This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
= 4(p + 2q)2(2p + 8q)-1, which will be used below.
407
LENGTH SPECTRA OF MANIFOLDS
3.
The trace formula
Let T be a discrete subgroup of G such that T \G is compact. Fix a GifG nvariant measure dx on T \G by requiring that for each f E C,(G), we have f (x)dx =
J r\G
(Z',, f(rx)dx. Let T be an irreducible unitary representation
of T on a finite dimensional vector space V, and denote by U the representation of G induced by T. Thus U acts on the Hilber t space H consisting of functions f : G --> V which satisfy (i) f (rx) = T(r)f (x) and (ii)
J r\G
(f (x), f (x))dz
< co where (., ) is the inner product on V. The action of G on H is by right translation. Thus (U(x)f)(y) = f(yx), x, y E G, f e H. U is a unitary representation of G. Under our assumption of compactness for T\G, it is well known that U is a discrete direct sum of irreducible unitary representations of G, each occurring with finite multiplicity. Denoting by 61(G) the set of equivalence
classes of irreducible unitary representations of G, we let nr(w, T) be the number of summands of U which lie in the class w. Then we can write U )1-GB(G) nr(w, T)w,, and nr(w, T) < co for each w.
For f E L1(G) let U(f) = J f(x)U(x)dx. U(f) is a bounded operator on H. G
As in [18], [7], we say that f is admissible if (i) the series E, f(y-lrx)T(r) converges absolutely, uniformly on compacts of G X G, to a continuous End (V)-valued function F(x, y, T) and (ii) the operator U(f) is of trace class. When f is admissible, we have the trace formula (3.1)
E nr(w, T) Trace U,(f) =
.G.(G)
fr\G
Trace F(x, x, T)dt ,
where U. is a representation of class w E &(G). Of course, U.(f) has a trace because U(f) does. As in [18], one rewrites the right side of (3.1) to get the Selberg trace formula nr(w, T) Trace U,(f) =
(3.2) mEB(G)
E Trace T(r) Vol (T,\G,)I,(f) 7EGr
where Cr is a complete set of representatives in T of the conjugacy classes of
elements of F, and G, is the centralizer of r in G, F,= T fl G,. Since T \G is compact, every element of T is semisimple, and G, is reductive, and T,\G, is compact. We fix a Haar measure dx, on G, in a manner analogous to the manner in which the Haar measure on G was fixed, following the Iwasawa decomposition of G,, and put dx, for the invariant measure on T,\G,. The volume Vol (T,\G,) is computed with respect to this measure. Finally, I,(f)!=
f(x-'rx)dx*, where dx* is the G-invariant measure on G,\G normalized fG,\G
so that dx = dx,dx7 .
408
RAMESH GANGOLLI
The use of (3.2) depends on having a stock of admissible functions. The following proposition was proved in [7]. Proposition 3.1. Let f E '1(K\G/K). then f is admissible. A similar assertion holds if f E W1(G) and is left and right K-finite. We shall only need this special case.
Let f E '1(K\G/K). Then U,(f) = 0 unless is of class one with respect to K, i.e., unless the restriction of U., to K contains the trivial representation of K. When U., is of class one, there is associated with it a unique positive definite elementary spherical function ¢,. say, A., E A. Then Trace U,,(f) = where ? is as in (2.1). (Cf. [6]). Thus, when f E '1(K\G/K), we get (3.3)
E nr((o, T) f(j.,) _ E Trace T(r). Vol (rr\Gr)Ir(f)
.Er(G,1)
7E Up
where (9(G, 1) stands for these elements in (9(G) which are of class one. We shall now compute the integrals I7(f) for f E W1(K\G/K), in a form suitable for use in § 4. An element x e G is said to be elliptic, if it is conjugate to some element of K and is then automatically semisimple. x E G is said to be hyperbolic, if it is semisimple but not elliptic. In all other cases x is said to be parabolic. When
G/I' is compact, r does not contain parabolic elements. It is well-known that r e r is elliptic if and only if it is of finite order. Both these properties are equivalent to the property that r has a fixed point on G/K. We assume throughout that r contains no nontrivial elliptic elements. Thus each r E r, r # 1, is hyperbolic. The integrals I7(f) can be computed for hyperbolic r quite simply, and can be expressed in terms of the Abel transform Ff of (2.2) when f is spherical. Let J be a Cartan subgroup of G with Lie algebra j, 0+ a set of positive roots for 0 _ ftc, jc). For any a e 0+ let , be the corresponding character of J. Put p, = 2 a, and PT = exp p(log h). We may assume that Pj is a well-defined character of J. Put 4,(h) = P(h) 11 aEm+ (1 - ba(h)-'), h e J, and let 0f be the invariant integral of f relative to J (cf. [9]). Thus (3.4)
0f (h) =
(h) f
f(x lrx)dx* .
J\G
Here a',(h) = sign IIaEIR (1 - ba(h)-'), the product being over the set 0 of real roots in 0+, i.e., those which are real on j, the Lie algebra of J. The Haar measure dh on J is normalized as mentioned in § 2 above, and dx* is the G-invariant measure on J\G such that dx = dh dx*. 0 is defined and smooth on T= J n G' = the regular points in J. For r E r, let Gr be its centralizer with Lie algebra gr, and let jr be a 0stable Cartan subalgebra of gr which is fundamental. Then one knowns t hat I7(f) and 0'r are related to each other, thanks to a theorem of Harish-Chandra
409
LENGTH SPECTRA OF MANIFOLDS
[10, p. 33]. If we let 0; be the set of positive roots of (gG, jr), and put H, _ rf a E ; H,, then we know that (3.5)
J,(f) = C,r fr(r ; H) ;
C, # 0
,
where 0 fr(r ; H) is the result of applying the differential operator H, to the function 0fr, and evaluating the result at T. All this is well-known and can be found, e.g., in [23]. The value of C, will be useful for us. It can be computed by using [10, Lemma 23], and [23, II, Chap. 8]. One should bear in mind that our normalizations of Haar measure differ from those used in [23, II, Chap. 8]. The value of C, is found to be pKr>-1 .(2)- r.2,-ar
(-1)mr[Wx,]
Cr =
aE 0 y K
(3.6)
x
fZ
(1 -
a E o9/gr
Here m, = -(dim G, - rank G, - dim K, + rank K,), n, = -(dim (G,/K - rank (G,/K,)), Wx, is the Weyl group of K, and [W,,] is its cardinality, 0r K stands for the compact roots in 0r , px, is the half sum of these roots, and
i9r is the complement of 0. in t+. Recall that we have assumed that rank (G/K) = 1. In this case there can be at most two nonconjugate Cartan subgroups. One of these is always noncompact, namely A = A,A,, and dim A0 = 1. When another nonconjugate Cartan subgroup exists, it is compact, and we may call it B. Thus there are two invariant integrals 0f and 0' .
We shall compute 0f for f E '1(K\G/K) and relate it to F f. Let a be a regular element of A, and let a = a,a,, ar E A, a, E A, Then (3.7)
F f(a,) _ ,(a,) f f(an)dn = Y
_
f (an)dn
fZv
Since f(an) = f(a,an) = f(an).
For regular a, the map n->a-In-lan is a diffeomorphism of N onto N whose Jacobian is computable. (See e.g. [11, Chapter X]). Thus
Ff(a,) = ,(a,)
fZ aEP+
(1 - a(a)-') f x f(n-lan)dn
(3.8)
_ ,(a')
I
fZ
aEP+
(1 - a(a)-') f"K f N
f(k-'n-1ank)dn dk .
since f is spherical.
The last integral can be transformed as in [10]. It equals f
f(x-1ax)dx, , Ap\G
410
RAMESH GANGOLLI
where dx = da, dx1*. Since Al is compact and carries normalized Haar measure,
this last integral equals f A\G f (x-lax)dx*. Also, if a E P+, so does a. Hence the
product ]]aE p_ (1 - a(a)-1) is real and has the same sign as ]]aEP+ (1 -a(a) ') a real
which of course is precisely ER(a). Using all this, we get
Ff(ap) = ,(a,)ER(a) fl (1 - a(a)-1) 5A,G f(x-lax)dx* a
(3.9)
fl (1 - Sa(ar)-1)-1o f(a) _ bo(ar)-' aEP_ where we have used the fact that for a E P_, a(a) = 1 so that a(a) = Sa(ar). Thus finally, we have (3.10)
of (a) _ P(ar)
fl (1 -
aEP_
a E A'
Now suppose that r E F, r # 1, so that r is hyperbolic. Let h = h(r) be an element of A to which r is conjugate. Then I,(f) = I.W. Let h = hph, ; then h, # 1, since r is hyperbolic. Clearly, ac is a Cartan subalgebra of gc. If a E
0+(ge, ac), then ba(h) = 1, so a(hdga(h,) = 1. Since a is real on a,, and purely imaginary on a, it follows that ga(h,) = 1. Since dim a, = 1, and a is real on A., we conclude that a - 1 on A,, and so a vanishes on a,. Thus a E P_. Therefore +(g°, ac) C P_. It follows that Gh C MAC, and A, is in the
center of Gh. Hence A is fundamental in Gh. The operator Hh, equals II {aEP_7;sa(h)=1} H. In particular, each H. occurring here is in a, Thus, in ap7 plying 11 h to (3.10), we need only worry about the factor g,(ar) fl.,,_ (1 -
a(ar)-1), since H h will not act on F f(a1) at all. The result of applying H h to this function and evaluating the result at h is seen to be equal to [W1 ]
{aEP_; ea(h)=1}
a PKn X ,(hr) X {aEP_;flea(h)#1} (1 - ba(h)-1)
.
Cf. [10, Lemma 24] for a similar computation. Using (3.5), (3.6), we have the following proposition. Proposition 3.2. Let r be a hyperbolic element of G, and let h = h(r) be
an element of A to which it is conjugate. Let h = hrh,, hr E A, h, E A,. Then
I,(f) = Ih(f) =
(3.11) C(h)
=
flaEP+ (1 - ba(h) 1)) 1
One should note that C(h) is actually positive. For later use, we shall examine C(h) a little more carefully. Since ,(hd = exp p(log h,) _ exp 2 f aE + a(log h,), we see that C(h) equals ER(h)
fl (exp 2a(log h) - a(hr)-1 exp - 2a(log
aEP+
411
LENGTH SPECTRA OF MANIFOLDS
Since any a is purely imaginary on a, we must have ea(h,)-1 = a(hp) ; if a is a real root, then of course ga(hp) = 1. Now it is well-known in our case that there is at most one real root in P+. Denote this root by ao when it exists. Then the factor corresponding to it is exp ao(log h) / 2 - exp - a,(log h) / 2. The remaining roots in P+ will be denoted by P° . These are all complex, and occur in conjugate pairs a, a. Thus we can find a subset Q+ of P+ so that P+ Qo UQo Now let a e Q+, and consider the factors corresponding to a and a in the above product. We have a(h,)-1 = ga(h1) = ga(h,). Let Oa(h,) be the argument of ga(h,). Thus ga(h,) = exp iea(h,). Then these two factors have the product exp a(log hp) + exp - a(log h) - 2 cos Oa(h,). Now all the numbers a(log h) are of the same sign, depending on which Weyl chamber hp lies in. Using this remark one quickly finds that
(3.12)
C(h) = exp - I p(log hp) I x (1 - exp - I ao(log hp) I) X fl (1 - 2 cos Oa(hr) exp - I a(log h) I
aGQ+
+ exp - 2 I a(log h) I)-1;
when P+ contains no real root, the factor corresponding to ao is, of course, absent.
4.
The length spectrum
As we have said in § 1, our results follow from applying the trace formula to suitable admissible functions, mainly to the fundamental solution of the heat equation on G/K. Let Q be the Casimir operator of G, and for t > 0 let g,(x) be the fundamental solution of the heat equation Qu = au/at on G/K, with u assumed spherical. The properties of g, are discussed in [4]. Let us briefly recall them. As a function on G, g, is spherical, nonnegative real valued, and gc+8 = g*g8,
for t, s > 0. g, is the fundamental solution in the sense that for any f e C(K\G/K), for example, the function U(x, t) = (g*f)(x) is the unique spherical solution of 2u = 8u/at such that u(x, t) - f (x) --> 0 uniformly on compact sets as t, 0. The function g, is in L1(K\G/K) for each t > 0, and k, can be computed. Indeed, k,(2) = exp - (<2, 2> +
)t. Since g, is integrable, k, is defined for all 2 such that 4°x is bounded, thus in the tube A + iCe, and the above formula for k, holds there. It follows, for example by using [21], that
g, e'1(K\G/K). In particular, g, is admissible. Since k,(2) is known, it is possible to compute the Abel transform Fg, by using the Fourier inversion formula. We get, remembering dim Ap = 1, (4.1)
Fgt(ap) =
(4;rt)-1/2 exp
- (t
+ Ilog apl2/(4t))
412
RAMESH GANGOLLI
Of course, a similar formula would hold when the dimension of A, > 1, but we would not be using it. Now applying (3.3) to gt, using (3.5) and (3.6) we find o
(G,1)
nr(w, T) exp
(4.2)
- (<2.,2.> +
)t . Thus we conclude that L(t) = )t t + 4 log h,(r) j2/t) = (47rt)-112C(h(r)) exp - ( t + 4l(r)2/t) t + 4s2/t)(L(t) - g8(1) Vol. (F\G))) = 0} fill t + 4li/t)(L(t) - g,(1) Vol. (I'\G)) c-o t + 4li/t)} we find 12 to be the supremum of > 0 ; lim ((47Ct)112 exp ( t + 4e2/t) L2(t)) = 0} t + 4lz/t)L2(t) is positive and equals e2 = I2 7117;1(7)=1211(r)-1C(h(r)) t + 4E,/ t) i>1 t + II2l(o)2/t) t) )t v Ee(G,1) > 0 ; so the whole sum approaches zero. Next, we know [4] that )t ¢2(x) Ic(2) I-z d2 , where c(2) is the Harish-Chandra c-function. It follows that oat(1) = [WI-1 f A exp - (<2, 2> + )t Ic(2)L-z d2 t t(r)s:v l(r)i(7)-1C(h(7)) exp - 4l(r)z/t 112 = (p + 2q)(2p + 8q)-1'2. Proof. We deal first with Q0(l). The result for Q1(l) will be deduced from . We put 2°: = {(p, 0) I p e Rn} and p, : = {(0, q) I q e Rn} which are of course Lagrangean subspaces. Then the (real representation of) unitary group U(n) naturally acts on A(n) : = A(R'n) transitively, and its isotropic subgroup at 2° is given by 0(n). Thus A(n) is diffeomorphic to U(n)/O(n). Now M = U(n)/O(n) has a structure of a compact symmetric space whose riemannian structure comes from the Killing form of the Lie algebra of U(n). In the present note we shall determine the cut locus and the first conjugate locus of a point of M, from which we may prove the assertion mentioned above. For compact simply connected symmetric spaces, it is known that the cut locus and the first conjugate locus of any point coincide with each other (see [2]). Note that 7r1(M) = Z for our manifold M = U(n)/O(n). Finally we shall determine all closed geodesics of M and calculate their intersection number with the oriented cycle Uk,_1 Ak(n) Received October 6, 1975.
Z Trace T(r) Vol (I',\G,) I,(gt)
rECr
On the right side we get from the term corresponding to r = 1, the contribution gt(1) (degree T). Vol (r\G). The remaining elements Cr are all hyperbolic since r is assumed torsion-free. Call the sum of these remaining terms JH(t). It can be shown (cf. Eaton [3], or [4]) that limt-o JH(t) = 0. This is actually done in Eaton [3] under the additional hypothesis that T is the trivial representation. But the expression for JH(t) when T is nontrivial is clearly dominated in absolute value by a multiple of the correspondiug expression when T is trivial, since gt > 0. Hence JH(t) -> 0 in our case also. If L(t) denotes the left side of (4.2), it follows that lim tn'2L(t) =
tn2gt(1))(Vol. r\G) (degree T) o
/
Here n = dim (G/K). It is shown in [4] that limt-o tnl2gt(1) exists and equals C'G, a constant which
depends only on G. Thus limt.o tn'2L(t) = C'G Vol. (I'\G) degree (T). Now introduce, for r > 0, the function (4.3)
N(r, T) _
nr(w, T) , 19. I<,
where Q. is the scalar with which the Casimir element acts in any representation of class w. When w is of class one, one can compute 12. and find that
D. = -
J
e-tr'dN(r, T),
showing that L(t) is the Laplace transform of N(r, T). Of course, the admissibility of gt shows that N(r, T) is finite for each r, and L(t) exists. Arguing as in [4], we now find by Karamata's theorem that, as r --> co, (4.4)
r-n'2N(r, T) - C'GP(2 + 11 Vol. (I'\G). degree (T) ,
which is analogous to a classical result of H. Weyl [24]. When T is trivial this result is implied by that of Minakshisundaram and Pleijel [15]. Of course (4.4)
LENGTH SPECTRA OF MANIFOLDS
413
is just a step away from Eaton's result.' When T is trivial, N(r, T) is just the Weyl function of the manifold T'\G/K. More precisely, if {(ai, ni)}i,1 is the spectrum of the Laplacian on T"\G/K, is easily seen that N(r, 1) = Z (i; 12d:srj ni. We shall write N(r) for N(r, 1). Clearly the knowledge of N(r) is equivalent to that of the spectrum of the Laplacian. In particular, the spectrum of the Laplacian on F \ G/ K determines Vol (F \ G). Cf. [4], [15]. This will be needed below.
We now turn to the consideration of the length spectrum of R = T'\G/K. For this purpose, we have to compute the terms in (4.2) explicitly, with T = 1 ; this will be done next, resulting in (4.7) below. Cearly G/K is the simply connected covering manifold of R, and we can identify F with the fundamental group 'r1(R). It is well-known that the free homotopy classes of closed paths on R are in a natural one-to-one correspondence with the set of conjugacy classes of F, and hence with the set Cr. For any r r= Cr, the corresponding free homotopy class always contains a periodic geodesic g, say, which has minimum length among all the paths in that class [2]. Let l(r) be the length of g,. Any closed path in this homotopy class can be lifted to a path of equal length on G/K which joins some point m r= G/K to the point rm. It follows that the length l(r) of g, is the minimum of the lengths of paths joining some point m e G/K to its image rm under r. In fact, l(r) = Inf,, G/K d(m, rm) where d(., ) is the Riemannian distance on G/K. Now, if m = xK with x r= G, we have d(m, rm) = d(xK, rxK) = d(K, x-'rxK) = a(x-'rx), where a is the function introduced in § 1. It follows that l(r) = infXEG a(x-'rx). Notice that l(r) depends only on the conjugacy class of r, as it should. Moreover, for the computation of l(r), we can replace r by any element h of G conjugate to r, even if h does not lie in F at all. This remark enables one to compute l(r) more explicitly. Recall that r is conjugate to an element h = h(r) E A. Let h = h,ht ; h acts as an isometry on G/K, with no fixed points. Since G/K is of negative curvature, it follows from [2], [16] that there is exactly one geodesic of G/K which is stabilized by h. This geodesic is characterized by the property that a point p e G/K is on the geodesic if and 2 Actually, one does not need to assume that P is torsion free. In that case the right side of (4.2) splits into three terms, namely Jc(t), JE(t), JH(t), coming respectively from
central, elliptic and hyperbolic elements in Cr. Cf. [4]. One sees that Jc(t) = gt(l) Vol. (P\G) E,Ezncr Trace T(1), where Z = center (G), so that Jim t' /2JG(t) = C'' Vol. (P\G) E Trace T(1) . t-o EZncr One can show as in [3] that limt-0 t'/2JE(t) = 0, so that one gets r n/2N(r,T) -
C'GF(n/2 + 1)-' Vol. (I'\G) E,Ezncr Trace T(-,), which implies that E,Ezncr Trace T(7) must be nonnegative. If now T is irreducible, then T(7) is a scalar for ; E Z n r by Schur's lemma, and T(r) = x('). Identity, where x is a character of the finite abelian
group z n r. If x is nontrivial character, it follows that E7EZncr T(1) = 0, so that r-1a2/N(r, T) - 0 if T jzncr is a nontrivial irreducible representation. When T is not irreducible, Framer x(7) = E deg Ti, where Ti runs over those irreducible summands of
T which restrict to the trivial character of z n P.
414
RAMESH GANGOLLI
only if d(p, hp) = inf EGix d(m, hm). Now it is easy to see that the geodesic Exp A, (where Exp is the exponential map of G/K from p to G/K) is stabilized by h, (recall here that dim AP = 1). Moreover, if p c Exp a,, then d(p, hp) =
6(h). This shows that inf,EGI, d(m, hm) = 6(h), so that l(r) = u(h(r)). Of course, u(h(r)) = I log hp(r) 1.
Note that l(r) = l(7-1), (indeed the geodesics in the homotopy class r-1 are just reverse to those in r), and l(ri) = fl(r) for any integer j > 1. Lemma 4.1. Let r E I', r # 1. Then I', is isomorphic to Z. Proof.
r is hyperbolic, and by conjugation, we may assume r E A, rp # 1.
Let r', r" E f', and suppose r' = r'rt, r" = rvrrt Since G, C MAp as we have seen above, and ri commutes with r, we have 7'E MAC. Thus It follows that the set of elements {r, r' E f',} is a subgroup of rr(rv)-1rp(rz) Ap. Clearly this is a discrete subgroup, hence it is isomorphic to Z. Let o, be a generator for it, and let a E I', be such that 3 = 3p3,. We claim that 3 generates f', freely. In fact let r' E F_ Then rv = 8i for some j E Z. We claim that r' = V. Indeed, r(O ' Thus r'O-i E r fl K, so that r'3-j = 1 since F contains no elliptic elements # 1. Hence r' = 3i and our 1
assertion follows. Remark. Using the negative curvature of G/K, this result could also have been deduced from the theorem of Preismann [17], which is more general. In our special case, the above proof is more direct. Definition 4.2. An element r E f', r 1, will be said to be primitive if r
is a generator of f',. Clearly every r E f', r # 1, can be written as o3 with j > 1 integral, and a primitive. The integer j is unique and will be denoted by j(r). We will next compute Vol. (f',\G,). We may again assume r E A. Then G, C MA,. In fact G, = M,Ap, where M, = M fl G,. Let r = rvr,. Each element of M, commutes with both r and rv, hence with rz. If follows that rz commutes with G,, so rz acts trivially on G,/K,. Thus the action of r on G,IK, is the same as the action of rp. Now it is clear that K, = K fl Gr = M,, and since G, = M,Ap we conclude that the action of I', on G, /K, is the same as the action of jai, j E Z} on AP, acting by left translation. Here we identify AP ~ G,/K,. We thus get (recalling that the measures have been so normalized that K, carries normalized Haar measure), (4.5)
Vol. (I',\G,) = Vol. (f',\G,/K,) = Vol. (A,/Jai, j E Z})
.
The last term is clearly equal to Ilog 8,1 = 1(8). Moreover, since r = OJcr), we have l(r) = j(7)l(3). Thus (4.6)
Vol. (I'r\Gr) = l(r)j(r)
1
.
Using all this in the trace formula, (3.3) with T = 1
415
LENGTH SPECTRA OF MANIFOLDS
L(t) =
(4.7)
mEa(G,1)
nr(w,
1) exp -
= g,(1) Vol. (z'\G) +
(<2w, 2 > +
E
7E Cr-[1}
1(r)j(r)-'I7(g )
Moreover, if r is conjugate to h = h(r) E A, we also know that I7(gc) = I h,(gt)
(4.8)
= (4'rt)-1"2C(h(r)) exp - (
because, as we have seen above, l(r) = log hp(r) 1.
It follows that for each t > 0, the series E7ECr-[1} l(r)j(r) exp is convergent; one sees from this that the numbers {l(r), r E C, - {1}} have no finite point of accumulation. In particular, one may indeed order them 0 < 11 < 12 , and the multiplicity mi of each li is finite. (This can also be inferred on general grounds of course.) One immediate consequence of (4.7) is that the length spectrum {li}i,1 of R is determined by the spectrum of the Laplacian, or what is the same, by the function L(t). For, as we saw before, L(t) determines the volume Vol. (T\G), 4l(r)2lt
and hence the first term on the right side of (4.7). Then the smallest of the numbers {l(r); r E C, - {1}}, which is of course 11, is seen to be equal to the supremum of the set > 0; lim ((47ct)1I2 exp (
c-.o
This means that 11 is determined by L(t). Moreover, it is seen that lim (47rt)112 exp (
=
E
1(r)j(r) 1C(h(r)) = 11
{Y; c (7) = t11
E
j(r)
[7; a (7) = L1}
which is positive. Call this number e1. One can now subtract off the contribution to L(t) from IT; l(r) = l1}, and putting
L2(t) = L(t) - g,(1) Vol. (F \G) - {(47ct)-1I2e1 exp - (
and that 1im, (47rt)1"2 exp (
416
RAMESH GANGOLLI
Proceeding in this way, we see that L(t) determines both the numbers {Ii}i, and {si}i,,, where si = li E[rECp;t(r)=zi j(r)-1C(h(r)). Conversely, a knowledge
of these numbers and of Vol. (r\G) clearly determines L(t), and hence the spectrum of the Laplacian ; indeed L(t) = g,(1) Vol. (r\G) + E (47rt)-'"Isi exp - (
When G = SL(2, R), C(h(r)) depends on r only via 1(r). In fact C(h(r)) _
2 cosh (l(r)/2f), and so si = 2li cosh
j(r)-1 Thus
in this case, knowledge of the sequence {(li, si)} is equivalent to the knowledge of the sequence {(li, )ii)}, where )7, ={rECr,acr)=t j(r)_1 Since {(li, si)} characterizes L(t), we see that in this special case {(li, 7)j)} characterizes L(t). This }
result was originally observed by Hiiber [12]. As we have seen in § 3, the expression for C(h(r)) is more complicated in the general case, and does not depend merely on l(r). Returning to the general case, we let Prr, be the set of primitive elements
in Cr - {1}. Then we can write (4.9)
L(t) = g,(1) Vol. (P\G) +
BEPrr, 1>1
l(3)Iar(g,)
where (4.10)
Ia3(g,) = (4ret)-1/2C(h(3!)) exp - (
The set {l(o) ; 3 E Pr,} can be ordered in a sequence 0 < r1 < r2 < ; let pi be the cardinality of the set {8 e Prr, ; 1(5) = ri}. We call the sequence {ri} the primitive length spectrum, and the sequence {(ri, pi)} the primitive length spectrum with multiplicity. One can ask to what extent these are determined by L(t). Obviously, the set {r1} is contained in the set {li}, which is determined by L(t). So one must try and decide from a knowledge of L(t) whether a given number l; is in the set {ri} or not, i.e., if it is a primitive length or not. Obviously, if l; is not a multiple of some smaller lk, it must be a primitive length. However, if 1, is a multiple of some smaller lk, it could happen that l; is also the length of some other primitive geodesic as well. The author has not been able to decide this question in general by using the above formula. However, when G = SL(2, R), one can answer this question. Indeed in this case, L(t) is characterized by {(li, r)i)} which we can assume known. Now l1 is obviously equal to r1, and ri, equals p1i since j(r) = 1 for all r such that l(r) = 11. Now consider 2r1. It must be one of the numbers {li}i>,. Suppose 2r1 = li,. Then
the numbers {ls ; s < i, - 1} must all be primitive lengths. Thus r, = is and 7)s = ps for all s < i1 - 1. We can now decide whether li, is a primitive length or not. For if lit = ri1, then we should have rii, = 2p1 + pi,, and pi, > 0. Thus, if riil > 2P, = I , we can conclude that li, is a primitive length, li, = ri, and pi, = rii, - 27)i On the other hand if 7)j1 = 2pi, then li, is not a primitive
LENGTH SPECTRA OF MANIFOLDS
417
length. Next, let li2 be the smallest member of the set {li}i>ii, which is an integral multiple of some number l; smaller than it. By the definition of lie, it is clear that the numbers {13 ; it < s < iz} are primitive lengths, and so , = p, for these. As to lie itself, we can decide whether it is a primitive length by comparing i2 with the sum Z ((k,j),jrk=ail j>1) 1 /j. If 12 is strictly larger, then lie is a primitive length, and the difference between i2 and this sum gives its
multiplicity. Proceeding in this way, we see that L(t) determines both the primitive length spectrum and its multiplicity. Finally, let Si = {k > 1, jr, = li for some j > 1}. Then we have mi = Z,kGSi Pk Hence the length spectrum
with multiplicity is also determined by L(t) in this case. When G is not SL(2, R), these questions are not settled by the present method, and a close look at the computations seems to indicate that in general L(t) probably would not determine the primitive length spectrum or the multiplicities. To return to our main topic, define for any l > 0, (4.11)
Q,(l) _ [{6 e Prr, ; 1(6) < l}]
,
Q,(1) = [{r E Cr - {1},1(r) < l}]
[S] stands for the cardinality of S. We shall now determine the asymptotic behaviour of the functions Q0(l), Q1(l) as l -> co. For h e A, with h, # 1 put (4.12)
C, (h) = exp - p(log h,) I jj+ (1 + exp - j a(log h,) I) -'
(4.13)
C_(h) = exp - j p(log h,) I jj (1 - exp - I a(log h,) p-'
(4.14)
C0(h) = exp - I p(log h,) I
IX
,
.GP+
and define (4.15) F(t) _ (42rt)-1/2(exp -
Z l(r)j(r)-'C(h(r)) exp - 4l(7)2/t rEer-{1)
and let F, F_, Fo be defined analogously by replacing C(h) by C+(h), C_(h), C0(h) in (4.15). Lemma 4.3. Let H(t) be any of the four functions F(t), F+(t), F_(t), F0(t),
and let, for r > 0, ft(r) = f e-r'H(t)dt. Then H(t) -> 0 as t --> 0, H(t) --> 1 as t -p co, and rH(r) --> 1 as r --> 0. Proof. We know that for r E Cr - {1}, l(r) = log h,(r) I is bounded away from zero. Hence, if p = sup.ep,,rEcr-{I} exp - ka(logh,(1))j, we conclude that p < 1. Let D = ((1 + p)/(1 Then for each r E Cr - {1}, (4.16)
C+(h(r)) < C(h(r)) < C_(h(r)) < D C+(h(r))
where we used the expression (3.12) for C(h). Therefore
418
RAMESH GANGOLLI
F+(t) < F(t) < F_(t) < DF+(t)
(4.17)
and similarly
Fo(t) < F_(t)
(4.18)
Now we know, by the remarks immediately following (4.2), that F(t) (called JH(t) there) approaches zero as t - 0. From (4.17), (4.18) it follows that F+(t), F_(t) and F0(t) all do the same. We next claim that F(t) - 1 as t - oo. In fact by
F(t) = 1 + (4.19)
nr(w, 1) . exp - (<2w, 2.> +
- gt(1) Vol. (1'\ G)
.
As t - 0o , each term in the sum approaches monotonely to zero, because <2w, 2.> +
gc(x) = [W(G, A,)]-' f A exp - (<2, 2> +
again by monotone convergence, we conclude that gt(1) - 0 as t
Now
(4.19) shows F(t) - 1 as t - oo. We will now show that F+(t) - 1 as t - 00 . The other functions F_, Fo can be treated similarly. Using (3.12) it is easy to see that C+(h(r))/C(h(r)) - 1 as l(r) = I log h,(r) I - oo . Lets > 0 be given, and choose and fix N so large that for l(r) > N, we have (4.20)
(1 - s)C(h(r)) < C+(h(r)) < (1 + s)C(h(r)) .
Let FN(t), Fl(t) be the tails of the series defining F(t), F+(t) beyond l(r) > N. Then one sees (4.21)
(1 - s)FN(t)
F`v(t) < (1 + s)FN(t)
For each fixed N, the sum (4irt)-111 exp
-
is a finite sum and approaches zero as t -f c . Since F(t) -f 1 as t -f 00 , it follows that FN(t) -f 1 as t -f oo. Thus from (4.21) we deduce
LENGTH SPECTRA OF MANIFOLDS
419
(1 - s) < Jimt-- FN(t) < liil FN(t) < 1 +
t--
Now by examining the sum F+(t) - FN(t) we can similarly conclude that limt__ (F+(t) - FN (t)) = 0. This together with the above shows that (4.23)
(1 - e) < lim F+(t) < lim F+(t) < 1 +
s
Since s is arbitrary, we conclude F+(t) -* 1 as t - o o. The first assertion of the lemma is proved by proceeding similarly for F_, Fo. Since F(t) is nonnegative and F(t) -+ 1 as t -p 0, Karamata's theorem [25] shows that
rP(r) -p 1 as r -p 0, where P(r) = f e-rtdF(t)
.
Also, the functions F+(t) - F(t), F_(t) - F(t) do not change sign, and approach 0 as t -+ o o. So by the same theorem, we must have r(P+(r) - P(r))
- 0, r(F'-(r) - P(r)) -p 0 as r -p 0. Finally, F0(t) - F(t) does not change
sign, and approaches 0 as t --p o o. So we get r(P0(r) - P_(r)) --p 0 as r --p 0. Since rP(r) - 1 as r --p 0, the proof is finished. Theorem 4.4. Let Q0(l), Q1(1) be the functions defined in (4.11). Then we
have
(4.25)
2 p l exp - (2 p l)Q0(l) -* 1
as l-p 0
2Ipl l exp - (2IpI l)Q1(l) - 1
as l - 0
where 2 IpI = 2
it. Recall first the notations of § 1. Let h(r) be in A, and h(r) conjugate to r e C, - {1}. log h,(r) is a multiple of Ho ; say it equals u,H0. Then 1(r) = I log h,(r) I = I ur I. I Ho 1. Also I p(log h,(r)) I = I ur I I p(H0) I. Then
I p(log hv(1)) I = 1(r)
-
I p(Ho) I / I Ho I
It can be computed easily that Ip(Ho)I IHoI = 2(p + 2q)(2p + 8q)-"2 = IpIHence 2IpI = (p + 2q)(2p + 8q) -1/2 and I p(log h,(r)) I = I pIl(r). Since each r equals 81(r) with 8 primitive, and l(r) = j(r)l(8), we have (4.26) Fo(t) _ (42rt)-1'2 exp - IPI2 t Z Z exp - (j I p I 1(8) + 6EPrp j>1 Thus
f
e-rtFo(t)dt 0
4j2l(8)2/t)
.
420
RAMESH GANGOLLI
_Z Z l(o) exp - j I P 110) aEPrr j>1
(4.27)
f (47t)exp - (( p 2 + r)t + 'j21(8)2/t)dt 0
f(4zrt)-1'2 exp (-x2t - Iy2/t)dt = (2x)-1 exp - xy to get
Use the formula Fo (r)
= 1 (r + I P
I2)-1/2
12 (r +
2)-112
Z Z 1(8) eXp - (jl(3)(P 1 + -%/r -+I P I2))
aEPrr j>1
2
(4.28) P
aEPrr
/r +
eXp -
1(8)
P
1 - exp - l(3)(pI + r -+p I2)
Let (4.29)
Pie)-1/2
Go(r) = 2(r +
aEPrr
1(8) exp - l(8)(p + /r + pI2)
which converges by comparison with Fo(r). The ratio of the corresponding terms in G0(r) and Fo(r) approaches 1 as 1(8) - co. So an argument similar to that of Lemma 4.3 shows that rG0(r) and rF0(r) have the same limit as r - 0. Since we know rF0(r) - 1 as r - 0, we conclude rG0(r) - 1 as r - 0. Now rG0(r) = 2 r(r + I p
r(r +
I2)-1/2
aEPrr IPI2)-1/2fo
(4.30)
l(8) exp - l(8) (I P I + Ir + I P
lexp
-
r IPI2 - IP1)
IPI2
Wr+
IpI
X(IJP r+ r-IPI)
p2-
0
Writing z = r + -I p I2 - I p I , we see that z - 0 as r - 0. Letting r - 0 in the above expression we conclude lim z fo exp - zl l exp - 2 I p I l dQ0(l)
Now Karamata's theorem gives us the first conclusion of the theorem. (See the note added in proof.) As to Q1(1), we have
(4.31)
Q0(l) = [{3; 8 e Pry, 1(8) < l}] G Q1(l) _ [{r E Cr - {1} ; 1(r) G l}. _ [{(8, 8 e Pry, j > 1, jl(8) < l}] l [SEPrr;b(a)Sll 18
=f
a
l
0Y
a
2 lQo(Y)dY
vow = Qo(l) + f0 Y
421
LENGTH SPECTRA OF MANIFOLDS
Since we know the asymptotic estimate for Q0(l), the estimate for Q1(l) follows easily from this expression. This finishes the proof of the main result. One notes that the asymptotic behaviour of Qo and Q1 depends only on the
metric structure of the covering manifold G/K and not on the particular manifold R (or what is the same, on r). This theorem generalizes a result of H. Huber [12] who treated the case G = SL(2, R). Huber's method is slightly different; it was followed by Berard-Bergery [1] to G = SO0(d, 1), d > 2; Our method generalizes the method of McKean [14] who works with G = SL(2, R). These authors use a
metric on G / K which gives it curvature - 1 in their cases. Our metric is somewhat different. This introduces an inessential discrepancy between the values of I p I which they get there and we get here. Huber also proved the remarkable formula [12, p. 26], 2,v/7cr(s)
(S -
1)r(s - y) 2s--1
(4.32)
+ rlS 1
-
+
2)
r(s)
r(s - 1)
nr((, 1)r( 2 (S - S-(2m)))F(2 (S -
Vol. (r/G)
2
Z
recr-{11
l(r)i(r)-1(cosh l(r) - l)-"'(cosh l(r))
1i2
where s±(2,.,) are the roots of S2 - S - J. = 0, and G = SL(2, R). J. is the eigenvalue of the Laplacian. One must bear in mind that Huber used the metric which gives curvature - 1 to G/K. HUber's proof of (4.32) utilizes methods involving the Green's function of the upper half-plane. Huber used the above formula together with the theorem of Ikehara to get the analogue of Theorem 4.4 for G = SL(2, R). A generalization of (4.32) for G = SO0(d, 1) is presented by Berard-Bergery in [1, p. 118], and is used there similarly to obtain Theorem 4.4 for G = SO0(d, 1). Both (4.32) and its generalization to SO0(d, 1) in [1] result from the traceformula by the choice of a suitable admissible function f,. One must, of course, compute f s and Ff.* In fact, let x e G, and x = ka,k', k, k' e K, a, e 4 be its polar decomposition. Put J Ho j = c (recall that this equals ,/2p _+8q). Let
Q e I be as in § 2, and put t = t(a,) = p(log a¢). Then t can be regarded as a coordinate on 4 Consider, for a complex S, the function fs(x) = (cosh t)-s where t = t(a,) and x = ka,k'. f s is clearly spherical. If Re s > p + 2q, one can show that f, E '1(K\G/K), so that f, is admissible. (4.32) and its generalization result from applying the trace formula to this f s. It is possible to compute the analogue of (4.32) for all the groups of rank (G/K) = 1 by computing j,, F f. directly. Since the main application of these formulas was to get
422
RAMESH GANGOLLI
Theorem 4.4 which we have obtained by other means, it does not seem worthwhile to give details of the derivation. We will content ourselves with quoting the result, which may amuse the reader : 72r(U), 1)7r(P+q+1)/2
(4.33)
rQ'(s - s-lam)))1 (2(s - s+lam)))
r(2s)I'('(s - q + 1)) = Vol. (r1G) + 7r(p+q+1)/2 21-s+(p+2q)/2 r(s - s(p + 2q)) r(ss)r(I(s - q + 1))
xE
, E Cr- (i}
1(7)1(7)-'C(h(7))(cosh l(r))-s+(P+2q)/2
where st(2) are the roots of the equation
S2 - s(p + 2q) + (p + 2q)2 + 2 (Ho)2 = 0 . Thus
st (2.) = (p + 2q) ± 2
-.1(Ho)2 = p(Ho) ± ia.(Ho)
The reader will easily check that when p = d - 1, q = 0 (which is appropriate for G = SO0(d, 1)), one gets from this the formula of [1, p. 1181. (4.32) results from p = 1, q = 0. The difference of metrics must be borne in mind. For the other groups G, the values of p, q are as follows : When G =
SU(d, 1), p = 2(d - 1) and q = 1 ; When G = Sp(d, 1), p = 4(d - 1) and
q= 3.When G=F,(-2o),p=8andq=7.
A final application of these methods which may be worth mentioning is the
following. Let x, y e G, and let for any r > 0, Q(x, y, r) be the number of elements 7 e I', such that a(y-'rx) < r. Q(x, y, r) is the number of points k on G/K which lie in a ball of radius r around the point yK. The computation of 1, alluded to above enables us to find the asymptotic behaviour of Q(x, y, r) as r--->oo ; (cf. [1]). Briefly, the method is as follows : Since fs is admissible, ,Er fs(xry 1) converges nicely and can be expanded as a series Z-E6(G,1) Z$=1 fs(2) (x) (y), where +;m, 1 < i < nr((o, 1), are
eigenfunctions of P. in L2(r\G/K), corresponding to the eigenvalue Q.. Now Z, f s(y-'rx) _ Z, (cosh a(y-'rx)/c)-s, with c = 8q as before, which
can be viewed as a Dirichlet series, convergent if Re s > p + 2q. On the right side, the computation of 1, allows one to conclude that this Dirichlet series has a single simple pole at s = p + 2q whose residue can be computed. Applying the theorem of Wiener-Ikehara one gets (4.34)
Q(x,Y, r)
2.7r (p+q+1)/2
e21P1,
r(2 (p + q + 1)).21pI Vol. (r\G)
2P +2q
as r - oo
.
LENGTH SPECTRA OF MANIFOLDS
423
We leave the details to the reader. A result analogous to Theorem 4.4 has been announced by Margulis [13]. See also Sinai [20]. These authors use ergodic theory. Margulis' result is the stronger one. His context is that of an arbitrary compact manifold of negative curvature, and he shows that Q0(l) - Cl-1 exp dl, for some positive d. In our special situation, we have been able to relate this constant d to the structure
of the manifold. Margulis' proofs have not appeared, as far as the author knows.
Added in proof. After this paper went to press, D. Hejhal pointed out to me that the proof of Theorem 4.4, as well as of the analogous theorem in McKean's paper, is based on an incorrect application of Karamata's theorem. However, the conclusion of the theorem is correct. There are several ways of filling the gap. One is to use Huber's method as indicated above, exploiting (4.33). The other is to use the heat kernel in the trace formula, and to study the behaviour of that formula for complex t in a sector. The third, and the most satisfactory, method is to study the Dirichlet series Z 1(6) exp -sl(d), s e C. By using the analytic properties of the Selberg zeta function (See R. Gangolli, Ill. J. Math. 21 (1977) 1-41), one can show that this series is
meromorphic in Re (s) > 2 1 p - e for some e > 0, and has a single simple pole at s = 2 1 p I with residue 2 p 1. Now Wiener-Ikehara's theorem yields Theorem 4.4. (This method is described for noncompact G/I' in a forthcoming paper of G. Warner and the author.) For yet another method, and a better result, see D. DeGeorge, Ann. Sci. Ecole Norm. Sup. 10(1977) 133-153. Bibliography L. Berard-Bergery, Laplacien et geodesiques fermees sur les formes despace hyperbolique compactes, Seminaire Bourbaki, 24ieme Annee, 1971-72, Exp. 406.
R. L. Bishop & B. O'Neill, Manifolds of negative curvature, Trans. Amer. Soc. 145 (1969) 1-48. T. Eaton, Thesis, University of Washington, Seattle, 1972. R. Gangolli, Asymptotic behaviour of spectra of compact quotients of certain symmetric spaces, Acta Math. 121 (1968) 151-192. , Spherical functions on semisimple Lie groups, Geometry and analysis on symmetric spaces, Marcel Dekker, New York, 1972. , Spectra of discrete uniform subgroups, Geometry and analysis on symmetric spaces, Marcel Dekker, New York, 1972.
R. Gangolli & G. Warner, On Selberg's trace formula, J. Math. Soc. Japan 27 (1975) 328-343. Harish-Chandra, Spherical functions on semisimple Lie groups. I, II, Amer.. J. Math. 80 (1958) 241-310, 533-613. -, Invariant eigendistributions on a semisimple Lie group, Trans. Amer. Math. Soc. 119 (1965) 457-508. Discrete series for semisimple Lie groups. II, Acta Math. 116 (1966) 1-111.
S. Helgason, Differential geometry and symmetric spaces, Academic Press, New York, 1962. H. Huber, Zur analytischen theorie hyperbolisher raumformen and bewegungsgruppen. I, Math. Ann. 138 (1959) 1-26.
424 1131
:[141
[15] [161
[17] 1181
1191
[20]
,[21]
RAMESH GANGOLLI
G. A. Margulis, Certain applications of ergodic theory to the investigation of manifolds of negative curvature, J. Functional Anal.i Prilozen 3 (1969) 89-90, (Russian). H. P. McKean, Selberg's trace formula as applied to a compact Rienzann surface,
Comm. Pure Appl. Math. 25 (1972) 225-246. S. Minakshisundaram & A. Pleijel, Some properties of the eigenfunctions of the Laplace operator on Riemannian manifolds, Canad. J. Math. 1 (1949) 242-256. V. Ozols, On the critical points of the displacement function of an isometry, J. Differential Geometry 3 (1969) 411-432. A. Preismann, Quelques proprietes globales des espaces de Riemann, Comment. Math. Helv. 15 (1943) 175-216. A. Selberg, Harmonic analysis and discontinuous subgroups in weekly symmetric Riemannian spaces with applications to Dirichlet series, J. Indian Math. Soc.
20 (1956) 47-87. Seminaire Berger, Varietes a courbures negative, Universite Paris VII, 1970/71.
Y. Sinai, The asymptotic behaviour of the number of closed geodesics on a
compact manifold of negative curvature, Izv. Akad. Nauk. SSSR Ser. Mat. 30 (1966) 1275-1296. P. Trombi & V. S. Varadarajan, Spherical transforms on semisimple Lie groups,
Ann. of Math. 94 (1971) 246-303.
[22]
[23]
N. Wallach, An asymptotic formula of Gelfand and Gangolli for the spectrum of I'\G, J. Differential Geometry 11 (1976) 91-101. G. Warner, Harmonic analysis on semisinzple Lie groups. I, II, Springer, Berlin, 1972.
[24]
H. Weyl, Das asymptotische Verteilungsgesetz der eigenschwingungen eines beliebig gestalteten elastischen korpers, Rend. Circ. Mat. Palermo 39 (1915)
[25]
D. Widder, The Laplace transform, Princeton University Press, Princeton, 1941. N. Wiener, Tauberian theorems, Ann. of Math. 33 (1932) 1-100. J. A. Wolf, Spaces of constant curvature, McGraw-Hill, New York, 1967.
1-50.
1261
127]
UNIVERSITY OF WASHINGTON, SEATTLE
J. DIFFERENTIAL GEOMETRY 12 (1977) 425-434
GLOBAL PROPERTIES OF SPHERICAL CURVES JOEL L. WEINER
Let a be a closed curve regularly embedded in Euclidean three-space satisfying suitable differentiability conditions. In addition, suppose a is nonsingular, i.e., free of multiple points. In 1968, B. Segre [4] proved the following about such curves. Theorem. If a is nonsingular and lies on a sphere, and 0 denotes any point of the convex hull of a with the condition that 0 (if lying on a) is not a vertex of a, then there are always at least four points of a whose osculating plane at
each of those points passes through 0. If 0 is a vertex of a then there are at least three points of a whose osculating plane at each of those points passes through 0. All terms used in the statement of the theorem are defined later in this paper. To quote H. W. Guggenheimer [2] who reviewed [4], "The 12-page proof
is rather complicated." Here we present a shorter and hopefully more transparent proof of this theorem. In addition, we need only require that the spherical curve a be of class C2 whereas Segre's proof requires a be of class C3. Also, we obtain, with no extra effort, a similar theorem which holds if a's only singularity is one double point ; in this case, the above mentioned minimums must be reduced by two. In the last section of this paper we characterize spherical curves with the following property : for every point 0 of the convex hull of a, other than a. vertex of a, there exists the same (necessarily even) number of distinct points of a whose osculating plane at each of those points passes through 0. The proofs of many results in this paper ultimately depend on ideas contained in a paper by W. Fenchel [1]. Throughout this paper we use the following conventions. By a curve we mean a regular C2 function a : D , E3, where D is an interval (with or without. end points) or a circle, and E3 is Euclidean three-space. We let a denote both the function and its configuration a(D) in E3. When D is a circle we say a is closed. If D is a closed interval we may sometimes refer to a as an arc. We say a point P in E3 is a multiple point of a if it is the image of k > 1 points of D. If k = 2 then P is called a double point. At a multiple point P we will think of P as k distinct points each traversed once by a as we traverse D once. If a has no multiple points, then we say a is nonsingular. Received July 16, 1975. This research was supported by NSF Grant GP 43030.
426
JOEL L. WEINER
1.
Geodesic curvature
Let a be an oriented spherical curve; i.e., a lies on a sphere S in E3 and has a preferred direction of traversal. Let S be oriented, say, with respect to the outward pointing normal. We denote by k the geodesic curvature of a as a curve in S. It is defined by k = (d2a/ds2) n, where s is the arc length parameter of a consistent with its orientation, and n is da/ds rotated +90° in the tangent plane to S at its point of contact with S. Since a is C2, k is a continuous function on a. At each point P of a there is in S a circle tangent to a which best approximates a near P. This circle w(P) is the osculating circle to a at P ; it is easy to see that w(P) is the intersection of the sphere S and the osculating plane ir(P) to a at P, when a is viewed as a curve in P. We have the following obvious lemma. Lemma 1. Let a be a spherical curve and P E a. Then k(P) = 0 if and only if 7r(P) goes through the center of S. We will need some lemmas about spherical curves proved by Fenchel [1]. Actually we state mild generalizations of these lemmas; see [1], [5] for their proofs. In these lemmas we speak of a set on the sphere being to the left of a curve. By this we mean that when the tangent vector to the curve in the preferred direction is rotated x-90° it points into the set. Also when we say a point P is between points A and B we mean that either A and B are antipodes or if A and B are not antipodes then P lies on the shorter geodesic arc through A and B. Lemma 2. A nonsingular spherical curve a with k > 0 and not identically zero connects two points A and B of a great circle r without otherwise meeting it. Then A and B are not antipodes of one another. In addition the region bounded by the curve and the smaller great circular arc AB of r and lying in a hemisphere is to the curve's left. Lemma 3. Let a be a nonsingular spherical curve with k > 0, and let r be an arbitrary great circle which meets a in at least two points. Then there is a subarc a, of a with the following characteristics: 1. The end points A and B of a, lie on r. 2. a, has otherwise no points in common with r. 3. All other points of intersection of a with r lie between A and B. Remark. If a, contains a point P for which k(P) > 0, then A and B are not antipodal by Lemma 2. In particular, more than a half circle of r is free of points of intersection with a. 2.
Fenchel's theorem
The convex hull of a point set M in Euclidean space is the smallest convex set containing M. Let Q be the convex hull of a spherical curve a. The next lemma characterizes the points of Q ; for its proof see [1, Satz All.
PROPERTIES OF SPHERICAL CURVES
427
Lemma 4. For 0 to be an element of 2 it is necessary and sufficient that there exists a plane 2 through 0 such that 0 is in the convex hull of a fl ,. Throughout this section we take 0 to be the center of the sphere S on which a lies. With this choice for 0, Lemmas 3 and 4 lead immediately to a theorem due to Fenchel [1, Satz II']. This theorem is restated to include the possibility that 0 is an element of the boundary of 2 as well as the interior of D. Theorem 1 (Fenchel). Suppose a is closed and nonsingular except perhaps for one double point. If 0 E 2, and a does not contain a great semicircular arc, then the geodesic curvature of a changes sign at least twice. The same lemmas can be used to prove the following extension of Theorem 1. This will be shown here. Theorem 2. Suppose a is closed and nonsingular. If 0 E Sl, and a does not contain a great semicircular arc, then the geodesic curvature of a changes sign at least four times. Remark. It is easy to construct examples of closed nonsingular spherical curves whose geodesic curvature changes sign only twice and which necessarily contain a great semicircular arc. It is a consequence of Lemma 2 that these curves lie in a hemisphere determined by the great semicircular arc. The remainder of this section is devoted to a proof of Theorem 2. Before we proceed we introduce some notation. If a is a non-closed spherical curve, and P, Q are two points of a, then by PaQ we mean the oriented arc running along a from P to Q. If P, Q are two points of the sphere S which are not antipodal, then PQ denotes the smaller great circular arc through P and Q oriented from P towards Q. To denote the larger great circular arc connecting P and Q, we write PAQ where A is on the great circle through P and Q but A PQ. By a Jordan curve we mean a nonsingular continuous image of a circle. Proof of Theorem 2.
Let a be a closed nonsingular curve lying on a sphere
S with center 0, and suppose that a contains no great semicircular arc. In particular, a's geodesic curvature k is not identically zero. Also suppose 0 E 2, the convex hull of a. By Theorem 1 we already know that k changes sign at least twice. We will show that the supposition that k changes sign only twice leads to a contradiction. Therefore suppose k changes sign twice at the points
A and B of a. Let a' and a' be the two curves into which a is separated by A and B, both oriented so that their geodesic curvature is nonnegative (and, of course, not identically zero). Suppose a' and a2 begin at A and end at B. By Lemma 2 there is a plane 2 through 0 such that 0 is in the convex hull
of 2 fl a. Let r = 2 fl s; it is, of course, a great circle. There are two cases to consider. Either 1. a meets r in at least three points and these points do not lie in an open half circle of r, or 2. a meets r in two points, which are necessarily antipodal.
428
JOEL L. WEINER
Case 1. Let C, D, E be distinct points at which a = a' U a2 meets r and which do not lie in an open half circle of r. We may suppose that C and D are points of a'; in fact, suppose C precedes D in a'. Since a' meets r in at least two points, Lemma 3 implies that there exists a subarc a; with the characteristics 1, 2, and 3 of that lemma. Also a7 is not a great semicircular arc. The remark following Lemma 3 implies that E must be a point of a2. We may assume that C and D are the end points of a'; if the new C, D, E lie in an open half circle of r so do the old C, D, E. Let H be the closed hemisphere determined by r and not containing a; ex-
cept for the end r pints C and D. Let L be the region to the left of the oriented Jordan curve' Ca'D U DC together with its boundary. Lemma 3 implies that a' C H U L. In particular A, B E H U L ; hence a2 must begin and end in H U L. The boundary of H U L is the Jordan curve a7 U DEC. Now if a2 is not contained in H U L, it must cross the boundary along DEC (excluding
the end points D and Q. Remember that a' and a2 meet only at A and B. We assume without loss of generality that a2 crosses DEC. If a2 did not cross
DEC, then it would be tangent to r at E. We could then rotate 2 a bit about the diameter of S through C or D so that a crosses r at points which we still call C, D, E and which still do not lie in an open half circle of r. Since a2 meets r at least twice, Lemma 3 implies the existence of a subarc c:1. Let a' begin at F and terminate at G. Characteristic 3 of a; implies that at least one of the points F and G is not between C and D. At this stage of the argument we suppose that F does not lie between C and D. The argument is similar if we suppose that G does not lie between C and D. Consider the oriented Jordan curve Aa1D U DF U Fa2A. If D and F are antipodal, then here DF is the half great circle not containing G.; see Fig. 1.
Fig.
1
Note that Da'B and Fa2B cannot cross the Jordan curve. That F012B does not cross DF is the only part of the preceding statement which may not be im-
PROPERTIES OF SPHERICAL CURVES
429
mediately clear. However Fa2B may only cross r along FG which is less than a half circle ; also DF is at most a half circle. Thus DF meets FG only at F. Thus Fa2B meets DF only at F. Now Da1B and Fa2B are on opposite sides of the Jordan curve near D and F, respectively. This is clear since a' is entering
H at D and al is leaving H at F. Thus B is both to the right and the left of the Jordan curve, which is a contradiction. Case 2. Let C and D be the two points in which a meets T. As already noted C and D are necessarily antipodal. This case can be reduced to Case 1 since there must be a great circle through C and D which intersects a at a third point E. Clearly C, D, E do not lie in an open half circle.
Remark. We do not use the fact that a' and a' join at A and B in a CZ fashion, but only that they begin and end at A and B, respectively. 3.
