Fact-checked by Grok 2 weeks ago

Penrose tiling

A Penrose tiling is a non-periodic of the generated by an aperiodic set of prototiles, named after the mathematician who introduced the concept in the . These tilings cover the plane without gaps or overlaps but cannot do so in a periodic manner, meaning no translation leaves the entire pattern unchanged. Penrose's original sets included up to six prototiles, but they were later simplified to two: either a pair of rhombi with acute angles of 36° and 72°, or a and a dart shape related by the . The discovery stemmed from Penrose's exploration of tilings with fivefold symmetry, inspired by quasiperiodic structures, and was first detailed in his paper on in . Subsequent refinements, such as the and kite-dart versions, demonstrated that just two prototiles suffice for aperiodicity, influencing fields like physics where similar patterns appear in atomic arrangements. Key properties include hierarchical structure, where larger patterns emerge from smaller ones via inflation rules based on the golden ratio \phi = (1 + \sqrt{5})/2, and diffraction patterns exhibiting sharp peaks indicative of underlying order despite the lack of periodicity. Penrose tilings have applications in , art, and , exemplifying how simple rules can produce complex, non-repeating beauty.

Fundamentals of Tilings

Periodic Tilings

A periodic tiling is a covering of the by non-overlapping tiles such that the entire pattern is invariant under translations by a set of vectors forming a , ensuring the design repeats indefinitely without gaps or overlaps. This generates the symmetries of the , allowing any portion of the pattern to be reproduced elsewhere by shifting along basis vectors. Mathematically, the of a periodic includes a , rank-two abelian isomorphic to \mathbb{Z}^2, generated by two linearly independent translation vectors that span the . This group acts freely on the plane, preserving the tiling's structure and ensuring uniformity across infinite repetitions. Classic examples include the square grid, where squares form a with basis vectors along the axes, creating a simple rectangular cell; the triangular , composed of equilateral triangles meeting six at each vertex, with a rhombus or hexagon as the cell; and the hexagonal , using regular hexagons with three meeting at each vertex, repeatable via a parallelogram cell. These monohedral , using a single tile shape, illustrate how periodic arrangements achieve complete plane coverage through modular repetition. Periodic tilings have appeared in human art since , with Mesopotamian clay cone mosaics from around 3000 BCE employing repeating geometric motifs for decorative walls, and floor mosaics from the 1st century BCE featuring tessellated patterns of squares and triangles. In medieval , such as the 14th-century decorations at the Topkapi Palace, artisans crafted intricate periodic girih tilings using interlocking stars and polygons, often based on square or hexagonal lattices, to adorn mosques and madrasas. These historical applications highlight periodic tilings' role in achieving aesthetic harmony through and repetition, contrasting with aperiodic tilings that lack such lattice-based periodicity.

Aperiodic Tilings

Aperiodic tilings represent a class of coverings where a of prototiles can cover the infinite without gaps or overlaps, but every such covering lacks periodicity, meaning no non-trivial vector exists that maps the entire onto itself. This property contrasts with periodic tilings, which repeat a fundamental domain indefinitely under , and highlights the counterintuitive possibility that local matching rules enforced by edges can force global disorder. The emerged from efforts to understand whether tile sets obeying strict adjacency rules necessarily produce repeatable patterns. The theoretical foundations trace to Hao Wang's 1961 introduction of Wang tiles—unit squares with colored edges that must match when adjacent—and his that any set of such tiles capable of the plane must admit a periodic . This assumption aligned with the intuition that local constraints imply global order, akin to the periodicity in crystallographic structures. However, in 1966, Robert Berger refuted the by proving the undecidability of the domino problem—determining whether a given tile set tiles the plane—and, in the process, constructed the first aperiodic set consisting of 20,426 Wang tiles, demonstrating that aperiodicity could arise from computability reductions simulating computations. Subsequent refinements reduced the size of aperiodic sets significantly. In 1971, Raphael Robinson streamlined Berger's approach and developed an aperiodic tile set of 6 prototiles, proving nonperiodicity through hierarchical structures that enforce infinite growth without repetition. Independently in the early 1970s, Robert Ammann devised several smaller aperiodic sets, including one with 16 tiles using linear markings (Ammann bars) to propagate non-periodic constraints across the plane, as later formalized and proven aperiodic. These constructions, building on tiles, established that aperiodicity requires only modest numbers of prototiles. The existence of aperiodic tile sets upends the classical view that finite local rules suffice for global periodicity, with profound implications for fields like quasicrystals in , where atomic arrangements mimic such tilings, and , where they encode undecidable problems. This challenges assumptions in , showing that complexity can emerge inevitably from simple rules without periodic symmetry.

Historical Development

Early Discoveries in Aperiodic Sets

In 1966, Robert Berger proved the undecidability of the domino problem—the question of whether a given set of Wang tiles can the plane—by constructing the first known aperiodic set of tiles, consisting of 20,426 Wang tiles with colored edges that enforce matching rules. This set tiles the plane completely but admits no periodic tiling, demonstrating that algorithmic determination of tilability is impossible in general. Berger's construction relied on simulating computations through tile placements, where edge colors propagate information to prevent periodic arrangements. Building on Berger's work, Raphael Robinson simplified the proof of aperiodicity and undecidability in by introducing a smaller aperiodic set of six prototiles, each a square with specific markings on the corners to enforce local matching rules and propagate hierarchical structures. These markings create "influences" that force tiles to align in increasingly larger squares, ensuring non-periodicity while allowing infinite tilings. Robinson's innovation reduced the complexity dramatically from Berger's thousands of tiles, shifting focus toward more manageable prototile sets that incorporate internal rules beyond simple edge colors. Subsequent efforts in the late and further minimized aperiodic sets while refining matching mechanisms. In , Robert Ammann developed an aperiodic set of 16 Wang tiles, using edge colors derived from substitutions in octagonal tilings to enforce aperiodicity through forced expansions. This was followed in 1996 by Jarkko Kari's construction of a 14-tile aperiodic Wang set, based on simulating Beatty sequences via finite automata encoded in tile colors, which generate rotations incompatible with periodicity. Shortly thereafter, Karel Culik II refined Kari's approach to produce the smallest known Wang tile set at the time, with 13 tiles over five colors, again using automaton-based rules to simulate non-periodic dynamics. In 2015, Emmanuel Jeandel and further reduced the minimum to 11 tiles over four colors, and proved by exhaustive search that no smaller aperiodic set exists. These reductions marked a conceptual from Berger's computationally intensive rectangular tiles to prototiles with sophisticated edge or marking rules that embed aperiodic order through local constraints and global hierarchies.

