Fact-checked by Grok 2 weeks ago

Structure constants

In , structure constants are the coefficients c_{ij}^k that define the Lie bracket operation in a \mathfrak{g} over a , typically \mathbb{R} or \mathbb{C}, with respect to a chosen basis \{e_i\}_{i=1}^n: [e_i, e_j] = \sum_{k=1}^n c_{ij}^k e_k. These constants are basis-dependent but fully characterize the algebra's bilinear, alternating structure, satisfying antisymmetry c_{ij}^k = -c_{ji}^k and the \sum_m (c_{ij}^m c_{mk}^l + c_{jk}^m c_{mi}^l + c_{ki}^m c_{mj}^l) = 0 for all indices i, j, k, l. Structure constants extend to more general algebras, where they parameterize the product x_i x_j = \sum_k c_{ij}^k x_k in an n-dimensional . In the context of Lie algebras, they enable classification into types like , semisimple, solvable, or based on properties such as the Killing form or derived series. In physics, particularly and gauge theories, they appear in commutation relations of generators, such as [t_b, t_c] = i \sum_e C_{bce} t_e, where the totally antisymmetric constants C_{bce} (often denoted f_{abc}) encode non-abelian symmetries like those of SU(3) in . Their values distinguish distinct Lie algebras—for instance, the non-zero constants in \mathfrak{su}(2) reflect its compact, nature—and facilitate computations in representations, including the where (t_b)_{\!ac} = i f_{abc}.

Fundamentals

General Definition

In , an F is a A over F equipped with a bilinear \mu: A \times A \to A, often denoted by (x, y) \mapsto xy, where bilinearity means that \mu is linear in each argument separately. This structure generalizes rings by incorporating the vector space properties over F, allowing from F to interact compatibly with the algebra's internal multiplication. To express this multiplication explicitly, choose a basis \{e_i\}_{i \in I} for A, where I is a finite or infinite . The product of basis elements is then given by e_i e_j = \sum_{k \in I} c_{ij}^k e_k, where the scalars c_{ij}^k \in F are called the structure constants of the with respect to this basis. Due to bilinearity, the multiplication of any two elements x = \sum_i x^i e_i and y = \sum_j y^j e_j (with coefficients x^i, y^j \in F) extends linearly as xy = \sum_{i,j,k} c_{ij}^k x^i y^j e_k. Thus, the structure constants fully determine the entire of the relative to the fixed basis. The multiplication operation itself is an intrinsic, coordinate-free property of the algebra, independent of any basis choice. However, the structure constants depend on the basis: under a , they transform according to the representation matrices of the new basis elements, abstracting the into numerical coefficients that encode the algebra's structure in a specific . This basis-dependent formulation facilitates computations while preserving the underlying abstract nature of the multiplication. In specialized cases, such as algebras, the bilinear operation takes the form of a skew-symmetric .

Historical Context

The concept of structure constants originated in the late with the work of Norwegian mathematician (1842–1899), who pioneered the theory of continuous transformation groups as a tool for analyzing symmetries in differential equations. Lie's investigations into infinitesimal transformations led to the formulation of the Lie bracket operation on the at the identity, where the coefficients defining this bracket in a basis are precisely the structure constants, encapsulating the of these groups. His seminal contributions, detailed in works such as Theorie der Transformationsgruppen (1888–1893), laid the foundational framework for what would become Lie algebras, though the explicit terminology and systematic use of structure constants developed later. A pivotal advancement occurred through the efforts of mathematician (1869–1951) between 1894 and 1900, who formalized the study of Lie algebras independently of the groups themselves and employed structure constants explicitly in the classification of simple Lie algebras. In his doctoral thesis, Sur la structure des groupes de transformations finis et continus (1894), Cartan utilized these constants to decompose algebras into root systems and Cartan subalgebras, confirming and extending Wilhelm Killing's earlier classifications by resolving inconsistencies and introducing invariant bilinear forms. Over the following years, including in his 1900 memoir on infinite continuous groups, Cartan refined these tools to handle both finite-dimensional and more general cases, establishing structure constants as central to algebraic classification and symmetry analysis. In the early , particularly during the and , German mathematician (1885–1955) significantly expanded the role of structure constants in , integrating them into the study of compact semisimple Lie groups. Weyl's four landmark papers from 1925–1926 demonstrated how real-valued structure constants in suitable bases facilitate the complete reducibility of representations and the computation of characters via the , bridging algebraic invariants with geometric and analytic properties. His work, including the 1927 Peter–Weyl theorem, emphasized the Killing form—derived from structure constants—to identify Cartan subalgebras and root systems, influencing applications in and . Following , structure constants gained prominence in and , where they underpinned symmetry groups describing fundamental interactions. A key milestone was the independent introduction of SU(3) flavor symmetry by (1929–2019) and in 1961 to organize the growing zoo of hadrons, using the group's structure constants—embodied in the —to predict particle multiplets and decay patterns under the "eightfold way." This framework, detailed in their 1961 papers, not only explained experimental data from accelerators but also paved the way for the , earning Gell-Mann the 1969 . Since the , advances in symbolic computation have transformed the handling of structure constants, enabling automated derivation and manipulation for high-dimensional Lie algebras beyond manual feasibility. Early algorithmic developments, such as those presented at the 1986 on Symbolic and Algebraic Computation, introduced methods for computing brackets and constants in systems like REDUCE and . By the 1990s, packages in and later incorporated these techniques for semisimple cases, supporting classifications and representation computations with polynomial-time efficiency over symbolic fields.

