Skip to main content
Log in

Planar Rosa: a family of quasiperiodic substitution discrete plane tilings with 2n-fold rotational symmetry

  • Published:
Natural Computing Aims and scope Submit manuscript

Abstract

We present Planar Rosa, a family of rhombus tilings with a 2n-fold rotational symmetry that are generated by a primitive substitution and that are also discrete plane tilings, meaning that they are obtained as a projection of a higher dimensional discrete plane. The discrete plane condition is a relaxed version of the cut-and-project condition. We also prove that the Sub Rosa substitution tilings with 2n-fold rotational symmetry defined by Kari and Rissanen do not satisfy even the weaker discrete plane condition. We prove these results for all even \(n\geqslant 4\). This completes our previously published results for odd values of n.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19

Similar content being viewed by others

References

  • Arnoux P, Furukado M, Harriss E, Ito S (2011) Algebraic numbers, free group automorphisms and substitutions on the plane. Trans Am Math Soc 363:4651–4699

    Article  MathSciNet  MATH  Google Scholar 

  • Arnoux P, Ito S, Sano Y (2001) Higher dimensional extensions of substitutions and their dual maps. J Anal Math 83:183–2062001

    Article  MathSciNet  MATH  Google Scholar 

  • Baake M, Grimm U (2013) Aperiodic order: a mathematical invitation. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Beenker FPM (1982) Algebraic theory of non-periodic tilings of the plane by two simple building blocks: a square and a rhombus. EUT report. WSK, Dept. of Mathematics and Computing Science

  • Bédaride N, Fernique Th (2015) When periodicities enforce aperiodicity. Commun Math Phys 335:1099–1120

    Article  MathSciNet  MATH  Google Scholar 

  • Conway J, Jones A (1976) Trigonometric Diophantine equations (on vanishing sums of roots of unity). Acta Arith 30:229–240

    Article  MathSciNet  MATH  Google Scholar 

  • Davis Philip J (1979) Circulant matrices. Wiley, New York

    MATH  Google Scholar 

  • De Bruijn NG (1981) Algebraic theory of Penrose’s nonperiodic tilings of the plane. I and II. Kon. Nederl. Akad. Wetensch. Proc. Ser. A

  • Gahler F, Rhyner J (1986) Equivalence of the generalised grid and projection methods for the construction of quasiperiodic tilings. J Phys A Math General. https://doi.org/10.1088/0305-4470/19/2/020

    Article  MATH  Google Scholar 

  • Grünbaum B, Shephard GC (1987) Tilings and patterns. Courier Dover Publications, Mineola

    MATH  Google Scholar 

  • Kari J, Lutfalla VH (2022) Substitution discrete plane tilings with \(2n\)-fold rotational symmetry for odd \(n\). Discrete Comput Geom

  • Kari J, Rissanen M (2016) Sub Rosa, a system of quasiperiodic rhombic substitution tilings with \(n\)-fold rotational symmetry. Discrete Comput Geometry

  • Kenyon R (1993) Tiling a polygon with parallelograms. Algorithmica 9:382–397

    Article  MathSciNet  MATH  Google Scholar 

  • Levitov LS (1988) Local rules for quasicrystals. Commun Math Phys 119:627–666

    Article  MathSciNet  Google Scholar 

  • Lutfalla VH (2021) An effective construction for cut-and-project rhombus tilings with global \(n\)-fold rotational symmetry. In: 27th IFIP WG 1.5 international workshop on cellular automata and discrete complex systems (AUTOMATA)

  • Lutfalla VH (2021) Substitution discrete planes. Theses, Université Sorbonne Paris Nord. hal:tel-03376430

  • Penrose R (1974) The role of aesthetics in pure and applied mathematical research. Bull Inst Math Appl 10:266–271

    Google Scholar 

  • Senechal M (1996) Quasicrystals and geometry. CUP Archive, Cambridge

  • Socolar JES (1990) Weak matching rules for quasicrystals. Commun Math Phys 129:599–619

    Article  MathSciNet  MATH  Google Scholar 

  • Solomyak B (1998) Nonperiodicity implies unique composition for self-similar translationally finite tilings. Discrete Comput Geom 20:265–279

Download references

Acknowledgements

We wish to thank Thomas Fernique for his help and proofreading.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Victor H. Lutfalla.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Technical details

Technical details

Lemma 34

(Decompposition of \(\mathbb {R}^n\) as the orthogonal direct sum of the spaces \(\mathcal {E}_n^k\)) For any even integer n. Define the planes \(\mathcal {E}_n^k\) for \(0\leqslant k < \tfrac{n}{2}\) as

$$\begin{aligned} \mathcal {E}_n^k := \left\langle \left( \cos \frac{(2k+1)i\pi }{n}\right) _{0\leqslant i< n}, \left( \sin \frac{(2k+1)i\pi }{n}\right) _{0\leqslant i < n} \right\rangle . \end{aligned}$$

The planes \(\mathcal {E}_n^k\) are pairwise orthogonal and their direct sum is \(\mathbb {R}^n\).

