Classical Groups

Total Page:16

File Type:pdf, Size:1020Kb

Classical Groups Chapter 2 Classical groups 2.1 Introduction In this chapter we describe the six families of so-called ‘classical’ simple groups. These are the linear, unitary and symplectic groups, and the three families of orthog- ¢¡ ¥¤ onal groups. All may be obtained from suitable matrix groups G by taking G Z £ G . Our main aims again are to define these groups, prove they are simple, and describe their automorphisms, subgroups and covering groups. We begin with some basic facts about finite fields, before constructing the general ¤ linear groups and their subquotients PSLn £ q , which are usually simple. We prove simplicity using Iwasawa’s Lemma, and construct some important subgroups using the geometry of the underlying vector space. This leads into a description of pro- jective spaces, especially projective lines and projective planes, setting the scene for ¤ proofs of several remarkable isomorphisms between certain groups PSLn £ q and other ¤§¦¨ ¤§¦¨ ¤©¦¨ £ £ groups, in particular PSL2 £ 4 PSL2 5 A5 and PSL2 9 A6. Covering groups and automorphism groups are briefly mentioned. The other families of classical groups are defined in terms of certain ‘forms’ on the vector space, so we collect salient facts about these forms in Section 2.4, before introducing the symplectic groups in Section 2.5. These are in many ways the easiest of the classical groups to understand. We calculate their orders, prove simplicity, discuss subgroups, covering groups and automorphisms, and prove the generic isomorphism ¤ ¦ ¤ ¤ ¦ ¨ ¨ £ £ Sp2 £ q SL2 q and the exceptional isomorphism Sp4 2 S6. Next we treat the unitary groups in much the same way. Most of the difficulties in studying the classical groups occur in the orthogonal groups. The first problem is that there are fundamental differences between the case when the underlying field has characteristic 2, and the other cases. The second problem is that even when we take the subgroup of matrices with determinant 1, and factor out any scalars, the resulting group is not usually simple. To avoid introducing too many complications at once, we first treat the case of orthogonal groups over fields of odd order, defining the three different types and cal- culating their orders. We use the spinor norm to obtain the (usually) simple groups, and describe some of their subgroup structure. Then we briefly describe the differ- ences that occur when the field has characteristic 2, and use the quasideterminant to obtain the (usually) simple group. 19 20 CHAPTER 2. CLASSICAL GROUPS The generic isomorphisms of the orthogonal groups all derive from the Klein cor- ¤ ¦ ¤ ¤ ¦ ¤ ¨ Ω ¨ Ω £ £ ¡ £ £ respondence, by which P q PSL4 q (and also P q PSU4 q ). This corre- ¦ 6 6 ¤ ¨ spondence also leads to an easy proof of the exceptional isomorphisms PSL4 £ 2 A8 ¤ ¦ ¤ ¨ £ and PSL3 £ 2 PSL2 7 . We complete the chapter by proving a simple version of Aschbacher’s theorem which gives a basic classification of the different types of maximal subgroups in a classical group. This proof relies heavily on representation theory, however, and can be omitted by those readers without the necessary prerequisites. 2.2 Finite fields All the classical groups are defined in terms of groups of matrices over fields, so before we can define the finite classical groups we need to know what the finite fields are. A field is a set F with operations of addition, subtraction, multiplication and division sat- ¤ ¢¤£¥¢§¦¨¢ isfying the usual rules. That is, F has an element 0 such that £ F 0 is an abelian ¡ ¤ £ © ¢¢ ¢ group, and F © 0 contains an element 1 such that F 0 1 is an abelian group, ¨ ¤ £ £ and x £ y z xy xz. It is an easy exercise to show that the subfield F0 generated by the element 1 in a finite field F is isomorphic to the integers modulo p, for some p, and therefore p is prime (called the characteristic of the field). Moreover, F is a vector space over F0, as the vector space axioms are special cases of the field axioms. As every finite dimensional vector space has a basis of n vectors, v1, . , vn, say, and ∑n every vector has a unique expression a v with a F , it follows that the field F i 1 i i i 0 has pn elements. Conversely, for every prime p and every positive integer n there is a field of order pn, which is unique up to isomorphism. [See below.] The most important