Fact-checked by Grok 2 weeks ago

Isometry

An isometry is a bijective mapping between metric spaces that preserves distances, meaning for any points x and y in the domain, the distance d(f(x), f(y)) equals d(x, y), where f is the isometry and d denotes the metric. In the context of Euclidean geometry, isometries are rigid transformations that maintain the shape and size of figures, including translations, rotations, reflections, glide reflections, and the identity transformation. These transformations preserve not only distances but also angles, lengths of segments, and congruence of triangles, ensuring that the intrinsic geometric properties of objects remain unchanged. Every isometry of the Euclidean plane can be expressed as the composition of at most three reflections, providing a fundamental classification into the five types mentioned. Isometries play a central role in understanding symmetry groups and the structure of geometric spaces, with applications extending to crystallography, computer graphics, and differential geometry.

Fundamentals

Introduction

An isometry is a mapping between spaces that preserves the distances between points, thereby maintaining the intrinsic geometric relations without distortion or scaling. This core concept underpins much of modern geometry and analysis, allowing for the study of shapes and structures in a way that emphasizes their essential properties rather than their positions. The idea of isometries traces its origins to Euclidean geometry, where Euclid, in his Elements around 300 BCE, introduced the notion of congruence through figures that could be superimposed via motion, implicitly relying on distance-preserving transformations to establish equality of size and shape. This intuitive understanding evolved significantly in the 19th century with the advent of non-Euclidean geometries; mathematicians like Nikolai Lobachevsky and János Bolyai independently developed hyperbolic geometry, extending the role of isometries to curved spaces where parallel lines diverge, yet distances remained rigorously preserved under transformations. The formal abstraction of isometries within general metric spaces was pioneered by Maurice Fréchet in his 1906 doctoral thesis, which laid the groundwork for treating arbitrary sets equipped with distance functions as unified mathematical objects. Isometries play a vital role in preserving geometric structure across diverse fields, enabling the analysis of symmetries and invariances that are fundamental to natural phenomena. In physics, they model motions, such as translations and rotations, which describe the of undeformable objects without altering internal distances. In , groups of isometries classify the periodic symmetries of crystal lattices, providing a for understanding atomic arrangements and material properties. Similarly, in , isometries facilitate transformations like rotations and reflections that maintain object integrity during rendering and . Global isometries are bijective by definition, ensuring a complete, reversible correspondence between spaces.

Definition in Metric Spaces

In metric spaces, an isometry is formally defined as a function f: (X, d_X) \to (Y, d_Y) between two spaces that preserves distances, meaning d_X(a, b) = d_Y(f(a), f(b)) for all points a, b \in X. This preservation ensures that the intrinsic geometry of distances in the domain space X is mirrored exactly in the image under f within the space Y. Isometries are always injective mappings. To see this, suppose f(a) = f(b) for some a, b \in X; then d_X(a, b) = d_Y(f(a), f(b)) = 0, which implies a = b by the properties of a . If an isometry is also surjective, it becomes bijective, and its inverse is likewise an isometry, as distances are preserved in both directions through . Such bijective isometries are often termed global isometries when considered on a single space. In contrast, a local isometry preserves distances only within neighborhoods of points, meaning for every x \in X, there exists a neighborhood U of x such that the restriction of f to U is an isometry onto its image in Y. This distinction allows local isometries to capture approximate geometric preservation without requiring global bijectivity, though they may fail to extend uniformly across the entire space. The collection of all bijective isometries from a (X, d) to itself, denoted \mathrm{Iso}(X), forms a group under , known as the of X. Composition of two isometries preserves distances, the map serves as the group , and inverses exist and are isometries, satisfying the group axioms. Basic examples illustrate these concepts in simple metric spaces. The identity map f(x) = x on any metric space (X, d) is the trivial isometry, preserving all distances exactly. On the real line \mathbb{R} equipped with the standard metric d(x, y) = |x - y|, translations f(x) = x + c for any constant c \in \mathbb{R} and reflections f(x) = -x + c are isometries, as both maintain absolute differences between points.

