ) (x,\xi ) = (\phi _3,\phi _1)_{L^2} \cdot (V_{\phi _2}f)(x,\xi ), \end{aligned} $$
(28)

which is some sort of reproducing kernel of short-time Fourier transforms in the background of ∗V.

2.3 Gelfand-Shilov Spaces

Before defining the Gelfand-Shilov spaces, we recall that the Schwartz space \(\mathcal S(\mathbb R^{d})\) consists of all (complex-valued) smooth functions \(f\in C^\infty (\mathbb R^{d})\) such that

$$\displaystyle \begin{aligned} \sup _{x\in \mathbb R^{d}}\big ( |x^\beta \partial ^\alpha f(x)| \big ) \le C_{\alpha ,\beta}, \end{aligned} $$
(29)

for some constants C α,β > 0, which only depend on the multi-indices \(\alpha ,\beta \in \mathbb N^{d}\). The Schwartz space possess several convenient properties, and is heavily used in mathematics, science and technology. For example, the Schwartz space is invariant under Fourier transformation. By duality the same holds true for its (L 2-)dual \(\mathcal S^{\prime }(\mathbb R^{d})\), the set of tempered distributions on \(\mathbb R^{d}\).

On the other hand, we observe that there are no conditions on the growths of the constants C α,β with respect to \(\alpha ,\beta \in \mathbb N^{d}\). This implies that in the context of the spaces \(\mathcal S(\mathbb R^{d})\) and \(\mathcal S^{\prime }(\mathbb R^{d})\), it is almost impossible to investigate important properties like analyticity or related regularity properties which are stronger than pure smoothness. For investigating such stronger regularity properties, we need to modify \(\mathcal S(\mathbb R^{d})\) and the estimate (29) by imposing suitable growth conditions on the constants C α,β. This leads to the definition of Gelfand-Shilov spaces, [26, 40].

We only discuss Fourier invariant Gelfand-Shilov spaces and their properties. Let \(0<s\in \mathbb R\) be fixed. We have two different types of Gelfand-Shilov spaces. The Gelfand-Shilov space \(\mathcal S_{s}(\mathbb R^{d})\) of Roumieu type with parameter s > 0 consists of all \(f\in C^\infty (\mathbb R^{d})\) such that

$$\displaystyle \begin{aligned} \sup _{x\in \mathbb R^{d}}\big ( |x^\beta \partial ^\alpha f(x)| \big ) \le Ch^{|\alpha + \beta |}(\alpha ! \beta !)^s, \end{aligned} $$
(30)

for some constants C, h > 0. In the same way, the Gelfand-Shilov space \(\Sigma _{s}(\mathbb R^{d})\) of Beurling type with parameter s > 0 consists of all \(f\in C^\infty (\mathbb R^{d})\) such that for every h > 0, there is a constant C = C h > 0 such that (30) holds. Hence, in comparison with the definition of Schwartz functions, we have limited ourself to constants C α,β in (29) which are not allowed to grow faster than those of the form

$$\displaystyle \begin{aligned} Ch^{|\alpha + \beta |}(\alpha ! \beta !)^s \end{aligned}$$

when dealing with Gelfand-Shilov spaces.

It can be proved that \(\mathcal S _s(\mathbb R^{d})\) and \(\Sigma _t(\mathbb R^{d})\) are dense in \(\mathcal S(\mathbb R^{d})\) when \(s\ge \frac 12\) and \(t>\frac 12\). We call such s and t admissible. On the other hand, for the other choices of s and t we have

$$\displaystyle \begin{aligned} \mathcal S _s(\mathbb R^{d})=\Sigma _t(\mathbb R^{d})=\{0\} , \quad \text{when}\quad s<\frac 12 ,\ t\le \frac 12. \end{aligned}$$

One has that \(\mathcal S _1(\mathbb R^{d})\) consists of real analytic functions, and that \(\Sigma _1(\mathbb R^{d})\) consists of smooth functions on \(\mathbb R^{d}\) which are extendable to entire functions on \(\mathbb C^{d}\). The topologies of \(\mathcal S_{s}(\mathbb R^{d})\) and \(\Sigma _{s}(\mathbb R^{d})\) are defined by the semi-norms

$$\displaystyle \begin{aligned} \Vert f\Vert _{\mathcal S_{s,h}}\equiv \sup \frac {|x^\beta \partial ^\alpha f(x)|}{h^{|\alpha + \beta |}(\alpha ! \beta !)^s}. \end{aligned} $$
(31)

Here the supremum should be taken over all \(\alpha ,\beta \in \mathbb N^{d}\) and \(x\in \mathbb R^{d}\). We equip \(\mathcal S_{s}(\mathbb R^{d})\) and \(\Sigma _{s}(\mathbb R^{d})\) by the canonical inductive limit topology and projective limit topology, respectively, with respect to h > 0, which are induced by the semi-norms in (31).

Let \(\mathcal S_{s,h}(\mathbb R^{d})\) be the Banach space which consists of all \(f\in C^\infty (\mathbb R^{d})\) such that \(\Vert f\Vert _{\mathcal S_{s,h}}\) in (31) is finite, and let \(\mathcal S_{s,h}^{\prime }(\mathbb R^{d})\) be the (L 2-)dual of \(\mathcal S_{s,h}(\mathbb R^{d})\). If \(s\ge \frac 12\), then the Gelfand-Shilov distribution space \(\mathcal S_{s}^{\prime }(\mathbb R^{d})\) of Roumieu type is the projective limit of \(\mathcal S_{s,h}^{\prime }(\mathbb R^{d})\) with respect to h > 0. If instead \(s> \frac 12\), then the Gelfand-Shilov distribution space \(\Sigma _{s}^{\prime }(\mathbb R^{d})\) of Beurling type is the inductive limit of \(\mathcal S_{s,h}^{\prime }(\mathbb R^{d})\) with respect to h > 0. Consequently, for admissible s we have

$$\displaystyle \begin{aligned} \mathcal S_{s}^{\prime}(\mathbb R^{d}) = \bigcap _{h>0} \mathcal S_{s,h}^{\prime}(\mathbb R^{d}) \quad \text{and}\quad \Sigma _{s}^{\prime}(\mathbb R^{d}) = \bigcup _{h>0} \mathcal S_{s,h}^{\prime}(\mathbb R^{d}). \end{aligned}$$

It can be proved that \(\mathcal S_{s}^{\prime }(\mathbb R^{d})\) and \(\Sigma _{s}^{\prime }(\mathbb R^{d})\) are the (strong) duals to \(\mathcal S_{s}(\mathbb R^{d})\) and \(\Sigma _{s}(\mathbb R^{d})\), respectively.

We have the following embeddings and density properties for Gelfand-Shilov and Schwartz spaces

$$\displaystyle \begin{aligned} \begin{alignedat}{3} \mathcal S _{s}(\mathbb R^{d}) &\hookrightarrow & \Sigma _{t}(\mathbb R^{d}) &\hookrightarrow & \mathcal S _t(\mathbb R^{d}) &\hookrightarrow \mathcal S(\mathbb R^{d}), \\ {} \mathcal S^{\prime}(\mathbb R^{d}) &\hookrightarrow & \mathcal S _t^{\prime} (\mathbb R^{d}) &\hookrightarrow & \Sigma _{t}^{\prime}(\mathbb R^{d}) &\hookrightarrow \mathcal S _{s}^{\prime}(\mathbb R^{d}), \qquad t>s\ge \frac 12, \end{alignedat} \end{aligned} $$
(32)

with dense embeddings. Here AB means that the topological spaces A and B satisfy A ⊆ B with continuous embeddings.

The Fourier transform possess convenient mapping properties on Gelfand-Shilov spaces and their distribution spaces. In fact, the Fourier transform extends uniquely to homeomorphisms on \(\mathcal S _s^{\prime }(\mathbb R^{d})\) and on \(\Sigma _s^{\prime }(\mathbb R^{d})\) for admissible s. Furthermore, \(\mathcal F\) restricts to homeomorphisms on \(\mathcal S _s(\mathbb R^{d})\) and on \(\Sigma _s(\mathbb R^{d})\).

One of the most important characterizations of Gelfand-Shilov spaces is performed in terms of estimates of the functions and their Fourier transforms. More precisely, in [8, 15] it is proved that if \(f\in \mathcal S^{\prime }(\mathbb R^{d})\) and s > 0, then \(f\in \mathcal S _s(\mathbb R^{d})\) (\(f\in \Sigma _s(\mathbb R^{d})\)), if and only if

$$\displaystyle \begin{aligned} |f(x)|\lesssim e^{-r|x|{}^{\frac 1s}} \quad \text{and}\quad |\widehat f(\xi )|\lesssim e^{-r|\xi |{}^{\frac 1s}}, \end{aligned} $$
(33)

for some r > 0 (for every r > 0). Here \(g_1 \lesssim g_2\) means that g 1(θ) ≤ c ⋅ g 2(θ) holds uniformly for all θ in the intersection of the domains of g 1 and g 2 and for some constant c > 0, and we write g 1 ≍ g 2 when \(g_1\lesssim g_2 \lesssim g_1\).

The analysis in [8, 15] can also be applied on the Schwartz space, from which it follows that an element \(f\in \mathcal S^{\prime }(\mathbb R^{d})\) belongs to \(\mathcal S(\mathbb R^{d})\), if and only if

$$\displaystyle \begin{aligned} |f(x)|\lesssim \langle x\rangle ^{-N} \quad \text{and}\quad |\widehat f(\xi )|\lesssim \langle \xi\rangle ^{-N}, \end{aligned} $$
(34)

for every N ≥ 0. Here and in what follows we let

$$\displaystyle \begin{aligned} \langle x\rangle = (1+|x|{}^2)^{\frac 12}. \end{aligned}$$

Remark 2.3

Several properties in Sects. 2.12.3 in the background of \(\mathcal S(\mathbb R^{d})\) and \(\mathcal S^{\prime }(\mathbb R^{d})\) also hold for the Gelfand-Shilov spaces and their distribution spaces. Let \(s\ge \frac 12\). By similar arguments which lead to Proposition 2.1 and (13), it follows that

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} (f,\phi ) &\mapsto V_\phi f & &:\, & &\mathcal S _s(\mathbb R^{d}) \times \mathcal S _s(\mathbb R^{d}) & &\to \mathcal S _s(\mathbb R^{2d}) {} \end{array}\end{aligned} $$
(35)

is continuous, which extends uniquely to continuous mappings

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} (f,\phi ) &\mapsto V_\phi f & &:\, & &\mathcal S _s^{\prime}(\mathbb R^{d}) \times \mathcal S _s(\mathbb R^{d}) & &\to \mathcal S _s^{\prime}(\mathbb R^{2d})\cap C^\infty (\mathbb R^{2d}) {} \end{array}\end{aligned} $$
(36)

and

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} (f,\phi ) &\mapsto V_\phi f & &:\, & &\mathcal S _s^{\prime}(\mathbb R^{d}) \times \mathcal S _s^{\prime}(\mathbb R^{d}) & &\to \mathcal S _s^{\prime}(\mathbb R^{2d}). {} \end{array}\end{aligned} $$
(37)

