Hostname: page-component-54dcc4c588-smtgx Total loading time: 0 Render date: 2025-09-24T14:16:22.325Z Has data issue: false hasContentIssue false

Metallic mean Wang tiles II: the dynamics of an aperiodic computer chip

Published online by Cambridge University Press:  24 September 2025

Sébastien Labbé*
Affiliation:
CNRS, LaBRI, UMR 5800, https://ror.org/057qpr032 Université de Bordeaux , Talence F-33400, France

Abstract

We consider a new family $(\mathcal {T}_n)_{n\geq 1}$ of aperiodic sets of Wang tiles and we describe the dynamical properties of the set $\Omega _n$ of valid configurations $\mathbb {Z}^2\to \mathcal {T}_n$. The tiles can be defined as the different instances of a square-shaped computer chip whose inputs and outputs are 3-dimensional integer vectors. The family include the Ammann aperiodic set of 16 Wang tiles and gathers the hallmarks of other small aperiodic sets of Wang tiles. Notably, the tiles satisfy additive versions of equations verified by the Kari–Culik aperiodic sets of 14 and 13 Wang tiles. Also configurations in $\Omega _n$ are the codings of a $\mathbb {Z}^2$-action on a 2-dimensional torus like the Jeandel–Rao aperiodic set of 11 Wang tiles. The family broadens the relation between quadratic integers and aperiodic tilings beyond the omnipresent golden ratio as the dynamics of $\Omega _n$ involves the positive root $\beta $ of the polynomial $x^2-nx-1$, also known as the n-th metallic mean. We show the existence of an almost one-to-one factor map $\Omega _n\to \mathbb {T}^2$ which commutes the shift action on $\Omega _n$ with horizontal and vertical translations by $\beta $ on $\mathbb {T}^2$. The factor map can be explicitly defined by the average of the top labels from the same row of tiles as in Kari and Culik examples. The proofs are based on the minimality of $\Omega _n$ (proved in a previous article) and a polygonal partition of $\mathbb {T}^2$ which we show is a Markov partition for the toral $\mathbb {Z}^2$-action. The partition and the sets of Wang tiles are symmetric which makes them, like Penrose tilings, worthy of investigation.

Information

Type
Analysis
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1 Introduction

Turing machines can be encoded into a finite set of Wang tiles (unit squares with labeled edges) in such a way that the Turing machine does not halt if and only if there exists a tiling of the plane by translated copies of the tiles respecting the condition that the common edge of adjacent tiles have the same label [Reference Berger7], see also [Reference Robinson49, Reference Ollinger43, Reference Jeandel and Vanier22]. As a consequence, the existence of a valid tiling of the plane with a given finite set of Wang tiles (called the domino problem) cannot be decided by an algorithm. Indeed, if the domino problem were decidable, we could use the algorithm solving the domino problem to solve the halting problem, which is a contradiction [Reference Turing60].

Therefore, we can think of Wang tiles as if their tilings are computing something. As observed by Wang, the undecidability of the domino problem implies the existence of aperiodic sets of Wang tiles [Reference Wang62]. Shortly after, Berger proved the undecidability of the domino problem and constructed the first known aperiodic set of Wang tiles [Reference Berger7]. Since then, aperiodic tilings has developed into an active subject of study with applications to the theory of quasicrystals [Reference Grünbaum and Shephard19, Reference Senechal53, Reference Baake and Grimm5, Reference Baake and Grimm6]. Thus, sets of Wang tiles (and their computations) can be classified into three cases:

  • Finite: the Wang tiles do not tile the plane,

  • Periodic: the Wang tiles tile the plane and one of the valid tiling is periodic,

  • Aperiodic: the Wang tiles tile the plane and none of the valid tilings are periodic.

The finite cases can be associated with computations that halt. The periodic cases can be associated with computations that do not halt and fall into an infinite loop. The aperiodic cases can be associated with computations that do not halt and never repeat.

For applications, computations that halt are usually preferred over computations that loop forever. Among computations that halt, the description of those “busy beavers” [Reference Brady9, Reference Aaronson1] running for the maximum number of steps before halting is an open question even for Turing machines made of only 6 rules [42] (it was recently solved for 5 rulesFootnote 1 ). In this article, we are interested in the description of computations that do not halt and never repeat. We focus on those that happen to be performed by small aperiodic sets of Wang tiles. We aim to reveal their links with dynamical systems and the coding of their orbits.

The Kari–Culik outliers

The smallest sets of aperiodic Wang tiles until 2015 were discovered by Kari and Culik in 1996. Kari [Reference Kari24] proved that a well-chosen set of 14 Wang tiles admits tilings of the plane, and that none of them is periodic. The proof that they are not periodic is cleverly short. It is based on an arithmetic interpretation of the edge labels of the Wang tiles. The tiles have labels $r,t,\ell ,b\in \mathbb {Q}$ satisfying an equation

(1.1)

for some $q\in \mathbb {Q}$ . We may interpret the Wang tile as a computation (the multiplication by q) with value t as an input and b as an output. The value $\ell $ is a carry input on the left and r is a carry output on the right. Kari [Reference Kari24] proposed a set of four tiles satisfying (1.1) with $q=2$ and ten tiles with $q=\frac {2}{3}$ . The proof of the nonexistence of a periodic tiling with those 14 tiles uses the fact that the equation $2^m 3^n = 1$ has only one solution over the integers ( $m=n=0$ ), see Figure 1. Based on the same idea, Culik [Reference Culik11] proposed a smaller aperiodic set of 13 tiles (four tiles satisfying (1.1) with $q=3$ and nine tiles with $q=\frac {1}{2}$ ). Note that generalizations of Kari–Culik tilings exist [Reference Eigen, Navarro and Prasad15] and that further results were obtained about their entropy [Reference Durand, Gamard and Grandjean14] and on a minimal subsystem [Reference Siefken54].

Figure 1 Averages of horizontal labels in a tiling with Kari’s 14 tiles are orbits under the map g on the interval $[\frac {2}{3},2]$ ; see [Reference Durand, Gamard and Grandjean14, Reference Kari27].

Among aperiodic tilings of the plane by Wang tiles, Kari and Culik sets seem like outliers. The aperiodicity of Penrose tiles [Reference Penrose44], Berger tiles [Reference Berger7], Robinson tiles [Reference Robinson49], Knuth tiles [Reference Knuth29], Ammann tiles [Reference Grünbaum and Shephard19, Reference Ammann, Grünbaum and Shephard3] can be explained by the hierarchical decomposition of their tilings. Often, aperiodic tilings have a self-similar structure [Reference Solomyak58, Reference Solomyak59, Reference Priebe and Solomyak46, Reference Praggastis45, Reference Akiyama and Arnoux2] and this is the case for recently discovered aperiodic geometrical tiles [Reference Socolar and Taylor57, Reference Smith, Myers, Kaplan and Goodman-Strauss55, Reference Smith, Myers, Kaplan and Goodman-Strauss56]. However, Kari and Culik tilings have positive entropy. Thus, they are not self-similar and do not possess a hierarchical decomposition [Reference Durand, Gamard and Grandjean14]. Note that the absence of hierarchical decomposition also follows from a cylindricity argument proposed by Thierry Monteil and explained in [Reference Durand, Gamard and Grandjean14, §4.2]. Moreover, except some extensions of Kari and Culik sets [Reference Eigen, Navarro and Prasad15, §Reference Baake and Grimm6], no other known aperiodic sets of tiles satisfy equations explaining their nonperiodicity.

The metallic mean family of aperiodic Wang tiles

The current article is the second article about a new family of aperiodic Wang tiles related to the metallic mean. Recall that the metallic mean $\beta $ is the positive root of the polynomial $x^2-nx-1$ where $n\geq 1$ is an integer [Reference de Spinadel13], that is,

$$\begin{align*}\beta = [n; n, n, \cdots] = n + \frac{1}{ n + \frac{1}{ n + \frac{1}{ n + \cdots}}} = n + \frac{1}{\beta}. \end{align*}$$

Metallic means were also called silver means in [Reference Schroeder52] and noble means in [Reference Baake and Grimm5].

Let us recall the main results proved in the first article of the series. For every integer $n\geq 1$ , the $n^{th}$ metallic mean Wang shift $\Omega _n$ is defined from a set $\mathcal {T}_n$ of $(n+3)^2$ Wang tiles. An illustration of the set $\mathcal {T}_3$ is shown in Figure 2. The labels of the Wang tiles are vectors in $\mathbb {N}^3$ . In Figure 2, we represent vectors as words for economy of space reasons. For instance, the vector $(1,1,4)$ is represented as $114$ . A finite rectangular valid tiling is shown in Figure 3 for the set $\mathcal {T}_3$ . More images of valid tilings with metallic mean Wang tiles are available in [Reference Labbé37].

Figure 2 The metallic mean Wang tile set $\mathcal {T}_n$ for $n=3$ .

Figure 3 A valid $15\times 15$ pattern with Wang tile set $\mathcal {T}_3$ .

It was shown in the previous article that the metallic mean Wang shift $\Omega _n$ is self-similar, aperiodic and minimal. We gather in the next theorem the main results already proved about $\Omega _n$ .

Theorem 1.1 [Reference Labbé37]

For every integer $n\geq 1$ ,

  1. (i) the metallic mean Wang shift $\Omega _n$ is self-similar, aperiodic and minimal,

  2. (ii) the inflation factor of the self-similarity of $\Omega _n$ is the n-th metallic mean, that is, the positive root of $x^2-nx-1$ .

Also, when $n=1$ , $\Omega _1$ is equivalent to the Wang shift defined from the 16 Ammann Wang tiles [Reference Grünbaum and Shephard19, p.595, Figure 11.1.13].

In order to describe the substitutive structure of the Wang shift $\Omega _n$ generated from the set $\mathcal {T}_n$ , it was needed in [Reference Labbé37] to introduce a larger set $\mathcal {T}_n'$ satisfying $\mathcal {T}_n\subseteq \mathcal {T}_n'$ . It was shown that the set $\mathcal {T}_n'$ is in bijection with the set of possible return blocks allowing to decompose uniquely the configurations of $\Omega _n$ . The return blocks are rectangular blocks of tiles with a unique junction tile (a tile where horizontal and vertical color stripes intersect) at the lower left corner. Also, it was proved in [Reference Labbé37] that in a valid configuration of $\Omega _n'$ , only the tiles from $\mathcal {T}_n$ appear. From this observation follows the self-similarity of $\Omega _n$ .

This article

In this article, we demonstrate that Kari and Culik tilings are not a complete oddity within aperiodic sets of tiles. In particular, we show for the first time that substitutive aperiodic sets of Wang tiles can also satisfy equations and even be defined by them, see Figure 4. This article is devoted to a family of aperiodic Wang tiles associated with the metallic mean numbers, the positive roots of the polynomials $x^2-nx-1$ where $n\geq 1$ is a positive integer. When $n=1$ , the family recovers the Ammann set of 16 Wang tiles [Reference Grünbaum and Shephard19].

Figure 4 A Venn diagram of aperiodic sets of Wang tiles. Aperiodicity of Kari [Reference Kari24] and Culik [Reference Culik11] sets of tiles and their extensions [Reference Eigen, Navarro and Prasad15] follows from the arithmetic equations satisfied by their matching rules. In this article, we show that the dashed region in the Venn diagram is nonempty, that is, there exists a family of substitutive (self-similar) aperiodic sets of Wang tiles whose matching rules satisfy arithmetic equations.

The labels of the Wang tiles are not numbers like in Kari and Culik sets, but rather integer vectors. Note that integers vectors were already used as labels of Wang tiles in [Reference Kari25, Reference Kari, Geffert, Karhumäki, Bertoni, Preneel, Návrat and Bieliková26], see also [Reference Kari27]. The equations satisfied by the tiles are derived from a function that expresses a relation between the labels of the Wang tiles. The function provides an independent definition of the family of metallic mean Wang tiles as the instances of an aperiodic computer chip. The family $(\Omega _n)_{n\geq 1}$ of metallic mean Wang shifts was introduced separately in [Reference Labbé37] where it was shown to be aperiodic as a consequence of its self-similarity.

Here, in this second article on the metallic mean Wang tiles, we prove that $\Omega _n$ is aperiodic for another reason. Namely, we show that the $\mathbb {Z}^2$ shift action on $\Omega _n$ is an almost 1-to-1 extension of a minimal $\mathbb {Z}^2$ -action by rotations on $\mathbb {T}^2$ . This reminds of a result proved for Penrose tilings [Reference Robinson48] and the two reasons for them to be aperiodic. Aperiodicity of Penrose tilings follows from its self-similarity [Reference Penrose44] and from their being a cut-and-project scheme [Reference de Bruijn12, Reference Baake and Grimm5].

For every integer $n\geq 1$ , we show that valid configurations in $\Omega _n$ are computing the orbits of a dynamical system defined by a $\mathbb {Z}^2$ -action $R_n$ on the 2-dimensional torus $\mathbb {T}^2$ . The dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ is defined by horizontal and vertical translation on $\mathbb {T}^2$ by the n-th metallic mean modulo 1. As for the Jeandel–Rao Wang shift [Reference Labbé33], the proof is based on a polygonal partition of $\mathbb {T}^2$ which we prove is a Markov partition for the toral $\mathbb {Z}^2$ -action. We also prove the existence of an almost one-to-one factor map $\Omega _n\to \mathbb {T}^2$ commuting the shift $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ with the toral $\mathbb {Z}^2$ -rotation $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ . Since $R_n$ is a free action, this provides a second reason for the Wang shift $\Omega _n$ to be aperiodic.

The factor map can be defined by taking averages of the dot product involving the top labels of the Wang tiles in the biinfinite row of tiles passing through the origin in a configuration. The existence of the factor map proves that the average changes from row to row by an irrational rotation by the n-th metallic mean number. This can be seen as an additive version of a multiplicative phenomenon known for Kari–Culik tilings. Recall that the average of top label values along a row is at the heart of Kari and Culik’s construction of aperiodic tilings where the average change by a rational multiplication from row to row [Reference Durand, Gamard and Grandjean14, Theorem 6].

The polygonal partition used to encode the toral $\mathbb {Z}^2$ -action is symmetric and is much more simple to define compared to the Markov partition associated with the Jeandel–Rao Wang shift. Moreover, the label of the polygonal atoms of the partition have a meaning in the sense that they define the linear inequalities describing their boundaries. The symmetry and simplicity of the partition was helpful to extend the family beyond the golden ratio. The results proved here for the metallic mean Wang tiles should serve as an inspiration to replace the labels of the Jeandel–Rao tiles by integer vectors satisfying equations. Understanding the matching rules of Jeandel–Rao tiles by means of arithmetic would open the door for discovering a vast family of aperiodic sets of Wang tiles beyond the family of metallic mean Wang tiles. See Section 11 for more open questions.

Structure of the article

In Section 2, we state the main results proved in this article. In Section 3, we present preliminary notions on dynamical systems, subshifts and Wang shifts. In Section 4, we recall the definition of the family of metallic mean Wang tiles. In Section 5, we show that instances of the $\theta _n$ -chip are the metallic mean Wang tiles. This proves Theorem A. In Section 6, we prove Theorem B and we present more equations satisfied by the metallic mean tiles and their tilings. In Section 7, we use the floor function on linear forms to construct valid tilings with the metallic mean Wang tiles and we prove Theorem C. In Section 8, we define an explicit factor map $\Omega _n\to \mathbb {T}^2$ and we prove Theorem D. In Section 9, we define the partition $\mathcal {P}_n$ for every integer $n\geq 1$ and we show that the metallic mean Wang shift is equal to the symbolic dynamical system defined by the coding of a toral $\mathbb {Z}^2$ -action by this partition. This shows that $\Omega _n$ is isomorphic as measure-preserving dynamical systems to a toral $\mathbb {Z}^2$ -action. We prove Theorem E and Theorem F in this section. In Section 10, we compute the renormalization of the partition $\mathcal {P}_n$ and $\mathbb {Z}^2$ -action $R_n$ using computations performed in SageMath when $n=3$ . We illustrate how the Rauzy induction of $\mathbb {Z}^2$ -actions and of polygonal partitions can be used to show the self-similarity of the symbolic dynamical system $\mathcal {X}_{\mathcal {P}_n,R_n}$ . In Section 11, we discuss some open questions raised by the current work.

2 Statements of the main results

An aperiodic computer chip

For every integer $n\geq 1$ , we define a finite subset $V_n\subset \mathbb {N}^3$ of vectors

$$\begin{align*}V_n = \{(v_0,v_1,v_2)\in\mathbb{N}^3\colon 0\leq v_0\leq v_1\leq 1 \text{ and } v_1\leq v_2\leq n+1\} \end{align*}$$

with nondecreasing entries where the middle entry is at most 1. We introduce a function

$$\begin{align*}\begin{array}{rccl} \theta_n:&V_n\times V_n & \to & \mathbb{Z}^3\\ &(u_0,u_1,u_2), (v_0,v_1,v_2) & \mapsto & (r_0,r_1,r_2), \end{array} \end{align*}$$

taking two vectors as input and returning one vector. Its image is defined by the rule

(2.1) $$ \begin{align} \left\{ \begin{array}{ll} &r_0=u_0,\\ &r_1=\begin{cases} v_2-n & \text{ if } u_0 = 0,\\ 1 & \text{ if } u_0 = 1, \end{cases}\\ &r_2=\begin{cases} v_1+u_0 & \text{ if } v_0 = 0,\\ u_2+1 & \text{ if } v_0 = 1. \end{cases} \end{array} \right. \end{align} $$

Notice that $(r_0,r_1,r_2)$ does not depend on $u_1$ . For every integer $n\geq 1$ , we construct a symmetric $\theta _n$ -chip, that is, a computer chip taking as inputs $u\in V_n$ on the left and $v\in V_n$ on the bottom and producing as outputs $\theta _n(u,v)$ on the right and $\theta _n(v,u)$ on the top (see Figure 5).

Figure 5 The $\theta _n$ -chip is a computer chip computing $\theta _n(u,v)$ and $\theta _n(v,u)$ from the left input u and bottom input v.

If $\theta _n(u,v)$ and $\theta _n(v,u)$ are in $V_n$ , then one can use multiple copies of the $\theta _n$ -chip and connect them to each other horizontally and vertically into an arbitrarily large rectangular cluster of $\theta _n$ -chips (see Figure 6).

Figure 6 A rectangular cluster of copies of the $\theta _n$ -chip.

We prove in this work the existence of arbitrarily large rectangular clusters of the $\theta _n$ -chip all of them performing correct computations. Also we show that no rectangular cluster of the $\theta _n$ -chip performs a periodic computation. Thus, we say that the $\theta _n$ -chip is an aperiodic computer chip. Perhaps we can say it is an aperiodic monochip, but we cannot say it is an aperiodic monotile as in [Reference Smith, Myers, Kaplan and Goodman-Strauss55, Reference Smith, Myers, Kaplan and Goodman-Strauss56] because the same chip with different inputs has to be considered a distinct Wang tile.

Instances of the chip are metallic mean Wang tiles

If we consider all possible values of inputs u and v in $V_n$ and if we restrict the outputs to be in the set $V_n$ , then we obtain a finite set of Wang tiles

(2.2)

which is the finite set of all possible instances of the $\theta _n$ -chip.

Theorem A. For every integer $n\geq 1$ , the Wang shift $\Omega _{\mathcal {C}_n}$ defined by the $\theta _n$ -chip is the $n^{th}$ metallic mean Wang shift $\Omega _n$ .

Something unexpected and surprising happens in the proof of Theorem A. The set $\mathcal {C}_n$ of instances of the $\theta _n$ -chip is exactly equal to the extended set $\mathcal {T}_n'$ of metallic mean Wang tiles introduced in [Reference Labbé37] in order to prove the self-similarity of $\Omega _n$ , see Proposition 5.1.

Tile labels satisfy Equations

The next result states that every tile in $\mathcal {C}_n$ satisfy a system of equations. While the equations associated with Kari’s [Reference Kari24] and Culik’s [Reference Culik11] aperiodic set of Wang tiles are multiplicative, the ones associated with $\mathcal {C}_n$ are additive.

Theorem B. Let $n\geq 1$ be an integer, $d=(0,-1,1)$ and $e=(1,0,0)$ . The set of Wang tiles defined by the $\theta _n$ -chip satisfy the following system of equations:

where $\langle \_,\_\rangle $ denotes the canonical inner product of ${\mathbb {Z}^{3}}$ .