Isometries in Euclidean Geometry

Classification

Isometries of are broadly classified into direct isometries, which preserve , and opposite isometries, which reverse it. Direct isometries maintain the of figures, while opposite isometries invert it, corresponding to transformations with +1 and -1, respectively, in their linear parts. In two dimensions, direct isometries consist of translations and . Translations shift every point by a fixed , preserving parallelism and distances without rotation. turn points around a fixed by a specified . Opposite isometries in 2D include reflections over a line and glide reflections, which combine a reflection over a line with a translation parallel to that line. In three dimensions, the classification expands to account for the additional spatial freedom. Direct isometries comprise translations, , and screw displacements, which combine a about an with a translation along the same . Opposite isometries include , glide reflections (a across a followed by translation parallel to the ), inversions through a point, improper (a composed with a perpendicular to the ), and rotary inversions (a followed by an inversion through a point on the ). A fundamental classification theorem states that every isometry of \mathbb{R}^n is a of at most n+1 reflections, with isometries arising from an even number of reflections and isometries from an number. In , this limits compositions to at most three reflections, directly yielding the four types listed above. In , up to four reflections are possible, encompassing the extended set of transformations. Fixed-point properties further distinguish these types. Translations have no fixed points in any dimension. Rotations fix a single point (the center) in or a line (the ) in 3D. Reflections fix a line in or a in 3D. Glide reflections and screw displacements generally have no fixed points unless the translational component is zero, reducing to pure reflections or rotations. Inversions and rotary inversions fix a single point (the center).

Examples and Group Structure

In , concrete examples of isometries illustrate their role in preserving distances and . A in the by an angle θ around the origin is given by the matrix \begin{pmatrix} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{pmatrix}, which maps any point (x, y) to (\cos \theta \cdot x - \sin \theta \cdot y, \sin \theta \cdot x + \cos \theta \cdot y). provide another fundamental example; in three dimensions, a by a \mathbf{v} = (v_1, v_2, v_3) shifts every point \mathbf{x} to \mathbf{x} + \mathbf{v}, maintaining all distances without altering . Reflections, such as over the x-axis in the , map (x, y) to (x, -y) and reverse orientation while preserving distances. The full group of isometries in n-dimensional , denoted E(n), forms a ℝ^n ⋊ , where is the of linear isometries (rotations and reflections) and ℝ^n represents the translation ; the action is by conjugation, reflecting the of linear transformations followed by translations. The orientation-preserving , consisting of rotations and translations, is the ℝ^n ⋊ SO(n), where SO(n) is the special . Subgroups of E(n) capture discrete symmetries in geometric objects. For instance, the rotation group of the , a , is the A5 of order 60, extended by reflections to the full icosahedral group of order 120, which acts as a finite subgroup of E(3). In two dimensions, wallpaper groups classify the 17 distinct symmetry patterns of periodic tilings under translations, , reflections, and glide reflections. Three-dimensional analogues, , number 230 and describe crystal lattice symmetries, combining translations with operations. These group structures find applications in analyzing symmetries of physical systems. In molecular chemistry, the isometry group of a molecule like benzene (D6h symmetry) determines its vibrational modes and stability under Euclidean transformations. Similarly, in tiling theory, wallpaper groups underpin the classification of Islamic geometric patterns and quasicrystals, ensuring periodic or aperiodic arrangements preserve distances.

Isometries in Normed Spaces

General Isometries

In this section, we consider distance-preserving maps (isometric embeddings) between normed spaces; surjective such maps are bijective isometries in the sense of the article's general definition. In normed linear spaces over the real numbers, an isometry is a function f: X \to Y between normed spaces (X, \|\cdot\|_X) and (Y, \|\cdot\|_Y) that preserves distances induced by the norms, meaning \|f(x) - f(y)\|_Y = \|x - y\|_X for all x, y \in X. This condition is equivalent to norm preservation up to translation, as \|f(x) - f(0)\|_Y = \|x\|_X whenever f(0) = 0. Unlike isometries in metric spaces more generally, those in normed spaces leverage the vector space structure, though the maps need not be linear or continuous without additional assumptions. Not all isometries between normed spaces are surjective or bijective. In finite-dimensional spaces, surjectivity follows from injectivity by dimension arguments, but infinite-dimensional examples abound. For instance, the right T: \ell_p \to \ell_p ( $1 \leq p < \infty) defined by T(x_1, x_2, x_3, \dots) = (0, x_1, x_2, \dots) preserves the \ell_p-norm since \|T x\|_p^p = \sum_{n=1}^\infty |x_n|^p = \|x\|_p^p, yet its range excludes sequences starting with nonzero first components, such as (1, 0, 0, \dots). Such embeddings highlight how isometries can be proper inclusions in infinite-dimensional settings like \ell_p spaces. A foundational result characterizing many isometries is the , which states that every surjective isometry f: X \to Y between real normed spaces with f(0) = 0 is linear. More generally, any surjective isometry is affine, meaning f(x) = L(x) + b for some linear isometry L and translation vector b \in Y. The theorem's proof outline relies on showing that such f preserves convex combinations: first, it maps extreme points of the unit ball to extreme points; second, using the fact that midpoints in normed spaces satisfy \| (x+y)/2 \| \leq (\|x\| + \|y\|)/2 with equality under isometry preservation, f extends affinely to rational combinations and, by continuity, to all points. This result, originally established in 1932, underscores that linearity emerges from surjectivity and origin fixation in real scalar fields, but fails over complex numbers or without surjectivity. Examples of isometries in specific normed spaces illustrate these properties. In \mathbb{R}^n equipped with the Euclidean norm, all isometries are compositions of orthogonal linear maps and translations, aligning with the Mazur–Ulam conclusion since the space is finite-dimensional and strictly convex. In contrast, for the \ell_\infty norm on sequence spaces, coordinate permutations \pi: \ell_\infty \to \ell_\infty defined by (\pi x)_i = x_{\pi(i)} are linear isometries, as \|\pi x\|_\infty = \sup_i |x_{\pi(i)}| = \sup_i |x_i| = \|x\|_\infty, and signed variants (multiplying by \pm 1) also preserve the norm. Unlike Euclidean spaces, where all surjective isometries fixing the origin are linear orthogonal transformations, general normed spaces admit non-affine isometries when surjectivity fails, particularly in non-strictly convex norms like \ell_1. For example, consider the map f: \mathbb{R} \to \mathbb{R}^2 with the \ell_\infty norm on the codomain, given by f(t) = (t, \sin t); then \|f(t) - f(s)\|_\infty = \max(|t - s|, |\sin t - \sin s|) = |t - s|, since |\sin t - \sin s| \leq |t - s| by the mean value theorem, making f an isometry onto its image but non-affine (hence non-linear). Such constructions exploit the "flat" unit balls in norms like \ell_1 or \ell_\infty, allowing non-linear embeddings that preserve distances without extending affinely to the whole space.

