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

Purely local growth of a quasicrystal

Thomas Fernique    Ilya Galanov
Abstract

Self-assembly is the process in which the components of a system, whether molecules, polymers, or macroscopic particles, are organized into ordered structures as a result of local interactions between the components themselves, without exterior guidance. In this paper, we speak about the self-assembly of aperiodic tilings. Aperiodic tilings serve as a mathematical model for quasicrystals - crystals that do not have any translational symmetry. Because of the specific atomic arrangement of these crystals, the question of how they grow remains open. In this paper, we state the theorem regarding purely local and deterministic growth of Golden-Octagonal tilings using the algorithm initially introduced in [FG20]. Showing, contrary to the popular belief, that local growth of aperiodic tilings is possible.

1 Introduction.

Quasicrystals are physical solids with aperiodic atomic structure and symmetries forbidden in classical crystallography. They were discovered by Dan Shechtman in 1982 who subsequently won the Nobel prize for his discovery in 2011.

It took Shechtman two years to publish his discovery, the paper [SBGC84] with the report of analuminum-magnesium alloy with, as it was shown via electron diffraction, 1010-fold or icosahedral symmetry, i.e., rotational symmetry with angle 2π102\pi\over 10 which is forbidden in periodic crystals by crystallographic restriction theorem. At the same time, sharp Bragg peaks of the diffraction pattern suggested long-range order in the new-found material. That was a clear violation of the fundamental principles of solid-state physics at the time. Two mentioned observations suggested that atoms in the material are structured in a non-periodic manner. The paper published in December 1984 coauthored by Levine and Steinhardt [LS84] named the phenomenon as a quasicrystallinity and the novel substance as a quasiperiodic crystal or quasicrystal.

To represent the peculiar atomic structure of quasicrystals Levine and Steinhardt in [LS84] proposed the Penrose tilings [Pen74]. A tiling is a covering of a Euclidian plane by given geometric shapes without gaps and overlaps. The set of basic shapes is called a prototile set and the elements are called prototiles or tiles.

Definition 1.

A prototile set is called aperiodic if admits only aperiodic tilings.

Many of the properties of Penrose tilings can be derived from a so-called cut-and-project scheme. This method, first introduced by DeBruijn [DB81] in 1981, is based on the discovery that Penrose tilings can be obtained by projecting certain points from higher dimensional lattices to a 22-dimensional plane. The method was subject of many generalizations, see [BG13] for a comprehensive overview. Apart from Penrose tilings, it allows generating various other aperiodic tilings with symmetries forbidden in classical crystals but found in quasiperiodic ones.

Moreover, DeBruijn showed that the collection of 33-patterns of a Penrose tiling uniquely determines the cutting slope in 55-dimensional space. In general, the ability to fix the slope in higher dimensional space solely through finite patterns is referred to as local rules. This subject was studied my many researchers [Bee82, Lev88, Soc90, LPS92, Le95, Le97, Kat95, LP95, BF13, BF15a, BF15b, BF17, BF20]. In particular, a complete characterization of local rules for planar octagonal tilings is given in [BF17]. Up to date aperiodic tilings and tilings generated via cut-and-project method is one of the main mathematical models to represent the atomical structure of a quasicrystal.

The discovery of Schechtman inspired other groups to search for quasicrystals. Within a few years, there had been reported quasicrystals with various other forbidden symmetries, including decagonal [Ben85], pentagonal [BH86], octagonal [WCHK87], and dodecagonal [INF85]. Despite the abundance of quasicrystals synthesized in labs, there was not a consensus whether quasicrystallinity is a fundamental state of matter or the quasicrystals appear only as metastable phases under very specific (and unnatural) conditions.

The common view was that aperiodic atomic structure is too complicated to be stable. Roger Penrose once said “For this reason, I was somewhat doubtful that nature would actually produce such quasicrystalline structures spontaneously. I couldn’t see how nature could do it because the assembly requires non-local knowledge”. The discussion on what governs the stability of quasicrystals is still ongoing. Possible stability mechanisms include energetic stabilization and entropy stabilization, see [DB06] for an overview.

Energy stabilization scenario suggests that quasicrystals can indeed be a state of minimal energy of the system, as in the case of classical crystals, and that short-range atomic interactions suffice to provide an aperiodically ordered structure. This is the case where the tilings with local rules are used as the main model.

Entropy stabilization scenario suggests that quasicrystals are always metastable phases and that aperiodic atomic order is governed by phason flips (a local rearrangement of atoms that leave the free energy of the system unchanged) and structural disorder even if the state is not energetically preferred. This is the case of random tiling model [LH99], [CKP00].

Energy stabilization scenario would allow quasicrystals to be formed naturally. For instance, in [NINE15] the growth of quasicrystals was directly observed with high-resolution transmission electron microscopy. Edagawa and his team produced a decagonal quasicrystal consisting of Al70.8Ni19.7Co9.5\textrm{Al}_{70.8}\textrm{Ni}_{19.7}\textrm{Co}_{9.5}. The growth process featured frequent errors-and-repair procedures and maintained nearly perfect quasiperiodic order at all times. Repairs, as concluded, were carried via phason flips, which is qualitatively different from the ideal growth models.

Another argument supporting the theory that quasicrystals can be stabilized via short-range interactions was given by Onoda et al in [OSDS88] [OSDS88, Soc91, HSS16]. They found an algorithm for growing a perfect Penrose tiling around a certain defective seed using only the local information. A defective seed is a pattern made from Penrose rhombuses which is not a subset of any Penrose tiling. The finding broke the belief that non-local information is essential for building quasicrystalline types of structure.

Crystals can grow, both classic ones and the quasicrystals, local interactions between particles guide them into their respected places along the crystal lattice. Let us view the growth of crystals from the viewpoint of pure geometry. The question transforms into the following: how to algorithmically assemble the atomic structure of a growing crystal using only the local information? Classical crystals exhibit the unit cell which makes the question trivial: as long as we can see an instance of the unit cell, we know the structure. The question becomes much harder as we proceed to quasicrystals as there is no translational symmetry and no unit cell.

Here we aim to understand whether it is possible to grow an aperiodic tiling via a local self-assembly algorithm. The meaning of the locality constraint is the following: at each step, the algorithm must have access only to a finite neighborhood around a randomly chosen vertex of a seed - a finite pattern of a tiling we are trying to expand. Given the local neighborhood, the algorithm must identify the set of vertices which are to be added to the seed (or it may decide that there is not enough information to add a tile or a vertex and do nothing). Finally, the algorithm must not store any information between the steps.

The algorithm we propose, initially announced in [FG20], is a generalization of the OSDS rules algorithm. Similarly, we add the vertices if and only if they are uniquely determined by the neighbourhood but, and that is the crucial difference, we allow the forced vertices to be distanced from the growing pattern and not share any edges with it. This allows us to jump over the undefined areas and, as simulations suggest, and grow the cylinder set of a given pattern or, as it was called in [Gar89], the empire.

In this paper we prove the algorithm as able to grow particular aperiodic tilings with local rules namely the Golden-Octagonal tilings initially defined in [BF15b].

Theorem.

For Golden-Octagonal tilings, for any finite pattern 𝒫\mathcal{P}, there exists a finite seed 𝒮𝒫\mathcal{S}\supset\mathcal{P} and a growth radius rr, such that the self-assembly Algorithm 1 builds the cylinder set (or empire) of 𝒫\mathcal{P}.

