Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-05-31T05:17:04.850Z Has data issue: false hasContentIssue false

Fourier transformable measures with weak Meyer set support and their lift to the cut-and-project scheme

Published online by Cambridge University Press:  17 February 2023

Nicolae Strungaru*
Affiliation:
Department of Mathematical Sciences, MacEwan University, 10700 – 104 Avenue, Edmonton, AB T5J 4S2, Canada and Institute of Mathematics “Simon Stoilow”, Bucharest, Romania
Rights & Permissions [Opens in a new window]

Abstract

In this paper, we prove that given a cut-and-project scheme $(G, H, \mathcal {L})$ and a compact window $W \subseteq H$, the natural projection gives a bijection between the Fourier transformable measures on $G \times H$ supported inside the strip ${\mathcal L} \cap (G \times W)$ and the Fourier transformable measures on G supported inside ${\LARGE \curlywedge }(W)$. We provide a closed formula relating the Fourier transform of the original measure and the Fourier transform of the projection. We show that this formula can be used to re-derive some known results about Fourier analysis of measures with weak Meyer set support.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

After the discovery of quasicrystals [Reference Shechtman, Blech, Gratias and Cahn39], it has become clear that we need to better understand the process of diffraction. Mathematically, the diffraction pattern of a solid can be viewed as the Fourier transform $\widehat {\gamma }$ of the autocorrelation measure $\gamma $ of the structure (see [Reference Hof13] for the setup and the monographs and see [Reference Baake and Grimm3, Reference Baake and Grimm4] for a general review of the theory). The measure $\gamma $ is positive-definite, and therefore it is Fourier transformable as a measure [Reference Argabright and de Lamadrid1, Reference Berg and Forst8, Reference Moody, Strungaru, Baake and Grimm31] with positive Fourier transform $\widehat {\gamma }$ . It is this measure $\widehat {\gamma }$ , which models the diffraction of our solid.

Structures with pure point diffraction, that is, structures for which $\widehat {\gamma }$ is a pure point measure, are now very well understood. Building on the earlier work of Gil deLamadrid–Argabright [Reference de Lamadrid and Argabright10], Solomyak [Reference Solomyak40, Reference Solomyak and Patera41], Lee–Moody–Solomyak [Reference Lee, Moody and Solomyak20], Baake–Moody [Reference Baake and Moody7], Baake–Lenz [Reference Baake and Lenz6], Gouere [Reference Gouéré11, Reference Gouéré12], Moody–Strungaru [Reference Moody and Strungaru30], and Meyer [Reference Meyer27], pure point diffraction was characterized in [Reference Lenz, Spindeler and Strungaru22, Reference Lenz, Spindeler and Strungaru23]. The focus now shifted toward models with mixed diffraction spectra, especially those with a large pure point part.

The best mathematical models for Delone sets with a large pure point spectrum and (generic) positive entropy are Meyer sets. They have been introduced in the pioneering work of Meyer [Reference Meyer26], and popularized in the area of Aperiodic Order by Moody [Reference Moody and Moody28, Reference Moody, Axel, Dénoyer and Gazeau29] and Lagarias [Reference Lagarias18, Reference Lagarias, Baake and Moody19]. They are usually constructed via a cut-and-project scheme (or simply CPS) and can be characterized via harmonic analysis, discrete geometry, algebra, and almost periodicity [Reference Meyer26, Reference Moody, Axel, Dénoyer and Gazeau29, Reference Strungaru, Baake and Grimm46]. The basic idea behind a CPS is to project points from a higher-dimensional lattice, which lie within a bounded strip of the real space, into the real space (see Definition 2.8 for the exact definition). If the cross section of the strip (called the window) is regular, then the resulting model set is pure point diffractive [Reference Baake and Moody7, Reference Hof and Patera14, Reference Richard and Strungaru35, Reference Schlottmann, Baake and Moody38]. Recent work proved pure point diffractivity for a larger class of weak model sets [Reference Baake, Huck and Strungaru5, Reference Keller and Richard15Reference Keller, Richard and Strungaru17, Reference Strungaru48].

As subsets of regular model sets, Meyer sets still exhibit a large pure point spectrum [Reference Strungaru43Reference Strungaru47, Reference Strungaru49] and a highly ordered continuous spectrum [Reference Strungaru43, Reference Strungaru45, Reference Strungaru47, Reference Strungaru49]. The long-range order of the spectrum of Meyer sets can be traced to that of a covering regular model set [Reference Strungaru47, Reference Strungaru49], and can be derived from the Poisson summation formula for the lattice in the CPS [Reference Baake and Grimm3, Reference Lev and Olevskii25, Reference Richard and Strungaru34, Reference Richard and Strungaru35].

One would expect it to be possible to relate the diffraction of a Meyer set (or more generally a measure with Meyer set support) directly to the lattice ${\mathcal L}$ in the CPS. It is the goal of this paper to establish this connection. Let us briefly explain our approach.

Fix a CPS $(G, H, {\mathcal L})$ and a compact set $W \subseteq H$ . It is easy to see that

$$\begin{align*}\gamma= \sum_{x \in {\LARGE \curlywedge}(W)} c_x \delta_x \qquad \longleftrightarrow \qquad \eta= \sum_{x \in {\LARGE \curlywedge}(W)} c_x \delta_{(x,x^{\star})} \end{align*}$$

establishes a bijection between translation bounded measures supported inside ${\LARGE \curlywedge }(W)$ and translation bounded measures supported inside ${\mathcal L} \cap (G \times W)$ . We first show in Proposition 3.6 that $\gamma $ is positive-definite if and only if $\eta $ is positive-definite. Since each Fourier transformable measure supported inside a Meyer set can be written as a linear combination of positive-definite measures supported inside a common model set, we establish in Theorem 4.1 that $\gamma $ is Fourier transformable if and only if $\eta $ is Fourier transformable, and relate their Fourier transform (see (4.1)).

We complete the paper by discussing in Section 5 how these results can be used to re-derive the known properties of diffraction for measures with weak Meyer set support, and potentially used to prove new results.

2 Definitions and notations

Throughout the paper, G denotes a second countable locally compact Abelian group (LCAG). By $C_{\mathsf {u}}(G)$ , we denote the space of uniformly continuous and bounded functions on G. This is a Banach space with respect to the sup norm $\|. \|_{\infty }$ . As usual, we denote by $C^{}_{0}(G)$ the subspace of $C_{\mathsf {u}}(G)$ consisting of functions vanishing at infinity, and by $C_{\mathsf {c}}(G)$ the subspace of compactly supported continuous functions. Note that $C_{\mathsf {c}}(G)$ is not complete in $(C_{\mathsf {u}}(G), \|. \|_{\infty })$ .

In the spirit of [Reference de Lamadrid and Argabright10], we denote by

$$ \begin{align*} K_2(G) := \mbox{Span} \{ \varphi *\psi : \varphi,\psi \in C_{\mathsf{c}}(G) \} \,. \end{align*} $$

Given two LCAG’s G and H and two functions $g : G \to {\mathbb C}, h : H \to {\mathbb C}$ , we denote by $g \otimes h : G \times H \to {\mathbb C}$ their tensor product

$$ \begin{align*} ( g \otimes h) (x,y)= g(x) h(y) \,. \end{align*} $$

It is obvious that whenever $\varphi \in C_{\mathsf {c}}(G), \psi \in C_{\mathsf {c}}(H)$ we have $\varphi \otimes \psi \in C_{\mathsf {c}}(G \times H)$ . Moreover, if $\varphi \in K_2(G)$ and $\psi \in K_2(H)$ , we have $\varphi \otimes \psi \in K_2(G \times H)$ .

