<<

MORSE THEORY AND IT’S APPLICATION TO THEORY

RAOUL BOTT

Abstract. The following is a LaTeX version of Bott’s 1960 book on Morse theory which is no longer in print. Some proofs are supplemented and some ambiguous notations are fixed.

Contents 1. Introduction 1 2. Morse Theory of Smooth Functions on a 4 3. The Morse Inequalities 9 4. Embedded in an 11 5. Topology of Flag Manifolds 16 6. The Structure of the Space Ωp,q(M) 19 7. The Index Theorem 21 8. Critical Manifolds 29 9. The Stable Homotopy Groups of the Unitary Group 31 Acknowledgement 35 10. References 36

1. Introduction To get a first idea about Morse Theory, we consider a simple example. Let T be a 2-dimensional , resting on its tangent plane V as indicated in fig.1. The distance of the points of T from the tangent place V is a real analytic, hence smooth (C∞) f on T . We set T a = {x ∈ T | f(x) ≤ a} T a is empty for a < 0, {p} for a = 0, homeomorphic with a 2-cell for 0 < a < f(q), homeomorphic with he product of a circle and a line segment for f(q) < a < f(p), is homeomorphic with the figure indicated in fig.2 for f(q) < a < f(s), and the whole torus for a ≥ f(s). As f grows from 0 to f(s), T a grows in successive steps from a to the whole torus. From the topological point of view, something new comes at the level of p, q, r, and s. These points are the critical points of f, the points where df = 0.

Here, we touch on a first essential idea of Morse Theory. There is a close relation between the behavior of f as a smooth function, especially with respect to its critical points, and the topological structure of T . Let us make this more precise. For that 1 2

Figure 1.

Figure 2. purpose, we introduce the concept of attaching an r-cell to a X. More generally, let Y be a second topological space, Z a subspace of Y and f : Z → X a continuous map. Let X˜ be the topological space obtained as follows: in the disjoint union of X and Y , we identify the points s of Z with their f(s) and provide the resulting space with the quotient topology. We say that X˜ is obtained from X by attaching Y to X according to the pair (Z, f). In particular, if y is an r-cell er and Z its boundarye ˙r, we simply say that er is attached to X ˜ S according to the map f and write X = X f er. Furthermore, we introduce the index of a non-degenerate critical point of a smooth function f defined on a smooth manifold M n. As we said above, a critical point of f is a point x ∈ M n, such that at x, df = 0, or expressed in local coordinates (x1, ..., xn): ∂f = 0, i = 1, ..., n. ∂xi A critical point x of f is called non-degenerate if  ∂2f  Hf = ∂xi∂xj is a nonsingular n × n matrix. It is easily proved that the number of positive and negative eigenvalues of Hf is independent of the local coordinates. The last num- ber is naturally called the index of x as a critical point of f. In the case of our example, the index of p, q, r, and s is 0, 1, 1, 2 respectively. From the homotopic point of view, the building up of t in successive steps can be described as follows: If there are no critical points between a and b then T a is of the same homotopic type as T b. If there is a single critical point of index k in this range, then T b is obtained by attaching a k-cell. MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 3

Figure 3. The building up of T by successive attaching of cells corresponding to the critical points of f

The main result of the simplest part of Morse theory states that the situation as described for the special function f on the special smooth manifold T is a general one. This is expressed by the following theorem. Theorem 1.1. Let f be a smooth function on the compact smooth manifold M, such that f has only non-degenerate critical points. Set M a = {x ∈ M, f(x) ≤ a}. Then, if f −1(a ≤ s ≤ b) does not contain any critical point of f, then M a and M b are homeomorphic. If f −1(a ≤ s ≤ b) contains only one critical point x of index b a S a λ, a < f(x) < b, then M ∼ M eλ, eλ being attached to M by a conveniently chosen map. As always, “smooth” means “C∞” and “manifold” means “connected manifold”. Let the situation be as in the theorem, and let ai ∈ R, i = 1, ..., m be the critical values of f, which we suppose to correspond 1-to-1 to the critical points of f and

M a1 ⊂ M a2 ⊂ ...M am .

n It can be shown, precisely in the same way as it can be shown that πk(S ) = 0 for ai ai−1 0 ≤ k < n, that πk(M ,M ) = 0 for 0 ≤ k ≤ λi − 2, where λi denotes the index of the critical point corresponding to ai, also that the homomorphism

ai−1 ai πλi−1(M ) → πλi−1(M ) is onto. Thus from the exact homotopy sequence we can conclude:

ai−1 ai πλi−1(M ) = πλi−1(M ) 0 ≤ k ≤ λi − 1. Further, it is easy to see that

r ai ai−1 H (M ,M ; Z) = 0 0 ≤ r ≤ λi and

r ai ai−1 H (M ,M ; Z) = Z r = 0, λi.

The second part of Morse theory is analogous to the first one, but deals with a different, more complicated situation. Let M be a compact . Let P , Q be a couple of points on M. We denote by µP,Q(M) the set of sectionally smooth , joining P and Q, and parametrized proportionally to the length of arc by a parameter t, 0 ≤ t ≤ 1. 4 RAOUL BOTT

Let L(u) be the length of the u ∈ µP,Q(M). If we set for every pair of curves u, v ∈ µP,Q(M) L(u, v) = max dist(u(t), v(t)) + |L(u) − L(v)| t∈[0,1] then it is easily seen that L(u, v) is a distance function on µP,Q(M). Now the result of Morse Theory in the previous theorem has an analogue in this case. The role of f by L and the in µP,Q(M) take the place of the critical points. More precisely, the following theorem holds: Theorem 1.2. Let M be a compact Riemannian manifold, P and Q a pair of points on M, µP,Q(M) the metric space of sectionally smooth curves from P to Q, parametrized proportionally to length of arc. Set c ΩP,Q(M) = {u ∈ µP,Q(M), L)(u) ≤ c}, where L(u) denotes the length of u. If there is no of length l, a ≤ l ≤ b, then b a ΩP,Q(M) ∼ ΩP,Q(M). If there is just one such geodesic g of length l, a < l < b, then b a [ ΩP,Q(M) ∼ ΩP,Q(M) eλ, where λ denoting the number of conjugate points of P along g which lie between P and Q.

Since πk(ΩP,Q(M)) = πk+1(M) ([1], p.55), this theorem yields information on the homotopy structure of M, provided that the “geodesic structure” of M is suf- ficiently known. This is the case for special kinds of manifolds, in particular sym- metric spaces.

2. Morse Theory of Smooth Functions on a Manifold We start with the first part of Theorem 1.1.

Theorem 2.1. Let M be a compact smooth manifold and f : M → R a smooth function on M. Set M a = {x ∈ M, f(x) ≤ a}. if df(p) 6= 0 for all points p ∈ M with a ≤ f(p) ≤ b, then M a and M b are homeomorphic. We recall that a local 1-parameter group of diffeomorphisms of M is a (smooth) map Φ: M × R → M, such that

(1)Φ |M×{t} is a diffeomorphism of M; (2)Φ( m, t1 + t2) = Φ(Φ(m, t1), t2) for |t1|, |t2|, and |t1| + |t2| < . If p ∈ M is not a fixed point of Φ, then there passes just one orbit curve of Φ through p, Φ determines in a natural way a vector field Φ˙ on M by setting g(Φ(p, t)) − g(p) D ˙ g = (Φ(˙ g))(p) := lim , Φ(p) t→0 t p ∈ M, g a smooth function on M. Φ˙ is zero at the fixed points of φ and tangent to the orbits of φ at the other points of M. MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 5

Lemma 2.2. Let M be a compact smooth manifold and V a vector field on M. Then there exists a (uniquely determined) local 1-parameter group Φ of diffeomorphisms of M such that Φ˙ = V For the proof of Lemma 2.2, see [2, p.5]. We now give the proof of Theorem 2.1.

Proof. We introduce a Riemannian metric on M. This can be done for example in the following way. Let {U1, ..., Um} be a finite open covering of M with coordinate systems, and {ϕ1, ..., ϕm} a partition of unity corresponding to {U1, ..., Um}. Let i i x1, ..., xn be coordinates in Ui and x ∈ Ui. for every pair (X,Y ) of tangent vectors to M at x, we define ϕi(X,Y ) be setting  ∂ ∂  i , i = δkl ∂xk ∂xl and extending by linearity. Then m X (X,Y ) = ϕi(X,Y ) i=1 defines a global Riemannian metric on M. This metric enables us to define a special vector field ∇f on M, the of f, by

(∇f(p),Y (p)) = df|p(Y (p))

Obviously, ∇f(p) = 0 is equivalent to df|p = 0. Now let φ be the local 1-parameter group of diffeomorphisms with the property that φ˙ = −∇f. Since by assumption, df|p 6= 0 for all x ∈ M with a ≤ f(x) ≤ b, by Lemma 2.2, there passes through every point of this subset of M a unique integral curve of −∇f. This is also true if we 0 1 replace a by a conveniently chosen a < a. Since f(m) − f(φ(m, 2 )) is continuous, this function has a positive minimum on the compact set {x ∈ M, a0 ≤ f(x) ≤ 0 b}. Therefore, every integral curve of −∇f which meets M b meets M a and also conversely. 0 A homeomorphism h : M b → M a is constructed as follows. For y ∈ M a , we set 0 h(y) = y. Now let y ∈ M b − M a . Suppose the integral curve of −∇f through y a a0 b intersects M , M , and M in ya, ya0 , and yb respectively. We determine h(y) on this curve C by setting

l(yaya0 ) l(h(y)ya0 ) = l(yya0 ) l(ybya0 ) where l denotes the length of arc along C.  For the proof of the second part of Theorem 1.1, we need some lemmas.

Lemma 2.3 (Morse). Let f : Rn → R be a smooth function on Rn, such that f(0) = 0, fxi (0) = 0, i = 1, ..., n, and det(fxixj (0)) 6= 0. Then in a neighborhood U of 0, coordinates (y1, ..., yn) can be introduced such that in U 2 2 2 2 2 f = −y1 − y2 − ... − yλ + yλ+1 + ... + yn, where λ is the index of f at 0. 6 RAOUL BOTT

Figure 4.