2 Definitions and the main result.

In this section we define and discuss properties of cut-and-project method, arguably the most versatile method to generate and to study aperiodic tilings. We define notion of tilings with local rules and introduce our self-assembly algorithm. In the end of the section, we formalize the mail result.

Octagonal tilings are simply the tilings made of rhombuses with four distinct edge-directions. This gives us six rhombus prototiles in total:

Definition 2.

Consider the prototile set:

{λvi+μvj,0λ,μ1},0i,j3,ij,\{\lambda\vec{v_{i}}+\mu\vec{v_{j}},\quad 0\leq\lambda,\mu\leq 1\},\qquad 0\leq i,j\leq 3,\qquad i\neq j,

where vi\vec{v_{i}} and vj\vec{v_{j}} are noncollinear vectors in 2\mathbb{R}^{2}. Any tiling of 2\mathbb{R}^{2} with the above prototile set is called octagonal.

Our to-go method for generating octagonal tilings is canonical cut-and-project scheme. Here we are closely following chapter 77 of [BG13].

Here we define and discuss properties of canonical cut-and-project method, our to-go method to generate octagonal tilings. Here we closely follow Chapter 77 of [BG13].

Definition 3.

A ndn\to d cut-and-project scheme consists of a physical space Ed{E\simeq\mathbb{R}^{d}}, an internal space EndE^{\perp}\simeq\mathbb{R}^{n-d}, a lattice n\mathbb{Z}^{n} in E×E=nE\times E^{\perp}=\mathbb{R}^{n} and the two natural projections π:nE\pi:\mathbb{R}^{n}\to E and π:nE\pi^{\prime}:\mathbb{R}^{n}\to E^{\perp} along with conditions that π|n\pi|_{\mathbb{Z}^{n}} is injective and that π\pi^{\prime} is dense in EE^{\perp}.

d{\mathbb{R}^{d}}d×nd{\mathbb{R}^{d}\times\mathbb{R}^{n-d}}nd{\mathbb{R}^{n-d}}π(n){\pi(\mathbb{Z}^{n})}n{\mathbb{Z}^{n}}π(n){\pi^{\prime}(\mathbb{Z}^{n})}\supsetπ\scriptstyle{\pi}π\scriptstyle{\pi^{\prime}}\supset\supsetdense\scriptstyle{\leavevmode\nobreak\ dense}\scriptstyle{*}11\scriptstyle{1-1} (1)

There is a bijection between π(n)\pi(\mathbb{Z}^{n}) and π(n)\pi^{\prime}(\mathbb{Z}^{n}), hence there is a well-defined map called the star map between EE and EE^{\perp}:

xx:=π((π|n)1(x)).x\to x^{*}:=\pi^{\prime}\left((\pi_{|\mathbb{Z}^{n}})^{-1}(x)\right).

A bounded subset WW of internal space with non-empty interior is called a window or acceptance domain. In the canonical cut-and-project scheme, the window is chosen to be W=π([0,1]n)W=\pi([0,1]^{n}). The window acts as a filter for points which are to be projected. Using the star map, the set of vertices of a tilings or the projection set can be written as

Λ(W):={xπ(n)xW}.\Lambda(W):=\{x\in\pi(\mathbb{Z}^{n})\mid x^{*}\in W\}.

Summary of a ndn\to d canonical cut-and-project scheme is depicted in the diagram 1.

Definition 4.

A lift is an injective mapping of vertices of a pattern to a subset of n\mathbb{Z}^{n} done in the following manner. Let {vi}i=0n\{v_{i}\}_{i=0}^{n} be set of edges of a rhombus tiling 𝒯\mathcal{T} up to a translation. First, we map each viv_{i} to a basis vector eie_{i} of n\mathbb{Z}^{n}. Afterward, an arbitrary vertex is mapped to the origin. Then a vertex x=i=1naivix=\sum_{i=1}^{n}a_{i}v_{i} of a pattern is lifted to i=14aiei\sum_{i=1}^{4}a_{i}e_{i}.

Thus vertices of every octagonal tiling can be lifted to 4\mathbb{Z}^{4}. Since we are interested in tilings with local rules, first, we pick a subset of octagonal tilings whose lifted vertices are close to a plane, in other words the ones which can be seen as digitizations of surfaces in higher dimensional spaces:

Definition 5.

An octagonal tiling is called planar if there exists a 2-dimensional affine plane EE in 4\mathbb{R}^{4} such that the tiling can be lifted into the stripe E+[0,1]4E+[0,1]^{4}. Then EE is called the slope of the tiling.

Now, for planar octagonal tilings we define the notion of local rules using an rr-altas, the collection of all rr-patterns, as follows:

Definition 6.

Vertices of a tiling which are at most rr edges away from a given vertex xx are called the rr-pattern of xx.

Definition 7.

The set of all rr-patterns of a tiling up to a translation is called the rr-atlas and denoted by 𝒜𝒯(r)\mathcal{A}_{\mathcal{T}}(r).

Definition 8.

A planar tiling 𝒯\mathcal{T} is said to admit local rules if there exists r>0r>0 such that, any rhombus tiling 𝒯\mathcal{T}^{\prime} with 𝒜𝒯(r)𝒜𝒯(r)\mathcal{A}_{\mathcal{T}^{{}^{\prime}}}(r)\subset\mathcal{A}_{\mathcal{T}}(r), 𝒯\mathcal{T}^{{}^{\prime}} is also a planar tiling with the slope parallel to the slope of 𝒯\mathcal{T}.

Definition 9.

Consider a patch 𝒫\mathcal{P} of a planar octagonal tiling with slope EE and window WW. The cylinder of 𝒫\mathcal{P}, denoted by [𝒫][\mathcal{P}], is the intersection of all the planar octagonal tilings which contain π𝒫\pi\mathcal{P} as a subset and whose slope is a translation of EE. We denote by W(𝒫)W(\mathcal{P}) the intersection of all the translations of WW containing π𝒫\pi^{\prime}\mathcal{P}.

One has W(𝒫)=W([𝒫])W(\mathcal{P})=W([\mathcal{P}]). The vertices of π[𝒫]\pi^{\prime}[\mathcal{P}] are uniformly dense in W([𝒫])W([\mathcal{P}]). Given a seed 𝒮\mathcal{S}, nothing more than [𝒮][\mathcal{S}] can be grown by a deterministic algorithm no matter whether it is local or not.

Data: Growth radius rr, rr-atlas 𝒜\mathcal{A} and an seed pattern 𝒮=𝒮0\mathcal{S}=\mathcal{S}_{0}
for k0k\geq 0 do
 𝒮k+1𝒮k\mathcal{S}_{k+1}\leftarrow\mathcal{S}_{k};
 for xx vertex of 𝒮k\mathcal{S}_{k} do
    𝒱x\mathcal{V}_{x}\leftarrow vertices of 𝒮k\mathcal{S}_{k} within distance rr of xx;
    𝒜x\mathcal{A}_{x}\leftarrow patterns in 𝒜\mathcal{A} translated by xx which contain 𝒱x\mathcal{V}_{x};
    x\mathcal{F}_{x}\leftarrow the vertices in common for all patterns in 𝒜x\mathcal{A}_{x};
    𝒮k+1𝒮kx\mathcal{S}_{k+1}\leftarrow\mathcal{S}_{k}\cup\mathcal{F}_{x};
    
   end for
 
end for
Algorithm 1 The self-assembly algorithm.

Now we are prepared to state the main result:

Theorem 1.

