Notes by Maksim Maydanskiy.

Similar documents
MORSE HOMOLOGY. Contents. Manifolds are closed. Fields are Z/2.

L19: Fredholm theory. where E u = u T X and J u = Formally, J-holomorphic curves are just 1

LECTURE 28: VECTOR BUNDLES AND FIBER BUNDLES

LECTURE 6: J-HOLOMORPHIC CURVES AND APPLICATIONS

Determinant lines and determinant line bundles

LECTURE 15: COMPLETENESS AND CONVEXITY

LECTURE 3 Functional spaces on manifolds

Introduction to Index Theory. Elmar Schrohe Institut für Analysis

LECTURE 11: TRANSVERSALITY

TOEPLITZ OPERATORS. Toeplitz studied infinite matrices with NW-SE diagonals constant. f e C :

LECTURE: KOBORDISMENTHEORIE, WINTER TERM 2011/12; SUMMARY AND LITERATURE

Geometric Composition in Quilted Floer Theory

Elliptic Regularity. Throughout we assume all vector bundles are smooth bundles with metrics over a Riemannian manifold X n.

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

Lecture 4: Harmonic forms

Holomorphic line bundles

1.4 The Jacobian of a map

Lecture Fish 3. Joel W. Fish. July 4, 2015

MILNOR SEMINAR: DIFFERENTIAL FORMS AND CHERN CLASSES

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM

THE UNIFORMISATION THEOREM OF RIEMANN SURFACES

Index theory on manifolds with corners: Generalized Gauss-Bonnet formulas

DIFFERENTIAL GEOMETRY, LECTURE 16-17, JULY 14-17

CHAPTER 8. Smoothing operators

MASLOV INDEX. Contents. Drawing from [1]. 1. Outline

Practice Qualifying Exam Questions, Differentiable Manifolds, Fall, 2009.

Nodal symplectic spheres in CP 2 with positive self intersection

5 Compact linear operators

Introduction to the Baum-Connes conjecture

DIFFERENTIAL GEOMETRY. LECTURE 12-13,

LECTURE 10: THE PARALLEL TRANSPORT

An introduction to cobordism

1 )(y 0) {1}. Thus, the total count of points in (F 1 (y)) is equal to deg y0

Math 215B: Solutions 3

A Crash Course of Floer Homology for Lagrangian Intersections

THE POINCARE-HOPF THEOREM

MATH 263: PROBLEM SET 1: BUNDLES, SHEAVES AND HODGE THEORY

J-holomorphic curves in symplectic geometry

Cup product and intersection

SPECTRAL THEORY EVAN JENKINS

THE GUTZWILLER TRACE FORMULA TORONTO 1/14 1/17, 2008

THE HODGE DECOMPOSITION

Morse homology. Michael Landry. April 2014

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M =

1. Classifying Spaces. Classifying Spaces

Dirac Operator. Göttingen Mathematical Institute. Paul Baum Penn State 6 February, 2017

Real Analysis Prelim Questions Day 1 August 27, 2013

Hyperkähler geometry lecture 3

CHAPTER 3. Gauss map. In this chapter we will study the Gauss map of surfaces in R 3.

SYMPLECTIC GEOMETRY: LECTURE 5

GEOMETRIC QUANTIZATION

Lecture 8. Connections

Topics in Geometry: Mirror Symmetry

NOTES ON DIFFERENTIAL FORMS. PART 5: DE RHAM COHOMOLOGY

DEVELOPMENT OF MORSE THEORY

Math Solutions to homework 5

The Dirichlet-to-Neumann operator

SYMPLECTIC MANIFOLDS, GEOMETRIC QUANTIZATION, AND UNITARY REPRESENTATIONS OF LIE GROUPS. 1. Introduction

L 2 Geometry of the Symplectomorphism Group

Chern forms and the Fredholm determinant

Mathematical foundations - linear algebra

Embedded contact homology and its applications

MATH 583A REVIEW SESSION #1

LECTURE 9: THE WHITNEY EMBEDDING THEOREM

LECTURE 2. (TEXED): IN CLASS: PROBABLY LECTURE 3. MANIFOLDS 1. FALL TANGENT VECTORS.