Proof. We start with the construction of smooth functions aij(x), i, j = 1, ..., n, such that n X f = aij(x)xixj, i,j=1 where aij(x) = aji(x). Using f(0) = df(0) = 0, we find in an elementary way: n Z 1 ∂ X Z 1 f(x) = (f(xt))dt = x ∂ (f(xt))dt ∂t i i 0 i=1 0 n n Z 1 X 1 X ∂ = x [∂ f(xt)(t − 1)] − x (∂ f(xt))(t − 1)dt i i 0 i ∂t i i=1 i=1 0 n n X X  Z 1  = − ∂j(∂if(xt)(t − 1))dt xixj j=1 i=1 0 n X = aij(x)xixj. i,j=1

We may assume: a11(0) 6= 0(at least one element aij 6= 0, in the case this element is aii with i 6= 1, we simply permute, in the case this element is Aij with i 6= j, we introduce new coordinatesx ¯1, ..., x¯n by settingx ¯i = xi + xj,x ¯j = xi − xj,x ¯k = xk fork 6= i, j). Firstly, let a11(0) > 0. In a neighborhood of 0, we can set

 P  √ a1αxα y˜1 = a11 x1 + α = 2, ..., n a11 y˜k = xk k = 2, ..., n √ where a11 denotes the positive root of a11. It follows:

n X 2 X aijxixj =y ˜1 + bαβy˜αv˜β. α,β=2  P   P  √ a1αxα √ a1αxα dy˜1 = (d a11) x1 + + a11 dx1 + d a11 a11 MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 7

√ X dy˜1(0) = a11dx1 + cαdxα

dy˜α(0) = dxα, α = 2, ..., n √ dy˜1 ∧ ... ∧ dy˜n = a11dx1 ∧ ... ∧ dxn 6= 0

 ∂y˜  Hence we have det ∂x (0) 6= 0. 2  ∂y˜  1 Since det(bαβ(0)) = det(aij(0)) det ∂x (0) and det(aij(0)) = 2n detfxixj (0) 6= 0, we also have det(bαβ) 6= 0 In the case that a11 < 0, we set X 2 X α β aijxixj = −y1 + y y . α,β

By induction, we get functions y1, ..., yn such that n X 2 f = iyi , i = ±1, i=1 and dy1 ∧ ... ∧ dyn(0) 6= 0. Since (y1, ..., yn) can be used as coordinates in a neigh- borhood of 0, the number of negative i’s equals the index of f at 0. This completes the proof. 

Lemma 2.4. Let X be a topological space, er an r-cell, Is an s-cell, given by s Is = {(t1, ..., ts) ∈ R , 0 ≤ ti ≤ 1} and let er ×Is be attached to X by a map f :e ˙r ×Is → X. If we set g = f|e˙ ×(0,...,0), S S S r then X f (er Is) ∼ X g er. S S S Proof. As illustrated in fig. 5, X f (er Is) can be deformed into X g er by a standard deformation. 

Figure 5.

Theorem 2.5. Let M be a compact smooth manifold, f : M → R a smooth function on M. Set M a = {x ∈ M, f(x) ≤ a}. If f −1(a ≤ x ≤ b) conatins exactly one non- b a S degenerate critical point p of index λ, a < f(p) < b, then M ∼ M eλ, where eλ is attached to M a by a conveniently chosen map. Proof. We may assume that f(p) = 0. From Theorem 2.1, it follows that it is  a S sufficient to prove the existence of a number , 0 <  ≤ b, such that M ∼ M eλ. By Lemma 2.3, there is a neighborhood U of p and local coordinates such that in U, f is given by 2 2 2 2 2 f = −y1 − y2 − ... − yλ + yλ+1 + ... + yn, 8 RAOUL BOTT

Furthermore, we may assume M to be provided with a Riemannian metric ds2 which is Euclidean in U. If we set:   \ 2 2 A = {y ∈ M U, y1 + ... + yλ ≤ ρ}    M∗ = M − A   S   then M = A M∗. For conveniently chosen  > 0, ρ > 0, this just mean that M  is obtained from M∗ by attaching a product eλ × In−λ to M∗ in the way described in Lemma 2.4. By this Lemma, we find   [ M ∼ M∗ eλ.  a It can be shown in the same way as in the proof of Theorem 2.1 that M∗ and M are homeomorphic. It only has to be shown that the gradient of f is not zero and that any integral curve starting from the boundary of M a meets the boundary of  2 M∗. Using the fact that ds is Euclidean in U, this is easily proved. 

Figure 6. The case of n = 2, λ = 1

Figure 7. The case of the torus, considered in the Introduction

A smooth function f on a smooth manifold M is called non-degenerate if all critical points of f are non-degenerate; and such a function is called strictly non- degenerate if f(p) 6= f(q) for every pair of critical points p and q. In Theorem 2.1 and Theorem 2.5, we only considered strictly non-degenerate functions. It is natural to question whether there always exist non-degenerate and strictly non-degenerate functions on a given manifold. The answer is given by the following. Theorem 2.6 (Morse-Thom). let M be a compact smooth manifold, Z the space of smooth functions on M, provided with the compact open topology. Then the subset of non-degenerate functions is everywhere dense in Z. MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 9

For the proof of this theorem, see [3], p.155. An immediate consequence of Theorem 2.6 is the following theorem. Theorem 2.7. Let the situation be as in Theorem 2.6. Then the subset of strictly non-degenerate functions is also dense in Z. Proof. It is sufficient to prove the following local result: n Let f be a function on R ,(x1, ..., xn) with only one non-degenerate critical point at (0, ..., 0). Then there exists a function f 0 on Rn such that: (1) |f − f 0| <  everywhere on Rn and f = f 0 outside a neighborhood of (0, ..., 0); (2) f 0 also has only one (non-degenerate) critical point, and this is again the point (0, ..., 0); (3) f 0(0) = f(0) + . Here  denotes an arbitrarily but sufficiently small positive number. Set n n X 2 A = {(x1, ..., xn) ∈ R , xi ≤ 1} i=1 n n X 2 B = {(x1, ..., xn) ∈ R , 1 ≤ xi ≤ 2} i=1 n n X 2 C = {(x1, ..., xn) ∈ R , xi ≥ 2} i=1 Let g be a function on Rn with g(x) = 0 if x ∈ C, 0 ≤ g(x) ≤ 1 if x ∈ B, and g(x) = 1 if x ∈ A. Taking for f 0(x) the function f(x) + g(x),  > 0, we can see that f 0 fulfills the conditions of (1), (2), (3) provided that  is sufficiently small. The only thing to be verified is that f 0 has no critical points on B. But on B, we have uniform bounds α and β, α > 0, such that kdfk ≥ α and kdgk ≤ β. From kdf 0k ≥ kdfk − kdgk ≥ α − β it follows that f 0 has no critical points on B provided that  is sufficiently small. 0 Hence f has only one (non-degenerate) critical point and that is the point 0. 

3. The Morse Inequalities Let M be a compact manifold, which can be built up by successively attaching cells, int he way described in section 2. Then it can be shown (see [8]) that there is a CW-complex K such that its cells are in dimension preserving 1-1 correspondence with the attaching cells, and the of K is the homology of M (with respect to any group of coefficients). Accepting this result, the well known Morse inequal- ities follow immediately from the results of section 2 if we take R as a domain of coefficients. In fact, let n X C = Ci i=0 the (naturally graded) vector space of chains of K, n X Z = Zi i=1 10 RAOUL BOTT the space of cycles, n X B = Bi i=1 the space of boundaries and n X H = Hi i=1 the real homology group of K. By definition, we have the exact sequences: (1)0 → Z → C →δ B → 0 (2)0 → B → Z → H → 0 where δ reduces the degree by 1. Set dimCi = ci, dimZi = zi, dimBi = bi, dimHi = hi (the i-th of K), i = 0, 1, 2, .... Combining (1) and (2), we find:

zi + bi−1 = ci, bi + hi = zi, ci − hi = bi + bi−1 i = 0, 1, ... (b−1 = 0)

Since bi ≥ 0, i = 0, 1, ..., these relations lead to the following sequence of inequali- ties:

c0 ≥ h0

c1 − c0 ≥ h1 − h0

c2 + c1 − c0 ≥ h2 − h1 + h0 ......

Let f be a non-degenerate smooth function on the compact smooth manifold M. By Theorem 2.7, there is a strictly non-degenerate function f 0 on M such that f and f 0 have the same number of critical points of index i, i = 0, 1, 2, .... Combination of Theorem 2.5, the result stated at the beginning of this section, and the above set of inequalities leads to

Theorem 3.1. Let f be a non-degenerate smooth function on the compact smooth manifold M. Let ci be the number of critical points of index i of f and hi, i = 0, 1, 2, ... the i-th Betti number of M. Then there exists a sequence of non-negative integers b−1, b0, b1, ... such that

ci − hi = bi + bi−1 i = 0, 1, 2, ... Therefore, there is a sequence of inequalities:

c0 ≥ h0

c1 − c0 ≥ h1 − h0

c2 + c1 − c0 ≥ h2 − h1 + h0 ......

These inequalities are known as the Morse inequalities. They imply that a (non- degenerate) smooth function on a compact smooth manifold necessarily has a cer- tain number of critical points of index i, i = 0, 1, ... . MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 11

4. Manifolds Embedded in an Euclidean Space In this section, the results of section 2 are applied to the special case that M n is n+k embedded smoothly in a Euclidean space R ,(x1, ..., xn+k) and f is the distance function from the points of M to a fixed point p, p ∈ Rn+k. We denote this function n by lp = lp(x), x ∈ M . n Let q ∈ M be a critical point of lp. This is equivalent to saying that pq is perpendicular to the TMq of M at q. Therefore, q is also a critical n point of the function lr, where r denotes any point of the line pq, r∈ / M . We choose a convenient coordinate system as follows. For q, we choose q = 0 = (0, ..., 0) and TM0 = {xn+1 = xn+2 = ... = xn+k = 0}. In a neighborhood of 0, n+k n x1, ..., xn can be used as a system of local coordinates on M, and in R , M can be given locally by xn+l = gl(x1, ..., xn), l = 1, ..., k → n Since p0 is perpendicular to M , the coordinate of p can be taken as (0, ..., 0, p1, ..., pk). Finally, we set (0, ..., 0, tp1, ..., tpk) = tp for 0 < t. Let, as usual, Hltp(0) denote the Hessian of ltp at 0. A straightforward calcula- tion gives: k 2 ! 1 X pl ∂ gl Hltp(0) = I − (0) , tkpk kpk ∂xi∂xj l=1 Pk 2 1/2 where I denotes the n × n identity matrix and kpk = ( l+1 pl ) . Using a well known result on quadratic forms (see [4], p. 158), we conclude that it is possible to choose a base in M0, such that

k 2 1 X pl ∂ gl I and − (0) kpk kpk ∂xi∂xj l=1 are simultaneously reduced to diagonal form, in such a way that 1 I kpk is reduced to the identity itself. With respect to this base, Hltp(0) is given by a matrix  1  a11 + t 0  ..  Hltp(0) =  .  . 1 0 ann + t We see (1) the coefficients in the diagonal are strictly decreasing functions of t; (2) only for a finite number of values of t, t1, ..., tm, Hltp(0) is non-singular, and t is a degenerate critical point of ltp; (3) for 0 < t  1, Hltp(0) is positive definite.