Lie Algebras

Definition in Lie Algebras

A Lie algebra \mathfrak{g} over a F (typically \mathbb{R} or \mathbb{C}) is a equipped with a called the Lie bracket [\cdot, \cdot]: \mathfrak{g} \times \mathfrak{g} \to \mathfrak{g} that satisfies two axioms: skew-symmetry, [x, y] = -[y, x] for all x, y \in \mathfrak{g}, and the , [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0 for all x, y, z \in \mathfrak{g}. These properties ensure that the bracket captures an infinitesimal version of the non-commutative multiplication in associated Lie groups. Given a basis \{e_i\}_{i=1}^n for the finite-dimensional \mathfrak{g}, the structure constants c_{ij}^k \in F (for i, j, k = 1, \dots, n) are defined by expressing the Lie bracket of basis elements as a of the basis: [e_i, e_j] = \sum_{k=1}^n c_{ij}^k e_k. The skew-symmetry of the bracket immediately implies that the structure constants are antisymmetric in the lower indices: c_{ij}^k = -c_{ji}^k for all i, j, k. These constants completely determine the structure, as the bracket of any two elements can be computed via linearity. The , when applied to basis elements, imposes a on the structure constants. Specifically, for all indices i, j, k, m, \sum_{l=1}^n \left( c_{ij}^l c_{lk}^m + c_{jk}^l c_{li}^m + c_{ki}^l c_{lj}^m \right) = 0. This relation ensures the associativity-like property of the bracket holds across the entire algebra. In certain contexts, particularly for semisimple or algebras over \mathbb{R}, alternative notations for the structure constants are used to exploit additional symmetries. For instance, in the \mathfrak{su}(n), the generators T_a (with a = 1, \dots, n^2 - 1) satisfy [T_a, T_b] = i f_{abc} T_c, where the f_{abc} are real structure constants that are totally antisymmetric in all three indices. This normalization is common in physics applications, such as , and highlights the complete antisymmetry arising from the properties of compact real forms.

Key Properties

The structure constants c_{ij}^k of a , defined by the Lie bracket [e_i, e_j] = c_{ij}^k e_k in a basis \{e_i\}, satisfy the antisymmetry relation c_{ij}^k = -c_{ji}^k for all indices i, j, k, which follows directly from the antisymmetry of the Lie bracket [X, Y] = -[Y, X]. In addition, these constants obey the , which imposes a further constraint in cyclic permutations of the indices. For simple Lie algebras over the complex numbers \mathbb{C}, the structure constants can be chosen to be totally antisymmetric, meaning c_{ijk} = c_{[ijk]} (with lowered indices via the Killing form, as detailed below), when the basis is with respect to an . This total antisymmetry simplifies computations and reflects the underlying symmetry of the algebra. For compact real Lie algebras, such as those underlying compact Lie groups like SU(n) or SO(n), the structure constants are real and totally antisymmetric in an appropriate . The structure constants also encode the adjoint representation of the , where the action of basis elements on the algebra itself is given by (\mathrm{ad}_{e_i})_j^k = c_{ij}^k, representing the linear maps \mathrm{ad}_{e_i}: e_j \mapsto [e_i, e_j]. In matrix form, the generators of the have elements ( \mathrm{ad}_{e_i} )_{jk} = - c_{ik}^j (up to index conventions and signs depending on the basis). A key invariant bilinear form associated with the Lie algebra is the Killing form (or Cartan-Killing form), defined as B(X, Y) = \mathrm{tr}(\mathrm{ad}_X \mathrm{ad}_Y) for X, Y in the , which is symmetric and ad-invariant. In terms of the structure constants and basis, the components are B_{ab} = \sum_{c,d} c_{ac}^d c_{bd}^c, providing a on the that can be used to raise and lower indices. For semisimple Lie algebras, the Killing form is non-degenerate, which implies that the algebra decomposes into a of simple ideals and allows the definition of root systems relative to a , where roots are linear functionals determined by the adjoint action and the form's inner product structure. This non-degeneracy is a hallmark property that distinguishes semisimple algebras from solvable or ones and underpins the Cartan-Weyl of such algebras.

Examples

su(2) and so(3)

The \mathfrak{su}(2) consists of $2 \times 2 anti-Hermitian traceless matrices and admits a given by the generators T^a = -\frac{i}{2} \sigma^a for a = 1, 2, 3, where \sigma^a are the : \sigma^1 = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \sigma^2 = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}, and \sigma^3 = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}. The commutation relations in this basis are [T^a, T^b] = \epsilon^{abc} T^c, where \epsilon^{abc} is the totally antisymmetric with \epsilon^{123} = 1, so the structure constants are f^{abc} = \epsilon^{abc}. The \mathfrak{so}(3) consists of $3 \times 3 real antisymmetric matrices and has a of rotation generators L_i for i = 1, 2, 3, explicitly given by L_1 = \begin{pmatrix} 0 & 0 & 0 \\ 0 & 0 & -1 \\ 0 & 1 & 0 \end{pmatrix}, \quad L_2 = \begin{pmatrix} 0 & 0 & 1 \\ 0 & 0 & 0 \\ -1 & 0 & 0 \end{pmatrix}, \quad L_3 = \begin{pmatrix} 0 & -1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 0 \end{pmatrix}. The commutation relations are [L_i, L_j] = \epsilon^{ijk} L_k, yielding structure constants f^{ijk} = \epsilon^{ijk}. The Lie algebras \mathfrak{su}(2) and \mathfrak{so}(3) are , with the isomorphism mapping T^a \mapsto L_a preserving the structure constants \epsilon^{abc}. In physics applications, such as in and rotations in , normalization conventions often employ Hermitian generators J^a = \frac{1}{2} \sigma^a for \mathfrak{su}(2), leading to [J^a, J^b] = i \epsilon^{abc} J^c and structure constants incorporating a factor of i, or rescale by factors of 2 to match trace normalizations like \operatorname{Tr}(J^a J^b) = \frac{1}{2} \delta^{ab}. Since both algebras are three-dimensional, their adjoint representation is the three-dimensional defining representation of \mathfrak{so}(3), where the generators act as the explicit $3 \times 3 matrices above; the matrix elements of the satisfy (\operatorname{ad}_{L_i})_{jk} = \epsilon_{ijk}, directly encoding the structure constants.

su(3)

