Fact-checked by Grok 2 weeks ago

General linear group

In , the general linear group of degree n is the group of n × n invertible matrices with entries in a given K (or more generally, a with identity), under the operation of . This group, denoted \mathrm{GL}_n(K) or \mathrm{GL}(n, K), consists of all such matrices with nonzero and forms a basic example of a when K is or numbers. It plays a central role in linear algebra, , and .

Definition and basic properties

Vector space formulation

The general linear group of a V, denoted \mathrm{GL}(V), consists of all invertible linear transformations from V to itself, which are also known as the automorphisms of V. These transformations preserve the vector space structure, mapping V bijectively onto itself while maintaining addition and scalar multiplication. The group operation is defined by of these linear maps, where the composition of two invertible transformations is again invertible and linear. For a finite-dimensional V of n \geq 1, \mathrm{GL}(V) is non-abelian when n \geq 2, meaning that the composition of transformations does not generally commute. The of the group is the map on V, which leaves every unchanged, and every element T \in \mathrm{GL}(V) has an T^{-1} that is also a linear satisfying T \circ T^{-1} = T^{-1} \circ T = . When V is a of n over a K, the group is commonly denoted \mathrm{GL}(n, K). This notation emphasizes the dependence on both the and the underlying , and it abstracts the structure away from specific bases. In this setting, matrix representations can be used to coordinatize elements of \mathrm{GL}(n, K) once a basis is chosen, though the group itself is defined independently of coordinates. To verify that \mathrm{GL}(V) forms a group, note that it is closed under : if T_1, T_2 \in \mathrm{GL}(V), then T_1 \circ T_2 is invertible with T_2^{-1} \circ T_1^{-1}. Associativity follows directly from the associativity of on the set of all maps from V to V. The identity map serves as the neutral element, and inverses exist by definition for each element. As an algebraic variety over an , \mathrm{GL}(n, K) has n^2, reflecting its as an open subset of the of all n \times n matrices.

Matrix representation and invertibility

The general linear group \mathrm{GL}(V) of a finite-dimensional V of n over a K is to the matrix group \mathrm{GL}(n, K), consisting of all n \times n invertible matrices with entries in K. This isomorphism arises from choosing a basis for V, which allows linear endomorphisms of V to be represented by matrices in M_n(K), the ring of all n \times n matrices over K; the invertible ones correspond precisely to automorphisms of V. The group operation in \mathrm{GL}(n, K) is , which mirrors the of linear maps: for matrices A, B \in \mathrm{GL}(n, K), the product AB represents the of the corresponding automorphisms and is also invertible. A matrix A \in M_n(K) belongs to \mathrm{GL}(n, K) it admits a two-sided A^{-1} \in M_n(K) such that A A^{-1} = A^{-1} A = I_n, the ; this condition is equivalent to the linear map represented by A being bijective. Matrices without inverses, such as the or any singular matrix with linearly dependent rows or columns, are excluded from \mathrm{GL}(n, K). For n=1, \mathrm{GL}(1, K) is isomorphic to the multiplicative group K^\times of nonzero elements in K, since 1×1 invertible matrices are simply the nonzero scalars. Over the real numbers \mathbb{R}, the group \mathrm{GL}(2, \mathbb{R}) includes matrices representing rotations (like those in the special orthogonal subgroup) combined with scalings, such as \begin{pmatrix} a & -b \\ b & a \end{pmatrix} for a^2 + b^2 \neq 0, which scale by \sqrt{a^2 + b^2} and rotate by \tan^{-1}(b/a).

Determinant characterization

The general linear group \mathrm{GL}(n, K) consists of all n \times n matrices A with entries in the field K such that \det(A) \neq 0. The non-vanishing of the is the precise condition for invertibility over K, as a matrix is invertible its is non-zero. This characterization provides an explicit algebraic test for membership in the group and is independent of the choice of basis.

