Fact-checked by Grok 2 weeks ago

Group representation

In mathematics, a group representation is a homomorphism from an abstract group G to the general linear group \mathrm{GL}(V) of invertible linear transformations on a vector space V over a field k, allowing the group's structure to be analyzed through concrete linear actions. This framework, often over the complex numbers for finite groups, equates to a module over the group algebra k[G], where elements of G act linearly on V. Equivalence of representations occurs via conjugation by an invertible linear map, preserving the group's action up to similarity. Representation theory, the broader study encompassing these objects, originated in 1896 with Ferdinand Georg Frobenius's work on group characters, motivated by Richard Dedekind's inquiries into group determinants from multiplication tables. Key developments include Maschke's theorem, which asserts that representations of finite groups over fields whose characteristic does not divide the group order are semisimple (decomposable into direct sums of irreducibles), and the Artin-Wedderburn theorem describing the structure of semisimple algebras. Central concepts include irreducible representations (those with no nontrivial invariant subspaces), characters (traces of representation matrices, forming class functions constant on conjugacy classes), and operations like (extending representations from subgroups) and tensor products. The theory applies across mathematics and physics, illuminating symmetries in (e.g., via unitary representations), classifying finite simple groups through character tables, and connecting to , , and via tools like the Jordan-Hölder theorem on uniqueness. For infinite groups, such as groups, representations extend to continuous homomorphisms, underpinning and models. Modern advances leverage computational methods and , introduced by and in the 1940s, to unify representations of groups, algebras, and quivers.

Introduction

Historical overview

The origins of group representation theory can be traced to the early , when began studying groups in the context of solving polynomial equations, introducing early notions of group actions on sets in his 1812 memoir on substitutions. further advanced this in the by linking groups to extensions in his work on the solvability of equations by radicals, laying foundational ideas for understanding symmetries through group structures. By the late , made pivotal contributions, motivated by Richard Dedekind's 1896 query on the group determinant; in his 1896 papers, Frobenius developed the theory of group characters, computed the character table for the S_3, and established the orthogonality relations for characters, marking the birth of modern for finite groups. Key early 20th-century advancements built on Frobenius's framework. William Burnside's 1897 book Theory of Groups of Finite Order (revised in 1911) synthesized theory and applied early ideas to classify groups, while his 1904 on the solvability of groups of order p^a q^b demonstrated the power of . Issai , in his 1901 dissertation and subsequent papers from 1905 to 1911, introduced irreducibility criteria for s, proved on endomorphisms of irreducible s, and extended to integral s, solidifying the algebraic foundations. Heinrich Maschke complemented this in 1898 by proving that s of finite groups over the complex numbers are completely reducible (Maschke's ), enabling the decomposition into irreducibles. In the mid-20th century, the theory expanded to infinite and continuous groups. Hermann Weyl's 1925 book The Theory of Groups and and papers from 1925–1926 developed unitary representations of compact Lie groups, introducing the highest weight construction for semisimple Lie algebras, while his 1931 work The Classical Groups unified finite and continuous cases through . Emil Artin's 1927 contributions in utilized group characters to define Artin L-functions, bridging with number theory by generalizing Dirichlet characters to non-abelian Galois groups. In the , Claude Chevalley's 1941 lectures and 1946 book Theory of Lie Groups pioneered aspects of the study of algebraic groups over arbitrary fields, proving that semisimple algebraic groups have the same as their Lie algebras. Modern computational aspects emerged in the 1960s with the advent of systems, enabling algorithmic computation of tables and irreducible representations. Joachim Neubüser's 1960 paper initiated systematic computational , leading to programs for and group computations by the late 1960s, such as those used in classifying finite simple groups. A brief application in physics appeared with Eugene Wigner's 1931 book Group Theory and Its Application to the of Atomic Spectra, where representations classified atomic states under symmetry groups.
YearMilestoneContributorKey Publication/THEOREM
1812Early permutation group studiesCauchyMemoir on substitutions
1830sPermutation groups in GaloisWorks on solvability by radicals
1896Invention of group characters and orthogonalityFrobenius"Über Gruppencharaktere"
1898Complete reducibility over ℂMaschkeMaschke's theorem
1904Solvability via charactersBurnsideBurnside's p^a q^b theorem
1905–1911Irreducibility and SchurPapers on linear substitutions
1925Unitary representations of Lie groupsWeylTheory of Groups and
1927Characters in ArtinArtin L-functions
1941Algebraic groups and representationsChevalleyLectures on Lie groups
1960Computational initiationNeubüserFirst computational paper on groups

Motivations

Group representations provide a powerful for abstracting and analyzing group actions by linearizing symmetries into concrete operations. Rather than studying groups in isolation, representations embed them as subgroups of general linear groups acting on vector spaces, enabling the use of linear algebra to compute and understand group behaviors. This approach transforms potentially intractable problems about group elements and relations into manageable manipulations, such as finding eigenvalues or solving systems of equations. At their core, groups encode symmetries arising in , physics, and , and representations illuminate these by revealing invariant subspaces—subspaces preserved under the . These subspaces correspond to irreducible components of the , allowing the decomposition of complex actions into simpler building blocks that respect the underlying . Such decompositions highlight how group elements act consistently on geometric objects or algebraic structures, providing insight into the invariants that remain unchanged. Representation theory unifies diverse mathematical themes, notably through its connections to , as pioneered by in the 1890s. Hilbert's finiteness theorem demonstrated that the ring of polynomial invariants under a linearly reductive on a is finitely generated, laying groundwork for studying symmetries via representations. This links to the decomposition of tensor products of representations into direct sums of irreducibles, which captures how combined symmetries behave and facilitates the classification of group elements through traces of representation matrices—quantities that are invariant under similarity and thus serve as signatures of conjugacy classes. More broadly, representations enable the reduction of intricate problems to linear algebra, such as solving systems of equations that are invariant under group actions or computing groups via induced representations. By focusing on linear actions, this simplifies the analysis of group symmetries across disciplines, turning abstract algebraic questions into concrete computational tasks.

