A Lie algebra is a vector space \mathfrak{g} over a field K (typically \mathbb{R} or \mathbb{C}) equipped with a bilinear operation [\cdot, \cdot]: \mathfrak{g} \times \mathfrak{g} \to \mathfrak{g}, called the Lie bracket, that is alternating (i.e., [x, x] = 0 for all x \in \mathfrak{g}) and satisfies the Jacobi identity [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0 for all x, y, z \in \mathfrak{g}.[1] This structure captures the algebraic essence of infinitesimal transformations and symmetries in a non-associative manner, distinguishing it from associative algebras.[2]Lie algebras are intimately connected to Lie groups, smooth manifolds that form groups under a compatible multiplication operation; the Lie algebra of a Lie group G is defined as the tangent space at the identity element e \in G, endowed with the Lie bracket derived from the commutator of left-invariant vector fields.[3] This correspondence, known as the Lie algebra of a Lie group, linearizes the nonlinear geometry of the group, facilitating the study of continuous symmetries through linear algebra.[4] For matrix Lie groups, such as GL(n, \mathbb{R}) or SO(n), the Lie algebra consists of matrices A satisfying certain conditions, with the bracket given by [A, B] = AB - BA.[5]The theory of Lie algebras originated in the late 19th century with the work of Sophus Lie, who developed it as part of his study of transformation groups, now called Lie groups, to analyze continuous symmetries in differential equations.[6] Key examples include the Heisenberg algebra, modeling quantum mechanics, and classical algebras like \mathfrak{sl}(n, \mathbb{C}), the special linear Lie algebra of trace-zero matrices.[7] Over the complex numbers, finite-dimensional semisimple Lie algebras are completely classified into four infinite families (types A_n, B_n, C_n, D_n) and five exceptional types (G_2, F_4, E_6, E_7, E_8), a result due to Killing and Cartan that underpins much of modern representation theory.[8]Lie algebras play a central role in diverse fields, including differential geometry for studying foliations and connections, algebraic geometry via algebraic groups, and theoretical physics for symmetry principles in particle physics (e.g., SU(3) for quantum chromodynamics) and general relativity.[9] Their representations, homomorphisms into \mathfrak{gl}(V) (general linear Lie algebra of endomorphisms of a vector space V), provide tools for decomposing complex systems into irreducible components, with profound implications for quantum field theory and integrable systems.
History and Motivations
Historical Development
The concept of Lie algebras originated from Sophus Lie's pioneering work on continuous transformation groups in the 1870s, where he sought to generalize Galois's discrete group theory to continuous symmetries for solving differential equations. Lie's investigations led him to the idea of infinitesimal transformations, which generate these groups and form the algebraic structure now known as a Lie algebra. In his key publications during the 1880s, such as the 1888 paper on transformation groups, Lie formalized these infinitesimal generators as vector fields satisfying certain commutation relations, laying the groundwork for the modern theory.[10][11]Building on Lie's ideas, Wilhelm Killing independently developed the structure of Lie algebras in his studies of non-Euclidean geometry during 1888–1890. In a series of memoirs published in the Mathematische Annalen, Killing classified the simple Lie algebras over the complex numbers, introducing the concept of root systems to describe their structure and identifying the four infinite families along with the five exceptional types. His work provided the first systematic classification, though it contained some gaps that were later addressed.[12]Élie Cartan refined and completed Killing's classification in his 1894 doctoral thesis, Sur la structure des groupes de transformations finis et continus, offering a rigorous proof for simple Lie algebras over the complex numbers using the Killing form and Cartan subalgebras. Cartan extended this in 1913–1914 by classifying the real forms of these complex Lie algebras, distinguishing compact, split, and other forms based on their geometric properties. These advancements solidified the algebraic framework.In the mid-20th century, Harish-Chandra advanced the representation theory of semisimple Lie algebras through a series of papers in the 1950s, including his 1951 work on the universal enveloping algebra and 1953 studies on unitary representations, introducing tools like Harish-Chandra modules to analyze infinite-dimensional representations. Following the development of quantum mechanics in the 1920s, Lie algebras evolved into essential tools for describing symmetries in physical systems, such as angular momentum in quantum theory and gauge symmetries in particle physics models like the Standard Model from the 1970s onward. Lie algebras serve as infinitesimal approximations to Lie groups, capturing the local structure of continuous symmetries in these applications.[13][14][15]
Connections to Lie Groups and Symmetry
Lie algebras arise naturally as the tangent spaces to Lie groups at their identity element, providing a linear algebraic framework to study the infinitesimal generators of continuous symmetries represented by the group. In this context, the Lie bracket operation encodes the commutation relations among these infinitesimal transformations, allowing for the algebraic description of symmetry operations that are otherwise captured geometrically by the manifold structure of the Lie group. This perspective shifts the focus from the global, nonlinear nature of Lie groups to their local, linearized approximations, facilitating computations in areas such as differential geometry and representation theory.[16][17]A key motivation for Lie algebras stems from their role in Noether's theorem, which establishes a profound link between continuous symmetries of physical systems and conserved quantities. Specifically, symmetries generated by Lie group actions on the Lagrangian of a system correspond to Lie algebra elements that yield conservation laws through the theorem's machinery, such as angular momentum conservation from rotational invariance. This connection underscores how Lie algebras capture the "infinitesimal" aspects of symmetries, enabling the derivation of conserved currents and charges in classical and quantum mechanics.[18][19]In physics, Lie algebras find prominent applications in describing symmetry groups like the rotation group SO(3), whose Lie algebra so(3) consists of skew-symmetric matrices representing infinitesimal rotations in three-dimensional space, essential for understanding rigid body dynamics and quantum spin. Similarly, the Lie algebra of the Lorentz group, so(3,1), underpins the symmetries of spacetime in special relativity, where its generators correspond to boosts and rotations that preserve the Minkowski metric, facilitating the formulation of relativistic invariants and particle representations.[20]Historically, the study of Lie algebras evolved from the geometric analysis of Lie groups to emphasize algebraic classification, driven by efforts to systematize infinite-dimensional transformation groups into finite-dimensional structures amenable to root system decompositions and Cartan subalgebras. This algebraic turn, initiated by figures like Sophus Lie and Wilhelm Killing, proved crucial for classifying simple Lie algebras over the complex numbers, providing tools for deeper insights into symmetry structures beyond the original group-theoretic motivations.[21]
