Fact-checked by Grok 2 weeks ago

Irreducible representation

In , an irreducible representation of a group G is a \rho: G \to \mathrm{GL}(V) from G to the general linear group of invertible linear transformations on a V (typically over the complex numbers \mathbb{C}) such that the only s of V invariant under the action of \rho(G) are the trivial subspace \{0\} and V itself. This property distinguishes irreducible representations as the "indecomposable" or fundamental units that cannot be broken down into simpler, nontrivial subrepresentations. Irreducible representations play a pivotal role in understanding group symmetries across and physics, serving as the basic building blocks for all finite-dimensional representations of finite groups. By Maschke's theorem, over a whose does not divide the of G, every finite-dimensional representation decomposes uniquely (up to and ordering) as a of irreducible representations. For finite groups, the number of distinct irreducible representations (up to ) equals the number of conjugacy classes in G, and the sum of the squares of their dimensions equals the of G. These representations are analyzed through their characters, which are the traces of the matrices \rho(g) for g \in G and provide relations that facilitate and . Examples include the n-dimensional representation of the O(n) on \mathbb{R}^n, which is irreducible, and the one-dimensional trivial representation where every group element acts as the identity. In applications, such as and , irreducible representations classify operations and predict physical states invariant under group actions. For groups and their algebras, similar principles hold under appropriate conditions, extending the theory to continuous symmetries.

Fundamentals

Historical Development

The concept of irreducible representations emerged in the late as part of the foundational work in group theory, particularly through the efforts of . In 1896, Frobenius introduced the notion of group characters in his seminal paper "Über Gruppencharaktere," where he analyzed the group determinant—a associated with a —and demonstrated that it factors into irreducible factors corresponding to the group's irreducible representations. This work extended earlier ideas from abelian groups and laid the groundwork for classifying irreducible representations of s by showing that the number of such representations equals the number of conjugacy classes. Over the following years, from 1896 to 1903, Frobenius further developed , establishing relations and proving that characters of irreducible representations form an for class functions, which enabled the complete of any into irreducibles. Shortly thereafter, in 1898, Heinrich Maschke proved that, under appropriate conditions on the field, every finite-dimensional representation of a decomposes uniquely (up to ) as a of irreducible representations—a result now known as Maschke's theorem. Building on Frobenius's foundations and Maschke's theorem, advanced the theory in the early by shifting focus from representations to abstract linear representations over the complex numbers. In his 1904 paper "Über die Darstellung der endlichen Gruppen durch gebrochene lineare Substitutionen," Schur provided criteria for irreducibility, including , which states that any of an irreducible representation is a scalar multiple of the . He extended this in subsequent works, such as the 1905 and 1907 papers "Untersuchungen über die Darstellung der endlichen Gruppen durch gebrochene lineare Substitutionen," where he utilized Maschke's theorem and developed methods to construct irreducible representations explicitly using Young's tableaux for symmetric groups. By 1911, Schur's contributions had solidified the framework for s, emphasizing the role of characters in determining irreducibility and multiplicity. In the 1920s, Hermann Weyl extended these ideas from finite and permutation groups to continuous groups, particularly compact Lie groups and Lie algebras, marking a transition to modern representation theory. Weyl's four groundbreaking papers from 1925 to 1926 introduced the Weyl character formula, which computes the character of any irreducible representation in terms of its highest weight, integrating root systems and the Weyl group. He also developed the "unitarian trick" to prove complete reducibility for semisimple Lie groups by averaging over the compact form and collaborated with Fritz Peter on the 1926 Peter-Weyl theorem, establishing that irreducible characters form an orthonormal basis for L² functions on compact groups. These advancements bridged discrete and continuous settings, influencing applications in quantum mechanics and symmetric spaces.

Notation and Terminology for Group Representations

In , a representation of a 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 consisting of all invertible linear endomorphisms of V. This homomorphism assigns to each element g \in G an invertible linear \rho(g) \in \mathrm{GL}(V) such that \rho(gh) = \rho(g) \rho(h) for all g, h \in G, and \rho(e) = I_V where e is the and I_V is the identity transformation on V. Equivalently, a representation provides a linear of G on V, meaning each group element acts via a linear transformation preserving the vector space structure. Upon choosing a basis for V, the representation \rho yields a matrix representation, which is a homomorphism \rho: G \to \mathrm{GL}(n, F) where n = \dim V, with each \rho(g) expressed as an invertible n \times n matrix over F. Standard notation includes \rho(g) for the image of g \in G, and the dimension of the representation, denoted \dim(\rho) or simply n, equals \dim V. If V is equipped with an inner product (e.g., a Hermitian inner product over \mathbb{C}), the representation is called unitary if each \rho(g) preserves the inner product, meaning \rho(g) is a unitary operator for every g \in G. Two representations \rho: G \to \mathrm{GL}(V) and \sigma: G \to \mathrm{GL}(W) are equivalent if there exists an invertible T: V \to W (a ) such that \sigma(g) = T \rho(g) T^{-1} for all g \in G. More generally, a T: V \to W is an intertwining between \rho and \sigma if it commutes with the , i.e., T \rho(g) = \sigma(g) T for all g \in G; equivalence holds if such a T is invertible. While representations of groups are a special case, they can be reformulated as modules over the group F[G], where the action of group elements extends linearly to the algebra, though the focus here remains on the homomorphic view for groups.