Linear Isometries

In normed linear spaces X and Y, a linear isometry is a bounded linear operator T: X \to Y that preserves the norm, meaning \|Tx\| = \|x\| for all x \in X. In finite-dimensional Euclidean spaces equipped with the standard inner product, linear isometries are precisely those represented by orthogonal matrices A satisfying A^T A = I, where I is the identity matrix. Such operators preserve inner products, as norm preservation implies inner product preservation via the polarization identity: for real spaces, \langle x, y \rangle = \frac{1}{4} \left( \|x + y\|^2 - \|x - y\|^2 \right), and a similar formula holds for complex spaces involving additional terms with iy. In general normed spaces, including infinite-dimensional Banach spaces, linear isometries do not necessarily preserve inner products unless the space is Hilbert. For instance, in \ell_p spaces with $1 \leq p < \infty and p \neq 2, surjective linear isometries take the form T((a_n)) = (\epsilon_n a_{\pi(n)}), where \pi is a permutation of the indices and \epsilon_n = \pm 1; these operators preserve the \ell_p norm but fail to preserve the \ell_2 inner product. Similar structures appear in spaces like c_0 and c, where isometries involve coordinate permutations and sign changes without inducing an inner product structure. The Mazur-Ulam theorem, which states that surjective isometries between real normed spaces are affine, extends naturally to the linear case: a surjective linear isometry fixing the origin is precisely a norm-preserving linear operator, and in Hilbert spaces, such operators are unitary, satisfying T^* T = I = T T^*. In complex Hilbert spaces, unitary operators have all eigenvalues lying on the unit circle in the complex plane, reflecting their norm-preserving and phase-shifting nature.