Formal Definition
Abstract Definition over a Field
A Lie algebra \mathfrak{g} over a field K is a vector space over K equipped with a bilinear operation [\cdot, \cdot]: \mathfrak{g} \times \mathfrak{g} \to \mathfrak{g}, called the Lie bracket, that satisfies two axioms: antisymmetry, [\mathbf{x}, \mathbf{y}] = -[\mathbf{y}, \mathbf{x}] for all \mathbf{x}, \mathbf{y} \in \mathfrak{g}, and the Jacobi identity, [\mathbf{x}, [\mathbf{y}, \mathbf{z}]] + [\mathbf{y}, [\mathbf{z}, \mathbf{x}]] + [\mathbf{z}, [\mathbf{x}, \mathbf{y}]] = 0 for all \mathbf{x}, \mathbf{y}, \mathbf{z} \in \mathfrak{g}. Over fields of characteristic not equal to 2, antisymmetry implies [\mathbf{x}, \mathbf{x}] = 0 for all \mathbf{x} \in \mathfrak{g}.[8] Bilinearity means that for all a, b \in K and \mathbf{x}, \mathbf{y}, \mathbf{z} \in \mathfrak{g}, [\mathbf{x}, a\mathbf{y} + b\mathbf{z}] = a[\mathbf{x}, \mathbf{y}] + b[\mathbf{x}, \mathbf{z}] and [a\mathbf{x} + b\mathbf{y}, \mathbf{z}] = a[\mathbf{x}, \mathbf{z}] + b[\mathbf{y}, \mathbf{z}].[22]This definition is standard over fields of characteristic zero, such as the real numbers \mathbb{R} or complex numbers \mathbb{C}, where the axioms hold without modification.[23] In positive characteristic, complications arise; for instance, in characteristic 2, antisymmetry does not imply [\mathbf{x}, \mathbf{x}] = 0, so the axiom is often replaced by the explicit condition [\mathbf{x}, \mathbf{x}] = 0 for all \mathbf{x} \in \mathfrak{g}.[1]The trivial Lie algebra over K is the vector space with the zero bracket, [\mathbf{x}, \mathbf{y}] = 0 for all \mathbf{x}, \mathbf{y}, which satisfies the axioms vacuously.[5] Lie algebras over a field can also arise from associative algebras by defining the bracket as the commutator [\mathbf{x}, \mathbf{y}] = \mathbf{xy} - \mathbf{yx}.[5]
Dimension and Basis
Lie algebras are defined as vector spaces equipped with a bilinear Lie bracket operation, and thus inherit the linear structure of vector spaces over their base field. For a Lie algebra \mathfrak{g} over a field K, the dimension \dim \mathfrak{g} is the dimension of \mathfrak{g} as a K-vector space, which is finite or infinite depending on the algebra.[6]In the finite-dimensional case, suppose \dim \mathfrak{g} = n < \infty. A basis for \mathfrak{g} is any ordered set of n linearly independent vectors \{e_1, \dots, e_n\}. The Lie bracket on \mathfrak{g}, being bilinear, is completely determined by its values on basis elements. Specifically, for each pair i, j, there exist unique scalars c_{ij}^k \in K (for k = 1, \dots, n) such that[e_i, e_j] = \sum_{k=1}^n c_{ij}^k e_k.These scalars c_{ij}^k are called the structure constants of \mathfrak{g} with respect to the basis \{e_i\}. They satisfy antisymmetry c_{ij}^k = -c_{ji}^k (from the antisymmetry of the bracket) and must obey the Jacobi identity when expressed in this basis, ensuring the algebraic structure is preserved under linear combinations.[6] The structure constants depend on the choice of basis; under a change of basis via an invertible matrix P, they transform according to the rule c'_{ij}^k = \sum_{p,q,r} P_{ip} P_{jq} (P^{-1})_{kr} c_{pq}^r, reflecting the isomorphism class of the Lie algebra.[24]A generating set for a Lie algebra \mathfrak{g} is a subset S \subseteq \mathfrak{g} such that the smallest Lie subalgebra containing S equals \mathfrak{g}. In the finite-dimensional case, generating sets can have cardinality ranging from the minimal number of generators (which is at most \dim \mathfrak{g} and often smaller) up to \dim \mathfrak{g}. In particular, any basis is a generating set, since the Lie subalgebra it generates contains its span, which equals \mathfrak{g}.[6]While the theory is most developed for finite-dimensional Lie algebras, infinite-dimensional examples exist, such as the Witt algebra or Virasoro algebra in conformal field theory, where the dimension is the cardinality of a Hamel basis over the base field; detailed treatment of such cases, including generalized notions of bases and constants, is deferred to discussions of extensions and infinite structures.[8]
Basic Examples
Abelian Lie Algebras
An abelian Lie algebra is a Lie algebra \mathfrak{g} over a field k in which the Lie bracket vanishes identically, that is, [\mathbf{x}, \mathbf{y}] = 0 for all \mathbf{x}, \mathbf{y} \in \mathfrak{g}. This condition implies that the bracket operation is both alternating and satisfies the Jacobi identity trivially, reducing the structure to that of a vector space equipped with the zero multiplication.[25] Consequently, abelian Lie algebras are commutative under the Lie bracket, highlighting their simplicity as the most basic nontrivial examples beyond mere vector spaces.[26]Key properties of abelian Lie algebras stem from the triviality of the bracket. Every subspace of an abelian Lie algebra is itself a subalgebra, and since the bracket is zero, all such subalgebras are automatically abelian.[27] Similarly, any subspace that is an ideal—meaning [\mathfrak{g}, I] \subseteq I—is abelian, as the left-hand side is zero. Regarding derivations, a derivation of a Lie algebra is a linear map D: \mathfrak{g} \to \mathfrak{g} satisfying D([\mathbf{x}, \mathbf{y}]) = [D\mathbf{x}, \mathbf{y}] + [\mathbf{x}, D\mathbf{y}] for all \mathbf{x}, \mathbf{y} \in \mathfrak{g}; in the abelian case, this Leibniz rule holds vacuously for every linear endomorphism, so the derivation algebra \mathrm{Der}(\mathfrak{g}) coincides with the full endomorphism algebra \mathrm{End}_k(\mathfrak{g}).[28]Examples of abelian Lie algebras abound due to their direct correspondence with vector spaces. Any vector space V over k becomes an abelian Lie algebra by defining [\mathbf{x}, \mathbf{y}] = 0 for all \mathbf{x}, \mathbf{y} \in V, with the dimension of the Lie algebra matching \dim V.[29] A concrete instance is the quotient of the 3-dimensional Heisenberg Lie algebra by its 1-dimensional center, yielding a 2-dimensional abelian Lie algebra.[30] Abelian Lie algebras exist in arbitrary finite or infinite dimensions and are classified up to isomorphism solely by their dimension as vector spaces, with no further structural invariants needed.[25]Abelian Lie algebras play a foundational role as building blocks in more complex constructions, such as semidirect products with non-abelian factors.[26]
