Chapter 1 Linear Groups

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 1 Linear Groups Chapter 1 Linear groups We begin, as we shall end, with the classical groups—those familiar groups of matrices encountered in every branch of mathematics. At the outset, they serve as a library of linear groups, with which to illustrate our theory. Later we shall find that these same groups also serve as the building-blocks for the theory. Definition 1.1 A linear group is a closed subgroup of GL(n; R). Remarks: 1. We could equally well say that: A linear group is a closed subgroup of GL(n; C). For as we shall see shortly, GL(n; C) has an isomorphic im- age as a closed subgroup of GL(2n; R); while conversely it is clear that GL(n; R) can be regarded as a closed subgroup of GL(n; C). 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) : det T = 0 f 2 6 g where M(n; k) denotes the space of all n n real or complex matrices (according as k = R or C). × 3. We define a norm X on M(n; k) by k k tr(X X) if k = R; X 2 = X 2 = ( 0 X ij tr(X X) k = C; k k i;j j j ∗ if where as usual X0 denotes the transpose and X∗ the conjugate transpose. This is just the usual Euclidean norm, if we identify M(n; k) with kN , where 424–III 1–1 424–III 1–2 2 2 N = n , by taking as the coordinates of the matrix X its n entries Xij. The requirement that a linear group should be closed in GL(n; k) refers to this metric topology. In other words, if T (i) is a sequence of matrices in G tending towards the matrix T GL(n; k), ie 2 T (i) T 0; k − k ! then T must in fact lie in G. Examples: 1. The general linear group GL(n; R) 2. The special linear group SL(n; R) = T GL(n; R) : det(T ) = 1 f 2 g 3. The orthogonal group O(n) = T GL(n; R): T 0T = I f 2 g In other words O(n) = T : Q(T v) = Q(v) v Rn ; f 8 2 g where 2 2 n Q(v) = x + ::: + x (v = (x1; : : : ; xn) R ); 1 n 2 ie O(n) is the subgroup of GL(n; R) leaving the quadratic form Q invariant. 4. The special orthogonal group SO(n) = O(n) SL(n) \ 5. The complex general linear group GL(n; C). This calls for some explana- tion, since GL(n; C) is not a group of real matrices, as required by Defini- tion 1. However, we can represent each complex matrix Z M(n; C) by a real matrix RZ M(2n; R) in the following way. (Compare2 the “realifica- tion” of a representation2 discussed in Part I Chapter 11.) If we “forget” scalar multiplication by non-reals, the complex vector space V = Cn becomes a real vector space RV of twice the dimension, with basis (1; 0;:::; ); (i; 0;:::; 0); (0; 1;:::; 0); (0; i; : : : ; 0);:::; (0; 0;:::; 1); (0; 0; : : : ; i): 424–III 1–3 Moreover each matrix Z M(n; C), ie each linear map 2 Z : V V ! defines a linear map RZ : RV RV; ! ie a matrix RZ M(2n; R). 2 Concretely, in passing from Z to RZ each entry Zj;k = Xj;k + iYj;k is replaced by the 2 2 matrix × Xj;k Yj;k − ! Yj;k Xj;k The map Z RZ : M(n; C) M(2n; R) 7! ! is injective; and it preserves the algebraic structure, ie R(Z + W ) = RZ + RW • R(ZW ) = (RZ)(RW ) • R(aZ) = a(RZ) a R • 8 2 RI = I • R(Z∗) = (RZ)0. • It follows in particular that RZ is invertible if and only if Z is; so R restricts to a map Z RZ : GL(n; C) GL(2n; R): 7! ! Whenever we speak of GL(n; C), or more generally of any group G of complex matrices, as a linear group, it is understood that we refer to the image RG of G under this injection R. The matrix X GL(2n; R) belongs to GL(n; C) if is built out of 2 2 matrices of the2 form × x y ! y− x 424–III 1–4 This can be expressed more neatly as follows. Let 0 1 0 1 1− 0 B C B 0 1 C iI J = B C 7! B − C B 1 0 C B C B .. C @ . A Since any scalar multiple of the identity commutes with all matrices, X M(n; C) = (iI)X = X(iI): 2 ) Applying the operator