Equivalently, if we let $\ell =(\ell _0,\ell _1,\ell _2)$ , $b=(b_0,b_1,b_2)$ , $r=(r_0,r_1,r_2)$ and $t=(t_0,t_1,t_2)$ , the equations in the theorem say that tiles in ${\mathcal {C}_{n}}$ satisfy $\ell _0=r_0$ , $b_0=t_0$ and

(2.3) $$ \begin{align} {\frac{t_2-t_1+\ell_2-\ell_1}{n}}-\ell_0 = {\frac{b_2-b_1+r_2-r_1}{n}}-b_0 \end{align} $$

which reminds of Equation (1.1).

Like Kari’s and Culik’s tiles, these equations behave well with tilings and more equations can be deduced for valid tilings of a rectangle, see Section 6. In particular, Equation (6.2) says that in a tiling of a cylinder of height k, the average of the inner product with $\textstyle \frac {1}{n}d$ of the top labels of the cylinder is obtained from the average of the inner product with $\textstyle \frac {1}{n}d$ of the bottom labels of the cylinder by k rotations on the unit circle by a fixed angle. The angle is equal to the frequency of columns in the cylinder containing junction tiles and vertical strip colored tiles, which is a rational number. Therefore, the existence of a cyclic rectangle is not directly forbidden from these equations. Note that we know from the self-similarity of $\Omega _n$ that the frequency of columns containing junction tile in every valid configuration in $\Omega _n$ is equal to $\beta ^{-1}$ , which is an irrational number [Reference Labbé37].

It remains an open problem to deduce the aperiodicity of the Wang shift $\Omega _n$ from the equations satisfied by the labels of $\theta _n$ -chip as this is nicely done for Kari and Culik sets of tiles. See Section 11 for related open questions.

Existence of valid tilings

Valid configurations in $\Omega _n$ can be constructed using the floor function on linear forms. Let $\Lambda _n:[0,1)^2\to \mathbb {Z}^3$ be defined as

$$\begin{align*}\Lambda_n(x,y) = \left( \begin{array}{r} \lfloor y-\beta^{-1}+1\rfloor\\ \lfloor \beta^{-1}x + y-\beta^{-1}+1\rfloor\\ \lfloor \beta x + y-\beta^{-1}+1\rfloor \end{array} \right). \end{align*}$$

where $\beta $ is the $n^{th}$ metallic mean, that is, the positive root of the polynomial $x^2-nx-1$ . For every $(x,y)\in \mathbb {R}^2$ , let

be a Wang tile where $\{x\}=x-\lfloor x\rfloor $ is the fractional part of a number $x\in \mathbb {R}$ .

Theorem C. For every integer $n\geq 1$ and every $(x,y)\in [0,1)^2$ , the configuration

is a valid tiling of the plane by the set of metallic mean Wang tiles $\mathcal {T}_n$ .

This construction reminds of the proof of existence of tilings with Kari and Culik tiles based on the balanced representation of real numbers and first difference of Beatty sequences [Reference Kari24, Reference Culik11], see also [Reference Eigen, Navarro and Prasad15, Reference Siefken54].

A factor map defined from averages of tile labels

In Kari–Culik tilings [Reference Kari24, Reference Culik11], there is a well-defined notion of average [Reference Durand, Gamard and Grandjean14] of the top tile labels along a bi-infinite horizontal row. The change of value from one row to the next row is described by a piecewise rationally multiplicative map. In this context, metallic mean Wang shifts also behave like Kari–Culik tilings. It involves the consideration of the average of specific inner products and irrational rotations instead of multiplications, see Figure 7 which can be compared with Figure 1.

Figure 7 A $10\times 5$ valid rectangular tiling with the set $\mathcal {T}_n$ with $n=3$ . The numbers indicated in the right margin are the average of the inner products $\langle \frac {1}{n}d,v\rangle $ over the vectors v appearing as top (or bottom) labels of a horizontal row of tiles and where $d=(0,-1,1)$ . We observe that these numbers increase by $\frac {3}{10}\ \pmod 1$ from row to row. The number $\frac {3}{10}$ is equal to the frequency of columns containing junction tiles (a junction tile is a tile whose labels all start with 0). Observe that this is a cylindrical tiling (left and right outer labels of the rectangle match) which simplifies the equations involved because the left and right carries cancel.

We show that the average of the dot products of the vector $\frac {1}{n}d=\frac {1}{n}(0,-1,1)$ with the top labels of a given row in a valid configuration $\mathbb {Z}^2\to \mathcal {T}_n$ in $\Omega _n$ is well-defined and takes a value in the interval $[0,1]$ (see Equation (8.1)). By symmetry of the set $\mathcal {T}_n$ , the same holds for the right labels of a given column. By considering the row and column going through the origin of a configuration, the two averages define a map $\Phi _n:\Omega _n\to \mathbb {T}^2$ (see Equation (8.2)). We prove that this map is a factor map from the Wang shift to the 2-torus.

Theorem D. Let $d=(0,-1,1)$ , $n\geq 1$ be an integer and $\Omega _n$ be the $n^{th}$ metallic mean Wang shift. The map

(2.4)

is a factor map, that is, it is continuous, onto and commutes the shift $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ with the toral $\mathbb {Z}^2$ -rotation $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ by the equation $ \Phi _n\circ \sigma ^k = R_n^k\circ \Phi _n $ for every $k\in \mathbb {Z}^2$ where

$$\begin{align*}\begin{array}{rccl} R_n&:\mathbb{Z}^2\times\mathbb{T}^2 & \to & \mathbb{T}^2\\ &(k,x) & \mapsto & R_n^k(x):=x + \beta k \end{array} \end{align*}$$

and $\beta =\frac {n+\sqrt {n^2+4}}{2}$ is the $n^{th}$ metallic mean, that is, the positive root of the polynomial $x^2-nx-1$ .

As a consequence of Theorem D, we deduce that $\Omega _n$ is aperiodic because $\beta $ is irrational and $R_n$ is a free $\mathbb {Z}^2$ -action, see Corollary 8.3. Note that since $\beta -\beta ^{-1}=n$ , we have $\beta =\beta ^{-1}\ \pmod 1$ .

Theorem D is an analogue of a result known for Kari and Culik aperiodic Wang tilings which satisfy equations involving balanced representations of real numbers and orbits of piecewise rationally multiplicative maps, see also Theorem 16 in [Reference Eigen, Navarro and Prasad15] and Proposition 3 in [Reference Siefken54]. Here the result applies to all of the configurations in the Wang shift $\Omega _n$ .

A symbolic dynamical system and a Markov partition

The Wang shift $\Omega _n$ can be independently described as a symbolic representation of the dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ by encoding its orbits with an appropriate topological partition of $\mathbb {T}^2$ . The partition of $\mathbb {T}^2$ naturally emerges from the set of preimages of the map and from Theorem C.

Since $\Lambda _n$ is defined as the floor of linear forms, for every tile $t\in \mathcal {T}_n$ , the set

is a polygonal open region in the unit square. It satisfies that $\mathcal {P}_n=\{P_t\mid t\in \mathcal {T}_n\}$ is a topological partition of $\mathbb {T}^2$ made of $(n+3)^2$ atoms. The polygonal partition $\mathcal {P}_n$ is the refinement of two polygonal partitions and , the second one being the image of the first under a symmetry by the positive diagonal. The partition can be constructed by drawing the following geodesics on the torus $\mathbb {T}^2$ :

  • two closed geodesics of slope $0$ and $\infty $ going through the origin $(0,0)$ ,

  • a closed geodesic of slope 0 going through the point $(0,\beta ^{-1})$ ,

  • a geodesic of slope $-\beta ^{-1}$ from $(0,\beta ^{-1})$ to $(1,0)$ ,

  • a geodesic of slope $-\beta $ from $(0,\beta ^{-1})$ to $(1,0)$ wrapping around the unit square fundamental domain n times.

See an illustration of $\mathcal {P}_n$ when $n=3$ in Figure 8. Every open region defined by the complement of the geodesics can be identified with a pair of vectors in $V_n$ and a unique tile in $\mathcal {T}_n$ with such top and right labels. As opposed to the four topological polygonal partitions associated with Jeandel-Rao tilings [Reference Labbé33], $\mathcal {P}_n$ can be computed only from and without considering the and partitions. This is because the set $\mathcal {T}_n$ of tiles is NE-deterministic, see Theorem 5.3.

Figure 8 The partition and its image under a symmetry with the positive diagonal. Their refinement is $\mathcal {P}_3$ which is a partition of the unit square into 36 polygonal atoms. Here $\beta $ is the third metallic mean, that is, the positive root of $x^2-3x-1$ .

The encoding of $\mathbb {Z}^2$ -orbits under $R_n$ by the topological partition $\mathcal {P}_n$ are 2-dimensional configurations whose topological closure is the symbolic dynamical system $\mathcal {X}_{\mathcal {P}_n,R_n}$ . We prove that $\mathcal {X}_{\mathcal {P}_n,R_n}=\Omega _n$ , and since $\Omega _n$ is a subshift of finite type by definition, we have the following theorem.

Theorem E. For every integer $n\geq 1$ , the symbolic dynamical system $\mathcal {X}_{\mathcal {P}_n,R_n}$ corresponding to $\mathcal {P}_n,R_n$ is equal to the metallic mean Wang shift $\Omega _n$ :

$$\begin{align*}\Omega_n = \mathcal{X}_{\mathcal{P}_n,R_n}. \end{align*}$$

In particular, $\mathcal {P}_n$ is a Markov partition for the dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ .

Markov partitions were originally defined for one-dimensional dynamical systems $\mathbb {Z}\overset {T}{\curvearrowright }\mathbb {T}^2$ and were extended to $\mathbb {Z}^d$ -actions by automorphisms of compact Abelian group in [Reference Einsiedler and Schmidt16]. Following [Reference Labbé33, Reference Labbé34], we use the same terminology and extend the definition proposed in [Reference Lind and Marcus40, §6.5] for dynamical systems defined by higher-dimensional actions by rotations, see Definition 9.1.

The maximal equicontinuous factor and an isomorphism

Using Theorem E and applying the results already proved for Jeandel–Rao Wang shift [Reference Labbé33], we have the following additional topological and measurable properties for the factor map. We refer the reader to the preliminary Section 3 for the notions and vocabulary on topological and measure-preserving dynamical systems that are used in the statement. A similar result holds for Penrose tilings [Reference Robinson48].

Theorem F. The Wang shift $\Omega _n$ and the $\mathbb {Z}^2$ -action $R_n$ have the following properties:

  1. (i) $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ is the maximal equicontinuous factor of $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ ,

  2. (ii) the factor map $\Phi _n:\Omega _n\to \mathbb {T}^2$ is almost one-to-one and its set of fiber cardinalities is $\{1,2,8\}$ ,

  3. (iii) the shift-action $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ on the metallic mean Wang shift is uniquely ergodic,

  4. (iv) the measure-preserving dynamical system $(\Omega _n,\mathbb {Z}^2,\sigma ,\nu )$ is isomorphic to $(\mathbb {T}^2,\mathbb {Z}^2,R_n,\lambda )$ where $\nu $ is the unique shift-invariant probability measure on $\Omega _n$ and $\lambda $ is the Haar measure on $\mathbb {T}^2$ .

3 Preliminaries on dynamical systems, subshifts and Wang shifts

This section follows the preliminary section of the chapter [Reference Labbé36] and article [Reference Labbé33].

3.1 Topological dynamical systems

Most of the notions introduced here can be found in [Reference Walters61]. A dynamical system is a triple $(X,G,T)$ , where X is a topological space, G is a topological group and T is a continuous function $G\times X\to X$ defining a left action of G on X: if $x\in X$ , e is the identity element of G and $g,h\in G$ , then using additive notation for the operation in G we have $T(e,x)=x$ and $T(g+h,x)=T(g,T(h,x))$ . In other words, if one denotes the transformation $x\mapsto T(g,x)$ by $T^g$ , then $T^{g+h}=T^g T^h$ . In this work, we consider the Abelian group $G=\mathbb {Z}\times \mathbb {Z}$ .

If $Y\subset X$ , let $\overline {Y}$ denote the topological closure of Y and let $\overline {Y}^T:=\cup _{g\in G}T^g(Y)$ denote the T-closure of Y. A subset $Y\subset X$ is T -invariant if $\overline {Y}^T=Y$ . A dynamical system $(X,G,T)$ is called minimal if X does not contain any nonempty, proper, closed T-invariant subset. The left action of G on X is free if $g=e$ whenever there exists $x\in X$ such that $T^g(x)=x$ .

Let $(X,G,T)$ and $(Y,G,S)$ be two dynamical systems with the same topological group G. A homomorphism $\theta :(X,G,T)\to (Y,G,S)$ is a continuous function $\theta :X\to Y$ satisfying the commuting property that $S^g\circ \theta =\theta \circ T^g$ for every $g\in G$ . A homomorphism $\theta :(X,G,T)\to (Y,G,S)$ is called an embedding if it is one-to-one, a factor map if it is onto, and a topological conjugacy if it is both one-to-one and onto and its inverse map is continuous. If $\theta :(X,G,T)\to (Y,G,S)$ is a factor map, then $(Y,G,S)$ is called a factor of $(X,G,T)$ and $(X,G,T)$ is called an extension of $(Y,G,S)$ . Two dynamical systems are topologically conjugate if there is a topological conjugacy between them.

A measure-preserving dynamical system is defined as a system $(X,G,T,\mu ,\mathcal {B})$ , where $\mu $ is a probability measure defined on the Borel $\sigma $ -algebra $\mathcal {B}$ of subsets of X, and $T^g:X\to X$ is a measurable map which preserves the measure $\mu $ for all $g\in G$ , that is, $\mu (T^g(B))=\mu (B)$ for all $B\in \mathcal {B}$ . The measure $\mu $ is said to be T -invariant. In what follows, when it is clear from the context, we omit the Borel $\sigma $ -algebra $\mathcal {B}$ of subsets of X and write $(X,G,T,\mu )$ to denote a measure-preserving dynamical system.

The set of all T-invariant probability measures of a dynamical system $(X,G,T)$ is denoted by $\mathcal {M}^T(X)$ . A T-invariant probability measure on X is called ergodic if for every set $B\in \mathcal {B}$ such that $T^{g}(B)=B$ for all $g\in G$ , we have that B has either zero or full measure. A dynamical system $(X,G,T)$ is uniquely ergodic if it has only one invariant probability measure, that is, $|\mathcal {M}^T(X)|=1$ . One can prove that a uniquely ergodic dynamical system is ergodic. A dynamical system $(X,G,T)$ is said strictly ergodic if it is uniquely ergodic and minimal.

Let $(X,G,S,\mu ,\mathcal {A})$ and $(Y,G,T,\nu ,\mathcal {B})$ be two measure-preserving dynamical systems. We say that the two systems are isomorphic (mod 0) if there exist measurable sets $X_0\subset X$ and $Y_0\subset Y$ of full measure (i.e., $\mu (X_0)=1$ and $\nu (Y_0)=1$ ) with $S^g(X_0)\subset X_0$ , $T^g(Y_0)\subset Y_0$ for all $g\in G$ and there exists a bi-measurable bijection $\phi _0:X_0\to Y_0$ ,

  • which is measure-preserving, that is, $\mu (\phi _0^{-1}(B))=\nu (B)$ for all measurable sets $B\subset Y_0$ ,

  • satisfying $\phi _0\circ S^g(x)=T^g\circ \phi _0(x)$ for all $x\in X_0$ and $g\in G$ .

The role of the set $X_0$ is to make precise the fact that the properties of the isomorphism need to hold only on a set of full measure. In this case, we call $\phi _0$ an isomorphism (mod 0) with respect to $\mu $ and $\nu $ . We also refer to an everywhere defined measurable map $\phi :X\to Y$ as an isomorphism (mod 0) with respect to $\mu $ and $\nu $ if $\phi (x)=\phi _0(x)$ with $x\in X$ for some $\phi _0$ and $X_0$ as above. When $\phi $ is also a factor map, some authors say that $\phi $ is a topo-isomorphism in order to express both its topological and measurable nature [Reference Fuhrmann, Gröger and Lenz18].

3.2 Maximal equicontinuous factor

A metrizable dynamical system $(X,G,T)$ is called equicontinuous if the family of homeomorphisms $\{T^{g}\}{}_{g\in G}$ is equicontinuous, that is, if for all ${\varepsilon }>0$ there exists ${\delta }>0$ such that

$$\begin{align*}{\mathrm{dist}}(T^{g}(x), T^{g}(y)) {<} {\varepsilon} \end{align*}$$

for all $g\in G$ and all $x,y\in X$ with ${\mathrm {dist}}(x,y)<{\delta }$ . According to a well-known theorem [Reference Aujogue, Barge, Kellendonk and Lenz4, Theorem 3.2], equicontinuous minimal systems defined by the action of an Abelian group are rotations on groups.

We say that $\theta :(X,G,T)\to (Y,G,S)$ is an equicontinuous factor if $\theta $ is a factor map and $(Y,G,S)$ is equicontinuous. We say that $(X_{\mathrm {max}}, G, T_{\mathrm {max}})$ is the maximal equicontinuous factor of $(X,G,T)$ if there exists an equicontinuous factor $\pi _{\mathrm {max}}:(X,G,T)\to (X_{\mathrm {max}}, G, T_{\mathrm {max}})$ , such that for any equicontinuous factor $\theta :(X,G,T)\to (Y,G,S)$ , there exists a unique factor map $\psi :(X_{\mathrm {max}}, G, T_{\mathrm {max}})\to (Y,G,S)$ with $\psi \circ \pi _{\mathrm {max}}=\theta $ . The maximal equicontinuous factor exists and is unique (up to topological conjugacy), see [Reference Aujogue, Barge, Kellendonk and Lenz4, Theorem 3.8] and [Reference Kurka31, Theorem 2.44].

Let $\theta :(X,G,T)\to (Y,G,S)$ be a factor map. We call the preimage set $\theta ^{-1}(y)$ of a point $y\in Y$ the fiber of $\theta $ over y. The cardinality of the fiber $\theta ^{-1}(y)$ for some $y\in Y$ has an important role and is related to the definition of other notions, see [Reference Aujogue, Barge, Kellendonk and Lenz4]. In particular, the factor map $\theta $ is almost one-to-one if $\{y\in Y:\mathrm {card}(\theta ^{-1}(y))=1\}$ is a $G_\delta $ -dense set in Y (that is a countable intersection of open sets which is dense in Y). In that case, $(X,G,T)$ is an almost one-to-one extension of $(Y,G,S)$ . The set of fiber cardinalities of a factor map $\theta :(X,G,T)\to (Y,G,S)$ is the set $\{\mathrm {card}(\theta ^{-1}(y)) : y \in Y\}\subset \mathbb {N}\cup \{\infty \}$ , see [Reference Fiebig17]. The set of fiber cardinalities of the maximal equicontinuous factor of a minimal dynamical system is invariant under topological conjugacy, see for instance [Reference Labbé33, Lemma 2.2].

