Fact-checked by Grok 2 weeks ago

Compact group

A compact group is a topological group G that is compact as a topological space, meaning the underlying topology is Hausdorff, the group multiplication and inversion maps are continuous, and every open cover of G has a finite subcover. This structure combines algebraic group properties with the geometric constraint of compactness, ensuring that G is both totally bounded and complete in its metric realization. Compact groups play a central role in and due to their rich structural properties. Every compact group admits a unique (up to positive scalar multiple) bi-invariant , which is finite and turns G into a when normalized. Consequently, the space of square-integrable functions L^2(G) with respect to this measure decomposes orthogonally into finite-dimensional irreducible unitary representations via the Peter-Weyl theorem, which states that the matrix coefficients of these representations are dense in C(G), the continuous functions on G. This theorem implies that all continuous unitary representations of compact groups are completely reducible, generalizing the decomposition of representations for finite groups. Examples of compact groups abound in both abstract and concrete settings. The circle group \mathbb{T} = \mathbb{R}/\mathbb{Z}, equipped with addition modulo 1, is the archetypal abelian compact group. Classical Lie groups such as the U(n) and the O(n) are compact matrix groups under , arising naturally in linear algebra and . More generally, profinite groups like the p-adic integers \mathbb{Z}_p provide non-Lie examples, and any with the topology is compact. For abelian compact groups, identifies them with abelian groups, establishing a profound correspondence between compact and structures. These properties make compact groups indispensable in applications ranging from , where symmetry groups like SU(2) model spin, to and via their connections to profinite completions.

Definition and properties

Definition

In , a compact group is a G in which the underlying is compact, meaning that every open cover of G admits a finite subcover. Compact groups are typically assumed to be Hausdorff, ensuring that the topology separates points. For a Hausdorff compact group, equivalent formulations arise in the metrizable case: the induced uniformity makes G a complete and totally bounded . More generally, arbitrary products of compact groups are compact by , facilitating the study of infinite-dimensional examples via inverse limits. The term and concept of compact groups were introduced by in 1925, initially in the context of for compact Lie groups such as the SU(n). This framework was later generalized beyond Lie groups. A basic non-trivial example is the circle group \mathbb{T} = U(1), consisting of complex numbers of modulus 1 under multiplication, which is compact as a of \mathbb{C}. Compactness ensures the existence of a bi-invariant on G.

Basic topological and algebraic properties

In a compact topological group G, the inversion map g \mapsto g^{-1} is a continuous homeomorphism, and the multiplication map (g, h) \mapsto gh from G \times G to G is continuous and uniformly continuous with respect to the respective uniform structures on G and G \times G. The continuity of these operations follows from the definition of a topological group, but compactness ensures additional regularity: specifically, the uniform continuity of multiplication arises because G \times G is compact and the map is continuous, implying that preimages of entourages (basic neighborhoods in the uniform structure) are open and thus contain compact sets whose finite covers yield uniform bounds. A sketch of the proof for uniform continuity involves showing that for any entourage W in the uniformity of G, the preimage under multiplication is an open set in G \times G; since G \times G is compact, this preimage admits a finite cover by basic entourages, establishing the uniform property. The of any element g \in G, defined as \{ h g h^{-1} \mid h \in G \}, is the continuous image of the G under the map h \mapsto h g h^{-1}, and thus compact. Since G is Hausdorff, this image is also closed, making each a compact closed subset of G. If G is , its compactness implies that G is finite, so every is finite. Compact groups have no small subgroups, meaning there exists a neighborhood V of the e such that no nontrivial of G is contained in V. This contrasts with the local structure of noncompact groups, where neighborhoods of the identity approximate elements but do not form subgroups globally. The proof relies on : if every neighborhood contained a nontrivial subgroup, repeated generation would yield a proper closed infinite whose leads to a via finite or arguments. A key consequence of compactness is that for any neighborhood U of the identity e, the collection of left translates \{ g U \mid g \in G \} forms an open cover of G. By compactness, there exists a finite subcover, so G = \bigcup_{i=1}^n g_i U for some finite set \{ g_1, \dots, g_n \} \subset G. This finite covering property underscores the "discreteness at infinity" in compact groups, where local neighborhoods suffice to cover the entire space finitely.

Examples

Abelian compact groups

Abelian compact groups form an important subclass of compact groups, characterized by their commutative operation, which simplifies their structural analysis through tools like Pontryagin duality. These groups arise naturally in harmonic analysis and topological group theory, where their duals provide insights into discrete structures. A fundamental result is Pontryagin duality, which establishes that every compact abelian group G is topologically isomorphic to the Pontryagin dual of some discrete abelian group. Specifically, the Pontryagin dual \hat{G} of G consists of all continuous homomorphisms from G to the circle group \mathbb{T} = \mathbb{R}/\mathbb{Z}, equipped with the compact-open topology, and this dual is discrete. Conversely, the dual of a discrete abelian group is compact abelian. This duality interchanges compactness and discreteness, enabling a complete classification via the structure of discrete abelian groups, which decompose as direct sums of cyclic groups. The algebraic and topological structure of compact abelian groups reflects this duality: every such group G decomposes as a topological direct product G \cong G_0 \times D, where G_0 is the connected component of the identity (a compact connected abelian group) and D is totally disconnected. The connected part G_0 is divisible in the case of tori. Examples of G_0 include finite-dimensional tori \mathbb{T}^n, which occur when G_0 is a Lie group, and more generally solenoid groups, which are non-Lie connected compact abelian groups. The totally disconnected part D is profinite, meaning it is the inverse limit of an inverse system of finite abelian groups under continuous surjective homomorphisms. In general, compact abelian groups themselves can be viewed through this lens, with the theorem that they arise as inverse limits of finite abelian groups capturing the profinite component's role in the overall decomposition. Representative examples illustrate this structure. The n-torus \mathbb{T}^n = (\mathbb{T})^n, for finite n, is a connected compact abelian Lie group, serving as the dual of the discrete group \mathbb{Z}^n; it exemplifies the connected divisible case and appears in applications like multidimensional Fourier analysis. Profinite completions provide totally disconnected examples, such as the p-adic integers \mathbb{Z}_p for a prime p, which is the inverse limit \varprojlim \mathbb{Z}/p^n\mathbb{Z} and the dual of the Prüfer p-group \mathbb{Z}(p^\infty); this group is torsion-free and metrizable. An uncountable torsion-free example is the additive group of the p-adic integers \mathbb{Z}_p, which is homeomorphic to the Cantor set and highlights the existence of non-Lie compact abelian structures beyond finite products.

Compact Lie groups

A compact is a endowed with a compact , where a is defined as a smooth manifold G equipped with a group structure such that the multiplication map m: G \times G \to G and the inversion map i: G \to G are smooth. The compatibility between the manifold structure and the group operations ensures that the topology on G serves as both a manifold topology and a topology, making every compact a compact . This compactness imposes strong restrictions on the group's structure, distinguishing compact s from non-compact ones like \mathrm{SL}(n, \mathbb{R}). Prominent examples of compact groups include the classical series: the special orthogonal groups \mathrm{SO}(n) consisting of n \times n real orthogonal matrices with determinant 1, the special unitary groups \mathrm{SU}(n) of n \times n complex unitary matrices with determinant 1, the unitary groups \mathrm{U}(n), and the compact groups \mathrm{Sp}(n) acting as quaternionic isometries on \mathbb{H}^n. Additionally, there are five exceptional compact simple groups: G_2, F_4, E_6, E_7, and E_8, which arise from unique systems and have dimensions 14, 52, 78, 133, and 248, respectively. These groups, along with their products and finite covers like the spin groups \mathrm{Spin}(n), illustrate the diversity within this class. Compactness yields distinctive algebraic properties for these groups. On the Lie algebra \mathfrak{g}, the Killing form K(X, Y) = \operatorname{tr}(\operatorname{ad}_X \operatorname{ad}_Y) is negative definite for any nonzero X \in \mathfrak{g} when \mathfrak{g} is semisimple, providing an Ad-invariant inner product that endows \mathfrak{g} with a positive definite metric via -\langle X, Y \rangle = -K(X, Y). Furthermore, all adjoint orbits under the action \operatorname{Ad}: G \to \mathrm{GL}(\mathfrak{g}) are closed, as the image of the compact group G under the continuous adjoint map is compact and hence closed in the Hausdorff topology of \mathfrak{g}. These features facilitate the study of representations and structure, with classification relying on root systems associated to maximal tori.

Totally disconnected compact groups

A totally disconnected compact group is a compact topological group in which the connected component of the is trivial, meaning that the only connected subgroups are the trivial one. Equivalently, such a group admits a basis of neighborhoods of the consisting of open subgroups, ensuring that every neighborhood of the contains no nontrivial connected subsets. The structure of totally disconnected compact groups is captured by their identification as profinite groups, which are inverse limits of finite discrete groups. A fundamental theorem states that every totally disconnected compact group is profinite, possessing a basis of neighborhoods of the identity formed by open normal subgroups of finite index. This profinite nature implies that these groups are Stone spaces in their dual formulation, with the topology arising from the construction. Representative examples include the profinite completion of the integers, denoted \hat{\mathbb{Z}}, which is the \varprojlim_n \mathbb{Z}/n\mathbb{Z} and serves as the universal profinite quotient of \mathbb{Z}. Another key example is the general linear group \mathrm{GL}_n(\mathbb{Z}_p) over the p-adic integers \mathbb{Z}_p for a prime p, which is compact and totally disconnected as a p-adic without a nontrivial . Additionally, closed groups of locally finite trees, such as certain rigid trees, yield compact totally disconnected subgroups when restricted to fixed-point-free actions preserving the tree structure. These groups find significant applications in , particularly through their role as s, which are profinite and thus totally disconnected compact, governing the structure of algebraic extensions via the Krull topology. For instance, the of the rationals \mathrm{Gal}(\bar{\mathbb{Q}}/\mathbb{Q}) exemplifies how such structures encode infinite , with fixed fields corresponding to open normal subgroups. on these groups exists and is normalized on compact open subgroups, facilitating integration over profinite completions in .