It follows that (14) makes sense after each \(\mathcal S\) in (15) are replaced by \(\mathcal S _s\). Let \(\phi \in \mathcal S _s(\mathbb R^{d})\setminus \{ 0 \}\) be fixed. Then by similar arguments which lead to (19) give that the mappings

$$\displaystyle \begin{aligned} V_\phi ^* : \mathcal S _s(\mathbb R^{2d}) \to \mathcal S _s(\mathbb R^{d}), \qquad V_\phi ^* : \mathcal S _s^{\prime}(\mathbb R^{2d}) \to \mathcal S _s^{\prime}(\mathbb R^{d}) \end{aligned} $$
((19)′)

are continuous. For P ϕ in (20) we have that (21) still holds true and that (22) can be completed with

$$\displaystyle \begin{aligned} P_\phi (\mathcal S _s(\mathbb R^{2d})) = V_\phi (\mathcal S _s(\mathbb R^{d})) \quad \text{and}\quad P_\phi (\mathcal S _s^{\prime}(\mathbb R^{2d})) = V_\phi (\mathcal S _s^{\prime}(\mathbb R^{d})). \end{aligned} $$
(38)

We also have that the twisted convolution in (26) is continuous from \(\mathcal S _s(\mathbb R^{2d})\times \mathcal S _s(\mathbb R^{2d})\) to \(\mathcal S _s(\mathbb R^{2d})\) and uniquely extendable to a continuous map \(\mathcal S _s(\mathbb R^{2d})\times \mathcal S _s^{\prime }(\mathbb R^{2d})\) or \(\mathcal S _s^{\prime }(\mathbb R^{2d})\times \mathcal S _s(\mathbb R^{2d})\) to \(\mathcal S _s^{\prime }(\mathbb R^{2d})\), and that the formulae (25)–(28) still hold true after each \(\mathcal S\) is replaced by \(\mathcal S _s\) in the attached assumptions.

If instead \(s>\frac 12\), then similar facts hold true with Σs in place of \(\mathcal S _s\) above, at each occurrence.

Remark 2.4

In similar ways as characterizing Gelfand-Shilov spaces in terms of Fourier estimates (see (33)), we may also use the short-time Fourier transform to perform similar characterizations. Moreover, the short-time Fourier transform can in addition be used to characterize spaces of Gelfand-Shilov distributions.

In fact, let \(\phi \in \mathcal S _s(\mathbb R^{d})\setminus \{ 0 \}\) (\(\phi \in \Sigma _s(\mathbb R^{d})\setminus \{ 0 \}\)) be fixed and let f be a Gelfand-Shilov distribution on \(\mathbb R^{d}\). Then the following is true:

  1. 1.

    \(f\in \mathcal S _s(\mathbb R^{d})\) (\(f\in \Sigma _s(\mathbb R^{d})\)), if and only if

    $$\displaystyle \begin{aligned} |V_\phi f(x,\xi )|\lesssim e^{-r(|x|{}^{\frac 1s} +|\xi |{}^{\frac 1s})} \end{aligned} $$
    (39)

    for some r > 0 (for every r > 0);

  2. 2.

    \(f\in \mathcal S _s^{\prime }(\mathbb R^{d})\) (\(f\in \Sigma _s^{\prime }(\mathbb R^{d})\)), if and only if

    $$\displaystyle \begin{aligned} |V_\phi f(x,\xi )|\lesssim e^{r(|x|{}^{\frac 1s} +|\xi |{}^{\frac 1s})} \end{aligned} $$
    (40)

    for every r > 0 (for some r > 0).

We refer to [31, Theorem 2.7] for the characterization 1. concerning Gelfand-Shilov functions and to [51, Proposition 2.2]) for the characterization 2. concerning Gelfand-Shilov distributions.

2.4 Weight Functions

A weight or weight function on \(\mathbb R^{d}\) is a positive function \(\omega \in L^\infty _{loc}(\mathbb R^{d})\) such that \(1/\omega \in L^\infty _{loc}(\mathbb R^{d})\). The weight ω is called moderate, if there is a positive weight v on \(\mathbb R^{d}\) and a constant C ≥ 1 such that

$$\displaystyle \begin{aligned} \omega (x+y) \le C\omega (x)v(y),\qquad x,y\in \mathbb R^{d}. \end{aligned} $$
(41)

If ω and v are weights on \(\mathbb R^{d}\) such that (41) holds, then ω is also called v-moderate. We note that (41) implies that ω fulfills the estimates

$$\displaystyle \begin{aligned} C^{-1}v(-x)^{-1}\le \omega (x)\le C v(x),\quad x\in \mathbb R^{d}. \end{aligned} $$
(42)

We let \(\mathcal P _E(\mathbb R^{d})\) be the set of all moderate weights on \(\mathbb R^{d}\).

We say that v is submultiplicative if

$$\displaystyle \begin{aligned} v(x+y) \le v(x)v(y) \quad \text{and}\quad v(-x)=v(x),\qquad x,y\in \mathbb R^{d}. \end{aligned} $$
(43)

We observe that if \(v\in \mathcal P _E(\mathbb R^{d})\) is even and satisfies

$$\displaystyle \begin{aligned} v(x+y)\le Cv(x)v(y), \qquad x,y\in \mathbb R^{d}, \end{aligned} $$
(44)

for some constant C > 0, then for v 0 = C 1∕2 v, one has that \(v_0\in \mathcal P _E(\mathbb R^{d})\) is submultiplicative and v ≍ v 0 (see e.g. [17, 19, 28]).

We also recall from [29] that if v is positive and locally bounded and satisfies (44), then \(v(x)\le C_0e^{r_0|x|}\) for some positive constants C 0 and r 0. In fact, if \(x\in \mathbb R^{d}\),

$$\displaystyle \begin{aligned} r=\sup _{|x|\le 1}\log v(x),\quad c=\log C \end{aligned}$$

and n is an integer such that n − 1 ≤|x|≤ n, then (44) gives

$$\displaystyle \begin{aligned} v(x) = v(n\cdot (x/n))\le C^nv(x/n)^n \le C^ne^{rn} = e^{(r+c)n} \le e^{(r+c)(|x|+1)}, \end{aligned}$$

which gives the statement.

Therefore, if v is a submultiplicative weight, then

$$\displaystyle \begin{aligned} v(x)\lesssim e^{r|x|},\qquad x\in \mathbb R^{d}, \end{aligned} $$
(45)

for some r ≥ 0. Hence, if \(\omega \in \mathcal P _E(\mathbb R^{d})\), then (41) and (45) imply

$$\displaystyle \begin{aligned} \omega (x+y) \lesssim \omega (x) e^{r|y|},\quad x,y\in \mathbb R^{d} \end{aligned} $$
(46)

for some r > 0. In particular, (42) shows that for any \(\omega _0\in \mathcal P_E(\mathbb R^{d})\), there is a constant r > 0 such that

$$\displaystyle \begin{aligned} e^{-r|x|}\lesssim \omega _0(x)\lesssim e^{r|x|},\quad x\in \mathbb R^{d}. \end{aligned}$$

If (41) holds, then there is a smallest positive even function v 0 such that (41) holds with C = 1. We remark that this v 0 is given by

$$\displaystyle \begin{aligned} v_0(x) = \sup _{y\in \mathbb R^{d}} \left ( \frac {\omega (x+y)}{\omega (y)}, \frac {\omega (-x+y)}{\omega (y)} \right ) , \end{aligned}$$

and is submultiplicative (see e.g. [19, 27, 49]). Consequently, if ω is a moderate weight, then it is also moderated by a submultiplicative weight. In the sequel, v and v j for j ≥ 0, always stand for submultiplicative weights if nothing else is stated.

We also remark that in the literature it is common to define submultiplicative weights as (43) should hold, without the condition v(−x) = v(x), i.e. that v does not have to be even (cf. e.g. [17, 19, 25, 28]). However, in the sequel it is convenient for us to include this property in the definition.

There are several subclasses of \(\mathcal P _E(\mathbb R^{d})\) which are interesting for different reasons. Though our results later on are formulated in background of weights in \(\mathcal P _E(\mathbb R^{d})\), we here mention some subclasses which especially appear in time-frequency analysis. First we observe the class \(\mathcal P ^{0} _E(\mathbb R^{d})\), which consists of all \(\omega \in \mathcal P _E(\mathbb R^{d})\) such that (46) holds for every r > 0.

The class \(\mathcal P _E^0(\mathbb R^{d})\) is important when dealing with spectral invariance for matrix or convolution operators on \(\ell ^2(\mathbb Z^{d})\) (see e.g. [30]). If \(v\in \mathcal P _E(\mathbb R^{d})\) is submultiplicative, then \(v\in \mathcal P _E^0(\mathbb R^{d})\), if and only if

$$\displaystyle \begin{aligned} \lim _{n\to \infty} v(nx)^{\frac 1n} = 1 \end{aligned} $$
(47)

(see e.g. [23]). The condition (47) is equivalent to

$$\displaystyle \begin{aligned} \lim _{n\to \infty} \frac {\log (v(nx))}n = 0, \end{aligned} $$
((47)′)

and is usually called the GRS condition, or Gelfand-Raikov-Shilov condition.

A more restrictive condition on v compared to (47) is given by the Beurling-Domar condition

$$\displaystyle \begin{aligned} \sum _{n=1}^\infty \frac {\log (v(nx))}{n^2} <\infty . \end{aligned} $$
(48)

This condition is strongly linked to non quasi-analytic classes which contain non-trivial compactly supported elements (see e.g. [29]). Any subexponential submultiplicative weight satisfies the Beurling-Domar condition. That is, suppose that θ ∈ (0, 1) and that \(v(x)=e^{r|x|{ }^\theta }\), \(x\in \mathbb R^{d}\), then (48) is fulfilled. We let \(\mathcal P _{\operatorname {BD}}(\mathbb R^{d})\) be the set of all weights which are moderated by submultiplicative weights which satisfy the Beurling-Domar condition.

Finally we let \(\mathcal P (\mathbb R^{d})\) be the set of all weights on \(\mathbb R^{d}\) which are moderated by polynomially bounded functions. That is, \(\omega \in \mathcal P (\mathbb R^{d})\), if and only if there are positive constants r and C such that

$$\displaystyle \begin{aligned} \omega (x+y) \le C\omega (x)(1+|y|)^r,\qquad x,y\in \mathbb R^{d}. \end{aligned}$$

Here we observe that v(x) = (1 + |x|)r is submultiplicative.