Segre's theorem
Generally, if P is a point of a curve a then at P a passes through the osculating plane to a at P. However if this does not happen we call P a vertex of a. Thus by a vertex of a curve a we mean a point P of a with the property that near P a lies on one side of the osculating plane to a at P. Theorem 3. Let a be a closed curve on the sphere S and let 0 E SQ, a's convex hull. Then (i) if a is nonsingular and 0 is not a vertex of a, there exist at least four points of a whose osculating plane at each of those points passes through 0,
(ii) if a is nonsingular and 0 is a vertex of a, there exist at least three points of a whose osculating plane at each of those points passes through 0, (iii) if a's only singularity is one double point and 0 is not a vertex of a,
there exist at least two points of a whose osculating plane at each of those points passes through 0. The idea behind the proof lies in the observation that Theorem 3 follows trivially from Theorems 1 and 2 by means of Lemma 1 if 0 is the center of S. So if 0 is not the center of S we let a* be the projection of a into a sphere X centered at 0 and apply Theorems 1 and 2 to a* to get the required num-
ber of points of a* whose osculating plane at each of those points passes through 0. If 0 E a, then a* is not a closed curve but one can still show that a* has the required number of points whose osculating plane at each of those points passes through 0. Finally we observe by Lemma 5 that an osculating plane at a point of a* passes through 0 if and only if the osculating plane at the corresponding point of a does so. We now introduce the notation which will be used in the proofs of Lemma 5 and Theorem 3. Let a be a closed curve on S, and SQ the convex hull of a. Suppose that 0 is any element of D and X is a sphere centered at 0. Let p : S --> X be the projection of S into X through 0. When 0 E a, p is understood to be defined only on S - {0}. Denote the image of P E S under p : S - > .Z by P*.
430
JOEL L. WEINER
If 0 is in interior of S, we let a* denote the image of a under p. If 0 e a, note first that p(a) is contained in a hemisphere H with boundary r*, where r* is the intersection of the tangent plane to S at 0 with 1. Assume 0 is not a multiple point of a ; then the limits of P* as P approaches 0 along a first from one side and then the other are two antipodal points on r*. We adjoin these points to p(a) and denote the resulting arc by a*. When 0 is a multiple point of a, we adjoin points of r* to p(a) as above to get a collection of arcs denoted by a*. Then let Q* be the convex hull of a*. Let ir(P) and 7r*(P*) denote the osculating planes to a at P and a* at P*, respectively. Lemma 5. Suppose P zt- 0. Then ir(P) passes through 0 if and only if ft*(P*) goes through 0. Moreover, if rr(P) passes through 0, then P is a vertex of a if and only if P*is a vertex of a*. Proof. The projection p : S - is a C°° difeomorphism of S onto its image. Thus the order of contact between two curves on S and their images under p on _Y is preserved (except if the contact is at 0 E a). Let w(P) and w*(P*) denote the osculating circles to a at P and a* at P*, respectively. Suppose 7r(P) passes through 0. Since w(P) lies in ir(P) which passes through 0, its image under p is a (great) circle on _Y if 0 a and is a half (great) circle on _Y if 0 E a. Let w(P)* denote the circle in which p((O(P)) lies on J. Since the order of contact is preserved, w(P)* _ w*(P*). Thus both ir(P) and it*(P*) contain w(P)*. Hence 7r(P) = 2r*(P*) passes through 0. The converse is proved in an identifical fashion. Now suppose ir(P) passes through 0. Then, by the above, ir(P) = 7r*(P*). If a lies on one side of ir(P) near P, clearly a* lies on one side of it*(P*) near P* and conversely. That is, P is a vertex of a if and only if P* is a vertex of a*. Proof of Theorem 3. We separate the proof into two cases according as 0 e a or not. Suppose 0 a. Then it is clear that 0 E Q* since 0 e Q. Thus we may apply Theorems 1 and 2 to a* lying on 1. If a is nonsingular, so is a* ; thus a* has at least four points where its geodesic curvature is zero. If a has just one double point, so does a* ; thus a* has at least two points where its geodesic curvature is zero. By Lemma 1, at each of these points of a* the osculating plane passes through 0. Hence by Lemma 5 the osculating planes at the corresponding points of a pass through 0. Thus we have proved (i) and (iii) for the case 0 a. Suppose 0 e a and 0 is not a multiple point of a. Assume now a is oriented. By means of p we orient a*. Denote the beginning of a* by A and the end by B. Let w be the osculating circle to a at 0. Its image under p including end points, denoted by w*, is a half great circular arc of 1. It is easy to see that uI* also begins at A and ends at B. Also w* and a* are tangent at A and B. If 0 is not a vertex of a, then a* is on opposite sides of m* in H near A and B ; see Fig. 2. If 0 is a vertex of a, then a* is on the same side of m* in H
PROPERTIES OF SPHERICAL CURVES
Fig.
431
2
near A and B. Let k* be the geodesic curvature of a*. Then using Lemma 2 and the idea of parity, one can show the following hold : 1. k* changes sign at least twice if 0 is not a vertex of a and a is nonsingular, 2. k* changes sign at least twice if 0 is a vertex of a and a is nonsingular,
k* changes sign at least once if 0 is not a vertex of a and a's only singularity is one double point. Again apply Lemmas 1 and 5, in that order, to prove (i), (ii), and (iii) for the case where 0 e a and 0 not a multiple point of a. If 0 is the double point of a the proof of (iii) is immediate. Corollary. Let a be a C3 closed nonplanar curve in E3 with no pair.of directly parallel tangents. Then a has at least four vertices. For the proof of this corollary see Segre [4, p. 263] where the same result is proven for C4 curves. Our results allow his proof to go through for C3 curves. Actually the corollary follows immediately from Theorem 2 and the remark following Theoerm 2 since the tangent indicatrix of a nonplanar curve cannot lie in a hemisphere. 3.
4. A characterization In this section we find a characterization for a (possibly singular) closed curve a lying on the sphere S and having the property that for each point 0 in its convex hull Q except for vertices of a there exists the same (necessarily even) number of distinct points of a whose osculating plane at each of those points passes through 0. The next lemma is especially important in this section. It follows by means of stereographic projection from a similar fact for plane curves due to Kneser ; see [3, p. 48] for Kneser's theorem and its proof. When we say that the circle cu lies between the (disjoint) circles m' and cue on the sphere S we mean that w is in the connected component of S - ((u' U cue) whose boundary is a' U cue. Lemma 6. Let a be spherical arc with monotone geodesic curvature k. Let
P, Q, and R be three points of a with Q between P and R. Then w(Q) is be-
432
JOEL L. WEINER
tween co(P) and co(R) if it is not equal to co(P) or c,(R). Moreover, w(Q) = c,(P) (respectively, co(R)) only if k(Q) = k(P) (respectively, k(R)). At this point we make some additional assumptions about the closed spher-
ical curve a which will hold throughout the remainder of this section. First, we require that there exists at most a finite number of points of a at which the geodesic curvature k takes on an extreme value. This is equivalent to requiring that a has at most a finite number of vertices since the vertices of a occur at the extremes of k. Secondly, we assume k is strictly monotone between the vertices of a. This second condition rules out the possibility of a having an arc of points with the same osculating plane. Let B denote the closed ball whose boundary S contains the closed curve a. Clearly Q C B. Theorem 4. Suppose a has n vertices. If 0 E B, then there exist at most n points of a whose osculating plane at each of those points passes through 0. Proof. Let V1, V2, , V, denote the vertices of a as they occur in making one circuit of a. Using the notation of § 2, we set ai = ViaVi+1 for i = 1, 2, , n, where V,t+1 = V1. We will show for each integer i, where 1 < i < n, there exists at most one point P E ai such that 0 E 7r(P). This immediately implies the theorem. Suppose, to the contrary, that ai contains two points P and Q such that 0 E 7r(P) fl 7r(Q). In particular, 7r(P) fl 7r(Q) # 0; hence co(P) fl w(Q) * 0. This is impossible by Lemma 6 since k is strictly monotone on ai. Remark. Note that vi E ai-1 fl ai for i = 1, 2, , n, where a° = an. Hence if 0 E B and, in addition, 0 E 7r(Vi), then there exist strictly less than n points of a whose osculating plane at each of those points passes through 0. Corollary. Suppose a has n vertices. If 0 E Q, then there exist at most n points of a whose osculating plane at each of those points passes through 0. Let V1, V21 , V, be the vertices of a. Note that n is necessarily even since it is the number of extreme points of the geodesic curvature of a. Theorem 5. Suppose (o(Vi) fl a = {Vi} for i = 1, 2, , n. Then for every 0 E Q - {V1, V2i , Vn} there exist exactly n points P11 P21 ., Pn of a such that 0 E 7r(Pi) for i = 1, 2, , n, and conversely. Proof. Let B' = B - U2=17r(Vi). Also let B'm be the set of points 0 in B' with the property that there exist exactly m points P11 P21 . . , P,n of a such that 0 E 7r(Pi) for i = 1, 2, . , m. Let Q' = Q - {V11 V21 ... , Vn}. For i = 1, 2, , n, the assumption w(Vi) fl a = {Vi} implies Q fl ;r(Vi) = {Vi}. Thus Q' is a connected subset of B'. The theorem is proved by showing that for any nonnegative integer m, B' is an open and closed subset of B'. This implies Q' c B' for some nonnegative integer m. Then we show m = n. The fact that B'is both open and closed in B' follows in three steps Step 1. B' c interior U,m,Sj B. Let 0 E B' and suppose there exist m points P1, P2, , P,n, of a such that 0 E 7r(Pi) and Pi is not a vertex of a for
433
PROPERTIES OF SPHERICAL CURVES
i = 1, 2,
, m. We will show for each integer i, where 1 < i < m, there
exists a neighborhood Ni of Pi in a with the property that Ui = UPEN, 7r(P) fl B' is an open set of B' containing 0. Moreover, we may assume N1, N2, .. , N are mutually disjoint. It is then clear that u = n Ui is a neighborhood 1
of 0 in U,n,S; B';.
Consider the point Pi. Since Pi is not a vertex there exists an open neighborhood Ni of Pi in a on which k is strictly monotone. By Lemma 6, Ni does not contain Pj, where j i. Let P' and P2' be the boundary points of Ni. It follows from Lemma 6 that UP E w(P) is an open set of S ; it is the component of S - [w(Pi) U w(Pi')] containing Pi. Then Ui = UPE,, 7r(P) f1 B' is an open set of B'. In fact Ui is the component of B' - [7r(P') U 7r(Pz')] containing Pi. Clearly 0 E Ui since Pi E Ni. Step 2. B'm is closed in B'. Let Oi, i = 1, 2, , be a sequence of points in Bm approaching 0 E B'. Thus for each i = 1, 2, , there exist exactly m , m. By taking points Pil, Pie, , Pig, of a such that Oi E 7r(Pi;) for j = 1, 2, subsequences if necessary, we may assume that Pi; approaches a point P5 as i
approaches infinity for j = 1, 2, , m. By continuity 0 E 7r(P;) for j = 1, 2, , m. Thus there are at least m points of a whose osculating plane at each of those points passes through 0 unless P; = Pk for some j k. Suppose this ; then in any neighborhood of P; = Pk there exist the distinct points PiJ. Pik, for i sufficiently large. Since Oi e 7r(Pi,) f1 7r(Pik), w(Pi,) fl w(Pik) 0. By
Lemma 6, P; = Pk is a vertex of a. But this contradicts the assumption 0
Ui=17r(Vi). Thus P;
Pk for all j
k between 1 and m inclusive. By Step
1 there exist at most m points P P2, , P of a with 0 E 7r(P;). Step 3. Bm is open in B'. This step follows immediately from Step 1 and Step 2 since B'm = 0 for m > n by Theorem 4. We now know that Q' c B' where m < n. Suppose m < n. We will show this leads to a contradiction. Let o E a n d.'. Since 0 E Q', there exist m points Pl, P2, , m. In the notation of the proof , P with 0 E ir(Pi) for i = 1, 2, of Theorem 4, there exists an are ai for some integer between 1 and n inclusive with the following property : there exists no point Q E ai such that 0 E 7r(Q).
Thus w(Vi) and w(Vi+,) do not have 0 between them. Hence, say, w(Vi) and 0 are separated by w(V2}). In particular Vi and 0 are on opposite sides of w(Vi+). Thus a must meet w(Vi+) at points other than Vi,,. The converse follows from the remark following the proof of Theorem 4. q.e.d. It may still be that for every point 0 of Q' there exists the same number of points of a whose osculating plane at each of those points passes through 0
even though w(Vi) n a {Vi} for some integer i, 1 < i < n. For this to happen the following must be true : if, say, V1 is a vertex of a and w(V) intersects a in more than V1, then there must be another vertex Vi for some integer i, 2 < i < n, such that ir(Vi) = 7r(V1). Also, for points P near V, and Q near Vi, ir(P) and 7r(Q) must be on opposite sides of rr(V,) = 7r(Vi).
434
JOEL L. WEINER References
W. Fenchel, Uber Krummung and Windung geschlossener Raumkurven, Math. Ann. 101 (1929) 238-252. [2] H. W. Guggenheimer, Rev. #4787, Math. Rev. 39 (1970) 871. [ 3 ] -, Differential geometry, McGraw-Hill, New York, 1963. [ 4 ] B. Segre, Alcune proprietd differenziali in grande delle curve chiuse sghembe, [1]
Rend. Mat. (6) 1 (1968) 237-297.
[5 ] J. L. Weiner, A theorem on closed space curves, Rend. Mat. (3) 8 (1975) 789-804. UNIVERSITY OF HAWAII
J. DIFFERENTIAL GEOMETRY 12 (1977) 435-441
THE DIMENSION OF BASIC SETS JOHN M. FRANKS
Let f : M -p M be a C' diffeomorphism of a compact connected manifold M. A closed f-invariant set A C M is said to be hyperbolic if the tangent bundle of M restricted to A is the Whitney sum of two Df-invariant bundles, i.e., if TM = Eu(A) +O E3(A), and if there are constants C > 0 and 0 < 2 < 1 such that lD fn(V) l < CA" I v I I D f -n(V) I < CA" I V I
for v E Es, n > 0, for v E E", n > 0
The diffeomorphism f is said to satisfy Axiom A if (a) the non-wandering set
Q(f) = {x E M : U n U..>0 f "(U) # 0 for every neighborhood U of x} of f is a hyperbolic set, and (b) Q(f) equals the closure of the set of periodic points of f. If f satisfies Axiom A, one has the spectral decomposition theorem of U Al where Ai are pairwise disjoint, Smale [9] which says Q(f) = A, U f-invariant closed sets and f J,, is topologically transitive.
These Ai are called the basic sets of f, and it is the object of this article to investigate restrictions on their dimensions imposed by the homotopy type of f and the fiber dimensions of the bundles E8 and E. In [11] S. Smale showed that any diffeomorphism can be isotoped to a diffeomorphism satisfying Axiom A with all basic sets of dimension zero. This disproved earlier conjectures that some homotopy classes might contain only diffeomorphisms with a basic set of positive dimension. Theorem 1 below shows that if one restricts either the fiber dimensions of the bundles E" or the total number of basic sets for f, then there are indeed homotopy classes all of whose diffeomorphisms (subject to these restrictions) have basic sets of positive dimension. In Theorem 2 we investigate diffeomrphisms with a single infinite basic set, the others being isolated periodic orbits. It is a pleasure to acknowledge valuable conversations with R. F. Williams. We consider diffeomorphisms which in addition to Axiom A satisfy the no-
cycle property [10] which we now define. If Ai is a basic set of f then its stable and unstable manifolds ([5] or [9]) are defined by W3(Ai) = {x E M I d(f n(x), Ai) - 0 as n - oo} Communicated by R. Bott, July 11, 1975. This research was supported in part by NSF Grant GP42329X.
436
JOHN M. FRANKS
Wu(Ai) = {x E M d(f -'i(x), Ai) - 0 as
n,
co } .
One says Ai G A; if Wu(A;) fl WI(Ai) zf- 0. If this extends to a total ordering on the basic sets Ai, then f is said to satisfy the no-cycle property and we re-index so that Ai G A; when i < j. If Ai is a basic set of f : M , M then we define the index ui of Ai with respect to f to be the fiber dimension of Eu(Ai). All homology and cohomology will be singular with real coefficients unless otherwise stated. Theorem 1. If f : M , M satisfies Axiom A and the no-cycle property and Hk(M) zf- 0, then there is a basic set Ai satisfying dim Ai > Ik - u,,I where ui is the index of A. Hence, if f has fewer basic sets than nonzero cohomology groups, it must have a basic set of positive dimension, or equivalently : Corollary 1. If f has only basic sets of dimension zero, then there is a basic set Ai with index u1 = k for each k such that Hk(M) zf- 0. Theorem 2. Suppose f : M , M satisfies Axiom A and the no-cycle property and has one infinite basic set A, the others being isolated periodic orbits. If f*: Hk(M) --> Hk(M) has an eigenvalue which is not a root of unity, then dim A > I n - 2k I where n = dim M. It A is an attractor, then dim A > max {(n - k), k}. We note that M. Shub [8] has shown that whenever f * : H*(M) , H*(M) has an eigenvalue which is not a root of unity, then f must have at least one infinite basic set. In case M is the n-dimensional torus Tn we can strengthen Theorem 2 because either f * : HI(T11) --> HI(T11) has an eigenvalue which is not a root of unity or f* : H*(Tn) --> H* (T11) is quasi-unipotent (i.e., has only roots of unity as eigenvalues). Corollary 2. If f*: Tn , Tn satisfies Axiom A and the no-cycle property and has only one basic set A which is infinite, then either f*: H*(Tn) --> H*(Tn)
is quasi-unipotent or dim A > n - 2. It is not difficult to construct diffeomorphisms on Tn with a single infinite basic set of dimension n, n - 1, but the author does not know if there is a diffeomorphism of T3 which is not unipotent on homology and with a single infinite basic set of dimension one (dimensions 2 and 3 can be realized in this case). The hypothesis that f* not be quasi-unipotent on cohomology is necessary since it is easy to construct f : Tn , Tn homotopic to the identity with a single infinite basic set of dimension zero. We review briefly the filtrations of [10] associated with a diffeomorphism which satisfies Axiom A and the no-cycle property. It is possible to find submanifolds (with boundary and of the same dimension as M),
M=MZD ... DM,DMo=0, such that
437
DIMENSION OF BASIC SETS
Mi
Ai =
/U f(Mi) c int M1 , I
I
mEZ
f'(M1 - Mi_1)
Wu(Ai) U Mi_1 = Mi-1 U n fm(Mi) mZ0
Henceforth f : M -p M will be a diffeomorphism of a compact manifold D satisfying Axiom A and the no-cycle property and M = Mz D M1_, D M, = 0 will be a filtration for f. The proofs of Theorems 1 and 2 use the following proposition which may be of some independent interest. Proposition 1. Suppose f : M -p M satisfies Axiom A and the no-cyclic property and Ai C Mi - Mi_1 is a basic set of f. Let S = {k I fk Hk(Mi, Mi_1) Hk(Mi, Mi_1) has a nonzero eigenvalue}. Then dim Ai > max S - min S. We procede now with a sequence of lemmas leading to the proofs of the results above. We will use closed local stable and unstable manifolds of a point x E A, denoted Ws (x) and Wu(x) (see [5] or [9]). Since it is not in general true that dim (X X Y) = dim X + dim Y it is necessary to use the concept of cohomological dimension over R [3] defined as :
follows : If X is a compact Housdorff space, then dim, X = sup {k I Hk(X, A ; R) # 0} where A runs over all closed subspaces of X and ft' is Cech cohomology
with real coefficients. By a result of [7, p. 152] dim, X < dim X. Lemma 1. Suppose Ai C Mi - Mi_1 is a basic set for f and Mi, Mi_1 are
the elements of a filtration for f. If k > dim, Ws (Ai), then the map f*: Hk(Mi, Mi_1) - Hk(Mi, Mi_1) is nilpotent. Proof. This is essentially the same as [4, Lemma 6] which drew heavily on [1]. Let X = Wu(Ai) U Mi_1 and let Ilk denote Cech cohomology with real coefficients. We use the closed local unstable manifolds of [5]. The inclusion (Wu(Ai), aWu(Ai)) - (X, W) is a relative homeomorphism where W =
cl(X - Wu(Ai)). Hence by a standard result [12, p. 266], HN(WI. (A), aWu (Ai)) = Hk(X, W)
By definition of dim,, Hk(WE (Ai), aWu(Ai)) = 0 ,
when k > dim, Wu(Ai). Since W is compact and X C {f n,o f n(int Mi_1)} U Ai it follows that f -(W) C Mi_1 for some m > 0. The diagram
(X, Mi-0 - (X, W)
f'
If.
(X, Mi-1)
438
JOHN M. FRANKS
commutes. Thus the map (f-)*: Hk(X, Mi_1) -* IIk(X, Mi_1) factors through Hk(X, W) so that (f'")* = (f*)- = 0 when k > dimR Wu(Ai). Now if f*: Hk(Mi, Mt_1) -* Hk(Mi, Mi_1) is not nilpotent, there is a subspace V # 0 with f *(V) = V. By [1, Lemma 1], the map h* is one-to-one on V where h*: Hk(Mi, Mi_1) = Hk(M1, Mi_1) -* Hk(X, Mi_1) is induced by the inclusion h : (X, Mt_1) -* (Mi, Mi_1). Thus we have a commutative diagram Hk(Mi, Mi-1) ('*) Hk(Mi, Mi-1) h*
H k(X,/Mi-1)
h* IHk(Xy.,
Mi-1)
But, (f*)mh*(V) = h*(f*)'nV = h*(V) 0, which is a contradiction if k > dimR Wu(Ai), since (f*): Hk(X, MT_1) -* E (X, M7_1) is zero in this case. Thus it must be the case that f*: Hk(Mi, Mi_1) -* Hk(Mi, Mi_1) is nilpotent when k > dimR Wu(Ai). q.e.d. If A is a basic set and x e A, we let Ws(x) = Ws(x) n A and WE (x) = WE (x) fl A. While it is true [9] that x e A has a neighborhood homeomorphic to WE(x) x WE (x), it appears to be an open question whether or not dim A = dim WE(x) + dim WE (x). For the cohomological dimension over R however we have the following. Lemma 2. Suppose A is a basic set for f, u = fiber dim Eu(A), and s = fiber dim ES(A). Then (a) dimR WE (A) = dimR WE(x) + u, (b) dimR W1,(A) = dimR WE (x) + s, (c) dimR A = dimR Ws (x) + dimR WE(x), where x is any point of A and e > 0 is sufficiently small.
Proof. We will use the following results from [13, Theorem 2.2 and Lemma 2.1]. If X and Y are compact Hausdorff spaces, then (1) dimR (X X Y) = dimR X + dimR Y, and (2) if n = dimR X, there exists a point p e X such that if U is any sufficiently small neighborhood of p in X, then Hn(X, X - U) # 0. Also if Y is a compact subset of X, then consideration of the exact sequence of the triple (X, Y, A), where A is a closed subset of Y,
Hn(X A) > IIn(Y A)
Hn+i(X, Y)
,
shows that dimR X > dimR Y.
We begin the proof of (a) by showing that dimR W(x) is independent of x E A. If y e A, then using the canonical coordinates [9, p. 781] for A and the fact that WS(orb (y)) is dense in A it is easy to show that WE(x) is homeomorphic to a compact subset of f-(WE(y)) for some m. This implies WE(x) is homeomorphic to a subset of WE(y) since f- is a diffeomorphism. Thus dimR WE(x) < dimR Wa(y) and the same argument shows dimR WW(y) < dimR WE(x).
DIMENSION OF BASIC SETS
439
By results of [6] there is a continuous map cp : A -* Emb (D, M) such that cp(z)(D) = Wa (z) where D is the disk of dimension u. The map i : Wa (x) X D -* Wa (A) given by J(y, t) = o(y)(t) is a homeomorphism onto a compact neighborhood K, of x in Wsu(A). But it is not possible that dimR Wsu(A):> dim K,, because the sets K,, cover Wu(A) and by (2) above together with excision at least one of them must have dimension over R equal to that of Wa (A). Thus dimR WE (A) = dimR Ws(X) + u for all x e A and (a) is proven. Applying this result to ;-1 proves (b). To prove (c) we consider the canonical coordinate map p : Ws (x) X Ws (x) A which is a homeomorphism onto a compact neighborhood J, of x in A. By (1) above dimR J, = dimR Ws (x) + dimR Ws (x). Since J. c A, dimR J,x G dimR A and again using (2) above and excision, it follows that dimR A = dimR Jx
for some x (and hence for all x since dimR Ws (x) and dimR Wa (x) are independent of x). Thus (c) is proven. q.e.d. Lemma 3. If A3 - A2 j - Al is a sequence of vector spaces exact at A2i ai : Ai -* Ai are linear maps commuting with i and j, and 2 is an eigenvalue of a2, then 2 is also an eigenvalue of either a3 or a1.
i
This is [4, Lemma 2] ; the proof is not difficult and will not be repeated here. Lemma 4. If 2 is an eigenvalue of f,* : Hk(M) -* Hk(M), then there is an Mi in the filtration for f such that fk : Hk(Mi, Mi_1) -* Hk(Mi, Mi_1) has .l as an eigenvalue. Proof. Consider the exact cohomology sequence of the triple
Hk(M, M,) - H'(M, M;-1) - Hk(M,, M,-1) There is a map f* induced by f on each of these groups, and these maps commute with the maps of the sequence. We now apply Lemma 1 to this sequence when j = 1. In this case the sequence is
Hk(M, M) _ H1(M) -* Hk(Mi Mo) ,
so either 2 is an eigenvalue of f* on Hk(M1, Mo) or an eigenvalue of f * on Hk(M, M). If the latter we set j = 2 and reapply Lemma 1 to show 2,, is an eigenvalue of f * on either Hk(M2i M) or Hk(M, M2). Continuing this procedure it follows that 2 is an eigenvalue of f* on Hk(Mi, M2_1) for some i, since Hk(M, M) = Hk(M, M) = 0. Proof of Proposition 1. Let k1= maxS. Then by Lemma 1, k1 G dimR WE (Ai) and by Lemma 2, dimR Ws (Ai) = dimR E (x) + u;, where x e Ai and ui =
fiber dim Eu(Ai), so k1 - ui G dimR WE(x). Let k = min S and let M; =
cl(M - M;). Then since f,* : Hk(Mi, M2_1) -* Hk(Mi, Mi_1) has a nonzero
eigenvalue, its adjoint f*,: Hk(Mi, M2_1) -* Hk(Mi, Mi_) has the same eigenvalue. Suppose M is orientable and n = dim M. Then [1, Lemma 4] shows gn_k : Hn-k(Mi 1, Mi) -* Hu k(M;_1, Mi) is similar to either f *k : Hk(Mi, Mi_1) -* Hk(Mi, Mi_1) or to - f *k. In either case gn-k
that if g = f : M -* M, 1
440
JOHN M. FRANKS
has a nonzero eigenvalue. Since g has the same basic sets as f (with Ws(f ; Ai) Wu(g;' Ai)) and M =1110 Ml D ... D11h = 0 is a filtration for g, we can apply to g the argument which showed k1 - ui < dimR W:(x). We have then that (n - k) - fiber dim Eu(g ; Ai) < dimR WE (g ; x) or (n - k) - si < dimR rV (f ; x) where si = fiber dim Es(f ; Ai). Adding this inequality to the one for k1 we have k1
- ui + (n - k) - si < dimR ' (x) + dimR Wu(x)
.
Since n = ui + si, k1 - k < dimR A by Lemma 2. That is, max S - min S < dimR Ai < dim Ai. In case M is not orientable, we let - : M - M be an oriented double cover of M and f : M --+ M a lift of f. If Ai = 7r-'(Ai) and Mi = 2r-'(Mi), then the Ai have all the properties of basic sets for f except they may not be topologically transitive. But f together with the nontrivial covering transformation on M will be transitive, and this is sufficient for everything we have done. So exactly as above, we use the filtration Mi and prove the result for Ai (-,r*: H J(Mi,
Mi_) - HJ(Mi, Mi_1) is surjective-see [1, Theorem 1]). Since dim Ai =
dim A, this completes the proof. Proof of Theorem 1. If 2 # 0 is an eigenvalue of f*: Hk(M) - HI(M) then by Lemma 4 there is an i such that 2 is an eigenvalue of f * : H'(Mi, Mi_) -> Hk(Mi, Mi_). Now if ui = fiber dim Eu(Ai), then from the proof of Proposition 1 we have k - ui < dimR WE(x) and ui - k = (n - k) - si < dimR Wu(x) for x E A. Since dim Ai > dimR Ai = dimR WI(x) + dimR Wu(x)
>max{(k-ui),(ui-k)}=Ik-uiI, the proof is complete. Proof of Theorem 2. If Ai C Mi - Mi_1 is a periodic orbit of period p, then fP fixes each point of Ai and Df2P preserves an orientation on EU(A). Let g = f2P. Since dim Ai = 0, it follows from the proof of Theorem 1 or from [1, Theorem 1] that g,* : Hk(Mi, Mi_1) Hk(Mi, Mi_1) is nilpotent unless k = fiber dim Eu(Ai). Now let L(g) _ k=o (-1)k tr (g,*) _ (-1)1 tr (gu*) where u = fiber dim Eu(Ai). By Lefschetz fixed point theory (see [4, Lemma 3] and [2, Theo-
rem 4.1]). L(g) = Z2,,, I(g; q) where I(g;q) denotes the index of q under g, which by a result of [9, p. 767] is (-1)u. Hence (- 1)u tr (gu)* = L(gm) = (- 1)up for all m > 0. That is, tr (gu)* = p for all m > 0, and it follows that the only nonzero eigenvalue of gu is 1, with multiplicity p. This is because the nonzero eigenvalues with multiplicity of a matrix A are determined by the poles of exp (Em=1 (tr Am)zm/m) (see [1] or [9]) and hence gu has the same nonzero eigenvalues as the p x p identity matrix. Consequently every nonzero eigenvalue of f*: H*(Mi, Mi_1) - H*(Mi, Mi_1) is a root of unity when Ai is
DIMENSION OF BASIC SETS
441
finite. This argument is essentially a reproof of a result of M. Shub [8]. Suppose now that M is orientable. If 2 is an eigenvalue of fk : Hk(M) --> Hk(M) which is not a root of unity, then it follows by Poincare duality (see [1, Lemma 4]) that f* : H,l_k(M) -+ Hn_k(M) has an eigenvalue ±2-1 and hence fn_ : H k(M) -+ H k(M) has an eigenvalue which is not a root of unity. Hence, if A C M, - MS_, is the infinite basic set, then f* : Hj(M8, M,-,) Hj(M,, MS_,) has an eigenvalue which is not a root of unity when j = k and when j = n - k. This follows from Lemma 4 and the fact shown above that f*: H*(Mi, Mti_1) -+ H*(Mi, Mi_1) has only roots of unity and zero as eigenvalues when i s. Thus by Proposition 1, dimA > (n - k) - k if n - k
> kanddimA> k - (n - k) if k> n - k so in any case dimA> In-2kj. If A is an attractor, then the filtration can be chosen such that (Ms, M,-1) = (M1, M° = 0) so f * : H°(Ms, MS _) = H°(M) - H°(MI) is nontrivial and it follows from Proposition 1 that dim A > max {(n - k), k}. This proves the theorem in the case M is orientable. If M is not orientable, let ir : M -+ M be an oriented two-fold covering of M and let f : M -+ M cover f. The map r* : Hk(M) -+ Hk(M) is surjective (see [1, Theorem 1]) so 7r* : Hk(M) , Hk(M) is injective and it follows that f*: Hk(M) , Hk(M) has an eigenvalue which is not a root of unity. Now if Ai = it-I(Ai) it may be that f : Ai -+ Ai is not topologically transitive, but the proof for the orientable case applied to f : M -+ M (using the filtration Mi = 7r-'(Mi)) still shows that if A =7r-'(A) then dim A > I n - 2kI and that if A is an attractor then dim A > max {(n - k), k}. Since dim A = dim A, the result follows. References
R. Bowen, Entropy versus homology for certain diffemorphisms, Topology 13 (1974) 61-67. A. Dold, Fixed point index and fixed point theorem for Euclidean neighborhood retracts, Topology 4 (1965) 1-8. E. Dyer, On the dimension of products, Fund. Math. 47 (1959) 141-160. J. Franks, Morse inequalities for zeta functions. M. Hirsch & C. Pugh, Stable manifolds and hyperbolic sets, Proc. Sympos. Pure Math., Vol. IV, 1970, 133-163. M. Hirsch, J. Palis, C. Pugh & M. Shub, Neighborhoods of hyperbolic sets, Invent.
Math. 9 (1970) 121-134. W. Hurewicz & H. Wallman, Dimension theory, Princeton University Press, Princeton, 1941. M. Shub, Morse-Smale diffeomorphisms are unipotent on homology, Proc. Sympos. Dynamical Systems, Salvador, Academic Press, New York, 1973.
S. Smale, Differentiable dynamical systems, Bull. Amer. Math. Soc. 73 (1967) 747-817. The 0-stability theorem, Proc. Sympos. Pure Math., Vol. IV, 1970, 289-297Stability and isotopy in discrete dynamical systems, Proc. Sympos. Dynamical Systems, Salvador, Academic Press, New York, 1973. E. Spanier, Algebraic topology, McGraw-Hill, New York, 1966. NORTHWESTERN UNIVERSITY
J. DIFFERENTIAL GEOMETRY 12 (1977) 443-460
ISOMETRY OF RIEMANNIAN MANIFOLDS TO SPHERES KENTARO YANO & HITOSI HIRAMATU
1.
Introduction
Let M be a differentiable connected Riemannian manifold of dimension n. We cover M by a system of coordinate neighborhoods {U ; xh}, where and in , n}, and denote the sequel indices h, i, j, k, - run over the range {1, 2, by gji, 17j, Kkjih, Kji and K the metric tensor, the operator of covariant differentiation with respect to the Levi-Civita connection, the curvature tensor, the Ricci tensor and the scalar curvature of M respectively. An infinitesimal transformation vh on M is said to be conformal if it satisfies (1.1)
Yvgji = vjvi + vivj = 2pgji
(vi = gihvh)
for a certain function p on M, where 2v denotes the operator of Lie derivation with respect to the vector field v (see [6]). When we refer in the sequel to an infinitesimal conformal transformation v, we always mean by p the function appearing in (1.1). When p in (1.1) is a constant (respectively, zero), the infinitesimal transformation is said to be homothetic (respectively, isometric). We also denote by YDP the operator of Lie derivation with respect to the vector field pi defined by
pi = gihph = Iip
(1.2)
where (1.3)
I7 = 9ihI h ,
ph = I7hp
gih being contravariant components of the metric tensor. We use gji and gih to lower and raise the indices respectively. The problem of finding conditions for a Riemannian manifold admitting an infinitesimal conformal transformation v to be isometric to a sphere has been extensively studied. For the history of this problem, see [7] and [8]. But in almost all the results on this problem the condition K = constant or Y, K = 0 is not assum0 has been assumed. As results in which the conditon ed, Sawaki and one of the present authors [12] (see also [11]) proved the following two theorems, in which and the remainder of this section, unless stated Communicated July 26, 1975.
444
KENTARO YANO & HITOSI HIRAMATU
otherwise, M will always denote a compact oriented Riemannian manifold of dimension n > 2 admitting an infinitesimal nonhomothetic conformal transformation v. Theorem A. M is isometric to a sphere if v satisfies (1.4)
Y, IY,,(IIGII- n -2 4K" + 2(n n1)(n2 2) 42vK] = 0
where (1.5)
Gji = Kji - 1nKgji
(1.6)
II G112 = GjiGji
,
4 = gjiF"Fj denoting the Laplacian. Theorem B. M is isometric to a sphere if v satisfies (1.7)
I ZII2 - n
+ 24K + 8(n + 1)42 K] = 0 ,
where
Zkjih
(1.8)
1
= Kkjih -
(1.9)
n(n - 1) K(3kgji - ajgki) Zk,,IZk ih .
IIZII2 =
Recently Amur and Hegde [2] (see also [3]) proved the following two theorems.
Theorem C. M is conformal to a sphere if v satisfies 2Dp2VK = 0 and (1.10)
J
(GpipZ
+
12 2v2DPKl dV > 0 ,
n
/
where 2Dp denotes the operator of Lie derivation with respect to pi and dV the volume element of M. Theorem D. M is conformal to a sphere if v satisfies 2Dp2vK=O, 2v2DPK
>0and2vJIGI12=0.
Very recently the present authors [9] proved the following two theorems. Theorem E. M is isometric to a sphere if v satisfies 2, IIG112 = 0 and (1.11)
J
KpipidV
>
1
2n(n - 1) J,'
[2np2K2 + (n + 2)pK2vK + (2vK)2]dV
.
ISOMETRY OF RIEMANNIAN MANIFOLDS
445
Theorem F. M is isometric to a sphere if v satisfies. Yv IJZ III = 0 and (1.11). All the above theorems have been obtained by applying the following Theorem G of Tashiro [5]. The purpose of the present paper is to continue the joint work of the present authors [9] and to prove some propositions on isometry of Riemannian
manifolds to spheres, in which the operator of Lie derivation t plays an important role. In the sequel, we need the following theorems. Theorem G (Tashiro [51). If a complete Riemannian manifold M of dimension n > 2 admits a complete infinitesimal nonhomothetic conformal transformation v such that (1.12)
PA - 1 n4Pgji = 0
then M is isometric to a sphere. Theorem H (Yano and Obata [10]. See also Obata [4]). If a complete Rie-
mannian manifold M of dimension n > 2 admits a nonconstant function p satisfying
(1.13)
V 1p1 - 1JPg11 = 0 , n
2D,K = 0 ,
then M is isometric to a sphere. We remark here that if a Riemannian manifold M of dimension n is isometric to a sphere, then M admits not only an infinitesimal nonhomothetic conformal transformation v satisfying (1.1) and (1.12) but also a nonconstant function p satisfying (1.13). 2.
Lemmas
In this section we prove some lemmas which we need in the next section. M is supposed to be a compact oriented Riemannian manifold of dimension n in all the lemmas except in Lemmas 4, 5, 6, 9 where M is supposed to be only a Riemannian manifold. Lemma 1. If M admits an infinitesimal conformal transformation v, then, for the function p appearing in (1.1) and for an arbitrary function f on M, we have
(2.1)
Proof.
Jpfdv
n J,x 2vfdV .
Since np = Ptvt, by Green's theorem (see [7]) we have
446
KENTARO YANO & HITOSI HIRAMATU
dV
0 = SM V(fvt)dV = SM which proves (2.1). Lemma 2. In M we have
f 'D fhdV = (2.2)
M
J .x
° 'DhfdV
+n
pfdV ,
= f x (Fif)(Fih)dV
=-f JM f4hdV=-f h4fdV M for any functions f and h on M, where IDf denotes the operator of Lie derivation with respect to the vector field Vif on M. Proof. This follows from
0 = f vi(fVih)dV = f (Vif)(Vih)dV + f f4hdV M x
,
J lx
0 = f Mvi(hvif)dV = Lemma 3.
x
(vih)(vif)dV + f h4fdV
.
M
In M we have
f
(2.3)
M
24KdV = -2 fM ppiViKdV
for any function p on M, K being the scalar curvature of M.
Proof. We have (2.3) by putting f = K and h = p2 in (2.2). Lemma 4 (Yano [7]). M, we have (2.4)
S VKkjih
For an infinitesimal conformal transformation v in
= -okV jpi + oJVkpi - Vkplgji + I7jp"gki
(2.5)
Y0Kji = -(n - 2)Vjpi - dpgji
(2.6)
Y,,K = -2(n - 1)4p - 2pK
.
Proof. We can prove these by using (1.1) and the following formulas on Lie derivatives :
v{jhi} = 5j pi + ( pj - gjip" , 2VKk ji" = 1 k2v{ j'ti} - I jSV{k'ti} {jh i} denoting Christoffel symbols formed with gji. Lemma 5. For an infinitesimal conformal transformation v in M, we have
(2.7)
°tvGji = -(n - 2)(1jpi
- 1n dpgji
447
ISOMETRY OF RIEMANNIAN MANIFOLDS
°z' vZk jih = -( I jpi + 3T kpi - vkPhgji + V jP'Lgki (2.8) +I
2
-4P(okgji - 3.% ) n
where Gji and Zkjih are defined by (1.5) and (1.8) respectively. Proof. These follow from Lemma 4. Lemma 6. If M admits an infinitesimal conformal transformation v, then for any function f on M we have
42vf = 1,;4f + 2p4f - (n - 2)pi7if .
(2.9)
Proof.
For an infinitesimal conformal transformation v, we have (see [71)
9kjvkvjvh + Kzhvi + n - 2 Vh(vtvt) = 0
(2.10)
n
Thus we obtain (2.9) by using (2.10) and the identity
gjiFjvtivhf - KhZvif = vh(4f) , which holds for any function f on M. Lemma 7. If M admits an infinitesimal conformal transformation v, then YVYDPKdV
(2.11)
m
n f n+2 if
n
n+2 fm (2.12)
fm p42vKdV ,
$M
and consequently (2.13)
fm
J
Y[v,D,]KdV
n n 2 J M per" 4 KdV + 2(n -} 2)
JM
p4LvKdV ,
where Dp denotes the vector field pi, and [v, Dp] the commutator of vector fields v and Dp. Proof. Using Lemmas 1, 3 and 6, we have
fm pYv4KdV = fm p4Y,KdV - 2 fm p24KdV + (n - 2) fm ppTTiKdV
=
f M
p4Y,KdV + (n + 2) f pYD,KdV
448
KENTARO YANO & HITOSI HIRAMATU
= f p4Y KdV - n
n
'V 'DPKdV ,
2 SM
which proves (2.11). (2.12) follows immediately from Lemma 2. Lemma 8. In M we have, for any function p on M, (2.14)
(2.15)
fm KjipjpidV = - 2 fm Kjipjp'dV 4JM
Proof.
(2.16)
fx p(YDpKji)gjidV
pYDPKdV -
1f
4 Jar p(
jidV DKkjih)gkg p
From the definition of K it follows that fm pYDpKdV = fm
pyDp(Kjigji)dV
pKjiYDpg'idV
fm p(YDpKji)gjidV + SM
.
On the other hand, since pi is a gradient, we have (2.17) (2.18)
YDpgji = -2Vjpi
YDpgji = 2Vjpi ,
Vj(ppiKji) = Kjipjp2 + pKjiV'pi + 2pp'ViK
where we have used FjK ji = IF jK. Using (2.16), (2.17) and (2.18), we have (2.14). We also have
(2.19)
fm p2DPKdV = fm pYDp(Kk jingkhgji)dV
= fm p(YDpKkjih)gkhgjidV - 4 f IV pKjiVjpidV from which and (2.18), (2.15) follows immediately. Lemma 9. In M we have, for any function p on M,
Kjipjpi + n (dp)2 + 2Dp4p(2.20)
Proof.
-(Fjpi -
n
Jpgji)(vjpi
14YDPp
-
Using Ricci formula we have 42',,p = gkivkv j(pipi) = 2gkiV k(piV Jpi)
n4pgiil 1
449
ISOMETRY OF RIEMANNIAN MANIFOLDS
= 2gk'(VkVJpi)pi + 2(v>Pi)(V1pi)
=
2gki(VivkP> - Kkijhph)pi + 2(v1Pi)(VJpi)
from which we find (2.20). Lemma 10. In M we have, for any function p on M,
SM Kp1p1dV + n n 1 fm YDPJpdV (2.21)
-f
(vp,
-
4Pgji(Vipi
n
4Pgii)dV ,
-n
or
fm KlipjpidV - n (2.22)
-J a2
Proof.
(v3Pi
n
1 fm (dp)W
- n i pgji
k7i
1
P
- n zPgui)dV
.
These follow from Lemmas 2 and 9.
Lemma 11. A sphere S" of dimension n > 2 admits a nonconstant function p such that (2.23)
v j pi
=0 - -Jpgji n
and consequently (2.24)
(2.25)
42p +
n
1
1
Kd p= 0,
v;vid p+
n
1
1
KV,pi = 0
vivizp - 1 42Pgii = 0 . n
Proof. It is known [11] that Sn admits a nonconstant function p such that (2.23) holds. This shows that the vector field pi defines an infinitesimal non-
homothetic conformal transformation on Sn with the associated function (1 /n)d p. Since K is a positive constant, using (2.6) in which v and p are replaced by ph and (1 / n)d p respectively we have the first equation of (2.24) and
therefore J p + (1 /(n - 1)) pK = c (c: constant), which implies the second equation of (2.24). From (2.23) and (2.24) we obtain (2.25). 3.
Propositions
In this section, we prove a series of propositions in which the operator of Lie derivation YDP plays an important role. M is supposed to be a compact
450
KENTARO YANO & HITOSI HIRAMATU
oriented Riemannian manifold of dimension n admitting an infinitesimal conformal transformation v in all the propositions and corollaries except : in Pro-
position 4 where M is supposed to be a complete Riemannian manifold of dimension n > 2, in Propositions 5, 7 and Corollary 5 where M is supposed to be a complete Riemannian manifold of dimension n > 2 admitting a complete infinitesimal nonhomothetic conformal transformation v, in Propositions 6, 12 and 13 where M is supposed to be only a Riemannian manifold, and in Propositions 8, 10 and Corollaries 1, 3 where M is supposed to be a compact oriented Riemannian manifold of dimension n. Proposition 1. For M we have (3.1) fm
1
G,ip5pidV -I 1 fM
f"'
0
2n
n2
The M of dimension n > 2 admits a nonhomothetic v such that the equality in (3.1) holds if and only if M is isometric to a sphere. Proof. By using (1.5), (2.6), Lemmas 1 and 2 and the identity
f Vi(pp'K)dV = fly KpipidV + f pK4pdV + f ppiViKdV = 0 ,
(3.2)
we have
f
M
K,ip3 pidV - n n 1 fm
= fM G,ipjpidV + =
fM G,i pjpidV
n-1 n
M
-
n
n
p)ZdV
f m KpipidV f
n n
pYDPKdV M
n
1 fM
p)2dV
fm pKJpdV
(4p)2dV
= fMGjip'pidV +n21 fM
= f G,ip'pidV +
1
f"'
n2
12n f 12n f
m
(d
M
Thus from Lemma 10 we obtain
f
M
(3.3)
G,ip'pidV + -1
n2
f YvYDPKdV - i f YDPY KdV 2n M
-fm (Fjpi - n d pgji} (V ipi
M
- n d pg'i)dV
ISOMETRY OF RIEMANNIAN MANIFOLDS
451
which implies (3.1). If the equality in (3.1) holds, then from (3.3) and Theorem G it follows that M is isometric to a sphere. Conversely, if M is isometric to a sphere, M admits an infinitesimal nonhomothetic conformal transformation v such that the equality in (3.1) holds because, for a sphere, G;i = 0 and K is a positive constant. Proposition 1 is a generalization of Theorem C. Proposition 2. If the dimension n of M is greater than 2, then
SM2v2MG dV - (n - 2) J
(3.4)
0.
,
ar
The M of dimension n > 2 admits a nonhomothetic v such that the equality in (3.4) holds if and only if M is isometric to a sphere. Proof. First of all we have -Tv G I = 2(-TvG.')G'i - 4p I GII'. 2
Substituting (2.7) in the above equation we find
-Tv 11GII2 = -2(n - 2)G,iVjpi - 4p GII2 ,
because of G;igjz = 0 or (3.5)
KjiV'p2
=- n
2 2 p I G 1I2 -
2(n
I
1
2)
-T,, G I I2 + 1 Kd p.
Using (2.18) and (3.5) we have P'(ppiKji) = K.jip'p2
-
2
P' IIGII'
--2(n
1
n
2
2) p-Tv I I G I12 + 1 p-TDPK + pKd p .
Integrating both sides of the above equation over M and using (2.6) and Lemmas 1 and 2, we obtain JMK;ipipzdV - n n
2
n
+
M(dp)'dV
2 Jx p' IIGII' dV -
2n J M
2
n
1J
2JM
2n(n
1
2)JM Y,Y, IIGIII, dV
YvYDPKdV - 1 f"' pKJpdV - n n 1 f p)'dV m (d p' G j j' dV - 2n(n1-
2)
f"' Yvtv
11 G I IZ dV
452
KENTARO YANO & HITOSI HIRAMATU
+ 2n SM YvYDPKdV - 2n fm
YDPYvKdV
or, by Lemma 10, II G IIZ dV - (n - 2) fm 2-'[V,DP]KdV
fm
= 2n(n - 2) f"' (v1pi +4nfm p2I
- n 4Pg;i) (V ipi - n d pgii )dV
G I IZ dV,
which together with Theorem G gives the proposition. Remark 1. Proposition 2 is a generalization of Theorem D. Using (2.13) and Lemma 1 we have SM (3 6)
f Y,Yv4KdV
1
n+2JM
- 2(n + 1) f n(n+2) M
Therefore Proposition 2 is essentially equivalent to Theorem A. Using (2.6), (3.2) and Lemmas 1 and 2 we have n SM KpipidV
SM Y[ (3.7)
- 2(n
1
1)
f
[2np2K2 + (n + 2)pKtvK + (IvK)Z]dV , M
which implies that Proposition 2 is essentially equivalent to Theorem E. Proposition 3. For M we have (3.8) SM
Yv1v
IZIIZ
dV - 4
fm
0.
The M of dimension n > 2 admits a nonhomothetic v such that the equality in (3.8) holds if and only if M is isometric to a sphere. Proof. First of all we have
yv IIZII2 = 2(YvZkj )Zkiih - 4p IIZIIZ Substituting (2.8) in the above equation we find yv IIZIIZ = -8G5iV'pi - 4p IIZIIZ
because of Zk,ik = G,i and G;ig'i = 0, or
453
ISOMETRY OF RIEMANNIAN MANIFOLDS
(3.9)
2 p II Z II2 - 1 Y,, it Z I2 + n K4 p .
KjiV''pi
8
Using (2.18) and (3.9) we have
V'(pp Kji) = Kj,pipi - 1 p2 JZIJ2
- 1 Pyv Al, + 1 pYDPK + 1 pKdp
.
Integrating both sides of the above equation over M and using (2.6) and Lemmas 1 and 2, we obtain fm KjipjpidV
- n n 1 fm (4p)zdV JarY"Y'IZI zdV
farp2jjZIj2dV-
2
1
+ 2n Jar
YVYDPKdV -
2n
far YDPYVKdV
,
or, by Lemma 10, p
f if
2' 2v II Z II2 dV - 4 Jar'[,,,DP]KdV 8n far
(Fjpi -
ndpgji)(V3p. - n4pgi )dV + 4n f
p2 JJZJI2dV ,
M
which together with Theorem G gives Proposition 3. Remark 2. Using (3.6), (3.7) and (3.8) we see that Proposition 3 is essentially equivalent to Theorems B and F. Proposition 4. M admits a nonconstant function p satisfying (3.10)
2'DPgji = 2pgji
I
eL DPK = 0
cp being a function on M, if and only if M is isometric to a sphere. Proof. If M admits a nonconstant function p satisfying (3.10), then, by Theorem H, M is isometric to a sphere because (3.10) is equivalent to (1.13). Conversely if M is isometric to a sphere, then M admits a nonconstant function p satisfying (2.23) and hence (3.10) because K is a positive constant for a sphere. Proposition 5. M admits a transformation v such that YDPgji = 2pgji cp being a function on M, if and only if M is isometric to a sphere.
454
KENTARO YANO & HITOSI HIRAMATU
Proof. This follows immediately from Theorem G. Ackler and Hsiung [1] proved this proposition for a special case in which the manifold M is compact and oriented and both SfVK = 0 and SfDPK = 0 hold. Proposition 6. For any function p on M we have (3.11)
Kjipjp1 + n (4p)2 + SfDPQp -
2
J- D,ap < 0 .
The complete M of dimension n > 2 admits a nonconstant function p such that the equality in (3.11) holds and SfDPK = 0 if and only if M is isometric to a sphere. Proof. This follows from Theorem H and Lemma 9. Proposition 7. M admits a transformation v such that the equality in (3.11) holds if and only if M is isometric to a sphere. Proof. This follows from Theorem G and Lemma 9. Proposition 8. For any function p on M we have (3.12)
J p(SDpKji)gjidV +
2(n
-
1)
n
f pJ2pdV > 0 . x
The M of dimension n > 2 admits a nonconstant function p such that 9DPK = 0 and the equality in (3.12) holds if and only if M is isometric to a sphere. Proof. Using Lemmas 2, 8 and 10 we have tit
p(YDpKji)gjidV +
(3.13)
2 rM (Fjpi
2(n
-
1)
fm
n
pd2pdV
- n d pgjz)(vjp2 - n d pgii)dV
,
which together with Theorem H gives Proposition 8. Corollary 1. M of dimension n> 2 admits a nonconstant function p such
that 9DPK = 0 and (3.14)
YDpKji
= - 2(nn- 1) d2pg;i 2
if and only if M is isometric to a sphere. Proof. If M is isometric to a sphere, then M admits a nonconstant function p such that (2.23) holds. Therefore using (2.24) we have 1
2
n
n
YDpKji = -KYDpgji = 2 n2
Kdpg;i = -
KV jpi
2(n - 1) dZpg;i . n2
455
ISOMETRY OF RIEMANNIAN MANIFOLDS
The "only if" part of the corollary is an immediate consequence of Proposition 8. Remark 3. By (2.25) in Lemma 11, (3.14) in Corollary 1 can be replaced by
2(n - 1) PjVidp .
YDPKji
(3.15)
n
Proposition 9. For M we have (3.12), and the M of dimension n > 2 admits a nonhomothetic v such that the equality in (3.12) holds if and only if M is isometric to a sphere. Proof. This follows from (3.13) and Theorem G.
Corollary 2. M of dimension n > 2 admits a nonhomothetic v such that (3.14) holds if and only if M is isometric to a sphere. Proof. This follows from Lemma 11 and Proposition 9. Remark 4. By (2.25) in Lemma 11, (3.14) in Corollary 2 can be replaced by (3.15). Proposition 10. For any function p on M we have p(yDPKkjih)gkhgjidV + fm
SM
(3.16)
P
+ 4(n - 1) j pdzpdV > 0
DpKdV .
JM
n
The M of dimension n > 2 admits a nonconstant function p such that IDPK = 0 and the equality in (3.16) holds if and only if M is isometric to a sphere. Proof. Using Lemmas 2, 8 and 10, we have SM pl
DpKk jih )gkhgjidV + SM p
DPKdV +
4(n
- 1) n
pd2pdV
SM
(3.17)
4 f", (17jpj - n dpgji)(P'pi - n dpgji)dV , which together with Theorem H gives the proposition. Corollary 3. M of dimension n > 2 admits a nonconstant function p such
that fDPK = 0 and (3.18)
o 7 °L
4
n2
2
p(gkhg ji - gjhgki)
if and only if M is isometric to a sphere. Proof. If M is isometric to a sphere, then M admits a nonconstant function p such that (2.23) holds. Since K is a positive constant and
456
KENTARO YANO & HITOSI HIRAMATU
Kkjih = n(n
1-
K(
1)
(gkhgji - gjhgki)
for a sphere, using (2.24) we obtain / + gkhvjpi - vjphgki - gjhvkpi) - n(n2- 1) K(Fkphgji
n
(VkFhd pgji + gkhvjviJp - vjvh4pgki - gjhvkvid p)
which together with (2.25) gives (3.18). The "only if" part of the corollary is an immediate consequence of Proposition 10. Remark 5. As is seen in the proof of Corollary 3, (3.18) in Corollary 3 can be replaced by ° 'DpKkjih
(3.19)
2
n
(vkvhdpgji + gkhvjvidp - vjvhJPgki - gjhvkvidp)
Proposition 11. For M we have (3.16). The M of dimension n > 2 admits a nonhomothetic v such that the equality in (3.16) holds if and only if M is isometric to a sphere. Proof. This follows from (3.17) and Theorem G. Corollary 4. M of dimension n > 2 admits a nonhomothetic v such that ° 'DpKkjih
(3.20)
_ - (
1
)
1
[DPK
4(n -
n
Zp
gkhgji -gjhgki)
holds if and only if M is isometric to a sphere. Proof. This follows from Lemma 11 and Proposition 11. Remark 6. In Corollary 4, we see, by using Lemma 11, that (3.20) can be replaced by DpKk jih 1
(3.21)
hgji - gjhgki)
n(n - 1) n
(Vkvhdpgji + gkhvjviJp - vjvhdpgki - gjhvkvidp)
Proposition 12. If M of dimension n > 2 admits an infinitesimal conformal transformation v, then (3.22)
(
Dp°L vGji)gji < 0 .
