Fact-checked by Grok 2 weeks ago

Linear algebraic group

A linear algebraic group over a field k is a smooth affine algebraic variety over k that carries a group structure compatible with its algebraic structure, typically realized as a closed subgroup of the general linear group \mathrm{GL}_n(k) for some positive integer n, where the group operation is matrix multiplication and the defining equations are polynomial. These groups generalize classical Lie groups to an algebraic geometry setting, allowing study over arbitrary fields rather than just the reals or complexes, and they form the foundation for much of modern representation theory and arithmetic geometry. The theory of linear algebraic groups emerged in the mid-20th century, with foundational contributions from Claude Chevalley in the 1950s, who developed the structure theory over algebraically closed fields of characteristic zero, and Armand Borel, who extended it to positive characteristic. Key texts, such as James E. Humphreys' Linear Algebraic Groups (1975) and T.A. Springer's Linear Algebraic Groups (1968, revised 1998), formalized the subject by integrating algebraic geometry and Lie theory. Prominent examples include the general linear group \mathrm{GL}_n, consisting of all invertible n \times n matrices; the special linear group \mathrm{SL}_n, defined by determinant one; the orthogonal group \mathrm{O}_n; and the symplectic group \mathrm{Sp}_{2n}, all of which arise as subgroups preserving specific bilinear forms. Simpler cases are the additive group \mathbb{G}_a = (k, +) and the multiplicative group \mathbb{G}_m = k^\times, which embed into \mathrm{GL}_1. Structurally, every linear algebraic group G has a Lie algebra \mathfrak{g}, obtained as the tangent space at the identity equipped with a Lie bracket from the group law, and over algebraically closed fields of characteristic zero, G is closely tied to its Lie algebra via the exponential map. Groups are classified as solvable, semisimple, or reductive based on their unipotent radical and derived subgroup: reductive groups, such as \mathrm{GL}_n and \mathrm{SL}_n, have no nontrivial unipotent normal subgroups and have a structure theorem expressing them in terms of a maximal torus and unipotent root groups isomorphic to \mathbb{G}_a. Semisimple groups, like the simple groups \mathrm{SL}_n (for n \geq 2) or exceptional types such as E_8, are direct products of such simple components, classified by Dynkin diagrams into four infinite families (A, B, C, D) and five exceptional ones (E_6, E_7, E_8, F_4, G_2). Borel subgroups, maximal connected solvable subgroups, play a central role, being conjugate and parametrizing flag varieties like G/B. Linear algebraic groups underpin diverse applications, from the study of finite groups of Lie type (e.g., Chevalley groups) in theory to in , where they model descent data and provide tools for solving Diophantine equations. In , their finite-dimensional representations decompose into irreducibles labeled by dominant weights, facilitating connections to quantum groups and modular forms. Over non-algebraically closed fields, phenomena like anisotropic tori and Galois actions introduce additional complexity, as explored in works on the Kneser-Tits problem.

Introduction and History

Overview and motivation

A linear algebraic group over a field k is defined as a closed subgroup of the general linear group \mathrm{GL}_n(k) for some positive integer n, where the subgroup is closed in the Zariski topology and carries a compatible group structure via matrix multiplication. This setup endows the group with the structure of an affine algebraic variety, allowing the application of algebraic geometry tools to study its properties. The on affine space \mathbb{A}^n_k, the spectrum of the k[x_1, \dots, x_n], is generated by the closed sets consisting of zeros of ideals of , making it coarser than classical topologies and suited to algebraic rather than analytic phenomena. Regular functions on such varieties are precisely the functions that restrict to the variety, forming the coordinate ring that encodes the . This framework provides a purely algebraic setting for groups, contrasting with the differential structure of Lie groups over \mathbb{R} or \mathbb{C}. The study of linear algebraic groups is motivated by their role as algebraic counterparts to groups, facilitating the analysis of s and symmetries in both continuous and discrete settings. In particular, they arise naturally in the context of linear of abstract groups, where the image of a representation into \mathrm{GL}_n(k) often forms such a , bridging with . This connection is essential for understanding finite groups of type through reduction modulo primes. The systematic development of linear algebraic groups began in the with Claude Chevalley's work, where they were first employed to construct Chevalley groups over arbitrary fields, providing a uniform algebraic framework for classical simple groups.

Historical development

The theory of linear algebraic groups traces its roots to the late , when the of simple algebras laid the groundwork for understanding the associated groups. initiated this effort in his seminal papers published between 1888 and 1890 in Mathematische Annalen, where he classified the finite-dimensional simple algebras over the complex numbers, identifying four infinite families (corresponding to linear, orthogonal, symplectic, and exceptional types) and introducing the Cartan-Killing form as a key invariant. refined and rigorously proved this classification in his 1894 doctoral thesis, Sur la structure des groupes de transformations finis et continus, establishing the complete list of simple algebras over \mathbb{C} and extending the results to real forms by 1914, which provided the infinitesimal structure essential for later developments in continuous transformation groups. In the early 20th century, the focus shifted to continuous groups and their representations, with contributions from Ludwig Maurer, , and . Maurer advanced the study of continuous groups around 1900–1910, developing parameterizations and generators that bridged and aspects of Lie's original . Cartan extended his work on Lie algebras to infinite-dimensional continuous groups and their applications in geometry during the 1910s–1920s. Weyl, in a series of papers from 1925–1926, provided the first complete of representations of compact continuous groups, proving complete reducibility and deriving character formulas, which were pivotal for and ; his 1939 book The Classical Groups further generalized these ideas to non-compact cases. The mid-20th century marked the transition to algebraic groups proper, beginning with Ellis Kolchin's 1948 paper "Algebraic Matric Groups and the Picard-Vessiot Theory of Homogeneous Linear Ordinary Equations," which introduced foundational for solvable algebraic matrix groups and proved the Lie-Kolchin characterizing connected solvable subgroups of (n). advanced the theory in the , particularly through his 1955 work constructing simple algebraic groups over finite fields (now known as Chevalley groups), which linked Lie algebras to finite groups of Lie type, and his 1956–1958 seminar notes on semisimple groups. , collaborating with Chevalley, systematically developed the structure theory in the , including his 1956 paper on linear representations and Borel subgroups, culminating in his influential 1969 book Linear Algebraic Groups (revised 1991), which established key results on reductive groups and their . In the 1960s, integrated algebraic groups into his scheme-theoretic , as detailed in Séminaire de Géométrie Algébrique du Bois-Marie (SGA 3) (1963–1964), where group schemes were defined over arbitrary schemes, enabling the study of algebraic groups over rings and resolving issues with nilpotents and positive characteristic. This reformulation facilitated deeper connections to arithmetic geometry. Post-1960 developments included Jacques Tits and Borel's 1965 work on , which provided a combinatorial for the of reductive groups over local fields via BN-pairs. Recent work continues to emphasize ties between linear algebraic groups and the .

Definitions and Basic Concepts

Formal definition over algebraically closed fields

An affine over an k is defined as the zero set V(I) of an I in the k[x_1, \dots, x_n], where I is taken to be (the vanishing ideal I(V(I))), as justified by the Hilbert Nullstellensatz, which establishes a between and affine algebraic sets. The points of the variety correspond to maximal in the coordinate ring k[V] = k[x_1, \dots, x_n]/I(V). The on such varieties is generated by the closed sets, which are themselves zero loci of polynomials, providing a coarse suited to . A linear algebraic group G over k is a smooth affine algebraic variety that is a closed subgroup of \mathrm{GL}(n, k) for some n, where closedness is with respect to the Zariski topology on the ambient space \mathbb{A}^{n^2}. Equivalently, G is an affine variety equipped with morphisms of varieties m: G \times G \to G for multiplication and i: G \to G for inversion, satisfying the group axioms, with the identity morphism from the singleton variety to G. The coordinate ring k[G] of G is the finitely generated, reduced k-algebra k[x_1, \dots, x_n]/I(G), where I(G) is the vanishing ideal of G, and the group structure induces a comultiplication \Delta: k[G] \to k[G] \otimes_k k[G] making k[G] into a Hopf algebra, with the dual maps corresponding to the group operations. The multiplication m and inversion i are morphisms of varieties because they are given by polynomial functions on the ambient affine space, hence regular maps that restrict to G \times G and G, respectively, preserving the algebraic structure; this follows from the fact that the defining equations of G are preserved under these operations due to the subgroup property. For instance, if G is defined by polynomial equations f_j(g) = 0 for g \in \mathrm{GL}(n,k), then f_j(gh) = 0 and f_j(g^{-1}) = 0 whenever f_j(g) = f_j(h) = 0, ensuring the images lie in G. The dimension of G, as an , is the of k[G], equivalently the transcendence degree of the function field k(G) over k, or the dimension of the at the . This dimension is finite since k[G] is finitely generated, and for closed embeddings into \mathrm{GL}(n,k), it satisfies \dim G \leq n^2, with equality for G = \mathrm{GL}(n,k).

Extension to arbitrary fields and schemes

The notion of a linear algebraic group over an arbitrary k extends the classical definition by considering the group of rational points G(k), which consists of the k-points of the G satisfying the group law defined by the scheme morphisms. These points form an abstract group, but to capture the full algebraic structure, one uses to relate G over k to its base change G_{k_s} over a separable closure k_s, where k_s / k is Galois. A k-form of a group H defined over k_s is a group G over k such that G_{k_s} \cong H, and such forms are classified by the set H^1(k, \Aut(H)), where the action is via the \Gal(k_s / k). This ensures that properties like connectedness and are preserved under base change. In the language of scheme theory, a linear algebraic group over k is an affine group scheme of finite type over \Spec(k), meaning it is a representable functor G: (\text{Comm } k\text{-Alg})^{\op} \to \text{Groups} from the opposite category of commutative k-algebras to groups, represented by an affine scheme \Spec(A) with A a finitely generated k-algebra. The group structure arises from compatible morphisms for multiplication, inversion, and the identity, satisfying the group axioms functorially. Over arbitrary k, including non-algebraically closed cases, this framework allows uniform treatment without assuming separability. For non-perfect fields (those where the Frobenius is not surjective), linear algebraic groups are defined to be and geometrically reduced, ensuring that the relative at every point has dimension equal to the relative dimension of the and that base change to the yields a reduced scheme with no elements. Smoothness over such fields implies the existence of a maximal k- in connected reductive groups, as guaranteed by Grothendieck's , and avoids pathologies like non-separated quotients. Reducedness excludes infinitesimal group , focusing on those of classical interest. The coordinate ring O(G) = k[G] of an affine linear algebraic group G over k is a commutative Hopf k-algebra, equipped with a comultiplication \Delta: O(G) \to O(G) \otimes_k O(G) induced by the multiplication morphism m: G \times_k G \to G, a counit \varepsilon: O(G) \to k from the identity section, and an antipode S: O(G) \to O(G) from the inversion morphism. These structures encode the group law algebraically, with coassociativity and compatibility axioms holding functorially; for example, in \GL_n, \Delta(T_{ij}) = \sum_\ell T_{i\ell} \otimes T_{\ell j}. This Hopf algebra perspective dualizes the group scheme and facilitates representation theory over arbitrary k. A prominent example of a non-split form arises with anisotropic tori over the real numbers \mathbb{R}, such as the norm torus defined by the equation z \overline{z} = 1 in \mathbb{C}^\times, which is a form of \mathbb{G}_m over \mathbb{R} but contains no non-trivial split subtorus and is compact as a real Lie group. Such tori illustrate Galois descent in action, as their splitting fields are quadratic extensions of \mathbb{R}, and they appear in forms of orthogonal or unitary groups.