It follows that the index of Hltp(0) is a decreasing (integer valued) function of 1 t , which only jumps at the points t1p, ..., tmp, and at these points this jump just equals ν(Hltip(0)) = the dimension if the nullity of Hltip(0), i + 1, ...m. Since by (3), the index of Hltp(0) is zero in a neighborhood of 0, we can state Theorem 4.1. Let the smooth manifold M be embedded smoothly in the Euclidean space R. For the point p ∈ R, p∈ / M, let lp(x) denote the distance from the points 12 RAOUL BOTT x ∈ M to p. Let q be a critical point of lp(x) (degenerate or not). Then the index of the Hessian Hlp(q) is given by X index Hlp(q) = ν(Hl(1−t)p+tq). 0

2 Setting Lp(x) = lp(x), we immediately deduce from 2 2 ∂ Lr ∂ lr (0) = 2lr (0), i, j = 1, ..., n r∈ / M, ∂xi∂xj ∂xi∂xj

Corollary 4.2. The statement of Theorem 4.1 remains true if we replace lp(x) by Lp(x). Example: Let M be the unit sphere S2 embedded in R3. Let p be a point in 3 2 R , p∈ / S . The critical points of Lp are obviously the intersection points q1 and 2 q2 of p0 with S . In figure 8, HLp(q1) has index 0 (Lx is degenerate for no point between p and q1) and HLp(q2) has index 2, for there is only one point x between p and q2 such that HLx(q2) is degenerate, which is 0 and ν(HL0(q2)) = 2.

Figure 8.

n Let (x1, ..., xn) be a system of coordinates in R , and let the straight line g = g(t) be given by

xi = pi + tqi, i = 1, ..., n. A variation of g will be a smooth family of lines

V (s, t): xi(s) = pi(s) + tqi(s), −∞ < s < ∞, with pi(0 = pi) and qi(0) = qi, i + 1, ..., n.

Figure 9. MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 13

A variation of g induces in a natural way a vector field J(t) along g:

dpi(s) dqi(s) J(t) = + ds s=0 ds s=0 Such a field, induced by a variation of g will be called a Jacobi field along g. d2J Jacobi fields are also characterized by the property dt2 = 0 all along g. It follows, that the Jacobi fields along g form a vector space of dimension 2n, which will be denoted by Jg. Let M m be a proper smooth submanifold of Rn, and let g be perpendicular to M at g(0). A variation of g relative to M is a variation V (s, t) of g with the following properties: (1) V (s, 0) ∈ M (2) all lines V (s, t) are perpendicular to M at V (s, 0).

Figure 10.

Let Jg(M) be the set of Jacobi fields along g which are induced by variations of g relative to M. In the sequel to this section, it will become obvious that Jg(M) is a linear subspace of Jg. So far, at every point g(t0) of g, there is a natural restriction homomorphism

rt0 : Jg(M) → Tg(t0)R.

Definition 4.3. g(t0), t0 6= 0, is called a focal segment of M of multiplicity ν,

ν > 0, if and only if dim Ker rt0 = ν. The fundamental result of this section is the following

Theorem 4.4. Let the notations be as introduced before. Then g(t0) is a focal segment of M of multiplicity ν if and only if g(0) is a degenerate critical point of

Lg(t0)(x) of nullity ν.

Proof. Let ν be the nullity of Lg(t0)(x). We may assume g(t0) to be the point (0, ..., 0). If (u1, ..., um) is a system of local coordinates on M in a neighborhood of g(0), and as usual n X 2 L = xi (uα), i=1 then the number ν equals m minus the rank of the quadratic form m 2 X ∂ L α β HL|g(0) = U V , ∂uα∂uβ α,β=1 g(0) 14 RAOUL BOTT where U = (U α) and V = (V β) are tangent vectors to M at g(0). Setting

∂xi i ∂ i i = Λα and (Λα) = Λα,β, i = 1, ..., n; α, β = 1, ..., m, ∂uα ∂uβ we find: n n ∂L X ∂xi X = 2 x (u ) = 2 x Λi ∂u i α ∂u i α α i=1 α i=1 n n ∂2L X X = 2 Λi Λi + 2 x Λi ∂u ∂u α β i α,β α β i=1 i=1 If we denote by (U, V ) the inner product, induced by the of M in Rn, and set m n X X xi n(U, V ) = Λi U αV β, l α,β α,β=1 i=1 where l denotes the distance from g(0) to 0, we find

HL|g(0)(U, V ) = 2{(U, V ) + l · n(U, V )}. Let V (s, t) = p(s) + tq(s) be a variation of g relative to M. From the definition of Jg(M), it follows immediately (1) p(s) ∈ M; n X i (2) qi(s)Λα(p(s)) = 0, α = 1, ..., m for all values of s. i=1 Differentiation of (2) with respect to s gives n m n X dqi X X ∂uβ (3) Λi (p(s)) + q (s)Λi (p(s)) (p(s)) = 0, α = 1, ..., m ds α i α,β ∂s i=1 β=1 i=1 where uβ = uβ(s) is the curve on M, described by xi = pi(s). Setting in particular s = 0, we get n m n X dqi X X pi(0) ∂uβ (4) (0)Λi (g(0)) + Λi (g(0)) (g(0)) = 0, α = 1, ..., m. ds α l α,β ∂s i=1 β=1 i=1 The Jacobi field along g induced by V (s, t) can be written as

ηi(t) = ηi(0) + tη˙i(0), i = 1, ..., n with

dpi = ηi(0) ds s=0

dqi =η ˙i(0), i = 1, ..., n. ds s=0

∂uβ α Obviously, η(0) ∈ Tg(0)M and ηβ = ∂s , β = 1, ..., m. Now let U = (U ) be any tangent vector to M m in g(0). Multiplying (4) with U α, we get by summation: (η(0)˙ ,U) − n(η(0),U) = 0. Conversely, let η(t) be a Jacobi field along g, such that

(a) η(0) ∈ Tg(0)M; MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 15

(b)(˙η(0),U) − n(η(0),U) = 0 for all U ∈ Tg(0)M.

Let pi(s) be a curve on M with pi(0) = pi, and

dpi = ηi(0) ds s=0

Consider (3) as a system of differential equations for q1, ..., qn with s as independent variable. There exists a solution with

dqi qi(0) = ηi(0) and =η ˙i(0), i = 1, ..., n, ds s=0

dqi since by (b), (4) is satisfied for these values of qi(0) and ds . Let qi(s) be such s=0 a solution. For this solution, we have: n X i qi(s)Λα(p(s)) = constant. i=1 Since n X i qi(0)Λα(p(0)) = 0, i=1 this constant is zero. It follows that there exists an element of Jg(M) inducing η(t) along g. Therefore, the properties (a) and (b) are characteristic for the elements of Jg(M).

The elements η ∈ Jg(M), for which rt0 (η)(0) = 0 are characterized as those elements of Jg, which satisfy the following conditions:

(a) η(0) ∈ Tg(0)M; (b)( η(0),U) + l · n(η(0),U) = 0 for all U ∈ Tg(0)M; (c) η(0) + l · η˙(0) = 0. From this, it is immediate that

dim Ker rt0 = ν(HLg(0)(p)) = ν. This proves Theorem 4.4.  Remark 4.5. If we define a linear transformation

T∗ : Tg(0)M → Tg(0)M by setting (T∗U, V ) = n(U, V ), the conditions (a) and (b) for the elements of Jg(M) can be replaced by the following ones:

(a’) η(0) ∈ Tg(0)M; ⊥ (b’)˙ η(0)−T∗(η(0)) ∈ Tg(0)M (the orthogonal complement of Tg(0)M in Tg(0)R). This leads to m + (n − m) = n linearly independent conditions for an element of Jg to be in Jg(M). Therefore:

dim Jg(M) = n.

From this it follows easily that dim Ker rt0 = ν is equivalent to

dim(TRg(t0)/rt0 (Jg(M))) = ν. Combination of Corollary 4.2 and Theorem 4.4 gives 16 RAOUL BOTT

Theorem 4.6. Let M be a proper smooth submanifold of Rn. Let a ∈ Rn, a∈ / M, and let b ∈ M be a critical point of the function La(x). Let ν(t) be the multiplicity of the focal segment, and zero otherwise. Then the index of b equals X ν(t). 0

5. Topology of Flag Manifolds This section is devoted to a sketch of some typical topological applications of elementary Morse theory, as this theory was developed in the preceding sections.

As usual, we denote by U(n) the unitary group in n complex variables. U(n1) × U(n2) × ... × U(nk), n1 + ... + nk = n, can be considered as a subgroup of U(n) in a canonical way. The quotient space is the complex flag manifold of type (n1, ..., nk), to be denoted as W (n1, ..., nk). For k = 2, we get a Grassmann variety, in particular for n1 = 1, n2 = n − 1, a complex projective space. As a first application, we shall sketch the proof of the following

i Theorem 5.1. H (W (n1, ..., nk), Z) = 0 for i odd, and W (n1, ..., nk) has no tor- sion. The tangent space to U(n) as the identity I of U(n) we denote as R. As a real vector space, it can be identified with the vector space of skew-symmetric complex n × n matrices, i.e., those matrices (aij) with aij = −a¯ji. If we set (X,Y ) = −Tr(X · Y ) for every pair of elements X, Y in R, we immediately see that we get a positive definite Riemannian metric m on R, which makes R into an Euclidean space. U(n) operates on R by the adjoint action AdU(n), defined by −1 AdU X = UXU ,U ∈ U(n),X ∈ R.

