1. Overview Over the Basic Notions in Symplectic Geometry Definition 1.1

Total Page:16

File Type:pdf, Size:1020Kb

1. Overview Over the Basic Notions in Symplectic Geometry Definition 1.1 AN INTRODUCTION TO SYMPLECTIC GEOMETRY MAIK PICKL 1. Overview over the basic notions in symplectic geometry Definition 1.1. Let V be a m-dimensional R vector space and let Ω : V × V ! R be a bilinear map. We say Ω is skew-symmetric if Ω(u; v) = −Ω(v; u), for all u; v 2 V . Theorem 1.1. Let Ω be a skew-symmetric bilinear map on V . Then there is a basis u1; : : : ; uk; e1 : : : ; en; f1; : : : ; fn of V s.t. Ω(ui; v) = 0; 8i and 8v 2 V Ω(ei; ej) = 0 = Ω(fi; fj); 8i; j Ω(ei; fj) = δij; 8i; j: In matrix notation with respect to this basis we have 00 0 0 1 T Ω(u; v) = u @0 0 idA v: 0 −id 0 In particular we have dim(V ) = m = k + 2n. Proof. The proof is a skew-symmetric version of the Gram-Schmidt process. First let U := fu 2 V jΩ(u; v) = 0 for all v 2 V g; and choose a basis u1; : : : ; uk for U. Now choose a complementary space W s.t. V = U ⊕ W . Let e1 2 W , then there is a f1 2 W s.t. Ω(e1; f1) 6= 0 and we can assume that Ω(e1; f1) = 1. Let W1 = span(e1; f1) Ω W1 = fw 2 W jΩ(w; v) = 0 for all v 2 W1g: Ω Ω We claim that W = W1 ⊕ W1 : Suppose that v = ae1 + bf1 2 W1 \ W1 . We get that 0 = Ω(v; e1) = −b 0 = Ω(v; f1) = a and therefore v = 0. Now let v 2 W . We have that Ω(v; e1) = c and Ω(v; f1) = d. And therefore v = (−cf1 + de1) + (v + cf1 − de1) : | {z } | {z } 2W Ω 1 2W1 Ω This shows the claim. Now we go on: we pick e2; f2 2 W1 with Ω(e2; f2) = 1. Let W2 = span(e2; f2). With the same arguments as above we obtain a decomposition V = U ⊕ W1 ⊕ ::: ⊕ Wn: 1 2 MAIK PICKL Ω Note that the ei; fi always come in pairs, because if one of the Wi would be one dimensional it would actually lie in U. But now we have our desired decomposition. Remark. We will call n the rank of the bilinear map Ω. Definition 1.2. A skew-symmetric bilinear map is called symplectic (or non-degenerate) if U = f0g (U as above). The map Ω is then called a linear symplectic structure on V and (V; Ω) is called a symplectic vector space. Corollary 1.1.1. As an immediate Corollary of Theorem 1.1 we get that dim(V ) = 2n since dim(U) = 0. Choosing a basis as in the Theorem we have that 0 id Ω(u; v) = uT v: −id 0 The basis e1; : : : ; en; f1; : : : ; fn is called a symplectic basis. Definition 1.3. We have different kind of subspaces of a symplectic vector space. Let W be a subspace of (V; Ω). Denote W Ω := fv 2 V jΩ(v; w) = 0 8w 2 W g Ω 1) A subspace W ⊂ V is called symplectic if W \ W = ; if and only if ΩjW is non-degenerate. For example span(ei; fi) is a symplectic subspace. Ω 2) A subspace W ⊂ V is called isotropic if W ⊂ W if and only if ΩjW ≡ 0. For example span(e1; e2) is isotropic. 3) A subspace is called coisotropic if W Ω ⊂ W if and only if W Ω is isotropic. For example codimension 1 subspaces are coisotropic. 1 4) A subspace is called Lagrangian if it is isotropic and dim(W ) = 2 dim(V ) if and only if W Ω = W . Definition 1.4. Let (V; Ω); (V 0; Ω0) be two symplectic vector spaces. A linear map ' : V ! V 0 is called a symplectomorphism if Ω = '∗Ω0. If a symplectomorphism exists (V; Ω) and (V 0; Ω0) are called symplectomorphic. Lemma 1.2. A linear map ' : V ! V is a symplectomorphism iff the graph Γ' := f(v; '(v)) 2 V × V jv 2 V g is a Lagrangian subspace of (V ×V; (−Ω)×Ω), where (−Ω)×Ω((v; v0); (w; w0)) = −Ω(v; w)+ Ω(v0; w0). Ω Proof. Let ' be a symplectomorphism. Then we have Γ' ⊂ Γ' , since (−Ω) × Ω((v; '(v)); (w; '(w))) = −Ω(v; w) + Ω('(v);'(w)) = −Ω(v; w) + Ω(v; w) = 0: The dimension of Γ' is equal to the dimension of V , which finishes the first direction. Now let Γ' be a Lagrangian subspace it follows that 0 = −Ω(v; w) + Ω('(v);'(w)); ∗ for all v; w 2 V . This shows that ' is an isomorphism and Ω = ' Ω. AN INTRODUCTION TO SYMPLECTIC GEOMETRY 3 We now turn to manifolds. Let M be a manifold and ! be a differential 2-form, i.e. for p 2 M we have that !p : TpM ×TpM ! R is a skew-symmetric bilinear map and wp varies smoothly in p. We say ! is closed if it satisfies d! = 0, i.e. ! is a cycle in the de Rham-complex. Definition 1.5. The 2-form ! is symplectic if ! is closed and !p is a symplectic bilinear map on TpM for all p 2 M. If ! is symplectic we obtain dim(M) = dim(TpM) = 2n: A symplectic manifold is a pair (M; !) where M is a manifold and ! is a symplectic form. Example 1.1. 