Matrix Lie Algebras
Matrix Lie algebras provide concrete realizations of abstract Lie algebras through the vector space of square matrices equipped with the commutator as the Lie bracket. The general linear Lie algebra \mathfrak{gl}(n, K) consists of all n \times n matrices over a field K of characteristic zero, with the Lie bracket defined by [A, B] = AB - BA for A, B \in \mathfrak{gl}(n, K).[26] This bracket satisfies the bilinearity, skew-symmetry, and Jacobi identity, turning \mathfrak{gl}(n, K) into a Lie algebra, as the commutator inherits these properties from the associative matrix multiplication.[26]The dimension of \mathfrak{gl}(n, K) is n^2, corresponding to the basis of matrix units E_{ij}, where E_{ij} has a 1 in the (i,j)-entry and zeros elsewhere. The Lie bracket on these basis elements is given explicitly by[E_{ij}, E_{kl}] = \delta_{jk} E_{il} - \delta_{li} E_{kj},where \delta denotes the Kronecker delta; this formula follows directly from computing the matrix products E_{ij} E_{kl} - E_{kl} E_{ij}.[26] As an associative algebra under matrix multiplication, \mathfrak{gl}(n, K) induces its Lie algebra structure via the commutator, providing a foundational example for studying more specialized Lie algebras.Important subalgebras of \mathfrak{gl}(n, K) include the special linear Lie algebra \mathfrak{sl}(n, K), comprising matrices with trace zero, which is closed under the commutator since \operatorname{tr}([A, B]) = 0.[26] Another example is the orthogonal Lie algebra \mathfrak{so}(n, K), consisting of skew-symmetric matrices (satisfying A^T = -A), which preserves the standard bilinear form and has dimension n(n-1)/2.[8] These subalgebras inherit the Lie structure from \mathfrak{gl}(n, K) and serve as models for classical Lie algebras. The set of diagonal matrices forms an abelian subalgebra of \mathfrak{gl}(n, K), where the bracket vanishes.[26]
Algebraic Operations and Structures
Subalgebras, Ideals, and Homomorphisms
A subalgebra of a Lie algebra \mathfrak{g} over a field K is a subspace \mathfrak{h} \subseteq \mathfrak{g} that is closed under the Lie bracket, meaning [\mathfrak{h}, \mathfrak{h}] \subseteq \mathfrak{h}.[31] This structure inherits the bilinear, alternating, and Jacobi properties from \mathfrak{g}, forming its own Lie algebra.[31]An ideal \mathfrak{i} of \mathfrak{g} is a subspace such that [\mathfrak{g}, \mathfrak{i}] \subseteq \mathfrak{i}.[32] Due to the antisymmetry of the Lie bracket, [x, y] = -[y, x] for all x, y \in \mathfrak{g}, left and right ideals coincide, and every ideal is two-sided.[33] For instance, in the Lie algebra \mathfrak{sl}(n, K) of trace-zero matrices, the subspace of strictly upper triangular matrices forms an ideal.[26]A Lie algebra \mathfrak{g} is simple if it is non-abelian and has no nontrivial ideals, i.e., the only ideals are \{0\} and \mathfrak{g} itself.[34] Simple Lie algebras play a foundational role in the structure theory of semisimple Lie algebras, as the latter decompose into direct sums of simples.[34]A Lie algebra homomorphism \phi: \mathfrak{g} \to \mathfrak{g}' is a linear map preserving the bracket: \phi([x, y]) = [\phi(x), \phi(y)] for all x, y \in \mathfrak{g}.[35] The kernel of such a \phi is an ideal in \mathfrak{g}, and the image is a subalgebra of \mathfrak{g}'.[36]
Lie Algebra Homomorphisms and Isomorphisms
A Lie algebra homomorphism between two Lie algebras \mathfrak{g} and \mathfrak{h} over the same field k is a linear map \phi: \mathfrak{g} \to \mathfrak{h} that preserves the Lie bracket, satisfying \phi([x, y]) = [\phi(x), \phi(y)] for all x, y \in \mathfrak{g}.[37] Such maps preserve the bilinear and skew-symmetric nature of the bracket, ensuring compatibility with the vector space structure and the Jacobi identity.[37]An isomorphism is a bijective Lie algebra homomorphism, meaning it has an inverse that is also a homomorphism, thereby preserving all structural properties including dimension and basis relations.[37] Isomorphisms identify Lie algebras up to structural equivalence, allowing classification problems to focus on invariant properties like solvability or semisimplicity.Automorphisms are isomorphisms \phi: \mathfrak{g} \to \mathfrak{g}, forming the automorphism group \mathrm{Aut}(\mathfrak{g}) under composition, which acts on \mathfrak{g} by preserving its Lie structure.[37] For finite-dimensional Lie algebras over \mathbb{R} or \mathbb{C}, \mathrm{Aut}(\mathfrak{g}) is itself a Lie group whose Lie algebra consists of derivations.Derivations are special linear endomorphisms D: \mathfrak{g} \to \mathfrak{g} satisfying the Leibniz rule D([x, y]) = [D(x), y] + [x, D(y)] for all x, y \in \mathfrak{g}, representing infinitesimal automorphisms generated by one-parameter subgroups of \mathrm{Aut}(\mathfrak{g}).[38]The kernel of a Lie algebra homomorphism \phi: \mathfrak{g} \to \mathfrak{h} is an ideal in \mathfrak{g}, and the image \phi(\mathfrak{g}) is a subalgebra of \mathfrak{h}.[37] By the first isomorphism theorem, there is a unique isomorphism \mathfrak{g}/\ker(\phi) \cong \mathrm{im}(\phi), inducing quotient Lie algebras when the kernel is an ideal.[37]
Direct and Semidirect Products
The direct product of two Lie algebras \mathfrak{l} and \mathfrak{m} over a field k, denoted \mathfrak{l} \oplus \mathfrak{m}, is the direct sum of the underlying vector spaces equipped with the Lie bracket defined componentwise:[(x_1, m_1), (x_2, m_2)] = ([x_1, x_2]_{\mathfrak{l}}, [m_1, m_2]_{\mathfrak{m}})for all x_1, x_2 \in \mathfrak{l} and m_1, m_2 \in \mathfrak{m}.[39] This structure makes \mathfrak{l} \oplus \mathfrak{m} a Lie algebra, where both \mathfrak{l} (identified with \mathfrak{l} \oplus \{0\}) and \mathfrak{m} (identified with \{0\} \oplus \mathfrak{m}) are ideals.[6] The direct product is solvable if and only if both \mathfrak{l} and \mathfrak{m} are solvable, as the derived series of the product terminates precisely when those of the factors do.The semidirect product \mathfrak{l} \rtimes_{\rho} \mathfrak{m} arises when there is a Lie algebra homomorphism \rho: \mathfrak{m} \to \mathrm{Der}(\mathfrak{l}) specifying an action of \mathfrak{m} on \mathfrak{l} by derivations. It is defined on the direct sum of vector spaces \mathfrak{l} \oplus \mathfrak{m} by the bracket[(x, m), (x', m')] = ([x, x'] + \rho(m)(x'), [m, m'])for x, x' \in \mathfrak{l} and m, m' \in \mathfrak{m}.[40] Here, \mathfrak{l} (as \mathfrak{l} \oplus \{0\}) forms an ideal, while \mathfrak{m} (as \{0\} \oplus \mathfrak{m}) is a subalgebra, and the action \rho encodes how elements of \mathfrak{m} "twist" the bracket in \mathfrak{l}. The semidirect product is solvable if both \mathfrak{l} and \mathfrak{m} are solvable, since the derived series remains contained within the solvable factors under the action. When \rho is the zero map, this reduces to the direct product.A concrete example is the Lie algebra of affine transformations in one dimension over \mathbb{R}, which is the semidirect product \mathbb{R} \rtimes \mathbb{R}. Identifying the first \mathbb{R} with translations and the second with scalings (via the derivation d/dx acting on the translation basis element), the nontrivial bracket reflects the non-commutativity of scaling and translating.[41]
