Lie Algebras, Their Representation Theory and GL Minor Thesis

Total Page:16

File Type:pdf, Size:1020Kb

Lie Algebras, Their Representation Theory and GL Minor Thesis Lie algebras, their representation theory and GLn Minor Thesis Greta Panova June 20, 2008 Contents 1 Introduction 1 2 Lie Algebras 2 2.1 Definitions and main theorems . 2 2.2 Representations of Lie algebras . 7 2.3 Representations of sl(2;F ) ..................................... 9 3 Lie Groups and their Lie algebras 11 4 Representations of semisimple Lie algebras 15 4.1 Root space decomposition . 15 4.2 Roots and weights; the Weyl group . 19 5 The special linear Lie algebra sl(n; C) and the general linear group GLn(C) 22 5.1 Structure of sl(n; C) ......................................... 22 5.2 Representations of sl(n; C)...................................... 24 5.3 Weyl's construction, tensor products and some combinatorics . 25 n 5.4 Representations of GLn(C ) .................................... 27 5.5 Some explicit construction, generators of SλV ........................... 28 6 Conclusion 30 A Schur functors, Weyl's construction and representations of GLn(C) 30 1 Introduction The goal of this minor thesis is to develop the necessary theory of Lie algebras, Lie groups and their representation theory and explicitly determine the structure and representations of sln(C) and GLn(C). My interest in the representations of GL(V ) come from their strong connection to combinatorics as developed in Chapter 7 and its appendix in [3]. Even though the irreducible holomorphic representations of GL(V ) can be derived without passing through the general theory of Lie algebras, as Schur did in his dissertation, any attempt to generalize the result or to obtain analogous results for the unitary or symplectic groups will pass through the theory of Lie algebras. Here we will develop the basic theory of Lie algebras and their representations, focusing on semisimple Lie algebras. We will develop most of the necessary theory to show facts like complete irreducibility of representations of semisimple Lie algebras, develop the theory necessary to decompose the lie algebras into root spaces and use these root spaces to decompose representations into weight spaces and list the irreducible 1 representations via weights. We will establish connections between Lie groups and Lie algebras, which will, for example, enable us to derive the irreducible representations of GL(V ) through the ones for gl(V ). In our development of the basic theory of Lie algebras we will follow mostly [2], while studying Lie groups, roots and weights, sl(n; S) we will follow [1]. We will encounter some combinatorial facts which will be taken for granted and whose proofs are found in [3]. 2 Lie Algebras In this section we will follow [2]. We will develop the basic theory of Lie algebras and later we'll establish how they arise from Lie groups and essentially motivate their existence. 2.1 Definitions and main theorems We will, of course, start with the definition of a Lie algebra. Definition 1. A Lie algebra is a vector space over a field F endowed with a bracket operation L × L ! L denoted (x; y) ! [xy], which satisfies the following axioms: (L1) [xy] is bilinear, (L2) [xx] = 0 for all x 2 L, (L3) [xy] satisfies the Jacobi identity [x[yz]] + [y[zx]] + [z[xy]] = 0 for all x; y; z 2 L. A homomorphism (isomorphism) of Lie algebras will be a vector space homomorphism (resp. isomor- phism) that respects the bracket operation, a subalgebra would a subspace closed under the bracket operation. An important example of Lie algebra is the general linear algebra gl(V ), which coincides as a vector space with End V (or Mn - space of n×n matrices) and has a bracket operation defined as [XY ] = XY −YX (it is a straightforward check to verify the axioms). Next we can define the special linear algebra sl(V ) as the subspace of End V of trace 0 endomorphisms and the same bracket operation, it is clearly a subalgebra of gl(V ). Similarly we can define other subalgebras of gl(V ) by imposing conditions on the matrices, e.g. uppertriangular, strictly uppertriangular, diagonal. Another important example of a gl(U) subalgebra, where U is a vector space endowed with a bilinear operation (denoted as multiplication), is