Fact-checked by Grok 2 weeks ago

G -module

In abstract algebra, a G-module (or G-module over the integers) is an abelian group M equipped with a left action of a group G on M by group endomorphisms, meaning there is a homomorphism G → Aut(M) where Aut(M) denotes the automorphism group of M as an abelian group. Equivalently, a G-module is a module over the integral group ring ℤ[G], where elements of ℤ[G] act on M via the ring's multiplication extended from the G-action. This structure captures compatible group actions on additive groups and forms the foundation for studying symmetries in algebraic settings. In the context of representation theory, the notion generalizes to R-G-modules for a commutative ring R, where M is an R-module and G acts by R-linear endomorphisms; when R is a field k, this yields a k[G]-module, or vector space over k with a linear G-action, often simply called a representation of G over k. Key examples include the regular representation, where M = ℤ[G] with left multiplication by group elements, and the trivial representation, where g · m = m for all gG and mM. Subrepresentations correspond to G-invariant subgroups or subspaces, and irreducible G-modules—those with no nontrivial submodules—are fundamental for decomposing general modules into direct sums via Maschke's theorem when G is finite and k has characteristic not dividing |G|. G-modules play a central role in group cohomology and homology, where the cohomology groups Hn(G, M) measure extensions and obstructions related to the G-action on M, with applications in , , and the study of group extensions. Characters, defined as the trace of the action maps, provide invariants for classifying representations up to , particularly for finite groups. The category of G-modules admits tensor products, direct sums, and Hom functors that respect the action, enabling the development of tailored to group symmetries.

Fundamentals

Definition

A -module, also known as a module over the group G, is an M equipped with a left \rho: G \times M \to M that is compatible with the abelian group structure of M [12]. Specifically, for all g, h \in G and m, n \in M, the action satisfies \rho(g, m + n) = \rho(g, m) + \rho(g, n) and \rho(g, \rho(h, m)) = \rho(gh, m), with the of G acting as the identity map on M [12]. This action turns M into a in the sense that elements of G act as group automorphisms on M [2]. Equivalently, the action defines a group homomorphism \rho: G \to \Aut(M), where \Aut(M) denotes the of the M . This perspective emphasizes that the group action is a of G by automorphisms preserving the additive structure of M . Such G-modules are in one-to-one correspondence with modules over the \mathbb{Z}G, providing an algebraic framework for studying s . A of G-modules between two G-modules M and N is a f: M \to N that is G-equivariant, meaning f(\rho(g, m)) = \rho(g, f(m)) for all g \in G and m \in M . These preserve the group action and form the arrows in the of G-modules, often denoted \Mod_G or _G\Mod . The of G-modules is an , where kernels, cokernels, and exact sequences are defined in the standard way for module categories . This structure allows for the application of techniques to G-modules, treating them as objects in a rich categorical framework .

Relation to Group Rings

The group ring \mathbb{Z}G consists of all formal sums \sum_{g \in G} a_g g, where a_g \in \mathbb{Z} and only finitely many a_g are nonzero, equipped with addition componentwise and multiplication defined by extending the group operation linearly: \left( \sum a_g g \right) \left( \sum b_h h \right) = \sum_{g,h \in G} a_g b_h (gh). This structure makes \mathbb{Z}G a ring with unity $1 = e, where e is the identity element of G. Every G-module M (an abelian group with a compatible G-action) corresponds bijectively to a left \mathbb{Z}G-module structure on M, given by the action \left( \sum_{g \in G} a_g g \right) \cdot m = \sum_{g \in G} a_g (g \cdot m) for all m \in M. This correspondence is an equivalence of categories between the category of G-modules and the category of left \mathbb{Z}G-modules. This equivalence is functorial in nature. The forgetful functor from the category of left \mathbb{Z}G-modules to the category of s, which forgets the \mathbb{Z}G-action, admits a left adjoint given by : for an A, the induced module is \mathbb{Z}G \otimes_{\mathbb{Z}} A, and a right adjoint given by : \mathrm{Hom}_{\mathbb{Z}}(\mathbb{Z}G, A). For a general R with unity, the construction generalizes by replacing \mathbb{Z} with R: the RG consists of formal R-linear combinations \sum_{g \in G} a_g g with finite , and G-modules over R (i.e., R-modules with compatible G-) are equivalent to left RG-modules via the analogous .

