Affine Lie algebras are infinite-dimensional Lie algebras that serve as a natural generalization of finite-dimensional semisimple Lie algebras, constructed as central extensions of loop algebras derived from a finite-dimensional simple Lie algebra over the complex numbers.[1][2] They form a distinguished subclass of Kac–Moody algebras, characterized by generalized Cartan matrices of affine type, which are symmetrizable, indecomposable matrices with determinant zero and all proper principal minors positive definite.[3][2]For a finite-dimensional simple Lie algebra g equipped with a nondegenerate invariant symmetric bilinear form (\cdot, \cdot), the untwisted affine Lie algebra \hat{g} is explicitly realized as the vector space \hat{g} = (\mathbb{C}[t, t^{-1}] \otimes g) \oplus \mathbb{C} K \oplus \mathbb{C} d, where K is a central element and d is a derivation acting by degree on the loop algebra \mathbb{C}[t, t^{-1}] \otimes g.[4][5] The Lie bracket is defined by [a t^m, b t^n] = [a, b] t^{m+n} + m \delta_{m, -n} (a, b) K for a, b \in g, with [K, \hat{g}] = 0 and [d, a t^m] = m a t^m.[4] This structure yields a Cartan subalgebra \hat{h} = h \oplus \mathbb{C} K \oplus \mathbb{C} d, where h is a Cartan subalgebra of g, and a root system \hat{\Delta} consisting of real roots m\delta + \alpha (for m \in \mathbb{Z}, \alpha \in \Delta \setminus \{0\}) and imaginary roots m\delta (for m \in \mathbb{Z} \setminus \{0\}), with \delta the basic imaginary root satisfying \delta(d) = 1 and \delta|_{\ h \oplus \mathbb{C} K} = 0.[4][5]Affine Lie algebras admit a triangular decomposition \hat{g} = \hat{n}^- \oplus \hat{h} \oplus \hat{n}^+ and possess an affine Weyl group \hat{W} = W \ltimes T, where W is the Weyl group of g and T is the group of translations by the coroot lattice.[5] Their representations, particularly integrable highest weight modules, play a central role in mathematics and physics, including the study of modular forms, the Geometric Langlands program, conformal field theory via the Sugawara construction, soliton equations like the KdV hierarchy, and quantum integrable models.[1][2] These algebras were developed in the 1960s and 1970s, independently by Victor Kac and Robert Moody in 1968 as a special case of Kac–Moody algebras, building on foundational work in infinite-dimensional Lie theory.[1]
Introduction
Overview
Affine Lie algebras are infinite-dimensional Lie algebras that generalize finite-dimensional semisimple Lie algebras and constitute the simplest non-trivial instances of Kac-Moody algebras. They are obtained by forming a central extension of the loop algebra associated to a finite-dimensional simple Lie algebra over the complex numbers.[1] This construction preserves many structural features of their finite-dimensional counterparts while introducing infinite dimensionality through the polynomial or Laurent polynomial dependence on a formal variable.[5]A defining characteristic of affine Lie algebras is their generalized Cartan matrix, which is of affine type—indefinite, symmetrizable, with determinant zero, and all proper principal minors positive.[6] This matrix encodes the Serre relations that generate the algebra via Chevalley generators, yielding a rich algebraic structure analogous to semisimple Lie algebras, including a triangular decomposition and a Weyl group action, despite the infinite dimension.[1]Each affine Lie algebra \hat{\mathfrak{g}} corresponds to a finite-dimensional simple Lie algebra \mathfrak{g}. Untwisted affine Lie algebras have rank equal to that of \mathfrak{g} plus two, due to the central element and scaling derivation in their construction, while twisted ones have rank one greater. This enlargement supports the extended root system including the basic imaginary root \delta.[5] Affine Lie algebras are classified into untwisted types, derived directly from loop algebras, and twisted types, obtained via finite-order automorphisms of \mathfrak{g}, each aligned with an affine Dynkin diagram extending the finite one.[6]
Historical development
The origins of affine Lie algebras trace back to the late 1960s, when Victor Kac and Robert Moody independently developed the general framework of Kac-Moody algebras, within which affine Lie algebras emerged as a special class of untwisted types. Moody's 1967 paper introduced Lie algebras constructed from generalized Cartan matrices, extending finite-dimensional semisimple Lie algebras to infinite dimensions while preserving key structural properties like root systems. Independently, Kac's 1968 work classified simple irreducible graded Lie algebras of finite growth, providing a parallel foundation that highlighted their connections to root-graded structures and infinite-dimensional generalizations.Building on these foundations, the 1970s saw the full classification of Kac-Moody algebras, including affine cases, through refinements of generalized Cartan matrices into finite, affine, and hyperbolic types based on their symmetrizability and properties.[7] This period solidified the algebraic theory, with affine Lie algebras distinguished by their realization as central extensions of loop algebras and their indefinite but bounded root multiplicities.In the 1980s, significant advancements integrated affine Lie algebras into vertex operator algebras, with Igor Frenkel's early constructions linking them to vertex operators and representations in 1980, and Richard Borcherds formalizing the vertex algebra framework in 1986 to unify these structures with modular invariance.[8] Concurrently, their role in physics gained prominence through applications in conformal field theory and string theory, where they model symmetries in two-dimensional systems and current algebras.[9]A key milestone was the 1983 publication of Victor Kac's book Infinite Dimensional Lie Algebras, which systematically presented the theory of affine Lie algebras and established them as a cornerstone of infinite-dimensional Lie theory.