Penrose's Innovations

, a mathematical , developed his groundbreaking aperiodic tilings during the 1970s, motivated by the challenge of achieving fivefold rotational symmetry in plane tilings, which is impossible in periodic lattices due to crystallographic restrictions. His work was influenced by the intricate tessellations of artist , whose prints explored impossible geometries and non-repeating patterns, inspiring Penrose to seek tile sets that enforced aperiodicity while exhibiting local fivefold symmetry. In 1974, Penrose published his initial aperiodic tile set in the paper "The rôle of aesthetics in pure and applied mathematical research," consisting of six prototiles based on pentagonal shapes with specific markings to enforce matching rules and prevent periodic arrangements. This set represented a significant simplification over earlier abstract aperiodic constructions, such as those using large numbers of tiles, by prioritizing aesthetic and symmetry-driven designs that were visually compelling yet mathematically rigorous. Penrose later refined this approach, reducing the number of prototiles while preserving aperiodicity. A major breakthrough came in with Penrose's publication "Pentaplexity" in , where he introduced a minimal set of two rhombi (known as the P3 tiling) whose edge markings ensured only non-periodic s of the , further streamlining the construction to emphasize proportions and fivefold . Around the same time, he independently formulated the and pair (P2 tiling), another two-tile system that achieves the same aperiodic properties through and shapes with matching rules. These innovations built on complex precursors from prior aperiodic set research but made the concept accessible through simple, symmetry-focused geometries. Penrose's tilings gained widespread recognition through his collaboration with science writer , who popularized them in articles starting in 1977 and in the 1989 book Penrose Tiles to Trapdoor Ciphers, bridging mathematical theory with public interest. Notably, Penrose's work predated the 1982 experimental discovery of quasicrystals by —real materials exhibiting aperiodic order with fivefold symmetry—providing a theoretical framework that helped interpret these findings as projections of higher-dimensional lattices.

Specific Penrose Tilings

Original Pentagonal Tiling (P1)

The original pentagonal tiling, designated as P1, was developed by as an early approach to constructing aperiodic tilings exhibiting fivefold . It relies on a set of six prototiles featuring specific edge markings known as junctures, which dictate compatible adjacencies to prevent periodic arrangements. These junctures consist of curved lines or arcs drawn on the edges of the prototiles, ensuring that only matching configurations can join, thereby enforcing local rules that propagate globally. The prototiles consist of irregular pentagons along with derived shapes such as the , star (pentagram), , and additional pentagon variants, all incorporating the in their proportions for compatibility. These six distinct prototiles arise from basic forms with rotational symmetries and specific juncture placements, allowing for precise control over assembly. Assembly begins with small clusters of these prototiles, where edges must align such that junctures form continuous paths without breaks or overlaps. Compatible prototiles fit together to form supertiles—larger composite units with outlines approximating regular decagons—demonstrating local pentagonal . For instance, a basic supertile might consist of five prototiles surrounding a central one, their junctures interlocking to create a star-like decagonal boundary. A small patch of such a reveals radiating fivefold symmetries, with prototiles arranged in rosettes that echo the decagonal form. The tiling's aperiodicity is enforced through , where supertiles serve as building blocks for even larger decagonal structures, creating an of scales centered on decagonal stars. This self-similar , driven by the juncture rules, precludes translational periodicity, as any periodic would disrupt the required decagonal alignments at higher levels. Penrose's innovations in this laid the for subsequent aperiodic sets by demonstrating how local constraints could globally prohibit .

Kite and Dart Tiling (P2)

The and tiling, designated as P2, is one of Roger Penrose's aperiodic tilings using just two prototiles: the and the . The is a convex with interior angles measuring 72°, 72°, 72°, and 144°, arranged such that the 144° angle is opposite the longest diagonal. The is a non-convex ( at one vertex) with interior angles of 72°, 36°, 72°, and 216°, where the 216° reflex angle creates the indentation. Both prototiles feature edges of two distinct lengths, with the ratio of longer to shorter sides equal to the φ = (1 + √5)/2 ≈ 1.618, typically normalized so short edges measure 1 and long edges measure φ; this proportion embeds fivefold symmetry locally while ensuring global aperiodicity. To generate valid tilings, the prototiles must adhere to strict edge-matching rules that align and vertices correctly, thereby forbidding periodic configurations. These rules are commonly implemented via decorations on the tiles, such as bicolored vertices (e.g., black and white markings) that must match at shared edges or directed arrows along edges indicating orientation and type, ensuring that only compatible juxtapositions occur—such as a 72° vertex of a meeting the 216° vertex of a in specific ways. Without these enforcements, periodic tilings become possible, but the rules restrict assemblies to aperiodic ones, promoting hierarchical growth. Tiling construction begins with small patches that expand hierarchically, forming characteristic supertile motifs like the "house" (a kite paired with a dart along a long edge, resembling a simple roofed structure) and the "star" (five darts converging at a central vertex, evoking a pentagram outline). These basic units combine to create larger, self-similar clusters, such as the infinite star pattern starting from five darts or the infinite sun from five kites, yielding an unending cascade of non-repeating scales that produce the illusion of fivefold rotational symmetry across the plane without true long-range order. The ratio of kites to darts in large regions approximates φ, reinforcing the tiling's intrinsic scaling properties. Although the and prototiles are non-convex, the P2 tiling is topologically equivalent to Penrose's original pentagonal tiling (P1) and the tiling (P3), meaning they produce diffeomorphic coverings of the with identical local classes under the matching rules, but P2's darts allow for smoother, more fluid boundaries in finite approximations compared to the rigid pentagons of P1.