ISOMETRY OF RIEMANNIAN MANIFOLDS
457
The complete M of dimension n > 2 admits a complete infinitesimal nonhomothetic conformal transformation v such that the equality in (3.22) holds if and only if M is isometric to a sphere. Proof. By using (2.7) we have
(YvGji)gji = 0
,
and consequently
(IDp2VGji)gji = -(YvGji)YDpgji = 2(2vGji)Fjpi
_ -2(n - 2)(Fjpi - 1 dpgjiFjPi n
_ -2(n - 2)(Fjpi
-
dpgji)(V'pi
-
n
dpgii)
n which together with Theorem G gives the proposition. Proposition 13. For M of dimension n > 2 we have
(3.23)
(2'Dp2VZkjih - 2pYDpZkjih)gkhgji < 0
The complete M of dimension n > 2 admits a complete nonhomothetic v such that the equality in (3.23) holds if and only if M is isometric to a sphere. Proof. From (2.8) it follows that
YvZkjih = -gkhvjpi + gjhvkpi - Vkphgji + vjphgki 2 n
+ dp(gkhgji - gjhgki) + 2pZkjih
,
and therefore that (YvZkjih)gkhgii = 0 . Using this we obtain (YD yVz"jih)gkh9ji
= 4(yvZkjih)gjipkph 1 i = -4(n - 2) Fjpi - -dpgjiFjp
n
+ 8pZkjing'ipkph .
On the other hand, since Zkjingkhgji = 0 we have (YDpZkjih)gkhgji = 4Zkjingj'Fkph
Thus
-
1
n
apgii)
458
KENTARO YANO & HITOSI HIRAMATU
(YDPYVZkjih - 2'PYDpZkjih)gkhgji
_ -4(n -
2)(Pjpi
-
4Pgji )('Pi P
-
n
4pgii)
n
which together with Theorem G gives the proposition. Corollary 5. M admits a transformation v such that eZ. DpeZ. VGji = 0
(3.24)
or
2p°Z' DpZkjih = 0
(3.25)
if and only if M is isometric to a sphere. Proof. This follows from Propositions 12 and 13. Proposition 14. For M we have (3.26)
JM
P(YDpGji)gjidV - n
J
Y[v,Dp7KdV > 0
The M of dimension n > 2 admits a nonhomothetic v such that the equality in (3.26) holds if and only if M is isometric to a sphere. Proof. We have, by using Gjigji = 0,
p(1DpGji)gji = -pGjiyDpgji = 2pGjiVjpi (3.27)
= 2pK jiV jpi -
2
n
PK4P
or, using (2.18), 2
p(
DpGji)gj' = Vj(ppiKji) - Kjipjpi - 2 pYDPK
- n pKdp
.
Integrating both sides of the above equation over M and using (2.6), we find
fm K jipjpidV - n n 1 fm (d p)2dV 2 Jar
p(YDpGji)gjidV - 2 f"' P2DpKdV
- 1n J pKJ pdV M
2 J a1
n-1 n
f (d p)2dV Jar
p(YDpGji)g'idV + 2n
f" YVYDpKdV
ISOMETRY OF RIEMANNIAN MANIFOLDS
+ 2n
459
SM (4p)2vKdV ,
or, by Lemmas 2 and 10,
1 SM 2'[v,DP]KdV (17jpi - n 4Pgji)(ViPi - n 4Pgii)dV 2 SM
SM p(2DPGji)gjidV (3.28)
which together with Theorem G gives the proposition. Corollary 6. M of dimension n > 2 admits a nonhomothetic v such that 1
p22DPGji =
(3.29)
nZ
if and only if M is isometric to a sphere. Proof. This is an immediate consequence of Proposition 14. Corollary 7. M of dimension n > 2 admits a nonhomothetic v such that
'DPGji = -
(3.30)
1
n(n
2)
[4K - 2(n n 1) dYvK]gji
if and only if M is isometric to a sphere. Proof. This follows from Lemma 7 and Proposition 14. Proposition 15. For M we have (3.31)
P(yDpZkjih)gk''gjidV -
fm
n JM
'[v,Dp]KdV > 0 .
The M of dimension n > 2 admits a nonhomothetic v such that the equality in (3.31) holds if and only if M is isometric to a sphere. Proof. We have, by using Zkjih,gk't = Gji and Gjigji = 0, p(YDPZkjih)gkhgji = -2pGjiyDpgji which together with
p(YDPGji)gji = -pGji2Dpgji implies
p(2'DpZk jih)gkhgji = 2p(yDPGji)gji
Integrating both sides of the above equation over M and using (3.28), we obtain M
P(YDpZkji3a)gkhgjidV
-2
n SM
Y[v,DPI KG.Y
460
KENTARO YANO & HITOSI HIRAMATU
4 J" (vpi
- n Jpgji )(17ip1 - n JPgji" Idv
,
which together with Theorem G gives the proposition. Corollary 8. M of dimension n > 2 admits a nonhomothetic v such that (3.32)
D,Zkjih =
p
2
nz(n - 1)
(_T[v,Dp]K)(gkhgji - gjhgki)
if and only if M is isometric to a sphere. Proof. This is an immediate consequence of Proposition 15. Corollary 9. M of dimension n > 2 admits a nonhomothetic v such that cfDoZkjih
(3.33)
2
n(n - 1)(n + 2) kvJK
- 2(n n 1) d-TvKj (gkhgji - gjhgki) ,
if and only if M is isometric to a sphere. Proof. This follows from Lemma 7 and Proposition 15. Bibliography [ 11
L. L. Ackler & C. C. Hsiung, Isometry of Riemannian manifolds to spheres, Ann.
Mat. Pura Appl. 99 (1974) 53-64. K. Amur & V. S. Hedge, Conformality of Riemannian manifolds to spheres, J. Differential Geometry 9 (1974) 571-576. [ ] , Some conditions for conformality of Riemannian manifolds to spheres, Tensor 28 (1974) 102-106. [ 4 ] M. Obata, Certain conditions for a Riemannian manifold to be isometric with a sphere, J. Math. Soc. Japan 14 (1962) 333-340. [ 5 ] Y. Tashiro, Complete Riemannian manifolds and some vector fields, Trans. Amer. Math. Soc. 117 (1965) 251-275. [ 6 ] K. Yano, The theory of Lie derivatives and its applications, North-Holland, [2]
Amsterdam, 1957.
[ 7 ] -, Integral formulas in Riemannian geometry, Marcel Dekker, New York, 1970. [8]
, Conformal transformations in Riemannian manifolds, Differentialgeometrie im Grossen, Berichte Math. Forschungsinst., Oberwolfach, Vol. 4, 1971, 339351.
K. Yano & H. Hiramatu, Riemannian manifolds admitting an infinitesimal conformal transformation, J. Differential Geometry 10 (1975) 23-38. [10] K. Yano & M. Obata, Conformal changes of Riemannian metrics, J. Differential Geometry 4 (1970), 53-72. [11] K. Yano & S. Sawaki, Riemannian manifolds admitting a conformal transformation group, J. Differential Geometry 2 (1968) 161-184. [12] , Riemannian manifolds admitting an infinitesimal conformal transformation, Kodai Math. Sem. Rep. 22 (1970) 272-300. [ 91
TOKYO INSTITUTE OF TECHNOLOGY KUMAMOTO UNIVERSITY, JAPAN
J. DIFFERENTIAL GEOMETRY 12 (1977) 461-471
SOME ALMOST HERMITIAN MANIFOLDS WITH CONSTANT HOLOMORPHIC SECTIONAL
CURVATURE LIEVEN VANHECKE
B. Smyth proved in [3] Theorem A. Let M be a complex hypersurface of a Kahlerian manifold M of constant holomorphic sectional curvature p. If M is of complex dimension > 2, then the following statements are equivalent :
(i) M is totally geoddsic in Al-, (ii) M is of constant holomorphic sectional curvature, (iii) M is an Einstein manifold, and at one point of M all sectional curwhen p > 0 (resp. G 0). vatures of M are > 4I p (resp. G 1p) 4 Considering nearly Kahler manifolds, S. Sawaki and K. Sekigawa proved in [2] the following generalization of this theorem. Theorem B. Let M be a complex hypersurface of a nearly Kahler manifold M with constant holomorphic sectional curvature p. If M is of complex dimension > 2, then the following statements are equivalent :
(i) M is totally geodesic in M, (ii) M is of constant holomorphic sectional curvature, (iii) at every point m e M all the sectional curvatures of M satisfy K(x, y) > 4p{1 + 3g(x, Jy)z} , where x, y are any orthonormal vectors of T,,,,(M). An almost Hermitian manifold with J-invariant Riemann curvature tensor is called an RK-manifold [6]. RK-manifolds with pointwise constant type form a particularly nice class of almost Hermitian manifolds, and many properties for Kahler manifolds can be generalized to this class [4], [5], [6], [7]. An RKmanifold with pointwise constant type and pointwise constant holomorphic sectional curvature is an Einstein manifold. The main purpose of this paper is to generalize the theorem of Smyth to complex hypersurf aces of such manifolds satisfying an interesting condition. This is done in § 3 following the same arguments as in [2], [3]. In § 1 we give some generalizations of theorems for RK-manifolds [6] to almost Hermitian manifolds. In § 2 we state some differential-geometric proCommunicated by K. Yano, July 26, 1975.
462
LIEVEN VANHECKE
perties of a complex hypersurf ace of an almost Hermitian manifold satisfying a certain condition, and finally in § 4 we give some properties for the holomorphic bisectional curvature [1]. We remark that, if necessary, the complex hypersurface is supposed to be connected. 1. Let M be a C°° differentiable manifold which is almost Hermitian, that is, the tangent bundle has an almost complex structure J and a Riemannian metric g such that g(JX, JY) = g(X, Y) for all X, Y E y(M) where y(M) is the Lie algebra of C`° vector fields on M. We suppose that dim M = n = 2m, and we denote by F the Riemannian connection on M. Let R be the Riemann curvature tensor, S the Ricci tensor defined by
(1)
n
S(x, Y) _
R(x, e21 y, ez) i=1
where x, y r= T, ,,(M), m r= M and {e1} is an orthonormal local frame field, and
K(x, y) the sectional curvature for a 2-plane spanned by x and y. We denote by H(x) the holomorphic sectional curvature of the 2-plane spanned by x and Jx. The sectional curvature of the antiholomorphic plane spanned by x and y, where g(x, y) = g(x, Jy) = 0, is called the antiholomorphic sectional curvature. An almost Hermition manifold such that the Riemann curvature tensor R is J-invariant, that is,
(2)
R(JX, JY, JZ, JW) = R(X, Y, Z, W)
,
yX, Y, Z, W e y(M)
is said to be an RK-manifold [6]. For such a manifold we have
(3) (4)
K(x, y) = K(Jx, Jy) S(x, Y) = S(Jx, Jy)
,
,
K(x, Jy) = K(Jx, y)
S(x, Jy) + S(Jx, y) = 0
We say further that an almost Hermitian manifold is of constant type at m r= M provided that for all x e Tm(M) we have
(5)
A(x, Y) = A(x, z)
with
(6)
A(x,y) = R(x,y,x,y) - R(x,y,Jx,Jy)
whenever the planes defined by x, y and x, z are antiholomorphic and g(y, y) = g(z, z). If this holds for all m r= M, we say that M has (pointwise) constant type. Finally, if X, Y E y(M) with g(X, Y) = g(X, JY) = 0, 2(X, Y) is con-
stant whenever g(X, X) = g(Y, Y) = 1, then M is said to have global constant type. The following theorems are generalizations of theorems given in [6]. The proofs are easy verifications.
ALMOST HERMITIAN MANIFOLDS
Theorem 1. Then
463
Let M be an almost Hermitian manifold and x, y e
R(x, y, x, y) = 32{3Q(x + Jy) + 3Q(x - Jy) - Q(x + y)
- Q(x - y) - 4Q(x) - 4Q(y)}
6 y) - 32(Jx, Jy)} + 1{13,(x,
(7)
+
6
{2(x, JY) + 2(Jx, Y}
,
where Q(x) = R(x, Jx, x, Jx). Theorem 2. Assume M is almost Hermitian, and let x, y e that g(x, x) = g(y, y) = 1 and g(x, Jy) = cos 8 > 0. Then
be such
K(x, y) = 8 {3(1 + cos 8)'H(x + Jy) + 3(1 - cos 8)''H(x - Jy)
- H(x + y) - H(x - y) - H(x) - H(y)}
(8)
+
6
{131(x, y) - 32(Jx, Jy)} +
6
{2(x, Jy) + 2(Jx, Y)}
,
if g(x, y) = 0. Theorem 3. Suppose M has constant holomorphic sectional curvature P at with g(x, x) = g(y, y) = 1 and g(x, y) = 0. a point m e M, and let x, y E Then
{1 + 3g(x, Jy)Z} +
K(x, y) = (9)
16{13,1(x, Y) - 32(Jx, Jy)}
4
+
16{2(x, Jy) + 2(Jx, Y)} .
Theorem 4. Let M be an almost Hermitian manifold with pointwise constant holomorphic sectional curvature p and pointwise constant type a. Then M is an Einstein manifold with (10)
2S(x,
x) = (m + 1)i + 3(m - 1)a
for g(x, x) = 1, and M is a space of constant holomorphic sectional curvature if and only if M has global constant type a. The definition of a in the theorem is given by (11)
2(x, y) = a ,
if g(x, x) = g(y, y) = 1 where x and y span an antiholomorphic plane.
464
LIEVEN VANHECKE
The following theorem is proved in [6]. Theorem 5. Assume M is an RK-manifold. Then M has (pointwise) constant type if and only if there exists a C°-function cr such that (12)
A(X, Y) = a{g(X, X)g(Y, Y) - g(X, Y)2 - g(X, JY)2}
for all X, Y e x(M). Furthermore, M has global constant type if and only if (13) holds with a constant function x. 2. For our purpose we need some considerations on complex hypersurfaces of an almost Hermitian manifold. We follow the notation of [2] and refer to that paper for the proofs of the given properties. See also [3]. Let M be an almost Hermitian manifold of complex dimension m + 1, and denote the almost complex structure and the Hermitian metric of M by J and g respectively. Moreover, let M be a complex hypersurface of M i.e., suppose that there exists a complex analytic mapping f : M -+ M. Then for each m e M we identify the tangent space Tm(M) with f*(Tm(M)) C T f1x,(M) by means of f Since f * o g = g' and J o f * = f * 0 J' where g' and J' are the Hermitian metric and the almost complex structure of M respectively, g' and J' are respectively identified with the restrictions of the structures g and J to the subspace f*(Tm(M)).
As is known, we can choose the following special neighborhood '&(m) of m for a neighborhood lC(f(m)) of f(m). Let {f/ ; m } (i = 1, 2, , 2m + 2) be a system of coordinate neighborhoods of M. Then {Qe ; mi} is a system of coordinate neighborhoods of M such that m2m+1= m2m+2 = 0 where mi = mi of.
By P we always mean the Riemannian covariant differentiation on M, and by N a differentiable unit vector field normal to M at each point of °I?(m). If X and Y are vector fields on the neighborhood Qe(m), we have (12)
I1Y = VXY + h(X, Y)N + k(X, Y)JN ,
where VSY denote the component of FXY tangent to M, V is the covariant differentiation of the almost complex Hermitian manifold M, and h and k are symmetric covariant tensor fields of degree 2 on °&(m). We have further (13) (14)
IXN = -AX + s(X)JN , Px(JN) = -BX + t(X)N ,
where AX and BX are tangent to M. A, B, s and t are tensor fields on 1&(m) of type (1,1) and (0,1) respectively, and A and B are symmetric with respect to g and satisfy (15)
(16)
h(X, Y) = g(AX, Y) k(X, Y) = g(BX, Y)
Now let M be a complex hypersurface satisfying the condition
ALMOST HERMITIAN MANIFOLDS
465
h(X, Y) = k(X, JY)
(17)
for any vector fields X and Y on QI(m) at every point m E M. It is easy to verify that this condition is independent of the choice of N. For such a hypersurface we have (18)
JA = -AJ ,
JB = -BJ ,
where JA and JB are symmetric with respect to g. Condition (17) is equivalent to (19)
B=JA.
Moreover we have Lemma 6 [2]. In a complex hypersurface M of M satisfying (17), at any , m) of point p E 0&(m) there exists an orthonormal basis {ei, Jez} (i = 1, 2, T,(M) with respect to which the matrix A is diagonal of the form
where Aez = 2 ez and AJe2 = -2jei. Lemma 7 [2]. If R and R are the Riemannian curvature tensors of k and a complex hypersurface M of k satisfying (17) respectively, then for any vector fields X, Y, Z, W on 0&(m) we have the following Gauss equation : R(X, Y, Z, W) = R(X, Y, Z, W) (20)
- {g(AX, Z)g(AY, W) - g(AX, W)g(AY, Z)} - {g(JAX, Z)g(JAY, W) - g(JAX, W)g(JAY, Z)}
.
Lemma 8 [2]. Let M be a complex hypersurface of k and satisfy condition (17). (i) If {x, y} is a 2-plane tangent to M at a point of 01'(m), then (21)
K(x, y) = K(x, y) - {g(Ax, x)g(Ay, y) - g(Ax, y)Z} - {g(JAx, x)g(JAy, y) - g(JAx, y)2}
where x, y form an orthonormal basis of the 2-plane.
LIEVEN VANHECKE
466
(ii)
(22)
If x is a unit vector tangent to M at a point of Gll(m), then H(x) = H(x) + 2{g(Ax, x)2 + g(JAx, x)2}
.
Proposition 9 [2]. Let M be a complex hypersurface of 111 of (pointwise) constant holomorphic sectional curvature u. If M is of complex dimension > 2 and satisfies condition (17), then at each point of M there exists a holomorphic plane whose sectional curvature in M is u, and therefore if M is of (pointwise) constant holomorphic sectional curvature u, then p = u. Finally this proposition gives Theorem 10 [2]. Let M be a complex hypersurface of M of constant holomorphic sectional curvature. If M is of complex dimension > 2 and satisfies condition (17), then the following statements are equivalent:
(i) M is totally geodesic in M, (ii) M is of constant holomorphic sectional curvature. 3. Let 1V1 be an almost Hermitian manifold, and M a complex hypersurface of 111 satisfying condition (17). It follows at once from (20), (19) and (18) that (23)
R(JX, JY, JZ, JW) - R(X, Y, Z, W) = R(JX, JY, JZ, JW) - R(X, Y, Z, W)
for any vector fields X, Y, Z, W on Old(m). Hence Theorem 11. Let M be an almost Hermitian manifold, and M a complex
hypersurface of k satisfying condition (17). If k is an RK-manifold, then M is also an RK-manifold. Further we have also (24)
R(X, Y, Z, W) - R(X, Y, JZ, JW) = R(X, Y, Z, W) - R(X, Y, JZ, JW)
for any vector fields X, Y, Z, W on Gll(m). Hence (25)
. (X, Y) = 2(X, Y) ,
and from (25) and Theorems 5, 11 we obtain Theorem 12. Let 1i21 be an RK-manifold of (pointwise) constant type a, and M a complex hypersurface satifying condition (17). Then M has (pointwise) constant type a. We need only this theorem for RK-manifolds, but it is easy to prove that this is still valid for a general almost Hermitian manifold. With the help of Theorem 4 we obtain an equivalent version of Theorem 10 for manifolds with (pointwise) constant type. Theorem 13. Let M be an almost Hermitian manifold of (pointwise) constant type, and M a complex hypersurface of complex dimension > 2 satisfy-
ALMOST HERMITIAN MANIFOLDS
467
ing condition (17). Then the following statements are equivalent:
(i) M is totally geodesic in k, (ii) M has global constant type and pointwise constant holomorphic sectional curvature. The following theorem is an immediate consequence of (9) and (12). Theorem 14. Let M be an RK-manifold with (pointwise) constant holomorphic sectional curvature p and (pointwise) constant type a. If x, y E T (M), M E M and g(x, x) = g(y, y) = 1, g(x, y) = 0, then (26)
K(x, y) = pf l + 3g(x, Jy)}z + -3,a{1 - g(x, Jy)2}
We prove now the main theorem of this paper. Theorem 15. Let M be a complex hypersurface of an RK-manifold M with constant holomorphic sectional curvature / and constant type a. If M is of complex dimension > 2 and satisfies condition (17), then the following statements are equivalent :
(i) M is totally geodesic in M, (ii) M is of constant holomorphic sectional curvature (or equivalently, M has global constant type and pointwise constant holomorphic sectional curvature), (iii)
(27)
at every point m E M, all the sectional curvatures of M satisfy
K(x, y) > p{1 + 3g(x, Jy)'} + 4a
,
if a > 0 ,
K(x, y) <
,
if a< 0
or (28)
4
{1 + 3g(x, Jy)z} + 3a
,
where x, y are orthonormal vectors spanning the 2-plane of Tm(M). Proof. (i) is equivalent to (ii) by Theorem 10 and Theorem 13.
Next, if M is of constant holomorphic sectional curvature p, then p = F by Proposition 9, and therefore we have (26) which implies (27) and (28). Finally, we prove that (iii) implies (i). Therefore consider an orthonormal basis as in Lemma 6, and set
Then from (21) follows (30)
K(x, y) = K(x, y) + 22
.
In the case a > 0, from (27) and the expression (26) for M we obtain
- 4 ag(x, JY)' > 22i ,
468
LIEVEN VANHECKE
which implies 2i = 0 (i = 1, , m). It follows then from Lemma 6 that A is identically zero at each point of M, so that M is totally geodesic in Al. In the same way we can treat the case a G 0. Following the same arguments we obtain Theorem 16. Let M be a complex hypersurface of an RK-manifold Al with pointwise constant holomorphic sectional curvature p vnd vanishing con-
stant type. If M is of complex dimension > 2 and satisfies condition (17), then the following statements are equivalent :
(i) M is totally geodesic in Al, (ii) M has pointwise constant holomorphic sectional curvature, (iii) at every point m e M, all the sectional curvatures of M satisfy (31)
K(x, y) > 4p{1 + 3g(x, Jy)Z}
where x, y are orthonormal vectors which span the 2-plane of Consider again an almost Hermitian manifold M of constant holomorphic sectional curvature p and (pointwise) constant type a. We know from Theorem 4 that k is an Einstein manifold with (32)
S=pg,
2p=(m+ 1)p+3(m-1)a.
Now let M be a complex hypersurface of M which satisfy condition (17), and consider further the basis {ei, Jei} of Lemma 6. Then it follows with the help of (9), (21) and (22) that (33)
(34)
H(ei) = p - 222 , S(ei, ei) = 2(m + 1)p + 2(m - 1)a - 222
If M is an Einstein manifold, then we have (35)
S=pg,
p=p-222 A2 = 221
(36)
where (37)
422=422=(m+
1)p+3(m-1)a-2p
Moreover (38)
H(ei) = p - m
2 1(p
+ 3a) = p - 2(m -
1)v ,
denoting the antiholomorphic sectional curvature. Hence Theorem 17. Let M be an almost Hermitian manifold with constant holomorphic sectional curvature p and (pointwise) constant type a, and let M be
469
ALMOST HERMITIAN MANIFOLDS
a complex Einstein hypersurface satisfying condition (17). If p is the Ricci curvature of M, then
(i) P<'s(m+ 1)/i+ 2(m- 1)a,
(ii) there exists a basis lei, Jei} as in Lemma 6 such that A' = 21 where
421 =(m+1)f +3(m-1)a-2p, (iii)
at each point of M there exists a holomorphic plane whose sectional
curvature is p - 2(m - 1)v where v = fe + 3a. It follows further from (37) that if 2i = 0 at one point, then 2i = 0 on the whole manifold M (supposed to be connected). Hence Theorem 18. Theorem 15 (iii) may be replaced by
(iii) M is an Einstein manifold, and at one point of M all sectional curvatures of M satisfy (27) or (28). Theorem 16 (iii) may be replaced by
(iii) M is an Einstein manifold, and at one point of M all sectional curvatures of M satisfy (31). In relation with Theorem 15 and formula (32) it is easy to verify Theorem 19. Let M be a complex hypersurface of an RK-manifold M with constant holomorphic sectional curvature u and constant type. If M is of complex dimension > 2 satisfies condition (17) and is totally geodesic in M, then the following statements are equivalent : (i) the antiholomorphic sectional curvature of M (or of k on M) is zero,
(ii) k=konM,
(iii) k(x, x) = 1 for g(x, x) = 1, where k (resp. k) denotes the Ricci tensor of M (resp. M). 4. Let a (resp. a') be a holomorphic 2-plane defined by the unit vector x (resp. y). Then the holomorphic bisectional curvature H(a, a') is defined by
[1]
H(a, a') = R(x, Jx, y, Jy)
(39)
It is easy to verify that H(a, a) depends only on a and (40)
Using (6) we obtain
H(a, a) = R(x, Y, X, Y) + R(Jx, Y, Jx, y) - 2(x, y) - 2(Jx, y)
which together with (7) gives Theorem 20. Let M be an almost Hermitian manifold, and a (resp. a') a holomorphic 2-plane in m E M defined by a unit vector x (resp. y). Then
H(a, a) =
6{Q(x + Jy) + Q(x - Jy) + Q(x + Y)
+ Q(x - y) - 4Q(x) - 4Q(y)}
(41)
-
{2(x, Y) + 2(Jx, JY) + .l(Jx, Y) + 2(x, Jy)} . 8
470
LIEVEN VANHECKE
If g(x, Jy) = cos B and g(x, y) = cos 0, then
(42)
H(a, a') = -{(1 + cos B)'H(x + Jy) + (1 - cos 0)2H(x - Jy) + (1 + cos ¢)2H(x + y) + (1 - cos ¢)2H(x - y)
- H(x) - H(y)} - s{2(x, Y) + A(Jx, Jy) + A(Jx, Y) + 2(x, Jy)} .
Using (12) we obtain Theorem 21. Let M be an RK-manifold with pointwise constant holomorphic sectional curvature la and pointwise constant type a. Then (43)
(44) (45)
H(a, a') = 2 (a - a) + I (a + a) (cos' B + cos' ¢)
2(p-a)
if p+a>0, if p+a<0.
The value 2(p - a) is attained when a is perpendicular to a', whereas the value p is attained when a = a'. In [7] we proved the following proposition. Proposition 22. Let M be an RK-manifold with pointwise constant type a. Then M is a space of constant curvature if and only if the holomorphic sectional curvature is equal to a. From this and Theorem 21 we obtain Theorem 23. Let M be an RK-manifold with pointwise constant type and pointwise constant holomorphic sectional curvature. Then M is a space of constant curvature if and only if the holomorphic bisectional curvature H(o-, o') vanishes when a is perpendicular to oa'.
From (43) it follows that H(a, o') is constant at a point m if p + a = 0, and we see in consequence of (9) and (12) that this means that the antiholomorphic sectional curvature is equal to - 2p. Hence Theorem 24. Let M be an RK-manifold as in Theorem 23. Then the holomorphic bisectional curvature is constant at a point m of M if and only if the antiholomorphic sectional curvature is - 2p where p is the holomorphic sectional curvature at m (or if and only if the holomorphic bisectional curvature is equal to the holomorphic sectional curvature). Finally, consider again an almost Hermitian manifold M, and let M be a complex hypersurface which satisfies condition (17). Then we obtain, from (20) and (18), (46)
R(x, Jx, y, Jy) = R(x, Jx, y, Jy) + 2g(Ax, y)2 + 2g(Ax, JY)'
,
which proves Theorem 25. Let M be an almost Hermitian manifold, and M a complex hypersurface satisfying condition (17). Then the holomorphic bisectional cur-
ALMOST HERMITIAN MANIFOLDS
471
vature of M does not exceed that of M. In particular, if M is a complex Euclidean space, then the holomorphic bisectional curvature of M is nonpositive.. Bibliography [1] [2] [3]
S. I. Goldberg & S. Kobayashi, Holomorphic bisectional curvature, J. Differential Geometry 1 (1967) 225-233. S. Sawaki & K. Sekigawa, Almost Hermitian manifolds with constant holomorphic sectional curvature, J. Differential Geometry 9 (1974) 123-134.
B. Smyth, Differential geometry of complex hypersurf aces, Ann. of Math. 85 (1967) 246-266.
[4] L. Vanhecke, Mean curvatures for holomorphic 2p-planes in some almost Hermitian manifolds, Tensor 30 (1976) 193-197.
[5] -, Mean curvatures for antiholomorphic p-planes in some almost Hermitian.
[6] [7]
manifolds, Kodai Math. Sem. Rep. 28 (1976) 51-58. , Almost Hermitian manifolds with J-invariant Riemann curvature tensor, Rend. Sem. Mat. Univ. e Politec. Torino 34 (1975-76) 487-498. , The axiom of coholomorphic (2p±1)-spheres for some almost Hermitian manifolds, Tensor 30 (1976) 275-281. CATHOLIC UNIVERSITY OF LOUVAIN HEVERLEE, BELGIUM
J. DIFFERENTIAL GEOMETRY 12 (1977) 473-480
TOTALLY REAL SUBMANIFOLDS BANG-YEN CHEN, CHORNG-SHI HOUR & HUEI-SHYONG LUE 1.
Introduction
Among all submanifolds of a Kaehler manifold M, there are two typical classes : one is the class of complex submanifolds and the other is the class of totally real submanifolds. A submanifold M of a Kaehler manifold k is said to be complex (resp. totally real) if each tangent space of M is mapped into itself (resp. the normal space) by the complex structure of M. In [3] Chen and Ogiue studied some fundamental properties of totally real submanifolds. In particular, some characterizations of totally real submanifolds and some classifications of totally real submanifolds in complex space forms are obtained [3]. (See also [7]). In this paper, we shall continue to study fundamental properties of totally real submanifolds. In particular, we shall obtain two reduction theorems for totally real submanifolds in complex space forms and also study totally real submanifolds with parallel mean curvature vector. 2.
Basic formulas
Let M2 be a 2m-dimensional' Kaehler manifold with complex structure J and metric tensor g. Let 17 (resp. R) be the Levi-Civita connection (resp. the We denote by F (resp. R) the induced Levi-Civita curvature tensor) of connection (resp. the curvature tensor) of an n-dimensional totally real submanifold M. Then the second fundamental form o of the immersion is given by
o'(X, Y) = PxY - FxY,
(2.1)
etc. are vector fields in M. The mean curvature vector H is then given by H = (1 In) Tr v. For a normal vector field , we write where X, Y,
,
Fx = -AFX + Dx ,
(2.2)
denotes the tangential (resp. normal) component Then we have g(a(X, Y), ) = g(AFX, Y). Let RD denote the curva-
where -AFX (resp. of
Communicated by K. Yano, August 4, 1975. The first author was partially supported by NSF Grant GP-36684. 1 In this paper we consider only real dimensions of manifolds.
474
BANG-YEN CHEN, CHORNG-SHI HOUH Sc HUEI-SHYONG LUE
ture tensor of the normal connection D, i.e., RD(X, Y) = D1Dy - DYD1 D[x,y . Then the Gauss, Codazzi and Ricci equations are given respectively by
g(R(X, Y)Z, W) = g(R(X, Y)Z, W) + g(a(X, Z), a(Y, W))
(2.3)
- g(a(Y, Z), a(X, W)) , (R(X, Y)Z)' = (I1a)(Y, Z) - (I ya)(X, Z)
(2.4)
(2.5)
g(R(X, Y)e, Y)) = g(RD(X, Y)e, Y))
-
A,l (X), Y)
where (Vxa)(Y, Z) = Dxa(Y, Z) - a(F,Y, Z) - a(Y, F,Z), (R(X, Y)Z)`" is the normal component of A(X, Y)Z, and , Y), , etc. are normal vector fields of M in M2'a.
A Kaehler manifold MZ"t is a complex space form of constant holomorphic sectional curvature c, denoted by Ml-(c), if its curvature tensor A satisfies
R(X, Y)Z = 4 {g(Y, Z)X - g(X, Z)Y + g(JY, Z)JX
(2.6)
- g(JX, Z)JY + 2g(X, JY)JZ} , where X, Y, Z, , etc. are vector fields in k2"L. If the ambient space k1is a complex space form 12'"(c), then (2.3), (2.4) and (2.5) reduce respectively to g(R(X, Y)Z, W) = g(a(X, W), a(Y, Z)) - g(a(X, Z), a(Y, W)) (2.7) (2.8)
+ 4 {g(X, W)g(Y, Z) - AX, Z)g(Y, W)} ,
(Ix6)(Y, Z) = (Pya)(X, Z) g(RD(X, Y)e, Y)) = g([A,, A,l (X), Y)
(2.9)
+ 4 {g(JY, )g(JX, i1) - AM )g(JY, 01 .
A normal vector field is called a parallel section in the normal bundle TIM if D = 0. A unit normal vector field is called an isoperimetric section if Tr A, is constant. A subbundle Q of the normal bundle TI(M) is holomorphic if Q is invariant under J, i.e., if JQ C Q. A subbundle Q of T1M is said to be parallel if Q is invariant under parallel translation, i.e., if Dx is also a section in Q for every local section in Q. It is clear that a unit normal vector field is parallel if and only if the line bundle generated by is parallel. For a subbundle Q of T1M, there exists a unique subbundle Q° of T1M such that Q and Q° are orthogonal and Q O Q° = T'M. We called Q° the complementary subbundle of Q. It is clear that for a totally real submanifold M in M, the complementary subbundle (J(TM))° of J(TM) is always holomor-
TOTALLY REAL SUBMANIFOLDS
475
phic, and a subbundle Q is parallel if and only if its complementary subbunble Q° is parallel. The complementary subbundles of holomorphic subbundles of
T'M is called a coholomorphic subbundles of T'M. It is clear that a subbundle Q of T'M is coholomorphic if and only if Q is the direct sum of J(TM) and a holomorphic subbundle of T'M. 3.
Reduction theorems
Let M be an n-dimensional totally real submanifold of a 2m-dimensional Kaehler manifold MZ"t. If there exists a 2r-dimensional parallel holomorphic subbundle Q of T'M, then for any section in Q and vector fields X, Y in M we have g(o'(X, Y), E) = g(rxY, E) = g(IxJY, JE) g(DxJY, JE) = -g(JY, DXJE) = 0 , from which we see that 64IQ (restriction of oy to Q) vanishes. Thus we have Lemma 1. Let M be an n-dimensional totally real submanifold of a 2mdimensional Kaehler manifold MZ"t. If Q is a 2r-dimensional parallel holomorphic subbundle of T'M, then or IQ - 0.
Let Q be a holomorphic subbundle of T'M. Then the coholomorphic subbundle Q` contains J(TM) as its subbundle and Q is parallel if and only if Q° is parallel. Hence from Lemma 1 we obtain Lemma 2. Let M be an n-dimensional totally real submanifold of a 2mdimensional Kaehler manifold MZ"t. If Q is a parallel coholomorphic subbundle of T'M, then Im 6r C Q, where Im u = {or(X, Y) : X, Y E TM}.
In particular, if the ambient space is a complex space form, then from Lemma 2 we have Theorem 1. Let M be an n-dimensional totally real submanifold of a 2mdimensional complex space form M2m(c). Then M is a totally real submanifold of a 2(n + s)-dimensional totally geodesic complex submanifold M'1`11(c) of M2m(c) if and only if there exists an (n + 2s)-dimensional parallel coholomorphic subbudle of T'M. The "only if" part is trivial, and the "if" part follows from Lemma 2 and an argument similar to the proof of Proposition 9 given in [1] with only slight modifications. In the following theorem, we shall give some necessary and sufficient conditions for the codimension of totally real submanifolds which can be reduced to minimal one. Theorem 2. Let M be an n-dimensional totally real submanifold of a 2mdimensional complex space form M2m(c). Then M is contained in a 2n-dimensional totally geodesic complex submanifold MZ"(c) of M'-(c) if and only if one of the following statements holds:
476
BANG-YEN CHEN, CHORNG-SHI HOUH & HUEI-SHYONG LUE
(i ) J(TM) is parallel. (ii) Im a C J(TM). (J(TM))c is parallel. Proof. Since J(TM) is coholomorphic, Theorem 1 implies that M is a totally real submanifold of a 2n-dimensional totally geodesic complex submanifold M2n(c) of Mz" (c) if and only if statement (i) holds. Hence it suffices to prove the equivalences of (i), (ii) and (iii). The equivalence of (i) and (iii) are clear. Moreover, Lemma 1 shows that (iii) implies (ii). Thus the theorem follows from Lemma 3. "Im a C J(TM)" implies "(J(TM))c is parallel." Proof. Let g be any section of the holomorphic subbundle (J(TM))c and Ima C J(TM). Then we have aI(J(TM)) = 0 and hence (iii)
0 = g(G(X, Y), JO) = g(GY, Je) = g(FxJY, -g(DxJY, ) = g(JY, Dxe)
Since this is true for all X and Y E TM, (J(TM))c is parallel. q.e.d. As an application of Theorems 1 and 2, we have the following. Corollary 1. Let M be an n-dimensional totally real, totally umbilical submanifold (n > 2) of a 2m-dimensional complex space form M2" (c), c 0.
(i) If H - 0, then M is contained in a 2n-dimensional totally geodesic complex submanifold M2n(c) of M2-(c). (ii) If H Erz 0, then M is contained in a 2(n + 1)-dimensional totally geo-
desic complex submanifold kI(I`)(c) of k'-(c). Proof. Since M is totally real and totally umbilical in M2" (c), Theorem 1 of [4] implies either (1) M is totally geodesic in M2-(c) or (2) the mean curvature vector H is nonzero and parallel. If Case (1) holds, then Theorem 2 shows that M is contained in a 2n-dimensional totally geodesic complex submanifold M27L(c) of M2-(c). If Case (2) holds, then H and JH span a holomorphic plane subbundle of T'M, say, V. From (2.9) of Ricci, we find that V is perpendicular to J(TM). Hence J(TM) O+ V is an (n + 2)-dimensional coholomorphic subbundle of T'M. Now, since H is parallel, for any vector fields X, Y in M we have (3.1) (3.2)
JFxY + Jc(X, Y) _ -AJY(X) + Dx(JY) -AJH(X) + Dx(JH) _ -JAH(X)
,
On the other hand, by the total umbilicity of M we find a(X, Y) = g(X, Y)H. Thus (3.1) and (3.2) give (3.3)
Dx(JH), D1(JY) E J(TM) O+ V
From these, we see that the subbundle J(TM) O+ V is a parallel coholomorphic
subbundle of T'M. Hence, by Theorem 1, M is contained in a 2(n + 1)-dimensional totally geodesic complex submanifold M2cn+"(c) of M'-(c).
TOTALLY REAL SUBMANIFOLDS
4.
477
Submanifolds with parallel mean curvature vector
In this section we shall assume that M is a 2n-dimensional Kaehler manifold MIT. Then from (2.1) and (2.2) we immediately have (4.1)
JA,YX = 6(X, Y)
D1(JY) = J171Y ,
JDx
(4.2)
From the first equation of (4.1) we obtain
RD(X, Y)(JZ) = JR(X, Y)Z
(4.8)
Moreover, (4.1) implies Lemma 4. Let M be an n-dimensional totally real submanifold of a 2ndimensional Kaehler manifold Men. Then the normal bundle TLM admits a parallel nontrivial (local) section if and only if the tangent bundle admits a parallel nontrivial (local) section. Remark 1. Lemma 4 implies that the sectional curvature of every plane section containing J vanishes if is parallel. In particular, if M is of constant sectional curvature, then the normal bundle admits no nontrivial parallel sections unless M is flat. In the following, we shall assume that the ambient space Al", is a complex space form M2"(c). If is a parallel section in TLM, then (2.8) of Codazzi reduces to
(FxA,)Y = (FyA,)X.
(4.4)
, Let e1, , en, 1, , be a local field of orthonormal vectors in M2'n(c), , e,,, are tangent to M (and hence 1, ,n defined along M, such that e1,
are normal to M). We put hi; = g(Akei, e;), where Ak = A,,,. Let Ki; denote the sectional curvature of the plane section 7r(ei, e;). Then K1 = g(R(ei, e;)e;, ei). By (2.7) of Gauss, (4.4) of Codazzi and an argument similar to Smyth [6] we may prove Lemma 5. Let M be an n-dimensional totally real submanifold of Men(c). If M admits a parallel isothermal section , then (4.5)
24(TrAD = Ei
, e, orthonormal eigenvectors of A, with where 4 is the Laplacian on M, e1, A,,,, and Ki; the sectional curvature of the plane section eigenvalues 2, 7r(ei, e;).
If M is compact and of nonnegative sectional curvatures, then Lemma 2, together with Hopf's lemma, gives (4.6)
Ki; +
(hi;)2}(A, - 21j) = 0
i
1
478
BANG-YEN CHEN, CHORNG-SHI HOUR & HUEI-SHYONG LUE
VAe = 0 .
(4.7)
Without loss of generality we may assume that
21 = ... _ 2vl = P1 *I
2v1+v2 = P2
2V1+...+Vr-i+1 = ... = 2n = Pr ,
, p,r are all distinct. We put
where p1i
[Pt] _ {v1 + ... + vt_1 + 1, ..., v1 + ... + vt_1 + vt} From (4.6) it follows that (4.8)
hk5 = Kij = 0
,
k = 1, ... , n, i E [Pt], j
(4.9)
Ak = I
0
[0
s
, enof Ae we find
Thus with respect to the eigenvectors e1i
[B;
{p81, t
0 ... 0 B2 .
.
0
0 Bkj
where Bk is a vt x vt-matrix. Now let us consider the orthonormal frame field given by (4.10)
e1, ... , en, Jet = S1, ... , Je. = S. ,
, e,,, are still eigenvectors of A,. With respect such frame fields, Ak's still have the forms (4.9). Moreover Ak's satisfy [3] where e1i
(4.11)
hjk=hkj=hik,
i,f,k=1,...,n.
By using (4.8) and (4.11) we obtain Lemma 6. Let M be a compact totally real submanifold in M2n(c). If M has nonnegative sectional curvature and admits a parallel isoperimetric section then with respect to the frame field (4.10), Ak's are given in the following forms :
0
479
TOTALLY REAL SUBMANIFOLDS
v,+ ... +vs_,GkGv,+... +v,,1GsGr, where B3 is a v, x v8-matrix. By using Lemma 6 we may prove Theorem 3. Let M be a compact n-dimensional manifold with nonnegative sectional curvature immersed in a 2n-dimensional complex space form M2n(c) as a totally real submanifold. Then the normal bundle TJ-M admits no parallel isoperimetric section unless MZn(c) is flat. Proof. We assume that $ is a parallel isoperimetric section. We shall divide the proof into two cases. Case a. 2, = ... = 2n = 2. In this case, M is umbilical with respect
Hence we may choose e , en in such a way that Je, = . Lemma 4 then implies that K,; = 0 for j > 1. On the other hand, (2.7) of Gauss gives K,; = c/4 + Zx {h;,h;; - (hl;)'}. Thus by (4.11) we find c = 0. Case b. .l, # 2i for some i. In this case, (4.8) implies K,; = 0 for j e [.ii]. On the other hand, Lemma 6 gives Zk {h;,h;j - (hl;)2} = 0. Thus (2.7) of Gauss implies c = 0. Corollary 2. Under the hypothesis of Theorem 3, if the mean curvature vector H is parallel, then either (i) c > 0 and M is minimal in MZn(c) or (ii) M2n(c) is flat, i.e., c = 0. Proof.
Since Tr AH = IH I2, the parallelism of H implies that either H = 0 is a parallel isoperimetric section. If H = 0, then M. is minimal in MZn(c), and the sectional curvatures of M is G 4c. Thus by the hypothesis we have c > 0. If H # 0, and H/CHI is an isoperimetric section, then Theorem 3 implies that k111(c) is flat. Remark 2. If M2n(c) is flat, then there exist compact submanifolds of MZn(c) which satisfy the assumptions of Theorem 3 and also admit parallel isoperimetric section. For example, let S' be a unit circle in the complex plane C'. Then S' X S' is a such totally real surface in C2. In view of Theorem 3, it is interesting to study totally real submanifolds or H / I H I
of the complex number space Cn which admits a parallel isoperimetric section. The proofs of the following two theorems are similar to that of Theorem 2 in [6]. So we just only give the necessary outlines of the proofs. Theorem 4. Let M be a compact n-dimensional totally real submanifold
imbedded in C. If M has nonnegative sectional curvature, and it admits a parallel isoperimetric section , then M is a product submanifold M, x . X MT, where M, is a compact v,-dimensional totally real submanifold imbedded in some Cvt, and Mt is contained in a hypersphere of Outline of proof. The assumption of the theorem implies that PAF = 0. C2t.
Thus the distinct eigenspaces T , TT of A, define parallel distributions of M. By the de Rham decomposition theorem, M is a product of Riemannian X MT, where the tangent bundle of MS corresponds to T. manifold M, x By Lemma 6 and a lemma of Moore [5] we see that M = M, x x MT
480
BANG-YEN CHEN, CHORNG-SHI HOUH & HUEI-SHYONG LUE
is a product submanifold imbedded in Cn = C"1 X X C. Moreover, Lemma 6 implies that each of Mt's is a totally real submanifold imbedded in some Cvt :
M, X ... X M,
imbedding
r
3,
Cv1 X ... X Cvr
is a parallel Let be the component of in the subspace C. Then normal section of M, in Cv,, and Mt is umbilical with respect to art( ). From these it follows that Mt is contained in a hypersphere of Cv° (see, for instance, [2])
Theorem 5. Let M be a compact n-dimensional totally real submanifold imbedded in Cn. If M has nonnegatve sectional curvature and parallel mean x M,r, where curvature vector H, then M is a product submanifold M, x Mt is a compact vt-dimensional totally real submanifold imbedded in some Cvt, and Mt is also a minimal submanifold of a hypersphere in Cvl. Outline of proof. Since the mean curvature vector H is parallel and there exists no compact minimal submanifold in Cn, H/IHI is a parallel isoperimetric X M,r such section. By Theorem 4, M is a product submanifold M, x
that Mt is totally real in some Ct't and Mt is umbilical with respect to the component 7rt(H) of H in the subspace Cvl. Since 7rt(H) is parallel and is the mean curvature vector of Mt in Cvt, Mt is a minimal submanifold of a hypersphere in Cv,. References T. E. Cecil, Geometric applications of critical point theory to submanifolds of complex projective space, Nagoya Math. J. 55 (1974) 5-31. [ 2 ] B. Y. Chen, Geometry of submanifolds, M. Dekker, New York, 1973. [ 3 ] B. Y. Chen & K. Ogiue, On totally real submanifolds, Trans. Amer. Math. Soc. 193 (1974) 257-266. , Two theorems on Kaehler manifolds, Michigan Math. J. 21 (1974) 225[4] [1]
229.
[51 J. D. Moore, Isometric immersions of Riemannian products, J. Differential Geometry 5 (1971) 159-168. [6] B. Smyth, Submanifolds of constant mean curvature, Math. Ann. 205 (1973) 265-280.
[71 K. Yano, Totally real submanifolds of a Kaehlerian manifolds, J. Differential Geometry 11 (1976) 351-359.
MICHIGAN STATE UNIVERSITY WAYNE STATE UNIVERSITY NATIONAL TSINGHUA UNIVERSITY, TAIWAN
J. DIFFERENTIAL GEOMETRY 12 (1977) 481-491
GEOMETRY OF HOROSPHERES ERNST HEINTZE & HANS-CHRISTOPH IM HOF
1.
Introduction
Let M be a Hadamard manifold, i.e., a connected, simply connected, complete riemannian manifold of nonpositive curvature. To be more precise, as-
sume that the sectional curvature K of M satisfies -b2 < K < -a2, where 0 < a < co and 0 < b < oo. If p E M and z is a point at infinity (cf. EberleinO'Neill [4], which we give as a general reference for Hadamard manifolds), there exists a horosphere through p with center z. This is defined as follows : Denote the geodesic ray from p to z by r, and consider the geodesic spheres through p with center 7(t), t > 0. As t goes to infinity, these spheres converge to the horosphere. More precisely, the horospheres are the level surfaces of the Busemann function F = lim F, where Ft is defined by F,(p) = d(p, 7(t)) - t. In the flat case (a = b = 0), horospheres are just affine hyperplanes, and in the case of constant negative curvature, using the Poincare model we see that horospheres are euclidean spheres internally tangent to the boundary sphere, minus the point of tangency. The main purpose of this paper is to
show that, to a certain extent, the geometry of horospheres in M may be compared with that in the spaces of constant curvature - a2 and -b2, respectively. We give two examples : 1. (Theorem 4.6). If ° is a horosphere and h denotes the distance in
° with respect to the induced metric, then for all p, q E Ye
a sinh ad(p, q) < h(p, q) < b sinh b d(p, q) , 2
2
where d is the distance function of M. 2. (Theorem 4.9). If r is a geodesic tangent to a horosphere X', and if p, q are the projections of r(± co) onto , then
b
In particular, we are obliged to H. Karcher for suggesting us a method to prove Theorem 4.6.
482
ERNST HEINTZE & HANS-CHRISTOPH IM HOF
We recall that we allow a = 0 and b = oo, corresponding to the possibilities that there is only one or no curvature bound beside K G 0. In these cases all our inequalities have to be interpreted in the obvious way, and eventually become meaningless, e.g., h(p, q) < 2/a for a = 0. Here we give a brief account of the content of the paper. In § 2 we prove a comparison theorem for stable Jacobi fields, which is crucial for all the following estimates. Moreover, this theorem is of its own interest, especially because in contrast to former authors ([5, p. 1171, [1, p. 132]) we get the optimal bounds. § 3 is concerned with the C2-differentiability of Busemann functions. Combining this with the comparison theorem for stable Jacobi fields, we get estimates for the differential of the flow which moves M towards a given point at infinity. In § 4, the main part of the paper, we formulate and prove a series of geometric comparison theorems involving distances on horospheres and their relationship to geodesics, as indicated by the examples above. Notation. All nontrivial geodesics are assumed to have unit speed. For p, q E M we denote by rpq the unique geodesic from p to q, and by d(p, q) the distance between p and q. If m denotes a point different from p and q, then
< (p, q) denotes the angle in [0, ir] subtended by r,np(O) and r,nq(0). Let M(me) denote the points at infinity. If P E M and z r= M(oo), there exists a unique geodesic from p to z, denoted by rp=. Therefore the definition of angles
Stable Jacobi fields
Definition 2.1. Let r : [0, co) --> M be a geodesic ray, and Y a Jacobi field along r. Y is said to be stable if II Y(t) II is bounded for t > 0. Lemma 2.2 Let r : [0, c) --> M be a geodesic ray, and let V E Mp, p = 7(0). Then there existis a unique stable Jacobi field Y along r with Y(0) = v,
and we shall denote this stable Jacobi field by Y. Proof. (i) Uniqueness follows immediately from the fact that in a Hadamard manifold the length of a Jacobi field is a convex function. (ii)
Denote by Y, the unique Jacobi field along r with YJ0) = v and
Y,(n) = 0. Applying Rauch's comparison theorem to Y,,, with the flat case), we get II Yn(0) -
it <
t
(comparison
li Ye(t) - Y..(t) I
Now by the convexity argument above, II Ym(t) is monotone decreasing in the
interval [0, m], so that
IIYn(0)-Y1.(0)II<1IvII, n
for n <m.
GEOMETRY OF HOROSPHERES
483
Thus {Y' (0)} is a Cauchy sequence with limit w, say. If Y denotes the Jacobi field along r with Y (0) = v and Yv(0) = w, it follows immediately that Y,,, as the limit of the Y,a, is stable. This completes the proof of the lemma. Proposition 2.3. Let Y be a perpendicular Jacobi field along the geodesic r : R --> M with Y(O) = 0. Assume as usual that the curvature K of M satisfies
-b2
A proof, using an idea of Bishop-Crittenden [2], may be found in Im HofRuh [6]. The assumption of positively pinched curvature in [6] is not essential. The proof may be changed in an obvious way by replacing sin and cos by sinh and cosh, respectively. As a consequence we get Theorem 2.4 (Comparison theorem for stable Jacobi fields). Let r : [0, 00) M be a geodesic ray, and Y2 the stable Jacobi field along r with r(O). Then v E M,(u) , v I
IvIe-bt
< 11Yv(t)II <
Iv11e-at
Proof. If Y. denotes the Jacobi field along r with Yn(O) = v and Yn(n) = 0 as above, then Y,a(t) -> Yv(t) for fixed t. By the last proposition (applied
to Z,a(t) = Y,,(n - t)), we get
sinh b(n - t) < II Y,a(t)1i < sinh a(n - t) sinh bn
I Y,(0)11 -
sinh an
for 0 < t < n. Thus e-bt
sinh a(n - t) = e- at = lim sinh b(n - t) < II Y,,(t) II < lim sinh an n-sinh bn I v Il
This completes the proof. 3.
Radial fields and radial flows
In this section we fix a point z e M(oo) and consider the corresponding radial field Z defined by Z(p) = pz(O). It is said to be radial in analogy to the radial field Zq, which is given by Zq(p) = j'pq(0) for a fixed point q E M and
p# q
In the following we will strongly need that Z is continuously differentiable, a fact which has been proved by P.Eberlein in an unpublished paper [3]. For the convenience of the reader we give here a new proof, which is also con-
siderably shorter. It is interesting to note that L. Green [5, p. 118] could show that actually Z is of class C2, provided VR is bounded and the curvature is strictly 4-pinched.
484
ERNST HEINTZE & HANS-CHRISTOPH IM HOF
Proposition 3.1 (Eberlein [3]). Let M be a Hadamard manifold, Z the radial field in direction of z e M(oo), and F a Busemann function at z. Then Z
_ -grad F, Z is C', and V,;Z = Yv(0) for all v e MP, where p e M, and Y v.
is the stable Jacobi field along rPz with
The basic idea of the proof, going back to P. Eberlein, is to carry over statements for radial fields in the direction of finite points to the given field Z. If
q e M, and Z' is the corresponding radial field, then we have Z4 = -grad Fq, where F' denotes the distance to q. Now if v e MP, an easy 2-parameter variation argument shows F,Zq = - V grad F4 = Y'(0)--, where Y is the Jacobi field along rPq with Y(O) = v, Y(d(p, q)) = 0, and L denotes the component orthogonal to tPq(0). If q is replaced by a point at infinity, Fq has to be replaced by the Busemann function F, and Y by the stable Jacobi field "vanishing at z". Proof. Let r be a geodesic with r(co) = z, and p.,, = r(n), n e N. If F,2 is
defined by Fn(p) = d(pn, p) - n, then F = lim F. is a Busemann function with respect to z, and Z,n = - grad F.2 is the radial field in the direction of p.. Z,2 is defined and C°° on M - {pj. We will show (i) the fields Z. converge uniformly on compact sets to Z, and (ii) for any vector field V on M the covariant derivatives FVZ,2 converge uniformly on compact sets to Yv, where Yv(p) = Y;.(0) and Yv is the stable Jacobi field along rPz with Y (0) = v = V(p). This proves the uniform convergence of the first and second derivatives of the functions F,n on compact sets. Thus F is C2, grad F = lim grad F. _
-17 grad F = -lim w grad Fn = Y'(0).
-Z, Z is Cl, and
Let K C M be compact and n, E N, such that p,n K for all n > n0. (i) Let p e K and n > no. Then IIIZ,n - ZHI (p) = 1t,,.(0) - tPz(0)JI goes
to zero uniformly on K, if the angles
Next, consider for p e K and n > n, VvZ,
- Yv I (p) = Fv grad F,n - Y'(0) = Yp-L(0) - Y,(0) I
where v = V(p), YPP,: is the Jacobi field along lPPn with Y,,,(0) = v and YPP (d(p, p.n)) = 0, and L denotes the component orthogonal to also depends on n. But an easy computation shows 11 Y,
(0)
- YPPn(0) = 11 V - vj-11 < d(p, P-..)
I1 v I
which
I
d(P, Pa)
which goes to zero uniformly on K as n tends to infinity. Thus it is enough to show
Y' .(0) - Yv(0)1 - 0 uniformly on K. For T > 0 let X Pn be the
Jacobi field pp along rPPn with TPn(0) = V(p), X(T) = 0, and define XP analogously. Then
GEOMETRY OF HOROSPHERES
485
I YPP.(0) - Yv(0) II < II YPp'(0) - XppJ0) II + II XT"(0) - XP (0) II
+ II XT'(0) - yv(0) By Rauch's comparison theorem II Y,p (o) - XPPn(o) II <
T
II Yppn(T) II < T II V(p) II
if n is sufficiently large so that d(p, p,,,) > T. The same argument yields Xpi(0) - Yv(0) II 5 I V(p) T. Thus the problem is reduced to show that, for fixed T, the difference Xpz(0) II goes to zero uniformly on K, I
as n tends to infinity. Using a lower curvature bound on KT = {p E M I d(p, K) < T} it is clear that d(q,(p), q(p)) -- 0 uniformly on K, where qn(p) = rpp, (T), q(p) = rpz(T). By the differentiable dependence of Jacobi fields and their derivatives on the boundary values, the result now follows. The radial flow. Now we want to study the flow generated by the vector field Z, which we call the radial flow (with respect to a fixed z E M(oo)) and
denote by * or {*J. Since the geodesics going to z are the integral curves of Z, this vector field i s obviously complete, and Jr is given by Jr _ ;70 (P o (1 R X Z) :
R X M -* M, where (P denotes the geodesic flow, and 'r the canonical projection. Proposition 3.1 implies immediately that i is C'. The following properties of Ia* are infinite versions of the lemma of Gauss and the comparison theorem of Rauch. Proposition 3.2. (i) If a vector u E MP is parallel to Z(p), then i,,k(u) is parallel to Z(* ,(p)), and III,*(u)II = IIuII (ii) If a vector v E MP is orthogonal to Z(p), then i,,k(v) is orthogonal to Z(* ,(p)), and the following inequalities hold vi
e-" < I*,,(v)II < IIv1I e a`
fort>0.
