not as xies cola

Upload: miguel-orrillo

Post on 14-Apr-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/29/2019 Not as Xies Cola

    1/177

    On the Geometry of Grassmannians and the

    Symplectic Group: the Maslov Index and Its

    Applications.

    Paolo Piccione

    Daniel Victor Tausk

    Notes for a short course given at the XI Escola de Geometria Diferencial,Universidade Federal Fluminense, Niteroi, RJ, Brazil

  • 7/29/2019 Not as Xies Cola

    2/177

  • 7/29/2019 Not as Xies Cola

    3/177

    Table of Contents

    Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

    1 Symplectic Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.1 A Short Review of Linear Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.2 Complex Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.3 Complexification and Real Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    1.3.1 Complex structures and complexifications . . . . . . . . . . . . . . . . 1 4

    1.4 Symplectic Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

    1.4.1 Isotropic and Lagrangian subspaces . . . . . . . . . . . . . . . . . . . . . . 2 2

    1.4.2 Lagrangian decompositions of a symplectic space . . . . . . . . . 2 6

    Exercises for Chapter 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    2 The Geometry of Grassmannians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    2.1 Differentiable Manifolds and Lie Groups . . . . . . . . . . . . . . . . . . . . . . . 32

    2.1.1 Classical Lie Groups and Lie Algebras . . . . . . . . . . . . . . . . . . . 3 42.1.2 Actions of Lie Groups and Homogeneous Manifolds . . . . . . 38

    2.1.3 Linearization of the Action of a Lie Group on a Manifold . . 4 2

    2.2 Grassmannians and Their Differentiable Structure . . . . . . . . . . . . . . . 44

    2.3 The Tangent Space to a Grassmannian . . . . . . . . . . . . . . . . . . . . . . . . . 47

    2.4 The Grassmannian as a Homogeneous Space . . . . . . . . . . . . . . . . . . . 50

    2.5 The Lagrangian Grassmannian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

    2.5.1 The submanifolds k(L0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

    Exercises for Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

    3 Topics of Algebraic Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    3.1 The Fundamental Groupoid and Group . . . . . . . . . . . . . . . . . . . . . . . . . 6 5

    3.1.1 Stability of the homotopy class of a curve . . . . . . . . . . . . . . . . 70

    3.2 The Homotopy Exact Sequence of a Fibration . . . . . . . . . . . . . . . . . . 73

    3.2.1 Applications to the theory of classical Lie groups . . . . . . . . . 84

    3.3 Singular Homology Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1

    3.3.1 The Hurewiczs homomorphism. . . . . . . . . . . . . . . . . . . . . . . . 1 0 2

    E x e r c i s e s f o r C h a p t e r 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 0 6

    iii

  • 7/29/2019 Not as Xies Cola

    4/177

    iv TABLE OF CONTENTS

    4 The Maslov Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

    4.1 Index of a Symmetric Bilinear Form . . . . . . . . . . . . . . . . . . . . . . . . . . 109

    4.1.1 The evolution of the index of a one-parameter family of

    symmetric bilinear forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 6

    4.2 Definition and Computation of the Maslov Index . . . . . . . . . . . . . . .121

    E x e r c i s e s f o r C h a p t e r 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 3 2

    5 Some Applications to Differential Systems . . . . . . . . . . . . . . . . . . . . . . . . 134

    5.1 Symplectic Differential Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

    5.2 The Maslov Index of a Symplectic Differential System . . . . . . . . . 138

    5.3 The Maslov Index of semi-Riemannian Geodesics and Hamil-

    t o n i a n S y s t e m s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 4 1

    5.3.1 Geodesics in a semi-Riemannian manifold . . . . . . . . . . . . . . 142

    5.3.2 Hamiltonian systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 4 5

    5.3.3 Further developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

    E x e r c i s e s f o r C h a p t e r 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 4 9

    A Answers and Hints to the Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 5 1

    From Chapter 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

    From Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

    From Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

    From Chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

    From Chapter 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

    Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

    Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

  • 7/29/2019 Not as Xies Cola

    5/177

    Introduction

    It has become evident through many mathematical theories of our century that

    Geometry and Topology offer very powerful tools in the study of qualitative and

    also quantitative properties of differential equations. The main idea behind these

    theories is that some equations, or better, some classes of equations can be studied

    by means of their symmetries, where by symmetry we mean generically any alge-

    braic or geometric structure which is preserved by their flow. Once such invariant

    structures are determined for a class of differential equations, many properties of

    the solutions of the class can be read off from the geometry of the curve obtained

    by the flow, taking values in the space (typically a Lie group) of all structure-

    preserving morphisms.

    A simple, but instructive, example is given by the Sturmian theory for second

    order ordinary differential equations in IR. The Sturm oscillation theorem dealswith equations of the form (px)+ rx = x, wherep and r are functions,p > 0,and is a real parameter. The theorem states that, denoting by C1o [, ] the spaceofC1-functions on [, ] vanishing at and , the index of the symmetric bilinear

    form B(x, y) = b

    a[pxy + rxy] dt in C1o [a, b] is equal to the sum over t

    ]a, b[

    of the dimension of the kernel of the bilinear form ta [pxy + rxy] dt in C1o [a, t].The classical proof of the Sturm oscillation theorem (see for instance [4, Chap-

    ter 8]) is obtained by showing that the two quantities involved in the thesis can be

    obtained as the winding numberof two homotopic closed curves in the real projec-

    tive line.

    The class of differential equations that we are interested in consists in the so

    called symplectic differential systems; these are linear systems in IRn IRnwhose flow preserve the canonical symplectic form, given by

    (v, ), (w, )

    =

    (v) (w). Recall that a symplectic form is a nondegenerate skew-symmetricbilinear form on a (necessarily even dimensional) vector space. These differential

    systems appear naturally in a great variety of fields of pure and applied mathe-

    matics, and many areas of mathematics and physics, like Calculus of Variations,Hamiltonian systems, (Pseudo-)Riemannian Geometry, Symplectic Geometry, Me-

    chanics and Optimal Control Theory produce examples of symplectic systems as

    basic objects of investigation. For instance, MorseSturm systems are special

    cases of symplectic systems; such systems are obtained from the Jacobi equation

    along any pseudo-Riemannian geodesic by means of a parallel trivialization of the

    tangent bundle of the pseudo-Riemannian manifold along the geodesic. More in

    general, symplectic systems are obtained by considering the linearized Hamilton

    equations along any solution of a (possibly time-dependent) Hamiltonian problem,

    v

  • 7/29/2019 Not as Xies Cola

    6/177

    vi INTRODUCTION

    using a symplectic trivialization along the solution of the tangent bundle of the

    underlying symplectic manifold. Another large class of examples where the the-

    ory leads naturally to the study of symplectic systems is provided by Lagrangianvariational theories in manifolds, possibly time-dependent, even in the case ofcon-

    strained variational problems. Indeed, under a suitable invertibility assumption

    called hyper-regularity , the solutions to such problems correspond, via the Le-

    gendre transform, to the solutions of an associated Hamiltonian problem in the

    cotangent bundle.

    The fundamental matrix of a symplectic system is a curve in the symplectic

    group, denoted by Sp(2n, IR), which is a closed subgroup of the general lineargroup GL(2n, IR), hence it has a Lie group structure. This structure is extremelyrich, due to the fact that symplectic forms on a vector space are intimately related

    to its complex structures, and such relation produces other invariant geometric and

    algebraic structures, such as inner products and Hermitian products.

    Many interesting questions can be answered by studying solutions of symplec-

    tic systems whose initial data belong to a fixed Lagrangian subspace ofIRnIRn.Recall that a Lagrangian subspace of a symplectic space is a maximal subspace on

    which the symplectic form vanishes. Such initial conditions are obtained, for in-

    stance, in Riemannian or pseudo-Riemannian geometry when one considers Jacobi

    fields along a geodesic that are variations made of geodesics starting orthogonally

    at a given submanifold. Since symplectic maps preserve Lagrangian subspaces,

    the image of the initial Lagrangian by the flow of a symplectic system is a curve

    in the set of all Lagrangian subspaces of IRn IRn. The set is a smooth(indeed, real-analytic) submanifold of the Grassmannian Gn(IR

    n IRn) of alln-dimensional subspaces ofIRn

    IRn; is called the Lagrangian Grassmannian

    of the symplectic space IRn IRn.The original interest of the authors was the study of conjugate points along

    geodesics in a semi-Riemannian manifold and their stability (see [29, 34]), with

    the aim of developing an infinite dimensional Morse Theory (see [ 30, 12, 26]) for

    semi-Riemannian geodesics. A few decades ago a new integer valued homological

    invariant, called the Maslov index, was introduced by the Russian school (see for

    instance [1] and the references therein) for closed curves in a Lagrangian submani-

    fold M of the space IR2n endowed with its canonical symplectic structure. The no-tion of Maslov index has been immediately recognized as an important tool in the

    study of conjugate points, and it has has been thoroughly investigated and extended

    in several directions by mathematical-physicists, geometers and analysts. There is

    nowadays a very extensive literature about the subject, and it is almost impossibleto acknowledge the work of all the many authors who have given significant con-

    tributions to the field. Our list of references ([6, 9, 11, 14, 15, 16, 24, 28, 36, 42])

    is far from being exhaustive.

    Periodic or non periodic solutions of Hamiltonian systems, like for instance

    geodesics in a semi-Riemannian manifold, define a curve in the symplectic group,

    or in the Lagrangian Grassmannian, hence they define a Maslov index. Roughly

    speaking, the Maslov index gives a sort ofalgebraic count of the conjugate points

    along a solution; here are some of the main properties of this invariant:

  • 7/29/2019 Not as Xies Cola

    7/177

    INTRODUCTION vii

    it is always finite (even when the number of conjugate points is infinite);

    it is stable by small perturbations of the data;

    it coincides with the geometric index in the case of a causal (timelike orlightlike) Lorentzian geodesic;

    it is related to the analytic index (or, more in general, to the relative index)of the solution, which is the index of the second variation of an associated

    Lagrangian action functional;

    it is related to the spectral properties of the associated Hamiltonian secondorder differential operator.

    Conjugate and focal points appear naturally in Optics, both classical and rela-

    tivistic, and the Maslov index provides a new topological invariant. For instance,

    the optics of light rays in a general-relativistic medium, like vacuum, dust, orplasma, magnetized or not, etc., can be described using a Hamiltonian formal-

    ism. As the underlying spacetime model one assumes an arbitrary 4-dimensional

    Lorentzian manifold (M, g), and the trajectories of light rays are projections ontoM of solutions in the cotangent bundle T M of some Hamiltonian function H :T M IR. Typically, the explicit form of the function H involves the spacetimemetric and a number of tensor fields by which the medium is characterized. For a

    comprehensive discussion of this subject we refer to Perlick [33].