The su(3) , underlying the SU(3), is an 8-dimensional real realized in the physics convention by Hermitian traceless generators T^a = \frac{1}{2} \lambda^a (for a = 1, \dots, 8), where the elements are i T^a (anti-Hermitian), and the \lambda^a are the . These satisfy the commutation relations [T^a, T^b] = i f^{abc} T^c, with over repeated indices implied, and the structure constants f^{abc} are real and totally antisymmetric in all indices. The are explicitly: \lambda^1 = \begin{pmatrix} 0 & 1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 0 \end{pmatrix}, \quad \lambda^2 = \begin{pmatrix} 0 & -i & 0 \\ i & 0 & 0 \\ 0 & 0 & 0 \end{pmatrix}, \quad \lambda^3 = \begin{pmatrix} 1 & 0 & 0 \\ 0 & -1 & 0 \\ 0 & 0 & 0 \end{pmatrix}, \lambda^4 = \begin{pmatrix} 0 & 0 & 1 \\ 0 & 0 & 0 \\ 1 & 0 & 0 \end{pmatrix}, \quad \lambda^5 = \begin{pmatrix} 0 & 0 & -i \\ 0 & 0 & 0 \\ i & 0 & 0 \end{pmatrix}, \quad \lambda^6 = \begin{pmatrix} 0 & 0 & 0 \\ 0 & 0 & 1 \\ 0 & 1 & 0 \end{pmatrix}, \lambda^7 = \begin{pmatrix} 0 & 0 & 0 \\ 0 & 0 & -i \\ 0 & i & 0 \end{pmatrix}, \quad \lambda^8 = \frac{1}{\sqrt{3}} \begin{pmatrix} 1 & 0 & 0 \\ 0 & 1 & 0 \\ 0 & 0 & -2 \end{pmatrix}. This basis is normalized such that \operatorname{Tr}(T^a T^b) = \frac{1}{2} \delta^{ab}. The structure constants f^{abc} fully determine the algebra, with all non-zero values (up to antisymmetric permutations) listed in the following table for the ordered indices a < b < c in the standard physics convention:
abcf^{abc}abcf^{abc}
1231367\sqrt{3}/2
147-1/2458\sqrt{3}/2
156\sqrt{3}/2678\sqrt{3}/2
246-1/2
257-1/2
345-1/2
These values reflect the symmetries of su(3), where the first three generators resemble the su(2) subalgebra (with f^{123} = 1), extended by additional generators introducing mixing terms. Note that signs may vary depending on the specific basis choice, but the listed values are consistent with common conventions in particle physics. In addition to the antisymmetric f^{abc}, su(3) features symmetric constants d^{abc} appearing in the anticommutator \{T^a, T^b\} = \frac{1}{3} \delta^{ab} I + 2 d^{abc} T^c, where I is the 3×3 identity matrix. The non-zero d^{abc} (totally symmetric, up to permutations) include d^{118} = d^{228} = d^{338} = \frac{1}{\sqrt{3}}, d^{146} = d^{157} = d^{256} = d^{344} = d^{355} = \frac{1}{2}, d^{247} = d^{366} = d^{377} = -\frac{1}{2}, and d^{448} = d^{558} = d^{668} = d^{778} = -\frac{1}{2\sqrt{3}}, d^{888} = -\frac{1}{\sqrt{3}}. These d^{abc} quantify deviations from orthogonality in the product of generators and are essential for computations involving symmetric invariants. The adjoint representation of su(3) is 8-dimensional, acting on the space of generators themselves via (T^a_{\text{adj}})_{bc} = -i f^{abc}, which encodes the algebra's action on its own basis. This representation corresponds to the defining root diagram of su(3), the A_2 Dynkin diagram with rank 2, where the Cartan subalgebra is spanned by T^3 and T^8. The six roots lie in the plane with coordinates (I_3, Y) (where Y = \frac{2}{\sqrt{3}} (T^8 \text{ eigenvalue})), positioned at \pm (1, 0), \pm \left(\frac{1}{2}, \frac{\sqrt{3}}{2}\right), and \pm \left(\frac{1}{2}, -\frac{\sqrt{3}}{2}\right), forming a hexagonal lattice that visualizes weight spaces and multiplets.

su(N)

The Lie algebra \mathfrak{su}(N) is the real vector space of N \times N traceless anti-Hermitian matrices, with dimension N^2 - 1. In the physics convention, it is generated by Hermitian traceless matrices T^a via the basis elements i T^a, satisfying the commutation relations [T^a, T^b] = i \sum_c f^{abc} T^c, where f^{abc} are the real, totally antisymmetric structure constants. For N > 3, a standard basis generalizes the of \mathfrak{su}(3), comprising diagonal Cartan generators H_k (spanning the (N-1)-dimensional ) and off-diagonal ladder operators corresponding to , such as symmetric and antisymmetric combinations like \frac{1}{\sqrt{2}} (|i\rangle\langle j| + |j\rangle\langle i|) and \frac{-i}{\sqrt{2}} (|i\rangle\langle j| - |j\rangle\langle i|) for i < j. These generators are orthonormal with respect to the Killing form, normalized such that \operatorname{Tr}(T^a T^b) = \frac{1}{2} \delta^{ab}. In the Chevalley basis adapted to \mathfrak{su}(N) (isomorphic to \mathfrak{sl}(N, \mathbb{C})), the roots are \alpha_{ij} = e_i - e_j for $1 \leq i \neq j \leq N, where \{e_k\} form an orthonormal basis of \mathbb{R}^N with \sum_k e_k = 0; positive roots correspond to i < j. The basis includes Cartan elements H_i dual to simple roots and root vectors E^{\alpha_{ij}} (with E^{-\alpha_{ij}} = (E^{\alpha_{ij}})^\dagger), such that the structure constants f^{abc} arise from commutators like [E^{\alpha}, E^{\beta}] = N_{\alpha, \beta} E^{\alpha + \beta} if \alpha + \beta is a root (or zero otherwise), and [H, E^{\alpha}] = \alpha(H) E^{\alpha}. For \mathfrak{su}(N), which is simply laced, the integers N_{\alpha, \beta} are \pm 1 or 0, determined by root string lengths via N_{\alpha, \beta}^2 = \frac{1}{2} q(p+1) where p, q count steps in the \beta-string through -\alpha. In practice, N_{\alpha, \beta} = \pm 1 when \alpha + \beta is a root and the roots are adjacent in the sense. The full f^{abc} incorporate these via index contractions with the metric. An explicit construction of the structure constants follows from the trace in the fundamental representation: f^{abc} = -i \operatorname{Tr}([T^a, T^b] T^c), directly yielding the antisymmetric coefficients from the normalized basis. This formula ensures consistency with the Lie algebra structure and allows computation for arbitrary N. For non-zero components, combinatorial expressions arise from index permutations; for instance, in the fundamental representation, f^{S_{nm} S_{kn} A_{km}} = \frac{1}{2} (and cyclic variants) where S_{nm} labels symmetric generators, A_{km} antisymmetric ones, and indices follow S_{nm} = n^2 + 2(m-n) - 1 for n < m. Recurrence relations facilitate extension from lower dimensions: embedding \mathfrak{su}(N) into \mathfrak{su}(N+1) adds a new row and column of zeros to generators, with new structure constants incorporating the prior ones plus terms like \sqrt{\frac{1}{2k(k-1)}} for Cartan interactions, enabling scalable computation of non-zero f^{abc} components such as those coupling off-diagonal ladders to diagonals, e.g., f^{S_{nm} A_{nm} D_m} = -\sqrt{\frac{m-1}{2m}}. These expressions highlight the combinatorial nature, with non-zeros determined by root differences \alpha_a - \alpha_b.

