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!
has zero normal component on the boundary JB" _ 1 ; and v is called a boundary conformal Killing vector field on the manifold M". Thus we obtain > A AI>,).H-{p,) /(P A 7W = tLwdx™ , > * . 0, /ds)(a) ^ 0. Then 0 = (s — a)^ where •f is a C°° function and i/r(a) ^ 0. Consequently k — \s — a\ \\ty\\ is not differentiable at s — a. On the other hand if (d(fi/ds)(a) = 0, then 0 = (s — a)2X where X is a C°° function. Consequently k = (s — a) 2 \\X\\ has a zero derivative at s — a. We will show that the geodesic curvature can be extended by using the function 6 defined on [0, L] by sin<9 = dr/ds , is denned by y>u= {hljt,u, — hz^u), = g(u, v) = g(Ju, Jv) —
VECTOR FIELDS AND INFINITESIMAL TRANSFORMATIONS.
427
5.4 N. — On a compact orientable Riemannian manifold Mn of constant nonpositive scalar curvature R with boundary B"-1, an infinitesimal boundary conformal motion leaving the boundary Bn~l invariant is a motion. THEOREM
COROLLARY 5.4 N. — If a compact orientable Riemannian manifold M" of constant scalar curvature R with boundary B^1 admits an infinitesimal nonhomothetic boundary conformal motion leaving the boundary B"-1 invariant, then R > o.
For the case of empty boundary B"--1, theorem 5.4 N and corollary 5 . 4 N are due to YANO [13]. Now let v be a conformal Killing vector field on an Einstein manifold M". Making use of equations (5.17), (5.18), (5.1), (2.15) and the fact that R is constant, we then obtain (5.21)
VjVt^^l^gij,
and therefore, in consequence of the first equation of (2.25), (5.22)
V i O / + Vyi=2A
where (5.23)
*A =
R
n(n—-1)
From equations (2.21), (5.22) follows immediately (5.24)
V,Wj + VjW,= o,
so that Wt is a Killing vector field on the manifold M", where we have placed (5.2 5)
w, = Vi—3»,/A.
If vt is a boundary conformal Killing vector field with zero normal component on the boundary Bn~\ then <&t is also due to equations (5.22), (5.23). By applying corollary 5.4 N we thus obtain THEOREM 5.5 N. — If a compact orientable Einstein manifold M" with boundary Bn~l and R>o admits an infinitesimal nonhomothetic boundary conformal motion leaving the boundary B"-1 invariant, generated by a vector field v with zero normal component on the boundary Bn~*, then the vector field v can be decomposed into
(5.26)
vt = w' + $tllt
where 1 = — R/n (n — 1) < o , w' is a Killing vector field with zero normal component on the boundary Bn~l, and
28
428
C.-C. HSIUNG.
For the case of empty boundary B"-1, theorem 5.5 N was obtained by LICHNEROWICZ ([7] or [8], p. i36) by using de Rham's decomposition of a vector field on a compact orientable manifold. Let [u, u*] be the Lie product of two vector fields u and u* on a Riemannian manifold M", so that (5.27)
[u, u*] = uu* — u*u.
Since in terms of the local coordinates x\ ..., x" we can express u and u* as (5-8)
»= »'^'
"* = " * ' ^ '
from equation (5.27) it follows that r
.1
• <* (
*/
_ / . du*' ~ \ dx'
d
\
n
•
d
(
,
d
du^ \ d dx' J dx>
Thus the contravariant and covariant components of the vector [u, u*] are given by (5.29) [ u, u*]> = ul V, u*'— un V, W, (5. 3o) [ u, u*]; = w Viu* — u*1 ViUj. From equations (2.4), (5.29) it follows that (5.3i)
[u, u*]J = Luu*J.
If u is a conformal Killing vector field satisfying equation (2.15), by equations (2.4), (5.3o) we obtain (5.32)
[u, u*]j = Luu*— 2*u}-
Now suppose that on an Einstein manifold M" with boundary B"-1 there exist two infinitesimal nonhomothetic boundary conformal motions leaving the boundary B"-1 invariant and generated by two vector fields v and v* respectively. Then we have equations (2.i5), (5.21), (5.22), (5.24), (5.26) and similar equations for the vector field v\ which will be denoted by the same numbers with a star. By means of equations (2.8), (5.32), (2.5), (2.i5)*, (2.24) we obtain (5.33)
L[ „, „,, gu = V, (L,v)— 2 0» v)) + V, (Lvv] — 2 0> v\)
On the other hand, from equations (2.8), (5.24), (5—4)* it follows that (5.34)
Llv g = o,
Lw* g = o,
VECTOR FIELDS AND INFINITESIMAL TRANSFORMATIONS.
429
and therefore (5.35)
[L„,, Lu,*] g = (LwLw*—Lif,*Llv)g
= o.
Since w is a Killing vector field, we have, in consequence of equations ( 2 . n ) , (2.12) for w, L„Tl/t = o,
(5.36)
which and equations (5.32) for O = o, (2.8), (2.5), (5.24)* lead immediately to (5.3 7 )
Llw,wt]gl/
= Vi(Lu,wi})+Vj{Llvwti) = Llv(V,w*- + Vy-i»*)+ 2 w*sLwTi) = o.
From equations (2.4), (2.7), (5.32) andV k ®i= V A , it follows that (5.38)
V, (Llv * ) = V( (i»/ <&,) = L w *i = [ IB, V * ],.
Similarly, making use of equations (2.5), (5.36), (5.21), (5. 3o), (5.21)* we can easily obtain (5.3 9 ) (5.40) (5.40
V;(LH,*1) = L w (V / *i) = L w (A* f l r i ;)= l(L„$)g,j, [*,**]*=/(****—<»;*), L ^ . s ^ ^ V , - ^ , <»*],•-r-V/[«>, $*],•= o.
Now we observe that if v and v* are two vector manifold Mn with zero normal components on then by using local boundary geodesic coordinates (2.15), (2.15)*, (3.5) we can easily see that on (5.42)
fields on the Einstein the boundary B"-1, and equations (5.3o), the boundary B n _ 1 ,
[v, »*], = o.
Similarly, if v and v* are two boundary conformal Killing vector fields on the Einstein manifold Mn with zero normal components on the boundary Bn~\ then by noticing the relations Vt
V.CP**!—»!**)= o.
Furthermore, equations (5.38), (5.39), (5.21) imply that [w, V*Sf] can be considered asV4>, where <]> is related to a nonhomothetic boundary conformal Killing vector field v on the manifold M" by the equation L„g = 1
C.-C. HSIUNG.
43o
boundary conformal motions on the manifold Mn leaving the boundary B"-1 invariant form a Lie algebra L, which can be decomposed into the direct sum (5.44)
L = L, + L,
with the relations (5.45)
[LLL^CLL
[Li.L.lcL,,
(5.46)
dim L, ^ dim L 2 — i ,
[L„L,]cL„
d i m L 1 ^ - ( d i m L — i),
where L is the subalgebra of L defined by the infinitesimal motions on the manifold M" leaving the boundary B"-1 invariant, and L,_ the vector space o/"V$, <S> being given in equation (2.i5) defining all the nonhomothetic infinitesimal boundary conformal motions on the manifold Mn. Equations (5.46) are obtained from the fact that if w0, «i, . . . , &v form a basis of L2, then the q elements [w0, w,] (i = i, . . . , q) of Lt are linearly independent. For the case of empty boundary Bnl, theorem 5.6 N is due to LICHNEROWICZ ([7] or [8], p. i38), and was proved again by YANO [13] by a different method, which is extended in this paper. 6. Projective Killing vector fields and infinitesimal projective motions. On a Riemannian manifold M" with boundary Bn~l let a vector field v generate an infinitesimal projective motion so that equation (2. i4) holds. Substituting equation (2.i4) in equation (2.io), contracting with respect to i and I, changing k to i and noticing that (6.i)
ViPi=VjPt,
we can easily obtain (6.2)
LvRtJ = (i — n)VJp„
which and equations (2.5), (2. i o) yield immediately (6.3)
Lv(ykRtj)
= (i — n) V* V , p — a pkRij —
piRjk—pjRu.
If Mn is an Einstein manifold, then V,uR;y = o and equation (6.3) can be reduced by equation (5.i) to (6.4)
(i — n) V* V, pi =c(ipk
gtJ + p, gJk + pj gkl),
where c is a constant defined by (6.5)
c = Rjn.
VECTOR FIELDS AND INFINITESIMAL TRANSFORMATIONS.
43l
From equation (6./+) and the Ricci identity (I.24) for ph it follows (6.6)
— pi IV,jt = Y~
(P* 9tj—Pi 9a)-
by multiplying equation (6.4) by g1' we obtain
Since pk=Vkp, (6.7)
V*[(i —n)V'V»p — a ( n + i)cp] = o,
which implies (6.8)
(i —n)V'V|(p — p 0 ) = a ( n + i)c(p —po),
where p0 is constant.
If
> o or R < o and in local boundary 1—n
geodesic coordinates V t p = o on the boundary B"-1, then by lemma 3 . 1 , p — p 0 = o or p , = o on the manifold Mn. Thus from equation (2.i4) we have L„T'/k=o, that is, the infinitesimal projective motion generated by the vector field v is an infinitesimal affine collineation, which is a motion by theorem 4.2 N if on the boundary Bn~l the vector field v has zero normal component and satisfies equation (4.3). An application of corollary 5.1 thus gives 6.1 N. — On a compact orientable Einstein manifold Mn with a totally geodesic or concave boundary Bn~l and negative constant scalar curvature R, there exists no infinitesimal projective motion, other than the identity, leaving the boundary B"-' invariant such that on the boundary Bn~l its generating vector field v has zero normal component and satisfiesVip — o and equation (4.3) in local boundary geodesic coordinates. THEOREM
For the case of empty boundary B"-1, theorem 6.1 N is due to and
NAGANO
YANO
[11].
Now let us consider a projective Killing vector field v on a Riemannian manifold M" with boundary J3"- 1 . Then by subtracting equation (3.14) multiplied by 2/(72 + 1 ) from equation (3.17) and making use of equation (2.21) we can immediately obtain (6.9)
—(Qp, „ ) + £ = £ (3„, 3») +(dp, cto) = f *)
[ vl(V, vt— V,i>,) + £ = i Pi V,v>] dA„_,. fin—1
"
"T* *•
Thus, from equations (6.9), (I.20), (2.21), (4.i), and theorem 4.1 T, follows THEOREM 6.2 T. — On a compact orientable Riemannian manifold M" with boundary Bn~\ if a projective Killing vector field v satisfies Rt P V ^ o,
432
C.-C. HSIUNG. n
and on the boundary B ~' has zero tangential component and satisfies § v = o, then it must satisfy RtjV'v> = o, and is a parallel vector field, that is, Vv = o on the manifold M", for a totally geodesic boundary Bn-\ In particular, on the manifold M" if the Ricci curvature Rt/v'vj is negative definite definite everywhere, then there exists no such nonzero projective Killing vector field v. Similarly, we have, in consequence of theorem 4.1 N, THEOREM 6.2 N. — On a compact orientable Riemannian manifold M" with boundary B"~l, if a projective Killing vector field v with zero normal component on the boundary -B" -1 satisfies RtjVivJ'^o and n
(6.io)
^vi(V1Di—Vtv1)=o
on B"-1
in local boundary geodesic coordinates, then it must satisfy RijV'vi = o, and is a parallel vector field for a totally geodesic boundary B n _ 1 . In particular, on the manifold Mn if the Ricci curvature Ri/v'v' is negative definite everywhere, then there exists no such nonzero projective Killing vector field v. For the case of empty boundary B"-1, theorems 6.2 T and 6.2 N are due to COUTY [2]. For the corresponding theorems on the nonexistence of a Killing vector field on a Riemannian manifold Mn with boundary B"-1, see [6]. Now let v be a projective Killing vector field on an Einstein manifold Mn. Then, from equations (I.27), (5.i), it follows immediately (6.11)
Qv = ^ v .
By means of equations (2.21), (I.20), (6. n ) we thus obtain v(6.12)
'
n v = -^8dv+ £~l\div. 2 it iR{n + 1 )
On t h e other hand, in consequence of equations (I.16), (2.20), and (6.2), (5.i) we have, respectively, (6.13) (6.14)
(d§v)i=—(n Lvgi/
=
+ i)pi, -2£jp±Vjpt.
From equations (6.12), (6.13), (6.14) it is readily seen that (6.i5)
U(8dv)=o,
VECTOR FIELDS AND INFINITESIMAL TRANSFORMATIONS.
433
so that 8dv is a Killing vector field. Since a Killing vector field is a special projective Killing vector field, equation (6.12) implies that d§v is a projective Killing vector field. For later use we need the formula (6.16)
Lv„T'/k=-n(*^_i)(p/di
+ pk3t/),
which can easily be obtained from equations (2.9), (6.4), (6.6), and also shows by definition that ddv is a projective Killing vector field. An application of theorems 6.1 N and 6.2 T together with a use of equations (6.1), (6.14), (6.16) thus gives 6.3. — If a compact orientable Einstein manifold Mn with boundary B' - and positive constant scalar curvature R admits a projective Killing vector field v such that on the boundary Bn~l it satisfies one of the following two sets of conditions in local boundary geodesic coordinates : THEOREM
1 1
n
(6.17)
Pi = o,
(6.18)
vt= o,
Vip = o, V,p = o,
^ (Vi w,-+ v > 0 = o, <5u = o
(J^0>
then the vector field v has a decomposition given by equation (6.12), where 8dv is a Killing vector field, §dv is an exact projective Killing vector field, and on the boundary JE?"-1 they both satisfy the conditions (6.17) or (6.18) at the same time as the vector field v. For the case of empty boundary Bn 1, theorem 6.3 is due to and
YANO
NAGANO [11]. REFERENCES.
[1] CABTAN (H.). — Notions d'algebre differentielle; application aux groupes de Lie et a u x varietes ou opere un groupe de Lie, Collogue de Topologie [ig5o, Bruxelles], p . 15-27. — Liege, G. Thone; Paris, Masson, ig5i. [2] COUTY (R.). — Transformations inflnitesimales projectives, C. R. Acad. Sc., Paris, t. 247, ig58, p . 804-806. [3] EISENHART (L. P.). — Riemannian geometry. — Princeton, Princeton University Press, ig49[4] HSIUNG (C.-C). — Curvature and Betti numbers of compact Riemannian manifolds with boundary, Rend. Sent. Mat., Univ. e Politec. Torino, t. 17, ig57-ig58, p . g5-i3i. [5] HSIUNG (C.-C). — A note of correction, Rend. Sem. Mat., Univ. e Politec. Torino, t. 2 1 , i g 6 i - i g 6 2 , p . 127-129.
[6] HSIUNG (C.-C). — Curvature and homology of Riemannian manifolds with boundary, Math. Z., t . 82, io63, p . 67-81. [7] LIOHNEROWICZ (A.). — Transformations inflnitesimales conformes de certaines varietes riemanniennes compactes, C. R. Acad. Sc. Paris, t. 241, ig55, p.
726-72g.
232 434
C.-C. HSIUNG.
[8] LICHNBROWICZ (A.)- — Giomitrie des groupes de transformations. — Paris, Dunod, ig58 (Travaux et Recherches mathGmatiques, 3). [9] de RHAM (G.) and KODAIRA (K.). — Harmonic integrals. — Princeton, Institute for Advanced Study, ig5o (Mimeographed notes). [10] SATO (L.). — On conformal Killing tensor fields, Bull. Yamagata University (Nat. Sc.), t. 3, ig56, p. 175-180. [11] YANO (K.) and NAGANO (T.). — Some theorems on projective and conformal transformations, Koninkl. Nederl. Akad. Wetensch., Proc, Serie A, t. 60 (Indag. Math., t. 19), 1957, p. 451-458. [12] YANO (K.). — The theory of Lie derivatives and Us applications. — Amsterdam, North Holland publishing Company, 1957 (Bibliotheca mathematica, 3). [13] YANO (K.). — Some integral formulas and their applications, Michigan Math. J., t. 5, 1 g58, p. 63-73. [14] YANO (K.). — Harmonic and Killing vector fields in compact orientable Riemannian spaces with boundary, Annals of Math., t. 69, 1959, p. 588-596. (Manuscrit recu le 4 novembre 1963.) C.-C.
HSIUNG,
Department of Mathematics, Lehigh University, Bethlehem, Penn. (iStats-Unis).
ACCADEMIA NAZIONALE DEI LINCEI Estratto dai Rendiconti della Classe di Scienze fisiche, matematiche e naturali Serie VIII, vol. XXXVII, fasc. 5. - Novembre 1964
G e o m e t r i a . — On the congruence CHUAN-CHIH
HSIUNG,
<
of hyper surfaces.
)
p r e s e n t a t a ** d a l S o c i o
B.
N o t a ( , ) di
SEGRE.
INTRODUCTION.—The purpose of this paper is to continue our former work [1] to derive a new integral formula for a pair of immersed compact hypersurfaces x (M) , x* (M) in a Euclidean space under a volume-preserving diffeomorphism. B y using this integral formula further conditions are found for x (M) , x* (M) to have the same second fundamental form, and a combination of this result with a former one of ours [1] thus gives a new congruence theorem on immersed compact hypersurfaces in a Euclidean space. In order to simplify the presentation of our work, as in [1] we shall consider tensor products of multivectors and exterior differential forms. Differentiation of multivectors will be taken in the sense of equation (2.3), and differentiation of exterior differential forms will be exterior differentiation; multiplication of matrices will be by the usual row-by-column law. Throughout this paper the range of all Latin indices is always from 1 to n inclusive. 1. LEMMAS.—Let V be a real vector space of dimension n (2i 2), and G and H two bilinear real-valued functions over V x V so that G and H are completely determined by the values gik = G (e{ , ek) and hik = H (e{ , ek), 1 2= 1', k sS n, where e\ , • • • e„ , form a basis of the space V. Under a change of basis (1.1)
e, - > e, = t{ek ,
the matrices || gik || and J hik \ are changed to T || gik j|'T and T || hik J'T respectively, where the repeated index k indicates the summation over its range, T = | / , j | , and T is the transpose of T. Consider the determinant (1.2)
det (ga + Uik) = det {gik) + nk? (gti , hik) + . • • + X" det (hit) .
Since det (git + ~Khik) will be multiplied by (det T ) 2 under the change (1.1) of basis, the ratio of any two coefficients in the polynomial in X on the right side of equation (1.2) is independent of the choice of the basis e1- e„. In particular, if G is nonsingular, t h a t is, if det (git) 4= o, the quotient 0-3)
HG=P(^,^)/det(^„)
depends only on G and H . = o for i=%=k), then
For example, if glt = S<4 ( = 1 for i = k ;
HG=2^«,7 i
W
-
(*) Supported in part by NSF grant GP-1567. (**) Nella seduta del 14 novembre 1964.
Chuan-Chih
Hsiung.
#
259
CHUAN-CHIH HsiUNG, On the congruence of hyfiersurfaces
Since the construction of H G is linear in H , it can be generalized to a bilinear function H over V x V with values in a vector space W, and HG is then an element in W. H G can be called the contraction of H relative to G, or, in the language of tensors, is a vector in W constructed from a covariant tensor H of order two with values in W relative to a nonsingular covariant tensor G of order two. The following two lemmas will be needed in t h e proof of our main theorem. LEMMA I.I.—Let~V be a real vector space of dimension n ( ^ 2 ) , and G and H symmetric positive definite bilinear real-valued functions over V X V completely determined by the values gik = G (e{ , e^ and hik = H (c,- , ek), 1 ^ i , k :£ n, where ex , • • •, c„ form a basis of the space V. Denote (1.4)
g=tet(gik)
,
h = det(hii).
Then (1-5)
H G S (%)'/-,
where the equality holds when and only when hik = pgik for a certain p. This lemma is .due to Girding [2]. LEMMA 1.2.—Let / 1 ( ••-,/„ be n mutually orthogonal unit vectors, and 5 i >' • • 1 £ » i rf 1' ' ' i "<]"• > z = l >' ' ' > n> Hnear differential forms at a point p of a Riemannian manifold M of dimension n (S; 2) imbedded in a Euclidean space E"+OT of dimension n -\- m for m ^ 1. Denote /l
/=
(1.6)
/„ w =
Then at the point
;/
,
S = |5 7) =
u
||Y)*
1 2= i , k:
p of the manifold M rifw--1
(1.7)
.
— (» — I ) / £ 7 ) / M > " - 2 = O.
This lemma is due to Chern and Hsiung; for its proof see [1, p . 283].
2.
HYPERSURFACES IN EUCLIDEAN SPACE.—Let
M be a C2-Rieman-
nian manifold of dimension n ( 2: 2), and consider an immersion x: M -> E" + , that is, a C 2 -mapping x of M into a Euclidean space E " + 1 of dimension n + 1, such that the induced linear mapping x^ on the tangent spaces of M is univalent everywhere. Then x (p), p e M , is a vector in E" + and will be called the position vector of the hypersurface x ( M ) . Let el>- • -,e„ be mutually orthogonal unit vectors in the tangent space of M a t a point p so that they form a frame. Since the mapping x„ is univalent, we identify et with **(e,-). Let co1 • • • co" be t h e coframe dual to e x • • • e„ so that the volume element of M is (2.1)
dV = C O ! A - - - A<°"-
200
Lincei - Rend. Sc. fis. mat. e nat. - Vol. XXXVII - novembre 1964
Let e„ + 1 be the unit normal vector of x (M) at x (p), and introduce the matrices (2.2)
CO = = J C O 1 , • • • , C O * [I
Then we have the matrix equations (2.3)
dx = coe
,
de = Q.e + 8c«+i,
where Q
(2-4)
I t o * I! 1
1 f^ i , k fin ,
1 [i
coj ,
do> = co A £2
,
(2.6)
a A 0=o, ,
dQ=
Q
AQ
,
where O. is denned by (2-7)
£2 = I o>i +1 , • • •, o v u || ,
(2.8)
^e„ + 1 = Oe .
It is well-known that the matrix D. gives the connection form of the induced Riemannian metric by x, and that the second fundamental form II (p) of the hypersurface x (M) at the point x (p) is the negative of the scalar product (dx , de„+{) of the two vectors dx and den+1 in the Euclidean space r»+l From equations (2.2), (2.3), (2.7), (2.8) it follows immediately that (dx , de„+{) = 2* co' co'
(2-9)
+i
where the multiplication of linear differential forms is commutative in the ordinary sense. In the explicit form the second equation of (2.5) can be written as (2.10)
co''A""+1= O ,
which enables us to put (2.11)
co"+!= b ,co>
where btJ- is symmetric in i , j , that is, (2.12)
K- = *u-
Since (e,- , e„ + i) = o , 1 ^ 1 ^ n, by exterior (2.3), (2.4), (2.7), (2.8) we can easily have (2-13)
Chuan-Chih Hsiung.
«i + i = - « ? + 1 .
differentiation
and equations
CHUAN-CHIH HSIUNG, On the congruence ofhypersurfaces
261
Substituting equations (2.11), (2.13) in equation (2.9) we thus reduce the second fundamental form of the hypersurface x (M) a t the point x (p) to the form (2.14)
II = buto*t*>.
On the other hand, using equations (2.3), (2.11), (2.14) we obtain the second differential of x in the ordinary sense: (2.15)
d%x = (du' + to* coj) e. + II en+1,
from which it follows that (2.16)
( e » + i / x ) = II .
Bu putting O - c »+0 =
(2.17)
y»+i,
we have the one-rowed matrix (2.18)
y„+1'Q
=
\\y„+1otJ-o>J\\.
Let xl(p) be the orthogonal projection of the position vector x(p) onto the direction along the normal vector e„+1 of ac(M) at the point x(p). Then we have (2.19)
x>- = y „ + 1 en+1 ,
and therefore the following quadratic differential form: (2.20)
Q = (a:1 , d2 x) = y„+1 II.
3. INTEGRAL FORMULAS.—Consider two immersions x,x* of a C 2 -Riemannian manifold M of dimension n ( 3 : 2) into a Euclidean space E" and a diffeomorphism,/ as given by the commutative diagram M——>x(M)
X
C E"+1
V
x*\ \
* **(M)C E"+I.
Then § 2 can be applied to the hypersurface * ( M ) , and for the corresponding quantities and equations for the hypersurface x*(M) we shall use the same symbols and numbers with a star respectively. Suppose t h a t / i s volume-preserving, that is, by definition it maps the volume element of one immersed hypersurface into t h a t of the other. As a consequence of this definition / exists only if M is oriented, and / is then orientation-preserving. Now over the abstract manifold M there are two induced Riemannian metrics with the same volume element, namely, (dx (fi) , dx (p)) and (dx*(p) , dx*(p)) = (d(Jox) (p) , d (fox) (p)) .
202
Lincei - Rend. Sc. fis. mat. e nat. - Vol. XXXVII - novembre 1964
Thus the notion of frames e1- • • e„ having measure I and an orientation coherent with that of M has a sense in both metrics. A t a point peM. any such frame can be obtained from a fixed one by a linear transformation of determinant 1. The condition for the frames el • • • en to be of measure I is (3-0
( c l A • • • A e„ , «i A • • • A e„) = 1 •
Differentiating equation (3.1) and using the second equation of (2.3) we can easily obtain (3.2)
Tr fl = £ to* = o .
For an (m X «)-matrix a = | a0-1 and an (n X ^>)-matrix 6 = | bik |J, whose elements are vectors in the space E" + 1 , we shall use the notation (a, b) to denote the matrix of real numbers given by
(3-3)
(«,*) = || £ («,,., 6 y ,)|.
In order to derive a new integral formula for two compact hypersurfaces ac(M),x*(M) under a volume-preserving diffeomorphism/, we have to construct some exterior differential forms globally defined over the manifold M . For this purpose we introduce the following matrices: (3-4) (3-5)
G = ( C , ' C ) = 'G = | M > A = | | M l , . . . , a ) „ | = coG,
(3.6)
v = '0e
(3-7)
Y=(x,'C) = |y1,..-,j„||,
(3.8)
r = Ye ,
(3.9)
y = ri>*»-i = <]*! A - - - A e „ .
,
» * = '0*C)
Since <\i is an exterior differential form of degree n — 1 globally defined over M, for a compact manifold M Stokes's theorem gives (3-io)
di> =
o.
M
Making use of equations (3.7), (2.3), (3.4), (3.5), (2.17) we obtain (3.11)
dY = (dx,'e)
+ (x,d'e)
= A + Y ' Q + y„+1'Q ,
from which and equation (3.8) it follows that (3.12)
dr = [A + Y (Q + '£}) + y„+l '6] e + Y8 e„+l .
By equations (3.6), (2.3) and the second equation of (2.6)* a similar calculation gives (3-13)
dv* = — '0* ('Q* + Q) e — '0* 0 e„+1 .
CHUAN-CHIH HSIUNG, On the congruence of hypersurfaces
263
Using equations (3.11), (3.12) and noting that (3.14)
d («*»-») = (» — 1) (dv*) v*»-z,
from equation (3.9) we have (3.15)
d*r=[A+Y('£l+Q)
+ yH+1'Q]ev*"-i 1
+ Y6e„ + 1 v*"-
— (n—i)r
—
(n—i)r'Q*(
'0* 6e„ +1 a*"" 2 .
Since d (e1 A • • • A «„) contains no term in c x A • • • A e„ in consequence of equation (3.2), equating the terms in ex[\- • • i\en on both sides of equation (3.15) we thus obtain (3.16)
(flty) C l A • • • A eH = (A + y„+1 <6) ev*'-l
+
+ Y [('Q. + D) c«* — (n — 1) c '6* f'Q* + Q) c] i>*»-2 . By putting / = e , \ = '6*, 7) = Q and r, = ' i i * respectively in Lemma 1.2 of of § I we can reduce equation (3.16) to (3.17)
(d<\>) ei A • • • Ae„ = [A + JVU '6 + Y ('£2 — 'O*)] w » - » .
Thus the integral formula (3.10) implies that for a pair of compact immersed hypersurfaces x ( M ) , **(M) in the space E" + under a volume-preserving diffeomorphism / , the integral, over the manifold M, of the coefficient of «iA • • • A«„ on the right side of equation (3.17) is zero. 4. THEOREMS.—Let M and M* be two ^-dimensional (n ^ 2) C 2 -Riemannian manifolds with fundamental tensors G and G* respectively, and f: M ->• M* a C 2 -mapping. Then on M there are two connections: the Levi-Civita connection defined by its Riemannian metric and the connection induced by the mapping / from the Levi-Civita connection of M*. The difference of these two connections is a tensor field A of contravariant order I and covariant order 2, and the construction in § 1 gives a vector field A G . / i s called an isometry, if G * = G , and an almost isometry relative to G if AG = o. It is obvious t h a t an isometry is also an almost isometry, since in this case A = o. THEOREM 4.1.—Let M be a (Z%-Riemannian manifold of dimension n ( 5 : 2), x , x* : M - > E " + two immersed compact hypersurfaces, in a Euclidean space E" of dimension n + I, with fundamental tensors G , G* and positive definite second fundamental forms II, II*, whose coefficient tensors being denoted by B, B*, respectively, and f: x (M) -> x* (M) a volume-preserving diffeomorphism. Then II = 1 1 * , if (4.1)
det B = det B*
(4.2)
GB* i£ GB ,
and f is an almost isometry relative to B*, that is, (4.3)
AB* = o .
264
Lincei - Rend. Sc. fis. mat. e nat. - Vol. XXXVII - novembre 1964
P r o o f .—In order to prove this theorem we have to find d<\i, which is the coefficient of ex A • • • A «„ on the right side of equation (3.17). For this purpose we observe that each term on the right side of equation (3.17) is of the same type as the form Tzev*"~1, where (44)
7C = ||7r ll - • -,7*J
is a one-rowed matrix of linear differential forms. By using equations (2.2), (2.4)*, (3.6) we have (4.5)
*ev —
1
= (2«,.,)(5cor+1-y)-1 =
= (X s, r .. ,-n 71^ A w.*"+1 A • • • A w*„"+1) cx A • • • A e„, where e,- ...,• is equal to + 1 or —I according as i\,---,i„ form an even or odd permutation of 1 , • • •, n, and is equal to zero otherwise, and the summation is extended over all 1\, • • •, z„ from 1 to n. Assuming (4.6)
71. = kiy to' ,
(4-7)
H = hi, ,
i J = 1 ,• •
-,n,
and using equations (2.11)*, (2.1), (1.2), (1.3), (2.14)* and elementary properties of determinants, from equation (4.5) we can easily obtain (4-8)
7iet>*«-i = n\ H B *(det B*) (e^--
• he„) dV.
By putting 7 t = A and n=y„+1'B in equation (4.8), and recalling 0 0 , = ^ u>' and equation (2.14), we therefore have (4-9)
Acu*»-i = n\ G B *(det B*) («! A • • • A e„)dV,
(4.10)
y„+1'Qev*»-1
= n\ y„+1BB*(det
B*) (^A • • •
f\e„)dV.
Since w = co*, from the first equations of (2.5), (2.5)* we have (4-i 1)
w A ( Q — Q*) = o ,
so that we can write (4.12)
=
a*. 00
a*. c o ' A c o '
— 0
C O * --C0** 1 t
Equations (4- i 0 , (4- 12) imply (4-i3)
'j
which gives the symmetry of d)j in the subscripts i ,j, (4-I4)
that is,
h
a*..= a . .
From the properties of the forms Q , O* and the definition of the tensor A it follows that the components of A are akiy. For each fixed k denote (4-15)
A*=4.
i,j
=
i,...,n.
CHUAN-CHIH HSIUNG, On the congruence of hypersurfaces
265
On the other hand, a use of equations (3.5), (3.5)*, (3-7), (4-12) yields readily the matrix (4-16)
Y (*2 k
By putting H = y^A (4.17)
« * ) = \yk a{j,...,
J\.
in equation (4.8) we obtain
y (;Q _
=
n
! -^ AB* (det B*) (ex A • • • A «„) a'V,
since AB* = o due to condition (4.3). reduce equation (3.17) to (4.18)
yk 4
Thus equations (4.17), (4-6), (4.7)
di/ = n\ (GB* + V„+i BB*) (det B*) dV.
Hence for a pair of compact immersed submanifolds x (M) , x*(M) under a volume-preserving diffeomorphism / with AB* = o the integral formula (3.10) becomes (4-19)
/ (GB* + y„+i BB*) (det B*) dV = o . Si
In particular, when the two hypersurfaces x (M) , ac*(M) are identical, by definition B B * = 1 and the formula (4.19) is reduced to (4-20)
/ (GB + v„+1) (det B) dV = o . M
Subtracting equation (4.20) from equation (4.19) and noticing condition (4.1) we obtain r
(4-2i)
/ [(GB* — GB) + yH+1 (B B * — 1)] (det B) dV = o . M
Since by the assumption of the theorem the compact hypersurface ac(M) has positive definite second fundamental form II, we can choose the common origin of the position vectors x (p) for all points / € M in the Euclidean space E" + I to be in ac(M) so t h a t y „ + 1 > o . Moreover, by Lemma 1.1 of § 1 we have (4.22)
BB* —
1^0,
where the equality holds when and only when, for a certain p , B = pB*, from which it follows that II = II* due to condition (4.1). Thus by condition (4.2) the integrand on the left side of equation (4.21) is nonnegative, and the validity of equation (4.21) gives immediately that the equality in (4.22) holds. Hence the theorem is proved. By combining Theorem 4.1 and a former one of ours [1] and noticing equation (2.20) we are readily led to
266
Lincei - Rend. Sc. fis. mat. e nat. - Vol. XXXVII - novembre 1964 T H E O R E M 4 . 2 . — U n d e r the same
the diffeomorphism
f is a rigid
assumptions
motion,
as those
if the following
in
further
Theorem
4.1,
conditions
are
satisfied: (4-23) and f (4.24)
BG. ^ BG , is an almost
isometry
relative
to G*, that
is,
AG* = o .
BIBLIOGRAPHY.
[1] S. S. CHERN and C. C. HsiUNG, On the isometry of compact submanifolds in Euclidean space, «Math. Annalen », 149, 278-285 (1963). 0 [2] L. GARDING, An inequality for hyperbolic polynomials, « J. Math. Mech. », 8, 957-965 (1959)-
R O M A , 1965 — Dott. G. Bardi, Tipografo dell'Accademia Nazionale dei Lincei
R e p r i n t e d from t h e P H O C E E D I N O S O F T H E N A T I O N A L A C A D E M Y O F S C I E N C E S
Vol. 54, N o . 6, p p . 1509-1513.
D e c e m b e r , 1965.
ON THE GROUP OF CONFORMAL TRANSFORMATIONS COMPACT RIEMANNIAN MANIFOLD*
OF A
BY CHUAN-CHIH HSIUNG DEPARTMENT OF MATHEMATICS, LEHIGH UNIVERSITY, B E T H L E H E M , PENNSYLVANIA
Communicated by S. S. Chern, October 11, 1965 1. Introduction.—It is known that Riemannian manifolds of constant scalar curvature R are important in the study of conformal transformations of compact Riemannian manifolds. In this direction a basic result is due to Yamabe, 1 namely, every Riemannian metric on a compact manifold Mn of dimension n > 2 can be deformed conformally to a Riemannian metric of constant R. Lichnerowicz2 showed that the largest connected group C0(M") of conformal transformations of a compact Riemannian manifold Mn with nonpositive constant R coincides with the largest connected group Io(Mn) of isometries of M", and Yano and Nagano3 jointly showed that if C0(Mn) ^ I0(Mn) for a complete Einstein space Mn (n > 2), then If" is globally isometric to a sphere. Very recently by making use of this joint result of Yano and Nagano, Lichnerowicz4 further obtained. THEOREM. Let M" (n > 2) be a compact Riemannian manifold with positive constant R and R^R^ = const., where i, j = 1, ..., n, and Ri} is the Ricci tensor of Mn. If C0(Mn) ?* I0(Mn), then M" is globally isometric to a sphere. This theorem contains the following two corollaries as special cases. 6 COROLLARY 1 (Goldberg and Kobayashi ). For a compact homogeneous Rien n mannian manifold M (n > 3) if Co(M ) ^ Io(Mn), then Mn is globally isometric to a sphere. 6 n COROLLARY 2 (Ba ). On a compact Riemannian manifold M (n > 2) with finite n Poincare group, if C0(M ) is noncompact and operates transitively, then Mn is globally conformal to a sphere. The purpose of this paper is to extend the above theorem of Lichnerowicz as follows; the method used is the same as that of Lichnerowicz. MAIN THEOREM. Let Rim (i,j,k,l — 1,. .. ,n) be the Riemann tensor of a compact Riemannian manifold Mn (n > 2) with positive constant R, and suppose that pvQi = c = const., T2p LP
(n - l)g] Q
(1)
2»(p + g)fl«'+'-»
=
J
n'+i-^n
-
(2) 1
l)'- '
where p,q are nonnegative integers and not both zero, and P = RimRim, n
Q = RU'RUV.
n
(3)
/ / Co(M ) ?£ Io(M ), then M" is globally isometric to a sphere. It should be noted that when p = 0, q = 1, or p = 1, q - 0, equation (2) is an identity, and for the first special case our main theorem is reduced to the theorem of Lichnerowicz. Furthermore, we still have the open question: When p = q = 0, is our main theorem still true? 2. Notations and Formulas.—Let Mn be a Riemannian manifold of dimension n (>2) and class C°, ||gftf|| the symmetric matrix of the positive definite metric of 1509
1510
MATHEMATICS:
C.-C. HSIUNG
PROC. N. A. S.
Mn, and ||sitf || the inverse matrix of \\gi}\\. Throughout this paper all Latin indices take the values 1 , . . . ,n. We shall follow the usual tensor convention that indices can be raised and lowered by using gij and gi}, respectively, and that repeated indices imply summation. In this section we shall list few known formulas (for the details of their derivations see Lichnerowicz' book,2 pp. 124-134, or the author's paper7), which will be needed in the proof of our main theorem. Let v be a vector field defining an infinitesimal conformal transformation on Mn, and denote by the same symbol v the 1-form corresponding to the vector field v by the duality defined by the metric of M". Then we have 2 Lvgtj == VM + Vfii =
nLJRnm
gtfiv, n = efVFtSv - e/V' p ,8v + g]kVlVhdv - gnVkVh8v,
(4) (5)
where Lv is the operator of the infinitesimal transformation v, V the operator of covariant derivation of M", 8 the operator of coderivation, and ekh = 1 for h — k, and = 0 for h ^ k. From equations (4) and (5) it follows immediately nLJtijx, = —2RijhlSv + guVjV^v - gaV}Vi8v + gnVtfibv — gnVhVfiv,
(6)
nLJtu, = -guvA8v + (n — 2)VvVJv,
(7)
where A = dd + 8d is the Laplace-Beltrami operator on Mn. A necessary and sufficient condition for a vector field v to define an infinitesimal conformal transformation on a compact manifold M" is that it satisfy Av + (l --jd8v
= Qv,
(8)
where Q is the operator of Ricci: Q:
vi-*2Rijv1.
For an infinitesimal transformation v on a manifold Mn, we have A8v = R/(n - l)8v - n/[2(n - 1)]LVR.
(9)
Let £/(,) and rmp) be two tensor fields of the same order p (^w) on a compact orientable manifold Mn, where I(p) denotes an ordered subset {ih. . .,iv) of the set {1,. .. ,n J of positive integers less than or equal to n. Then the local and global scalar products (£,TJ) and (£,r;) of the tensor fields £ and 77 are defined by (lv) = -, ^ ' W ) , (f,u) = f
(Z,v)dAn,
(10) (11)
where dAn is the element of area of the manifold Mn at a point. From equations (10) and (11) it follows that (£,£) is nonnegative, and that (£,£) = 0 implies that % = 0 on the whole manifold M". For any 1-form £ on a compact orientable manifold M" we have
VOL.
54, 1965
MATHEMATICS:
(A{
C.-C.
HSIUNG
+ ( l - ? ) di* - Q£,£ ) ^ 0,
1511
(12)
where the equality holds when and only when £ defines an infinitesimal conformal transformation on M". 3. A Lemma.—The following lemma is due to Lichnerowicz.4 LEMMA 1. Let v be a vectorfielddefining an infinitesimal conformal transformation on a compact orientable Riemannian manifold M" of constant scalar curvature R. Then (tf«V 4 V^ + - ^ - , So) ^ 0. \ n{n - 1) /
(13)
Proof: By applying the integral formula (12) to the 1-form dSv we have l2^1 ~ ^ Adbv - Qd8v, ddv) ^ 0.
(14)
On the other hand, covariant differentiation gives = <( 2(W ~
V* Sv ( ^ — — [Ad&v - QdSv)
1)
Ad5v - Qd5y,d5v)>
-
(15)
From equations (14) and (15) and the well-known Green's theorem that for any vector field £ on a compact orientable Riemannian manifold Mn,
I
V'fcdA,
o,
<M<
we thus obtain (—
AASv - 8Qd8v,8v) ^ 0.
(16)
Since the manifold M" is of constant scalar curvature R, it is known that V4fl« = 0,
(17)
and for the infinitesimal transformation v equation (9) is reduced to A5v =
So. n — 1
(18)
From equations (17) and (18) follow immediately SQdSv = -2Vi(RiiVfiv)
= -2RilViViSv,
(19)
2
R AASv =
— 8v. (20) [n — l ) 2 Substituting equations (19) and (20) in (16), we hence obtain the required inequality (13).
1512
MATHEMATICS:
C.-C. HSIUNG
PKOC. N. A. S.
4. Proof of the Main Theorem.—Without loss of generality we may assume our manifold Mn to be orientable, since otherwise we need only to take an orientable twofold covering space of Mn. On the manifold Mn consider the covariant tensor field T of order 2(2p + q): P
•* MiSiii- • • tpiitiplpUiti
• • • vqVq
=
Q
H r=X
ttirirkrlr
1 1 tCu3r>, s=l
^^7 7Tv n (0jrkr9irir - QhuQu*,) n gw n +'>{n — l)v r = i .=1
(21)
v
From equation (21) it is easily seen t h a t the length of T is [2(2p + g)]!| T| * = [2(2p + g)]!
^
^
_
(22)
which is constant by the assumption of the theorem. Since Co(M") ^ Io(Mn), we may assume v to define an infinitesimal nonisometric conformal transformation on M", so t h a t to j* 0. Then we have L,\ T\2
= 2(LVT,T) + i ^ - ± - D \f\Uv n
= 0.
(23)
From equation (23) it follows immediately t h a t 2 ( 2 y + g)
«L,r,r>, to) = -
(I rI %,«»).
(24)
n
*JtL
On the other hand, from equations (6) and (7) we obtain v 2v " j1j1ic1i1.
. • iPjvKvlvuwi-
• -u„vQ
=
"V
n
p
11
"W*l;
r=\
"•
1 + - 2 [J?wi*iii---R<,.,j,_it,J!I..l(«f4,i,VirVj;rSw w =i + 9h ,V*rVirto - gJTlTVkrVirSv)R \
pP
R",u
a=i gWl.Vj>V(rto 9
ir-HJT+llCr+llr+1' • • Riplpkplp]
q1
II
Ru,v,
8= 1
11 R RiirrjrlCr T).tiiTT 2-i -\— II 2J [Ruivi- • •Ru..it..1[~9u.*A&v n r=i
n
s=i
+ (n - 2)V , 2(2p + g)fl"+'to
*
«
(25)
By means of equations (21), (25), (18), (3), (2), and (22) an elementary calculation yields
«L,r,r),to) = - ^ (|r| 2 to,to) -P^Q*
[~4P , (w - 2 ) g l /
# 2 to
\
VOL.
54, 1965
MATHEMATICS:
C.-C.
HSIUNG
1513
By comparing equations (24) and (26), noticing that bv ^ 0, and making use of Lemma 1, we thus have -* ilj'lfclh- ' • ipjpkplpUlVl-
• -UqVq ~
^'
\" * )
Multiplying equation (27) by ghh
n gtr'rgirtr r=2
R s=l
g".^
and using equation (21) we obtain Rhkl = RghkJn, which implies that M" is an Einstein space. Hence, by the joint theorem of Yano and Nagano mentioned in the Introduction, M" is globally isometric to a sphere, and our main theorem is proved. * Supported by National Science Foundation grant GP-4222. 1 Yamabe, H., "On a deformation of Riemannian structures on compact manifolds,'' Osaka Math. J., 12, 21-37 (1960). 2 Lichnerowicz, A., Giomitrie des Groupes de Transformations (Paris: Dunod, 1958), p. 134. 3 Yano, K., and T. Nagano, "Einstein spaces admitting a one-parameter group of conformal transformations," Ann. Math., 69, 451-461 (1959). 4 Lichnerowicz, A., "Sur les transformations conformes d'une variety riemannienne compacte," Compt. Rend., 259, 697-700 (1964). 6 Goldberg, S. I., and S. Kobayashi, "The conformal transformation group of a compact homogeneous Riemannian manifold," Bull. Amer. Math. Soc, 68, 378-381 (1962). 6 Ba, Boubakar, thesis, Paris, 1964. 7 Hsiung, C. C , "Vector fields and infinitesimal transformations on Riemannian manifolds with boundary," Bull. Soc. Math. France, 92, 411-434 (1964).
STRUCTURES AND OPERATORS ON ALMOST-HERMITIAN MANIFOLDSO) BY
CHUAN-CHIH HSIUNG
Introduction. It is well known (see, for instance, [6, §9.2]) that for a Kahlerian structure the real Laplace-Beltrami operator A commutes with real operators L and A defined in §3, and the complex Laplace-Beltrami operator • is real and equal to \ A. The purpose of this paper is to study the converse of these three properties of Kahlerian structures. §§1 and 2 contain fundamental notations and definitions of various almostHermitian structures, as well as real and complex operators, on an almost-Hermitian manifold. In §3 we shall prove THEOREM 3.1. For an almost-Hermitian structure of dimension n ( ^ 2), if A commutes with the operator Lor A with respect to all forms of any degree p (0 ^ p ^ n — 2), then the structure is Kahlerian.
It is known that the commutativity of A with L or A plays a crucial role in the proof of the Hodge's well-known theorem concerning the relationship between the effective harmonic forms and Betti numbers of compact Kahlerian manifolds. Thus Theorem 3.1 kills a possibility of extending the Hodge's theorem to more general manifolds. It should also be noted that recently Weil [7] used the commutativity of some operators to characterize Chern's generalization of Kahlerian structures. From Weil's result and our Theorem 3.1 here it seems natural to characterize a structure by using the commutativity of some operators. In §4 we study the realization of the complex operator • by establishing the following theorems. THEOREM 4.1. The complex operator • for an almost-Hermitian structure is real with respect to every form of decree 0 if and only if the structure is almostsemi-Kdhlerian. Moreover, with respect to every form of degree 0, if • for an almost-Hermitian structure is real, then • = ^A.
Presented to the Society, January 27, 1965 under the title Characterizations of Kahlerian structures; received by the editors July 19,1963 and, in revised form, June 2, 1965. C1) This work was done at the University of California, Berkeley, California under National Science Foundation Research Grant NSF G-19137, and at Lehigh University, Bethlehem, Pennsylvania under Research Grant AF-AFOSR 62-206 from the Air Force Office of Scientific Research and under National Science Foundation Grant GP-1567.
136
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS
137
THEOREM 4.2. If for an almost-Hermitian structure the relation • = JA holds for all forms of degrees 0 and 1, f/zen ffte structure is Kdhlerian.
Kodaira and Spencer [4] have shown that if the relation D = iA holds for an almost-Hermitian structure, then the structure is integrable. Theorem 4.2 was a conjecture for some time, and was proved very recently by A. W. Adler [1] by a different method under a stronger assumption that the relation • = £A holds for a Hermitian structure and all forms of degrees 0, 1 and 2. For more general case we shall have THEOREM 4.3. For an almost-Hermitian structure, if the complex operator • is real with respect to all forms of degrees 0 and 1, then it is also with respect to all forms of degree 2.
It seems that the conclusion of Theorem 4.3 could be extended to all forms of degree p ( > 2). However, due to the complication of the calculation the author is unable to show it. Throughout this paper, the dimension of a manifold M" is understood to be n ( 2: 2), and all forms and structures are of class at least C 2 . The author is indebted to the referee for his suggestion in regard to the proof of Theorem 3.1 in the present form. 1. Notations and real operators. Let M" be a Riemannian manifold of dimension n ( 2: 2), || gtJ | with gtJ = gJt the matrix of the positive definite metric of the manifold M", and \\g'J\\ the inverse matrix of \\gij\\- Throughout this paper all Latin indices take the values l,---,n unless stated otherwise. We shall follow the usual tensor convention that indices can be raised and lowered by using glJ and gtJ respectively; and that when a Latin letter appears in any term as a subscript and superscript, it is understood that this letter is summed for all the values 1, •••,«. Moreover, if we multiply, for example, the components atj of a tensor of type (0,2) by the components bJk of a tensor of type (2,0), it will always be understood that j is to be summed. Let Jt be the set {1, ••-,«} of positive integers less than or equal to n, and I(p) denote an ordered subset {ii,---,ip} of the set 91 for p ^ n. If the elements iu---,ip are in the natural order, that is, if i t < ••• < ip, then the ordered set I(p) is denoted by I0(p). Furthermore, denote the nondecreasingly ordered p-tuple having the same elements as I(p) by (p)>, and let I(p; s \j) be the ordered set I(p) with the s-th element is replaced by another element j of % which may or may not belong to I(p). We shall use these notations for indices throughout this paper. When more than one set of indices is needed at one time, we may use other capital letters such as J, K, L, ••• in addition to /. At first we define
138
C.-C. HSIUNG
fO,
[March
if<JG>)>#
(1.1) e ^ = -I 0,
if J(p) or K(p) contains repeated integers,
[ + 1 or — 1, if the permutation taking J(p) into K(p) is even or odd. By counting the number of terms- it is easy to verify that (.1.2)
£
s
l...n
I(p)K(n-p)
£
(1-3)
K(p+8) e J(p)
~ P !£ K(»-p)> -
P ! £ K ( p + «)-
On the manifold M", let V denote the covariant derivation with respect to the affine connection T, with components T)k in local coordinates JC 1 ,*".*". of the Riemannian metric g, and let <£ be a differential form of degree p given by (1.4)
4> = j l 4>Hp)dxHp) = KiP)dxIoW
,
where <£/(p) is a skew-symmetric tensor of type (0,p), and we have placed dxUp) = dxh A ••• A dxip.
(1.5) Then we have
# = (#) / 0 ( p + 1 ) dx I ° ( ' + 1 ) ,
(1.6) where (1-7)
(#)/(p + 1 ) = ^
C D V A W
Denote eJ(„) = 8J1(n7(det(gl7))1^.
(1.8)
Then by using orthonormal local coordinates x 1 ,"-,x"andlthe relation (1.2) we can easily obtain •7)
e
l(p)K(n-p)e
— P!£K(n-p).
The dual operator * is defined by (for this see, for instance, [5]) (1-10) where d-11)
* 0 = (*<£)IO( B -P)^ IO( ''" P) >
(^)/(n-P) = ^^(P)r(n-P)^(P).
From equations (1.10), (1.11) it follows that for the scalar 1 (1.12)
*1 = (det (g^dx1
A - A dxn,
1966]
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS
139
which is just the element of area of the manifold M". By using orthonormal local coordinates x1, •••,x" we can easily verify that **> = ( - l)p("-p>^.
(1.13)
Denote the inverse operator of * by * - 1 . Then from equation (1.13) it is seen that on forms of degree p *- 1 =(-iy<"-p>* .
(1.14)
The codifferential operator d is defined by (1.15)
<50 = (-l) p+n+1 *- 1 rf*<£.
Making use of equations (1.6), (1.7), (1.10), (1.11) we obtain immediately (1.16)
^ = (^) / 0 ( p _ 1 ) ^ J "C- 1 ) ,
where (1-17)
W)/(p-i)=- V J ^ V D .
By means of equations (1.13), (1.14) it is easy to verify that (1.18)
A*=*A,
where A is the Laplace-Beltrami operator defined by (1.19)
A = 5d + dd.
For a form
V J V,«£ /(p) + I
(1.20) 1
+
V
i- §l(p;%\a,l\V)R
i.it,
where (1.21)
VJ = gJkVk,
(1.22)
*',u = dr'jjdx'- drjtidxk+ r^r;, - rjiri,
(1.23)
RJk = R°Jks.
2. Complex structures and operators. On a Riemannian manifold M" with metric tensor gtJ, if there exists a tensor FtJ of type (1,1) satisfying (2.1)
F/Fjk=-eki}
then f y is said to define an almost-complex structure on the manifold M", and the manifold M" is called an almost-complex manifold. From equation (2.1) it
140
C.-C. HSIUNG
[March
follows that the almost-complex structure F^ induces an automorphism J of the tangent space of the manifold M" at each point with J 2 = — I, I being the identity operator, such that, for any tangent vector vk, (2.2)
J:vk-*Fikv\
If an almost-complex structure FtJ further satisfies (2.3)
gtjFk FkJ = ga,
then F / i s said to define an almost-Hermitian structure on the manifold Ml and the manifold M" is called an almost-Hermitian manifold. From equations (2.1), (2.3) it follows that the tensor Ffj. of type (0,2) defined by (2.4)
FtJ = gjtF,"
is skew-symmetric. Thus on an almost-Hermitian manifold we have the associated differential form (2.5)
co = Fijdx1 A dx>.
By using the multiplication of matrices, from equation (2.1) we readily see that a necessary condition for the existence of an almost-complex structure on a Riemannian manifold M" is that the dimension n of the manifold M" be even. It should also be remarked that an almost-complex manifold is always orientable, and the orientation depends only on the tensor FtJ. An almost-Hermitian structure FtJ defined on a manifold M" is called an almostKahlerian structure and the manifold M" an almost-Kahlerian manifold, if the associated form co is closed, that is, (2.6)
dco = 0.
From equations (2.5), (2.6) it follows that an almost-Kahlerian structure F( satisfies (2-7)
Fhij EE V»Fy + V ^ + VjFhi = 0.
The tensor FhtJ is obviously skew-symmetric in all indices. An almost-Hermitian structure FtJ (respectively manifold) satisfying (2.8)
F,3-VjF,'=0
is called an almost-semi-Kahlerian structure (respectively manifold). In particular, the structure F/ is Kahlerian if VjF,.* = 0. In this case, by means of equation (2.1) it is easily seen that the torsion tensor ttJk = Fjh(dFtkl8xh - dFhkldx') - Fih(8FJkl8xh -
dFhkl8xJ)
vanishes, so that the integrability condition of the almost-complex structure FtJ
1966]
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS
141
is satisfied. But in general when * y * = 0 , the almost-Hermitian structure FtJ is defined to be Hermitian. Multiplying equation (2.4) by F *' we obtain FtjFhi = - s).
(2.9)
By taking covariant differentiation of both sides of equation (2.9), noticing that F,JVkFtJ
(2.10)
= 0,
and making use of equations (2.7), (2.8) it is easily seen that F„.,F ij = 2F / ,'> i .
(2.11)
Thus an almost-semi-Kahlerian structure F,J satisfies FhiJFiJ = 0.
(2.12)
Multiplication of equation (2.11) by F£ and a use of equation (2.9) give (2.13)
-±-FhiJFiJFkh.
Fk=
From equations (2.7), (2.8), (2.13) we hence reach that an almost-Kdhlerian structure or manifold is also almost-semi-Kahlerian. We now consider an almost-Hermitian manifold M" with an almost-Hermitian structure Ff so that equation (2.3) holds, and shall introduce complex operators (compare [6, Chapter IX]) on the manifold M". At first we define (2.14)
n S = ^(gf
-s/(-r>F/)
Z
1,0
and its conjugate^) tensor (2.15)
n / ' = n , ' = \- (g/ + V( - i ) F / ) . 0,1
Z
1,0
Let p + a = p, p ^ 0, a ^ 0, and set 1 1
I(p)
— e/(p)
l l mi 1,0
J,
(2.16)
n
s
m
l l r> 1,0
' . . . lTT l
n,
0,1
1 ""e ^ e
Ko(p)So(
0,1
Then for a form (f> given by equation (1.4) we have
(2.17)
n«A = (nA 0 ( p ) dx'° ( p ) , P ,ff
\ P ,
I
(2) Throughout this paper a bar over a letter or symbol denotes the conjugate of the complex number or operator defined by the letter or symbol.
253
142
C.-C. HSIUNG
[March
where (2.18)
(n*)
=
n
w
'Xc
I(p)
We next define a complex covariant differentiator (2.19)]
#,= n/Vj, 1,0
and the corresponding contravariant differentiator
(2.20)
2l = giksik= n / v j ' = n ] /V'', 0,1
1,0
where VJ" = ^'V,..
(2.21)
The conjugate operators of @t and 3>l are (2.22)]
9t = n , % o.i
(2.23)
!»' = n / V ' . 1,0
Furthermore, for a form 0 given by equation (1.4) we define the complex analogues of the real operators d and 8:
(#)/(P+D = f 2 \p+a=p
n
p,
/l(p+l)
(2.24)
=
2
n
I(P+1)
^;G(F))
p+tr = p p + 1 ,
W)«P-D = ( £
n s n *)
\p+o=p p,(r-i
p,
//(p-D
(2.25) p+
p,ff
The conjugate operators of d and X) have the forms (#WD=
( £
n d n 0)
\p+CT=P P.IT+1
P,
/ K.P+1)
(2.26) =
E p+ff^p
n / ( p + i / ° p ^-0j o (p), p a+1
1966]
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS
(^)KP- 1 )= ( 2
n s
\p+
143
n^) p ,a
' Up— 1)
(2.27) p+a=p
p,a
Now we introduce a complex Laplace-Beltrami operator (2.28)
• = 1)5 + 55
and its conjugate operator (2.29)
n = D5 + 5!D.
3. Proof of Theorem 3.1. Let M" be an almost-Hermitian manifold with an almost-Hermitian structure F/ SO that equation (2.3) holds. Then on the manifold M" we can define the real operators L and A as follows: (3.1)
L4> = 4>Aco
for any form (j), and A = *~1L*.
(3.2)
From the known fact that A commutes with * and * _ 1 , it follows from equation (3.2) immediately that if A commutes with L, it also commutes with A. Therefore it is sufficient to prove the theorem under the commutativity of A with A, that is, under the condition that for all forms 4> of any degree p (0 ^ p ^ n — 2) (3.3)
A(<£ A (a) = (A<£) A oo.
By means of equation (1.20), (2.5) for any form <> / of degree p (0 5j p 5S n — 2) given by (3.4)
4> =
fa{P)dxHp\
we can easily obtain [(A<£) A c o ] / ( p ) j , , 2 = - (V^V> / ( P ) )F,,, 2 +
i
4>I(p:mR'tFJlj2
s= l
(3.5) i
P
Vl(p;S\a,i\b)r
s
jab
P
(j)A03= \
sJ2
s=l P
(3.6)
-
1
P
\
D faipjiu)?}!!, + £
dx
Hp)
s
J2
A dx A dx ' ,
I
144
C.-C. HSIUNG
[March
[A(<£ Aca)]/(p),-u.2= - V'V,(>/(p)F„,2) p +
2
+
^I(p)FhaR"h
+< >
t I(p)F"J2R''ji
1 +
r <^/(p;S|a.i|6) F JiJ2 R a i*i*
^ s<(
P
P (j>l(p;S\a)Fbj2R''
+
2 s=l
+
isji
^
+
i,J2
<
t>I(p;S\a)Fij2R''ji
^
P q<s
1 ~
P ^I(p;i\a,l\b)FiJ2R"
•£ s
(3.7)
jli,
P ~
£
-
2 s= l
+F
ibR"ju2)
jii
P
1 —
S q<s
( ^ / ( p ^ i a ^ i f t j F y j j i ? ' '4J2
L-'.P
—
—
ab < >
t I(p;S\a,i\b)Fj1iR"
2 s
<
l>Hp:3\a)(.Pj1bR''
J2i
J2is +
F
bl
R
"
j2j,)
1 ...,p +
2 s
' / ' / ( p ^ l a J I f t ) ^ i,R"jU2+
T,
where T denotes the remaining terms. In particular, for a form
°/
where we have placed
y V V FV V 1
J J1J2
= UG. •
71J2"
1966]
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS 1
145
P
+ 2 £ F w ,K''* Wl + 2 £
F; l6 K'*\.
s= l P s= l
(3-9)
- ' l ' ' F W l R ' - W l - ' " l " F Wa K'-'', ll( q<s
s
- 2 2
P
FisbR'-Jlj2-
5= 1
S
FJlisR'sJ2
5= 1
Similarly, for a form $, all of whose components are zero except <£i(p) = xk for an arbitraryfixedlc(l^fc^H) and set I(p), we also have T= 0. Now at a general point P of the manifold M" let us choose orthogonal geodesic local coordinates x1, •••,x" so that (3-10)
gi{F) = &ij,
lt
where 5tj are Kronecker deltas. Then for this form
[(A0)Aa>] J(rtWl = x» £
FJd2R\,
5= 1
[A(> Aco)-]Hp)Jlj2= (3.12)
2VtFJlj2-xkV'VJFJlj2
+ xk 2 F,.uVR'v. + x*G,.uV 5= 1
From equations (3.11), (3.12), (3.9), (3.3), it follows immediately that VkFJlh = 0 for all k, j u j 2 at the point P, and hence the theorem is proved. 4. Realization of the complex operator • • Throughout this section the complex operator • is defined with respect to an almost-Hermitian structure F^ on an almost-Hermitian manifold M"unless stated otherwise. As was mentioned in the introduction, if the structure FtJ is Kahlerian, then the complex operator • is real and equal to £A. In this section we shall study the converse of this relationship between the operators • , A and the structure FtJ.
146
C-C. HSIUNG
[March
At first, we apply the operator Q to any form £ of degree zero. From equations (2.28), (2.24), (2.27), (2.16), (2.19), (2.22), (2.14), (2.15), (1.20) it follows that (A l \
p+ff = l
p.ff
1,0
= A{ + 1 V*iV( - F / Vt£ + V( - 1) V A since Vft Vfi = V,- VA<^, which implies that Fj Vh V ^ = 0. Thus, with respect to every £,, • is real (and therefore D = iA), if and only if V*iVVy£ = 0 for every £, or equivalently if and only if V*Fj*= 0 by choosing £ to be x*for an arbitrary k. Hence we obtain Theorem 4.1 in consequence of equation (2.8). We next apply the operator • to any form r\ of degree one. From equations (2.24),-, (2.28), (2.16) it follows that (4.2)
{Un\ = -(Ix + 1 2 + I 2 )^'(Hi +112)®kr\kl - I I I ,
where we have placed i, = e™.1"1 n
n
1,0 l
2
—
b
lL
iit
n 1X
mi
m2 1,0
1,0
(4-3)
S
»P U I - / 2 )
0,1 £
(rir 2 ) )
Hi = «?£}„ n f l l " n . ^ , , , TT
_
p"l6l
1,0
1,0
TT
«1 TT
1,0
0,1
"lpfcfcl .
in = nti9j (n/i + n / ' W v 1,0
\l,0
0,1
/
O'I •••jp) indicating that the indices j u •••,jp are in the natural order. By means of equations (2.14), (2.15), (4.3) an elementary calculation yields
it = n /• n ,/* - n hJi n /2 1,0
(4.4a)
0,1
= 1 (g, V » -
1,0
0,1
1,0
0,1
ftlV*
+iW1
0,1
0,1
~ FiihFh Oi < J2),
i2 = n /« n ,/* - n,/« n ,'* o\ <;2) 1,0
1,0
1,0
1,0
+ V( - ix^/'-f/ 2 + s W - giJlFhJl - giti2Fihy] Oi < J 2 ),
1966]
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS
k
iit = nu 1,0
n ^ - nj2 n > ut<j2) 1,0
1,0
1,0
= - r f . V - * ' - e.ke-kl ^
+
(4.4b)
Hi = \
LSji 6jz
^(
147
k
6j2
_ 1)(g F
oji
' * J2 * Ji
+ g^Fj"
J ^
{gjgj?
+ F-kF-kl - F kF.kl
-gjhu1
Jl
J2
- ZJ*F^ - SH'FJ^I
+ Fj?FJ2k> -Fj'Fu1)
(A
4III = V,, V% ~ Ft?F/ V„ V \ - Fit' V ; f i Vhr,j - V( - Wit"VtV^
+ V f l jy Vi,, +
Fk'VhV%).
From equations (4.4a,b), (2.19), (2.22) we thus obtain (Ii + I 2 +"l2)^'(Hi + H 2 ) ^ , , = y ( * , ' V " S . / V 2 ) • &[gMl
~ gj'sn1
+ V( - DGf/ ^
= y {.(gliig,li*-g,/1gtJx)&(gj!'gj2it
- * y > * * ) ] • V ^ , O i
+ V(-1)^*,^*)VAI]
+ FSV'F^VMJ
+ V(-l)(^1tViVtJ/7.-2F/V^V^
+ VF^Vrf, -
V'F^flu
+
FkJVkVhnj)l
which with equations (4.2), (4.4) give immediately 4(0/),, = - 2V i V i f ,, + [ V j , V i , y - f / V i F / V t ) | i i + fIl*JF«[Vlk, (4.5)
Vfoj-F/V'FjVtfj
+ f fl ' V ( /V Vfy + V ( - 1){ V'F/ V ^ t - (V.F,-,* + VhFk)VknJ + 2 F / V Vrf,,
where (4.6)
[ V t o V j = V » V , - V,V».
For the proofs of Theorem 4.2 and 4.3, we need the following LEMMA 4.1. / / the complex operator • for an almost-Hermitian structure on an almost-Hermitian manifold M" is real with respect to all forms ofldegrees 0 and 1, then the structure tensor FtJ satisfies
148
C.-C. HSIUNG
(4.7)
V , F / + VyF * = 0,
(4.8)
[March F/Rhm.
FtR'ju, =
Proof. We first choose orthogonal geodesic local coordinates xl,---,xn at a general point P of the manifold M", so that equations (3.10) hold. Then at the point P for the form r\ = xhdxl
(4.9)
for any fixed distinct h and i,
all the second covariant derivatives of any of its components being vanishing, Theorem 4.1 and the assumption that Im (D»7) = 0, Im denoting the imaginary part, imply immediately equation (4.7). Thus for general local coordinates x1, •••,x" and a general form r\ of degree 1, from equation (4.5) we can reduce the condition Im (OvDu = 0 to (4.10)
2FJkVJVkrlil - F tl *[V,, W
- F*'[Vk, V fl ]i,, = 0. (p)
By using the Ricci identity for any tensor >j(4f of type (p,q)
(4.11)
[V () V t ]^ (s) ^>= £ <W(p:Slo)Ri*fltI - £ * , { , V ( r t « V s=l
t= l
and the Bianchi identity R\}k + RhJki + RhkiJ = 0,
(4.12) and noticing that
2F/V'Vtf <1 = F'*[Vy,V 4 ], ;il , an elementary calculation from equation (4.10) gives readily equation (4.8). Hence the lemma is proved. It should be noted that due to equation (4.7) Fu is a Killing tensor (for this see, for instance, [8]), and it is well known that on the 6-sphere S6 there exists an almost-Hermitian structure [2], whose tensor is a Killing tensor [3]. It is also well known that the relation (4.8) holds automatically for a Kahlerian structure Ft J. In fact, by means of the identity (4.11) we have (4.13)
0 = [ V„ V J F / = Ft'RJM -
FbjRbm.
Multiplication of equation (4.13) by gthgJm leads immediately to equation (4.8). For the proof of Theorem 4.2, we first use the identity (4.11) and equations (4.8), (2.1), (4.7) to obtain
[ V„ V tl V = r,aR\ = F,,*FV[ V», VJ*,, (4.14)
- Fi V'F(1* Vhr,j + F,,' V ^ V% = F , ' V t f / V ' F , / - V'F^) = 2F/V*F,/Vrfj.
1966]
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS
149
Then from equations (4.5), (1.20), (4.13) and Theorem 4.1 it follows that (4.15)
2(13,),, = (Ai0fl + FjVFtl
V*,.
As before, by choosing orthogonal geodesic local coordinates x \ • • •, x" at a general point P of the manifold M" and considering the form r\ given by equation (4.9), from equation (4.15) we readily see that the condition D = iA implies that F/VhFhj=0
(4.16)
for each
iui,h.
k
Multiplication of equation (4.16) by Ft and use of equation (2.1) thus give V*F(1* = 0 for each ilth, k, which completes the proof of Theorem 4.2. Finally we apply the operator • to any form £ of degree two. From equations (2.24), -,(2.28), (2.16) it follows that (DOhb = - (IV! + IV2 + 1 ^ + fV2) W(Vi + V2 + V2) ®Ulk2) (4.17)
- (vij + vi^/vii! + vn2 + v n j s ) ' ^ ,
where we have placed lTV v
l
m m m i 2 3 iiii2
— —
r TT > "m! 1,0
bo
(4.18) — omim2"i °iiii2
Tv 2
x y
V
a i°2 a 3 Uij2Ji)
1
— ~
2
— bj>0l°26i — (JU2Jl)
V
TT "mi 1,0 TT
P E
1X
TT xx
r2
TT xx
r2
m2 1,0
r i
"' 11
V V
VT VI I
m m Pfc i 2 iii2
—
—
TT
(4.20) v l
2
fc
iii2
VTT V I I I
— P" 1 " 2 — siji
VTT Vll2
1 1 — — P Zi" jl *
(4.21)
TT
r
mi
'
1,0 1J - mi 1,0
1,0
m2 1,0
TT X 1
a2 1.0
1,0 TT 11 ai
TT
xx
lii 0,1
ai U1
n\ 0,1
c
'iMiixii). (rir 2 )si >
" 2 TT «3-*(*i*i) l l «3 £(«1«2«3)' 1,0
IT "» a2 1,0
xx
xx
Ul
TT x l
«2
1,0
TT "i x l ai 1,0 A 1
TT
xx
m2
XX
b'iJJUuii (rxr2r3)'
m3 1,0
1,0
1,0
(4.19)
1TT A
TT »i-*(*i*2) • »i fc(lliii2)i»> 0,1
xx
'V'-Ji fc (rir2)'
b
riSl'
" 2 p( i i*2) e (»i"2)>
TT (,, . B»p(*»*2) 11 £„,„, .
0,1
By means of equations (2.14), (2.15), (4.18) an elementary but rather lengthy calculation gives IVX + IV2 + IVi + iV2 = X^2h) .
(4.22)
where Mifh3 1S skew-symmetric in j l t j 2 , J3 and given by (A')X\
}JU2J3 _ »l'2
„ h„_ JlpJl o n 012 61
_
oMo. J*o. 1* — a, }loJ*a. •" 61 6(2 o i l 0I1 61 © 12
+ g,2Sig/2gir + s . / W V 3 - gi2Jigh'2gij3-
150
C.-C. HSIUNG
[March
From equations (4.19), (4.22) it follows that (4.24) (IVi + IV2 + IVx + N2)St(y1 where
+ V2 +Y2) = Wji^XAi
A, = n,,* n,.2*< n,3*2 1,0
(4.25)
1,0
(fc^fej),
1,0
A2= n* n* nf1,0
1,0
nit» n,> n„**
0,1
1,0
+ n / « n,2*> n, 3 1,0
+A2+ T2),
1,0
k
1,0
0,1
(fe1
0,1
Substituting equations (2.14), (2.15) in equations (4.25) we can easily obtain 8 W 0 * i + A2 + A2) =
MliVVSngj^Sj!2
4- v.kFklF,
k2
+ v.kiF-kF.kl
4- e,k2F-kF-ki
and therefore ^ V i
+ ^2 + i2) n t * 1,0
(4.27) Making use of equations (2.19), (2.22), (4.23), (4.24), (4.27) we thus have
4 imKiVi + iv 2 + iVi + iv.mvt (4.28)
=
-WfW&ih
+ v 2 + v 2 ** t CtwJ
+ FfWi&n-FfV'V&H
fillVhV^i!-fil'VVh[iil.
+
Similarly, by means of equations (4.20), (4.21), (4.3), (4.4), (2.19), (2.22) we obtain
(4.29)
(Vli + V I 2 ) ^ = i - [*,**,/' - f t * * , *
+ V (- i X f t W - ir./'O] v4, VJCVH! + vn2 + v n ^ ' c ^ j
= vjfo**, * - gfgp'fi'kMl
(4.30) =
T
[VfcVC,-, - V ( - l)(F fc ''V,VC yi + V ^ ' V * ^ ) ] ,
1966]
STRUCTURES AND OPERATORS ON HERMITIAN MANIFOLDS
151
from which it follows that 4 ImCCVIi + V I ^ / V I l ! + VII 2 + W l ! ) ^ * ^ ) ] (4.31)
= - F*Vfl V ^ „ - V ^ V 4 : , , , +
F/VhV^kll
+ V<2F»'V*C(ll + f „* V 4 V'C Hl - *„* V t V*C« 2 . Substitution of equations (4.28), (4.31) in equation (4.17) thus yields 4 Im(DOMl= VVV»C,lb-(VJlF<*+ (4 . 32 )
+
V,F(l*)V„C',2
< V'F<>* + v i a O Vrf',, +
2F/WACM2
+ g ^ C V * v j c „ a + giJFi2hlvj,vh]t;iil. On the other hand, by using equation (4.8) and the identities (4.11), (4.12) we can obtain 2F/V>V„CMJ
=
^[Vj.VJC.b
= 2Ft/RihC„i ^[V.pVjC.b (4.33)
2FtjRjtC,lt,
+
FtlJRX,2-FnuRlhJtl,
= -
F^tuR'jhh
= FJ tlkRk}ilt2=
- FJktklRtJilh
^ [ V ^ V J U
= - FtjRfCu
+
W C V ^ V j : , ^
-
= 0,
FfiWw
Fh%KHH-Fk%lRkhli2,
* tf J ? fa*[V J ,VJC!i 1 = - FtWChk
+
FttWw,.
Theorem 4.1 and equations (4.7), (4.33) thus reduce the right side of equation (4.32) to zero, and hence Theorem 4.3 is proved. REFERENCES
1. A. W. Adler, Classifying spaces for KShler metrics IV: The relation A = 2 # n , Math. Ann. (to appear). 2. C. Ehresmann, Sur les variitis presque complexes, Proceedings of the International Congress of Mathematicians, Cambridge, Mass., 1950, Vol. 2, Amer. Math. Soc., Providence, R. I., 1952, pp. 412-419. 3. T. Fukami and S. Ishihara, Almost-Hermitian structure on S6, T6hoku (Math. J. 7 (1955), 151-156. 4. K. Kodaira and D. C. Spencer, On the variation of almost-complex structure, Algebraic Geometry and Topology, a Symposium in Honor of S. Lefschetz, Princeton Univ. Press, Princeton, N. J., 1957, pp. 139-150. 5. G. De Rham and K. Kodaira, Harmonic integrals, Mimeographed^ notes, Institute for Advanced Study, Princeton, N. J., 1950.
C.-C. HSIUNG
152
6. M. Schiffer and D. C. Spencer, Function of finite Riemann surfaces, Princeton Univ. Press, Princeton, N. J., 1954. 7. A. Weil, Un thior&me fondamental dz Chern en giomitrie riemannienne, Bourbaki seminar notes, Paris, 1961-1962. 8. K. Yano and S. Bochner, Curvature and Betti numbers, Annals of Mathematics Studies, No. 32, Princeton Univ. Press, Princeton, N. J., 1953. LEHIGH UNIVERSITY, BETHLEHEM, PENNSYLVANIA
AFFINE DIFFERENTIAL GEOMETRY OF CLOSED HYPERSURFACESt By C H U A N - C H I H H S I U N G and J A M A L K .
SHAHIN
[Received 8 October 1965—Revised 24 February 1966]
Introduction I t is well known t h a t about four decades ago Blaschke (1) made a very extensive study of the affine differential geometry of closed convex surfaces in ordinary 3-space. I n order to simplify Blaschke's work, in 1957 Santalo (9) used E. Cartan's (2) method of moving frames and exterior differential calculus to obtain the fundamental equations of the geometry, by means of which some integral formulae for a closed surface were derived to characterize affine spheres or ellipsoids, as in 3-space a proper affine sphere is an ellipsoid ((1) 212). The purpose of this paper is to generalize Santalo's method to establish the affine differential geometry of closed orientable locally convex hypersurfaces Mn in an affine space An+1 of dimension n + 1 ( > 3). Section 1 is concerned with the group of the unimodular affine transformations and its structural equations in a Euclidean space En+1 of dimension n+ 1 ( ^ 3). I n § 2 we determine the Frenet affine frames in an invariant way, and thereby obtain the fundamental Frenet formulae in the theory of orientable locally convex hypersurfaces Mn in an affine space An+1. Section 3 contains the integrability conditions of the system of the fundamental Frenet formulae obtained in §2. I n § 4 we have the canonical expansion of the hypersurface Mn at the origin of the space An+1, together with the first and second affine fundamental forms of the hypersurface Mn at a general point. A verification t h a t these two fundamental forms are identical with Blaschke's is also given. In §5, we introduce, associated with a general point of the hypersurface Mn, the element of affine volume, affine mean curvatures, affine distance P from the origin, and the Laplacian of P. I n § 6 relations between some affine and metric invariants are given. Section 7 is devoted to the derivation of some integral formulae for a closed hypersurface Mn in the affine space An+1. f This research was partially supported by National Science Foundation grants GP-1567 and GP-4222. Proc. London Math. Soc. (3) 17 (1967) 715-35
716
C H U A N - C H I H H S I U N G AND JAMAL K.
SHAHIN
The last section contains various characterizations of affme hyperspheres by means of the integral formulae of § 7. The authors wish to acknowledge their indebtedness to the referee for his valuable suggestions in regard to the weakness of some conditions contained in this paper. 1. The affine group and its structural equations I n a Euclidean space En+1 of dimension n + 1 (n > 2), we shall consider the group of the unimodular affme transformations, t h a t is, the group (1-1)
X'a=*aJ>Xfi + ba,
where the determinant \a/\ = 1. Throughout this paper all italic indices take the values 1, ..., n, Greek indices the values 1, ..., n+1, and we shall follow the usual tensor convention t h a t when a letter appears in any term as a subscript and also a superscript, it is understood t h a t the letter is summed for all the possible values unless we state otherwise. Now consider a frame Xe1...en+1 of this group formed by a point X and n+1 vectors ev ..., en+1 in the space En+1 with the point X as origin, such that the determinant (1.2)
| e i , . . . , e m + 1 | = l.
Then we can write (1.3)
dX = co%,
(1.4)
dea = a)Jee,
where d denotes the exterior differentiation. The relative components oj a , o)J of the unimodular affine group can be computed by using equations (1.2), (1.3), (1.4). In fact, we have (1.5)
a>a = \e1,...,ea_1,
(1.6)
(x>J = \e1,...,efi_1,
dX,ea+1,...,en+1\, dea, es+v
...,en+1\.
By exterior differentiation of equation (1.2), and use of equation (1.6), we are readily led to (1.7)
S <
= 0.
a
Moreover, since ddX = ddex — 0, from equations (1.3), (1.4) we can easily obtain the equations of structure: (1.8) (1.9)
doja = o)/,Acofia, ^ /
= V A V .
where A denotes the exterior product.
A F F I N E D I F F E R E N T I A L GEOMETRY 717 2. Frenet affine frames Let X = X^u1, ...,un) be the vector equation of a hypersurface Mn of class C 2 in an affine space An+1 of dimension n+1 (n ^ 2). To study the hypersurface Mn we consider the submanifold of the manifold of frames Xex...en+1 as defined in § 1 such t h a t e 1; ...,en are tangent vectors of the hypersurface Mn at the point X. Denoting by the same symbols the forms on this submanifold of frames induced by the identity mapping, from equation (1.5) we have w^1
(2.1)
= 0,
and therefore, from equation (1.8), (2.2)
dcon+1 = u>k/\(x)kn+1 = 0.
By a lemma of E. Cartan ((2) 11) on exterior algebra, equation (2.2) implies (2.3)
0^+1 = V o '
(bij
=
bH).
I n order to determine the affine Frenet frames associated with the hypersurface Mn at the point X, we apply to the frame Xev..en+1 a unimodular transformation which leaves invariant the tangent space of the hypersurface Mn at the point X: (2.4)
e* = ajep,
a ^
= 0,
\aj\
= 1.
Then by using equations (1.3) and (2.4), and the equation, denoted by (1.3)*, corresponding to equation (1.3) for the frame Xe*...e* +1 , we obtain co{ = afco*).
(2.5)
Similarly, from equations (1.4), (1.4)*, (2.4) we have (2.6)
a,f»+! = a/a»>+Va„ +1 m+1 .
By introducing the matrices Q = K"+\...,
A = (ai),
co =
(w\...,con),
B={bij),
and the corresponding matrices Q*, w*, B* for the frame Xe*...e* +1 , equations (2.3), (2.3)*, (2.5), (2.6) can be written respectively as (2.8) (2.9)
Q = wB,
a* =
O* =
ntA/an+1n+1,
where lA denotes the transpose of the matrix A. From equations (2.8), (2.9) it follows immediately t h a t (2.10)
ABtA/an+1n+1
= B*,
718 CHUAN-CHIH HSIUNG AND JAMAL K. SHAHIN which implies t h a t (2.11)
| 5 * | = |^l'l-B|/(«Wi n + 1 ) n -
If the hypersurface Mn is orientable, then by considering an orientationpreserving transformation (2.4) of frames in the tangent space of the hypersurface Mn at the point X we have (2.12)
an+1^
> 0.
Since \A \ \B\ =fr 0, it follows from equation (2.11) t h a t (2.13)
sign | ^ | = sign 15* |
is an affine invariant. From now on in the remainder of this paper we shall always assume every hypersurface Mn to be orientable and locally convex; the latter condition is equivalent to the condition that the matrix B has either positive eigenvalues only or negative eigenvalues only. At first we assume t h a t the matrix B has only positive eigenvalues. Then by an elementary theorem in linear algebra the matrix B is positive. From equations (2.12), (2.3) it follows t h a t B/an+1n+1 is a positive symmetric matrix. Thus by a well-known theorem in linear algebra again, there exists an orthogonal matrix C such that A = \B\~1/nln+2)C and (2.14)
B* = ABlA/an+1n+i
= I,
where I is the identity matrix. From equation (2.4) we thus have an+1n+1 = \B\inn+2), and the second equation of (2.8) becomes tof^1
(2.15)
= a>*\
Since the o>** form a basis in the dual space of the tangent space of the hypersurface Mn at the point X, we may write (2.16)
la/co*\
so t h a t from equation (2.15), li:in+1 = 1 for i = j , and = 0 for i ^ j . I n order to simplify the computations, we first study the special transformation (2-17)
ef* = e*.
e**x = cB+1*e* + e*+1>
which will serve to define a unique affine normal by the condition (2.18)
w **«+i
= 0.
By applying equation (1.6) to the vectors ef*, ..., e*^, and making use of equations (2.17), (1.4), (1.2), we can easily obtain (2.19)
«**«+!
= S Cn+1 *o>"
+ «•"+*,
A F F I N E D I F F E R E N T I A L GEOMETRY so that the condition (2.18) becomes (2.20)
<
f =
719
- i ; w ^ i
A comparison of equations (2.16), (2.20) gives immediately (2-21)
cB+1* = - I n + w » + i ,
and hence the affine normal e**x is uniquely determined. On the other hand, from equations (2.18), (1.9), (2.6)*, (2.15) we have (2.22)
do**?*1 = to*** A cof «+* = £ a,**} A « « = 0, i
so t h a t we may write (2.23)
*>*« = lfa*i,
where Z/ = Z/. When the affine normal e**x has been determined, the only remaining admissible transformations are (2-24)
e, = # « , * * ,
^ i = e«i,
where J7 = (//) is an orthogonal matrix. Under the transformation (2.24), the forms <*>*** = w*i are related t o their transformed forms ' by, in consequence of equations (2.5), (2.25)
w**=//>'.
On the other hand, by means of equations (1.4), (2.18), (2.23) we have (2.26) Since den+1 = de*fv from equations (2.26), (2.25) it follows that (2-27)
//<W = Iffiw*.
Thus we obtain (2.28)
K + i ' - ' ^ + i " ) = (&,...,&*)FL*F,
where L = (£/). Since L is symmetric, there exists an orthogonal matrix F such t h a t the matrix FLlF is diagonal. Thus the frame Xev..en+1 is completely determined in an invariant way. If the matrix B has only negative eigenvalues then i t is negative, and we make the frame change ex -> — ex, e2, ..., e„ fixed, e n + 1 -> — e n+1 .
720 CHUAN-CHIH HSIUNG AND JAMAL K. SHAHIN By equation (2.10) with / -1
0
.
.
.
.
0
\
'-:•-.;} \ 0 . . . . 0 -1 / B is changed into —(ABlA), which has only positive eigenvalues. By applying the same procedures as above we may also determine the invariant Frenet frame Xev..en+1. By combining the above two cases we hence obtain the following fundamental Frenet formulae in the theory of orientable locally convex hypersurfacesf Mn in an affine space An+1 of dimension n + 1 (n ^ 2): (2.29)
dX = ai 1 ^,
(2.30)
dea = a>Jefi,
(2.31)
den+i
(2.32)
...
(2.33)
Wf™+1
(2.34)
-
1^0)%
n+l _ 0
^n+X
=
(o\ Kia)i
=
>
(2.35) i
(2.36)
<»i = kk^k'
where the ^ ' s are called the affine principal curvatures of the hypersurface Mn at the point X. If the KVS are all equal, the point X is called an affine umbilic. A hypersurface is called an affine hypersphere if all its points are affine umbilics. From equations (2.35), (2.36) we have n
(2.37)
^""-S^Wi=2
A comparison of equations (2.36), (2.37) gives immediately
(2.38)
V =
-2Vi=2
f For the local theory of hypersurfaces of other types see H . Flanders, 'Local theory of affine hypersurfaces', J. d'Analyse Math&matique, 15 (1965) 353-87, which appeared during the revision of this paper.
A F F I N E D I F F E R E N T I A L GEOMETRY 721 Moreover, by differentiating equation (2.33) and making use of equations (1.8), (1.9), (2.32), (2.33), (2.36) we can easily obtain
(2-39)
ink-y+w-w = o.
I n particular, when i = k equation (2.39) becomes (2.40)
IJ = \{y
+ lj)
(i not summed).
Finally, substitution of equation (2.36) in equations (1.8), (1.9) gives (2.41)
dot =
liki<Jha)k,
dco/ = lihmlmkia>h A cok + KW A (J.
(2.42)
3. The integrability conditions B y applying t h e equations (1.8), (1.9) of structure to the Frenet formulae (2.29), ..., (2.36), and making use of equations (2.37), ..., (2.42), we can obtain the integrability conditions of the system of the Frenet formulae. Let the covariant derivatives / f (i = 1, ...,n) of a function/(w 1 , ...,un) on the hypersurface Mn be defined by the relation (3-1)
df=ftia>*.
