Symplectic Vector Spaces, Lagrangian Subspaces, and Liouville's Theorem

Total Page:16

File Type:pdf, Size:1020Kb

Symplectic Vector Spaces, Lagrangian Subspaces, and Liouville's Theorem SYMPLECTIC VECTOR SPACES, LAGRANGIAN SUBSPACES, AND LIOUVILLE’S THEOREM CONNER JAGER Celestial Mechanics Junior Seminar - Nicolas Templier Abstract. The purpose of this paper is to prove Liouville’s theorem on volume-preserving flows from a symplectic linear algebra point of view. Along the way, we will prove a series of properties about symplectic vector spaces that elucidate the usefulness of this approach with regard to the correspondence between the configuration space and the symplectic phase space. Though only briefly discussed, this paper motivates how symplectic manifolds provide the natural abstraction of the phase space of a dynamical system through the simplified linear example: symplectic vector spaces. 0. Motivation In the study of celestial mechanics, there arises a natural connection between symplectic manifolds and the phase space of a dynamical system. Intuitively, we may think of the two defining properties of symplectic manifolds as follows: (1) each point is associated with a vector in the tangent space of some embedded manifold and a covector in the cotangent space, and (2) it is equipped with a 2-form satisfying certain “natural” constraints. Indeed, in classical mechanics, we generate systems of objects under force that are naturally symplectic in nature: (1) each object is associated with a state consisting of a position in some embedded configuration space and momentum, and (2) we have a set of laws governing how the forces act upon the object (the Hamiltonian function). In fact, as pointed out in Cohn’s article “Why symplectic geometry is the natural setting for classical mechanics,” symplectic manifolds provide precisely this desired abstraction of phase spaces. A fundamental example of this intimate connection is the classic n-body problem, where we seek to determine the orbits of n objects in R3 by their mutual gravitational interactions. By identifying the initial phase space, that is, identifying a vector in the symplectic vector space R6n characterizing the initial position and momentum for each of the n bodies, and determining the energy (Hamiltonian) function, can construct a 6 dimensional phase flow that fully characterizes how the phase space (ie. position and momentum) evolves through time [C, A]. Understanding the properties of symplectic manifolds provides a clean way to analyze these dynamical systems. Though the manifold under which the system of study operates need not be Euclidean, as the n- body problem elucidates, very often the system involves the motion of particles through three space, and thus the manifold is Euclidean [C]. As such, this paper will limit its scope to symplectic vector spaces. This is a very strong simplifying assumption since, as we will later prove, every symplectic vector space is isomorphic to the R2n and therefore is characterized fully by its dimension. 1 2CONNERJAGER 1. Symplectic Vector Spaces It is, of course, important to first establish some basic definitions and principles concerning sym- plectic vector spaces. Definition 1.1. A symplectic vector space is a (finite-dimensional) vector space V equipped with a bilinear form ω (called the symplectic form of V ) that is antisymmetric: ω(u, v)= ω(v, u) • − non-degenerate: if u = 0, then there exists a vector v such that ω(u, v) = 0. • If we also assume the characteristic of the field of V is = 2, we can deduce that ω(v, v) = 0 for all vectors v V from the antisymmetry of ω. For