From a standard property of the trace, we readily deduce that AdU(n) leaves m invariant. Therefore, AdU(n) gives a representation of U(n) as a group of orthogonal transformations of R. We intend to study the orbits of U(n) in R. Let X ∈ R. It is easy to see that the subgroup of U(n) leaving X invariant is of the type U(n1) × ... × U(nk), n1 + ... + nk = n. In fact, let V be an n-dimensional complex vector space, on which U(n) operates as the group of unitary matrices and R as the group of skew symmetric matrices. In the same way as in the real case for symmetric matrices, there is an orthogonal decomposition of V , such that X leaves the factors of this decomposition invariant. The elements of U(n) which leaves X invariant by the adjoint actions are those unitary matrices which leave these factors invariant. Therefore, for a suitable base, the subgroup of U(n) leaving X invariant is a group U(n1) × ... × U(nk), n1 +...+nk = n canonically embedded in U(n). The orbit of X is the flag manifold W (n1, ..., nk). It is clear that all types of flag manifolds appear in this way. Let X ∈ R. The orbit of AdU(n) through X will be denoted by MX . Lemma 5.2. Let M be any orbit of U(n) in R, A ∈ R, a point of an orbit of maximal dimension, and let AB be orthogonal to M at B. A segment BC of the MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 17 line AB is a focal segment of multiplicity ν of LB(M) if and only if dimMA = dimMC = ν This lemma is a very special case of a more general theorem, which states, for example, the same results for the case that U(n) is replaced by an arbitrary connected Lie group. The proof of Lemma 5.2 is very analogous to the considerations in section 9 and will therefore be omitted.

From Lemma 5.2, Theorem 4.4, and the fact that the dimension of a complex flag manifold is always even, we conclude that the index of every critical point of a function LP (M) is always even. Applying the results of section 3 to a general point P , we obtain the statement of Theorem 5.1 (The existence of such a point P can be proved easily in this case; it follows also from the following considerations). As a second example, we sketch how elementary Morse theory can be used to obtain the Betti numbers of a complex flag manifold. It follows from the definition of m that the Jacobi bracket [X,Y ] = XY = YX, X,Y ∈ R has the property (1) ([X,Y ],Z) = (X, [Y,Z])

Lemma 5.3. The tangent space TX to MX at X is given by

(a) TX = {Z = [Y,X],Y ∈ R} and the normal space to MX at X by

(b) NX = {Y ∈ R, [Y,X] = 0}. Proof. (a) For every element Y ∈ R, etY is a 1-parameter family subgroup of U(n), to which Y is the tangent vector in I. We get all vectors Z of TX by taking

Ad tY X − X (2) Z = lim e t→0 t for all elements Y ∈ R. But instead of (2), we can write (1 + tY + ...)X(1 − tY + ...) − X Z = lim t→0 t t(YX − XY ) + t2(...) = lim t→0 t = YX − XY = [Y,X]. (b) [X,Y ] = 0 is equivalent to (A, [X,Y ]) = 0 for all A ∈ R. This on its turn is by (1) equivalent to ([A, X],Y ) = 0 for all A ∈ R and, by (a), just means that Y ∈ NX .  Lemma 5.4. If at one of its points, a line is perpendicular to an orbit, then that line is perpendicular to all orbits which it intersects.

Proof. Let the line B + tA be perpendicular to MB at B. By Lemma 5.3, this means that ([X,B],A) = 0 for all X ∈ R. Since ([X,A],A) = (X, [A, A]) = 0 for all X ∈ R, we find: ([X,B + tA],A) = ([X,B],A) + t([X,A],A) = 0 for all X ∈ R, and this is precisely the formal expression of our assertion.  18 RAOUL BOTT

With respect to a suitable basis in V , every element of R can be written as   iθ1 a12 ··· a1n  .  −a¯12 iθ2 .     . .. .   . . .  −a¯1n ······ iθn then X R = h ⊕ e12 ⊕ ... ⊕ e(n+1)n = h ⊕ ekl where 0 ······ 0     iθ1 0  . . .    . .. a . h =  ..  , e =  kl  , k 6= 1  .  kl . . .   . −a¯ .. .  0 iθn   kl   0 ······ 0 

Let h∗ denote the subset of h for which the real number θ1, ...θn are all different. In geometric language, h∗ consists of those points in the vector space h, which do not lie on any of the hyperplanes θi − θj = 0, i, j = 1, ..., n, i < j. Obviously, h∗ consists of “almost all” points of h.

Lemma 5.5. If X ∈ h∗, then NX = h. P P Proof. It is readily verified that in this case, [X, ekl] = ekl. From R = h ⊕ P P P ekl,we deduce [X,R] = [X, h] ⊕ [X, ekl] = ekl, since [X, h] = 0 by definition of the bracket. By lemma 5.3, [X,R] = TX , NX = h. 

Lemma 5.6. (a) Let M be any orbit of AdU(n) in R and P ∈ h∗. Then the critical points of the function LP (M) are precisely the points of M ∩ h. (b) In particular, these points are independent of the choice of P in h∗.

Proof. (a) Let L be a critical point of LP (M). The line P L is perpendicular to M at L. From Lemma 5.4, it follows that P L is perpendicular to MP at P . Since P ∈ h∗, and , by Lemma 5.5, we also have the direction of P L lies in h, so we see that P L lies in h, and in particular that L ∈ h. (b) For every point X ∈ h, we have [X, h] = 0. If L ∈ M ∩ h, then every line through L in h is perpendicular to M at L (Lemma 5.3), in particular P L. Since M is always compact, LP (M) certainly has critical points. Therefore, every orbit intersects h. 

Let P ∈ h∗ and L a (non-degenerate) critical point of LP (M). By Theorem 4.4 and the index λ(L) of L is given by X λ(L) = ν(Fi), i where the points Fi are the focal points on the segment P L and ν(Fi) are their multiplicities as focal points. From Lemma 5.2, we see that the focal points Fi are the points where the segment P L intersects orbits of lower dimension, and furthermore that, if F is such a point, then

ν(F ) = dimMP − dimMF . MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 19

But since P L lies in h, these points Fi are precisely the points of intersection with the hyperplanes θi − θj = 0, i < j. It is obvious, that dimMP − dimMF equals two times the number of θi − θj = 0, i < j, which vanish at F . The problem to determine the Betti numbers of a flag manifold is now reduced to a problem of Euclidean geometry. In fact, let M be any orbit, that is some type of flag manifold. The intersection points L1, ..., Ls of M with h can be determined explicitly by an algebraic procedure. Taking a general point P in h∗, the intersec- tions of the segments P L1, ..., P Ls with the hyperplanes θi − θj = 0, i < j, can be determined also, and therefore the index of every critical point of LP (M). Since the homology of M vanishes in odd dimensions, the i-th Betti number of M is just the number of critical points Lj of index i. Needless to say that similar methods apply to many other situations.

6. The Structure of the Space Ωp,q(M) Let M be a compact Riemannian manifold and p, q a couple of points on M. We denote Ωp,q(M) the set of sectionally smooth curves joining p with q, i.e. the set of piecewise smooth mappings f : [0, 1] → M, with f(0) = p, f(1) = q, parametrized proportionally to the length of arc (constant speed); by L(c), c ∈ Ωp,q(M) the a length of c; by Ωp,q(M) the subset of Ωp,q(M) determined by a Ωp,q(M) = {c ∈ Ωp,q(M)|L(c) ≤ a}; by ρ(x, y) the distance of pair of points x, y on M, i.e. the infimum of L(d), d ∈ Ωx,y(M). A simple argument (see [1], pg. 45) shows that ρ(c, c0) = max ρ(c(t), c0(t)) + |L(c) − L(c0)| 0≤t≤1 is a distance function on Ωp,q(M). With respect to the topology induced by this metric, L is a continuous function on Ωp,q(M). We consider Ωp,q(M) provided with this topology.

Figure 11.

ρ(x, y) is a continuous, but in general not a differentiable function on M × M. However, there exists a number ρ > 0 such that for ρ(x, y) < ρ, there is just one geodesic arc from x to y, of length ρ(x, y). For this ρ, ρ2(x, y), restricted to the part of M × M determined by ρ(x, y) < ρ, is a differentiable function of (x, y). ([1], pg. 49). In the sequel, Ωp,q(M) will be considered for a fixed pair of points p, q on M. a a So we can write Ω instead of Ωp,q(M) and Ω instead of Ωp,q(M) . Let n be such a2 2 that b = n+1 < ρ . If we set M × ... × M(n times) = M∗ and 2 2 ρ (p, x1) + ... + ρ (xn, q) = φ(x1, ..., xn) = φ(x), 20 RAOUL BOTT

b b then φ is a differentiable function on M∗ , where M∗ is determined by b M∗ = {x ∈ M∗, φ(x) ≤ b}. a b The following theorem relates Ω very strongly with M∗ . a b Theorem 6.1. Ω and M∗ are of the same homotopy type, i.e. there exist maps a b b a α :Ω → M∗ and β : M∗ → Ω , such that β ◦ α and α ◦ β are homotopic with the a b identity map of Ω and M∗ respectively. Proof. The proof is carried out in four steps.

(a) Definition of α. For c ∈ Ωa, we define i α(c) = {c(t ), ..., c(t )}, t = , i = 1, ..., n. 1 n i n + 1 b L(c) We have to verify: α(c) ⊂ M∗ . The length of the arc c(ti)c(ti+1) of c equals n+1 , hence L(c) ρ(c(t ), c(t )) ≤ . i i+1 n + 1 From the definition of φ, it follows that L2(c) a2 φ(α(c)) ≤ (n + 1) ≤ = b. (n + 1)2 n + 1 So for every a b a b c ∈ Ω , α(c) ∈ M∗ or α(Ω ) ⊂ M∗ .