2n • The prototype for a symplectic manifold is R with linear coordinates x1; : : : ; xn; y1; : : : ; yn. The form n X ! = dxi ^ dyi i=1 is symplectic and the set @ @ @ @ (p);:::; (p); (p);:::; (p) @x1 @xn @y1 @yn is a symplectic basis of TpM. • In a similar spirit we have that Cn is a symplectic manifold with linear coordinates z1; : : : ; zn and symplectic form n i X ! = dz ^ dz : 2 k k k=1 • Let M = S2 regarded as the set of unit vectors in R3. We have a form !(u; v) = hp; u × vi 2 ? for u; v 2 TpS = fpg . This form is closed since it is of top degree. It is non- degenerate since for example u 6= 0 and v = u × p gives !(u; v) 6= 0. • The higher dimensional spheres S2n with n > 1 don't carry a symplectic structure. This follows from the general fact, that the even dimensional De-Rham cohomology k groups HdR(M) of a compact symplectic manifold M are not zero. • Any non-orientable Manifold doesn't carry a symplectic structure, since !n is a non- vanishing form and therefore gives a orientation on any symplectic manifold. Definition 1.6. Let (M1;!1); (M2;!2) be two 2n dimensional symplectic manifolds. Let ∗ ' : M1 ! M2 be a diffeomorphism. Then ' is a symplectomorphism if !1 = ' !2. Theorem 1.3. Let (m; !) be a 2n dimensional symplectic manifold, and let p 2 M be any point in M. Then there is a coordinate chart (U; ' = x1; : : : ; xn; y1; : : : ; yn) centered at p (i.e. '(p) = 0) s.t. n X ! = dxi ^ dyi; i=1 locally on U. Such a chart is called a Darboux chart. 4 MAIK PICKL Example 1.2. The Cotangent Bundle Let X be any n-dimensional manifold and let M = T ∗X be it's cotangent bundle. Let ∗ (U; x1; : : : ; xn be a chart on X. Then the differentials (dx1)x;:::; (dxn)x form a basis of Tx X ∗ Pn for all x 2 U. For ξ 2 Tx X we can write ξ = i=1 ξi(dxi)x. This induces a coordinate chart on T ∗U ⊂ T ∗X ∗ 2n T U ! R (x; ξ) 7! (x1; : : : ; xn; ξ1; : : : ; ξn): 0 0 For two charts (U; ' = x1; : : : ; xn); (V; = x1; : : : ; xn) on X for x 2 U \ V we have n −1 X @( ◦ ' )i (dx0 ) = ('(x))(dx ) i x @x j x j=1 j ∗ and therefore for ξ 2 Tx X we have n n X X 0 0 (1) ξ = ξi(dxi)x = ξj(dxj)x i=1 j=1 −1 0 Pn @( ◦ ' )i where ξj = i=1 ξi ('(x)). This is a smooth function and thus the cotangent @xj bundle is smooth. We now have the following 2-form on T ∗U, n X ! = dxi ^ dξi: i=1 This 2-form is closed since we have a 1-form on T ∗U, n X α = ξidxi i=1 with −dα = !. We claim that α is independent of the choice of coordinates. This fol- ∗ lows directly from formula (1), since for 2 coordinate charts (T U; x1; : : : ; xn; ξ1; : : : ; ξn); ∗ 0 0 0 0 (T V; x1; : : : ; xn; ξ1; : : : ; ξn) we have that n n X X 0 0 0 α = ξidxi = ξjdxj = α i=1 j=1 on T ∗U \ T ∗V . So α gives a 1-form on T ∗X called the tautological form or Liouville 1-form, the 2-form ! is called the canonical symplectic form. Example 1.3. Complex projective Space n PC n Consider the linear coordinates z1; : : : ; zn on C n f0g, where zj = xj + iyj. We will use the complex valued forms dzj := dxj + idyj; dzj := dxj − idyj and the differential operators @ 1 @ @ @ 1 @ @ := − i ; := + i @zj 2 @xj @yj @zj 2 @xj @yj AN INTRODUCTION TO SYMPLECTIC GEOMETRY 5 acting on the space of complex valued smooth functions on Cn n f0g. In this notation we can write the differential of a smooth function f : Cn n f0g ! C as df = @f + @f, where n n X @f X @f @f := dz ; @f := dz : @z j @z j j=1 j j=1 j We also need the 2-form n X @2f @@f = dzj ^ dzk @zj@zk j;k=1 We can now define a 2-form on Cn n f0g, by i !~ := @@ log(jzj2) FS 2 n i X = jzj2dz ^ dz − z z dz ^ dz : 2jzj4 k k j k j k j;k=1 Then there exists a unique real-valued 2-form ! on n−1 s.t.