Isometries on Manifolds

Definition and Types

An isometry between two Riemannian manifolds (M, g_M) and (N, g_N) is a smooth map f: M \to N that preserves the Riemannian metric tensor, meaning the pullback satisfies f^* g_N = g_M. Equivalently, for every point p \in M and tangent vectors X, Y \in T_p M, the inner product is preserved: g_M(X, Y) = g_N(df_p(X), df_p(Y)), where df_p: T_p M \to T_{f(p)} N is the differential of f. This definition ensures that f preserves lengths, angles, and the overall geometric structure induced by the metrics. Local isometries are smooth maps that preserve the metric in sufficiently small neighborhoods, specifically local diffeomorphisms \phi: (M, g_M) \to (N, g_N) such that \phi^* g_N = g_M locally around each point. For instance, covering maps between Riemannian manifolds, such as the universal cover of a manifold, are local isometries because they lift the metric exactly in local charts while possibly being non-injective globally. In contrast, global isometries are bijective smooth maps with smooth inverses (i.e., diffeomorphisms) that satisfy the pullback condition everywhere, ensuring a one-to-one correspondence between the entire manifolds. Isometric immersions preserve the metric but may self-intersect, whereas isometric embeddings are injective immersions that yield global isometries when combined with surjectivity. Among global isometries, which are precisely the diffeomorphisms preserving the metric, one distinguishes orientation-preserving types—those that preserve the orientation of tangent spaces—and orientation-reversing types, which reverse it, often corresponding to reflections or improper rotations in the isometry group. For example, the isometries of the n-sphere S^n with its round metric form the orthogonal group O(n+1), acting by rotations and reflections on the embedding in \mathbb{R}^{n+1}, where the special orthogonal subgroup SO(n+1) consists of the orientation-preserving isometries. Similarly, the isometries of the hyperbolic plane \mathbb{H}^2 (e.g., with the upper half-plane metric g = \frac{1}{y^2}(dx^2 + dy^2)) are given by Möbius transformations of the form z \mapsto \frac{az + b}{cz + d} with a, b, c, d \in \mathbb{R} and ad - bc = 1, with the orientation-preserving ones forming PSL(2, \mathbb{R}), while the full isometry group also includes orientation-reversing transformations such as reflections, forming a group isomorphic to PGL(2, \mathbb{R}).

Properties and Theorems

Isometries of a Riemannian manifold possess several fundamental properties that stem from their definition as metric-preserving diffeomorphisms. A key result is the , which establishes that the group of all isometries of a Riemannian manifold, endowed with the compact-open topology, forms a finite-dimensional whose dimension is at most n(n+1)/2, where n is the dimension of the manifold. This theorem underscores the smooth structure inherent in the isometry group, ensuring that local flows of infinitesimal isometries () integrate to global one-parameter subgroups of isometries. Isometries preserve a wide array of geometric structures derived from the Riemannian metric. Specifically, they map geodesics to geodesics, as the minimizing property of geodesic segments is invariant under distance-preserving maps. Moreover, since isometries pull back the metric tensor and thus the Levi-Civita connection, they preserve the Riemann curvature tensor pointwise, implying that sectional, Ricci, and scalar curvatures remain unchanged along corresponding points. Volumes are also preserved by orientation-preserving isometries, as the Riemannian volume form, defined via the determinant of the metric, transforms covariantly under such maps. Additionally, isometries conjugate Killing vector fields to Killing vector fields, since the Lie derivative condition for Killing fields—vanishing on the metric—is preserved under metric isomorphisms. Rigidity results highlight the constrained nature of isometries in certain manifolds. In spaces of constant sectional curvature, such as , spheres, or , any isometry is uniquely determined by its value and first derivative (its 1-jet) at a single point, reflecting the high degree of symmetry and the fact that the isometry group acts simply transitively on the tangent bundle. This uniqueness follows from the for complete simply connected manifolds of constant curvature, where the is a global diffeomorphism, allowing reconstruction of the map from local data. A notable distinction arises between local and global isometries on manifolds. While every local isometry preserves the metric in a neighborhood, it may not extend to a global isometry of the entire manifold; for instance, the universal cover of a compact hyperbolic surface admits local isometries from the surface that fail to extend globally due to the non-trivial fundamental group. Symmetric spaces provide concrete examples of these properties in action. These are Riemannian manifolds where the isometry group acts transitively, with the isotropy representation at any point being orthogonal, and the geodesic symmetries (reflections through a point) generate local isometries. In such spaces, the isometry group is a Lie group that decomposes as a semidirect product of the maximal compact subgroup and the vector space of parallel translations, ensuring transitive action and preservation of all curvature invariants.