Group operations and homomorphisms

Linear algebraic groups are affine algebraic varieties equipped with group structure, where the multiplication m: G \times G \to G, defined by m(g, h) = gh, and the inversion i: G \to G, defined by i(g) = g^{-1}, are both morphisms of algebraic varieties. These operations ensure that the group law is compatible with the algebraic structure, allowing G to function as both a group and an algebraic variety. A \phi: G \to H between linear algebraic groups G and H is a that preserves the group operations, satisfying \phi(gh) = \phi(g)\phi(h) and \phi(g^{-1}) = \phi(g)^{-1} for all g, h \in G. The of \phi, defined as \ker \phi = \{ g \in G \mid \phi(g) = e_H \}, where e_H is the in H, is a closed of G. The image \phi(G) is a closed of H. An isogeny is a surjective homomorphism \phi: G \to H with finite kernel; such maps are finite morphisms of degree equal to the order of the kernel. Isomorphisms are bijective homomorphisms whose inverses are also homomorphisms, equivalently bijective with bijective differential at the identity. In positive characteristic, an isogeny is separable if its kernel is reduced (étale) or if the induced map on tangent spaces is injective; inseparability arises when the kernel contains non-reduced components, as in the Frobenius morphism. The center Z(G) of a linear algebraic group G is the closed subgroup consisting of all elements that commute with every element of G, given by Z(G) = \{ z \in G \mid zg = gz \ \forall g \in G \}. The derived subgroup G', or , is the smallest closed such that G/G' is abelian; it is generated by all commutators [g, h] = ghg^{-1}h^{-1} for g, h \in G. For connected semisimple groups, G' = G. Associated to any \phi: G \to H is its at the , a d\phi_e: \Lie(G) \to \Lie(H) between the algebras, defined via the at the . This map preserves the Lie bracket and is injective if \phi is separable; for , d\phi_e is an of algebras.

Examples

Classical matrix groups

The general linear group \mathrm{GL}(n,k) over a k consists of all invertible n \times n matrices with entries in k, defined as the Zariski-open subset of the of all n \times n matrices where the is nonzero. This group has dimension n^2, as its is the full matrix \mathfrak{gl}(n,k). It serves as the ambient space for many classical examples and is itself a fundamental linear algebraic group. The \mathrm{SL}(n,k) is the of the morphism \det: \mathrm{GL}(n,k) \to \mathbb{G}_m, where \mathbb{G}_m is the , consisting of all matrices in \mathrm{GL}(n,k) with 1. Its defining equation is \det(A) = 1 for A \in M_n(k), and it has dimension n^2 - 1, reflecting the single constraint on \mathrm{GL}(n,k). For n \geq 2, \mathrm{SL}(n,k) is a linear algebraic group over algebraically closed fields, meaning it has no nontrivial proper connected subgroups. The orthogonal group \mathrm{O}(n,k) (assuming \mathrm{char}(k) \neq 2) preserves the standard nondegenerate symmetric bilinear form, defined by the equation X^T X = I_n for X \in M_n(k), where I_n is the n \times n identity matrix. This group has dimension n(n-1)/2, corresponding to the number of independent entries above the diagonal in an orthogonal matrix. The special orthogonal group \mathrm{SO}(n,k) is the kernel of the determinant map on \mathrm{O}(n,k), satisfying the same equation with the additional condition \det(X) = 1, and thus shares the same dimension n(n-1)/2. The symplectic group \mathrm{Sp}(2n,k) preserves a nondegenerate alternating on k^{2n}, standardly given by the equation X^T J X = J, where J = \begin{pmatrix} 0 & I_n \\ -I_n & 0 \end{pmatrix} and I_n is the n \times n . This defining relation ensures the form \omega(x,y) = x^T J y remains invariant under the . The of \mathrm{Sp}(2n,k) is n(2n+1), arising from the constraints on the matrix entries that maintain the symplectic structure.

Elementary and other simple examples

One of the simplest non-trivial examples of a linear algebraic group is the \mathbb{G}_m, defined over a field k as the of the k[T, T^{-1}], which corresponds to the affine line \mathbb{A}^1_k minus the origin. This group has dimension 1 and consists of invertible elements k^\times, with the group operation given by multiplication in k. It serves as the building block for more complex tori, which are split as direct products of copies of \mathbb{G}_m. Another elementary example is the additive group \mathbb{G}_a, the spectrum of the Hopf algebra k[T], representing the affine line \mathbb{A}^1_k itself. Here, the group operation is polynomial addition on k, and \mathbb{G}_a is unipotent of dimension 1, with all elements satisfying g - 1 nilpotent. Any connected one-dimensional linear algebraic group over an algebraically closed field is isomorphic to either \mathbb{G}_m or \mathbb{G}_a. For a higher-dimensional unipotent example, consider the H_3 over k, a three-dimensional algebraic group of class 2. It can be realized as the with coordinates (x, y, z) \in k^3 and multiplication (x, y, z) \cdot (x', y', z') = (x + x', y + y', z + z' + x y'), where is one-dimensional, spanned by elements with x = y = 0. This structure illustrates the nilpotency typical of unipotent groups, with the also Heisenberg-type. Among simple linear algebraic groups, the projective special linear group \mathrm{PSL}_n(k) for n \geq 2 provides a classical example, which is and of semisimple n-1. Exceptional simple groups include those of types E_6, E_7, and E_8, with s 6, 7, and 8, respectively, and corresponding dimensions 78, 133, and 248; these arise from exceptional Lie algebras and have no classical analogs. Their Dynkin diagrams distinguish them from series like A_n. Non-split examples highlight field dependence: over the real numbers \mathbb{R}, the of the Hamilton quaternion algebra \mathbb{H} (with basis $1, i, j, k and relations i^2 = j^2 = k^2 = ijk = -1) forms a non-split inner form of \mathrm{SL}_2(\mathbb{C}), where the norm-1 elements yield the \mathrm{SU}(2). This contrasts with the split form \mathrm{M}_2(\mathbb{R}) over \mathbb{R}, demonstrating how quaternion algebras ramify at infinite places to produce anisotropic groups.

Structural Elements

Connected components and radical

In the Zariski topology on a linear algebraic group G, the connected components coincide with the irreducible components of G as an . These components all have the same dimension as G, and G is a finite of them. The identity component G^0 is the unique connected component containing the e; it is an irreducible affine that forms an open of G with finite index. As a , G^0 is stable under conjugation by elements of G, and the cosets gG^0 for g \in G partition G into its s. Over an , G^0 is , and it acts transitively on each irreducible component of homogeneous spaces under G. When G is defined over a subfield k, G^0 is also defined over k if k is perfect. Moreover, over perfect fields, G^0 is Zariski dense in G. The unipotent radical R_u(G) of G is defined as the largest unipotent of G; it is unique, characteristic (stable under automorphisms of G), closed, connected, nilpotent, and solvable. Equivalently, R_u(G) is the identity component of the intersection of the unipotent radicals of all Borel subgroups of G containing a fixed . Over a k, if G is a k-group, then R_u(G) is defined over k. The quotient G / R_u(G) is a reductive linear algebraic group. The solvable radical R(G) of G is the largest connected solvable ; it is unique, closed, and contains R_u(G) as its unipotent part. It can be characterized as the identity component of the intersection of all Borel subgroups of G. Over a , R(G) is defined over the base field and, in the reductive case, takes the form of a central . The G / R(G) is then semisimple. A subgroup L of G is a maximal reductive subgroup such that G = R_u(G) \rtimes [L](/page/Levi) via a ; it is unique up to conjugation and contains a of G. The reductive quotient G / R_u(G) is isomorphic to L, providing a structural that reduces the study of general linear algebraic groups to reductive ones. Over algebraically closed fields, the existence of such Levi subgroups follows from the of parabolic subgroups.

Lie algebra and its properties

The Lie algebra of a linear algebraic group G over an k, denoted \Lie(G), is defined as the at the e of G. This captures the structure of G, serving as a to the group near the . In characteristic zero, \Lie(G) can also be identified with the space of k-derivations \Der_k(k[G], k) of the coordinate ring k[G], where derivations are left-invariant under the . This identification equips \Lie(G) with a given by the of derivations, making it a over k. A key connection between the Lie algebra and the group is provided by the \exp: \Lie(G) \to G, which sends an element X \in \Lie(G) to the group element obtained via the \exp(X) = I + X + \frac{X^2}{2!} + \cdots, assuming a of G. This map is well-defined for elements and plays a central role in characteristic zero, where it relates the algebraic and structures. However, in positive characteristic, the exponential map is not always surjective onto unipotent elements, and its behavior is more restricted due to issues with formal power series convergence and p-nilpotency. The group G acts on \Lie(G) via the adjoint representation \Ad: G \to \Aut(\Lie(G)), defined by conjugation: for g \in G and X \in \Lie(G), \Ad(g)X = gXg^{-1}. The differential of this representation at the identity yields the adjoint action \ad: \Lie(G) \to \End(\Lie(G)), where \ad(X)Y = [X, Y] for Y \in \Lie(G). In characteristic zero, the Killing form \kappa(X, Y) = \Tr(\ad(X) \circ \ad(Y)) on \Lie(G) is an invariant bilinear form that provides important structural information, such as nondegeneracy for semisimple Lie algebras. This form arises naturally from the trace in the adjoint representation and aids in distinguishing solvable and semisimple components. For a closed subgroup H \leq G, the Lie algebra \Lie(H) is a Lie subalgebra of \Lie(G), obtained as the subspace of derivations vanishing on the ideal defining H. This inclusion preserves the Lie bracket and reflects the embedding of infinitesimal structures. Semisimple elements in \Lie(G) correspond to those whose adjoint action is diagonalizable over an algebraic closure. In positive characteristic p > 0, several classical properties of algebras fail to hold for those arising from linear algebraic groups. For instance, Cartan's criterion for solvability or , which relies on the form in the , does not generally apply, as the Killing form may degenerate even for simple algebras. Additionally, 's theorem on simultaneous triangularization of representations breaks down, exemplified by the \mathfrak{sl}_2(k) becoming solvable in characteristic 2 without upper-triangularizable representations. These limitations necessitate alternative approaches, such as restricted algebras incorporating p-operations.

Element Decompositions

Unipotent elements

In a linear algebraic group G defined over an k, an element g \in G is unipotent if all eigenvalues of g (with respect to any faithful rational representation of G) are equal to 1. This condition is independent of the choice of representation and characterizes unipotency intrinsically within the group. For the general linear group \mathrm{GL}(n, k), an g is unipotent if and only if g - I is a , meaning there exists a positive integer m \leq n such that (g - I)^m = 0, where I denotes the . In this setting, unipotent elements correspond to matrices that are conjugate to strictly upper triangular matrices with 1s on the diagonal. The set of all unipotent elements in a reductive linear algebraic group G forms a closed subvariety known as the unipotent cone. In characteristic zero, this cone corresponds closely to the image under the of the nilpotent cone in the of G, and the unipotent elements generate unipotent s. The centralizer C_G(u) of a unipotent u \in G in a connected reductive group is itself connected and has a structure given by a together with its unipotent . The normalizer N_G(u) is finite over C_G(u), often involving components related to the action on the centralizer. In positive characteristic, the definition of unipotency remains the same—all eigenvalues equal to 1—but the relationship to the differs significantly; for instance, not all unipotent elements arise as exponentials of elements in the , due to the limited surjectivity of the .