Construction
Definition as central extensions
Affine Lie algebras can be defined as central extensions of loop algebras associated to finite-dimensional simple Lie algebras. Given a finite-dimensional simple Lie algebra \mathfrak{g} over \mathbb{C}, the corresponding loop algebra L(\mathfrak{g}) is constructed as the tensor product \mathfrak{g} \otimes \mathbb{C}[t, t^{-1}], where \mathbb{C}[t, t^{-1}] denotes the Laurent polynomials in t.[5] The Lie bracket on L(\mathfrak{g}) is defined by [x \otimes t^m, y \otimes t^n] = [x, y] \otimes t^{m+n} for x, y \in \mathfrak{g} and integers m, n \in \mathbb{Z}, extending the bracket of \mathfrak{g} bilinearly.[5] This structure captures the infinite-dimensional nature of loops on the underlying finite-dimensional algebra.To obtain the affine Lie algebra, one forms a central extension of L(\mathfrak{g}) by a one-dimensional center \mathbb{C} K, where K is the canonical central generator. The Lie algebra is then \hat{\mathfrak{g}} = L(\mathfrak{g}) \oplus \mathbb{C} K \oplus \mathbb{C} d, incorporating an additional derivation generator d = -t \frac{d}{dt}.[3] The central extension is specified by a 2-cocycle \omega: L(\mathfrak{g}) \times L(\mathfrak{g}) \to \mathbb{C} K given by the trace in the Killing form of \mathfrak{g}, explicitly \omega(x \otimes t^m, y \otimes t^n) = m \delta_{m+n,0} \kappa(x, y) K, where \kappa is the Killing form on \mathfrak{g}.[5] This cocycle ensures the extension is nontrivial and basic, preserving the simplicity properties in the infinite-dimensional setting.The Lie brackets involving the new generators are defined as follows: [x \otimes t^m, K] = 0 for all x \in \mathfrak{g} and m \in \mathbb{Z}, reflecting the centrality of K; [K, d] = 0, making K and d commute; and [d, x \otimes t^m] = m (x \otimes t^m), which implements the \mathbb{Z}-grading on L(\mathfrak{g}) by degrees in t.[3] The full bracket on \hat{\mathfrak{g}} combines the loop algebra structure with the cocycle term: [x \otimes t^m, y \otimes t^n] = [x, y] \otimes t^{m+n} + m \delta_{m+n,0} \kappa(x, y) K. The element d thus plays the role of a grading operator, decomposing \hat{\mathfrak{g}} into weight spaces \hat{\mathfrak{g}} = \bigoplus_{m \in \mathbb{Z}} \mathfrak{g} \otimes t^m \oplus \mathbb{C} K \oplus \mathbb{C} d, where the span of \mathbb{C} K \oplus \mathbb{C} d sits in degree zero.[5]
Loop algebras and untwisted affine Lie algebras
Loop algebras provide a foundational construction for untwisted affine Lie algebras, serving as infinite-dimensional Lie algebras derived from finite-dimensional simple Lie algebras. Given a finite-dimensional simple Lie algebra \mathfrak{g} over \mathbb{C}, the loop algebra \mathfrak{g}[t, t^{-1}] consists of elements that are Laurent polynomials in t with coefficients in \mathfrak{g}, formally \mathfrak{g}[t, t^{-1}] = \bigoplus_{n \in \mathbb{Z}} t^n \mathfrak{g}. The Lie bracket on the loop algebra extends the bracket of \mathfrak{g} by bilinearity, defined as [t^m \otimes x, t^n \otimes y] = t^{m+n} \otimes [x, y] for x, y \in \mathfrak{g} and m, n \in \mathbb{Z}.[5]Untwisted affine Lie algebras arise as central extensions of these loop algebras, specifically the unique (up to isomorphism) simply-laced central extension \hat{\mathfrak{g}} of \mathfrak{g}[t, t^{-1}] by a one-dimensional center. This extension incorporates the coroot lattice of \mathfrak{g} extended by the affine root lattice, ensuring the structure captures the infinite-dimensional analog of the finite-dimensional root system without additional twisting. The full algebra is realized as \hat{\mathfrak{g}} = \mathfrak{g}[t, t^{-1}] \oplus \mathbb{C} K \oplus \mathbb{C} d, where K generates the center and d is a derivation, with the bracket relations extending those of the loop algebra via a canonical 2-cocycle on \mathfrak{g}[t, t^{-1}].[5]The structure of \hat{\mathfrak{g}} admits a natural \mathbb{Z}-grading derived from the loop variable, given by \hat{\mathfrak{g}} = \bigoplus_{n \in \mathbb{Z}} \mathfrak{g}_n \oplus \mathbb{C} K \oplus \mathbb{C} d, where \mathfrak{g}_n = t^n \mathfrak{g} for n \neq 0 and \mathfrak{g}_0 = \mathfrak{g}. This grading is induced by the action of d, which satisfies [d, x] = n x for x \in \mathfrak{g}_n, reflecting the polynomial degrees.[5][10]A key automorphism property of untwisted affine Lie algebras stems from the circle action generated by e^{2\pi i \mathrm{ad} d}, which acts trivially on each graded component due to integer eigenvalues of \mathrm{ad} d and fixes both K and d. This action preserves the Lie bracket and the central extension structure, underscoring the periodic nature inherited from the loop construction and ensuring the algebra's invariance under full circle rotations.[5]
Twisted affine Lie algebras and Dynkin diagrams
Twisted affine Lie algebras arise from non-trivial finite-order automorphisms of finite-dimensional simple Lie algebras, providing a generalization of the untwisted construction. Let \mathfrak{g} be a finite-dimensional simple complex Lie algebra, and let \sigma: \mathfrak{g} \to \mathfrak{g} be a diagram automorphism of order m \geq 2. The subalgebra of fixed points \mathfrak{g}^\sigma = \{ x \in \mathfrak{g} \mid \sigma(x) = x \} is reductive and isomorphic to the Lie algebra associated with the folded Dynkin diagram obtained by identifying nodes under \sigma. The twisted loop algebra L_\sigma(\mathfrak{g}) consists of Laurent polynomials \sum_{i \in \mathbb{Z}} x_i \otimes [t^i](/page/T.I.) \in \mathfrak{g} \otimes \mathbb{C}[t, t^{-1}] such that \sigma(x_i) = \omega^{-i} x_i, where \omega = e^{2\pi i / m} is a primitive m-th root of unity. This algebra is then centrally extended by a one-dimensional center \mathbb{C} c using the standard 2-cocycle on loop algebras, \psi\left( \sum x_i t^i, \sum y_j t^j \right) = \sum_{i} i (x_i | y_{-i}), where (\cdot | \cdot) is the invariant bilinear form on \mathfrak{g} normalized so that short coroots have squared length 2; the derivation element d = -t \frac{d}{dt} is adjoined to yield the full twisted affine Lie algebra \hat{\mathfrak{g}}_\sigma = L_\sigma(\mathfrak{g}) \oplus \mathbb{C} c \oplus \mathbb{C} d.[11]The classification of twisted affine Lie algebras corresponds to the non-identity diagram automorphisms of the finite simple Lie algebras, which exist only for types A_{2n}, A_{2n-1}, D_{2n+1}, D_4, and E_6 with orders m=2 or $3, leading to three infinite families and two exceptional cases of non-simply-laced affine types. Untwisted affine Lie algebras, arising from the identity automorphism \sigma = \mathrm{[id](/page/id)}, have simply-laced Dynkin diagrams and correspond to the loop algebra construction without twisting. In contrast, twisted cases produce diagrams with multiple bonds, reflecting the action of \sigma on roots and distinguishing short and long roots in \mathfrak{g}^\sigma. For instance, the automorphism \sigma of order 2 on D_{2n+1} swaps the two end nodes of the Dynkin diagram, yielding the twisted type D_{2n+1}^{(2)} whose underlying finite algebra is B_n.[11][12]Affine Dynkin diagrams for twisted algebras are derived by folding the untwisted affine diagram under the automorphism \sigma, identifying nodes in the same orbit and adjusting bond multiplicities according to the order m. The general procedure starts with the finite Dynkin diagram of \mathfrak{g}, folds it via \sigma to obtain the diagram of \mathfrak{g}^\sigma, then adds an extra affine node connected to an end node of this folded diagram, with the connection strength determined by the Cartan matrix entries. For untwisted cases, this yields simply-laced diagrams like A_n^{(1)} by adding a node symmetrically to the cycle. Twisted diagrams, however, result from non-trivial folding: for example, D_{2n+1}^{(2)} has a diagram with nodes labeled 0 to n, where node 0 connects doubly to node 1, and the chain from 1 to n-1 has single bonds, ending with a double bond to two short roots at n. The full list of twisted affine Dynkin diagrams is given below:
Type
Finite Origin
Order m
Diagram Description
A_{2n}^{(2)} (n≥2)
A_{2n}
2
Linear chain of n+1 nodes with double bonds connecting the affine node to node 1 and specific doubles in the chain based on folding.
