Fact-checked by Grok 2 weeks ago

Rotational symmetry

Rotational symmetry is a fundamental geometric property where a figure or object appears identical to itself after around a fixed point, known as the center of . In the discrete case, this occurs for rotations by multiples of 360°/n, where n is the finite of symmetry; in the continuous case, the figure is invariant under any ( ). This invariance under distinguishes it from other symmetries like , as rotations preserve the of the figure. For discrete rotational symmetry, the order n quantifies the symmetry, indicating the number of distinct rotations—ranging from 0° () to multiples of 360°/n—that map the figure onto itself. For example, an has order 3, remaining unchanged under rotations of 120° and 240°; a , by contrast, exhibits continuous rotational symmetry of order. In group theory, the set of discrete rotations forming rotational symmetry constitutes a Cn, closed under composition, with the and inverses ensuring the group's structure. Properties include associativity of rotations and a single fixed point at . Beyond two dimensions, rotational symmetry generalizes to m-dimensional via the special SO(m), where SO(3) in three dimensions underlies the conservation of in physical laws. These symmetries are crucial in , , and physics for analyzing invariant structures.

Fundamentals

Definition

Rotational symmetry is a geometric property exhibited by a figure or object that remains invariant—appearing unchanged—when rotated by specific angles around a center of rotation. In two dimensions, this center is a fixed point; in three dimensions, it is a fixed axis. This transformation preserves the distances and angles between points, mapping the object precisely onto itself without altering its overall appearance. In the Euclidean plane, such rotations are isometries that fix the center point while moving all other points along circular arcs. In three-dimensional space, rotations are isometries that fix all points along the axis while moving other points along circular paths perpendicular to the axis. For instance, a square demonstrates this symmetry, as it looks identical to its original orientation after a 90-degree rotation about its center. This form of symmetry differs fundamentally from other types, such as , which involves mirroring the object across a line or plane, potentially reversing its orientation and eliminating — the property of existing in non-superimposable mirror-image forms. In contrast, rotational symmetry maintains the object's and does not require a mirror line; objects with pure rotational symmetry can thus be chiral. Similarly, it is distinct from translation symmetry, which shifts the entire object along a straight line without fixing any point, resulting in no central invariance. At its core, rotational symmetry relies on the concept of invariance under a , where the —a —leaves the object's essential features unaltered, even if individual points are repositioned. This property is fundamental in and extends to both two-dimensional figures in the and three-dimensional objects in space, provided the rotation axis passes through the designated center.

Order of Symmetry

The order of rotational symmetry, denoted as n, is a quantitative measure that indicates the number of times a figure can be rotated by equal angles around its center while appearing unchanged, completing a full 360° rotation after n such steps. This finite value applies to discrete rotational symmetries, where the object maps onto itself only for specific rotation angles. For instance, an has an order of 3, as it coincides with itself after rotations of 120°, 240°, and 360°. The smallest rotation angle \theta that preserves the figure is given by \theta = \frac{360^\circ}{n} = \frac{2\pi}{n} \text{ radians}. This angle divides the full circle evenly, and higher orders correspond to finer divisions, such as n = 6 for a regular hexagon with \theta = 60^\circ. The order thus characterizes the periodicity of the symmetry, with n = 1 indicating no non-trivial rotational symmetry beyond the full turn. In contrast, objects exhibiting continuous rotational symmetry, such as a , possess an order because they remain invariant under rotations by any arbitrary , not limited to discrete positions. This order reflects the absence of preferred orientations, allowing the figure to overlay itself perfectly for every possible rotation within 360°.

