This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Distortion Reversal in Aperiodic Tilings

Louisa Barnsley Mathematical Sciences Institute
Australian National University
Canberra, ACT, Australia
[email protected]
Michael Barnsley Mathematical Sciences Institute
Australian National University
Canberra, ACT, Australia
[email protected]
 and  Andrew Vince Department of Mathematics
University of Florida
Gainesville, FL, USA
[email protected]
Abstract.

It is proved that homeomorphic images of certain two-dimensional aperiodic tilings, such as Ammann-A2 tilings, are recognizable, in both mathematical and practical senses. One implication of the results is that it is possible to search for distorted aperiodic structures in nature, where they may be hiding in plain sight.

Key words and phrases:
Ammann tiling, aperiodic tiling, distortion
1991 Mathematics Subject Classification:
52C20, 05B45
This work was partially supported by a grant from the Simons Foundation (#322515 to Andrew Vince).

1. Introduction

Herein we prove that homeomorphic images of certain two-dimensional self-similar aperiodic tilings are recognizable, in both mathematical and practical senses. In the practical sense our results say it is possible to search for and recognize distorted aperiodic structures, for example in natural and physical settings. We recall that the mathematical discovery of aperiodic tilings by Penrose in 1972 and Ammann in 1975 preceded the discovery by Shechtman in 1982 of quasicrystals, and the subsequent examples of quasicrystals in meteorites in 2009 [13].

In this paper we focus on Ammann-A2 tilings [2], but the ideas are more generally applicable. Our results illustrate that aperiodic tilings can be retrieved from their images distorted by unknown homeomorphisms. For example, we show how distorted Ammann-A2 tilings, such as the ones in Figure 1, are recognizable both practically and mathematically. In Figure 2 an algorithm (without knowledge of the distorting homeomorphism) is applied to the distorted tiling to successively retrieve a patch of the original Ammann-A2 tiling.

The situation is a bit strange: the unknown distortion may be such as to change the Hausdorff dimension of the boundaries of the tiles in a very rough way, or to transform an Ammann-A2 tiling to a tiling by triangles. Nevertheless, the key properties of the tiling may still be discernable.

Refer to caption
Figure 1. This illustrates a patch of an Ammann-A2 tiling, top left, and three different homeomorphisms of it. Properties of the Ammann-A2 tiling may be determined from a distorted image.

There is a substantial literature on the occurrence of two-dimensional tilings in physics [14]. Many of these are distortions of tilings by hexagons. Many naturally occurring tilings have an average of six faces per tile, and typical tiles are hexagons. Ammann-A2 tiles are hexagonal and comprise among the simplest of self-similar tilings. Despite this, they do not seem to show up in nature. Are they hiding in plain sight? We provide some tools which may be applied to this question.

2. Organization and Main Results

Let PP be a finite aperiodic prototile set. This means that there exist tilings of the plane by copies of the tiles in PP, but every such tiling is non-periodic, i.e., has no translational symmetry. Such tilings have been objects of fascination, both in research and recreational mathematics, since their introduction in the 1960’s. The most well-known such prototile set consists of the two Penrose tiles. To develop the ideas in this paper we concentrate, because of their simplicity, on tilings based on the set of two Ammann-A2 tiles. There are uncountably many corresponding Ammann-A2 tilings, obtained either by matching rules or by substitution rules. Basic properties of such Ammann-A2 tilings are provided in Section 3.

Refer to caption
Figure 2. Successive images, obtained by amalgamation, converge towards patches of an Ammann-A2 tiling.

Many non-periodic tilings have a hierarchical structure, meaning that if TT is such a tiling, then there exist a sequence of tilings T=T¯0,T¯1,T¯2,T=\overline{T}_{0},\overline{T}_{1},\overline{T}_{2},\dots, such that, for all n1n\geq 1, each tiling in T¯n\overline{T}_{n} is the non-overlapping union of tiles in T¯n1\overline{T}_{n-1}. For the Ammann-A2 tilings, the tiling T¯n\overline{T}_{n}, appropriately scaled, is again an Ammann-A2 tiling. Specifically, there is a constant ss such that Tn:=snT¯nT_{n}:=s^{n}\overline{T}_{n} is an Ammann-A2 tiling for all nn. There is a well-defined procedure, which we call amalgamation, taking TnT_{n} to Tn+1T_{n+1}, and denoted Tn+1=a(Tn)T_{n+1}=a(T_{n}). Amalgamation and hierarchy for the Ammann-A2 tilings is explored in Section 4.

Here is the question that is the subject of this paper. Suppose that we are given a distorted version TT^{\prime} of an Ammann-A2 tiling TT. Specifically, there is an unknown (not revealed to us) homeomorphism hh of the plane such that T=h(T)T^{\prime}=h(T). Is it possible to retrieve the undistorted Ammann-A2 tiling TT from its distorted image TT^{\prime} - without knowledge of hh? As long as the distance between any point xx in the plane and its image h(x)h(x) is bounded, we answer this question in the affirmative. Suppose that amalgamation can be carried over to the distorted tiling TT^{\prime}. Then Section 5 contains the following somewhat surprising result: if Tn=an(T),n0,T^{\prime}_{n}=a^{n}(T^{\prime}),\,n\geq 0, are the tilings obtain by successive application of the amalgamation operator aa applied to the distorted image TT^{\prime}, then the sequence {Tn}\{T_{n}\} of tilings converges, in the sense of Theorem 1 in Section 5 and Remark 1, to the original tiling TT. This is illustrated in Figure 2.

The above distortion reversal result is predicated on being able to amalgamate tiles of the distorted tiling TT^{\prime}, a tiling in which metric and geometric properties of the original tiling are lost. Only the combinatorial properties of TT are carried over to TT^{\prime}. In Section 6 we obtain, completely combinatorially, the needed amalgamation operator aa applied to TT^{\prime}. This is done in Algorithm A in Section 6; see Theorem 2. Of course, the plane and tilings of it, are unbounded. We would like to perform the distortion reversal efficiently on a large patch of the tiling TT^{\prime}, say on a patch that contains the disk DRD_{R} of radius RR centered at the origin. Algorithm B, a variation of Algorithm A, accomplishes this. Theorem 3 in Section 6 states that each application of the combinatorial amalgamation applied to such a patch can be performed in running time O(R2)O(R^{2}).

An alternative method for reversing the distortion in an Ammann-A2 tiling is provided in Section 7. It uses the combinatorial amalgamation operator applied to the distorted tiling TT^{\prime} to generate an infinite binary string cc, called the code of the tiling. The code is the input for an iterated function system based method to retrieve the original tiling TT. This method, encapsulated in Theorem 4 of Section 7, has the advantage that there is no restriction on the homeomorphism hh. It has the disadvantage that there is no sequence of increasingly accurate images converging to the original tiling TT; it simply produces TT or, in finite time, an arbitrarily large patch of TT.

