Fact-checked by Grok 2 weeks ago

Isometry group

In , the isometry group of a X, denoted \operatorname{Isom}(X), is the set of all bijective maps f: X \to X that preserve distances, i.e., d(f(x), f(y)) = d(x, y) for all x, y \in X, forming a group under . These maps, called isometries, capture the symmetries of the space by maintaining its geometric structure without distortion. In Euclidean geometry, the isometry group of \mathbb{R}^n consists of all transformations of the form f(x) = Ax + b, where A is an (satisfying A A^T = I) and b \in \mathbb{R}^n, encompassing translations, rotations, reflections, and glide reflections. This group is a of the O(n) and the additive group \mathbb{R}^n, reflecting both rigid motions and orientation-reversing symmetries. For the \mathbb{H}^n, the isometry group is O^+(n,1), which includes orientation-preserving hyperbolic translations, rotations, and boosts, as well as orientation-reversing isometries like reflections. Isometry groups play a central role in studying geometric symmetries, classifying spaces up to congruence, and analyzing discrete subgroups for tilings and crystallographic patterns. In , they are groups when the space is smooth, enabling the investigation of homogeneous spaces and symmetric spaces as quotients by closed subgroups. Applications extend to physics, where they model invariances in , and to for transformations.

Definition and Preliminaries

Formal Definition

In mathematics, the isometry group of a metric space (X, d), denoted \operatorname{Isom}(X), is the set of all bijections f: X \to X such that d(f(x), f(y)) = d(x, y) for all x, y \in X. This collection consists precisely of the distance-preserving maps that are also surjective and injective, ensuring they are structure-preserving transformations of the space. The set \operatorname{Isom}(X) forms a group under the operation of function composition. Closure holds because if f, g \in \operatorname{Isom}(X), then for all x, y \in X, d((g \circ f)(x), (g \circ f)(y)) = d(g(f(x)), g(f(y))) = d(f(x), f(y)) = d(x, y), so g \circ f \in \operatorname{Isom}(X). The identity map \mathrm{id}_X, defined by \mathrm{id}_X(x) = x for all x \in X, serves as the neutral element, as it preserves distances trivially and is bijective. For inverses, if f \in \operatorname{Isom}(X), then f^{-1} is also bijective and an isometry, since for all a, b \in X, d(f^{-1}(a), f^{-1}(b)) = d(f(f^{-1}(a)), f(f^{-1}(b))) = d(a, b), using the bijectivity of f and the distance-preserving property of f. Associativity follows from the associativity of function composition. In metric spaces equipped with an orientation, such as Euclidean spaces, isometries are classified as direct (orientation-preserving) or opposite (orientation-reversing). The direct isometries form a normal subgroup of index 2 in \operatorname{Isom}(X), and the full isometry group is their semidirect product with \mathbb{Z}/2\mathbb{Z}, where the action of the generator (a reflection) conjugates direct isometries appropriately. For the trivial metric space X with at most one point, \operatorname{Isom}(X) consists solely of the identity map, as there are no nontrivial bijections or distances to preserve.

Isometries of Metric Spaces

An between (X, d_X) and (Y, d_Y) is a bijective f: X \to Y satisfying d_Y(f(x), f(y)) = d_X(x, y) for all x, y \in X. This condition ensures that f acts as an of the structures, preserving not only distances but also the induced . Specifically, open balls are mapped to open balls of the same radius: the image f(B_X(x, r)) equals B_Y(f(x), r) for any x \in X and r > 0, implying that f preserves and closedness of sets. Consequently, isometries are homeomorphisms, as their inverses also preserve distances and thus inherit the same topological properties. Every between spaces is continuous, a direct consequence of preservation, which implies that for any , choosing \delta = \epsilon suffices in the \epsilon-\delta definition of . Moreover, isometries are precisely the bijective maps with Lipschitz constant exactly 1: d_Y(f(x), f(y)) \leq d_X(x, y) holds with equality, making them non-expansive and embedding the faithfully. The converse—that every continuous with continuous inverse is an —does not hold in general spaces but can under completeness assumptions, such as when the spaces are complete and the is uniformly continuous in a manner compatible with the metrics. Isometries serve as the foundational elements of the isometry group, exhibiting behaviors analogous to rigid motions without relying on structure. Archetypal examples include , which shift all points by a fixed while preserving separations; rotations, which fix a central point and cyclically permute distances around it; and reflections, which fix a "mirror" while inverting distances across it—all defined intrinsically through preservation. Regarding fixed points, individual isometries do not generally possess them; for instance, a in a non-compact space like \mathbb{R}^n fixes no points. However, in complete metric spaces satisfying conditions like the existence of a conical bicombing, certain isometries admit fixed points or invariant functionals, particularly when acting on bounded orbits or in spaces such as Banach spaces or CAT(0) spaces. These properties highlight isometries' role in maintaining geometric integrity across diverse metric environments.