For Golden-Octagonal tilings, for any finite pattern 𝒫\mathcal{P} centered at the origin, there exists a seed 𝒮𝒫\mathcal{S}\supset\mathcal{P} and a growth radius rr, such that at step hh of the self-assembly Algorithm 1, it builds at least [𝒫]B0(h)[\mathcal{P}]\cap B_{0}(h).

3 Preliminaries

First, in this section, we define the notion of subperiod, one of the most useful tools which allow us to describe and characterize octagonal cut-and-project tilings with local rules. Then we state several lemmas in preparation to prove the main result in the next section.

3.1 Subperiods

Definition 10.

For a planar 424\to 2 tiling with a slope EE, a subperiod of type i{1,2,3,4}i\in\{1,2,3,4\} is the smallest non-zero vector in 4\mathbb{R}^{4} which belongs to EE and which has exactly 33 integer coordinates along with an irrational ii-th coordinate.

In this paper our focus is mainly on a particular family of octagonal tilings: the Golden-Octagonal tilings are the planar 424\to 2 tilings generated via canonical cut-and-project method whose slope EE is generated by:

u=(10φφ),v=(01φ1),u=\begin{pmatrix}-1\\ 0\\ \varphi\\ \varphi\end{pmatrix},\quad v=\begin{pmatrix}0\\ 1\\ \varphi\\ 1\end{pmatrix},

where φ\varphi is the golden ratio. Its set of subperiods {qi}\{q_{i}\} is written as:

q1=(1φ011),q2=(1φ10),q3=(01φ1),q4=(1101φ).q_{1}=\begin{pmatrix}1-\varphi\\ 0\\ 1\\ 1\end{pmatrix},\quad q_{2}=\begin{pmatrix}1\\ \varphi\\ 1\\ 0\end{pmatrix},\quad q_{3}=\begin{pmatrix}0\\ 1\\ \varphi\\ 1\end{pmatrix},\quad q_{4}=\begin{pmatrix}1\\ 1\\ 0\\ 1-\varphi\end{pmatrix}.

Consider a subperiod qiq_{i} of a planar tiling, we denote qi\lfloor q_{i}\rfloor (respectively qi\lceil q_{i}\rceil) an integer vector which is equal to qiq_{i} everywhere except for the irrational ii-th coordinate, instead of which we take its floor (respectively ceiling). For example, for Golden-Octagonal tiling expressions for q2\lfloor q_{2}\rfloor and q2\lceil q_{2}\rceil are written as

q2=(1φ10)=(1110),q2=(1φ10)=(1210).\lfloor q_{2}\rfloor=\begin{pmatrix}1\\ \lfloor\varphi\rfloor\\ 1\\ 0\end{pmatrix}=\begin{pmatrix}1\\ 1\\ 1\\ 0\end{pmatrix},\quad\lceil q_{2}\rceil=\begin{pmatrix}1\\ \lceil\varphi\rceil\\ 1\\ 0\end{pmatrix}=\begin{pmatrix}1\\ 2\\ 1\\ 0\end{pmatrix}.

For a point A4A\in\mathbb{Z}^{4}, πAW\pi^{\prime}A\in W, it is obvious that either π(A+qi)\pi^{\prime}(A+\lfloor q_{i}\rfloor) or π(A+qi)\pi^{\prime}(A+\lceil q_{i}\rceil) is in WW as well. Iterating these sort-of-speech jumps along a subperiod gives us the following definition:

Definition 11.

For a planar octagonal tiling 𝒯\mathcal{T} with local rules with a window WW and vertex X𝒯X\in\mathcal{T}, we define a subperiod line of vertex XX of type ii as the set:

Qi(X)={y4|πyW,yj=xj+qj,j{0,1,2,3}{i}}.Q_{i}(X)=\{y\in\mathbb{Z}^{4}\leavevmode\nobreak\ |\leavevmode\nobreak\ \pi^{\prime}y\in W,\quad y_{j}=x_{j}+q_{j},\quad j\in\{0,1,2,3\}\setminus\{i\}\}.
Refer to caption
Refer to caption
Figure 1: A pattern of Golden-Octagonal tiling is on the left and the same pattern projected to perpendicular space is on the right. Dotted arrows represent qi\lfloor q_{i}\rfloor, solid arrows - qi\lceil q_{i}\rceil. Subperiods lines of a vertex in the center are depicted with different colors.
Definition 12.

Consider a pattern 𝒫\mathcal{P}, we say that a vertex X𝒫X\in\mathcal{P} sees its ii-th subperiod line if there exists r>0r>0 such that:

BX(r)𝒫Qi(X).B_{X}(r)\cap\mathcal{P}\cap Q_{i}(X)\neq\emptyset.

We say that a vertex XX sees NN vertices along its ii-th subperiod if there exists r>0r>0 such that:

𝐜𝐚𝐫𝐝(BX(r)𝒫Qi(X))N.\mathbf{card}(B_{X}(r)\cap\mathcal{P}\cap Q_{i}(X))\geq N.

A notable feature of the vertices of canonical cut-and-project tilings is that after the projection to the perpendicular space they are uniformly distributed in the window [Els85],[Sch98], [BG13], [HKSW14]. In this paper, we use the weak version of mentioned result addressing only the distribution of vertices along subperiod lines:

Lemma 1.

Consider a planar octagonal tiling 𝒯\mathcal{T} with at least one subperiod, a vertex A𝒯A\in\mathcal{T} and let Qi(A)Q_{i}(A) be any of its subperiod lines. For any open line segment (B,C)W(B,C)\subset W belonging to the line which contains πQi(A)\pi^{\prime}Q_{i}(A), there exists r>0r>0 inversely proportional to the length of the segment (B,C)(B,C), such that in rr-neighborhood of AA there exists a vertex XQi(A)X\in Q_{i}(A) such that πX(B,C)\pi^{\prime}X\in(B,C).

3.2 Coincidences of subperiods

The following lemma highlights the interconnection between subperiods of different types in Golden-Octagonal tilings. We note that the lemma can be generalized to all the tilings with local rules. However, in this paper, we only prove it for Golden-Octagonal tilings:

Lemma 2.

For Golden-Octagonal tiling, given RR\in\mathbb{N}, there exists ll\in\mathbb{N}, such that for any two integer open line segments (A,A+ei)(A,A+e_{i}), (B,B+ej),ij(B,B+e_{j}),i\neq j, with the following properties: their orthogonal projections to perpendicular space intersect at some point XWX\in W and AB1<R\|A-B\|_{1}<R, there is another open line segment (C,C+ek)(C,C+e_{k}), ki,jk\neq i,j, whose projection intersect both of the first two intervals in XX and AC1<lR\|A-C\|_{1}<lR. Moreover, each of the three intervals has an endpoint that projects into WW.

Proof.

Let {i,j,k}={1,2,3}\{i,j,k\}=\{1,2,3\}, without loss of generality let

A=(0,0,0,0),B=(b1,b2,b3,b4),b1,b2,b3,b4.A=(0,0,0,0),\quad B=(b_{1},b_{2},b_{3},b_{4}),\quad b_{1},b_{2},b_{3},b_{4}\in\mathbb{Z}.