R, X RM(n; C) = JX = XJ: 2 ) Converseley, if JX = XJ then it is readily verified that X is of the required form. Thus RM(n; C) = X M(2n; R): JX = XJ ; f 2 g and in particular GL(n; C) = T GL(2n; R): JX = XJ f 2 g 6. The complex special linear group SL(n; C) = T GL(n; C) : det T = 1 f 2 g Note that the determinant here must be computed in M(n; C), not in M(2n; R). Thus T = i = SL(1; C); 2 although 0 1 RT = − ! SL(2; R): 1 0 2 7. The unitary group U(n) = T GL(n; C): T ∗T = I f 2 g where T ∗ denotes the complex transpose of T . In other words U(n) = T : H(T v) = H(v) v Cn ; f 8 2 g 424–III 1–5 where H is the Hermitian form 2 2 n H(v) = x1 + ::: + xn (v = (x1; : : : ; xn) C ): j j j j 2 Since R(T ∗) = (RT )0, a complex matrix is unitary if and only if its real counterpart is orthogonal: U(n) = GL(n; C) O(2n) \ 8. The special unitary group SU(n) = U(n) SL(n; C) \ 9. The quaternionic general linear group GL(n; H). The quaternions q = t + xi + yj + zk (t; x; y; z R); 2 with multiplication defined by i2 = j2 = k2 = 1; jk = kj = i; ki = ik = j; ij = ji = k − − − − form a division algebra or skew field. For the product of any quaternion q = t + xi + yj + zk with its conjugate q¯ = t xi yj zk − − − gives its norm-square q 2 =qq ¯ = t2 + x2 + y2 + z2; k k so that if q = 0, 6 q¯ q 1 = : − q 2 k k We can equally well regard H as a 2-dimensional algebra over C, with each quaternion taking the form q = z + wj (z; w C); 2 and multiplication in H being defined by the rules jz =zj; ¯ j2 = 1 − 424–III 1–6 Surprisingly perhaps, the entire apparatus of linear algebra extends almost unchanged to the case of a non-commutative scalar field. Thus we may speak of an n-dimensional vector space W over H, of a linear map t : W W , etc. ! A quaternionic vector space W defines a complex vector space CW by “forgetting” scalar multiplication by non-complex quaternions (ie those in- volving j or k), in just the same way as a complex vector space V defines a real vector space RV . If W has quaternionic dimension n, with basis e1; e2; : : : ; en , then CW has complex dimension 2n, with basis f g e1; je1; e2; je2; : : : ; en; jen : f g Moreover each matrix Q M(n; H), ie each linear map 2 Q : W W ! defines a linear map CQ : CW CW; ! ie a matrix CQ M(2n; C). 2 Concretely, in passing from W to CW each entry Qr;s = Zr;s + iWr;s is replaced by the 2 2 complex matrix × Zr;s Wr;s − ! Wr;s Zr;s The map Q CQ : M(n; H) M(2n; C) 7! ! is injective; and it preserves the algebraic structure, ie C(Q + Q0) = CQ + CQ0 • C(QQ0) = (CQ)(CQ0) • C(aQ) = a(CQ) a C • 8 2 CI = I • C(Q∗) = (CQ)∗. • 424–III 1–7 In this last relation, Q∗ denotes the quaternionic matrix with entries (Q∗)rs = Qsr: To identify M(n; H) as a subspace of M(2n; C), consider the automor- phism of H 1 q q~ = jqj− : 7! In terms of its 2 complex components, q = z + wj q~ =z ¯ +wj: ¯ 7! (Note that q~ =q ¯; indeed, the map q q¯ is not an automorphism of H, but an anti-automorphism.)6 7! Let J denote the diagonal matrix j 0 0 1 J = 0 j M(n; H); B C B .. C 2 @ . A and consider the map 1 Q Q~ = JQJ − : M(n; H) M(n; H): 7! ! We see from above that C(Q~) = CQ; where X¯ denotes (for X M(n; C)) the matrix with entries 2 ¯ Xr;s = Xr;s: Now 0 1 0 1 1− 0 B C B 0 1 C J CJ = B C ; 7! B − C B 1 0 C B C B .. C @ . A which we take the liberty of also denoting by J (as we did earlier, when defining the embedding of GL(n; C) in GL(2n; R), although there we re- garded J as a real matrix rather than a complex one). Thus 1 M(n; H) X M(2n; C): JXJ − = X¯ ⊂ f 2 g 424–III 1–8 1 Conversely, it is readily verified that if JXJ − = X¯, then X is constructed from 2 2 matrices of the form specified above, and so arises from a quater- nionic× matrix.
