1 Introduction
The theory of aperiodic order is concerned with (generalizations of) Delone sets displaying some long-range order while at the same time being far from periodic. Until recently, the theory has mostly been studied in Euclidean space or abelian locally compact groups—we refer the reader to [Reference Baake and Grimm1] for an extensive bibliography in the abelian setting. With any (weighted) Delone set in an abelian locally compact group, one can associate a topological dynamical system, the so-called hull system of the Delone set, and these systems play a central role in the dynamical approach to the study of aperiodic order. A crucial question in this context is how properties of the dynamical system are reflected in the structure of the underlying Delone set and vice versa. For example, it is a famous result of Lagarias and Pleasants in [Reference Lagarias and Pleasants30] that the hull of a linearly repetitive Delone set in ${\mathbb R}^n$ is uniquely ergodic.
Recently, the scope of the theory of aperiodic order was extended to the realm of arbitrary (that is, not necessarily abelian) locally compact groups and, even more generally, proper homogeneous metric spaces, cf. [Reference Björklund and Hartnick7, Reference Björklund, Hartnick and Pogorzelski10–Reference Björklund, Hartnick and Pogorzelski12] in collaboration with Björklund. Further recent developments in this area can be found in [Reference Beckus and Pogorzelski5, Reference Björklund and Hartnick8, Reference Björklund and Hartnick9, Reference Björklund, Hartnick and Stulemeijer13, Reference Cordes, Hartnick, Tonić and Machado17, Reference Fish22, Reference Hrushovski28, Reference Machado36–Reference Machado38]. A particular focus has been on (aperiodic) Delone sets in locally compact second countable (lcsc) groups and, more specifically, in constructing interesting examples of such sets. While the original focus of this program was on Delone sets arising from a non-abelian version of Meyer’s cut-and-project construction, our most recent work [Reference Beckus, Hartnick and Pogorzelski4] introduces symbolic substitution techniques to a non-abelian setting. Using these techniques, we are able to construct the first examples of aperiodic linearly repetitive Delone sets in non-abelian Lie groups. By analogy with the Lagarias–Pleasants theorem mentioned above, it is then natural to ask whether the hull dynamical systems of these Delone sets are uniquely ergodic. We will see that the answer to this question is positive for all examples provided in [Reference Beckus, Hartnick and Pogorzelski4], but the proof of this fact will depend on a very peculiar property of our examples, namely exact polynomial growth of the ambient Lie group. Our construction in the companion paper [Reference Beckus, Hartnick and Pogorzelski4] produces linearly repetitive Delone sets inside a certain class of Lie groups called homogeneous Lie groups (see [Reference Folland and Stein23] or [Reference Fischer and Ruzhansky21, Ch. 3] for a more modern treatment). Lie groups from this class turn out to be nilpotent (hence, in particular, amenable) and admit a canonical quasi-isometry class of left-invariant metrics called homogeneous metrics, which induce the given topology.
If G is a homogeneous Lie group with homogeneous metric d, then a subset $\Lambda \subseteq G$ is called a Delone set if there exist $R> r > 0$ such that $d(x,y) \geq r$ for all distinct $x,y \in \Lambda $ and such that every $x \in G$ is at distance at most R from $\Lambda $ . It is called linearly repetitive if there exists a constant $c_\Lambda \geq 1$ such that every pattern of radius $\rho $ of $\Lambda $ occurs in every ball of radius $c_\Lambda \rho $ (this is made precise in Definition 4.6 below). By definition, the hull of a linearly repetitive Delone set is its orbit closure with respect to the Chabauty–Fell topology on the space of closed subsets of G. We then have the following generalization of the Lagarias–Pleasents theorem, which we expect to draw interesting connections to questions of spectral convergence via dynamical systems, cf. [Reference Beckus, Bellissard and De Nittis3, Reference Beckus and Pogorzelski5].
Theorem 1.1. (Unique ergodicity from linear repetitivity, homogeneous case)
Let G be a homogeneous Lie group with a homogeneous metric d. Then, the hull of every Delone set in G that is linearly repetitive with respect to d is minimal and uniquely ergodic.
We emphasize that there are plenty of examples to which Theorem 1.1 applies. Indeed, in [Reference Beckus, Hartnick and Pogorzelski4], we were able to construct non-periodic repetitive Delone sets, which are linearly repetitive with respect to a homogeneous metric on the ambient group, in 133 of the 149 families of indecomposable $1$ -connected nilpotent $7$ -dimensional real Lie groups classified in [Reference Gong25]. In fact, these examples were a key motivation for us to establish Theorem 1.1.
We will actually prove a slightly stronger result than Theorem 1.1, which applies to all weighted Delone sets which are almost linearly repetitive (this is made precise below). Minimality of the hull holds in an even wider context, see Proposition 2.4 below, and can be established by adapting ideas from [Reference Frettlöh and Richard24] to the setting of weighted Delone sets. The main contribution of the present article lies in proving unique ergodicity, and it is in this part of the proof that we use the assumption that G is a homogeneous Lie group with homogeneous metric d. As explained in [Reference Beckus, Hartnick and Pogorzelski4], these assumptions imply that the pair $(G,d)$ has exact polynomial growth in the sense of the following definition (Definition 4.3 below).
Definition. Let G be an lcsc group with left-Haar measure $m_G$ and let d be an adapted (that is, left-invariant, proper, and locally bounded) metric on G. We say that G has exact polynomial growth with respect to d if there are constants $C> 0$ and $q \geq 0$ such that
where $B_t$ denotes the open ball of radius t around the identity of G.
Every compactly generated lcsc group which has exact polynomial growth with respect to some metric d as above obviously has polynomial growth in the sense of geometric group theory. Conversely, by a celebrated result of Breuillard [Reference Breuillard14], every compactly generated lcsc group of polynomial growth admits an adapted metric of exact polynomial growth, see Theorem 4.13. If G is moreover a connected Lie group, then this metric can be chosen to be continuous (in fact, Riemannian and, in particular, real-analytic), but it seems to be an open problem whether this is the case in general. It turns out that exact polynomial growth is actually sufficient to establish Theorem 1.1.
Theorem 1.2. (Unique ergodicity from linear repetitivity, exact case)
If an lcsc group G has exact polynomial growth with respect to an adapted metric d, then the hull of every (weighted) Delone set in G that is (almost) linearly repetitive with respect to d is minimal and uniquely ergodic.
We establish Theorem 1.2 in Theorem 4.9 below; note that by the previous remarks, Theorem 1.1 is merely a special case. The assumption that G has exact polynomial growth with respect to d implies that metric balls around the identity form a Følner sequence in G, and even a strong Følner exhaustion sequence in the sense of Definition 3.2 below. Linear repetitivity with respect to d can then be reformulated in terms of this Følner sequence, and this leads to the notion of (almost) tempered repetitivity of a Delone set with respect to a Følner sequence (see Definition 3.5 below). Using this notion, one can reformulate Theorem 1.2 in terms of Følner sequences rather than metrics, and this metric-free formulation actually holds in much greater generality.
Theorem 1.3. (Unique ergodicity from tempered repetitivity)
Let $(T_m)$ be a strong Følner exhaustion sequence in an amenable unimodular lcsc group G and let $\Lambda $ be a weighted Delone set in G. If $\Lambda $ is almost tempered repetitive with respect to $(T_m)$ , then the hull of $\Lambda $ is minimal and uniquely ergodic.
Theorem 1.2 is indeed a special case of Theorem 1.3, since linear repetitivity with respect to an adapted metric of exact polynomial growth is equivalent to tempered repetitivity with respect to the corresponding strong Følner exhaustion sequence of balls (see Proposition 4.7 below).
It seems to us that Theorem 1.3 is the most natural generalization of the Lagarias–Pleasants theorem to the general amenable case. By Proposition 3.4 below, strong Følner exhaustion sequences exist in all unimodular, amenable lcsc groups, but it is not clear at all in which cases (beyond the case of exact polynomial growth) they can be chosen to be balls with respect to a suitable metric; hence, the need for a metric free formulation. Theorem 1.3 indicates that, in this metric-free context, tempered repetitivity is the correct replacement for linear repetitivity; hence, we would like to advertise this notion.
While Theorem 1.3 is formulated in terms of Delone dynamical system, it is easy to deduce a corresponding unique ergodicity result in the symbolic setting. To formulate such a result, let $\mathcal A$ be a finite alphabet and $\Gamma $ be a countable amenable group. We refer to an element of $\mathcal {A}^{\Gamma }$ as a coloring of $\Gamma $ by $\mathcal A$ and equip $\mathcal A^\Gamma $ with its natural compact metrizable topology, given by the product topology. Given a coloring $C \in \mathcal {A}^{\Gamma }$ , we refer to its orbit closure in $\mathcal A^\Gamma $ as the hull of C. Similarly to the case of Delone sets, one can also define the notions of a symbolically tempered repetitive coloring (with respect to a Følner sequence in $\Gamma $ ) or a symbolically linearly repetitive coloring (with respect to a left-invariant metric on $\Gamma $ ), see Definition 5.3 below. In fact, the situation is slightly simpler than in the Delone setting, since for countable groups, the notions of Følner sequence and strong Følner sequence are equivalent by [Reference Pogorzelski and Schwarzenberger47, Lemma 2.7(d)]. In any case, the following result can then be deduced easily from Theorem 1.3, using the fact that colorings of $\Gamma $ can be identified with certain weighted Delone sets in $\Gamma $ (see §5 for details).
Theorem 1.4. (Unique ergodicity in the symbolic setting)
Let $\Gamma $ be a countable amenable group and let $\mathcal {A}$ be a finite set. If $\mathcal {C} \in \mathcal {A}^{\Gamma }$ is symbolically tempered repetitive with respect to a Følner exhaustion sequence, then the hull of $\mathcal {C}$ is minimal and uniquely ergodic. In particular, if $\Gamma $ has exact polynomial growth with respect to an adapted metric $d_{\Gamma }$ and if $\mathcal {C}$ is symbolically linearly repetitive with respect to $d_{\Gamma }$ , then the hull of $\mathcal {C}$ is minimal and uniquely ergodic.
Let us now discuss some of the ingredients of the proof of Theorem 1.3. As in the cases of Theorems 1.1 and 1.2, the minimality result of Theorem 1.3 can be obtained by adapting arguments from [Reference Frettlöh and Richard24], whereas the unique ergodicity result requires new ideas. Our approach is based on the observation by Damanik and Lenz [Reference Damanik and Lenz19] that in the Euclidean setting, linear repetitivity implies the validity of a uniform sub-additive convergence theorem for certain functions defined on the collection $\mathcal {B}({\mathbb R}^n)$ of all n-dimensional boxes in ${\mathbb R}^n$ , which in turn can be used to ensure unique ergodicity. This observation has been refined in many different directions, see e.g. [Reference Besbes, Boshernitzan and Lenz6, Reference Frettlöh and Richard24, Reference Lenz32–Reference Lenz, Schwarzenberger and Veselić35, Reference Pogorzelski43, Reference Pogorzelski45, Reference Pogorzelski and Schwarzenberger47]. A different approach is via strongly almost periodic translation bounded measures, see [Reference Lee, Lenz, Richard, Sing and Strungaru31]. In Theorem 3.1 below, we establish a general version of such a convergence theorem for a large class of amenable groups, which is general enough to allow us to deduce Theorem 1.3 (and thereby also Theorems 1.1, 1.2, and 1.4). While the precise statement of the convergence theorem is rather technical, let us highlight two of the main features of the theorem and its proof.
-
• The theorem works for general amenable unimodular lcsc groups and for a rather general class of weight functions. To achieve this, the study of n-dimensional boxes in ${\mathbb R}^n$ is replaced by the Ornstein–Weiss machinery of $\varepsilon $ -quasi-tilings [Reference Ornstein and Weiss41] and more specifically its variant from [Reference Pogorzelski and Schwarzenberger47].
-
• There is no requirement that the Delone set under consideration has finite local complexity (FLC). A consequence of this is that Theorem 1.2 also holds for almost linearly repetitive Delone sets (and Theorem 1.3 also holds for almost tempered repetitive Delone sets). This is achieved using ideas from Frettlöh and Richard [Reference Frettlöh and Richard24]. Roughly speaking, to incorporate non-FLC situations, one has to identify not only patches which are equal up to a translation, but also patches which are close up to a translation. The fact that being $\delta $ -close for a small $\delta>0$ is not an equivalence relation creates all kinds of technical difficulties.
It is well known in the abelian setting that uniform sub-additive convergence theorems have applications beyond unique ergodicity of Delone dynamical systems, and this is no different in our more general setting. We demonstrate the strength of our convergence theorem by two such applications: one to Banach densities and one to the integrated densities of states (IDS). We deduce the corresponding theorems along the lines of the corresponding abelian situations.
Upper and lower Banach densities appear as an important combinatorial quantity in many mathematical areas, see for instance a recent example [Reference Downarowicz, Huczek and Zhang20] for countable amenable groups. For sets with enough symmetry, the upper and the lower Banach density may coincide, and one can just refer to the Banach density of a set. In analogy to the abelian realm, this situation occurs for almost tempered repetitive weighted Delone sets. For the definition of the upper and lower Banach density of a weighted Delone set, we refer to §6. Given a weighted Delone set $\Pi $ , we denote by $\delta _\Pi $ the associated Dirac comb which is a Radon measure on the group.
Corollary 1.5. Let G be an amenable unimodular lcsc group with Haar measure $m_G$ . Suppose that $\Lambda $ is a weighted Delone set that is almost tempered repetitive with respect to a strong Følner exhaustion sequence $(T_m)$ of G. Then, the Banach density
exists uniformly in $\Pi \in \mathcal {H}_\Lambda $ .
The implications of our convergence theorem concerning the IDS are a bit more technical to state. We give a sample result in the setting of finitely generated amenable groups $\Gamma $ whose elements are labeled by colors taken from a finite set $\mathcal {A}$ . Since each coloring $\mathcal {C} \in \mathcal {A}^{\Gamma }$ can be interpreted as a weighted Delone set in $\Gamma $ , it makes sense to define a version of tempered repetitivity for a coloring. In this context, our convergence theorem allows us to verify a criterion from [Reference Pogorzelski and Schwarzenberger47] which then shows that the IDS for certain pattern-equivariant operators on graphs can be uniformly approximated by finite volume analogs. The precise class of considered operators is defined in Definition 6.2.
Corollary 1.6. Let $\Gamma $ be a finitely generated amenable group and let $\mathcal {C} \in \mathcal {A}^{\Gamma }$ be a coloring of the group by a finite set $\mathcal {A}$ . Suppose that $\mathcal {C}$ is symbolically tempered repetitive as a weighted Delone set with respect to some strong Følner exhaustion sequence $(T_m)$ . Then for every $\mathcal {C}$ -invariant self-adjoint operator $H: \ell ^2(\Gamma ) \to \ell ^2(\Gamma )$ and for its integrated density of states $N_H: {\mathbb R} \to [0,1]$ , we get
where the $N_{H_m}$ denote the empirical eigenvalue distribution functions of restrictions $H_m$ of H to a subspace of $\ell ^2(T_m^{-1})$ .
This article is organized as follows. In §2, we introduce Delone dynamical systems and recall the equivalence of minimality of these systems and (almost) repetitivity of the underlying set. Section 3 is devoted to the main convergence theorem, cf. Theorem 3.1, which is stated and proven for so-called admissible weight functions defined on hulls of almost tempered weighted Delone sets. From this, we derive in §4 the unique ergodicity of the hulls of almost tempered repetitive weighted Delone sets, as stated in Theorem 1.3. Moreover, we derive Theorems 1.2 and 1.1 from Theorem 1.3, cf. Theorem 4.9. Section 5 is devoted to symbolic systems over a finite set and the proof of Theorem 1.4. The aforementioned applications including the proofs of the Corollaries 1.5 and 1.6 are carried out in §6. Appendices A and B contain proofs of some technical lemmas on almost sub-additive weight functions and topological aspects of dynamical systems induced by weighted Delone sets.
2 Preliminaries on Delone dynamical systems
2.1 Delone dynamical systems
We recall that a subset P of a metric space $(X,d)$ is called:
-
1. uniformly discrete if there exists $r>0$ such that $d(x,y) \geq r$ for all distinct $x,y \in P$ ;
-
2. relatively dense if there exists $R>0$ such that X is an R-neighborhood of P;
-
3. Delone if it is uniformly discrete and relatively dense.
If P is a Delone set, $\sigma \geq 1$ and $\alpha : P \to [\sigma ^{-1}, \sigma ]$ is a function, then the pair $(P, \alpha )$ is called a weighted Delone set. We consider Delone sets as weighted Delone sets with constant weight $1$ .
Let G be an lcsc group and let d be a metric on G. Following [Reference Cornulier and de la Harpe18], we say that a metric d on G is:
-
• left-invariant if $d(gh, gk) = d(h,k)$ for all $g,h, k \in G$ ;
-
• proper if $B_t := \{g \in G \mid d(g,e) < t\}$ is relatively compact for all $t> 0$ ;
-
• locally bounded if every $g\in G$ admits a neighborhood of finite diameter with respect to d;
-
• adapted if it is left-invariant, proper, and locally bounded.
By [Reference Cornulier and de la Harpe18, Proposition 2.A.9], any continuous proper left-invariant metric generates the topology on G. We observe that if d and $d'$ are continuous metrics on G which are coarsely equivalent (in the sense of [Reference Cornulier and de la Harpe18, p. 11]), then they define the same class of Delone sets. We will need the following facts concerning such metrics (see [Reference Cornulier and de la Harpe18, §1.D]).
Proposition 2.1. Every lcsc group admits a continuous proper left-invariant metric d. Any such metric is automatically adapted and generates the underlying topology. Moreover, any two adapted metrics on G are coarsely equivalent (even if they are not continuous).
In view of the proposition, we say that a subset $P \subseteq G$ is uniformly discrete, relatively dense, or Delone if it has the corresponding property with respect to some (hence any) continuous adapted metric. For the remainder of this section, we fix an adapted metric d on G. Later on (see §4 below), we will have to choose an adapted metric with very peculiar properties, but for the moment, the precise choice of metric will not matter to us. In view of Proposition 2.1 and the previous considerations, we assume as well that d is continuous, and we make this assumption throughout this and the next section. We then denote by
the open, respectively closed, ball around the identity so that $B_s \subseteq \overline {B}_s \subseteq B_t$ for all $s < t$ . The open and closed balls around arbitrary centers $g \in G$ are defined as $B_t(g) = gB_t$ and $\overline {B}_t(g) = g \overline {B}_t $ , respectively. We emphasize that $\overline {B}_r$ may be larger than the closure $\overline {B_r}$ of $B_r$ . For example, in the non-Archimedean case, $\overline {B_r} = B_r$ since the latter is compact. Due to our continuity assumption on d, we know that $B_r$ is open and $\overline {B}_r$ is closed; neither of these statements would be true for a general (that is, possibly discontinuous) adapted metric.