Properties

Submodules and Quotients

In the category of G-modules, a N of a G-module M is an abelian of M that is closed under the G-, meaning g \cdot n \in N for all g \in G and n \in N. This ensures that N itself inherits a well-defined G-module structure from M. Given a submodule N \subseteq M, the quotient M/N is defined as the set of cosets \{m + N \mid m \in M\} with the induced operation. The G-action on M/N is given by g \cdot (m + N) = (g \cdot m) + N for g \in G and m \in M, which is well-defined precisely because N is G-. This makes M/N a [G](/page/G)-module in a natural way. A of G-modules \cdots \to M_{i-1} \xrightarrow{f_{i-1}} M_i \xrightarrow{f_i} M_{i+1} \to \cdots with G-equivariant homomorphisms f_i: M_i \to M_{i+1} (i.e., f_i(g \cdot m) = g \cdot f_i(m) for all g \in G, m \in M_i) is if the underlying sequence of abelian groups is exact, meaning \operatorname{im} f_{i-1} = \ker f_i for each i. In particular, a short $0 \to N \to M \to Q \to 0consists ofG-equivariant maps where the first is injective, the second surjective, and \operatorname{im}(N \to M) = \ker(M \to Q)$. Direct sums and products of G-modules are formed componentwise with respect to the diagonal G-action: for G-modules M_i (i \in I), the direct sum \bigoplus_{i \in I} M_i has elements finite sums \sum m_i with m_i \in M_i and g \cdot (\sum m_i) = \sum (g \cdot m_i), while the \prod_{i \in I} M_i has elements (m_i)_{i \in I} with g \cdot (m_i) = (g \cdot m_i)_{i \in I}. These constructions preserve the G-module structure exactly as in the underlying of abelian groups.

Morphisms and Categories

The category of G-modules, denoted \mathrm{Mod}_G or G-\mathrm{Mod}, consists of all abelian groups equipped with a left action of the group G as objects, with morphisms given by G-equivariant group homomorphisms, i.e., additive maps f: M \to N satisfying f(g \cdot m) = g \cdot f(m) for all g \in G and m \in M. This category is equivalent to the category of left modules over the group ring \mathbb{Z}G, and as such, it inherits the structure of an abelian category: every monomorphism is the kernel of its cokernel, every epimorphism is the cokernel of its kernel, and kernels and cokernels exist and coincide with their images. In \mathrm{Mod}_G, the kernel of a morphism f: M \to N is the usual kernel \{m \in M \mid f(m) = 0\} as an abelian group, which is automatically a G-submodule since equivariance implies g \cdot \ker f \subseteq \ker f; similarly, cokernels are quotients by G-submodule images. The F: \mathrm{Mod}_G \to \mathrm{Ab}, which sends a G-module to its underlying and a G- to the corresponding , is : a of G-modules is if and only if the underlying of is , as the G- does not affect the computation of kernels and cokernels in the underlying category. The Hom-sets \mathrm{Hom}_G(M, N) are the of all G- from M to N, and for fixed M, the covariant \mathrm{Hom}_G(M, -): \mathrm{Mod}_G \to \mathrm{Ab} is left , preserving finite limits such as kernels of $0 \to N' \to N \to N'', yielding $0 \to \mathrm{Hom}_G(M, N') \to \mathrm{Hom}_G(M, N) \to \mathrm{Hom}_G(M, N''). Dually, \mathrm{Hom}_G(-, N) is contravariant and left . The category \mathrm{Mod}_G has enough projective objects, meaning that for every G-module M, there exists a surjection P \twoheadrightarrow M from a projective G-module P; the free \mathbb{Z}G-modules \mathbb{Z}G \otimes_{\mathbb{Z}} A \cong \bigoplus_A \mathbb{Z}G for abelian groups A are projective, and every projective G-module is a direct summand of a free one, allowing projective resolutions for computing derived functors. Similarly, \mathrm{Mod}_G has enough injective objects, so every G-module embeds into an injective one, facilitating injective resolutions; injective G-modules can be constructed as direct sums or products of indecomposable injectives over \mathbb{Z}G. These properties ensure that \mathrm{Mod}_G supports the full machinery of , including Ext and Tor functors.