Group-Theoretic Properties

Group Structure and Operations

The isometry group \operatorname{Iso}(X) of a metric space (X, d) forms a group under the operation of function composition. Closure holds because the composition of two isometries f, g \in \operatorname{Iso}(X) satisfies d(g \circ f(x), g \circ f(y)) = d(f(x), f(y)) = d(x, y) for all x, y \in X, preserving distances. Associativity follows directly from the associativity of function composition on the set of all maps from X to itself. The identity element is the identity map \operatorname{id}_X, which trivially preserves distances. For inverses, since every isometry is bijective, the inverse map f^{-1} satisfies d(f^{-1}(u), f^{-1}(v)) = d(f(f^{-1}(u)), f(f^{-1}(v))) = d(u, v) for all u, v \in X, confirming that f^{-1} \in \operatorname{Iso}(X). When X is a topological metric space, \operatorname{Iso}(X) can be endowed with a natural structure. Specifically, the topology of on compact subsets—where a sequence of isometries \{f_n\} converges to f if \sup_{x \in K} d(f_n(x), f(x)) \to 0 for every compact K \subset X—makes \operatorname{Iso}(X) a . In this , the group operations of and inversion are continuous: convergence in this sense preserves uniform limits on compacts, ensuring that limits of compositions and inverses remain isometries. For proper spaces, this often renders \operatorname{Iso}(X) locally compact and second countable. Homomorphisms between isometry groups arise naturally from isometric embeddings of metric spaces. Given an isometric embedding \phi: X \to Y between metric spaces (X, d_X) and (Y, d_Y), it induces a group homomorphism \psi: \operatorname{Iso}(X) \to \operatorname{Iso}(Y) whose image consists of isometries preserving \phi(X) setwise, defined by \psi(f) = \phi \circ f \circ \phi^{-1} for f \in \operatorname{Iso}(X), where \phi^{-1} is understood on the image \phi(X). This map preserves the group operation since \psi(f_1 \circ f_2) = \phi \circ (f_1 \circ f_2) \circ \phi^{-1} = (\phi \circ f_1 \circ \phi^{-1}) \circ (\phi \circ f_2 \circ \phi^{-1}) = \psi(f_1) \circ \psi(f_2). If \phi is an isometry (bijective), then \psi is an isomorphism. The center of \operatorname{Iso}(X), denoted Z(\operatorname{Iso}(X)), comprises all isometries that commute with every element of the group, i.e., Z(\operatorname{Iso}(X)) = \{g \in \operatorname{Iso}(X) \mid g \circ h = h \circ g \ \forall h \in \operatorname{Iso}(X)\}. This is always a , as the center of any group is under conjugation. In non-abelian cases, such as the isometry group of \mathbb{R}^n for n \geq 2, the center is often trivial, consisting only of the , due to the semi-direct product structure \operatorname{Iso}(\mathbb{R}^n) \cong [O(n](/page/O(n)) \ltimes \mathbb{R}^n where orthogonal transformations do not generally commute with translations. subgroups of \operatorname{Iso}(X) are subgroups invariant under conjugation by all elements; examples include the subgroup of translations in isometry groups, which is . A key operation in \operatorname{Iso}(X) is conjugation, defined by g \circ f \circ g^{-1} for g, f \in \operatorname{Iso}(X). This relabels the geometric features of f: the fixed-point set of the conjugate is the image under g of the fixed-point set of f, i.e., \{x \in X \mid (g \circ f \circ g^{-1})(x) = x\} = g(\{y \in X \mid f(y) = y\}). Similarly, orbits under the action of f are mapped to orbits under the conjugate, preserving their structure via g. Conjugation thus acts as a "relabeling" that maintains the intrinsic while shifting positions in X.

