Complexification, Complex Structures, and Linear Ordinary Differential

Total Page:16

File Type:pdf, Size:1020Kb

Complexification, Complex Structures, and Linear Ordinary Differential Complexification, complex structures, and linear ordinary differential equations Jordan Bell [email protected] Department of Mathematics, University of Toronto April 3, 2014 1 Motivation The solution of the initial value problem 0 n x (t) = Ax(t); x(0) = x0 2 R ; where A is an n × n matrix over R, is x(t) = exp(At)x0. If we want to compute the solution and if A is diagonalizable, say A = P DP −1, we use exp(At) = exp((P DP −1)t) = P exp(Dt)P −1: n Thus if the matrix A has complex eigenvalues, then although exp(At)x0 2 R , 0 −1 it may not be the case that P −1x 2 n. For example, if A = , then 0 R 1 0 −i 0 −i i 1 i 1 D = ;P = ;P −1 = : 0 i 1 1 2 −i 1 1 For x = , 0 0 1 i 1 1 1 i P −1x = = : 0 2 −i 1 0 2 −i This is similar to how Cardano's formula, which expresses the roots of a real cubic polynomial in terms of its coefficients, involves complex numbers and yet the final result may still be real. In the following, unless I specify the dimension of a vector space, any state- ment about real vector spaces is about real vector spaces of finite or infinite dimension, and any statement about complex vector spaces is about complex vector spaces of finite or infinite dimension. 1 2 Complexification 2.1 Direct sums If V is a real vector space, a complex structure for V is an R-linear map J : 2 V ! V such that J = −idV . If V is a real vector space and J : V ! V is a complex structure, define a complex vector space VJ in the following way: let the set of elements of VJ be V , let addition in VJ be addition in V , and define scalar multiplication in VJ by (a + ib)v = av + bJ(v): One checks that for α; β 2 C and v 2 VJ we have (αβ)v = α(βv), and thus that 1 VJ is indeed a complex vector space with this definition of scalar multiplication. Let V be a real vector space, and define the R-linear map J : V ⊕V ! V ⊕V by J(v; w) = (−w; v): 2 J = −idV ⊕V . J is a complex structure on the real vector space V ⊕ V . The complexification of V is the complex vector space V C = (V ⊕ V )J . Thus, V C has the same set of elements as V ⊕ V , the same addition as V ⊕ V , and scalar multiplication (a + ib)(v; w) = a(v; w) + bJ(v; w); which gives (a + ib)(v; w) = a(v; w) + b(−w; v) = (av; aw) + (−bw; bv) = (av − bw; aw + bv): If the real vector space V has dimension n and if fe1; : : : ; eng is a basis for V , then f(e1; 0);:::; (en; 0); (0; e1);:::; (0; en)g is a basis for the real vector space V ⊕ V . Let v 2 V C. Using the basis for the real vector space V ⊕ V , there exist a1; : : : ; an; b1; : : : ; bn 2 R such that v = a1(e1; 0) + ··· an(en; 0) + b1(0; e1) + ··· + bn(0; en) = a1(e1; 0) + ··· + an(en; 0) + b1J(e1; 0) + ··· + bnJ(en; 0) = (a1 + ib1)(e1; 0) + ··· + (an + ibn)(en; 0); where in the last line we used the definition of scalar multiplication in V C. One checks that the set f(e1; 0);:::; (en; 0)g is linearly independent over C, and therefore it is a basis for V C. Hence C dimC V = dimR V: 1One should also verify that distributivity holds with this definition of scalar product; the other properties of a vector space are satisfied because VJ has the same addition as the real vector space V . 2 2.2 Complexification is a functor If V; W are real vector spaces and T : V ! W is an R-linear map, we define T C : V C ! W C by T C(v1; v2) = (T v1; T v2); this is a C-linear map. Setting ιV (v1; v2) = (v1; 0) and ιW (w1; w2) = (w1; 0), T C : V C ! W C is the unique C-linear map such that the following diagram commutes2 C V C T W C ιV ιW V T W Complexification is a functor from the category of real vector spaces to the category of complex vector spaces: C (idV ) (v1; v2) = (idV v1; idV v2) = (v1; v2) = idV C (v1; v2); C so (idV ) = idV C , and if S : U ! V and T : V ! S are R-linear maps, then (T ◦ S)C(v1; v2) = (T (Sv1);T (Sv2)) = T C(Sv1; Sv2) = T