Generalizations over rings and fields

Over finite fields

The general linear group over a \mathbb{F}_q with q elements, denoted GL(n, q), consists of all invertible n \times n matrices with entries in \mathbb{F}_q. The of this group is given by the formula |GL(n, q)| = \prod_{k=0}^{n-1} (q^n - q^k). This product arises from counting the number of ordered bases for the \mathbb{F}_q^n, where the first basis vector can be any of the q^n - 1 nonzero vectors, the second any vector not in the span of the first (totaling q^n - q choices), and so on for subsequent vectors. This combinatorial interpretation aligns with the matrix perspective: each invertible matrix corresponds to a linear transformation that maps the standard basis to another ordered basis of \mathbb{F}_q^n. For small cases, explicit computations yield |GL(2, 2)| = 6, and this group is isomorphic to the symmetric group S_3 via its action permuting the three nonzero vectors in \mathbb{F}_2^2. Similarly, |GL(2, 3)| = 48. The group GL(n, q) acts on \mathbb{F}_q^n by , and this action is transitive on the set of nonzero vectors: for any two nonzero vectors u, v \in \mathbb{F}_q^n, there exists g \in GL(n, q) such that gu = v, as one can extend \{u\} to a basis and map it to a basis containing v. The action also induces on the set of lines (1-dimensional subspaces) in \mathbb{F}_q^n, with the \mathbb{P}^{n-1}(\mathbb{F}_q) serving as the space. In , elements of GL(n, q) generate linear error-correcting codes through their action on code subspaces or via sections, providing constructions with controlled minimum distance for reliable data transmission over noisy channels. For instance, the in the variety of GL(n, q) yield codes whose parameters relate directly to the group's order and sizes.

Over arbitrary s

The general linear group over an arbitrary R with identity, denoted \mathrm{GL}_n(R), consists of all n \times n matrices with entries in R whose determinants lie in the of units R^\times, under the operation of . This definition captures the automorphisms of the free R- R^n, as a matrix A \in M_n(R) with \det(A) \in R^\times admits an in M_n(R), ensuring it induces a . Over such rings, the determinant condition generalizes the field case but introduces subtleties due to the potential lack of unique factorization or algorithms in R. For principal ideal domains like the ring of integers \mathbb{Z}, the group \mathrm{GL}_n(\mathbb{Z}) comprises precisely those integer matrices with determinant \pm 1. The units of \mathbb{Z} are \{ \pm 1 \}, so this condition ensures invertibility over \mathbb{Z}. The kernel of the determinant map \det: \mathrm{GL}_n(\mathbb{Z}) \to \{ \pm 1 \} is the special linear group \mathrm{SL}_n(\mathbb{Z}), consisting of matrices with determinant exactly 1, which plays a central role in arithmetic geometry and modular forms. The structure of \mathrm{GL}_n(R) depends profoundly on the Bass stable rank of R, defined as the smallest d such that every unimodular row of greater than d over R reduces to a shorter unimodular row via elementary operations. For a R with \mathrm{sr}(R) = d, it holds that \mathrm{GL}_n(R) = \mathrm{E}_n(R) for all n > d, where \mathrm{E}_n(R) is the generated by all elementary matrices (those obtained by adding multiples of one row/column to another). This equality reflects stabilization in algebraic , as \mathrm{E}_n(R) exhausts the stable general linear group \mathrm{GL}(R) = \varinjlim_n \mathrm{GL}_n(R), and the stable rank quantifies how quickly this stabilization occurs. Rings like domains have stable rank 2, while rings in sufficiently many variables can have higher ranks, complicating the generation of \mathrm{GL}_n(R) for small n. A representative example arises over polynomial rings R = k, where k is a . The units k^\times are the nonzero constant polynomials, isomorphic to k^\times, so \mathrm{GL}_1(k) \cong k^\times. For n \geq 2, \mathrm{GL}_n(k) includes matrices with constant nonzero determinants; since k has stable rank 2, the elementary subgroup E_n(k) equals SL_n(k) for n ≥ 2, and thus generates the special linear part. Suslin's stability theorem confirms that this holds more generally for polynomial rings in multiple variables, with generation by elementary matrices for sufficiently large n. Unlike over fields, where all finitely generated projective modules are free, over general commutative rings there can exist non-free finitely generated projective modules (e.g., over Dedekind domains that are not domains, such as the in quadratic number fields with class number greater than 1, where non-principal ideals give rank-1 projectives). The group GL_n(R) describes the s of the R^n, while the endomorphism rings and groups of non-free projectives are studied using more advanced tools in algebraic .