Mathematical Framework

Discrete Rotational Symmetry

Discrete rotational symmetry describes the property of an object or figure that remains under rotations by multiples of a fixed \frac{2\pi}{n} radians around a central point or , where n is a positive greater than or equal to 1. This symmetry is formalized mathematically as the C_n, which is the generated by a single element of n. The elements of C_n are the n distinct rotations r^k for k = 0, 1, \dots, n-1, where r denotes the generator by \frac{2\pi}{n}, and r^n is the . As an abstract group, C_n satisfies the group axioms: it is closed under composition, meaning the composition of any two rotations in the group yields another rotation in the group; it contains the identity element corresponding to a 0° (or $2\pi) rotation; and every element has an inverse, given by the rotation in the opposite direction by the same angle. These properties ensure that the set of discrete rotations forms a well-defined algebraic structure, with the order of the group equal to n, referencing the order of rotational symmetry discussed earlier. For instance, a square exhibits C_4 symmetry, invariant under rotations by 90°, 180°, and 270° around its center. In two dimensions, discrete rotational symmetry around a point is represented by the standard applied to coordinate vectors. For a rotation by \theta = \frac{2\pi k}{n}, the transformation is given by \begin{pmatrix} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{pmatrix}, which preserves distances and orientations in the plane. This matrix form arises from the linear transformation that rotates the basis vectors, ensuring the figure maps onto itself for each k. Such symmetries contribute to the point groups in two and three dimensions. In , pure discrete rotational symmetries generate the cyclic subgroups of rosette groups, which describe symmetries fixing a point in the , such as those of regular polygons. In contrast, point groups incorporating discrete rotations, like those of polyhedra (e.g., the icosahedral group with 5-fold axes), extend to finite rotations around one or more axes, but the fundamental single-axis case remains cyclic C_n.

Continuous Rotational Symmetry

Continuous rotational symmetry refers to the invariance of an object or system under rotations by any arbitrary angle θ within the interval [0, 2π), distinguishing it from discrete cases by allowing full rotational freedom without a minimal nonzero angle. In two dimensions, this symmetry is embodied by the special SO(2), which consists of all 2×2 orthogonal matrices with 1, representing rotations around the origin in the . Similarly, in three dimensions, continuous rotational symmetry is captured by the group SO(3), comprising all 3×3 orthogonal matrices with 1, which describe rotations in . The mathematical structure underlying continuous rotational symmetry is that of a , where the group operations are and the parameter space—typically the angle θ—is continuous and infinite-dimensional in the sense of admitting a manifold . For SO(2), the so(2) is one-dimensional, generated by rotations, reflecting the single degree of freedom in planar rotations. In contrast, so(3) for SO(3) is three-dimensional, corresponding to rotations about three independent axes, and exhibits a non-abelian structure due to the non-commutativity of successive rotations in 3D. A fundamental representation of continuous rotations in 2D utilizes the complex plane, where multiplication by the unit complex number e^{i\theta} = \cos\theta + i\sin\theta effects a counterclockwise rotation by angle θ around the origin. This exponential form arises from Euler's formula and directly parametrizes elements of SO(2). In 3D, rotations in SO(3) can be represented using unit quaternions, which provide a compact, singularity-free parametrization via the map from the unit sphere in four dimensions to the rotation group, leveraging the double-covering homomorphism from SU(2) to SO(3). These representations highlight the smooth, continuous nature of the symmetry, enabling interpolation between rotations and facilitating computations in fields requiring arbitrary angular transformations.

Multiple Axes and Invariance

In , objects can possess rotational about multiple axes that intersect at a single point, such as of the object, leading to a richer set of invariance properties compared to single-axis cases. For instance, a exhibits threefold rotational about axes passing through opposite vertices (four such axes, with rotations of 120° and 240°), fourfold about axes through the centers of opposite faces (three axes, with 90°, 180°, and 270° rotations), and twofold about axes through the midpoints of opposite edges (six axes, with 180° rotations). These multiple axes collectively generate the full rotational group of the , known as the octahedral group O, which has 24 and is isomorphic to the S_4. The invariance under combined rotations around these axes arises from the group structure formed by composing individual rotations, where the overall is generated by cyclic subgroups corresponding to each type, often involving semidirect products rather than direct products due to non-commutativity. For polyhedral symmetries, this results in finite groups such as the tetrahedral group T (order 12, isomorphic to A_4) for the , featuring threefold axes through vertices and twofold axes through edges, and the icosahedral group I (order 60, isomorphic to A_5) for the or , with fivefold axes through vertices, threefold through faces, and twofold through edges. The composition of successive s R_1(\theta_1) around \mathbf{u}_1 and R_2(\theta_2) around \mathbf{u}_2 is represented by the matrix product R = R_2 R_1, which yields another rotation in the special orthogonal group SO(3), but generally R_1 R_2 \neq R_2 R_1 unless the axes coincide. In two dimensions, rotational symmetries are confined to a single axis perpendicular to the plane, limiting the structure to cyclic or dihedral groups that are abelian, whereas three-dimensional space permits multiple non-collinear axes to intersect, enabling non-abelian polyhedral rotation groups like T, O, and I. This dimensionality difference restricts 2D objects to invariances under rotations about one effective axis, while 3D allows for the complex interplay of axes that defines higher-order symmetries in platonic solids.