C(SC(v1; v2)); so (T ◦ S)C = T C ◦ SC. 2.3 Complexifying a complex structure If V is a real vector space and J : V ! V is a complex structure, then 2 (J C) (v1; v2) = J C(Jv1; Jv2) 2 2 = (J v1;J v2) = (−v1; −v2) = −(v1; v2); C 2 so (J ) = −idV C . Let Ei = fw 2 V C : J Cw = iwg;E−i = fw 2 V C : J Cw = −iwg: If w 2 V C, then one checks that w − iJ Cw 2 Ei; w + iJ Cw 2 E−i; and 1 1 w = w − iJ Cw + w + iJ Cw : 2 2 It follows that V C = Ei ⊕ E−i: 2See Keith Conrad's note: http://www.math.uconn.edu/∼kconrad/blurbs/linmultialg/complexification.pdf 3 2.4 Complex structures, inner products, and symplectic forms If V is a real vector space of odd dimension, then one can show that there is no 2 linear map J : V ! V satisfying J = −idV , i.e. there does not exist a complex structure for it. On the other hand, if V has even dimension, let fe1; : : : ; en; f1; : : : ; fng be a basis for the real vector space V , and define J : V ! V by Jej = fj; Jfj = −ej: Then J : V ! V is a complex structure. If V is a real vector space of dimension 2n with a complex structure J, let e1 6= 0. Check that Je1 62 spanfe1g. If n > 1, let e2 62 spanfe1; Je1g: Check that the set fe1; e2; Je1; Je2g is linearly independent. If n > 2, let e3 62 spanfe1; e2; Je1; Je2g: Check that the set fe1; e2; e3; Je1; Je2; Je3g is linearly independent. If 2i < 2n then there is some ei+1 62 spanfe1; : : : ; ei; Je1; : : : ; Jeig: I assert that fe1; : : : ; en; Je1; : : : ; Jeng is a basis for V . Using the above basis fe1; : : : ; en; Je1; : : : ; Jeng for V , let fi = Jei, and define h·; ·i : V × V ! R by hei; eji = δi;j; hfi; fji = δi;j; hei; fji = 0; hfi; eji = 0: Check that this is an inner product on the real vector space V . Moreover, hJei; Jeji = hfi; fji = δi;j = hei; eji ; and 2 2 hJfi; Jfji = J ei;J ej = h−ei; −eji = hei; eji = δi;j = hfi; fji ; and hJei; Jfji = hfi; −eji = − hfi; eji = 0 = hei; fji ; and hJfi; Jeji = h−ei; fji = − hei; fji = 0 = hfi; eji : 4 Hence for any v; w 2 V , hJv; Jwi = hv; wi : We say that the complex structure J is compatible with the inner product h·; ·i, i.e. J :(V; h·; ·i) ! (V; h·; ·i) is an orthogonal transformation. A symplectic form on a real vector space V is a bilinear form ! : V × V ! R such that !(v; w) = −!(w; v), and such that if !(v; w) = 0 for all w then v = 0; we say respectively that ! is skew-symmetric and non-degenerate. If a real vector space V has a complex structure J, and h·; ·i is an inner product on V that is compatible with J, define ! by !(v; w) = v; J −1w = hv; −Jwi = − hv; Jwi ; which is equivalent to !(v; Jw) = hv; wi : Using that the inner product is compatible with J and that it is symmetric, !(v; w) = − hv; Jwi = − Jv; J 2w = − hJv; −wi = hw; Jvi = −!(w; v); so ! is skew-symmetric. If w 2 V and !(v; w) = 0 for all v 2 V , then − hv; Jwi = 0 for all v 2 V , and thus Jw = 0. Since J is invertible, w = 0. Thus ! is nondegenerate. Therefore ! is a symplectic form on V .3 We have !(Jv; Jw) = − Jv; J 2w = − J 2v; J 3w = − h−v; −Jwi = − hv; Jwi = !(v; w): We say that J is compatible with the sympletic form !, namely, J :(V; !) ! (V; !) is a symplectic transformation. On the other hand, if V is a real vector space with symplectic form ! and J is a compatible complex structure, then h·; ·i : V × V ! R defined by hv; wi = !(v; Jw) is an inner product on V that is compatible with the complex structure J. Suppose V is a real vector space with complex structure J : V ! V and that h : VJ × VJ ! C is an inner product on the complex vector space VJ . Define g : V × V ! R by4 1 1 g(v ; v ) = h(v ; v ) + h(v ; v ) = (h(v ; v ) + h(v ; v )) : 1 2 2 1 2 1 2 2 1 2 2 1 3 Using the basis fe1; : : : ; en; f1; : : : ; fng for V , fi = Jei, we have 2 !(ei; fj ) = − hei; Jfj i = − ei;J ej = − hei; −ej i = hei; ej i = δi;j ; and !(ei; ej ) = − hei; Jej i = − hei; fj i = 0;!(fi; fj ) = 0: A basis fe1; : : : ; en; f1; : : : ; fng for a symplectic vector space that satisfies these three condi- tions is called a Darboux basis.