Semisimple elements and tori

In the theory of linear algebraic groups over a k, an element g \in G(k) of a linear algebraic group G is defined to be semisimple if, in every rational (V, \rho_V) of G, the linear operator \rho_V(g) is semisimple, meaning it is diagonalizable over the algebraic closure \overline{k}. This property is independent of the choice of representation and aligns with the semisimple part in the Jordan-Chevalley . Over perfect fields, semisimple elements form a G_s of G, though this subgroup is not necessarily closed in the ; for example, in the group B_2 of upper triangular $2 \times 2 matrices with determinant 1 over an of not 2, G_s is not closed. A fundamental property of semisimple elements is their containment in tori: every semisimple s \in G_s lies in some maximal of G. Moreover, the centralizer Z_G(s) of a semisimple s is connected and reductive. In connected solvable groups, the semisimple elements generate the derived subgroup, and their centralizers contain maximal tori. For reductive groups, regular semisimple elements—those whose centralizers are tori—form a Zariski-open dense of G, and their conjugates generate G. A in a linear algebraic group is a connected commutative algebraic T that becomes isomorphic, after base change to a finite separable k'/k, to a finite of copies of the \mathbb{G}_m. Equivalently, T is a connected consisting entirely of semisimple elements and is diagonalizable over \overline{k}, meaning its rational representations decompose as direct sums of one-dimensional representations. The X^*(T) of a T is a finitely generated , and T is smooth with dimension equal to the of X^*(T). Split tori, which are isomorphic to \mathbb{G}_m^r directly over k, serve as building blocks; every is a quotient of an induced torus, constructed as a product of twists (\mathbb{G}_m)_{k_i/k} for separable extensions k_i/k. Maximal tori play a central role in the structure of reductive and semisimple groups. A maximal torus in a connected linear algebraic group G is a torus not properly contained in any larger torus. In a connected reductive group, any two maximal tori are conjugate under an element of G(k_s), where k_s is the separable closure of k, and the centralizer of a maximal torus T is T itself. For example, in GL_n(k), the group of invertible n \times n matrices, the diagonal matrices form a split maximal torus of dimension n, with normalizer the monomial matrices and Weyl group the symmetric group S_n. Tori normalize unipotent subgroups and appear in the Levi decomposition of parabolic subgroups, underscoring their role in classifying representations and root systems.

Jordan-Chevalley decomposition

In a linear algebraic group G over an k, every element g \in G(k) admits a unique g = g_s g_u = g_u g_s, where g_s is semisimple and g_u is unipotent, with g_s, g_u \in G(k). This Jordan-Chevalley decomposition generalizes the classical Jordan canonical form for matrices and holds for any linear algebraic group, not just matrix groups like \mathrm{GL}_n(k). The proof proceeds by embedding G as a closed subgroup of \mathrm{GL}_n(k) and considering the regular representation or a faithful representation \phi: G \to \mathrm{GL}_V(k). For g \in G(k), the image \phi(g) decomposes uniquely as \phi(g) = \phi(g)_s \phi(g)_u = \phi(g)_u \phi(g)_s via the matrix Jordan form, where \phi(g)_s is diagonalizable over k and \phi(g)_u has all eigenvalues 1. Since the semisimple and unipotent parts are polynomials in \phi(g), they lie in \phi(G(k)), and by faithfulness of \phi, there exist unique g_s, g_u \in G(k) mapping to them. Uniqueness follows from the uniqueness in \mathrm{GL}_V(k) and the embedding properties. In characteristic 0, this relies directly on the Jordan canonical form over algebraically closed fields. The decomposition is preserved under group homomorphisms and automorphisms: if \psi: G \to H is a of linear algebraic groups, then \psi(g_s) and \psi(g_u) are the semisimple and unipotent parts of \psi(g), respectively. This functoriality extends the abstract Jordan decomposition in the to the group level via the in characteristic 0. In positive characteristic p > 0, the decomposition still exists over algebraically closed k, but the proof adapts using the fact that semisimple elements are diagonalizable over k and unipotent elements satisfy g^{p^m} = 1 for some m, without relying on the full Jordan form, which may fail. Chevalley's version refines this for semisimple groups by incorporating restricted root systems in the Lie algebra, ensuring the decomposition aligns with the p-structure via the Frobenius map and ensures compatibility with the Chevalley basis. For rationality over non-closed fields, if G is defined over a perfect field F \subseteq k and g \in G(F), then g_s, g_u \in G(F), so the decomposition is defined over F. This fails over imperfect fields, where separability issues arise, but holds for perfect F due to the polynomial nature of the parts. This property facilitates and applications in the study of forms of algebraic groups.

Key Subgroups

Maximal tori

A torus T in a linear algebraic group G defined over a field k is a connected diagonalizable subgroup, meaning it is isomorphic to a closed subgroup of the diagonal matrices in \mathrm{GL}_n(k) for some n. Over an algebraically closed field, every torus T is isomorphic to (\mathbb{G}_m)^r for some integer r \geq 0, where \mathbb{G}_m = k^\times is the multiplicative group and r = \dim T is the dimension of the torus. The character lattice X^*(T) = \mathrm{Hom}(T, \mathbb{G}_m) of a torus T is then a free abelian group of rank r, consisting of the algebraic group homomorphisms from T to \mathbb{G}_m. A maximal torus in G is a torus that is maximal among all tori with respect to inclusion, or equivalently, a maximal connected abelian subgroup consisting entirely of semisimple elements. In a connected linear algebraic group G over an algebraically closed field k, every maximal torus has the same dimension, called the rank of G, and any two maximal tori are conjugate under an element of G(k). Moreover, every semisimple element of G lies in some maximal torus. Over a general k, a T may not be , meaning it need not be isomorphic to (\mathbb{G}_m)^r over k itself; instead, there exists a finite L/k, called a splitting field for T, over which T_L becomes isomorphic to (\mathbb{G}_m)^r. The absolute rank or split rank of T is this dimension r, while the k-rank of G is the dimension of a maximal k-split torus in G, i.e., the largest r such that G contains a subgroup isomorphic to (\mathbb{G}_m)^r over k. Any torus decomposes uniquely as a product of its maximal k-split part and its k-anisotropic kernel. For a maximal torus T in a connected reductive group G, the normalizer N_G(T) and centralizer C_G(T) satisfy C_G(T)^0 = T, and the W(G, T) = N_G(T) / C_G(T) is a that acts faithfully on the character lattice X^*(T) by of the basis elements corresponding to the roots. This action is independent of the choice of up to isomorphism.

Unipotent subgroups

A unipotent U of a linear algebraic group G is a closed consisting entirely of unipotent elements, i.e., elements whose eigenvalues are all 1 in every rational representation of G. Such arise as the unipotent radicals of parabolic and are generated by unipotent elements of G. In characteristic 0, every unipotent is connected. Unipotent subgroups are nilpotent algebraic groups, meaning that the lower central series U = \gamma_1(U) \supset \gamma_2(U) \supset \cdots, defined by \gamma_{i+1}(U) = [U, \gamma_i(U)], terminates at the after finitely many steps. The [U, U] is itself a unipotent subgroup, and this property holds over any . Nilpotency implies the existence of a central series $1 = Z_0(U) \subset Z_1(U) \subset \cdots \subset Z_m(U) = U, where each Z_{i+1}/Z_i(U) is central in U/Z_i(U). Over a perfect field, a connected unipotent subgroup U admits a composition series with successive quotients isomorphic to the additive group \mathbb{G}_a, the one-dimensional unipotent group. Over a perfect field, a connected unipotent subgroup U of dimension d admits a composition series with successive quotients isomorphic to ℊ_a, making it a nilpotent group of dimension d. In particular, it is unipotent and, in characteristic 0, isomorphic as a variety to affine space 𝔸^d via the exponential map. In characteristic 0, the \exp: \operatorname{Lie}(U) \to U is an of algebraic varieties, identifying the with the group via a . The group multiplication on U corresponds to the Baker-Campbell-Hausdorff formula on \operatorname{Lie}(U), which expresses the product \exp(x) \exp(y) = \exp(Z(x,y)) for x, y \in \operatorname{Lie}(U), where Z(x,y) is a convergent in the . This facilitates the study of representations and structure theorems for unipotent subgroups.

Borel subgroups and parabolic subgroups

In a linear algebraic group G defined over an k, a B is a maximal connected closed solvable . Such a B contains a T (as detailed in the section on maximal tori) and can be decomposed as a B = T \ltimes U, where U is the unipotent radical of B consisting of unipotent elements (as covered in the section on unipotent subgroups). All of G are conjugate under the action of G(k). A fundamental consequence of this structure is the Bruhat decomposition, which expresses G as a G = \bigcup_{w \in W} B w B, where W = N_G(T)/T is the of G consisting of cosets of the normalizer N_G(T) of T in G modulo T. Each B w B is a locally closed subvariety of G, known as a Bruhat cell, with \dim(B w B) = \dim B + \ell(w), where \ell(w) denotes the of w in W with respect to a set of representatives for the simple reflections in W. A parabolic subgroup P of G is a proper closed connected that contains some B. Equivalently, P is parabolic if the G/P is a . Every parabolic subgroup P admits a P = L \ltimes R_u(P), where L is a connected reductive Levi subgroup (a factor of P) and R_u(P) is the unipotent radical of P, a unipotent . All Levi factors of P defined over k are conjugate by elements of R_u(P)(k). The quotient G/P, known as a (partial) flag variety, is a smooth projective parameterizing flags of subspaces stabilized by P. Its is given by \dim(G/P) = \dim G - \dim P. For a B, the full flag variety G/B has equal to the dimension of the unipotent radical of the opposite Borel. Given a Borel subgroup B = T \ltimes U containing the maximal torus T, the opposite Borel subgroup B^- is the unique Borel subgroup containing T such that B \cap B^- = T. It decomposes as B^- = T \ltimes U^-, where U^- is the unipotent radical opposite to U, generated by the unipotent subgroups corresponding to the negative roots relative to those in U. In a connected reductive group G, the subgroup generated by U and U^- equals the derived subgroup [G, G]. In general, the opposite Borels B and B^- play symmetric roles in decompositions of G.

Reductive and Semisimple Groups

Definitions and characterizations

A reductive linear algebraic group G over a k is a connected affine group such that its unipotent R_u(G) is trivial, meaning R_u(G) = \{e\}. Equivalently, G is reductive if the unipotent of G / Z(G)^0 is trivial, where Z(G)^0 denotes the of of G. A semisimple linear algebraic group is a reductive group with finite , or equivalently, one whose solvable R(G) is trivial. Thus, semisimple groups form a subclass of reductive groups, distinguished by the absence of nontrivial connected solvable subgroups and a finite . Reductive groups admit a structural characterization as G = Z(G)^0 \cdot [G, G], an almost where [G, G] is the derived subgroup, which is semisimple, with finite . Over a of characteristic zero, a connected linear algebraic group G is reductive if and only if its \mathfrak{g} = \mathrm{Lie}(G) is reductive. Prominent examples of reductive groups include the general linear group \mathrm{GL}_n, which has a one-dimensional center and thus is not semisimple, and classical groups such as the \mathrm{SO}_n and \mathrm{Sp}_{2n}. The \mathrm{SL}_n exemplifies a semisimple group, possessing finite center for n \geq 2.