Examples and Occurrences

Geometric Figures

Rotational symmetry in geometric figures is prominently exhibited by regular polygons in two dimensions and Platonic solids in three dimensions, where rotations about a map the figure onto itself while preserving distances and angles. These shapes serve as canonical examples due to their uniform construction from congruent regular polygonal faces meeting identically at each vertex. In two dimensions, an possesses rotational symmetry of order 3, meaning it appears unchanged after rotations of 120°, 240°, and 360° about its ; visually, each such cycles the three vertices to the position of the next, forming a closed path that traces the perimeter. exhibits order 4 rotational symmetry, invariant under 90°, 180°, 270°, and 360° rotations about its , where vertices map sequentially around the boundary, and opposite sides align perfectly after 180° turns. Similarly, a regular pentagon has order 5 symmetry, remaining superimposed after rotations of 72°, 144°, 216°, 288°, and 360° about its , with each advancing the five vertices to adjacent positions in a star-like or circumferential path. The circle represents the limiting case of infinite rotational symmetry, appearing identical under any angle of about its , as every point on the maps continuously to another without discrete steps. Extending to three dimensions, Platonic solids demonstrate rotational symmetries along multiple axes passing through their centers, with orders determined by the figure's regularity. The tetrahedron features four axes of order 3, each passing through a vertex and the centroid of the opposite face; a 120° rotation about such an axis permutes the three adjacent vertices cyclically while fixing the opposite face's orientation. The cube and its dual, the octahedron, share rotational symmetries including three axes of order 4 (through opposite face centers, allowing 90°, 180°, and 270° rotations that cycle four edges or faces), four axes of order 3 (through opposite vertices, cycling three faces), and six axes of order 2 (through midpoints of opposite edges, swapping pairs of faces). The dodecahedron and icosahedron possess even richer symmetries, with six axes of order 5 (through opposite vertices, rotating by 72° increments to map five adjacent faces or vertices), ten axes of order 3 (through face centers), and fifteen axes of order 2 (through edge midpoints). In each case, rotations map vertices to vertices and faces to faces, preserving the overall structure. Archimedean solids, which incorporate regular polygons of multiple types in a vertex-transitive arrangement, inherit similar high rotational symmetries from the Platonic solids but exhibit semi-regularity, allowing rotations that permute faces of different shapes while maintaining uniformity at vertices.

Natural and Artistic Patterns

In nature, rotational symmetry manifests prominently in biological structures, particularly through radial arrangements that enhance functionality. , or sea stars, exemplify discrete rotational symmetry of order 5, with their five arms radiating from a central disk, a pentaradial pattern that evolved in adult echinoderms for efficient environmental interaction. Similarly, many flowers display rotational symmetry in their petals, typically of orders 3 to 8; for instance, lilies often exhibit order 3 rotational symmetry with six identical tepals (three petals and three sepals), while sunflowers approximate continuous rotational symmetry through densely packed florets arranged in spirals. The shell provides an example of approximate continuous rotational symmetry via its growth, where each chamber expands outward in a self-similar curve that maintains angular consistency over rotations, approximating invariance under arbitrary angles. Artistic and cultural creations frequently incorporate rotational symmetry to evoke harmony and introspection. Mandalas, originating in Hindu and Buddhist traditions, often feature discrete rotational symmetry of orders 4 to 12, with intricate patterns radiating from a center to symbolize the universe's cyclical nature and aid meditation. In Islamic art, rosette patterns in geometric tiles demonstrate rotational symmetry, such as 8-fold or 10-fold arrangements in mosque decorations, where star-like motifs repeat under specific rotations to create infinite, non-figural designs that reflect divine order. Leonardo da Vinci's Vitruvian Man (c. 1490) approximates order 4 rotational symmetry through the superposition of a circle and square, with the human figure's limbs positioned to align under 90-degree rotations, illustrating Renaissance ideals of proportional balance in the human form. Radial rotational symmetry in often serves evolutionary roles, particularly in sessile or slow-moving , by enabling sensing that aids in predator avoidance; for example, the uniform arm distribution in allows threat detection from any angle without directional bias. Aesthetically, rotational symmetry holds cultural significance across societies, symbolizing balance and cosmic equilibrium in art, as seen in mandalas where symmetrical repetition fosters a of and spiritual unity. Real-world instances of rotational symmetry are typically approximate due to environmental imperfections, deviating from ideal discrete orders. Snowflakes, for instance, ideally possess order 6 rotational symmetry from the of ice crystals, but vapor fluctuations during formation cause slight irregularities, resulting in unique, non-perfect patterns while retaining overall sixfold invariance.