Exterior differentiation immediately (3-2)
and
use
of equations
(2.41), (2.36)
yield
d(likia>*) = (likrf + limHhkm)o>h A a A
By equating the corresponding sides of equations (2.42), (3.2) we are thus led to (3.3) yti (3-4)
- IJj = l^lmi
+ IJlf
- l ^ l j
+1J) + Ki
(i,j not summed),
kkj,n ~ kh\k = hhmLkj ~ kkmlmhj + kmj(hnm ~ hkm)>
where at least one of h and k is different from both i and j . Similarly, differentiating equation (2.34) and making use of equations (1.9), (2.32), (2.34), (2.36), we have (3.5) (3.6)
(KJ - «%£* = (K"—>c%/ i
K J = (KI — K^IH*
(j, Tc + i and not summed), (i,j not summed).
4. The canonical expansion and the fundamental forms Now let the origin 0 of the coordinate system in the affine space An+X be taken on the hypersurface Mn. Then by Taylor's theorem the position vector (4.1)
X =
X(u1,...,un)
of a point X on the hypersurface Mn in a neighbourhood of the origin 0
722
CHUAN-CHIH HSIUNG AND JAMAL K. SHAHIN
can be expanded in the series (4.2) X = (dX)0 + \{d*X)0 + UdaX)0 +..., where the subscript 0 denotes the value at the origin 0. By successive differentiation and the use of the Frenet formulae (2.29), (2.30), (2.33) we have (
dX = <»%, d2X = (dcoi + c^'to/)^ + £ w V e B + 1 ,
(4.3)
d3X = [d{d
1
fc
+ S (3w do + ZyA>Ww )e.n+l> Now suppose that (4.4)
(dX)0 = **(e,) 0 .
Then from equations (4.2), (4.3), (4.4) we can easily obtain the following canonical expansion of the hypersurface Mn at the origin 0: (4.5)
*"+1s|K)0,...,(en)0,X| = ^ S («*)s + | S ^ * * a ^ a * + . . . .
Any combination of to1, ..., w™ is an affine invariant differential form. We shall call (4.6)
?
= 2 (co*)* i
the first affine fundamental form of the hypersurface Mn in the affine space An+1; Blaschke ((1) 168) called it the quadratic fundamental form. Since the form
i
^ H I ^ W w ' , i
which we call the second affine fundamental form of the hypersurface Mn in the affine space An+1, and was called by Blaschke ((1) 168) the cubic fundamental form. I t might be desirable to express these invariant forms
A F F I N E D I F F E R E N T I A L GEOMETRY
723
given by the vector equation (4.1). Then we can write (4.8) a/ = gfduK Substituting equation (4.8) in equations (2.29) and (2.30), and comparing the resulting equations with dX = Xidui,
(4.9)
det =
fiej£duf,
we obtain immediately (4-10)
Xt = gjei,
(4-11)
fie^ = aijkek + 9jien+v
where X{ = BX/du1, and aitk are not needed in this paper. Partial differentiation of equation (4.10), and use of equations (2.30), (4.11), give (4-12)
Xtj = bi?ek + S
gfgfe^i,
k
where bitk are not needed either. From equations (4.10), (4.12) we obtain (4-13)
Atj =
\X1,...,Xn,Xtj\
= ^ n k"
k
k
G 0
0 1
.,e,n+l\
= 9 S 9ik9fk> k
where G = (gr/) and g = \G\. Thus (4.14)
A = \Aii\ = iT+*.
Similarly, from equations (4.10), (4.3) it follows t h a t \X1,...,Xn,d'X\=g?. Hence we can express tp in term of Blaschke's notation as follows: (4.15)
T
=
\X1,...,Xn,d*X\/A™»+».
Similarly, we have (4.16)
4, =
\X1,...,Xn,d*X\/AV<"+»-§d
5. Affine invariants At a point X of the hypersurface Mn, define the element of affine area dQ. to be (5.1)
dQ =
co1A...A(on,
and the ith affine mean curvature t o be the i t h elementary symmetric function of the affine principal curvatures K1, ..., Kn divided by the
724
CHUAN-CHIH H S I U N G AND JAMAL K.
SHAHIN
number of terms, namely, (^Wi= S Kh--Kji (1 < » < » ) . W h<-
(5.3)
P(Z) = ( - l ) » - i | Z - Z , e 1 , . . . , e J .
If Z is taken to be the origin 0 then (5.4)
P=(-l)-i|X>ei,...,en|.
For a closed convex hypersurface Mn the choice of the affine normal vector determined by the transformation (2.14) means t h a t the vector always points to the inside of the hypersurface Mn, and by such a choice the affine distance P from any point 0 inside the hypersurface Mn is always negative; the latter can be easily seen from equation (6.13). Thus the element of affine volume of the hypersurface Mn at a point X, formed by the elemental hyperpyramid with the origin 0 as the vertex and with the hyperparallelpiped as the base, formed by the vectors to1e1, ..., conen, is defined to be (5.5)
d7 = ^£|X>e1,...,el>1A...Ao>" 1
n+l'
-Pdto,
so t h a t the affine volume enclosed by a closed hypersurface Mn is given by
(5.6)
V=
?— f PdQ..
By using equations (2.29), (2.30), (2.35), (2.33), applying the ordinary rule for differentiation of determinants, and noticing that n is even, we can obtain (5.7)
dP=S(-l)i+1^l^;e1,...,ei,...,em+1|, i
where the circumflex over ei indicates t h a t the vector ei is to be deleted. Thus by equation (3.1) for P we have (5-8)
P i = (-l)«+i|Z,e1,...,«<,...,en+1|,
from which it follows t h a t if (5.9)
en+1 = XX,
t h a t is, if the affine normal vector en+1 is in the direction of the position
A F F I N E D I F F E R E N T I A L GEOMETRY 725 vector X, then Pt = 0, so t h a t the affine distance P is constant (for this result for the case n = 2 see Blaschke ((1) 111)). The absolute differential and the covariant partial derivatives of the intrinsic vector (Pv ...,Pn) are defined by the equation (5.10)
DPt = dP, -
= i>>.
PJOJJ
From equations (2.29), (2.30), (2.34), (5.8), (2.36) an elementary calculation gives (5.11)
dPti = (1 + P « V - S PJtk*"*i
Thus from equations (5.10), (5.11), (2.36), we have (5.12)
DPt = ( 1 + P « W - S ( J « ' + W > * > j
and therefore, in consequence of equation (2.40), (5.13)
Pa = 1 + PK* - 2 S
tyPj
(i not summed).
i
Hence, by using equations (5.13), (2.38), (5.2), we obtain the Laplacian ofP: (5.14) IIP ^ZPU = 71(1+?^). i
6. Relations between affine and metric invariants Let a hypersurface Mn in the affine space An+1 be given by the vector equation (4.1). Then from equations (5.1), (4.8), (4.14) we can easily express the element of affine area dQ. in terms of the local coordinates « {, s : (6.1)
dQ. = gdu1 A ... Adun = A1/in+2)du1
A ... Adun.
Let the coefficients of the first and second metric fundamental forms at a point X of a hypersurface Mn given by the vector equation (4.1) in the Euclidean space En+1 be Eti and Li} respectively. Then at the point X of the hypersurface Mn the element of metric area is (6.2)
da = EHu1A
...
Adun,
and the n metric principal curvatures are the roots of the determinantal equation (6.3)
\Lii-XEij\
= 0,
so t h a t the wth metric mean curvature of the hypersurface Mn at the point X is given by (6.4) where L = \Lij\, E = \Eti|. (6.5)
Kn =
L/E,
On the other hand, it is known t h a t X^x ... xXn
= S*N,
726
C H U A N - C H I H H S I U N G AND JAMAL K.
SHAHIN
where the left-hand side is the vector product of the n vectors Xx, ..., Xn in the Euclidean space En+1, and N is the unit metric normal vector of the hypersurface Mn at the point X. Thus we have (6.6)
X^X,
x...xXn)
= EW.Xtj
= EiLq,
where the dot between two vectors in the Euclidean space En+1 denotes the Euclidean scalar product of the two vectors. Since (6.7)
Xu-(Xix
... xX n ) = |X v
...,Xn,Xu\,
from equations (6.6), (4.13) we obtain immediately (6.8)
Ay =
E*Lip
A =
En/2L.
so t h a t (6.9)
By using equations (6.1), (6.2), (6.4), (6.9) we thus have the relation between the element dQ of affine area and the element da of metric area of the hypersurface Mn at the point X: dQ. = (Kn)Vin+u da.
(6.10)
Now let us consider the affine distance P from the origin 0 in the Euclidean space En+1 to the tangent space of the hypersurface Mn at the point X. From equations (4.10), (5.4), by replacing Xti in equation (4.13) by X we obtain immediately (6.11)
\X,X1,...,Xn\
=
(-l)"-igP.
Taking scalar products of both sides of equation (6.5) with the vector X we have \X,X1,...,Xn\^(-l)npE*,
(6.12)
where p = X.N is the metric distance from the origin 0 in the Euclidean space En+1 to the tangent space of the hypersurface Mn at the point X. From equations (6.11), (4.14), (6.4), (6.9), (6.12) we thus obtain (6.13)
P = -p/W»+»>.
As a generalization of the affine volume V given by equation (5.6) we consider the affine invariants Jm = {~\)m\
(6.14)
Pmd&. JM*
From equations (6.10), (6.13) it follows immediately t h a t (6.15)
Jm=
f
pm(Knyi-m)/in+2)
d(J
JM*
I n particular, we have J0 = Q, Jx = — (n+ 1)7. I n general, J is an affine
A F F I N E D I F F E R E N T I A L GEOMETRY 727 invariant of the hypersurface Mn with respect to the origin 0, from which the affine distance P and the metric distance p are defined. For closed convex surfaces M2 the invariants Jm satisfy certain inequalities. For example, J02 < —±TTJX, where equality holds for ellipsoid ((1) 198). I t would be interesting to extend this so-called affine isoperimetric inequality to a general n > 2. 7. Integral formulae In this section we shall apply the fundamental Frenet formulae given in §2 to obtain some integral formulae for a closed orientable hypersurface Mn in an affine space An+1. By means of the relation d2X = 0 and the ordinary rule for differentiation of determinants, we have the differential form (7.1)
d|X,e„+1,dX,
...,dX,den+1,
...,den+11
n—r
r—1
1 n 1
= ( - ) ~ \dX,..., dX, den+1,...,den+i, n—r+l
+ \X,dX,...,
en+1 \
r—1
dX, den+1,...,
n—r
den+i \
(r=l,...,n).
r
Using the elementary identity I .I = n\/[(n — i)\ i\] and equations (2.29), (2.31), (5.1), (5.2), we obtain (7.2)
\dX,...,dX,den+1,...,den+1, e
n—r+l
»+il
r—1
= (n-r+l)\
(r-l)\
X
Kh ... tci'-W1 A ... A at**-'-* A a>h A ... A aA->
S ii<...
•\eH'
• • • > ein-,+i>
= —^-r- SKH ... K*'-IdQ = n\
e
h'
•••>ejr-i'en+l\
Hr_xdQ.
Similarly, we have, in consequence of equation (5.4), (7.3)
\X,dX,...,dX,den+l,...,den+1\ ~-
„
-
n—r
-».
= (-1)"-%!
„
H^dCl.
--
r
Substitution of equations (7.2), (7.3) in equation (7.1) gives immediately (_ 1
(7.4)
-
)n-l
'-.—
d\X,en+1,dX,...,dX,den+1,...,den+1 n-r
= Hr_1dQ + HrPd£l.
r—1
728 CHUAN-CHIH HSITTNG AND JAMAL K. SHAHIN Integrating both sides of equation (7.4) over the hypersurface Mn, and applying Stoke's theorem, we obtain the integral formulae (7.5)
f (Hr_1 + HrP)d£l = 0
(r = l,...,m).
At all points of the hypersurface Mn, suppose t h a t Hv ..., Hr>0, 1 ^ r ^ n, and t h a t there exists a point 0 in the space An+X such t h a t P < 0. Let Hrm, HrU be respectively the minimum and maximum values of Hr on the hypersurface Mn. By putting (7.6)
jrr_x= f
H^dQ,
from equations (5.6), (7.5), we then have (7.7)
( n + 1 ) ^ F < Je^
< (n +
l)HmV,
where the equaUties hold only if Hr is constant, that is, only if Mn is an affine hypersphere, by Theorem 8.2 for s = r. I n order to extend the integral formula (7.5) for r = 1 we consider the differential form Pm\X,en+vdX, ...,dX\. By using equations (2.29), •
.
'
n-l
(2.31), (5.8), (3.1) for P , (7.4) for r = 1, we obtain (7-8)
(n-l)l
~7-~d(Pr-\X,en+vdX>...,dX\) m-1
= mP™-12 (^.i)2 da + nPm(l + HjP) d£l. Integrating both sides of equation (7.8) over the hypersurface Mn, and applying Stoke's theorem, we arrive at the integral formula (7.9)
m[
Pm-1Y1{PifdQ.
+ n[
Pm{l + HxP)dQ. = Q.
For m = 1, equation (7.9) becomes, in consequence of equation (5.6), f H1P*dn={n + l)V--{ S(-P<)S<*Q. JM" n JM" i If H1> 0 on the hypersurface Mn, by applying Schwarz's inequality for integrals and using equations (7.5) for r = 1, (7.6), (7.10), we obtain (7.10)
(7.11)
Q » < J T 1 f H1P*dQ = tf1\{n + l)V--[ S(P«)"dnl, JMn L njjn' i ' J where the equality holds only if P is constant, t h a t is, by Theorem 8.2 for s = 1, only if Mn is an affine hypersurface, as Hx is constant.
A F F I N E D I F F E R E N T I A L GEOMETRY Similarly, we have (7J2)
729
7^=T)Td(P"\X,e n + 1 ,de n + 1 ,...,de n + } \) n-l 2
= mHnP™-i S [(Pi) /**] dQ + nP™{Hn_x + HJP) dQ. i
Integrating both sides of equation (7.12) over the hypersurface Mn, and applying Stoke's theorem, we obtain the following integral formula, which is a generalization of the formula (7.5) for r = n: (7.13)
m\
^ P - ^ Z K / y y ^ d n + wf Pm(Hn^
JM"
i
+ HnP)dQ = 0.
JM"
I t should be noted t h a t the formula (7.5) for any r not equal to 1 or n cannot be easily extended in the same way as above. For n = 2, the formulae (7.9), (7.13) are due to Santalo (9). Furthermore, for this special case, by putting m—l for m in equation (7.9) and subtracting the resulting equation from equation (7.13), we have (7.14)
2f
(HtP*-l)P^1i^i
+ m{
Pm~V2(Pi)2
+ ^(Pjj) 2 ]dQ.
- (TO - 1 ) f P m " 2 [(P 1 ) 2 + (P 2 ) 2 ] dQ. = 0. JM* 2
If, on the surface M , H2 — 1/P 2 and K1, K2 > 0, then from equation (7.14) for TO = 1 it follows that P is constant and therefore H2 is also. Hence by Theorem 8.2 for n = 2, s = 2, the surface M2 is an affine sphere or an ellipsoid (Grotemeyer (4)). 8. Theorems At first we state the following three lemmas, which will be needed for the proofs of our main theorems. The proofs of the lemmas are omitted here, but can be found in ((5) 52, 104, 105). LEMMA 8.1. Let Sr be the rth elementary symmetric function ofn non-zero real numbers K1, ..., Kn, and suppose that H0 = 1 and Hr = Sr/l r = 1, ..., n. (8.1)
I for
Then H^H^-Hf^O
(r = l , . . . , » - l ) ,
where the equality for any value of r implies that K1 = ... = Kn. LEMMA 8.2. Let K1, ..., Kn, H0, Hr be defined as in Lemma 8.1. / / Hs_t, ..., Hg_x, Hs > 0, 1 s£ i s£ s < n, then (8.2)
H^/H^
< ... < E8_JES^ 1
s=
Hs^/Hs,
where the equality at any stage implies that K = ... = K71.
730
CHUAN-CHIH HSIUNG AND JAMAL K. SHAHIN
LEMMA 8.3. Let K1, ..., Kn, H0, Hr be defined as in Lemma
8.1.
//
Hx, ..., Hs > 0, 1 < s < n, then (8.3)
HX>HJ>
...>HV*,
where the equality at any stage implies that K1 = ... = Kn. I n all theorems of this section it is understood t h a t Mn is a closed orientable locally convex hypersurface of class C 2 in an affine space An+1 of dimension n+l (n > 2). THEOBEM 8.1. Suppose that there exists an integer s, 1 ^ s < n, such that Hs > 0, and either P < — Hs_x/Hs or P > — Hs_1/Hs at all points of the hypersurface Mn. Then Mn is an affine hypersphere.
Proof. Since Hs > 0, the conditions P < — Hs_x/h\s and P > — HS_JHS are respectively equivalent to HgP + Hg^^O and HsP + Hs_1'^0. Equation (7.5) for r = s, together with either of these two inequalities, implies t h a t (8.4)
P =
-Hs„1/Hs.
For s < n, substituting equation (8.4) in equation (7.5) for r = s+1, we obtain (8.5)
f (l/H„)(H>-Ha_1HM)dQ
= 0.
Due to the inequality in Lemma 8.1 for r = s, the integrand on the lefthand side of equation (8.5) is non-negative, and therefore equation (8.5) holds when and only when, a t all points of the hypersurface Mn, (8.6)
H'-H^H^-O.
From Lemma 8.1 it follows t h a t K1 = ... = Kn at all points of the hypersurface Mn, and thus Mn is an affine hypersphere. For s = n, substituting equation (8.4) in equation (7.5) for r = n— 1, we obtain (8.7)
f {\/Hn)(Hn_S-HnHn_2)d£l
= 0.
By applying Lemma 8.1 for r = n— 1 with the same argument as above, we can show t h a t Mn is also an affine hypersphere. Hence the proof of the theorem is complete. 8.2. Suppose that there exist a point 0 in the space An+1, and an integer s, 1 < s < n, such that at all points of Mn the function P is of the same sign, #,. > 0 for all i = 1, ..., s, and Hs is constant. Then Mn is an affine hypersphere. THEOBEM
A F F I N E D I F F E R E N T I A L GEOMETRY 731 Proof. Case 1. s < n. By the inequality (8.1) for r = 1, ..., s, and the assumption t h a t Hi> 0 for i = 1, ..., s, we obtain H1/H0>H2/H1>...>HS+1/HS, and, in particular, (8.8)
HXHS > Hs+V
where the equality implies that K1 = . . . = = Kn. From equation (7.5) for r = 1 and the assumption that H1> 0 and P is of the same sign a t all points of the hypersurface Mn, i t follows t h a t P must be negative. Multiplying both sides of the inequality (8.8) by P, integrating over the hypersurface Mn, and applying equation (7.5) for r = 1 and r = s+ 1, we can readily obtain, in consequence of the assumption t h a t Hs is constant, -Hs(
dft=f
HflsPdQ*:
f Hs+1PdQ
= -Hsf dQ,
from which it follows that (8.9)
f P(H1Hs-Hs+1)dQ
= 0.
Due to the inequality (8.8), the integrand on the left-hand side of equation (8.9) is non-positive, and therefore HXHS — Hs+1 = 0. From Lemma 8.1 it follows t h a t K1 = ... = Kn a t all points of the hypersurface Mn, and therefore the theorem is proved for s < n. Case 2. s = n. By using the assumption that Ht > 0 for i = 1, ..., n, from Lemma 8.3 we have the inequalities (8.10)
H1 > HJ > ... ^ Hn_v^-»
> Hnv* = c,
where c is a positive constant. By means of equation (7.5) for r = n, and the inequalities (8.10), we obtain (8.11)
f HnPdn
= -f
Hn_xdQ.^-cn~1\
dQ..
On the other hand, making use of equation (7.5) for r = 1, the inequalities (8.10), and the fact that P < 0, we have (8.12)
f HnPdQ. = cn-x[ RVnPd£l>cn-A JM"
JMn
HXPdQ = -c™" 1 f dQ. JM"
JMn
Combination of the inequalities (8.11) and (8.12) yields immediately (8.13)
f
P^^-H^dQ^O.
JM"
Due t o the inequalities (8.10), the integrand on the left-hand side of
732 CHUAN-CHIH HSIUNG AND JAMAL K. SHAHIN equation (8.13) is non-negative, and therefore Hn1/n = Hv which b y Lemma 8.3 implies t h a t K1 = ... = Kn at all points of the hypersurface Mn. Hence the proof of the theorem is complete. The formulae (7.5), and Theorems 8.1, 8.2, were obtained by Hsiung (6) for the metric case, and by Hsiung (7) and Feeman and Hsiung (3) jointly for the Riemannian case, which was further generalized by Katsurada (8). 8.1. The hypersurface Mn with the relation Ht = — 1/P > 0 is an affine hyper sphere. COROLLARY
This is an immediate consequence of Theorem 8.1 for « = 1, and also can be proved as follows. From equation (7.9), P is constant and therefore Hx is also. Hence b y Theorem 8.2 for s = 1, Mn is a n affine hypersphere. For n = 2, Corollary 8.1 is due to Grotemeyer (4). 8.2. The hypersurface Mn with the relation Hx > — 1/P > 0 (or 0 < Hx ^ — 1/P) is an affine hypersphere. COROLLARY
This is also an immediate consequence of Theorem 8.1 for 8=1. But by the assumption and equation (5.14) we have AP > 0 (or A P < 0). Using a lemma of Bochner-Hopf ((11) 26-31) we therefore obtain AP = 0, so t h a t H1 = — 1/P. Hence this corollary is reduced to Corollary 8.1. THEOREM 8.3. / / there are integers i and s, 1 < i < s < n, and constants Cj > 0 for i < j < s — 1 such that at all points of the hypersurface Mn, Hit ..., Hs> Oand
(8.14)
Hs=^CjHp 3
then Mn is an affine hypersphere. Proof. By equation (8.2) we have, for i ^j < s— 1, (8.15)
Hi/Hs-H^JH^ = (Ht/H^iH^Hs-H^Hi)
> 0,
where the equality holds only when K1 = ... = Kn. From equations (8.14), (8.15) it follows t h a t i
i
or fl,-,-S^H>0.
(8.16) i
where the equality holds only when K1 = ... = Kn.
A F F I N E D I F F E R E N T I A L GEOMETRY
733
On the other hand, by means of equation (7.5) we obtain (8.17)
f
(Hs_1-XciHj_1)dn
= -(
p(Hs-XcjH))d£l
= 0.
Equations (8.17), (8.16) imply t h a t •"s-l
=
XJ
C
J-"J-1
i
at all points of the hypersurface Mn. Hence Mn is an affine hypersphere. CoitOLLAitY 8.3. / / there are integers i and s, 1 < i < s < n, and a constant c such that Hi; ..., Hs > 0 and Hs = c i ^ at all points of the hypersurface Mn, then Mn is an affine hypersphere. THEOREM 8.4. Suppose that there are integers i and s, 0 < i < s < n, and constants Cj > 0 for i < j ^ s—1, such that at all points of the hypersurface Mn, Hit ..., Hs+1 > 0 and
(8.18)
tfs=2c^.. i
If P is of fixed sign throughout the hypersurface Mn then Mn is an affine hypersphere. Proof. By equation (8.2) we have, for i ^j (8.19)
^ s—1,
Hj/Hs - Hi+1/Hs+1 = m/H^iH^H,-
HHJE})
1
< 0,
n
where the equality holds only when K = ... = K . From equations (8.18), (8.19) it follows that 1 = Scfl/J, < i
or (8.20)
S e ^ ^ i . 3
5
w
- S ^
i < ° .
+
3
where the equality holds only when K1 = ... = Kn. On the other hand, by means of equation (7.5) we obtain (8.21)
f p(HH.1--2cfHj+1)dn JM"
\
3
=- \ /
( f l , - S c ^ ) d Q = 0.
JM»\
j
I
Since, from the assumption of the theorem and inequality (8.20),
is of fixed sign throughout the hypersurface Mn, equation (8.21) imphes that 3
n
at all points of the hypersurface M .
Hence Mn is an affine hypersphere.
734
CHUAN-CHIH HSIUNG AND JAMAL K. SHAHIN THEOREM 8.5. If there exist an integer s, 1 < s < n, and a constant c such that Hs = cHs_x > 0 at all points of the hypersurface Mn, then Mn is an affine hyper sphere. Proof. Since Hs > 0, c cannot be zero, and Hs_1 must be of fixed sign. By equation (7.5) we have, for 1 < s < n, H^H^-cH^)
= HS^-HS_2HS
> 0,
so t h a t Hs_1 — cHs_2 is of fixed sign, and vanishes identically only when K1 = ... = Kn. On the other hand, equation (7.5) gives f
(Hs_l-cHs_2)d£l
= f
PicH^-HJdQ^O,
which implies t h a t Hs_1 = cHs_2 at all points of the hypersurface Hence Mn is an affine hypersphere.
Mn.
THEOREM 8.6. Suppose that there exist an integer s, 1 < s ^ n, and a constant c such that at all points of the hypersurface Mn, Hi> 0 for i = 1, ..., s, and
(8.22)
fl^x1''"-1'
> c > H,1".
If P is of fixed sign throughout the hypersurface Mn then Mn is an affine hypersphere. Proof. Since P is of fixed sign, and Hx > 0, equation (7.5) for r = 1 implies t h a t P < 0. By equation (8.3) we have Ht ^ c, and therefore, in consequence of equations (8.22), (7.5), (8.23) - f c«- 1 # 1 PdQ ^ - f c*PdQ. > f H^dQ.
^ - f
c—lHxPd£i.
Thus the equality holds everywhere in (8.23), so t h a t f P{H1-c)dQ
= 0,
which implies that Ht = c. Hence, by Theorem 8.2 for s = 1, Mn is an affine hypersphere. THEOREM 8.7. Suppose that there exist an integer s, 1 < s ^ n, and a constant c such that at all points of the hypersurface Mn, -ffs_1; Hs > 0 and
(8.24)
H^H,
>c>
Hs_2/Hs_v
If P is of fixed sign throughout the hypersurface Mn then Mn is an affine hypersphere.
AFFINE DIFFERENTIAL
GEOMETRY
735
Proof. From equation (7.5) for r = s, it follows t h a t P < 0. On the other hand, by using equations (7.5), (8.24), we have (8.25)
f
Hs_2dQ. = - f
JM"
H^PdQ
> - \
cH^Pi
JMn
JjU"
= f cEs_xdQ. > [ Hs_sdD.. Thus the equality holds everywhere in (8.25), so t h a t (8.26)
f
P(Hs_1-cHs)dQ.
= 0.
Since P(fl g _ 1 — oHg) ^ 0, equation (8.26) implies t h a t i/ s _j = ci/ s at all points of the hypersurface ilf™. Hence, by Theorem 8.5, Mn is an affine hypersphere. Theorems 8.3-8.7 are due to Stong (10) in the Riemannian case. REFERENCES 1. W. BLASCHKE, Vorlesungen Uber Dijferentialgeometrie, II (Berlin, 1923). 2. E. CABTAN, Les systemes differentiels exterieurs et leurs applications geometrique (Paris, 1945). 3. G. F . FBBMAN and C. C. HSITJNG, 'Characterization of Riemann n-spheres', American J. Math. 81 (1959) 691-708. 4. K. P. GBOTBMEYEB, 'Eine kennzeichnende Eigenschaft der Affinspharen', Arch. Math. 3 (1952) 307-10. 5. G. H . HABDY, J . E . LITTLBWOOD, and G. POLYA, Inequalities (Cambridge, 1934).
6. C. C. HSIUNG, 'Some integral formulas for closed hypersurfaces', Math. Scandinavica 2 (1954) 286-94. 7. 'Some integral formulas for closed hypersurfaces in Riemannian space', Pacific J. Math. 6 (1956) 291-99. 8. Y. KATSUBADA, 'Generalized Minkowski formulas for closed hypersurfaces in Riemann space', Ann. Mat. Pura Appl. (4) 57 (1962) 283-93. 9. L. A. SANTOL6, 'Geometria diferencial afln y cuerpos convexos', Math. Notae 16 (1957) 20-42. 10. R. E. STONG, 'Some characterizations of Riemann n-spheres', Proc. American Math. Soc. 11 (1960) 945-51. 11. K. YANO and S. BOCHNBB, Curvature and Betti numbers (Princeton, 1953).
Lehigh University Bethlehem, Pa U.S.A.
ON THE GROUP OF CONFORMAL TRANSFORMATIONS OF A COMPACT RIEMANNIAN MANIFOLD. II B Y CHTJAN-CHIH HSIUNG
1. Introduction. Let Mn be a connected Riemannian manifold of dimension n, and C0(Mn), I0(Mn) the largest connected groups of conformal transformations and isometries of M" respectively. I n a previous paper [2], the author established THEOREM 1. Let Rhuk, Ra(h, i, j , k = 1, • • • , n) be respectively the Riemann and Ricci tensors of a compact Riemannian manifold M"(n > 2) with positive constant scalar curvature R, and suppose that
PBQ" = C = const. ,
(1) (2)
C
2v{p + g)R2'v+Q-l) nv+Q~\n - I)'-1 '
2p , (n ~ l)g P + Q
where p, q are nonnegative integers and not both zero, and P = RktikRkiit
(3) n
,
Q = tf'Ri, .
n
If C0(M ) i* I0{M ), then M" is isometric to a sphere. It should be noted that when p = 0, q = 1, or p = 1, q = 0, Equation (2) is an identity, and for the first special case Theorem 1 is due to Lichnerowicz [3]. Furthermore, we still have the open question: When p = q = 0, is Theorem 1 still true? On the other hand, Obata [4] obtained THEOREM 2. Let M"(n > 2) be a complete Riemannian manifold with metric tensor gu , and V the operator of covariant derivation of Mn. If M" admits a nonconstant function p such that V,-V,p = —c2pgij, where c is a positive constant, then M" is isometric to an n-sphere of radius 1/c.
Very recently, by making use of Theorem 2, Yano [5] proved THEOREM 3. Suppose that a compact orientable Riemannian manifold M"(n > 2) with constant R admits an infinitesimal nonhomothetic conformal transformation v so that
(4)
Lcg{i = 2<j>gij ,
<j> ^ const. ,
where L, is the operator of the infinitesimal transformation v. If
J
r.-^V dAn > o,
•I M
Received July 1, 1966. The research was partially supported by the National Science Foundation grant GP-4222. 337
338
C. C. H S I U N G
where
(5)
Tti = Rti - f gtt ,
It is obvious that Theorem 4 is a generalization of the special case of Theorem 1 where p = 1, q = 0 or p = 0, q = 1. The purpose of this paper is to generalize Theorem 4 to 5. Suppose that a compact Riemannian manifold Mn(n > 2) with constant R admits an infinitesimal nonhomothetic conformal transformation v. If THEOREM
(6)
a*L,P + b(2a + nb)LvQ = const. ,
where a and b are constants such that c = 4a2 + 2(n -
(7)
2)ab
+ TI(TI -
2)b2
>
0,
then Mn is isometric to a sphere. It is obvious that when a = 0 or b = 0, Theorem 5 becomes Theorem 4. 2. Lemmas. Throughout this section Mn will always denote a compact orientable Riemannian manifold of dimension n > 2. Let A be the LaplaceBeltrami operator on Mn. Then for any scalar field / on Mn (8)
A/ = - V ' W .
Thus we have (9)
f
A/ dAn = 0
from the well-known theorem of Green: (10)
[
V% dAn = 0,
where & is any vector field on M". LEMMA 1. If a nonconstant scalar field
339
TBANSFORMATIONS OF A RIEMANNIAN MANIFOLD
Proof. From Equations (8), (9) we obtain (11)
0= f
A(02) dAK = 2 [ = 2 \
(0 A* - <£>,) dAn (k
Equation (11) gives Lemma 1 immediately. LEMMA
2. Let v be an infinitesimal conjormal transformation on Mn so that
(12)
L,gti = 200,-, .
Then (13)
LJtknk = 2(j}Rhijk - ghkVi4>i +ff*,-V*
(14)
£„£„• = 0„ Atf> - (n - 2 ) V A ,
(15)
Z,,Z2 = 2(n - 1) A<£ - 2#0.
Lemma 2 can be proved by a straightforward computation (see, for instance, [1; 425-426]). LEMMA 3. / / M" has constant R and admits an infinitesimal nonhomothetic conformal transformation v so that (4) holds, then (16)
A
(17)
R > 0,
(18)
f
Equation (16) follows from Equation (15) due to the constancy of R, Equation (17) from Lemma 1, and Equation (18) from Equations (9), (16), (17). 3. Proof of Theorem 5. On the manifold Mn consider the covariant tensor field of order 4: (19) where (20)
Whijk = aTkiik + bghkTu ,
Thiik — Rhijk — ~, _ ... (gughk — QikQhUt
and a and b are constants satisfying (7). Then (21)
WhiikWhiik = a2P + (2a + nb)bQ - - [ - ^ r + (2a + nb)b R2, ib \_Tt
X
which and the constancy of R imply that (22)
L,(WhiikWhiik)
= a2L,P + b(2a + nb)LvQ.
__
340
C. C. HSITJNG
By assuming the infinitesimal nonhomothetic conformal transformation v to be denned by (4), from Equations (19), (20), (5), (12), (13), (14), (16) we can easily obtain (23)
L,Whiik
= 2a
ag^Vrti
4a + (3n — 4)6 „ , __ .
Multiplying both sides of Equation (23) by Whiih and making use of Equations (19), (20), (5), (8), (16), (21), (7), an elementary calculation yields WhiikL,Whiih
(24)
= 2$WhiikWhiik
- cT'V^
.
By substituting Equation (24) in the well-known formula Lv(WhiikWhiik)
(25)
= 2WhtikL,Whiik
- HWhiikWhiik
,
we thus have <j>L,(WhiikWhiik) = -H2WhiikWhiik
(26)
- 2o«r"'V^ .
Since the manifold M" is of constant R, it is known that for n > 2 (27)
V'X,- = 0,
and therefore (28)
V'Tn = 0.
Thus
(29)
v'(2W) = r^V + *r«vy.
Without loss of generality we may assume our manifold Mn to be orientable, since otherwise we need only to take an orientable twofold covering space of M". Substituting Equation (26) in Equation (29), integrating the resulting equation over the manifold M" and using Equation (10) we obtain (30)
2c I J M*
TM
dAn=
[
Jit'
.
J M»
On the right side of Equation (30), the first integral vanishes due to Equations (6), (22), (18), and the second integral is nonnegative since its integrand is so. Hence the integral on the left side of Equation (30) is nonnegative, and Theorem 5 follows from Theorem 3 immediately. REFERENCES
1. C. C. HSIUNG, Vectorfieldsand infinitesimal transformations on Riemannian manifolds with boundary, Bulletin de la Soci6t6 Mathfanatique de France, vol. 92(1964), pp. 411-434.
TRANSFORMATIONS OF A RIEMANNIAN MANIFOLD
341
2. C. C. HSIUNG, On the group of conformal transformations of a compact Riemannian manifold, Proceedings of the National Academy of Sciences, TJ. S. A., vol. 54 (1965), pp. 15091513. 3. A. LICHNEROWICZ, Sur les transformations conformes d'une variiti riemannienne compacte, Comptes Rendus de l'Academie des Sciences, Paris, vol 259(1964), pp. 697-700. 4. M. OBATA, Certain conditions for a Riemannian manifold to be isometric with a sphere, Journal of the Mathematical Society of Japan, vol. 14(1962), pp. 333-340. 5. K. YANO, On Riemannian manifolds with constant scalar curvature admitting a conformal transformation group, Proceedings of the National Academy of Sciences, U. S. A., vol. 55(1966), pp. 472-476. L E H I G H UNIVERSITY
J . DIFFERENTIAL GEOMETRY 1 (l%7) 8»-97
CURVATURE AND CHARACTERISTIC CLASSES OF COMPACT RIEMANNIAN MANIFOLDS YUK-KEUNG CHEUNG & CHUAN-CHIH HSIUNG In memory of Professor Vernon G. Grove
Introduction During the past quarter century the development of the theory of fibre bundles has led to a new direction in differential geometry for studying relationships between curvatures and certain topological invariants such as characteristic classes of a compact Riemannian manifold. Along this direction the first and simplest result is the Gauss-Bonnet formula [2], [3], which expresses the Euler-Poincare characteristic of a compact orientable Riemannian manifold of even dimension n as an integral of the n-th sectional curvature or the Lipschitz-Killing curvature times the element of area of the manifold. Later, Chern [5] obtained curvature conditions respectively for determining the sign of the Euler-Poincare characteristic and for the vanishing of the Pontrjagin classes of a compact orientable Riemannian manifold. Recently, Thorpe [8] extended a special case of Chern's conditions by using higher order sectional curvatures, which are weaker invariants of the Riemannian structure than the usual sectional curvature. The purpose of this paper is to further extend the conditions of both Chern and Thorpe. In §1, for a Riemannian manifold the equations of structure are given, and higher order sectional curvatures and related differential forms are defined. §2 contains the differential forms expressing, respectively, the Euler-Poincare characteristic and the Pontrjagin classes of compact orientable Riemannian manifolds in the sense of de Rham's theorem. In §3, we first define some general curvature conditions, and then use them to extend the above mentioned results of Chern and Thorpe. The proofs of the main results (Theorems 3.1 and 3.2) of this paper are easily deduced from several lemmas. 1. Higher order sectional curvatures Let M be a Riemannian manifold of dimension n (and class C°°), and Vx, V* respectively the spaces of tangent vectors and covectors at a point x of the manifold M. By taking an orthonormal basis in Vx and its dual basis in V*, Communicated April 20, 1967. The work of the second author was partially supported by the National Science Foundation grant GP-4222.
90
YUY-KEUNG CHEUNG & CHUAN-CHIH HSIUNG
over a neighborhood U of the point x on the manifold M, we then have a family of orthonormal frames xe1 • • • en and linear differential forms m^, • • •,a>n such that < e , ,
(1.1)
i
Throughout this paper all Latin indices take the values 1, • • •, n unless stated otherwise. The equations of structure of the Riemann metric are d(0t = 2 o)j A (Oji,
(1.2)
(Oij + o)jt = 0 ,
* dwij = 2 <»ik A (Oicj + Qij,
On + Qji = 0 ,
k
and the Bianchi identities are 2 oij A Qji = 0 , (1.3)
J
dQij + J^Qik A (i)kj — 2 Oik A £*./ = 0 , *
k
where the wedge A denotes the exterior multiplication. In terms of a local coordinate system w\ • • • ,un in the neighborhood U let (1.4)
et = £ ATf3/3M*. A-
Then (1-5)
$ ^ = —- 2 Stjkla)k A a>;, 2 *,;
where (l-O)
>Jijt; = RpqrsXiXjXrkXl
,
repeated indices implying summation over their ranges, and Rvqrs being the Riemann-Christoffel tensor. Throughout this paper, for indices we shall use /(p) to indicate the ordered set of p integers i\, • • •, iv among 1, • • - , « . When more than one set of indices is needed at one time, we shall use other capital letters such as J, H,R,S, • • • in addition to /. Now for an even p < n, we define the following p-f orm: (1.7)
61
A • • • A Qjp.ljp ,
where 8J,$ is + 1 (respectively — 1), if the integers ix, • • •, ip are distinct and J(p) is an even (respectively odd) permutation of lip); it is zero in all
CURVATURE AND CHARACTERISTIC CLASSES
91
other cases. Clearly, 9tj — Q{j. These forms 8I(V), except for constant factors, were first used by Chern in [3], [5]. For an even n, 81...n is intrinsic and called the Gauss curvature form of the manifold M, and the p-th Gauss curvature form studied by Eells [6] is closely related to d1...p. By using equation (1.5), equation (1.7) can be written in the form (1.8)
@I(p) =
-
.„ ,
2p/Zp\
Zl dnp)SjiHhxht '
- -
Sjp-ljpHp-lhp(0H(p),
Hip)
where we have placed (1-9)
a>H
For each p-dimensional plane P in the tangent space Vx of the manifold M at a point x, the Lipschitz-Killing curvature at the point x of the p-dimensional geodesic submanifold of the manifold M tangent to P at the point x is called the p-th sectional curvature of the manifold M at the point x with respect to the p-dimensional P, and is given (see for instance [1, p. 257]) in terms of any orthonormal basis eh, • • •, eip of P by (\ 101
K,
(i-iyj)
n.l(p)(r)
(P\ — ( ~ ^ ) ? / ,sRw,ssw
ff
...R
———-—o. n p ) o H ( p ) i\ T i r . i H S ^ • XJl- • -X'-'X^ii
v
^rp-irpsp-isp • -X^o .
ijr*ia
zp
From the geometric structure it is obvious that KI(P)(P) is independent of the choice of the orthonormal basis ei%, • • •, etp of P. For p = 2, K[(py(P) is the usual /?iemannian sectional curvature of the manifold M at the point x with respect to the plane P, and for p = n (even), it is the Lipschitz-Killing curvature of the manifold M at the point x. By using equation (1.6), equation (1.10) is readily reduced to (1.11)
KI(p)(p)
^ ^Hv)^np)^]i.jthihim - 'Sjp-ljphp-ihp •
=
2.
Characteristic classes
Let V be a vector space of dimension n over the real field R, and V* its dual space. Then there is a pairing of V and V* into R, which we denote by <X, X'> e R , X e V, X' e V*. The Grassmann algebra of A(V) of V is a graded algebra admitting a direct sum decomposition (2.1)
A ( F ) = A°(*0 + f\\V) + • • • +
An(V),
where Ar(V) is the subspace of all homogeneous elements of AC 7 ) of degree
92
YUY-KEUNG CHEUNG & CHUAN-CHIH HSIUNG
r. From A(V) and the Grassmann algebra A{V*) of V* we form their tensor product A(Y)<S> A(V*), which is generated as a vector space by products of the form £
(f
Suppose now a scalar product be given in V. We will be interested in the subspace A2k(V)® A2*(K*) of A(V)
A = 2 (etl A • • • A e««) <8>tf( „ , , u>
where £'HU) e A2k(V*),
and 2 denotes the summation over all the different
combinations of i1( • • •, i2* among 1, • • - , « . We call (2.4)
MI2=Lf?
w
6 A«(K*)
the square of the measure of A ; it is clearly independent of the orthonormal basis e 1; • • •, e„. These algebraic notions can be applied naturally to the space Vx of tangent vectors and the space V* of covectors at a point x of a differentiable manifold M of dimension n (and class C°°). Suppose that the equations of structure of the Riemann metric on the manifold M be given by equations (1.2). Then introduce (2.5)
Q= S ( e t A e j ) ® f l w .
For an even positive integer p < n, we can easily find (2.6)
£*' 2 = 2 [(e 4l A • • • Aeip) ® (QilUA • • •
AQtp.llpy\.
For a fixed set l(p), from equations (2.6), (1.7) it follows that the coefficient of the term e 4l A • • • A e i p in Qp/2 is equal to pi 6I(P), so that (2.7)
DP'2 = p\ 2 [(e4l A • • • A e4„) ® @ / ( p ) ].
In particular, for an even n, equation (2.7) becomes (2.8)
Q^2 = n\(e1A
••• A e J ® ^ . . . , .
Thus we can formulate the Gauss-Bonnet formulas as follows (see for instance [3]):
CURVATURE AND CHARACTERISTIC CLASSES
93
Theorem 2.1. Let M be a compact orientable Riemannian manifold of even dimension n. Then the Euler-Poincare characteristic x(M) of M is given by the inetgral
(2.9)
X(M)
= 2J—f^r 2V
f ©!.....
(rc/2)! J M
From equations (2.7), (2.4) we obtain immediately |£*I 2 =[(2*)!] 2 £(07C2*)) 2 >
(2.10)
and therefore a theorem of Chern [5] can be stated in the following form: Theorem 2.2.' The differential form
defines the k-th Pontrjagin class1 Pk of a compact orientable manifold in the sense of de Rham's 3.
Riemannian
theorem.
Relationships between curvatures and characteristic classes
Let M be a Riemannian manifold of dimension n, p an even positive integer < n, and alh = ahl, i, h = 1, • • • , « , given smooth real-valued functions on the manifold M at a point x. Denote (3-1)
A1(p)yH(p)
= \aiahp\
(or, /S = 1, • • -,p),
where the rows and columns of the determinant j aiah? | are arranged in the natural order of a and B, respectively. At the point x on the manifold M we then consider the following curvature condition: (3-2)
dntfSjijzh!^
• • • Sjp-ijphp-ihp = 2 p /
for all I(p), H(p) 6 ( 1 , • • • , « ) , the manifold M at the point Chern in [5]. From equations Lemma 3.1. For a fixed set (3.3)
where KP is a smooth real-valued function on x. For p = 2, this condition has been used by (3.2), (1.8), (1.9), (1.11) follows immediately of indices I(p) condition (3.2) implies
"i(v) — —r K p 2-i PI
(3.4)
KI
^KP),H
H{p)
KPAI(P)I(P)
and also equation (3.3) implies condition (3.2). 1
For the definition see also [4].
KpAI(p)tH(p)
,
>
YUY-KEUNG CHEUNG & CHUAN-CHIH HSIUNG
94
In particular, when atj = 5 U , then (3.5)
AI(p}H(p)
= dI(p)tH(py
,
where (3.6)
^i(p),HiP) = \Stahf\
(«, /3 = 1, • • -,p),
and therefore equations (3.3), (3.4) are reduced to (3.7)
^i(P)
— Kpi(P) ,
(3-8)
KIm = ( - \y>2Kv •
Thus, from equation (3.8) we have Lemma 3.2. Condition (3.2) with atj = 8tj implies that the p-th sectional curvature KI(P)(P) at the point x of the manifold M is constant, that is, independent of the p-dimensional plane P at the point x. On the other hand, from equations (1.8), (1.11) it follows immediately that equation (3.7) implies equation (3.8). The converse is also true as was proved by Thorpe [8], so that we can state, altogether, Lemma 3.3. Equations (3.7) and (3.8) are equivalent. For the converse of Lemma 3.2, we notice that [7, p. 238] (3.9) where 5/ ( p ) i W ( p ) is defined exactly in the same way as d%pf, so that equation (3.7) can be written as ®I(P)
=
rKP 2 J 8l(p),H(p)U>H(p) P\ H(p)
(3.10) = Pi
-Kp 2] ^I(p),H(p)0)H(p) • Hip)
A comparison of equation (3.10) with equation (1.8) yields immediately condition (3.2) with atj = 8tj. By using Lemma 3.3 and combining the result with Lemma 3.2 we hence obtain Lemma 3.4. The p-th sectional curvature KI(P) of the manifold M at a point x is constant if and only if condition (3.2) with ais — dtj holds. Lemma 3.5. On a Riemannian manifold M of dimension n, if condition (3.2) holds for some even p and q with p + q < n, then (3.11)
" / ( p + 9) — —
~-~KpKq
(P + q)\
2J
•^I
H(p + q)
so that condition (3.2) also holds for p + q with
KP+Q
=
KPK9.
CURVATURE AND CHARACTERISTIC CLASSES
95
Proof of Lemma 3.5. Let the set I(p + q) have distinct elements, and (h(P), hiQ)) be a partition of I(p + q), where l^p) = ( i u , • • •, ilp) and I2(q) — (*2i> • • •, ho)- Then, from equation (1.7), ft,
- —
""T*'
(3.12)
V
^'i(P)'2(?)0
A ... A O
(P + q)l «" 2 > A "421122 A
' " " A Mi2,q-li2q >
where 2 denotes the summation over all such partitions of I(p + q) into (/l.'s)
(hip), I2(<1))- For a fixed I(p + q), let /(p + 9) be an even permutation of I(p + q) such that / 1; • • • , / „ € / ^ p ) , / p + 1 , • • •, /„+, e 72(.?). By denoting J'(q) — O'p+i' •" •>/?+«) a n c i using equation (1.8), from equation (3.12) we then obtain "/(p + <7) = ~7 ; r r Zi "Ap)^iniia A • • • A W (P + )!
C'LV
(3-13)
A W ? 4 „ , „ A ••• Afi <2 , s -» 2S (p + )! (/,y>
On the other hand, by the Laplace expansion of the determinant/4,, ( p +5)H(!)+g) according to the first p rows we have Zj
(3.14)
•^J(p + q),H(p+q)(OH(p
+ q)
# + «> —
ZI ^ ^ ( P ) , H ( 9 ) a ' H ( ) ) ) A #(p>
Z J ^J'(Q),H'(Q)0)H'(q)
>
H'(q)
where H'(q) — (hp+1, • • •, hp+q). Substituting equation (3.3) in equation (3.13) and using equation (3.14) we arrive at equation (3.11), and an application of Lemma 3.1 hence completes the proof of Lemma 3.5. By repeatedly applying Lemma 3.5 we can easily obtain Corollary 3.5.1. Let px, • • • ,pk be even positive integers, and (mlt • • • ,mk) k
a k-tuple of nonnegative integers such that q — 2 miPi < n. On a Riemannian manifold M of dimension n, if condition (3.2) holds for plt • • • pk, then if also holds for q with >Cq =
(>CPl)mi---(>Cpk)m«-
Corollary 3.5.2. On a Riemannian manifold M of even dimension n if condition (3.2) holds for some even p dividing n, then (3.15)
0 i . . . n = (Kp)n/P I atJ I 0»! A ••• A (On,
where o>! A • • • f\ (onis the element of area of the manifold M.
96
YUY-KEUNG CHEUNG & CHUAN-CHIH HSIUNG
Combination of Theorem 2.1 with Corollary 3.5.2 gives immediately Theorem 3.1. On a compact orientable Riemannian manifold M of even dimension n if condition (3.2) holds at every point x for an even p dividing n, and (— l)n/2(Kp)n/p | atj \ keeps a constant sign, then this sign is the sign of the Euler-Poincare characteristic x(M) of the manifold M. Moreover, under this hypothesis, %(M) = 0 only when (KPY/V | a{j \ vanishes identically. This theorem was obtained by Chern [5] for p = 2, and by Thorpe [8] for a
ij
— "ij •
For studying Pontrjagin classes we need Lemma 3.6. Equation (3.3) can be written in the following form: (3.16)
0 / ( p ) — KpdJH A • • • A u>ip ,
where wia are linear forms defined by (3.17)
d»i„ = 2 aiah0)h
(a=l,
• • •, P) •
h
Proof. Let px, p2 be any two positive integers such that p1 + p2 = p. Then by the Laplace theorem we can expand the determinant AI(P} >H(P) according to the first p1 row. By using this expansion it is easily seen that all r M terms of 2 AHV)}H(p)(i)H(p) are equal so that we have
(3.18)
ZI ^I(P),H(P)0)H(P) I»
H(
=
/
: — 2 Pl'-Pl'- H
^I(P1),H(p1)0)H(p1)
2
H'(p2)
^I'(P3).H'(P2)0)H'(P2)
>
where (3.19)
I'(p2) = (iPi+1,
••-, ip),
H'(p2) = (hPi+1,
•••,hp).
Repeatedly applying the same argument as above to both factors of the righthand side of equation (3.18) yields immediately equation (3.16). Now we are in a position to prove the following theorem concerning the general curvature conditions for the vanishing of the Pontrjagin classes. Theorem 3.2. On a compact orientable Riemannian manifold M of dimension n if condition (3.2) holds at every point x for an even p
97
CURVATURE AND CHARACTERISTIC CLASSES
(3.20)
where
&I(2k)
2
—7— (2k)\
2 ] &J(p) A &J'(2k-p) (J,
,
J')
denotes the summation over all such partitions of I(2k) into
(J(p), J'(2k — p)). By using condition (3.2) for p and Lemmas 3.1 and 3.6, equation (3.20) is reduced to (3.21)
eI(2k)
=
p ! (
^~
p ) !
«, E ^
A ••• A ^
Aew-
p )
,
where d>^ are linear forms denned by equation (3.17), so that 6I(2k-, A 8i(2k) is a sum, each term of which contains an exterior factor (3.22)
&h A • • • A &iv A &h A • • • A &,v ,
where each subscript jzl(2k). Since 2A: < 2p, at least two of these j's must be equal, and therefore each of such factors (3.22) is zero. Thus 01<2k) A @i(2k) = 0 for all /(2&). By Theorem 2.2 we hence obtain Pk(M) = 0 for all k with pj2
(2*-1? < k < 2*p; i = 1, 2, • • . ) .
Hence Pk(M) = 0 for all k > p/2, and the theorem is proved. References [ 1]
C. B. Allendoerfer, Global theorems in Riemannian geometry, Bull. Amer. Math. Soc. 54 (1948) 249-259. [ 2 ] C. B. Allendoerfer & A. Weil, The Gauss-Bonnet theorem for Riemannian polyhedra, Trans. Amer. Math. Soc. 53 (1943) 101-129. [ 3 ] S. S. Chern, A simple intrinsic proof of the Gauss-Bonnet formula for closed Riemannian manifolds, Ann. of Math. 45 (1944) 747-752. [ 4] , Topics in differential geometry, Mimeographed notes, Institute for Advanced Study, Princeton, 1951. [ 5] , On curvature and characteristic classes of a Riemannian manifold, Adh. Math. Sem. Univ. Hamburg 20 (1956) 117-126. [ 6 ] J. Eells, Jr., A generalization of the Gauss-Bonnet theorem, Trans. Amer. Math. Soc. 92 (1959) 142-153. [ 7 ] F . D . Murnaghan, The generalized Kronecker symbol and its application to the theory of determinants, Amer. Math. Monthly 32 (1925) 233-241. [ 8 ] J. A . Thorpe, Sectional curvatures and characteristic classes, Ann. of Math. 80 (1964) 429-443. D R E X E L INSTITUTE OF TECHNOLOGY LEHIGH UNIVERSITY
Math. Zeitschr. 105, 307-312 (1968)
The Group of Conformal Transformations of a Compact Riemannian Manifold CHUAN-CHIH HSIUNG* and JONG-DUNG LIU
1. Introduction Let C(M"), 1(M") be respectively the groups of conformal transformations and isometries of a Riemannian manifold M" of dimension n>2, C0(M"), 70(M") the connected components of the identities of C(M"), I{M"), and Rijki and R^ the Riemann and Ricci tensors of the manifold M", respectively. Recently Barbance [1] obtained the following interesting Theorem 1. For a compact Riemannian manifold M" (n > 2) with positive constant scalar curvature R and constant R,J Rtj, condition C(M") + I(M")
(1.1)
implies that M" is isometric to a sphere. Theorem 1 is due to Lichnerowicz [5] when the condition (1.1) is replaced by C 0 (M"R/ 0 (M").
(1.2)
The purpose of this paper is to generalize Theorem 1 to Theorem2. Let M" (n>2) be a compact Riemannian manifold with positive constant R, and suppose that a2 P + b(2a + nb)Q = const.,
(1.3)
where P = R^klRijkl, Q = RijRtJ, and a, b are constant such that o+nb#0.
(1.4) (1.5)
Then condition (I.I) implies that M" is isometric to a sphere. Theorem 2 is reduced to Theorem 1 for a = 0, and is due to Hsiung [3] for b = 0 and the condition (1.1) replaced by (1.2). We need the following theorem of Obata [6] in our proof of Theorem 2. Theorem 3. A compact Einstein manifold M" of constant R admits a nonconstant function p such that Ap = nRp, (1.6) if and only if the manifold M" is isometric to an n-sphere of radius 1/|/R, where A is the Laplace-Beltrami operator. * The work of this author was partially supported by the National Science Foundation grant GP-4222. 22*
308
C.-C. Hsiung and J.-D. Liu:
2. Notations and Formulas Let M" be a Riemannian manifold of dimension n>2 and class C°°, \\gij\\ the symmetric matrix of the positive definite metric tensor g of M", and ||g'J|| the inverse matrix of \\gij\\. Throughout this paper all indices take the values 1,..., n, and we shall follow the usual tensor convention that indices can be raised and lowered by using g'j and gtJ respectively, and that repeated indices imply summation. In this section we shall list some well-known formulas for our latter use; for the details of their derivation see for example [2, pp.89 —90], where the Riemann tensor differs from ours in sign. Let / be a conformal transformation of M". Then the metric tensor g and its transformed tensor g* are related by /*(go) = g& = e 2 ' g u . (Z1) where a is any function on the manifold M". From Eq.(2.1) we have /*(gu) = g*u=e-2.gu (2 .2) For any tensor on M", the corresponding tensor defined by the canonical Riemannian connection corresponding to g* will be denoted by the same letter with a star. Then e~2aRfjki = Rijki-gii^}k-gjk^ii + gik^i + Sji^ik-(giigjk-gikgji)A^, (2.3) where we have put Axa=ViaVia, (2.4) (J..= V.Vi(j-ViaVjo, V being the operator of covariant derivation of M" with respect to g. By means of Eqs.(2.2), (2.3) we can have Rfk = Rjk + (2-n)cjk + gjklAa + (2-n)Al(T], (2.5) R* = e-2a[R where A is given by
+ 2(n-l)Aa
+ {n-l){2-n)A1a],
(2.6)
Aa=-ViVia.
(2.7)
3. Proof of Theorem 2 Due to condition (1.1) we may assume that there is a nonisometric conformal transformation/of the manifold M" so that §2 can be used here. On the manifold M" consider the covariant tensor field of order 4 obtained by Hsiung in [4]: Wijkl = a Rim~
n{n_x){gngjk-gikgji)
+bgii(Rjk-—gjk),
(3-1)
where a and b are constants satisfying (1.5). Substituting Eqs.(2.3), (2.5), (2.6) in the equation corresponding to Eq.(3.1) for f*{Wijkl), we obtain, after an elementary simplification, e~2aWi^kl=Wijkl
+ a^-gil(jjk-gjk(7il
+ gii(Jik-—
+ gikajl
(giigjk-gikgji)(A + Aia)j
+ bgil[(2-n)oJk
+ —(2-n){A(T +
Alo)gjl^.
(3.2)
Conformal Transformations of a Compact Riemannian Manifold
309
By putting u = e ", we have a=— Inu.
(3.3)
Substituting Eq.(3.3) in Eqs.(2.4), (2.7) yields immediately A^ = (Axu)lu2,
<Jjk=-(VkV}u)/u,
A<J=-(AU)/U-(A1U)/U2.
(3.4)
Therefore Eq.(3.2) becomes, in consequence of Eqs.(3.4), WiJkl = u~2^Wijkl + ~tijk[(u)], where we have placed r tijki(u) = a[gilVkVjU + ,(
+ bgil(n-2)[VkVjU
(3.5)
2Au i gjkVlViu-gikVlVjU-gJlVkViu+—^-(gilgjk-gikgJl)\ Au
\
+ — gjkJ.
On the other hand, from Eq.(3.1), we have WiiklWiJkl = a2 P + b(2a + nb)Q
[^— + b(2a + nb)]R2, n L n—1 J
(3.7)
which is equal to a positive constant A2 by the assumption of Theorem 2. Theref
°re
f*{A2)
= A*2 = A2.
(3.8)
But from Eq.(3.5) we obtain, by use of Eq.(2.2), A*2 = WlJklW*^k, = AA2+—tljkl(u)WlJk,
+ ^tlJkl(u)^k\u)\.
(3.9)
Combining Eq.(3.8) with Eq.(3.9) thus gives u-3A2 = uA2 + 2tiJkl(u)WiJk,
+ ~tijkl(u)tijkl(u). (3.10) u Now we divide our discussion into two cases. First Case. A = 0. In this case, since |jg£j-|| and \\gfj\\ are positive definite, we have WiJti=WtJkl = 0, (3.11) which and Eq.(3.5) imply tijkl{u)=0. (3.12) Multiplying Eqs.(3.1), (3.6) by g", summing for i, I, and using Eqs.(3.11), (3.12) and condition (1.5), we obtain R Au Rjk=—gjk, VkVjU + — gjk = 0, (3.13) the first of which showing that M" is an Einstein manifold. Since the conformal transformation / is nonisometric, u can not be reduced to a constant.
310
C.-C. Hsiung and J.-D. Liu:
By applying Ricci identity to V{u, multiplying the resulting equation by g ' \ and summing for i, k we have VtVjVtu-VJViVlu
= RjlViu.
(3.14)
From Eqs.(3.13) it follows that Rji = gikRjk = ^glkgj*
= 8lj^,
(3-15)
. Au Au — gIJ=-Vj—, n n where 5)=\ for i=j, and = 0 for i=t=j. Substituting Eqs.(3.15) in Eq.(3.14) and noticing, due to Eq.(2.7), V'V}Viu=-Vt
-VjV'ViU^VjAu, we obtain R
n-\ P;M =
ViAU,
or
V
>(-£TU-AU)=0>
since R is constant. Therefore there exists a constant u0 such that Au =
—(u — u0), n— 1 so that there is a nonconstant function u — u0 satisfying
(3.16)
n
A(u — u0) =
— (u — u0), (3.17) n— 1 which shows, by Theorem 3, that M" is isometric to a sphere. Second Case. yl=t=0. We shall show that in this case u = \, so that every conformal transformation / of the manifold M" is isometric, which contradicts condition (1.1). Thus this case will never occur, and Theorem 2 is proved. Since R is constant and n>2, it is known that
from which and Eq.(3.1) it is easily seen gugjtWJkl so that
= 0,
1Jkl
gtl^jW
gii{VkVju)Wiikl=VklVj{uguWi'kl)'\.
= 0,
(3.18) (3.19)
By means of the well-known theorem of Green that for any vector field £ on a compact orientable Riemannian manifold M",
IWdA^O,
(3.20)
Conformal Transformations of a Compact Riemanman Manifold
311
dA„ being the element of area of the manifold M" at a point with respect to the tensor g, from Eq.(3.19) we thus obtain Uu(^ju)W,Jk,dAn
= 0.
(3.21)
M"
Using Eq.(3.21) and the first equation of (3.18), from Eq.(3.6) hence follows itijkl(u)WiikldAn
= 0.
(3.22)
Since u > 0 , integration of (3.10) over M" and use of Eq.(3.22) yield immediately _ 3 , . . f ,. ,,-.. f J u dA„^ ) udA„. (3.23) Remark. The equality in (3.23) holds only when $ -tijkl(u)tiikl(u)dAn M"
= 0.
(3.24)
U
Eq.(3.24) implies tiJkl(u) = 0, which and Eq.(3.10) give u~3 = u and therefore « = 1. Substituting Eqs.(3.4) in Eq.(2.6) we have R* = u2^R + 2(l-n)~ + n(l-n)^f].
(3.25)
(3.26)
Let p be a real number. Since Vi(up + 1Viu)=~up+1Au
{p+l)upA1u,
+
an application of Green's formula (3.20) yields J up + 1AudA„ = (p+l) | upAxudAn. M»
(3.27)
M"
Multiplying both sides of Eq.(3.26) by up, integrating them over M", and using Eq.(3.27), we obtain \RupdA„= Since Atu^0
j RupJr2dAn
+ {2p + 2 + n){\-n)\upArudAn.
(3.28)
and R is constant, we have J updAn^
| up+2dA„
for 2p + 2 + n^0.
(3.29)
For n 2:4, substitution of p = — 1 and p = — 3 in Eq. (3.29) gives J u~3dA„S M"
J u _ 1 ^ „ ^ f «di4„. M"
(3.30)
M"
By comparing (3.30) with (3.23) we obtain J u~3dA„=
J udi„,
and therefore u = l by the above remark.
(3.31)
312
Hsiung and Liu: Conformal Transformations of a Compact Riemannian Manifold
For n = 3, putting p= — 1 in Eq.(3.29) we have J u~ldAn^
| udAn.
M"
(3.32)
M"
By assuming the area of the manifold M" to be 1, an application of Schwarz inequality to u~* and u* yields j" u~1dA„ | udAn^ M"
M"
\ dA„ = l.
(3.33)
M"
Comparing (3.33) with (3.32) we obtain judAn^l.
(3.34)
Since n — Z, from Eq.(2.1) it follows that Ig5l = "~ 6 Inl-
and therefore that
dA% = u~3 dA„,
where dA* denotes the element of area of the manifold M" at a point with respect to the tensor g*. Thus \dA*=
\u~*dAn=
\dAn=\,
(3.35)
since the area of the manifold M" is fixed. Substitution of Eq.(3.35) in (3.34) glV6S
\udAn^
\ u'3dA„.
(3.36)
By comparing (3.36) with (3.23) we obtain Eq.(3.31) and therefore w=l by the above remark. Hence the proof of Theorem 2 is complete. Bibliography 1. Barbance, C : Transformations conformes d'une variete riemanniene compacte. C. R. Acad. Sci. Paris 260, 1547-1549 (1965). 2. Eisenhart, L. P.: Riemannian geometry, 2nd edition. Princeton: Princeton University Press 1949. 3. Hsiung, C. C : On the group of conformal transformations of a compact Riemannian manifold. Proc. Nat. Acad. Sci. U.S.A. 54, 1509-1513 (1965). 4. — On the group of conformal transformations of a compact Riemannian manifold. II. Duke Math. J. 34, 337-341 (1967). 5. Lichnerowicz, A.: Sur les transformations conformes d'une variete riemannienne compacte. C.R. Acad. Sci. Paris 259, 697-700 (1964). 6. Obata, M.: Certain conditions for a Riemannian manifold to be isometric with a sphere. J. Math. Soc. Japan 14, 333-340 (1962). Professor Chuan-Chih Hsiung Department of Mathematics and Astronomy Lehigh University Bethlehem, Pa. 18015, USA (Received October 19,1967)
J . DIFFERENTIAL GEOMETRY 2 (1968) 9-24
ISOMETRIES OF COMPACT SUBMANIFOLDS OF A RIEMANNIAN MANIFOLD CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
Introduction Let M be a differentiable manifold of dimension n > 2, x, x* : M —> R two immersed compact submanifolds in a Riemannian manifold of dimension n + m (m > 0) admitting a special infinitesimal nonisometric conformal transformation f, and f: x(M) —»x*(M) a volume-preserving diffeomorphism. Conditions are found for (i) / to be an isometry and (ii) hypersurfaces x{M), x*(M) to have the same second fundamental form when m = 1. In particular, when R is a Euclidean space E, these conditions were obtained by Chern and Hsiung [1] for case (i), and by Hsiung [3] for case (ii), as in these cases, in E the position vector of a point with respect to a fixed point 0 generates such a special infinitesimal nonisometric homothetic transformation. In order to simplify the presentation of our work, as before we shall consider tensor products of multivectors and exterior differential forms. Differentiation of multivectors will be taken in the sense of equation (2.22), and differentiation of exterior differential forms will be exterior differentiation; multiplication of matrices will be by the usual row-by-column law. 1.
Lemmas
Let V be a real vector space of dimension n ( > 2 ) , and let G and H be two bilinear real-valued functions over V XV so that G and H are completely determined by the values gtk = G(e4, ek) and hik = H(eu ek), 1 < i, k < n, where the vectors eu • • • ,en constitute a basis for the space V. Under a change of basis (1.1)
et-+et=t\et,
the matrices \gik\ and \hik\ are changed to T\gikYT and T\hlkyT respectively, where the repeated index k is summed over its range, T = \\tf\\, and l T is the transpose of T. Consider the determinant Communicated June 20, 1967. The work of the first author was partially supported by the National Science Foundation grant GP-4222.
10
(1.2)
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
| gik + Xhik | = | gik | + n?.P(gik, hik) + • • • + r-1 hik I •
Since \glk + Xhik | will be multiplied by j Tf under the change (1.1) of basis, the ratio of any two coefficients in the polynomial in X on the right side of equation (1.2) is independent of the choice of the basis e1 • • • en for V. In particular, if G is nonsingular, that is, if \gik\^0, the quotient (1.3)
H0=P{gik,hik)l\gik\
depends only on G and H. For example, if gik = 8ik ( — 1 for i = k; = 0 for i ^ k), then HG — 2 hu/n. i
Since the construction of HG is linear in H, it can be generalized to a bilinear function H over V X V with values in a vector space W, and HG is then an element in W. HG can be called the contraction of H relative to G, or, in the language of tensors, Ha is a vector in W constructed from a covariant tensor H of order two with values in W relative to a real-valued nonsingular covariant tensor G of order two. The following two lemmas will be needed in the proof of our main theorems. Lemma 1.1. Let V be a real vector space of dimension n ( > 2 ) , and G and H symmetric positive definite bilinear real-valued functions over V X V completely determined by the values gik — G(et, ek) and hik = H(et, ek), 1 < i, k < n, where ex, • • •, en form a basis of the vector space V. Denote (1.4)
S=|*»|,
h =
\hlk\.
Then (1.5)
Ha > (hlgf*
,
where the equality holds when and only when hik = cgik for a certain c. The inequality (1.5) is one in a class of inequalities for hyperbolic polynomials given by Garding [2]. A direct and elementary proof of the inequality (1.5) was given by Rhodes [4] by making use of the generalized arithmeticmean—geometric-mean inequality. A still simpler proof of the inequality (1.5) is obtained as follows: Since both matrices G and H are positive definite, we can choose the matrix T of a change of basis such that G be the unit matrix and H a diagonal matrix. Then by the generalized arithmetic-mean—geometricmean inequality we obtain HG = ±Zhu> n
(hn • • • hnny*
= (h/gf*
,
i
where the equality holds when and only when hik = cgilc for a certain c. Lemma 1.2. Let e1} • • •, en be any n vectors in a vector space V of dimen-
ISOMETRIES OF COMPACT SUBMANIFOLDS
11
sion n, and &, • • •, £„, rfc (i, k = 1, • • •, n) any elements of the dual space V* of V such that
£ ri = o .
(1.6) Denote (1.7)
e = i = fe,
f =Ui,
J? = | | J # | ,
• • •> ? » | ,
1 < i, k < n .
Then (1.8)
yew"'1 - (n - l)efyewn~* = 0 ,
where the product of elements in each space is the exterior product. This lemma is due to Chern and Hsiung; for its proof see [1, p. 283].
2. Submanifolds of a Riemannian manifold Throughout this paper the ranges of indices are given as follows, unless stated otherwise: 1 < i, j , k, • • •
(2.1)
< n ,
1 0 ) .
Let R be a C2-Riemannian manifold of dimension n-\- m with fundamental tensor G, and Rx, R* respectively the tangent space and its dual space at a point xeR. By taking C2-bases ex • • • en+m and CD1 • • • con+m in the tangent spaces and their dual spaces over a neighborhood U of the point jcei?, we then have (2.2)
dx = a)"ea ,
(2.3)
dea = wfaep ,
where d denotes exterior differentiation, the repeated indices imply summation, and the w^ are linear differential forms. Since a1, • • •, wn+m span the space of linear differential forms over the neighborhood U, we can write (2.4)
where y"fa are the components of the affine connexion C of the Riemannian metric G with respect to the frame xex • • • en+m.
12
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
Let x1, • • •, xn+m be a system of local coordinates in the neighborhood U of the point x. Then (2.5)
dx> = a>'e> ,
where el (/3 — 1, • • • , « + rri) are the components of ea. Put (2.6)
ffl;
= />/*• ,
where J"^ are the components of the affine connexion C in the local coordinates x1, • • -,xn+m. Substituting equation (2.5) in equation (2.6), and comparing the resulting equation with equation (2.4), we then find
(2.7)
rf, = / > ; .
Let M be an abstract manifold of dimension n, and consider an immersion x.M ~+R, that is, a C2-mapping x of M into R such that the induced linear mapping x# on the tangent spaces is univalent everywhere. By choosing eu • • •, en to be in the tangent space of the manifold x(M) at the point x{p), p e M, and denoting by the same symbols the forms on this submanifold of frames ex • • • en induced by the mapping x, at the point x{p) of the manifold x(M) we have, from equation (2.2), (2.8)
con+s = 0
(s = 1, •• -,m) ,
and therefore (2.9)
dx = w^i ,
(2.10)
det = a>\e, .
Let f be a vector field on the manifold R, so that (2.11)
i = Ve. •
By using equations (2.3), (2.4), (2.8) we then obtain (2.12)
d$ = {Vi^)
where Ft denotes the covariant derivative in the direction et, and V£" is given by (2.i3)
r ( f = r , i + ZYu »
f Bit being defined by (2.14)
<*£" = f.iffl* •
ISOMETRIES OF COMPACT SUBMANIFOLDS
13
More generally, for a mixed tensor ^ we have (2.15)
pt?f = f;,4 + fjrli -
£rt
Since
df-
(2.16)
_9T -d** 3**
substituting equation (2.5) in equation (2.16), and comparing the resulting equation with (2.14), we obtain (2.17) which holds for any tensor. From equations (2.7), (2.17) it follows that if e1 • • -en is a natural frame, then V becomes the covariant differentiation in local coordinates. Thus, by the definition of tensors, (2.18)
V&.t = 0 ,
where gaP are the components of the fundamental tensor G of the manifold R. Moreover, on R if f generates an infinitesimal conformal transformation, then, (2.19)
f 7 ^ f + Vfi. = 2Pg.
where p is a real-valued function on R. In particular, the infinitesimal conformal transformation generated by the vector field § is homothetic or isometric according as p is a constant function or p = 0. On R, f is said to generate an enlarging infinitesimal conformal transformation, if (2.19) holds with positive p. From equation (2.19) it follows that p =
F„f", and n + m therefore, by means of the well-known theorem of Green, that if R is compact 0, where dV is the volume element of the and orientable, then fpdV manifold R. Hence if R is compact and orientable, then it does not admit any enlarging infinitesimal conformal transformations. We are indebted to K. Yano for calling our attention to this point. Now choose en+1 • • • en+m to be an orthonormal frame in the normal bundle of x(M). By introducing the matrices a) = J a)1, ••-, (on\\ ,
(2.20)
(2.21)
e = ^n + n
14
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
equations (2.9), (2.10) then become (2.22)
dx = coe,
de = Qe + da ,
where
(2.23)
fl
= KI
0 = K + 'll-
Exterior differentiation of the first equation of (2.22) gives (2.24)
do = co A Q,
a Ad = 0 .
The second equation of (2.24) can be written explicitly as (2.25)
cu* A o>rs = 0 .
By applying a well-known lemma of E. Cartan, from equation (2.25) we then have (2.26)
a>rs = Atf'to',
A?;* = Aft' •
These functions Aft' defined on the submanifold x{M) are coefficients in the second fundamental form 77 of x(M). In fact, from equations (2.21) and (2.26) we obtain the second differential of x in the ordinary sense: (2.27)
d2x = {do? + coW^et + A2/swicoJen+s .
If v is any normal vector of x{M) at x(p), then the second fundamental relative to v is given by (2.28)
form
77, = < „ , (Px> =
where < , > denotes the scalar product in the tangent space of the manifold 7?, and the multiplication of linear differential forms is commutative in the sense of the symmetric algebra. For each normal field v, since 77„ is a globally defined symmetric covariant tensor field of order two on x(M), we may form the contraction (77„)Gn of 77„ relative to the induced fundamental tensor Gn on x{M). Therefore there is exactly one normal vector field H over x(M) defined by (2.29)
,
independent of the particular v chosen. This normal vector field H is called the mean curvature vector. It is easy to verify that H = A??sziJe A submanifold is called minimal if its mean curvature vector is identically zero.
15
ISOMETRIES OF COMPACT SUBMANIFOLDS
When m = 1 we may take v to be the unit normal vector en+1 so that from equation (2.28) the second fundamental form of the hypersurface x(M) at the point x(p) becomes II = A^w'w1 .
(2.30) 3.
Integral formulas
Consider two immersions A;, A:* of a C2-Riemannian manifold M of dimension n ( > 2) into a Riemannian manifold R of dimension n + m (m > 0), and a diffeomorphism / as given by the commutative diagram M —
-
X{M)
^ \ X
C R
f ^
**(M) C # .
Then §2 can be applied to the submanifold x(M), and for the corresponding quantities and equations for the submanifold x*(M) we shall use the same symbols and numbers with a star respectively. Suppose that / is volumepreserving; that is, by definition it maps the volume element of one immersed submanifold into that of the other submanifold. As a consequence of this definition f exists only if M is oriented, and f is then orientation-preserving. Now over the abstract manifold M there are two induced Riemannian metrics with the same volume element, namely,
dx*(p)>
=
d(foX)(p)>
.
Thus the notion of frames ex • • • en having measure 1 and an orientation coherent with that of M has a sense in both metrics. At a point p e M any such frame can be obtained from a fixed one by a linear transformaton of determinant 1. The volume element dV of the submanifold x(M) is given by (3.1)
dV = o1 A ••• A co",
where m1 • • • wn is the coframe dual to the frame ex • • • en of measure 1. The condition for the frame ex • • • en to be of measure 1 is (3.2)
<e x A • • • A en, ex A • • • A en> = 1 ,
where the left-hand side is the scalar product of multivectors defined by the scalar product in the tangent space Rx of the manifold R. Differentiating equation (3.2) and using the second equation of (2.22), we obtain, in consequence of <e 4 , en+s> = 0 ,
16
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
(3.3)
TrQ = 2 a>\ = 0 . i
For an (m X «)-matrix a — |[«4j?|[ and an (nxp)-matrix b — \\bjk\, whose entries are vectors in the tangent space Rx, we shall use the notation < a , by to denote the matrix of real numbers given by
(3.4)
<«,*> = IE <**>***> Ij
In order to derive a new integral formula for two compact submanifolds x(M), x*(M) under a volume-preserving diffeomorphism /, we shall assume that on the manifold R there is a vector field £. Put (3.5)
< £ , en+s> =
yn+s,
and introduce the matrix (3.6)
= <£, Ja> .
h = \\yn+1, • • •, yn+m\
By using equations (2.23), (2.26) we then have the (1 Xn)-matrix (3.7)
m = \\yn+sA?;W\\ .
On the other hand, consider the orthogonal projection £L(x(p)) of the vector $(x(p)) in the normal space of the submanifold x(M) at x(p), so that we have f 1 = yn+sen+s
(3.8)
•
Thus substitution of equation (3.8) in equation (2.28) gives the quadratic differential form (3.9)
/7f± = < £ x ,
.
In order to construct some exterior differential forms globally defined over the submanifolds x(M), x*(M), we further introduce the following matrices: G = <e, le> = 'G=
(3.10) (3.11) (3.12) (3.13)
G* -
*(«), /*(««)> = lG* = \g% A = |o)i, • • •, mn\\ = coG , A*=
(3.16)
K , . . . , a . * | = a.G*,
u — Ae ,
(3.14) (3.15)
\\gik\\ ,
Y =
< f , *e>
u* = A*e , =
r=Ye,
\\ylt • • -,yn\\
,
ISOMETRIES OF COMPACT SUBMANIFOLDS
0 = ru*n-1 =
(3.17)
17
Aen,
1
(3.18)
8= IfflVifx, •••,a) Fif»l •
The diffeomorphism / is an isometry if and only if G = G*. Since $ is an exterior differential form of degree n — 1 globally defined over M, for a compact M Stokes' theorem gives
(3.19)
Cd$ = 0.
Making use of equations (2.12), (3.18), (3.10), (2.22), (3.15), (3.16) we obtain dY = < d § , £ e> + < £ , = 0 + Y£fl + hl6 ,
(3.20)
from which, together with (3.16), it follows that dr = [d + Y(lQ + Q) + hfS\e + Yda .
(3.21)
By equations (3.11), (2.22)*, (3.13), (2.24) we have dG* =
(3.22)
+
[
= Q*G* + G* Q* ,
(3.23)
dA* = (dco*)G* - codG* = - A*'Q* ,
so that du* = - A*(lQ* + Q)e - A*da .
(3.24)
By using equations (3.17), (3.21), (3.24) and the relation (3.25)
^(K*"-1)
= (n -
l)(du*)u*"-2,
we obtain (3.26)
d0=[d
+ Y('Q + Q) + h^eu*"-1 + Ydau*n-1-(n-l)rA*dau*n-2
- {n - V)rA*QQ* + Q)eu*n~i .
On the other hand, from the second expression of equation (3.17) we have (3.27)
dO ={d<j>~)ex A - A « « + ( - l ) " _ 1 M « i A • • • A e») .
Since the frames ex • • • en are of measure 1, d{ex A • • • A en) contains no terms in ex A • • • A en in consequence of (3.3). Comparison of equation (3.26) with equaton (3.27) gives
18
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
(3.2o)
(d&e* A • • • A «» = Oeu*"-1 + h'deu*"-1 + Y[('Q + Q)eu* - (n - l)eA*(tQ* + Q)e]u*n-2.
By putting w — «*, f = 'A*, y = lQ* and y = Q in equation (1.8) in turn we can reduce equation (3.28) to (3.29)
(dfier A • • • A en = [0 + h'0 + YQQ - ^ * ) ] e « * " - 1 .
Thus application of the integral formula (3.19) yields that for a pair of compact immersed submanifolds x(M), x*(M) under a volume-preserving diffeomorphism / in a Riemannian manifold R, on which there is a vector field £, {the coefficient of ex A • • • A en in (3.30)
i [6 + hle + YQQ - tQ*)}eu*n-1} = 0 .
By the same method as above we can also derive an integral formula for two compact immersed hyper surf aces x(M), x*(M) in the above Riemannian manifold R under a volume-preserving diffeomorphism /. For this purpose we shall assume m = 1 in the remainder of this section, and introduce the following additional matrices: v = lde,
(3.31) (3.32)
¥ =
IT*"-
1
v* =
c
0*e,
= cpei A • • • A en .
As before, for a compact manifold M Stokes' theorem gives (3.33)
f
On the other hand, from the second equation of (2.22) we have (3.34)
dO = Q A 6 ,
so that by means of equations (3.31), (2.22), (3.34)* a calculation similar to the one above for du* yields (3.35)
dv* = -
£
0*Cfl* + Q)e -
'8*6en+1.
Using equations (3.20), (3.21) and the relation (3.25) for »*, from equation (3.32) we obtain (3.36)
dW = [d + Y('Q + Q) + y^Wev**1-1 + Yde^v*7"1
- (n - \yo*{*Q* + Q)ev*n-*
- (n - 1M'0* A 0)en+1v*'->.
ISOMETRIES OF COMPACT SUB MANIFOLDS
19
Since d(ex A • • • A en) contains no terms in e1 A • • • A en in consequence of equation (3.3), equating the terms in ex A • • • A e„ on both sides of equation (3.36) we thus have (3 37)
{d
A
y d)ev n 1 ' ' ' A ' " = {° + ^ *~ + Y[('Q + Q)ev* - (n - lye^Q* + Q)e]v*n~* .
By putting w = »*, ? = td*, q = lQ* aud rj = Q m equation (1.8) in turn we can reduce equation (3.37) to (3.38)
(d
Thus application of the integral formula (3.33) yields that for a pair of compact immersed hyper surf aces x(M), x*(M) under a volume-preserving diffeomorphism / in a Riemannian manifold R, on which there is a vector field £,
(3.39)
I {the coefficient of et A • • • A en in ^ [^ + y» + i'0 + 5 T 0 - ' f l * ) ] ^ " - 1 } = o . 4.
Theorems
Let M and M* be two ^-dimensional (n > 2) C2-Riemannian manifolds with fundamental tensors G and G* respectively, and / : M —> M* a ^ - m a p ping. Then on M there are two connexions; the Levi-Civita connexion defined by its Riemannian metric and the connexion induced by the mapping f from the Levi-Civita connexion of M*. The difference of these two connexions is a tensor field A of contravariant order 1 and covariant order 2, and the construction in §1 gives a vector field AG. / i s called an isometry if G* = G, and an almost isometry relative to G if Aa = 0. It is clear that an isometry is also an almost isometry, since in this case A = 0. Theorem 4.1. Let M be a differentiable manifold of dimension n > 2, x, x* : M —> R two immersed compact submanifolds, w/f/i fundamental tensors G, G* respectively1, in a Riemannian manifold R, which admits an enlarging infinitesimal nonisometric conformal transformation f, and f: x(M) —> x*(M) a volume-preserving diffeomorphism. Then f is an isometry, if it is an almost isometry relative to G*, that is, (4.1)
Aa, =
0,
and the following inequality holds at every p e M : 1 It should be remarked that in §3 we have also used G, G* to denote the matrices of these two tensors.
20
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
(4.2)
QG < Q0> ,
where Q is the coefficient tensor of the quadratic form
x = l^i, • • -, 7rJ|
is a one-rowed matrix of linear differential forms on M. By using equations (2.21), (3.14), (3.13) we have (4.4)
xeu*"-1 = ( 2 T T ^ X S (ofej)-1 j * = ( L *ii-••<„*«! A o»g A • • • A O i
A • • • A en ,
where etl...in is equal to + 1 or — 1 according as iu • • •, in form an even or odd permutation of 1, • • •, n, and is equal to 0 otherwise, and the summation is extended over all iu • • •, in from 1 to n. Assuming (4.5)
Ki = hij(oJ
(4.6)
H =
,
\\hiJ\\,
recalling cof = gfjO)J, \gfj\ = 1, and using elementary properties of determinants and equations (3.1), (1.2), (1.3) we can obtain (4.7)
TTCK*"-1
=
R!
He,(ei A • • • A en)dV .
On the other hand, from equations (1.2), (1.3) we have, in consequence of equation (2.19) for the submanifold x(M),
(4.8)
i = -^(Fih In
+ V£d8*i3
= pGG, ,
where p > 0. Thus substitution of n = 0 and n = Wd in equation (4.7) in turn and use of equations (3.18), (4.8), (3.7), (3.9) give immediately (4.9)
deu*"-1 = n\ PGa,(ei A • • • A en)dV ,
ISOMETRIES OF COMPACT SUBMANIFOLDS
(4.10)
21
Wdeu*71-1 = n\ gG„(*i A • • • A en)dV .
Since m = a>*, from the first equations of (2.24), (2.24)* we have (4.11)
a A (0 - S*) = 0 ,
so that we can write m\ - wfk = a\y
(4.12)
.
Equations (4.11), (4.12) imply (4.13)
a*,a>« A « ' = 0 1
which gives the symmetry of a\s in the subscripts i, j , that is, (4.14)
a*, = a% .
From the properties of the forms Q, Q* and the definition of the tensor A defined at the beginning of this section it follows that the components of d are a*y, so that we can write
(4.15)
J* = K-1 .
A use of equations (2.23), (2.23)*, (3.15), (4.12) yields readily the matrix (4.16)
Y('Q - (J2*) = |y»afX> • • •, yXX'll .
By putting H = ykAk in equation (4.7) we obtain (4.17)
YQO - 'Q^eu*"-1 = n\ ykA%{ex A • • • A en)dV = 0,
since A%. = 0 due to condition (4.1). Thus equations (4.9), (4.10), (4.17) reduce the integral formula (3.30) to (4.18)
j(PGa,
+ Qa,)dV = 0 .
M
In particular, when the two submanifolds are identical, by definition G0, = 1 and the formula (4.18) becomes (4.19)
j(p + Q6)dV = 0 .
Subtracting equation (4.19) from equation (4.18) we obtain (4.20)
[[p(Ga. - 1) + (Q0. - Q0)]dV = 0 .
22
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
Since p > 0 by the condition of the enlarging infinitesimal conformal transformation £, and Gg, > 1 by Garding's inequality (1.5), from the condition (4.2) of our theorem it follows that the integrand in (4.20) is nonnegative, and therefore that (4.21)
GG, = 1 .