Other Contexts

Associative Algebras

In associative algebras over a field K, given a basis \{e_i\}_{i=1}^n, the multiplication is defined by e_i e_j = \sum_k c_{ij}^k e_k, where the coefficients c_{ij}^k \in K are the structure constants. Unlike the case in Lie algebras, these constants need not satisfy antisymmetry, so c_{ij}^k may equal c_{ji}^k in general, reflecting the non-commutative but associative product. A concrete example arises in the general linear algebra \mathfrak{gl}(n, K), the associative algebra of n \times n matrices over K. It admits a basis of matrix units \{E_{pq}\}_{p,q=1}^n, where E_{pq} is the matrix with a 1 in the (p,q)-entry and zeros elsewhere. The multiplication is given by E_{pq} E_{rs} = \delta_{qr} E_{ps}, so the structure constants are \delta_{qr} when the product expands in the basis, with all other coefficients zero. This explicit form highlights how the constants encode the bilinear map of matrix multiplication without imposing skew-symmetry. The universal enveloping algebra U(\mathfrak{g}) of a Lie algebra \mathfrak{g} over K provides another instance where structure constants play a key role. For a basis \{x_i\} of \mathfrak{g} with Lie bracket [x_i, x_j] = \sum_k c_{ij}^k x_k, the algebra U(\mathfrak{g}) is the quotient of the free associative algebra on the x_i by the ideal generated by x_i x_j - x_j x_i - [x_i, x_j], yielding relations x_i x_j - x_j x_i = \sum_k c_{ij}^k x_k. Thus, the Lie algebra's structure constants directly encode the non-commutative aspects of the associative multiplication in U(\mathfrak{g}). In representation theory, structure constants also determine the multiplication in coordinate rings of algebraic varieties, such as those arising in the study of representation rings for reductive groups. For instance, in the representation ring of a group like GL_n, the constants appear as multiplicities in tensor product decompositions, which geometrically correspond to structure constants in the convolution algebra related to the coordinate ring of the affine Grassmannian.

Hopf Algebras

In Hopf algebras, which extend bialgebras by incorporating an antipode map S: H \to H that acts as a convolution inverse to the identity, structure constants encode both the multiplicative and comultiplicative structures relative to a chosen basis. Given a basis \{e_i\} for the Hopf algebra H over a field k, the algebra structure is defined by the multiplication e_i e_j = \sum_k c_{ij}^k e_k, where c_{ij}^k are the structure constants for the product. Dually, the coalgebra structure is captured by the coproduct \Delta(e_i) = \sum_{j,k} f_i^{jk} e_j \otimes e_k, with structure constants f_i^{jk} determining the tensor decomposition, alongside the counit \varepsilon(e_i) and the antipode S(e_i) = \sum_m s_i^m e_m. These constants fully specify the Hopf algebra axioms, including coassociativity of \Delta and compatibility between the algebra and coalgebra operations. A canonical example arises in the group algebra k[G] of a finite group G over k, where the basis consists of group elements \{g \mid g \in G\}. The multiplication follows g \cdot h = gh, yielding structure constants c_{gh}^m = \delta_{m, gh} for the product, reflecting the group law. The coproduct is group-like, \Delta(g) = g \otimes g, so the corresponding constants simplify to f_g^{hk} = \delta_h^g \delta_k^g, with counit \varepsilon(g) = 1 and antipode S(g) = g^{-1}. This structure makes k[G] a cocommutative , useful for encoding group representations. Quantum groups provide q-deformed analogs, where structure constants twist the classical Lie algebra relations. For the quantized universal enveloping algebra U_q(\mathfrak{su}(2)), generated by E, F, K, K^{-1} with relations like KE = q^2 E K and EF - FE = \frac{K - K^{-1}}{q - q^{-1}}, the coproduct deforms as \Delta(E) = E \otimes K + 1 \otimes E, \Delta(F) = F \otimes 1 + K^{-1} \otimes F, and \Delta(K) = K \otimes K, introducing q-dependent structure constants in a basis of monomials that generalize the undeformed \mathfrak{su}(2) case. These twisted coproducts ensure the Hopf algebra is neither commutative nor cocommutative. Hopf algebras with such structure constants underpin tensor categories, where the comultiplication induces a monoidal structure on representations, and quasi-triangular forms—featuring a universal R-matrix—impart braiding, as in the representations of U_q(\mathfrak{su}(2)). This braided structure facilitates applications in quantum invariants and integrable systems.