the algebra of derivations Der U. It is first of all the vector space of linear maps δ : U ! U, such that δ(XY ) = Xδ(Y ) + δ(X)Y . In order to verify that this is a subalgebra we need to check that [δδ0] 2 DerU. For the sake of exercise we will do that now. We need to check that [δδ0] = δδ0 − δ0δ satisfies the derivation condition (it is clear it's a linear map), so for X; Y 2 U we apply the derivation rule over and over 1. Now if we let U = L with the bilinear operation defined by the bracket of L, we can talk about Der L. Some of these derivations arise as endomorphism ad x : L ! L given by ad x(y) = [xy]. It is a derivation, since by the Jacobi identity ad x(yz) = [x[yz]] = −[y[zx]] − [z[xy]] = [y[xz]] + [[xy]z] = [y ad x(z)] + [ad(y)z]. We can now make a Definition 2. The map L ! Der L given by x ! ad x, where ad x(y) = [xy] for all y 2 L, is called the adjoint representation of L. Continuing with the definitions we need to define ideals of a Lie algebra. The notion is the same as ring ideals, just that multiplication is given by the bracket, in other words I is an ideal of L if for any x 2 L and y 2 I [xy] 2 I. Two important ideals are the center Z(L) = fz 2 Lj[xz] = 0 for all x 2 Lg and the derived algebra [LL] consisting of all linear combinations of [xy]. Sums and products (given by the bracket) of ideals are clearly also ideals. A Lie algebra with no nontrivial ideals (i.e. 0 and itself) and with [LL] 6= 0 is called simple, in this case [LL] 6= 0 necessarily implies [LL] = L. We define the normalizer of a subalgebra K of 1[δδ0](XY ) = δ(δ0(XY )) − δ0(δ(XY )) = δ(δ0(X)Y + Xδ0(Y )) − δ0(δ(X)Y + Xδ(Y )) = δ(δ0(X)Y ) + δ(Xδ0(Y )) − δ0(δ(X)Y ) − 0 0 0 0 0 0 0 0 0 0 δ (Xδ(Y )) = δ(δ (X))Y +δ (X)δ(Y )+δ(X)δ (Y )+Xδ(δ (Y ))−δ (δ(X))Y −δ(X)δ (Y )−δ (X)δ(Y )−Xδ (δ(Y )) = X(δδ (Y )− δ0δ(Y )) + (δδ0(X) − δ0δ(X))Y = X[δδ0](Y ) + [δδ0](X)Y 2 L as NL(K) = fx 2 Lj[xK] ⊂ Kg and the centralizer of a subset X of L as CL(X) = fx 2 Lj[xX] = 0g, both of them turn out to be subalgebras of L by the Jacobi identity 2 We need to define what we mean by a representation of a Lie algebra since this is the ultimate goal of this paper. A representation of L is a homomorphism φ : L ! gl(V ). The most important example of which is the already mentioned adjoint representation ad L :! gl(L). Since linearity is clear, all we need to check is that it preserves the bracket, which again follows from the Jacobi identity 3. We proceed now to some very important concepts in the theory of Lie algebras. We call a Lie algebra L solvable if the sequence of ideals L(0) = L; L(1) = [LL];:::;L(n) = [L(n−1)L(n−1)] eventually equals 0. Note for example that simple algebras are not solvable, as L(n) = L for all n. Here is the place to introduce the subalgebra of gl(n; F ) t(n; F ) of upper triangular n × n matrices and the subalgebra n(n; F ) of strictly upper triangular matrices. We notice that [t(n; F )t(n; F )] ⊂ n(n; F ), and that taking the commutator of two matrices A and B with entries ai;j; bi;j = 0 if j − i ≤ k for k > 0 will result in C = AB − BA with ci;j = 0 for j − i ≤ k + 1 and continuing this way we will eventually get a 0 matrix (e.g. for k > n), so these two algebras are solvable. We note here some easy facts. If L is solvable, so are all its subalgebras and homomorphic images. Also, if I is a solvable ideal of L and L=I is solvable, then so is L. This can be shown by considering the quotient map π : L ! L=I. Since (L=I)(n) = 0, then so is 0 = (π(L))(n) = π(L(n)), so L(n) ⊂ Ker π = I. Since I(k) = 0 for some k, then L(n+k) = (L(n))(k) = 0. As a consequence of this fact we have that if I and J are solvable ideals of L, then since (I + J)=J ∼ I=(I \ J) is solvable as a quotient of I, by the what we just proved I + J is also solvable. This allows us to make one of the most important definitions, that of semisimplicity. If S is a maximal (not contained in any other such) solvable ideal of L, then for any other solvable ideal I we would have S + I solvable, so by the maximality of S, S + I = S and hence I ⊂ S, showing that S must be unique. We denote S by Rad L, the radical of L and make the Definition 3. A Lie algebra L is called semisimple if its radical is 0, Rad L = 0. We now define another very important notion to play a role later.