3. Ammann-A2 Tilings

In this paper a tile is defined as a set in the plane homeomorphic to a closed disk. A tiling of the plane is a set of non-overlapping tiles whose union is 2{\mathbb{R}}^{2}. By non-overlapping is meant that the intersection of any two distinct tiles has empty interior. For all tilings in this paper, the intersection of any two distinct tiles is connected (possibly empty). A patch of a tiling is a finite number of tiles whose union is a topological disk.

The classical Ammann-A2 hexagon GG, sometimes referred as the golden bee, is depicted in Figure 3. It is the only polygon, other than any right triangle and any parallelogram with side lengths in the ratio 2:1\sqrt{2}:1, that is the non-overlapping union of two smaller similar copies of itself [12]. These two smaller copies are isometric to sGsG and s2Gs^{2}G, where s=1/τs=1/\tau and τ\tau is the square root of the golden ratio, i.e., τ\tau is the positive real root of x4x21=0x^{4}-x^{2}-1=0.

Refer to caption
Refer to caption
Figure 3. Golden bee GG

An Ammann-A2 tiling is a tiling of the plane by non-overlapping isometric copies of sGsG and s2Gs^{2}G, which we will refer to as the large and small tiles, respectively. The tiling must obey matching rules dictated by the elliptical markings on the upper tiles of the left panel in Figure 4. Other alternative but equivalent sets of markings are also possible, for example the markings shown in the right panel of Figure 4. Although the decorations on small and large tiles at the top of the left panel and the decorations on the small and large tiles of the right panel differ, matching according to either set of decorations define the Ammann-A2 tilings, as observed in [7, p. 551]; see also [1, 2, 4, 10].

Refer to caption
Refer to caption
Figure 4. Two equivalent matching rules for the Amman tilings. As explained in the paragraph above, the matching rules dictated by the decorated pair of small and large tiles at the upper left and the matching rules dictated by the decorated pair of small and large tiles at the right both define the Ammann tilings. As explained in Section 4, the figure at the lower left shows how a small Ammann tile meets its large partner.

Hereafter we will call an Ammann-A2 tiling simply an Ammann tiling. A portion of an Amman tiling is shown in Figure 5. Two tilings TT and TT^{\prime} are said to be congruent if there is an isometry UU such that T=U(T)T^{\prime}=U(T). Much has been written on Ammann tilings both in mathematics journals, for example [4, 7] and in recreational sources, notably [8, 11]. There are uncountable many Ammann tilings, each being non-periodic. Every Ammann tiling is repetitive and every pair of Ammann tilings are locally isomorphic. A tiling TT is repetitive, also called quasiperiodic, if, for every finite patch PP of TT, there is a real number RR such that every ball of radius RR contains a patch congruent to PP. Two tilings are locally isomorphic if any patch in either tiling also appears in the other tiling.

The Ammann tilings may be defined, as an alternative to matching rules, by substitution rules, see [4, 6]. Another method of construction appears in [3] and is used in Section 7 of this paper.

Refer to caption
Refer to caption
Figure 5. Left: a circular patch of an Ammann-A2 tiling, with markings as in Figure 4. Right: the same patch, with tiles in various colors, and part of a possible continuation using faded tiles.

4. Amalgamation and the Hierarchy

Let TT be an Ammann tiling. For each small tile 𝗌\mathsf{s} in TT, there is a partner, defined as the unique large tile 𝖻\mathsf{b} in TT such that the union of 𝗌\mathsf{s} and 𝖻\mathsf{b}, called the amalgamation of 𝗌\mathsf{s} and 𝖻\mathsf{b}, is a hexagon congruent to GG. Any two distinct small tiles have distinct partners. The tiling obtained from TT by amalgamating each small tile with its partner is denoted A(T)A(T). The scaled tiling a(T):=sA(T)a(T):=sA(T) is an Ammann tiling called the amalgamation of TT.

Proposition 4.1.

If TT is an Ammann tiling, then so is a(T)a(T).

Proof.

The upper left and right panels of Figure 4 show two different decorations on the small and large tiles, call them the D1D1 and the D2D2 decorations, respectively.

Each small tile in a(T)a(T) (left tile in the right panel of Figure 4) is a (scaled by ss) large tile in TT (right tile in the left panel); and each large tile in a(T)a(T) (right tile in the right panel) is the scaled amalgamation of a small and large tile in TT (bottom of the left panel).

As previously noted, the matching rules dictated by the D1D1 and D2D2 decorations are equivalent in that they both define the Ammann tilings. Therefore if TT is an Ammann tiling, then so is a(T)a(T). ∎

Related comments on matching rules for the Ammann tilings are given in [1, 2, 4]. In particular see [1] for a discussion of replacement of markings and the notion of ghost markings. Equivalence of different sets of markings is discussed in [10].

The sequence of tilings H(T):={T,A(T),A2(T),}H(T):=\{T,A(T),A^{2}(T),\dots\}, consisting of Ammann tilings at larger and larger scales, will be referred to as the hierarchy of TT. In general, if tt is any tile at any level of the hierarchy, then tt is the non-overlapping union of tiles in TT. Denote this tiling of tt by 𝒯(t)\mathcal{T}(t). The tiling 𝒯(t)\mathcal{T}(t) is sometimes referred to as a supertile in the tilings literature; see for example [5, 7, 9]. Any large tile tt in An(T)A^{n}(T) is congruent to τn1(G)\tau^{n-1}(G). For such a large tile TT in An(T)A^{n}(T), denote 𝒯(t)\mathcal{T}(t) by 𝒯n\mathcal{T}_{n} which, by Proposition 4.2 below, is well-defined, independent of tt. The tiling 𝒯6\mathcal{T}_{6} appear in Figure 6.

Refer to caption
Figure 6. The supertile 𝒯6\mathcal{T}_{6}.
Proposition 4.2.

If pp and qq are congruent tiles at any level of the hierarchy, then 𝒯(p)\mathcal{T}(p) and 𝒯(q)\mathcal{T}(q) are congruent.

Proof.

It is true at level 11; A(T)A(T) is the non-overlapping union of a small and a large Ammann tile in a unique way. Proceeding by induction on the level, assume that the statement of the proposition is true at level n1n-1, and consider congruent tiles pp and qq of An(T)A^{n}(T). Then either (1) pp and qq are small tiles of An(T)A^{n}(T), in which case they are large tiles of An1(T)A^{n-1}(T), or (2) pp and qq are large tiles of An(T)A^{n}(T), in which case they are the amalgamation of a large and a small tile of An1(T)A^{n-1}(T). In case (1) we have that 𝒯(p)\mathcal{T}(p) and 𝒯(q)\mathcal{T}(q) are congruent by the induction hypothesis. In case (2) the decomposition of pp into a non-overlapping union of a large and small tile of An1(T)A^{n-1}(T) is the same as for qq. Moreover, by the induction hypothesis, the decompositions of large and small tiles of An1(T)A^{n-1}(T) is unique. Therefore we again have that 𝒯(p)\mathcal{T}(p) and 𝒯(q)\mathcal{T}(q) are congruent. ∎

Lemma  1 below is needed to prove a main result in Section 6. We first introduce a useful combinatorial notion. Define a planar map MM to be a 2-cell embedding of a locally finite simple graph Γ\Gamma in the plane. By locally finite is meant that the degree of each vertex is finite, and by simple is meant no loops or multiple edges. The faces of MM are the closures of the connected components of 2Γ{\mathbb{R}}^{2}\setminus\Gamma. By 22-cell embedding is meant that each face of MM is homeomorphic to a closed disk. Two planar maps MM and MM^{\prime} with underlying graphs Γ\Gamma and Γ\Gamma^{\prime}, respectively, are isomorphic if there is a graph isomorphism taking Γ\Gamma to Γ\Gamma^{\prime} that preserves faces.