Rhombus Tiling (P3)

The rhombus tiling, designated as P3 in Penrose's classification, employs two prototiles: a thin rhombus with interior angles of 36° and 144°, and a thick rhombus with angles of 72° and 108°. Both prototiles are rhombi with equal side lengths, and their diagonals stand in the golden ratio \phi = \frac{1 + \sqrt{5}}{2} \approx 1.618, where the longer diagonal exceeds the shorter by this irrational proportion. This geometric relation ties the tiles intrinsically to five-fold symmetry, as the angles derive from divisions of the 360° circle by 5 and 10, facilitating non-periodic arrangements. To enforce aperiodicity, matching rules govern tile adjacency, typically implemented via directed arrows or indentations on edges that must align precisely—such as head-to-tail matching or compatible markings. These constraints prevent periodic repetitions and naturally generate hierarchical supertiles, including the characteristic "sun" (a central thick surrounded by five thin ones) and "" (a five-pointed of alternating rhombi) patterns, which serve as building blocks for larger structures. The convex nature of these offers practical advantages for physical realizations, as they can be more readily fabricated from materials like wood or metal without cuts, unlike alternative representations. Moreover, the is mutually equivalent to the and (P2), achievable by subdividing each thick into a and a half-kite, and each thin into two half-kites and a , preserving the aperiodic properties under this . A notable example is finite patches of the that approximate the pattern, a supertile exhibiting prominent five-fold bounded by straight edges, demonstrating how local rules propagate global aperiodicity.

Core Mathematical Features

Golden Ratio Integration

The , denoted φ and defined as φ = (1 + √5)/2 ≈ 1.6180339887, is a fundamental that satisfies the equations φ = 1 + 1/φ and φ² = φ + 1. This self-similar property, where φ appears in its own expansion as [1;1,1,1,...], underscores its role in generating hierarchical structures without periodic repetition, a key factor in the aperiodicity of Penrose tilings. The irrationality of φ prevents commensurate ratios in tile counts or dimensions, ensuring that no can extend infinitely across the plane. In the kite and dart prototiles (P2 tiling), the governs the edge lengths and diagonals, with short sides normalized to length 1 and long sides to length φ. The features two adjacent short sides and two long sides, with its long diagonal of length φ and short diagonal of length 1; the mirrors this but is , sharing the same side and diagonal s of 1:φ. These proportions derive from subdividing a regular , where the diagonal-to-side is precisely φ, enabling the tiles to approximate fivefold locally. For the rhombus prototiles (P3 tiling), the thin has acute angles of 36° and obtuse angles of 144°, while the thick has 72° and 108° angles, both sets derived from divisions of the 72° central angle in a regular pentagon (360°/5 = 72°). All edges are of equal length, but the diagonals follow the 1:φ ratio, with the longer diagonal spanning φ times the shorter one, mirroring the and geometry. For explicit ratios in the , if the width (short diagonal) is set to 1, the length (long diagonal) equals φ, ensuring compatibility in assemblies. The integration of φ extends to scaling in Penrose tilings, where enlarges linear dimensions by a factor of φ and areas by φ² ≈ 2.618, preserving the overall structure while introducing finer hierarchical details. This scaling reflects the nature of φ, allowing infinite subdivision without periodicity.

Local Pentagonal Symmetry

Penrose tilings demonstrate local pentagonal symmetry through clusters of tiles that form decagonal or star-shaped patterns exhibiting fivefold of 72 degrees around specific vertices, while the entire structure remains aperiodic without a repeating . These local configurations arise naturally from the tile geometries and matching rules, allowing finite patches to mimic the symmetry of a regular or , but such patterns cannot extend infinitely without defects in a periodic setting. In the rhombus tiling variant (P3), vertex figures reveal this symmetry through 7 distinct types of local arrangements, each involving pentagonal coordination where tiles meet at angles derived from multiples of 72 degrees. These vertex types ensure that the local environment around each point maintains fivefold rotational invariance up to a finite radius, contributing to the tiling's quasiperiodic character. Fivefold rotational symmetry proves incompatible with translational periodicity in the plane because the irrational rotation angles, rooted in the golden ratio, cannot align with a lattice's rational translational vectors, leading to unavoidable mismatches over large distances. Diffraction patterns from Penrose tilings further highlight this local , producing discrete Bragg peaks arranged in a 10-fold symmetric array that reflects the underlying without implying long-range order.

Construction Techniques

Inflation Procedures

The procedure for Penrose tilings is a generative method that constructs larger supertiles by subdividing each prototile into a cluster of smaller copies of the prototiles, scaled by a linear factor of the \phi = \frac{1 + \sqrt{5}}{2} \approx 1.618, resulting in an area increase by \phi^2 \approx 2.618 per . This self-similar rule ensures that repeated applications produce increasingly complex tilings while inherently enforcing the aperiodic structure through the geometric constraints of the prototiles. In the kite and dart tiling (P2), the inflation rule specifies that each kite supertile is decomposed into two smaller kites and two smaller half-darts, while each dart supertile is decomposed into one smaller kite and two smaller half-darts. These substitutions maintain edge-to-edge matching and cover the original tile without gaps or overlaps, with the overall pattern expanding hierarchically (the half-darts pair with adjacent tiles to form full darts). For the rhombus tiling (P3), which uses a thin rhombus (with angles $36^\circ and $144^\circ) and a thick rhombus (with angles $72^\circ and $108^\circ), the inflation proceeds as follows: the thin rhombus subdivides into two smaller thin rhombi and one smaller thick rhombus, whereas the thick rhombus subdivides into two smaller thick rhombi and one smaller thin rhombus. This process similarly scales by \phi linearly, preserving the decagonal symmetry and aperiodic ordering. The hierarchical nature of inflation allows tilings to be generated iteratively from a finite seed patch, such as a single prototile or a small valid cluster, through infinite subdivisions that fill the plane densely while prohibiting periodic arrangements due to the irrational scaling factor \phi. This method underscores the self-similarity inherent in Penrose tilings, where each level reveals patterns that mirror the whole at larger scales.

