Fact-checked by Grok 2 weeks ago

Discrete group

In , a discrete group is a topological group endowed with the discrete . Discrete groups frequently arise as discrete subgroups of larger Lie groups, providing a bridge between algebraic and geometric structures; for instance, the \mathrm{SL}_2(\mathbb{Z}) serves as a classic example, acting discontinuously on the hyperbolic plane via Möbius transformations. In geometric group theory, discrete groups are typically understood as countable groups with this , enabling the study of their large-scale geometry through tools like word metrics and Cayley graphs, which model the group's combinatorial structure. These graphs, with vertices representing group elements and edges corresponding to generators, reveal quasi-isometric invariants that classify groups up to coarse equivalence, a cornerstone for analyzing fundamental groups of manifolds and solving problems in low-dimensional . The significance of discrete groups extends to their role in symmetries of discrete objects, such as lattices, tilings, and polyhedra, where they generate fundamental domains and orbifolds in quotient spaces. Finitely generated and finitely presented discrete groups, defined by finite sets of generators and relations, are particularly tractable, allowing computations of asphericity via associated 2-complexes. Overall, discrete groups underpin advancements in areas like and theory, influencing .

Definition

Topological definition

A G is called if it is equipped with the , in which every of G is open. Equivalently, G is if the e has a neighborhood containing no other points of G, meaning e is isolated and, by the homogeneity of , every point is isolated. In the on any set, the group operations of and inversion are automatically continuous, as all functions between are continuous. A subgroup H of a G is if H, endowed with the induced from G, has the , so that every \{ h \} for h \in H is open in H. This means each point of H is isolated in the relative topology, ensuring no accumulation points within H. Unlike general topological groups, where the topology may allow points to accumulate and connected components to form, discreteness enforces complete separation of points while preserving the continuity of the group operations.

Equivalent conditions

A topological group G equipped with the discrete topology is discrete if and only if the identity element e is an isolated point in G, meaning there exists an open neighborhood U of e such that U \cap G = \{e\}. Since left translations by elements of G are homeomorphisms, this condition at the identity extends to every point in G being isolated, confirming that the subspace topology on G is the full discrete topology. For a subgroup H of a topological group G, an equivalent characterization of discreteness is that every sequence in H consisting of distinct elements has no convergent subsequence in G. This follows from the fact that in the subspace topology on H, discreteness implies no limit points within H, and thus no non-constant convergent sequences in H, with convergence in G restricted to H yielding the same property. Such sequences diverge in G, ensuring that H inherits the discrete topology without accumulation points from the ambient space. In the special case where the topological group G is metrizable, a H is there exists \varepsilon > 0 such that for every h \in H, the open ball of radius \varepsilon centered at h contains no other elements of H. This uniform separation condition leverages the to verify of points in H, equivalent to the topological definition via the first-countability of metric spaces, where sequences suffice to detect limit points. These characterizations typically assume that the topological group G is Hausdorff, as non-Hausdorff topologies can lead to pathologies where subgroups fail to be closed or where fails to propagate uniformly. Standard treatments of , including those involving subgroups, therefore impose the to ensure well-behaved convergence and quotient structures.

Properties

Topological properties

A discrete group, equipped with the discrete topology, is totally disconnected as a , meaning that its only connected subsets are ; this follows from the fact that every is both open and closed in the discrete topology. Consequently, the connected component of the consists solely of the identity itself. Every group is locally compact, since each point admits a compact neighborhood—namely, the singleton set containing that point, which is compact in the . Moreover, compact neighborhoods in a group are precisely the finite subsets, as any infinite subset lacks a finite subcover for its open cover by singletons. Finite groups are therefore compact, while infinite ones are not. Discrete groups are always Hausdorff, as the discrete topology separates distinct points with disjoint open singletons. Their topology is zero-dimensional, possessing a basis of clopen sets (the singletons). In particular, any countable discrete group is second-countable, rendering it a zero-dimensional Lie group. Discrete groups admit a Bohr compactification, which is a compact Hausdorff group bG together with a continuous homomorphism \sigma: G \to bG that is injective on G and dense in bG, such that every continuous homomorphism from G to a compact group factors through bG. For discrete G, the image \sigma(G) is dense in the compact group bG, extending the discrete topology continuously to this compactification.