Invariants and Orbits

The \operatorname{Iso}(X) of a (X, d) acts on X by evaluation, where each f \in \operatorname{Iso}(X) maps x \in X to f(x). This is faithful, meaning the of the \operatorname{Iso}(X) \to \operatorname{Sym}(X) is trivial, as distinct isometries differ at some point. The orbits under this partition X into , where two points x, y \in X lie in the same if there exists f \in \operatorname{Iso}(X) such that f(x) = y, i.e., if x and y are congruent via an . For a fixed x \in X, the stabilizer \operatorname{Stab}(x) = \{f \in \operatorname{Iso}(X) \mid f(x) = x\} is the of isometries fixing x. The orbit-stabilizer theorem applies here: the of the \operatorname{Orb}(x) equals the of \operatorname{Stab}(x) in \operatorname{Iso}(X), or |\operatorname{Orb}(x)| = [\operatorname{Iso}(X) : \operatorname{Stab}(x)] = |\operatorname{Iso}(X)| / |\operatorname{Stab}(x)| when the group is finite. This relation quantifies how the size of the symmetry group at a point determines the extent of the orbit. Isometries preserve key geometric invariants of the space. By definition, distances are invariant: d(f(x), f(y)) = d(x, y) for all f \in \operatorname{Iso}(X) and x, y \in X. In inner product spaces, such as spaces, isometries also preserve angles, as the inner product satisfies \langle f(u), f(v) \rangle = \langle u, v \rangle for vectors u, v. More generally, in Riemannian manifolds, isometries preserve curvature tensors, including , ensuring that local geometric properties like remain unchanged under the . Fundamental domains provide a way to classify and represent the of \operatorname{Iso}(X). A fundamental domain F \subset X for the action is a that intersects each in exactly one point (or minimally, up to identifications), allowing the quotient space X / \operatorname{Iso}(X) to be modeled via F. For subgroups of isometries, such as Fuchsian groups acting on spaces, Dirichlet fundamental domains—defined as intersections of half-spaces \{z \in X \mid d(z, z_0) \leq d(\gamma z, z_0)\} for \gamma \in \Gamma and center z_0—are convex and locally finite, facilitating the study of orbit structures and group generation from pairings. As an example, consider the \mathbb{Z}^n equipped with the Euclidean induced from \mathbb{R}^n. The S_n, acting by coordinates, consists of isometries since the is invariant under permutations. The under this action correspond to congruence classes of vectors modulo , i.e., points with the same of coordinates, such as all permutations of (1, 2, 2) forming one . Applying the , for a vector like (1, 2, 3) with distinct entries, the is trivial, so |\operatorname{Orb}((1,2,3))| = n!.

Classifications and Decompositions

Polar Decomposition

In Hilbert spaces, the polar decomposition theorem provides a factorization for bounded linear operators that highlights their isometric and positive components. Specifically, for a bounded linear operator T on a Hilbert space \mathcal{H}, there exists a unique decomposition T = U |T|, where U is a partial isometry and |T| = \sqrt{T^* T} is the positive self-adjoint operator defined by the absolute value of T. For an isometry f: \mathcal{H} \to \mathcal{H}, satisfying \|f(x)\| = \|x\| for all x \in \mathcal{H} (equivalently, f^* f = I), the positive part simplifies to |f| = I, yielding f = u where u is a partial isometry with initial projection I; if f is surjective, u is unitary (orthogonal). This structure underscores that linear isometries on Hilbert spaces are essentially partial unitaries, preserving the inner product structure up to domain considerations. For general isometries in normed spaces, particularly affine isometries, the polar decomposition adapts to separate the orthogonal (norm- and angle-preserving) component from a translational shift, reflecting the semidirect product structure of the isometry group. In Euclidean space \mathbb{R}^n equipped with the standard metric, any isometry f: \mathbb{R}^n \to \mathbb{R}^n admits the form f(x) = Q(x - a) + b, where Q \in O(n) is an orthogonal matrix (satisfying Q^T Q = I and preserving norms and angles), and a, b \in \mathbb{R}^n are fixed vectors accounting for the translational aspects; this is equivalent to a composition of a translation, an orthogonal transformation, and another translation. This decomposition arises because fixing a point (e.g., the origin after translation) reduces the isometry to a linear orthogonal map, with the full group forming the semidirect product O(n) \ltimes \mathbb{R}^n. A proof sketch for the linear case relies on the (SVD), which coincides with the in finite dimensions: for a linear L on \mathbb{R}^n, the SVD L = U \Sigma V^T has all singular values in \Sigma equal to 1, implying L = U V^T is orthogonal (unitary in the real case). Extending to affine isometries involves conjugating by a to linearize around a fixed point, applying the linear decomposition, and translating back, ensuring the overall map preserves distances. In infinite-dimensional Hilbert spaces, the continuous analog uses the on |L|, but for isometries, it again yields the identity positive operator. This decomposition applies primarily to linear or affine isometries in normed or inner product spaces, where the metric derives from a norm; in general spaces, such as non-complete or non-Euclidean ones, not all isometries admit an orthogonal-translation split, as the absence of a structure prevents affine representations. The roots of these ideas trace to 19th-century work on rigid motions, notably Chasles' 1830 theorem decomposing general displacements into rotations and translations (or screws in ).