Recommended publications
  • Affine Springer Fibers and Affine Deligne-Lusztig Varieties
    Affine Springer Fibers and Affine Deligne-Lusztig Varieties Ulrich G¨ortz Abstract. We give a survey on the notion of affine Grassmannian, on affine Springer fibers and the purity conjecture of Goresky, Kottwitz, and MacPher- son, and on affine Deligne-Lusztig varieties and results about their dimensions in the hyperspecial and Iwahori cases. Mathematics Subject Classification (2000). 22E67; 20G25; 14G35. Keywords. Affine Grassmannian; affine Springer fibers; affine Deligne-Lusztig varieties. 1. Introduction These notes are based on the lectures I gave at the Workshop on Affine Flag Man- ifolds and Principal Bundles which took place in Berlin in September 2008. There are three chapters, corresponding to the main topics of the course. The first one is the construction of the affine Grassmannian and the affine flag variety, which are the ambient spaces of the varieties considered afterwards. In the following chapter we look at affine Springer fibers. They were first investigated in 1988 by Kazhdan and Lusztig [41], and played a prominent role in the recent work about the “fun- damental lemma”, culminating in the proof of the latter by Ngˆo. See Section 3.8. Finally, we study affine Deligne-Lusztig varieties, a “σ-linear variant” of affine Springer fibers over fields of positive characteristic, σ denoting the Frobenius au- tomorphism. The term “affine Deligne-Lusztig variety” was coined by Rapoport who first considered the variety structure on these sets. The sets themselves appear implicitly already much earlier in the study of twisted orbital integrals. We remark that the term “affine” in both cases is not related to the varieties in question being affine, but rather refers to the fact that these are notions defined in the context of an affine root system.
    [Show full text]
  • Simultaneous Diagonalization and SVD of Commuting Matrices
    Simultaneous Diagonalization and SVD of Commuting Matrices Ronald P. Nordgren1 Brown School of Engineering, Rice University Abstract. We present a matrix version of a known method of constructing common eigenvectors of two diagonalizable commuting matrices, thus enabling their simultane- ous diagonalization. The matrices may have simple eigenvalues of multiplicity greater than one. The singular value decomposition (SVD) of a class of commuting matrices also is treated. The effect of row/column permutation is examined. Examples are given. 1 Introduction It is well known that if two diagonalizable matrices have the same eigenvectors, then they commute. The converse also is true and a construction for the common eigen- vectors (enabling simultaneous diagonalization) is known. If one of the matrices has distinct eigenvalues (multiplicity one), it is easy to show that its eigenvectors pertain to both commuting matrices. The case of matrices with simple eigenvalues of mul- tiplicity greater than one requires a more complicated construction of their common eigenvectors. Here, we present a matrix version of a construction procedure given by Horn and Johnson [1, Theorem 1.3.12] and in a video by Sadun [4]. The eigenvector construction procedure also is applied to the singular value decomposition of a class of commuting matrices that includes the case where at least one of the matrices is real and symmetric. In addition, we consider row/column permutation of the commuting matrices. Three examples illustrate the eigenvector construction procedure. 2 Eigenvector Construction Let A and B be diagonalizable square matrices that commute, i.e. AB = BA (1) and their Jordan canonical forms read −1 −1 A = SADASA , B = SBDB SB .