Algebraic properties

A discrete topological group inherits its algebraic structure from the underlying group operation, with the discrete topology ensuring that certain closure properties hold under compatible topologies. Specifically, any of a discrete topological group is itself discrete in the , as the discrete topology on the ambient group restricts to the discrete topology on the . Similarly, if N is a closed of a discrete topological group G, the G/N equipped with the is also discrete, since the cosets are open sets in G and the quotient map is open. In connected topological groups, normal discrete subgroups exhibit particularly restrictive algebraic behavior. A discrete normal subgroup N of a connected topological group G must be central, meaning that every element of N commutes with every element of G. This follows from the fact that conjugation maps g \mapsto gng^{-1} for n \in N are continuous homomorphisms from the connected group G to the discrete group N, hence constant, implying gng^{-1} = n for all g \in G. Consequently, such subgroups are abelian, as central subgroups of any group are abelian. The discreteness of a group also imposes constraints on its when viewed as a of a larger with countability axioms. In particular, every discrete of a second-countable is countable. This arises because the second-countability of the ambient group allows a countable basis, and the discreteness ensures that the subgroup injects into a countable collection of cosets via a fundamental domain of finite measure. Regarding compactness, finite discrete groups are compact, as they are finite spaces in the topology. However, infinite discrete groups cannot be compact in the Hausdorff sense, since the open cover by singletons admits no finite subcover. This finiteness condition highlights the interplay between the algebraic infinitude of the group and its topological , preventing infinite discrete groups from being compact without altering the topology.

Examples

Discrete subgroups of the real line

The additive group of real numbers (\mathbb{R}, +) endowed with its standard metric topology provides a simple setting to illustrate discrete subgroups. In this context, a subgroup H \subseteq \mathbb{R} is discrete if it inherits the discrete topology from the subspace topology, meaning every point in H is isolated; equivalently, for every h \in H, there exists \epsilon > 0 such that the open ball B(h, \epsilon) \cap H = \{h\}. This isolation condition holds precisely when the elements of H are separated by a positive minimal distance. A fundamental classification theorem states that every proper subgroup of (\mathbb{R}, +) is either dense in \mathbb{R} or cyclic, generated by a single nonnegative real number a \geq 0. The dense subgroups fail to be discrete, as their elements accumulate everywhere without . In contrast, the cyclic subgroups a\mathbb{Z} = \{ka \mid k \in \mathbb{Z}\} (with a > 0) are discrete, since consecutive elements are separated by a, allowing an open ball of radius a/2 around each point to contain no other subgroup elements. The case a = 1 yields the integers \mathbb{Z} as a prototypical example, where the minimal between distinct points is 1, ensuring in the subspace topology induced from \mathbb{R}. The trivial subgroup \{0\} is also discrete, as its sole element is isolated by any positive-radius ball centered at 0 that intersects the subgroup only at itself. More generally, for any integer n > 0, the subgroup n\mathbb{Z} is discrete with spacing n, generated by n and forming a lattice-like structure along the line. A key non-example is the rationals \mathbb{Q}, which is dense in \mathbb{R} because between any two reals there lies a rational, violating the isolation condition and rendering \mathbb{Q} non-discrete. This distance-based verification aligns with equivalent conditions for discreteness in metric topological groups like \mathbb{R}, where closedness and proper separation suffice to confirm the discrete subspace topology.

Discrete symmetry groups

Discrete symmetry groups arise as discrete subgroups of the isometry groups of Euclidean spaces, acting properly discontinuously to produce tilings with periodic symmetries. These groups capture the rigid motions—translations, rotations, reflections, and glide reflections—that preserve infinite patterns in the plane or space, fundamental to understanding geometric repetition and order. Frieze groups consist of discrete isometries of the that preserve an infinite horizontal strip, such as a band of repeating motifs along a line. They are generated by translations along the strip and additional symmetries like rotations or reflections perpendicular to it, with the translation isomorphic to the integers \mathbb{Z}. There are precisely seven groups, classified up to conjugacy by their geometric symmetries despite some algebraic isomorphisms. These include: (1) translations only (); (2) glide reflections (step); (3) translations and vertical reflections (sidle); (4) translations and 180° rotations (spinning hop); (5) glide reflections and 180° rotations (spinning sidle); (6) translations and horizontal reflections (); and (7) translations, horizontal and vertical reflections (all symmetries). Examples feature translations for basic repetition and glide reflections combining translation with reflection over a parallel line, enabling patterns like those in key designs or modern borders. Extending to full plane symmetries, are the 17 discrete subgroups of the Euclidean plane's that act periodically via a of translations, enabling infinite without gaps or overlaps. Classified by the International Union of Crystallography, they combine five types—, rectangular, centered rectangular, square, and hexagonal—with symmetries restricted by the crystallographic limitation theorem, allowing only of orders 1, 2, 3, 4, or 6. Key elements include translations forming a rank-2 abelian , about points, reflections over lines, and glide reflections; notation like p4mm encodes cell type, maximal order, and mirror presence. These groups underpin the of two-dimensional periodic structures, from Islamic geometric art to molecular layers, where each dictates compatible shapes and orientations. In three dimensions, crystallographic groups, or space groups, generalize this to discrete subgroups of the isometry group, generating periodic lattices for crystal structures. There are 230 such groups when distinguishing enantiomorphic pairs (mirror images), or 219 affine types otherwise, enumerated independently in the 1890s by W. Barlow, E.S. Federov, and A. Schönflies using extensions of two-dimensional classifications. They incorporate 14 Bravais lattices with 32 point groups, augmented by screw rotations and glide planes, finite in number due to the crystallographic restriction. These groups describe the symmetries of atomic arrangements in solids, emphasizing translational periodicity in all directions for tiling . A notable example beyond is the \mathrm{PSL}(2, \mathbb{Z}), the projective over the integers, which forms a discrete subgroup of \mathrm{PSL}(2, \mathbb{R}). It acts on the \mathbb{H}^2 via Möbius transformations, producing a fundamental domain like the standard modular region bounded by geodesics. This action yields a non-Euclidean with triangular faces, highlighting discrete symmetries in curved geometries. In contrast, the full rotation group \mathrm{SO}(3) of three-dimensional Euclidean space is a connected Lie group, homeomorphic to the real projective space \mathbb{RP}^3, and thus not discrete, as its continuous paths connect any two rotations without isolated elements.