Recommended publications
  • Group Homomorphisms
    1-17-2018 Group Homomorphisms Here are the operation tables for two groups of order 4: · 1 a a2 + 0 1 2 1 1 a a2 0 0 1 2 a a a2 1 1 1 2 0 a2 a2 1 a 2 2 0 1 There is an obvious sense in which these two groups are “the same”: You can get the second table from the first by replacing 0 with 1, 1 with a, and 2 with a2. When are two groups the same? You might think of saying that two groups are the same if you can get one group’s table from the other by substitution, as above. However, there are problems with this. In the first place, it might be very difficult to check — imagine having to write down a multiplication table for a group of order 256! In the second place, it’s not clear what a “multiplication table” is if a group is infinite. One way to implement a substitution is to use a function. In a sense, a function is a thing which “substitutes” its output for its input. I’ll define what it means for two groups to be “the same” by using certain kinds of functions between groups. These functions are called group homomorphisms; a special kind of homomorphism, called an isomorphism, will be used to define “sameness” for groups. Definition. Let G and H be groups. A homomorphism from G to H is a function f : G → H such that f(x · y)= f(x) · f(y) forall x,y ∈ G.
    [Show full text]
  • Categories of Sets with a Group Action
    Categories of sets with a group action Bachelor Thesis of Joris Weimar under supervision of Professor S.J. Edixhoven Mathematisch Instituut, Universiteit Leiden Leiden, 13 June 2008 Contents 1 Introduction 1 1.1 Abstract . .1 1.2 Working method . .1 1.2.1 Notation . .1 2 Categories 3 2.1 Basics . .3 2.1.1 Functors . .4 2.1.2 Natural transformations . .5 2.2 Categorical constructions . .6 2.2.1 Products and coproducts . .6 2.2.2 Fibered products and fibered coproducts . .9 3 An equivalence of categories 13 3.1 G-sets . 13 3.2 Covering spaces . 15 3.2.1 The fundamental group . 15 3.2.2 Covering spaces and the homotopy lifting property . 16 3.2.3 Induced homomorphisms . 18 3.2.4 Classifying covering spaces through the fundamental group . 19 3.3 The equivalence . 24 3.3.1 The functors . 25 4 Applications and examples 31 4.1 Automorphisms and recovering the fundamental group . 31 4.2 The Seifert-van Kampen theorem . 32 4.2.1 The categories C1, C2, and πP -Set ................... 33 4.2.2 The functors . 34 4.2.3 Example . 36 Bibliography 38 Index 40 iii 1 Introduction 1.1 Abstract In the 40s, Mac Lane and Eilenberg introduced categories. Although by some referred to as abstract nonsense, the idea of categories allows one to talk about mathematical objects and their relationions in a general setting. Its origins lie in the field of algebraic topology, one of the topics that will be explored in this thesis. First, a concise introduction to categories will be given.