Definitions

Reducible and Irreducible Representations

In the context of a representation \rho: [G](/page/G) \to \mathrm{[GL](/page/GL)}(V) of a group [G](/page/G) on a finite-dimensional V over a k, a W \subseteq V is if \rho(g)W \subseteq W for all g \in [G](/page/G). The representation \rho is reducible if it admits a proper nontrivial W, meaning $0 \subsetneq W \subsetneq V; otherwise, \rho is irreducible. Equivalently, \rho is reducible if there exists a proper subrepresentation, that is, the restriction \rho|_W: [G](/page/G) \to \mathrm{[GL](/page/GL)}(W) for some proper W \neq 0, V. Under conditions ensuring complete reducibility, such as Maschke's theorem for finite groups over fields where the characteristic does not divide the group order, if \rho is reducible with W, then there exists a complementary U \subseteq V such that V = W \oplus U, where both W and U are invariant under \rho, and the restrictions \rho|_W and \rho|_U are subrepresentations of \rho. This decomposition allows the representation to be "reduced" by studying the smaller subrepresentations separately, motivating the focus on irreducible representations as the fundamental building blocks in . The criterion is central to identifying reducibility, and reduction can often be achieved explicitly using operators. Suppose W is a given and P_0: V \to W is any onto W (satisfying P_0^2 = P_0 and \mathrm{im}\, P_0 = W). For a G with |G| invertible in k (i.e., \mathrm{char}\, k \nmid |G|), define the averaged P = \frac{1}{|G|} \sum_{g \in G} \rho(g) \circ P_0 \circ \rho(g)^{-1}. This P is a onto W that intertwines the action of G, meaning \rho(h) \circ P = P \circ \rho(h) for all h \in G, so \ker P provides an invariant complement to W. Such projections enable the block-diagonalization of the representing matrices in a basis respecting the V = W \oplus \ker P. Maschke's theorem establishes that, under the same field characteristic condition, every finite-dimensional representation of a is completely reducible: it decomposes as a of irreducible representations. This result, proved using averaging techniques similar to the projection construction above, underscores the importance of irreducibility by guaranteeing that all representations can be built from irreducibles without "entangled" components.

Decomposable and Indecomposable Representations

In , a representation \rho: G \to \mathrm{GL}(V) of a group G over a k, equivalently a over the group kG, is called decomposable if V is isomorphic to a W \oplus U of two nonzero kG-submodules, meaning both W and U are under the action of G. Conversely, the representation is indecomposable if it cannot be expressed as such a of two nonzero subrepresentations. This notion arises naturally in the study of modules over , where every finite-dimensional decomposes uniquely (up to and ordering) into a of indecomposable modules by the Krull-Schmidt theorem, provided the endomorphism rings of the indecomposables are s. A key characterization is that a finite-dimensional representation V is indecomposable if and only if its endomorphism ring \mathrm{End}_{kG}(V) is a , meaning it has a unique (the non-units form an ideal). The concepts of decomposability and indecomposability differ from reducibility and irreducibility, as the latter concern the existence of a single proper rather than a complemented decomposition. Indecomposability is a weaker condition than irreducibility: every irreducible representation is indecomposable (since it has no proper subrepresentations at all), but the converse fails in non-semisimple settings. For instance, over an of zero, finite-dimensional representations of finite groups are completely reducible by Maschke's , so every representation decomposes into a of irreducibles, and thus indecomposability coincides with irreducibility. However, in p > 0 dividing the group , or for representations of algebras without categories, indecomposable representations that are reducible (i.e., possessing proper invariant subspaces without invariant complements) abound. A classic example occurs in for p-groups. Consider the G = \mathbb{Z}/p\mathbb{Z} over the field k = \mathbb{F}_p of p, with the two-dimensional V = k^2 where the g acts via the matrix \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix}. The W = \operatorname{span}\{(1,0)\} is under this , since g \cdot (1,0)^T = (1,0)^T, making V reducible. However, there is no complement to W in V, as any potential complement would not be preserved by g; thus, V is indecomposable. This Jordan block-like structure exemplifies how, in non-semisimple categories such as p-restricted representations of finite-dimensional algebras (including group algebras in characteristic p), decompositions may fail even when invariant subspaces exist. Such examples highlight the role of indecomposables in classifying beyond the semisimple case.