Haar measure

Existence and uniqueness

A Haar measure on a compact group G is defined as a regular \mu on G that is left-invariant, meaning \mu(gA) = \mu(A) for all g \in G and Borel sets A \subseteq G, non-zero and finite on compact sets (with \mu(G) < \infty), and positive on non-empty open sets. This measure induces a left-invariant integral on continuous functions f: G \to \mathbb{C}, satisfying \int_G f(g) \, d\mu(g) = \int_G f(hg) \, d\mu(g) for all h \in G and integrable f. The fundamental theorem on Haar measure for compact groups states that there exists a unique (up to positive scalar multiple) left-invariant regular Borel measure \mu on G that is finite and positive on compact sets, and it can be normalized so that \mu(G) = 1, making it a probability measure. For compact groups, this normalized Haar measure is also right-invariant, hence bi-invariant. Existence follows from the Riesz representation theorem applied to the space of continuous functions C(G) on the compact group G, which is equipped with the sup norm. Since G is compact, C(G) separates points, and one constructs a positive linear functional \Lambda: C(G) \to \mathbb{C} that is left-invariant by approximating it via finite sums over group elements and using partitions of unity or mean values of translates. Specifically, for f \in C(G), define the mean value over a finite set \{a_1, \dots, a_n\} as \frac{1}{n} \sum_{i=1}^n f(ga_i), and take the limit in the uniform topology using compactness to obtain a translation-invariant functional, which represents a regular measure by Riesz. Uniqueness up to scalar multiple is established by showing that if \mu and \nu are two left-invariant regular Borel measures on G, then there exists c > 0 such that \nu = c \mu. This relies on the fact that for any continuous f \geq 0 with \int f \, d\mu = 1, the translates f_h(g) = f(h^{-1}g) span a dense , and invariance implies \int f_h \, d\nu = \int f \, d\nu for all h, so by density and continuity, \nu is a multiple of \mu. \mu(G) = 1 then fixes the constant.

Properties and normalization

One key property of the Haar measure on a compact group G is its bi-invariance: the left-invariant Haar measure \mu is also right-invariant, meaning \mu(Ag) = \mu(A) for all measurable A \subseteq G and g \in G. This follows from the fact that compact groups are unimodular, so the modular function \Delta: G \to (0, \infty) satisfies \Delta(h) = 1 for all h \in G. It is conventional to normalize the Haar measure on a compact group G such that \mu(G) = 1, making it a probability measure. For a closed subgroup H \subseteq G, the quotient space G/H inherits a unique Haar measure \nu from \mu, defined via the disintegration formula \int_G f(g) \, d\mu(g) = \int_H \left( \int_{G/H} f(hx) \, d\nu(x) \right) d\mu_H(h) for suitable integrable f: G \to \mathbb{C}, where \mu_H is the normalized Haar measure on H. This induced measure \nu is also normalized so that \nu(G/H) = 1, satisfying \mu(G) = \nu(G/H). The finite total measure enables an analog of Fubini's theorem for products: on G \times G equipped with the \mu \times \mu, integrals of measurable functions f: G \times G \to \mathbb{C} satisfy \int_{G \times G} f(g_1, g_2) \, d(\mu \times \mu)(g_1, g_2) = \int_G \left( \int_G f(g_1, g_2) \, d\mu(g_2) \right) d\mu(g_1) whenever the iterated integrals exist. Bi-invariance implies a simple change-of-variables formula. For a right translate, the general relation for left Haar measures is \int_G f(g) \, d\mu(g) = \int_G f(gh) \Delta(h)^{-1} \, d\mu(g) for integrable f: G \to \mathbb{C} and h \in G. In compact groups, \Delta \equiv 1, so this reduces to \int_G f(g) \, d\mu(g) = \int_G f(gh) \, d\mu(g), confirming right invariance directly. The finiteness of \mu(G) ensures that the L^p spaces on G are well-behaved: for $1 \leq p < q \leq \infty, L^q(G) \subseteq L^p(G) with continuous inclusion, and the dual of L^p(G) is L^{p'}(G) where $1/p + 1/p' = 1. This structure underpins harmonic analysis on compact groups, facilitating decompositions like the .

Structure of compact groups

General structure theorem

The general structure theorem for compact groups, primarily due to the work of , , , and , characterizes their algebraic and topological form in terms of Lie and profinite components. Specifically, every connected compact Hausdorff group is a . This resolves the compact case of , affirming that connectedness and compactness suffice for the group to admit a compatible Lie group structure, with smooth manifold topology and Lie algebra. More globally, every compact group arises as an extension of a compact by a profinite group: there exists a closed normal compact Lie subgroup L (the connected component of the identity) such that the quotient G/L is a totally disconnected compact group, hence profinite. This structure implies that compact groups are pro-Lie groups, meaning they are inverse limits of Lie groups. To see this, given any neighborhood U of the identity in a compact Hausdorff group G, there exists a compact normal subgroup H \subseteq U such that G/H is a Lie group (in fact, linear over \mathbb{C}). Iterating over a basis of neighborhoods yields a system of surjective homomorphisms from G onto Lie groups with kernels forming a basis of neighborhoods, establishing the inverse limit description. Profinite groups themselves fit as the totally disconnected case, being inverse limits of finite discrete groups. The proof outline leverages the no small subgroups (NSS) property: a topological group has NSS if there exists a neighborhood of the identity containing no nontrivial proper subgroup. Locally compact groups with NSS are precisely the Lie groups. For compact G, the Peter–Weyl theorem provides faithful finite-dimensional unitary representations, allowing linearization and approximation. One constructs open normal subgroups by quotienting out small kernels where the image inherits NSS (via metric approximations and continuity arguments), ensuring the quotients are Lie; the compactness ensures the kernels are compact and normal. This approximation process yields the Lie-by-profinite extension and the pro-Lie inverse limit. A key corollary is that the dimension of a compact group G is well-defined as the dimension of its maximal connected Lie subgroup (the connected component of the identity), which coincides with the dimension of the associated . This dimension is invariant under the approximations and finite quotients in the structure theorem, providing a measure of the "Lie part" even for infinite-dimensional profinite extensions. For example, the additive group of p-adic integers \mathbb{Z}_p has dimension 0, as it is profinite with trivial connected component.

Structure of compact Lie groups

Compact Lie groups exhibit a canonical decomposition that separates their abelian and semisimple components. For a connected compact Lie group G with Lie algebra \mathfrak{g}, the Lie algebra decomposes as \mathfrak{g} = \mathfrak{z}(\mathfrak{g}) \oplus [\mathfrak{g}, \mathfrak{g}], where \mathfrak{z}(\mathfrak{g}) is the center of \mathfrak{g} (an abelian Lie algebra) and [\mathfrak{g}, \mathfrak{g}] is semisimple. At the group level, G = Z(G)^0 \times [G, G], where Z(G)^0 denotes the connected component of the identity in the center Z(G) of G (a torus) and [G, G] is the commutator subgroup, a closed connected semisimple Lie subgroup with finite center. In the semisimple case, where Z(G)^0 is trivial, the adjoint representation of G is faithful, with kernel precisely Z(G), which is finite. Consequently, the adjoint form G / Z(G) has trivial center and acts faithfully via the adjoint representation; if the Lie algebra of G is simple, then G / Z(G) is a simple Lie group. Simply connected compact Lie groups admit a product decomposition into a torus and a product of simple simply connected compact Lie groups, reflecting the direct sum structure of their semisimple Lie algebras. This structure underscores the reductive nature of compact Lie algebras and facilitates the study of representations and homomorphisms.

Classification and geometry of compact Lie groups

Classification by rank and type

The classification of simple compact Lie algebras over the real numbers, which underpin the structure of simple compact Lie groups, divides them into four infinite families of classical types and five exceptional types, as established by the work of Killing and Cartan. The classical families are A_n (corresponding to the special unitary Lie algebra su(n+1)), B_n (odd orthogonal so(2n+1)), C_n (compact symplectic sp(n)), and D_n (even orthogonal so(2n)), while the exceptional families are G_2, F_4, E_6, E_7, and E_8. The rank of a simple compact Lie algebra is the dimension of its , or equivalently, the dimension of the maximal torus in the corresponding . For the classical types , , , and , the rank is n (with n ≥ 1 for A_n, n ≥ 2 for B_n, n ≥ 3 for C_n to exclude isomorphisms C_1 ≅ A_1 and C_2 ≅ B_2, and n ≥ 4 for D_n to exclude D_2 ≅ A_1 × A_1 (not simple) and D_3 ≅ A_3). The exceptional types have fixed ranks: 2 for , 4 for , 6 for , 7 for , and 8 for . This distinction between classical (infinite families tied to matrix groups) and exceptional (finite, non-matrix-like) types highlights the organizational structure of the classification. Low-rank isomorphisms include A_1 ≅ B_1 ≅ C_1 and B_2 ≅ C_2. Each simple complex semisimple Lie algebra admits a unique compact real form up to isomorphism, ensuring that the compact Lie groups associated with these algebras are determined uniquely by their underlying complex structure, modulo covering groups. For instance, the Lie algebra su(2) of type A_1 is isomorphic to so(3) of type B_1, corresponding to the groups SU(2) and Spin(3), which are double covers of SO(3). The following table summarizes the types, associated Lie algebras, ranks, dimensions (of the Lie algebra), and representative simply connected groups (noting low-rank isomorphisms: A_1 ≅ B_1 ≅ C_1 ≅ su(2); B_2 ≅ C_2; D_3 ≅ A_3):
TypeLie AlgebraRankDimensionRepresentative Group
A_nsu(n+1)nn(n+2)SU(n+1) (e.g., SU(2) for n=1)
B_nso(2n+1)nn(2n+1)Spin(2n+1) (e.g., Spin(5) for n=2)
C_nsp(n)nn(2n+1)Sp(n) (e.g., Sp(3) for n=3)
D_nso(2n)nn(2n-1)Spin(2n) (e.g., Spin(6) for n=3)
G_2g_2214G_2
F_4f_4452F_4
E_6e_6678E_6
E_7e_77133E_7
E_8e_88248E_8