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]
  • 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]
  • Functor and Natural Operators on Symplectic Manifolds
    Mathematical Methods and Techniques in Engineering and Environmental Science Functor and natural operators on symplectic manifolds CONSTANTIN PĂTRĂŞCOIU Faculty of Engineering and Management of Technological Systems, Drobeta Turnu Severin University of Craiova Drobeta Turnu Severin, Str. Călugăreni, No.1 ROMANIA [email protected] http://www.imst.ro Abstract. Symplectic manifolds arise naturally in abstract formulations of classical Mechanics because the phase–spaces in Hamiltonian mechanics is the cotangent bundle of configuration manifolds equipped with a symplectic structure. The natural functors and operators described in this paper can be helpful both for a unified description of specific proprieties of symplectic manifolds and for finding links to various fields of geometric objects with applications in Hamiltonian mechanics. Key-words: Functors, Natural operators, Symplectic manifolds, Complex structure, Hamiltonian. 1. Introduction. diffeomorphism f : M → N , a vector bundle morphism The geometrics objects (like as vectors, covectors, T*f : T*M→ T*N , which covers f , where −1 tensors, metrics, e.t.c.) on a smooth manifold M, x = x :)*)(()*( x * → xf )( * NTMTfTfT are the elements of the total spaces of a vector So, the cotangent functor, )(:* → VBmManT and bundles with base M. The fields of such geometrics objects are the section in corresponding vector more general ΛkT * : Man(m) → VB , are the bundles. bundle functors from Man(m) to VB, where Man(m) By example, the vectors on a smooth manifold M (subcategory of Man) is the category of m- are the elements of total space of vector bundles dimensional smooth manifolds, the morphisms of this (TM ,π M , M); the covectors on M are the elements category are local diffeomorphisms between of total space of vector bundles (T*M , pM , M).
    [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]
  • Commentary on Thurston's Work on Foliations
    COMMENTARY ON FOLIATIONS* Quoting Thurston's definition of foliation [F11]. \Given a large supply of some sort of fabric, what kinds of manifolds can be made from it, in a way that the patterns match up along the seams? This is a very general question, which has been studied by diverse means in differential topology and differential geometry. ... A foliation is a manifold made out of striped fabric - with infintely thin stripes, having no space between them. The complete stripes, or leaves, of the foliation are submanifolds; if the leaves have codimension k, the foliation is called a codimension k foliation. In order that a manifold admit a codimension- k foliation, it must have a plane field of dimension (n − k)." Such a foliation is called an (n − k)-dimensional foliation. The first definitive result in the subject, the so called Frobenius integrability theorem [Fr], concerns a necessary and sufficient condition for a plane field to be the tangent field of a foliation. See [Spi] Chapter 6 for a modern treatment. As Frobenius himself notes [Sa], a first proof was given by Deahna [De]. While this work was published in 1840, it took another hundred years before a geometric/topological theory of foliations was introduced. This was pioneered by Ehresmann and Reeb in a series of Comptes Rendus papers starting with [ER] that was quickly followed by Reeb's foundational 1948 thesis [Re1]. See Haefliger [Ha4] for a detailed account of developments in this period. Reeb [Re1] himself notes that the 1-dimensional theory had already undergone considerable development through the work of Poincare [P], Bendixson [Be], Kaplan [Ka] and others.