In the rest of this section, we review some of the basic concepts which are important for this paper. For a more general review of these, we recommend [Reference Baake and Grimm3, Reference Baake and Grimm4].

2.1 Measures

In the spirit of Bourbacki [Reference Bourbaki9], by a measure, we understand a linear functional on $C_{\mathsf {c}}(G)$ which is continuous with respect to the inductive topology. This notion corresponds to the classical concept of a Radon measure (see [Reference Richard and Strungaru35, Appendix]). For the case $G={\mathbb R}^d$ , a clear exposition of this is given in [Reference Baake and Grimm3].

Definition 2.1 A linear functional $\mu : C_{\mathsf {c}}(G) \to {\mathbb C}$ is called a Radon measure (or simply a measure) if for each compact set $K \subseteq G$ there exists a constant $C_K$ such that, for all $\varphi \in C_{\mathsf {c}}(G)$ with $\mbox {supp}(\varphi ) \subseteq K$ , we have

$$ \begin{align*} \left| \mu(\varphi) \right| \leq C_K \| \varphi \|_{\infty} \,. \end{align*} $$

We will often write $\int _G \varphi (t) d \mu (t)$ instead of $\mu (\varphi )$ .

A measure $\mu $ is called positive if for all $\varphi \in C_{\mathsf {c}}(G)$ with $\varphi \geq 0$ we have $\mu (\varphi ) \geq 0$ .

By the Riesz representation theorem [Reference Rudin37], a positive Radon measure is simply a positive regular Borel measure. Moreover, each Radon measure is a linear combination of (at most four) positive Radon measures [Reference Richard and Strungaru35, Appendix].

Next, we review the total variation of a measure.

Definition 2.2 Given a measure $\mu $ , we can define [Reference Pedersen32, Reference Reiter and Stegeman33, Reference Richard and Strungaru35] a positive measure $\left | \mu \right |$ , called the total variation of $\mu $ , such that, for all $\varphi \in C_{\mathsf {c}}(G)$ with $\varphi \geq 0$ , we have

$$ \begin{align*} \left| \mu \right| (\varphi) = \sup \{ \left|\mu(\psi) \right| : \psi \in C_{\mathsf{c}}(G), \mbox{ with } |\psi| \leq \varphi \} \,. \end{align*} $$

We are now ready to introduce the concept of translation boundedness for measures and norm almost periodicity.

Definition 2.3 Let $A \subseteq G$ be a fixed precompact set with nonempty interior. We define the A-norm of $\mu $ via

$$ \begin{align*} \| \mu \|_A := \sup_{x \in G} \left| \mu \right| (x+A) \,. \end{align*} $$

A measure $\mu $ is called translation bounded if $\| \mu \|_A <\infty $ .

Remark 2.4 ([Reference Baake and Moody7, Reference Spindeler and Strungaru42])

Different precompact sets $A_1, A_2$ with nonempty interior define equivalent norms. Therefore, the definition of translation boundedness does not depend on the choice of A.

This allows us to define

$$ \begin{align*} {\mathcal M}^{\infty}(G) := \{ \mu : \mu \mbox{ is a translation bounded measure} \} \,. \end{align*} $$

Then $({\mathcal M}^{\infty }(G), \|. \|_A)$ is a normed space. It is in fact a Banach space [Reference Richard and Strungaru36].

Next, we review the definition of norm almost periodicity as introduced in [Reference Baake and Moody7].

Definition 2.5 Let $ A \subseteq G$ be a fixed precompact set with nonempty interior. A measure $\mu \in {\mathcal M}^{\infty }(G)$ is called norm almost periodic if, for each $\varepsilon>0$ , the set

$$ \begin{align*} P_{\varepsilon}^A (\mu) := \{ t \in G : \| T_t \mu -\mu \|_A < \varepsilon \} \, \end{align*} $$

of $\varepsilon $ -norm almost periods of $\mu $ is relatively dense.

As discussed above, different precompact sets define equivalent norms. This means that while the set of $\varepsilon $ -norm almost periods on $\mu $ depends on the choice of A, the almost periodicity of $\mu $ is independent of this choice.

Any norm almost periodic measure is strongly almost periodic [Reference Baake and Moody7], and the two concepts are equivalent for measures with Meyer set support [Reference Baake and Moody7]. In general, norm almost periodicity is an uniform version of strong almost periodicity [Reference Spindeler and Strungaru42, Theorem 4.7]. The class of norm almost periodic pure point measure was studied in detail and characterized in [Reference Strungaru, Baake and Grimm46].

Let us next recall positive-definiteness for functions and measures. For more details, we recommend [Reference Berg and Forst8, Reference Moody, Strungaru, Baake and Grimm31].

Definition 2.6 A function $f : G\longrightarrow {\mathbb C}$ is called positive-definite if, for all $ n \in \mathbb N$ and all $x_1,\ldots , x_n\in G$ , the matrix $\left (f (x_k - x_l)\right )_{k,l=1,\ldots , n}$ is positive Hermitian. This is equivalent to

$$\begin{align*}\sum_{k,l=1}^n \overline{c_l} f (x_k - x_l) c_k \geq 0 \qquad \forall n \in \mathbb N, x_1,\ldots, x_n\in G, c_1, \ldots , c_n \in {\mathbb C} \,. \end{align*}$$

A measure $\mu $ is called positive-definite if, for all $ \varphi \in C_{\mathsf {c}}(G),$ we have

$$ \begin{align*} \mu (\varphi * \widetilde{\varphi}) \geq 0 \,.\end{align*} $$

This is equivalent to $\mu *\varphi *\tilde {\varphi }$ being a positive-definite function for all $\varphi \in C_{\mathsf {c}}(G)$ [Reference Berg and Forst8, Reference Moody, Strungaru, Baake and Grimm31].

We complete the subsection by reviewing the notion of Fourier transformability for measures. For a more detailed review of the subject, we recommend [Reference Moody, Strungaru, Baake and Grimm31].

Definition 2.7 A measure $\mu $ on G is called Fourier transformable if there exists a measure $\widehat {\mu }$ on $\widehat {G}$ such that, for all $\varphi \in K_2(G)$ , we have $|\check {\varphi }| \in L^1( |\widehat {\mu } |)$ and

2.2 Cut-and-project schemes and Meyer sets

In this part, we review some notions related to the cut-and-project formalism. For more details, we recommend [Reference Baake and Grimm3, Reference Moody and Moody28, Reference Moody, Axel, Dénoyer and Gazeau29].

Definition 2.8 By a CPS, we understand a triple $(G, H, {\mathcal L})$ consisting of a second countable LCAG G, an LCAG H, and a lattice ${\mathcal L} \subseteq G \times H$ such that:

  1. (i) $\pi _H({\mathcal L})$ is dense in H.

  2. (ii) The restriction $\pi _G|_{\mathcal L}$ of the first projection $\pi _G$ to ${\mathcal L}$ is one to one.

Given a CPS $(G,H, {\mathcal L})$ , we will denote by $L:= \pi _G({\mathcal L})$ . Then, $\pi _G$ induces a group isomorphism between ${\mathcal L}$ and L. Composing the inverse of this with the second projection $\pi _H$ , we get a mapping

$$ \begin{align*} \star : L \to H, \end{align*} $$

which we will call the $\star $ -mapping. We then have

$$ \begin{align*} {\mathcal L} = \{ (x,x^{\star}) : x \in L \} \,. \end{align*} $$

Given a CPS $(G, H, {\mathcal L})$ and a subset $W \subseteq H$ we can define

$$\begin{align*}{\LARGE \curlywedge}(W):= \{ x \in L : x^{\star} \in W \} \,. \end{align*}$$

When W is precompact, we will call ${\LARGE \curlywedge }(W)$ a weak model set. If furthermore W has nonempty interior ${\LARGE \curlywedge }(W)$ is called a model set.