Representations and Modules
Definition of Representations
A representation of a Lie algebra \mathfrak{g} over a field k is a Lie algebra homomorphism \rho: \mathfrak{g} \to \mathfrak{gl}(V), where V is a vector space over k and \mathfrak{gl}(V) denotes the Lie algebra of all endomorphisms of V equipped with the commutator bracket [A, B] = AB - BA.[33] This means that \rho is a linear map preserving the Lie bracket, so \rho([x, y]) = [\rho(x), \rho(y)] for all x, y \in \mathfrak{g}.[33]Equivalently, a representation can be defined as a bilinear map \mathfrak{g} \times V \to V, denoted (x, v) \mapsto x \cdot v or \rho(x)v, satisfying the Leibniz rule [x, y] \cdot v = x \cdot (y \cdot v) - y \cdot (x \cdot v) for all x, y \in \mathfrak{g} and v \in V.[42] This action turns V into a \mathfrak{g}-module, where the module structure is given precisely by this bilinear operation compatible with the Lie bracket on \mathfrak{g}.[43] In this view, representations of \mathfrak{g} correspond bijectively to modules over the universal enveloping algebra U(\mathfrak{g}), which is the associative algebra generated by \mathfrak{g} subject to the relations from the Lie bracket.[44]A representation \rho: \mathfrak{g} \to \mathfrak{gl}(V) is called irreducible if the only \mathfrak{g}-invariant subspaces of V are \{0\} and V itself, where a subspace W \subseteq V is \mathfrak{g}-invariant if \rho(x)w \in W for all x \in \mathfrak{g} and w \in W.[43]In representations of semisimple Lie algebras, particularly those with a chosen Cartan subalgebra \mathfrak{h}, the concepts of weight spaces and roots provide a decomposition framework. A weight is an element \lambda \in \mathfrak{h}^*, the dual space of \mathfrak{h}, and the corresponding weight space is the subspace V_\lambda = \{ v \in V \mid \rho(h)v = \lambda(h) v \ \forall h \in \mathfrak{h} \}.[45] Roots are the nonzero weights appearing in the decomposition of the adjoint action of \mathfrak{h} on \mathfrak{g}, forming the root system that underlies the structure of such representations.[45]
Adjoint Representation
The adjoint representation of a finite-dimensional Lie algebra \mathfrak{g} over a field of characteristic zero is defined by the linear map \mathrm{ad}: \mathfrak{g} \to \mathfrak{gl}(\mathfrak{g}) given by \mathrm{ad}_x(y) = [x, y] for all x, y \in \mathfrak{g}, where [\cdot, \cdot] denotes the Lie bracket.[46] This map is a Lie algebra homomorphism because it preserves the bracket: [\mathrm{ad}_x, \mathrm{ad}_y] = \mathrm{ad}_{[x,y]}, as follows from the Jacobi identity.[6] Thus, \mathrm{ad} equips \mathfrak{g} with a representation on itself by endomorphisms, making it a canonical example of a Lie algebra representation.[46]The kernel of the adjoint representation is the center Z(\mathfrak{g}) = \{ x \in \mathfrak{g} \mid [x, \mathfrak{g}] = 0 \}, consisting of elements that commute with everything in \mathfrak{g}.[6] The image \mathrm{ad}(\mathfrak{g}) lies in the derivation algebra \mathrm{Der}(\mathfrak{g}), the space of all linear maps D: \mathfrak{g} \to \mathfrak{g} satisfying D([x,y]) = [Dx, y] + [x, Dy] for all x, y \in \mathfrak{g}, since the adjoint action preserves the bracket by the Jacobi identity.[6] For a semisimple Lie algebra, the adjoint representation is faithful and surjective onto \mathrm{Der}(\mathfrak{g}), meaning every derivation is inner (of the form \mathrm{ad}_x for some x \in \mathfrak{g}).[6]Associated to the adjoint representation is the Killing form, a symmetric bilinear form B: \mathfrak{g} \times \mathfrak{g} \to k defined byB(x, y) = \operatorname{tr}(\mathrm{ad}_x \circ \mathrm{ad}_y),where \operatorname{tr} is the trace in the endomorphism algebra \mathfrak{gl}(\mathfrak{g}).[47] This form is invariant under the adjoint action, satisfying B([x, z], y) + B(z, [x, y]) = 0 for all x, y, z \in \mathfrak{g}, which follows from the cyclicity of the trace and the properties of the bracket.[47] For semisimple Lie algebras over \mathbb{C}, the Killing form is nondegenerate, providing a key tool for studying the structure.[47]
Universal Enveloping Algebra
The universal enveloping algebra U(\mathfrak{g}) of a Lie algebra \mathfrak{g} over a commutative ring k with unit is defined as the quotient of the tensor algebra T(\mathfrak{g}) = \bigoplus_{n=0}^\infty \mathfrak{g}^{\otimes n} (with \mathfrak{g}^{\otimes 0} = k) by the two-sided ideal I generated by all elements of the form x \otimes y - y \otimes x - [x, y] for x, y \in \mathfrak{g}.[48] This construction embeds \mathfrak{g} into U(\mathfrak{g}) via the canonical inclusion i: \mathfrak{g} \hookrightarrow U(\mathfrak{g}), turning U(\mathfrak{g}) into an associative unital algebra that "envelops" the Lie structure of \mathfrak{g}.[48] The map i is a Lie algebra homomorphism, and U(\mathfrak{g}) is universal among all associative algebras receiving a Lie homomorphism from \mathfrak{g}: for any Lie homomorphism \phi: \mathfrak{g} \to A into an associative algebra A, there exists a unique algebra homomorphism \tilde{\phi}: U(\mathfrak{g}) \to A extending \phi.[48]A fundamental result characterizing the structure of U(\mathfrak{g}) is the Poincaré–Birkhoff–Witt (PBW) theorem, which provides an explicit basis when k has characteristic zero. If \{e_1, \dots, e_n\} is a basis for the finite-dimensional Lie algebra \mathfrak{g}, then the images in U(\mathfrak{g}) of the monomials e_1^{k_1} e_2^{k_2} \cdots e_n^{k_n} for all nonnegative integers k_1, \dots, k_n form a basis for U(\mathfrak{g}) as a k-vector space.[49] Equivalently, \operatorname{gr} U(\mathfrak{g}) \cong S(\mathfrak{g}), the symmetric algebra on \mathfrak{g}, where \operatorname{gr} denotes the associated graded algebra with respect to the natural filtration on U(\mathfrak{g}) by degree.[49] This theorem ensures that U(\mathfrak{g}) is free as a left or right U(\mathfrak{g})-module generated by 1, and it highlights how the Lie bracket relations deform the commutative multiplication of the symmetric algebra into an associative one.[49]Representations of the Lie algebra \mathfrak{g} on a vector space V correspond bijectively to left modules over U(\mathfrak{g}), where the action of elements of \mathfrak{g} on V extends uniquely to an action of all of U(\mathfrak{g}) satisfying the universal property.[48] In particular, the adjoint representation of \mathfrak{g} on itself extends to a U(\mathfrak{g})-module structure via left multiplication on U(\mathfrak{g}).[48] For semisimple Lie algebras over \mathbb{C}, the center Z(U(\mathfrak{g})) consists of central elements that act as scalars on irreducible representations, and the Harish-Chandra isomorphism identifies Z(U(\mathfrak{g})) with the ring of invariants S(\mathfrak{h})^W, where \mathfrak{h} is a Cartan subalgebra and W is the Weyl group.