    [Show full text]
  • Arxiv:1302.5859V2 [Math.RT] 13 May 2015 Upre Ynfflosi DMS-0902661
    STABILITY PATTERNS IN REPRESENTATION THEORY STEVEN V SAM AND ANDREW SNOWDEN Abstract. We develop a comprehensive theory of the stable representation categories of several sequences of groups, including the classical and symmetric groups, and their relation to the unstable categories. An important component of this theory is an array of equivalences between the stable representation category and various other categories, each of which has its own flavor (representation theoretic, combinatorial, commutative algebraic, or categorical) and offers a distinct perspective on the stable category. We use this theory to produce a host of specific results, e.g., the construction of injective resolutions of simple objects, duality between the orthogonal and symplectic theories, a canonical derived auto-equivalence of the general linear theory, etc. Contents 1. Introduction 2 1.1. Overview 2 1.2. Descriptions of stable representation theory 4 1.3. Additional results, applications, and remarks 7 1.4. Relation to other work 10 1.5. Future directions 11 1.6. Organization 13 1.7. Notation and conventions 13 2. Preliminaries 14 2.1. Representations of categories 14 2.2. Polynomial representations of GL(∞) and category V 19 2.3. Twisted commutative algebras 22 2.4. Semi-group tca’s and diagram categories 24 2.5. Profinite vector spaces 25 arXiv:1302.5859v2 [math.RT] 13 May 2015 3. The general linear group 25 3.1. Representations of GL(∞) 25 3.2. The walled Brauer algebra and category 29 3.3. Modules over Sym(Ch1, 1i) 33 3.4. Rational Schur functors, universal property and specialization 36 4. The orthogonal and symplectic groups 39 4.1.
    [Show full text]
  • The Algebra Generated by Three Commuting Matrices
    THE ALGEBRA GENERATED BY THREE COMMUTING MATRICES B.A. SETHURAMAN Abstract. We present a survey of an open problem concerning the dimension of the algebra generated by three commuting matrices. This article concerns a problem in algebra that is completely elementary to state, yet, has proven tantalizingly difficult and is as yet unsolved. Consider C[A; B; C] , the C-subalgebra of the n × n matrices Mn(C) generated by three commuting matrices A, B, and C. Thus, C[A; B; C] consists of all C- linear combinations of \monomials" AiBjCk, where i, j, and k range from 0 to infinity. Note that C[A; B; C] and Mn(C) are naturally vector-spaces over C; moreover, C[A; B; C] is a subspace of Mn(C). The problem, quite simply, is this: Is the dimension of C[A; B; C] as a C vector space bounded above by n? 2 Note that the dimension of C[A; B; C] is at most n , because the dimen- 2 sion of Mn(C) is n . Asking for the dimension of C[A; B; C] to be bounded above by n when A, B, and C commute is to put considerable restrictions on C[A; B; C]: this is to require that C[A; B; C] occupy only a small portion of the ambient Mn(C) space in which it sits. Actually, the dimension of C[A; B; C] is already bounded above by some- thing slightly smaller than n2, thanks to a classical theorem of Schur ([16]), who showed that the maximum possible dimension of a commutative C- 2 subalgebra of Mn(C) is 1 + bn =4c.
    [Show full text]
  • Note on the Decomposition of Semisimple Lie Algebras
    Note on the Decomposition of Semisimple Lie Algebras Thomas B. Mieling (Dated: January 5, 2018) Presupposing two criteria by Cartan, it is shown that every semisimple Lie algebra of finite dimension over C is a direct sum of simple Lie algebras. Definition 1 (Simple Lie Algebra). A Lie algebra g is Proposition 2. Let g be a finite dimensional semisimple called simple if g is not abelian and if g itself and f0g are Lie algebra over C and a ⊆ g an ideal. Then g = a ⊕ a? the only ideals in g. where the orthogonal complement is defined with respect to the Killing form K. Furthermore, the restriction of Definition 2 (Semisimple Lie Algebra). A Lie algebra g the Killing form to either a or a? is non-degenerate. is called semisimple if the only abelian ideal in g is f0g. Proof. a? is an ideal since for x 2 a?; y 2 a and z 2 g Definition 3 (Derived Series). Let g be a Lie Algebra. it holds that K([x; z; ]; y) = K(x; [z; y]) = 0 since a is an The derived series is the sequence of subalgebras defined ideal. Thus [a; g] ⊆ a. recursively by D0g := g and Dn+1g := [Dng;Dng]. Since both a and a? are ideals, so is their intersection Definition 4 (Solvable Lie Algebra). A Lie algebra g i = a \ a?. We show that i = f0g. Let x; yi. Then is called solvable if its derived series terminates in the clearly K(x; y) = 0 for x 2 i and y = D1i, so i is solv- trivial subalgebra, i.e.