    [Show full text]
  • The Unitary Representations of the Poincaré Group in Any Spacetime
    The unitary representations of the Poincar´e group in any spacetime dimension Xavier Bekaert a and Nicolas Boulanger b a Institut Denis Poisson, Unit´emixte de Recherche 7013, Universit´ede Tours, Universit´ed’Orl´eans, CNRS, Parc de Grandmont, 37200 Tours (France) [email protected] b Service de Physique de l’Univers, Champs et Gravitation Universit´ede Mons – UMONS, Place du Parc 20, 7000 Mons (Belgium) [email protected] An extensive group-theoretical treatment of linear relativistic field equa- tions on Minkowski spacetime of arbitrary dimension D > 2 is presented in these lecture notes. To start with, the one-to-one correspondence be- tween linear relativistic field equations and unitary representations of the isometry group is reviewed. In turn, the method of induced representa- tions reduces the problem of classifying the representations of the Poincar´e group ISO(D 1, 1) to the classification of the representations of the sta- − bility subgroups only. Therefore, an exhaustive treatment of the two most important classes of unitary irreducible representations, corresponding to massive and massless particles (the latter class decomposing in turn into the “helicity” and the “infinite-spin” representations) may be performed via the well-known representation theory of the orthogonal groups O(n) (with D 4 <n<D ). Finally, covariant field equations are given for each − unitary irreducible representation of the Poincar´egroup with non-negative arXiv:hep-th/0611263v2 13 Jun 2021 mass-squared. Tachyonic representations are also examined. All these steps are covered in many details and with examples. The present notes also include a self-contained review of the representation theory of the general linear and (in)homogeneous orthogonal groups in terms of Young diagrams.
    [Show full text]
  • Low-Dimensional Representations of Matrix Groups and Group Actions on CAT (0) Spaces and Manifolds
    Low-dimensional representations of matrix groups and group actions on CAT(0) spaces and manifolds Shengkui Ye National University of Singapore January 8, 2018 Abstract We study low-dimensional representations of matrix groups over gen- eral rings, by considering group actions on CAT(0) spaces, spheres and acyclic manifolds. 1 Introduction Low-dimensional representations are studied by many authors, such as Gural- nick and Tiep [24] (for matrix groups over fields), Potapchik and Rapinchuk [30] (for automorphism group of free group), Dokovi´cand Platonov [18] (for Aut(F2)), Weinberger [35] (for SLn(Z)) and so on. In this article, we study low-dimensional representations of matrix groups over general rings. Let R be an associative ring with identity and En(R) (n ≥ 3) the group generated by ele- mentary matrices (cf. Section 3.1). As motivation, we can consider the following problem. Problem 1. For n ≥ 3, is there any nontrivial group homomorphism En(R) → En−1(R)? arXiv:1207.6747v1 [math.GT] 29 Jul 2012 Although this is a purely algebraic problem, in general it seems hard to give an answer in an algebraic way. In this article, we try to answer Prob- lem 1 negatively from the point of view of geometric group theory. The idea is to find a good geometric object on which En−1(R) acts naturally and non- trivially while En(R) can only act in a special way. We study matrix group actions on CAT(0) spaces, spheres and acyclic manifolds. We prove that for low-dimensional CAT(0) spaces, a matrix group action always has a global fixed point (cf.
    [Show full text]
  • The General Linear Group
    18.704 Gabe Cunningham 2/18/05 [email protected] The General Linear Group 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 multiplication. It is easy to see that GLn(F ) is, in fact, a group: matrix multiplication is associative; the identity element is In, the n × n matrix with 1’s along the main diagonal and 0’s everywhere else; and the matrices are invertible by choice. It’s not immediately clear whether GLn(F ) has infinitely many elements when F does. However, such is the case. Let a ∈ F , a 6= 0. −1 Then a · In is an invertible n × n matrix with inverse a · In. In fact, the set of all such × matrices forms a subgroup of GLn(F ) that is isomorphic to F = F \{0}. It is clear that if F is a finite field, then GLn(F ) has only finitely many elements. An interesting question to ask is how many elements it has. Before addressing that question fully, let’s look at some examples. ∼ × Example 1: Let n = 1. Then GLn(Fq) = Fq , which has q − 1 elements. a b Example 2: Let n = 2; let M = ( c d ). Then for M to be invertible, it is necessary and sufficient that ad 6= bc. If a, b, c, and d are all nonzero, then we can fix a, b, and c arbitrarily, and d can be anything but a−1bc. This gives us (q − 1)3(q − 2) matrices.