Cartan-Dieudonné Theorem

The Cartan–Dieudonné theorem states that every in the O(n) over the real numbers, acting on an n-dimensional , can be expressed as the composition of at most n reflections in hyperplanes. Specifically, proper (those with +1, i.e., rotations in SO(n)) are products of an even number of such reflections, while improper ones ( -1) require an odd number. This result characterizes the as being generated by reflections, providing a fundamental decomposition for isometries preserving the metric. Élie Cartan formalized an early version of the theorem in the context of during the early 20th century, with a detailed treatment appearing in his work on spinors and quadratic forms. later generalized it to nondegenerate symmetric bilinear forms over arbitrary fields. A \sigma_v across the orthogonal to a nonzero vector v is given by the formula \sigma_v(x) = x - 2 \proj_v(x) = x - 2 \frac{x \cdot v}{\|v\|^2} v, where \proj_v(x) denotes the orthogonal of x onto the line spanned by v. Compositions of an even number of such reflections yield elements of SO(n), corresponding to rotations, while odd compositions produce improper isometries. These reflections are represented by Householder matrices, which are symmetric and orthogonal with -1. The proof proceeds by on the dimension n. For the base case n=1, the group O(1) consists of the and a single (multiplication by -1). Assuming the result for dimensions less than n \geq 2, consider an f \in O(n). If f fixes a (i.e., has eigenvalue +1 with multiplicity at least n-1), it restricts to an orthogonal transformation on that hyperplane, which by is a product of at most n-1 reflections. Otherwise, there exists a s such that s \circ f fixes a hyperplane, reducing the problem to the previous case and yielding at most n reflections overall. reflections are used constructively to align vectors and create fixed hyperplanes iteratively. The theorem extends to pseudo-orthogonal groups O(p,q) preserving indefinite quadratic forms on pseudo-Euclidean spaces of signature (p,q), where every element is a product of reflections across hyperplanes orthogonal to non-isotropic vectors, with the number bounded by the dimension p+q. This generalization applies to spaces like Minkowski spacetime, maintaining the generation by reflections but requiring care with null directions.

Examples in Specific Spaces

Euclidean Isometry Groups

The isometry group of n-dimensional space \mathbb{R}^n, denoted E(n), consists of all distance-preserving transformations of the space. This group is structured as a E(n) = T(n) \rtimes O(n), where T(n) \cong \mathbb{R}^n is the normal subgroup of and O(n) is the comprising rotations and reflections. The arises because translations and orthogonal transformations do not commute in general; specifically, conjugating a translation by an orthogonal transformation rotates the translation . Elements of E(n) take the general affine form f(\mathbf{x}) = Q \mathbf{x} + \mathbf{b}, where Q \in O(n) and \mathbf{b} \in \mathbb{R}^n, which preserves the \|\mathbf{f(x)} - \mathbf{f(y)}\| = \|Q(\mathbf{x} - \mathbf{y})\| = \|\mathbf{x} - \mathbf{y}\| since Q is orthogonal. The orientation-preserving subgroup of E(n), known as the special Euclidean group SE(n) or E^+(n), is the semidirect product SE(n) = T(n) \rtimes SO(n), where SO(n) is the special orthogonal group of proper rotations (determinant 1). This subgroup excludes reflections and improper rotations, focusing on rigid motions that maintain . The linear part Q \in SO(n) in the affine representation of elements in SE(n) incorporates a proper , though the full also includes the \mathbf{b}. A fundamental classification of elements in SE(3), the case for , is given by Chasles' theorem, which states that every orientation-preserving is a screw displacement: a about an combined with a parallel to that . This helical motion generalizes pure (zero translation) and pure (zero ), providing a unified geometric description for motions in . The theorem highlights the one-parameter family of possible screw axes and pitches characterizing such transformations. Finite subgroups of E(n) are discrete and compact, arising primarily from symmetries of polytopes or lattices, and are conjugate to finite subgroups of O(n) with trivial translation components. In two dimensions, these include the cyclic groups C_k ( by multiples of $2\pi/k) and groups D_k ( and reflections of k-gons). In three dimensions, prominent examples are the rotation groups of the Platonic solids: the tetrahedral group A_4 (order 12, symmetries of the ), octahedral group S_4 (order 24, or ), and icosahedral group A_5 (order 60, or ), all finite subgroups of SO(3). These groups generate the full symmetry including reflections when embedded in O(3), illustrating the discrete rotational structure within the broader framework.

Isometries of the Hyperbolic Plane

The isometry group of the hyperbolic plane, denoted Isom(ℍ²), consists of all transformations that preserve the hyperbolic metric. The orientation-preserving subgroup, Isom⁺(ℍ²), is isomorphic to the projective special linear group PSL(2, ℝ), which acts on the upper half-plane model ℍ² = {z ∈ ℂ | Im(z) > 0} via Möbius transformations of the form z ↦ (az + b)/(cz + d), where a, b, c, d ∈ ℝ and ad - bc = 1. These transformations preserve the hyperbolic metric ds² = (dx² + dy²)/y² in the upper half-plane model, ensuring distances and angles are maintained. Isometries in the hyperbolic plane are classified based on their fixed points and action. Elliptic isometries fix a unique point in ℍ² and rotate around it, analogous to rotations but in a space of constant negative curvature. Parabolic isometries, also called horocyclic, fix exactly one point on the boundary at infinity and translate along horocycles. Hyperbolic isometries fix no points in ℍ² but two points on the boundary, acting as translations along the unique geodesic connecting those points; unlike Euclidean translations, these have no fixed points in the interior and generate infinite-order elements. In the Poincaré disk model, where ℍ² is represented as the open unit disk { (x, y) ∈ ℝ² | x² + y² < 1 } with ds² = 4(dx² + dy²)/(1 - x² - y²)², isometries are fractional linear transformations that preserve this and the disk . These transformations map (circular arcs orthogonal to the ) to and maintain the conformal . Discrete subgroups of Isom⁺(ℍ²), known as Fuchsian groups, admit fundamental domains such as ideal triangles (with vertices on the ) or geodesic strips, which tile the hyperbolic plane under the without overlap except on boundaries. This contrasts with isometry groups, where translations fix no interior points but hyperbolic isometries introduce elements of infinite order without interior fixed points, reflecting the unbounded nature of .