By [Reference Björklund and Hartnick7], a subset $P \subseteq G$ is uniformly discrete with respect to d (or any other continuous adapted metric on G) if and only if there exists an open relatively compact subset $U \subseteq G$ such that $\#(gU \cap P)\leq 1$ for all $g \in G$ , and relatively dense if and only if there exists a compact subset $K \subseteq G$ such that P is left-K-syndetic in the sense that $P. K = G$ , where for two subsets $A,B\subseteq G$ , the notation $A.B$ refers to the Minkowski product of A and B. With a slight abuse of notation, we write $gA$ for $\{g\}.A$ , where $g\in G$ and $A\subseteq G$ . We then say that P is U-uniformly discrete and K-relatively dense, respectively, and refer to a set with both of these properties as a $(U,K)$ -Delone set. Finally, we denote by
the collection of weighted Delone sets with uniform parameters $U, K, \sigma $ . Note that G acts on $\operatorname {Del}(U,K,\sigma )$ by $g.(P, \alpha ) := (gP, g_*\alpha )$ , where $g_*\alpha (q) := \alpha (g^{-1}q)$ .
We will identify each $(P, \alpha ) \in \operatorname {Del}(U,K,\sigma )$ with the associated Dirac comb
and thereby think of $\operatorname {Del}(U,K,\sigma )$ as a G-invariant subset of the space of Radon measures on G. We will always equip the latter space with the weak- $*$ topology. Then by [Reference Baake and Lenz2, Reference Björklund, Hartnick and Pogorzelski11], the subspace $\operatorname {Del}(U,K,\sigma )$ is compact and the G-action on this space is jointly continuous. The topology induced on the space $\operatorname {Del}(U,K,1)$ of (unweighted) Delone sets is precisely the Chabauty–Fell topology (see [Reference Björklund, Hartnick and Pogorzelski11] and also [Reference Müller and Richard39, §2.1] and the references therein).
We will need yet another description of the topology on $\operatorname {Del}(U,K,\sigma )$ . Given a relatively compact subset $S \subseteq G$ and two Radon measures $\mu $ and $\nu $ on G, we define
In particular, using our identification of weighted Delone sets and their corresponding Dirac combs, we define
for weighted Delone sets $\Lambda , \Pi $ . Then, given $\Lambda \in \operatorname {Del}(U,K,\sigma )$ , the sets
form a neighborhood basis of $\Lambda $ in $\operatorname {Del}(U,K,\sigma )$ , as $\delta $ and S vary over all positive real numbers and all relatively compact subsets of G, respectively (cf. Proposition B.1 in Appendix B). In particular, Delone sets $\Pi _n$ converge to $\Lambda $ in $\operatorname {Del}(U,K,\sigma )$ if and only if $\lim _{n \to \infty } d_S(\Pi _n, \Lambda ) = 0$ for all relatively compact S.
Given $\Lambda \in \operatorname {Del}(U,K,\sigma )$ , the orbit closure $\mathcal {H}_{\Lambda } := \overline {\{ g . \Lambda :\, g \in G\}}$ is compact and the G-action on $\mathcal {H}_{\Lambda }$ is jointly continuous, that is, $G\curvearrowright \mathcal {H}_{\Lambda }$ is a topological dynamical system over G. We refer to $\mathcal H_\Lambda $ as the hull of $\Lambda $ and to $G\curvearrowright \mathcal {H}_{\Lambda }$ as a Delone dynamical system.
2.2 Almost repetitivity and minimality
In this subsection, we give a combinatorial characterization of minimality of Delone dynamical systems, generalizing results of [Reference Frettlöh and Richard24] in the unweighted case. Throughout this subsection, we fix a weighted Delone set $\Lambda = (P, \alpha ) \in \operatorname {Del}(U,K,\sigma )$ with associated Dirac comb $\delta _\Lambda $ . If $S \subseteq T$ are two relatively compact subsets of G and $\nu $ is a Radon measure on T, then we denote by $\nu |_S$ the restriction of $\nu $ to the Borel $\sigma $ -algebra of S.
Given a relatively compact subset $S \subseteq G$ , we refer to the pair $\Lambda _S:=({\delta _\Lambda }|_{S}, S)$ as the S-patch of $\Lambda $ . If $p = (\mu , S)$ is a patch of $\Lambda $ , we refer to $\mu $ as the underlying measure and S as the support of p (which is not to be confused with the support of $\mu $ ). For $T \subseteq S$ , we define the restriction of p to T by $p|_{T} := (\mu |_{T}, T)$ .
Two patches $p = (\mu , S)$ and $q = (\nu , T)$ of $\Lambda $ are called equivalent if there exists $g \in G$ with $g.p := (g_*\mu , gS) = (\nu , T)$ , and an equivalence class of patches is called a pattern. A pattern is said to be of size S for a relatively compact set S if the supports of its patches are translates of S.
Given $\delta \geq 0$ , we say that two patches $p = (\mu , S)$ and $q = (\nu , T)$ of $\Lambda $ are $\delta $ -similar if there exists $g \in G$ such that $gS = T$ and $d_T(g_*\mu , \nu ) < \delta $ . By definition, two patches are $0$ -similar if and only if they are equivalent. In particular, $0$ -similarity is an equivalence relation, whereas $\delta $ -similarity is not a transitive relation for $\delta>0$ .
Given a relatively compact subset $T \subseteq G$ and $\delta \geq 0$ , we say that a pattern $[p]$ of $\Lambda $ with representative $p = (\mu , S)\ \delta $ -occurs in T if there is some $g \in G$ such that $gS \subseteq T$ and p is $\delta $ -similar to $\Lambda _{gS}$ . If $\delta =0$ , we simply say that $[p]$ occurs in T. Similarly, a patch $p=(\mu ,S)$ ( $\delta $ -)occurs in a patch $q=(\nu ,T)$ if there is some $g\in G$ such that $gS\subseteq T$ and p is ( $\delta $ -)similar to $q|_{gS}$ .
Definition 2.2. A function $\varrho : (0,1) \times [1, \infty ) \to [1, \infty )$ is called an almost repetitivity function for $\Lambda $ if for every $\delta \in (0,1)$ and all $R \geq 1$ , every pattern of $\Lambda $ of size $B_R\ \delta $ -occurs in $B_{\varrho (\delta , R)}(h)$ for all $h \in G$ . If such a function exists, then $\Lambda $ is called almost repetitive.
Note that $\varrho (\delta ,R)\geq R$ for all $\delta>0$ .
Remark 2.3. Some authors define linear repetitivity by asking the above condition to hold for all $R\geq r_0$ , where $r_0>0$ is an arbitrary constant, whereas we insist here that $r_0 =1$ . While the two possible definitions are a priori different, this difference is irrelevant for us. Namely, we can always rescale the metric by a constant to achieve that the condition holds with $r_0=1$ . The reader is invited to check that all our results are invariant under such rescalings, and hence all our theorems hold true for either definition. In view of this fact, we will always assume $r_0 = 1$ to keep the notation simple.
Proposition 2.4. Let $\Lambda $ be a weighted Delone set in G. Then, $\Lambda $ is almost repetitive if and only if the associated Delone dynamical system $G \curvearrowright \mathcal H_\Lambda $ is minimal.
Proof. We closely follow the proof of [Reference Frettlöh and Richard24, Theorem 3.11].
Suppose the hull of $\Lambda = (P, \alpha )$ is minimal. Let $R \geq 1$ and fix $\Pi = (Q,\beta ) \in \mathcal {H}_{\Lambda }$ . Assume that $0 <\varepsilon < R$ . We further choose $R^{\prime } \geq R$ such that $B_{R^{\prime }} \cap Q = \overline {B}_R \cap Q$ . We set
Then, $\mathcal {U}_{R,\varepsilon } ( \Pi )$ is a neighborhood of $\Pi $ in the weak- $\ast $ -topology, cf. Proposition B.1. By minimality of $\Lambda $ , every element in $\mathcal {H}_{\Lambda }$ has a dense orbit. This gives
By compactness of the hull, we find finitely many $x_1, x_2, \ldots , x_{\ell }$ such that $\mathcal {H}_{\Lambda } \subseteq \bigcup _{j=1}^{\ell } x_j \, \mathcal {U}_{R,\varepsilon } ( \Pi )$ . We now choose $\varrho _1(R)> 0$ such that $x_j \in B_{\varrho _1(R)}$ for all $1 \leq j \leq \ell $ . This means that for all $h \in G$ , there is some $h^{*} \in B_{\varrho _1(R)}$ such that $h.(P,\alpha ) = h^{*}.(D,\gamma )$ and $(D,\gamma )$ is some element in $\mathcal {U}_{R,\varepsilon }(\Pi )$ . This also gives $D \cap B_{R^{\prime }} = h^{*\, -1} h\, P \cap B_{R^{\prime }}$ . Since the patches $(\delta _{(Q \cap B_{R^{\prime }}, \beta _{| Q \cap B_{R^{\prime }}})}, B_{R^{\prime }})$ and $( \delta _{(D \cap B_{R^{\prime }}, \gamma _{| D \cap B_{R^{\prime }}})}, B_{R^{\prime }})$ are $\varepsilon $ -similar, the left-invariance of the metric tells us that also $( \delta _{(Q \cap B_{R^{\prime }}, \beta _{| Q \cap B_{R^{\prime }}})}, B_{R^{\prime }})$ is $\varepsilon $ -similar to the $R^{\prime }$ -patch of P centered at $h^{-1}h^{*}$ . We will now show that $h^{-1}h^{*} B_R B_{\varepsilon } \subseteq B_{\varrho _1(R) + 2R}(h^{-1})$ . The condition $\varepsilon < R$ gives $B_R B_{\varepsilon } \subseteq B_{2R}$ . Further, we have $h^{*}B_{2R} \subseteq B_{\varrho _1(R)}B_{2R} \subseteq B_{\varrho _1(R) + 2R}$ . Hence, we obtain $h^{-1}h^{*}B_RB_{\varepsilon } \subseteq B_{\varrho _1(R) + 2R}(h^{-1})$ , as claimed. Since $\Pi $ was chosen arbitrarily, this shows that for all possible $B_R$ -patterns $[p]$ of $\Lambda $ and for each $h \in G$ , there is some R-patch of $\Lambda $ contained in $B_{\varrho _1(R) + 2R}(h)$ which is $\varepsilon $ -similar to a representative of $[p]$ . A straightforward compactness argument shows that there is a finite number of patterns $[p_k]$ such that every $B_{R}$ -patch of $\Lambda $ is $\varepsilon $ -similar to a representative of one of the $[p_k]$ . Hence, repeating the above procedure finitely many times, we can set $\varrho (R) := \max _{k} \varrho _k(R) + 2R$ and we find that $\Lambda $ is almost repetitive with almost repetitivity function $\varrho $ .
Conversely, assume by contradiction that $\Lambda = (P,\alpha )$ is almost repetitive but $\mathcal {H}_\Lambda $ is not minimal. Let $\Pi = (Q,\beta ) \in \mathcal {H}_{\Lambda }$ and $\{g.\Pi :\,g\in G\}$ is not dense in $\mathcal {H}_{\Lambda }$ . This implies $\Lambda \notin \mathcal {H}_{\Pi }$ as well as $\mathcal {H}_{\Pi } \subsetneq \mathcal {H}_{\Lambda }$ . We fix a compact neighborhood $\mathcal {V} \subseteq \mathcal {H}_{\Lambda }$ of $\Lambda $ containing the set $\{\Theta \in \mathcal {H}_{\Lambda }:\, d_{B_{1/\varepsilon }} ( \Theta ,\, \Lambda ) < \varepsilon \} $ for some small $\varepsilon> 0$ such that $\mathcal {V} \cap \mathcal {H}_{\Pi } = \emptyset $ . We now take a compact set $K \subseteq G$ with $B_{1/\varepsilon }(e) \subseteq \mathring {K}$ and define ${T_{K, \varepsilon }( \Lambda ) = \{g \in G:\, d_{\mathring {K}}(g . \Lambda , \Lambda ) < \varepsilon \}}$ .
Claim. The set $T_{K, \varepsilon }( \Lambda )$ is right-relatively dense.
Assuming the validity of the claim for a moment, we see
for some compact set $K^{\prime } \subseteq G$ . By compactness of $\mathcal {V}$ and the continuity of the G-action, we find that the set $K^{\prime }.\mathcal {V}$ is compact. Hence, we arrive at $\mathcal {H}_{\Lambda } \subseteq K^{\prime }.\mathcal {V}$ . We get that $\Pi = x.\Theta $ for some $x \in K^{\prime }$ and some $\Theta \in \mathcal {V}$ . Thus, $\Theta = x^{-1}\Pi $ contradicting the fact that ${\mathcal {V} \cap \mathcal {H}_{\Pi } = \emptyset }$ .
It remains to show the claim. For this, we proceed along the lines of [Reference Frettlöh and Richard24, proof of Lemma 3.6, implication (iii) ⇒ (i)]. To this end, find some $r \geq 1$ such that $\mathring {K} \subseteq B_r$ and set $R:= \varrho (\varepsilon , r)$ , where $\varrho $ is an almost repetitivity function of $\Lambda $ . We fix a countable cover of G of the form $G = \bigcup _{i=1}^{\infty } B_R(g_i)$ . By almost repetitivity, for all $i \in {\mathbb N}$ , we find $x_i \in G$ such that $x_i^{-1}B_r \subseteq B_{R}(g_i)$ and $d_{B_r}( x_i\Lambda , \Lambda ) = d_{x_i^{-1}B_r} ( \Lambda , x_i^{-1}\Lambda ) < \varepsilon $ . Now to show the claim, it clearly suffices to show that $T_{r,\varepsilon }:= \{x_i:\, i \in {\mathbb N}\} \subseteq T_{K, \varepsilon }(\Lambda )$ is right-relatively dense in G. For each $b \in B_r$ , we have $x_i^{-1}b \subseteq B_R(g_i)$ for all $i \in {\mathbb N}$ . This shows that ${G = \bigcup _{i=1}^{\infty } B_{2R}(x_i^{-1}b) = \bigcup _{i=1}^{\infty } x_i^{-1} B_{2R}(b)}$ . For some fixed $b \in B_r$ , we define
This set is clearly compact. We will show $K^{\prime } T_{r,\varepsilon } = G$ . Take $h \in G$ . This gives $h^{-1}b \in x_i^{-1}B_{2R}(b)$ for some $i \in {\mathbb N}$ . This yields
Since h was chosen arbitrarily, the proof of the claim is finished.
While almost repetitivity arises naturally in the study of minimality of Delone dynamical systems, a more classical notion is the following one.
Definition 2.5. A function $\varrho : [1,\infty ) \to [1, \infty )$ is called a repetitivity function for $\Lambda $ if for all $R \geq 1$ , every pattern of $\Lambda $ of size $B_R$ occurs in $B_{\varrho (R)}(h)$ for all $h \in G$ . If such a function exists, then $\Lambda $ is called repetitive.
Remark 2.6. Similarly as for the almost repetitivity function, we have $\varrho (R)\geq R$ and we assume $R\geq 1$ .
We say that $\Lambda = (P, \alpha )\in \operatorname {Del}(U,K,\sigma )$ has finite local complexity (FLC) if for ever compact set S, the set $\{(x^{-1}. \Lambda )|_S :\, x\in P\}$ is finite.
Proposition 2.7. Let $\Lambda \in \operatorname {Del}(U,K,\sigma )$ be of finite local complexity. If $\Lambda $ is almost repetitive, then $\Lambda $ is repetitive.
Proof. Without loss of generality, we assume that $e\in U$ . Since $\Lambda = (P,\alpha )$ is of finite local complexity, the set
is finite for each $r>0$ . In particular, if $r>0$ is such that the topological boundary $\overline {B_r}\setminus B_r$ is not intersecting $x^{-1}P$ for any $x\in P$ , then there is a $\delta>0$ such that ${x^{-1}P\cap \partial _{B_{\delta _1}}(B_{r}) = \emptyset} $ holds for all $x\in P$ , where
is the $B_{\delta _1}$ boundary of the topological boundary of $B_r$ .
Let $R>0$ be fixed. Since $\Lambda $ is K-relatively dense and of finite local complexity, we can choose an $R_1>0$ and a $\delta _1>0$ such that:
-
(a) for each $x\in G$ , there is a $y_x\in P$ satisfying $x\in B_{R_1}(y_x)$ ;
-
(b) for all $y\in P$ , we have $y^{-1}P\cap \partial _{B_{\delta _1}}(B_{R+R_1}) = \emptyset $ .
Consider an R-patch $p:=\Lambda _{B_R(x)}$ for $x\in G$ . Let $y_x\in P$ be such that $x\in B_{R_1}(y_x)$ . Now consider the patch
By construction, p occurs in $p'$ . Let $0<\delta <\min \{\delta _1,\sigma ^{-1}\}$ be such that $B_\delta \subseteq U$ and
The latter is possible as $P(R+R_1)$ is finite.
Let $h\in G$ . Since $\Lambda $ is almost repetitive, $p'\ \delta /2$ -occurs in $B_{\varrho (\delta /2,R+R_1)}(h)$ , where $\varrho $ is an almost repetitivity function for $\Lambda $ . Thus, there is a $g\in G$ such that $gB_{R+R_1}\subseteq B_{\varrho (\delta /2,R+R_1)}(h)$ and $d_{B_{R+R_1}}( \delta _{p'}, \delta _{g^{-1}\cdot q_h}) <\delta /2$ , where $q_h:=\Lambda _{B_{R+R_1}(g)}$ . Thus,
holds. By construction, $\delta _{p'}$ has support on $\{e\}$ , namely, $\delta _{p'}(e)>\sigma ^{-1}$ . Hence, there is a unique $y_g\in P$ and $g'\in B_{\delta /2}$ such that $g=y_g g'$ using that $\delta <\sigma ^{-1}$ , P is $B_\delta $ -uniformly discrete, and $d_{B_{R+R_1}} ( \delta _{p'}, \delta _{g^{-1}.q_h})<\delta /2$ . Due to condition (b), we conclude $\delta _{y_g^{-1}\cdot \Lambda }|_{B_{R+R_1}} = \delta _{y_g^{-1}\cdot \Lambda }|_{B_{R+R_1}(g')}$ and so
By construction, $(y_g^{-1}\cdot \Lambda )_{B_{R+R_1}}$ occurs in $B_{\varrho (\delta /2,R+R_1)+\delta /2}(h)$ since $g'\in B_{\delta /2}$ . Invoking that $\delta <\sigma ^{-1}$ , P is $B_\delta $ -uniformly discrete, $g'\in B_{\delta /2}$ , and $d_{B_{R+R_1}}(\delta _{p'}, \delta _{g^{-1}\cdot q_h} )<\delta /2$ , we derive
Thus, $(y_g^{-1}\cdot \Lambda )_{B_{R+R_1}}=p'$ follows by the choice of $\delta $ and since $(y_g^{-1}\cdot \Lambda )|_{B_{R+R_1}}, p'\in P(R+R_1)$ . Since p occurs in $p'$ and $(y_g^{-1}\cdot \Lambda )_{B_{R+R_1}}$ occurs in $B_{\varrho (\delta /2,R+R_1)+\delta /2}(h)$ , we conclude that p occurs in $B_{\varrho (\delta /2,R+R_1)+\delta /2}(h)$ .
Hence, the map $\varrho ':(0,\infty )\to (0,\infty )$ defined by $\varrho '(R):=\varrho (\delta /2,R+R_1)+\delta /2$ defines a repetitivity function. Note here that in the previous considerations, $\delta $ and $R_1$ depend on R but not on h.