fact about finite fields which we need is that the set of all non-zero elements forms a multiplicative group, which is cyclic. For the polynomial ring F x over any field F is a Euclidean domain and therefore a unique factorization domain. In particular a polynomial of degree n has at most n roots. If the multiplicative e ¦ ¦ group F of a field of order q has exponent e strictly less than q 1, then x 1 has ¦ q ¦ 1 roots, which is a contradiction. Therefore the exponent of F is q 1, so F contains elements of order q ¦ 1, since it is abelian, and therefore it is cyclic. q ¨ q Note also that all elements x of F satisfy x x, and so the polynomial x ¦ x ¨ ¤ n ¦ α factorizes as ∏ £ x . Moreover, the number of solutions to x 1 in F is the α F ¤ ¢ ¦ ¦ greatest common divisor £ n q 1 of n and q 1. A useful consequence of this for 2 £ λ 2 the field of order q is that for every µ q there are exactly q 1 elements q ¨ λλ ¨ λ1 q satisfying µ. ¦ ¡ ¨ Observe that if f is an irreducible polynomial of degree d over the field p p ¤ ¡ d £ of order p, then p x f is a field of order p . Conversely, if F is a field of order d p , let x be a generator for F , so that the minimum polynomial for x over p is an ¦ ¡ ¤ ¨ £ irreducible polynomial f of degree d, and F p x f . One way to see that such ¨ a field exists is to observe that any field of order q pd is a splitting field for the q polynomial x ¦ x. Splitting fields always exist, by adjoining roots one at a time until q the polynomial factorises into linear factors. But then the set of roots of x ¦ x is closed ¨ ¨ ¨ ¨ q q ¤ q q q under addition and multiplication, since if x x and y y then £ xy x y xy and ¨ ¨ ¤ q q q £ £ £ £ x y x y x y. Hence this set of roots is a subfield of order q, as required. 2.3. THE GENERAL LINEAR GROUPS 21 ¨ To show that the field of order q pd does not depend on the particular irreducible ¨ ¡ ¤ £ polynomial we choose, suppose that f1 and f2 are two such, and Fi p x fi . Since ¤ q q ¦ ¦ f2 £ t divides t t, and t t factorizes into linear factors over F1, it follows that ¨ ¤ ¡ F1 contains an element y with f2 £ y 0. Hence the map x y extends to a field homomorphism from F2 to a subfield of F1. Moreover, the kernel is trivial, since fields have no quotient fields, so this map is a field isomorphism, since the fields are finite. ¨ ¨ ¨ If also f1 f2 then any automorphism of F F1 F2 has this form, so is defined by the image of x, which must be one of the d roots of f1. Hence the group of auto- ¡ p morphisms of F has order d. On the other hand, the map y y (for all y F) is an ¤ automorphism of F, and has order d. Hence Aut £ F is cyclic of order d. 2.3 The general linear groups Let V be a vector space of dimension n over the finite field q of order q. The general ¤ linear group GL £ V is the set of invertible linear maps from V to itself. Without much n loss of generality, we may take V as the vector space q of n-tuples of elements of q , ¤ ¤ £ ¢ and identify GL £ V with the group (denoted GLn q ) of invertible n n matrices over q . ¨ ¤ There are certain obvious normal subgroups of G GLn £ q . For example, the ¨ λ £ λ centre, Z say, consists of all the scalar matrices In, where 0 q and In is the ¦ n ¢ n identity matrix. Thus Z is a cyclic normal subgroup of order q 1. The quotient ¡ ¤ G Z is called the projective general linear group, and denoted PGLn £ q . ¨ ¤ ¤ ¤ £ £ Also, since det £ AB det A det B , the determinant map is a group homomor- ¤ phism from GLn £ q onto the multiplicative group of the field, so its kernel is a normal ¤ £ subgroup of index q ¦ 1. This kernel is called the special linear group SLn q , and ¤ consists of all the matrices of determinant 1. Similarly, we can quotient SLn £ q by the ¤ subgroup of scalars it contains, to obtain the projective special linear group PSLn £ q , ¤ sometimes abbreviated to Ln £ q . Now an invertible matrix takes a basis to a basis, and is determined by the image of an ordered basis. The only condition on this image is that the ith vector is linearly independent of the previous ones—but these span a space of dimension i ¦ 1, which ¤ ¤ ¤¥¤¦¤¦¤ ¤ i 1 ¤ n n n 2 n n 1 £ £ ¦ £ ¦ £ ¦ £ ¦ ¡ has q ¡ vectors in it, so the order of GLn q is q 1 q q q q q q , ¤ ¤¥¤¦¤¦¤ ¤ ¨ © n § n 1 2 2 n £ ¦ £ ¦ £ ¦ which can also be expressed as q ¡ q 1 q 1 q 1 . ¤ ¤ ¤ £ £ ¦ The orders of SLn £ q and PGLn q are equal, being GLn q divided by q 1.