    [Show full text]
  • Hamiltonian and Symplectic Symmetries: an Introduction
    BULLETIN (New Series) OF THE AMERICAN MATHEMATICAL SOCIETY Volume 54, Number 3, July 2017, Pages 383–436 http://dx.doi.org/10.1090/bull/1572 Article electronically published on March 6, 2017 HAMILTONIAN AND SYMPLECTIC SYMMETRIES: AN INTRODUCTION ALVARO´ PELAYO In memory of Professor J.J. Duistermaat (1942–2010) Abstract. Classical mechanical systems are modeled by a symplectic mani- fold (M,ω), and their symmetries are encoded in the action of a Lie group G on M by diffeomorphisms which preserve ω. These actions, which are called sym- plectic, have been studied in the past forty years, following the works of Atiyah, Delzant, Duistermaat, Guillemin, Heckman, Kostant, Souriau, and Sternberg in the 1970s and 1980s on symplectic actions of compact Abelian Lie groups that are, in addition, of Hamiltonian type, i.e., they also satisfy Hamilton’s equations. Since then a number of connections with combinatorics, finite- dimensional integrable Hamiltonian systems, more general symplectic actions, and topology have flourished. In this paper we review classical and recent re- sults on Hamiltonian and non-Hamiltonian symplectic group actions roughly starting from the results of these authors. This paper also serves as a quick introduction to the basics of symplectic geometry. 1. Introduction Symplectic geometry is concerned with the study of a notion of signed area, rather than length, distance, or volume. It can be, as we will see, less intuitive than Euclidean or metric geometry and it is taking mathematicians many years to understand its intricacies (which is work in progress). The word “symplectic” goes back to the 1946 book [164] by Hermann Weyl (1885–1955) on classical groups.
    [Show full text]
  • Morse Theory for Codimension-One Foliations
    TRANSACTIONS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 298. Number 1. November 1986 MORSE THEORY FOR CODIMENSION-ONE FOLIATIONS STEVEN C. FERRY AND ARTHUR G. WASSERMAN ABSTRACT. It is shown that a smooth codimension-one foliation on a compact simply-connected manifold has a compact leaf if and only if every smooth real- valued function on the manifold has a cusp singularity. Introduction. Morse theory is a method of obtaining information about a smooth manifold by analyzing the singularities of a generic real-valued function on the manifold. If the manifold has the additional structure of a foliation, one can then consider as well the singularities of the function restricted to each leaf. The resulting singular set becomes more complicated, higher dimensional, and less trivially placed. In exchange, one obtains information about the. foliation. In this paper we consider codimension-one smooth foliations and show that, in this case, the occurrence of one type of singularity, the cusp, is related to the existence of compact leaves. Specifically, we show that a smooth codimension-one foliation on a compact simply-connected manifold has a compact leaf if and only if every real-valued function on M has a cusp. The singularities of a function on a foliated manifold are very similar to the singularities of a map to the plane. See [F, H-W, L 1, T 1, Wa, W]. However, Levine [L 1] has shown that the only obstruction to finding a map of a compact manifold to the plane without a cusp singularity is the mod 2 Euler characteristic of the manifold.