3.3 Subshifts and shifts of finite type

In this section, we introduce multidimensional subshifts, a particular type of dynamical systems [Reference Lind and Marcus40, §13.10], [Reference Schmidt51, Reference Lind39, Reference Hochman20]. Let $\mathcal {A}$ be a finite set, $d\geq 1$ , and let $\mathcal {A}^{\mathbb {Z}^d}$ be the set of all maps $x:\mathbb {Z}^d\to \mathcal {A}$ , equipped with the compact product topology. An element $x\in \mathcal {A}^{\mathbb {Z}^d}$ is called configuration and we write it as $x=(x_{\boldsymbol{m}})=(x_{\boldsymbol{m}}:{\boldsymbol{m}}\in \mathbb {Z}^d)$ , where $x_{\boldsymbol{m}}\in \mathcal {A}$ denotes the value of x at ${\boldsymbol{m}}$ . The topology on $\mathcal {A}^{\mathbb {Z}^d}$ is compatible with the metric defined for all configurations $x,x'\in {\mathcal {A}}^{\mathbb {Z}^d}$ by ${\mathrm {dist}}(x,x')=2^{-\min \left \{\Vert {\boldsymbol{n}}\Vert \,:\, x_{\boldsymbol{n}}\neq x^{\prime }_{\boldsymbol{n}}\right \}}$ where $\Vert {\boldsymbol{n}}\Vert = |n_1| + \dots + |n_d|$ . The shift action $\sigma :{\boldsymbol{n}}\mapsto \sigma ^{\boldsymbol{n}}$ of the additive group $\mathbb {Z}^d$ on $\mathcal {A}^{\mathbb {Z}^d}$ is defined by

(3.1) $$ \begin{align} (\sigma^{\boldsymbol{n}}(x))_{\boldsymbol{m}} = x_{{\boldsymbol{m}}+{\boldsymbol{n}}} \end{align} $$

for every $x=(x_{\boldsymbol{m}})\in \mathcal {A}^{\mathbb {Z}^d}$ and ${\boldsymbol{n}}\in \mathbb {Z}^d$ . If $X\subset \mathcal {A}^{\mathbb {Z}^d}$ , let $\overline {X}$ denote the topological closure of X and let ${\overline {X}^{\sigma }}:=\{\sigma ^{\boldsymbol{n}}(x)\mid x\in X, {\boldsymbol{n}}\in \mathbb {Z}^d\}$ denote the shift-closure of X. A subset $X\subset \mathcal {A}^{\mathbb {Z}^d}$ is shift-invariant if ${\overline {X}^{\sigma }}=X$ . A closed, shift-invariant subset $X\subset \mathcal {A}^{\mathbb {Z}^d}$ is a subshift. If $X\subset \mathcal {A}^{\mathbb {Z}^d}$ is a subshift we write $\sigma =\sigma ^X$ for the restriction of the shift action (3.1) to X. When X is a subshift, the triple $(X,\mathbb {Z}^d,\sigma )$ is a dynamical system and the notions presented in the previous section hold.

A configuration $x\in X$ is periodic if there is a nonzero vector ${\boldsymbol{n}}\in \mathbb {Z}^d\setminus \{{\boldsymbol{0}}\}$ such that $x=\sigma ^{\boldsymbol{n}}(x)$ and otherwise it is nonperiodic. We say that a nonempty subshift X is aperiodic if the shift action $\sigma $ on X is free.

For any subset $S\subset \mathbb {Z}^d$ let $\pi _S:\mathcal {A}^{\mathbb {Z}^d}\to \mathcal {A}^S$ denote the projection map which restricts every $x\in \mathcal {A}^{\mathbb {Z}^d}$ to S. A pattern is a function $p\in \mathcal {A}^S$ for some finite subset $S\subset \mathbb {Z}^d$ . To every pattern $p\in \mathcal {A}^S$ corresponds a subset $\pi _S^{-1}(p)\subset \mathcal {A}^{\mathbb {Z}^d}$ called cylinder. A nonempty set $X\subset \mathcal {A}^{\mathbb {Z}^d}$ is a subshift if and only if there exists a set $\mathcal {F}$ of forbidden patterns such that

(3.2) $$ \begin{align} X = \{x\in\mathcal{A}^{\mathbb{Z}^d} \mid \pi_S\circ\sigma^{\boldsymbol{n}}(x)\notin\mathcal{F} \text{ for every } {\boldsymbol{n}}\in\mathbb{Z}^d \text{ and } S\subset\mathbb{Z}^d\}, \end{align} $$

see [Reference Hochman20, Prop. 9.2.4]. A subshift $X\subset \mathcal {A}^{\mathbb {Z}^d}$ is a subshift of finite type (SFT) if there exists a finite set $\mathcal {F}$ such that (3.2) holds. In this article, we consider shifts of finite type on $\mathbb {Z}\times \mathbb {Z}$ , that is, the case $d=2$ .

3.4 Wang shifts

A Wang tile is a tuple of four colors $(a,b,c,d)\in I\times J\times I\times J$ where I is a finite set of vertical colors and J is a finite set of horizontal colors, see [Reference Wang62, Reference Robinson49]. A Wang tile is represented as a unit square with colored edges:

For each Wang tile $\tau =(a,b,c,d)$ , let , , , denote respectively the colors of the right, top, left and bottom edges of $\tau $ .

Figure 9 The set of 3 Wang tiles introduced in [Reference Wang62] using letters $\{A,B,C,D,E\}$ instead of numbers from the set $\{1,2,3,4,5\}$ for labeling the edges. Each tile is identified uniquely by an index from the set $\{0,1,2\}$ written at the center each tile.

Figure 10 A finite $3\times 3$ pattern on the left is valid with respect to the Wang tiles since it respects Equations (3.3) and (3.4). Validity can be verified on the tiling shown on the right.

Let $\mathcal {T}=\{t_0,\dots ,t_{m-1}\}$ be a set of Wang tiles such as the one shown in Figure 9. A configuration $x:\mathbb {Z}^2\to \{0,\dots ,m-1\}$ is valid with respect to $\mathcal {T}$ if it assigns a tile in $\mathcal {T}$ to each position of $\mathbb {Z}^2$ so that contiguous edges of adjacent tiles have the same color, that is,

(3.3)
(3.4)

for every ${\boldsymbol{n}}\in \mathbb {Z}^2$ where ${\boldsymbol{e}}_1=(1,0)$ and ${\boldsymbol{e}}_2=(0,1)$ . A finite pattern which is valid with respect to $\mathcal {U}$ is shown in Figure 10.

Let $\Omega _{\mathcal {T}}\subset \{0,\dots ,m-1\}^{\mathbb {Z}^2}$ denote the set of all valid configurations with respect to $\mathcal {T}$ . Together with the shift action $\sigma $ of $\mathbb {Z}^2$ , $\Omega _{\mathcal {T}}$ is a subshift that we call a Wang shift. Furthermore, $\Omega _{\mathcal {T}}$ is a subshift of finite type (SFT) of the form (3.2) since $\Omega _{\mathcal {T}}$ is the subshift defined from the finite set of forbidden patterns made of all horizontal and vertical dominoes of two tiles that do not share an edge of the same color. Reciprocally, every subshift of finite type can be encoded into a Wang shift following a well-known construction (see [Reference Mozes41, p. 141-142]).

To a configuration $x\in \Omega _{\mathcal {T}}$ corresponds a tiling of the plane $\mathbb {R}^2$ by the tiles $\mathcal {T}$ where the unit square Wang tile $t_{x({\boldsymbol{n}})}$ is placed at position ${\boldsymbol{n}}$ for every ${\boldsymbol{n}}\in \mathbb {Z}^2$ , as in Figure 10. In this article, we consider tilings from the symbolic point of view. In particular, we represent tilings of plane by Wang tiles symbolically by configurations $\mathbb {Z}^2\to \mathcal {T}$ .

A configuration $x\in \Omega _{\mathcal {T}}$ is periodic if there exists ${\boldsymbol{n}}\in \mathbb {Z}^2\setminus \{0\}$ such that $x=\sigma ^{\boldsymbol{n}}(x)$ . A set of Wang tiles $\mathcal {T}$ is periodic if there exists a periodic configuration $x\in \Omega _{\mathcal {T}}$ . Originally, Wang thought that every set of Wang tiles $\mathcal {T}$ is periodic as soon as $\Omega _{\mathcal {T}}$ is nonempty [Reference Wang62]. This statement is equivalent to the existence of an algorithm solving the domino problem, that is, taking as input a set of Wang tiles and returning yes or no whether there exists a valid configuration with these tiles. Berger, a student of Wang, later proved that the domino problem is undecidable and he also provided a first example of an aperiodic set of Wang tiles [Reference Berger7]. A set of Wang tiles $\mathcal {T}$ is aperiodic if the Wang shift $\Omega _{\mathcal {T}}$ is a nonempty aperiodic subshift. This means that in general one cannot decide the emptiness of a Wang shift $\Omega _{\mathcal {T}}$ .

4 The family of metallic mean Wang tiles

In this section, we recall from [Reference Labbé37] the definition of the set $\mathcal {T}_n$ of metallic mean Wang tiles and the extended set $\mathcal {T}_n'$ which satisfies $\mathcal {T}_n\subset \mathcal {T}_n'$ . The extended set $\mathcal {T}_n'$ was used to prove the self-similarity of the Wang shift $\Omega _n$ defined over $\mathcal {T}_n$ .

For every integer $n\in \mathbb {Z}$ , we write $\overline {n}$ to denote $n+1$ and $\underline {n}$ to denote $n-1$ :

$$ \begin{align*} \overline{n} &:= n+1,\\ \underline{n} &:= n-1. \end{align*} $$

For every Wang tile $\tau =(a,b,c,d)$ , we define its symmetric image under a symmetry by the positive diagonal as $\widehat {\tau }=(b,a,d,c)$ :

4.1 The tiles

For every integer $n\geq 1$ , let

$$\begin{align*}V_n = \{(v_0,v_1,v_2)\in\mathbb{Z}^3\colon 0\leq v_0\leq v_1\leq 1\text{ and }v_1\leq v_2\leq n+1\}. \end{align*}$$

be a set of vectors having nondecreasing entries with an upper bound of 1 on the middle entry and an upper bound of $n+1$ on the last entry. The label of the edges of the Wang tiles considered in this article are vectors in $V_n$ . To lighten the figures and the presentation of the Wang tiles, it is convenient to denote the vector $(v_0,v_1,v_2)\in V_n$ more compactly as a word $v_0v_1v_2$ . For instance the vector $(1,1,1)$ is represented as $111$ .

To help the reading of the tiles and tilings, we assign a color to the vectors according to the following rule: a vector $v\in 00\mathbb {N}$ is drawn in blue, a vector $v\in 01\mathbb {N}$ is drawn in yellow and a vector $v\in 11\mathbb {N}$ is drawn in white. Overlap between blue and yellow regions will be shown in green.

For every integer $n\geq 1$ and for every $i,j\in \mathbb {N}$ such that $0\leq i\leq n$ and $0\leq j\leq n$ , we have the following white tiles:

For every $i,n\in \mathbb {N}$ such that $0\leq i\leq n$ , we have the following blue, yellow, green and antigreen tiles:

For every $n\in \mathbb {N}$ and $k,\ell ,r,s\in \{0,1\}$ such that $k\leq \ell $ and $r\leq s$ , we have the following junction tiles (the gray region will be drawn in a blue or yellow color depending on the specific values of $k,\ell ,r,s$ according to the same rule as above):

Junction tiles play a similar role as junction tiles in [Reference Mozes41].

4.2 The extended set $\mathcal {T}_n'$ of metallic mean Wang tiles

In this section, we give the definition of the family of extended sets of Wang tiles $(\mathcal {T}_n')_{n\geq 1}$ .

From the above, we define the following sets of tiles:

$$ \begin{align*} W_n &= \left\{ w_n^{i,j} \mid 1\leq i\leq n, 1\leq j\leq n \right\} &&(n^2\text{ white tiles}),\\ B^{\prime}_n &= \left\{ b_n^i \mid 0\leq i \leq n \right\} &&(n+1\text{ horizontal blue tiles}),\\ Y_n &= \left\{ y_n^i \mid 1\leq i\leq n \right\} &&(n\text{ horizontal yellow tiles}),\\ G_n &= \left\{ g_n^i\mid 0\leq i\leq n \right\} &&(n+1\text{ horizontal green tiles}),\\ A_n &= \left\{ a_n^i \mid 1\leq i \leq n \right\} &&(n \text{ horizontal antigreen tiles}). \end{align*} $$

Finally, we have a set of 9 junction tiles:

We may observe that $\widehat {W_n}=W_n$ and $\widehat {J^{\prime }_n}=J^{\prime }_n$ are closed under reflection. Also, $\widehat {B^{\prime }_n}$ are $n+1$ vertical blue tiles, $\widehat {Y_n}$ are n vertical yellow tiles, $\widehat {G_n}$ are $n+1$ vertical green tiles and $\widehat {A_n}$ are n vertical antigreen tiles.

The extended set of metallic mean Wang tiles $\mathcal {T}_n'$ can be described in terms of the white, yellow, green, blue, antigreen and junction tiles seen before.

Definition 4.1 (Extended set of metallic mean Wang tiles [Reference Labbé37])

Let

$$\begin{align*}\mathcal{T}_n'= W_n\cup Y_n \cup \widehat{Y_n} \cup G_n \cup \widehat{G_n} \cup B^{\prime}_n \cup \widehat{B^{\prime}_n} \cup A_n \cup \widehat{A_n} \cup J^{\prime}_n. \end{align*}$$

The set $\mathcal {T}_n'$ defines the extended metallic mean Wang shift $\Omega ^{\prime }_n=\Omega _{\mathcal {T}_n'}$ .

The set $\mathcal {T}_n'$ contains $n^2+2(n+1+n+n+1+n)+9=n^2+8n+13$ Wang tiles. The set of Wang tiles $\mathcal {T}_n'$ for $n=4$ is shown in Figure 11.

Figure 11 Extended metallic mean Wang tile sets $\mathcal {T}_n'$ for $n=4$ . The junction tiles $j_n^{0,0,1,1}$ and $j_n^{1,1,0,0}$ are shown with a $\times $ -mark in their center.

4.3 The family $\mathcal {T}_n$ of $(n+3)^2$ Wang tiles

In this section, we give the definition of the family of sets of Wang tiles $(\mathcal {T}_n)_{n\geq 1}$ . The set $\mathcal {T}_n$ is a subset of $\mathcal {T}_n'$ defined as follows. Let

$$ \begin{align*} B_n &= B_n' \setminus \left\{ b_n^n\right\} &&(\text{subset of}\ n \text{ horizontal blue tiles}),\\ J_n &= J_n' \setminus \left\{ j_n^{1,1,0,0}, j_n^{0,0,1,1} \right\} &&(\text{subset of 7 junction tiles}). \end{align*} $$

Definition 4.2 (Metallic mean Wang tiles[Reference Labbé37])

For every positive integer n, we construct the set of Wang tiles

$$\begin{align*}\mathcal{T}_n= W_n\cup Y_n \cup \widehat{Y_n} \cup G_n \cup \widehat{G_n} \cup B_n \cup \widehat{B_n} \cup J_n. \end{align*}$$

The set of tiles defines the Metallic mean Wang shift $\Omega _n=\Omega _{\mathcal {T}_n}$ .

The subset $\mathcal {T}_n$ contains $n^2+2(n+n+1+n)+7=(n+3)^2$ Wang tiles. They are shown in Figure 12 for $n=1,2,3,4,5$ .

Figure 12 Metallic mean Wang tile sets $\mathcal {T}_n$ for $n=1,2,3,4,5$ .

5 The $\theta _n$ -chip and metallic mean Wang tiles

In this section, we relate the $\theta _n$ -chip with metallic mean Wang tiles. The proposition below provides an independent characterization of the extended set $\mathcal {T}_n'$ of metallic mean Wang tiles as instances of the $\theta _n$ -chip, see Equation 2.2.

Proposition 5.1. For every $n\geq 1$ , the set of instances of the computer chip is equal to the extended set of metallic mean Wang tiles, that is, $\mathcal {C}_n=\mathcal {T}_n'$ .

Proof. ( $\subseteq $ ) Let be a Wang tile such that $u=(u_0,u_1,u_2)\in V_n$ , $v=(v_0,v_1,v_2)\in V_n$ , $\theta _n(u,v)\in V_n$ and $\theta _n(v,u)\in V_n$ . We proceed case by case:

  • If $u_0=1$ and $v_0=1$ , then $1=u_1\leq u_2$ , $1=v_1\leq v_2$ and

    $$ \begin{align*} \theta_n(u,v)&=(u_0,1,u_2+1)=(1,1,u_2+1)\in V_n,\\ \theta_n(v,u)&=(v_0,1,v_2+1)=(1,1,v_2+1)\in V_n. \end{align*} $$

    Thus, $0\leq u_2\leq n$ and $0\leq v_2\leq n$ and $\tau \in W_n$ is a white tile.

  • If $u_0=0$ and $v_0=1$ , then

    $$ \begin{align*} \theta_n(u,v)&=(u_0,v_2-n,u_2+1)=(0,v_2-n,u_2+1)\in V_n,\\ \theta_n(v,u)&=(v_0,1,u_1+v_0)=(1,1,u_1+1)\in V_n, \end{align*} $$

    where $0\leq u_2\leq n$ , $n\leq v_2\leq n+1$ and $0\leq u_1\leq 1$ . There are four possibilities according to the values of $v_2\in \{n,n+1\}$ and $u_1\in \{0,1\}$ that we consider case by case:

    • If $v_2=n$ and $u_1=0$ , then is a blue horizontal stripe tile with $0\leq u_2\leq n$ .

    • If $v_2=n$ and $u_1=1$ , then is an antigreen horizontal tile with $1\leq u_2\leq n$ .

    • If $v_2=n+1$ and $u_1=0$ , then is a green horizontal overlap tile with $0\leq u_2\leq n$ .

    • If $v_2=n+1$ and $u_1=1$ , then is a yellow horizontal stripe tile with $1\leq u_2\leq n$ .

  • If $u_0=1$ and $v_0=0$ , the possibilities are the symmetric image of the previous case. Thus, $\tau \in \widehat {B_n}\cup \{\widehat {b_n^n}\} \cup \widehat {A_n}\cup \widehat {G_n}\cup \widehat {Y_n}$ is a blue, antigreen, green or yellow vertical tile.

  • If $u_0=0$ and $v_0=0$ , then

    $$ \begin{align*} \theta_n(u,v)&=(u_0,v_2-n,v_1+u_0)=(0,v_2-n,v_1)\in V_n,\\ \theta_n(v,u)&=(v_0,u_2-n,u_1+v_0)=(0,u_2-n,u_1)\in V_n, \end{align*} $$
    where $0\leq u_2-n\leq u_1\leq 1$ and $0\leq v_2-n\leq v_1\leq 1$ . In particular, $(v_2-n,v_1),(u_2-n,u_1)\in \{(0,0),(0,1),(1,1)\}$ . In all cases, we have is a junction tile.

( $\supseteq $ ) Proving $\mathcal {C}_n\supseteq \mathcal {T}_n'$ is not necessary to conclude the proof, since $\mathcal {C}_n\subseteq \mathcal {T}_n'$ and $\mathcal {T}_n'$ is a finite set. Indeed, the set $\mathcal {T}_n'$ contains $\#\mathcal {T}_n'=n^2+8n+13$ elements. Also, in the proof that $\mathcal {C}_n\subseteq \mathcal {T}_n'$ made above, we exhibited $n^2$ white tiles, $2(n+1)$ blue tiles, $2n$ antigreen tiles, $2(n+1)$ green tiles, $2n$ yellow tiles and $9$ junction tiles in $\mathcal {C}_n$ . Therefore, $\mathcal {C}_n$ contains $n^2+2(n+1+n+n+1+n)+9=n^2+8n+13$ elements. We conclude that $\mathcal {C}_n=\mathcal {T}_n'$ .

Alternatively, $\mathcal {C}_n\supseteq \mathcal {T}_n'$ can be proved directly. One may check that for every , we have $\{r,t,\ell ,b\}\subset V_n$ , $r=\theta _n(\ell ,b)$ and $t=\theta _n(b,\ell )$ . Thus, $\tau \in \mathcal {C}_n$ .

We may now prove the first main result.

Theorem A

For every integer $n\geq 1$ , the Wang shift $\Omega _{\mathcal {C}_n}$ defined by the $\theta _n$ -chip is the $n^{th}$ metallic mean Wang shift $\Omega _n$ .

Proof. From Proposition 5.1, we have $\mathcal {C}_n=\mathcal {T}_n'$ . It was shown in [Reference Labbé37] that the tiles in the difference set $\mathcal {T}_n'\setminus \mathcal {T}_n$ do not appear in valid configurations of $\Omega _{\mathcal {T}_n'}$ , so that $\Omega _{\mathcal {T}_n'}=\Omega _{\mathcal {T}_n}$ . Thus, we conclude the equalities

$$\begin{align*}\Omega_{\mathcal{C}_n} = \Omega_{\mathcal{T}_n'} = \Omega_{\mathcal{T}_n} = \Omega_n. \end{align*}$$

Now, we show that the computation performed by $\theta _n$ is invertible. Let

$$\begin{align*}\begin{array}{rccl} \psi_n:&V_n\times V_n & \to & \mathbb{Z}^3\\ &(r_0,r_1,r_2), (t_0,t_1,t_2) & \mapsto & (\ell_0,\ell_1,\ell_2), \end{array} \end{align*}$$

be the function defined by

(5.1) $$ \begin{align} \left\{ \begin{array}{ll} &\ell_0=r_0,\\ &\ell_1=\begin{cases} t_2-t_0 & \text{ if } r_0 = 0,\\ 1 & \text{ if } r_0 = 1, \end{cases}\\ &\ell_2=\begin{cases} t_1+n & \text{ if } t_0 = 0,\\ r_2-1 & \text{ if } t_0 = 1. \end{cases} \end{array} \right. \end{align} $$

The following proposition states that the south and west colors of tiles in $\mathcal {C}_n$ can be deduced from the right and top colors using the map $\psi _n$ .

