LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM

Similar documents
SYMPLECTIC GEOMETRY: LECTURE 5

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

A LITTLE TASTE OF SYMPLECTIC GEOMETRY: THE SCHUR-HORN THEOREM CONTENTS

LECTURE 11: SYMPLECTIC TORIC MANIFOLDS. Contents 1. Symplectic toric manifolds 1 2. Delzant s theorem 4 3. Symplectic cut 8

A little taste of symplectic geometry

LECTURE 4: SYMPLECTIC GROUP ACTIONS

LECTURE 15: COMPLETENESS AND CONVEXITY

Delzant s Garden. A one-hour tour to symplectic toric geometry

pacific journal of mathematics Vol. 181, No. 2, 1997 MOMENT MAPS ON SYMPLECTIC CONES Suzana Falc~ao B. de Moraes and Carlos Tomei We analyze some conv

ON NEARLY SEMIFREE CIRCLE ACTIONS

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

LECTURE 1: LINEAR SYMPLECTIC GEOMETRY

A Convexity Theorem For Isoparametric Submanifolds

Symplectic Reduction and Convexity of Moment Maps. Theo van den Hurk

Hamiltonian Toric Manifolds

LECTURE 8: THE MOMENT MAP

LECTURE 5: SOME BASIC CONSTRUCTIONS IN SYMPLECTIC TOPOLOGY

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 15-16: PROPER ACTIONS AND ORBIT SPACES

IGA Lecture I: Introduction to G-valued moment maps

Nicholas Proudfoot Department of Mathematics, University of Oregon, Eugene, OR 97403

arxiv:math/ v1 [math.sg] 23 Nov 2005

Transversality. Abhishek Khetan. December 13, Basics 1. 2 The Transversality Theorem 1. 3 Transversality and Homotopy 2

LECTURE 11: TRANSVERSALITY

Geometry of the momentum map

Geometry and Dynamics of singular symplectic manifolds. Session 9: Some applications of the path method in b-symplectic geometry

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

A DANILOV-TYPE FORMULA FOR TORIC ORIGAMI MANIFOLDS VIA LOCALIZATION OF INDEX

arxiv: v1 [math.sg] 6 Nov 2015

Intersection of stable and unstable manifolds for invariant Morse functions

B 1 = {B(x, r) x = (x 1, x 2 ) H, 0 < r < x 2 }. (a) Show that B = B 1 B 2 is a basis for a topology on X.

Cohomology of the Mumford Quotient

Many of the exercises are taken from the books referred at the end of the document.

12 Geometric quantization

Polar orbitopes. Leonardo Biliotti, Alessandro Ghigi and Peter Heinzner. Università di Parma - Università di Milano Bicocca - Ruhr Universität Bochum

Bredon, Introduction to compact transformation groups, Academic Press

Differential Topology Final Exam With Solutions

LECTURE 5: COMPLEX AND KÄHLER MANIFOLDS

Reduced phase space and toric variety coordinatizations of Delzant spaces

LECTURE 16: LIE GROUPS AND THEIR LIE ALGEBRAS. 1. Lie groups

Exercises in Geometry II University of Bonn, Summer semester 2015 Professor: Prof. Christian Blohmann Assistant: Saskia Voss Sheet 1

A CONSTRUCTION OF TRANSVERSE SUBMANIFOLDS

CUT LOCI AND DISTANCE FUNCTIONS

Notes for Functional Analysis

M4P52 Manifolds, 2016 Problem Sheet 1

Topological properties

Morse Theory and Applications to Equivariant Topology

Lecture III: Neighbourhoods

Differential Geometry qualifying exam 562 January 2019 Show all your work for full credit

CHAPTER 1. TOPOLOGY OF ALGEBRAIC VARIETIES, HODGE DECOMPOSITION, AND APPLICATIONS. Contents

CHAPTER 7. Connectedness

HAMILTONIAN TORUS ACTIONS IN EQUIVARIANT COHOMOLOGY AND SYMPLECTIC TOPOLOGY

Symplectic convexity theorems

Res + X F F + is defined below in (1.3). According to [Je-Ki2, Definition 3.3 and Proposition 3.4], the value of Res + X

Monomial equivariant embeddings of quasitoric manifolds and the problem of existence of invariant almost complex structures.