Core Definitions

Group homomorphisms to general linear groups

In the framework of , a of a group G on a V over a F is defined as a \rho: G \to \mathrm{GL}(V), where \mathrm{GL}(V) denotes the general linear group of all invertible linear endomorphisms of V. This homomorphism encodes how elements of G act linearly on V, preserving the group operation through composition of endomorphisms. The associated is then expressed as g \cdot v = \rho(g)(v) for all g \in G and v \in V, ensuring that the action respects both the vector space structure and the group multiplication: (gh) \cdot v = g \cdot (h \cdot v) and e \cdot v = v, where e is the identity in G. The of the representation, denoted \dim V, is commonly referred to as the of the representation, which quantifies its complexity and plays a central role in theorems. A particularly simple case is the trivial representation, where \rho(g) = \mathrm{Id}_V (the endomorphism) for every g \in G, meaning the action fixes every vector in V invariantly. Representations can also be classified by : a representation is faithful if the homomorphism \rho is injective, G as a of \mathrm{GL}(V) without , whereas non-faithful representations have a non-trivial kernel, effectively representing a . As an illustrative abstract example, consider the cyclic group C_n = \langle \sigma \mid \sigma^n = e \rangle. A representation \rho: C_n \to \mathrm{GL}(V) is fully determined by the image \rho(\sigma), which must be an invertible endomorphism of order dividing n, i.e., \rho(\sigma)^n = \mathrm{Id}_V, and extended by \rho(\sigma^k) = \rho(\sigma)^k for k = 0, \dots, n-1. This setup highlights how the homomorphism property constrains the possible actions without specifying a basis or matrix form. Representations of this type are often analyzed up to equivalence, where two are equivalent if there exists an invertible linear map intertwining their actions.

Vector space representations

A vector space representation of a group G assigns to each g \in G an invertible linear transformation \rho(g): [V](/page/V.) \to [V](/page/V.) on a finite-dimensional [V](/page/V.) over a F, typically \mathbb{C} or \mathbb{R}, such that the map \rho: G \to \mathrm{GL}(V) is a . This means \rho(g) preserves vector addition and scalar multiplication, and the assignment respects the group operation via \rho(gh) = \rho(g) \rho(h) for all g, h \in G. To obtain a concrete matrix form, select a basis \{e_1, \dots, e_n\} for V, where n = \dim_F V. The action of \rho(g) is then encoded by an n \times n (a_{ij}(g)) with entries in F, satisfying \rho(g) e_j = \sum_{i=1}^n a_{ij}(g) e_i for each j = 1, \dots, n. This facilitates computations, as the property translates to : A(gh) = A(g) A(h), where A(g) = (a_{ij}(g)). The choice of basis affects the specific matrices but not the underlying representation. If P is an invertible n \times n matrix representing a change of basis, the transformed matrices are given by similarity A'(g) = P^{-1} A(g) P for all g \in G, preserving the linear action on V. Equivalent representations in this sense yield isomorphic modules over the group algebra F[G]. Over F = \mathbb{C}, a representation is unitary if each \rho(g) is a unitary linear operator, meaning \rho(g)^* = \rho(g)^{-1} with respect to a Hermitian inner product on V, thereby preserving the inner product: \langle \rho(g) v, \rho(g) w \rangle = \langle v, w \rangle for all v, w \in V. For compact groups, every finite-dimensional complex representation is equivalent to a unitary one, obtained by averaging an inner product over the group. Although finite-dimensional spaces form the core setting, the concept extends to infinite-dimensional Hilbert spaces, where unitary representations ensure continuity and completeness in the analysis of group actions, as in on non-compact groups.

Fundamental Properties

Equivalence of representations

In , two representations \rho: [G](/page/G) \to \mathrm{[GL](/page/GL)}(V) and \sigma: [G](/page/G) \to \mathrm{[GL](/page/GL)}(W) of a group [G](/page/G) on spaces [V](/page/V.) and [W](/page/V.) over the same are said to be equivalent if there exists an invertible T: [V](/page/V.) \to [W](/page/V.) such that \sigma(g) = T \rho(g) T^{-1} for all g \in [G](/page/G). This condition implies that the representations are related by a in the spaces, preserving the group action up to . is an on the set of representations, partitioning them into classes where representations within the same class are structurally indistinguishable. The invertible map T is known as an intertwining operator (or G-) between the representations, satisfying the commutation relation T \rho(g) = \sigma(g) T for all g \in G. The space of all such intertwining operators forms the Hom space \mathrm{Hom}_G(V, W), which is a vector space under pointwise addition and scalar multiplication. When T is invertible, the representations are isomorphic, meaning V and W are isomorphic as modules over the group algebra \mathbb{F}[G], where the group action is extended linearly. This module-theoretic perspective underscores that equivalent representations capture the same abstract G-module structure. Representations are classified up to equivalence, with uniqueness holding in the sense that any two equivalent representations yield the same , often used to study invariants like or types. For decompositions, two representations are equivalent their direct summands match up to equivalence and multiplicity; for instance, \rho \oplus \sigma is equivalent to \rho' \oplus \sigma' precisely when the pairs \{\rho, \sigma\} and \{\rho', \sigma'\} consist of equivalent components with the same multiplicities. This ensures that non-matching decompositions, such as differing irreducible summands, yield non-equivalent representations.

Subrepresentations and quotients

In the context of a representation \rho: [G](/page/G) \to \mathrm{[GL](/page/GL)}(V) of a group [G](/page/G) on a V, a subrepresentation is defined as a W \subseteq V such that \rho(g)W \subseteq W for all g \in [G](/page/G). This condition ensures that the restriction of \rho to W, denoted \rho|_W: [G](/page/G) \to \mathrm{[GL](/page/GL)}(W), defines a valid representation on W. Such a subspace W is also called a [G](/page/G)-invariant subspace or simply an , emphasizing the preservation of the subspace under the . Given a subrepresentation W \subseteq V, the quotient space V/W inherits a natural representation structure from \rho, known as the quotient representation. This is defined by \overline{\rho}(g)(v + W) = \rho(g)v + W for all g \in G and v \in V, where the bar denotes the induced map on the quotient. The quotient representation captures the action of G on the "cosets" of W within V, providing a way to factor out the subrepresentation while preserving the linear group action. These constructions fit into the framework of s of representations. Specifically, for a subrepresentation W \subseteq V, there is a short $0 \to W \xrightarrow{i} V \xrightarrow{q} V/W \to 0, where i is the and q is the canonical map, both G-equivariant (i.e., commuting with the representation actions). This sequence is exact in the category of representations, meaning \ker i = 0, \mathrm{im} i = \ker q = W, and \mathrm{im} q = V/W, thus encoding the relationship between the subrepresentation, the original representation, and the . Representations can be classified based on their subrepresentation structure, particularly through the notions of and semisimple modules (or representations, viewed as modules over the group algebra). A representation is if it admits no proper nonzero subrepresentations, meaning the only invariant subspaces are \{0\} and V itself. In contrast, a representation is semisimple if it decomposes as a of representations; here, subrepresentations and quotients behave particularly well, as every subrepresentation has a complementary , ensuring that quotient maps have precisely the expected kernels without additional complications from non-split extensions. This distinction highlights how semisimple representations allow for clean decomposition into building blocks via subrepresentations and quotients. The kernel of a representation \rho: G \to \mathrm{GL}(V) is the normal subgroup \ker \rho = \{ g \in G \mid \rho(g) = \mathrm{Id}_V \}, consisting of group elements that act trivially on V. Since \rho is a , this kernel is normal in G, and the representation factors through the G / \ker \rho. This kernel provides insight into the "ineffective" part of the and relates subrepresentations to the broader structure of .