By applying Lemma 1.1 of Garding again, equation (4.21) holds only when G — cG* for a certain c. Since G and G* are positive definite with \G\ = | G* | = 1, we must have c = 1 and therefore G = G*, which implies that / is an isometry. Formula (4.19) gives Corollary 4.1. Suppose that a Riemannian manifold R admits an enlarging infinitesimal nonisometric conformal transformation. Then any complete minimal submanifold of R is noncompact. In particular, when R is a Euclidean space, this corollary is due to Chern and Hsiung [1]. By the same method as above we can have Theorem 4.2. Let M be a differentiable manifold of dimension n>2, x, x*: M —> R two immersed compact hypersurfaces with fundamental tensors G, G* and positive definite second fundamental forms JI, II*, whose coefficient tensors are denoted by A, A* respectively, in a Riemannian manifold R which admits an enlarging infinitesimal nonisometric conformal transformation £, and f: x(M) —> x*(M) a volume-preserving diffeomorphism. Then II — II*, if f is an almost isometry relative to A*, that is, (4.22)
AA. = 0 ,
and the following conditions are satisfied: (4.23)
1^1 = 1^*1,
(4.24)
GA, > GA ,
(4.25)
yn+1 =
where g1 is the orthogonal projection of the vector £ onto the direction along the unit normal vector en+1 of x{M) at the point x(p). In particular, when R is a Euclidean space, this theorem is due to Hsiung [3]. Proof. As in the proof of Theorem 4.1, we observe that each term on the right side of equation (3.38) is of the same type as the form nev*71'1, where JC is defined by equation (4.3). By equations (2.21), (3.31), (2.23)* we have Kev n 1
* ~ = ( 2 *««i)(E o.r+1«y)"_1
(4.26)
*
J
= ( 2 eii-i»Tii A < B + 1 A • • • A o>t:+1)ex A • • • A en .
23
ISOMETRIES OF COMPACT SUBMANIFOLDS
Making use of equations (4.5), (4.6), (2.26)*, (2.30)*, (3.1), (1.2), (1.3) and elementary properties of determinants, from equation (4.26) we can obtain Ttev*"-1 = n\HA,\A*\(e1A---A
(4.27)
en)dV .
Thus substitution of n = Q and -K = yn+id in equation (4.27) in turn and use of equations (3.18), (3.7), (3.9), (2.30) and the equation obtained from equation (4.8) by replacing G* by A* give immediately (4.28) (4.29)
Oev*"-1 = n! pGAt\A*\(eiA---A 1
y^Oev*"-
en)dV ,
= n! yn+1AA,\A*\(e1A---A
en)dV .
On the other hand, by the same argument as that used in deriving equation (4.17) we have (4.30)
YQQ - 'Q^ev*71-1 = n! ykAkA,\A*\(Je1A--A
en)dV = 0 ,
since JA. = 0 due to condition (4.22). Thus equations (4.28), (4.29), (4.30) reduce the integral formula (3.39) to (4.31)
J(PGA,
+ yn+1AAt)\A*\dV
= 0.
M
In particular, when the two hypersurfaces x(M), **(M) are identical, AA, = 1 and the formula (4.31) becomes (4.32)
j(PGA
+ yn+1)\A\dV
= 0.
M
Subtracting equation (4.32) from equation (4.31) and using condition (4.23) we obtain (4.33)
J[P(GA, - GA) + yn+1(AAt -l)]\A\dV
= 0.
M
Since p > 0, and AA, > 1 by Garding's inequality (1.5), from the conditions (4.24), (4.25) of our theorem it follows that the integrand in (4.33) is nonnegative, and therefore that (4.34)
AA.=
1.
By applying Lemma 1.1 of Garding again, equation (4.34) holds only when A = cA* for a certain c. Since by assumption A and A* are positive definite with equal determinants, we must have c = 1 and therefore A = A*, which implies that U = 77*.
320
24
CHUAN-CHIH HSIUNG & BURGESS HAROLD RHODES
Bibliography [ 1 ] S. S. Chern & C. C. Hsiung, On the isometry of compact submanifolds in Euclidean space, Math. Ann. 149 (1963) 278-285. [2] L. Garding, An inequality for hyperbolic polynomials, J. Math. Mech. 8 (1959) 957965. [ 3 ] C. C. Hsiung, On the congruence of hypersurfaces, Atti Accad. Naz. Lincei Rend. CI. Sci. Fis. Mat. Natur. (8) 37 (1964) 258-266. [ 4 ] B. H. Rhodes, On some inequalities of Garding, Acad. Roy. Belg. Bull. CI. Sci. (5) 52 (1966) 594-599. LEHIGH
UNIVERSITY
J . DIFFERENTIAL GEOMETRY 2 (1968) 185-190
ON THE GROUP OF CONFORMAL TRANSFORMATIONS OF A COMPACT RIEMANNIAN MANIFOLD. Ill CHUAN-CHIH HSIUNG
1.
Introduction
Let gi}, Rhijk, Rtj = Rkijk be respectively the metric, Riemann and Ricci tensors of a Riemannian manifold Mn of dimension n, and denote (l.D
P = RWRMjk,
Q =
RiWij.
Throughout this paper all Latin indices take the values 1, • • •, n unless stated otherwise, and repeated indices imply summation. In a recent paper [2] the author proved Theorem 1. Suppose that a compact Riemannian manifold Mn(n>2) with constant scalar curvature R = gtjRa admits an infinitesimal nonhomothetic conformal transformation v, and let Lv be the operator of the infinitesimal transformation v. If (1.2)
a?LvP + b(2a + nb)LvQ = const.,
where a and b are constants such that (1.3)
c
EE
4a2 + 2(n - 2)ab + n(n - 2)b* > 0 ,
then Mn is isometric to a sphere. In particular, when a = 0 or b = 0, Theorem 1 is reduced to a result of Yano [4], which is a generalization of some results of Lichnerowicz [3] and the author [1]. Yano pointed out that condition (1.3) is equivalent to that a and b are not both zero. Very recently, Yano and Sawaki [5] obtained the following theorem similar to Theorem 1: Theorem 2. Suppose that a compact Riemannian manifold Mn (n > 2 ) with constant R admits an infinitesimal nonhomothetic conformal transformation v. // (1.4)
LvLv[(n - 2)a*P + 4b(2a + b)Q] < 0 ,
Communicated November 28, 1967. Supported partially by the National Science Foundation grant GP-7513.
322 186
CHUAN-CHIH HSIUNG
where a and b are constants such that a + b =£ 0, then Mn is isometric to a sphere. The purpose of this paper is to generalize both Theorems 1 and 2 to Theorem 3. Suppose that a compact Riemannian manifold Mn (n > 2 ) with constant R admits an infinitesimal nonhomothetic conformal transformation v. If (1.5)
+C~_4^0)<0,
LvLv(c?P
where c
= 4a* + (n-
2)[2a £ bt + (b, - b2 + bz - bt + b, - bef
(1.6) - 2bA - 2b2b4 + 2bbb6 + (n - 1) Z bU > 0 , a and b's being constants, then Mn is isometric to a sphere. It is obvious that Theorem 3 is reduced to a generalization of Theorem 1 when b2 = • • • = be = 0 , and to Theorem 2 when b1=
...=bt
= b/(n - 2),
K = b6 = 0 .
We need the following theorem of Yano [5] to prove Theorem 3 : Theorem 4. Suppose that a compact orientable Riemannian manifold Mn (n > 2) with constant R admits an infinitesimal nonhomothetic conformal transformation v so that (1.7)
Lvgij = 2^>gtj,
and let V denote the operator of covariant derivation of Mn with respect to gij. If
(1.8)
fTfjf^dA>0,
where (1.9)
Tij = Rij
gtj, n 0* = V1^ — gijV$, and dA is the element of area of Mn at a point, then Mn is isometric to a sphere. 2. n
Lemmas
Throughout this section M will always denote a compact orientable Riemannian manifold of dimension « > 2 . Let d be the Laplace-Beltrami operator on Mn. Then, for any scalar field / on Mn,
GROUP OF CONFORMAL TRANSFORMATIONS
187
Af=-ViF1f.
(2.1) Thus we have (2.2)
f J / dA = 0 Mn
from the well-known Green's formula:
(2.3)
JV'£«
where £4 is any vector field on Mn. Lemma 1. // a nonconstant scalar field
= 2 U<j>A<j> -
ftfddA
(2.4)
= 2 f (kp - p$t)dA , which gives Lemma 1 immediately. Lemma 2. Let v be an infinitesimal conformal transformation on Mn so that (2.5)
Lvgu =
2^gi}.
Then ,„ ,. (2.6)
LvRhijk
(2.7)
= 2<j>Rhijk — ghkFj^t + ghjFk
LvRi}
(2.8)
— gt/k
LVR = 2(« - 1)J0 - 2R$ .
Lemma 2 can be proved by a straightforward computation. Lemma 3. If Mn has constant R and admits an infinitesimal nonhomothetic conformal transformation v so that (1.7) holds, then (2.9) (2.10)
fy
=
Rf/(n-l), R>0.
188
CHUAN-CHIH HSIUNG
Equation (2.9) follows from equation (2.8) due to the constancy of R, and equation (2.10) from Lemma 1. Lemma 4 (Yano and Sawaki [5]). // Mn admits an infinitesimal conformal transformation v so that equation (2.5) holds, then, for any scalar field f on Mn, (2.11) Proof. (2.12)
C6fdA
=
-—CLvfdA.
Substituting fvt for f4 in the Green's formula (2.3) we obtain fWvt
dA = -
Mn
CvtF*f dA=-
f LJ dA .
Mn
Mn
On the other hand, since Lvgij = FtVj + V}Vi, from equation (2.5) we have FlVi = nj>, which and equation (2.12) yield the required equation (2.11) immediately. 3.
Proof of Theorem 3
On the manifold Mn consider the covariant tensor field of order 4 : (3.1)
Whi,t
=
aThiJk
+ bl8hkTu
KghjTik
~
— b&ikThj + b5ghiTjk
—
+
b&iThk
b(gjkThi,
where (3.2)
Tfrij/c = R-hijk — —;
—KgijShk — SikShj) ,
n(n — 1) and a and b's are constants satisfying (1.6). Then (3.3) W^Whm = a*P + 1=** Q - 1 [ " . + ±=*f\R>, n—2 n \n — 1 n —2 I where c is defined by (1.6). From equation (3.3) it follows that (3.4)
Lv(W^Whijk)
= Lv [&P + c~^*
Q
By assuming the infinitesimal nonhomothetic conformal transformation v to
GROUP OF CONFORMAL TRANSFORMATIONS
189
be defined by (1.7), from equations (3.1), (3.2), (1.9), (2.5), (2.6), (2.7), (2.9) we can easily obtain LvWMjk
= 2a$RMjk
- [a + (n -
+ [a + (n - 2)bjgh/k^i
- [a + in - 2)b^gi}V k(j>h
+ [a + {n - 2)b^gikV^h
- (n - 2)b^ghiVk^j + (n -
- -J*—gijgKt[4a n(n — 1) (3.5)
2)b^ghkV^
fR gttg n{n — 1)
+
" ~~
n(n — 1)
+ On - 4)(fo, + M [4a +
(3„ _ 4)(fo2 + fcj]
b6
— 2btfj)ghjRik
2)b^jkV4h
+ 2b^>gijRhk
n
~
be
n(n — 1)
— 2bi(j>glkRhj
+ 2bi(j>ghiRjk
—2btff>gjkRki .
Multiplying both sides of equation (3.5) by Whiik and making use of equations (3.1), (3.2), (1.9), (3.3), (1.6) and R\jk = 0, an elementary but lengthy calculation yields W^kLvWhijk
(3.6)
= 2<j>WM^Whijk - cTWjfc
.
By substituting equation (3.6) in the well-known formula (3.7)
Lv(W^*WMjk)
+ 2WM'*LvWMJt
-
tyWM»>WMjt,
we thus have (3.8)
W
M
% « J = -4fiWWWMjk - IcfTWfa .
Since the manifold Mn is of constant R, it is known that fo n < 2 (3.9)
P'Rij = 0 ,
and therefore (3.10)
FiTiJ = 0.
Thus (3.11)
rj(.TtJM*) = Tttfp
+ 4>TuVifi .
Without loss of generality we may assume our manifold Mn to be orientable, since otherwise we need only to take an orientable twofold covering space of
190
CHUAN-CHIH HS1UNG
Mn. Substituting equation (3.8) in equation (3.11), integrating the resulting equation over the manifold Mn and using equation (2.3) we obtain 2c j TtrfifidA (3.12)
**
^Lv{WM^WMjlc)dA
= n
*
+ 4(ftWWWMjkdA
.
On the right side of equation (3.12), the second integral is nonnegative since its integrand is so, and the first integral is equal to, by Lemma 4 and equations (3.3), (1.5), - 1 (LvLv(W^«WMjk)dA
= - 1 {LvLv[a'P
+
C 4
~ fQ]dA
> 0.
Hence the integral on the left side of equation (3.12) is nonnegative, and Theorem 3 follows from Theorem 4 immediately. References [ 1 ] C. C. Hsiung, On the group of conformal transformations of a compact Riemannian manifold, Proc. Nat. Acad. Sci. U.S.A. 54 (1965) 1509-1513. [2] , On the group of conformal transformations of a compact Riemannian manifold. II, Duke Math. J. 34 (1967) 337-341. [ 3 ] A. Lichnerowicz, Sur les transformations conformes d'une variete riemannienne compacte, C. R. Acad. Sci. Paris 259 (1964) 697-700. [ 4 ] K. Yano, On Riemannian manifolds with constant scalar curvature admitting a conformal transformation group, Proc. Nat. Acad. Sci. U.S.A. 55 (1966) 472-476. [ 5 ] K. Yano & S. Sawaki, Riemannian manifolds admitting a conformal transformation group, J. Differential Geometry 2 (1968) 185-190. LEHIGH
UNIVERSITY
Math. Z. 115, 235-251 (1970) © by Springer-Verlag 1970
Submanifolds of Spheres SIHAM BRAIDI and CHUAN-CHIH HSIUNG*
1. Introduction Recently, Simons [3] made an important contribution to the study of minimal submanifolds immersed in a Riemannian manifold by using the derivation of the linear elliptic second order differential equation satisfied by the second fundamental form of each minimal submanifold. By applying an integral inequality of Simons, Chern, do Carmo and Kobayashi [1] jointly obtained two basic theorems for determining all n-dimensional submanifolds M minimally immersed in a unit sphere S"+p of dimension n+p such that the square of the length of the second fundamental form A of M is equal to
\\A\\2 = n/(2-jY
(1.1)
Nomizu and Smyth [2] jointly studied complete hypersurfaces M of S n+1 such that M has constant mean curvature and nonnegative sectional curvature, and satisfies the condition (1.1). The purpose of this paper is to continue the joint work of Chern, do Carmo and Kobayashi by extending their two basic theorems to compact oriented n-dimensional immersed submanifolds M of SB+P satisfying a certain condition on the second fundamental form A of M, which reduces to the condition (1.1) when M is minimally immersed in S"+p. In § 2 we compute the Laplacian of the second fundamental form of a general immersed submanifold of a locally symmetric space. In § 3 we give the basic inequality for the traces of symmetric matrices used by Chern, do Carmo and Kobayashi, and extend the integral inequality, mentioned above, and another one of Simons to compact oriented submanifolds immersed in a space of constant sectional curvature. § 4, the last section, contains the main theorems of this paper. 2. Laplacian of the Second Fundamental Form In this section we shall compute the Laplacian of the second fundamental form of a submanifold immersed in a locally symmetric space. Let M be an n-dimensional Riemannian manifold immersed in an (n + p)dimensional Riemannian manifold N. Choose a local field of orthonormal frames c 1; ...,en+p in N such that, restricted to M, the vectors et, ...,en are * The work of the second author was partially supported by NSF grants GP-7513, GP-11965.
236
S. Braidi and Chuan-Chih Hsiung:
tangent to M (and, consequently, the remaining vectors e „ + 1 , . . . , en+p are normal to M). Unless stated otherwise, we shall make of the following convention on the ranges of indices: l^A,B,
C,... ^n + p,
l^i,j,k,...
gn,
n+l^ce,P,y,...
^n + p,
and when a letter appears in any term as a subscript and a superscript, it is understood that this letter is summed over its range. Let co1, ...,con+p be the coframe field dual to e1,...,e„+p chosen above. Then the structure equations of N are given by dcoA=-coAAcoB,
a)i + ooBA = Q>,
(2.1)
A
dcoi = -corAco% + 4> ,
KABCD + KABDC = 0.
(2.2)
Restriction of these forms to M gives (2.3)
Since 0=dcox = -co" ACO', by Cartan's lemma we may write a)f = /i?X.
Htj = hJt-
(2-4)
From these formulas, we obtain dco1 = - co) A CQ>, co) + co{ = 0, dm) = - co\ A co) + Q),
(2.5) k
l
Q) = $ R'jkl co Aco ,
(2.6)
R'JHI = K'Jkl + £ (hU hjt - h\x h)k),
(2.7)
a
dco«p=-co°yACo} + Q*fi, Crf = \R%ua?K(ol,
(2.8)
R'w = K'pu + Y1
(2-9)
i
The forms (coj) define the Riemannian connection of M or, more precisely, a connection in the tangent bundle T= T(M) of M (and therefore also a connection in the cotangent bundle T* = T* (M) of M), the forms (coJS) define a connection in the normal bundle T1 = TL (M) of M, and hfj co' co1 ea is called the second fundamental form of the immersed manifold M in N. Sometimes the second fundamental form is denoted by its componentsft?-.— V hf, ex is called n i
the mean curvature vector or the mean curvature normal, and an immersion is said to be minimal if its mean curvature vector vanishes identically, i.e., if V A?, = 0 for all a.
Submanifolds of Spheres
237
By taking the exterior differential of (2.4) and defining hxjk by h}Jk cok=dhfj - hft to) - hfj co\ + hfj co£.
( 2 - 10 )
we obtain (^+|K"ij,)^A(»i=0! Kjk — h*kj = K*ikj =
—
(2.11)
K'ijk •
(2.12)
Similarly, by taking the exterior differential of (2.10) and defining hfJkl by hl}kl col = dh?jk-hljk to\-h?lk co'j-hfj! tok + h?jk to},
(2.13)
(h"jki ~ \ Km Rmm ~ i Kj Rmm + J K R%kl) cokAcol = 0,
(2.14)
we have X
m
hfjki ~ h jik = hfm R
Jki
m
+ hmj R iki — n;j- R'/IM.
(2.15)
With respect to the connection in the bundles T* T defined above, we have a covariant differentiation which maps a section of T 1 (g) r * ® - - - ® T* into a k
section of T 1 ® T* (g) • • • (g) T*. The second fundamental form /if- is a section of fc+T the vector bundle T 1 (g) T* (g) T*, and fcyk is the covariant derivative of h^. Similarly, h*Jkl is the covariant derivative ofhaJk. We may also consider -K01^, as a section of the bundle TL®T*®T*
(2.16)
This covariant derivative of Kxijk must be distinguished from the covariant derivative of KABCD as a curvature tensor of N, which will be denoted by KABCD.E. Restricted to M, K1^^ is given by «"«*;« = * V « - x "w* *ft K*w h% ~ K*m hii + * % K* •
(2-17)
In this section, we shall assume that N is locally symmetric, i.e., KABCD.E = 0. The Laplacian A hfj of the second fundamental form hfj is defined by Wj^Kkk-
(2.18)
k
Taking the covariant derivative of (2.12) and substituting the resulting equation in (2.18), we obtain A^=YjKkjk-^mk k
k
k
k
(2.19)
On the other hand, from (2.15) we have Km—Kikj+Km
Rmijk+Kii
Rmkjk—Hi
Wpjk •
(2.20)
238
S. Braidi and Chuan-Chih Hsiung:
We first replace hlikj in (2.20) by hlkij-KxkikJ according to (2.12), and then substitute the right side of (2.20) for hiijk in (2.19) so that ^ hij = E C»«y- K\ikj~ K'Ukt + Km Rmijk + hTml R\jk -fefcR'fi]k).
(2.21)
From (2.7), (2.9), (2.17) and (2.21), it follows * K = E Wkij-K"lJI, htk + 2K'fitt h<jk-K*kl)k hfj + lK'w
hi,
k
+ Kmkik h*mj + KmkJk h^ + 2 KmiJk h^,k) + E (h*mihljhlk +
(2.22)
2himhk
P,m,k
-KtWmkHj-KjhlM, and hence we obtain *,Uj,k
• K'm
h
*j hh+2Kmktk
Kj Mj + 2KmiJk h*mk hi)
(2.23)
a, P, i, j , k, I
+htjKlhfjhil-hiKihlJhfl]. 3. Integral Formulas In this section we assume that the ambient space N in § 2 is of constant sectional curvature c. Then KABCD = c(SAC3BD-dADSBC), (3.1) where 8AC are the Kronecker deltas, and (2.23) reduces to ZkljAk}j= a, i,j
nc^hfjhfj-c^iZhti)2
E hfjh^j + a, i, j , k
+
ct,i,j
E
a
i
(KtikKM-kijhfjhith&d
(3.2)
at,fi,i,j,k,l
-
E
WkWj-hf.hijWhfj-hfihtj).
x,P,i,j,k,l
(3.2) was obtained jointly by Chern, do Carmo and Kobayashi [1] for a minimal M in N. For each a, let Hx denote the symmetric matrix (h-j). Then (3.2) can be written as 1K^J= E htjh>kkiJ + Yi[ncTrH];-c(TrHa)2l a, i,j
a, i,j,
fc
a
+ E {Tr(tfa ff, -Hf Hf - [Tr(ff. H„)]2
(3 . 3)
+ TrJJ,Tr(ff« # , # . ) } , where Tr f^ denotes the trace of the matrix U\. (3.3) was obtained by Nomizu and Smyth [2] for a hypersurface M, i.e., for p= 1.
Submanifolds of Spheres
239
Now set
(3.4) (3.5) (3.6)
so that S^p is a symmetric (p x p)-matrix and can be assumed to be diagonal for a suitable choice of en+1,..., e„+p, and S is the square of the length of the second fundamental form h"j of M. In general, for a symmetric matrix A = (atj) we have JrA2 = Z(atj)2.
(3.7)
Ui
Clearly, T r A 2 = T r ( T - 1 ^ T ) 2 for any orthogonal matrix T. Thus (3.3) can be written in the form: tx,i,j,k
a
+ XTr(tf a ^-tf,H a ) 2
(3.8)
«./>
- c X (TrffJ 2 + X Tr ^ Tr (if. ff, H J . For later developments we need the following algebraic lemma. Lemma 3.1. Let A and B be symmetric (n x n)-matrices. Then -Tr(AB-BA)2^2TrA2TrB2,
(3.9)
and the equality holds for nonzero matrices A and B if and only if A and B can be transformed simultaneously by an orthogonal matrix into scalar multiples of A and B respectively, where '1 B= | ° -
0 1
|.
(3.10)
0/ \ 0 0/ Moreover, if Ay, A2, A3 are symmetric (n x n)-matrices such that -Tr(AxAll-Al)Ax)2
= 2TrAJTrB2,
l g a , / J £ 3 , a*j8,
(3.11)
then at least one of the matrices Ax must be zero. Proof. We may assume that B is diagonal, and denote its diagonal entries by bu ...,b„. Suppose that A = (aij). Then by a simple calculation we obtain -Tr(AB-BA)2=-ZaiJ(bj-bi)aji(bi-bj)
= Za2j(bi-bj)2^2YiaUbf + b ^2^a2jYb2k=2TrA2JrB2. i.j
k
(3.12)
332 240
S. Braidi and Chuan-Chih Hsiung:
Now assume that A and B are nonzero matrices and that the equality holds in (3.9). Then the equality must hold everywhere in (3.12). From the fourth equality in (3.12) it follows that au = 0,
i=l,...,n,
(3.13)
and from the third equality in (3.12) it follows that bi + bj=0,
if a o + 0.
(3.14)
Since A is nonzero and satisfies the condition (3.12), at least one of atJ (ij=j) is nonzero, and therefore without loss of generality we may assume that a12 + 0. Thus we obtain, from (3.14), that b1 = —b2, and, from the fourth equality in (3.12), that b3 = --- = b„=0. Since B =1=0, we must have by = — ft2=t=0, and can conclude, from the fourth equality in (3.12), that atJ=0 for (i,j) + (l, 2). To prove the last statement of the lemma, since the three nonzero symmetric matrices AUA2,A3 satisfy (3.11), from the second statement, which we have just proved, we can see that one of these matrices can be transformed to a scalar multiple of A as well as to a scalar multiple of B by orthogonal matrices. But this is impossible, since A and B are not orthogonally equivalent, q.e.d. By applying Lemma 3.1, we obtain - XTr(ff.tf,-iJ / ,ff a ) 2 + £ S a 2 - n c S ^ZStSp
+
ZSt-ncS (3.15)
a
<x
= p2 al+p(p-\)a2
— ncS,
where Pffi = I X = s>
L ) 0-2= Z 5 « s *
(3.16)
£(Sa-S,)2£0,
(3.17)
It can be easily seen that p2{p-l)(a2-a2)= and therefore that p2al+p{p-\)a2=p{2p-l)(jl-p(p-\)(al-a2)
^p{2p-\)o2 = {2
)s 2 .
(3.1»)
From (3.8), (3.15) and (3.18), we thus have -YttftjAHl^W,a, i,j
£ hljkluj, a, i,j, k
(3.19)
Submanifolds of Spheres
241
where we have put Wp=\(2--\
S-nc]
L\
PI
S + c^iTrHf-YTrH.TT^HpHJ. J
(3.20)
x,p
Theorem 3.1*. Let M be a compact oriented n-dimensional manifold immersed in an (n+p)-dimensional Riemannian space N of constant sectional curvature c. Then J[^-Z(TrHa)J(TrHa)]*1^0, (3.21) M
x
where *1 denotes the volume element of M. Lemma 3.2. If M is a compact oriented n-dimensional manifold immersed in an (n+p)-dimensional Riemannian manifold N, then f X hfjAMj *1= - J Af <x,i,j
M
£
[htjkf *1 £ 0 .
(3.22)
a,i,j,k
Proof of Lemma 3.2. From the defining Eq. (2.18) of the Laplacian, we have
Integration of (3.23) over M and application of Stokes' theorem to the left side yield immediately that the integral of the left side and hence that of the right side vanish. Proof of Theorem 3.1. By taking covariant differentiation with respect to the connection in the bundles T*, T defined in §2 and using (2.12), (2.18) we obtain WjKkdj=Htjjhlkt+h}jhiklj, (3.24) I
WjjhU)^
£
z,i,j,k
(h^h^+htjjht.d
a,i,j,k
= I
(HjjiiWkk + HljjHkkd
= Z(TrHJA(TrHx)+ a.
£
(3.25) hfjjh^.
a,i,j,k
Integration of (3.24), (3.25) over M and application of Stokes' theorem to the left side yield 1 I htJ WkkiJ *1= J D (Tr Hx) A (Tr HJ *1. (3.26) M
a,i,j,k
M
a
1 The authors wish to thank Morio Obata for his suggestion in regard to the present general form of this theorem.
242
S. Braidi and Chuan-Chih Hsiung:
Thus from (3.19), (3.22) and (3.26) it follows immediately K^-ZfTrHJzKTrtf,)]*! M
(3.27)
a
M
a, i,j
M
tz,i,j,k
which is just (3.21). Theorem 3.1 is due to Simons [3] for minimal M in N, but the proof given above is an extension of that given jointly by Chern, do Carmo and Kobayashi [1] for Simons' case. Corollary. Under the same conditions as in Theorem 1, if ^-ECTrffJ/lfTrfggO a
everywhere on M, then Wp — £ (Tr HJ A (Tr HJ = 0 everywhere on M. a
It should be noted that if M is minimally immersed in N and not totally geodesic, then S * 0 and Tr Hx=0 for all a, and therefore Wp - £ (Tr HJ zl (Tr HJ = 0 is reduced to S = nc {2
1. Moreover, for p = l, Wp can be expressed in
terms of the first three mean curvatures of the hypersurface M in the space N. For this purpose let St, i=l,..., n, be the i-th elementary function of the eigenvalues kx,...,Xn of the matrix H„+1. Then the i-th mean curvature Mt of the hypersurface M at a point x is defined by (») « , - * . and we can easily obtain
= (S^-2S2){Sl-2S2-nc)
+ cSf-S1(Sf
- 3 S 1 S 2 + 3S3)
= - S 2 S 2 + 4S| + c ( l - n ) S 2 + 2ncS 2 -3S 1 S 3
(3.28)
" ( " ^ [-nM 1 2 M 2 + 2(n-l)M 2 2 -2cM 2 + 2 c M 2 - ( n - 2 ) M 1 M 3 ] . Now letfe(x) be the average of all the sectional curvatures of an n-dimensional Riemannian manifold M at a point x. Then n(n — 1) k(x) is the scalar curvature of M at x, and we have, in orthonormal local coordinates, k(x)=-—i-—5i>VkRiJkl. n(n — 1)
(3.29)
243
Submamfolds of Spheres
Theorem 3.2. Let M be an n-dimensional manifold immersed in an (n + p)dimensional space N of constant sectional curvature c. Then, for xeM, 1 k(x) = c + n(w-l)
E(TrtfJ 2 -S].
Proof. By using (2.7) and (3.1) we obtain, from (3.29), k{x)=-
1 5il 5* [c (5ik 8j, - 5tl 5jk) + X (/& h)x - *?, h%)] n(n — 1) a 1 n(n — 1) {cn(l-n)+ £ [ ( « « / - n ^ ] }
= c+
1 2 -S]. n ( n - l ) LE(TrHJ t
Theorem 3.3. Let M be a compact oriented n-dimensional manifold immersed in an (n + p)-dimensional space N with constant sectional curvature c. Then the scalar curvature of M satisfies the inequalities k(x)^c +
n(n-l)
for every xeM,
lc-k{x)+ / „ |c-fc(x)+ . K „ \ n(n-1)/ n(n-l)
,
c K - X T r H , Tr(H aff„H J - £ (Tr H J A(Tr tfa)
+
(2-1) „ > - D2
where K = ^ (Tr H a ) 2 . Proo/ Theorem 3.3 follows immediately from Theorems 3.1 and 3.2. Theorems 3.2 and 3.3 were obtained by Simons [3] for the case where c = 1 and M is minimally immersed in JV.
4. Main Theorems Throughout this section we shall assume that M is a compact oriented n-dimensional manifold immersed in an (n + p)-dimensional space N of constant sectional curvature 1, and that M is not totally geodesic in N and satisfies UK-WrHJA(TrHa)-]*l = 0,
(4.1)
244
S. Braidi and Chuan-Chih Hsiung:
which and (3.27) imply hfjk = 0
for all a,i,j,k.
(4.2)
Therefore by use of (2.18), (4.2) we have Ahfj=0, and the terms on both sides of (3.19) vanish. It follows that all inequalities in (3.15), (3.17) and (3.19) are equalities. Since in (3.15) use was made of the inequality — Tr (Hx Hf — Hp HJ2 ^ 2Tr#2Tr^2,wehave - T r ( H , / / , - # , tfJ2 = 2 T r t f 2 T r H 2 ,
a*/?.
(4.3)
The equality in (3.17) means p(p-l)(
(4.4)
From (4.3) and Lemma 3.1, we conclude that at most two of the matrices Hx are nonzero, in which case they can be assumed to be scalar multiples of A and B defined by (3.10). We now consider the cases p = \ and p ^ 2 separately. Case p — \. We set and choose our frame field in such a way that hu = 0
for i+j.
(4.5)
Lemma 4.1. After a suitable renumbering of the basis elements el5 ...,e„, we have either (i)
h1=---=h„ = constant,
or (ii)
hi=--- = hm = X = constant, l<m
Proof. Setting i=j in (2.10) and noting (4.2), (4.5), we obtain 0=dh„
(4.6)
which shows that ht is constant. Since hiJk = 0and dhu = 0, (2.10) implies 0 = h„ co'j+hu co't^ihi-hj) co), from which it follows that co) •= 0 whenever ht + hj. Thus, if hi + hj, then, by (2.2) a n d (3.1),
.
,
•
•
0 = dco) = — cok A co* — a)'n + 1 A co"+ + co' A o^.
Submamfolds of Spheres
245
The first term on the right side of the above equation vanishes, since colk 4= 0 and a) 4= 0 would imply ht = hj=hk, contradicting the hypothesis. Hence 0 = -C0J, + 1 A CDnj + 1 + COi A ft/ = hik hjf ft/ A CO1 + CO1 A ft)J = (/Z;/l J +l)ftJ i AO)-',
which shows that if ht + hj, then hthj= — 1. Set h^ — X. Then we have either hl = --- = h„ = X, which proves part (i) of our Lemma4.1, or, by renumbering the indices of ex,..., e„ if necessary, h1 = --- = hm = X and hj 4=A for j > m . In the latter case, h± hj = —1 for; = m 4-1, ...,n, and thereforeh m+1 = ••• = hn= —r = M» proving part (ii). q. e. d. In case (i) of Lemma 4.1, M is totally umbilical. It is known that a totally umbilical hypersurface of an (n + l)-sphere £ " + 1 is locally (globally if it is complete) a sphere; in particular, it is a great sphere of £ " + 1 if it is totally geodesic. In case (ii) of Lemma 4.1, the two distributions defined by co1 = • • • = of = 0 and com+1=-- = co" = 0 are both integrable by Frobenius' theorem, and give a local decomposition of M. Then every point of M has a neighborhood U which is a Riemannian product Vt x V2 with dim Vl = m, dim V2 = n — m. From (2.7) we see that the curvatures of Vl and V2 are respectively given by ^
(
= (l + ^
2
)(^^-^^),
2
Rm = (1 + M )($*Sji~St,8jJ,
for \^i,j,k,lSm,
(4.7)
for m +1SUj,Kl£n.
(4.8)
If m ^ 2 (resp. n — m^2), then Vx (resp. V2) is a space of constant sectional curvature 1 + A2 (resp. 1 + /z2). If m = 1 (resp. n — m = 1), then Vx (resp. F2) is a curve and hence also a space of constant curvature. Now let h denote n times the constant mean curvature of M. Then h = YJhi = mX + {n — m)n.
(4.9)
i
Since in this case the condition W^ holds automatically, Eq. (4.9) and the relation X \x — — 1 give h + -\/h2 +4m(n — m) -*= s . 2m or A—
Replacing en+1 we hence have
/i-i//i2+4m(n-mj J«= ^h^ : , 2(n — m)
,,. ,„, (4.10)
h — i/h2 + 4m{n — m) _ h + -\/h2 + 4m(n — m) r , /i = — . 2m 2{n — m) by — e„ +1 if necessary, we may assume (4.10). In summary,
Theorem 4.1. Let M be a compact oriented immersed hypersurface satisfying J \_W1 — (Tr Hn+l)A(Tr Hn+1J] *1 = 0 in an (n + l)-dimensional space N of M
246
S. Braidi and Chuan-Chih Hsiung:
constant sectional curvature 1. Then M is either an n-sphere in N or locally a Riemannian direct product M^>U=VlxV2 of spaces V± and V2 of constant sectional curvature, dimF 1 = m ^ l and dim V2=n — mS;l. In the latter case, with respect to an adapted frame field, the connection form (cog) of N, restricted to M, is given by -Xco1
(o\...(Olm
0 -Xof
a%...a% „m + l
•<+1
-fl(Dm
+ 1
(4.11)
0 <
-p.co"
.. .fiC0"
0
afm+i... •
XufL...X(om
/" co
m+1
where X, /J, are given by (4.10). Theorem 4.1 was obtained jointly by Chern, do Carmo and Kobayashi [1] for a minimal hypersurface M without the compactness and orientation conditions, and by Nomizu and Smyth [2] for a complete hypersurface M with constant mean curvature and nonnegative sectional curvature. Now let S"(r) denote a p-dimensional sphere with radius r in a (p + 1)dimensional Euclidean space R p + 1 , m and n be positive integers such that m
(4.14)
by the argument in the proof of Lemma 4.1. On the other hand, Sm(r) (resp. S"~m(s)) has constant sectional curvature 1+X2 (resp. 1+/J.2) which is equal to 1/r2 (resp. 1/s2) so that l + X2 = l/r2,
l+n2 = l/s2.
(4.15)
Submanifolds of Spheres
247
Thus by (4.13), (4.14), (4.15), without loss of generality, we have X=-s/r,
n = r/s,
(4.16)
from which it follows immediately that — ms/r + (n — m)r/s = h,
(4.17)
where h is n times the constant mean curvature of M m '" _m (r, s) and given by (4.9). By solving (4.13), (4.17) simultaneously, we therefore obtain, without loss of generality, /-5 h +2mn — hyh +4m(n — m) 2 T = 2(h2 + n2) ' , (4.18) 2 2 2 + 2n -2mn + hyh + 4m(n-m) 2_h S
W'+n2)
~
'
For simplicity, we shall denote, respectively, by r 0 , s 0 the r, s given by (4.18). Hence Mm-"~m(r0,s0) is a hyper surface in S B+1 . Since TrH,, is constant, it is obvious that (TrH„ +1 ) zl(TrH„ +1 ) = 0. By substituting (4.16), together with the respective multiplicities m and n — m, in (3.28), an elementary computation gives immediately that W1=0. Hence the condition (4.12) holds for M m -"- m (r 0 , s0). Now let f0,fi, ...,/ m be an orthonormal frame field for ]Rm+1 such that /o is normal to Sm(r0), and (j)0,^1, ...,(j)m the dual frame field. Similarly, for S"~m{s0) in R"" m + 1 , we choose an orthonormal frame field fm+u ...,f„+1 such t h a t / B + 1 is normal to S"-m(s0), and let <j)m + 1, ...,4>n+1 be the dual frame field. Let {4>i)A, B = 0 _! , „+J be the connection form for IR"+2 with respect to the dual frame field (<j)A)A=0A „ +1 . Then these forms <j>i, restricted to Mm-"-m(r0, s0), satisfy
<£° = 4," + 1 =0, 0?=-0o=-—0s,
i=l,...,m, (4.19)
0i
+ 1
=-^
+1
= — 4>\
j=
0 B = - ^ =O
m+l,...,n.
for 4 = 0 , 1 , . . . , m a n d B = m + l , . . . , n + l .
The image of the imbedding Mm-"-m(r0,s0) -* IR"+2 lies in the unit sphere S" +1 . Now we take a new frame field e0,eu...,en+1 for R " + 2 such that et=ft, +1 i = 1,..., n, e 0 is normal to S" , and e„ +1 is normal to Mm-"~m(r0, s0) so that e
O
=
r
o / o + S 0 Jn + 1>
e,=/j, e
17
Math. Z., Bd. 115
«+l
=S
0i0
(=1, ...,n, —r
ofn + l-
(4.20)
248
S. Braidi and Chuan-Chih Hsiung:
Let co°, co1,..., ffl"+1 be the dual frame field. Then col = 4>\ t u »+
1
i=l,...,n, 0
(4.21)
+1
=s o ( /) -ro(/)" .
The connection form (co£)AtB=0,1, ...,n+i for JRn+2 with respect to the dual frame field (coA) is given by = r0(/.9 + s 0 ^ + 1 ,
ffl/9=_Gy0
j=l,...,n,
co) = (j>),
i,j=l,...,n,
(ol+i=-co"+i=s0(t)i0-r0
(4.22)
i=l,...,n.
Substitution of (4.19) in (4.21), (4.22) shows immediately that the connection form K ) x , B = i t . . . , „ + i of S" +1 , restricted to Mm-"~m(r0,s0), coincides with the form in (4.11). Hence we obtain Theorem 4.2. The submanifolds Mm,n~m(r0,s0) in Sn+1 are the only compact oriented hypersurfaces ofSn+1, which have constant mean curvature, are not totally umbilical but satisfy condition (4.12). Under an extra minimal immersion condition on the hypersurfaces of S" +1 Theorem 4.2 is due to Chern, do Carmo and Kobayashi [1] jointly. Case p ^ 2 In this case Eq. (4.4) implies fff =
ff2.
(4.23)
Since, by Lemma3.1, at most two of Ha, a = n+l, ...,n + p, are different from zero, we may first assume only one of them, say Hx, is different from zero. Then from (3.16) we have a1 = SJp and G2=Q, contradicting (4.23). Therefore we can assume that #n + l = ^ >
Ha=0
H
n + 2=I^B,
/l,/U=t=0,
for oc = n + 3,
(4.24)
where A and B are defined by (3.10), and therefore (2.4) are reduced to a^+1 = Xco2, +2
col
l
= nco ,
cof = 0
con2+1=Xco1, +2
2
co2 =-p.co ,
co?+l=0
for i = 3, ...,n,
+2
co1 = Q
for <x = n + 3,...,n + p,
for i = 3,...,n,
i=\,...,n.
On the other hand, by using (4.2), we obtain, from (2.10), dh?j=h?lcoli + htJco\-h!Jco}.
(4.25)
Submanifolds of Spheres
249
Setting a = n + 1 , i= 1, j=2 and using (4.24), we see that dX = dh\ J 1 =0, which means that X is constant. Setting a = n + 1 , i= 1 and y'^3, we see that w?=0
for;^3.
(4.26)
Setting a = n + l, i = 2 and j ^ 3 , we see that co)=0
tor
j = 3.
(4.27)
Similarly, setting a = n + 2 and i = ; ' = l , we see that p. is constant. From (4.25), (4.26), (4.27), (2.6) and (3.1), it follows that if j = 3, then 0 = da)) = — co{ A cokj + co1 A oyi' = co1 A a9. Since co1, ...,a>" are orthonormal, a>1A<x>J=0 implies coJ = 0 for j2:3. Thus dim M = 2. Substituting (4.24) in (3.16), we obtain pa1=2(X2 + p2),
p(p-l)
and therefore p2 (p - 1 ) {a2 - a2) = 4 [(p - 1 ) X4 - 21 2 p2 + (p - 1 ) / ] , which becomes, due to (4.4), ( p - l ) A 4 - 2 / l 2 / i 2 + ( p - l ) ^ 4 = 0.
(4.28)
Since the discriminant of (4.28) must be nonnegative, we have l-(p-l)2^0, which and the assumption p =2 imply that p must be 2. Hence dim N = 4, and from (4.28) it follows that X2 = p2. Moreover, since TrH„+1 = T r # „ + 2 = 0 , Af is minimally immersed in N, and the condition W^,=0 becomes S = f . But S = 4X2 from (3.6), so X2 = p.2 = | . Replacing e 3 by — e3 if necessary, we may assume that - 1 = ^ = 1/^3. On the other hand, setting a = 3, i = j= 1 in (4.25), we have 21 col=
<x>\ = 2o3\.
(4.29)
i"
By the second equation of (2.6), (2.7) and (3.1) with c = 1, we obtain the curvature of M, namely, „,,„,-, ,s i ^ i 1 •> ,.„„, J fi12 = (l-/l 2 -/i 2 )co 1 Aco 2 =ico 1 Aco 2 . (4.30) In summary, we hence arrive at Theorem 4.3. Let M be a compact oriented n-dimensional manifold immersed in an (n + p)-dimensional space N of constant sectional curvature 1, such that M is not totally geodesic in N but satisfies
j[%-£(7rtfjzl(7rtfj]*l=().
250
S. Braidi and Chuan-Chih Hsiung:
If p^.2, then n = p = 2, M is minimally immersed in N, and with respect to an adapted dual orthonormal frame field co1,..., co4, the connection form (cog) ofN, restricted to M, is given by 0
co\
p.co2
col
0
lico1
/xco2
Xco2
Xco1
0
2co\
Xco1
XcO2
2co\
0
— fXCO1
-A = /*=l/v/3.
(4.31)
Theorem 4.3 was obtained jointly by Chern, do Carmo and Kobayashi [1] for a minimal immersed manifold M in N without the compactness and orientation conditions. We shall now define the Veronese surface. Let (x,y, z), (u1, ...,w 5 ) be the natural coordinate systems in R 3 , IR5 respectively. Then the mapping defined by l/3
l/3
]/3 (4.32)
u* = —^(x2-y2), 2/3
us=-7(x2 6
+
y2-2z2).
defines an isometric immersion of S2 (y3) into SA. Since two points (x, y, z) and ( —x, — y, —z) of S 2 (j/3) are mapped into the same point of S*, this mapping defines an imbedding into S 4 of the real projective plane, which is called the Veronese surface. Since Eqs. (4.32) induce an action of the orthogonal group SO (3) on SA, so that the immersion S2 (~\/3) - » S 4 is equivariant. In other words, we obtain a representation of SO (3) into SO (5), which induces a representation of the Lie algebra so (3) into the Lie algebra so (5). A simple straightforward calculation shows that this induced representation maps a matrix of the form
(4.33)
eso(5).
Let (cos) be the Maurer-Cartan form for so (5), and set co' = co'5,
i=l,...,4.
(4.34)
Submanifolds of Spheres
251
Then the restriction of (u>i) t o the image of so(3) in so(5) is given by '
0
co\
\xco2
— /ICO 1
CO1
°>i
0
/ICO 1
\xca2
CO2
km2
A co1
0
2co\
0
2 co
0
0
0
0
0
1
Xco
1
1
-Xco ,-co
2
-co
2
-A = / x = l / j / 3 .
(4.35)
A comparison of (4.35) with (4.31) gives immediately Theorem 4.4. The Veronese surface in S 4 is the only compact oriented n-dimensional submanifold in Sn+P for p^.2, which is not totally umbilical but satisfies condition (4.1). Under an extra minimal immersion condition on the submanifolds in S"+p, Theorem 4.4 is due to Chern, do Carmo and Kobayashi [1] jointly. It should also be noticed that all other minimal submanifolds in S" + p for p>2 given by them jointly [1,§5] are examples where the submanifolds do not satisfy condition (4.1). Bibliography 1. Chern, S. S., do Carmo, M., Kobayashi, S.: Minimal submanifolds of a sphere with second fundamental form of constant length. To appear in Marshall Stone Jubilee volume. 2. Nomizu, K., Smyth, B.: A formula of Simons' type and hypersurfaces with constant mean curvature. J. Differential Geometry 3, 367-377 (1969). 3. Simons, J.: Minimal varieties in riemannian manifolds. Ann. of Math. 88, 62— 105 (1968). Prof. Chuan-Chih Hsiung and Dr. Siham Braidi Department of Mathematics Lehigh University Bethlehem, Pa. 18015, USA (Received February 19, 1970)
344
MINIMAL IMMERSIONS IN RIEMANNIAN SPHERES CHUAN-CHIH HSIUNG
1.
($£&)
Introduction
Let M* be a complete Riemannian manifold of dimension n+m (n>2, m>0) and class C \ and 0 a point on M*. Then we define a Riemannian sphere S0 O) of radius r of M* of 0 to be the locus of a point which has a constant geodesic distance r from 0. The purpose of this paper is to establish the following theorem: Theorem 1.1. Let x:M-^M* be an isometric immersion of class C2 as 1 normal coordinates ^ of a compact Riemannian manifold M of dimension n{>2) in a complete Riemannian manifold M* of dimension n+m ( O T > 0 ) and class C \ and A be the Laplace-Beltrami operator on M. If (1- 1)
A ^ = Xx,
for some constant X=fcQ, then X is necessarily positive and x realizes a minimal immersion in a Riemannian (1.2)
{n+m—1) - sphere Sn+m~1(r)
of
radius
r=VnJx
in M*, provided that such an {n+m —1) -sphere S"+m~1{r) of class C2 exists. Conversely, if there exist Riemannian {n+m —1) -spheres Sn+m~1{r) of radius r and class C2 in M*, and x realizes a minimal immersion in such a sphere gn+m,-i(j.^ as normai coordinates, then x satisfies (1.1) and (1.2). It should be remarked that in the above theorem if n/X is very small, then the existence of the Riemannian {n+m —I) -sphere Sn+m-1 {Vn/X)
does not
require any additional condition; otherwise, as examples, either one of the following two conditions can be used as a necessary condition: (a)
M* is simply connected with nonpositive sectional curvature.
Received February 11, 1970. The work was partially supported by the National Science Foundation grant GP-11965. 1) For the definition see p. 227, 223
345 c
224
- c - HSIUNG
(b) M* is compact and simply connected with the sectional curvature K(P)
of every plane section P satisfying the inequalities
where c is constant. Under condition (a), on M* there are no cut points along geodesies, and therefore every pair of points is joined by a unique minimizing geodesic. Under condition (b), the distance between every point O e M * and its cut locus C(0) is not less than c (for c=n, see [ 4 ] , [5]), and therefore n and X in the theorem can be related by w<Xc2. The above theorem is due to Takahashi [ 6 ] for a Euclidean M*. For other minimal immersions in Euclidean spheres see [ 9 ] , [ 2 ] . 2.
Notation and formulas
Let M and M* be Riemannian manifolds of dimensions n ( > 2 ) and n+m ( m > 0 ) with local coordirftes (w1, • • -,un) and (x1, • • -,xn+m), and metri ctensors ga and g*AB, respectively. Throughout this paper, unless stated otherwis ewe shall make use of the \>Alowing convention on the ranges of indices: 1 < A, B, C, E, • • • < n + m, 1 <, Uj,k, n + 1
• • • < n, • • •
and also the convention that repeated indices imply summation over their ranges. Let (2.1)
x:M-+M*
be an isometric immersion of class C 2 . Then in local coordinates, x can be represented by (2.2)
xA=xk(u\---,u"),
(.A =
and we have (2-3)
gij=g*ABx\txH,
where (2.4)
x\i = dx*/du\
j,
l,---,n+m),
346 RIEMANNIAN SPHERES
225
or in the vector form, (2.5)
x,i =
dx/dui.
N o w we introduce the generalized covariant differentiation, which is useful for studying submanifolds
manifolds. Let TABI
of Riemannian
b e a mixed
tensor of the second order in the x's, and a covariant vector in the u's, as indicated by the capital and small Latin
indices respectively. T h e n
following
A
Tucker [ 7 ] , the generalized covariant derivative of T m with respect t o t h e u's is defined as V / T % < = ^f- 1 du
(2.6)
+TiET°,
-nETAotx\
i-vf^BK,
where the Christoffel symbols VBC with capital respect to the g'w and
Latin indices a r e formed
the x's, and those H , with
respect to the gi} and the u's.
It should
small Latin
with
indices
b e noted that t h e definition
with
of t h e
generalized covariant differentiation can be applied t o any tensor in the u's and x's and
that
t h e generalized
covariant differentiation
of sums and
products
obey the ordinary rules. If a tensor is one with respect t o the u's only, so that only small Latin indices appear, its generalized covariant derivative is the same as its usual covariant derivative with respect to the u's. Moreover, in generalized covariant differentiation
t h e metric tensors
g*iB a n d gtj can be treated as
constants. Since xA is an invariant for transformations of the u's, its generalized covariant derivative is the same as its usual covariant
derivative with
respect
to the u's; so that (2.7)
Vtx = x, i =
dx/dui.
By (2. 6) the generalized covariant derivative of X7iX is (2. 8)
VjxA, t=^±L-T*jx\
A
h+T
cx\ ixc,,,
which is symmetric in the indices i and j . Let en+1, •••,en+m
be an orthonormal basis of the normal space at a point
x of x(M). T h e n (see, for instance, [ 8 , Chap. X ] ) : (2.9)
V<*,y=£{We«,
347 c
226
( 2 . 1 0 ) Vie« = dejdui where (giJ)
- c - HSIUNG
= -Q>«\ikgikx,
,• + £ &«tnei>, (a = n + l, • • • ,n + m),
is the inverse of the matrix (ga),
(2.11)
Q,.\ij = Cl«\ji,
and
#«/m + #/>«ii = 0.
T h e vector
(2.12)
^r=2^e„, a
where
(2.13)
^r„=Q.uyg-^,
is called the mean
curvature
x is said to be minimal
vector
of x{M)
at the point x, and the immersion
if the mean curvature vector ^
vanishes
identically,
i. e., if (2.14)
^„ = 0
for a—n + 1, • • -,n + m.
By definition, (2.15) Ax=-g"Vtx,j, and therefore from (2. 9 ) , ( 2 . 1 3 ) follows immediately
(2.16)
A-z=-S-^e
x is said to be harmonic Theorem 2 . 1 . manifold n+m 3.
[ 3 ] . An isometric
M of dimension is minimal
if A ^ = 0. T h u s we have immersion
x: M —> M* of a
n in a Riemannian
if and only if x is
manifold
harmonic.
Proof of Theorem 1.1
Substitution of (2. 15) in ( 1 . 1) gives (3.1)
x = — —- X A
^*e«,
a
and therefore we obtain, in consequence of (2. 10), ViX
= X, i = — - X •^•Cl«\Ug,ikX, A
J
a
(3.2) \
«
f
M* of
Riemannian dimension
348 RIEMANNIAN SPHERES
227
which implies
8/=4-S-^O.i«« J t ,
(3.3)
A.
a
where 8j is the Kronecker delta, putting j=i
in (3.3), summing for i, and
using (2.13), we have
S ( - ^ ) 2 = «X,
(3.4) which shows (3. 5)
X >0.
On the other hand, from
(3.1), (3.4), we obtain the length of x, as a
vector in the manifold M*, 1 * 1 = 4 " ( £ ( ^ ) 2 ) 1 / 2 = V « A = const. = r(say).
(3.6)
A.
^
Now suppose x to be the normal coordinates of a point P of M* with respect to a fixed point 0 on M* so that
(3. 7)
x = £s,
where £ is the unit tangent vector at 0 of the geodesic y joining 0 to P, and 5 is the arc length of the geodesic y. From (3.7), (3.6) it follows immediately that
\x\=s=r.
Thus the immersion x,
as normal coordinate, realizes an
immersion in the Riemanian (n + m—1) -sphere S^"'1 (r) of radius r
of M*
at 0. The unit tangent vector at P of the geodesic y is dx/ds=^. hand, by the well-known Gauss' Lemma unit
tangent vector is perpendicular
Sn0+m~1{r)
On the other
(see, for instance, [1, p. 147]) this
to the
Riemannian
(n + m—1) -sphere
so that we can take it to be the unit normal vector e„+m at P. Thus
from (3.7), on the Riemannian (n + m — l) -sphere SS+m~i (r) we have (3. 8)
x — ren+m,
and therefore (2. 10) for <x = n + m becomes (3. 9) which implies
- i - ViX=-On+m\ttgltx, r
j+XK**,,,*,, P
349 c
228
(3.10)
£ln+m\tj=-gij/r,
- c - HSIUNG
# B + m ,, u = 0,
(/3=n + l, •••n + ij
Transvecting the first equation of (3. 10) with g
m-1).
and substituting the resulting
equation in (2.13) we obtain, in consequence of (3. 6), (3. 11)
-<+*,= -n/r=
—V\n-
Substitution of (3. 11) in (3. 4) yields immediately (3.12)
" 5 ? ( - * • ) ' = 0,
which implies (3.13)
^t. = 0,
(a = n + l, •••,n +
m-l).
Hence x realizes a minimal immersion in the Riemannian (w + m — 1)-sphere S j + " - l ( r ) , where r=VnJ\
b
y (3-6)-
Conversely, assume that x as normal coordinates realizes a minimal immersion in a Riemannian (n+m — l) -sphere Sn+m~1(r) of class C2 of M*. Then we can take an orthonormal basis e„+1, • • •, en+m of the normal space at a point x of x(M) in M* in such that (3.14)
e„+m =
x/r,
which is perpendicular to 5™ +m_1 (r). In our case, equation (3.9) and therefore equation (3.10) are automatically satisfied. Moreover, since Q«n/ (a = n + l, • • •, n + m—l) are considered as the second fundamental tensors of the induced immersion in Sn+m~1(r), we have, by (2. 13) and our assumption of the minimal immersion, ^<,=Q,«\ijgti = 0,
(a = n+l,
• • -, n + m — 1),
and therefore, from (2. 16), (3.14), A-£= — X ~#.e.= — ^n+men+m= a
But from (2.13) and the first equation of (3. 10),
Hence A^c =
nx/r2,
which completes the proof of Theorem 1.1.
—-rfn+mx/r.
350 RIEMANNIAN SPHERES
229
REFERENCES [ 1] [ 2]
[ 3] [ 4] [ 5] [ 6] [7] [ 8] [ 9]
R. L. Bishop and R. J. Crittenden, Geometry of manifolds, Academic Press, New York, 1964. S. S. Chern, M. do Carmo and S. Kobayashi, Minimal submanifolds of a sphere •with second fundamental form of constant length, to appear in Marshall Stone Jubilee volume. J. Eells, Jr. and J. H. Sampson, Harmonic mappings of Riemannian manifolds, Amer. J. Math. 186 (1964) 109-160. W. Klingenberg, Ober Riemannsche Mannigfaltigkeiten mit positiver Kriimmung, Comment. Math. Helv. 35 (1961) 47-54. , Ober Riemannsche Mannigfaltigkeiten mit nach oben beschrankter Kriimmung, Annali di Mat. 60 (1963) 49-60. T. Takahashi, Minimal immersions of Riemannian manifolds, J. Math. Soc. Japan 18 (1966) 380-385. A. W. Tucker, On generalized covariant differentiation, Ann. of Math. 32 (1931) 451-460. C. E. Weatherburn, An introduction to Riemannian geometry and the tensor calculus, Cambridge, 1938. J. Simmons, Minimal varieties in riemannian manifolds, Ann. of Math. 88 (1968) 62-105.
Lehigh
University
ROCKY MOUNTAIN JOURNAL OF MATHEMATICS Volume 1, Number 3, Summer 1971
CURVATURE AND CHARACTERISTIC CLASSES OF COMPACT PSEUDO-RIEMANNIAN MANIFOLDS 1 CHUAN-CHIH HSIUNG AND JOHN J. LEVKO III
Introduction. In the last three decades various authors have studied the relationships between curvatures and certain topological invariants such as characteristic classes of a compact Riemannian manifold. One of the earliest results was the Gauss-Bonnet formula [1], [6], which expresses the Euler-Poincare characteristic of a compact orientable Riemannian manifold of even dimension n as the integral, over the manifold, of the nth sectional curvature or the Lipschitz-Killing curvature times the volume element of the manifold. Later, Chern [8] obtained curvature conditions respectively for determining the sign of the Euler-Poincare characteristic and for the vanishing of the Pontrjagin classes of a compact orientable Riemannian manifold. Recently, Thorpe [11] extended a special case of Chern's conditions by using higher order sectional curvatures, which are weaker invariants of the Riemannian structure than the usual sectional curvature, and Cheung and Hsiung [5] jointly further extended the conditions of both Chern and Thorpe. On the other hand, Avez [3] and Chern [9] used different methods to show that the Gauss-Bonnet formula is also true up to a sign on a compact orientable pseudo-Riemannian manifold. Very recently, from general remarks on connexions and characteristic homomorphisms of Weil [7, pp. 57-58], Borel [4] elegantly deduced this fact and expressed the Pontrjagin classes of a compact orientable pseudo-Riemannian manifold in terms of the curvature 2-forms. The purpose of this paper is to give an independent proof of Borel's result on the Pontrjagin classes and to extend the above mentioned joint work of Cheung and Hsiung to a compact orientable pseudoRiemannian manifold. §1 contains some fundamental formulas for a pseudo-Riemannian manifold such as the equations of structure, and the formulas for the higher order sectional curvatures and related differential forms. §2 is devoted to expressing the Euler-Poincare characteristic and the Pontrjagin classes of a compact orientable pseudo-Riemannian manifold in terms of the curvature 2-forms in the sense of de Rham's theorem. In §3, we extend the above mentioned joint results of Cheung and Hsiung Received by the editors October 23, 1969 and, in revised form, March 13, 1970. AMS 1970 subject classifications. Primary 53C50, 53C65, 57D20. S u p p o r t e d partially by the National Science Foundation grants GP-7513, GP11965. Copyright ° Rocky Mountain Mathematics Consortium
523
524
C.-C. HSIUNG AND J. J. LEVKO III
to the pseudo-Riemannian case by first establishing several lemmas and then deducing the proofs of the two main theorems of this section. 1. Fundamental formulas. Let Mn be a compact orientable manifold of dimension n ( ^ 2) with a pseudo-Riemannian metric gap(uK) of signature r, the number of positive eigenvalues of the matrix of the metric, where u 1 , • • •, u" are local coordinates of an arbitrary point x G. Mn; throughout this paper all Greek and Latin indices take the values 1, • • •, n unless stated otherwise. At any point x £ M" and over a neighborhood U of x we consider the spaces Vx and V* of tangent vectors and covectors respectively, and the family xei • • • en of orthonormal frames and linear differential forms o>', • • •, o)n with respect to an orthonormal basis in Vx and its dual basis in V*; that is, {e{, o>> = Sy = 1,
(1.1)
if i = j ,
= 0,
Xijij.
The pseudo-Riemannian metric of Mn is of the form
(1.2)
ds*=
£
(a,') 2 -
i=l
2
(o*)2.
i=r+l
For indices we use I(p) to indicate the ordered set of p integers i 1; • • •, ip among 1, • • •, n. When more than one set of indices is needed at one time, we shall use other capital letters in addition to I. The equations of structure of the pseudo-Riemannian metric are
L3
( )
db>{ = £
wJ
A
W
J>
W
J
+ to^i = 0,
L
do'j = 2 o>kj A <jk + Mj,
a , + o>< = o,
where the components for the curvature 2-form fty satisfy Clfj = aij (1.4)
ff,
= -ikj 0^.= ^
Then, for any even p^ define the p-form
(i.5)
(lSjgr), (r
n),
(lgi,jgn).
n and distinct set of integers iu
• • •, ip we
e/(P) = -i-2(-i)«c/)8^ ) , ft J J l A ••• A%,_,j;,, V-
](P)
where c(J) denotes the number of the curvature 2-forms Qjk with j > r for each combination (J1J2, ' ' ',jp-ijp)>and
COMPACT PSEUDO-WEMANNIAN MANIFOLDS
8/(pj (1.6)
=
+1> if J(p) is a n
even
525
permutation of I(p),
= — 1, if J(p) is an odd permutation of I(p), = 0,
otherwise.
In terms of a natural orfhonormal basis in local coordinates u 1 , • • •, u in the neighborhood U we have n
nij=±Rijkza>k
(1.7)
Aw*,
where repeated indices imply summation over their ranges, and Rijkt is the Riemann-Christoffel tensor. Thus (1.5) can be written as (1.8) ©;(P) = - r ^ - p Z
'
P!
2
8$ifl"j*,h,
""• K^y.,,-*„^H(P),
J(P),H(P)
where (1.9)
dVif(p) = ft)'1' A • • • Ac/'f = duhi A • • • A du''''
is the volume element of the p-dimensional submanifold of M" with local coordinates uh>, • • •,u'1''. Let P be any p-dimensional plane in the tangent space Vx of the manifold M n at a point x. Then the Lipschitz-Killing curvature at x of the p-dimensional geodesic submanifold of M" tangent to P at x is called the pth sectional curvature of M" at x with respect to P, and is given (see, for instance, [2, p. 257]) in terms of any orthonormal basis eiy • • • e,-,, of P by
(I.IO)
KHP)(P)
= - ^ § ^ 2
W,
*M. • • • i 7 ' - W . * P -
2. Characteristic classes. Let M" be a connected manifold of dimension n (i? 2) endowed with a pseudo-Riemannian metric g ^ of signature r. Consider any Riemannian metric hap on M n . In the tangent space Mx of Mn at each point x, g ^ defines the field T of symmetric linear transformations by means of hap. Now, consider the decomposition (2-1)
Mx =
£
Wi+
£
Wj,
where W, = {X G M, | TX(X) = XiX} = { X £ M I | gx(X, Y) = \ ^ ( X , Y), for all Y <= Mx} . Put r = 2)i; x, > o dim Wj. Since g and therefore T are nonsingular, r is constant and we have
526
C.-C. HSIUNG AND J. J. LEVKO III
n —r=
]£
dim Wj.
Suppose that (2.3)
W'n=
X
Wfo
W"~\=
2
W,-.
Then the distributions W:x—*Wrx and W n ' r : x —» W" - '* are differentiable, and this leads us to define (2.4)
gx\W'x = a,
- gxIW-'x = b,
where the distributions a : x —* ax and b : x-+bx Now, define (2.5)
are differentiable.
g= a+ b.
Then g is positive definite, and with respect to this g, g has eigenvalue 1 of multiplicity r and eigenvalue —1 of multiplicity n—r, that is, for each x €E M n , Mx = Wrx + Wn~rx is the eigenspace-decomposition with respect to g. By considering the tensor (2.6)
i(t) =
g+tb,
or (2.7)
l(t) = a + (1 + t)fe,
where f is a real parameter, we see that % defines a nonsingular pseudoRiemannian metric on M" for tj^—1, which is Riemannian for t > — 1 and is of signature r for t < — 1; in particular, H( — 2) = g. The inverse tensor of (2.6) is given by (2.8)
l"fi(t) = gali - ~—fr"5
.
Let Tyae(t) be the Christoffel symbols with respect to Zap(t). Then C*ap{t) = TyaP(t) - n a/8 (0) defines a tensor, where P ^ O ) are the Christoffel symbols with respect to SLap(0) = g,^. Let V ( " and V(0> be the covariant derivation operators in the Riemannian connexion associated with SLap(t) and g ^ = $>ap(0), r e spectively. By using normal coordinates at a point x on M" with respect to g ^ so that P a/ s(0) = 0 at the point x, we have
The cyclical permutation of the indices a,/3,a in (2.9) yields
527
COMPACT PSEUDO-MEMANNIAN MANIFOLDS
(2.10)
0 = V„«»M*) " W * ) C " „ ( t ) - ^ ( f ) C W * ) .
(2.11)
0 = v , < ° > U 0 - Xya(t)CyaB(t) -
Ky{t)CyaB(t).
Subtracting (2.10) from the sum of (2.9) and (2.11) we thus obtain ^i0Ke(t)).
(2.12) OaB{t) = lj^(f)(V„«»JU') + V ^ U t ) Now, for a contravariant vector t?M on Mn we have VB<"o" = Va<()>t^ + ((
C^ve, <(
VX<"VB<«>«" = Vx >(V« M + Vx^OaflO") = ( V ^ V ^ ' i ? " + C% V a <«V - 0 X t t Vp<°>c)
+ C^Cpa,^-CpXaC^X]> and therefore (Vx«>Va<" - V a e > V x « V = (VX(°»VQ(°) - V Q < 0 >V X (0 V
+ [ ( V X < ° > O Q / J ) - (VQ<°>(>X„) + C^O^ -
O^OxflJo".
Thus the Ricci identity gives *„*(*) = *,*(<)) + Vx(°>Oa/!(t) - V„<°>Ox/,(t) (Z.lo)
+ C J O O a ^ t ) - O a p (*)Ox^), where RM^aX(*) and R'W(O) a r e the Riemann-Christoffel tensors with respect to r v a/3 (t) and P ^ O ) respectively. Let dV(t) and dV(0) be the volume elements of M" associated with ZaB(t) and HaB(0) respectively. By using equation (2.7) and orthonormal local coordinates M1, • • •, un, we readily obtain dV(0) = |det(£^(0))p/2 du1 A • • • A du" = |det (a) det ( f c ) | « W A • • • A dun, (2 15)
dV(t)
=
|dCt (£
^ ( ^ ) ) | 1 / 2 dul A ' ' ' = |1 + t|("-^ 2 dV(0).
A
d
""
More generally, the volume elements dVt[(4k)(t) and dVn(4k)(0), respectively, associated with lhlhj(t) and i/,,/,/0), i, j = 1, • • % 4k, of the 4fedimensional submanifold of Mn, 4k = n, with the local coordinates (uh', • • •, lA*) are related by
528 (2.16)
C.-C. HSIUNG AND J. J. LEVKO III
dVH{4k)(t) = |det(£,„,„(*))| u W" A • • • A duh« =
\l+t\°i*dVH(4k)(0),
where s h's are greater than r, and the remaining h's are less than or equal to r. THEOREM 2.1. Let M" be a compact orientable manifold of even dimension n with a pseudo-Riemannian metric of signature r. Then the Euler-Poincare characteristic X(Mn) of the manifold M" is given by
(2 17)
x(Mn)
-
=-i^S-L *••-
where&i...n is given by (1.5). Theorem 2.1 is due to Allendorfer and Weil [1] and Chern [6] for the Riemannian case, and due to Avez [3] and Chern [9] for the pseudo-Riemannian case. Our proof is essentially the same as that of Avez. PROOF. By means of (1.8) we have =
®l--»
On/2„|
S
8 in„RJihhlh2
(2.18)
RJ'-1J„h„->hndVH(„).
•••
Now let®!...n(t) be the form©i...„ associated with£ a/3 (() given by (2.6). Then
ei...„(t) = n " S (-i)c^{(n,nflJ1J2W---«;„-u„W =
0n/2„\ *
8l",nR,''J-2A1h2(^)
S
ni
J(n),H(n)
•••*"-'w..-,*.(')dvH(B,(t), and from (2.13), (2.12), (2.8), (2.7), (2.15) it follows that (2.i9)
ei...»(*)=|i +
( | < - , »
where P(t) is a polynomial in t. Thus
(2-20) =
II + +|(n-r)/2
'
+
'
VW
(1 + t)» '
2
^ -
COMPACT PSEUDO-RIEMANNIAN MANIFOLDS
529
where Q(t) is a polynomial in t. Now, suppose that r is even. Since £a/3(t) for t> — 1 defines a Riemannian metric on M", it is known [1], [6] that f(t) for t > —1 is the Euler-Poincare characteristic X(Mn) of M". Thus from (2.20) we have (2.21)
Q(t) = (1 + *)<"+r>/2X(M")
for t> - 1 and therefore for all t. Substitution of (2.21) in (2.20) thus gives (2.22)
II + t|<"--->/2 # ) = X(M") | x + ^ B _ r y 2 , for all t.
For t < - 1 we have 1 + t < 0 so that |1 + t|<"-r"2 = (-l)< n - r > /2 (l + t)<»-')/2? and hence /(f) = (-1)<"-'>'2X(M"), which proves our formula (2.17) for even r. Finally, suppose that r is odd such that r = 2r' + 1. Then we have, fort > - 1 , (2-23)
X(M") =
T
n
<
^
• ^
-
^
,
which implies that Q(t) = 0 for t > — 1 and therefore for all t. Hence X(Mn) = 0 = ft) for all t, which shows that our formula (2.17) is also true for odd r, and completes the proof of Theorem 2.1. THEOREM 2.2. Let M" be a compact orientable n-dimensional manifold with a pseudo-Reimannian metric of signature r. Then the differential form
(2.24)
V4k -
L[(2fc)!] J
\J;
;
(2*/c!)2(2^
2 e, (2t) A 0,(2fc)
/(t,
defines the kth Pontrjagin class Pk of the manifold Mn in the sense of de Rham's theorem. Theorem 2.2 is due to Chern [8] for the Riemannian case and due to Borel [4] for the pseudo-Riemannian case. However the proof given below is different from that of Borel. PROOF. By means of (1.8) we can rewrite (2.24) as
530
C.-C. HSIUNG AND J. J. LEVKO III
(2.25)
••• R ^ i ^ . ^ ' 8nik)RJ'j2hik+Ii,2k+2
• • • RJ'2k~,J-2iliuk-,h4kdVH{4k).
Now let^4fc(£) be the form^fc associated with lap(t). Then
I72fc)'l2 »J(2fc)
*J/((2fc!ff1M,h,(t)
2
(2^1)2(2^)2^^
(2.26)
'
.
•8{((lfc))R-'72,,2/, + | / , ; , l + 2 ( i ) • • • Rhk~l
•••H^-'^W) h^-^k(t)dVHm(t),
and from (2.13), (2.12), (2.8), (2.7), (2.16), (2.25) it follows easily that ^4k(t) can be expressed in the following general form:
(2.27)
, M t ) = (1 + t)~4k " f I1 + *l1/2 i=0
or (2.28)
*4k(t) = (1 + t)-«(E(t)
+\1 + t\u*F(t)),
where Qi(t), E(t) and F(t) are polynomials in t. It is known [8] that ^4k(t) for t > — 1 defines the fcth Pontrjagin class Pk of Mn with real coefficients, so that (2.29)
Pk = V4k(t) + B4k
for t > — 1, where B4k is the group of the exact 4fc-forms of the manifold M n , which is obviously independent of t. Substitution of (2.28) in (2.29) thus gives (2.30) (1 + tykPk = E(t) + |1 + *|1/2F(t) + (1 4- t)4kB4k,
for t > - 1 .
Since E(t) and F(t) are polynomials in t, (2.30) implies that F(t) = 0 for £ > — 1 and therefore for all t. Hence from (2.28), (2.30) we see that (2.29) holds for all t, so that ^4k(t) defines the fcth Pontrjagin class Pk for all t, and in particular the case where t = — 2 gives our Theorem 2.2. 3. Relationships between curvatures and characteristic classes. Let Mn be a connected manifold of dimension n ( = 2) with a pseudo-
COMPACT PSEUDO-RIEMANNIAN MANIFOLDS
531
Riemannian metric g^p of signature r, p be an even positive integer ^ n, and aih = aM (i, h = 1, • • •, n) be given smooth real-valued functions on M". Denote (3.1)
Aj (p)>H(p) = det(o iah)l )
(o,/3 = 1, • • ;p),
where the rows and columns of det(ai„hj) are arranged in the natural order of a and /8, respectively. Consider the following curvature condition at x £ M": (3.2)
£
8 f l p > J l h l h l • • • «'"-';>„_,,,„= 2»% P A 7(P) , H(P) ,
;(p)
for all I(p), H(p) £ (1, • • •, n), where KP is a smooth real-valued function on M" at x. In the Riemannian case, this condition was first used by Chern [8] for p = 2, by Thorpe [11] for a general p but o-ih = Sih (in this case (3.2) implies that the Lipschitz-Killing curvature Kj(P)(P) is constant at x for every P and all I(p)), and then jointly by Cheung and Hsiung [5] for general p and a^. Furthermore, it is easy to see another geometric significance of the condition (3.2) for the Riemannian case, namely, if Mn is a hypersurface of a Euclidean space, then the symmetric tensor aih may be taken to be the second fundamental form of M". From (3.2), (1.8), (1.9), (1.10) follows immediately LEMMA 3.1. For a fixed set of indices I(p), condition (3.2) implies (3-3)
®/( P ) =
-jKpAI(phH(p)0>HipK
(3-4)
K, ( P ) =
(-1)P/2KPA/(P),H(P),
and also equation (3.3) implies condition (3.2), where (3.5)
a>"<»> =
In particular, when ay = 8y, then (3.6)
Ai(p)iH(p) — A/(P),H(P),
where (3.7)
A, (p) , H(P, = det(8,tth<j)
fo/8
= 1, • • •, p).
Therefore (3.3), (3.4) are reduced to (3.8)
(3-9) Thus, from (3.9) we have
@/(p) = KPO>'«» ,
K/(P)=(-1)*'V
532
C.-C. HSIUNG AND J. J. LEVKO III
LEMMA 3.2. Condition (3.2) with ay = Si, implies that the pth sectional curvature Ki^P) at the point x of the manifold M" is constant, that is, independent of the p-dimensional plane P at the point x.
On the other hand, from (1.8), (1.9), (1.10) it follows immediately that (3.8) implies (3.9). The converse is also true (can be proved in exactly the same way as given by Thorpe [11] for the Riemannian case), so that we can state, altogether, LEMMA 3.3. Equations (3.7) and (3.8) are equivalent. For the converse of Lemma 3.2, we notice that [ 10, p. 238] (3-10)
8 nP) = A 7(p),H(P),
so that (3.8) can be written as
P
(3.11)
HiP)
'
=
J
p\
TKPAI(P),H(P)<»H(P)-
A comparison of (3.11) with (1.8), (1.9) yields immediately condition (3.2) with a,} = 8 y . By combining this result with Lemma 3.2 and using Lemma 3.3 we hence obtain LEMMA 3.4. The pth sectional curvature KiiP} of the manifold Mn at a point x is constant if and only if condition (3.2) with a^ = 8y holds. LEMMA 3.5. On a pseudo-Riemannian manifold M" of dimension n, if condition (3.2) holds for some even p and q with p + q S n, then (3.12)
©,<„+,) =
K1#qAHv+q),H(p+q)<»H(l'+q),
*
so that (3.2) also holds for p + q withKp+q = KpKq. PROOF OF LEMMA 3.5. Let the set I(p + q) have distinct elements, and (h(p), h(q)) be a partition of I(p + q), where I\(p) = (in, • • •, i lp ) and I2(q) — (121, - ' ", kg)- Then, from (1.5),
\P "•" q)\ (3.13)
iIM
A •• • Afti,.,-,!,, Aft i21 i 22
A • • • A aM.lJif ,
COMPACT PSEUDO-RIEMANNIAN MANIFOLDS
533
where c(/i), c(I2) denote the numbers of the curvature 2-forms iljk with j > r for the combinations (tuii 2 , ' " ', »i,p-i*iP) and (121*22, • • •, t2,q-i*2)> respectively, and ^ ( / {} denotes the summation over all such partitions of l(p + q) into (Ii(p), h{q))- For a fixed I(p + q), let J(p + q) be an even permutation of I(p + q) such that ji, • • •, j p G Ii(p), and j p + 1 , • • ; j p + q G I2(q). By denoting J'(q) = (jp+u ' ' ", jp+q)> using (1.5) and noticing that altogether there are (ppq) such partitions of I(p + q) into (h(p), hi*]))' from (3.13) we then obtain
©'<»-> =
{p
+ q)]
tf2
( - i w J & h w , . A • • • A nji()_iii;, A(-l)c^8;f^ijli22 A • • • A a , , , . ^
(3.14) = 0; ( p ) A 0;.«,) .
On the other hand, by the Laplace theorem we can expand the determinant Aj(P+q)yH(v+q) according to the first p rows. By using this expansion it is easily seen that all (p +vq) terms of Aj{p+q)!H{p+q^(aHlp+q} are equal so that we have (3.15)
=
iP
^qf
AJ(P),H(P)CO""»
A AiW,X*),
where H'(q) = (h p + i, • • •, hp+q). Substituting (3.3) in (3.14) and using (3.15) we arrive at (3.12), and an application of Lemma 3.1 hence completes the proof of Lemma 3.5. By repeatedly applying Lemma 3.5 we can easily obtain COROLLARY 3.5.1. Let pu • • •, pk be even positive integers, and (mi, ' " "> mk) a k-tuple of nonnegative integers such that q = ^di=imipi § n. On a pseudo-Riemannian manifold M" of dimension n, if condition (3.2) holds for p 1 ; • • •, pk, then it also holds for q with
3.5.2. On a pseudo-Riemannian manifold Mn of even dimension n, if condition (3.2) holds for some positive even integer p dividing n, then COROLLARY
(3.16)
©i..„ = («„)""»detfo^co1 A • • • A w " ,
where to1 A • • • A a>" is the volume element
ofMn.
Combination of Theorem 2.1 with Corollary 3.5.2 gives immediately
534
C.-C. HSIUNG AND J. J. LEVKO III
THEOREM 3.1. On a compact orientable pseudo-Riemannian manifold Mn of even dimension n with a pseudo-Riemannian metric of signature r, if condition (3.2) holds at every point xfor a positive even integer p dividing n, and ( — l)[r/2)(*cp)n/p det(ay) keeps a constant sign, then this sign is the sign of the Euler-Poincare characteristic X(Mn) of Mn. Moreover, under this hypothesis, X(Mn) = 0 only when (K P )" /P det(a,j) vanishes identically.
For the Riemannian case, this theorem was obtained by Chern [8] for p = 2, by Thorpe [11] for atj = 8y, and jointly by Cheung and Hsiung [5] for a general p. For studying Pontrjagin classes we need LEMMA 3.6. Equation (3.3) can be written in the following form: (3.17)
© /(p) =
Kp&™
,
where d>'" are linear forms defined by (3.18)
&*« = atjlaf
(o = 1, • • ;p).
PROOF. Let p t , p 2 be any two positive integers such that p\ + p2 = p. By using pu p2 for p, q, from (3.15) we then have
(3.19)
A/(P),H(p,COH(p)
^-j-A/(P])>„(Pl)a>"
=
where (3.20)
l'{p2) = (iPl
+ 1,
• • •, ip),
H'(p2) = ( V + i, • • •, K).
Repeatedly applying the same argument as above to both factors on the right-hand side of (3.19) yields immediately (3.17). Now we are in a position to prove the following theorem concerning the general curvature conditions for the vanishing of the Pontrjagin classes. THEOREM 3.2. On a compact orientable pseudo-Riemannian manifold Mn of dimension n, if condition (3.2) holds at every point x for a positive even integer p^ n, then the kth Pontrjagin class Pk(Mn) of M" is zero for all k ^ p/2.
For the Riemannian case, this theorem is due to Chern [8] for p = 2, to Thorpe [11] for ay = Sy, and jointly due to Cheung and Hsiung [5] for a general p.
COMPACT PSEUDO-RIEMANNIAN MANIFOLDS
535
PROOF. First, we consider the case p ^ 2k < 2p. Let(/i(p), Z2(2/c —p)) be a partition of a fixed I(2k), and J(2k) an even permutation of I(2k) such that j u • • -, jp G h(p) and j p + 1 , • • •, j 2 k G I2(2k - p). By denoting/'(2fc — p) = (jp + i, • • -,j2fc), from (3.14) we have
(3-21)
0/<2*> =
S8j
W
Ae
T O
.
p ) )
where 5](/,/) denotes the summation over all such partitions of I(2k) into (J(p), J'(2k — p)). By using condition (3.2) for p and Lemmas 3.1 and 3.6, equation (3.21) is reduced to (3.22)
@I(2k) =
Kp
£
d>'<"> A ey(2fc-p),
(/./'I
where a>j°are linear forms defined by (3.18), so that 0/(2*) A 0/(2jy is a sum, each term of which contains an exterior factor (3.23)
fiW
A S-fa" ,
where all the superscripts j , j G I(2k). Since 2k < 2p, at least two of t h e / s and fs in (3.23) must be equal, so that each of such factors (3.23) is zero. Thus0/(2fc) A 0/(2*) = 0 for all I(2k). By Theorem 2.2 we hence obtain Pk{Mn) = 0 for all k with pj2^k< p. Finally, since condition (3.2) is assumed to hold for p, by Corollary 3.5.1 it also holds for 2'p (i = 1, 2, • • •). Applying the same arguments as above we therefore have Pk(Mn) = 0
(2{-ip g K
2*p; i = 1, 2, • • •)•
n
Hence Pk(M ) = 0 for all k = p/2, and the theorem is proved. BIBLIOGRAPHY 1. C. B. Allendoerfer and A. Weil, The Gauss-Bonnet theorem for Riemannian polyhedra, Trans. Amer. Math. Soc. 53 (1943), 101-129. MR 4, 169. 2. , Global theorems in Riemannian geometry, Bull. Amer. Math. Soc. 54 (1948), 249-259. MR 10, 266. 3. A. Avez, Formule de Gauss-Bonnet-Chern en metrique de signature quelconque, C. R. Acad. Sci. Paris 255 (1962), 2049-2051. MR 26 #2993. 4. A. Borel, Sur une generalisation de la formule de Gauss-Bonnet, An. Acad. Brasil. Ci. 39 (1967), 31-37. MR 36 #5866. 5. Y. K. Cheung and C. C. Hsiung, Curvature and characteristic classes of compact Riemannian manifolds, J. Differential Geometry 1 (1967), 89-97. MR 36 #827. 6. S. S. Chern, A simple intrinsic proof of the Gauss-Bonnet formula for closed Riemannian manifolds, Ann. of Math (2) 45 (1944), 747-752. MR 6, 106. 7. , Topics in differential geometry, Mimeographed Notes, Institute for Advanced Study, Princeton, N. J., 1951. MR 19, 764.
536
C.-C. HSIUNG AND J. J. LEVKO III
8. , On curvature and characteristic classes of a Riemannian manifold, Abh. Math. Sem. Univ. Hamburg 20 (1955), 117-126. MR 17, 783. 9. , Pseudo-Riemannian geometry and the Gauss-Bonnet formula, An. Acad. Brasil. Ci. 35 (1963), 17-26. MR 27 #5196. 10. F. D. Murnaghan, The generalized Kronecker symbol and its application to the theory of determinants, Amer. Math. Monthly 32 (1925), 233-241. 11. J. A. Thorpe, Sectional curvatures and characteristic classes, Ann. of Math. (2) 80 (1964), 429-443. MR 30 #546. LEHIGH UNIVERSITY, B E T H L E H E M , PENNSYLVANIA
18015
J . DIFFERENTIAL GEOMETRY 5 (1971) 383-403
COMPLEX LAPLACIANS ON ALMOST-HERMITIAN MANIFOLDS CHUAN-CHIH HSIUNG & JOHN J. LEVKO III
Introduction In [2] Hsiung (i) defined a new complex Laplacian n 2 for an almostHermitian structure, which is different from the one, denoted by Di, given by Kodaira and Spencer [3], (ii) verified for n 2 the well-known conjecture that if • 2 = d/2 for all 0- and 1-forms, where J is the real Laplacian, then the structure is Kahlerian, (iii) studied the conditions for n 2 to be real for all 0and 1-forms. Very recently, Ogawa [5] continued Hsiung's work to show that if either • 2 or • , is real for all 0- and 1-forms, then the structure is Kahlerian. The purpose of this paper is to introduce three more complex Laplacians n 3 , D 4, D 5 for an almost-Hermitian structure and to study the conditions for these Laplacians to be real, together with some relationships among all D ' s . We shall continue to use Hsiung's method [2] which is somewhat different from Ogawa's, and also for completeness we shall reprove Ogawa's result here. § 1 contains fundamental notation and real operators on a Riemannian manifold. In § 2 we define various almost-Hermitian structures first and then some complex operators for an almost Hermitian structure leading to the complex Laplacians • t, i = 1, • • •, 5. Some conditions for the tensor of an almostHermitian structure to be Kahlerian are also given for use in the proofs of our main theorems. § 3 is devoted to the computation of • 4f and • tq, i = 1, • • •, 5, for any 0-form £ and l-form -q on an almost-Hermitian manifold. In § 4 we show that for an almost-Hermitian structure if the complex Laplacian Qi, i = 1,2 or 4 is real with respect to all 0- and 1-forms, then the structure is Kahlerian. In § 5 we obtain the following relationships among the • 's: If for an almost-Hermitian structure the relation Im • 1 — Im Q { (i = 2 or 4) or Im D 2 = Im Uj (7 = 4 or 5) holds for all 0- and 1-forms, where Im denotes the imaginary part, then the structure is Kahlerian. Throughout this paper, the dimension of a manifold Mn is understood to be n > 2, and all forms and structures are of class at least C2.
Communicated July 29, 1970. Research partially supported by the National Science Foundation grant GP-11965.
384
CHUAN-CHIH HS1UNG & J O H N J . LEVKO III
1.