SOME REMARKS ON THE TOPOLOGY OF HYPERBOLIC ACTIONS OF R n ON n-manifolds

1 Smooth manifolds and Lie groups

A. CANNAS DA SILVA, V. GUILLEMIN, AND A. R. PIRES

A MARSDEN WEINSTEIN REDUCTION THEOREM FOR PRESYMPLECTIC MANIFOLDS

Morse theory and stable pairs

Eilenberg-Steenrod properties. (Hatcher, 2.1, 2.3, 3.1; Conlon, 2.6, 8.1, )

LECTURE 22: THE CRITICAL POINT THEORY OF DISTANCE FUNCTIONS

THEODORE VORONOV DIFFERENTIABLE MANIFOLDS. Fall Last updated: November 26, (Under construction.)

Notation. For any Lie group G, we set G 0 to be the connected component of the identity.

10. The subgroup subalgebra correspondence. Homogeneous spaces.

An Invitation to Geometric Quantization

REGULAR TRIPLETS IN COMPACT SYMMETRIC SPACES

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

arxiv:math/ v2 [math.sg] 19 Feb 2003

arxiv: v1 [math.sg] 26 Jan 2015

THE EULER CHARACTERISTIC OF A LIE GROUP

Solutions to the Hamilton-Jacobi equation as Lagrangian submanifolds

Hyperkähler geometry lecture 3

MANIN PAIRS AND MOMENT MAPS

Lecture 1. Toric Varieties: Basics

A real convexity theorem for quasi-hamiltonian actions. Florent Schaffhauser

Math 147, Homework 5 Solutions Due: May 15, 2012

LECTURE 14: LIE GROUP ACTIONS

The Symmetric Space for SL n (R)

THE POINCARE-HOPF THEOREM

Handlebody Decomposition of a Manifold

k=0 /D : S + S /D = K 1 2 (3.5) consistently with the relation (1.75) and the Riemann-Roch-Hirzebruch-Atiyah-Singer index formula

THE JORDAN-BROUWER SEPARATION THEOREM

274 Microlocal Geometry, Lecture 2. David Nadler Notes by Qiaochu Yuan

CHARACTERISTIC CLASSES

A little taste of symplectic geometry

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

GEOMETRIC QUANTIZATION

Symplectic Origami. A. Cannas da Silva 1,2, V. Guillemin 3, and A. R. Pires 3

A generic property of families of Lagrangian systems

Log-Concavity and Symplectic Flows

Math 249B. Nilpotence of connected solvable groups

APPLICATIONS OF HOFER S GEOMETRY TO HAMILTONIAN DYNAMICS

Cohomology of Quotients in Symplectic and Algebraic Geometry. Frances Clare Kirwan

NOTES ON DIFFERENTIAL FORMS. PART 5: DE RHAM COHOMOLOGY

Robustly transitive diffeomorphisms

LECTURE 8: THE SECTIONAL AND RICCI CURVATURES

Problems in Linear Algebra and Representation Theory

BRUHAT-TITS BUILDING OF A p-adic REDUCTIVE GROUP

Symmetric Spaces Toolkit

Transcription:

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM Contents 1. The Atiyah-Guillemin-Sternberg Convexity Theorem 1 2. Proof of the Atiyah-Guillemin-Sternberg Convexity theorem 3 3. Morse theory 6 4. Student presentation 9 1. The Atiyah-Guillemin-Sternberg Convexity Theorem The statement of AGS convexity theorem. Let (M, ω) be a compact connected symplectic manifold. In this and next lecture, we will study the special case where G = T is a compact connected abelian Lie group, acting in Hamiltonian fashion on M. Since T is abelian, from standard Lie theory we know that T T k for some k. It follows that t R k, and thus t R k. We have seen that in this case the moment map µ : M t is a T -invariant map. The main theorem we want to prove is Theorem 1.1 (Atiyah-Guillemin-Sternberg Convexity Theorem). Let (M, g, T, µ) be a compact connected Hamiltonian T-manifold. Then the image of µ is a convex polyhedron in t whose vertices are the image of the T -fixed points. Remark. In the general case of a Hamiltonian action of a compact Lie group G, the image of µ might be much more complicated. However, one can prove that the intersection of µ(m) with each Weyl chamber is a convex polyhedron. Examples. Example. The moment map for the standard rotational S 1 -action on S 2 is µ(x, y, z) = z. So the image of µ is the interval [ 1, 1]. Observe that the pre-images of the points ±1 are north/south poles of S 2, which are exactly the fixed points of the S 1 -action. 1