Structure Theory
Solvable and Nilpotent Lie Algebras
A Lie algebra \mathfrak{g} over a field K is defined to be solvable if its derived series terminates at the zero algebra after finitely many steps. The derived series is constructed recursively as \mathfrak{g}^{(0)} = \mathfrak{g} and \mathfrak{g}^{(k+1)} = [\mathfrak{g}^{(k)}, \mathfrak{g}^{(k)}] for k \geq 0, where [ \cdot, \cdot ] denotes the Lie bracket; thus, \mathfrak{g} is solvable if there exists some integer n such that \mathfrak{g}^{(n)} = \{0\}. This condition captures the idea of successive commutator subalgebras becoming trivial, analogous to solvable groups via the commutator subgroup series.[50]In contrast, a Lie algebra \mathfrak{g} is nilpotent if its lower central series reaches the zero subalgebra in finitely many steps. The lower central series is defined by \mathfrak{g}_0 = \mathfrak{g} and \mathfrak{g}_{k+1} = [\mathfrak{g}, \mathfrak{g}_k] for k \geq 0; hence, \mathfrak{g} is nilpotent if \mathfrak{g}_m = \{0\} for some positive integer m.[51] Every nilpotent Lie algebra is solvable, since the lower central series refines the derived series in the sense that each term in the former lies within the corresponding derived term, ensuring the derived series also terminates.[52] Over an algebraically closed field of characteristic zero, a finite-dimensional Lie algebra \mathfrak{g} is nilpotent if and only if the adjoint map \mathrm{ad}_x: \mathfrak{g} \to \mathfrak{g} is a nilpotent endomorphism for every x \in \mathfrak{g}, as stated by Engel's theorem.[50] This characterization links nilpotency directly to the nilpotency of individual adjoint operators, providing a useful criterion for verification.[53]Abelian Lie algebras, where the Lie bracket vanishes identically [\mathfrak{g}, \mathfrak{g}] = \{0\}, form the simplest examples of both solvable and nilpotent algebras, as their derived and lower central series terminate immediately at the first step.[52] A non-abelian example is the three-dimensional Heisenberg Lie algebra over a field K, with basis \{x, y, z\} and nontrivial bracket [x, y] = z while all other brackets among basis elements are zero; its lower central series is \mathfrak{g}_0 = \mathfrak{g}, \mathfrak{g}_1 = [ \mathfrak{g}, \mathfrak{g} ] = Kz, and \mathfrak{g}_2 = [ \mathfrak{g}, \mathfrak{g}_1 ] = \{0\}, confirming nilpotency (and thus solvability), whereas the derived series is \mathfrak{g}^{(0)} = \mathfrak{g}, \mathfrak{g}^{(1)} = Kz, \mathfrak{g}^{(2)} = \{0\}.[54] This algebra exemplifies how nilpotency can arise from a single central commutator extension of an abelian algebra.[50]
Semisimple Lie Algebras and Levi Decomposition
A Lie algebra \mathfrak{g} over a field of characteristic zero is said to have a radical, denoted \mathrm{rad}(\mathfrak{g}), which is the unique maximal solvable ideal of \mathfrak{g}. This radical coincides with the largest solvable ideal and, in the finite-dimensional case, contains the nilradical as its maximal nilpotent ideal. A Lie algebra is semisimple if and only if its radical vanishes, i.e., \mathrm{rad}(\mathfrak{g}) = \{0\}.[55][56]The Levi decomposition theorem provides a fundamental structural result for finite-dimensional Lie algebras over fields of characteristic zero. Specifically, any such Lie algebra \mathfrak{g} can be expressed as a semidirect product \mathfrak{g} = \mathfrak{s} \ltimes \mathrm{rad}(\mathfrak{g}), where \mathfrak{s} is a semisimple Lie subalgebra that complements the radical, meaning \mathfrak{g} = \mathfrak{s} + \mathrm{rad}(\mathfrak{g}) and \mathfrak{s} \cap \mathrm{rad}(\mathfrak{g}) = \{0\}. This decomposition highlights how the "non-solvable" part of \mathfrak{g} acts on its solvable radical via the Lie bracket.[56][57]Semisimple Lie algebras possess a particularly simple internal structure: they decompose as a direct sum of simple Lie ideals. That is, if \mathfrak{g} is semisimple, then \mathfrak{g} \cong \bigoplus_{i=1}^k \mathfrak{g}_i, where each \mathfrak{g}_i is a simple Lie algebra (possessing no nontrivial ideals) and the sum is direct with respect to the Lie bracket. This orthogonality of the simple factors underscores the absence of solvable or nilpotent components within semisimple algebras.[58]The Killing form offers an invariant bilinear form that distinguishes semisimple Lie algebras. Defined as B(X, Y) = \mathrm{Tr}(\mathrm{ad}_X \circ \mathrm{ad}_Y) for X, Y \in \mathfrak{g}, where \mathrm{ad} denotes the adjoint representation, the Killing form is symmetric and invariant under the adjoint action. On a semisimple Lie algebra, the Killing form is non-degenerate, meaning its radical (the set of elements pairing to zero with all others) is trivial; conversely, non-degeneracy of the Killing form implies semisimplicity.[58][56]
Cartan's Criteria for Solvability and Semisimplicity
Lie's theorem characterizes the representation theory of solvable Lie algebras over algebraically closed fields of characteristic zero. It states that if L is a solvable Lie algebra, then for any finite-dimensional representation \rho: L \to \mathfrak{gl}(V) on a vector space V, there exists a flag of subspaces $0 = V_0 \subset V_1 \subset \cdots \subset V_n = V such that each V_i is invariant under the action of \rho(L).[59] In particular, applied to the adjoint representation, a solvable Lie algebra admits a flag of ideals invariant under the adjoint action.[59]Cartan's criterion for solvability provides an algebraic test using the Killing form, defined as B(x, y) = \operatorname{tr}(\operatorname{ad}_x \operatorname{ad}_y), where \operatorname{ad} denotes the adjoint representation. A Lie algebra L over a field of characteristic zero is solvable if and only if B(x, y) = 0 for all x \in [L, L] and y \in L.[60] This condition implies that the Killing form degenerates on the derived algebra [L, L].[60]For semisimplicity, Cartan's criterion states that a Lie algebra L over a field of characteristic zero is semisimple if and only if its Killing form B is non-degenerate.[60] This non-degeneracy ensures that L has no non-trivial solvable ideals.[60]Weyl's theorem offers an equivalent characterization: a Lie algebra L over a field of characteristic zero is semisimple if and only if it contains no non-zero abelian ideals.[61] This highlights the absence of proper abelian subalgebras that are ideals as a defining feature of semisimplicity.[61]
Classification
Classification over Algebraically Closed Fields