Physical and Crystallographic Applications

In , the rotational symmetries compatible with periodic structures in three dimensions are restricted to rotation axes of orders 1, 2, 3, 4, and 6, resulting in 32 distinct s that describe the possible symmetry operations of . This limitation arises from the , which demonstrates that rotations of order 5 or higher cannot tile space periodically without gaps or overlaps, ensuring compatibility with translational symmetries. For instance, (SiO₂) belongs to point group 32 and features a principal 3-fold axis aligned with its c-crystallographic axis, allowing the crystal to appear identical after a 120° , which contributes to its piezoelectric properties. In physics, continuous rotational invariance of the or action principle implies the conservation of through , a foundational result linking symmetries to conserved quantities. This symmetry underpins the rotational dynamics of isolated systems, such as planetary orbits or motion, where total remains constant absent external torques. In , rotational invariance extends to the intrinsic property of , representing an internal for elementary particles like electrons, which transforms under rotations according to the particle's spin representation and contributes to the total of composite systems. The quantum mechanical manifestation of rotational invariance is captured by the commutation relation between the \hat{H} and the \hat{\mathbf{L}}: [\hat{H}, \hat{\mathbf{L}}] = 0, indicating that energy eigenstates can be simultaneously eigenstates of components, facilitating the in central potential problems like the . This ensures the conservation of in time evolution, as the SO(3) acts unitarily on the . In modern applications, rotational symmetry governs the structure of molecular orbitals, where the quantum number \ell determines the orbital's rotational character (e.g., s-orbitals with \ell=0 are spherically symmetric, while p-orbitals with \ell=1 exhibit directional lobes). In , fullerenes such as C₆₀ exemplify high-order rotational symmetry, possessing icosahedral (Iₕ) symmetry with 5-fold, 3-fold, and 2-fold rotation axes that dictate their closed-shell electronic structure and stability, enabling applications in carbon-based nanomaterials for electronics and . This symmetry influences the degenerate molecular orbitals of fullerenes, leading to unique optoelectronic properties exploited in photovoltaic devices.

Extensions and Relations

Combined with Other Symmetries

Rotational symmetry often combines with reflection symmetry to form dihedral groups, which describe the full set of symmetries for regular polygons in the plane. The dihedral group D_n consists of n rotations by multiples of $360^\circ/n around the center, paired with n reflections across axes passing through the center and vertices or midpoints of sides, yielding a total order of $2n. This structure arises as a semidirect product C_n \rtimes C_2, where C_n is the cyclic group of rotations and C_2 generates the reflections, with the reflection conjugating rotations to their inverses. For instance, the square's symmetries form D_4, encompassing four rotations (0°, 90°, 180°, 270°) and four reflections, for a group of order 8. In two dimensions, rotational symmetry combines with to produce frieze groups, which govern infinite strip patterns repeating along one direction, such as decorative borders. The frieze group denoted p2 features translations along the strip axis combined with 180° rotations about points midway between centers, without reflections or glides, enabling rotational motifs like alternating S-shapes in friezes. These seven frieze groups collectively incorporate rotations up to order 2 with translations, alongside possible reflections and glide reflections (translations composed with reflections parallel to the strip). Extending to three dimensions, rotational symmetry pairs with along the same to define screw axes, fundamental to crystallographic structures. A screw operation applies a by \theta about an followed by a by distance t parallel to that , denoted as n_m where n is the order (\theta = 360^\circ/n) and m/n is the fractional relative to the lattice repeat. Mathematically, for a point \mathbf{x} relative to the , the transformed position is given by \mathbf{x}' = R_{\theta} (\mathbf{x} - \mathbf{p}) + \mathbf{p} + t \hat{\mathbf{u}}, where R_{\theta} is the rotation matrix by \theta, \mathbf{p} a point on the axis, and \hat{\mathbf{u}} the unit vector along the axis; after n applications, the net effect is a full lattice translation. Glide operations, meanwhile, combine reflection with translation parallel to the reflection plane, further enriching these symmetries. In , these combinations—rotations with translations, reflections, screws, and glides—generate the full set of 230 groups, which classify all possible periodic crystal symmetries. These groups build upon the 32 point groups (pure rotations and reflections) by incorporating lattice translations, with screw axes and glide planes accounting for the majority of the 230 distinct types, as tabulated in the International Tables for Crystallography.