2 LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM Example. Consider the standard T 2 on CP 2 via The moment map is (e it 1, e it 2 ) [z 0 : z 1 : z 2 ] = [z 0 : e it 1 z 1 : e it 2 z 2 ] µ([z 0 : z 1 : z 2 ]) = 1 2 ( z1 2 z, z ) 2 2 2 z 2 The image is a triangle in R 2 with vertices (0, 0), ( 1, 0), (0, 1 ). Observe that these 2 2 vertices are the image of [1 : 0 : 0], [0 : 1 : 0] and [0 : 0 : 1] of CP 2, which are exactly the fixed points of the T 2 -action. One can easily extend this example to arbitrary dimension, in which case the image of the moment map is a simplex. Example. Hirzebruch surfaces. Application: Schur-Horn theorem. Let A be an Hermitian matrix whose eigenvalues are λ 1 λ 2 λ n. Arrange the diagonal entries of A in increasing order as a 1 a 2 a n. It was shown by Schur that for each 1 k n, k k a i λ i. i=1 Horn prove the converse: For any sequences λ k and a k satisfying all the above inequalities, there exists an Hermitian matrix A whose diagonal entries are a k s and whose eigenvalues are λ k s. We can give a symplectic geometric proof of Schur-Horn s theorem. For any λ = (λ 1,, λ n ) R n, let H λ be the set of all n n Hermitian matrices whose eigenvalues are precisely λ 1,, λ n. As was explained in Yuguo s presentation, H λ can be identified with a U(n)-coadjoint orbit and thus is a symplectic manifold. The coadjoint action of U(n) on H λ is a Hamiltonian action. This action restricts to a Hamiltonian action of the maximal torus T n T U(n) on H λ, whose moment map is µ : H λ R n, A = (a ij ) n n (a 11, a 22,, a nn ) =: a. The fixed points of the T action are the diagonal matrices, whose diagonal entries has to be λ σ = (λ σ(1),, λ σ(n) ), where σ is a permutation of (1, 2,, n). So according to the Atiyah-Guillemin-Sternberg convexity theorem, a vector a is the diagonal of a Hermitian matrix in H λ if and only if a lies in the convex hull of the points λ σ. This is equivalent to the Schur-Horn inequalities. Remark. Kostant extends Schur-Horn s theorem to more general coadjoint orbits. i=1

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM 3 Application: number of T -fixed points. As an application of the AGS convexity theorem, we can prove Proposition 1.2. Let (M, g) be a compact connected symplectic manifold admitting a Hamiltonian T -action. Suppose there exists m M such that T acts locally free at m. Then there must be at least k + 1 fixed points, where k = dim T. Proof. Since the action is locally free at m, the stabilizer T m is a finite subgroup of T, and thus t m = {0}. It follows that Im(dµ m ) = t 0 m = t. So dµ m is surjective, i.e. µ is a submersion near m. It follows that µ : M t is an open map near m. In particular, µ(p) is an interior point of µ(m). So µ(m) is a non-degenerate convex polytope in t R k, which has at least k + 1 vertices. Since each vertex is the image of a T -fixed point, the T action has at least k + 1 fixed points. 2. Proof of the Atiyah-Guillemin-Sternberg Convexity theorem We will follow the Guillemin-Sternberg approach to prove the convexity theorem. The Atiyah approach is sketched in Ana Canas de Silver s book. The equivariant Darboux theorems. Suppose G is a compact connected Lie group acting smoothly on M. Suppose m 0 is a fixed point of the G-action. We can endow with M a G-invariant Riemannian metric. Then sufficiently small geodesic balls around m 0 is a contractible invariant domain, and using the ordinary Poincare lemma it is easy to prove the following invariant Poincar e lemma: Any invariant closed k-form in a neighborhood of m 0 is the differential of an invariant k 1-form. As a consequence, we can prove Theorem 2.1 (Equivariant Darboux theorem). Let (M, ω i ), i = 1, 2 be symplectic G-spaces. Let m be a fixed point of G so that ω 1 (m) = ω 2 (m). Then there is an invariant neighborhood U of m and an equivariant diffeomorphism f of U into M so that f(m) = m and f ω 2 = ω 1. Proof. Apply Moser s trick as before. Details left as an exercise. Now suppose (M, ω) be a symplectic G-manifold, and let m be a G-fixed point. Then the isotropy action of G on T m M, g v = dg m (v), is a linear G-action, i.e. a representation, of G on T m M. Moreover, if we fix an G-invariant Riemannian metric on M, then the exponential map exp : T m M M is equivariant. It follows from the previous theorem that