Structure theorems including

One of the central structure theorems for linear algebraic groups concerns the Levi decomposition of parabolic subgroups. For a parabolic subgroup P of a reductive linear algebraic group G over a k, there exists a reductive L of P, called a Levi factor, such that P = L \rtimes R_u(P), where R_u(P) is the unipotent of P and the is . This decomposition holds over algebraically closed fields of zero for parabolic subgroups of reductive groups, and Levi factors are unique up to conjugation by elements of the unipotent . In positive , existence may fail for certain group schemes, such as \mathrm{[SL](/page/SL)}_n(W_2(k)), and conjugacy of Levi factors is not always guaranteed, though it holds when the dimension of the unipotent is less than the p. The proof of the Levi decomposition proceeds by induction on the dimension of G, using centralizers of . For a parabolic P containing a T, the centralizer C_G(T) intersects P in a Levi factor L, and the unipotent radical R_u(P) acts faithfully on L via the structure, ensuring the splitting. In characteristic zero, cohomological vanishing conditions, such as H^1(G, V) = 0 and H^2(G, V) = 0 for relevant modules V, guarantee the existence and conjugacy of Levi factors. Over fields of positive characteristic, restrictions arise from the and p-envelopes in the , which control the unipotent structure but may prevent splitting in non-reduced cases. A related theorem addresses solvable normal subgroups. For a connected reductive group G over a field of characteristic zero and any connected solvable N, the group decomposes as G = C_G(N) \rtimes N, where C_G(N) is the centralizer of N in G. This follows by on : the centralizer C_G(N) is reductive, normalizes N, and the action is faithful, yielding the . In positive , the decomposition holds for smooth groups over perfect s but requires adjustments for the unipotent part using p-envelopes to handle inseparability. Reductive groups also admit an almost direct product decomposition involving their center and derived subgroup. For a connected reductive group G over an , G = Z(G)^0 \cdot [G, G], where Z(G)^0 is the connected component of the (serving as the , a ), and [G, G] is the derived , with finite intersection. Here, [G, G] is semisimple, and the product is almost direct in the sense that the central Z(G)^0 intersects the semisimple part finitely. The proof uses on and centralizers: the centralizer of a yields the component, while the derived captures the semisimple structure, with finiteness ensured by counts. In positive , the persists for reductive groups, but the derived may involve Frobenius kernels, and intersections remain controlled by p-envelopes.

Classification

Root systems and Dynkin diagrams

In a semisimple linear algebraic group G over an of characteristic zero, with respect to a T, the \mathfrak{g} = \Lie(G) admits a space decomposition \mathfrak{g} = \Lie(T) \oplus \bigoplus_{\alpha \in \Phi} \mathfrak{g}_\alpha, where \Phi is the set of roots, each root space \mathfrak{g}_\alpha is one-dimensional, and for basis elements e_\alpha \in \mathfrak{g}_\alpha with \alpha > 0 and f_\alpha = e_{-\alpha} \in \mathfrak{g}_{-\alpha}, the Lie bracket satisfies [e_\alpha, f_\alpha] = h_\alpha with h_\alpha \in \Lie(T). This decomposition arises from the adjoint action of T on G, where roots correspond to the non-trivial weight spaces. The roots \Phi form a finite reduced root system in the real vector space X^*(T) \otimes \mathbb{R}, where X^*(T) is the character group of T, consisting of nonzero elements \alpha \in X^*(T) \otimes \mathbb{Q} such that \mathfrak{g}_\alpha \neq \{0\}. A subset \Delta \subset \Phi of simple roots is a basis for the root lattice over \mathbb{Z}, linearly independent over \mathbb{R}, such that every root is an integer linear combination of elements in \Delta with all coefficients nonnegative or all nonpositive. The Cartan matrix A = (a_{ij}) associated to \Delta = \{\alpha_i\} has entries a_{ij} = 2 (\alpha_i, \alpha_j) / (\alpha_j, \alpha_j), where (\cdot, \cdot) is the invariant bilinear form on X^*(T) \otimes \mathbb{R} induced from the Killing form on \mathfrak{g}; the diagonal entries are 2, and off-diagonal entries are nonpositive integers. This matrix is independent of the choice of simple roots and determines the root system up to isomorphism. The Dynkin diagram encodes the Cartan matrix as a graph with vertices corresponding to simple roots and edges (single, double, or triple) indicating the absolute values of off-diagonal entries, oriented arrows distinguishing root lengths when unequal. The irreducible root systems of rank n are classified by the following diagrams: A_n (linear chain of n vertices, corresponding to \mathfrak{sl}_{n+1}); B_n (chain of n vertices with a double bond and arrow at the short root end); C_n (chain with double bond and arrow at the long root end); D_n (chain of n-2 vertices branching to two at the end); E_6, E_7, E_8 (chain with a branch at the third vertex from the end, extending to lengths 6, 7, 8); F_4 (chain with double and triple bonds); and G_2 (two vertices connected by a triple bond with arrow). Extended Dynkin diagrams append an additional vertex connected to the highest root, yielding affine types used in the study of affine Kac-Moody algebras. The W is the finite group generated by reflections s_\alpha: \lambda \mapsto \lambda - \langle \lambda, \alpha^\vee \rangle \alpha for \alpha \in \Phi, where \alpha^\vee = 2\alpha / (\alpha, \alpha) is the coroot; it acts faithfully on X^*(T) \otimes \mathbb{R} and permutes . The longest element w_0 \in W is the unique element sending all simple roots to negative roots, with length equal to the number of positive roots. A choice of Borel subgroup B containing T determines a set of positive \Phi^+ = \{\alpha \in \Phi \mid \mathfrak{g}_\alpha \subset \Lie(U)\}, where U is the unipotent of B, consisting of those with nonnegative coefficients in the basis \Delta. The on \Phi^+ is defined by \ht(\alpha) = \sum k_i for \alpha = \sum k_i \alpha_i with \alpha_i \in \Delta and k_i \in \mathbb{N}, attaining its maximum at the highest .

Chevalley groups and split forms

Chevalley introduced a basis for the of a semisimple algebraic group that allows the construction of models over the s. For a with simple \alpha_i, the Chevalley basis consists of elements h_i (corresponding to simple coroots), e_\alpha for positive roots \alpha, and f_\alpha = e_{-\alpha} for negative roots, satisfying [e_\alpha, f_\alpha] = h_\alpha where h_\alpha = \sum k_i h_i with coefficients k_i, and the in the Lie brackets [e_\alpha, e_\beta] = N_{\alpha,\beta} e_{\alpha+\beta} (with N_{\alpha,\beta} \in \mathbb{Z}) are integers determined by the . This basis ensures that the admits a \mathbb{Z}-form, enabling the extension to characteristic zero and positive characteristics. Using this basis, Chevalley constructed a split reductive G over \mathbb{Z} for each root datum, generated by a maximal torus T and unipotent subgroups U_\alpha (one for each positive root \alpha), subject to relations mirroring the brackets, such as [x_\alpha(t), x_\beta(u)] = \prod x_{\alpha+\beta}(c t^k u^m) with constants c, k, m. The group G is smooth and affine over \Spec(\mathbb{Z}), with generic fiber the reductive group over \mathbb{Q}, and it possesses a BN-pair : a B = T U (with U the product of U_\alpha) and a normalizer N generated by T and elements n_\alpha satisfying the Tits axioms, including the Bruhat G = \bigcup B w B for w in the . Specializing to finite fields \mathbb{F}_q (with q = p^n, p prime not dividing certain denominators from the ), the groups G(\mathbb{F}_q) are the finite Chevalley groups, which coincide with the fixed points of a on the split form over the . For adjoint types (where the root datum has trivial center), these groups are simple except in small cases like A_1(\mathbb{F}_2), A_1(\mathbb{F}_3), and B_2(\mathbb{F}_2). The universal Chevalley group corresponds to the simply connected form (minimal kernel in the to the form), while the form is the by ; over \mathbb{F}_q, the simply connected versions may have non-trivial Schur multipliers (the second H_2(G(\mathbb{F}_q), \mathbb{Z})), which are typically cyclic of order 1, 2, 3, or 8 depending on the type and q, as computed for all irreducible root systems. Twisted forms arise by applying a Galois automorphism \sigma of the algebraic closure that acts on the via a (of order 2 or 3 for non-exceptional types), yielding groups like ^2A_n(\mathbb{F}_{q^2}) or ^2B_2(\mathbb{F}_{q^2}) as fixed points under the twisted \text{Fr}_\sigma, which permutes roots accordingly while preserving the BN-pair structure up to .

Classification over non-algebraically closed fields

Over a non-algebraically closed k, the isomorphism classes of reductive linear algebraic groups G with a given root datum are classified using . Specifically, the forms of a fixed reductive group \hat{G} over the \bar{k} are parametrized by the first set H^1(k, \hat{G}), which captures the k-twists of \hat{G}. This cohomology measures the extent to which \hat{G} fails to be over k, with trivial elements corresponding to the form. Inner forms arise from the center of the simply connected , classified by H^1(k, Z(\hat{[G](/page/G)})), where Z(\hat{[G](/page/G)}) is of the dual group \hat{[G](/page/G)}. These are isogenous to the split form and preserve the . Outer forms, in , are classified by H^1(k, \Aut(\hat{[G](/page/G)})), reflecting automorphisms of the root datum that do not arise from inner automorphisms; they include more general twists and are parametrized by torsors under the outer \Out(\hat{[G](/page/G)}). For semisimple groups, the full set of forms is the of H^1(k, \Aut(\hat{[G](/page/G)})) \to H^1(k, \hat{[G](/page/G)}). A reductive group G over k is quasi-split if it contains a defined over k, equivalently, if it admits a maximal torus over k. In this case, the minimal parabolic subgroup over k is unique up to conjugation, and the relative is generated by a of simple with no distinguished roots (i.e., \Delta_0 = \emptyset). Quasi-split forms represent the "most split" non-split possibilities, with a unique inner form up to . The anisotropic kernel of G with respect to a maximal split torus S is the derived subgroup of the centralizer Z_G(S); it is an anisotropic semisimple group. For semisimple G, the anisotropic kernel is anisotropic semisimple. Dimension bounds follow from the structure: the dimension of a maximal split torus S satisfies \dim S \geq \# k\Delta, where k\Delta is the set of k-simple roots, and the dimension of a Borel subgroup is \frac{1}{2}(\dim G + \dim S). Anisotropic groups over local fields like \mathbb{R} have compact groups of k-points. Tits provided a complete classification of semisimple groups over arbitrary fields using the geometry of buildings and apartments associated to maximal split tori. The building of G is a whose apartments correspond to maximal split tori, with chambers labeled by elements; the k-structure is captured by the on the via the Frobenius or \Gamma = \Gal(\bar{k}/k). The invariant is the index I(G) = (D, \Delta_0, \star), where D is the , \Delta_0 the set of distinguished simple roots (fixed by \Gamma), and \star indicates diagram automorphisms. This classifies all forms up to isomorphism, extending the split case. For example, over \mathbb{R}, the special linear group \SL_n has two real forms up to isomorphism: the split form \SL_n(\mathbb{R}), which is non-compact, and the compact form \SU(n), arising as an inner twist via the quaternion algebra (for even n) or outer automorphism. The compact form has trivial \mathbb{R}-points beyond \{\pm I\} in certain cases, illustrating anisotropy, while the split form has maximal rank tori of dimension n-1.