Among these weight classes we have

$$\displaystyle \begin{aligned} \mathcal P (\mathbb R^{d})\subsetneq \mathcal P _{\operatorname{BD}}(\mathbb R^{d}) \subsetneq \mathcal P _E^0(\mathbb R^{d}) \subsetneq \mathcal P _E(\mathbb R^{d}). \end{aligned} $$
(49)

In fact, it is clear that the ordering in (49) holds. On the other hand, if r > 0 and θ ∈ (0, 1), then due to

$$\displaystyle \begin{aligned} e^{r|x|{}^\theta } &\in \mathcal P _{\operatorname{BD}}(\mathbb R^{d})\setminus \mathcal P(\mathbb R^{d}), \\ {} e^{r|x|/\log (e+|x|)} &\in \mathcal P _E^0(\mathbb R^{d})\setminus \mathcal P _{\operatorname{BD}}(\mathbb R^{d}), \\ {} \text{and}\quad e^{r|x|} &\in \mathcal P _E(\mathbb R^{d})\setminus \mathcal P _E^0(\mathbb R^{d}), \end{aligned} $$
(50)

it also follows that the inclusions in (49) are strict.

We refer to [16, 28, 29, 49] for more facts about weights in time-frequency analysis.

2.5 Mixed Norm Spaces of Lebesgue Type

For every p, q ∈ (0, ] and weight ω on \(\mathbb R^{2d}\), we set

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Vert F\Vert _{L_{(\omega )}^{p,q}(\mathbb R^{2d})} &\equiv \Vert G_{F,\omega ,p}\Vert _{L^q(\mathbb R^{d})}, & \quad &\text{where} & \quad G_{F,\omega ,p}(\xi ) &= \Vert F( \cdot \, ,\xi )\omega ( \cdot \, ,\xi )\Vert _{L^p(\mathbb R^{d})}\end{array}\end{aligned} $$

and

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Vert F\Vert _{L^{p,q}_{*,(\omega )}(\mathbb R^{2d})} &\equiv \Vert H_{F,\omega ,q}\Vert _{L^p(\mathbb R^{d})}, & \quad &\text{where} & \quad H_{F,\omega ,q}(x) &= \Vert F(x,\, \cdot \, )\omega (x,\, \cdot \, )\Vert _{L^q(\mathbb R^{d})}, \end{array}\end{aligned} $$

when F is (complex-valued) measurable function on \(\mathbb R^{2d}\). Then \(L^{p,q}_{(\omega )}(\mathbb R^{2d})\) (\(L^{p,q}_{*,(\omega )}(\mathbb R^{2d})\)) consists of all measurable functions F such that \(\Vert F\Vert _{L_{(\omega )}^{p,q}}<\infty \) (\(\Vert F\Vert _{L^{p,q}_{*,(\omega )}}<\infty \)).

In similar ways, let Ω1, Ω2 be discrete sets, ω be a positive function on Ω1 × Ω2 and \(\ell _0^{\prime }(\Omega _1\times \Omega _2)\) be the set of all formal (complex-valued) sequences \(c=\{ c(j,k)\} _{j\in \Omega _1,k\in \Omega _2}\). Then the discrete Lebesgue spaces, i.e. the Lebesgue sequence spaces

$$\displaystyle \begin{aligned} \ell ^{p,q}_{(\omega )}(\Omega _1\times \Omega _2) \quad \text{and}\quad \ell ^{p,q}_{*,(\omega )}(\Omega _1\times \Omega _2) \end{aligned}$$

of mixed (quasi-)norm types consist of all \(c\in \ell _0^{\prime }(\Omega _1\times \Omega _2)\) such that \(\Vert c\Vert _{\ell _{(\omega )}^{p,q}(\Omega _1\times \Omega _2)}<\infty \) respectively \(\Vert c\Vert _{\ell _{*,(\omega )}^{p,q}(\Omega _1\times \Omega _2)}<\infty \). Here

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Vert c\Vert _{\ell _{(\omega )}^{p,q}(\Omega _1\times \Omega _2)} &\equiv \Vert G_{c,\omega ,p}\Vert _{\ell ^q(\Omega _2)}, & \quad &\text{where} & \quad G_{c,\omega, p}(k) &= \Vert F( \cdot \, ,k )\omega ( \cdot \, ,k)\Vert _{\ell ^p(\Omega _1)} \end{array}\end{aligned} $$

and

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Vert c\Vert _{\ell ^{p,q}_{*,(\omega )}(\Omega _1\times \Omega _2)} &\equiv \Vert H_{c,\omega ,q}\Vert _{\ell ^p(\Omega _1)}, & \quad &\text{where} & \quad H_{c,\omega, q}(j) &= \Vert c(j,\, \cdot \, )\omega (j,\, \cdot \, )\Vert _{\ell ^q(\Omega _2)}, \end{array}\end{aligned} $$

when \(c\in \ell _0^{\prime }(\Omega _1\times \Omega _2)\).

2.6 Convolutions and Multiplications for Discrete Lebesgue Spaces

Next we discuss extended Hölder and Young relations for multiplications and convolutions on discrete Lebesgue spaces. The Hölder and Young conditions on Lebesgue exponent are then

$$\displaystyle \begin{aligned} \frac 1{q_0} &\le \frac 1{q_1}+\frac 1{q_2}, {} \end{aligned} $$
(51)

respectively

$$\displaystyle \begin{aligned} \frac 1{p_0} &\le \frac 1{p_1}+\frac 1{p_2} - \max \left ( 1,\frac 1{p_1},\frac 1{p_2}\right ). {} \end{aligned} $$
(52)

Notice that, when p 1, p 2 ∈ (0, 1), then (52) becomes \( p_0 \geq \max \{ p_1, p_2 \} \), while for p 1, p 2 ≥ 1 it reduces to the common Young condition

$$\displaystyle \begin{aligned} 1 + \frac 1{p_0} \le \frac 1{p_1}+\frac 1{p_2}. \end{aligned}$$

The conditions on the weight functions are

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \omega _0(j) &\le \omega _1(j)\omega _2(j), & \qquad j &\in \Lambda , {} \end{array}\end{aligned} $$
(53)

respectively

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \omega _0(j_1+j_2) &\le \omega _1(j_1)\omega _2(j_2), & \qquad j_1,j_2 &\in \Lambda , {} \end{array}\end{aligned} $$
(54)

where Λ is a lattice of the form

$$\displaystyle \begin{aligned} \Lambda = \{ \, n_1e_1+\cdots +n_de_d\, ;\, (n_1,\dots ,n_d)\in \mathbb Z^{d}\, \} , \end{aligned}$$

where e 1, …e d is a basis for \(\mathbb R^{d}\).

Proposition 2.5

Let p j, q j ∈ (0, ], j = 0, 1, 2, be such that (51) and (52) hold, let \(\Lambda \subseteq \mathbb R^{d}\) be a lattice and let ω j be weights on Λ, j = 0, 1, 2. Then the following is true:

  1. 1.

    if (53) holds, then the map (a 1, a 2)↦a 1 ⋅ a 2 from ℓ 0( Λ) × ℓ 0( Λ) to ℓ 0( Λ) extends uniquely to a continuous map from \(\ell ^{q_1}_{(\omega _1)}(\Lambda ) \times \ell ^{q_2}_{(\omega _2)}(\Lambda )\) to \(\ell ^{q_0}_{(\omega _0)}(\Lambda )\) , and

    $$\displaystyle \begin{aligned} \Vert a_1\cdot a_2\Vert _{\ell ^{q_0}_{(\omega _0)}} \le \Vert a_1\Vert _{\ell ^{q_1}_{(\omega _1)}} \Vert a_2\Vert _{\ell ^{q_2}_{(\omega _2)}}, \qquad a_j\in \ell ^{q_j}_{(\omega _j)}(\Lambda ),\ j=1,2 \text ; \end{aligned} $$
    (55)
  2. 2.

    if (54) holds, then the map (a 1, a 2)↦a 1 ∗ a 2 from ℓ 0( Λ) × ℓ 0( Λ) to ℓ 0( Λ) extends uniquely to a continuous map from \(\ell ^{p_1}_{(\omega _1)}(\Lambda ) \times \ell ^{p_2}_{(\omega _2)}(\Lambda )\) to \(\ell ^{p_0}_{(\omega _0)}(\Lambda )\) , and

    $$\displaystyle \begin{aligned} \Vert a_1*a_2\Vert _{\ell ^{p_0}_{(\omega _0)}} \le \Vert a_1\Vert _{\ell ^{p_1}_{(\omega _1)}}\Vert a_2\Vert _{\ell ^{p_2}_{(\omega _2)}}, \qquad a_j\in \ell ^{p_j}_{(\omega _j)}(\Lambda ),\ j=1 ,2 \text . \end{aligned} $$
    (56)

The assertion 1. in Proposition 2.5 is the standard Hölder’s inequality for discrete Lebesgue spaces. The assertion 2. in that proposition is the usual Young’s inequality for Lebesgue spaces on lattices in the case when p 0, p 1, p 2 ∈ [1, ]. A proof of Proposition 2.5 is given in Appendix A in [54].

3 Modulation Spaces, Multiplications and Convolutions

In this section we introduce modulation spaces, and recall their basic properties, in particular in the context of Gelfand-Shilov spaces. Notice that we permit the Lebesgue exponents to belong to the full interval (0, ] instead of the most common choice [1, ], and general moderate weights which may have a (sub)exponential growth. Here we also recall some facts on Gabor expansions for modulation spaces.

Then we deduce multiplication and convolution estimates on modulation spaces. There are several approaches to multiplication and convolution in the case when the involved Lebesgue exponents belong to [1, ] (see [9, 17, 19, 32, 43, 48]). Here we consider the case when these exponents belong to (0, ) (see also [1, 2, 25, 41, 42, 50]). In addition, and in order to keep the survey style of our exposition, we focus on the bilinear case, and refer to [54] for extension of these results to multi-linear products.

3.1 Modulation Spaces

The (classical) modulation spaces, essentially introduced in [17] by Feichtinger are given in the following. (See e.g. [18] for definition of more general modulation spaces.)

Definition 3.1

Let p, q ∈ (0, ], \(\omega \in \mathcal P _E(\mathbb R^{2d})\) and \(\phi \in \Sigma _1 (\mathbb R^{d})\setminus \{ 0 \}\).

  1. 1.

    The modulation space \(M^{p,q}_{(\omega )}(\mathbb R^{d})\) consists of all \(f\in \Sigma _1 ^{\prime }(\mathbb R^{d})\) such that

    $$\displaystyle \begin{aligned} \Vert f\Vert _{M^{p,q}_{(\omega )}}\equiv \Vert V_\phi f\Vert _{L^{p,q}_{(\omega )}} \end{aligned}$$

    is finite. The topology of \(M^{p,q}_{(\omega )}(\mathbb R^{d})\) is defined by the (quasi-)norm \(\Vert \, \cdot \,\Vert _{M^{p,q}_{(\omega )}}\);

  2. 2.