Proof. (i) It is enough to show iJr (Z(p)) = Z(* ,(p)), but this is true, since the geodesics going to z are the integral curves of Z.
We recall that Z = -grad F, where F is a Busemann function at z, and that the horospheres centered at z are the level surfaces of F. Therefore the complements MP = {v E MP v L Z(p)} are the tangent spaces of the horospheres, and , maps horospheres onto parallel horospheres. This implies the first part of (ii). In order to prove the inequalities, we now compute ]t,K(v) for v E MP explicitely. By definition *,,(v) = 7r* o (pt,k o Z,F(v). We use the (ii)
identification TSM = SM +Q TM Q+ TM given by 7rs X 'r,F x K, where SM de-
notes the unit tangent bundle, 'rs : TSM SM is the canonical projection, rr,: TSM - TM is the differential of 7r: SM M, and K : TSM TM is the connection map. Then Z,k(v) = (Z(p), v, V7,Z) = (Z(p), Y2(0), Y;;(0)), where Yv is the stable Jacobi field along rpz with initial value Y,(0) = v, (compare Proposition 3.1). Therefore we get coz oZ*(v) = (Z(*,(p)), Y7,(t), Y'(t)) and
486
ERNST HEINTZE & HANS-CHRISTOPH IM HOF
*t,(v) = Y,,(t). By the comparison theorem for stable Jacobi fields we conclude
IIv1Ie-ac < Il*tx(v)II <
4.
IIvIIe-ac
Distances on horospheres
Generalities. Since Busemann functions, and therefore horospheres, are at least C2, the notions of distance and geodesic curves are defined with respect to the induced metric. As level surfaces of a Busemann function, horospheres are closed and therefore complete ; in particular, we always have minimal geodesics joining two points. In the case of constant negative curvature horospheres are flat, but in the other symmetric spaces of rank 1 and negative curvature this is no longer true. In these spaces horospheres may be represented as nonabelian nilpotent Lie groups with a left invariant metric, and therefore have curvatures of both signs (J. Wolf [9]) and even conjugate points (J. O'Sullivan [7]). In the following we will estimate some distances on horospheres arising in special geometric situations. Still assuming -b2 < K < -a2, we will use as comparison manifolds the spaces Ha and Hb of constant curvature -a2 and - b2, respectively. Two asymptotic geodesics. Let r0 be a geodesic, and denote by _Vt the horosphere trhough ro(t) with center r0(cc). Obviously we have _Vt = *t( where *t is the radial flow in the direction of r0(co). Now consider an asymptotic geodesic rl, and choose the origin rl(O) on moo. Then r,(t) E her, and we can define h(t) to be the fit-distance of ro(t) and rl(t). As a first application of Proposition 3.2 we give an estimate for h(t). Propositson 4.1. For t > 0 we have h(0)e-bt < h(t) < h(0)e-at
be a minimal °o geodesic joining r,(0) and r,(0). Proof. Let uo : [0, 11 Then ut = *t cu, is a curve on _Vt from ro(t) to rl(t), and we have h(t) < l(ut) = f'o II ut1I = f 1
< e-111 f,
e-ath(0)
.
0
0
The proof of the inequality on the left hand side is similar. Remark. Combining the above result with Theorem 4.6 below we immediately get, for d(t) = d(r0(t), r1(t)) and t > 0,
(2 aresinh 2 h(O)e-11 < d(t) < h(0)e-at As H. Karcher remarked, this can be improved by a different method to
d(0)e-"' < d(t) < (2 sinh
2
d(0) Ie-at
487
GEOMETRY OF HOROSPHERES
Two estimates for the Busemann function with geometric applications. We
consider a Busemann function F at an infinite point z. To compare F with Busemann functions in spaces of constant curvature, we study the restriction f = F o r for a given geodesic r. While f measures the deviation of r from a fixed horosphere with center z, the derivative grad F> measures the angle between r and the horospheres centered at z. In the following, f,, and f b denote functions defined analogously in the spaces Ha and Hb, respectively. Lemma 4.2. Given that f , la, f b are as described above. Assume f (O)
f .(O) = f b(0) and f'(0) = f a'(0) = f(0). Then f' (s) < f'(s) < f '(s) for s > 0 and f a(s) < AS) < f b(s) for s E R. Proof. For s > 0 consider the triangle 4 determined by p = r(0), q = r(s)
and z. The angles a = p(q, z) and p = 1 1(p, z) satisfy cos a = - f'(0) and cos R = f'(s). Let a be the geodesic ray from p to z, and denote by 4(t) the triangle determined by p, q and 6(t). The angle p(t) = 13(t) > Pb(t), so that Ra > P > 8b and therefore fa'(s) G f'(s) G fb(s). The second statement of the proposition follows by integration. (If s < 0, consider the inverse geodesic r_(s) = r(-s) and observe fp(s) > f'(s) > fb(s)). Lemma 4.3. Given that f, fa, fb are as before. Assume f (O) = f a(0) = fb(O) and f (l) = f a(l) = f b(l) . Then fb(s) < f(S) < fa(S)
for
s E [0, 1]
.
Proof. Fix s E [0, 1] and look at the triangles 41 = (r(0), r(s), z) and 42 = (r(1), r(s), z). In one of thet riangles, say in 41f the angle R at r(s) is not smaller than the corresponding angle Pa in Ha. Suppose for the moment that R equals Pa. Then Lemma 4.2, applied to 41, implies f (s) G f a(S). This is a fortiori if
R>Pa-
The proof of the inequality on the left hand side is similar. Remark. Since in the flat case f o is linear, the above lemma gives another proof of the convexity of F. For the geometric applications consider triangles 4 with two vertices p, q E M and one vertex z at infinity. Such a triangle gives rise to the following data: I = d(p, q), a = p(q, z), and R =
488
ERNST HEINTZE & HANS-CHRISTOPH IM HOF
da
Existence and uniqueness of 4a and 4b are obvious.
Let r be the geodesic ray with r(0) = p and r(l) = q, and assume that F is normalized such that f (O) = 0, where again f = For. Then f (l) = d, fl(l) = cos p, and Lemma 4.2 applies. Proposition 4.5. Given that 4 is as above. In Ha and Hb there exist unique triangles 4a and 4b (up to isometries) with l = la = lb and d = da = db, and for these triangles we have
ab < a < as and Pb < f < Na . Proof.
With the same notation as above, Lemma 4.3 implies fb(O) < f'(0) < fa(0)
and
fb(l) > f'(l) > fa(l)
These give the estimates for a and A, since f'(0) _ -cos a and f'(l) = cos f3. Distances on horospheres. Our next aim is to compare the '-distance h(p, q) of two points p, q on a given horosphere ' with their usual distance d(p, q). If moreover p and q lie on a different horosphere `, then their distance h'(p, q) may be different from h(p, q). However, the following theorem gives estimates independent of the chosen horosphere. and denote their-distance by h(p, q). Theorem 4.6. Assume p, q E Then
a sinh 2 d(p, q) < h(p, q) <
2 sinh 2 d(p, q)
Proof. First we prove the inequality on the left hand side. We choose in Ha two points pa and qa lying on a horosphere a, such that their 'a-distance ha(pa, qa) equals h(p, q). Let ra : [0, 11 , Ha be the geodesic from pa to qa, and u,: [0, 1] _ a the projection of ra onto 'a along the geodesics orthogonal to Via. Then l(pa) = ha(pa, qa) and ra(s) = +'a(-fa(s), pa (S)), where ,J*a denotes the radial flow associated with -lea, F, is the Busemann and f a = Fa o ra. function vanishing on be a minima] '-geodesic from p to q satisfying Now let p : [0, 11 , JJI 11 = l t2a ll,
and define the curve r : [0,11--+M from p to q by r(s) _
J(- f a(s), ,a(s)), where if,, denotes the radial flow associated with
ra(S) _ -fa(s) grad Fa + la'/a(s) ,
. Since
1(s) = -fa(s) grad F +
where t = - f a(s) > 0 and F denotes the Busemann function vanishing on , Proposition 3.2 implies 1I r(s) J < k ra(5)1, and hence l(r) < l(ra). Now we have
d(p, q) < l(r) < l(ra) = d(Pa, qa) . A computation in hyperbolic geometry shows (cf. [8, p. 1061): ha(Pa, qa) _ (2/a) sinh (a/2)d(pa, qa). Therefore we get
489
GEOMETRY OF HOROSPHERES
a sink 2 d(p, q) <
2
sink a d(pa, q.) = ha(pa, q.) = h(p, q)
In order to prove the other inequality we start with the geodesic r : [0, 1] - M from p to q and its projection p : [0, 1] - .XP. Then h(p, q) < l(u), and (with the same notation as above) r(s) = *(- f (s), p(s)). Now we choose two points Pb and qb in Hb lying on a horosphere 'b such that hb(pb, q,) = 1(y). Let
[0, 1] - 'b be the Pb-geodesic from pb to q, and consider the curve [0, 1] - Hb from Pb to q, defined by rb(s) = ib(- f (s), ,ub(s)). As before, Proposition 3.2 implies d(pb, qb) < l(rb) < l(r) = d(p, q), and therefore pb : rb :
h(p, q) < l(p) = hb(pb, qb) = b sinh 2 d(pb, qb) < b sink 2 d(p, q)
Projection of a geodesic onto a horosphere. Let .y' be a horosphere with center z, and F the Busemann function vanishing on -VP. Then the projection
: M - .VP along the geodesics going to z is defined by
= j o (F x 1 w),
where ,,rr is the radial flow in the direction of z. Now given a geodesic r, we
estimate the length of its projection curve p = o r and the .VP-distance between its endpoints. Proposition 4.7. Let .VP be a horosphere and r a geodesic starting on .VP. Assume R < ir12, where R denotes the angle between r(0) and grad,,,) F. Denote by l(s) the length of p 1 [0, s], and by h(s) the .XP-distance between p(0)
= r(0) and p(s). Then for s > 0 sin (3 1( 1 sin R < h(s) < l(s) < 1 a ` coth as + cos (3 f b ` coth bs + cos (3 1I
Proof. First we prove the inequality on the right hand side. Consider the same data as above in H, and fix s > 0. Lemma 4.2 implies f (s) > fo,(s) and (3(s) < Pa(s), where f denotes the restriction F o r, and (3(s) the angle between r(s) and grad,,,) F. Now we compute 11 ,u(s) lI. We decompose (s) into cos n(s) grad,,, F and an orthogonal part tJ-(s) of length sin n(s). Then a(s) _ 7)*(r(s)) = 77x(rJ-(s)) = irt,(rJ-(s)) for t = f (s). Therefore Proposition 3.2 implies 1I a(s) J < sin (3(s) e-afts'. Similarly we get II Fla(s) 11= sin Pa(s) e-of-11), which together with AS) > f a(s) and (3(s) < Ms) yields !I,u(s) 1I < 11 lia(s) 11 and, by integration, h(s) < ha(s), where ha,(s) is the corresponding function for Ha, as usual. Now
an easy computation in Ha gives ha(s) = a-1 sin R (coth as + cos p)-1. Next we prove the inequality on the left hand side. Consider the same data as
above in Hb, and assume h(s) < hb(s) for a certain s > 0. By Lemma 4.2 we have f(s) < fb(s). In Hb there is a unique point qb with Fb(gb) = fb(s) and rjb(gb) = pb(h(s)). Denote by rb the geodesic segment from pb(0) to q, and by sb its length. The assumption h(s) < hb(s) implies sb < s. Now consider the curve r' in M lying over an .XP-geodesic from p(O) to p(s) with For' _
490
ERNST HEINTZE & HANS-CHRISTOPH IM HOF
Fb o rb, and denote its length by s'. Proposition 3.2 implies (as in the proof of Theorem 4.6) s' < sb. By its construction the curve r' joins p(0) to a point q
with the properties F(q) = fb(s) and (q) = p(s). Since f(s) < fb(s), the convexity of the distance function d(p(0), ) implies s < d(p(0), q) < s', which contradicts s' < sb < s. Hence h(s) > hb(s). From now on we assume a > 0, i.e., the curvature of M is bounded away from zero. In this case the point p(co) is defined to be the intersection with ' of the unique geodesic from r(co) to z. Denote by l the length of p, and by h the -distance between p(O) and p(cc). Corollary 4.8. Assume p < it/2 as before. Then 1
sing
sin (8
1
-distance between p(- co )
and p(+ cc). Then
?b
.
Proof. The inequalities 2/b < l < 2 /a follow from Corollary 4.8, by observing R = 0. Here we prove h > 2/b.
For s > 0 there are points p = r(-s_) and q = r(s+), such that s_ + s+ = 2s and f (-s_) = f (s+). Consider the same situation in Hb, and look at the points pb = 7b(- s) and qb = rb(s). Then fb(-s) = fb(s), and Lemma 4.2 implies f (s+) < f b(s) (provided s+ < s- ; otherwise, we have f (- s_) < f,(-s)).
Denote by h(s) the '-distance between ri(p) and (q), and assume h(s) < 2hb(S).
Now choose points pb and qb on the horosphereb = Fb 1(f b(s)), such that b-distance between b(pb) and b(gb) is h(s), and join p' and qb by the the geodesic segment rb. The assumption h(s) < 2hb(s) implies that the fib-distance between pb and qb is strictly smaller than the fib-distance between Pb and qb. `
Therefore l(rb) = d(pb, qb) < d(pb, qb) = 2s. Now define a curve r' over an '-geodesic from ri(p) to (q) by F o r' = Fb orb As in the proof of the previous proposition we have l(r') < l(rb). Denote the endpoints of r' by p' and q'. Since (p') = gy(p), r1(q) = (q), and F(p') = F(q') = fb(s) > f(s+) = F(p) = F(q), we get 2s = d(p, q) < d(p', q') < l(r'). This contradicts l(r') l(rb) < 2s; hence h(s) 2hb(s) = 2 (b coth bs)-', according to Proposition 4.7. Passing to the limit we get h > 2/b. As an application of Theorem 4.9 we give the following example. Let T :
491
GEOMETRY OF HOROSPHERES
M --> M be a parabolic isometry (cf. [4]) with fixed point z, and " a horosphere with center z. Denote by hT the -displacement function of T restrict. ed to Proposition 4.10. Assume a > 0. If inf hT > 21a (sup hT < 2/b), then (does not for all x e M(oo), x z z, the geodesic r from x to T(x) intersects
intersect ").
Proof. Denote by a the geodesic from x to z, and assume that a(0) lies on the horosphere moo, which has center z and is tangent to r. Then T o a joins T(x) to T(z) = z, and T o a(0) lies on o (cf. [4, p. 83]). Let, be the horo-
sphere parallel too and containing the points a(t) and T o a(t), and denote by h(t) the `fit-distance between these two points. According to Theorem 4.9 we have 2/b G h(O) G 2/a, and Proposition 4.1 implies h(t) G h(0)e-at for t > 0 and h(t) > h(0)e-at for t G 0. Now for t > 0 t does not meet r and h(t) G h(0)e-at < h(O) G 2/a, whereas for t G 0, t intersects r and h(t) > h(0)e-at > h(O) > 2/b. This completes the proof. References D. V. Anosov & J. G. Sinai, Some smooth ergodic systems, Uspehi Mat. Nauk 22 (1967) 107-172; Russian Math. Surveys 22 (1967) 103-167.
R. L. Bishop & R. J. Crittenden, Geometry of manifolds, Academic Press, New York, 1964. P. Eberlein, Busemann functions are C2, unpublished. P. Eberlein & B. O'Neill, Visibility manifolds, Pacific J. Math. 46 (1973) 45-109. L. W. Green, The generalized geodesic flow, Duke Math. J. 41 (1974) 115-126. H. C. Im Holf & E. A. Ruh, An equivariant pinching theorem, Comment. Math. Helv. 50 (1975) 389-401. J. J. O'Sullivan, Manifolds without conjugate points, Math. Ann. 210 (1974) 295-311.
0. Perron, Nichteuklidische Elementargeometrie der Ebene, Teubner, Stuttgart, 1962.
J. A. Wolf, Homogeneity and bounded isometrics in manifolds of negative curva-
ture, Illinois J. Math. 8 (1964) 14-18. UNIVERSITY OF BONN
J. DIFFERENTIAL GEOMETRY 12 (1977) 493-497
SUR LES STRUCTURES DE COURBURE D'ORDRE 2 DANS R" JACQUES GASQUI
Une structure de courbure d'ordre 2 dans R' est un tenseur de type (0, 4) sur Rn, verifiant les identites classiques auxquelles satisfait le tenseur de courbure d'une metrique riemannienne sur Rn (cf. identites (2) et (3)). Etant donne
une structure a de courbure d'ordre 2 sur Rrz, nous nous proposons de resoudre le systeme d'equations aux derivees partielles a coefficients constants :
(1)
a'gtij
+
axiaxk
a'g,k
azgzk
a'gjt
axLax,
ax>ax,
axiaxk
=
-
aZ,kz
of les gz1 sont les composantes d'une 2-forme symetrique sur R. Si g est une metrique riemannienne, son tenseur R de courbure de type (0, 4) s'ecrit:
Rijka _
1
azgza
2
axjaxk
+
_
a'g,ix
aZgik
ax Lax,
axiax,
a,gia axLaxk
\I1
+ z g09(r>kr4t - II) P,2
avec
r kti> = 1 2 h ghk aghi ax'
aghj
agL 11
+ axi - Zx
Ainsi le premier membre de (1) est, a 2 pres, la partie principale de R. La condition de compatibilite de (1) est que a verifie la deuxieme identite de Bianchi par rapport a la connexion plate canonique sur Rn. Nos calculs se ramenent facilement a ceux que nous avions fait dans [1] pour 1'equation de Cartan pour les immersions isometriques. Si la condition de compatibilite est satisfaite, les resultats classiques sur les systemes a coefficients constants permettent d'obtenir des solutions de (1) sur tout ouvert convexe de R1. Les notations et le formalisme sont ceux de [2] et [1]. Je tiens a remercier H. Goldschmidt qui a bien voulu relire mon manuscrit et me faire part de ses remarques. Tons les objets avec lesquels on travaille seront supposes differentiables de classe C. On note respectivement x', , x et T* les coordonnees canoniques Communicated by D. C. Spencer, September 30, 1975.
JACQUES GASQUI
494
et le fibre cotangent de Rn. Si w est un tenseur de type (0, p) sur Rn (i.e., est une section de Ox ' T* sur Rn), on posera
(a
a
axiP
axis
1-
......ip
, iP compris entre 1 et n. Les tenseurs de type (0, p) symetriques ou alternes sur une partie de leurs arguments seront toujours reperes par leurs composantes par rapport a la base (dxil 0 . (9
pour tous entiers i de C°°(®P T*).
Une structure de courbure d'ordre 2, ou double 2-forme, sur Rn est une section a de A2T* Ox A2T* sur Rn verifiant
aijkl - aklij , aijkl + ajkil + akijl = 0
(2)
(3)
pour tous entiers i, j, k, 1. Si g est une 2-forme symetrique sur Rn, on verifie facilement que le tenseur w(g) defini par 0)(g)ijkl
a2gil
aXjaxk
+
_
a2gjk
aXiaXl
a2gik
_
axjaxi
a2gjl
axZaxk
est une structure de courbure d'ordre 2. En outre on a
(4)
aw(g)jklm + ax'
60,(g)kilm,
axj
+
aw(g)ijlM axk
=0
Les identites (4) reviennent a dire que w(g) verifie la deuxieme identite de Bianchi par rapport a la connexion plate canonique sur Rn (cf. [3]). On note G le sous-fibre de A2T* OO Z12 T* dont les sections sont les structures
de courbure d'ordre 2 et H le sous-fibre de T* Ox G dont les sections w verifient la deuxieme identite de Bianchi : (J)ijklrm, + Wjkil-m, + Wkijl'm. =
0.
Soit J2(S2T*) le fibre des 2-jets de sections du fibre S2 T* et soit cp: J2(S2T*) G le morphisme de fibres vectoriels sur Rn defini par (P(j2(g)(x)) = w(g)(x) ,
si g est une 2-forme symetrique define sur un voisinage de x E R. Si o(cp) design le symbole de o (cf. [2]), on a
6((,0)(0ijkl = Ciljk + Cjkil - Cikjl - Cjlik ,
SUR LES STRUCTURES DE COURBURE
495
pour tout C E SIT* ® SZT*. Le noyau de cp est une equation differentielle lineaire d'ordre 2 sur SZT* que 1'on notera R2. Posons N = I n(n + 1) et fixons une 2-forme symetrique B sur Rn a valeurs dans RN elle que les vecteurs
Bij(x)=B(
axi
a axe
avec1Gi<j
(x)
soient lineJairement independants en tout point x e Rn. Si r est un entier, notons SrT* ®R le produit tensoriel du fibre SrT* par le fibre trivial sur Rn de fibre RN, et B(' l'isomorphisme de fibres vectoriels sur Rn de S''T* ® RN sur S"rT* ® SZT* defini par B
Si C E
Soit P
(r)
(C)il,...,irjk = \Bjkl Cil,..., it
e
RN, ou < , > design le produit scalaire canonique sur RN. ® R`N , G le morphisme de fibres vectoriels sur Rn defini par
p(C)ijkl =
SZT* (3 R` P G
(5)
B(2)1
lid
SIT* ® SZT *
.
G
et
S1+1T* (3 RA'
(6)
T* ®SrT* ®RN 1B(r)
Sr+'T* ®SZT*
T* ® STT* ®S2T*
commutent, ou a est le morphisme de Spencer (cf. [2]). Reprenant alors les calculs du § 4 et des propositions 5.1 et 5.2 de [1], en remplacant TY par RN et p*, par p, la condition de "rang maximum" imposee a B permet d'affirmer que Ker (p) est un sous-fibre involutif de S2 T* (DR N. D'apres la commutativite de (5) et (6), le symbole de R2, noyau de v(cp), est aussi involutif. Si C E S3T* ® SZT*, on a
d(ip) ° a(C)ijkl'ne = Cijmkl + Cikljm - Cijikm - Cikmjl
En sommant 1'expression ci-dessus par permutation circulaire sur i, j, k, on
496
JACQUES GASQUI
obtient zero ; ainsi l'image de o1) o 3 est contenue dans H. Toujours d'apres la proposition 5.1 de [1], on a Im (p o 3) = H : combinant cc dernier resultat avec la commutativite de (5) et (6), on en deduit que l'image de o a est egale a H. Si R3 design le premier prolongement de 1'equation R2i on a, d'apres [2], la suite exacte
R3-R2 K>(T*(3 G)IH, oil K est le morphisme de fibres vectoriels defini comme suit : si p = j2(g)(x) E R2, avec g section de S2T* sur un voisinage de x e R", K(p) est la classe de i dans (Ty* (9 G.)/Hs, of A est 1'element de T © G, defini par
N2Jklr, =a(0)(g);kl'm,)(x) a Mais, d'apres (4), on a K = 0, donc R2 est formellement integrable. Des calculs precedents et du theoreme 4.3 de [2], on deduit que la condition de compatibilite que doit verifier le second membre a du systeme (1) est la deuxieme identite de Bianchi : a a
ajkI ,, +
a
as
akilm +
a axkaiac =
0
On a alors le Theoreme. Soit a une structure de courbure d'ordre 2 sur R" verifiant la deuxieme identite de Bianchi par rapport a la connexion plate canonique sur R". Si U est un ouvert convexe (resp. ouvert convexe borne) de R", it existe une 2-forme symetrique (resp. metrique riemannienne) g sur U telle que
w(g)=a. Demonstration. L'existence de solutions sur un ouvert convexe est un resultat classique sur les systemes a coefficients constants (cf. [4]). L'existence de solutions metriques riemanniennes sur un ouvert convexe borne U s'en deduit alors facilement : si g est une 2-forme symetrique sur U verifiant w(g) _
a, la forme g define sur U par
91 j = g¢; + A , avec A E R+ assez grand, est une metrique riemannienne sur U qui verifie encore
0)(g) = a.
SUR LES STRUCTURES DE COURBURE
497
References J. Gasqui, Sur l'existence d'immersions isometriques locales pour les varietes riemanniennes, J. Differential Geometry 10 (1975) 61-84. H. Goldschmidt, Existence theorems for analytic linear partial differential equations, Ann. of Math. 86 (1967) 246-270. R. S. Kulkarni, On the Bianchi identities, Math. Ann. 199 (1972) 175-204. B. Malgrange, Systemes differentiels d coefficients constants, Seminaire Bourbaki, 15ieme annee, 1962-63, Exp. 246. UNIVERSITE SCIENTIFIQUE ET MEDICALE DE GRENOBLE
J. DIFFERENTIAL GEOMETRY 12 (1977) 499-511
CURVATURES OF COMPLEX SUBMANIFOLDS OF C" PAUL YANG
0.
Introduction
Complex submanifolds Mn of a complex N-space CN from the viewpoint of hermitian geometry are distinguished by (a) the existence of N holomorphic imbedding functions f 1, fz, IN so that the kahler form is of the form iaa(Z I f 2I2), and as a consequence (b) the imbedding is minimal in the sense of riemannian geometry, and all the holomorphic sectional curvatures are nonpositive. In [2] Bochner demonstrated that the Poincare metric of constant negative curvature on the unit disc cannot be holomorphically imbedded in CN even locally. It seems therefore reasonable to pose the following Question. Does there exist a complete complex submanifold M" of CN with holomorphic sectional curvature bounded away from zero? In this paper we discuss partial results to this question. To begin with, we show in § 1 that a negative answer to this question would imply that there is no bounded complete complex submanifold of CN. In § 2, utilizing an elementary observation on the Gauss map we answer the question in the negative for hypersurfaces, and in § 3 we show that it suffices to consider the question for holomorphic curves (n = 1). In § 4 we recall the higher order curvature functions introduced by Calabi
and show that two such functions are enough to determine a holomorphic curve uniquely up to a rigid motion in CN, and thus providing a justification for a generalization of the theorem in § 2, in terms of the higher order curvature functions. In § 5, applying the method of extremal length we derive a criterion, which involves the curvature behavior at infinity of a simply connected metric riemann surface M for it to be conformally equivalent to the disc. It is subsequently used to sharpen the result in § 2. The last section contains curvature estimate for a piece of curve in CZ which
is a graph over a domain in C. The author would like to thank Professors H. Wu and S. S. Chern for their guidance and generous assistance. Communicated by S. Kobayashi, October 3, 1975. Partial results of this paper form essentially the second part of the author's thesis at the University of California, Berkeley written under the direction of Professor H. Wu.
500
PAUL YANG
1.
Boundedness of complex submanifolds in Cn
It is an open question whether a complete minimal submanifold of euclidean space can be realized in a bounded region. In the case of complex submanifolds we show that a negative answer to the question in the introduction would yield a negative answer here as well. This is accomplished by the following. Proposition. The unit ball BN can be holomorphically imbedded in C2N with the following properties : Let F : BN - C2N be the imbedding. Then (A) (B)
I dF(v) I> I v I for each v e T,(BN'), p e BN, the holomorphic sectional curvatures of BN are strongly negative, i.e.,
K(v) < -c < 0 for all v e T,(BN), p e BN. Proof.
Consider the map F : BN - C2N given by F(z,, , zN) = (z , , e). Relative to the coordinates z1, , zN, we have gi j =
K(v) _ -((E Kijxivzv'v'1vl)/I Ei gijvivj I4 KaiiivZv2vZv2
giivivj I' < -c < 0
ti
for some c. Condition (A) is satisfied, since F(BN) is a graph over BN. q.e.d. Next we recall the Gauss-Codazzi equation for computing the curvature of
a submanifold o : M - M: (1)
K(v) = R(v, Jv, v, Jv) = R(v, Jv, v, Jv) - + ,
where R = curvature tensor of submanifold M, R = curvature tensor of manifold M, B = second fundamental form of M, K = holomorphic sectional curvature of M, k = holomorphic sectional curvature of M. Recalling that the second fundamental form is complex linear (B(Jx, Y) = B(x, Jy) = JB(x, y)), we may rewrite (1) as K(v) = K(v) - 21 B(v, v) IZ
,
which expresses the curvature decreasing phenomenon of a complex submanifold of a kahler manifold. Thus we are in a position to conclude the argument. Suppose cp : Mn - BN is a complete holomorphic immersion. Then composing cp with the map F constructed in the proposition, we see clearly that F o cp is still a complete immersion (a consequence of (A)). Since Fcp(M) is a submanifold of F(B), the cur-
COMPLEX SUBMANIFOLDS OF C"
501
vature decreasing property implies that F o cp(M) has strongly negative holomorphic sectional curvature. 2.
Hypersurfaces
In this section we study the Gauss map of a complex hypersurface in C"' 1 An elementary computation shows that the kahler form of the metric induced by the Gauss map is the negative of the ricci form of M itself. It will follow that no complete hypersurf ace of C"+1 can have strongly negative holomorphic sectional curvature.
To fix notation, let M" > C"+' be a complex hypersurface. The Gauss map is defined analogously as the classical gauss map of a surface in R3. Let
, be a unit normal to M at p. , is determined up to a multiplication by ei', sop determines a point in P"C. (Actually we shall use the fact that p
ti
7r
goes through the Hopf fibration M > S" +' > P,zC, and the Fubini-Study metric comes from the standard metric on the sphere S'"+'.) The Gauss map is simply G(p) = If we represent M locally by the zero set of a holomorphic function f, then G(p) = [(af /az,)(p), , (af /az"}1)(p)] is a local representation of G, hence G is conjugate holomorphic. Let co = kahler form of Fubini-Study metric in P"C. Then lr*m =
Let S denote the ricci form of M, i.e., the (1,1) form corresponding to the ricci tensor. The claim is
7r*u,=-S.
(2)
Let 1 G a, (3, r G n + 1 ; 1 G i, j, h < n. Let e,(x) be a field of unitary frames with e1, , e,, tangent to M, and e,n+1 normal to M. Its dual coframe field consists of n + 1 complex valued linear differential forms Ba of type (1, 0), and the kahler metric for C"+1 is written as
ds'BB. The connection forms O
O+
satisfy
=0,
dO
O AO
.
The curvature forms Oa, are given by
dO. _ Oar _
Bar A BTU + (9.3
R,,,eO A 1 (- 0 in this case)
Restricting everything to the hypersurface we have from
0"+i=0
PAUL YANG
502
that 0 = don+1 =
of A oi, n+1
so that by Cartan's lemma where aik = aki Bi,n+l = Z aikOk , doi, = Z Bik A 0k.a - oi,n+1 A o,j,n+1
Therefore Oi; = Oi> - oi,n+l A Bj,n+1 = -oi,n+l A B7,n+1 The Ricci form is given by (3)
ZS.IkO;ABk=-Z ai;aiko;nok. 7,k,i
Since the Gauss map is given by G(p) = en+1(p), the induced kahler form is
(4)
G*w = den+1 A We-.,+1 = Z Bn+1,i A Bn+1,i
Z Bi,n+l A Oi,n+1 =
ai.iaikej A ok
Comparing (3) with (4) yields the claim. Theorem. Let p : Mn -> C7+1 be a complete complex hypersurface. Then (Mn, So) cannot have strongly negative holomorphic sectional curvature. Proof. Suppose not; say in fact K(v) < -c G 0, therefore the Ricci curvature is < -(n - 1)c. In view of (2) this means that the Gauss map is distance-increasing by a factor greater than (n - 1)c, hence (M, G) is complete with respect to the induced metric G*w. Since the Gauss map is equidimensional and antiholomorphic, the induced metric must have constant holomorphic sectional curvature +4, which would imply that M PnC, i.e., M is a compact manifold, a clear contradiction. Remarks. (1) It is easy to obtain quantitative version of this theorem, for example, using the same argument one can show that if So : Mn -> C7 is a complete complex hypersurf ace then for any point p a Mn, c > 0, there
exist a sequence of points pi and vectors vi E TP1Mn such that K(vi) > -c/dist (p, pi)', where dist (p, pi) can be either geodesic distance on Mn or the euclidean distance I p - pi 1. It is easy to see in either case that if the conclusion is false, then the induced metric G*w is complete. (2) The theorem goes through for a minimal hypersurface of RN+1 in a completely analogous manner. 3.
Reduction to holomorphic curves
The general case of arbitrary codimension can be reduced to the consideration of holomorphic curves, as shown in the following proposition. Proposition. Suppose Mn is a complete complex submanifold of CN with either of the following properties:
503
COMPLEX SUBMANIFOLDS OF Cn
the holomorphic sectional curvature K satisfies K(v) < -c < 0, CN is bounded. cp : Mn Then there exists an affine subspace LN-11+1 of CN such that L'-'+' (1 Mn is a nonsingular complete holomorphic curve with the same property (1) or (2). Proof. For each nonnegative integer i > 0, let Bi denote the set (1) (2)
{z E CN I I z I < i}. The set of affine subspaces L of CN
of dimension N - n + 1,
whose intersection with Bi (1 Mn is a nonsingular curve, is clearly open and dense in the space of all affine linear subspaces. L and S can be made into a complete metric space. A category argument then yields the existence of an LN-11+1 such that LN-h- n M" is a nonsingular curve, which being a closed subset of Mn is cleary complete. Property (1) is satisfied due to the curvaturedecreasing property. Property (2) is trivial. Remark. The properties (1) and (2) and completeness are inherited when one passes to the universal covering manifold. Therefore for the questions under consideration, it suffices, in view of the uniformization theorem, to take the unit disc as the underlying Riemann surface. 4.
Holomorphic curves
Consider a holomorphic curve cp: M' -f Cn. In terms of local coordinates z = x + iy, the hermitian metric induced from Cn may be written as ds2 = , f,) . As the metric is conformal, F I dz 12, where F = E i I f'12 and (p = (f 1, f2, the Laplace operator can be expressed simply as d = (4/F)(d2/dzdz), and the
Gauss curvature is found to be K = (-2/F)(d2/dzdz) logF = -24 log F. According to Calabi, there is a sequence of inductively defined nonnegative real analytic functions on M : Lemma (Calabi [31). Suppose that the image cp(M) lies in no hyperplane. Then we may define a sequence of functions {Fn} as follows :
F1=F,
Fo=1,
(5)
Fk+1 =
F Fkx
d log Fk) , 1 \ dzdz /
for 1 < k < n
Fk is nonnegative and vanishes only at isolated points. The succeeding function is defined by (5) away from these points, but extend to a real analytic function
on all of M.Fk=O fork>n+
1.
Let ds2 = F IdzI2 be a real analytic hermitian metric on M, and suppose that a sequence of functions Fk satisfying (5) can be defined with the same properties as in the above lemma. Then there exists a uniTheorem (Calabi [31).
que holomorphic isometric immersion of (M, ds2) into C" up to a motion of Cn. Simple computations give the following explicit formulas for these functions , f): in terms of the imbedding functions (f 1, F1 = E I fi I1 i
,
F2 =
i» I fi'f;. - filf
IZ
504
PAUL YANG 2
F, = Z
det
i>j>k
fi
fj
fk
fi'
f;
fk
fi'
f"
fT k
.
F = det (fci>)I2 I From these intrinsically defined functions it is possible to define higher order curvature functions Ko = 0, Kk = (Fk 1Fk_1)/(FFk) for k > 1. A simple computation then gives Kk = 24 log Fk. These intrinsic curvature functions have geometric meaning : they are the squared norms of higher order fundamental forms (Lawson [7]). In particular, K, = - EK where K is the Gauss curvature of ds2 as usual, and Kk = 0 if k > n. The curvature functions satisfy certain recurrence relations ; the ones of interest to us are
(6)
241ogKk=Kk+1-2K, + Kk_1-K1
k=1,...,n-1,
and, as a consequence,
(7)
41og K1K2
-2K,,_1 - (n - 1)K1
In terms of these curvature functions it is possible to formulate the following rigidity condition. Proposition. If gyp, cp' : M'--p C" are two holomorphic immersions satisfying K, = Ki, K, = K2, then F = F', and consequently o and cp' are congruent. Proof. If K1 (and hence K) vanishes identically, then Calabi's theorem (Lemma 1) implies that 0(M) and 0'(M) lie in a 1-dimensional linear subvariety of Cn. In this case checking the congruence of 0 and 0' is trivial. Let us then assume that neither K1 nor Ki vanishes identically. To prove 0 and 0' are congruent, it suffices to prove F = F' in some open set because of the analyticity of F and F'. Take a coordinate neighborhood on which all the Fk and Fk are nowhere zero, and denote this neighborhood by U. On f1 we have 2
F(K, - 3K1) = ddd- log K, =
d
dz log K.' = F'(K, - 3K)
We know K, - 3K1 = K2' - 3Ki. We will show in the following that K2 - 3K1 is not identically zero in U. Hence we may cancel the common factor to get
F = F', as desired.
Now suppose K2 - 3K1 = 0 in f1, and we will deduce a contradiction. This condition implies both 4 log K1 = 0 and K2 = 3K1. Hence 4log K, = 4log 3K1
= 0. Since K, = 24 log K2 + 2K2, we see that K, = 6K1. Now suppose at the kth stage, we have Kk = I k(k + 1)K1. With the help of the recurrence relation,
COMPLEX SUBMANIFOLDS OF Cn
Kx+,=24logKk+ 2Kk-Kk_1+K1f
505
1
we get
Kk+l = 24 log 2k(k + 1)K) + k(k + 1)K, - k(k - 1)K1 + K, = (k + 1) (k + 2)K1 In particular, 0 = K,, = In(n + 1)K1, which contradicts our assumption that K1 is not identically zero. q.e.d. The previous proposition indicates that perhaps one should consider higher order osculations in our study of the general question. In this direction it is easy to prove Theorem. If 0: D --> C1 is a complete holomorphic curve, then either
infK1 =0orinfK1
K,t_,=0.
Proof.
Suppose not ; then there is a curve 0: D - CN with the property that K1 > c > 0, K1, . , K,,_1 > c > 0 for some c. Consider the metric ds2
= K, Kn_lds2, which is a complete hermitian metric by the assumptions. Further, the relation 4logK1 ... K7L_1 = -2K7L_1 - (n - 1)K1 shows that
is' has strictly positive curvature. This is a contradiction to the fact that unit disc cannot carry such a metric (see Greene-Wu [4]). 5.
A condition for hyperbolicity
In Greene-Wu [4], it is stated that if M is a simply connected Riemann surface with hermitian metric g. Suppose there exist p, E M and c > 0 such that -c/d(p,, p)2+E < K(p) < 0 for some positive c. Then M is conformally equivalent to C. A complementary result is the following. Theorem. If M is a simply connected Riemann surface, and g a hermitian metric on M with nonpositive curvature satisfying : there exist po E M and a compact set G C M, such that K(p) G -c/d(p,, p)2 for all p 0 G. Then M is conformally equivalent to a disc. Remark. This result has been extended by Greene and Wu [5] to a condition for a complete manifold M of arbitrary dimension to be complete hyperbolic in the sense of Kobayashi. The proof employs the method of extremal length. We give a short convenient formulation. Definition. Let Q be a region in the plane, and T a set of rectifiable 1chains in Q. Consider the family of all conformal metrics ds = p I dz 1, where p is required to be Borel-measurable. Let L(y, p) =
for each y E T,
p dz j
A(Q, p) = f p2dxdy a
,
L(T, p) = inf L(y, p) rEr
PAUL YANG
506
Define the extremal length of the family I' (relative to Q) to be Aa (P) = sup L(F, p)2/A(.Q, p) ,
where p satisfies 0 < A(12, p) < oo
.
P
The following properties of extremal length are well-known ; see for instance Ahlfors and Sario [1] or Otsuka [8]. ( I ) A, (F) is a conformal invariant. More precisely, if cp is a conformal
map sending Q to Q', I' to F', then A, (P) = A,, (F').
(II)
Let D be a domain in the plane such that its complement with respect to the extended plane consists of mutually disjoint nonempty closed subsets
F, F'. Let D, be a sequence of domains increasing to D such that the complement of each D,, with respect to the extended plane consists of closed sets F,, F,, which decrease to F and F' respectively. Let F, be the set of rectifiable arcs in D, which join Fn to F,,, and I' be the set of rectifiable arcs in D which join F to F'. Then A D. (FJ increases to A D (F). (III) For a doubly connected region D in the plane bounded by simple closed curves ?' and 72 as below, consider the two families of curves in D :
Fl = {closed curves which separate -, and r,} , Fz = {"radial arcs" which join ?' to 72} Then we have A D (Fl) A D (F2) = 1. (IV) For a doubly connected region D in the plane bounded by a simple closed curve 7 and the component of D° containing oo , A D (F2) < oo implies that F U D is conformally a disc. Proof of the Theorem. Since M is simply connected, we may consider M as a subset of C. Let 0 correspond to po. Then the exponential map exp,, : T0(M) -> M is a diffeomorphism by the curvature assumption, and we have well-defined geodesic polar coordinates (r, 0) on M, (which is distinct from the usual z = rei") with the metric ds2 = dr2 + G(r, 0)d02. Let K(p) G -c/d(0, p)2
_ -c/r(p)2 for r(p) > r, > 0. Let r0=r(p)} r0Gr(p)
An upper estimate for the extremal length A DR (FR) will be derived. We remark that dr2 + G(r, 0)d02 is a conformal metric on the underlying Riemann surface, and therefore so is po Idz12 = (drz + G(r, 0)d2)/G(r, 0). It follows immediately from (III) that
COMPLEX SUBMANIFOLDS OF Cn
l \ DR (Fr) =
= inf A(DR, p)
1
ADR (FIR)
P
L(I'R, p)2
507
A(DR, po) inf L(1, po)2 rE FIR
It is clear that L(1, po) > 2,c for all ' E I'R so that AD. DR (I'R) < IA(DR, po) /
To estimate A(DR, po) we compare G(r, 8) against the solution of the Jacobi equation 0.
For r > ro,
-K =
(8)
c/r2 .
Let v/ G o = mine / G (ro, 8), ( G r)o = min, v/ G r(ro, 8) > 0. The solution of the ordinary differential equation
Yrr = c/r2 ,
Y(ro) = N -Go
,
Yr(ro) = (1/6r)o
is found to be
y(r) -
4c - 1) ,%/ -1 + 4cro 2(,8/1 + 4c)
+
r- ?(1 + 1 + 4c)
'(,/f + 4c + 1) r-'(1 1 + 4cro 2(1 - 1 + 4c)
- 1 + 4c)
.
It is easy to see that y(r) > A(ro, c, Go, (,,/ G r)o)r s (1 + 1 + 2c), so that G(r, 8) > A(ro, c, v/G o, (v/ G r)o)r 2 (1 + /1 + 2c). Hence
A(DR, po) = 2z R (r/G(r, 8))dr G A fro fro
1 + 1 + 2c < M < 0
which implies A DR (T2) < -1 /z. It follows from (II) that AD (I'2) < 00 , and M is conformally equivalent to a disc by (IV). Corollary. Let 0: M- C be a simply connected complete immersed curve, with 0(po) = 0. If K(p) < - c/I 0(p) I2 for large j gy(p) I, then M is con f ormally a disc. Proof. It is clear that 10(p) I < d(po, p). We now apply this criterion to an immersed holomorphic curve in C2. Corollary. If cP : M- C2 is a simply connected complete holomorphic curve,
then there is no po E M such that, for some constant c > 0 and compact set
GCM, p0EGand
PAUL YANG
508
K(p) < -c/d(po, p)2
for p I G .
Proof. Suppose that on the contrary, such po, c and G do exist. It follows , zn} be the finite from the theorem that M is conformally a disc. Let {z set of points in G where K vanishes. Then near each zi, K can be written as K(z) = -Iz - ziIZ'"h(z), where h is a positive function. Consider the metric (fl Iz - zij-Z"`)(-K)ds2. Since "the first factor is bounded away from zero outside the compact set G, it follows as before that it is a complete metric. But the curvature of the new metric is rj i I z - zi 12"i, hence it stays bounded away from zero outside G. Therefore an application of comparison theorem shows that M is again compact, which is a contradiction. q.e.d. The method of construction in the proof of the above corollary may also be used to obtain the following. Theorem. Let M be obtained from a compact Riemann surface M by deleting a finite number of closed connected subsets each of which has nonempty interior, and let (p: M -> CZ be a complete holomorphic curve. Then there is no po in M such that, for some constant c > 0 and compact G C M, K(p) -c/d(po, p)2 for p G. Proof. Suppose on the contrary, such po, c and G do exist. We choose f to be a meromorphic function on M satisfying : (a) the zeroes of I f 12 in M are precisely those of K, counted according to order of multiplicity, (b) the poles if any are contained in the interior of M - M, (c) min,,,.If(z)1Z > 0. Since the zeroes of K are finite in number; and aM is compact, such function
f can easily be found. Then we can prove as in the corollary above that -If I2Kds2 is a complete metric of curvature If 11. Thus by a previous argument
M would be compact, a contradiction.
6. A curvature estimate While the previous discussions center around global restrictions on the curvature of a complete immersed curve, there is, as in the case of minimal surfaces in R3, a semilocal restriction on the curvature. Theorem. Suppose a complex holomorphic curve in CZ is parametrized as a graph (z, f (z)) over a disc I z I < R, then we have the estimate
R < 4J(a)-1/Z ,
where a = minIKI .
The idea, due to E. Heinz [6], is simply to estimate the area of the surface over {IzI < r}. To proceed, we work with polar coordinates z = Proof.
re20, let D, = {(z, f (z)) : r < p}, and we have
(1 + f'f')dx A dy .
Area (DP) = 2 f f J
Izl
509
COMPLEX SUBMANIFOLDS OF Cn
Observe that
(d" - d') = rado - ar a dr
de =
ar
ae
an application of Stoke's theorem yields pd f 2r f'(pe'B)f'(pe20)d8
= f 0 p dp
(1 + f'(pe2')f'(pez'))d8
= f lzl dc(1 + f'f) =
JJ Izl
ddC(1 + f'f')
Consider the following function F(r), defined for 0 < r < R, F(r) = f or pdp f o, (1 + f'(pe20)f'(peie))d8 (= Area of (D,))
+ f'f')-3 < -a < 0. Differentiating F(r), we have
Let K
2,
F(r) = r f (1 + f'f')dO , 0
d
Zn
dB+rdr f0 (1+ff)dO, zl=,
F "(r) = E (1 + so that
F"(r) > f =f
IzI<, Izl
ddc(1 + f'!')
f" f"dx A dy > f
IzI<,
a(l + f'f')3dx A dy
Since, by Holder's inequality, dy)1/3
Rr2 < F(r) < (nr2)2/3 f /
z<,
(1 + f'f')3dx A
,
we obtain
F"(r) > a f f
zl<,
(1 + f' f')3dxdy >
;r2 4
(F(r))'
from which we deduce the differential inequality du
(F (p))2 = 2F'(p)F"(p) > 2F(p)
>
2
pa
d (F(p))' p
naP4 (F(p))3
for all 0< p< R.
PAUL YANG
510
Therefore
dP
(F'(p))2 >
2'.;r'
r'
F(p))'
or
>
(F'(r))2
(F(r))'
a
F(r) >
or
(F(r))2
2ir2r4
a
)1/2
`\ 2
;rr2
which implies
(9)
)>
(F(r)
drr
d(
ff-"'
Ri dr 7cR;
\1i2
\2
1
F(r)
for 0 < r < R
r2
)dr > (1 \2/
fx2
1
it J x, r2
dY
1_ 1
(a )1/21
1
F(R2)
F(R1)
F(R1)
1
_
1
1
1
/
\2
R1
R2
Letting R2 -> R we obtain 1iZ(
Ri
>
\2/ \R1
R
Hence 1
R
>
1
R1 -
(a )-1/2( 1 2
\ R2 /
or
R<(
Choose R1 = 2(2/a)1"2, we obtain R <
(a/2) /2R1
4(x/2)-1/2
- 1 ) -1
(a/2)1/2R2
as asserted.
References
[1]
L. Ahlfors & L. Sario, Riemann surfaces, Princeton University Press, Princeton,
[2]
S. Bochner, Curvature in Hermitian metric, Bull. Amer. Math. Soc. 53 (1947)
[3]
E. Calabi, Metric Riemann surfaces, Contributions to the theory of Riemann
1960.
179-195.
surfaces, Annals of Math. Studies, No. 30, Princeton University Press, Princeton, 1953.
R. Greene & H. Wu, Curvature and complex analysis, I, Bull. Amer. Math. Soc. 77 (1971) 1045-1049. [ 5 ] R. Greene & H. Wu, Some function-theoretic properties of noncompact Kdhler manifolds, Proc. Sympos. Pure Math. Vol. XXVII, Amer. Math. Soc., 1975, 33-41. [6] E. Heinze, Uber Flkchen mit Eineindeutiger Projektion auf eizze Ebenederen [4]
COMPLEX SUBMANIFOLDS OF Cn
L7] L8]
L9]
511
Kriimmung durch Ungleichungen Eingeschrdnkt sind, Math. Ann. 129 (1955) 451-454. H. B. Lawson, Holomorphic mappings and minimal surfaces, Proc. Conf. Differential Geometry, North Carolina, 1970. M. Otsuka, Dirichlet problem, extremal length and prime ends, Van Nostrand, Princeton, 1967. A. Vitter, On the curvature of complex hypersurfaces, Indiana Univ. Math. J. 23 (1974) 813-326. RICE UNIVERSITY
J. DIFFERENTIAL GEOMETRY 12 (1977) 513-522
ALMOST HYPERSURFACE STRUCTURES ROBERT H. BOWMAN & DONALD H. SINGLEY
1.
Introduction
This paper is a continuation of the paper, Almost submanifold structures, [1]. A second-order connection on a C' manifold M gives rise to a structure which resembles an immersion of M into another Riemannian manifold. We call this an almost submanifold (AS) structure, and call the structure an almost hypersurface (AH) structure if an additional condition analogous to the codimension of M being one is satisfied. As was noted in the previous paper, sometimes this AH structure in fact satisfies the integrability conditions for an immession of M into RT+1. However, this realization as an immersion is not always possible in cases where the vector field analogous to a normal vector field vanishes
at certain singular points. In the first part of this paper, we prove a more general integrability theorem, which considers isolated singular points, and we also study curvature properties of AH structures at singular points. The second part of this paper introduces the concept of a realizable structure, which resembles a generalized immersion, and we obtain certain curvature conditions necessary for such a structure to exist on M. Other curvature conditions imply that an AS structure cannot be an AH structure, and two such theorems are included in this section. The last section of this paper states an equivalence principle between AS structures and immersions of M into another Riemannian manifold, and the principle is used to give immediate proofs of two theorems which simply translate theorems about hypersurfaces to theorems about AH structures.
2.
Preliminary remarks and definitions
In this section, we will outline the results of [1] used in this paper. A second-order connection on M, as defined originally in [2], determines a vector bundle structure on 2M, such that IM _- TM O+ TM. The first factor of TM we call the horizontal bundle and identify with the tangent space of M ; the second we call the vertical bundle. Henceforth, the letters W, X, Y, and Z will always denote horizontal vector fields or horizontal vectors, A and B will denote sections of 2M, and will be a vertical vector field or vertical vector. The second-order connection described above determines a covariant differCommunited by K. Yano, September 15, 1975.
ROBERT H. BOWMAN & DONALD H. SINGLEY
514
entiation F of a section A of 2M. Now let < , > be a fiber metric on oir : 2M M, such that the vertical and horizontal subbundles are orthogonal. We may then define an operator V' in terms of F by
'Y = IxY = V1Y + a(X, Y) ,
17/
(1)
F,
= -dr(X) + axe = -d£(X) + Dx .
In the first equation of (1), the horizontal component VxY is the covariant derivative of the first-order connection induced by F, and the vertical component a(X, Y), which we call the second fundamental form, is bilinear. In the second equation of (1), D is a connection in the vertical bundle, and the horizontal vector field dr(X) is defined by
(2)
for all Y. The operator V' is called an almost submanifold or AS structure on 2M. We shall also assume that V' is Riemannian (see [1] for definitions) ; this implies that a(X, Y) = a(Y, X). We now define the first vertical space V,(x) of an AS structure at x E M by V,(x) = span {a(X, Y) I X, Y E Mx} .
If V,(x) has maximum dimension l at any point X E M, we will call l the pseudocodimension of M. If l = 1, we call the AS structure an almost hypersurf ace or AH structure. In this case, we define a normal vector , by orthonormalizing a(X, Y) where a 0, and by setting x = 0 where a - 0. Moreover, we define d(X) = dr(X) for this . It is proved in [1] that Dx and are orthogonal ; hence considering only tangent vectors and vectors in the first vertical space we redefine V' by
VXY = V1Y + a(X, Y) , (3)
vxe = -S(X) .
This d is very similar to a Weingarten map for an immersed hypersurface. We also define h(X, Y) = I a(X, Y) I and note that
(4)
a(X, Y) = h(X, Y) .
We define a singular point of an AH structure as a point x E M such that a,(X, Y) - 0 for all horizontal vectors X and Y. At such a point, d may not be C-, since (2) defining d can be rewritten as
and the length function is not C`° at zero. However, since the length function
515
ALMOST HYPERSURFACE STRUCTURES
is continuous at zero, d(X) will at least be contiunous for any differentiable vector field X. The curvature tensor of an AS structure V' is defined as
(5)
R'(X, Y)A = vxv'YA - v''F'A - 1
YIA
By a standard calculation, (5) implies the equation analogous to the Gauss equation,
(6)
- .
In the case of an AH structure, the curvature tensor R' is not well-defined on sections A of 2M, since the definition of R' given in (5) would involve taking
derivatives of d(X), which is not differentiable at singular points. In fact, we cannot even define R' on vectors in T(M) at singular points, since at a singular
point we cannot write a(X, Y) as f , where f is a differentiable function. However, the right-hand side of (6) does make sence at singular points, so for an AH structure with singular points, we define
(7)
(Fxh)(Y, Z)e - (VYh)(X, Z)
At a singular point, the expression in (7) is not defined. Finally, given two linearly independent horizontal vectors X and Y, we define the sectional curvature k'(X, Y) of an AS structure in the plane spanned by X and Y to be
(8)
k'(X, Y) = k(X, Y) + -
where k(X, Y) is the usual sectional curvature of M in the plane spanned by X and Y. Again, for an AH structure with singular points we define k'(X, Y) by the right-hand side of (8). This makes k'(X, Y) a C- operator on tangent vectors X and Y for any AH structure. 3.
Singular points of almost hypersurface structures
Away from singular points, a theorem in [1] shows that an AH structure on M is integrable-that is, can be identified with an immersion of M as a hypersurface of Euclidean space-if and only if R' - 0. However, as noted above, one of the integrability conditions, (7), breaks down at singular points. The following theorem summarizes the situation for a manifold with isolated singular points.