Recommended publications
  • Math 651 Homework 1 - Algebras and Groups Due 2/22/2013
    Math 651 Homework 1 - Algebras and Groups Due 2/22/2013 1) Consider the Lie Group SU(2), the group of 2 × 2 complex matrices A T with A A = I and det(A) = 1. The underlying set is z −w jzj2 + jwj2 = 1 (1) w z with the standard S3 topology. The usual basis for su(2) is 0 i 0 −1 i 0 X = Y = Z = (2) i 0 1 0 0 −i (which are each i times the Pauli matrices: X = iσx, etc.). a) Show this is the algebra of purely imaginary quaternions, under the commutator bracket. b) Extend X to a left-invariant field and Y to a right-invariant field, and show by computation that the Lie bracket between them is zero. c) Extending X, Y , Z to global left-invariant vector fields, give SU(2) the metric g(X; X) = g(Y; Y ) = g(Z; Z) = 1 and all other inner products zero. Show this is a bi-invariant metric. d) Pick > 0 and set g(X; X) = 2, leaving g(Y; Y ) = g(Z; Z) = 1. Show this is left-invariant but not bi-invariant. p 2) The realification of an n × n complex matrix A + −1B is its assignment it to the 2n × 2n matrix A −B (3) BA Any n × n quaternionic matrix can be written A + Bk where A and B are complex matrices. Its complexification is the 2n × 2n complex matrix A −B (4) B A a) Show that the realification of complex matrices and complexifica- tion of quaternionic matrices are algebra homomorphisms.
    [Show full text]
  • Symplectic Linear Algebra Let V Be a Real Vector Space. Definition 1.1. A
    INTRODUCTION TO SYMPLECTIC TOPOLOGY C. VITERBO 1. Lecture 1: Symplectic linear algebra Let V be a real vector space. Definition 1.1. A symplectic form on V is a skew-symmetric bilinear nondegen- erate form: (1) ω(x, y)= ω(y, x) (= ω(x, x) = 0); (2) x, y such− that ω(x,⇒ y) = 0. ∀ ∃ % For a general 2-form ω on a vector space, V , we denote ker(ω) the subspace given by ker(ω)= v V w Vω(v, w)=0 { ∈ |∀ ∈ } The second condition implies that ker(ω) reduces to zero, so when ω is symplectic, there are no “preferred directions” in V . There are special types of subspaces in symplectic manifolds. For a vector sub- space F , we denote by F ω = v V w F,ω(v, w) = 0 { ∈ |∀ ∈ } From Grassmann’s formula it follows that dim(F ω)=codim(F ) = dim(V ) dim(F ) Also we have − Proposition 1.2. (F ω)ω = F (F + F )ω = F ω F ω 1 2 1 ∩ 2 Definition 1.3. A subspace F of V,ω is ω isotropic if F F ( ω F = 0); • coisotropic if F⊂ω F⇐⇒ | • Lagrangian if it is⊂ maximal isotropic. • Proposition 1.4. (1) Any symplectic vector space has even dimension (2) Any isotropic subspace is contained in a Lagrangian subspace and Lagrangians have dimension equal to half the dimension of the total space. (3) If (V1,ω1), (V2,ω2) are symplectic vector spaces with L1,L2 Lagrangian subspaces, and if dim(V1) = dim(V2), then there is a linear isomorphism ϕ : V V such that ϕ∗ω = ω and ϕ(L )=L .