Other notable examples

Fuchsian groups are discrete subgroups of the projective special linear group \mathrm{PSL}(2, \mathbb{R}), which consists of orientation-preserving isometries of the hyperbolic plane \mathbb{H}^2. These groups, introduced by Henri Poincaré in his 1882 memoir Théorie des groupes fuchsiens, play a key role in the uniformization of Riemann surfaces and the study of hyperbolic geometry. A prominent subclass consists of Schottky groups, which are freely generated by hyperbolic elements pairing disjoint simple closed curves (circles) on the Riemann sphere, ensuring the group acts freely and properly discontinuously on \mathbb{H}^2. Kleinian groups extend this concept to three dimensions, defined as discrete subgroups of \mathrm{PSL}(2, \mathbb{C}), the group of orientation-preserving isometries of hyperbolic 3-space \mathbb{H}^3. Also originating from Poincaré's work in 1883, these groups are fundamental to the geometrization of 3-manifolds, where quotients \mathbb{H}^3 / \Gamma for a Kleinian group \Gamma yield hyperbolic structures on manifolds. Their limit sets on the sphere at infinity determine the and of the quotients, with applications in Thurston's hyperbolization . Free groups on a finite number of generators, when endowed with the topology, form basic examples of discrete groups, as every is open and the is properly discontinuous by definition. These groups arise naturally in combinatorial and as groups of surfaces with punctures. Profinite groups, constructed as inverse limits of finite groups each equipped with the topology, yield discrete topological groups precisely when the resulting group is finite, such as the of order n as \varprojlim \mathbb{Z}/p^k\mathbb{Z} for prime p dividing n. Infinite profinite groups, like the p-adic integers \mathbb{Z}_p, carry a non- compact topology. In contrast, not all subgroups of topological groups are ; for instance, the of the circle group \mathbb{T} = \mathbb{R}/\mathbb{Z} generated by an irrational rotation r_\alpha(x) = x + \alpha \mod 1 with \alpha irrational is dense in \mathbb{T}, hence non-.