Remark 2.8. Note that if $\Lambda = (P,\alpha )$ is repetitive, then $\Lambda $ necessarily has FLC since $B_{\varrho (R)}(h)\cap P$ is uniformly bounded in $h\in G$ . Furthermore, every repetitive $\Lambda $ is almost repetitive, which can be seen by setting $\varrho (\delta ,R):=\varrho (R)$ for $R \geq 1$ . In other words, for Delone sets, we have the equivalence
3 An abstract ergodic theorem
3.1 Statement of the theorem
Throughout this section, G will denote an amenable unimodular lcsc group. We fix a Haar measure $m_G$ on G. The goal of this section is to explain and prove the following theorem.
Theorem 3.1. (Uniform sub-additive convergence theorem)
Assume that:
-
• $\Lambda $ is a weighted Delone set in G with hull $\mathcal {H}_\Lambda $ ;
-
• $(T_m)$ is a strong Følner exhaustion sequence in G;
-
• $\Lambda $ is almost tempered repetitive with respect to $(T_m)$ ;
-
• w is an admissible weight function on G over $\mathcal {H}_\Lambda $ .
Then, there exists $I_w \in {\mathbb R}$ (depending on $\Lambda $ and w, but independent of $(T_m)$ ) such that
Theorem 3.1 will be proved in §3.4 after introducing all the required terminology. Tempered repetitivity with respect to a strong Følner exhaustion sequence will be defined in §3.2 and admissible weight functions will be discussed in §3.3.
3.2 Strong Følner exhaustion sequences and tempered repetitivity
We recall that amenability of G is equivalent to the existence of a Følner sequence, that is, a sequence $(T_m)$ of compact subsets of G of positive Haar measure such that for all compact subsets $K \subseteq G$ ,
In fact, one can choose a Følner sequence with additional properties. Given relatively compact subsets $L, S \subseteq G$ , we denote by
the L-boundary of S, as defined in [Reference Ornstein and Weiss41].
Definition 3.2. A Følner sequence $(T_m)$ is called a strong Følner exhaustion sequence if:
-
(i) $(T_m)$ is a strong Følner sequence in the sense of [Reference Pogorzelski and Schwarzenberger47], that is, for all compact subsets $K \subseteq G$ ,
$$ \begin{align*} \lim_{m \to \infty} \frac{m_G(\partial_{K}(T_m))}{m_G(T_m)} = 0; \end{align*} $$ -
(ii) $(T_m)$ is a strong exhaustion sequence, that is, $\{e\} \subseteq T_m \subseteq \mathring {T}_{m+1}$ and $\bigcup _{m} T_m = G$ .
Remark 3.3. We point out that by [Reference Pogorzelski, Richard and Strungaru46, Proposition 5.7], the notion of a strong Følner sequence is equivalent to the concept of a van Hove sequence. The latter has been used prominently in the literature on aperiodic order in locally compact abelian groups.
Proposition 3.4. Every amenable unimodular group G admits a strong Følner exhaustion sequence.
Proof. Building on the uniform Følner condition proven in [Reference Ornstein and Weiss41], one observes that G admits a strong Følner sequence $(S_n)$ , cf. [Reference Pogorzelski44, Lemma 2.8]. Moreover, the sequence $(S_n)$ can be turned into a strong Følner exhaustion sequence by the usual construction (cf. [Reference Greenleaf26]). By $\sigma $ -compactness of G, we can now choose an exhaustion of G by compact subsets $(K_n)$ containing the identity. Set $T_1 := S_1$ and choose $m_1$ such that $T_1 \subseteq \mathring {K}_{m_1}$ . Then choose $\ell \in {\mathbb N}$ large enough such that $m_G(\partial _{K_{m_1}}(S_{\ell })) \leq \tfrac 12 m_G(S_{\ell })$ . Thus, there is an $h \in G$ such that $K_{m_1}h \subseteq S_{\ell }$ . If we now set $T_2:= S_{\ell }h^{-1}$ , then $T_1 \subseteq \mathring {K}_{m_1} \subseteq T_2$ . One now proceeds by induction to construct the desired strong Følner exhaustion sequence $(T_m)$ .
From now on, $\mathcal T = (T_m)$ denotes a strong Følner exhaustion sequence in G and $\Lambda $ denotes a weighted Delone set in G.
Definition 3.5. The repetitivity index for $\Lambda $ with respect to $\mathcal {T}$ is the function $ \mathcal {R}_{\Lambda }^{\mathcal {T}}: [0,1) \times {\mathbb N} \to {\mathbb N} \cup \{ +\infty \}$ given by
where, by convention, $\inf \emptyset = \infty $ . We then define the repetitivity portion of $\Lambda $ with respect to $\mathcal {T}$ as
where we use the convention that $\zeta (\delta ) = 0$ if there is some $m \in {\mathbb N}$ such that $\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m) = \infty $ . We say that $\Lambda $ is almost tempered repetitive with respect to $\mathcal {T}$ if $\zeta (\delta )> 0$ for all $\delta> 0$ . We say that $\Lambda $ is tempered repetitive with respect to $\mathcal {T}$ if $\zeta (0)> 0$ .
Remark 3.6. Some comments on the definition are in order.
-
(i) Since $(T_m)$ displays asymptotic invariance from the left, we use patterns arising from the inverse sequence $(T_m^{-1})$ in the above definitions. In certain situations of interest, one can choose the $T_m$ to be symmetric, for instance, if they arise as certain closed balls with respect to a suitable adapted metric, cf. §4.2.
-
(ii) Almost tempered repetitivity is the condition on Delone sets which will enable the proof of our ergodic theorem. The terminology ‘tempered’ is chosen because in the literature, it refers to a growth condition of Følner sequences used in the proofs of ergodic theorems.
-
(iii) Since the repetitivity index is monotonically decreasing in $\delta $ , tempered repetitivity implies almost tempered repetitivity.
-
(iv) If $\Lambda $ is tempered repetitive, then $\mathcal {R}_{\Lambda }^{\mathcal {T}}(0,m) < \infty $ for all $m \in {\mathbb N}$ , and if $\Lambda $ is almost tempered repetitive, then $\mathcal {R}_{\Lambda }^{\mathcal {T}}(\delta ,m) < \infty $ for all $m \in {\mathbb N}$ and $\delta>0$ . This can be used to establish minimality of the associated Delone dynamical system.
Proposition 3.7. (Minimality)
A weighted Delone set $\Lambda $ is almost repetitive if and only if $\mathcal {R}_{\Lambda }^{\mathcal {T}}(\delta , m) < \infty $ for all $\delta> 0$ and all $m \in {\mathbb N}$ . In particular, every almost tempered repetitive Delone set is almost repetitive, and hence the associated Delone dynamical system is minimal.
Proof. We note that by the properties of a strong exhaustion sequence, we find for all $r> 0$ some $\ell (r) \in {\mathbb N}$ such that $\overline {B}_r \subseteq \mathring {T}_{\ell (r)}$ and for each $s \in {\mathbb N}$ , we find some $L(s) \in {\mathbb N}$ such that $T_s \subseteq B_{L(s)}$ . Moreover, by the left-invariance of the metric $d_G$ , we have ${B_l = B_l^{-1}}$ and $\overline {B}_l = \overline {B}_l^{-1}$ for all $l> 0$ . Thus, the condition $\mathcal {R}_{\Lambda }^{\mathcal {T}}(\delta , l) < \infty $ for all $\delta> 0$ and all $l \in {\mathbb N}$ implies for a given $R \geq 1$ , every $B_R$ -pattern $\delta $ -occurs in each ball of radius $L ( \mathcal {R}_{\Lambda }^{\mathcal {T}}(\delta , \ell (R)) )$ . Conversely, for $m \in {\mathbb N}$ , we have that every $\mathring {T}_m^{-1}$ -pattern must $\delta $ -occur in every ball of radius $\varrho (\delta , L(m))$ (with $\varrho $ denoting an almost repetitivity function), whence also in each $h\mathring {T}_M^{-1}$ with $M = \ell ( \varrho (\delta , L(m)))$ .
The following example shows that (almost) tempered repetitivity depends crucially on the underlying strong Følner exhaustion sequence.
Example 3.8. Consider the Delone set $D = {\mathbb Z}$ (seen as weighted Delone set with constant weight $1$ ) in $G = {\mathbb R}$ . Then, D is tempered repetitive with respect to the sequence ${\mathcal {T} = (\overline {B}_m)_m}$ , where $\overline {B}_l$ is the closed ball around $0$ with radius l with respect to the Euclidean metric. Now consider the sequence ${\mathcal {T}^{\prime } = (\overline {B}_{r_m})_m}$ with $r_1 := 1$ and ${r_{m+1} := 2^{r_m}}$ for $m \geq 1$ . Note that an open ball with integer radius $M \in {\mathbb N}$ and integer center contains $2M-1$ points of ${\mathbb Z}$ , while balls of the same size around a non-integer center contain $2M$ points of ${\mathbb Z}$ . This implies $\mathcal {R}_D^{\mathcal {T}^{'}}(M) \geq M+1$ for all $M \in {\mathbb N}$ . However, we have $\lim _m ({2r_m - 1})/({2r_{m+1} - 1}) = 0$ . Hence, D is not (almost) tempered repetitive with respect to $\mathcal {T}^{\prime }$ .
3.3 Weight functions and $\varepsilon $ -quasi-tilings
For the purpose of the following definition, we denote by $ \mathcal {RK}(G)$ the set of all relatively compact subsets of G. We also denote by X a compact metrizable space on which G acts jointly continuously; in our applications, $G \curvearrowright X$ will always be a Delone dynamical system.
Definition 3.9. A function $w: \mathcal {RK}(G) \times X \to {\mathbb R}$ is called an almost sub-additive weight function over X if:
-
(W1) $w(\emptyset , x) = 0$ for all $x \in X$ (normalization);
-
(W2) there is a compact $J \subseteq G$ and an $\eta> 0$ such that
$$ \begin{align*} | w(L,x) - w(K,x) | \leq \eta\cdot ( m_G(L \setminus K) + m_G(\partial_J(L)) + m_G(\partial_J(K)) ) \end{align*} $$for all $x \in X$ and all $K, L \in \mathcal {RK}(G)$ with $K \subseteq L$ (almost-monotonicity); -
(W3) there is a compact subset $B \subseteq G$ and a $\theta \geq 0 $ such that for any finite collection of pairwise disjoint $K_i \in \mathcal {RK}(G)$ , we have
$$ \begin{align*} w\bigg( \bigsqcup_{i} K_i,\,x \bigg) \leq \sum_{i} w(K_i, x) + \theta \cdot \sum_{i} m_G(\partial_B(K_i)) \end{align*} $$for all $x \in X$ (almost sub-additivity); -
(W4) there is a compact subset $I \subseteq G$ and a $\vartheta \geq 0$ such that for each $x \in X$ , $K\in \mathcal {RK}(G)$ , and $h\in G$ , we have
$$ \begin{align*} | w(K,x) - w(Kh^{-1},hx) | \leq \vartheta\cdot m_G(\partial_I(K)) \quad \operatorname{(almost-equivariance)}\kern-1.5pt. \end{align*} $$
Remark 3.10. If $w: \mathcal {RK}(G) \times X \to {\mathbb R}$ is any function satisfying properties (W1) and (W2), then for every relatively compact subset $L \subseteq G$ and every $x \in X$ , we have the boundedness property
From this inequality, we deduce that if $w: \mathcal {RK}(G) \times X \to {\mathbb R}$ is an almost sub-additive weight function and $\mathcal {RK}(G)_+$ denotes the set of relatively compact sets of positive Haar measure, then we can define functions $w^{+}, w^{-}: \mathcal {RK}(G)_+ \times X \to {\mathbb R}$ by
We will study these functions for a particular class of almost sub-additive weight functions over a Delone dynamical system.
Definition 3.11. Let $\mathcal H_\Lambda $ be the hull of a weighted Delone set $\Lambda $ in G. Then, an almost sub-additive weight function over $\mathcal H_\Lambda $ is called an admissible weight function provided:
-
(W5) if $(T_m)$ is any strong Følner exhaustion sequence, then for all $\varepsilon>0$ , there is a $\delta> 0$ and $m_0\in {\mathbb N}$ such that for all $m\geq m_0$ ,
$$ \begin{align*} d_{T_m^{-1}}(\Pi, \Phi) \leq \delta \Longrightarrow |w(T_m, \Phi) -w(T_m, \Pi)|\leq \varepsilon \cdot m_G(T_m), \end{align*} $$where $d_{T_m^{-1}}(\Pi , \Phi )$ is defined as in equation (2).
Remark 3.12. Let $(X,d)$ be a compact metric space on which G acts continuously. Then, we can study almost sub-additive weight functions that additionally satisfy the following.
-
(W5*) If $(T_m)_{m\in {\mathbb N}}$ is a strong Følner exhaustion sequence and d is a metric generating the topology of X, then for all $\varepsilon>0$ , there is an $m_0\in {\mathbb N}$ such that for all $m\geq m_0$ , one can find $\delta _m> 0$ such that
$$ \begin{align*} d(x,y) \leq \delta_m \Longrightarrow |w(T_m, x) -w(T_m, y)|\leq \varepsilon m_G(T_m). \end{align*} $$
This condition has the advantage that it can be defined for general (that is, not necessarily Delone) dynamical systems. While it is sufficient for the proof of the Ornstein–Weiss type lemma (Lemma 3.17), it is not sufficient for the proof of our main theorem.
Let us sketch an argument how axiom (W5) implies axiom (W5 $^\ast $ ) for $X:= \mathcal H_\Lambda $ . First, we can define a metric $d_{\ast }$ on $\mathcal {H}_\Lambda $ by
where $\Phi :=(P,\alpha ),\Pi :=(Q,\beta )\in \mathcal {H}_\Lambda $ , and $c_\ast>0$ is chosen in such a way that $B_{c_\ast }\subseteq U$ and $c_\ast <(2\sigma )^{-1}$ . Now, the metric $d_{\ast }$ generates the topology of $\mathcal H_\Lambda $ by Proposition B.1 in Appendix B, and axiom (W5) implies axiom (W5 $^*$ ) for this specific choice of metric. However, a straightforward computation shows that if axiom (W5 $^\ast $ ) holds for some metric d on $\mathcal {H}_\Lambda $ defining the topology, then it holds for any such metric. We thus conclude that indeed axiom (W5) implies axiom (W5 $^\ast $ ) for $X:= \mathcal H_\Lambda $ .
At this point, we have defined all the terms in the statement of Theorem 3.1. We now turn to its proof.
3.4 $\varepsilon $ -quasi-tilings and the proof of the abstract ergodic theorem
Our proof of Theorem 3.1 will follow the same strategy as the proofs in [Reference Frettlöh and Richard24, Reference Lagarias and Pleasants30]. The main difference is that we have to replace box tilings in Euclidean space with the $\varepsilon $ -quasi tile machinery which was developed by Ornstein and Weiss [Reference Ornstein and Weiss41]. We use the formulation as in [Reference Pogorzelski and Schwarzenberger47] and recall the necessary definition.
Definition 3.13. Let $A \subseteq G$ be a relatively compact subset and $\varepsilon> 0$ . Set $N(\varepsilon ):=\lceil -\log ({\varepsilon }/({1-\varepsilon }))\rceil $ and let $S_1, \ldots , S_{N(\varepsilon )}$ be compact subsets of G and let $C_1^A, \ldots , C_{N(\varepsilon )}^A$ be finite subsets.
We say that A is $\varepsilon $ -quasi tiled by the prototiles $S_i^{\varepsilon }$ with center sets $C_i^A$ if the following conditions are satisfied.
-
(T1) $S_i^\varepsilon C_i^A \subseteq A$ for $1\leq i\leq N(\varepsilon )$ .
-
(T2) For each $1\leq i\leq N(\varepsilon )$ and every $c\in C_i^A$ , there is a measurable set ${\widetilde {S}}_i^\varepsilon (c)\subseteq S_i^\varepsilon c$ satisfying:
-
• $(1-\varepsilon )m_G(S_i^\varepsilon c) \leq m_G({\widetilde {S}}_i^\varepsilon (c)) \leq m_G(S_i^\varepsilon c)$ ;
-
• $\bigcup _{c\in C_i^A}S_i^\varepsilon c = \bigsqcup _{c\in C_i^A} {\widetilde {S}}_i^\varepsilon (c)$ , where the latter union consists of pairwise disjoint sets.
-
-
(T3) $S_i^\varepsilon C_i^A \cap S_j^\varepsilon C_j^A =\emptyset $ for $1\leq i < j\leq N(\varepsilon )$ .
-
(T4) $m_G(\bigsqcup _{i=1}^{N(\varepsilon )} S_i^\varepsilon C_i^A) \geq (1-2\varepsilon ) m_G(A)$ .
The following theorem guarantees the existence of certain $\varepsilon $ -quasi-tilings. It is a weaker statement than [Reference Pogorzelski and Schwarzenberger47, Theorem 4.4(a)] and a slightly stronger version of [Reference Ornstein and Weiss41, Theorem 6]. Given relatively compact subsets $S, T \subseteq G$ and $\delta>0$ , we will say that T is $(S,\delta )$ -invariant if
Theorem 3.14. (Existence of $\varepsilon $ -quasi tilings)
Let $0<\varepsilon <1/10$ and $N(\varepsilon ):= \lceil -\log ({\varepsilon }/({1-\varepsilon }))\rceil $ . Then for every amenable unimodular lcsc group G, strong Følner exhaustion sequence $(S_l)_{l\in {\mathbb N}}$ , and natural number n, there exist sets $S_1^\varepsilon , \ldots , S_{N(\varepsilon )}^\varepsilon $ with the following properties:
-
(i) $S_i^\varepsilon \in \{S_l \,:\, l\geq \max \{i,n\}\}$ ;
-
(ii) $S_n\subseteq S_1^\varepsilon \subseteq \cdots \subseteq S_{N(\varepsilon )}^\varepsilon $ ;
-
(iii) there is some $\delta _0> 0$ such that for every $( S^{\varepsilon }_{N(\varepsilon )} S^{\varepsilon \, -1}_{N(\varepsilon )}, \delta _0 )$ -invariant compact set ${A \subseteq G}$ , there are finite sets $C_i^A \subseteq A$ such that A is $\varepsilon $ -quasi tiled by the prototiles $S_i^{\varepsilon }$ with center sets $C_i^A$ .
We give some remarks on $\varepsilon $ -quasi tilings.
Remark 3.15.
-
(i) According to condition (T3), the set A can be $\varepsilon $ -quasi tiled by $S_i^\varepsilon \,,\, 1\leq i\leq N(\varepsilon )\,$ such that tiles of different type do not overlap. However, tiles of the same type might overlap, see condition (T2). However, condition (T2) asserts that these sets are $\varepsilon $ -disjoint, which means one can remove from them portions of measure at most $\varepsilon $ to obtain disjoint sets.
-
(ii) In fact, these trimmed sets can be proven to maintain certain invariance conditions, cf. [Reference Pogorzelski and Schwarzenberger47, part (b) of Theorem 4.4]. More precisely, for a given compact $B \subseteq G$ and $0 < \zeta < \varepsilon $ , we can make sure that $m_G(\partial _B(S_i^{\varepsilon })) / m_G(S_i^{\varepsilon }) \leq \zeta ^2$ and $m_G(\partial _B(\tilde {S}_i^{\varepsilon }(c))) / m_G(\tilde {S}_i^{\varepsilon }(c)) \leq 4\zeta $ for all $1 \leq i \leq N(\varepsilon )$ and all $c \in C_i^m$ . We will need this latter fact only once in this paper, namely for the proof of Lemma 3.16 which is given in Appendix A.