    [Show full text]
  • Lecture 5: Semisimple Lie Algebras Over C
    LECTURE 5: SEMISIMPLE LIE ALGEBRAS OVER C IVAN LOSEV Introduction In this lecture I will explain the classification of finite dimensional semisimple Lie alge- bras over C. Semisimple Lie algebras are defined similarly to semisimple finite dimensional associative algebras but are far more interesting and rich. The classification reduces to that of simple Lie algebras (i.e., Lie algebras with non-zero bracket and no proper ideals). The classification (initially due to Cartan and Killing) is basically in three steps. 1) Using the structure theory of simple Lie algebras, produce a combinatorial datum, the root system. 2) Study root systems combinatorially arriving at equivalent data (Cartan matrix/ Dynkin diagram). 3) Given a Cartan matrix, produce a simple Lie algebra by generators and relations. In this lecture, we will cover the first two steps. The third step will be carried in Lecture 6. 1. Semisimple Lie algebras Our base field is C (we could use an arbitrary algebraically closed field of characteristic 0). 1.1. Criteria for semisimplicity. We are going to define the notion of a semisimple Lie algebra and give some criteria for semisimplicity. This turns out to be very similar to the case of semisimple associative algebras (although the proofs are much harder). Let g be a finite dimensional Lie algebra. Definition 1.1. We say that g is simple, if g has no proper ideals and dim g > 1 (so we exclude the one-dimensional abelian Lie algebra). We say that g is semisimple if it is the direct sum of simple algebras. Any semisimple algebra g is the Lie algebra of an algebraic group, we can take the au- tomorphism group Aut(g).
    [Show full text]
  • Arxiv:1705.10957V2 [Math.AC] 14 Dec 2017 Ideal Called Is and Commutator X Y Vrafield a Over Let Matrix Eae Leri Es Is E Sdfierlvn Notions
    NEARLY COMMUTING MATRICES ZHIBEK KADYRSIZOVA Abstract. We prove that the algebraic set of pairs of matrices with a diag- onal commutator over a field of positive prime characteristic, its irreducible components, and their intersection are F -pure when the size of matrices is equal to 3. Furthermore, we show that this algebraic set is reduced and the intersection of its irreducible components is irreducible in any characteristic for pairs of matrices of any size. In addition, we discuss various conjectures on the singularities of these algebraic sets and the system of parameters on the corresponding coordinate rings. Keywords: Frobenius, singularities, F -purity, commuting matrices 1. Introduction and preliminaries In this paper we study algebraic sets of pairs of matrices such that their commutator is either nonzero diagonal or zero. We also consider some other related algebraic sets. First let us define relevant notions. Let X =(x ) and Y =(y ) be n×n matrices of indeterminates arXiv:1705.10957v2 [math.AC] 14 Dec 2017 ij 1≤i,j≤n ij 1≤i,j≤n over a field K. Let R = K[X,Y ] be the polynomial ring in {xij,yij}1≤i,j≤n and let I denote the ideal generated by the off-diagonal entries of the commutator matrix XY − YX and J denote the ideal generated by the entries of XY − YX. The ideal I defines the algebraic set of pairs of matrices with a diagonal commutator and is called the algebraic set of nearly commuting matrices. The ideal J defines the algebraic set of pairs of commuting matrices.