516
ROBERT H. BOWMAN & DONALD H. SINGLEY
Theorem 1. Suppose that M is a simply connected and connected C° ndimensional manifold, which bears a Riemannian AH structure with R' - 0, and whose associated tensor sl is C° except on a set A of isolated singularities, which do not simply disconnect M. Then M is isometrically imbeddable as a C1 submanifold of R"+'. Moreover, this submanifold admits a second fundamental form which agrees with sl. Proof. The set M - A is a connected and simply connected C°° manifold such that d satisfies the Gauss-Codazzi equations. Hence by the fundamental theorem for hypersurf aces there is an isometric imbedding of M - A into Rn+1
(which we also denote by M - A) such that d is the second fundamental form. We may fill in the holes in M - A introduced in this procedure by taking limit points of M - A in Rn+' (this is possible since the imbedding is isometric and thus the holes must be single points) and denote the result by M again. If x e A, r : [to, t1] -p Iv! is a piecewise C`° curve through x (that is, C`° except at x), and N is the normal vector field on M - A (which exists locally at least), then set
(9)
N,(t) =
'
10
t=t',
,
where T(t) is the tangent vector to r except at 7(t') = x. If we take N(to) _ N,(,,,, then the right-hand side of (9) is continuous since d(T(t)) , 0 as t , t'. Thus there is a unique solution N(t) to (9) which must agree with the normal vector field except at 7(t') = x, so that we may define a normal vector at x by setting Nx = N(t'). That N(t') is independent of r may be seen as follows : Let 3: [so, sl] - M be a second broken C`° curve parametrized such that s' = t' and r(t') = a(s') = x. If the broken C`° curve e : [to, sl] , M is defined by (10)
e(t) =
r(t), {a(t),
if t0
t' = s'
by repeating the above argument, then N(t') = N(s') is the same for all of these curves. The normal field thus obtained is clearly C'. if neChoose a coordinate plane of R ' (employing an isometry of cessary) which contains Nx, such that the slope of the line containing N,x in the coordinate plane exists and is not zero. Consider the coordinate curve on M obtained by intersecting this coordinate plane with M. The orthogonal projection of N along the coordinate curve into the coordinate plane is continuous and thus determines a continuous field of normal lines to the coordinate curve in a neighborhood of x, whose slope exists and is not zero. If f is the equation of the coordinate (planar) curve and s(u) is the slope of the normal line through (u, f (u)), then
ALMOST HYPERSURFACE STRUCTURES
f'(u) = -1 /s(u)
(11)
517
,
and we see that the coordinate curve has a continuous derivative in a neighborhood of x. Since the coordinate curves through x are C1 and X E A is arbitrary, we see that M is a C1 submanifold. The second fundamental form may be globally defined by taking its value to be zero on A. The next theorem leads to examples of manifolds with no integrable AH structure :
Theorem 2. At any singular point of an AH structure on M, all sectional curvatures of the AH structure are the sectional curvatures of M itself. Proof. Since a - 0 at any singular point, the result follows immediately from (8). Corollary. On any even-dimensional ovaloid N. there is no global integrable AH structure whose first-order connection agrees with the Riemannian connection of N. Proof. Since the ovaloid N is homeomorphic to an even-dimensional sphere, whose Euler characteristic is nonzero, every tangent vector field on such an ovaloid must have a zero. So the vector of the AH structure must vanish at some singular point-that is, the set of singular points is nonempty. Moreover, for any AH structure, the set of singular points is closed. By Theorem 2, since all sectional curvatures of an ovaloid are positive, R' cannot be equal to zero at any singular point and hence, by the continuity of R', cannot be equal to zero in a whole neighborhood of a singular point. If N consists entirely of singular points, the AH structure is clearly not integrable. If N has nonsigular points, by taking a point on the boundary of the singular set, we see that R' is not zero at some nonsingular point, so that the AH structure is again not integrable. 4.
Realizable structures
Given an immersion of M into R'1"with second fundamental form h(X, Y) and a nonvanishing C- unit vector field on M, there is an AH structure on 2M, obtained by taking a(X, Y) = h(X, Y)&. Moreover, by Theorems 1 and 2, an AH structure comes from such an immersion, if and only if R' - 0 and the AH structure has no singular points. We may now ask which AH structures arise from an immersion plus a vector field g' which is not of unit length, where again we define a(X, Y) = h(X, (In particular, ' might have a zerh.) This leads to the following definition. Definition. An AH structure on a Riemannian manifold M will be said to be realizable if s/ = f.sd', where .s' is the Weingarten map coming from an isometric immersion of M into R' -1, and f is a C`° function on M. (Note that not of unit length, we obtain such an c/.) by defining a(X, Y) = h(X, Theorem 3. Let k'(X, Y) be the sectional curvature of a realizable AH
ROBERT H. BOWMAN & DONALD H. SINGLEY
518
structure in the plane spanned by X and Y, and k(X, Y) the sectional curvature of the same plane in the underlying manifold. Then (i) k'(X, Y) G k(X, Y) if k(X, Y) > 0, and k'(X, Y) > k(X, Y) if k(X, Y) < 0, where the inequlity is strict unless either k(X, Y) = 0 for the vectors X
and Y, or f = 0 at the point, (ii) at a point x, k'(X, Y) is a constant multiple of k,,(X, Y), independent of the vectors X and Y, (iii) wherever k(X, Y) is not equal to zero, the function f is determined up to sign by the equation k'(X,
f2 = 1 - k(X, Y)
(12)
Proof . By (2) and (8) together with the symmetry of A' and the Gauss equation for A' we have
_
+ f2(
k'(X, Y) - k(X, Y)
-( - )
The following lemma completes the proof. Lemma. Let a : Rq X Rq -> R3 be a symmetric bilinear mapping. If (14)
<01(X, X), a(Y, Y)> - < 0
for all X, Y E R", and the real algebraic variety V consisting of all X in R" such that a(X, X) = 0 has dimension r, then s > q - r. Proof. Extend a to a symmetric complex bilinear map of C4 X Cq -> C.
519
ALMOST HYPERSURFACE STRUCTURES
Consider the equation a(Z, Z) = 0, Z E C. This is equivalent to a set of s quadratic equations
a'(Z,Z) =0, ,a`(Z,Z) =0
.
The set S of solutions by the dimension theorem for complex algebraic varieties
has (complex) dimension d > p - s. If s > q, we are done. If s < q, we choose any regular point P E S. This point P has a neighborhood U which is a d-dimensional complex analytic submanifold of C. Thus there is a map f :
C' - U C Cq such that some d x d subdeterminant of the d x q (complex) Jacobian is not zero in a neighborhood of P. So we may choose d of the standard coordinates of Cq, which for convenience we relabel as Z, through Zd, over which U projects diffeomorphically onto an open subset U' of the complex subspace V = {Z1, , Zd, 0, , 0}. We let Re: Cq - Rq be the map which takes {Z1, , Zq} into {X1, , Xq}, where each Z; = X; + iY;. Then clearly Re(U) projects diffeomorphically down to Re(U'), so Re(U) has (real) dimension d, as well. But Re(U) is a subset of V by the following argument : Let Z = X + iY be any point in U. Then
0 = a(Z, 2) = a(X, X) - a(Y, k) + 2ia(X, k)
.
So a(X, X) - a(Y, k) and a(X, k) = 0. By (14), this implies that a(X, X) = a(Y, k) = 0; i.e., X and f are both in V. So, Re(U) is a subset of V, and hence r > d. Since d > q - s, we have s > q - r. The second theorem in this area is similar to a theorem of Otsuki [4] in the case of immersions in Riemannian spaces. Theorem 5. Let M be a manifold admitting an AS structure. If, for some point p E M, k'(X, Y) > k(X, Y) for all pairs of linearly independent vectors X and Y in a q-dimensional subspace Q of T,(M), then the pseudocodimen-
sion of the AS structure on M is > q - 1. Proof. (15)
The hypothesis implies, as in the previous theorem, that - < 0
.
for X and Y linearly independent. Assume that the dimension of the image
of a is < q - 1. The lemma stated in the last proof shows that there is a nonzero vector X E Q such that a(X, X) = 0. Consider the linear transformation given by Z a(X, Z). Since the dimension of the image of a is < q - 1, and since dimension (null space) + dimension (image) = dimension (domain), there is a vector Y, linearly independent from X, such that a(X, Y) = 0. Substituting X and Y into (15) we obtain a contradiction. Corollary. Let M be a Riemannian manifold. If the sectional curvatures of a 3-dimensional subspace of T,(M) are all negative, for some point p E M, then any realizable structure on M whose first-order connection agrees with
520
ROBERT H. BOWMAN & DONALD H. SINGLEY
the Riemannian connection on M must have a singular point at p. In particular, if the sectional curvatures of M are all negative, and the dimension of M is > 3,
then the only realizable structure on M is the trivial one, a(X, Y) - 0. Proof. This follows immediately from part (i) of Theorem 3. Remark. An example of a manifold with an AS structure whose pseudocodimension is > 1 is given by taking a manifold M admitting two metrics g and g'. Let a be the difference tensor of the two metrics, and let the connections in the vertical and horizontal bundles be (say) the Riemannian connections coming from g and g' respectively. To obtain the example, let Mn be an ovaloid, and g the first fundamental form of M. Since the second fundamental form of an ovaloid is positive definite, we may let g' be the second fundamental form of M. We recall that the rank of the difference tensor is defined as follows : Let {e.} be a local orthonormal frame for g, and let the components of the difference tensor ab, be defined by the equation (16)
a(eb, ee) _
a
abcea
Define a bilinear form B by setting (17)
B(es, e1) = Z auvauv U,v
We define the rank of a to be the rank of B as a bilinear form. (See [5] for a more intrinsic definition of B.) Lemma. The rank of a at x < the dimension of V,(x). Proof. Let q be the dimension of V1(x), and let the frame {ez} have the , e span the complement to V,(x). By (16), ab, = 0 property that eq+ for a > q. Hence, by (17), B(eb, e,) = 0 for b or c > q, so the rank of B is
< q. Using this lemma, we may restate a theorem in [5] as follows : Theorem. For an ovaloid N'n, if the dimension of V(x) < q on x in an open subset U of N, then U is at least (m - q)-umbilical, and the principal
curvature associated to the (m - q)-umbilical directions is constant (q # m
or m - 1). Now take an ovaloid Nn on which the principal curvatures are not equal on any open set, such as an ellipsoid n+1 j=1
XI/d=1
where as are all unequal. Then, for almost all points x in Nn, the dimension of V,(x) > n - 2, since otherwise Nn would be at least 2-umbilic on an open set. So the pseudocodimension of Nn is at least n - 1.
ALMOST HYPERSURFACE STRUCTURES
521
5. AH structures as immersions In many cases, we may simply regard R' as the curvature of an imaginary immersing space for M and note that proofs for immersed submanifolds go through almost verbatim to give proofs of similar theorems about AH or AS structures. As examples of this technique we quote the following two theorems. Theorem 6. Let M be a manifold admitting an AH structure, and d the
x-
associated tensor of this structure. Let t(x) be the type number of 4 at that is, the rank of the linear transformation 4 at x. Let R'(X, Y) and R(X, Y) be the curvature transformations arising from the curvature tensors R' and R, respectively. Then
(i) t(x) = 0 or 1 #j R'(X, Y) = R(X, Y) for all plane sections at x, (ii) t(x) > 2# t(x) = n - dim Tx , where Tx = {X e T,,(M) : R(X, Y) = R'(X, Y), yY e Tx(M)}
.
Theorem 7. Let f 1 and f be AH structures on 2M inducing the same metric on M, and let R f' ,(X, Y) = R f,(X, Y) for all plane sections of M. If the type number of 4 f, > 3, then W., = ±d f,. The proofs of Theorems 6 and 7 can be easily deduced from the fact that the relative curvature R(X, Y) - R'(X, Y) plays the same role as R(X, Y) does for a hypersurface immersed in Euclidean space, where the counterpart 2
of the curvature R' of the immersing space is equal to zero. We then note that the proofs of the analogous theorems in Euclidean space [3, Theorems (6.1) and (6.2), pp. 42 and 43] make no use of the special properties (such as homogeneity) of Euclidean space. Moreover, the proofs rely only on the Gauss equation, which is formally identical to (6). So those proofs go through to this case. Other theorems can be proved similarly. This equivalence principle can also be reversed; certain theorems about AS structures can yield theorems about submanifolds of Euclidean space. As an example, the reader may easily modify the proof of Theorem 4 to obtain the following extension of [3, Theorem 4.7, p. 28]. Theorem 8. Let M be an n-dimensional Riemannian manifold isometrically immersed in Rn+P, and let a(X, Y) be the second fundamental form of M.
Then p > m - d, if there is a point x of M such that (i) the tangent space T,(M) contains an m-dimensional subspace T. such that the sectional curvature for any plane in T. is nonpositive, and (ii) the set of vectors X e T. such that a(X, X) = 0 forms a real algebraic variety of dimension d.
522
ROBERT H. BOWMAN & DONALD H. SINGLEY
References [1] [2] [3] [4]
[5]
R. H. Bowman, Almost submanifold structures, J. Differential Geometry 9 (1974) 537-546. , Second order connections, J. Differential Geometry 7 (1972) 549-561. S. Kobayashi & K. Nomizu, Foundations of differential geometry, Vol. II, Interscience, New York, 1969. T. Otsuki, Isometric imbedding of Riemannian manifolds in a Riemannian
manifold, J. Math. Soc. Japan 6 (1954) 221-234. R. B. Gardner, Subscalar pairs of metrics and hypersurfaces with a nondegenerate second fundamental form, J. Differential Geometry 6 (1972) 437-458. ARKANSAS STATE UNIVERSITY ARKANSAS COLLEGE
J. DIFFERENTIAL GEOMETRY 12 (1977) 523-553
PERTURBATION THEORY FOR CONDITION (C) IN THE CALCULUS OF VARIATIONS JILL P. MESIROV 1.
INTRODUCTION 1.
Introduction
One classical method of solving the Dirichlet problem for d u = f on a bounded domain Q C Rn, is to minimize the Dirichlet integral J(u) = f 2I Fu IZ + f(x)u over an appropriate class of functions u on Q. The generalized variational Dirichlet problem is to study the critical points of a functional
J(u) = f P(u) where ?, the Lagrangian, is a nonlinear differential operator from sections, with prescribed boundary values, of a fiber bundle E over a compact manifold M with boundary, to sections of the trivial line bundle RM. The key step in finding a critical point which is a minimum, for example, is to show that the functional actually achieves its minimum value. For this we need some sort of compactness condition ; we use the Palais-Smale condition (C).
To state this precisely, we consider LP(E), a manifold modeled on the Sobolev space of sections whose distributional derivatives up through order k are in LP, with norm II P, A functional J : LP(E) -> R satisfies condition (C), if given any subset S of Lk (E) on which J1 is bounded but j I DJ I I is not bounded
away from zero, then there is a critical point of J in the closure of S. If J is CZ and bounded below, and satisfies condition (C), then J assumes a minimum on each component of LP(E), [13], [16]. The main question we consider is, if Jo satisfies the Palais-Smale condition (C), under what conditions on a perturbation -Y,' can we show that the perCommunicated by R. S. Palais, July 26, 1975. Supported in part by N.S.F. grant
MPS-74-19388.
524
JILL P. MESIROV
turbed functional J = Jo - Y' also satisfies condition (C)? With this in mind, we observe that a functional J having both of the following two properties satisfies condition (C) : (a) A functional J is pseudo-proper on LP if I J(S) I G a for some set S C Lp implies uIP,k G b for all u E S.
(b) A functional J is coercive on Lx, if given any bounded sequence (ui) in Lk such that
(DJ,, - DJu,)(uz - u,) - 0 , then (uz) has an Lk convergent subsequence. (See § 2 for a more precise definition.) The condition we call pseudo-proper, is classically sometimes called the coercive condition, and is written : j u jjP, k - co implies J(u) I --> co.. In practice this is how these conditions are used. The pseudo-proper condition on J provides a weakly compact set, from which we get a candidate for the minimum of J as a weak limit. The coercive condition is used to show the weak limit is in fact the strong limit of a convergent subsequence. In the literature, condition (C) is almost always verified by checking these or similar conditions. The pseudo-proper condition is also used in monotonicity methods of solving partial differential equations [7]. Our work is an investigation of the pseudo-proper and coercive conditions. If J0(u) = f 2'(u) is pseudo-proper (resp. coercive) on Lk, what conditions on
Yi (u) = f V(u) insure J = Jo - is still pseudo-proper (resp. coercive)? We treat the two conditions separately. This is more than a stability question. it is not difficult to show that if a perturbation K is "small" enough, then it can be. preserves the two conditions. We ask, rather, how large The contents of the paper are as follows : § 2 contains technical preliminaries and notation. (We suggest skipping this section, referring back to it as the need arises.) In § 3 we discuss the motivating example of geodesics in the presence of a bounded potential as a perturbation problem. Boundedness is too restrictive a condition for most applications. The point of the remainder of the paper is to investigate pseudo-properness, coercivity, and condition (C) under weaker assumptions on perturbations. §§ 4 through 7 contain the perturbation results for the pseudo-proper condition. We begin with an especially illuminating and useful special case. For u E L,(M, R) with "zero boundary values," let
J0(u) = f pu lz
Jo is the square of the Li norm of u and thus pseudo-proper. Let V : M X R
PERTURBATION THEORY FOR CONDITION (C)
525
, R be continuous, and 2, > 0 be the first eigenvalue of the Laplacian. Finally, let J(u) = J0(u) - f V(x, u). The essential content of Theorem 5.1 is Theorem 5.1'.
(a)
If
V(x, s) < cont. + Ks' for all s E R, and K < 2, then J is bounded below and pseudo-proper. (b) If V(x, s) > cont. + 2,s' , then J is not pseudo-proper. The theorem can be viewed as an asymptotic growth estimate on V, compare with [6, § 8]. Part (a) shows that if V grows at most quadratically, at a
rate bounded by 2, then it preserves the pseudo-properness of f I Fu I'; part (b) shows the sharpness of the bound on the growth rate found in (a). As an outgrowth of Theorem 5.1, we give an example of a functional which satisfies condition (C), but is not pseudo-proper.
The remainder of §§ 5 through 7 extends Theorem 5.1 to more general functionals J0, arising from kth order Lagrangians, and perturbations Y'. In the second half of § 5 we extend part (a) to general perturbations of order zero, i.e., to those which only depend on u and not any derivatives of u. In this context we also discuss geodesics in the presence of possibly unbounded potentials. (The essential step in our discussion of geodesics is to find an analogue of 21 in a setting where there is no linear structure.) The extension of part (a) to general perturbations of order k - 1 is carried out in § 6, while in § 7 we extend part (b) to perturbations of order k - 1. The coercive condition is dealt with in § 8. We consider coercive functionals
J0(u) = f 2(u), where 2 is a polynomial differential operator of order k and satisfies an auxiliary condition which insures 2 is smooth from LP (E) to L'(RM).
The perturbations -11-(u) = f V(u) are also polynomial, only depend on the
(k - 1)-jet of u, and also satisfy the auxiliary condition. Under these hypotheses we get an optimal result, namely that all such perturbations preserve coercivity. We also show that we can relax the auxiliary conditions on V depending on the relation between p and the dimension of M. Related conditions are also given for the case where the functionals act on sections of a vector bundle , in which case Lk(e) is a Banach space. Some of these results were announced in [9]. In a subsequent paper [11 ] we will continue our investigation of which functionals are pseudo-proper and coercive. There we show, for example, that if 2(u) is a quadratic polynomial
526
JILL P. MESIROV
in u and its derivatives, then modulof I u 12, J(u) =
J
2(u) is pseudo-proper if
and only if the bilinear form associated with 2 is uniformly strongly elliptic. The author would like to thank R. Palais for helpful and stimulating conversations. 2.
Preliminaries and notation
M denotes a compact connected C°° Riemannian manifold of dimension n > 1 with or without smooth boundary, aM. and , denote finite dimensional 0 < k < oo, the linear space of Ck C°° vector bundles over M, and sections of . Co is the linear space of C°° sections with compact support
in the interior of M. We define C`°(E) for E a finite dimensional C°° fiber bundle over Min a similar manner. RM is the trivial line bundle M X R over M. An RMC structure for a vector bundle consists of a Riemannian metric
for , whose norm we write as
I and inner product < , >, together with a Riemannian connection [3]. We denote by F the covariant derivative with respect to the connection, by Fj the jth covariant derivative, and by d the Laplacian with respect to the given metric. We use the sign convention giving du = u,;,; + uyy for the standard metric on R2. Choosing some RMC structure on TM and , we define norms IuIIp,k =
I
j=0
(f.
IF'ulP)l/p
for nonnegative integers k, and real p > 1. We define
as the Sobolev space of sections whose covariant derivatives up through order k are in If p = 2, is a Hilbert is a Banach space with respect to space with inner product
(u, v)k =j=0 Z fM
Different RMC structures will yield equivalent norms since M is compact. There are other norms equivalent to the one given above, we will use the one best suited to the problem at hand. We refer often to the standard Sobolev embedding and Rellich Theorems (see [4, pp. 22-23, 28, 31]). If pk > n, we can give Lk(E) the structure of a C`° infinite dimensional For more detailed expositions of possible Finsler manifold, modeled on precedures see [19], [2], [3], and [12]. One step necessary in one procedure is the construction of vector bundle neighborhoods. As we will need this for
527
PERTURBATION THEORY FOR CONDITION (C)
a-local argument in § 8, we give the definition here. For a proof of the existence of vector bundle neighborhoods see [12, § 12]. Definition 2.1. Let u E C°(E). A vector bundle neighborhood of u in E is a vector bundle $ over M such that $ is an open subbundle of E and u E C°($). We can now give a more precise definition of coercivity for functionals on Lk(E). Definition 2.2.
Let E be a C'° fiber bundle over a compact connected ndimensional C- manifold M, and pk > n. A functional J : LP(E) -> R, is called coercive if on any vector bundle neighborhood $ c E, if [DJS. - DJSj](s - sj) -> 0 , and Si IILk(E) is bounded, then si has an L"($) strongly convergent subsequence. Definition 2.3. A map P : C°°(E) C°°($) is called a polynomial di ff er-
ential operator of order k and weight at most w, if for each local representation of P, the jth component of Pu(x) is [Pu(x)], = F,(x, u,(x), ... , um(x), Daui(x)) ,
1 < j al < k ,
where each of the functions F; is a sum of terms of the form l(x, uA), ..., um(x))D11ui1(x) ... Dsqulq(x) + I j9 I < w, [ 12, p. 69]. (An intrinsic dewith 1 < I 1 I < k, and 181 I + finition may be found in [10, appendix II].)
Definition 2.4. A polynomial differential operator is said to be strict, if for some local representation of P the functions li are of the form 2i(x, u1(x), ... , um(x)) = a(x)ui1 ... ui,
.
(Note that this notion is invariant in the base variable of E (in x), but not in the fiber variable (i.e., in u).) In other words, a polynomial differential operator is a polynomial in the derivatives of u whose coefficients may depend on u, but not on the derivatives of u. For a strict polynomial, the coefficients are further restricted to be polynomials in u. Examples. (1) P(u) = a(x)eu(u')' is a polynomial differential operator of order 1 and weight 2. (2) P(u) = a(x)u3(u')2 is a strict polynomial differential operator of order 1 and weight 2. For the rest of the paper when considering LPk we assume 1 < p < co, since then bounded sets are weakly compact. All integration is with respect to the Riemannian measure on M. Notation. We designate an open set Q with compact closure contained in M by Q C C M. D" denotes the unit n-disk, i.e., {x E Rh j jxj < 1}. The tangent
JILL P. MESIROV
528
bundle of a manifold M is written TM. For the unit interval [0, 1], we use I. To designate the space of Lti maps from a manifold M into R we use either of the following : Lti(RM) or Lk(M, R). For an open set Q C R11 we use LP(Q) and Lk(12, R) interchangeably. Lk(E)af (resp. f) denotes the closure in Lti(E) (resp. of the set of g E Lti(E) (resp. which agree with f on some neighborhood of the boundary of M. The closure of Co in is while Co ( is the space of u E Co with compact support in Q. The uniform norm is written 11 . D" = al°llaxi= ax,'- in the usual multi-index notation. To designate weak convergence we use Finally, the k-jet of a section u is written jk(u). j
3.
An example for motivation
The motivation for our perturbation problem comes from classical mechanics. We are given the following data : N : differentiable manifold = configuration space K : Riemannian metric on TN = kinetic energy V : R-valued function on N = potential energy (By abuse of notation we may think of V as acting on TN.) L:
K - V = Lagrangian.
The motion of the system is an extremal of L at a point (p, v) E TN, where p is a position vector and v a velocity vector. We will show that the action integral
E(u) = f L(u) = f K(u, u') - V(u)dt satisfies condition (C), assuming that (i) K(u, u') _ l u'(t) Iu(t) is the metric on configuration space N, (ii) the potential V is a smooth bounded function on N, and (iii) u E Li(I, N)p,Q, i.e., Li paths from P to Q in N.
It is not difficult to show that the energy integral E°(u) = f K(u, u')dt, the "unperturbed functional", satisfies condition (C) by verifying pseudo-properness and coercivity [17]. Assuming this, we consider the perturbed functional E. As we observed in the introduction, it suffices to prove the following two claims to establish condition (C) for E. Claim 3.1. E is pseudo-proper on L'(I, N)P,Q. Proof. Since E° is pseudo-proper on Ll(I, N)P,Q, it is enough to show that if E is bounded on S C Li(I, N)p,Q, then E° is also bounded on S. Now V is bounded, so for all u E S
0 < E°(u) = E(u) + f V(u)dt < E(u) + (cont.) length (I) Claim 3.2.
E is coercive on Li(I, N)P,Q.
PERTURBATION THEORY FOR CONDITION (C)
529
Proof. We will show that if a sequence ui is bounded in Li, and (DE,, - DE,,,j)(ui - u;) > 0, then ui has a strongly L' convergent subsequence. Since E° is coercive and
(DE,,;, - DE.j)(ui - uj) = (DE°U - DE°uj)(ui - u,) I (DVu.
- DV,,,)(ui - uj)dt,
it is enough to show that, for a relabeled subsequence of ui, J
(DV,,,, - DV,,.)(ui - u,)dt - 0 .
The inclusion of Li into C° is compact, so there is a relabeled subsequence ui which converges uniformly. Since V is smooth this means DV, converges, which, along with the uniform boundedness of ui, gives the desired result. q.e.d.
Modifying the above, we get the following more general result. Theorem 3.3. Let J°: Lk(E)a f > R be a Cl functional on sections of a tri-
vial fiber bundle ;r: E > M over a compact manifold M with fiber N, and pk > dim M. Let V : N > R be a smooth uniformly bounded function on the fiber. If J° is pseudo-proper and coercive, then so is J(u) = J°(u) - J V(u). Moreover, if J° is bounded below it is enough to assume V is bounded from above.
II. 4.
PSEUDO-PROPERNESS
Almost J-boundedness
In § 3, we dealt with the pseudo-proper half of the perturbation problem by showing that for a uniformly bounded perturbation 'K (or one which is bounded above), if the perturbed functional J = J° - 'K is bounded on a set S C Lk(E)a f, then the original functional J° must also be bounded on S. This assumption on 'K is too strong. It is enough for V to be dominated by J°, which leads us to the concept of almost J-boundedness. Throughout §§ 4 through 7, unless otherwise stated, we do not assume
pk > n. Definition 4.1. Let J and /i' be functionals. Then V is almost J-bounded if there exist constants 0 < 1 and K such that for all u,
''(u) < OJ(u) + K . The following theorem is an obvious consequence of the definition.
530
JILL P. MESIROV
Theorem 4.2. Let J0: LP(E)a,, - * R be a functional which is bounded from below. If Yl' : Lj(E)af -p R is almost J-bounded, then the perturbed functional J = J0 - Yl' is bounded below and pseudo-proper. Using Theorem 4.2 and the pseudo-properness of the functional J0(u) _ 1I U IP,k, in subsequent sections we will give various conditions on perturbations Yl' which insure the pseudo-properness of the perturbed functional J(u) = J0(u) - 'Y/'(u) = I u I IP,k - Y/'(u)
.
Since the results extend in a straightforward manner to more general functionals J0 which dominate the norm uIIP,k, i.e., I u II
p,k <
cJ0(u) + K
,
(c, K : constants)
,
we will omit the details of the extension. 5.
0-order perturbations
We begin with an especially illuminating and useful example. Let M be a compact Riemannian manifold with nonempty boundary. For u (E Ll(M, R)0, let JM(u) = JJvu I2 = I u 112,1
.
Let V : M X R - * R, and define a functional Yl' by Y/' (u) = f V(x, U) .
Note that Yl' is 0-order since it depends on u, but not any derivatives of u. For j = 1, 2, let A; > 0 be the jth eigenvalue of the Laplacian, with corresponding orthonormal eigenfunctions ¢; E Li(M, R)0i -4o, = 2;c,. Finally, let J = J0 - 'Yl-. Theorem 5.1. (a) If -
V(x, s) < const. + Ks'
for all s E R, and K < 2, then J is bounded below and pseudo-proper. (b) If V(x, s) = const. + Ks2
,
and K > 2, then J is not pseudo-proper. Moreover, if K = 2 for any j = 1 , 2,
(c)
, then J does not satisfy condition (C). If V is continuous, and there are constants r > 21 and c such that
531
PERTURBATION THEORY FOR CONDITION (C)
V(x, S) > c + rs2
(5.2)
,
for all s > 0 (or s < 0), then J is not pseudo-proper. In fact (5.2) need only hold on some open set D C M. Remark. Part (a) of Theorem 5.1 says that if V grows at most quadratically with a rate of growth bounded by 11, then J will be pseudo-proper. We can rewrite this condition on V in terms of an asymptotic growth estimate, i.e., lim sup
si-
V(x, S)
< K < 21
IsI2
In part (b) we see that if V grows quadratically, then the 2, "growth constant" is a sharp bound for the pseudo-properness of J. Part (c) shows the sharpness of the restriction in part (a). In particular, if V grows faster than quadratically even on an open set D C M, e.g., V(x, s) > c, + c2 Is 11 where t > 2, then J will not be pseudo-proper. As a special case, V (X, s) = f (X) ± s3
satisfies (5.2) for s > 0 (for +), or s < 0 (for -). There are similar phenomena in the more general situation. Proof. (a) By Theorem 4.2, it is enough to show that V is almost J, bounded. First we recall that
J Ipur
2, =
inf
-
ueL2(M,R)o
lu 2
>0
.
f Now J
V(x, U) <
J
(const. + Ku2) < (const.) vol (M) + K I Fu I2
.
But K < 1, so we see V is almost f Vu 12-bounded. (b)
First observe that
J
J(u) = f I Vu I2 - Ku2 = - f u(d u + Ku) , and
DJ,,,(v) = 2 f Vu.Vv - Kuv = 2 f u(-4v) - Kuv . Now writing u in an eigenfunction expansion, u = E a1¢l, we see 11a12, J(u) = E (21 - K)a12, and
I u 112,1 =
JILL P. MESIROV
532
Ju z,1 =
Al
K
1-
)2
ale
.
Say K > A1, since 2, > oo, 2N < K < 2,+, for some N > 1. Let
/-K 1
K - .1N+1
cN+
If we define u; = ju, then J(u,) = 0 while 11 u,112,1 > 00 , and hence J is not pseudo-proper. Note that if K = 21, we can let u = 01. This establishes the first part of (b). If K = AN, let u, = jcN + (1 / j) pN+1. Then we see that I J(u,) I G 2N+1 - K, IVJu; II > 0, but 11 u, 1I2,1 > 00 . Hence we see J does not satisfy condition (C). (c) Replacing u by - u if necessary, we need only consider the case s > 0. Pick z 1 01. Then for any a e R,
(5.3)
1141 + z
l'2'1 =
f I V(ao, + z) I2 = a2A1 + f I Fz I2
Also by (5.2)
J(a¢1+z)
J(41 + z) < (A1 - T)a2 + f 1 Vz I2 - c1
Thus for fixed z 101, since y > A1, (5.5)
J(a51 + z) < const.
for all a e R .
There are essentially two cases : as a > oo, either (i) lim J(a51 + z) > const. or (ii) lim J(ao1 + z) = - co . The first corresponds to V(x, s) = 21s2, when J(a5) = 0, and the second toy > Al as can be seen from (5.4). Case (i). Say there is a z 101 such that lim sup J(a51 + z) > const. as a > cc. Then there are a, E R, a; > oo , s uch that J(a, p1 + z) > const.. If u, = a,o1 + z, then by (5.3) and (5.5) we see that 11 u,112,1 > o but I J(u,) const.. Thus J is not pseudo-proper. Case (ii). Assume that for all z 1 441, we have lim sup J(a51 + z) _ - o
as a > 00 . We can pick z, 1 1 such that z; < 1 and f I Vz, I2 > co . Let F,(a) = J(ao1 + z,). Since V is continuous and lz, Ih < 1, V(x, z,) is bounded. Thus f I Vz, I2 > co implies F,(0) > co as j > 0. Consequently for j sufficiently
PERTURBATION THEORY FOR CONDITION (C)
533
large, F;(0) > u, for some constant p. Moreover, F;(a) - - oo as a - oo. Therefore, since Fj is continuous, for each j sufficiently large there is an aj
such that Fj(a;¢, + z;) = p. Now let u; = ajo, + z;. Then J(u;) = u but u; II2,1 - oo, thus J is not pseudo-proper. It is enough for (5.2) to hold only on some open set, since if Q C M and J is not pseudo-proper on L;(Q),, then J is not pseudo-proper on L,I(M),. This follows from the fact that if u E L;(Q),, then it can be extended to u E Li(M)0
bylettingu - uonQandu - OonM - Q. Remark. In the case of V(x, s) = Ks' we can say more than part (b) of Theorem 5.1. In fact we can completely analyze J(u) =
J
I Fu I2 - Ku2 in terms
of K. J satisfies condition (C) if and only if K # A; for j = 1 , 2,
. Further-
more, if K < 2, the critical points of J are minima ; while if K > 2, and J satisfies condition (C) (so K # A), the critical points of J are "saddle points." We will show in Theorem 8.13 that J is coercive for all K, although this fact is not needed here. Proof of Remark. First we show that if J is bounded on a sequence u;, and II FJu; I12,, - 0, then the uj's are converging strongly in Li(M, R)0. In fact, we show I I u;112,, --> 0. If u; _), a1¢t, then II VJu; Jl2,, = E21 (1
- K I2a,z > min (1 -
K 2
K
21a12
since each term on the right hand side is positive. Now because 2,
cc,
1 - K12, --> 1 as 1--> co. Since K is not equal to any eigenvalue, 1 - K/,lt # 0 for any 1, and therefore there is a r > 0 such that 11 - K/,la 1 > r for every
l= 1,2, ..Hence wesee
IIFJu,112,E>-r2E21a12=rZlluil2,,
To study the nature of the critical points of J, we examine the second variation D2J,,(v, v) = 2 Fv Fv - Kv2 = 2 J
J
v(-4v - Kv)
Thus, if v =), b,o, then D2Jw(v, v) = E (2 - K)bl2 .
When K < ,l if v 0 then D2Ju(v, v) > 0, because .l, < 22 < . Hence, if u is a critical point then it must be a minimum. On the other hand, if 2 < K < 2j,,, then we can choose v to make D2J,,(v, v) positive or negative. q.e.d. With this insight we are ready to answer the question of which zero order
JILL P. MESIROV
534
perturbations are almost J, -bounded for J0(u) = 11 u 1 11,, = f I l7xu Iv, u E (Lk)o.
If M has no boundary we must modify Jo to J0(u) = f (Pku l P + I u l P), but the J theorems go through with little change. Let be a finite dimensional C`° vector bundle over a compact connected Cm finite dimensional manifold M with boundary. Choose an RMC structure for TM and . Let Jo : R be a functional defined as follows vlcu lP
J0(u) = f
.
Define constants 2(p, k) by flpxulP
2(p, k) =
inf
ue Lf(E)o
(' lulP
Note that 2(p, k) is the reciprocal of the norm of the continuous linear incluinto Lo(b)o, and thus is positive. sion of Theorem 5.6. Let V be a 0-order differential operator from to RM, i.e., V : -+ R. is a fiber bundle morphism, such that there exist constants A and a such that (5.7)
J
V(u) < A + aJ Jul-'
for all u E LIP() )o
.
If a < 2(p, k), then J(u) = Jo(u) - f V(u) is bounded below and preudo-proper on Proof. The natural modification of the proof of Theorem 5.1 part (a). Theorem 5.8. In the notation of Theorem 5.6, let Jo : f , R and V satisfy (5.7) for all u E f. Then there is a constant r > 0, such that if
a < r, then J(u) = JM(u) - f V(u) 2(p, k)12P. f. Moreover, is bounded below and pseudo-proper on Proof. It suffices to show there exist positive constants B, r such that
(5.9)
Ju
i JIV'`ulP.
535
PERTURBATION THEORY FOR CONDITION (C)
Then the argument is again similar to the proof of Theorem 5.1 part (a). Note this r and the r in the statement of the theorem are identical. In turn, (5.9) follows from the inequality (5.10)
u IP,O
Bl + 1 ui rl
(To obtain (5.9) from (5.10) use (a + b)P < 2P(aP + bP) for any a, b > 0.) Then one finds B = (2B1)P, and r = (rl/2)P. To prove (5.10), fix ¢ E Lk(e)a f. If u r= Lk(e)a f, then v -_ u - ¢ E Lk($)0. Now use the triangle inequality and the definition of 2(p, k) applied to 11 v I. The constant B1 depends only on
and r° > ,(p, k). Thus r = (rl/2)P > ,(p, k)/2P. Remarks. 1. As we remarked before for the case of aM = 0,
J Fku jP
f
is no longer a norm, so we must add f I u I P to J0(u). The same general theo-
rems are true in this case. J 2. The value of the constants ,(p, k) and r will crucially depend on the choice of norms and the RMC structure. Their existence is of course independent of such choices. 3. In the case of M = bounded domain in Rn, we can verify condition (5.7) on V by checking for the following pointwise conditions
(a)
V(x,s)
or
(b)
lim sup V (X' s) G sl_w
SIP
An appropriate version of the above should extend to the general fiber bundle setting. We now present some results in that direction. We consider, as in § 3, the classical mechanics case, i.e., geodesics in the presence of a potential. For the rest of this section, let N be a complete noncompact Riemannian manifold, and p the distance function induced by the metric (if N is compact all smooth potentials are bounded). Let I = [0, 1], and let J0(u) =
J
I u'(t) I Zdt
be a function defined for u E Li(I, N), where u(0) = P, u(1) = Q for fixed P, Q E N, i.e., Li paths from P to Q, written Li(I, N)P,Q. We seek a perturbation theorem of the nature of Theorems 5.1, 5.6 and 5.8. That is, we want to find asymptotic growth conditions on a potential function V : N - R such that
JILL P. MESIROV
536
J(u) = J0(u) - f V(u(t))dt a
is pseudo-proper.
Recall the condition from Theorem 5.1 was lim sup
V(x, S) s-z
where 2, is the first eigenvalue of the Laplacian, and is the "best" constant possible in the inequality
f IuIz<
1
f I pu1Z,
uELi(M,R)o.
Thus in order to handle the case of geodesics, we will find a function on N with
which to compare V corresponding to f(s) = sz on R, and an invariant constant corresponding to A,. For a fixed xo E N, define a function p,,: N - R by p,xo(x) = p(xo, x). Let So be the set of r E R such that for some Br E R,
f
p(x,, u(t))zdt < 1 J0(u) + Br o
for all u E Ll(I, N)P,Q. We will show that the constant To = sup (So) depends only on the metric on N. This constant ro plays the role of 2, in the perturbation theorem. One can show that ro < Go, but since we do not use this fact the proof is omitted. In order to be able to state the asymptotic growth conditions, we must introduce the notion of asymptotic equivalence. Let X be a connected topological space. Definition 5.11. For f : X - R, we say lim sup AX) = a
if for every s > 0 there is a compact set K. C X such that f (x) < a + e for x E X - K,, and no smaller a will do. Definition 5.12. Let V, W be continuous functions on X. V is asymptotically dominated by W if
lim sup ( ___
V(X)
W(x)
)
<1
.
V and W are asymptotically equivalent if each one asymptotically dominates the other.
PERTURBATION THEORY FOR CONDITION (C)
537
Claim 5.13. If V is dominated by W, then for all s > 0 there is a constant K_ > 0 such that
V(x) < K + (1 + e)W(x)
for allxGX. Definitions (5.11) and (5.12). q.e.d. Asymptotic equivalence relates to the notion of almost J-boundedness as
Proof. follows.
Theorem 5.14. Let M, N be Riemannian manifolds, N complete noncompact and connected, and M compact. Let J: Lk(M, N), f -p R, where pk > n. If W : N - R is almost J-bounded, and V : N -* R is dominated by W, then V is almost J-bounded.
Proof. We know J
W(u)
where 0<1 .
Since W dominates V, pick e > 0 such that (1 + e) < 1 /B. Then by (5.13) there is a K. > 0 such that
f V(u) < Ks + (1 + e) f W(u) < K, + K + 0(1 + e)J(u) for all u E Lk(M, N), f. But 0(1 + e) < 1, therefore V is almost J-bounded. q.e.d. The function with which we will compare V, corresponding to f (s) = s2 in the linear case, is p(x0, x)2. We first show that the "comparison" is independent of the specific point x0 E N which we might pick. Theorem 5.15. Given any points x0 and x, in N, then pro and pal are asymptotically equivalent. Proof. Given any e > 0, let K. = {x E N I p(x x) < p(x,, x0)/e}. Then K.
is compact, and for all x E N - K, p(x0, x)
P(x,x)
<1+
The argument is symmetric in x, and x0. q.e.d. We now give a perturbation theorem corresponding to Theorem 5.1 part
(a), for the space of paths L'1(I, N)p,Q. Recall that we seek conditions on V : N -p R such that Y, (u) = fo V (u(t))dt is almost Jo bounded where 1
J0(u) = f I u'(t) I'dt ,
it E Li(I, N)P,Q
Before we state and prove the theorem, we prove two lemmas. The first es-
538
JILL P. MESIROV
tablishes an inequality similar to (5.9), and the second establishes the existence of the invariant ro which corresponds to the first eigenvalue 2, Lemma 5.16. If xo E N, then there are constants B, r > 0 such that (5.17)
p(xo,
Jo
u(t))2dt <
1J r
I u'(t) J2dt + B 0
for all u e LI'(I, N)P,Q. Moreover, r > 1/2. Proof. For any path u E Li(I, N)P,Q, 1
p(u(0), u(t)) < f 1 u'(s) Ids < (f 1 I u'(s) I'ds)1'2 .
(5.18)
0
0
Therefore,
p(xo, u(t)) < p(xo, u(0)) + p(u(0), u(t)) < p(xo, P) + ($i u'(t) 2dt)
.
o
Consequently, p(xo, u(t))2dt < 2p(xo, P)2
Jo
+ 2 JoIu
(t) I2dt
.
Lemma 5.19. Let So be the set of r e R satisfying (5.17) for some B, E R. Then each of the following holds:
S,r0.
(a) (b) If ro = sup (So), then ro > 0 and is independent of xo. (c) If xo=P, then ro> land we can letB=0. Proof. (a) This is immediate from Lemma 5.16. (b) Again by Lemma 5.16 we know ro > 0. It remains to show ro is inde-
pendent of xo. Pick x, E M, x, # xo, with corresponding S, and r,. Note that if r e S, (resp. So), then P < r implies P E S, (resp. S). If r, # ro, say To < T we will obtain a contradiction. Pick s, > 0 such that
ro + a, < T,. Now let s = e,/ro > 0, so s > 0 and ro + roE < T i.e., ro < r,/(1 + s). With s thus chosen, since p(xo, .)' and p(x .)' are asymptotically equivalent, there is a K, > 0 such that (5.20)
Jo
p(xo, u(t))2 < J o Ks + (1 + s)
fl o
p(xi, u(t))2
.
Since To < r,/(1 + s), there is a 8 > 0 such that To < (r, - 8)/(1 + s). Now
= sup (S,), so there is a r in S, such that r + a > T that is, r > Ti - a which implies Ti - 8 E S, by an above note. Combining this with (5.20) we see 1
(5.21)
p(xo,
So
u(t))' < K, + B(1 + s) + (1 + 8) J(u) T, - a
PERTURBATION THEORY FOR CONDITION (C)
539
Now (5.21) implies (r, - 8)/(1 + E) E So, but ro < (r, - 8)/(1 + E) and ro = sup (So). This is a contradiction. (c) This follows immediately from (5.18). Theorem 5.22. If V : N -+ R is asymptotically dominated by ap(x, .)2 for any x e N, a < ro, then the functional
J(u) = fo ( u'(t)1' - V(u(t)))dt
u E Ll(I, N) p,Q
is bounded below and pseudo-proper. Proof. By Theorem 5.14 and 5.15 it is enough to show that ap(x, . )2 is almost Jo bounded for some x e N, where
J0(u) = J0ut(t)2dt. 11 This follows from Lemma 5.19. 6.
(k - 1)-order perturbations
In this section we investigate conditions on a perturbation V, which imply The JiI7 ku P on the space almost Jo boundedness for J0(u) =
results in this section generalize those in § 5, especially part (a) of Theorem 5.1. They apply to the case of arbitrary boundary values with an appropriate change of the constants involved, using the technique of Theorem 5.8. We analyze almost f 1 I7'`u IP-boundedness, for the case of perturbations V which are dominated by some strict polynomial differential operator P of order
at most k - 1, and homogeneous of degree at most p, i.e.,
V(u) G constant + P(u) Locally one can express this condition as
V(x, u, D°u) G constant + Z Aa(x)Dal.... Dazu a
where a = (a
,
a,.) is an r-tuple of multi-indices, 0 < I a1 I G k - 1, and
r G p. One should understand the above condition as an asymptotic growth condition on V. For example: 1. In the case k = 1 and p = 2, where V is a function of the zero jet and lim sup V (x Zs) G a(x) si__
Is
I
we have V(x, u) G b + a(x)u'- for some constant b.
540
2.
JILL P. MESIROV
In particular, there is no growth requirement in the "negative direction"
such as the - eu term in
V(x, u, u') = -eu + a(sin u)u2 + b(u')2 G au' + b(u')'
.
One might ask how necessary is the above restriction on V. We have shown
the necessity in the case k = 1, p = 2 in Theorem 5.1. In Theorem 7.1 we will discuss the necessity of the growth condition for arbitrary p and k. Given the above growth condition on V, we wish to show that V is almost f IFku Ip-bounded, that is, f V(u) G B
J
lVkuIp + coast .
for some B < 1. Clearly it is enough for us to consider the dominating polynomials P(u). Lemma 6.1. Let P be a strict polynomial differential operator from to RM, of order at most k, and homogeneous of degree r < p. Then P extends to into L'(RM). a C`° map of Proof. This is a local question, so we may assume M = Dn, = Dn X R'n, , um), then P(u) is a sum of terms of the and RM = Dn X R. If u = (ul, form A(x)Dl,ua,,
Dazua,z ,
where A E C`°(M, R), 0 < I cei I < k. The map u * D-ua,t is a linear differential operator of order I cei I from to RM, and thus extends to a continuous linear into Lk_ i j i (M, R). Therefore it suffices to show that multiplicamap of R) into '(M, R). The tion is a continuous multilinear map of O+i_1 Lk_
proof of this, given r G p, is a straightforward application of the Sobolev theorems and Holder's inequality, see, for example, [12, Theorem 9.4]. Theorem 6.2. Let P be a strict polynomial differential operator of order k from to RM, homogeneous of degree at most p. Then there is a constant
r > 0 such that for all u E (6.3)
f P(u) <- r f IF kU Ip
.
Moreover, if ro is the greatest lower bound of the set of r > 0 satisfying (6.3), and
ro<11 then f P(u) is almost $VkuPbounded for u e Lk(),.
PERTURBATION THEORY FOR CONDITION (C)
541
Proof. By Lemma 6.1, P extends to a continuous map from Lk(e)o into Lo(R,). Since P is homogeneous of degree p, there is a > 0 such that IPu111,0 < r11u11P,k = r f IVkulp
(If not, there are u, E LP(e)o with II u, IIP,k -p 0 and ''i1 P(u,) II,,o = 1. By continuity,
if u, -* 0 in LPk then P(u,) -* P(0) = 0 in Lo(Rry7).) q.e.d. For Theorem 6.2 to yield an effective procedure, one must find the constant ro more explicitly. We carry this out in the case where P acts on scalar valued functions on a bounded open set Q in Euclidean space. Let
JIDu
II uIIP,, _
be the norm on Lp(Q, R)o, for 0 < 1 < k. Define constants ,(p, k ; 1) for 0 <
1
,(p, k ; 1) = so 2(p, k ;
1)
inf
ueLP(Q)o
u IIP,k > 0 lullp,=
is the reciprocal of the norm of the continuous linear inclusion
of LP(Q)o into LP(Q)o.
Let V(u) = A(x)(Da'u)al (D"r'u)a', where A is smooth, ai < p. Then each of the following holds : There exist constants p1, , pv such that Z 1 / pi = 1, and
Theorem 6.4.
0 < I ai I < k - 1, and (a)
SaiPi
lim sup
SP
isi_-
= Ai < oo
, N. (In fact Ni is 0 or 1.) There is a smallest constant K such that for all
for i = 1, 2, (b)
(6.5)
> K,
f V(u) < r II U IP,k + const .
for all u E Lk(Q),. Moreover, if K < 1, then V(u) is almost II U IIP,k-bounded. K < I AI I_ II
Aj
)llpj
_ 2(p, k; Ia,1)
ai < p, then we can make K arbitrarily close to 0, but the con(d) If stant in (6.5) may go to + oo. Proof. (a) We first observe that if Z ai = p, then we let pi = p/ai and ai < p, then Z ai/p < 1, hence we can pick 1/pi > ai/p such Ni = 1. If 1/pi= 1, and that
542
JILL P. MESIROV
Islaipi
lim sup
sl--
(b) (c)
0.
ISIP
This follows immediately from Theorem 6.2, but is proved again inn(c). We observe that A(x)(Daiu)ai ...
f(' V(u) 1
(6.6)
1
< 1 A IIw f I (Dniu)ai ... (DQNu)ati
.
Now pick pi's as in (a), and constants pi, which we will choose later, such that f jr , pi = 1. Since Z 1 /pi = 1, we know that for any x1, , xN, xl ... xN = ,Ulxl ... UNXN < (pixi)PI /Pi + ... + (p1VxN)PN/PN
Combining this with (a) and (6.6) we find
f V(u)
A
(6.7) < II A Ii [
+ ... +
JD1uIP
IIAI! - [
NNfuN" f
+ const
PN Alul
P12(P, k ;
_ +.... +
I a,PN2(P, k
aN
j u lp,x + const
So our first approximation of K is (6.8)
K < J I A I I_ [
P 2(P, k ;
SNP
+ ... +
1
N1
PN2(P, k ; 1 aN )
1 a l)
Now we will pick the pi's to minimize the right hand side of (6.7). Using standard techniques we get (jIVfj
l
Cj/ /C1'i
,0i
,
where ci = j3 /2(P, k ; l ai I). Substituting back into (6.8) and using the fact that
Z 1/pi = 1, we get
K
j
I/pj
\2(p,k;Iajl)I
Note the constants in (6.7) depend on IIA 1-, and arise from the lim sup statement of (a). 0 for (d) is clear from the above and the fact that if Z ai < p then
i=
Remark. Condition (b) is stated as it is because K might be - co in which case we cannot use it in (6.5).
543
PERTURBATION THEORY FOR CONDITION (C)
It is straightforward to extend Theorem 6.4 to V's which are a sum
V(u) = E
(6.9)
(D°N(a.a)u)aN(a.a,
Aa,a(x)(Daiu)ai ...
where the Aa,o, are smooth, each a is an N(a, a)-tuple of multi-indices with
0 < l ail < k - 1, and each a is an N(a, a)-tuple such that E ai < p. The result of this extension is Corollary 6.10. With V as in (6.9) let Ka,a be the K of Theorem 6.4 corresponding to the Aa,a monomial. If E Ka,a < 1, then V is almost I u II P,,-bounded. Remark. It is clear how one extends Theorems 6.2, 6.4, 6.10, to kth order perturbations.
(k - 1)-order perturbations: Growth restrictions The results of this section continue the generalization of Theorem 5.1. First we focus on part (c), and consider the question of the necessity of restricting ourselves to perturbations dominated by polynomials which are homogeneous of degree at most p. The idea is to show that if the perturbation V grows faster than a polynomial in the derivatives, homogeneous of degree p, then 7.
J(u) = 1 l u l I pp, k- f V (u)
is not pseudo-proper. Next we generalize part (b) of Theorem 5.1 by showing that if V is homogeneous of degree p, then there is a constant K such that if
K > 1, then ul
P,k -
f V(u)
is not pseudo-proper. Let
Mu) = lull p,k = $VkuP on Lk(g)0, and
1/'(u) = f V(u) , where V is a (k - 1)-order differential operator from
to RM, continuous on
L)0.
Theorem 7.1. If there is a continuons functional 1/',(u) on set Q C C M, and 1jr e Co ( lQ) such that (7.2) and
Y/'(,i,jr) > c, + Y/',(,iiJr)
for all 2 > 0 ,
an open
544
JILL P. MESIROV
lim 2--
(7.3)
V,00 = + 2P
then J(u) = Jo(u) - 1/'(u) is not pseudo-proper on Proof. Pick an open set Q. C C M, with Q n Q. = 0, and 0j e Co ( IQ) such that (7.4)
II j IIPo,k- < 1
(7.5)
J0(0j) = 1I 0j IIP,k -> 0
,
We will show there are constants aj, such that J(u j) = 0 and 11 u j IIp,k = J0(u j) cc, where u j = 0j + a j,Ji, thereby establishing that J is not pseudo-proper. Observe that
(7.6)
Jo(u j) = fDo I Fk0j I P + I aj I P f , IF'* Iv = J0(Oj) + aj I PJO(*) > JO(Oj)
so Jo(u j) - co as j , co. Also if 0 represents the 0-section, (7.7)
Y/'(uj) = f120 V(0j)
_
+f
(0j) +
D
V(aj1G) +
(aj*) + c
V(0)
J.,Y-Qo-Q
,
where
c=
-L, V(0) - f
YI-9
is a constant independent of j. Since V is continuous, and 10j (7.8)
f V (0j)
V(0) + f
V(0)
1, we know
forallj=1,2,
.
Hence assuming aj > 0, we see by (7.7), (7.8), and (7.2) that (7.9)
-K + c + c, + -1-,(aj*) < 'Y''(uj) < K + 'Y''(aj*) + c .
Using a continuity argument, we now show there are aj > 0 such that J(uj) = 0, i.e., 'Y''(uj)/J0(uj) = 1. By (7.6), (7.7), and (7.9), we have (7.10)
-K+c+c,+ 1,`(aj*) < Jo(0j) + I ajNPJ0W
1''(uj)
J0(uj)
(ajy)+c Jo(oj)
Since J0(gj) , cc, pick j so large that J0(gj) > K + c + '/''(0), where 0
denotes the 0-section. Then letting aj = 0, we see
PERTURBATION THEORY FOR CONDITION (C)
545
K+c/+Ii'(0) <1 Jo(¢ j)
and hence for a j = 0, '1'-(uj)/J0(uj) < 1
.
For a fixed j, by condition (7.3) on Yi', we see that
-K+c+c,+Yi',(aj*) J0(0j) + aj l'JoM as a j --> co. Therefore, there is an a j > 0 such that 1 < 1" (uj)/JO(uj) .
Since Yl'(u j) /J0(u j) is a continuous function of a j, for j sufficiently large
there are aj > 0 such that Y/'(uj)/J0(uj) = 1
i.e., J(uj) = 0.
q.e.d. What sorts of functionals Yi' satisfy the conditions of Theorem 7.1? To answer this we give several examples. Example 1. Let M be a bounded open domain in Rn , andre = M X R so
we are considering real-valued functions on M. Say ''u) = J V(u). If there exist constants ca > 0 and not all zero, and qa > p, such that on some open
set 0 C M, V(u) > constant +
ca(D"u)9a
then f IV'`up - Y1 '(u) is not pseudo-proper on L3(M, R)0. This follows imme-
diately from Theorem 7.1, with Z ca(D"u)(1°
1,1
Example 2.
Since 11'(u) =
J
V(u) is only determined up to integration by
parts, we need to assume more than V(u) > constant + Z Pj(u) where the P, are strict polynomial differential operators homogeneous of degree greater than p. This can be seen clearly in the following : (a)
Say J0(u) =
= 3u'u'. Then
Jo
(u")2 on L'2([0, 1], R)0. Say 11'(u) =
r
J
V(u) for V(x, u, u')
546
JILL P. MESIROV
K(u) = 3
f1
Jo
u2u' = 10 (u3)' = 0 J
Hence Jo - -//- is pseudo-proper, even though YK is a cubic polynomial in u and u'.
Say J0(u) = f (u"')2 on L3([O, 1], R)o. Let Yl-(u) = f V(u) where V(x, u, u', u") = u3u". Then l (b)
'Y/-(u) = f u3u" = - 3 f u2(u')2 < 0
.
Hence Jo - 'Y1- is pseudo-proper, despite the fact that V is quartic. Example 3. Here we show the additional assumption needed on -//-, to eliminate the difficulties exhibited in Example 2. Say J0(u) =
set 12cM,
J
V ku;P as in Theorem 7.1, Y" (u) = f V(u), and on an open
V(u) > constant + P(u)
,
where P is a strict polynomial differential operator homogeneous of degree
q>p If there is a section' E Co (e 1t,) such that
P(*) > 0 ,
(7.11)
then f IVkuFP - -K(u) is not pseudo-proper on Lk(ee)o.
This is immediate from Theorem 7.1. One cannot satisfy (7.11) in Examples 2a and 2b. Now we consider the special case of the above, where the perturbation arises from a Lagrangian V which is a strict polynomial operator, homogeneous in u of degree p. We investigate the pseudo-properness of functionals of the form J'(u) = J0(u)
II u II1>>k -
JV(u)
>0.