A tiling TT can be considered a planar map. A vertex of TT is the intersection point of three or more distinct tiles of TT, assuming that the intersection is not empty, and an edge of TT is the intersection, if not empty or a single point, of two distinct tiles of TT. The faces of TT are the tiles of TT.

From an Ammann tiling TT, a tiling T^\widehat{T} is constructed as follows. Note that the degree (number of incident edges) of any vertex of an Ammann tiling TT is either 33 or 44. Color red each edge of TT (and its two incident vertices) that joins two vertices of degree 44. For any red vertex lying on only one red edge, remove the color red from that vertex and remove the incident edge. Let T^\widehat{T} be the tiling induced by the red edges and vertices, i.e., by removing all edges and vertices not colored red. For any face ff of T^\widehat{T}, it’s red boundary f\partial f together with all enclosed tiles of TT is a finite tiling which is denoted T(f)T(f).

Lemma 1.

With notation as above, if TT is an Ammann tiling, then T^\widehat{T} has the following properties.

  1. (1)

    If ff is a tile of T^\widehat{T}, then T(f)=𝒯4T(f)=\mathcal{T}_{4} or T(f)=𝒯5T(f)=\mathcal{T}_{5}.

  2. (2)

    T^=A5(T)\widehat{T}=A^{5}(T).

Proof.

In 𝒯n\mathcal{T}_{n}, any edge adjacent to exactly one tile is called a boundary edge. A boundary vertex is a vertex that lies on a boundary edge. In 𝒯n\mathcal{T}_{n}, color edges and vertices red by the same rule as used to color TT, but, in addition, color red any edge (and incident vertex) joining a vertex of degree 44 to a boundary vertex and color all boundary edges (and incident vertices) red. As previously, for any red vertex lying on only one red edge, remove the color red from that vertex and remove the incident edge. Denote by 𝒯n^\widehat{\mathcal{T}_{n}} the tiling induced by only the red edges and vertices. Figure 6 shows the tiling 𝒯6\mathcal{T}_{6} with red edges designated and the corresponding tiling 𝒯6^\widehat{\mathcal{T}_{6}}. Note that, for any red edge ee in 𝒯6\mathcal{T}_{6} not on the boundary, one incident tile is the reflection of the other incident tile in the line that extends ee. Also note that 𝒯6^\widehat{\mathcal{T}_{6}} has two faces, say ff and ff^{\prime} such that T(f)=𝒯4T(f)=\mathcal{T}_{4} and T(f)=𝒯5T(f^{\prime})=\mathcal{T}_{5}. We claim that the same is true for all 𝒯n,n6\mathcal{T}_{n},n\geq 6, i.e., (1) for any red edge ee of 𝒯n\mathcal{T}_{n} not on the boundary, one incident tile is the reflection of the other in the line that extends ee (and this implies that any red vertex of 𝒯n\mathcal{T}_{n}, not on the boundary, has degree 44), and (2) for any tile ff of 𝒯n^\widehat{\mathcal{T}_{n}} it is the case that T(f)=𝒯4T(f)=\mathcal{T}_{4} or T(f)=𝒯5T(f)=\mathcal{T}_{5}.

The claim will be proved by induction on nn. Assume that assertions (1) and (2) above are true for 𝒯n\mathcal{T}_{n} and 𝒯n^\widehat{\mathcal{T}_{n}}. Obtain 𝒯n+1\mathcal{T}_{n+1} by first subdividing each large tile in 𝒯n\mathcal{T}_{n} into one large and one small tile, then enlarging the resulting tiling by a factor of τ\tau. Now color the edges of 𝒯n+1\mathcal{T}_{n+1} as previously prescribed. If ee is a red edge in 𝒯n\mathcal{T}_{n} and ff and ff^{\prime} are the adjacent tiles, then either they are both small tiles in 𝒯n\mathcal{T}_{n}, and hence not subdivided, or they are both large tiles in 𝒯n\mathcal{T}_{n}, and hence subdivided in exactly the same way, so that the reflection property remains true in 𝒯n+1\mathcal{T}_{n+1}. Therefore, each red edge in 𝒯n\mathcal{T}_{n} induces either one or two red edges in 𝒯n+1\mathcal{T}_{n+1}. Let EE be this set of induced red edges and let 𝒯n+1\mathcal{T}^{\prime}_{n+1} be 𝒯n+1\mathcal{T}_{n+1}, but with only EE and the boundary edges (and incident vertices) colored red. Since each tile of 𝒯n^\widehat{\mathcal{T}_{n}} is either 𝒯4\mathcal{T}_{4} or 𝒯5\mathcal{T}_{5}, each tile in 𝒯n+1\mathcal{T}^{\prime}_{n+1} is either 𝒯5\mathcal{T}_{5} or 𝒯6\mathcal{T}_{6}. But when additional red faces are added to 𝒯n+1\mathcal{T}^{\prime}_{n+1} to obtain 𝒯n+1\mathcal{T}_{n+1}, no additional red edges are added to each 𝒯5\mathcal{T}_{5} because 𝒯5\mathcal{T}_{5} has no degree 44 internal (not on the boundary) vertices, and the additional red edges added to each 𝒯6\mathcal{T}_{6}, as shown in Figure 6, divides each 𝒯6\mathcal{T}_{6} into one 𝒯4\mathcal{T}_{4} and one 𝒯5\mathcal{T}_{5}. Our claim has now been proved.

To extend the result from 𝒯n^\widehat{\mathcal{T}_{n}} to T^\widehat{T}, let ff be any tile of T^\widehat{T}. If ff lies in the interior (no edge of ff on the boundary) of a tile in An(T)A^{n}(T) for some nn, then by the paragraphs above, T(f)T(f) is either 𝒯4\mathcal{T}_{4} or 𝒯5\mathcal{T}_{5}. So assume that there is an edge ee of ff that lies the on the boundary of some tile tnt_{n} of An(T)A^{n}(T) for all nn sufficiently large, say nn0n\geq n_{0}. Denote by TnT_{n} the subset of the tiling TT that lies in tnt_{n}. Then Tn0Tn0+1Tn0+2T_{n_{0}}\subset T_{n_{0}+1}\subset T_{n_{0}+2}\subset\cdots. The nested union nn0Tn\bigcup_{n\geq n_{0}}T_{n} is an Ammann tiling of a proper subset of the plane. But it is known [4] that this occurs only if it is a tiling of a half-space bounded by the line LL that extends ee or a quadrant of the plane bounded by two perpendicular rays L1L_{1} and L2L_{2}. Moreover, if TT is an Ammann tiling of the half plane, then the only extension to an Ammann tiling of the entire plane is obtained by reflecting TT in the line LL of TT. Similarly for a tiling obtained from a tiling TT of a quadrant by reflecting in the two perpendicular border lines L1L_{1} and L2L_{2}. In the half plane case, the edges and vertices of TT on LL will be colored red, and in the quadrant case, the edges and vertices of TT on L1L_{1} and L2L_{2} will be colored red. So again, T(f)T(f) is either 𝒯4\mathcal{T}_{4} or 𝒯5\mathcal{T}_{5}, and statement (1) of Lemma 1 is proved.