-
(iii) Given any strong Følner exhaustion sequence $(A_m)$ , the theorem implies that for a given $\varepsilon $ , one finds sets $S_i^{\varepsilon }$ ( $1 \leq i \leq N(\varepsilon )$ ) extracted from a strong Følner exhaustion sequence (which possibly is $(A_m)$ itself), along with $M \in {\mathbb N}$ , such that for all ${m \geq M}$ , one finds finite sets $C_i^m \subseteq A_m$ such that conditions (T1)–(T4) hold true for $A= A_m$ and $C_i^A = C_i^m$ .
-
(iv) The assertion (T4) given above is weaker than [Reference Pogorzelski and Schwarzenberger47, condition (iv) in Definition 4.1], which provides precise quantitive information on the portion of A covered by the sets $S_i^{\varepsilon } C_i^A$ for a fixed $1 \leq i \leq N(\varepsilon )$ . As shown in [Reference Pogorzelski and Schwarzenberger47, Remark 4.3], these estimates result in assertion (T4).
The proof of Theorem 3.1 rests on two lemmas whose proofs build on known techniques used in the context of sub-additive and almost-additive Ornstein–Weiss type lemmas, see e.g. [Reference Ceccherini-Silberstein, Coornaert and Krieger16, Reference Gromov27, Reference Krieger29, Reference Ornstein and Weiss41]. For the sake of self-containment, we give their proofs in Appendix A. The first lemma shows that weight functions are compatible with respect to $\varepsilon $ -quasi tilings of Følner sets.
Lemma 3.16. Let G be an amenable unimodular lcsc group. Let $0 < \varepsilon < 1/10$ and a strong Følner exhaustion sequence $(S_n)$ be given. Suppose further that $v:\mathcal {RK}(G) \to {\mathbb R}$ is a function for which there are $\eta (v), \theta (v) \geq 0$ as well as compact subsets $J, B \subseteq G$ such that:
-
(w1) $v(\emptyset ) = 0$ ;
-
(w2) $| v(L) - v(K) | \leq \eta (v)\, ( m_G(L \setminus K) + m_G(\partial _J(L)) + m_G(\partial _J(K)) ) $ for all $K, L \in \mathcal {RK}(G)$ with $K \subseteq L$ ;
-
(w3) $ v(\sqcup _i K_i) \leq \sum _i v(K_i) \, + \, \theta (v) \sum _i m_G(\partial _B(K_i)) $ for finitely many pairwise disjoint sets $K_i \in \mathcal {RK}(G)$ .
Then, for every compact subset $I \subseteq G$ , there is some $m_{I} \in {\mathbb N}$ such that for every $n \geq m_{I}$ , one finds $(I, \varepsilon )$ -invariant sets $S_i^{\varepsilon } \in \{S_l:\, l \geq \max \{n,i\}\}$ ( $1 \leq i \leq N(\varepsilon ) = \lceil ({-\varepsilon })/({\log (1-\varepsilon )}) \rceil $ ) as well as $\delta _0> 0$ such that each compact subset $A \subseteq G$ which is $( S_{N(\varepsilon )}^{\varepsilon }S_{N(\varepsilon )}^{\varepsilon \, -1}, \delta _0 )$ -invariant and at the same time $(L, \varepsilon )$ -invariant for $L \in \{J,B,I\}$ can be $\varepsilon $ -quasi tiled by prototiles $S_i^{\varepsilon }$ and finite center sets $C_i^A$ for $1 \leq i \leq N(\varepsilon )$ such that, in addition, one has
The second lemma is about convergence of the functions $w^+$ from equation (3).
Lemma 3.17. Let $w: \mathcal {RK}(G) \times X \to {\mathbb R}$ be an almost sub-additive weight function and $(T_m)$ be a strong Følner exhaustion sequence in G. Then, the limits
exist for all $x \in X$ . Moreover, the limits do not depend on the sequence $(T_m)$ . If, in addition, w satisfies condition (W5 $^\ast $ ) and if the action $G \curvearrowright X$ is minimal, then there is a value $I^+_w \in {\mathbb R}$ such that
Remark 3.18. If $-w$ instead of w is a sub-additive weight function (in particular, if w is actually additive), then the corresponding statements hold for $w^{-}$ instead of $w^+$ .
Proof of Theorem 3.1
Let $w: \mathcal {RK}(G) \times \mathcal {H}_{\Lambda } \to {\mathbb R}$ be an admissible weight function and define $w^+$ and $w^-$ by equation (3).
By Proposition 3.7, the dynamical system $G \curvearrowright X$ is minimal. However, the function w satisfies axiom (W5) and therefore also axiom (W5 $^\ast $ ) by Remark 3.12, and hence the strong form of Lemma 3.17 applies. We conclude that the limit $\nu ^{+} = \lim _{n \to \infty } w^{+}(T_m,\Pi )$ exists uniformly over $\Pi $ and is independent of the choices of $\Pi \in \mathcal {H}_{\Lambda }$ and the strong Følner exhaustion sequence.
Moreover, we define $\nu ^{-}(\Pi ) := \liminf _{n \to \infty } w^{-}(T_n, \Pi )$ for $\Pi \in \mathcal {H}_{\Lambda }$ . It follows from Lemma A.1 in Appendix A that $\nu ^{-} = \nu ^{-}(\Pi )$ is independent of the choice of $\Pi \in \mathcal {H}_{\Lambda }$ and that the limit inferior is realized by one subsequence $(n_l)$ of the integers simultaneously for all $\Pi $ . In light of that, it suffices to prove that
To this end, we fix a subsequence $(m_l)$ of the integers such that $\lim _{l \to \infty } w^{-}(T_{m_l}, \Lambda ) = \nu ^{-}$ . For the sake of simpler notation, we set $T_l^{\prime } := T_{m_l}$ for all $l \in {\mathbb N}$ . We assume by contradiction that $\nu ^{-} < \nu ^{+}$ . Let $0 < \varsigma < ({\nu ^{+} - \nu ^{-}})/{4}$ and choose $L_{\varsigma } \in {\mathbb N}$ such that $|\nu ^{-} - w^{-}(T^{\prime }_l,\, \Lambda )| < \varsigma $ for all $l \geq L_{\varsigma }$ . Then, for $l \geq L_{\varsigma }$ , one finds $g_l \in G$ such that ${w(T_l^{\prime } g_l, \Lambda )}/{m_G(T_l^{\prime })} \leq w^{-}(T_l^{\prime },\Lambda ) + \varsigma $ . Combined with the previous estimate and with the choice of $\varsigma $ , we conclude from $4 \varsigma < \nu ^{+} - \nu ^{-}$ that
for all $l \geq L_{\varsigma }$ . Writing $\Lambda = (P,\alpha )$ , we define the $S^{-1}_l$ -patches $q_l:= \Lambda _{S_l^{-1}}$ with inverse support $S_l := T_l^{\prime } g_l = T_{m_l} g_l$ . Using axiom (W5), we find some $0< \delta < 1$ and an $m_0 \in {\mathbb N}$ such that for all $l \geq m_0$ , the condition $d_{S^{-1}_{l}}(\Pi , \Phi ) \leq \delta $ implies $|w(S_{l}, \Pi ) - w(S_{l},\Phi )| < \varsigma m_G(S_{l})$ for $\Pi , \Phi \in \mathcal {H}_{\Lambda }$ . We fix this $\delta $ and take an arbitrary $h \in G$ . Since $\Lambda $ is almost tempered repetitive with respect to $\mathcal {T} = (T_m)$ , the pattern arising from the patches $q_l$ must $\delta $ -occur in $h^{-1}T^{-1}_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)}$ , where $\mathcal {R}^{\mathcal {T}}_{\Lambda }$ is the repetitivity index for $\Lambda $ with respect to $\mathcal {T}$ . This means that there must be some element $\widetilde {g}_{l,h} \in G$ such that $S_l \widetilde {g}_{l,h} \subseteq T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)}h$ and $d_{S_l^{-1}}( \Lambda ,\, \widetilde {g}_{l,h}\cdot \Lambda ) \leq \delta $ . This yields
for $l \geq m_0$ . Furthermore, increasing $m_0$ if necessary and invoking axiom (W4), we find that
for $l \geq m_0$ . We point out that $m_0$ only depends on parameters given by the weight function w and on $\varsigma $ but not on objects constructed in the proof. Given h as above, define ${A_{l,h}:= T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)}h \setminus S_l \widetilde {g}_{l,h}} $ . We will distinguish the following two cases.
-
(A) There is some $0 < \kappa _{\Lambda }(\delta ) < 1$ such that ${m_G(T_{m_l})}/{m_G(T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)})} \leq \kappa _{\Lambda }(\delta )$ for all $l \in {\mathbb N}$ .
-
(B) There is some subsequence $(T_{m_{l_k}})_k$ such that $\lim _{k \to \infty } ({m_G(T_{m_{l_k}})}/ {m_G(T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_{l_k})})}) = 1$ .
We will cover both of these cases separately and, in both cases, we will obtain $\nu ^{+} \leq \nu ^{+} - c(\varsigma )$ for some $c(\varsigma )> 0$ , which is clearly a contradiction. We start with case (A).
Case (A). Invoking the general relations
for sets $C,D,L \in \mathcal {RK}(G)$ and $g \in G$ , we find
for all relatively compact sets $K \subseteq G$ . This shows that the left-hand side of the above inequality tends to $0$ uniformly in h if l tends to infinity. In particular, the sequences $(A_{l,h})_l$ are strong Følner sequences for all $h \in G$ . Now, let $0 < \varepsilon < 1/10$ and choose some ${n_{\varepsilon } \geq \max \{m_0, L_{\varsigma }\}}$ such that
for all $n \geq n_{\varepsilon }$ . We also make sure that $n_{\varepsilon }$ is chosen large enough such that
where $L \in \{ J,B,I\}$ and $J,B,I$ are the compact subsets of G determined by w as of axioms (W2), (W3), and (W4). We fix $n \geq n_{\varepsilon }$ and we apply Lemma 3.16 to find prototile sets
as well as $M \in {\mathbb N}$ with $M \geq n$ such that for all $l \geq M$ , we have $m_l \geq n$ and for all $h \in G$ , we find a finite set $C_i^{l,h} \subseteq A_{l,h}$ for each $1 \leq i \leq N(\varepsilon )$ such that the properties (T1)–(T4) from Definition 3.13 are satisfied and at the same time, we obtain
where the constant $t_1 \geq 0$ only depends on the parameters $\eta , \theta , \vartheta $ given by the definition of the weight function w. We emphasize at this point that the fact that the parameters $n, M$ (and hence also l) can be chosen independently of h is justified by the observation that the validity of Lemma 3.16 depends on the invariance properties of $A_{l,h}$ with respect to the prototile sets, combined with the fact that $\lim _l m_G(\partial _{K}(A_{l,h}))/m_G(A_{l,h}) = 0$ uniformly over h for all $K \in \mathcal {RK}(G)$ .
Exploiting the sub-additivity property (W3) of w and recalling $A_{l,h}:= T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)}h \setminus S_l \widetilde {g}_{l,h}$ , we find with the estimates in equations (6) and (7), as well as with $l \geq M$ , that
for some constant $t_2 \geq 0$ which only depends on the sub-additivity parameter $ \theta $ of w and on $\delta $ . Recall that $l \geq M \geq m_0$ . Thus, we can combine the latter inequality with the inequality (5) and $m_G(S_l)=m_G(T^{\prime }_l)$ to obtain
Invoking the sub-additivity induced by the above $\varepsilon $ -quasi tilings and using the basic inequality $m_G(A_{l,h}) \leq m_G(T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)})$ , the latter inequality transforms to
Note that the tiling properties (T1)–(T4) from Definition 3.13 yield
Furthermore, inequality (4) leads to
for $l \geq M \geq L_{\varsigma }$ . Recall that we have chosen $n \geq n_{\varepsilon }$ and $S_i^{\varepsilon } \in \{ T_k: k \geq \max \{ i,n \} \}$ . Thus,
holds for all $1 \leq i \leq N(\varepsilon )$ , all $l \geq M$ , each $h \in G$ , and each $c \in C_i^{l,h}$ .
Since $\Lambda $ is almost tempered repetitive, the repetitivity portion $\zeta :[0,1) \to [0,1]$ of $\Lambda $ with respect to $\mathcal {T} = (T_m)$ is positive everywhere and, in particular, we have
We set $a_{\varepsilon } := (1-2\varepsilon )$ if $\nu ^+ \leq 0$ , and $a_{\varepsilon } := (1-\varepsilon )^{-1}$ if $\nu ^{+}> 0$ and $t_3 := t_1 + t_2$ . Since $a_{\varepsilon } \nu ^{+} - \nu ^{+}\geq 0$ , we finally derive from the above estimates
for all $l \geq M$ and all $h \in G$ , where the tempered repetitivity was used in the last step. Thus, taking the supremum over all $h \in G$ on the left-hand side of the latter inequality and sending $l \to \infty $ , we find that
By sending $\varepsilon \to 0$ , we obtain $\nu ^{+} \leq \nu ^{+} - 9/10 \zeta (\delta ) \varsigma $ and since both $\varsigma $ and $\zeta (\delta )$ are positive numbers, we have arrived at a contradiction. Hence, we conclude $\nu ^{+} = \nu ^{-}$ in case (A).
Case (B) is much easier to handle. We assume with no loss of generality that
Then, there is some $M_{\varsigma } \in {\mathbb N}$ such that $\sup _{h \in G} m_G(A_{l,h}) (m_G(T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)}))^{-1} < {\varsigma }/{5\eta }$ for all $l \geq M_{\varsigma }$ , where $\eta> 0$ is the boundedness constant of w as of axiom (W2). Increasing $M_{\varsigma }$ if necessary, we can assume that $\sup _{h \in G} m_G( \partial _L(A_{l,h})) (m_G(T_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m_l)}))^{-1} < ({1}/{10})\varsigma (\max \{1,\eta ,\theta \})^{-1}$ and ${m_G(\partial _L(S_l))}/{m_G(S_l)} < {1}/{10} \varsigma (\max \{1,\theta \})^{-1}$ for all ${L \in \{ J,B,I \}}$ and each $l \geq M_{\varsigma }$ , where $J,B,I \subseteq G$ are compact and $\theta \geq 0$ depend on the weight function $w$ . We now fix $l \geq M_{\varsigma }$ . By the choice of $\widetilde {g}_{l,h}$ , inequalities (4) and (5) together with $m_G(S_l)=m_G(T^{\prime }_l)$ imply
for all $l \geq M_{\varsigma }$ and each $h \in G$ . Then, $m_G(\partial _L(C)) = m_G(\partial _L(Cg))$ , the sub-additivity property (W3) with constant $\theta \geq 0$ (given by $w$ ), and the imposed invariance conditions imply
for all $l \geq M_{\varsigma }$ and all $h \in G$ . Invoking Remark 3.10 with constant $\eta> 0$ applied to $w(A_{l,h},\, \Lambda )$ , then the invariance condition on $A_{l,h}$ and the implications explained at the beginning of the proof of Case (B) yield
for all $l \geq M_{\varsigma }$ and all $h \in G$ . Thus, by taking first the supremum over all $h \in G$ and then sending $l \to \infty $ , we arrive at
which is a contradiction since $\varsigma> 0$ .
4 Unique ergodicity
4.1 Unique ergodicity from tempered repetitivity
We are now in a position to deduce Theorem 1.3 from our abstract sub-additive convergence theorem, Theorem 3.1. The idea to connect ergodicity properties of dynamical systems $G \curvearrowright X$ to the existence of limits of the form $\lim _{m \to \infty } {w(T_m,\Lambda )}/{m_G(T_m)}$ for some class of almost-additive functions $w:\mathcal {RK}(G) \times X \to {\mathbb R}$ and strong Følner exhaustion sequence $(T_m)$ is well established in the abelian case [Reference Damanik and Lenz19, Reference Frettlöh and Richard24, Reference Lagarias and Pleasants30, Reference Lenz32]. While the proof of the sub-additive convergence theorem was more complicated in the non-abelian case than in the abelian case, the application to the unique ergodicity statement is not much different, and hence we will be brief.
For the remainder of this subsection, we work in the setting of Theorem 1.3. Thus, $\mathcal T = (T_m)$ is a strong Følner exhaustion sequence in an amenable unimodular lcsc group G and $\Lambda $ is a weighted Delone set in G, which is almost tempered repetitive with respect to $\mathcal T$ . We then denote by $\mathcal H_\Lambda $ the hull of $\Lambda $ . To apply Theorem 3.1 in this context, we need to construct suitable weight functions.
Given $f \in C(\mathcal {H}_{\Lambda })$ , we define the mapping
Proposition 4.1. For every $f \in C(\mathcal {H}_{\Lambda })$ , the function $w_f$ given by equation (8) is an almost sub-additive weight function over $\mathcal H_\Lambda $ .
Proof. We have to check axioms (W1)–(W4): axiom (W1) holds by the convention of the empty integral being equal to $0$ . Furthermore, axiom (W2) holds with $J = \emptyset $ and ${\eta = \| f\|_{\infty }}$ . The additivity of the integral yields axiom (W3) (which holds as an equality) with $B = \emptyset $ and $\theta = 0$ . As for axiom (W4), a straightforward integral substitution allows us to choose $\vartheta = 0$ and $I = \emptyset $ .
The key step is now to show that for $f \in C(\mathcal {H}_{\Lambda })$ , the weight function $w_f$ is actually admissible, that is, that it also satisfies axiom (W5). We will use the fact that, by the same proof as in the abelian case [Reference Frettlöh and Richard24, Reference Müller and Richard39], a metric generating the weak- $*$ -topology on $\mathcal {H}_\Lambda $ is defined via
where $\Phi :=(P,\alpha ),\Pi :=(Q,\beta )\in \mathcal {H}_\Lambda $ and $c_\ast>0$ are such that $B_{c_\ast }\subseteq U$ and ${c_\ast <(2\sigma )^{-1}}$ . We will use this metric in the proof of the following lemma.
Lemma 4.2. For every $f\in \mathcal {C}(\mathcal {H}_\Lambda )$ , the weight function $w_f$ is admissible with respect to any strong Følner exhaustion sequence $(T_m)$ .
Proof. We fix $f\in \mathcal {C}(\mathcal {H}_\Lambda )$ , a strong Følner exhaustion sequence $(T_m)$ , and some $\varepsilon> 0$ throughout. We then have to show the existence of some $\delta>0$ and an $m_0\in {\mathbb N}$ such that for all $m\geq m_0$ , we have the implication
Since f is uniformly continuous on the compact space $\mathcal {H}_{\Lambda }$ , there is a $0<\delta <c_\ast $ such that the condition $d_{\ast }(\Pi ,\Theta ) \leq \delta $ implies $|f(\Pi ) - f(\Theta )| \leq {\varepsilon }/{3}$ . Since $(T_m)$ is a strong Følner exhaustion sequence and $B:=B_{1/\delta }\in \mathcal {RK}(G)$ , we find some $m_0 \in {\mathbb N}$ such that ${m_G(\partial _B(T_m))}/{m_G(T_m)}\leq {\varepsilon }/{3\|f\|_\infty }$ and $B \subseteq T_m$ for all $m\geq m_0$ . It follows that for all $\Theta \in \mathcal {H}_{\Lambda }$ ,
Note that since $e \in B$ and $B=B^{-1}$ , we deduce $Bg \subseteq T_m$ for $g \in T_m \setminus \partial _B(T_m)$ . Let $\Theta ,\Pi \in \mathcal {H}_\Lambda $ such that $d_{T^{-1}_m}(\Pi ,\Theta )\leq \delta $ . Thus, we conclude $d_{g^{-1}B^{-1}}(\Pi ,\Theta )\leq \delta $ , and hence $d_{B}(g.\Pi , g.\Theta )=d_{B^{-1}}(g.\Pi , g.\Theta )\leq \delta $ for all $g\in T_m\setminus \partial _B(T_m)$ . Since $\delta <c_\ast $ , $d_{\ast }(g.\Pi ,g.\Theta )<\delta $ follows by the definition of $d_\ast $ for all $g\in T_m\setminus \partial _B(T_m)$ . Thus, $g\in T_m\setminus \partial _B(T_m)$ yields $|f(g.\Pi )-f(g.\Theta )|\leq {\varepsilon }/{3}$ and we arrive at
Now, equation (9) follows from the triangle inequality.