    The aim of this booklet is to provide a complete, self-contained study of the

    geometry of the Grassmannian manifolds, the symplectic group and the Lagrangian

    Grassmannian. This study will lead us naturally to the notion of Maslov index, that

    will be introduced in the context of symplectic differential systems.

    These notes are organized as follows. In Chapter 1 we describe the algebraicsetup; we will study complex structures and symplectic structures on finite dimen-

    sional vector spaces. Special attention is given to the Lagrangian subspaces and to

    the Lagrangian decompositions of a symplectic space.

    Chapter 2 is entirely dedicated to Differential Geometry; we will study at

    the differentiable structure of the Grassmannian manifolds and of the Lagrangian

    Grassmannians. We will develop briefly the theory of Lie groups and their actions

    on differentiable manifolds, so that we will be able to describe the Grassmannians

    as homogeneous spaces.

    In Chapter 3 we will develop the algebraic topological framework of the the-

    ory, including the basics of homotopy theory and of singular homology theory for

    topological spaces. Using the long exact sequences in homotopy and in homology,and using the Hurewicz homomorphism we will compute the first homology and

    relative homology group of the Grassmannians.

    In Chapter 4 we will introduce the notion of Maslov index for curves in the

    Lagrangian Grassmannian, and we will present some methods to compute it in

    terms of the change of signature of curves of bilinear forms.

    In Chapter 5 we will show how symplectic differential systems are produced

    in several geometrical problems; more precisely, we will consider the case of the

    Jacobi equation along a semi-Riemannian geodesic, and the case of the linearized

  • 7/29/2019 Not as Xies Cola

    8/177

    viii INTRODUCTION

    Hamilton equation along the solution of a Hamiltonian problem in a symplectic

    manifold.

    At the end of each Chapter we have given a list of exercises whose solution is

    based on the material presented in the chapter. The reader should be able to solve

    the problems as he/she goes along; the solution or a hint for the solution of (almost)

    all the exercises is given in Appendix A.

  • 7/29/2019 Not as Xies Cola

    9/177

    INTRODUCTION ix

    Ognuno sta solo sul cuor della terra

    trafitto da un raggio di sole:

    ede subito sera.

  • 7/29/2019 Not as Xies Cola

    10/177

  • 7/29/2019 Not as Xies Cola

    11/177

    CHAPTER 1

    Symplectic Spaces

    1.1. A Short Review of Linear Algebra

    In this section we will briefly review some well known facts concerning the

    identification of bilinear forms and linear operators on vector spaces. These identi-

    fications will be used repeatedly during the exposition of the material, and, to avoid

    confusion, the reader is encouraged to take a few minutes to go through the pain ofreading this section.

    The results presented are valid for vector spaces over an arbitrary field K, how-ever we will mainly be interested in the case that K = IR or K = C. Moreover,we emphasize that even though a few results presented are also valid in the case of

    infinite dimensional vector spaces, in this chapter we will always assume that the

    vector spaces involved are finite dimensional.

    Let V and W be vector spaces. We denote by Lin(V, W) and by B(V, W) re-spectively the vector spaces of all the linear operators T : V W and ofbilinearoperators, called also bilinear forms, B : V W K; by V we mean the dualspace Lin(V, K) ofV. Shortly, we set Lin(V) = Lin(V, V) and B(V) = B(V, V).

    There is a natural isomorphism:

    (1.1.1) Lin(V, W) B(V, W),which is obtained by associating to each linear operator T : V W the bilinearform BT B(V, W) given by BT(v, w) = T(v)

    w

    .

    1.1.1. REMARK. Given vector spaces V,W,V1, W1 and a pair (L, M) of linearoperators, with L Lin(V1, V) and M Lin(W, W1), one defines another linearoperator:

    (1.1.2) Lin(L, M) : Lin(V, W) Lin(V1, W1)by:

    (1.1.3) Lin(L, M)

    T = M

    T

    L.

    In this way, Lin(, ) becomes a functor, contravariant in the first variable and co-variant in the second, from the category of pairs of vector spaces to the category

    of vector spaces. Similarly, given linear operators L Lin(V1, V) and M Lin(W1, W), we can define a linear operator B(L, M) : B(V, W) B(V1, W1)by setting B(L, M) B = B(L, M). In this way, B(, ) turns into a functor,contravariant in both variables, from the category of pairs of vector spaces to the

    category of vector spaces. This abstract formalism will infact be useful later (see

    Section 2.3).

    1

  • 7/29/2019 Not as Xies Cola

    12/177

    2 1. SYMPLECTIC SPACES

    The naturality of the isomorphism (1.1.1) may be meant in the technical sense

    ofnatural isomorphism between the functors Lin(

    ,

    ) andB(

    ,

    ) (see Exercise 1.1).

    To avoid confusion, in this Section we will distinguish between the symbols

    of a bilinear form B and of the associated linear operator TB , or between a linearoperator T and the associated bilinear form BT. However, in the rest of the bookwe will implicitly assume the isomorphism (1.1.1), and we will not continue with

    this distinction.

    Given another pair of vector spaces V1 and W1 and operators L1 Lin(V1, V),L2 Lin(W1, W), the bilinear forms BT(L1, ) and BT(, L2) correspond via(1.1.1) to the linear operators TL1 and L2T respectively. Here, L2 : W W1denotes the transpose linear operator ofL2 given by:

    L2() = L2, W.We will identify every vector space V with its bidual V and every linear

    operator T with its bitranspose T. Given T Lin(V, W) we will thereforelook at T as an element in Lin(W, V); if BT is the bilinear operator associatedto T, then the bilinear operator BT associated to T

    is the transpose bilinear

    operator BT B(W, V) defined by BT(w, v) = BT(v, w).Given B B(V), we say that B is symmetric if B(v, w) = B(w, v) for

    all v, w V; we say that B is anti-symmetric if B(v, w) = B(w, v) for allv, w V (see Exercise 1.2). The sets of symmetric bilinear forms and of anti-symmetric bilinear forms are subspaces ofB(V), denoted respectively by Bsym(V)and Basym(V).

    The reader is warned that, unfortunately, the identification (1.1.1) does not

    behave well in terms of matrices, with the usual convention for the matrix repre-sentations of linear and bilinear operators.

    If(vi)ni=1 and (wi)

    mi=1 are bases ofV and W respectively, we denote by (v

    i )ni=1

    and (wi )mi=1 the corresponding dual bases of V

    and W. For T Lin(V, W),the matrix representation (Tij) ofT satisfies:

    T(vj) =mi=1

    Tij wi .

    On the other hand, ifB B(V, W), the matrix representation (Bij) ofB is definedby:

    Bij = B(vi, wj);

    hence, for all T Lin(V, W) we have:Tij = T(vj)(wi) = BT(vj , wi) = [BT]ji .

    Thus, the matrix of a linear operator is the transpose of the matrix of the corre-

    sponding bilinear operator; in some cases we will be considering symmetric opera-

    tors, and there will be no risk of confusion. However, when we deal with symplectic

    forms (see Section 1.4) one must be careful not to make sign errors.

  • 7/29/2019 Not as Xies Cola

    13/177

    1.1. A SHORT REVIEW OF LINEAR ALGEBRA 3

    1.1.2. DEFINITION. Given T Lin(V, W), we define the pull-backassociatedto T to be map:

    T# : B(W) B(V)given by T#(B) = B(T, T). When T is an isomorphism, we can also define the

    push-forwardassociated to T, which is the map:

    T# : B(V) B(W)defined by T#(B) = B(T

    1, T1).1.1.3. EXAMPLE. Using (1.1.1) to identify linear and bilinear operators, we

    have the following formulas for the pull-back and the push-forward:

    (1.1.4) T#(B) = T TB T, T#(B) = (T1) TB T1.

    The identities (1.1.4) can be interpreted as equalities involving the matrix repre-sentations, in which case one must use the matrices that represent B, T#(B) andT#(B) as linear operators.

    For B B(V), the kernel ofB is the subspace ofV defined by:(1.1.5) Ker(B) =

    v V : B(v, w) = 0, w V

    .

    The kernel of B coincides with the kernel of the associated linear operator T :V V. The bilinear form B is said to be nondegenerate if Ker(B) = {0};this is equivalent to requiring that its associated linear operator T is injective, orequivalently, an isomorphism.

    1.1.4. EXAMPLE. If B B(V) is nondegenerate, then B defines an isomor-phism TB between V and V and therefore we can define a bilinear form [TB]#(B)in V by taking the push-forward of B by TB . By (1.1.4), such bilinear form isassociated to the linear operator (T1B )

    ; if B is symmetric, then [TB]#(B) is the

    bilinear form associated to the linear map T1B .

    1.1.5. DEFINITION. Let B Bsym(V) be a symmetric bilinear form in V.We say that a linear operator T : V V is B-symmetric (respectively, B-anti-symmetric) if the bilinear form B(T, ) is symmetric (respectively, anti-symmet-ric). We say that T is B-orthogonal ifT#[B] = B, i.e., ifB(T, T) = B.

    1.1.6. EXAMPLE. Given B Bsym and T Lin(V), the B-symmetry ofT isequivalent to:

    (1.1.6) TB T = (TB T);clearly, the B-anti-symmetry is equivalent to TB T = (TB T).

    When B is nondegenerate, we can also define the transpose of T relatively to

    B, which is the operator T Lin(V) such that B(T v , w) = B(v, T w) for allv, w V. Explicitly, we have(1.1.7) T = T1B T TB.

  • 7/29/2019 Not as Xies Cola

    14/177

    4 1. SYMPLECTIC SPACES

    Then, T is B-symmetric (resp., B-anti-symmetric) iffT = T (resp., iffT = T),and it is B-orthogonal iffT = T1.

    We also say that T is B-normal ifT commutes with T.

    Given a subspace S V and a bilinear form B B(V), the orthogonalcomplementS ofS with respect to B is defined by:

    (1.1.8) S =

    v V : B(v, w) = 0, w S

    .

    In particular, Ker(B) = V. The annihilator So of S is the subspace of V

    defined as:

    So =

    V : (w) = 0, w S

    .

    Observe that S = T1B (So).

    1.1.7. EXAMPLE. Assume that B Bsym(V) is nondegenerate and let T Lin(V); denote by T the B-transpose ofT. IfS V is an invariant subspace forT, i.e., ifT(S) S, then the B-orthogonal complement S of S is invariant forT. This follows from (1.1.7) and from the identity S = T1B (S

    o), observing thatthe annihilator So ofS is invariant for T.

    1.1.8. PROPOSITION. If B B(V) is nondegenerate and S V is a sub-space, then dim(V) = dim(S) + dim(S).

    PROOF. Simply note that dim(V) = dim(S)+dim(So) and that dim(S) =dim(So), and S = T1B (S

    o), with TB an isomorphism, because B is nondegen-erate.