Representations

Irreducible representations and weights

For a semisimple linear algebraic group G over an algebraically closed field k of characteristic zero, the finite-dimensional irreducible rational representations are parametrized by dominant weights in the character lattice X^*(T) of a maximal torus T \subset G. In characteristic zero, all finite-dimensional rational representations of semisimple groups are completely reducible into direct sums of irreducibles. A weight \Lambda \in X^*(T) is dominant if it lies in the closed fundamental Weyl chamber X^*(T)^+, defined as the set of \lambda such that \langle \lambda, \alpha^\vee \rangle \geq 0 for all simple coroots \alpha^\vee corresponding to a choice of positive roots. This parametrization follows from the highest weight theorem, which asserts that every irreducible representation has a unique highest weight that is dominant. Given a dominant weight \Lambda \in X^*(T)^+, the corresponding irreducible representation is denoted V(\Lambda), realized as the quotient of the induced module from the Borel subgroup by its unique simple submodule. It admits a highest weight vector v_\Lambda such that t \cdot v_\Lambda = \Lambda(t) v_\Lambda for t \in T and the unipotent radical U_\alpha of a positive root subgroup acts trivially on v_\Lambda for each positive root \alpha > 0, i.e., U_\alpha v_\Lambda = 0. The weights of V(\Lambda) all lie in the convex hull of the Weyl group orbit W \cdot \Lambda, and the representation is completely determined by the action on this cyclic vector generated by v_\Lambda. The character \chi_\Lambda of V(\Lambda), which encodes the formal sum of weights with multiplicities, is given by the : \chi_\Lambda = \frac{\sum_{w \in W} \varepsilon(w) e^{w(\Lambda + \rho) - \rho}}{\sum_{w \in W} \varepsilon(w) e^{w(\rho) - \rho}}, where W is the , \varepsilon(w) is the of w, \rho is half the sum of the positive roots, and e^\mu denotes the basis element for weight \mu. This formula, originally due to Weyl for compact Lie groups but valid in the algebraic setting via of s, allows explicit computation of dimensions and weight multiplicities for classical types. To compute the multiplicity m(\mu) of a weight \mu in V(\Lambda), the recursive Freudenthal multiplicity formula provides an efficient : for \mu \neq \Lambda, m(\Lambda, \mu) = \sum_{\alpha > 0} \frac{(\Lambda + \mu + 2\rho, \alpha)}{(\alpha, \alpha)} m(\Lambda, \mu + \alpha), with m(\Lambda, \Lambda) = 1 and m(\Lambda, \nu) = 0 if \nu \not\leq \Lambda in the dominance order. This formula, derived from the adjoint action of the on the tensor product with the , is particularly useful for verifying branching rules and decomposing tensor products in low ranks. In positive characteristic p > 0, the highest weight theory simplifies for restricted dominant weights \Lambda \in X_1(T)^+, where $0 \leq \langle \Lambda, \alpha^\vee \rangle < p for all simple coroots; the modules L(\Lambda) with such \Lambda are irreducible and form a complete set of simple rational G-modules of restricted highest weight. For general dominant \Lambda, the Frobenius endomorphism F: G \to G induces kernels G_r = \ker(F^r) for r \geq 1, which are infinitesimal group schemes whose representations parametrize the composition factors of L(\Lambda) via the Steinberg tensor product theorem: L(\Lambda) \cong L(\Lambda_0) \otimes L(\Lambda_1)^{(1)} \otimes \cdots \otimes L(\Lambda_s)^{(s)}, where \Lambda = \Lambda_0 + p \Lambda_1 + \cdots + p^s \Lambda_s with each \Lambda_i restricted, and V^{(j)} denotes the j-th Frobenius twist. This decomposition highlights the p-adic structure but introduces complications like non-vanishing cohomology absent in characteristic zero.

Induced and tensor representations

One fundamental construction of representations for a linear algebraic group G over an algebraically closed field k is induction from a subgroup H \leq G. Given a rational representation \rho: H \to \mathrm{GL}(V) of H, the induced representation \mathrm{Ind}_H^G V is the k[G] \otimes_{k[H]} V, the tensor product of the coordinate ring of G over that of H with V, equipped with the natural G-action. For parabolic subgroups P, this yields a rational G-representation whose underlying vector space includes the finite-dimensional space of global sections of the associated vector bundle on G/P. Induction from Borel subgroups B is especially significant for reductive groups, as it realizes finite-dimensional irreducible representations. For a character \chi_\lambda: B \to k^\times corresponding to a weight \lambda \in X(T) \otimes \mathbb{Z} (with T a maximal torus in B), the induced module \mathrm{Ind}_B^G k_\lambda (often denoted E(\lambda)) has nonzero global sections precisely when \lambda is dominant, and it has a unique simple quotient L(\lambda) of highest weight \lambda. This follows from the structure of parabolic inductions and the semisimplicity of representations in characteristic zero. Frobenius reciprocity provides a duality between induction and restriction functors, facilitating the study of these constructions. Specifically, for rational G-representations W and H-representations V, there is a natural isomorphism \mathrm{Hom}_G(W, \mathrm{Ind}_H^G V) \cong \mathrm{Hom}_H(\mathrm{Res}_H^G W, V), which preserves exactness and dimensions. This adjunction is central to decomposing induced modules and computing multiplicities in representation theory. A geometric incarnation of induction from Borel subgroups is given by the Borel–Weil theorem, which realizes irreducible representations as cohomology of line bundles on the flag variety. For a complex semisimple group G with Borel subgroup B and dominant integral weight \lambda, let L_\mu denote the line bundle on G/B associated to weight \mu, and let w_0 be the longest element of the Weyl group W. Then, the contragredient of the induced representation (\mathrm{Ind}_B^G L_{-w_0 \lambda})^\vee is isomorphic to the irreducible representation V(\lambda) of highest weight \lambda, where higher cohomology vanishes. This theorem, extended by Bott to account for Weyl group action in general cases, underscores the interplay between algebraic and geometric methods. Tensor products offer another construction, decomposing into direct sums of irreducibles via combinatorial coefficients. For the general linear group \mathrm{GL}_n(\mathbb{C}), if V(\lambda) and V(\mu) are irreducible polynomial representations labeled by partitions \lambda, \mu, their tensor product decomposes as V(\lambda) \otimes V(\mu) = \bigoplus_\nu c^\nu_{\lambda \mu} V(\nu), where the Littlewood–Richardson coefficients c^\nu_{\lambda \mu} are nonnegative integers counting semi-standard Young tableaux of skew shape \nu / \lambda with content \mu that are lattice permutations. These coefficients determine the representation ring of \mathrm{GL}_n and extend to other classical groups via branching rules. Restriction of representations to Levi subgroups of parabolic subgroups P = L U (with L the Levi factor) yields modules that decompose into direct sums over characters of the center of L, often analyzable as . These are finitely generated modules over the universal enveloping algebra of the Lie algebra of L, with finite-dimensional generalized eigenspaces for the center, finite multiplicity for each generalized weight, and L-finite action. Such modules facilitate parabolic induction, where \mathrm{Ind}_P^G (V \otimes \chi) for V a representation of L and \chi a character of L produces representations with explicit structure in characteristic zero. In positive characteristic p > 0, the tensor product theorem provides a decomposition for irreducible rational representations of semisimple algebraic groups. For a dominant weight \lambda \in X(T)^+, write \lambda = \lambda_0 + p \lambda_1 + \cdots + p^r \lambda_r in its p-adic expansion, with each $0 \leq \lambda_i < p. Then, the irreducible module L(\lambda) is isomorphic to the L(\lambda) \cong L(\lambda_0) \otimes L(\lambda_1)^{(1)} \otimes \cdots \otimes L(\lambda_r)^{(r)}, where each L(\lambda_i) is the irreducible module of restricted highest weight \lambda_i. This theorem, due to , exploits Frobenius morphisms and controls the complexity of representations in modular settings.

Applications

Geometric invariant theory and quotients

Geometric invariant theory (GIT) provides a framework for constructing quotients of algebraic varieties under actions of linear algebraic groups, particularly reductive ones, by associating invariants to orbits and forming moduli spaces that parametrize isomorphism classes. For a reductive linear algebraic group G acting on an affine variety X over an algebraically closed field k, the ring of invariants k[X]^G consists of regular functions on X that are fixed by the G-action. This ring is finitely generated as a k-algebra when G is reductive, ensuring that the categorical quotient X // G = \operatorname{Spec} k[X]^G exists as an affine variety. The natural projection \pi: X \to X // G is a G-invariant morphism that is constant on G-orbits, separates closed G-invariant subsets, and is universal among such morphisms with these properties. To extend this to projective varieties and ensure well-behaved quotients, GIT introduces notions of stability via linearizations of ample line bundles. A key tool is the , which characterizes semistable points in terms of one-parameter subgroups (1-PS). For a linearized action of G on a projective variety X = \mathbb{P}(V), a point x \in X is semistable if there exists no 1-PS \lambda of G such that \lim_{t \to 0} \lambda(t) \cdot x = 0. Points satisfying the stricter condition that the limit exists and the stabilizer is finite-dimensional are stable. The semistable locus X^{ss} admits a good quotient X^{ss} // G, which is a projective geometric quotient for the stable locus X^s, meaning orbits are closed and separated in the quotient. David Mumford developed GIT in 1965 to construct such quotients systematically, resolving issues with orbit closures and providing projective moduli spaces for stable objects under group actions. For affine actions, the existence of finite generation of invariants in positive characteristic relies on the geometric reductivity of reductive groups, established by Haboush's theorem: over any field, a reductive group is geometrically reductive, implying that the invariants k[X]^G are finitely generated for any affine X and thus affine quotients exist. This result, originally conjectured by Mumford, confirms that semisimple groups (hence reductive ones) yield well-defined categorical quotients without pathological behavior in mixed characteristic. A classic example arises from the action of \mathrm{SL}(2) on the space of binary forms of degree d, viewed as an affine variety in coefficients. The ring of invariants is generated by the discriminant for d=2 (a quadratic form) and more generally by covariants like the Aronhold invariants for higher degrees, allowing the categorical quotient to parametrize isomorphism classes of such forms up to \mathrm{SL}(2)-equivalence. This construction illustrates how GIT quotients capture moduli of plane curves or conics, with semistable points excluding those with repeated roots via the Hilbert-Mumford criterion applied to diagonal 1-PS.