Examples

Permutation representations

A permutation representation of a finite group G arises from a left of G on a finite set X. This action induces a linear representation \rho: G \to \mathrm{GL}(V) on the complex vector space V = \mathbb{C}^X of functions f: X \to \mathbb{C}, defined by (g \cdot f)(x) = f(g^{-1} x) for g \in G, f \in V, and x \in X. The space V admits a natural permutation basis consisting of the Dirac delta functions \{\delta_x \mid x \in X\}, where \delta_x(y) = 1 if y = x and 0 otherwise; the group action permutes these basis vectors according to the action on X. In this basis, the representing matrix \rho(g) is a with a 1 in position (x, g^{-1} x) for each x \in X, and 0s elsewhere. Consequently, the of \rho(g), which counts the number of 1s on the diagonal, equals the number of fixed points of g on X, i.e., \operatorname{tr} \rho(g) = |\{ x \in X \mid g x = x \}|. The permutation representation decomposes according to the G-s on X: since X is a of s, V is the of the subspaces spanned by the basis elements in each , yielding a of transitive permutation representations (one for each ). For the symmetric group S_n acting naturally on \{1, 2, \dots, n\}, the associated permutation representation admits a 1-dimensional subrepresentation known as the sign representation, given by \rho(\sigma) = \det(\rho_{\mathrm{perm}}(\sigma)) = \operatorname{sgn}(\sigma) = (-1)^{n - c(\sigma)}, where c(\sigma) is the number of cycles in \sigma (yielding +1 for even permutations and -1 for odd ones). This representation is the unique nontrivial 1-dimensional representation of S_n. A concrete example is the natural permutation representation of S_3 on the set \{1, 2, 3\}, which has dimension 3. The character of this representation takes value 3 on the , 1 on each of the three transpositions, and 0 on each of the two 3-cycles. By character orthogonality, it decomposes as the of the 1-dimensional trivial representation and the 2-dimensional of S_3.

Regular and induced representations

The regular representation of a finite group G over the complex numbers is the representation \rho: G \to \mathrm{GL}(\mathbb{C}[G]), where \mathbb{C}[G] is the group algebra with basis \{e_g \mid g \in G\} and G acts by left multiplication: \rho(h) \left( \sum_{g \in G} a_g e_g \right) = \sum_{g \in G} a_g e_{h g}. This construction endows \mathbb{C}[G] with a natural G-module structure, capturing the group's action on itself. The dimension of the regular representation is |G|, as the basis has one element per group element. In the basis \{e_g\}, the matrix of \rho(h) is a permutation matrix corresponding to the left multiplication by h, which permutes the basis elements by shifting indices: the entry in position (e_k, e_g) is 1 if k = h g and 0 otherwise. This permutation action highlights the as a faithful representation of G, embedding it into the on |G| letters. For finite G, the regular representation decomposes as a of all s of G, where each V_i appears with multiplicity equal to \dim V_i. This multiplicity follows from the of characters and implies the sum-of-squares \sum_i (\dim V_i)^2 = |G|, providing a key tool for classifying irreducibles. Given a subgroup H \leq G and a representation \sigma: H \to \mathrm{GL}(V) of H on a finite-dimensional complex V, the \mathrm{Ind}_H^G \sigma is the representation of G on the vector space \mathbb{C}[G] \otimes_{\mathbb{C}[H]} V, with dimension \dim V \cdot |G:H|. Equivalently, it can be described as the space of \mathbb{C}-valued functions f: G \to V satisfying f(h x) = \sigma(h) f(x) for all h \in H, x \in G, with G-action (g \cdot f)(x) = f(g^{-1} x). The action of G on the induced space is given explicitly by (\mathrm{Ind}_H^G \sigma)(g) \left( \sum_{t \in T} e_t \otimes v_t \right) = \sum_{t \in T} e_{g t} \otimes v_t, where T is a set of coset representatives for H in G, adjusted for elements where g t \in H u for some u \in T via the H-action. More generally, for an arbitrary element \sum a_h h \otimes v \in \mathbb{C}[G] \otimes_{\mathbb{C}[H]} V, the action is g \cdot (\sum a_h h \otimes v) = \sum a_h (g h) \otimes v, with identification under the right \mathbb{C}[H]-action. The Frobenius reciprocity theorem relates induction and restriction: for representations V of G and W of H, there is a natural \mathrm{Hom}_G(V, \mathrm{Ind}_H^G W) \cong \mathrm{Hom}_H(\mathrm{Res}_H^G V, W). In terms of characters, this yields \langle \chi_V, \chi_{\mathrm{Ind}_H^G W} \rangle_G = \langle \chi_{\mathrm{Res}_H^G V}, \chi_W \rangle_H, equating multiplicities under induction and restriction.

Reducibility

Reducible representations

A representation \rho: G \to \mathrm{GL}(V) of a group G on a finite-dimensional V over a k (such as \mathbb{C}) is said to be reducible if there exists a proper nontrivial W \subset V that is under the action of all \rho(g) for g \in G, meaning \rho(g)W \subseteq W for every g. This invariance implies that the representation restricts to a subrepresentation on W and induces a representation on V/W. A representation is completely reducible if it decomposes as a direct sum of irreducible representations, i.e., V = W_1 \oplus W_2 \oplus \cdots \oplus W_m where each W_i is an irreducible subrepresentation. Over the complex numbers \mathbb{C} for finite groups G, every finite-dimensional is completely reducible, as guaranteed by Maschke's , which states that any has a complementary , allowing full decomposition into irreducibles. This property holds because the characteristic of \mathbb{C} does not divide the order of G, ensuring the group algebra \mathbb{C}[G] is semisimple. Schur's lemma provides a key tool for understanding irreducible representations within reducible ones: if \rho is irreducible over an algebraically closed field like \mathbb{C}, then the endomorphism algebra \mathrm{End}_G(V) = \{ T \in \mathrm{End}(V) \mid T \rho(g) = \rho(g) T \ \forall g \in G \} consists precisely of scalar multiples of the identity, i.e., \mathrm{End}_G(V) \cong k \cdot \mathrm{Id}. This implies that irreducible subrepresentations are rigid, with no nontrivial intertwiners, which aids in decomposing reducible representations by identifying distinct irreducible factors. To explicitly decompose reducible representations, projection operators are constructed using group averages. The orthogonal projection onto the subspace of invariants V^G = \{ v \in V \mid \rho(g)v = v \ \forall g \in G \} is given by P = \frac{1}{|G|} \sum_{g \in G} \rho(g), which is idempotent (P^2 = P) and G-equivariant when |G| is invertible in k. For finite G over \mathbb{C}, more generally, the projection onto the isotypic component V_\sigma corresponding to an irreducible representation \sigma (the sum of all subrepresentations isomorphic to \sigma) is P_\sigma = \frac{\dim \sigma}{|G|} \sum_{g \in G} \chi_\sigma(g^{-1}) \rho(g), where \chi_\sigma is the character of \sigma; this operator satisfies P_\sigma^2 = P_\sigma and projects V onto V_\sigma while annihilating other components. These projections enable the explicit construction of the complete decomposition into isotypic components, each of which is a direct sum of copies of \sigma.