    [Show full text]
  • Material on Algebraic and Lie Groups
    2 Lie groups and algebraic groups. 2.1 Basic Definitions. In this subsection we will introduce the class of groups to be studied. We first recall that a Lie group is a group that is also a differentiable manifold 1 and multiplication (x, y xy) and inverse (x x ) are C1 maps. An algebraic group is a group7! that is also an algebraic7! variety such that multi- plication and inverse are morphisms. Before we can introduce our main characters we first consider GL(n, C) as an affi ne algebraic group. Here Mn(C) denotes the space of n n matrices and GL(n, C) = g Mn(C) det(g) =) . Now Mn(C) is given the structure nf2 2 j 6 g of affi ne space C with the coordinates xij for X = [xij] . This implies that GL(n, C) is Z-open and as a variety is isomorphic with the affi ne variety 1 Mn(C) det . This implies that (GL(n, C)) = C[xij, det ]. f g O Lemma 1 If G is an algebraic group over an algebraically closed field, F , then every point in G is smooth. Proof. Let Lg : G G be given by Lgx = gx. Then Lg is an isomorphism ! 1 1 of G as an algebraic variety (Lg = Lg ). Since isomorphisms preserve the set of smooth points we see that if x G is smooth so is every element of Gx = G. 2 Proposition 2 If G is an algebraic group over an algebraically closed field F then the Z-connected components Proof.
    [Show full text]
  • Modules and Lie Semialgebras Over Semirings with a Negation Map 3
    MODULES AND LIE SEMIALGEBRAS OVER SEMIRINGS WITH A NEGATION MAP GUY BLACHAR Abstract. In this article, we present the basic definitions of modules and Lie semialgebras over semirings with a negation map. Our main example of a semiring with a negation map is ELT algebras, and some of the results in this article are formulated and proved only in the ELT theory. When dealing with modules, we focus on linearly independent sets and spanning sets. We define a notion of lifting a module with a negation map, similarly to the tropicalization process, and use it to prove several theorems about semirings with a negation map which possess a lift. In the context of Lie semialgebras over semirings with a negation map, we first give basic definitions, and provide parallel constructions to the classical Lie algebras. We prove an ELT version of Cartan’s criterion for semisimplicity, and provide a counterexample for the naive version of the PBW Theorem. Contents Page 0. Introduction 2 0.1. Semirings with a Negation Map 2 0.2. Modules Over Semirings with a Negation Map 3 0.3. Supertropical Algebras 4 0.4. Exploded Layered Tropical Algebras 4 1. Modules over Semirings with a Negation Map 5 1.1. The Surpassing Relation for Modules 6 1.2. Basic Definitions for Modules 7 1.3. -morphisms 9 1.4. Lifting a Module Over a Semiring with a Negation Map 10 1.5. Linearly Independent Sets 13 1.6. d-bases and s-bases 14 1.7. Free Modules over Semirings with a Negation Map 18 2.
    [Show full text]
  • Självständiga Arbeten I Matematik
    SJÄLVSTÄNDIGA ARBETEN I MATEMATIK MATEMATISKA INSTITUTIONEN, STOCKHOLMS UNIVERSITET Geometric interpretation of non-associative composition algebras av Fredrik Cumlin 2020 - No K13 MATEMATISKA INSTITUTIONEN, STOCKHOLMS UNIVERSITET, 106 91 STOCKHOLM Geometric interpretation of non-associative composition algebras Fredrik Cumlin Självständigt arbete i matematik 15 högskolepoäng, grundnivå Handledare: Wushi Goldring 2020 Abstract This paper aims to discuss the connection between non-associative composition algebra and geometry. It will first recall the notion of an algebra, and investigate the properties of an algebra together with a com- position norm. The composition norm will induce a law on the algebra, which is stated as the composition law. This law is then used to derive the multiplication and conjugation laws, where the last is also known as convolution. These laws are then used to prove Hurwitz’s celebrated the- orem concerning the different finite composition algebras. More properties of composition algebras will be covered, in order to look at the structure of the quaternions H and octonions O. The famous Fano plane will be the finishing touch of the relationship between the standard orthogonal vectors which construct the octonions. Lastly, the notion of invertible maps in relation to invertible loops will be covered, to later show the connection between 8 dimensional rotations − and multiplication of unit octonions. 2 Contents 1 Algebra 4 1.1 The multiplication laws . .6 1.2 The conjugation laws . .7 1.3 Dickson double . .8 1.4 Hurwitz’s theorem . 11 2 Properties of composition algebras 14 2.1 The left-, right- and bi-multiplication maps . 16 2.2 Basic properties of quaternions and octonions .