    If B is either symmetric or anti-symmetric, then it is easy to see that S (S); the equality does nothold in general, but only ifB is nondegenerate.

    1.1.9. COROLLARY. Suppose that B B(V) is either symmetric or anti-symmetric; ifB is nondegenerate, then S = (S).

    PROOF. It is S (S); by Proposition 1.1.8 dim(S) = dim(S). If B B(V) is nondegenerate and S V is a subspace, then the restriction

    ofB to S S may be degenerate. We have the following:1.1.10. PROPOSITION. The restriction B|SS is nondegenerate if and only if

    V = S

    S.

    PROOF. The kernel of the restriction B|SS is SS; hence, ifV = SS,it follows that B is nondegenerate on S. Conversely, if B is nondegenerate on S,then S S = {0}. It remains to show that V = S + S. For, observe that themap:

    (1.1.9) S x B(x, )|S Sis an isomorphism. Hence, given v V, there exists x S such that B(x, ) andB(v, ) coincide in S, thus x v S. This concludes the proof.

  • 7/29/2019 Not as Xies Cola

    15/177

    1.2. COMPLEX STRUCTURES 5

    1.1.11. COROLLARY. Suppose that B B(V) is either symmetric or anti-symmetric; ifB is nondegenerate, then the following are equivalent:

    B is nondegenerate on S; B is nondegenerate on S.

    PROOF. Assume that B is nondegenerate on S. By Proposition 1.1.10 it isV = SS; by Corollary 1.1.9 we have V = S(S), from which it followsthat B is nondegenerate on S by Proposition 1.1.10. The converse is analogous,since (S) = S.

    1.1.12. EXAMPLE. Proposition 1.1.10 actually does not hold ifV is not finitedimensional. For instance, if V is the space of square summable sequences x =(xi)iIN of real numbers, i.e.,

    iINx

    2i < +, B is the standard Hilbert product

    in V given by B(x, y) =

    iINxiyi and S V is the subspace consisting of

    all almost null sequences, i.e., xi = 0 only for a finite number of indices i IN,then it is easy to see that S = {0}. What happens here is that the map (1.1.9) isinjective, but not surjective.

    1.1.13. REMARK. Observe that Proposition 1.1.10 is indeed true if we assume

    only that S is finite dimensional; for, in the proof presented, only the finiteness ofdim(S) was used to conclude that the map (1.1.9) is an isomorphism.

    As an application of Proposition 1.1.10 we can now prove that every symmetric

    bilinear form is diagonalizable. We say that a basis (vi)ni=1 of V diagonalizes the

    bilinear form B ifB(vi, vj) = 0 for all i = j, i.e., ifB is represented by a diagonalmatrix in the basis (vi)

    ni=1.

    1.1.14. THEOREM. Suppose thatK is a field of characteristic different from 2.Given B Bsym(V), there exists a basis (vi)ni=1 ofV that diagonalizes B.

    PROOF. We prove the result by induction on dim(V). If dim(V) = 1 theresult is trivial; assume dim(V) = n and that the result holds true for every vectorspace of dimension less than n. IfB(v, v) = 0 for all v V, then B = 0. For,

    0 = B(v + w, v + w) = 2 B(v, w),

    and the filed K has characteristic different from 2. Since the result in the case thatB = 0 is trivial, we can assume the existence ofv1 V such that B(v1, v1) = 0. Itfollows that B is nondegenerate on the one-dimensional subspace K v1 generatedby v1; by Proposition 1.1.10 we get:

    V = K v1 (K v1).By the induction hypothesis, there exists a basis (vi)

    ni=1 of(K v1)

    that diagonal-

    izes the restriction ofB; it is then easy to check that the basis (vi)ni=2 diagonalizes

    B.

    1.2. Complex Structures

    In this section we will study the procedure of changing the scalar field of a real

    vector space, making it into a complex vector space. Of course, given a complex

  • 7/29/2019 Not as Xies Cola

    16/177

    6 1. SYMPLECTIC SPACES

    vector space, one can always reduce the scalars to the real field: such operation

    will be called reduction of the scalars.

    Passing from the real to the complex field requires the introduction of an ad-ditional structure, that will be called a complex structure. Many of the proofs in

    this section are elementary, so they will be omitted and left as an exercise for the

    reader.

    For clarity, in this section we will refer to linear operators as IR-linear or C-linear, and similarly we will talk about IR-bases or C-bases, real or complex di-mension, etc.

    Let Vbe a complex vector space; we will denote by VIR the real vector spaceobtained by restriction of the multiplication by scalars C V Vto IR V V. Observe that the underlying set of vectors, as well as the operation of sum,coincides in Vand VIR. We say that VIR is a realification of V, or that VIR isobtained by a reduction of scalars from V.The endomorphism v iv ofVgiven by the multiplication by the imaginaryunit i =

    1 is C-linear, hence also IR-linear. The square of this endomorphismis given by minus the identity ofV. This suggests the following definition:

    1.2.1. DEFINITION. Let V be a real vector space. A complex structure in V isa linear operator J : V V such that J2 = J J = Id.

    Clearly, a complex structure J is an isomorphism, since J1 = J.Given a complex structure J on V it is easy to see that there exists a unique

    way of extending the multiplication by scalars IRV V ofV to a multiplicationby scalar C V V in such a way that J(v) = iv. Explicitly, we define:(1.2.1) (a + b i)v = av + b J(v), a, b

    IR, v

    V.

    Conversely, as we had already observed, every complex extension of multiplication

    by scalars for V defines a complex structure on V by J(v) = iv.We will henceforth identify every pair (V, J), where V is a real vector space

    and J is a complex structure of V, with the complex vector space Vobtained from(V, J) by (1.2.1). Observe that V is the realification VIR ofV.

    1.2.2. EXAMPLE. For every n IN, the space IR2n has a canonical complexstructure defined by J(x, y) = (y, x), for x, y IRn. We can identify (IR2n, J)with the complex vector space Cn by (x, y) x + iy. In terms of matrix repre-sentations, we have:

    (1.2.2) J = 0 II 0 ,

    where 0 and I denote respectively the 0 and the identity n n matrices.We have the following simple Lemma:

    1.2.3. LEMMA. Let(V1, J1) and (V2, J2) be real vector spaces endowed withcomplex structures. A IR-linear operator T : V1 V2 is C-linear if and only ifT J1 = J2 T. In particular, the C-linear endomorphisms of a vector space withcomplex structure (V, J) are the IR-linear endomorphisms ofV that commute withJ.

  • 7/29/2019 Not as Xies Cola

    17/177

    1.2. COMPLEX STRUCTURES 7

    PROOF. Left to the reader in Exercise 1.3.

    1.2.4. REMARK. Observe that ifJ is a complex structure on V, then also

    J isa complex structure, that will be called the conjugate complex structure. For Cand v V, the product of and v in the complex space (V, J) is given by theproduct of and v in the complex space (V, J), where is the complex conjugateof . The set of complex bases of (V, J) and (V, J) coincide; observe howeverthat the components of a vector in a fixed basis are conjugated when replacing Jby J.

    A C-linear operator T between complex spaces is still C-linear when replacingthe complex structures by their conjugates in both the domain and the counterdo-

    main. The representations ofT with respect to fixed bases in the complex structuresand the same bases in the conjugate complex structures are given by conjugate ma-

    trices.

    1.2.5. DEFINITION. A map T between complex vector spaces is said to beanti-linear, or conjugate linear, if it is additive and if T(v) = T(v) for all C and all v in the domain ofT.

    An anti-linear map is always IR-linear when we see it as a map between therealifications of the domain and the counterdomain. Moreover, a map is anti-linear

    if and only if it is C-linear when the complex structure of its domain (or of its

    counter domain) is replaced by the complex conjugate. In particular, the anti-linear

    endomorphisms of(V, J) are the IR-linear endomorphisms ofV that anti-commutewith J.

    We have the following relation between the bases of(V, J) and ofV:

    1.2.6. PROPOSITION. Let V be a (possibly infinite dimensional) real vectorspace andJ a complex structure on V. If(bj)jJ is a C-basis of(V, J), then theunion of(bj)j Jand

    J(bj)

    jJ

    is an IR-basis ofV.

    PROOF. Left to the reader in Exercise 1.4.

    1.2.7. COROLLARY. The real dimension of V is twice the complex dimensionof(V, J); in particular, a (finite dimensional) vector space admits a complex struc-ture if and only if its dimension is an even number.

    PROOF. We only need to show that every real vector space of even dimension

    admits a complex structure. This is easily established by choosing an isomorphism

    with IR2n and using the canonical complex structure given in Example 1.2.2.

    1.2.8. EXAMPLE. If(V, J) is a real vector space with complex structure, thenthe dual complex space of(V, J) is given by the set ofIR-linear maps : V Csuch that:

    (1.2.3) J(v) = i (v), v V.It is easy to see that (1.2.8) determines the imaginary part of when it is knownits real part; hence we have an IR-linear isomorphism:

    (1.2.4) (V, J) V,

  • 7/29/2019 Not as Xies Cola

    18/177

    8 1. SYMPLECTIC SPACES

    where : C IR denotes the real part operator. The isomorphism (1.2.4) there-fore induces a unique complex structure of V that makes (1.2.4) into a C-linear

    isomorphism. Such complex structure is called the dual complex structure, and itis easy to see that it is given simply by the transpose operator J.

    We conclude this section with a relation between the matrix representations of

    vectors and operators in real and complex bases. Let (V, J) be a 2n-dimensionalvector space with complex structure; a basis of V adapted to J , shortly a J-basis,is a basis of the form

    (1.2.5) (b1, . . . , bn, J(b1), . . . , J (bn));

    in this case, (bj)nj=1 is a complex basis of (V, J). For instance, the canonical basis

    ofIR2n, endowed with the canonical complex structure, is a J-basis correspondingto the canonical basis ofCn. In other words, the J-bases of a vector space areprecisely those with respect to which the matrix representations of J is that givenby (1.2.2). The existence ofJ-bases is given by Proposition 1.2.6.

    Let a J-basis of V be fixed, corresponding to a complex basis B = (bj)nj=1of (V, J). Given v V with coordinates (z1, . . . , zn) in the basis B, then itscoordinates in the (real) J-basis ofV are:

    v (x1, . . . , xn, y1, . . . , yn),where zj = xj + i yj , xj , yj IR. If T is a C-linear operator represented in thecomplex basis by the matrix Z = A + i B (A and B real), then its representationin the corresponding J-basis is:

    (1.2.6) T A BB A .1.2.9. REMARK. Formula (1.2.6) tells us that the map

    Z = A + B i

    A BB A

    is an injective homomorphism of the algebra of complex n n matrices into thealgebra of real 2n 2n matrices.