    The modulation space (of Wiener amalgam type) \(W^{p,q}_{(\omega )}(\mathbb R^{d})\) consists of all \(f\in \Sigma _1 ^{\prime }(\mathbb R^{d})\) such that

    $$\displaystyle \begin{aligned} \Vert f\Vert _{W^{p,q}_{(\omega )}}\equiv \Vert V_\phi f\Vert _{L^{p,q}_{*,(\omega )}} \end{aligned}$$

    is finite. The topology of \(W^{p,q}_{(\omega )}(\mathbb R^{d})\) is defined by the (quasi-)norm \(\Vert \, \cdot \,\Vert _{W^{p,q}_{(\omega )}}\).

For convenience we set \(M^{p,q}=M^{p,q}_{(\omega )}\) and \(W^{p,q}=W^{p,q}_{(\omega )}\) when the weight ω is trivial, i.e. when ω(x, ξ) = 1 for every \(x,\xi \in \mathbb R^{d}\). We also set

$$\displaystyle \begin{aligned} M^p_{(\omega )} \equiv M^{p,p}_{(\omega )} \ (\ = W^{p,p}_{(\omega )}\ ) \quad \text{and}\quad M^p \equiv M^{p,p} \ (\ = W^{p,p}\ ) . \end{aligned}$$

Remark 3.2

Modulation spaces possess several convenient properties. Let p, q ∈ (0, ], \(\omega \in \mathcal P _E(\mathbb R^{2d})\) and \(\phi \in \Sigma _1(\mathbb R^{d})\setminus \{ 0 \}\). Then the following is true (see [17,18,19,20, 25, 28] and their analyses for verifications):

  • the definitions of \(M_{(\omega )}^{p,q}(\mathbb R^{d})\) and \(W_{(\omega )}^{p,q}(\mathbb R^{d})\) are independent of the choices of \(\phi \in \Sigma _1 (\mathbb R^{d})\setminus \{ 0 \}\), and different choices give rise to equivalent quasi-norms;

  • the spaces \(M^{p,q}_{(\omega )}(\mathbb R^{d})\) and \(W^{p,q}_{(\omega )}(\mathbb R^{d})\) are quasi-Banach spaces which increase with p and q, and decrease with ω. If in addition p, q ≥ 1, then they are Banach spaces;

  • If p, q ≥ 1, then the \(L^2(\mathbb R^{d})\) scalar product, \(( \cdot \, ,\, \cdot \, )_{L^2(\mathbb R^{d})}\), on \(\Sigma _1 (\mathbb R^{d})\times \Sigma _1 (\mathbb R^{d})\) is uniquely extendable to dualities between \(M^{p,q}_{(\omega )}(\mathbb R^{d})\) and \(M^{p^{\prime },q^{\prime }}_{(1/\omega )}(\mathbb R^{d})\), and between \(W^{p,q}_{(\omega )}(\mathbb R^{d})\) and \(W^{p^{\prime },q^{\prime }}_{(1/\omega )}(\mathbb R^{d})\). If in addition p, q < , then the dual spaces of \(M^{p,q}_{(\omega )}(\mathbb R^{d})\) and \(W^{p,q}_{(\omega )}(\mathbb R^{d})\) can be identified with \(M^{p^{\prime },q^{\prime }}_{(1/\omega )}(\mathbb R^{d})\) respectively \(W^{p^{\prime },q^{\prime }}_{(1/\omega )}(\mathbb R^{d})\), through the form \(( \cdot \, ,\, \cdot \, )_{L^2(\mathbb R^{d})}\);

  • if ω 0(x, ξ) = ω(−ξ, x), then \(\mathcal F\) on \(\Sigma _1 ^{\prime }(\mathbb R^{d})\) restricts to a homeomorphism from \(M^{p,q}_{(\omega )}(\mathbb R^{d})\) to \(W^{q,p}_{(\omega _0)}(\mathbb R^{d})\).

  • The inclusions

    $$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Sigma _1 (\mathbb R^{d}) &\subseteq & M^{p,q}_{(\omega )}(\mathbb R^{d}),W^{p,q}_{(\omega )}(\mathbb R^{d}) &\subseteq &\Sigma _1 ^{\prime}(\mathbb R^{d}) \quad &\text{when} & \quad \omega \in \mathcal P _E(\mathbb R^{2d}), {} \end{array}\end{aligned} $$
    (57)
    $$\displaystyle \begin{aligned}\begin{array}{r*{20}l}[1ex] \mathcal S _1 (\mathbb R^{d}) &\subseteq & M^{p,q}_{(\omega )}(\mathbb R^{d}),W^{p,q}_{(\omega )}(\mathbb R^{d}) &\subseteq &\mathcal S _1 ^{\prime}(\mathbb R^{d}) \quad &\text{when} & \quad \omega \in \mathcal P _E^0(\mathbb R^{2d}) {} \end{array}\end{aligned} $$
    (58)

    and

    $$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \mathcal S(\mathbb R^{d}) &\subseteq & M^{p,q}_{(\omega )}(\mathbb R^{d}),W^{p,q}_{(\omega )}(\mathbb R^{d}) &\subseteq & \, \mathcal S^{\prime}(\mathbb R^{d}) \quad &\text{when} & \quad \omega \in \mathcal P (\mathbb R^{2d}) {} \end{array}\end{aligned} $$
    (59)

    are continuous. If in addition p, q < , then these inclusions are dense.

We recall from [49] that the embeddings (57)–(59), are essentially special cases of certain characterizations of the Schwartz space, Gelfand-Shilov spaces and their distribution spaces in terms of suitable unions and intersections of modulation spaces. In fact, let p, q ∈ (0, ] and s ≥ 1 be fixed and set

$$\displaystyle \begin{aligned} v_{r,t}(x,\xi ) = \begin{cases} e^{r(|x|{}^{\frac 1t}+|\xi |{}^{\frac 1t}))}, & t\in \mathbb R_+ \\ {} (1+|x|+|\xi |)^r, & t=\infty , \end{cases} \end{aligned} $$
(60)

where r > 0. Then

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Sigma _s(\mathbb R^{d}) &= \bigcap _{r>0}M^{p,q}_{(v_{r,s})}(\mathbb R^{d}) & &= \bigcap _{r>0}W^{p,q}_{(v_{r,s})}(\mathbb R^{d}), {} \end{array}\end{aligned} $$
(61)
$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \mathcal S _s(\mathbb R^{d}) &= \bigcup _{r>0}M^{p,q}_{(v_{r,s})}(\mathbb R^{d}) & &= \bigcup _{r>0}W^{p,q}_{(v_{r,s})}(\mathbb R^{d}), {} \end{array}\end{aligned} $$
(62)
$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \mathcal S(\mathbb R^{d}) &= \bigcap _{r>0}M^{p,q}_{(v_{r,\infty })}(\mathbb R^{d}) & &= \bigcap _{r>0}W^{p,q}_{(v_{r,\infty})}(\mathbb R^{d}), {} \end{array}\end{aligned} $$
(63)
$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \mathcal S^{\prime}(\mathbb R^{d}) &= \bigcup _{r>0}M^{p,q}_{(1/v_{r,\infty })}(\mathbb R^{d}) & &= \bigcup _{r>0}W^{p,q}_{(1/v_{r,\infty})}(\mathbb R^{d}), {} \end{array}\end{aligned} $$
(64)
$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \mathcal S _s^{\prime}(\mathbb R^{d}) &= \bigcap _{r>0}M^{p,q}_{(1/v_{r,s})}(\mathbb R^{d}) & &= \bigcap _{r>0}W^{p,q}_{(1/v_{r,s})}(\mathbb R^{d}) {} \end{array}\end{aligned} $$
(65)

and

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Sigma _s^{\prime}(\mathbb R^{d}) &= \bigcup _{r>0}M^{p,q}_{(1/v_{r,s})}(\mathbb R^{d}) & &= \bigcup _{r>0}W^{p,q}_{(1/v_{r,s})}(\mathbb R^{d}). {} \end{array}\end{aligned} $$
(66)

The topologies of the spaces on the left-hand sides of (61)–(66) are obtained by replacing each intersection by projective limit with respect to r > 0 and each union with inductive limit with respect to r > 0.

The relations (61)–(66) are essentially special cases of [49, Theorem 3.9], see also [31, 45, 46]. In order to be self-contained we here give a proof of (62).

Proof of (62)

Since

$$\displaystyle \begin{aligned} M^\infty _{(v_{2r,s})}(\mathbb R^{d}) \subseteq M^{p,q}_{(v_{r,s})}(\mathbb R^{d}),W^{p,q}_{(v_{r,s})}(\mathbb R^{d}) \subseteq M^\infty _{(v_{r,s})}(\mathbb R^{d}), \end{aligned}$$

it suffices to prove the result for p = q = . Let \(\phi \in \Sigma _1(\mathbb R^{d})\setminus \{ 0 \}\) be fixed. First suppose that

$$\displaystyle \begin{aligned} f\in M^\infty _{(v_{r,s})}(\mathbb R^{d})=W^\infty _{(v_{r,s})}(\mathbb R^{d}). \end{aligned}$$

Then it follows from the definition of modulation space norm that (39) holds for some r > 0. By Remark 2.4 it follows that \(f\in \mathcal S _s(\mathbb R^{d})\), and we have proved

$$\displaystyle \begin{aligned} \bigcup _{r>0}M^\infty _{(v_{r,s})}(\mathbb R^{d}) \subseteq \mathcal S _s(\mathbb R^{d}). \end{aligned} $$
(67)

Suppose instead that \(f\in \mathcal S _s(\mathbb R^{d})\). Then (39) holds for some r > 0, giving that \(f\in M^\infty _{(v_{r,s})}(\mathbb R^{d})\). Hence (67) holds with reversed inclusion, and the result follows. □

Example 3.3

Let p = q = 1 and ω = 1. Then \(M^{1,1} _{(\omega )} (\mathbb R^{d}) = M^{1} (\mathbb R^{d})\) is the Feichtinger algebra, probably the most prominent example of a modulation space. We refer to a recent survey [34] for a detailed account on \(M^{1} (\mathbb R^{d})\), and to [14, Lemma 11] for a list of its basic properties.