Discrete subgroups of Lie groups

Lattices and fundamental domains

In the context of s, a is defined as a \Gamma of a locally compact G such that the quotient G/\Gamma has finite . This finite-volume condition ensures that \Gamma is "maximal" in a measure-theoretic sense among subgroups, as the action of \Gamma on G partitions the into regions of bounded total volume. play a central role in the study of homogeneous spaces, where the quotient G/\Gamma often forms a manifold with rich geometric structure. Lattices are classified as uniform or non-uniform based on the compactness of the quotient. A uniform lattice \Gamma in G yields a compact quotient G/\Gamma, meaning the fundamental domain can be chosen to be compact, which implies \Gamma acts cocompactly on G. In contrast, a non-uniform lattice produces a non-compact quotient with finite volume, often featuring cusps or ends where the group action thins out. A classic example is the modular group \mathrm{SL}(2, \mathbb{Z}) acting as a non-uniform lattice in \mathrm{SL}(2, \mathbb{R}), where the quotient is the modular surface with a cusp at infinity. Associated with any \Gamma in a G is a fundamental domain D \subset G, an such that the \Gamma-translates \gamma D for \gamma \in \Gamma G without interior overlap and G a set of measure zero. The domain D has finite volume, matching that of G/\Gamma, and serves as a "fundamental region" for reduction theory, allowing explicit descriptions of orbits and invariants under the . For rank-one semisimple groups, such domains can be constructed explicitly using the geometry of symmetric spaces, facilitating the study of properties and volume computations. The provides a powerful tool for analyzing in groups like \mathrm{PSL}(2, \mathbb{R}), which act on . For a \Gamma yielding a M = \mathbb{H}^2 / \Gamma, the formula equates the trace of the (related to Laplacian eigenvalues on M) to a sum over lengths of closed geodesics in the fundamental domain, weighted by orbital contributions. This spectral-geometric duality reveals how the distribution of eigenvalues encodes the geometry of the fundamental domain, with applications to eigenvalue estimates and prime geodesic theorems on such manifolds. Rigidity theorems further highlight the structural stability of lattices in higher- semisimple groups. The Mostow-Prasad rigidity theorem states that for irreducible lattices \Gamma_1 and \Gamma_2 in semisimple groups G_1 and G_2 of rank at least two, with finite-volume quotients, any \Gamma_1 \cong \Gamma_2 extends to a unique (up to conjugation) G_1 \cong G_2. Originally proved by Mostow for cocompact lattices in 1968 and extended by Prasad to finite-volume cases in 1973, this result implies that the geometry and representation of such lattices are rigidly determined by their abstract group structure.

Arithmetic and congruence subgroups

Arithmetic groups are discrete subgroups of Lie groups that arise from algebraic groups defined over the rational numbers \mathbb{Q}. Specifically, a subgroup \Gamma of a semisimple Lie group G is arithmetic if it is commensurable with G(\mathbb{Z}), where G is an algebraic group defined over \mathbb{Q} embedded into G(\mathbb{R}), and commensurable means that the intersection of \Gamma and G(\mathbb{Z}) has finite index in both. These groups are lattices, meaning they are discrete subgroups with finite covolume in G. A canonical example is \mathrm{SL}(n, \mathbb{Z}) as a discrete subgroup of \mathrm{SL}(n, \mathbb{R}) for n \geq 2. Congruence subgroups form an important class of finite-index subgroups within arithmetic groups. A congruence subgroup of an arithmetic group \Gamma = G(\mathbb{Z}) is defined by congruence conditions modulo some integer m, specifically as the preimage under the reduction map G(\mathbb{Z}) \to G(\mathbb{Z}/m\mathbb{Z}). The principal congruence subgroup \Gamma(m) is the kernel of this map, consisting of elements that reduce to the identity modulo m. For instance, in \mathrm{SL}(2, \mathbb{Z}), \Gamma(m) = \{ A \in \mathrm{SL}(2, \mathbb{Z}) \mid A \equiv I \pmod{m} \}, which is normal and of index growing with m. Notable examples of arithmetic groups include the Bianchi groups, which are \mathrm{SL}(2, \mathcal{O}_d) where \mathcal{O}_d is the ring of integers in the imaginary quadratic field \mathbb{Q}(\sqrt{-d}) for square-free positive integers d, acting discretely on hyperbolic 3-space. Another class is the Hilbert modular groups, such as \mathrm{SL}(2, \mathcal{O}_K) for a totally real number field K of degree greater than 1, where \mathcal{O}_K is the ring of integers of K, embedding into products of \mathrm{SL}(2, \mathbb{R}). Margulis superrigidity provides a profound : for irreducible lattices \Gamma in a higher-rank semisimple G over \mathbb{Q} (with real rank at least 2 and not isogenous to \mathrm{SO}(1,n) \times K or \mathrm{SU}(1,n) \times K), \Gamma must be . This implies that such lattices exhibit strong rigidity in their representations and homomorphisms. The congruence subgroup problem investigates whether every finite-index subgroup of an group \Gamma = G(k) (with k a number field) contains a principal of finite index, or equivalently, if all finite-index subgroups are . For \mathrm{SL}(n, \mathbb{Z}) with n \geq 3, the problem has a positive solution: every finite-index subgroup is a . In general, for semisimple groups of \mathbb{Q}-rank at least 2, the problem is affirmatively resolved, though counterexamples exist in lower ranks, such as non- subgroups in \mathrm{SL}(2, \mathbb{Z}).