A_{2n-1}^{(2)} (n≥2)
A_{2n-1}
2
Linear chain of n nodes with double bond between nodes 1 and 2; remaining single bonds; affine node connected singly to the end node.
Linear chain from node 0 (affine, double to 1) to n-1 with single bonds, ending in double bond to two forked short root nodes.
D_4^{(3)}
D_4
3
Central node triple-bonded to three end nodes; affine node connected singly to one end.
E_6^{(2)}
E_6
2
Folded chain of 7 nodes (rank 7) with double bonds and a branch at node 3; affine node incorporated in the folding with appropriate double connections.
(Note: Full precise diagrams with node labels and marks appear in standard references such as Kac's Infinite Dimensional Lie Algebras.)[11][12][13]In the root system of a twisted affine Lie algebra, the imaginary roots are generated by the null root \delta, an element in the root lattice with (\delta, \alpha) = 0 for all real roots \alpha and (\delta, \delta) = 0, corresponding to the central element in the grading. The imaginary roots are then \{ k \delta \mid k \in \mathbb{Z} \setminus \{0\} \}, each with multiplicity equal to the dimension of \mathfrak{g}^\sigma, and they span the degenerate directions of the bilinear form inherited from the finite case. This structure distinguishes affine root systems from finite ones, enabling infinite-dimensional representations while preserving key finiteness properties like the Weyl group action on real roots.[11]
Algebraic structure
Cartan subalgebra and Weyl basis
In affine Lie algebras, the Cartan subalgebra \hat{\mathfrak{h}} is constructed as an extension of the Cartan subalgebra \mathfrak{h} of the underlying finite-dimensional simple Lie algebra \mathfrak{g}. Specifically, \hat{\mathfrak{h}} = \mathfrak{h} \oplus \mathbb{C} [K](/page/K) \oplus \mathbb{C} [d](/page/D*), where K is a central element representing the canonical central extension and d is the derivation operator corresponding to the loop variable grading.[11][14] The basis for \hat{\mathfrak{h}} consists of the elements \{h_i \mid i = 1, \dots, r\} from \mathfrak{h}, together with K and d, where r = \dim \mathfrak{h} is the rank of \mathfrak{g}. This yields a commutative subalgebra of dimension r + 2.[5]The Weyl basis for an affine Lie algebra \hat{\mathfrak{g}} is adapted from the finite-dimensional case and consists of root vectors e_\alpha associated to the real roots \alpha \in \hat{\Phi}^\mathrm{re}, where \hat{\Phi}^\mathrm{re} denotes the set of real roots in the affine root system. For each real root \alpha, there is a corresponding root space \hat{\mathfrak{g}}_\alpha spanned by e_\alpha, satisfying the commutation relation [h, e_\alpha] = \alpha(h) e_\alpha for all h \in \hat{\mathfrak{h}}. The negative root vectors are taken as f_\alpha = e_{-\alpha}, and the full basis includes the Cartan elements, forming a triangular decomposition \hat{\mathfrak{g}} = \hat{\mathfrak{n}}_- \oplus \hat{\mathfrak{h}} \oplus \hat{\mathfrak{n}}_+ analogous to the finite case. These root vectors generalize the Serre relations, ensuring the algebra is generated by the simple root vectors with the appropriate bracket relations.[11][5]A key subset of the Weyl basis is given by the Chevalley generators \{e_i, f_i \mid i = 0, 1, \dots, l\}, corresponding to the affine simple roots \{\alpha_i\}, where l is the rank of the affine algebra (one more than the finite rank). These satisfy the affine Serre relations derived from the generalized Cartan matrix A = (a_{ij}): [h_i, e_j] = a_{ij} e_j, [h_i, f_j] = -a_{ij} f_j, [e_i, f_j] = \delta_{ij} h_i, and the Serre relations (\mathrm{ad}_{e_i})^{1 - a_{ij}} (e_j) = 0 and (\mathrm{ad}_{f_i})^{1 - a_{ij}} (f_j) = 0 for i \neq j. The element h_0 in the Cartan basis is adjusted to incorporate the highest root of \mathfrak{g}.[14][11]Affine Lie algebras are infinite-dimensional, but they admit a \mathbb{Z}-grading induced by the derivation d, with [d, x_m] = m x_m for basis elements x_m in the m-th graded component, each of which is finite-dimensional and isomorphic to the root spaces of \mathfrak{g}. This grading structure ensures that the algebra decomposes into finite-dimensional pieces, facilitating computations in representations and root systems.[5][14]
Killing form and bilinear form
In affine Lie algebras, the Killing form is defined analogously to the finite-dimensional case as B(X, Y) = \mathrm{tr}_{\hat{\mathfrak{g}}}(\mathrm{ad}\, X \cdot \mathrm{ad}\, Y), where the trace is taken in a suitable representation of the infinite-dimensional algebra \hat{\mathfrak{g}}.[15] This form is symmetric and invariant under the adjoint action, but unlike the non-degenerate Killing form on finite-dimensional simple Lie algebras, it is degenerate on \hat{\mathfrak{g}}. The radical of the Killing form, consisting of elements Z such that B(Z, X) = 0 for all X \in \hat{\mathfrak{g}}, is the two-dimensional subspace \mathbb{C} K \oplus \mathbb{C} d, where K is the canonical central element and d is the derivation element generating the loop grading.[16]To obtain a non-degenerate invariantbilinear form, one employs a normalized basic inner product (\cdot | \cdot) on \hat{\mathfrak{g}} that extends the Killing form on the underlying finite-dimensional simple Lie algebra \mathfrak{g}. This form is defined on the loop algebra component by (x \otimes t^m | y \otimes t^n) = \delta_{m, -n} B(x, y) for x, y \in \mathfrak{g}, where B on \mathfrak{g} is normalized such that (\alpha_i | \alpha_i^\vee) = 2 for long simple roots \alpha_i, and extended to the full affine algebra by setting (K | K) = 0, (K | d) = 1, and (d | d) = 0, with pairings between loop elements and K or d vanishing.[16] The normalization ensures invariance under the Lie bracket and non-degeneracy when restricted appropriately, with the pairing (K | d) = 1 linking the central and derivation elements.[5]This basic inner product is closely tied to the dual Coxeter number g (also denoted h^\vee) of \mathfrak{g}, which scales the relationship between the Killing form and the normalized form on \mathfrak{g}: specifically, B(x, y) = 2g (x | y) for x, y \in \mathfrak{g}.[17] In the affine setting, g appears in properties such as the trace anomaly or level shifts in representations, where the effective level is adjusted by -g relative to the bare central charge from K.[18]The canonical central element K resides in the Cartan subalgebra of \hat{\mathfrak{g}} and is given explicitly by K = \sum h_i, where the sum is over a basis of coroots h_i of the finite-dimensional algebra \mathfrak{g} (corresponding to the simple roots). This element commutes with all of \hat{\mathfrak{g}} under the Lie bracket, reflecting its role in the central extension, and the normalization of the form on \mathfrak{g} together with (K | d) = 1 fixes the scale of the basic inner product.[5]