    [Show full text]
  • CLIFFORD ALGEBRAS and THEIR REPRESENTATIONS Introduction
    CLIFFORD ALGEBRAS AND THEIR REPRESENTATIONS Andrzej Trautman, Uniwersytet Warszawski, Warszawa, Poland Article accepted for publication in the Encyclopedia of Mathematical Physics c 2005 Elsevier Science Ltd ! Introduction Introductory and historical remarks Clifford (1878) introduced his ‘geometric algebras’ as a generalization of Grassmann alge- bras, complex numbers and quaternions. Lipschitz (1886) was the first to define groups constructed from ‘Clifford numbers’ and use them to represent rotations in a Euclidean ´ space. E. Cartan discovered representations of the Lie algebras son(C) and son(R), n > 2, that do not lift to representations of the orthogonal groups. In physics, Clifford algebras and spinors appear for the first time in Pauli’s nonrelativistic theory of the ‘magnetic elec- tron’. Dirac (1928), in his work on the relativistic wave equation of the electron, introduced matrices that provide a representation of the Clifford algebra of Minkowski space. Brauer and Weyl (1935) connected the Clifford and Dirac ideas with Cartan’s spinorial represen- tations of Lie algebras; they found, in any number of dimensions, the spinorial, projective representations of the orthogonal groups. Clifford algebras and spinors are implicit in Euclid’s solution of the Pythagorean equation x2 y2 + z2 = 0 which is equivalent to − y x z p = 2 p q (1) −z y + x q ! " ! " # $ so that x = q2 p2, y = p2 + q2, z = 2pq. If the numbers appearing in (1) are real, then − this equation can be interpreted as providing a representation of a vector (x, y, z) R3, null ∈ with respect to a quadratic form of signature (1, 2), as the ‘square’ of a spinor (p, q) R2.
    [Show full text]
  • Notes and Solutions to Exercises for Mac Lane's Categories for The
    Stefan Dawydiak Version 0.3 July 2, 2020 Notes and Exercises from Categories for the Working Mathematician Contents 0 Preface 2 1 Categories, Functors, and Natural Transformations 2 1.1 Functors . .2 1.2 Natural Transformations . .4 1.3 Monics, Epis, and Zeros . .5 2 Constructions on Categories 6 2.1 Products of Categories . .6 2.2 Functor categories . .6 2.2.1 The Interchange Law . .8 2.3 The Category of All Categories . .8 2.4 Comma Categories . 11 2.5 Graphs and Free Categories . 12 2.6 Quotient Categories . 13 3 Universals and Limits 13 3.1 Universal Arrows . 13 3.2 The Yoneda Lemma . 14 3.2.1 Proof of the Yoneda Lemma . 14 3.3 Coproducts and Colimits . 16 3.4 Products and Limits . 18 3.4.1 The p-adic integers . 20 3.5 Categories with Finite Products . 21 3.6 Groups in Categories . 22 4 Adjoints 23 4.1 Adjunctions . 23 4.2 Examples of Adjoints . 24 4.3 Reflective Subcategories . 28 4.4 Equivalence of Categories . 30 4.5 Adjoints for Preorders . 32 4.5.1 Examples of Galois Connections . 32 4.6 Cartesian Closed Categories . 33 5 Limits 33 5.1 Creation of Limits . 33 5.2 Limits by Products and Equalizers . 34 5.3 Preservation of Limits . 35 5.4 Adjoints on Limits . 35 5.5 Freyd's adjoint functor theorem . 36 1 6 Chapter 6 38 7 Chapter 7 38 8 Abelian Categories 38 8.1 Additive Categories . 38 8.2 Abelian Categories . 38 8.3 Diagram Lemmas . 39 9 Special Limits 41 9.1 Interchange of Limits .