Irreducible representations

An of a group G on a V over a k is a \rho: G \to \mathrm{GL}(V) that admits no proper nontrivial subrepresentations, meaning the only G- subspaces of V are \{0\} and V itself. In the context of theory, such a representation corresponds to a simple module over the group algebra k[G], where the module has no proper nontrivial submodules. A key criterion for irreducibility is given by , which characterizes the endomorphism ring of an . Specifically, if \rho: G \to \mathrm{GL}(V) is an over an k and \mathrm{End}_G(V) = \{\phi \in \mathrm{End}(V) \mid \phi \circ \rho(g) = \rho(g) \circ \phi \ \forall g \in G\} denotes the space of G-equivariant s of V, then \mathrm{End}_G(V) = k \cdot \mathrm{Id}_V, the scalar multiples of the identity operator. To prove this, first note that for any \phi \in \mathrm{End}_G(V), the image \mathrm{im}(\phi) is a G- of V. Since \rho is irreducible and \phi \neq 0, it follows that \mathrm{im}(\phi) = V, so \phi is surjective. Similarly, the \ker(\phi) is G-invariant, and since \dim V < \infty, surjectivity implies injectivity, hence \ker(\phi) = \{0\}. Thus, \phi is invertible, and the set of all such \phi forms a division algebra over k. By the assumption that k is algebraically closed, this division algebra must be k itself, so every \phi \in \mathrm{End}_G(V) is scalar. Conversely, if \mathrm{End}_G(V) = k \cdot \mathrm{Id}_V, then any nonzero \phi \in \mathrm{End}_G(V) is invertible, implying that no proper nontrivial G- exists, as the image of any such subspace under \phi would contradict irreducibility. This criterion provides a practical test for irreducibility: a representation is irreducible if and only if its endomorphism ring consists solely of scalars. For finite groups, Maschke's theorem guarantees that representations decompose into irreducibles under suitable conditions. Precisely, if G is finite and k is a field whose characteristic does not divide |G|, then every finite-dimensional representation of G over k is semisimple, meaning it is isomorphic to a direct sum of irreducible representations. The proof proceeds by showing that any subrepresentation has a complementary invariant subspace. Suppose W \subset V is a G-invariant subspace of a finite-dimensional representation (V, \rho). Equip V with an inner product \langle \cdot, \cdot \rangle, and define the projection P: V \to W by averaging over the group: more invariantly, the orthogonal projection onto W followed by group averaging P(v) = \frac{1}{|G|} \sum_{g \in G} \rho(g) P_0(\rho(g^{-1}) v), where P_0 is the orthogonal projection to W. Since \mathrm{char}(k) \nmid |G|, the averaging operator is well-defined and G-equivariant, and its image is W with kernel complementary to W. This yields V = W \oplus U with U G-invariant. Iterating this decomposition shows full semisimplicity. Moreover, for finite G, the number of distinct irreducible representations (up to isomorphism) over an algebraically closed field of characteristic not dividing |G| equals the number of conjugacy classes in G. This semisimplicity extends to compact groups via Weyl's unitary trick. For a compact Lie group G, every continuous finite-dimensional representation on a complex vector space is equivalent to a unitary representation with respect to some G-invariant Hermitian inner product, obtained by averaging an arbitrary inner product over G using the Haar measure. Since unitary representations preserve the inner product, any invariant subspace has an orthogonal complement that is also invariant, mirroring the finite-group case and implying complete reducibility into irreducibles. This trick reduces the study of representations of compact groups to the unitary case, where orthogonality relations simplify analysis.

Character Theory

Definition and basic properties of characters

In representation theory, the character of a representation \rho: G \to \mathrm{GL}(V) of a finite group G on a finite-dimensional complex vector space V is defined as the function \chi_\rho: G \to \mathbb{C} given by \chi_\rho(g) = \operatorname{tr}(\rho(g)) for each g \in G, where \operatorname{tr} denotes the trace of the matrix representing the linear operator \rho(g). This definition extends to a linear functional on the group algebra \mathbb{C}[G] by linearity. Characters possess several fundamental algebraic properties. First, \chi_\rho is a class function, meaning \chi_\rho(gh) = \chi_\rho(hg) for all g, h \in G, since the trace satisfies \operatorname{tr}(\rho(gh)) = \operatorname{tr}(\rho(g)\rho(h)) = \operatorname{tr}(\rho(h)\rho(g)) = \operatorname{tr}(\rho(hg)) and is invariant under simultaneous conjugation of the matrices. Additionally, \chi_\rho(e) = \dim V, as \rho(e) is the identity operator whose trace equals the dimension of the space. For representations that can be chosen unitary (which is always possible over \mathbb{C}), \chi_\rho(g^{-1}) = \overline{\chi_\rho(g)}, the complex conjugate, because the eigenvalues of \rho(g) are roots of unity and those of \rho(g^{-1}) are their conjugates. Characters exhibit multiplicativity with respect to direct sums and tensor products of representations. For representations \rho: G \to \mathrm{GL}(V) and \sigma: G \to \mathrm{GL}(W), the character of the direct sum satisfies \chi_{\rho \oplus \sigma}(g) = \chi_\rho(g) + \chi_\sigma(g) for all g \in G, following from the additivity of the trace on block-diagonal matrices. Similarly, the character of the tensor product is \chi_{\rho \otimes \sigma}(g) = \chi_\rho(g) \chi_\sigma(g), as the trace of the Kronecker product of matrices multiplies accordingly. The space of class functions on G carries a Hermitian inner product defined by \langle \chi, \psi \rangle = \frac{1}{|G|} \sum_{g \in G} \chi(g) \overline{\psi(g)} for characters \chi, \psi. This inner product is positive definite and induces a Hilbert space structure, with \langle \chi_\rho, \chi_\sigma \rangle = \dim \mathrm{Hom}_G(V, W) measuring the intertwining dimension between representations. The kernel of a character \chi, defined as \ker \chi = \{ g \in G \mid \chi(g) = \chi(e) \}, forms a normal subgroup of G. This subgroup consists of elements acting as scalar multiples of the identity on the representation space, up to the character's trace value. For a representation \sigma of a subgroup H \leq G, the character of the induced representation \chi_{\mathrm{Ind}_H^G \sigma} is given by the formula \chi_{\mathrm{Ind}_H^G \sigma}(g) = \frac{1}{|H|} \sum_{\substack{k \in G \\ k^{-1} g k \in H}} \sigma(k^{-1} g k) for g \in G, where the sum is over those k such that the conjugate k^{-1} g k lies in H (and \sigma is extended by zero outside H). This expression, known as the , arises from averaging the action over cosets.