Hall Polynomials

In the context of finite abelian p-groups, Hall polynomials arise as enumerative invariants that quantify the distribution of subgroups within a given group structure. For a prime p and partitions λ, μ, ν of integers representing the types of finite abelian p-groups (where the type corresponds to the invariant factors or elementary divisors in the decomposition into cyclic components), the Hall polynomial h_{\mu,\nu}^{\lambda}(q) counts the number of subgroups H of an abelian p-group G of type λ such that H has type μ and the quotient G/H has type ν. This count is independent of the specific choice of p and turns out to be a polynomial in q = p with non-negative integer coefficients, allowing evaluation at prime powers in more general settings. The original construction traces back to Philip Hall's work on the subgroup lattice of abelian p-groups, where such polynomials capture the combinatorial regularity observed in subgroup extensions. These polynomials serve as structure constants in the Hall algebra associated to the category of finite abelian p-groups (or equivalently, finite-length modules over a principal ideal domain with residue field of cardinality q). The Hall algebra H is a free abelian group on the basis consisting of isomorphism classes [G] of finite abelian p-groups, equipped with a bilinear multiplication derived from short exact sequences: specifically, [ \mu ] \cdot [ \nu ] = \sum_{\lambda} h_{\mu,\nu}^{\lambda}(q) [ \lambda ] , where μ corresponds to the subgroup type and ν to the quotient type in extensions $0 \to (type μ) \to (type λ) \to (type ν) \to 0. This algebra is associative and commutative, with the Hall polynomials providing the explicit coefficients that encode the extension data, thus structuring the ring of class functions on the category in a way analogous to multiplication in representation rings. A concrete illustration occurs for elementary abelian p-groups, which correspond to vector spaces over the finite field \mathbb{F}_q with q = p. Here, the types are given by partitions consisting of single parts (1^k for dimension k), and the Hall polynomials specialize to Gaussian binomial coefficients. For example, h_{(1^k),(1^{n-k})}^{(1^n)}(q) = \binom{n}{k}_q, which counts the number of k-dimensional subspaces in an n-dimensional vector space over \mathbb{F}_q, given explicitly by \binom{n}{k}_q = \prod_{i=0}^{k-1} \frac{q^{n-i} - 1}{q^{k-i} - 1}. This connection highlights the polynomials' role in linear algebra over finite fields, where they measure subspace flags and appear in the decomposition of induced modules from subspaces. In modular representation theory, Hall polynomials link to the analysis of induced representations for p-groups over fields of characteristic p, facilitating the computation of decomposition numbers in the group algebra kG, where k is algebraically closed of characteristic p. They contribute to understanding Brauer characters—traces of representations on simple modules—particularly in blocks with abelian defect groups, by providing combinatorial coefficients for the multiplicity of simple modules in induced or projectivized representations, thus informing the block structure and source algebras in finite group algebras.

Applications

In Physics

In Yang-Mills theories, which form the foundation of non-Abelian gauge interactions in particle physics, the structure constants of the underlying Lie algebra encode the nonlinear self-interactions of the gauge fields. The field strength tensor is given by F_{\mu\nu}^a = \partial_\mu A_\nu^a - \partial_\nu A_\mu^a + g f^{abc} A_\mu^b A_\nu^c, where A_\mu^a represents the gauge potential in the adjoint representation, g is the coupling constant, and f^{abc} are the totally antisymmetric structure constants satisfying [T^a, T^b] = i f^{abc} T^c for generators T^a. This commutator structure ensures the gauge invariance of the theory under local transformations, with the cubic term capturing the non-Abelian nature that distinguishes these models from electromagnetism. In quantum chromodynamics (QCD), the SU(3)_c gauge theory describing strong interactions, the structure constants f^{abc} dictate the gluon dynamics, particularly through the triple-gluon vertex that enables gluon self-couplings and contributes to phenomena like and . The QCD Lagrangian includes terms where gluons mediate quark interactions and self-interact via vertices proportional to g_s f^{abc}, with g_s the strong coupling; this leads to the distinctive feature of colored gluons carrying color charge themselves, unlike photons. Explicitly, the three-gluon interaction in involves a factor -i g_s f^{abc} [g^{\mu\nu} (p_1 - p_2)^\lambda + \text{cycl.}], underscoring the role of f^{abc} in momentum-dependent gluon scattering processes essential for jet production and hadronization in high-energy collisions. The electroweak theory, unifying weak and electromagnetic interactions under the SU(2)_L \times U(1)_Y gauge group, incorporates structure constants from the SU(2) sector to describe the charged and neutral current couplings of the W and Z bosons. The SU(2) structure constants f^{abc} = \epsilon^{abc} (with \epsilon^{abc} the Levi-Civita symbol) appear in the covariant derivative for left-handed fermion doublets, leading to W boson exchange in charged current processes like \beta-decay, while the Z boson, a mixture of the neutral SU(2) and U(1) fields, has couplings modulated by the weak mixing angle \theta_W but rooted in the same SU(2) algebra. These constants ensure the consistency of parity-violating interactions and the mass generation via spontaneous symmetry breaking by the , predicting precise ratios such as M_W / M_Z = \cos \theta_W. In the AdS/CFT correspondence, a conjectured duality between gravity in anti-de Sitter space and conformal field theories (CFTs) on its boundary, the structure constants of the CFT's operator algebra correspond to three-point coupling constants that match holographic computations in the bulk, offering insights into strongly coupled quantum gravity. For instance, in \mathcal{N}=4 super Yang-Mills theory dual to type IIB string theory on AdS_5 \times S^5, the CFT structure constants from chiral primary operator three-point functions align with Witten diagrams involving bulk scalar exchanges, providing exact non-perturbative agreements that validate the duality even beyond weak coupling regimes. This matching extends to higher-point correlators and has implications for quark-gluon plasma hydrodynamics and black hole physics.