4 LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM Theorem 2.2 (Equivariant Darboux theorem, version 2). Let (M, ω) be a symplectic G-manifold and m a fixed point of the G-action. Then there is an invariant neighborhood U of m in M and an equivariant diffeomorphism ϕ from (U, ω) into (T m M, ω m ) so that ϕ(m) = 0 and ϕ ω m = ω. One can identify the symplectic vector space (T m M, ω m ) with the complex space C n. So Theorem 2.3 (Equivariant Darboux theorem, version 3). Let (M, ω) be a symplectic G-manifold and m a fixed point of the G-action. Then there is an invariant neighborhood U of m in M and local complex coordinates z 1,, z n so that on U, the symplectic form can be written as ω = 1 dz k d z k, 2i and the G-action becomes a linear symplectic G-action on C n. k Darboux theorem for the moment map. Now let s turn back to the case G = T is a torus of dimension n. Let α 1,, α n t be the weights of the isotropy representation of T on T m M. In other words, e it (z 1,, z n ) = (e iα 1(t) z 1,, e iαn(t) z n ). Theorem 2.4. Let p U be a point whose coordinate is z. Then µ(p) = µ(m) + k z k 2 2 α k. Proof. Exercise. Remark. Locally a symplectic action is always Hamiltonian since the first cohomology vanishes. The local convexity. Now let s go back to the theorem. Let U be the invariant neighborhood given by the equivariant Darboux theorem above. Then the image of U under the moment map µ, near the point µ(m), is { } n µ(m) + s k α k 0 s k ε. In other words, we proved k=1 Proposition 2.5. Let (M, ω, T, µ) be a Hamiltonian T -space and m a T -fixed point. Then there exists a neighborhood U of m so that µ(u) is a cone with vertex µ(m).

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM 5 This local convexity theorem has a relative version: Let T 1 T be a subgroup, and m 1 M is a fixed point under the induced T 1 action. As we have seen in lecture 8, the moment map µ 1 of this T 1 -action is µ 1 = dι T µ. Note that the map dι T : t t 1 is nothing else but the projection π : t t 1 if we identify t 1 as a subspace of t. Applying the previous arguments, we can find a neighborhood U 1 of m 1 so that µ 1 (U 1 ) is the cone with vertex µ 1 (m 1 ) described as above. It follows that its preimage is Proposition 2.6. Let (M, ω, T, µ) be a Hamiltonian T -space, T 1 a subgroup of T, and m 1 a T 1 -fixed point. Let α1, 1, αn 1 1 be the weights of the isotropy representation of T 1 on T m1 M. Then there exists a neighborhood U 1 of m 1 so that µ(u 1 ), near µ(m 1 ), is { ( n1 ) } µ(m) + π 1 s k α k 0 s k ε. The global convexity. In the next section we shall prove k=1 Lemma 2.7 (Guillemin-Sternberg lemma). For any X g, the function µ X : M R has a unique local minimum/maximum. Proof of the Atiyah-Guillemin-Sternberg convexity theorem. Since M is compact, the image µ(m) is compact in g. Let ξ g be a point in the boundary of µ(m). Take a point m M so that µ(m) = ξ. Let T 1 = T m be the stabilizer of m and let α 1,, α n t 1 be the weights of the isotropy representation of T 1 on T m M. Then there exists a neighborhood U of m in M so that { ( n1 ) } µ(u) = ξ + π 1 s k α k 0 s k ε. k=1 We denote S = π 1 ({ s k α k s k 0}). Let S 1 be a boundary component of S. Since S 1 has codimension at least 1, one can choose X g so that and η, X = 0 for η S k η, X < 0 for η in the interior of S. Now suppose ξ, X = a. Then for any m U, µ X (m ) = µ(m ), X = ξ + η, X a, i.e. a is a local maximum of µ X. According to the Guillemin-Sternberg lemma, it is an absolute maximum. So µ(m), X = µ X (M) a. Applying this argument to all boundary components S k of S, we conclude that µ(m) sits in the cone ξ + S. It follows that µ(m) is a convex polyhedron.