Recommended publications
  • 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]
  • View This Volume's Front and Back Matter
    http://dx.doi.org/10.1090/gsm/039 Selected Titles in This Series 39 Larry C. Grove, Classical groups and geometric algebra, 2002 38 Elton P. Hsu, Stochastic analysis on manifolds, 2001 37 Hershel M. Farkas and Irwin Kra, Theta constants, Riemann surfaces and the modular group, 2001 36 Martin Schechter, Principles of functional analysis, second edition, 2001 35 James F. Davis and Paul Kirk, Lecture notes in algebraic topology, 2001 34 Sigurdur Helgason, Differential geometry, Lie groups, and symmetric spaces, 2001 33 Dmitri Burago, Yuri Burago, and Sergei Ivanov, A course in metric geometry, 2001 32 Robert G. Bartle, A modern theory of integration, 2001 31 Ralf Korn and Elke Korn, Option pricing and portfolio optimization: Modern methods of financial mathematics, 2001 30 J. C. McConnell and J. C. Robson, Noncommutative Noetherian rings, 2001 29 Javier Duoandikoetxea, Fourier analysis, 2001 28 Liviu I. Nicolaescu, Notes on Seiberg-Witten theory, 2000 27 Thierry Aubin, A course in differential geometry, 2001 26 Rolf Berndt, An introduction to symplectic geometry, 2001 25 Thomas Friedrich, Dirac operators in Riemannian geometry, 2000 24 Helmut Koch, Number theory: Algebraic numbers and functions, 2000 23 Alberto Candel and Lawrence Conlon, Foliations I, 2000 22 Giinter R. Krause and Thomas H. Lenagan, Growth of algebras and Gelfand-Kirillov dimension, 2000 21 John B. Conway, A course in operator theory, 2000 20 Robert E. Gompf and Andras I. Stipsicz, 4-manifolds and Kirby calculus, 1999 19 Lawrence C. Evans, Partial differential equations, 1998 18 Winfried Just and Martin Weese, Discovering modern set theory. II: Set-theoretic tools for every mathematician, 1997 17 Henryk Iwaniec, Topics in classical automorphic forms, 1997 16 Richard V.