By our assumption, there are points X(A,A+e1)X\in(A,A+e_{1}) and Y(B,B+E2)Y\in(B,B+E_{2}) written as:

X=(x,0,0,0),Y=(b1,y,b3,b4),x,y,X=(x,0,0,0),\quad Y=(b_{1},y,b_{3},b_{4}),\quad x,y\in\mathbb{R}\setminus\mathbb{Z},

such that πX=πY=X\pi^{\prime}X=\pi^{\prime}Y=X. Therefore, YXEY-X\in E and:

b1x\displaystyle b_{1}-x =λ\displaystyle=-\lambda
y\displaystyle y =μ\displaystyle=\mu
b3\displaystyle b_{3} =λφ+μφ\displaystyle=\lambda\varphi+\mu\varphi
b4\displaystyle b_{4} =λφ+μ\displaystyle=\lambda\varphi+\mu

The last two equations yield:

λ\displaystyle\lambda =b3+b4φ\displaystyle=-b_{3}+b_{4}\varphi
μ\displaystyle\mu =(b3b4)φ\displaystyle=(b_{3}-b_{4})\varphi

The first two equations then yield:

x\displaystyle x =b1b3+b4φ\displaystyle=b_{1}-b_{3}+b_{4}\varphi
y\displaystyle y =(b3b4)φ\displaystyle=(b_{3}-b_{4})\varphi

Therefore, XX and YY can be written as

X=(b1b3+b4φ,0,0,0),Y=(b1,(b3b4)φ,b3,b4).X=(b_{1}-b_{3}+b_{4}\varphi,0,0,0),\quad Y=(b_{1},(b_{3}-b_{4})\varphi,b_{3},b_{4}).

Now we search for the third intersection, let

C=(c1,c2,c3,c4),c1,,c4,C=(c_{1},c_{2},c_{3},c_{4}),\quad c_{1},\dots,c_{4}\in\mathbb{Z},

and

]C,C+e3[Z=(c1,c2,z,c4),z.]C,C+e_{3}[\ni Z=(c_{1},c_{2},z,c_{4}),\quad z\in\mathbb{R}\setminus\mathbb{Z}.

Again, since we demand that πZ=X\pi^{\prime}Z=X, we must have ZXEZ-X\in E:

c1x\displaystyle c_{1}-x =λ\displaystyle=-\lambda
c2\displaystyle c_{2} =μ\displaystyle=\mu
z\displaystyle z =λφ+μφ\displaystyle=\lambda\varphi+\mu\varphi
c4\displaystyle c_{4} =λφ+μ\displaystyle=\lambda\varphi+\mu

The second and fourth equations yield:

λ\displaystyle\lambda =(φ1)(c4c2)\displaystyle=(\varphi-1)(c_{4}-c_{2})
μ\displaystyle\mu =(b3b4)φ\displaystyle=(b_{3}-b_{4})\varphi

The first and third ones then yield:

x\displaystyle x =c1+(φ1)(c4c2)\displaystyle=c_{1}+(\varphi-1)(c_{4}-c_{2})
z\displaystyle z =c2+c4+c2φ\displaystyle=-c_{2}+c_{4}+c_{2}\varphi

Two expressions obtained for xx must be equal, that is:

c4c2=b4c_{4}-c_{2}=b_{4}
c1c4+c2=b1b3c_{1}-c_{4}+c_{2}=b_{1}-b_{3}

Consequently, the third intersection point ZZ is written as:

Z=(b1b3+b4,c2,c2φb4,b4+c2),Z=(b_{1}-b_{3}+b_{4},c_{2},c_{2}\varphi-b_{4},b_{4}+c_{2}),

where c2c_{2} can be chosen freely. Taking c2<Rc_{2}<R, since AB1<R\|A-B\|_{1}<R, it immediately follows that CA1<8R\|C-A\|_{1}<8R. Also, because the projection of each interval is a translation of an edge of WW and XX is in WW, it follows that at least one endpoint of each interval also projects into WW. Cases for other triples {i,j,k}\{i,j,k\} are treated in a similar way. ∎

3.3 Shape of rr-patterns

Consider an nn-pattern 𝒫\mathcal{P} of a Golden-Octagonal tiling centered at a vertex c4c\in\mathbb{Z}^{4}:

𝒫={x4|xc1<n,πxW}.\mathcal{P}=\{x\in\mathbb{Z}^{4}\ \ |\ \ \|x-c\|_{1}<n,\ \pi^{\prime}x\in W\}.

To describe the geometrical shape of π(𝒫)\pi(\mathcal{P}) we compute a section of a 33-sphere of radius rr in L1L_{1} distance with the generating plane of a Golden-Octagonal tiling. For example, one of the intersections of two of the 33-dimensional faces of the sphere and the plane can be written as:

{x1+x2+x3+x4=rx1+x2+x3+x4=r(x1,x2,x3,x4)T=λu+μv\displaystyle\begin{cases}x_{1}+x_{2}+x_{3}+x_{4}=r\\ -x_{1}+x_{2}+x_{3}+x_{4}=r\\ (x_{1},x_{2},x_{3},x_{4})^{T}=\lambda u+\mu v\end{cases}

Solving the system we obtain

x=r2+φ(01φ1),\displaystyle x=\frac{r}{2+\varphi}\begin{pmatrix}0\\ 1\\ \varphi\\ 1\end{pmatrix},

note that xx, the coordinate of one of the vertices of the section, is proportional to q3q_{3}, the third subperiod. Calculating the other vertices of the section we obtain the following:

Proposition 1.

There exists a set of constants {ci+}\{c_{i}\in\mathbb{R}^{+}\} for i=1,,4i=1,\dots,4, such that for any rr\in\mathbb{N}, vertices of an rr-pattern of a Golden-Octagonal tiling centered in origin, when projected to the physical space, are within uniformly bounded distance from an octagon defined as a convex hull of ±cirπqi\leavevmode\nobreak\ \pm c_{i}r\pi^{\prime}\vec{q}_{i}, where qiq_{i} is the ii-th subperiod.

3.4 Forcing vertices

The last lemma in the section addresses the idea of how two vertices in a tilings can force the third one.

Lemma 3.

Consider a Golden-Octagonal tiling with a window WW and three points B,A,C4B,A,C\in\mathbb{Z}^{4}, such that there exists a translation of the slope such that πB,πA\pi^{\prime}B,\pi^{\prime}A, and πC\pi^{\prime}C belong to consecutive (clockwise or anticlockwise) edges of the translated window WW^{\star}. Then, if πB,πCW\pi^{\prime}B,\pi^{\prime}C\in W, it follows that πAW\pi^{\prime}A\in W.

Proof.

Trivial in the case of Golden-Octagonal tilings since the window is convex polygon without right angles, see Figure 2. ∎

Refer to caption
Figure 2: We have πB,πA\pi^{\prime}B,\pi^{\prime}A, and πC\pi^{\prime}C belonging to three consecutive edges of a translated window WW. Note that there is no translation of the window WW such that πB,πCW\pi^{\prime}B,\pi^{\prime}C\in W but πAW\pi^{\prime}A\notin W.

4 Proof of Theorem 1

To state the main Lemma, we need to define so-called extreme worms, the vertices of a pattern that are closest to the boundary of the window after the projection:

Definition 13.