In Mathematics

In the classification of semisimple Lie algebras over an algebraically closed field of characteristic zero, structure constants are fundamental in determining the via the simple root system. The A = (a_{ij}), with entries a_{ij} = 2 \langle \alpha_i, \alpha_j \rangle / \langle \alpha_j, \alpha_j \rangle for simple roots \alpha_i, \alpha_j, encodes the off-diagonal structure constants in the through the (ad_{e_i})^{1 - a_{ij}} e_j = 0. These relations generate the algebra from the , enabling the classification of simple Lie algebras into finite types (, , , , , , , , ) based on irreducible . The , a non-degenerate invariant bilinear form for semisimple algebras given by B(e_i, e_j) = \sum_k c_{ik}^l c_{jl}^k, further confirms the semisimple structure and relates directly to these constants. In the representation theory of semisimple Lie algebras, structure constants influence the computation of weight systems and multiplicities in highest weight representations. Irreducible highest weight modules L(\lambda) have weights whose multiplicities m_\lambda(\mu) satisfy recursion relations, such as the Freudenthal formula m_\lambda(\mu) = \sum_{\alpha > 0} \frac{\langle \mu + \rho, \alpha \rangle}{\langle \lambda + \rho, \alpha \rangle} m_\lambda(\mu - \alpha), where the inner product is the Killing form derived from the structure constants, and \rho is half the sum of positive . Algorithms implementing these recursions, often using the root lattice and action, rely on the specific values of c_{ij}^k to resolve the branching and ensure integrality of multiplicities. For example, in type A_n, these computations yield the known partition-based multiplicities for symmetric powers. Structure constants also underpin invariant theory for semisimple Lie algebras, particularly in the construction and eigenvalues of Casimir operators, which generate the center of the universal enveloping algebra. The quadratic Casimir operator in a representation with generators T^a is given by C_2 = \sum_a T^a T^a in a basis orthonormal with respect to the Killing form; its eigenvalues, such as C_2(\lambda) = (\lambda, \lambda + 2\rho) normalized by the dual Coxeter number, are computed using the Killing form metric induced by the constants B_{ab} = -\sum_{c,d} f_{a c d} f_{b c d}. These operators classify invariants under the adjoint action and determine decomposition rules in tensor products of representations. Geometrically, structure constants appear in the Maurer-Cartan equations for Lie groups, providing a differential formulation of the algebra's bracket. For a matrix Lie group G, the left-invariant Maurer-Cartan form \omega = g^{-1} dg satisfies d\omega + \frac{1}{2} [\omega, \omega] = 0, where the wedge product bracket expands componentwise as [\omega \wedge \omega]^k = \sum_{i < j} c_{ij}^k \omega^i \wedge \omega^j. This equation describes the zero of the canonical flat on the trivial bundle G \times \mathfrak{g}, linking to the of the group manifold.

Basis Selection

Criteria for Choosing a Basis

In Lie algebras, the choice of basis significantly influences the form and utility of the structure constants c_{ij}^k, defined by [e_i, e_j] = c_{ij}^k e_k. A key desirable property is with respect to the B(X, Y) = \operatorname{Tr}(\operatorname{ad} X \cdot \operatorname{ad} Y), an that is non-degenerate on semisimple algebras. For compact real forms, such as \mathfrak{su}(N), the negative Killing form is positive definite, allowing selection of a basis \{t_i\} where B(t_i, t_j) = 2 \delta_{ij}, often achieved via the normalized by \operatorname{Tr}(t_i t_j) = 2 \delta_{ij}. This simplifies computations involving invariant tensors and ensures the structure constants are real and totally antisymmetric, c_{ijk} = -c_{jik}, reflecting the algebra's intrinsic properties independent of basis choice. Another important criterion is sparsity in the structure constants, meaning many c_{ij}^k = 0, which reduces the number of non-zero terms in commutation relations and enhances computational efficiency. Sparse bases are particularly valuable in numerical methods for Lie-group integrators, where the complexity of exponentiating Lie algebra elements scales with the density of non-zeros; for instance, in \mathfrak{so}(N), a basis exploiting sparse commutators like [C_{rs}, C_{kl}] = \delta_{sl} C_{rk} - \delta_{rk} C_{sl} lowers the cost from O(N^6) to O(N^3) operations. Such sparsity aids in applications like solving differential equations on Lie groups, where dense constants would inflate matrix multiplications. The impact of basis choice extends to practical computations in and physics. Structure constants determine the matrices (\operatorname{ad} e_i)_{jk} = c_{ij}^k, influencing the form of higher-dimensional representation matrices and their eigenvalues. In , for example, a suitable basis simplifies integrals over Clebsch-Gordan coefficients or selection rules in addition. Poor basis choices can lead to dense matrices requiring more storage and time, while optimized ones facilitate sparse linear algebra techniques. Under a , the structure constants transform tensorially to preserve the Lie bracket and . If the new basis elements are e'_i = S_i^a e_a with S, the transformed constants are given by c'_{ij}^k = S_i^a S_j^b (S^{-1})_l^k c_{ab}^l, ensuring the algebra's identities hold invariantly. This tensorial nature allows optimization by seeking bases that minimize non-zeros or align with symmetries. A illustrative example is the \mathfrak{so}(3), isomorphic to \mathfrak{su}(2). The Cartesian basis \{J_x, J_y, J_z\} yields real, totally antisymmetric structure constants [J_i, J_j] = \epsilon_{ijk} J_k, with all cyclic permutations non-zero, making it Hermitian and suitable for generators but less sparse. In contrast, the spherical () basis \{J_z, J_+, J_-\} with J_\pm = J_x \pm i J_y has only three non-zero relations: [J_z, J_\pm] = \pm J_\pm, [J_+, J_-] = 2 J_z, introducing sparsity ideal for diagonalizing J_z in irreducible representations, though at the cost of complex entries. This highlights how spherical bases excel in spectral decompositions despite lacking Hermiticity.