Root system and simple roots
The root space decomposition of an untwisted affine Lie algebra \hat{\mathfrak{g}} associated to a finite-dimensional simple Lie algebra \mathfrak{g} with root system \Delta is given by \hat{\mathfrak{g}} = \hat{\mathfrak{h}} \oplus \bigoplus_{\alpha \in \hat{\Delta}} \hat{\mathfrak{g}}_\alpha, where \hat{\mathfrak{h}} is the affine Cartan subalgebra and \hat{\Delta} is the affine root system.[5] The affine root system consists of \hat{\Delta} = \{ \alpha + n\delta \mid \alpha \in \Delta, n \in \mathbb{Z} \} \cup \{ n\delta \mid n \in \mathbb{Z} \setminus \{0\} \}, where \delta is the basic imaginary root corresponding to the central extension.[3] Each root space \hat{\mathfrak{g}}_\alpha is one-dimensional for real roots and has dimension equal to the rank of \mathfrak{g} for imaginary roots.[19]The roots in \hat{\Delta} are classified into real roots and imaginary roots. Real roots are those of the form \alpha + n\delta for \alpha \in \Delta and n \in \mathbb{Z}, each with positive squared length (\alpha + n\delta, \alpha + n\delta) > 0 under the invariant bilinear form, and they support one-dimensional root spaces generated by elements like e_{\alpha + n\delta}.[5] Imaginary roots are the nonzero multiples n\delta for n \in \mathbb{Z} \setminus \{0\}, which have zero squared length (n\delta, n\delta) = 0 and correspond to nilpotent subalgebras of dimension equal to the rank of the finite-dimensional algebra.[3] The bilinear form used here, which is nondegenerate on the real span of the roots, is defined in detail separately but induces the structure on the root spaces.[5]A choice of simple roots for the affine root system is the set \hat{\Pi} = \{\alpha_0, \alpha_1, \dots, \alpha_l \}, where \{\alpha_1, \dots, \alpha_l\} are the simple roots of the finite root system \Delta, and \alpha_0 = \delta - \theta with \theta the highest root of \mathfrak{g}.[19] This choice ensures that every root is a non-negative integer combination of the simple roots for the positive subsystem, and the affine Dynkin diagram extends the finite one by adding the node for \alpha_0.[3] The generalized Cartan matrix A = (a_{ij}) for $0 \leq i,j \leq l is defined by a_{ij} = 2 (\alpha_i | \alpha_j) / (\alpha_j | \alpha_j), which is symmetrizable and of affine type.[5]The finite Weyl group W of \mathfrak{g} acts on the affine roots by permuting the finite part \alpha, preserving the \delta component, and thus orbits under W consist of sets like \{ w(\alpha) + n\delta \mid w \in W \} for fixed \alpha \in \Delta and n \in \mathbb{Z}.[5] The full affine Weyl group extends this action by including translations, but the basic orbits on real roots are generated by these finite Weyl group actions.[3] Imaginary roots are fixed by the Weyl group action.[19]
Representations
Highest weight modules
Highest weight modules for affine Lie algebras generalize the corresponding notion from finite-dimensional semisimple Lie algebras, providing a framework for studying infinite-dimensional representations. A highest weight module is generated by a single vector v_\lambda, the highest weight vector, of weight \lambda \in \hat{\mathfrak{h}}^* (with \lambda|_{\mathfrak{h}} \in \mathfrak{h}^*, \lambda(K) = k, \lambda(d) = 0), annihilated by all positive root vectors e_\alpha for \alpha \in \hat{\Delta}^+. These modules are \mathbb{Z}-graded, with the grading induced by the action of the derivation d, and they form the building blocks for the representation theory of affine Lie algebras.[11]The universal highest weight module, known as the Verma module M(\lambda), is obtained by inducing from a one-dimensional \hat{\mathfrak{h}}-module on which \hat{\mathfrak{h}} acts via the weight \lambda and the positive part \hat{\mathfrak{n}}^+ acts trivially.[11] The character of the Verma module is given by\mathrm{ch} M(\lambda) = \frac{L(\lambda + \rho, 0)}{\prod_{\alpha \in \Delta^+} (1 - e^{-\alpha})} \frac{1}{\eta(q)^l},where L(\mu, 0) = \sum_{w \in W} \epsilon(w) e^{w \mu} is the Weyl numerator for the finite-dimensional root system evaluated at \mu, \prod_{\alpha \in \Delta^+} (1 - e^{-\alpha}) is the finite-dimensional denominator, \eta(q) is the Dedekind eta function, and l is the rank of the underlying finite-dimensional Lie algebra.[11] This formula reflects the separation between the finite-dimensional "classical" weights and the infinite-dimensional loop-like contributions captured by the eta function. For generic \lambda, the finite part corresponds to the irreducible finite-dimensional representation.Verma modules are typically reducible, containing proper submodules generated by singular vectors, which are nonzero vectors annihilated by \hat{\mathfrak{n}}^+ but not scalar multiples of the highest weight vector.[11] The existence of such singular vectors occurs when \lambda satisfies certain reflection conditions related to the affine Weyl group, specifically when \langle \lambda + \rho, \alpha_i^\vee \rangle = m for some integer m \leq 0 and simple coroot \alpha_i^\vee, leading to embeddings M(\lambda) \hookrightarrow M(s_i \cdot \lambda) via the Weyl group reflection s_i.[11] These conditions determine the submodule lattice and the structure of irreducible quotients, with integrability requiring \lambda to lie in the dominant Weyl chamber to ensure finite-dimensional weight spaces and positive energy representations.In the classical limit as the level k \to \infty, highest weight modules of the affine Lie algebra recover the finite-dimensional irreducible representations of the underlying finite-dimensional simple Lie algebra. This limit highlights the deformation aspect of affine structures, where the central extension parameter k scales the representations toward their finite-type counterparts.