the purpose of this paper, we will assume that the ∈ field of V is R. [HZ] Example 1.1. The canonical example of a symplectic vector space is R2n with the bilinear form ω (u, v)= Ju,v ,where , is the Euclidean inner product (ie. dot product) and J is the 2n 2n 0 · · × matrix: 0 I J = n I 0 − n We can easily verify that ω0 is antisymmetric and non-degenerate: Let u =(u ,...,u ),v =(v ,...,v ). • 1 2n 1 2n ω (u, v)=u v + + u v u v u v 0 n+1 1 ··· 2n n − 1 n+1 −···− n 2n ω (v, u)=v u + + v u v u v u = ω (u, v) 0 n+1 1 ··· 2n n − 1 n+1 −···− n 2n − 0 Let u =0 = u =(u ,...,u )whereu = 0 for some i.Thus,ω (u, e ) = u =0 = • ⇒ 1 2n i | 0 i | i ⇒ ω (u, e ) = 0 for the standard basis vector e . 0 i i Definition 1.2. The symplectic orthogonal complement of some subspace E V is defined as ⊆ E⊥ := v V : ω(u, v)=0, u E { ∈ ∀ ∈ } It is critical to note that in general, E and E⊥ need not be complements and E⊥ need not itself be a symplectic vector space. As a counterexample, consider when E = span(v) for some nonzero vector v V . For any x, y E,x = c v, y = c v and thus ∈ ∈ 1 2 ω(x, y)=c c ω(v, v)=0 1 2 · (In this case, E E and ω is degenerate on E.) ⊆ ⊥ From these observations, we will define special classes of subspaces of a symplectic vector space. [L] SYMPLECTIC VECTOR SPACES, LAGRANGIAN SUBSPACES, AND LIOUVILLE’S THEOREM 3 Definition 1.3. Let V be a symplectic vector space. A subspace E V is called ⊆ isotropic if E E • ⊆ ⊥ Lagrangian if E = E • ⊥ symplectic if ω =0 • |E We will now prove the following theorem that allows gives us an isomorphism between any sym- 2n plectic vector space and the canonical symplectic space (R ,ω0). We will make use of the following propositions. The proofs of these propositions are straightforward, and they are left to the reader as an exercise (see [L]). Proposition 1.1. dim V =dimE +dimE⊥ Proposition 1.2. (E⊥)⊥ = E Proposition 1.3. Let E V . The following conditions are equivalent: ⊆ (1) E is symplectic (2) E⊥ is symplectic (3) E E = 0 ∩ ⊥ { } (4) V = E E ⊕ ⊥ Theorem 1.1. Given a symplectic vector space (V,ω), there exists a basis e1,...,en,f1,...,fn of V such that ω(ei,ej)=0 ω(fi,fj)=0 ω(ei,fj)=δij Proof. Choose any vector e V . By the non-degeneracy of ω, there exists some vector f V such ∈ ∈ that ω(e, f )=c = 0. Using the linearity of ω in the second component, we can normalize f to f = f /c such that ω(e, f) = 1. Suppose e and f were linearly dependent. We could then write f = λe,butthenω(e, f)=λω(e, e) = 0, a contradiction. Thus, e and f span a subspace E V ⊆ of dimension 2. By our construction, E is symplectic: every non-zero vector v E has a non-zero ∈ component of either e or f,soω applied to v and f or e, respectively, is non-zero and thus ω is |E non-degenerate. If dim V = 2, we are done; if not, we use Proposition 1.3 and induction to generate the desired basis on E .SinceV = E E and ω(v, w) = 0 for v E,w E , this completes ⊥ ⊕ ⊥ ∈ ∈ ⊥ the proof. We leave it to the reader to show that the desired conditions on the basis hold, and that the matrix representation of ω with respect to this basis is J. One important consequence of Theorem 1.1 is that the symplectic vector spaces have even dimension. [HZ] 4CONNERJAGER 2. Lagrangian Subspaces The goal of this section is to highlight the importance of Lagrangian subspaces. As we will see, we can think of the configuration space as an embedded subspace of the symplectic phase space. This embedded space is Lagrangian. The following theorem, which shows how we can generate a symplectic vector space from a given vector space, proves this point. Theorem 2.1. For any vector space V with dual V , the vector space V V is