Applications

In geometry and crystallography

In geometry, discrete groups play a fundamental role in describing the symmetries of periodic structures, particularly in where they classify the possible arrangements of atoms in . The symmetry groups of crystals in three dimensions are known as space groups, which are discrete subgroups of the motion group \mathbb{E}(3) that act properly discontinuously on \mathbb{R}^3. There are exactly 230 such space groups, encompassing all combinations of types, symmetries, screw axes, and glide planes that preserve the crystal . Discrete groups also arise in the construction of manifolds and orbifolds through quotient spaces. When a discrete group \Gamma acts freely and properly discontinuously on a manifold M (i.e., \Gamma is torsion-free), the quotient M/\Gamma is a manifold without singularities. If \Gamma has torsion elements, the quotient M/\Gamma forms an , which generalizes manifolds by allowing isolated singular points corresponding to finite stabilizers. In hyperbolic geometry, discrete groups such as Fuchsian and Kleinian groups generate important classes of surfaces and three-manifolds. A Fuchsian group is a discrete subgroup of \mathrm{PSL}(2,\mathbb{R}), acting on the hyperbolic plane \mathbb{H}^2, and its quotient by a fundamental domain yields a hyperbolic surface of finite area. Similarly, Kleinian groups, discrete subgroups of \mathrm{PSL}(2,\mathbb{C}) acting on hyperbolic three-space \mathbb{H}^3, produce hyperbolic three-manifolds via fundamental domains, providing models for low-dimensional topology. The Bieberbach theorems characterize groups on . These theorems state that for a \Gamma of the \mathbb{E}(n) properly discontinuously and cocompactly on \mathbb{R}^n, the translation \Gamma_\mathrm{trans} is a isomorphic to \mathbb{Z}^n of full rank, and \Gamma is a \mathbb{Z}^n \rtimes F where F is a finite point group faithfully on the . Such groups, known as Euclidean crystallographic groups, underpin the symmetry of periodic tilings in n-dimensions. Beyond periodic crystals, groups model aperiodic structures in physical materials like quasicrystals, which exhibit long-range order without translational periodicity. Quasicrystals are described by patterns arising from tilings generated by operations, often involving icosahedral or five-fold rotational symmetries incompatible with periodicity. A prominent example is the of the plane, constructed via inflation rules that substitute larger prototiles with clusters of smaller ones, yielding aperiodic tilings with rotational symmetries and modeling the atomic arrangements in real quasicrystalline alloys.

In number theory and arithmetic geometry

In , discrete groups play a central role in the study of modular forms, which are automorphic forms defined on quotients of the upper half-plane \mathbb{H} by congruence subgroups \Gamma of \mathrm{SL}(2, \mathbb{Z}). These forms, such as , exhibit transformation properties under the action of \Gamma, enabling the construction of spaces of cusp forms whose dimensions are governed by arithmetic invariants like the level and weight. Arithmetic subgroups like \mathrm{SL}(2, \mathbb{Z}) provide the foundational discrete structure for these quotients, linking holomorphic properties to number-theoretic phenomena such as the distribution of primes. Discrete groups also connect to L-functions and zeta functions through cohomology of arithmetic groups like \mathrm{GL}(n, \mathbb{Z}). The residue of the of a number field at s=1 involves the class number, which is interpreted via the idele class group in and reflects the structure of ideals. For higher-rank groups, automorphic L-functions attached to representations of \mathrm{GL}(n, \mathbb{A}_\mathbb{Q})/\mathrm{GL}(n, \mathbb{Q}) encode analytic continuations and functional equations, where \mathrm{GL}(n, \mathbb{Q}) embeds discretely and relates to cohomology classes that compute special values. In arithmetic geometry, étale fundamental groups serve as profinite discrete groups that classify Galois representations associated to varieties over number fields. The étale fundamental group \pi_1^{\text{ét}}(X_{\overline{K}}) of a scheme X over a number field K is profinite and captures unramified covers, with continuous representations \rho: \pi_1^{\text{ét}}(X_{\overline{K}}) \to \mathrm{GL}_n(\overline{\mathbb{Q}}_\ell) corresponding to l-adic cohomology sheaves. These groups, introduced by Grothendieck, replace topological fundamental groups and enable the study of arithmetic invariants like the conductor through their profinite topology. The establishes deep links between representations of arithmetic discrete groups and automorphic forms on adelic quotients. It conjectures correspondences between irreducible Galois representations of \mathrm{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) (a ) and cuspidal automorphic representations of \mathrm{GL}(n, \mathbb{A}_\mathbb{Q})/\mathrm{GL}(n, \mathbb{Q}), where \mathrm{GL}(n, \mathbb{Q}) embeds discretely. These correspondences preserve L-functions and imply reciprocity laws, unifying with via the adelic . The Shafarevich theorem, proved by Faltings, asserts the finiteness of isomorphism classes of abelian varieties over a number field with bounded , relying on discrete group actions in . Specifically, it bounds the set of such varieties up to isomorphism by controlling torsors under the action of the , whose profinite structure ensures only finitely many isomorphism classes satisfy the conductor condition. This finiteness extends to semiabelian varieties and underpins results on the arithmetic of Jacobians.