First, we may as well assume that Y/'(u) is not bounded above, because in this case J,(u) is pseudo-proper for all nonnegative 7). Thus for some u E L'(ee)o, ,Yl-(u) > 0. Let
K = sup K(u)
,
Jo(u)
where the sup is taken over all u E LM), for which P'(u) > 0. Then K > 0,
PERTURBATION THEORY FOR CONDITION (C)
547
and K < 00 since by Theorem 6.4, (6.7) and the homogeneity of -Y/', there is a constant c < oo such that 17(u) < cJ0(u). K corresponds to the reciprocal of the first eigenvalue of the Laplacian, which plays a crucial role in Theorem 5.1. Theorem 7.12. If ri > 1 /K, then J,(u) is not pseudo-proper. Moreover, this holds for ri = 1/K if K is attained for some u. is homogeneous, there is a u, EA 0 in Lf(g)0 Proof. Since 1 / ri < K, and such that 1 /ri < '11'(un)/J0(un). Thus we see
J,(u,) = J0(un) - ri'Y/'(un) < 0 .
(7.13)
The argument is now similar to Theorem 7.1. It is possible to construct a smooth one-parameter family ¢, E Lk(e)o such
that u - 0, and < 1, as t> oo, (c) un+0tEA 0. Consider the function f (t) = J,,(u,, + cc), which is a continuous function of t since the ct's vary smoothly. Since 0, - 0, we have f(O) = J,(un) < 0. As t > o, f (t) > oc because the continuity of V and (a) imply 'Y/'(u, + 0) is bounded, while (b) implies J0(u, + ct) > o c. Thus there is a c such that f (r) (a) (b)
=0, i.e., a c, withMMun+o jIp,x#0andJ,(un+cz)=0. Using the homogeneity of 'Y/' in the usual way, we get a sequence v; _ j(u, + 0;) showing J, is not pseudo-proper. Theorem 7.14. If K = sup -11'(u)/J0(u) taken over all nonzero u E Lf(e)o is nonnegative, then J,(u) is pseudo-proper for 0 < ri < 1 /K. Proof. Since 7 (u) < KJ0(u) for all u E Lll(e)o, we see r'11^(u) < r2KJ0(u) q.e.d. for ri > 0. If r2K < 1, this implies ri,Y/'(u) is almost J, -bounded. It is likely that a more precise version of the above two theorems is true which includes the case when K might be negative. Note that the above theorems generalize almost immediately to the case where V is an in homogeneous polynomial of degree p, since the terms of lower degree homogeneity always preserve the pseudo-proper condition (see Theorem 6.4 (d)). III. 8.
COERCIVITY
Perturbing coercive functionals
We consider the following problem. Say a functional J0(u) = f Y (u) is co-
ercive on Lk(E), where Y is a differential operator from E to R. of order k. If V is an operator of order k - 1, what conditions on a perturbation '//'(u) = f V(u) insure that J(u) = J0(u) - 'Y/'(u) is also coercive on Lk(E)? We will
548
JILL P. MESIROV
only consider the case where Y is a polynomial differential operator of weight pk, and V is also a polynomial (not necessarily strict) differential operator. The weight restriction on Y is placed to insure that Y extends to a C'° map from Lk(E) into Lo(R,M) (see [12, pp. 69-77]). In this case Theorem 8.6 gives the optimal result : as long as we stay within the weight restriction imposed by the original Lagrangian Y, all lower order polynomial perturbations preserve coercivity. Since we are working in a fiber bundle setting, we must assume pk > dim M throughout this part of the section. In the case of a vector bundle , the spaces are Banach spaces, and
thus the pk > dim M assumption is not required. In this situation, if we assume the perturbation V is polynomial in u, i.e., V is a strict polynomial differential operator of order at most k - 1 and degree at most p, then we can get explicit algebraic conditions on V to preserve the coercivity condition (see Theorem 8.13). From Definition 2.2 of coercivity we can reduce our considerations to vector bundles
over compact M. Further, if a perturbation Y ,-(u) = f V(u) satisfies
the following condition : (8.1)
if s; - s in
then (DY', - D'Y1'8)(si - s,)
-0,
and Jo is a coercive functional, then it follows that J = Jo - '//- is also coercive. In fact, using a partition of unity argument, it is not difficult to show that it is enough to check condition (8.1) locally. Thus we can reduce the perturbation problem of coercivity to the verification of (8.1), on vector bundles = Q X RI where Q C R" is a bounded open domain. Indeed, by composing V with coordinate functions we reduce further to the case = Q X R. Therefore we will state and prove our theorems in this setting with the understanding that they hold in the general fiber bundle case. Consider a functional J,: Lk(Q, R) > R which is coercive, and perturbations of the form 11-(u) =
J
V(jr(u)), where pk > n, V is say C', and 0 < r < k - 1.
We are looking for conditions on V, so that it satisfies (8.1), where, in coordinates, D,11 is of the following form : (8.2)
DY
e
J o<
(si, D°si) (DT7))
and1
si - s in Lk, then D°si -> D's uniformly for 0 < a < r. This together with the continuity of DV gives condition (8.1).
PERTURBATION THEORY FOR CONDITION (C)
Corollary 8.4.
is C', then J. Corollary 8.5.
549
If V only depends on the 0-jet of u, i.e., V : R - R, and is coercive.
If V depends on the (k - 1)-jet of u, Q C Rt (i.e., n = 1),
and V is C', then J. -
is coercive.
We now restrict to the case of polynomial perturbations V(j,r(u)), 0 < r <
k - 1, on u e LP(Q, R) where Q c Rh and pk > n. As we said before our result is the best possible since we want to perturb polynomial Lagrangians of weight pk on Lk by lower order polynomial Lagrangians keeping J. smooth from Lk(E) into LI (RM). Theorem 8.6. Let V be a polynomial differential operator of order at most
k - 1 on Lk(Q, R), where Q c Rh and pk > n. If the weight (V) < pk, then Y-(u) = f V(u) satisfies condition (8.1). Thus J = Ja - Yl- is coercive. Proof. It is enough to consider V of the form
V(u) = f(x, u)D"u ... D"°u
,
where f is smooth, and the ai are multi-indices, not necessarily distinct, 1 <
ai I < k - 1, and weight (V)
ai I < pk, since V is a sum of such
terms. In fact, for ease in exposition we will do the case of only two distinct multi-indices since there is no great difference in the proof for the more general case. Thus, let V(u) = f(x, u)(Dru)s(D'u)t
1 < r j, 181
hi - h in Lakllrl
(8.7) and
gi- ginLpk/a Writing o^f(x, u)/oo^u = f2(x, u), we have (D-11-,,, - D'11-
)(ui - uj)
J (f2(x, ui)hisgit - f2(x, uj)hjsgjt). (ui - uj) (8.8)
+ s $[f(x, u)hz-1git - f (x, uj)h3-'gjt] (hi - hj)
+ t $[f(x, ui)hisgz-1 - f(x, uj)hjsgj-1] . (gi - gj)
550
JILL P. MESIROV
We will show that each of the three terms of (8.8) tends to 0. For the first term we observe
J (f2(x, ui)hisgit - f2(x, uj)hjsgjt)(ui - uj) IIf2(x, ui)hisgit - f2(x, uj)hjsgjt 11 ,, II ui - uj 11
.
But ui - u in LP and pk > n, so by the Sobolev and Rellich theorems, the ui e C°, and ui - u uniformly ; therefore I ui - uj Ilw - 0. Thus it is enough to show that f2(x, ui)hisg2t I'll,, is bounded. Since f is smooth and the ui's are uniformly bounded, there is a constant K > 0 such that (8.9)
f l f2(x, ui)hisgit I <- K f I hisgit I
By (8.7) we know that the hi are bounded in Loklr' and the gi are bounded in Lok"'. Since 1/d
1/c
f
Ihi"g2tl < (f JhLlsc)
(f Igiltd)
,
where 1 /c + 1 /d = 1, expression (8.9) is bounded as long as sc < pk/ I r l and td < pk/I31, i.e., as long as
1 = 1 + 1 > Slrl+tlAl c
d
pk
which is true since weight (V) = s I ;r I + t 13 I < pk. We show the second term tends to 0 in a similar way (8.10)
f [f(x, u)hz-Igit - f(x, uj)h;-'gjtj(hi
- hj))
< I f(x, ui)hi-'git - f(x, u j)h;-'gjt 11 ,0 11 hi - hj
where 1 / r + 1 / q = 1. By (8.7) we know that 1I hi - h j q,° - 0 as long as IIr,, is q < pk/ I r I. It remains to find conditions on r such that I f (x, bounded. Since f is smooth and the ui's uniformly bounded, there is a constant
Kr > 0 such that (8.11)
f If(x, ui)h?-lgi` IT < Kr f I hz-lgit r
Since ltrd)1/d
f I hsL-lgit IT <( 1\1J
I
hi(s-1)rc)1/c1\I
1J
I gi
551
PERTURBATION THEORY FOR CONDITION (C)
for 1 /c + 1 /d = 1, using (8.7) we see expression (8.11) is bounded as long as (s - 1)rc < pk / I r l and trd < pk / 181, that is, as long as 1
=
1
+ 1 > (s - 1) TI dr
cr
r
pk
+ _L151
pk
.
Combining this with the previous condition on q, we see expression (8.10) tends to 0 if 1
q
+ 1 > I7I + (S - 1) Irl + t ICI r
pk
pk
pk
that is, if pk > s I r 1 + t I a = weight (V) .
The third term tends to zero by the same argument used for the second
term. q. e. d. Theorem 8.6 tells us that as long as we remain within the weight restriction imposed by the original functional, all lower order polynomial perturbations preserve coercivity. However, by "milking" the Sobolev theorems we can obtain more precise conditions on perturbations V which insure the preservation
of coercivity. For example, if p > n, from Theorem 8.3 we know that any C' function V of jk_,(u) will preserve coercivity. We now briefly explain how one gets the stronger results. First observe that using the full power of the Sobolev and Rellich theorems we can replace (8.7) by :
If ui u in Lk, Iri < k - 1, and p(k - I r) > n, then Drui - Dru uniformly, and hence in Lo for all a. If, however, p(k - r) < n, then Drui - Dru in Lo for all a < pn/[n - p(k - I rI)l Using this fact, and the methods used in the proof of Theorem 8.6 we can prove the following theorem, from which Theorem 8.6 follows as a corollary. Let V(u) = f (x, u)(DT'u)S1 ... (DTNu)SN on Lf (Q, R) where Q C R", pk > n,
and the Ti are distinct multi-indices 1 < I ri < k - 1. Let ai = pn / [n pp(k - I riI)].
Theorem 8.12.
If E (si / ai) < 1, where the sum is taken only over those
i for which p(k - I ri I) < n, then r(u) = J V(u) satisfies condition (8.1) and therefore preserves coercivity. Remark. One can interpret Theorem 8.12 as saying that as long as r(u) is well defined, i.e., as long as V(u) E Lo(Q, R), then Yl'(u) satisfies condition (8.1). What about the case where V is a strict polynomial operator, i.e., polynomial in u? In this setting we do not have to assume pk > n, and using the
same method as in Theorems 8.6 and 8.12 we obtain the following result for vector bundles . which extends to the spaces
552
JILL P. MESIROV
Let Q, ai, and ri be as for Theorem 8.12, except that now 0 < I ri I < k Let
V(u) = f(x)(Dnlu)sl ... (Drvu)" , Theorem 8.13.
1.
u e Lk(Q, R)
If Z (si l ai) < 1, where the sum is taken only over those
i for which p(k - rip < n, then K(u) = f V(u) satisfies condition (8.1) and thus preserves coercivity. J Theorem 8.13 has many implications for functionals arising from strict polynomial Lagrangians, acting on real or vector valued functions on a compact manifold. For example, for the special case which we considered in Theorem 5.1, J(u) = f I Fu IZ - V (x, u)
,
Theorem 8.13 implies that for any quadratic strict polynomial perturbation V, J is a coercive functional regardless of the dimension of the manifold M. Bibliography S. Agmon, Lectures on elliptic boundary value problems, Math. Studies No. 2, Van Nostrand, Princeton, 1965. H. I. Elliason, Geometry of manifolds of maps, J. Differential Geometry 1 (1967) 169-194.
Variation integrals in fiber bundles, Global Analysis (Proc. Sympos. Pure Math. Vol. XVI (Berkeley, Calif., 1968), Amer. Math. Soc., 1970, 67-89. A. Friedman, Partial differential equations, Holt, Rienhart and Winston, New York, 1969. W. B. Gordon, Physical variational principals which satisfy the Palais-Smale condition, Bull. Amer. Math. Soc. 78 (1972) 712-716. P. Hartman & G. Stampacchia, On some non-linear elliptic differential-functional equations, Acta Math. 115 (1966) 271-310. J. L. Lions, Quelques methodes de resolution des problemes aux limites non lineaires, Dunod, Paris, 1969. J. L. Lions & E. Magenes, Non-homogeneous boundary value problems and applications, Vol. I, Springer, Berlin, 1972. J. P. Mesirov, Calculus of variations: perturbations preserving condition (C), Bull. Amer. Math. Soc. 80 (1974) 1260-1264. [10]
[11] [12]
--, Perturbation theory for the existence of critical points in the calculus of variations, Dissertation, Brandeis University, 1974.
, A cl assificati on for po lynom ia l L agrangians, to appear in Indiana Univ. Math. J. R. S. Palais, Foundations of global non-linear analysis, W. A. Benjamin, New
York, 1968. [13]
Luste r nik - S ch n i re lman th eory on Banac h manifolds, Topology 5 (1966)
115-132. [14] [15]
, M orse th eory on Hilb ert manifolds, Topology 5 (1963) 299-340. R. S. Palais & S. Smale, A generalized Morse theory, Bull. Amer. Math. Soc. 70
[16]
J. Schwartz, Generalizing the Lusternick-Schnirelman theory of critical points,
(1964) 165-171.
PERTURBATION THEORY FOR CONDITION (C)
[17] 118] .[19]
553
Comm. Pure Appl. Math. 17 (1964) 307-315. S. Smale, Morse theory and a non-linear generalization of the Dirichlet problem,
Ann. of Math. (2) 80 (1964) 382-396. G. Stampacchia, On some regular multiple integral problems in the calculus of variations, Comm. Pure Appl. Math. 16 (1963) 383-421.
K. Uhlenbeck, The calculus of variations and global analysis, Dissertation, Brandeis University, 1968. UNIVERSITY OF CALIFORNIA, BERKELEY
J. DIFFERENTIAL GEOMETRY 12 (1977) 555-564
THE MANIFOLD OF THE LAGRANGEAN SUBSPACES OF A SYMPLECTIC VECTOR SPACE TAKASHI SAKAI
1.
Introduction
Let (Ell', a) be a real 2n-dimensional symplectic vector space with symplectic form a, i.e., a is a nondegenerate skew-symmetric bilinear form on E. Then an n-dimensional subspace 2 of E will be called a Lagrangean subspace if a 0 holds. The set A(E) of all Lagrangean subspaces of (El", a) has a structure of n(n + 1)-dimensional compact connected regular algebraic variety. If we put dk(d) : = {Fe E A(E)I dim (2 fl Fe) = k} for 2 E A(E), then A°(d) is a cell (i.e., diffeomorphic to Rn1n+11i') for any 2 E A(E). Moreover Z (2) : = Uk,1 A"(2) is an algebraic subvariety of A(E), and defines an oriented cycle of codimension one, whose Poincare dual is a generator of H'(A(E), Z) = Z and defines the Maslov-Arnold index [1], [3], [4]. This index plays an important role in the proof of Morse index theorem in the calculus of variations [4]. In the present note, we shall give a differential geometric characterization of Z (2), i.e., by introducing an appropriate riemannian metric on A(E) we shall show that T, (2) is the cut locus of some p e A(E) and A°(2) is the interior set of P. In
fact, take a basis {ei, f;} (1 < i, j < n) of E such that a(ei, e;) = a(fi, f;) = 0 and a(ei, f;) = -8i;. Then with respect to this basis (E, a) may be identified with (RIn, °), where R'n = {(p, q) I p, q E Rn} is 2n-dimensional euclidean space
with the canonical inner product < , >, and a°((p, q), (p', q')) : =
556
TAKASHI SAKAI
2.
2.1.
) be the Lie algebra of U(n) (resp. 0(n)). We put
Let ( (resp.
B.ij: =Eij -Eji
Preliminaries
,
Cij: _
(Eij + Eji) ,
Ai: = 1/fCii
,
where Eij denotes the n x n-matrix whose r-th row and s-th column are given by di,djs. Then 0 may be considered as a real Lie algebra with basis {Bij(1
< i < j < n), C.ij(1 < i < j < n), A.i(1 < i < n)}, and we have the vector space direct sum ( = 9J2 + , where we put 9X: _ {Ai(1 < i < n), Ci;(i < j)} and : = {Bij(i < j)}. Now we define an inner product Q on @ by Q(X, Y) : - 2 trace XY. Then {Ai, Bij(i < j), C.ij(i < j)} forms an orthonormal basis of C with respect to Q. We shall give the Lie multiplication table.
0, [Ai, A it [Ai, Bjk] = 'v/-2{dij Cik - dikCij} [A i, Cjk] =- v1 2 {dijBik + dikBjj}
(2.1)
[Bij, Bkll = -dikBjl + diiBik + 6JkBil - djlBik [B'ij, Ckl] = -dikCil - dilCjk + djkCil + djlCik [Cij, CklJ = -dikBjl - d'iIBjk - djkBil - djlBik
If we define an involutive automorphism s : U(n)
U(n) by ( A B
-AB)J
(A - A) denotes the real representation of an element A)' where of U(n), then the fixed point set of s is O(n) and dsei,,, = id.,, dSe y = -id, does hold. Since 9) is simple, it contains no nonzero ideal of 0. Finally Q is a ds-invariant, ad ()-invariant positive definite bilinear form on 0. We may define a riemannian structure g on U(n) /O(n) by restricting Q to M X 9 and then translating with U(n). Thus (M = U(n)/O(n), g) is a riemannian symA
(-B
metric space with an oily (orthogonal involutive Lie algebra) (C, ds, Q). Note that t : = {A.i(1 < i < n)} forms a Cartan subalgebra of the oily (Ci, ds, Q) (i.e., maximal abelian subalgebra in V), and the center of Csi is generated by
c : = (A, +
+ A)/fn. Now let r : U(n) - U(n)/O(n) be the canonical
projection and put o = ;r(e) which may be identified with 2, As usual we shall identify 9 and the tangent space T0M via the map dhre. We denote by rg the left translation on U(n)/O(n) by an element g e U(n). 2.2. Lemma. Let 17 denote the covariant differentiation of Levi-Civita connection of M with respect to g, and let R(X, Y)Z: = v[S YIZ - [F,, FYIZ be the curvature tensor. Then the curvature tensor at o is given as follows :
R(Ai, Aj)Ak = R(Ai, Aj)Ckl = 0
,
R(Ai, C3k)AI = 2{dijd'ilCkl + dikdilCjl - (dijdkl + dikdjl)Cil} ,
MANIFOLD OF LAGRANGEAN SUBSPACES
" fl ij
R(Ai, Cjk)Cim =
(2.2)
/
R(Cij, CUM. =
557
8ijoimCkd + 3ik3i1Cjm. + 8i1OifnCjl
- (oij3km + oikojm)Cii - (3ijdki + oikoji)Cim} "{-(ojkoim + 3j18km,)Ci7n - (oikolm + 8il8km)Cjm (oiiojm+(3 iioim)Ckm+(dik3,im+d jkdim)CZn }
R(Cij, Cki)Cpq
-(jAq + 8ji(kq)Cip
9
(3jk5ip + 3jldkp)Ciq
- (oikoiq + oiiokq)Cjp - Oikdip + oii3kp)Cjq l jilip)Ckq ((oiiojq + ojioiq)Ckp + (/ loikdjq + (3jk.3iq)Cip + Oikojp + (jk(ip)Czq
Proof. Direct calculation by the formula R(X, Y)Z = [[X, Y], Z] for X, Y, Z E St, [5]. 2.3. Proposition. We denote by K(a) the sectional curvature for the plane
section a. Then we have 0 < K(a) < 4 for all a. Proof. Since M is homogeneous, we may restrict our attention to T0M. Let {U, V} be an orthonormal basis for a, and let 1' be a Cartan subalgebra containing U. Then there exists an h E SO(n) such that Ad h(X) = W. We may assume a = {Ad (h)X, qAd (h)Y}, where X = Ei aiAi, Y = E QiAi + j p
K(a) = Q(R(U, V)U, V) = Q(R(X, Y)X, Y) = 2 E (ap - aq)2Pq 2 Max I ap - agl2 < 4 p
p
where the equality holds if and only if X = (Ap - Aq)/f and Y = Cpq for some p < q. 2.4. Now we shall review the notion of cut locus and conjugate locus of a point of a riemannian manifold. Let (M, g) be a compact riemannian manifold, and let Exp denote the exponential mapping. Let X be a unit tangent
vector at x E M. Then t - Exp tX is a geodesic emanating from x with the initial direction X and parametrized by the are length. t0X (resp. Exp t0X) is called a tangential conjugate point (resp. conjugate point) of x along the geodesic t - Exp tX, if there exists a nonzero Jacobi field J(t) along t - Exp tX such that J(0) = J(to) = 0. Next, t0X (resp. Exp t0X) will be called a tangential cut point (resp. cut point) of x along t - Exp tX, if the geodesic segment t -p Exp tX (0 < t < s) is a minimal geodesic for any s < to, but t -p Exp tX
(0 < t < s) is not a minimal geodesic for any s > to. Then it is easy to see the following : Assume that Exp t0X is a cut point of x along a geodesic t , Exp tX, which is not a conjugate point. Then there exists a unit vector Y E T yM, Y # X, such that Exp t0X = Exp toY. (Tangential) conjugate locus (resp. (tangential) cut locus) of x is defined as the set of (tangential) conjugate
TAKASHISAKAI
558
points (resp. (tangential) cut points) of x along all the geodesics emanating from x. Finally the interior set Int (x) of x is defined as M\cut locus of x. Let a(X) be a positive number such that a(X)X is the tangential cut point of x along t - Exp tX. Then the exponential mapping Exp maps {tX ; X E TIM, g(X, X) = 1, 0 < t < a(X)} diffeomorphically onto Int (x). Thus Int (x) is a cell for any x E M. Now for our manifold M = U(n)/O(n), note that every geodesic t - Exp tX emanating from o with the initial direction direX(X E 9J2, g(X, X) = 1) may
be expressed in the form t - exp tX . o, where exp denotes the exponential mapping from the Lie algebra Ci to U(n), [5]. Then we get 2.5. Proposition. Tangential conjugate locus (resp. tangential cut locus) of o is a hypersurface of the revolution about the line generated by c = and may be obtained by rotating the tangential con+ A,t) (A, + jugate points (resp. tangential cut points) in a about the above line by the action of Ad (SO(n)).
Proof. We put : = {C Jk(1 < j < k < n)}. Since dir o Ad hX = drh, o dir(X) for h E SO(n) and X E 9JZ, Ad (SO(n)) acts on 9J2 as an isometry group and transfers tangential conjugate (resp. cut) points into tangential conjugate (resp. cut) points. By (2.1) we have ad 00a = and consequently Ad (SO(n))9C = V. Moreover since c = (A, + - + An)/.,/-n belongs to the center of Ci, Ad (SO(n)) leaves c invariant and maps the orthogonal complement of c in a onto the orthogonal complement of c in Jt. 3.
Conjugate locus
In this section we shall determine the tangential conjugate locus of o. By Proposition 2.5, it suffices to consider the tangential conjugate locus along a
geodesic t , Exp t(Ei aiAi), ),i az = 1. 3.1. Proposition. Let X = Zi aiAi be a unit vector in lt. Then the symmetric linear transformation of Jt which is defined by V - R(X, V)X has the following eigenvalues: 0 with the eigenspace 2f and 2(aJ - ak)' with eigenvector CJk(1 < j < k < n). The first tangential conjugate point of o along a
geodesic t - Exp t(Ei aiAi) is given by (Min,<, 7r/(f I a; - ak )) ',i aiAi Proof. (3.1)
The first assertion is clear, because by Lemma 2.2 we have
R(X, AJ)X = 0,
R(X, CJk)X = 2(ag - ak)2CJk
Next take an orthonormal basis {Ai, CJ,} of T0M = 9JZ. By parallel translating {Ai, CJ,} along t -p Exp tX, we have an orthonormal frame field {Ai(t), CJk(t)} along the above geodesic. Let J(t) : = 'i ai(t)Ai(t) + EJ
MANIFOLD OF LAGRANGEAN SUBSPACES
559
(d2/dt2)bjk(t) + 2(aj - ak)2bik(t) = 0
(d'"ldt2)ai(t) = 0 ,
with ai(0) = bjk(0) = 0. So we have
J(t) = t Ei aiAi(t) + E bjk(sin j
I aj - akI t) CJ k(t)
for some constants ai, bjk. Then J(to) = 0 holds for some to > 0 if and only
-2 1otj-akI to=0forsome j
Cut locus
First we shall give the following lemma. 4.1. Lemma. Let t0X be the tangential cut point of o along a geodesic t -
Exp tX, where X = E aiAi E W, 7, a = 1. Then either t0X is a tangential j3iAi E 2X, conjugate point of o along t - Exp tX or there exists a unit Y = X, such that Exp t0X = Exp t0Y. Proof. Suppose that t0X is not a conjugate point. Then there exists a unit vector Z E M such that Exp t0X = Exp t0Z and Z X. We shall show [X, Z] = 0. In fact, suppose [X, Z] # 0. We may assume Z X. Since M is a symmetric space, Exp t0X = it exp t0X holds and we have exp t0X = exp (t0Z)h for some h E 0(n). Then exp (- t0X) exp (sZ) exp (t0X) = Ad h-' exp (sZ), and consequently Ad (exp (-t0X))Z = Ad h- 1Z holds. But Ad (exp (-tOX))Cjk Y
= Cjk cos ajk + BIk sin ajk with ajk = ziAi + Ej
/(aj - ak)to. So if we put Z =
[X, Z] = -,/-2N G E zjk(aj - ak)Bjk j
i
= Adh-'(E ziAi + E zjkCjk) E w i
j
Since [X, Z] # 0, and Z 91, we get for some j < k, sin ajk = 0 with aj ak *- 0, i.e., 8/ (aj - ak) = irm, where m is a nonzero integer. Thus by the proof of Proposition 3.1, t0X is a tangential conjugate point of o which is a contradiction. So we have [X, Z] = 0. Let 2X' be a Cartan subalgebra which contains X, Z. Then there exists an element k E SO(n) such that X = Ad (k)X and Z = Ad (k)Y for some Y, Y X. Then we have rk Exp t0X = Exp to Ad (k)X = Exp t0X = Exp t0Z = Exp to Ad (k)Y = rk Exp t0Y ,
560
TAKASHISAKAI
and consequently we get Exp t0X = Exp t0Y for some Y e &C, Y
X.
q.e.d. Now we shall determine the tangential cut locus of o. By Proposition 2.5, it suffices to consider the tangential cut point t0X of o along a geodesic t >
Exp tX, where X = E aiAi e 2C, E a = 1. Then by Lemma 4.1, t0X is a conjugate point of o or there exists a unit vector Y (X) in &C such that Exp t0X = Exp t0Y. Generally, by a direct calculation, Exp tX = Exp tY holds for some unit vector Y = E ifli if and only if (
2 (ai - pi)t = min
)
for mi E Z
holds. So first, for a given X we shall search for the minimum positive number to such that (*) holds for some unit Y E &C, Y X. 4.2. Lemma. to = Min1G2G,t n/ (2 f2 Jail). Proof. We shall use the vector notation ; a= (a1i . , an), m = (m1i , mn) E Zn - {0}. Then (*) is equivalent to
t=
(4.1)
2,v/
"ImI2
=a-
(4.2)
al
I
2la,m>m. m
So, if we determine mo 0 such that the value of t defined by (4.1) takes the minimum positive value, then 9 is automatically determined by (4.2). Now we put a: = Max1S1s Jail = Max,.,., I J/Jm12. Then we get I I Im12
< la,Im1I + ... + IanIImnI < alm1l + ... + Im,I `a mi + ... + ynn mi -f- ... + Mn,
-
So Max,G Zn_ [o) I I I m 12 = a, and the equality holds only in the following case : Let a = ail I = = I aik 1, then mil = E1 sgn ail, , mi, = Sk sgn ai,, (s1, , Ek = 0 or 1) and other mi's are equal to zoro. Thus we have to = Min,nez-to} n mJ2/(2 [ la, ml) = n/(2fa) with a = Max1<15 Jail.
4.3. Remark.
If a = I ail J
= . . . = aik I, then 9 which is determined by ii
ii
(4.2) with above mi's are given by 9 = (a1f , ±a? , , ±ai5, ±aix, a.). So there exist 2k - 1 Y = /3iAi( X) such that Exp toX = Exp t0Y holds.
Now the tangential cut point t0X of o along t > Exp tX is given by
to : = Min It > 01 tX is the first tangential conjugate point of o along t > Exp tX or there exists a unit Z e t (Z X) such that Exp tX = Exp tZ} = Min It > 01 tX is the first tangential conjugate point of o along t -> Exp tX or there exists a unit Y e W (Y X) such that Exp tX = Exp tY}.
561
MANIFOLD OF LAGRANGEAN SUBSPACES
Since tX is the first tangential conjugate point of o along t - Exp tX if and only if t = Min f
(X = E aiAi) is given by t° (=t°) = Mini ;r/(2 [ ail). Note that the cut point t°X is the first conjugate point if and only if Max;
i.e., there exist some j < k such that Ia; = ak = a and a; + ak = 0 hold. Thus we get 4.4. Proposition.
The tangential cut point t°X of o along a geodesic
t - Exp tX, where X =
aiAi and g(X, X) = 1,
is given
by to =
/(2 [ IaiI). Minl
4.5. Theorem. For X = E aiAi, where E ai' = 1, we put a(X) : =
Max,
have Ux-, Al(n) = Cut locus of o, A°(n) = Interior set of o. Proof. First we shall show that for a unit vector X e '?X with a(X) _ ai, ... = ai,1, dim ((Exp t°(X)X) fl p0) is equal to k. In fact, we may assume
a(X) = I a, I = ... = I ak I > ak+l I > ...
a,, I. Then we have
Exp t°(X)X = exp t°(X)X 2°
X, cos fJa,t°(X)
(p,4)Ip = x,n cos /a"t°(X)
x, sin /a,t°(X)
4=
... xn)eR"
(xl
X" sin /a"t°(X)
= (p, 4)
_
0
xk+l cos (ak+l7r/2a(X))
p
x cos
J
±x,
q
_
±Xk xk+l sin (ak+l r/2a(X)) xn, sin (a,'1r/2a(X))
J
562
TAKASHI SAKAI
From this our assertion is obvious. Since r, (h e SO(n)) leaves ,a° invariant, we get {Exp Ad (SO(n))t°(X)X I X =
a,Ai( a
=
11
with a(X) = l aid = ... = I aik l} C Ak(n) . Similary it is easy to show that Int (o) = {Exp Ad (SO(n)) tX 10 < t < t°(X), X E X, g(X, X) = 1} C A°(n).
But by Propositions 1.4 and 4.4, "
M = U Ak(n) D k=°
U
({Exp Ad (SO(n))t°(X)X}
XEW,IXI=1
U {Exp Ad (SO(n))tX IO < t < t°(X)j)
= Cut locus of o U Int (o) = M. Thus the proof is completed. 4.5. Corollary. Diameter of M = / n
(2,v" -2). Injectivity radius of M
_ ;02v"T) = (Diameter of M)/drank M. Proof. Diameter of M = MaxE.2=1 Mint
Closed geodesics
we put X(m) : _ 5.1. Theorem. For m : = (m1, , m") E Z" - {0}, Ei (mill ml)Ai e X. Then each of the following holds: (i) c(t) : t - Exp tX(m), 0 < t < I ml7r/ f T, is a closed geodesic of length with the initial point o. Its multiplicity is equal to the greatest comI m 17r/ ,v'mon divisor of m1, , m". (ii) Every closed geodesic of M with the initial point o may be expressed in the form t-Exp tAd (h)X(m), where h e SO(n) andX(m) _ Ei (mill mI)Ai, m E Z" - {0}. (iii) The intersection number of a clesed geodesic t -+ Exp t Ad (h)X(m), 0 < t < I mI 7r/ f , with the oriented codimension one cycle Uk'=, Al(n) is given by T, mi. Proof. 1'. c(t) : t -+ Exp t(E aiAi), where E a' = 1 and 0 < t < t1 i is a geodesic loop # exp t1(Z aiAi) E O(n) + sin 2 ait1)aij) E O(n) # v T ait1 = mi, mi e Z t1 =7rmil(v Tai) _7r Imllf , ai = mill m I with m = (m1, ,m")E Z" -{0} ((cos /Tait1
563
MANIFOLD OF LAGRANGEAN SUBSPACES
Next, since exp (7r I m IX(m) /
-2) = ((cos mi7r)li j) e 0(n), we get aiAi) m = d7r(Ad ((cos mi2r)oij))(Z aiAi)
e(7r ml /V) = dzexp(n
= d7r(E aiAi) = c(0) ,
that is, c(t), 0 < t < 7r I m I / ti[J is a closed geodesic. 2°.
Let k be the greatest common divisor of m , mn; i.e., m = kp, , p,t) e Zn - {0} and p , p,, are relatively prime. Then we
p : = (p
get by 1 °, c(7r Ip I / -,/) = c(0), c(7r l p I / f) = c(0), and consequently c(t),
0 < t < 7r I m I / f, is a closed geodesic of multiplicity k. Conversely let c(t) : t -> Exp tX(m), 0 < t < t, = r I m I l v"- -2, m e Zh - {0}, be a closed geodesic of multiplicity k. Then from c(t,/k) = c(0), c(t,/k) = c(0), we get ir ImI/(N'2k) = ir IpI/f , pi/IPI = mi/ImI for some p EZn - {0} with relatively prime p , p,,. That is m = kp and the greatest common divisor
of m
, m,, is equal to k. Thus we have shown (i). (ii) is obvious from 1 ° and the fact that TZ = Ad (SO(n))22t. 3'. Let c(t) : t > Exp t Ad (h)X(m), 0 < t < 7r I m / 2 , be a closed geodesic. To show (iii), it suffices to consider the case h = e. Then the intersection number of c(t) with the oriented cycle Uk=, A11(n) is given by Z,wnpa#{o} sgn ge()nt0 (t) where gacc)
,
is the following symmetric form on a subspace c(t) fl Po of c(t)
Let q,,,)c(t) be the symmetric bilinear form on c(t) defined by
cos ti/ ait\ l\I`xi sin
yi cos ti/ ait\\ (yj sin wait
= (,vl- _2 /I mD E mixiyi sin' (/mit/ mI) Then go(c)np0C(t) is
,
(ai = mill mI)
defined as the restriction of q,,,)c(t) to the subspace
c(t) fl po of c(t), [4]. Now c(t) fl po # {0} if and only if cos ti1_2_ait = 0 for at least one ai. Now put T : = {(mi, r) I 1 < i < n, 1 < r < I mi I, r integer} and on T : (mi, r) - (mj, s) (4 consider the following equivalence relation " 2ow 1)). 21)) =t7,,(: _7rIm1 tir(: =7r ImI/('V ". We denote by [(mi, r)] the equivalence class of (mi, r) with respect to sor some (mi, r) e T, and c(ti,r) Then c(t) fl po # {0} holds if and only if t =
fl po = {(0, q) I q = 1(0, ... , ±xi,, .. , ±xik, ... , 0), where i ... , ik are dennc(ti,r) : , (mik, rk)}. Thus we have termined by [(mi, r)] = {(mi,, r), (ti/ 2 /ImI) Ek= mi;xi;yi,, so that sgn ga(t,,,)nPoC(ti.r) = Ek=, sgn mi., and sgn mi. _ consequently the intersection number is equal to mi, because of #T = Y,7_, I mil. i=, (sgn m) ImiI = 5.2. Corollary. Two closed geodesics t > Exp t Ad (h)X(m), 0 < t <
TAKASHISAKAI
564
Iml r/1 and t --> Exp t Ad (k)X(n), 0 G t G I n I rwhere h, k e SO(n), are homotopieally equivalent if and only if i=1 mi = Zi-1 ni. References
[1] [2] [3] [4]
V. I. Arnold, On a characteristic class entering in quantization condition, Funct. Anal. Appl. 1 (1967) 1-13. R. Crittenden, Minimum and conjugate points in symmetric spaces, Canad. J. Math. 14 (1962) 320-328. J. J. Duistermaat, Fourier integral operators, Lecture Notes, Courant Institute, New York, 1973. , On the Morse index in variational calculus, Advances in Math. 21 (1976) 173-195.
[5]
S. Helgason, Differential geometry and symmetric spaces, Academic Press, New York, 1962.
Added in proof. After the present note had been submitted, the following articles on cut loci of compact symmetric spaces appeared. [61 H. Naitoh, On cut loci and first conjugate loci of the irreducible symmetric Rspaces and the irreducible compact hermitian symmetric spaces, Hokkaido Math. [7]
J. 6 (1977) 230-242. T. Sakai, On cut loci of compact symmetric spaces, Hokkaido Math. J. 6 (1977) 136-161.
, On the structure of cut loci in compact riemannian symmetric spaces, to appear in Math. Ann. [9] , Cut loci of compact symmetric spaces, to appear in Proc. United StatesJapan Sem. Differential Geometry, 1977. [10] M. Takeuchi, On conjugate loci and cut loci of compact symmetric spaces, I, II, preprints, 1977. [8]
HOKKAIDO UNIVERSITY, SAPPORO UNIVERSITY OF BONN
J. DIFFERENTIAL GEOMETRY 12 (1977) 565-566
ON A THEOREM OF AL'BER ON SPACES OF MAPS VAGN LUNDSGAARD HANSEN
In [1, Theorem 7] (see also [2, Theorem 32]) Al'ber proved Theorem 1 (Al'ber). Let V be a compact connected Riemannian manifold, and let M be a compact connected Riemannian manifold with strictly negative sectional curvature. Let CZ(V, M) denote the space of maps of class CZ of V into M equipped with the CZ-topology. Then one of the following holds for any (path-) component K in CI(V, M) : (1)
K has the homotopy type of a point, and contains a unique harmonic
map,
(2) K has the homotopy type of a circle, and all the harmonic maps in K map V with the same value of the Dirichlet integral into the same closed geodesic of M,
(3) K has the homotopy type of M, and each harmonic map in K maps V into a single point of M. Theorem 1 can also be proved by the methods developped by Eells and Sampson [3] and is very close to being stated explicitely in Hartman [6]. The purpose of this note is to point out that the topological aspect of Theorem 1 is a simple consequence of classical knowledge about the fundamental. group of a Riemannian manifold with negative sectional curvature and the following elementary result in homotopy theory. Lemma. Let X be a locally compact connected CW-complex, and Y a space of type (ir, 1). Let C(X, Y) denote the space of continuous maps of X into Y equipped with the compact-open topology. For any based map f : X
-> Y denote by C(X, Y; f) the (path-) component in C(X, Y) containing f, and denote by C(x; f) the centralizer of f,(zri(X)) in jr1(Y). Then C(X, Y; f} is a space of type (C(zr; f), 1). We recall that a connected CW-complex Y is called a space of type (7c, 1), if rc is a group, 7ci(Y) = 0 for i > 2, and jr1(Y) -- ir. We recall also that if A
is a subset of the group G, then the centralizer of A in G is the subgroup CA(G) = {g E G l ag = ga, all a e Al. A proof of the lemma can be found in Gottlieb [4, Lemma 2].
It is a classical result of Hadamard-Cartan that a complete Riemannian manifold M with nonpositive sectional curvature is a space of type (;r1(M), 1).
From the classical results of Preissmann [8] it also follows that the fundaCommunicated by J. Eells, Jr., October 25, 1975.
566
VAGN LUNDSGAARD HANSEN
mental group of a compact manifold M with strictly negative sectional curvature has Property C. A group it is said to have property C if any centralizer in it is either the identity subgroup, or an infinite cyclic group, or rr. The necessary Riemannian geometry to prove the results above concerning the fundamental group of a Riemannian manifold with negative sectional curvature can also be found in Gromoll, Klingenberg and Meyer [5, § 7.21. The following theorem contains the topological aspect of the theorem of Al'ber. Theorem 2. Let X be a locally compact connected CW-complex, and Y a space of type ('r, 1) where it has property C. Then any component in C(X, Y) has the homotopy type of either a point, or a circle, or Y. Theorem 2 follows by observing that the component determined by the based map f : X - Y is a space of type (C(rr; f), 1) and that C(rr; f) is either the identity subgroup, or an infinite cyclic group, or rr,(Y). We should also remark that the space of maps CZ(V, M) has the same homotopy type as C(V, M) by well-known approximation theorems or, for an elegant proof, by Palais [7, Theorem 13. 141. Al'ber's proof of Theorem 1 involves fairly advanced calculus of variations, and it is necessary for his proof that the domain is compact. So apart from a completely topological setting we obtain also in Theorem 2 a slight generalization of the topological part of Theorem 1 in so far that we only need the domain to be locally compact. It would be interesting to have an example of a manifold M, which is a space of type ('r, 1) where rr has property C, but where M does not admit a Riemannian metric with strictly negative sectional curvature. References [1] [2] [3] [4]
S. I. Al'ber, Spaces of mappings into a manifold with negative curvature, Dokl. Akad. N auk SSSR 178 (1968) 13-16; Soviet Math. Dokl. 9 (1968) 6-9. , The topology of functional manifolds and the calculus of variations in the large, Russian Math. Surveys 25 (1970) 51-117. J. Eells & J. H. Sampson, Harmonic mappings of Riemannian manifolds, Amer. J. Math. 86 (1964) 109-160. D. H. Gottlieb, Covering transformations and universal fibrations, Illinois J. Math. 13 (1969) 432-437. D. Gromoll, W. Klingenberg & W. Meyer, Riemannsche Geometrie im Grossen, Lecture Notes in Math. Vol. 55, Springer, Berlin, 1968. P. Hartman, On homotopic harmonic maps, Canad. J. Math. 19 (1967) 673-687.
[5] [6] [7] R. S. Palais, Foundations of global non-linear analysis, Math. Lecture Note Series, [8]
Benjamin, New York, 1968. A. Preissmann, Quelques proprietes globales des espaces de Riemann, Comment. Math. Helv. 15 (1943) 175-216. UNIVERSITY OF COPENHAGEN
J. DIFFERENTIAL GEOMETRY 12 (1977) 567-581
THE MORSE INDEX THEOREM IN THE CASE OF TWO VARIABLE END-POINTS JOHN BOLTON
1.
Introduction
Let W be a C`° complete positive-definite Riemannian manifold, and let P, Q be submanifolds of W. If r: [0, b] -+ W is a geodesic of W intersecting P and Q orthogonally at r(0) and r(b) respectively, then r may be thought of as a "stationary point" of the length function L acting on the space of paths from P to Q. If Q, is the space of continuous piecewise-smooth vector fields along r, which are orthogonal to r and have initial vector tangential to P and final vector tangential to Q, then the Morse index form I : Sl, X Q, -+ R is a symmetric bilinear map which is interpreted as the hessian of L. The index of I is the dimension of a maximal subspace of D, on which I is negative definite, so this is a measure of the number of essentially different directions in which r can be deformed to obtain shorter paths from P to Q lying arbitrarily close to r. If Q is a point, the Morse index theorem says that the index of I is equal to the sum of the orders of the focal points of P along r. (See e.g., [2, Chapter 11].)
In this paper we prove a Morse-type index theorem in the general case by defining the notion of a (P, Q)-focal point of signed order, and then obtaining an expression for the index of I as the sum of an initial term together with the signed orders of the (P, Q)-focal points. This is obtained in Theorem A in § 4. Ambrose [1] and Morse [3] also have extensions to the general case. However the author feels that the present approach has advantages for two reasons. First, the initial term is easily computed because it depends only on the second fundamental forms S, T of P, Q respectively with respect to r'(0), r'(b) respectively. Secondly, the definition of (P, Q)-focal point is very natural and rather easier than, for instance, Ambrose's corresponding notion of a "conjugate point of P and Q".
The method of proof of Theorem A follows [1] and [2] in that an index function i is defind on [0, b] and the discontinuities of i are analysed. Unlike [1] and [2] however the index function in our case is not necessarily nondecreasing. This makes it unlikely that the ad-hoc subdivisions of [0, b] used in Received October 27, 1975.
JOHN BOLTON
568
this paper can be avoided by using methods similar to those employed by Osborn in [4]. The simple nature of the initial term in Theorem A makes it interesting to obtain upperbounds on c E R+ in order that there be no (P, Q)-focal points on [0, c]. This is the motivation behind Theorem B, which is stated and proved in § 5. This theorem is similar to some of the comparison theorems proved by Warner in [5], although the proof is rather different. 2.
The index form
Notation will be as in § 1, with the additional assumptions that r is parameterized by arc length and prolonged so that its domain of definition is R. If B is a C--manifold and if b E B, then Bb will be the tangent space of B at b. For each t E R, let r'(t) be the tangent vector of r at 7(t), and let Wt = {X E Wrct> : <X, r'(t)> = 0} .
For each t E R, R(t) will be the Ricci transformation of Wt into itself given by
R(t)X = R(r'(t), X)r'(t) where R is the curvature tensor of W. Let V be the vector space of parallel vector fields along r, which are orthogonal to r. Then the evaluation map V - Wt which sends X to X(t) is a linear isomorphism which will be used to identify Wt with V. For t > 0, Qr' will be the vector space of continuous piecewise-smooth maps X : [0, t] --+ V with X(O) E P7P) and X(t) E Q,(b), and X will be the derivative of X. Then It : Q; X QY - R will be given by t
It(X, Y) = f
o
- X (ti ), Y(ti)>
+ <X(t) - TX(t), Y(t)> - <X(0) - SX(0), Y(0)> , where the sum is over the jumps ti of X in ]0, t[. It is a symmetric bilinear map and is the Morse index form arising from the variational problem with end conditions S at 0, T at t as described below. Suppose e2 is a submanifold of W intersecting r orthogonally at r(t), and suppose that the second fundamental form of 2 with respect to r'(to) is equal to T (so, in particular 2rcto> = Q711)). Consider a 1-parameter family of curves
rs(0 < s < s) from P to 2 converging to r = ro as s --+ 0. Let X be the associated transverse vector field, i.e., X(t) is tangential to the curve s at s = 0. If L(s) is the length of rs, then L ds s=o
=0
dzL ds2
= Ito(X, X) 's=o
rs(t)
THE MORSE INDEX THEOREM
569
It is in this manner that I has the interpretation of the hessian of L as mentioned in the introduction. If i(t), a(t), n(t) are the index, augmented index and nullity of I', it is well known [1, p. 65] that a(t), n(t) and i(t) are finite, so that a(t) = n(t) + i(t). To prove Theorem A we study the way in which i(t) changes as t goes from 0 to b. In § 3 it is shown that a(t) is upper semi-continuous, and i(t) is lower semi-continuous. Thus a(t) and i(t) are continuous (and hence locally constant)
at all points where n(t) = 0. The jump discontinuities of i(t) and a(t) at a point with n(t) # 0 are evaluated in Propositions 1 and 2, and the proof of Theorem A is completed in § 4, where an expression for i(0+) is obtained. 3.
Jumps of i and a on ]0, b]
We use i(L), a(L), n(L) to denote the index, augmented index and nullity of a symmetric bilinear map L. If X is a scalar or vector valued function (resp. a vector field along y), then X will be its derivative (resp. covariant derivative). The following lemma is the tool used in the analysis of the discontinuities of i(t) and a(t). Lemma 1. Let U be a finite-dimensional vector space, and let SB(U) be the vector space of real-valued symmetric bilinear maps on U. Let K : ]c, d[
SB(U) be continuously differentiable at t, e ]c, d[, and let N be the null space of K(t,). Then 3s > 0 such that vp e ]0, s[ (i) i(K(t, + p)) > i(K(to)) + i(K(to) IN X N), (ii) i(K(t, - p)) > i(K(to)) + i(-K(t0) I N X N). Proof. (i) Equip U with a scalar product, and if Z is a subspace of U let (Z), be the unit sphere of Z. Let C be a subspace of U of dimension i(K(t,)) on which K(to) is negative definite, and let D be a subspace of N of dimension i(K(t,) IN X N) on which K(to) is negative definite. Since K is continuously differentiable at to, as, > 0 and an open neighborhood B of (D), in (D O C), on which K(to + p)(X, X) < 0 yX E B, y0 < p < s,. Now (D O+ C)1\B is compact, so 3s, > 0 such that K(t, + p)(Y, Y) < 0 yY E (D O+ C)1\B, yO < it is clear that K(t, + p)(X, X) < 0, yX E (D O+ C) p < s,. Ifs = min Vp E ]0, sL (ii)
Apply (i) to L, where L(to + p) = K(t, - p) for all suitably small p.
Let to E ]0, b]. Following standard practice we construct a finite dimensional subspace B of QL° such that i(I'° I B X B) = i(t) and a(I'0 I B X B) = a(t,). If X is a smooth vector field along 7 with X = RX then X is called a Jacobi
field. The set / of Jacobi fields which are everywhere orthogonal toy is a vector space. If X E / has X(0) E P,,,, and X(0) - SX(0) I P,(o), then X is called a P-Jacobi field. These arise as the transverse vector fields associated with variations of 7 through geodesics intersecting P orthogonally (see [2, p. 2221).
A finite sequence {ui} with 0 < u, <
< u. < t, is strongly normal in
570
JOHN BOLTON
[0, t0] under the following conditions : (i) Each nontrivial P-Jacobi field has no zeros in 10, u1]. (ii) Each nontrivial Jacobi field X with X(to) E Q7(b) and X(to) - TX(to) L Qr(b) has no zeros in [u,,, t0[. (iii) For i = 2, zero in [uti_1i ui+1]
, n - 1, each nontrivial Jacobi field has at most one
It follows from the Rauch comparison theorem and the extension due to Warner [5, Cor. 4.2] that strongly normal sequences exist, and moreover there are a finite sequence {ui} and e > 0 such that {ui} is strongly normal in [0, t] for all t e [to - e, to + E]. For all such t we set Bt = {X E SQ; : X is smooth with X = RX except possibly at
u" ... , u.; SX(O) - 1(0) 1 Pr(o)' TX(t) - !(I)
I QT(b)}
.
Theorem (For proof see [1, p. 68]). Assume n > 1. For each t e [ta to + e], i(t) = i(P I Bt X Bt) and a(t) = a(P I B` X Bt). Let H = Wu1 +O O+ W. The evaluation map eve : Bt - H given by evt(X) = (X(u1), ..., X(u,,)) is a linear isomorphism, so the map J: ]toE, to + e[ - SB(H) given by J(t)(x, y) = It(evt 1(x), evc 1(y))
is well-defined. Moreover, by the above theorem, i(t) = i(J(t)) and n(t) _ n(J(t)). In the following lemma we do the computation necessary to apply Lemma 1. Lemma. J is smooth, and the derivative J(t) of J at to is given by
J(t)(x, y) =
(1)
ax Y
J(to)(x Y) _ (\ ah at
I
Now a
a
atL(ah as 'Y)
- \ at
ah
-
(M -=-)]
as '
Y)
+ (6h as '
at
571
THE MORSE INDEX THEOREM
ax,
a
aax
ar a
= C a RX Y> ah
ay
a
at
RY) = 0 by symmetry of R . - Ca`Y ah '
Also, (aX/ah)(t,, u.n) = 0, so from (1)
(2)
Ato)(x, Y) = [(--a-
a
at
1' >
+
ay )J
\ ah '
(t0, t0)
.
Writing C(h) _ (aX /at)(h, h) and D(h) = X(h, h), we see that if U E Q, (b) then
.
Differentiating this with respect to h, and then putting Y(h, h) = U we get
a2x + at,
a
ax Y) at
ah
/T r ax + ax 1,
(h, h) _ \\
LL
ah
at J
Y> (h, h) .
This together with (2) gives
(3)
J(to)(x, y) = [
Now if N
then <X, N> I (h, h) = 0 so that
_ [
and this gives the answer needed to prove the lemma. From the definition of It, it is clear that X E S),' is in the null space P if and only if each of the following two conditions holds.
t of
(i) X is a P-Jacobi field. (ii) X(t) - TX(t) 1 Q,(b) If dim 6t 0, then we call t a (P, Q)-focal point of order n(t) = dim /t. Notice that if Q is a point then a (P, Q)-focal point is usually called a focal point of P along y, while if both P and Q are points then a (P, Q)-focal point is just a conjugate point of y(0) along y. If t E 10, b] and X, Y E /t, then
J(t)(evt(X), evt(Y)) =
572
JOHN BOLTON
fine H. Let n+(t) (resp. n_(t)) be the dimension of a maximal subspace of /t on which the symmetric bilinear map (X, Y) H
ing proposition are immediate, while (iii) and (iv) use the fact that a(t) +
i(-J(t)) = dim H. Proposition 1. Let t E ]0, b]. Then 3e > 0 such that Vp E 10, e[
i(t + p) > i(t) + n_(t), (ii) i(t - p) > i(t) + n+(t), (i) (iii)
(iv)
a(t + p) < a(t) - n+(t), a(t - p) < a(t) - n_(t).
It follows that i and a are locally constant at any t with n(t) = 0. We call t a nondegenerate (P, Q)-focal point if n_(t) + n+(t) = n(t) > 0. Clearly, if W has positive sectional curvatures at the (P, Q)-focal point t, then t is nondegenerate, while if Q is a point then all (P, Q)-focal points are nondegenerate. Proposition 2. If t is a nondegenerate (P, Q)-focal point, then the inequalities of Proposition 1 are equalities, and t is an isolated (P, Q)-focal point. Proof. Let e be as in Proposition 1 and let p E ]0, e[. From Proposition 1 we have
a(t) - n+(t) > a(t + p) > i(t + p) > i(t) + n_(t) a(t) - n_(t) > a(t - p) > i(t - p) > i(t) + n+(t)
.
Since a(t) = i(t) + n(t), the hypothesis of Proposition 2 implies that all the above inequalities are equalities. The result now follows. 4.
Calculation of i(0+)
If t is sufficiently small and positive, then i(t) (resp. a(t)) is equal to
i(It I /t x /) (resp. a(It I /t x fit)) where Ift = {X E / : X(O) E Pr(o), X(t) E Qr(a)}
.
For a proof of this see [1, p. 64].
If X, Y E /t, then
(4)
It(X, Y) = <1(t) - TX(t), Y(t)> - <1(0) - SX(O), Y(0)>
.
If P and Q were both hypersurfaces, each /t = / and the right hand side
THE MORSE INDEX THEOREM
573
of (4) would be defined on / for all t e R. This would make it possible to compute i(0+) by using Lemma 1, so we begin this section by considering this case.
Let P, be hypersurfaces of W, which intersect r orthogonally at r(0), r(b) respectively. Let S, T be the second fundamental forms of P, Q with respect to r'(0), r'(b), respectively, and let a (t) (resp. a(t)) be the index (resp. aug mented index) of the corresponding index form It. We compute 'T(01) in terms of S, T and R(0), and later use this to compute i(0-) in the general case. Lemma 3. Let N be the null space of Then 3s > 0 such that Vp E ]0, e[
(i) (ii)
i(p)>i(S-T)+i(L1NX N),
a(p) < i(3 - T) + a(L N x N),
where L E SB(V) is given by
L(X, Y) =
.
Let J : R - SB(/) be given by
J(t)(X, Y) = <x(t) - TX(t), Y(t)> - <X(0) - 3X(0),
Y(O)>
As already remarked, 3e > 0 such that for 0 < t < e, i(J(t)) = i (t) and a(J(t)) = a(t). Clearly
J(0)(X, Y) =
N={XE/:3(X)=T(X)}. We now show that (a) i(J(0) I N X N) = i(L I N X N), (b) a(J(0) I N X N) = a(L J N X N). Let
U={XE/:X(0)=SX(0)}, and
N,={XE/:X(0)=0}. Then / = U O+ N and l : U ---> V given by l(X) = X(0) is a linear isomorphism. Let N, = l-1(N). Then N = N, O+ N2. Also j(0) is positive definite on
N and
J(0)(X, Y) = L(l(X), l(Y))
Further, if X E N Y E N2, then
for X, Y E N, .
574
JOHN BOLTON
J(0)(X, Y) = -
We now return to the general case in which the end manifolds P, Q are not necessarily hypersurfaces. For any subspace B C V, p,: V -p B will be the orthogonal projection onto
B, and B-- will be ker PB. Let U = P,(o) n Q,(b), and write S - T as an abbreviation for the map
Pt,o(SIU-TIU):U- U. We will construct symmetric linear maps S, T on V such that (i)
Prrcu, o S I P,(o) = Sand PQ,,,, o T I Q,(b)
= T,
index (resp. null space) of S - T = index (resp. null space) of S - T, (iii) T I U depends only on S and T. P, Q will then be chosen to have S, T as second fundamental forms with respect to 7'(0), r'(b) respectively. It is clear that (i) implies that i (t) > i(t) and a(t) > a(t) for all t 0. We later show that for sufficiently small positive t, i(t) < i(t). This will yield the desired expression for i(0+) which, by Lemma 3, depends only on S, T and
(ii)
R(O).