Notation and real operators
Let M" be a Riemannian manifold of dimension n(>2),\\gij\\ with gtj = gj( be the matrix of the positive definite metric of M", and \\gis || be the inverse matrix of \\gi}\\. Throughout this paper all Latin indices take the values 1, • • • ,n unless stated otherwise. We shall follow the usual tensor convention that indices can be raised and lowered by using gij and giS respectively, and also that when a Latin letter appears in any term as a subscript and superscript, it is understood that this letter is summed over its range. Moreover, if we multiply, for example, the components atj of a tensor of type (0, 2) by the components bjk of a tensor of type (2,0), it will always be understood that / is to be summed. Let Jf be the set {1, • • - , « } of positive integers less than or equal to n, and let /(p) denote an ordered subset {iu • • •, ip} of the set Jf for p < n. If the elements ix, • • •, ip are in the natural order, that is, if it < • • • < ip, then the ordered set I(p) is denoted by /„(/?). Furthermore, denote the nondecreasingly ordered ptuple having the same elements as I(p) by (p)>, and let l(p; s \ j) be the ordered set I(p) with the s-th element is replaced by another element / of Jf, which may or may not belong to I(p). We shall use these notations for indices throughout this paper. When more than one set of indices is needed at one time, we may use other capital letters such as J,K,L, • • • in addition to / . At first we define 0, if (p)> ± <X(p)>, J.<J» |0> if J(p) or K(p) contains repeated integers, «K(P) - + 1 or — 1, if the permutation taking J(p) into K(p) is even or odd. By counting the number of terms it is easy to verify that
a
T\ •Al
/(p)J(n-p).l"-n e
l".n
e
I(p)K(n-p)
— n | J(n-p) — V • sK(.n-p) >
On the manifold Mn, let V denote the covariant derivation with respect to the affine connection r, with components r)k in local coordinates x1, • • • ,xn, of the Riemannian metric g, and let $ be a differential form of degree p given by
(1.4)
where ^ / ( p ) is a skew-symmetric tensor of type (0, p), and we have placed (1.5) Then we have
dxI(p) = dx*i A ••• A dxlP .
385
COMPLEX LAPLACIANS
d
(1.6) where (1.7)
n!
Denote (1.8)
= £/V;r( d e t (Sij))
eIm
1/2
Then by using orthonormal local coordinates xl, • • • ,xn and relation (1.2) we can easily obtain \L-y)
e
Hp)K(n-p)e
— P • ^K(.n-p) •
The dual operator * is defined by (see, for instance, [6]) (1.10)
* <j> = (* ^ W p A 7 0 ' " " 3 " ,
where (1.11)
(*
JW
reJ(p)Kn-p)9'
From (1.10), (1.11) it follows that for the scalar 1 (1.12)
* 1 = (det (gtjWdxl
A • • • A dxn ,
which is just the element of area of the manifold Mn. By using orthonormal local coordinates x\ • • •, xn we can easily verify that (1.13)
**0 = (_1)J><»-I»0
Denote the inverse operator of * by *_1. Then from (1.13) it is seen that on forms of degree p (1.14)
*- l = (_l)p<»-»>* .
The codifferential operator 8 is defined by (1.15)
5$ = (-iy+n+1*-ld*
.
Making use of (1.6), (1.7), (1.10), (1.11) we obtain immediately (1.16) where
8$ = (^) 7 O ( P VJC 7 O < 2 '" 1 > ,
386
CHUAN-CHIH HSIUNG & J O H N J . LEVKO III
(1.17)
(<¥W» = - ^ / « - i ) •
For a form $ of degree p defined by (1.4) we can obtain
(1.18) +
ZJ 9l(r;s\a,i\l))R s<(
ish '
where J is the Laplace-Beltrami operator defined by (1.19) and
J = 3d + dd ,
(i.20)
(i.2i)
v> =
g'*rt,
Rtjkl = erydx1 - drydx* + r%ru - r%nk,
(1-22)
Rjk = 2.
R°jks.
Complex structures and operators
On a Riemannian manifold Mn with metric tensor gtj, if there exists a tensor i V of type (1,1) satisfying (2.1)
Ft%* = - e f ,
then Fij is said to define an almost-complex structure on the manifold Mn, and the manifold Mn is called an almost-complex manifold. From (2.1) it follows that the almost-complex structure FJ induces an automorphism J of the tangent space of the manifold Mn at each point with P = —1,1 being the identity operator, such that, for any tangent vector vk, (2.2)
/ : v" -^Ffv1 .
If an almost-complex structure FtJ further satisfies (2.3)
S « i W = 8uu ,
then Fi] is said to define an alrnost-Hermitian structure on the manifold Mn, and the manifold Mn is called an almost-Hermitian manifold. From (2.1), (2.3) it follows that the tensor Ftj of type (0,2) defined by (2.4)
FtJ = gjkFt"
is skew-symmetric. Thus on an almost-Hermitian manifold we have the associated differential form
387
COMPLEX LAPLACIANS
(2.5)
co = Fijdx* A dx' .
By using the multiplication of matrices, from (2.1) we readily see that a necessary condition for the existence of an almost-complex structure on a Riemannian manifold M" is that the dimension n of the manifold Mn be even. It should also be remarked that an almost-complex manifold is always orientable, and the orientation depends only on the tensor F / . An almost-Hermitian structure F / defined on a manifold Mn is called an almost-Kahlerian structure and the manifold Mn an almost-Kahlerian manifold, if the associated form co is closed, that is, (2.6)
dco = 0 .
From (2.5), (2.6) it follows that an almost-Kahlerian structure Ftj satisfies (2.7)
Fhij = VhFi} + VtFjh + FjFhi = 0 .
The tensor Fhij is obviously skew-symmetric in all indices. An almost-Hermitian structure FJ (respectively manifold) satisfying (2.8)
Fi = -PjFt> = 0
is called an almost-semi-Kahlerian structure (respectively manifold). In particular, the structure FJ is Kahlerian if F j F / = 0. In this case, by means of (2.1) it is easily seen that the torsion tensor ttjk = Fjh(dFik/dxh
- dF^/dx1)
- Fih(dFjk/dxh
-
dFhk/dx^)
vanishes, so that the integrability condition of the almost-complex structure F / is satisfied. But in general when ti}k = 0, the almost-Hermitian structure FJ is defined to be Hermitian Multiplying (2.4) by Fhi we obtain (2.9)
FtjFM = -s* .
By taking covariant differentiation of both sides of (2.9), noticing that (2.10)
FW.F.j = 0 ,
and making use of (2.7), (2.8) it is easily seen that (2.11)
Fhi3F" = 2FhiFi .
Thus an almost-semi-Kahlerian structure F t J satisfies (2.12)
FMiF" = 0
388
CHUAN-CHIH HSIUNG & JOHN J . LEVKO III
Multiplication of (2.11) by Fkh and use of (2.9) give (2.13)
Fk = -±FMjF*JFk*
.
From (2.7), (2.8), (2.13) we hence conclude that an almost-Kdhlerian structure or manifold is also almost-semi-Kdhlerian. In the proofs of our theorems we shall need the following lemmas. Lemma 2.1. An almost-Hermitian structure F satisfying (2.14)
PtFJk = V}Fik
is Kdhlerian. Proof. From the skew-symmetry of Fu we have (2.15)
ViFjk + ViFkj = 0.
Taking the sum of (2.15) and the two similar equations obtained from it by cyclic permutation of the indices i,j,k, and making use of (2.14) we obtain ViFjk + VtFkj + VjFki — 0, which together with (2.15) implies immediately V,Fkt = 0. Lemma 2.2. An almost-Hermitian structure F satisfying F"VkVkFiS
(2.16)
= 0
is Kdhlerian. Proof. From (2.9) we have 0 = PWk(FtjF")
= 2 ( F " P F t F w + V^^F**)
,
which together with (2.16) gives VkFuVkF^ = 0 and therefore F 4 F„ = 0. Lemma 2.3 (5. Koto [4]). An almost-Hermitian structure F satisfying (2.17) (2.18)
VtFf + VjFf = 0 , Rhi=
-^RnjuF^F^
is Kdhlerian. Proof. (2.17) can be written as (2.19)
VtFjk = VkFtj .
Multiplying (2.19) by Fi}, using (2.10) and taking the covariant derivative Pt of the resulting equation, we obtain, in consequence of (2.19), (2.20)
FWfiFj,
+ VJFijVF3
= 0•
COMPLEX LAPLACIANS
389
On the other hand, using (2.19) and the relation -FijVjVtFkl from the Ricci identity it follows respectively that (2.21)
V.V^FJU = rt?jFtl
(2.22)
FWtFjFkl
+ RajuFka
- RattiFf
= -iFU(R*kiiFal
+ RamFka)
=
FWiVjFkl,
, .
Similarly, the Bianchi identity leads to 2RhtJkF*> = RhmFij RhJlkF" = {Rhm + Rhikj)F^ - 2RhktjFiJ
,
and therefore to (2.23)
RhmF"
= -lRKttJFi
.
Substituting (2.21) in (2.20) and using (2.22), (2.23), (2.1) we can obtain (2.24)
V*FitVJF*i = Rkl + F^R^F^
- RaliJFk«)
.
Interchanging k, I in (2.24) and subtracting the resulting equation from (2.24) we have (2.25)
RattjFfF** = RallJFk"F*J ,
and therefore (2.24) is reduced to (2.26)
FkFi3FlF^
= Rkl - \Rali}Fk-F^
,
which together with (2.18) implies (2.27)
VfrfF' = 0 .
Multiplying (2.27) by gkl we hence obtain VkFiS = 0. Lemma 2.4. For an almost-Hermitian structure F, condition (2.28)
Ft*R„ =
Fk'R*tJl
implies condition (2.18). Proof. Since F Rkjhl — %F (Rkjhl — Rljhk) — ^F (Rkjhl + Rkhlj)
=
%F Rkijh
390
CHUAN-CHIH HSIUNG & JOHN J . LEVKO III
by the Bianchi identity, from (2.28) we obtain (2.29)
Fj*Rhk = ^R}hklF"
.
Multiplying (2.29) by F / and using (2.1) lead immediately to (2.18). We now consider an almost-Hermitian manifold Mn with an almost-Hermitian structure F, and shall follow Spencer (compare [7, Chapter IX]) to introduce complex operators on the manifold Mn. At first we define
(2.30)
n^=
J(et^-V-lFt0
1,0
and its conjugate1 tensor (2.3D
n«' = n v = K V + 0,1
j=iFS)
1,0
A simple calculation gives the following identities:
= n**,
1,0
1,0
1,0
nv n/ = 0 , nv nv = nv-
(2.32)
1,0
0,1
0,1
0,1
0,1
Let p + a = p, p > 0, a > 0, set (2.33)
1
'•' 11 «i 0,1
and define [] Hv)JW
1
-°
-°
1 1 n„ 0,1
£
JJO(P)SO(»)
'
to be the identity for p — a = 0 and to be zero for either
p < 0 or a < 0. Then for a form
(2.34)
p,o
p,a
where (2-35)
( n 0 / < « = n iu>j'ip>*j.<» • p,a
p^a
We next define a complex covariant differentiator 1 Throughout this paper a bar over a letter or symbol denotes the conjugate of the complex number or operator defined by the letter or symbol.
COMPLEX LAPLACIANS
(2.36)
391
®i=X\ Wi . 1,0
and the corresponding contravariant differentiator
(2.37)
& = gik3ik = n ft' = n w • 0,1
1,0
The conjugate operators of &i and Q)1 are (2.38)
2, = n ,'P, , 0,1
& = WfV3 .
(2.39)
1,0
Now we define the complex analogues of the real operators d and 8 defined by (1.7), (1.15) respectively:
(2.40)
d, = s
n
(in.
(2.4D
d2 = s
n
dn .
p + o = p p + 2,a — 1
(2.42)
a, = n
n
p+a=p
(2.43)
/9,c
^n ,
p^a — 1
82 = E
n
/>,c
^n •
p + a = p p + l,c — 2
p,ff
The conjugate operators of d1; d2 and 51; 52 have the forms:
(2.44)
dx = s
n
<* n >
(2.45)
d2 = s
n
dn .
p + a = p p —l,tf+2
(2.46)
3, = 2
(3 — 1 , ( 7
p +ff= P
(2.47)
p,a
fl 5 f l ,
g2 = n
n
|0,<7
^n •
Furthermore, for a p-form 54 given by (1.4) we define (2.48)
(di0)7(P+i) = (2d2 + ^1 — d2)j(P+1} ,
(2.49)
(#^)/ ( P _i, = (252 + a, - ?,)„„_„ ,
(2.50)
(2.5D
(3^)7 (P+ i, =
S
II
p + o=p
p + l,a
(^)/ (J ,-i, = - s
/
W P ,
^O
W
>
n «(,-» , ' ,<, "®^/. w .
p+o = p
together with their conjugate operators:
p,o
392
CHUAN-CHIH HSIUNG & J O H N J . LEVKO III
(2.52)
(3i07 ( p + i, = (2^ 2 + d,-
(2.53)
(£0)/<„-i, = (282 + 3, - 52)7(p_1) ,
(2.54)
(5^)/< P+1) =
(2.55)
(^)/(,-i, = -
S
d2)I(p+u
,
/(P+»^,(P>^^.(„
I!
>
II * / < P - I / O ( P , ^ O ( P > •
2
It is known that (see [3], [5]) (2.56)
•$! = — * 3[ * ,
•$., = — * 92 * ,
and that (see [3]) if the structure F of the manifold Mn is Kahlerian, then d2(j> = d$ — 0 for any form 0, and therefore 9j = dx. Now we introduce the following complex Laplace-Beltrami operators: (2.57)
Ui = Qidi + diQi,
(2.58)
n3 = M
+
(5.59)
D4 = M
+ d& ,
(2.60)
D6 = *A + & •
(i=l,2),
^ ,
It should be noted that Qi was first defined by Kodaira-Spencer [3], and D 2 by Hsiung [2]. From [3] we know that d = 3 t + 3\. In order to apply 32 + 32) let £ be any 0-form. The we have, in consequence of (2.50), (2.36), (2.32), (2.30), (2.6i)
o£)tl = n
4l ^f
= Wi£
- S^iFtjrfi
,
which together with (1.6), (1.7) gives (2.62)
d$ = (3, + 32)f .
Similarly, for any 1-form v, using (2.50), (2.36), (2.33), (2.34), (2.35), (2.32) we can obtain
(a,?)*,*. = n if n «,*(^7* - ^ ) + o v ru* - n if n ^x7^* 1,0
(2.63)
1,0
1,0
i
.
which together with (1.6), (1.7) gives (2.64)
dij = i(9 2 + djij .
0,1
1,0
0,1
393
COMPLEX LAPLACIANS
The almost-complex structure F of the manifold Mn is said [3] to be (completely) integrable if and only if d\ = 0. Now by means of (2.61), (2.50), (2.30), • • •, (2.36) and the relation (2.65)
PtV£ = V}V£
for any 0-form £, an elementary but lengthy calculation gives 4(9
(2 66)
^li2
= (F
^ W _ *WV)^£ + V-IOW-JW)^ •
// d\ is real for any 0-form f, then by taking f = JE1 for any arbitary i with respect to any local coordinates xl, • • • ,xn, from (2.66) we obtain (2.14), and therefore by Lemma 2.1 the structure F is Kahlerian. 3.
Expressions for • 's
In this section we shall give expressions for • 4f and n i^» where i = 1, • • •, 4, and £ and ij are respectively any 0- and 1-forms on an almost-Hermitian manifold Mn with an almost-Hermitian structure F. 3.1. Laplacian n 2 . In [2, pp. 146-147] we obtained (3.1)
4D 2 f = 2Jf + P F f c ' ( - F / P * S + J^Wtf 4(D2)?)4l = -FSW^Ftfj
- FjWtFsJF^
- 2Wjjtl
(3.2)
+ W„ViW
+ =l{piFSPjVil
Ft*F^PkVj
+
+
,
F^F^W^V^
- (FjFh* +
PuF/)Pk^
+ 2FjWJPkVh - Fh*[Pj, Pkty - F*Wt, PtM > where (3.3) 3.2.
Wh,Pt] = Laplacian Di-
VhFt-VtFK.
Atfirstwe notice that as a result of (2.65) we have
(3.4)
F.Wjf = 0 .
By using (2.57), (2.53), (2.45), • • •, (2.48), (2.33), ((2.40),(2.41), (2.43),(2.1), (2.30), (2.32), (2.34), (2.35), (3.4), (1.17), (1.18) we can obtain (3.5)
2D,f = 25 n d$ = J ? + V ^ T P ^ F ^ . 1,0
In order to compute Dtf, from (2.48), (2.52), (2.40), (2.41), (2.43), (2.45), (2.46), (2.47) we first see that
(3.6) 91 = 2 n ^ n + n ^ n + n ^ n - n ^ n > 2,0
0,1
2,0
1,0
1,1
0,1
0,2
1,0
fori-forms,
394
CHUAN-CHIH HSIUNG & J O H N J . LEVKO III
:
(3.7)
9l = 2l\8U
+ USU
0,1
2,0
1,0
+ l\SU-U§U,
2,0
0,1
1,1
for
1,0
2-forms .
0,2
Next, by means of (1.6), (1.7), (2.33), (2.34), (2.35), (1.2), (1.3), (2.30), (2.31), we obtain
n d n v = tn «,* n *.'^» n s - ?i n *')?, + n <,* n <,%^ - n <,* n U^VJ + n «.* n iWi n *' - ?* n / w * / o ( 2 '
1,1
0,1
1,0
1,0
1,0
(3.8)
0,1
0,1
0,1
1,0
1,0
0,1
0,1
0,1
= \{VtlVu - PuVil + F^Fu\FkVj
0,1
- FlVk)
+ V ^ f o ^ F , , ' - J \ , V + Fh"Ft2l(FkF^ - ^F**))
9 ( l l f 1,0 t 2,0
in 1,0
i1,0 n
i1,0 n
in 1,0
3
+ n u n 4l *(^* - ri9,)]d^«<» 1,0
(3-9)
1,0
= hWtWJ*'
- FkFtJ) + F,,*(F»F(l* - FhFk')]
+ 2Filr]t2 - 2Ptjh + 2FiJFi^(FkVj
- FjVk)
+ A/=T[ 9 ,(F,,F 4 I * - FtlFtJ + FSFtWM
-
FlFS)
+ 2FiJ(FhVj - FjVh) + 2FhW}Vh - PtlVj)]}dX'°«
n * n ? = r u * n iW* n i3 - Vi n *';M*7O<2> 2,0
0,1
1,0
(3.10)
1,0
0,1
= UVJIF^F^J
0,1
- FhFk0 + Fu*(ytlFt> - F*F(l*)] + F^Fi2WiFkj
- FkF^]}dx^
,
- u d U v = U u t u 0,2
1,0
(3.11)
0,1
0,1
1,0
= HvjlFi'iFuF^
1,0
- FkFtJ) + F^(VtFtJ + F^FuKFM
4([1 dd n ?)it = i>s*i}ViV*FJ + 1,0
(3.12)
- rtlFk>)l
- F.F,*) W
( , )
,
F^FJVm,
1,0
+ F^P^ftjj + V^l(v/i/kFkj
- FhFkVk + FiilFk^FlF"rli + PkFkJPilVj +
+ Fkiphp% + / y r ^ v ) .
Pif.Wyj
395
COMPLEX LAPLACIANS
Substitution of (3.8), • • •, (3.11) in (3.6) thus gives 3.7 = WuVu - FHVU + V ^ l f c / W -
(3 13)
fifiS*
Now put AM = Fiji, + V ^ G ? / * / , , ' + F , , ' ^ , , ) , BJ* = eX" - *?#" + FhlFsk2 - FsklFi^ ,
(3.14)
fkikt
Jif
tj
c *iF
*2
kl
Then (3.15)
3 , , = i(Ailh
- Ahil)dx!°™ .
By means of (3.13), (2.33), (2.34), (2.35), (2.30), (2.31), (1.16), (1.17), elementary but rather lengthy calculations give
0,1
1,1
= 2 [1 F'kNfc - efc*« + F,*'Ftl»* - F^'F,*')^*,] 0,1
+ + + + +
(3.16)
+ F^CF'F.Fy - FV S F/) + FWFtFtJ - FsFhF^)] (2FtllF°Fs* - F,T'F 4l *)F t7l + F ^ F f y ^ F . ' - 2F.,F/) F*VixFxiV>-q, - F^VFr'Vfl. + 2FsFsrju 2Fil*FsWlVk + S^lfalFWFtWtFS - F«F,0 F ll *F'F,'(F I F^ - F t F,') + FV.F,/ - F ^ F , ' Fil^Fs\VsFlFki - PF.F.0] + (2FSF4/ r^F.Wy, s l s 2F F/FjVtl + F'FtJPjV, - F^Fs (FkF^F Vj + F'FtVjVl) 2FiJFsFsVj - 2FsWjVh} ,
- 8 ( 1 1 * II 3."?)*, 1,0
2,0
= n^'tW'S + V-lCf>J;M*l4l] 1,0
=
ftU^W/V - F*F,') + ^ / F / P F A F ^ - F t F,0 + F'F*(ytFu' - FtlF»>) + Ftl\VVJFk> - F'F»F,') + Fs\FsFkFtJ - F°FhFki)] + 2FhWsF/(FkVl - Fm)
396
(3.17)
CHUAN-CHIH HSIUNG & JOHN J . LEVKO III
+ + + + +
F^VrfWff + F S F/) + F^F^iF^ - 2FkVl) F*VtlFtiV% + 2F°FsVil - 2P°PilVs 2Ftl*Fsl(F°FkVl - F°FlVk) =i{vMFtl*VFr'>{Vf,' - VJFt*) + FifP'FWM s FlFkO + IF.WFtWM - FkFt0 + F FsFtJ WUFJ + FJFWVtFJ - F'FM)] + F'F^F^ l s VuF.W'Vi + Fi^Fs FkFliF rlj F.JF^F/F^ 2F°FsWilVj - FjVh) - 2F«FilWkFlW% 2Fs"(F°FilVk - FWkVil) + 2FtlW*FjVs - F°FsVj)} ,
-8(nan3i7)<. 0,1
(3.18)
2,0
= ViWK'Wf*' - P*F^ + W V W ( f W - F,F»')] + F tF F > ' *Wfl* - V ^ + FHrPsFrk(FsVk - FkVs) + J=i{vMF,'
3 19
( - )
+
= 2 n F-KB** -
0,2
v^ic^^j
1,0
= toWF^WtF,' - F S F^) + FttF'VFWM i/^-LrtF'F'FtWM - P*F,*) + Ftl'P'Fr*(rjFS - VtF.*)\ .
- FkF^)]
Substituting (3.6), (3.7), (3.12), (3.16), • • •, (3.19) in (2.57) and using (2.32) and (3.20)
2Fj«PWkVtl = F'*W„ Fk]Vil ,
we can obtain, after some elementary simplification,
4(0,7),, = 4K2n an+ n3n + n«n-n«n)3i9 + n^ri9]i 1 0,1
2,0
= VJWFXW
(3.21)
+ + + +
1,0
2,0
0,1
1,1
1,0
0,2
- FkFh0 + VFtfWtFJ -
1,0
1,0
W)
+ Fs«(P°FhFki - PVkFtl') + FtllPJP*Ftr\ Fs*P°Vj(FhFkJ - 2FkF^) + F*yFtll(rkyi - FlVk) FillFlFkWkr]l + F^P'FfPtf, - 2P'P,Vtl [F„FilV + Fi/F«[Fi,FJ^ V ^ T ^ / F ^ / / - PF S F,/ + F , / S F / ) + 2F S F/F^ 4l
397
COMPLEX LAPLACIANS
- 2(FsFiJ + VJJWtf + F°WS,FAVH - F«W„ PJVJ - Fu'W., VM) • 3.3. Laplacian • 3. In the same way as above we can compute • t£ and D if], i = 3,4, 5, but we shall omit the details in this section and §§3.4, 3.5. We find that (3.22) (3.23)
u£ = Drf , k
1d.fi = -? V
4(38£tf)
2(dAv)u = 2K2 n s n + r u n + r u n - r u r m ^ 0,1
2,0
= F.iVFJWflt (3.25)
1,0
2,0
0,1
1,1
1,0
- Fkrjj) + FtWFrWfl,
- 2FsFsVil + FsFilVs -
0,2
-
FsFJFrVj)
Ftl*F,lVV^
+ i/=l\F,FilW.Vi - 2FjVs) + PSFSK2FJVU - Pitfj) - FtJFsFjVs + FJ(2FsFjVil - FsFilVj)] , 4(n3vK = VjF^FfF^ + 2FJFsFil"(FJrlk - FkVj) + F^WWri, - F«FkWrVj) + 2Ft;FsF^FjVs s - W*FsVh + 2F FilVs - FhF% + Fil^FkKFlF% - 2F"FlVj) + 2FsFiJ(FsVj - 2FjVs) + FuFkWkr,j + 2Fs'(2FsF3rjh - FsFuVi) + F.JiFjF^ - 2FWjVk) + Ft'rtlF%] . 3.4. (3.27) (3.28) (3.29) 3
Laplacian • 4. For D 4I, D fl we obtain the following equations: 2(3,0,, = Ftg - J=\Ftl>V£
,
•«£=•,£, 2Q2r, = -Fkr,k + V^lFkW%
,
4 ( 3 ^ ) , , = FtllFtF^F% - FtlF"Vk + FtllFk^,P% + V^liWP'T,, + FWn + F.JF/%)
,
,
CHUAN-CHIH HSIUNG & J O H N J . LEVKO HI
398
-Witfu (3.31)
-4(n
(3.32)
3.5.
= VJF^FW.F.J - WtlFt0 +
Laplacian D 5- Finally, for the remaining Laplacian • 5 wefirsthave
(3.33)
n,f = Dif ,
8lV = - , , P UtJ~U
(3.34)
1,0
i'V% • 1,0
Adding (3.8) to (3.9) gives 8(dlV)ilU = 4PilVi, - 4F ls7ll + 7u[Ftl"(PuF^ - FkFtJ) + Fu\FkF^ - PtlFtW
(3.35)
+ v^foP/V^FVOW -
PM
+ PifiJ - I W ] + 4FhWk%1 - 4FhWkVi,} . Now put (jJb)
,
+ V-l(3yjF^F^P^
+ y/ifij
+ 4F tl 'F, 7(l ) .
Then (3.37)
8dlV = (G4lll - GWlW°(2> .
As in the derivation of (3.16), (3.17) we can obtain (3.38)
- 2 ( [ ] 8 n dlV)h = ft PKef'eJ? - 4 # * + F/'F,,*' - F ^ ' F , * ' ) ^ ] ,
399
COMPLEX LAPLACIANS
(3.39)
-4(n s n dlV)it = n vim + V - I C ^ G . ^ J , 1,0
where B^ thus have
2,0
1,0
1 2
and C* ^ are defined in (3.14). After some calculations we can
16(0^ = i6[(n 8 n + n « n n ? + n ^ l i , 1,0
2,0
0,1
= nWF^WM
1,1
1,0
FtlrFs"FsFrl(FkF^
- FkF,0 +
+ 4F i l *(PF,iV - V'VtF,' + FkFlF^)]
- ViFS
+ %Fu*vfk>v% - wv,-nu + wk, rtl]if (340)
+ 4FulF*'Wi, Fk\rlj + V=Tfo,[4F(l*F«F,«(F»F{* - FtFki) + FjVFrWf,' k
l
+ 3F, VFtl (ylFlt*
-
- FkF^
Vfk*) + 4PtlV*Fk>
l
+ 4F tl *F, P(|7 4 F l * - FtFkJ)] + l
+ AF^F^{FrFs
l
+ FsFr WiV* ~ Wifj" l
+ 4F**[Fll, Fk]Vj + 4F <1 [F l , W 4.
WF'V^ +
FjF^F^
+ 4F'*[F„F,]^ 1 } .
Realization of • 's
Theorem 4.1. The complex Laplacian n«> i = 1, • • •, 5, /or an almostHermitian structure is real with respect to every 0-form if and only if the structure is almost-semi-Kahlerian. Moreover, with respect to every 0-form, if n « i = 1, • • •, 5, for an almost-Hermitian structure is real, then r j 4 = A/2 for i = 1, • • - , 5 . Proof. The theorem follows immediately from (3.1), (3.5), (3.22), (3.28), (3.33) and (2.8) by choosing the 0-form f to be xk for an arbitary k with respect to any local coordinates x\ • • •, xn. Theorem 4.2. For an almost-Hermitian structure, if the Laplacian • t, i = 1,2 or 4, is real with respect to all 0- and l-forms, then the structure is Kdhlerian. Kodaira and Spencer [3] have shown that if the relation (4.1)
Di = A/2
holds for an almost-Hermitian structure, then the structure is integrable. The particular case of Theorem 4.2 in which (4.2)
Di = A/2
( i = 1,2 or 4)
holds was a conjecture for some time; it was proved by Hsiung [2] for / = 2 and by A. W. Adler [1] for i = 1 by a different method under a stronger as-
400
CHUAN-CHIH HSIUNG & J O H N J . LEVKO III
sumption that (4.1) holds for a Hermitian structure and all 0-, 1- and 2-forms. Theorem 4.2 was proved by Hsiung [2] and Ogawa [5] for i — 2, and by Ogawa [5] for i = 1 by a somewhat different method. Proof, (i) i = 2. In [2, p. 148] Hsiung proved that under the assumption of the theorem the structure F satisfies2 (2.17) and (2.28). Then the theorem follows immediately from Lemmas 2.4 and 2 . 3 ; this was pointed out to one of the authors by H. Wakakuwa. (ii) i = 1. Using the Ricci and Bianchi identities and (2.23) we can easily obtain (4-3)
FhWj,
W
= Ftl*Rt%
,
F*'Wt, FtMj ~ FS*W» Ft]Vtl = - iF"lR\lklVj
(4.4) (4.5)
VJTtlF" ~ VtytF"
= FJRh«
sa
- \F WilSa
, .
By assumption, for any 1-form rj, Im n ^ = 0 which is reduced to, in consequence of Theorem 4.1, (2.8), (3.21), (4.3), (4.4), (4.5), (4.6)
2(FsFtJ + FhFsWjVs
+ FhkRk^
+ WfiJ
- R^F^Vj
= 0.
By choosing (4.7)
rj = dxh ,
for an arbitary h
with respect to any local coordinates x\ • • •, xn, from (4.6) it thus follows that (4.8)
VSVSF^ + F^RS
- R^FS
= 0 .
Multiplying (4.8) by Fhil and using (2.1) we obtain (2.16), and therefore the structure F is Kahlerian by Lemma 2.2. (iii) i = 4. At a general point P of the manifold Mn we choose orthogonal geodesic local coordinates x\ • • •, xn so that (4.9)
8ij(P)
= 3tJ,
ruP)
= o,
where 8iS are Kronecker deltas. By using Theorem 4.1, and choosing rj to satisfy (4.7) first and then (4.10)
rj = xhdxl ,
for any fixed distinct h and /
with respect to the geodesic local coordinates xl, • • •, xn, from (3.32) the condition Im (CU??) = 0 for any 1-form rj is reduced to (4.11)
V'V,Ftlh - V'Viff
By mistake, (2.28) was printed as F^R'ju
= 0 ,
= FfWiu
in [2, p. 148].Q
401
COMPLEX LAPLACIANS
(4.12)
F*F4l, + IV iff
+ VlFi? = 0 .
Interchanging I, h in (4.12) and adding the resulting equation to (4.12) we obtain (4.13)
Viff + VfJ
= 0.
From (4.11), (4.13) it thus follows that VjV3F^ = 0 ,
(4.14)
and hence by Lemma 2.2 the structure is Kahlerian. 5. Relationships among • 's Theorem 5.1. // for an almost-Hermitian structure the relation (5.1)
ImDi = ImD<
(f = 2 or 4)
holds for all 0- and l-forms, then the structure is Kahlerian. Proof, (i) i = 2. From (3.5), (3.1) and condition (5.1) for any 0-form f, we have VhFh'V£ = 0 .
(5.2)
By choosing f = xi for an arbitary i with respect to any local coordinates x\ • • -,xn, from (5.2) follows immediately (2.8), which together with(3.2), (3.21), (3.20) reduces condition (5.1) for any 1-form -q to
(5.3)
0W
+ ViFWrf - (F'Vtf,* - WMv,
=0.
Choosing t] to satisfy (4.7) first and then (4.10) with respect to the local coordinates x\ • • •, x" defined by (4.9) we therefore obtain (4.11), (4.13), and hence the structure is Kahlerian for the same reasoning given in the proof (iii) of Theorem 4.2. (ii) / = 4. As in part (i), from (3.5), (3.28), (3.1) and condition (5.1) for any 0-form f, we obtain (2.8), which together with (3.21), (3.32) reduces condition (5.1) for any 1-form -n to (5-4)
O W - PFW)F^ = 0 .
By choosing r] to satisfy (4.10) with respect to the local coordinates x\ • • •, xn defined by (4.9), we have (5-5)
FlFhil - VhFHl = 0 .
Thus by Lemma 2.1 the structure is Kahlerian. Theorem 5.2. // for an almost-Hermitian structure either the relation
402
CHUAN-CHIH HSIUNG & J O H N J . LEVKO III
(5.6)
Im Q 2 = Im D 4
or (5.7)
Re D , = Re D ,
holds for all I-forms, where Re denotes the real part, then the structure is Kdhlerian. Proof. From (3.1), (3.32), by the same argument as in the proof of Theorem 5.1 for i = 4 it is easily seen that conditions (5.6), (5.7) imply ?hFtll = VhFhl = 0 ,
(5.8) (5.9)
/VW
- F^FjFtll = 0 ,
respectively. By multiplying (5.9) by Fkh, we can reduce (5.9) to (5.8). Hence by Lemma 2.1, the structure is Kahlerian under either (5.6) or (5.7). Theorem 5.3. // for an almost-Hermitian structure the relation (5.10)
Im D , = Im D s
holds for all 0- and l-forms, then the structure is Kdhlerian. Proof. From (3.33), (3.5), (3.1) and condition (5.10) for any 0-form f we obtain (2.8). Then by the same argument as in the proof of Therem 5.1 for i = 2, (2.8), (3.2), (3.40) reduce condition (5.10) for any 1-form i? to (5 11)
F
iTFrWkFsh
- VsFk») + 3 F S * P / V ( F ; F / - PkF^)
+ AFJF.WWtFf (5.12)
FtlrF'l(VTF,K
Multiplying (5.12) by F^Ff (5.13)
- Vf^)
= 0,
+ FsFr») = 0 .
and use of (2.1) give VtF^ + FkF/ = 0 .
Substituting (5.13) in (5.11) we can easily obtain (5.14)
2F^FslVsVkF^
- F/P'F^P.Ff
= 0.
Multiplying (5.14) by Fhil and using (2.1), (2.8), (5.13) we therefore have (5.15)
FsFhlPsF^
= 0,
which implies that VsFiyl = 0. Hence the structure is Kahlerian. q.e.d. Finally, it should be remarked that there are no theorems involving the Laplacian D 3 similar to Theorems 4.2,5.1,5.2,5.3. However, we have the following two theorems, the proofs of which are omitted.
403
COMPLEX LAPLACIANS
Theorem 5.4. (5.16)
// for an almost-Hermitian structure the relation ImD3 = ImD1
+ }Im
(M)
holds for all 1-forms, then the structure is Kdhlerian. Theorem 5.5. (5.17)
// for an almost-semi-K'dhlerian structure the relation
Im D 3 = Im • * + i Im ($&)
(i = 2 or 4)
holds for all 1-forms, then the structure is Kdhlerian. References [1] [2] [3] [4] [ 5] [ 6] [7]
A. W. Adler, Classifying spaces for Kdhler metrics IV: The relation A =2-\Z\* Math. Ann. 160 (1965) 41-58. C. C. Hsiung, Structures and operators on almost-Hermitian manifolds, Trans. Amer. Math. Soc. 122 (1966) 136-152. K. Kodaira & D. C. Spencer, On the variation of almost-complex structure, Algebraic Geometry and Topology, A Symposium in Honor of S. Lefschetz, Princeton University Press, Princeton, 1957, 139-150. S. Koto, Some theorems on almost Kdhlerian spaces, J. Math. Soc. Japan 12 (1960) 422-433. Y. Ogawa, Operators on almost Hermitian manifolds, J. Differential Geometry 4 (1970) 105-119. G. de Rham & K. Kodaira, Harmonic integrals, Mimeographed notes, Institute for Advanced Study, Princeton, 1950. M. Schiffer & D. C. Spencer, Functionals of finite Riemann surfaces, Princeton University Press, Princeton, 1954. LEHIGH UNIVERSITY
TRANSACTIONS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 163, January 1972
CONFORMALITY AND ISOMETRY OF RIEMANNIAN MANIFOLDS TO SPHERESO BY
CHUAN-CHIH HSIUNG(2) AND LOUIS W. STERN Abstract. Suppose that a compact Riemannian manifold Mn of dimension n>2 admits an infinitesimal nonisometric conformal transformation v. Some curvature conditions are given for M" to be conformal or isometric to an /r-sphere under the initial assumption that L„R = 0, where LV is the operator of the infinitesimal transformation v and R is the scalar curvature of Mn. For some special cases, these conditions were given by Yano [10] and Hsiung [2].
1. Introduction. Let M" be a Riemannian manifold of dimension n ^ 2 and class C 3 , (gfj) the symmetric matrix of the positive definite metric of Mn, and (g") the inverse matrix of (g„), and denote by Vi; RMjk, Rij = Rlciik and R = gijRi} the operator of covariant differentiation with respect to gtj, the Riemann tensor, the Ricci tensor and the scalar curvature of Mn respectively. Let d be the operator of exterior derivation, S the operator of coderivation, and A = dS + Sc?the LaplaceBeltrami operator. Throughout this paper all Latin indices take the values 1 , . . . , n unless stated otherwise. We shall follow the usual tensor convention that indices can be raised and lowered by using gij and gtj respectively, and that repeated indices imply summation. Let v be a vector field defining an infinitesimal conformal transformation on M". Denote by the same symbol v the 1-form corresponding to the vector field v by the duality defined by the metric of Mn, and by Lv the operator of the infinitesimal transformation v. Then we have (1-1)
Lvgi, = ViVj + VjVi = 2pgi,.
The infinitesimal transformation v is said to be homothetic or an infinitesimal isometry according as the scalar function p is constant or zero. On a compact orientable Riemannian manifold, an infinitesimal homothetic transformation is necessarily an infinitesimal isometry; see [9]. We also denote by Ldp the operator of the infinitesimal transformation generated by the vector field pl defined by (1.2)
Pl = gijPi,
Pi = Vy/>.
Received by the editors June 29, 1970. AMS 1970 subject classifications. Primary 53A30, 53C20; Secondary 54H15. Key words and phrases. Infinitesimal nonisometric conformal transformations, scalar curvature, lengths of Riemann and Ricci curvature tensors. O The contents of this paper were published as a Research Announcement in Bull. Amer. Math. Soc. 76 (1970), 1253-1256. (2) The work of this author was partially supported by NSF grant GP-11965. Copyright © 1972, American Mathematical Society
65
C.-C. HSIUNG AND L. W. STERN
66
[January
Let f/(p) and -qKv) be two tensor fields of the same order pikn on a compact orientable manifold Mn, where I(p) denotes an ordered subset {/j,..., ip} of the set { 1 , . . . , n} of positive integers less than or equal to n. Then the local and global scalar products <£, rj} and (f, 17) of the tensor fields £ and t\ are defined by (1-3)
<^7?> = ( l / j P ! ) ^ V ( p ) ,
(1.4)
(t,v)=
!
<£,*!> dV,
where dV is the element of volume of the manifold Mn at a point. We also define (1-5)
U\\ =pl
O.
From (1.3) and (1.4) it follows that (f, f) is nonnegative, and that (f, | ) = 0 implies that £ = 0 on the whole manifold Mn. In the last decade or so various authors have studied the conditions for a Riemannian manifold Mn of dimension « > 2 with constant scalar curvature R to be either conformal or isometric to an n-sphere. Very recently Yano, Obata, Hsiung and Mugridge (see [13], [10], [4]) have been able to extend some of these results by replacing the constancy of R by LuR = 0, where u is a certain vector field on Mn. The purpose of this paper is to continue their work, in particular Yano's [10], by establishing the following theorems. To begin we denote by (C) the following condition: (C) A compact Riemannian manifold Mn of dimension n > 2 admits an infinitesimal nonisometric conformal transformation v satisfying (1.1) with p # 0 such thatZ,vi? = 0. THEOREM
I. An orientable Mn is conformal to an n-sphere if it satisfies condition
(C) and (1-6)
(p^-^Rp*^
(1.7)
Lv(a2A+^^
^0,
BJ = 0,
where A and B are defined by A = RhiikRMik,
(1-8)
B = R»RU,
and a, c are constant such that c = 4a2 + (n-2)\2aZ (1.9)
l
bi + (z i= 1
i=1
(-l)«-liY 7
-2(b1b3 + b2bi-b5b6)
+ (n~l)
7 bf\ > 0, (=1
J
Us being any constants^). (3) An elementary calculation shows that c^O, where equality holds if and only if bl=---=bi, b5 = ba = 0,a=-(n-2)bi..
1972]
CONFORMALITY AND ISOMETRY OF RIEMANNIAN MANIFOLDS
67
For the case a # 0 , c-4a2 = 0 and the case a = 0, c - 4 a 2 # 0 , Theorem I is due to Yano [10]. THEOREM II. A manifold M* is conformal to an n-sphere if it satisfies condition (C) and any one of the following three sets of conditions:
(1.10)
V\V)(Rf)
= Rpgu
(fis a scalar function),
(1.11)
QdP = (2/») d{RP),
V.VX/Jp) = *V,VyP,
(1.12)
A,^u = a ^i;
(°= is a scalar function),
where Q is the operator of Ricci defined by, for any vector field u on Mn, Q:ui-+2Rijui.
(1.13)
For constant R, conditions (1.11) and (1.12) in Theorem II will lead to the conclusion that Mn is isometric to an n-sphere of radius («(«— l)/R)112; for this see [12]. III. A manifold Mn with constant R is isometric to an n-sphere of radius (n(n—l)/R)112 if it satisfies conditions (C) and (1.10). THEOREM
Theorem III is due to Lichnerowicz [6] when condition (1.10) is replaced by the following one: (1.14)
v is the gradient of a scalar function/, i.e., vf = V ; /.
For constant R, it is easily seen that condition (1.14) is a special case of condition (1.10). In fact, in this case by using (1.1) condition (1.10) becomes Vtt>y+Vyt;f = 2VfV;/, which is satisfied by vi = Vlf+ui where ux is any vector field generating an infinitesimal isometry. IV. A manifold Mn is isometric to an n-sphere if it satisfies condition (C),LdpR = 0, and THEOREM
AaBb = c = const,
(1.15) (1-16)
(la ( n - l ) M \-A+—B—)
C
=
2a(a + b)R2(a + "-1) „•+»-!(,,-1)-*'
where A, B are given by (1.8), and a, b are nonnegative integers and not both zero. For constant R, Theorem IV is due to Lichnerowicz [6] for a=0, b = 1 and due to Hsiung [2] for general a and b. The following known theorems will be needed in the proofs of our Theorems I-IV. A (YANO AND NAGANO [11]). If a complete Einstein space Mn of dimension n>2 admits an infinitesimal nonisometric conformal transformation, then Mn is isometric to an n-sphere. THEOREM
68
C.-C. HSIUNG AND L. W. STERN
[January
B (OBATA [7]). If a complete Riemannian manifold Mn of dimension n ^ 2 admits a nonconstant function p such that V4V,p= -c2pgit, where c is a positive constant, then Mn is isometric to an n-sphere of radius 1/c. THEOREM
C (TASHIRO [8]). If a complete Riemannian manifold Mn of dimension n>2 admits a nonconstant function p such that (1.17) V,V,P = -(l/n)gtAp, THEOREM
then Mn is conformal to an n-sphere. D (YANO [10]). An orientable manifold Mn is conformal to an n-sphere if it satisfies condition (C) and THEOREM
(i.i8)
( i ? i ; / J y - ( i / « ( « - i))i?V, i) = o.
2. Notation and formulas. In this section we shall list some known formulas (for the details of their derivations see Lichnerowicz' book [5, pp. 124-134] or Hsiung's paper [1]) which will be needed in the proofs to follow. Let u b e a vector field defining an infinitesimal conformal transformation on a Riemannian manifold M" so that (1.1) holds. Then we have (2.1)
P
=
-80/11,
(2.2) LvR\ik = -«gV,/,y + «JVlpfc-g„VfcVhp+gffcVyVV, where Vh=gihVu and 4 = 1 for h = k and = 0 for h + k. From (1.1) and (2.2) it follows immediately that (2.3) (2.4) (2.5)
LvRhm
= 2pRhm-ghkVipj+ghjViplc-gijVllpk
+ gili:Vhpj,
LvR{j = gi}Ap - (» - 2) ViPj, LVR =
2(n-l)AP-2RP.
For any scalar field / and vector field u on Mn, we have (2.6) (2.7)
A/ = - V % / ; (Au\=
-V^jUi
+ i(Qu)i,
where Q is the operator of Ricci defined by (1.13). A necessary and sufficient condition for a vector field v to define an infinitesimal conformal transformation on a compact manifold Mn is that it satisfy (2.8)
Av + (l-2/n)d8v=
Qv.
For an infinitesimal transformation v on a manifold M", we have (2.9)
ASy = ( l / ( « - l))R8v-(n/2(n-
l))LvR.
For any 1-form f on a compact orientable manifold Mn we have (2.10)
(A£ + (l -2/n) d8£- Q£, 0 ^ 0,
where the equality holds when and only when £ defines an infinitesimal conformal transformation on Mn.
1972] CONFORM ALITY AND ISOMETRY OF RIEMANNIAN MANIFOLDS
69
On the manifold M* consider the following tensors:
(2.ii) (2.12) (?
r„ = Thm =
R^-imRgu,
Rhm-{.lKn-\))R(giighK-gikghj),
W
„
Mjk = aThm + b±gnkTu - b2ghjTik + b3guThk - bigikThj + bbgMTjk - be gjkThi,
where a and b are constants. It is easily seen that (2.14)
gUTu = 0,
g**Thm = Tti.
Moreover, by (1.3), (1.5) and (2.13) we have (2.15)
C
\\W\\ = a>A + «- 2
-^B-1n\n—\n
(^L+SZ^)^, — 2)
where c is defined by (1.9). 3. Lemmas. Throughout this section Mn will always denote a compact orientable Riemannian manifold of dimension n ^ 2. LEMMA
3.1. Iffis
a scalar field on M" and A/=0, then f is constant.
Proof. From (2.6) and our assumption A/=0, it follows that A(/ 2 )= - V'V^/ 2 ) = -2(V'/)(V i /). By substituting V4(/2) for & in the well-known Green's formula (3.1)
f
V%dV=0,
-»M"
where & is any vector field on M", we therefore have
(3.2)
0=1" h(P)dV=-l\
(V'JWtfidV,
which implies that V f /=0 since (V'/)(V,/) is nonnegative. LEMMA
3.2. /"or an orientable Mn satisfying LdpR = 0 and (C) defined in §1, we
Acme
(3.3)
(RUVtViP +Rip/Kin-1),
P)
^ 0.
For constant 7?, Lemma 3.2 is due to Lichnerowicz [6]. Proof. By applying the integral formula (2.10) to the 1-form dp we have (3.4)
((2(n - l)/n)AdP - Qdp, dp) ^ 0.
On the other hand, covariant differentiation gives
(3 5)
VWWn-lMMp-QdpM = «2(n-l)/n)MP-Qdp,
dP>-«2(n-l)/n)AAP-8Qdp,
P >.
70
C.-C. HSIUNG AND L. W. STERN
[January
From (3.4), (3.5) and Green's formula (3.1) we thus obtain (3.6)
((2(n-l)/n)AAp-82/>, P) ^ 0.
Due to the assumption LvR = 0, (2.5) is reduced to (3.7)
AP =
RP/(n-l).
Since LdoR = 0 implies (3.8)
p'V f * = 0,
substitution of p2VfJ? for | f in Green's formula (3.1) gives (3.9)
(PAR, P) = 0.
On the other hand, by the second Bianchi identity we have (3.10)
V'Rti = $VtR,
which together with (3.8) implies (3.11)
/>'V>Ry = 0.
From (1.13), (3.11) and (3.7) follow immediately (3.12)
8QdP = - 2 V i ( J R > ) =
-2R"WiPl;
(3.13)
AAP = (l/(« - l) a )/J a p + (l/(« - 1))/>A/J.
Substituting (3.12), (3.13) in (3.6) and making use of (3.9), we hence obtain the required inequality (3.3). 4. Proofs of theorems. Proof of Theorem I. From (2.13) and the condition LvR = 0 it follows that (4.1)
LV\\W\\ = Lv(a2A +
C
-^-B^-
By means of (2.13), (2.12), (1.1), (2.3), (2.4), (3.7) we can easily compute LvWMjk (for the details see [3, p. 189]), and then multiplying both sides of the resulting expression by WMik and making use of (2.13), (2.12), (1.1), (2.15), (1.9) and R\jk = 0 an elementary but lengthy calculation yields (4.2)
WM'kLv Whm = 29\W\-
cT"ViPj.
Substitution of (4.2) in the well-known formula (4.3)
Lv || W || = 2 WM*LV WMk -%p\W ||
thus gives (4.4)
pZ,J^|| = - V | | » ' | | - 2 c p r ' V l f t .
392 1972]
CONFORMALITY AND ISOMETRY OF RIEMANNIAN MANIFOLDS
71
A straightforward computation and use of (2.6), (3.7), (2.11), (3.10) can easily show that V'(/? Pft ) = (VtR)ppl +
(4.5) (4.6)
nTaPp')
RPlpt-Rapa/(n-l),
= RiiP'p1 + PT"ViPj + ((« - 2)/2«)(Vjfl)pPi 1
By substituting (4.5) for (V^pp and (4.4) for pT'^tp, Mn and making use of (3.1), we thus obtain
(l/n)RPip\
in (4.6), integrating over
2c(tf„py-(l/n(n-l))*V.l) = 4(\\W\\, P*) + (LV\\W\\, P) + c(pipi-(l/(n-l))RP\
R).
Since (|| W\, p2) is nonnegative, from (4.7), (4.1) and our assumption (1.6), (1.7) we obtain (1.18). Hence by Theorem D, M* is conformal to an «-sphere. Proof of Theorem II. First suppose (1.10) holds. Then from (1.10), (2.6) it follows that A(Rf) = —nRp, implying, together with (3.7), that (4.8)
A(p + (l/»(«-l))/?/) = 0.
Thus by Lemma 3.1, p + (l/«(n— l))^?/is a constant. Using (1.10), (3.7) we therefore obtain ViPi = - ( l / « ( « - l ) ) V t V , ( * / ) = -(!/«)*« V Hence by Theorem C, M* is conformal to an n-sphere. Next suppose (1.11) holds. From the definition of A it follows that (4.9)
dAp = Adp,
which, together with (3.7) and the first equation of (1.11), implies (4.10) A dp + (1 - 21n) dhdP-Qdp = 0. Thus by the necessary and sufficient condition (2.8) we see that dp generates an infinitesimal conformal transformation on Mn so that Ldpgij = 2
V i P y =
where ^ # 0 in consequence of (3.7). From (4.11) and the second equation of (1.11) it follows that (4.12)
V4V,(*P) = R
Thus the condition (1.10) is satisfied for v = dp a n d / = p , and hence Mn is conformal to an w-sphere. Finally suppose (1.12) holds. Then (2.4) becomes (4.13)
agiy = g w A / > -(»-2)V 1 p y .
Multiplying (4.13) by g1' and using (3.7) we obtain (4.14)
« = (2/n)RP.
72
C.-C. HSIUNG AND L. W. STERN
[January
Substitution of (4.14), (3.7) in (4.13) thus gives (1.17), and hence Theorem C completes the proof of our theorem. Proof of Theorem III. It is exactly the same as that of Theorem II for condition (1.10) except that the application of Theorem C should be replaced by that of Theorem B. Proof of Theorem IV. Without loss of generality we may assume our manifold Mn to be orientable as otherwise we need only to take an orientable twofold covering space of Mn. On the manifold Mn consider the covariant tensor field T of order 2(2a + b): T •* (4.15) =
hiiii1k1---haiaiakauivyutlvb o. >> 1 i Rh,irjrk, 1 1 Rusvs— r=l s=l
n a
Jfa + b a. + &/"„ y\a 1 1 \Si,i,Sh,kT~ n l \ ~) r=l
Si,k,Sh,ir)
1) | [ s= l
Susvs-
From (4.15) an elementary calculation gives the length of T: [2(2a + b)]\(T, T} = AaBb-2aR*a
(4.16)
+b
>/na +
b
(n-l)a.
Thus by condition (1.15), LvR = 0 and the extension of formula (4.3) to the tensor T we immediately obtain (4.17)
LV
which implies (4.18)
«LVT, r>, P) = 2(2a + b)(P(T, T>, P).
On the other hand, from (2.3), (2.4) we obtain T T
'-'•o±
hiiiiiki,---haiajakauivi---ubvi, a = 2ap ] ^ [ Kirirkr r=l
b Yl s=l
R
USV,
a ~ 2 , [-^ftlUyifcl ' ' • - ^ f t , - 1 i r - i y r - l f c r - l r=l • («-ft r fc r V i r V J f p - g
M r
VJrVkrp+giriykVhrP
-girkr
Vjr
VhrP) b
v^-19)
• Rhr+l(r a.
+ lJr + lfcr +1 •
b
• • Rhaialaka] [ I RUsVs s=l
+ 1 1 ^-IhirUkr Z {-^UlUl" ' • • ^ u , - ! » , - ! r=l s=l
•[^u 5 ^A / 3 -(n-2)V„ s V U s P ]-i? U s + 1 „ s + 1 - • ./? U6V J a+
2(2a + /j)JR » y u "
l"— V
r=l
_
» s=l
By means of (4.15), (4.19), (3.7), (1.16), (1.8), (4.16) an elementary calculation yields (4.20)
«L„r, T}, P) = 2a(P<.T, J>, P) AaBb [2(2a + b)]
e + ^)(-v,v, P+ ^. f ).
1972]
CONFORMALITY A N D ISOMETRY OF RIEMANNIAN MANIFOLDS
73
By comparing (4.18) and (4.20), noticing that p^O, and making use of Lemma 3.2, we thus have
Multiplying (4.21) by r=2
s= l
and using (4.15) we obtain Rhh = Rghh/n, which implies Mn is an Einstein space. Hence, by Theorem A, Mn is isometric to an «-sphere, and our theorem is proved. BIBLIOGRAPHY 1. C. C. Hsiung, Vector fields and infinitesimal transformations on Riemannian manifolds with boundary, Bull. Soc. Math. France 92 (1964), 411-434. M R 31 #2693. 2. , On the group of conformal transformations of a compact Riemannian manifold, Proc. Nat. Acad. Sci. U.S.A. 54 (1965), 1509-1513. M R 32 #6372. 3. , On the group of conformal transformations of a compact Riemannian manifold. I l l , J. Differential Geometry 2 (1968), 185-190. M R 38 #1637. 4. C. C. Hsiung and L. R. Mugridge, Conformal changes of metrics on a Riemannian manifold, Math. Z. 119 (1971), 179-187. 5. A. Lichnerowicz, Geometrie des groupes de transformations, Dunod, Paris, 1958. 6. — , Sur les transformations conformes d'une variete riemannienne compacte, C. R. Acad. Sci. Paris 259 (1964), 697-700. M R 29 #4007. 7. M. Obata, Certain conditions for a Riemannian manifold to be isometric with a sphere, J. Math. Soc. Japan 14 (1962), 333-340. M R 25 #5479. 8. Y. Tashiro, Complete Riemannian manifolds and some vector fields, Trans. Amer. Math. Soc. 117 (1965), 251-275. M R 30 #4229. 9. K. Yano, On harmonic and Killing vector fields, Ann. of Math. (2) 55 (1952), 38-45. M R 13, 689. 10. , On Riemannian manifolds admitting an infinitesimal conformal transformation Math. Z. 113 (1970), 205-214. M R 41 #6114. U . K . Yano and T. Nagano, Einstein spaces admitting a one-parameter group of conformal transformations, Ann. of Math. (2) 69 (1959), 451-461. M R 21 #345. 12. K. Yano and M. Obata, Sur le groupe de transformations conformes d'une variete de Riemann dont le scalaire de courbure est constant, C. R. Acad. Sci. Paris 260 (1965), 2698-2700. M R 31 #697. 13. , Conformal changes of Riemannian metrics, J. Differential Geometry 4 (1970), 53-72. M R 41 #6113. LEHIGH UNIVERSITY, BETHLEHEM, PENNSYLVANIA
18015
Differential Geometry, in honor of K. Yano Kinokuniya, Tokyo, 1972, 145-161.
ISOMETRIES OF COMPACT HYPERSURFACES WITH BOUNDARY IN A RIEMANNIAN SPACE CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
1.
Introduction
In 1927 S. Cohn-Vossen [1] proved the following important rigidity theorem on surfaces: An isometry between two closed convex surfaces in a Euclidean 3-space R3 is either a motion or a motion and a reflection, or, equivalently, preserves the second fundamental form. In 1943 G. Herglotz [4] gave an ingenious proof of this theorem by introducing an interesting integral formula. It is natural to ask whether the converse of this theorem would be true or not. The first major result in this direction was given in 1957 by V. G. Grove [3] who proved the affirmative of the converse but with an extra condition as follows: A diffeomorphism between two closed convex surfaces in R3 preserving the second fundamental form and the Gaussian curvature is an isometry. Grove followed Herglotz's method to prove this theorem by deriving an integral formula similar to Herglotz's, and later H. F. Munzner [6] gave a simple derivation of Grove's integral formula. Very recently R. B. Gardner [2] has extended Grove's result to two closed convex hypersurfaces in a Euclidean (« + l)-space Rn+1. The main purpose of the present paper is to further extend Gardner's result to two compact convex hypersurfaces with boundaries under a general boundary condition in a locally flat Riemannian (n + l)-space. § 2 contains the characterization of the 1-forms defining the Levi-Civita connection of a given Riemannian metric on a Riemannian ^-manifold Mn. In § 3 the difference tensor of the connection forms of two metrics on the manifold Mn is defined, and the difference between the Riemann tensors as well as that between Ricci tensors associated with the two metrics are expressed in terms of this.difference tensor. In § 4 we define subscalar pairs of metrics on Mn, and derive an integral formula for a pair of Riemannian metrics on a compact oriented Riemannian n-manifold Mn with boundary dMn. § 5 deals with pairs of metrics inducing the same Research partially supported by the National Science Foundation grant GP11965.
146
CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
volume element on Mn; an integral formula for such a pair of metrics is deduced from the formula obtained in § 4, and curvature conditions are found for the equality of a pair of metrics, under a boundary condition, which induce the same volume element and form a subscalar pair on Mn with boundary 3M". § 6 is devoted to a proof of our main theorem by means of the last result of § 5 mentioned above, and to the establishment of two local theorems on the isometry of two convex hypersurfaces of a Riemannian (n + l)-space with constant sectional curvature. The authors wish to thank editors for their pointing out that for later application no boundary condition on the metrics need be used in Theorem 4.1 and that the condition "locally flat" should be imposed in Theorem 6.1. 2.
Connection forms of a metric
Throughout this paper we shall suppose that all Latin indices take the values 1, • • • , « , and Greek indices the values 1, • • •, n + 1 unless stated otherwise, and we shall follow the usual tensor convention that when a letter appears as a subscript and a superscript in any term, then it is understood that this letter is summed over its range. Let / be a Riemannian metric on a Riemannian n-manifold Mn, and (o)i) be a local coframe in terms of which | | $ | | is the matrix of 1-forms defining the Levi-Civita connection of /. We recall that if (2.1)
I =
gija>W,
then (j>) can be characterized by the following equations: (2.2)
dm1 = a>j A <j>) ,
(2.3)
dgij = gik
,
where A denotes the exterior product; (2.3) expresses the constancy of gtJ with respect to the covariant differentiation defined by the connection forms <j>) given by (2.2). Moreover, the Riemann and Ricci tensors Rijki and Rtj associated with this connection of / are defined by (2.4) (2.5)
6) = dfi - fj A
,
where (2.6)
Rlm
+ R*ii* = 0 .
A a>1 ,
COMPACT HYPERSURFACES
147
Throughout this paper we shall also use the tensor gi} and its inverse tensor gij to lower and raise indices of tensors respectively. 3.
Difference tensors
Let /, /* be two Riemannian metrics on a Riemannian n-manifold Mn referred to the same local coordinates (x1), and $), <j>*) be their respective Levi-Civita connection forms with respect to the coframe (o)i) dual to (d/dx1). Then define 1-forms 0) on Mn as follows:
(3.1)
where the tensor K)k is called the difference tensor of the connection forms $) and $*) of the metrics / and /*. For the metric /* we shall use (a)* to denote the equation similar to equation (a) for the metric /, where a is a pair of numbers such as 2.2, 2.3 or others. Proposition 3.1. K)k is symmetric in j and k. Proof. From (2.2,) (2.2)* and (3.1) it follows immediately that 0 = mj A (#• - <j>*)) = K%coj A wk , which implies that K)k is symmetric in / and k. We next derive an interchange formula. Proposition 3.2. Put 0 = ||0}||,0 = ||0}||,0* = ||0*j-||, where 0),O) and 0*) are respectively given by (3.1), (2.4) and (2.4)*. Then (3.2)
d* - 6 = DT0 - 0 A 0 ,
where Dj is the covariant differentiation with respect to the metric I so that D,0 = K%^ = -(dK)k
A a>1 - £ * , # - K\tf, + Kljk^f) A w« ,
the subscript " , " denoting the covariant derivative with respect to the metric I. Proof. Differentiating (3.1) and using (2.2) we obtain (3.4)
d # _ d^*) = (dK)k - K)^)
A a)* .
By means of (2.4), (2.4)*, (3.3) and (3.4), a straightforward calculation thus gives (3.5)
6*) - 0) = K)Kla>« Am1-
which is (3.2).
q.e.d.
( # - ^ ) A ( # - <j>*ld ,
148
CHUAN-CHIH HSIUNG & SITANSU S. M1TTRA
Substituting (3.1), (2.4), (2.4)* in (3.5) and equating the coefficients of a)k A a)1 on both sides, we can easily obtain (3.6)
R**Jkl - R*Jtl = K\ktl - £',•,* - Ky^
+ Kj.K^
.
By contraction in i and / from (3.6) we thus have, in consequence of Proposition 3.1, (3.7)
KJ M - K'Jltt = R*jk 4.
- Rjk - (KTjKi, - Kj.KjJ
.
Subscalar pairs of metrics
From the difference tensor K)k in § 3 define a tensor Cki by (4.D
C „ = K)kK{t - K*uK>Jt .
Then by Proposition 3.1 it is easily seen that Cki is symmetric in k and /. Consider the symmetric quadratic form (4.2)
C = C»iO)V .
/ and /* are said to form a subscalar pair of metrics if C is positive and semi-definite. Theorem 4.1. Let Mn be a compact oriented Riemannian n-manifold with boundary dMn, and I and I* be a pair of Riemannian metrics on Mn. Then f(tr 7 Ric* - tr 7 Ric - tr 7 QdV Mn
(4.3)
=
C J
f:(__1}k-1(dstgi.yn{Vk_Wk) k=l
dMn
co1 A
• • • A ft)*"1 A cu*-" A
• • • A con ,
where dV is the volume element of Mn with respect to I, Ric and Ric* are the Ricci tensors formed with respect to I and I* respectively, and for any two nonsingular matrices A, B (4.4)
tiAB = tr(A-,B)
,
tr denoting the trace and v and w are two vector fields on Mn with contravariant components vl and wl defined by (4.5)
vl = gikK)k
,
w* = giJK% .
149
COMPACT HYPERSURFACES
For an empty boundary dM" Theorem 4.1 is due to Gardner [2]. Proof. From (4.5), (3.7) it follows immediately that div7 (v — w) = v1^ — witi (4 6)
= 8jk[R*Jk
'
- Rj* - {KTjKl, - KjhK\J]
.
By using (4.1), (4.2), (4.4) and noting that we shall use the same notation to denote a quadratic form and its matrix, (4.6) can be rewritten as (4.7)
div7 (v - w) = tr 7 Ric* - tr 7 Ric - tr 7 C .
Now let us recall that for any vector field X on Mn, div X is denned by (see, for instance, [5, pp. 281-282]) (4.8)
(div X)dV = Lx(dV)
= (d o ix + tx o d)(dV) = d(ix(dV))
,
where Lx is the Lie derivative in the direction of X, and cx is the interior product with respect to X. Thus for any vector field X on Mn, Stokes' theorem gives
(4.9)
J(divX)dV=
f
Mn.
3Mn
lx(dV)
.
Let / be a 0-form, and <wr, w'r be r-forms. Then it is known that (see, for instance, [5, p. 35]) iz(f) = 0 ,
(4
cAcoJ =
Wl (Z)
,
cx(a>r A o0 = cx{wr) A w's + ( - \Ya,r A txWs) . By applying (4.10) to the known formula (4.11)
dV = ( d e t ^ ' V A • • • A a>»
we can easily obtain
,, , 0 , (4.12)
a)1 A • • • A cu*"1 A o)*+1 A • • • A con , where Z = Y!tlX{d/dxt. In particular, for X = v — w from (4.9) and (4.12) it follows immediately that
150
CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
/
(4.13)
= Jf k2(-D*- 1 (detft,) ,/8 (i>*-w*) dMn
o)1 A • • • A a)*'1 A a>k + 1 A • • • A
Metrics inducing the same volume element
Let / and /* be a pair of Riemannian metrics on a compact oriented Riemannian /z-manifold Mn, and choose a local coframe (of) such that / = gijotm3 ,
(5.1)
I* = gfjofw1 .
Then by (2.3) and (2.3)* satisfied by the Levi-Civita connection forms <j>) and $*) of / and /*, we easily obtain, in consequence of (3.1), J detg^\ (5 2)
{det8i3
'
=
d e t g ^ [ t r {g*jkdgf) d e t
_
t r {gjkdg
)}
^
= 2^^L{
- #) = _2-jLS« * > ' . det gu
Proposition 5.1. // two Riemannian metrics I and I* on a compact oriented Riemannian n-manifold M" induce the same volume element, then (5.3)
K*j = 0.
Proof. By (4.11) and (4.11)*, the assumption dV = dV* implies that detg^ = det gfj so that (5.2) becomes K^m3 = 0, thus giving (5.3). q.e.d. In this case, from (4.1), (4.2), (5.3) it follows that C = K%Ktl(o"a>1 .
(5.4)
Theorem 5.2. Let Mn be a compact oriented Riemannian n-manifold with boundary dMn, and let I and I* be a pair of Riemannian metrics on Mn inducing the same volume element. / / / = /* on dMn, then (5.5)
f[(tr 7 . - tr 7 )(Ric* - Ric) + (tr,, + tij)C]dV = 0 . Mn
For an empty boundary dMn Theorem 5.2 is due to Gardner [2].
151
COMPACT HYPERSURFACES
Proof. Under the hypotheses of our theorem we can apply Theorem 4.1 to get (4.3). Furthermore, by interchanging the roles of / and /* in (4.3) and using the hypothesis that / = /* on dM", from (3.1), (4.1), (4.2), (4.5) it follows immediately that f[tr 7 »Ric - tr 7 ,Ric* - tr 7 , C]dV* Mn
(5.6)
=
C 2(_1)i:-i(det^)v2(_^
+
w *)
J k=\ dMn a1 A
• • • A a)*"1
A cok+1 A
• • • A to" ,
where dV* is the volume element on Mn with respect to /*. Since dV* — dV, by adding (4.3) to (5.6) we hence obtain (5.5). q.e.d. By means of (4.5), (5.3), (5.4), the integral.formula (5.5) can be written in tensor notation as follows:
(5.7) J*[(g*« - *")(**« - Ru) + (g*tj + gfOKt&jW
=0.
Mn
Now we prove the following algebraic lemma. Lemma 5.3. / / A is a positive definite symmetric real matrix of order n, then (5.8)
t r ^ + t r ^ " 1 > In ,
where the equality holds if and only if A is the identity matrix. Proof. Since A is a positive definite symmetric real matrix, we can diagonalize A to the form A = ||a^|| where atj = 0 for i =£ / and au > 0. By the known fact that for any positive real number p, /* + l/fi>2, where the equality holds if and only if /j. = 1, we thus obtain t r ^ + t r ^ " 1 = 2 (a„ + l/a i 4 ) > 2n , i
where the equality holds if and only if au = 1 for all i. Hence our lemma is proved. Theorem 5.4. Let Mn be a compact oriented Riemannian n-manifold with boundary dMn, and I and I* be a pair of Riemannian metrics on Mn. Then I = I* on the whole Mn, if I = I* on dMn and ( i ) Ric — Ric* = X(I — I*), where A is a negative constant, (ii) / and I* induce the same volume element, (iii) / and I* form a subscalar pair.
152
CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
When dMn is empty, Theorem 5.4 is due to Miinzner for a 2-sphere M in the Euclidean space R3, and due to Gardner [2] for a negative Einstein metric I and an arbitrary Einstein metric /* on Mn. Proof. Under the hypotheses of our theorem we can have (5.5). Since by assumption (iii), C is positive semi-definite, from the known fact in linear algebra that the trace of a positive semi-definite matrix is nonnegative it follows that the third term of the integrand of (5.5) is nonpositive, so that n
(5.9)
f (tr7 - tr 7 .)(Ric - K\c*)dV < 0 , Mn
which is reduced by assumption (i) to (5.10)
f(tr 7 - tr7,)(7 - I*)dV > 0 . Mn
By assuming (5.11)
/ = &,o)W ,
/* =
ft>V
,
from (4.5) an elementary calculation shows that (5.12)
(tr7 - tr7.)(Z - /*) = 2n - g%gij - gug*i3 .
Since / and /* are both symmetric and positive definite, so is (5.13)
A = 1*1" = \\g*g^\\ .
Thus by Lemma 5.3 we have (5.8), which is equivalent to (5.14)
g*g*J + gtJg**i > In .
Combination of (5.10), (5.12), (5.14), (5.8) shows that the equality holds in (5.8). Since this is possible only when A is the identity matrix, from (5.13) we obtain that / = /* on the whole Mn. 6.
Isometries of convex hypersurfaces
Let Mn be a compact oriented hypersurface of a Riemannian (n + 1)space iVB+1. If Mn has a definite (positive or negative) second fundamental form, then Mn is defined to be a convex hypersurface of Nn+\ An immersion
X:Mn-^W
COMPACT HYPERSURFACES
153
is defined to be a convex immersion if Mn is a convex hypersurface of Nn+1. In this section we shall assume that X is a convex immersion and that Mn has a negative definite second fundamental form. We choose a local affine frame {ex, • • •, en+l) on the image X(Mn) such that ex, • • -,en are tangent fields of M" and e B+1 is the unit normal to Mn. Since Mn is convex, we can choose the local coframe (a>\ • • •, o)n+1) such that the negative definite second fundamental form / / of Mn is diagonalized as follows:
- / / = £ Wf .
(6.1)
i
n
Restricted to M we now get con+1 = 0 ,
(6.2)
and the structure equations of Nn+1 are (6.3)
dm" = co? A to". , ' dcol = (o'p A w° + i i ? " ^ ^ A wJ ,
where R"m is the Riemann tensor of Nn+1 satisfying (6.4)
R'frl
+ R«,ir = 0 .
The 1-forms a)% are the connection forms of Nn+1, and enable us to define the generalized covariant differentiation which is useful for studying submanifolds in a Riemannian space. In fact, let (6.5)
u — Wea
be a vector field on Nn+1. Then define its generalized covariant differential to be (6.6)
Du = (Dw)ea ,
where (6.7)
Du" = du" + u?m% .
In particular, for the vectors e„ themselves, (6.6) gives (6.8)
Dea = atfe, .
The generalized covariant differentiation can be defined for mixed tensor fields of higher orders. For definiteness consider a mixed tensor field A of order 3:
154
(6.9)
CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
A = 2 Aifi. ® a/ ® a)4 , A*
and define its generalized covariant differential to be (6.10)
DA'ti = d/4j f + A'H(o"r - A%a/f - A"fJm{ = A%>ico' .
Now suppose that the first fundamental forms / and I of M" and Nn+l are given respectively by (6.11)
I = aaf(o"m? .
/ = gijofo)* ,
Then it is easily verified that gi} and aaf can be treated as constants in the generalized covariant differentiation, i.e., (6.12)
Da.f = 0,
Dgi) =
0.
From (6.10) and (6.12) we thus obtain da
(6
13)
*i> = a.& + a/sr°»« > dgn = gum) + gjk«>ki •
In view of (6.13) and the first equation of (6.3) it follows that the LeviCivita connection forms of / and I are m) and <w£ respectively. Differentiating (6.2) and using the first equation of (6.3) we obtain (6.14)
d(on" = of A co?+1 = 0 .
By a lemma of Cartan on exterior algebra, (6.14) implies that (6.15)
hll =
hji.
Next we note that the relation between gtj and aaf is given by (6.16)
8tj = et-e} =
amfeie>i,
where the dot denotes the inner product of two vectors in the space Nn+1, and e\ are the components of the vector et. We also note that the second fundamental form 11 of Mn is defined to be (6.17)
~II =
DXDen+l,
where (6.18)
DX = dX = m% .
Then from (6.17), (6.18), (6.8), (6.16) it follows that
COMPACT HYPERSURFACES
(6.19)
- / / = aa,{DX)"{Den+iy
155
= g„coV„+1 .
On the other hand, since aa$ is constant with respect to the generalized covariant differentiation, we have 0 = D(en+1-ei)
= &X+1 + < + 1 >
so that (6.20)
= -<"? + 1 •
ftX+i
Substitution of (6.15), (6.20) in (6.19) thus gives (6.21)
- / / = -Aj/uV .
By comparing (6.21) with (6.1) we obtain (6.22)
htj = -8U
,
where 8tj are Kronecker deltas. Thus (6.22) reduces (6.15) to < + 1 = -co1 ,
(6.23)
which further reduces (6.20) to (6.24)
cokn+1 = 2 8ik^ • i
From now on we shall assume that Nn+1 is of constant sectional curvature c. Then (6.25)
R"M = c($an
- d"xafr) ,
where d° are Kronecker deltas. In particular, (6.26)
R*+\jk
= 0 .
Differentiating (6.23) and using (6.3), (6.26), (6.23) we obtain (6.27)
da? = - S ^ A
Addition of (6.27) to (6.3) or (6.28)
dco1 = a)1 A a)
gives (6.29)
dm* = coj A # ,
156
CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
where (6.30)
# = i(o,* - a>0 •
Since $ is skew-symmetric and htj = — 5^, (2.3) is satisfied for —//, and hence $) are the Levi-Civita connection forms of —// of Mn. Similarly, subtracting (6.29) from (6.28) we have (6.31)
Zco>A Wi
+ co{) = 0 ,
i
so that by Cartan's lemma we can write (6.32)
J(o)J + a>{) = P%a>* ,
where P)k is symmetric in / and k. Since Pljk is obviously symmetric in i and j from (6.32), it is symmetric in all three indices. Next we calculate the Levi-Civita connection forms of the third fundamental form / / / of M". By definition and (6.8), (6.16), (6.24) we have (6.33)
/ / / = Den+i-Den+1
= £ g"a>W ,
so that the metric tensor of / / / is gij. Differentiating ik
(6.34)
8ijg
= 5kj
and using the second equation of (6.13) we can easily obtain (6.35)
df
= -giko>{
- g3kcoi ,
which together with (6.27) shows that (6.36)
flj = -m{
are the Levi-Civita connection forms of / / / . By definition and (6.36), (6.3), (6.23), (6.24), the Riemann tensor Sjikt with respect to the metric / / / of Mn is given by *S*„,a»* Aco1 = dM - Vi A -H (6
"37)
= S 8'%^
A co1 - *#,«<»* A col .
i
Equating the coefficients of the general term a>k A wl on both sides of (6.37) we therefore obtain (6.38)
S'ttl = gil8Jk - gik5n
- & Jkl
)
157
COMPACT HYPERSURFACES
from which follows immediately the Ricci tensor Sik of Mn with respect to the metric / / / : (6.39)
Sik = S\kl
= - ( n - l)g» - Z
R'JUJ
•
j
Theorem 6.1. Let Mn and M*n be two compact oriented Riemannian n-manifolds with boundaries dMn and dM*n respectively, and let X: Mn-+Nn+1
,
X*: M*n-^ Nn+1
be two convex immersions in a locally flat Riemannian Nn + 1. Suppose that there exists a diffeomorphism
(n + l)-space
h: Mn -^M*n preserving the second fundamental form and the Gauss-Kronecker curvature. If h restricted to dMn is an isometry of dMn onto dM*n, then h is an isometry of Mn onto M*n. For Euclidean space Nn+1 and empty boundaries dMn and dM*n, this theorem is due to Grove [3] for n = 2, and to Gardner [2] for n > 2. Proof. We shall prove this theorem by using Theorem 5.4 for the respective third fundamental forms / / / and / / / * of Mn and M*n. To this end we first observe that since by the hypothesis of this theorem h restricted to dMn is an isometry of dMn onto dM*n, from (6.11), (6.11)*, (6.33), (6.33)* it follows that (6.40)
/// = ///*
on dMn .
Next, since the space Nn+l is locally flat, i.e., of zero sectional curvature, by putting c = 0 in (6.25) we have R"fri = 0. Thus subtraction of (6.39)* from (6.39) readily shows that the difference between the Ricci tensors Ric 7 / Z and Ric 7/7 « with respect to / / / and / / / * is StJ - S* = - ( « - l)(g" - g*«) , or (6.41)
R i c / / 7 - Ric777» = - ( « - 1)(//Z - ///*) ,
n - 1> 0.
Similar to (4.12), the volume elements of M" and M*n with respect to the metrics / / / and / / / * are given respectively by ^ (6.42) t£
dV = (det g ' 0 v V A • • • A o»" , dV* = (det gwy/'a,*1 A • • • A w*n .
158
CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
Since M" and M*n have the same second fundamental form, from (6.1), (6.1)* it follows that £ («/)2 = 2 ("*0 2 •
(6.43)
Without loss of generality we can suppose Mn and M*n to have the same local coordinates (xl) so that under the diffeomorphism h corresponding points have the same local coordinates, and further suppose that for all i, et(x) and ef (x) are unit tangent vectors of the ^-curves on M* and M*n respectively. Then from (6.18), (6.18)* we see that w* and o>*4 are multiples of dxl for all /, and therefore (6.43) holds only when a)i differs from cu** for all / at most by a sign. In other words, by changing the sign of the normal e*+1 of M*n if necessary, it is always possible to choose the affine frames (ef) such that co*4 = co*
(6.44)
(/ = 1, • • - , « ) .
Since the Gauss-Kronecker curvature K of Mn with respect to / is the product of the n eigenvalues of / / relative to /, i.e., the product of the n roots of the determinant equation in X: (6.45)
det (// - XI) = 0 ,
or
det (8U + Xgtj) = 0 ,
it is easily seen that (6.46)
K = (-l)"detg« .
Similarly, the Gauss-Kronecker curvature K* of M*n with respect to /* is (6.47)
K* = ( - l ) » d e t £ * ' J ' .
Thus the assumption that K = K* gives det giJ = det g*iJ, which together with (6.44) and (6.42) implies that (6.48)
dV = dV* .
Finally, the difference tensor of the connection forms •$•{ and c4/ of the forms / / / and —// given respectively by (6.36) and (6.30) is (6.49)
f{ -
Similarly, (6.50)
f*{ - # = - P * > * .
159
COMPACT HYPERSURFACES
Subtracting (6.50) from (6.49) we thus obtain the difference tensor of the connection forms ty( and f*{ of the forms / / / and / / / * : (6.51)
H - +*{ = flJX ,
where (6.52)
H% = P*% - P%
is symmetric in all three indices. Since / / / and ///* induce the same volume element, it follows from Proposition 5.1 that (6.53)
H\j = 0 ,
and therefore that for / / / and ///* the quadratic form C given by (5.4) becomes H)kHil(okcol =
£
fl*t#>W
,
which is positive semi-definite by diagonalizing the matrix \\H)k\\ for each i so that the (/,/') entry of the diagonal matrix || 2«,^ H^H^W is £ \ (#j 7 ) 2 Hence / / / and ///* form a subscalar pair of metrics, and by this together with (6.40), (6.41), (6.48) we can apply Theorem 5.4 to conclude that / / / = ///* or giJ = g*ij, which implies that gtj = gfj or / = /*. Corollary 6.2. Theorem 6.1 is still valid if in its statement the GaussKronecker curvature is replaced by the volume element. Proof. This corollary follows immediately from (4.11), (4.11)*, (6.44), (6.46), (6.47) and Theorem 6.1. Finally we shall establish two local theorems on isometry by using the Ricci tensors with respect to the first and second fundamental foons / and / / of a convex hypersurface Mn of a Riemannian (n + l)-space Nn+1 of constant sectional curvature c. To this end let us first recall that the LeviCivita connection forms of / and —// are respectively m) and (j>) given by (6.30). By means of (2.4), (6.3), (6.23), (6.24), (6.25) we can easily obtain the Riemann tensor i?*^; of the hypersurface M" with respect to / : (6.54)
R*JH = -Sitgil
+ 8jl8ik + c{dian - 8\ajlc) ,
from which follows the Ricci tensor Rjk of Mn with respect to / : (6.55)
RJk = Rtju = -5Jte S g« + g>* - c(n - l)ajk . i
Being the sum of the n eigenvalues of / / relative to /, i.e., the sum of the n roots of the determinant equation (6.45) in X, the mean curvature H of Mn is given by
160
CHUAN-CHIH HSIUNG & SITANSU S. MITTRA
(6.56)
nH = t r z / / = - t r 7 ( - / / ) = - 2 *« , i
which reduces (6.55) to (6.57)
Rjk = nH8jk + gi* - c(n - \)ajk .
Similarly, after an elementary calculation we can obtain the Riemann tensor of Mn with respect to —//: Ti
pmpj
•*• j k l —
l
pmpj
ilL mk
l
ik1 ml
+ *(*«*" - 8jkgu - 8ugik + 8Jlg«) + ic(8ilcajl - 5{au - d\ajk + 8{aik) ,
(6.58)
where P)k is defined by (6.32). By putting Bjk = PTjPmk — P%Plim
(6-59)
we thus have the Ricci tensor T]te of Mn with respect to —//: jk
—
=
1
Ti
jkl
B
(6.60)
= i* - i ( " - 2)g^ + \nH8jk + Jc[(2 - n)ajk -8iZ a«l • i
Proposition 6.3. (n > 2), and let
Let Mn and M*n be two Riemannian
X: Mn -^Nn+1 ,
n-manifolds
X*: M*n-> Nn+1
be two convex immersions in a Riemannian (n + l)-space Nn+1 of constant sectional curvature c. Suppose that there exists a diffeomorphism h: Mn -^M*n preserving the second fundamental form and the Ricci tensor with respect to the first fundamental form. Then h is an isometry of Mn onto M*n Proof. By (6.57) and (6.57)*, the hypothesis Rjk = R% implies (6.61)
nH8Jk + gi" = nH*8jk + g*» .
Taking the trace of both sides of the matrix equation (6.61) and making use of (6.56), (6.56)* we thus obtain that H = H* which reduces (6.61) to gi" = g*'k. Hence / = /*. Proposition 6.4. Let Mn and M*n be two Riemannian n-manifolds (n > 3), and let
161
COMPACT HYPERSURFACES
be two convex immersions in a Riemannian (n + l)-space Nn+1 of constant sectional curvature c. Suppose that there exists a difjeomorphism h:
Mn->M*n
preserving the second fundamental form and the quadratic form Bijco'-co1. Then h is an isometry of Mn onto M*n. Proof. Since / / = //*, Tjk = Tjk*. By (6.60) and (6.60)*, the hypothesis Bjk — B% implies (6.62)
nH8jk - (n - 2)g» = nH*5jk - (n - 2)g**« .
Taking the trace of both sides of the matrix equation (6.62) and making use of (6.56), (6.56)* we thus obtain that H = H*, which and n > 3 reduce (6.62) to gJk = g*Jk, and hence the proposition is proved. For Euclidean space Nn+1 Propositions 6.3 and 6.4 are due to Gardner [2]. References t 1 ] S. Cohn-Vossen, Zwei Sdtze iiber die Starrheit der Eifldchen, Nachr. Ges. Wiss. Gottingen (1927) 125-134. [ 2 ] R. B. Gardner, Subscalar pairs of metrics and hypersurfaces with a nondegenerate second fundamental form, J. Differential Geometry 6 (1972). [ 3 ] V. G. Grove, On closed convex surfaces, Proc. Amer. Math. Soc. 8 (1957) 777-786. [ 4 ] G. Herglotz, Vber die Starrheit der Eifldchen, Abh. Math. Sem. Univ. Hamburg 15 (1943) 127-129. [ 5 ] S. Kobayashi & K. Nomizu, Foundations of differential geometry, Vol. I, Interscience, New York, 1963. [ 6 ] H. F. Munzner, Eine Bemerkung zum Kongruenzsatz von Grove, Arch. Math. 18 (1967) 525-528. LEHIGH UNIVERSITY BETHLEHEM, PA. 18015
U.S.A.
412
COLLOQUIUM XXVI
ADMITTING
MATHEMATICUM
DfiDlfi A M. WLADYSLAW SLEBODZltfSKI
RIEMANNIAN MANIFOLDS CERTAIN CONFORMAL CHANGES
1972
OF
METRIC
BY
CHUAN-CHIH HSITTNG- (BETHLEHEM, PA.) AND L A E R Y E. M U G E I D G E (EASTON, PA.)
1. Introduction. Let M be an n-dimensional differentiable connected Eiemannian manifold with metric tensor g. In order to distinguish g from other metrics on M, we denote the manifold M with g by (M, g). If g* is another metric on M, and there is a function a on M such that g* = e2ag, then g and g* are said to be conformally related or conformal to each other, and such a change of metric g -> g* is called a conformal change of Eiemannian metric. Let (M, g) and (M', g') be two Eiemannian manifolds, a n d / : M -> M' a diffeomorphism. Then g* = / " V is a Eiemannian metric on M. When g* and <jr are conformally related, that is, when there exists a function a on M such that g* = e2o#, we call / : (M, g) -> (Jf', #') a conformal transformation. In particular, if a is constant, then / is called a homothetic transformation or a homothety; if cr = 0, then / is called an isometric transformation or an isometry. The group of all conformal (homothetic or isometric) transformations of M onto itself is called a conformal transformation (a homothetic transformation or an isometry) group and is denoted by C(M) (E(M) or I(M)). The connected components of the identity of C(M), H(M) and I(M) are denoted by C0(M), HQ(M) and I0(M), respectively. If a vector field v on M defines an infinitesimal conformal transformation on (M, g), then v satisfies Lvg = 2og, where Lv denotes the Lie derivative with respect to v, and Q is a function on M. The vector field v defines an infinitesimal homothetic transformation or an infinitesimal isometry according as Q is constant or zero. In the last decade or so various authors have studied the conditions for a Eiemannian manifold M of dimension n > 2 with constant scalar curvature R and admitting either an infinitesimal non-isometric or a non-homothetic conformal transformation to be either conformal or isometric
413 136
C.-C. H S I U N 6 AND L. R. M U G E I D G B
to an w-sphere. Very recently Yano, Obata and Sawaki [7]- [9] and Hsiung and Mugridge [4] have been able to extend some of these results by replacing the constancy of R by LVR = 0, where v is a certain vector field on M. The purpose of this paper is to extend further the joint work of Yano and Sawaki [9] and that of Hsiung and Mugridge [4]. Throughout this paper all Latin indices take the values X, ..., n unless stated otherwise, and we shall follow the usual tensor convention that indices can be lowered and raised by using, respectively, gi3- and g%i, the elements of the inverse of the matrix (g^), and that repeated indices imply summation. In the proofs of the theorems in Section 3 we need the following known theorems: THEOREM A (Ishihara and Tashiro [5], Tashiro [6]). If a complete Riemannian manifold M of dimension n ^ 2 admits a non-constant function g such that ViVjQ+—AQgij n
= 0,
then M is conformal to an n-sphere. THEOREM B (Yano and Obata [8]). A complete Riemannian manifold M of dimension n > 2 is isometric to an n-sphere, if it admits a non-constant scalar field u such that (1.1) LduR = 0 and (1.2)
Vtu1 +
—Augij=0
hold, where Ldu is the Lie derivative with respect to the vector field ul defined by (1.3)
Uj = VjU,
ul =
gijUj,
V being the covariant derivative with respect to g. 2. Notation and formulas. Let {M,g) be a Eiemannian manifold of dimension n > 2 and class C3. In this section we shall list some well-known formulas for our later use; for the details of their derivation see, for example, [1], p. 89-90, where the Eiemann tensor differs from ours in sign. Suppose that (M, g) admits a conformal change of metric
(2.1)
4' = e2"9n,
where a is a C3 function on M. Then (2.2)
g*ij = e~2°giS.