Extensions and Generalizations

Approximate Isometries

In metric spaces, an approximate isometry, also known as an ε-isometry, is a map f: X \to Y between metric spaces (X, d_X) and (Y, d_Y) such that there exists \varepsilon > 0 with |d_X(a, b) - d_Y(f(a), f(b))| \leq \varepsilon for all a, b \in X. These maps preserve distances up to an additive error bounded by \varepsilon, making them useful for studying spaces where exact preservation is not required but bounded distortion is. Unlike exact isometries, ε-isometries need not be bijective, though they can approximate the geometry of the original space closely when \varepsilon is small. A more flexible notion in coarse geometry is that of a quasi-isometry, which allows both multiplicative and additive distortions. Specifically, a map f: X \to Y is a (\lambda, \varepsilon)-quasi-isometry if \lambda \geq 1, \varepsilon \geq 0, and it satisfies \frac{1}{\lambda} d_X(a, b) - \varepsilon \leq d_Y(f(a), f(b)) \leq \lambda d_X(a, b) + \varepsilon for all a, b \in X, along with a quasi-surjectivity condition: there exists R > 0 such that every point in Y is within distance R of the image f(X). Quasi-isometries capture the large-scale structure of metric spaces, ignoring bounded-scale features, and form an : if f: X \to Y is a quasi-isometry, then there exists a quasi-inverse g: Y \to X such that both d(f \circ g(y), y) and d(g \circ f(x), x) are bounded for all x \in X, y \in Y. This equivalence preserves asymptotic properties, such as connectivity at infinity or growth rates. Quasi-isometries play a central role in , where they classify up to their coarse geometry. For a G with finite generating set S, the word metric d_S(g, h) = |g^{-1}h|_S—the minimal length of a word in S \cup S^{-1} representing g^{-1}h—endows G with a structure. Different finite generating sets S and S' induce quasi-isometric word metrics on G, as the identity map between (G, d_S) and (G, d_{S'}) is a quasi-isometry. A key application is the quasi-isometric invariance of Gromov hyperbolicity: a X is \delta-hyperbolic if, for all x, y, z, w \in X, the Gromov product satisfies (x|y)_z \geq \min\{(x|w)_z, (y|w)_z\} - \delta, where (x|y)_z = \frac{1}{2}(d(z,x) + d(z,y) - d(x,y)); if f: X \to Y is a quasi-isometry, then X is hyperbolic if and only if Y is. This invariance implies that hyperbolicity is a geometric property of groups, independent of the choice of generating set, and enables the study of group actions on hyperbolic spaces via their boundaries. To quantify how far two metric spaces are from being isometric, the Gromov-Hausdorff distance provides a metric on the space of compact metric spaces up to isometry. For compact metric spaces X and Y, d_{GH}(X, Y) = \inf \{ d_H(\phi(X), \psi(Y)) \}, where the infimum is over all metric spaces Z and isometric embeddings \phi: X \to Z, \psi: Y \to Z, and d_H(A, B) = \max\{\sup_{a \in A} d_Z(a, B), \sup_{b \in B} d_Z(b, A)\} is the between subsets A, B \subseteq Z. If d_{GH}(X, Y) < \varepsilon, then there exists an \varepsilon-isometry between X and Y up to isometric embeddings into a common space. This distance extends to non-compact spaces via pointed versions and is used to study convergence of sequences of metric spaces in applications like manifold approximation and shape analysis. Bilipschitz maps offer a refinement of quasi-isometries focused on multiplicative control without additive terms. A map f: X \to Y is bilipschitz with constants c, C > 0 if c \, d_X(a, b) \leq d_Y(f(a), f(b)) \leq C \, d_X(a, b) for all a, b \in X; if bijective with bilipschitz inverse, it is a bilipschitz equivalence. Every bilipschitz map is a quasi-isometry (with \lambda = \max\{1/c, C\} and \varepsilon = 0), but the converse fails, as quasi-isometries allow additive errors that grow irrelevant at large scales. Bilipschitz equivalences preserve more refined geometric features, such as Assouad-Nagata dimension, and are crucial in embedding theory for metric spaces.

In Other Structures

In the field of compressed sensing, the (RIP) generalizes the notion of isometry to matrices acting on sparse vectors. A A \in \mathbb{R}^{m \times n} satisfies the RIP of order k with constant \delta_k if, for all k-sparse vectors x (those with at most k nonzero entries), (1 - \delta_k) \|x\|_2^2 \leq \|A x\|_2^2 \leq (1 + \delta_k) \|x\|_2^2, ensuring that A behaves like an isometry on the subspace of k-sparse signals, with singular values bounded near 1. This property, introduced by Candès, Romberg, and , guarantees stable recovery of sparse signals from underdetermined measurements via , such as basis pursuit. In C*-algebras, isometries are defined as surjective linear maps \phi: A \to B between C*-algebras that preserve the operator norm, i.e., \|\phi(a)\| = \|a\| for all a \in A. Such maps are precisely the *-isomorphisms, which additionally preserve the involution and algebraic structure, as established in foundational work on operator algebras. For example, spatial isometries arise from unitary representations, composing a unitary operator with the identity map to yield norm preservation. From a category-theoretic , isometries appear as in the category of spaces enriched over the monoidal poset ([0, \infty], \geq, +), where objects are sets equipped with as "hom-objects" and morphisms are distance-non-increasing maps. In this framework, due to Lawvere, an is a bijective enriched that admits an , corresponding exactly to a distance-preserving (isometry) with distance-preserving . This enriched structure unifies with categorical limits and colimits, such as Cauchy completions. Discrete isometries arise in theory, particularly for the \mathbb{Z}^n endowed with the Euclidean norm. An isometry here is a linear f: \mathbb{Z}^n \to \mathbb{Z}^n that preserves both the lattice structure and distances, equivalent to an element of the modular O(n, \mathbb{Z}) = \{ A \in GL(n, \mathbb{Z}) \mid A^T A = I \}. These maps induce relations on the lattice points, preserving volumes and modulo lattice translations, and play a key role in classifying coincidence site lattices in . For instance, in \mathbb{Z}^2, rotations by 90 degrees generate such isometries, forming finite subgroups isomorphic to cyclic or dihedral groups. Examples of isometries in other structures include distance-preserving maps on graphs and transformations in probability measures. A graph isometry is a bijection \phi: V(G) \to V(H) between vertex sets of graphs G and H such that the shortest-path distance satisfies d_G(u, v) = d_H(\phi(u), \phi(v)) for all vertices u, v, generalizing metric isometries to combinatorial settings; trees admit rich automorphism groups of such isometries. In probability theory, an isometry with respect to total variation distance on Borel probability measures is a surjective map T such that \| \mu - \nu \|_{TV} = \| T_\# \mu - T_\# \nu \|_{TV} for all measures \mu, \nu, where T_\# denotes pushforward; these are characterized as compositions of measurable bijections with their inverses preserving the distance.