We can now prove unique ergodicity for almost tempered repetitive Delone sets $\Lambda $ in G.
Proof of Theorem 1.3
Suppose that $\Lambda $ is almost tempered repetitive with respect to the strong Følner exhaustion sequence $(T_m)$ . Let $f \in C(\mathcal {H}_{\Lambda })$ . Since $w_f$ , as considered above, is an admissible weight function, we deduce from Theorem 3.1 that there is a number $I(f) \in {\mathbb R}$ such that
for all $\Pi \in \mathcal {H}_{\Lambda }$ . We claim that for an arbitrary G-invariant Borel probability measure $\nu $ on $\mathcal {H}_{\Lambda }$ , we get $I(f) = \int _{\mathcal {H}_{\Lambda }} f\, d\nu $ . Indeed, combining the latter limit relation with the dominated convergence theorem, Fubini’s theorem, and the G-invariance of $\nu $ , we arrive at
This shows that for any two G-invariant Borel probability measures $\nu _1$ and $\nu _2$ , we obtain
for all $f \in C(\mathcal {H}_{\Lambda })$ . This, in turn, is equivalent to $\nu _1 = \nu _2$ . Therefore, the hull of $\Lambda $ is uniquely ergodic. It is also minimal by Proposition 3.7.
4.2 Metrics of exact polynomial growth and linear repetitivity
In the previous subsection, we have established Theorem 1.3 which asserts that every tempered repetitive weighted Delone set with respect to a strong Følner exhaustion sequence in an lcsc amenable group is uniquely ergodic and minimal. While the notion of a Delone set was originally defined using a (continuous) adapted metric on G, it does not depend on any choice of metric on G, but only on G as a topological group. The same metric independence thus holds for the statement of the theorem. In the present subsection, we would now like to relate tempered repetitivity to the more classical notion of linear repetitivity; this notion, unlike the previous ones, does depend on a choice of metric. To obtain a metric version of Theorem 1.3, we will have to choose a metric with the additional property of exact polynomial volume growth. As we will see in Theorem 4.13 below, such metrics exist on all compactly generated groups of polynomial growth (e.g., word metrics with respect to compact generating sets), but it seems to be unknown in which generality they can be chosen to be continuous. For this reason, we insist for the remainder of this section that d is an adapted metric on G, which is not necessarily continuous. We still use the same notation as before concerning open and closed balls, but we emphasize that without the continuity assumption, $B_t$ need not be open and $\overline {B}_t$ need not be closed in the topology of G. However, left-invariance of d still implies that
Since d is locally bounded, we may choose $r_0$ such that $B_{r_0}$ contains an open identity neighborhood. We claim that then
Indeed, the former follows from the fact that every $x \in \overline {B_r}$ admits an open neighborhood which is contained in $xB_{r_0} \subseteq \overline {B_r} \overline {B_{r_0}} \subseteq \overline {B_{r+r_0}} \subseteq \overline {B_s}$ . For the latter, if $x \in \overline {B_r}$ , then there exist $x_n \in B_r$ such that $x_n \to x$ and hence $x_n^{-1}x \in B_{r_0}$ for some sufficiently large $n_0$ , that is, $x \in x_{n_0}^{-1}B_{r_0} \subseteq B_rB_{r_0} \subseteq B_{r+r_0}$ .
If d is continuous, then $B_t$ is open and $\overline {B}_t$ is closed, and we have
Also, if $K \subseteq G$ is compact and d is continuous, then there exists $r_K> 0$ such that ${K \subseteq B_{r_K}}$ . This latter property is invariant under coarse equivalence, and hence holds for all adapted pseudo-metrics on G, even if they are discontinuous by Proposition 2.1. We will be interested in metrics with the following special property (see [Reference Nevo40, §4.4] for background information).
Definition 4.3. (Groups with exact polynomial growth with respect to a metric)
Let G be an lcsc group and let d be an adapted metric on G. We say that G has exact polynomial growth with respect to d if there are constants $C> 0$ and $q \geq 0$ such that
Note that exact polynomial growth is a property of the pair $(G,d)$ , not just of the (topological) group G. It is also not invariant under replacing d by a coarsely equivalent metric, and hence G may have exact polynomial growth with respect to some adapted metrics, but not for others. Note that exact polynomial growth implies that for all $r>0$ , we have
The following observation is crucial for our purposes. Though it is well known to experts (see e.g. [Reference Nevo40, Proposition 4.13] or [Reference Breuillard14, Corollary 1.10] for a slightly different formulation), we give a proof for the sake of being self-contained.
Proposition 4.4. (Balls as Følner sequences)
Let G be an lcsc group that has exact polynomial growth with respect to some adapted metric d. Assume that either:
-
(a) d is continuous and $T_n := \overline {B}_{r_n}$ , where $(r_n)_{n \in \mathbb N}$ is any strictly monotone sequence with $\lim _n r_n = \infty $ ; or
-
(b) $T_n := \overline {B_{r_n}}$ , where $r_{n+1} - r_n> r_0$ for all $n \in \mathbb N$ , where $r_0>0$ is chosen such that $B_{r_0}$ contains an open identity neighborhood.
Then $(T_n)_{n \in \mathbb N}$ is a strong Følner exhaustion sequence.
Proof. To deal with both cases simultaneously, we set $r_0 :=0$ in Case (a). Note first that in either case, the sets $T_n$ are compact by continuity, respectively properness of d (whereas closed balls need not be topologically closed, let alone compact, in Case (b)). Moreover, we have $T_n \subseteq \mathring {T}_{n+1}$ by equations (12) and (11), respectively, and hence $(T_n)$ is a strong exhaustion sequence in either case.
Moreover, by the previous remarks, we can find for every compact subset $K \subseteq G$ some $r_K> 0$ with $K \subseteq B_{r_K}$ . This yields
We claim that if $x \in B_{r_n-r_K}$ , then $x \not \in \partial _K T_n$ . Indeed, otherwise, we would find $y \in B_{r_K}$ and $z \in G \setminus T_n$ such that $x = yz$ and hence $z = y^{-1}x \in B_{r_n} \subseteq T_n$ . However, we deduce that in Case (a), we have $\partial _KT_n \subseteq B_{r_K} \overline {B}_{r_n} \subseteq \overline {B}_{r_K + r_n} \subseteq B_{r_n + r_K + 1}$ . In Case (b), we can use equation (11) to similarly deduce
Thus, in either case, we have
Using equation (13), we obtain
This finishes the proof.
We will see that if one restricts attention to strong Følner exhaustion sequences which arise from metrics of exact polynomial growth as above, then our notion of (almost) tempered repetitivity reduces to the more classical and purely metric notion of (almost) linear repetitivity. Let us recall the relevant definitions.
Definition 4.5. Let G be an lcsc group and let d be an adapted metric on G. A weighted Delone set $\Lambda $ in G is called almost linearly repetitive with respect to d if it is almost repetitive and there exists some function $c_{\Lambda }:(0,\infty ) \to (0,\infty )$ such that
is an almost repetitivity function.
For weighted Delone sets with FLC, it makes sense to work with the following definition.
Definition 4.6. A weighted Delone set $\Lambda $ on G is called linearly repetitive if it is repetitive and there exists a constant $c_{\Lambda }$ such that $\varrho (R) := c_{\Lambda } R,\, R\in [1,\infty )$ is a repetitivity function.
We emphasize that (almost) linear repetitivity is a metric notion, whereas (almost) tempered repetitivity is a notion relative to a given strong Følner exhaustion sequence. However, we have the following relation.
Proposition 4.7. (Linear repetitivity versus tempered repetitivity)
Assume that G is an lcsc group that has exact polynomial growth with respect to some adapted metric d and let $r_0> 0$ such that $B_{r_0}$ is an identity neighborhood. Then, for a weighted Delone set $\Lambda $ in G, the following statements are equivalent:
-
(i) $\Lambda $ is almost linearly repetitive with respect to d;
-
(ii) $\Lambda $ is almost tempered repetitive with respect to every strong Følner exhaustion sequence of the form $\mathcal {T} = (\overline {B_{r_n}})$ with the sequence $(r_n)$ satisfying
$$ \begin{align*} r_0 \leq \inf_n (r_{n+1} - r_n) \leq \sup_{n}( r_{n+1} - r_n) < \infty; \end{align*} $$ -
(iii) $\Lambda $ is almost tempered repetitive with respect to the strong Følner exhaustion sequence $(\overline {B_{nr_0}})$ .
If d is continuous, then we can replace closures of balls by closed balls, the condition in statement (ii), by
and the strong Følner exhaustion sequence in statement (iii) by $(\overline {B}_{n})$ .
If $\Lambda $ has FLC, then we can drop the word ‘almost’ from any of the above equivalences.
Remark 4.8. Note that due to part (b) of Proposition 4.4, the sequences appearing in assertions (ii) and (iii) of Proposition 4.7 are indeed strong Følner exhaustion sequences. In the case where the metric is additionally continuous, we find that the closed balls give rise to strong Følner exhaustion sequences by part (a) of Proposition 4.4. Note that the condition on the sequence $(r_n)$ given in assertion (ii) implies that $r_n < r_{n+1}$ for all $n \in {\mathbb N}$ , as well as $r_n \nearrow \infty $ as $n \to \infty $ . However, even if d is continuous, these two monotonicity conditions are not sufficient for $\Lambda $ being tempered repetitive with respect to $\mathcal {T}$ , as can be inferred by Example 3.8.
Proof. Assume that G has exact polynomial growth with respect to d with parameters say $C> 0$ and $q \geq 0$ , and let $(t_m)$ be any strictly increasing sequence such that $t_m \nearrow \infty $ as $m \to \infty $ . We recall from equation (13) that for all $c> 0$ , there exists a constant $\gamma> 1$ such that for all $m \in {\mathbb N}$ ,
Then, for every $c^{\prime }> 0$ , we find a constant $\beta \geq 1$ such that for all $m \in {\mathbb N}$ and $c \in \{1, c'\}$ ,
The same argument also holds for $\overline {B}_{ct_m}$ instead of $\overline {B_{c t_{m}}}$ (one can set $r_0 = 1$ in this situation).
(i) $\Longrightarrow $ (ii). Assume that $\Lambda $ is almost linearly repetitive with almost repetitivity function $\varrho (\delta ,R) = c_{\Lambda }(\delta ) \cdot R$ (where $c_{\Lambda } = c_{\Lambda }(0)$ in the FLC case) and let $\mathcal T = (T_m)$ , where ${T_m:= \overline {B_{r_m}}}$ . We set $a:= \inf _m (r_{m+1} - r_m)> 0$ , as well as $A := \sup _m (r_{m+1} - r_m) < \infty $ . Then, by the given almost linear repetitivity, we get $\mathcal {R}_{\Lambda }^{\mathcal {T}}(\delta , m) \leq a^{-1} c_{\Lambda }(\delta ) r_{m}$ for all $m \in {\mathbb N}$ . Further, we have that $T_{\mathcal {R}_{\Lambda }^{\mathcal {T}}(\delta , m)} \subseteq \overline {B_{A \mathcal {R}_{\Lambda }^{\mathcal {T}}(\delta , m)}} $ for all $m \in {\mathbb N}$ . In line with the remark made at the beginning of the proof (applied with $t_m = r_m$ and with $c^{\prime } = A a^{-1} c_{\Lambda }(\delta )$ ), we find $\beta = \beta (\delta ) \geq 1$ such that
Hence, the repetitivity portion of $\Lambda $ is bounded from below by $\zeta (\delta ) \geq \beta (\delta )^{-2} \cdot (Aa^{-1})^{-q} c_{\Lambda }(\delta )^{-q}$ , which is positive. This shows that $\Lambda $ is almost tempered repetitive with respect to $\mathcal {T}$ . In the continuous case, the same argument works with $T_m := \overline {B}_{r_m}$ instead of $T_m:= \overline {B_{r_m}}$ .
(ii) $\Longrightarrow $ (iii). This is obvious since the sequence $(r_n)$ with $r_n = r_0n$ (or $r_n = n$ in the continuous case) for all $n \in {\mathbb N}$ satisfies the growth condition given in assertion (ii).
(iii) $\Longrightarrow $ (i). Since the statements are invariant under rescaling the metric, we may assume that $r_0 = 1$ to simplify notation. Assume that $\Lambda $ is almost tempered repetitive with respect to $\mathcal {T} = (\overline {B_n}) $ . Now let $R \geq 1$ and $m := \lceil R \rceil + 1$ such that $m \in {\mathbb N}$ with $m \geq 2$ and $m-2 < R \leq m - 1$ . By assumption, there is a function $\zeta :[0,1) \to [0,1]$ with $\zeta (\delta )>0$ for all $\delta> 0$ (and in the FLC case, $\zeta (0)> 0$ ) such that
Using the considerations made at the beginning of the proof (applied with $t_m =m$ and $c^{\prime } = 1$ ), we obtain $\beta (\delta ) \geq 1$ such that for all $m \in {\mathbb N}$ ,
Note that $m = (m-2) + 2 \leq R + 2R = 3R$ . By the first inclusion given in equation (11) and $R+1 \leq m$ , we get $B_R \subseteq \mathring {\overline {B_{R+1}}} \subseteq \mathring {\overline {B_m}}$ . It follows from the definition of the repetitivity index that every $B_R$ -pattern $\delta $ -occurs in $h \mathring {T}^{-1}_{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m)} = \mathring {\overline {B_{{\mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m)}}(h)}} \subseteq B_{ \mathcal {R}^{\mathcal {T}}_{\Lambda }(\delta , m) + 1}(h) $ for all $h \in G$ , where the latter inclusion follows from the second inclusion in equation (11). We can now choose $\kappa (\delta )> 0$ large enough such that $( \beta (\delta )^2/\zeta (\delta ))^{1/q} 3R + 1 \leq \kappa (\delta )R$ for all $R \geq 1$ , and then $\varrho (\delta , R):= \kappa (\delta ) R$ is an almost repetitivity function for $\Lambda $ . The special case where d is a continuous metric follows along similar lines, using the inclusions given in equation (12).
This now allows us to restate Theorem 1.3 in purely metric terms and thereby establish Theorem 1.2 from §1.
Theorem 4.9. Assume that G is an lcsc group that has exact polynomial growth with respect to an adapted metric d. If a weighted Delone set $\Lambda $ in G is almost linearly repetitive with respect to d, then its hull $\mathcal H_\Lambda $ is minimal and uniquely ergodic.
4.3 Examples of groups with metrics of exact polynomial growth
We have seen in the previous subsection that adapted metrics of exact polynomial growth give rise to strong Følner exhaustion sequences, and allow us to translate the abstract notion of tempered repetitivity into the more concrete notion of linear repetitivity. We now provide some explicit examples of such metrics. Our starting point is the Cygan–Korányi metric on the Heisenberg group.
Example 4.10. Consider the three-dimensional Heisenberg group $G = {\mathbb C} \times {\mathbb R}$ with group law given by
and note that for $z \in {\mathbb C}$ and $t \in {\mathbb R}$ , we have
The Cygan–Korányi norm $\| \cdot \|: G \to [0, \infty )$
defines a group norm on the Heisenberg group, cf. [Reference Capogna, Danielli, Pauls and Tyson15, p. 18, inequality (2.12)], which means that the formula
defines a left-invariant metric $d_G$ on G, called the Cygan–Korányi metric. Since the Cygan–Korányi norm is bounded above and below by suitable powers of the Euclidean norm on ${\mathbb C} \times {\mathbb R} \cong {\mathbb R}^3$ , and the metric $d_G$ is proper and defines the group topology. Thus, $d_G$ is continuous and adapted. Using that the Haar measure $m_G$ on G is the three-dimensional Lebesgue measure, a simple integration via polar coordinates using the substitution $s = r^2$ yields
Since the above integral is equal to the area of the quarter circle with radius $R^2$ , which is $(R^2)^2 \pi / 4$ , we obtain
and hence,
Therefore, the Heisenberg group has exact polynomial growth with respect to the Cygan–Korányi metric $d_G$ .
The previous example can be generalized to cover the whole class of homogeneous Lie groups. As explained in [Reference Fischer and Ruzhansky21, Ch. 3], each homogeneous Lie group G comes equipped with a class of continuous adapted metrics, called homogeneous metrics, which are mutually quasi-isometric, and a natural family of dilation automorphisms $(D_\unicode{x3bb} )_{\unicode{x3bb}>0}$ such that ${B_t = D_t(B_1)}$ for balls with respect to any homogeneous metric d. By [Reference Fischer and Ruzhansky21, p. 100, equation (3.6)], this implies that
where $\kappa $ is a constant depending on G (called the homogeneous dimension of the homogeneous group). We have seen in the case of the Heisenberg group that $\kappa = 4$ . It follows immediately from equation (14) that homogeneous groups have exact polynomial growth with respect to homogeneous metrics, and this property also carries over to lattices in such a group (see [Reference Beckus, Hartnick and Pogorzelski4]).
Proposition 4.11. Let G be a homogeneous Lie group. If d is a homogeneous metric on G, then G has exact polynomial growth with respect to d. Similarly, if $\Gamma $ is a lattice in G, then $\Gamma $ has exact polynomial growth with respect to $d|_{\Gamma \times \Gamma }$ .
At this point, we can easily deduce Theorem 1.1 from §1.
Proof of Theorem 1.1
In view of Theorem 4.9 and the fact that linear repetitivity implies almost linear repetitivity, we only have to show that every homogeneous metric d is adapted and that G has exact polynomial growth with respect to d. The former follows from [Reference Fischer and Ruzhansky21, Proposition 3.1.37] and the latter follows from Proposition 4.11.
Note that Proposition 4.11 generalizes Example 4.10 since the Cygan–Korányi metric is a homogeneous metric on the Heisenberg group. Our next goal is to characterize all groups which admit an adapted metric of exact polynomial growth.
In geometric group theory, a finitely generated group $\Gamma $ is said to have polynomial growth provided for some (hence any) finite generating set S, there exist constants $C\geq 1$ and $k \in \mathbb N$ such that $|S^n| \leq C \cdot n^k$ for all $n \in \mathbb N$ . More generally, if G is a compactly generated lcsc group, then we say that G has polynomial growth if for some (hence any) compact symmetric generating set $\Omega $ and Haar measure $m_G$ on G, there exist constants $C\geq 1$ and $k \in \mathbb N$ such that
We warn the reader that while finitely generated discrete groups of polynomial growth are virtually nilpotent by a celebrated theorem of Gromov, this need not be the case in the compactly generated lcsc groups of polynomial growth. However, such groups are still virtually compact-by-solvable, and hence amenable. A more precise structure theory for compactly generated lcsc groups of polynomial growth was developed only fairly recently by Breuillard [Reference Breuillard14]. One consequence of this structure theory is the following volume growth formula, which is a special case of [Reference Breuillard14, Theorem 1.1].