Group-Theoretic Representation

In group theory, rotational symmetry is formalized through the action of rotation groups, which capture the structure of transformations preserving orientation and distances in Euclidean space. These groups provide a unified algebraic framework for both discrete and continuous rotations, enabling the study of symmetries via abstract algebraic tools rather than geometric descriptions alone. The representation theory of these groups further allows symmetries to be realized as linear transformations on vector spaces, facilitating applications in diverse fields such as quantum mechanics. Rotational groups are realized as Lie subgroups of the O(n), which consists of all n \times n orthogonal matrices preserving the inner product, with the special SO(n) specifically comprising the orientation-preserving rotations (those with 1). For instance, in three dimensions, SO(3) parameterizes all possible rotations around the origin. Irreducible representations of these groups decompose the space of functions or states into fundamental building blocks invariant under rotations; in quantum applications, the irreducible representations of SO(3) correspond to multiplets, where the dimension of the representation is $2\ell + [1](/page/1) for integer or \ell, underpinning the classification of particle states and selection rules in . Finite rotational symmetries correspond to discrete subgroups, such as the cyclic group C_n generated by a rotation by $2\pi/n radians around a fixed axis, while continuous symmetries are modeled by infinite Lie groups like SO(3), which is compact and non-abelian. Point groups, which include rotations about multiple axes, admit finite discrete subgroups classified by their character tables—tabular summaries of traces of representation matrices under group elements. For the point group C_{3v}, which describes symmetries of an equilateral triangle with vertical mirror planes (e.g., rotations by 0, $120^\circ, $240^\circ and three reflections), the character table is as follows:
C_{3v}E$2C_3$3\sigma_vFunctions
A_1111z, z^2
A_211-1R_z
E2-10(x,y), (xz, yz)
This table reveals the one-dimensional totally symmetric representation A_1 and the two-dimensional irreducible E, essential for decomposing vibrational modes in molecular symmetry analysis. For finite rotational groups like C_n, there exists an to a of the S_n (the group of permutations of n elements), via , which embeds any finite group into a acting on itself by left multiplication; this realizes rotations as permutations of equivalent positions in a symmetric object. In the continuous case, on compact groups such as SO(3) generalizes classical through the Peter-Weyl theorem, decomposing square-integrable functions into matrix-valued coefficients over irreducible representations, enabling harmonic expansions for signals on the rotation manifold. The formalization of symmetry groups, including rotational ones, traces to the late , with Arthur Cayley's development of abstract in the 1850s providing the algebraic foundation, and Felix Klein's (1872) classifying geometries by their groups, thereby integrating rotations into a broader that influenced modern .