symplectic, with ∗ ⊕ ∗ symplectic form ω((v ,f ), (v ,f )) = f (v ) f (v ) 1 1 2 2 2 1 − 1 2 Proof. It suffices to show that ω is non-degenerate and skew-symmetric since the bilinearity is inherited from the linearity of f ,f . Choose any non-zero (v, f) V V .Ifv = 0, then there 1 2 ∈ ⊕ ∗ must be some linear functional g V such that g(v) = 0. The vector (0,g) V V satisfies the ∈ ∗ ∈ ⊕ ∗ non-degeneracy condition: ω((v, f), (g, 0)) = g(f) f(0) = g(f) =0 − If v = 0, then f =0 = f(w) = 0 for some w V . The vector (w, 0) V V satisfies the ⇒ ∈ ∈ ⊕ ∗ non-degeneracy condition: ω((v, f), (w, 0)) = 0 f(w)= f(w) =0 − − Finally, we note that ω((v ,f ), (v ,f )) = f (v ) f (v )= ω((v ,f ), (v ,f )) 2 2 1 1 1 2 − 2 1 − 1 1 2 2 Corollary 2.1. E = V 0 and E =0 V are Lagrangian subspaces of the symplectic vector ⊕ n∗ n ⊕ ∗ space V V = E E . ⊕ ∗ ⊕ Proof. ( ) Let ω be the symplectic form on V V as defined above. Choose any v V . ⊆ ⊕ ∗ ∈ ω((v, 0), (v, 0)) = 0(v) 0(v)=0 = (v, 0) E⊥ − ⇒ ∈ ( ) Let (w, f) E . For any v V , ⊇ ∈ ⊥ ∈ ω((w, f), (v, 0)) = 0(w) f(v)=0 = f(v)=0 v V = f =0 = (w, f) E − ⇒ ∀ ∈ ⇒ ⇒ ∈ Thus, E = E⊥. A similar argument shows E is also Lagrangian. Corollary 2.1 tells us that not only can we generate a symplectic vector space from a generic vector space, but the generating vector space will be embedded in the symplectic space as a Lagrangian subspace. [L] We will now prove the converse. Theorem 2.2. Any symplectic vector space (V,ω) contains a Lagrangian subspace E V such ⊆ that V = E E .
Recommended publications
  • 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]
  • A Guide to Symplectic Geometry
    OSU — SYMPLECTIC GEOMETRY CRASH COURSE IVO TEREK A GUIDE TO SYMPLECTIC GEOMETRY IVO TEREK* These are lecture notes for the SYMPLECTIC GEOMETRY CRASH COURSE held at The Ohio State University during the summer term of 2021, as our first attempt for a series of mini-courses run by graduate students for graduate students. Due to time and space constraints, many things will have to be omitted, but this should serve as a quick introduction to the subject, as courses on Symplectic Geometry are not currently offered at OSU. There will be many exercises scattered throughout these notes, most of them routine ones or just really remarks, not only useful to give the reader a working knowledge about the basic definitions and results, but also to serve as a self-study guide. And as far as references go, arXiv.org links as well as links for authors’ webpages were provided whenever possible. Columbus, May 2021 *[email protected] Page i OSU — SYMPLECTIC GEOMETRY CRASH COURSE IVO TEREK Contents 1 Symplectic Linear Algebra1 1.1 Symplectic spaces and their subspaces....................1 1.2 Symplectomorphisms..............................6 1.3 Local linear forms................................ 11 2 Symplectic Manifolds 13 2.1 Definitions and examples........................... 13 2.2 Symplectomorphisms (redux)......................... 17 2.3 Hamiltonian fields............................... 21 2.4 Submanifolds and local forms......................... 30 3 Hamiltonian Actions 39 3.1 Poisson Manifolds................................ 39 3.2 Group actions on manifolds.......................... 46 3.3 Moment maps and Noether’s Theorem................... 53 3.4 Marsden-Weinstein reduction......................... 63 Where to go from here? 74 References 78 Index 82 Page ii OSU — SYMPLECTIC GEOMETRY CRASH COURSE IVO TEREK 1 Symplectic Linear Algebra 1.1 Symplectic spaces and their subspaces There is nothing more natural than starting a text on Symplecic Geometry1 with the definition of a symplectic vector space.