Historical development

The concept of the general linear group emerged in the mid-19th century alongside the development of theory. In 1858, published his seminal memoir on the theory of matrices, where he formalized the operations of , multiplication, and the role of the in characterizing nonsingular matrices, thereby laying the groundwork for the group of invertible linear transformations. and collaborated closely on , and the term "general linear group" emerged in the late to describe the collection of all invertible linear substitutions acting on vector spaces. By the late 19th century, attention turned to linear groups over finite fields. Eliakim Hastings Moore, in the 1890s, explored the structure of finite vector spaces and the actions of linear transformations upon them, establishing foundational results on the of such groups and linking them to Galois fields. Building on this, Leonard Eugene Dickson in 1901 provided the first systematic treatment of linear groups over arbitrary fields in his monograph Linear Groups, with an Exposition of the Galois Field Theory, where he derived the cardinality of GL(n, q) as the product ∏_{k=0}^{n-1} (q^n - q^k). Parallel developments in arose through Sophus Lie's work in the late , who conceptualized the general linear group GL(n, ℝ) as a continuous group acting on , integrating it into his theory of differential equations and symmetries. In the early 1900s, advanced this framework by developing the associated 𝔤𝔩(n), the tangent space at the comprising all n × n matrices under the bracket, which facilitated the classification of semisimple Lie algebras. The mid-20th century saw the abstraction of the general linear group within modern algebra. In the 1930s, reformulated using the action of Galois groups on vector spaces, incorporating GL(n, K) as the group of automorphisms of separable extensions, thereby embedding linear groups into the study of field extensions without reliance on primitive elements. , in the 1940s as part of the Bourbaki collective, generalized the structure of GL(n) to division rings via the Dieudonné determinant, establishing it as a foundational object in the theory of algebraic groups over arbitrary fields. In the postwar era, applications proliferated across number theory and topology. Robert Langlands initiated his program in the 1960s, positing deep correspondences between n-dimensional Galois representations and automorphic representations of GL(n, ℚ\ℝ × ∏ ℚ_p), unifying disparate areas like class field theory and modular forms. Concurrently, in the 1970s, Daniel Quillen connected GL(n, k) to stable homotopy theory through algebraic K-theory, showing that the higher K-groups K_*(k) are the homotopy groups of the classifying space BGL(k)^+, linking linear algebra to the stable stems of spheres.

Key subgroups

Special linear group

The special linear group \mathrm{SL}(n, K), where K is a , is defined as the of the \det: \mathrm{GL}(n, K) \to K^\times, consisting of all n \times n invertible matrices over K with equal to 1. This makes \mathrm{SL}(n, K) a of \mathrm{GL}(n, K), and since the map is surjective, the of \mathrm{SL}(n, K) in \mathrm{GL}(n, K) equals the of K^\times, which is infinite when K is an infinite . For n \geq 2 and fields K with more than three elements, \mathrm{SL}(n, K) is a , meaning it equals its own [ \mathrm{SL}(n, K), \mathrm{SL}(n, K) ]. This property highlights its simple structure in group-theoretic terms, excluding small exceptional cases like \mathrm{SL}(2, \mathbb{F}_2) and \mathrm{SL}(2, \mathbb{F}_3). Moreover, \mathrm{SL}(n, K) is generated by elementary transvections, specifically the matrices E_{ij}(\lambda) = I + \lambda e_{ij} for i \neq j and \lambda \in K, where e_{ij} denotes the standard matrix unit with a 1 in the (i,j)-entry and zeros elsewhere. These generators reflect the group's close relation to the structure of vector spaces over K. Notable examples include \mathrm{SL}(2, \mathbb{R}), which acts on the upper half-plane \mathcal{H} = \{ z \in \mathbb{C} \mid \operatorname{Im}(z) > 0 \} via Möbius transformations z \mapsto \frac{az + b}{cz + d} for \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \mathrm{SL}(2, \mathbb{R}). This action preserves the hyperbolic metric and underlies much of the geometry of the hyperbolic plane. Similarly, \mathrm{SL}(2, \mathbb{Z}) is known as the modular group, acting on \mathcal{H} with a fundamental domain given by the region \{ z \in \mathcal{H} \mid |\operatorname{Re}(z)| \leq 1/2, |z| \geq 1 \}, which tiles \mathcal{H} under the group action. The associated Lie algebra \mathfrak{sl}(n, K) comprises all n \times n matrices X over K with \operatorname{tr}(X) = 0, forming a Lie subalgebra of \mathfrak{gl}(n, K) under the commutator bracket. On \mathfrak{sl}(n, K), the trace form \operatorname{tr}(XY) serves as an analog to the Killing form, being proportional to it via the relation B(X,Y) = 2n \operatorname{tr}(XY), where B is the Killing form; this non-degenerate bilinear form underscores the semisimple nature of \mathfrak{sl}(n, K) for n \geq 2.