    [Show full text]
  • On the Complexification of the Classical Geometries And
    On the complexication of the classical geometries and exceptional numb ers April Intro duction The classical groups On R Spn R and GLn C app ear as the isometry groups of sp ecial geome tries on a real vector space In fact the orthogonal group On R represents the linear isomorphisms of arealvector space V of dimension nleaving invariant a p ositive denite and symmetric bilinear form g on V The symplectic group Spn R represents the isometry group of a real vector space of dimension n leaving invariant a nondegenerate skewsymmetric bilinear form on V Finally GLn C represents the linear isomorphisms of a real vector space V of dimension nleaving invariant a complex structure on V ie an endomorphism J V V satisfying J These three geometries On R Spn R and GLn C in GLn Rintersect even pairwise in the unitary group Un C Considering now the relativeversions of these geometries on a real manifold of dimension n leads to the notions of a Riemannian manifold an almostsymplectic manifold and an almostcomplex manifold A symplectic manifold however is an almostsymplectic manifold X ie X is a manifold and is a nondegenerate form on X so that the form is closed Similarly an almostcomplex manifold X J is called complex if the torsion tensor N J vanishes These three geometries intersect in the notion of a Kahler manifold In view of the imp ortance of the complexication of the real Lie groupsfor instance in the structure theory and representation theory of semisimple real Lie groupswe consider here the question on the underlying geometrical
    [Show full text]
  • Patterns of Maximally Entangled States Within the Algebra of Biquaternions
    J. Phys. Commun. 4 (2020) 055018 https://doi.org/10.1088/2399-6528/ab9506 PAPER Patterns of maximally entangled states within the algebra of OPEN ACCESS biquaternions RECEIVED 21 March 2020 Lidia Obojska REVISED 16 May 2020 Siedlce University of Natural Sciences and Humanities, ul. 3 Maja 54, 08-110 Siedlce, Poland ACCEPTED FOR PUBLICATION E-mail: [email protected] 20 May 2020 Keywords: bipartite entanglement, biquaternions, density matrix, mereology PUBLISHED 28 May 2020 Original content from this Abstract work may be used under This paper proposes a description of maximally entangled bipartite states within the algebra of the terms of the Creative Commons Attribution 4.0 biquaternions–Ä . We assume that a bipartite entanglement is created in a process of splitting licence. one particle into a pair of two indiscernible, yet not identical particles. As a result, we describe an Any further distribution of this work must maintain entangled correlation by the use of a division relation and obtain twelve forms of biquaternions, attribution to the author(s) and the title of representing pure maximally entangled states. Additionally, we obtain other patterns, describing the work, journal citation mixed entangled states. Finally, we show that there are no other maximally entangled states in Ä and DOI. than those presented in this work. 1. Introduction In 1935, Einstein, Podolski and Rosen described a strange phenomenon, in which in its pure state, one particle behaves like a pair of two indistinguishable, yet not identical, particles [1]. This phenomenon, called by Einstein ’a spooky action at a distance’, has been investigated by many authors, who tried to explain this strange correlation [2–9].
    [Show full text]
  • Majorana Spinors
    MAJORANA SPINORS JOSE´ FIGUEROA-O'FARRILL Contents 1. Complex, real and quaternionic representations 2 2. Some basis-dependent formulae 5 3. Clifford algebras and their spinors 6 4. Complex Clifford algebras and the Majorana condition 10 5. Examples 13 One dimension 13 Two dimensions 13 Three dimensions 14 Four dimensions 14 Six dimensions 15 Ten dimensions 16 Eleven dimensions 16 Twelve dimensions 16 ...and back! 16 Summary 19 References 19 These notes arose as an attempt to conceptualise the `symplectic Majorana{Weyl condition' in 5+1 dimensions; but have turned into a general discussion of spinors. Spinors play a crucial role in supersymmetry. Part of their versatility is that they come in many guises: `Dirac', `Majorana', `Weyl', `Majorana{Weyl', `symplectic Majorana', `symplectic Majorana{Weyl', and their `pseudo' counterparts. The tra- ditional physics approach to this topic is a mixed bag of tricks using disparate aspects of representation theory of finite groups. In these notes we will attempt to provide a uniform treatment based on the classification of Clifford algebras, a work dating back to the early 60s and all but ignored by the theoretical physics com- munity. Recent developments in superstring theory have made us re-examine the conditions for the existence of different kinds of spinors in spacetimes of arbitrary signature, and we believe that a discussion of this more uniform approach is timely and could be useful to the student meeting this topic for the first time or to the practitioner who has difficulty remembering the answer to questions like \when do symplectic Majorana{Weyl spinors exist?" The notes are organised as follows.