Deflation and Decomposition Methods

Deflation serves as the inverse process to in Penrose tilings, enabling the systematic of a given tiling by identifying and grouping clusters of smaller prototiles to form larger supertiles according to the reverse of the inflation substitution rules. This method allows analysts to reduce any valid Penrose tiling iteratively, confirming its adherence to the underlying aperiodic structure by converging to the original set of prototiles, such as kites and or thin and thick rhombi. Unlike , which builds hierarchical supertiles from basic tiles by subdivision, groups tiles to verify consistency across scales, often revealing the unique prototile set regardless of the starting configuration. In the kite and dart tiling (P2), deflation proceeds by identifying supertile boundaries and grouping: a larger kite is formed from two smaller kites and two smaller half-darts (pairing to one full dart), while a larger dart is formed from one smaller kite and two smaller half-darts; these groupings are scaled by the factor \phi^{-2} in area, where \phi = (1 + \sqrt{5})/2 is the golden ratio, ensuring the areas match and the process tiles the plane without gaps or overlaps. For the rhombus tiling (P3), a similar approach applies: a larger thin rhombus is composed of two smaller thin rhombi and one smaller thick rhombus, and a larger thick rhombus of one smaller thin rhombus and two smaller thick rhombi, again scaled by \phi^{-2}, with supertile boundaries delineated by tracing edges that align with the inflation rules in reverse. Repeated application of these steps on any finite or infinite valid tiling converges to the fundamental prototile set, demonstrating the self-similar nature of the structure. A key decomposition technique involves breaking down Penrose prototiles into Robinson triangles, named after Raphael M. Robinson, who introduced them as an auxiliary tool for analyzing aperiodic tilings. These are two types of isosceles triangles: the "acute" triangle with angles 36°, 72°, and 72°, and the "obtuse" triangle with angles 108°, 36°, and 36°. In the rhombus tiling (P3), each thin rhombus is composed of two acute triangles joined along their long edges, while each thick rhombus consists of one acute and one obtuse triangle similarly joined; for the kite and dart tiling (P2), the kite divides into two acute triangles, and the dart into one acute and one obtuse. To decompose a full tiling, one draws the short diagonals across each prototile to separate them into these triangles, then applies deflation rules to the triangles themselves—such as grouping for the reverse of subdividing the acute triangle into two acute and two obtuse smaller ones, and the obtuse into two acute and one obtuse—scaled appropriately to maintain the hierarchy. This triangular breakdown facilitates precise boundary identification and subdivision at each level. The and methods, particularly via Robinson triangles, provide a practical means to verify that local matching rules in Penrose tilings propagate consistently to global scales, as iterative reduction ensures no inconsistencies arise in the hierarchical structure without invoking full aperiodicity arguments. By repeatedly , one can trace how edge markings or constraints enforced locally extend through supertile levels, confirming the tiling's validity and uniqueness within the aperiodic family.

Matching Rules and Enforcement

Matching rules in Penrose tilings consist of local constraints on how tile edges or vertices must align to form valid configurations, ensuring that only aperiodic tilings can cover the without gaps or overlaps. These rules can be realized through several mechanisms, including edge vectors where directional arrows on tile edges must to zero at every , color matchings requiring adjacent edges to share the same color designation, or physical notches that permit only in approved orientations. In the kite and dart tiling (P2), matching rules are enforced by specific geometric features: the dart's concave edge fits precisely into the kite's corresponding convex edge, while additional markings—such as aligned arcs or lines on edges—dictate permissible adjacencies. These rules designate distinct vertex types, including "sun" vertices where five kites converge and "star" vertices where five darts meet, alongside mixed configurations that maintain local fivefold symmetry. Local prohibitions, such as barring two darts from adjoining along their long edges, further constrain placements. Such local rules enforce an infinite hierarchical structure, as satisfying them at one scale necessitates consistent patterns at progressively larger scales, inherently precluding periodic arrangements. serves as a tool to confirm adherence by breaking down the into constituent prototiles. Consequently, every valid under these rules adheres to the underlying framework, guaranteeing aperiodicity across all possible configurations.

Properties and Implications

Global Aperiodicity Proofs

Global aperiodicity proofs for Penrose tilings demonstrate that no valid tiling by kite and dart (P2) or thin and thick rhombi (P3) prototiles can exhibit translational periodicity, relying on the inherent role of the golden ratio \phi = \frac{1 + \sqrt{5}}{2} and the hierarchical structure induced by inflation and deflation rules. A foundational argument assumes a periodic tiling exists with some finite translation vector, implying that the entire plane can be covered by repeating a fundamental domain. However, the inflation procedure scales tile sizes by powers of \phi, which is irrational, leading to supertile positions that cannot align periodically without contradicting the irrational scaling factors. Specifically, if a tiling were periodic, the relative frequencies of tile types (such as the ratio of thick to thin rhombi approaching \phi : 1) would be rational, but the irrationality of \phi ensures this ratio remains irrational, deriving a contradiction. The hierarchy argument employs deflation, the inverse of inflation, to show infinite descent is possible only in non-periodic tilings. In a valid Penrose tiling, repeated deflation decomposes larger supertiles into smaller ones indefinitely, creating an infinite hierarchy of scales without reaching a base case that could repeat periodically. If periodicity held, deflation would eventually map the fundamental domain to a finite set of prototiles that tile periodically at every level, but the scaling by \phi^{-n} (where \phi^{-1} = \phi - 1) disrupts this due to the irrationality, preventing closure under a finite period. This infinite regress implies no finite translational symmetry can govern the entire tiling. Formal results characterize the space of all Penrose tilings as a fiber bundle over the 2-torus with Cantor set fibers, where the fiber is a compact, totally disconnected perfect metric space with no isolated points, confirming the absence of any non-trivial translational symmetries. This topological structure, arising from the matching rules and substitution dynamics, ensures that while individual tilings cover the plane with uniform density (e.g., the area proportion of tile types stabilizes to values involving \phi), vertex positions exhibit quasiperiodic order rather than strict periodicity.