Next, let us review the concept of a Meyer set, which plays a fundamental role in this paper.

Definition 2.9 A set $\Lambda \subseteq G$ is called a Meyer set if $\Lambda $ is relatively dense and $\Lambda -\Lambda -\Lambda $ is uniformly discrete.

For equivalent characterizations of Meyer sets, see [Reference Lagarias18, Reference Lagarias, Baake and Moody19, Reference Meyer26, Reference Moody and Moody28, Reference Strungaru, Baake and Grimm46]. Of importance to us will be the following result.

Theorem 2.10 ([Reference Strungaru, Baake and Grimm46])

Let $\Lambda \subseteq G$ be relatively dense. Then $\Lambda $ is Meyer if and only if it is a subset of a (weak) model set.

Moreover, if $\Lambda $ is Meyer, it is a subset of a weak model set in a CPS $(G,H, {\mathcal L})$ with metrizable and compactly generated H.

We should emphasize here that the key for all results below is that fact that a Meyer set is a subset of a model set, and relative denseness plays no role. Because of this, in the spirit of [Reference Strungaru49], we will refer to an arbitrary subset of a (weak) model set as a weak Meyer set. It is obvious that a subset of a weak Meyer set is a weak Meyer set and that a measure is supported inside a Meyer set if and only if its support is a weak Meyer set.

Given a CPS $(G,H, {\mathcal L})$ , the map

(2.1) $$ \begin{align} L \ni x \to (x,x^{\star}) \in {\mathcal L} \, \end{align} $$

is a group isomorphism, and hence it induces an isomorphism between the spaces of (bounded) functions on L and ${\mathcal L}$ , respectively. Since ${\mathcal L}$ is a discrete group, the space of (translation bounded) measures on ${\mathcal L}$ can be identified with the space of (bounded) functions on ${\mathcal L}$ . On another hand, L is typically dense in G, and many functions on L do not induce pure point measures on G.

For us, of interest will be measures supported inside weak model sets ${\LARGE \curlywedge }(W)$ . Since ${\LARGE \curlywedge }(W)$ is uniformly discrete [Reference Moody and Moody28], the space of (translation bounded) measures on ${\LARGE \curlywedge }(W)$ can be identified with the space of (bounded) functions on ${\LARGE \curlywedge }(W)$ , and corresponds via the above isomorphism with the spaces of (translation bounded) measures or (bounded) functions on ${\mathcal L}$ , respectively, that are supported inside $G \times W$ .

Our focus in this paper is on these two spaces. We will study them as spaces of measures, and we will be interested in the relation between the Fourier theory of these two spaces, and the behavior of the Fourier transform with respect to the isomorphism induced by (2.1). For this reason, let us introduce the following notations.

Given a CPS $(G,H, {\mathcal L})$ and a compact set W, we denote by

$$ \begin{align*} {\mathcal M}^{\infty}({\LARGE \curlywedge}(W)) &:= \{ \mu \in {\mathcal M}^{\infty}(G) : \mbox{supp}(\mu) \subseteq {\LARGE \curlywedge}(W)\} \,;\\ {\mathcal M}^{\infty}({\mathcal L}; W) &:=\{ \nu \in {\mathcal M}^{\infty}(G \times H) : \mbox{supp}(\nu) \subseteq \left( {\mathcal L} \cap (G \times W) \right) \} \,. \end{align*} $$

The isomorphism (2.1) induces a bijection $f: {\LARGE \curlywedge }(W) \to {\mathcal L} \cap (G \times W)$ . This induces a bijective map $\mathbb {L}_{G,H,{\mathcal L}, W}: {\mathcal M}^{\infty }({\LARGE \curlywedge }(W)) \to {\mathcal M}^{\infty }({\mathcal L}; W)$ , taking a measure on ${\LARGE \curlywedge }(W)$ into its pushforward via f, defined by

$$\begin{align*}\mathbb{L}_{G,H,{\mathcal L}, W}(\mu)= \sum_{(x,x^{\star}) \in {\mathcal L}} \mu(\{x \}) \delta_{(x,x^{\star})} \,, \end{align*}$$

with inverse $\mathbb {P}_{G,H,{\mathcal L}, W}: {\mathcal M}^{\infty }({\mathcal L} ;W) \to {\mathcal M}^{\infty }({\LARGE \curlywedge }(W))$

$$\begin{align*}\mathbb{P}_{G,H,{\mathcal L}, W}(\nu)= \sum_{(x,x^{\star}) \in {\mathcal L}} \nu(\{(x,x^{\star}) \}) \delta_{x} \,. \end{align*}$$

Let us note here in passing that $\mathbb {P}_{G,H,{\mathcal L},W}$ is simply the pushforward via $f^{-1}$ .

We will refer to these mappings as the lift operator and the projection operator, respectively. When the CPS and window are clear from the context, we will simply write $\mathbb {L}(\mu )$ and $\mathbb {P}(\nu )$ , respectively, instead of $\mathbb {L}_{G,H,{\mathcal L}, W}(\mu )$ and $\mathbb {P}_{G,H,{\mathcal L}, W}(\nu )$ , respectively.

The main results in this paper are that these operators are bijections between the subspaces of Fourier transformable (or cones of positive-definite) measures, and relate their Fourier transforms.

To understand the connection between the Fourier transforms, let us recall the notion of dual CPS. Given a CPS $(G, H, {\mathcal L})$ , we can define

$$ \begin{align*} {\mathcal L}^0 := \{ (\chi, \psi) \in \widehat{G} \times \widehat{H} : \chi(x)\psi(x^{\star}) =1 \, \forall x \in L \} \,. \end{align*} $$

Then, $(\widehat {G}, \widehat {H}, {\mathcal L}^0)$ is a CPS [Reference Baake, Huck and Strungaru5, Reference Moody and Moody28, Reference Moody, Axel, Dénoyer and Gazeau29, Reference Strungaru, Baake and Grimm46]. We will refer to this as the CPS dual to $(G, H, {\mathcal L})$ .

3 Positive-definite measures with weak Meyer set support

In this section, we show that $\mathbb {L}_{G,H,{\mathcal L}, W}$ and $\mathbb {P}_{G,H,{\mathcal L}, W}$ take positive-definite measures to positive-definite measures.

Let us start with the following obvious lemma, which follows immediately from Definition 2.6 and the fact that the function from (2.1) is a group isomorphism.

Lemma 3.1 Let $(G, H, {\mathcal L})$ be a CPS, and let $f : L \to {\mathbb C}$ be a function. Define ${g :{\mathcal L} \to {\mathbb C}}$ via

$$\begin{align*}g(x,x^{\star}):= f(x) \,. \end{align*}$$

Then f is positive-definite on L if and only if g is positive-definite on ${\mathcal L}$ .

Let us recall now the following result of [Reference Lenz and Strungaru24], which we will use often in the paper.

Proposition 3.2 ([Reference Lenz and Strungaru24, Proposition 2.4])

Let G be a LCAG, let $\mu $ be a discrete measure on G, and let

$$\begin{align*}f(x):= \mu (\{ x \}) \,. \end{align*}$$

Then, the following are equivalent:

  1. (i) The measure $\mu $ is a positive-definite measure on G.

  2. (ii) The measure $\mu $ is a positive-definite measure on $G_{d}$ .

  3. (iii) The function f is a positive-definite function on G.

  4. (iv) The function f is a positive-definite function on $G_{d}$ .