Connection Between Irreducibility and Indecomposability

In , every irreducible representation of a group G on a V is indecomposable. This follows because irreducibility means V has no proper nonzero G- subspaces, so it cannot decompose as a V = U \oplus W where both U and W are nonzero G- subspaces; such a would contradict the absence of proper subrepresentations. Indecomposability, however, is a weaker : a is indecomposable if it is not isomorphic to a of two nonzero subrepresentations, but it may still admit proper subrepresentations that do not have complements. The converse—that every indecomposable representation is irreducible—does not hold in general but is true under semisimple conditions on the representation category. For instance, for finite groups over the complex numbers \mathbb{C}, Maschke's theorem implies that the group algebra \mathbb{C}[G] is semisimple, so every finite-dimensional is completely reducible into a of irreducibles; thus, indecomposables coincide with irreducibles. Similarly, this holds for representations of finite groups over fields of characteristic not dividing |G|. Counterexamples arise in , where the characteristic p of the field divides |G|, leading to nonsemisimple categories with indecomposable but reducible . A classic example is the two-dimensional permutation of S_3 over \mathbb{F}_2: it has a one-dimensional trivial sub but cannot be expressed as a of two nonzero sub, making it indecomposable yet reducible. Such modules are common for p-groups, where only the trivial is irreducible, but higher-dimensional uniserial modules are indecomposable with nontrivial . The Krull-Schmidt theorem provides a framework for in these settings: for artinian modules (such as finite-dimensional representations over a ), any into indecomposable summands is unique up to and reordering of factors. This uniqueness aids in classifying representations, even when indecomposables exceed irreducibles. A key structural consequence of irreducibility is that the algebra \operatorname{End}_G(V) forms a , meaning every nonzero has an inverse; this reflects the absence of nontrivial invariant subspaces and underpins many orthogonality relations.

Examples

Trivial and One-Dimensional Representations

The trivial representation of a G is the one-dimensional representation \rho: G \to \mathrm{GL}(1, \mathbb{C}) given by \rho(g) = [1](/page/1) for all g \in G. This representation acts on the complex vector space \mathbb{C} by sending every group element to multiplication by the scalar , and it is always irreducible because the underlying space has dimension , admitting no proper nontrivial subspaces. More generally, any one-dimensional of a G over \mathbb{C} is a \rho: G \to \mathbb{C}^\times, where \mathbb{C}^\times denotes the of nonzero complex numbers. Such representations are matrix-valued functions taking values in $1 \times 1 matrices, and they are irreducible by virtue of their : the only subspaces of a one-dimensional are the zero and the full space itself, both of which are trivially . For finite abelian groups, all irreducible representations over \mathbb{C} are one-dimensional; this follows from the commutativity of the group, which implies that every irreducible representation must be scalar (by applied to commuting operators) and thus one-dimensional. In this context, the irreducible representations coincide with the characters of the group, forming the dual group \hat{G} under pointwise multiplication, and there are exactly |G| such representations. For a one-dimensional representation \rho, the character function satisfies \chi(g) = \rho(g) for all g \in G, as the trace of the $1 \times 1 matrix \rho(g) is simply the scalar itself. A concrete example is the sign representation of the S_3, which sends even permutations to 1 and permutations (transpositions) to -1; this is a nontrivial one-dimensional irreducible representation distinct from the trivial one. For the C_n = \langle r \rangle of n, the irreducible representations are the n one-dimensional characters \rho_k(r^m) = \exp(2\pi i k m / n) for k = 0, 1, \dots, n-1, where \rho_0 is the trivial representation and the others provide the full set of roots of unity characters.

Irreducible Representations over the Complex Numbers

In the complex numbers, an algebraically closed field of characteristic zero, every finite-dimensional representation of a finite group G is completely reducible into a direct sum of irreducible representations. For such groups, the irreducible representations over \mathbb{C} are classified by character theory: the number of distinct irreducible representations (up to isomorphism) equals the number of conjugacy classes in G, and the dimension of each irreducible representation divides |G|. A key formula arising from the orthogonality of characters states that for the character \chi of an irreducible representation, \sum_{g \in G} |\chi(g)|^2 = |G|. This relation underscores how character values determine the representation's dimension \chi(1) and confirm irreducibility. A concrete non-abelian example is the S_3 of order 6, which has three conjugacy classes (, the class of transpositions, and the class of 3-cycles) and thus three irreducible representations over \mathbb{C}: the one-dimensional trivial representation, the one-dimensional sign representation, and the two-dimensional representation. The dimensions 1, 1, and 2 all divide 6, consistent with the general . The representation acts on the two-dimensional \{(x,y,z) \in \mathbb{C}^3 \mid x + y + z = 0\} of the natural permutation representation, and it is faithful. Explicit matrices in this representation, with respect to a suitable basis, are as follows for the generators (123) and (12): \rho((123)) = \begin{pmatrix} -\frac{1}{2} & -\frac{\sqrt{3}}{2} \\ \frac{\sqrt{3}}{2} & -\frac{1}{2} \end{pmatrix}, \quad \rho((12)) = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}. These matrices extend by group relations to the full representation, which is irreducible over \mathbb{C}. Another illustrative non-abelian example is the quaternion group Q_8 = \{\pm 1, \pm i, \pm j, \pm k\} of order 8, which has five conjugacy classes ({1}, {-1}, {\pm i}, {\pm j}, {\pm k}) and thus five irreducible representations over \mathbb{C}: four one-dimensional representations and one two-dimensional representation. The dimensions 1, 1, 1, 1, and 2 all divide 8. The one-dimensional representations factor through the abelianization Q_8 / \{\pm 1\} \cong \mathbb{Z}_2 \times \mathbb{Z}_2, providing the trivial representation and three nontrivial ones where elements outside the kernels act by -1. The two-dimensional representation is faithful and arises from the inclusion of Q_8 in the unit quaternions; with respect to the basis {1, j} of the quaternions as a complex vector space, the matrices are: \rho(i) = \begin{pmatrix} i & 0 \\ 0 & -i \end{pmatrix}, \quad \rho(j) = \begin{pmatrix} 0 & -1 \\ 1 & 0 \end{pmatrix}, \quad \rho(k) = \begin{pmatrix} 0 & -i \\ -i & 0 \end{pmatrix}, with \rho(-x) = -\rho(x) for x = i, j, k and \rho(\pm 1) = \pm I_2. This representation is irreducible over \mathbb{C}.