References

  1. [1]
    [PDF] 7 Symmetry and Group Theory - Penn Math
    More generally, we say that a shape has rotational symmetry of order n if it can be rotated by any multiple of 360 /n without changing its appearance. We can ...
  2. [2]
    [PDF] The Mathematics of Symmetry
    If we perform the basic 72 degree rotation 5 times, we bring the shape back to its starting position. We say that this shape has 5-fold rotational symmetry.
  3. [3]
    Symmetry - Department of Mathematics at UTSA
    Dec 14, 2021 · Rotational symmetry. The triskelion has 3-fold rotational symmetry. Rotational symmetry is symmetry with respect to some or all rotations in ...
  4. [4]
    Symmetry around a Point in the Plane
    ### Summary of Rotational Symmetry
  5. [5]
    Rotations - Brown Math
    Rotational Symmetry. Mathematical Definition: Rotation through an angle (counterclockwise) around the origin: Written Definition: A rotation fixes one point ...
  6. [6]
    Symmetry - Ximera - The Ohio State University
    Aug 30, 2025 · A shape has a reflection symmetry over a particular line when we reflect the shape over the line · A shape has rotation symmetry when we can ...
  7. [7]
    About Symmetries of the Plane
    Translation -- Slide the pattern for some distance along a straight line. Rotation -- Leave one point in the same place, and rotate the pattern through some ...
  8. [8]
    Introduction & Symmetry Operations - Tulane University
    Aug 27, 2013 · A symmetry operation is an operation that can be performed either physically or imaginatively that results in no change in the appearance of an object.
  9. [9]
    Rotational Symmetry - Definition, Order, Examples - Cuemath
    Rotational symmetry is defined as a type of symmetry in which the image of a given shape is exactly identical to the original shape or image in a complete turn.
  10. [10]
    Rotational Symmetry - BYJU'S
    The rotational symmetry of a shape explains that when an object is rotated on its own axis, the shape of the object looks the same.
  11. [11]
    Rotational Symmetry -- from Wolfram MathWorld
    Rotational symmetry of degree n corresponds to a plane figure being the same when rotated by 360/n degrees, or by 2pi/n radians.
  12. [12]
    Order of Rotational Symmetry - Basic Mathematics
    The order of rotational symmetry of a geometric figure is the number of times you can rotate the geometric figure so that it looks exactly the same as the ...<|control11|><|separator|>
  13. [13]
    Symmetry of a Circle | Corollary and Proof | Worksheets - Cuemath
    A circle has rotational symmetry. The order of the symmetry in a circle is infinite. That means if we rotate the circle by any degree of angle along its ...
  14. [14]
    Order of Rotational Symmetry of a Circle - onlinemath4all
    A circle has an infinite 'order of rotational symmetry'. In simplistic terms, a circle will always fit into its original outline, regardless of how many times ...
  15. [15]
  16. [16]
    [PDF] Rotation Matrices in two, three and many dimensions
    A real orthogonal matrix R with det R = 1 provides a matrix representation of a proper rotation. The most general rotation matrix R represents a ...
  17. [17]
    [PDF] High Symmetry Groups - MIT
    The point group of the Sphere is given the label K, and this is the point group used for free atoms in the gas phase. We are usually dealing with molecules, and ...<|separator|>
  18. [18]
    [PDF] Chapter 7 Continuous Groups, Lie Groups, and Lie Algebras
    The group SO(2) is simple enough that the full benefits of an in- finitesimal generator are not readily apparent. We will see in the next section, where we ...
  19. [19]
    [PDF] 7 Continuous Group - Xie Chen
    ... group. For example, the Lie algebra of the SO(2) group is so(2) and the Lie algebra of SO(3) is so(3). There are two important structures of this algebra:.
  20. [20]
    [PDF] The Quaternions and the Spaces S3, SU(2), SO(3), and RP
    Thus, SO(3) and RP3 are at the same time groups, topological spaces, and manifolds, and in fact they are Lie groups (see Marsden and. Ratiu [15] or Bryant [6]).
  21. [21]
    [PDF] Lie groups and Lie algebras (Winter 2024)
    Lie groups, by contrast, are about so-called continuous symmetries (more aptly called smooth sym- metries), such as translational or rotational symmetries of a ...
  22. [22]
    [PDF] MATH431: Complex Numbers - UMD MATH
    Sep 27, 2021 · Theorem 1.6. 1. Multiplication by the unit complex number eθı = cos θ + sin θı rotates the complex plane counterclockwise about the origin by θ ...Missing: operator | Show results with:operator
  23. [23]
    [PDF] Unit Quaternions and Rotations in SO(3) - UPenn CIS