Next, we prove a slight generalization of [Reference Lenz and Strungaru24, Lemma 2.10] and [Reference Baake and Grimm3, Lemma 8.4].

Lemma 3.3 Let $\gamma $ be a positive-definite pure point measure on G, and let L be any subgroup of G. Then, the function $g : L \to {\mathbb C}$ defined via

$$ \begin{align*} g(x):= \gamma ( \{ x \}) \, \end{align*} $$

is a positive-definite function on L.

Proof Define $f : G \to {\mathbb C}$ via

$$ \begin{align*} f(x):= \gamma ( \{ x \}) \,. \end{align*} $$

Then f is a positive-definite function on G Proposition 3.2. Definition 2.6 immediately gives that the restriction $g=f|_L$ to the subgroup L is a positive-definite function on L.

We will also need the following result.

Lemma 3.4 Let G be any group, and let $H \leq G$ be a subgroup. Let $f : H \to {\mathbb C}$ be a positive-definite functions. Then, the function $g : G \to {\mathbb C}$ defined via

$$ \begin{align*} g(x):= \left\{ \begin{array}{l c} f(x), & \mbox{ if } x \in H, \\ 0, & \mbox{ otherwise } \end{array} \right. \end{align*} $$

is positive-definite on G.

Proof Let $n \in \mathbb N, x_1,\ldots , x_n\in G$ and $c_1,..,c_n \in {\mathbb C}$ . Note that $g(x_k-x_l) =0$ whenever $x_k -x_l \notin H$ .

On G define the standard equivalence $\pmod {H}$ as

$$ \begin{align*} x \equiv y\quad \pmod{H} \Leftrightarrow x-y \in H \,. \end{align*} $$

This induces an equivalence relation on the set $\{ x_1,\ldots , x_n \}$ , and hence we can partition this set in equivalence classes $F_1,\ldots , F_m$ .

To make the computation clearer, define $c : G\to {\mathbb C}$

$$\begin{align*}c(x):= \left\{ \begin{array}{cc} c_j, & \mbox{ if } x=x_j, \\ 0, & \mbox{ otherwise. } \end{array} \right.\, \end{align*}$$

Then,

$$ \begin{align*} \sum_{k,l=1}^n g(x_k-x_l)c_k \overline{c_l} &=\sum_{k,l=1}^n g(x_k-x_l)c(x_k) \overline{c(x_l)} \\ &= \sum_{i=1}^m \sum_{j=1}^m \left( \sum_{x \in F_i} \sum_{y \in F_j} g(x-y) c(x) \overline{c(y)} \right)\\ &= \sum_{i=1}^m \left( \sum_{x,y \in F_i} g(x-y)c(x)\overline{c(y)} \right) \,. \end{align*} $$

Now, fix some $1 \leq i \leq m$ , and let $F_i:=\{ z_1,.., z_q \}$ . Then,

$$ \begin{align*} \sum_{x,y \in F_i} g(x-y)c(x)\overline{c(y)}&= \sum_{r,s=0}^q g(z_r-z_s)c(z_r)\overline{c(z_s)} \\ &= \sum_{r,s=1}^q g(z_r-z_s)c(z_r)\overline{c(z_s)} \\ &= \sum_{r,s=1}^q f\left((z_r-z_1) -(z_s-z_1)\right)c(z_r)\overline{c(z_s)} \geq 0 \end{align*} $$

by the positive-definiteness of f applied to $m; y_1:=z_1 - z_1; y_2:=z_2 - z_1 ;\ldots; y_q:=z_q - z_1 \in H$ and $c^{\prime }_1=c(z_1),\ldots ,c^{\prime }_q=c(z_q)$ .

Therefore, for each i, we have $\sum _{x,y \in C_i} g(x-y)c(x)\overline {c(y)} \geq 0$ , and hence

$$ \begin{align*} \sum_{k,l=1}^n g(x_k-x_l)c_k \overline{c_l}=\sum_{i=1}^m \left( \sum_{x,y \in C_i} g(x-y)c(x)\overline{c(y)} \right) \geq 0.\\[-40pt] \end{align*} $$

Remark 3.5 One can also prove Lemma 3.4 by using Fourier analysis. Indeed, since f is positive-definite, the measure $\mu :=f \theta _{H_{\mathsf {d}}}$ is a positive-definite measure on the discrete group $H_{\mathsf {d}}$ [Reference Argabright and de Lamadrid1, Corollary 4.3]. Then, it is Fourier transformable on $H_{\mathsf {d}}$ and its Fourier transform is positive [Reference Argabright and de Lamadrid1, Reference Berg and Forst8]. As $H_{\mathsf {d}}$ is closed in the discrete group $G_{\mathsf {d}}$ , by [Reference Argabright and de Lamadrid1, Theorem 4.2], the measure $\nu :=g \theta _{G_{\mathsf {d}}}$ is Fourier transformable on $G_{\mathsf {d}}$ and has positive Fourier transform. Then, $\mu $ is positive-definite [Reference Argabright and de Lamadrid1, Theorem 4.1]. Therefore, by Proposition 3.2, g is positive-definite on G.

We are now ready to prove the following result.

Proposition 3.6 Let $(G,H, {\mathcal L})$ be a CPS, let ${\LARGE \curlywedge }(W)$ be a weak model set, and let $f : G \to {\mathbb C}$ be a function which vanishes outside ${\LARGE \curlywedge }(W)$ . Let

$$ \begin{align*} \gamma &=\sum_{x \in {\LARGE \curlywedge}(W)} f(x) \delta_x, \\ \eta&= \sum_{(x,x^{\star}) \in {\mathcal L}} f(x) \delta_{(x,x^{\star})}= \mathbb{L} (\gamma) \,. \end{align*} $$

Then $\gamma $ is a positive-definite measure on G if and only if $\eta $ is a positive-definite measure on $G \times H$ .

Proof $\Rightarrow $ : Denote as usual $L := \pi _G({\mathcal L})$ . Define $g: L \to {\mathbb C}$ via

$$ \begin{align*} g(x):= \gamma ( \{ x \}) \,, \end{align*} $$

that is, $g= f|_{L}$ .

Then, by Lemma 3.3, g is a positive-definite function on L and hence, by Lemma 3.1, the function $h: {\mathcal L} \to {\mathbb C}$

$$ \begin{align*} h(x,x^{\star})=g(x) \end{align*} $$

is a positive-definite function on ${\mathcal L}$ . Therefore, by Lemma 3.4, the function $j: G \times H \to {\mathbb C}$

$$ \begin{align*} j(z):= \left\{ \begin{array}{l c} h(z), & \mbox{ if } z \in {\mathcal L,} \\ 0, & \mbox{ otherwise } \end{array} \right. \end{align*} $$

is positive-definite on $G \times H$ . The claim follows from Proposition 3.2.

$\Leftarrow :$ Since $\eta $ is positive-definite, by Lemma 3.3, the function $h: {\mathcal L} \to {\mathbb C}$ defined by

$$ \begin{align*} h(x,x^{\star}):= \eta(\{(x,x^{\star}) \}) = f(x) \,, \end{align*} $$

is positive-definite on ${\mathcal L}$ and hence, by Lemma 3.1, the restriction $g=f|_L$ is positive-definite on L. As f is zero outside ${\LARGE \curlywedge }(W) \subseteq L$ , it follows from Lemma 3.4 that f is a positive-definite function on G. The claim follows now from Proposition 3.2.

Remark 3.7

  1. (a) In Proposition 3.6, the positive-definiteness of $\eta $ and $\gamma $ is equivalent to the positive-definiteness of the function f.