Maximal tori and root systems

In a compact connected Lie group G, a maximal torus is defined as a maximal connected abelian subgroup T \subseteq G. The dimension of such a T equals the rank of G, which is the dimension of a maximal abelian subalgebra in the Lie algebra \mathfrak{g} of G. Every element of G lies in some maximal torus, and all maximal tori in G are conjugate under the action of G. The normalizer N_G(T) of a maximal torus T in G is the set of elements g \in G such that g T g^{-1} = T, and the centralizer C_G(T) coincides with T itself. The quotient N_G(T)/C_G(T) \cong N_G(T)/T forms the Weyl group W of G with respect to T, which acts on T and plays a key role in the structure of G. At the Lie algebra level, let \mathfrak{t} be the Lie algebra of T, and consider the complexified Lie algebra \mathfrak{g}_\mathbb{C} = \mathfrak{g} \otimes \mathbb{C} with Cartan subalgebra \mathfrak{h} = \mathfrak{t}_\mathbb{C}. The adjoint representation of T on \mathfrak{g} exponentiates to a diagonalizable action on \mathfrak{g}_\mathbb{C}, yielding a root system \Phi \subset \mathfrak{h}^*, where the roots \alpha \in \Phi are the nonzero linear functionals on \mathfrak{h} such that the root spaces \mathfrak{g}_\alpha = \{ X \in \mathfrak{g}_\mathbb{C} \mid \mathrm{ad}(H)X = \alpha(H)X \ \forall H \in \mathfrak{h} \} are nonzero. The roots \Phi lie in the real subspace i \mathfrak{t}^* \subset \mathfrak{h}^*. The root space decomposition of \mathfrak{g}_\mathbb{C} is given by \mathfrak{g}_\mathbb{C} = \mathfrak{h} \oplus \bigoplus_{\alpha \in \Phi} \mathfrak{g}_\alpha, where each \mathfrak{g}_\alpha is one-dimensional for semisimple G. This decomposition reflects the semisimple structure of \mathfrak{g} and underpins the geometric properties of G. A choice of Borel subgroup B \subseteq G, which is a maximal connected solvable subgroup containing T, induces a partial ordering on \Phi by selecting a set of positive roots \Phi^+, consisting of those roots that are positive with respect to a suitable Weyl chamber in \mathfrak{t}. The simple roots \Delta \subset \Phi^+ form a basis for the real span of \Phi such that every root in \Phi is an integer linear combination of elements of \Delta, with coefficients nonnegative for roots in \Phi^+ and nonpositive for those in -\Phi^+. This choice of Borel subgroup and associated simple roots provides a fundamental datum for analyzing the geometry and representations of G.

Weyl groups and Dynkin diagrams

In the theory of compact Lie groups, the Weyl group W associated to a root system \Phi with simple roots \Delta is the finite subgroup of the orthogonal group on the real vector space spanned by \Phi, generated by the reflections s_\alpha for each \alpha \in \Delta, where the reflection s_\alpha acts on a vector \lambda by s_\alpha(\lambda) = \lambda - \langle \lambda, \alpha^\vee \rangle \alpha. These reflections satisfy s_\alpha^2 = id and generate W as a Coxeter group with presentation determined by the angles between simple roots. The group W is finite because it preserves the root system and acts faithfully on it. The Weyl group W acts orthogonally on the Cartan subalgebra \mathfrak{t} (the real span of a maximal torus), preserving the inner product, and permutes the roots via w(\beta) = \beta - 2 \frac{(\beta, \alpha)}{(\alpha, \alpha)} \alpha for w = s_\alpha and roots \beta. This action extends to conjugation on the , reflecting the normalizer structure W \cong N_G(T)/T, where T is a maximal torus and N_G(T) its normalizer in the compact group G. Among the elements of W, the longest element w_0 is the unique one of maximal length (minimal number of simple reflections in its decomposition) that maps the positive root system \Phi^+ to the negative roots \Phi^-, satisfying w_0(\lambda) = -\lambda on the weight lattice for dominant weights \lambda. The structure of W is encoded combinatorially by Dynkin diagrams, which classify the irreducible finite root systems up to isomorphism and thus the semisimple compact Lie groups up to local isomorphism. A Dynkin diagram is a graph with vertices corresponding to the simple roots \alpha_i \in \Delta, and edges between vertices i and j determined by the Cartan integers a_{ij} = \langle \alpha_i, \alpha_j^\vee \rangle = 2 \frac{(\alpha_i, \alpha_j)}{(\alpha_j, \alpha_j)}, which are integers satisfying a_{ii} = 2, a_{ij} \leq 0 for i \neq j, and a_{ij} a_{ji} \in \{0,1,2,3\}. Specifically, no edge if a_{ij} = 0 (orthogonal roots), a single undirected edge if a_{ij} = a_{ji} = -1 (120° angle), a double edge (with arrow from longer to shorter root) if |a_{ij}| = 2 or |a_{ji}| = 2 but not both (135° or 150° angles), and a triple edge for a_{ij} = -3, a_{ji} = -1 (as in type G_2). These integers are defined intrinsically for compact connected semisimple Lie groups via degrees of maps in the fundamental group or intersection numbers in the normalizer, without reference to the Lie algebra. The connected Dynkin diagrams of finite type are precisely those without cycles, multiple edges beyond the specified, or subdiagrams of extended type, yielding the classical series A_n (linear chain of n vertices, n \geq 1), B_n (linear with short double arrow at end, n \geq 2), C_n (linear with long double arrow at end, n \geq 3), D_n (linear forking into two at end, n \geq 4), and the exceptional types E_6, E_7, E_8 (branched trees with 6, 7, 8 vertices), F_4 (linear with double and triple segments), and G_2 (two vertices with triple arrow). The order of W for an irreducible root system is the product of the degrees of the basic polynomial invariants of W acting on \mathfrak{t}; for example, in type A_n these degrees are $2, 3, \dots, n+1, yielding |W| = (n+1)! (isomorphic to the symmetric group S_{n+1}), while for B_n or C_n the degrees $2, 4, \dots, 2n give |W| = 2^n n!.

Topology of compact Lie groups

Connected compact Lie groups decompose as a product of a torus (the connected component of the center) and a semisimple connected compact Lie group; the following focuses on the semisimple case, where \pi_1(G) is finite.

Fundamental group

The fundamental group \pi_1(G) of a connected semisimple compact Lie group G is finite. This finiteness follows from the topological structure of compact Lie groups, where the exponential map from the Lie algebra to the group induces a covering that reveals \pi_1(G) as a discrete subgroup of finite order. For a connected semisimple compact Lie group G with Lie algebra \mathfrak{g} and Cartan subalgebra \mathfrak{h}_\mathbb{R}, \pi_1(G) is isomorphic to the quotient of the integral lattice \Gamma_I = \{ X \in \mathfrak{h}_\mathbb{R} \mid \exp(2\pi i X) = e \} by the coroot lattice \Gamma_C = \mathrm{span}_\mathbb{Z} \{ \tau_{\alpha_i} \mid \alpha_i \in F \}, where F is a basis of simple coroots. Equivalently, \pi_1(G) can be described as the cokernel of the inclusion of the coroot lattice into the cocharacter lattice of a maximal torus, yielding a finite abelian group. The order of \pi_1(G) equals the index [P : X(T)], where P is the weight lattice and X(T) is the character lattice of a maximal torus T \subset G. Every connected semisimple compact Lie group G admits a finite-sheeted universal covering map \hat{G} \to G from a simply connected compact Lie group \hat{G}, with the kernel of this homomorphism being a finite central discrete subgroup isomorphic to \pi_1(G). This covering encodes the topological structure of G, distinguishing it from its universal cover while preserving the Lie algebra isomorphism \mathfrak{g} \cong \hat{\mathfrak{g}}. Representative examples illustrate these properties. The special unitary group SU(n) for n \geq 2 is simply connected, so \pi_1(SU(n)) = 0. In contrast, the special orthogonal group SO(n) for n \geq 3 has \pi_1(SO(n)) = \mathbb{Z}/2\mathbb{Z}, with the spin group \mathrm{Spin}(n) serving as its simply connected double cover. The symplectic group \mathrm{Sp}(n) is also simply connected, yielding \pi_1(\mathrm{Sp}(n)) = 0.