References

  1. [1]
    [PDF] Isometries
    An isometry is a mapping of the plane where the segment connecting any two points is congruent to the segment connecting their images. It means 'preserving ...
  2. [2]
    [PDF] Classifying Isometries - MIT Mathematics
    An isometry is a geometric transformation that preserves distances between pairs of points. We present methods to classify isometries in the Euclidean plane, ...
  3. [3]
    [PDF] 2. SYMMETRY IN GEOMETRY 2.1. Isometries - User Web Pages
    Any function which satisfies this property is called an isometry. This comes from the ancient Greek words “isos”, meaning equal, and “metron”, meaning measure.<|control11|><|separator|>
  4. [4]
    Isometry -- from Wolfram MathWorld
    An isometry is a bijective map between metric spaces that preserves distances, where d(f(x),f(y))=d(x,y). In a plane, it's a linear transformation preserving ...
  5. [5]
    Non-Euclidean geometry - MacTutor History of Mathematics
    In the Bolyai - Lobachevsky type of geometry, straight lines have two infinitely distant points. In the Riemann type of spherical geometry, lines have no (or ...
  6. [6]
    Maurice Fréchet (1878 - 1973) - Biography - MacTutor
    In his dissertation of 1906, discussed above, he started a whole new area with his investigations of functionals on a metric space and formulated the abstract ...
  7. [7]
    [PDF] 3. The Motion of Rigid Bodies - DAMTP
    A rigid body is a collection of points with fixed distances. Its motion is translation plus rotation, described by a time-dependent orthogonal matrix.
  8. [8]
    [PDF] Group theory applied to crystallography - Mathematics
    Apr 29, 2008 · One of the crucial roles of space groups in crystallography is that they are used to classify the different crystal patterns. The idea here is ...
  9. [9]
    Orthogonality of isometries in the conformal model of the 3D space
    ... isometries that can be useful, for example, in applications of computational geometry, including computer graphics, robotics, and molecular geometry.
  10. [10]
    [PDF] Lecture notes on metric space and Gromov-Hausdorff distance
    Sep 29, 2017 · Definition 1.4 (Isometry). Given two metric space (X, dX), (Y,dY ), a func- tion f : X → Y preserves the distance ...<|control11|><|separator|>
  11. [11]
    [PDF] 21 2. Isometries and Rigid Motions of the Real Line Suppose two ...
    b) Every distance preserving function f : X → Y is injective. Hence, every isometry f : X → Y is a bijection.
  12. [12]
    [PDF] Length Spaces - ETH Zurich
    then it is called an isometry. We denote the isometry group of a metric space X by Isom(X). The map f : X →. ¯. X is a local isometry if for every x ∈ X there ...
  13. [13]
    Isometry groups of proper metric spaces
    Oct 28, 2013 · For every metric space (Y, ), Iso(Y, ) stands for the group of all. (bijective) isometries of (Y, ). Isometry groups (equipped with the topology ...
  14. [14]
    [PDF] MA 398 Homework 1: Metric Spaces go the distance!
    An isometry of a metric space (X,d) is a function f : X → X such that for all x,y ∈ X, d(f(x), f(y)) = d(x,y). (1) Give two examples of ...
  15. [15]
    [PDF] isometries of the plane - UChicago Math
    Jul 21, 2009 · What is an isometry? An isometry is a distance-preserving transformation. In this paper, we consider isometries of the plane C.Missing: mathematics | Show results with:mathematics
  16. [16]
    [PDF] Euclidean isometries and surfaces - UChicago Math
    Before we attempt to classify all the euclidean isometries, we begin with several familiar examples of isometries: translations, reflections, and rotations.Missing: congruency | Show results with:congruency
  17. [17]
    [PDF] Lecture Notes in Modern Geometry 1 The euclidean plane
    Now using the three reflections theorem, we prove the classification theorem for euclidean isometries. ... Let's first understand the geometry of such isometries.
  18. [18]