For any tensor with respect to g, the corresponding tensor with respect to g* will be denoted by the same letter with a star.
414
CONFORMAL CHANGES OF METRIC
137
If w e denote b y Ehiik a n d B^ t h e E i e m a n n tensor a n d t h e Eicci tensor of (M, g), respectively, we have (2.3)
e
Bjiijk = EMjk — ghkaij — gijahk-\-ghj0ik
+ gikahj —
where we have p u t (2.4)
cry = V ^ a - V ^ V f f f ,
Axa = V*
B y means of equations (2.2) a n d (2.3), we can h a v e (2.5)
B% = Bij + (2-n)aij
+ gij[Aa +
B* = e-2a[B + 2(n-l)A
(2.6)
+
(2^n)A1a], {n-l)(2-n)A1o],
where (2.7)
Aa =
-V%o.
On t h e manifold M consider t h e tensor fields (2.8)
Tij = Btj
(2.9) (2.10)
Thijk = Bhijk —
gfj,
B # _ , v (9hk9ij — 9hj 9ik) t
Whijk = aThiik + fti^Ty - b2ghiTik + 6 3 ^ ^ - h9ikThj + h9hiTik -
b6gikThi,
where t h e a a n d ft's a r e constant. (2.10) is t h e covariant tensor field of order 4 defined b y Hsiung [3]. I t should b e noted t h a t T„ = ghkThijk,
(2.11)
g{% = 0,
(2.12)
gMTMik = 0 ,
(2.13)
VZ^-^-Vy-R,
(2.12) a n d (2.13) being consequences of t h e first a n d second Bianchi identities, respectively. Substituting (2.3), (2.5) a n d (2.6) in t h e equation corresponding t o (2.9) for Tlijk, we obtain, after a n elementary simplification, (2.14)
e-2aTlijk
= Ihm - ghk aH - gtj ahk + ghj aik + 2 +9ik<*h3
(9hk9ij-9hj9ik)(^
+ ^i
415
138
C.-C. H S I U N G AND L. R. M U G E I D G K
Multiplication of b o t h sides of (2.14) b y g*hk a n d use of (2.8), (2.2) a n d (2.4) give immediately (2.15)
T*j^Tij + (n-2)Pij,
where (2.16)
Pii=
-aij-
— n
(Aa+A1a)gij.
B y substituting (2.14) a n d (2.15) in t h e equation corresponding t o (2.10) for Wlijk, we obtain e-2°W*hm
(2.17) =
W
Mik + a \-ghk ati -gi} ahk +ghj aik +gik ahj - — (ghkgi:i -ghjg{k){ Aa + A1a)\ +
+ (n - 2) Ihg^Pv
- b2ghjPik + b^P^-b^g^P^
+ b5ghiPjk -
b6gjkPM].
On t h e other hand, if we set (2.18) u = e~a, it follows t h a t (2.19) a = —lau. Substituting (2.19) in (2.4) a n d (2.7) yields immediately (2.20)
ai} = -(V^lu,
Axa = (A^/u2,
-(A^/u-iA^/u2.
Aa =
Therefore (2.6), (2.14), (2.16) a n d (2.17) become, in consequence of (2.20) and (2.4), (2.21)
R* = u*R-2{n-l)uAu-n(n—l)uiui,
(2.22)
u-T%ijk = TMjk + U"1 [ghk ViUj + g{j Vhuk - ghj Wfuk 2 - 9ik vh^i H
=
{9hk9n ~ 9hj9ik) Ait] n Thijk + ghkPij + gijPhk — ghjPik ~~ 9ikPhi 1
(2-23)
Pii==u-^iUj+^Lg^
u2W*Mjk = WMjk +
(2-24)
1 -tMjk, u
where we have placed
+
\nZ^ + bs) gvP™~ t l ^ + b*J 9<*PM +
bsghiPjk-b6gjkPt
416 CONFORMAL CHANGES OF METRIC
139
From (2.23) and the second equation of (2.11), it follows immediately (2.26)
TqF*
^u^TyVWu.
We also have, from (2.15), (2.22) and (2.24), respectively, (2.27)
T*jT*{i = tt*[T0Tw + 2 ( » - 2 ) T < , P * + (»-2) , P<,.P < '],
(2.28)
T*hiikT*hijk = u* {TMjkThiik + 8T«,F« + 4{n - 2)P<,P«],
(2.29)
W*hiikW*Mik = u* lwhijk WMjk + - thijkWhijk + ^ hm thm) •
Substituting (2.10) and (2.25) in (2.29) and making use of (2.12) and g^Pij = 0, an elementary but lengthy calculation gives (2.30)
W*hmW*hiik = uiiWhmWMjk
+ 2c(n-2)TijPij
+
c(n~2fPijP%
where i
6
(n-1)
yb2i-2{b1b3
+
b2bi-b5b6).
i=\
An elementary calculation shows that c > 0 , where the equality holds if and only if 6X = ... = 64, b5 = b6 = 0 and a = (2—n)b1. 3. Lemmas. LEMMA 3.1. Suppose that a compact orientable Biemannian manifold M admits a conformal change of metric (2.1). Then, for an arbitrary number p,
fup-%juiujdV+
(3.1)
(V+^-PW
M
M
[up-1(u-1LduSr-uLauB)dV-
= - ~ 2n J M
-(n+p-2)\
fup-2uiujV{uidV I J L
M
+
f 2n(n—l) J x
(Bup~l-B*up'3)uiuidV-
M
-— f' u'-'iuyfdV
417 140
C.-C. H S I U N G AND L. R. M U G R I D G E
In particular, if p = 2 — n, then fu~n+1TijuiujdV
(3.2)
u-n+iPiiFlidV
+ f
M
M
§u-n+1(u-lLduR*-uLduR)dV.
= ~Y~ M
Proof. From (2.23) it follows that (3.3)
= VfttjVV-^tt)1/"-
eP^P"
Multiplying (3.3) by up~l, integrating over M, using the equations obtained by directly computing Vi(up~1ujViuj) and Vi{up~1uiAu), applying the well-known Green's formula
(3.4)
= 0,
JV^7 M
£i being any vector field on M, and substituting « ' V1' V« uf = R^u*
(3.5)
- u* V« Au,
in the resulting equation we obtain fup+1PijP*'dV
(3.6)
= ~{p-l)
M
n—1 c pp11 •i l • (uu--uuVViAudV iAudV nn JJ
(up-1Bi/utuidV--
-
(u'-'vfu'ViUjdVM
J
M
V—l + ^-^ [ nn J
M
up-1uiu*dV.
M
On the other hand, solving (2.21) for Au, we have (3.7)
Au = — £ \jfh
—•{uR-u-1R*-n(n —
l)u-1uiui).
_L 1
Substituting (3.7) in (3.6), and using (2.8) and LauB=uiVtB,
(3.8)
LduR* =
u%R*,
an elementary computation leads readily to the required formula (3.1). LEMMA 3.2. Suppose that a compact orientable Riemannian manifold M admits a conformal change of metric (2.1). Then
(3.9)
f(u-n*-ul)dV M
c(n-2)2
= -°-~~n
JLduRdV + M
+ c(n-2)2
CuPijPtldV, M
418
CONFORMAL
{u-n+3X-u-n-lX*)dV
CHANGES
OF METRIC
141
j «- B L < l u .B*dF +
= -
M
M
+ c(n-2f
Cir^PijPVdV,
where X = WhijkWMik,
(3.11)
X* =
T ^ W * *
<m c *s defined by (2.31). P r o o f . F r o m (2.30) a n d (2.26) it follows t h a t u~3X*-uX
(3.12)
= c(n-2)(2TijViu}
+
{n-2)uPijPii).
Integrating (3.12) over M, using V*
(3.13)
=u^iTii-\-Ti^iui,
a n d (2.13), (3.8), a n d applying Green's formula (3.4), we can easily obtain (3.9). Similarly, we can derive (3.10) b y multiplying (3.12) b y u n+2, integrating t h e resulting equation over M, using (3.13), (2.13), (3.8), (3.2), and applying Green's formula (3.4). 4 . Theorem. T H E O R E M 4.1. If a compact Biemannian manifold M of dimension n > 3 admits a conformal change of metric (2.1) with a < 4 such that LduB*>unLduB,
(4.1)
uaX = ((u-l)(p+l)X*,
(4.2)
c>0,
hold, where u = e~a,
LduB<0.
Finally, M is isometric to an n-sphere under condition tion to conditions (4.1) and (4.2). P r o o f . A t first we notice (4.4)
J(uX-ir3X*)dVM
(1.1), in addi-
j(u-n+zX-u-n~l??)dV M
=
f(un-2~-l)(u-n+3X~u-n-1X*)dV.
419 142
C.-C. H S I U N G A N D L. R. M U G R I D G E
Furthermore, j'(un~2-l)(u-n+3X-%-n~lX*)dV-
(4.5)
Ju-n^{un-2-l)(u^a-l)X*dV
M
M
n+3 a
= fu~
n 2
Ju-n+3-a{un-2-l)(u-l)
- {u - -l){u°?i-A*)dV=
M
M
the last two steps being, respectively, due to (4.2) and u>0,
(un-2-l)(u-l)^0,
Since a ^ 4, we have (u11-2-1) (u*'a-1) imply §(uX-%-3X*)dV-
(4.6)
M
2*^0.
> 0 so that (4.4) and (4.5)
f(u-n+3A-u~n-1X*)dV^0. M
On the other hand, using (3.9), (3.10) and (4.1), we obtain f\ul-u-3X*)dV-
(4.7) M
=
n
~2)
({u-n+3X-u~n~ll*)dV M
f {LduR-u-nLduR*)dV-c(n-2)2 M
f (u + u-n+3)PijPiidV
<0.
M
Thus, from c > 0 and assumption (4.1), it is readily seen that the equality holds in both (4.6) and (4.7), and, therefore, that (4.8)
{(u + u-^PqPVdV = 0, M
which together with (2.23) gives (1.2). Hence, by Theorem A, M is conformal to an w-sphere. When a = 4 and cp = 0, the left-hand side of (3.9) is reduced to zero and the right-hand side together with condition (4.3) gives immediately (1.2). Finally, if (1.1) holds, then M is isometric to an n-sphere by the same argument as above and Theorem B, q.e.d. Theorem 4.1 is due to Yano and Sawaki [9] when (4.9) or (4.10)
a = 0, 61
=
b2 = ... = b6 = 0
...=ft4,
b5=b6=0,
and condition (4.1) is replaced by LduR = LduR* = 0. Theorem 4.1 is also due to Hsiung and Mugridge [4] when cp = 0, a = 4 and condition (4.1) is replaced by / RAudV = 0, and due to Yano and Obata [8] M
with an additional condition (4.9) or (4.10).
420
OONFORMAL CHANGES OF METRIC
143
BEFEBENCES [1] [2] [3] [4] [5] [6] [7] [8] [9]
L. P . E i s e n h a r t , Biemannian geometry, 2nd ed., Princeton University Press, Princeton 1949. S . I . G o l d b e r g and K. Y a n o , Manifolds admitting a non-homothetic eonformal transformation, Duke Mathematical Journal 37 (1970), p . 655-670. C. C. H s i u n g , On the group of eonformal transformations of a compact Biemannian manifold. Ill, Journal of Differential Geometry 2 (1968), p . 185-190. — and L. R. M u g r i d g e , Conformal changes of metrics on a Biemannian manifold, Mathematische Zeitschrift 119 (1971), p . 179-187. S. I s h i h a r a and Y. T a s l i i r o , On Biemannian manifolds admitting a concircular transformation, Mathematical Journal of Okayama University 9 (1959), p . 19-47. Y. T a s h i r o , Complete Biemannian manifolds and some vector fields, Transactions of the American Mathematical Society 117 (1965), p . 251-275. K. Y a n o , On Biemannian manifolds admitting an infinitesimal conformal transformation, Mathematische Zeitschrift 113 (1970), p . 205-214. — and M. Ob a t a, Conformal changes of Biemannian metrics, Journal of Differential Geometry 4 (1970), p . 53-72. K. Y a n o and S. S a w a k i , Notes on conformal changes of Biemannian metrics, Kodai Mathematical Seminar Reports 22 (1970), p . 480-500.
LEHIGH UNIVERSITY LAFAYETTE COLLEGE
Beeu par la Bedaction le 10. 12. 1971
J . DIFFERENTIAL GEOMETRY 6 (1972) 595-598
THE SIGNATURE AND G-SIGNATURE OF MANIFOLDS WITH BOUNDARY CHUAN-CHIH HSIUNG
1.
The signature theorem
Let M be a compact oriented C°° manifold1 of dimension 4k with boundary B of dimension 4k — 1. The oriented B is called a reflecting boundary of M if it admits an orientation-reversing involution it. A simple example of the reflecting boundaries of M is a (4k — l)-sphere. For convenience and simplicity, we shall always denote such as (B, it) a reflecting boundary with its involution together but with its dimension omitted. The following problem seems to be of interest: Is there any manifold M with a reflecting boundary (B, it) on which the involution it cannot be extended to the interior of the manifold M ? Now let M with boundary B be a C°° homeomorphic copy of (M, B) with the same orientation, and (x be the homeomorphism so that n(M, B) = (M, B). Then we can define the double of M with a reflecting boundary (B, it) to be a C°° closed oriented manifold N such that N = M U M and that B = it(B) by identifying (nt(x) e B with x for all x e B. Thus on the double N we can define a homeomorphism v: N —* N by: (1.1)
(/!(*) , v(x) — i [fi \x) ,
for*eM , for x € M .
To see that this is well-defined, at first we notice that v(x) = fi(x) e B for x e B. Since * is identified with (iit(x) e B, v(x) = v([nt(x)) = (i~\nn(x)) = it(x) e B, and therefore fi(x) is identified with it(x); this is indeed true by the definition of our identification and the assumption it2 = 1. Clearly, v is an involution. (It should be noted that the definition of doubling a manifold M here is somewhat different from the ordinary one under which M and M are of opposite orientations so that every point of B is a fixed point under the involution v; for the latter see, for instance, [3].) Alternatively, we may regard doubling the manifold M with a reflecting boundary (B, it) as finding a C°° homeomorphism
h:M->N Communicated February 3, 1972, and, in revised form, June 25, 1972. partially supported by the National Science Foundation grant GP-33944. 1 Throughout this paper all manifolds are differentiable.
Research
596
CHUAN-CHIH
HSIUNG
where N is a C*° closed manifold with an involution v:N-*N such that vh = hit: B —* N'. Assume h maps M into a fundamental domain of the involution v in such a way that B is mapped onto itself. We shall identify M with the fundamental domain henceforth, so that we may regard the double N as composed of two halves M and M with the same orientation such that M is mapped onto M by v, M f| M = B, and v\B = it. A Riemannian metric on the double N, for which the involution v is an isometry, is said to be symmetric, and the restriction of a symmetric metric on JV to M is also called a symmetric metric on M. Now there arises the problem of deriving a C° symmetric Riemannian metric on N from a C°° Riemannian metric on M. For this problem at first we are naturally tempted to prolong to N a differentiable metric g on some manifold containing M by setting (1.2)
g(x) = g(v(x))
for x s M. Although (1.2) is well defined, the difficulty is that the resulting metric will, in general, not be differentiable across B. However, on the other hand, for a given C°° Riemannian metric g everywhere defined on N we may obtain from it a C" symmetric metric g by setting (1.3)
g(x) = i[g(x) + g(v(x))] ,
for x e N .
Now we consider a C" symmetric Riemannian metric g o n J V . Then by Hirzebruch signature theorem [5] the signature of N is given by
(1.4)
sign (N) = JLk(Pl, • • •, Pk){Qg) , N
where Pj is the /'-th Pontrjagin class of N, and {L t (p 1; • • •, pk)(Qg)} is the Hirzebruch's multiplicative sequence of polynomials with each pt expressed in terms of the curvature 2-forms Qjk of the Riemannian metric g by a theorem of Chern [2]. However, J Lk(py, •••, Pk)(Qs) = ( J + J ) L*(Pl> • • • » Pk)(Qg)
(1.5)
"
u
k
= 2 J Lk(pu
••-,pk)(Qg)
,
M
since Lk(pu • • -,pk)(Qg) depends only on the Riemmanian metric g, and the metric g is symmetric. On the other hand, the following theorem was first observed by S. P. Novikov and proved jointly by Atiyah and Singer [1, Prop. (7.1)]:
597
SIGNATURE AND G-SIGNATURE
Theorem 1.1. Suppose that two compact oriented manifolds Mr and M2 have a common boundary B with opposite orientations. Then (1.6)
sign (Af, U M2) = sign (M„ B) + sign (M2, B) ,
where sign (Mt,B) denotes the signature of the manifold Mt with boundary B for i = 1,2. By applying Theorem 1.1 to our case we obtain (1.7)
sign (N) = sign (M, B) + sign (M, B) = 2 sign (M, B) ,
since M and M are homeomorphic with the same orientation. Combination of (1.4), (1.5), (1.7) thus gives
(1.8)
sign (M,B) = JLk(Pl, • • •, P*)(0.) •
Hence we arrive at Theorem 1.2. 77ze signature of a compact oriented C°° manifold M of dimension Ak with a reflecting boundary B of dimension Ak — 1 is given by (1.8) where g is a symmetric metric on M. When B is empty, Theorem 1.2 reduces to Hirzebruch signature theorem. 2.
The G-signature and signature-defect
We first state the following theorem of Atiyah and Singer [1, p. 588], which is a generalization of Theorem 1.1 on the additivity property of the signature: Theorem 2.1. Suppose that two compact oriented manifolds Ml and M2 have a common boundary B with opposite orientations, and that a compact Lie group acts differentiably on Mx and M2 preserving the orientations. Then (2.1)
sign (G; Af! U M2) = sign (G;M1,B)
+ sign (G; M2, B) ,
where sign(G; Mj U M2) and sign(G; Mt,B) denote, respectively, the Gsignatures of Mx U M2 and Mi with boundary B for i = 1,2. As in § 1 let M be a compact oriented manifold of dimension 21 with a reflecting boundary (B,n), and N = M U M be the double of M with an involution v defined by (1.1). Then we naturally intend to extend any automorphism g of M to an automorphism of N by (2.2)
g(x) ,
for x e M ,
l
for x e M ,
v{gv~ (x)) ,
so that (2.3)
g{vn{x)) = vgv~\vTt{x)) — vgx(x)
for x s B .
On the other hand, for x e B since x = vx(x) by our identification for the reflecting boundary (B,K), we have
598
(2.4)
CHUAN-CHIH HSIUNG
g(x) = vKg(x) ,
which, together with (2.3), implies immediately that to well define an automorphism of N by (2.2) it is necessary that (2.5)
ng = gn
on B .
Now let M be a compact oriented manifold of dimension 2/ without boundary, and suppose that there is a compact Lie group G acting on M preserving the orientation. For expressing sign (g; M) for an element g of G, in [1] Atiyah and Singer obtained the G-signature Theorem (6.12), Corollary (6.13), Proposition (6.15), Corollary (6.16) (which was proved by Conner and Floyd [4, Cor. (27.4)] by a different method), and Proposition (6.18). If M has a reflecting boundary (B, n), by using our method of doubling a manifold in § 1 and Theorem 2.1, we can easily show that the above mentioned expressions of Atiyah and Singer also hold for sign (g; M, B) provided that on the boundary B, g has no fixed point and satisfies the condition (2.5). Very recently Hirzebruch [6] defined the signature-defect of a finite group acting effectively on a connected compact oriented manifold M without boundary, and obtained some interesting relationships between number theory and the signature-defect at some special points of a four-dimensional M. Now let M be a connected compact oriented manifold of dimension 4 with a reflecting boundary (B,7t), and suppose that there is a finite group G acting by orientation-preserving diffeomorphisms effectively on M and freely on B such that TCG = Gn on B. By following Hirzebruch we can easily generalize his definition of the signature-defect of a group action on a manifold without boundary to the G-action on M, and show that his relationships [6, § 5] between number theory and the signature-defect also hold for the signature-defect of the G-action on the manifold M with boundary {B, n) by using the conditions of G on (B, n) and the extention in § 1 of Proposition (6.18) of Atiyah and Singer [1]. References [ 1 ] M. F. Atiyah & I. M. Singer, The index of elliptic operators. Ill, Ann. of Math. 87 (1968) 546-604. [ 2 ] S. S. Chern, On curvature and characteristic classes of a Riemannian manifold, Abh. Math. Sem. Univ. Hamburg 20 (1956) 117-126. [ 3 ] P. E. Conner, The Neumann's problem for differential forms on Riemannian manifolds, Mem. Amer. Math. Soc. No. 20, 1956. [ 4 ] P. E. Conner & E. E. Floyd, Differentiable periodic maps, Springer, Berlin, 1964. [ 5 ] F. Hirzebruch, Topological methods in algebraic geometry, 3rd ed., Springer, Berlin, 1966. [ 6] , The signature theorem: Reminiscences and recreation, Prospects in Math., Annals of Math. Studies, No. 70, Princeton University Press, Princeton, 1971, 3-31. LEHIGH UNIVERSITY UNIVERSITY OF WARWICK
425 A C K L E S , L.
L.
- HSIUNG, CHUAN-CHIH
1974 Annali di Matematica pura ed applicata (IV), Vol. IC, pp. 53-64
Isometry of Riemannian Manifolds to Spheres. LYNN
L. ACKLEE (Fitehburg, Mass.) - CHUAN-CHIH
HSIUNG
(Bethlehem, Penn.) (*) (**)
Summary. - Some known conditions for a compact Riemannian n-manifold M", n > 2, which has constant scalar curvature M and admits an infinitesimal nonisometric conformal transformation, to be isometric to an n-sphere are generalized to manifolds with nonconstant B.
1. - Introduction. Let Mn be a Biemannian manifold of dimension n > 2 and class G3, (gif) the symmetric matrix of the positive definite metric on M", and (gij) the inverse matrix of (ga), and denote by V,, Bhijk, Bij = Bkijk and B = giiBii the operator of covariant differentiation with respect to g^, the Eiemann tensor, the Eicci tensor and the scalar curvature of M" respectively. Let d be the operator of exterior derivation, & the operator of coderivation, and A = dd + dd the Laplace-Beltrami operator. Throughout this paper all Latin indices take the values 1,..., n unless stated otherwise. We shall follow the usual tensor convention that indices can be raised and lowered by using g*> and gtj respectively, and that repeated indices imply summation. Further let (1.1) (i-2)
Tti =
Rtl-^Rgilt
T*uk = - f t w —™7^—T\ R (9ij9hk—giicgH)
•
7h\7b — JL J
It is easily seen that (1-3)
9"Ttl=0,
ft*TMth=Ttl.
The tensor Ttj measures the deviation of Mn from being an Einstein space, and Thm from being a space of constant sectional curvature. Let v be a vector field defining an infinitesimal conformal transformation on M". Denote by the same symbol v the 1-form corresponding to the vector field v by the duality defined by the metric on M", and by Lv the operator of the infinitesimal
(*) E n t r a t a in Eedazione il 20 giugno 1972. (**) The work of t h e second author was partially supported by NSP grant GP-11965.
426 54
LYNN L. ACKLER - CHTTAN-CHIH HSIUNG: Isometry of Biemannian, etc.
transformation v. Then we have (1.4)
Lvgit = VtVj + VjVt = 2Qg„ .
The infinitesimal transformation v is said to be nomothetic or an infinitesimal isometry according as the scalar function Q is constant or zero. On a compact orientable Riemannian manifold, an infinitesimal homothetic transformation is necessarily an isometry; see [14]. Further we denote by Lde the operator of the infinitesimal transformation generated by the vector field Q' defined by (1.5)
Q'^g^Qi,
Qi=V,Q.
Let f/(, and t]I(p) be two tensor fields of the same order p < n on a compact orientable manifold Mn, where I(p) denotes an ordered subset {iu ..., iv) of the set {1,..., n} of positive integers less than or equal to n. Then the local and global scalar products
(1.7)
^^J"p)VnvU
(S,ri)=jdr,
where dV is the element of volume of the manifold Mn at a point. We also define (i.8)
u\\=p\(§jy.
From (1.6) and (1.7) it follows that (f, f) is nonnegative, and that (f, f) = 0 implies that | = 0 on the whole manifold Mn. In the last decade or so various authors have studied the conditions for a Eiemannian manifold Mn of dimension n > 2 with constant scalar curvature B to be either conformal or isometric to an w-sphere; for these see specially [8], [13]. Very recently YANO, OBATA, HSIUNG, MUGRIDGE and STERN (see [18], [16], [6], [7]) have been able to extend some of these results by replacing the constancy of B by one or two conditions such as LUB = 0, where u is a certain vector field on M". The purpose of this paper is to continue their work, in particular YANO'S [16] and HSIUNG and STERN'S [7], by establishing the following theorems. To begin we denote by (C) the following condition: (C)
(1.9)
A compact Eiemannian manifold Mn of dimension n > 2 admits an infinitesimal nonisometric conformal transformation v satisfying (1.4) with Q =£ 0 such that
LvB=0,
427 L Y N N L. A C K L E R - C H U A N - C H I H H S I U N G : Isometry
of Biemannian,
etc.
55
and (1.10)
LnB = 0.
THEOREM: 1. - If an orientable M" satisfies condition (C) and any one of the following equivalent conditions: (1.11)
Ldegij = 2(pgij
(1.12)
(Tii,QiQi)>0,
(LIS)
(q> is a scalar
function),
(BMe«e,__L_e.B.li)>0,
then Mn is isometric to an n-sphere. (1.11) is equivalent t o t h e condition t h a t do define a n infinitesimal conformal transformation on M". For constant B, Theorem 1 with condition (1.12) is due t o Y A N O [15] a n d OBATA [11].
Moreover, YANO [16] proved t h a t a n Mn satisfying
condition (1.13) a n d condition (C) without (1.10) is conformal t o an w-sphere. COROLLARY 1. - An orientable M" is isometric to an n-sphere it if satisfies dition (C) and any one of the following three sets of conditions: (1.14)
V,V,(-B/) = BoBn
(1.15)
Qdo = - d(Bo),
(1.16)
LvBtj
= ajr.j
(/ is a scalar
function),
(a is a scalar
function),
con-
where Q is the operator of Bicci defined by, for any vector field u on Mn, Q:ui-+2R„ui.
(1.17)
For constant B, Corollary 1 is due t o LICHNEROWICZ [10] for condition (1.14) due t o YANO a n d OBATA [17] a n d B I S H O P a n d GOLDBERG [1] for condition (1.15),
and due to Y A N O a n d OBATA [17] for condition (1.16). Moreover, H S I U N G and STERN [7]
proved t h a t an Mn in Corollary 1 without satisfying t h e condition (1.10) is conformal t o an w-sphere. THEOREM 2. - M" is isometric to an n-sphere if it satisfies
(1.18)
L^a'A+^^BJ=0,
where (1.19)
A == Bhi'"Bhiik,
B=
BOB,,,
condition
(C) and
428 56
L Y N N L. A C K L E R - C H U A N - C H I H H S I U N G : Isometry
of Biemannian,
etc.
and a and c are constants such that (1.20)
c = 4a 2 + ( » - 2 ) [ 2 a i > < + ( 2 ( - I ) ' - 1 * . ) 2 L
1=1
» 1=1
'
-2(M. +&A-&A) + (»-i)inl >o, 6's fteiwgf am/ constants ('). Theorem 2 is d u e t o H S I U N G [3] for constant B a n d either a = 0, c — 4a 2 # 0, B = const or a i= 0, c —4« 2 = 0, J . = const., due t o YANO [15] for constant B a n d either a = 0, c — 4a 2 == 0 or a == 0, c — 4a 2 = 0, due t o H S I U N G [4] for constant B a n d 62 = ... = be = 0, due t o Y A N O a n d SAWAKI [19] for constant B and bx = ... = bt = = 6/(TO — 2 ) , 65 = be = 0, due t o
H S I U N G [5] for constant B,
a n d due t o
HSIUNG
and STERN [7] for either a = 0, c — 4a 2 ^ 0, 5 = const, or a == 0, c — 4a 2 = 0, A = const. Moreover, YANO [16] proved t h a t an M" satisfying all conditions in Theorem 2 for either a = 0, c — 4a 2 ^ 0 or a == 0, c — 4a 2 = 0 is conformal t o an n-sphere if (1.10) is replaced b y a certain integral inequality. I n t h e proofs of t h e above theorems we need t h e following theorem. THEOREM: A (OBATA [11]). - If a complete Biemannian TO>2 admits a noneonstant function Q such that
(1.21)
manifold Mn of dimension
v 1 v i e = -o«effHj
where c is a positive constant, then M" is isometric to an n-sphere of radius 1/c.
2. - Notation and formulas. I n this section we shall list some known formulas (for details of their derivations see Lichnerowicz' book [9, pp. 124, 134] or Hsiung's paper [2]) which will be needed in t h e proofs t o follow. Let v be a vector field defining an infinitesimal conformal transformation on a Biemannian manifold M" of dimensionTO> 2 so t h a t (1.4) holds. Then we have (2.1)
Q= i
~dvjn, i
(2.2)
Lvr jk=d iQk
(2.3)
LvB\jk
+
d'kQi~glkoi,
= V * ( £ . r \ , ) - V,(E./\»).
Substitution of (2.2) in (2.3) gives immediately
(2.4)
LXW = - K^iQi + ajv* e< - ff „v»g* +
ff„vie*,
(*) An elementary calculation shows that c > 0 where equality holds if and only if & ! = . . . = 64, bs=b6=0, a= -(n-2)b1.
429
L Y N N L. ACKXER - C H U A N - C H I H H S I U N G : Isometry
of Riemannian,
etc.
57
where <5£ = 1 for h = k a n d = 0 for h # k. From (2.1) and (2.4) it follows immediately t h a t (2.5)
LvRhiik
= 2oRhiik — ghkViQi + ghiViQk — gtjVhQi +
(2.6)
LvRii
(2.7)
£ t jB = 2 ( w - l ) A e — 2 E e .
g^hQ,,
= gijLQ — (n — 2)V,Qit
On M" for a n y scalar field / and a n y vector field u we have —ViVif,
(2.8)
Af =
(2.9)
(Au)t = — d ' V , « , + i(g«)«,
where Q is t h e operator of Eicci defined b y (1.17). A necessary and sufficient condition for a vector field v t o define a n infinitesimal conformal transformation on a compact manifold Mn is t h a t it satisfy (2.10)
Av + (1 — 2jn)ddv = Qv.
3. — Lemmas. Throughout this section M" will always denote a compact orientable Eiemannian manifold of dimension w > 2 . LEMMA 3.1. - If a nonoonstant scalar field y> on an M" satisfies Ayi = ky>, where k is constant, then k is positive. PROOF. - From (2.8) it follows t h a t A(ip2) = — V'V^v 2 ) = —2(yV'V < v' + WiW*) = 2ipAy> — 2y>iy>'. By substituting V 4 (y 2 ) for £,- in t h e well-known Green's formula
(3.1)
(v'f„l) = 0 ,
where f < is a n y vector field on Mn, we therefore have (3.2)
d=(A(y>*),l)
which gives Lemma 3.1 immediately.
=
2(kV*-y>iy>i,l),
430 58
LYNN L. ACKXER - OHUAN-OHIH HSIUNG: Isometry of Riemannian, etc. LEMMA
3.2. - If f is a scalar field on Mn and Af = 0, then f is constant.
PKOOF. - Replacing v by / in (3.2) and using Af= 0 we obtain ((Aif){A*f), l) = 0, which implies that V,/ = 0 since (V,/)(V'7) is nonnegative. Hence the lemma.
LEMMA 3.3. - (YANO [16]). If v defines an infinitesimal nonisometric conformal transformation on M" such that (1.9) holds, then (3 3)
-
(B«e'e'~S(jr=ij^M) +
+ 2 (v*' + ^h) <#*" v*< + n^=T) *»»«) = ° • - By noticing
(3.4)
V< [(v,g, + i z % „ ) <>'] = [ v ' V . j , + i V,Jg] ' +
+ 2 /v4g, + i z%„, v,g# + i Jw„ Furthermore, using the Ricci identity for the vector g{ and the relation (3.5)
WiQi =
AtQl,
we obtain
or V i V <ei = i2„ ? i — V , ^ .
(3.6)
Substituting (3.6) in (3.4), integrating the resulting equation over Mn and applying Green's formula (3.1) we have (3.7)
UuQ'e'-'^Q^iAQ,
lj + 2(viQi + ~AQgit, ViQi + i Aog^ = 0 .
On the other hand, integration of A(OAQ)
= (ve)'—e'V,de—V(ev,/ig),
431 LYNN L. AOICLEB - OHUAN-CHIH HSIUNG: Isometry of Biemannian, etc.
59
over M" and use of Green's formula again give (3.8)
(Q*VIAQ,1)
=
({AQ);1).
Substituting (3.8) in (3.7) we obtain (3.9)
(RijQiei-7^Z-±(Aey,
l) + 2(vtet+±AQgtl,
V, e , + ^
W o
) = 0 .
On the other hand, (1.9) and (2.7) imply (3.10)
AQ = QBI(n — l),
by which (3.9) is hence reduced to the required formula (3.3). LEMMA
(3.11) PROOF.
3.4. - If Mn satisfies condition (0), then (Tt„ QiQl) + 2 {vi6l + ^
^
QBg„, V,g, + ^ - ^ oBg^ = 0 .
- Prom (3.10) it follows that
(3.12)
Q'V.AQ
= ^
(Q,B
+
QV,R)Q>
.
On the other hand, by a known formula for Lie derivatives we have (3.13)
LdeB =
Q'VSB
,
so that (1.10) becomes (3.14)
Q'V,B = 0 ,
by which (3.12) is reduced to (3.15)
Q>V1AQ = QiQlBI(n — l).
Substitution of (3.10) and (3.15) in (3.7) hence gives the required formula (3.11).
4. - Proofs of theorems. PROOF OF THEOREM 1. - At first, assume condition (1.11) holds. Then from the equation corresponding to (2.7) for (1.11) we obtain
(4.1)
LdQB = 2{n — l)A(p — 2
432 60
LYNN L. ACKLER - OHUAN-CHIH HSITJNG: Isometry of Biemannian, etc.
and therefore (1.10) is reduced to (4.2)
A
J.
On the other hand, by the definition of LdQ and (3.5) we have Laegtj = 2V^ 3 , and therefore, in consequence of (1.11), (4.3)
ViO,=
Multiplication of (4.3) by gtl gives (4.4)
AQ = —
ncp.
Thus from (4.4) and (3.10) we obtain
which, together with (3.14), implies (4.6)
^
=
___L_CEJe
+
gjB).
Substituting (4.2) and (4.4) in (4.6) we have (4.7)
QAB
= 0.
Since v defines an infinitesimal nonisometric oonformal transformation on a compact orientable Biemannian manifold M", Q is nonconstant and nonzero. Therefore AB = 0 and hence, by Lemma 3.2, B is constant. Substitution of (4.5) in (4.3), (4.4) gives (4.8)
^ , =
- ^ 3 ^ , ,
(4.9)
Ap = pB/{n — l).
Since Bj(n — 1) is constant and Q is nonconstant, B is positive by applying Lemma 3.1 to (4.9), and hence M" is isometric to an w-sphere of radius (i?/[w(»i —1)])* by applying Theorem A to (4.8). Next, we need to show the equivalence of conditions (1.11), (1.12), (1.13). To this end, at first from (4.8) it follows that
(4.10)
(v<e, + j
^ e % „ v, e , + n^=Y)Q^)
=o•
433 LYNN L. ACKLEE - OHTTAN-CHIH HSITJNG: Isometry of Riemannian, etc.
61
Thus, we have (Tit, g,g,) = 0 by Lemma 3.4, and
^ — r , e2-B2, l ) = 0
UtrfQ' \
n(n — 1) ^
/
by Lemma 3.3. Hence under condition (0), (1.11) implies (1.12) and (1.13). Now assume that either (1.12) or (1.13) holds. Then from Lemma 3.4 or Lemma 3.3 we have
which implies (4.8). Thus dp defines an infinitesimal conformal transformation on Mn, and hence under condition (C) either (1.12) or (1.13) implies (1.11). PROOF OF COROLLARY 1. - (i) Suppose (1.14) holds for a scalar function/. Then by means of (2.8) it is easily seen that A(Bf) = — UQB, which, together with (3.10), implies
(4.11)
A\Q+
*
JB/1 = 0 y
so that by Lemma 3.2, (4.12)
e
+
1
Bf = const. n(n — 1)
Differentiating (4.12) twice and using (1.14) we obtain (4.8). Thus dp defines an infinitesimal conformal transformation on Mn and the corollary follows from Theorem 1. (ii) Suppose (1.15) holds. From the definition of A it follows that dAp = Ado, which, together with (3.10) and (1.15), implies Adp + (1 — 2jn)dddp — Qdp~0. Thus by the necessary and sufficient condition (2.11), dp defines an infinitesimal conformal transformation on M", and the corollary follows from Theorem 1. (iii) Finally suppose (1.16) holds. Then (2.6) becomes <xgu=AQgii— (n — 2)V<e,
or
V(g, =
(Ap — a.)git.
Hence by Theorem 1, we have the corollary. PROOF OF THEOREM
2. - On the manifold Mn consider the covariant tensor
field of order 4: (4.13)
WMk = aTtuk + bighk Tti - b2ghl TiH + btg„ Thk — hga Thi + b„ghi Tjlc — b„gjk Tht,
434 62
LTNN L. AOKLUE - OHUAN-CHIH HSITJNG: Isometry of Biemannian, etc.
where Ttj and Thm are defined by (1.1) and (1.2), and a and 6's are constants satisfying (1.20). Then C
(4.14)'
,| T"F || = a M+^B-l(^ " w —2 w\w — 1 + ^ w^ — 2V /
y
,'
where c is denned by (1.20). (4.14) and (1.9) imply that (4.15)
Lv\\W\\^Lv(a*A +
°~^By
From (4.13), (1.1), (1.2), (1.4), (2.5), (2.6), (3.10) we can easily obtain (4.16)
LvWhii!c = 2aoRAiik-[a
+ (n-2)b1]ghhViQt
+ [a +
(n-2)b,]guVtQt-
— [a + (» — 2)63]sr(,V*eft + [« + (» — 2)64]ff,»V,e» — — (n — 2)b5ghiVkQt+(n QR
n(n — 1)
— 2)biglkViQh —
«r«0»[4a + (Sn-4)(6 1 + 63)] +
,eB „g»9Ml*a + (3»-4)(6, + 6 4 ) ] n(n — 1) 3W 4 3^ 4 + ~n(n-l)h"QBghig"' n(n-l) h^R9l"ghi + 26i W » B « +
— 2bioghlRik + 2b3ogiiRhk—2biQgikRh)
+
2bbqghiRlk~
— 2beogtkRki. Multiplying both sides of (4.16) by W™* and making use of (4.13), (1.1), (1.2), (2.8), (3.10), (4.14), (1.20) and R'ijk = 0, an elementary but lengthy calculation yields (4.17)
WhilkLvWhiih = 2o\\W\\— cT"Vj0i.
By substituting (4.17) in the well-known formula (4.18)
LV\\W\\
=2W"^Whm-8Q\\W\\,
we thus obtain (4.19)
QL*WI
=-V|W||—2oeI"'V,e<.
By the second Bianchi identity we have (4.20)
ViRll = iVtR,
435 L Y N N L. A O K L E B - C H U A N - C H I H H S I U N G : Isometry
of Riemannian,
etc.
63
which, together with (3.14), implies (4.21)
Q'V'BH
= 0.
A straightforward computation a n d use of (1.1), (4.21), (3.14) give (4.22)
V ' ( Z > g ' ) = TUQ*Q> + e T „ V ' e ' .
Without loss of generality we m a y assume our manifold M" t o b e orientable, since otherwise we need only t o take an orientable twofold covering space of M". Substituting (4.19) in (4.22), integrating t h e resulting equation over t h e manifold M" and using (3.1) we obtain
(4-23)
4o(2 , H , e<e ,) = (i.[|W||,e) + 4 ( | i r | | , e « ) .
On the right side of (4.23) the first integrand vanishes due t o (1.18), (4.15), and t h e second integral is nonnegative since its integrand is so. Hence t h e integral on t h e left side of (4.23) is nonnegative, a n d Theorem 2 follows from Theorem 1 immediately.
BIBLIOGEAPHY [1] E. L. BISHOP - S. I. GOLDBERG, A characterization of the Euclidean sphere, Bull. Amer. Math. Soc, 72 (1966), pp. 122-124. [2] C. C. HSIUNG, Vector fields and infinitesimal transformations on Riemannian manifolds with boundary, Bull. Soc. Math. Prance, 92 (1964), pp. 411-434. [3] C. C. HSIUNG, Ore the group of conformal transformations of a compact Riemannian manifold, Proc. Nat. Acad. Sci. U.S.A., 54 (1965), pp. 1509-1513. [4] C. C. HSIUNG, On the group of conformal transformations of a compact Riemannian manifold, I I , Duke Math. J., 34 (1967), pp. 337-341. [5] C. C. HSIUNG, On the group of conformal transformations of a compact Riemannian manifold, I I I , J. Differential Geometry, 2 (1968), pp. 185-190. [6] C. C. HSIUNG - L. E. MUGBIDGE, Conformal changes of metrics on a Riemannian manifold, Math. Z., 119 (1971), pp. 179-187. [7] C. C. HSIUNG - L. W. STEKN, Conformality and isometry of Riemannian manifolds to sphres, Bull. Amer. Math. Soc, 76 (1970), pp. 1253-1256 (announcement); Trans. Amer. Math. Soc, 161 (1972), pp. 65-73. [8] J. LELONG-FERRAND, Transformations conformes et quasiconformes des varietes riemanniennes; application a la demonstration d'une conjecture de A. Lichnerowicz, C. E. Acad. Sci. Paris, 269 (1969), pp. 583-586. [9] A. LICHNEROWICZ, Oeometrie des groupes de transformations, Dunod, Paris, 1958. [10] A. LICHNEROWICZ, Sue Irs transformation conformes d'une variete riemannienne compacte, C. E. Acad. Sci. Paris, 259 (1964), pp. 697-700. [11] M. OBATA, Certain conditions for a Riemannian manifold to be isometric with a sphere, J. Math. Soc. JapaD, 14 (1962), pp. 333-340.
436 64
L Y N N L. A C K L E R - C H U A N - C H I H HSITJNG: Isometry
of Biemannian,
etc.
[12] M. OBATA, Quelques inegalites integrates sur une variete riemannienne compacte, C. E. Acad. Soi. Paris, 264 (1967), pp. 123-125. [13] M. OBATA, The conjectures on conformal transformations of Biemannian monifolds, Bull. Amer. Math. Soc, 77 (1971), pp. 265-270 (announcement); J. Differential Geometry, 6 (1971), pp. 247-258. [14] K. YANO, On harmonic and Killing vector fields, Ann. of Math., 55 (1952), pp. 38-45. [15] K. YANO, On Biemannian manifolds with constant scalar curvature admitting a conformal transformation group, Proc. Nat. Acad. Soi. U.S.A., 55 (1966), pp. 472-476. [16] K. YANO, On Biemannian admitting an infinitesimal conformal transformation, Mat. Z., 113 (1970), pp. 205-214. [17] K. YANO - M. OBATA, Sur le group de transformations conformes d'une variete de Biemann dont le scalaire de courbure est constant, C.E. Acad. Soi. Paris, 260(1965), pp. 2698-2700. [18] K. YANO - M. OBATA, Conformal changes of Biemannian metrics, J. Differential Geometry, 4 (1970), pp. 53-72. [19] K. YANO - S. SAWAKI, Biemannian manifolds admitting a conformal transformation group, J. Differential Geometry, 2 (1968), pp. 161-184.
J . DIFFERERTIAL GEOMETRY 9 (1974) 177-193
THE TOTAL ABSOLUTE CURVATURE OF CLOSED CURVES IN RIEMANNIAN MANIFOLDS F. BRICKELL & C. C. HSIUNG
1.
Introduction
This paper is concerned with some extensions of the theorems of Fenchel [3], Milnor [5], Fary [2] on the total absolute curvature of closed curves in euclidean space. We obtain a theorem for closed curves in a complete simply connected riemannian n-manifold with nonpositive sectional curvature, and this leads to more precise results when the curvature is constant. Throughout this paper the summation convention for repeated indices is used, and all indices take the values 1, • • •, n unless stated otherwise. 2.
Some geometry of shells
Let 0 be a point on a closed C°° curve C embedded in a riemannian nmanifold M, and suppose that C lies in a normal neighborhood of 0. Then C can be expressed in terms of its arc length s as exp0 r(sK(s) ,
0 < s < L ,
where r(s) > 0, £0) is a unit vector in the tangent space T0M, and L is the total length of C. The functions r, £ are C°° functions, and we extend them by continuity to the closed interval 0 < s < L. Lemma 1. The extended functions r, £ possess right-hand and left-hand derivatives of all orders at s = 0 and s = L respectively. They have the particular values r(0) = r{L) = 0 ; (dr/ds)(0) = -(dr/ds)(L)
= 1;
C(0) = _C(L) = (dC/ds)(0) . Proof. Choose a system of normal coordinates determined by an orthonormal frame at 0. Let d(s), i = 1, • • • , « , be the components of £ with respect to this frame, and C;(s) be the values of the coordinate functions on C. Then Communicated November 12, 1972, and, in revised form, August 1, 1973.
178
F . BRICKELL & C. C. HSIUNG
Ci(s) = r(sXt(s) ,
0<s
,
where the C°° functions ct can be regarded as periodic functions of period L. Because c4(0) = 0, we can express ci = sAt where the At are C°° functions. Consequently, for s > 0,
-(5c/)*=
>
i , / = l , •••,/!.
The statements in the lemma about the value s = 0 follow from these formulas and the relations (dct/ds)(0) = AM
,
2 ((dcj/dsXO))* = 1 .
Similar arguments can be used to justify the statements about s — L. q.e.d. Let Q denote the set of points (y, s) in R2 (throughout this paper 7?* denotes the i-dimensional space of real numbers) such that 0 < y < r(s), 0 < s < L, and define /: Q —» M by /(y, s) = exp0 yC(^) . We call (Q, f) the shell on C with vertex 0. The curve s —> Q(s) which lies on the unit sphere in T0M is called the indicatrix of the shell. Denote by Q, the
(0,L)
C : y = r(j)
subset of fi where y > s > 0. Since from Lemma 1, dr/ds is continuous and (dr/ds)(0) = 1, the function r(» is strictly monotonic increasing on some interval 0 < s < a. Similarly, r(s) is strictly monotonic decreasing on some interval f} < s < L. Now for s e [a, /3], r(s) > 0 and so r(s) has a positive minimum b on [a, fl. Consider any e such that 0 < s < min{r(a), r{fi),b}. The equation r(s) = e will have just two solutions for J s [0, L], thatis, for £ sufficiently small the line y = e will meet the boundary of Q in just two points. Essentially our method is to apply the Gauss-Bonnet theorem to the induced
179
TOTAL ABSOLUTE CURVATURE
metric on (Q„ /), and then to let s —»• 0. The key result is Theorem 1. Unfortunately there are some technical difficulties due to the singularities of f. We will make use of the structure equations for a riemannian «-manifold expressed in polar coordinates. Choose an orthonormal frame at 0, and let u1, • • •, un denote the normal coordinates determined by the frame. Extend the frame to a moving frame Xu • • • ,Xn on the normal neighborhood by parallel translation along the geodesies through 0. Denote the dual moving coframe by 0l, • • • ,dn, and let dj = —6{ be the components of the connection form with respect to these frames. Define the mapping F: Rn+1 -* M by w*(F(/, a\ ••-, an)) = tai,i=l, •••,«. It is shown in [10, p. 27] that (1)
F*6l = aldt + y3* ,
F*6) =
'j
'
where the forms £*, $ do not involve dt. These forms are zero for / — 0, and satisfy the differential equations dp/St = da1 + a*ft ,
(2) (3)
dfildt = {WmoF)a*p
,
where R*jkl are the components of the curvature tensor with respect to the moving frames. For our purposes it is useful to note that ( 4)
p = t{X) o F)das ,
a'{I) ° F) = a1 ,
where the X) are determined by ff1 — X)du}. Both equalities follow from the first relation in (1). We denote the components of the vector £(s) with respect to the frame Xu • • • ,Xn by £i(s), i = 1, • • •, n. The mapping / is expressed in terms of the normal coordinates u1, • • -,un by «*(/(y,s)) = yC,i(s). Consequently, at any point (y,s),
(5)
/ 3 = c ^ , dy
dul
Ul = ds
y<^d. ds dU1
Because £ is a unit vector, it follows that the vectors f^d/dy, f*d/ds are linearly dependent iff f^d/ds = 0. Therefore / is an immersion except for points on the lines y = 0 or s = a where a is any number such that (d£/ds)(a) = 0. The latter singularities correspond to the points C(a) at which the curve C is tangent to the geodesic from 0. In order to use the structure equations we introduce the function ty: R2 —* n+1 R defined by (y,s) ->• (y,Zi(s), ••-,C„Cs)) .
180
F . BRICKELL & C. C. HSIUNG
We define C°° functions <%, mH = — coi} in R2 by
Consequently f = Fof. (6)
Tp/3* = a)tds ,
i p # = co^ds ,
and note that the relations (4) imply that (7)
co, = y{X) o f)dZj/ds ,
Zffl o /) = C, .
It follows from (5) and (7) that, at any point (y, s), (
8
)
f*z-
=
ZJXJ ,
f^—
dy
=
COJXJ .
ds
The functions to,, coij are zero for y = 0 and satisfy differential equations which are consequences of the structure equations (2) and (3). The equations are (9)
dcojdy = d£t/ds + ZjWji ,
(10)
—dwjt/dy = RjikiCKa)i •
We make one immediate deduction. It follows from (9) that (d/dy)^^ Therefore, because ^o^ is zero for y = 0, we can deduce that (11)
= 0.
Ci«»i = 0 .
The riemannian metric o n M induces a metric 2 * (f*di)2 in .R2. (1), (6), (11) can be used to show that this metric is dy2 + h2ds2
where
h
= (?-•)'
It is positive definite except at the singularities of /, which are therefore the zeros of h. The Gaussian curvature K is defined at the nonsingular points and it is easy to check that (12)
K = -(d2h/dy2)/h
.
In fact the integrand KdA, where dA is the area element hdyds, still makes sense at the singularities of / on £?,. This will follow from Lemma 4. We will need to compare K with the sectional curvature of M. We denote by KM the sectional curvature of the plane section spanned by f^d/dy, f*d/ds. This function is defined at the nonsingular points of /, and it follows from (8) that KM = —Rjik£j<»iCk(0i/h2 . Consequently we obtain from (9) and (10) that
TOTAL ABSOLUTE CURVATURE
d^/dy2
(13)
= RjinZ&ai
181
,
and therefore KM = -w^cojdy^/h2
.
In fact, the integrand KMdA still makes sense at the singularities of / on Q. This will follow from the next lemma. But before we give the proof we introduce a convenient notation. A sequence of functions such as o)u • • • ,(on will be regarded as the components of a function cu with values in Rn, and we will make use of the standard norm and scalar product on Rn. Thus, for example, (14)
h = \\a>\\ ,
Lemma 2. (15)
KM = -(a>,d2a>/dy2y/h2 .
The function AM defined by
AM = — <
where h =£ 0 ; Au = 0 otherwise,
is continuous on Q. Proof. AM is obviously continuous where h j= 0, and at these points
M*l<£ Hawaii • Clearly this inequality holds at all points of Q. (13) shows that the functions d2
dh/dy =
We will calculate this derivative at the points where h = 0. According to the first equality in (7), o)4 = y^ where ^ = Q)of)dQj/ds. Consequently, for y > 0, h = y \\n\\. Now ^(0) = 8), and therefore //(0, s) = d£/ds. It follows that for points on the line y = 0 the right-hand derivative
9/1/^ = 1^(0^)11 = 11^/^11 . The other points where h = 0 lie on the lines s = a where a is any number such that (d£/ds)(a) = 0. At these points dh/dy = 0. It follows from the formula h = y \\/n\\ that dh/dy is continuous at the points on the line y = 0 where /u^O. The other points at which h — 0 are the points {a, a) such that {dC,/ds)(a) = 0. To demontrate continuity at these points we use the inequality
182
F . BRICKELL & C. C. HSIUNG
(17)
\dh/dy\<\\da>/dy\\.
For h i=- 0 this inequality is a consequence of (16), and it is easy to check that it also holds where h = 0. The functions da)i/dy are of course continuous, and are zero at (a, a). Therefore it follows from (17) that dh/dy is continuous at (a, a). Lemma 4. The function d2h/dy2 is continuous on Q,. Proof. It is convenient to put (18)
A=
-d'h/df
.
We find from (16) that at points where h =£ 0
(19)
A=A
M
- /r'flMi 8 Hawaii 2 - ,dw/dyy).
The points of Q, at which h = 0 lie on the lines s = a where a is such that (dQ/ds)(a) = 0. Consequently A = 0 at these points. The continuity of A at points where h =£ 0 is obvious from (19). According to Lemma 2 the continuity of A at the zeros of h will follow from that of A„ — A. To establish this we obtain from the first equality in (7)
where [fi{] is the matrix inverse to [X{of], The right-hand side is linear in the functions wt with coefficients which are bounded on Qt. Consequently an inequality \\dw/dy\\ < A \\a>\\ = Ah is valid on Q, where A is some constant depending on e. Therefore using (19), we find (20)
0 < AK - A < \\da>/dy\\2/h < A2h ,
and this inequality implies that AM — A is continuous at the zeros of h. We have to consider also the geodesic curvature K of the curve C: s —> (r(s), s) in (Q,f). This curvature is defined at the points where / is an immersion, and it is easy to check that (21)
K = dh/dy - (d2r/ds2)/h .
Let KM denote the geodesic curvature of the curve C in M, t the unit tangent vector of C, and y the second fundamental form of (Q,f). Then from the wellknown formula
TOTAL ABSOLUTE CURVATURE
183
K\ = K2 + the square of the length of y , it is clear that \KX\ > \K\. Our aim is to extend K to an integrable function, defined almost everywhere on the interval [0, L] such that the inequality continues to hold. It is convenient to define k(s) — h(r(s),s), 0 — w{r{s),s) so that K = ||0||. As the parameter s is the arc length along C it follows that (22)
k2 + (dr/ds)2 = 1 .
The following lemma establishes properties of the function k. Lemma 5. (i) k is absolutely continuous on [0, L]. (ii) k is differentiable at the points where it is nonzero. It is differentiable at a zero s = a iff (d
—\TZ < 6 < JK .
It follows from (22) that cos 6 = k. In the induced metric on (£?,/), d is the angle between the tangent vector to C and the s-axis. The required properties of 6 are established in Lemma 6. The function 6 is absolutely continuous on [0, L\. It is differentiable at a point a iff k is differentiable at a. Proof. The function sin -1 is uniformly continuous on the closed interval [—1,1], and so there exists a number X > 0 such that |sin -1 a — sin_1/3] < \n when \a — /3| < X, — 1 < a, B < 1. The function r is C°° on [0, L], and so there exists a number B > 0 such that \d2r/ds2\ < B on [0,L]. Let s1} ,s2 be such that 0 < s17 s2 < L, \sl ~ s2\ < X/B. We will write du r't, kt for the values of the functions 0, r' — dr/ds, k at the points st, i — 1,2. From the definition of 6 we obtain sin (82 — !> = r'2kx — r[k2 = A:1(',2 — r[) — r'i(K — kj ,
184
F. BRICK.ELL & C. C. HSIUNG
and therefore |sin(0 2 -€,)]<
M-ril
+ \k2~
k,\ .
It follows from the mean value theorem that | sin 82 — sin dx | < I. Consequently \62 — d^ < \K, and so we may use the inequality \62 — 6^ < \n\ sin (82 — ^ ) | to conclude that
\0* - 0i\ < Wi
~ <\ + \k* ~ W •
The absolute continuity of 0 now follows from that of the function k (Lemma 5) and the C°° function dr/ds. To investigate the differentiability of 6 we consider first the points where k ^ 0. At these points cos d =£ 0, and the formula sin 6 — dr/ds gives dd/ds = (d2r/ds2)/k
(23)
.
On the other hand, in a neighborhood of a zero of k the function sin 6 =£0, and the formula cos 6 = k gives dd/ds = -{dk /ds) /'(dr/ds) iff dk/ds exists. Since an absolutely continuous function is differentiable almost everywhere [4, p. 205], it follows from Lemma 3, Lemma 6 and the formulas (21) and (23) that the definition (24)
K = dh/dy -
dd/ds
extends the domain of K. to [0, L] except for a set of Lebesgue measure zero. Further, the proofs of the lemmas show that K = 0 where k = 0. Consequently the inequality \KM\ > \K\ holds almost everywhere on [0, L], We are now able to prove the key theorem. The statement of this theorem makes use of the meanings which we have attached to curvatures at the singularities of /. Theorem 1. Let 0 be a point on a closed C° curve C imbedded in a riemannian n-manifold M, and suppose that C lies in a normal neighborhood of 0. Construct the shell (Q, f) on C with vertex 0. Let K be the Gaussian curvature of the induced metric on (Q, f), and use dA for its area measure. Denote by K the geodesic curvature of C considered as a curve in (Q,f), and let s be its arc length. Then [LKds = T: + I — [ KdA , Jo
Jo
185
TOTAL ABSOLUTE CURVATURE
where L is the length of C, and I is the length of the indicatrix of the shell. Proof. Let su L — s2 be the values of s at which the line y = e meets the curve y = r(s). By definition
f KdA =lim f
Jo
>~°Ja,
-—dyds
dy2
= \im(-[L~S2^L(r(s),s)ds -° \ J.i dy
+ {L~S*-M-(s,s)ds) . J** dy I
It follows from Lemma 3 that f KdA = - [L®L(iis),s)ds Ja Jo dy
+ I.
The proof is completed by substituting for dh/dy from the definition (24). For, according to Lemma 6, 6 is absolutely continuous and therefore [4, p. 207] [L-^°-ds = 6{L) - 0(0) . Jo ds Lemma 1 shows that the right-hand side is equal to — K. 3.
An inequality for the total absolute curvature
Theorem 2. Let M be a complete simply connected riemannian manifold with a nonpositive sectional curvature function Ku. Then the geodesic curvature KM of any closed C° curve C imbedded in M satisfies the inequality (25)
f \KM\ds > 2n - f KM dA , Jc Ja
where {Q, f) is any shell on C. Proof. It is well known that such a manifold M is a normal neighborhood of each of its points so that the shell (Q,f) is defined. According to Lemma 1 the indicatrix of the shell joins a pair of antipodal points on a unit sphere and therefore its length / > K. Consequently as \KX | > |/c| we obtain, using Theorem 1,
Jc
\KM\ds > \ |K| ds > \ Kds = it + I — | KdA > 2x — \ KdA . Jo
Jo
Ja
Ja
The proof is completed by (Kx — K)dA = (AM — A)dyds > 0, which follows from (12), (14), (15), (18), (20) (on M wherever K and KM are defined the formula KM — K > 0 is due to J. L. Synge [8]).
186
F. BRICKELL & C. C. HSIUNG
It should be remarked that by a completely different method J. Szenthe [9] has shown that
\>cM\ds> 2K, which is obviously weaker than our inequality
(25). For KM =£ 0, from (25) we prove immediately the following conjecture of N. H. Kuiper: Corollary. The total absolute curvature of any closed C°° curve imbedded in a complete simply connected riemannian manifold with negative sectional curvature is greater than 2K. We can obtain more precise results in the case when M has constant sectional curvature. But we will need some information about imbedded closed curves in the hyperbolic and euclidean planes. As in euclidean space we will say that a subset S of hyperbolic space is convex if {P, Q} c S implies that the geodesic segment joining P to Q also lies in S. An imbedded closed curve C in the hyperbolic plane is said to be convex if it is the boundary of a convex subset of the plane. In our case C will be of class C°% and then an equivalent condition is that C lie entirely on one side of each tangent geodesic. Theorem 3. Let M be a hyperbolic {resp. euclidean) space, and denote its sectional curvature by v. Then the geodesic curvature tcM of any closed C°° curve C imbedded in M satisfies the inequality \ KM | ds > In — vA , where A is the area of any shell on C. If the equality occurs for some shell on C, then C is imbedded as a closed convex curve in a hyperbolic (resp. euclidean) plane in M. Proof. The first statement is a special case of Theorem 2. To prove the second statement we suppose that the equality occurs for a shell (Q,f) on C with vertex 0. According to the proof of Theorem 2 the indicatrix of this shell has length K and is therefore a great semi-circle on the unit sphere in T0M. Consequently C lies in the totally geodesic surface S through 0 tangent to the plane of this great semi-circle. The surface S with its induced metric is a hyperbolic (resp. euclidean) plane in M, and the geodesic curvature of C, considered as a curve in S, is equal to KU. We will need Lemma 7. KM has a constant sign. Proof. It is clear that KM = K at the nonsingular points of /, that is, where h ^ 0. Suppose that KM(a) > 0. Then KM > 0 on some open interval / containing a. The function h is not everywhere zero on /, because that implies the restriction of C to / is a geodesic and so KM(<X) = 0. Therefore K = KM > 0 on some subinterval of/. Similarly, the supposition KM(CC) < 0 implies that K < 0 on some interval of values of s. Consequently, unless KM has a constant sign, there is a strict inequality
TOTAL ABSOLUTE CURVATURE
Jo
187
I K I ds > icds , Jo
and so, from the proof of Theorem 2, a strict inequality in the statement of Theorem 3. It follows from Lemma 7 that C is convex. This fact is well known in the euclidean case [ l , p . 21], and we prove it for the hyperbolic case in the appendix (Theorem 5). Theorem 3 is an extension of a well-known theorem of Fenchel [3]. The following theorem is an extension of the theorems of Fary [2] and Milnor [5]. Theorem 4. Let M be a hyperbolic or euclidean space of dimension three, and denote its sectional curvature by v. Let KM be the geodesic curvature of a C°° knot C in M. Then for some shell on C | KM | ds >
ATZ
— vA ,
where A is the area of the shell. Proof. Let D denote a closed imbedded C°° curve in M such that there is no shell on D which satisfies the stated inequality. Therefore, for any point 0 on D, | K.M j ds < An — vA , where A is the area of the shell on D with vertex 0. On the other hand, the proof of Theorem 2 contains the inequality | KU | ds >
K
+ I — vA ,
where / is the length of the indicatrix of the shell. Consequently / < 3K. Our aim is to show that this condition implies that D is unknotted. We will prove first that the condition implies there is a plane through 0 (hyperbolic or euclidean respectively) which meets D in just one other point and is transversal to D at both points of intersection. Then Theorem 6, which is proved in the appendix, shows that D is unknotted. Consider an oriented great circle G on the unit sphere I in T0M. G lies in a plane y in T0M and gives an orientation to ;-. Choose an orientation for T0M. This orientation is determined by the orientation of y and a unit vector normal to y. The measure of a set of oriented great circles on I is, by definition, the area measure of the corresponding unit normals considered as points of 2. We attach to each great circle G the number n(G) of its points of intersection with the indicatrix of the shell. Applying Crofton's theorem [1, p. 33] we find that
188
F . BRICKELL & C. C. HSIUNG
(26)
\n{G)dG = AKI < 12K .
It is clear that n(G) > 1, which, together with (26) and the following Lemma 8, implies that the set of great circles on 2 with n(G) = 1 has strictly positive measure because otherwise U(G)dG > hdG = 12* . Lemma 8. The measure of the set of great circles on 2 with n(G) = 2 is zero. Proof. The indicatrix £: s —• £(s) joins antipodal points P, Q on I. Any great circle on I meeting £ in just two points either goes through P and Q or is tangent to £ for some s, 0 < s < L ("tangent to C" is to include going through a singular point where dt^/ds = 0). Consider the set 5\ of oriented great circles on I through P and Q. Since their normals lie on a great circle on 2, obviously S^ has measure zero. Consider the set S2 of great circles on I tangent to £ for some s, 0 < s < L. If / = hXiiO) is the normal to the plane of such a great circle, then (27)
kUs) = 0 ,
l&t/ds = 0 .
We will show that S2 has measure zero by using Sard's theorem [6]. Consider the projection I x / —» 2" where / is the open interval (0, L). Let % be the restriction of this projection to the 2-dimensional submanifold N of points (/, s) such that l£i = 0. A point p = (I, s) e N is critical for % iff d/ds is tangent to N at p. This is so iff lid^ds = 0. Consequently the set of critical values of X is the set of unit vectors / satisfying (27). Sard's theorem implies that this set (and hence S2) has measure zero. The set of great circles with n(G) = 2 is contained in the union Sx U 52 and therefore has measure zero, q.e.d. Each oriented great circle G determines an oriented plane a in M (hyperbolic or euclidean respectively) whose tangent plane at 0 contains G. The planes corresponding to the great circles with n(G) = 1 are transversal to D at 0, and meet D in just one other point. We have to show that some of these planes are transversal to D at both points of intersection. Let P denote a general point on D distinct from 0, and a an oriented plane containing P and 0. Consider those planes which are not transversal to D at P. Lemma 9. As P varies on D, the union of the corresponding great circles on I has measure zero. Proof. We will show that this union is the set S2 introduced in the proof of Lemma 8. Consider an oriented plane a containing P and 0, and suppose
189
TOTAL ABSOLUTE CURVATURE
that its normal at 0 is /4*4(0). Then its normal at P is /4x4(P). The plane a contains the geodesic OF and therefore (28)
ltUs) = 0 •
It follows from (8) that a is not transversal to D iff karris), s) = 0. For spaces of constant nonpositive sectional curvature v the functions a>4 are proportional to d£i/ds. To show this we will integrate (13) which simplifies to d2(Oi/dy2 —
—vwi.
Imposing the initial conditions co4 = 0, and dwi/dy = dd/ds when y = 0, we find a>i =
, sinh (V — vy)—p^ , V—v ds
a>i = ydd/ds
,
v < 0, v= 0 .
Consequently a is not transversal to D iff (29) hdZi/ds = 0 . As (28) and (29) are the same as (27), it follows that the set of great circles which we are considering is the set S2. q.e.d. Consequently by Lemma 9, almost all great circles on I with n(G) = 1 determine planes which are transversal to D at both points of intersection. 4.
Appendix
In this appendix we prove two theorems which have been used in our previous work. Theorem 5. Let H be the hyperbolic plane of constant sectional curvature — 1, and C be a closed C°° curve imbedded in H. Then C is convex iff it can be oriented so that its geodesic curvature K > 0. Proof. Let (x, y) be the standard coordinates on R2. We will take as our model of H the subset of R2 where x > 0, together with the metric (dx)2 + {dyf x2 The sign of the geodesic curvature of a curve in H depends on a choice of an orientation for H. We choose the orientation defined by the chart (x, y). Then it is easy to derive the formula (30)
K
= (x'y" - y'x" - xy')/x2 ,
190
F . BRICKELL & C. C. HSIUNG
where the primes denote differentiation with respect to the arc length s. It follows at once that the lines y = constant are geodesies in H, and our proofs will make use of this fact. If a curve C is tangent to such a geodesic, then in a neighborhood of the point of contact we can use the coordinate x as a parameter on C. It is not difficult to verify the formula
We are now ready to begin the proof of the theorem. To show that the stated condition is necessary we suppose that C is convex, but that it cannot be oriented to have its geodesic curvature K > 0. Therefore there are points on C at which K > 0, and also points where tc < 0. Consider a point where K > 0, and choose the arc length parameter so that this point has parameter value zero. Let a be the supremum of those values of s for which n(t) > 0, 0 < t < s. Then ie(a) = 0, and in any neighborhood of a there are values of s for which K(S) < 0. Let P denote the point on C of parameter a. We can use a proper isometry of H to move C so that it becomes tangent at P to the x-axis with x' > 0 at P. The curve C remains convex, and its curvature is unaltered. Because ic(s) > 0 for 0 < s < a, the formula (31) shows that some part of C lies above the x-axis. Therefore, because C is convex and the x-axis is a geodesic, C lies entirely above the x-axis. Of course, if Q is any other point on C, we can again move C so that it becomes tangent to the jc-axis at Q with x' > 0 at Q. But if Q is near to P, then we need only move C a small amount. Consequently, if Q is near enough to P, then C will still lie above the x-axis. We can choose Q so that K < 0 at Q. Because y' = 0 at Q, it follows from the formula (30) that y" < 0 at Q. This inequality contradicts the previous conclusion that C lies above the x-axis. Therefore C can be oriented so that K> 0. To prove the converse implication we begin with a technical lemma. Lemma 10. Let C be a C°° closed curve imbedded in the hyperbolic plane H with geodesic curvature K>0. Suppose that a geodesic y = constant is tangent to C at a point P of arc length parameter a, and that x'ia) > 0. Then there exist numbers a, b such that a < a < b , y(s) > y(a) y(a) > y{a) ,
for
a< s < b ,
y(b) > y(a) .
Proof. Let [a', b'] be the largest closed interval containing a on which K, = 0. On this interval the curve C coincides with the geodesic y = y(«). We can choose a, b to satisfy the conditions a < a', V < b and x' > 0 on the interval [a, b]. The fact that a, b also satisfy the conditions stated in the lemma is an easy consequence of the formula (31). q.e.d.
TOTAL ABSOLUTE CURVATURE
191
To complete the proof of the theorem we suppose that K > 0 on C, but that C is not convex. Consequently there are points of C on both sides of some tangent geodesic. We can move C so that this tangent geodesic becomes the ;c-axis with x' > 0 at the point of contact, P. Lemma 10 gives us information about the shape of C near to P. But there are also points on C for which the y-coordinate is strictly negative. It follows that there is at least one local maximum for y on C at which x' > 0. Lemma 10 implies that such a local maximum cannot exist. Therefore C must be convex. In the next theorem we will be dealing with hyperbolic or euclidean space M of dimension three. Let D be a closed piecewise C°° curve imbedded in M. A plane (hyperbolic or euclidean respectively), which meets D in just two points and its transversal to D at each point of intersection, will be said to be transversal to D. In this definition a plane is transversal to D at a point of intersection of two C° arcs if it is transversal to both arcs and also separates them. It is not difficult to see that if a plane p is transversal to D, then any plane sufficiently near to p is also transversal to D. If D admits a transversal plane through each of its points, then we will say that it has the transversal property. Theorem 6. Let C be a closed piecewise C°° curve embedded in hyperbolic or euclidean space M of dimension three. If C has the transversal property, then C is a trivial knot. Proof. (The idea of the proof is due to T. Poston.) Let D be a closed piecewise C°° curve embedded in M, and suppose that D has the transversal property. We will describe a procedure for splitting D into two new closed curves which also have the transversal property. Choose a point Q on D and a plane q through Q transversal to D. Let R denote the second point of intersection. The plane q divides D into two arcs rit r2, and we complete these arcs to closed curves Du D2 by adding on the geodesic segment QR. To show that the new curves Du D2 also have the transversal property we restrict our attention to D1. First of all note that if P lies on the segment QR, then a suitable small variation of the plane q will provide a plane through P transversal to D^ UP is not on this segment, then there is a plane p through P, transversal to D, which does not go through Q or R. Up meets D again in the arc ru then it cannot meet the segment QR and is therefore a transversal plane for Dx. On the other hand, if p meets D again in the arc r2, then it must also meet the segment QR. It will meet this segment transversally and so will again be a transversal plane for D,. We now outline the small amount of knot theory which we need. For details we refer to [7]. The procedure we have just described factorizes the knot D into the product of the knots D 1; D2. Associated with any knot is a nonnegative integer called the genus of the knot. It satisfies the simple relation. (32)
genus D = genus Dl + genus D2 .
192
F . BRICKELL & C. C. HSIUNG
This relation, together with the fact that a knot is trivial iff it has genus zero, is all that we use. We will say that the closed curve D satisfies condition A if, for any splitting of D into closed curves D„ D2, at least one of the curves is a trivial knot. As a product of a finite number of trivial knots is a trivial knot, Theorem 6 is a consequence of the following lemmas. Lemma 11. A piecewise C°° knot C which has the transversal property can be split into a product of knots each satisfying condition A. Proof. Either C itself satisfies conditon A or it can split into a product of knots Cu C2 which have the transversal property and which are not trivial knots. The relation (32) shows that Cx and C2 will each have a genus strictly less than that of C. If C„ C2 both satisfy condition A, then the proof is complete. Otherwise we can continue by splitting Cx or C2 (or both) into products. This process will terminate after a finite number of steps because the genus of a knot is a nonnegative integer. Lemma 12. A piecewise C°° knot D which satisfies condition A is a trivial knot. Proof. We will say that an arc F of D is trivial, if there is a transversal plane through its end points Q, R, and the closed curve formed by r and the geodesic segment QR is a trivial knot. We first note that any point P on D lies in the interior of a trival arc. For consider a plane p through P transversal to D, and let P' denote its second point of intersection with D. The plane p divides D into two arcs at least one of which is trivial. Rotate p about P' to increase the length of the trivial arc. If the rotation is sufficiently small, the new arc will also be trivial. It will, of course, include P in its interior. Because D is compact, we can cover D by a finite number of trivial arcs. We can also arrange that the cover is minimal in the sense that no arc is contained in the union of the remaining arcs of the cover. Let rit • • •, rn be such a minimal cover, and let Qt, Rt be the end points of I V Divide D into two arcs by a transversal plane qx through Qu Rlr and construct a new knot D' by replacing the arc rx by the geodesic segment QXRX. The relation (32) shows that if D' is a trivial knot, then so is D. Apart from the segment QXRX the knot D' lies entirely on one side of qx, and we will say that it lies in a half-space KX. Now consider the knot D'. Some of the arcs F2, • • • ,Tn will be arcs of D', and the remainder will have just one end point on rx. Let r„ denote an arc with an end point Qa on 7 \ . There is a plane qa through Qa, Ra transversal to D. The plane qa must intersect the segment <2i^i> and we will denote the point of intersection by Q'a. One of the arcs QxRa or RxRa is contained in r„. The union of this arc with the segment QiQi or g ^ respectively, gives an arc r'a olD'. It is not difficult to show that f'„ is a trivial arc of D'. In fact, for this pur-
TOTAL ABSOLUTE CURVATURE
193
pose by assuming Qx [to be on F a we have to show that the closed curve E consisting of the segment Q'jQ^ the part of r„ from Qi to Ra and the segment RjQ'c is a trivial knot. We are given that F a is a trivial arc of D. This means that the closed curve F consisting of the arc ra and the segment RaQa is a trivial knot. Consider the closed curve G consisting of the arc ra, the segment RjQ'a and t n e segment Q'jQa. G is isotopic to F because we can deform the broken geodesic segment RjQ'jQ* into the segment RjQa keeping within the transversal plane qa. Since F is a trivial knot, G is also a trivial knot and therefore genus G = 0. Since the knot G is the product of the knot E and the knot consisting of the arc of r„ from Qa to Qu the segment <2i2i and the segment Q'aQa, the relation (32) implies that genus E = 0, and hence that £ is a trivial knot. We carry out the above modifications where necessary, and obtain a cover of D' consisting of the segment Qfii and trivial arcs r'2, • • •, F'n. We can now construct another new knot D" by replacing r'2 by a geodesic segment lying in a transversal plane q2. As before D" lies in a half-space n2, and if D" is trivial then so is D'. After n such steps we will arrive at a knot S consisting entirely of geodesic segments. S will lie on the boundary of the convex set formed by the intersection of the half-spaces K1} • • • ,nn, and will therefore be a trivial knot. Consequently D is a trivial knot. References [ 1] [2] [ 3] [4] [5] [6] [7] [ 8] [9] [10]
S. S. Chern, Curves and surfaces in Euclidean space, Studies in Global Geometry and Analysis, Studies in Math., Vol. 4, Math. Assoc. Amer., 1967, 16-56. I. Fary, Sur la courbure totale d'une courbe gauche faisant un noeud, Bull. Soc. Math. France 77 (1949) 128-138. W. Fenchel, Uber Kriimmung und Windung geschlossener Raumkurven, Math. Ann. 101 (1929) 238-252. L. M. Graves, The theory of functions of real variables, Second edition, McGrawHill, New York, 1956. J. Milnor, On the total curvature of knots, Ann. of Math. 52 (1950) 248-257. A. Sard, The measure of the critical values of differentiable maps, Bull. Amer. Math. Soc. 48 (1942) 883-890. H. Schubert, Die eindeutige Zerlegbarkeit eines Knotens in Primknoten, S.-B. Heidelberger Akad. Wiss. Math.-Natur. K l . No. 3 (1949) 57-104. J. L. Synge, The first and second variations of the length in Riemannian space, Proc. London Math. Soc. 25 (1926) 247-264. J. Szenthe, On the total curvature of closed curves in Riemannian manifolds, Publ Math. Debrecen 15 (1968) 99-105. J. A. Wolf, Spaces of constant curvature, McGraw-Hill, New York, 1967. UNIVERSITY OF SOUTHAMPTON LEHIGH UNIVERSITY
454 HSIUNG, C. C.
1975 Annali di Matematica pura ed applicata (IV), Vol. CII, pp. 103-107
A Remark on Pinched Manifolds with Boundary (*). OHUAN-OHIH HSITJNG (Bethlehem, P a . , U.S.A.) (**)
Dedicated to Professor BENIAMINO SEGEE
Summary. - The well-known homeomorphism- and diffeomorphism-sphere theorems for compact simply connected pinched manifolds without boundary are generalized to pinched manifolds with boundary.
1. - Pinched manifolds without boundary. Let M be a Riemannian manifold with a Riemannian metric g, and | a twodimensional plane through a point P e ¥ in t h e tangent space of M a t P. M is said t o be 6-pinched with respect to g if t h e sectional curvature R(P, f), with respect to g, of M a t every point P for every f satisfies (1)
0<
6A
where A is a positive constant. For a suitable normalization of g, (1) m a y be expressed as (2)
0 < d
M. B E R G E R [1] and
e)
W . K L I N G E N B E R G [7], [8] proved
THEOREM 1. - If a compact simply connected n-dimensional Riemannian manifold without boundary is d-pinched with d>\ with respect to a metric, then it is homeomorphic to an n-sphere 8". Theorem 1 improves t h e result of H . E . RAUCH [9] who initiated t h e problem with c r > 0 . 7 4 . . . . Since then various authors have been able t o determine t h e pinching constant d for replacing « homeomorphic »in Theorem 1 b y « diffeomorphic ». D. GROMOLL [5], E . CALABI, and Y. SHIKATA [11] obtained a pinching constant <5n
depending on n with limit 1 as n ->• oo, and M. SITGIMOTO, K. SHIOHAMA and H . KARCHER [12] jointly showed d = 0.87. I n this direction t h e most general result so far is t h e following due t o E . A. H U H [10]. THEOREM 2. - If a compact simply connected n-dimensional Riemannian manifold without boundary is 0.80-pinched with respect to a metric, then it is diffeomorphic to 8n. (*) Entrata in Redazione il 16 maggio 1973. (**) Partially supported by the National Science Foundation grant GP-33944.
455 104
CHUAN-OHIH HSIUNG:
A remark on pinched manifolds with boundary
2. — Doubling of a manifold. How let M be a compact oriented 0°° manifold of dimension n with boundary B of dimension n — 1. The oriented B is called a reflecting boundary of M if it admits an orientation-reversing involution n (this definition was first given in [6]). For convenience and simplicity, we shall always denote by (B, n) a reflecting boundary B with its involution n together but with its dimension omitted. A simple example of the reflecting boundary of M is an (n — l)-sphere. In general, a simply connected reflecting boundary may be a homological sphere and at the same time not diffeomorphic to a standard sphere; an example of such a boundary is the boundary of the 12-dimensional manifold of E. BBIESKOEN [2], which is defined by 6
2 «5 + 4 + a£ = e
(for small e # 0),
7
<-l
where ^ , . . . , « , are the coordinates of a point in a 7-dimensional complex space, and z( is the conjugate of zt. Let M with boundary S be a 0™ homeomorphic copy of (M, B) with the same orientation, and p, be the homeomorphism so that p(M, B) = (M, S). Then we can define the double of M with a reflecting boundary (B, n) to be a Ga closed oriented manifold JV such that J = I u I and that B = n{B) by identifying /j,7i{x)eS with a; for all a;eB. Thus on the double N we can define a homeomorphism r: N -s* Jf by H(x),
iov
xeM,
JJT^X) ,
for x e iff.
To see that this is well-defined, at first we notice that v{x) =/x{x) e S for every x e B. Since a; e B is identified with ^ ( a ) e B , v(/m(x)) = /jr^/mix)) = n(x) e B, and therefore n{x) must be identified with n(x); this is indeed true by the definition of our identification and the assumption T I 2 = 1 . Clearly, v is an involution. (It should be noted that the definition of doubling a manifold M here is somewhat different from the ordinary one under which M and M are of opposite orientations so that every point of £ is a fixed point under the involution v.) Alternatively, we may regard doubling the manifold M with a reflecting boundary (.B, n) a finding a 0™ homeomorphism h: M-+N
456 CHtrAN-CHiH
HSIUNG:
A remark on pinched manifolds with boundary
105
where JT is a C° closed manifold with an involution
such that vh—hn:B->N. Assume h maps M into a fundamental domain of the involution v in such a way that B is mapped onto itself. We shall identify M with the fundamental domain henceforth, so that we may regard the double N as composed of two halves M and M with the same orientation such that M is mapped onto M by v, Mnl= B, and v\B = n. A Eiemannian metric on the double N, with respect to which the involution v is an isometry, is said to be symmetric, and the restriction of a symmetric metric on N to M is also said to be symmetric on M. S"ow there arises the problem of deriving a CT symmetric Eiemannian metric on N from a Ga Eiemannian metric on M. For this problem at first we are naturally tempted to prolong to N a differentiable metric g on some manifold containing M by setting (3)
g(x) = g{v(x))
for oce M. Although (3) is well defined, the difficulty is that the resulting metric will, in general, not be differentiable across B. However, on the other hand, for a given C°° Eiemannian metric g everywhere defined on N, we may obtain from it a Ca symmetric metric g by setting (4)
g(<>>) = i[g(sv) + g(v(as))'] ,
iovweN.
Hence on the double N we can always consider a C° symmetric Eiemannian metric. It should be remarked that if the manifold M has empty boundary B, then the double N of M consists of two homeomorphic copies M and M, and therefore every C° Eiemannian metric of M is symmetric.
3. — Main theorem. The purpose of this note is to generalize Theorems 1 and 2 to pinched manifolds with a reflecting boundary by establishing the following theorem. THEOREM. - Let M be a compact oriented simply connected C Eiemannian manifold of dimension n with a reflecting boundary (B, ri) such that the homology groups Ei(B) = 0, i= 1, ..., n — 2. If M is d-pinched with d> i (respectively d= 0.80) with respect to a O™ symmetric metric g, then M is homeomorphic (respectively diffeomorphic) to a compact subset 0 Q 8n with boundary D such that the absolute homology groups Hf(C) = 0 , i = 1,..., n — 1, and O = 8n when and only when B is empty.
457 106
OfftXAN-CHiH HsitTNG: A remark on pinched manifolds with boundary
It is obvious that the above theorem reduces to Theorems 1 and 2 when the boundary B is empty. PKOOF. - As before we define the double N of (M, B) to be a C° closed oriented manifold N = M u M, where M with boundary S is a C° homeomorphic copy of (M, B) with the same orientation, and B is identified with n(B) so that M n M = B. Let g denote the C° symmetric metric on the double N, whose restriction to M is g. Then it is easily seen that the double N is d-pinched with d > J or d = 0.80 with respect to g, since M is so with respect to g by the hypothesis of the theorem. For the case where d > J (respectively <5 = 0.80) by applying Theorem 1 (respectively Theorem 2) to the double N we thus obtain a homeomorphism (respectively diffeomorphism) f:N-> 8". Denote f(M)=G, f(M) = C, f(B) = D, so that G\JC= 8" and Cr\C=B. Then by the definition of the double N and the hypothesis of the theorem we have (5)
Ht{G) = H,(C)
for all i,
(6)
Ht(D)=0,
i=l,...,n
— 2.
Now consider the Mayer-Vietoris sequence of the proper triad (8n; G, C): (7)
... ->ff,(D) ^Et(G)
+ #,(£) -*>#,(«») -^HUD)
->•••
in which the homomorphism zl is a composition of the boundary operator d and the homomorphisms induced by inclusion maps, the homomorphism
ueHi(D),
where
hl9:Hi(B)^Hi(C),
h^-.H^^E^C)
are induced by inclusion maps h^.B-^G,
h2:B-+C.
Since the sequence (7) is well-known to be exact (see, for instance [4, p. 39]), from (6) and -ff<()Sn)= 0 for i= 1,..., n — 2 it follows immediately that f o r l < » < « — 2,
ff,(C)=0,
»=!,...,» —2.
458 CHXTAN-CHIH H S I U N G : A remark on 'pinched manifolds
with boundary
107
On t h e other hand, by means of H^^S") = 0 and t h e exactness of t h e sequence (7) we readily see t h a t xp for i = n — 1 is onto. Since D is t h e boundary of both G and 0, hlmu = 0 and h2^u = 0 for every u e H^^D), so t h a t for i = n — 1 the image of y> is zero. Thus b y using (5) we obtain (9)
ff.-i(0)
= 0.
Hence combination of (8), (9) completes t h e proof of our theorem. I t should be remarked t h a t J . CHBEGEE and D . GEOMOLL [3] jointly obtained a result similar t o our Theorem with our pinching condition on M and boundary conditions on B replaced by t h e condition t h a t every sectional curvature of M be only positive and t h e condition t h a t t h e boundary 2? be nonempty totally and convex. Finally, for t h e validity of our Theorem it is natural t o ask what are t h e most general boundary conditions on B and metric g on M, t o which our pinching condition is referred.