Theorem 4.12. (Breuillard’s volume growth formula [Reference Breuillard14, Theorem 1.1])
Let G be an lcsc group of polynomial growth. Then, there exists $d(G) \in \mathbb N_0$ and for every compact symmetric generating set $\Omega $ , there exists a constant $c(\Omega )> 0$ such that
Note that polynomial growth is a property of a (compactly generated) lcsc group G, whereas the definition of exact polynomial growth also involves a choice of metric d. Nevertheless the two notions are closely connected.
Theorem 4.13. (Polynomial growth versus exact polynomial growth, cf. [Reference Breuillard14])
For a compactly generated lcsc group G, the following conditions are equivalent:
-
(i) G has polynomial growth;
-
(ii) G admits an adapted metric d of exact polynomial growth.
Proof. (ii) ${\implies}$ (i). Since the balls $B_n$ exhaust G and G is compactly generated, it is generated by some ball $B_N$ ; rescaling the metric by a positive constant if necessary, we may thus assume that G is generated by $B_1$ . Now, let $q>0$ be chosen such that d-balls satisfy the condition $\lim _{t \to \infty } ({m_G(B_t)}/{C t^{q}}) = 1$ and pick an integer $k> q$ . Then, we can find $C'> 0$ such that $m_G(B_t) < C' t^{k}$ for all $t\geq 1$ and hence
which shows that G has polynomial growth.
(i) ${\implies}$ (ii). Let $\Omega $ be a compact generating set of G. It then follows from Theorem 4.12 that the word metric $d_\Omega $ with respect to $\Omega $ has exact polynomial growth, and this metric is obviously adapted.
Note that the above proof of the implication (i) ${\implies}$ (ii) does not produce a continuous adapted metric of exact polynomial growth, unless G is discrete to begin with. The remainder of this subsection is devoted to the question for which groups of polynomial growth can one actually produce a continuous adapted metric of exact polynomial growth and will not be needed in what follows. We start our discussion with a simple example.
If G is a connected Lie group of polynomial growth, then any choice of inner product on the Lie algebra of G will induce a left-invariant Riemannian metric on G, and the corresponding metric d will be continuous and adapted. This metric will have the additional property of being geodesic in the sense that for all $x,x' \in G$ , there exists an isometry $\gamma $ from an interval $I = [a,b] \subseteq \mathbb R$ to a path in G such that $\gamma (a) = x$ and $\gamma (b) = x'$ (by a suitable version of the Hopf–Rinow theorem). For an arbitrary compactly generated lcsc group G, one only knows that there exists a continuous adapted metric d on G, which is large-scale geodesic in the sense of [Reference Cornulier and de la Harpe18, pp. 10], but it is unclear under which conditions this metric can be chosen to be geodesic in the strict sense.
Theorem 4.14. (Exact polynomial growth with respect to a continuous metric)
Assume that a compactly generated lcsc group G of polynomial growth admits a geodesic (rather than just a large-scale geodesic) continuous adapted metric d. Then, G has exact polynomial growth with respect to any such metric. In particular, a connected Lie group admits a continuous adapted metric of exact polynomial growth if and only if it has polynomial growth.
For the remainder of this section, we assume that d is a geodesic continuous adapted metric on a compactly generated lcsc group G of polynomial growth. The fact that d is geodesic has the following consequence concerning balls in G.
Lemma 4.15. The equality $\overline {B}_n = \overline {B}_1^n$ holds for all $n \in \mathbb N$ .
Proof. Let $x \in \overline {B}_n$ with $\alpha := \|x\|$ and $m := \lceil \alpha \rceil - 1 \leq n-1$ . Choose a geodesic $\gamma $ in G with $\gamma (0) = e$ and $\gamma (\alpha ) = x$ , and for $k = 1, \ldots , n$ , define $y_k := \gamma (k-1)^{-1}\gamma (k)$ . Then, by the left-invariance of the metric d and the fact that $d(e, y_k) = d(\gamma (k-1),\gamma (k)) = 1$ , we get
and hence $ \overline {B}_n \subseteq \overline {B}_1^n$ . The converse inclusion is immediate from left-invariance and the triangle inequality.
In view of Breuillard’s volume growth formula, we can prove Theorem 4.14.
Proof of Theorem 4.14
We apply equation (15) to $\Omega := \overline {B}_1$ and use that $m_G(\overline {B}_r)= m_G(B_r)$ for all $r> 0$ . For $q=d(G)$ and $c=c(\Omega )$ as in Theorem 4.12, we obtain with Lemma 4.15 that
Now, for an arbitrary sequence $(t_n)$ with $\lim _{n \to \infty } t_n = \infty $ , we have
The converse inequality follows analogously.
Remark 4.16. (Beyond connected Lie groups)
The simple argument used to establish Theorem 4.14 still carries through if d is merely asymptotically geodesic (see [Reference Nevo40, Theorem 14.3]), but it no longer works if d is only assumed to be large-scale geodesic in the sense of [Reference Cornulier and de la Harpe18]. In this case, we can still find positive integers $a,b,c$ such that for all $x,x' \in G$ , there exist $k \in \mathbb N$ with $n \leq a d(x,x') + b$ and $x_0, \ldots , x_k \in G$ with $x = x_0$ , $x' = x_n$ , and $d(x_{i-1}, x_i) \leq c$ for all $i \in \{1, \ldots , k\}$ , but this yields only the weaker statement
instead of Lemma 4.15. In the notation of Theorem 4.12, this leads to the conclusion that
which, in the language of [Reference Nevo40], means that G has strict polynomial growth with respect to d. Our argument thus shows that every compactly generated lcsc group G of polynomial growth admits a continuous adapted metric d such that G has strict polynomial growth with respect to d. It seems to be an open problem for which classes of compactly generated lcsc group of polynomial growth can one find a metric, which is both continuous and of exact polynomial growth, see the discussion in [Reference Nevo40, §4.4] and the references therein, in particular, [Reference Breuillard14] and [Reference Pansu42].
5 Symbolic systems
In this section, we describe how the previous considerations carry over to symbolic systems over a finite alphabet and we give the proof of Theorem 1.4.
In the following, $\Gamma $ will denote an amenable, countable, discrete group. For the Haar measure on $\Gamma $ , we fix the normalized counting measure $m_{\Gamma }$ which satisfies $m_{\Gamma }(\{e \}) = 1$ for the identity element $e \in \Gamma $ . We will additionally assume that $\Gamma $ is a uniform lattice in an amenable lcsc group H, that is, $\Gamma $ is a discrete subgroup of H whose quotient $H/\Gamma $ is compact in the quotient topology. This point of view is useful for constructing Delone sets in the ambient group H via symbolic considerations over $\Gamma $ . Moreover, we do not lose any generality because we can always consider $\Gamma $ as a lattice in itself. We will also assume that H carries adapted metric $d_H$ and its restriction to $\Gamma $ gives rise to an adapted $d_{\Gamma }$ on $\Gamma $ (which is automatically continuous since $\Gamma $ is discrete).
We further fix a finite set $\mathcal {A}$ , called alphabet, which carries the discrete topology. Then, $\mathcal {A}^{\Gamma }$ becomes a compact metrizable space when endowed with the product topology. We will call the elements in $\mathcal {A}^{\Gamma }$ colorings and the elements in $\mathcal {A}$ are called colors. For any coloring $\mathcal {C} \in \mathcal {A}^{\Gamma }$ , we can define its translation by $\gamma \in \Gamma $ as $(\gamma. \mathcal {C})(x) = \mathcal {C}(\gamma ^{-1}x)$ for all $x \in \Gamma $ . This leads to the notion of the hull of $\mathcal {C}$ , defined as
where the closure is taken in the product topology. Clearly, the hull is a compact metrizable space as well, and $\Gamma $ acts on it via translations. We say that the hull $\Omega _{\mathcal {C}}$ is uniquely ergodic if it admits a unique $\Gamma $ -invariant probability measure. Every $\mathcal {C} \in \mathcal {A}^{\Gamma }$ gives rise to a canonical definition of (colored) patches and patterns. Writing $\mathcal {F}(\Gamma )$ for the collection of all finite subsets of $\Gamma $ , for each $S \in \mathcal {F}(\Gamma )$ , we denote the S-patch of $\mathcal {C}$ as the restriction $\mathcal {C}_{|S}$ of the map $\mathcal {C}$ to the set S. We say that for $S,T \in \mathcal {F}(\Gamma )$ , the corresponding patches of $\mathcal {C}$ are equivalent if there is some $\gamma \in \Gamma $ such that $\gamma S = T$ and $\mathcal {C}(\gamma s) = \mathcal {C}(s)$ for all $s \in S$ . Given an S-patch of $\mathcal {C}$ , we denote its equivalence class as the pattern of that S-patch. We write $[S]_{\mathcal {C}}$ for the collection of all patterns arising from all $\gamma S$ -patches in $\mathcal {C}$ . For $p \in [S]_{\mathcal {C}}$ and $T \in \mathcal {F}(\Gamma )$ , we say that p occurs in T if there are some representative $\mathcal {C}_{|xS}$ of p and $\gamma \in \Gamma $ such that $\gamma xS \subseteq T $ and $\mathcal {C}(\gamma xs) = \mathcal {C}(xs)$ for all $s \in S$ . We point out that $[S]_{\mathcal {C}}$ is finite for some $\mathcal {C} \in \mathcal {A}^{\Gamma }$ since $\mathcal {A}$ is finite, that is, $\mathcal {C}$ is of finite local complexity with respect to the above equivalence relation of patterns.
We will now relate symbolic colorings with weighted Delone sets in the ambient group H. To this end, we fix $\sigma \geq 1$ along with an injective map $\iota : \mathcal {A} \to [\sigma ^{-1}, \sigma ]$ and define the map
where U open and K are chosen such that $\Gamma $ is $(U,K)$ -Delone in H.
Lemma 5.1. The map $\mathcal {I}$ is a continuous, $\Gamma $ -equivariant embedding.
Proof. The injectivity of $\mathcal {I}$ is clear from the injectivity of $\iota $ and uniform discreteness of $\Gamma $ in H. As for the $\Gamma $ -equivariance, take $\gamma _0 \in \Gamma $ , $\mathcal {C} \in \mathcal {A}^{\Gamma }$ , and $B \subseteq H$ Borel. Then,
It remains to show that $\mathcal {I}$ is continuous. To this end, suppose that $(\mathcal {C}_n)\subseteq \mathcal {A}^\Gamma $ is a sequence and $\mathcal {C}$ is an element in $\mathcal {A}^\Gamma $ such that $\lim _{n \to \infty } \mathcal {C}_n = \mathcal {C}$ . Fix a compact neighborhood W of $\{e\}$ in H with $W \cap \Gamma = \{e\}$ . Since all values of $\mathcal {I}$ are measures supported on $\Gamma $ , it suffices to show that $\lim _{n \to \infty } \mathcal {I}(\mathcal {C}_n)(\varphi ) = \mathcal {I}(\mathcal {C})(\varphi )$ for all $\varphi \in C_c(H)$ with the support of $\varphi $ being contained in $KW$ , where $K \subseteq \Gamma $ is finite. Indeed, given a finite set K and fixing such $\varphi $ , we find $n_K \in {\mathbb N}$ such that we have ${\mathcal {C}_n}_{|K} = \mathcal {C}_{|K}$ holds. Therefore, $\mathcal {I}(\mathcal {C}_n)(\varphi ) = \mathcal {I}(\mathcal {C})(\varphi )$ follows for all $n \geq n_K$ . This finishes the proof.
Proposition 5.2. For any $\mathcal {C} \in \mathcal {A}^{\Gamma }$ , the dynamical systems $\Gamma \curvearrowright \Omega _{\mathcal {C}}$ and $\Gamma \curvearrowright \mathcal {H}_{\mathcal {I}(\mathcal {C})}$ are topologically isomorphic.
Proof. It suffices to show that the restriction $\tau $ of $\mathcal {I}$ to $\Omega _{\mathcal {C}}$ gives rise to a $\Gamma $ -equivariant homeomorphism onto $\mathcal {H}_{\mathcal {I}(\mathcal {C})}$ . It follows from Lemma 5.1 (with $H=\Gamma $ ) that $\tau $ is an injective $\Gamma $ -map. Moreover, $\tau $ is continuous as a restriction of the map $\mathcal {I}$ from Lemma 5.1. The restriction of $\tau $ to the orbit $\{\gamma .\mathcal {C}:\, \gamma \in \Gamma \} \subseteq \Omega _{\mathcal {C}}$ is bijective onto its image $\{\gamma .\mathcal {I}(\mathcal {C}):\, \gamma \in \Gamma \} \subseteq \mathcal {H}_{\mathcal {I}(\mathcal {C})}$ . Since $\tau (\Omega _{\mathcal {C}})$ must be compact and hence closed as an image of a compact set under a continuous map with values in a Hausdorff space, we must have $\tau (\Omega _{\mathcal {C}}) = \mathcal {H}_{\mathcal {I}(\mathcal {C})}$ . So $\tau $ is surjective as well. Since $\tau $ is a bijective continuous mapping between two compact Hausdorff spaces, $\tau $ is indeed a homeomorphism.
We transfer the repetitivity concepts from Definitions 2.5 and 3.5 to symbolic systems. Since colorings $\mathcal {C} \in \mathcal {A}^{\Gamma }$ are of finite local complexity with respect to the pattern equivalence relation described above, we can find symbolic analogs of (linear) repetitivity and tempered repetitivity. Recall that H carries an adapted metric $d_H$ that restricts to an adapted metric $d_{\Gamma }$ on $\Gamma $ .
For $r> 0$ , we denote by $B^{H}_r(h)$ the collection of points in H with $d_{H}$ -distance from the identity $h \in H$ being less or equal than r. We analogously define $B^{\Gamma }_r(\gamma )$ for $\gamma \in \Gamma $ . If $h=e$ or $\gamma = e$ , we just write $B^{H}_r$ and $B^{\Gamma }_r$ , respectively.
Definition 5.3. ((Linear) Repetitivity for symbolic systems)
In the situation above, we say that $\mathcal {C} \in \mathcal {A}^{\Gamma }$ is symbolically repetitive with respect to $d_{\Gamma }$ if for every $r \geq 1$ , there is some $R= R(r)> 0$ such that every pattern $p \in [B_r^{\Gamma }]_{\mathcal {C}}$ occurs in $B_{R}(\gamma )$ for all $\gamma \in \Gamma $ . We say that $\mathcal {C}$ is symbolically linearly repetitive if it is symbolically repetitive and there is a $C \geq 1$ such that one can choose $R(r) = Cr$ for all $r\geq 1$ .
We now fix a Følner exhaustion sequence, that is, a Følner sequence $\mathcal {T} = (T_m)$ in $\Gamma $ with $e \in T_m \subsetneq T_{m+1}$ for all $m \in {\mathbb N}$ . Then, $\mathcal {T}$ is also a strong Følner sequence by [Reference Pogorzelski and Schwarzenberger47, Lemma 2.7(d)], and hence a strong Følner exhaustion sequence in $\Gamma $ .
Definition 5.4. (Tempered repetitivity of symbolic systems)
In the situation described above, we define the repetitivity index $\mathcal {R}^{\mathcal {T}}_{\mathcal {C}}: {\mathbb N} \to {\mathbb N} \cup \{+ \infty \}$ for $\mathcal {C}$ with respect to $\mathcal {T}$ as
We say that $\mathcal {C}$ is symbolically tempered repetitive with respect to $\mathcal {T}$ if
Proposition 5.5. For $\mathcal {C} \in \mathcal {A}^{\Gamma }$ , we have the following assertions.
-
(a) If $\mathcal {C}$ is symbolically (linearly) repetitive with respect to $d_{\Gamma }$ , then $\mathcal {I}(\mathcal {C})$ is (linearly) repetitive with respect to $d_{H}$ .
-
(b) Let $H= \Gamma $ and let $\mathcal {T}$ be a Følner exhaustion sequence. Then, $\mathcal {C}$ is symbolically tempered repetitive with respect to $\mathcal {T}$ if and only if $\mathcal {I}(\mathcal {C})$ is tempered repetitive with respect to $\mathcal {T}$ .
Remark 5.6. It is not difficult to come up with a more general statement for assertion (b) covering also cases where $H \neq \Gamma $ . This is however a technical point that we do not need for the purposes of this work.
Proof. We first prove statement (a). To this end, we fix some $r \geq 1$ along with some r-patch $P_{\mathcal {C}} := (\mu ,\, B^H_r(g) )$ of $\mathcal {I}(\mathcal {C})$ , where $g \in H$ and $\mu = \sum _{\gamma \in \Gamma \cap B_r^H(g)} \iota (\mathcal {C}(\gamma )) \delta _{\gamma }$ , where $\iota :\mathcal {A} \to (0,\infty )$ is as defined in Lemma 5.1. Let $h \in H$ . Since $\Gamma \leq H$ is cocompact and $d_H$ is locally bounded, there is an $r_0> 0$ such that $\Gamma B^H_{r_0} = H$ . Hence, there are $\gamma _g,\gamma _{h} \in \Gamma $ such that $g \in B^H_{r_0}(\gamma _g)$ and $h \in B^H_{r_0}(\gamma _h)$ . We write $q_{\mathcal {C}}:= \mathcal {C}_{|B^H_r(g) \cap \Gamma } \in \mathcal {A}^{B^H_r(g) \cap \Gamma }$ for the corresponding patch in $\mathcal {C}$ . Then, $q_{\mathcal {C}}$ occurs in
Note that we also used here the fact that $B^H_r B^H_s \subseteq B^H_{r+s}$ for $r,s \geq 0$ by the left-invariance of the metric $d_H$ . Since $d_{\Gamma }$ is the restriction of $d_H$ and since $\mathcal {C}$ is symbolically (linearly) repetitive, there is some $R= R(r + r_0)> 0$ (which in the case of linear repetitivity can be chosen as $R(r) = Cr$ for some universal $C \geq 1$ ) such that the pattern of $\tilde {q}_{\mathcal {C}}$ occurs in $B^{\Gamma }_{R(r + r_0)}(\gamma _h)$ . Thus, we find by virtue of Lemma 5.1 that the pattern of $P_{\mathcal {C}}$ must occur in $B^H_{R(r + r_0) + r_0}(h)$ . We now set $R^{\prime }(r) = r_0 + R(r + r_0)$ (and in the linearly repetitive situation, we have $C^{\prime }:=r_0+C(1+r_0) \geq C$ such that $r_0 + C(r + r_0) \leq C^{\prime }r$ for all $r \geq 1$ and define $R^{\prime }(r) = C^{\prime } r$ ). In any case, we see that the pattern of $P_{\mathcal {C}}$ occurs in $B^H_{R^{\prime }(r)}(h)$ . This finishes the proof.
The validity of assertion (b) is clear from the definitions and from Lemma 5.1.
We are now in position to prove the following which contains all assertions claimed in Theorem 1.4 as subcases.
Theorem 5.7. Let $\Gamma $ be a countable amenable group that is co-compactly embedded as a lattice in an lcsc group H carrying an adapted metric $d_H$ . Let $\mathcal {A}$ be a finite set and assume that $\mathcal {C} \in \mathcal {A}^{\Gamma }$ . Then the following assertions hold.