Concerning statement (2), the tiles in both A5(T)A^{5}(T) and T^\widehat{T} are congruent copies of τ5(G)\tau^{5}(G) and τ4(G)\tau^{4}(G). Let tT^t\in\widehat{T}. If 𝗌\mathsf{s} is an small tile in Aj(T),j<5,A^{j}(T),\,j<5, that is contained in tt, and 𝖻\mathsf{b} is its partner, then the shapes dictate that 𝖻\mathsf{b} is also contained in tt. Therefore T^=A5(T)\widehat{T}=A^{5}(T). ∎

5. Bounded Distortion Homeomorphism

Definition 1.

A homeomorphism hh of the plane has bounded distortion if there is a constant CC such that

|xh(x)|C|x-h(x)|\leq C

for all x2x\in{\mathbb{R}}^{2}. In this case, hh is said to have bounded distortion CC.

Lemma 2.

Let hh be a homeomorphism of the plane with bounded distortion CC. If ϕr\phi_{r} is a similarity transformation with scaling ratio rr, then

|xϕ1hϕ(x)|1rC|x-\phi^{-1}\circ h\circ\phi(x)|\leq\frac{1}{r}\,C

for all x2x\in{\mathbb{R}}^{2}.

Proof.

Let x2,y=ϕr(x),z=h(y),w=ϕr1(z)x\in{\mathbb{R}}^{2},\,y=\phi_{r}(x),\,z=h(y),\,w=\phi_{r}^{-1}(z). Then

|xϕr1hϕr(x)|\displaystyle|x-\phi_{r}^{-1}\circ h\circ\phi_{r}(x)| =|xw|=|ϕr1(y)ϕr1(z)|=1r|yz|\displaystyle=|x-w|=|\phi_{r}^{-1}(y)-\phi_{r}^{-1}(z)|=\frac{1}{r}\,|y-z|
=1r|yh(y)|1rC\displaystyle=\frac{1}{r}\,|y-h(y)|\leq\frac{1}{r}\,C

for all x2x\in{\mathbb{R}}^{2}. ∎

Let 𝕏\mathbb{X} denote the set of tilings as defined in Section 3. For a tiling T𝕏T\in\mathbb{X}, let T\partial T denote the union of the boundaries of the tiles of TT. Define a distance function on 𝕏\mathbb{X} as follows.

d(T,T)=dH(T,T).d(T,T^{\prime})=d_{H}(\partial T,\partial T^{\prime}).

where dHd_{H} is the Hausdorff distance. Intuitively, the dd-distance between tilings is small if the tilings are almost the same over the entire plane. The next corollary follows immediately from Lemma 2.

Corollary 1.

If hh a homeomorphism of bounded distortion CC and ϕr\phi_{r} is a similarity transformation with scaling ratio rr, then

d(ϕr1hϕr)(T),T)1rCd\big{(}\phi_{r}^{-1}\circ h\circ\phi_{r})(T),T\big{)}\leq\frac{1}{r}\,C

for all tilings T𝕏T\in\mathbb{X}.

If TT is an Ammann tiling and hh is a homeomorphism of the plane, then hh acts on TT. Let T=h(T)T^{\prime}=h(T) be the image of an Ammann tiling TT under a homeomorphism hh. The small and large tiles in TT^{\prime} are defined to be the images of the small and large tiles in TT. If 𝗌\mathsf{s}^{\prime} is a small tile in TT^{\prime} corresponding to a small tile 𝗌\mathsf{s} in TT, then the partner of 𝗌\mathsf{s}^{\prime} is the image under hh of the partner of 𝗌\mathsf{s}. Therefore the amalgamation a(T)a(T^{\prime}) and the hierarchy H(T)H(T^{\prime}) can be defined for TT^{\prime} exactly as they are for TT.

Theorem 1.

Let TT be an Ammann tiling, hh a bounded distortion homeomorphism of the plane, and T=h(T)T^{\prime}=h(T). Then

limkd(ak(T),ak(T))=0.\lim_{k\rightarrow\infty}d\big{(}a^{k}(T^{\prime}),a^{k}(T)\big{)}=0.
Proof.

Let H(T)={T=T0,T1,T2,}H(T)=\{T=T_{0},T_{1},T_{2},\dots\} be the hierarchy of TT and H(T)={T=T0,T1,T2,}H(T^{\prime})=\{T^{\prime}=T_{0}^{\prime},T_{1}^{\prime},T_{2}^{\prime},\dots\} the hierarchy of TT^{\prime}. Then

ak(T)\displaystyle a^{k}(T) =skTk\displaystyle=s^{k}T_{k}
ak(T)\displaystyle a^{k}(T^{\prime}) =skTk=skh(Tk)=skh(skak(T))=(skhsk)(ak(T))\displaystyle=s^{k}T^{\prime}_{k}=s^{k}h(T_{k})=s^{k}h(s^{-k}a^{k}(T))=(s^{k}\circ h\circ s^{-k})\big{(}a^{k}(T)\big{)}
=(τkhτk)(ak(T))\displaystyle=(\tau^{-k}\circ h\circ\tau^{k})\big{(}a^{k}(T)\big{)}

where we recall that τ=s1\tau=s^{-1}. By Corollary 1 we have

d(ak(T),ak(T))=d((τkhτk)(ak(T)),ak(T))Cτk.d\big{(}a^{k}(T^{\prime}),a^{k}(T)\big{)}=d\big{(}(\tau^{-k}\circ h\circ\tau^{k})\big{(}a^{k}(T)),a^{k}(T)\big{)}\leq\frac{C}{\tau^{k}}.

From this it follows that

limkd(ak(T),ak(T))=0.\lim_{k\rightarrow\infty}d\big{(}a^{k}(T^{\prime}),a^{k}(T)\big{)}=0.\qed
Remark 1.

Theorem 1 does not state that ak(T)a^{k}(T^{\prime}) becomes arbitrary close to TT as kk\rightarrow\infty. It does imply, however, that an arbitrarily large finite patch of the tiling ak(T)a^{k}(T^{\prime}) becomes arbitrarily close to a patch PkP_{k} of TT as kk\rightarrow\infty. This follows from Proposition 4.1 and the fact that any two Ammann tilings are locally isomorphic.