References

  1. [1]
    Discrete Group -- from Wolfram MathWorld
    A discrete group is a topological group with the discrete topology. Often in practice, discrete groups arise as discrete subgroups of continuous Lie groups ...Missing: definition | Show results with:definition
  2. [2]
    On the Geometry and Topology of Discrete Groups: An Overview
    Mar 4, 2024 · For a discrete group, we simply mean a countable group with the discrete topology. Definition 1. A discrete group Γ is said to be finitely ...
  3. [3]
    [PDF] A Crash Course in Topological Groups - Cornell Mathematics
    Nov 8, 2011 · topological groups. Any group taken with the discrete topology. Any (arbitrary) direct product of these with the product topology. Note that ...
  4. [4]
    None
    ### Summary of Discrete Topological Group/Discrete Subgroup Definition
  5. [5]
    [PDF] TOPOLOGICAL GROUPS: A Topological Background
    If we let V = ∩n i=1Ui, we have F ∩ V = ∅, and C ⊂ V , as desired. D. Page 11. TOPOLOGICAL GROUPS. 11. Theorem 4.1.
  6. [6]
    [PDF] Fuchsian Groups: Intro - UCSD Math
    topological group. A discrete subgroup of a topological group G is a subgroup of G which, as a topological space, inherits the discrete topology from G.
  7. [7]
    [PDF] Part I Basic Properties of Topological Groups
    May 19, 2020 · This chapter contains basic results about the point-set topology of topological groups. Whenever we use the Hausdorff condition, ...
  8. [8]
    [PDF] 2.4 Topological Groups - UPenn CIS
    Since Lie groups are topological groups (and manifolds), it is useful to gather a few basic facts about topological groups. Definition 2.11. A set, G, is a ...
  9. [9]
    [PDF] Introduction to Topological Groups
    Apr 12, 2015 · These notes provide a brief introduction to topological groups with a special emphasis on Pontryagin-van Kam- pen's duality theorem for locally ...
  10. [10]
    [PDF] Topological Groups - MDPI
    Since each discrete topological group is locally compact and balanced, Theorem 40 implies ... groups of locally finite buildings are totally disconnected, locally ...
  11. [11]
    [PDF] topology.pdf - Stacks Project
    Sep 28, 2025 · is Hausdorff, quasi-compact, and totally disconnected. By Lemma 14 ... that G → G/N is a surjection to a finite discrete topological group.
  12. [12]
    [PDF] Lie groups - IME-USP
    (f) Any finite or countable group endowed with the discrete topology becomes a 0-dimensional Lie group. Such examples are called discrete Lie groups. (g) We ...
  13. [13]
    [PDF] Bohr Compactifications - OSU Math
    Theorem. Let G be an LCA group. If G is compact, then bG is discrete in the compact-open topology. Similarly, if G is discrete, then bG is compact.
  14. [14]
    [PDF] Bohr Compactifications of Discrete Structures
    Mar 22, 1999 · The Bohr compactification and the Bohr topology are well-known for groups, but they can easily be generalized to arbitrary structures.
  15. [15]
    [PDF] 5.1 Lattices in locally compact groups
    Feb 20, 2024 · Let Γ be a discrete subgroup of a second countable locally compact group G. Then Γ is countable and G contains a measurable set of unique Γ- ...
  16. [16]
    [PDF] Math 100A Week 7: Discrete subgroups of isometries
    Nov 15, 2024 · Definition. A subgroup Λ of the additive group Rn is said to be discrete if there is ϵ > 0 so that λ ∈ Λ and λ ̸= 0 implies |λ| ≥ ϵ. Lemma 2.
  17. [17]
    [1312.7067] Subgroups of the additive group of real line - arXiv
    Dec 26, 2013 · This property of subgroups of the additive group of reals is special and well known (see Abels and Monoussos [4]).
  18. [18]
    [PDF] Frieze Groups and Crystallographic Groups
    The seven frieze groups catalog all symmetry groups that leave a design invariant under all multiples of just one translation. However, there are 17 additional ...
  19. [19]
    [PDF] classification of the 17 wallpaper groups
    These 17 classes, called the wallpaper groups, provide an interesting way to apply basic group theory to both geometry and art.Missing: discrete periodic tilings
  20. [20]
    [PDF] On Three-Dimensional Space Groups - arXiv
    There are 219 three-dimensional crystallographic space groups (or 230 if we distinguish between mirror images). They were independently enumerated in the ...<|separator|>
  21. [21]
    [PDF] Pseudomodular surfaces - UCSB Math