Diagonal and unitary subgroups

The diagonal subgroup D(n, K) of the general linear group \mathrm{GL}(n, K) over a K consists of all invertible diagonal matrices, i.e., those with nonzero entries on the diagonal. This is isomorphic to the (K^\times)^n, where K^\times is the of the , reflecting the independence of the diagonal entries. It serves as a in \mathrm{GL}(n, K), meaning it is a (a connected abelian algebraic group) that is maximal among such subgroups, and all maximal tori in \mathrm{GL}(n, K) are conjugate under the . Over the complex numbers, the unitary group U(n) is the compact subgroup of \mathrm{GL}(n, \mathbb{C}) defined by U(n) = \{ A \in \mathrm{GL}(n, \mathbb{C}) \mid A^* A = I \}, where A^* denotes the conjugate transpose of A, and I is the identity matrix. This condition ensures that elements of U(n) preserve the standard Hermitian inner product on \mathbb{C}^n. As a Lie group, U(n) is compact and has real dimension n^2. Similarly, over the reals, the orthogonal group O(n) is the compact subgroup of \mathrm{GL}(n, \mathbb{R}) given by O(n) = \{ A \in \mathrm{GL}(n, \mathbb{R}) \mid A^T A = I \}, where A^T is the transpose; it preserves the standard Euclidean inner product on \mathbb{R}^n. The group O(n) has two connected components, distinguished by the determinant (\det A = \pm 1), with the special orthogonal group \mathrm{SO}(n) comprising the component where \det A = 1; both have Lie algebra dimension n(n-1)/2. The Cartan decomposition provides a key structural insight into \mathrm{GL}(n, \mathbb{R}), expressing it as \mathrm{GL}(n, \mathbb{R}) = [O(n)](/page/Orthogonal_group) A [O(n)](/page/Orthogonal_group), where A is the of positive diagonal matrices (a Cartan subgroup). This arises from the polar form of matrices and highlights the role of the as a maximal compact . Associated with the maximal torus (such as the diagonal subgroup), the Weyl group of \mathrm{GL}(n, K) is the quotient of the normalizer of the torus by the torus itself, yielding the symmetric group S_n. It acts on the torus by permuting the diagonal entries, generated by reflections corresponding to adjacent transpositions, and is realized via monomial matrices (permutation matrices with nonzero entries on the permuted positions).

Projective linear group

The projective linear group \mathrm{PGL}(n, R), also known as the projective general linear group, is the of the general linear group \mathrm{GL}(n, R) by its center, the scalar matrices R^\times I_n. This captures the action of \mathrm{GL}(n, R) on , where scalar multiples act trivially.

Affine and semilinear groups

The affine group \mathrm{Aff}(n, k), or general affine group over a k, is the \mathrm{GL}(n, k) \ltimes k^n, combining linear transformations with translations. It consists of all invertible affine transformations x \mapsto Ax + b where A \in \mathrm{GL}(n, k) and b \in k^n. The general semilinear group \Gamma \mathrm{L}(n, k) extends \mathrm{GL}(n, k) by incorporating field automorphisms, forming the \mathrm{GL}(n, k) \rtimes \mathrm{Aut}(k). Its elements are semilinear transformations T(v) = \sigma(Av) for \sigma \in \mathrm{Aut}(k), A \in \mathrm{GL}(n, k).