Orthogonality and decomposition

One of the key features of character theory for finite groups is the orthogonality relations satisfied by the characters of irreducible representations. These relations arise from the inner product on the space of class functions F_c(G, \mathbb{C}), defined by \langle \chi, \psi \rangle = \frac{1}{|G|} \sum_{g \in G} \chi(g) \overline{\psi(g)}, where G is a finite group and \chi, \psi are class functions (constant on conjugacy classes). For characters \chi_V and \chi_W of irreducible representations V and W, this inner product equals \dim \Hom_G(V, W), which is 1 if V \cong W and 0 otherwise. Thus, the column orthogonality relation states that \sum_{g \in G} \chi_V(g) \overline{\chi_W(g)} = |G| \delta_{V,W}, where \delta_{V,W} is the Kronecker delta. This orthogonality implies that distinct irreducible characters are linearly independent. The row orthogonality relation provides another perspective, focusing on sums over irreducible characters for fixed group elements. For elements g, h \in G, \sum_V \chi_V(g) \overline{\chi_V(h)} = |C_G(g)| \quad \text{if } g \text{ and } h \text{ are conjugate, and } 0 \text{ otherwise}, where the sum is over a complete set of irreducible representations V and C_G(g) is the centralizer of g in G. Since the size of the conjugacy class of g is |G|/|C_G(g)|, this relation shows that the columns of the character table (indexed by conjugacy classes) are orthogonal when appropriately weighted by class sizes. These orthogonality relations establish the completeness of the irreducible characters: they form an orthonormal basis for the vector space of class functions F_c(G, \mathbb{C}) with respect to the inner product above. The dimension of this space equals the number of conjugacy classes, which matches the number of irreducible representations by basic properties of characters. This basis property allows any class function, including the character of an arbitrary representation, to be uniquely expressed as a linear combination of irreducible characters. A central application is the decomposition of any finite-dimensional representation \rho: G \to \GL(V) into a direct sum of irreducible representations. The multiplicity m_i of the irreducible representation with character \chi_i in the decomposition \chi = \sum_i m_i \chi_i is given by the inner product m_i = \langle \chi, \chi_i \rangle = \frac{1}{|G|} \sum_{g \in G} \chi(g) \overline{\chi_i(g)}. Since characters determine representations up to isomorphism over \mathbb{C}, this formula completely classifies the representation via its character table projection. Character tables tabulate these values for all conjugacy classes and irreducible characters, facilitating computations. For example, the symmetric group S_3 has three conjugacy classes: the identity (size 1), transpositions like (1\,2) (size 3), and 3-cycles like (1\,2\,3) (size 2). Its three irreducible representations yield the following character table:
RepresentationId (size 1)(1 2) (size 3)(1 2 3) (size 2)
Trivial (C_+)111
Sign (C_-)1-11
Standard (C_2)20-1
The orthogonality relations can be verified directly on this table, confirming the basis property and enabling decompositions, such as the regular representation of S_3 as C_+ \oplus C_- \oplus C_2 \oplus C_2.

Branches

Representations of finite groups

In the theory of group representations, finite groups exhibit particularly tractable behavior over fields of characteristic zero, such as the complex numbers \mathbb{C}. Every finite-dimensional representation of a finite group G over \mathbb{C} is completely reducible, meaning it decomposes as a direct sum of irreducible representations. This result, known as , relies on the fact that the group order |G| is invertible in \mathbb{C}, allowing an averaging projection onto invariant subspaces. Consequently, the representation theory of finite groups over \mathbb{C} reduces to classifying the irreducible representations, which are finite in number and uniquely determined up to isomorphism. The structure of the group algebra \mathbb{C}[G] further illuminates this classification. By the Artin-Wedderburn theorem, since \mathbb{C}[G] is a semisimple Artinian algebra, it decomposes as a direct sum of matrix algebras over \mathbb{C}: \mathbb{C}[G] \cong \bigoplus_{i=1}^r M_{n_i}(\mathbb{C}), where the n_i are the dimensions of the distinct irreducible representations of G, and r is the number of such irreducibles, equal to the number of conjugacy classes in G. This isomorphism underscores the complete reducibility and provides a algebraic foundation for understanding all representations as modules over \mathbb{C}[G]. Over fields of positive characteristic p dividing |G|, the situation changes significantly. Representations are no longer necessarily completely reducible, and the number of irreducible representations over an algebraically closed field of characteristic p equals the number of p-regular conjugacy classes in G (those consisting of elements whose orders are coprime to p). This result, due to Brauer, restricts the theory: there are fewer irreducibles than over \mathbb{C}, and their dimensions do not necessarily divide |G|, complicating decomposition compared to the characteristic-zero case. Character tables, which tabulate the values of irreducible characters on conjugacy classes, play a central role in classifying representations of finite groups. Constructing these tables computationally is feasible via algorithms that exploit orthogonality relations and modular arithmetic. A seminal method, introduced by Dixon in 1968, computes irreducible characters by iteratively building power character tables and resolving ambiguities using probabilistic techniques over finite fields. This approach efficiently handles groups of moderate order, enabling explicit verification of decompositions and symmetries. A concrete example is the quaternion group Q_8 = \{\pm 1, \pm i, \pm j, \pm k\} of order 8. It has five irreducible representations over \mathbb{C}: four one-dimensional representations corresponding to the abelian quotient Q_8 / \{\pm 1\} \cong \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z\}, and one faithful two-dimensional representation realized in the quaternions via the standard embedding. The dimensions satisfy $1^2 + 1^2 + 1^2 + 1^2 + 2^2 = 8 = |Q_8|, confirming completeness. Representations over the rationals \mathbb{Q} retain complete reducibility, as \mathbb{Q} has characteristic zero and |G| is invertible therein, so Maschke's theorem applies. However, the irreducible \mathbb{Q}-representations differ from those over \mathbb{C}: they are realized by summing Galois orbits of complex irreducibles under the action of \mathrm{Gal}(\mathbb{Q}(\zeta)/\mathbb{Q}), where \zeta is a root of unity, often yielding higher-dimensional modules that are indecomposable over \mathbb{Q} but split over \mathbb{C}. For instance, cyclic groups of prime order have one-dimensional \mathbb{Q}-irreducibles, but non-abelian groups like Q_8 require a four-dimensional irreducible in addition to the one-dimensional ones to capture the full structure rationally.