6 LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM Finally a point µ(m) is a vertex if and only if n 1 = n for the point m, i.e. the stabilizer of m is T itself. So m is a fixed point of the T -action. 3. Morse theory It remains to prove the Guillemin-Sternberg lemma. Morse-Bott functions. Let M be a compact manifold and f C (M) a smooth function. Recall that a point m M is called a critical point of f if df m = 0. The set of all critical points of f is denoted by Crit(f). Now let m Crit(f) be a critical point of f. Consider the Hessian map Hess m (f) : T m M T m M R defined by (1) Hess m (f)(x m, Y m ) = X(Y f)(m), where X and Y are any vector fields whose value at m are X m and Y m respectively. Lemma 3.1. For m Crit(f), Hess m (f) is well-defined, symmetric and bilinear. Proof. Symmetry follows from the fact X(Y f)(m) Y (Xf)(m) = [X, Y ]f(m) = df m, [X, Y ] m = 0, Since X(Y f)(m) = X m (Y f), the right hand side of (1) is independent on the choice of X and is bilinear on X m. By symmetry, X(Y f)(m) = Y (Xf)(m), the right hand side of (1) is also independent of the choice of Y and is bilinear on Y m. Recall that a critical point m Crit(f) is called non-degenerate if the bilinear form Hess m (f) is non-degenerate. A function f C (M) is called a Morse function if Crit(f) is discrete, and each m Crit(f) is non-degenerate. Obviously any Morse function on a compact manifold has only finitely many critical points. Definition 3.2. A function f C (M) is called a Morse-Bott function if Crit(f) is a manifold (with different components, C i, which may have different dimensions), and for each m M the Hessian Hess m (f) is non-degenerate in the direction transverse to Crit(f) (equivalently, the kernel of Hess m (f) is T m Crit(f)). The index. Now suppose f is a Morse-Bott function. Let n ± (m) be the number of positive/negative eigenvalues of Hess m (f) at m Crit(f). They are of course constant along each connected component C i of Crit(f). We shall denote them by n ± j and call them the index/coindex of C j. Now we equip M with a Riemannian metric. Note that the gradient vector field f vanishes exactly on Crit(f). Using the Riemannian metric g one can identify

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM 7 Hessf with 2 f. Let φ t be the negative gradient flow of f. Then for any m C j, one has the decomposition T m M = T m C j E + m E m, where E ± are spanned by the eigenspace of positive/negative eigenvalues of 2 f respectively. Observe that as t, every point moves to point in Crit(f) under the flow φ t. For each j the consider the set of points that flow to C j as t, W s (C j ) = {m M φ t (m) C j as t + }. Fact: W s (C j ) is a submanifold of M of dimension n + j + dim(c j). Moreover, at each m C j, T m W s (C j ) = T m C j E + m. We will call W s (C j ) the stable submanifold of C j. Similarly by studying the negative gradient flow as t, one gets the unstable submanifold of C j, W u (C j ) = {m M φ t (m) C j as t }, which is a submanifold of dimension n j + dim(c j), whose tangent space at m C j is T m W u (C j ) = T m C j Em. We shall use the following topological observation: If N is a submanifold of M of codimension at least 2, then M \ N is connected. We have seen Note that n j n + j = codim(w u (C j )), n j = codim(w s (C j )). means each point in C j is a local minimum. Proposition 3.3. Let f be a Morse-Bott function. If none of these n j s equals to 1, then there exists a unique j such that n j = 0. In other words, there is a unique C j on which f is local minimum, and thus an absolute minimum. Proof. By the assumption, if n j 0, then n j 2. The union of all these stable submanifolds has codimension at least 2. It follows that their complement, M c, is a nonempty connected subset in M. Since M is the union of all stable submanifolds, there exists at least one critical component C j with index n j = 0. Note that W s (C j ) is an open subset of M. Moreover, since M c is connected, it cannot be a union of more than one disjoint open subset. It follows that such C j is unique. So f attains its absolute minimum on C j. Remark. By the same way one can show that if none of the n + j s equals to 1, then there exists a unique C k on which f is local maximum, and thus an absolute maximum.