Proposition 5.2. We have

(5.2)

Proof. Let $\ell ,b\in V_n$ and suppose that $r=(r_0,r_1,r_2)=\theta _n(\ell ,b)$ and $t=(t_0,t_1,t_2)=\theta _n(b,\ell )$ . From Equation (2.1), we have

(5.3) $$ \begin{align} \left\{ \begin{array}{ll} &r_0=\ell_0,\\ &r_1=\begin{cases} b_2-n & \text{ if } \ell_0 = 0,\\ 1 & \text{ if } \ell_0 = 1, \end{cases}\\ &r_2=\begin{cases} b_1+\ell_0 & \text{ if } b_0 = 0,\\ \ell_2+1 & \text{ if } b_0 = 1, \end{cases} \end{array} \right. \quad \text{and } \quad \left\{ \begin{array}{ll} &t_0=b_0,\\ &t_1=\begin{cases} \ell_2-n & \text{ if } b_0 = 0,\\ 1 & \text{ if } b_0 = 1, \end{cases}\\ &t_2=\begin{cases} \ell_1+b_0 & \text{ if } \ell_0 = 0,\\ b_2+1 & \text{ if } \ell_0 = 1. \end{cases} \end{array} \right. \end{align} $$

The above holds if and only if

$$\begin{align*}\left\{ \begin{array}{ll} &\ell_0=r_0,\\ &\ell_1=\begin{cases} t_2-t_0 & \text{ if } r_0 = 0,\\ 1 & \text{ if } r_0 = 1, \end{cases}\\ &\ell_2=\begin{cases} t_1+n & \text{ if } t_0 = 0,\\ r_2-1 & \text{ if } t_0 = 1, \end{cases} \end{array} \right. \quad \text{and } \quad \left\{ \begin{array}{ll} &b_0=t_0,\\ &b_1=\begin{cases} r_2-r_0 & \text{ if } t_0 = 0,\\ 1 & \text{ if } t_0 = 1, \end{cases}\\ &b_2=\begin{cases} r_1+n & \text{ if } r_0 = 0,\\ t_2-1 & \text{ if } r_0 = 1. \end{cases} \end{array} \right. \end{align*}$$

if and only if $\ell =(\ell _0,\ell _1,\ell _2)=\psi _n(r,t)$ and $b=(b_0,b_1,b_2)=\psi _n(t,r)$ . Thus, from Equation (2.2), we have

As a consequence of Proposition 5.2, there is a bijection between the south-west and the north-east colors for the tiles in $\mathcal {C}_n$ . Using the vocabulary of [Reference Kari and Papasoglu28], we may state the following result. A set $\mathcal {T}$ of Wang tiles is called SW-deterministic if there do not exist two different tiles in $\mathcal {T}$ that would have same colors on their bottom and left edges, respectively. In other words, for all colors $C_1$ and $C_2$ there exists at most one tile in $\mathcal {T}$ whose bottom and left edges have colors $C_1$ and $C_2$ , respectively. NW-, NE- and SE-deterministic sets of Wang tiles are defined analogously. Thus, we obtain a conceptual proof for a result already obtained in [Reference Labbé37].

Theorem 5.3 [Reference Labbé37, Lemma 4.3]

For every integer $n\geq 1$ , the set of Wang tiles $\mathcal {C}_n$ is NE-deterministic and SW-deterministic.

Proof. The set of Wang tile $\mathcal {C}_n$ is SW-deterministic by definition and NE-deterministic from Proposition 5.2.

6 Equations satisfied by the Wang tiles and their tilings

In this section, we show that the set $\mathcal {C}_n$ of Wang tiles satisfy a system of equations. Moreover, we show that the rectangular tilings (of sizes $h\times 1$ , $\infty \times 1$ and $h\times k$ ) generated by them satisfy equations. While the equations associated with Kari’s [Reference Kari24] and Culik’s [Reference Culik11] aperiodic sets of Wang tiles are multiplicative, the ones associated with ${\mathcal {C}}_{n}$ are additive.

In the next theorem, we show that tiles in $\mathcal {C}_n$ satisfy $\ell _0=r_0$ , $b_0=t_0$ and the equation

$$\begin{align*}\frac{t_2-t_1+\ell_2-\ell_1}{n}-\ell_0 = \frac{b_2-b_1+r_2-r_1}{n}-b_0 \end{align*}$$

which reminds of Equation (1.1).

Theorem B

Let $n\geq 1$ be an integer, $d=(0,-1,1)$ and $e=(1,0,0)$ . The set of Wang tiles defined by the $\theta _n$ -chip satisfy the following system of equations:

where $\langle \_,\_\rangle $ denotes the canonical inner product of $\kern1pt{\mathbb {Z}^{3}}$ .

Proof. Let $\ell =(\ell _0,\ell _1,\ell _2)$ , $b=(b_0,b_1,b_2)$ , $r=(r_0,r_1,r_2)$ and $t=(t_0,t_1,t_2)$ . We always have $r_0=\ell _0$ and $t_0=b_0$ . Thus, $\langle e, \ell \rangle =\ell _0=r_0=\langle e, r\rangle $ and $\langle e, b\rangle =b_0 =t_0=\langle e, t\rangle $ . Moreover,

$$ \begin{align*} \langle d, b\rangle &= b_2-b_1,\\ \langle d, \ell\rangle &= \ell_2-\ell_1. \end{align*} $$

The proof of the remaining equality is split in four cases. We use Equation (5.3) in the computations below.

  • If $(b_0,\ell _0)=(0,0)$ , then

    $$ \begin{align*} \langle d, t+\ell\rangle &= (t_2-t_1) + (\ell_2-\ell_1) = (\ell_1+b_0)-(\ell_2-n) + (\ell_2-\ell_1) = b_0 + n = n\\ \langle d, r+b\rangle &= (r_2-r_1) + (b_2-b_1) =(b_1+\ell_0)-(b_2-n) + (b_2-b_1) = \ell_0 + n = n\\ n\langle e, \ell-b\rangle &= n(\ell_0-b_0) = 0 \end{align*} $$
  • If $(b_0,\ell _0)=(0,1)$ , then $\ell _1=1$ and

    $$ \begin{align*} \langle d, t+\ell\rangle &= (t_2-t_1) + (\ell_2-\ell_1) =(b_2+1)-(\ell_2-n) + (\ell_2-\ell_1) = b_2 + n\\ \langle d, r+b\rangle &= (r_2-r_1) + (b_2-b_1) =(b_1+\ell_0)-(1) + (b_2-b_1) = b_2\\ n\langle e, \ell-b\rangle &= n(\ell_0-b_0) = n \end{align*} $$
  • If $(b_0,\ell _0)=(1,0)$ , then $b_1=1$ and

    $$ \begin{align*} \langle d, t+\ell\rangle &= (t_2-t_1) + (\ell_2-\ell_1) =(\ell_1+b_0)-(1) + (\ell_2-\ell_1) = \ell_2\\ \langle d, r+b\rangle &= (r_2-r_1) + (b_2-b_1) =(\ell_2+1)-(b_2-n) + (b_2-b_1) = \ell_2 + n\\ n\langle e, \ell-b\rangle &= n(\ell_0-b_0) = -n \end{align*} $$
  • If $(b_0,\ell _0)=(1,1)$ , then $b_1=\ell _1=1$ and

    $$ \begin{align*} \langle d, t+\ell\rangle &= (t_2-t_1) + (\ell_2-\ell_1) =(b_2+1)-(1) + (\ell_2-\ell_1) = b_2 + \ell_2 - \ell_1\\ \langle d, r+b\rangle &= (r_2-r_1) + (b_2-b_1) =(\ell_2+1)-(1)+ (b_2-b_1) = \ell_2 + b_2 - b_1\\ n\langle e, \ell-b\rangle &= n(\ell_0-b_0) = 0 \end{align*} $$

In all the four cases, we have $\langle d, t+\ell \rangle =\langle d, r+b\rangle +n\langle e, \ell -b\rangle $ .

The two sets in the statement of Theorem B are not equal. For instance satisfy the equations when $n=4$ , but it is not a tile in $\mathcal {C}_n$ .

Equation (1.1) behaves well with valid tiling of an horizontal strip by Wang tiles associated with the same multiplication factor $q\in \mathbb {Q}$ . The same holds with tiles in $\mathcal {C}_n$ which are related to some addition of a certain value modulo 1.

The equation satisfied by the tiles proved in Theorem B extends to an equation for $h\times k$ rectangular valid tilings.

Lemma 6.1. Let $n,h,k\geq 1$ be integers and $d=(0,-1,1)$ and $e=(1,0,0)$ . Let

$$\begin{align*}\{(r^{(i,j)},t^{(i,j)},\ell^{(i,j)},b^{(i,j)})\}_{1\leq i\leq h,1\leq j\leq k} \end{align*}$$

be a family of tiles in $\mathcal {C}_n$ forming a valid tiling of a $h\times k$ rectangle, see Figure 13. Let

$$\begin{align*}R=\textstyle\frac{1}{k}\sum_{j=1}^k r^{(h,j)},\quad T=\textstyle\frac{1}{h}\sum_{i=1}^h t^{(i,k)},\quad L=\textstyle\frac{1}{k}\sum_{j=1}^k\ell^{(1,j)}\quad\text{ and }\quad B=\textstyle\frac{1}{h}\sum_{i=1}^h b^{(i,1)} \end{align*}$$

be the average of the right, top, left and bottom labels of the rectangular tiling. Then the following equation holds

(6.1) $$ \begin{align} \frac{1}{k}\left\langle \textstyle\frac{1}{n}d, T - B\right\rangle -\langle e,L\rangle = \frac{1}{h} \left\langle \textstyle\frac{1}{n}d, R - L \right\rangle - \langle e, B \rangle. \end{align} $$

Figure 13 An $h\times k$ rectangular tiling of tiles from $\mathcal {C}_n$ .

Proof. From Theorem B, we have $\langle e, \ell ^{(i,j)}\rangle =\langle e, r^{(i,j)}\rangle $ , $\langle e, b^{(i,j)}\rangle =\langle e, t^{(i,j)}\rangle $ and

$$\begin{align*}\langle \textstyle\frac{1}{n}d, t^{(i,j)} - b^{(i,j)}\rangle - \langle e , \ell^{(i,j)} \rangle = \langle \textstyle\frac{1}{n}d, r^{(i,j)} - \ell^{(i,j)} \rangle - \langle e , b^{(i,j)} \rangle, \end{align*}$$

for every integers i and j such that $1\leq i\leq h$ and $1\leq j\leq k$ . We have

$$ \begin{align*} \frac{1}{k}\left\langle \textstyle\frac{1}{n}d, T - B\right\rangle -\langle e,L\rangle &= \frac{1}{k} \left\langle \textstyle\frac{1}{n}d, \frac{1}{h}\sum_{i=1}^ht^{(i,k)} - \frac{1}{h}\sum_{i=1}^hb^{(i,1)}\right\rangle -\langle e, \textstyle\frac{1}{k}\sum_{j=1}^k\ell^{(1,j)} \rangle\\ &= \frac{1}{kh}\sum_{i=1}^h \left\langle \textstyle\frac{1}{n}d, t^{(i,k)} - b^{(i,1)}\right\rangle -\frac{1}{k}\sum_{j=1}^k\langle e, \ell^{(1,j)} \rangle\\ &= \frac{1}{kh}\sum_{i=1}^h \left\langle \textstyle\frac{1}{n}d, \sum_{j=1}^k t^{(i,j)} - \sum_{j=1}^k b^{(i,j)} \right\rangle -\frac{1}{k}\sum_{j=1}^k\langle e, \textstyle\frac{1}{h}\sum_{i=1}^h\ell^{(i,j)} \rangle\\ &= \frac{1}{kh}\sum_{i=1}^h\sum_{j=1}^k \left( \left\langle \textstyle\frac{1}{n}d, t^{(i,j)} - b^{(i,j)} \right\rangle -\langle e, \ell^{(i,j)} \rangle\right)\\ &= \frac{1}{kh}\sum_{i=1}^h\sum_{j=1}^k \left( \left\langle \textstyle\frac{1}{n}d, r^{(i,j)} - \ell^{(i,j)} \right\rangle - \left\langle e, b^{(i,j)} \right\rangle\right) \\ &= \frac{1}{kh}\sum_{j=1}^k \left\langle \textstyle\frac{1}{n}d, \sum_{i=1}^hr^{(i,j)} - \sum_{i=1}^h\ell^{(i,j)} \right\rangle - \frac{1}{h}\sum_{i=1}^h \left\langle e, \textstyle\frac{1}{k}\sum_{j=1}^kb^{(i,j)} \right\rangle \\ &= \frac{1}{kh}\sum_{j=1}^k \left\langle \textstyle\frac{1}{n}d, r^{(h,j)} - \ell^{(1,j)} \right\rangle - \frac{1}{h}\sum_{i=1}^h \left\langle e, b^{(i,1)} \right\rangle \\ &= \frac{1}{h} \left\langle \textstyle\frac{1}{n}d, \frac{1}{k}\sum_{j=1}^kr^{(h,j)} - \frac{1}{k}\sum_{j=1}^k\ell^{(1,j)} \right\rangle - \left\langle e, \textstyle\frac{1}{h}\sum_{i=1}^h b^{(i,1)} \right\rangle \\ &= \frac{1}{h} \left\langle \textstyle\frac{1}{n}d, R - L \right\rangle - \left\langle e, B \right\rangle.\\[-44pt] \end{align*} $$

Equation (6.1) is a simple consequence of the equations satisfied by the tiles, but it has important implications. If $L=R$ , then $\left \langle \textstyle \frac {1}{n}d, R - L \right \rangle =0$ and $k\langle e,L\rangle $ is an integer. Thus, the average of the inner product with $\textstyle \frac {1}{n}d$ of the top labels is obtained from the average of the inner product with $\textstyle \frac {1}{n}d$ of the bottom labels by k rotations on the unit circle by a fixed angle:

(6.2) $$ \begin{align} \langle\textstyle\frac{1}{n} d,T\rangle = \langle\textstyle\frac{1}{n} d,B\rangle - k\langle e,B\rangle \quad\pmod 1. \end{align} $$

If $\Omega _n$ admits a periodic tiling, then there exists an $h\times k$ rectangular tiling of tiles from $\mathcal {C}_n$ such that $L=R$ and $B=T$ . From Equation (6.1), we get that $\left \langle e,L \right \rangle =\left \langle e,B \right \rangle $ . This equation means that the frequency of rows with no junction tiles is equal to the frequency of columns with no junction tiles. This holds if and only if h times the number of rows with no junction tile is equal to k times the number of columns with no junction tiles. Copies of the $h\times k$ rectangular tiling can be used to tile periodically a $hk\times hk$ square respecting all matching rules containing as many rows with no junction tile as columns with no junction tile. But this is not sufficient to prove that no periodic tiling exist.

Kari’s [Reference Kari24] and Culik’s [Reference Culik11] equations allow to show in a few lines that their sets of Wang tiles admit no periodic tiling. Proving the same for $\Omega _n$ directly from the equations remains an open question.

7 Valid tilings obtained from floors of linear forms

In this section, we present a method to construct valid tilings in $\Omega _n$ . It is based on the integer-floor value of three specific linear form over two variables.

Let $n\geq 1$ be an integer and let $\beta $ be the positive root of $x^2- nx-1$ . We denote the negative root by $\beta ^*$ which satisfies $\beta \beta ^*=-1$ and $\beta +\beta ^*=n$ . We consider the matrix

$$\begin{align*}M_n=\left(\begin{array}{cc} 0 & 1 \\ \beta^{-1} & 1 \\ \beta & 1 \end{array}\right) \end{align*}$$

and the map $\lambda _n:\mathbb {R}^2\to \mathbb {R}^3$ defined by

$$\begin{align*}\lambda_n(x,y)=M_n\cdot \left(\begin{array}{cc} \{x\} \\\{y\} \end{array}\right) + \left(\begin{array}{cc} \beta^*+1 \\ \beta^*+1 \\ \beta^*+1 \end{array}\right) \end{align*}$$

where $\{x\}=x-\lfloor x\rfloor $ is the fractional part of x. Since $\lambda _n(x,y)=\lambda _n(x+1,y)=\lambda _n(x,y+1)$ , it is also well-defined on the torus $\lambda _n:\mathbb {T}^2\to \mathbb {R}^3$ . Then, we define a coding function $\Lambda _n$ as the coordinate-wise floor of $\lambda _n$ when restricted to the domain $[0,1)^2$ . More precisely, we have

$$\begin{align*}\begin{array}{rcl} \Lambda_n:[0,1)^2 & \to & \mathbb{Z}^3\\ (x,y) & \mapsto & \left( \begin{array}{r} \lfloor y+\beta^* +1\rfloor\\ \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ \lfloor \beta x + y+\beta^* +1\rfloor \end{array} \right), \end{array} \end{align*}$$

see Figure 14.

Figure 14 The preimage sets of the map $(x,y)\mapsto \Lambda _n(x,y)$ defines a partition of $[0,1)^2$ which is the refinement of the three partitions on the left. The above images are when $n=3$ .

Recall that, for every integer $n\geq 1$ , we have

$$\begin{align*}V_n = \{(v_0,v_1,v_2)\in\mathbb{Z}^3\colon 0\leq v_0\leq v_1\leq v_2\leq n+1\text{ and } v_1\leq 1 \}. \end{align*}$$

Lemma 7.1. For every $(x,y)\in [0,1)^2$ , $\Lambda _n(x,y)\in V_n$ .

Proof. Let $(x,y)\in [0,1)^2$ . Since $\beta>1$ , we have

$$\begin{align*}0 < \beta^*+1 \leq y+\beta^*+1 \leq \beta^{-1}x + y+\beta^*+1 \leq \beta x + y+\beta^*+1 < \beta + 1+\beta^*+1 = n + 2. \end{align*}$$

Thus, taking the floor function, we obtain

$$\begin{align*}0 \leq \lfloor \beta^* +1\rfloor \leq \lfloor y+\beta^*+1\rfloor \leq \lfloor \beta^{-1}x + y+\beta^*+1\rfloor \leq \lfloor \beta x + y+\beta^*+1\rfloor < n+2. \end{align*}$$

Therefore, if $(v_0,v_1,v_2)=\Lambda _n(x,y)$ , we have $0\leq v_0\leq v_1\leq v_2\leq n+1$ . Also

$$\begin{align*}\beta^{-1}x + y+\beta^*+1 <\beta^{-1} + 1+\beta^*+1 =1+1= 2. \end{align*}$$

Thus,

$$\begin{align*}v_1 = \lfloor \beta^{-1}x + y+\beta^*+1\rfloor \leq1. \end{align*}$$

We conclude $\Lambda _n(x,y)=(v_0,v_1,v_2)\in V_n$ .

The following lemma shows a relation between $\Lambda _n$ and the map $\theta _n$ defined in Equation (2.1).