Irreducible Representations over Finite Fields

Irreducible representations of finite groups over finite fields of characteristic p, where p is prime, form the subject of modular representation theory, which addresses cases where p divides the group order and classical theorems like Maschke's semisimple decomposition fail. In this context, the group algebra kG (with k algebraically closed of characteristic p) may not be semisimple, leading to indecomposable but reducible modules. Brauer theory organizes these representations into p-blocks, which are the indecomposable two-sided ideals of kG annihilating the simple modules, partitioning the irreducible representations according to linked composition factors in projectives. A key example illustrates the differences from characteristic zero: for the S_3 over a field of characteristic 3 (such as \mathbb{F}_3), there are two irreducible representations, both one-dimensional—the trivial representation and the sign representation—corresponding to the two p-regular conjugacy classes (the and the class of transpositions). The two-dimensional irreducible representation from the case collapses into the of these two one-dimensional representations, as the action of 3-cycles becomes unipotent and the module decomposes. For cyclic groups C_n over \mathbb{F}_p, the irreducible representations remain one-dimensional when p \nmid n, matching the characteristic zero count of n distinct characters. However, when p \mid n, such as for C_p, there is only one irreducible representation, the trivial one, since all non-identity elements are p-singular. In this case, projective indecomposables play a central role; the unique projective indecomposable module is the group algebra itself (regular module), which has composition length p with p-1 trivial factors in its socle and head. The number of irreducible modular representations equals the number of p-regular conjugacy classes, which may differ from the number of ordinary irreducible representations over \mathbb{C}. Ordinary characters reduce modulo p via a lift to a p-adic ring, but these reductions are typically not irreducible and decompose into modular irreducibles. The decomposition matrix D = (d_{ij}) encodes this relation, where each ordinary irreducible character \chi_i expresses as \chi_i = \sum_j d_{ij} \phi_j on p-regular elements, with \phi_j the Brauer characters of modular irreducibles and d_{ij} \geq 0 integers (usually 0 or 1, but not always). For S_3 in characteristic 3, labeling rows by the ordinary characters (trivial \chi_1, \chi_2, \chi_3) and columns by modular Brauer characters \phi_1 (trivial), \phi_2 (), D = \begin{pmatrix} 1 & 0 \\ 0 & 1 \\ 1 & 1 \end{pmatrix}, reflecting the standard character's decomposition.

Properties

Schur's Lemma

Schur's lemma is a fundamental result in representation theory that characterizes the linear maps, known as intertwiners, between irreducible representations of a group. Consider a group G acting via representations \rho: G \to \mathrm{GL}(V) and \sigma: G \to \mathrm{GL}(W) on finite-dimensional vector spaces V and W over a field k. An intertwiner is a linear map T: V \to W such that T \rho(g) = \sigma(g) T for all g \in G. If \rho and \sigma are irreducible but not equivalent, then the space of intertwiners \mathrm{Hom}_G(V, W) = \{0\}. If \rho and \sigma are equivalent irreducible representations, then \mathrm{Hom}_G(V, W) is isomorphic to a division algebra over k; in particular, over an algebraically closed field such as \mathbb{C}, this space is one-dimensional, consisting of scalar multiples of isomorphisms. A key special case arises when V = W and \rho = \sigma, so T is an endomorphism commuting with the representation. For an irreducible representation \rho over an algebraically closed field k, any such T satisfying [T, \rho(g)] = 0 for all g \in G must be a scalar multiple of the identity: T = \lambda I for some \lambda \in k. This follows from the general version, as the endomorphism ring \mathrm{End}_G(V) is then a division algebra over an algebraically closed field, hence one-dimensional and consisting of scalars. To sketch the proof, suppose T: V \to W is a nonzero intertwiner between irreducible representations over any . The \ker T and \mathrm{im} T are G-invariant subspaces of V and W, respectively. By irreducibility, \ker T = \{0\} (so T is injective) and \mathrm{im} T = W (so T is surjective), hence T is an . For the endomorphism case over an , if T is not a scalar, consider T - \lambda I for some eigenvalue \lambda of T; its is a proper nonzero , contradicting irreducibility. Thus, all eigenvalues coincide, and T = \lambda I. An important application concerns unitary representations over \mathbb{C}. For a finite-dimensional irreducible unitary representation \pi: G \to U(V) of a group G, any two \pi-invariant Hermitian inner products on V differ by a scalar multiple. Indeed, if \langle \cdot, \cdot \rangle_1 and \langle \cdot, \cdot \rangle_2 are such inner products, the associated sesquilinear form map T: V \to V defined by \langle Tv, w \rangle_1 = \langle v, w \rangle_2 is an intertwiner (positive definite by unitarity), hence T = \lambda I with \lambda > 0. This establishes the uniqueness of the unitary form up to scaling.