    [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]
  • Matrix Lie Groups
    Maths Seminar 2007 MATRIX LIE GROUPS Claudiu C Remsing Dept of Mathematics (Pure and Applied) Rhodes University Grahamstown 6140 26 September 2007 RhodesUniv CCR 0 Maths Seminar 2007 TALK OUTLINE 1. What is a matrix Lie group ? 2. Matrices revisited. 3. Examples of matrix Lie groups. 4. Matrix Lie algebras. 5. A glimpse at elementary Lie theory. 6. Life beyond elementary Lie theory. RhodesUniv CCR 1 Maths Seminar 2007 1. What is a matrix Lie group ? Matrix Lie groups are groups of invertible • matrices that have desirable geometric features. So matrix Lie groups are simultaneously algebraic and geometric objects. Matrix Lie groups naturally arise in • – geometry (classical, algebraic, differential) – complex analyis – differential equations – Fourier analysis – algebra (group theory, ring theory) – number theory – combinatorics. RhodesUniv CCR 2 Maths Seminar 2007 Matrix Lie groups are encountered in many • applications in – physics (geometric mechanics, quantum con- trol) – engineering (motion control, robotics) – computational chemistry (molecular mo- tion) – computer science (computer animation, computer vision, quantum computation). “It turns out that matrix [Lie] groups • pop up in virtually any investigation of objects with symmetries, such as molecules in chemistry, particles in physics, and projective spaces in geometry”. (K. Tapp, 2005) RhodesUniv CCR 3 Maths Seminar 2007 EXAMPLE 1 : The Euclidean group E (2). • E (2) = F : R2 R2 F is an isometry . → | n o The vector space R2 is equipped with the standard Euclidean structure (the “dot product”) x y = x y + x y (x, y R2), • 1 1 2 2 ∈ hence with the Euclidean distance d (x, y) = (y x) (y x) (x, y R2).
    [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]
  • The Fundamental Homomorphism Theorem
    Lecture 4.3: The fundamental homomorphism theorem Matthew Macauley Department of Mathematical Sciences Clemson University http://www.math.clemson.edu/~macaule/ Math 4120, Modern Algebra M. Macauley (Clemson) Lecture 4.3: The fundamental homomorphism theorem Math 4120, Modern Algebra 1 / 10 Motivating example (from the previous lecture) Define the homomorphism φ : Q4 ! V4 via φ(i) = v and φ(j) = h. Since Q4 = hi; ji: φ(1) = e ; φ(−1) = φ(i 2) = φ(i)2 = v 2 = e ; φ(k) = φ(ij) = φ(i)φ(j) = vh = r ; φ(−k) = φ(ji) = φ(j)φ(i) = hv = r ; φ(−i) = φ(−1)φ(i) = ev = v ; φ(−j) = φ(−1)φ(j) = eh = h : Let's quotient out by Ker φ = {−1; 1g: 1 i K 1 i iK K iK −1 −i −1 −i Q4 Q4 Q4=K −j −k −j −k jK kK j k jK j k kK Q4 organized by the left cosets of K collapse cosets subgroup K = h−1i are near each other into single nodes Key observation Q4= Ker(φ) =∼ Im(φ). M. Macauley (Clemson) Lecture 4.3: The fundamental homomorphism theorem Math 4120, Modern Algebra 2 / 10 The Fundamental Homomorphism Theorem The following result is one of the central results in group theory. Fundamental homomorphism theorem (FHT) If φ: G ! H is a homomorphism, then Im(φ) =∼ G= Ker(φ). The FHT says that every homomorphism can be decomposed into two steps: (i) quotient out by the kernel, and then (ii) relabel the nodes via φ. G φ Im φ (Ker φ C G) any homomorphism q i quotient remaining isomorphism process (\relabeling") G Ker φ group of cosets M.
    [Show full text]