(b) Definition of β. If we set p = x0 and q = xn+1, from the definition of φ, we derive immediately b for x = (x1, ..., xn) ∈ M∗ : n √ X 2 a ρ (xi, xi+1) ≤ b, or, for i = 0, ..., n : ρ(xi + xi+1) ≤ b = √ < ρ. n + 1 i=0

Hence for i = 0, ..., n, xi and xi+1 can be joined by a unique geodesic arc. The union of these arcs determines an element of Ω, which, by definition, is β(x). We only have to verify that β(x) ∈ Ωa, i.e. that L(β(x)) ≤ a. This is an immediate consequence of the Schwarz inequality: n X √ p √ a L(β(x)) = ρ(xi + xi+1) ≤ n + 1 φ(x) ≤ n + 1√ = a. n + 1 i=0

a a (c) Construction of a deformation Dτ :Ω → Ω , 0 ≤ τ ≤ 1, with D0 = identity and D1 = β ◦ α. a i τ Let c ∈ Ω and let for i = 0, ..., n + 1, xi = c(ti), ti = n+1 . Set xi = c((1 − τ)ti + τti+1) and define Dτ (c) as the (conveniently parametrized) element of a τ τ Ω , consisting of the (unique determined) geodesic arc px0 , the arc x0 x1 of c, the τ τ geodesic arc x1x1 , ..., the arc xnq of c. Obviously, D0 is the identity and D1 = β ◦α. It has to be verified that Dτ is actually a deformation. This is straightforward but not completely trivial. We refer to [1], pg. 51. MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 21

Figure 12.

b b (d) Construction of a deformation ∆τ : M∗ → M∗ , with ∆0 = identity and ∆1 = α ◦ β. b For x ∈ M∗ let si ∈ [0, 1] be such that β(x)(si) = xi, i = 0, ..., n + 1. Set τ τ τ τ τ β(x)(ti) = yi, β(x)((1 − τ)si + τti) = xi , x = (x1 , ..., xn) and ∆τ (x) = x . We τ b have to verify x ∈ M∗ . τ τ τ τ Since the distance from xi to xi+1 at most equal the distance from xi to xi+1 along β(x), we find τ τ ρ(xi , xi+1) ≤ L(β(x))[(1 − τ)si+1 + τti+1 − (1 − τ)si − τti]

= L(β(x))[(1 − τ)(si+1 − si) + τ(ti+! − ti)] τ τ If we set the distance from xi to xi+1 along β(x) as δi, this inequality becomes L(β(x)) ρ(xτ , xτ ) ≤ (1 − τ)δ + τ . i i+1 i n + 1 From this follows, by the definition of φ: n τ X 2 τ τ φ(x ) = ρ (xi , xi+1) i=0 n ! n ! X L2(β(x)) L(β(x)) X ≤ (1 − τ)2 δ2 + τ 2 + 2τ(1 − τ) δ i n + 1 n + 1 i i=0 i=0 L2(β(x)) L2(β(x)) = (1 − τ)2φ(x) + τ 2 + 2τ(1 − τ) n + 1 n + 1  L2(β(x))  L2(β(x)) = φ(x) + φ(x) − (1 − τ)2 − φ(x) − . n + 1 n + 1

L2(β(x)) Since, by Schwarz inequality, φ(x) − n+1 ≥ 0, we find for 0 ≤ τ ≤ 1: φ(xτ ) ≤ φ(x) ≤ b. τ b This means that x ∈ M∗ . ∆0 is the identity and ∆1 = α ◦ β by definition. For the straightforward proof that ∆τ is actually a deformation, we again refer to [1]. 

7. The Index Theorem b We use the notations of Section 6. On M∗ , we shall study the function, which some may refer as the energy function: n X 2 φ(x) = ρ (xi, xi+1) i=0 22 RAOUL BOTT

b Theorem 7.1. x = (x1, ..., xn) ∈ M∗ is a critical point of φ if and only if

(i) px1x2..xnq is a geodesic from p to q; (ii) ρ(p, x1) = ρ(x1, x2) = ... = ρ(xnq).

Figure 13.

Proof. We assume the First Variation Formula, which can be stated in the following way (see [3]). Let a, b ∈ M, a 6= b, with ρ(a, b) < ρ (see section 6), let g be the unique geodesic segment from a to b. X0 and X1 are the unit tangent vector to g at a and b, respectively. U0 and U1 are tangent vectors to M at a and b, respectively. The function ρ(x, y) is differentiable in a neighborhood of (a, b) on M × M. So we can speak of the directional derivative of ρ(x, y) in the direction of the tangent vector (U0,U1) at (a, b) as

hX1,U1i − hX0,U0i. Furthermore, if we fix a and consider the function ρ(a, x), x 6= a, on M, the directional derivative in direction U1 is simply hX1,U1i. We denote by si the length of the geodesic segment gi from xi to xi+1. On gi, − + denote the unit tangent vectors to gi at xi, xi+1 by si , si , respectively, i = 1, ..., n. Then the directional derivative of φ in direction (U1, ..., Un) is: + − + − 2s0hs0 ,U1i − 2s1hs1 ,U1i + 2s1hs1 ,U2i − ... − 2snhsn ,Uni. For this expression, we can write:

+ − + − + − 2[h(s0s0 − s1s1 ),U1i + h(s1s1 − s2s2 ),U2i + ... + h(sn−1sn−1 − snsn ),Uni] x is a critical point of φ, if and only if this expression vanishes for all tangent vectors (U1, ...Un). This means precisely that x is a critical point of φ, if and only if px1x2...xn is a geodesic from p to q and ρ(p, x1) = ρ(x1, x2) = ... = ρ(xn, q).  Pn Remark 7.2. If we consider the continuous function L = i=0 ρ(xi, xi+1), then L b is differentiable in a neighborhood of all points x = (x1, ..., xn) ∈ M∗ with xi 6= xj for all i 6= j. In the same way as in the proof of Theorem 7.1, we see that such a point x is a critical point of L, if and only if px1x2...xnq is a geodesic from p to q. The index theorem gives an expression for the index of a critical point of φ. To derive this expression, we need several preliminaries. Lemma 7.3. Let A be a quadratic form on the real vectors space Y. then the index of A can be characterized as the maximal dimension of a subspace of V on which the restriction of A is positive. MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 23

Proof. Let d de this dimension, λ the index of A, and 1, ..., n a coordinate system on V , such that A is given by λ n X 2 X 2 i + j . i=1 j=λ+1

Since A is negative definite on the subspace of V spanned by x1, ..., xλ, d ≥ λ. Conversely, suppose A to be negative definite on a subspace W of V of dimension d0 > λ. W has a subspace of dimension at least 1 in common with the subspace spanned by xλ+1, ..., xn on which A is semi-positive definite. This is impossible, thus λ ≥ d. Combination gives λ = d.  Lemma 7.4. Let A(t), 0 ≤ t ≤ 1, be a continuous family of quadratic forms on the n-dimensional real vector space V n, with the following properties:

(i) A(t1) ≤ A(t2) for t1 ≤ t2; (ii) A(t) is degenerate for a finite number of t-values, 0 < t < 1: α1, ..., αr and possibly for t = 0, but not for t = 1. Pr Then: λ(A(0)) − λ(A(1)) = k=1 ν(A(αk)).

Proof. Let us look at the point αi. From condition (i) and Lemma 7.2, we deduce

(2) λ(A(t)) ≤ λ(A(αi)) for t ≥ αi. n On the other hand, if A(αi) is negative definite on the subspace W of V , then A(t) is negative definite on W for |t − αi| < . From Lemma 7.2, it follows

(3) λ(A(t)) ≥ λ(A(αi)) for |t − αi| < . Combining (2) and (3) we find:

(4) λ(A(t)) = λ(A(αi)) for αi ≤ t ≤ αi + . Denoting by µ(A(t)) the number of positive terms in a reduction of A(t) to diagonal form, we prove in the same way for αi −  < t < αi: 0 (5) µ(A(t )) = µ(A(αi)). Combination of (4) and (5) gives: 0 0 λ(A(t ) − λ(A(t)) = n − µ(A(t )) − λ(A(t)) = n − µ(A(αi))λ(A(αi)) = ν(A(αi)) the nullity of A(αi). Application of this last result to the points α1, ..., αr clearly gives the statement of the lemma. 

Let g(t) be a constant-speed geodesic on M. A family of geodesics gα(t), −∞ < α < ∞, all parametrized proportionally to arc-length from gα(0), is called a variation of g(t) if

(i) gα(t) depends differentiably on α; (ii) g0(t) = g(t). A variation of g(t) induces a vector field U(t) along g by

∂V (α, t) U(t) = . ∂α α=0 A vector field along g, induced by a variation of g is called a Jacobi field along g. Clearly, this definition generalizes the concept of Jacobi field along a straight line in a Euclidean space, as defined in Section 4. A vector field along the segment s of g is called a Jacobi field along s, if and only if it is the restriction to s of a Jacobi 24 RAOUL BOTT

field along g. If a and b are the endpoints of s, the vector space of Jacobi fields Vs along g which vanishes at both a and b is denoted by ab. We admit the following properties of Jacobi fields which are either trivial or can be proved by standard methods. (i) Let U 0(t) be the covariant derivative of U(t) along g. Then there is one and 0 only one Jacobi field along g with given initial values U(t0) and U (t0); Proof sketch. Jacobi fields are determined by the Jacobi equation, which is a second-order differential equation, so (i) is trivial.  (ii) If the segment g(t), 0 ≤ t ≤ 1, of g lies completely in a neighborhood in which every two points can be joined by a unique geodesic arc, then there is one and only one Jacobi field along g(t) with prescribed boundary values at g(0) and g(1); Proof sketch. g is the unique geodesic from g(0) to g(1), so g(0) and g(1) are in the injective radius of each other. In particular, they are not conju- gate. Then for U(g(0)) = 0, the linear transformation Tg(0)M → Tg(1)M, U˙ (g(0)) 7→ U(g(1)) is non-singular. So combining with (i), there exists unique Jacobi field with U(g(0)) = 0, U(g(1)) = b. Similarly, there exists unique Jacobi field with U(g(0)) = a, U(g(1)) = 0. Summing up these two Jacobi fields yields the desired Jacobi field. 