Examples and Constructions

Trivial and Regular Modules

The trivial G-module is defined as any abelian group M equipped with a G-action such that g \cdot m = m for all g \in G and m \in M. This action renders every element of G acting as the identity map on M, making it the simplest possible G-module structure. A standard example is the integers \mathbb{Z} viewed as a trivial G-module, where the action ignores group elements entirely. The regular G-module is the group ring \mathbb{Z}G endowed with the left multiplication action, where g \cdot \left( \sum_{h \in G} n_h h \right) = \sum_{h \in G} n_h (g h) for g \in G and coefficients n_h \in \mathbb{Z}. This module captures the full structure of G acting on its formal linear combinations. For a finite group G, when considered over the complex numbers \mathbb{C}, the regular module \mathbb{C}G decomposes as a direct sum \bigoplus_{\rho} (\dim \rho) \cdot \rho, where the sum runs over all irreducible representations \rho of G. For a G and H \leq G, the on the G/H is the \mathbb{Z}[G/H] generated by the left cosets, with G acting by permutation: g \cdot (kH) = (g k) H for g \in G and coset representative k \in G. This yields a G-module isomorphic to the induced module from the trivial H-module \mathbb{Z}, providing a concrete . In the context of vector spaces, an example arises with the O(n) acting on \mathbb{R}^n via , where g \cdot v = g v for g \in O(n) and v \in \mathbb{R}^n. This defines a O(n)- structure that preserves the standard inner product, illustrating a faithful linear action.

Induced and Coinduced Modules

In of groups, given a group G and a H \leq G, important constructions allow one to build G-s from H-s. The induced module provides a way to extend an H-module to a G-module via over the . Specifically, for an H-module N, the induced module is defined as \operatorname{Ind}_H^G N = \mathbb{Z}G \otimes_{\mathbb{Z}H} N, where the G-action is given by g \cdot (x \otimes n) = gx \otimes n for g, x \in G and n \in N. This construction leverages the bimodule structure of \mathbb{Z}G over \mathbb{Z}G and \mathbb{Z}H. Dually, the coinduced module extends H-modules using Hom spaces. For an H-module N, the coinduced module is \operatorname{Coind}_H^G N = \operatorname{Hom}_{\mathbb{Z}H}(\mathbb{Z}G, N), equipped with the G-action (g \cdot f)(x) = f(x g^{-1}) for g, x \in G and f \in \operatorname{Hom}_{\mathbb{Z}H}(\mathbb{Z}G, N). This action ensures compatibility with the H-module structure on N, reflecting the right G-module structure on \mathbb{Z}G. These participate in adjunctions with the restriction functor \operatorname{Res}_H^G, which forgets the [G](/page/G)-action to yield an [H](/page/H+)-module. The functor is left to restriction: \operatorname{Hom}_G(\operatorname{Ind}_H^G N, M) \cong \operatorname{Hom}_H(N, \operatorname{Res}_H^G M) for any [G](/page/G)-module M and [H](/page/H+)-module N. Symmetrically, is right to restriction: \operatorname{Hom}_G(M, \operatorname{Coind}_H^G N) \cong \operatorname{Hom}_H(\operatorname{Res}_H^G M, N). These isomorphisms, known as Frobenius adjunctions or in certain contexts, underpin many applications in . When the index [G : H] is finite, the induced and coinduced modules are naturally isomorphic as G-modules. This isomorphism arises from the finite sum over coset representatives and preserves the module structures, facilitating computations in both directions.