Center and universal cover

The center Z(G) of a compact Lie group G consists of all elements g \in G that commute with every element of G, forming a closed normal subgroup that is itself a compact Lie group. For connected compact Lie groups, if G is abelian, then Z(G) = G, which is a torus; if G is semisimple, then Z(G) is finite. In general, Z(G) can be computed as the kernel of the adjoint representation \mathrm{Ad}: G \to \mathrm{Aut}(\mathfrak{g}), where \mathfrak{g} is the Lie algebra of G, or equivalently as the intersection of the kernels \bigcap_{g \in G} \ker(\mathrm{Ad}_g). A concrete example is the special unitary group \mathrm{SU}(n), whose center is the cyclic group of order n consisting of scalar matrices e^{2\pi i k / n} I_n for k = 0, \dots, n-1. For connected compact Lie groups with finite fundamental group (e.g., semisimple ones), the universal cover \tilde{G} is also compact and simply connected, with its center Z(\tilde{G}) being a discrete subgroup such that Z(\tilde{G}) = \pi^{-1}(Z(G)), where \pi: \tilde{G} \to G is the covering map. The quotient by the centers yields isomorphic adjoint groups: \tilde{G} / Z(\tilde{G}) \cong G / Z(G). For semisimple cases, if \tilde{G} is the simply connected cover, then Z(\tilde{G}) is finite and isomorphic to the quotient of the weight lattice P by the root lattice Q of the root system associated to \mathfrak{g}, i.e., Z(\tilde{G}) \cong P / Q. More generally, for a semisimple compact Lie group G, Z(G) is a finite quotient of P / Q, reflecting the choice of covering corresponding to a subgroup of the fundamental group. In terms of dual lattices, this can be expressed using the coweight lattice P^\vee and coroot lattice Q^\vee, where the isomorphism aligns with the pairing between weights and coroots.

Representation theory

Peter–Weyl theorem

The Peter–Weyl theorem establishes the foundation for the representation theory of compact groups by providing a complete orthogonal decomposition of the Hilbert space of square-integrable functions on the group. For a compact topological group G equipped with its unique normalized \mu, the space L^2(G) decomposes as an orthogonal direct sum over all equivalence classes of finite-dimensional irreducible unitary representations \pi of G: L^2(G) = \bigoplus_\pi \left( V_\pi^* \otimes V_\pi \right), where V_\pi is the representation space of \pi, and the summands correspond to the spaces of matrix coefficients of \pi. Choosing an orthonormal basis \{e_i\} for V_\pi, the normalized matrix coefficients u_{ij}^\pi(g) = \sqrt{\dim V_\pi} \langle \pi(g) e_j, e_i \rangle, \quad 1 \leq i,j \leq \dim V_\pi, over all such \pi, i, and j, form a complete orthonormal basis for L^2(G). This decomposition implies that every function in L^2(G) can be uniquely expanded as an infinite linear combination of these matrix coefficients. A key component of the theorem is the orthogonality of characters, where the character \chi_\pi(g) = \operatorname{tr} \pi(g) of an irreducible representation \pi satisfies \int_G \overline{\chi_\pi(g)} \chi_\sigma(g) \, d\mu(g) = \delta_{\pi \sigma} for irreducible unitary representations \pi and \sigma. This relation extends Schur orthogonality from finite groups to the compact case and follows directly from the inner product structure on matrix coefficients. Among its consequences, the theorem ensures that the left regular action of G on L^2(G) by translation is unitary, as the matrix coefficients transform appropriately under this action. Additionally, the finite-dimensional irreducible representations separate points on G, meaning that for any distinct g, h \in G, there exists an irreducible representation \pi such that \pi(g) \neq \pi(h); consequently, the algebra generated by matrix coefficients is dense in the continuous functions C(G) with respect to the uniform norm. The proof relies on averaging operators to project onto isotypic components in the regular representation and invokes Schur's lemma to establish orthogonality between distinct irreducibles, with completeness following from density arguments using the Peter–Weyl approximation property for continuous functions.

Unitary representations and orthogonality

For a compact group G equipped with a bi-invariant Haar measure \mu, every continuous finite-dimensional representation \pi: G \to \mathrm{GL}(V) on a complex vector space V is equivalent to a unitary representation. This equivalence is achieved by defining a G-invariant inner product on V via averaging: \langle v, w \rangle = \int_G \langle \pi(g)v, \pi(g)w \rangle_0 \, d\mu(g), where \langle \cdot, \cdot \rangle_0 is any inner product on V; the resulting representation preserves this inner product, making \pi unitary. Given an orthonormal basis \{e_i\} for the Hilbert space of a unitary representation \pi, the matrix coefficients are the continuous functions u^\pi_{ij}: G \to \mathbb{C} defined by u^\pi_{ij}(g) = \langle \pi(g) e_j, e_i \rangle. These functions form the building blocks for the L^2-decomposition of G and play a central role in the , which establishes an orthonormal basis of matrix coefficients for irreducible representations. The Schur orthogonality relations quantify the inner products of these coefficients. For two unitary irreducible representations \pi and \sigma of dimensions d_\pi and d_\sigma, respectively, with matrix coefficients u^\pi_{ij} and u^\sigma_{kl}, \int_G \overline{u^\pi_{ij}(g)} u^\sigma_{kl}(g) \, d\mu(g) = \delta_{\pi\sigma} \delta_{i k} \delta_{j l}, where \delta denotes the and the is normalized so that \mu(G) = 1. If \pi \not\cong \sigma, the integral vanishes, ensuring orthogonality between distinct irreducibles; within the same representation, it yields the stated normalization. These relations, originally derived for , hold more generally for and underpin the completeness of the representation theory. The linear span of all matrix coefficients from finite-dimensional unitary representations is dense in the space C(G) of continuous complex-valued functions on G with the uniform norm. This density follows from the Peter–Weyl theorem and implies that the coefficients separate points on G, providing a Fourier-like analysis for non-abelian compact groups.

Irreducible representations

In the representation theory of compact groups, the irreducible unitary representations exhibit several fundamental properties that distinguish them from those of non-compact groups. A continuous unitary representation of a compact group G on a Hilbert space is called irreducible if there are no closed proper invariant subspaces. Unlike in the non-compact case, every irreducible unitary representation \pi of G is finite-dimensional. This finiteness arises from the compactness of G, which implies that the image \pi(G) is a compact subgroup of the unitary group, and thus the representation cannot sustain infinite-dimensional irreducibility. A key consequence is the complete reducibility of all unitary representations. Every continuous unitary representation of a compact group G decomposes as an orthogonal direct sum of finite-dimensional irreducible unitary representations. This decomposition is unique up to ordering and isomorphisms, allowing any such representation to be expressed as \rho \cong \bigoplus_{\pi \in \widehat{G}} m_{\pi} \pi, where \widehat{G} denotes the set of equivalence classes of irreducible unitary representations, and m_{\pi} is the multiplicity of \pi in \rho. Schur's lemma provides a characterization of the endomorphisms of an irreducible representation. For an irreducible unitary representation \pi: G \to U(H) of a compact group G on a Hilbert space H, the algebra of bounded linear operators on H that commute with \pi(g) for all g \in G consists precisely of the scalar multiples of the identity operator. In other words, \{ T \in B(H) \mid T \pi(g) = \pi(g) T \ \forall g \in G \} = \mathbb{C} I. This result follows from the unitarity and irreducibility, ensuring that the commutant is one-dimensional over \mathbb{C}. The multiplicities m_{\pi} in the decomposition can be computed using characters, leveraging the orthogonality relations for irreducible representations. The character of a representation \rho is the function \chi_{\rho}(g) = \operatorname{tr}(\rho(g)), and for compact G equipped with its normalized Haar measure \mu, the multiplicity of an irreducible \pi in \rho is given by m_{\pi} = \int_G \chi_{\rho}(g) \overline{\chi_{\pi}(g)} \, d\mu(g). This formula stems from the fact that the irreducible characters \{\chi_{\pi} \mid \pi \in \widehat{G}\} form an orthonormal set in L^2(G) with respect to the inner product \langle f, h \rangle = \int_G f(g) \overline{h(g)} \, d\mu(g). Specifically, \langle \chi_{\pi}, \chi_{\sigma} \rangle = \delta_{\pi \sigma}, which directly yields the multiplicity as the inner product \langle \chi_{\rho}, \chi_{\pi} \rangle. The irreducible unitary representations of a compact group are in one-to-one correspondence with their characters. Two irreducible unitary representations \pi and \sigma are equivalent if and only if \chi_{\pi} = \chi_{\sigma}. Moreover, the characters are class functions, meaning \chi_{\pi}(g) = \chi_{\pi}(hgh^{-1}) for all g, h \in G, so they are constant on conjugacy classes and thus parametrize the irreducibles via their values on these classes. This correspondence underpins the Peter–Weyl theorem, which asserts that the matrix coefficients of these representations form an orthonormal basis for L^2(G).

Representation theory of connected compact Lie groups

Representations of maximal tori

In compact connected Lie groups, maximal tori are abelian subgroups isomorphic to (U(1))^r, where r is the rank of the group. The characters of a maximal torus T form the character group \operatorname{Hom}(T, U(1)), which is isomorphic to the integer lattice \mathbb{Z}^r. This group, known as the weight lattice and denoted \Lambda, consists of all continuous homomorphisms from T to the circle group U(1). The irreducible representations of T are one-dimensional and are precisely the characters of T; each such representation is labeled by a weight \lambda \in \Lambda. For t \in T expressed in coordinates t = (e^{2\pi i \theta_1}, \dots, e^{2\pi i \theta_r}) with \theta = (\theta_1, \dots, \theta_r) \in \mathbb{R}^r / \mathbb{Z}^r, the corresponding character is given by \chi_\lambda(t) = \prod_{j=1}^r e^{2\pi i \langle \lambda, \theta_j \rangle} = e^{2\pi i \langle \lambda, \theta \rangle}, where \langle \cdot, \cdot \rangle denotes the standard pairing between the weight lattice and the cocharacter lattice. To classify representations of the full Lie group, the weights are analyzed relative to the root system. The fundamental Weyl chamber is an open cone in the real span of \Lambda defined by the inequalities \langle \lambda, \alpha^\vee \rangle > 0 for all positive coroots \alpha^\vee corresponding to a choice of positive . Its closure contains the dominant weights, which are the elements \lambda \in \Lambda satisfying \langle \lambda, \alpha^\vee \rangle \geq 0 for all positive coroots \alpha^\vee. Each Weyl group orbit in the weight lattice intersects the closure of the fundamental Weyl chamber in exactly one dominant weight, providing the labels for highest weight irreducible representations in highest weight .