6. Combinatorial Amalgamation

Let 𝕋\mathbb{T} be the set of all images of Ammann tilings of the plane under homeomorphisms of the plane. Essential to Theorem 1 is being able to apply the amalgamation operator a:𝕋𝕋a:\mathbb{T}\rightarrow\mathbb{T}. Let TT be an Ammann tiling and T=h(T)T^{\prime}=h(T), where hh is a homeomorphism of the plane. In this section it is proved that a(T)a(T^{\prime}) can be determined combinatorially, without knowledge of the homeomorphism hh.

Theorem 2.

Let TT be an Ammann tiling, hh a homeomorphism of the plane and T=h(T)T^{\prime}=h(T). Then Algorithm A below computes a(T)a(T^{\prime}), without knowledge of hh.

The proof of Theorem 2 appears after a few simple lemmas. With terminology from Section 4, the following lemma is clear.

Lemma 3.

If TT is a tiling of the plane and hh is a homeomorphism of the plane, then as planar maps TT and h(T)h(T) are isomorphic.

Two planar maps M1M_{1} (left) and M2M_{2} (right) are shown in Figure 7 (ignoring the dots and arrows). Recalling the definition of 𝒯n\mathcal{T}_{n} from Section 4, the following two lemmas are apparent by inspection.

Lemma 4.

The two planar maps M1M_{1} and M2M_{2} in Figure 7 are isomorphic to 𝒯4\mathcal{T}_{4} and 𝒯5\mathcal{T}_{5}, respectively, considered as planar maps.

Lemma 5.

Each of the two planar maps in Figure 7, ignoring the arrows, has trivial automorphism group.

Refer to caption
Figure 7. Rule for locating the partner of a small tile.

Each arrow in Figure 7 points from a face bounded by a 44-cycle to an adjacent face. A consequence of Lemma 5 is that, if MM is any planar map isomorphic to M1M_{1} or M2M_{2} in Figure 7, then arrows can be uniquely assigned to MM such that there is an arrow from face f1f_{1} to f2f_{2} in MM if and only if there is an arrow from ϕ(f1)\phi(f_{1}) to ϕ(f2)\phi(f_{2}) in M1M_{1} or M2M_{2}, where ϕ\phi is the isomorphism.

Algorithm A Input: A tiling TT^{\prime} that is the image of an Amman tiling, distorted by an

unknown homeomorphism.

Output: The amalgamated tiling a(T)a(T^{\prime}).

Perform the following steps to obtain the tiling a(T)a(T^{\prime}) from TT.

  1. (1)

    Color red each edge (and its two incident vertices) of TT^{\prime} that joins two vertices of degree 44. But, for any red vertex lying on only one red edge, remove the color red from that vertex and remove the incident edge.

  2. (2)

    Let T^\widehat{T^{\prime}} be the tiling induced by just the red edges and vertices of TT^{\prime}, i.e., by removing all edges and vertices not colored red.

  3. (3)

    For every tile fT^f\in\widehat{T^{\prime}}, let T(f)T^{\prime}(f) be the set of tiles of TT^{\prime} contained in tt. Viewed as a planar map, each T(f)T^{\prime}(f) is isomorphic to one of the planar maps in Figure 7. (This is shown to be the case in the proof below of Theorem 2.) For every fT^f\in\widehat{T^{\prime}} do:

    • For every tile 𝗌\mathsf{s} of T(f)T^{\prime}(f) that is bounded by a 44-cycle, do:

      • Locate the partner 𝖻\mathsf{b} of 𝗌\mathsf{s} using the relevant arrow in Figure 7.

      • Replace the two tiles 𝗌\mathsf{s} and 𝖻\mathsf{b} in TT^{\prime} by their union to form a single tile in A(T)A(T^{\prime}), and hence in a(T)a(T^{\prime}) after scaling by s=1/τs=1/\tau.

Proof of Theorem 2.

The notation in this proof follows the notation in Section 4 just prior to Lemma 1. Let 𝗌\mathsf{s}^{\prime} be any small tile in TT^{\prime}. It must be shown that Algorithm A pairs 𝗌\mathsf{s}^{\prime} with its large partner tile in TT^{\prime}. The tile 𝗌\mathsf{s} is contained in a unique tile fA5(T)=T^f\in A^{5}(T)=\widehat{T}, the last equality by statement (2) of Lemma 1. Therefore 𝗌:=h(𝗌)\mathsf{s}^{\prime}:=h(\mathsf{s}) is contained in f:=h(f)A5(T)f^{\prime}:=h(f)\in A^{5}(T^{\prime}). Since fT^f\in\widehat{T}, we know that fh(T^)f^{\prime}\in h(\widehat{T}).

By Lemma 3, TT and h(T)h(T), as well as T^\widehat{T} and h(T^)h(\widehat{T}), are isomorphic as planar maps. Therefore, by statement (1) of Lemma 1, if 𝒯(f)\mathcal{T}(f^{\prime}) denotes the set of all tiles in TT^{\prime} contained in ff^{\prime}, then 𝒯(f)\mathcal{T}(f^{\prime}) is isomorphic to 𝒯4\mathcal{T}_{4} or 𝒯5\mathcal{T}_{5}. By Lemma 4, we know that 𝒯(f)\mathcal{T}(f^{\prime}) is isomorphic to one of the planar maps in Figure 7.

Note that the boundary of any face in TT^{\prime} is a 44-cycle if the corresponding tile is small and a 55 or 66-cycle if the corresponding tile is large. Therefore, by Lemma 5 and the remarks following it, the partner of 𝗌f\mathsf{s}^{\prime}\in f^{\prime} is uniquely determined by the arrows in Figure 7. This completes the proof of Theorem 2. ∎

Algorithm A acts on the tiling TT^{\prime} of the entire plane and hence does not run in finite time. From a practical point of view, it may be asked whether it is possible to efficiently compute the amalgamation of the distorted tiling TT^{\prime} on a finite subset of the plane, for instance on a patch containing the disk DRD_{R} of radius RR centered at the origin. We tweak Algorithm A to obtain a combinatorial Algorithm B that applies to a finite patch of a tiling. Algorithm B computes a(T)a(T^{\prime}) on DRD_{R} and runs in time quadratic in RR. This is the content of Theorem 3 below. Figure 8 illustrates Algorithm A above and Algorithm B below, showing the red edges of a distorted Ammann tiling before and after the red edges that do not belong to cycles are removed.

The notation T|RT^{\prime}|R and a(T|R)a(T^{\prime}|R) in Theorem 3 are defined as follows. If TT^{\prime} is a tiling, T|RT^{\prime}|R the set of tiles of TT^{\prime} with non-empty intersection with DRD_{R}. If TT^{\prime} is an Ammann tiling or a homeomorphic image of an Ammann tiling, the amalgamation a(T|R)a(T^{\prime}|R) is obtained by replacing every small tile tt in T|RT|R by the union of tt and its partner - even if the partner of tt has empty intersection with DRD_{R}.

Lemma 6.