Integrable representations at level k
In the theory of representations of affine Lie algebras, a highest weight module is said to be at level k if the central element K in the Cartan subalgebra acts as multiplication by the positive integer k on every vector in the module.[20] For such modules, the highest weight \lambda must satisfy the admissibility condition (\lambda \mid \theta) \leq k, where \theta is the highest root of the underlying finite-dimensional simple Lie algebra \mathfrak{g}, ensuring the representation is well-defined and finite-dimensional in certain senses.[21] These level k modules generalize the finite-dimensional representations of \mathfrak{g} and play a crucial role in applications to conformal field theory and vertex operator algebras.The integrable highest weight representations at level k are precisely the irreducible highest weight modules L(\lambda) where \lambda is a dominant integral weight for \mathfrak{g} lying inside the fundamental alcove of the affine root system. Specifically, these are the weights \lambda satisfying $0 \leq (\lambda + \rho \mid \alpha_i^\vee) \leq k for all simple coroots \alpha_i^\vee of the affine Lie algebra, where \rho is the Weyl vector (half-sum of positive roots).[20] This classification ensures that the representation is integrable, meaning that for every real root, the corresponding \mathfrak{sl}_2-triple acts via a finite-dimensional representation of \mathrm{SL}(2, \mathbb{C}).[21] The number of such representations grows polynomially with k, reflecting the bounded region of the alcove.The characters of these integrable modules are given by the Weyl-Kac character formula:\mathrm{ch}\, L(\lambda) = \frac{\sum_{w \in \hat{W}} \epsilon(w) e^{w(\lambda + \rho)}}{\eta(q)^l P(q)},where \hat{W} is the affine Weyl group, \epsilon(w) is the sign of w, \eta(q) is the Dedekind eta function, l is the rank of \mathfrak{g}, and P(q) is the denominator involving string functions accounting for the multiplicities of imaginary roots.[20] This formula provides a closed-form expression for the generating function of weights and multiplicities, analogous to the finite-dimensional Weyl character formula but extended to the infinite-dimensional setting.The tensor product of two integrable representations at level k decomposes as a direct sum of other integrable representations at the same level k, governed by fusion rules whose coefficients are nonnegative integers independent of the underlying field theory. These rules form a fusion ring and can be computed explicitly for low ranks using combinatorial methods.[22]
Vertex operator algebras
Vertex operator algebras provide an axiomatic framework that captures the algebraic structure underlying representations of affine Lie algebras, particularly through the vacuum module. For a simple finite-dimensional Lie algebra \mathfrak{g}, the vacuum module V^k(\mathfrak{g}) at positive integer level k is constructed as the quotient of the universal enveloping algebra of the affine Lie algebra \hat{\mathfrak{g}} by the maximal ideal generated by the central element minus k times the identity and the positive part of the Borel subalgebra. This module carries a natural structure of a graded vertex operator algebra (VOA), where the grading is by L_0-eigenvalues (conformal weights), the vacuum vector is the highest weight vector of weight 0, and the conformal vector implements the Virasoro algebra action. The vertex operators Y(a, z): V^k(\mathfrak{g}) \to V^k(\mathfrak{g})[[z, z^{-1}]]\{z\} for a \in V^k(\mathfrak{g}) satisfy the key axioms of a VOA, including the Jacobi identity z_0^{-1} \delta\left(\frac{z_1 - z_0}{z_0}\right) Y(a, z_1) Y(b, z_0) - (-z_0)^{-1} \delta\left(\frac{z_0 - z_1}{-z_0}\right) Y(b, z_0) Y(a, z_1) = z_1^{-1} \delta\left(\frac{z_1 - z_0}{z_1}\right) Y(Y(a, z_1 - z_0) b, z_0) and the translation property [L_{-1}, Y(a, z)] = \frac{d}{dz} Y(a, z), ensuring the locality and associativity properties essential for operator product expansions.A defining feature of the affine VOA V^k(\mathfrak{g}) is the operator product expansion (OPE) governing the vertex operators, which encodes the commutation relations of the affine algebra. For root vectors e_\alpha, e_{-\alpha} corresponding to a root \alpha \in \Delta, the OPE takes the form Y(e_\alpha, z) Y(e_{-\alpha}, w) \sim \frac{(e_\alpha e_{-\alpha})(w)}{z - w} + \ regular\ terms, where (e_\alpha e_{-\alpha}) is the evaluation of the bilinear form on the module at w, reflecting the singular part of the correlation functions. This OPE structure extends to all elements and underpins the locality axiom, (z - w)^{2\ell(a,b) - \resdeg Y(a,z) Y(b,w)} Y(a, z) Y(b, w) = (z - w)^{2\ell(b,a) - \resdeg Y(b,w) Y(a,z)} Y(b, w) Y(a, z) + o(1), with locality degree \ell(a,b). The Virasoro algebra acts on V^k(\mathfrak{g}) via the Sugawara construction, yielding operators L_n = \frac{1}{2(k + g^\vee)} \sum_{m \in \mathbb{Z}} :J_m^a J_{n-m}^a :, where \{J_n^a\} are the modes of the affine current fields J^a(z) = \sum J_n^a z^{-n-1} spanning the Lie algebra basis with Killing form pairing, and :\ : denotes normal ordering. The central charge of this Virasoro action is c = \frac{k \dim \mathfrak{g}}{k + g^\vee}, where g^\vee is the dual Coxeter number of \mathfrak{g}, determining the anomaly in two-dimensional conformal theories.Affine VOAs at positive integer levels exhibit strong structural properties, including locality, which ensures that vertex operators commute up to regular terms, and rationality, meaning that every admissible module is completely reducible and finitely generated as a module over the VOA. Specifically, for k \in \mathbb{Z}_{>0}, V^k(\mathfrak{g}) is rational, with the category of its modules being semisimple and modular invariant under the action of the modular group SL(2,\mathbb{Z}). This rationality follows from the unitarizability and the fusion rules being given by the Verlinde formula, restricting the admissible levels to integrable representations. These properties make affine VOAs prototypical examples of rational chiral algebras in conformal field theory.[23]
Advanced topics
Affine Weyl group