Consider a patch 𝒫\mathcal{P} of a planar octagonal tiling with slope EE and window WW. For α0\alpha\geq 0, we denote by 𝒫α\mathcal{P}_{\alpha} the vertices of 𝒫\mathcal{P} which correspond to points at distance at least α\alpha from the boundary of W(𝒫)W(\mathcal{P}):

𝒫α:={x𝒫|d(πx,W(𝒫))α},\mathcal{P}_{\alpha}:=\{x\in\mathcal{P}\leavevmode\nobreak\ |\leavevmode\nobreak\ d(\pi^{\prime}x,\partial W(\mathcal{P}))\geq\alpha\},

and respectively

[𝒫]α:={x[𝒫]|d(πx,W(𝒫))α}.[\mathcal{P}]_{\alpha}:=\{x\in[\mathcal{P}]\leavevmode\nobreak\ |\leavevmode\nobreak\ d(\pi^{\prime}x,\partial W(\mathcal{P}))\geq\alpha\}.

We have [𝒫]0=[𝒫][\mathcal{P}]_{0}=[\mathcal{P}], [𝒫]α=[𝒫α][\mathcal{P}]_{\alpha}=[\mathcal{P}_{\alpha}], and [𝒫]α[𝒫]β[\mathcal{P}]_{\alpha}\subset[\mathcal{P}]_{\beta} for α>β\alpha>\beta.

Definition 14.

Consider a patch 𝒫\mathcal{P} of a planar octagonal tiling with slope EE and window WW. For α0\alpha\geq 0, we denote by Extiα(𝒫)Ext_{i}^{\alpha}(\mathcal{P}) the vertices of [𝒫][\mathcal{P}] whose orthogonal projection on EE^{\bot} belongs to the ii-th edge of W(𝒫)W(\mathcal{P}) and is at distance at least α\alpha from each line which contains another edge of W(𝒫)W(\mathcal{P}). Also, we denote i=08Extiα([𝒫])\cup_{i=0}^{8}Ext_{i}^{\alpha}([\mathcal{P}]) as Extα([𝒫])Ext^{\alpha}([\mathcal{P}]). We call such vertices extreme worms of 𝒮\mathcal{S}.

Lemma 4.

For Golden-Octagonal tilings, for any real number α>0\alpha>0, there is a seed radius mNm\in N and growth raduis rNr\in N, such that the self-assembly algorithm provided with an nn-pattern 𝒮\mathcal{S} centered at the origin, where nmn\geq m, as a seed, and an rr-atlas, builds at step hh at least ([𝒮]αExtα([𝒮])B(0,h)([\mathcal{S}]_{\alpha}\cup Ext^{\alpha}([\mathcal{S}])\cap B(0,h).

Proof.

Given an α>0\alpha>0, consider an mm-pattern 𝒮\mathcal{S} such that Extiα([𝒮])𝒮Ext_{i}^{\alpha}([\mathcal{S}])\cap\mathcal{S}\neq\emptyset for i=1,,8i=1,\dots,8, which satisfy the following constraints:

αsinω81+|1W(𝒮α)|>|πe1|,\displaystyle\frac{\alpha}{\sin\omega_{81}}+|\partial_{1}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{1}|, αsinω23+|2W(𝒮α)|>|πe2|,\displaystyle\frac{\alpha}{\sin\omega_{23}}+|\partial_{2}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{2}|, (2)
αsinω23+|3W(𝒮α)|>|πe3|,\displaystyle\frac{\alpha}{\sin\omega_{23}}+|\partial_{3}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{3}|, αsinω45+|4W(𝒮α)|>|πe4|,\displaystyle\frac{\alpha}{\sin\omega_{45}}+|\partial_{4}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{4}|,
αsinω45+|5W(𝒮α)|>|πe1|,\displaystyle\frac{\alpha}{\sin\omega_{45}}+|\partial_{5}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{1}|, αsinω67+|6W(𝒮α)|>|πe2|,\displaystyle\frac{\alpha}{\sin\omega_{67}}+|\partial_{6}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{2}|,
αsinω67+|7W(𝒮α)|>|πe3|,\displaystyle\frac{\alpha}{\sin\omega_{67}}+|\partial_{7}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{3}|, αsinω81+|8W(𝒮α)|>|πe4|,\displaystyle\frac{\alpha}{\sin\omega_{81}}+|\partial_{8}W(\mathcal{S}_{\alpha})|>|\pi^{\prime}e_{4}|,

where iW(𝒮α)\partial_{i}W(\mathcal{S}_{\alpha}) is the ii-the edge of W(𝒮α)W(\mathcal{S}_{\alpha}) numerated clockwise starting from the leftmost edge, and ωij\omega_{ij} is the angle between iW(𝒮α)\partial_{i}W(\mathcal{S}_{\alpha}) and jW(𝒮α)\partial_{j}W(\mathcal{S}_{\alpha}), see Figure 3. Note that for any nn-pattern, with n>mn>m, the constraints above hold as well.

Refer to caption
Figure 3: Window WW is depicted in blue, the exact position of WW we do not know. Window W(𝒮α)W(\mathcal{S}_{\alpha}) is depicted in black, with edges numerated clockwise starting from the leftmost one, ωij\omega_{ij} denotes the angle between edges ii and jj.

Base of induction. Since 𝒮\mathcal{S} is an mm-pattern, the first nontrivial step of the algorithm is with k=nm1k=nm1. Let the growth radius rr be such that BX(r)𝒮=𝒮B_{X}(r)\cap\mathcal{S}=\mathcal{S} for all X𝒮X\in\mathcal{S}. Consequently, all the vertices in m+1\mathcal{H}_{m+1} are forced.

Step of induction. We need to show that given vertices at step kk, all the new vertices at step k+1k+1 are forced:

k:=([𝒮]αExtα([𝒮]))B0(k)([𝒮]αExtα([𝒮]))B0(k+1)=:k+1.\mathcal{H}_{k}:=\left([\mathcal{S}]_{\alpha}\cup Ext^{\alpha}([\mathcal{S}])\right)\cap B_{0}(k)\to([\mathcal{S}]_{\alpha}\cup Ext^{\alpha}([\mathcal{S}]))\cap B_{0}(k+1)=:\mathcal{H}_{k+1}.

It is easy to see using Proposition 1 that there are no vertices in k+1k\mathcal{H}_{k+1}\setminus\mathcal{H}_{k} that see less than three subperiod lines and that all the vertices which see exactly three subperiod lines are necessarily projected near the corners, see Figure 4. We divide k+1k\mathcal{H}_{k+1}\setminus\mathcal{H}_{k} into two sets: vertices which see at least N(α)N(\alpha) vertices along each of its subperiod lines and, therefore, when projected on EE are relatively far away from corners of the growing pattern, and vertices which see N(α)N(\alpha) vertices only along three of its subperiod lines, respectively whose vertices land near corners when projected on EE.

Refer to caption
Figure 4: The seed 𝒮\mathcal{S} is depicted with red. The vertices of k\mathcal{H}_{k} are at a bounded distance from a convex polygon with vertices written as ±ciπqi\pm c_{i}\pi^{\prime}\vec{q_{i}} for i=1,2,3,4i=1,2,3,4.

Case 1. It’s the easier case of the two. Let OO be the intersection of perpendicular bisectors of W(𝒮)W(\mathcal{S}) and let SijS_{ij} denote the sector which contains ωij\omega_{ij}. Let Ak+1kA\in\mathcal{H}_{k+1}\setminus\mathcal{H}_{k}, for example, be a vertex with πAS23\pi^{\prime}A\in S_{23}, see Figure 5. Consider subperiod lines Q2(A)Q_{2}(A) and Q3(A)Q_{3}(A). Assuming that we have chosen a big enough growth radius rr, using Lemma 1 we state that in rr-neighborhood of AA there exists two vertices BQ2(A)B\in Q_{2}(A) and CQ3(A)C\in Q_{3}(A) such that (πAB,πAC)=ω23\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC})=\omega_{23}. Since it is impossible to position window WW in a way that B,CWB,C\in W and AWA\notin W, we conclude that BB and CC force AA, see Figure 6. We treat subcases when πA\pi^{\prime}A projected to other sectors similarly.