Familiar examples arise when p = q = 2 and ω = 1. Then \(M_{(\omega )} ^{2,2} (\mathbb R^{d}) =M^{2} (\mathbb R^{d}) = L^2 (\mathbb R^{d})\), and

$$\displaystyle \begin{aligned} M_{(\omega_s)} ^{2,2} (\mathbb R^{d}) = H^s (\mathbb R^{d}), \qquad s \in \mathbb{R}, {} \end{aligned}$$

where ω s(ξ) = 〈ξs, and \(H^s (\mathbb R^{d})\) is the Sobolev space (also known as the Bessel potential space) of distributions \(f \in \mathcal S^{\prime }(\mathbb R^{d}) \) such that

$$\displaystyle \begin{aligned} \|f\|{}_{H^s}^2 := \int_{\mathbb R^{d}} \langle \xi \rangle^{2s} | \widehat{f} (\xi)|{}^2 d \xi < \infty, {} \end{aligned}$$

cf. [28, Proposition 11.3.1]. Furthermore, if v s(x, ξ) = 〈(x, ξ)〉s, then \( M_{(v_{s})} ^{2,2} (\mathbb R^{d}) = Q_s (\mathbb R^{d})\), \(s \in \mathbb {R},\) [7, Lemma 2.3]. Here Q s denotes the Shubin-Sobolev space, [44].

Finally we remark that modulation spaces can be conveniently discretized in terms of Gabor expansions. In order for explaining some basic issues on this, in a similar way as in Subsection 1.5 in [54], we limit ourself to the case when the involved weights are moderated by subexponential functions. That is, we suppose that ω in \(M^{p,q}_{(\omega )}(\mathbb R^{d})\) satisfies

$$\displaystyle \begin{aligned} \omega (x+y,\xi +\eta ) \lesssim \omega (x,\xi )e^{r(|x|{}^{\frac 1s}+|\xi |{}^{\frac 1s})}, \end{aligned} $$
(68)

for some s > 1 and r > 0. We observe that this implies that

$$\displaystyle \begin{aligned} \Sigma _s(\mathbb R^{d}) \subseteq M^{p,q}_{(\omega )}(\mathbb R^{d}) \subseteq \Sigma _s^{\prime}(\mathbb R^{d}), \end{aligned} $$
(69)

in vew of (42), (61) and (66). For more general approaches we refer to [19, 27, 28, 42, 50].

Since s > 1, it follows from Sections 1.3 and 1.4 in [33] that there are \(\phi ,\psi \in \Sigma _s (\mathbb R^{d})\) with values in [0, 1] such that

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \operatorname{supp} \phi &\subseteq \Big [-\frac 34,\frac 34\Big ] ^d,& \quad \phi (x) &= 1& \quad &\text{when}& \quad x&\in \Big [-\frac 14,\frac 14\Big ]^d {} \end{array}\end{aligned} $$
(70)
$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \operatorname{supp} \psi &\subseteq [-1,1]^d,& \quad \psi (x) &= 1& \quad &\text{when}& \quad x&\in \Big [-\frac 34,\frac 34\Big ]^d {} \end{array}\end{aligned} $$
(71)

and

$$\displaystyle \begin{aligned} \sum _{j\in \mathbb Z^{d}}\phi ( \cdot \, -j) =1. \end{aligned} $$
(72)

Let \(f\in \Sigma _s^{\prime }(\mathbb R^{d})\). Then xf(x)ϕ(x − j) belongs to \(\Sigma _s^{\prime }(\mathbb R^{d})\) and is supported in \(j+[-\frac 34,\frac 34]^d\). Hence, by periodization it follows from Fourier analysis that

$$\displaystyle \begin{aligned} f(x)\phi (x-j) = \sum _{\iota \in \pi \mathbb Z^{d}}c(j,\iota )e^{i\langle x,\iota\rangle }, \qquad x\in j+[-1,1]^d, \end{aligned} $$
(73)

where

$$\displaystyle \begin{aligned} c(j,\iota ) = 2^{-d}(f,\phi ( \cdot \, -j)e^{i\langle \, \cdot \,,\iota\rangle }) = \left ( \frac \pi 2\right )^{\frac d2} V_\phi f(j,\iota ), \qquad j\in \mathbb Z^{d},\ \iota \in \pi \mathbb Z^{d} . \end{aligned}$$

Since ψ = 1 on the support of ϕ, (73) gives

$$\displaystyle \begin{aligned} f(x)\phi (x-j) = \left ( \frac \pi 2\right )^{\frac d2} \sum _{\iota \in \pi \mathbb Z^{d}}V_\phi f(j,\iota )\psi (x-j)e^{i\langle x,\iota\rangle }, \qquad x\in \mathbb R^{d}, \end{aligned} $$
((73)′)

By (72) it now follows that

$$\displaystyle \begin{aligned} f(x) &= \left ( \frac \pi 2\right )^{\frac d2} \sum _{(j,\iota )\in \Lambda}V_\phi f(j,\iota )\psi (x-j)e^{i\langle x,\iota\rangle }, \qquad x\in \mathbb R^{d}, {} \end{aligned} $$
(74)

where

$$\displaystyle \begin{aligned} \Lambda &= \mathbb Z^{d}\times (\pi \mathbb Z^{d}), {} \end{aligned} $$
(75)

which is the Gabor expansion of f with respect to the Gabor pair (ϕ, ψ) and lattice Λ, i.e. with respect to the Gabor atom ϕ and the dual Gabor atom ψ. Here the series converges in \(\Sigma _s^{\prime }(\mathbb R^{d})\). By duality and the fact that compactly supported elements in \(\Sigma _s(\mathbb R^{d})\) are dense in \(\Sigma _s^{\prime }(\mathbb R^{d})\) we also have

$$\displaystyle \begin{aligned} f(x) = \left ( \frac \pi 2\right )^{\frac d2} \sum _{(j,\iota )\in \Lambda}V_\psi f(j,\iota )\phi (x-j)e^{i\langle x,\iota\rangle }, \qquad x\in \mathbb R^{d}, \end{aligned} $$
(76)

with convergence in \(\Sigma _s^{\prime }(\mathbb R^{d})\).

Let T be a linear continuous operator from \(\Sigma _s(\mathbb R^{d})\) to \(\Sigma _s^{\prime }(\mathbb R^{d})\) and let \(f\in \Sigma _s(\mathbb R^{d})\). Then it follows from (74) that

$$\displaystyle \begin{aligned} (Tf)(x) = \left ( \frac \pi 2\right )^{\frac d2} \sum _{(j,\iota )\in \Lambda}V_\phi f(j,\iota )T(\psi ( \cdot \, -j)e^{i\langle \, \cdot \,,\iota\rangle })(x) \end{aligned}$$

and

$$\displaystyle \begin{aligned} T(\psi ( \cdot \, -j)e^{i\langle \, \cdot \,,\iota\rangle })(x) = \left ( \frac \pi 2\right )^{\frac d2} \sum _{(k,\kappa )\in \Lambda}(V_\phi (T(\psi ( \cdot \, -j)e^{i\langle \, \cdot \,,\iota\rangle })))(k,\kappa ) \psi (x-k)e^{i\langle x,\kappa\rangle }. \end{aligned}$$

A combination of these expansions show that

$$\displaystyle \begin{aligned} (Tf)(x) = \left ( \frac \pi 2\right )^{\frac d2} \sum _{(j,\iota )\in \Lambda} (A\cdot V_\phi f)(j,\iota )\psi (x-j)e^{i\langle x,\iota\rangle }, \end{aligned} $$
(77)

where A = (a(j, k))j,k ∈ Λ is the Λ × Λ-matrix, given by

$$\displaystyle \begin{aligned} a(\boldsymbol j ,\boldsymbol k ) =& \left ( \frac \pi 2\right )^{\frac d2} (T(\psi ( \cdot \, -j)e^{i\langle \, \cdot \,,\iota\rangle }) , \phi ( \cdot \, -k)e^{i\langle \, \cdot \,,\kappa\rangle } )_{L^2(\mathbb R^{d})} \\ &\qquad \qquad \qquad \qquad \qquad \text{when} \quad \boldsymbol j =(j,\iota ) \ \text{and} \ \boldsymbol k =(k,\kappa ). \end{aligned} $$
(78)

By the Gabor analysis for modulation spaces we get the following restatement of [54, Proposition 1.8]. We refer to [17, 19,20,21, 25, 27, 28, 50] for details.

Proposition 3.4

Let s > 1, p, q ∈ (0, ], \(\omega \in \mathcal P _E (\mathbb R^{2d})\) be such that (68) holds for some r > 0, \(\phi ,\psi \in \Sigma _s (\mathbb R^{d}) \) with values in [0, 1] be such that (70), (71) and (72) hold true, and let \(f\in \Sigma _s^{\prime }(\mathbb R^{d})\) . Then the following is true:

  1. 1.

    \(f\in M^{p,q}_{(\omega )}(\mathbb R^{d})\) , if and only if \(\Vert V_\phi f\Vert _{\ell _{(\omega )}^{p,q}(\mathbb Z^{d}\times \pi \mathbb Z^{d})}<\infty \);

  2. 2.

    \(f\in M^{p,q}_{(\omega )}(\mathbb R^{d})\) , if and only if \(\Vert V_\psi f\Vert _{\ell _{(\omega )}^{p,q}(\mathbb Z^{d}\times \pi \mathbb Z^{d})}<\infty \);

  3. 3.

    the quasi-norms

    $$\displaystyle \begin{aligned} f\mapsto \Vert V_\phi f\Vert _{\ell _{(\omega )}^{p,q}(\mathbb Z^{d}\times \pi \mathbb Z^{d})} \quad \mathit{\text{and}}\quad f\mapsto \Vert V_\psi f\Vert _{\ell _{(\omega )}^{p,q}(\mathbb Z^{d}\times \pi \mathbb Z^{d})} \end{aligned}$$

    are equivalent to \(\Vert \, \cdot \,\Vert _{M^{p,q}_{(\omega )}}\).

The same holds true with \(W^{p,q}_{(\omega )}\) and \(\ell _{*,(\omega )}^{p,q}\) in place of \(M^{p,q}_{(\omega )}\) respectively \(\ell _{(\omega )}^{p,q}\) at each occurrence.