LECTURE 5: SOME BASIC CONSTRUCTIONS IN SYMPLECTIC TOPOLOGY

Max-Planck-Institut für Mathematik Bonn

THREE APPROACHES TO MORSE-BOTT HOMOLOGY

Math 6455 Nov 1, Differential Geometry I Fall 2006, Georgia Tech

Differential Topology Final Exam With Solutions

An Invitation to Geometric Quantization

Moduli spaces of graphs and homology operations on loop spaces of manifolds

Lecture III: Neighbourhoods

Symplectic topology and algebraic geometry II: Lagrangian submanifolds

Convex Symplectic Manifolds

MATH 205C: STATIONARY PHASE LEMMA

1 Math 241A-B Homework Problem List for F2015 and W2016

Morse Theory for Lagrange Multipliers

The Symmetric Space for SL n (R)

Finite-dimensional spaces. C n is the space of n-tuples x = (x 1,..., x n ) of complex numbers. It is a Hilbert space with the inner product

REAL INSTANTONS, DIRAC OPERATORS AND QUATERNIONIC CLASSIFYING SPACES PAUL NORBURY AND MARC SANDERS Abstract. Let M(k; SO(n)) be the moduli space of ba

EXOTIC SMOOTH STRUCTURES ON TOPOLOGICAL FIBRE BUNDLES B

Some aspects of the Geodesic flow

FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASS 48

j=1 u 1jv 1j. 1/ 2 Lemma 1. An orthogonal set of vectors must be linearly independent.

COMPUTABILITY AND THE GROWTH RATE OF SYMPLECTIC HOMOLOGY

Your first day at work MATH 806 (Fall 2015)

LECTURE 26: THE CHERN-WEIL THEORY

CONSIDERATION OF COMPACT MINIMAL SURFACES IN 4-DIMENSIONAL FLAT TORI IN TERMS OF DEGENERATE GAUSS MAP

LINEAR ALGEBRA KNOWLEDGE SURVEY

Bordism and the Pontryagin-Thom Theorem

Lefschetz Fibrations and Pseudoholomorphic Curves. Michael Usher, MIT

Classification of (n 1)-connected 2n-dimensional manifolds and the discovery of exotic spheres

INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD

Fredholm Theory. April 25, 2018

Measures on spaces of Riemannian metrics CMS meeting December 6, 2015

NOTES ON NEWLANDER-NIRENBERG THEOREM XU WANG

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

The Spinor Representation

The topology of positive scalar curvature ICM Section Topology Seoul, August 2014

Transcription:

SPECTRAL FLOW IN MORSE THEORY. 1 Introduction Notes by Maksim Maydanskiy. Spectral flow is a general formula or computing the Fredholm index of an operator d ds +A(s) : L1,2 (R, H) L 2 (R, H) for a family A s of (almost) self-adjoint operators on a Hilbert space H. The simplest version of such an operator is one where H = R n. This appears in Morse theory when studying the spaces of gradient trajectories. Thus Morse theory is a natural setting to consider such operators, with the view towards other geometric applications in various flavours of Floer theory. These notes are based on the two talks on spectral flow that I gave at the Stanford student symplectic geometry seminar. The content is not original, but rather a compilation of various sources. There is a number of references for the basic setup of Morse theory, of which the most detailed and concise is the note of Alex Ritter [7]. The approach to spectral flow is based on the book of Kronheimer and Mrowka [4]. The goal, to paraphrase a famous saying, was to give the simplest exposition to spectral flow possible, but not simpler. 2 Setup. 2.1 Basic setup We start with a manifold M of dimension n with f a morse function on M and g a Riemannian metric on it. We denote by f g the gradient of f with respect to g that is g(f g, ) = df( ). This is a vector field on M. Let p and q be critical points of f. We consider the space of smooth paths P = {γ : R M lim s + = p, lim s = q}. Such a path is a gradient trajectory if dγ(s) f ds g(γ(s)) = 0. For any γ P, even one that doesn t solve this equation, the left hand side of it defines a map S g (γ) : R γ T M. Putting the spaces γ T M for all γ P together gives a bundle B P and we view S g as a section of this bundle. The gradient trajectories are then the intersection of S g with zero. The modern or function-analytic or moduli space approach to Morse homology studies this space of trajectories and uses it and related moduli spaces to define algebraic invariants of M. The advantage of this approach (as opposed to classical or geometric one) is that it generalizes to other settings (see a nice table in [7, Lecture 10]). To understand how this zero set of the section F g behaves it helps to consider zero-sets of sections of finite dimensional bundles first. When we later consider the infinite-dimensional version some of the words will get translated. We mention these translation in parenthesis. With this in mind, let B n+m P n be a fiber bundle and s its section. We identify P with the zero section (and hence any point p P with it s image in B). In that