Orthogonality Relations for Characters

The character of a representation \rho: G \to \mathrm{GL}(V) of a finite group G over \mathbb{C} is defined as the function \chi_\rho: G \to \mathbb{C} given by \chi_\rho(g) = \mathrm{tr}(\rho(g)) for each g \in G, where \mathrm{tr} denotes the trace. This function is a class function, meaning it is constant on conjugacy classes of G, and it determines the representation up to isomorphism when \rho is irreducible. The space of class functions on G is equipped with the inner product \langle \chi, \psi \rangle = \frac{1}{|G|} \sum_{g \in G} \chi(g) \overline{\psi(g)}, where \overline{\psi(g)} is the complex conjugate. This inner product induces a Hermitian structure, and for characters of irreducible representations, it yields key orthogonality properties. The first orthogonality relation states that if \rho and \sigma are irreducible representations of G, then \frac{1}{|G|} \sum_{g \in G} \chi_\rho(g) \overline{\chi_\sigma(g)} = \delta_{\rho \sigma}, where \delta_{\rho \sigma} = 1 if \rho \cong \sigma and $0 otherwise. This shows that the characters of distinct irreducibles are orthogonal with respect to the inner product. The second orthogonality relation concerns sums over irreducible characters: for g, h \in G, \sum_{\rho} \chi_\rho(g) \overline{\chi_\rho(h)} = \frac{|G|}{|C_G(g)|} \delta_{\mathrm{cl}(g), \mathrm{cl}(h)}, where the sum runs over all irreducible representations \rho (up to ), C_G(g) is the centralizer of g in G, and \delta_{\mathrm{cl}(g), \mathrm{cl}(h)} = 1 if g and h are conjugate and $0 otherwise. These relations can be proved using applied to the of G. The decomposes as a of each irreducible representation \rho with multiplicity equal to its dimension \dim \rho = \chi_\rho(1). Consider the operator on this space induced by left multiplication by a fixed element; implies that intertwiners between distinct irreducibles vanish, leading to the of traces (characters) when summing over the group. For the second relation, conjugation by group elements preserves the decomposition, and centralizer sizes account for the action within classes. As consequences, the irreducible characters form an orthonormal basis for the space of class functions on G with respect to the inner product, and they constitute a complete set, spanning all such functions. This basis property enables the decomposition of any representation via its and facilitates computations like the number of irreducibles equaling the number of conjugacy classes.

Complete Reducibility for Compact Groups

In the theory of group representations, compact groups possess a fundamental property known as complete reducibility, which states that every finite-dimensional continuous over the numbers is a of irreducible representations. This result, often attributed to , extends the discrete case of Maschke's theorem for finite groups to the continuous setting and relies crucially on the topological structure of compact groups. The proof begins with the construction of a G-invariant inner product on the representation space V. Given any inner product ⟨⋅,⋅⟩₀ on V, define a new inner product by averaging over the group using the normalized Haar measure μ (with μ(G)=1): \langle v, w \rangle = \int_G \langle \rho(g)v, \rho(g)w \rangle_0 \, d\mu(g), where ρ: G → GL(V) is the representation. This integral converges due to compactness, and the resulting inner product is Hermitian positive definite and invariant under ρ(g) for all g ∈ G, making the representation unitary. With this invariant inner product, any invariant subspace U ⊆ V has an orthogonal complement U^⊥ that is also invariant, allowing the iterative decomposition of V into a direct sum of irreducible subrepresentations. To explicitly decompose V, one projects onto its isotypic components (multiplicities of each irreducible ρ). The orthogonal projection operator onto the ρ-isotypic component is given by P_\rho = (\dim \rho) \int_G \overline{\chi_\rho(g)} \, \rho(g) \, d\mu(g), where χ_ρ is the of ρ; this generalizes the finite-group formula and is G-equivariant by . These projections are mutually orthogonal and sum to the identity on V, yielding the complete decomposition V ≅ ⨁_ρ (V_ρ ⊗ Hom_G(V_ρ, V)), where the sum is finite. This property connects to the Peter–Weyl theorem, which asserts that the Hilbert space L²(G) decomposes as a direct sum ⊕_ρ (V_ρ ⊗ V_ρ^*) over all irreducible representations ρ, with matrix coefficients forming an orthonormal basis; this infinite-dimensional analog underscores the complete reducibility in the finite-dimensional case. In contrast, non-compact groups like SL(2,ℝ) admit finite-dimensional representations that are indecomposable but not completely reducible, such as certain extensions of the trivial representation.