  2. (b) Denoting by

    $$ \begin{align*} \mathcal{PD}({\LARGE \curlywedge}(W)) &:= \{ \mu \in {\mathcal M}^{\infty}({\LARGE \curlywedge}(W)) : \mu \mbox{ is positive-definite } \}, \\ \mathcal{PD}({\mathcal L}; W) &:=\{ \nu \in {\mathcal M}^{\infty}({\mathcal L}, W) : \nu \mbox{ is positive-definite } \} \,. \end{align*} $$

    Proposition 3.6 says that

    $$ \begin{align*} \mathbb{L}( \mathcal{PD}({\LARGE \curlywedge}(W)) ) &=\mathcal{PD}({\mathcal L}; W), \\ \mathbb{P} (\mathcal{PD}({\mathcal L}; W) )&= \mathcal{PD}({\LARGE \curlywedge}(W)) \,.\end{align*} $$

4 The lift of Fourier transformable measures

We can now prove that, given a CPS $(G, H,{\mathcal L})$ and a compact set K, the lifting operator induces a bijection between the space of Fourier transformable measures supported inside ${\LARGE \curlywedge }(W)$ and the space of Fourier transformable measures supported inside $ {\mathcal L} \cap (G \times W)$ .

Theorem 4.1 Let $(G,H, {\mathcal L})$ be a CPS, and let $W \subseteq H$ be compact. Let $\gamma $ be a translation bounded measure supported inside ${\LARGE \curlywedge }(W)$ , and let

$$\begin{align*}\eta:= \mathbb{L}_{G,H,{\mathcal L}, W}(\gamma) \,. \end{align*}$$

Then $\gamma $ is Fourier transformable if and only if $\eta $ is Fourier transformable.

Moreover, if $\varphi \in K_2(H)$ is any function so that $\varphi \equiv 1$ on W, then, for all $\psi \in C_{\mathsf {c}}(\widehat {G})$ , we have $\psi \otimes \hat {\varphi } \in L^1(\widehat {\eta })$ and

(4.1) $$ \begin{align} \widehat{\gamma}(\psi)= \widehat{\eta}( \psi \otimes \check{\varphi})=: (\widehat{\eta})_{\check{\varphi}}(\psi) \,. \end{align} $$

Proof $\Longrightarrow $ By [Reference Strungaru47, Lemma 8.3], there exist a compact set $W \subseteq K$ and four positive-definite measures $\omega _{1},\omega _{2},\omega _{3},\omega _{4}$ supported inside ${\LARGE \curlywedge }(K)$ such that

$$\begin{align*}\gamma=\omega_1-\omega_2+i\omega_3-i \omega_4 \,. \end{align*}$$

Then, we have

$$ \begin{align*} \eta&= \mathbb{L}_{G,H,{\mathcal L}, W}(\gamma) =\mathbb{L}_{G,H,{\mathcal L}, K}(\gamma)=\mathbb{L}_{G,H,{\mathcal L}, K}(\omega_1-\omega_2+i\omega_3-i \omega_4 )\\ &= \mathbb{L}_{G,H,{\mathcal L}, K}(\omega_1)-\mathbb{L}_{G,H,{\mathcal L}, K}(\omega_2)+i\mathbb{L}_{G,H,{\mathcal L}, K}(\omega_3)-i\mathbb{L}_{G,H,{\mathcal L}, K}(\omega_4) \,. \end{align*} $$

Now, by Proposition 3.6, for all $1 \leq j \leq 4$ , the measure $\mathbb {L}_{G,H,{\mathcal L}, K}(\omega _j)$ is positive-definite. Therefore, as a linear combination of positive-definite measures, $\eta $ is Fourier transformable.

$\Longleftarrow $ . Our argument is similar to [Reference Richard and Strungaru34].

First, fix an arbitrary $\varphi \in K_2(H)$ so that $\varphi \equiv 1$ on W. We split the rest of the argument into two steps.

Step 1: We show that $(\widehat {\eta })_{\check {\varphi }}$ is a measure.

Let us first note that for all $\psi \in K_2(G)$ , we have $\psi \otimes \varphi \in K_2(G \times H)$ . Therefore, since $\eta $ is Fourier transformable, we have

(4.2) $$ \begin{align} \left| \check{\psi} \otimes \check{\varphi} \right| \in L^1(|\widehat{\eta}|) \,. \end{align} $$

We now show that for all $\phi \in C_{\mathsf {c}}(\widehat {G})$ , we have $\left | \phi \otimes \check {\varphi } \right | \in L^1(|\widehat {\eta }|)$ and that

$$\begin{align*}(\widehat{\eta})_{\check{\varphi}}(\phi):= \widehat{\eta}( \phi \otimes \check{\varphi}) \, \end{align*}$$

defines a measure.

Let $K \subseteq \widehat {G}$ be a fixed compact set. Then, there exists some $\psi \in K_2(G)$ such that $\check {\psi } \geq 1_K$ [Reference Berg and Forst8, Reference Moody, Strungaru, Baake and Grimm31].

Now, for all $\phi \in C_{\mathsf {c}}(\widehat {G})$ with $\mbox {supp}(\phi ) \subseteq K$ , we have

(4.3) $$ \begin{align} |\widehat{\eta}| \left(|\phi \otimes \check{\varphi}| \right)=\int_{\hat{G} \times \hat{H}} \left| \phi(s)\right| \cdot \left| \check{\varphi}(t) \right| \mbox{d} |\widehat{\eta}|(s,t) \leq \| \phi \|_{\infty} \int_{\hat{G} \times \hat{H}} \left| \check{\psi}(s)\right| \cdot \left| \widehat{\varphi}(t) \right| \mbox{d} |\widehat{\eta}|(s,t) < \infty, \end{align} $$

and hence $(\widehat {\eta })_{\check {\varphi }}$ is well defined.

Moreover, for all $\phi \in C_{\mathsf {c}}(\widehat {G})$ with $\mbox {supp}(\phi ) \subseteq K$ , it follows from (4.3) that

$$\begin{align*}\left| (\widehat{\eta})_{\check{\varphi}}(\phi) \right| \leq C_K \| \phi \|_{\infty}, \end{align*}$$

where

This shows that $(\widehat {\eta })_{\check {\varphi }}$ is a measure.

Step 2: We show that for all $\phi \in K_2(G)$ , we have

and

Let $\phi \in K_2(G)$ be arbitrary.

Since G is second countable, so is $\widehat {G}$ [Reference Reiter and Stegeman33]. In particular, $\widehat {G}$ is $\sigma $ -compact [Reference Reiter and Stegeman33]. Therefore, there exists a sequence $K_n$ of compact sets with $K_n \subseteq (K_{n+1})^{\circ }$ such that

$$\begin{align*}\widehat{G}= \bigcup_{n} K_n.\\[-20pt] \end{align*}$$

Let $\psi _n \in C_{\mathsf {c}}(\widehat {G})$ be so that $1_{K_n} \leq \psi _n \leq 1_{K_{n+1}}$ .

Then, $\psi _n \widehat {\phi } \in C_{\mathsf {c}}(\widehat {G})$ and by the definition of

, we have

Now, for all n, we have by (4.2)

Therefore, by the dominated convergence theorem [Reference Reiter and Stegeman33, Theorem 3.2.51], we have

(4.4)

Next, by the monotone convergence theorem [Reference Reiter and Stegeman33], we have

Note that for each n, we have

Since

, we get

This shows that

. Therefore,

is dominated by

and converges pointwise to

. Thus, by (4.4) and the dominated convergence theorem, we get

Finally, by the Fourier transformability of $\eta $ , we have

Therefore, we proved that for all $\phi \in K_2(G)$ , we have

and

This proves that $\gamma $ is Fourier transformable and

$$\begin{align*}\widehat{\gamma}= (\widehat{\eta})_{\check{\varphi}} \,, \end{align*}$$

completing the proof.