Proof

Remark that the fact that the direct sum of the spaces \(\mathcal {E}_n^k\) is \(\mathbb {R}^n\) is a consequence of the fact that they are pairwise orthogonal. Indeed if they are pairwise orthogonal then their sum is direct (trivial intersection) and the dimension of the sum is the sum of the dimensions so the dimension is n and the sum is \(\mathbb {R}^n\).

Now we prove that they are pairwise orthogonal. Take \(0\leqslant k_1< k_2 < \tfrac{n}{2}\). \(\mathcal {E}_n^{k_1}\) and \(\mathcal {E}_n^{k_2}\) are orthogonal, indeed :

  • \(\left\langle \left( \cos \frac{(2k_1+1)i\pi }{n}\right) _{0\leqslant i< n} \bigg | \left( \cos \frac{(2k_2+1)i\pi }{n}\right) _{0\leqslant i < n} \right\rangle = 0\), indeed we have

    $$\begin{aligned}&\left\langle \left( \cos \frac{(2k_1+1)i\pi }{n}\right) _{0\leqslant i< n} \bigg | \left( \cos \frac{(2k_2+1)i\pi }{n}\right) _{0\leqslant i < n} \right\rangle \\&\qquad = \sum \limits _{i=0}^{n-1} \cos \frac{(2k_1+1)i\pi }{n}\cos \frac{(2k_2+1)i\pi }{n}\\&\qquad = \sum \limits _{i=0}^{n-1}\frac{1}{2}\left( \cos \frac{2i(k_1+k_2+1)\pi }{n}+ \cos \frac{2i(k_2-k_1)\pi }{n}\right) \\&\qquad = \left( \sum \limits _{i=0}^{n-1}\frac{1}{2} \cos \frac{2i(k_1+k_2+1)\pi }{n}\right) + \left( \sum \limits _{i=0}^{n-1}\frac{1}{2}\cos \frac{2i(k_2-k_1)\pi }{n}\right) \\&\qquad = 0 \end{aligned}$$

    Note that the important hypothesis is \(0\leqslant k_1< k_2 < \tfrac{n}{2}\) so \(0< (k_2 - k_1), (k_1+k_2+1) < n\) and \(\sum \limits _{i=0}^{n-1}\frac{1}{2} \cos \frac{2i(k_1+k_2+1)\pi }{n} = 0\) (and same with \((k_2-k_1\))).

  • similarly \(\left\langle \left( \sin \frac{(2k_1+1)i\pi }{n}\right) _{0\leqslant i< n} \bigg | \left( \sin \frac{(2k_2+1)i\pi }{n}\right) _{0\leqslant i < n} \right\rangle = 0\) with

    $$\begin{aligned} 2\sin \frac{(2k_1+1)i\pi }{n}\sin \frac{(2k_2+1)i\pi }{n} = \cos \frac{2i(k_1+k_2+1)\pi }{n}- \cos \frac{2i(k_2-k_1)\pi }{n}. \end{aligned}$$
  • similarly we have \(\left\langle \left( \cos \frac{(2k_1+1)i\pi }{n}\right) _{0\leqslant i< n} \bigg | \left( \sin \frac{(2k_2+1)i\pi }{n}\right) _{0\leqslant i < n} \right\rangle = 0\) with

    $$\begin{aligned} 2\cos \frac{(2k_1+1)i\pi }{n}\sin \frac{(2k_2+1)i\pi }{n} = \sin \frac{2i(k_1+k_2+1)\pi }{n}- \sin \frac{2i(k_2-k_1)\pi }{n}. \end{aligned}$$
  • similarly we have \(\left\langle \left( \sin \frac{(2k_1+1)i\pi }{n}\right) _{0\leqslant i< n} \bigg | \left( \cos \frac{(2k_2+1)i\pi }{n}\right) _{0\leqslant i < n} \right\rangle = 0\).

Lemma 35

(The billiard word \(w\) of slope \(\gamma\) is well-defined) Let n be an even integer at least 4. Let \(\gamma :=(\cos \tfrac{i\pi }{n})_{0\leqslant i < n/2}\). Let \(\Gamma _\frac{1}{2} := \langle \gamma \rangle + (\frac{1}{2} ,\frac{1}{2} , \dots ,\frac{1}{2} ) = \{ (\frac{1}{2} + t\cos \tfrac{i\pi }{n})_{0\leqslant i < n/2} |\ t \in \mathbb {R}^+\}\). Let \(H_{j,k}:= \{ x \in \mathbb {R}^{n/2}|\ x_j = k\}\) with \(j<\frac{n}{2}\) and \(k> 1\).