    [Show full text]
  • [Math.GR] 9 Jul 2003 Buildings and Classical Groups
    Buildings and Classical Groups Linus Kramer∗ Mathematisches Institut, Universit¨at W¨urzburg Am Hubland, D–97074 W¨urzburg, Germany email: [email protected] In these notes we describe the classical groups, that is, the linear groups and the orthogonal, symplectic, and unitary groups, acting on finite dimen- sional vector spaces over skew fields, as well as their pseudo-quadratic gen- eralizations. Each such group corresponds in a natural way to a point-line geometry, and to a spherical building. The geometries in question are pro- jective spaces and polar spaces. We emphasize in particular the rˆole played by root elations and the groups generated by these elations. The root ela- tions reflect — via their commutator relations — algebraic properties of the underlying vector space. We also discuss some related algebraic topics: the classical groups as per- mutation groups and the associated simple groups. I have included some remarks on K-theory, which might be interesting for applications. The first K-group measures the difference between the classical group and its subgroup generated by the root elations. The second K-group is a kind of fundamental group of the group generated by the root elations and is related to central extensions. I also included some material on Moufang sets, since this is an in- arXiv:math/0307117v1 [math.GR] 9 Jul 2003 teresting topic. In this context, the projective line over a skew field is treated in some detail, and possibly with some new results. The theory of unitary groups is developed along the lines of Hahn & O’Meara [15].
    [Show full text]
  • Hartley's Theorem on Representations of the General Linear Groups And
    Turk J Math 31 (2007) , Suppl, 211 – 225. c TUB¨ ITAK˙ Hartley’s Theorem on Representations of the General Linear Groups and Classical Groups A. E. Zalesski To the memory of Brian Hartley Abstract We suggest a new proof of Hartley’s theorem on representations of the general linear groups GLn(K)whereK is a field. Let H be a subgroup of GLn(K)andE the natural GLn(K)-module. Suppose that the restriction E|H of E to H contains aregularKH-module. The theorem asserts that this is then true for an arbitrary GLn(K)-module M provided dim M>1andH is not of exponent 2. Our proof is based on the general facts of representation theory of algebraic groups. In addition, we provide partial generalizations of Hartley’s theorem to other classical groups. Key Words: subgroups of classical groups, representation theory of algebraic groups 1. Introduction In 1986 Brian Hartley [4] obtained the following interesting result: Theorem 1.1 Let K be a field, E the standard GLn(K)-module, and let M be an irre- ducible finite-dimensional GLn(K)-module over K with dim M>1. For a finite subgroup H ⊂ GLn(K) suppose that the restriction of E to H contains a regular submodule, that ∼ is, E = KH ⊕ E1 where E1 is a KH-module. Then M contains a free KH-submodule, unless H is an elementary abelian 2-groups. His proof is based on deep properties of the duality between irreducible representations of the general linear group GLn(K)and the symmetric group Sn.
    [Show full text]
  • 14.2 Covers of Symmetric and Alternating Groups
    Remark 206 In terms of Galois cohomology, there an exact sequence of alge- braic groups (over an algebrically closed field) 1 → GL1 → ΓV → OV → 1 We do not necessarily get an exact sequence when taking values in some subfield. If 1 → A → B → C → 1 is exact, 1 → A(K) → B(K) → C(K) is exact, but the map on the right need not be surjective. Instead what we get is 1 → H0(Gal(K¯ /K), A) → H0(Gal(K¯ /K), B) → H0(Gal(K¯ /K), C) → → H1(Gal(K¯ /K), A) → ··· 1 1 It turns out that H (Gal(K¯ /K), GL1) = 1. However, H (Gal(K¯ /K), ±1) = K×/(K×)2. So from 1 → GL1 → ΓV → OV → 1 we get × 1 1 → K → ΓV (K) → OV (K) → 1 = H (Gal(K¯ /K), GL1) However, taking 1 → µ2 → SpinV → SOV → 1 we get N × × 2 1 ¯ 1 → ±1 → SpinV (K) → SOV (K) −→ K /(K ) = H (K/K, µ2) so the non-surjectivity of N is some kind of higher Galois cohomology. Warning 207 SpinV → SOV is onto as a map of ALGEBRAIC GROUPS, but SpinV (K) → SOV (K) need NOT be onto. Example 208 Take O3(R) =∼ SO3(R)×{±1} as 3 is odd (in general O2n+1(R) =∼ SO2n+1(R) × {±1}). So we have a sequence 1 → ±1 → Spin3(R) → SO3(R) → 1. 0 × Notice that Spin3(R) ⊆ C3 (R) =∼ H, so Spin3(R) ⊆ H , and in fact we saw that it is S3. 14.2 Covers of symmetric and alternating groups The symmetric group on n letter can be embedded in the obvious way in On(R) as permutations of coordinates.