    [Show full text]
  • Lecture 2: Spectral Theorems
    Lecture 2: Spectral Theorems This lecture introduces normal matrices. The spectral theorem will inform us that normal matrices are exactly the unitarily diagonalizable matrices. As a consequence, we will deduce the classical spectral theorem for Hermitian matrices. The case of commuting families of matrices will also be studied. All of this corresponds to section 2.5 of the textbook. 1 Normal matrices Definition 1. A matrix A 2 Mn is called a normal matrix if AA∗ = A∗A: Observation: The set of normal matrices includes all the Hermitian matrices (A∗ = A), the skew-Hermitian matrices (A∗ = −A), and the unitary matrices (AA∗ = A∗A = I). It also " # " # 1 −1 1 1 contains other matrices, e.g. , but not all matrices, e.g. 1 1 0 1 Here is an alternate characterization of normal matrices. Theorem 2. A matrix A 2 Mn is normal iff ∗ n kAxk2 = kA xk2 for all x 2 C : n Proof. If A is normal, then for any x 2 C , 2 ∗ ∗ ∗ ∗ ∗ 2 kAxk2 = hAx; Axi = hx; A Axi = hx; AA xi = hA x; A xi = kA xk2: ∗ n n Conversely, suppose that kAxk = kA xk for all x 2 C . For any x; y 2 C and for λ 2 C with jλj = 1 chosen so that <(λhx; (A∗A − AA∗)yi) = jhx; (A∗A − AA∗)yij, we expand both sides of 2 ∗ 2 kA(λx + y)k2 = kA (λx + y)k2 to obtain 2 2 ∗ 2 ∗ 2 ∗ ∗ kAxk2 + kAyk2 + 2<(λhAx; Ayi) = kA xk2 + kA yk2 + 2<(λhA x; A yi): 2 ∗ 2 2 ∗ 2 Using the facts that kAxk2 = kA xk2 and kAyk2 = kA yk2, we derive 0 = <(λhAx; Ayi − λhA∗x; A∗yi) = <(λhx; A∗Ayi − λhx; AA∗yi) = <(λhx; (A∗A − AA∗)yi) = jhx; (A∗A − AA∗)yij: n ∗ ∗ n Since this is true for any x 2 C , we deduce (A A − AA )y = 0, which holds for any y 2 C , meaning that A∗A − AA∗ = 0, as desired.
    [Show full text]
  • Irreducible Representations of Complex Semisimple Lie Algebras
    Irreducible Representations of Complex Semisimple Lie Algebras Rush Brown September 8, 2016 Abstract In this paper we give the background and proof of a useful theorem classifying irreducible representations of semisimple algebras by their highest weight, as well as restricting what that highest weight can be. Contents 1 Introduction 1 2 Lie Algebras 2 3 Solvable and Semisimple Lie Algebras 3 4 Preservation of the Jordan Form 5 5 The Killing Form 6 6 Cartan Subalgebras 7 7 Representations of sl2 10 8 Roots and Weights 11 9 Representations of Semisimple Lie Algebras 14 1 Introduction In this paper I build up to a useful theorem characterizing the irreducible representations of semisim- ple Lie algebras, namely that finite-dimensional irreducible representations are defined up to iso- morphism by their highest weight ! and that !(Hα) is an integer for any root α of R. This is the main theorem Fulton and Harris use for showing that the Weyl construction for sln gives all (finite-dimensional) irreducible representations. 1 In the first half of the paper we'll see some of the general theory of semisimple Lie algebras| building up to the existence of Cartan subalgebras|for which we will use a mix of Fulton and Harris and Serre ([1] and [2]), with minor changes where I thought the proofs needed less or more clarification (especially in the proof of the existence of Cartan subalgebras). Most of them are, however, copied nearly verbatim from the source. In the second half of the paper we will describe the roots of a semisimple Lie algebra with respect to some Cartan subalgebra, the weights of irreducible representations, and finally prove the promised result.
    [Show full text]
  • DIFFERENTIAL GEOMETRY FINAL PROJECT 1. Introduction for This
    DIFFERENTIAL GEOMETRY FINAL PROJECT KOUNDINYA VAJJHA 1. Introduction For this project, we outline the basic structure theory of Lie groups relating them to the concept of Lie algebras. Roughly, a Lie algebra encodes the \infinitesimal” structure of a Lie group, but is simpler, being a vector space rather than a nonlinear manifold. At the local level at least, the Fundamental Theorems of Lie allow one to reconstruct the group from the algebra. 2. The Category of Local (Lie) Groups The correspondence between Lie groups and Lie algebras will be local in nature, the only portion of the Lie group that will be of importance is that portion of the group close to the group identity 1. To formalize this locality, we introduce local groups: Definition 2.1 (Local group). A local topological group is a topological space G, with an identity element 1 2 G, a partially defined but continuous multiplication operation · :Ω ! G for some domain Ω ⊂ G × G, a partially defined but continuous inversion operation ()−1 :Λ ! G with Λ ⊂ G, obeying the following axioms: • Ω is an open neighbourhood of G × f1g Sf1g × G and Λ is an open neighbourhood of 1. • (Local associativity) If it happens that for elements g; h; k 2 G g · (h · k) and (g · h) · k are both well-defined, then they are equal. • (Identity) For all g 2 G, g · 1 = 1 · g = g. • (Local inverse) If g 2 G and g−1 is well-defined in G, then g · g−1 = g−1 · g = 1. A local group is said to be symmetric if Λ = G, that is, every element g 2 G has an inverse in G.A local Lie group is a local group in which the underlying topological space is a smooth manifold and where the associated maps are smooth maps on their domain of definition.