Refer to caption
Figure 5: If AA belongs to the sector containing ω23\omega_{23}, applying Lemma 1 we choose r>r> such that there exists BQ3(A)BA(r)kB\in Q_{3}(A)\cap B_{A}(r)\cap\mathcal{H}_{k} and CQ2(A)BA(r)kC\in Q_{2}(A)\cap B_{A}(r)\cap\mathcal{H}_{k} such that (πAB,πAC)=ω23\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC})=\omega_{23}. The latter means that BB and CC force AA.
Refer to caption
Figure 6: We have BQ3(A)B\in Q_{3}(A), CQ2(A)C\in Q_{2}(A) along with (πAB,πAC)=ω23\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC})=\omega_{23}. Since it is impossible to translate window WW in such a way that B,CWB,C\in W but AWA\notin W, we conclude that vertices BB and CC force AA.

Case 2. Let us say that Akk+1A\in\mathcal{H}_{k}\setminus\mathcal{H}_{k+1} is the vertex such that πA\pi A is near a corner of k\mathcal{H}_{k} pointed by ±q2\pm q_{2}. The latter means that AA sees each of its subperiod lines except for Q3(A)Q_{3}(A). The proof that AA is forced by its neighborhood depends on the position of πA\pi^{\prime}A relative to the window W(𝒮α)W(\mathcal{S}_{\alpha}). We consider four subcases, see Figure 7:

Refer to caption
Figure 7: The subregions of W(𝒮α)W(\mathcal{S}_{\alpha}) corresponding to subcases 2a2c2a-2c are depicted respectively with blue, green, and red. Subcase 2d2d concerning the continuation of Extα([𝒮])Ext^{\alpha}([\mathcal{S}]) is depicted with red dashed line.

Case 2a. Let AA be such that

πAW(𝒮α)((πExt2α(𝒮)πe3)(πExt6α(𝒮)πe3))\pi^{\prime}A\in W(\mathcal{S}_{\alpha})\setminus((\pi^{\prime}Ext^{\alpha}_{2}(\mathcal{S})\oplus\mathbb{R}\pi^{\prime}e_{3})\cup(\pi^{\prime}Ext^{\alpha}_{6}(\mathcal{S})\oplus\mathbb{R}\pi^{\prime}e_{3})) (3)

This subcase is very similar to Case 11 but instead of all four subperiod lines, vertex AA sees just three. This downside is mitigated by the fact that distance between πA\pi^{\prime}A and 3W(𝒮α)\partial_{3}W(\mathcal{S}_{\alpha}) or 7W(𝒮α)\partial_{7}W(\mathcal{S}_{\alpha}) is greater than some positive constant ε(α)\varepsilon(\alpha). By choosing r=r(ε)r=r(\varepsilon) big enough, by Lemma 1 we are guaranteed to find in the rr-neighborhood of AA two vertices BB and CC belonging to different subperiod lines such that (πAB,πAC)\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC}) equals to one of ωij\omega_{ij}. Consequently, AA is forced by BB and CC, see Figure 8.

Refer to caption
Figure 8: Let AA be in subregion 3 of W(𝒮α)W(\mathcal{S}_{\alpha}). Along subperiods lines Q1(A),Q2(A)Q_{1}(A),Q_{2}(A), and Q4(A)Q_{4}(A), we find two vertices BB and CC such that (πAB,πAC)\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC}) equals to one of ωi\omega_{i}. Similar to Case 11, BB and CC force AA.

Case 2b. Let AA be such that

πAW(𝒮α)((πExt2α(𝒮)π([0,e3]))(πExt6α(𝒮)π([0,e3]))).\pi^{\prime}A\in W(\mathcal{S}_{\alpha})\cap((\pi^{\prime}Ext^{\alpha}_{2}(\mathcal{S})\oplus\pi^{\prime}([0,-e_{3}]))\cup(\pi^{\prime}Ext^{\alpha}_{6}(\mathcal{S})\oplus\pi^{\prime}([0,e_{3}]))). (4)

Since we are near the corner of the growing pattern, using Proposition 1 we state that given a large enough growth radius r1r_{1}, we see vertices from both extreme worms directed by πq2\pi q_{2} in the r1r_{1}-neighborhood of AA. First, let πA\pi^{\prime}A be closer to the 3W(𝒮α)\partial_{3}W(\mathcal{S}_{\alpha}) than to 7W(𝒮α)\partial_{7}W(\mathcal{S}_{\alpha}) and let BB be a vertex from the r1r_{1}-neighborhood which belongs to the extreme worm Ext2α([𝒮])kExt^{\alpha}_{2}([\mathcal{S}])\cap\mathcal{H}_{k}, so we have πA\pi^{\prime}A and πB\pi^{\prime}B near abutting edges of the W(𝒮α)W(\mathcal{S}_{\alpha}).

Consider the unit interval (A,A+e3)(A,A+e_{3}), when projected to the perpendicular space, because of (2), it will necessarily intersect the line in the perpendicular space which contains πExt2α([𝒮])\pi^{\prime}Ext^{\alpha}_{2}([\mathcal{S}]), see Figure 9. Depending on the position of BB, we choose one of two intervals π(B,B+e2),π(B,Be2)\pi^{\prime}(B,B+e_{2}),\pi^{\prime}(B,B-e_{2}) which intersects π(A,A+e3)\pi^{\prime}(A,A+e_{3}), let XWX\in W be the intersection point. By Lemma 2 we can choose kk so that in the krkr-neighborhood of AA there is a vertex C1[𝒮]C_{1}\in[\mathcal{S}] such that interval π(C1,C1e4)\pi^{\prime}(C_{1},C_{1}-e_{4}) also intersects π(A,A+e3)\pi^{\prime}(A,A+e_{3}) in the point XX. Now consider the vertex C2=C1e3C_{2}=C_{1}-e_{3}, by our initial assumptions it belongs to [𝒮]α[\mathcal{S}]_{\alpha} but may not belong to k\mathcal{H}_{k}. However, by Proposition 1, there are vertices from Q4(C2)Q_{4}(C_{2}) in k\mathcal{H}_{k}, let CQ4(A)kC\in Q_{4}(A)\cap\mathcal{H}_{k} be the one closest to AA. By Lemma 3, BB and CC forces AA.

If πA\pi^{\prime}A is closer to the 7W(𝒮α)\partial_{7}W(\mathcal{S}_{\alpha}) than to 3W(𝒮α)\partial_{3}W(\mathcal{S}_{\alpha}), the proof remains the except for BB must be chosen along Ext6α([𝒮])Ext_{6}^{\alpha}([\mathcal{S}]) instead of Ext2α([𝒮])Ext_{2}^{\alpha}([\mathcal{S}]) and interval (A,Ae3)(A,A-e_{3}) must be taken instead of (A,A+e3)(A,A+e_{3}).