Connections to Lie theory and number theory

Linear algebraic groups defined over the real or complex numbers give rise to Lie groups through their points over these fields. For a linear algebraic group G defined over \mathbb{C}, the group G(\mathbb{C}) is a complex Lie group, and when G is simply connected, semisimple, and connected, it coincides with the simply connected complex Lie group having the same Lie algebra. Similarly, for G defined over \mathbb{R}, the real points G(\mathbb{R}) form a real Lie group, which in the semisimple case is a covering group of the real points of an algebraic group, with the covering related via the universal cover of the corresponding complex group. These identifications allow the structure theory of Lie groups to inform the analytic properties of algebraic groups over \mathbb{R} and \mathbb{C}. A key structural feature for real semisimple Lie groups arising from linear algebraic groups is the Cartan decomposition. For such a group G(\mathbb{R}), there exists a Cartan involution \theta on the Lie algebra \mathfrak{g}_\mathbb{R}, yielding a decomposition \mathfrak{g}_\mathbb{R} = \mathfrak{k} \oplus \mathfrak{p}, where \mathfrak{k} is the Lie algebra of a maximal compact subgroup K and \mathfrak{p} is the orthogonal complement with respect to the Killing form made positive definite by \theta. At the group level, this extends to G(\mathbb{R}) = K \exp(\mathfrak{p}), a diffeomorphism that generalizes the polar decomposition of matrices and facilitates the study of invariant metrics and representations. For complex semisimple Lie groups G(\mathbb{C}) from linear algebraic groups, the Iwasawa decomposition provides a useful uniformization. It decomposes G(\mathbb{C}) = K A N, where K is a maximal compact subgroup, A is a maximal split torus, and N is the unipotent radical of a Borel subgroup, with the decomposition being a diffeomorphism. This structure, analogous to the real case but adapted to the complex setting, aids in analyzing the topology and harmonic analysis on these groups, reducing problems to the compact factor K. In the specific case of G = \mathrm{GL}_n, the decomposition takes the form \mathrm{GL}_n(\mathbb{R}) = N A K with K = \mathrm{O}_n(\mathbb{R}), A the diagonal positive matrices, and N upper triangular unipotent. Turning to number theory, arithmetic subgroups of linear algebraic groups play a central role in the study of automorphic forms and related arithmetic objects. For a linear algebraic group G defined over the rationals \mathbb{Q}, an arithmetic subgroup \Gamma is the intersection of G(\mathbb{Q}) with a lattice in G(\mathbb{R}), such as \Gamma = G(\mathcal{O}) for a ring of integers \mathcal{O}. These subgroups are discrete in G(\mathbb{R}) and act properly discontinuously on symmetric spaces associated to G, yielding quotients that parametrize arithmetic data like modular curves. A prototypical example is G = \mathrm{SL}_2, where \Gamma = \mathrm{SL}_2(\mathbb{Z}) is an arithmetic subgroup whose action on the upper half-plane produces the modular curve X(1), and modular forms of weight $2k for \Gamma are holomorphic functions f: \mathbb{H} \to \mathbb{C} satisfying f\left( \frac{az + b}{cz + d} \right) = (cz + d)^{2k} f(z) for \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \Gamma, holomorphic at cusps. These forms generate L-functions with Euler products, linking to Dirichlet series and arithmetic invariants like class numbers. The Langlands program establishes profound connections between linear algebraic groups and number theory by conjecturing a correspondence between automorphic representations and Galois representations. For a reductive linear algebraic group G over a number field F, the program posits that irreducible automorphic representations \pi of the adelic group G(\mathbb{A}_F) (factorizing as tensors \pi = \otimes_v \pi_v over places v) correspond to continuous homomorphisms \rho: \mathrm{Gal}(\overline{F}/F) \to {}^L G(\mathbb{C}), where {}^L G is the Langlands dual group. The local components \pi_v at finite places match the Frobenius semisimplification of \rho via L-parameters, with global compatibility ensured by matching Artin and automorphic L-functions L(s, \pi) = L(s, \rho). For G = \mathrm{GL}_n, this reduces to a bijection between cuspidal automorphic representations and n-dimensional Galois representations, realizing number-theoretic objects like modular forms as geometric Galois data. In the local setting over p-adic fields, points G(F) for a non-archimedean local field F (residue characteristic p) form p-adic Lie groups central to the local Langlands correspondence. These groups admit the Bruhat-Tits building B(G, F), a simplicial complex on which G(F) acts, with vertices parametrized by special parahoric subgroups and edges reflecting valuations on the root datum. The building encodes the structure of hyperspecial maximal compact subgroups and filtration spaces, facilitating the classification of smooth irreducible representations of G(F). The local Langlands conjecture, proven for certain groups like \mathrm{GL}_n, asserts a bijection between such representations and irreducible representations of the Weil-Deligne group of F, with the building providing a geometric framework for supercuspidal induction and depth-zero types in the correspondence. This local picture underpins the global Langlands program by providing the place-by-place matching required for the adelic automorphic forms.

Computational aspects and modern uses

Computational software systems such as GAP and Magma provide essential tools for constructing and manipulating linear algebraic groups, particularly Chevalley groups and root systems. In GAP, the CHEVIE package implements algorithms for finite reflection groups and reductive algebraic groups over finite fields, enabling the computation of Weyl group representations and character tables for groups like SL_n(q). Similarly, Magma's intrinsic functions support the creation of root systems and root data for semisimple groups, allowing users to compute Dynkin diagrams, positive roots, and Galois actions on groups defined over number fields. These capabilities facilitate the explicit construction of split forms of linear algebraic groups, such as the special linear group SL_n over finite fields, by generating generators and relations from root system data. Recognition algorithms for black-box presentations of linear algebraic groups, where elements are accessed only via multiplication oracles, have advanced significantly for classical types. These methods identify groups isomorphic to PSL_n(q) or other projective special linear groups by probabilistic testing of subgroup structures, often running in polynomial time relative to the group order. For classical groups, the Aschbacher-O'Nan-Scott theorem provides a structural foundation, classifying maximal subgroups into geometric, irreducible, or almost simple classes, which guides constructive recognition by reducing the problem to verifying specific embedding dimensions and field characteristics. Such algorithms output standard generators, enabling further computations like order calculation or representation construction, with Las Vegas guarantees for correctness. In modern machine learning, representations of linear algebraic groups underpin equivariant neural networks, which enforce symmetry in models for physical systems. Post-2020 developments include E(3)-equivariant architectures that use irreducible representations of rotation and translation groups to process molecular structures, improving generalization in density functional theory predictions. Similarly, Clifford group equivariant networks leverage spin representations for O(3)-invariant tasks, such as protein folding, by parameterizing layers with group actions to preserve equivariance under transformations. These approaches draw on highest weight theory for decomposing tensor products, linking back to irreducible representations while enhancing efficiency in high-dimensional data. Quantum computing applications exploit Lie group structures from linear algebraic groups for algorithm design and complexity analysis. Recent work employs Lie algebras to characterize variational quantum circuits, deriving explicit generators for the unitary group and analyzing barren plateaus through concentration of measure on group manifolds. For instance, free-fermion encodings map to Lie subalgebras, enabling simulation of symmetry-protected phases with reduced qubit overhead. These methods also inform quantum advantage proofs by quantifying entanglement via representation theory of reductive groups. In particle physics, reductive linear algebraic groups model gauge symmetries, with spontaneous symmetry breaking generating masses via Higgs mechanisms. The Standard Model's SU(2) × U(1) electroweak sector, a reductive group over the complexes, breaks to U(1) electromagnetism, yielding the W and Z boson masses while preserving photon masslessness. Grand unified theories extend this to larger reductive groups like SU(5), where breaking patterns unify forces and predict proton decay, though unobserved, constraining model parameters. Numerical methods for matrix groups over the reals emphasize stability in Lie group integrators to preserve manifold structure. Lie-group exponential maps ensure solutions remain in subgroups like SO(n), avoiding drift from rounding errors in Euler discretizations. For GL(n, ℝ), symmetry-adapted solvers use Cayley transforms for orthogonal projections, maintaining positive definiteness and conditioning in optimization tasks. These techniques achieve backward stability, with error bounds scaling as machine epsilon times the condition number of the representation.

Relation to Lie groups

Linear algebraic groups defined over the real numbers \mathbb{R} or complex numbers \mathbb{C} yield Lie groups via their points over these fields: the set G(\mathbb{R}) forms a real Lie group, and G(\mathbb{C}) forms a complex Lie group, embedded as closed subgroups of the general linear group \mathrm{GL}(n, \mathbb{R}) or \mathrm{GL}(n, \mathbb{C}). Over \mathbb{C}, every connected semisimple Lie group is algebraic, meaning it arises precisely as the complex points of a linear algebraic group. Over \mathbb{R}, connected semisimple real Lie groups are covering groups of real algebraic groups, with the algebraic structure providing a polynomial description of the group's defining relations. These Lie groups admit a maximal compact subgroup K, which is compact and lies in the center of the group's unitary representations; for instance, in the special linear group \mathrm{SL}(n, \mathbb{R}), K = \mathrm{SO}(n). Linear algebraic groups serve as an "integral form" or algebraic backbone for their associated Lie groups, capturing the structure in terms of polynomial equations over the base field. A prime example is \mathrm{SL}(n, \mathbb{R}), which acts as a real form of the complex Lie group \mathrm{SL}(n, \mathbb{C}): the former consists of real matrices of determinant 1, providing an algebraic model that restricts the complex group's analytic freedom to real coefficients while preserving the Lie algebra \mathfrak{sl}(n, \mathbb{R}). This real form is obtained via complex conjugation as an involution on \mathrm{SL}(n, \mathbb{C}), yielding fixed points that form the real algebraic group. Such forms highlight how algebraic groups encode the "discrete" or polynomial aspects of Lie groups, facilitating arithmetic and geometric applications. The representation theory of these real Lie groups draws heavily on the algebraic data, as pioneered by Harish-Chandra. Specifically, the Harish-Chandra embedding theorem allows the construction of discrete series representations—square-integrable irreducible unitary representations—from the algebraic structure of the group, embedding the representation into induced modules over the complexified algebraic group. For reductive real algebraic groups, this approach shows that discrete series exist if and only if the group admits a compact Cartan subgroup, linking analytic unitarity to algebraic compactness conditions. Infinitesimally, linear algebraic groups and their Lie group realizations share the same Lie algebra, defined as the tangent space at the identity equipped with the commutator bracket, ensuring local equivalence near the identity element. However, the algebraic framework is more restrictive: linear algebraic groups are Zariski-closed subgroups defined by polynomial ideals, imposing global polynomial constraints absent in general Lie groups. Consequently, not all Lie subgroups of algebraic groups are algebraic; a canonical counterexample is the one-parameter subgroup of the 2-torus T^2 = \mathbb{R}^2 / \mathbb{Z}^2 generated by the flow (t \mapsto (t, \alpha t \mod 1)) for irrational \alpha, which is dense and non-closed, hence non-algebraic despite being a Lie subgroup. This illustrates the analytic flexibility of Lie groups beyond algebraic varieties.