    [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]
  • Arxiv:1505.02430V1 [Math.CT] 10 May 2015 Bundle Functors and Fibrations
    Bundle functors and fibrations Anders Kock Introduction The notions of bundle, and bundle functor, are useful and well exploited notions in topology and differential geometry, cf. e.g. [12], as well as in other branches of mathematics. The category theoretic set up relevant for these notions is that of fibred category, likewise a well exploited notion, but for certain considerations in the context of bundle functors, it can be carried further. In particular, we formalize and develop, in terms of fibred categories, some of the differential geometric con- structions: tangent- and cotangent bundles, (being examples of bundle functors, respectively star-bundle functors, as in [12]), as well as jet bundles (where the for- mulation of the functorality properties, in terms of fibered categories, is probably new). Part of the development in the present note was expounded in [11], and is repeated almost verbatim in the Sections 2 and 4 below. These sections may have interest as a piece of pure category theory, not referring to differential geometry. 1 Basics on Cartesian arrows We recall here some classical notions. arXiv:1505.02430v1 [math.CT] 10 May 2015 Let π : X → B be any functor. For α : A → B in B, and for objects X,Y ∈ X with π(X) = A and π(Y ) = B, let homα (X,Y) be the set of arrows h : X → Y in X with π(h) = α. The fibre over A ∈ B is the category, denoted XA, whose objects are the X ∈ X with π(X) = A, and whose arrows are arrows in X which by π map to 1A; such arrows are called vertical (over A).
    [Show full text]
  • Cotangent Bundles
    This is page 165 Printer: Opaque this 6 Cotangent Bundles In many mechanics problems, the phase space is the cotangent bundle T ∗Q of a configuration space Q. There is an “intrinsic” symplectic structure on T ∗Q that can be described in various equivalent ways. Assume first that Q is n-dimensional, and pick local coordinates (q1,... ,qn)onQ. Since 1 n ∗ ∈ ∗ i (dq ,... ,dq )isabasis of Tq Q,wecan write any α Tq Q as α = pi dq . 1 n This procedure defines induced local coordinates (q ,... ,q ,p1,... ,pn)on T ∗Q. Define the canonical symplectic form on T ∗Q by i Ω=dq ∧ dpi. This defines a two-form Ω, which is clearly closed, and in addition, it can be checked to be independent of the choice of coordinates (q1,... ,qn). Furthermore, observe that Ω is locally constant, that is, the coefficient i multiplying the basis forms dq ∧ dpi, namely the number 1, does not ex- 1 n plicitly depend on the coordinates (q ,... ,q ,p1,... ,pn)ofphase space points. In this section we show how to construct Ω intrinsically, and then we will study this canonical symplectic structure in some detail. 6.1 The Linear Case To motivate a coordinate-independent definition of Ω, consider the case in which Q is a vector space W (which could be infinite-dimensional), so that T ∗Q = W × W ∗.Wehave already described the canonical two-form on 166 6. Cotangent Bundles W × W ∗: Ω(w,α)((u, β), (v, γ)) = γ,u−β,v , (6.1.1) where (w, α) ∈ W × W ∗ is the base point, u, v ∈ W , and β,γ ∈ W ∗.
    [Show full text]
  • The Hypoelliptic Laplacian on the Cotangent Bundle
    JOURNAL OF THE AMERICAN MATHEMATICAL SOCIETY Volume 18, Number 2, Pages 379–476 S 0894-0347(05)00479-0 Article electronically published on February 28, 2005 THE HYPOELLIPTIC LAPLACIAN ON THE COTANGENT BUNDLE JEAN-MICHEL BISMUT Contents Introduction 380 1. Generalized metrics and determinants 386 1.1. Determinants 386 1.2. A generalized metric on E· 387 · 1.3. The Hodge theory of hE 387 1.4. A generalized metric on det E· 389 1.5. Determinant of the generalized Laplacian and generalized metrics 391 1.6. Truncating the spectrum of the generalized Laplacian 392 1.7. Generalized metrics on determinant bundles 393 1.8. Determinants and flat superconnections 394 1.9. The Hodge theorem as an open condition 395 1.10. Generalized metrics and flat superconnections 395 1.11. Analyticity and the Hodge condition 397 1.12. The equivariant determinant 397 2. The adjoint of the de Rham operator on the cotangent bundle 399 2.1. Clifford algebras 400 2.2. Vector spaces and bilinear forms 400 2.3. The adjoint of the de Rham operator with respect to a nondegenerate bilinear form 402 2.4. The symplectic adjoint of the de Rham operator 403 2.5. The de Rham operator on T ∗X and its symplectic adjoint 404 ∗ 2.6. A bilinear form on T ∗X and the adjoint of dT X 407 2.7. A fundamental symmetry 408 2.8. A Hamiltonian function 409 2.9. The symmetry in the case where H is r-invariant 412 2.10. Poincar´e duality 412 2.11. A conjugation of the de Rham operator 413 2.12.
    [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]