    1.3. Complexification and Real Forms

    In this section we show that any real vector space can be extended in a canon-ical way to a complex vector space, by mimicking the relation between IRn andCn;such an extension will be called a complexification of the space. We also show that,

    given a complex space, it can be seen as the complexification of several of its real

    subspaces, that will be called the real forms of the complex space. We will only

    consider the case of finite dimensional spaces, even though many of the results

    presented will hold in the case of infinite dimensional spaces, up to minor modifi-

    cations. Some of the proofs are elementary, and they will be omitted and left to the

    reader as Exercises.

  • 7/29/2019 Not as Xies Cola

    19/177

    1.3. COMPLEXIFICATION AND REAL FORMS 9

    1.3.1. DEFINITION. Let Vbe a complex vector space; a real form in Vis a realsubspace

    V0 of

    V(or, more precisely, a subspace of the realification

    VIR of

    V) such

    that: VIR = V0 i V0.In other words, a real form V0 in Vis a real subspace such that every v V

    can be written uniquely in the form v = v1 + i v2, with v1, v2 V0.To a real form V0 we associate maps:

    (1.3.1) : V V0, : V V0, c : V V,given by (v1 + i v2) = v1, (v1 + i v2) = v2 and c(v1 + i v2) = v1 i v2,for all v1, v2 V. We call , and c respectively the real part, imaginary part,and conjugation operators associated to the real form V0. All these operators areIR-linear; the operator c is also anti-linear. For v V, we also say that c(v) is theconjugate ofv relatively to the real form V0, and we also write:

    c(v) = v.

    1.3.2. DEFINITION. Let V be a real vector space. A complexification ofV is apair (VC, ), where VC is a complex vector space and : V VC is an injectiveIR-linear map such that (V) is a real form in VC.

    The result of the following Proposition is usually known as the universal prop-

    erty of the complexification:

    1.3.3. PROPOSITION. LetV be a real vector space, (VC, ) a complexificationofV andWa complex vector space. Then, given an IR-linear map f : V WIR,there exists a unique C-linear map f : VC

    Wsuch that the following diagram

    commutes:

    (1.3.2) VC

    f

    !!V

    OO

    f// W

    PROOF. Left to the reader in Exercise 1.5.

    As corollary, we obtain the uniqueness of a complexification, up to isomor-

    phisms:

    1.3.4. COROLLARY. Suppose that(VC1 , 1) and(VC2 , 2) are complexifications

    of V. Then, there exists a unique C-linear isomorphism : VC1 VC2 such thatthe following diagram commutes:

    VC1 // VC2

    V

    1

    ``AAAAAAA 2

    >>~~~~~~~~

    PROOF. Left to the reader in Exercise 1.6.

  • 7/29/2019 Not as Xies Cola

    20/177

    10 1. SYMPLECTIC SPACES

    IfV is a real vector space, we can make the direct sum V V into a complexspace by taking the complex structure J(v, w) = (

    w, v). Setting (v) = (v, 0),

    it is easy to see that (V V, ) is a complexification ofV, that will be called thecanonical complexification of the real vector space V.

    By Corollary 1.3.4, we do not need to distinguish between complexifications

    of a vector space; so, from now on, we will denote by VC the canonical complex-ification ofV, or, depending on the context, we may use the symbol VC to denotesome other complexification ofV, which is necessarily isomorphic to the canonicalone.

    The original space V will then be identified with (V), so that we will alwaysthink of an inclusion V VC; since (V) is a real form in VC, then VC is a directsum ofV and i V:

    VC = V i V.1.3.5. EXAMPLE. The subspace IRn Cn is a real form in Cn, hence Cn is a

    complexification ofIRn.

    1.3.6. EXAMPLE. The space Mn(IR) of real n n matrices is a real form inthe space Mn(C) of complex n n matrices.

    A less trivial example of a real form in Mn(C) is given by u(n), which is thespace ofanti-Hermitian matrices, i.e., matrices A such that A = A, where Adenotes the conjugate transpose matrix of A. In this example, i u(n) is the spaceof Hermitian matrices, i.e., the space of those matrices A such that A = A. It iseasy to see that Mn(C) = u(n) i u(n), and so u(n) is a real form in Mn(C) andMn(C) is a complexification ofu(n).

    1.3.7. EXAMPLE. IfVis a complex vector space and if (bj)n

    j=1 is a complexbasis ofV, then the real subspace V0 ofVIR given by:

    V0 = nj=1

    j bj : j IR, j

    is a real form in V.Actually, every real form of Vcan be obtained in this way; for, if V0 V

    is a real form, then an IR-basis (bj)nj=1 of V0 is also a C-basis of V. It follows

    in particular that the real dimension of a real form V0 is equal to the complexdimension ofV.

    Example 1.3.7 tells us that every complex space admits infinitely many real

    forms; in the following proposition we give a characterization of the real forms ina complex space. We recall that a bijection of a set is said to be an involution if2 = = Id.

    1.3.8. PROPOSITION. LetVbe a complex space. Then there exists a bijectionbetween the set of real forms in Vand the set of the anti-linear involutive automor-

    phisms ofV. Such bijection is obtained by: associating to each real form V0 V its conjugation operator c (see

    (1.3.1));

  • 7/29/2019 Not as Xies Cola

    21/177

    1.3. COMPLEXIFICATION AND REAL FORMS 11

    associating to each anti-linear involutive automorphism c ofVthe set ofits fixed points

    V0 =

    {v

    V: c(v) = v

    }.

    The above result suggests an interesting comparison between the operation of

    realification of a complex space and the operation of complexification of a real

    space. In Section 1.2 we saw that, roughly speaking, the operations of realification

    and of addition of a complex structure are mutually inverse; the realification is a

    canonical procedure, while the addition of a complex structure employs an addi-

    tional information, which is an automorphism J with J2 = Id. In this section wehave the opposite situation. The complexification is a canonical operation, while

    its inverse operation, which is the passage to a real form, involves an additional

    information, which is an anti-linear involutive automorphism.

    Let us look now at the complexification as a functorial construction. Let V1and V2 be real spaces; from the universal property of the complexification (Propo-sition 1.3.3) it follows that each linear operator T : V1 V2 admits a uniqueC-linear extension TC : VC1 VC2 . We have the following commutative diagram:

    VC1TC VC2

    V1

    TV2

    The operator TC is called the complexification ofT; more concretely, we have thatTC is given by:

    TC(v + i w) = T(v) + i T(w), v, w

    V1.

    It is immediate that:

    (1.3.3) (T1 T2)C = TC1 TC2 , IdC = Id,and, when T is invertible

    (1.3.4) (TC)1 = (T1)C.

    The identities (1.3.3) imply that the complexification V VC, T TC is a func-tor from the category of real vector spaces with morphisms the IR-linear operatorsto the category of complex vector spaces with morphisms the C-linear operators.

    Given a linear operator T : V1 V2, it is easy to see that:(1.3.5) Ker(TC) = (Ker(T))C, Im(TC) = Im(T)C;

    in the context of categories, the identities (1.3.5) say that the complexification is an

    exact functor, i.e., it takes short exact sequences into short exact sequences.

    If U V is a subspace, it is easy to see that the complexification iC of theinclusion i : U V is injective, and it therefore gives an identification of UC witha subspace of VC. More concretely, the subspace UC of VC is the direct sum ofthe two real subspaces U and i U ofVC; equivalently, UC is the complex subspaceof VC generated by the set U VC. However, not every subspace of VC is thecomplexification of some subspace of V. We have the following characterization:

  • 7/29/2019 Not as Xies Cola

    22/177

    12 1. SYMPLECTIC SPACES

    1.3.9. LEMMA. Let V be a real vector space and letZ VC be a complexsubspace of its complexification. Then, there exists a real subspace U

    V with

    Z= UC if and only ifZis invariant by conjugation, i.e.,c(Z) Z,

    where c is the conjugation relative to the real form V VC. IfZ= UC, then suchU is uniquely determined, and it is given explicitly by U = Z V.

    PROOF. Left to the reader in Exercise 1.7.

    Given real vector spaces V1 and V2, observe that the map:

    (1.3.6) Lin(V1, V2) T TC Lin(VC1 , VC2 )is IR-linear; we actually have the following:

    1.3.10. LEMMA. The map (1.3.6) takes Lin(V1, V2) isomorphically onto a realform in Lin(VC1 , V

    C2 ).

    PROOF. Since (VC2 )IR = V2 i V2, it is easy to see that:(1.3.7) Lin

    V1,

    VC2IR

    = Lin(V1, V2) iLin(V1, V2).

    From the universal property of the complexification, it follows that the restriction

    operator

    (1.3.8) Lin

    VC1 , VC2

    S = S|V1 LinV1, VC2 IRis an isomorphism. From (1.3.7) and (1.3.8) it follows:

    (1.3.9) LinVC1 , VC2 = Lin(V1, V2) Lin(V1, V2),where the two summands on the right of (1.3.9) are identified respectively with the

    image of (1.3.6) and with the same image multiplied by i.

    From Lemma 1.3.8 it follows in particular that the dual V = Lin(V, IR)can be identified with a real form of the dual of the complexification (VC) =Lin(VC,C) (compare with Example 1.2.8).

    Along the same lines of Lemma 1.3.9, in the next lemma we characterize the

    image of (1.3.6):

    1.3.11. LEMMA. Let V1, V2 be real vector spaces. Given a C-linear map

    S:

    VC1 VC2 , the following statements are equivalent:

    there exists an IR-linear map T : V1 V2 such thatS= TC; Spreserves real forms, i.e., S(V1) V2; Scommutes with conjugation, i.e., c S= S c, where c denotes the

    conjugation operators in VC1 andVC2 with respect to the real forms V1 and

    V2 respectively.

    When one (hence all) of the above conditions is satisfied, there exists a unique

    T Lin(V1, V2) such thatS= TC, which is given by the restriction ofS.

  • 7/29/2019 Not as Xies Cola

    23/177

    1.3. COMPLEXIFICATION AND REAL FORMS 13

    1.3.12. EXAMPLE. Let V1, V2 be real vector spaces; choosing bases for V1and V2, the same will be bases for the complexifications V

    C1 and V

    C2 (see Ex-

    ample 1.3.7). With respect to these bases, the matrix representation of a linearoperator T : V1 V2 is equal to the matrix representation of its complexificationTC : VC1 VC2 (compare with the result of Section 1.2, and more in particularwith formula (1.2.6)). In terms of matrix representations, the map (1.3.6) is simply

    the inclusion of the real matrices into the complex matrices.

    1.3.13. EXAMPLE. The real form in Lin(VC1 , VC2 ) defined in the statement

    of Lemma 1.3.10 corresponds to a conjugation operator in Lin(VC1 , VC2 ); given

    S Lin(VC1 , VC2 ), we denote by Sits conjugate operator. Explicitly, Sis givenby:

    S= c S c.For, using Proposition 1.3.8 and Lemma 1.3.11, it suffices to observe that S c S c defines an anti-linear involutive automorphism of Lin(VC1 , VC2 ) whosefixed point set is the image of (1.3.6). Observe that we have the identity:

    S(v) = S(v), v VC1 .In terms of bases, the matrix representation ofSis the complex conjugate of thematrix representation ofS.