Induced representations and branching

In the representation theory of connected compact Lie groups, finite-dimensional irreducible representations are constructed by inducing one-dimensional representations (characters) of a maximal torus T to the full group G via a Borel subgroup B containing T. Let \pi: T \to \mathbb{C}^\times be a character of T, corresponding to a weight \lambda \in \mathfrak{t}^*, extended trivially to the unipotent radical N of B to yield a representation of B. The induced representation \operatorname{Ind}_B^G \pi acts on the space of smooth functions f: G \to \mathbb{C} satisfying the twisted equivariance condition f(gb) = \pi(b)^{-1} f(g) for all g \in G and b \in B, with G acting by left translation: (g \cdot f)(x) = f(g^{-1} x). This construction realizes \operatorname{Ind}_B^G \pi as the space of global sections of the line bundle L_\lambda over the flag variety G/B, associated to the principal B-bundle G \to G/B. By the Borel–Weil theorem, for a dominant \lambda (i.e., \langle \lambda, \alpha^\vee \rangle \geq 0 for all positive coroots \alpha^\vee), the space of global holomorphic sections H^0(G/B, L_{-\lambda}) is isomorphic to the dual of the V_\lambda with highest \lambda, providing a concrete realization of every of G. For non-dominant \lambda, the decomposes into a of irreducibles corresponding to the orbit of \lambda, shifted by the Weyl vector. This process aligns with the Bruhat decomposition G = B W B (with W the ), facilitating the analysis of sections over the flag variety. Representations of maximal tori, which decompose into of one-dimensional characters labeled by the lattice, serve as the starting point for this . Branching rules describe the decomposition of an \sigma of G upon restriction to a closed K \subset G, denoted \operatorname{Res}_G^K \sigma = \bigoplus m_i \tau_i, where \{\tau_i\} are irreducibles of K and m_i are multiplicities. For K = T a , \operatorname{Res}_G^T \sigma decomposes into spaces V_\mu of dimension equal to the multiplicity m_\mu of the \mu, each carrying a one-dimensional of T; the \mu lie in the of the orbit of the highest , with m_\mu computed via combinatorial formulas like Kostant's multiplicity . In general, for closed K \subset G, branching is governed by the geometry of the embedding K \hookrightarrow G and involves finite multiplicities since K is compact. These decompositions are crucial for reducing representations to , such as in symmetric spaces or dual pairs. Frobenius reciprocity establishes a duality between and restriction, adapted to the torus setting via the W = N_G(T)/T. For a \pi of T (extended to B) and an irreducible \sigma of G, the multiplicity equals \langle \operatorname{Ind}_B^G \pi, \sigma \rangle_G = \langle \pi, \operatorname{Res}_G^T \sigma \rangle_{T}^W, where the right-hand inner product is the W- ( over W): \langle \pi, \eta \rangle_{T}^W = \frac{1}{|W|} \sum_{w \in W} \langle \pi, w \cdot \eta \rangle_T, with integrals over T using the . This holds because irreducible characters of G are W- functions on T, ensuring the reciprocity captures the in weight decompositions. For general closed subgroups H \subset G, the untwisted form \langle \operatorname{Ind}_H^G \rho, \psi \rangle_G = \langle \rho, \operatorname{Res}_G^H \psi \rangle_H applies directly, linking dimensions and multiplicities across levels. Multiplicities in induced representations and their branching decompositions are quantified by the Weyl dimension formula, which gives the dimension of the irreducible module V_\lambda with highest weight \lambda: \dim V_\lambda = \prod_{\alpha > 0} \frac{\langle \lambda + \rho, \alpha \rangle}{\langle \rho, \alpha \rangle}, where the product runs over positive \alpha and \rho is the Weyl vector (half-sum of positive ). This formula arises as the leading term in the character expansion and previews the full by encoding the "volume" of the weight under W-action; for instance, it determines the total multiplicity of T-characters in \operatorname{Res}_G^T V_\lambda, as \dim V_\lambda = \sum_\mu m_\mu. In branching to general K, multiplicities m_i satisfy \dim \sigma = \sum m_i \dim \tau_i, with explicit computation often requiring case-by-case analysis via Littlewood-Richardson coefficients for classical groups.

Weyl character formula

The Weyl character formula provides an explicit expression for the character of any of a connected compact , parameterized by a dominant highest weight λ in the weight lattice. This formula, derived from the structure of the of the group's , allows computation of characters without constructing the explicitly and plays a central role in the of such groups. Let G be a connected compact with \mathfrak{g}, T \subset G with \mathfrak{t}, W = N_G(T)/T, and \Delta \subset \mathfrak{t}^* relative to a choice of positive roots \Delta^+ \subset \Delta. For a dominant integral weight \lambda \in \mathfrak{t}^*, the character \chi_\lambda of the corresponding , restricted to T, is given by \chi_\lambda(t) = \frac{\sum_{w \in W} \varepsilon(w) \, e^{\langle w(\lambda + \rho), \log t \rangle}}{\sum_{w \in W} \varepsilon(w) \, e^{\langle w\rho, \log t \rangle}}, where \rho \in \mathfrak{t}^* is the half-sum of the positive roots \rho = \frac{1}{2} \sum_{\alpha \in \Delta^+} \alpha, \varepsilon(w) is the sign of the element w, \langle \cdot, \cdot \rangle denotes the pairing \mathfrak{t}^* \times \mathfrak{t} \to \mathbb{R}, and \log: T \to \mathfrak{t} is the logarithm (defined locally near the identity and extended analytically). This expression extends to all of G by the property of characters and , as conjugacy classes intersect T densely. The formula is valid precisely when \lambda is dominant integral, ensuring the representation is finite-dimensional and irreducible. The denominator in the formula is the Weyl denominator function, which admits a product expression over the positive roots: \sum_{w \in W} \varepsilon(w) \, e^{\langle w\rho, \log t \rangle} = \prod_{\alpha \in \Delta^+} \left( e^{\langle \alpha, \log t \rangle} - 1 \right) in a suitable formal normalization, or more commonly in trigonometric form as \prod_{\alpha \in \Delta^+} 2 \sinh \left( \frac{1}{2} \langle \alpha, \log t \rangle \right) when identifying T with a quotient of \mathbb{R}^r via the exponential map. This product form highlights the vanishing of the denominator at the identity (corresponding to the augmentation), reflecting the trace-zero property for non-trivial representations. Equivalently, the Weyl character formula can be expressed using alternants (determinantal forms) with respect to a basis of the weight space. Fixing an orthonormal basis \{ \mu_1, \dots, \mu_r \} for \mathfrak{t}^* adapted to the roots, the numerator is the determinant \sum_{w \in W} \varepsilon(w) \, e^{\langle w(\lambda + \rho), \log t \rangle} = \det \left( e^{\langle (\lambda + \rho)_{\sigma(i)}, \log t \rangle} \right)_{1 \leq i,j \leq r}, where the (\lambda + \rho)_i are the coordinates of \lambda + \rho in the basis and \sigma runs over permutations; the full character on T is then the ratio of such determinants for \lambda + \rho and \rho. This determinant representation underscores the antisymmetric nature of the Weyl group action and facilitates computations in specific cases.

Example: SU(2)

The SU(2) provides a concrete illustration of the for compact Lie groups of rank one, with its of type A₁ and isomorphic to ℤ/2ℤ. This group is the simply connected compact form associated to the su(2), where the consists of a single pair of roots ±α, and the action reflects the reflection across the perpendicular to α. The irreducible representations of SU(2) are labeled by dominant weights corresponding to non-negative half-integers j = 0, 1/2, 1, 3/2, ..., each yielding a finite-dimensional unitary of 2j + 1. These representations arise from highest weights mα/2 with m = 2j an even non-negative in the coroot , ensuring integrality for the simply connected group. Applying the to these representations, the character of the irrep with label j, evaluated on a group element g = diag(e^{iθ/2}, e^{-iθ/2}) in a , is given by \chi_j(\theta) = \frac{\sin\left(\left(j + \frac{1}{2}\right)\theta\right)}{\sin\left(\frac{\theta}{2}\right)}. This formula captures the trace of the representation matrix, which simplifies due to the rank-one structure and alternates under the Weyl group action of sign change on the weight. A notable example is the adjoint representation of SU(2), which corresponds to the spin-1 irrep (j=1) and realizes the double cover SU(2) → SO(3), where it induces the standard 3-dimensional representation of the rotation group SO(3). This connection highlights how half-integer spins in SU(2) project to integer spins in SO(3), with the adjoint action preserving the Lie algebra structure.