The affine Weyl group \hat{W} of an affine Lie algebra is defined as the semidirect product \hat{W} = [W](/page/W) \ltimes Q^\vee, where W is the finite Weyl group associated to the underlying finite-dimensional simple Lie algebra, and Q^\vee is the coroot lattice generated by the coroots \alpha_i^\vee of the simple roots \alpha_i.[24][25] This group is infinite and acts faithfully on the real vector space \mathfrak{h}_\mathbb{R} underlying the Cartan subalgebra \mathfrak{h}. It is generated by the simple reflections r_i for i = 0, 1, \dots, l, where l is the rank of the finite Lie algebra, and these generators satisfy the relations of the affine Coxeter system corresponding to the affine Dynkin diagram, including r_i^2 = 1 and braid relations (r_i r_j)^{m_{ij}} = 1 with m_{ij} determined by the diagram edges.[24][25]The affine Weyl group acts on the dual space \hat{\mathfrak{h}}^* of the extended Cartan subalgebra \hat{\mathfrak{h}} = \mathfrak{h} \oplus \mathbb{C} c \oplus \mathbb{C} d, where weights \lambda \in \hat{\mathfrak{h}}^* have fixed level k = \lambda(c). This action is level-preserving and given by the twisted (or dot) action: for w \in W,w \cdot \lambda = w(\lambda + \rho) - \rho + k \Lambda_0,where \rho is the half-sum of the positive roots (Weyl vector) and \Lambda_0 is the basic imaginary weight with \Lambda_0(c) = 1 and vanishing pairings with roots and d.[24][26] The full group action combines this with translations by elements of Q^\vee, preserving the lattice of integrable weights at level k.A key feature of this action is the decomposition of \hat{\mathfrak{h}}^* into alcoves, which are fundamental domains bounded by affine hyperplanes \{\lambda \mid (\lambda \mid \alpha_i^\vee) = m\} for integers m and simple coroots \alpha_i^\vee (including the affine one). For integrable highest weight representations at positive integer level k, the fundamental alcove is the region\Delta = \{ \lambda \mid 0 < (\lambda \mid \alpha_i^\vee) < k \ \forall i = 1, \dots, l, \ (\lambda \mid \theta^\vee) > -k \},where \theta is the highest root of the finite Lie algebra, and the affine Weyl group acts simply transitively on the set of alcoves, with integrable weights lying in the orbit of dominant weights within \Delta.[24][27]The translation subgroup consists of elements t^\alpha for \alpha \in Q^\vee, acting on weights as t^\alpha(\lambda) = \lambda + \alpha (where \alpha \subset \mathfrak{h}^*, extended trivially on the \mathbb{C} \Lambda_0 \oplus \mathbb{C} \delta components). In the level-preserving dot action, this becomes t^\alpha \cdot \lambda = \lambda + \alpha - ((\lambda + \rho) \mid \alpha) \delta. These translations commute with the finite Weyl group action in the semidirect product and generate shifts between parallel alcoves, facilitating the description of weight lattices modulo the root lattice.[24][25]
Character formulas
The Weyl-Kac character formula provides a closed-form expression for the characters of irreducible highest weight modules of affine Lie algebras, generalizing the classical Weyl character formula to the infinite-dimensional setting. For a dominant integral weight \lambda, the character \mathrm{ch} L(\lambda) of the irreducible module L(\lambda) is given by\mathrm{ch} L(\lambda) = \frac{\sum_{w \in \hat{W}/W} \epsilon(w) e^{w(\lambda + \rho)}}{\prod_{\alpha > 0} (1 - e^{-\alpha})},where \hat{W} is the affine Weyl group, W is the finite Weyl group of the underlying finite-dimensional Lie algebra, \epsilon(w) is the sign of the Weyl group element w, \rho is the Weyl vector (half the sum of positive roots), and the product runs over all positive roots \alpha of the affine root system. This formula holds formally as a power series in the variables e^{-\alpha}, accounting for the infinite number of roots in the affine case through the structure of the cosets \hat{W}/W.The denominator in the formula satisfies the affine denominator identity, which is the analogue of the Weyl denominator formula for finite-dimensional Lie algebras:\prod_{\alpha > 0} (1 - e^{-\alpha}) = \sum_{w \in \hat{W}} \epsilon(w) e^{w \rho}.For untwisted affine Lie algebras, this identity specializes to a product involving the classical Weyl denominator for the finite root system and the Dedekind eta function \eta(q) raised to the power of the dimension of the finite-dimensional Lie algebra, reflecting the contribution from imaginary roots. This identity underpins many combinatorial applications, such as the Macdonald identities, and ensures the rationality of the characters in the formal variable q = e^{-\delta}, where \delta is the basic imaginary root.For non-integrable highest weight modules, where \lambda is not a dominant weight at a positive integer level, the characters decompose into sums involving string functions. The string function \sigma_{\mu,\xi}(q) associated to an integrable module L(\mu) and a weight \xi in the fundamental Weyl chamber is the generating function \sigma_{\mu,\xi}(q) = \sum_{n=0}^\infty m_{\mu}(\xi - n \delta) q^n, where m_{\mu}(\cdot) denotes weight multiplicities. The full character then expresses as \mathrm{ch} L(\lambda) = \sum_{\xi \in \max(\lambda)} \sigma_{\lambda,\xi}(e^{-\delta}) e^{\xi}, with the sum over maximal weights in the Weyl orbit; this decomposes the infinite-dimensional graded components into cosets modulo the coroot lattice, adapting the Weyl-Kac formula via recursive relations on folded root systems.[28]Under the principal gradation, where the Cartan subalgebra is graded primarily by the derivation element, the characters of integrable modules specialize to expressions involving Jacobi theta functions. For instance, in the principal gradation of an affine Lie algebra \hat{\mathfrak{g}}, the character of a highest weight module decomposes as a finite linear combination of theta functions \theta_{\Lambda}(q, z) = \sum_{m \in \mathbb{Z}} q^{m^2/2} z^m associated to the finite-dimensional weights, modulated by the level and adjusted for the affine extension; this form highlights the modular properties intrinsic to the theta series while aligning with the general Weyl-Kac structure.[29]
Modular invariance
The characters of integrable highest weight modules for an affine Lie algebra at positive integer level k exhibit modular invariance under the action of the modular group \mathrm{SL}(2,\mathbb{Z}), which is generated by the transformations T: \tau \mapsto \tau + 1 and S: \tau \mapsto -1/\tau. The character is defined as \chi_\lambda(\tau) = \mathrm{Tr}_{L(\lambda)} q^{L_0 - c/24}, where q = e^{2\pi i \tau}, L_0 is the zeroth Virasoro mode from the Sugawara construction, and c = k \dim \mathfrak{g} / (k + g) is the central charge with g the dual Coxeter number. Under T, the transformation is diagonal: \chi_\lambda(\tau + 1) = e^{2\pi i [(\lambda, \lambda + 2\rho)/2(k + g) - c/24]} \chi_\lambda(\tau), where \rho is the Weyl vector. Under S, the characters mix via the unitary modular S-matrix: \chi_\lambda(-1/\tau) = \sum_\mu S_{\lambda\mu} \chi_\mu(\tau).[29]The modular S-matrix elements are given byS_{\lambda\mu} = \frac{1}{\sqrt{|W|(k + g)^r}} \sum_{w \in W} \epsilon(w) \exp\left( \frac{2\pi i (w(\lambda + \rho), \mu + \rho)}{k + g} \right),where W is the finite Weyl group of the underlying finite-dimensional Lie algebra, r is the rank, and \epsilon(w) is the sign of w. This formula arises from expressing the characters as ratios of theta functions associated to the root lattice, whose modular properties follow from the transformation laws of theta functions under \mathrm{SL}(2,\mathbb{Z}). The span of the normalized characters \{\chi_\lambda / \eta(\tau)^{\dim \mathfrak{g}}\}_{\lambda \in P_k^+} (with \eta the Dedekind eta function) forms a finite-dimensional unitary representation of the modular group, proving the modular invariance; this representation is irreducible for simply-laced algebras and decomposes otherwise into known factors.