BEFEEENCES [1] M. BERGER, Les varietes riemanniennes (l/4:)-pincees, C. E. Acad. Soi. Paris, 250 (1960), pp. 442-444; Ann. Scuola Norm. Sup. Pisa, (3), 14 (1960), pp. 161-170. [2] E. BRIESKORN, Beispiele zur Differentialtopologie von Singularitaten, Invent. Math., 2 (1966-67), pp. 1-14. [3] J. CHEEGER - D. G-ROMOIX, On the structure of complete manifolds of nonnegative curvature, Ann. of Math., 96 (1972), pp. 413-443. [4] S. ELLENBERG - N. STUENROD, Foundations of Algebraic Topology, Princeton University Press, Princeton, 1952. [5] D. GROMOLL, Differenzierbare Strukturen und Metriken positiver Krummung auf Spharen, Math. Ann., 164 (1966), pp. 353-371. [6] C. C. HSIUNG, The signature and G-signature of manifolds with boundary, J. Differential Geometry, 6 (1972), pp. 595-598. [7] W. KLINGENBERG, Contributions to Biemannian geometry in the large, Ann. of Math., 69 (1959), pp. 654-666. [8] W. KLINGENBERG, ifber Miemannsche Mannigfaltigkeiten mit positiver Krummung, Comment. Math. Helv., 35 (1961), pp. 47-54. [9] H. E. RAUCH, A contribution to differential geometry in the large, Ann. of Math., 54 (1951), pp. 38-55. [10] E. A. RUH, Krummung und differenzierbare Struhtur auf Spharen, II, to appear in Math. Ann. [11] Y. SHIKATA, On the differentiable pinching problem, Osaka Math. J., 4 (1967), pp. 279-287. [12] M. SUGIMOTO - K. SHIOHAMA - H. KABCHER, On the differentiable pinching problem,
Math. Ann., 195 (1971), pp. 1-16.
459 BULLETIN OF THK INSTITUTE OF MATHEMATICS ACADEMIA SINICA Volume 3, Number 1, June 1975
THE GENERALIZED POINCARE CONJECTURE ON HIGHER DIMENSIONAL MANIFOLDS WITH BOUNDARY BY
CHUAN-CHIH HSIUNG To Ky Fan and Sze-Tsen Hu, on the occasion of their sixtieth birthdays Abstract. The generalized Poincar6 conjecture is proved for some higher dimensional combinatorial manifolds with boundary which may not be connected.
1. Introduction. The well-known Poincare' conjecture states that every closed simply connected 3-manifold is homeomorphic to a 3-sphere S3, and the generalized Poincar6 conjecture that every closed ^-manifold which has the homotopy type of an w-sphere S" is homeomorphic to S". The generalized conjecture for n > 5 was Proved first by S. Smale and immediately after by J. R. Stallings and E. C. Zeeman, while the cases w = 3, 4 are still open; their results are the contents of the following two theorems: THEOREM A. Let M" be a closed C* n-manifold which has the homotopy type of S", n > 5. Then M" is homeomorphic to S".
B. Let M" be a closed connected combinatorial nmanifold, n > 5. If M" has the homotopy type of S", then M" is homeomorphic to S". THEOREM
Theorem A was first proved by Smale [4, 5]. Stallings [7] obtained a Proof of Theorem B (and hence Theorem A) for n>7, and Zeeman [9, 10] showed that Stallings' proof can be adapted to cover the cases n = 5, 6. Smale [5] gave a different proof of Theorem B by using a constructed differentiable structure. The purpose of this paper is to extend Theorem B to a compact oriented combinatorial manifold with boundary. Received by the editors April 18, 1975. Research partially supported by the National Science Foundation grant GP-43665.
177
178
CHUAN-CHIH HSIUNG
Let Mn
2. Doubling of a manifold.
be a compact -1
combinatorial w-manifold with boundary B" 1
[June
oriented
of dimension n — 1, let
be a homeomorphic copy of {Mn,
M" with boundary B"-
B"-1)
with the opposite orientation, and let /i be the homeomorphism so that fi{Mn, B"-1) = (Mn, B"-1). l
M" with boundary B"~
T h e n we can define the double of
to be a closed oriented combinatorial wn
manifold N" such that N = M"uMn ^(j)eB""
1
with x for all s e B "
-1
and S " - 1 = J B»- 1 , by identifying .
(For another type of doubling
a manifold see the author's paper [2].) can define a homeomorphism
T h u s on the double N" we
v : N" —> N" by
v[x) = tx {x) , l
= fi~ {x) ,
for
x e M" ,
for
jeM".
To see that this is well-defined, we first notice t h a t v (x) = ii {x) e B " - 1 for every l e B * " 1 . Since xeB"-1 is identified with M(X) e B " " 1 , y(a?) = v ( , a ( # ) ) = A ~ I ( / ' ( ^ ) ) =xeBn~1, and therefore # ( # ) must be identified with x; this is indeed the definition of our identification. Clearly, v is an involution with all points of B as its only fixed points. Alternatively, we may regard doubling the manifold M" with boundary B"~l as finding a homeomorphism h : M" -> N", where N" is a closed combinatorial ^-manifold with an involution v : N" -» N". Assume that h maps M" into a fundamental domain of the involution v in such a way that B"'1 of the involution v.
is mapped onto the set of fixed points
We shall henceforth identify M" with t h e
fundamental domain, so that we may regard the double N" as composed of two halves M" and Mn
with opposite orientations such
that M" is mapped onto M" by v, and M" nM" = B—1. It should be remarked that if t h e manifold M" has empty boundary B"-1, then t h e double N" of M" consists of two homeomorphic copies M" and
M".
3. Main theorem. T h e purpose of this Paper is to extend Theorem B to a compact oriented combinatorial manifold with
1975]
GENERALIZED POINCARE CONJECTURE
179
boundary by establishing the following theorem. Let M" be a compact oriented connected combinatorial n-manifold, n>5, with boundary B"~1 of dimension n — 1. Let N" be the double of the manifold M", let xk{Mn) be the kth [absolute) homotopy group of M", and let nk{N", M") be the kth relative homotopy group of N" mod M". If N" is simply connected, and the boundary operation THEOREM.
(3.1)
d :x*+1{N;
,
M")~>K,(M")
k = l,--,
in/2],
is an isomorphism onto, then M" is homeomorphic to a compact subset C C S" with boundary D"~l such that the absolute homology groups (3.2)
Hk(C»)=0,
the relative homology
l,---,n-l,
groups
Hk(C», D"-1) = 0 ,
(3.3) and
k =
C" = S" when and only when
k = l,---, B"~l
is
n-l,
empty.
Since xk{X, X) = 0 for any space X and any k, it follows that this theorem reduces to Theorem B when the boundary B"~1 is empty, and that the generalized Poincare" conjecture for the cases n = 3, 4 can also be extended to this form as follows: The above theorem is also true for n = 3, 4. It should be also remarked that for connected and simply connected B"-1 this theorem is due to Smale [6, p. 135] and can be easily proved by using the Stallings' theorem [8, p. 251] on the invertibility of ^-cobordisms. Proof.
Consider the homotopy sequence of the pair ^
nk+1{N\
M»)
(Nn,
M"):
-^^(APJ-^^JV)
- ^ K„ (N», M") -±+ *»-,(M") — i * • • • in which 9 is the boundary operator, and «"*, .7* are the homomorphisms induced by the inclusion maps i : M"^N", j : Nn-*(N", M"). Since the sequence (3.4) is exact (see, for instance, [3, p. 115]), it follows from (3.1) that i* ith{Mn) = 0 and therefore t h a t /* : 7ik(Nn) -> Kk{N", M") is an isomorphism into. On the other hand, because of (3/n W e have 7* Kk{N") = 0. T h u s we obtain nk{Nn) = 0,
180
CHUAN-CHIH HSIUNG
[June
k — 2,- • •, [n/2], and hence, in consequence of t h e simple connectedness of N", the Poincare duality and the Hurewicz isomorphism, (3.5) (3.6)
x»(N") = 0 ,
k = !,•••,
n - 1 ,
HUN")=0,
k = l,--;n-l.
1
Let (M», B"- ) be a homeomorphic copy of {M", B"-1) the opposite orientation such that (3.7)
MnuM"
= Nn,
M"n M» = B—1
with
( = B—' by identification).
Then (3.8)
HUM") = HUM") ,
k =
l,---,n-l.
Now consider the Mayer-Vietoris sequence of t h e proper triad (N"; M", M"): > Hk+1(N")
-U, HUB"-1)
-^Hk{M")
-t->Ht[N')
-U
+ HUM") Ht-UB*-1)
>•••
in which the homomorphism A is a composition of the boundary operator 9 and the homomorphisms induced by inclusion maps, the homomorphism 4> is induced by inclusion maps, and the homomorphism V is defined by iru = (hi u, —h2 u) ,
ueHkiB"-1)
,
where hh : H^B"'1)
>HUM")
,
ht
> HUM")
are induced by inclusion maps ht.B"-1
>M",
hi.B"-1
>M".
Since t h e sequence (3.9) is well known to be exact (see, for instance [1, P. 39]), from (3.6) it follows immediately that -f is an isomorphism onto for l
— 2. By using (3.8) we thus obtain a homomorphism
1
hljf : HUB— ) ~->Hk{M") and an isomorphism-onto ( & v hj -+HUM") (3.10)
+ HUM") HUB-1)=0,
: HUB—1)
for k = 1,- • •, n - 2. This is possible only if HUM")=0,
k =
\,---,n-2.
On the other hand, since V for i = n — 1 is onto, and B"*1 is the boundary of both M" and M", it follows that hi. u = 0 and
1975]
GENERALIZED POINCARE CONJECTURE
181
ht K = 0 for every «Gff„-i(5" _ I ), so that for i = n — 1 the image of ty is zero. Thus by (3.8) we obtain (3.11)
H„-x (M«) = 0 .
Since, by (3.5), Nn has the homotopy type of S", applying Theorem B to N" we obtain a homeomorphism f:N*^> S". Set
f(M') = C",
ftB-1)
= D"-1,
so that C is a compact subset of S" with boundary D"'1. Thus (3.2) follows from (3.10), (3.11), and use of the Lefschetz duality gives (3.3). Hence the theorem is proved. REFERENCE 1. S. Eilenberg and N. Steenrod, Foundations of algebraic topology, Princeton University Press, Princeton, N. J., 1952. 2. C. C. Hsiung, The signature and G-signature of manifolds with boundary, J. Differential Geometry 6 (1972), 595-598. 3. S. T. Hu, Homotopy theory, Academic Press, New York, 1959. 4. S. Smale, The generalised Poincart conjecture in higher dimensions, Bull. Amer. Math. Soc. 66 (1960), 373-375. 5. , Generalized Poincart's conjecture in dimensions greater than four, Ann. of Math. 74 (1961), 391-406. 6. , A survey of some recent developments in differential topology, Bull. Amer. Math. Soc. 69 (1963), 131-145. 7. J. R. Stallings, Polyhedral homotopy-spheres, Bull. Amer. Math. Soc. 66 (1960), 485488. 8. , On infinite processes leading to differentiability in the complement of a point, Differential and combinatorial topology (A symposium in honor of Marston Morse), Princeton University Press, Princeton, N. J., 1965, pp. 245-254. 9. E. C. Zeeman, The generalised Poincart conjecture, Bull. Amer. Math. Soc. 67 (1961), 270. 10. , The Poincart conjecture for » ; > 5 , Topology of 3-manifolds and related topics, Prentice-Hall, Englewood Cliffs, N.J., 1962, pp. 198-204.
LEHIGH UNIVERSITY, BETHLEHEM, PENNSYLVANIA
isoi5. U. S. A ,
464 Annali di Matematica pura ed applicata (IV), Vol. CIX, pp. 289-304
Congruence Theorems for Compact Hypersurfaces of a Riemannian Manifold (*). OHUAN-OHIH HSIUNG
- TIMOTHY P. Lo (Bethlehem, Penn.) (**)
Summary. - Let Mm, Mm be two m-dimensional compact oriented hypersurfaces of class C3 immersed in a Riemannian manifold Bm+1 of constant sectional curvature. Suppose that Bm+1 admits a one-parameter continuous group G of conformal transformations satisfying a certain condition (which holds automatically when G is a group of isometric transformations). Suppose further that there is a 1 — 1 transformation TxeG between Mm and Mm such that P=TT(p)P for each Pe Mm and each PeMm. If the r-th mean curvature for any r, Kr<m, of Mm at each point PeMm is equal to that of Mm at the corresponding point P= TZ(P)P, together with other conditions, then Mm and Mm are congruent mod G. This is a generalization of a joint theorem of H. HOPF and Y. KATSTJRADA [5] in which G is a group of isometric transformations.
0. - Introduction. Let Mm, Mm be two compact oriented hypersurfaces of class O3 immersed in a Euclidean space Em+1 of dimension m + 1 ( > 3 ) . Suppose t h a t there is a parallel transformation / in a fixed direction e in Em+1 (respectively a central transformation with a fixed center 0 in Em+1): Mm^-Mm such t h a t t h e line joining every point PeMm to its image PeMm is parallel to e (respectively passes through 0). Let Hr (respectively Hr) be the r-th mean curvature of Mm (respectively Mm) at P (respectively P). I n the last two decades various authors have been able to show t h a t under t h e condition Hr(P) = HT(P) (0.1)
(respectively \OP\'Hr(P) = \OP\'Er(P), from P to 0)
where
\OP\ denotes t h e distance
for l < r < w and every P , together with other conditions, / is a translation (respectively homotheticity). The pioneer authors were H . H O P F and Voss [3], who jointly studied the case where m = 2, r = l for parallel transformations. Later HSITJNG [6] a n d Voss [11] independently extended this joint work of Hopf a n d Voss
(*) Entrata in Bedazione il 13 Giugno 1975. (**) The first author was partially supported by the National Science Foundation grant GP-33944.
465 290
CHUAN-CHIH HSITJNG
-
TIMOTHY
P. Lo: Congruence theorems, etc.
to the case of general m with r = 1 and all r respectively. On the other hand, for central transformations Hsu [7] studied the case where m = 2, r = 1; STONG [9] the case of general m with r = 1; and AEPPLI [1] the case of general m with all r. In order to extend the above results to two compact oriented hypersurfaces Mm, Mm of class O3 immersed in a Eiemannian Manifold Bm+1, suppose that Bm+1 admits a one-parameter continuous group 0 of transformations, and further that there is a 1-1 transformation TreO between Mm and Mm such that P=T T . ( P ) P for each P e Mm and each P e Mm. Consider an additional family of hypersurfaces &™= Tr(p)Mm for each PeMm, and let Hr (respectively Hr, 3r) be the r-th mean curvature of Mm (respectively Mm, Mm) at P (respectively P) KATSUEADA [8] and HOPF and KATSUEADA [4] have shown that under the condition (0.2)
H1(P) = S1(P)
for each PeM™,
together with other conditions, Mm and Mm are congruent mod TT, i.e., T(P) = const. Eecently, HOPF and KATSUEADA [5] obtained the congruence theorems by replacing condition (0.2) by (0.3)
Hr(P) = Hr{P)
for l < r < m and every P ,
and further assuming that Bm+l is of constant sectional curvature and G is a group of isometric transformations. These theorems are generalizations of all of those mentioned above for Euclidean Bm+1. The purpose of this paper is to generalize the latter congruence theorems of Hopf and Katsurada to the case where O is a group of conformal transformations satisfying a special condition which holds automatically for isometric transformations. In § 1 we define the vector product of m vectors of a Eiemannian manifold P m + 1 of dimension m + l > 3 at a point P and the generalized covariant differentiation on a hypersurface in Bm+1. In § 2 principal curvatures, mean curvatures and fundamental forms are defined and discussed on a hypersurface Mm immersed in a Eiemannian manifold Bm+1. Certain contravariant tensors of the second order associated with the second fundamental form on Mm are introduced, and a basic vector-valued m-form on Mm is expressed in terms of the combined operator of the vector product on Bm+1 and the exterior product of differentials on Mm. In §3 we assume that the manifold Bm+X admits a one-parameter continuous group 0 of conformal transformations, and discuss a one-parameter family of hypersurfaces of Br*1 containing Mm, Mm, which are in a 1-1 correspondence with corresponding points along the orbits of the transformations of Q. § 4 is devoted to the derivation of our basic integral formulas for a pair of compact oriented hypersurfaces Mm, Mm on a Eiemannian manifold Bm+1 of constant sectional curvature discussed in §3. § 5 contains our congruence theorems.
466 OHUAN-CHIH HSITJNG
-
TIMOTHY
P. Lo: Congruence theorems, etc.
291
1. — Vector product and generalized covariant differentiation. Throughout this paper we shall follow the usual tensor convention that repeated indices imply summation, and we shall suppose that all Latin indices take the values 1,..., m + 1 and Greek indices the values 1,..., m unless stated otherwise. Let Bm+1 be a Biemannian manifold of dimension m + 1 > 3 and class (73, and 1 x ,..., xm+1 a set of local coordinates of a point P on Bm+1. Let aijdxidx' define the Biemannian metric on Bm+1 so that (afj) is a positive definite symmetric matrix. Let Ai,..., Am be m vectors of Bm+1 at P, and A\ be the contravariant components of Aa in the local coordinates x1,..., xm+1 of P. Let J.! X... X An be the vector product of the m vectors Alt..., Am, which is denned to be the vector of Bm+1 at P whose j-th contravariant component is S' (1.1)
n
(A1x...xAm)'={-l) a-*
a
a^-i
8' a
••• «m+i^-i
where d\ are the Kronecker deltas, and a is the determinant \a{j\. Let J be a vector of Bm+1 at P with contravariant components J' in x1,..., xm+1. Since the scalar product of any two vectors Aa and Aft is denned by (1.2)
Aa-A0=aiiAlxA'l!,
it follows that (1.3)
J - C ^ X . .xAm) =
(-l)mVa\J,A1,,. • ? -"-ml
j
where \J, Ax,..., Am\ is a determinant, the entries of each of whose columns are the contravariant components of the vector indicated. Thus by (1.2) it is readily seen that the vector A1X...xAm is orthogonal to each of the m vectors Alf..., Am. Now consider a 0 3 hypersurface Mm of the Biemannian manifold Bm+1. The position vector X of a point P on Mm can be given by the vector equation X= F(u1,...,um) where the m parameters u1, ...,um take values in a simply connected domain S of the m-dimensional real number space. The components of Ffa1,..., um) are of class G3, and the rank of the Jacobian {dXjdux) is m at all points of 8. Let the first fundamental form of the hypersurface Mm at the point P be (1.4)
I — gatsduadu0,
where the matrix (<jra/9) is positive definite, so that |
gap =
aij(dxildu")(dxildue).
467 292
OHUAN-CHIH HSITJNG - TIMOTHY P . L o : Congruence theorems,
etc.
The element of area of Mm at P is given by (1.6)
Vgdu1A...Adum,
dA =
where d denotes the exterior differentiation, and A the exterior multiplication. Let Pex... em+l be a n orthonormal frame of JBm+1 at P , so t h a t eu ..., em+1 form an ordered set of mutually orthogonal unit vectors of Bm+1 at a point P . Thus
d-7)
V«,= «»«?«?=*«,
where d(i are t h e Kronecker deltas. The position vector X of the point P is defined to be t h e vector of Bm+1 at P with the local coordinates x1,..., xm+1 of P as its cont r avariant components. To study t h e hypersurface Mn, we choose et,..., em to be tangent to Mm at P so t h a t em+1 is the unit normal to Mm at P . Thus (1-8)
em+1-X,«=0
(a = l , . . . , m ) ,
where (1.9)
Xa=ax/3w«,
and we can write (1.10)
6iX...X
where c is a function of the it's. I n order to find an expression for c, we consider two matrices (1.U)
?=(tf),
V=(v4),
where
the superscript of the element
<+1=(-l)"+'+1B'a-*
(r = l , . . . , m + l ) ,
where Br is the determinant of the m-th order obtained by deleting the r-th column from t h e matrix cp. Substitution of (1.13) in (1.7) for i = j = m + l gives (1-14)
ca=P,
468 CHUAN-CHIH HSITJNG - TIMOTHY P . L o : Congruence theorems, etc.
293
where -B2 ...
(~i)mBm+1 m+l
(1.15)
B =
which is equal to t h e sum of t h e products of t h e corresponding determinants of t h e m-th order of t h e two matrices (1.11). B y an elementary theorem on determinants (see, for instance, [2, p . 102]), from (1.7) it follows immediately t h a t
*=ltfvjl = i.
(1.16)
Thus for an orientable hypersurface Mm, due to (1.10), (1.14) and (1.15), we can choose t h e direction of t h e unit normal vector em+1 such t h a t c = l so t h a t (1.10) becomes (1.17)
e± X . . . X em — em+i.
Taking t h e scalar product of t h e vector em+1 with each side of (1.17) and using (1.3), we can easily obtain (1.18)
«i»
Similarly, by choosing t h e orientation of t h e frame Pex ...em in t h e tangent space of a n orientable hypersurface Mm a t a point P t o b e t h a t of t h e frame P X ( 1 . . . X_„ we can have (1.19)
XAx...xX>m = Vgem
Now we are in a position t o introduce t h e generalized covariant differentiation D, which is useful for studying submanifolds of Eiemannian manifolds. Let A}x be a mixed tensor of t h e second order in t h e <»'s a n d a covariant vector in t h e M'S, as indicated b y t h e Latin a n d Greek indices. Following T U C K B E [10], t h e generahzed covariant derivative of A)x with respect t o t h e it's is defined b y (1.20)
DfA*. = dA}jdup+rklAfx^ - r^ALti, - r*A%
where T\k with Latin indices are t h e Ohristoffel symbols formed with respect to a 0 a n d t h e a?'s, a n d F^y with Greek indices with respect t o gxfl a n d t h e w's: (1.21) (1.22)
pi _ 1 „ (San ik ~2a \dx* <*
2g
+
[du?^
dakl __ daik\ 9a!' dec') ' d
469 294
CHUAN-CHIH HSIUNG
-
TIMOTHY
P. Lo: Congruence theorems, etc.
(a*') and (g*?) being the inverse matrices of (a«) and (<jra/3) respectively. It should be noted that this definition of generalized covariant differentiation can be applied to any tensor in the w's and a?'s, and that the generalized covariant differentiation of sums and products of tensors obeys the ordinary rules. If a tensor is one with respect to the w's only, so that only Greek indices appear, its generalized covariant derivative is the same as its covariant derivative with respect to the w's. Furthermore, in generalized covariant differentiation the fundamental tensors au and gxp can be treated as constants. Since xl is an invariant for transformation of the w's, its generalized covariant derivative is the same as its covariant derivative with respect to the it's, which is denoted by V, so that (1.23)
DaX* = VXX< = dX'jdu".
2. - Curvatures and fundamental forms. Consider an oriented hypersurface Mm immersed in a Biemannian manifold Rm+l so that we can apply § 1. Let N be the unit normal vector of Mm at a point P. Then the first fundamental form I of Mm at P and the second fundamental form I I of Mm at P with respect to N are defined to be (2.1)
l=dXdX,
11=
-dX-dN,
or (1.4) and (2-2)
11= b^du'du*
(bx„=bxl!).
Since I is positive definite, the eigenvalues klt..., fcm of I I relative to I or the roots hlf...,lcm of the determinant equation 1 6 ^ — ^ 1 = 0 are all real and called the principal curvatures of Mm at P . The r-th mean curvature BT of Mm at P is defined to be the r-th elementary symmetric function of fcx,..., fcra divided by the number of terms, that is,
(2.3)
(rW= 2 K-K \
where I (2.4)
/
(Kr<»),
a 1 <...
1 is the binomial coefficient, and for convenience we put Ht = l .
In particular, H„ is called the Gauss-Kronecker curvature of M™ at P . If Hm is nonzero, then the roots of the determinant equation |A6a/?+ gxf!\ = 0 are the negative values of g 1 = l/fc l( ..., qm = ljkm, which are called the radii of principal curvature.
470 CHTTAN-CHIH HSIXJNG - TIMOTHY P . L o : Congruence theorems, etc.
295
By noticing (& ... e J " 1 = \bxjl\lg, we can write g~l\)Aag-\-gae\ = I M _ 1 I ^ W + gM(Qi ••• Sm)'1 = _ •?> J r 2 (product of t h e Qa.,'8 taken m — r at a time)
= 2*r[ 2 (^-ej- 1 !, r-0
L«i
<*r
J
which thus gives, in consequence of (2.3),
(2.5)
r'^
+ whl^l
k,
when Hm^0. Let J1* be t h e adjoint matrix of a square matrix 2 ? = (-Fa/3) so t h a t FF*= \F\(S^). The general element (A&^-f- jraj3)* of t h e adjoint matrix of t h e matrix (A&^+gr^) is a polynomial in A of degree m — 1, so t h a t we can write m
/ * i — 1 \
r ^ ^ + f e ) * = r ?/ r _ 1 ( P _ i ) c ^ -
(2.6)
which defines t h e contravariant tensor df of t h e second order for a fixed r, l < r < m . I n particular, we have (2.7)
c?=g«f,
Since t h e form I is positive definite, it is well known t h a t b y a suitable real nonsingular linear transformation of t h e local coordinates u1, ...,um we can reduce gap and 6a/3 a t a n y point P of Mm to s. ^=<W,
(2.8)
f ka ^ = ( o
for x = /5, for a ^ / 3 .
By means of (2.8), from (2.6) it follows t h a t (2.9)
cf = 0
foree^/S,
cf=-~
for a = /3, T
OrCr
and therefore we can easily show (2.10)
(m - 1)! cf =
where bvx= gPybx/s, and (2.11)
£ ai ... Km =VffSgn ( a 1 ? . . . , « „ ) .
-
K:\,
471 296
CHUAN-CHIH HSIUNG
- TIMOTHY P. Lo: Congruence theorems, etc.
sgn(a 1; ..., o^) denoting the sign of the permutation on the m numbers 1,..., m into «D •••> «m) which is zero if two of the a's are equal. From ey _Vm we define the contravariant tensor (2.12)
f*-*=
g f " " ' . . . ^ " ^ ^ m = g-tigafa,...,
am).
A use of (1.6) and (2.12) gives immediately duXiA...Adu"m=e°Cl-"mdA.
(2.13)
Let A1,..., Am be m vectors at a point P of the hypersurface Mm, and suppose that the components of each vector A" (a = l,..., m) in the local coordinates x1, ...,xm+1 of P are differentiable functions of the m variables u1,..., um. By using the combined operator x of the vector product x of vectors and the exterior product A of differentials, we can define the vector (2.14)
A1x...xA'9dAi+^x...
xdA™ =
= ( i 1 X . . . X l i X - ° w 4 - ' ' + 1 X...XDimA™)dui'+1A-~Adui°>,
(j = 1,...,m).
I t is obvious that the vector denned in (2.14) is independent of the order of dAi+1,...,dAm. From (1.6), (1.7) and (1.19) we can easily obtain (2.15)
dXx...XdX=m\NdA.
3. - Infinitesimal conformal transformations and lemmas. Consider an (m+l)-dimensional (m>2) oriented Biemannian manifold Em+1 of class 0s, which admits an infinitesimal conformal transformation (3.1)
X=X+£(x)dt,
where X is the position vector of a point P on Bm+1, whose contravariant components are the local coordinates x1,..., x™*1 of P , f is an infinitesimal conformal vector, and <5T is an infinitesimal displacement. Suppose that the orbits of the transformations generated by £ cover Rm+1 simply and that f is continuous and nonzero everywhere. Then we can choose a local coordinate system such that the orbits of the transformation are the a;1-coordinate curves so that (3.2)
1 = ^ ,
where <5X is a Kronecker vector whose contravariant components are the Kronecker
472 CHXTAN-CHIH HSIUNG
-
TIMOTHY
P. Lo: Congruence theorems, etc.
297
deltas d\,...,d™+1. Thus (3.1) becomes (3.3)
l=X+S1dr,
and Bm+1 admits a one-parameter continuous group G of transformations which are 1-1 mappings of Bm+1 onto itself given by (3.3) in the special coordinate system. Now consider two oriented hypersurfaces Mm and Mm of class C% imbedded in Bm+1 and given by Mm:X=X(u), (3.4) _ _ Mm: X = X(u) + d^u), where u1,..., um are local coordinates of Mm, and r is a continuous function at each point of Mm. We shall assume that Mm and Mm do not contain a piece of a hypersurface of the form f(%*,..., xm+1) = 0 covered by the orbits of the transformations of G. Between Mm and Mm, whose corresponding points are along the orbits of the transformations of G, we can define a parametric linear family of hypersurfaces as follows: (3.5) Clearly, M^= (3.6)
Jf™ = (l-t)M™+tM™,
(0
Mm and M<*>=Mm. From (3.4) it follows that Mfa is given by Xit)(u)=(l-t)X(u)
+ tX(u),
(0<«1),
or (3.7)
Xu)(u) = X(u) + dltr(u),
(0<<<1) .
Let (gxp), (gxp), (da) be the symmetric matrices of the positive definite metrics of Mm, Mm, Bm+1 respectively, and denote their determinants and inverse matrices °y 9) 9, a a n d (9^), (jf'')) (a") respectively. We shall follow the usual tensor convention that on each manifold the indices of tensors can be raised and lowered respectively by using the contravariant and covariant tensors of the respective metric tensor. We shall also use the same symbols and numbers, as for Mm, with the subscript (t) respectively for the corresponding quantities and equations for Jf(™. Let N be the unit normal vector of Mm at the point X(u), and we shall denote the derivative with respect to t by the accent. Then it is well known that (see, for instance, [12, Chaper VIII]) (3.8)
vj-r- =
(3.9)
D.* =: - M * X y ,
(3.10)
*>,»«»--DpK=-
b«X,
where bx/) are given by (2.2), and Bm
Zm^K
**W>
are Biemann symbols for the tensor a(j.
473 298
CHUAN-CHIH HSITJNG
LEMMA
-
TIMOTHY
P. Lo: Congruence theorems, etc.
3.1. - The vector N[ty lies in the tangent space of M^ and satisfies K=-(ifir
(3-11)
where the dot denotes the scalar product of two vectors in Bm+1, and (3.12)
r a = drldu",
Xm
= aZ m /3«".
- Since JVr(j)-X(()a= 0 and N{1)-N^t) = l, by differentiation with respect to t we have PROOF.
(3-13)
^ • % + ^ ) ' 4 « =
(3.14)
0
'
Nm-N'm=Q.
Prom (3.14) we may write (3.15)
IT'm=^Xm,
where c*3 are to be determined. Differentiating (3.7) with respect to u" and t, and substituting the resulting equation in (3.13), we obtain (3-16)
^
Substitution of (3.15) in (3.16) and use of (1.5)(() give c"= - (V#(())»roT«> from which and (3.15) follows immediately the required (3.11). LEMMA 3.2. - The expression \/gw(81-Nu)) is independent of t, or precisely, (3-17)
v W i •-»"«>) =
Vg(^-N),
where g>(()= \g(M,\. - It follows easily from (1.19)(j), (1.3), (3.7) and (1.19). From (3.17) and (1.6) we obtain readily PROOF.
LEMMA 3.3.
(3.18) LEMMA
(3.19)
(d1-ITm)dAV)=(i1-N)dA. 3.4.
V A * + V , a „ = 2eo„,
where Q is a nonzero scalar function of as1, ...,xm+1 (3.20)
2Qati = datfjdx1.
satisfying
474 -
CHTJAN-CHIH HSIUNG
TIMOTHY
P. Lo: Congruence theorems, etc.
299
- Since dx is an infinitesimal conformal vector field on Bm+1, by definition we have (3.19). On the other hand, from the definition of covariant differentiation, it follows that Vidlj = amVid\=aikr\i, and similarly, V i ^ u = a t t . J ^ . By means of (1.21) we can easily obtain ajir^i+ a(kT\i= da^/dx1. PROOF.
LEMMA 3.5.
( ^ 0 ) ' = DK + A-^o*.**" •
(3.21) PROOF.
- Prom D2T{()= DxN(t)dux and (1.20), it follows immediately
(3.22)
DS*m = dN{t) +
r;kNlt)Xft)tXdu".
Differentiation of (3.22) with respect to t and use of (3.7) give (3.23)
(Z>^ 0 )'=
dN
'
+ Nlt^d^ridu"
.
Similarly, we have (3.24)
DN« = **« +
N&Xf^du*.
Substracting (3.24) from (3.23) we thus obtain (3.21). LEMMA 3.6. - The quantity Vffii)&
Vg^bU)a$=:(l—t)Vgbafi+tVgbap.
PROOF.
(3.26)
- Prom (1.3), (1.19), (3.8) it follows that (1-t) Vgbafi = Va(l -t)\X,lt
..., X,m, DXX^\.
Similarly, (3.27)
(
tV§baff= Vat\Xtl, . . . , - ? „ , Z>« Xfi\ = m
v
T
\X lt ..., X<m, DaX p-\- ^Ta^l + 2 1-^,1 » •••> -^,*-i> ^1 *> -^,*+l > •••» -^v»! -Da-Z',/»|) fc=l
'
in consequence of (3.4). By adding (3.27) to (3.26) and using (3.7) we thus obtain (1— t) VgbX0+tVgbclff=
Va\Xu)1, ...,X U)<m , DaXlt)j>\ = Vg^blmp
.
475 300
CHTJAN-CHIH HSITJNG
-
P. Lo: Congruence theorems, etc.
TIMOTHY
4. - Integral formulas. We are now in a position to derive our basic integral formulas for a pair of compact oriented hypersurfaces Mm, Mm with boundaries BMm, dMm in a Biemannian manifold Bm+1 of constant sectional curvature K so that (4-1)
Rm = •£(<*««« -
a a
u sk) •
We first use (1.20) and §2 to compute the following differential m-form: (4.2)
D[N,(ly(d1rXDN{t)X...XDNit)xDX(t)X...xDXm)] r—1
=
m—r
= DN'm • ( V XDNit)X...XDNit)
xDXmX...XDXm)
+ N[t) • [(MJ x XDN{t)X...xDN(t)
XDX{t)X...xDX(l)]
r—1
+
m—r
+ N[t) • (d.Dr XDNmX...Xl)N{t)
XBXit)X...XDXit))
r—1
+
+
m—r
r— 1
+
m—r
(r-l)N[ty(61TXD'N(t)XBN{t)X...XDN(t)XBXmX...XDX{t)). r— 2
m—r
Denote the four terms on the right side of (4.2) by (I), (II), (III), and (IV) respectively. Taking the generalized covariant differential of (3.9)((), we obtain (4.3)
D»jy(0 = [ - ( ^ 6 ( W ) < I ( ( ) , , -
bmx^DdX(t)JduxAdu3.
The second term on the right side of (4.3) becomes, in consequence of (3.8)((), (4-4)
- b^b^g^N^du'Adu",
which vanishes since with respect to a and d the coefficient of dif/\du,d is symmetric whereas du*/\dua is skew-symmetric. Thus by means of (3.10)(() and (4.1), from (4.3) it follows that (4.5)
D»2TW = - UDdbmx - BJ,(m) g^X^du'Adu6 =
2RimN(t)Xlt)jX^xX(l)/)gJ)X{t)tYduc'/\du
so that (4.6)
=
(IV) = 0.
= 0,
476 CHTJAN-CHIH HSIXJNG
- TIMOTHY P. Lo: Congruence theorems, ete.
301
Substituting (3.11)((), (3.9)(0, (1.23)(() in (4.2) and using (2.10)((), (1.19)((), (2.13)(() we can easily obtain (4.7)
(III) = ( - l p V - 1 ) ! (*x •
Naf-c.Tfd&dAn.
Similarly, by (3.9)((), (1.23)((), (2.8)(j), (2.3)(() and Lemma 3.3 we have 61r-(DNmX...XJ>NmXBZmX...XDXm)={-l)rrmlHll)r(d1-N)dA.
(4.8)
m—r
T
Since (BX(f))' = dX'^ = b^dx, differentiation of (4.8) with respect to t and use of Lemma 3.5 and (4.2) give (4.9)
(I) = {-iy-^rE[0r(d,-N)dA
+
+ ra^r^NU
ra du* X DNwx...XDNw
X DXwX'...xDXM)
r—1
•
m—r
Let the second term on the right side of (4.9) be denoted by (I)2« Then (4.10)
(I) + (II) = ( - l J - i - ^ T i f f U V J y ) dA + ( I ) , + (II).
Since B8[= rnX\tXyduv,
from (1.20), (3.11), (3.9)((), (2.10)(() we obtain
(II) = ( - i y - H m - 1 ) ! T T „ ( 0 X • N w ) a i t r l n X l m j l N ^ d A w
(4.11)
.
Now put (4.12)
d^XN^+^X^.
By taking the scalar product of (4.12) and N(t) we have X= d1-N{t). Similarly, by taking the scalar product of (4.12) and X(t)^ and using (1.5)w we can obtain
K= {81-X
Next we put (4.14)
riN'^fiNtv+f'Xt^.
By taking the scalar product of (4.14) and JV(() and using (1.21) and Lemma 3.4 we have „ _ r« V AT 20 - Annali di Matematica
1
(daik 4-
daik
—
d(lil
\
N1 Kk - - f^* Ns N" - n
477 302
-
CHUAN-CHIH HSITJNG
P. Lo: Congruence theorems, etc.
TIMOTHY
since with respect to j and 1c, N^N^ is symmetric whereas (da^ldsc? — danldxk) is skew-symmetric. Similarly, we can obtain yf so that - T ^ , = QNfo + ahj(rhMN>t)Xfthfl)g$Xi^
(4-15)
.
Thus from (4.9), (4.13) and (4.15) we have (4.16)
(I)2 = r(dt • JSTm)N(t) • {x.avI«lN*mX'mtygftXmjld«' xDNit)x...xDNit)xDXmx...xDNm) r—1
X
+
m—r
{QNmTydtfxDNm x... x DNW x DXm X...X J>ZW) = r—1
= ( I )2a+( I )26)
m—r
S a y-
Since c"^.= e^ r , as before a simple calculation gives (4.17)
(I)2a = (_ I ) ' " >
- 1 ) ! TT.(ai • Nm)
a^X^N^dA^.
From (1.21), (3. 20) it follows that aisrkl + a ^ = da^Jdx1 = 2 e % ,
(4-18)
which, together with (4.11) and (4.17), implies immediately (4.19)
(II) + ( I ) 2 a = 0 .
Finally, we can obtain (4.20)
(I) 2 b =
(-inm-l)lTT.Q{d1-Xm)i$taA{fi.
Substituting (4.6), (4.7), (4.10), (4.16), (4.19), (4.20) in (4.2) and using (3.18) we have (4-21)
fe~^D[N
{dlT
% DNM
= mr^-^H'^dA
% ... g D N m % DX{t) g . . ^ DXit))] = r—1
' '
m—r
i
+ rTaTf{d1-Nyg g$id&rdA-rTraQg~i(d1-XWfi)<$Tgiw&A
By setting i
(4-22)
Cf=g*jgrfc*ldt, o l
(4-23)
Br=-
rTT.ff-*/(*i • Xm)
.
478 CHUAN-CHIH HSIXJNG - TIMOTHY P . L o : Congruence theorems, etc.
303
integrating (4.21) with respect to t over t h e interval [0,1], and interchanging D and the integral sign due to t h e continuity of t h e integrand, we obtain I
(4 24
- )
( 1
p ^-pf-^v^i-r %-py») a... a.pjr(,) Sflx(<) %... xpxw)dt = 0 jyi
= -T{d1-N)(HT—Hr)dA+
{d1-NyrarfiGfdA+
oBrdA ,
(l
,
where Hr is t h e r - t h mean curvature of Mm. Integrating (4.24) over t h e hypersurface Mm and applying Stoke's theorem we thus arrive a t (4.25)
-
(r(d1-N)(Sr-Sr)dA Mm
Mm _
(
+ [(d1-Nyrar0CfdA
+ LBrdA
=
Mm
1
1
= (i» / ~—1)! Ci Jf [J fa
5. — Congruence theorems. THEOREM 5.1. - Let Bm+1 be an (m +1)-dimensional oriented Biemannian manifold of class Cs having constant sectional curvature and admitting a continuous infinitesimal conformal vector field f which generates a one-parameter group G of conformal transformations. Suppose that the orbits of the transformations of G cover Bm+1 simply, and particularly all the transformations of G, except the identity transformation, have no fixed points. Let Mm, Mm be two compact oriented hypersurfaces of class G3 without boundary immersed in Bm+1, and suppose that there is a 1-1 transformation TTeG between Mm and Mm such that P = Tr^g)P for each P e Mm and each P e Mm, and such that the set of points, at which the orbits of Tr are tangent to Mm or Mm, has no inner points. If both second fundamental forms of Mm, Mm are positive definite, the r-th mean curvatures Hr,Hr{l<,r<m) of Mm, Mm are equal at each pair of corresponding m m points of M , M and jgBT dA^O where g and Br are defined by (3.19) and (4.23) Mm
_
respectively, then Mm and Mm are congruent modCr. When Q — 0, this theorem is due t o H . H O P F a n d Y. KATSURADA [5].
479 304
CHTJAN-CHIH HSITJNG
-
TIMOTHY
P. Lo: Congruence theorems, etc.
- Since the second fundamental forms b^u'tiP and b^vfv? of Mm and Mm are positive definite, b^u"^ (0<«<1) is also so by Lemma 3.6. Thus 2 c wr w PROOF.
for all t and r, and therefore £ Vfu"')/ are positive definite. Since dMm = 0 and dx • N =£ 0 almost everywhere on Mm, from the assumptions of the theorem it follows immediately that the validity of (4.25) implies that on Mm we have Ta = 0 or T = constant almost everywhere, and everywhere by continuity. Hence the theorem is proved. 5.2. - Assume that two hypersurfaces Mm and Mm have nonempty boundaries dMm and dMm of dimension m — 1 and satisfy all other conditions in Theorem 5.1. If dMm and dMm are coincident, then Mm and Mm are also so. THEOREM
PROOF. - The proof of this theorem is exactly the same as that of Theorem 5.1, except that for this theorem the vanishing of the right side of the integral formula (4.25) is due to the following reason: Since the boundaries dMm, dMm are nonempty and coincident, 1 = 1 on the boundary. Prom (3.4) it thus follows that r = 0 on 3Mm.
BIBLIOGRAPHY
[1] A. A E P P L I , Einige Ahnlichkeits- und Symmetrie-satze fiir differenzierbare Flaehen im JRaum, Comment. Math. Helv., 33 (1959), p p . 174-195. [2] L. P . EISBNHABT, Riemannian geometry, Princeton University Press, Princeton, 1949. [3] H. H O P F - K. Voss, Sin Satz aus der Flachentheorie im Orossen, Arch. Math., 3 (1952), pp. 187-192. [4] H. H O P F - Y. KUTSUBADA, Some congruence theorems for closed hypersurfaces in Hiematm spaces. - I I : Method based on a maximum principle, Comment. Math. Helv., 4 3 (1968), pp. 217-223. [5] H . H O P F - Y. KATSUBADA, Some congruence theorems for closed hypersurfaces in Biemann spaces. - I l l : Method based on Voss' proof, Comment. Math. Helv., 46 (1971), p p . 478-486. [6] C. C. HSITJNG, Some global theorems on hypersurfaces, Canad. J . Math., 9 (1957), p p . 5-14. [7] C. S. Hsu, Characterization of some elementary transformations, Proc. Amer. Math. S o c , 10 (1959), p p . 324-328. [8] Y. KATSUBADA, Some congruence theorems for closed hypersufaces in Biemann spaces. I : Method based on Stokes' theorem, Comment. Math. Helv., 43 (1968), p p . 176-194. [9] R. E . STONG, Some global properties of hypersurfaces, Proc. Amer. Math. S o c , 11 (1960), pp. 126-131. [10] A. W. TUCKEB, On generalized comriant differentiation, Ann. of Math., 32 (1931), pp. 451-460. [11] K. Voss, Einige differentialgeometrische Kongruenzsdtze fiir geschlossene Flaehen und Hyperfldchen, Math. Ann., 131 (1956), p p . 180-218. [12] C. E . WEATHEBBUEN, An introduction to Biemannian geometry and the tensor calculus, Cambridge University Press, Cambridge, 1950.
Can. J. Math., Vol. XXVIII, No. 1, 1976, pp. 63-72
ISOMETRY OF RIEMANNIAN MANIFOLDS TO SPHERES, II NEILL H. ACKERMAN AND C. C. HSIUNG
1. Introduction. Let Mn be a Riemannian manifold of dimension n ^ 2 and class C3, (gtj) the symmetric matrix of the positive definite metric of Mn, and (gi}) the inverse matrix of (gtj), and denote by V*, i?»<#> Rtj = Rktjit a n d R = gijRtj the operator of covariant differentiation with respect to gtj, the Riemann tensor, the Ricci tensor and the scalar curvature of Mn respectively. Let d be the operator of exterior differentiation, 8 the operator of codifferentiation, and A = dS + 8d the Laplace-Beltrami operator. Throughout the paper all indices take the values 1, . . . , n unless stated otherwise and can be raised and lowered by using gij and gtj respectively, and repeated indices indicate summation. Let » b e a vector field defining an infinitesimal conformal transformation of Mn, and Lv the Lie derivative with respect to v. Then we have (1.1)
Lvgfj = Vflij + V > 4 = 2pgij.
The infinitesimal transformation v is said to be homothetic or an infinitesimal isometry according as the scalar function p is constant or zero. We also denote by Ldp the Lie derivative with respect to the vector field p' defined by (1.2)
p< = gVpj,
=
Pi
VjP.
Let £/(j,) and TJ up) be two tensor fields of the same order p ^ n on a compact orientable manifold Mn, where I(P) denotes an ordered subset {t'i, . . . , iv\ of the set {1, . . . , n\ of positive integers less than or equal to n. Then the local and global scalar products (£, ?j) and (£, r;) of the tensor fields £ and i\ are defined by
(1.3)
<$,i»> = ^ £ / { p W
(1.4)
(*, v)= f
<«, v)dV,
where dV is the element of volume of the manifold Mn at a point. We also define (1.5)
| | { | | =/>•<{,*>.
Received November 22, 1974 and in revised form, February 18, 1975. The research of the second author was partially supported by the National Science Foundation grant GP-43605. 63
N. H. ACKERMAN AND C. C. HSIUNG
64
From (1.3) and (1.4) it follows that (£, £) is nonnegative, and that (£, £) = 0 implies J = 0 on the whole manifold M". In the last decade or so many authors have studied the conditions for a Riemannian manifold Mn of dimension n > 2 to be either conformal or isometric to an ra-sphere. Very recently K. Amur and V. S. Hedge [2] weakened one of the two conditions LVR = 0 and LdpR — 0 studied jointly by Hsiung and Stern [6], and Yano and Hiramatu [11] removed the condition from some of these results of Hsiung and Stern and some other known results. The purpose of this paper is to continue the joint work of Yano and Hiramatu to obtain the following theorems by removing both conditions LVR = 0 and LdpR = 0 from the joint results of Hsiung with Stern [6] and Ackler [1]. In the following Theorems 1 and 2, Mn will denote a compact Riemannian manifold of dimension n > 2 with metric gtj, which admits an infinitesimal nonisometric conformal transformation v satisfying (1.1) with p 9* 0. 1. An oriented manifold Mn is isometric to an n-sphere if it satisfies one of the following three equivalent conditions: THEOREM
( P + ^ [nRptp* - (LVR + nRP)AP], l j ^ 0, (1.6)
\P~l
p(nLdpR + ALVR), l ) ^ 0,
( P + ^ [L„ Ldp]R, l ) ^ 0, where D
(1.7)
7- r 2 . , c — 4a 2
P = PL{a A + -^^B \_LV, L,dp\ — L^L,^
1 / 2o2
- - {^^
. c — 4a 2 \ _2~|
+ - ^ Y J * J,
L,dpLv,
A and B are defined by (1.8)
A = RhijhRnm,
B =
R»RU,
and a, c are constants such that c = 4a 2 + ( » - 2 ) (1.9)
L
2a £ bt+ ( £ ( - 1 ) *"»&,) * Vi 1 7 '- 1 " - 2(&i&, + &2&4 - &6&6) + ( „ _ ! )
£ bA > 0,
b's being arbitrary constants. An elementary calculation shows that c ^ 0 where equality holds if and only if bi = . . . = bit bb = b6 = 0, a = — (n — 2)bx. For LVR = 0, Theorem i (referred to the first inequality of (1.6) for P = 0 and (nRptp1 — (LVR + nRp)Ap, 1) ^ 0) with "isometric" replaced by "con-
RIEMANNIAN MANIFOLDS
65
formal" is due to Yano [10] for either o ^ 0 , c - 4a 2 = 0 or a = 0, c — 4a 2 ^ 0, and due to Hsiung and Stern [6] for general a and b's. Theorem 1 (referred to the first inequality of (1.6) for P = 0 and {nRptp1 - (LVR + nRp)Ap, 1) ^ 0) is due to Yano and Hiramatu [11] for a ^ 0, c — 4a2 = 0 or a = 0, c — 4a2 -^ 0. For constant R, Theorem 1 (referred to the second inequality of (1.6) for P = 0) is due to Lichnerowicz [8] for a = 0, c 9^ 0, B = constant, due to Hsiung [3] for a ^ 0, c — 4a 2 = 0, A = constant, due to Yano [9] for either a = 0, c ?± 0, or a ^ 0, c — 4a2 = 0, due to Hsiung [5] for bi = . . . = b6 = 0, due to Yano and Sawaki [13] for b\ = . . . = &4 = b/(n — 2), 65 = &6 = 0, and due to Hsiung [5] for general a and b's. For LVR = 0, Ld„R = 0, Theorem 1 (referred to the second inequality of (1.6) for P = 0) is due to Ackler and Hsiung [1]. THEOREM
2. A manifold Mn is isometric to an n-sphere if it satisfies
(1.10)
Lv(A"Bb)
n(l.ii) m
+ 1) .-(2a +i ( " - l ) Mj - 2aa(a-+ b)R^ "- , cy _ 1 ) 0 -i A B n +b l(n
= 0,
(1-12)
/
7?2
\
\
where A, B are given by (1.8), and a, b are nonnegative integers and not both zero. For constant R and AaB", Theorem 2 is due to Lichnerowicz [7] for a = 0, b = 1, and due to Hsiung [3] for general a and 6. For constant A"Bb and L„i? = 0, LdpR = 0, Theorem 2 is due to Hsiung and Stern [6]; in this case condition (1.12) is satisfied automatically since (1.13)
(i?-V 4 V,p + ^
L
, P ) fc 0,
which is due to Hsiung and Stern [6], and due to Lichnerowicz [8] for constant R. In the proofs of the above theorems we need the following theorems. A (Yano and Nagano [12]). If a complete Einstein space Mn of dimension n > 2 admits an infinitesimal nonisometric conformal transformation, then M" is isometric to an n-sphere. THEOREM
THEOREM B (Tashiro [8]). If a complete Riemannian manifold M" of dimension n > 2 admits a complete vector field v satisfying (1.1) with p ^ const, and
(1.14)
V(VjP
=
-gijAp/n,
then Mn is isometric to an n-sphere.
N. H. ACKERMAN AND C. C. HSIUNG
66
2. Notation and formulas. In this section we shall list some well known formulas which will be needed in the proofs to follow. Let v be a vector field defining an infinitesimal conformal transformation on a Riemannian manifold M" of dimension n ^ 2 so that (1.1) holds. Then we have (2.1)
p=
(2.2)
LvR\jk
Vfl*/n, = -ekhViPj
+ e/V«p* - gtjV.p" + gtkVjPh,
where p" = Vp> and tkh = 1 for h = k ande*/1 = 0 for h ^ k. From (1.1) and (2.2) it follows immediately that (2.3)
LvRnijk
= 2pRhijk — ghkViPj + gnjV{pk — gijVhpk +
(2.4)
LvRtj
(2.5)
LVR = 2(n - l)Ap - 2i?p.
gtkVhpj,
= g(jAp - ( n - 2) VtPj,
For any scalar field / on Mn, we have (2.6) A/ = - V ' V , / . On the manifold Afn consider the following tensors: (2.7)
Tti =
Rtj-±Rgu,
(2.8)
FAJJ-J; = J? ft< j fc — —7
TT^(gMgi3
—
ghjgik)i
n(n — l ) (2.9)
Whim = aTh(jk + bigmTtj — bzghjTik + bzg(jThk ~ bigikThj
+ bbghiTjk
—
bsgjkThti
where a and b's are constants satisfying (1.9). From (2.7) and (2.8) it follows immediately that (2.10)
gl'Tu
= 0, ghkThtik = Tti,
which, together with (2.9), imply that (2.11)
g'Y'W^
= 0,
ghlgikWhtjk
= 0,
ghiglkWhijk
= 0.
Moreover by (1.3), (1.5) and (2.9) we have \\W\\=a*A+C--^B-±{^ + C - ^ - ) R \ n —2 n \n — 1 n—2 / where A, B, and c are defined by (1.8) and (1.9). (2.12)
3. L e m m a s . Throughout this section Mn will always denote a compact oriented Riemannian manifold of dimension n > 2.
RIEMANNIAN MANIFOLDS
67
LEMMA 3.1 (Yano [5, (2.11), (2.12); or 11, Lemma 4]). If p is a scalar field on Mn, then (3.1)
\RtjpV
- * - = ^ (A P ) 2 , l ) + 2\viPj
+ lgiAP,
ViPt + -giAp)
= 0,
l ) + 2[VfPj + ^gijAp, ViPj + \giAp)
= 0.
or equivalently (3.2)
[RijpV
-
TL
~piVtAP,
For a proof of Lemma 3.1 one may also see [1, p. 58]. LEMMA 3.2 (Yano [10]). / / Mn of dimension n > 2 admits an infinitesimal nonhomothetic conformal transformation v satisfying (1.1) with p ?^ const, and either one of the following two conditions: (3.3)
(R&V
- 5 - = ^ (Ap)2, l ) ^ 0,
(3.4)
[R^PV
- ^—^pV.Ap,
l ) ^ 0,
then Mn is isometric to an n-sphere. Proof. This follows from Lemma 3.1 and Theorem B. Substitution of (2.5) in (3.3), (3.4) and use of (3.5)
LdpR =
p'ViR
and Lemma 3.2 yield LEMMA 3.3. / / Mn of dimension n > 2 admits an infinitesimal nonhomothetic conformal transformation v satisfying (1.1) with p s* const, and either one of the following two conditions: (Loi?
- ^~TY)
+
2Rp)2
'
l
(3.6)
[RVPV
)
= °-
(3 ,7)
( * „ P V - \ Rptp* - \ PLdpR - ^ p'ViUR,
l ) £ 0,
then Mn is isometric to an n-sphere. LtR
Condition (3.7) is due to Yano and Hiramatu [11]. In particular, when = 0 and LdpR = 0, conditions (3.3) and (3.4) are reduced to
(3.8)
(3.9)
^ - ^ l ) ^
1
) ^
0
'
(r„, PiPj) ^ 0,
so that in this special case Lemma 3.3 is due to Ackler and Hsiung [1].
68
N. H. ACKERMAN AND C. C. HSIUNG
LEMMA 3.4. If Mn admits an infinitesimal nonisometric conformal transformation v satisfying (1.1) with p ^ O , then for any scalar field f on Mn (3.10)
(LJ, 1) = - » ( / P , 1).
Proof. From the definition of Lv and (2.1) we have (3.11)
V«(/V) =
Lvf+nfp.
Integration of (3.11) over M" and use of the well-known Green's formula (3.12)
(V%, 1) = 0,
where £4 is any vector field on Mn, give (3.10) immediately. LEMMA 3.5. For any scalar fields f and h on Mn, (3.13)
(Ldfh, 1) = (Ldhf, 1) = (VtfV%
1) = (fAh, 1) = (hAf, 1).
Proo/. (3.13) follows from ( V , ( / V ' A ) , 1) = (VifV%
1) - (/AA, 1) = 0,
(V,(*V«/), 1) = ( V i A V / , 1) - (hAf, 1) = 0. LEMMA 3.6. If Mn admits an infinitesimal conformal transformation v satisfying (1.1), then (3.14)
(p*AR, 1) = (2PLdpR, 1).
Proof. (3.14) follows from (3.13) and (3.5) by p u t t i n g / = R and h = p2 in (3.13). LEMMA 3.7. 7/ Af" admits an infinitesimal conformal transformation v satisfying (1.1), then (3.15)
( L ^ p i ? , 1) = - | (P2Ai?, 1),
(3.16)
(LdpLvR, 1) = «LVR)AP, 1),
(3.17)
([L„ L , , ^ , 1) = - | (p2Ai?, 1) - ((LVR)AP, 1).
Proof. By Lemmas 3.4 and 3.6 we have (L^R,
1) = - «( P L dp i?, 1) = - | (p2Ai?, 1),
which proves (3.15). By putting / = p and h = LVR in (3.13) we readily obtain (3.16), and (3.17) follows from (3.15) and (3.16).
RIEMANNIAN MANIFOLDS
69
LEMMA 3.8. For any scalar field p on M", (3.18)
| (p2AR, 1) - (i?PAp, 1) + {RPiP\ 1) = 0,
(3.19)
i (LJ.dpR, 1) + (2?pAPl 1) - ( W , 1) = 0.
Proof. Integration of (3.20)
ViiRpp*)
= pp'ViR
+ RpiP* - RpAp
over M" and use of (3.12), (3.14) and (3.5) give (3.18). (3.19) follows from (3.15) and (3.18). LEMMA 3.9. If M" admits an infinitesimal conformal transformation v satisfying (1.1), then (3.21) sf = - 38 = ([Lv,Lip]R,
1),
where (3.22) s/ = (nRpip* - (L,R + nRp)Ap, 1), 3 = (nLipR + ALVR, p). Proof. This proof is due to H. Hiramatu. From (3.19) we have stf = (LvLipR — (LvR)Ap, 1) which together with (3.16) gives immediately stf = ([L„ Ldp]R, 1). On the other hand, by p u t t i n g / = LipR in Lemma 3.4 a n d / = p, h = L„i? in Lemma 3.5 we obtain 31 — — ([Z,„, Ldp]R, 1). 4. Proof of t h e theorems. Proof of Theorem 1. By means of (2.9), (2.8), (2.7), (1.1), (2.3), (2.4) and (2.5) we can easily obtain LvWhijk = 2apRhijk - [a + (n - 2)&i]gMV4p, + [a + (w - 2)&2]gwViP* - [a + (n - 2)bl]gijVhPk + [a + (n - 2)bi]gikVhp - (n - 2)b6ghiVjPk + (n (4.1)
2)begjkVhPt
, 1 ^{2a[pR+ (»-l)Ap]+ (n-l)[2PR+(n-2)Ap]\ n(n — l ) • [-gijghk{bi + b„) + gikghjQ>2 + bi)}
+
+ -gMg»\2pR + (» - 2)Ap](-& 6 + bs) + n
2P(blgMRi}
— b2ghjRik + bigijRhic — b4gikghj + br,ghtRjk — b6gjkRhi). Multiplying both sides of (4.1) by Whilk and making use of (2.7),. . . , (2.11), (1.9) and Riilk = 0 we have, by an elementary but lengthy calculation, (4.2)
W*"*L,Whtit = 2p\\W\\ -
cTVVtPj.
70
N. H. ACKERMAN AND C. C. HSIUNG
Substitution of (4.2) in the well known formula (4.3)
L,\\W\\ = 2Wt»L.Whtit
- 8P\\W\\
thus gives (4.4)
2cpTi*VtPj.
pL.\\W\\ = -±p*\\W\\ -
Integrating (4.4) over Mn we obtain (4.5)
-2c(PT^VtP},
1) = (L,\\W\\, p) + 4 ( | | ^ | | , p«).
The equivalence of the three conditions given by (1.6) is obvious from Lemma 3.9. For proving Theorem 1 we assume that the second inequality of (1.6) holds. Applying covariant differentiation and using (2.6), (3.6), (4.5) we obtain V 4 ( i W - \RpPi - ~P2Vii? (4.6)
= 2?„pV - \ W
-±pVtLvR)
- \ PLdpR - i p'V^R + pT1^^,
+
in [(n ~ *)LdpR + 2pAR + AL'Rl
On the other hand, integrating (4.6) over M", applying Green's formula (3.12) and substituting (4.5), (3.14) in the resulting equation we have \RiiPlp' — ~RPiP* — ~ pLi9R — r-p'ViLti?, 11 (4.7) = ~ {LV\\W\\, p) + | (\\W\\, p2) - ^ {nLdpR + ALVR, p). Since (||W||, p2) is nonnegative, from (4.7), (2.12), (1.7) and the second inequality of (1.6) we obtain (3.7). Hence by Lemma 3.3, Mn is isometric to an w-sphere. A proof of Theorem 1 based on the first condition of (1.6) can be obtained by following the proof of Theorem I in [6]. In fact, substituting (3.20) for pp*V4i? and (4.4) for pTijV iPj in [6, (4.6)] and using (3.12), (2.5) and the first condition of (1.6) we can easily reach (3.8). Proof of Theorem 2. Without loss of generality we may assume our manifold Mn to be oriented, as otherwise we need only to take an orientable two-fold covering space of Mn. On Mn consider the covariant tensor field T of order 2(2a + b): a t^a\
=
11
6 Rhririrkr
11
Ru.v. Ra+t
n
a+0
a l l
(n - l ) a i
iShrkr&irjr ~
ghrjrgirkr)
11
£««!>.•
488
RIEMANNIAN MANIFOLDS
71
From (4.8) it is easily seen that the length of T is (4.9)
[2(2a + 6)]! <7\ T) = AaB» - ^ *
_
1)B,
which, together with (1.10), implies (4.10)
[2(2a + &)]! LV{T, T) = n a
+ t
^
1 } «L v R
2ia+b)
.
Thus by the extension of formula (4.3) to the tensor T we immediately obtain (4.11)
L,(T, T) = 2(LVT, T) - 4(2a + b)P(T, T),
from which and (4.10) it follows that 2(2a + b)P{T, T) -
n«+»{n_^[2{2a
.P) •
+ h)]l
On the other hand from (2.3) and (2.4) we have LvThltjlkl
_ _ ,haijaka-u1v1..
ubvb
a
&
= 2ap | | Rhrirlrkr
| |
r=l
RU3V,
s-1
a
~
A^l [Rllliljlkl
•• •
Rhr-lir-lir-lkr-liShrkr^ir^jrP
T=\
— ghrjr^U^krP
+
girir^kr^hrP
~
girkr^jr^
krp)
& ' Rhr+llr+ljr+lkr+1
(4.13)
• • • -K-haialaka'
1 1
-^UsVs
s=l a i l l
6 -Kflrirjrkr
r-1
'
'
X^UWX
• • •
^Ue-lVe-l
s=l
• [gusvAP — in — 2)VUsV„sp]i?„!+1„s+1 . . . Rubvb] 2 (2a + b)Ra+bP + L„R"+6 A
w a+ V - l) a
U ^*'^»'*' —
6
girkrghrjr)
1 1 S-1
£ « « Cs •
By means of (4.8), (4.9), (4.13) and (1.8) an elementary calculation yields [2(2a + b)}\ {L„T, T) (4 14)
= 2a
P^2a
+ W
^ ^ 0a+l7?2
(a + b) -^f
r r ^ AP +
bAaB°-LRAp,
72
N. H. ACKERMAN AND C. C. HSIUNG
from which and (1.11) it follows readily that {{LVT,T),P)
=
2a((T,T),P2)
AaB" (4a (n - 2)b\ ( it ~ [2(2o + b)]l \ ~A + —BT~) \ R V'V*P
(4.15)
+
R ) n Ap' "/•
Substituting (4.12) in (4.15) and using (1.11), (2.5), (1.12) we can easily show that
(4.16)
« r , r), P 2 ) ^ o .
This means that (T, T) = 0 which implies (ft.1.1
)
=
1 h1iij1k1...Kaia}akaU1v1..-Uht>b
"•
Multiplying (4.17) by r-2
s-1
and using (4.8) we obtain Rtljl = Rgt jjn which implies that Mn is an Einstein space. Hence by Theorem A, M" is isometric to an ra-sphere. REFERENCES 1. L. L. Adder and C. C. Hsiung, Isometry of Rtemannian manifolds to spheres, Ann. Mat. Pura Appl. 99 (1974), 53-64. 2. K. Amur and V. S. Hedge, Conformality of Riemannian manifolds to spheres, J. Differential Geometry 9 (1974), 571-576. 3. C. C. Hsiung, On the group of conformal transformations of a compact Riemannian manifold, Proc. Nat. Acad. Sci. U.S.A. 54 (1965), 1509-1513. 4. On the group of conformal transformations of a compact Riemannian manifold. II, Duke Math. J. 34 (1967), 337-341. 5. On the group of conformal transformations of a compact Riemannian manifold. Ill, J. Differential Geometry 2 (1968), 185-190. 6. C. C. Hsiung and L. W. Stern, Conformality and isometry of Riemannian manifolds to spheres, Trans. Amer. Math. Soc. 161 (1972), 65-73. 7. A. Lichnerowicz, Sur les transformations conformes d'une variete riemannienne compacte, C.R. Acad. Sci. Paris 259 (1964), 697-700. 8. Y. Tashiro, Complete Riemannian manifolds and some vector fields, Trans. Amer. Math. Soc. 117 (1965) 251-275. 9. K. Yano, On Riemannian manifolds with constant scalar curvature admitting a conformal transformation group, Proc. Nat. Sci. U.S.A. 55 (1966), 472-476. 10. On Riemannian manifolds admitting an infinitesimal conformal transformation, Math. Z. 113 (1970), 205-214. 11. K. Yano and H. Hiramatu, Riemannian manifolds admitting an infinitesimal conformal transformation, J. Differential Geometry 10 (1975), 23-38. 12. K. Yano and T. Nagano, Einstein spaces admitting a one-parameter group of conformal transformations, Ann. of Math. 69 (1959), 451-461. 13. K. Yano and S. Sawaki, Riemannian manifolds admitting a conformal transformation group, J. Differential Geometry 2 (1968), 161-184.
Lehigh University, Bethlehem, Pennsylvania
490
Sonderabdruck aus ARCHIV DER MATHEMATIK Vol. XXVII, 1976
BIRKHAUSER VERLAG, BASEL UND STUTTGART
FaSC.
5
A R e m a r k on Cobordism of Manifolds with Boundary By CHTJAN-CHIH HSITTNG *)
1. Introduction. The following theorem of Thorn [4] is well-known: Theorem 1.1. The Pontrjagin numbers boundary are cobordism invariants.
of a compact oriented manifold
without
The purpose of this note is t o study cobordism of manifolds with boundary b y extending the above theorem of Thorn to the boundary case. All manifolds considered are t o be of class C°°, and the dimension of a manifold will always be denoted by a superscript such as n in Mn, a manifold of dimension n. Let M\, M\ be compact oriented manifolds with reflecting boundaries - B " - 1 , - B " - 1 respectively, so t h a t there are orientation-reversing involutions j t i , m on -B" _ 1 , B%_1 respectively; this definition was originally given b y the author [2]. For convenience we shall always denote a reflecting boundary with its involution together such as (-BJ - 1 , 7ti). Two pairs {M\, B^'1) and (M^^Z'1) are said to be cobordant or to belong t o t h e same cobordism class if there is a compact oriented manifold Mn+1 with boundary 5 » such t h a t J3» = J / ? U Z)» U (—-M»), where D» has B^1 U ( B^1) _1 1 as its boundary and is e m p t y when i?™ a n d B^ both are, and such t h a t there is an orientation-reversing n on Dn such t h a t jr | 2 ? ? - 1 = zrj, i = 1, 2 ; for this definition cf. [3]. This " c o b o r d a n t " relation is an equivalence relation since it is clear t h a t the relation is reflexive and symmetric; t h a t it is transitive can be seen from the following obvious construction: If ilf? 2 +1 has boundary M\ U D%2 U {—Ml), and M^1 has boundary M\ U D%3 U ( — M%) with other conditions satisfied, then JfJ 2 + 1 a n d i f f f 1 are identified along t h e common boundary M\, and the resulting structure can be smoothed out to give a compact oriented C°° manifold whose boundary is M" U Z>"2 U U D23 U (— -M3). The equivalence cobordism classes of compact oriented w-manifolds with reflecting boundary form an abelian group, which is called the cobordism group in dimension n, if we use the operation induced b y taking the disjoint union, with inverse given by reversing orientation. I t is obvious t h a t for compact oriented manifolds without boundary our cobordism classes and their groups are the ones defined by Thorn. Now we are in a position t o state our theorem: *) Eesearch partially supported by the National Science Foundation grant GP-43665.
491
552
C.-C. HSIUNG
ARCH. MATH.
Theorem 1.2. / / two compact oriented manifolds with reflecting boundaries are cobordant, then they have equal Pontrjagin numbers. 2. Curvature forms; Pontrjagin classes and numbers. Let M be a Riemannian manifold of dimension n, and Vx, V* respectively the spaces of tangent vectors and covectors at a point x of M. By taking an orthonormal basis in Vx and its dual basis in V*, over a neighborhood U of x on M, we then have a family of orthonormal frames and linear differential forms coi, ..., con such that <e$, coy) = dy ( = 1 for i=j, and = 0 otherwise), and the Riemannian metric is of the form
ds*=X coh Throughout this section all Latin indices take the values l,,..,n unless stated otherwise. The structure equations of the Riemannian metric of M are du>i = 2 W ; A mH >
w
y + on = 0 >
i
da>ij = ^ one A co/cj + Qij,
£2ij + Qji = 0,
k
where the wedge A denotes the exterior multiplication. The forms Qy depend only on the Riemannian metric of M and are called the curvature 2-forms of the metric. Throughout this section for indices we shall use I(p) to indicate the ordered set of p integers ii, ..., ip among 1, ..., n. When more then one set of indices is needed at one time, we shall use other capital letters such as J, H,... in addition to I. Now for an even p^n, we define the following p-ioim: @nv) = ^ y <3j(£) Qhh
A -- A Q
"
h-ih '
where <5/^ is + 1 (respectively — 1), if the integers h, ...,ip are distinct and J(p) is an even (respectively odd) permutation of I(p); it is zero in all other cases. Clearly, @i} = Qij. Finally, the forms &nP) as well as Qy depend only on the Riemannian metric of M. A theorem of Chern [1] can now be stated in the following form: Theorem 2.1. The differential form ,
2 n
[(2i)!] 2
r
R
R
defines the j-th Pontrjagin class pj of a compact oriented Riemannian manifold without boundary in the sense of de Eham's theorem, where 2 denotes the summation over all (i)
the different combinations of i\, ..., i^j among 1, ..., n. From now on we shall identify the Pontrjagin class pj with the form (2.1) which depends only on the Riemannian metric of the manifold. For a compact oriented manifold M of dimension 4k without boundary and for each set of j \ , ..., jr such
492 Vol. XXVII, 1976
Cobordism of Manifolds
553
that j \ + • • • + jr = k we have an integer (2-2)
Ph
,AM) =
hh-PirM
Thus we have n (k) such integers if n (k) is the number of distinct partitions of k. All these integers are called the Pontrjagin numbers of M. 3. Doubling of a manifold. Let M be a compact oriented G°° manifold of dimension n with boundary B of dimensionTO—1. Let M with boundary S be a C°° homeomorphic copy of (M, B) with the same orientation, and /n be the homeomorphism so that fx (M, B) = (M, B). Then we can define the double of M with a reflecting boundary (B, n) to be a C°° closed oriented manifold N such that N~MuM and that S = n(B) by identifying fj,n{x) e S with x for all x e B; this definition was given by the author in [2]. Thus on the double N we can define a homeomorphism v : 2V-> N by V
^
=
i/u(x), \/u-Hx),
for for
xeM, xeM.
To see that this is well-defined, at first we notice that v(x) = JU (x) e B for every x e B. Since x e B is, identified with /j,n(x) e B, v{fxji{x)) = fjr1(fin{x)) = n(x) e B, and therefore /u (x) must be identified with n (x); this is indeed true by the definition of our identification and the assumption jfi = 1. Clearly, v is an involution. (It should be noted that the definition of doubling a manifold M here is somewhat different from the ordinary one under which M and M are of opposite orientations so that every point of B is a fixed point under the involution v.) Alternatively, we may regard doubling the manifold M with a reflecting boundary (B, n) as finding a C°° homeomorphism
where N is a C°° closed manifold with an involution
such that vh = hn: B->N. Assume h maps M into a fundamental domain of the involution v in such a way that B is mapped onto itself. We shall identify M with the fundamental domain henceforth, so that we may regard the double N as composed of two halves M and M with the same orientation such that M is mapped onto M by v, MC\M = B, and v\B = n. A Riemannian metric on the double N, with respect to which the involution v is an isometry, is said to be symmetric, and the restriction of a symmetric metric on N to M is also said to be symmetric on M. To derive a C°° symmetric Riemannian metric on N from a C°° Riemannian metric on M, at first we are naturally tempted to prolong to N a differentiable metric g on some manifold containing M by setting (3.1)
g(x)=g(v(x))
for xeM. Although (3.1) is well defined, the difficulty is that the resulting metric will, in general, not be differentiable across B. However, on the other hand, for a
493 554
C.-C. HSIUNG
ARCH. MATH.
given C°° Riemannian metric g everywhere defined on N, we may obtain from it a C°° symmetric metric g by setting g(x) = \[?{x)
+ #(v (a;))],
for
xeN.
Hence on the double N there always exists a C°° symmetric Riemannian metric. It should be remarked that if the manifold M has empty boundary B, then the double N of M consists of two homeomorphic copies M and M, and therefore every (7°° Riemannian metric of M is symmetric. 4. Proof oi theorem 1.2. Let M\k, JJ/|* be compact oriented manifolds with reflecting boundaries B\k~x, -Bf*-1 respectively, so that there are orientation-reversing involutions m, 7C2 on 5 j * _ 1 , - B | i _ 1 respectively. Suppose that {M\k, Bf*-1) and (if|*, -B|* _I ) are cobordant. Then there is a compact oriented manifold Mik+1 with boundary Bik such that Bu
M\k U D™ u ( - Mf),
=
where Dik has B4k~l U (— -B|*_1) as its boundary, and is empty when B\*_1, £ ^ _ 1 both are, and such that there is an orientation-reversing involution n on D4fc such that 7i\Blk-1
(4.1)
7i\B^-1
= m,
ik+1
= n2.
4fc
Let M with boundary 5 be a C°° homeomorphic copy of (Mik+1, Bik) with the same orientation, and fi be the homeomorphism so that fi(M4k+1, Bik) = (Mik+1, Sik). Then we can have a compact oriented manifold Mik+1 U MAk+l with boundary Nfk \J (-N*k) such that D*k = nD*k by identifying fin{x) e B*k with x for all a; e D « . From (4.1) it follows that Nf = Jf|* U M$h (i = 1, 2) is closed since - S p - 1 = m B^-1 due to the identification of pnt (x) e .Bf*-1 with a; for all x e Bf*-1, and therefore that Ntk is the double of Mfk with reflecting boundary Bf1"1. Thus the closed manifolds N\k and IV|* a r e cobordant and hence have the same Pontrjagin numbers by Theorem 1.1 of Thorn. On the other hand, we consider a C°° symmetric Riemannian metric
jrW.gt)
= hhiffi)-p*M)>
i = i.2,
where ji -f • • • -f- jr — k, and $(
J>j; (gt) • • • pi (gt) = / J + J Wfor,)• • • ^ (gr4) = 2 / ^ (9i) • • • p)r(9i) ,
due to the fact that plh (g{) • • • p)r (gt) depends only on the Riemannian metric gt, and the metric
\p)x(9l) -pi(an) N{k
= j>?(g z ) N**
-pi(g2).
494 Vol. XXVII, 1976
Cobordism of Manifolds
555
From (4.2), (4.3) it thus follows t h a t for each partition j i , . . . , j r of k
(4.4)
J ^ f a ) •••pi(gx) = jp% (g2)
-pi(g2).
Since the Pontrjagin numbers of Mjk are independent of the metrics gt, and t h e Pontrjagin classes of the boundary Bf-1 of M\k vanish, (4.4) implies t h a t Mik have the same Pontrjagin numbers, and hence Theorem 1.2 is proved. References [1] S. S. CHEEN, On curvature and characteristic classes of a Riemannian manifold. Abh. Math. Sem. Univ. Hamburg 20, 117-126 (1956). [2] C. C. HSIUNG, The signature and G-signature of manifolds with boundary. J. Differential Geometry 6, 595-598 (1972). [3] R. E. STONG, Manifolds with reflecting boundary. J. Differential Geometry 9, 465—474 (1974). [4] R. THOM, Quelques propriety globales des varietes differentiables. Comment. Math. Helv. 28, 17-86 (1954). Eingegangen am 2. 6. 1975 Anschrift des Autors: Chuan-Chih Hsiung Department of Mathematics Lehigh University Bethlehem, Pennsylvania 18015, USA
J . DIFFERENTIAL GEOMETRY 12 (1977) 133-151
INTEGRAL FORMULAS FOR CLOSED SUBMANIFOLDS OF A RIEMANNIAN MANIFOLD C. C. HSIUNG, JONG DIING LIU & SLTANSU S. MITTRA
Dedicated to Professor Buchin Su on his 75th birthday 1.
Introduction
In 1903, H. Minkowski [11] obtained the following two integral formulas for a closed convex surface S in a Euclidean 3-space E3: (1.1)
f (1 + PH)dV = 0 , Js
{ (H + PK)dV = 0 , Js
where H and K are respectively the mean curvature and the Gaussian curvature of S at a point P whose position vector with respect to the origin 0 of E3 is x, dV is the area element of S at P, and p is the scalar product <*, e) of x and the unit normal vector e of S at P. In 1954 C. C. Hsiung [5] extended formulas (1.1) to a closed oriented hypersurface Mm in a Euclidean (m + l)-space Em+1 {m > 2) and obtained characterizations of hyperspheres in Em+1. In 1956 C. C. Hsiung [6] and in 1959 G. F. Feeman and C. C. Hsiung [3] extended Hsiung's integral formulas to the case in which £ m + 1 is a Riemannian space Nm+1 of constant sectional curvature, and obtained characterizations of umbilical hypersurfaces in Nm+1. In 1962, Y. Katsurada [7] extended the aforesaid results to a closed oriented hypersurface in Nm+l by introducting an infinitesimal conformal vector field £ to replace the position vector field x. In 1968 and 1969, Y. Katsurada, H. Kojyo and T. Nagai [8], [9], [10] obtained integral formulas for a closed oriented submanifold Mm of dimension m ( > 2) in a Riemannian n-manifold Nn (n > m) of constant sectional curvature with respect to an infinitesimal conformal vector field f and a special unit normal vector field e of Mm, and conditions for Mm to be umbilical with respect to e. In 1971 B. Y. Chen and K. Yano [1] studied the case in which the field e is more general but Nn is Euclidean and f is the position vector field x. The purpose of the present paper is to extend the results of Chen and Yano to the general case in which Nn is Riemannian and f is an infinitesimal conformal vector field so that all known results are special cases of ours. Communicated April 19, 1972, and, in revised form, May 28, 1976. The work of the second author was done during his visit to Lehigh University and partially supported by the National Science Council of the Republic of China.
134
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
In § 2 we first define the vector product of two tangent vectors of a Riemannian n-manifold Nn at a point P, and then discuss orthonormal frames Peteu ••• ein on JV» at P. § 3 contains the fundamental definitions and formulas for a submanifold Mm of dimension m ( > 2) immersed in Nn (« > m). In particular, some formulas are reduced to simpler forms when Nn is of constant sectional curvature. Suppose that Nn admits a continuous infinitesimal conformal vector field £, and let e be a unit normal vector field over M * parallel in the normal bundle of Mm. In § 4 we derive integral formulas for a closed oriented Mm in JV" with respect to f and e, and in § 5 we obtain various conditions for MTO to be umbilical with respect to e. We wish to thank Y. Katsurada for her discussion with one of us about some computation involving the infinitesimal conformal vector field £. 2.
Vector product and orthonormal frames
Throughout this paper unless stated otherwise the ranges of indices are given as follows: 1 < i,}, k, • • • < m , (2.1)
1 <
a, /3, y, • • • <
n ,
m + 1 < A, B, C, • • • < n ,
(m < ri) .
We shall also follow the usual tensor convention that when a letter appears in any term as a subscript and a superscript, it is understood that this letter is summed over its range. Let Nn be a Riemannian manifold of dimension n ( > 3) and class C3, (x\ • • •, xn) local coordinates of a point P in Nn, and aaijdxadxf' a Riemannian metric of Nn, where aaP = apa and the matrix (aafl) is positive definite so that the determinant \aa?\ = a is positive. Let Ax, • • •, An_x be n — 1 tangent vectors of the manifold Nn at the point P, and Aai the contravariant components of At in the local coordinate system x1, • • • ,xn. Let Ax X • • • X An_! denote the vector product of the n — 1 vectors Ax, • • •,An_l, which is defined to be the tangent vector of the manifold Mn at P whose /3-th contravariant component is (see, for instance, Feeman and Hsiung [3]) 3i a\A\
a
(2.2)
(A,
X An_y =
n-lnl-l/2
{-\y- a
Si ••• 8> oa2A" • • • aanAx
where <5£ are the Kronecker deltas. Let T be a tangent vector of the manifold Nn at the point P with contravariant components T" in x\ • • •, xn. From the
135
INTEGRAL FORMULAS
definition of the scalar product of any two vectors At and Ap namely, (2.3)
=
aatA\A),
it follows that the scalar product of the two vectors T and A1 x • • • X An_1 is given by (2.4)
(T,Axx
••• xAn_1}
= (-ir-W\T,A1,
••-,An-i\,
where \T, Ax, • • •, An_i\ is a determinant, the elements of each of whose columns are the contravariant components of the vector indicated. Thus by (2.4) it is readily seen that the vector ^ 4 1 x - - - X ^ 4 n _ i i s orthogonal to each of the n — 1 vectors Ax, • • •, An_x. Now consider an orthonormal frame Pex • • • en on Nn at P, where e1; • • •, en form an ordered set of n mutually orthogonal unit tangent vectors of the manifold Nn at P so that (2.5)
<e„, efy = arle^f
= dafl ,
where 8af are the Kronecker deltas. The position vector x of the point P is defined to be the tangent vector of the manifold N™ at the point P whose contravariant components are the local coordinates x1, • • •, xn of the point P. Let av • • • ,an be distinct and suppose that 1 < ax, • • • , a n < n. Then we can write (2-6)
eai x • • • X «.„_, = cean ,
where c is a function of the x's. In order to find an expression for c, we consider the two matrices (2.7)
$ = (#) ,
f = M)
,
(i = 1, • • • , « - 1) ,
where (2.8)
# - aafe"t ,
f f = ef ,
the superscript of the element <j>\ or f? indicating the row to which the element belongs, and the subscript indicating the column. From (2.2) and (2.6) it is easily seen that (2.9)
cel9 = ( - ir+rBra-wi
,
( r = 1, • • •, n) ,
where Br is the determinant of the matrix of (n — l)th order obtained by deleting the f-th column from the matrix <j>. Substitution of (2.9) in (2.5) for a = /3 = an gives
136
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
c2 = B
(2.10) where
B1
• ( - _iy-iBn • ai
•
<
B = (2.11)
-B2
C
e1
•
pn e
o»-i
which is equal to the sum of the products of the corresponding determinants of the (n — l)th order of the two matrices (2.7). By an elementary theorem on determinants (see, for instance, [2, p. 102]), from (2.5) it follows immediately that (2.12)
B = \fiM\ =
l,
which, together with (2.10), implies that (2.13)
c = ±1 .
If the orientations of e1; • • •, en are so chosen that (2.14)
\eu---,en\>0,
then by taking the scalar product of the vector ean with each side of (2.6) and using (2.4), (2.13), we can easily obtain (2.15)
\ev
•
P
=
/7-V2
and therefore (2.16)
e
ai
X
• • • X £«„_! — « « ! . . . « / « , ;
where 8ai...an = + 1 or —1 according as the permutation of alf • • -,an into 1, • • •, n is even or odd. 3. m
Immersed submanifolds
Let x: M —»• N be an m-dimensional (2 < m < n) submanifold of class C3 immersed in a Riemannian «-manifold A^" denned in § 2. For simplicity we shall write x(Mm) as Mm. Let (u1, • • •, um) be local coordinates of a point P on Mm. Then (3.1)
n
x" = x"(ul, • • •, wm) ,
(a = 1, • • •, n) ,
are of class C3, and the first fundamental form of Mm at P is defined to be
499 INTEGRAL FORMULAS
(3.2)
/ = (dx, dx} = g^ditdu1
137
,
where d denotes the exterior differentiation, and the matrix (gi3) is positive definite so that the determinant \gtJ\ = g > 0. Let x°4 denotes the covariant derivative of x" with respect to gtj. Then it is known that (3.3)
x"tt = dx*/du* ,
(3.4)
gtj = a^x^j
.
The element of volume of Mm at P is given by (3.5)
dV = Vg^du1 A ••• A dum ,
where A denotes the exterior multiplication. Now we are in a position to introduce the generalized covariant differentiation, which is useful for studying submanifolds of Riemannian manifolds. Let A°fi be a mixed tensor of the second order in the JC'S and a covariant vector in the w's, as indicated by the Greek and Latin indices. Then following A. W. Tucker [13], the generalized covariant derivative of A°fi with respect to the M'S is defined by
(3.6)
r,A'ft = dA^/du^ + r%Arfrt, - r^A^
- r%A^ ,
where the Christoffel symbols rafr with Greek indices are formed with respect to the aa? and the x's as follows: (3.7)
r% = ! a - f e
+ d-BiL -
d
-<\,
(a°?) being the inverse matrix of (aaf), and those r% with Latin indices are formed with respect to the gtj and the M'S in a similar way. It should be noted that this definition of generalized covariant differentiation can be applied to any tensor in the M'S and the JC'S, and that the generalized covariant differentiation of sums and products of tensors obeys the ordinary rules. If a tensor is one with respect to the M'S only, so that only Latin indices appear, its generalized covariant derivative is the same as its covariant derivative with respect to the M'S. Furthermore, in generalized covariant differentiation, the fundamental tensors aaf and gtJ can be treated as constants. Since x° is an invariant for the transformation of M'S, its generalized covariant derivative is the same as its covariant derivative with respect to the M'S, SO that (3.8)
FiX" = x°t = dx'/du* .