3.2 Multiplications and Convolutions in Modulation Spaces

As a first step for approaching multiplications and convolutions for elements in modulation spaces, we reformulate such products in terms of short-time Fourier transforms. Let \(\phi _0,\phi _1 ,\phi _2\in \Sigma _1 (\mathbb R^{d})\) be fixed such that

$$\displaystyle \begin{aligned} \phi _0=(2\pi )^{-\frac {d}2}\phi _1\phi _2 \end{aligned} $$
(79)

and let \(f_1,f_2\in \Sigma _1 (\mathbb R^{d})\). Then the multiplication f 0 = f 1 f 2 can be expressed by

$$\displaystyle \begin{aligned} F_0(x,\xi ) &= \big ( F_1(x,\, \cdot \, ) * F_2(x,\, \cdot \, ) \big ) (\xi ). {} \end{aligned} $$
(80)

where

$$\displaystyle \begin{aligned} F_j &= V_{\phi _j}f_j,\qquad j=0,1,2. {} \end{aligned} $$
(81)

In fact, by Fourier’s inversion formula we get

$$\displaystyle \begin{aligned} &\ \big ( (V_{\phi _1}f_1)(x,\, \cdot \, ) * (V_{\phi _2}f_2)(x,\, \cdot \, ) \big ) (\xi ) \\ &= (2\pi )^{-d} \iiint f_1(y_1)\overline{\phi _1(y_1-x)} f_2(y_2)\overline{\phi _2(y_2-x)} e^{-{i\langle y_1,\xi -\eta\rangle }} e^{-i\langle y_2,\eta\rangle } \, dy_1dy_2d\eta \\ &= \int f_1(y)\overline{\phi _1(y-x)} f_2(y)\overline{\phi _2(y-x)}e^{-i\langle y,\xi\rangle }\, dy =(2\pi )^{\frac d2}(V_{\phi _1\phi _2}(f_1f_2))(x,\xi ). \end{aligned} $$

We also observe that we may extract f 0 = f 1 f 2 by the formula

$$\displaystyle \begin{aligned} f_0 = (\Vert \phi _0\Vert _{L^2})^{-1}V_{\phi _0}^*F_0 , \end{aligned} $$
(82)

provided ϕ 0 is not trivially equal to 0.

In the same way, let \(\phi _0,\phi _1,\phi _2\in \Sigma _1 (\mathbb R^{d})\) be fixed such that

$$\displaystyle \begin{aligned} \phi _0=(2\pi )^{\frac d2}\phi _1*\phi _2, \end{aligned} $$
(83)

and let \(f_1,f_2,g\in \Sigma _1 (\mathbb R^{d})\). Then the convolution f 0 = f 1 ∗ f 2 can be expressed by

$$\displaystyle \begin{aligned} F_0(x,\xi ) &= \big ( F_1( \cdot \, ,\xi ) *F_2( \cdot \, ,\xi ) \big ) (x). {} \end{aligned} $$
(84)

where F j are given by (81), and that we may extract f 0 = f 1 ∗ f 2 from (82).

Next we discuss convolutions and multiplications for modulation spaces, and start with the following convolution result for modulation spaces. For multiplications of elements in modulation spaces we need to swap the conditions for the involved Lebesgue exponents compared to (51) and (52). That is, these conditions become

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \frac 1{p_0} &\le \frac 1{p_1}+\frac 1{p_2},& \qquad \frac 1{q_0} &\le \frac 1{q_1}+\frac 1{q_2} - \max \left ( 1,\frac 1{p_0},\frac 1{q_1},\frac 1{q_2} \right ), {} \end{array}\end{aligned} $$
(85)

or

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \frac 1{p_0} &\le \frac 1{p_1}+\frac 1{p_2},& \qquad \frac 1{q_0} &\le \frac 1{q_1}+\frac 1{q_2} - \max \left ( 1,\frac 1{q_1},\frac 1{q_2} \right ). {} \end{array}\end{aligned} $$
(86)

The conditions on the weight functions are

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \omega _0(x,\xi _1+\xi _2) &\le \omega _1(x,\xi _1)\omega _2(x,\xi _2), & \qquad x,\xi _1,\xi _2 &\in \mathbb R^{d}, {} \end{array}\end{aligned} $$
(87)

respectively

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \omega _0(x_1+x_2,\xi ) &\le \omega _1(x_1,\xi )\omega _2(x_2,\xi ), & \qquad x_1,x_2,\xi &\in \mathbb R^{d} . {} \end{array}\end{aligned} $$
(88)

Theorem 3.5

Let p j, q j ∈ (0, ) and \(\omega _j\in \mathcal P _E(\mathbb R^{2d})\) , j = 0, 1, 2, be such that (85) and (87) hold. Then (f 1, f 2)↦f 1 f 2 from \(\Sigma _1 (\mathbb R^{d})\times \Sigma _1(\mathbb R^{d})\) to \(\Sigma _1 (\mathbb R^{d})\) is uniquely extendable to a continuous map from \(M^{p_1,q_1}_{(\omega _1)}(\mathbb R^{d}) \times M^{p_2,q_2}_{(\omega _2)}(\mathbb R^{d})\) to \(M_{(\omega _0)}^{p_0,q_0}(\mathbb R^{d})\) , and

$$\displaystyle \begin{aligned} \Vert f_1f_2\Vert _{M_{(\omega _0)}^{p_0,q_0}} \lesssim \Vert f_1\Vert _{M_{(\omega _1)}^{p_1,q_1}}\Vert f_2\Vert _{M_{(\omega _2)}^{p_2,q_2}} , \quad f_j\in M_{(\omega _j)}^{p_j,q_j}(\mathbb R^{d}),\ j=1,2 \text . \end{aligned} $$
(89)

Theorem 3.6

Let p j, q j ∈ (0, ) and \(\omega _j\in \mathcal P _E(\mathbb R^{2d})\) , j = 0, 1, 2, be such that (86) and (87) hold. Then (f 1, f 2)↦f 1 f 2 from \(\Sigma _1 (\mathbb R^{d})\times \Sigma _1(\mathbb R^{d})\) to \(\Sigma _1 (\mathbb R^{d})\) is uniquely extendable to a continuous map from \(W^{p_1,q_1}_{(\omega _1)}(\mathbb R^{d}) \times W_{(\omega _2)}^{p_2,q_2}(\mathbb R^{d})\) to \(W_{(\omega _0)}^{p_0,q_0}(\mathbb R^{d})\) , and

$$\displaystyle \begin{aligned} \Vert f_1f_2\Vert _{W_{(\omega _0)}^{p_0,q_0}} \lesssim \Vert f_1\Vert _{W_{(\omega _1)}^{p_1,q_1}} \Vert f_2\Vert _{W_{(\omega _2)}^{p_2,q_2}} , \quad f_j\in W_{(\omega _j)}^{p_j,q_j}(\mathbb R^{d}),\ j=1,2 \text . \end{aligned} $$
(90)

The corresponding results for convolutions are the following. Here the conditions on the involved Lebesgue exponents are swapped as

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \frac 1{p_0} &\le \frac 1{p_1}+\frac 1{p_2} - \max \left ( 1,\frac 1{q_0},\frac 1{p_1},\frac 1{p_2} \right ), & \qquad \frac 1{q_0} &\le \frac 1{q_1}+\frac 1{q_2} {} \end{array}\end{aligned} $$
(91)

or

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \frac 1{p_0} &\le \frac 1{p_1}+\frac 1{p_2} - \max \left ( 1,\frac 1{p_1},\frac 1{p_2} \right ), & \qquad \frac 1{q_0} &\le \frac 1{q_1}+\frac 1{q_2} {} \end{array}\end{aligned} $$
(92)

Theorem 3.7

Let p j, q j ∈ (0, ) and \(\omega _j\in \mathcal P _E(\mathbb R^{2d})\) , j = 0, 1, 2, be such that (88) and (92) hold. Then (f 1, f 2)↦f 1 ∗ f 2 from \(\Sigma _1 (\mathbb R^{d})\times \Sigma _1(\mathbb R^{d})\) to \(\Sigma _1 (\mathbb R^{d})\) is uniquely extendable to a continuous map from \(M^{p_1,q_1}_{(\omega _1)}(\mathbb R^{d}) \times M^{p_2,q_2}_{(\omega _2)}(\mathbb R^{d})\) to \(M_{(\omega _0)}^{p_0,q_0}(\mathbb R^{d})\) , and

$$\displaystyle \begin{aligned} \Vert f_1*f_2\Vert _{M_{(\omega _0)}^{p_0,q_0}} \lesssim \Vert f_1\Vert _{M_{(\omega _1)}^{p_1,q_1}}\Vert f_2\Vert _{M_{(\omega _2)}^{p_2,q_2}} , \quad f_j\in M_{(\omega _j)}^{p_j,q_j}(\mathbb R^{d}),\ j=1,2 \text . \end{aligned} $$
(93)

Theorem 3.8

Let p j, q j ∈ (0, ) and \(\omega _j\in \mathcal P _E(\mathbb R^{2d})\) , j = 0, 1, 2, be such that (88) and (91) hold. Then (f 1, f 2)↦f 1 ∗ f 2 from \(\Sigma _1 (\mathbb R^{d})\times \Sigma _1(\mathbb R^{d})\) to \(\Sigma _1 (\mathbb R^{d})\) is uniquely extendable to a continuous map from \(W^{p_1,q_1}_{(\omega _1)}(\mathbb R^{d}) \times W^{p_2,q_2}_{(\omega _2)}(\mathbb R^{d})\) to \(W_{(\omega _0)}^{p_0,q_0}(\mathbb R^{d})\) , and

$$\displaystyle \begin{aligned} \Vert f_1*f_2\Vert _{W_{(\omega _0)}^{p_0,q_0}} \lesssim \Vert f_1\Vert _{W_{(\omega _1)}^{p_1,q_1}}\Vert f_2\Vert _{W_{(\omega _2)}^{p_2,q_2}} , \quad f_j\in W_{(\omega _j)}^{p_j,q_j}(\mathbb R^{d}),\ j=1,2 \text . \end{aligned} $$
(94)

We observe that Theorems 3.2–3.5 in [54] are multi-linear versions of the previous results. In particular, Theorems 3.5 and 3.6 are Fourier transformations of Theorems 3.7 and 3.8. Hence it suffices to prove the last two theorems, cf. [54]. To shed some ideas of the arguments, we give a proof in the unweighted case of Theorem 3.7. We will use Proposition A.1 from Appendix A, which is a special case of [54, Proposition 3.6].

Proof of Theorem 3.7

Suppose \(f_j\in \mathcal S(\mathbb R^{d})\), \(\phi _j\in \mathcal S(\mathbb R^{d})\), j = 0, 1, 2 be such that

$$\displaystyle \begin{aligned} f_0=f_1*f_2 \quad \text{and}\quad \phi _0=(2\pi) ^{\frac d2}\phi _1*\phi _2\neq 0, \end{aligned}$$

and let F j be the same as in (81). Then

$$\displaystyle \begin{aligned} F_0(x,\xi ) = (V_{\phi _1}f_1( \cdot \, ,\xi )*V_{\phi _2}f_2( \cdot \, ,\xi ))(x), \end{aligned}$$

in view of (84).

We have

$$\displaystyle \begin{aligned} 0\le \chi _{k_1+Q}*\chi _{k_2+Q} \le \chi _{k_1+k_2+Q_{d,2}}, \qquad k_1,k_2\in \mathbb Z^{d}, \end{aligned}$$

where Q d,r is the cube

$$\displaystyle \begin{aligned} Q_{d,r} =[0,r]^d \quad \text{and}\quad Q= Q_{d,1}= [0,1]^d, \end{aligned}$$

and χ E is the characteristic function with respect to the set E.