Cartan-Weyl Basis

The Cartan-Weyl basis provides a canonical choice of basis for a complex semisimple Lie algebra \mathfrak{g}, leveraging its root space decomposition \mathfrak{g} = \mathfrak{h} \oplus \bigoplus_{\alpha \in \Phi} \mathfrak{g}_\alpha, where \mathfrak{h} is a Cartan subalgebra of dimension equal to the rank l of \mathfrak{g}, and \Phi is the root system with one-dimensional root spaces \mathfrak{g}_\alpha = \mathbb{C} E_\alpha. The basis consists of l commuting generators H_i \in \mathfrak{h} spanning the Cartan subalgebra and root vectors E_\alpha for each root \alpha \in \Phi^+ (positive roots), together with E_{-\alpha} for the corresponding negative roots. The structure constants in the Cartan-Weyl basis are determined by the following commutation relations, which encode the Lie bracket: [H_i, H_j] = 0, \quad [H_i, E_\alpha] = \alpha_i E_\alpha, where \alpha_i = \alpha(H_i) is the i-th coordinate of the root \alpha with respect to the basis dual to \{H_i\}. For root vectors, [E_\alpha, E_{-\alpha}] = H_\alpha, \quad [E_\alpha, E_\beta] = N_{\alpha,\beta} E_{\alpha+\beta} if \alpha + \beta \in \Phi and \alpha + \beta \neq 0, with N_{\alpha,\beta} = 0 otherwise; here H_\alpha \in \mathfrak{h} satisfies \beta(H_\alpha) = 2(\beta, \alpha)/(\alpha, \alpha) for the invariant bilinear form (\cdot, \cdot) normalized so that short roots have squared length 2. The coefficients satisfy antisymmetry N_{\alpha,\beta} = -N_{\beta,\alpha} and N_{\alpha,\beta} = N_{\beta,-\alpha-\beta}. A special case is the Chevalley basis, which is an integral form of the Cartan-Weyl basis obtained by generating root vectors from a choice of simple roots \{\alpha_i\} via repeated Lie brackets, ensuring integer structure constants. This basis satisfies the Serre relations (\ad_{E_{\alpha_i}})^{1 - \langle \alpha_i, \alpha_j^\vee \rangle} E_{\alpha_j} = 0, \quad (\ad_{F_{\alpha_i}})^{1 - \langle \alpha_j, \alpha_i^\vee \rangle} F_{\alpha_j} = 0 for distinct simple roots \alpha_i, \alpha_j, where F_{\alpha_i} = E_{-\alpha_i}, \ad_X Y = [X, Y], and \alpha_j^\vee = 2 \alpha_j / (\alpha_j, \alpha_j) is the coroot; analogous relations hold for the Cartan generators. These relations guarantee the basis respects the of the . The Cartan-Weyl basis offers key advantages for computations involving structure constants: the adjoint action of \mathfrak{h} is diagonal on root vectors, while the full decomposes into irreducible blocks corresponding to \alpha-strings (chains of roots differing by multiples of \alpha), rendering the structure constants block-diagonal and simplifying matrix representations. The multiplicities N_{\alpha,\beta} are explicitly determined up to sign by the geometry, with |N_{\alpha,\beta}|^2 = \frac{1}{2} (\alpha + \beta, \alpha + \beta) \left( \frac{2(\alpha, \beta)}{(\alpha, \alpha)} + 1 \right) in the standard normalization, or equivalently N_{\alpha,\beta} = \pm \sqrt{ \frac{(\alpha + \beta, \alpha + \beta)}{(\alpha, \alpha)} } \cdot k where k accounts for string lengths in non-simply laced cases.