Over algebraically closed fields of characteristic zero, such as the complex numbers \mathbb{C}, the classification of finite-dimensional semisimple Lie algebras reduces to the classification of their simple ideals, as every semisimple Lie algebra is a direct sum of simple ones.[61] This classification, originally established by Wilhelm Killing and refined by Élie Cartan, identifies all such simple Lie algebras up to isomorphism through their root systems and associated Dynkin diagrams.[62] The process hinges on the structure theory, where a semisimple Lie algebra \mathfrak{g} admits a Cartan subalgebra \mathfrak{h} such that \mathfrak{g} = \mathfrak{h} \oplus \bigoplus_{\alpha \in \Phi} \mathfrak{g}_\alpha, with \Phi the root system in the dual space \mathfrak{h}^*.[63]The root system \Phi consists of nonzero linear functionals \alpha \in \mathfrak{h}^* (roots) for which the root spaces \mathfrak{g}_\alpha are one-dimensional. A choice of positive roots \Phi^+ determines simple roots \Delta \subset \Phi^+, a basis for \Phi over \mathbb{Z} such that every root is an integer linear combination of elements in \Delta. The Weyl group W is the finite group generated by reflections s_\alpha: \beta \mapsto \beta - 2 \frac{(\beta, \alpha)}{(\alpha, \alpha)} \alpha across root hyperplanes, acting on \mathfrak{h}^* and preserving \Phi. These simple roots encode the structure via the Cartan matrix A = (a_{ij}) with a_{ij} = 2 \frac{(\alpha_i, \alpha_j)}{(\alpha_j, \alpha_j)}, whose realization as a Dynkin diagram—a graph with nodes for simple roots and edges indicating non-perpendicularity and length ratios—classifies the possible root systems.[64]The irreducible root systems, and thus the simple Lie algebras, fall into four infinite classical families and five exceptional cases, distinguished by their Dynkin diagrams:
Type A_n (n \geq 1): A chain of n nodes, corresponding to \mathfrak{sl}(n+1, \mathbb{C}).
Type B_n (n \geq 2): A chain of n nodes with a double bond (length ratio 2:1) at the end, corresponding to \mathfrak{so}(2n+1, \mathbb{C}).
Type C_n (n \geq 3): A chain of n nodes with a double bond (length ratio 1:2) at the end, corresponding to \mathfrak{sp}(2n, \mathbb{C}).
Type D_n (n \geq 4): A chain of n-2 nodes branching into two at the end, corresponding to \mathfrak{so}(2n, \mathbb{C}).
The exceptional types are E_6, E_7, E_8 (extended chains with a branch), F_4 (chain with double and triple bonds), and G_2 (two nodes with triple bond). These diagrams are the only connected, simply-laced or folded graphs satisfying the classification criteria (no loops, nodes of degree at most 3, etc.), ensuring all finite-dimensional simple complex Lie algebras are isomorphic to one of these.[63]For each such Lie algebra, a Chevalley basis provides an integral structure: it consists of basis elements e_\alpha, f_\alpha \in \mathfrak{g}_\alpha, \mathfrak{g}_{-\alpha} for \alpha \in \Phi^+, and h_i for simple coroots, with all structure constants integers under the bracket relations [e_\alpha, f_\alpha] = h_\alpha, [h_i, e_\alpha] = a_{i\alpha} e_\alpha, etc. This basis, constructed by Claude Chevalley, allows realization over \mathbb{Z} and extension to any field, preserving the classification.[65][66]
Real Forms and Complexification
A real Lie algebra \mathfrak{g} over \mathbb{R} can be extended to a complex Lie algebra via its complexification \mathfrak{g}_\mathbb{C} = \mathfrak{g} \otimes_{\mathbb{R}} \mathbb{C}, where the Lie bracket on \mathfrak{g}_\mathbb{C} is defined by extending the bracket on \mathfrak{g} using complex bilinearity. This construction identifies \mathfrak{g}_\mathbb{C} with the direct sum \mathfrak{g} \oplus i\mathfrak{g} as real vector spaces, with the bracket [X + iY, X' + iY'] = [X,X'] - [Y,Y'] + i([X,Y'] + [Y,X']) for X,Y,X',Y' \in \mathfrak{g}. The complexification preserves the structure of \mathfrak{g} while allowing the application of tools from complex semisimple Lie algebra theory, such as root decompositions relative to a Cartan subalgebra.[67]For semisimple complex Lie algebras, the complexification process relates real forms to their complex counterparts. Every semisimple complex Lie algebra \mathfrak{g} admits at least one real form, meaning a real subalgebra \mathfrak{h} \subseteq \mathfrak{g} (viewed as a real vector space) such that the complexification \mathfrak{h}_\mathbb{C} \cong \mathfrak{g}. In particular, every semisimple complex Lie algebra is the complexification of a unique (up to isomorphism) compact real form, which is the Lie algebra of a compact semisimple Lie group. A classic example is \mathfrak{su}(2), the Lie algebra of $3 \times 3 skew-Hermitian matrices with trace zero, which is a compact real form of \mathfrak{sl}(2,\mathbb{C}).[68][69][70]The classification of real simple Lie algebras, completed by Élie Cartan in 1914, identifies all real forms of the simple complex Lie algebras classified by Killing and Cartan over \mathbb{C}. For a given simple complex Lie algebra \mathfrak{g}, the real simple forms are finite in number and determined by involutive automorphisms of \mathfrak{g} (Cartan involutions). For instance, \mathfrak{sl}(3,\mathbb{C}) has exactly three inequivalent real simple forms: the split form \mathfrak{sl}(3,\mathbb{R}), the compact form \mathfrak{su}(3), and the non-compact non-split form \mathfrak{su}(2,1). This classification extends the complex case by accounting for the possible real structures compatible with the root system of \mathfrak{g}.[68][69][71]Real forms are further distinguished by the signature of the Killing form B(X,Y) = \operatorname{tr}(\operatorname{ad}_X \operatorname{ad}_Y). For the compact real form of a semisimple complex Lie algebra, the Killing form is negative definite, reflecting the compactness of the associated Lie group. In contrast, the split real form, which maximizes the dimension of the abelian part in a Cartan decomposition, has an indefinite Killing form whose signature is ( \dim \mathfrak{g} - r, r ), where r is the rank of \mathfrak{g}; equivalently, the number of positive eigenvalues equals \dim \mathfrak{g} - r. For example, in \mathfrak{sl}(2,\mathbb{R}) (the split form of \mathfrak{sl}(2,\mathbb{C})), the Killing form has signature (2,1). These signatures provide an invariant to differentiate forms within the classification.[69][70][72]
Relation to Lie Groups
Lie Algebra of a Matrix Lie Group