    The theory presented in this section can be easily generalized to the case of

    multi-linear operators, anti-linear operators and operators with mixed multi-

    linearity, like sesquilinear operators. The latter case has special importance:

    1.3.14. DEFINITION. Given complex vector spaces V1, V2 and V, we say thata map B : V1 V2 Vis sesquilinear if for all v1 V1 the map B(v1, ) isanti-linear and for all v2 V2 the map B(, v2) is C-linear.

    IfV1 = V2 and if a real form is fixed in V, we say that a sesquilinear map Bis Hermitian (respectively, anti-Hermitian) ifB(v1, v2) = B(v2, v1) (respectively,B(v1, v2) = B(v2, v1)) for all v1, v2 V1.

    A Hermitian form in a complex space Vis a sesquilinear Hermitian map B :V V C; ifB is positive definite, i.e., B(v, v) > 0 for all v = 0, we also saythat B is a positive Hermitian product, or simply an Hermitian product, in V.

    In the same way that we define the complexification TC for an IR-linear op-erator, we can define the complexification BC of an IR-multilinear operator B :V1 Vp V as its unique extension to a C-multi-linear operator BC :VC1 VCp VC. Similarly, we can associate to an IR-linear operator itsunique extension TC : VC1 VC2 to an anti-linear operator, and to an IR-bilinearoperator B : V1V2 V its unique sesquilinear extension BCs : VC1 VC2 VC.

    In Exercise 1.8 the reader is asked to generalize the results of this section,

    in particular Proposition 1.3.3, Lemma 1.3.10 and Lemma 1.3.11, to the case of

    multi-linear, conjugate linear or sesquilinear operators.

  • 7/29/2019 Not as Xies Cola

    24/177

    14 1. SYMPLECTIC SPACES

    1.3.15. EXAMPLE. If V is a real vector space and B Bsym(V) is a sym-metric bilinear form on V, then the bilinear extension BC of B to VC is sym-

    metric; on the other hand, the sesquilinear extension BCs of B is a Hermitianform on VC. Similarly, the bilinear extension of an anti-symmetric bilinear formis anti-symmetric, while the sesquilinear extension of an anti-symmetric form is

    anti-Hermitian.

    The notions of kernel (see (1.1.5)), nondegeneracy and orthogonal complement

    (see (1.1.8)) for symmetric and anti-symmetric bilinear forms generalize in an ob-

    vious way to sesquilinear Hermitian and anti-Hermitian forms. IfB is symmetric(or anti-symmetric), it is easy to see that the condition of nondegeneracy of B isequivalent to the nondegeneracy of either BC or BCs . Moreover, ifB Bsym(V)is positive definite, i.e., B(v, v) > 0 for all v = 0, then its sesquilinear extensionBCs is also positive definite. Observe that the C-bilinear extension BC will be

    nondegenerate, but it is not positive definite (see Exercise 1.9).For instance, the canonical inner product ofIRn is given by:

    x, y =nj=1

    xjyj .

    Its sesquilinear extension defines the canonical Hermitian product in Cn, given by

    (1.3.10) z, wCs =nj=1

    zjwj ,

    while its C-bilinear extension is given by:

    z, wC =nj=1

    zjwj .

    1.3.16. REMARK. In the spirit of Definition 1.1.5, given a complex space Vand a Hermitian form B in V, we say that a C-linear operator T Lin(V) is B-

    Hermitian (respectively, B-anti-Hermitian) ifB(T , ) is a Hermitian (respectively,anti-Hermitian) form. We also say that T is B-unitary ifB(T , T ) = B.

    Given a real vector space V, B Bsym(V) and if T Lin(V) is a B-symmetric (respectively, B-anti-symmetric) operator, then its complexification TC

    in Lin(VC) is a BCs-Hermitian (respectively, BCs-anti-Hermitian) operator.IfT is B-orthogonal, then TC is BCs-unitary.

    1.3.1. Complex structures and complexifications. The aim of this subsec-

    tion is to show that there exists a natural correspondence between the complex

    structures of a real space V and certain direct sum decompositions of its complex-ification VC.

    Let V be a real vector space and let J : V V be a complex structure in V; wehave that JC is a C-linear automorphism of the complexification VC that satisfies(JC)2 = Id. It is then easy to see that VC decomposes as the direct sum ofthe two eigenspaces ofJC corresponding to the eigenvalues i and i respectively;

  • 7/29/2019 Not as Xies Cola

    25/177

    1.3. COMPLEXIFICATION AND REAL FORMS 15

    more explicitly, we define:

    Vh = v VC : JC(v) = iv,Va =

    v VC : JC(v) = iv.

    Then, Vh and Va are complex subspaces ofVC, and VC = Vh Va; the projec-tions onto the subspaces Vh and Va are given by:

    (1.3.11) h(v) =v iJC(v)

    2, a(v) =

    v + iJC(v)

    2, v VC.

    We call the spaces Vh and Va respectively the holomorphic and the anti-holomor-phic subspaces ofVC. Next proposition justifies the names of these spaces (seealso Example 1.3.18 below):

    1.3.17. PROPOSITION. Let V be a real vector space and J a complex struc-ture in V. Then, the projections h and a given in (1.3.11) restricted to V definerespectively a C-linear isomorphism of (V, J) onto Vh and a C-anti-linear iso-morphism of(V, J) onto Va.

    Proposition 1.3.17 tells us that, if we complexify a space V that already pos-sesses a complex structure J, we obtain a complex space VC that contains a copyof the original space (V, J) (the holomorphic subspace) and a copy of (V, J)(the anti-holomorphic subspace). Observe also that the holomorphic and the anti-

    holomorphic subspaces ofVC are mutually conjugate:

    Va = c

    Vh

    , Vh = c

    Va

    ,

    where c denotes the conjugation ofV

    C

    relative to the real form V.In our next example we make a short digression to show how the theory of thissubsection appears naturally in the context of calculus with functions of several

    complex variables.

    1.3.18. EXAMPLE. The construction of the holomorphic and the anti-holomor-

    phic subspaces appears naturally when one studies calculus of several complex

    variables, or, more generally, when one studies the geometry of complex manifolds.

    In this example we consider the space Cn, that will be thought as the real space

    IR2n endowed with the canonical complex structure. The real canonical basis ofCn (IR2n, J) will be denoted by:

    x1 , . . . , xn , y1 , . . . , yn;this is a basis ofIR2n adapted to J, and the corresponding complex basis ofCn isgiven by:

    x1, . . . ,

    xn

    .

    We now consider another complex space, given by the complexification (IR2n)C C2n. We denote by J the multiplication by the scalar i in Cn, while in C2n suchmultiplication will be denoted in the usual way v i v. Let JC : C2n C2n be

  • 7/29/2019 Not as Xies Cola

    26/177

    16 1. SYMPLECTIC SPACES

    the complexification ofJ, which defines the holomorphic and the anti-holomorphicsubspaces ofC2n.

    By Proposition 1.3.17, the projections h and a defined in (1.3.11) map thecanonical complex basis ofCn respectively into a basis of the holomorphic sub-

    space and a basis of the anti-holomorphic subspace ofC2n. These bases are usually

    denoted byzj

    nj=1

    andzj

    nj=1

    ; using (1.3.11) we compute explicitly:

    zj=

    1

    2

    xj i

    yj

    ,

    zj=

    1

    2

    xj+ i

    yj

    .

    Observe that the vector zj

    is conjugate to the vector zj

    .

    The notation xj

    , yj

    for the canonical basis of IR2n is justified by the iden-

    tification of vectors in IR2n with the partial derivative operators on differentiablefunctions f : IR2n

    IR. The complexification ofIR2n is therefore identified with

    the space of partial derivative operators acting on complex differentiable functionsf : IR2n C; in this notation, the CauchyRiemann equations, that characterizethe holomorphic functions, are given by setting equal to 0 the derivatives in thedirections of the anti-holomorphic subspace:

    (1.3.12)

    zjf = 0, j = 1, . . . , n .

    Observe that f satisfies (1.3.12) if and only if its differential at each point is aC-linear operator from Cn (IR2n, J) to C.

    We now show that the decomposition into holomorphic and anti-holomorphic

    subspace determines the complex structure:

    1.3.19. PROPOSITION. LetV be a real vector space and consider a direct sumof the complexification VC = Z1 Z2, where Z1 andZ2 are mutually conjugatesubspaces ofVC. Then, there exists a unique complex structure J on V such thatZ1 = Vh; moreover, for such J, it is also Z2 = Va.

    PROOF. The uniqueness follows from the fact that Vh is the graph ofJwhenwe use the isomorphism VC V V. For the existence, consider the uniqueC-linear operator in VC that has Z1 and Z2 as eigenspaces corresponding to theeigenvalues i and i respectively. Clearly, such operator commutes with the con-

    jugation and its square equals Id. From Lemma 1.3.11 it follows that it is of theform JC for some complex structure J : V V.

    Let now T be a C-linear endomorphism of(V, J), i.e., an IR-linear endomor-phism ofV such that TJ = JT; let TC be its complexification. It is easy to seethat the the holomorphic and the anti-holomorphic subspaces of VC are invariantby TC; moreover, we have the following commutative diagrams:

    (1.3.13)

    VT V

    h |V

    = =h |VVh

    TCVh

    VT V

    a |V

    = =a |VVa

    TCVa

  • 7/29/2019 Not as Xies Cola

    27/177

    1.3. COMPLEXIFICATION AND REAL FORMS 17

    It follows from Proposition 1.3.17 that the vertical arrows in the diagram on the left

    areC-linear isomorphisms of(V, J) with Vh and the vertical arrows in the diagram

    on the right are C-linear isomorphisms of(V, J) in Va.Let now (bj)

    nj=1 be a complex basis of (V, J) and let (bj , J(bj))

    nj=1 be the

    corresponding real basis ofV adapted to J. The latter is also a complex basis forVC (see Example 1.3.7). By Proposition 1.3.17, the vectors uj , uj defined by:

    (1.3.14) uj =bj iJ(bj)

    2 Vh, uj = bj + iJ(bj)

    2 Va, j = 1, . . . , n

    form a complex basis of (V, J). If T is represented by the matrix Z = A + B i,with A, B real matrices, in the basis (bj)

    nj=1 ofV

    C (hence it is represented by the

    matrix (1.2.6) with respect to the real basis ofV), then it follows from (1.3.13) thatthe matrix representation ofTC with respect to the basis (uj , uj)

    nj=1 ofV

    C is given

    by:

    (1.3.15) TC

    Z 00 Z

    .