Proof outline of character formula

The Weyl character formula expresses the character \chi_\lambda of the irreducible representation V_\lambda of a compact connected semisimple G with highest weight \lambda as \chi_\lambda(g) = \frac{\sum_{w \in W} \varepsilon(w) e^{w(\lambda + \rho)}}{\sum_{w \in W} \varepsilon(w) e^{w \rho}}, where W is the , \varepsilon(w) is the sign of w, and \rho is half the sum of the positive . In the algebraic approach using highest weight theory, the irreducible module V_\lambda is constructed as the quotient of the M_\lambda by its unique maximal proper submodule, where the is the induced module from the with highest weight \lambda. This construction ensures V_\lambda is finite-dimensional for dominant integral \lambda, with weights forming a W-orbit under the action of lowering operators corresponding to negative roots. The weight multiplicities m_\mu in V_\lambda, which determine the character as the formal \mathrm{ch}(V_\lambda) = \sum_\mu m_\mu e^\mu, are given by Kostant's formula: m_\mu = \sum_{w \in W} \varepsilon(w) P(w(\lambda + \rho) - (\mu + \rho)), where P is the Kostant partition function counting the number of ways to write a weight as a non-negative combination of positive roots. This alternating over the accounts for the relations imposed by the submodule, ensuring multiplicities are non-negative s. To compute the character explicitly as a formal sum, one may use Freudenthal's recursion formula, which relates m_\mu to multiplicities of lower weights via half the sum of adjacent root differences, or the Bernstein-Gelfand-Gelfand (BGG) resolution, a projective resolution of V_\lambda by that alternates over the to yield the via an . These methods confirm the is a in the exponentials e^\alpha for roots \alpha, invariant under W. An analytic proof proceeds via the on the flag variety G/T or the Atiyah-Bott fixed-point index theorem applied to the Borel-Weil-Bott , where V_\lambda appears as the of a L_\lambda on G/T, with the given by localization at the |W| fixed points (the Weyl orbits). The index theorem yields the alternating sum directly as the contribution from these points, matching the denominator's vanishing order. The proof unfolds in four main steps. First, for the maximal torus case (G = T), the representation is one-dimensional with trivial character e^\lambda, and Weyl group action is absent. Second, for general G, induce the one-dimensional representation from the torus via the Borel subgroup to obtain the full module, whose character is the induced character formula involving the torus restriction. Third, apply Weyl's integration formula to reduce integrals over G to integrals over T: \int_G f(g) \, dg = \frac{1}{|W|} \int_T f(t) |\delta(t)|^2 \, dt, where \delta(t) = \prod_{\alpha \in R^+} 2 \sin(\alpha(t)/2), allowing computation of inner products and orthogonality for characters. Fourth, form the alternating sum \sum_{w \in W} \varepsilon(w) \chi_\lambda(w t) to project onto the W-anti-invariant part, which equals the numerator times the denominator, verifying the formula by uniqueness of such rational functions with prescribed poles.

Duality

Pontryagin duality for abelian cases

For a compact G, the Pontryagin dual \hat{G} is defined as the set of all continuous group homomorphisms from G to the circle group \mathbb{T} = \mathbb{R}/\mathbb{Z}, equipped with the , which makes \hat{G} a locally compact . This topology ensures that \hat{G} is Hausdorff and turns the pointwise multiplication of characters into a continuous group operation. A fundamental result is the Pontryagin duality theorem, which asserts that G is topologically isomorphic to its double dual \hat{\hat{G}} via the evaluation map ev: G \to \hat{\hat{G}} given by ev(g)(\chi) = \chi(g) for g \in G and \chi \in \hat{G}. For compact abelian G, the dual \hat{G} is , reflecting the duality between and discreteness in the category of locally compact abelian groups. A classic example is the circle group \mathbb{T}, whose dual is the integers \mathbb{Z} with the topology, where characters are given by \chi_n(z) = z^n for z \in \mathbb{T} and n \in \mathbb{Z}. Every compact G is topologically isomorphic to a C \times P, where C is a compact connected abelian group and P is a totally disconnected compact abelian group. The is then \hat{G} \cong D_{tf} \times D_t, where D_{tf} is a discrete torsion-free abelian group and D_t is a discrete torsion abelian group. This structure highlights how Pontryagin duality interchanges connected components with torsion-free discrete factors and totally disconnected parts with torsion discrete groups. Pontryagin duality underpins Fourier analysis on compact abelian groups by identifying functions on G with their Fourier transforms on \hat{G}. Specifically, the characters in \hat{G} form an orthonormal basis with respect to the Haar measure on G, allowing the Fourier series expansion of integrable functions f: G \to \mathbb{C} as f(g) = \sum_{\chi \in \hat{G}} \hat{f}(\chi) \chi(g), where \hat{f}(\chi) = \int_G f(h) \overline{\chi(h)} \, dh. This framework generalizes classical Fourier analysis on the circle to arbitrary compact abelian settings, facilitating the study of representations and harmonic functions.

Tannaka–Krein reconstruction

The Tannaka–Krein reconstruction theorem provides a duality that allows a compact group G to be recovered from the tensor category \operatorname{Rep}(G) of its continuous finite-dimensional unitary representations, via the forgetful functor to the category of vector spaces \operatorname{Vect}. Specifically, given a symmetric monoidal category \mathcal{C} equipped with a fiber functor \omega: \mathcal{C} \to \operatorname{Vect} that forgets the group action, the theorem asserts that if \mathcal{C} \cong \operatorname{Rep}(G) for some compact group G, then G can be reconstructed as the group of natural automorphisms of \omega, or more algebraically, as the automorphism group of the functor in the presence of rigidity conditions like those imposed by irreducible representations. This duality, originally established for non-abelian compact groups, generalizes the abelian Pontryagin duality by handling the full representation category rather than just characters. Central to the reconstruction is the construction of a from the matrix coefficients of representations in \operatorname{Rep}(G). The space of matrix coefficients, consisting of functions g \mapsto \langle \pi(g) v, w \rangle for \pi \in \operatorname{Rep}(G), v, w in the representation space, forms a dense C_r(G) of the continuous functions C(G) on G, equipped with a Hopf algebra structure via the comultiplication induced by the of representations: \Delta(f)(g,h) = f(gh). The compact group G is then recovered as the \operatorname{Spec}(A) of the character algebra A, the universal commutative Hopf algebra coacting on \operatorname{Rep}(G), ensuring uniqueness up to isomorphism under the category equivalence. This algebraic perspective applies particularly well to matrix groups, where representations are realized in \mathrm{GL}(n, \mathbb{C}), allowing explicit computation of the dual structure. The theorem's non-abelian nature extends its utility beyond classical groups to deformed or quantum settings, such as compact quantum groups, where reconstruction proceeds from the of corepresentations (unitary representations on Hilbert spaces with coactions). In this framework, pioneered by Woronowicz, a compact quantum group is defined via its Hopf C^*-algebra, and the Tannaka–Krein duality reconstructs it from the rigid C^*-tensor of corepresentations, mirroring the classical case but incorporating non-commutative geometry. Applications include classifying ergodic actions and coactions, as seen in quantum analogs like SU_q(2), where the representation determines the underlying quantum structure uniquely.

Relations to non-compact groups

Real and complex forms

Compact s are intimately connected to non-compact s through the processes of and the theory of real forms. Every compact G admits a unique G^\mathbb{C}, which is a containing G as a maximal compact real . This is universal in the sense that any holomorphic from G to a factors uniquely through G^\mathbb{C}. For instance, the \mathrm{SU}(2) complexifies to the \mathrm{SL}(2,\mathbb{C}), where \mathrm{SU}(2) sits as a maximal compact . Real forms provide a framework for relating compact and non-compact real Lie groups to a common complex structure. Given a complex semisimple G^\mathbb{C} with \mathfrak{g}^\mathbb{C}, a real form is a real subgroup H (or its \mathfrak{h}) such that \mathfrak{g}^\mathbb{C} = \mathfrak{h} \oplus i \mathfrak{h} as real vector spaces. Compact real forms correspond to those H that are compact, characterized by the Killing form being negative definite on \mathfrak{h}. Such forms arise as the fixed points of an anti-holomorphic \theta on G^\mathbb{C} with \theta^2 = \mathrm{id}, where the compact real form is \{ g \in G^\mathbb{C} \mid \theta(g) = g \}. For non-compact real forms, the Cartan decomposition plays a central role. If G is a non-compact real semisimple with maximal compact K, its decomposes as \mathfrak{g} = \mathfrak{k} \oplus \mathfrak{p}, where \mathfrak{k} is the of K, \mathfrak{p} is the with respect to the Killing form, and \mathfrak{p} is invariant under the adjoint action of K. The Killing form B is negative definite on \mathfrak{k} and positive definite on \mathfrak{p}, with B(\mathfrak{k}, \mathfrak{p}) = 0. This decomposition stems from a Cartan involution \theta on \mathfrak{g}, which is an automorphism of order 2 such that B(X, \theta Y) is positive definite, with \mathfrak{k} = \{ X \mid \theta(X) = X \} and \mathfrak{p} = \{ X \mid \theta(X) = -X \}. Globally, G = K \exp(\mathfrak{p}) as a diffeomorphism. Examples illustrate these relations clearly. For the complex group \mathrm{SL}(n,\mathbb{C}), the compact real form is the special unitary group \mathrm{SU}(n), while the non-compact real form \mathrm{SL}(n,\mathbb{R}) admits the Cartan decomposition with maximal compact \mathrm{SO}(n), where \mathfrak{sl}(n,\mathbb{R}) = \mathfrak{so}(n) \oplus \mathfrak{p} and \mathfrak{p} consists of symmetric traceless matrices. In contrast, the orthogonal group \mathrm{SO}(n) is the compact real form of \mathrm{SO}(n,\mathbb{C}), with non-compact real forms like \mathrm{SO}(p,q).