Set

$$\displaystyle \begin{aligned} G(x,\xi ) &= (|V_{\phi _1}f_1( \cdot \, ,\xi )|*|V_{\phi _2}f_2( \cdot \, ,\xi )|)(x), \\ {} a_j(k,\kappa ) &= \Vert V_{\phi _j}f_j\Vert _{L^\infty ((k,\kappa )+Q_{2d,1})}, \quad j=1,2, \end{aligned} $$

and

$$\displaystyle \begin{aligned} b(k,\kappa ) &= \Vert G\Vert _{L^\infty ((k,\kappa )+Q_{2d,1})} \end{aligned} $$

Then

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Vert V_{\phi _0}^*F_0\Vert _{M^{p_0,q_0}} &\asymp \Vert P_{\phi _0}F_0\Vert _{\mathsf{W} (\ell ^{p_0,q_0})} \lesssim \Vert F_0\Vert _{\mathsf{W} (\ell ^{p_0,q_0})} \notag \\ {} &\le \Vert G\Vert _{\mathsf{W} (\ell ^{p_0,q_0})} \asymp \Vert b\Vert _{\ell ^{p_0,q_0}}, {} \end{array}\end{aligned} $$
(95)

and

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \Vert f_j\Vert _{M^{p_j,q_j}} &\asymp \Vert a_j\Vert _{\ell ^{p_j,q_j}} {} \end{array}\end{aligned} $$
(96)

in view of (A.5) and Proposition A.1 in Appendix A (see also [25, Theorem 3.3])).

By (84) we have

$$\displaystyle \begin{aligned} G(x,\lambda ) &\le \sum _{k_1,k_2\in \mathbb Z^{d}} a_1(k_1,\lambda )a_2(k_2,\lambda ) (\chi _{k_1+Q}*\chi _{k_2+Q})(x) \\ {} &\le \sum _{k_1,k_2\in \mathbb Z^{d}} a_1(k_1,\lambda )a_2(k_2,\lambda ) \chi _{k_1+k_2+Q_{d,2}}(x). \end{aligned} $$
(97)

We observe that

$$\displaystyle \begin{aligned} \chi _{k_1+k_2+Q_{d,2}}(x)=0 \quad \text{when}\quad x\notin l+Q_d,\ (k_1,k_2)\notin \Omega _l, \end{aligned}$$

where

$$\displaystyle \begin{aligned} \Omega _l = \{ \, (k_1,k_2)\in \mathbb Z^{2d}\, ;\, l_j-2 \le k_{1,j} +k_{2,j}\le l_j+1\, \} , \end{aligned}$$

and

$$\displaystyle \begin{aligned} k_j=(k_{j,1},\dots ,k_{j,d})\in \mathbb Z^{d}, \qquad j=1,2, \quad \text{and}\quad l=(l_1,\dots ,l_d)\in \mathbb Z^{d}. \end{aligned}$$

Hence, if x = l in (97), we get

$$\displaystyle \begin{aligned} &b(l,\lambda ) \le \sum _{(k_1,k_2)\in \Omega _l} a_1(k_1,\lambda )a_2(k_2,\lambda ) \\ &\qquad \qquad \qquad \qquad \qquad \le \sum _{m\in I}(a_1( \cdot \, ,\lambda )*a_2( \cdot \, ,\lambda ))(l-2e_0+m), \end{aligned} $$
(98)

where \(e_0=(1,\dots ,1)\in \mathbb Z^{d}\) and I = {0, 1, 2, 3}d.

If we apply the \(\ell ^{p_0}\) quasi-norm on (98) with respect to the l variable, then Proposition 2.5 (2) and the fact that I is finite set give

$$\displaystyle \begin{aligned} \Vert b( \cdot \, ,\lambda )\Vert _{\ell ^{p_0}} \le & \left \Vert \sum _{m\in I} ( a_{1}( \cdot \, ,\lambda ) *a_{2}( \cdot \, ,\lambda )) ( \cdot \, -2e_0+m)\right\Vert{\ell ^{p_0}} \\ &\le \sum _{m\in I} \Vert ( a_{1}( \cdot \, ,\lambda ) *a_{2}( \cdot \, ,\lambda )) ( \cdot \, -2e_0+m)\Vert{\ell ^{p_0}} \\ &\qquad \qquad \asymp \Vert a_{1}( \cdot \, ,\lambda ) *a_{2}( \cdot \, ,\lambda )\Vert _{\ell ^{p_0}} \\ &\qquad \qquad \qquad \qquad \le \Vert a_{1}( \cdot \, ,\lambda )\Vert _{\ell ^{p_1}} \Vert a_{2}( \cdot \, ,\lambda )\Vert _{\ell ^{p_2}}. \end{aligned} $$

By applying the \(\ell ^{q_0}\) quasi-norm and using Proposition 2.5 (1) we now get

$$\displaystyle \begin{aligned} \Vert b\Vert _{\ell ^{p_0,q_0}} \lesssim \Vert a_{1}\Vert _{\ell ^{p_1,q_1}}\Vert a_{2}\Vert _{\ell ^{p_2,q_2}}. \end{aligned}$$

This is the same as

$$\displaystyle \begin{aligned} \Vert G\Vert _{L^{p_0,q_0}} \lesssim \Vert F_1\Vert _{L^{p_1,q_1}}\Vert F_2\Vert _{L^{p_2,q_2}}. \end{aligned}$$

A combination of this estimate with (95) and (96) gives that f 1 ∗ f 2 is well-defined and that (93) holds.

The uniqueness now follows from that (93) holds for \(f_1,f_2\in \mathcal S(\mathbb R^{d})\), and that \(\mathcal S(\mathbb R^{d})\) is dense in \(M^{p,q}(\mathbb R^{d})\) when p, q < . □

4 Gabor Products and Modulation Spaces

In this section we give an illustration how the multiplication properties for modulation spaces can be used when treating certain nonlinear problems. We consider the Gabor product which is connected to such multiplication properties. It is introduced in [14] in order to derive a phase space analogue to the usual convolution identity for the Fourier transform. We will prove a formula related to (80), and then use results from previous section to extend the Gabor product initially defined on \( M^1 (\mathbb R^{2d}) \times M^1 (\mathbb R^{2d}) \) to some other spaces. Finally, we show how the Gabor product gives rise to a phase-space formulation of the qubic Schrödinger equation.

Definition 4.1

Let \(\phi \in M^1 (\mathbb R^{d}) \setminus \{ 0 \}\), and let \(F_1 , F_2 \in M^1 (\mathbb R^{2d})\). Then the Gabor product ϕ is given by

$$\displaystyle \begin{aligned} &\quad \left(F_1 \natural_{\phi}\, F_2 \right) (x, \xi) \\ &= (2\pi )^{-d} e^{-i\langle x,\xi\rangle } \iiint_{\mathbb R^{3d} } \overline{\widehat{\phi} \left(\zeta -\xi \right)} e^{i\langle x,\zeta\rangle } F_1 (y,\eta ) F_2 (y, \zeta - \eta )\, dyd\eta d\zeta . \end{aligned} $$
(99)

In the proof of [14, Lemma 13] it is justified that the Gabor product in (99) is well-defined, and that

$$\displaystyle \begin{aligned} \natural _\phi : M^1 (\mathbb R^{2d})\times M^1 (\mathbb R^{2d}) \to M^1 (\mathbb R^{2d}) \end{aligned}$$

is a continuous map.

The Gabor product is particularly well-suited in the context of the STFT.

Theorem 4.2

Let \(\phi , \phi _1, \phi _2 \in M^1 (\mathbb R^{d}) \backslash \left \{0 \right \}\) . Then

$$\displaystyle \begin{aligned} ( \phi _2,\phi _1 )_{L^2 (\mathbb R^{d})} V_{\phi} (f_1 \cdot f_2) = (V_{\phi _1} f_1) \natural_{\phi} (V_{\overline \phi _2} f_2), \quad f_1, f_2 \in M^1 (\mathbb R^{d}). {} \end{aligned} $$
(100)

Moreover, \( V_{\phi } (f_1 \cdot f_2) \in M^1 (\mathbb R^{2d})\).

Proof

We have

$$\displaystyle \begin{aligned} ((V_{\phi _1}f_1) &\natural _\phi (V_{\overline \phi _2}f_2))(x,\xi ) \end{aligned} $$
(101)
$$\displaystyle \begin{aligned}{}[1ex] &= (2\pi )^{-d}e^{-i\langle x,\xi\rangle } \iint _{\mathbb R^{2d}} \overline {\widehat \phi (\zeta -\xi )}e^{i\langle x,\zeta\rangle }G(y,\zeta )\, dyd\zeta , {} \end{aligned} $$
(102)

where

$$\displaystyle \begin{aligned} G(y,\zeta ) &= \int _{\mathbb R^{d}} (V_{\phi _1}f_1)(y,\eta ) (V_{\overline \phi _2}f_2)(y,\zeta -\eta )\, d\eta . \notag \end{aligned} $$

By Parseval’s formula we get

$$\displaystyle \begin{aligned} G(y,\zeta ) &= \int _{\mathbb R^{d}} (V_{\phi _1}f_1)(y,\eta ) (V_{\overline \phi _2}f_2)(y,\zeta -\eta )\, d\eta \\ {} &= \int _{\mathbb R^{d}} \mathcal F (f_1\overline {\phi _1( \cdot \, -y)})(\eta ) \mathcal F (f_2{\phi _2( \cdot \, -y)})(\zeta -\eta )\, d\eta \\ {} &= (\mathcal F (f_1\overline {\phi _1( \cdot \, -y)})\, ,\, \mathcal F (\overline {f_2\, \phi _2( \cdot \, -y)}e^{i\langle \, \cdot \,,\zeta\rangle }) )_{L^2(\mathbb R^{d})} \\ {} &= (f_1\overline {\phi _1( \cdot \, -y)}\, ,\, \overline {f_2\, \phi _2( \cdot \, -y)}e^{i\langle \, \cdot \,,\zeta\rangle } )_{L^2(\mathbb R^{d})} \\ {} &= \int _{\mathbb R^{d}}f_1(z)\overline {\phi _1(z-y)} f_2(z)\phi _2(z-y)e^{-i\langle z,\zeta\rangle }\, dz . \end{aligned} $$

By inserting this into (102) and using Fubini’s theorem we get

$$\displaystyle \begin{aligned} ((V_{\phi _1}f_1) &\natural _\phi (V_{\overline \phi _2}f_2))(x,\xi ) \\ {} &= (2\pi )^{-d}e^{-i\langle x,\xi\rangle } \iint _{\mathbb R^{2d}} \overline {\widehat \phi (\zeta -\xi )}e^{-i\langle z-x,\zeta\rangle }f_1(z)f_2(z) H(z) \, dzd\zeta , \end{aligned} $$

where

$$\displaystyle \begin{aligned} H(z) &= \int _{\mathbb R^{d}}\phi _2(z-y)\overline {\phi _1(z-y)}\, dy = (\phi _2,\phi _1)_{L^2}. \end{aligned} $$