    On the other hand, the matrix representation of TC with respect to the basis(bj , J(bj))

    nj=1 is again (1.2.6) (see Example 1.3.12). This shows in particular that

    the matrices in (1.2.6) and in (1.3.15) are equivalent (or conjugate, i.e., represent-

    ing the same operator in different bases).

    We summarize the above observations into the following:

    1.3.20. PROPOSITION. LetV be a real vector space andJ a complex structurein V. IfT is a C-linear endomorphism of(V, J), then:

    the trace ofT as an operator on V is twice the real part of the trace of Tas an operator on (V, J);

    the determinant of T as an operator on V is equal to the square of theabsolute value of the determinant of T as an operator on (V, J).

    More explicitly, ifA, B and real n n matrices, Z = A + B i andC is the matrixgiven in (1.2.6), then we have the following identities:

    tr(C) = 2tr(Z), det(C) = |det(Z)|2 ,where tr(U), det(U) denote respectively the trace and the determinant of the ma-trix U, and(), || denote respectively the real part and the absolute value of thecomplex number.

    1.3.21. REMARK. Suppose that V is endowed with a positive definite innerproduct g and that J : V V is a complex structure which is g-anti-symmetric,Then, we have J#g = g, i.e., J is g-orthogonal. The operator JC on VC willthen be anti-Hermitian (and unitary) with respect to the Hermitian product gCs

    in VC (see Remark 1.3.16). It is easy to see that the holomorphic and the anti-holomorphic subspaces ofJ are orthogonal with respect to gCs:

    gCs(v, w) = 0, v Vh, w Va.

  • 7/29/2019 Not as Xies Cola

    28/177

    18 1. SYMPLECTIC SPACES

    Using g and J, we can also define a Hermitian product gs in V by setting:

    gs(v, w) = g(v, w) + ig(v,Jw), v, w

    V.

    Actually, this is the unique Hermitian form in (V, J) that has g as its real part.We have the following relations:

    gCs

    h(v), h(w)

    =gs(v, w)

    2, gCs

    a(v), a(w)

    =

    gs(v, w)

    2, v, w V;

    they imply, in particular, that if (bj)nj=1 is an orthonormal complex basis of(V, J)

    with respect to gs, then the vectors

    2 uj ,

    2uj , j = 1, . . . , n, (see (1.3.14)) form

    an orthonormal real basis ofVC with respect to gCs . Also the vectors bj and J(bj),j = 1, . . . , n, form an orthonormal real basis ofV with respect to g, and thereforethey form a complex orthonormal basis of VC with respect to gCs . We concludethen that ifZ = A + B i (A, B real matrices), then the matrices in formulas (1.2.6)

    and (1.3.15) are unitarily equivalent, i.e., they represent the same complex operatorin different orthonormal bases.

    1.4. Symplectic Forms

    In this section we will study the symplectic vector spaces. We define the notion

    of symplectomorphism, which is the equivalence in the category of symplectic

    vector spaces, and we show that symplectic vector spaces of the same dimension

    are equivalent.

    1.4.1. DEFINITION. Let V be a real vector space; a symplectic form on V is ananti-symmetric nondegenerate bilinear form : V V IR. We say that (V, )is a symplectic vector space.

    1.4.2. REMARK. If Basym(V) is a possibly degenerate anti-symmetricbilinear form on V, then defines an anti-symmetric bilinear form on the quo-tient V /Ker(); it is easy to see that is nondegenerate, hence (V /Ker(), ) isa symplectic space.

    We start by giving a canonical form for the anti-symmetric bilinear forms; the

    proof is similar to the proof of Theorem 1.1.14.

    1.4.3. THEOREM. LetV be ap-dimensional vector space and Basym(V)an anti-symmetric bilinear form on V. Then, there exists a basis ofV with respectto which the matrix of (as a bilinear form) is given by:

    (1.4.1) 0n In 0nrIn 0n 0nr0rn 0rn 0r

    ,where r = dim(Ker()), p = 2n + r, and 0, 0 and I denote respectivelythe zero matrix, the zero square matrix and the identity matrix.

    PROOF. In first place, it is clear that, if a basis as in the thesis is found, then

    the last r vectors of this basis will be a basis for Ker(), from which we get r =dim(Ker()) and p = 2n + r.

  • 7/29/2019 Not as Xies Cola

    29/177

    1.4. SYMPLECTIC FORMS 19

    For the proof, we need to exhibit a basis (bi)pi=1 ofV such that:

    (1.4.2) (bi, bn+i

    ) =

    (bn+i

    , bi) = 1, i = 1, . . . , n ,

    and (bi, bj) = 0 otherwise. We use induction on p; if p 1 then necessarily = 0 and the result is trivial.

    Lets assumep > 1 and that the result is true for all vector spaces of dimensionless thanp. If = 0 the result is trivial; lets assume then that v, w V are chosenin such a way that (v, w) = 0, for instance (v, w) = 1. Then, it is easy to seethat is nondegenerate when restricted to the two-dimensional plane generated byv and w; from Proposition 1.1.10 it follows that:

    V = (IRv + IRw) (IRv + IRw).We now use the induction hypothesis to the restriction of to the (p 2)-dimen-sional vector space (IRv + IRw)

    , and we obtain a basis (b2, . . . , bn, bn+2, . . . , bp)of (IRv + IRw) in which takes the canonical form. This means that equality(1.4.2) holds for i = 2, . . . , n, and (bi, bj) = 0 otherwise. The desired basis forV is then obtained by setting b1 = v and bn+1 = w.

    1.4.4. COROLLARY. If (V, ) is a symplectic space, then V is even dimen-sional, and there exists a basis (bi)

    2ni=1 of V with respect to which the matrix of

    as a bilinear form is given by:

    (1.4.3)

    0 II 0

    ,

    where 0 andI denote respectively the zero and the identity n n matrices.1.4.5. DEFINITION. We say that (bi)2ni=1 is a symplectic basis of (V, ) if the

    matrix of as a bilinear form in this basis is given by (1.4.3).

    Observe that the matrix of the linear operator : V V is given by thetranspose of (1.4.3), i.e., it coincides with the matrix given in (1.2.2).

    Corollary 1.4.4 tells us that every symplectic space admits a symplectic basis.

    We now define sub-objects and morphisms in the category of symplectic spaces.

    1.4.6. DEFINITION. Let (V, ) be a symplectic space; We say that S is a sym-plectic subspace ifS V is a subspace and the restriction |SS is nondegenerate.Hence, (S, |SS) is a symplectic space.

    Let (V1, 1) and (V2, 2) be symplectic spaces; a linear operator T : V1

    V2

    is a symplectic map ifT#(2) = 1, i.e., if

    2(T(v), T(w)) = 1(v, w), v, w V1.We say that T is a symplectomorphism if T is an isomorphism and a symplecticmap.

    A symplectomorphism takes symplectic bases in symplectic bases; conversely,

    if T : V1 V2 is a linear map that takes some symplectic basis of V1 into somesymplectic basis ofV2, then T is a symplectomorphism.

  • 7/29/2019 Not as Xies Cola

    30/177

    20 1. SYMPLECTIC SPACES

    In terms of the linear operators 1 Lin(V1, V1 ) and 2 Lin(V2, V2 ), amap T

    Lin(V1, V2) is symplectic if and only if:

    (1.4.4) T 2 T = 1.1.4.7. REMARK. Observe that the right hand side of equality (1.4.4) is an iso-

    morphism, from which it follows that every symplectomorphism T is an injectivemap. In particular, the image T(V1) is always a symplectic subspace ofV2.

    1.4.8. EXAMPLE. We define a symplectic form in IR2n by setting:

    (1.4.5)

    (v1, w1), (v2, w2)

    = v1, w2 w1, v2,for v1, v2, w1, w2 IRn, where , denotes the canonical inner product of IRn.We say that (1.4.5) is the canonical symplectic form of IR2n; the canonical basisofIR2n is a symplectic basis for , hence the matrix of (as a bilinear map) withrespect to the canonical basis of IR2n is (1.4.3).

    The existence of a symplectic basis for a symplectic space (Corollary 1.4.4)

    implies that every symplectic space admits a symplectomorphism with (IR2n, ),hence the proof of every theorem concerning symplectic spaces can be reduced to

    the case of(IR2n, ).We can also define a canonical symplectic form in IRn IRn by setting:

    (v1, 1), (v2, 2)

    = 2(v1) 1(v2),where v1, v2 IRn and 1, 2 IRn. Again, the canonical basis ofIRn IRnis a symplectic basis for the canonical symplectic form of IRn IRn.

    1.4.9. REMARK. Denoting by (dq1, . . . , dqn, dp1, . . . , dpn) the canonical ba-

    sis ofIR

    2n

    (dual of the canonical basis ofIR

    2n

    ), the canonical symplectic form ofIR2n is given by:

    =ni=1

    dqi dpi.

    It follows easily:

    n = . . . = (1)n(n1)2 dq1 . . . dqn dp1 . . . dpn.Hence, n is a volume form in IR2n; for all symplectomorphism T of(IR2n, ) wetherefore have:

    T#(n) = n = det(T) n,

    from which it follows det(T) = 1. In general, not every linear map T with

    det(T) = 1 is a symplectomorphism of(IR2n

    , ); when n = 1 the symplectic form is a volume form, hence T is a symplectomorphism if and only ifdet(T) = 1.

    The symplectomorphisms of a symplectic space (V, ) form a group by com-position.

    1.4.10. DEFINITION. Let (V, ) be a symplectic space; the symplectic groupof (V, ) is the group of all symplectomorphisms of (V, ), denoted by Sp(V, ).We denote by Sp(2n, IR) the symplectic group ofIR2n endowed with the canonicalsymplectic form.

  • 7/29/2019 Not as Xies Cola

    31/177

    1.4. SYMPLECTIC FORMS 21

    Using a symplectic basis of (V, ), a map T Lin(V) is a symplectomor-phism if and only if the matrix M that represents T in such basis satisfies:

    (1.4.6) MM = ,

    where is the matrix given in (1.4.3). Writing

    (1.4.7) T

    A BC D

    ,

    then (1.4.6) is equivalent to the following relations:

    (1.4.8) DA BC = I, AC and BD are symmetric,where A,B,C,D are n n matrices, I is the n n identity matrix, and meanstranspose (see Exercise 1.10). A matrix of the form (1.4.7) satisfying (1.4.8) will

    be called a symplectic matrix.

    We define direct sum of symplectic spaces.

    1.4.11. DEFINITION. Given symplectic spaces (V1, 1) and (V2, 2), we de-fine a symplectic form = 1 2 on V1 V2 by setting:

    ((v1, v2), (w1, w2)) = 1(v1, w1) + 2(v2, w2), v1, w1 V1, v2, w2 V2.The space (V1 V2, 1 2) is called the direct sum of the symplectic spaces(V1, 1), (V2, 2).