Applications

In Physics and Chemistry

In quantum mechanics, irreducible representations (irreps) of symmetry groups provide a fundamental framework for classifying quantum states according to their transformation properties under group operations. For instance, the irreps of the rotation group SO(3) correspond directly to the angular momentum quantum number l = 0, 1, 2, \dots, where each irrep has dimension $2l + 1 and labels the degeneracy of energy levels in central potential problems, such as the hydrogen atom. This classification arises from the complete reducibility of representations for compact groups like SO(3), ensuring that physical observables, such as the Hamiltonian, commute with symmetry operations and thus preserve irrep labels. A key application is in deriving selection rules for transition probabilities, where matrix elements \langle \psi | O | \phi \rangle between states \psi and \phi vanish unless the irreps of the states are contained in the decomposition of the involving the irrep of the operator O, as determined by Clebsch-Gordan coefficients. This principle, encapsulated in the Wigner-Eckart theorem, enforces symmetries in processes like atomic s and , preventing forbidden interactions. The foundational role of in was established by in his 1931 monograph, which systematically applied irreps to atomic spectra and symmetry principles. In chemistry, irreps of molecular s classify vibrational modes, enabling the analysis of symmetry-adapted for polyatomic molecules. For (H₂O), which belongs to the C_{2v} , the three vibrational modes transform as the irreducible representations A₁ (symmetric stretch), A₁ (bending), and B₂ (asymmetric stretch), all one-dimensional due to the group's Abelian nature. Character tables of these s further determine spectroscopic activity: a mode is () active if it matches the symmetry of the (e.g., A₁, B₁, B₂ in C_{2v} for H₂O, allowing all modes to be IR active), while Raman activity requires matching the tensor symmetries (e.g., A₁ and B₂ for H₂O). This -based selection ensures that only specific modes contribute to observed spectra, aiding in molecular identification and structural elucidation.

Irreducible Representations of Lie Groups

Finite-dimensional irreducible representations of compact s are classified by dominant integral weights within the s associated to their s. For a semisimple compact G with \mathfrak{g}, the \Delta decomposes the \mathfrak{h}^* of the \mathfrak{h}, and the positive \Delta^+ define a Weyl chamber. A weight \lambda \in \mathfrak{h}^* is dominant if it lies in the closure of the fundamental Weyl chamber, satisfying \langle \lambda, \alpha \rangle \geq 0 for all positive \alpha \in \Delta^+, where \langle \cdot, \cdot \rangle denotes the induced by the Killing form. Each such dominant integral weight \lambda parametrizes a unique finite-dimensional irreducible representation L(\lambda), up to , which is highest weight with highest weight vector annihilated by the positive spaces. A concrete example arises for the \mathrm{SU}(2), whose \mathfrak{su}(2) has of type A_1. The irreducible representations are labeled by non-negative half-integers j = 0, 1/2, 1, 3/2, \dots, with dimension $2j + 1. These correspond to symmetric powers of the defining on \mathbb{C}^2, where the highest weight is $2j times the fundamental weight, and the representation space decomposes into weight spaces with weights -2j, -2j+2, \dots, 2j. This classification extends the general highest weight theory, illustrating how root multiplicities and orbits determine the full weight structure. Infinitesimal representations of Lie groups are studied through representations of their Lie algebras, where a representation \rho: \mathfrak{g} \to \mathfrak{gl}(V) captures the local action near the identity. For semisimple \mathfrak{g}, Casimir operators—central elements in the universal enveloping algebra U(\mathfrak{g}), such as the quadratic Casimir C = \sum_i x_i x^i with respect to an bilinear form—act as scalars on irreducible representations, providing invariants like the eigenvalue \langle \lambda + 2\rho, \lambda \rangle for highest \lambda, where \rho is half the sum of positive . These operators distinguish representations and facilitate computations of dimensions via Weyl's character . Weyl's unitary trick enables the construction of finite-dimensional irreducible representations for non-compact semisimple Lie groups by complexifying the Lie algebra and embedding into a compact real form, where representations are unitary and thus completely reducible. For instance, the complexification of \mathfrak{sl}(n, \mathbb{R}) yields \mathfrak{sl}(n, \mathbb{C}), whose representations restrict to those of the compact \mathrm{SU}(n). This method, originally developed by Hermann Weyl, leverages the averaging operator over the compact group with respect to the Haar measure to project onto invariant subspaces, ensuring irreducibility in the highest weight modules upon restriction.

The Lorentz Group

The , denoted SO(3,1), is the group of spacetime symmetries preserving the Minkowski metric in , and its irreducible representations play a central role in classifying fields and particles in relativistic . The finite-dimensional irreducible representations of SO(3,1) are labeled by pairs of non-negative half-integers or integers (m, n), where the dimension of the is (2m+1)(2n+1). These representations are non-unitary due to the non-compact nature of the group but are crucial for describing tensor fields and spinors. They arise from the of the complexified Lorentz algebra so(3,1)_{\mathbb{C}} with (2,\mathbb{C}) \oplus (2,\mathbb{C}), where the (m, n) corresponds to the of the spin-m and spin-n representations of the two (2,\mathbb{C}) factors. The universal double cover of the proper orthochronous Lorentz group SO^+(3,1) is SL(2,\mathbb{C}), which facilitates the classification: the (m, n) representations transform under SL(2,\mathbb{C}) \times SL(2,\mathbb{C}) as the bifundamental (2m+1) \otimes (2n+1), with m and n determining the eigenvalues of the Casimir operators associated with each sl(2,\mathbb{C}) factor, given by m(m+1) and n(n+1). The Lie algebra generators consist of rotation generators J_i and boost generators K_i (for i=1,2,3), satisfying the commutation relations [J_i, J_j] = i \epsilon_{ijk} J_k, \quad [K_i, K_j] = -i \epsilon_{ijk} J_k, \quad [J_i, K_j] = i \epsilon_{ijk} K_k, which reflect the boost-rotation mixing characteristic of the Lorentz algebra. A representative example is the vector representation, corresponding to (1/2, 1/2), which acts on 4-component Minkowski vectors and has dimension 4. Another key example is the Dirac spinor representation, obtained as the direct sum (1/2, 0) \oplus (0, 1/2), also 4-dimensional, describing 4-component spinor fields that combine left- and right-handed Weyl spinors. For physical applications requiring unitarity, such as quantum fields, the relevant irreducible representations are infinite-dimensional unitary ones, classified by using the method of induced representations from the little group of the , though the Lorentz subgroup structure yields the principal series. These unitary representations form continuous series parameterized by a complex \nu with \operatorname{Re} \nu = 1/2 (principal series) or within complementary series bounds, acting on Hilbert spaces of functions on the or , and are essential for massless fields with . The principal series includes representations induced from characters of the maximal compact subgroup SO(3), ensuring irreducibility and unitarity for the non-compact boosts.