Lemma 7.2. If $x,y\in [0,1)$ , then

$$\begin{align*}\Lambda_n(x,y) =\theta_n\big( \Lambda_n(\{x+\beta^*\},y), \Lambda_n(\{y+\beta^*\},x) \big). \end{align*}$$

Proof. Let $x,y\in [0,1)$ . We want to show that if $\ell _0, \ell _1, \ell _2,b_0,b_1,b_2\in \mathbb {Z}$ are such that

$$\begin{align*}\Lambda_n(\{x+\beta^*\},y) = \left( \begin{array}{r} \lfloor y+\beta^*+1\rfloor\\ \lfloor \beta^{-1}\{x+\beta^*\} + y+\beta^*+1\rfloor\\ \lfloor \beta \{x+\beta^*\} + y+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{r} \ell_0\\ \ell_1\\ \ell_2 \end{array} \right) \end{align*}$$

and

$$\begin{align*}\Lambda_n(\{y+\beta^*\},x) = \left( \begin{array}{r} \lfloor x+\beta^*+1\rfloor\\ \lfloor \beta^{-1}\{y+\beta^*\} + x+\beta^*+1\rfloor\\ \lfloor \beta \{y+\beta^*\} + x+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{r} b_0\\ b_1\\ b_2 \end{array} \right), \end{align*}$$

then $\Lambda _n(x,y)=\theta _n\left ((\ell _0,\ell _1,\ell _2),(b_0,b_1,b_2)\right )$ . Let $r_0,r_1,r_2\in \mathbb {Z}$ be such that

$$\begin{align*}\Lambda_n(x,y)= \left( \begin{array}{r} \lfloor y+\beta^*+1\rfloor\\ \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ \lfloor \beta x + y+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{r} r_0\\ r_1\\ r_2 \end{array} \right). \end{align*}$$

We want to show that the variables satisfy the definition of the function $\theta _n$ given in Equation (2.1). We have $r_0= \lfloor y+\beta ^*+1\rfloor =\ell _0$ . Therefore, the first equation defining the map $\theta _n$ is satisfied.

Assume that $\ell _0=\lfloor y+\beta ^*+1\rfloor =0$ . Then $-\beta ^{-1}=\beta ^*\leq y+\beta ^*<0$ . Also $0\leq \beta ^{-1}x <\beta ^{-1}$ . Thus, $-\beta ^{-1}\leq \beta ^{-1}x + y+\beta ^*<\beta ^{-1}$ . We have

$$ \begin{align*} r_1 &=\lfloor \beta^{-1}x + y+\beta^*\rfloor+1\\ &=\lfloor \beta(\beta^{-1}x + y+\beta^*)\rfloor +1 &(\text{because } {-}\beta^{-1}\leq\beta^{-1}x + y+\beta^*<\beta^{-1})\\ &=\lfloor \beta (y+\beta^*) + x\rfloor +1 \\ &=\lfloor \beta (y+\beta^*+1) + x+\beta^*\rfloor+1-n &(\text{because }\beta+\beta^*=n)\\ &=\lfloor \beta \{y+\beta^*\} + x+\beta^*\rfloor+1-n\\ &=b_2 -n \end{align*} $$

Assume that $\ell _0=\lfloor y+\beta ^*+1\rfloor =1$ . Then $0\leq y+\beta ^*< 1$ . Also, we have $y<1$ , so that $y+\beta ^*< 1+\beta ^*$ . Moreover, $0\leq \beta ^{-1}x<\beta ^{-1}$ . Thus, $0< \beta ^{-1}x + y+\beta ^*< \beta ^{-1} + 1 +\beta ^* =1$ . We have

$$\begin{align*}r_1 =\lfloor \beta^{-1}x + y+\beta^*\rfloor+1 = 0 + 1 = \ell_0. \end{align*}$$

Therefore, the second equation defining the map $\theta _n$ is satisfied.

Assume that $b_0=\lfloor x+\beta ^*+1\rfloor =0$ . This implies that $-1\leq x+\beta ^*<0$ , which implies $x<\beta ^{-1}$ . Thus, $0\leq \beta x< 1$ . We need to consider the cases $\ell _0=0$ and $\ell _0=1$ separately. First, suppose that $\ell _0=\lfloor y+\beta ^*+1\rfloor =0$ . Then $-1\leq y+\beta ^*< 0$ . Thus, $-1\leq \beta x + y+\beta ^*<1$ . We have

$$ \begin{align*} r_2 &=\lfloor \beta x + y+\beta^*+1\rfloor\\ &=\lfloor \beta^{-1}(\beta x + y+\beta^*)\rfloor +1 &(\text{because } -1\leq(\beta x + y+\beta^*)<1)\\ &=\lfloor \beta^{-1}(\beta x + y+\beta^*) +\beta^{-1}+\beta^*\rfloor +1\\ &=\lfloor \beta^{-1}(1+y+\beta^*) +x+\beta^*\rfloor +1\\ &=\lfloor \beta^{-1}\{y+\beta^*\} +x+\beta^*\rfloor +1\\ &= b_1 = b_1 + 0 = b_1+\ell_0. \end{align*} $$

Secondly, suppose that $\ell _0=\lfloor y+\beta ^*+1\rfloor =1$ . Then $0\leq y+\beta ^*< 1$ , which implies $\{y+\beta ^*\}=y+\beta ^*$ . Thus, $0\leq \beta x + y+\beta ^*<2$ . We have

$$ \begin{align*} r_2 &=\lfloor \beta x + y+\beta^*+1\rfloor\\ &=\lfloor \beta x + y+\beta^*-1\rfloor+2\\ &=\lfloor \beta^{-1}(\beta x + y+\beta^*-1) \rfloor+2 &(\text{because } -1\leq(\beta x + y+\beta^*-1)<1)\\ &=\lfloor \beta^{-1}(y+\beta^*) +x +\beta^*\rfloor+2\\ &=\lfloor \beta^{-1}\{y+\beta^*\} +x+\beta^*\rfloor+2\\ &= b_1 + 1= b_1+\ell_0. \end{align*} $$

Assume that $b_0=\lfloor x+\beta ^*+1\rfloor =1$ . This implies that $0\leq x+\beta ^*<1$ , which implies $\{x+\beta ^*\}=x+\beta ^*$ . We have

$$ \begin{align*} r_2 &=\lfloor \beta x + y+\beta^*+1\rfloor\\ &=\lfloor \beta x +\beta\beta^*+1+ y+\beta^* +1\rfloor &(\text{because }\beta\beta^*=-1)\\ &= \lfloor \beta (x+\beta^*) + y+\beta^*+1\rfloor + 1\\ &= \lfloor \beta \{x+\beta^*\} + y+\beta^*+1\rfloor + 1\\ &= \ell_2 + 1 = \ell_2+b_1. \end{align*} $$

Therefore, the third equation defining the map $\theta _n$ is satisfied.

For every $(x,y)\in \mathbb {R}^2$ , let

which can be interpreted geometrically as a Wang tile:

Lemma 7.3. If $(x,y)\in \mathbb {R}^2$ , then

  • ,

  • ,

  • is an instance of a $\theta _n$ -chip tile.

Proof. We observe that is the image of under the tile reflection $t\mapsto \widehat {t}$ by the positive slope diagonal.

From Lemma 7.1, for every $(x,y)\in [0,1)^2$ , we have $\Lambda _n(x,y)\in V_n$ . Therefore, for every $(x,y)\in \mathbb {R}^2$ ,

$$\begin{align*}\Lambda_n\left(\{x\},\{y\}\right),\quad \Lambda_n\left(\{y\},\{x\}\right),\quad \Lambda_n\left(\{x+\beta^*\},\{y\}\right),\quad \Lambda_n\left(\{y+\beta^*\},\{x\}\right) \in V_n. \end{align*}$$

From Lemma 7.2, for every $(x,y)\in \mathbb {R}^2$ , we have

$$\begin{align*}\Lambda_n(\{x\},\{y\}) =\theta_n\big( \Lambda_n(\{x+\beta^*\},\{y\}), \Lambda_n(\{y+\beta^*\},\{x\}) \big). \end{align*}$$

Also

$$\begin{align*}\Lambda_n(\{y\},\{x\}) =\theta_n\big( \Lambda_n(\{y+\beta^*\},\{x\}), \Lambda_n(\{x+\beta^*\},\{y\}) \big). \end{align*}$$

Thus, .

Here is another characterization of the set of Wang tiles $\mathcal {T}_n$ .

Proposition 7.4. The following holds:

Proof. First, recall from Proposition 5.1 that

(7.1) $$ \begin{align} \mathcal{C}_n = \mathcal{T}_n' = \mathcal{T}_n\cup \{j_n^{0,0,1,1}, j_n^{1,1,0,0}\} \cup \left\{ a_n^i, \widehat{a_n^i} \mid 1\leq i \leq n \right\} \cup \left\{ b_n^n,\widehat{b_n^n} \right\} \end{align} $$

where

Let

First we show that $U_n\subseteq \mathcal {T}_n$ . It follows from Lemma 7.3 that $ U_n \subset \mathcal {C}_n$ . Thus, using Equation (7.1), the goal is to show that

(7.2) $$ \begin{align} U_n \cap\left( \{j_n^{0,0,1,1}, j_n^{1,1,0,0}\} \cup \left\{ a_n^i, \widehat{a_n^i} \mid 1\leq i \leq n \right\} \cup \left\{ b_n^n,\widehat{b_n^n} \right\} \right)=\varnothing. \end{align} $$

Suppose that there exists $(x,y)\in [0,1)^2$ such that . Then $\Lambda _n(x,y)=000$ and $\Lambda _n(y,x)=011$ . More precisely, we have

$$ \begin{align*} \Lambda_n(x,y)&= \left( \begin{array}{r} \lfloor y+\beta^*+1\rfloor\\ \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ \lfloor \beta x + y+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{r} 0\\ 0\\ 0 \end{array} \right),\\ \Lambda_n(y,x)&= \left( \begin{array}{r} \lfloor x+\beta^*+1\rfloor\\ \lfloor \beta^{-1}y + x+\beta^*+1\rfloor\\ \lfloor \beta y + x+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{r} 0\\ 1\\ 1 \end{array} \right). \end{align*} $$

In particular,

$$\begin{align*}0 = \lfloor \beta x + y+\beta^*+1\rfloor \geq \lfloor \beta^{-1}y + x+\beta^*+1\rfloor =1, \end{align*}$$

which is a contradiction. The same contradiction is obtained if . Therefore, these two junction tiles are not in $U_n$ .

Suppose that there exists $(x,y)\in [0,1)^2$ such that for some integer i satisfying $1\leq i\leq n$ . Then $\Lambda _n(x,y)=00\overline {i}$ and $\Lambda _n(y,x)=112$ . More precisely, we have

$$ \begin{align*} \Lambda_n(x,y)&= \left( \begin{array}{r} \lfloor y+\beta^*+1\rfloor\\ \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ \lfloor \beta x + y+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{c} 0\\ 0\\ i+1 \end{array} \right),\\ \Lambda_n(y,x)&= \left( \begin{array}{r} \lfloor x+\beta^*+1\rfloor\\ \lfloor \beta^{-1}y + x+\beta^*+1\rfloor\\ \lfloor \beta y + x+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{c} 1\\ 1\\ 2 \end{array} \right). \end{align*} $$

In particular, $\lfloor y+\beta ^*+1\rfloor =0$ implies that $-\beta ^{-1}\leq y+\beta ^*<0$ . Also $0\leq \beta ^{-1}x<\beta ^{-1}$ , so that $-\beta ^{-1}\leq \beta ^{-1}x+y+\beta ^*<\beta ^{-1}$ . Therefore,

$$\begin{align*}0=\lfloor \beta^{-1}x + y+\beta^*+1\rfloor =\lfloor \beta(\beta^{-1}x + y+\beta^*)\rfloor+1 =\lfloor \beta y + x -1\rfloor+1 =\lfloor \beta y + x \rfloor. \end{align*}$$

On the other hand, using $\lfloor a+b\rfloor \leq \lfloor a\rfloor +\lfloor b\rfloor +1$ for every $a,b\in \mathbb {R}$ , we obtain

$$ \begin{align*} 2=\lfloor \beta y + x+\beta^*+1\rfloor \leq\lfloor \beta y + x\rfloor+\lfloor\beta^*+1\rfloor+1 = 0 + 0 + 1 = 1, \end{align*} $$

which is a contradiction. A similar contradiction is obtained if we suppose that such that . Therefore, there is no antigreen tile in $U_n$ .

Suppose that there exists $(x,y)\in [0,1)^2$ such that . Then $\Lambda _n(x,y)=00\overline {n}$ and $\Lambda _n(y,x)=111$ . More precisely, we have

$$ \begin{align*} \Lambda_n(x,y)&= \left( \begin{array}{r} \lfloor y+\beta^*+1\rfloor\\ \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ \lfloor \beta x + y+\beta^*+1\rfloor \end{array} \right) = \left( \begin{array}{c} 0\\ 0\\ n+1 \end{array} \right). \end{align*} $$

In particular, using $\beta =n+\beta ^{-1}$ and $x<1$ , we obtain

$$ \begin{align*} n+1 &=\lfloor \beta x + y+\beta^*+1\rfloor\\ &=\lfloor (n+\beta^{-1}) x + y+\beta^*+1\rfloor\\ &\leq\lfloor n+\beta^{-1} x + y+\beta^*+1\rfloor\\ &=\lfloor \beta^{-1} x + y+\beta^*+1\rfloor+n = 0 + n = n, \end{align*} $$

which is a contradiction. A similar contradiction is obtained if we suppose that such that . Therefore, the blue tiles $b_n^n$ and $\widehat {b_n^n}$ are not in $U_n$ . This shows that Equation (7.2) holds. Thus, $U_n\subseteq \mathcal {T}_n$ .

Now we show that $\mathcal {T}_n\subseteq U_n$ . We have $J_n\subset U_n$ since

We have $B_n\subset U_n$ since

We have $G_n\subset U_n$ since

We have $Y_n\subset U_n$ since

We have $W_n\subset U_n$ since

Therefore, $J_n\cup B_n \cup G_n \cup Y_n\cup W_n \subseteq U_n$ . Since $\widehat {U_n}=U_n$ , we also have $\widehat {B_n} \cup \widehat {G_n} \cup \widehat {Y_n}\subseteq U_n$ . We conclude that $\mathcal {T}_n \subseteq U_n$ and $\mathcal {T}_n = U_n$ .

This allows to construct valid configurations $\mathbb {Z}^2\to \mathcal {T}_n$ from any starting point $(x,y)$ on the torus. See Figure 15.

Figure 15 For every $(x,y)\in [0,1)^2$ the map $\mathbb {Z}^2\to \mathcal {T}_n$ defined by is a valid tiling of the plane by the set of Wang tiles $\mathcal {T}_n$ .

Theorem C

For every integer $n\geq 1$ and every $(x,y)\in [0,1)^2$ , the configuration

is a valid tiling of the plane by the set of metallic mean Wang tiles $\mathcal {T}_n$ .

Proof. Let $(x,y)\in [0,1)^2$ and $(i,j)\in \mathbb {Z}^2$ . We have $c_{(x,y)}(i,j)\in \mathcal {T}_n$ from Proposition 7.4. Also the right color of the tile $c_{(x,y)}(i,j)$ is $\Lambda _n(\{x+i\beta ^{-1}\},\{y+j\beta ^{-1}\})$ which is equal to the left color of the tile $c_{(x,y)}(i+1,j)$ . Finally, the top color of the tile $c_{(x,y)}(i,j)$ is $\Lambda _n(\{y+j\beta ^{-1}\},\{x+i\beta ^{-1}\})$ which is equal to the bottom color of the tile $c_{(x,y)}(i,j+1)$ . Therefore, $c_{(x,y)}$ is a valid configuration of Wang tiles from the set $\mathcal {T}_n$ .

The set $\{c_{(x,y)}\colon (x,y)\in [0,1)^2\}$ is not a subshift because it is not topologically closed. Indeed, if $(x_0,y_0)$ lies on the boundary of the partition, there is more than one configuration associated with it. The configuration $c_{(x_0,y_0)}$ is one of them, but $\lim _{(x,y)\to (x_0,y_0)}c_{(x,y)}$ might be a different configuration if the limit is taken coming from another direction. The same issue happens with the representation of numbers in base 10. For example, the number 1 has two base-10 representations, one being $1.000000\dots $ and the other $0.999999\dots $ .

This implies that the set $\{c_{(x,y)}\colon (x,y)\in [0,1)^2\}$ is not the set of all valid configurations of $\mathcal {T}_n$ . In other terms, $c:(x,y)\mapsto c_{(x,y)}$ is not surjective in the set $\Omega _n$ of all valid configurations of $\mathcal {T}_n$ . One way to solve this issue is to take the topological closure

$$\begin{align*}C = \overline{\left\{c_{(x,y)}\colon (x,y)\in[0,1)^2\right\}} \end{align*}$$

which is a nonempty subshift satisfying $C\subseteq \Omega _n$ . Since $\Omega _n$ is minimal [Reference Labbé37], we conclude the equality $C=\Omega _n$ must hold.

A standard approach is to create the subshift C as the symbolic extension of a dynamical system defined on the 2-torus $\mathbb {T}^2$ . This is what we do in the next two sections.

8 An explicit factor map

The goal of this section is to introduce a factor map $\Omega _n\to \mathbb {T}^2$ explicitly defined from the average of inner products of the labels of the Wang tiles in a configuration, see Equation (8.2). Then, we prove Theorem D using this explicit factor map.

First, it is convenient to make some observation on the inner product with the vector $d=(0,-1,1)$ of the tile labels. In the statement below, we use the indicator function $\mathbb {I}_{A}\colon \mathbb {R}\to \{0,1\}$ of a subset $A\subset \mathbb {R}$ defined as

$$\begin{align*}\mathbb{I}_A(x)= \begin{cases} 1 & \text{ if } x\in A,\\ 0 & \text{ if } x\notin A. \end{cases} \end{align*}$$

Lemma 8.1. Let $n\geq 1$ be an integer and $d=(0,-1,1)$ . If $x,y\in [0,1)$ , then

$$\begin{align*}\langle d, \Lambda_n(x,y)\rangle = \lfloor nx\rfloor + \mathbb{I}_{[1-\{nx\},1)}(\{\delta_{x} + y\}) \end{align*}$$

where $\delta _{x}=1-\beta ^{-1}(1-x)$ .

Proof. Let $x,y\in [0,1)$ . Observe that $\delta _{x}=1-\beta ^{-1}(1-x)=\beta ^{-1}x +\beta ^*+1$ . We have

$$ \begin{align*} \langle d, \Lambda_n(x,y)\rangle &= \lfloor \beta x + y+\beta^*+1\rfloor - \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ &= \lfloor (n+\beta^{-1}) x + y+\beta^*+1\rfloor - \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ &= \lfloor nx+\delta_{x}+y\rfloor - \lfloor \delta_{x}+y\rfloor\\ &= \left( \lfloor nx\rfloor +\lfloor \delta_{x}+y\rfloor +\lfloor \{nx\}+\{\delta_{x}+y\}\rfloor\right) - \lfloor \delta_{x}+y\rfloor\\ &= \lfloor nx\rfloor +\lfloor \{nx\}+\{\delta_{x}+y\}\rfloor\\ &= \lfloor nx\rfloor + \begin{cases} 0 & \text{ if } \{nx\}+\{\delta_{x}+y\} < 1,\\ 1 & \text{ if } \{nx\}+\{\delta_{x}+y\} \geq 1. \end{cases} \end{align*} $$

The conclusion follows.

As illustrated in Figure 7 for a finite rectangular pattern, the average of the values of $\langle \frac {1}{n}d,v\rangle $ for labels v appearing along an horizontal line can be considered for valid configurations $w:\mathbb {Z}^2\to \mathcal {T}_n$ . For some reason (in order to have the equality $\phi _n(c_{(x,y)}) = y$ in Proposition 8.2), it is convenient to consider the average of the top label of the tiles on the horizontal row passing through the origin. Assuming that the limit exists for every configuration, this leads to a map from the Wang shift to the interval $[0,1]$ defined as follows:

(8.1)

where denotes the top label of the Wang tile t.

We show in the next proposition that $\phi _n$ is well-defined and that it recovers the parameter y of a configuration $c_{(x,y)}$ .