    [Show full text]
  • SCHUR-WEYL DUALITY Contents Introduction 1 1. Representation
    SCHUR-WEYL DUALITY JAMES STEVENS Contents Introduction 1 1. Representation Theory of Finite Groups 2 1.1. Preliminaries 2 1.2. Group Algebra 4 1.3. Character Theory 5 2. Irreducible Representations of the Symmetric Group 8 2.1. Specht Modules 8 2.2. Dimension Formulas 11 2.3. The RSK-Correspondence 12 3. Schur-Weyl Duality 13 3.1. Representations of Lie Groups and Lie Algebras 13 3.2. Schur-Weyl Duality for GL(V ) 15 3.3. Schur Functors and Algebraic Representations 16 3.4. Other Cases of Schur-Weyl Duality 17 Appendix A. Semisimple Algebras and Double Centralizer Theorem 19 Acknowledgments 20 References 21 Introduction. In this paper, we build up to one of the remarkable results in representation theory called Schur-Weyl Duality. It connects the irreducible rep- resentations of the symmetric group to irreducible algebraic representations of the general linear group of a complex vector space. We do so in three sections: (1) In Section 1, we develop some of the general theory of representations of finite groups. In particular, we have a subsection on character theory. We will see that the simple notion of a character has tremendous consequences that would be very difficult to show otherwise. Also, we introduce the group algebra which will be vital in Section 2. (2) In Section 2, we narrow our focus down to irreducible representations of the symmetric group. We will show that the irreducible representations of Sn up to isomorphism are in bijection with partitions of n via a construc- tion through certain elements of the group algebra.
    [Show full text]
  • The Partition Algebra and the Kronecker Product (Extended Abstract)
    FPSAC 2013, Paris, France DMTCS proc. (subm.), by the authors, 1–12 The partition algebra and the Kronecker product (Extended abstract) C. Bowman1yand M. De Visscher2 and R. Orellana3z 1Institut de Mathematiques´ de Jussieu, 175 rue du chevaleret, 75013, Paris 2Centre for Mathematical Science, City University London, Northampton Square, London, EC1V 0HB, England. 3 Department of Mathematics, Dartmouth College, 6188 Kemeny Hall, Hanover, NH 03755, USA Abstract. We propose a new approach to study the Kronecker coefficients by using the Schur–Weyl duality between the symmetric group and the partition algebra. Resum´ e.´ Nous proposons une nouvelle approche pour l’etude´ des coefficients´ de Kronecker via la dualite´ entre le groupe symetrique´ et l’algebre` des partitions. Keywords: Kronecker coefficients, tensor product, partition algebra, representations of the symmetric group 1 Introduction A fundamental problem in the representation theory of the symmetric group is to describe the coeffi- cients in the decomposition of the tensor product of two Specht modules. These coefficients are known in the literature as the Kronecker coefficients. Finding a formula or combinatorial interpretation for these coefficients has been described by Richard Stanley as ‘one of the main problems in the combinatorial rep- resentation theory of the symmetric group’. This question has received the attention of Littlewood [Lit58], James [JK81, Chapter 2.9], Lascoux [Las80], Thibon [Thi91], Garsia and Remmel [GR85], Kleshchev and Bessenrodt [BK99] amongst others and yet a combinatorial solution has remained beyond reach for over a hundred years. Murnaghan discovered an amazing limiting phenomenon satisfied by the Kronecker coefficients; as we increase the length of the first row of the indexing partitions the sequence of Kronecker coefficients obtained stabilises.
    [Show full text]