    [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]
  • Book: Lectures on Differential Geometry
    Lectures on Differential geometry John W. Barrett 1 October 5, 2017 1Copyright c John W. Barrett 2006-2014 ii Contents Preface .................... vii 1 Differential forms 1 1.1 Differential forms in Rn ........... 1 1.2 Theexteriorderivative . 3 2 Integration 7 2.1 Integrationandorientation . 7 2.2 Pull-backs................... 9 2.3 Integrationonachain . 11 2.4 Changeofvariablestheorem. 11 3 Manifolds 15 3.1 Surfaces .................... 15 3.2 Topologicalmanifolds . 19 3.3 Smoothmanifolds . 22 iii iv CONTENTS 3.4 Smoothmapsofmanifolds. 23 4 Tangent vectors 27 4.1 Vectorsasderivatives . 27 4.2 Tangentvectorsonmanifolds . 30 4.3 Thetangentspace . 32 4.4 Push-forwards of tangent vectors . 33 5 Topology 37 5.1 Opensubsets ................. 37 5.2 Topologicalspaces . 40 5.3 Thedefinitionofamanifold . 42 6 Vector Fields 45 6.1 Vectorsfieldsasderivatives . 45 6.2 Velocityvectorfields . 47 6.3 Push-forwardsofvectorfields . 50 7 Examples of manifolds 55 7.1 Submanifolds . 55 7.2 Quotients ................... 59 7.2.1 Projectivespace . 62 7.3 Products.................... 65 8 Forms on manifolds 69 8.1 Thedefinition. 69 CONTENTS v 8.2 dθ ....................... 72 8.3 One-formsandtangentvectors . 73 8.4 Pairingwithvectorfields . 76 8.5 Closedandexactforms . 77 9 Lie Groups 81 9.1 Groups..................... 81 9.2 Liegroups................... 83 9.3 Homomorphisms . 86 9.4 Therotationgroup . 87 9.5 Complexmatrixgroups . 88 10 Tensors 93 10.1 Thecotangentspace . 93 10.2 Thetensorproduct. 95 10.3 Tensorfields. 97 10.3.1 Contraction . 98 10.3.2 Einstein summation convention . 100 10.3.3 Differential forms as tensor fields . 100 11 The metric 105 11.1 Thepull-backmetric . 107 11.2 Thesignature . 108 12 The Lie derivative 115 12.1 Commutator of vector fields .
    [Show full text]
  • MAT 531 Geometry/Topology II Introduction to Smooth Manifolds
    MAT 531 Geometry/Topology II Introduction to Smooth Manifolds Claude LeBrun Stony Brook University April 9, 2020 1 Dual of a vector space: 2 Dual of a vector space: Let V be a real, finite-dimensional vector space. 3 Dual of a vector space: Let V be a real, finite-dimensional vector space. Then the dual vector space of V is defined to be 4 Dual of a vector space: Let V be a real, finite-dimensional vector space. Then the dual vector space of V is defined to be ∗ V := fLinear maps V ! Rg: 5 Dual of a vector space: Let V be a real, finite-dimensional vector space. Then the dual vector space of V is defined to be ∗ V := fLinear maps V ! Rg: ∗ Proposition. V is finite-dimensional vector space, too, and 6 Dual of a vector space: Let V be a real, finite-dimensional vector space. Then the dual vector space of V is defined to be ∗ V := fLinear maps V ! Rg: ∗ Proposition. V is finite-dimensional vector space, too, and ∗ dimV = dimV: 7 Dual of a vector space: Let V be a real, finite-dimensional vector space. Then the dual vector space of V is defined to be ∗ V := fLinear maps V ! Rg: ∗ Proposition. V is finite-dimensional vector space, too, and ∗ dimV = dimV: ∗ ∼ In particular, V = V as vector spaces. 8 Dual of a vector space: Let V be a real, finite-dimensional vector space. Then the dual vector space of V is defined to be ∗ V := fLinear maps V ! Rg: ∗ Proposition. V is finite-dimensional vector space, too, and ∗ dimV = dimV: ∗ ∼ In particular, V = V as vector spaces.