Applications and Extensions

Crystallography and Symmetry Groups

In crystallography, isometry groups provide the foundational framework for describing the symmetries of crystal structures, particularly through their discrete subgroups known as space groups. These space groups are the 230 distinct discrete subgroups of the isometry group E(3) that act properly discontinuously and cocompactly on \mathbb{R}^3, incorporating combinations of , , , and more complex operations such as screws (rotation combined with along the axis) and glides ( combined with parallel to the reflection plane). Such groups ensure that the periodic arrangement of atoms in a crystal lattice remains invariant under these transformations, enabling the classification of all possible crystal symmetries in three dimensions. The development of space group theory emerged in the late , with independent enumerations by Arthur Schönflies and Evgraf Stepanovich Federov, who both identified the full set of 230 groups by 1891 through systematic analysis of possible symmetry operations compatible with translational periodicity. Schönflies approached the problem via extensions, while Federov emphasized geometric constructions, and their correspondence helped resolve minor discrepancies in early lists. This enumeration laid the groundwork for modern structural , allowing scientists to map observed patterns to specific atomic arrangements. A key constraint in this framework is the , which limits rotational symmetries in two-dimensional to orders 1, 2, 3, 4, or 6 due to the requirement that must map the discrete points onto themselves without gaps or overlaps. For instance, a 5-fold would rotate vectors by angles incompatible with the linear combinations defining the , leading to non-periodic structures; thus, only these orders preserve the 's translational invariance. This theorem extends implications to groups, restricting symmetries to the 32 crystallographic classes. Bieberbach's theorems further characterize these discrete isometry subgroups, stating that every discrete cocompact subgroup \Gamma of E(n) possesses a normal subgroup of translations of finite index, with the quotient \Gamma / T isomorphic to a finite subgroup of O(n), ensuring the fundamental domain is a compact parallelohedron tiling \mathbb{R}^n. The first theorem identifies the translation lattice as the kernel of the action, while the second guarantees uniqueness up to for a given group. These results characterize crystallographic groups as the cocompact subgroups of E(n), each containing a finite-index of translations. Central to space group actions is the preservation of the crystal lattice \Lambda \subset \mathbb{R}^3 under isometries, formalized by the condition f(\Lambda) = \Lambda for any symmetry operation f in the group, meaning f maps lattice points to lattice points while maintaining distances and orientations. This invariance ensures that the lattice serves as the translational skeleton, upon which point group operations act to generate the full space group, as seen in examples like the primitive cubic lattice preserved by the Pm\overline{3}m group.

Generalizations to Other Structures

In Riemannian geometry, isometries are diffeomorphisms between Riemannian manifolds (M, g) and (N, h) that preserve the , satisfying \phi^* h = g, where \phi^* h denotes the metric. This condition ensures that the inner product on tangent spaces is preserved: for all p \in M and X, Y \in T_p M, g_p(X, Y) = h_{\phi(p)}(d\phi_p(X), d\phi_p(Y)). The group of all such isometries, denoted \mathrm{Isom}(M, g), forms a under composition, acting smoothly on M by the Myers–Steenrod theorem, provided M has finitely many connected components. The concept of isometry groups extends to discrete structures like , where the isometry group consists of automorphisms that preserve the , defined as the length of the shortest between . A is a on the that preserves adjacency, and since are determined by adjacency relations, such maps automatically preserve distances. For example, the of a with n is the of order 2, consisting of the identity and the reversal map, both of which maintain between . Conformal isometries generalize isometries by preserving angles rather than distances, leading to structures like the Möbius group on . The Möbius group \mathrm{Möb}(S^n) is the full \mathrm{Conf}(S^n) of the n-, comprising projective transformations that preserve the and act as hypersphere-preserving maps. These transformations are conformal, maintaining local angles and shapes, and include orientation-preserving elements isomorphic to \mathrm{SO}^+(1, n+1). On of constant , subgroups of the Möbius group fixing specific sphere complexes yield true isometries of the underlying . Modern extensions of isometry groups appear in non-Riemannian settings like Finsler geometry and Alexandrov spaces, where smoothness is replaced by bounds. In Finsler geometry, an between manifolds (M, F) and (N, \bar{F}) is a \phi: M \to N such that \bar{F} \circ d\phi = F, preserving the Finsler norm on tangent vectors. This generalizes the Riemannian case by allowing asymmetric metrics. In Alexandrov spaces, which are metric spaces with bounded below or above in a synthetic sense, isometries are bijective distance-preserving maps that maintain and convexity properties, with the isometry group \mathrm{Iso}(X) exhibiting bounded dimension relative to the space's structure. For instance, if the dimension of \mathrm{Iso}(X) achieves the maximum possible for an n-dimensional Alexandrov space, then X is isometric to a . These developments, building on foundational work in metric , enable the study of singular spaces without requiring differentiability.