Dual pairs and theta correspondence

In the context of , a reductive dual pair consists of two reductive subgroups K and K' of a \mathrm{Sp}(W) over \mathbb{R}, where K and K' act as mutual centralizers. When K is compact, such pairs provide a mechanism to relate finite-dimensional representations of K to infinite-dimensional unitary representations of the non-compact group K'. The theta correspondence, also known as the Howe correspondence, arises from the oscillator representation (or Weil representation) \omega_\psi of the \mathrm{Mp}(W), the double cover of \mathrm{Sp}(W), realized on the space of functions on a associated to W. For an \pi of the compact group K, the theta lift \theta_{K,K'}(\pi) is defined as the unique irreducible quotient of the representation of K' obtained by projecting \omega_\psi|_{K' \times K} onto the \pi-isotypic component and taking the algebraic dual or a similar construction to ensure unitarity. This lift is non-zero under stable range conditions, such as when the dimension parameters satisfy certain inequalities ensuring persistence of the correspondence. Howe duality asserts that the oscillator representation \omega_\psi decomposes multiplicity-freely as \omega_\psi \cong \bigoplus_{\pi \in \mathcal{H}(K)} \pi \otimes \theta_{K,K'}(\pi), where \mathcal{H}(K) is the set of irreducible representations of K that appear in the decomposition, often all finite-dimensional irreducibles when K is compact and the pair is in the stable range. This duality interchanges the roles of K and K', establishing a between \mathcal{H}(K) and the corresponding modules for K'. A classical example is the dual pair (O(n), \mathrm{Sp}(2m, \mathbb{R})) in \mathrm{Sp}(2nm, \mathbb{R}), where O(n) is compact; here, the theta lift of the standard of O(n) yields the holomorphic series representation of \mathrm{Sp}(2m, \mathbb{R}). Another prominent type I example is (U(n), U(p,q)) in \mathrm{Sp}(2n(p+q), \mathbb{R}), with U(n) compact; the correspondence lifts unitary representations of U(n) to those of the non-compact U(p,q), with explicit descriptions in terms of Langlands parameters when p+q = n. These constructions have significant applications in the of non-compact groups, where theta lifts from compact subgroups generate families of unitary representations, such as supercuspidal or discrete series types. In the study of automorphic forms, the local extends to global settings via see-saw dual pairs, relating automorphic representations on adelic quotients of compact and non-compact groups. Additionally, it informs branching laws by providing multiplicity formulas for restrictions of representations from non-compact groups to compact subgroups through iterated theta lifts.