Group schemes and Hopf algebras

Linear algebraic groups can be generalized to the broader framework of affine group schemes over a base field k, where an affine group scheme G is defined as the spectrum of a Hopf algebra A over k. The Hopf algebra structure on A encodes the group operations: the comultiplication \Delta: A \to A \otimes_k A corresponds to the multiplication map, the counit \varepsilon: A \to k to the unit map, and the antipode S: A \to A to the inversion map. This construction allows for a functorial description of the group, where G represents the functor from the category of k-algebras to groups given by G(R) = \Hom_{k\text{-alg}}(A, R) for any commutative k-algebra R. Linear algebraic groups correspond precisely to those affine group schemes that are of finite type over k, meaning A is finitely generated as a k-algebra. The representable functor perspective unifies the geometric and algebraic viewpoints: morphisms from an affine scheme \Spec R to G in the category of schemes over k are in natural isomorphism with algebra homomorphisms A \to R, preserving the group structure via the Hopf algebra maps. This equivalence extends the classical notion of points of an algebraic group, allowing group schemes to capture infinitesimal or non-reduced structures that linear algebraic groups over algebraically closed fields may not exhibit. For instance, in positive characteristic, affine group schemes of finite type include linear algebraic groups but also more general objects like finite flat group schemes, which are flat and of finite presentation over k. Finite flat group schemes provide concrete examples beyond smooth linear algebraic groups. The scheme of nth roots of unity, denoted \mu_n, is the affine group scheme \Spec k/(x^n - 1), representing the functor R \mapsto \{r \in R^\times \mid r^n = 1\} on k-algebras R. In characteristic p > 0, the Frobenius kernel \alpha_p is another finite flat group scheme, given by \Spec k/(x^p), which represents the additive group functor R \mapsto (R, +) but truncated at p-torsion elements, as it is the kernel of the Frobenius endomorphism on the additive group \mathbb{G}_a. These schemes are flat over k and of order p or n, illustrating how Hopf algebras can model group-like objects with nilpotent elements. Representations of an affine group scheme G = \Spec A are equivalently comodules over the Hopf algebra A, where a finite-dimensional representation corresponds to a right comodule structure \rho: V \to V \otimes_k A satisfying coassociativity and counit properties derived from the coalgebra structure of A. The category of finite-dimensional comodules over A is thus tensorial and equivalent to the representation category of G, generalizing the module category for linear algebraic groups. This coalgebra perspective highlights the duality between the group scheme and its Hopf algebra, where coactions encode how G acts on vector spaces. Over non-algebraically closed fields, defining linear algebraic groups requires descent data to ensure consistency under base change. For an affine G over a K/k with \Gamma, descent data consists of a \Gamma-action on the Hopf algebra A_K compatible with the Hopf structure, allowing reconstruction of a A over k such that G = \Spec A descends properly. More generally, faithfully flat descent applies to arbitrary base changes, ensuring that over k can be obtained from those over a faithfully flat extension via cocycle conditions on the Hopf algebra maps. This framework is essential for studying linear algebraic groups over fields like or finite fields, where classifies torsors and forms.

Tannakian categories and abelian varieties

A Tannakian category over a k is defined as a rigid abelian tensor category \mathcal{C} equipped with a \omega: \mathcal{C} \to \mathrm{Vec}_k, which is an exact faithful k-linear tensor to the category of finite-dimensional k-vector spaces, such that the endomorphism ring of the unit object is k. The automorphism \mathrm{Aut}^\otimes(\omega) of the is a pro-algebraic over k, representing the tensor automorphisms of \omega. The reconstruction theorem states that for a neutral Tannakian category \mathcal{C} over k, there exists an affine group scheme G over k such that \mathcal{C} \cong \mathrm{Rep}_k(G) as tensor categories, via the equivalence induced by \omega, where G = \mathrm{Aut}^\otimes(\omega). This duality recovers the linear algebraic group G from its representation category, generalizing Tannaka's original theorem for compact groups to the algebraic setting. Abelian varieties provide a commutative specialization of linear algebraic groups, serving as the projective analogues of the additive group \mathbb{G}_a. An A over a k is a complete connected commutative group variety, meaning it is a projective algebraic variety equipped with a group structure given by morphisms of varieties. Each abelian variety A of dimension g admits a dual abelian variety \hat{A}, which parametrizes the degree-zero Picard group \mathrm{Pic}^0(A) consisting of translation-invariant invertible sheaves on A. The Picard variety of A is the component of the identity in the Picard scheme, often isomorphic to \hat{A}, and encodes the algebraic structure of line bundles on A. For a projective curve C of g \geq 1 over k, the \mathrm{Jac}(C) is the of dimension g parametrizing the degree-zero line bundles on C, with a principal . The group law on \mathrm{Jac}(C) can be realized using divisors: ample line bundles correspond to effective theta divisors \Theta_L, and the addition of points translates via the relation \Theta_{L \otimes M} = t_{a+b}^* \Theta_L + t_a^* \Theta_M - 2\Theta_L for points a, b representing L, M, up to linear equivalence. Isogenies between abelian varieties are surjective group homomorphisms with finite kernels, preserving the group structure and having degree a power of the , such as \deg() = n^{2g} for the multiplication-by-n . For a prime \ell \neq \mathrm{char}(k), the \ell-adic Tate module T_\ell A = \varinjlim A[\ell^n](k^{\mathrm{sep}}) is a \mathbb{Z}_\ell- of $2g, carrying a continuous representation of the \mathrm{Gal}(\bar{k}/k) that encodes p-adic and arithmetic properties of A.