    [Show full text]
  • Spacetime May Have Fractal Properties on a Quantum Scale 25 March 2009, by Lisa Zyga
    Spacetime May Have Fractal Properties on a Quantum Scale 25 March 2009, By Lisa Zyga values at short scales. More than being just an interesting idea, this phenomenon might provide insight into a quantum theory of relativity, which also has been suggested to have scale-dependent dimensions. Benedetti’s study is published in a recent issue of Physical Review Letters. “It is an old idea in quantum gravity that at short scales spacetime might appear foamy, fuzzy, fractal or similar,” Benedetti told PhysOrg.com. “In my work, I suggest that quantum groups are a valid candidate for the description of such a quantum As scale decreases, the number of dimensions of k- spacetime. Furthermore, computing the spectral Minkowski spacetime (red line), which is an example of a dimension, I provide for the first time a link between space with quantum group symmetry, decreases from four to three. In contrast, classical Minkowski spacetime quantum groups/noncommutative geometries and (blue line) is four-dimensional on all scales. This finding apparently unrelated approaches to quantum suggests that quantum groups are a valid candidate for gravity, such as Causal Dynamical Triangulations the description of a quantum spacetime, and may have and Exact Renormalization Group. And establishing connections with a theory of quantum gravity. Image links between different topics is often one of the credit: Dario Benedetti. best ways we have to understand such topics.” In his study, Benedetti explains that a spacetime with quantum group symmetry has in general a (PhysOrg.com) -- Usually, we think of spacetime scale-dependent dimension. Unlike classical as being four-dimensional, with three dimensions groups, which act on commutative spaces, of space and one dimension of time.
    [Show full text]
  • Rotation Matrix - Wikipedia, the Free Encyclopedia Page 1 of 22
    Rotation matrix - Wikipedia, the free encyclopedia Page 1 of 22 Rotation matrix From Wikipedia, the free encyclopedia In linear algebra, a rotation matrix is a matrix that is used to perform a rotation in Euclidean space. For example the matrix rotates points in the xy -Cartesian plane counterclockwise through an angle θ about the origin of the Cartesian coordinate system. To perform the rotation, the position of each point must be represented by a column vector v, containing the coordinates of the point. A rotated vector is obtained by using the matrix multiplication Rv (see below for details). In two and three dimensions, rotation matrices are among the simplest algebraic descriptions of rotations, and are used extensively for computations in geometry, physics, and computer graphics. Though most applications involve rotations in two or three dimensions, rotation matrices can be defined for n-dimensional space. Rotation matrices are always square, with real entries. Algebraically, a rotation matrix in n-dimensions is a n × n special orthogonal matrix, i.e. an orthogonal matrix whose determinant is 1: . The set of all rotation matrices forms a group, known as the rotation group or the special orthogonal group. It is a subset of the orthogonal group, which includes reflections and consists of all orthogonal matrices with determinant 1 or -1, and of the special linear group, which includes all volume-preserving transformations and consists of matrices with determinant 1. Contents 1 Rotations in two dimensions 1.1 Non-standard orientation
    [Show full text]
  • Arxiv:1709.02742V3
    PIN GROUPS IN GENERAL RELATIVITY BAS JANSSENS Abstract. There are eight possible Pin groups that can be used to describe the transformation behaviour of fermions under parity and time reversal. We show that only two of these are compatible with general relativity, in the sense that the configuration space of fermions coupled to gravity transforms appropriately under the space-time diffeomorphism group. 1. Introduction For bosons, the space-time transformation behaviour is governed by the Lorentz group O(3, 1), which comprises four connected components. Rotations and boosts are contained in the connected component of unity, the proper orthochronous Lorentz group SO↑(3, 1). Parity (P ) and time reversal (T ) are encoded in the other three connected components of the Lorentz group, the translates of SO↑(3, 1) by P , T and PT . For fermions, the space-time transformation behaviour is governed by a double cover of O(3, 1). Rotations and boosts are described by the unique simply connected double cover of SO↑(3, 1), the spin group Spin↑(3, 1). However, in order to account for parity and time reversal, one needs to extend this cover from SO↑(3, 1) to the full Lorentz group O(3, 1). This extension is by no means unique. There are no less than eight distinct double covers of O(3, 1) that agree with Spin↑(3, 1) over SO↑(3, 1). They are the Pin groups abc Pin , characterised by the property that the elements ΛP and ΛT covering P and 2 2 2 T satisfy ΛP = −a, ΛT = b and (ΛP ΛT ) = −c, where a, b and c are either 1 or −1 (cf.