Spectral and Diffraction Analysis

The analysis of Penrose tilings treats the tiling as a point set, typically the positions of vertices or centers, and computes its to reveal the pattern's structural order. In the infinite limit, the consists of a pure point measure, characterized by sharp Bragg peaks without diffuse scattering. These peaks occur at wavevectors belonging to a forming a decagonal , where the positions are generated as integer combinations of basis vectors in directions rotated by multiples of \pi/5, with amplitudes scaled by powers of the \phi = (1 + \sqrt{5})/2. This decagonal arrangement reflects the underlying 10-fold inherent in the tiling's local pentagonal features. The intensity of the pattern is formally given by the I(\mathbf{k}) \propto \left| \sum_{j} e^{2\pi i \mathbf{k} \cdot \mathbf{x}_j} \right|^2, where \mathbf{x}_j are the positions in the point set, and the is taken in the via the measure. For Penrose tilings constructed as model sets from a 5-dimensional cubic , this yields Bragg peaks whose and relative intensities are determined by the \mathbb{Z}[\phi] in the physical space, ensuring a hierarchically structured with intensities decaying according to the window function's . Spectral properties of the associated arise from the inflation (substitution) rule governing the tiling's . The for the thin and thick rhombs in the P3 Penrose tiling is \begin{pmatrix} 2 & 1 \\ 1 & 1 \end{pmatrix}, with eigenvalues \phi^2 and $1/\phi^2. These Pisot-Vijayaraghavan eigenvalues imply a pure point for the translation dynamics on the of Penrose tilings, meaning the eigenfunctions form a basis of continuous functions on the tiling space, which in turn guarantees the pure point nature of the . This spectral and diffraction behavior demonstrates how Penrose tilings emulate the long-range order of crystals in producing discrete Bragg peaks, yet incorporate forbidden rotational symmetries, thereby providing a mathematical bridge between aperiodic geometry and physical quasicrystal phenomena.

Extensions and Applications

Connections to Quasicrystals

In 1982, Dan Shechtman observed a diffraction pattern with 10-fold symmetry in a rapidly solidified aluminum-manganese alloy containing 14% manganese, challenging the prevailing crystallographic dogma that crystals must exhibit translational periodicity. This discovery, published in 1984, revealed an icosahedral phase with long-range orientational order but no translational symmetry, which was later interpreted through the lens of Penrose tilings as a physical manifestation of aperiodic order. Quasicrystals are mathematically modeled as cut-and-project sets derived from higher-dimensional periodic lattices, where a slice of the lattice is projected onto lower-dimensional space, yielding structures that approximate the aperiodic arrangements of Penrose tilings. This method, introduced shortly after Shechtman's observation, provides a rigorous framework for understanding the quasiperiodic atomic arrangements in these materials, linking the discrete, non-repeating patterns of Penrose tilings to the diffuse yet sharp diffraction patterns observed experimentally. Decagonal quasicrystals, such as those in the Al-Co-Ni system, serve as direct physical analogs to the pentagonal (P3), where atomic clusters arrange in a quasiperiodic fashion mimicking the tiling's fivefold symmetry and local pentagonal motifs. High-resolution imaging and structural refinements of Al-Co-Ni alloys confirm this correspondence, with the material's decagonal superstructure aligning to rhombic or pentagonal Penrose frameworks, demonstrating stable aperiodic order under thermodynamic conditions. In recognition of this , was awarded the 2011 for the of quasicrystals, explicitly crediting the foundational of Penrose tilings in providing the conceptual and mathematical basis for interpreting their aperiodic structures. In 2023, researchers Zhi Li and Latham Boyle demonstrated that Penrose tilings are mathematically equivalent to a form of quantum error-correcting code, where the non-repeating patterns enable safeguarding through local indistinguishability, allowing tolerance to localized errors without revealing global structure. This connection, published on in November 2023, opens applications of Penrose tilings in and . The Ammann-Beenker tiling is an aperiodic tiling of the plane featuring eightfold rotational symmetry, constructed using two prototiles: and with acute angles of 45 degrees. These tiles are related by the , $1 + \sqrt{2}, analogous to the role of the in Penrose tilings, and aperiodicity is enforced through matching rules that generate a hierarchical structure without translational periodicity. The tiling can be generated via substitution rules, where larger supertiles are composed of smaller ones in a self-similar fashion, ensuring non-periodic arrangements. Other variants of Penrose tilings include the sphinx tiling, which uses two prototiles—a larger sphinx-shaped hexiamond and a smaller companion —and relies on rules to produce aperiodic coverings with fivefold . Similarly, the F-tiling employs two irregular pentagonal prototiles, the F-tile and another complementary shape, also admitting only non-periodic tilings through substitution methods that build infinite hierarchies. These variants, like the standard kite-and-dart or Penrose sets, demonstrate how minimal prototile counts can force aperiodicity via geometric constraints. Generalizations to three dimensions extend Penrose principles to icosahedral , using two rhombohedral prototiles: a prolate (obtuse) and an acute one, whose edges correspond to projections from a five-dimensional . Matching rules on faces ensure that valid tilings of are aperiodic, forming structures analogous to two-dimensional Penrose tilings but with icosahedral . These 3D tilings maintain hierarchical , with supertiles decomposing into smaller copies under . Substitution systems for aperiodic tilings trace back to Berger's 1966 construction of an aperiodic set with over 20,000 prototiles, which used hierarchical substitutions to simulate undecidable computations and force non-periodicity. Robinson later refined this approach in 1971, developing a set of six square prototiles with markings that enforce aperiodicity through macro-tile hierarchies, where larger squares contain smaller ones in a non-repeating . Kari's contributions include a 1996 aperiodic set of 14 tiles, where edge colors simulate multiplication of irrational Beatty sequences via Mealy machines, ensuring only non-periodic tilings are possible. These structures, including the Ammann-Beenker and Penrose variants, enforce aperiodicity through similar hierarchical rules that propagate constraints globally, preventing periodic repetitions. However, they differ in underlying : Penrose tilings exhibit decagonal (fivefold) , while Ammann-Beenker features octagonal (eightfold) , and 3D extensions introduce icosahedral , each tied to distinct irrational ratios like the or silver mean. , Robinson, and Kari's systems emphasize computational undecidability via tile markings, contrasting with the geometric focus of Ammann and Penrose but sharing the core mechanism of infinite, non-repeating hierarchies.