-
(a) If $\mathcal {C}$ is symbolically tempered repetitive with respect to some Følner exhaustion sequence, then the hull $\Omega _{\mathcal {C}}$ is minimal and uniquely ergodic.
-
(b) If $\mathcal {C}$ is symbolically linearly repetitive with respect to the metric $d_{\Gamma }$ induced by $d_H$ and if H has exact polynomial growth with respect to $d_H$ , then the hull
$$ \begin{align*} \mathcal{H}_{\mathcal{I}(\mathcal{C})} = \overline{\{ h.\mathcal{I}(\mathcal{C}):\, h \in H \}}^{w*} \end{align*} $$is minimal and uniquely ergodic, where $\mathcal {I}(\mathcal {C})$ is defined according to the map of Lemma 5.1. -
(c) If $\mathcal {C}$ is symbolically linearly repetitive with respect to some adapted metric $d_{\Gamma }$ on $\Gamma $ , and if $\Gamma $ has exact polynomial growth with respect to $d_{\Gamma }$ , then the hull ${\Omega }_{\mathcal {C}}$ is minimal and uniquely ergodic.
Proof. We first prove the assertion (a). By Proposition 5.5(b), we find that $\mathcal {I}(\mathcal {C})$ is tempered repetitive as a weighted Delone set over $\Gamma $ with respect to some Følner exhaustion sequence. It follows from Theorem 1.3 that the dynamical system $\Gamma \curvearrowright \mathcal {H}_{\mathcal {I}(\mathcal {C})}$ is minimal and uniquely ergodic. By Proposition 5.2, these features also carry over to the system $\Gamma \curvearrowright \Omega _{\mathcal {C}}$ .
To prove the assertion (b), we note first that $\mathcal {I}(\mathcal {C})$ is linearly repetitive with respect to $d_H$ as a weighted Delone set in H by Proposition 5.5(a). Since H has exact polynomial growth with respect to $d_H$ , we infer from Theorem 4.9 that the system $H \curvearrowright \mathcal {H}_{\mathcal {I}(\mathcal {C})}$ is minimal and uniquely ergodic.
Assertion (c) follows from assertion (b) in the case $H = \Gamma $ , while observing with Proposition 5.2 that the dynamical systems $\Gamma \curvearrowright \Omega _{\mathcal {C}}$ and $\Gamma \curvearrowright \mathcal {H}_{\mathcal {I}(\mathcal {C})}$ are topologically isomorphic.
6 Two applications
This section is devoted to the proofs of Corollaries 1.5 and 1.6.
If $(T_m)$ is a (left) Følner sequence, then $\lim _{m} ({m_G(\partial _K(T_m))}/{m_G(T_m)}) =0$ for every compact $K \subseteq G$ . To prove Corollary 1.5 and Corollary 1.6, we need that for a suitable boundary notion, a similar fraction goes to zero for $(T_m^{-1})$ which will be a (right) Følner sequence. In light of this, define the (right)-K-boundary of S by
The following is a simple modification of [Reference Pogorzelski44, Lemma 2.3] for $\partial _K$ replaced by $\tilde {\partial }_K$ .
Lemma 6.1. Let $K,L,S,T\subseteq G$ . Then, the following assertions hold true:
-
(a) $\tilde {\partial }_K(S\cup T) \subseteq \tilde {\partial }_K(T) \cup \tilde {\partial }_K(S)$ ;
-
(b) $\tilde {\partial }_K(S\setminus T) \subseteq \tilde {\partial }_K(T) \cup \tilde {\partial }_K(S)$ ;
-
(c) $\tilde {\partial }_K(T)\subseteq \tilde {\partial }_L(T)$ if $K\subseteq L$ ;
-
(d) $\tilde {\partial }_K(\tilde {\partial }_L(T)) \subseteq \tilde {\partial }_{KL}(T)$ ;
-
(e) if $e\in K$ , then $xK\subseteq \tilde {\partial }_{K^{-1}}(T)\cup T$ holds for each $x\in T$ .
Furthermore, if $(T_m)$ is a strong Følner exhaustion sequence, then $(T_m^{-1})$ is an exhaustion sequence and for every compact $K \subseteq G$ ,
Proof. Assertions (a)–(c) are immediate from the definition of $\tilde {\partial }_K(T)$ . As for assertion (d), we note that
The claim now follows from
Assertion (e) is proven as follows. Let $x\in T$ and $g\in xK$ . If $g\in T$ , there is nothing to prove. Otherwise, $g\in G\setminus T$ and so $g\in (G\setminus T)K$ as $e\in K$ . Since $g\in TK$ , we derive $g\in \partial _{K^{-1}}(T)$ .
Suppose $(T_m)$ is a strong Følner exhaustion sequence. Then, clearly, $(T_m^{-1})$ is an exhaustion sequence. Since G is unimodular, $m_G(S)=m_G(S^{-1})$ holds for all $S\in \mathcal {RK}(G)$ . Thus,
holds for each compact $K \subseteq G$ and $m \in {\mathbb N}$ . Hence, $\lim _{m\to \infty } ({m_G(\tilde {\partial }_K(T_m^{-1}))}/{m_G(T_m^{-1})}) = 0$ follows as $(T_m)$ is a strong Følner sequence and $m_G(T_m)=m_G(T_m^{-1})$ .
6.1 Equality of upper and lower Banach density
The upper and lower Banach densities for a weighted Delone set $\Pi $ along a Følner sequence $(T_m)$ are defined by
The above limit inferior and limit superior are actually limits by Lemma 3.17 and Remark 3.18 as the almost sub-additive weight function $w(S,\Pi ):=\delta _{\Pi }(S^{-1})$ is actually additive, see below. Invoking Theorem 3.1, we show in Corollary 1.5 that the upper Banach density and the lower Banach density coincide if $\Pi $ is an almost tempered repetitive weighted Delone set.
Proof of Corollary 1.5
Let $\Lambda $ be a weighted Delone set in G that is almost tempered repetitive with respect to a strong Følner exhaustion sequence $(T_m)$ . We seek to prove that
exists uniformly in $\Pi \in \mathcal {H}_\Lambda $ . Due to Theorem 3.1, it suffices to show that $w:\mathcal {RK}(G)\times \mathcal {H}_\Lambda \to \mathbb {R},\, w(K,\Pi ):= \delta _{\Pi }(K^{-1})$ is an admissible (almost sub-)additive weight function. Recall that $\Pi $ is a Radon measure on G. Thus, w satisfies conditions (W1) and (W3) with $B = \emptyset $ and $\theta =0$ . Recall that $h_*\delta _{\Pi }(S):=\delta _{\Pi }(h^{-1}S)$ holds for any $h\in G$ and $S\in \mathcal {RK}(G)$ . Let $K\in \mathcal {RK}(G)$ , $h\in G$ , and $\Pi \in \mathcal {H}_\Lambda $ , then
holds implying condition (W4) with $I = \emptyset $ and $\vartheta =0$ .
Recall that for each $\Pi =(Q,\beta )\in \mathcal {H}_\Lambda $ , Q is a $(U,K)$ -Delone set where $e\in U$ holds without loss of generality. Let Q be an arbitrary left-U-uniformly discrete subset of G. It is straightforward to show that there exists a $V\subseteq G$ open (only depending on U) satisfying $e\in V$ , $\overline {V}^{-1}=\overline {V}$ , and $x\overline {V}\cap y\overline {V}= \emptyset $ for all $x,y\in Q$ and every left-U-uniformly discrete set Q. By assertion (e) of Lemma 6.1, one obtains that for each $S \in \mathcal {RK}(G)$ and $x\in S$ , we have $x\overline {V}\subseteq \tilde {\partial }_{\overline {V}}(S)\cup S$ . Thus,
follows.
Let $L,K\in \mathcal {RK}(G)$ with $K\subseteq L$ and $\Pi =(Q,\beta )\in \mathcal {H}_\Lambda $ . Since $\delta _{\Pi }(S)= \sum _{x\in Q} \beta (x)\delta _x(S)$ , we obtain
where we used $\beta (x)\leq \sigma $ and $\tilde {\partial }_{\overline {V}}(L^{-1}\setminus K^{-1}) \subseteq \tilde {\partial }_{\overline {V}}(L^{-1}) \cup \tilde {\partial }_{\overline {V}}(K^{-1})$ . By the symmetry of $\overline {V}$ and the unimodularity of G, we observe that
and likewise for K. Hence, $w$ satisfies axiom (W2) with $\eta := {\sigma }/{m_G(\overline {V})}$ and $J:=\overline {V}$ . It remains to prove condition (W5).
Let $\varepsilon>0$ and suppose that V is chosen as before. Fix $0 < \delta < 1$ such that $B_{2\delta }\subseteq V$ and $\delta <\min \{({\sigma ^{-1}}/{2}), ({m_G(\overline {V})}/{2 + 4\sigma })\varepsilon \}$ . Due to Lemma 6.1, there is an $m_0\in \mathbb {N}$ such that $m_G(\tilde {\partial }_{\overline {V}^3}(T_m^{-1}))\leq \delta m_G(T_m^{-1})$ holds for $m\geq m_0$ . We use properties (a)–(d) subsequently. Let $m\geq m_0$ and $\Pi =(Q,\beta ), \Phi =(P,\alpha )\in \mathcal {H}_\Lambda $ be such that $d_{T_m^{-1}}(\Pi ,\Phi )<\delta $ where as defined in the equalities (1) and (2), we have
Set $S_1:=T_m^{-1}\setminus \tilde {\partial }_{\overline {V}}(T_m^{-1})$ . Thus, for each $x\in S_1 \cap P$ , $x\overline {V} \subseteq T_m^{-1}$ follows. Let ${x\in S_1 \cap P}$ . Since Q is U-uniformly discrete, $B_{\delta }\subseteq V$ and
we conclude that there is a unique $y_x\in T_m^{-1} \cap Q$ such that $y_x\in B_{\delta }(x)$ and $|\alpha (x)-\beta (y_x)|<\delta $ . Let $S_2:=T_m^{-1}\setminus \tilde {\partial }_{\overline {V}^2}(T_m^{-1})$ . Then, $y\in S_2 \cap Q$ implies $y\overline {V}\subseteq S_1$ . Thus, $|\delta _{\Pi }(B_\delta (y))-\delta _{\Phi }(B_\delta (y))|<\delta <\sigma ^{-1}$ implies as before that there is a unique $x_y\in P\cap S_1$ satisfying $x_y\in B_\delta (y)$ and $|\alpha (x_y)-\beta (y)|<\delta $ . Consequently, ${S_2 \kern1.7pt{\cap}\kern1.7pt Q\kern1.7pt{\subseteq}\kern1.7pt \{y_x\,{:}\, x\kern1.7pt{\in}\kern1.7pt S_1 \kern1.7pt{\cap}\kern1.7pt P\}}$ . Hence,
follows. With this at hand, we derive
By Lemma 6.1(c) and the choice $m_0$ , we have $m_G(\tilde {\partial }_{\overline {V}}(T_m^{-1}))\leq \delta m_G(T_m^{-1})$ . Since $\sharp (S\cap D )\leq {1}/{m_G(\overline {V})} (m_G(\tilde {\partial }_{\overline {V}}(S)) + m_G(S))$ for any U-uniformly set D and any $S\in \mathcal {RK}(G)$ , the previous considerations lead to
Since $\tilde {\partial }_I(\tilde {\partial }_J(S))\subseteq \tilde {\partial }_{IJ}(S)$ holds by Lemma 6.1(d), the estimate
is derived using $m_G(T_m^{-1})=m_G(T_m)$ which holds as G is unimodular. Thus, w satisfies condition (W5).
6.2 Uniform convergence of the IDS
We briefly sketch the model outlined in [Reference Lenz, Schwarzenberger and Veselić35] and in [Reference Pogorzelski and Schwarzenberger47]. To be consistent with the previous sections, we will consider Cayley graphs with a left-invariant metric as well as pattern equivalence defined by left translations. (In the mentioned papers, the authors consider translations from the right but also deal with left Følner sequences.)
We suppose that $\Gamma $ is an amenable countable group generated by a finite symmetric set $S \subseteq \Gamma $ . For the Haar measure on $\Gamma $ , we fix the normalized counting measure $m_{\Gamma }$ which satisfies $m_{\Gamma }(\{e \}) = 1$ for the identity element $e \in \Gamma $ .
Recall for finitely generated groups, Følner sequences $(F_n)$ consist of non-empty finite sets such that for all $s \in S$ , one has
This definition is equivalent to the notion of strong Følner sequences in the above sense by [Reference Pogorzelski and Schwarzenberger47, Lemma 2.7(d)]. The generating set S comes with a natural left-invariant word metric. Precisely, for $g,h \in \Gamma $ , one defines
In the following, we write
for $g \in \Gamma $ and $n \in {\mathbb N}$ .
Each group $\Gamma $ generated by a symmetric set S comes with a Cayley graph with vertices given by $\Gamma $ and two elements $g,h \in \Gamma $ are connected by an edge if there exists some $s \in S$ such that $gs = h$ . We write $\operatorname {Cay}(\Gamma , S)$ for this graph.
We consider a coloring $\mathcal {C}\in \mathcal {A}^\Gamma $ of the group $\Gamma $ which is nothing but a map ${\mathcal {C}: \Gamma \to \mathcal {A}}$ , where $\mathcal {A}$ is a finite set whose elements are called colors. As demonstrated in §5, by assigning to each color in $a \in \mathcal {A}$ a different positive number $\iota (a)> 0$ , we can identify $\mathcal {C}$ with the weighted Delone set $I(\mathcal {C}):= (\Gamma , \delta _{\mathcal {C}})$ , where $\delta _{\mathcal {C}} := \sum _{x \in \Gamma } \iota (\mathcal {C}(x)) \delta _{\{ x\}} $ on $\Gamma $ . Invoking Lemma 5.1 and Proposition 5.2, we consider a coloring $\mathcal {C}$ both as mappings $\Gamma \to \mathcal {A}$ and as weighted Delone sets in the above form. Given $\mathcal {C}$ , we define its hull by
where the closure is taken with respect to the product topology in $\mathcal {A}^\Gamma $ induced from the discrete topology on $\mathcal {A}$ . We now consider a coloring $\mathcal {C}$ that is symbolically tempered repetitive with respect to a strong Følner exhaustion sequence as of Definition 5.4. It follows from Theorem 5.7(a) that $\Gamma \curvearrowright \Omega _{\mathcal {C}}$ is uniquely ergodic. Given a finite set ${E \in \mathcal {F}(\Gamma )}$ , we define the E-patch for $\mathcal {C}$ as in §5 above.
Since characteristic functions over the cylinder sets associated with patches are continuous, we derive from Theorem 3.1 (in combination with Proposition 5.2) that for every colored patch $\mathcal {E} := \mathcal {C}_{|E}$ with $E \subseteq \Gamma $ finite, the limits
exist for all $\mathcal {D} \in \Omega _{\mathcal {C}}$ , where for $\mathcal {D}^{\prime } \in \Omega _{\mathcal {C}}$ , one has $1_{\mathcal {E}}(\mathcal {D}^{\prime }) = 1$ if and only if $ \mathcal {D}_{|E}^{\prime }= \mathcal {E}$ and $1_{\mathcal {E}}(\mathcal {D}^{\prime }) = 0$ otherwise. This yields
where we write $\mathcal {C}_{|x^{-1}E} \cong \mathcal {E}$ if the patches $\mathcal {C}_{|x^{-1}E}$ and $\mathcal {E}$ are equivalent. With no loss of generality, we will assume that the identity e is contained in E since, otherwise, we can pick $a \in E$ and repeat the above argument with the coloring $\mathcal {D} := a.\mathcal {C}$ and the patch $\mathcal {C}_{|a^{-1}E}$ instead of $\mathcal {C}$ and $\mathcal {E}$ . Furthermore, the sequence $(T_m^{-1})$ is a right-Følner sequence, see Lemma 6.1. Thus,
Consequently, we arrive at
for every colored patch $\mathcal {E} := \mathcal {C}_{|E} $ with $E \subseteq \Gamma $ finite. In the setting of [Reference Pogorzelski and Schwarzenberger47, §5] with right translations of patches replaced by left translations of patches, the latter condition is equivalent to
for all patches $\mathcal {E} := \mathcal {C}_{|E} $ with $E \subseteq \Gamma $ finite, where $\#^l_{\mathcal {E}}(T_m^{-1})$ counts the number of occurrences of the pattern of $\mathcal {E}$ in $T_m^{-1}$ . Recall that $(T^{-1}_m)$ is a right Følner sequence. The setting in [Reference Pogorzelski and Schwarzenberger47] is reversed: there, one deals with right occurrences of colored patterns, with a right-invariant word metric and with left Følner sequences.
We turn to operators on the given left Cayley graph.
For a subset $W \subseteq \Gamma $ , we write $\ell ^2(W, \mathcal {S})$ for the space of functions $u: W \to \mathcal {S}$ satisfying $\sum _{W \in \Gamma } \|u(g)\|^2_{\mathcal {S}} < \infty $ , where $\mathcal {S}$ is a finite dimensional Hilbert space with norm $\| \cdot \|_{\mathcal {S}}$ . For a finite set $L \subseteq \Gamma $ and an operator $H: \ell ^2(\Gamma , \mathcal {S}) \to \ell ^2(\Gamma , \mathcal {S})$ , we set
where $i_L: \ell ^2(L, \mathcal {S}) \to \ell ^2(\Gamma , \mathcal {S})$ and $p_L: \ell ^2(\Gamma , \mathcal {S}) \to \ell ^2(L, \mathcal {S})$ are the canonical inclusion and projection operators defined as
For $g,h \in \Gamma $ , we write $H(g,h) := p_{\{g\}}H i_{\{h\}}$ .
On $\ell ^2(\Gamma , \mathcal {S})$ , we now define the class of operators which is of interest to us.
Definition 6.2. Let $H:\ell ^2(\Gamma , \mathcal {S}) \to \ell ^2(\Gamma , \mathcal {S})$ be a self-adjoint operator with $\Gamma , S$ and $\mathcal {S}$ as above. Further, suppose that $\mathcal {C}: \Gamma \to \mathcal {A}$ is a coloring. We say that H has finite hopping range if there is some $M \in {\mathbb N}$ such that $H(g,h) = 0$ whenever $d_S(g,h) \geq M$ . We say that H is $\mathcal {C}$ -invariant if there is some $N \in {\mathbb N}$ such that $H(xg, xh) = H(g,h)$ for $x,g,h \in \Gamma $ that satisfy
If H has finite hopping range and is $\mathcal {C}$ -invariant, then we call $R:= \max \{M;N\}$ the overall range of H.
We now define the eigenvalue counting function for a self-adjoint operator ${A:\mathcal {S} \to \mathcal {S}}$ defined on finite dimensional Hilbert space $\mathcal {S}$ . Specifically, the cumulative eigenvalue counting function of A is defined by
where the sum is taken over the set of all eigenvalues of A less or equal than E and $\operatorname {mult}(\unicode{x3bb} )$ is the multiplicity of the eigenvalue $\unicode{x3bb} $ .
In the following, we fix once and for all a self-adjoint operator $H:\ell ^2(\Gamma , \mathcal {S}) \to \ell ^2(\Gamma , \mathcal {S})$ that has finite hopping range and is $\mathcal {C}$ -invariant with respect to a coloring $\mathcal {C}: \Gamma \to \mathcal {A}$ with overall range $R \in {\mathbb N}$ .