    [Show full text]
  • Dual Pairs in the Pin-Group and Duality for the Corresponding Spinorial
    DUAL PAIRS IN THE PIN-GROUP AND DUALITY FOR THE CORRESPONDING SPINORIAL REPRESENTATION CLÉMENT GUÉRIN, GANG LIU, AND ALLAN MERINO Abstract. In this paper, we give a complete picture of Howe correspondence for the setting (O(E, b), Pin(E, b), Π), where O(E, b) is an orthogonal group (real or complex), Pin(E, b) is the two-fold Pin-covering of O(E, b), and Π is the spinorial representation of Pin(E, b). More pre- cisely, for a dual pair (G, G′) in O(E, b), we determine explicitly the nature of its preimages (G, G′) in Pin(E, b), and prove that apart from some exceptions, (G, G′) is always a dual pair in e f e f Pin(E, b); then we establish the Howe correspondence for Π with respect to (G, G′). e f Contents 1. Introduction 1 2. Preliminaries 3 3. Dual pairs in the Pin-group 6 3.1. Pull-back of dual pairs 6 3.2. Identifyingtheisomorphismclassofpull-backs 8 4. Duality for the spinorial representation 14 References 21 1. Introduction The first duality phenomenon has been discovered by H. Weyl who pointed out a correspon- dence between some irreducible finite dimensional representations of the general linear group GL(V) and the symmetric group Sk where V is a finite dimensional vector space over C. In- d deed, considering the joint action of GL(V) and Sd on the space V⊗ , we get the following decomposition: arXiv:1907.09093v1 [math.RT] 22 Jul 2019 d λ V⊗ = M λ σ , [ ⊗ (Vλ,λ) GL(V) ∈ π [ where GL(V)π is the set of equivalence classes of irreducible finite dimensional representations d of GL(V) such that Hom (V , V⊗ ) , 0 and σ is an irreducible representation of S .
    [Show full text]
  • Algebraic Topology
    Algebraic Topology Len Evens Rob Thompson Northwestern University City University of New York Contents Chapter 1. Introduction 5 1. Introduction 5 2. Point Set Topology, Brief Review 7 Chapter 2. Homotopy and the Fundamental Group 11 1. Homotopy 11 2. The Fundamental Group 12 3. Homotopy Equivalence 18 4. Categories and Functors 20 5. The fundamental group of S1 22 6. Some Applications 25 Chapter 3. Quotient Spaces and Covering Spaces 33 1. The Quotient Topology 33 2. Covering Spaces 40 3. Action of the Fundamental Group on Covering Spaces 44 4. Existence of Coverings and the Covering Group 48 5. Covering Groups 56 Chapter 4. Group Theory and the Seifert{Van Kampen Theorem 59 1. Some Group Theory 59 2. The Seifert{Van Kampen Theorem 66 Chapter 5. Manifolds and Surfaces 73 1. Manifolds and Surfaces 73 2. Outline of the Proof of the Classification Theorem 80 3. Some Remarks about Higher Dimensional Manifolds 83 4. An Introduction to Knot Theory 84 Chapter 6. Singular Homology 91 1. Homology, Introduction 91 2. Singular Homology 94 3. Properties of Singular Homology 100 4. The Exact Homology Sequence{ the Jill Clayburgh Lemma 109 5. Excision and Applications 116 6. Proof of the Excision Axiom 120 3 4 CONTENTS 7. Relation between π1 and H1 126 8. The Mayer-Vietoris Sequence 128 9. Some Important Applications 131 Chapter 7. Simplicial Complexes 137 1. Simplicial Complexes 137 2. Abstract Simplicial Complexes 141 3. Homology of Simplicial Complexes 143 4. The Relation of Simplicial to Singular Homology 147 5. Some Algebra. The Tensor Product 152 6.