A matrix Lie group G is a Lie group that can be realized as a closed subgroup of the general linear group \mathrm{GL}(n, \mathbb{R}) for some n, consisting of invertible n \times n real matrices. The Lie algebra \mathfrak{g} of such a group G is defined as the vector space of all n \times n real matrices X such that \exp(tX) \in G for all sufficiently small real numbers t > 0.[73] This construction identifies \mathfrak{g} concretely with elements of the matrix algebra M_n(\mathbb{R}) that are tangent to G at the identity matrix I_n.[74]The Lie algebra \mathfrak{g} is isomorphic to the tangent space T_{I_n}G at the identity, which serves as the space of left-invariant vector fields on G. These left-invariant vector fields are generated by curves through the identity in G, and the Lie bracket on \mathfrak{g} is induced by the commutator of vector fields, providing a bilinear operation [X, Y] = XY - YX for X, Y \in \mathfrak{g}.[75] For matrix Lie groups, this bracket corresponds directly to the matrix commutator, ensuring that \mathfrak{g} inherits a Lie algebra structure from the group's smooth manifold topology.[76]A representative example is the special unitary group \mathrm{SU}(2), the Lie group of $2 \times 2 complex unitary matrices with determinant 1. Its Lie algebra \mathfrak{su}(2) consists of all $2 \times 2 complex matrices X satisfying X^\dagger = -X (skew-Hermitian) and \operatorname{Tr}(X) = 0, with the standard basis given by the Pauli matrices (up to factors of i).[77] Another example is the special linear group \mathrm{SL}(2, \mathbb{R}), the Lie group of $2 \times 2 real matrices with determinant 1; its Lie algebra \mathfrak{sl}(2, \mathbb{R}) comprises all $2 \times 2 real matrices with trace zero, equipped with the commutator bracket.[78] These examples illustrate how the Lie algebra captures the infinitesimal symmetries of the group near the identity.
Exponential Map and Logarithm
The exponential map provides a fundamental connection between a Lie algebra \mathfrak{g} and its associated Lie group G, mapping elements of the algebra to group elements via a smooth homomorphism. For matrix Lie groups, where \mathfrak{g} consists of matrices, the exponential map is defined explicitly as\exp(X) = \sum_{k=0}^{\infty} \frac{X^k}{k!},which converges for all X \in \mathfrak{g} since the series is the standard matrix exponential. This map generalizes to arbitrary Lie groups by defining \exp(X) as the time-1 flow of the left-invariant vector field generated by X \in \mathfrak{g}, ensuring it is well-defined and smooth. The exponential map is a Lie group homomorphism from the additive group of \mathfrak{g} (viewed as \mathbb{R}^n) to G, and it plays a crucial role in "delinearizing" the tangent space at the identity.[79][80]A key property of the exponential map is that it is a local diffeomorphism near the origin of \mathfrak{g}. Specifically, the differential d\exp_0: \mathfrak{g} \to T_e G \cong \mathfrak{g} is the identity map, implying that \exp restricts to a diffeomorphism from a neighborhood of $0 \in \mathfrak{g} onto a neighborhood of the identity e \in G. This local invertibility allows the construction of one-parameter subgroups: for each X \in \mathfrak{g}, the curve t \mapsto \exp(tX) is a smooth homomorphism from \mathbb{R} to G with derivative X at t=0. Global behavior varies by group type; for simply connected nilpotent Lie groups, \exp is a global diffeomorphism, providing a bijection between \mathfrak{g} and G. In contrast, for compact connected Lie groups, \exp is surjective but generally not injective, as the group is covered by the image with possible discrete kernel related to the fundamental group.[16][79][80]The logarithm serves as the local inverse to the exponential map near the identity. For elements g \in G sufficiently close to e, there exists a unique X \in \mathfrak{g} such that \exp(X) = g, and this X = \log(g) is smooth in a neighborhood of e. The logarithm is multi-valued globally in general, but its principal branch is well-defined locally, facilitating computations like solving differential equations on the group. This inverse property underscores the exponential map's role in parameterizing small group elements via algebra elements.[79][80]The exponential map intertwines the adjoint representations of the group and algebra. The adjoint action of G on \mathfrak{g} is given by \mathrm{Ad}_g(X) = g X g^{-1} for g \in G and X \in \mathfrak{g} (in the matrix case), while the algebra's adjoint is \mathrm{ad}_X(Y) = [X, Y]. A fundamental relation is \mathrm{Ad}_{\exp(X)} = \exp(\mathrm{ad}_X), where the right-hand side is the exponential of the linear operator \mathrm{ad}_X: \mathfrak{g} \to \mathfrak{g}. This identity links conjugation in the group to nested commutators in the algebra, with applications in computing conjugacy classes and stability analysis.[16][79]
Extensions and Generalizations
Derivations and Automorphisms
A derivation of a Lie algebra \mathfrak{L} over a field F of characteristic zero is a linear endomorphism \delta: \mathfrak{L} \to \mathfrak{L} satisfying the Leibniz rule \delta([x, y]) = [\delta(x), y] + [x, \delta(y)] for all x, y \in \mathfrak{L}. This condition ensures that derivations capture infinitesimal symmetries preserving the Lie bracket structure. The space \operatorname{Der}(\mathfrak{L}) of all such derivations forms a Lie algebra under the commutator bracket [\delta_1, \delta_2] = \delta_1 \circ \delta_2 - \delta_2 \circ \delta_1, which itself satisfies the derivation property.The adjoint representation provides a canonical family of derivations: for each x \in \mathfrak{L}, the map \operatorname{ad}_x: y \mapsto [x, y] is a derivation, known as an inner derivation. The image \operatorname{ad}(\mathfrak{L}) under the adjoint map \operatorname{ad}: \mathfrak{L} \to \operatorname{Der}(\mathfrak{L}) is a Lie subalgebra of \operatorname{Der}(\mathfrak{L}), and the kernel of \operatorname{ad} is the center of \mathfrak{L}. Outer derivations are the cosets in the quotient Lie algebra \operatorname{Der}(\mathfrak{L}) / \operatorname{ad}(\mathfrak{L}), measuring derivations not arising from the algebra's own elements. For semisimple Lie algebras, all derivations are inner, so \operatorname{Der}(\mathfrak{L}) = \operatorname{ad}(\mathfrak{L}) and there are no nontrivial outer derivations.An automorphism of \mathfrak{L} is an invertible linear map \sigma: \mathfrak{L} \to \mathfrak{L} preserving the bracket, i.e., \sigma([x, y]) = [\sigma(x), \sigma(y)] for all x, y \in \mathfrak{L}. The set \operatorname{Aut}(\mathfrak{L}) of all automorphisms forms a group under composition, acting as the symmetry group of the Lie algebra. Inner automorphisms are those induced by the adjoint action, \operatorname{Int}(\mathfrak{L}) = \{ e^{\operatorname{ad}_x} \mid x \in \mathfrak{L} \}, a normal subgroup of \operatorname{Aut}(\mathfrak{L}), with the outer automorphism group given by the quotient \operatorname{Out}(\mathfrak{L}) = \operatorname{Aut}(\mathfrak{L}) / \operatorname{Int}(\mathfrak{L}). In the context of Lie groups, one often considers the Lie algebra of continuous automorphisms, which for a connected Lie group G with Lie algebra \mathfrak{L} is \operatorname{Der}(\mathfrak{L}). Derivations and automorphisms play a role in constructing semidirect products of Lie algebras.