Structure and Homological Algebra

Invariants and Coinvariants

In a G-module M, the submodule of invariants, denoted M^G, consists of those elements fixed by the entire group action: M^G = \{ m \in M \mid g \cdot m = m \ \forall \, g \in G \}. This forms a submodule of M on which G acts trivially via the identity map. The module of coinvariants, denoted M_G, is the quotient of M by the submodule generated by all elements of the form g \cdot m - m for g \in G and m \in M: M_G = M / \langle g \cdot m - m \mid g \in G, \, m \in M \rangle. This construction yields the largest quotient module of M on which G acts trivially, effectively identifying elements within each G-orbit. When is finite, a key connection between invariants and coinvariants arises via the norm map N: [M^G](/page/M&G) \to M_G, defined by N(m) = \sum_{g \in [G](/page/G)} g \cdot m modulo the relations generating the coinvariants submodule (which simplifies to |[G](/page/G)| \cdot m in the quotient since m is fixed). This map plays a central role in explicit computations involving these constructions, such as in group and the study of fixed-point subspaces.

Resolutions and Ext Functors

In , a projective of a left \mathbb{Z}[G](/page/G)-module M is an \cdots \to P_1 \to P_0 \to M \to 0, where each P_i is a projective \mathbb{Z}[G](/page/G)-module, typically since modules over group rings are projective. These resolutions provide a framework for computing derived functors associated to M, enabling the study of extensions and in the category of G-modules. Projective \mathbb{Z}[G](/page/G)-modules, such as modules of rank n generated by basis elements with diagonal G-action, serve as building blocks for such resolutions. The Ext functors \operatorname{Ext}^n_{\mathbb{Z}G}(M, N) for G-modules M and N are the right derived functors of the Hom functor \operatorname{Hom}_{\mathbb{Z}G}(-, -), measuring the n-fold extensions of M by N. They are computed by applying \operatorname{Hom}_{\mathbb{Z}G}(-, N) to a projective resolution P_\bullet \to M of M, yielding \operatorname{Ext}^n_{\mathbb{Z}G}(M, N) \cong H^n \bigl( \operatorname{Hom}_{\mathbb{Z}G}(P_\bullet, N) \bigr), where the cohomology is taken after deleting the identity map P_0 \to M. This construction captures obstructions to lifting homomorphisms and is central to understanding module extensions in group representations. Dually, the Tor functors \operatorname{Tor}_n^{\mathbb{Z}G}(M, N) are the left derived functors of the M \otimes_{\mathbb{Z}G} N, relating to the of over the . Given a projective P_\bullet \to M of M, they are given by \operatorname{Tor}_n^{\mathbb{Z}G}(M, N) \cong H_n \bigl( P_\bullet \otimes_{\mathbb{Z}G} N \bigr), with computed after deleting the augmentation P_0 \to M. These functors quantify the failure of exactness in and arise naturally in the study of bilinear forms under G-actions. For a finite group G, the standard bar resolution provides an explicit projective resolution of the trivial \mathbb{Z}G-module \mathbb{Z}, where G acts trivially. It is the complex \cdots \to P_1 \to P_0 \to \mathbb{Z} \to 0 with P_n the free \mathbb{Z}G-module on basis elements [g_0 | \cdots | g_n] for g_i \in G, and differential d_n [g_0 | \cdots | g_n] = \sum_{i=0}^n (-1)^i [g_0 | \cdots | \hat{g_i} | \cdots | g_n] + \sum_{i=0}^{n-1} (-1)^i [g_0 | \cdots | g_i g_{i+1} | \cdots | g_n], where the hat denotes omission; this resolution is exact and functorial, facilitating computations of Ext and Tor groups involving the trivial module.