Implementations in Art and Design

Penrose tilings have inspired numerous artistic works, particularly those drawing from the visual motifs of , whose tessellations influenced Roger Penrose's development of aperiodic patterns in the . Escher, in turn, incorporated modified Penrose rhombi into his final tessellation artwork, creating intricate, non-repeating designs that blend mathematical precision with . This mutual influence led to exhibitions such as "Nadir and Zenith in the World of Escher," which featured Penrose tilings alongside Escher's prints to highlight their shared exploration of symmetry and infinity. Digital generations of Penrose patterns have further expanded this artistic legacy, with artists using software to produce infinite, non-periodic motifs for prints and installations, as seen in works exhibited at the Bridges Conference on mathematical art. In architecture, Penrose tilings offer a visually dynamic alternative to periodic patterns, often implemented using the rhombus variant for its relative ease in physical construction. Early installations appeared in the 1980s, including floor pavings in academic buildings that showcased the tilings' fivefold symmetry. Notable examples include the courtyard tiling at University's Bachelor Hall, completed in the late 1980s as an artistic and educational feature demonstrating aperiodic order. More recent applications extend to facades and public spaces, such as the stone and clay Penrose flooring outside the Mathematical Institute at the , which creates an unpredictable yet harmonious surface. Similarly, the entrance paving at the Simons Center for Geometry and Physics at employs Penrose tiles to evoke mathematical elegance in a built environment. Roger Penrose secured a in 1979 for a set of two rhombus-shaped tiles designed to produce non-periodic coverings, intended for applications in , , and decorative surfaces to ensure unique, non-repeating patterns that resist straightforward replication. This , titled "Set of tiles for covering a surface," emphasized the tiles' ability to form aperiodic tilings with fivefold symmetry, influencing subsequent designs in both industrial and artistic contexts. The invention's focus on practical enforcement of aperiodicity through tile shapes has been cited in discussions of for geometric designs, underscoring its role in protecting innovative patterns. Contemporary implementations leverage digital fabrication for broader accessibility, including 3D-printed sculptures and jewelry that translate Penrose patterns into tangible forms. For instance, designers like Omri Revesz created the "Penrose" jewelry collection in 2016 using to form interlocking, aperiodic motifs from , allowing for lightweight, customizable pieces that capture the tiling's infinite variety. Artists such as Debora Coombs and Duane Bailey have produced 3D-printed Penrose tiling sculptures, extending the two-dimensional patterns into volumetric art that explores spatial aperiodicity. Software tools facilitate these creations, with open-source projects utilizing libraries like for generating of Penrose tilings, enabling precise digital prototyping for prints, jewelry, and larger installations up to 2025. These modern uses highlight the tiling's enduring appeal in blending , craftsmanship, and aesthetic .