Let RR be a positive real number. If TT is an Ammann tiling and hh is a homeomorphism of bounded distortion CC, then

  1. (1)

    h(A5(T))|RDR+2C+τ6h(A^{5}(T))|R\subset D_{R+2C+\tau^{6}},

  2. (2)

    the number of tiles of h(T)h(T) contained in the disk DRD_{R} is less than τπ(1545)(R+C)2.\tau\pi(\frac{15}{4}-\sqrt{5})(R+C)^{2}.

Proof.

Concerning statement (1), if xx is any point in a tile th(A5(T))|Rt^{\prime}\in h(A^{5}(T))|R, and yy is any point in tDRt^{\prime}\cap D_{R}, then by the triangle inequality and the fact that hh is of bounded distortion CC, we have

|x|\displaystyle|x| |y|+|xy|R+diam(t)=R+diam(h(t))R+diam(t)+2C\displaystyle\leq|y|+|x-y|\leq R+diam(t^{\prime})=R+diam(h(t))\leq R+diam(t)+2C
R+diam(τ5G)+2C=R+τ6+2C,\displaystyle\leq R+diam(\tau^{5}G)+2C=R+\tau^{6}+2C,

where diamdiam denotes the diameter of the set.

Concening statement (2), because hh is of bounded distortion CC, we have DRh(DR+C)D_{R}\subseteq h(D_{R+C}). Each small tile in an Ammann tiling is congruent to s2Gs^{2}G. Therefore there are less than area(DR+C)/area(s2G)=τπ(1545)(R+C)2area(D_{R+C})/area(s^{2}G)=\tau\pi(\frac{15}{4}-\sqrt{5})(R+C)^{2} Ammann tiles in DR+CD_{R+C}. ∎

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8. Left: a distorted Amman tiling with edges that connect vertices of degree 4 marked in red. Right: the same tiling after removal of red edges that do not lie on red cycles.

Algorithm B Input: A tiling TT^{\prime} that is the image of an Amman tiling, distorted by an

unknown homeomorphism of bounded distortion CC, and a positive real

number RR.

Output: A subset T′′T^{\prime\prime} of the amalgamated tiling a(T)a(T^{\prime}) that contains a(T|R)a(T^{\prime}|R).

  1. (1)

    Set R^=R+2C+τ6\widehat{R}=R+2C+\tau^{6} Perform the following steps to obtain the tiling T′′T^{\prime\prime} from the tiling T|R^T^{\prime}|\widehat{R}:

  2. (2)

    Color some edges and vertices of T|R^T^{\prime}|\widehat{R} red according to the following rules:

    • Color red each edge (and its two incident vertices) of T|R^T^{\prime}|\widehat{R} that joins two vertices of degree 44.

    • Remove the color red from any red vertex or edge not lying on a red cycle in T|R^T^{\prime}|\widehat{R}.

  3. (3)

    Let T^\widehat{T^{\prime}} be the tiling induced by the red edges and vertices of T|R^T^{\prime}|\widehat{R}.

  4. (4)

    For every tile tT^t\in\widehat{T^{\prime}}, let T(t)T^{\prime}(t) be the set of tiles of T|R^T^{\prime}|\widehat{R} contained in tt. Viewed as planar map, each T(t)T^{\prime}(t) is isomorphic to one of the planar maps in Figure 7. For every tT^t\in\widehat{T^{\prime}} do:

    • For every tile 𝗌\mathsf{s} of T(t)T^{\prime}(t) that is bounded by a 44-cycle, do:

      • Locate the partner 𝖻\mathsf{b} of 𝗌\mathsf{s} using the the relevant arrow in Figure 7.

      • Replace the two tiles 𝗌\mathsf{s} and 𝖻\mathsf{b} in T|R^T^{\prime}|\widehat{R} by their union to form a single tile in T′′T^{\prime\prime}.

Theorem 3.

Let TT be an Ammann tiling, hh a homeomorphism of the plane of bounded distortion, and T=h(T)T^{\prime}=h(T). Then Algorithm B above computes a(T|R)a(T^{\prime}|R) in time quadratic in RR, without knowledge of hh.

Proof.

In addition to what was shown in the proof of Theorem 2, it must be verified that the finite tiling T^\widehat{T^{\prime}} constructed in Algorithm B is contained in DR^D_{\widehat{R}}, which follows from statement (2) of Lemma 1 and startement (1) of Lemma 6.

Concerning the running time of Algorithm B, it is clear that the number of steps is O(n)O(n), where nn is the sum of the number of tiles in T|R^T^{\prime}|\widehat{R}. By statement (2) of Lemma 6, the number of tiles in T|R^T^{\prime}|\widehat{R} is less than

τπ(1545)(R^+C)2=τπ(1545)(R+3C+τ6)2=O(R2).\tau\pi(\frac{15}{4}-\sqrt{5})(\widehat{R}+C)^{2}=\tau\pi(\frac{15}{4}-\sqrt{5})(R+3C+\tau^{6})^{2}=O(R^{2}).\qed

7. Direct Retrieval of an Ammann tiling from a Distorted Image

Let TT be an Ammann tiling and T=h(T)T^{\prime}=h(T), where hh is a homeomorphism of the plane. In this section we extract from TT^{\prime}, without using hh, an infinite binary string θ=θ1θ2θ3\theta=\theta_{1}\,\theta_{2}\,\,\theta_{3}\cdots. A tiling T(θ)T(\theta) will be combinatorially constructed from θ\theta such that T=T(θ)T=T(\theta).

Definition 2.

Let 𝕊\mathbb{S} be the set of all pairs (T,t)(T^{\prime},t^{\prime}), where T=h(T)T^{\prime}=h(T) and t=h(t)t^{\prime}=h(t) for some Ammann tiling TT, large tile tTt\in T, and homeomorphism hh of the plane. Let {1,2}={c1c2c3:ci{1,2}for alli}\{1,2\}^{\infty}=\{c_{1}\,c_{2}\,c_{3}\cdots:c_{i}\in\{1,2\}\;\text{for all}\,i\}, i.e., the set of infinite binary stings. Also let {1,2}\{1,2\}^{\ast} denote the set of finite binary strings. To define a map g:𝕊𝕊g:\mathbb{S}\rightarrow\mathbb{S} and a function c:𝕊{1,2}c:\mathbb{S}\rightarrow\{1,2\}^{\infty}, called the code of (T,t)(T^{\prime},t^{\prime}), let (T,t)𝕋(T^{\prime},t^{\prime})\in\mathbb{T}. Define (T~,t~)=g(T,t)(\widetilde{T},\widetilde{t})=g(T^{\prime},t^{\prime}) as follows. There are two cases.

Case 1. If tt^{\prime} has a partner in TT^{\prime}, then let T~=A(T)\widetilde{T}=A(T^{\prime}), the amalgamation of TT^{\prime}, and let t~\widetilde{t} be union of tt^{\prime} and its partner, a large tile in T~\widetilde{T}. Define g(T,t)=(T~,t~)g(T^{\prime},t^{\prime})=(\widetilde{T},\widetilde{t}) and c~(T,t)=1\widetilde{c}(T^{\prime},t^{\prime})=1.