    [Show full text]
  • INTRODUCTION to ALGEBRAIC GEOMETRY 1. Preliminary Of
    INTRODUCTION TO ALGEBRAIC GEOMETRY WEI-PING LI 1. Preliminary of Calculus on Manifolds 1.1. Tangent Vectors. What are tangent vectors we encounter in Calculus? 2 0 (1) Given a parametrised curve α(t) = x(t); y(t) in R , α (t) = x0(t); y0(t) is a tangent vector of the curve. (2) Given a surface given by a parameterisation x(u; v) = x(u; v); y(u; v); z(u; v); @x @x n = × is a normal vector of the surface. Any vector @u @v perpendicular to n is a tangent vector of the surface at the corresponding point. (3) Let v = (a; b; c) be a unit tangent vector of R3 at a point p 2 R3, f(x; y; z) be a differentiable function in an open neighbourhood of p, we can have the directional derivative of f in the direction v: @f @f @f D f = a (p) + b (p) + c (p): (1.1) v @x @y @z In fact, given any tangent vector v = (a; b; c), not necessarily a unit vector, we still can define an operator on the set of functions which are differentiable in open neighbourhood of p as in (1.1) Thus we can take the viewpoint that each tangent vector of R3 at p is an operator on the set of differential functions at p, i.e. @ @ @ v = (a; b; v) ! a + b + c j ; @x @y @z p or simply @ @ @ v = (a; b; c) ! a + b + c (1.2) @x @y @z 3 with the evaluation at p understood.
    [Show full text]
  • Optimization Algorithms on Matrix Manifolds
    00˙AMS September 23, 2007 © Copyright, Princeton University Press. No part of this book may be distributed, posted, or reproduced in any form by digital or mechanical means without prior written permission of the publisher. Index 0x, 55 of a topology, 192 C1, 196 bijection, 193 C∞, 19 blind source separation, 13 ∇2, 109 bracket (Lie), 97 F, 33, 37 BSS, 13 GL, 23 Grass(p, n), 32 Cauchy decrease, 142 JF , 71 Cauchy point, 142 On, 27 Cayley transform, 59 PU,V , 122 chain rule, 195 Px, 47 characteristic polynomial, 6 ⊥ chart Px , 47 Rn×p, 189 around a point, 20 n×p R∗ /GLp, 31 of a manifold, 20 n×p of a set, 18 R∗ , 23 Christoffel symbols, 94 Ssym, 26 − Sn 1, 27 closed set, 192 cocktail party problem, 13 S , 42 skew column space, 6 St(p, n), 26 commutator, 189 X, 37 compact, 27, 193 X(M), 94 complete, 56, 102 ∂ , 35 i conjugate directions, 180 p-plane, 31 connected, 21 S , 58 sym+ connection S (n), 58 upp+ affine, 94 ≃, 30 canonical, 94 skew, 48, 81 Levi-Civita, 97 span, 30 Riemannian, 97 sym, 48, 81 symmetric, 97 tr, 7 continuous, 194 vec, 23 continuously differentiable, 196 convergence, 63 acceleration, 102 cubic, 70 accumulation point, 64, 192 linear, 69 adjoint, 191 order of, 70 algebraic multiplicity, 6 quadratic, 70 arithmetic operation, 59 superlinear, 70 Armijo point, 62 convergent sequence, 192 asymptotically stable point, 67 convex set, 198 atlas, 19 coordinate domain, 20 compatible, 20 coordinate neighborhood, 20 complete, 19 coordinate representation, 24 maximal, 19 coordinate slice, 25 atlas topology, 20 coordinates, 18 cotangent bundle, 108 basis, 6 cotangent space, 108 For general queries, contact [email protected] 00˙AMS September 23, 2007 © Copyright, Princeton University Press.