References

  1. [1]
    [PDF] Algebraic Groups - James Milne
    Dec 20, 2015 · An algebraic group is a matrix group defined by polynomial conditions. More abstractly, it is a group scheme of finite type over a field.<|control11|><|separator|>
  2. [2]
    [PDF] Linear Algebraic Groups: a Crash Course - OSU Math
    Jan 24, 2011 · Definition 1.2 A linear algebraic group is an affine variety that is an algebraic group. Example 1.3 The multiplicative group Gm = k ...
  3. [3]
    [PDF] a Linear Algebraic Group? - Skip Garibaldi
    Definitions. A linear algebraic group over a field F is a smooth affine variety over F that is also a group, much like a topological group is a topological ...
  4. [4]
    [PDF] Humphreys-Linear-algebraic-groups-1998.pdf - CSE, IIT Bombay
    Humphreys. Page 11. Page 12. Preface to the First Printing. Over the last two decades the Borel-Chevalley theory of linear algebraic groups (as further ...Missing: original | Show results with:original<|control11|><|separator|>
  5. [5]
    [PDF] Lie Algebras, Algebraic Groups, and Lie Groups - James Milne
    May 5, 2013 · Lie algebras are an essential tool in studying both algebraic groups and Lie groups. Chapter I develops the basic theory of Lie algebras, ...
  6. [6]
    [PDF] Structure Theory of Reductive Groups through Examples
    Dec 13, 2011 · A systematic study of linear algebraic groups was initiated by C. Chevalley and A. Borel in the mid. 1950s, a few decades after the ...
  7. [7]
    [PDF] A Centennial: Wilhelm Killing and the Exceptional Groups
    While Lie was motivated by applications to differential equations, Killing was led to his problem from his work in geometry. Let ~ be a simple Lie algebra over ...
  8. [8]
    Essays in the history of Lie groups and algebraic groups, by Armand ...
    Feb 12, 2003 · Early workers were aware of global examples (such as SLn(C)), but the theory was mostly developed for what we today would call “germs” of Lie.Missing: continuous | Show results with:continuous
  9. [9]
    None
    ### Summary of Historical Development of Lie Theory (19th to Mid-20th Century)
  10. [10]
    New frontiers in Langlands reciprocity | EMS Magazine
    In this survey, I discuss some recent developments at the crossroads of arithmetic geometry and the Langlands programme.
  11. [11]
    [PDF] Lectures on Linear Algebraic Groups
    We work over an algebraically closed base field throughout. Linear algebraic groups over more general base fields are best treated using the theory of group ...
  12. [12]
    [PDF] Linear Algebraic Groups I (Stanford, Winter 2010)
    Feb 14, 2010 · A group variety G over k is called linear algebraic if it is affine. Remark 1.1.6. If G is an algebraic k-group scheme, then one can show that ...
  13. [13]
    None
    Below is a merged summary of the segments from T.A. Springer's *Linear Algebraic Groups* (1998), consolidating all information into a single, comprehensive response. To retain the maximum amount of detail efficiently, I will use a structured format with tables where appropriate, followed by a narrative summary for additional context. The response avoids redundancy while preserving all key points, definitions, theorems, examples, and URLs from the provided summaries.
  14. [14]
    [PDF] Algebraic Groups Lecture Notes - UW Math Department
    Dec 1, 2014 · approach (over an algebraically closed field k, say) where we have algebraic maps giving the group operations S × S → S, σ: S → S, e: e → S.<|control11|><|separator|>
  15. [15]
    [PDF] REDUCTIVE GROUP SCHEMES Brian Conrad
    Jun 1, 2011 · Introduction. These notes present the theory of reductive group schemes, simplifying the original proofs via tools developed after 1963 (see ...
  16. [16]
    [PDF] Algebraic Groups
    The first eight chapters of the book study general algebraic group schemes over a field. They culminate in a proof of the Barsotti–Chevalley theorem realizing.
  17. [17]
    [PDF] Classical Groups, Representations, and Invariants
    Part 1: Linear Algebraic Groups. Lecture 1. Classical Groups and Linear Algebraic Groups. Definition of a Linear Algebraic Group. Let GL(n, C) be the group of ...
  18. [18]
    [PDF] linear algebraic groups
    However, for non-perfect fields, the absolute unipotent radical could be larger; for example, consider a (non- trivial) finite, purely inseparable extension k0/ ...
  19. [19]
    [PDF] (Linear) Algebraic Groups 1 Basic Definitions and Main Examples ...
    Let G be a connected linear algebraic group of dimension 1. Then, either G ∼= Gm or G ∼= Ga. 4 Lie Algebra (associated to algebraic groups) (Heline).
  20. [20]
    [1702.04831] Cohomology of unipotent group schemes - arXiv
    Feb 16, 2017 · The second half of this paper investigates in detail the cohomology of Frobenius kernels (U_3)_{(r)} of the Heisenberg group U_3 \subset GL_3.
  21. [21]
    [PDF] UNIPOTENT ALGEBRAIC GROUPS - Tufts University
    We call a linear algebraic group unipotent if every element is unipotent. ... In particular, it is a subgroup of the ordinary Heisenberg group of dimension 2m + 3 ...
  22. [22]
    [PDF] Algebraic Groups, Lie Groups, and their Arithmetic Subgroups
    Apr 1, 2011 · This work is a modern exposition of the theory of algebraic groups (affine group schemes),. Lie groups, and their arithmetic subgroups. BibTeX ...
  23. [23]
    [PDF] e8, the most exceptional group - Skip Garibaldi
    The five exceptional simple Lie algebras over the complex number are included one within the other as g2 ⊂ f4 ⊂ e6 ⊂ e7 ⊂ e8. The biggest one, e8, is in many ...
  24. [24]
    [PDF] Algebraic Groups and Number Theory - OpenScholar
    ... Borel ... This line of research emerged fairly recently as an independent area of mathematics, often called the arithmetic theory of (linear) algebraic groups.
  25. [25]
    [PDF] Linear Algebraic Groups
    Algebraic groups. Let K be an algebraically closed field. An algebraic K-group G is an algebraic variety over K, and a group, such that the maps µ : G × G ...
  26. [26]
    [PDF] Reductive Groups
    In the 1950s Chevalley became interested in algebraic groups as a link between complex. Lie algebras and finite groups. In a fundamental paper, Chevalley (1955) ...
  27. [27]
    [PDF] Linear Algebraic Groups Florian Herzig
    Definition 38. G is unipotent if G = Gu. Example. Un is unipotent, and so is Ga (as Ga. ∼. = U2) ...
  28. [28]
    [PDF] LINEAR ALGEBRAIC GROUPS
    Linear algebraic groups are matrix groups defined by polynomials; a typi- cal example is the group SLn of matrices of determinant one. The theory of algebraic ...
  29. [29]
    [PDF] Algebraic Groups I. Existence of Jordan decomposition
    In this handout, we address the key issue in the proof of existence of Jordan decomposition: if k is algebraically closed and G is a smooth closed subgroup of ...<|control11|><|separator|>
  30. [30]
    [PDF] Algebraic Groups: Part V - McGill University
    It explains the special role played by tori and Borel subgroups in the study of linear algebraic groups. ... (2) Every semisimple element of G lies in a maximal ...
  31. [31]
    [PDF] basics on reductive groups - Yale Math
    Borel subgroups. Definition 2.1. A subgroup B ⊂ G is called a Borel subgroup if it is maximal. (w.r.t. inclusion) of all connected solvable subgroups ...<|control11|><|separator|>
  32. [32]
    [PDF] BRUHAT DECOMPOSITION AND APPLICATIONS G. Lusztig
    Jun 22, 2010 · Let k be a field. We assume that k is algebraically closed, unless otherwise specified. Let G be a connected reductive algebraic group over k.
  33. [33]
    [PDF] GENERAL PARABOLIC SUBGROUPS OF GLn(C)
    We now define Borel subgroup. Definition 1.1. A subgroup B of a group G is called Borel if it is a maximal solvable connected algebraic subgroup of G.
  34. [34]
    [PDF] LEVI DECOMPOSITIONS OF A LINEAR ALGEBRAIC GROUP
    Finally, if H is reductive and P is a parabolic subgroup, one knows for any linear representation V of H that Hi(H, V) ≃ Hi(P, V) for i ≥ 0. Using this fact, we ...
  35. [35]
    [PDF] 1 Let G be a complex semisimple linear algebraic group, with a ...
    opposite Borel subgroup B− is w0Bw0. Its Lie algebra is b. − = h ⊕ M β∈R− gβ. 2. Let X = G/B; this is a smooth variety of dimension N. The T-fixed points ...
  36. [36]
    [PDF] Some Transitive Linear Actions of Real Simple Lie Groups
    Key Words and Phrases: Transitive linear action, reductive group, actions of ... semisimple and so G = Z(G)0 · [G, G], where Z(G)0 is the connected component.<|control11|><|separator|>
  37. [37]
    Linear Algebraic Groups - SpringerLink
    This book presents foundational material on algebraic groups, Lie algebras, transformation spaces, and quotient spaces.
  38. [38]
    [PDF] Semisimple Lie Algebras, Algebraic Groups, and Tensor Categories
    May 9, 2007 · Borel, A. 1991. Linear algebraic groups, volume 126 of Graduate Texts in Mathematics. Springer-Verlag, New York. Bourbaki, N. Lie. Groupes ...<|control11|><|separator|>
  39. [39]
    [PDF] Classification of root systems
    Sep 8, 2017 · In this section we recall the construction of a root system from a semisimple. Lie algebra given in Chapter 6 of the text. We will also state ...
  40. [40]
    [PDF] Lie Algebras
    Apr 23, 2004 · In particular, The. Dynkin diagrams characterize all possible irreducible root systems. Proof. Since ∆ is a basis of E and ∆0 is a basis of ...
  41. [41]
    [PDF] Lecture 17 — Cartan Matrices and Dynkin Diagrams
    Nov 4, 2010 · We'll see that to every root system there corresponds a semisimple Lie algebra, so it's important to know all the root systems. Last time we ...Missing: groups | Show results with:groups
  42. [42]
    Introduction to Lie Algebras and Representation Theory - SpringerLink
    In stockThis book is designed to introduce the reader to the theory of semisimple Lie algebras over an algebraically closed field of characteristic 0, with emphasis on ...
  43. [43]
    Sur certains groupes simples - Project Euclid
    1955 Sur certains groupes simples. C. Chevalley · DOWNLOAD PDF + SAVE TO MY LIBRARY. Tohoku Math. J. (2) 7(1-2): 14-66 (1955). DOI: 10.2748/tmj/1178245104.
  44. [44]
    SCHUR MULTIPLIERS OF THE KNOWN FINITE SIMPLE GROUPS1 ...
    In [9] Steinberg shows that, for Chevalley groups, if the cardinality of the field is large enough, the p-part of the multiplier is trivial. The same holds for ...
  45. [45]
    Variations on a theme of Chevalley. - Project Euclid
    1959 Variations on a theme of Chevalley. Robert Steinberg · DOWNLOAD PDF + SAVE TO MY LIBRARY. Pacific J. Math. 9(3): 875-891 (1959).
  46. [46]
    [PDF] Reductive Groups over Fields - Stanford Mathematics
    Jan 19, 2020 · For a field k, a linear algebraic group over k is a smooth affine k-group scheme (equivalently, a smooth closed k-subgroup of GLn ). Remark 1.1.
  47. [47]
    [PDF] Borel-Weil-Bott theorem and geometry of Schubert varieties
    Let G be a semisimple, simply-connected, complex algebraic group with a fixed Borel subgroup B, a maximal torus H ⊂ B, and associated Weyl group W. (Recall that ...
  48. [48]
    [0706.3251] On tensor products of polynomial representations - arXiv
    Jun 22, 2007 · ... representations of GL(n,\mathbb{C}) is isomorphic to another. As a consequence we discover families of Littlewood-Richardson coefficients ...
  49. [49]
    [PDF] The Steinberg Tensor Product Theorem for GL(m|n) - OU Math
    The Steinberg Tensor Product Theorem is a fundamental result in the modular representation theory of algebraic groups. The purpose of the present article is to ...
  50. [50]
    Geometric Invariant Theory - SpringerLink
    $$199.99 Free delivery 14-day returnsOct 29, 2012 · “Geometric Invariant Theory” by Mumford/Fogarty (the first edition was published in 1965, a second, enlarged edition appeared in 1982) is ...
  51. [51]
    Reductive groups are geometrically reductive - Annals of Mathematics
    Reductive groups are geometrically reductive. Pages 67-83 from Volume 102 (1975), Issue 1 by William J. Haboush ... Authors. William J. Haboush. ← Previous ...
  52. [52]
    [PDF] Structure Theory of Semisimple Lie Groups
    This article provides a review of the elementary theory of semisimple Lie al- gebras and Lie groups. It is essentially a summary of much of [K3]. The four.
  53. [53]
    [PDF] Modular Functions and Modular Forms
    It is easy to define modular functions and forms, but less easy to say why they are important, especially to number theorists. Thus I will begin with a rather ...
  54. [54]
    AN ELEMENTARY INTRODUCTION TO THE LANGLANDS ...
    This program emerged around 1967 in the form of a series of conjectures, and it has subsequently influenced recent research in number theory in much the same ...
  55. [55]
    None
    ### Summary of p-Adic Groups, Bruhat-Tits Buildings, and Local Langlands Correspondence
  56. [56]
    The development version of the CHEVIE package of GAP3
    I describe the current state of the development version of the CHEVIE package, which allows to transform the theories of Coxeter groups, reductive algebraic ...
  57. [57]
    Root Systems and Root Data - Magma Computational Algebra System
    Sep 26, 2025 · Root systems are used to classify the semisimple Lie algebras. The closely related concept of a root datum is used to classify the groups of Lie ...Missing: linear | Show results with:linear
  58. [58]
    Introduction - Magma Computational Algebra System
    This chapter describes Magma functions for computing with finite real root systems. A root system describes the reflections in a reflection group (Chapter ...Missing: software | Show results with:software
  59. [59]
    Effective black-box constructive recognition of classical groups
    We describe a black-box Las Vegas algorithm to construct standard generators for classical groups defined over finite fields. We assume that the field has ...
  60. [60]
    [PDF] Black box classical groups - University of Oregon
    If a black box simple group is known to be isomorphic to a classical group over a field of known characteristic, a Las Vegas algorithm is used to produce.
  61. [61]
    [PDF] Black-box recognition of finite simple groups of Lie type by statistics ...
    Given a black-box group G isomorphic to some finite simple group of Lie type and the characteristic of G, we compute the standard name of G by a Monte.
  62. [62]
    General framework for E(3)-equivariant neural network ... - Nature
    May 18, 2023 · Here we propose an E(3)-equivariant deep-learning framework to represent density functional theory (DFT) Hamiltonian as a function of material structure.Results · Equivariance Of The... · The Neural Network...Missing: post- | Show results with:post-
  63. [63]
    [PDF] Clifford Group Equivariant Neural Networks
    We introduce Clifford Group Equivariant Neural Networks: a novel approach for constructing O(n)- and E(n)-equivariant models. We identify and study the.
  64. [64]
    Theoretical Aspects of Group Equivariant Neural Networks - arXiv
    Apr 10, 2020 · Group equivariant neural networks have been explored in the past few years and are interesting from theoretical and practical standpoints.Missing: post- | Show results with:post-
  65. [65]
    A Lie algebraic theory of barren plateaus for deep parameterized ...
    Aug 22, 2024 · In this work, we present a general Lie algebraic theory for BPs that can be applied to any deep, unitary, parametrized quantum circuit ...
  66. [66]
    Characterization of variational quantum algorithms using free fermions
    Mar 30, 2023 · We study variational quantum algorithms from the perspective of free fermions. By deriving the explicit structure of the associated Lie algebras.
  67. [67]
    Lie groups for quantum complexity and barren plateau theory - arXiv
    Jul 30, 2025 · ... quantum algorithms. In this review, we introduce the theory of Lie groups and their algebras to analyze two fundamental problems in quantum ...
  68. [68]
    [PDF] Introduction to Symmetry and Broken Symmetry in Physics - HAL
    Mar 13, 2025 · In particle physics this is the Higgs mechanism, in condensed matter its non-relativistic version is known as the Anderson-Higgs model. ...Missing: reductive | Show results with:reductive
  69. [69]
    [PDF] Symmetries and Group Theory in Particle Physics
    ... symmetry, now called flavor SU(3) invariance, which allowed the inclusion of ... breaking, which we illustrate in the frame of a couple of simple ...
  70. [70]
    [PDF] Lie-group methods - HAL
    Jun 8, 2016 · Definition 2.13 A real matrix Lie group is a smooth subset G ⊆ Rn×n, closed ... is more stable when numerical methods are implemented with small ...
  71. [71]
    [PDF] computational-aspects-of-lie-groups-algorithms-and-applications-in ...
    Jan 30, 2025 · In quantum information science, efficient algorithms for computing Lie group exponentials help in designing quantum gates and simulating ...
  72. [72]
    A method of Lie-symmetry GL(n,R) for solving non-linear dynamical ...
    The Lie-group numerical method has a stabilizing effect to retain the solution on the associated manifold, which thus naturally has a regularization effect to ...
  73. [73]
    [PDF] Lie Groups and Algebraic Groups - UCSD Math
    We show how to put a Lie group structure on a closed subgroup of the general linear group and determine the Lie algebras of the classical groups. We develop the ...
  74. [74]
    [PDF] Author(s): Armand Borel and Harish—Chandra
    The following proposition strengthens 1.7 , and extends to reductive real algebraic groups well-known properties of semi-simple Lie groups. 1.10. PROPOSITION.
  75. [75]
    [PDF] Lie Groups. Representation Theory and Symmetric Spaces
    In this Chapter we discuss elementary properties of Lie groups, Lie algebras and their relationship. We will assume a good knowledge of manifolds, vector.
  76. [76]
    [PDF] Basic Theory of Affine Group Schemes - James Milne
    Mar 11, 2012 · This is a modern exposition of the basic theory of affine group schemes. Although the emphasis is on affine group schemes of finite type ...
  77. [77]
    Introduction to Affine Group Schemes - SpringerLink
    These turn out to be equivalent to the ring-theoretic objects called Hopf algebras, with which we can then con struct new examples. Study of their ...
  78. [78]
    [PDF] Tannakian Categories
    May 13, 2011 · The functor A. SpecA defines an equivalence of categories between the category of k-bialgebras and the category of affine group schemes over k.
  79. [79]
    [PDF] Abelian Varieties - James Milne
    Mar 16, 2008 · These notes are an introduction to the theory of abelian varieties, including the arithmetic of abelian varieties and Faltings's proof of ...