Applications

Group Cohomology

Group cohomology is a fundamental tool in homological algebra that studies the structure of groups through their actions on abelian groups, utilizing G-modules as coefficients. For a discrete group G and a left ℤG-module M, the nth cohomology group is defined as
H^n(G, M) = \Ext^n_{\mathbb{Z}G}(\mathbb{Z}, M),
where ℤ denotes the trivial ℤG-module with G acting via the identity map. This definition arises from applying the right derived functors of the Hom functor to a projective resolution of the trivial module ℤ over the group ring ℤG.
In low dimensions, these groups admit concrete interpretations. The zeroth cohomology group is the subgroup of invariants,
H^0(G, M) = M^G = \{ m \in M \mid g \cdot m = m \ \forall g \in G \},
consisting of elements fixed by the entire group action. The first cohomology group classifies crossed homomorphisms from G to M modulo principal ones: a crossed homomorphism f: G → M satisfies f(gh) = f(g) + g · f(h) for all g, h ∈ G, and principal crossed homomorphisms are those of the form g ↦ g · m - m for some fixed m ∈ M. Thus,
H^1(G, M) = Z^1(G, M) / B^1(G, M),
where Z^1 and B^1 denote the groups of crossed and principal crossed homomorphisms, respectively. This group relates to conjugacy classes of group extensions, as elements of H^1(G, M) parametrize automorphisms of extensions classified by H^2(G, M), up to inner automorphisms induced by the module.
A key technical tool is Shapiro's lemma, which relates of to that of the full group via coinduced . For a H ≤ G and an ℤH- N, the coinduced is
\Coind_H^G N = \Hom_{\mathbb{Z}H}(\mathbb{Z}G, N),
with G-action given by (g · f)(x) = f(x g) for f ∈ \Coind_H^G N and x ∈ ℤG. Shapiro's lemma states that
H^*(G, \Coind_H^G N) \cong H^*(H, N)
naturally in both G and N, providing a powerful method to compute by reducing to .

Representation Theory

In , a left over the group algebra KG, where K is a of characteristic zero and G is finite, is equivalent to a \rho: G \to \mathrm{GL}(V) of G on a finite-dimensional V over K. The structure arises from the action g \cdot v = \rho(g) v for g \in G and v \in V, making V a KG- via the identification KG \cong \bigoplus_{g \in G} K g. This equivalence holds because the group algebra encodes the linear transformations induced by group elements, and in characteristic zero, such representations are completely reducible. A fundamental result in this context is Maschke's theorem, which asserts that if the characteristic of K does not divide |G|, then every finite-dimensional KG- is , decomposing uniquely (up to ) as a of irreducible submodules. This implies that any submodule has a complementary invariant submodule, ensuring that representations are direct sums of irreducibles without extensions. The theorem relies on the invertibility of |G| in K, allowing averaging projectors over the group to split exact sequences. For example, over \mathbb{C}, all finite-dimensional representations of finite groups are semisimple. Characters provide a key tool for classifying these representations. The character \chi_\rho of a representation \rho is the class function \chi_\rho(g) = \mathrm{tr}(\rho(g)), which is constant on conjugacy classes and determines the representation up to when K = \mathbb{C}. The inner product of characters \chi, \psi is defined as \langle \chi, \psi \rangle = \frac{1}{|G|} \sum_{g \in G} \chi(g) \overline{\psi(g)}, and for irreducible characters, the orthogonality relations hold: \langle \chi_i, \chi_j \rangle = \delta_{ij}, with the sum of squares of irreducible degrees equaling |G|. Column orthogonality further states that for conjugacy classes C_k, \sum_i \chi_i(g_k) \overline{\chi_i(g_l)} = |G| \delta_{kl} / |C_k|. These relations form the basis for character tables, which classify all irreducibles and compute decomposition multiplicities. Induced representations extend characters from subgroups. For a subgroup H \leq G and a representation V of H over K, the induced representation \mathrm{Ind}_H^G V is the [KG](/page/KG)-module with underlying space K[G] \otimes_{K[H]} V, equipped with the diagonal . Its character is \chi_{\mathrm{Ind}_H^G V}(g) = \frac{1}{|H|} \sum_{t \in G, t^{-1} g t \in H} \chi_V(t^{-1} g t). Frobenius reciprocity links induction and restriction: for representations V of H and W of G, the multiplicity \langle \mathrm{Ind}_H^G V, W \rangle_G = \langle V, \mathrm{Res}_H^G W \rangle_H, where inner products count homomorphisms or trace averages. This adjunction facilitates decomposition of induced irreducibles and of branching rules.