    Mar 25, 2021 · The key to representation of rotations in SO(3) by unit quaternions is a certain group homomorphism called the adjoint representation of SU(2). ...
  24. [24]
    [PDF] Rotational Symmetries Of The Cube - maths.nuigalway.ie
    Example 1: Axis with faces. In this example we are doing the axis of rotation through axis A (see figure 8). If we do a clockwise rotation of 90° through the A ...Missing: multiple | Show results with:multiple
  25. [25]
    [PDF] Symmetry Groups of the Platonic Solids - George Sivulka
    Jun 1, 2018 · The Rotational Symmetry Group of a poly- hedron is the group of rotational symmetries of the rigid solid, denoted in this paper as SR(X).
  26. [26]
    CoordinateTransformations - Intelligent Motion Lab
    The composition of two rotation matrices R1R2 is also a rotation matrix. In general, however, R1R2≠R2R1 unless the two are rotations about a shared axis.
  27. [27]
    [PDF] Rotation group - OU Math
    Feb 19, 2010 · ... SO(3) is a Lie group called Spin(3). The group Spin(3) is isomorphic to the special unitary group. SU(2); it is also diffeomorphic to the unit 3 ...
  28. [28]
    Regular Polygon -- from Wolfram MathWorld
    A regular polygon is an n-sided polygon in which the sides are all the same length and are symmetrically placed about a common center.
  29. [29]
    Platonic Solid -- from Wolfram MathWorld
    Platonic solids are convex polyhedra with equivalent faces of congruent regular polygons. There are five: cube, dodecahedron, icosahedron, octahedron, and ...
  30. [30]
    1.2: Counting Symmetries - Mathematics LibreTexts
    Mar 5, 2022 · Of course, some objects have an infinite number of symmetries. A circle is a good example of this: every rotation is a symmetry, and there are ...
  31. [31]
    [PDF] Symmetry Groups of Platonic Solids - Brown Math
    Sep 17, 2007 · (1). The angles (α,β,γ) are essentially the Euler angles of the rotation. Geo- metrically, you get a general rotation by rotating various ...
  32. [32]
    Archimedean Solid -- from Wolfram MathWorld
    Archimedean solids are convex polyhedra with nonintersecting regular polygons of two or more types, arranged in the same way about each vertex, and have high  ...<|control11|><|separator|>
  33. [33]
    Genetically speaking, starfish have no arms—only a head | Science
    Nov 1, 2023 · Starfish, however, are echinoderms—animals with lines of symmetry radiating out from a center point, like spokes on a wheel. Echinoderms evolved ...
  34. [34]
    Rotational Symmetry - EscherMath
    Mar 10, 2009 · The Clematis shown has 8-fold rotational symmetry (45 degrees). It has 8 flower petals arranged around the center of the flower. The top left ...
  35. [35]
    What's special about the shape of a Nautilus shell? - EarthSky
    Apr 18, 2013 · Cutaway of a nautilus shell showing the chambers arranged in an approximately logarithmic spiral. ... It has been rotated 180 degrees.
  36. [36]
    Rotational Symmetry - Interactives and Examples | CK-12 Foundation
    Sep 23, 2025 · What makes mandalas special is their rotational symmetry. If you turn a mandala around its centre - by half a turn, a quarter turn, or even ...
  37. [37]
    Application-based principles of islamic geometric patterns - Nature
    Feb 1, 2023 · A rotation symmetry group detection technique for the characterization of Islamic Rosette Patterns. Pattern Recognit Lett. 2015;68(111):117 ...
  38. [38]
    Leonardo's Vitruvian Man Drawing: A New Interpretation Looking at ...
    Apr 9, 2015 · The relationship established between the circle's diameter and the side of the square is a consequence of the geometric relationship probably discovered by ...<|separator|>
  39. [39]
    The manoeuvrability hypothesis to explain the maintenance of ...
    Jul 12, 2012 · From the principles developed so far it follows that asymmetry or radial symmetry could have evolved only in animals which do not locomote or ...
  40. [40]
    Balance, symmetry, and emphasis - Smarthistory
    Balance is an even use of elements throughout a work of art. Symmetry is a very formal type of balance consisting of a mirroring of portions of an image.
  41. [41]
    Snowflake Science - SnowCrystals.com
    Why six? The six-fold symmetry you see in a snow crystal arises from the arrangement of water molecules in the ice crystal lattice. As this ice crystal ...
  42. [42]
    2.4: Crystallographic Point Groups - Chemistry LibreTexts
    Apr 7, 2024 · The five cyclic crystallographic point groups 1, 2, 3, 4, and 6 are both chiral and polar. There are five noncentrosymmetric point groups ...
  43. [43]
    [PDF] Handout 3: Crystallographic Plane & Space Groups