    If (V, ) is a symplectic space, two subspaces S1, S2 V are said to be -orthogonal if(v1, v2) = 0 for all vi Si, i = 1, 2. IfV = S1 S2 with S1 andS2 -orthogonal, then it is easy to see that both S1 and S2 are symplectic subspacesof(V, ); in this case we say that V is the symplectic direct sum of the subspaces

    S1 and S2.Observe that the notion of direct sum for symplectic spaces is not meant as a

    sum in a categorical sense, i.e., it is not true that a symplectic map on a direct sum

    V1 V2 is determined by its restriction to V1 and V2 (see Exercise 1.15).1.4.12. EXAMPLE. If Ti : Vi Vi , i = 1, 2, are symplectic maps, then the

    map T = T1 T2 : V1 V2 V1 V2 defined by:T(v1, v2) = (T1(v1), T2(v2)), vi Vi, i = 1, 2,

    is also symplectic. If both T1 and T2 are symplectomorphisms, then also T is asymplectomorphism.

    One needs to be careful with the notion of direct sum of symplectic spaceswhen working with symplectic bases; more explicitly, the concatenation of a sym-

    plectic basis (bi)2ni=1 ofV1 and a symplectic basis (b

    j)2mj=1 ofV2 is not a symplectic

    basis of V1 V2. In order to obtain a symplectic basis of V1 V2 we need torearrange the vectors as follows:

    (b1, . . . , bn, b1, . . . , b

    m, bn+1, . . . , b2n, b

    m+1, . . . , b

    2m).

    Similar problems are encountered when dealing with symplectic matrices: the sim-

    ple juxtaposition of along the diagonal of an element ofSp(2n, IR) and an element

  • 7/29/2019 Not as Xies Cola

    32/177

    22 1. SYMPLECTIC SPACES

    ofSp(2m, IR) does notproduce an element ofSp(2(n+ m), IR); in order to obtaina symplectic matrix it is necessary to perform a suitable permutation of the rows

    and the columns of such juxtaposition.

    1.4.1. Isotropic and Lagrangian subspaces. In this subsection we consider

    a fixed symplectic space (V, ), with dim(V) = 2n.

    1.4.13. DEFINITION. A subspace S V is said to be isotropic if|SS = 0.Observe that S is isotropic if and only if it is contained in its orthogonal S

    with respect to ; from Proposition 1.1.8 we have:

    (1.4.9) dim(S) + dim(S) = 2n,

    from which it follows that the dimension of an isotropic subspace is at most n.

    Observe that the notion of isotropic subspace is, roughly speaking, opposite tothe notion of symplectic subspace; for, by Proposition 1.1.10, S is a symplecticsubspace iffS S = {0}.

    We have the following:

    1.4.14. LEMMA. Let L V be a subspace; the following statements areequivalent:

    L is maximal isotropic, i.e., L is isotropic and it is not properly containedin any other isotropic subspace ofV;

    L = L; L is isotropic anddim(L) = n.

    PROOF. IfL is maximal isotropic, then L L

    and for v L

    the subspaceL + IR v is isotropic and it contains L. It follows that L = L + IR v, hencev L and L = L. If L = L, then L is isotropic, and from (1.4.9) it followsthat dim(L) = n. Finally, if L is isotropic and dim(L) = n, then L is maximalisotropic, because the dimension of an isotropic subspace is at most n.

    1.4.15. DEFINITION. A subspace L V is said to be Lagrangian subspace ifit satisfies one (hence all) of the statements in Lemma 1.4.14.

    1.4.16. EXAMPLE. The subspaces {0} IRn and IRn {0} are Lagrangiansubspaces ofIR2n endowed with the canonical symplectic structure. Given a linearmap T Lin(IRn), then its graph Graph(T) = {v + T(v) : v IRn} is aLagrangian subspace of IR2n endowed with the canonical symplectic structure ifand only ifT is symmetric with respect to the canonical inner product ofIRn.

    1.4.17. EXAMPLE. If S V is an isotropic subspace, then the kernel of therestriction of to S is the subspace (S) S = S (see Corollary 1.1.9). Itfollows that defines by passing to the quotient a symplectic form in S/S(Remark 1.4.2).

    In the following definition we relate symplectic forms and complex structures

    on V:

  • 7/29/2019 Not as Xies Cola

    33/177

    1.4. SYMPLECTIC FORMS 23

    1.4.18. DEFINITION. A complex structure J : V V is said to be compatiblewith the symplectic form if (

    , J

    ) is an inner product in V, i.e., a symmetric,

    positive definite bilinear form on V. More explicitly, J is compatible with if:

    (Jv,w) = (v,Jw), v, w V,and if(Jv,v) > 0 for all v = 0.

    1.4.19. EXAMPLE. The canonical complex structure of IR2n (Example 1.2.2)is compatible with the canonical symplectic structure of IR2n. The inner product(, J) is simply the canonical inner product of IR2n. It follows that every sym-plectic space admits a complex structure compatible with the symplectic form: it

    is enough to define J by the matrix (1.2.2) with respect to any fixed symplectic ba-sis. Such basis will then be an orthonormal basis with respect to the inner product

    (

    , J

    ).

    Lets assume that J is a given complex structure on V which is compatiblewith , and lets denote by g the inner product (, J); J is a symplectomorphismof(V, ) (see Exercise 1.16) and the following identity holds:

    g(J, ) = .A compatible complex structure Jcan be used to construct a Lagrangian which

    is complementary to a given Lagrangian:

    1.4.20. LEMMA. If L V is a Lagrangian subspace and J is a complexstructure compatible with , then V = L J(L).

    PROOF. It suffices to observe that L and J(L) are orthogonal subspaces withrespect to the inner product g.

    1.4.21. COROLLARY. Every Lagrangian subspace admits a complementary

    Lagrangian subspace.

    PROOF. It follows from Lemma 1.4.20, observing that J(L) is Lagrangian,since J is a symplectomorphism (Exercise 1.16).

    We can define a complex valued sesquilinear form gs (see Definition 1.3.14) inthe complex space (V, J) by setting:

    (1.4.10) gs(v, w) = g(v, w) i (v, w).

    It is easy to see that gs is a positive Hermitian product in (V, J).Recall from Remark 1.3.16 that a C-linear endomorphism is gs-unitary whengs(T, T) = gs; in this situation we also say that T preserves gs. We have thefollowing:

    1.4.22. PROPOSITION. Let T Lin(V) be an IR-linear map; the followingstatements are equivalent:

    T is C-linear in (V, J) andgs-unitary; T is orthogonal with respect to g andT Sp(V, ).

  • 7/29/2019 Not as Xies Cola

    34/177

    24 1. SYMPLECTIC SPACES

    PROOF. IfT is C-linear and gs-unitary, then T preserves gs, hence it preservesseparately its real part, which is g, and its imaginary part, which is

    . Hence T

    is an orthogonal symplectomorphism.Conversely, ifT is an orthogonal symplectomorphism, then the following iden-

    tities hold:

    T g T = g, T T = , = g J,considering g and as linear operators in Lin(V, V) (see Example 1.1.3)). Itfollows easily that J T = T J, i.e., T is C-linear. Since T preserves both thereal and the imaginary part of gs, we conclude that T is gs-unitary.

    1.4.23. EXAMPLE. The canonical complex structure J of IR2n (see Exam-ple 1.4.8) is compatible with its canonical symplectic structure (Example 1.2.2),

    and the inner product g corresponds to the canonical inner product ofIR2n. If weidentify (IR2n, J) with Cn (Example 1.2.2), the Hermitian product gs coincides

    with the canonical Hermitian product ofCn given in (1.3.10).

    1.4.24. REMARK. Observe that if(V, J) is a complex space endowed with aHermitian product gs, then the real part ofgs is a positive inner product g on V andthe imaginary part ofgs is a symplectic form on V; moreover, defining as minusthe imaginary part ofgs, it follows that J is compatible with and g = (, J).

    1.4.25. REMARK. If V is a real vector space, g is a positive inner producton V and J is a complex structure which is g-anti-symmetric (or, equivalently,g-orthogonal), then we get a symplectic form on V by setting = g(J, ). Thecomplex structure J will then be compatible with , and g = (, J). Again, wealso get a Hermitian product gs in (V, J) defined by (1.4.10).

    We have the following relation between Lagrangian subspaces and the Hermit-

    ian product gs:

    1.4.26. LEMMA. A subspace L V is Lagrangian if and only if it is a realform which is preserved by gs, i.e., V = L J(L) andgs(L L) IR.

    PROOF. It follows from Lemma 1.4.20 and the observation that the imaginary

    part ofgs equals . As a corollary, we now prove that the group of gs-unitary isomorphisms of

    (V, J) act transitively on the set of Lagrangian subspaces of (V, ):

    1.4.27. COROLLARY. Given any pair of Lagrangian subspaces L1, L2 of V,

    there exists a C-linear isomorphism T of (V, J) which is gs-unitary and such thatT(L1) = L2.

    PROOF. Let (bj)nj=1 be an orthonormal basis of L1 with respect to the inner

    product g; since L1 is a real form of (V, J), it follows that (bj)nj=1 is a complex

    basis of (V, J) (see Example 1.3.7). Moreover, since gs is real on L1, it followsthat (bj)

    nj=1 is an orthonormal basis of (V, J) with respect to gs. Similarly, we

    consider a basis (bj)nj=1 of L2 which is orthonormal with respect to g, and we

    obtain that (bj)nj=1 is a gs-orthonormal basis of(V, J). It follows that the C-linear

  • 7/29/2019 Not as Xies Cola

    35/177

    1.4. SYMPLECTIC FORMS 25

    isomorphism T defined by T(bj) = bj , for all j = 1, . . . , n, is unitary and satisfies

    T(L1) = L2.

    It follows that also the symplectic group acts transitively on the set of La-

    grangian subspaces:

    1.4.28. COROLLARY. Given any pair L1, L2 of Lagrangian subspaces of thesymplectic space (V, ), there exists a symplectomorphism T Sp(V, ) such thatT(L1) = L2.

    PROOF. It follows from Corollary 1.4.27, observing that every gs-unitary mapis a symplectomorphism (Proposition 1.4.22).

    1.4.29. REMARK. For later use, we will mention a mild refinement of the

    result of Corollary 1.4.27. Given Lagrangian subspaces L1, L2

    V and chosen

    orientations O1 and O2 respectively on the spaces L1 and L2, it is possible to finda C-linear and gs-unitary endomorphism T of (V, J) such that T(L1) = L2 andsuch that T|L1 : L1 L2 is positively oriented. To see this, it suffices to choosein the proof of Corollary 1.4.27 the g-orthonormal bases (bj)

    nj=1 and (b

    j)nj=1 ofL1

    and L2 respectively in such a way that they are positively oriented.