Case 2. If tt^{\prime} has no partner in TT^{\prime}, then let T~=A2(T)\widetilde{T}=A^{2}(T^{\prime}), the second amalgamation of TT^{\prime}. Then tt^{\prime} is a small tile in A(T)A(T^{\prime}) that has a partner. In T~\widetilde{T} let t~\widetilde{t} be the union of tt^{\prime} and its partner in A(T)A(T^{\prime}). Let g(T,t):=(T~,t~)g(T^{\prime},t^{\prime}):=(\widetilde{T},\widetilde{t}). Note that t~\widetilde{t} is a large tile in T~\widetilde{T}. Define c~(T,t)=2\widetilde{c}(T^{\prime},t^{\prime})=2.

Define a sequence of tilings T0=T,T1,T2,T^{\prime}_{0}=T^{\prime},T^{\prime}_{1},T^{\prime}_{2},\dots and tiles t0=t,t1,t2,t^{\prime}_{0}=t^{\prime},t^{\prime}_{1},t^{\prime}_{2},\dots, recursively by (Tn+1,tn+1)=g(Tn,tn)(T^{\prime}_{n+1},t^{\prime}_{n+1})=g(T^{\prime}_{n},t^{\prime}_{n}). Then define the code of (T,t)(T^{\prime},t^{\prime}) to be the binary string

c(T,t)=c~(T0,t0)c~(T1,t1)c~(T3,t3),c(T^{\prime},t^{\prime})=\widetilde{c}(T^{\prime}_{0},t^{\prime}_{0})\;\widetilde{c}(T^{\prime}_{1},t^{\prime}_{1})\;\widetilde{c}(T^{\prime}_{3},t^{\prime}_{3})\cdots,

and note that, as done in Section 6, the amalgamation used to determine the code c(T,t)c(T^{\prime},t^{\prime}) of TT^{\prime} can be done combinatorially without knowing the homeomorphism.

Theorem 4 below states that c(T,t)c(T^{\prime},t^{\prime}) completely determines the original Ammann tiling TT, and can be used to generate TT. An intuitive way to see this is to start with a patch P0P_{0} consisting of a single large Ammann tile tt. Let c(T,t)=c1c2c(T,t)=c_{1}\,c_{2}\,\cdots. If c1=1c_{1}=1, embed P0P_{0} in a patch P1P_{1} as shown at the start of the bottom row in Figure 9; if c1=2c_{1}=2, embed P0P_{0} in a patch P1P_{1} as at the start of the top two rows in Figure 9. Continue in this way to get a nested sequence of patches as in Figure 9. The nested union is TT.

Remark 2.

It is well known that there are a couple of special cases where the nested union {Pn;n0}\cup\{P_{n};n\geq 0\} is a tiling that fills a half plane or a quadrant of the plane. Therefore, in this section such tilings will also be classified as Ammann tilings. If TT is an Ammann tiling of the half plane, then a tiling of the entire plane satisfying the matching rules can be obtained by reflecting TT in the border line of TT. These tilings will not be classified as Amman tilings in this section. Similarly for a tiling obtained from a tiling TT of a quadrant by reflecting in the two border rays.

A more rigorous formulation of the above method of retrieving the original tiling from the code uses the following combinatorial construction. Recall that s=1/τs=1/\tau, where τ\tau is the square root of the golden ratio. The fact that the golden bee GG is the non-overlapping union of two small similar copies of itself can be stated more precisely as follows.

G=f1(G)f2(G),G=f_{1}(G)\cup f_{2}(G),

where f1(G)f_{1}(G) and f2(G)f_{2}(G) are non-overlapping and

(7.1) f1(xy)=(0ss0)(xy)+(s0),f2(xy)=(s200s2)(xy)+(01).f_{1}\begin{pmatrix}x\\ y\end{pmatrix}=\begin{pmatrix}0&-s\\ s&0\end{pmatrix}\begin{pmatrix}x\\ y\end{pmatrix}+\begin{pmatrix}s\\ 0\end{pmatrix},\qquad f_{2}\begin{pmatrix}x\\ y\end{pmatrix}=\begin{pmatrix}s^{2}&0\\ 0&-s^{2}\end{pmatrix}\begin{pmatrix}x\\ y\end{pmatrix}+\begin{pmatrix}0\\ 1\end{pmatrix}.
Refer to caption
Figure 9. The starts of three possible sequences of embeddings, from left to right, as described in the text. The initial large tile, the patch P0P_{0}, is shown shaded and has a fixed location in the plane.

For σ=σ1σ2σk{1,2}\sigma=\sigma_{1}\,\sigma_{2}\,\dots\,\sigma_{k}\in\{1,2\}^{*} let

e(σ):=i=1kσie(\sigma):=\sum_{i=1}^{k}\sigma_{i}

and let σ:=σk\sigma_{*}:=\sigma_{k} denote the last term in the string σ\sigma. We introduce the notation

fσ\displaystyle f_{\sigma} :=fσ1fσ2fσk\displaystyle:=f_{\sigma_{1}}\circ f_{\sigma_{2}}\circ\cdots\circ f_{\sigma_{k}}
fσ\displaystyle f_{-\sigma} :=fσ11fσ21fσk1.\displaystyle:=f_{\sigma_{1}}^{-1}\circ f_{\sigma_{2}}^{-1}\circ\cdots\circ f_{\sigma_{k}}^{-1}.

and

θ|k:=θ1θ2θk\theta|k:=\theta_{1}\,\theta_{2}\,\dots\theta_{k}

for θ=θ1θ2,{1,2}\theta=\theta_{1}\,\theta_{2},\cdots\in\{1,2\}^{\infty}. By convention, θ|0\theta|0 is the empty string \emptyset; e()=0e(\emptyset)=0, and ff_{\emptyset} is the identity.

Let

W(θ,k):={σ{1,2}:e(σ)>e(σ)e(θ|k)0}.W(\theta,k):=\{\sigma\in\{1,2\}^{*}\,:\,e(\sigma_{*})>e(\sigma)-e(\theta|k)\geq 0\}.

Intuitively, the finite strings σW(θ,k)\sigma\in W(\theta,k) are those whose sum differs from the sum of the elements of θ|k\theta|k by zero or one. This forces each tile t(θ,k,σ)t(\theta,k,\sigma) in Definition 3 below to be congruent to either a large or small Ammann tile.

Definition 3.

For each θ{1,2}\theta\in\{1,2\}^{\infty}, a tiling T(θ)T(\theta) is constructed in three steps.