    [PDF] ISOMETRY ON LINEAR n-NORMED SPACES
    Introduction. Let X and Y be metric spaces. A mapping f : X → Y is called an isometry if it satisfies dY (f(x),f(y)) = dX(x, y) for all x, y ∈ X, ...
  19. [19]
    isometry in nLab
    May 15, 2024 · An isometry is a function that preserves a metric, either in the sense of a metric space or in the sense of a Riemannian manifold.
  20. [20]
    [PDF] arXiv:2008.03668v2 [math.FA] 1 Mar 2022
    Mar 1, 2022 · In the framework of infinite dimensional Hilbert spaces this is not valid, since the right shift operator A is an isometry (A∗A = I) but not ...<|separator|>
  21. [21]
    [1306.2380] The Mazur-Ulam theorem - arXiv
    Jun 10, 2013 · Abstract:A short proof of the Mazur-Ulam theorem concerning isometries of real normed spaces. Subjects: Metric Geometry (math.MG).Missing: statement | Show results with:statement
  22. [22]
    The Mazur–Ulam theorem - ScienceDirect.com
    Recall that a map α : X → Y between real normed spaces is said to be affine if it satisfies α ( ( 1 − t ) x 1 + t x 2 ) = ( 1 − t ) α ( x 1 ) + t α ( x 2 ) ...
  23. [23]
    [PDF] Isometries of Finite-Dimensional Normed Spaces†
    An isometry between the normed spaces (E,k k1) and (F,k k2) is a linear application T : E → F such that, for all x ∈ E, kTxk2 = kxk1. It is clear that an ...
  24. [24]
    Surjective linear isometries on $\ell_\infty(\mathbb{N})$ - MathOverflow
    Sep 3, 2021 · Given an isometry, after composing with a diagonal-standard isometry, we can suppose that f(1X)=1X, i.e., f∈G, where G is the stabilizer of 1X.What are the matrices preserving the $\ell^1$-norm?A space isometric to $\ell_\infty^2$More results from mathoverflow.netMissing: norm | Show results with:norm
  25. [25]
    Example of a non-linear isometry? - Mathematics Stack Exchange
    Aug 24, 2011 · Is there a simple example of an isometry between normed vector spaces that is not an affine map? linear-algebra · Share.What matrices preserve the $L_1$ norm for positive, unit norm ...What are the linear isometries on Rn, equipped with the l1 norm?More results from math.stackexchange.com
  26. [26]
    [PDF] KEITH CONRAD - 1. Introduction An isometry of Rn is a function h ...
    Every isometry of Rn can be uniquely written as the composition t◦k where t is a translation and k is an isometry fixing the origin. ... Mathematics 22 (1972), ...
  27. [27]
    [PDF] MATH 423 Linear Algebra II Lecture 28: Inner product spaces.
    Theorem 2 (Polarization Identity) Suppose V is an inner product space with an inner product (·,·) and the induced norm ‖·‖. (i) If V is a real vector space, ...
  28. [28]
    [PDF] arXiv:2110.11786v1 [math.FA] 22 Oct 2021
    Oct 22, 2021 · In the current paper, we survey many results characterizing the group of surjective linear isometries on various Banach spaces. Our emphasis is ...
  29. [29]
    100. Isometries on Hilbert Spaces - Project Euclid
    Isometrics on Hilbert Spaces. 437. The preceding lemma will illustrate the fact that an isometry is precisely decomposed as the direct sum of a unitary operator ...
  30. [30]
    [PDF] lecture 2: the riemannian metric
    For the standard metrics on Rm and Tm: if we regard Tm = Rm/Zm, then the projection π : (Rm,g0) → (Tm,g0) is a local isometry but not a global isometry.
  31. [31]
    [PDF] 3 Riemannian metric - Durham University
    Riemannian metric g is called a Riemannian manifold (M,g). ... A diffeomorphism f : (M,g) → (N,h) of two Riemannian manifolds is an isometry if. Df(p) : TpM → Tf( ...
  32. [32]
    [PDF] Chapter 16 Isometries, Local Isometries, Riemannian Coverings and ...
    Definition 16.3. Given a Riemannian manifold (M,g), a vector field X is a Killing vector field iff the Lie deriva- tive of ...
  33. [33]
    [PDF] Riemannian manifolds - IME-USP
    4 Theorem (Myers-Steenrod) The isometry group Isom(M,g) of a Riemannian manifold. (M,g) has the structure of a Lie group with respect to the compact-open ...