Generalizations

Topological G-Modules

A topological G-module is defined as a topological abelian group M together with a continuous action of a topological group G on M, meaning the map G \times M \to M given by (g, m) \mapsto g \cdot m is continuous and satisfies the axioms g \cdot (m_1 + m_2) = g \cdot m_1 + g \cdot m_2, (g_1 g_2) \cdot m = g_1 \cdot (g_2 \cdot m), and e \cdot m = m for all g, g_1, g_2 \in G, m, m_1, m_2 \in M, and identity e \in G. This structure generalizes the algebraic notion of a G-module to settings where continuity ensures compatibility with the topologies on G and M. Morphisms between topological G-modules are continuous group homomorphisms that are G-equivariant, i.e., f(g \cdot m) = g \cdot f(m) for all g \in G, m \in M. The category of topological G-modules, denoted \mathbf{T}\mathbf{-}\mathbf{Mod}, has these objects and morphisms; it is abelian when the topologies are complete and metrizable, allowing for kernels, cokernels, and exact sequences in the topological sense. In such cases, the category supports , including Ext functors computed via projective or injective resolutions of topological G-modules. Prominent examples include actions of Lie groups on Banach spaces, where G acts via continuous linear operators preserving the norm topology on M, as in the study of smooth representations. Another key instance arises in , where profinite Galois groups G = \mathrm{Gal}(\overline{K}/K) act continuously on discrete G-modules such as the of units or ideals in the of extensions, enabling the computation of via continuous cochains. For compact topological groups G, the Peter-Weyl theorem provides a foundational decomposition: every unitary of G on a M (a complete , hence a topological under addition) is a of finite-dimensional irreducible representations, with matrix coefficients spanning a dense of L^2(G). This highlights how topological G-modules over compact G capture on G through orthogonal decompositions.

Modules over Semigroups

In the context of semigroup actions, an S-module for a S is defined as an (M, +) equipped with a S × M → M, denoted (s, m) ↦ s · m, satisfying s · (m_1 + m_2) = s · m_1 + s · m_2, (s t) · m = s · (t · m), and s · 0 = 0 for all s, t ∈ S and m_1, m_2 ∈ M. This structure generalizes the notion of a G-module by providing a distributive without requiring invertibility of elements in S, thus omitting conditions like g^{-1} · (g · m) = m. The absence of inverses leads to key differences in homological properties and categorical structure compared to group modules. When S is a monoid (a with an ), the of S-modules is equivalent to the of left modules over the monoid ring ℤ*, where ℤ* is the on S equipped with (∑ a_s s)(∑ b_t t) = ∑_{s,t ∈ S} (a_s b_t)(s t). In this setting, idempotents e ∈ S (satisfying e^2 = e) play a role analogous to units, acting as local identities on modules where e · m = m for elements in the image of e, facilitating decompositions and projections in the module . This equivalence mirrors the relationship between G-modules and modules over the G, as noted in discussions of algebraic representations. A concrete example is the action of the additive ℕ (natural numbers including 0) on the ℤ, where n · m = n m ( in ℤ, or equivalently, adding m to itself n times). This satisfies the required properties, as distributivity holds over addition in ℤ and the semigroup operation composes via repeated scaling. Another illustrative construction is the Reynolds for averaging over a semigroup , which generalizes the group averaging to compact abelian semigroups S acting on function spaces like C(S); for a positive T generating a semigroup, the Reynolds R f = ∫ e^{-t} T_t f , dt projects onto fixed points while preserving positivity and contractivity. In general, the category S-Mod of S-modules lacks sufficient projective objects, unlike the category of G-modules for finite groups G, where free modules provide projective resolutions for all objects; this deficiency arises particularly for semigroups without identity, complicating homological computations.