  1. (1)

    A single tile. For each integer k1k\geq 1 and each σ[N]\sigma\in[N]^{\ast}, construct a single tile t(θ,k,σ)t(\theta,k,\sigma) that is similar to GG:

    t(θ,k,σ):=(f(θ|k)fσ)(sG).t(\theta,k,\sigma):=(f_{-(\theta\,|\,k)}\circ f_{\sigma})(sG).
  2. (2)

    A patch of tiles. Form a patch T(θ,k)T(\theta,k) of tiles given by:

    T(θ,k):={t(θ,k,σ):σW(θ,k)}.T(\theta,k):=\left\{t(\theta,k,\sigma)\,:\,\sigma\in W(\theta,k)\right\}.
  3. (3)

    A tiling. The tiling T(θ){T}(\theta), depending only on θ{1,2}\theta\in\{1,2\}^{\infty}, is the union of the patches T(θ,k)T(\theta,k), which is known [3] to be a nested union:

    T(θ):=k1T(θ,k).{T}(\theta):=\bigcup_{k\geq 1}T(\theta,k).

The set {1,2}\{1,2\}^{\infty} is called the parameter set. For each parameter θ\theta, the tiling T(θ)T(\theta) is called the θ\mathbf{\theta}-tiling. Modulo Remark 2, a θ\theta-tiling is an Ammann tiling.

Theorem 4.

Let T=h(T)T^{\prime}=h(T) be the image of Ammann tiling TT under a homeomorphism hh of the plane. Let tt^{\prime} be any large tile in TT^{\prime}, and let θ:=c(T,t)\theta:=c(T^{\prime},t^{\prime}) be the code of (T,t)(T^{\prime},t^{\prime}). Then T=T(θ)T=T(\theta), independent of which large tile tTt^{\prime}\in T^{\prime} is chosen.

Proof.

Let t=h1(t)Tt=h^{-1}(t^{\prime})\in T, and note that c(T,t)=c(T,t)c(T^{\prime},t^{\prime})=c(T,t). Let θ:=c(T,t)\theta:=c(T,t). It must be shown that T(θ)=TT(\theta)=T for any large tile tTt\in T. Since an Ammann tiling is determined by its code, it suffices to show that, for any large tile tTt\in T, there is a large tile pT(θ)p\in T(\theta) such that that c(T(θ),p)=θc(T(\theta),p)=\theta.

A tile t(θ,k,σ)T(θ)t(\theta,k,\sigma)\in T(\theta) is a large tile congruent to sGsG in T(θ)T(\theta) if e(σ)e(θ|K)=0e(\sigma)-e(\theta\,|\,K)=0 and a small tile congruent to s2Gs^{2}G if e(σ)e(θ|K)=1e(\sigma)-e(\theta\,|\,K)=1. From the definitions, this implies that t(θ,k,σ)T(θ)t(\theta,k,\sigma)\in T(\theta) is a small tile if and only if σ=2\sigma_{*}=2, i.e., σ=σ1σ2σj1 2\sigma=\sigma_{1}\,\sigma_{2}\cdots\sigma_{j-1}\,2, and the tile t(θ,k,ω)t(\theta,k,\omega), where ω=σ1σ2σj1 1\omega=\sigma_{1}\,\sigma_{2}\cdots\sigma_{j-1}\,1, is the large partner of t(θ,k,σ)t(\theta,k,\sigma). Let p=t(θ,k,σ)p=t(\theta,k,\sigma), where σ=σ1σ2σj\sigma=\sigma_{1}\,\sigma_{2}\cdots\sigma_{j}, be an arbitrary tile in T(θ)T(\theta). Then winding through Definition 3, the code of c(T(θ),p)c(T(\theta),p) is given by c(T(θ),p)=σjσj1σ1θk+1θk+2c(T(\theta),p)=\sigma_{j}\,\sigma_{j-1}\cdots\sigma_{1}\,\theta_{k+1}\,\theta_{k+2}\cdots.

Consider two cases. If the original large tile tTt\in T has a partner, then θ1=1\theta_{1}=1, i.e., c(T,t)=1θ2θ3c(T,t)=1\theta_{2}\,\theta_{3}\cdots. In T(θ)T(\theta), let p=t(θ,1,1)p=t(\theta,1,1). Then the code of c(T(θ),p)c(T(\theta),p) is given by c(T(θ),p)=1θ2θ3c(T(\theta),p)=1\theta_{2}\,\theta_{3}\cdots. If the original large tile tTt\in T has no partner, then θ1=2\theta_{1}=2, i.e., c(T,t)=2θ2θ3c(T,t)=2\theta_{2}\,\theta_{3}\cdots. In T(θ)T(\theta), let p=t(θ,1,2)p=t(\theta,1,2). Then the code of c(T(θ),p)c(T(\theta),p) is given by c(T(θ),p)=2θ2θ3c(T(\theta),p)=2\theta_{2}\,\theta_{3}\cdots. In either case c(T(θ),p)=θc(T(\theta),p)=\theta. ∎

Remark 3.

An arbitrarily large patch of the original tiling TT can be obtained in finite time and efficiently from the distorted tiling TT^{\prime} by first using the combinatorial amalgamation Algorithm B to obtain a finite concatenation θ|k\theta|k of a code θ\theta for TT^{\prime} and then by using the procedure in Definition 3 to construct the patch T(θ,k)T(\theta,k).

aknowlegements

We thank Vanessa Robbins for interesting conversations in relation to this work.

References

  • [1] S. Akiyama, A note on aperiodic Ammann tiles, Discrete and Computational Geometry,48 (2012), 702-710.
  • [2] R. Ammann, B. Grunbaum, G. C. Shephard, Aperiodic tiles, Discrete and Computational Geometry 8(1992), 1-25.
  • [3] M. F. Barnsley, A. Vince, Self-similar polygonal tilings, Amer. Math. Monthly 124 (2017), 905-921.
  • [4] B. Durand, A. Shen, N. Vereshchagin, On the structure of Ammann-A2 tilings, Discrete and Computational Geometry 63 (2020), 577-606.
  • [5] N. P. Frank, L. Sadun, Fusion tilings with infinite local complexity, Topology Proceedings 43 (2012),
  • [6] D. Frettlöh, E. Harriss, F. Gähler: Tilings encyclopedia, https://tilings.math.uni-bielefeld.de/
  • [7] B. Grünbaum and G. S. Shephard, Tilings and Patterns, W. H. Freeman and Company, New York, 1987.
  • [8] M. Gardner, Extraordinary nonperiodic tilings that enriches the theory of tiles, Scientific American 236 (1977), 110-121.
  • [9] C. Goodman-Strauss, Matching rules and substitution tilings, Annals of Mathematics 147 (1998), 181-223.
  • [10] A. Korotin, Local rules for self-similar tilings of the plane by polygons, Honours thesis (in Russian), 2016.
  • [11] M. Senechal, The mysterious Mr. Ammann, Math. Intelligencer 26 (2004), 10-21.
  • [12] J. H. Schmerl, Dividing a polygon into two similar polygons, Discrete Math., 311 (2011), 220-231.
  • [13] P. J. Steinhardt, The Second Kind of Impossible: The Extraordinary Quest for a New Form of Matter, Simon & Shuster, 2018.
  • [14] D. Weaire, N. River, Soap, cells and statistics - random patterns in two dimensions, Contemporary Physics, 50 (2009), 199-239.