Refer to caption
Figure 9: In Case 2b2b when πA\pi^{\prime}A projects to the subregion defined by 4 depicted in green, we consider an intersection XX of π(A,A+e3)\pi^{\prime}(A,A+e_{3}) with either π(B,Be2)\pi^{\prime}(B,B-e_{2}) or π(B,B+e2)\pi^{\prime}(B,B+e_{2}). Using Lemma 2 we find interval (C,C+e4)(C,C+e_{4}) with πC1W(𝒮)\pi^{\prime}C_{1}\in W(\mathcal{S}), whose projection intersects π(A,A+e3)\pi^{\prime}(A,A+e_{3}) also in XX. Then, we consider C2=C1e3C_{2}=C_{1}-e_{3} and its subperiod line Q4(C2)Q_{4}(C_{2}). Let CQ4(C2)C\in Q_{4}(C_{2}) be vertex from k\mathcal{H}_{k} closest to AA. Now, pair BB and CC forces AA.

Case 2c. Let AA be such that

πAW(𝒮α)(\displaystyle\pi^{\prime}A\in W(\mathcal{S}_{\alpha})\cap( (πExt2α(𝒮)πe3)(πExt2α(𝒮)π[0,e3])\displaystyle(\pi^{\prime}Ext^{\alpha}_{2}(\mathcal{S})\oplus\mathbb{R}\pi^{\prime}e_{3})\setminus(\pi^{\prime}Ext^{\alpha}_{2}(\mathcal{S})\oplus\pi^{\prime}[0,-e_{3}])
\displaystyle\cup (πExt6α(𝒮)πe3)(πExt6α(𝒮)π[0,e3])),\displaystyle(\pi^{\prime}Ext^{\alpha}_{6}(\mathcal{S})\oplus\mathbb{R}\pi^{\prime}e_{3})\setminus(\pi^{\prime}Ext^{\alpha}_{6}(\mathcal{S})\oplus\pi^{\prime}[0,e_{3}])), (5)

Here we use the Lemma 1 two times consecutively. First, we choose a radius r1r_{1} such that in the r1r_{1}-neighborhood of AA there is a vertex BQ3(A)B\in Q_{3}(A) which projects into the subregion of W(𝒮α)W(\mathcal{S}_{\alpha}) defined by (3), the one associated with the previous Case 2b2b. Using the same reasoning as in Case 2b2b, and possibly increasing the radius needed by r1r_{1}, we conclude that BB is forced by its local neighborhood.

Second, we choose r2r_{2} such that in the r2r_{2}-neighborhood of AA contains a vertex CkQ4(A)C\in\mathcal{H}_{k}\cap Q_{4}(A) which does not belong to the subregions (5) and (4) associated with Cases 2b2b and 2c2c. Since (πAB,πAC)=ω34=ω78\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC})=\omega_{34}=\omega_{78}, we conclude AA is forced by BB and CC, both of which are forced by a local neighborhood of AA.

Refer to caption
Figure 10: Let AA be in subregion 4 of W(𝒮α)W(\mathcal{S}_{\alpha}). First, we choose CQ4(A)kC\in Q_{4}(A)\cap\mathcal{H}_{k} such that πC\pi^{\prime}C does not belong to (4) and (5). Along subperiod line Q3(A)Q_{3}(A) we find a vertex BB with πB\pi^{\prime}B in (4), which is forced as in Case 2b2b. Since (πAB,πAC)\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC}) equals to ω34\omega_{34}, BB and CC forces AA.

Case 2d.

This is the last case, it is left to consider

A(Ext2α([𝒮])Ext6α([𝒮]))(k+1k)A\in(Ext_{2}^{\alpha}([\mathcal{S}])\cup Ext_{6}^{\alpha}([\mathcal{S}]))\cap(\mathcal{H}_{k+1}\setminus\mathcal{H}_{k})

For example, let AExt2α([𝒮])(k+1k)A\in Ext^{\alpha}_{2}([\mathcal{S}])\cap(\mathcal{H}_{k+1}\setminus\mathcal{H}_{k}), the proof once again depends on the position of πA\pi^{\prime}A in the W(𝒮α)W(\mathcal{S}_{\alpha}). If πA\pi^{\prime}A is closer to 1W(𝒮α)\partial_{1}W(\mathcal{S}_{\alpha}) than to 3W(𝒮α)\partial_{3}W(\mathcal{S}_{\alpha}), then in the rr-neighborhood of AA, given that rr is big enough, there exists BExt2α([𝒮])B\in Ext_{2}^{\alpha}([\mathcal{S}]) which is closer to 3W(𝒮α)\partial_{3}W(\mathcal{S}_{\alpha}) than AA, as it is guaranteed by Lemma 1. Since AA is near a corner of the growing pattern directed by q2\vec{q_{2}}, it sees every subperiod except for the third, let BQ1(A)B\in Q_{1}(A) be any vertex along the first subperiod which is already in place. Again, since (πAB,πAC)=ω12\angle(\pi^{\prime}\overrightarrow{AB},\pi^{\prime}\overrightarrow{AC})=\omega_{12}, BB and CC force AA, see Figure 12.

On the other hand, if πA\pi^{\prime}A is closer to 3W(𝒮α)\partial_{3}W(\mathcal{S}_{\alpha}) than to 1W(𝒮α)\partial_{1}W(\mathcal{S}_{\alpha}), then using Lemma 1 we find a vertex BB which is closer to 1W(𝒮α)\partial_{1}W(\mathcal{S}_{\alpha}), when projected to the perpendicular space, than AA. Then, we consider a vertex C=Ae3C^{\prime}=A-e_{3}, which belongs to W(𝒮α)W(\mathcal{S}_{\alpha}), as it is guaranteed by 2, and its subperiod line Q4(A)Q_{4}(A). Since CC^{\prime} also sees every subperiod except for the third, we can choose CQ4(A)kC\in Q_{4}(A)\cap\mathcal{H}_{k}. By Lemma 3, BB and CC force AA, see Figure 12.

If AExt6α(𝒮)A\in Ext_{6}^{\alpha}(\mathcal{S}) instead of Ext2α(𝒮)Ext_{2}^{\alpha}(\mathcal{S}), the proof remains the same except that BB is taken along Ext6α([𝒮])Ext_{6}^{\alpha}([\mathcal{S}]) and C=A+e3C^{\prime}=A+e_{3}.

That concludes the proof of Case 22 for vertices near the two corners of the growing pattern directed by πq2\pi q_{2}. The crucial property we used during the proof is that for vertices near a corner that do not see their ii-th subperiod line and are projected close to ii-th edge of W(Sα)W(S_{\alpha}) (as in Case 2b2b and 2c2c), to apply Lemma 3, the extreme worm passing through the corner must be of type i1(mod8)i-1\pmod{8} or i+1(mod8)i+1\pmod{8}. Fortunately for us, by Proposition 1, this property is satisfied for every corner of the growing pattern. Consequently, the proof for vertices near other corners remains the same up to a rearrangement of indices.

Taking as seed any nn-pattern, with n>mn>m does not change the reasoning since constraints 2 remain satisfied.

Theorem 1 immediately follows from Lemma 4.