    1.4.30. REMARK. Given a Lagrangian subspace L0 V, then it is alwayspossible to find a basis (bj)

    2nj=1 ofV which is at the same time symplectic, adapted

    to J, and such that (bj)nj=1 is a basis of L0. For, if(bj)

    nj=1 is a g-orthonormal

    basis ofL0, then the basis defined in (1.2.5) satisfies the required properties; more-over, such basis is g-orthonormal and the complex basis (bj)

    nj=1 of (V, J) is gs-

    orthonormal. We therefore obtain a basis that puts simultaneously all the objects

    (V, ,J,g ,gs, L0) in their canonical forms.

    In the spirit of Remark 1.4.24 and Remark 1.4.25, one can ask himself whether

    given a real space V endowed with a symplectic form and a positive inner productg, it is possible to construct a complex structure Jand a Hermitian product gs whichare naturally associated to g and . If one requires the condition = g(J, ),then this is clearly impossible in general, because there exists a unique operator

    H Lin(V) such that = g(H, ), and such H does not in general satisfyH2 = Id.

    We conclude the subsection with a result in this direction:

    1.4.31. PROPOSITION. Let(V, ) be a symplectic space andg a positive inner

    product in V. Then there exists a unique complex structure J in V which is g-anti-symmetric (or, equivalently, g-orthogonal) and compatible with .

    PROOF. The uniqueness is the hard part of the thesis, which we now prove.

    Suppose that J is a given g-anti-symmetric complex structure in V which is com-patible with , and let H Lin(V) be the unique operator such that = g(H, ).Then, H is a g-anti-symmetric isomorphism ofV.

    The compatibility ofJ with is equivalent to the condition that g(HJ, ) be asymmetric bilinear form on V which is negative definite. By the usual identification

  • 7/29/2019 Not as Xies Cola

    36/177

    26 1. SYMPLECTIC SPACES

    of linear and bilinear maps, we see that the g-anti-symmetry property ofH and J,together with the g-symmetry ofHJ are expressed by the following relations:

    g J = J g, g H = H g, g H J = J H g,from which it follows easily that H J = J H.

    We now consider the complexifications JC, HC Lin(VC) and the uniquesesquilinear extension gCs ofg to VC; clearly, gCs is a positive Hermitian productin VC, with respect to which HC and JC are anti-Hermitian operators (see Exam-ple 1.3.15 and Remark 1.3.16); moreover, HC JC = JC HC and (JC)2 = Id.

    Since HC is gCs-anti-Hermitian, then HC can be diagonalized in a gCs-ortho-normal basis of VC (see Exercise 1.18); its eigenvalues are pure imaginary (nonzero, because HC is invertible), and since HC commutes with the conjugation, itfollows that eigenspaces of HC corresponding to two conjugate eigenvalues are

    mutually conjugate (see Lemma 1.3.11). We can then write a gCs

    -orthogonal de-composition:

    VC =r

    j=1

    Zi j r

    j=1

    Zi j ,

    where j > 0 for all j, Zi the eigenspace ofHC corresponding to the eigenvaluei; also, Zi is the conjugate ofZi.

    Since JC commutes with HC, it follows that the eigenspaces of HC are invari-ant by JC. The restriction ofJC to each Zij is an anti-Hermitian operator whosesquare is Id, from which it follows that such restriction is diagonalizable, and itspossible eigenvalues are i and i. The restriction ofgCs(JC HC , ) to Zij , that

    coincides with the restriction of ij g

    Cs

    (J

    C

    , )) must be Hermitian and negativedefinite, from which it follows that the unique eigenvalue of the restriction of JCto Zij must be equal to i.

    We conclude that the restriction of JC to Zij is the operator of multiplicationby i, and the restriction of JC to Zij is the operator of multiplication by i;such conditions determine JC, which shows the uniqueness ofJ.

    For the existence, simply consider the unique complex structure J on V whoseholomorphic space coincides with

    j Zij (see Proposition 1.3.19).

    1.4.2. Lagrangian decompositions of a symplectic space. In this subsection

    we study the properties of Lagrangian decompositions of a symplectic space, that

    will be fundamental in the study of the Lagrangian Grassmannian in Section 2.5.

    Throughout this subsection we will fix a symplectic space (V, ), with dim(V) =2n. We start with a definition:

    1.4.32. DEFINITION. A Lagrangian decomposition of(V, ) is a pair (L0, L1)of Lagrangian subspaces ofV with V = L0 L1.

    1.4.33. EXAMPLE. The pair (IRn {0}, {0} IRn) is a Lagrangian decom-position ofIR2n endowed with the canonical symplectic structure. More generally,ifL V is a Lagrangian subspace and J is a complex structure on V compatible

  • 7/29/2019 Not as Xies Cola

    37/177

    1.4. SYMPLECTIC FORMS 27

    with , then (L, J(L)) is a Lagrangian decomposition of(V, ) (see Lemma 1.4.20and the proof of Corollary 1.4.21).

    Given a Lagrangian decomposition (L0, L1) of(V, ), we define a a map:

    L0,L1 : L1 L0by setting

    (1.4.11) L0,L1(v) = (v, )|L0for all v L1; it is easy to see that L0,L1 is an isomorphism (see Exercise 1.19).

    1.4.34. REMARK. The isomorphism L0,L1 gives us an identification of L1with the dual space L0, but the reader should be careful when using this identifi-cation for the following reason. The isomorphism L0,L1 induces an isomorphism(L0,L1)

    : L0 L0 L1; however, (L0,L1) does not coincide with L1,L0 ,but with its opposite:(1.4.12) (L0,L1)

    = L1,L0 .IfL V is a Lagrangian subspace, we also define an isomorphism:

    L : V /L L,by setting L(v + L) = (v, )|L.

    Given a Lagrangian decomposition (L0, L1) of(V, ), we have the followingcommutative diagram of isomorphisms:

    (1.4.13) L1L0,L1

    ""EE

    EEEE

    EE

    q

    L0

    V /L0

    L0

  • 7/29/2019 Not as Xies Cola

    38/177

    28 1. SYMPLECTIC SPACES

    PROOF. Let (bi)ni=1 be a basis ofL0 and let (b

    i)ni=1 be the basis of L

    0 which

    corresponds to (bi)ni=1 by the given isomorphism. Using Lemma 1.4.35 we can

    find symplectic bases (bi)2ni=1 and (bi)2ni=1 of V in such a way that (bi)2ni=n+1 and(bi)

    2ni=n+1 are bases of L1 and L

    1 respectively; to conclude the proof one simply

    chooses T such that T(bi) = bi, i = 1, . . . , 2n.

    1.4.37. COROLLARY. If L0 V is a Lagrangian subspace, then every iso-morphism ofL0 extends to a symplectomorphism of V.

    PROOF. Choose a Lagrangian L1 complementary to L0 (see Corollary 1.4.21)and apply Corollary 1.4.36.

    The technique of extending bases of Lagrangians to symplectic bases of the

    symplectic space may be used to give an alternative proof of Corollary 1.4.28.

    Roughly speaking, Corollary 1.4.28 tells us that Lagrangian subspaces are indis-

    tinguishable from the viewpoint of the symplectic structure; our next Propositiontells us that the only invariant of a pair (L0, L1) of Lagrangian subspaces is thedimension of their intersection L0 L1:

    1.4.38. PROPOSITION. Given three Lagrangian subspaces L0, L , L V with

    dim(L0 L) = dim(L0 L), there exists a symplectomorphism T of(V, ) suchthatT(L0) = L0 andT(L) = L

    .

    PROOF. By Corollary 1.4.37, there exists a symplectomorphism of(V, ) thattakes L0 into itself and L0 L onto L0 L; we can therefore assume without lossof generality that L0 L = L0 L.

    Set S = L0 L = L0 L; clearly S is isotropic and L0, L , L S. Wehave a symplectic form in S/Sobtained from by passing to the quotient (see

    Example 1.4.17).Denote by q : S S/S the quotient map; it is easy to see that q(L0),

    q(L) and q(L) are Lagrangian subspaces of(S/S,); moreover,

    q(L0), q(L)

    and

    q(L0), q(L)

    are both Lagrangian decompositions of S/S and hence there

    exists a symplectomorphism T of (S/S,) such that T

    q(L0)

    = q(L0) and

    T

    q(L)

    = q(L) (see Corollary 1.4.28). The required symplectomorphism T Sp(V, ) is obtained from the following Lemma.

    1.4.39. LEMMA. Let L0 V be a Lagrangian subspace and let S L0be any subspace. Consider the quotient symplectic form on S/S; then, givenany symplectomorphism T of (S/S,) with T(q(L0)) = q(L0), there exists asymplectomorphism T of (V, ) such that T(S) = S (hence also T(S) = S),T(L0) = L0, and such that the following diagram commutes

    ST|S //

    q

    S

    q

    S/S

    T

    // S/S

    where q : S S/S denotes the quotient map.

  • 7/29/2019 Not as Xies Cola

    39/177

    EXERCISES FOR CHAPTER 1 29

    PROOF. Write L0 = S R; hence L0 = So Ro, where So and Ro are theannihilators ofS and R respectively. Let L1 be any complementary Lagrangian to

    L0 in V (Corollary 1.4.21). We have:L1 =

    1L0,L1

    (So) 1L0,L1(Ro).We obtain a direct sum decomposition V = V1 V2 into -orthogonal subspacesgiven by:

    V1 = S 1L0,L1(Ro), V2 = R 1L0,L1(So),from which it follows that V is direct sum of the symplectic spaces V1 and V2.

    Observe that S = V2 S, hence the quotient map q restricts to a symplec-tomorphism ofV2 into S

    /S; therefore, we have a unique symplectomorphism T

    ofV2 such that the diagram:

    V2T //

    q|V2

    V2

    q|V2

    S/ST

    // S/S

    commutes. Since T preserves q(L0) it follows that T preserves R; we then define

    T by setting T|V1 = Id and T|V2 = T (see Example 1.4.12). 1.4.40. REMARK. We claim that one can actually choose the symplectomor-

    phism T in the thesis of Proposition 1.4.38 in such a way that T restricts to a posi-tively oriented isomorphism ofL0; namely, ifdim(L0 L) = dim(L0 L) = 0then this claim follows directly from Corollary 1.4.36. For the general case, we

    observe that in the last part of the proof of Lemma 1.4.39 one can define T|V1

    to

    be any symplectomorphism ofV1 which preserves S (while T|V2 = T is kept un-changed); since S is Lagrangian in V1, using Corollary 1.4.37, we get that T|S canbe choosen to be any isomorphism A ofSgiven a priori (and T|R does not dependon A). Since dim(S) 1, this freedom in the choice of A can be used to adjustthe orientation ofT|L0 .

    Exercises for Chapter 1

    EXERCISE 1.1. Show that the isomorphism between the spaces Lin(V, W)and B(V, W) given in (1.1.1) is natural in the sense that it gives a natural iso-m