Representations of Lie groups

A representation of a Lie group G is a smooth homomorphism \rho: G \to \mathrm{GL}(V), where V is a finite-dimensional complex vector space, encoding the linear action of G on V. For infinite-dimensional settings, particularly unitary representations, \rho acts on a Hilbert space \mathcal{H} by preserving the inner product, i.e., \langle \rho(g)v, \rho(g)w \rangle = \langle v, w \rangle for all g \in G and v, w \in \mathcal{H}. These representations capture the continuous symmetries of G, extending finite-group theory to manifolds with group structure, and are central to applications in quantum mechanics and geometry. The infinitesimal counterpart arises via the derivative d\rho: \mathfrak{g} \to \mathfrak{gl}(V) at the identity, where \mathfrak{g} is the Lie algebra of G, yielding a Lie algebra representation that linearizes the group action. This differential map preserves the Lie bracket, \rho([X,Y]) = [d\rho(X), d\rho(Y)], and facilitates analysis of representations through algebraic tools like root systems. For compact Lie groups, the Peter-Weyl theorem provides a Fourier-like decomposition: the space L^2(G) of square-integrable functions on G (with respect to the Haar measure) decomposes as the completed direct sum \hat{\oplus}_\pi (V_\pi^* \otimes V_\pi), where the sum runs over equivalence classes of irreducible finite-dimensional representations \pi with representation spaces V_\pi, and the summands consist of matrix coefficients. This orthogonality implies that irreducible representations are finite-dimensional and that L^2(G) is spanned by these coefficients, enabling harmonic analysis on non-abelian groups. In the semisimple case, irreducible finite-dimensional representations of a complex semisimple Lie algebra \mathfrak{g} (and thus of the corresponding simply-connected Lie group) are classified by dominant integral weights via highest weight theory, originally developed by Cartan and Weyl. Specifically, each such representation corresponds uniquely to a dominant weight \lambda in the weight lattice, with highest weight space one-dimensional and annihilated by positive root vectors; the weights lie in the convex hull of the Weyl group orbit of \lambda. This parametrization, known as the , determines the representation up to isomorphism and extends to compact or reductive groups through analytically integral dominant weights on the Cartan subalgebra. A canonical example is the special unitary group \mathrm{SU}(2), whose irreducible representations are the spin representations labeled by j = 0, 1/2, 1, 3/2, \dots, each of dimension $2j + 1 with basis states |j, m\rangle for m = -j, \dots, j. These arise from the ladder operators J_\pm = J_1 \pm i J_2 acting on the highest weight state |j, j\rangle, generating the full space while preserving the commutation relations [J_i, J_j] = i \epsilon_{ijk} J_k. Characters of representations on compact Lie groups are class functions \chi_\rho(g) = \mathrm{Tr}(\rho(g)), integrated against test functions f \in C(G) via the Haar measure dg (normalized to \int_G dg = 1): for instance, the multiplicity of an irrep \pi in \rho is \int_G \overline{\chi_\pi(g)} \chi_\rho(g) \, dg. This inner product leverages the bi-invariant Haar measure, unique up to scalar for compact G, to orthogonalize characters and decompose representations analytically.

Generalizations

Representations over rings and modules

The group ring R[G] associated to a commutative ring R with identity and a group G is the free left R-module with basis \{ g \mid g \in G \}, consisting of all formal finite sums \sum_{g \in G} r_g g where r_g \in R. Addition is defined componentwise, and multiplication is extended bilinearly from the group operation: (r g)(s h) = (r s)(g h) for r, s \in R and g, h \in G, making R[G] an associative unital R-algebra. A representation of G over R is equivalently a left R[G]-module M, or a unital ring homomorphism R[G] \to \mathrm{End}_R(M), where the group action arises from the module structure via \left( \sum r_g g \right) m = \sum r_g \rho(g)(m) for m \in M and a corresponding representation \rho: G \to \mathrm{Aut}_R(M). This generalizes the classical case of representations over fields, where modules are vector spaces, but extends to more general rings where modules may not be free or semisimple. In modular representation theory, attention shifts to cases where R is a field k of characteristic p > 0 dividing the order |G| of a finite group G, so k[G] is not semisimple. Here, Maschke's theorem fails, as the group algebra lacks the property that every module is a direct sum of simples, leading to non-split extensions and indecomposable modules beyond the simples. Blocks of k[G] are the indecomposable two-sided ideals corresponding to primitive central idempotents, partitioning the simple k[G]-modules and capturing linked representations via their projective covers. The decomposition matrix D relates the irreducible characters over a field of characteristic zero (e.g., \mathbb{C}) to the irreducible Brauer characters over k, with entries d_{ij} giving the multiplicity of the j-th simple k[G]-module in the reduction modulo p of the i-th ordinary irreducible module; these matrices are integral with non-negative entries and determine the Cartan matrix of composition multiplicities in projectives. Projective k[G]-modules play a central role in characteristic p, as every finitely generated module admits a projective cover, and the indecomposable projectives are in bijection with the simple modules, each being the unique indecomposable projective with a given simple socle or head. For a simple module S, its projective cover P_S has head S and is determined up to , with the set of all indecomposables generating the category in blocks of defect greater than zero. In blocks with full defect (defect group a Sylow p-subgroup), the projectives encode the non-semisimple structure, and Green's relates projectives over G to those over subgroups containing normalizers of defect groups. Examples illustrate these concepts over integral domains like \mathbb{Z}. For a cyclic group G = \langle g \rangle of prime order p, the integral group ring \mathbb{Z}[G] has representations as \mathbb{Z}[G]-lattices, which decompose as M \cong M_s \oplus \bigoplus_{i=1}^m \mathbb{Z} y_i, where s = 1 + g + \cdots + g^{p-1} annihilates the torsion submodule M_s, an o-module with o = \mathbb{Z}[\theta] for a primitive p-th root of unity \theta, isomorphic to a direct sum of ideals in o. Torsion arises in (g-1)M / (\theta - 1)M_s, of type (p, \dots, p) with invariants including the \mathbb{Z}-rank and ideal classes, yielding $2h + 1 non-isomorphic indecomposables where h is the class number of o. Brauer characters provide a character theory for modular representations, defined for a k[G]-module M on p-regular elements g \in G (those of order coprime to p) as the trace of the action, lifted to a complex-valued function via a map from eigenvalues in k^\times to roots of unity. For an indecomposable projective P, the Brauer character \eta_P vanishes on non-p-regular elements and equals a \mathbb{Z}-linear combination of ordinary characters via the transpose decomposition matrix, enabling decomposition of modular representations from characteristic zero data. In the case of p-groups, Brauer characters lift directly to ordinary characters of the same degree, reflecting the uniserial structure of projectives over cyclic p-group algebras.