Infinite-dimensional general linear group

The infinite-dimensional general linear group arises in two primary settings: the algebraic context over vector spaces and the analytic context over Hilbert spaces. In the analytic setting, for a complex Hilbert space H, the group \mathrm{GL}(H) consists of all bounded linear operators on H that admit bounded inverses, forming a group under composition. This group is equipped with the operator norm topology inherited from the Banach algebra B(H) of all bounded operators on H. \mathrm{GL}(H) is an open subset of B(H), since for any invertible T \in B(H), the set of S \in B(H) satisfying \|T^{-1}(T - S)\| < 1 consists of invertible operators, with inverses given by the convergent Neumann series. The Calkin algebra, defined as the quotient B(H)/K(H) where K(H) is the ideal of compact operators, has \mathrm{GL}(H) mapping onto its group of invertible elements; the preimage under this quotient map yields the Fredholm operators, whose index is analyzed via the Atiyah–Singer index theorem in geometric applications, associating the index to topological invariants like the Euler characteristic. Algebraically, for an infinite-dimensional V over a k, \mathrm{GL}(V) denotes the group of all invertible linear endomorphisms of V. However, due to the lack of finite bases, stability considerations lead to the infinite general linear group \mathrm{GL}(\infty, k) = \varinjlim_n \mathrm{GL}(n, k), the direct (inductive) limit under the stabilization embeddings \mathrm{GL}(n, k) \hookrightarrow \mathrm{GL}(n+1, k) that pad matrices with a $1 in the bottom-right corner. This colimit embeds all finite-dimensional \mathrm{GL}(n, k) densely and plays a central role in algebraic , where K_1(k) = \mathrm{GL}(\infty, k)^{\mathrm{ab}} is the abelianization. Unlike its finite-dimensional counterparts, \mathrm{GL}(H) for infinite-dimensional separable H is not a in the classical finite-dimensional sense, though it forms an open of the Banach Lie group B(H). A key is its contractibility in the norm topology, established by Kuiper's theorem, implying that \mathrm{GL}(H) is simply connected and homotopy-equivalent to a point. This contrasts with the finite-dimensional case, where \mathrm{GL}(n, \mathbb{C}) has nontrivial groups. Examples illustrate these structures: for the separable H = \ell^2(\mathbb{Z}), the bilateral S e_n = e_{n+1} (where \{e_n\} is the ) is unitary, hence lies in \mathrm{GL}(H) with norm $1. The unilateral shift on H = \ell^2(\mathbb{N}) is a in the preimage of the invertibles in the Calkin algebra, with Fredholm -1, demonstrating the to ; more generally, the additivity of the under reflects the group near the component.

Linear monoids

The full linear monoid over a R, denoted M_n(R), consists of all n \times n matrices with entries in R, equipped with the operation of , which is associative and has the as the unit element. This structure forms a , and the general linear group GL(n, R) arises as the submonoid of invertible (unit) elements within it. In contrast to the group of invertibles, M_n(R) includes non-invertible matrices, making it a richer algebraic object that captures all linear endomorphisms of the free R-module R^n. As a semigroup under multiplication, M_n(R) exhibits notable structural features beyond those of groups. Idempotents in M_n(R) are matrices E satisfying E^2 = E, which, over integral domains, correspond to projections onto direct summands of R^n. Zero divisors abound, consisting of non-zero matrices A and B such that AB = 0, reflecting the presence of non-trivial kernels and cokernels in the associated linear maps. A key is the of a , defined as the minimal number of generators of the submodule, which satisfies \operatorname{rank}(AB) \leq \min(\operatorname{rank}(A), \operatorname{rank}(B)) and provides a partial on idempotents via comparability of their images. When R = K is a , M_n(K) is isomorphic to the ring of K-linear endomorphisms \operatorname{End}_K(K^n) under composition, forming a simple Artinian . By the Wedderburn-Artin theorem, its unique decomposition as a semisimple is M_n(K) \cong M_n(K) itself, highlighting its indecomposability into smaller matrix components over the division K. In the infinite-dimensional setting, the full linear monoid analogue is B(H), the set of all bounded linear operators on a complex H, forming a under operator composition with the identity operator as unit. This structure generalizes finite-dimensional endomorphisms while incorporating norm considerations absent in the finite case. Matrix semigroups like subsemigroups of M_n(R) have applications in , where they model state transformations in weighted or rational automata to analyze language recognition and growth rates. In , they arise in the study of linear dynamical systems, aiding the determination of and via semigroup-generated trajectories.