(iii) If the Jacobi field U(t) is perpendicular to g(t) for t = t1 and for t = t2, t1 6= t2, then U(t) is perpendicular to g(t) for all values of t. Proof sketch. Byg ¨ = 0 and U¨ + R(U, g˙)g ˙ = 0, we have d hU,˙ g˙i = hU,¨ g˙i = hR(U, g˙)g, ˙ g˙i = 0. dt ˙ d Thus hU, g˙i = dt hU, g˙i is constant, then hU, g˙i is linear. Since hU, g˙i is 0 for t = t1 and for t = t2, t1 6= t2, it must be constant. 

Figure 14.

Let U ⊂ M be an open set in which every two points can be joined by a unique geodesic arc. Let a, b ∈ U, a 6= b, and let g(t), 0 ≤ t ≤ 1, be the constant-speed geodesic from a to b. The function l = ρ(x, y) is a differentiable function on U × U in a neighborhood of (a, b). Let Ca and Cb be differentiable cells of codimension 1, passing through a and b, respectively, and perpendicular to g(t) at a and b, ∗ respectively. C = Ca × Cb is a submanifold of U × U. Let l be the restriction of l MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 25

∗ ∗ to C. It follows from (4), that (a, b) is a critical point of l , and the Hessian Hla,b is well-defined. ∗ We admit the following expression for Hla,b (this is the so-called Second Vari- ation Formula, see [3]): ∗ 0 0 (6) XY Hla,b = (Xb,Yb ) − (Xa,Ya) + nb(Xb,Yb) − na(Xa,Ya), where the symbols have the following meaning: Xa and Ya are tangent vectors to Ca at a, Xb and Yb are tangent vectors to Cb at b.(Xa,Xb) and (Ya,Yb) can be considered as tangent vectors of C at (a, b). By property (ii) of Jacobi fields, there 0 is a unique Jacobi field Y (t) along g(t) with Y (0) = Ya and Y (1) = Yb. Y (t) denotes the covariant derivative of Y along g(t). Finally, na(Xa,Ya) and nb(Xb,Yb) denote the second fundamental forms at a and b with respect to g(t), evaluated at Xa, Ya and Xb, Yb respectively. For a definition of this for which plays no further role in these notes, we refer to [5], pg. 257, and the reference given there. Furthermore, if we fix a and consider the function l = ρ(a, x), x 6= a, and also ∗ ∗ its restriction l to Cb, then b is a critical point of l , and ∗ 0 (7) XbYbHl = (Xb,Yb ) + nb(Xb,Yb), 0 where Yb is the covariant derivative of the Jacobi field which is 0 at a and Yb at b.

Figure 15. Caption

Again, we use the notations of section 6, but replace p, q by c, d. Suppose cx1x2...xnd is a geodesic from c to d such that xi 6= xj for i 6= j. From the remark b after Theorem 7.1, it follows that x = (x1, ..., xn) ∈ M∗ is a critical point of the function n X L = ρ(xi, xi+1). i=0 Let Ci, i = 1, ..., n be differentiable cells of codimension 1, passing through xi and orthogonal to g(t) at xi. We set C1 ×...×Cn = N and consider N as a submanifold b ∗ of M∗ . Clearly, x ∈ N. Let L be the restriction of L to N. Obviously, x is a critical point of L∗. From (6) and (7), we deduce the following expression for the ∗ Hessian HLx: n ∗ X + − (8) UV HLx = Ui(Vi − Vi ). i=1 where U = (U1, ..., Un), V = (V1, ..., Vn), Ui and Vi being tangent vectors to Ci in − + xi, i = 1, ..., n. Vi and Vi denote the values at xi of the covariant derivative along 26 RAOUL BOTT xi−1xi and xixi+1, respectively, of the Jacobi field determined by Vi−1, Vi and Vi, Vi+1, respectively, with V0 = Vn+1 = 0. If V is in the null space of HL∗, then it follows from (8) and property (i) that the Jacobi fields determined by V0 and V1 along rx1, by V1 and V2 along x1x2,... form a global Jacobi field along g(t), 0 ≤ t ≤ 1, which vanishes at r and s. Conversely, by (8) and property (iii), such a global field determines a vector V in the null space of HL∗. Proposition 7.5. Let g(t), 0 ≤ t ≤ 1, be a geodesic on M, g(0) = r, g(1) = s. Let b x = (x1, ..., xn) ∈ M∗ be a subdivision of g(t), x0 = r, xn+1 = s and xi 6= xj for ∗ ∗ Vg(t) i 6= j. Let Lx be defined as above. Then ν(HLx) = dim rs . Again we use the notations of Section 6. Let g(t), 0 ≤ t ≤ 1, be a constant-speed b geodesic, with g(0) = p, g(1) = q. Let g(t) be subdivided by x = (x1, .., xn) ∈ M∗ , set p = x0, q = xn+1 and suppose xi 6= xj for i 6= j for i, j = 1, ..., n + 1. We claim that there are only a finite number of t-values, 0 < t < 1, which we call α , ...., α , such that dimVg 6= 0. (see [3] for the proof) 1 r pg(αi) Let g(s1) be a point of g(t) between xn and q, s1 6= αi, i = 1, ..., r. For s1 ≤ t ≤ 1, we define the functions L(t) and L∗(t) in the same way with respect to p and g(t) as the functions L and L∗ are defined with respect to p and q. From the remark after Theorem 7.1, it follows that x is a critical point of L(t) ∗ and therefore of L (t) for all t, s1 ≤ t ≤ 1. Using the triangle inequality, we find for t1 ≤ t2: 0 0 L(t2)(x ) ≤ L(t1)(x ) + ρ(g(t1), g(t2)) for x0 in a neighborhood of x. Since

L(t2)(x) = Lx(t1)(x) + ρ(g(t1), g(t2)) we find HL(t1)|x ≥ HL(t2)|x. By restriction (Lemma 7.2): ∗ ∗ HL (t1)|x ≥ HL (t2)|x. Applying Lemma 7.3 we find: ∗ ∗ X ∗ λ(HL |x) = λ(HL (s1)|x) + ν(HL (t)(x)),

s1≤t≤1 and by Proposition 7.4: ∗ ∗ X Vg (9) λ(HL |x) = λ(HL (s1)|x) + dim pg(t) . s1≤t≤1 b Now we replace the subdivision x of g(t) by a subdivision y = (y1, ..., yn) ∈ M∗ of the segment pg(s1) of g(t), in such a way that the t-value of yn is smaller than that the t-value of xn. b ∗ On M∗ there is a path Y from x to y such that L (s) non-degenerate at all of its points. Since HL∗(s) depends on s continuously and is non-degenerate for all the points of Y , we find ∗ ∗ λ(HLy(s)) = λ(HLx(s)). We choose a point g(s2) 6= α1, ..., αr between yn and g(s1), and have the same reasoning for the segment g(s1)g(s2) as we did above for the segment g(s1)q. We can go this way, such that after a finite number of steps we arrive at a point sl on MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 27 g(t), lying close enough to p that the function L(sk) has an absolute, non-degenerate b minimum at the subdivision z = (z1, ..., zn) ∈ M∗ , obtained by repeated application ∗ of our procedure. It follows that HLz(sk) and therefore HLz(sk) is positive definite; ∗ i.e., λ(HLz(sk)) = 0. Repeated application of (9) finally gives ∗ X Vg (10) λ(HL (x)) = dim pg(t) . 0

ρ(p, x1) = ρ(x1, x2) = ... = ρ(xn, q).

Figure 16. According to Theorem 7.l1 and the remark at the end of it, x is a critical point b L (which is a function differentiable in neighborhood of x on M∗ ) and of φ. Again, let Ci, i = 1, ..., n be differentiable cells, perpendicular to g(t) in xi, and ∗ ∗ N = C1 × ... × Cn. Let L and φ be the restrictions of L and φ to N respectively. x is a critical point of both L∗ and φ∗. We intend to prove the following relations: ∗ ∗ (11) λ(Hφx) = λ(HLx) = λ(HLx) = λ(Hφx). For the proof, we need the following more general lemma. Let M be a differentiable manifold and N a differentiable submanifold of M. Let p ∈ N be a critical point of the differentiable function f on M. Then p is also ∗ a critical point of the function f = f|N . We will compare the indices at p only, the subscript p will be omitted. Lemma 7.6. Under the conditions just described, λ(Hf ∗) ≤ λ(Hf). Further, if there exists a differentiable map τ : U → N of some neighborhood U of p on M into N such that: (i) τ(p) = p; (ii) d(τ|N ) is an automorphism of the tangent space TpN of N at p; (iii) f(τ(x)) ≤ f(x) for all x ∈ U; then λ(Hf ∗) = λ(Hf). ∗ Proof. We may identify Hfp with the restriction of Hfp to Np. It follows trivially from lemma 7.2 that under a restriction the index of a quadratic form can only ∗ ∗ decrease. Hence λ(HL) ≥ λ(Hf ). Now let τ (Hf) be the quadratic form on Mp defined by (τ ∗(Hf))(x, y) = Hf(dτ(x), dτ(y)). ∗ ∗ We can define τ (Hf ) similarly since dτ(Np) ⊂ Np. We have the following se- quence: (12) λ(τ ∗(Hf ∗)) = λ(Hf ∗) ≤ λ(Hf) ≤ λ(τ ∗(Hf)). 28 RAOUL BOTT

The first relation follows from the fact that dτ|Np is an automorphism of Np; that is, from condition (ii). The second we already used. The third is a consequence of condition (iii). Indeed, let F (x) = f(x) − f(τ(x)), x ∈ U. Then F ≥ 0, with F (p) = dF (p) = 0. Hence HFp is non-negative. But this Hessian is precisely Hf − τ ∗(Hf), whence Hf − τ ∗(Hf), and this clearly implies λ(Hf) ≤ λ(τ ∗(Hf)). On the other hand, we also have (13) λ(τ ∗(Hf ∗)) = λ(τ ∗(Hf)).