    [Show full text]
  • Arxiv:1005.0156V3
    FOUR PROBLEMS REGARDING REPRESENTABLE FUNCTORS G. MILITARU C Abstract. Let R, S be two rings, C an R-coring and RM the category of left C- C C comodules. The category Rep (RM, SM) of all representable functors RM→ S M is C shown to be equivalent to the opposite of the category RMS . For U an (S, R)-bimodule C we give necessary and sufficient conditions for the induction functor U ⊗R − : RM→ SM to be: a representable functor, an equivalence of categories, a separable or a Frobenius functor. The latter results generalize and unify the classical theorems of Morita for categories of modules over rings and the more recent theorems obtained by Brezinski, Caenepeel et al. for categories of comodules over corings. Introduction Let C be a category and V a variety of algebras in the sense of universal algebras. A functor F : C →V is called representable [1] if γ ◦ F : C → Set is representable in the classical sense, where γ : V → Set is the forgetful functor. Four general problems concerning representable functors have been identified: Problem A: Describe the category Rep (C, V) of all representable functors F : C→V. Problem B: Give a necessary and sufficient condition for a given functor F : C→V to be representable (possibly predefining the object of representability). Problem C: When is a composition of two representable functors a representable functor? Problem D: Give a necessary and sufficient condition for a representable functor F : C → V and for its left adjoint to be separable or Frobenius. The pioneer of studying problem A was Kan [10] who described all representable functors from semigroups to semigroups.
    [Show full text]
  • Maxwell's Equations
    Maxwell’s equations Daniel Henry Gottlieb August 1, 2004 Abstract We express Maxwell’s equations as a single equation, first using the divergence of a special type of matrix field to obtain the four current, and then the divergence of a special matrix to obtain the Electromagnetic field. These two equations give rise to a remarkable dual set of equations in which the operators become the matrices and the vectors become the fields. The decoupling of the equations into the wave equation is very simple and natural. The divergence of the stress energy tensor gives the Lorentz Law in a very natural way. We compare this approach to the related descriptions of Maxwell’s equations by biquaternions and Clifford algebras. 1 Introduction Maxwell’s equations have been expressed in many forms in the century and a half since their dis- covery. The original equations were 16 in number. The vector forms, written below, consist of 4 equations. The differential form versions consists of two equations; see [Misner, Thorne and Wheeler(1973); see equations 4.10, 4.11]. See also [Steven Parrott(1987) page 98 -100 ] The ap- plication of quaternions, and their complexification, the biquaternions, results in a version of one equation. William E. Baylis (1999) equation 3.8, is an example. In this work, we obtain one Maxwell equation, (10), representing the electromagnetic field as a matrix and the divergence as a vector multiplying the field matrix. But we also obtain a remarkable dual formulation of Maxwell’s equation, (15), wherein the operator is now the matrix and the field is now the vector.