Using the fact that $\mathbb {L}$ is a bijection with inverse $\mathbb {P}$ , we get the following corollary.

Corollary 4.2 Let $(G,H, {\mathcal L})$ be a CPS, and let $W \subseteq H$ be compact. Let $\eta $ be a translation bounded measure supported inside ${\mathcal L} \cap (G \times W)$ , and let $\gamma = \mathbb {P}_{G,H,{\mathcal L}, W}(\eta )$ . Then $\eta $ is Fourier transformable if and only if $\gamma $ is Fourier transformable.

Moreover, if $\varphi \in K_2(H)$ is any function so that $\varphi \equiv 1$ on W, then, for all $\psi \in C_{\mathsf {c}}(\widehat {G})$ , we have $\psi \otimes \hat {\varphi } \in L^1(\widehat {\eta })$ and (4.1) holds.

5 Applications

In this section, we will discuss the relation (4.1) and how can it be used to (re)derive some results from [Reference Strungaru47].

To make the things easier to follow, we introduce the notion of strongly admissible functions.

5.1 Strongly admissible functions for CPS

Let us start with the following definition.

Definition 5.1 Given a group H of the form $H={\mathbb R}^d \times H_0$ , with a LCAG $H_0$ , a function $f : H \to {\mathbb C}$ is called strongly admissible if there exists $g\in C_{\mathsf {u}}({\mathbb R}^d)$ and $\varphi \in C_{\mathsf {c}}(H_0)$ such that:

  • $\| (1+|x|^{2d}) g \|_{\infty } < \infty $ .

  • $f= g \otimes \varphi $ .

Next, given a CPS $(G,H,{\mathcal L})$ , we will denote by ${\mathcal M}_{{\mathcal L}}(G \times H)$ , the space of ${\mathcal L}$ -periodic measures on $G \times H$ . Note that by [Reference Lenz and Richard21, Proposition 6.1]

$$\begin{align*}{\mathcal M}_{{\mathcal L}}(G \times H) \subseteq {\mathcal M}^{\infty}(G \times H) \,. \end{align*}$$

We will see below that given a Fourier transformable measure $\gamma $ with weak Meyer set support, Theorem 4.1 can be used to create a CPS $(G, H={\mathbb R}^d \times H_0, {\mathcal L})$ , an ${\mathcal L}^0$ -periodic measure $\rho (=\widehat {\eta })$ and a strongly admissible function f on $\widehat {H} ={\mathbb R}^d \times \widehat {H_0}$ such that, equation (4.1) yields

$$\begin{align*}\gamma= (\rho)_{f} \,. \end{align*}$$

This motivates us to closely look at the properties of $(\rho )_f$ , for a CPS $(G, H={\mathbb R}^d \times H_0, {\mathcal L}), \rho \in {\mathcal M}_{{\mathcal L}}(G \times H)$ and strongly admissible functions f.

Let us start with the following simple observation which also explains the name “strongly admissible.”

Given a CPS $(G, H={\mathbb R}^d \times H_0, {\mathcal L})$ , a measure $\rho \in {\mathcal M}_{{\mathcal L}}(G \times H)$ , and strongly admissible function f, it is obvious that the function f is admissible for $(G, H, {\mathcal L},\rho )$ in the sense of [Reference Lenz and Richard21, Definition 3.1]. Therefore, by [Reference Lenz and Richard21, Proposition 6.3], we can define a translation bounded measure $\rho _f$ on G via

$$\begin{align*}\rho_f(\phi):= \rho (\phi \otimes f) \qquad \forall \phi \in C_{\mathsf{c}}(G) \,. \end{align*}$$

This measure is strongly almost periodic by [Reference Lenz and Richard21, Theorem 3.1]. In fact, the strong admissibility of f immediately implies that $\rho _f$ is norm almost periodic.

Indeed, let $(G, H={\mathbb R}^d \times H_0, {\mathcal L})$ , let $f=g \otimes \varphi $ be strongly admissible, and let $\rho \in {\mathcal M}_{{\mathcal L}}(G \times H)$ . Pick any compact set $ \mbox {supp}(\varphi ) \subseteq W \subseteq \widehat {H_0}$ , and let $K, K_1 \subseteq G$ be compact sets in G with $K \subseteq K_1^{\circ }$ . Then, a standard computation similar to [Reference Strungaru47, Lemma 5.2] shows that

$$\begin{align*}\| (\rho)_f \|_{K} \leq C \| \varphi \|_{\infty} \| (1+|x|^{2d}) g \|_{\infty} \| \rho \|_{K_1 \times [-\frac{1}{2},\frac{1}{2}]^d \times W } \,, \end{align*}$$

where

$$ \begin{align*} C:= \left( \sum_{n \in {\mathbb Z}^d} \sup_{z \in n+[-\frac{1}{2},\frac{1}{2}]^d } \frac{1}{1+|z|^{2d}} \right) < \infty \,. \end{align*} $$

This immediately gives the following stronger version of [Reference Strungaru47, Lemma 5.2].

Fact 5.2 Let $(G, H={\mathbb R}^d \times H_0, {\mathcal L})$ be a CPS, let $\rho \in {\mathcal M}_{{\mathcal L}}(G \times H)$ , and let $f \in C^{}_{0}(H)$ be strongly admissible. Then, $\rho _f$ is a norm almost periodic measure.

5.2 Fourier transform of measures with weak Meyer set support

Fix an arbitrary Meyer set $\Lambda $ and a Fourier transformable measure $\gamma $ with $\mbox {supp}(\gamma ) \subseteq \Lambda $ .

By Theorem 2.10 and the structure theorem of compactly generated groups, there exists a CPS $(G, {\mathbb R}^d \times {\mathbb Z}^m \times \mathbb K, {\mathcal L})$ with compact $\mathbb K$ and a compact $W \subseteq {\mathbb R}^d \times {\mathbb Z}^m \times \mathbb K$ such that

$$\begin{align*}\Lambda \subseteq {\LARGE \curlywedge}(W) \,. \end{align*}$$

By eventually enlarging W, we can assume without loss of generality that

$$\begin{align*}W=W_0 \times F \times \mathbb K \end{align*}$$

for compact $W_0 \subseteq {\mathbb R}^d$ and finite $F \subseteq {\mathbb Z}^m$ .

Set $H_0={\mathbb Z}^m \times \mathbb K$ . It is easy to see that we can find function $\varphi \in C_{\mathsf {c}}^{\infty }({\mathbb R}^d)\cap K_2({\mathbb R}^d)$ and $\psi \in K_2(H_0)$ with the following properties:

  • $\phi := \varphi \otimes \psi \equiv 1$ on W.

  • $\widehat {\psi } \in C_{\mathsf {c}}(\widehat {H_0})$ .

It follows that

(5.1) $$ \begin{align} f:=\check{\phi}= \check{\varphi} \otimes \check{\psi} \end{align} $$

is a strongly admissible function of $\widehat {H} ={\mathbb R}^d \times \widehat {H_0}$ .

Next, define $\eta :=\mathbb {L}_{G, {\mathbb R}^d \times H,{\mathcal L}, W}(\gamma )$ . Then, by Theorem 4.1, $\eta $ is Fourier transformable. Moreover, since $\mbox {supp}(\eta ) \subseteq {\mathcal L}$ , the measure $\rho =\widehat {\eta }$ is ${\mathcal L}^0$ -periodic by [Reference de Lamadrid and Argabright10, Proposition 6.1]. Finally, (4.1) gives

(5.2) $$ \begin{align} \widehat{\gamma}=(\rho)_{f} \,. \end{align} $$

Fact 5.2 then gives the following result.