    [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]
  • SYMPLECTIC GEOMETRY and MECHANICS a Useful Reference Is
    GEOMETRIC QUANTIZATION I: SYMPLECTIC GEOMETRY AND MECHANICS A useful reference is Simms and Woodhouse, Lectures on Geometric Quantiza- tion, available online. 1. Symplectic Geometry. 1.1. Linear Algebra. A pre-symplectic form ω on a (real) vector space V is a skew bilinear form ω : V ⊗ V → R. For any W ⊂ V write W ⊥ = {v ∈ V | ω(v, w) = 0 ∀w ∈ W }. (V, ω) is symplectic if ω is non-degenerate: ker ω := V ⊥ = 0. Theorem 1.2. If (V, ω) is symplectic, there is a “canonical” basis P1,...,Pn,Q1,...,Qn such that ω(Pi,Pj) = 0 = ω(Qi,Qj) and ω(Pi,Qj) = δij. Pi and Qi are said to be conjugate. The theorem shows that all symplectic vector spaces of the same (always even) dimension are isomorphic. 1.3. Geometry. A symplectic manifold is one with a non-degenerate 2-form ω; this makes each tangent space TmM into a symplectic vector space with form ωm. We should also assume that ω is closed: dω = 0. We can also consider pre-symplectic manifolds, i.e. ones with a closed 2-form ω, which may be degenerate. In this case we also assume that dim ker ωm is constant; this means that the various ker ωm fit tog ether into a sub-bundle ker ω of TM. Example 1.1. If V is a symplectic vector space, then each TmV = V . Thus the symplectic form on V (as a vector space) determines a symplectic form on V (as a manifold). This is locally the only example: Theorem 1.4.
    [Show full text]
  • M7210 Lecture 7. Vector Spaces Part 4, Dual Spaces Wednesday September 5, 2012 Assume V Is an N-Dimensional Vector Space Over a field F
    M7210 Lecture 7. Vector Spaces Part 4, Dual Spaces Wednesday September 5, 2012 Assume V is an n-dimensional vector space over a field F. Definition. The dual space of V , denoted V ′ is the set of all linear maps from V to F. Comment. F is an ambiguous object. It may be regarded: (a) as a field, (b) vector space over F, (c) as a vector space equipped with basis, {1}. It is always important to bear in mind which of the three objects we are thinking about. In the definition of V ′, we use (c). The distinctions are often pushed into the background, but let us keep them in clear view for the remainder of this paragraph. Suppose V has basis E = {e1,e2,...,en}. ′ If w ∈ V , then for each ei, there is a unique scalar ω(ei) such that w(ei) = ω(ei)1. Then w = ω(e1) ω(e2) · · · ω(en) . (1) {1}E The columns (of height 1) on the right hand side are the coefficients required to write the images of the basis elements in E in terms of the basis {1}. (The definition of the matrix of a linear map that we gave in the last lecture demands that each entry in the matrix be an element of the scalar field, not of a vector space.) The notation w(ei) = ω(ei)1 is a reminder that w(ei) is an element of the vector space F, while ω(ei) is an element of the field F. But scalar multiplication of elements of the vector space F by elements of the field F is good old fashioned multiplication ∗ : F × F → F.
    [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]
  • Manifolds, Tangent Vectors and Covectors
    Physics 250 Fall 2015 Notes 1 Manifolds, Tangent Vectors and Covectors 1. Introduction Most of the “spaces” used in physical applications are technically differentiable man- ifolds, and this will be true also for most of the spaces we use in the rest of this course. After a while we will drop the qualifier “differentiable” and it will be understood that all manifolds we refer to are differentiable. We will build up the definition in steps. A differentiable manifold is basically a topological manifold that has “coordinate sys- tems” imposed on it. Recall that a topological manifold is a topological space that is Hausdorff and locally homeomorphic to Rn. The number n is the dimension of the mani- fold. On a topological manifold, we can talk about the continuity of functions, for example, of functions such as f : M → R (a “scalar field”), but we cannot talk about the derivatives of such functions. To talk about derivatives, we need coordinates. 2. Charts and Coordinates Generally speaking it is impossible to cover a manifold with a single coordinate system, so we work in “patches,” technically charts. Given a topological manifold M of dimension m, a chart on M is a pair (U, φ), where U ⊂ M is an open set and φ : U → V ⊂ Rm is a homeomorphism. See Fig. 1. Since φ is a homeomorphism, V is also open (in Rm), and φ−1 : V → U exists. If p ∈ U is a point in the domain of φ, then φ(p) = (x1,...,xm) is the set of coordinates of p with respect to the given chart.
    [Show full text]