Given a finite non-empty set $F \subseteq \Gamma $ and $E \in {\mathbb R}$ , we define
where $F^R := \{x \in F:\, B^S_{R+1}(x) \subseteq F\}$ . Then, the functions
are elements in the Banach space of bounded right continuous functions on ${\mathbb R}$ with supremum norm, denoted by $(\operatorname {BRC}({\mathbb R}),\, \| \cdot \|_{\infty } )$ . They denote the empirical eigenvalue distribution of the operators $H_F$ for non-empty finite sets $F \subseteq \Gamma $ . Note that dividing by $\operatorname {dim}(\mathcal {S})$ gives the correct normalization to obtain numbers in $[0,1]$ since
for finite $F \subseteq \Gamma $ .
Proof of Corollary 1.6
For the proof of Corollary 1.6, suppose that the coloring ${\mathcal {C}:\Gamma \to \mathcal {A}}$ is symbolically tempered repetitive with respect to $(T_m)$ . Let $H:\ell ^2(\Gamma , \mathcal {S}) \to \ell ^2(\Gamma , \mathcal {S}) $ be a self-adjoint, $\mathcal {C}$ -invariant finite hopping range operator with overall range $R \in {\mathbb N}$ . Due to the existence of the frequencies as displayed in equation (16) and taking into account the reversed model as in [Reference Pogorzelski and Schwarzenberger47, §§5 and 7], we obtain from in [Reference Pogorzelski and Schwarzenberger47, Theorem 7.4] that there is some $N_H \in \operatorname {BRC}({\mathbb R})$ such that
where $H_m = H[(T_m^{-1})^R]$ and the $N_{H_m}$ are the empirical eigenvalue distributions as of equation (17).
Remark 6.3. Also shown in [Reference Pogorzelski and Schwarzenberger47, Theorem~7.4] is that $N_H$ is the distribution function of a probability measure $\mu _H$ on ${\mathbb R}$ .
Acknowledgments
We thank Daniel Lenz for an inspiring discussion on linear repetitive Delone sets and related sub-additive ergodic theorems. Moreover, we are thankful to Christoph Richard for conversations on almost linear repetitivity. We are grateful to Michael Björklund and Amos Nevo for discussions on Følner conditions for balls in groups of polynomial growth. We thank Hung Ninh Ngoc for carrying out explicit volume computations in the Heisenberg group with respect to the Cygan–Korányi metric. This work was partially supported by the Deutsche Forschungsgemeinschaft (BE 6789/1-1 to S.B.)
A Appendix. Weight functions and convergence
In this appendix, we give the proofs of Lemmas 3.16 and 3.17.
Proof of Lemma 3.16
Fix $0 < \varepsilon < 1/10$ , a compact $I \subseteq G$ , as well as $\zeta = \varepsilon / 16$ . Fix a strong Følner exhaustion sequence $(S_l)$ . We first choose $m_I \in {\mathbb N}$ large enough such that $m_G(\partial _I(S_k)) \leq \varepsilon m_G(S_k)$ for all $k \geq m_I$ . Fix $n \in {\mathbb N}$ with $n \geq m_I$ . Using Theorem 3.14, we find prototiles $S_i^{\varepsilon } \in \{ S_l:\, l \geq \max \{i, n \} \}$ for $1 \leq i \leq N(\varepsilon )$ and $\delta _0> 0$ such that every $( S_{N(\varepsilon )}^{\varepsilon } S_{N(\varepsilon )}^{\varepsilon \, -1},\, \delta _0 )$ -invariant compact subset $A \subseteq G$ can be $\varepsilon $ -quasi tiled while satisfying the properties (T1)–(T4) given in the statement of Theorem 3.14. So, in the following, we assume that A is $( S_{N(\varepsilon )}^{\varepsilon } S_{N(\varepsilon )}^{\varepsilon \, -1},\, \delta _0 )$ -invariant and also $(L, \varepsilon )$ -invariant, where $L \in \{J,B,I\}$ and we fix finite center sets $C_i^A \subseteq A$ giving rise to an $\varepsilon $ -quasi tiling. By Remark 3.15(ii), we can also make sure that
for $L \in \{J,B\}$ and for all $1 \leq i \leq N(\varepsilon )$ and $c \in C_i^A$ .
We define
as well as $A_{\varepsilon } := A \setminus \bigcup _{i=1}^{N(\varepsilon )} \bigcup _{c \in C_i^A} \tilde {S}_i^{\varepsilon }(c)$ . We obtain
Furthermore, we use the sub-additivity property (w3) of v, combined with the general inclusions
for general sets $C,D,E \in \mathcal {RK}(G)$ and the triangle inequality to obtain
Next, we use the almost monotonicity condition (w2) to find that
In view of conditions (w1) and (w2) (and the above set inclusions for boundaries), we also have
Plugging this into the above inequality and using $m_G(S_i^{\varepsilon }c \setminus \tilde {S}_i^{\varepsilon }(c)) < \varepsilon $ along with the invariance conditions on the prototiles, cf. equality (A.1) and $\zeta ^2 \leq 4\zeta $ , we arrive at
By the tiling property (T4) of Theorem 3.14 and using the fact that $\sum _i \sum _c m_G(S_i^{\varepsilon }c) \leq 2m_G(A)$ , we get
We finally use the additional invariance assumptions on A, that is, $m_G(\partial _J(A)) \leq \varepsilon m_G(A)$ and $m_G(\partial _B(A)) \leq \varepsilon m_G(A)$ . Hence, the choice $\zeta = \varepsilon / 16$ finally leads to
For the proof of Lemma 3.17, we need another lemma first.
Lemma A.1. Let $w:\mathcal {RK}(G) \times X \to {\mathbb R}$ be an almost sub-additive weight function satisfying (W5 $^\ast $ ) and assume that $G \curvearrowright X$ is minimal. Then, for every strong Følner exhaustion sequence $(T_m)$ and for each $\varepsilon> 0$ , there is some $m_0 \in {\mathbb N}$ such that
for all $x,y \in X$ and for all $m \geq m_0$ .
Proof. Let $\varepsilon> 0$ . We now choose $m_0 \in {\mathbb N}$ such that property (W5 $^\ast $ ) of Remark 3.12 holds applied with $\varepsilon / 3$ instead of $\varepsilon $ . If necessary, we increase $m_0$ such that $m_G(\partial _I(T_m))/m_G(T_m) \leq \varepsilon /(6 \vartheta )$ holds for all $m \geq m_0$ , where $I \subseteq G$ is the compact subset of property (W4) that only depends on w. Let $m \geq m_0$ and fix $x,y \in X$ . By definition of $w^{+}$ and property (W4) of Definition 3.9, we find and element $h^m_x \in G$ such that
Moreover, since the action $G \curvearrowright X$ is minimal, we find $j_m \in G$ such that $d_{X}(h^m_x x, j_m y ) < \delta _m$ (with $\delta _m> 0$ chosen as in property (W5 $^\ast $ )), and by property (W5 $^\ast $ ), we arrive at
Putting everything together and using again property (W4), we observe
By definition of $w^+$ and $m_G(\partial _I(T_m))/m_G(T_m) < \varepsilon (6\vartheta )^{-1}$ , we conclude $w^{+}(T_m, x) \leq w^{+}(T_m, y) + \varepsilon $ for all $m \geq m_0$ . Interchanging the roles of x and y yields the statement for $w^{+}$ . The assertion for $w^{-}$ can be proven in the very same manner and we leave the details to the reader.
We are ready to give the proof of Lemma 3.17.
Proof of Lemma 3.17
It suffices to show
for all $x\in X$ . Invoking Remark 3.10, we see that both the above limsup and the liminf are contained in the interval $[-\eta , \eta ]$ for all $x \in X$ , where $\eta \geq 0$ is the constant as of axiom (W2) in Definition 3.9. So fix $x \in X$ . We find a subsequence $(S_l)$ of $(T_m)$ such that
By property (W4), there is a compact $I \subseteq G$ and $\vartheta \geq 0$ such that the almost-equivariance property for w is satisfied. Let $0<\varepsilon <1/10$ and $N(\varepsilon ):=\lceil -\log ({\varepsilon }/{1-\varepsilon })\rceil $ . We apply Lemma 3.16 with $B=B$ , $I=I$ , $J=J$ , $\eta (v) = \eta $ , and $\theta (v) = \theta $ , where J and $\eta $ describe the almost monotonicity property (W2), and B and $\theta $ describe the almost sub-additivity property (W3) of w. We choose $\ell \in {\mathbb N}$ large enough such that $m_G(\partial _I(S_l)) \leq \varepsilon m_G(S_l)$ for all $l \geq \ell $ . Hence, we find a collection of prototiles
taken from the sequence $(S_l)$ and there is some $M \in {\mathbb N}$ such that for each $m \geq M$ , the set $T_m$ can be $\varepsilon $ -quasi tiled by the prototiles $S_i^{\varepsilon }$ with center sets $C_i^m$ for $1 \leq i \leq N(\varepsilon )$ and
for all $y \in X$ . Further, combining the previous estimate and property (W4) of a weight function gives
for all $h \in G$ and $m \geq M$ . Increasing M if necessary and using $m_G(\partial _I(S_l)) \leq \varepsilon m_G(S_l)$ for $l\geq \ell $ together with $\sum _i \sum _c m_G(S_i^{\varepsilon }c) \leq 2m_G(T_m)$ , we arrive at
for all $h \in G$ and $m \geq M$ .
Let $S^\varepsilon :=S_{i_0}^\varepsilon $ be chosen such that $w^+(S^\varepsilon ,x)=\max _{1\leq i\leq N(\varepsilon )}w^+(S_i^\varepsilon ,x)$ . Then,
for all $1 \leq i \leq N(\varepsilon )$ . In addition, the tiling properties (T1)–(T4) listed in Theorem 3.14 lead to
Combined with the previous estimate, this yields
for all $m \geq M$ and $h \in G$ . (Note that the sign of $w^{+}(S^{\varepsilon }, x)$ determines the factor in front of it.) Taking the supremum over $h\in G$ and the limsup in m, we derive
We recall that by definition, we have that $S^{\varepsilon } = S_l$ for some $l \geq \ell $ . Since $\ell $ was chosen arbitrarily but large enough and since $(S_l)$ is a subsequence of $(T_m)$ , where $(w^{+}(T_m, x))_m$ attains its limit inferior, we can send $\ell $ to infinity to see
Sending $\varepsilon \to 0$ gives what was claimed above.
We show next that the limits do not depend on the choice of the Følner sequence. To this end, fix two nested Følner sequences $(T_m)$ and $(T^{\prime }_l)$ . Then, for each $\varepsilon < 1/10$ , each $\ell \in {\mathbb N}$ , and large m, the set $T_m$ can be $\varepsilon $ -quasi tiled by the same means as above by prototiles taken from $(T^{\prime }_l)_{l \geq \ell }$ . Repeating exactly the same arguments as before leads to inequality (A.2) with $(S_l)$ replaced by $T^{\prime }_{l}$ . Sending $\varepsilon \to 0$ shows that the limit with respect to one sequence is less than or equal to the limit with respect to the other sequence. By symmetry, the independence follows.
If $w$ is even a topological weight function, $(T_m)$ is a strong Følner exhaustion sequence, and if the action $G \curvearrowright X$ is additionally minimal, then the limits $\lim _{m \to \infty } {w^{+}}(T_m, x)$ must coincide for all $x \in X$ and the convergence must be uniform by Lemma A.1.
B Appendix. Topology of weighted Delone sets
This section is devoted to a particular useful neighborhood basis for the weak- $*$ -topology on spaces of weighted Delone sets.
We fix an lcsc group G, an open subset $U \subseteq G$ , compact subset $K \subseteq G$ , and $\sigma \geq 1$ . We then consider weighted Delone sets in $\operatorname {Del}(U,K,\sigma )$ , that is, subsets of G which are left-U-uniformly discrete and left-K-syndetic with weights in the interval $[\sigma ^{-1}, \sigma ]$ .
Let $S \in \mathcal {RK}(G)$ . Recall the notion
for two weighted Delone sets $\Lambda = (P,\alpha )$ and $\Pi = (Q,\beta )$ .
Proposition B.1. Let $\Lambda $ be a weighted Delone set in G. Then for all $S \in \mathcal {RK}(G)$ and $\delta> 0$ , the set
is a weak- $*$ -neighborhood of $\Lambda $ . Moreover, $\{ \mathcal {U}_{S,\delta }(\Lambda ): S \in \mathcal {RK}(G),\,\delta> 0 \}$ defines a weak- $*$ -neighborhood basis of $\Lambda $ .
Proof. Let $0 < \delta < \sigma ^{-1}/2$ , $S\in \mathcal {RK}(G)$ , and $\Lambda =(P,\alpha )\in \operatorname {Del}(U,K,\sigma )$ . Define $S':=S.B_1\in \mathcal {RK}(G)$ . Since $\sigma \geq 1$ , $B_{2\delta }(x)\subseteq S'$ holds for all $x\in S$ . There is no loss in generality in assuming that $e\in U$ and $\delta>0$ is small enough such that $B_{2\delta }(x) \cap B_{2\delta }(y)=\emptyset $ for any $x,y\in D$ , where D is some U-uniformly discrete set.
Set $K:=S'\setminus \bigcup _{x\in P\cap S'} B_{\delta }(x)$ . By Urysohn’s lemma, there is a $\phi \in {\mathcal C}_c(G)$ such that $0\leq \phi \leq 1$ , $\phi (x)=0$ for all $x\in P$ and $\phi |_K = 1$ . Furthermore, for each $x\in P\cap S'$ , there are $\phi _x\in {\mathcal C}_c(G)$ such that $0\leq \phi _x\leq 1$ , $\phi _x$ is supported in $B_{2\delta }(x)$ and $\phi _x|_{B_\delta (x)}\equiv 1$ . Then, the set
is a weak- $*$ -neighborhood of $\Lambda $ in $\operatorname {Del}(U,K,\sigma )$ . We will show first that $\mathcal {V}\subseteq \mathcal {U}_{S,\delta }(\Lambda )$ showing that $\mathcal {U}_{S,\delta }(\Lambda )$ is indeed a weak- $\ast $ -neighborhood of $\Lambda $ .
Let $\Pi =(Q,\beta )\in \mathcal {V}$ . By definition of $\phi $ , we have $\delta _{\Lambda }(\phi )=0$ . Thus, $|\delta _{\Pi }(\phi )| = |\delta _{\Lambda }(\phi )-\delta _{\Pi }(\phi )|<\delta $ implies $ Q\cap S' \subseteq \bigcup _{x\in P\cap S'} B_\delta (x) $ as $\delta <\sigma ^{-1}$ and $\phi _{|K} =1 $ .
By the choice of $\phi _x$ , we have $\delta _{\Lambda }(\phi _x)=\alpha (x)$ for all $x \in P \cap S^{\prime }$ . In particular, for each $x\in P\cap S$ , $|\delta _{\Lambda }(\phi _x)-\delta _{\Pi }(\phi _x)|<\delta $ and the condition $B_{2\delta }(x) \cap B_{2\delta }(y) = \emptyset $ for $x,y \in Q$ with $x \neq y$ implies that there is a unique $z_x\in Q\cap B_{2\delta }(x) \subseteq Q\cap S'$ . Since $Q\cap S' \subseteq \bigcup _{x\in P\cap S'} B_\delta (x)$ , we conclude $z_x\in Q\cap B_{\delta }(x)$ . Furthermore, $\phi _x|_{B_\delta (x)}\equiv 1$ yields $\phi _x(x) = 1 = \phi _x(z_x)$ . Thus,
holds for each $x\in P\cap S$ . If $y\in Q\cap S$ , there is a unique $x\in P\cap S'$ such that $y \in B_{\delta }(x)$ . Thus, $\delta _{\Lambda }(B_{\delta }(x)) = \delta _{\Lambda }(B_{\delta }(y))$ and $\delta _{\Pi }(B_{\delta }(x)) = \delta _{\Pi }(B_{\delta }(y))$ for these choices of x and y. Hence,
for all $y \in Q \cap S$ and the unique $x \in P \cap S^{\prime }$ with $y \in B_{\delta }(x)$ . Consequently, we have proven that $\Pi \in \mathcal {V}$ implies $d_S(\Lambda ,\Pi )< \delta $ . Thus, $\mathcal {U}_{S,\delta }(\Lambda )$ is a weak- $\ast $ -neighborhood of $\Lambda $ .
Let $\Lambda =(P,\alpha )\in \operatorname {Del}(U,K,\sigma )$ , $\phi \in {\mathcal C}_c(G)$ and $\varepsilon>0$ . To show that $\{ \mathcal {U}_{S,\delta }(\Lambda ): S \in \mathcal {RK}(G),\, \delta> 0 \} $ defines a neighborhood basis, it suffices to show that there is an ${S\in \mathcal {RK}(G)}$ and a $\delta>0$ such that
Without loss of generality, suppose that $e\in U$ and $B_{2\varepsilon }(x) \cap B_{2\varepsilon }(y) = \emptyset $ for all $x,y \in D$ with $x \neq y$ and for all U-uniformly discrete set D. Let $S\in \mathcal {RK}(G)$ be chosen such that it contains the compact support of $\phi $ and $S\cap P = S.B_{2\varepsilon } \cap P$ . Set $N:=\sharp (S.B_\varepsilon \cap P)$ . Since $\phi \in {\mathcal C}_c(G)$ , there is a $\delta>0$ such that $\delta <\min \{\varepsilon ,\sigma ^{-1}, ({\varepsilon }/{2N\|\phi \|_{\infty }})\}$ and $|\phi (x)-\phi (y)|< {\varepsilon }/{2N\sigma }$ holds for all $x,y\in G$ with $d_G(x,y)<\delta $ .
Let $\Pi =(Q,\beta )\in \mathcal {U}_{S.B_\varepsilon ,\delta }(\Lambda )$ . By the choice of S, each point $x \in P \cap S.B_{\varepsilon }$ is necessarily contained in $P \cap S$ . Hence, for each $x\in P\cap S.B_\varepsilon = P \cap S$ , the condition $|\delta _{\Lambda }(B_\delta (x)) - \delta _{\Pi }(B_\delta (x))|<\delta \leq \varepsilon $ implies that there is a unique $y_x\in Q\cap S.B_{\varepsilon }\cap B_\delta (x)$ such that $|\alpha (x)-\beta (y_x)|<\delta \leq {\varepsilon }/{2N\| \phi \|_{\infty }}$ . In particular, we have shown that $\{y_x\,: x\in P\cap S.B_\varepsilon \} \subseteq Q\cap S.B_\varepsilon $ . We claim the latter two sets are actually equal.
For indeed, if there was a $y\in Q\cap S.B_\varepsilon $ that is not equal to some $y_x$ , then $B_\delta (y)\cap P\subseteq S.B_{2\varepsilon }\cap P$ and this would lead to $B_\varepsilon (y)\cap P=\emptyset $ . Thus,
as $\Pi \in \mathcal {U}_{S.B_\varepsilon ,\delta }(\Lambda )$ , which is a contradiction to $\Pi \in \operatorname {Del}(U,K,\sigma )$ . Furthermore, ${\Pi \in \mathcal {U}_{S.B_\varepsilon ,\delta }(\Lambda )}$ yields
for all $x \in P \cap S.B_{\varepsilon }$ . Thus,
using $d_G(x,y_x)<\delta $ .