Categorical representations

In , a representation of a group G on a \mathcal{C} is defined by a \rho: G \to \Aut(\mathcal{C}), where \Aut(\mathcal{C}) denotes the of strict automorphisms of \mathcal{C} (s of categories that preserve the skeletal structure strictly), such that \rho preserves the group operation: \rho(gh) = \rho(g) \circ \rho(h) and \rho(e) = \id_{\mathcal{C}}. More generally, to accommodate non-strict s, one considers functors to the group of autoequivalences of \mathcal{C}, which are equivalence-preserving transformations up to natural . This generalizes classical representations, where \mathcal{C} is the of vector spaces or modules, by abstracting the action to the level of the entire rather than individual objects. The of representations of G in the , denoted \Rep(G, \Set), consists of G-sets and equivariant maps, which arise as representable functors from the delooping BG (the one-object with morphisms given by G) to \Set. These actions on objects of \Set extend naturally to the categorical framework, providing a foundational example where the representation recovers the group's permutation action on discrete structures. Examples of categorical representations abound in structured categories. In abelian categories, such as the of coherent sheaves on an X, a group G acting on X induces a representation via pullback functors f_g^*: \Sh(X) \to \Sh(X) for g \in G, which form autoequivalences preserving the abelian structure. Similarly, in topological categories, where objects carry topology and morphisms are continuous, the functor \rho must consist of continuous autoequivalences to respect the topological enrichment. Enriched representations over a monoidal like \Vect_k (vector spaces over a field k) generalize linear representations by requiring the autoequivalences to be k-linear and monoidal, preserving tensor products up to isomorphism. Tannakian duality provides a reconstruction theorem for certain categorical representations: for a neutral Tannakian category \mathcal{T} over a k equipped with a fiber functor \omega: \mathcal{T} \to \Vect_k, the category \Rep(G) of finite-dimensional representations of an affine G is equivalent to \mathcal{T}, allowing recovery of G as the \Aut^\otimes(\omega) of the fiber functor. This duality underscores how the categorical structure encodes the group via tensor-preserving actions. In higher dimensions, categorical representations extend to 2-groups (categorical groups), where a representation is a 2-functor from the 2-group to the 2-category of autoequivalences of \mathcal{C}, yielding a monoidal 2-category of such representations that generalizes the 1-categorical case.