Proposition 8.2. For every integer $n\geq 1$ , the following holds:

  1. (i) for every $(x,y)\in [0,1)^2$ , $\phi _n(c_{(x,y)}) = y$ ,

  2. (ii) $\phi _n:\Omega _n\to [0,1]$ is continuous,

  3. (iii) $\phi _n:\Omega _n\to [0,1]$ is onto,

  4. (iv) if $\beta $ denotes the positive root of the polynomial $x^2-nx-1$ , then

    $$ \begin{align*} \phi_n(\sigma^{{\boldsymbol{e}}_1}w) &= \phi_n(w),\\ \phi_n(\sigma^{{\boldsymbol{e}}_2}w) &= \phi_n(w) +\beta^{-1} \quad\pmod 1. \end{align*} $$

Proof. (i) Let $R_\alpha (x)=\{x+\alpha \}$ be the rotation by angle $\alpha $ on the interval $[0,1)$ . If $\alpha $ is irrational, then for every $x\in [0,1)$ the sequence $(R^i_\alpha (x))_{i\in \mathbb {Z}}$ is uniformly distributed modulo 1 [Reference Kuipers and Niederreiter30, Exercise 2.5]. Therefore, using Weyl’s equidistribution theorem for Riemann-integrable functions [Reference Kuipers and Niederreiter30, Corollary 1.1], for every $(x,y)\in [0,1)^2$ , we have

(ii) Now we want to show that the rule $\phi _n$ defines a continuous map $\Omega _n\to \mathbb {T}$ . Since $\Omega _n$ is minimal [Reference Labbé37], we have that the orbit ${\overline {\{c_{(0,0)}\}}^{\sigma }} =\{\sigma ^k c_{(0,0)} \mid k\in \mathbb {Z}^2\} =\{c_{\beta ^{-1}k\ \pmod {\mathbb {Z}^2}} \mid k\in \mathbb {Z}^2\}$ is a dense subset of $\Omega _n$ . Therefore, $\{c_{(x,y)} \mid x,y\in [0,1)\}$ is dense in $\Omega _n$ . Let $w\in \Omega _n$ . There exists a sequence $(x^{(\ell )},y^{(\ell )})_{l\in \mathbb {N}}$ with $x^{(\ell )},y^{(\ell )}\in [0,1)$ such that $w=\lim _{\ell \to \infty } c_{(x^{(\ell )},y^{(\ell )})}$ .

Notice that the limit $(x^{(\infty )},y^{(\infty )})=\lim _{\ell \to \infty } (x^{(\ell )},y^{(\ell )})\in [0,1]^2$ exist. This essentially follows from [Reference Labbé33, Lemma 3.4] allowing to define another factor map, see Equation (9.2). Indeed, suppose on the contrary that the sequence $(x^{(\ell )},y^{(\ell )})_{l\in \mathbb {N}}$ has two distinct accumulation points $(p_1,q_1)$ and $(p_2,q_2)$ . Recall that is a topological partition of $\mathbb {T}^2$ . Since the orbits under the $\mathbb {Z}^2$ -action $R_n$ are dense, there exists $(i,j)\in \mathbb {Z}^2$ such that and where $t_1$ and $t_2$ are two distinct tiles in $\mathcal {T}_n$ . Therefore, for sufficiently large $\ell \in \mathbb {N}$ , we have

which is a contradiction.

We split the proof according to the behavior of $\lim _{\ell \to \infty }ny^{(\ell )}$ , and more precisely if it converges to an integer and if so from above or from below (the fact that it converges from above or from below when it converges to an integer follows from the existence of the configuration w because the boundary of the topological partition contains the vertical and horizontal lines passing through integers points). We proceed as above using Weyl equidistribution theorem. We have

This shows that the rule $\phi _n$ defines a map $\Omega _n\to [0,1]$ and that this map is continuous.

(iii) If $y\in [0,1)$ , then $y= \phi _n(c_{(0,y)})$ . If $y=1$ , then $y= \phi _n(\lim _{y\to 1^-}c_{(0,y)})$ . Thus, the map $\phi _n$ is onto.

(iv) Since the map $\phi _n$ is continuous, we only need to show the equalities for a dense subset of $\Omega _n$ . Let $(x,y)\in [0,1)^2$ . We have

$$\begin{align*}\phi_n(\sigma^{{\boldsymbol{e}}_1} c_{(x,y)}) = \phi_n(c_{(\{x+\beta^{-1}\},y)}) = y = \phi_n(c_{(x,y)}). \end{align*}$$

Moreover, we have

$$\begin{align*}\phi_n(\sigma^{{\boldsymbol{e}}_2}c_{(x,y)}) = \phi_n(c_{(x,\{y+\beta^{-1}\})}) = \{y+\beta^{-1}\} = \phi_n(c_{(x,y)}) +\beta^{-1} \quad\pmod 1.\\[-40pt] \end{align*}$$

Since $\phi _n(\sigma ^{{\boldsymbol{e}}_1}w) = \phi _n(w)$ for every configuration $w\in \Omega _n$ , the factor map $\phi _n$ is far from being injective. We may improve this as follows. We use the symmetry of the tiles in $\mathcal {T}_n$ to define an involution on $\Omega _n$ . If $w\in \Omega _n$ is a configuration, then its image under a reflection by the positive diagonal is the configuration $\widehat {w}\in \Omega _n$ defined as

$$\begin{align*}\begin{array}{rccl} \widehat{w}:&\mathbb{Z}^2 & \to & \mathcal{T}_n\\ &(i,j) & \mapsto & \widehat{w_{j,i}}. \end{array} \end{align*}$$

This allows to define a map from the Wang shift to the $2$ -dimensional torus

(8.2) $$ \begin{align} \begin{array}{rccl} \Phi_n:&\Omega_n & \to & \mathbb{T}^2\\ &w & \mapsto & (\phi_n(\widehat{w}), \phi_n(w)). \end{array} \end{align} $$

The first coordinate $\phi _n(\widehat {w})$ computes the average of the inner product with d of the right-hand labels of the Wang tiles in the column containing the origin of the configuration w. We show in the next theorem that $\Phi _n$ is a factor map.

Theorem D

Let $d=(0,-1,1)$ , $n\geq 1$ be an integer and $\Omega _n$ be the $n^{th}$ metallic mean Wang shift. The map

(8.3)

is a factor map, that is, it is continuous, onto and commutes the shift $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ with the toral $\mathbb {Z}^2$ -rotation $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ by the equation $ \Phi _n\circ \sigma ^k = R_n^k\circ \Phi _n $ for every $k\in \mathbb {Z}^2$ where

$$\begin{align*}\begin{array}{rccl} R_n&:\mathbb{Z}^2\times\mathbb{T}^2 & \to & \mathbb{T}^2\\ &(k,x) & \mapsto & R_n^k(x):=x + \beta k \end{array} \end{align*}$$

and $\beta =\frac {n+\sqrt {n^2+4}}{2}$ is the $n^{th}$ metallic mean, that is, the positive root of the polynomial $x^2-nx-1$ .

Proof. From Proposition 8.2, $\phi _n$ is continuous. Thus, $\Phi _n$ is also continuous.

Let $(x,y)\in [0,1)^2$ . Using Lemma 7.3, for every $(i,j)\in \mathbb {Z}^2$ , we have

Thus, the identity $\widehat {c_{(x,y)}}=c_{(y,x)}$ holds. We obtain

$$\begin{align*}(x,y) = ( \phi_n(c_{(y,x)}), \phi_n(c_{(x,y)}) ) = ( \phi_n(\widehat{c_{(x,y)}}), \phi_n(c_{(x,y)}) ) =\Phi_n(c_{(x,y)}). \end{align*}$$

Therefore, $\Phi _n$ is onto.

Let $w\in \Omega _n$ be a configuration. Let $k=(k_1,k_2)\in \mathbb {Z}^2$ . Using Proposition 8.2, we have

$$ \begin{align*} \Phi_n\circ\sigma^k (w) &= \left(\phi_n(\widehat{\sigma^k w}), \phi_n(\sigma^k w)\right)\\ &= \left(\phi_n(\sigma^{(k_2,k_1)} \widehat{w}), \phi_n(\sigma^{(k_1,k_2)} w)\right)\\ &= \left(\phi_n(\widehat{w})+\beta^{-1}k_1, \phi_n(w)+\beta^{-1}k_2\right) \quad\pmod{\mathbb{Z}^2}\\ &= (\phi_n(\widehat{w}), \phi(w)) + \beta^{-1}(k_1,k_2) \quad\pmod{\mathbb{Z}^2}\\ &= \Phi_n(w) + \beta^{-1}k \quad\pmod{\mathbb{Z}^2}\\ &= R_n^k\circ \Phi_n (w).\\[-41pt] \end{align*} $$

Corollary 8.3. For every $n\geq 1$ , $\Omega _n$ is aperiodic.

Proof. By contradiction, suppose that $\Omega _n$ contains a periodic configuration w such that $\sigma ^k(w)=w$ for some $k\in \mathbb {Z}^2\setminus \{(0,0)\}$ . The image $\Phi _n(w)\in \mathbb {T}^2$ must be a periodic point for the $\mathbb {Z}^2$ -action $R_n$ because, using Theorem D, we have

$$\begin{align*}\Phi_n(w) =\Phi_n(\sigma^k(w)) = R_n^k(\Phi_n(w)) = R_n^k(\Phi_n(w)). \end{align*}$$

The $\mathbb {Z}^2$ -action $R_n$ has no periodic point, since the metallic mean $\beta $ is an irrational number. Thus, we must have $k=0$ , which is a contradiction. The subshift $\Omega _n$ is nonempty. Thus, $\Omega _n$ is aperiodic.

Remark 8.4. Note that Corollary 8.3 cannot be considered as a totally independent proof of aperiodicity of $\Omega _n$ . Recall that aperiodicity of $\Omega _n$ was proved in [Reference Labbé37] from the self-similarity of $\Omega _n$ . Indeed, Corollary 8.3 uses Theorem D which depends on Proposition 8.2. In the proof of Proposition 8.2, we use the minimality of $\Omega _n$ which was proved in [Reference Labbé37] and deduced from its self-similarity.

In other words, the following question remains open.

Question 8.5. Can the aperiodicity of $\Omega _n$ be proved independently of its self-similarity?

9 The factor map is an isomorphism (mod 0)

The goal of this section is to show more properties of the factor map $\Phi _n:\Omega _n\to \mathbb {T}^2$ introduced in the previous section. Based on the approach presented in [Reference Labbé33], we prove Theorem E and Theorem F.

Let $n\geq 1$ be an integer. We consider the continuous $\mathbb {Z}^2$ -action $R_n$ defined on $\mathbb {T}^2=\mathbb {R}^2/\mathbb {Z}^2$ by

$$\begin{align*}\begin{array}{rccl} R_n&:\mathbb{Z}^2\times\mathbb{T}^2 & \to & \mathbb{T}^2\\ &({\boldsymbol{n}},{\boldsymbol{x}}) & \mapsto & R_n^{\boldsymbol{n}}({\boldsymbol{x}}):={\boldsymbol{x}} + \beta{\boldsymbol{n}} \end{array} \end{align*}$$

where $\beta =\frac {n+\sqrt {n^2+4}}{2}$ is the positive root of the polynomial $x^2-nx-1$ . We say that $R_n$ is a toral $\mathbb {Z}^2$ -rotation and it defines a dynamical system that we denote $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ . In this section, we encode this dynamical system symbolically using a partition associated with the Wang tiles $\mathcal {T}_n$ .

Recall that

$$\begin{align*}\begin{array}{rcl} \Lambda_n:[0,1)^2 & \to & \mathbb{Z}^3\\ (x,y) & \mapsto & \left( \begin{array}{r} \lfloor y+\beta^* +1\rfloor\\ \lfloor \beta^{-1}x + y+\beta^*+1\rfloor\\ \lfloor \beta x + y+\beta^* +1\rfloor \end{array} \right). \end{array} \end{align*}$$

From Lemma 7.1, we have in fact that $\Lambda _n$ is a map $[0,1)^2 \to V_n$ . Therefore,

is a partition of $[0,1)^2$ . Its symmetric image is

which is another partition of $[0,1)^2$ , where $\eta :(x,y)\mapsto (y,x)$ . Also, we let

where ${\boldsymbol{e}}_1=(1,0)$ and ${\boldsymbol{e}}_2=(0,1)$ . These partitions are illustrated for $n=1,2,3,4$ in Figure 16, Figure 17, Figure 18 and Figure 19. We may observe in these figures a nice property of the partitions: is the same partition (with different indices) as (this is related to the fact that the set of Wang tiles $\mathcal {T}_n$ is both NE-deterministic and SW-deterministic, see Theorem 5.3).

Figure 16 The partitions , , and .

Figure 17 The partitions , , and .

Figure 18 The partitions , , and .

Figure 19 The partitions , , and .

We now want to construct the refined partition whose atoms are defined as follows. For each $(v_1,v_2,v_3,v_4)\in (V_n)^4$ , we define the interior of the intersection

$$\begin{align*}P_{(v_1,v_2,v_3,v_4)} = \operatorname{\mathrm{Interior}}\left( \Lambda_n^{-1}(v_1) \cap \eta\circ\Lambda_n^{-1}(v_2) \cap R^{{\boldsymbol{e}}_1}(\Lambda_n^{-1}(v_3)) \cap R^{{\boldsymbol{e}}_2}(\eta\circ\Lambda_n^{-1}(v_4))\right). \end{align*}$$

It follows from Proposition 7.4 that the quadruples $\tau $ for which $P_\tau $ has nonempty interior define a set which is equal to the set of Wang tiles $\mathcal {T}_n$ :

$$\begin{align*}\mathcal{T}_n = \left\{\tau\in (V_n)^4 \mid P_\tau \neq\varnothing\right\}. \end{align*}$$

Recall that, for some finite set A, a topological partition of a compact metric space M is a finite collection $\{P_a\}_{a\in A}$ of disjoint open sets $P_a\subset M$ such that $M = \bigcup _{a\in A} \overline {P_a}$ . Naturally, the set $\mathcal {T}_n$ defines a topological partition

$$\begin{align*}\mathcal{P}_n = \{P_\tau\}_{\tau\in\mathcal{T}_n} \end{align*}$$

of $\mathbb {R}^2/\mathbb {Z}^2$ which is the refinement of the four partitions (the right color), (the top color), (the left color) and (the bottom color).

9.1 Symbolic dynamical system $\mathcal {X}_{\mathcal {P}_n,R_n}$

We now define the symbolic dynamical system associated with the toral $\mathbb {Z}^2$ -rotation $R_n$ generated by the partition $\mathcal {P}_n$ . We adapt [Reference Lind and Marcus40] to the 2-dimensional setting as it was done in [Reference Hochman20] and [Reference Labbé33].

If $S\subset \mathbb {Z}^2$ is a finite set, we say that a pattern $w\in \mathcal {A}^S$ is allowed for $\mathcal {P}_n,R_n$ if

(9.1) $$ \begin{align} \bigcap_{{\boldsymbol{k}}\in S} R_n^{-{\boldsymbol{k}}}(P_{w_{\boldsymbol{k}}}) \neq \varnothing. \end{align} $$

Let $\mathcal {L}_{\mathcal {P}_n,R_n}$ be the collection of all allowed patterns for $\mathcal {P}_n,R_n$ . The set $\mathcal {L}_{\mathcal {P}_n,R_n}$ is the language of a subshift $\mathcal {X}_{\mathcal {P}_n,R_n}\subseteq \mathcal {A}^{\mathbb {Z}^2}$ defined as follows, see [Reference Hochman20, Prop. 9.2.4],

$$\begin{align*}\mathcal{X}_{\mathcal{P}_n,R_n} = \{x\in\mathcal{A}^{\mathbb{Z}^2} \mid \pi_S\circ\sigma^{\boldsymbol{n}}(x)\in\mathcal{L}_{\mathcal{P}_n,R_n} \text{ for every } {\boldsymbol{n}}\in\mathbb{Z}^2 \text{ and finite subset } S\subset\mathbb{Z}^2\}. \end{align*}$$

We say that $\mathcal {X}_{\mathcal {P}_n,R_n}$ is the symbolic dynamical system corresponding to $\mathcal {P}_n,R_n$ .

For each $w\in \mathcal {X}_{\mathcal {P}_n,R_n}\subset \mathcal {A}^{\mathbb {Z}^2}$ and $m\geq 0$ there is a corresponding nonempty open set

$$\begin{align*}D_m(w) = \bigcap_{\Vert{\boldsymbol{k}}\Vert\leq m} R_n^{-{\boldsymbol{k}}}(P_{w_{\boldsymbol{k}}}) \subseteq \mathbb{T}^2. \end{align*}$$

The closures $\overline {D}_m(w)$ of these sets are compact and decrease with m, so that $\overline {D}_0(w)\supseteq \overline {D}_1(w)\supseteq \overline {D}_2(w)\supseteq \dots $ . It follows that $\cap _{m=0}^{\infty }\overline {D}_m(w)\neq \varnothing $ . In order for points in $\mathcal {X}_{\mathcal {P}_n,R_n}$ to correspond to points in $\mathbb {T}^2$ , this intersection should contain only one point. This leads to the following definition. A topological partition $\mathcal {P}_n$ of $\mathbb {T}^2$ gives a symbolic representation of $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ if for every $w\in \mathcal {X}_{\mathcal {P}_n,R_n}$ the intersection $\cap _{m=0}^{\infty }\overline {D}_m(w)$ consists of exactly one point ${\boldsymbol{x}}\in \mathbb {T}^2$ . We call w a symbolic representation of ${\boldsymbol{x}}$ .

Markov partitions were originally defined for one-dimensional dynamical systems $\mathbb {Z}\overset {T}{\curvearrowright }\mathbb {T}^2$ and were extended to $\mathbb {Z}^d$ -actions by automorphisms of compact Abelian group in [Reference Einsiedler and Schmidt16]. Following [Reference Labbé33, Reference Labbé34], we use the same terminology and extend the definition proposed in [Reference Lind and Marcus40, §6.5] for dynamical systems defined by higher-dimensional actions by rotations.

Definition 9.1. A topological partition $\mathcal {P}$ of $\mathbb {T}^2$ is a Markov partition for $\mathbb {Z}^2\overset {R}{\curvearrowright }\mathbb {T}^2$ if

  • $\mathcal {P}$ gives a symbolic representation of $\mathbb {Z}^2\overset {R}{\curvearrowright }\mathbb {T}^2$ and

  • $\mathcal {X}_{\mathcal {P},R}$ is a shift of finite type (SFT).

9.2 Proofs of main results

First, we have the following result.

Lemma 9.2. The dynamical system $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\mathcal {X}_{\mathcal {P}_n,R_n}$ is minimal and $\mathcal {X}_{\mathcal {P}_n,R_n}$ is aperiodic.

Proof. Since $R_n^{{\boldsymbol{e}}_1}$ and $R_n^{{\boldsymbol{e}}_2}$ are linearly independent irrational rotations on $\mathbb {R}^2/\mathbb {Z}^2$ , we have that $R_n$ is a free $\mathbb {Z}^2$ -action. Thus, from [Reference Labbé33, Lemma 5.2], $\mathcal {X}_{\mathcal {P}_n,R_n}$ is minimal and aperiodic.

Each atom of the partition $\mathcal {P}_n$ is invariant only under the trivial translation. Therefore, from [Reference Labbé33, Lemma 3.4], $\mathcal {P}_n$ gives a symbolic representation of the dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ . Thus, we can define the following function:

(9.2) $$ \begin{align} f_n:\mathcal{X}_{\mathcal{P}_n,R_n}\to\mathbb{T}^2 \end{align} $$

be such that $f_n(w)$ is the unique point in the intersection $\cap _{m=0}^{\infty }\overline {D}_m(w)$ .

Proposition 9.3. Let $n\geq 1$ be an integer. The map $f_n:\mathcal {X}_{\mathcal {P}_n,R_n}\to \mathbb {T}^2$ is a factor map satisfying

$$\begin{align*}f_n\circ\sigma^k = R_n^k\circ f_n \end{align*}$$

for every $k\in \mathbb {Z}^2$ .

Proof. The result is an application of Proposition 5.1 from [Reference Labbé33].

From the minimality of the Wang shift $\Omega _n$ proved separately in [Reference Labbé37], we may now prove Theorem E using the same method as in [Reference Labbé33].

Theorem E

For every integer $n\geq 1$ , the symbolic dynamical system $\mathcal {X}_{\mathcal {P}_n,R_n}$ corresponding to $\mathcal {P}_n,R_n$ is equal to the metallic mean Wang shift $\Omega _n$ :

$$\begin{align*}\Omega_n = \mathcal{X}_{\mathcal{P}_n,R_n}. \end{align*}$$

In particular, $\mathcal {P}_n$ is a Markov partition for the dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ .

Proof. From Proposition 8.1 in [Reference Labbé33], we have that $\mathcal {X}_{\mathcal {P}_n,R_n}\subseteq \Omega _n$ for every integer $n\geq 1$ . It was proved in [Reference Labbé37] that the Wang shift $\Omega _n$ is minimal for every integer $n\geq 1$ . Thus, $\mathcal {X}_{\mathcal {P}_n,R_n}=\Omega _n$ .

Each atom of the partition $\mathcal {P}_n$ is invariant only under the trivial translation. Therefore, from [Reference Labbé33, Lemma 3.4], $\mathcal {P}_n$ gives a symbolic representation of $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ . Since $\mathcal {X}_{\mathcal {P}_n,R_n}=\Omega _n$ is a shift of finite type, we conclude that the partition $\mathcal {P}_n$ is a Markov partition for the dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ .

In fact, we can show that the factor map $f_n$ is equal to the map $\Phi _n$ explicitly defined in Section 8 from the average of the labels of Wang tiles on the row and column containing the origin. It follows from the next lemma.

Lemma 9.4. For every $(x,y)\in [0,1)^2$ , we have $f_n(c_{(x,y)})=(x,y)$ .

Proof. Let $v_1,v_2,v_3,v_4\in V_n$ . Observe that

For every $k\in \mathbb {Z}^2$ , we have

so that

Therefore, for every $m\in \mathbb {N}$ , we have

$$\begin{align*}(x,y) \in \bigcap_{\Vert k\Vert\leq m} R_n^{-k}(\overline{P_{c_{(x,y)}(k)}}) = \overline{D_m}(c_{(x,y)}). \end{align*}$$

Since $\mathcal {P}_n$ gives a symbolic representation of the dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ , we have that $\cap _{m=0}^{\infty }\overline {D}_m(c_{(x,y)})$ is a singleton and

$$\begin{align*}\cap_{m=0}^{\infty}\overline{D}_m(c_{(x,y)})=\{(x,y)\}. \end{align*}$$

Therefore, $f(c_{(x,y)})=(x,y)$ .

Proposition 9.5. The factor map $f_n:\Omega _n\to \mathbb {T}^2$ is equal to the factor map $\Phi _n:\Omega _n\to \mathbb {T}^2$ explicitly defined in Equation (8.2):

$$\begin{align*}f_n=\Phi_n. \end{align*}$$

Proof. From Lemma 9.4, we have $ f_n(c_{(0,0)}) =(0,0)$ . Also, observe that the configuration $c_{(0,0)}$ is symmetric: $\widehat {c_{(0,0)}}=c_{(0,0)}$ . Thus, we have

$$\begin{align*}\Phi_n(c_{(0,0)}) =(\phi_n(\widehat{c_{(0,0)}}), \phi_n(c_{(0,0)})) =(\phi_n(c_{(0,0)}), \phi_n(c_{(0,0)}))=(0,0). \end{align*}$$

Let $w\in \Omega _n$ be any configuration. Since $\Omega _n$ is minimal [Reference Labbé37], there exists a sequence $(k_\ell )_{\ell \in \mathbb {N}}$ such that $k_\ell \in \mathbb {Z}^2$ such that $w=\lim _{\ell \to \infty }\sigma ^{k_\ell }(c_{(0,0)})$ . From Proposition 9.3 and Theorem D, $f_n$ and $\Phi _n$ are factor maps commuting the shift map with the $\mathbb {Z}^2$ -action $R_n$ on the torus $\mathbb {T}^2$ . Thus, we obtain

$$ \begin{align*} \Phi_n(w) &= \Phi_n\left(\lim_{\ell\to\infty}\sigma^{k_\ell}(c_{(0,0)})\right)\\ &= \lim_{\ell\to\infty}\Phi_n\circ\sigma^{k_\ell}(c_{(0,0)})\\ &= \lim_{\ell\to\infty}R_n^{k_\ell}\circ\Phi_n(c_{(0,0)})\\ &= \lim_{\ell\to\infty}R_n^{k_\ell}\left((0,0)\right)\\ &= \lim_{\ell\to\infty}R_n^{k_\ell}\circ f_n(c_{(0,0)})\\ &= \lim_{\ell\to\infty}f_n\circ\sigma^{k_\ell}(c_{(0,0)})\\ &= f_n\left(\lim_{\ell\to\infty}\sigma^{k_\ell}(c_{(0,0)})\right) = f_n(w).\\[-42pt] \end{align*} $$

The factor map $\Phi _n$ between the dynamical system $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ and the $\mathbb {Z}^2$ -action $R_n$ on the torus $\mathbb {T}^2$ satisfies additional properties. In particular, $\Phi _n$ is an isomorphism of measure-preserving dynamical systems. Their proofs follow the structure of similar results proved in [Reference Labbé33] for Jeandel–Rao tilings.

Theorem F

The Wang shift $\Omega _n$ and the $\mathbb {Z}^2$ -action $R_n$ have the following properties:

  1. (i) $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ is the maximal equicontinuous factor of $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ ,

  2. (ii) the factor map $\Phi _n:\Omega _n\to \mathbb {T}^2$ is almost one-to-one and its set of fiber cardinalities is $\{1,2,8\}$ ,

  3. (iii) the shift-action $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ on the metallic mean Wang shift is uniquely ergodic,

  4. (iv) the measure-preserving dynamical system $(\Omega _n,\mathbb {Z}^2,\sigma ,\nu )$ is isomorphic to $(\mathbb {T}^2,\mathbb {Z}^2,R_n,\lambda )$ where $\nu $ is the unique shift-invariant probability measure on $\Omega _n$ and $\lambda $ is the Haar measure on $\mathbb {T}^2$ .

Proof. From Theorem E, we have $\mathcal {X}_{\mathcal {P}_n,R_n}=\Omega _n$ .

(i) From Proposition 9.3, the factor map $f_n:\mathcal {X}_{\mathcal {P}_n,R_n}\to \mathbb {T}^2$ commutes the actions $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\mathcal {X}_{\mathcal {P}_n,R_n}$ and $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ . From [Reference Labbé33, Proposition 5.1], $f_n$ is one-to-one on $f_n^{-1}(\mathbb {T}^2\setminus \Delta _{\mathcal {P}_n,R_n})$ where

$$ \begin{align*} \Delta_{\mathcal{P}_n,R_n}:=\bigcup_{{\boldsymbol{k}}\in\mathbb{Z}^2}R_n^{\boldsymbol{k}} \left(\bigcup_{\tau\in\mathcal{T}_n}\partial P_\tau\right) \subset \mathbb{T}^2 \end{align*} $$

is the set of points whose orbit under the $\mathbb {Z}^2$ -action $R_n$ intersect the boundary of the topological partition $\mathcal {P}_n=\{P_\tau \}_{\tau \in \mathcal {T}_n}$ . From [Reference Labbé33, Corollary 5.3] (which is a consequence of [Reference Aujogue, Barge, Kellendonk and Lenz4, Lemma 3.11]), $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ is the maximal equicontinuous factor of $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\mathcal {X}_{\mathcal {P}_n,R_n}$ .

(ii) We have that $\{y\in \mathbb {T}^2:\mathrm {card}(f_n^{-1}(y))=1\}=\mathbb {T}^2\setminus \Delta _{\mathcal {P}_n,R_n}$ is a countable intersection of open sets and is dense in $\mathbb {T}^2$ . Thus, it is a $G_\delta $ -dense set in $\mathbb {T}^2$ . Therefore, the factor map $f_n:\mathcal {X}_{\mathcal {P}_n,R_n}\to \mathbb {T}^2$ is almost one-to-one. From Proposition 9.5, we have $f_n=\Phi _n$ .

Suppose that ${\boldsymbol{x}}\in \Delta _{\mathcal {P}_n,R_n}$ . We have $\mathrm {card}(f_n^{-1}({\boldsymbol{x}}))\geq 2$ . If $\mathrm {card}(f_n^{-1}({\boldsymbol{x}}))>2$ , then we may show that there exists ${\boldsymbol{n}}\in \mathbb {Z}^2$ such that ${\boldsymbol{x}}=R_n^{\boldsymbol{n}}({\boldsymbol{0}})$ . If ${\boldsymbol{x}}=R_n^{{\boldsymbol{n}}}({\boldsymbol{0}})$ for some ${\boldsymbol{n}}\in \mathbb {Z}^2$ , then the set $f_n^{-1}({\boldsymbol{x}})$ contains 8 different configurations of the form $\lim _{\varepsilon \to 0}c_{\varepsilon {\mathbf{v}}}$ for some ${\mathbf{v}}\in \mathbb {R}^2\setminus \Theta ^{\mathcal {P}_n}$ where $\Theta ^{\mathcal {P}_n}=\mathbb {R}\cdot \{(1,0),(0,1),(1,-\beta ),(1,\beta ^*)\}$ . If ${\boldsymbol{x}}\in \Delta _{\mathcal {P}_n,R_n}$ but not in the orbit of ${\boldsymbol{0}}$ under $R_n$ , then $\mathrm {card}(f_n^{-1}({\boldsymbol{x}}))=2$ . We conclude that $\{\mathrm {card}(f_n^{-1}({\boldsymbol{x}}))\mid {\boldsymbol{x}}\in \mathbb {T}^2\}=\{1,2,8\}$ .

(iii) The dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ is minimal. We have that $\lambda (\partial P)=0$ for each atom $P\in \mathcal {P}_n$ where $\lambda $ is the Haar measure on $\mathbb {T}^2$ . The partition $\mathcal {P}_n$ gives a symbolic representation of the dynamical system $\mathbb {Z}^2\overset {R_n}{\curvearrowright }\mathbb {T}^2$ . Thus, from [Reference Labbé33, Proposition 6.1], the dynamical system $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\mathcal {X}_{\mathcal {P}_n,R_n}$ is uniquely ergodic.

(iv) Since the dynamical system $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\mathcal {X}_{\mathcal {P}_n,R_n}$ is uniquely ergodic, it admits a unique shift-invariant probability measure $\nu $ on $\Omega _n$ . From [Reference Labbé33, Proposition 6.1], the measure-preserving dynamical system $(\Omega _n,\mathbb {Z}^2,\sigma ,\nu )$ is isomorphic to $(\mathbb {T}^2,\mathbb {Z}^2,R_n,\lambda )$ where $\lambda $ is the Haar measure on $\mathbb {T}^2$ .

10 Renormalization and Rauzy induction of $\mathbb {Z}^2$ -rotations

Another consequence of Theorem E is that the symbolic dynamical system $\mathcal {X}_{\mathcal {P}_n,R_n}$ is self-similar because this was proved in [Reference Labbé37] for the Wang shift $\Omega _n$ . The Rauzy induction of polygonal partitions and of toral $\mathbb {Z}^2$ -rotations defined in [Reference Labbé34] can be used to compute the self-similarity of the symbolic dynamical system $\mathcal {X}_{\mathcal {P}_n,R_n}$ . We illustrate below how this can be done for a fixed value of an integer $n\geq 1$ .

For some postive integer $n\geq 1$ , we define the positive root $\beta $ of the polynomial $x^2-nx-1$ . Computations will be done in the number field generated by this root. We perform the computations below with $n=3$ , but it works with other integers. For instance, the computation of the self-similarity for $n=7$ from the Rauzy induction is done in about 200 seconds on a recent laptop.

We define a function that computes the atoms $\Lambda _n^{-1}(v)$ for every $v\in V_n$ . Note that in SageMath, an entry equal to [-1,7,3,4] represents the inequality $7x_1+3x_2+4x_3\geq 1$ .

We define the set $V_n$ and we check that the sum of the area of the polygons $\{\Lambda _n^{-1}(v)\}_{v\in V_n}$ is 1.

For readability reason, we define a map which concatenates the entries of a vector into a string.

We define the $\mathbb {Z}^2$ -action $R_n$ on $\mathbb {R}^2/\mathbb {Z}^2$ as two polyhedron exchange transformations on the unit square.

We construct the partition (ignoring the atom with empty interior) and the three other partitions from it.

We compute the refinement of the and partitions and of the and partitions.

In general, we would need to compute the refinement of the two partitions. But here, since they are equal up to relabeling, we may take one as the refinement and compute the bijection of the labels between them.

We compute the set of Wang tiles defined by the refinement of the four partitions , , and :

We perform the Rauzy induction on the square window $[0,\beta ^{-1}]\times [0,\beta ^{-1}]$ using the algorithms induced_partition and induced_transformation defined in [Reference Labbé34]. First, we perform the induction on the domain restricted to the inequality $x\leq \beta ^{-1}$ .

Secondly, we perform the induction on the domain restricted to the inequality $y\leq \beta ^{-1}$ .

We rescale the induced partition by the factor $-\beta $ and translate it back to the unit square in the positive quadrant. Then we apply each generator of the $\mathbb {Z}^2$ -action once on the rescaled induced partition.

We check that the resulting partition is equal to the initial partition. We check that the induced action is equal to the initial action.

The self-similarity computed by this Rauzy induction is the product of the above 2-dimensional substitutions by the bijection of the labels.

The computed self-similarity s123 is:

The above self-similarity can be illustrated with the Wang tiles computed above as follows:

We may observe that the self-similarity computed here from the Rauzy induction on polygonal partition on $\mathcal {P}_3$ and toral $\mathbb {Z}^2$ -action $R_3$ is the same as the self-similarity proved for the Wang shift $\Omega _3$ in [Reference Labbé37].

11 Open questions

For almost twenty years, the Kari and Culik sets of Wang tiles were the smallest known aperiodic sets of Wang tiles. In 2015, Jeandel and Rao performed an exhaustive search on all sets of Wang tiles of cardinality up to 11 [Reference Jeandel and Rao21] and proved that sets of Wang tiles of cardinality at most 10 either do not tile the plane or tile the plane and one of the valid tilings is periodic. Moreover, they provided a list of 36 sets of 11 Wang tiles considered to be candidates for being aperiodic. One of the candidates was intriguing because Fibonacci numbers appeared in the structure of the transducers involved in the computation of valid tilings. Jeandel and Rao focused on the intriguing candidate, shown in Figure 20, and they proved it to be aperiodic. The set of valid configurations over these 11 tiles forms a subshift that we call the Jeandel–Rao Wang shift.

Figure 20 The Jeandel–Rao aperiodic set of 11 Wang tiles.

In [Reference Labbé33], it was proved that a minimal subshift within the Jeandel–Rao Wang shift is the coding of a dynamical system defined by the following $\mathbb {Z}^2$ -action $R_0$ on the $2$ -dimensional torus $\mathbb {R}^2/\Gamma _0$ , where $\Gamma _0=\left (\begin {smallmatrix} \varphi & 1\\ 0 & \varphi +3\end {smallmatrix}\right )\mathbb {Z}^2$ is a lattice in $\mathbb {R}^2$ involving the golden ratio $\varphi =\frac {1+\sqrt {5}}{2}$ :

$$\begin{align*}\begin{array}{rccl} R_0:&\mathbb{Z}^2\times\mathbb{R}^2/\Gamma_0 & \to & \mathbb{R}^2/\Gamma_0\\ &({\boldsymbol{k}},{\boldsymbol{x}}) & \mapsto &{\boldsymbol{x}}+{\boldsymbol{k}}. \end{array} \end{align*}$$

The symbolic coding is obtained through a polygonal partition $\mathcal {P}_0$ of a fundamental domain of $\mathbb {R}^2/\Gamma _0$ . The partition was proved to be a Markov partition for $R_0$ after comparing the substitutive structure computed from the Rauzy induction of $R_0$ and $\mathcal {P}_0$ [Reference Labbé34] with the substitutive structure of the associated Wang shift [Reference Labbé32, Reference Labbé35].

Intuitively, this means that the Jeandel–Rao Wang tiles shown in Figure 20 correspond to computing the orbit of points in the plane $\mathbb {R}^2$ under the translations by $+1$ horizontally and $+1$ vertically modulo the lattice $\Gamma _0$ . How this is possible is still a mystery. The link between the 11 Jeandel–Rao Wang tiles themselves and the golden ratio or toral rotation $R_0$ remains unclear. Unlike the Kari example, the values 0, 1, 2, 3, 4 of the labels of the Jeandel–Rao Wang tiles are five distinct symbols rather than arithmetic values. They do not satisfy a known equation.

In general, the following questions can be raised.

Question 1. Let $\mathcal {T}$ be a set of Wang tiles such that the Wang shift $\Omega _{\mathcal {T}}$ is aperiodic.

  • Is it multiplicative (Kari-Culik-like)? More precisely, can we replace the labels of the tiles in $\mathcal {T}$ by arithmetic values in such a way that an equation similar to (1.1) is satisfied?

  • Is it additive (metallic mean-like)? More precisely, can we replace the labels of the tiles in $\mathcal {T}$ by integer vectors computed from floors of linear forms as in Proposition 7.4 and satisfying additive equations as in Theorem B?

Does there exist an aperiodic set of Wang tiles which is neither multiplicative nor additive?

Solving Question 1 for Jeandel–Rao Wang tiles would improve our understanding of the Jeandel–Rao Wang shift. Hopefully it would allow to generate more examples maybe not related to the golden ratio and that are not self-similar. Remember that the computations made by Jeandel and Rao took one year using 100 cpus to explore exhaustively the sets of 11 Wang tiles [Reference Jeandel and Rao21]. Finding new examples by exploring all sets of 12, 13 or 14 Wang tiles becomes soon out of reach. We need to understand what is happening in order to find other examples and characterize them.

Question 2. If an aperiodic set of Wang tiles is additive (metallic mean-like) with labels given by integer vectors satisfying equations, can we use the equations to directly prove that the Wang shift $\Omega _{\mathcal {T}}$ is aperiodic following the short arithmetical argument for the nonperiodicity of Kari’s tile set?

Finding an answer to Question 2 for the Ammann set of 16 Wang tiles was the original motivation of the author which led to the discovery of the family of metallic mean Wang tiles. As we discussed in Section 6, Question 2 remains open even for the Ammann 16 Wang tiles and the family of metallic mean Wang tiles.

In general, we may ask the following question.

Question 3. For which invertible matrix $M\in \mathrm {GL}_2(\mathbb {R})$ does there exist a set of Wang tiles $\mathcal {T}$ such that the Wang shift $\Omega _{\mathcal {T}}$ is isomorphic, as a measure-preserving dynamical system, to the toral $\mathbb {Z}^2$ -rotation $R:\mathbb {Z}^2\times \mathbb {T}^2\to \mathbb {T}^2$ defined by $R^{\boldsymbol{k}}({\boldsymbol{x}})={\boldsymbol{x}}+M{\boldsymbol{k}}$ on the 2-dimensional torus $\mathbb {T}^2=(\mathbb {R}/\mathbb {Z})^2$ ?

The Markov partition associated with Jeandel–Rao tiles and action $R_0$ on $\mathbb {R}^2/\Gamma _0$ is related to the golden ratio [Reference Labbé33]. In this contribution, we describe a family of $\mathbb {Z}^2$ -actions related to the metallic-mean quadratic integers. Can we find examples related to other numbers?

Question 4. For which $\mathbb {Z}^2$ -actions defined by rotations on a $2$ -dimensional torus does there exist a Markov Partition? When is this partition smooth/polygonal?

As for toral hyperbolic automorphisms, we can expect that smooth Markov partitions are associated with algebraic integers of degree 2 and that the partition is piecewise linear in this case [Reference Cawley10]. Markov partitions for typical toral hyperbolic automorphisms have fractal boundaries [Reference Bowen8].

The relation with toral hyperbolic automorphisms does not come out of nowhere. Indeed, the self-similarity of $\Omega _n$ proved in [Reference Labbé37] has an incidence matrix of size $(n+3)^2\times (n+3)^2$ . Its eigenvalues are all quadratic integers, 0 or $\pm 1$ . This incidence matrix acts hyperbolically as a toral automorphism on a subspace of $\mathbb {R}^{(n+3)^2}$ thus admits a Markov partition with piecewise linear boundaries. A link between this Markov partition and the partition $\mathcal {P}_n$ can be expected, because this is what happens for $1$ -dimensional sequences. Indeed, the Markov partition associated with the toral automorphism $\left (\begin {smallmatrix}1&1&1\\1&0&0\\0&1&0 \end {smallmatrix}\right )$ is a suspension of the Rauzy fractal [Reference Rauzy47] as nicely illustrated in a talk by Timo Jolivet [Reference Jolivet23].