∗ ∗ ∗ ) To see this remark that τ (Hf ) is just the restriction of τ (Hf to Np. Hence (14) λ(τ ∗(Hf ∗)) ≤ λ(τ ∗(Hf)). Suppose A is a subspace on which τ ∗(Hf) is negative definite. Then A does not intersect the kernel of dτ, therefore B = dτ(A) ⊂ Np has the same dimension as A. By definition, τ ∗(Hf ∗) is negative definite on B and since dimB = dimA, we find: (15) λ(τ ∗(Hf ∗)) ≥ λ(τ ∗(Hf)). Combining (14) and (15), we get (13). From (13), it follows that the inequalities in (12) must in fact be equalities. In particular, λ(Hf) = λ(Hf ∗). This completes the proof of Lemma 7.5, and now we have all the necessary preparations for the proof of (11).  1 2 Proof of (11). Let F = φ − n+1 L . By the Cauchy inequality F ≥ 0. Further, near x, F is differentiable and F (x) = 0. Hence dF (x) = 0, and the Hessian of F is not 1 2 negative at x. Thus Hφx ≥ n+1 HL , whence

(16) λ(Hφx) ≤ λ(HLx). ∗ Our next steps is to show that λ(HLx) = λ(HLx) by means of Lemma 7.5. We have to construct a map τ : U → N of a suitable neighborhood U of x with the L(x) properties (i), (ii), and (iii) of Lemma 7.5. For thsi purpose, let , 0 <  < 2(n+1) ,  be a real number and let Ci be the parallel translates of Ci along g(t) by a distance − , in the direction from p to q. Similarly, let Ci be the -translate along g(t) in ± the opposite direction. The geodesic g meets Ci transversely. Hence there is a 0 0 neighborhood U of x, such that for all y ∈ U the polygon β(y) (Section 6) also ± ± meets the Ci transversely, say at points ηi (y). These points, taken in the obvious − + − order p, η1 , η1 , η2 , ..., defines a polygon η(y). We now define the components of τ(y) to be the intersection of η(y) with Ci. The transformation τ is well defined and differentiable on some smaller vicinity U of x, again because τ(x) = x, so that η(y) will meet the Ci’s transversely for y close enough to x.

Figure 17. MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 29

In figure 17, the situation is described graphically. τ maps U into N, further τ(x) = x and L(τ(y)) ≤ L(y) as follows from the triangle inequality. This all requirements of Lemma 7.5 are satisfied except possibly (ii). We therefore still L(x) have to show that τ|N has a non-singular differential for 0 <  < 2(n+1) , and clearly depends continuously on . However, for  = 0, τ0|N is the identity. Hence for  sufficiently small, d(τ|N ) is non-singular. Applying Lemma 7.5, we have established ∗ (17) λ(HLx) = λ(HLx). From (7) we deuce n 2L(x) X ), (18) UV Hφ∗ = U (V + − V x n + 1 i i i i=1 where the symbols have the same meaning as in (8). From (8) and (18) it follows ∗ ∗ (19) λ(HLx) = λ(Hφx). Finally, the first part of Lemma 7.5 gives: ∗ (20) λ(Hφx) ≤ λ(Hφx) Combining (16), (17), (19), (20), we proved (11).  From (10) and (11), we deduce the Index Theorem. Theorem 7.7 (The Index Theorem). Let g(t), 0 ≤ t ≤ 1, be a geodesic on M, b g(0) = p, g(1) = q and let x = (x1, ..., xn) ∈ M∗ be the subdivision of g(t) given by ρ(p, x1) = ρ(x1, x2) = ... = ρ(xn, q). Then x is a critical point of the function 2 2 2 φ = ρ (p, x1) + ρ (x1, x2) + ... + ρ (xn, q), and X Vg λ(Hφx) = dim pg(t), 0

8. Critical Manifolds Definition 8.1. Let f be a smooth function on the manifold M. The connected submanifold V of M will be called a non-degenerate critical manifold of f, if (1) every point x ∈ V is a critical point of f; (2) for all x ∈ V , the null space of Hfx is precisely the tangent space to V . This definition generalizes the concept of a non-degenerate critical point.

From the definition, it follows that also the index λ(Hfx) is the same for all points x ∈ V . Therefore, we can speak of the index of f on V . It will be denoted by λf (V ) or simply by λ(V ) if there is no danger of confusion. Lemma 8.2. Let f be a smooth function on M, and p ∈ M a critical point of f. If in a convenient neighborhood of p, we set ρ(p, y) = φ(y) and hdφ, dfi = F , then p is a critical point of F and HFp = 4Hfp.

Proof. By direct verification.  30 RAOUL BOTT

Lemma 8.3. Let f be a smooth function on M which assumes its absolute minimum on the non-degenerate critical manifold V ⊂ M. If there are no other critical points of f on M a and if M a is compact, then M a ∼ V . Proof. Let p ∈ V , and let f 0 be the restriction of f to the “geodesic plane” per- pendicular to V at p. The gradient of f 0 is transverse to an -sphere in this plane 0 of  small enough. This follows from lemma 8.2, since Hfp is non-degenerate by the definition of a non-degenerate critical manifold. From the compactness of V , we deduce the existence of a global tubular neighborhood of N of V in M such that ∇f is transverse to the boundary of N. In the same way as in the proof of theorem 2.1, a deformation of M a in N can be constructed. Combining this with  the obvious fact that N ∼ V , we get the statement of the lemma.  Lemma 8.4. Let f be a smooth function on M such that for a ≤ x ≤ b, there is only one critical value x = c, a < c < b. Suppose furthermore that f −1(c) is the non-degenerate critical manifold V . If M b is compact, then: b a M ∼ M ∪ eλ1 ∪ ... ∪ eλr . with λi ≥ λ(V ).

Proof. Let N 1 and N 2 be closed, sufficient;y small tubular neighborhood of V in b M, of radius 1, 2, respectively, 1 < 2. Let φ(x) be a smooth function on M with the following properties: φ(x) = 1 on N 1 φ(x) = 0 outside N 2

Figure 18.

Let π : N 2 → V be the fibre projection and g a (strongly) non-degenerate function on V . We set f 0 = f + αρ(x)π∗(g), α > 0, and assert that for α sufficiently small, f 0 has non-degenerate critical points only, all lying on V and precisely the critical points of g. We consider f −1(a < x < b) ⊂ M. Outside N 2 , f 0 has no critical point. Between N 1 and N 2 and onto their boundaries, we have: |df| > A, d|ρ(x)π∗(g)| < B, and therefore: |df 0| ≥ |df| − αd|ρ(x)π∗(g)| > A − αB > 0 for α sufficiently small. By looking at a restriction to a fibre of N 1 , we see that on N 1 , f 0 has only critical points on V and there only at the critical points of g. In MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 31 these points, the null space of Hf is just the tangent space to V (by definition) and that of π∗(g), the tangent space to the fibre of N 1 through that point (by direct verification). It follows that the critical points are non-degenerate and that their indices are at least λ(f)p = λ(V ).  9. The Stable Homotopy Groups of the Unitary Group In this final section, we shall apply the results of the preceding sections to the calculation of the stable homotopy group of the unitary groups. This means the following: The exact homotopy sequence of the fibering U(n)/U(n − 1) = S2n−1 gives us πk(U(n)) = πk(U(n − 1)), 0 ≤ k ≤ 2n − 2, from which it follows that k + 2 π (U(n)) = π (U(m)), n, m ≥ . k k 2 This group is called the k-th stable homotopy group of the unitary group(s). It will be denoted by πk(U). We start with a simple result concerning Jacobi fields on a compact connected Lie group, considered as a Riemannian manifold by providing it with a left and right invariant Riemannian structure. Since we need only the result for SU(n), we shall state and prove it for this group, but the general case is no more difficult. The reason, that we consider SU(n) rather than U(n) is that Ω(SU(n)) is connected but Ω(U(n)) is not. The way in which we use this will become clear in the sequel. The Lie algebra L of SU(n) is provided with a metric m0 in a natural way, namely with the restriction to L of the metric m, introduced in Section 5. For Y ∈ L, V (ρ, t) = eρY · etX · e−ρY , −∞ < ρ < +∞ defines a variation of the geodesic g(t) = etX , X ∈ L. On g(t), we consider in particular the segment s : 0 ≤ t ≤ 1, starting at the identity e of SU(n) and ending at a = g(1). V (ρ, t) induces a Jacobi field along g(t), given by

∂ ρY −ρY (1) UY (t) = (e · g(t) · e ) . ∂ρ ρ=0 Writing instead of eρY · g(t) · e−ρY , eρY · e−Adg(t)ρY · g(t), we see that

(2) UY (t) = (Y − Adg(t)(Y ))g(t). Vg The result we shall prove is, that all elements of ea can be written in the form (1) for a suitable choice of Y . Let J denote the vector space of Jacobi fields along s, which vanishes at e (but not necessarily at a). As remarked in Section 7, a Jacobi field along g(t) is completely determined by its value at e and the value of its covariant derivative along g(t) at e. From this, it is obvious that dim J = dim SU(n) = dim L. Let P ⊂ L be the subspace of L, consisting of all those elements of L which induce the zero Jacobi field along s. From (2) and from

Ad tX Y − Y [X,Y ] = lim e , t→0 t 32 RAOUL BOTT it follows that (3) P = {Y ∈ L, [X,Y ] = 0}. Consider the sequence α β⊕γ (4) 0 −→ P −→ L ⊕ P −→ J −→ 0, where α denotes the injection of P into L, β the homomorphism which attaches to Y ∈ L, UY ∈ J , and where γ is defined as follows: γ(Z), Z ∈ P , will be the Jacobi field along g(t) induced by the variation (5) W (ρ, t) = et(ρZ+X), −∞ < ρ < +∞ of g(t). Since W (ρ, 0) = e for all ρ, γ(Z) ∈ J for all Z ∈ P . We shall prove that the sequence (4) is exact. For this, it is sufficient to prove: β(Y ) + γ(Z) = 0, Y ∈ L, Z ∈ P implies: Y ∈ P , Z = 0. Indeed, if we prove this, then β ⊕ γ is automatically surjective, since dim kernel(β ⊕ γ)) = dim P , dim image(β ⊕ γ)) = dim L + dim P − dim P = dim L = dim J . Now, (3) allows us to write instead of (5), W (ρ, t) = etρZ · etX ,

∂ tX (6) W (ρ, t) = tZe . ∂ρ ρ=0 If β(Y ) + γ(Z) = 0, then, by (2) and (6):

(7) Y − AdetX (Y ) + tZ = 0, 0 ≤ t ≤ 1. Denoting by ( , ) the inner product with respect to m0, we find:

(Y,Z) − (AdeX (Y ),Z) + (Z,Z) = 0.