References

  1. [1]
    [PDF] The General Linear Group Related Groups - Eastern Illinois University
    The general linear group will be considered as the group of linear transfonnations of a vector space onto itself under composition of map- pings and as the ...
  2. [2]
    [PDF] MATH 103A Modern algebra I
    non-abelian examples, such as the general linear group GL2(R) of invertible 2×2 matrices. Each matrix A ∈ GL2(R) yields a transformation of the plane R2 ...
  3. [3]
    [PDF] Linear Algebraic Groups
    The dimension of GLn(K) is n2, and it is connected. In the case n = 1, the usual notation for GL1(K) is Gm. The only connected algebraic groups of dimension 1 ...
  4. [4]
    [PDF] Some notes on linear algebra - Columbia Math Department
    We let GL(V ) be the group of isomorphisms from V to itself. If V is finite dimensional, with dimV = n, then GL(V ). ∼. = GLn(k) by choosing a basis of V .
  5. [5]
    [PDF] Chapter 2 Linear groups
    The general linear group GL(V ) is the set of invertible linear maps from V to itself. Without much loss of generality, we may take V as the vector space Fn q ...
  6. [6]
    [PDF] The General Linear Group
    Feb 18, 2005 · Definition: Let F be a field. Then the general linear group GLn(F) is the group of invert- ible n × n matrices with entries in F under matrix ...
  7. [7]
    [PDF] Matrix groups
    A matrix group, or linear group, is a group G whose elements are invertible n×n matrices over a field F. The general linear group GL(n,F) is the group.
  8. [8]
    [PDF] Chapter 1 Linear groups
    2. By GL(n, k) (where k = R or C) we mean the group of invertible n × n. matrices, ie. GL(n, k) = {T ∈ M(n, k) : detT 6= 0}
  9. [9]
    [PDF] Introduction to Representations of GL(n) - Theorem of the Day
    The general linear group is the group of all n × n non-singular matrices. Notice that this is indeed a group; it satisfies the group axioms,. • The product ...
  10. [10]
    [PDF] LECTURE II 1. General Linear Group Let Fq be a finite field of order ...
    General Linear Group. Let Fq be a finite field of order q. Then GLn(q), the general linear group over the field. Fq, is the group of invertible n × n matrices ...
  11. [11]
    [PDF] (January 14, 2009) [06.1] Given a 3-by-3 matrix M with integer ...
    Jan 14, 2009 · [06.4] Show that GL(2, F2) is isomorphic to the permutation group S3 on three letters. There are exactly 3 non-zero vectors in the space F2.Missing: S_3 | Show results with:S_3
  12. [12]
    GL(2,3) - GroupNames
    G = GL2(𝔽3) order 48 = 24·3. General linear group on 𝔽32. Order 48 #29 ... Polynomial with Galois group GL2(𝔽3) over ℚ. action, f(x), Disc(f). 8T23, x8-4x7 ...
  13. [13]
    [PDF] Action of GL(2,q) on non zero vectors over GF(q) - m-hikari.com
    Sep 12, 2020 · In this paper, we will test transitivity, determine the ranks and subdegrees of GL(2,q) acting on a set of non zero vectors over GF(q). Keywords ...
  14. [14]
    [PDF] Group Actions and Finite Fields - UC Berkeley math
    Feb 9, 2011 · Example 1 (The General Linear Group). GLn(Fq) consists of those n×n ... Let V be a vector space over Fq of dimension n, and fix an ...
  15. [15]
    [PDF] Error Correcting Codes From General Linear Groups - arXiv
    Mar 8, 2023 · ... finite fields. The main goal of our paper is to ... When q needs to be emphasized, we will use Fq instead of F. The general linear group ...
  16. [16]
    Principal series for general linear groups over finite commutative rings
    We construct, for any finite commutative ring R, a family of representations of the general linear group G L 𝑛 ⁢ ( 𝑅 ) whose intertwining properties mirror ...
  17. [17]
    On the normal structure of the general linear group over a ring