case if s and the zero section are transverse, their intersection is a submanifold of P of dimension n m by implicit function theorem (it is empty if m > n). Transversality at a point p in Im(s) P is the statement that T p Im(s) T p P = T p B = T p P T p F p, where F p is the fiber of B over p. Of course T p Im(s) is the Im(Ds), where Ds is the derivative (or linearizarion ) of s, Ds : T p P T p B. Composing Ds with projection to T p F p (taking vertical part of the linearization ) gives the linear map (operator) Ds vert : T p P T p F p. The section is transverse to the zero section at p if and only if Ds vert is surjective. If it is, the kernel of it is the tangent space to the moduli space Im(s) P. Note that the dimension of that moduli space is the index of Ds vert - dimension of kernel minus dimension of cokernel. Of course if Ds vert is surjective cokernel is 0, but the index makes sense even if the map is not surjective and is constant - equal to n m - essentially by rank-nulity theorem. We want to repeat this argument in our - infinite dimensional - situation. To be able to do so we need to overcome several technical difficulties. We now mention some of them. 1) To be able to apply implicit function theorem in infinite-dimensional setting we need B to be a Banach bundle over Banach manifold P. However, the space of smooth paths in P is not Banach (ultimately, because C (R) is not Banach), so we need to work with appropriate completions. Our first hunch is to take L 1,2 (R, M) (we need once differentiable to make sense and L 1,2 is a Hilbert space, which is handy). However, because R is noncompact this space does not really make sense. Indeed, usually to define L k,2 (N, M) one either embeds M into some R N or uses charts on M which works fine for compact domain N (see, for example, [7, Lecture 11]). However, for a non-compact domain, in both of these approaches a constant map γ : R p may be assigned infinite norm and excluded from L 2 (R.M). This is a problem in general. of d(γ(s)) ds For us, the solution is to notice that there is no issue defining L 1,2 loc (R, M) or L 1,2 (R, R n ). Moreover by Sobolev embedding, paths in L 1,2 loc (R, M) are also in C(R, M). So it makes sense to talk about such a path converging to a point. Now for a path γ(s) converging to a fixed p for large s we land close enough to p that we can write γ(s) = exp v(s) for a unique v T p M (here exp v is the endpoint of the geodesic of the metric g starting from p and tangent to v). Identifying T p M = R n we require v(s) to be in L 1,2 (R, R n ) (this is independent of the T p M = R n identification). That is the space of paths P that we actually work with. It is locally modelled on L 1,2 (R, R n ) and so is Banach. 2) Just as in the finite dimensional case, to get the moduli spaces of trajectories to be a manifold we need the section F g to be transverse to the zero-section. To achieve this we need to perturb the metric g. To that end, we cross B P with G - the space of all C k smooth metrics on M and consider the universal section F given by the same formula as F g but now for varying g. One proves that this universal section is transverse to the zero section, and uses infinite dimensional version of Sard s theorem to conclude that F g is transverse for a generic (co meagre and hence second category) set of gs. However, since g is not smooth, F is also not smooth, but rather only C k ([7, Lecture 13]). Even the finite-dimensional Sard theorem only holds when the smoothness k is large enough (see [2, p.69] for an amazing (counter)example of what happens when this smoothness bound does not hold). The same issue appears in the infinite dimensional We are still doing finite-dimensional Morse theory. It s the abstract opeartor as a section setup that is infinite-dimensional. 2

version and imposes the requirement k > max(0, indexf q ). In the Morse theory case it is enough to take k > n. However, one would still want to have generic smooth metrics to work with. An additional lemma ( [7, Lecture 5], see also [5, p. 52] to get that. Alternatively, one can avoid going through C k metrics altogether by using a Floer space - a Banach space made out of smooth functions (see [1] also mentioned in [5, Remark 3.2.7, p.53]. 2.2 Linearizing the vertical part of F g. We saw above that the operator in question is the vertical part of DF g at a gradient trajectory γ. Thus we want to compute it in local coordinates with the goal of then proving it Fredholm and computing its index. This is Ritter HW 12 and lecture 13 and then lecture 14 [7]. The chart around γ is given by the geodesic flow with respect to the metric g, so for a vector field v along γ we compute DF vert g (v) = t t=0 ( s w f g(w)) = ( t s w t f g(w)) t=0 = s v ( v f g) v. Here w(s, t) = exp v(s) tv(s). Using parallel frame e i write v = v i e i and compute s v = s v i e i, v f g = v j ej f g(γ) = (A s ) i j vj e i. Thus in these coordinates L 1,2 (R, R n ) L 2 (R, R n ) DF vert g (v) = ( s + A s )v We note that lim s + = Hess p and lim s = Hess q, non-degenerate symmetric matrices. Thus our linearised vertical part is amenable to spectral flow techniques. 3 The spectral flow. We have now reduced our problem to studying the Fredholmness and index properties of F As = s + A s : L 1,2 (R, R n ) L 2 (R, R n ). We assume that lim s + A s = A + and lim s A s = A with A ± invertible. Recall that for a symmetric matrix Index A is the numbers of negative eigenvalues. Our goal is to prove the spectral flow theorem: Index F A = Index A + Index A. Our exposition is an adaptation of the approach of Kronheimer and Mrowka [4, Chapter 14]. 3.1 Constant case Invertibility. Consider the case when A s = A is independent of t. In fact, we will start even simpler, with F λ = s + λ : L 1,2 (R, R) L 2 (R, R). Firstly, if λ = 0 the operator is not Fredholm!. If you are used to compact domains this may be surprising, as it is elliptic. In particular s : L 1,2 (S 1, R) L 2 (S 1, R) is Fredholm. Things are different on the non-compact domain R. 3