References

  1. [1]
    Pentaplexity A Class of Non-Periodic Tilings of the Plane
    Penrose, R. Pentaplexity A Class of Non-Periodic Tilings of the Plane. The Mathematical Intelligencer 2, 32–37 (1979).
  2. [2]
    [PDF] Automatic Generation of Penrose Empires - Computer Science
    May 19, 2000 · Penrose tilings are infinite tilings that cannot tile the plane in a periodic manner. They are of interest to researchers because they may ...
  3. [3]
    [PDF] What is... Penrose's Tiles? - OSU Math Department
    In 1973 and 1974 Roger Penrose (Emeritus Rouse Ball Professor of Mathematics at Oxford) discovered three sets of aperiodic tiles. We will work with the kite (K).<|control11|><|separator|>
  4. [4]
    [PDF] The Ammann–Beenker Tilings Revisited - HAL
    Oct 22, 2015 · Penrose, The Role of aesthetics in pure and applied research, Bull. Inst. Maths. Appl. 10 (1974). [8] N. Wang, H. Chen, K. Kuo, Two ...
  5. [5]
    [PDF] Decagonal and Quasi-Crystalline Tilings in Medieval Islamic ...
    Feb 23, 2007 · The most famous example of a quasi- crystalline tiling is the Penrose tiling (20, 22), a. two-tile tessellation with long-range quasi- periodic ...
  6. [6]
    [PDF] Introduction to hierarchical tiling dynamical systems
    Coincidentally, in 1982 Alan Mackay [77] published the diffraction pattern of a Penrose tiling, shown in its original form in figure 1.4. Once Shechtman's ...
  7. [7]
    [PDF] A guide to Penrose tilings - arXiv
    Oct 29, 2023 · The paper. [Pen78] contains an account of how Penrose discovered his sets with 6 and 2 tiles, and a nice Escheresque non-periodic tiling with ...
  8. [8]
    [PDF] Tilings∗ - MIT Mathematics
    A tiling problem asks us to cover a given region using a given set of tiles, com- pletely and without any overlap. Such a cov- ering is called a tiling. Of ...
  9. [9]
    Tiling the plane, periodic and aperiodic - Paul Bourke
    A periodic tiling is one where it is possible to make a parallelogram (generally larger than the tiles) that can be repeated to produce the same tiling.
  10. [10]
    [PDF] 3 TILINGS - CSUN
    Periodic tiling of En: A tiling, not necessarily monohedral, whose symmetry group contains an n-dimensional lattice. This definition can be adapted to in- clude ...
  11. [11]
    Periodic and Non-Periodic Tiling | - | Graham Shawcross PhD Architect
    Oct 12, 2012 · Periodic tilings, like hexagons, are fixed. Non-periodic tilings tile non-repeating but can also tile periodically. Only hexagons are truly  ...
  12. [12]
    Tessellation in Architecture from Past to Present - ResearchGate
    Based on Roman mosaics, the tessellation has an important place in architecture since the ancient times. Through the history, different patterns have been used ...
  13. [13]
    Undecidability and nonperiodicity for tilings of the plane
    Undecidability and nonperiodicity for tilings of the plane ... Article PDF. Download to read the full article text. Use our pre-submission checklist.
  14. [14]
    [PDF] Lecture 2: Domino Problem, Part I: Wang tiles.
    Dec 16, 2014 · Refuted by Berger (Wang's student) in 1966: he exhibited an aperiodic set of 20426 Wang tiles. Robinson (1971): aperiodic set of 56 Wang ...
  15. [15]
    Undecidability and Nonperiodicity for Tilings of the Plane. - EuDML
    Robinson, Raphael M.. "Undecidability and Nonperiodicity for Tilings of the Plane.." Inventiones mathematicae 12 (1971): 177-209.
  16. [16]
    [PDF] Yet Another Proof of the Aperiodicity of Robinson Tiles - LIPN
    Robinson, Undecidability and nonperiodicity for tilings of the plane, Invent. Math. 12 (1971), 177-–209. 1One checks that this can be done in a unique way ...
  17. [17]
    [PDF] Aperiodic hierarchical tilings - Chaim Goodman-Strauss
    He gave such a set, of some 20,000 tiles. By 1971 R. Robinson had produced a much smaller set of six prototiles and streamlined Berger's general result as ...
  18. [18]
    [PDF] An aperiodic set of 11 Wang tiles - HAL Inria
    Jun 22, 2015 · Wang tiles are square tiles with colored edges. A tiling of the plane by Wang tiles consists in putting a Wang tile in each cell of the grid ...
  19. [19]
    A small aperiodic set of Wang tiles - ScienceDirect.com
    A new aperiodic tile set containing only 14 Wang tiles is presented. The construction is based on Mealy machines that multiply Beatty sequences of real ...
  20. [20]
    An aperiodic set of 13 Wang tiles - ScienceDirect.com
    A new aperiodic tile set containing only 13 tiles over 5 colors is presented. Its construction is based on a recent technique developed by Kari.
  21. [21]
    How Surreal Artist MC Escher Influenced Physicist Roger Penrose
    Jan 9, 2023 · Penrose tiling really new ways of thing of seeing visual representation of fiery fundamental geometric and algebraic transformations and things.
  22. [22]
    The Roger Penrose puzzle | Museum Escher in The Palace
    In 1974 Penrose invented an aperiodic tiling which consists of only two basic shapes.
  23. [23]
    Tilings and Quasi-Crystals; a Non-Local Growth Problem?
    Penrose, 1974. R. Penrose. The rôle of aesthetics in pure and applied mathematical research. Bull. Inst. Math. Applications, 10 (no. 7/8) (1974), pp. 266-271.<|control11|><|separator|>
  24. [24]
    [PDF] Penrose Tiles To Trapdoor Ciphers - Index of /
    The Spectrum Series of the Mathematical Association of America was so named to reflect its purpose: to publish a broad range of books including biographies,.
  25. [25]
    Penrose tiling observed in a quasi-crystal - Nature
    Jul 4, 1985 · High-resolution electron microscope images of this alloy, reported here, reveal the localized pentagonal pseudo-symmetry of the structure.
  26. [26]
    [PDF] 32 Pentaplexity* - UT Math
    The tilings have a number of remarkable properties, some of which were described in. Gardner's article. I shall give here a brief account explain- ing how these ...
  27. [27]
    [PDF] kite dart - Brown Math
    Nov 5, 2007 · Theorem 5.1 Any Penrose tiling with 5-fold symmetry is either P1 or P2. Proof: Let Q be a Penrose tiling with 5-fold symmetry. Let B be some ...
  28. [28]
    [PDF] Penrose Tiles - Linda Green
    These are the “kite” and “dart” Penrose tiles. What do you notice about ... approaches the Golden Ratio φ = 1 + /5. 2. = 1.6180339887.... ▷ Why is this ...
  29. [29]
    [PDF] Lectures on Penrose Tilings, Prof. Alexander F. Ritter. - People
    Sep 8, 2014 · A set F of tiles is called aperiodic if every tiling of the plane using copies of tiles from F is always non-periodic. That is, no matter how ...
  30. [30]
    [PDF] Penrose Tiling - University of Utah Math Dept.
    Jan 16, 2020 · Here is a way to make a nonperiodic tiling with rhombus tiles. As we have seen, the triangle and rhombuses may tile the plane periodically or.
  31. [31]
    [PDF] A Game of Life on Penrose Tilings - Cornell Mathematics
    A tiling is periodic if its symmetry group includes translations in two nonparallel directions. Otherwise the tiling is nonperiodic. Page 11. nonperiodic ...
  32. [32]
    Penrose Rhomb - Tilings Encyclopedia
    Discovered in 1973 and 1974 by R. Penrose in - at least - three versions (Rhomb, Penrose kite-dart and Penrose Pentagon boat star), all of them forcing ...Missing: Eureka | Show results with:Eureka
  33. [33]
    Penrose Tiles - Steve Dutch
    In 1974, Penrose discovered an aperiodic tiling that uses only two shapes, nicknamed kites and darts. When people use the term "Penrose tiling", this is usually ...
  34. [34]
    [PDF] Algebraic theory of Penrose's non-periodic - Brown Math
    There are two kinds of hexagons, which we shall call D-hexagons and Q- hexagons, respectively. In the case that the singular line is a 0-line, the D ...
  35. [35]
    [PDF] The Empire Problem in Penrose Tilings - Computer Science