    [Show full text]
  • Symmetries of the Space of Linear Symplectic Connections
    Symmetry, Integrability and Geometry: Methods and Applications SIGMA 13 (2017), 002, 30 pages Symmetries of the Space of Linear Symplectic Connections Daniel J.F. FOX Departamento de Matem´aticas del Area´ Industrial, Escuela T´ecnica Superior de Ingenier´ıa y Dise~noIndustrial, Universidad Polit´ecnica de Madrid, Ronda de Valencia 3, 28012 Madrid, Spain E-mail: [email protected] Received June 30, 2016, in final form January 07, 2017; Published online January 10, 2017 https://doi.org/10.3842/SIGMA.2017.002 Abstract. There is constructed a family of Lie algebras that act in a Hamiltonian way on the symplectic affine space of linear symplectic connections on a symplectic manifold. The associated equivariant moment map is a formal sum of the Cahen{Gutt moment map, the Ricci tensor, and a translational term. The critical points of a functional constructed from it interpolate between the equations for preferred symplectic connections and the equations for critical symplectic connections. The commutative algebra of formal sums of symmetric tensors on a symplectic manifold carries a pair of compatible Poisson structures, one induced from the canonical Poisson bracket on the space of functions on the cotangent bundle poly- nomial in the fibers, and the other induced from the algebraic fiberwise Schouten bracket on the symmetric algebra of each fiber of the cotangent bundle. These structures are shown to be compatible, and the required Lie algebras are constructed as central extensions of their linear combinations restricted to formal sums of symmetric tensors whose first order term is a multiple of the differential of its zeroth order term.
    [Show full text]
  • Monads and Algebras - an Introduction
    Monads and Algebras - An Introduction Uday S. Reddy April 27, 1995 1 1 Algebras of monads Consider a simple form of algebra, say, a set with a binary operation. Such an algebra is a pair hX, ∗ : X × X → Xi. Morphisms of these algebras preserve the binary operation: f(x ∗ y) = f(x) ∗ f(y). We first notice that the domain of the operation is determined by a functor F : Set → Set (the diagonal functor FX = X ×X). In general, given an endofunctor F : C → C on a category C, we can speak of the “algebras” for F , which are pairs hX, α : FX → Xi of an object X of C and an arrow α : FX → X. Morphism preserve the respective operations, i.e., F f FX - FY α β ? f ? X - Y Now, in algebra, it is commonplace to talk about “derived operators.” Such an operator is determined by a term made up of variables (over X) and the operations of the algebra. For example, for the simple algebra with a binary operation, terms such as x, x ∗ (y ∗ z), and (x ∗ y) ∗ (z ∗ w) determine derived operators. From a categorical point of view, such derived operators are just compositions of the standard operators and identity arrows: id X X- X id ×α α X × (X × X) X - X × X - X α×α α (X × X) × (X × X) - X × X - X To formalize such derived operators, we think of another functor T : C → C into which we can embed the domains of all the derived operators. In particular, we should have embeddings I →.
    [Show full text]
  • 1 Symplectic Vector Spaces
    Symplectic Manifolds Tam´asHausel 1 Symplectic Vector Spaces Let us first fix a field k. Definition 1.1 (Symplectic form, symplectic vector space). Given a finite dimensional k-vector space W , a symplectic form on W is a bilinear, alternating non degenerate form on W , i.e. an element ! 2 ^2W ∗ such that if !(v; w) = 0 for all w 2 W then v = 0. We call the pair (W; !) a symplectic vector space. Example 1.2 (Standard example). Consider a finite dimensional k-vector space V and define W = V ⊕ V ∗. We can introduce a symplectic form on W : ! : W × W ! k ((v1; α1); (v2; α2)) 7! α1(v2) − α2(v1) Definition 1.3 (Symplectomorphism). Given two symplectic vector spaces (W1;!1) and (W2;!2), we say that a linear map f : W1 ! W2 is a symplectomorphism if it is an isomorphism of vector ∗ spaces and, furthermore, f !2 = !1. We denote by Sp(W ) the group of symplectomorphism W ! W . Let us consider a symplectic vector space (W; !) with even dimension 2n (we will see soon that it is always the case). Then ! ^ · · · ^ ! 2 ^2nW ∗ is a volume form, i.e. a non-degenerate top form. For f 2 Sp(W ) we must have !n = f ∗!n = det f · !n so that det f = 1 and, in particular, Sp(W ) ≤ SL(W ). We also note that, for n = 1, we have Sp(W ) =∼ SL(W ). Definition 1.4. Let U be a linear subspace of a symplectic vector space W . Then we define U ? = fv 2 W j !(v; w) = 0 8u 2 Ug: We say that U is isotropic if U ⊆ U ?, coisotropic if U ? ⊆ U and Lagrangian if U ? = U.
    [Show full text]