The billiard word \(w\) of line \(\Gamma _\frac{1}{2}\) is well-defined , i.e., the line \(\Gamma _\frac{1}{2}\) never intersects two distinct hyperplanes at once, i.e.,

for any \(j,j'\in \{0,1, \dots , \frac{n}{2} -1\}\), \(k,k' \in \mathbb {N}\), if \((j,k)\ne (j',k')\) then \(\Gamma _\frac{1}{2} \cap H_{j,k} \cap H_{j',k'} = \emptyset\).

Proof

If \(j=j'\) the result is trivial because in that case \(k\ne k'\) and therefore \(H_{j,k} \cap H_{j',k'} = \emptyset\).

If \(j\ne j'\), without loss of generality we assume \(j<j'\). For contradiction, assume that the intersection is non-empty, i.e., assume that there exists \(x\in \Gamma _\frac{1}{2} \cap H_{j,k} \cap H_{j',k'}\). By definition of \(\Gamma _\frac{1}{2}\), \(x = t\gamma + (\frac{1}{2} ,\frac{1}{2} ,\dots ,\frac{1}{2} )\) for some \(t \in \mathbb {R}^+\). By definition of \(H_{j,k}\) and \(H_{j',k'}\), \(x_j = k\) and \(x_{j'} = k'\).

So we get \(t = (k-\frac{1}{2} )/\cos (\tfrac{j\pi }{n}) = (k'-\frac{1}{2} )/\cos (\tfrac{j'\pi }{n})\) which implies \((k-\frac{1}{2} )\cos (\tfrac{j'\pi }{n}) - (k'-\frac{1}{2} )\cos (\tfrac{j\pi }{n}) = 0\). Since the angles are rational with \(\pi\) with \(0\leqslant \tfrac{j\pi }{n}< \tfrac{j'\pi }{n} < \tfrac{\pi }{2}\) and the coefficients are rational, this is a trigonometric diophantine equation with two terms so we can apply known results on trigonometric diophantine equations Conway and Jones (1976). Using this known result we obtain that the only possible case is \(\tfrac{j\pi }{n}=0\) and \(\tfrac{j'\pi }{n}=\tfrac{\pi }{3}\). We now have \((k-\frac{1}{2} )\cdot \frac{1}{2} - (k'-\frac{1}{2} ) = 0\), i.e., \(\tfrac{k}{2} + \tfrac{1}{4} = k'\) which is impossible since k and \(k'\) are integers. We get the expected contradiction so this means that the intersection is empty.

Lemma 36

(The eigenvalue matrix \(N_{n}\) is almost orthogonal) Let n be an even integer greater than 2. Let \(N_{n}\) be defined as

$$\begin{aligned} N_{n} := \left( \lambda _{n,j,i} \right) _{0\leqslant i,j< \frac{n}{2}} = \left( \eta _j \cos \tfrac{(2i+1)j\pi }{n}\right) _{0\leqslant i,j < \frac{n}{2}}, \end{aligned}$$

with \(\eta _0:=1\) and \(\eta _j:= 2\) for \(0<j<\tfrac{n}{2}\) (Definition 12).

The matrix \(N_{n}\) is almost orthogonal in the sense that \(N_{n}\cdot D^\textsf{T}= D\cdot N_{n}^\textsf{T}= \frac{n}{2}\textrm{Id}_{\frac{n}{2}}\) with the matrix D defined below.

In particular, \(N_{n}\cdot \gamma ^\textsf{T}= (\frac{n}{2},0,0,\dots 0)^\textsf{T}\) with \(\gamma := \left( \cos (\tfrac{i\pi }{n})\right) _{0\leqslant i < \frac{n}{2}}\) the optimal rhombus frequency vector (Definition 15).

Proof

Let us first recall that here n is an even integer.

Definitions: We define four matrices A, B, C and D by:

$$\begin{aligned} A&:= \left( a_i \cos \frac{i(2j+1)\pi }{2n}\right) _{0\leqslant i,j< n} \qquad \qquad \qquad a_i = {\left\{ \begin{array}{ll} \sqrt{\frac{1}{n}} \text { if } i=0 \\ \sqrt{\frac{2}{n}} \text { otherwise}\end{array}\right. }\\ B&:= \left( b_i \cos \frac{i(2j+1)\pi }{n}\right) _{0\leqslant i< \frac{n}{2},0\leqslant j< n} \qquad \qquad b_i = {\left\{ \begin{array}{ll} \sqrt{\frac{1}{n}} \text { if } i=0 \\ \sqrt{\frac{2}{n}} \text { otherwise}\end{array}\right. }\\ C&:= \left( c_i \cos \frac{i(2j+1)\pi }{n}\right) _{0\leqslant i,j< \frac{n}{2}} \qquad \qquad \qquad c_i = {\left\{ \begin{array}{ll} \sqrt{\frac{2}{n}} \text { if } i=0 \\ \frac{2}{\sqrt{n}} \text { otherwise}\end{array}\right. }\\ D&:= \left( \cos \frac{(2i+1)j\pi }{n}\right) _{0\leqslant i,j< \frac{n}{2}} \end{aligned}$$