    The present paper is devoted to the study of normal subgroups of the general linear group over a ring and the centrality of the extension St (n, R) →
  18. [18]
    Stable rank of rings and dimensionality of topological spaces
    Bass, "K-theory and stable algebra," Publ. Math.,22, 5–60 (1964) ... Vasershtein, "On the stabilization of the general linear group over rings," Matem.
  19. [19]
    [PDF] A3 Rings and Modules
    In Z the units are only 1 and −1 whilst every non-zero element is a unit in Q, R, C. (b) R[x] is an integral domain and the units are the non-zero constant ...
  20. [20]
    [PDF] The general linear group of polynomial rings over regular rings
    In this note we shall prove for two types of regular rings A that every element of is a product of an element of (the group of elementary matrices) and an ...Missing: non- | Show results with:non-
  21. [21]
    Highlights in the History of Spectral Theory - jstor
    1858 Arthur Cayley, A memoir on the theory of matrices, Philos. Trans. Roy. Soc. London, 148. (1858) 17-37, (Math. Papers, II, 475-496). 1870 Theodor ...
  22. [22]
    MOORE ON GENERAL ANALYSIS—I - Project Euclid
    MOORE ON GENERAL ANALYSIS—I. General Analysis. Part I. The Algebra of Matrices. By Eliakim Hastings Moore with the cooperation of Raymond Walter Barnard.
  23. [23]
    Linear groups, with an exposition of the Galois field theory
    Oct 31, 2007 · Linear groups, with an exposition of the Galois field theory. by: Dickson, Leonard E. (Leonard Eugene), 1874-. Publication date: 1901.
  24. [24]
    [PDF] Sophus Lie and the Role of Lie Groups in Mathematics
    The group of linear transformations which leaves x² + y² + z² – c²+2. Page 8. is called the Lorentz group. This group together with the translations. (x, y, z ...
  25. [25]
    [PDF] A History of Complex Simple Lie Algebras - SFA ScholarWorks
    Dec 16, 2023 · Élie Cartan's influence on the development of the theory of Lie algebras, though ... Let L be a Lie subalgebra of gl(V ), where dimV = n < ∞.
  26. [26]
    [PDF] GALOIS THEORY
    The group of the general equation of degree n is the symmetric group on n letters. The general equation of degree n is. ~~ not solvable by radicals if n > 4 ...
  27. [27]
    [PDF] DIEUDONNÉ'S DETERMINANTS AND STRUCTURE OF GENERAL ...
    In this partially expository article we revisit the construction of the Dieudonné de- terminant and structure of general linear groups over division rings.
  28. [28]
    [PDF] Introduction to the Langlands Program - Mathematics
    An important aspect of the Langlands theory for GLn is a reciprocity conjecture that implies the Artin Conjecture (end of §5). We state this conjecture of ...
  29. [29]
    On the Cohomology and K-Theory of the General Linear ... - jstor
    This paper computes the cohomology group H*(GLJk, F1) and determines higher-dimensional algebraic K-groups, showing an isomorphism of K*(k) with homotopy ...
  30. [30]
    [PDF] 14 The Special Linear Group SL(n, F) - Brandeis
    SL(n, F) denotes the kernel of the homomorphism det : GL(n, F) ³ F× = {x ∈ F |x 6= 0} where F is a field. Note that the determinant homomorphism has section s ...
  31. [31]
    [PDF] Linear Algebraic Groups and K-Theory
    These groups are not necessarily algebraic – for example, the group. SLn(D) as defined by the Dieudonné determinant, is in general not algebraic, when D is ...
  32. [32]
    [PDF] Introductory Lectures on SL(2,Z) and modular forms.
    There is an important action of SL(2, R) on the upper half plane U = {z = x + iy | y > 0}, as fractional linear (Mobius) transformations: τ ↦→ aτ + b cτ + ...
  33. [33]
    [PDF] sl2(z) - keith conrad