    [Show full text]
  • IMAGE of FUNCTORIALITY for GENERAL SPIN GROUPS 11 Splits As Two Places in K Or V Is Inert, I.E., There Is a Single Place in K Above the Place V in K
    IMAGE OF FUNCTORIALITY FOR GENERAL SPIN GROUPS MAHDI ASGARI AND FREYDOON SHAHIDI Abstract. We give a complete description of the image of the endoscopic functorial transfer of generic automorphic representations from the quasi-split general spin groups to general linear groups over arbitrary number fields. This result is not covered by the recent work of Arthur on endoscopic classification of automorphic representations of classical groups. The image is expected to be the same for the whole tempered spectrum, whether generic or not, once the transfer for all tempered representations is proved. We give a number of applications including estimates toward the Ramanujan conjecture for the groups involved and the characterization of automorphic representations of GL(6) which are exterior square transfers from GL(4); among others. More applications to reducibility questions for the local induced representations of p-adic groups will also follow. 1. Introduction The purpose of this article is to completely determine the image of the transfer of globally generic, automorphic representations from the quasi-split general spin groups to the general linear groups. We prove the image is an isobaric, automorphic representation of a general linear group. Moreover, we prove that the isobaric summands of this representation have the property that their twisted symmetric square or twisted exterior square L-function has a pole at s = 1; depending on whether the transfer is from an even or an odd general spin group. We are also able to determine whether the image is a transfer from a split group, or a quasi-split, non-split group. Arthur's recent book on endoscopic classification of representations of classical groups [Ar] establishes similar results for automorphic representations of orthogonal and symplectic groups, generic or not, but does not cover the case of the general spin groups we consider here.
    [Show full text]
  • 1 the Spin Homomorphism SL2(C) → SO1,3(R) a Summary from Multiple Sources Written by Y
    1 The spin homomorphism SL2(C) ! SO1;3(R) A summary from multiple sources written by Y. Feng and Katherine E. Stange Abstract We will discuss the spin homomorphism SL2(C) ! SO1;3(R) in three manners. Firstly we interpret SL2(C) as acting on the Minkowski 1;3 spacetime R ; secondly by viewing the quadratic form as a twisted 1 1 P × P ; and finally using Clifford groups. 1.1 Introduction The spin homomorphism SL2(C) ! SO1;3(R) is a homomorphism of classical matrix Lie groups. The lefthand group con- sists of 2 × 2 complex matrices with determinant 1. The righthand group consists of 4 × 4 real matrices with determinant 1 which preserve some fixed real quadratic form Q of signature (1; 3). This map is alternately called the spinor map and variations. The image of this map is the identity component + of SO1;3(R), denoted SO1;3(R). The kernel is {±Ig. Therefore, we obtain an isomorphism + PSL2(C) = SL2(C)= ± I ' SO1;3(R): This is one of a family of isomorphisms of Lie groups called exceptional iso- morphisms. In Section 1.3, we give the spin homomorphism explicitly, al- though these formulae are unenlightening by themselves. In Section 1.4 we describe O1;3(R) in greater detail as the group of Lorentz transformations. This document describes this homomorphism from three distinct per- spectives. The first is very concrete, and constructs, using the language of Minkowski space, Lorentz transformations and Hermitian matrices, an ex- 4 plicit action of SL2(C) on R preserving Q (Section 1.5).
    [Show full text]