Hence, by evaluating the integral with respect to ζ, and using Fourier’s inversion formula, we get

$$\displaystyle \begin{aligned} ((V_{\phi _1}f_1) &\natural _\phi ((V_{\overline \phi _2}f_2)))(x,\xi ) \\ {} &= (2\pi )^{-\frac d2}e^{-i\langle x,\xi\rangle }(\phi _2,\phi _1)_{L^2} \int _{\mathbb R^{d}}\overline {\phi (z-x)}e^{i\langle x-z,\xi\rangle }f_1(z)f_2(z) \, dz \\ {} &= (\phi _2,\phi _1)_{L^2} V_\phi (f_1f_2)(x,\xi ), \end{aligned} $$

which gives (100), and the result follows. □

The formula (100) is closely related to (80). In fact, the windows \( \phi _j \in \Sigma _1 (\mathbb R^{d})\), j = 0, 1, 2, in (80) should satisfy the condition (79), while (100) is valid for arbitrary non-zero elements from \( M^1 (\mathbb R^{d})\). For example, when ϕ = ϕ 1 = ϕ 2 and \(\Vert \phi \Vert _{L^2 (\mathbb R^{d})}=1\), then (100) reduces to

$$\displaystyle \begin{aligned} V_\phi (f_1 \cdot f_2) = (V_\phi f_1) \natural_\phi (V_\phi f_2), \qquad f_1, f_2 \in M^1 (\mathbb R^{d}), {} \end{aligned} $$
(103)

while (80) does not allow such choice of windows.

One of the main goals of [14] are extensions of the Gabor product to some function spaces \(\mathcal {F}_j (\mathbb R^{2d})\), j = 0, 1, 2, so that ϕ maps \(\mathcal {F}_1 \times \mathcal {F}_2\) into \(\mathcal {F}_0\), with:

$$\displaystyle \begin{aligned} \|F_1 \natural_\phi F_2 \|{}_{\mathcal{F}_0} \leq C \|F_1\|{}_{\mathcal{F}_1} \|F_2\|{}_{\mathcal{F}_2}. {} \end{aligned} $$
(104)

This can be considered as a phase space form of the Young convolution inequality.

Next we discuss continuity of the Gabor product on certain spaces involving superpositions of short-time Fourier transforms. In the end we deduce properties similar to [14, Theorem 29]. Instead of modulation spaces of the form \(M^{p,q}_{(\omega )}(\mathbb R^{d})\), p, q ∈ [1, ), \(\omega \in \mathcal P _E(\mathbb R^{2d})\), here we consider modulation spaces of Wiener amalgam types \(W^{p,q}_{(\omega )}(\mathbb R^{d})\), and allow the “quasi-Banach” choice for Lebesgue parameters, i.e. p and q are allowed to be smaller than one.

Thus, in what follows we assume that p, q ∈ (0, ), \(\omega \in \mathcal P _E(\mathbb R^{2d})\) is v-moderate, and consider \(L^{p,q}_{*,(\omega )}(\mathbb R^{2d})\) spaces rather than \(L^{p,q}_{(\omega )}(\mathbb R^{2d})\) which are treated in [14].

We need some additional notation. Let s > 1, \(N\in \mathbb {N}\) be given, and let

$$\displaystyle \begin{aligned} \mathcal{G}= \left \{ \phi _n = \overline{\phi _n}\, ;~ n \in \mathbb{N} \right \} \subseteq \Sigma_s (\mathbb R^{d}), \end{aligned}$$

be an orthonormal basis of \( L^2 (\mathbb R^{d})\). Then let \(\mathcal {V}_{\mathcal {G},\omega }^{(N),p,q}(\mathbb R^{2d})\) be the closure of

$$\displaystyle \begin{aligned} {\mathcal V}_{\mathcal{G}}^{(N)} (\mathbb R^{2d}) = \left \{ \sum_{n=1}^N V_{\phi _n}f_n~;~f_n \in \Sigma _1(\mathbb R^{d}) \right \} {} \end{aligned} $$
(105)

with respect to the \(L^{p,q}_{*,(\omega )}(\mathbb R^{2d})\) norm. In particular, if N = 1, ϕ = ϕ 1 and p, q ≥ 1, then this reduces to the closure

$$\displaystyle \begin{aligned} {P_\phi }(L^{p,q}_{*,(\omega )}(\mathbb R^{2d}) ) = V_\phi (W^{p,q} _{(\omega )} (\mathbb R^{d}) ) \end{aligned}$$

of

$$\displaystyle \begin{aligned} P_\phi (\Sigma _1(\mathbb R^{2d}) ) = V_\phi (\Sigma _1(\mathbb R^{d}) ) \end{aligned}$$

with respect to the \(L^{p,q}_{*,(\omega )}(\mathbb R^{2d})\) norm.

By [14, Theorem 26], it follows that for every \( F \in \mathcal {V}_{\mathcal {G},\omega }^{(N),p,q}(\mathbb R^{2d})\) there exist \(f_n \in W_{(\omega )}^{p,q}(\mathbb R^{d})\), n = 1, 2, …, N, and such that

$$\displaystyle \begin{aligned} F = \sum_{n=1}^N V_{\phi _n}f_n~. {} \end{aligned} $$
(106)

Theorem 4.3

Let p j, q j ∈ (0, ) and \(\omega _j\in \mathcal P _E(\mathbb R^{2d})\) be v j –moderate, j = 0, 1, 2, and such that (86) and (87) hold, and let \(\phi \in \Sigma _s (\mathbb R^{d}) \) , s > 1. Then the Gabor product ♮ ϕ from \( {\mathcal V}_{\mathcal {G}}^{(N)} \left (\mathbb R^{2d}\right ) \times {\mathcal V}_{\mathcal {G}}^{(N)} \left (\mathbb R^{2d}\right ) \) to \(W_{(v)}^{1,1}(\mathbb R^{2d})\) , extends uniquely to a continuous map from \(\mathcal {V}_{\mathcal {G},\omega _1}^{(N),p_1,q_1} (\mathbb R^{2d}) \times \mathcal {V}_{\mathcal {G},\omega _2}^{(N),p_2,q_2}(\mathbb R^{2d})\) to the closure of \({P_\phi }(L^{p_0,q_0}_{*,(\omega _0 )}(\mathbb R^{2d}) )\) , and

$$\displaystyle \begin{aligned} \|F_1 \natural_{\phi} F_2 \|{}_{L_{*, (\omega_0)}^{p_0,q_0}} \lesssim \|F_1 \|{}_{L_{*, (\omega_1)}^{p_1,q_1}} ~ \|F_2 \|{}_{L_{*, (\omega_2)}^{p_2,q_2}}~, {} \end{aligned} $$
(107)

for all \(F_j \in \mathcal {V}_{\mathcal {G}, \omega _j}^{(N),p_j,q_j}(\mathbb R^{2d}) \) , j = 1, 2.

In particular, if F j = V ϕ f j , j = 1, 2, and \(\| \phi \|{ }_{L^2}= 1\) , then (107) reduces to

$$\displaystyle \begin{aligned} \| V_\phi f_1 \natural_{\phi} V_\phi f_2 \| _{L_{*, (\omega_0)}^{p_0,q_0}}= \| f_1 f_2 \|{}_{W_{(\omega _0)}^{p_0,q_0}} \lesssim \|f_1 \|{}_{W_{(\omega _1)}^{p_1,q_1}} ~ \|f_2 \|{}_{W_{(\omega _2)}^{p_2,q_2}}~. {} \end{aligned} $$
(108)

We omit the proof which is a slight modification of the proof of Theorem 29 in [14].

We end the paper by formally demonstrating how the Gabor product arises in a phase space version of the cubic Schrödinger equation. Consider the elliptic nonlinear Schrödinger equation (NLSE) given by

$$\displaystyle \begin{aligned} i \frac{\partial \psi}{\partial t} + \Delta \psi + \lambda | \psi|{}^{2} \psi =0, {} \end{aligned} $$
(109)

subject to the initial condition:

$$\displaystyle \begin{aligned} \psi (x,0)= \varphi (x). {} \end{aligned}$$

Here λ = ±1 stands for an attracting (λ = +1) or repulsive (λ = −1) power-law nonlinearity, and the Laplacian is given by

$$\displaystyle \begin{aligned} \Delta = \sum_{j=1}^d \frac{\partial^2}{\partial x_j^2}. \end{aligned}$$

Thus we consider ψ = ϕ(x, t) with \( x \in \mathbb R^{d},\) and t in an open interval \(I \subseteq \mathbb {R}\).

Using the following intertwining relations

$$\displaystyle \begin{aligned} V_\phi (x_j \psi) = - D_{\xi _j} V_{\phi} \psi, \qquad V_\phi ( D_{x_j} \psi ) = \left( \xi _j +D_{x_j} \right ) V_\phi \psi, {} \end{aligned}$$

j = 1, ⋯ , d, and assuming that ϕ is a real-valued window, we obtain upon application of the STFT V ϕ to (109) that

$$\displaystyle \begin{aligned} i \frac{\partial F}{\partial t} - \sum_{j=1}^d \left(\xi _j + D_{x_j} \right)^2 F + \lambda \widetilde F \natural_\phi F \natural_\phi F =0. {} \end{aligned} $$
(110)

Here, \(D_{x_j}=-i \frac {\partial }{\partial x_j}\),

$$\displaystyle \begin{aligned} F(x,\xi ,t) =& V_\phi (\psi( \cdot \, ,t)) (x,\xi ) \\ &= (2\pi )^{-\frac d2} \int _{\mathbb R^{d}} \psi(y,t) \overline{\phi (y-x)} e^{-i\langle y,\xi\rangle } \, dy, \quad x,\xi \in \mathbb R^{d}, \ t \in \mathbb{R}, \end{aligned} $$

and \(\widetilde F \) is given by

$$\displaystyle \begin{aligned} \widetilde F(x, \xi ) = \overline{F (x, - \xi )}. {} \end{aligned} $$
(111)

By considering (110) the phase-space formulation of the initial value problem may be well-posed for more general initial distributions. This means that the phase-space formulation “contains” the solutions of the standard NLSE, but it is richer, as it admits other solutions. We refer to [11,12,13], where phase-space extensions are explored in several different contexts.

Let us conclude by noticing that (110) contains the triple product. Thus, its qualitative analysis calls for a multilinear extension of Theorems 3.6 and 4.3. Then the conditions (86) and (87) become more involved, see [54]. Such analysis demands a more technical tools and arguments and goes beyond the scope of this survey article.