Since (Y,Z) = (AdeX (Y ), AdeX (Z)), this leads to

(8) AdeX (Y ){AdeX (Z) − Z} + (Z,Z) = 0. Z is in P , therefore by (2), the first term on the left side of (8) vanishes and we find: (Z,Z) = 0, hence Z = 0, hence, by (7), Y ∈ P . This proves the exactness of (4). Vg Finally, to prove that the elements of ea can be written in the form of (1), it −1 Vg is sufficient to prove that (β ⊕ γ) ( ea) ⊂ L. But the pair (Y,Z)(Y ∈ L, Z ∈ P ) lies in the inverse image if and only if (7) holds for t = 1. This gives again Z = 0. Proposition 9.1. Let SU(n) be provided with a left and right invariant Riemann- ian structure, let g(t), 0 ≤ t ≤ 1, g(0) = e, g(1) = p be a geodesic segment on SU(n). Vg Then every element of ep is induced by a variation of g(t) of the form eρY · g(t) · e−ρY , −∞ < ρ < +∞, where Y denotes an appropriate element of the Lie algebra of SU(n). Our argument gives the same result for any compact Lie group, but we do not need this fact, in its turn a very special case of the “variational completeness” of symmetric spaces (see [6]; also [5] for the proof given above). We consider the situation of Section 7, with M = SU(2m), p = e and q = −e. b b According to Theorem 7.1, the critical points of φ on Mn = M∗ correspond to the n-tuple (x1, ..., xn), such that ex1x2...xn(−e) is a geodesic segment on SU(2m), MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 33 and such that ρ(e, x1) = ρ(x1, x2) = ... = ρ(xn, −e). Let y = (y1, ..., yn) be such an n-tuple. Then all the transforms of y by the natural operation of SU(2m) on M∗ (induced by the adjoint action of SU(2m) on itself) are also critical points of φ. It b b follows that the critical points of φ on N∗ = M∗ minus its boundary consist of a b series of compact submanifolds of N∗ , on each of which φ is constant. We shall prove that these critical manifolds are non-degenerate in the sense of Section 8. Let   iα1 0  ..  X =  .  0 iα2m be an element of the Lie algebra of U(2m), considered as the space of skew- symmetric matrices (see Section 5). X lies in the Lie algebra of SU(2m) if and only if

2m X αk = 0. k=1 In that case, the geodesic

eiα1t 0  tX  ..  e =  .  0 eiα2mt lies on the maximal torus D, consisting of the diagonal matrices of SU(2m). b Now let C be a critical manifold of φ on N∗ , let x ∈ C, and let g be a geodesic −1 −1 ex1x2...xn(−e). There is an element ω ∈ SU(2m), such that ωx1ω ∈ D. ωgω −1 −1 passes through ωx1ω lying on D, so ωgω ⊂ D. Therefore, every critical manifold contains at least one point x = (x1, ..., xn), such that ex1...xn(−e) ⊂ D. To prove our statement, we have to show, that for a point p on a critical manifold C of φ, the zero space of Hφp is precisely the tangent space to C at p. A trivial calculation shows that every tangent vector to C at p certainly lies in the zero space. It therefore suffices to show that ν(Hφp) = dim C. By the same kind of argument as ∗ used in Section 7 to prove that λ(Hφp) = λ(Hφp), it can be proved (in the general ∗ Vs case) that ν(Hφp) = ν(Hφp). By Proposition 7.5, this nullity equals dim e(−e), where s denotes the geodesic segment ex1...xn(−e). If we denote by G the subgroup of SU(2m), leaving s fixed, we have: dim C = dim SU(2m) − dim G. On the other Vs hand, if follows from Proposition 9.1, that dim e(−e) = dim L − dim P . Since dim L = dim SU(2m), and since it follows easily from (3) that dim P = dim G, we have proved our theorem that C is a non-degenerate critical manifold. From the proof, it follows also that to determine the indices of the critical mani- fold C, it is sufficient to determine the indices of geodesic segments g(t), 0 ≤ t ≤ 1, from e to −e, lying on D and considered as representing a critical point x of φ on b N∗ . By Theorem 7.7 (the index theorem), this index equals

X V(g) dim eg(t), 0

V(g) provided that dim eg(t) differs from 0 only for a finite number of t-values, 0 ≤ t ≤ 1. The argument, giving us the equality V(s) dim e(−e) = dim SU(2m) − dim G, tells us also, that V(g) dim eg(t) = dim Gg(t) − dim G, where Gg(t) denotes the subgroup of SU(2m), leaving the point g(t) fixed. Let the segment s : 0 ≤ t ≤ 1 on g(t) be given by

e2πiα1t 0   ..  g(t) =  .  0 e2πiα2mt Since g(t) ⊂ SU(2m), 2m X αk = 0, k=1 and since g(1) = −e, 2β − 1 α = k , k = 1, ..., 2m, β integer. k 2 k

Conversely, every sequence of rational numbers α1, ..., α2m, fulfilling these con- ditions gives us a geodesic segment from e to −e on D. 1 1 If we take α1 = ... = αm = 2 , αm+1 = ... = α2m = − 2 , then it can be checked easily that for all values of t, 0 < t < 1, dim Gg(t) = dim G. Therefore, this α-sequence gives a geodesic ex1...xn(−e), such that λ(x) = λ(C1) = 0, x ∈ C1. Consequently, on this C1, φ assumes its absolute minimum a0.

C1 = SU(2m)/S(U(m) × U(m)) = U(2m)/(U(m) × U(m)) is a Grassmann variety. If we take for example the sequence 3 1 1 α = , α = ... = α = , α = ... = α = − , 1 2 2 m−1 2 m 2m 2 1 then on the corresponding geodesic, there is for 0 < t < 1 just one t-value t0 = 2 , V(g) 1 such that dim 1 6= 0 (there is just one value of t, 0 < t < 1, namely t = 2 , eg( 2 ) 3 1 such that 2 t and − 2 t are congruent modulo 1). V(g) dim 1 = dim S(U(m + 2) × U(m − 2)) − dim S(U(1) × U(m − 2) × U(m + 1)) eg( 2 ) = 2m + 2. It can be checked that the index of all geodesic g corresponding to α-sequence in 1 which not only αk = ± 2 appears, is at least this number. In fact, it is sufficient to remark that in such a case, there is at least one point g(t ) with dim Vg 6= 0, 0 eg(t0) 0 < t0 < 1 (and on the other hand only a finite number), and this dimension equals

S(U(n1) × ... × U(nk)) − S(U(m1) × ... × U(ml)), where m1 +...+ml = 2m is a refinement of n1 +...+nk = 2m. Then this expression is at least 2m + 2. (see Milnor’s Morse Theory, p.131) MORSE THEORY AND IT’S APPLICATION TO HOMOTOPY THEORY 35

Since two geodesic segments on D from e to −e and corresponding to α-sequences 1 with all αk = ± 2 clearly lies in the same critical manifold C, we deduce from Theorem 6.1, Lemma 8.2, Lemma 8.3 and the results above that for every a > a0

a Ωe(−e)(SU(2m)) = SU(2m)/S(U(m) × U(m)) ∪ ed1 ∪ ed2 ∪ ... ∪ edi with d1, d2, ..., di ≥ 2m + 2, and i depending on a. Therefore

πk(Ωe(−e)(SU(2m))) = πk(SU(2m)/S(U(m) × U(m)))

= πk(U(2m)/U(m) × U(m)), 3 ≤ k ≤ 2m,

Since, as stated at the start of Section 1,

πk(Ωe(−e)(SU(2m))) = πk+1(SU(2m)), we can conclude:

(9) πk+1(SU(2m)) = πk+1(U(2m)) = πk(U(2m)/U(m)×U(m)), 3 ≤ k ≤ 2m.

At the very beginning of this section, we remarked that

πk(U(2m)/U(m)) = 0

Using the exact homotopy sequence of the fibering

(U(2m)/U(m),U(2m)/U(m) × U(m),U(m)) we find:

(10) πk(U(2m)/U(m) × U(m)) = πk−1(U(m)), 1 ≤ k ≤ 2m − 2.

Combining (9) and (10), we get for large m:

πk+1(U(2m)) = πk−1(U(m)), k ≥ 3.

Since π0(U) = π2(U) = 0 and π1(U) = π3(U) = Z ([7], p. 132), we have

πk+2(U) = πk(U), k ≥ 0, and explicitly

π2k(U) = 0,

π2k+1(U) = Z, k = 0, 1, 2, ...

Acknowledgement. I typed this book as part of my Summer 2020 REU on Morse theory. I want to thanks my mentor Professor Peter Petersen for recommending this book, resolving my confusions, and proofreading this document. I also want to thanks Zhenyi Chen for our fruitful discussions on Morse Theory over the two months. 36 RAOUL BOTT

10. References [1] H. Seifert and W. Threlfall: Variationsrechnung im Grossen, Teubner, Leipzig 1938. [2] K. Nomizu: Lie Groups and Differential Geometry, Publications of the Math- ematical Society of Japan, vol. 2, Tokyo 1956. [3] M. Morse: The in the Large, A. M. S. Colloquium Publications, vol. 18, New York 1934. [4] W. Graeub: Linear Algebra, Die Grundlehren der Mathematichen Wissenschaften, vol. 97, Springer, Berlin 1958. [5] R. Bott: An Applicaition of the Morse Theory to the Topology of Lie Groups, Bulletin dela Societe Mathematique de France, vol. 84(1956), pp. 251-282. [6] R. Bott and H. Sameloon: Application of the Theory of Morse to Symmetric Spaces, American Journal of Mathematics, vol. 80(1958), pp. 964-1029. [7] N. Steenrod: The Topology of Fibre Bundles. Princeton 1951. For the materials of these notes, see also [8] R. Bott: The Stable Homotopy Groups of the Classical Groups, Annuls of Mathematics (2), vol. 70(1959), pp. 313-337.