Overview of the proof: the matrix A is a Discrete Cosine Transform matrix, sometimes called DCT-III, which is know to be orthogonal. From this we prove that B is semi-orthogonal and C is orthogonal. From the fact that C is orthogonal we get \(N_{n}\cdot D^\textsf{T}= D\cdot N_{n}^\textsf{T}= \tfrac{n}{2}\textrm{Id}_{\frac{n}{2}}\) because \((N_{n}\cdot D^\textsf{T})_{i,k} = \tfrac{n}{2}(C\cdot C^\textsf{T})_{i,k}\).

Details and computations: We consider as known the fact that A (DCT-III) is orthogonal. As B is a rectangular matrix and not a square matrix, it cannot be orthogonal, however we say it is semi-orthogonal for \(B\cdot B^\textsf{T}= \textrm{Id}_{\frac{n}{2}}\). This holds due to the fact that A is orthogonal and

$$\begin{aligned} (B\cdot B^\textsf{T})_{i,k} = \sum \limits _{j=0}^{n-1}B_{i,j}B_{k,j} = \sum \limits _{j=0}^{n-1}A_{2i,j}A_{2k,j} = (A\cdot A^\textsf{T})_{2i,2k}. \end{aligned}$$

Now remark that \(B_{i,j}=B_{i,n-1-j}\) because

$$\begin{aligned} \cos \left( \frac{i(2(n-1-j)+1)\pi }{n}\right) = \cos \left( 2i\pi - \frac{i(2j+1)\pi }{n}\right) = \cos \left( \frac{i(2j+1)\pi }{n}\right) , \end{aligned}$$

So \((B\cdot B^\textsf{T})_{i,k} = (C\cdot C^\textsf{T})_{i,k}\) because \(C_{i,j}=\sqrt{2}B_{i,j}\) and

$$\begin{aligned} (B\cdot B^\textsf{T})_{i,k} = \sum \limits _{j=0}^{n-1} B_{i,j}B_{k,j} = 2\sum \limits _{j=0}^{\frac{n}{2}-1} B_{i,j}B_{k,j} = \sum \limits _{j=0}^{\frac{n}{2}-1} \sqrt{2}B_{i,j}\sqrt{2}B_{k,j} = \sum \limits _{j=0}^{\frac{n}{2}-1} C_{i,j}C_{k,j}. \end{aligned}$$

So C is orthogonal. Now remark that C is somewhat similar to \(N_{n}^\textsf{T}\).

We actually have \((C^\textsf{T}\cdot C)_{i,j} = \tfrac{2}{n} (N_{n}\cdot D^\textsf{T})_{i,j}\) , for any \(0\leqslant i,j < n\), indeed

$$\begin{aligned} \left( C^\textsf{T}\cdot C\right) _{i,j}&= \sum \limits _{k=0}^{\frac{n}{2}-1} C_{k,i}C_{k,j} = \sum \limits _{k=0}^{\frac{n}{2}-1} c_k^2 \cos \left( \frac{k(2i+1)\pi }{n}\right) \cos \left( \frac{k(2j+1)\pi }{n}\right) \\&= \sum \limits _{k=0}^{\frac{n}{2}-1} \eta _k\frac{2}{n} \cos \left( \frac{k(2i+1)\pi }{n}\right) \cos \left( \frac{k(2j+1)\pi }{n}\right) = \tfrac{2}{n} \left( N_{n}\cdot D^\textsf{T}\right) _{i,j}, \end{aligned}$$

recall that \(N_{n}=(\eta _j \cos \tfrac{(2i+1)j\pi }{n})_{0\leqslant i,j < \frac{n}{2}}\) with \(\eta _0 = 1\) and \(\eta _j= 2\) for \(j>0\), so we indeed have \(c_j^2 = \tfrac{2 \eta _j}{n}\).

In particular we have \(N_{n}\cdot \gamma ^\textsf{T}= (\tfrac{n}{2},0,0\dots , 0)^\textsf{T}\).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kari, J., Lutfalla, V.H. Planar Rosa: a family of quasiperiodic substitution discrete plane tilings with 2n-fold rotational symmetry. Nat Comput 22, 539–561 (2023). https://doi.org/10.1007/s11047-022-09929-8

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11047-022-09929-8

Keywords

Navigation