As an illustration, consider f(t) = 1 t for t > 1 smoothly cut of to zero on [0, 1] and zero on (, 0]. Then f is in L 2 (R, R), but is not in the image of any L 1,2 function (as it s integral is ln(t) + C on t > 1 and is not square integrable. The same thing happens for all functions decaying as t α for α ( 3/2, 1/2) and so the cokernel of s is not finite dimensional. However if λ 0 the operator is in fact invertible. Proposition 3.1 ([4]14.1.2). F λ = s + λ : L 1,2 (R, R) L 2 (R, R) is invertible. Proof. The proof is based on Fourier transform. Recall that F T (u) = û is the function given by û(ξ) = + u(s)eitξ ds. Plancherel theorem says that F T is an isomorphism of L 2 (R, C) to itself. Its more general version says that F T of L 1,2 (R, C) is the completion of Cc (R, C) in the norm û(ξ) F T (L 1,2 ) = + (1 + ξ2 ) u(ξ) 2 dξ. Of course the reason we like F T is that it takes s + λ to D - the multiplication by iξ + λ. Thus in the transformed form it has inverse D 1 which is multiplication by (iξ + λ) 1. We compute: D 1 û F T (L 1,2 = + (1 + ξ 2 ) (iξ + λ) 1 u(ξ) 2 dξ = + 1 + ξ 2 λ 2 + ξ 2 u(ξ) 2 dξ max (1, 1λ ) 2 û(ξ) L 2. So the inverse of F λ (given by Fourier transforming, dividing and Fourier transforming back) lands in L 2 (R, R) so F λ is invertible as a map L 1,2 (R, C) L 2 (R, C. Finally note that the space of real functions is Fourier transformed to the space of those functions that have u(ξ) = u( ξ) and this condition is unaffected by multiplying or dividing by (iξ + λ), so both F λ and its inverse restrict to that subspace, proving our proposition. Now if we consider F A : L 1,2 (R, R n ) L 2 (R, R n ) with constant A then by changing basis on the target R n we get A = diag(λ 1,..., λ n ). So if A is not invertible then F A is not Fredholm, and if A is invertible then all F i = s + λ i : L 1,2 (R, R) L 2 (R, R) are invertible and so is F A. Theorem 3.2. If A s = A is constant and A invertible, then F A is invertible. 3.2 The general case Fredholmness. To begin, we consider the case where A s is eventually constant. That is, that there is S > 0 such that A s = A + for s > S and A s = A for s < S. If A s is not constant we can not expect F A to be invertible (after all its index should be the spectral flow, not 0). But if it is Fredholm it should be invertible up to compact operators. That is F is Fredholm if and only if there us a Q : L 2 (R, R n ) L 1,2 (R, R n ) such that F P I and P F I are compact. We shall now glue the inverses of F A+ and F A to get the almost-inverse P. To that effect, let G 1 be the inverse of F A and G 2 be the inverse on F A+. We take a partition of unity η 1 + η 2 = 1 subordinate to the cover U 1 = (, 1), U 2 = ( 1, + ) of R (recall this means that η i is zero outside U i ). Take also smooth µ 1 to 4