Infinite-Dimensional Lie Algebras
Infinite-dimensional Lie algebras generalize finite-dimensional ones by allowing vector spaces of countable or uncountable dimension, leading to richer structures but also significant analytical challenges, such as the need for topological completions to ensure well-defined operations.[81] Unlike their finite-dimensional counterparts, which admit complete classifications over algebraically closed fields of characteristic zero, infinite-dimensional Lie algebras often lack such global classifications and require additional structure like gradings or central extensions to study their representations and symmetries.A prominent example is the Witt algebra, which consists of vector fields on the circle S^1 with Laurent polynomial coefficients, realized over \mathbb{C} with basis elements L_n = -z^{n+1} \frac{d}{dz} for n \in \mathbb{Z} satisfying the commutation relations [L_m, L_n] = (m - n) L_{m+n}.[82] This algebra arises as the Lie algebra of infinitesimal conformal transformations in two dimensions and serves as a foundational structure for more complex extensions. Its unique nontrivial central extension, the Virasoro algebra, introduces a central element c with relations [L_m, L_n] = (m - n) L_{m+n} + \frac{c}{12} (m^3 - m) \delta_{m, -n}, capturing quantum anomalies in physical systems.[82] Another key class is the family of Kac-Moody algebras, defined via generalized Cartan matrices and generated by elements satisfying Serre relations, extending finite-dimensional semisimple Lie algebras to infinite dimensions through affine or hyperbolic types.To equip infinite-dimensional Lie algebras with a suitable topology, they are often modeled on complete normed spaces, such as Banach spaces over \mathbb{R} or \mathbb{C}, ensuring the Lie bracket remains continuous and bilinear.[81] This topological framework is essential for defining exponentials and analyzing representations, as Fréchet or nuclear topologies may also be employed for smoother manifolds underlying associated Lie groups. Representations of these algebras, particularly affine Kac-Moody types, involve infinite weights due to the extended root systems, with Verma modules serving as universal highest-weight modules induced from characters on the Cartan subalgebra.[83] These modules are cyclic and infinite-dimensional, decomposing into weight spaces with multiplicities determined by Weyl group orbits, providing building blocks for irreducible representations.In applications, the Virasoro algebra underpins the symmetry algebra of two-dimensional conformal field theories, where the central charge c parametrizes the theory's anomaly and unitary minimal models correspond to specific rational values of c.[84] Kac-Moody algebras appear in these theories as current algebras, generating integrable representations at positive integer levels that realize affine symmetries, while in string theory, they describe the gauge symmetries of the heterotic string worldsheet, with the Virasoro algebra ensuring conformal invariance and anomaly cancellation.[85]
Lie Rings and Category-Theoretic Views
A Lie ring is an abelian group L (equivalently, a \mathbb{Z}-module) equipped with a bilinear operation [ \cdot, \cdot ]: L \times L \to L, called the Lie bracket, that satisfies antisymmetry [x, x] = 0 for all x \in L and the Jacobi identity [x, [y, z]] + [y, [z, x]] + [z, [x, y]] = 0 for all x, y, z \in L.[86] This structure generalizes Lie algebras by replacing the underlying field with the ring of integers, allowing the theory to extend to settings without division, such as integral models of Lie structures. Sub-Lie rings and ideals are defined analogously to those in Lie algebras, preserving the bracket operation, and concepts like nilpotency and solvability carry over via the additive group structure.[87]Prominent examples of Lie rings include integral forms of Lie algebras, which are \mathbb{Z}-submodules of a Lie algebra over \mathbb{Q} or \mathbb{R} that span the full space upon tensoring with the field and respect the bracket. For instance, the special linear Lie ring \mathfrak{sl}_n(\mathbb{Z}) consists of n \times n integer matrices with trace zero, equipped with the commutator bracket [A, B] = AB - BA, serving as an integral model for the Lie algebra \mathfrak{sl}_n(\mathbb{Q}).[88] Another class comprises current Lie rings, constructed as tensor products of a finite-dimensional Lie algebra with the additive group of a commutative ring, such as smooth functions on a manifold valued in the Lie algebra, where the bracket is pointwise; these arise in the study of symmetries in field theories and extend finite-dimensional structures to infinite settings while maintaining the Lie ring axioms over \mathbb{Z}.[89]The category \mathbf{LieAlg}_k of Lie algebras over a field k has objects given by Lie algebras (vector spaces over k with compatible brackets) and morphisms by linear maps preserving the bracket, i.e., Lie algebra homomorphisms \phi: L \to L' satisfying \phi([x, y]) = [\phi(x), \phi(y)]. This category admits various functors relating it to the category of Lie groups \mathbf{LieGrp}, such as the infinitesimalization functor that assigns to a Lie group its Lie algebra via left-invariant vector fields, and the integration functor attempting to recover Lie groups from Lie algebras, though the latter is not always an equivalence due to topological obstructions. These functors facilitate the study of Lie algebras as tangent spaces to Lie groups, enabling algebraic tools to analyze continuous symmetries.[48]A key categorical feature is the adjunction involving the forgetful functor U: \mathbf{LieAlg}_k \to \mathbf{Vect}_k, which forgets the bracket structure and views a Lie algebra as its underlying vector space. This forgetful functor has a left adjoint, the free Lie algebra functor \mathrm{FreeLie}: \mathbf{Vect}_k \to \mathbf{LieAlg}_k, which constructs the free Lie algebra on a vector space V as the quotient of the free associative algebra by the ideal generated by relations enforcing antisymmetry and the Jacobi identity, with the universal property ensuring any linear map from V to a Lie algebra extends uniquely to a Lie homomorphism. This adjunction underscores the algebraic generation of Lie structures and parallels similar constructions in other algebraic categories, providing a framework for universal properties in Lie theory.[90]