At a point P on Mm we can choose e m+1 , • • •, en of the orthonormal frame Pet- • -en on Nn denned in § 2 to be unit normal vectors of Mm. Then we can have (see, for instance, [16, Chapter X])
500 138
C. C. HSIUNG, JONG DUNG LIU & SITANSU S. MITTRA
(3.9)
PiXj = Z 0AHjeA , A
(3-10)
QA^ =
(3-11)
FteA = -QAmgk]Xj
+ 2 9BA\ieB , B
iJ
where (g ) is the inverse matrix of (jgi}), and (3-12)
=
'*A\ij
(3.13)
'''A\]i
>
SABH + SBAli = 0 ,
so that SAA]i = 0. Thus being defined to be —
(3.14)
The equations of Gauss and Mainardi-Codazzi of M m in Nn are (see, for instance, [2, p. 162]) (3.15)
R-hijk
— Z J \'° A\hk°J A\lj
^C\ij,k
— &C\ik,j
+ Rc,frSX<",l>.X,iXljX,k
'*A\ti,j'°A\ilt)
==
ZJ \^BC\k^B\ij
(3.16)
;
vBcljiJBjik)
* T KaprSe0xtiXjX^k
where the Riemann symbols /? f t ^ s = ghiRlijk the g{j and the u's are defined by
R\jk = ^k-dn±
(3.17)
,
for M m formed with respect to
+ r\trik - rikrij,
and the Riemann symbols RafrS for Nn formed with respect to the aaf and the x's can be similarly defined. In particular, if the manifold Nn is of constant sectional curvature C, from the definition it follows that (3.18)
Rafr, = C(aaSafr — aara?l) ,
and therefore (3.15), (3.16) are reduced, in consequence of (3.4), to (3.19)
Rhijk
= Z J \@ A\hk^
A\ij
~ ^A\kj^A\ik)
+ C(ghkgij
ghjgik)
A
(3.20)
"C\ij,k
— "G\ik,j
=
Z J \P^BC\k"B\ij
&BC\j"B\ik)
B
Moreover, by using (3.11), (3.9), (3.20) we can easily obtain
•
•>
139
INTEGRAL FORMULAS
d2eA =
diP^e^u1) - 0A]ikQBlljgkJ
= S
(P^BAH B \
+ S ScAH&Bc^eBdu1 A du* . 0
1
m
The principal curvature of M at P with respect to a normal vector eA (m + 1 < A < ri) are the eigenvalues /^(e J , • • •, km(eA) of the matrix (/2 AW ) relative to the matrix (gtj), i.e., the roots of the determinant equation (3.22)
det {QMij - Xgij) = 0
in X, and the rth mean curvature of M m at P with respect to eA is defined to be the rth elementary symmetric function of kt(eA), • • •, km(eA) divided by the number of terms, i.e.,
(m)Kr(eA) =
Vr/
(3.23)
2
M O • • • K^A) ,
*!<-
(
fn\
J = m\j(r\{m
— r)\). For convenience, we assume that K0(eA) =
= 1. F e M m is called an umbilical point of M m with respect to e 4 if k^e^) = • • • — km(eA) at P, and M m is called an umbilical submanifold of iV" with respect to a vector field eA if every point of M m is an umbilical point with respect to eA at that point. It is well known that a closed oriented hypersurface in a Euclidean space Em+1 consisting entirely of umbilical points with respect to the unique normal vector field is a hypersphere,. If ka is a real simple root of (3.22), then (3.24)
(QMij - ka8ij)paf
= 0 ,
(/ = 1, • • •,m) ,
define, to within a factor, m quantities paf, i = 1, • • •, m, which are the contravariant components of a real vector in the tangent space of Mm at P, called a principal vector of Mm at P corresponding to the principal curvature ka, as is seen by changing the coordinates and making use of the tensor properties of QA[ij and gtJ. If kt is another real simple root of (3.22), we have a second vector pb |* defined by (3.25)
{QMij - fcjg^Pal* = 0 ,
(/ = 1, • • •, m) .
Multiplying (3.24) by p„\J and (3.25) by pj-*, summing for / in each case and subtracting, we have, since ka ^ kb by hypothesis, (3.26)
SuPatPA3
= 0 ,
140
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
that is, the two vectors pa\l and pb\j are orthogonal. Hence, as is well known, the m principal vectors px\l, • • •, pm\l corresponding to the m principal curvatures kx, • • • ,km with respect to the unit normal vector eA of Mm at P are mutually orthogonal. Lemma 3.1. By a suitable choice of the local coordinates ul, • • •, um of m M at a point P we have (3.27)
VteA = —kiXA -\- J^ 9BMieB
,
(i = 1, • • •, m, not summed) ,
B
where ku • • • ,km are the principal curvatures of Mm at P with respect to eA. Proof. Choose the local coordinates u1, • • •, um of M™ at P such that xA, • • •, xi7n to be the m principal vectors px\l, • • •, pmf of Mm at P corresponding to ku • • •, km, so that gtj = 0 for / ^fc / at P. The contravariant components x°a and paf of the principal vector pa\ in the x's and the M'S respectively are connected by the relation (3.28)
*•„ = x">iPaf
.
Multiplying (3.28) by aaflx% and summing for a we obtain gat = g 6i p 0 |* from which it follows that (3.29)
pa\* = Si .
Substituting (3.29) in (3.24) gives (3.30)
&A\ij — ki£i] >
(/ = 1, • • -,m, not summed) .
From (3.30) and (3.11) follows immediately (3.27). q.e.d. Let f be an infinitesimal conformal vector field on the manifold Nn, and L( the Lie derivative with respect to £. Then on Nn we have (3.31)
L(aaf = £,,,, + £,,„ = 2pa.f ,
where p is a function of x1, • • •, xn. The field f is said to be homethetic or isometric according as p is constant or zero. Lemma 3.2. // the local coordinates x1, • • •, xn on Nn are so chosen that the Kronecker vector 8t, whose contravariant components are the Kronecker deltas 8\, • • •, <5™, generate an infinitesimal conformal vector field on Nn, then onNn (3.32)
Ltla.f
= 2paap = da^/dx1 .
Proof. From the definition of cavariant differentiation with respect to the x's it follows that (3.33)
"la,p
^a r "l.j9
^«r* 10 '
141
INTEGRAL FORMULAS
and similarly 8lf,a = a^F\a. By means of (3.7) we readily have aarF[f + afrrrlir = daaf/dx\ which together with (3.31) gives (3.32). q.e.d. If the vector 5j generates an infinitesimal conformal vector field on Nn, then using (3.33) we immediately obtain that on Nn (3.34)
d(8l) = dlrdxr = dl^du1
= ri^du1
,
which together with (3.7) implies
(3.35)
aASl = I f - ^ '
+ ^iL - ^ W « ' . dx1
2\dxr 4.
dx" ) '
Integral formulas
Let x: Mm —» N" be an m-dimensional (2 < m < «) submanifold of class C immersed in a Riemannian w-manifold W , which is of constant sectional curvature and admits a continuous infinitesimal conformal vector field f, so that §§2 and 3 can be applied. In this section we shall derive some integral formulas for closed oriented M m with respect to a fixed unit normal vector field, em+1 say, on M m . For this purpose we choose the orientation of the orthonormal frame Pe1 • • • en of Nn at a point P defined in § 3 such that (2.14) and therefore (2.15) hold, and we also choose the local coordinates JC1, • • •, xn and w1, • • • ,um ol Nn and M m at P respectively such that the Kronecker vector 81 be the infinitesimal conformal vector | , and that x;1, • • •, x>m be the m principal vectors py\l, • • •, p m |* of Mm at P with respect to e m+1 , so that at P 3
(4.1) (4.2)
gtJ = 0, et = *,[/] ,
(i#j), (i = 1, • • •, m) ,
where (4.3)
[i] = Vfe .
Now we are in a position to evaluate the following exact differential m-form for 1
= (I) + ( - l ) " - ' ( i - 1)(II) + ( - l ) » - l ( I I I ) + (-i)™- 1 (iv) + ( - i ) ™ - 1 s (V)„,
where we have used d2x = 0 and put
142
(4.5)
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
(I) = Va\d8!,dx,
• • -,dx,dem+1, • • • ,dem+1,em+1, < —, — TO—
(4.6) (4.7)
i
• -,e„\ ,
i—1
• • •, dx, d2em+1, dem+l,
• • -,dem+1,em+1,.-
m—i
t —2
(II) = i/a\dvdx, (III) = Vald^dx,
• • -,dx, dem+1, • • -,dem+1,em+2, m-i
--,en\
,
• • -,en\ ,
i
(IV) = VQ |o 1} dx, • • •, dx, dem+1, • • •, aem+1, em+1, aem+2, (4.o)
m_i
j_i
(V)„ = Vfl|5 1 ; dx, • ••, dx,dem+1,
••
-,dem+l,
By means of (3,27) for ,4 = m + 1, (4.2), (4.3), (2.4), (2.3), (2.16), (3.35), (3.32), (3.4), (3.5), (4.1) we obtain (I) = (-ly-Walddt,
[h]ehdu^, • • •, [in_t]e/m_tdu>*-' i
' jm-i
+ iUm-i
e
:!m
+ li im-i
+ 1dU
~'-
,
+1
,
' ' ' ' ^ J m - i L / m - i J ^ J m - i " ^ ™ ' ^m + 1! ' ' ' > &n I
= (_1)™ + V _ ,•)!(/ _ Dla^dSi A ^ [ / J • • • [jm-J •^.-^-<
(4.10)
+
i ••• ^ . - A '
1
A • • • A <*<_,
= (_l)»+«(m - ,•)!(,• - 1)!
• • • kjm-,duil
- (-n»^(m-Q!(i2[/ m ]
A • • • A <*«>-'>
l)lr-1 . . . r /
i 3a./, _. ,/. 9* 1
• ^ . - ^ - * + . • • • **«-><*"'" A d V ' A • • • A ^ M ^ - 1 = ( - 1)*-V>(™ - 0 ! (i - 1 ) ! (m - i + 1) 2 L—I
* * . • + , - • • * * _ Av • _
Jm — i + 1
Jm—1
It should be remarked that in the summation on /i, • • •, jm in (4.10) for fixed Jm-i+i, •••>7m-i>m — i other y's are together and their order is immaterial, and the remaining / can take any one of the other (m — i + 1) /'s, namely, 7i> • • • > 7m-€> 7m> so that we get the factor m — i + 1. From (4.10) and (3.23) follows immediately
505
INTEGRAL FORMULAS
(4.11)
(I) = ( - D ' - ' / n ! pK^ie^ddV
143
.
Substituting (3.21) for A = m + 1 in (4.6) gives readily (4.12)
(II) = 0 .
Using the same method as above we can easily obtain (4.13)
(III) = ( - l ) m + 4 m ! iSuen^Kt{e^ddV
(4.14)
(IV) = ( - l)m+i(m
,
- 0 ! (i - 1)! <3lf em+2>
3m—i + 1 ) * " 13m
The vector field 8X can be decomposed into two parts: (4.15)
3, = <S1U + 8m ,
where 5 ^ is tangent to M m , and 5n„ normal to Mm. Let e and e be two unit normal vector fields over Mm coplanar with 8lln. Then (4.16)
8m = <31In, e>e + <51|K, e>e .
Now suppose that the unit normal vector field em+1 is parallel in the normal bundle of M™, i.e., by the definition, dem+1 is tangent to Mm everywhere. Then by choosing em+l = e and e m+2 = e everywhere o n M m and using (4.15) and (4.16) we obtain (4.17)
<<51; e m+a > = 0 ,
(a = 3, • • - , « - m) ,
and therefore (4.18)
(V)„ = 0,
(3
Combination of (4.4), (4.11), • • •, (4.14), (4.18) gives d(*/a\81,dx,
• • •,dx, dem+1, • • •, dem+1, em+1, • • -,en\) m—i
(4.19)
=
i—1
( _ 1 } i - i m ! [pK^e^)
+ <3„ « m+1 >X,(e» +1 ) + F ^ ^ W (z = 1, • • -,m)
, .
where Fie (4.20)
)-
(m
~ ° ! (l' ~ m
•
1}!
<S e
\
T
k-
Jm-i+li—iJm-l
Integrating (4.19) over an oriented Mm and applying Stokes' theorem we hence arrive at
144
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
Theorem 4.1. Let x:Mm-^-Nn be a closed oriented m-dimensional (2 < m < n) submanifold of class C3 immersed in a Riemannian n-manifold Nn, which is of constant sectional curvature and admits a continuous infinitesimal conformal vector field f. // e m+1 and em+2 are unit normal vector fields over Mm such that em+1 is parallel in the normal bundle of Mm, and em+l, em+2 are coplanar with the normal component of f, then [pKi-i(em.+i) + <,S,em+iyKi(em^)]dV (4.21)
= -|
J M™
Ft(em+1)dV
,
J M™.
(i — 1, • • -,m) , where p is given by (3.32). Remarks. 1. If n — m — 1, then Fi(em+l) = 0, i — 1, • • -,m, hold automatically, and formulas (4.21) are due to Hsiung [5] for Euclidean Nn with | generated by the position vector x of a general point of Mm with respect to a fixed point 0 in Nn, due to Hsiung [6] and Feeman and Hsiung [3] for a Riemannian Nn and a special f, and due to Katsurada [7] for a Riemannian Nn and a general f. 2. For Euclidean Nn and general n with the position vector field x as f, formulas (4.21) are due to Chen and Yano [1], and due to Yano [14], [15] under some additional conditions. 3. For Euclidean iV", the condition of the parallelism of em+l in the normal bundle of Mm can be replaced by the condition that Mm be immersed in a hypersphere of Af" centered at the origin of ./V". 4. For a special em+1, formulas (4.21) are due to Katsurada and Kojyo [13], and Katsurada [8]. 5.
Characterizations of umbilical submanifolds
In this section we use integral formula (4.21) to derive various conditions for a submanifold of a Riemannian manifold to be umbilical with respect to a given normal vector field. For this purpose we first state the following three lemmas which will be needed for the proofs of our main theorems. The proofs of the lemmas are omitted here, but can be found in [4, pp. 52, 104-105]. Lemma 5.1. Let K^e^, i = 1, • • • ,m, be given by (3.23). Then
(5.1)
Ktie^
- K^ieJKUeA)
> 0,
(i = 1, • • •, m - 1) ,
where the equality implies that k^e^) = • • • = km(eA). Lemma 5.2. // Ki(eA), K^^e^, • • •, Kt_j_l(eA) > 0, 1 < /' < i < m, then (5 2)
Kj~i(eA)
->. Ki_2(eA)
>...->
K-i-j-i(eA)
Kt{eA)
~ Ki^ieJ
~~
Kt_j(eA)
~~
145
INTEGRAL FORMULAS
where the equality at any stage implies that k^e^ = • • • = Lemma 5.3. // Kx(eA), • • •, Kj(eA) > 0, /' < m, then (5.3)
Kx(eA) > K2(eAy<> > £ 3 ( 0 1 / 3 > ••• > K,{eJ"
km(eA).
,
where the equality at any stage implies that k^e^ = • • • = km(eA). In the remainder of this section we shall use the following notation: Nn: A Riemannian n-manifold (n > 2) having constant sectional curvature and admitting a continuous infinitesimal conformal vector field £ so that Ljtf,,,, = 2pa„p where aall is the Riemannian metric tensor of Nn. Mm : A closed oriented m-dimensional (n > m > 2) submanifold of class C3 immersed in Nn. e: A unit normal vector field on M m parallel in the normal bundle of M m . k^K^Ft, and p: ^i(e),i
p = pK^JKt
.
For i<m, substituting (5.4) in (4.21), where i is replaced by i + 1, gives (5.5)
f
-£-(X 4 s - K^K^dV
= 0 .
Due to (i) and (5.1), the integrand of (5.5) is nonnegative, and therefore (5.5) holds only when, at all points of M m , Kf — ^ j . ^ + j = 0. From Lemma 5.1 it follows that kx = • • • = km at all points of Mm, and hence Mm is umbilical with respect to e. For i = m, substituting (5.4) in (4.21) where i is replaced by / — 1, we obtain (5.6)
f J»t™
f(X Km
B
. / - Km_2Km)dV
= 0 .
508 146
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
By applying Lemma 5.1 with the same argument as above, we can show that Mm is also umbilical with respect to e. Theorem 5.2. Mm is umbilical with respect to e if at all points of Mm for an integer i, 1 < i < m, ( i ) p, Ki+1, Ki, Ki_1 > 0, (ii) p > —pKi-.JKt, (iii) F i + 1 = 0. For Euclidean Nn with the position vector field x as ?, Theorem 5.2 is due to Chen and Yano [1]. It should be remarked that we may have a similar theorem by assuming p < 0 instead of p > 0. Proof. By (ii) and Lemma 5.2 we have (5.7)
p > -pK^/Ki
> -pKt/Kt+1
.
(4.21), with i replaced by i + 1, and (5.7) imply that the equality holds in (5.7), and hence M m is umbilical with respect to e by Lemma 5.2. Theorem 5.3. Mm is umbilical with respect to e if at all points of Mm for an integer s, 1 < s < m, ( i ) p is of the same sign, (ii) K, > 0 , i = 1, ...,s, (iii) Ks is constant, (iv) p is of the same sign, (v) F1 = Fs+l = 0 for 1 < s < m, and Fx = Fs — 0 for s — m. For Euclidean Nn and n = m + 1, Theorem 5.3 is due to Hsiung [5]. Proof. Case 1. s < m. By (ii) and inequality (5.1) for i= 1, •••, s we obtain Ki/K0 > KzlK.! > • • • > Ks+l/Ks
,
and, in particular, (5.8)
K^s
> Ks+1 ,
where the equality holds only when k1 = • • • = km in view of Lemma 5.1. Here we assume p > 0. Then from (4.21) for / = 1 and assumptions (i), (ii), (v) it follows that p is negative. (For the case p < 0, the arguments in the proof of our theorem will be exactly the same, except that p would be positive.) Multiplying both sides of inequality (5.8) by p, integrating over M™, and applying (4.21) for i = 1 and / = s + 1, we can readily obtain, in consequence of (iii) and (v), -
f
PKsdV
= f
from which it follows that
pK^dV
< [
PKs+1dV
= -
f
PKsdV
,
147
INTEGRAL FORMULAS
(5.9)
f
piKJZ. - Ks+1)dV = 0 .
Since p is negative, from (5.8) we see that the integrand in (5.9) is nonpositive and therefore must be zero. Thus the equality holds in (5.8) so that kt = • • • = km everywhere by Lemma 5.1. Hence Mm is umbilical with respect to e. Case 2. s = m. From (ii), (iii) and Lemma 5.3 it follows that Kx > K2"2 > ••• > K , . ! 1 ' * - " > Km1/m = c ,
(5.10)
where c is a positive constant. By means of (4.21) for i = m, assumption (v) and inequalities (5.10), we obtain (5.11)
f
PKmdV
= -
f
9Km_xdV
< -c™" 1 f
pdV .
On the other hand, using (4.21) for i = 1, (v), (5.10) and the fact that p < 0, we have f
pKmdV = [
pcmdV = cm-1 f
pKm1/mdV
(5.12)
> c91"1 f
pK.dV = -c" 1 " 1 f
pdV
Combination of (5.11) and (5.12) shows immediately that the equality holds in (5.12) and therefore that (5.13)
f
p(Kmv™ - KJdV = 0 .
Since p < 0, (5.10) implies that the integrand of (5.13) is nonnegative and therefore that Kx — Km1/m. Thus by Lemma 5.3, ^ = • • • = km at all points of Mm. Hence the proof of Theorem 5.3 is complete. Theorem 5.4. Mm is umbilical with respect to e if at all points of Mm for two integers i and s,i < i < s < m, ( i ) Kt, Ki+1, • • •, Ks > 0, (ii) Ks = 2}=i CjK-p f°r some constants Cj > 0, i < j < s — 1, (iii) p is of the same sign, (iv) Fj = 0,j=l,...,s-1. Proof. We observe (5.14)
* —
J- 1 — K±. (K-S_i _ J^j-A
In view of Lemma 5.2, the right side of (5.14) is nonnegative for i < j < s — 1. Thus
148
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
(5.15)
K,IK,7>Ki_lIKt_l,
where the equality holds only when kx = • • • = km. By (ii) and (5.15) we obtain 1 = Z CJKJ/K,
> S CJKJ.JK..,
,
or (5.16)
K.-1-LCJKJ.^O, J=i
where the equality holds only when kx = • • • = km. Thus by means of (4.21), (iv) and (ii) we obtain (5.17)
f J Mm
pU^ - Z c,K3\dV \
j~t
= - [
J
P(K,-
J M™
\
'£ CjKj)dV = 0 . j =i
J
(5.16), (5.17), (iii) show immediately that the equality holds in (5.16). Hence Mm is umbilical with respect to e. Theorem 5.5. Mm is umbilical with respect to e if at all points of Mm for two integers i and s, 0 < / < s < m, (i) Kt,---,K,+1>0, (ii) Ks = TiY=i CjKj,-for some constants c} > 0, i < j < s — 1, (iii) p is of the same sign, (iv) Fj = 0,j= 1, . . . , s - 1. Proof. By Lemma 5.2 we have (5.18)
^J — ^J+1 = ^J (^s+1 — !^i+±\ < 0 , Ks Ks+i Ks+l\ Ks Kj 1
where the equality holds only when kt = • • • km. From (ii), (5.18) it follows that i
1 = 2 CJKJ/K,
< Z] CjK ' j+1/Ks+1 ,
or (5.19)
Ks+1 - Z CjKJ+1 < 0 ,
where the equality holds only when kx = • • • = km. Thus by means of (4.21), (iv) and (ii) we obtain
149
INTEGRAL FORMULAS
(5.20)
f
P(KS+1
- Z CjKj+l)dV
= - f
p(K, - L
CjKj)dV
= 0.
(5.19), (5.20), (iii) show immediately that the equality holds in (5.19). Hence Mm is umbilical with respect to e. Theorem 5.6. Mm is umbilical with respect to e if at all points of Mm for an integer i, 1 < / < m, (i) Kt>0, (ii) Ki = cKi_1, for some constant c, (iii) p is of the same sign, (iv) F^ = Ft = 0. Proof. Due to (i), c cannot be zero and JKJ_J must be of a fixed sign. Using (ii) and Lemma 5.1 we have •^i-i^i-i
—
cKt_2) = Kt_^ — KiKi.2 > 0 ,
so that (5.21)
Ki_i — cKt_2
is of fixed sign ,
and vanishes identically only when k, = • • • = km. Thus by means of (4.21), (iv) and (ii) we obtain (5.22)
f
piK^ - cKt_2)dV = - [
p(Kt - cK^dV
= 0.
(5.21), (5.22), (iii) imply immediately that K^^ = cKt_2. Hence Theorem 5.6 is proved. Corollary 5.6. Mm is umbilical with respect to e if at all points of Mm (i) Km>0, (ii) 2?=i ( V W = constant, (iii) p is of the same sign, (iv) Fm_, = Fm = 0. Proof. By (ii) and the definition (3.23) of Kt we obtain m
mKm_JKm
= 2 (1/&*) = constant ,
so that Km = cKm_! ,
for some constant c .
Hence Corollary 5.6 is an immediate consequence of Theorem 5.6 for i = m. Theorem 5.7. Mm is umbilical with respect to e if at all points for an integer s, 1 < s < m, and a constant c (i) Kt>0fori=l,...,s,
150
C. C. HSIUNG, J O N G DUNG LIU & SITANSU S. MITTRA
(ii) K,_^-» > c > Ksl's, (iii) p is of the same sign, (iv) p is of the same sign, ( v ) Ft=F2 = Fs = 0. Proof. As in the proof of Theorem 5.3 we may assume p > 0. Then due to (iii), (v) and (i) for i = 1, (4.21) for i = 1 implies p < 0. By (5.3), (ii) we have Kx > i£ s _ 1 1/(s_1) > c, and therefore, in consequence of (ii), (4.21) for i = s and i = 1, -
c'-lpKxdV
[
> -
cspdV > -
f
J M™
f
J M™
(5.23)
= f
pcs~HV
pKs_,dV > [
J M™
= -
pKsdV
J jtf™
J Mi*
c'-1pK1dV .
f J Si™
Thus the equality holds everywhere in (5.23), so that f J
piK, - c)dV = 0 , M™
which implies that K^ = c. Hence, by Theorem 5.3 for s = 1, Mm is umbilical with respect to e. Theorem 5.8. Mm is umbilical with respect to e if at all points of Mm for an integer s, 1 < s < m, and a constant c (i) Ks^,Ks>0, (ii) K..JK, >c> K,_t/K,_lt (iii) p is of the same sign, (iv) p is of the same sign, (v) F s _ 1 = JFs = 0. Proof. As before we may assume p > 0. Then due to (i), (iii) and (v), (4.21) implies p < 0. By using (ii), (4.21) for / = s — 1 and i = s we have f , ^ _ ..
PKs_2dV
= -
J Mm
f
pK,_xdV > -
J M"»
J M"i
f
cPKsdV
J M™
J M™
Thus the equality holds everywhere in (5.24), so that (5.25)
f
p(Ks_, - cKs)dV = 0 .
Since p(Xs_! — cKs) < 0, (5.25) implies that K,^ = cKs at all points of Mm. Hence, by Theorem 5.6 for i = s, Mm is umbilical with respect to e.
INTEGRAL FORMULAS
151
Theorems 5.4, 5.5, 5.6 and Corollary 5.6 are due to Chen and Yano [1] for Euclidean Nn with the position vector field x as f. Theorems 5.4, • • •, 5.8 are due to Strong [12] for n = m + 1 with the position vector field x as f. References [ 1]
B. Y. Chen & K. Yano, Integral formulas for submanifolds and their applications, J. Differential Geometry 5 (1971) 467-477. [ 2 ] L. P. Eisenhart, Riemannian geometry, Princeton University Press, Princeton, 1949. [ 3 ] G. F . Feeman & C. C. Hsiung, Characterizations of Riemann n-spheres, Amer. J. Math. 81 (1959) 691-708. [ 4 ] G. H . Hardy, J. E. Littlewood & G. Polya, Inequalities, Cambridge University Press, Cambridge, 1934. [ 5 ] C. C. Hsiung, Some integral formulas for closed hypersurfaces, Math. Scand. 2 (1954) 286-294. [ 6] , Some integral formulas for closed hypersurfaces in Riemannian space, Pacific J. Math. 6 (1956) 291-299. [ 7 ] Y. Katsurada, Generalized Minkowski formulas for closed hypersurfaces in Riemann space, Ann. Mat. Pura Appl. 57 (1962) 283-293. [ 8] , Closed submanifolds with constant v-th mean curvature related with a vector field in a Riemannian manifold, J. F a c . Sci. Hokkaido Univ. Ser. I, 20 (1969) 171-181. [ 9 ] Y. Katsurada & H . Kojyo, Some integral formulas for closed submanifolds in a Riemann space, J. Fac. Sci. Hokkaido Univ. Ser. I, 20 (1968) 90-100. [10] Y. Katsurada & T . Nagai, On some properties of a submanifold with constant mean curvature in a Riemann space, J. F a c . Sci. Hokkaido Univ. Ser. I, 20 (1968) 79-89. [11] H. Minkowski, Volumen und Oberflache, Math. Ann. 57 (1903) 447-495. [12] R. E. Stong, Some characterizations of Riemann n-spheres, Proc. Amer. Math. Soc. 11 (1960) 945-51. [13] A. W. Tucker, On generalised covariant differentiation, Ann. of Math. 32 (1931) 451-460. [14] K. Yano, Integral formulas for submanifolds and their applications, Canad. J. Math. 22 (1970) 376-388. [15] , Submanifolds with parallel mean curvature vector of a Euclidean space or a sphere, Kodai Math. Sem. Rep. 23 (1971) 144-159. [16] C. E. Weatherburn, An introduction to Riemannian geometry and the tensor calculus, Cambridge University Press, Cambridge, 1950. LEHIGH UNIVERSITY NATIONAL TSING H U A UNIVERSITY AND A C A D E M I A SLNICA PENNSYLVANIA BUREAU OF CORRECTION
A GENERALIZATION OF THE RIGIDITY THEOREM OF COHN-VOSSEN CHUAN-CHIH HSIUNG AND JONG DUNG LIU 1. Introduction The well-known rigidity theorem of S. Cohn-Vossen [1] states that an isometry between two closed convex surfaces in a Euclidean 3-space is trivial; that is, it is either a motion or a motion and a reflection, or, equivalently, it preserves the second fundamental form with a possible change of sign. In 1943 G. Herglotz [6] gave an ingenious proof of this theorem by introducing an interesting integral formula. The converse of this theorem was proved by V. G. Grove [5], extended by R. B. Gardner [4] to two closed convex hypersurfaces in a Euclidean (n+l)-space, and further extended jointly by C. C. Hsiung and S. S. Mittra [7] to two compact convex hypersurfaces with boundaries in a locally flat Riemannian (n + l)-space. The purpose of the present paper is to extend Cohn-Vossen's theorem to two closed convex surfaces in a Riemannian 3-manifold by establishing the following theorem; our proof is a generalization of that of Herglotz. THEOREM 1.1. Let S, S* be two compact oriented immersed surfaces, in a Riemannian 3-manifold R3 of constant sectional curvature, with second fundamental forms II, II* with respect to unit normal vectors e3, e3*, respectively. Suppose that there is a continuous infinitesimal conformal vector field £ of R3 over a subdomain D containing S and S*, the form II is positive definite, and there is an onto-isometry f:S~* S*. If the scalar product £.e3 (respectively £,.e3*) in R3 is of the same sign over S (respectively S*), thenII = II* underf
When R3 is a Euclidean manifold E3, it is easily seen that the position vector x (respectively x*) of a general point of S (respectively S*) with respect to a fixed point 0 in E3 generates an infinitesimal homothetic vector field over S (respectively S*). If we choose this vector field x (respectively x*) to be our £, and if S, S* are convex, then we can choose O in the convex set bounded by S (respectively S*) so that £. e3 (respectively £,. e3*) is of the same sign over S (respectively S*). Hence the rigidity theorem of Cohn-Vossen is a special case of our theorem. The following elementary lemma will be needed in the proof of Theorem 1.1. LEMMA 1.1.
Let
Xx2 + 2fixy + vy2, X* x2+ 2[i* xy + v* y2 be two positive definite quadratic forms with X*v* — fi*2 = Xv — /i 2 . Received 1 July, 1976. The second author was visiting Lehigh University and partially supported by the National Science Council of the Republic of China. [J. LONDON MATH. SOC. (2), 15 (1977), 557-565]
558
CHUAN-CHIH HSIUNG AND JONG DUNG LIU
Then X* — X
n* — fi <0, li* —H v* —v
where the equality holds when and only when the two forms are identical. In §2 we first define the vector product of 2 tangent vectors of a Riemannian 3-manifold R3 at a point P, and then discuss orthonormal frames Peh eheh on R3 at P. §3 contains the fundamental definitions and formulas for a surface S immersed in a Riemannian 3-manifold R3. In particular, some formulas are reduced to simpler forms, when R3 is of constant sectional curvature. §4 is devoted to the proof of Theorem 1.1.
2. Vector product and orthonormal frames Throughout this paper unless stated otherwise the ranges of indices are given as follows: 1 <*,./,*,...< 3; 1 < « , / ? , y , . . . < 2. (2.1) We shall also follow the usual tensor convention that when a letter appears in any term as a subscript and a superscript, it is understood that this letter is summed over its range. Let R3 be a Riemannian 3-manifold of class C3, x'(i = 1, 2, 3) local co-ordinates of a point P in R3, and atj dx' dx} a Riemannian metric of R3, where ai} = a^ and the matrix (a y ) is positive definite so that the determinant |a£j-| = a is positive. Let Au A2 be two tangent vectors of the manifold R3 at the point P, and A^, A2' the contravariant components of Alt A2 in the local co-ordinate system x1, x2, x3 respectively. Let Ax x A2 denote the vector product of the two vectors Au A2, which is defined to be the tangent vector of the manifold R3 at the point P whose j-ih contravariant component is (see, for instance, [3])
{AxxA2y
V
an A^
ai2 At'
ai3 At'
an A2'
ai2 A2'
ai3 A2'
(2.2)
where 8/ are the Kronecker deltas. Let T be a tangent vector of the manifold R3 at the point P with contravariant components T' in x1, x2, x3. From the definition of the scalar product of any two vectors Ax and Ap, namely, i.J^flijA'M
(2-3)
it follows that the scalar product of the two vectors T and A1 x A2 is given by T.A1xA2
= ai\T,A1,A2\,
(2.4)
where \T, Alt A2\ is a determinant, the elements of each of whose columns are the contravariant components of the vector indicated. Thus by (2.4) it is readily seen that the vector AtxA2 is orthogonal to each of the two vectors Au A2.
A GENERALIZATION OF THE RIGIDITY THEOREM OF COHN-VOSSEN
559
Now consider an orthonormal frame Pe^ e2 e3 on R3 at P, where eu e2, e3 form an ordered set of 3 mutually orthogonal unit tangent vectors of the manifold R3 at P so that ei-ej = ahkeiheJk = Sij, (2.5) where <5y are the Kronecker deltas. The position vector x of the point P is denned to be the tangent vector of the manifold R3 at the point P whose contravariant components are the local co-ordinates x1, x2, x3 of the point P. Let J'I, i2, (3 be distinct and suppose that 1 < iu i2, i3 < 3. Then we can write eilxeh
= ceh,
(2.6)
where c is a function of the co-ordinates x'. In order to find an expression for c, we consider the two matrices <£ = «>/), / = aijeJ, ^J = eJ, (2.8) the superscript of the element >/ or ij/J indicating the row to which the element belongs, and the subscript indicating the column. From (2.2) and (2.6) it is easily seen that c e , / = ( - l ) r + , # , a - * (r=l,2,3), (2.9) where Br is the determinant of the matrix of the second order obtained by deleting the r-th column from the matrix 0. Substitution of (2.9) in (2.5) for i = j = i3 gives c2 = B,
(2.10)
where B1 1
eu e^
-B2
(-1)" B
e,,2
eu3 ei23
2
ei2
(2.11)
which is equal to the sum of the products of the corresponding determinants of the second order of the two matrices (2.7). By an elementary theorem on determinants (see, for instance, [2; p. 102]), from (2.5) it follows immediately that B = \4>j' ^PJ\ = h
(2.12)
which, together with (10), implies that c=+l.
(2.13)
If the orientations of eu e2, e3 are so chosen that \eu e2, e3\ > 0,
(2.14)
then by taking the scalar product of the vector eh with each side of (2.6) and using (2.4), (2.13), we can easily obtain etl x eh = (5,-, i2 h eh,
(2.15)
ki. e2, e3\ =
(2.16)
and therefore where 5h or odd.
h h
is +1 or — 1 according as the permutation of iu i2, i3 into 1, 2, 3, is even
CHUAN-CHIH HSIUNG AND JONG DUNG LIU
560
3. Immersed surfaces Let S be a C3 surface immersed in a Riemannian 3-manifold R3 defined in §2, and let (w1, K2) be local co-ordinates of a point P on S. Then x> = xl{u\ u\
(1 = 1,2,3),
(3.1)
are of class C3, and the first fundamental form of S at P is defined to be I =dx.dx = g^du'du",
(3.2)
where d denotes the exterior differentiation, and the matrix (ga/)) is positive definite, so that \gxp\ = g > 0. Since the functions x' are invariants for transformations of the local co-ordinates u1, u2 on S, their first covariant derivatives with respect to gap are the same as their partial derivatives with respect to u1, u2, that is, x\a = dxi/du',
(3.3)
g^ = aijx\xx\l>.
(3.4)
and therefore we have The element of area of S at P is given by dA = J(g)duiAdu2,
(3.5)
where A denotes the exterior multiplication. Now we are in a position to introduce the generalized covariant differentiation. Let A'jx be a mixed tensor of the second order in the co-ordinates x and a covariant vector in the co-ordinates u, as indicated by the Latin and Greek indices. Then, following A. W. Tucker [8], the generalized covariant derivative of A'jX with respect to the co-ordinates u is defined by
v, A% = dA'jjeue+r'„ A"J. xlj - r*yi A \ X *«,, - n „ A1}V
(3.6)
where the Christoffel symbols T'Jk with Latin indices are formed with respect to the a,j and the co-ordinates x as follows:
p J
.,(dan
+
dakl
da,k\
*"^(i? ^ - w ) -
(3 7)
'
x
(a' ) being the inverse matrix of (atJ), and those r fy with Greek indices are formed with respect to the gxP and the co-ordinates « i n a similar way. It should be noted that this definition of generalized covariant differentiation can be applied to any tensor in the co-ordinates u and x, and that the generalized covariant differentiation of sums and products of tensors obeys the ordinary rules. If a tensor is one with respect to the u' only, so that only Greek indices appear, its generalized covariant derivative is the same as its covariant derivative with respect to these uK Furthermore, in generalized covariant differentiation the fundamental tensors atj and gxp can be treated as constants. Since x' is an invariant for transformation of the uJ, its generalized covariant derivative is the same as its covariant derivative with respect to the u? so that V^x^x'^^dx'/du*. (3.8) At a point P on S we can choose e3 of the orthonormal frame Pet e2 e3 defined
A GENERALIZATION OF THE RIGIDITY THEOREM OF COHN-VOSSEN
561
in §2 to be a unit normal vector of S. Then we can have ?/>*,« = bx/le3,
(3.9)
b,f, = (y0xj.e3,
(3.10)
^e3=bxPg^xiy,
(3.11)
where (g"") is the inverse matrix of (gxf), and bxfl = bfx. Thus at P the second fundamental form //, defined by / / = — dx.de3, the mean curvature H and the Gaussian curvature K of S are given by II ^b^du'du13, (3.12) H = (2g)~1(gi1b22-g12b2l-g21b12+g22b11),
(3.13)
K = *_1(*n*22-*«*«)•
(3.14)
The equations of Gauss and Mainardi-Codazzi for the surface S are Kfyt = (Ks bPy - bxy bfi}) + RhiJk x*>a x\„ x\y x\t, b*fi, y ~ bxy, /5 = RhiJk e3 where the Riemann symbols Rxfyt the M are defined by
R'pys -
., du>
X
(3.15)
,a x ,P x ,y >
(3.16)
= gzx R"PyS for S formed with respect to the gxfi and
-
, y1 du
+^pfiy r"pt—rppar'pl/,
(3.17)
and the Riemann symbols RhiJk for R3 formed with respect to the atJ and the x can be similarly defined. Equations (3.9), ..., (3.16) can be found, for instance, in [9; Chapter VIII]. In particular, if the manifold R3 is of constant sectional curvature C, from the definition it follows that R~njk = C(ahk au - ahJ aik), (3.18) and therefore (3.15), (3.16) are reduced, in consequence of (3.4), to Rtfyi = (bx}bPy-bxybpil) bxfiy-bxyiP
+ C(,gx3gPy-gxygfii),
(3.19)
= 0.
(3.20) 3
Let £ be an infinitesimal conformal vector field on a subdomain D of R , and L{ the Lie derivative with respect to £,. Then on D we have L
i au = S (, J+, >• = 2Pau> (3 •21) 3 where p is a function of x , x , x . The field £, is said to be homothetic or isometric according as p is constant or zero. 1
2
3 . 1 . If the local co-ordinates x1^2,x3 on R3 are so chosen that the Kronecker vector Su whose contravariant components are the Kronecker deltas 5^,3^,3^, generate an infinitesimal conformal vector field on a subdomain D ofR3, then on D LSl atJ = 2pa,j = da^dx1. (3.22) LEMMA
562
CHUAN-CHIH HSIUNG AND JONG DUNG LIU
Proof. From the definition of covariant differentiation with respect to the coordinates x' it follows that &u,j = aik5"u] = alk Tklp (3.23) and similarly 5ljt, = ajk Fku. By means of (3.7) we readily have aikTk1J + aJkTkli = da^/dx1, which together with (3.21) gives (3.22). If the vector <5X generates an infinitesimal conformal vector field on a subdomain D of R3, then using (3.23) we immediately obtain that on D d{d^)=diUkdxk=diukxktJu" = r ' n ^ t f ,
(3.24)
which together with (3.7) implies that / da, i
8aki
da,k \ ,
4. Proof of Theorem 1.1 Without loss of generality we may assume that the corresponding points of the two surfaces S and S* under / have the same local co-ordinates u1 and u2. Thus §2 and §3 can be applied to the manifold R3 and the surface S, and for the corresponding quantities and formulas for S* we shall use the same symbols and numbers with a star, respectively. We may choose the orientation of the normal vector e3 such that (2.14) holds, and choose the local co-ordinates x1, x2, x3 and ul, u2 on D and S respectively such that the Kronecker vector 8t is our infinitesimal conformal vector £, and e ei = *,i/V£n> i = xt2ly/g22, (4.1) so that gn = 0, g*12 = 0, (4.2) the second equation of (4.2) being due to the fact that/is an isometry. First, we want to evaluate the following exact differential 2-form: d(J(a)\Su W, e3\) = rf + &-<#,
(4.3)
where W = -^—
(Iil-){b*11du1+b\2du2)x1
s/ = y/(fi)\d8uW,e3\,
(4.5)
@ = J{a)\8udW,e3\,
(4.6)
V = y/(a)\8ltW,de3\.
(4.7)
A GENERALIZATION OF THE RIGIDITY THEOREM OF COHN-VOSSEN
563
By using (2.4), (2.15), (3.25) we have s/ = ddx .Wxe3 [-(llJg)(b*11du1+b*12du2)x,2
= ay (<«!)'A
+ (1/VS) @*12 rf"1 + *>*22 du2) XA]J 1
/galJ
,
8a
kJ
da
lk\
.
, ,*
k
^
k
f2
•* ,1 ^
. ,„
1 2 " ^ " ^ f 2 ^ f 2 ^ 11/'
We can cancel some terms in the above equation by interchanging the dummy indices / and k, and then obtain, in consequence of (3.22), (3.4), (4.2), (3.13)* and (3.5), st = 2pH*dA.
(4.8)
Direct exterior differentiation of (4.4) and use of (3.9), (3.20)* give
=
dW
~Jg\j(ff)
(V l 12 V 2 ll) 1
« ** - » ** *'
+ / ( ^ i i - ) ( V „ . 6 * 2 2 - V „ I Z . * 1 2 ) x 2 dul*du2 + (*)e3
and therefore 38=0.
(4.9)
From (3.11) and (4.2) it follows that on S de3 =
<7-2
(611di/1+fe12rfM2)*i—
£
gil
g
(b12du1+b22du2)x
2.
(4.10)
By using (2.4), (2.15), (4.10) and (3.5) we obtain <8
=5i.Wxde3 = <5i • [-£~*(6*ii du1 +b*12 du2) A (Z>21 du1 +b22 du2) e3 +g~i(b*2idu1
+b*22du2) A^nrtw 1 +b12du2) e3]
= -(l/g)(S1.e3)(b*11b22-2b*12b12+b*22b11)dA.
(4.11)
Combination of (4.8), (4.9), (4.11) and (4.3) gives (dy/(a)\Su W,e3\) = 2 p H * ^ + (l/g)(^.c 3 )(** 1 1 2> 22 -2ft* 12 Z> 12 +Z,* 22 b^dA. (4.12) Integrating (4.12) over S and applying Green's theorem to the left side of the equation, we thus arrive at the integral formula
2J" [PH*dA+jj— s
s
(51.e3)(b*llb22-2b*12b12+b*22b11)dA = 0.
(4.13)
564
CHUAN-CHIH HSIUNG AND JONG DUNG LIU
On the other hand, since gaf = g*afi, and Rx^d and R*xfiyd depend only on the g^ from (3.19), (3.19)* and the assumption that the second fundamental form// of S is positive definite it follows that b*11b*22-b*i=b11b22-b\2
= b>0.
(4.14)
= 0,
(4.15)
Thus (4.13) can be rewritten as 2 f [pH*dA+ [[ — (S^eJdA where
s
s J
=b*11b22-2b*12b12+b*22bll = 2b-
b*n-b11
b*12-b12
b 2\—b2l
b
22
(4.16)
— b22
In particular, when S* and S are identical, J = 2b and H* = H, and therefore (4.15) takes the form lU
pHdA+2^
f—(S 1 .e 3 )dA = 0.
(4.17)
Subtracting (4.17) from (4.15) we obtain
2JjpH*dA-2JjpHdA = jj _1_ — (2b-J)(S1.e3)dA g
7/7
**11-*11
**12-*12
'6*21-^21
**22~*22
OV^)^-
(4.18)
Since by assumption dt.e3 is of the same sign over S, without loss of generality we may assume 51.e3>0 over S. Thus, by Lemma 1.1, the integrand on the right side of (4.18) is non-positive so that U pH*dA^ s
[[ pHdA. s
(4.19)
By interchanging the roles of S and S* in (4.3), ..., (4.19) we can have
| J pHdA ^j(pH*dA. s
(4.20)
s
Combination of (4.19) and (4.20) yields f f pHdA = j jpH*dA, s s
(4.21)
from which it follows that the determinant on the right side of (4.18) vanishes, and
A GENERALIZATION OF THE RIGIDITY THEOREM OF COHN-VOSSEN
565
therefore by Lemma 1.1 we obtain &*n =£11. b*l2=bl2,
b*12=b22.
Hence the proof of Theorem 1.1 is complete. References 1. S. Cohn-Vossen, " Zwei Satze ttber die Starrheit der Eiflachen ", Nachr. Ges. Wiss. Gbttingen (1927), 125-134. 2. L. P. Eisenhart, Riemannian geometry (Princeton University Press, Princeton, 1949). 3. G. F. Feeman and C. C. Hsiung, " Characterizations of Riemann n-spheres ", Amer. J. Math., 81 (1959), 691-708. 4. R. G. Gardner, " Subscalar pairs of metrics and hypersurfaces with a nondegenerate second fundamental form ", J. Differential Geometry, 6 (1972), 437-458. 5. V. G. Grove, " On closed convex surfaces ", Proc. Amer. Math. Soc, 8 (1957), 777-786. 6. G. Herglotz, " Uber die Starrheit der Eiflachen ", Abh. Math. Sem. Univ. Hamburg, 15 (1943), 127-129. 7. C. C. Hsiung and S. S. Mittra, " Isometries of compact hypersurfaces with boundary in a Riemannian space ", Differential Geometry, in honour of K. Yano (Kinokuniya, Tokyo, 1972), 145-161. 8. A. W. Tucker, " On generalized covariant differentiation ", Ann. of Math., 32 (1931), 451-460. 9. C. E. Weatherbum, An introduction to Riemannian geometry and the tensor calculus (Cambridge University Press, Cambridge, 1950).
Lehigh University National Tsing Hua University, and Academia Sinica.
CHUAN-CHIH HSIUNG AND LARRY R . MUGRIDGE
EUCLIDEAN AND CONFORMAL
INVARIANTS
OF SUBMANIFOLDS
1. I N T R O D U C T I O N
Let x: Mn -*• Nn+m be an n ( ^ 2)-dimensional submanifold immersed in an (« + m)-dimensional Riemannian manifold Nn+m. For simplicity we shall write x(Mn) as M " . Let gtj be the Riemannian metric tensor on Mn induced by the immersion x, e a unit normal vector of M" at a point x, and Qy the second fundamental tensor of Mn at x with respect to e. Then the eigenvalues Ai(e),..., hn(e) of the matrix (£lti) relative to the matrix (g (i ), i.e., the roots of the determinant equation (1.1)
det(Q„ - \gtj) = 0
in A, are called the principal curvatures of Mn at x with respect to e, and the rth mean curvature of Mn at x with respect to e is defined to be the rth elementary symmetric function of h^e), ..., hn(e) divided by the number of terms, i.e.,
(1.2)
("W)=
2
K
r = 1, ..., n,
where I t is a binomial coefficient. It is well known (Haantjes [5]) that every conformal mapping/on a Euclidean space En+m can be decomposed into a product of similarity transformations (i.e., Euclidean motions and homotheties) and inversions {77,}. Let Mn -> En+m be an n-dimensional submanifold immersed in En+m. Then a quantity on Mn is a conformal invariant if it is invariant under all the conformal mappings of En+m, for which the center of every inversion does not lie on Mn. The purpose of this paper is to establish the following two theorems. THEOREM 1. Let x:Mn~^-Nn + m be an n-dimensional submanifold immersed in an (n + ni)-dimensional Riemannian manifold Nn + m, and {en + 1, ..., en + m} an orthonormal basis of the normal space at a point x of Mn. Then the normal vector of Mn n +m
(1.3)
2
#ife>.
a —n + 1
The work of the first author was partially supported by NSF Grant GP-43665. Geometriae Dedicata 8 (1979) 31-38. 0046-5755/79/0081-0031$01.20 Copyright © 1979 by D. Reidel Publishing Co., Dordrecht, Holland, and Boston, U.S.A.
32
CHUAN-CHIH HSIUNG A N D LARRY R . MUGRIDGE
and the scalar
(1.4)
2 " H2(ea)
are invariant under a Euclidean motion in the normal space at the point x of Mn. Moreover, if Nn+m is of constant sectional curvature, then the invariant (1.4) depends only on the Riemannian metric on Mn induced by the immersion x. The invariability of the vector (1.3) is well known, but we give its proof here just for completeness. The vector (1.3) is called the mean curvature vector of Mn in Nn+m, and its magnitude
[
Tl/2
n+m
2
o=n+l
(#i(0) 2J
is called the mean curvature of Mn in Nn+m. If (1.5) vanishes at every point of Mn, then Mn is said to be minimal. For n = 2 and Nn+m = E3, the invariant (1.4) is the well-known Gaussian curvature. H. Weyl [8] obtained some invariants similar to the invariant (1.4). THEOREM 2. Let x:Mn-+En + m be an n-dimensional submanifold immersed in an (n + m)-dimensional Euclidean space En+m, dVn the volume element of Mn at a point x, and{en+1, ..., en+m} an orthonormal basis of the normal space of Mn at x. Then
{
n +m
2
U B
[(HAea)Y - H2(ea)]\
a=n+l
dVn
J
n
is a conformal invariant ofM
in E
n+m
.
Theorem 2 is due to W.Blaschke [1] for n = 2, m = 1, and due to B.Y. Chen [2] for n = 2 and a general m. Moreover, for a compact oriented M2 by using the well-known Gauss-Bonnet formula it follows that JMa H\ dV2 is a global conformal invariant (J.H.White [9], B.Y.Chen [2]). It should be noted that for Mn in Theorem 2, S.S.Chern and J.Simons [3] have also obtained a conformal invariant, but it is of one differential order higher than that of the invariant (1.6).
2. N O T A T I O N A N D FORMULAS
Let Mn and Nn+m be Riemannian manifolds of dimensions n (>2) and n + m (m > 0) with local coordinates (u1, ..., un) and (x1, ..., xn+m) and metric tensors g{j and aAB, respectively. Throughout this paper, unless stated
INVARIANTS OF SUBMANIFOLDS
33
otherwise we shall make use of the following convention on the ranges of indices: 1 < A, B, C, ... 1 < i,j,k, ... n + 1 < a, ft, y, ...
sS n + m, < n, < n + m,
and also the convention that repeated indices imply summation over their ranges. Let x:Mn^Nn+m
(2.1)
be an isometric immersion of class C 2 , and the immersed submanifold x(Mn) be simply written as Mn. Then in local coordinates, x can be represented by (2.2)
xA = xA(ux, ..., un),
(A = 1, . . . , n + m),
and we have (2.3)
gtj = a^x^j
= <x,,, Xj},
where < , > denotes the inner product in the tangent space at a point x of Nn + m, (2.4)
xfi = 8xAldui,
or, in the vector form, (2.5)
xri = dx/dti'.
Now we introduce the generalized covariant differentiation, which is useful for studying submanifolds of Riemannian manifolds. Let TAX be a mixed tensor of the second order in the JC'S, and a covariant vector in the u's, as indicated by the capital and small Latin indices respectively. Then following Tucker [6], the generalized covariant derivative of TAX with respect to the w's is defined as O 6\ (4.0)
V TA — \ jl Bi —
i)TBiA - ,
J- VA TC V D -t- I CD1 Bix,J ~
PC TA \-0 ffc TA BD1 Cix,j — A ij * Bk >
L
where the Christoffel symbols TAC with capital Latin indices are formed with respect to the aAB and the JC'S, and those T% with small Latin indices with respect to the gtj and the M'S. It should be noted that the definition of the generalized covariant differentiation can be applied to any tensor in the w's and x's and that the generalized covariant differentiation of sums and products obey the ordinary rules. If a tensor is one with respect to the u's only, so that only small Latin indices appear, its generalized covariant derivative is the same as its usual covariant derivative with respect to the w's. Moreover, in generalized covariant differentiation the metric tensors aAB and gtj can be treated as
34
C H U A N - C H I H HSIUNG A N D LARRY R. MUGRIDGE
constants. Since xA is an invariant for transformations of the M'S, its generalized covariant derivative is the same as its usual covariant derivative with respect to the M'S and is the same as its usual partial derivative with respect to the M'S, SO that (2.7)
V,x = x , = dx/du*.
Let {en+1 en+m} be an orthonormal basis of the normal space at a point x of Mn. Then we can have (see, for instance, [7, Chapter X])
(2.8)
V,*,, = 2 * W « , a
(2.9)
Q.,„ =
(2.10)
Vtea = - Qa]ikgkix,f + 2 */i«u««.
where (gif) is the inverse matrix of (gtj), and (2.ii)
aaW = n.lH,
(2.12)
&aBi, + &Baii = 0,
so that &aaH = 0. Thus being defined to be —
H ( 0 = £!„,„ djAiii*.
The equations of Gauss of Mm in Nn+m are (see, for instance, [4, p. 162]) (2.14)
Rhljk
— 2^ (Qtt\hk&aHJ a
~ &tt\hj&a\ik)
+
RABCD^hX~i*,jX~k»
where the Riemann symbols Rhm = gMRlm for Mm formed with respect to the gtj and the M'S are defined by O 1^ (Z.ID)
npft Dh _ 'I A ijk — -Qy£
aph
and the Riemann symbols RABCD for Nn+m formed with respect to the a^ and the x's are similarly defined. In particular, if Nn+m is of constant sectional curvature K, then (2.16)
RABCD
= K(aADaBC - aAcaBD),
and therefore (2.14) reduces, in consequence of (2.3), to (2.17)
Rhm - y (Qa\hk&a\ii - iia|Wfi«|(fc) + K(ghkgti - ghigik)-
INVARIANTS OF SUBMANIFOLDS
35
3. PROOF OF THEOREM 1
Let {en+1, . . . , en+m} be another orthonormal basis of the normal space at a point x of the submanifold Mn of the manifold Nn+m. Then we have n+m
(3.1)
ea =
2
c e
« »>
<* = n + 1, ...,n + m,
c
«*«'
<x = n + I, ...,n + m,
»=n + l n+m
(3.2)
ea =
2 lfn
+1
where (eg) is an orthogonal matrix so that n+m
(3.3)
^
<** = 8«>
«,j8 = n + 1,...,« + w,
* » = 8»>
«,|8 • = « + 1, . . . , » + m.
(3.4)
2
From (2.9) and (3.1) it follows that the second fundamental tensor 0.a]ii, corresponding to the unit normal vector ea, are given by n+m
(3.5)
Qa|W =
2
<%&»&> « = n + I, ...,n + m.
»=n + l
Let G be the determinant of (gw), ghiJk a second minor of G whose elements are common to the hth row, ith row, jth column and kth column in G, and gM.ik t jj e a ig e b ra i c complement of the minor gMJk of G, i.e., (— 1)*+'+'+* times the determinant of order n — 2 which remains when the rows and columns containing ghUk are deleted from G. Then by Laplace's development of a determinant we have (3.6)
G = 2 *«.**"•* t
for any two fixed distinct h, i which are not summed. Thus from (1.1) and (1.2) it follows that n
(3.7)
nH^eJ = 2
(3.8)
(?W(e a ) = I
Q
«i«£">
2
a = n + I, ...,n + m,
(Q.i«n«i* - n.l«Q«i«)«M-*, a = M + 1, . . . , « + W.
36
C H U A N - C H I H HSIUNG A N D LARRY R.
MUGRIDGE
By using (3.1), (3.7), (3.5), (3.4) we obtain nn
n+m
2
nn + +m m
«#i(e«)ea = 2
a=n + l
// n+m n+m
2
2
ij = l ntv = n + l\a=n n
=2
\
c*cl)Q7Mg»e, +l
1
n+m
n +m
2 v&wfe*-
2 «^i(e>„,
so that (1.3) is independent of the choice of the orthonormal basis l^n + l> • • • > ^n + mi-
Similarly, from (3.8), (3.5), (3.4) it follows that n+m
/
\
2 (aW*-) a =n + l
<*/ 1
-G u
n
n +m
2
2
h
/ n+m
2 <**) \a=n + l x
1
n
= VJr
2
\
/
("«|«^v|ifc
—
^Mhk^vw)S
'
n +m
2
h,i.y,fc = l «,v = n + l ft
W i f c A i * - Qm^Qvw)*"-*
= 2 KWo, i=n + l
\~
so that (1.4) is independent of the choice of the orthonormal basis The last statement of Theorem 1 is readily shown by substituting (2.17) in (3.8) and noticing that RMjk depends only on the metric gpq of Mn. Hence the proof of Theorem 1 is complete.
4. P R O O F OF THEOREM 2
Let gxj be the Riemannian metric tensor on Mn. The first fundamental form of Mn at a point x is defined to be (4.1)
I = dx • dx,
where dx is a vector-valued linear form on Mn, and the dot denotes the inner product of two vectors in En+m. From (4.2)
dx = x ( du\
(4.1) and (2.3) with aAB = SAB it follows that (4.3)
l = gijdutdu>.
INVARIANTS OF SUBMANIFOLDS
37
Since in En+m the generalized covariant differentiation is the same as the usual covariant differentiation, we have (4.4)
dea = V,e8d««,
and by (4.2), (4.4) and (2.10) we can write the second fundamental form of Mn at x with respect to ea as (4.5)
II(e„) = - d x . d e a .
Thus the eigenvalues Ai(ea), . . . , hn(ea) of the matrix (ClaH]) relative to the matrix (gw) defined in Section 1 are the eigenvalues of II(e„) relative to I. By Theorem 1, (1.6) is invariant under a Euclidean motion. It is easy to see that (1.6) is also invariant under a homothety x - > p x , where p is a nonzero constant. Hence it suffices to show that (1.6) is invariant under an inversion 77 on En+m, whose center does not lie on the submanifold Mn. Choose the center of the inversion -n to be the origin of a coordinate system in the Euclidean space En+m, and let x, x be the position vectors of a pair of corresponding points of the submanifold Mn and its image submanifold Mn under IT. Then the definition of an inversion implies (4.6)
x = {c2r~2)x,
r2 = x • x,
where c is the radius of the inversion -n. By (4.6) we readily obtain (4.7)
dx = (c2c2) dx - 2(c2r ~3 dr)x,
(4.8)
dx • dx = (cV- 4 ) dx • dx.
From (4.7) it is easy to see that (4.9)
ea = 2r~2(x • ea)x - ea,
a = n + 1, . . . , n + m,
from an orthonormal basis of the normal space of Mn at x. By means of (4.7) and (4.9) we obtain (4.10)
dx • dea = 2c2r ~\x • ea) dx • dx - (c2r~2) dx • dea
and therefore, in consequence of (4.8), (4.11)
dx • dett + Adx • dx dx • dea + l-2(Xr2ea)
- A p \ dx • dx
Let dVn be the volume element of the submanifold Mn at a point x, and ^i(e«)> • • • >fin(e«)the principal curvatures of Mn at x with respect to ea. From (4.8), (4.11), (4.1) and (4.5) it follows that (clr)2»dVn,
(4.12)
dVn =
(4.13)
Uett) = -c-2r2ht(ea)
- 2c" 2 (x • ea),
/=!,...,«.
38
C H U A N - C H I H HSIUNG AND LARRY R. MUGRIDGE
Thus by (1.2) and its corresponding equation for Mn we have (4.14)
HAea) = -c-^H^e.)
(4.15)
H2(ea) = c-^HzieJ
- 2c~2(x • ea), + Ac^r\x
• e^H^e*) + Ac'Xx • eaf,
which imply immediately (4.16)
(H^))2
- H2(ea) = c-V*[(//i(0) 2 - H2(ea)].
Hence using (4.12) we obtain [(^ife)) 2 - H2(ea)] \
2 a=n+l
{
dVn
J n+m
2 a=n+l
~\n!2
[(Hx(ea))2 - H2(ea)}\
dVn,
J
and Theorem 2 is proved. It should be noted that a hypersphere of En+1 has vanishing invariant (1.6) since the principal curvatures Ai(e), ..., hn{e) of the hypersphere with respect to the unique unit normal vector e at every point x are equal. BIBLIOGRAPHY 1. Blaschke,W., Vorlesungen iiber Differentialgeometrie. Ill, Springer, Berlin (1929). 2. Chen.B.Y., 'An Invariant of Conformal Mappings', Proc. Amer. Math. Soc. 40, 563-564 (1973). 3. Chern,S.S. and Simons,J., 'Some Cohomology Clases in Principal Fiber Bundles and their Application to Riemannian Geometry', Proc. Nat. Accad. Sci. U.S.A. 69, 791794 (1971). 4. Eisenhart.L.P., Riemannian Geometry, Princeton University Press, Princeton (1949). 5. Haantjes, J.,' Conformal Representations of an n-dimensional Euclidean Space with a Non-definite Fundamental Form on Itself, Nederl. Akad. Wetensch. Proc. Ser. A, 40, 700-705 (1939). 6. Tucker.A.W., 'On Generalised Covariant Differentiation', Ann. of Math. 32, 451-460 (1931). 7. Weatherburn,C. E., An Introduction to Riemannian Geometry and the Tensor Calculus, Cambridge University Press, Cambridge (1950). 8. Weyl.H., 'On the Volume of Tubes', Amer. J. Math. 61, 461-472 (1939). 9. White,J.H., 'A Global Invariant of Conformal Mappings in Space', Proc. Amer. Math. Soc. 38, 162-164 (1973).
Authors' addresses: Chuan-Chih Hsiung Lehigh University Bethlehem, Pa. 18015, U.S.A.; and Larry R. Mugridge Kutztown State College Kutztown, Pa. 19530, U.S.A. (Received July 1, 1977)
CHUAN-CHIH HSIUNG A N D KYU SAM PARK
SOME UNIQUENESS THEOREMS ON TWO-DIMENSIONAL R I E M A N N I A N M A N I F O L D S I M M E R S E D IN A GENERAL EUCLIDEAN
SPACE
(Dedicated to Professor Buchin Su on his 80th birthday)
1.
INTRODUCTION
Let M be a differentiable manifold of dimension n 5s 2, and X: M„ -> £ , m a differentiable map of Mn into a Euclidean space En + m of dimension n + m with m > 0. The map X is called an immersion, and Mn, or rather Mn together with map X, is called an immersed submanifold of £„ + m if the functional matrix of X is of rank n everywhere. When m = 1, an immersed submanifold Mn of the space En+m is called an immersed hypersurface. Let us consider now an oriented immersed submanifold Mn. For each point PtMn there is a unique linear space N of dimension m normal to X(Mn) at the point X(P). For any unit normal vector er(P) in the normal space N at the point X{P), we define the first fundamental form I, which is positive definite, and the second fundamental form H r associated with er{P) to be (1.1)
I = d*dX,
II r = dX-de r ,
where dX and de are vector-valued linear differential forms on M , and the dot denotes the scalar product of two vectors in En+m. The eigenvalues h^ej,..., hn(er) of II r relative to I are called the principal curvatures of the manifold Mn associated with er(P\ and the fcth mean curvature Hik)(er) of Mn at X(P) with respect to er(P) is defined to be the kth elementary symmetric function of h{{ej,..., hn{er) divided by the number of terms: (1.2)
f"W>(er)= W
where I
I
hh{er)...hik(er),
(k = 1,... ,n),
ii<-
J is a binomial coefficient. In particular, we have the first and wth
mean curvatures H(1)(er) = (l/«)I? =1 /j i (e r ) and Hw(er) = /t,(e r )... hn(er); the latter is also called the Gauss-Kronecker curvature of Mn with respect to er(P), which is simply the Gaussian curvature when m = 1. It is well known that for a manifold M„ immersed in an E . for any n
n+ m
m > 0, the normal vector n+ m
(1.3)
£
tf(1>(e>r
r = n+ 1
Geometriae Dedicata 12 (1982) 35-51. 0046-5755/82/0121-0035$02.55. Copyright © 1982 6y D. Reidel Publishing Co., Dordrecht, Holland, and Boston, U.S.A.
J
36
C H U A N - C H I H H S I U N G A N D KYU SAM P A R K
is invariant under a Euclidean motion in the normal space N of Mn at the point X(P). The invariant normal vector (1.3) is called the mean curvature vector of Mn at the point X, and its magnitude n+m
(1.4)
£
[(H (1) (e r )) 2 ] 1/2
is called the mean curvature of Mn at the point X. Recently, C. C. Hsiung and L. R. Mugridge [6] extended the Gaussian curvature of a M 2 immersed in E3 by proving that the scalar n+m
(1.5)
£
H™(er)
r = n+l
is invariant under a Euclidean motion in N at the point X. The new invariant (1.5) will be called the generalized Gaussian curvature of Mn at X. A normal vector er is said to be parallel in the normal bundle of Mn if der is tangent to Mn everywhere. For simplicity we shall write Hr and Kr for the first mean curvature Hw(er) and the Gauss-Kronecker curvature H{n\er) with respect to er, respectively. Let MN be a compact oriented Riemannian manifold immersed in a Euclidean space E , . By a normal frame Xe ,, ... e , on the manifold M r
n+m
J
n+i
n+m
n
we mean a point X of the manifold Mn and an ordered set of mutually orthogonal unit vectors en+l,... ,en + m normal to the manifold Mn at the point X. The manifold Mn is called a star manifold if there exists a point 0, called a pole, in E„, „ and class C2 field of normal frames Xe . , ... e„. over the *^
7
it
T
nt
n
T
1
it
T
m
manifold M n such that the Gauss-Kronecker curvature H{n\er) of Mn and the support function X-er with respect to the pole 0 are positive for every normal vector er, n + 1 < r ^ n + m, at every point of the manifold M . This normal frame Xe„, . ...e ,„ is called a fundamental normal frame of the star manifold Mn at the point X. An n-dimensional star manifold with boundary is an rc-dimensional compact subset of an n-dimensional star manifold. In 1903, Minkowski [7] established his well-known uniqueness theorem on compact convex surfaces in E3, which states that a diffeomorphism between two compact convex surfaces in E3 such that the two surfaces have parallel inner normal vectors and the same Gaussian curvature at every pair of corresponding points is a translation in E3. Since then several authors have given different proofs of this uniqueness theorem (for example, see [4] for a brief history). In 1957, S. S. Chern [1] proved the theorem by deriving some integral formulas. In 1958 C. C. Hsiung [4] extended the uniqueness theorem to convex surfaces with boundary. In 1959, S. S. Chern [2] gave a proof of Minkowski's uniqueness theorem for convex hypersurfaces in En+1 for any n>2.
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
37
The purpose of the present paper is to extend Minkowski's uniqueness theorem to two-dimensional Riemannian manifolds M 2 immersed in E2 + for any m > 0 and to obtain a new uniqueness theorem for the immersed submanifolds M2 as follows (for generalizations of Christoffel's uniqueness theorem which is similar to Minkowski's theorem see [5]): THEOREM I. Let M2 and M* be compact oriented two-dimensional Riemannian manifolds immersed in a Euclidean space E2 + m of dimension 2 + mfor any m>0. Suppose that (i) M2 is a star manifold with respect to a fundamental normal frame e3,...,e2 + m, (ii) them—\ normal vector fields e3,... ,e2 + m-i are parallel in the normal bundle ofM2 and (iii) the mean curvatures of M2 with respect to the normal vectors e3,...,e2 + m are all equal, that is, Hr = Hs(r,s = 3,... ,2 + m). If there exists an orientation-preserving diffeomorphism f of M2 onto M* such that at each pair of corresponding points (iv) M2 and M* have the same normal frame e3,... ,e2 + m, (v) M2 and M* have the same generalized Gaussian curvature, that is, ^r = ^Kr — S^ 3 m X*, where K* denotes the Gauss-Kronecker curvature of M* with respect to e , (vi) the mean curvatures of M* with respect to the normal vectors e3,..., e2+m are all equal, that is, H* = H* (r,s = 3,... ,2 + m), and (vii) the support function X*er of M\ with respect to the same pole 0 is positive for r=3,... ,2 + m, thenf is a translation in E2 + m. It is clear that Theorem I with m = 1 is Minkowski's uniqueness theorem for surfaces in E3. THEOREM II. Let M2 and M* be compact two-dimensional oriented Riemannian manifolds immersed in a Euclidean space E2 + mfor any m > 0. If there exists an orientation-preserving diffeomorphism f of M 2 onto M* such that at each pair of corresponding points M2 and M* have (i) a common normal frame e3,... ,e2 + m, (ii) the same mean curvature vector, that is, ^=™Hrer — ^ J T ^ r * ^ ' an( ^ (iii) the same nonzero generalized Gaussian curvature, that is, 'L**™Kr = = £,2=+3mK* + 0, and (iv) there exists a positive integer r (3 ^ r < 2 + m)for which Kr>0 throughout M2, then f is a translation in E, , .
38
C H U A N - C H I H H S I U N G A N D KYU SAM P A R K
It should be remarked that the manifold M2 in Theorem II need not be a star manifold with respect to the normal frame e3,... ,e2 + m. 2. I M M E R S E D S U B M A N I F O L D S IN E U C L I D E A N SPACE
Suppose that a Euclidean space E2 + m is oriented. By a frame Xel...e2 + m'm the space E2+m we mean a point XeE2+m and an ordered set of mutually orthogonal unit vectors ex,..., e2 + m with an orientation coherent with that of the space E2 + m so that the determinant \e{,..., e2 + m\ is equal to + 1 . To avoid confusion we shall use the following ranges of indices hereafter: (2.1)
Ka,j8<2,
3 < r , s , t<2 + m,
l^i,j,k^2
+ m.
Then we have (2.2) ei-ej = 5iJ, where d{, are the Kronecker deltas. Let F(2, m) be the space of all frames in the space E2 + m, so that dimF(2,m) = ~(2 +m)(2 + m+I). In F(2,m) we introduce the linear differential forms
(2.3)
dA" = 5 > k ,
de ( =5>iA.
i
J
where (2.4)
0 ^ + 0^ = 0.
Since d(dX) = 0 and d(de,.) = 0, from (2.3) we have (2.5)
do; = Xco'j A co'p,
dm'.j = £ 0 4 A
J
k
where A denotes the exterior product. As explained in Section 1, an immersed two-dimensional submanifold in the space E2 + m is an abstract manifold M2 and a differentiable map X: M2 -* E2 + m such that the induced map X on the tangent space is injective everywhere. Analytically, our map can be defined by a vector-valued function X(P), PeM2. Our assumption implies that the differential dX(P) of X{P), which is a linear differential form on M2 with value in E2+m, has as values a linear combination of exactly two vectors tx, t2. Since X^ is injective, we identify the tangent space of M 2 at the point P with the plane formed by tl,t2. A linear combination of the vectors tl,t2 is called a tangent vector, and a vector orthogonal to t1 and t2 is called a normal vector. The immersion of M2 in E2 + m gives rise to a bundle B, whose bundle space is the subset of M2 x F(2, m) consisting of (P,X(P)eie2e3
...e2
+ m)eM2
x F(2,m)
such that el,e2 are tangent vectors, and e3,..., e2 + m are normal vectors at the point X(F).
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
39
Consider the inclusion map <>/ and the projection p: B -t M 2 x F{2, m) ^ F(2, m). By putting (2.6)
coi = (p4>)*uj'.,
a>iJ = {p(l))*a>'ij,
from (2.4) and (2.5) we have (2.7)
o>i.. + o;.i = 0,
(2.8)
dco; = ^oi. A co,,,
dco0. = X > ; ) i A cokj..
J
k
Here co; and (ou are linear differential forms in the bundle B. From the definition of the bundle B it follows that (2.9)
(or = 0 (r = 3, . . . , 2 + m)
and that co, and co2 are linearly independent. Thus for the immersion X: M2 -> ->£ 2 + m wehave 2
(2.10)
2+m
dX=5>ae«,
de ; = £ co; .e..
From (2.8) and (2.9) it follows that EcoaACOra
= 0
(r = 3, . . . , 2 + m),
from which by Cartan's lemma on exterior algebra we have .(2.11)
<»„ = ! > „ / » „ >
Kv = K^
(r = 3,...,2 + m).
If the determinant det (^ra/)) =^ 0 for some r, by introducing the matrix (Ara(() inverse to the matrix (Ariiji) we have (2.12)
<^ = I > m / ) « V
*nt = Kik
(f = 3 , . . . , 2 + m).
By means of (2.2), (2.10), (2.11) and (2.12), Equations (1.1) can be written as (2.13) a
at,/?
a,/?
Thus by the definitions of Hr and X r we obtain (2.14)
Kr = AniAr22
(2.15)
Hr = %Arll
- An2Ar2l
= det(,!„,) = l/det(Aa/IX
+ /l r22 ) = |trace(/l m / ) ).
Through a point in a Euclidean space E2 + m\et A^,... ,A2 + m_lbe2 + m — — 1 differentiable vector functions of 2 variables ux,u2, and let J be any
40
C H U A N - C H I H H S I U N G AND KYU SAM PARK
vector. The scalar product of the vector J and the vector product Ax x ... x ^2 + m-i °f the vectors Alt... ,A2 + m_l is given by (2.16)
J-(Alx...xA2
+ m_1)
= (-ir-1\J,Al,...,A2
+ m_1\,
where \j,Alt... ,A2+m_l\ is a determinant, the elements of each of whose columns are the contravariant components of the vector indicated. It is obvious that the vector Al x ... x A2+m_1 is orthogonal to each of the vectors A1,... ,A2+m_l and that an interchange of any two vectors in the vector product changes the vector product only by a sign. Thus we have (2.17)
etx
... xerx
... x e2 + m = ( - l)m + rer,
where the circumflex over er indicates that the vector er is to be deleted. The vector product of vectors and the exterior product of differential forms can be combined to define a convenient operation x as follows (for example, see [3] or [5]): At x ... x Ai_l x dAi x A.+ l x ... x AJj_' -l i x dA.xi A j+i x A
(2.18)
A
.
x ... x A2 + m_1 =(Al x ... x At_t x Aixt x Aj+l
x Ai+l
x ... x Aj_x x AjAj
c
x
J
x ... x A2 + m_l)du " A du" ,
where a repetition of an index means summation over the range of the index, a ; , a.. = 1,2, and At x. = dA./du*'. It is readily seen that the vector (2.18) is independent of the order of the vectors dAt, d,4.. Let dA be the area element of an immersed submanifold M2 in the space E2+m. Then by the means of the operation x we obtain (2.19)
d Z x d X x e 3 x ... xerx
... x e2 + m = ( -
l)m+r2erdA.
On the other hand, by (2.10), (2.17) and (2.18) we have dX x dX x e3 x ... x er x ... x e2 + m= (— l)m+r2er<x>1 A a>2. Comparison of this equation with (2.19) yields (2.20)
dA = a>1 A o 2 .
From (2.11), (2.14), (2.15) and (2.20) it follows that (2.21)
KrdA = corl Acor2,
(2.22)
2Hr dA = con A co2 - cor2 /\ col
(r = 3,... ,2 + m).
3. I N T E G R A L F O R M U L A S FOR A P A I R OF IMMERSED M A N I F O L D S
Let M be a compact differentiable manifold of dimension 2, and let M2 and M* be compact oriented immersed manifolds in E2+m given by X:M-> ->E 2 + m and X*:M - » E 2 + m , respectively. Then Section 2 can be applied
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
41
to the manifold M2, and for the corresponding quantities and equations for the manifold M* we shall use the same symbols and numbers with a star respectively. Suppose that there is an orientation-preserving diffeomorphism / of M 2 onto M* such that at each pair of corresponding points M 2 and M* have a common normal frame and hence parallel tangent spaces. Without loss of generality we may assume that (3.1)
e* = et
( i = l , . . . , 2 + m).
From (2.10), (2.10)* and (3.1) it follows that (3.2) < = fflra (r = 3, . . . , 2 + m ; a = l , 2 ) . Suppose also that M2 and M* have the same nonzero generalized Gaussian curvature at each pair of corresponding points, so that
(3.3)
2X* = EKr=*0. r
r
From (3.2), (2.21) and (2.21)* we obtain (3.4) and hence
(3.5)
K*dA* = KrdA
{r = 3,... ,2 + m),
YK?dA* = HKrdAr
r
By (3.3) and (3.5) we thus have (3.6)
dA* = dA,
which together with (3.4) gives (3.7)
K = Kr
{r = 3,...,2
+ m).
For the pair of immersed manifolds M 2 and M* under the assumption that the m — 1 normal vector fields e3,... ,e2+m_1 are parallel in the normal bundle of M 2 and there exists a fixed integer t (3 < t ^ 2 + m) for which Kt > 0, we introduce a linear differential form 2+m
(18)
C 2 + m = j:{-lY+1\X*,X,dX,e3,...,er,...,e2
Zl z +m
= X(-Dr+1or. r=3
To calculate dC 2 + m we first calculate dDr as follows: dDr = d\X*,X,dX,e3,...,er,...,e2 + J(3.9)
2+m
= (l) + (li)- E(ni) s ,
+ m\-
42
C H U A N - C H I H H S I U N G A N D KYU SAM PARK
where (3.10)
(L)=\dX*,X,dX,e3,...,er,...,e2
+ m\,
(3.11)
(ll) = \X*,dX,dX,e3,...,er,...,e2
+ m\,
(3.12)
(Ul)s=\X*, X,dX,e3,...
,des,...
,e„ ... ,e2
+ m\.
By means of (2.10), (2.10)*, (3.1), (2.16), (2.17), (2.18) and (3.10) we have {I)=-\X,dX*,dX,e3,...,er,...,e2
+ m\
m
= ( - l) A--(dX* x dX x e3 x ... x er x ... x e2+m) = ( - l)mX-(el
x ... x er x ... x e2 + m)(co* A a>2 - cof A to,)
r
= ( - l) (X-e r )(a)*
ACO2-CO* A
co,).
From (2.12), (2.12)*, (3.2), (2.21), (2.14), (2.14)* and (3.7) it follows that for 3^t=^2 + m CO* A U>2 — CO* A CO,
~ (^*21<1 +
A
*2
= ( A * 1*,22 ~}*XlKl\
(3.14)
2
0
A
U , l 1W,1 +
A
,12 W ,2)
"/i,*21A,12+A*22/l,n)Kl
A
" ^
= [ - (A* , - Arl ,)(A* 2 - A,„) + (/* 2 - ^ , 2 ) ^ 2 1 - ;-,2.) + + Wl 1 *,*22 - ;-M 2^*2 1 ) + ( ; ",. 1 *,22 - K 1 2*,2 1 ) ] *
x (co„ A cor2) = [ - dettf*, - A,.,) + det(x^) + + det (/,„)] K,dA = - det(/*„ - X,„)K,dA + 2dA. Substituting (3.14) in (3.13) gives (3.15)
(I) = ( - 1)' + > p r det {**, - A ^ K , dA + ( -
l)r2PrdA,
where p r = X e r , and 3 < r, t < 2 + m. By means of (2.10), (2.16), (2.19) and (3.11) we obtain (3.16)
(II) = ( - l)mX*-( - l) m + r 2e r dA = (- \)r+l2p*
dA,
where p* = X*-er. The vector-valued functions X and X* can be written
(3.17)
2+m
2+m
Jf=[Vi.
X * = £ **<>,..
i= 1
i=l
If e 3 , . . . , e2+m _, are parallel in the normal bundle of M 2 , then from the definition it follows immediately that (3.18)
ct)„ = 0
(s,r = 3,...,2 + m),
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
43
so that (3.19)
d e s = £ co„ea (s = 3,...,2 + m). a=l
Using the above method together with (2.22), (3.12), (3.17) and (3.19), we can calculate (III)s for s < r as follows: (III)s = ( - l)m + sX*-(X x dX x des x e3 x ... x es x ... x erx x ... x e 2 + m)-_= ( - l ) m + sA'*-[Ars(co1 A t o s 2 - f f l 2 A c o s l ) ( - l ) s - 1 e 1 x (3.20)
x ... x e r x ... xe1
+m
+ Xr(co1 A co s2 _
2
-co2Aa>sl){-\y~ e1
x ... x es x ... x e2 + J
=
+1
= ( - 1)' (X***,-)• [(*/, - W K A ws2 i
-W2 r
= ( - l) 2(Xs*Xr - X*Xs)Hs
AWsi)] =
dA.
A similar calculation yields the same expression of (III)s for s > r. Therefore for s 4=- r we obtain (3.21)
(lll)s = {-lY2{X*Xr-X*Xs)HsdA
(s = 3,...,2+ m).
Substituting (3.15), (3.16) and (3.21) in (3.9) gives dDr = ( - 1)-+ % det (A*„ - XtJKt r
(3.22)
+ ( - l) 2p r dA + ( - l)
r+ x
dA+ 2p* dA+
2+m
+2
Y ( - i) r+ ' w * , - X:XS)HS dA, s=3
which together with (3.8) implies 2+m
dC 2 + m = E P r d e t ^ - A ^ c U r=3
2+m
(3.23)
- 2 ^
2+m
P,dA + 2YP?dA
r=3 2+m2+m
+ 2 Z
+
r=3
X (Xs*Xr-Z*Zs)HsdA
r= 3 s= 3
Since 2+m 2+m
Z I (*,**,-*,**,)/*,
s=3
2+m
= Z (X*X - X * X , ) ( H - H r ) d A ,
44
C H U A N - C H I H H S I U N G A N D KYU SAM P A R K
we can rewrite (3.23) as 2+m r=3
2+m
(3-24)
2+m
- 2 X pr
r=3
+2 X
{X;Xr-X?X,)(Ht-Hr)dA.
r,s = 3 s>r
By interchanging the roles of M2 and M* in (3.24) we obtain 2+m
dC2%m= E ^ d e t ^ - A ^ ) / ^ r=3 2+m
(3-25)
2+m
- 2 X pr*d^ + 2 J P r d i + r=3 2+m
r=3
+2 E
(XsX*-XrX:)(Hf-H*)dA,
r,s= 3 s>r
where 2+m
(3.26)
C* +m = I
(-lrMA-.A-'.dJf*,^,...^,...,^^!.
r=3
Addition of (3.24) and (3.25) yields 2+m
d(C 2+m + C*2+m) = E (pr + p*) det (A*, - Xm)Kt AA + r=3 2+m
(3.27)
+2 £
(x*X,-X r *X s )((tf s -tf r )-
r>s= 3 s>r
- (H*-//*)) d4. Integrating the differential form (3.27) over the compact manifold M2 and using Green's theorem, we obtain the integral formula r 2+m
I J
(p, + Pr*)det(A*/I-AlB/,)K,(L4 +
r=3
(3.28) +2
r 2+ + mm /-2 E (XfXr - X*Xs){{Hs - Hr) - {H* - H*)) dA = 0. =3 J r,s r,s=3 Mi s>r
In order to derive an integral formula for a pair of manifolds M2 and M | satisfying the conditions of Theorem II, we introduce a linear differential
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
45
form (3.29) where (3.30)
A2 + m = \e3,...,e2
+ m,W,dX\,
W = X* - X.
Then we have 2+m
(3.31)
d , 4 2 + m = X (i)r + (ii), r=3
where (3.32)
(i\ = \e3,...,de„...,e2
(3.33)
(ii) = \e3,...,e2
+ m,W,dX\,
+ m,dW,dX\.
From (2.10), (2.16), (2.17), (2.18) and (2.22) we can easily obtain (3.34)
(i\=-2(W-er)HrdA.
By substituting (3.30) in (3.33) we can write (3.35) where
(ii) = (ii) a -(ii) 6 , (Z)a =
(336)
\e3,...,e2+m,dX*,dX\,
(ii)iJ = | e 3 , . . . , e 2 + m , d X , d X | .
Using (2.10), (2.10)* and (3.14) we can readily have (")a = ( - l ) 2 + " - 1 v' 3l > 4 * - * e2+m * AX* * dX] = L^4 (3.37)
= dtt{X^-Xuf)KtdA (3.38)
+ 2dA,
(ii)» = 2 ( U .
A combination of (3.31) with (3.34), (3.35), (3.37), (3.38) gives 2+m
(3.39)
d,42 + m = - 2 X (W-e r )H r dA - det (A*, - ltJKt
r=3
By integrating the differential form (3.39) over the compact manifold M 2 we obtain the integral formula C 2+m
X (W-er)Hr
(3.40)
det(A* / J -A M )K r d^ = 0.
r=3 M2
Similarly, we introduce the linear differential form (3.41)
A*2 + m = \e3,...,e2
+ m,W,dX*\,
46
C H U A N - C H I H H S I U N G AND KYU SAM P A R K
and obtain its differential 2+m
(3.42)
dA*+m = - 2 X {Wer)H* dA + det (A*, - X^K,
dA,
r=3
and therefore the integral formula 2+m
(3.43)
2 £(WAer)Hr*d^-
det(A* -A t e .)K r cW = 0.
r=3
Subtracting (3.40) from (3.43), we thus arrive at the integral formula 2+m
X
(3.44)
(Wer)(H*-Hr)dA-
dct{X^-XJKtdA
= Q.
Mi
4. P R O O F S O F THEOREMS I A N D II
We need the following elementary lemma in the proofs of Theorems I and II: LEMMA 4.1. Let (4.1)
ax + 2bxy + cy2,
a*x + 2b*xy + c*y2
be two positive definite quadratic forms such that (4.2)
ac-b2
= a*c* - b*2.
Then (4.3)
•a
b*-b
b
c* — c
^0,
where the equality holds if and only if the two forms are identical. Lemma 4.1 is well known, so its proof is omitted here. Proof of Theorem I. Because of assumptions (i), (ii), (iv), (v) in Theorem I we have the integral formula (3.28), whose second term vanishes due to assumptions (iii) and (v) so that (3.28) is reduced to 2+m
(4.4)
I
(P, + P:)det(A* -XtJKtdA
= 0.
On the other hand, assumptions (i) and (vii) immediately imply that over the whole manifold M2,T,2+™(pr + p*) > 0 and Kt > 0, which together with (4.3) show that the integrand of the integral (4.4) is nonpositive, and therefore that (4.4) holds if and only if (4.5)
det(A* /I -A tt/ ,) = 0.
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
47
Thus, by Lemma 4.1 again we obtain (4-6)
4*, = ^
(0,0=1,2).
From (4.6), (3.2), (2.12), (2.12)*, (2.10), (2.10)* it follows readily that (4.7)
d{X* -X)
= 0.
Hence X* — X is a constant vector, that is,/is a translation. Proof of Theorem II. Because of assumptions (i), (iii) and (iv) in Theorem II we have the integral formula (3.44), whose first term vanishes due to assumption (ii) so that (3.44) is reduced to (4.8)
det(A*txP -XtJK tdAt txP'
= 0.