    field K, with ring of integers OK , all finite-index subgroups of SLn(OK ) (n ≥ 3) are congruence subgroups if and only if K has at least one real embedding.Missing: perfect | Show results with:perfect
  34. [34]
    [PDF] In this lecture, we discussed the basics of the Lie group/Lie algebra cor
    Jan 26, 2024 · The Lie algebra of SLn(K) is the set of trace zero matrices in Mn(K). Proof. By the definition of the Lie algebra associated to a matrix group,.
  35. [35]
    [PDF] Lecture 10 — Trace Form & Cartan's criterion
    Oct 12, 2010 · For the killing form on gln(F), consider the basis of gln(F) ... Given a Lie algebra g over F, let g = F ⊗F g be a Lie algebra over F.
  36. [36]
    [PDF] basics on reductive groups - Yale Math
    For example, the subgroup of all diagonal matrices in GLn is a maximal torus. Theorem 2.4. All maximal tori are conjugate. From now and until the end of the ...
  37. [37]
    [PDF] 1 The Classical Groups - UCSD Math
    We note that if we choose a basis for V and if V is n-dimensional then the matrices of the elements of End(V ) form a Lie algebra which we will denote by gl(n,F) ...
  38. [38]
    [PDF] CLASSICAL GROUPS 1. Orthogonal groups These notes are about ...
    the Lie algebra structure is the commutator of matrices defined above. The descriptions above make it clear that O(n) is a (closed) Lie subgroup of GL(n,R), ...
  39. [39]
    Bruhat, Cartan and Iwasawa decompositions in GLn(R), O(p, q) and ...
    The Iwasawa Decomposition states that G= KB. This says that every flag can be generated by an orthonormal basis. The Cartan decomposition states that G=KAK.
  40. [40]
    [PDF] The diagonal matrices of GLn(R) form a Cartan subgroup.
    If αi is the value of the ith diagonal element of a matrix, then the roots of GLn(R) are the linear forms αi − αj for i 6= j. For example the roots system of ...
  41. [41]
    [PDF] Lecture 9
    The Weyl group W of GL(n,C) is the symmetric group Sn. It acts on T, Λ and Φ by permuting the coordinates. Let W be a group with a fixed set I of generators ...Missing: S_n | Show results with:S_n
  42. [42]
    [PDF] 2. Groups 2.1. Groups and monoids. Let's start out with the basic ...
    We denote this set by Mn(R). Matrix multiplication is associative, so. Mn(R) is a semigroup and in fact a monoid with identity element e = diag(1 ...
  43. [43]
    [PDF] introduction to reductive monoids - louis solomon
    It is this elementary observation which connects the idempotents in Mn to the group structure of GLn and allows one to build a theory of reductive monoids on ...
  44. [44]
    [PDF] matrix semigroups over semirings
    We study properties determined by idempotents in the following families of matrix semigroups over a semiring S: the full matrix semigroup Mn(S), the semigroup.Missing: M_n( | Show results with:M_n(
  45. [45]
    [PDF] An Introduction to Wedderburn Theory & Group Representations
    Wedderburn theory includes basic definitions, the semisimplicity of matrix algebras, Wedderburn's theorem, group representation theory, and characters.<|control11|><|separator|>
  46. [46]
    bounded operators on a Hilbert space form a C∗ - C * - -algebra
    Mar 22, 2013 · In this entry we show how the algebra B(H) B ⁢ ( H ) of bounded linear operators on an Hilbert space H H is one of the most natural examples ...
  47. [47]
    [PDF] On the Size of Finite Rational Matrix Semigroups - arXiv
    After the preliminaries (section 2) and the proof of Theorem 1 (section 3), we discuss applications in automata theory (section 4). In particular we show that ...
  48. [48]
    [PDF] Computational Problems in Matrix Semigroups
    This thesis deals with computational problems that are defined on matrix semigroups, which play a pivotal role in Mathematics and Computer Science.