Construction of S and T. Let C (resp. D) be the orthogonal complementary subspace of P,(,) (resp. Q,(b)) in P,(0) + Q,(b). Then Pr(o) + Qr(b) = U O C O D, and C ( D is orthogonal to U.
Define S1f T,: V -p V by the requirements that (a) Im T1 C D and Im S1 C C, (b)
PCoD o (S o Prr,o, - T o PQr (b)) = T 1 - S1.
Let S*, T* be the adjoints of S1i T1, and let 2 E R. Put
S = S, + S* + 2pr1a; + S o Pr, m, T =T1+
T1*
- 2PQrl + T ° PQr cb
It is clear that S and T are symmetric and that (i) and (iii) are satisfied. Also, if N is the null space of S - T, then yX E N
(S - T)X = (S, - T1)X + (S* - T*)X + SX - TX = -PC D(SX - TX) + SX - TX + (S* - T*)X = pvSX - pUTX + (S* - T*)X =0, since XENC U. It is a consequence of (i) that i(S - T) > i(S - T), so the following lemma, together with a countup of dimensions, shows that (ii) is also true.
THE MORSE INDEX THEOREM
575
Lemma . Let S - T be positive definite on G C U. Then, for suitably large 2, S - T is positive definite on G 1) UJ-. (We henceforth assume that S/T are defined using such a 2.) Proof. Recall that (B)1 denotes the unit sphere of a normed vector space B. If Z E (G)1, then by (i)
<(S - T)X, X> = <(S - T)X, X> > 0 .
Thus there is an open neighborhood D of (G)1 in (G O UL), such that (S - T)X, X > 0, yX E D. If Y E (G O+ UL)1f then there are Y1 E G, Y2 E C, Y3 E D, Y, E (P,(o) + Q,(b))L such that Y = Y1 + Y2 + Y3 + Y4. Thus \(S - T )Y, Y> = 2(\Y2, Y2> + \Y3, Y3> + 2
where K is a continuous function on (G G UL),. If H = (G G UL)1\D, then H is compact so we may choose s, p E R+ such that E = inf {
p = sup {j K(Y) }
.
YEH
Choose 2 > pls. Then
<(9 -T)X,X>>0,
VXE(CeUL)1,
and the lemma is established. So far, i, n, n+, n_ have been integer-valued functions defined on the positive real numbers. We now extend their domains of definition to the nonnegative reals as follows.
Let i(0) = i(S - T) and n(0) = n(S - T). Let n+(0) (resp. n_(0)) be the dimension of a maximal subspace of N on which pN o (R - TT I N) is positive
(resp. negative) definite, where N is the null space of S - T. If n(0) # 0, then we say that 0 is a (P, Q)-focal point of order n(0), while 0 is a nondegenerate (P, Q)-focal point if n_(0) + n+(0) = n(0) > 0. Notice that these definitions are independent of the choice of 2 used in the definition of S and T. Also, if W has positive sectional curvatures at r(0), then n+(0) = 0 and n_(0) = n(0), while if P,(o) fl Q,cb> = {0}, then i(0) = n(0) = 0. Proposition 3. If n(0) = 0 or if 0 is a nondegenerate (P, Q)-focal point, then 3s > 0 such that there are no (P, Q)-focal points on 10, s[ and Vp E 10, e[,
i(p) = i(0) + n_(0). Proof. Let s > 0 be as in Lemma 3. Then that lemma, together with property (ii) of S and T, shows that Vp E 10, s[
i(0) + n_(0) > a(p) > 1(fc) > i(0) + n_(0) .
576
JOHN BOLTON
Thus the above inequalities are equalities, so there are no (P, Q)-focal points on 10, s[. We already know that a(u) G a(le), so it remains to show that for sufficiently small positive p, i(p) > i(0) + n_(0). Let V1, V2 be subspaces of Pro>, Qrcb> respectively, such that v, f1 V2 = {0} and V, +O V2 = P, (o) + Q,(b) . Define L : U > V as follows. For each X E U, there are unique elements v, E V1, v2 E V2 such that
S,X+T,X=v2-v,. Set
LX=v,+S,X=v2-T,X. It follows from the Rauch comparison theorem for submanifolds [5, p, 3511
that 3s, > 0 such that for all h E ] -s s,[\{O} and all X E U there are unique Jacobi fields Xh, X, along r such that (a) Xh is a P-Jacobi field with X7z(2h) = X + 2hLX, (b) r, (h) E Q,cb>, 7 (h) - T.,Th(h) 1 Qr b>, h(2 h) = X + 2hLX. .
Now define J: ]-s s,[ > SB(U) by
J(h)(X, Y) = <Xh(2h) - h(2h), Yh(2h)> J(0)(X, Y) _ <(S - T)X, Y> .
for h # 0 ,
Notice that for h E ]0, s,[, J(h)(X, Y) = Ih( h, Dh) where Xh 1 [0, 2h] _ Xh [0, 2h], 3h [2h, h] = Xh I [2h, h], and similarly for h. In the following lemma we do the calculation necessary for applying Lemma 1. Lemma. J is smooth, and
J(0)(X, Y) = <(R - TT)X, Y> for X, Y in the null space N of S - T. Proof of Lemma. Let m be the dimension of P, and d the dimension of V. The following ranges of indices will be used:
Gd.
Gm,
We shall also employ the summation convention whereby repeated indices are summed over their respective ranges.
Let X, Y E U, and pick an orthonormal basis e , ed for V such that X = xe, for some x E R, and {e , e j spans Pro>. Let u, and va be the Jacobi fields with u1(0) = ei, ui(0) = Se,, and va(0) = 0, 'i (0) = ea. Since va(0) = 0, the vector fields wo. given by
wa(t) = va(t)/t are smooth, and {u,(t),
.
, uM(t),
for t # 0, wa(0) = 'i (0) = ea wd(t)} are linearly independent
THE MORSE INDEX THEOREM
577
for t E ] -Elf El[. Thus we can uniquely define smooth functions aA R by the requirement that
ai(h)ui(2h) + a,(h)w,(-h) = X + lhLX Clearly a,(0) = 0, so the functions b.: I -Elf El[ , R given by
ba(h) = a,(h)/h
for h * 0, b,(0) = 6a(0)
are smooth. Also
(5)
ai(h)ui(2h) + 2ba(h)v,(2h) = X + 2hLX ,
for h E I -E1, El[
However, by definition of s, we have that yh E ]
(6)
yt E R,
Xh,(t) = ai(h)ui(t) + 2ba(h)va(t)
So, if we define X0(t) = ai(0)ui(t) + 2ba(0)va(t), then X(h, t) = Xh(t) is a smooth vector-valued function of two variables. Differentiating (5) with respect to h at h = 0 we have
(7)
ai(0)ui(0)
2ai(0)ui(0)
2b,(0)v,(0)
2LX
Putting h = 0 in (5) we get that
(8)
a1(0) = x, ai(0) = 0
for i = 2,
-,m
so from (7)
(9)
c
(0)e1 + 19X + b,(0)e, = 2LX
Thus, if K(t) _ <(aX/at)(t, 2t), Y(t, 2t)>, then K is smooth, and from (6)
K(0) =
k(O)
= 2(
However, b,(0)e, is orthogonal to P,(,,, so
Thus from (10) (11)
k(0) = 2(
.
578
JOHN BOLTON
Using similar techniques it can be shown that if
H(h) _ < h(2h), h(2h)>
for h # 0
H(O) =
then H is smooth and (12)
II(0) = 2(
.
Since J(h)(X, Y) = K(h) - H(h) we see that J is smooth and J(0)(X, Y) =
-
J(0)(X, Y) =
so that i(J(p)) < i(p).
However, as already remarked, J(p)(X, Y) Thus
i(0) + n_(0) < i(J(p)) < i(p) < a(p) < a(ft) < i(0) + n_(0) and the proof of the proposition is complete. Propositions 1, 2 and 3 are combined to give the main result of the paper : Theorem A. Let P, Q be submanifolds of W, and let r be a geodesic of W intersecting P and Q orthogonally at r(0) and r(b) respectively. If P, Q have only nondegenerate (P, Q)-focal points on [0, b], then these (P, Q)-focal points are finite in number. Further, the index i(b) and the augmented index a(b) of the index form of this configuration are given by
i(b) = i(S - T) +O
n+(t)
O
where S, T are the second fundamental forms of P, Q with respect to 7'(0), r'(b), respectively.
5. A comparison theorem In view of Theorem A it is desirable to obtain an estimate of the distance
from P to the first (P, Q)-focal point. If S - T is positive definite (e.g., if
579
THE MORSE INDEX THEOREM
fl QT(b) _ {0}), then methods similar to those employed by Warner [5, Proof of Theorem 3.21 may be used to yield such information. However, these methods depend on Jt being positive definite for small t, and as we have seen, this is not the case for general S, T. P1(O
In this section we illustrate a method of finding such estimates using the idea of translates of S as employed by Ambrose in [1]. The principal drawback in our use of this construction is that we must assume that P is a hypersurface of W. Let t, be the first focal point of P along r, and let t E [0, t0[. For each X e V,
there is a unique P-Jacobi field ' such that X (t) = X. Let S,(X) = n"(t). This defines a map S,: V - V which may be shown to be an element of the space SL(V) of symmetric linear maps from V to V (See [1, p. 54]). Notice that if P is not a hypersurf ace, then the above breaks down at t = 0. Lemma 4. The map S : [0, to[ - SL(V) given by S(t) = S, is smooth and satisfies the Riccati equation
S(t) = R(t) - S2(t)
.
Proof. The smoothness of S follows from the theory of solutions of ordinary differential equations. Let ' be a P-Jacobi field. Then (S.')(t) so by differentiating we get (S.
')(t) + (-)(t) =
A.
,
which gives (S.1')(t)
= (R,-)(t) - (SS. ')(t)
.
Hence the lemma is proved. If L e SL(V), let L# e SL(Q,(b,) be given by L#(X) = pQ,,,,LX. Theorem B. Assume that P is a hypersurface of W and that (i) each eigenvalue of S has modulus < A, (ii) each eigenvalue of S# - T has modulus > Q > 0, (iii) for each positive t, each eigenvalue of R(t) has modulus < 0. Then the first (P, Q)-focal point occurs at or after t,, where t, is the smallest positive solution of the equation cot 01"t = 2 10 12(0 + A2 + AD)
t=QA-1(A+ 2)-1 Proof.
if 0 > 0
if 0=0.
For B > 0, let v8 be the smallest positive solution of cot 6"2t = AB-u2 ,
and let v0 = A-'. It follows from the Rauch comparison theorem for submani-
580
JOHN BOLTON
folds [5, Corollary 4.2(a)] that the first focal point of P occurs at or after and this occurs after tl. Thus S is defined on [0, t,]. Let
rm
D={(B,t)ER(>0) X R(>0):t
,
where K = cot-' (-110-112)
Then g is continuous and negative on D, and ag/at = -0 - g2. Lemma. If X, Y are unit vectors in V and if t < v , then (d/dt)<S,X, Y> < I (ag/at)(0, t) I
Proof of Lemma. L E SL(V) let
Let 11
1 be the norm on V associated with <
and for
IIL1=sup {IILZ;I:ZEVand llZll= 1}. Then =I
To establish the lemma it is enough to show that I I St I I < I g(0, t) I for 0 > (9, t < re. Since II S0I I < I g(0, 0) 1 it suffices to show that if 0 > 0, t2 E [0, z, [, Z E V are such that IIZII = 1, 0 < ISt211= I g(6, t2)I, then (d/dt) iIStZlI < ag/atl at (0, t2). However, this is clear because in this case d dt
11 StZ I
I =
at t = t2
at(0,t2).
Returning to the proof of Theorem B, we note that if t2 E [0, t1[ is a (P, Q)focal point, then 3X (E Q,(b) such that IIXII = 1 and Si2X = TX. However, from the lemma it is clear that if Y (E Q,,6), then <SOX - St2X, Y> < 19091 0) - g((9, t2) I < 19091 0) - g((9, t1) I = Q
Since S0 = S, we now have a contradiction of hypothesis (ii). This completes the proof of Theorem B. Similar theorems may be proved using the above methods.
THE MORSE INDEX THEOREM
581
References
W. Ambrose, The index theorem in Rienzannian geometry, Ann. of Math. 73 (1961) 49-86. (21 R. L. Bishop & R. J. Crittenden, Geometry of manifolds, Academic Press, New [1]
[3]
(4] [5]
York, 1964. M. Morse, The calculus of variations in the large, Amer. Math. Soc. Colloq. Publ. Vol. 18, 1934. H. Osborn, The Morse index theorem, Proc. Amer. Math. Soc. 18 (1967) 759-762.
F. W. Warner, Extension of the Rauch comparison theorem to subnzanifolds, Trans. Amer. Math. Soc. 122 (1966) 341-356. UNIVERSITY OF DURHAM, ENGLAND
J. DIFFERENTIAL GEOMETRY 12 (1977) 583-594
e-FOLIATIONS OF CODIMENSION TWO JOSIAH MEYER
Introduction A codimension-q foliation .
of the manifold M is an e-foliation if a fram-
ing of its normal bundle Q can be chosen to be invariant under the linear holonomy of each leaf. These structures occur as one extreme case in a general theory of transverse H-structures for foliations analogous to the theory of G-structures for manifolds. In codimension one, e-foliations are defined by a nonsingular closed 1-form. There is a strong structure theorem for such foliations of compact manifolds due essentially to G. Reeb [7] (also see [1, (5.5)]). L. Conlon [1] has investigated the properties of e-foliations in higher codimension and has proven a partial analogue of Reeb's theorem in codimension two. We view an e-foliation as a foliation with transverse structure modeled by a parallelizable manifold. In this spirit, we define a Lie foliation as a foliation with transverse structure modeled by a Lie group. Evidently, every Lie foliation is an e-foliation. It is easy to see that the two notions coincide in codimension one, but differ in codimension greater than two. The main result of this paper is that every codimension-two e-foliation of a closed manifold is a Lie foliation. This additional structure enables us to answer some questions left open by Conlon and essentially complete the structure theory for e-foliations of codimension two. In § 1 we define e-foliations and Lie foliations as special cases of a general notion of transverse structures for foliations. § 2 is devoted to the proof of our main result and the remainder of the paper consists of remarks on the struc-
ture of e-foliations in codimension two. In particular, an example of a Lie foliation modeled on the affine group of transformations of R1 is constructed,
and a theorem of D. Tischler [8] is used to draw several easy but pleasant corollaries of our main theorem. Unless otherwise specified, all manifolds and mappings considered are assumed to be differentiable of class C. I would like to express my gratitude to my advisor, Lawrence Conlon, for his help in this work and indeed for his own work from which this is derived. Communicated by S. Sternberg, September 30, 1975.
JOSIAH MEYER
584
The unified treatment of transverse structures for foliations sketched in § 1 of this paper was suggested by Professor Conlon. 1.
Transverse structures for foliations
In order to develop a notion of a natural geometric structure for foliations,
it is convenient to reformulate the description of a foliation in terms of a Haefliger cocycle by building into the cocycle the notion of a modeling manifold for the transverse structure of the foliation. Definition. Let NQ be a smooth q-dimensional manifold. An N4-cocycle on the manifold M is a collection of triples {(U., gay)}a gEA such that (i) {U,},,,,, is an open cover of M, (ii) each f. is a submersion of U. onto an open subset of NQ, (iii)
g., is a local diffeomorphism of N4 such that for each x E U fl Up
fa(x) = g,(f,(x)) We say the N4-cocycle {(U., ga,)} represents the foliation F of M if the bundle E of tangents to the foliation is given locally by E Ua = ker f a * C T(U.). In this case, any additional structure supported by the manifold Nq, which is preserved by the local diffeomorphisms g,,, can be interpreted as a transverse structure for . For example, if we can find an N4-cocycle repre-
such that the g,,'s preserve some given Borel measure on NQ, we say that F admits a transverse measure [6]. From this point of view, Conlon's notion of a transverse H-structure for a senting
foliation [1] may be defined as follows. Definition. Let H be a Lie subgroup of GL(q, R). A codimension-q foliation . is said to admit a transverse H-structure if an Nq-cocycle representing °- can be chosen such that the g,,'s are local H-diffeomorphisms of some
H-structure on the manifold N. We study the extreme case where H is the trivial subgroup e. An e-structure for the manifold NQ is a framing of its tangent bundle (or an absolute parallelism on Ne). Definition. An e-foliation is a foliation which admits a transverse e-structure.
In this case, we have a framing of the normal bundle Q C T*(M) of F by sections "parallel along the leaves of " [1] (a section a E r(Q) is parallel along the leaves of F if for each a, co I U = f,* for some 1-form on Ne). Via a choice of Riemannian metric on M, we may view Q as the subbundle of T(M) orthogonal to the bundle of tangents of F, Q = E-L. Then a section X E r(Q) C 36(M) is parallel along the leaves of 9 if fa*X is a well-defined
vector field on UU.) C N4 for each a. A useful equivalent formulation in terms of a Bott-basic connection gives : X E r(Q) is parallel along the leaves of .F if for every Y E T'(E), [X, Y] E T'(E), [1]. The simplest example of an e-foliation is the foliation of the total space
e-FOLIATIONS OF CODIMENSION TWO
585
of a fibration p : M , N onto a parallelizable manifold by fibers. Indeed by [1, (4.4)] these are the only e-foliations of compact manifolds which admit a closed leaf. As a final example of a transverse structure for a foliation, we ask that the transverse structure be modeled by a Lie group. Definition. Let G be a Lie group. A foliation .F is a Lie foliation modeled on the Lie group G if it can be represented by a G-cocycle f (Ua, fa, g,,)} where the gap's are (restrictions of) left translation by elements of G.
Since the underlying manifold of G supports an e-structure given by left invariant vector fields, every Lie foliation is an e-foliation. If (a1i , aq) is an e-structure for .F obtained by pulling back an e-structure for the manifold G given by left invariant 1-forms, we have dwi =
Z
1<j
Cjkw j A (Lk ,
where c;k are constants of structure for G. Conversely, if (a)1, , mq) is a framing of the normal bundle of a foliation and satisfies (*), then ,F is a Lie foliation modeled on G [4, (5.1)]. The equivalent formulation for Q C T(M) is :.F is a Lie foliation modeled on G if and only if a framing (X1, , Xq) of Q can be chosen so that the Q-component of the brackets [Xi, X,] satisfies [X1, X,]Q = x=1 C,Xk where Q. are "dual" constants of structure for G. 2.
The main theorem
The starting point of our analysis is the following. (2.1) Theorem (Conlon [1]). Let . be a codimension-two e-foliation of a closed manifold M. Then the universal covering space of M has the form NI A X R1 where A is the universal covering space of a typical leaf A of F. Furthermore, the lifted foliation of M is the foliation of A X R2 by leaves of the form A X point. We remark that if we assume that . is a codimension-two Lie foliation, the above is a special case of (5.1) of [4]. (2.2) Theorem. Every codimension-two e-foliation ,F of a closed manifold M is a Lie foliation. The proof will be a series of lemmas and observations. We choose a Riemannian metic on M invariant under the natural parallelism along the leaves
of .F (i.e., since F admits a transverse e-structure, it admits a transverse 0(2)-structure) and view Q as a subbundle of T(M). Let (Y1, Y2) be an estructure for . consisting of orthonormal vector fields Yi e T (Q) C X (M), and let (Y1, Y2) be a lifted e-structure for , Yi e T(Q) C (NI). Conlon's analysis continues with the observation that the group of covering transformations 2v1(M) maps leaves of to leaves of , and preserves the
586
JOSIAH MEYER
induced e-structure (Y YZ) on M. In particular, this defines a representation p : 7r,(M) -+ Diff+(R2).
The crucial observation for the present results is that the transverse e-structure (Y Y2) induces a framing of the tangent bundle of R2 (i.e., an absolute parallelism or e-structure for the manifold R2) by the complete vector fields Xi, i = 1, 2, given by projecting the corresponding Yi along the leaves of Thus for cp E 7r,(M), p(w) E Diff+ (R2) is an automorphism of the absolute parallelism (X1, X2) on R2, i.e., we have a representation p : 7r,(M) -+ Aut (X X2)
< Diff+ (R2). The advantage of this is that the group Aut (X X2) is particularly amenable to study. Indeed we have the following : (2.3) Theorem (Kobayashi [5]). Let G be the group of automorphisms of an absolute parallelism on the manifold M. Then G is a Lie group. Furthermore G acts freely on the manifold M, and the orbits of this action are (regular) closed submanifolds of M. In particular, dim G < dim M. Following Conlon, we observe that the kernel of the representation p may be identified with the fundamental group of a typical leaf A E 9'. We designate the group ker (p) by 7r,(A), and notice that the quotient group 7r,(M) /7r,(A) may be identified with the relative homotopy set 7r,(M, A).
Let G be the group of automorphisms of the absolute parallelism (X X2) on R2 induced by the e-structure (Y Y2) for the foliation 97. Then dim G < 2 and 7r,(M, A) ~ p(7r,(M)) < G. To prove our theorem, we must show that an e-structure (Y;, Y2) for 9 can be chosen to satisfy [Yi, Y2]Q = C,Yi + C2Y2. We divide the argument into three cases corresponding to the possible dimensions of the group G associated to our original choice of e-structure. (2.4) Lemma. If dim G = 2, then f-- is a Lie foliation modeled on the Lie group G. Proof. Since G acts freely on R2 and the orbits of this action are closed in R2, the mapping of G into R2 which takes g E G to w9(°)(° E R2) is a diffeomorphism of the underlying manifold of G onto R2. Let (Z Z2) be a framing of the tangent bundle of R2 by vector fields invariant under the action of G. Then [Z Z2] = C,Z, + C2Z2 where C,, are constants of structure for G, and
since p(7r,(M)) < G it follows that (Z Z2) is associated to an e-structure (Yi, Y2) for the foliation such that [Yi, Y2]Q = C,Yi + C2Y2 (2.5) Lemma. If dim G = 0, then there is a smooth fibration s : M , T2, and -9-- is the foliation of M by the fibers of s. In particular, .F is a Lie foliation modeled on the Lie group R2.
Proof. A foliation given by the fibers of a fibration onto a Lie group is always a Lie foliation modeled on the universal covering group. Since T2 is the only compact parallelizable 2-manifold, by (4.4) of [1] it suffices to show has a closed leaf. Since the orbits of G and hence of p(7r,(M)) are closed regular 0-dimensional submanifolds of R2, the union of all leaves of 5t covering a particular leaf A
e-FOLIATIONS OF CODIMENSION TWO
587
of F is a closed regular submanifold of M. Hence the leaf A is proper, and A is closed by [1, (4.2)]. q.e.d. To complete the proof of (2.2) we may now assume that dim G = 1. For this case, the proof will be accomplished by a series of lemmas. Let Go be the connected component of the identity in G, and let X E E(R2) be the complete vector field that generates the action of Go on R2. Let ( , be the Riemannian metric on R2 for which (X1, X) is an orthonormal frame, and let X-' E X (R2) be the unit normal to X with respect to ( > such that (Xp , XP) determines the same orientation as (X,P, X2p).
(2.6) Lemma. X1 is invariant under the action of G on R2. Proof. Since Go is a 1-dimensional normal subgroup of G and X is a basis of the Lie algebra of Go, for g E G we have ad (g)X = f g X for some 0 # f g E R. It follows that the diffeomorphism cpg satisfies cpg*pX, = Since cpg preserves the fields X, and X2, it preserves the orientation of R2 determined by (X X). Since for P E R2 at least one of (XP, XP) or (XP, X2P) is a framing of T,(R2), it follows that fg > 0 for all g E G. Hence (cpg*PXp, Xg,g«,) determines the same orientation as (X,Pg The lemma follows from the definition of X1 and the observation that cog preserves the metric ( , >. (2.7) Corollary. X1 is complete as a vector field on R2. Proof. By definition, X1 is a nowhere zero section of the normal bundle
of the codimension-one foliation F, of R2 given by the orbits of the action of Go. Since X1 is invariant under the action of G, it is invariant under the action of Go, and hence is parallel along the leaves of F,. Since p(r,(M)) is a subgroup of G, it acts on R2 as a group of diffeomorphisms preserving the foliation F, and the field X1. Hence the codimensionone foliation p-'(.Fo) of M (where p is the projection p : M - R2) and the field p-1(X1) (i.e., p-1(X1) E T(0) and p,K(p-1(X--)) = X1) are invariant under the action of rt,(M) on M. This defines a codimension-one e-foliation Y(= r(p-1(.T p))) of M which is integral to the foliation F (i.e., the leaves of .' are tangent to the leaves of 9) and whose e-structure (r*(p-1(X1))) is carried to the field X- via lifting to M and projection along the leaves of In particular, X1 is a complete vector field on R2. (2.8) Corollary. There is a system of coordinates (x, y) on R2 such that
a/ax=Xl and a/ay=X. Proof. X1 and X are complete everywhere linearly independent vector fields. By (2.6), [X, X1] = 0, hence they are coordinate vector fields. q.e.d. Notice that with respect to this coordinate system, the action of Go on R2 is given by translation in the y-direction. (2.9) Lemma. G/Go - Z. Proof. GIG, acts freely on R1(= Go/R2) as a group of automorphisms of the Lie group structure on R1 given by projecting the vector field X1 along
588
JOSIAH MEYER
the leaves of Hence G/Go is realized as a subgroup of the Lie group R. Since the orbits of G are closed in R2, it follows that the orbits of G/Go
are closed in R' and hence that G/Go - Zr, r = 0 or 1. Since M is compact and p : 111 > R2 induces a continuous map p : M = 7c1(M)\M > p(r1(M))\R2, we conclude that p(2r1(M))\R2 is compact. The inclusion of p(r1(M)) into G induces a continuous map p(2r1(M))\R2 > G\R2. Hence G\R2 is compact.
Since G\R2 - (G/Go)\(Go\R2) - (G/Go)\R', G/Go is nontrivial. Hence G/Go = Z. q.e.d. By replacing X1 by some constant multiple cX-- we can assume that the action of a generator g E G/Go on R1 = Go\R2 is given by g(x) = x + 1. (2.10) Lemma. G is a semi-direct product of R by Z(G = R X,Z). More precisely, G - {(t, n) I t E R, n E Z} with group operation given by (t1i n1) a (t2i n2)
(t1 + a7 t2i n1 + n2) where a # 0 is a real number. Proof. The lemma follows easily from the exact sequence of Lie groups :
1) G j
) 0. ) R(- Go) ) Z(- GIG,) Let g E G/Go be a generator and choose g E G such that j(g) = g. Then 2: Z > G given by A(n) = gn splits the exact sequence and is a Lie group homomorphism. Since R = Go is a normal subgroup of G, we can define a
0
homomorphism 9: Z > Aut (R) by cp(n) : t,-,, gntg-n. By identifying Aut (R) with the multiplicative group of real numbers, we find some a # 0 such that cp(l) is multiplication by a. It follows that ,r: R X, Z > G defined by i1(t, n) = tgn is a Lie group isomorphism. (2.11) Lemma.
For (t, n) E G - R X, Z and (x, y) the coordinate system of (2.8), cp,,,,,(x, y) _ (x + n, any + t + cn) where c. is a constant depending on n and the above choice of g. Proof. We have already noted that coct,o,(x, y) = (x, y + t). Let c(x) be the function defined by cp(o,t,(x, 0) _ (g(x), c(x)) _ (x + 1, c(x)). Since a/ax = X1 is invariant under the action of G, it follows that c(x) is the constant function c1. Then S°ro,u(x Y) = W(O,ucPcv,o (x 0) = (P(ay,o)S'ro,u(x, 0) = (x + 1, ay + c1) (P(0,70(x, y) = oP(O,n-1)(x + 1, ay + c1)
(x + n, any +
an-1c1
+ an-2c1 + ... + c1)
(x+n,ally +c.) q.e.d. (a/ay x,) = an(a/ay)(p(t,n)((x, y)). Recall We can now compute cP x that from (2.6) we had written cpg*pXp = fgX,g(p) where fg > 0, hence a > 0 and we can define the field ax(a/ay). Let Z1 = X1 = a/ax and Z2 = ax(a/ay). Then Z1 and Z2 are everywhere The formula follows from writing cp t,n, = c°ct,o, °co(o,n)
e-FOLIATIONS OF CODIMENSION TWO
589
linearly independent complete vector fields on RI, which are invariant under the action of G. Since p(;r,(M)) < G, the absolute parallelism (Z1, Z) on RI is associated to an e-structure (Yi, Y2) for 97. Finally since [Z Zz] = log (a)Z1f it follows that [Yi, Y2]Q = log (a)Yi. This completes the proof of (2.2). (2.12) Corollary. Let -F be a codimension-two e-foliation of a closed manifold M. Then is defined by a 2-form w = w1 A wz such that either
do,,=0=do)zor do),=w1Aw2and do)z=0. Proof. There are only two simply connected 2-dimensional Lie groups, Rz and the 2-dimensional affine group I. 3.
An example of a Lie foliation modeled on the affine group I
In [1], Conlon had remarked that a codimension-two e-foliation which admits a closed leaf is a Lie foliation modeled on the Lie group Rz (in Conlon's terminology, a foliation with a "strong transverse e-structure"). In addition, assuming 7r1(M, A) was abelian, he showed that a codimension-two e-foliation of a closed manifold M always admits a C° strong transverse e-structure (i.e., there exists an Rz-cocycle of class C° representing the foliation such that the gaa's are restrictions of translations in RI). Notice that for a codimension-two Lie foliation of a closed manifold M, modeled on the Lie group G, we have realized rc,(M, A) as a subgroup of G such that the space 7r1(M, A)\G is compact. Since G is either Rz or the affine group I, and since s?X does not contain an abelian subgroup which compactifies it, we have the following corollary of (2.2). (3.1) Corollary. A codimension-two e-foliation of a closed manifold M is a Lie foliation modeled in the Lie group Rz if and only if 7,(M, A) is abelian. The loss of differentiability in Conlon's approach resulted from applications of a theorem of Sacksteder where one must introduce a possibly new differentiable structure on the manifold. Conlon also conjectured that every codimension-two e-foliation of a closed manifold M admits a strong transverse e-structure and noted that a codimension-two e-foliation for which the group rc,(M, A) was not abelian would give a counterexample to this conjecture. By our present results, such a foliation would be a Lie foliation modeled on the Lie group W. We have the following example. Let co : T2 -> TI be the diffeomorphism induced by the linear mapping Rz -> Rz given by multiplication by the matrix (1
2), and let M3 be the
manifold T2 x [0, 1]/(p, 0) - (cp(p), 1). Then R3 = 1(13 and rc1(M3) is the group of diffeomorphisms generated by x
x+1
z
y z
y=
x
zy z
x
y+ 1 z
590
JOSIAH MEYER
and
x+y
x
Pa y = x+2y z-1
z
Let X, Y, 2 E 36(R3) be given by Z\
X-\3+2
/
2
1-5)-z(aIx
\3
2
-1
+(1
+
ax
ay
AIT)ayI 2
Then the fields f, Y and 2 are everywhere linearly independent and are invariant under the action of ;r,(M). Hence the fields X = 7r*X, Y = T*Y and Z = 7r*2 E X(M), (where it : R3 -p M3 is the projection), are well defined. Let 9 be the codimension-two foliation of M3 given by integral curves to
the vector field X. We compute [X, Y] _ [7r*X, '*Y] = 7r*[X, f] = 0 and [X, Z] = log (3
+2 5 )X ,
and conclude that (Y, Z) is an e-structure for 97. Furthermore since [Y, Z] = log( 3
2)Y,
,F is a Lie foliation modeled on 2f. For the coordinates
2-1 Y)
x'=( 2-1x+y), y'=(x+
z'
z
on R3(= M3), ." is the foliation of R3 by lines parallel to the x'-axis (3M Riz.) x R2(,,,,Z') as in (2.1)), and the induced action of 7,,(M') on R',.,,., is given by PQp1)
'/)
= (Y'
=
1
y
+
AT5
2
P(az) (z') Z/
and
e-FOLIATIONS OF CODIMENSION TWO
P(1p3)
\z /
=
2
z'
591
y
-1
That is,ic 1(M) = ic1(M, A) = &'(M)) = Z2 X* Z' where *(1) is multiplication
by ((3 - /-5)/2). Since Y E f'(Q) is parallel along the leaves of F, the bundle E O+ span (Y)
is the tangent bundle of a codimension-one foliation, of M. It is interesting to note that in the example the codimension-one foliation , is an e-foliation of M3 (it is the foliation of M3 by the fibers of a bundle T' - > M3 -> S1). However the codimension-one foliation F, admits a transverse H-structure where H is a discrete subgroup of GL(1), but does not admit a transverse e-structure. Indeed E. Fedida [3] points out that this foliation has every leaf dense in M3, but not all leaves are diffeomorphic-some leaves are diffeomorphic to R' and some to S1 X R1. 4.
A structure theorem for codimension-two e-foliations
Let 9 be a codimension-two e-foliation of the closed manifold M. Then of ft is the foliation of M by the fibers of a fibration p : M > G (G = RI or VI) . To study a leaf A E 3 , we study the orbit of a leaf A e I covering A under the action of 'r1(M). This is just the inverse image under p of the orbit of a point g e G under left translation by elements of the subgroup ir1(M, A) < G. In particular, the closure A in M of a leaf A E F corresponds to an orbit of the closed subgroup ir1(M, A) < G. the lifted foliation
If the Lie group ir1(M, A) is 2-dimensional, then 7r1(M, A) = G and the orbits of 7r1(M, A) are dense in G, and it follows that the leaves of 3 are dense in M. If ir1(M, A) is 0-dimensional, then 7r1(M, A) = 7r1(M, A) and it follows as in (2.5) that every leaf of F is closed, and 3 is the foliation of M by fibers of a smooth fibration s : M > T2. In this case ir1(M, A) = Z2. If ir1(M, A) is 1-dimensional, then the arguments of (2.10) show that ir1(M, A) =_ R X, Z. Furthermore, each leaf L of the codimension-one e-foliation _7 of M in the proof of (2.7) is closed in M and is itself e-foliated (in codimension one) by the leaves of F. Then by Reeb's theorem Y is the foliation of M by fibers of a smooth fibration s : M > S1 and the group ir1(M, L)
Z1. Each leaf A e F is dense in a leaf L e 9 and 7r1(L, A) = Z1, k > 2. It follows that ir1(M, A) = Z' Xc Z1. We have the following theorem. (4.1) Theorem. Let F be a codimension-two e-foliation of a closed manifold M. Then one of the following holds : (i) every leaf is closed and is the fiber of a smooth fibration s : M > T' and the group ir1(M, A) = ZZ,
592
JOSIAH MEYER
(ii) the closure of every leaf is a fiber of a smooth fibration s : M -> S' and the group 7r,(M, A) = Zk Xs, Z', k > 2, (iii) every leaf is dense. Remark. If F is a Lie foliation modeled on R2, then 7r,(M, A) = Zk, k > 2. If F is a Lie foliation modeled on W, then we can choose an e-structure (Y Y2) E T(Q) C X(M) such that [Y Y2]Q = Y,. In particular ,, is a codimension-one e-folation of M, and for L E F Y, we have 7r,(M, A) _ 7r,(L, A) Xy, 7r,(M, L). For case (iii) of (4.1), we have ;r,(M, A) = 7r,(L, A) Xs,
Zk, k > 2 where 7r,(L, A) is a torsion free abelian group. We do not know if 7r,(L, A) is necessarily finitely generated, or indeed if there exists a Lie foliation modeled on ?C of a closed manifold with every leaf dense. We also remark that Conlon had proven a C° version of (4.1) under the as-
sumption that ;r,(M, A) was abelian, [1, (4.5)], and that (4.1) is a special case of a theorem of Fedida on the structure of Lie foliations [3]. Added in proof. K. M. de Cesare [On transversely parallelizable, codimension-two foliations, preprint] has announced that a Lie foliation modelled on the Lie group of a closed manifold cannot admit a dense leaf and has stated a refinement of Theorem (4.1). We also remark that P. Molino [Etude des feuilletages transversalement complets et applications, Ann. Sci. Ecole Norm. Sup., to appear] has stated a similar structure theorem for codimension-two e-foliations as a particular case of a structure theorem for transversally complete foliations. 5.
Tischler's theorem and some corollaries
In this section, we use (2.12) to draw some immediate but pleasant corollaries of the celebrated theorem of D. Tischler. (5.1) Theorem (Tischler [8]). Let M be a closed manifold. Suppose M admits m linearly independent nonvanishing closed 1-forms. Then M is a fiber bundle over Tn. (5.2) Corollary. If M supports a codimension-two e-foliation, then M is a fiber bundle over S'. In particular dimR H'(M, R) > 1. (5.3) Corollary. M supports a Lie foliation modeled on R2 if and only if M is a fiber bundle over T2. In particular dimR H'(M, R) > 2. Tischler also shows that a codimension-one e-foliation of a compact manifold M with dimR H'(M, R) = 1 must have every leaf closed [8, Theorem 2]. Since a codimension-two e-foliation is always tangent to a codimension-one e-foliation, we have (5.4) Corollary. If M supports a codimension-two e-foliatiotion and dimR H'(M, R) = 1, then F is a Lie foliation modeled on W, and the closure of every leaf of F is a fiber of a smooth fibration s : M -> S'. Notice also in this case, the fiber F of the above fibration is itself e-foliated
e-FOLIATIONS OF CODIMENSION TWO
593
in codimension one, hence by (5.1) there is a fibration p : F - S' but the manifold M does not fiber over T2. 6.
Codimension-two e-foliations of 3-manifolds
a codimensionLet M3 be a closed oriented 3-dimensional manifold, and two e-foliation of M3. is closed, then M3 is a T2-bundle over (6.1) Proposition. If no leaf of S', and r,(M3) is the semi-direct product of Z2 by Z'. is closed, every leaf is diffeomorphic to R' and Proof. Since no leaf of
r (M3) = r,(M3, A). In particular r,(M3) is a subgroup of either R2 or I. By (5.1) there is a fibration p: M3 > S'. The fiber F is a closed oriented surface hence is determined by its genus. From the exact homotopy sequence
for fibrations we have 0 > r,(F) > r,(M3) > r,(S') > 0. Hence r,(F) is isomorphic to a subgroup of either R2 or a. In particular the commutator subgroup of r,(F) is abelian. From the classification of the fundamental groups of surfaces, it follows that F is a surface of genus < 1. Since All = R3, F S2. Hence F = T2 and we have the exact sequence
0 > Z2(= r,(T2)) > r,(M3) > Z(= r,(S')) > 0 . q.e.d. The following is a special case of [2, Corollary 4]. is closed, then M3 (6.2) Corollary. If r,(M3) is abelian and no leaf of = T3. is a Lie foliation modeled on R2, by (5.3) we have a fiProof. Since bration S' -> M3 > T2. By a standard spectral sequence argument we conclude that this bundle is nontrivial if and only if rank H'(M3, Z) = 2. Since by (6.1), r,(M3) = Z3 = H'(M3, Z) it follows that M3 = T3. is closed, then (6.3) Corollary. If r,(M3) is nonabelian and no leaf of rank H'(M3, Z) = 1.
0 and r,(M3) < ill. r,(M3) Z Proof. We have 0 > Z2 Write a = {(s, t) E R' j (s t) o (s2i t) = (s, + atls2i t, + t), 1 # a > 0}. The commutator subgroup C of r,(M3) is nontrivial and consists of elements of the form (s, 0). Since C < i(Z2) and i(Z') is abelian, it follows that i(Z2) is generated by elements (s 0), (s 0) for s,, s2 rationally independent real numbers.
Let (s, t) E r,(M3) be such that AS, t) is a generator of Z, (t # 0). Then r,(M3) is generated by (s 0), (s2, 0), (s, t). Computing the commutators of the generator we get (s,(1 - at), 0) and (s2(1 - a'), 0) are rationally independent
elements of the commutator subgroup of r,(M3) and it follows that rank H'(M3, Z) < 1. q.e.d. is the foliation of M3 by admits a closed leaf, then by [1, (4.4)] If fibers of a fibration S' > M3 - T2. In summary we have (6.4) Proposition. Let M3 be an oriented 3-manifold, and 9-- a codimen-
594
JOSIAH MEYER
sion-two e-foliation of M3. If $ = dimR H'(M3, R), then 1 < 81 < 3 and each of the following holds. (i) $ = 1 if and only if is a Lie foliation modeled on a (5.4). In this case the closure of every leaf of is a fiber of a smooth fibration T2 -> M3 > S1. (ii) Nl = 2 if and only if 9°o is the foliation of M3 by fibers of a nontrivial
boundle S' > M3 > T2. (iii) pl = 3 if and only if M3 = T3. Remark. In case (i) above, we can choose an e-structure (Y1, Y2) for 9 satisfying [Y1, Y2]Q = Y1. Then S is the foliation of M3 by fibers of a fibration TZ > M3 -> S', and the foliation 3Yz has every leaf dense in M3. A theorem of J. Plante's [6] asserts that for M3 as above, if p1(M3) < 1 and the group ir1(M3) has non-exponential growth, then every transversely
oriented foliation of M3 has a compact leaf. In our case, since F,, has no closed leaf, it follows that ir1(M3) must have exponential growth.
Plante also remarks that construction of § 3 of this paper with the matrix (1 1
1) 2
replaced by an integer matrix of determinant ±1 whose eigenvalues
are on the unit circle but different from one yields a 3-manifold satisfying the hypotheses of his theorem. In particular, these are T2 bundles over S` which do not admit codimension-two e-foliations. References [1]
L. Conlon, Transversally parallelizable foliations of codimension two, Trans. Amer. Math. Soc. 194 (1974) 79-102.
[2]
, Foliations and locally free transformation groups of codimension two, Mich.
Math. J. 21 (1974) 87-96. [ 3 ] E. Fedida, Feuilletages du plan-feuilletages de Lie, Thesis, Universite Louis Pasteur-Strasbourg, 1973.
R. Hermann, On the differential geometry of foliations, Ann. of Math. (3) 72 (1960) 445-457. [ 5 ] S. Kobayashi, Theory of connections, Ann. Mat. Pura Appl. 43 (1957) 119-194. [6] J. F. Plante, Foliations with measure preserving holonomy, Ann. of Math. 102 (1975) 327-361.
[41
[7] G. Reeb, Sur certaines proprietes topologiques des varietes feuilletees, Actualites
Sci. Indust. No. 1183; Publ. Inst. Math. Univ. Strasbourg No. 11, Hermann, Paris, 952, 91-154. [8] D. Tischler, On fibering certain foliated manifolds over S1, Topology 9 (1970) 153-154. UNIVERSITY OF IOWA ELMIRA COLLEGE
J. DIFFERENTIAL GEOMETRY 12 (1977) 595-598
COMPACT REAL HYPERSURFACES OF A COMPLEX PROJECTIVE SPACE MASAFUMI OKUMURA
Introduction Let M be an n-dimensional real hypersurface of a complex projective space CP(n+1)12 of complex dimension (n + 1)/2, and H the Weingarten map of the immersion i : M -> It is known [1] that if a compact minimal hypersurface M of CP(n+1)'2 satisfies trace H2 G n - 1, then trace H2 = n - 1, and up to isometrics of CP(n+I)/2 M is a certain distinguished minimal hypersurface Mp,q for some p and q. The purpose of the present paper is to generalize the above result in such a way that we have an integral inequality which is still valid even if the imCP'n+11/2.
mersion i is not necessarily minimal. Two main tools for this purpose are Lemma 1.1, to be stated in § 1, and the following integral formula established by Yano [3], [41: (0.1)
f
{Ric (X, X) + 2 I L(X)g 2 - PX 2
(div X)'}*1 = 0
,
J zl1
where X is an arbitrary tangent vector field on M, * 1 is the volume element of
M, and I YI denotes the length with respect to the Riemannian metric of a vector field Y on M. In § 1 we explain the model space Ml,q, and in § 2 we present some formulas to be used in § 3. Finally in § 3 we prove our main result. 1.
Submersion, immersion and the model Mp,q
Let S11+2 be an odd-dimensional sphere of radius 1 in a Euclidean (n + 3)the Riemannian submersion with totally geodesic fibres, which is defined by the Hopf fibration Sn+2 > Cp(n+I112. The almost complex structure J of CP(n+1V2 is nothing but
space En", CP(a+'i2 the complex projective space, and
the fundamental tensor of the submersion ;r, and the Riemannian metric G With respect to (J, G), of CP(n+I)'2 is induced naturally from that of S11+2.
CP'n+1112 is a Kaehlerian manifold of constant holomorphic sectional curvature 4. The curvature tensor R of CP(n+1) 2 is given by Received October 31, 1975.
596
MASAFUMI OKUMURA
R(X, Y)Z = G(Y, Z)X - G(X, Z)Y + G(JY, Z)JX - G(JX, Z)JY - 2G(JX, Y)JZ , where X, Y and Z are tangent vector fields on CP(11 +1)/2 For a real hypersurf ace M of CP111+1'v2 and the circle bundle M over M we can construct a Riemannian submersion 7r compatible with the Hopf fibration 7r in such a way that M is a hypersurf ace of Sn12 and that for 7r: M - M, the following diagram commutes :
M
t
> Sn+2
7r I
> CP(n+1)/2 .
In this case i is an isometry on the fibres. We take the family of generalized Clifford surfaces Mi,S = ST X SS, where r + s = n + 1. Regarding E"+3 as a complex I (n + 3)-space, we choose the spheres to lie in complex subspaces. Then we get fibrations S' > M2p+1,2q+i - MM,q compatible with the Hopf fibration, where 2(p + q) = n - 1. MM,,, thus obtained are remarkable classes of real hypersurfaces of CP(n+l'i2 Remark. In [1], Mr,S always means ST X SS which is immersed in Sn+2 minimally. But in this paper we do not assume that M,.,S is minimal. A fundamental relation between M and M is the following [2].
Lemma. 1.1. In order that the Weingarten map H of M is covariant constant, it is necessary and sufficient that the Weingarten map H of M commutes with the fundamental tensor F of 7r. From this lemma we know that if the Weingarten map H commutes with the fundamental tensor F of 7r, M must be M,.,S and consequently M must be MPC, for some p, q. 2.
Local formulas for a real hypersurface
Let X be a vector field over a real hypersurface M of CP(n+1)"2 and N the
unit normal local field to M. Then the transforms JX and JN of X and N respectively by the almost complex structure J of CP(n+1)'2 can be expressed by
(2.1)
JX = FX + u(X)N,
JN = _U ,
where F is the fundamental tensor of the submersion 7r: M -> M, [2]. F, u and U thus obtained define, respectively, antisymmetric linear transformation
of the tangent bundle T(M), a 1-form and a vector field on M. In terms of the induced Riemannian metric g we have
COMPACT REAL HYPERSURFACES
(2.2)
597
g(U, X) = u(X)
Iterating J to X and N we can easily see that (2.3)
FZX = -X + g(U, X) U
(2.4)
FU=O,
(2.5)
g(U, U) = I
,
.
The second fundamental form h and the corresponding Weingarten map H of T(M) are defined and related to covariant differentiation P and 17 in M and M respectively by the following formulas : (2.6)
V xY = FxY + h(X, Y),
(2.7)
IxN = -HX ,
and h(X, Y) = g(HX, Y)N = g(X, HY)N. Since the Riemannian connection of CP(n+"VZ leaves the almost complex structure J invariant, (2.1), (2.6) and (2.7) imply that (2.8) (2.9)
(V F)Z = g(U, Z)HY - g(HY, Z)U , VyU = FHY.
Lemma 2.1. In order that the Weingarten map H of M commutes with the fundamental tensor F of r., it is necessary and sufficient that the vector field U is an infinitesimal isometry.
Proof. We compute the Lie derivative L(U)g of the Riemannian metric g with respect to U, and obtain (L(U)g)(X, Y) = g(17xU, Y) + g(VyU, X)
= g(FHX, Y) + g(FHY, X) = g((FH - HF)X, Y)
,
because of the fact that H is symmetric and F is antisymmetric. Thus we
have proved Lemma 2.1. Let R and Ric be respectively the curvature tensor and the Ricci tensor of M. Then from (1.1) we have
R(X, Y)Z = g(Y, Z)X - g(X, Z)Y + g(FY, Z)FX - g(FX, Z)FY - 2g(FX, Y)FZ + g(HY, Z)HX - g(HX, Z)HY, Ric (X, Y) = (n + 2)g(X, Y) - 3g(U, X)g(U, Y) (2.11) + (trace H)g(HX, Y) - g(HZX, Y) . (2.10)
3.
A generalization of Lawson's theorem
Here we prove a theorem which is a generalization of Lawson's theorem
MASAFUMI OKUMURA
598
stated in the beginning of the introduction. First we apply (0.1) to the vector field U. Since F is antisymmetric and H is symmetric, (2.9) implies that div U = trace FHr= 0 and consequently (0.1) becomes (3.1)
J
{Ric (U, U) - IFUI2}*1 = -2 f atIL(U)g12*1 < 0
,
where equality holds if and only if U is an infinitesimal isometry. On the other
hand, (2.3) (2.5) (2.9) and (2.11) imply that (3.2) (3.3)
Ric (U, U) = n - 1 + (trace H)g(HU, U) - g(H2U, U) 1 FU12 = trace FHt(FH) = -trace F2H2 = trace H2 - g(H2U, U)
Substituting (3.2) and (3.3) into (3.1), we have (3.4)
fm in - 1 + (trace H)g(HU, U) - trace H2}* 1 < 0 ,
where equality holds if and only if U is an infinitesimal isometry. Thus combinning Lemma 1.1 with Lemma 2.1 gives Theorem. Let M be a compact orientable real hypersurface of CP"'+1)'2 over which the following inequality (3.5)
in - 1 + (trace H)g(HU, U) - trace H2}* 1 > 0
holds. Then, up to isometries of CPll+1'12, M is Mp,q for some p and q. Corollary 1. Let M be a compact orientable real hypersurface of If the Weingarten map H of M satisfies (3.6)
CP("+1)'2.
trace H2 < n - 1 + (trace H)g(HU, U)
then, up to isometries of CP«}1'i2, M is Mc.,q for some p and q. Corollary 2, [1]. Let M be a compact orientable minimal hypersurface of CPl,,+11'2 over which trace H2 < n - 1 holds. Then, up to isometries of CPI" +1)/2' M is Mp,q for some p, q. Bibliography [1] [2] [3] [4]
H. B. Lawson, Jr., Rigidity treorems in rank-1 symmetric spaces, J. Differential Geometry 4 (1970) 349-357. M. Okumura, On some real hypersurfaces of a complex projective space, Trans. Amer. Math. Soc. 212 (1975) 355-364. K. Yano, On harmonic and Killing vector fields, Ann. of Math. 55 (1952) 38-45. , Integral formulas in Riemnannian geometry, Dekker, New York, 1970. SAITAMA UNIVERSITY, JAPAN
J. DIFFERENTIAL GEOMETRY 12 (1977) 599-618
CURVATURE AND SPECTRUM OF COMPACT RIEMANNIAN MANIFOLDS P. GUNTHER & R. SCHIMMING
0.
Introduction
Let M be a compact and orientable Riemannian manifold of class C"" with positive definite metric g and dim M = n. The eigenvalues 2, corresponding to the eigenvalue problem dug + Aw = 0 for alternating, not necessarily homogeneous, differential forms cw which are regular everywhere on M, form a moLet Q3;, notonically increasing (in a strict sense) sequence Spec (M) = be the finite-dimensional eigenspace belonging to A1. The projection operator PIP' which gives the homogeneous part of degree p : (oIP' = PIPla) of any differential form w maps 3i onto UP) = P(P' 3i. In WI) we choose an orthonorP), and introduce mal basis co , 1 Gel /G di %im
(0.1)
01P) (x) = Z <(P
co >(x)
yx E M ,
,
t
is defined by (1.6), (1.5). Now or O?P)(x) = 0 in case 3iP) = {0}, where the following asymptotic expansion holds (for this cf. [19], [18], [3], [5], [10], [91):
e 'it J >(x) o (4;ct)-n 2
(0.2)
k=0
(Sp Vk ))(x)(2t)k
where the V(P)(x, ) form a system of double differential forms' of degree p, defined in the neighborhood of the diagonal of M X M by recursion formulas (see (2.1), (2.2)). One gets the expansion (0.2) by applying the parametric method to Green's form for the heat equation of the manifold M. Integrating (0.2) over M yields (0.3)
i=0
e-'it dim
%,
P) ,_-(4,ct)-n/2 t-»+0
k=0
(2t)k f (Sp V( ))(x)dv(x) . 71
According to the well-known theorems of Hodge [16] and De Rham [6] dim Q50Ov) Received June 12, 1975, and, in revised form, December 15, 1976. ' In the normalization adopted here the V(P)(x, ) coincide with the coefficients of the Riesz kernel forms in [13], [15], [24], [25]. For the Riesz kernel forms and their coefficients further cf. [8], [4]. Note that only double differential forms of "bi-degree" (p, p) appear.
600
P. GUNTHER & R. SCHIMMING
= RP coincides with the p-th Betti number, while for i > 1 a direct decomposition P> holds, where &?> consists of exact forms and A(P) consists of coexact forms ; °> = {0}, Azn> _ {0}. The "theoreme de telescopage" (this name was given to the theorem in [3]) by McKean Jr. and Singer [18] states dim *P> = dim (S+,(P+1> for p = 0, 1, , n - 1. Introducing these facts into (0.3) and taking the alternating sum with respect to p one gets for even n (cf. [18], [31): (0.4)
(-1)PRP = (22r)-"i2
X(M) _
P=0
P=0
(0.5) P=o
(-1)P
fm
(- 1)P f (Sp Vn(ii)(x)dv(x) ii
(Sp V(P))(x)dv(x) = 0
,
dk #
,
n 2
where X(M) denotes the Euler-Poincare characteristic of M. For odd n one has X(M) = 0, and (0.5) is valid for all k > 0. Now it is highly desirable to find relations between the V (P)(x, )-or more exactly their traces-on the one hand and the curvature of M on the other hand. In this direction many partial results are known (cf. H. P. McKean Jr. and I. M. Singer [18], M. Berger [2], E. Combet [5], T. Sakai [22], V. K. Patodi [20], H. Donelly [7], where they considered small values of k and obtained some conclusions on isospectral manifolds). V. K. Patodi [21] has shown that for even n and k < 2In the integrand in (0.5) vanishes while the integrand in (0.4) just equals Chern's invariant occuring in the generalization of the Gauss-Bonnet integral theorem.' Another proof of these facts was given by P. B. Gilkey [111; he studied the general nature of the invariants considered here and gave an interesting characterization of the Pontrjagin and the Chern forms. Generalizations of Gilkey's results which are closely connected with the index-theory of elliptic differential
operators can be found in the papers : P. B. Gilkey [12] and M. Atiyah, R. Bott, V. K. Patodi [1]. In the present paper new results concerning the above mentioned problems are presented. Our method is quite different from those used by the authors listed above. By a coincidence form we mean a double differential form the coefficients
of which are defined only on the diagonal of M x M. Every double form U1111(x, ) defined in a neighborhood of the diagonal gives a coincidence form
UIP>(x, x). Using the metric, the Ricci tensor and the curvature tensor (our convention for the sign of the curvature tensor coincides with that of J. A. Schouten [23]) we define some coincidence forms in the following way:' This is just the content of a conjecture of H. P. McKean, Jr. and I. M. Singer [18];
see also the paragraph after the corollary to Theorem II. A detailed proof of the Gauss-Bonnet integral formula can be found, e.g., in the book of R. Sulanke and P. Wintgen [26], where there is also an extensive list of literature. The multiplication of double differential forms is the exterior multiplication with respect to both groups of the variables, which is commutative and denoted by A.
CURVATURE AND SPECTRUM
G">(x)
G(o>(x)
601
G(P)(x) =
=
1
(0.6)
T'(1)(x) = (0.7)
[G(1>(x)]P
P
VxEM,
dx E M,
T(2) (X) = 2R2.7aa(x)dxi A yf(3)(x)
= 2R',ijRh
A (x)dxi A dx1 A
A die A
By means of these forms we further construct, for k > 1, Z(2k) _ (- 1)k [qyLY(2)]k
Z(1) _ _7Tr(1) ' YY
2"k! (0.8)
-1)k
Z(2k-1) -
r]Ti'(1)
2k-'(k - 1) !