Corollary 5.3 ([Reference Strungaru47, Theorem 7.1])

Let $\gamma $ be a measure with weak Meyer set support. Then, $\widehat {\gamma }$ is norm almost periodic.

5.3 Generalized Eberlein decomposition

In this subsection, we show a pseudo-compatibility of the mapping $\rho \to (\rho )_f$ of (5.3), for ${\mathcal L}$ periodic $\rho \in {\mathcal M}_{{\mathcal L}}(G \times H)$ and strongly admissible f, with respect to the Lebesgue decomposition. We explain this, as well as our meaning of “pseudo-compatibility” below.

First, it is easy to see that the map satisfies:

  • if $\rho $ is pure point, then $(\rho )_f$ is pure point;

  • if $\rho $ is absolutely continuous, then $(\rho )_f$ is absolutely continuous;

  • if $\rho $ is singular continuous, then $(\rho )_f$ can have all three spectral components;

and hence does not preserves the Lebesgue decomposition. On another hand, for each $\alpha \in \{\mathsf {pp}, \mathsf {ac}, \mathsf {sc}\}$ , one can defined an operator $P_{\alpha }$ on the space ${\mathcal M}_{{\mathcal L}}^{\infty }(G \times H)$ with the property that for all strongly admissible f and all $\rho \in {\mathcal M}_{{\mathcal L}}^{\infty }(G \times H)$ , we have

(5.3) $$ \begin{align} \left( P_{\alpha} (\rho)\right)_{f}= \left((\rho)_{f}\right)_{\alpha} \,. \end{align} $$

This can be done simply by first showing that

$$\begin{align*}L_{\alpha}(\sum_{j=1}^m c_j \psi_j \otimes \phi_j) := \sum_{j=1}^m c_j \left( \rho_{\phi_j} \right)_{\alpha}(\psi_j) \end{align*}$$

for all $c_1,\ldots ,c_m \in {\mathbb C}, \psi _1,\ldots , \psi _m \in C_{\mathsf {c}}(\widehat {G}), \phi _1,\ldots , \phi _m \in C_{\mathsf {c}}(\widehat {H})$ is well defined, linear, and continuous with respect to the inductive topology. Therefore, $L_{\alpha }$ can be uniquely extended to a measure $P_{\alpha } (\rho )$ , which is ${\mathcal L}$ invariant and satisfies (5.3).

Now, exactly as above, let $\gamma $ be a Fourier transformable measure supported inside a Meyer set $\Lambda $ , and let $(G,H, {\mathcal L}), \eta , \varphi , \phi , \psi $ be as in Section 5.2. Let f be as in (5.1), and let $\rho = \widehat {\eta }$ .

Then, for each $\alpha \in \{\mathsf {pp}, \mathsf {ac}, \mathsf {sc}\}$ , the measure $P_{\alpha } (\rho )$ is the Fourier transform of some measure $\mu $ supported on ${\mathcal L}^0$ [Reference Richard and Strungaru36].

Define

$$\begin{align*}\nu:= \sum_{x \in L} \phi(x^{\star}) \mu(\{ (x,x^{\star}) \}) \delta_x \,. \end{align*}$$

Then, $\mbox {supp}(\nu ) \subseteq {\LARGE \curlywedge }(\mbox {supp}(\phi ))$ and, exactly as in the proof of Theorem 4.1, we get

$$ \begin{align*} \widehat{\nu} &= \left( \widehat{\mu} \right)_{f}= \left( P_{\alpha} (\widehat{\eta}) \right)_{f} =\left((\widehat{\eta})_{h}\right)_{\alpha}= \left( \widehat{\gamma} \right)_{\alpha} \,. \end{align*} $$

Therefore, we get the following corollary.

Corollary 5.4 ([Reference Strungaru47, Theorem 4.1])

Let $\gamma $ be a Fourier transformable measure supported inside a Meyer set $\Lambda $ . Then, there exist a model set $\Gamma \supseteq \Lambda $ and three Fourier transformable measures $\gamma _{\mathsf {s}},\gamma _{\mathsf {0s}}, \gamma _{\mathsf {0a}}$ supported inside $\Gamma $ such that

5.4 Discussion

We have seen in this section that the Fourier transform of a measure $\gamma $ with weak Meyer set support can be describe via (5.2) as the projection in the dual CPS of a ${\mathcal L}^0$ -periodic measure via a strongly admissible function. We used this result to (re)derive properties of $\widehat {\gamma }$ , and we expect that this connection will lead to some new applications in the future. Indeed, while now we know quite a few properties of the Fourier transform of measures with weak Meyer set support [Reference Aujogue2, Reference Strungaru43Reference Strungaru49], we know much more about fully periodic measures in LCAG (see, for example, [Reference Richard and Strungaru36]). Moreover, the strong admissibility of f is likely to transfer many properties from $\rho $ to $\rho _f$ . It is also worth pointing out that, while the strong admissibility of f was sufficient to derive the conclusions in this section, in fact f can be chosen of the form

$$\begin{align*}f := g \otimes P \otimes \psi :{\mathbb R}^d \times \mathbb T^m \times \widehat{\mathbb K} \to {\mathbb C} \,, \end{align*}$$

where $g =\hat {\varphi } \in {\mathcal S}({\mathbb R}^d)$ is the Fourier transform of some $\varphi \in C_{\mathsf {c}}^{\infty }({\mathbb R}^d)$ ; P is a trigonometric polynomial, that is, a sum of characters, that is, $P= \sum _{j=1}^m \chi _j$ for some $\chi _1,\ldots , \chi _j \in \widehat {\mathbb T^m}$ and $\psi \in C_{\mathsf {c}}(\widehat {\mathbb K})$ is the characteristic function of $\{ 0 \}$ . These properties are much stronger than strong admissibility, and have the potential to lead to nice applications in the future.

Acknowledgment

We are grateful to Michael Baake and Christoph Richard for many insightful discussions which inspired this manuscript. We would like to thank the two anonymous referees for some suggestions that improved the quality of this manuscript.

Footnotes

This work was supported by NSERC with grant 2020-00038, and the author is grateful for support.