References

  1. [1]
    StructureConstants - Maple Help - Maplesoft
    The structure constants are defined to be the Lie brackets of the basis elements of the Lie algebra by . •
  2. [2]
    Structure Constant - an overview | ScienceDirect Topics
    Structure constants are parameters that characterize the strength of interactions, determining the algebra structure of an algebra.
  3. [3]
    [PDF] Lecture 14 Lie Algebras and Non Abelian Symmetries
    The expression above is the defining property of the group G and is called the algebra of the group. The set of constants Cbce are called structure constants ...
  4. [4]
    [PDF] MATH 433 Applied Algebra Lecture 16
    Page 9. Algebra over a field. Definition. An algebra A over a field F (or F-algebra) is a vector space with a multiplication which is a bilinear operation on A ...
  5. [5]
    [PDF] Algebras over a field
    Oct 14, 2014 · Roughly speaking, an algebra over a field F is just a ring R with F contained in the center of R. In particular R is an F-vector space, ...
  6. [6]
    [PDF] A Closed Structure Constant Formula for the Universal Enveloping ...
    Given any algebra A, over any field, with basis B = {ei} the structure constants of A. with respect to the basis B are the scalars ck. i,j such that.
  7. [7]
    [PDF] ALGEBRAS 1. Definitions and Examples Let k be a ... - Keith Conrad
    Our definition of a k-algebra ... The exact same argument shows that the only algebraic division algebra over an algebraically closed field is the field itself.
  8. [8]
  9. [9]
    Élie Cartan (1869 - 1951) - Biography - MacTutor
    Élie Cartan worked on continuous groups, Lie algebras, differential equations and geometry. His work achieves a synthesis between these areas. He is one of ...Missing: constants | Show results with:constants
  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 ...Missing: 1920s 1930s
  11. [11]
    Hermann Weyl - Stanford Encyclopedia of Philosophy
    Sep 2, 2009 · Hermann Weyl was a great and versatile mathematician of the 20 th century. His work had a vast range, encompassing analysis, algebra, number theory, topology, ...
  12. [12]
    SYMSAC86: Symposium on Symbolic and Algebraic Computation
    Oct 1, 1986 · This paper describes a new computer algebra system design based on ... Symbolic algorithms for Lie algebra computation · Author Picture ...
  13. [13]
    [PDF] Algorithms for Lie Algebras of Algebraic Groups
    Lie algebras are called after Sophus Lie (1842 – 1899), a Norwegian nineteenth century mathematician who realized that continuous transformation groups ...
  14. [14]
    [PDF] Lecture 2 — Some Sources of Lie Algebras
    Sep 14, 2010 · Over a given field F, define the algebra of dual numbers D to be. D := F[ ]/( 2) = 1a + b | a, b ∈ F, 2 = 0l. We then define the Lie algebra Lie ...
  15. [15]
    [PDF] Physics 218 Useful relations involving the generators of su(N)
    structure constants,. (Fa)bc = −ifabc . (32). It is also convenient to ... which is equivalent to the Jacobi identity, fabefecd + fcbefaed + fdbeface ...<|control11|><|separator|>
  16. [16]
    [PDF] Contents - Physics Courses
    We see that the structure constants are independent of the representation D, but are dependent on the coordinatization of G. Note that all the business in the ...
  17. [17]
    [PDF] Note on Representations, the Adjoint rep, the Killing form, and ...
    the Lie algebra, with ... Once we normalize our generators, as discussed below, so that the killing form is a multiple of the identity and the structure constants.
  18. [18]
    [PDF] 11 Group Theory
    ab = fb ca = fb ac. (11.79). Combining (11.77, 11.78, & 11.79), we see that the structure constants of a compact Lie group are totally antisymmetric fb ac ...
  19. [19]
    [PDF] 18.745: lie groups and lie algebras, i - MIT Mathematics
    ij ∈ k are the structure constants. Then the algebra U(g) can be ... The Killing form of a Lie algebra g is the form. Bg(x, y) = tr(adxady). The ...<|control11|><|separator|>
  20. [20]
    [PDF] The Cartan-Killing Form
    after relabeling j → k and k → j. The expression for the Cartan metric tensor is basis-dependent. ... In general, the structure constants of a complex Lie algebra ...<|control11|><|separator|>
  21. [21]
    None
    ### Summary of Key Content from the Document
  22. [22]
    [PDF] Notes on group theory, v5 (S. Naculich, July 2024)
    for SU(2), the Lie algebras are the same, with structure constants cabc = abc. 3.2 SU(3). For SU(3), the generators in the defining representation are the Gell- ...
  23. [23]
  24. [24]
    [PDF] Lie Algebra Representation Theory – SU(3)
    Dec 3, 2013 · The structure constants fabc are real and totally antisymmetric. To proof this, one can express the structure constants by evaluating the right- ...
  25. [25]
    [PDF] Physics 251 Properties of the Gell-Mann matrices Spring 2017
    The explicit form of the non-zero su(3) structure constants are listed in Table 1. Table 1: Non-zero structure constants1 fabc of su(3). abc fabc abc fabc.Missing: 3x3 | Show results with:3x3
  26. [26]
    [PDF] Simple Lie algebras
    Finally, all the (n + 1)n roots of su(n + 1) are given by β(ij) = ei − ej , 1 ≤ i, j ≤ n + 1 and β(ij) is a positive root when i>j and obviously β(ij) ...
  27. [27]
  28. [28]
    On the Deformation Theory of Structure Constants for Associative ...
    Jul 14, 2010 · Definition 2.1. The structure constants are said to define deformations of the algebra A generated by given DDA if all fjk are left zero ...<|control11|><|separator|>
  29. [29]
    [PDF] 18.745 F20 Lecture 12: The Universal Enveloping Algebra of a Lie ...
    ij ∈ k are the structure constants. Then the algebra U(g) can be described as the quotient of the free algebra kh{xi}i by the relations xixj − xjxi = ∑ k.
  30. [30]
    [PDF] Structure Constants for Hecke and Representation Rings - UMD MATH
    They prove that the non-vanishing of a representation ring structure constant implies the non-vanishing of the corresponding Hecke algebra structure constant.
  31. [31]
    [PDF] A primer of Hopf algebras - OSU Math
    Summary. In this paper, we review a number of basic results about so-called Hopf algebras. We begin by giving a historical account of the results obtained ...<|control11|><|separator|>
  32. [32]
    [PDF] Symmetric Functions and Hall Polynomials - UC Berkeley math
    This book, 'Symmetric Functions and Hall Polynomials', by IG Macdonald, is a second edition, and includes topics such as Abelian groups and Hall polynomials.
  33. [33]
    Conservation of Isotopic Spin and Isotopic Gauge Invariance
    The possibility is explored of having invariance under local isotopic spin rotations. This leads to formulating a principle of isotopic gauge invariance.Missing: URL | Show results with:URL
  34. [34]
    [PDF] 2. Yang-Mills Theory - DAMTP
    As we will see, Yang-Mills is an astonishingly rich and subtle theory. It is built upon the mathematical structure of Lie groups.
  35. [35]
    [PDF] 9. Quantum Chromodynamics - Particle Data Group
    Jun 5, 2018 · Quantum Chromodynamics (QCD), the gauge field theory that describes the strong interactions of colored quarks and gluons, is the SU(3) component ...
  36. [36]
    [PDF] Lecture 8 Quantum Chromo Dynamics
    There can be three or four gluon vertices: The three gluon vertex has a complicated factor: −gsf abc. [gµν(q1 − q2)λ + gνλ(q2 − q3)µ + gλµ(q3 − q1)ν] f.Missing: triple- | Show results with:triple-
  37. [37]
    [PDF] 10. Electroweak Model and Constraints on New Physics
    Dec 1, 2023 · ... [1–4] is based on the gauge group. SU(2)×U(1), with gauge bosons Wi. µ, i = 1,2,3, and Bµ for the SU(2) and U(1) factors, respectively, and the ...
  38. [38]
    [PDF] Introduction to Lie Algebras and Representation Theory
    Write down the structure constants relative to the usual basis of R3. 2. Verify that the following equations and those implied by (L1) (L2) define a Lie algebra ...
  39. [39]
    An algorithm for computing weight multiplicities in irreducible ... - arXiv
    Mar 23, 2016 · An algorithm for computing weight multiplicities in irreducible modules for complex semisimple Lie algebras. Authors:Mikaël Cavallin.
  40. [40]
    [PDF] Lie Algebras - Harvard Mathematics Department
    Feb 5, 2004 · The Maurer-Cartan equation is of central importance in geometry and physics, far more important than the Campbell-Baker-Hausdorff formula itself ...
  41. [41]
    [PDF] The Lie Algebras su(N) An Introduction
    Mar 3, 2018 · It is conceived to give directly a concrete idea of the ( ) su N algebras and of their laws. The detailed developments, the numerous references ...
  42. [42]
    [PDF] Lie-group methods - HAL
    Jun 8, 2016 · if sparsity of structure constants is disregarded. Similar construction can be applied in other Lie algebras. In principle, we can go on to ...
  43. [43]
    [PDF] Geometric Structures on Lie Algebras
    ... Lie algebra given by the structure constants relative to a basis 13 and we make a change of basis by. B = 13T, then we immediately know the structure ...
  44. [44]
    None
    ### Summary of Bases for Lie Algebras, Cartesian vs Others, Structure Constants, Killing Form
  45. [45]
  46. [46]
    [PDF] The computation of structure constants according to Jacques Tits
    is based on a straightforward transcription of the Jacobi identity. ... corresponding Lie algebra; (3) calculate the set of structure constants for the system.