References

  1. [1]
    [PDF] Length Spaces - ETH Zurich
    Let G be a subgroup of the isometry group of a metric space X. The action of G on X is said to be proper if for each x ∈ X there exists an > 0 such that ...
  2. [2]
    [PDF] Part IB - Geometry (Definitions) - Dexter Chua
    Definition (Isometry group). The isometry group Isom(Rn) is the group of all isometries of Rn, which is a group by composition. Definition (Special ...<|control11|><|separator|>
  3. [3]
    [PDF] Hyperbolic Geometry - OSU Math
    Every linear transformation in O+(n, 1) is an isometry of Hn. In fact, these are the only isometries of Hn. Theorem 5.1. The isometry group of Hn is O+(n, 1) ...
  4. [4]
    [PDF] Euclidean isometries and surfaces - UChicago Math
    Definition 2.2. A isometry group Γ of the euclidean plane is called discontinuous if no P ∈ R2 has a Γ-orbit with a limit point. By limit ...
  5. [5]
    [PDF] THE ISOMETRY GROUP OF SEMI-RIEMANNIAN MANIFOLDS
    The main goal of this work is to prove that the isometry group of a semi-Riemannian manifold is a Lie group, and to do so in a self-contained way, assuming no ...
  6. [6]
    6.4 Examples of isometry groups and homogeneous spaces - Fiveable
    Euclidean group E(n) encompasses all isometries of n-dimensional Euclidean space · Euclidean group E(n) has important applications in physics and robotics.
  7. [7]
    [PDF] NOTES FOR MATH 5510, FALL 2017, V 1 1. Metric Spaces 2 1.1 ...
    Dec 3, 2017 · This means that the set of all isome- tries is a group under composition. Definition 2.1. Let Isom(X, d) = {f : X → X : f is an isometry of (X, ...<|control11|><|separator|>
  8. [8]
    [PDF] Isometries of the plane and linear algebra - Keith Conrad
    inverse of an isometry being an isometry, isometries form a group under composition. We will describe the elements of this group and show how the group law ...
  9. [9]
    [PDF] Chapter 7. Isometries and Symmetry Groups
    7.2 Theorem: The set of isometries on Rn is a group under composition. Proof: The identity map I : Rn → Rn is an isometry because I(x) − I(y) = kx − yk.
  10. [10]
    [PDF] Direct and Opposite Isometries - User Web Pages
    Direct isometries maintain orientation, while opposite isometries change it. Translations and rotations are direct; reflections and glide reflections are  ...
  11. [11]
    [PDF] arXiv:1801.09962v1 [math.AG] 30 Jan 2018
    Jan 30, 2018 · isometries, denoted by Iso2 is isomorphic to the semidirect product O(R, 2) o. R. 2. The isometries preserving also the orientation of the ...
  12. [12]
    [PDF] Continuity and Homeomorphisms - ScholarWorks@GVSU
    A function f from a metric space (X, dX) to a metric space (Y,dY ) is an isometry if f is a bijection and. dY (f(a),f(b)) = dX(a, b). (14.1) for all a, b ∈ X.
  13. [13]
    [PDF] Metric topology III: Introduction to functions and continuity
    Any isometry is continuous; for any metric space X and any fixed p0 ∈ X, the function X → R, p 7→ d(p, p0) is continuous. In both these examples one may ...
  14. [14]
    [PDF] Review of metric spaces
    Aug 30, 2005 · Certainly an isometry is continuous. The usual definition of the completion Y of a metric space X is that Y is a complete metric space with an.
  15. [15]
    [PDF] KEITH CONRAD - 1. Introduction An isometry of Rn is a function h ...
    It is clear that the three kinds of isometries pictured above (translations, rotations, reflections) are each invertible (translate by the negative vector, ...
  16. [16]
    [2207.00963] A metric fixed point theorem and some of its applications
    Jul 3, 2022 · A general fixed point theorem for isometries in terms of metric functionals is proved under the assumption of the existence of a conical bicombing.
  17. [17]
    [PDF] Geometry & Groups 2 Metric spaces
    Jan 8, 2012 · If an isometry from (X,d) to (X0,d0) exists, then the spaces are called isometric. The inverse of an isometry is clearly an isometry, and so are ...
  18. [18]
    [PDF] A Course in Metric Geometry Dmitri Burago Yuri Burago Sergei Ivanov
    May 2, 2018 · This book covers metric spaces, length spaces, and introduces Riemannian and hyperbolic geometries, aiming to give a detailed exposition of ...
  19. [19]
    [PDF] isometry groups of non-positively curved spaces: structure theory
    The full isometry group of such a space is a Lie group. Polyhedral complexes of piecewise constant non-positive curvature, such as trees or Euclidean buildings.
  20. [20]
    algebraic and topological properties of the group of isometries on ...
    Under the standard composition of operators and the strong operator topology, the isometry group G(E) is a topological group.
  21. [21]
    [2502.16712] Group homomorphisms induced by isometries - arXiv