be 1 on U 1 and zero for s > 2 and µ 1 to be 1 on U 2 and zero for s < 2. Define P 1 : L 2 (R, R n ) L 1,2 (R, R n ) by P 1 v = µ 1 G 1 η 1 v and similarly for P 2. Finally let P = P 1 + P 2. We will have several occasions to use the product rule s (f(s)v) = f(s)( s )v + f(s)v, or written another way f(s)( s )v = s (f(s)v) + f(s)v We compute: F A P 1 v = ( s + A s )µ 1 G 1 η 1 v = µ 1 G 1 η 1 v + µ 1 s (G 1 η 1 v) + A s µ 1 G 1 η 1 v = µ 1 G 1 η 1 v + µ 1 ( s + A )(G 1 η 1 v) µ 1 A G 1 η 1 v + µ 1 A s G 1 η 1 v = µ 1 η 1 v + µ 1 G 1 η 1 v + µ 1 (A s A )G 1 η 1 v = η 1 v + µ 1 G 1 η 1 v + µ 1 (A s A )G 1 η 1 v Since µ 1 is supported on [ 2, 1], µ 1 is supported on s 1 and A s A is supported on s S. The difference F A P 1 η 1 is supported on [ S 2, 1] and since L 1,2 (R) L 2 ([ S 2, 1]) is compact the difference is compact as well. Similarly F A P 2 η 2 is compact, so that F A P η 1 + η 2 = F A P I is compact, as wanted. In the other direction we get P 1 F A w = µ 1 G 1 η 1 ( s + A s )w = µ 1 G 1 ( s (η 1 w) η 1 v + η 1 A s w) = µ 1 G 1 ( s + A )(η 1 w) + η 1 (A s A )w η 1 w = η 1 w + µ 1 G 1 (η 1 (A s A )w η 1 w) As η 1 and η 1 (A s A ) are both supported on [ S 1, 1] and the difference P 1 F A η 1 is compact. Similarly P 2 F A η 2 and hence P F A I are compact. This proves that F A is Fredholm. 3.3 The diagonal case Index. We now turn to computing the index of F A. To start, consider the case n = 1 and A s = λ(s). Then F A v = 0 is the ODE s v = λv which we can solve explicitly s ln(v) = λ(s), v(s) = Ae s 0 λ(x) dx. This is in L 2 (R, R) if and only if λ + > 0 and λ < 0. Since we know that F As is Fredholm, the cokernel of F As is the kernel of the adjoint of F As which (by integration by parts) is just F As, and that is one dimensional if λ + < 0 and λ > 0 and empty otherwise. Putting this two together proves that index of F As is given by the spectral flow in this case. Now in higher dimensions, but still keeping A s = diag(λ 1 (s),..., λ n (s)) we see that F A (v) = 0 splits into n coordinate ODE s s v i = λ i v i, and so the dimension of the kernel is the number of λ i switching from negative to positive. Just as before, since 5

F As is Fredholm its cokernel is the kernel of F As and so has dimension equal to the number of λ i switching from positive to negative. This proves the index=spectral flow for the case A s = diag(λ 1 (s),..., λ n (s)). 3.4 The general case Index. The general case now follows readily by homotopying it to the diagonal case. Namely we need two facts 1) Fredholm index is invariant under homotopy through Fredholm operators. 2) Any path A s of symmetric matrices with A and A + invertible is homotopic through other such paths to the path Ãs = A + λβ(s)a for some λ R and β(s) a function with β = 0 for s < 1 and β(s) = 1 for s > 1. In particular Ãs is diagonal in some coordinate system on the target R n and so index= spectral flow holds for it. The first one is a standard fact about Fredholm operators. For the second one we first observe that there is a λ R such that A + has the same index (as a symmetric matrix) as A +λ - we just need to bump the right number of eigenvalues of A across 0. Lemma 3.3. Any two invertible symmetric matrices A and B of the same index are homotopic through invertible symmetric matrices. Proof. An invertible symmetric matrix A has an orthonormal basis in which it is diagonal (which can be chosen to give the canonical orientation on R n ) and conversely a choice of such a basis together with n non-zero real numbers (to serve as eigenvalues in that basis) specifies A. Given A and B first rotate a basis of A to a basis of B keeping the λ s constant and in such a way that the negative eigenvalue eigenvectors of A are taken to negative eigenvalue eigenvectors of B (this is possible since indexes of A and B agree, and the group SO(n) is path connected). Then simply rescale λs of A to λs of B linearly. Since the corresponding λs are of the same sign, they never become zero during this process. Considering the corresponding symmetric matrix for each labeled basis gives a path of invertible symmetric matrices from A to B. Now we pre-homotope A s to be constant A for s < S and A + for s > S (for some S) and then just use the homotopy from A +λ to A + given by the lemma above to homotope A s to a path between A and A + λ. Finally,we homotope that path to à linearly. The index stays fixed throughout and hence is equal to spectral flow from A to A + λ which is the same as that from A to A +, which completes the proof. References [1] Floer [2] Hirsh [3] Kerman [4] Kronheimer-Mrowka [5] McDuff-Salamon 6

[6] Robbin-Salamon [7] Ritter [8] Salamon 7