8 LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM Proof of the Guillemin-Sternberg lemma. Obviously the Guillemin-Sternberg lemma follows from Theorem 3.4. Let (M, ω, T, µ) be a compact connected Hamiltonian T -space. Then for any X g, µ X is a Morse-Bott function, and all of its indeces n ± j s are even. Proof. Let X g and m Crit(µ X ). Let T 1 be the closure of the 1-parameter subgroup {exp(tx) t R}. Then m is a fixed point for the action of T 1. Let α 1,, α n be the weights for the isotropy representation of T 1 on T m M. Then as we have seen, the moment map of this T 1 action is locally of the form µ 1 (p) = µ 1 (m) + z k 2 2 α k. On the other hand, µ 1 = π µ, where π is induced by the inclusion T 1 T. So µ X (p) = µ X 1 (p) = µ X 1 (m) + z k 2 2 α k(x) = µ X 1 (m) + α k (X) (x 2 k + y 2 2 k). It follows that the critical points of µ X near m are defined locally by the equations z j = z j+1 = = z d = 0, where j is chosen so that α j (X) = = α n (X) = 0 and α 1 (X) 0,, α j 1 (X) 0, and d is half the dimension of M. In other words, the critical set Crit(µ X ) is a smooth submanifolds near each point m Crit(µ X ). So µ X is a Morse-Bott function. Moreover, the Hessian of µ X at m has eigenvalues In particular, all indeces are even. α 1 (X), α 1 (X),, α n (X), α n (X), 0,, 0. Connectedness of level sets. One can say more from the Morse-Bott theory. Lemma 3.5. Let f be a Morse-Bott function such that none of the n ± j s equals 1. Then for any regular value r of f, f 1 (r) is a connected submanifold of M. Proof. According to the previous proposition, we see that there exists a unique C j on which f attains its absolute minimum, and there exists a unique C k one which f attains its local maximum. Let M 0 = (M \ Crit(f)) C j C k be the complement of the union of all other critical levels. It is an open dense connected subset of M. Consider the negative gradient flow φ t on M 0. Since r is a regular value, f 1 (r) is a smooth submanifold of M, and min f < r < max f. So each orbit intersects with f 1 (r) at a unique point. So the map f 1 (r) R M 0, (m, t) φ t (m) is a diffeomorphism. Since M 0 is connected, f 1 (r) must be connected.

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM 9 Proposition 3.6. Let f be a Morse-Bott function such that none of the n ± j s equals 1. Then for any r, f 1 (r) is connected. Proof. We may assume r is neither the minimum nor the maximum, because the connectedness is already proved for that two cases. Let M 1 be set of points which tends to the minimum of f as t, and tends to the maximum as t. Then f 1 (c) M 1 is a connected submanifold. It remains to prove that f 1 (c) M 1 is dense in f 1 (c). Let m f 1 (c), and U be an arbitray small neighborhood of m. Then U M 1 is connected because we only delete a submanifold of codimension at least 2. Since c is neigher a local minimum nor a local maximum, one can find points m 1, m 2 in U M 1 such that f(m 1 ) < c and f(m 2 ) > c. So one can find a point m 3 in U M 1 with f(m 3 ) = c. This completes the proof. Corollary 3.7. Under the assumption of AGS, µ X (r) is connected for any X and any r. 4. Student presentation