References

  1. [1]
    [PDF] Group Homology and Cohomology
    Let G be a group. A (left) G-module is an abelian group A on which G acts by additive maps on the left; if g e G and a e A, we write ga for the action of.
  2. [2]
    Section 59.57 (0A2H): Group cohomology—The Stacks project
    Feb 14, 2018 · A G-module, sometimes called a discrete G-module, is an abelian group M endowed with a left action a : G \times M \to M by group homomorphisms ...
  3. [3]
    Modules and Group Cohomology - William Stein
    Modules and Group Cohomology. Let $ A$ be a $ G$ module. This means that $ A$ is an abelian group equipped with a left action of $ G$ ...
  4. [4]
    [PDF] 1 Basic definitions
    1 Basic definitions. Representations. A representation of a group G over a field k is a k-vector space V together with an action of G on V by linear maps.
  5. [5]
    [PDF] Group Representations and Character Theory - UChicago Math
    Aug 26, 2011 · We call a representation ρ the trivial representation if ρ = 1. Definition 3.5. An FG-module is a vector space V over a field F together with.
  6. [6]
    [PDF] Modules and Wedderburn Theory - Alexander Rhys Duncan
    Apr 13, 2023 · The structure of a representation of G over the field k is equivalent to the structure of a left kG- module. Finite-dimensional representations ...
  7. [7]
    [PDF] An Introduction to the Cohomology of Groups
    Specifying a ZG-module M is the same thing as specifying an abelian group M on which G acts, i.e. there is a homomorphism G → Aut(M). We denote by Z the ZG- ...<|separator|>
  8. [8]
    [PDF] math 101b: algebra ii part d: representations of groups - Brandeis
    The group ring KG. The main idea is that representations of a group G over a field K are “the same” as modules over the group ring KG = K[G]. Also, left ...
  9. [9]
    [PDF] Some module theory - Purdue Math
    A left (right) ZG module is the same thing as an abelian group with a left (right) action by G. A homomorphism of R-modules f : M ! N is an abelian group homo-.<|control11|><|separator|>
  10. [10]
    [PDF] 30 Semisimple Algebras: Introduction - Brandeis
    A G-submodule of a G-module V is a vector subspace. W ⊆ V with the property ... This is an isomorphism of G-modules by the definition of the action of G on V ∗.
  11. [11]
    [PDF] A Group Action on an R-Module and G-Module
    Jun 26, 2024 · Throughout this work, all rings are commutative, with unit only in specific places we will mention it; all modules are unitary, all group.
  12. [12]
    [PDF] 23 Tate cohomology
    Nov 27, 2017 · A G-module is an abelian group A equipped with a. G-action compatible with its group structure: g(a + b) = ga + gb for all g ∈ G, a, b ∈ A.1.
  13. [13]
    [PDF] AN INTRODUCTION TO HOMOLOGICAL ALGEBRA
    ... Exact Couples. ) Homology and Cohomology. Definitions and First Properties ... functor. The cohomology of sheaves, the Grothendieck spectral sequence, lo ...
  14. [14]
    [PDF] Introduction to representation theory by Pavel Etingof, Oleg Golberg ...
    Very roughly speaking, representation theory studies symmetry in linear spaces. It is a beautiful mathematical subject which has many.
  15. [15]
    Orthogonal Group -- from Wolfram MathWorld
    The orthogonal group O(n) is the group of n×n orthogonal matrices. These matrices form a group because they are closed under multiplication and taking inverses.<|control11|><|separator|>
  16. [16]
    [PDF] representation theory, lecture 0. basics - Yale Math
    It follows that, for a B-module N, the space A ⊗B N carries a natural structure of a left A-module (induced module). Also A is a B-A-bimodule. So. HomB(A, N) ...
  17. [17]
    [PDF] Notes for 'Representations of Finite Groups' - University of Oregon
    Aug 13, 2012 · 4 Induced representations ... In particular, the number of irreducible representations of a finite group equals the number of conjugacy classes.