  34. [34]
    [PDF] ISOMETRIES OF THE HYPERBOLIC PLANE - UChicago Math
    Now we will see that there exist Möbius transformations that map a given ge- odesic to another given geodesic and a point on that geodesic to another given.
  35. [35]
    [PDF] 5 The hyperbolic plane - UC Davis Math
    Using geodesics we can now show that any isometry is a Möbius transformation as above. So suppose that F : D → D is an isometry. Take a Möbius isometry G taking.
  36. [36]
    Smooth semi-Lipschitz functions and almost isometries between ...
    Nov 1, 2020 · On the other hand, the term almost isometry is sometimes used with a different meaning, in connection with coarse equivalence of metric spaces.
  37. [37]
    [PDF] arXiv:1904.03504v3 [math.MG] 17 Sep 2019
    Sep 17, 2019 · Theorem 3.4. Let MAI denote the category of discrete metric spaces of bounded geometry with almost isometries as morphisms. Then the mapping f 7 ...
  38. [38]
    [PDF] HYPERBOLIC GROUPS - M. Gromov - IHES
    In fact, every 0-hyperbolic space isometrically embeds into tree and an arbitrary 8-hyperbolic space can be approximated by trees (see §6). (a). 1.2 First ...Missing: original | Show results with:original
  39. [39]
    [PDF] AN INTRODUCTION TO QUASI-ISOMETRY AND HYPERBOLIC ...
    Jonathan Block, Advisor. A Quasi-Isometry is map between metric spaces that allows us to distort distances but preserves the large scale geometry of the spaces.
  40. [40]
    [PDF] Metric structures for Riemannian and non-Riemannian spaces - IHES
    Page 1. Metric Structures for Riemannian and Non-riemannian Spaces. M. Gromov edited by J. Lafontaine and P. Pansu. Page 2 ...Meme ceux qui furent favorables ...
  41. [41]
    [PDF] Lipschitz Maps in Metric Spaces - eScholarship
    Definition 1.1.2. A mapping f : X → Y is called bi-Lipschitz, or C-bi-Lipschitz to empha- size the constant, if there is a constant C ≥ 1 such that. C−1d(x1,x2 ...
  42. [42]
    Bilipschitz equivalence is not equivalent to quasi-isometric ... - arXiv
    Apr 23, 2009 · We show that certain lamplighter groups that are quasi-isometric to each other are not bilipschitz equivalent. This gives a positive answer to a question.
  43. [43]
    [PDF] The Restricted Isometry Property and Its Implications for ...
    Feb 27, 2008 · This technique known as “compressed sensing” or. “compressive sampling” relies on properties of the sensing matrix such as the restricted ...Missing: seminal | Show results with:seminal
  44. [44]
    [PDF] Isometries of operator algebras
    Isometries between operator algebras can be anti-isomorphisms or isomorphisms, and are composites of a unitary multiplication and a map preserving the C*- ...
  45. [45]
    Isometries between C*-algebras - Project Euclid
    A linear isometry T between C*-algebras A and B, locally reduces to a Jordan triple isomorphism by a projection, and preserves the Jordan triple product.
  46. [46]
    [PDF] METRIC SPACES, GENERALIZED LOGIC, AND CLOSED ... - GitHub
    Key words and phrases: Metric spaces, enriched categories, logic. Commentary (©) F. William Lawvere, 2002. Permission to copy for private use granted. 1 ...<|control11|><|separator|>
  47. [47]
    Indices of Coincidence Isometries of the Hyper Cubic Lattice $Z^n
    Aug 29, 2006 · Abstract: The problem of computing the index of a coincidence isometry of the hyper cubic lattice \mathbb{Z}^{n} is considered.Missing: congruence | Show results with:congruence
  48. [48]
    [PDF] ON ISOMETRIC EMBEDDINGS OF GRAPHS - UCSD Math
    If G is a finite connected graph with vertex set V and edge set E, a standard way of defining a distance dc on G is to define dG(x, y) to be the number of edges ...
  49. [49]
    Isometries of probability measures with respect to the total variation ...
    Mar 15, 2022 · Isometries of probability measures with respect to total variation distance are characterized as surjective isometries, and are push-forwards ...