References

Argabright, L. N. and de Lamadrid, J. G., Fourier analysis of unbounded measures on locally compact abelian groups, Memoirs of the American Mathematical Society, American Mathematical Society, Providence, 145, 1974.CrossRefGoogle Scholar
Aujogue, J.-B., Pure point/continuous decomposition of translation-bounded measures and diffraction. Ergodic Theory Dynam. Systems 40(2020), 309352.CrossRefGoogle Scholar
Baake, M. and Grimm, U., Aperiodic order. Volume 1: A mathematical invitation, Cambridge University Press, Cambridge, 2013.CrossRefGoogle Scholar
Baake, M. and Grimm, U. (eds.), Aperiodic order. Volume 2: Crystallography and almost periodicity, Cambridge University Press, Cambridge, 2017.CrossRefGoogle Scholar
Baake, M., Huck, C., and Strungaru, N., On weak model sets of extremal density . Indag. Math. 28(2017), 331.CrossRefGoogle Scholar
Baake, M. and Lenz, D., Dynamical systems on translation bounded measures: pure point dynamical and diffraction spectra , Ergodic Theory Dynam. Systems 24(2004), 18671893.CrossRefGoogle Scholar
Baake, M. and Moody, R. V., Weighted Dirac combs with pure point diffraction . J. Reine Angew. Math. (Crelle) 573(2004), 6194.Google Scholar
Berg, C. and Forst, G., Potential theory on locally compact abelian groups, Springer, Berlin–New York, 1975.CrossRefGoogle Scholar
Bourbaki, N., General topology, parts I and II, Hermann, Paris, 1966.Google Scholar
de Lamadrid, J. G. and Argabright, L. N., Almost periodic measures, Memoirs of the American Mathematical Society, American Mathematical Society, Providence, 85, no. 428, 1990.CrossRefGoogle Scholar
Gouéré, J.-B., Diffraction and palm measure of point processes . C. R. Math. Acad. Sci. Paris 336(2003), 5762.CrossRefGoogle Scholar
Gouéré, J.-B., Quasicrystals and almost periodicity . Comm. Math. Phys. 255(2005), 655681.CrossRefGoogle Scholar
Hof, A., On diffraction by aperiodic structures . Comm. Math. Phys. 169(1995), 2543.CrossRefGoogle Scholar
Hof, A., Uniform distribution and the projection method . In: Patera, J. (ed.), Quasicrystals and discrete geometry, Fields Institute Monographs, 10, American Mathematical Society, Providence, RI, 1998, pp. 201206.Google Scholar
Keller, G. and Richard, C., Dynamics on the graph of the torus parametrisation . Ergodic Theory Dynam. Systems 28(2018), 10481085.CrossRefGoogle Scholar
Keller, G. and Richard, C., Periods and factors of weak model sets . Israel J. Math. 229(2019), 85132.CrossRefGoogle Scholar
Keller, G., Richard, C., and Strungaru, N., Spectrum of weak model sets with Borel windows, to appear in Canad. Math. Bull. arXiv:2107.08951 Google Scholar
Lagarias, J., Meyer’s concept of quasicrystal and quasiregular sets . Comm. Math. Phys. 179(1996), 365376.CrossRefGoogle Scholar
Lagarias, J. C., Mathematical quasicrystals and the problem of diffraction . In: Baake, M. and Moody, R. V. (eds.), Directions in mathematical quasicrystals, CRM Monograph Series, 13, American Mathematical Society, Providence, RI, 2000, pp. 6193.Google Scholar
Lee, J.-Y., Moody, R. V., and Solomyak, B., Pure point dynamical and diffraction spectra . Ann. Henri Poincaré 3(2002), 10031018.CrossRefGoogle Scholar
Lenz, D. and Richard, C., Pure point diffraction and cut and project schemes for measures: the smooth case . Math. Z. 256(2007), 347378.CrossRefGoogle Scholar
Lenz, D., Spindeler, T., and Strungaru, N., Pure point diffraction and mean, Besicovitch and Weyl almost periodicity. Preprint, 2020. arXiv:2006.10821 Google Scholar
Lenz, D., Spindeler, T., and Strungaru, N., Pure point spectrum for dynamical systems and mean almost periodicity. To appear in Ergodic Theory Dynam. Syst., 2023. arXiv:2006.10825 CrossRefGoogle Scholar
Lenz, D. and Strungaru, N., Note on the set of Bragg peaks with high intensity . Ann. Henri Poincaré 17(2016), 673687.CrossRefGoogle Scholar
Lev, N. and Olevskii, A., Quasicrystals and Poisson’s summation formula . Invent. Math. 200(2015), 585606.CrossRefGoogle Scholar
Meyer, Y., Algebraic numbers and harmonic analysis, North-Holland, Amsterdam, 1972.Google Scholar
Meyer, Y., Quasicrystals, almost periodic patterns, mean-periodic functions and irregular sampling . Afr. Diaspora J. Math. 13(2012), 745.Google Scholar
Moody, R. V., Meyer sets and their duals . In: Moody, R. V. (ed.), The mathematics of long-range aperiodic order, NATO ASI Series, C489, Kluwer, Dordrecht, 1987, pp. 403441.Google Scholar
Moody, R. V., Model sets: a survey . In: Axel, F., Dénoyer, F., and Gazeau, J. P. (eds.), From quasicrystals to more complex systems, Centre de Physique des Houches, 13, Springer, Berlin, 2000, pp. 145166.CrossRefGoogle Scholar
Moody, R. V. and Strungaru, N., Point sets and dynamical systems in the autocorrelation topology . Canad. Math. Bull. 47(2004), no. 1, 8299.CrossRefGoogle Scholar
Moody, R. V. and Strungaru, N., Almost periodic measures and their Fourier transforms . In Baake, M. and Grimm, U. (eds.), Aperiodic order. Volume 2: crystallography and almost periodicity, Cambridge University Press, Cambridge, 2017, pp. 173270.CrossRefGoogle Scholar
Pedersen, G. K., Analysis now, Springer, New York, 1989, Revised printing, 1995.CrossRefGoogle Scholar
Reiter, H. and Stegeman, J. D., Clasical harmonic analysis and locally compact groups, Clarendon Press, Oxford, 2000.CrossRefGoogle Scholar
Richard, C. and Strungaru, N., Pure point diffraction and Poisson Summation . Ann. Henri Poincaré 18(2017), 39033931.CrossRefGoogle Scholar
Richard, C. and Strungaru, N., A short guide to pure point diffraction in cut-and-project sets . J. Phys. A: Math. Theor. 50(2017), no. 15, 154003.CrossRefGoogle Scholar
Richard, C. and Strungaru, N., On the Fourier transformability of measures supported inside lattices, in preparation.Google Scholar
Rudin, W., Real and complex analysis, McGraw-Hill, New York, 1987.Google Scholar
Schlottmann, M., Generalized model sets and dynamical systems . In: Baake, M. and Moody, R. V. (eds.), Directions in mathematical quasicrystals, CRM Monograph Series, American Mathematical Society, Providence, RI, 2000, pp. 143159.Google Scholar
Shechtman, D., Blech, I., Gratias, D., and Cahn, J. W., Metallic phase with long-range orientational order and no translation symmetry . Phys. Rev. Lett. 53(1984), 183185.CrossRefGoogle Scholar
Solomyak, B., Dynamics of self-similar tilings . Ergodic Theory Dynam. Systems 17(1997), 695738.CrossRefGoogle Scholar
Solomyak, B., Spectrum of dynamical systems arising from Delone sets . In: Patera, J. (ed.), Quasicrystals and discrete geometry, Fields Institute Monographs, 10, American Mathematical Society, Providence, RI, 1998, pp. 265275.Google Scholar
Spindeler, T. and Strungaru, N., On norm almost periodic measures . Math. Z. 299(2021), 233255.CrossRefGoogle Scholar
Strungaru, N., Almost periodic measures and long-range order in Meyer sets . Discrete Comput. Geom. 33(2005), 483505.CrossRefGoogle Scholar
Strungaru, N., On the Bragg diffraction spectra of a Meyer set . Canad. J. Math. 65(2013), 675701.CrossRefGoogle Scholar
Strungaru, N., On weighted Dirac combs supported inside model sets . J. Phys. A: Math. Theor. 47(2014), 335202.CrossRefGoogle Scholar
Strungaru, N., Almost periodic pure point measures . In: Baake, M. and Grimm, U. (eds.), Aperiodic order. Volume 2: crystallography and almost periodicity, Cambridge University Press, Cambridge, 2017, pp. 271342.CrossRefGoogle Scholar
Strungaru, N., On the Fourier analysis of measures with Meyer set support . J. Funct. Anal. 278(2020), Article no. 108404, 30 pp.CrossRefGoogle Scholar
Strungaru, N., Model sets with precompact Borel windows. Preprint, 2020. arXiv:2012.06912 Google Scholar
Strungaru, N., Why do Meyer sets diffract? Preprint, 2021. arXiv:2101.10513Google Scholar