  18. [18]
    Cohomology of Groups - Kenneth S. Brown - Google Books
    As a second year graduate textbook, Cohomology of Groups introduces students to cohomology theory (involving a rich interplay between algebra and topology)
  19. [19]
    [PDF] Cohomology of groups - Purdue Math
    B• is called the bar or standard resolution. The first name comes from the ... If G is a finite group of order N, then for any G-module M and n > 0, we ...
  20. [20]
    [PDF] Group Cohomology - Lecture Notes - Berkeley Math
    We are aiming for: Theorem 2.3. Let G be a group, M a ZG-module. Then there is a bijection: equivalence classes of extensions of G by M.
  21. [21]
    [PDF] Math 210B. Group cohomology and group extensions
    In particular, if H1(G, M) = 0 then all automor- phisms of E as a group extension are necessarily of the trivial type arising from conjugation by an element of ...
  22. [22]
    [PDF] Lectures on the Cohomology of Finite Groups
    (4) If Fp is given the trivial G–module structure, then H∗(G, Fp) is the ordinary mod p cohomology ring of BG, and so it has an action of the Steenrod algebra ...<|control11|><|separator|>
  23. [23]
    [PDF] A Course in Finite Group Representation Theory
    The representation theory of finite groups has a long history, going back to the 19th century and earlier. A milestone in the subject was the definition of ...
  24. [24]
    [PDF] Isaacs_Character_theory.pdf
    ' Preface. Character theory provides a powerful tool for proving theorems about finite groups. 1n fact.. there are some important results, such as Frobenius ...
  25. [25]
    [PDF] induced representations for finite groups - Math
    Frobenius ReCiProCitY. 1.1. Restriction functor. Let G be a a finite group. Let H be the a subgroup of G. Denote by Rep(G), resp. Rep(H), the categories of ...
  26. [26]
    [PDF] 30 Galois cohomology and the invariant map for local fields
    Feb 12, 2018 · Let G be a topological group. A topological G-module A is an abelian topological group on which G acts continuously. In other words, the map ...
  27. [27]
    [PDF] Algebraic Cohomology of Topological Groups Author(s)
    Let G be a topological group. A topological G-module is an abelian top- ological group A together with a continuous map G x A -4 A satisfying the usual ...
  28. [28]
    [PDF] GROUP AND GALOIS COHOMOLOGY Romyar Sharifi
    We turn briefly to an even more interesting example. DEFINITION 1.2.7. A group extension of G by a G-module A is a short exact sequence of groups. 0 → A.
  29. [29]
    [PDF] REPRESENTATION THEORY
    Representations. Let G be a topological group. Topological vector spaces are taken over a topo- logical field k (which we fix). We denote by End(V ), Aut(V ) ...
  30. [30]
    3.5 Profinite groups and infinite Galois theory - Kiran S. Kedlaya
    If G is profinite, by a G -module we mean a topological abelian group M with a continuous G -action . M × G → M . In particular, ...
  31. [31]
    [PDF] The Peter-Weyl Theorem for Compact Groups x1 Preliminaries.
    De nition 1: A (unitary) representation of G is a continuous homomorphism from G to the unitary group U(H ) on a (complex) Hilbert space H equipted with the ...
  32. [32]
  33. [33]
    (PDF) Semigroup Cohomology and Applications - ResearchGate
    ... is the same as for groups: Hn(S, A) = Extn. ZS(Z, A). Here Sis a semigroup, Aa left S-module (i. e. a left module over the integral. semigroup ring Z ...
  34. [34]
    Positive Reynolds operators and moment theory | Semigroup Forum
    In this paper, positive moment operators on the space C(S) are considered, where S is a compact abelian semigroup possessing a separating set of continuous.
  35. [35]
    Homological properties of modules over semigroup algebras
    The main result about the projectivity of c 0 ( S ) states that for a weakly cancellative inverse semigroup S, c 0 ( S ) is projective if and only if S is ...