    Sep 5, 2001 · A Fuchsian group is a discrete subgroup of PSL(2, R). As such it acts discontinuously on H2 (the upper half plane model of the hyperbolic ...
  22. [22]
    [PDF] Lie groups and Lie algebras (Winter 2024)
    Let us focus, in particular, on the groups SO(3),SU(2),SL(2,R), and their topology. The Lie group SO(3) consists of rotations in 3-dimensional space. ... SO(3) is ...
  23. [23]
    [PDF] Fuchsian Groups and Fundamental Regions - UChicago Math
    Fuchsian Groups are discrete subgroups of PSL2(R), the group of orientation preserving isometries of the hyperbolic plane, and are important in hyperbolic ...
  24. [24]
    Théorie des groupes fuchsiens : Poincaré, Henri, 1854-1912
    Feb 11, 2010 · by: Poincaré, Henri, 1854-1912. Publication date: 1882. Topics: Automorphic functions. Publisher: Uppsala : Almqvist & Wiksells.
  25. [25]
    [PDF] Fuchsian Schottky groups are classical Schottky groups 1 Introduction
    But most groups on the boundary of. Schottky space have an empty ordinary set, so Schottky space strictly contains classical Schottky space. However, this ...Missing: source | Show results with:source
  26. [26]
    [PDF] A brief survey of the deformation theory of Kleinian groups 1 ... - arXiv
    A Kleinian group is a discrete subgroup of PSL2(C) = SL2(C)/{±I}, which we view as acting both on the Riemann sphere C by Möbius transformations and on real ...
  27. [27]
    Three dimensional manifolds, Kleinian groups and hyperbolic ...
    A group whose limit set is a circle is called the Fuchsian group (because of. Poincaré's modesty), and groups whose limit sets are Jordan curves are called.
  28. [28]
    [PDF] A conjecture on the Hall topology for the free group - l'IRIF
    Abstract. The Hall topology for the free group is the coarsest topology such that every group morphism from the free group onto a finite discrete group.<|control11|><|separator|>
  29. [29]
    [PDF] INVERSE LIMITS AND PROFINITE GROUPS - OSU Math
    Then since the discrete topology on a finite group is compact, Hausdorff, and totally disconnected, we have by Lemma 2.5 that G is compact and Hausdorff.
  30. [30]
    [PDF] dynamical systems on the circle - UChicago Math
    Theorem 3.1. Let rα : S1 → S1 be the rotation by α ∈ (0, 1). If α is irrational, the orbit of any point in S1 under rα is dense in S1.
  31. [31]
    [PDF] LATTICES IN LIE GROUPS - LSU Math
    A lattice Γ ⊂ G is an irreducible lattice if Γ is not reducible. Page 15. LATTICES IN LIE GROUPS. 15. Definition 7. If G ⊂ ...
  32. [32]
    [PDF] Lie groups, algebraic groups and lattices - arXiv
    Sep 30, 2012 · The Lie algebra of a Lie group G is defined by. L(G) = {X ∈ Md(R) ... A discrete subgroup Γ of a Lie group G is called a lattice if. µ(Γ ...<|separator|>
  33. [33]
    [PDF] DEFICIENCIES OF LATTICE SUBGROUPS OF LIE GROUPS
    If G is a connected Lie group, then a lattice in G is a discrete subgroup Γ such that G/Γ has finite volume. It is uniform if G/Γ is compact. Lubotzky proved.
  34. [34]
    Fundamental Domains for Lattices in (R-)rank 1 Semisimple Lie ...
    This paper describes a fundamental domain for a lattice in a rank one semisimple Lie group over real numbers, where a lattice has finite invariant volume.<|separator|>
  35. [35]
    [PDF] An introduction to Selberg's trace formula
    is generated by a primitive hyperbolic element P. Thus, if P is a hyper- bolic element, P = p for some integer k, and P, is a unique primitive hyperbolic ...
  36. [36]
    [PDF] Selberg's Trace Formula: An Introduction - University of Bristol
    The aim of this short lecture course is to develop Selberg's trace formula for a compact hyperbolic surface M, and discuss some of its applications. The.
  37. [37]
    None
    Below is a merged summary of the information from all provided segments about arithmetic groups, based on the arXiv:math/0106063 document. To retain all details efficiently, I will use a structured table format in CSV style, followed by a concise narrative summary that integrates the key points. The table captures the core concepts, definitions, examples, and references across all segments, while the narrative provides a cohesive overview.
  38. [38]
    [PDF] jonathan pfaff and jean raimbault - HAL
    associated arithmetic lattices are (up to commensurability) the so-called Bianchi groups. ΓD = SL2(OD) where OD is the ring of integers in Q(. √. −D). Note ...