    May 8, 2006 · The darkened lines are the forced Ammann bars; when three forced bars form a particular isoceles triangle, a kite is forced.<|control11|><|separator|>
  36. [36]
    [PDF] Penrose tilings - OU Math
    Apr 20, 2008 · The kite and dart are then formed out of two golden triangles. This can be seen in Figure 5. Figure 5 also shows the ratio of kites to darts and ...
  37. [37]
    [PDF] Penrose Tilings as a Model of Quasicrystals - MrBertman.com
    Figure 12 shows some examples of these periodic tilings. In 1973, Roger Penrose, an English mathematician, discovered that there was a set of 6 tiles, known as ...Missing: Eureka | Show results with:Eureka<|control11|><|separator|>
  38. [38]
    Patch frequencies in rhombic Penrose tilings - PMC - PubMed Central
    ... Penrose rhombus tiling. The 2-boundary dual to Inline graphic is the ... Frequencies of vertex configurations in Penrose tilings, all belonging to Inline graphic ...
  39. [39]
    [PDF] Quasicrystals started out as a mathematical game. But one day the ...
    As we shall see, the golden ratio and pen- tagonal symmetry are embedded in the design of Penrose tilings in many won- drous ways. Any Penrose tiling can be ...
  40. [40]
    Penrose Tilings - UBC Math
    Penrose tilings of the plane. In 1973 Roger Penrose found a set of six tiles that force aperiodicity. Later, in 1974 he was able to reduce the set to two tiles.Missing: publications | Show results with:publications
  41. [41]
    Impossible Cookware and Other Triumphs of the Penrose Tile
    Apr 25, 2014 · Penrose made several versions of his aperiodic tile sets. One of his most famous is known as the “kite” and the “dart.” The kite looks like ...
  42. [42]
    [PDF] Penrose Tiling - Andrew Glassner
    Actually, Penrose found two distinct sets of aperiodic tiles, one called “kites and darts” and the other called “rhombs.” Each set is made of only two tiles.
  43. [43]
    [PDF] Aperiodic Hierarchical Tilings Chaim Goodman-Strauss
    Raphael Robinson gave the first really small set (and a simpler version of Berger's proof) in a lovely 1971 paper. Robinson's small aperiodic set: Page 14 ...
  44. [44]
    Mathematical diffraction of aperiodic structures - RSC Publishing
    Jul 16, 2012 · Mathematical diffraction of aperiodic structures. M. Baake and U. Grimm, Chem. Soc. Rev., 2012, 41, 6821 DOI: 10.1039/C2CS35120J. To request ...
  45. [45]
    Mathematical diffraction of aperiodic structures | Request PDF
    Mathematical diffraction of aperiodic structures. July 2012 · Chemical Society Reviews. Michael Baake · Uwe Grimm. Kinematic diffraction is well suited ...
  46. [46]
    A primer of substitution tilings of the Euclidean plane - ScienceDirect
    Another imperfect substitution uses the unmarked Penrose rhombs. ... eigenvalues of the substitution matrix play a critical role. We can consider ...
  47. [47]
    [PDF] Consequences of Pure Point Diffraction Spectra for Multiset ... - arXiv
    Oct 23, 2009 · Delone multisets that can be represented by substitution tilings using a concept of “legal cluster.” This allows us to move freely between both ...
  48. [48]
    [PDF] The DiSCovery of QuaSiCrySTalS - Nobel Prize
    The pentagonal Penrose tiling (figure 7) is a self-similar pattern with 5-fold symmetry, long- range order and no translational periodicity. de Bruijn applied ...
  49. [49]
    Metallic Phase with Long-Range Orientational Order and No ...
    Nov 12, 1984 · We have observed a metallic solid (Al-14-at.%-Mn) with long-range orientational order, but with icosahedral point group symmetry, which is inconsistent with ...
  50. [50]
    Quasicrystals: A New Class of Ordered Structures | Phys. Rev. Lett.
    Dec 24, 1984 · A quasicrystal is the natural extension of the notion of a crystal to structures with quasiperiodic, rather than periodic, translational order.
  51. [51]
    Decagonal quasicrystals – What has been achieved? - ScienceDirect
    The underlying tilings resemble either rhomb or pentagon Penrose tilings, hexagon–boat–star (HBS) tilings or Masakova tilings [38] that are based on decagonal ...
  52. [52]
    Atomic structure of an Al – Co – Ni decagonal quasicrystalline surface
    Jul 7, 2004 · Furthermore, the surface of the quasicrystal has symmetry different from the Penrose tiling. As mentioned previously, on one terrace of the Al ...
  53. [53]
    Dan Shechtman – Banquet speech - NobelPrize.org
    Dan Shechtman's speech at the Nobel Banquet in the Stockholm City Hall, 10 December 2011. Your Majesties, Your Royal Highnesses, ...<|control11|><|separator|>
  54. [54]
    Aperiodic tiles | Discrete & Computational Geometry
    Jul 1, 1992 · Download PDF · Discrete & Computational Geometry Aims and ... Cite this article. Ammann, R., Grünbaum, B. & Shephard, G.C. Aperiodic tiles.
  55. [55]
    [PDF] The Ammann–Beenker Tilings Revisited - HAL
    Oct 22, 2015 · This paper introduces two tiles whose tilings form a one-parameter family of tilings which can all be seen as digitization of two-dimensional.Missing: original | Show results with:original
  56. [56]
    (PDF) Riddles of the sphinx tilings - ResearchGate
    Apr 27, 2023 · The sphinx tile, a hexiamond composed of six equilateral triangles in the shape of a sphinx, has interesting and complex tiling behavior due ...Missing: original paper
  57. [57]
    Theory of Matching Rules for the 3-Dimensional Penrose Tilings
    Since their invention by Roger Penrose, his well-known aperiodic and five-fold symmetric tilings of the planes have motivated numerous works [1-3].
  58. [58]
    [PDF] The icosahedral quasiperiodic tiling and its self-similarity
    May 19, 2016 · We investigate a 3-dimensional analogue of the Penrose tiling, a class of 3-dimensional aperiodic tilings whose edge vectors are the vertex.
  59. [59]
    Markus Rissanen | 2013 Bridges Conference
    Statement. Inspired by the famous aperiodic Penrose tiling I started some years ago to study non-periodic rhombic tilings with rotational symmetric ...<|separator|>
  60. [60]
    The Penrose Tiling in the Bachelor Hall Courtyard of Miami University
    In 1964 a Harvard University doctoral student discovered such a set that used more than 20,000 differently shaped tiles. Later he was able to reduce the number ...Missing: date | Show results with:date
  61. [61]
  62. [62]
    Penrose Tiling – SCGP - Simons Center for Geometry and Physics
    Dec 1, 2015 · The paving pattern outside the ground entrance to the Simons Center for Geometry and Physics follows a design invented by Roger Penrose in the early 1970s.Missing: 1980s installations
  63. [63]
    US4133152A - Set of tiles for covering a surface - Google Patents
    A set of tiles for covering a surface is composed of two types of tile. Each type is basically quadrilateral in shape and the respective shapes are such thatMissing: UK | Show results with:UK
  64. [64]
    Mathematical shapes and intellectual property – a legal ... - The IPKat
    Jun 23, 2023 · In 1974, British Nobel Laureate Sir Roger Penrose proposed solving the problem using two different tile shapes and proved that the resulting ...Missing: 1979 | Show results with:1979
  65. [65]
    Maison 203 previews new 3D printed jewelry
    Jun 30, 2016 · Penrose, the new collection of 3D printed jewelry created by the Israeli designer Omri Revesz for Maison 203, recaptures the tiling invented ...Missing: modern | Show results with:modern
  66. [66]
    Three Dimensional Penrose Tiling by Debora Coombs and Duane ...
    Jan 28, 2021 · Debora Coombs and Duane Bailey discuss their three dimensional Penrose tiling sculptures ... Self-assembling material pops into 3D. Steve Mould• ...
  67. [67]
    This project generates a SVG file with a Penrose tiling ... - GitHub
    This project generates a SVG file with a Penrose tiling (an aperiodic tiling of the plane with two tiles), using the cairo-fortran bindings as a fpm ...Missing: tools | Show results with:tools