Question 5. What is the relation between the Markov partition for the hyperbolic toral automorphism defined from the incidence matrix of the self-similarity of $\Omega _n$ and the Markov partition $\mathcal {P}_n$ associated with $\mathbb {Z}^2\overset {\sigma }{\curvearrowright }\Omega _n$ ?

The symmetric properties of $\Omega _n$ and of the partition $\mathcal {P}_n$ make them a good object of study to tackle these questions in more generality.

Acknowledgments

The author would like to thank the referees for the extensive and helpful comments that have helped to improve the presentation. The author is also thankful to Hugo Parlier for his comments on an earlier draft of the introduction and to Vincent Delecroix for making the author realize that it is not the Birkhoff ergodic theorem which is needed in the proof of Proposition 8.2 but rather simply the Weyl’s equidistribution theorem.

Competing interests

The authors have no competing interests to declare.

Funding statement

This work was partly funded from France’s Agence Nationale de la Recherche (ANR) projects CODYS (ANR-18-CE40-0007) and IZES (ANR-22-CE40-0011). It was also supported by grants from the Symbolic Dynamics and Arithmetic Expansions (SymDynAr) Project, co-funded by ANR (ANR-23-CE40-0024) and FWF (I 6750), the Austrian Science Fund.

Reproducibility statement

All results proved in this article are proved by hand. Computations performed in Section 10 are based on the open-source mathematical software SageMath [50] and the optional package slabbe [Reference Labbé38]. All SageMath input/output blocks in this article were created using the sageexample environment with SageTeX version 2021/10/16 v3.6 and with the following software versions:

The fact that these software are open-source means that anyone is free to use, reproduce, verify, adapt for their own needs all of the computations performed therein according to the GNU General Public License (version 2, 1991, http://www.gnu.org/licenses/gpl.html).

The contents of all of the sageexample environments from the tex source are gathered in the file demos/arXiv_2403_03197_doctest.sage autogenerated by SageTeX when running pdflatex. This file is included in the slabbe package and available at https://gitlab.com/seblabbe/slabbe/. It allows to make sure that future releases of the package do not break the code included in this article. It is possible to reproduce all computations present in this article and check that all outputs are correct, by doctesting this file, that is, by running the command sage -t demos/arXiv_2403_03197_doctest.sage. It should output All tests passed! and [58 tests, 11.75s wall] (most probably with a different timing).

References

Aaronson, S., ‘The busy beaver frontier’, SIGACT News 51 (2020), 3254. http://doi.org/10.1145/3427361.3427369.CrossRefGoogle Scholar
Akiyama, S. and Arnoux, P. (eds), Substitution and Tiling Dynamics: Introduction to Self-inducing Structures, CIRM Jean-Morlet Chair, Marseille, France, Fall 2017, Vol. 2273 (Cham: Springer, 2020), xix+453 pp.Google Scholar
Ammann, R., Grünbaum, B. and Shephard, G. C., ‘Aperiodic tiles’, Discrete Comput. Geom. 8 (1992), 125.10.1007/BF02293033CrossRefGoogle Scholar
Aujogue, J.-B., Barge, M., Kellendonk, J. and Lenz, D., ‘Equicontinuous factors, proximality and Ellis semigroup for Delone sets’, in Mathematics of Aperiodic Order, Vol. 309 of Progr. Math., Birkhäuser/Springer, Basel (2015), 137194. http://doi.org/10.1007/978-3-0348-0903-0_5.CrossRefGoogle Scholar
Baake, M. and Grimm, U., Aperiodic Order. Vol. 1, Vol. 149 of Encycl. Math. Appl. (Cambridge University Press, Cambridge, 2013), xvi+531 pp. http://doi.org/10.1017/CBO9781139025256.CrossRefGoogle Scholar
Baake, M. and Grimm, U. (eds), Aperiodic Order. Vol. 2: Crystallography and almost periodicity, Vol. 166 of Encycl. Math. Appl. (Cambridge University Press, Cambridge, 2017). http://doi.org/10.1017/9781139033862.CrossRefGoogle Scholar
Berger, R., ‘The undecidability of the domino problem’, Mem. Amer. Math. Soc. No. 66 (1966), 72.Google Scholar
Bowen, R., ‘Markov partitions are not smooth’, Proc. Amer. Math. Soc. 71 (1978), 130132. http://doi.org/10.2307/2042234.Google Scholar
Brady, A. H., ‘The busy beaver game and the meaning of life’, in The Universal Turing Machine, a Half-Century Survey, (1988), 259277.10.1093/oso/9780198537748.003.0009CrossRefGoogle Scholar
Cawley, E., ‘Smooth Markov partitions and toral automorphisms’, Ergodic Theory Dynam. Systems 11 (1991), 633651. http://doi.org/10.1017/S0143385700006404.CrossRefGoogle Scholar
Culik, K. IIAn aperiodic set of 13 Wang tiles’, Discrete Math. 160 (1996), 245251. http://doi.org/10.1016/S0012-365X(96)00118-5.Google Scholar
de Bruijn, N. G., ‘Algebraic theory of Penrose’s nonperiodic tilings of the plane. I, II’, Indag. Math. 43 (1981), 3952, 53–66.Google Scholar
de Spinadel, V. W., ‘The family of metallic means’, Vis. Math. 1 (1999), 1 HTML document; approx. 16.Google Scholar
Durand, B., Gamard, G. and Grandjean, A., ‘Aperiodic tilings and entropy’, Theoret. Comput. Sci. 666 (2017), 3647. http://doi.org/10.1016/j.tcs.2016.12.013.Google Scholar
Eigen, S., Navarro, J. and Prasad, V. S., ‘An aperiodic tiling using a dynamical system and Beatty sequences’, in Dynamics, Ergodic Theory, and Geometry, Vol. 54 of Math. Sci. Res. Inst. (Cambridge University Press, Cambridge, 2007), 223241. http://doi.org/10.1017/CBO9780511755187.009.CrossRefGoogle Scholar
Einsiedler, M. and Schmidt, K., ‘Markov partitions and homoclinic points of algebraic ${\mathbf{Z}}^d$ -actions’, Tr. Mat. Inst. Steklova 216 (1997), 265284.Google Scholar
Fiebig, D., ‘Factor maps, entropy and fiber cardinality for Markov shifts’, Rocky Mountain J. Math. 31 (2001), 955986. http://doi.org/10.1216/rmjm/1020171674.CrossRefGoogle Scholar
Fuhrmann, G., Gröger, M. and Lenz, D., ‘The structure of mean equicontinuous group actions’, Israel J. Math. 247 (2022), 75123. http://doi.org/10.1007/s11856-022-2292-8.CrossRefGoogle Scholar
Grünbaum, B. and Shephard, G. C., Tilings and Patterns (W. H. Freeman and Company, New York, 1987), xii+700 pp.Google Scholar
Hochman, M., ‘Multidimensional shifts of finite type and sofic shifts’, in Combinatorics, Words and Symbolic Dynamics, Vol. 159 of Encycl. Math. Appl., Cambridge Univ. Press, Cambridge (2016), 296358. http://doi.org/10.1017/CBO9781139924733.010.Google Scholar
Jeandel, E. and Rao, M., ‘An aperiodic set of 11 Wang tiles’, Adv. Comb. 2021 (2021), 37. http://doi.org/10.19086/aic.18614.Google Scholar
Jeandel, E. and Vanier, P., ‘The undecidability of the Domino Problem’, in Substitution and tiling dynamics: introduction to self-inducing structures, Vol. 2273 of Lecture Notes in Math., Springer, Cham (2020), 293357. http://doi.org/10.1007/978-3-030-57666-0_6.Google Scholar
Jolivet, T., ‘Toral automorphisms, Markov partitions and fractals’ (2012). https://jolivet.org/timo/docs/slides_guangzhou_toral.pdf.Google Scholar
Kari, J., ‘A small aperiodic set of Wang tiles’, Discrete Math. 160 (1996), 259264. http://doi.org/10.1016/0012-365X(95)00120-L.CrossRefGoogle Scholar
Kari, J., ‘The tiling problem revisited (extended abstract)’, in Machines, Computations, and Universality. 5th International Conference, MCU 2007, Orléans, France, September 10–13, 2007. Proceedings, Springer, Berlin (2007), 7279. http://doi.org/10.1007/978-3-540-74593-8_6.CrossRefGoogle Scholar
Kari, J., ‘On the undecidability of the tiling problem’, in Geffert, V., Karhumäki, J., Bertoni, A., Preneel, B., Návrat, P. and Bieliková, M. (eds), SOFSEM 2008: Theory and Practice of Computer Science (Springer, Berlin, 2008), 7482.CrossRefGoogle Scholar
Kari, J., ‘Piecewise affine functions, Sturmian sequences and Wang tiles’, Fundam. Inform. 145 (2016), 257277. http://doi.org/10.3233/FI-2016-1360.Google Scholar
Kari, J. and Papasoglu, P., ‘Deterministic aperiodic tile sets’, Geom. Funct. Anal. 9 (1999), 353369. http://doi.org/10.1007/s000390050090.CrossRefGoogle Scholar
Knuth, D. E., The Art of Computer Programming. Vol. 1: Fundamental Algorithms, 2nd printing (Addison-Wesley, Reading, London, Don Mills, 1969), xxi+634 pp.Google Scholar
Kuipers, L. and Niederreiter, H., Uniform Distribution of Sequences, Pure and Applied Mathematics (Wiley, New York, 1974), xiv+390 pp.Google Scholar
Kurka, P., Topological and Symbolic Dynamics, Vol. 11 of Cours Spécialisés, Société Mathématique de France, Paris (2003), xii+315 pp.Google Scholar
Labbé, S., ‘A self-similar aperiodic set of 19 Wang tiles’, Geom. Dedicata 201 (2019), 81109. http://doi.org/10.1007/s10711-018-0384-8.Google Scholar
Labbé, S., ‘Markov partitions for toral ${\mathbb{Z}}^2$ -rotations featuring Jeandel–Rao Wang shift and model sets’, Ann. H. Lebesgue 4 (2021), 283324. http://doi.org/10.5802/ahl.73. Google Scholar
Labbé, S., ‘Rauzy induction of polygon partitions and toral ${\mathbb{Z}}^2$ -rotations’, J. Mod. Dyn. 17 (2021), 481528. http://doi.org/10.3934/jmd.2021017. Google Scholar
Labbé, S., ‘Substitutive structure of Jeandel–Rao aperiodic tilings’, Discrete Comput. Geom. 65 (2021), 800855. http://doi.org/10.1007/s00454-019-00153-3.CrossRefGoogle Scholar
Labbé, S., ‘Three characterizations of a self-similar aperiodic 2-dimensional subshift’, Preprint (2020), arXiv:2012.03892, 46 pp., book chapter in preparation.Google Scholar
Labbé, S., ‘Metallic mean Wang tiles I: self-similarity, aperiodicity and minimality’, Forum Math. Sigma 13 (2025), e133. http://doi.org/10.1017/fms.2025.10069.CrossRefGoogle Scholar
Labbé, S., ‘Optional SageMath Package slabbe (Version 0.8.0)’, https://pypi.python.org/pypi/slabbe/ (2025).Google Scholar
Lind, D., ‘Multi-dimensional symbolic dynamics’, in Symbolic Dynamics and its Applications, Vol. 60 of Proc. Sympos. Appl. Math., Amer. Math. Soc., Providence, RI (2004), 6179. http://doi.org/10.1090/psapm/060/2078846.Google Scholar
Lind, D. and Marcus, B., An Introduction to Symbolic Dynamics and Coding (Cambridge University Press, Cambridge, 1995), xvi+495 pp. http://doi.org/10.1017/CBO9780511626302.Google Scholar
Mozes, S., ‘Tilings, substitution systems and dynamical systems generated by them’, J. Analyse Math. 53 (1989), 139186. http://doi.org/10.1007/BF02793412.CrossRefGoogle Scholar
OEIS Foundation Inc., ‘Entry A060843 in the On-Line Encyclopedia of Integer Sequences: Busy Beaver problem: $a(n)=$ maximal number of steps that an $n$ -state Turing machine can make on an initially blank tape before eventually halting’, (2023). https://oeis.org/A060843. Google Scholar
Ollinger, N., ‘Two-by-two substitution systems and the undecidability of the domino problem’, in Logic and Theory of Algorithms, Vol. 5028 of Lecture Notes in Comput. Sci., Springer, Berlin (2008), 476485. http://doi.org/10.1007/978-3-540-69407-6_51.CrossRefGoogle Scholar
Penrose, R., ‘Pentaplexity. A class of non-periodic tilings of the plane’, Math. Intell. 2 (1979), 3237. http://doi.org/10.1007/BF03024384.CrossRefGoogle Scholar
Praggastis, B., ‘Numeration systems and Markov partitions from self-similar tilings’, Trans. Amer. Math. Soc. 351 (1999), 33153349. http://doi.org/10.1090/S0002-9947-99-02360-0.Google Scholar
Priebe, N. and Solomyak, B., ‘Characterization of planar pseudo-self-similar tilings’, Discrete Comput. Geom. 26 (2001), 289306.CrossRefGoogle Scholar
Rauzy, G., ‘Nombres algébriques et substitutions’, Bull. Soc. Math. France 110 (1982), 147178. http://doi.org/10.24033/bsmf.1957.Google Scholar
Robinson, E. A. Jr.The dynamical properties of Penrose tilings’, Trans. Amer. Math. Soc. 348 (1996), 44474464. http://doi.org/10.1090/S0002-9947-96-01640-6.CrossRefGoogle Scholar
Robinson, R. M., ‘Undecidability and nonperiodicity for tilings of the plane’, Invent. Math. 12 (1971), 177209. http://doi.org/10.1007/BF01418780.CrossRefGoogle Scholar
Sage Developers, SageMath, the Sage Mathematics Software System (Version 10.6) (2025). http://www.sagemath.org.Google Scholar
Schmidt, K., ‘Multi-dimensional symbolic dynamical systems’, in Codes, Systems, and Graphical Models (Minneapolis, MN, 1999), Vol. 123 of IMA Vol. Math. Appl. (Springer, New York, 2001), 6782. http://doi.org/10.1007/978-1-4613-0165-3_3.Google Scholar
Schroeder, M., Fractals, Chaos, Power Laws. Minutes from an Infinite Paradise (W.H. Freeman and Company, New York, 1991).CrossRefGoogle Scholar
Senechal, M., Quasicrystals and Geometry (Cambridge University Press, Cambridge, 1995).Google Scholar
Siefken, J., ‘A minimal subsystem of the Kari–Culik tilings’, Ergodic Theory Dynam. Systems 37 (2017), 16071634. http://doi.org/10.1017/etds.2015.118.CrossRefGoogle Scholar
Smith, D., Myers, J. S., Kaplan, C. S. and Goodman-Strauss, C., ‘An aperiodic monotile’, Comb. Theory 4 (2024), Paper No. 6, 91. http://doi.org/10.5070/C64163843.Google Scholar
Smith, D., Myers, J. S., Kaplan, C. S. and Goodman-Strauss, C., ‘A chiral aperiodic monotile’, Comb. Theory 4 (2024), Paper No. 13, 25. http://doi.org/10.5070/C64264241.Google Scholar
Socolar, J. E. S. and Taylor, J. M., ‘An aperiodic hexagonal tile’, J. Combin. Theory Ser. A 118 (2011), 22072231. http://doi.org/10.1016/j.jcta.2011.05.001.CrossRefGoogle Scholar
Solomyak, B., ‘Dynamics of self-similar tilings’, Ergodic Theory Dynam. Systems 17 (1997), 695738. http://doi.org/10.1017/S0143385797084988.CrossRefGoogle Scholar
Solomyak, B., ‘Nonperiodicity implies unique composition for self-similar translationally finite tilings’, Discrete Comput. Geom. 20 (1998), 265279. http://doi.org/10.1007/PL00009386.CrossRefGoogle Scholar
Turing, A. M., ‘On computable numbers, with an application to the Entscheidungsproblem’, Proc. London Math. Soc. 42(2) (1936), 230265. http://doi.org/10.1112/plms/s2-42.1.230.Google Scholar
Walters, P., An Introduction to Ergodic Theory, Vol. 79 of GTM (Springer-Verlag, New York, Berlin, 1982), ix+250 pp.10.1007/978-1-4612-5775-2CrossRefGoogle Scholar
Wang, H., ‘Proving theorems by pattern recognition – II’, Bell Syst. Tech. J. 40 (1961), 141. http://doi.org/10.1002/j.1538-7305.1961.tb03975.x.CrossRefGoogle Scholar
Figure 0

Figure 1 Averages of horizontal labels in a tiling with Kari’s 14 tiles are orbits under the map g on the interval $[\frac {2}{3},2]$; see [14, 27].

Figure 1

Figure 2 The metallic mean Wang tile set $\mathcal {T}_n$ for $n=3$.

Figure 2

Figure 3 A valid $15\times 15$ pattern with Wang tile set $\mathcal {T}_3$.

Figure 3

Figure 4 A Venn diagram of aperiodic sets of Wang tiles. Aperiodicity of Kari [24] and Culik [11] sets of tiles and their extensions [15] follows from the arithmetic equations satisfied by their matching rules. In this article, we show that the dashed region in the Venn diagram is nonempty, that is, there exists a family of substitutive (self-similar) aperiodic sets of Wang tiles whose matching rules satisfy arithmetic equations.

Figure 4

Figure 5 The $\theta _n$-chip is a computer chip computing $\theta _n(u,v)$ and $\theta _n(v,u)$ from the left input u and bottom input v.

Figure 5

Figure 6 A rectangular cluster of copies of the $\theta _n$-chip.

Figure 6

Figure 7 A $10\times 5$ valid rectangular tiling with the set $\mathcal {T}_n$ with $n=3$. The numbers indicated in the right margin are the average of the inner products $\langle \frac {1}{n}d,v\rangle $ over the vectors v appearing as top (or bottom) labels of a horizontal row of tiles and where $d=(0,-1,1)$. We observe that these numbers increase by $\frac {3}{10}\ \pmod 1$ from row to row. The number $\frac {3}{10}$ is equal to the frequency of columns containing junction tiles (a junction tile is a tile whose labels all start with 0). Observe that this is a cylindrical tiling (left and right outer labels of the rectangle match) which simplifies the equations involved because the left and right carries cancel.

Figure 7

Figure 8 The partition and its image under a symmetry with the positive diagonal. Their refinement is $\mathcal {P}_3$ which is a partition of the unit square into 36 polygonal atoms. Here $\beta $ is the third metallic mean, that is, the positive root of $x^2-3x-1$.

Figure 8

Figure 9 The set of 3 Wang tiles introduced in [62] using letters $\{A,B,C,D,E\}$ instead of numbers from the set $\{1,2,3,4,5\}$ for labeling the edges. Each tile is identified uniquely by an index from the set $\{0,1,2\}$ written at the center each tile.

Figure 9

Figure 10 A finite $3\times 3$ pattern on the left is valid with respect to the Wang tiles since it respects Equations (3.3) and (3.4). Validity can be verified on the tiling shown on the right.

Figure 10

Figure 11 Extended metallic mean Wang tile sets $\mathcal {T}_n'$ for $n=4$. The junction tiles $j_n^{0,0,1,1}$ and $j_n^{1,1,0,0}$ are shown with a $\times $-mark in their center.

Figure 11

Figure 12 Metallic mean Wang tile sets $\mathcal {T}_n$ for $n=1,2,3,4,5$.

Figure 12

Figure 13 An $h\times k$ rectangular tiling of tiles from $\mathcal {C}_n$.

Figure 13

Figure 14 The preimage sets of the map $(x,y)\mapsto \Lambda _n(x,y)$ defines a partition of $[0,1)^2$ which is the refinement of the three partitions on the left. The above images are when $n=3$.

Figure 14

Figure 15 For every $(x,y)\in [0,1)^2$ the map $\mathbb {Z}^2\to \mathcal {T}_n$ defined by is a valid tiling of the plane by the set of Wang tiles $\mathcal {T}_n$.

Figure 15

Figure 16 The partitions , , and .

Figure 16

Figure 17 The partitions , , and .

Figure 17

Figure 18 The partitions , , and .

Figure 18

Figure 19 The partitions , , and .

Figure 19

Figure 20 The Jeandel–Rao aperiodic set of 11 Wang tiles.