  39. [39]
    [PDF] The congruence subgroup problem - Institute for Mathematics
    Evidently C(S, G) provides a measure of the failure of the family of S-arithmetic groups coinciding with the family of S-congruence subgroups. From the ...<|control11|><|separator|>
  40. [40]
    978-3-662-06431-3.pdf
    There are precisely 230 of these, the 230 space groups. They determine the types and positions of symmetry elements that are possible for a crystal structure.
  41. [41]
    3. Crystal Structure
    Combining lattic type, screw axis, and glide plane symmetries with the point group symmetries gives 230 different possi- ble symmetries for the arrangement of ...
  42. [42]
    [PDF] Introduction to Orbifolds - arXiv
    Notice that the same orbifold can appear as a quotient by a discrete group and also as a quotient by a finite group, such as the pillow case in Example 1.1.8.
  43. [43]
    [PDF] William P. Thurston The Geometry and Topology of Three-Manifolds
    Recall that a Fuchsian group (of type I) is a Kleinian group Γ whose limit set. LΓ is a geometric circle. Examples are the fundamental groups of closed, ...
  44. [44]
    [PDF] Kleinian groups and 3-Manifolds - UChicago Math
    Feb 28, 2014 · A Kleinian group is a finitely generated and discrete group of conformal symmetries of the sphere, where.
  45. [45]
    [PDF] Connective Bieberbach Groups - Purdue Math
    Let G be a Bieberbach group and let α ∈ Aut(G). Then the semidirect product G oα Z is a Bieberbach group if and only if there exists m ≥ 1 such that αm ∈ Inn(G) ...
  46. [46]
    On Bieberbach's Analysis of Discrete Euclidean Groups - jstor
    A euclidean group is discrete if and only if it is a subdirect product of a point group and a discrete space group such that the group of all elements of the ...
  47. [47]
    [PDF] Crystallography of Quasicrystals
    What can we know about a quasicrystal structure and what do we want to know, why, and what for, this is the guideline. The part “Structures” presents examples ...
  48. [48]
    Tilings: mathematical models for quasicrystals | Aperiodic Crystals
    This chapter discusses tilings as mathematical models for quasicrystals. In a first approximation quasicrystals may be described as being space filling with ...
  49. [49]
    Quasicrystals—The impact of N.G. de Bruijn - ScienceDirect
    Nov 15, 2013 · In [28] de Bruijn made a study of the inflation-deflation rules of the Penrose tilings providing details left out by Gardner [33]. In several of ...
  50. [50]
    Automorphic forms on SL2(R), by A. Borel, Cambridge Tracts in ...
    The rich interaction of themes is already apparent in the simplest modular forms, the Eisenstein series on congruence subgroups Γ of SL2(Z). For Γ = SL2(Z), the.
  51. [51]
    Modular Forms, a Computational Approach
    It introduces congru- ence subgroups of SL2(Z) and modular forms as functions on the complex upper half plane. We discuss q-expansions, which provide an ...<|separator|>
  52. [52]
    [PDF] Arithmetic Geometry and Analysis on Foliated Spaces - arXiv
    Feb 1, 2008 · ζ-functions of curves over finite fields this suggests that the Dedekind zeta function should be cohomological in nature. The rest of this ...
  53. [53]
    [PDF] arXiv:math/0404394v4 [math.NT] 4 May 2005
    These L-functions are associated to GL(n,Q)\GL(n,AQ), and they are more precisely denoted L(s, π, ρ) in which the. Langlands L-group LG = GL(N,C) and ρ : LG → ...
  54. [54]
  55. [55]
    [PDF] arXiv:2005.07919v1 [math.NT] 16 May 2020
    May 16, 2020 · Étale cohomology and associated Galois representations. Grothendieck defined étale cohomology as an algebraic replacement of singular cohomology ...
  56. [56]
    Introduction to the Langlands program, by J. Bernstein and S ...
    Jan 20, 2004 · In the arithmetic theory of automorphic forms one can begin with a group G as above of the form G(R), where G is a reductive linear ...
  57. [57]
    [PDF] arXiv:1202.2110v2 [math.RT] 6 Nov 2014
    Nov 6, 2014 · Abstract. The Langlands Program relates Galois representations and auto- morphic representations of reductive algebraic groups.
  58. [58]
    [PDF] arXiv:2001.06855v2 [math.AG] 8 Feb 2025
    Feb 8, 2025 · The Shafarevich conjecture for abelian varieties, proved by Faltings and Zarhin (see. [FWG+92, VI §1, Theorem 2]) asserts the finiteness of ...