Refer to caption
Figure 11: If d(A,3W(𝒮α))>d(A,1W(𝒮α))d(A,\partial_{3}W(\mathcal{S}_{\alpha}))>d(A,\partial_{1}W(\mathcal{S}_{\alpha})), using Lemma 1 we find BExt2α([𝒮])kB\in Ext_{2}^{\alpha}([\mathcal{S}])\cap\mathcal{H}_{k} BB such that d(B,1W(𝒮α))>d(A,1W(𝒮α))d(B,\partial_{1}W(\mathcal{S}_{\alpha}))>d(A,\partial_{1}W(\mathcal{S}_{\alpha})), and, since AA sees every subperiod except for the 33rd, we find vertex CQ1(A)kC\in Q_{1}(A)\cap\mathcal{H}_{k}. Pair BB and CC forces AA.
Refer to caption
Figure 12: If d(A,3W(𝒮α))<d(A,1W(𝒮α))d(A,\partial_{3}W(\mathcal{S}_{\alpha}))<d(A,\partial_{1}W(\mathcal{S}_{\alpha})), again using Lemma 1 we find BExt2([𝒮])αkB\in Ext_{2}([\mathcal{S}])^{\alpha}\cap\mathcal{H}_{k} such that d(B,1W(𝒮α))<d(A,1W(𝒮α))d(B,\partial_{1}W(\mathcal{S}_{\alpha}))<d(A,\partial_{1}W(\mathcal{S}_{\alpha})). Then, consider a vertex C=Ae3[𝒮α]C^{\prime}=A-e_{3}\in[\mathcal{S}_{\alpha}] and its subperiod line Q4(C)Q_{4}(C). Using Lemma 1 we find vertex CQ4(C)kC\in Q_{4}(C^{\prime})\cap\mathcal{H}_{k}. Pair BB and CC forces AA.

References

  • [Bee82] F. Beenker. Algebraic theory of non periodic tilings of the plane by two simple building blocks: a square and a rhombus. Technical Report TH Report 82-WSK-04, Technische Hogeschool Eindhoven, 1982.
  • [Ben85] L. Bendersky. Quasicrystal with one-dimensional translational symmetry and a tenfold rotation axis. Physical review letters, 55:1461–1463, 10 1985.
  • [BF13] N. Bédaride and Th. Fernique. Aperiodic Crystals, chapter The Ammann–Beenker tilings revisited, pages 59–65. Springer Netherlands, Dordrecht, 2013.
  • [BF15a] N. Bédaride and Th. Fernique. No weak local rules for the 4p-fold tilings. Discrete & Computational Geometry, 54:980–992, 2015.
  • [BF15b] N. Bédaride and Th. Fernique. When periodicities enforce aperiodicity. Communications in Mathematical Physics, 335:1099–1120, 2015.
  • [BF17] N. Bédaride and Th. Fernique. Weak local rules for planar octagonal tilings. Israël journal of mathematics, 222:63–89, 2017.
  • [BF20] N. Bédaride and Th. Fernique. Canonical projection tilings defined by patterns. Geometriae Dedicata, 208:157–175, 2020.
  • [BG13] M. Baake and U. Grimm. Aperiodic Order, volume 149. Cambridge University Press, 2013.
  • [BH86] P. Bancel and P. A. Heiney. Icosahedral aluminium - transition-metal alloys. Physical review. B, Condensed matter, 33:7917–7922, 07 1986.
  • [CKP00] H. Cohn, R. Kenyon, and J. Propp. A variational principle for domino tilings. Journal of the American Mathematical Society, 14, 08 2000.
  • [DB81] N. De Bruijn. Algebraic theory of Penrose’s nonperiodic tilings of the plane. Nederl. Akad. Wetensch. Indag. Math., 43:39–52, 1981.
  • [DB06] M. De Boissieu. Stability of quasicrystals: Energy, entropy and phason modes. Philosophical Magazine A-physics of Condensed Matter Structure Defects and Mechanical Properties - PHIL MAG A, 86:1115–1122, 02 2006.
  • [Els85] V. Elser. Comment on "quasicrystals: A new class of ordered structures". Phys. Rev. Lett., 54:1730–1730, Apr 1985.
  • [FG20] Th. Fernique and I. Galanov. Local growth of planar rhombus tilings. J. Phys.: Conf. Ser., 1458:012001, 2020.
  • [Gar89] M. Gardner. Penrose Tiles to Trapdoor Ciphers. Freema, 1989.
  • [HKSW14] Alan Haynes, Henna Koivusalo, Lorenzo Sadun, and James Walton. Gaps problems and frequencies of patches in cut and project sets. Mathematical Proceedings of the Cambridge Philosophical Society, -1, 11 2014.
  • [HSS16] C. T. Hahn, J. E. S. Socolar, and P. J. Steindhardt. Local growth of icosahedral quasicrystalline tilings. Physical review B, 94:014113, 2016.
  • [INF85] T. Ishimasa, H. U. Nissen, and Y. Fukano. New order state between crystalline and amorphous in Ni\textrm{N}i - Cr{C}r particles. Physical review letters, 55:511–513, 08 1985.
  • [Kat95] A. Katz. Beyond Quasicrystals: Les Houches, March 7–18, 1994, chapter Matching rules and quasiperiodicity: the octagonal tilings, pages 141–189. Springer Berlin Heidelberg, Berlin, Heidelberg, 1995.
  • [Le95] T. Q. T. Le. Local rules for pentagonal quasi-crystals. Discrete & Computational Geometry, 14:31–70, 1995.
  • [Le97] T. Q. T. Le. Local rules for quasiperiodic tilings. In The mathematics of long-range aperiodic order (Waterloo, ON, 1997), volume 489 of NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., pages 331–366. Kluwer Acad. Publ., Dordrecht, 1997.
  • [Lev88] L. Levitov. Local rules for quasicrystals. Communications in Mathematical Physics, 119:627–666, 1988.
  • [LH99] C. L. Henley. Random tiling models. Quasicrystals: The State of the Art, pages 459–560, 11 1999.
  • [LP95] T. Q. T. Le and S. Piunikhin. Local rules for multi-dimensional quasicrystals. Differential Geometry and its Applications, 5:10–31, 1995.
  • [LPS92] T. Q. T. Le, S. Piunikhin, and V. Sadov. Local rules for quasiperiodic tilings of quadratic 22-planes in 𝐑4{\bf R}^{4}. Communications in mathematical physics, 150:23–44, 1992.
  • [LS84] D. Levine and P. J. Steinhardt. Quasicrystals: A new class of ordered structures. Physical Review Letters, 53:2477–2480, 1984.
  • [NINE15] K. Nagao, T. Inuzuka, K. Nishimoto, and K. Edagawa. Experimental observation of quasicrystal growth. Phys. Rev. Lett., 115, 8 2015.
  • [OSDS88] G. Onoda, P. Steinhardt, D. DiVincenzo, and J. Socolar. Growing perfect quasicrystals. Phys. Rev. Lett., 60:2653, 1988.
  • [Pen74] R. Penrose. The role of aesthetics in pure and applied mathematical research. The Institute of Mathematics and its Applications Bulletin, 10:266–271, 1974.
  • [SBGC84] D. Shechtman, I. Blech, D. Gratias, and J. Cahn. Metallic phase with long-range orientational symmetry and no translational symmetry. Phys. Rev. Let., 53:1951–1953, 1984.
  • [Sch98] M. Schlottmann. Cut-and-project sets in locally compact abelian groups. pages 247–264, 1998.
  • [Soc90] J. E. S. Socolar. Weak matching rules for quasicrystals. Communications in Mathematical Physics, 129:599–619, 1990.
  • [Soc91] J. Socolar. Growth rules for quasicrystals. In D. DiVincenzo and P. Steinhardt, editors, Quasicrystals: the State of the art, pages 213–238. World Scientific, 1991.
  • [WCHK87] N. Wang, H. Chen, and K. H. Kuo. Two-dimensional quasicrystal with eightfold rotational symmetry. Physical review letters, 59:1010–1013, 09 1987.