[29]A key consequence of this modular invariance is the Verlinde formula, which computes the fusion multiplicities (dimensions of intertwining spaces) N_{\lambda\mu}^\nu among integrable modules:N_{\lambda\mu}^\nu = \sum_{\rho \in P_k^+} \frac{S_{\lambda\rho} S_{\mu\rho} S_{\nu\rho}^*}{S_{0\rho}},where the sum runs over integrable weights at level k, and S^* denotes complex conjugate (with S unitary). This formula derives from the modular S-matrix via residue calculus on the characters or homological interpretations of conformal blocks, providing a purely algebraic expression for fusion rules without reference to vertex operator algebra structure. It was originally conjectured in the physics literature and rigorously proved using representation theory of affine Lie algebras.[30]Modular invariance extends to partition functions, which are bilinear forms \sum_{\lambda \in P_k^+} |\chi_\lambda(\tau)|^2 summing over all integrable representations (or modular-invariant combinations thereof). These functions are invariant under \mathrm{SL}(2,\mathbb{Z}) because the S-matrix is unitary (S^\dagger = S^{-1}) and the T-action is diagonal with phases ensuring overall invariance. Such partition functions classify consistent toroidal compactifications in physical applications but mathematically correspond to positive definite Hermitian forms on the space of characters.[29]
Applications
In physics: Conformal field theory
Affine Lie algebras play a central role in two-dimensional conformal field theories (CFTs), particularly as the symmetry algebras underlying Wess-Zumino-Witten (WZW) models. These models describe sigma models on the manifold of a compact Lie group G with Lie algebra \mathfrak{g}, augmented by a topological Wess-Zumino term that ensures conformal invariance at the quantum level. The level k of the model, an integer for compact groups, quantifies the coupling to this term and determines the representation theory of the theory.[31]The chiral symmetry of the WZW model is generated by left- and right-moving currents transforming in representations of the affine Lie algebra \hat{\mathfrak{g}}_k at level k. These currents satisfy the operator product expansion (OPE)J^a(z) J^b(w) \sim \frac{k \delta^{ab}}{(z-w)^2} + \frac{i f^{abc} J^c(w)}{z-w},where f^{abc} are the structure constants of \mathfrak{g}, encoding the current algebra structure of the affine Kac-Moody symmetry. This algebra extends the Virasoro algebra, yielding a central chargec = \frac{k \dim \mathfrak{g}}{k + h^\vee},with h^\vee the dual Coxeter number of \mathfrak{g}, which governs the anomaly and ensures the theory is unitary for positive integer k.[31][32]Primary fields in the WZW model transform in the integrable highest-weight representations of \hat{\mathfrak{g}}_k, labeled by weights [\lambda](/page/Lambda) satisfying (\lambda, \theta) \leq k, where \theta is the longest root of \mathfrak{g}. These fields \Phi_\lambda obey the OPEJ^a(z) \Phi_\lambda(w) \sim \frac{t^a(\lambda) \Phi_\lambda(w)}{z-w} + \cdots,with t^a(\lambda) the representation matrices, and their correlation functions are constructed using conformal blocks derived from the representation characters, facilitating computations of scattering amplitudes and partition functions.[31]WZW models exemplify rational CFTs, where the affine Lie algebra at integer level k generates a rational vertex operator algebra (VOA) with a finite number of irreducible modules, leading to unitary theories with modular-invariant partition functions classified by ADE diagrams for groups like SU(2). This rationality underpins applications in integrable systems and condensed matter physics, such as the quantum Hall effect.[31][33]
In mathematics: Geometric representation theory
Geometric representation theory realizes affine Lie algebras through geometric objects such as affine Grassmannians and quiver varieties, providing tools to study their representations via algebro-geometric methods. The geometric Satake equivalence establishes a monoidal equivalence between the category of representations of the Langlands dual group and the category of perverse sheaves on the affine Grassmannian associated to a complex reductive group G. The affine Grassmannian \mathrm{Gr}_G, defined as the quotient G(\mathbb{C}((t)))/G(\mathbb{C}[]), parametrizes G-bundles on the formal disk with certain modification data, thereby geometrizing the irreducible representations of the dual affine Lie algebra \hat{\mathfrak{g}}^\vee. This equivalence, initially conjectured by Drinfeld and proven for complex coefficients by Mirković and Vilonen, extends the classical Satake isomorphism to a categorical level and has been generalized to arbitrary commutative rings.[34]For integrable representations at positive integer level k, the mirabolic affine Grassmannian—a subvariety parametrizing bundles with fixed trivialization on a formal neighborhood—plays a central role, with its category of perverse sheaves equivalent to the semisimple category of level-k representations via the mirabolic category. This construction links the geometry of loop groups to the finite-dimensional weight spaces in affine representation theory.[35]Quiver representations offer another geometric model, where modules over an affine Lie algebra \hat{\mathfrak{[g](/page/G)}} correspond to representations of the associated affine Dynkin quiver Q, with dimension vectors serving as weights in the root system of \hat{\mathfrak{[g](/page/G)}}. Ringel's Hall algebra approach constructs the positive part of the universal enveloping algebra from the extension groups of quiver representations, yielding structure constants that match the Serre relations for affine Kac-Moody algebras; this is extended to tame affine quivers, providing a combinatorial basis for representations.[36]Loop group actions on the Beilinson-Drinfeld Grassmannian, a relative version of the affine Grassmannian over a curve, geometrize opers—flat connections on principal G-bundles with nilpotent residues—and relate to the affine Weyl group through the monodromy and spectral curves of these objects. Opers parametrize solutions to integrable hierarchies like the Drinfeld-Sokolov system, with the Grassmannian's geometry reflecting the Weyl group orbits in the loop algebra.Nakajima quiver varieties furnish a hyper-Kähler resolution of the affine quotient of the representation space of a framed quiver, realizing the Grothendieck ring of integrable representations of the affine Lie algebra at positive levels through their intersectioncohomology. These varieties, constructed as GIT quotients stabilized by moment map levels, provide a geometric framework for crystal bases and tensor product multiplicities in affine representation theory.