    Crystallographic symmetry involves long-range order, with 2, 3, 4, 6-fold rotations. There are 32 point groups, 230 space groups (3D) and 17 plane groups (2D).
  44. [44]
    The symmetry of crystals. The crystallographic restriction theorem
    Crystals can only have 2-fold, 3-fold, 4-fold, or 6-fold rotation axes. 5-fold, 7-fold, 8-fold, and higher-fold symmetries are not possible.
  45. [45]
    Crystals - Introduction - The Quartz Page
    Nov 30, 2016 · The top view shows that although the prism of the crystal is six-sided, it has a three-fold rotational symmetry. Quartz thus belongs to the ...
  46. [46]
    10 Crystal Morphology and Symmetry – Mineralogy - OpenGeology
    Squares have 4-fold rotational symmetry (rotation of 90°), and hexagons have 6-fold rotational symmetry (rotation of 60°). Box 10-2 contains more examples of ...
  47. [47]
    7.4: Rotational invariance and conservation of angular momentum
    Feb 27, 2021 · Angular momentum = ( r × p ) = constant. This proves the Noether' theorem that the angular momentum about any axis is conserved if the ...
  48. [48]
    17 Symmetry and Conservation Laws - Feynman Lectures - Caltech
    Symmetry with respect to rotations around the x-, y-, and z-axes implies the conservation of the x-, y-, and z-components of angular momentum. Symmetry with ...
  49. [49]
    [PDF] ANGULAR MOMENTUM - MIT OpenCourseWare
    Dec 16, 2013 · ... angular momentum operators commute with the Hamiltonian. Central potential Hamiltonians: [ L. ˆ i ,H ] = 0 . (1.70). We have seen that L. ˆ i ...
  50. [50]
    A Quantum Mechanical MP2 Study of the Electronic Effect of ... - MDPI
    Sep 29, 2024 · Although fullerene C60 has an icosahedral symmetry rather than a perfect spherical geometry, its electronic energy scheme can be derived by ...<|control11|><|separator|>
  51. [51]
    Quantum scattering of icosahedron fullerene C60 with noble-gas ...
    Apr 23, 2024 · The icosahedral anisotropy of the molecular system is reproduced by expanding the potential in terms of symmetry-allowed spherical harmonics.
  52. [52]
    [PDF] Fullerenes: Synthesis and Applications
    Jun 30, 2018 · Fullerene molecules are constituted entirely of carbon with 20 hexagonal and 12 pentagonal rings as their basis of an icosahedral symmetry in a ...
  53. [53]
    [PDF] dihedral groups - keith conrad
    The elements of Dn are rotations or reflections; there is no “mixed rotation-reflection”: the product of a rotation ri and a reflection rjs (in either order) is ...
  54. [54]
    [PDF] Classifications of Frieze Groups and an Introduction to ...
    May 10, 2019 · Now we will show by applying the composition rtr to a group with a rotational symmetry that this will be a translation.
  55. [55]
    [PDF] Script Unit 4.4 (Screw Axes in Crystal Structures).docx
    Along this axis the atoms are symmetry related to each other by applying a screw rotation: if we rotate first this atom by 60 degrees and then translate it ...
  56. [56]
    (International Tables) Volume A home page
    Volume A of the series, Space-group symmetry, contains diagrams and tables of data for the 17 plane groups, the 230 space groups and the 32 crystallographic ...
  57. [57]
    [PDF] Groups and their Representations Karen E. Smith
    Representation theory studies how abstract groups can be realized as groups of rigid transformations, often as linear transformations of Euclidean space.
  58. [58]
    [PDF] 11 Group Theory
    point in n-dimensional space is the group O(n) of rotations and reflec- tions. The rotations in n-space form the special orthogonal group SO(n). Linear ...
  59. [59]
    [PDF] Representations of the Rotation Groups SO(N)
    Dec 14, 2019 · This can be written as 4 → 3⊕1, which is to say that the 4-dimensional representation of SO(4) breaks up into 1- and 3-dimensional ...
  60. [60]
    Character table for point group C3v
    Character table for point group C3v. C3v, E, 2C3 (z), 3 v. linear functions, rotations. quadratic functions. cubic functions. A1, +1, +1, +1, z, x2+y2, z2, z3, ...
  61. [61]
    15.3: Permutation Groups - Mathematics LibreTexts
    Aug 16, 2021 · By now we've seen several instances where a group can appear through an isomorphic copy of itself in various settings. The simplest such example ...
  62. [62]
    [PDF] 2. harmonic analysis on compact groups. - MIT Mathematics
    These notes recall some general facts about Fourier analysis on a compact group. K. They will be applied eventually to compact Lie groups, particularly to ...
  63. [63]
    [PDF] The Evolution of Group Theory: A Brief Survey - Israel Kleiner
    Mar 14, 2004 · And Klein in his Lectures on the Icosahedron of 1884 "solved" the quintic equation by means of the symmetry group of the icosahedron.