References

  1. [1]
    Irreducible Representation -- from Wolfram MathWorld
    An irreducible representation of a group is a group representation that has no nontrivial invariant subspaces. For example, the orthogonal group O(n) has an ...
  2. [2]
    [PDF] representation theory of finite groups - Yale Math
    Definition 2.14. A non-zero G-representation V is irreducible if the only subspaces V which are preserved by the G-action are 0 and V .<|separator|>
  3. [3]
    [PDF] Representation Theory1
    Irreducible representation: A representation that is not reducible; i.e. it is impossible to find a similarity transformation that reduces all of its matrices ...
  4. [4]
    [PDF] Understanding irreducible representations - UChicago Math
    Representations take a group or Lie algebra's action into a vector space. A representation can be decomposed into irreducible representations.
  5. [5]
    [PDF] The origin of representation theory - UConn Math Department
    Abstract. Representation theory was created by Frobenius about 100 years ago. We describe the background that led to the problem which motivated Frobenius to ...<|control11|><|separator|>
  6. [6]
    [PDF] Introduction to representation theory - MIT Mathematics
    Jan 10, 2011 · (which we will explain below), Frobenius created representation theory of finite groups. ... irreducible representations of a finite group can be ...
  7. [7]
    Frobenius, Schur, and the Berlin Algebraic Tradition - SpringerLink
    Schur, I., Über die Darstellung der endlichen Gruppen durch gebrochene lineare Substitutionen, J. Reine Angew. Math. 127 (1904), 20–50. MATH Google Scholar.
  8. [8]
  9. [9]
    Hermann Weyl and representation theory | Semantic Scholar
    ... 1925–26, Weyl wrote four epochal papers in representation theory of Lie groups ... representation theory of Lie groups via the four papers mentioned above.
  10. [10]
    [PDF] Hermann Weyl and Representation Theory
    After Weyl communicated the above results to Issai Schur, the latter responded on 10th November 1924 in a long 14-page let- ter wherein Schur mostly talked ...<|control11|><|separator|>
  11. [11]
    Hermann Weyl - Stanford Encyclopedia of Philosophy
    Sep 2, 2009 · It was also during the first years of the 1920s that Weyl came to appreciate the power and utility of group theory, initially in connection with ...
  12. [12]
    [PDF] Representation Theory - UC Berkeley math
    The dimension of any complex irreducible representation of a finite group divides the order of the group. Calling G the group and V the representation, we ...
  13. [13]
    [PDF] A brief introduction to group representations and character theory
    Sep 28, 2018 · An F-representation of G consists of an F-vector space V and a group homomorphism ρ : G → GL(V ). The dimension of a representation (V,ρ) is the ...
  14. [14]
    [PDF] 5. Representations of groups
    Once again, a matrix representation of the group is obtained, of dimension equal to the dimension of the vector space V. The vector space is said to carry the ...
  15. [15]
    [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. Moreover, if ρ = σ, then HomG ...<|control11|><|separator|>
  16. [16]
    [PDF] 1 Basic notions of representation theory - MIT OpenCourseWare
    A representation V = 0 of A is irreducible (or simple) if the only subrepresenta tions of V are 0 and V . Definition 1.13. Let V1,V2 be two representations of ...Missing: textbook | Show results with:textbook
  17. [17]
  18. [18]
    [PDF] Izmir lectures 2014
    Any indecomposable representation in Mod kQ has local endomorphism ring, any representation in ModkQ is the direct sum of indecomposable representations ...
  19. [19]
    [PDF] Modular Representation Theory and the CDE Triangle - Arun Debray
    May 18, 2015 · Moreover, when char(K) = p does divide |G|, there are representations that are reducible, but indecomposable. Example 1.1.2. Let V be the ...
  20. [20]
    [PDF] A Course in Finite Group Representation Theory
    It is natural to reduce these representations modulo a prime ideal, at which point we have modular representations.
  21. [21]
    [PDF] representation theory for finite groups - UChicago Math
    Aug 29, 2014 · What is Representation Theory? Group representations describe elements of a group in terms of invertible linear transformations.
  22. [22]
    [PDF] Representation theory of finite groups III
    Jul 14, 2015 · The group S3 has two one-dimensional representations, namely the trivial one and the sign ... irreducible representations of S3. For symmetric.
  23. [23]
    [PDF] Digression Cyclic Groups Character Tables for Cyclic Groups
    Because all the operations are in the different classes, there are N irreducible representations and they are all one dimensional. This means that the.
  24. [24]
    [PDF] Linear Representations of Finite Groups - Auburn University
    A CG-module is faithful if the representation it affords has trivial kernel (i.e., if the representation is injective). Prove that if there exists a simple ...