References

  1. [1]
    [PDF] 1. Haar measure on compact groups
    1.1. Compact groups. Let G be a group. We say that G is a topological group if G is equipped with hausdorff topology such that the multiplication (g, h) 7− → ...Missing: mathematics | Show results with:mathematics
  2. [2]
    [PDF] Properties of topological groups and Haar measure
    groups: Definition 1.8. A locally compact topological group is a Hausdorff topo- logical group for which each point has a compact neighborhood. Proposition ...
  3. [3]
    [PDF] CHAPTER 6 Representations of compact groups
    The real orthogonal group On(R) := U(n)∩GLn(R) is another example of a compact group. Because of the definition of group of t.d. type, there are many such ...
  4. [4]
    [PDF] The Peter-Weyl Theorem for Compact Groups x1 Preliminaries.
    Theorem 17: Let G be a compact group. (1) Every irreducible representation of G is nite dimensional. (2) If is the left-regular representation of G ...
  5. [5]
    [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 ...
  6. [6]
    [PDF] TOPOLOGICAL GROUPS The purpose of these notes is to give a ...
    A topological space X is called locally compact if for every x ∈ X, there is a compact neighborhood U ⊂ X of x. Before proving a few basic properties of locally ...
  7. [7]
    [PDF] Representations of Compact Topological Groups
    It is easy to show that G is compact. Profinite groups are ubiquitous in mathematics. For example, the p-adic integers Zp for a prime p form a profinite group, ...
  8. [8]
    [PDF] Harmonic analysis on compact abelian groups 1. Approximate ...
    Mar 23, 2013 · [1.1] Topological groups As expected, a topological group G is a group with a topology, such that the group operation G × G → G and the ...
  9. [9]
    [PDF] Some algebraic properties of compact topological groups - People
    Compact topological groups arise in many areas of mathematics. Probably the first example one thinks of is S1, the circle group. When the theory of Lie.
  10. [10]
    [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 ...
  11. [11]
    Abstract Harmonic Analysis
    Standard concepts and notationare used without explanation. Hewitt and Ross, Abstract harmonic analysis, vol. I. 1. Page 9. Chapter I. Prebnucaries. The symbols ...
  12. [12]
    [PDF] duality and structure of locally compact abelian groups ..... for the ...
    Our aim is to describe the principal structure theorem for locally compact abelian groups, and to acquaint the reader with the Pontryagin - van Kampen duality ...
  13. [13]
    Abstract Harmonic Analysis: Volume II - SpringerLink
    Abstract Harmonic Analysis. Book Subtitle: Volume II: Structure and Analysis for Compact Groups Analysis on Locally Compact Abelian Groups. Authors: Edwin ...
  14. [14]
    [PDF] math 210c. compact lie groups - Harvard Mathematics Department
    First definitions and examples. Definition 1.1. A topological group is a topological space G with a group structure such that the multiplication map m : G × ...
  15. [15]
    [PDF] 18.199 Talk 1 : A Crash Course on Lie Groups
    Definition 1.1. A set G is a Lie group if it is a group and a ... If G is a compact Lie group, then the sectional curvature is positive semi-definite.
  16. [16]
    [PDF] Lecture Notes on Compact Lie Groups and Their Representations
    In this introductory chapter, we essentially introduce our very basic objects of study, as well as some fundamental examples. We also establish some.<|separator|>
  17. [17]
    [PDF] Compact Lie Groups - University of Oregon
    May 5, 2022 · ... Weyl Group ... Hermann Weyl wrote an amazing series of papers in which he extended the theory of group representations from finite ...
  18. [18]
    [PDF] The Cartan-Killing Form
    Theorem 2: If g is a compact semisimple Lie algebra, then for any nonzero X ∈ g, it follows that (X, X) < 0 and conversely. Proof: Since Adg (for g ∈ G) is the ...
  19. [19]
    [PDF] The Killing Form, Reflections and Classification of Root Systems 1 ...
    Theorem 1. For g the Lie algebra of a compact semisimple Lie group, the. Killing form K is strictly negative definite on g, i.e. for any X 6 ...
  20. [20]
    [PDF] An introduction to totally disconnected locally compact groups
    Jan 11, 2018 · A totally disconnected locally compact (tdlc) group is the group of components G/G◦, where G◦ is the connected component of the identity in a ...
  21. [21]
    [PDF] An introduction to totally disconnected locally compact groups
    Feb 21, 2019 · A totally disconnected locally compact (tdlc) group is the group of left cosets of a connected locally compact group, and is studied as both ...
  22. [22]
    [PDF] Infinite Galois Theory
    May 1, 2016 · With these two theorems, we have that profinite groups are equivalent to the Hausdorff, totally disconnected, and compact topological groups. 11 ...
  23. [23]
    [PDF] An introduction to totally disconnected locally compact groups
    Sep 14, 2020 · 3.3 Examples of totally disconnected locally compact groups . ... The structure of totally disconnected, locally compact groups.
  24. [24]
    [PDF] Admissibility of representations of totally disconnected locally ...
    Jan 14, 2016 · Definition 1.4 (Compact and Locally Compact groups). A compact group is a compact and Hausdorff topological group. Likewise a locally ...<|separator|>
  25. [25]
    [PDF] infinite galois theory (draft, ctnt 2020)
    Theorem 5.1. The topology on a Galois group Gal(L/K) is totally disconnected: the only nonempty connected subsets are points. Proof. Let C be a nonempty ...
  26. [26]
    254A, Notes 3: Haar measure and the Peter-Weyl theorem - Terry Tao
    Sep 27, 2011 · In this post, we will construct Haar measure on general locally compact groups ... Theorem 3 (Existence and uniqueness of Haar measure) Let {G} be ...
  27. [27]
    a functional analysis proof of the existence of haar measure on ...
    the existence and uniqueness of Haar measure are established separately. For compact groups, a simple proof of the existence and uniqueness of Haar measure.
  28. [28]
    [PDF] Basic Representation Theory - LSU Math
    2 is called the modular function for the group G. If ∆ is identically one, the group G is said to be unimodular. Thus a left Haar measure on G is right ...<|control11|><|separator|>
  29. [29]
    [PDF] 13 Haar measures and the product formula - MIT Mathematics
    Oct 25, 2016 · One defines a right Haar measure analogously, but in most cases they coincide and in our situation we are working with an abelian group (the ...
  30. [30]
    [PDF] 1 Construction of Haar Measure - UCSD Math
    This completes the proof. 16. Page 17. 1.1 Exercises. Exercise 1.4. Let m be the normalized Haar measure of a compact group G. For f ∈ C(G) or L. 1. (G) show ...
  31. [31]
    on haar measure in locally compact t2 spaces. - jstor
    measure on G. COROLLARY 5.9. If H is a compact closed subgroup of a locally compact topological group G, then there exists a unique Haar measure on G/H. (Note.
  32. [32]
    [PDF] Haar Measure on LCH Groups
    Dec 3, 2016 · Haar measure is a translation invariant measure on LCH groups, used in math, physics, and statistics. It is a translation invariant measure.
  33. [33]
    [PDF] 2015.84624.Topological-Transformation-Groups.pdf
    Jul 14, 1970 · TOPOLOGICAL TRANSFORMATION GROUPS. Deane Montgomery and Leo Zippin. INTERSCIENCE PUBLISHERS, INC., NEW YORK. INTERSCIENCE PUBLISHERS LTD ...
  34. [34]
    [PDF] lecture 28: the structure of compact lie groups
    Let G be a compact Lie group, then g is reductive. More explicitly, g = g/ ⊕ Z(g), where Z(g) is abelian, and g/ is semisimple ...
  35. [35]
    [PDF] Structure Theory of Semisimple Lie Groups
    (c) the Killing form of g0 is negative definite. Proof. If G is compact connected with Lie algebra g0, then Ad(G) is compact; hence (a) implies ( ...<|separator|>
  36. [36]
    [PDF] arXiv:math/0506118v1 [math.RT] 7 Jun 2005
    Jun 7, 2005 · This allows us to use results about Cartan subalgebras (such as root decomposition, properties of root systems, etc.) when studying compact Lie ...
  37. [37]
    [PDF] Classification of Real Forms of Semisimple Lie Algebras
    So let us list real forms of simple Lie algebras that we know so far. 1. Type An−1. We have the split form sln(R), the compact form su(n), and ...
  38. [38]
    [PDF] Classification of Compact Simple Lie Algebras - ETH Zürich
    May 12, 2018 · In this reprt, the compact Lie algebras are classified via the classification of complex simple Lie algebras.
  39. [39]
    [PDF] 1. Representations of compact Lie groups This is to write down ...
    Representations of compact Lie groups. This is to write down some facts which are needed to do the home- work. The main point is Proposition 1.4.
  40. [40]
    [PDF] Classification of root systems
    Sep 8, 2017 · In this section we recall the construction of a root system from a semisimple. Lie algebra given in Chapter 6 of the text. We will also state ...
  41. [41]
    [PDF] 18.755 (Lie Groups and Lie algebras II) Notes - MIT
    2.5.3 Cartan matrix and Dynkin Diagrams . ... – Compact inner class (only one since Dynkin diagram has no nontrivial auto) so(2p + 1, 2q) where p + q = n ...
  42. [42]
    [PDF] A Note on the Cartan Integers
    note we show how these integers may be defined for a compact, connected, semisimpie Lie group without recourse to Lie algebra. They turn out to be ...
  43. [43]
    [PDF] the classification of simple complex lie algebras - UChicago Math
    Aug 24, 2012 · In particular, every Lie group has an associated Lie algebra consisting of its tangent space at the identity equipped with a bracket operation.
  44. [44]
    [PDF] What is the order of the Weyl group of E8? We'll do this by 4 different
    So the order of the Weyl group is. 135 · 27 · 8! = |27 · S8| ×. # norm 4 elements. 2 × dim L. Remark 312 Conway used a similar method to calculate the order of ...
  45. [45]
    [PDF] math 210c. compact lie groups - Mathematics
    Lie theory, we need to address the Lie group meaning of the integrality condition on. `a ∈ X(T). ∗. Q ? Our goal is to show `a(X(T)) ⊂ Z (with no finiteness ...
  46. [46]
    [PDF] Lie Groups. Representation Theory and Symmetric Spaces
    In this Chapter we discuss elementary properties of Lie groups, Lie algebras and their relationship. We will assume a good knowledge of manifolds, vector.
  47. [47]
    [PDF] Notes on compact Lie groups
    Mar 8, 2021 · The fundamental example is the subgroup of diagonal matrices in U(n) or SU(n). Exercise 6.1. Let T ⊂ G be a maximal torus with Lie algebra t := ...
  48. [48]
    [PDF] 28.3 Compact Lie algebras - Berkeley Math
    28.3 Compact Lie algebras. The Lie algebras we have constructed are defined over the reals, but are the split rather than the compact forms.
  49. [49]
    Die Vollständigkeit der primitiven Darstellungen einer ...
    Die Vollständigkeit der primitiven Darstellungen einer geschlossenen kontinuierlichen Gruppe ... Article PDF. Download to read the full article text. Use our pre- ...
  50. [50]
    [PDF] Lectures on Representation Theory Dragan Milicic
    Therefore, there exists an inner product on V such that (π, V ) is a unitary representation. 1.7. Orthogonality relations. Let (ν, U) and (π, V ) be two ...
  51. [51]
    [PDF] complete reducibility of representations of compact groups
    In this lecture, we will show that every finite dimensional continuous real or complex representation of a compact topological group is completely reducible.
  52. [52]
    [PDF] representations of compact groups
    Representations of compact groups are unitarizable, and decompose into direct sums of finite-dimensional irreducibles. Continuous convolution operators are ...
  53. [53]
    [PDF] Representation theory of compact groups and complex reductive ...
    Mar 30, 2011 · We have Schur's Lemma. Lemma 1.3. If V,W are irreducible representations, then HomG(V,W) is. 1-dimensional if V. ∼. = W and 0-dimensional ...
  54. [54]
    [PDF] 0.1. Weights for a torus T. We saw that the space of weights for the ...
    Take G to be a connected compact Lie group. Fix a maximal torus T ⊂ G. Let. V be a representation space for G. By restriction, V is a representation space for.
  55. [55]
    [PDF] Highest-weight Theory: Borel-Weil - Columbia Math Department
    representations, using induction on group representations, rather than at the. Lie algebra level. In this section G will be a compact Lie group, T a maximal.
  56. [56]
    [PDF] AIM 2003 Branching Law Notes - University of Utah Math Dept.
    Jul 30, 2003 · A basic problem in the representation theory of a compact Lie group is to calculate the restriction of an irreducible representation of K to a ...Missing: rules | Show results with:rules
  57. [57]
    [PDF] Induced Representations and Frobenius Reciprocity
    In the first problem set, one exercise will be to prove Frobenius reciprocity in the Lie algebra case, and in the Lie group case, for compact Lie groups with ...Missing: torus | Show results with:torus
  58. [58]
    Theorie der Darstellung kontinuierlicher halb-einfacher Gruppen ...
    Theorie der Darstellung kontinuierlicher halb-einfacher Gruppen durch lineare Transformationen. I. Published: December 1925. Volume 23, pages 271–309, (1925) ...
  59. [59]
    [PDF] The Weyl Character Formula - Columbia Math Department
    We have seen that irreducible representations of a compact Lie group G can be constructed starting from a highest weight space and applying negative roots.
  60. [60]
    [PDF] Math 210C. Calculation of some root systems
    The classical groups are certain non-trivial compact connected Lie groups with finite cen- ter: SU(n) for n ≥ 2, SO(2m + 1) for m ≥ 1, Sp(n) for n ≥ 1, and SO( ...<|control11|><|separator|>
  61. [61]
    [PDF] arXiv:hep-th/9604029v1 5 Apr 1996
    Appendix : Weyl's Character Formula for SU(2) - Schwinger's. Derivation. Below we reproduce Schwinger's derivation of the Weyl's character formula of SU(2) [2].<|control11|><|separator|>
  62. [62]
    [PDF] arXiv:1610.00547v6 [quant-ph] 17 May 2017
    May 17, 2017 · The adjoint representation for d = 2, i.e.. Ad : SU(2) → SO(3) has a particularly nice form. Any matrix from SU(2) can be written in a form.
  63. [63]
    [PDF] The Weyl Integral and Character Formulas
    We have seen that irreducible representations of a compact Lie group G can be constructed starting from a highest weight space and applying negative roots.
  64. [64]
    [PDF] Locally compact abelian groups - Part III: Pontryagin Duality
    The dual of a discrete group is a compact group and the dual of a compact group is a discrete group. Page 22. Pontryagin duality theorem. There is a natural ...
  65. [65]
    Tannaka-Krein theorem in nLab
    Jul 1, 2025 · The Tannaka-Krein theorem (in the narrow sense) is a particular Tannaka duality theorem for compact topological groups.
  66. [66]
    [PDF] Tannaka-Krein duality for compact quantum group coactions (survey)
    Jun 4, 2019 · Tannaka [30] showed that a compact group G can be reconstructed if the set ReppGq of its continuous finite dimensional representations is known ...
  67. [67]
    [PDF] WORONOWICZ TANNAKA-KREIN DUALITY AND FREE ...
    To finish the proof of Theorem 1.1 we have to show that the compact quantum group G is unique up to isomorphism. Let G be another compact quantum group ...
  68. [68]
    [PDF] Notes on complex Lie groups - ETH Zürich
    Nov 7, 2024 · Every compact Lie group admits a complexification, unique up to canonical isomorphism. Theorem 1.8 (Cartan). Let Gc be a complex Lie group and ...
  69. [69]
    [PDF] NOTES ON THE LOCAL THETA CORRESPONDENCE Stephen S ...
    Jul 6, 1996 · Reductive dual pairs, the theta correspondence, the Howe duality principle, restriction of the metaplectic cover, the Weil representation for ...
  70. [70]
    TRANSCENDING CLASSICAL INVARIANT THEORY Let SP2n(R ...
    TRANSCENDING CLASSICAL INVARIANT THEORY. ROGER HOWE. 1. INTRODUCTION. Let SP2n(R) = SP2n = Sp be the real symplectic group of rank n. Let Sp denote the two ...
  71. [71]
    First Occurrence for the Dual Pairs (U(p, q), U(r, s)) | Cambridge Core
    Nov 20, 2018 · We prove a conjecture of Kudla and Rallis about the first occurrence in the theta correspondence, for dual pairs of the form \left( U\left( ...