    Feb 23, 2025 · We investigate the continuous group homomorphisms induced by isometries of AP(G) into AP(H). Among others, the following results are proved.
  22. [22]
    [PDF] isometry groups of homogeneous spaces with positive sectional ...
    Definition. Let (M,g) be a Riemannian manifold with isometry group ˆG and isotropy group ˆHp for p ∈ M. If the isotropy representation of ...
  23. [23]
    [PDF] The geometry of conjugation in Euclidean isometry groups - arXiv
    Abstract. We describe the geometry of conjugation within any split subgroup H of the full isometry group G of n-dimensional Euclidean space.<|control11|><|separator|>
  24. [24]
    [PDF] Extending group actions on metric spaces - Carolyn R. Abbott
    Oct 1, 2018 · Combining this with (13), we obtain the required inequality. Page 15. Extending group actions on metric spaces 639. Proposition 3.7. For any ...
  25. [25]
    [PDF] Chapter 3: Transformations Groups, Orbits, And Spaces Of Orbits
    That is, this subspace is a fundamental domain for the action of the symmetric group on Rn. b.) Compute the isotropy groups which arise. By our basic principle ...
  26. [26]
    [PDF] Riemannian Geometry
    Note that local isometries preserve curvatures, in the sense that they pre- serve Levi-Civita connections. Sectional curvature is preserved as a number.
  27. [27]
    [PDF] Riemannian Manifolds: An Introduction to Curvature
    with Euclidean plane geometry, which you studied in school. Its elements are points, lines, distances, angles, and areas. Here are a couple of typical.
  28. [28]
    Fundamental domains - SpringerLink
    Jun 29, 2021 · In this chapter, we pursue a general construction of nice fundamental domains for the action of a discrete group of isometries.
  29. [29]
    RELATIVE SELF-ADJOINT OPERATORS IN HILBERT SPACE
    The polar decomposition theorem for A implies the existence of a unique elementary operator R relative to which A is self-adjoint and having the further ...
  30. [30]
    [PDF] ME/SE 740 Lecture 13 Chasle's Theorem
    Chasle's Theorem. Rigid body motion in the plane. Consider a rigid body motion from frame B1 to frame E. If this motion is not a pure translation, there is a ...
  31. [31]
    [PDF] The Cartan–Dieudonné Theorem - UPenn CIS
    The following fact is the key to the proof that every isometry can be decomposed as a product of reflections. Lemma 7.1.3 Let E be any nontrivial Euclidean ...
  32. [32]
    [PDF] Double covers of pseudo-orthogonal groups - prof. Andrzej Trautman
    The group On0,n00 is known to be generated by reflections in hyperplanes orthogonal to non-isotropic vectors (Cartan-Dieudonné). Every reflection is an ...
  33. [33]
  34. [34]
    [PDF] Adjoint and Coadjoint Orbits of the Special Euclidean Group
    May 3, 2015 · We give a geometric description of the adjoint and coadjoint orbits of the special Euclidean group. We implement the method of little ...
  35. [35]
    [PDF] Affine Transformations - UCSD CSE
    An affine transformation is a function R" →R" with the form. X1 input ... • Affine transformations take the form of x→ Ax+b. +b b3×1. X3 × 1. A3×3. X3 ...<|control11|><|separator|>
  36. [36]
    [PDF] Relativistic Chasles' theorem and the conjugacy classes of ... - arXiv
    Feb 1, 2013 · Mathematically, Euler's and Chasles' theorems establish the existence of a certain type of representative for each conjugacy class of the group ...
  37. [37]
    [PDF] on euler's rotation theorem
    (Mozzi-Chasles) If m is an orientation preserving isometry then m is a screw displacement, that is, a rotation about a line l followed (or preceded) by a ( ...
  38. [38]
    [PDF] The Platonic solids and finite rotation groups
    They are called the cyclic subgroup and the p-dihedral subgroup. Definition ... We also took a quick look at the two other finite subgroups of the rotation group ...
  39. [39]
    [PDF] classifying the finite subgroups of so3 - The University of Chicago
    Aug 29, 2020 · A finite subgroup of SO3 is either a cyclic group, a dihedral group, the tetrahedral group, the octahedral group or the icosahedral group. The ...Missing: Euclidean | Show results with:Euclidean
  40. [40]
  41. [41]
    [PDF] Hyperbolic space and its isometries
    The two models are equivalent under any Möbius transformation that maps the upper half-plane onto the unit disk. We will denote either one of these models by H2 ...
  42. [42]
    [PDF] ISOMETRIES OF THE HYPERBOLIC PLANE - UChicago Math
    In this paper, I will explore basic properties of the group PSL(2, R). These include the relationship between isometries of H. 2, Möbius transforma- tions, and ...
  43. [43]
    [PDF] Elements of Hyperbolic geometry