Other uses
Affine root systems arising from affine Lie algebras play a significant role in combinatorics, particularly in the study of crystal bases and path models for representation characters. Crystal bases provide a combinatorial framework for understanding the structure of irreducible highest weight modules over quantum affine algebras, where the affine root system encodes the branching rules and tensor product decompositions via Kashiwara operators. For instance, the Young wall model realizes these crystal bases explicitly for quantum affine algebras, offering a bijection between certain combinatorial objects and the basis elements of the modules. Similarly, path models, such as extensions of Littelmann paths to affine settings, model the crystal graphs of basic representations, allowing computation of characters through weighted paths in the affine root lattice. These tools facilitate the enumeration of tableaux and walls that correspond to the Weyl group orbits in affine types, bridging combinatorial enumeration with character formulas.[37][38]In number theory, affine Kac-Moody groups defined over p-adic fields extend classical reductive groups and connect to arithmetic structures, including relations with Drinfeld modules in function field settings. These groups, constructed as loop groups over p-adic fields, admit smooth representations analogous to those of finite-dimensional p-adic groups, with applications to the local Langlands program. The associated affine root systems govern the Iwahori-Hecke algebras and spherical functions on these groups, providing tools for studying unitary representations and their characters. Drinfeld's realizations of quantum affine algebras further link these structures to Drinfeld modules, where the infinite-dimensional Lie algebra aspects inform the theory of elliptic modules and t-motives over global function fields.[39][40]Automorphic forms on affine Kac-Moody groups over local fields incorporate Eisenstein series defined using the geometry of affine buildings, establishing analogies with classical modular forms. These Eisenstein series are constructed via induced representations from minimal parabolic subgroups, with their constant terms computed using the affine Weyl group action on the building's apartments. The affine building, a generalization of the Bruhat-Tits building, serves as the space for these forms, where the spherical Hecke algebra acts compatibly with the affine root system. This framework links to modular forms through the constant term expansions and intertwining operators, offering insights into the spectral decomposition of L²-spaces on these groups.[39]In the theory of cluster algebras, affine types provide monoidal categorifications via Hall algebras of quiver representations, where the affine Dynkin diagrams determine the exchange relations and cluster variables. The Hall algebra of coherent sheaves on the affine projective line, for example, categorifies the cluster algebra of affine type A, with simple modules corresponding to cluster monomials. Quantum affine algebras arise naturally in this context, with their positive part bases matching the Laurent phenomenon in affine cluster algebras through categorification functors. This connection, established via quiver Hecke algebras, enables positivity results for cluster variables in affine types and links to canonical bases in representation theory.[41][42]
Examples
Affine \mathfrak{sl}_2
The untwisted affine Lie algebra \hat{\mathfrak{sl}}_2 associated to the simple Lie algebra \mathfrak{sl}_2 provides a fundamental example of an affine Kac-Moody algebra of type A_1^{(1)}. It arises as a central extension of the loop algebra \mathfrak{sl}_2 \otimes \mathbb{C}[t, t^{-1}] and plays a key role in understanding the general structure of untwisted affine algebras. This algebra is generated by elements that extend the basis of \mathfrak{sl}_2, which consists of e, f, and h satisfying [h, e] = 2e, [h, f] = -2f, and [e, f] = h, with the invariant bilinear form normalized such that (h, h) = 2.The basis of \hat{\mathfrak{sl}}_2 includes the tensor products e \otimes t^n, f \otimes t^n, h \otimes t^n for all integers n \in \mathbb{Z}, along with the central element K and the degree operator d. The Lie bracket on the loop algebra part is given by[X \otimes t^m, Y \otimes t^n] = [X, Y] \otimes t^{m+n} + m \delta_{m+n, 0} (X, Y) Kfor X, Y \in \mathfrak{sl}_2, while additional relations include [d, X \otimes t^n] = n (X \otimes t^n), [K, \hat{\mathfrak{sl}}_2] = 0, and [d, K] = 0. Representative commutation relations from the finite algebra extend directly, such as [h \otimes t^m, e \otimes t^n] = 2 e \otimes t^{m+n} and [h \otimes t^m, f \otimes t^n] = -2 f \otimes t^{m+n}. The generalized Cartan matrix for this algebra is\begin{pmatrix}
2 & -1 \\
-1 & 2
\end{pmatrix},with simple roots \alpha_0 = \delta - \alpha and \alpha_1 = \alpha, where \alpha is the simple root of \mathfrak{sl}_2 and \delta is the basic imaginary root satisfying \delta(d) = 1 and \delta(K) = 0.The root system of \hat{\mathfrak{sl}}_2 consists of real roots \pm 2 + n \delta for n \in \mathbb{Z} and nonzero imaginary roots n \delta for n \in \mathbb{Z} \setminus \{0\}. The real roots correspond to the Weyl group orbit of the simple roots under the affine Weyl group action, preserving the nondegenerate restriction of the invariant form to the real root hyperplanes, while the imaginary roots lie in the radical of the form.For integrable representations at a positive integer level k \in \mathbb{Z}_{>0}, the highest weights are the dominant integral weights \lambda satisfying (\lambda, \theta) \leq k, where \theta is the highest root of \mathfrak{sl}_2. These are precisely the weights j \varpi_1 for $0 \leq j \leq k, where \varpi_1 is the fundamental weight with \varpi_1(h) = 1. Such representations are irreducible, highest weight modules that admit a vertex operator algebra structure at level k.
The affine \hat{\mathfrak{su}}(2)_k at positive integer level k is the compact real form of the untwisted affine Lie algebra corresponding to the finite-dimensional Lie algebra \mathfrak{su}(2). In this realization, the generators are taken to be compact, meaning they satisfy the commutation relations of the affine algebra with a bilinear form that extends the negative definite Killing form of \mathfrak{su}(2) to the real subspace, while the central extension introduces degeneracy along the imaginary root direction. This structure ensures unitarity in the associated representations, making it particularly suitable for applications in conformal field theory.The integrable representations of \hat{\mathfrak{su}}(2)_k are the highest-weight modules labeled by spins j/2 where j = 0, 1, \dots, k, corresponding to the finite number of weights inside the level-k alcove of the weight lattice. These modules are unitary and finite-dimensional at each graded level, with the highest-weight vector annihilated by the positive-mode generators.[31]The fusion rules for these representations, which govern the tensor product decomposition in the category of integrable modules, are given byN_{j_1 j_2}^{j_3} = 1if |j_1 - j_2| \leq j_3 \leq j_1 + j_2, j_3 \equiv j_1 + j_2 \pmod{2}, and j_3 \leq k, and N_{j_1 j_2}^{j_3} = 0 otherwise. This truncation of the classical SU(2) Clebsch-Gordan coefficients arises from the Verlinde formula applied to the modular S-matrix of the theory.[43]The characters of these integrable representations, which encode the graded dimensions, are explicitly\chi_j(q) = \frac{q^{j(j+1)/4(k+2)} - q^{(k-j)(k-j+2)/4(k+2)}}{(q^{1/2};q)_\infty (q^{k+2};q)_\infty / (q^{(k+2)/2};q)_\infty},where (a;q)_\infty = \prod_{n=0}^\infty (1 - a q^n) is the q-Pochhammer symbol; this is the principally specialized form of the Weyl-Kac character formula specialized to the integrable sector. These characters satisfy modular invariance under the action of SL(2,\mathbb{Z}) when combined into partition functions.In physics, \hat{\mathfrak{su}}(2)_k underlies the SU(2)_k Wess-Zumino-Witten (WZW) model, a rational conformal field theory with central charge c = 3k/(k+2), where the primary fields transform in the integrable representations and exhibit the above fusion rules. This model describes non-Abelian anyons in certain fractional quantum Hall states and serves as a building block for integrable spin chains via coset constructions, such as the critical SU(2) Heisenberg chains at integrability points.[31]