A W(2) _
k-1
A [W(2)]k-2 1/\ L J
2
L
Thus the coincidence forms ', Z are defined explicitly in terms of the curvature tensor of (M, g). Now we can state our theorems. Theorem I. For k > 0 and 0 < l < n - 1 one has n-d
E (- 1)PTY kP (x, x) A \
G(`-P)
P=O
0,
(0.9)
fork
2
LL
for k = r_n - l -!- 1
(_ 1)n ZZcn-a)
2
LL
By taking the trace Theorem I becomes Theorem II. Fork > 0 and 0 < l < n - 1 one has n-I
/
1)P( 72
l (0.10)
P)(sp Vti ))(x)
fork
0
L
(-1)n-'(Sp Z(n-a))(x)
,
for k =
2
n-l 2
-r '
1
.
1
The traces of the Z(11) which appear in Theorem II are given more explicitly by
Corollary to Theorem H.
For k > 1 one has
P. GUNTHER & R. SCHIMMING
602
Sp Z(2k) = - 1 Sp
Z(2k-1)
k
(0.11)
(-1)k(2k) !
Rv2k-li2k]i2k-li2k]
22kk !
For l = 0 and even n Theorem II just gives the above mentioned result of V. K. Patodi [21] and P. B. Gilkey [11], i.e., the conjecture of H. P. McKean Jr. and I. M. Singer [18]. (After seeing our manuscript Professor P. B. Gilkey kindly communicated to us that he was able to prove by means of his method our formulas (0.10) and also the upper bound for the number of invariants following from Theorem V.) For l = 0 and odd n Theorem II and the following two theorems give no new information if one takes the duality properties into consideration. Using Theorem II we obtain the following asymptotic formula. Theorem III. For 0 < l < n - 1 one has n-l
i=0 p=0
(0.12)
(-1)p
n
- Ple-z J 1
- 1[(n-1+1)/27-n/2{(-1)"-12C(n-1+1)/21-n1l-n/2(Sp Z(n-L))(x) + O(t)}
.
Integrating over M and applying the "telescopage" theorem at last one gets Theorem IV. For O < l < n - 1 one has 1)p(n l
=
(0.13)
P ) R + 2:_1)P(n : =0
11, L dip>
//
(-1)n-12C(n-d+1)/27-n7 n/2
tC(n.-L+1)/27-n/2)
Xf
(Sp Z(n-L))(x)dv(x) + O(t)}
The main tool for the proof of these theorems, to be given in §§ 1-2, is a rearrangement of the recursion system defining the double differential forms V(P), called "expansion to transport forms". This method has already been
applied in other investigations connected with Huygens' principle (cf. R. Schimming [24], [25], P. Gunther [13]). An analogous expansion applies also to other geometrical objects forming a graded algebra and allowing the definition of a Laplacian 4, e.g., tensors or spinors. The expansion to transport forms possesses a dual formulation to be treated in § 3, and enables us to draw some more conclusions. As a matter of fact the coincidence forms Vkp)(x, x) are universal in the sense that their components are polynomials of the components giJ, gi;, Ri;h,k, Fim.RiJhk of the metric tensor and its inverse, the curvature tensor, and Fit the covariant derivatives of the latter, the polynomial coefficients being inde-
-
CURVATURE AND SPECTRUM
603
pendent of n = dim M. The traces Sp Vk) then are scalar invariants and polynomials in gig, V2y ViRi;hk with coefficients depending on n. With regard to the duality this first gives 1 + [n/2] invariants for fixed k. The following
theorem proven in § 3 reduces this number to 1 + min {k, [n/2]} and at the same time explains the way in which the Sp VkP) depend on n and p. Theorem V. There are universal scalar invariants ak for 0 < r < k with integers r and k,i.e., polynomials in the contra variant fundamental tensor and the covariant derivatives of the curvature tensor with coefficients independent of the dimension n, such that min(k,p,n-P]
y
Sp VkP) =
(0.14)
( n - 2Y}6k
\p-r
r=°
yk > 0,
ypwith0Gp
For k > 1 one has ak
(0.15)
k
=
(- 1)k S'p Z(2k)
For small values of k the invariants 6k can be given explicitly as follows :
°=1, 0
0 = Sp W(o) =
R
61 = Sp W,,,)
,
R
12
2 = Sp
1 11 VLVLR + (0.16)
4
30
1
180
Ri'kmRi'jkm -
1
180
RiiRi + 1 R2} '
72
62=SpW(1)
4-
6 VLVLR
- 12 Rijk-Rijkm
1 Ri'Ri.; - 6 R2}
62 = SpZI4) = 8 {RijL RijLna - 4Ri'Rij + R2} For the derivation of (0.16) and of course (0.15), developments in a normal coordinate system are used (cf. [13], [14], [15]). The formulas (0.16) agree with a result of V. K. Patodi [20]. In another paper the present results will be applied to Kahlerian manifolds. 1.
Conventions and preliminaries
For two alternating differential forms °p(P),(P) of degree p with the local representation (1.1)
(P (P) = °pi1...i9dxi1 A ... A dxi,
dxi®
P. GUNTHER & R. SCHIMMING
604
+(P) = +il...iPdxi, A ... A
(1.2)
(P(P'(x) Ox *(P'(g) will denote the double differential form with the local representation (P(P)(x) 0 .1.(P)() (1.3)
A ... A
=
A ... Adg°P .
The trace of a double differential form U(P) given by (1.4)
U(P)(x, E)
= Ui,...iPa,....P(x, )dxil A ... A dxiPdr' A ... A dg°P
is the scalar quantity (1.5)
(Sp UcP))(x)
= p! gi,a,(x) ...
x)
Further we define
= Sp ((P (P) ©*(P))(x)
(1.6)
(1.7)
(0 (P) +(P)) = f
where the volume element is given by dv(x) _ *1. Let the dual of the form (1.1) be defined in the usual way by (1.8)
A ... A dxj-
1
(n-p)!
Then instead of (1.7) one can write (1.7')
((P (P),
W (P)) = J
M
(P) (P(P) A **(P)
The terms "orthogonal" or "orthonormal" with respect to differential forms refer to the scalar product defined by (1.7) or (1.7'). The Laplacian d = do + Od for alternating differential forms obeys a product rule (1.9)
d(a(P) A p(q') = (da(P)) A pcq) + a(P) A dp(q) - 2L(a(P', p(Q))
where a(P), p(q) are two forms with p + q < n, and (1.10)
L(a(P', (q)) = Fia(P' A 1ip(Q) + L,(a(P', p(Q))
As one knows, the expression L0(a(P', p(q') is bilinear in the components of the forms a(P', p(q) with the curvature tensor as the coefficients of its terms. We introduce
CURVATURE AND SPECTRUM
(1.11)
:Qjl =
605
A dxi2
and the operation' (1.12)
e;Qp) = Mp;i2...1,dx'2 A ... A dx',
.
Then the well-known representation of 4 (cf., for example, [6]) leads to (1.13)
LO(a(P), 3(q)) = (-1)P.Qi1 A e;(a(P)) A el(3(q))
and this formula remains valid if a(P), (3(q) are double forms. Lemma 1.1. For three forms a(P), p(q), Y(r) with p + q + r < n one has L0(a(P) /p(4) A 7(r)J)
(1.14)
N
(P)
)) A (r)
qr
(P)
(r) A (q)
An analogous formula holds for double forms a(P), p(q), placed by + 1. Proof. Use (1.13) and the formula (1.15)
(r) with (-1)qT re-
A r(r)) = et(3(q)) A Y(T) + (-1l)gp(q> A et(r(T)) Ncase None
,
which is easy to verify. In the of double forms has to take into consideration the commutativity of their multiplication. In the following the transport form T(') plays an important role. Let x E M and E M be two points of M sufficiently near to each other, and denote by T(x, ) the square of their geodesic distance. Then the double differential form T(') satisfies the differential equation
L,(I', Tn)) = 0
(1.16)
and the initial condition (1.17)
Tu>(e, ) = Gu>(e)
In (1.16) the differentiation refer to the variable x, while remains fixed. (1.16) can be interpreted as ordinary differential equations for the coefficients of T(') along the geodesic lines issuing from . Thus T(17(x, ) is defined in a
neighborhood of the diagonal of M x M and is of class C. We further define T(P) by (1.18)
T(P) = 1 p! [T(i)]P
T(o) = 1
'The definition (1.12) is, as far as we know, due to E. Kahler [17]. Also the relation (1.15) was already given in [17], but it should be noted that our inner multiplication defined in §3 differs from the inner multiplication of Kahler.
P. GUNTHER & R. SCHIMMING
606
and easily get the composition rule
T(P) A T() = (P + g1 TcP+q'
-
(1.19)
Lemm 1.2. (1.20)
p
J
For p > 2 one has
dxT(P) - djc1> A T(P-1) - Lx(T(1) T'>) A
T(P-2)
Proof. For p = 2 the relation (1.20) readily follows from (1.18) and (1.9). Now assume it to be true for a certain p. Then substitute it in (1.21) JXT(P+1) =
1
p+l
{4 T(l) A T(P) + T(l) A d,xT(P) - 2L.x(TT1> T(P))}
together with (1.22)
Lx(T(1), T(P))
= L,,(T 1>, T(l)) A
T(p-1)
,
which in turn follows from Lemma 1.1. Thus we arrive at our relation with p replaced by p + 1. Lemma 1.3. The transport forms obey the duality relation Tcn-P>(x )
(1.23)
(The *-operation for double forms refers to both groups of indices and variables as long as no other convention is made.) Proof. Since both sides of (1.23) satisfy the same differential equation (1.16) of geodesic parallel translation (1.24)
Lx(f', Tln-P)) = 0 ,
Ljf', *T(P)) = 0 ,
all one has to do is to perform the routine verification of (1.23) for the initial values, i.e., for the coincidence forms Lemma 1.4. For a double differential form U(P> of degree p, 0 G p G n,
and 0 < q G n, one has (1.25)
Sp (U(P) A TO) _
(n - pUCP> g
)Sp
.
Proof. The relation (1.25) obviously holds for q = 0 and arbitrary p. Assuming it to be true for a certain q and arbitrary p one has Sp (UcP> A T(q+l)) = (1.26)
1
q+ 1
Sp ((U(P) A Ta)) A T())
= (n - p - 11Sp (UcP> q
A Ta>) .
607
CURVATURE AND SPECTRUM
An elementary consideration shows
Sp (Uip) A T11) = (n - p) Sp
(1.27)
U`P)
Substituting (1.27) in (1.26) yields our relation with q replaced by q + 1. In particular, taking p = 0 and U101 = 1 in (1.25) one gets Sp T(q) =
(1.28)
(n) q
2.
Expansion to transport forms
The recursion system for the determination of the double forms VkP), occurring in the asymptotic expansion (0.2), reads as follows (cf., for example, [13])
L,(r, IJop)) + MVp) = 0
(2.1)
(2.2)
Ly(r, Vkp)) + [M + 2k]VkP)
JXVkp)1 ,
k>1
where M(x, ) = 24,zr(x, ) - n. The differential equations (2.1) are to be adjoined by the initial conditions (2.3)
V0P)(e, e)
=
The Vk) with k > 1 are uniquely determined by (2.2) and the regularity condition of the corresponding coincidence forms. Note that in general the V(P) are defined not over the whole, but only in a suitable neighborhood of the diagonal, of M X M. Theorem 2.1. There are regular double differential forms Wk q) of degree
q, 0 < q < n, k > 0, defined in a neighborhood of the diagonal of M X M such that VkP)
(2.4)
=
min[2k,p)
Wk(q)AA T(P-q) q==0
Proof. (2.5)
For the sake of a formal simplification we set
Wkq)=O,Tq)=0,
yq<0,k>0.
Now, to solve (2.1), (2.2), by introducing P
(2.6)
VkP) = g==0 r Wk(q) A A T[p-q)
into (2.2) and making use of L,(r, T(P-q)) = 0, we obtain L,xr, VkP)) + [M + 2k]VkP) + J Vkp)1
608
P. GUNTHER & R. SCHIMMING
_
(2.7)
P
q=o
{Lx(T, Wkq)) + [M + 2k]W(q)} A VP-q)
+ Y, {axWkq)1 A
T(P-q)
q=o
+ Wk4)1
A
4xT(P-q) - 2Lx(W(q)1
VP-q))}
Into the last sums substitution of (1.20), as well as the conclusion following from Lemma 1.1, gives (2.8)
Lx(Wk9)1, T(P-q)) = L.,(Wk21 T(1)) A
T(P-q-1)
.
By using a trivial change of the second summation index and considering (2.5) at last, we can convert (2.2) to P
(2.9)
o
{Lx(I,Wkq)) + [M + 2k]Wkq) + 4,W(q)1 - 2Lx(W( 11) Wk-, 1) A 4T1 -Wkq 12) A Lx(T(l) T(1))} A
T(1))
T(P-q) = 0.
One gets an analogous equation for k = 0. In order to establish (2.1), (2.2) we impose on the forms Wk(q) the following system:
(2.10)
LJT, Woq)) + MWoq) = 0 Lx(-P, Wkq)) + [M + 2k]Wk11)
(2.11)
_ -dxWk2, + 2LJ Wkq 11)
T(1))
-
Wkq 11) A dxT(1)
+ Wkq 1-2) A Lx(T(1) T(1))
The initial conditions (2.3) are fulfilled if we require (2.12)
1
(2.13)
e) = 0 ,
,
dq > 1
.
(2.10), (2.11) represent a recursion system with respect to the increasing k for
the W. In particular, the equations with q > 2k give a separate recursion system for the Wk(q) with q > 2k. This separate system is solved by taking into account the initial conditions (2.13) and requiring (2.14)
Wk q) = 0,
dk > 0, dq with q > 2k
.
Accordingly, (2.6) changes to (2.4). (2.10), (2.11) for q < 2k together with the initial condition (2.12) for k = 0 and with the regularity condition of the coincidence forms Wkq)(, ) for k > 1 now uniquely determine the Wk(q) with
q < 2k. Thus Theorem 2.1 is proved. Theorem 2.2. For integer l and 0 < l < n one has
609
CURVATURE AND SPECTRUM n-L
for 2k
10 (- 1)n-ZW (n-Z)
(2.15) E (-l)PV(P) A T(n-l-P) = P-0
x
Proof. Applying Theorem 2.1 (in its modification (2.6)) and using (1.19) and (2.14) we get n-Z
5 (-1)PV(P) A TI"-'-P) = k
P=0
n-I p
5 (-1)PW(q) A T(P q) A
T(n-t-P)
7C
P=0 q=0
n-Z p n- l- q _E,t(-1)P P=0 q=0 P-q
(q) AT(n-Z-q) Wk
An interchange of the successive summations in the last double sum gives n-an-a
E E (-1)P
q-0 p°q
n -l- q 1.31x(4) A T(n-Z-q) P
q
which reduces to (- 1)n
by a well-known summation formula for the binomial coefficients. Thus we establish our assertion on account of (2.14). Remark. By means of (2.15) the Wkq) are expressed in terms of the V(q),
and in addition to this one obtains certain remarkable relations among the double forms Vk0) themselves. The coincidence forms Z(P) defined in the introduction will turn out to be the coincidence values of some of the W(11). Lemma 2.1. The double differential form T(1) obeys 7C
(2.16)
0
0
(2.17) (4
(2.18)
2
where the covariant differentiations Vi refer to the first argument x, and the coincidence form T(1) was defined in (0.7). Proof. Covariant differentiations of the defining equation of T(1), LX(r, T(1)) =
(FjF)V'T(1) = 0
give
(V VjF)FjT(I) + (vjF)FiFjT(1) = 0 , (37ipipjr)pjTa) + 2(pipjF)(pipjT(1)) + (pjr)pipipjT(1) = 0 Passage to the coincidence values x --> into account
.
establishes (2.16), (2.17) if one takes
(vjr)(e, ) = 0 , (viVjl')(e, ) =
(ViVjVxr)(e, 0 = 0
Now if we set T(1)(x, ) = ti.(x, g)dxid°
,
610
P. GUNTHER & R. SCHIMMING
then, as is well known, the Laplacian of T(1) can be expressed as Q,,Tc1>(x,
{-I'Ijtia(x, ) + R%(x)t1 (x, e)}dxidg°
and the corresponding coincidence form reduces to
Ri( )t1 (, e)dxidg° =
Ria(e)dxidga =
which completes the proof. Theorem 2.3. For integer 1, 0 < l < n - 1, the coincidence forms Z(P) defined by (0.8) obey (2.19)
W
Z(
n ti+1>i27(s
Proof. (a) We start with the case where n - 1 is even. Put n - 1 = 2k. Then [(n - 1 + 1)/2] = k. In view of Wk(q) = 0 for q > 2k the recursion system (2.10), (2.11) gives (2.20)
-
Lx(I Wk2k>) + [M + 2k]Wk2k) = Wki 2) A L,x(Tc1) T11))
yk> 1
.
Now passing to the coincidence values on both sides and considering
M(e, ) = 0
L.x(P, Wk2k))(e, ) = 0
,
we get (2.21)
2kWk2k)(,
yk > 1
(Wk-l 2) A Lx(T')Tc1>))(e,
,
which is solved by in view of (2.12), (2.22)
Wk2k)(e, 0
=
1
2kk!
[L.z(T(') Tu))(e, )]k
,
yk > 1
.
Owing to (2.16), here one can replace the operator L by the operator Lo. Further using formula (1.13) we readily get Lo(T(1> T
Q'L(s)gjj( )gjp(s)ds' A )d a A
Hence (2.23)
jiJk2k)(
, e) _
(2k 011 [Y''(2)(
)]k = Z(3k)(e)
yk>1,
which is our assertion in the considered case. (b)
Now let us turn to the case n - 1 = 1, where [(n - 1 + 1)/2] = 1.
By (2.11), (2.5) and (2.14) we have, for the determination of Wit)
CURVATURE AND SPECTRUM
(2.24)
T(1))
L.,,(j', Wit>) + [M + 2] Wit) =
611
- W(0) A 4 T(1)
Passing to the coincidence values one can replace L by Lo owing to (2.16), but the resulting expression vanishes because the first argument in Lo is a form of degree 0. On account of (2.12) we get (2.25)
e)
and hence, by Lemma 2.1, (2.26)
Will)(ee, e)
O)(e) =
(c) At last we consider the case where n - 1 is odd. Put n - 1 = 2k - 1 with k > 2. Then [(n - l + 1)/2] = k. By considering Wk(q) = 0 for q > 2k, from (2.11) it follows
L,z(j', Wk2k-1>) + [Al +
(2.27)
= 2Ly(YY k2k1 2)
2k]W,(2k-1)
T()) - Wk2ki 2> A Jj(')
Wk-13) A Lx(TO) T(1))
Passing to the coincidence values on both sides and taking account of the results of (a) and (b) we get -(2 - 1)k
2kWk2k-1)(ee, e) = 2k
(k
(2.28)
1
- 1)
!
Lo((
0
(2)) k-1
(-1)k-1
A Y'(1)(0
T2k-2(k - 1) ! Wk2ki 3)(e, ) A ? (2)(e)
Now owing to Lemma 1.1 and (1.13) Lo((
(2.29)
(2))k-1 T(u)(e e)
(k - 1)La(?'12) T(1))(ee, ) A (
(k - 1)Q
c2))k
A ej(T'(2))(ee) A e1(G(1))(ee) A (Y'12)11-2(e)
_ (k - 10(3)(e) A (P'(2))k-2(e) Substituting this in (2.28) one has 2kWk2k-1>(e, e) = -Wk2ki 3)(e, ) A (2.30)
+
[ 2k ((k 1)k 1)
A
(2)
(e)
- (k - 10'(3)(e)] A [Y(2>(e)]k-2
P. GUNTHER & R. SCHIMMING
612
Solving (2.30) by an elementary induction with respect to k, the beginning case k = 1 being given by (2.26), we get
rtjcu()
1)k
2k 1(k - 1) !
(2.31)
A )7!'(2)() Y-
L
which completes the proof of Theorem 2.3. Now we are in a position to prove Theorems I-IV mentioned in the introduction. Proof of Theorem I. Pass to the coincidence values on both sides of formula (2.15) of Theorem 2.2 and apply formula (2.19) of Theorem 2.3. Proof of Theorem II. Take the trace on both sides of formula (0.9) of Theorem I and consider Lemma 1.4 and the fact that the G(P> are the coincidence values of the T(P). be the differentials appearProof of the corollary to Theorem II. Let dxi, ing in a local representation of a double form. Let ej be the operation defined by (1.12) with respect to the differentials dxi, and ea an analogous operation both operations commute with each other. with respect to the differentials Then for a double differential form U(P> one can write Sp U(P) =
(2.32)
1
P
Sp (g"eiea(UcP))
Now an easy computation using (1).15) shows that O i.eiers([Y'
(2)]k)
= kgi'ea(eti(Y' (z)) A\
ei(?j1'(2)) A [Y-(2)]k-1
= kg i'e
(2.33)
[W(2)]k-1)
+ k(k -
1)gi'ei(?Y'(2)) A ea(Y' (2)) A
[l'(2)]k-2
Further, from the definition (0.7) it follows
gile.eiff (2)) = -4p(') ,
9iaei(T* (2)) A e (y'(2)) = 2Y-(3)
Thus by (2.32), (2.33) one has the assertion of the corollary : Sp Z(2k) _
(- 1)k Sp [p (2)]k k
k!
(-1) k-1
L
Sp
Y" A Y' (2) - k ([.i
2k -1k !
I Sp Z(2k-1)
k
2
1
T-(3) A [?r(2)Ik-2 L
613
CURVATURE AND SPECTRUM
Proof of Theorem III.
Multiply the expansion (0.2) by (-1)P (n
pl
0 < p < 1, and add the resulting expression. Substituting the results of Theorem II in the right of the obtained equation we arrive at (0.12). Proof of Therem IV. First, one has to integrate formula (0.12) over M. For i > 1 one takes into consideration n-l
(-1)P n P=O n-i
\
(p)
p l dim
E (-1)P
n
pl{dim
(
P) + dim Si P)}
1
n_z
_ E (-1)P (n \
1
pl dimA2p ' + E (-1)P(n p=o
n-\ )
n-z
E (-1)p{(n-Tp)
_ E(-1)P(n
p
1
p ) dim l
\\
- (n - p 1
1l dim S p)
J
1
which establishes Theorem IV. 3.
Dualisation and Conclusions
Definition. Let X(P), Y(P+q) be double differential forms of degrees p, p + q (q > 0) respectively. A double form X(P) V Y(P+q) of degree q is defined by (3.1)
X(P) V y(p+q) _ *(X(P) A *y(p+q))
.
The following theorem gives an expansion dual to the expansion to transport forms of Theorem 2.1. Theorem 3.1. For 0 < p < n and the double forms Wk(q) considered in § 2 one has (3.2)
IJ(kn-p) =
min(2k,p)
Wk(") V T(n-P+q) q=0
Proof.
Taking the dual of both sides of (2.4) yields minl2k,p}
(3.3)
*(Wkq) A T(p-q)
VkP) _ q=o
By Lemma 1.3 we can write (3.4)
T(p-q) = *T(n-P+q)
Further we use the well-known relation
614
P. GUNTHER & R. SCHIMMING
*VkP) =
(3.5)
ykn-P)
which originates from the duality properties of Green's form of the heat equation. Substituting (3.4) and (3.5) in (3.3) one can apply the above definition and thus derive the assertion of Theorem 3.1. Lemma 3.1. For a double differential form UIP) of degree p one has (3.6)
Sp (*U(P)) = Sp U(P)
The proof is a matter of routine. Lemma 3.2. For a double differential form U(P) of degree p, 0 < p < n, one has
ql Sp
Sp (U(q) V T( n-P+q)) = (n
(3.7)
p
U(q)
q
The proof follows from our definition and Lemmas 3.1, 1.3 and 1.4. Theorem 3.2. For the double forms Wk(I) considered in § 2 one has, dk
>0,dpwith0
min[2
- q\ Sp Wkq = :n ,P} (n - q1
-P}
(n p
q=o
p-
q=u
/I
Sp
k
qJ
Proof. We take the trace on both sides of (3.2) in Theorem 3.1, and obtain by applying Lemma 3.2 min{2k,p1
Sp ykn P) = E
(3.9)
q=o
( n- ql Sp W(,,)
p-q
Further we replace p by n - p and take the trace once more on both sides of (2.4) in Theorem 2.1. This yields (3.10)
Sp
Vkn-P) =
min{2En PI (n - ql Sp Wk(q) q=o
J
The comparison between (3.9) and (3.10) leads to the assertion. Lemma 3.3. For integers p and r with 0 < r < p < n and 0 < r < 2n one has (3.11)
(p
- rr/
q
(-1)q+T (p
- q)(q
r
r)
The proof is carried out by induction with respect to r. Now we come to the Proof of Theorem V. To begin with we consider for fixed k > 0 the linear equation system for quantities a k', 0 < r < [2n] :
615
CURVATURE AND SPECTRUM
(3.12) Sp Wk = r (-1)p+r(\p o
r
Jar
-
n
`dP with 0 < P< 121
,
The matrix of the coefficients of the linear system has triangular shape with principal diagonal consisting of numbers equal to one. Hence the ak, 0 < r < [2n,] are uniquely determined as linear combinations of the Sp Wkp>, 0 < p < [2n], with integer coefficients. We now show that the quantities 6k thus defined obey r
min[p,[n/2][
(3.13) Sp Wk(P) =
E r=o
(-1)p+r
dpwith 0
p - r )6k
To this end we form the sums
L (n - q) {Sp l Wkq' I = q=o ``p
(3.14)
-
-q
P
min [q;[n/2]{
E
(- 1)r+q
r=o
r
q- r
k
Obviously Ip = 0 for 0 < p < [2n] owing to (3.12). On the other hand for p > 2n we get, setting p' = n - p and formally taking into account Theorem 3.2,
(n - /\q Spp (n 1 r+q kq) Z Z,(-) _Eq=o \ q-r r/P-- qq) 7 r=oq=r np [n/2]
P-
IP
p
Here we can apply (3.12) and Lemma 3.3 so that
(_1),,+r (n =q)(
I
q q- r
p'
q=O r=o
r
)6k _ [n] (n - 2r)6k r=o
p
r
(n'-2r)6r- P (n-2r)6r=0. r=0 p- Y p - r k
r=0
The equations Ip = 0 for 0 < p < n, however, form a linear homogenous system for the quantities between the braces in (3.14), the matrix of which / has triangular shape and has the numbers (n 0 P) 0 in the diagonal. This \ establishes (3.13). Further we show
6k=0
(3.15)
fork
To this end let k < [2n]. In (3.13) we choose p = 2[2n] to begin with. Because of Wk(q) = 0 for q > 2k we get [n/2] rr=O
1 r (-)r\2[2n] -r
6k =
(-1)[n/2]akn/2]
=0.
616
P. GUNTHER & R. SCHIMMING
Assume Qk = 0 to be shown already for [2n] - l + 1 < r < [2n], where ad-
ditionally [2n] - l > k, otherwise the proof is finished. Then in (3.13) we set p = 2([2n] - 1). Again with p > 2k one has (3.16)
r
min[2[n/2]-21,[n/2]}
(2[2n]
- 2l -
Ja'
=0
By the induction hypothesis the summation index r can be limited according
to r < [2n] - 1. Then, however, kn/2]_i = 0 follows immediately from (3.16). Thus (3.15) is proved. Return now to (3.9) written in the equivalent form Sp VIP)
(3.17)
E (n - ql Sp Wk(q) = q-o p - q
Substituting (3.13) in (3.17) taking account of (3.15), interchanging the successive summations and applying Lemma 3.3, we thus get (0.14) of Theorem V. It remains to show that the ak are universal invariants. To this end we think the coefficients of the double forms Wxq)(x, ) as functions of x expanded to finite Taylor's series about with a remainder of a sufficiently high order. Performing the expansion in a normal coordinate system with origin the individual Taylor terms can be expressed according to a standard procedure as
polynomials, with coefficients independent of the dimension, in the metric tensor and its inverse, the curvature tensor and the covariant derivatives of
the latter.' This can be arranged in such a way that only the gig and occur, not the g,5. For the lowest double form W0(°) it can be seen from the formula (in normal coordinates) Fit
(3.18)
Woo)(x, sue)
=
g = Det (gay)
For the higher double forms Wk(q) it can be seen by an induction with respect to k using the recursion system (2.11). The gi5 occur only in the coincidence
value, that is, only in the Taylor term of zero order, of the transport form TT'). Since no terms gj1 occur in the coincidence forms no factors n = gi''gij can appear in the traces Sp Wkq). Thus the Sp Wkq) turn out to be universal invariants, and the same holds for the because the linear equation system (3.12) for the determination of the Qk by the Sp Wk(q) is independent
of n. This completely establishes Theorem V. Proof of the corollary to Theorem V. Owing to (3.15) we can write (3.13) in the form Sp
WkP)
_
(-1)P+''( r
p -r)Qk
5 In [141 explicit formulas are given by means of which the derivatives of the metric tensor in normal coordinates are expressed by the curvature tensor and its covariant derivatives.
617
CURVATURE AND SPECTRUM
Now put p = 2k. Then the sum on the right-hand side contains at most one term, while to the left-hand side we apply (2.23) so that we get (0.15) and, in addition,
a[k,r]+l _ ... =6k=0
for k> [n] 2
References
[1] [2] [3]
[4] [5] [6] [71
[8]
[9]
M. Atiyah, R. Bott & V. K. Patodi, On the heat equation and the index theorem, Invent. Math. 19 (1973) 279-330. M. Berger, Le spectre des varietes Riemanniennes, Rev. Roumaine Math. Phys. Appl. 13 (1968) 915-931. M. Berger, P. Gauduchon & E. Mazet, Le spectre d'une variete Riemannienne, Lecture Notes in Math. Vol. 194, Springer, Berlin, 1971. E. Combet, Solutions elementaires des d'Alembertiens generalises, Memor. Sci. Math. Vol. 160, Gauthier-Villars, Paris, 1965. , Parametrix et invariants stir les varietes compactes, Ann. Sci. Ecole Norm. Sup. 3 (1970) 247-271. G. de Rham, Varietes diflerentiables, Hermann, Paris, 1955. H. Donnelly, Symmetric Einstein spaces and spectral geometry, Indiana Univ. Math. J. 24 (1974) 603-606. G. F. D. Duff, Harmonic p-tensors on normal hyperbolic Riemannian spaces, Canad. J. Math. 5 (1953) 57-80. A. Friedman, The wave equation for differential forms, Pacific J. Math. 11 (1961) 1267-1279.
M. P. Gaffney, Asymptotic distributions associated with the Laplacian for forms, Comm. Pure Appl. Math. 11 (1958) 535-545. [11] P. B. Gilkey, Curvature and the eigenvalues of the Laplacian for elliptic complexes, Advances in Math. 10 (1973) 344-382. , Local invariants of a pseudo-Riemannian manifold, Math. Scand. 36 (1975) [12] [10]
[13]
109-130. P. GUnther, Einige Sdtze fiber huygenssche Diflerentialgleichungen, Wiss. Z. KarlMarx-Univ. Leipzig, Math.-Natur. Reihe 14 (1965) 497-507.
[14] -, Spinorkalkiil and Normalkoordinaten, Z. Angew. Math. Mech. 55 (1975) [15] [16] [17]
[18] [19]
205-210. P. GUnther & V. Wiinsch, Maxwellsche Gleichungen and Huygenssches Prinzip.
I, Math. Nachr. 63 (1974) 97-121.
W. V. D. Hodge, Theorie and Anwendungen harmonischer Integrale, B. G. Teubner, Leipzig, 1958. (The German translation from the English edition.) E. Kahler, Innerer and iiusserer Differentialkalkiil, Abh. Deutsch. Akad. Wiss.
Berlin Kl. Math. Phys. Tech 4 (1960). H. P. McKean, Jr. & I. M. Singer, Curvature and the eigenvalues of the Laplacian, J. Differential Geometry 1 (1967) 43-69. S. Minakshisundaram & A. Pleijel, Some properties of the eigenfunctions of the Laplace-operator on Riemannian manifolds, Canad. J. Math. 1 (1949) 242-256.
V. K. Patodi, Curvature and the fundamental solution of the heat operator, J. Indian Math. Soc. 34 (1970) 269-285. [21] -, Curvature and the eigenforms of the Laplace-operator, J. Differential Geometry 5 (1971) 233-249. [22] T. Sakai, On eigenvalues of Laplacian and curvature of Riemannian manifolds, Tohoku Math. J. (2) 23 (1971) 589-603. [23] J. A. Schouten, Ricci-Calculus, 2nd ed., Springer, Berlin, 1954. [24] R. Schimming, Riemannsche Metriken mit ebener Symmetrie and das Huygens[20]
sche Prinzip, Thesis, Leipzig, 1971.
618
P. GUNTHER & R. SCHIMMING
[25]
, Zur Giiltigkeit des Huygensschen Prinzips bei einer speziellen Metrik, Z. Angew, Math. Mech. 51 (1971) 201-208.
[26]
R. Sulanke & P. Wintgen, Differentialgeometrie and Faserbiindel, Deutsch. Verlag Wiss., Berlin, 1972. KARL-MARX UNIVERSITY, LEIPZIG
J. DIFFERENTIAL GEOMETRY 12 (1977) 619-627
THE SECOND FUNDAMENTAL FORM OF A PLANE FIELD BRUCE L. REINHART
This paper arose out of attempts to understand geometrically the meaning of various foliation invariants introduced in the last few years. Because these invariants are associated to the normal bundle, the characteristics of the normal plane field are important. Since the normal plane field is not integrable, one is led to study the Riemannian geometry of arbitrary plane fields, which is done by generalizing the second fundamental form. Concepts such as mean curvature and minimality can then be introduced for a plane field, and it can be shown that a totally geodesic plane field has vanishing second fundamental form. This is of interest because the normal plane field to an R-foliation is totally geodesic. Given a foliation of a Riemannian manifold, a foliation connection is chosen in the normal bundle which is as compatible as possible with the Riemannian connection. Certain formulas are developed for the components of the connection and curvature forms, then used to prove a number of results, including : the leaf classes hi for odd i depend only on the second fundamental form of the normal plane field, the Godbillon-Vey class in higher codimension is given by a formula analogous to that of Reinhart and Wood [8], and the reductions modulo the integers of certain leaf classes of dimension 4j - 1 are independent of the choice of framing (they are defined only for framed foliations). A method for calculating the cohomology of truncated relative Weil algebras is essential to obtain the results. Such a method has been given in general by Kamber and Tondeur [9], [10], while more recently Guelorget and Joubert [5], building on their work, have given very explicit formulas for the case needed here, the general linear algebra modulo the orthogonal algebra. Finally, some examples are given of vector fields in euclidean 3-space such that the normal plane fields are not integrable, and their second fundamental forms have certain prescribed properties. 1.
Plane fields
A smooth p-plane field on a smooth Riemannian n-manifold is assumed. The inner product will be symbolized by < , >, while the Riemannian covariReceived December 13, 1975, and, in revised form, August 25, 1976. Research supported by the National Science Foundation under grant MPS 73-08938A02.
620
BRUCE L. REINHART
ant derivative, connection form, and curvature form are denoted by 17*, 9*, , p} is a local orthonormal basis for the plane field, , q = n - p} is a local orthonormal basis for the normal plane {XZ I i = 1, field, and {¢°, (oil is the corresponding coframe. U, V, W denote arbitrary tangent vectors of the frame and X, Y, Z arbitrary normal vectors belonging to the frame. The second fundamental form T of the plane field is defined by the formulas : and Q* . {Va I a = 1,
TX=0,
,
J, U, X>
,
The first and third formulas imply immediately that T is tensorial with respect to X, while the tensoriality with respect to U and V follows from
In the case that the plane field is integrable, T is exactly the second fundamental form of the leaves as immersed submanifolds. In any case, it is a symmetric 2-form with values in the normal bundle. In a completely analogous way, we define the second fundamental form S of the normal plane field : <sXY, U> = 2
sU=o,
,
<SX U, Y>
<5XY, U> ,
<sXY,z>=<sXU,v>=o.
Each of these tensors has a well-defined trace
Z. T,aV.,
E i SXiX Z
which is a normal (respectively tangent) vector of the original plane field called the mean curvature vector. A plane field will be said to be minimal if the trace of its second fundamental form is the zero vector, and to be totally geodesic if each geodesic that is tangent to it at one point remains tangent for its entire length. The following proposition generalizes a well-known property of submanifolds. Proposition 1. If a plane field is totally geodesic, then its second fundamental form is identically 0. Proof. Since each unit vector U belonging to the plane field is the initial vector of a geodesic,
In particular, the hypotheses are satisfied by the normal plane field to a
SECOND FUNDAMENTAL FORM OF A PLANE FIELD
621
foliation with bundle-like metric (or R-foliations), so the second fundamental form of the normal plane field vanishes in this case. In a way completely ananlogous to the definition of the second fundamental form, vector valued antisymmetric 2-forms B and A are defined. The equation
[X, Y] = VXY - F*X shows that A = 0 if and only if the normal plane field is integrable. Similarly, B = 0 if and only if the given plane field is integrable. Hence these are known as the integrability tensors. Foliations
2.
For the remainder of the paper, it will be assumed that the given plane field is integrable, hence defines a foliation of codimension q. Then B = 0 and T is the usual second fundamental form of the leaves. Furthermore, there is a connection in the normal bundle which is well-adapted to the foliation and the metric [5] in the sence that
FXY =17 Y
,
.
The corresponding connection and curvature forms will be denoted by B and Q. In terms of a local frame we have
VyXi = Zj Bj2(Y)Xj , VVXi = j Oji(V)Xj , Qji = d8j1 + Zk 0k2 A Bjk B*ji = -B*i.j
j w' A &j2 Q*jZ = -.Q*ii .
dwi =
The first two equations are the definition of Bji, the next two are proved by Guelorget and Joubert [5], and the last two are a restatement of the fact that the Lie algebra of the orthogonal group consists of skew-symmetric matrices. Bji and Sji are neither symmetric nor skew-symmetric with respect to their indices, but useful formulas can be given for their symmetric and antisymmetric parts. Let B,sji = 2 (Bji + Big) and BAji = 2(O - Big) so that Bji = Bsji + BAj2.
In a similar manner, Qj' = Qs ji + QAjk. Proposition 2. 8?(X) = BAj2(X) = B*ji(X) 0A/(V) = B* ji(V) +
DA/
,
'+ d0Aj2 '++ Ek BAk2 ^ / \ BAjk + LJk BSk' / \ BSjk ^
+2Lk
0s j71 - BSk2 ^ / \ BAjk)
:-S/ - dBSii + 2Z k (BAki A BSjk + BSkz A BAjk) Proof. The first formula is obvious, while the second and third formulas follow from
622
BRUCE L. REINHART 0.7Z(V)
=
=
The property of 0 which makes it interesting for study of characteristic classes is the Bott vanishing theorem [2], which is based on the following property of the curvature :
(1)
Q;i(Va, Va) = 0 . 3.
Characteristic classes
The real valued characteristic classes of a vector bundle are represented in de Rham cohomology by products of differential forms ci which locally can be written
ci = tr (S)1) - Cl,... Z a9a1 A Qa3a2 A . . ai
.
S) l,-
,
where S); is the curvature form of any connection in the bundle. If i is odd, ci is cohomologous to zero. Guelorget and Joubert [5] give the following local formula for global differential forms hi, i odd, such that dhi = ci :
(2)
hi = i tr
f dtO8{tSQB
j
1
+ 2A +
(t2 - 1)082}i-1
0
where the multiplication of Lie algebra valued forms is interpreted in the same manner as in the formula for ci. ci is a scalar valued form of degree 2i, while
hi is a scalar valued form of degree 2i - 1. Formula (1) implies that for the normal bundle of a foliation of codimension q, any product of ci's of total dimension greater than 2q is 0, hence that the characteristic ring vanishes above dimension 2q (see Bott [2]). The characteristic classes of the foliation are by definition the cohomology of the cochain complex consisting of products of the hi (for i odd) and ci (for all i) with the operation of exterior differentiation and the Bott vanishing relations [1], [3], [6]. In particular, the Godbillon-Vey class h1c1q will be of interest in this paper. The formula (1) also
implies that the restriction of hi to any leaf is a closed form. This defines a ring of leaf classes on any leaf [7]. The aim of this section is to relate these classes to the geometry of the foliation by means of the Guelorget-Joubert formulas.
SECOND FUNDAMENTAL FORM OF A PLANE FIELD
623
Lemma 1. hi = Za+,e+r=i-i Ba,er tr
(Qsa2Ae0szT+i
where
(a+P+r+ 1)! !(a+1)(a+3)... (a+2r+ 1)
BaPr=(-2)T a! Proof. cients
Expand formula (2) by the multinomial formula to obtain coeffi-
Bapr =
i. a! f3! r!
f ta(t, - 1)rdt , o
which can then be evaluated by integration by parts. In particular h1 = tr (Bs) so that h1(V) = (3)
i
Bsii(V) _
i
h,(X) = 0 . Thus hl is determined completely by the trace of the second fundamental form of the normal plane field, that is, its mean curvature vector. If the mean curvature vector is nonzero at a point, then in the neighborhood of that point it may be written as rcV, where V1 is a unit vector. Henceforth, it will be supposed that in any local basis {Va} for the tangent plane to a foliation, V, is so chosen at any point where it is possible to do so. Then h,(V) = K
formula for the value on an i-tuple of tangent vectors, since again only the pure Bs term occurs. That is, we need to consider only the term
(-2)i-li !
(4)
tr (Bs2i-1)
1.3.5 ... (2i-1)
which still depends only on the second fundamental form of the normal plane field.
Lemma 2.
If k # 0, then for a > 1
Xk) = -k{
= 1irz
if
V .>j
(-1)r(-1)1414+1))/2 kq+1 fjk_1q {
where the summation is overall permutations of {a
, aq}.
624
BRUCE L. REINHART
Proof.
Bsii(V*vj, - v*x.V.)
desii(V., X.)
c1(V,, X.) _
Tti
= -,\ + + , 1, v*x. a> from which the first formula follows. Since h1(V,) = hi(X,) = 0, the formula for h1c1q involves only h1(V1)c1(V«l, X1) ... c, (V,,, Xq)
which immediately implies the result. The tensor w*xkV1i Va> = t(X,, Va) Will be called the torsion of the normal plane field, by analogy to the codimension-1 case. Theorem. The form h1 defining the first leaf class is the metric dual of the mean curvature vector of the normal plane field. More generally, the leaf classes hi for odd i depend only on the second fundamental form of the normal plane field. The Godbillon-Vey class h1c1q depends on the mean curvature vector and the torsion tensor of the normal plane field, and on the second fundamental form of the leaves. If the normal plane field is minimal, h1 and h1c1q vanish, while if it is totally geodesic, all the leaf and foliation classes vanish. Proof. The first statement follows from the formulas (3) and the fact that the mean curvature vector of the normal plane field is tangent to the leaves. The second statement follows from the fact that the leaf classes are defined by restriction of hi to the leaves, so that formula (4) applies. At points where x 0, the Godbillon-Vey form is given by Lemma 2, while at points where k = 0, h1(V) = 0 for all V, so the Godbillon-Vey form vanishes. (Recall that h1(X) always vanishes.) This proves the third statement and the first part of
the fourth statement. If the normal field is totally geodesic, 08 = 0, while every term in the formula of Lemma 1 contains a positive power of 0,s. This completes the proof of the theorem. 4.
Foliations with trivial normal bundle
In general, the bases {Va, Xi} and {¢", mi} can be constructed only locally, but the case where they exist globally is of special interest. The ci are defined as before, but when the normal bundle is trivial, they are all cobundaries. In the classifying space for the general linear group, however, they are coboundaries only on the total space, not on the base space, so a universal formula for hi, i even, with dhi = ci can be given only on the principal bundle. By using a section, the hi can be pulled back to forms on the base space, also denoted by hi, but which depend on the choice of section. A formula for hi is [5]
hi = i tr
J
dtB{tQ + (t2 - t)62}i-1 0
,
SECOND FUNDAMENTAL FORM OF A PLANE FIELD
625
which by the same methods as in the proof of Lemma 1 is shown to be
hi =
Aap tr (Q202P+1)
where Aa0 = 1 and for R > 0
a ! (a + 2 + 1) ! The cochain complex consisting of cochains hi and ci, with the operation of exterior differentiation and the Bott vanishing relations gives rise to a set of characteristic classes which partially overlap those defined in the case of arbitrary normal bundle. For example, in codimension 2, in the general case the classes are c2i h1c12, and h1c2 while in the case of trivial normal bundle, the classes are h1c12, h1c2i h2c2, h1h2c12 and h1h2c2. Also, the hi for even i give
rise to leaf classes by restricting to a leaf, just as they do for i odd. The object of this section is to obtain certain information about the way these classes depend on the geometry of the foliation and on the choice of frame. Proposition 3. The leaf class hi for i even depends on the framing, the integrability tensor of the normal plane field, and the second fundmental form of the normal plane field. If the normal field is integrable and admits a Riemannian parallel framing, the leaf class is 0. If it admits a Riemannian parallel framing, the leaf class is independent of the choice among such framings. Proof. The only term in the formula for hi which affects the leaf class is
_ (i - 1) !
i!
(2i - 1) !
tr 02i-1
On the other hand OSaaalOSasaZ A ... A 0'3a'-2i -I
_
BSa1"Z A 0saaa3 A ... (- 1)(i-1)(2i-1) E BSaaa-1a1 A .
AOSaai-1a1
.
. A BSaa9 A Osa,- = 0 ,
if i is even, since the last sum equals the first sum. Hence writing 0 = BA + 0S and expanding gives a formula in which every term contains a positive power of 0A. The formulas for 0A and 0S given in the last section then imply the result. Let k be a real number such that icci is an integer class in the classifying space. Then Chern and Simons [4] have defined a cohomology class (Kh), E H22-1(M ; R/Z)
on the base manifold M of any smooth vector bundle. This class is called a
626
BRUCE L. REINHART
Pontrjagin character. It is associated to Iehi in the sense that its pull back to the principal bundle is the reduction modulo Z of the real class defined by Ichi. Hence the restriction to any leaf of (1ehi)1 is the reduction modulo Z of the leaf class. Since (1chi)1 does not depend on any choice of frame, neither does the reduction of the leaf class. Thus the following proposition has been proved. Proposition 4. The reducation modulo the integers of the leaf class hi (i even) does not depend on the choice of framing, and is in fact the restriction to the leaf of the Pontrjagin character P; (j = i/2) of the normal bundle. Examples
5.
The simplest nonintegrable plane fields are normal p-plane fields to a vector field in euclidean three-space, but even in this case there are examples of totally geodesic plane fields, minimal fields which are not totally geodesic, and fields with nonzero principal curvatures of the same or of opposite sign. Any Killing vector field is an R-foliation, hence gives S = 0. In the example
-y
a
ax
+xay
az
the straight lines normal to the orbits may be seen fairly easily. The nonzero components of A are given by +(1 + x2 + Y2)-1/2(1 + x2)-1/2(1 + Yz)-1/z .
A minimal plane field which is not totally geodesic is given by the normal field to
a a.
-x ay
az
The principal curvatures (that is, the eigenvalues of S) are +(2 + 2x2)-1 and the nonzero components of A have the same values. The tangent vector fields to the lines of curvature (that is, the corresponding eigenvectors) are +2-1/2
/2
ax
+ (2 + 2x2)
2X2)-1/2
ay
+ x(2 +
az
Perturbing slightly two-dimensional foliations by surfaces of positive (respectively negative) curvature will give examples of plane fields which are not integrable and which have principal curvatures of different absolute value and the same (respectively opposite) signs.
SECOND FUNDAMENTAL FORM OF A PLANE FIELD
627
References [1]
]2[ [3] [4]
N. Bernstein & B. I. Rosenfeld, Szzr les classes characteristiques des feuilletages, J. Functional Analysis 6 (1972) 68-69. R. Bott, On a topological obstruction to integrability, Proc. Sympos. Pure Math, Vol. 16, Amer. Math. Soc., 1970, 127-131. R. Bott & A. Haefliger, On the characteristic classes of r-foliations, Bull. Amer. Math. Soc. 78 (1972) 1039-1044. S. S. Chern & J. Simons, Characteristic forms and geometric invariants, Ann. of
Math. (2) 99 (1974) 48-69. S. Guelorget & G. Joubert, Algebre de Weil et classes characteristiques d'un feuilletage, C. R. Acad. Sci. Paris Ser. A-B, 277 (1973) All-A14. [ 61 F. W. Kamber & P. Tondeur, Characteristic classes of modules over a sheaf of Lie algebras, Notices, Amer. Math. Soc. 19 (1972) A-401. [5]
[7]
[8] [9]
B. Reinhart, Holonomy invariants for framed foliations, Geometrie differentielle, Colloque, Santiago de Compostela, 1972, 47-52; Lecture Notes in Math. Vol. 392, Springer, Berlin, 1974, 47-52. B. Reinhart & J. Wood, A metric formula for the Godbillon-Vey invariant for foliations, Proc. Amer. Math. Soc. 38 (1973) 427-430.
F. W. Kamber & P. Tondeur, Cohomologie des algebres de Weil relatives
tronquees, C. R. Acad. Sci. Paris Ser. A-B, 276 (1973) A459-A462. [10] -, Characteristic invariants of foliated bundles, Manuscripta Math. 11 (1974) 51-89.
UNIVERSITY OF MARYLAND
J. DIFFERENTIAL GEOMETRY 12 (1977) 629-633
A NOTE ON A THEOREM OF NIRENBERG TAKAAKI NISHIDA
The abstract forms of the nonlinear Cauchy-Kowalewski theorem are investigated in [1] and [2] in a little different formulations. We note here that the Nirenberg's formulation and proof in [1] can be simplified to give an improved abstract nonlinear Cauchy-Kowalewski theorem in a scale of Banach spaces, which contains both theorems in [1] and [2]. The proof follows that of Nirenberg exactly except one point. Definition. Let S = {B,},,, be a scale of Banach spaces, and let all B,, for p > 0 be linear subspaces of Bo. It is assumed that B, C B,,, J ' < IIP for any p' < p, where HI IIP denotes the norm in B. Consider in S the initial value problem of the form (1)
du
dt
= F(u(t), t)
(2)
tJ
,
u(0) = 0 .
Assume the following conditions on F :
(i)
For some numbers R > 0, ri > 0, po > 0 and every pair of numbers
p, p' such that 0 < p' < p < p, (u, t)
(3)
F(u, t) is a continuous mapping of
{uEB..;1u11P
(ii) For any p' < p < po and all u, v c B, with for any t, I t < ri, F satisfies
(4)
11 F(u, t) - F(v, t) I
I
Iu
.
< R, 11 v
< R, and
u-v11'/(P-P),
where C is a constant independent of t, u, v, p or p'. (iii) F(0, t) is a continuous function of t, I t I < ri with values in B,, for every P < Po and satisfies, with a fixed constant K, (5)
IIF(O,t)I1P
0
Theorem. Under the preceding hypotheses there is a positive constant a such that there exists a unique function u(t) which, for every positive p < Po and tI < a(po - p), is a continuously differentiable function of t with values in BP, Iu(t)I1,o < R, and satisfies (1), (2). Communicated by L. Nirenberg, February 12, 1976.
TAKAAKI NISHIDA
630
Remark. The assumption (ii) on F is simpler than that in [1] or [2]. Proof. Let B be the Banach space of functions u(t) which, for every nonnegative p < po and It I < a(po - p), are continuous functions of t with values in BP, and have the norm
(6)
sup a
M[u] =
Itl
(
i u(t) P
a(po
P)
p)
- i) < + oo
Iti
1
We seek a solution of
(7)
u(t) =
rt J0
F(u(r), r)dr
with finite norm M[u] with a suitably small. Our solution will be obtained as the limit of a sequence u, defined recursively by
u0=0,
(8)
uk+1 = uk+vk
where
(9)
for I tI < ak(p0 - P)
Iluk(t)IIP < R
and vk is defined by (10)
vk(t) =
ft
F(uk(r), r)dr - uk(t) ,
0
i.e., t
uk+1(t) = f 0 F(uk(r), r)dr .
Here, for every p < po and Its < ak(po - p), uk(t) and vk(t) are continuous functions of t with values in B,, for which Mk[vk] are finite, where (11)
Mk[v] =
o
sup a
i
v(t)11,
(ak(Po - P)
-1
It 0 aP< -P)
the numbers ak being defined by (12)
-(k+2)-Z),
ak+l=ak(1
k=0,1,2,...
so that (13)
a=ao]-[(1-(k+2)-Z)>0, 0
and ao will be chosen suitably small later.
A THEOREM OF NIRENBERG
631
Let us imagine that ui are determined for i G k with Mi[ui] < + co and Ilui(t) IIP < 2R for It! < ai(po - p). By the assumption (i), vk(t) is well defined. Set 2k = Mj[vk] < + oo
(14)
.
Then for t j < ak+l(po
Ilvk(t)I
- p) I
aklak+l - 1
and it follows that for tI < ak+1(p0 - p) uk+1(t) IIP G
2k
ak/ax+1
-1
ux(t)
IIP
,
and so, by recursion, k
(15)
I uk+1(t) IIP < Z A>l (ajl a,+1 - 1) .
We will require that k
Z Ajl (ajl a,+1 - 1) < 2R .
(16)
0
Then for I tJ < ak+l(po - p) we
have II uk+1(t) II P
< 2R and so F(uk+l(t), t) is
defined.
Our aim is to estimate Ak so that Ak -> 0 as k -> + co, and (16) holds for any k > 0. By (8) and (10) we have vk+1(t) = ft [F(uk+1(T)1 T) - F(uk(T), T)]dT .
Thus for Itl < ak+1(po - p), we see from the assumption (ii) that it vk+1(t)IIP < C
('`
f
Jo
Ilvx(T)IIP(r) dzl
p(T) - p
for some choice of p(T) < po - T I /ax We may set p(T) = 1(po - I tI /ak+1 + p). Then we find by virtue of (14) (assuming, say, t > 0) t
Ivk+1(t)
IIP
T(ak+l(po - p) - T)-zdT
4Cak+1?ik 0
G 4Cak+'Akt f (ak+l(po - p) - T)-ZdT 0
632
TAKAAKI NISHIDA
= 4Ca
i
t a x+1(p0 - p)
/(al(pot - P)
/
Consequently 2k+l = Mk+1Lvk+1] < 4Cak}12k
t
sup 0
- p)lt -
ak+1(po
ak+1(P0 - p)
1
ak+l(po - p)/t - 1
< 4Cak+lak < 4Ca02x
Hence fork=0,1,2, (17)
2k_i < 4CaO2k
Now we can choose a0. Using the assumption (iii) we know that , l o = MO
f t F(0, r)d,r
J
L LJO
sup
-
(a0(P0 - P)
ItI
tl
Itl
0
11 < a0K
/
We shall require that for j = 0, 1, 2, 2j < 24a0K(j
(18)
2)-4
Assuming (18) to be true for 2k we find from (12) and (17) 2k+1 < 4Ca02'a0K(k + 2)-4 l4
< 24a0K(k + 3)-' 4Cao
k + 21) < 24a0K(k
+ 3)-4
provided ao < a' independent of k. We have to verify (16). From (12) and (18)
u kk
2S/(aj/aj+1 - 1)
0
k
k
Z 2j/(l - aj+1/aj) = E 25(j +
2)1
0
0
k <24a0KE(j+2)-1<24a0KE(j+2)-'<2R
0
0
provided a0 < a". If we choose a0 < a' and a0 < a", we find the functions uk are defined for all k with (19)
for Itl
lluk(t)IUP < 2R ,
- p)
Furthermore, from (14) we have for Itl < a(po - p) < ak(po l uk+1(t) - uk(t) Il P <
2k/(a/c(PO - P)
ItI
- p)
- 11 < 2kA a(po - P) -
M[uk+1 - uk] < 2k
11
e
A THEOREM OF NIRENBERG
633
Since Z 2k < + co, it follows that uk converges to some u(t) in B. From (19) we have 11 u(t)11 G 2R for ItI < a(po - p). The limit u(t) is the unique solution of (7) and therefore of (1) and (2) because of the same arguments in [1]. Added in proof. The theorem can be generalized for the integral equation Of the form
u(t) = uu(t) + f F(t - s, s, u(s))ds , the proof of which will appear in the appendix of the paper by T. Kano and T. Nishida, Sur les ondes surfaces de l'eau avec une justification mathematique
de l'equation de l'eau peu profonde, J. Math. Kyoto Univ., 1978. References
[ 11
[2]
L. Nirenberg, An abstract form of the nonlinear Cauchy-Kowalewski theorem, J. Differential Geometry 6 (1972) 561-576. L. V. Ovsjannikov, A nonlinear Cauchy problem in a scale of Banach spaces, Dokl. Akad. Nauk SSSR, 200 (1971) ; Soviet Math. Dokl. 12 (1971) 14971502.
KYOTO UNIVERSITY
Recommend Documents
Sign In