    We are able to characterize the full isometry group of Isom(H2). Theorem 1.14. The isometry group Isom(H2) is generated by PSL(2, R) and the reflexion z ...
  44. [44]
    [PDF] Chapter 9 - Poincaré's Disk Model for Hyperbolic Geometry
    For Poincaré's Disk Model we take the set of points that lie inside the unit circle, i.e., the set. Ж 2 = {(x, y) | x2 + y2 < 1}. Note that points on the ...
  45. [45]
    [PDF] Hyperbolic geometry MA 448 - University of Warwick
    Mar 1, 2013 · 3 Isometries of D and H. 41. 3.1 Classification of elements of Aut(ˆC). . . . . . . . . . . . . . . 41. 3.1.1 Classification by fixed points ...<|control11|><|separator|>
  46. [46]
    [PDF] Math 6640 – Hyperbolic Geometry Course Notes, Fall 2023
    Dec 7, 2023 · The full group of isometries is obtained by adding one additional generator: the map z 7→ −¯z. To see this is the full group of isometries, one ...
  47. [47]
    Spin Space Groups: Full Classification and Applications | Phys. Rev. X
    Aug 28, 2024 · For more than one hundred years, the 230 space groups have been the standard mathematical description of symmetries in solid material ...
  48. [48]
    Fundamental concepts for the study of crystal symmetry
    There are 14 of them among the 230 space groups of dimension three; every other space group is one of their subgroups. The list of the 230 space groups has ...
  49. [49]
    Arthur Schönflies (1853 - 1928) - Biography - MacTutor
    By 1891 Schönflies had found the complete list of 230 such groups. His ... space groups was made independently by E S Fedorov. Schönflies corresponded ...Missing: Federov | Show results with:Federov
  50. [50]
    1.1. Historical introduction - Wiley Online Library
    The presentations by Fedorov and Schoenflies of the 230 space groups were not yet appropriate for use in structure determina- tions with X-rays. The ...Missing: Schönflies Federov
  51. [51]
    [PDF] Crystal Math - Physics Courses
    Euclidean group: S ⊂ E(3), and a general space group operation { g \\ t. } ... Of the 230 three-dimensional space groups, 157 are nonsymmorphic and contain ...
  52. [52]
    [PDF] Crystallographic Point Groups and Space Groups
    Jun 8, 2011 · The rotational symmetries of a discrete lattice are limited to 2-, 3-,. 4-, and 6-fold. Proof. Suppose R is a rotational symmetry of the lattice ...
  53. [53]
    [PDF] Periodicity, Quasiperiodicity, and Bieberbach's Theorem on ... - People
    A czystallographic group is a discrete, cocompact group of isometries of n- dimensional Euclidean space. All terms in this definition are explained in ...
  54. [54]
    [PDF] Connective Bieberbach Groups - Purdue Math
    A crystallographic group of dimension k ≥ 1 is a discrete co-compact subgroup of the isometry group Iso(Rk) = Rk o O(k) of the Euclidean space R k. In his ...
  55. [55]
    Mathematics of 2-Dimensional Lattices
    Dec 7, 2022 · Any orientation-preserving isometry f is a composition of translations and ... in terms of distances, if f maps a lattice Λ to Λ. , then f ...
  56. [56]
    [PDF] Mathematics of 2-dimensional lattices - Vitaliy Kurlin
    For example, the area of the unit cell U spanned by any basis of a lattice Λ is an isometry invariant because a change of basis is realised by a 2 × 2 matrix ...
  57. [57]
    None
    Below is a merged response summarizing the definition of Riemannian isometry and isometry group across all provided segments. To retain all information in a dense and organized manner, I will use a table in CSV format for key details (definition, metric tensor condition, isometry group as Lie group, page references, and useful URLs), followed by a narrative summary that consolidates additional context and nuances. This approach ensures all details are preserved while maintaining clarity.
  58. [58]
    [PDF] Automorphisms of graphs - vlsicad page
    An automorphism of a graph G is a permutation g of the vertex set of G with the property that, for any vertices u and v, we have ug ∼ vg if and only.
  59. [59]
    [PDF] The Möbius group - and Geometry
    Möbius geometry is the geometry of the group of Möbius transfor- mations, that is, hypersphere preserving (point) transformations, acting on the n-sphere Sn as ...
  60. [60]
    [PDF] Isometries, submetries and distance coordinates on Finsler manifolds
    Jun 20, 2014 · The mapping ϕ: M → N is called a local Finslerian isometry if every point has a neighbourhood on which ϕ is a Finslerian isometry.
  61. [61]
    [1109.4878] Isometry groups of Alexandrov spaces - arXiv
    Sep 22, 2011 · For an Alexandrov space (with curvature bounded below), we determine the maximal dimension of its isometry group and show that the space is ...