References

  1. [1]
    [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.
  2. [2]
    None
    ### Summary of Group Representations Introduction
  3. [3]
    [PDF] representation theory its rise and its role in number theory
    Jul 3, 2018 · Initially, the group is finite, as in the researches of Dedekind and. Frobenius, two of the founders of the subject, or a compact Lie group, as ...
  4. [4]
    [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 ...
  5. [5]
    Charles W. Curtis. Pioneers of Representation Theory
    In addition to Frobenius, the book focuses mainly on three other “pioneers”: William Burnside, Issai Schur, and Richard Brauer. Curtis states that his interest ...
  6. [6]
    [PDF] Hermann Weyl and Representation Theory
    In 1925–26, Weyl wrote four path-breaking papers in representa- tion theory which apart from solving fundamental problems, also gave birth to the subject of ...Missing: 1931 | Show results with:1931
  7. [7]
    [PDF] On the history of Artin's L-functions and conductors Seven letters ...
    Jul 23, 2003 · In the year 1923 Emil Artin introduced his new L-functions belonging to. Galois characters. But his theory was still incomplete in two respects.
  8. [8]
    extended artin-schreier theory of fields - Project Euclid
    Dedicated to the memory of Gus Efroymson. Introduction. This survey is concerned with recent developments in the. Artin-Schreier theory of fields.
  9. [9]
    [PDF] Computational Group Theory
    Oct 5, 1998 · Computational group theory has a history going back more than 80 years. It is a collaborative effort of researchers in a wide range of areas ...
  10. [10]
    [PDF] Group Representations and Harmonic Analysis from Euler to ...
    May 3, 1996 · theory of finite groups occurs with Artin L func- tions, which Artin introduced in the 1920s. An. Artin L function over the rationals Q encodes.<|control11|><|separator|>
  11. [11]
    [PDF] Math 210B. Why study representation theory? - Mathematics
    Why study representation theory? 1. Motivation. Books and courses on group theory often introduce groups as purely abstract algebraic objects, but in practice ...
  12. [12]
    [PDF] Group Representations and Character Theory - UChicago Math
    Aug 26, 2011 · The primary motivation for the study of group representations is to simplify the study of groups. Representation theory offers a powerful ...
  13. [13]
    [PDF] The Fundamental Theorem of Invariant Theory
    Nov 15, 2006 · 1890: Hilbert's finiteness theorem: C[V]G is finitely generated for a wide class of groups (linearly reductive groups). Gordan: “Das ist ...
  14. [14]
    [PDF] Linear Representations of Finite Groups - Auburn University
    A group homomorphism ρ : G → GL(V ) is called a linear K-representation of G in V (or just a representation of G for short).
  15. [15]
    [PDF] Representation Theory - UC Berkeley math
    Groups arise in nature as “sets of symmetries (of an object), which are closed under compo- sition and under taking inverses”. For example, the symmetric ...Missing: history | Show results with:history
  16. [16]
    [PDF] representation theory workshop
    (1) For G any group, there is always a trivial representation given by V = C ... Then we get a representation of G by composing the homomorphisms. G → G/N → GL(V ).
  17. [17]
    [PDF] Part II - Representation Theory - Dexter Chua
    The dimension (or degree) of a representation ρ : G → GL(V ) is dimF(V ). Recall that ker ρ < G and G/ ker ρ. ∼. = ρ(G) ≤ GL(V ). In ...<|separator|>
  18. [18]
    [PDF] m3/4/5p12 group representation theory - People
    (3) Character theory: the character of a group representation, character tables and ... A representation is faithful if ρV is injective. Exercise 1.1. For what ...
  19. [19]
    [PDF] Group representation theory, Lecture Notes - Imperial College London
    Jun 3, 2024 · A representation of a group G is a pair (V,ρ) where V is a vector space and ρ : G → GL(V ) is a group homomorphism. Remark 2.4.2. We can also ...
  20. [20]
    Linear Representations of Finite Groups - SpringerLink
    In stockBook Title: Linear Representations of Finite Groups · Authors: Jean-Pierre Serre · Series Title: Graduate Texts in Mathematics · Publisher: Springer New York, NY.
  21. [21]
    [PDF] Representation Theory and its Applications in Physics
    Equivalence of representations, as stated in Definition 1.1.1. is an equivalence relation on the representations. Proof. We need to prove that the ...
  22. [22]
    [PDF] Representation Theory of Groups - Algebraic Foundations 1.1 B
    It is easy to check that the notion of equivalence of representations defines an equivalence relation on the set of representations of G. It follows from ...
  23. [23]
    None
    ### Summary of Equivalence of Group Representations from https://www.math.lsu.edu/~sengupta/notes/finitegrouprepresent.pdf
  24. [24]
    [PDF] A brief introduction to group representations and character theory
    Sep 28, 2018 · 1 Reflecting my personal taste, these brief notes emphasize character theory rather more than general representation theory. 1. Definition and ...
  25. [25]
    [PDF] representation theory. lecture notes - vera serganova
    Aug 30, 2005 · An operator T ∈ HomG (V,W) is called an intertwining operator. It is clear that HomG (V,W) is a vector space.
  26. [26]
    [PDF] Representation Theory of Finite Groups - arXiv
    this representation is reducible and is an injective map (also called faithful representation). ... [D1] Dornhoff L., “Group representation theory. Part A ...
  27. [27]
    [PDF] Representations of finite groups - UBC Math
    Nov 10, 2016 · There is an English translation, titled Linear representations of finite groups. It was published by Springer-. Verlag in 1977 and contains ...
  28. [28]
    [PDF] Lie Group, Lie Algebra and their Representations
    Jun 8, 2010 · Definition 53. If S ⊆ V is a g-submodule then V/S is also a representation of g, the quotient representation. Definition 54. Let g be a Lie ...
  29. [29]
    [PDF] 20 Representation theory - UC Berkeley math
    First of all, if the group G acts on some set X, then we can write X as the disjoint union of the orbits of G on X, and G is transitive on each of these orbits.
  30. [30]
    [PDF] Introduction to representation theory - MIT Mathematics
    Jan 10, 2011 · Introduction to representation theory. Pavel Etingof, Oleg Golberg, Sebastian Hensel,. Tiankai Liu, Alex Schwendner, Dmitry Vaintrob, and ...
  31. [31]
    [PDF] Representation Theory1
    Equivalent representations: Two representa- tions D and D0 are equivalent if they are re- lated by a similarity transformation (invert- ible matrix) S: i.e., D0 ...
  32. [32]
  33. [33]
    [PDF] maschke's theorem - abigail sheldon - UChicago Math
    The basic concept in representation theory is that of a group representation. Let G be a group and let F be R or C. Recall that GL(n,F) denotes the group of ...
  34. [34]
    [PDF] A Basic Note on Group Representations and Schur's Lemma
    Abstract. Here we look at some basic results from group representation theory. We discuss Schur's Lemma in the context of R[G]-modules and note.
  35. [35]
    [PDF] Notes on Representations of Finite Groups
    We now give an alternate proof, as it gives us the opportunity to introduce the regular representation and see more representation theory techniques in action.
  36. [36]
    [PDF] Lectures on representations of finite groups and invariant theory
    Chapter I. Representation theory of finite groups. 5. I.1. Basic definitions and examples. 5. I.2. Invariant subspaces and complete reducibility.
  37. [37]
    Ueber den arithmetischen Charakter der Coefficienten der ...
    Maschke, H. Ueber den arithmetischen Charakter der Coefficienten der Substitutionen endlicher linearer Substitutionsgruppen. Math. Ann. 50, 492–498 (1898).
  38. [38]
    [PDF] CHARACTER THEORY OF FINITE GROUPS Chapter 1 - SLMath
    Notation: We write ker(χ) = {g ∈ G | χ(g) = χ(1)}. This normal subgroup is called the kernel of χ, and χ is faithful if ker(χ) = 1.
  39. [39]
    None
    Summary of each segment:
  40. [40]
    [PDF] MASCHKE'S THEOREM OVER GENERAL FIELDS Let G be a finite ...
    This theorem is about groups with order equal to a power of p, not groups with the weaker property of having order divisible by p. Proof. We begin by showing ...
  41. [41]
    [PDF] An Introduction to Wedderburn Theory & Group Representations
    If Char(F)=0 or (Char(F),|G|)=1, then FG is a semisimple algebra. Theorem 4.2. The group algebra CG is isomorphic to. Mn1 (C) ⊕···⊕ Mnr (C).Missing: complex | Show results with:complex
  42. [42]
    [PDF] Representations of a finite group in positive characteristic
    Brauer's theorem is easy to state: an element of G is called p-regular if its order is not divisible by p. A p-regular conjugacy class is a con- jugacy class ...
  43. [43]
    Representations in characteristic p - MathOverflow
    Jan 24, 2010 · Brauer's proof that the number of similarity classes of irreducible representations of G over an algebraically closed field of characteristic p ...
  44. [44]
    Computing Irreducible Representations of Groups - jstor
    In a previous paper [5] I have described a method of computing the character table of a finite group using modular arithmetic. Here I want to point out that one ...<|separator|>
  45. [45]
    [PDF] On The Representations and Characters of Quaternions Group Q8
    In this work, we recall first some results on the theory of reprentations and character of finite groups, and by the end we arrive to determine the irreducible ...
  46. [46]
    Constructing Rational Representations of Finite Groups
    This paper outlines a rst ap- proach to the problem of how irreducible ratio- nal representations can be constructed by comput- ing and analysing endomorphisms ...
  47. [47]
    [PDF] 18.745: lie groups and lie algebras, i - MIT Mathematics
    We have previously defined (finite dimensional) representations of Lie groups and (iso)morphisms between them. We can do the same for Lie algebras: Definition ...
  48. [48]
    [PDF] Introduction to Lie Groups and Lie Algebras Alexander Kirillov, Jr.
    Theorem 6.14 (Lie theorem about representations of solvable Lie algebra). Let ρ: g → gl(V ) be a complex representation of a solvable Lie algebra g. Then ...
  49. [49]
    [PDF] Harmonic Analysis on Compact Lie Groups: the Peter-Weyl Theorem
    The Peter-Weyl theorem says that representations of compact Lie groups behave very much like representa- tions of finite groups, with the analytic issues ...
  50. [50]
    [PDF] Structure Theory of Semisimple Lie Groups
    As a consequence of (e), the Weyl group acts on the weights, pre- serving multiplicities. The extreme weights are those in the orbit of the highest weight.
  51. [51]
    [PDF] Chapter 4 SU(2) - Rutgers Physics
    We have formed a representation with orthonormal basis {|j, mi,m = j, j − 1,j − 2, ··· , −j}, which is 2j + 1 dimensional, with 2j an integer. We found the ...
  52. [52]
    [PDF] Math 210B. Characters for compact Lie groups
    In particular, all of the “classical matrix groups” that are closed subgroups of GLn(R) or GLn(C) are Lie groups, such as: SLn(R), SLn(C),. Sp2g(R), Sp2g(C), ...
  53. [53]
    [PDF] arXiv:1604.01679v2 [math.QA] 12 Jul 2017
    Jul 12, 2017 · categories. 1. Introduction. Given a group G and a monoidal category C, a strict action of G on C is a group morphisms from G to AutStrict.
  54. [54]
    [PDF] Tannakian Categories
    May 13, 2011 · (Tannakian duality) Let K be a topological group. The category Rep. R .K/ of con- tinuous representations of K on finite-dimensional real ...
  55. [55]
    [PDF] Categorical representations of categorical groups - arXiv
    Sep 17, 2004 · We show that the categorical representations of a categorical group form a monoidal 2-category, by direct analogy with the way in which the ...