  25. [25]
    [PDF] Notes on Representations of Finite Groups
    Via a series of exercises, we prove that the dimension of any irreducible representation divides the order of the group. This section assumes familiarity ...
  26. [26]
    [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 ...
  27. [27]
    [PDF] Modular Representation Theory - UC Berkeley math
    P is a projective indecomposable if and only if P/Rad(P) ∼= Soc(P). Definition 5.2. A finite-dimensional algebra is a Frobenius algebra if there exists a sym-.
  28. [28]
    [PDF] Introduction to modular representation theory of finite groups
    The main aim of this paper is to give a short presentation of the theory of modular representations for finite groups. We will go through some fundamental ...
  29. [29]
    [PDF] Modular Representations of Symmetric Groups - Math (Princeton)
    Aug 7, 2011 · It turns out that the Brauer characters of two irreducible representations are equal if and only if the representations are isomorphic, and ...
  30. [30]
    [PDF] The Modular Representation Theory of Cyclic Groups of Prime ...
    Jul 31, 2018 · One of the fundamental results in representation theory is the following theorem, given by Maschke in 1898. Theorem 1.1. Let V be a ...Missing: original citation
  31. [31]
    Schur's lemma in nLab
    ### Summary of Schur's Lemma from nLab
  32. [32]
    [PDF] Schur's lemma and its applications
    Suppose that π: G → GL(V ) is a finite dimensional irreducible unitary representation. If both h−, −i1 and h−, −i2 are π-invariant hermitian inner products on V ...
  33. [33]
    [PDF] complete reducibility of representations of compact groups
    This is a important and useful theory, but it will take us too far afield to prove it here. A proof can be found in [1, Chapter 7, §1, No. 2, Theorem 1].
  34. [34]
    [PDF] lecture 14. compact groups, peter-weyl theorem. complete
    Assume that G is a compact matrix Lie groups, T : G → GL(V ) is its complex representation. Then T is completely reducible. 2. Connected Lie groups. Proposition ...
  35. [35]
    [PDF] CHAPTER 6 Representations of compact groups
    irreducible subspace of H. Proof. Let h·,·i be a G-invariant inner product on H. ... Hence to complete the proof of the theorem it suffices to prove that ...
  36. [36]
    [PDF] Representation Theory And Quantum Mechanics
    This group is abelian. Therefore, from Corollary 1 of Schur's Lemma, we know that all of the irreducible representations of U(1) must be 1-dimensional.<|control11|><|separator|>
  37. [37]
    [PDF] Irreducible Tensor Operators and the Wigner-Eckart Theorem
    The selection rule m = q + m′ in the Clebsch-Gordan coefficient (55) means that many of the integrals vanish, so we have exaggerated the total number of ...
  38. [38]
    [PDF] Group theory
    All irreducible representations of C2v are one- dimensional, and only three vibrational modes (3N-6) exist in H2O. The symmetry of these modes will be obtained ...
  39. [39]
    [PDF] A Second Application of Symmetry
    Oct 5, 2015 · Symmetry is used to interpret IR and Raman spectra, determine symmetry of motions, and identify IR/Raman active vibrations using character ...
  40. [40]
    [PDF] Introduction to Lie Groups and Lie Algebras Alexander Kirillov, Jr.
    All highest weight representations with highest weight λ are of the form Mλ/W for some. W ⊂ Mλ. It is easy to see that Mλ/W is irreducible iff W is a ...
  41. [41]
    [PDF] arXiv:math/0506118v1 [math.RT] 7 Jun 2005
    Jun 7, 2005 · In this article we describe the structure and representation theory of compact Lie groups. ... Then the irreducible highest weight representation ...
  42. [42]
    Theorie der Darstellung kontinuierlicher halb-einfacher Gruppen ...
    Theorie der Darstellung kontinuierlicher halb-einfacher Gruppen durch lineare Transformationen. I ... Article PDF. Download to read the full article text. Use our ...
  43. [43]
    [PDF] Chapter 9 - Classification of finite representations of the Lorentz group
    The finite-dimensional representations of the Lorentz group are not unitary. However, field operators Ψα(x, t) do not require direct prob- ability amplitude ...
  44. [44]
    [PDF] arXiv:hep-th/0611263v2 13 Jun 2021
    Jun 13, 2021 · Casimir operators vanish for all helicity representations). • The (complex) Lorentz algebra so(3,1) is isomorphic to the direct sum of two ...
  45. [45]
    [PDF] ON UNITARY REPRESENTATIONS OF THE INHOMOGENEOUS ...
    Lorentz group can be decomposed into normal irreducible representations. It should be stated, however, that the representations in which the unit operator.
  46. [46]
    Infinite irreducible representations of the Lorentz group - Journals
    All these representations are infinite except D-κ, κ* which is finite if 2κ, 2κ* are both integral. For suitable values of (κ, κ*), Dκ, κ* or D+κ, κ* is ...<|control11|><|separator|>