Hence a repetition of the arguments used in the proof of Theorem I shows that / i s a translation. q.e.d. Theorems I and II can be easily extended to two-dimensional Riemannian manifolds with boundary immersed in a Euclidean space E2+m as follows: THEOREM I'. Let M2 and M* be compact oriented two-dimensional Riemannian manifolds, with boundaries B2 and B* respectively, immersed in a Euclidean space E2+mfor any m > 0. Suppose that there exists an orientationpreserving diffeomorphismf of M 2 onto M*, and that M2,M* and fall satisfy the conditions of Theorem I. If f restricted to B2 is a translation in E2+m carrying B2 onto B*, then f is a translation carrying the whole manifold M2 onto the whole manifold M*. THEOREM II'. Let M2 and M* be compact oriented two-dimensional Riemannian manifolds, with boundaries B2 and B* respectively, immersed in a Euclidean space E2 + mfor any m > 0. Suppose that there exists an orientationpreserving diffeomorphismf of M 2 onto M*, and that M2,M* and fall satisfy the conditions of Theorem II. If the diffeomorphismf restricted to the boundary B2 is a translation inE2 + m carrying B2 onto B*, then f is a translation carrying the whole manifold M2 onto the whole manifold M*. Proof of TheoremT. (3.28) now becomes
Since M2 has a boundary B2, the integral formula
\(C2+m + C*+m)= |
2
if(pr + pr*)det(A*/i-^)K(d^+
M2
(4.9)
C 2+m
+2
X (XfXr - X*Xs)((Hsrts= 3 Mi s>r
Hr) - {H* -
H*))dA.
C H U A N - C H I H H S I U N G AND KYU SAM PARK
48
From (3.8) and (3.26) it follows that ^ 2 + m + ^2 + m —
(4.10)
2+m
= £ (-iy+1\X*,X,dX-dX*,e3,...,er,...,e2
+ m\.
On B2, since / is a translation, we obtain dX — dX* = 0 and therefore C2+m + C| + m = 0. Thus the integral formula (4.9) is reduced to (3.28) again, and Theorem I' is proved by the arguments in the remainder of the proof of Theorem I. Proof of Theorem II'. Since M2 has a boundary B2, by using (3.39) and (3.42) we can easily see that the integral formula (3.44) now becomes (•^ 2 + m
^ 2 + m)
2+m
(4.11)
=2
£ (W-er)(Hr-H*)dA
+
r=3 M2
+2
det(A* - O K t ( L 4 . 1W2
From (3.29) and (3.41) it follows that (4.12)
A*+m -A2
+m
= \e3,...,e2+m, W, dX* - dX\.
On B2, since / is a translation, we obtain dX* — dX = 0 and therefore A*+m — A2+m = 0. Thus the integral formula (4.11) is reduced to (3.44) again, and Theorem II' is proved by the arguments in the remainder of the proof of Theorem II. 5. EXAMPLE
As an example of a Riemannian manifold M2 satisfying conditions (i), (ii) and (iii) of Theorem I we consider a two-sphere immersed in a Euclidean space E 4 . Consider a unit sphere (5.1)
S2 = {(x,y,z)\x2 + y2 + z2 = l}
parametrized by (5.2)
0(u, v) = (sin u cos v, sin u sin v, cos u),
0 < u < it, 0 < v < In.
Suppose that S2 is immersed in £ 4 by an immersion X: S2 -* E4 which is
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
49
defined by (5.3)
X{x,y,z)J^{x + y),X(x-y), "7f'7f }
Then we have (X ° cj)) (u, v) = I —- sin u (cos v + sin v), V
(5.4)
1 . , . , 1 1 —= sin u (cos v — sin t>), —= cos u, —/= cos M
Jl
J2 Jl
with partial derivatives: {X °4>)u = \ —T= cos u (cos i; + sin v), 1 1 1 . , —;= cos u(cos j; — sin v), 7= sin u, 7= sin u , '2 x/2 x/2
(5.5)
(X ° >)„= ( —?= sin u(cos v - sin u), —P= sin u(cos v + sin v), 0,0 /2 The vectors {X°4>)u and CX" 0 ^ are orthogonal to each other, and their lengths are given by (5.6)
||(A"°0)J = 1,
| | ( X ° 0 ) J = sinM.
Now we choose the unit tangent vectors el,e2 of the immersed surface X ° 0 as follows: (5.7)
ei
=(*°
and normal vectors e 3 , e 4
«2 = ^ ( * ° 0 . -
e 3 = ^(sin u(cos v + sin v), (5.8)
sin u(cos t> — sin v), cos u — 1, cos u + 1), e 4 = ^(sin w(cos v + sin u), sin u(cos v — sin t>), cos u + 1, cos u — 1).
It is easy to see that the unit vectors el,e2,ei,eA that (5.9)
\el,e2,e3,eA\
= l.
are mutually orthogonal and
50
C H U A N - C H I H H S I U N G A N D KYU SAM P A R K
Furthermore, the partial derivatives of e3 and e4 are: e
3u = « I A A
(5.10)
e3v = e2 sin
ujjl,
r
r
e
*u = eiN2>
e
A» = ei
sin
" / V 2-
Thus we have 1
s
,
i
n
"
J
1
J
s m M
J
(5.11)
de, = —f=e. du + ——e2 dv, de. = -7=e1du + -^e2dv. 3 /2 x/2 x/2 N/2 On the other hand, from (5.3), (5.4) and (5.8) it follows that (5.12)
&{X°$) = (X°
By (1.1), (5.11) and (5.12), for the immersed surface X°
A1(er) = A 2 (e r )=l/ > /2
(r = 3,4),
which together with (1.2) gives (5.15)
Kr=l/2,
Hr=l/J2
(r = 3,4).
From (5.4) and (5.8) follow the support functions X • er: (5.16)
(Xo
(X°4>Ye4 = 1/^/2.
Hence the two-sphere S2 immersed in EA satisfies condition (i) of Theorem I by (5.15) and (5.16), condition (ii) by (5.11), and condition (iii) by (5.15). BIBLIOGRAPHY 1. Chern, S. S.: 'A Proof of the Uniqueness of Minkowski's Problem for Convex Surfaces', Amer. J. Math. 79 (1957), 949-950. 2. Chern, S. S.: 'Integral Formulas for Hypersurfaces in Euclidean Space and Their Applications to Uniqueness Theorems', J. Math. Mech. 8 (1959), 947-955. 3. Hsiung, C. C. : 'Curvature and Betti Numbers of Compact Riemannian Manifolds with Boundary',Rend. Sent. Mat. Univ.ePolitec. Torino 17(1957-1958), 95-131. 4. Hsiung, C. C.: 'A Uniqueness Theorem for Minkowski's Problem for Convex Surfaces with Boundary', Illinois J. Math. 2 (1958), 71-75. 5. Hsiung, C. C : 'Some Uniqueness Theorems on Riemannian Manifolds with Boundary', Illinois J. Math. 4 (1960), 526-540.
TWO-DIMENSIONAL RIEMANNIAN MANIFOLDS
51
6. Hsiung, C. C. and Mugridge,L. R.: 'Euclidean and Conformal Invariants of Submanifolds', Geom. Dedicata 8 (1979), 31-38. 7. Minkowski, H. : 'Volumen und Oberflache', Math. Ann. 57 (1903), 447-495.
Authors' Addresses Chuan-Chih Hsiung, Lehigh University, Bethlehem, PA 18015, U.S.A. Kyu Sam Park, Kutztown State College, Kutztown, PA 19530, U.S.A. (Received September 8, 1980)
BULLETIN OF THE INSTITUTE OF MATHEMATICS ACADEMIA SINICA Volume 14, Number 3, September 1986
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE BY
CHUAN-CHIH HSIUNG ( » ? & ) Abstract. The purpose of this paper is to solve the long standing unsolved problem: Does there exist a complex structure on the 6-sphere?
Introduction. The notion of a complex manifold is a natural outgrowth of that of a differentiable manifold. Its importance lies to a large extent in the fact that the complex manifolds include the complex algebraic varieties and the Riemann surfaces as special cases, and furnish the geometric basis for functions of several complex variables. Its development has led to clarifications of classical algebraic geometry and to new results and problems. A fundamental problem in the theory of complex manifolds is to characterize the orientable manifolds of even dimension 2tt which can be given a complex struture. By using characteristic classes and cohomology operations on the classes, in particular, the Steenrod squaring and reduced power operations, various necessary conditions are known for a manifold to be almost complex and therefore to be complex since a complex structure is naturally an almost complex structure. These conditions suffice to show that among the even-dimensional spheres only the 2-sphere S 2 and the 6-sphere S 6 are almost complex. In fact, the absence of an almost complex structure on S4* for k ;> 1 and S2" for n ;> 4 was proved by Wu [10] and jointly by Borel and Serre [2] respectively. Kirchhoff [6], [7] has shown that if S" admits an almost complex structure, then Sn+1 admits an absolute parallelism, and Adams [1] that S"+1 admits an absolute parallelism only for Received by the editors October 11, 1985.
231
232
CHUAN-CHIH HSIUNG
[September
n + 1 = 1, 3 and 7. The result of Adams combined with that of Kirchhoff implies the results of Wu, Borel and Serre. On the other hand, for a given almost complex structure, necessary conditions are known in order that it defines a complex structure (Ehresmann and Libermann [4], Eckmann and Frohlicher [3]). These conditions are also sufficient, if the almost complex structure is analytic. The purpose of this paper is to solve the long standing unsolved problem: Does there exist a complex structure on S6? Our results are contained in the following theorem and corollaries. If an almost complex structure J on a In-manifold M (n ;> 2) is a complex structure, then THEOREM.
Riemannian
2n
(1)
Fi,' F(i' Rijizt
+ FiJ F,ti Riji.k + F, s ' FtJ Ru,tt
= 0,
for any distinct ix, iit i3, k, where Ft' and Rkijt are respectively the tensor of J and the Riemann curvature tensor of M 2 " with respect to a geodesic local coordinate system, all indices take the values 1,- • •, 2n, and the repeated indices imply summation. 1. A Riemannian 2n-manifold MiH («;> 2) of constant nonzero sectional curvature admits a complex structure if and only if Min admits a flat metric. COROLLARY
COROLLARY 2. There does not exist a complex structure on a Riemannian 2n-manifold M2n (n ^ 2) satisfying: ( a ) M2" does not admit a flat metric. ( b ) M2n is of constant nonzero sectional curvature. COROLLARY
3. There does not exist a complex structure on Sin
for « ^ 2. 1. Notation and Riemannian metrics. Let Mm be a Riemannian manifold of dimension m ( ^ 2), Us^H with gi} = #,•,• be the matrix of a positive difinite metric of Mm, and ||0'">|| be the inverse matrix of ||g, ; ||. In this section all Latin indices take the values l,---, m unless stated otherwise. We shall follow the usual tensor convention that indices can be raised and lowered by using Q'J and Qij respectively, and also that when a Latin letter appears in any
1986]
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
233
term as a subsecript and superscript, it is understood that this letter is summed over its range. Moreover, if we mutiply, for example, the components atj of a tensor of type (0, 2) by the components b,k of a tensor of type (2, 0), it will always be understood that j is to be summed. Let N be the set {l,---, m\ of positive integers less than or equal to m, and let I(p) denote an ordered subset {*i,-"» *p) of the set N for p <m. If the elements iu•••, ip are in the natural order, that is, if *\ <•••<*>, then the ordered set I(p) is denoted by Io(P). Furthermore, denote the nondecreasingly ordered p-tuple having the same elements as I(p) by Q(p)>. We shall use these notations for indices throughout this paper. When more than one set of indices is needed at one time, we may use other capital letters such as J, K, L,- •• in addition to I. At first we define (0, if (/>)> ¥>
a
ON
f'(jV(«-P)£l"'"> e
.1)
l—n>
e
J , | cJ(m-P)
/(i»K(i»-p) — P \
t
Klm-p),
On the manifold Mm, let V denote the covariant derivation with respect to the affine connection r, with components r}» in local coordinates x1,---, xm, of the Riemannian metric gtj, and let 0 be a differential form of degree p given by
0 = - J r */(,> dx1^ = 0/§(#, dxW\ P\
(1.4)
where 0/c>) is a skew-symmetric tensor of type (0, p), and we have placed (1.5) dxI(-» = dx'i A---A dx'p, the wedge A denoting the exterior differentiation. (1.6) where
rf0
= («fe)i f(>+0
dx^P+v,
Then we have
234
CHUAN-CHIH HSIUNG
(1.7)
[September
(<*# W o = ~ r «/<S?i> Vntjip).
Let T\ be a tensor of type (1, 1) on Mm, and let 9/3a?' denote t h e partial derivative with respect to x'. T h e n the Ricci identity gives (1.8)
(V, V, -VjV,)
T\ = T\ R'sji - T'i «*»„.
where /?*,-,» is the Riemann curvature tensore Mm given by
(1.9)
R\jk = dr'Jdx" - arl/dxi + r\, ri, - r'fk r\},
and satisfying the following algebraic identity: Rhijk + Rhjki + Rhkij = 0 .
(1.10)
If Mm is of constant sectional curvature R, then Rhijk =
(1.11)
R(elQii-e)Qi*),
or (1.12)
Rnjk = ff(ff** gr.y - Qkj On),
where (1.13)
72*,;* = ght
R'ijt.
If this constant R is zero, then Mm is said to be locally Euclidean or locally flat, and (1.11) becomes (1.14)
R>,Jt = 0 .
In this case, there exists a local coordinate system (x*') at every point of Mm such that the new Riemannian metric tensor satisfies Q*k(x*') = const. T h u s every coordinate neighborhood of Mm can be mapped isometrically into a Euclidean »?-space Em. Conversely, if every coordinate neighborhood of Mm can be mapped isometrically into an Em, then (1.14) holds. T h e new metric g*h is called a flat metric. 2. Complex structures.
Consider a 2«-dimensional real manifold
M2n of class C° covered by a system of coordinate h
(x ),
neighborhoods
w h e r e t h e Latin indices h, i, ,/',••-run over the range 1, 2 , - - - ,
n, 1 , 2 , • • • , » .
Introduce
in each
coordinate
neighborhood
(a?*)
1986]
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
235
x
complex coordinates (z ) defined by (2.1)
2' = ^ +
/ ^ / ,
where the Greek indices X, ft, v, X',--- run over the 1, 2,---, n. We call xh real coordinates and zx complex coordinates of a point with respect to these systems of coordinates respectively. M2n is said to admit a complex structure and is called a complex manifold if there exists a system of complex coordinate neighborhoods (zx) covering the whole manifold M2* such that in the intersection of two coordinate neighborhoods (zx) and (zx') we have (2.2)
**'= / ' ' ( * ' ) ,
Iditf'lJ-O,
where fx'(zx) are holomorphic (or analytic) functions of complex variables zl, z2,---, z", and \d%zx'\ denotes the Jacobian. When we write (2.2) in the form (2.3)
x"' = xh\xk),
h' = 1, 2,- • •, n, 1, 2,- • •, n,
xh' are real analytic functions of x" and
\dhx"'\ = \dzzv\\dxzx'\
>0,
so that a complex manifold is of class C°° and orientable. The spaces in the following examples have complex structures. EXAMPLES. 1.
The complex number space C whose points are the ordered tt-tuples of complex numbers (z1,---, zn). In this case, each of f1' of (2.2) is the identity function. 2. Any Riemannian 2«-manifold Min which has a flat metric or, equivalently, is locally Euclidean. In this case there is an isometric mapping / of every coordinate neighborhood of MiH into a Euclidean 2«-space EiH. Let Ein carry the usual complex structure so that z= (z1,---, z") represents the complex coordinate system on Ein. Then the complex coordinate system on M2" is «(/(/»)) where p e M2n. 3. StH — [P], where S2* is the standard sphere of an arbitrary radius r in an E2"*1, and P is the north pole ((),•••, 0, **). In this
236
CHUAN-CHIH HSIUNG
[September
case let / : S2" — {P) —> Ein be the stereographic projection which is known to be bidifferentiable. Then / induces a flat metric on S 2 " - {P}. 3. Almost complex structures. On a Riemannian manifold M , if there exists a tensor Ft1 of type (1, 1) satisfying m
(3.1)
F,l F,* = - ei,
(i, j , k = 1, • • •, m),
j
then Fi is said to define an almost complex structure on Mm, and Mm is called an almost complex manifold. From (3.1) it follows that the tensor Fi} induces an automorphism / of the tangent space of A/"* at each point with / 2 = — /, / being the identity operator, such that, for each tangent vector »*, J(vt)
(3.2)
= Fltv'.
Throughout this paper an almot complex structure with tensor Fi3 will be simply written as an almost complex structure Fi1. By using the multiplication of matrices, from (3.1) we readily see that a necessary condition for the existence of an almost complex structure on a Riemannian manifold Mm is that the dimension m of Mm be even, and therefore that an almost complex manifold is always orientable, and the orientation depends only on the tensor Fi'. Every complex manifold carries a natural almost complex structure. To show this, we consider the space C of tt-tuples of complex numbers («',•••, z") satisfying (2.1). With respect to the coordinate system (a?1,- • •, as", x1,- • -, a?") we define an almost complex structure / on C by (3 3)
J(3/dxlJ J(d/dxx)
= d/dxr, = - d/dxl,
x = 1,- • •, » .
From (3.2) it follows that the matrix of the tensor Fi1 in this case is given by
(3.4) where i, ] — !,"•,
W>-(-°.; '}} n, !,•••, », and X, u — !»• • •> »»•
1986]
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
237
4. Complex operators. We consider a Riemannian manifold Mm (m ^ 2) with an almost complex structure F^, and shall follow Spencer (compare [9, Chapter IX]) to introduce operators on the manifold Mm. At first we define
(4.D
n < ' = I-•(«*- v - \ ~F,*) O
1,0
1
and its conjugate tensor
(4.2)
II'' = I P = \ M + ^ ^ F,'). 0,1
£
1,0
In this section all Latin indices take the values 1,-• •, m. A simple calculation gives the following identities:
IP IP=IP> 1,0
(4.3)
1,0
1,0
IPIP=°> 1,0
0,1
ny rii*=IP0,1
0,1
0,1
Let p + o = p, p ^ 0, a > 0, set
(4.4)
"
"
*•'
•n-/»---n-/'^w.), 0,1
0,1
and define I I / c ^ * 5 to be the identity for p = a= 0 and to be zero for either p < 0 or a < 0. Then for a form 0 given by (1.4) we have
(4.5)
n> = (n>), f rf^'0^.
where
(4.6)
(n^-iw^^w
Now for a p-lorm we define the complex analogue of the real operator d defined by (1.7): 1 Throughout this paper a bar over a letter or symbol denotes the conjugate nf the complex number or operator defined by the letter or symbol.
238
CHUAN-CHIH HSIUNG
(47)
A= E
(4.8)
rf«=
[September
n d n.
D n <*np+ o=* p+S,o-l
p,o
The conjugate operators of di, d% have the forms:
(4.9)
A=
s n * n.
p + o=i> (J,a+1
(4.io)
rf2=
/>,o
E n <*n p + a=p p—l,a+%
p,a
Furthermore, for a />-form 4> given by (1.4) we define (4.11) ( 9 # W » = [(2rf» + di - d»)0]/(#+». It is known that (see [8, p. 140]) the almost complex structure Fi' on Mm is a complex structure if and only if 92 = 0 .
(4.12)
5. Calculations of d for 0- and 1-forms. In this section we shall apply the operator d to 0- and 1-forms, The calculations with respect to an almost Hermitian structure Fi3' were made by the author jointly with J. J. Levko [5, p. 397 and pp. 393-395] in 1971; the results actually hold for a general almost complex structure Fi' and the condition "almost Hermitian" is extraneous. We shall follow the notation in §4. Let (a?1,---, xm) be the local coordinates of the manifold Mm, and let V denote the covariant derivation with respect to a general Riemannian metric Qu of Mm. For any 0-form £, rf2 = rfj = 0, and di £ = II rf£ since 1,0
II is the identity operator. From (4.11) it follows that 0,0
(5.1)
(Sf),, = | - «
- -/~=T FtJ) V} f.
Now let ri be any 1-form on Mm, given by (5.2)
r, = m dx'.
In order to compute Q-q, from (4.11), (4.7), (4.8), (4.10) we first see that W-*V
8,0
0,1
8,0
1,0
1,1
0,1
0,3
1,0
for 1-foms.
1986]
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
239
Next, by means of (1.6), (1.7), (4.4), (4.5), (4.6), (1.2), (1.3) (4.1), (4.2), we obtain
n d n n=[n*,» ILV (V. I P - v, nyH+ n.i*n./^i-n' a *n.vv^ 1.1
U
Oil
1,0
0,1
0,1
»
1,0
0,1
0,1
'
1,0
0,1
+ II'.' II'.' (Vi I P - V*Ip'K~U*'°(2) 1.0
(5 4)
'
0,1
»
0,1
0,1
'
-I
k
= j
{V,^.-2 - V,-2 riil +Fil F,ti (V«y ~ Vj Vk) + V-1
[Vj (V,, FiJ - V,, Ftxi + Fil'F,1'(VkF,J-V,Ft'))
n d n n = [IP* IP' (v. I P - ** IPKa,o
L
1,0
i,o
i,o
i.o
/
+ ILV II'.' ^ 7* - V.^y) "W«w 1,0
(5.5)
i,o
\
1,0
-J
o + F * , * ( V * F , / - V.^iV)] + 2V,-, v,t -2V,-2 v.-, + 2Fil'Fiak(VkriJ - Vjvt) + V^l hyCV*. F , / - V,-, i V + Fi*F,tt(ykF,iV,Fhi)) + 2F,X' (V,-2 „, - V,- 7,,)
n <* n * = n I P ; (v» I P - v ' I P ) *> ^°c2) S,0
0,1
1,0
1,0
0,1
»
0,1
'
8
(5.6)
+ F,1* + ( V , l / V - V * F , l * ) ] + v ^T^[V,- l F,V'-V,- 2 F,- l > + F.-/F,-,' (V,F*>' - V»F,*)]} ^ 7 « C 2 ) , /
- n <* n ^- ii'.' ii'/ (v* n >s - ^ I P ) * ^ ( 2 ) 0,2
(57)
1,0
0,1
0,1
»
X,0
1,0
= |{*ylA*(v l ,F,'-r.F, i >)
'
240
CHUAN-CHIH HSIUNG
[September
H-iVCVtiV-V^iV)] + / - T Vj [V,, Ft J - ViM i V + F ^ F , , ' (V, iV - V*F,>)]}
s
' = l ^ ' ^ -
v
' . " .
+ T/-I[>,-(V,1F/-V,-2F,/)
+ F, i >V,*, 1 -F, l 'V,J7, 1 ]}«to / . w . Substituting (df),-, given by (5.1) for ??,-, in (5.8) we obtain the real part of 32
Re 3 s f - -1 [ F / (V (l Fiti - V,, F,,') + FiJ V] FiJ - Ft J V] F,-2'](V,- f) ^ V " .
By taking £ = a;* for any arbitrary k, we have V,- £ = dxk/dxi and therefore 32 £ = 0 implies that
= e*
(5.10) FjKVix Ft J - Vi, FiJ) + fij Vj F,* - Ft J Vj Ft* = 0 . Multiplying (5.10) by Fth gives (5.11)
V,, F , / - V,, F,,* = (F,-2* Ft J - Ft* Ft J) V, Fk».
It should be remarked that the tensor on the left-hand side of (5.10) is the Nijenhuis torsion tensor whose vanishing gives the integrability condition of the almost complex structure Ft'. 6. Calculations of 9 for 2-forms. In this section we shall apply 3 to 2-forms. The calculation is similar to that given in §5 but is extremely lengthy; an enormous number of terms will be involved. het c be any 2-form on Mm, given by (6.1)
C = Cv, dx'i A dx'>.
As in §5 we first see that v
(62)
8,0
2,1
0,8 '
1,1
+(2n + n)<*n + (n-n)<*n x
2,1
1,2 '
0,2
» 8,0
1,2
for
'
2,0
2-forms.
1986 |
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
241
For the symbols ej and the almost complex structure F , ' we define the following:
(6 5)
A
W
» = e'»'.'.
e \V*
fl#',-«^\',JV'*•'.*'*>.*»•
(6.6) 1 S 8
1 2 S
By means of (4.1), (4.2), (4.4), (4.5), (4.6), for a 3-form
(6.7) we obtain
,C3),
(6.8)
8(n *),.«,= ( ^w - i / : r T ^S) */
,(3),
8 ( n *),,<„ = < $; + ^ ^ cztfS) */.
(S)»
(6.9)
8 ( n *),.«, = ( < 9 'S - i / : = T ^
*'.
where {
'
gfgj = itfg] - M ,
Otfgj = £ f 8) - Cfgj,
Substitution of (6.8), (6.9) in (6.2) and use of (6.10) give (611)
8(9C) /o(3) = +ltW + vIiay
where
(6-12)
thm =4 (A'$ - V-l C$)(d I I c)/o(3),
+
/-i(B;:;:; + c;:s)](rfn^.„, J
8,0
242
CHUAN-CHIH HSIUNG
[September
Similarly for t h e 2-form C given by (6.1) we have
d n f=(d n 4.o, ** ro .
(6.U)
i,i
»
i,i
*
d n c - U n 4.c» <*»'»c8).
(6.i5) where
8 rf
( nc) ; o C 3 )
(6.16) e
v
ni
*iLv£» ^
189
(6.i7)
eA eA T Ah 8
l i s
2
) i ; i y 2 J, 18
= **:*}V[(4: ,;» - *;* e;» + # • )<•,,,,] 9
8
8
S
8 2
i
+ T/=T.;ff v» l (,#c, 1 i i ). 12 8
8 2
Xh\z and j« Y being defined by 2 8
(6.18)
8 2
xft = ft," F»," - i V . /?*,'•, 2 8
(6.19)
^
= e;* ft,'. + e[* F„tii - e'j Fk%>* - e'f * » , " • 3 2
8
2
2
3
Now for simplifying the calculations we choose geodesic local coordinates a? 1 ,•••, a!* at a general point P of the manifold Mm so that (6.20)
dgij(P)/dx"
= 0,
which is equivalent to (6.21)
r$»CP) = 0 .
On the other hand, for a tensor Xi of type (0,1) on Mm we have (6.22)
Vy Xi = dXi/dx) n
n
,
Vk V y Xi =
(6.23)
d2 Xf :
dXh rft k
dX> dX —
8x
k-
1 is
XtTli, dXn rh
dXi
—— I ik
dx>
Xi (——— rtj — /*« rjt — A* rhj). \ dxk I
n»
— 7h- J jk
dx
1986]
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
243
From (6.23) and a similar equation with * and j interchanged it follows that
(6.24)
V»(Vi Xi - Vi X,) d*x, _ d*Xj = dx' 8xh dx' dx* , dXh rn _ dXh rh , dXj „h dx1
dx'
dx"
dXi „» dxk
Now let us consider a special 1-form n of (5.2) where Vi = ej for an arbitrary k: (6.25)
v = dx*.
Then the 2-form (5.8) becomes (6.26)
dn = (. . dx'i A dx'>,
where ^ (6.27)
= 4T ^ < - ^ < + i / ^ T (Vi, F,2k - Vh Ft>
a
1
2
By using (6.20), (6.21), (6,22), (6.24) we can easily show that in geodesic local coordinates x1,---, xm at a general point P of the manifold Mm, V,t e\(P) = 0, Vy(V-l£>, - V , a £ * ) ( P ) = 0 , 2
(6.28)
1
<<* *>:-*<: ^ + ^ ) ( ^ v ' ' ^ ^ r ^ v ' . ^ < ) ( P ) = o ' (e;» e't'-e't> e'f-tfXFjJV^V,,) 2
8
8
8
2 3
-Fh*VtlVit) '1
)(J>) = 0, 2
^ ' / ' (*",,' V,-, V, ej - Fy,' V,, Vj e) ) ( P ) = 0 . 8 2
1
2
Our purpose is to find conditions for 82 y(P) = 0. For this purpose, in order to simplify our calculations without loss of generality we may reduce (6.27) to, in consequence of (6.11), (6.12), (6.13), (6.16), (6.17), (6.28), (6.29)
C,,,, = j / - 1 (V*. Fi2" - V,, F,,»).
244
CHUAN-CHIH HSIUNG
[September
Substituting (6.29) in (6.16), (6.17) gives, for £,-,,„ given by (6.29),
( W
« ^=1
,;•;•,'• [F,,h v», F,,L (V,, F,,> - V,, Fh>) 1 2 3
+ V», V*a FA3* + FA,'I F»/» V*3 V,, F y ,»], where (6.31) (6.32)
( r f l l f l ,„ = *».<» + i / : i r i ^o(3),. 16ff/i( = ^ // ' [o VJ * , ft.'CV, F*3* - Vhi Fj") » » a.n + F„3' Vftl(V»a Fjk - Vj * » / ) ] , (3
32F
(6 33)
'° <3 >
= e
w ! 8 [ F * ^ 2 V *' F»>h(-Vi> Fh" ~ Vh *V.*) + 2(V», V*2 F»,» + F»Ji F»,'> V», V;-2 F ; i *)].
Since II is the conjugate operator of II, from (6.15), (6.17), (6.29), 0.2
3,0
(6.31) it follows that (6.34)
d n f = ( - 0>.c» + / = T V/.«>)
Substituting (6.31), (6.34) in (6.13) yields
(6.35)
/«C«
+ • / - 1 [0).c»(tfft„ -
^„<«K
C^)
By (6.3), (6.11), (6.12), (6.30) we obtain 2Re +Itin = S£iy. 8(V*3 Ff - V} F*3*)(V*1 FHJ - Vhi FHJ) 18 3
+ 3 eW> FuJ Vj(V»2 F*3* - V*3 F»,») 1 8 3
+ 6 s"1"2"3 F* 'i Fh h(Vt F.-*—V.- F,- *) • (V,, FHJ - V*3 Fh') +
3ell.yFl,lhFl,^Fll3h 1 23
•^.(Vy./V-Vy./V).
1986]
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
245
Similarly, a very lengthy computation and use of (6.32), (6.5), (6.4) and (6.31), (6.6), (6.3) give
- 1 6 I W 1 1 A'§ + 5D'$) = 33 e*;*'\v», F / - V,- F*S*)(V», F*a> - Vh, F»,') - 36 ,;A;» F*J V*,(V*a F / - V, F»,») (6 37)
+ 30 ,W» ^
v
.(VA2 F*3* - V»3 F,.*)
+ 30 **£> F»,'t F . , ' . ^ F,a* - V,, F / )
•'(V/JF»l'-V*1F/lO + 15 «W» FuJ* F»,'.(V, F*3* - V*3 F / ) •YVI.FI,'-V/.Fi/)
+ 30 •,*•£• F»/. F»,». F, 3 '. V,,(Vi, F,,»-V, s F,a»), 16V>.c M (53$ + 3 C $ ) = 9 «;•*»'• (V,- F*t» - V», F/)(V* a F*3> - V»3 F»,*) i n
+ 36 «;•;»*• F*/ V*a(V», F / - V, F*5*) 18 3
<6-38)
- 6 tf£ F* t ' Vy(V», F*3* - V*3 F*a») + 18 . ^ F»/i F»/«(Vi F,,» - Vi, F / )
"(V/lF»,'-V»,F/lO + 15 «|JJJ F»,'. F*a'
!~Re(dC)/ o[3) = 0,V2/8 + totxitiv
where (6.40)
e,llt{l == «;**•*• (V*8 F / - V, F*a*)(V*1 F„a> - V*2 FHJ). 12 8
(6.41)
a),,,,,, = # • ' • F»,' V,-(V*a F*3* - V*3 F*a*). 1 2 &
7. Proof of the theorem.
At first, it should be remarked that
246
CHUAN-CHIH HSIUNG
September]
all the equatior s in §§1—6 with m = 2« can be applied to this section and the next section. From (4.12), (6.39) it follows that if the almost complex structure Fts on the manifold M 2 " is a complex structure, then (5.11) holds and (7.1)
0,V2,-3 +
Substituting (5.11) in (6.41) gives °>'ihh = Hiliiii + ifi^Va-
(7.2) where (7.3)
ffMl,,
= e^Y- i V Vi, Fi* VjiFnJ* F*,>. - FH^ F*,'0, 12 8
(7.4)
JT,,,,,, = ,,• • a Fhh F,h FH^VJ,
Vjt Fh" - Vh Vh Fh>).
Similarly, by substituting (5.11) in each factor on the right-hand side of (6.40) we can easily show that (i.O)
^'i'2'3
==
"iihh-
In order to compute jKij,-2,-3, we first expand out the right-hand side of (7.4) and then use the Ricci identity (1.8) and the identity (1.10); the final result is - -§• * w . = *V> Fhh
Rkhhh
+ Fhh Fi3h R>fihji
+ Fi3h F(lh
R>li)i]t.
Combination of (7.1), (7.2), (7.5), (7.6) gives condition (I) and theorem is therefore proved. 8. Proofs of the corollaries. (8.1) Then Fjj^Fji,
Fi} = gJt W,\
At first, Let F" = Q" Ft1.
since otherwise from (3.1) - 2n = Fa F^ = Fij F" > 0 ,
a contradiction. Thus on the manifold M2" there is a nonvanishing tensor field of type (0, 2): (8.2)
e,j = F) - F)i.
1986]
NONEXISTENCE OF A COMPLEX STRUCTURE ON THE SIX-SPHERE
247
If Mu is of constant sectional curvature R, by means of (1.12). (8.1), (8.2) we can reduce condition (I) to (8.3)
R(Ftl> f,,,, + F,lk £,,,, + Fi,» £ M | ) = 0.
Multiplying (8.3) by F"i and using (3.1) we obtain (8.4)
(ft - 1) R£,llt = 0 .
Since &,/, is nonvanishing, (8.4) holds for n ^ 2 if and only if 2? = 0. If R T^ 0, then i V satisfies condition (I) only when Mu admits another metric with respect to which M2" is of zero sectional curvature. Hence Corollary 1 is proved. If M" is of constant sectional curvature R and does not admit a flat metric, then R cannot vanish, and therefore (8.4) cannot hold for ft ^ 2. Hence Corollary 2 is proved. A 2«-sphere S2* does not admit a flat metric since otherwise its Euler-Poincar6 characteristic is zero by the Gauss-Bonnet formula. Since S2" is of constant positive sectional curvature with respect to the metric induced by the standard imbedding in a Euclidean space Ein+i, Corollary 3 follows immediately from Corollary 2. REFERENCES 1. J. F. Adams, On the non-existence of elements of Hopf invariant one, Ann. of Math., 72 (1960) 20-104. 2. A. Borel and J. P. Serre, Groupes de Lie et puissances rSduites de Steenrod, Amer. J. Math., 75 (1953) 409-448. 3. B. Eckmann and A. Frehlicher, Sur VintSgrabllitt des structures presque complexes, C.R. Acad. Sci. Paris, 232 (1951) 2284-2286. 4. C. Ehresmann and P. Libermann, Sur les structures predsque hermitiennes isotropes, C.R. Acad. Sci. Paris, 232 (1951) 1281-1283. 5. C. C. Hsiung and J. J. Levko III, Complex Laplacians on almost-Hermitian manifolds, J. Differential Georretry, 5 (1971) 383-403. 6. A. Kirchho:', Sur Vexistence de certains champs tensoriels sur les sphSres d. n dimensions, C.R. Acad. Sci. Paris,225 (1947) 1258-1260. 7. , Bet 'rage zur topologischen linearen algebra, Compositio Math., 11 (1953) 1-36. 8. K. Kodaira and D. C. Spencer, On the variation of almost-complex structure, Algebraic Geometry and Topology, A Symposium in Honor of S. Lefschetz, Princeton University Press, Princeton, 1957, 139-150. 9. M. Schiffer and D. C. Spencer, Functionals of finite Riemann surfaces, Princeton University Press, Princeton, 1954. 10. W. T. Wu, Sur les classes caracttristiques des strutures fibrees sphtriques, Actualities Sci. Indust., No. 1183, Hermann, Paris, 1952. Department of Mathematics Lehigh University Bethlehem, Pennsylvania 18015 U. S. A.
T R A N S A C T I O N S OF T H E AMERICAN MATHEMATICAL SOCIETY Volume 305, Number 1, January 1988
EULER-POINCARE CHARACTERISTIC AND HIGHER ORDER SECTIONAL CURVATURE. I CHUAN-CHIH HSIUNG AND KENNETH MICHAEL SHISKOWSKI ABSTRACT. The following long-standing conjecture of H. Hopf is well known. Let M be a compact orientable Riemannian manifold of even dimension n > 2. If M has nonnegative sectional curvature, then the Euler-Poincare characteristic x ( ^ ) *s nonnegative. If M has nonpositive sectional curvature, then x(M) is nonnegative or nonpositive according as n = 0 or 2 mod 4. This conjecture for « = 4 was proved first by J. W. Milnor and then by S. S. Chern by a different method. The main object of this paper is to prove this conjecture for a general n under an extra condition on higher order sectional curvature, which holds automatically for n = 4. Similar results are obtained for Kahler manifolds by using holomorphic sectional curvature, and F. Schur's theorem about the constancy of sectional curvature on a Riemannian manifold is extended.
Introduction. The first and simplest result relating local and global invariants in differential geometry is the Gauss-Bonnet formula. It expresses the Euler-Poincare characteristic x ( ^ 0 of a compact orientable Riemannian manifold M of even dimension n as an integral of the nth sectional curvature (or the Lipschitz-Killing curvature) times the volume element of M. Of course, for n = 2, the «th sectional curvature is the usual sectional curvature. Over the last three decades, the authors of [5, 6, 9, 10, and 24] have obtained various curvature conditions determining the sign of x ( ^ 0 - However, the following long-standing conjecture remains open. H. HOPF'S CONJECTURE. / / M has nonnegative sectional curvature, then x(M) is nonnegative. If M has nonpositive sectional curvature, then x(M) is nonnegative when n = 0 mod 4, and nonpositive when n = 2 mod 4. This conjecture cannot be established (see [7 and 13]) just by use of the GaussBonnet formula. For n = 4, the conjecture was proven by J. W. Milnor, and then by S. S. Chern in [5], using a different method. In [1], R. L. Bishop and S. I. Goldberg obtained a similar result for Kahler manifolds, using holomorphic sectional curvature instead of sectional curvature. Recently, D. L. Johnson [12] proved the Hopf conjecture for Kahler manifolds of real dimension 6.
Received by the editors December 5, 1983 and, in revised form, October 29, 1986. Presented to the Society, January 27, 1984, under the title Euler-Poincare characteristic. 1980 Mathematics Subject Classification (1985 Revision). Primary 53C20, 53C55, 53C65, 57R20. ©1988 American Mathematical Society 0002-9947/88 $1.00 + $.25 per page 113
114
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
In this paper, we relate the sign of x(M) to the sign of higher order sectional curvatures, studying higher order holomorphic sectional curvatures when M is Kahler. The main results are these. THEOREM A. Let M be an oriented compact Riemannian manifold of even dimension n, and p be an even integer with 2 < p < n. If the pth and (n — p)th curvature operators of M are both nonnegative or both nonpositive, then the Euler-Poincare characteristic x(M) > 0. / / one is nonnegative and the other nonpositive, then X(M)<0. THEOREM B. Let M be a connected Kahler manifold of real dimension 2n with n > 2, and let p be an even integer with 2 < p < 2n. If M has pointwise constant pth holomorphic sectional curvature, then M has constant pth holomorphic sectional curvature. THEOREM C. Let M be a Kahler manifold of real dimension 2w with n > 2, and let p be an even integer with 2 < p < 2n. If M has constant zero pth holomorphic sectional curvature, then M has constant zero qth holomorphic sectional curvature for all even integers q with q> p. IfM is also compact, then the Euler-Poincare characteristic X(M) is zero.
In fact, a more general result than Theorem A is proved in §2. However, for n = 4, Theorem A can be shown to reduce to the Hopf conjecture. Theorem B is well known when p = 2. Restated for a Riemannian manifold M with usual sectional curvature in place of holomorphic sectional curvature, Theorem B is the famous theorem of F. Schur, established for 2 < p < In by J. R. Thorpe [24]. As originally stated, above, Theorem B was proved by R. M. Naviera [18], using methods different from ours. Theorem C is proved in §5. A similar result for a Riemannian M was obtained by J. R. Thorpe in [24]. 1. Curvature operators. Let V be an n-dimensional real inner product space with inner product ( , ). For p an integer, 1 < / > < « , let AP(V) or simply Ap denote the space of p-vectors of V. We call a p-vector a decomposable if it can be written in the form ux A • • • Aup with u, e V. Any basis E = {eu...,en} for V induces a basis {e, A • • • Ae ; |1 < ^ < • • • < / < n) for Ap consisting of decomposable p-vectors. An arbitrary /^-vector a is of the form a = V a, ... e, A • • • Ae, , where summation extends over all 1 < /', < • • • < /' ^ n. The coefficients a, ...,, skew-symmetric in their indices, are called the Pliicker coordinates of a with respect to the basis E. We define an inner product for A ' on decomposable /^-vectors by
dct[(u,,Vj)],
where ut, v e V. It is easily seen that the basis for A ' induced by an orthonormal basis for V is itself orthonormal. We identify the Grassmann manifold G of oriented ^-dimensional subspaces P of V with the submanifold of kp consisting of decomposable ^-vectors of length one
EULER-POINCARE CHARACTERISTIC. I
115
by P <=> ux A • • •
Aup,
where {u1,...,u} is an oriented orthonormal basis for P. Elements of G will be called /j-planes. The following result is well known and will be needed later. For a proof, see Hodge and Pedoe [11, pp. 309ff]. LEMMA
1.1
( T H E GRASSMANN
QUADRATIC
/J-RELATIONS).
Choose
a
basis
{e^.. .,en) for V. Then the p-vector a is decomposable if and only if its Plucker coordinates satisfy P+ \
(i)
E (-i)*«*
•*
,- «,-,• ,• = o ,
for all 1 < ik, Ji < n where the symbol "over ik indicates that the subscript ik is to be deleted. Let 9tp, or simply 0t for fixed p, denote the vector space of all selfadjoint linear transformations on Ap with inner product (T,U) = tmce(T°U). Elements of Si are called />th (order) curvature operators on V. We associate to each R e S? its pth (order) sectional curvature function aR: G -» R defined by oR(P) = (R(P), P) for all P e G, where R is the set of real numbers. Then R e S$p satisfies the Bianchi identities if p+\
(2)
E ( ( - l ) ' * , ( « i A ••• Auk A ••• AUp
+ 1),UkAup
+ 2A
••• Au2p)
= 0,
k=l
for all ( . [ , . . . , u2p e V. Let 9$ denote the space of elements of !% which satisfy the Bianchi identities, and let Sf denote the space of elements R e !% for which oR = 0. Then from Stehney [23, §1] we have the following. LEMMA 1.2. With respect to the inner product defined above on SP, & — 3$ © £f is an orthogonal decomposition of 0t. We will need the following consequence of Lemma 1.2. LEMMA 1.3. Forp an integer with 0 < / > < « , let *p: Ap -> A"~p be the Hodge star operator. Then the map 2: @p -> 0tn~p is orthogonal with respect to the above inner product where fi(5) = *pS*(n - p) for all S <E ®p. PROOF. First, 0 is well defined since the adjoint of *p is Thus for S, T e @p we have (*pS*(n
- p),*pT*(n
- p)) = t r a c e ( > 5 * ( « - p)°*pT*{n
- p))
= trace((-l)'("^,V(S»T)*(«-/-)) = trace(v(5°r)(v)_1) = trace(5»7) = Hence the map 0, is orthogonal.
(S,T).
116
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
1.4. The map ft: &p -* <%"~p preserves the orthogonal decomposition given in Lemma 1.2, i.e., fi preserves the Bianchi identities as well as the operators with identically zero sectional curvature. COROLLARY
PROOF. In order to show that Q preserves the orthogonal decomposition, it is sufficient to show that for all 5 e £fp, *pS*(n - p) e Sf "~p since we know by Lemma 1.3 that fl is orthogonal. Now for an (n - /?)-plane a in V, *(n - p)(a) is a /7-plane in V. Thus for all S e6^p,
(*pS*(n - p)(a), a> = (-l)p("'p)(S(*n - p)a, *(n - p)a) = 0, so that *pS*(n - p) e 9">-p. Although the statements above are true in a general inner product space setting, our main interest is the />th curvature operator arising on a tangent space Mm of a Riemannian manifold M at a point m. Specifically, let R denote the Riemannian curvature tensor at m. Then for even integers p > 2, the pth curvature operator R is defined in Thorpe [26] by (RP(ui
A
• • •
=
A
" , ) >
y
i
~,p/i i ^ ^
A
• • •
£
Av
p)
sgn(a)sgn( B)R( «„(1), « a(2) , vm, vp(T>)
P- a,f}<ESp • • • R(U*(p-l)>ua(p)>
V
P(p-l)'
V
P(p))>
where «,-, Vj e Afm, and Sp is the permutation group on ( 1 , . . . , p). Also, Thorpe [26] establishes that Rp satisfies the Bianchi identities. The generalized Gauss-Bonnet theorem expresses the Euler-Poincare characteristic x ( M ) of a compact oriented Riemannian manifold M of even dimension n as the integral (3)
x{M) = yj c
n
KdV,
J
M
where K is the Lipschitz-Killing curvature of M, c„ is the volume of the Euclidean unit n-sphere, and dV is the volume element of M. The following useful theorem is a composite of the two theorems and Corollary 2 from Thorpe [26]. We write a2k for aR . THEOREM 1.5. Let M be an oriented Riemannian manifold of even dimension n. Then we have the following. (a) The Lipschitz-Killing curvature Katm^M satisfies
(4)
K(m) = (R„{Mm),Mm)
= pX(n ~ p)l trace(*(» -
p)R„.p*pRp).
(b) Let n = 4k. Then R2k*(2k)= ±*(2k)R2k if and only if o2k(*(2k)P) = ±a2kP for all 2k-planes P. Signs above are to be taken consistently, withplus implying that K > 0, minus that K < 0. (c) Let M be compact with n = 4k, R2k = + *(2k)R2k, and Pk denoting the kth Pontryagin class of M. Then the plus sign gives x(M)>~^\Pk\>0,
EULER-POINCARE CHARACTERISTIC. I
117
while the minus sign gives x(M) < 0 and Pk = 0. Moreover, x(M) = 0 if and only if M is 2k-flat, i.e., a2k = 0. Now we consider the case of a Kahler manifold M with an almost complex structure J and Riemannian curvature operator R. DEFINITION 1.6. Also denote by J the extension of the structure / from Mm to Ap{Mm) defined by J(ux A • • • Aup) = JulA ••• AJup for «, e Mm, extending linearly to Ap(Mm). Note that this gives J2 = (-1)". LEMMA 1.7. (a) Let R be the pth curvature operator on M. Then Rp = Rp° J = J°Rp. (b) The adjoint of J is (-l)pJ. (c) For even p, J and *p commute. PROOF, (a) This follows directly from the definition of Rp and the fact that for a Kahler manifold R = R°J = J ° R. (b) This follows directly from the definitions of ( , ) and J on Ap(Mm), and the fact that the adjoint of J on Mm is -J. (c) Let p be even. Now the Hodge star operator *p: Ap(Mm) -* A"~p(Mm) can be defined by
for £, T) e Ap(Mm), where Mm is the oriented /i-plane which is the tangent space of A / a t m . Let £, r\ e Ap{Mm). Since JMm = Mm and the adjoint of J is J, we have
[ U W ) ( 0 ] AT) = / ( [ V ( / 0 ] A /!,) = / « / £ , /T,)MJ
V ' which implies */>^ = J*PDEFINITION 1.8. For p even, a /^-plane 2 is said to be holomorphic if JQ = Q. Also, pih holomorphic sectional curvature is pi\\ sectional curvature restricted to holomorphic />-planes. Hence J * ^
=
THEOREM 1.9. Let V be an even-dimensional real inner product space with a complex structure J and Hermitian inner product ( , ). Also, let T: A2(V) -» A2(V) be a self-adjoint linear operator satisfying the Bianchi identity and T = T° J = J °T. Then we have the following. (a) T is uniquely determined by its holomorphic sectional curvature, i.e., if T has zero holomorphic sectional curvature, then T = 0. (b) / / T has constant holomorphic sectional curvature H, then for all X,Y,Z,W e V we have
(5)
(T(X A Y),Z A W)
= (H/4)[(X,Z),(Y,W)
- (X,W)(Y,Z) -(X,JW)(Y,JZ)
+
(X,JZ)(Y,JW) +
{X,JY)(Z,JW)\.
118
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
PROOF. This theorem is merely a restatement of Kobayashi-Nomizu [14, vol. II, Chapter IX, Propositions 7.1 and 7.3]. COROLLARY 1.10. Let Vand T satisfy the hypotheses of Theorem 1.9 with dim V = 2n and T having constant holomorphic sectional curvature H. Let { elt er,,..., en, en*} be an orthonormal basis of V, where ejf = Jei for i = 1,..., n. Then from this basis of V, we have an orthonormal basis B of A2(V) given by B = {ei A ei 11 < ix < i2 < n* where 1 < 1* < ••• < n < n*}. Furthermore, for Q e B we have TJ
f
HQ + — (6) V
T(Q)
'
= {
V X. /
£
Q',
if Q is holomorphic,
Q'CB,Q'*Q
\
Q,
\H(Q + JQ),
hohmorphic
if Q is not holomorphic.
PROOF. This obviously follows from (5) and the fact that B is an orthonormal basis of A2(V). Finally, we mention the relationship between the A:th Chern form and the (2k)th curvature operator on a compact Kahler manifold given in §2 of Gray [10]. THEOREM 1. Let M be a compact Kahler manifold of real dimension In with almost complex structure J. For m e M, let {ex,elt,...,en, e„*} be an orthonormal basis of Mm where e,» = Jef for i = \,...,n, let yk be the kth Chern form {via de Rham's theorem), and let R2k be the (2k)th curvature operator. Then
(4*)*fc! (7)
, .
. .
{2kl
(R2k(eh A e;. A • • • Aeik A e,..), y\ A • • • Aj2k),
L 1 < I, < • • •
forallj\,...,j2k
eMm.
PROOF. This follows directly from (2.4), Lemma 2.1, the definition of (2A:)th curvature operator in §2 of Gray [10], and the fact that Gray's (2/c)th curvature operator is [(2k)\/2k\R2k.
2. Euler-Poincare characteristic and curvature operators. Let 11^: Stp -» 88p be the orthogonal projection associated with the orthogonal decomposition of Lemma 1.2. 2.1. Let M be an oriented compact Riemannian manifold of even dimension n, and p be an even integer with 2 < p < n. Let Rp and Rn-p be the pth and (n — p)th curvature operators on M. If at each m e M there exists Rp e &p such that I I K = R with Rp and R„_p both positive semidefinite or negative semidefinite operators, then the Euler-Poincare characteristic x(M) > 0. If semidefinite is replaced by definite at at least one point of M, then x(M) > 0. IfRp and RnzJ> are of opposite signs as operators, then x(M) < 0 or x(M) < 0 according as both Rp and Rn_p are semidefinite or definite (at at least one point of M). THEOREM
EULER-POINCARE CHARACTERISTIC. I
119
PROOF. Assume that at each m e M the above holds with operators Rp and Rn_p both positive semidefinite. The other cases can be proven similarly. Since II R = Rp, there exists Sp e 9" with Rp = Rp + S,. Let {i> 1; ..., vk } be an orthonormal basis of Ap(Mm) consisting of eigenvectors of Rp, and let {X v ..., Xk} be the corresponding eigenvalues of Rp with respect to the given basis. Let K(m) be the Lipschitz-Killing curvature of M at m. Then by Corollary 1.4, Lemma 1.2, and (4) we have
K{m) =
[p\(n-Py./n\}tvacc{[*(n-p)Rn_p*p}[Rp-Sp])
= [p\(n - p)\/n\}[tT
= [p\(n -p)\/n\]
L
- trace(*(« -
p)Rn_p*pSp)]
p)R„-p*pRp) _
(*{n-p)Rn_*PRp{v1),v1)
i= i
k
= [p\(n -p)\/n\\
I
Xi(*(n-p)Rn_p*p(v,),vi)
i-i
k
= [p\(n-p)\/n\]
£ AX^n-^VO.V^^O, /=i
since X, > 0 for i = l,...,k, and i l , , ^ is positive semidefinite gives (Rn_p(a), a) > 0 for all a e A"~p. Hence by the generahzed Gauss-Bonnet formula (3) we have x ( ^ 0 > 0. REMARK 2.2. (a) Let the dimension of V be four. Then Thorpe [27] has shown that R e .S?2 has nonnegative (positive) sectional curvature if and only if R = H2R for some positive semidefinite (definite) R e ^ 2 . (b) D. Meyer [16] has shown that a compact Riemannian manifold with positive definite 2nd (usual) curvature operator is a homology sphere. (c) Thorpe [24] characterizes the pth curvature operators Rp of constant ;?th sectional curvature K as R = K • I where / is the identity operator. These are positive or negative definite according as K > 0 or K < 0. 3. Curvature operators of the form Rp = cAp on a Kahier manifold. Following Stehney [22] we can define a selfadjoint operator Rp = cAp on AP(V) for c e R and a selfadjoint operator A on V by cAp(ux A • • • A up) = c{Aul A • • • AAup) for all Uj e V, extending linearly to AP(V). Moreover, R = cAp satisfies the Bianchi identities. THEOREM 3.1. Let V be a real inner product space of dimension In, and J be a complex structure on V. For p an even integer with 2 < p < 2n, let Rp = cAp be an operator on AP(V) as defined above with Ap = J ° Ap = Ap ° J. Also, let {v1,...,v2n} be an orthonormal basis of V consisting of eigenvectors of A with corresponding eigenvalues Xv..., X2n. Then either R = 0 or A has exactly p nonzero eigenvalues A , , . . . , X, (1 < i'x < • • • < i < In) with vt A • • • A y, a holomorphic
120
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
p-plane, so that for all y e AP(V) Rp(y) = cA,-, • • • \ip(y, vh A • • • Avip)vh A ••• Avip. PROOF. Let B = {v^ A • • • AVj \1 ^j\ < • • • <jp^2n} for the above orthonormal basis of AP(V) consisting of eigenvectors of Rp with corresponding eigenvalues cX, • • • X,. Assume that Rp is not identically zero. Then c * 0 and there exist 1 < j x < • • • < i < In so that A;- • • • X, =#0. Thus ^4 has nonzero eigenvalues X,,..., X, with 1 < /'i < • • • < ip < 2w. Also, the corresponding eigenvector i>(. A • • • Ay, of Rp is a holomorphic /7-plane since Rp = J ° Rp. Now assume there exists X,• =fc 0 for some / =£ / , , . . . , / „ . Then A, • • • X, X, is not zero, and vt A • • • Aot _ A Vj is an eigenvector of R with eigenvalue X, • • • X, A •. Since Rp = J° Rp, vix A • • • At, A t . is also a holomorphic />-plane. Now y, A • • • A Vj and u, A • • • A vt A Vj are two holomorphic /7-planes which intersect in a holomorphic (p — l)-plane. This is a contradiction since p — 1 is odd. Hence Xy = 0 for all j + iv..., ip. Thus all other eigenvalues of Rp are zero other than cX^ • • • X,, and Rp must be of the stated form since B is an orthonormal basis of AP(V). COROLLARY 3.2. Let M be a Kahler manifold of real dimension In. For p an even integer with 2 < p < In, if the pth sectional curvature of M at m e M is constant at value K, then K = 0. PROOF. Assume the above is given with K + 0. Then from Thorpe [24], the pth curvature operator Rp at m is of the form Rp = K • Ip for the identity /. This contradicts Theorem 3.1 since / has In nonzero eigenvalues. Hence, K = 0. COROLLARY 3.3. Let M be a compact Kahler manifold of real dimension In. If at each m G M there exists c e R , an even integer p with 2 ^ p < n, and a self adjoint operator A on Mm so that the pth curvature operator Rp of M at m e M is of the form R = cAp, then the Euler-Poincare characteristic x(M) = 0, and the nth Chern class y„ of M is zero. PROOF. Let the above at m e M be given. Then at m e M, the (2/>)th curvature operator R 2 = c2A2p by Lemma 2.5 of Stehney [22]. In the following we assume Rp is written in the form of Theorem 3.1. CASE 1. Let 2 < p < n. Then 2 <2p <2n and Theorem 3.1 imply that R2p = 0 at m e M, since A cannot have both exactly p and 2p nonzero eigenvalues. Hence by the trace formula (4) for the Lipschitz-Killing curvature K at m, we have K(m) = 0. Also, (4) gives R2n{m) = 0. 2 2 2 2 CASE 2. Let p = n. Then R2n = c A " = (c det A)I " a t m e M for the identity /. But det A = 0 since A has exactly n nonzero eigenvalues or we must have c = 0 by Theorem 3.1. For either situation, we have R2n = 0 at m, and formula (4) again gives K(m) = 0. Now since the above occurs at each m e M, K = 0 and R2n = 0 on M. Hence the generalized Gauss-Bonnet formula (3) gives x ( ^ ) = 0. a r , d formula (7) implies that Y„ is zero also.
EULER-POINCARE CHARACTERISTIC. I
121
COROLLARY 3.4. Let the hypothesis of Corollary 3.3 be given with p constant on M. Then the kth Chern classes yk of M are zero for k > p, and yp/2 is zero if and only if M is p-flat. Also, M is q-flat for all q > 2/>. PROOF. From the proof of Corollary 3.3 it follows that the (2/>)th curvature operator R2p = 0 on M. Thus the gth curvature operator Rq = 0 for all q > 2p by Thorpe [24, Theorem 6.4] and [26, Lemma of §3]. Hence we have by formula (7) that yk is zero for k > p. Also, Rq = 0 on M for all q > 2p implies that M is g-flat for all q > 2p. Now if M is />-flat, then the above Lemma of Thorpe [26] gives Rp = 0 on M, and formula (7) implies that yp/2 is zero. On the other hand, let yp/2 be zero and assume M is not /?-flat. Suppose at some m0 G M, Rp is not identically zero and is of the nonzero form stated in Theorem 3.1. In the notation of Theorem 3.1, the holomorphic /?-plane u, A • • • A u( can be written as ex A ex» A • • • Aer A er. where r = p/2 for an orthonormal basis {e x ,e x .,...,e n ,e n »} of AfOTo with e}. = Jej for all j = l,...,n. Now using formula (7) and Theorem 3.1 we get at m0 that
{r\)(2Tr)ryp/2{el,el„...,er,er,)
0 = = ^TT-
I lsiji<
(Rp{eh
•••
A ey.. A • • • AeJr A eJf),
<jr
fj A e l t A • • • Aer A e r . )
But cX, • • • A, ¥= 0 by Theorem 3.1. Thus we have a contradiction, and hence M is 'l
'/>
J
/>-flat. 3.5. Le/ the hypothesis of Corollary 3.3 be given with p constant on M and n < p < 2n. If the (In — p)th curvature operator R2n^p of M has constant zero holomorphic sectional curvature, then the Euler-Poincare characteristic x(M) = 0, and the nth Chern class y„ is zero. PROOF. If Rp = 0 at m e M, then by formula (4) the Lipschitz-Killing curvature at m is K(m) = 0. Also, if Rp =£ 0 at m e M, then Theorem 3.1, together with its notation and proof, gives that for all \p(Mm), COROLLARY
Rp(y) = CA„ • • • \,p(y,vlt
A ••• Avip)vh A ••• Avlp.
Now using formula (4) and the basis of Mm in Theorem 3.1 we have [(2n)]/(p\(2n = c\,
- p)\)]K(m) = trace(*(2« - p)R2„_p*pRp) E <*(2" - p)R2n-p*pRp{»jl A • • • Kv]f),vh m
•••\if(R2.-P( p[»ll
A ••• A 0 / J ) , *p[vti A •••
A • • • Avjp)
Avip}).
122
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
Then K(m) = 0 since R2„-p has constant zero holomorphic sectional curvature, and *p(Vj A ••• AVj) is a holomorphic (2n — ^-plane by Theorem 3.1 and Lemma 1.7(c). Thus K = 0 on M, and R2n = 0 on M by formula (4). Hence the generalized Gauss-Bonnet formula (3) gives x(M) = 0, and formula (7) implies that yn is zero. 4. Constancy of holomorphic sectional curvature. The purpose of this section is to prove the following theorem. THEOREM 4.1. Let M be a connected Kdhler manifold of real dimension In -with n > 2, and let p be an even integer with 2 < p < 2n. If M has pointwise constant pth holomorphic sectional curvature, then M has constant pth holomorphic sectional curvature.
Before we prove Theorem 4.1, we need the following lemma. LEMMA 4.2. Let V be a real inner product space of dimension In for n > 2 with complex structure J and Hermitian inner product with respect to J. For an even integer p with 2 < p < 2«, let [e^e^,.. .,e„,en») be an orthonormal basis of V with e,» = Jei for i = \,...,n, and let B = {eti A • • • Aet 11 < ix < • • • < ip s$ n*}. Also, let T: AP(V) -» AP(V) be a selfadjoint operator, satisfying the Bianchi identities T = J °T = T ° J, with constant holomorphic sectional curvature H. Then for any holomorphic p-plane Q e B,
[H \H (8)
(T(Q),Q') = 0
ifQ' = Q, if Q' is holomorphic, and Q'' C\ Q is a holomorphic (p — 2)-plane, */ Q' is not holomorphic, but Q' n Q contains a holomorphic (p — 2)-plane.
PROOF. The case when p — 2 easily follows from Corollary 1.10, and so in the remainder of the proof we assume that 2 < p < 2n. For 2k = p — 2 and any 1 < j \ < • • • <jk*in,letA= e^ A e^ A • • • Ae J( A e].. be the holomorphic (p - 2)-plane formed from the given basis of V stated above. Then Ax is a holomorphic (2 K — 2&)-plane since the inner product is Hermitian. Now define a second-order curvature operator R on A -1 by
(R(X A Y),Z A W) = (T(A AX AY),A L
AZ
AW)
for all X, Y, Z, W e A , extending linearly. Then R satisfies the hypothesis of Corollary 1.10, and using formula (6) with the basis of Ax taken from the given basis of V we easily get formula (8) for T. PROOF OF THEOREM 4.1. Let F(M) be the principal 0(2n)-bundle of orthonormal frames on M, where 0(2n) is the group of orthogonal 2n X 2« matrices. Let II: F(M) -* M be the projection map. The curvature form S2 = [S2, ] of the Riemannian connection of M is a smooth 2-form on F(M) with values in the Lie algerbra o(2n) of real skew-symmetric 2n X 2n matrices. If v,w are tangent vectors at z = (m; ex,...,e2n) e F(M), with
EULER-POINCARE CHARACTERISTIC. I
123
R the Riemannian curvature operator, we have for v' = n*(<;) and w' = TL*(w) that G, 7 (Z)(U,H>) = (R(e,
For I = {i1,...,ip}
A «?,), v' A
w').
with iv ...,ip
integers between 1 and In, we define
«/ = ( i / p O E « / a A / - 2 A - - - A a w / where J = {j\,---,jp} for y 1 ,..., jp integers between 1 and 2n, with 8j = 1 (or -1) if the integers iv ...,ip are distinct and J is an even (or odd) permutation of /, with 8} = 0 otherwise, and with the sum Ej taken over all selections J from { 1 , . . . , In}. Then 0, is a p-iorm on F(M) for all /. Also DO, = 0 for all / since DQ = 0 where D is covariant differentiation. Specifically, Da = da° h for an arbitrary form a where d is exterior differentiation and h takes the horizontal part. The 1-forms oi; are defined on F(M) for / = 1,..., 2« and a tangent vector v at z = (w; e 1 ; . . . , e2«) G JF(M) by co.(z)(u) = (II*(y), e,-> where ( , > denotes inner product. Note that £>
= (Rp{eliA E
••• A ^ J . n ^ o j A ••• An,( D/ ,)> < ^ ( ^ A ••• A ^ ) , ^ A ••• Aejp) XuA A ••• A wy ! ( « ! , . . . , u,,).
Hence, at z = (m; elt..., e2„) s ^(M) and for I = {i^,..-, (9)
0, =
£
ip},
< U , ( e / i A • • • A * , J , ^ A • • • A^>co A A • • • A « v
Now define the subbundle U(M) of F(M) consisting of the unitary frames on M, i.e., z = (m: Jex, . . . , Jen, ex, . . . , e„) is an element of U(M) if { Jex,..., Jen, ex,...,en} is an orthonormal basis of Mm. Then U(M) has structure group £/(«) of unitary matrices on C" considered as the subgroup of 0(2 w) consisting of the matrices commuting with the matrix
where I„ is the n X n identity matrix. Note that this representation of U(n) into 0(2«), called the real representation of U(n), is given by
"»-{-*
A)
for A + iB e U(n), where both A and 5 are real n X n matrices. From §§3 and 4, Chapter IX in [14], it follows that the Riemannian connection of M is a connection in U(M), i.e., the connection form and curvature form £2 of the Riemannian connection of M in F(M), when restricted to U(M) c F(M), take values in the Lie algebra u(n) considered as its real representation in o(2n). We may
124
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
restrict 0„ w, and all formulas above involving them to U(M) without any alteration. In the following portion of the proof, we only work in U(M) and treat all forms used previously as defined on U(M). Let / = {J\J= {h,...,jp} with 1
4
\eit
if
H -
if ./*>»•
A<".
00
For all /, 7 e / we define C functions RfJ on t/(M) by /? 7 , y (z) = (Rp(E,), Es). J ^Jf is said to be holomorphic if Ej(z) is a holomorphic />-plane for all z G U(M). Now let w, = «,-, A • • • A ^ for all 7 <=,/. Then on t/(M) for I e , / , formula (9) becomes
(io)
e, = I *„«,.
Let # be the C°° function on t/(M) obtained by taking the C°° function H' on M representing the pointwise constant /7th holomorphic sectional curvature and composing it with the projection IT: U(M) -> M. For /? = 2/c, let 70 = {1, 2 , . . . , A:, n + 1, n + 2 , . . . , n + k} e / . Then formulas (8) and (10) give for all z e U(M) (11)
0,o(z) = # ( z )«,„(*) + \H(Z) + 0- £
£
«•>/(*) + I
yefl
*>,(*)
*/„./(*)«/(*)>
jeC
where ^ = {J e / 1 J + I0, J is holomorphic, and J (~) I0 is holomorphic of order p — 2}, B = {J &Jf\J is not holomorphic, J D I0 contains a holomorphic subset of order p - 2}, C= {/e/|/n/
does not contain a holomorphic subset of order p - 2}. Note that J= A L> B U C U {I0}. Thus Du,j = 0 for all J e / , and (11) gives (12)
0
'/„ +' ^7e/4I «/J +7eCI [ K ; ) A 4
0 = D$ln = DH A «,, L
Since H is constant on the fibers of U(M), dH is a horizontal 1-form and equals DH. Letting dH = tf2lHlui and DRloJ = I?k'LlR,nJ^uk for C°° functions //,-, Rt j , k o n U(M), (12) thus becomes (13)
0 = E " , « , A «,„ + ^ ( X +
E £ JeCykmJ
R
iu,j,kukA
U
J • '
" , " , A «,
EULER-POINCARE CHARACTERISTIC. I
125
It follows from the definitions of A and C that the intersection of {to, A u,o \ ir £ I0} and {to, A
B = [ehA
• •• Ae,2n J l < i\ < • • • < i2n_2 < n*, 1 < 1* < • • • < n < « * } .
Then for Q e B, formula (6) holds. In particular, if H = 0, then T is identically zero. PROOF. Let *2, *(2« - 2) be the respective Hodge star operators. Then applying Corollary 1.10 to *(2n — 2)T*2 and using Corollary 1.4 together with Lemma 1.7 we obtain the above result. Next we restate the corollary in §2 of Thorpe [26] as follows.
LEMMA 5.2. Let M be a Riemannian manifold of dimension n, and p,q be positive even integers with p + q < n. Let ( e 1 ; . . -,ep+q) be an orthonormal basis for a (p + q)-plane P at m G M, and B = {e, A • • • A el, 11 < i1 < • • • < ip < p + q). Then
(14)
*' + « ( p ) " 7^?iv
L
MG)A*,(G A ),
where Q ± is the oriented orthogonal complement of Q in P, and R , R , Rp+q are the pth, qth, (p + q)th curvature operators of M, respectively. THEOREM 5.3. Let M be a Kahler manifold of real dimension 2n for n > 2, andp be an even integer with 2 < p < 2n. Let the pth holomorphic sectional curvature of M at m 0 e M be identically zero. Then the qth holomorphic sectional curvature of M at m0 is identically zero for all even q > p. In particular, at w 0 e M the Lipschitz-Killing curvature K(m0) = 0, andM is (2n — 2)-flat atm0. PROOF. Throughout the proof we work only at m 0 e M. Let Rr be the rth curvature operator of M for even integers r with 2 < r < 2n. In order to prove the theorem, it is sufficient to show that constant zero rth holomorphic sectional curvature of M implies constant zero (r + 2)nd holomorphic sectional curvature of M for an even integer r with 2 < r < 2n. Note that constant
126
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
zero (2n — 2)nd holomorphic sectional curvature of M implies that M is (In — 2)flat, by Lemma 5.1. Assume constant zero rth holomorphic sectional curvature of M for an even integer r with 2 < r < In. Let P be a holomorphic (r + 2)-plane with an orthonormal basis {e^e^,.. .,es,es*} where ei„=Jej for all i and 2s = r + 2. Then P = ex A e,„ A • • • ACjA e s ,. Extend this basis of P to an orthonormal basis {el,el»,..., e„, en.} of Mm , where e,-» = /e, for all ;'. Let 2? = { e,- A • • • A e(. 11 < /j < • • • < / r < .?*}, B' = [eh A • • • Ae,.Jl < ix < • • • < ir < «*}, and 5 " = {e,^ A e,211 < ij < i 2 < «*} where 1 < 1* < • • • < n < n*. Then Z?, 5 ' , B" are orthonormal bases of Ar(P), Ar(Mm ), and A2(Mm ), respectively. Let II : Ar(MmJ -* Ar(P) be the orthogonal projection map. Note that J° TLP = TlP°J since P is holomorphic, and J(Ar(P)) = Ar(P). We can define an rth-order curvature operator on P by R'r= UP°(Rr\ Ar^P)). Then R'r satisfies the Bianchi identity, R'r = J° R'r = R'r° J, and R'r has constant zero holomorphic sectional curvature since Rr does. From Lemma 5.1 it follows that R'r is identically zero. Hence for all Q e B we have that Up(Rr(Q)) = 0. By Lemma 5.2 and formula (14), (15)
R,+2(P)
=j
^
-
\r "I" L).
E Rr{Q) A ^ ( G 1 ) . Q^g
For each Q e 2? we have
*,(e)= I <*,(e),e'>e\ * 2 (e x ) = E («2(ex),e">e", g"£fl" SO
«r(e)AJR2(e^)=
E
[<«,(e),e , x«2(e ± ),e">(e'Ae")].
Letting /?'" = B" n A 2 (P) and using the above with formula (15), we obtain
(Rr+2(P),P)
= TI^~ ^
+
E ^
(Rr(Q),Q')(R2(Q±),Q"XQ'
A g",P>.
Q.Q'<=B' Q"eB'"
But n,,(/?,((?)) = 0 for all g e B, i.e., ? f (2).e'> = 0 for all 2 , 2 ' e B. Hence (Rr+2(P), P) = 0, so that constant zero rth holomorphic sectional curvature of M implies constant zero (r + 2)nd holomorphic sectional curvature of M. The following Theorem is a corollary of Theorem 5.3. THEOREM 5.4. Let M be a compact Kdhler manifold of real dimension 2n for n > 2. For an even integer p with 2 < p < 2«, let the pth holomorphic sectional curvature of M be identically zero. Then the Euler-Poincare characteristic x(M) = 0 and the (n — \)st Chern class y 1 is zero.
579 EULER-POINCARE CHARACTERISTIC. I
127
PROOF. By Theorem 5.3 we know that the Lipschitz-Killing curvature K of M is identically zero, and so the generalized Gauss-Bonnet theorem gives x ( ^ 0 = 0Also by Theorem 5.3 we know that M is (2« — 2)-flat, i.e., R2„~2 ' s identically zero. Hence from Theorem 1.11 and formula (7), it follows that yn_1 is zero. Our last theorem involves the vanishing of Chern classes for a compact Kahler manifold M. THEOREM 5.5. Let M be a compact Kahler manifold of real dimension In for n ^ 2. For an even integer p with 2 < p < In, let Rp(Q) = 0 for all holomorphic p-planes Q, where Rp is the pth curvature operator of M. Then the Chern classes y of M are all zero for all q > p/2. PROOF. By Theorem 1.11 and formula (7), it is sufficient to show that if XQ) = 0 for all holomorphic r-planes Q with r an even integer such that 2 ^ r < In, then Rr+2(Q') = 0 for all holomorphic (r + 2)-planes Q'. Let P be a holomorphic (r + 2)-plane with orthonormal basis { ex, ex»,..., es, es,} where e,. = Je, for all (' and 2s = r + 2. Then P = ex A ex, A • • • Ac,A es„. Let B = {e(i A • • • Ae, |1 < il < ••• < ir^s* with 1 < 1* < • • • < s < s*}. Then by Lemma 5.2, formula (15) holds. Assume that Rr(Q) = 0 (or all holomorphic r-planes Q, and let R
Q' = Xx A JXl A • • • AX,/\JX,A
Y A Z,
where 2t = r - 2. Then 0 = Rr{Xl /\JXl
A ••• M , A JX, A (Y-JZ)
A(JY + Z))
= Rr( XY A JXX A • • • A X, A JXt A Y A JY) + Rr{X\
A
JX\
A
•' •
t\X,/\JX,/\ZAJZ)
+ 2Rr{Xx AJXX A ••• AX, AJX, A Y A Z) =
2Rr(Q').
Hence Rr(Q') = 0 for any /--plane Q' containing a holomorphic (r — 2)-subplane. From this, it is clear that Rr(Q) = 0 for all Q e B since each of these contains a holomorphic (r — 2)-subplane. Thus formula (15) implies that Rr+2(P) = 0. REFERENCES 1. R. L. Bishop and S. I. Goldberg, Some implications of the generalized Gauss-Bonnet theorem, Trans. Amcr. Math. Soc. 11 (1964), 508-535. 2. On the second cohomologv group of a Kahler manifold of positive curvature, Proc. Amer. Math. Soc. 16(1965), 119-122. 3. S. Block and D. Giesieker, The positivity of the Chern classes of an ample vector bundle. Invent. Math. 12(1971), 112-117. 4. J. P. Bourguignon and H. Karcher, Curvature operators: pinching estimates and geometric examples, Ann. Sci. Ecole Norm. Sup. 11 (1978), 71-92. 5. S. S. Chern, On the curvature and characteristic classes of a Riemannian manifold, Abh. Math. Sem. Univ. Hamburg 20 (1956), 117-126. 6. Y. K. Cheung and C. C. Hsiung, Curvature and characteristic classes of compact Riemannian manifolds, J. Differential Geom. 1 (1967), 89-97. 7. R. Gcroch, Positive sectional curvature does not imply positive Gauss-Bonnet integrand, Proc. Amer. Math. Soc. 54 (1976), 267-270.
128
CHUAN-CHIH HSIUNG AND K. M. SHISKOWSKI
8. A. Gray, A generalization of F. Schur's theorem, J. Math. Soc. Japan 21 (1969), 454-457. 9. , Some relations between curvature and characteristic classes, Math. Ann. 184 (1970), 257-267. 10. , Chern numbers and curvature, Amer. J. Math. 100 (1978), 463-476. 11. W. V. Hodge and D. Pedoe, Methods of algebraic geometry, Cambridge Univ. Press, Cambridge, 1947. 12. D. L. Johnson, Curvature and Euler characteristic for six-dimensional Kahler manifolds, Illinois J. Math. 28 (1984), 654-675. 13. P. Klembeck, On Geroch's counterexample to the algebraic Hopf conjecture, Proc. Amer. Math. Soc. 59 (1976), 334-336. 14. S. Kobayashi and K. Nomizu, Foundations of differential geometry, Vol. I, 1963; Vol. II, 1969, Wiley, New York. 15. R. S. Kulkarni, On the Bianchi identities, Math. Ann. 199 (1972), 175-204. 16. D. Meyer, Sur les varietes Riemanniennes a operateur de courbure positif, C.R. Acad. Sci. Paris Ser. A-B 272 (1971), 482-485. 17. S. Mori, Projective manifolds with ample tangent bundles, Ann. of Math. (2) 110 (1979), 590-606. 18. A. M. Naveira, Characterisation des varietes a courbures sectionnelles holomorphes generalisees constantes, J. Differential Geom. 9 (1974), 55-60. 19. , On the higher order sectional curvatures, Illinois J. Math. 19(1975), 165-172. 20. I. M. Singer and J. A. Thorpe, The curvature of ^-dimensional Einstein spaces, Global Analysis, Papers in Honor of K. Kodaira, Univ. of Tokyo Press and Princeton Univ. Press, 1969, pp. 355-365. 21. Y. T. Siu and S. T. Yau, Compact Kahler manifolds ofpositive bisectional curvature, Invent. Math. 59 (1980), 189-204. 22. A. Stehney, Courbure d 'ordrep et les classes de Pontrjagin, J. Differential Geom. 8 (1973), 125-134. 23. , Extremal sets ofpth sectional curvature, J. Differential Geom. 8 (1973), 383-400. 24. J. A. Thorpe, Sectional curvature and characteristic classes, Ann. of Math. (2) 80 (1965), 429-443. 25. , On the curvature of Riemannian manifolds, Illinois J. Math. 10 (1966), 412-417. 26. , Some remarks on the Gauss-Bonnet integral, J. Math. Mech. 18 (1969), 779-786. 27. , On the curvature tensor of a positively curved A-manifold, Proc. 13th Biennial Sem. Canad. Math. Congress, Vol. 2, 1971, pp. 156-159. DEPARTMENT OF MATHEMATICS, LEHIGH UNIVERSITY, BETHLEHEM, PENNSYLVANIA 18015 DEPARTMENT OF MATHEMATICS, EASTERN MICHIGAN UNIVERSITY, YPSILANTI, MICHIGAN 48197
Conformal Invariants of Submanifolds. II CHUAN-CHIH HSIUNG & JOHN J. LEVKO
1. Introduction. Let x:Mn —> Nn+m be an n (> 2)-dimensional submanifold immersed in a Riemannian (n + m)-manifold Nn+m. For simplicity we shall write x(Mn) as Mn. Let g^ be the Riemannian metric tensor on Mn induced by the immersion x, e a unit normal vector of Mn at a point x, and fijj the second fundamental tensor of Mn at x with respect to e. Then the eigenvalues hi(e), ...,hn(e) of the matrix (fiy) relative to the matrix (gij), i.e., the roots of the determinant equation (1.1)
det(fiy - Xgij) = 0
in A, are called the principal curvatures of M " at x with respect to e, and the r th mean curvature of Mn at x with respect to e is defined to be the r th elementary symmetric function of hi(e),... ,hn(e) divided by the number of terms, i.e., (1.2)
("W(e)= ^
'
Yl
Me)---Me),
r = l,...,n,
ti<...
where (") is a binomial coefficient. It is convenient to define Ho(e) = 1. It is well known (Haantjes [5]) that every conformal mapping / on a Euclidean (n + m)-space En+m can be decomposed into a product of similarity transformations (i.e., Euclidean motions and homotheties) and inversions {7Tj}. Let Mn —• En+m be an n-dimensional submanifold immersed in En+m. Then a quantity on Mn is a conformal invariant if it is invariant under all the conformal mappings of En+m, for which the center of every inversion does not lie on Mn. Hsiung and Mugridge [6], [7] have established the following three theorems: Theorem 1. Let x:Mn —• Nn+m be an n-dimensional submanifold immersed in a Riemannian (n + m)-manifold Nn+m, and { e n + i , . . . ,en+m} be an orthonormal basis of the normal space at a point x of Mn. Then the normal vector of Mn n+m
(1.3)
J2 #i( e «) e « a=n+l 181 Indiana University Mathematics Journal © , Vol. 37, No. 1 (1988)
182
C.-C. HSIUNG & J. J. LEVKO
and the scalar n+m
(1-4)
J2 ^ ( O a=n+l
are invariant under a Euclidean motion in the normal space at the point x of Mn. Moreover, if Nn+m is of constant sectional curvature, then the invariant (1-4) depends only on the Riemannian metric on Mn induced by the immersion x. The invariability of the vector (1.3) is well known. The vector is called the mean curvature vector of Mn in Nn+m, and its magnitude n+m
(i.5)
(#i(o) 2 ] 1 / 2
[£ a=n+l
is called the mean curvature of M " in Nn+m. If (1.5) vanishes at every point of Mn, then Mn is said to be minimal. For n — 2 and Nn+m = E3, the invariant (1.4) is the well-known Gaussian curvature. H. Weyl [8] obtained some invariants similar to the invariant (1.4). Theorem 2. Let x:Mn —> E n + m be an n-dimensional submanifold immersed in a Euclidean (n + m)-space En+m, e be a unit normal vector of Mn at a point x, 5 m _ 1 be the (n— l)-sphere of the unit normal vectors of Mn at x, and dVn and dcrm-i be the volume elements of Mn and Sm~1 atx, respectively. Then (1.6)
[H2{e)}n'2 do-m_-\
K2 = ijm
dVn
is a local conformal invariant of Mn immersed in En+m, (1.7)
/
and
K2
is a global conformal invariant of a compact oriented Mn in En+m, (1.8)
where
H2(e)=[H1(e)]2-H2(e).
Theorem 3. Let x: Mn —> En+m, dVn and W2 be defined as in Theorem 2, and let { e n + 1 , . . . ,en+m} be an orthonormal basis of the normal space of Mn at x. Then n+m
(1.9)
[ J3 H2(ea)]n/2dVn a=n+l
is a conformal invariant of Mn in
En+m.
Conformal Invariants of Submanifolds
183
Theorems 2 and 3 are due to W. Blaschke [1] for n = 2, m = 1, and due to B. Y. Chen [2] for n = 2 and a general m . Moreover, for a compact oriented M2 by using the well-known Gauss-Bonnet formula it follows that JM2 H2 dV2 is a global conformal invariant (J. H. White [9], B. Y. Chen [2]). It should be noted that for Mn in Theorem 3, S. S. Chern and J. Simons [3] have also obtained a conformal invariant, but it is of one differential order higher than the invariants (1.6) and (1.9). It should also be noted that a hypersphere of En+1 has vanishing invariants (1.6) and (1.9) since the principal curvatures h\(e),...,hn(e) of the hypersphere with respect to the unique unit normal vector e at every point x are equal. The purpose of this paper is to generalize Theorem 2 as follows. Theorem
(1-10) (1.11)
4.
Under the assumptions of Theorem 2,
K2i = U^ \K2i+1\ = ^J
i
[H2i{e)]n,(2i) dam^\ dVn,
i
| f t 2 i + i ( e ) r / ( 2 m ) dam^ dVn,
where 1 < i < [^] and
(1.12)
Hj[e)= [ffi(e)f-flj(e)-j;
[
[ifi(e)]VtW, j = 3,...,n,
with (1.13)
Wi = 0,
are local conformal invariants of Mn immersed in En+m, (1-14)
/ J Mn
K2i
and
f JMn
and
\K2i+1\
are global conformal invariants of a compact oriented Mn in En+m . It should be noted that for i = 1 the invariant (1.10) and the first invariant of (1.4) are given in Theorem 2, and that a hypersphere of En+1 has vanishing invariants (1.10) and (1.11). 2. Proof of Theorem 4. Let (2.1)
x:Mn ->
En+m
C.-C. HSIUNG & J. J. LEVKO
184
be an isometric immersion of class C2, and the immersed submanifold x{Mn) be simply written as Mn. The first fundamental form of M " at a point x defined to be (2.2)
l=
dx-dx,
where dx is a vector-valued linear form on Mn, and the dot denotes the inner product of two vectors in En+m. Moreover, the second fundamental form of Mn at x with respect to a general unit normal vector e is defined to be (2.3)
11(e) = -dx
-de,
which is a vector-valued quadratic form on Mn. The eigenvalues hi(e),...,hn(e) of 11(e) relative to I are the principal curvatures of Mn at x with respect to e, and the r t h mean curvature Hr(e) of Mn at x with respect to e is given by (1.2). It is obvious that K2% and i^2i+i are invariant under similarity transformations in En+m, so that it suffices to show that K2i and |K 2 i+i| are invariant under an inversion n on En+m, whose center does not lie on the submanifold Mn. Choose the center of the inversion ir to be the origin of a coordinate system in En+m, and let x, x be the positive vectors of a pair of corresponding points of the submanifold Mn and its image submanifold Mn under TT . Then the definition of an inversion implies x = (c2r~2)x,
(2.4)
r2=x-x,
where c is the radius of the inversion n. From (2.4) we readily obtain (2.5) (2.6)
(c2r~2)dx-2(c2r-3dr)x,
dx = dx • dx = (c^r-
) dx- dx.
Let {e„+i,..., e n + m } be an orthonormal basis of the normal space at a point x of Mn. By using (2.5) it is easy to verify that (2.7)
ea = 2r~2(x-ea)x-ea,
a — n + 1,... ,n + m,
form an orthonormal basis of the normal space of Mn at x. Similarly, e = 2r _ 2 (a;-e)a;-e
(2.8)
is a unit normal vector Mn at x. Since e can be written as n+m
(2.9)
e=
^T a=n+l
n+m
aaea,
^ a=n+l
a 2 = 1,
Conformal Invariants of Submanifolds
185
we have n+m
(2.10)
e=
^2
aaea.
Thus, if the vector e moves over the sphere 5 m _ 1 of Mn at x, then the vector e moves over the (m — l)-sphere 5 m _ 1 of the unit normal vectors of Mn at x. By means of (2.4), (2.5) and (2.8) we obtain dx-de = 2c2r~i(x -e)dx-dx
(2.11)
— (c2r~2) dx • de,
and therefore, in consequence of (2.6), 2 c2 ' , ( 2{x-e) r2\ , , ' 2 A ~ ) dx-dx dx • de + Xdx • dx = — $ dx • de + I r Let dVn be the volume element of the submanifold Mn at a point x, and hi(e),... ,hn(e) the principal curvatures of Mn at x with respect to e. From (2.6), (2.12), (2.2), and (2.3) for ea replaced by e, it follows that
(2.12)
dVn=(f)2ndVn,
(2.13)
hi(e) = -c-2r2hi(e)-2c~2(x-e),
(2.14)
i=
l,...,n.
To complete our proof we need the following two lemmas: Lemma is given by (2.15)
1.
The s -th mean curvature Hs{e) of Mn at x with respect to e
c2sHs(e) = (-iy£r-t(S)r2t(x.ey-tHt(e),
s=
l,...,n.
v t /
t=o
Proof. (2.14) can be rewritten as hi(e) = -c'2r2[A
(2.16)
+ hi(e)],
where (2.17)
2r~2(x-e).
A =
Then by (2.16) we have, for ii < ... < is and s < n, (2.18)
M e ) - M e ) = (-l)'c-2'r3'^fft(e)^-*,
where
(2.19)
Y, i1<...
<7
'( e )=Lli) E ^
'
ii<...
Me).-Me),
< < «,
C.-C. HSIUNG & J. J. LEVKO
186
we can easily obtain (2.15). Lemma (2.21)
2. Hi{e) = (-iyc-2ir2iHi(e),
t = 2,...,n.
Proof. From (2.15) it follows that
(2.22) and therefore that (2.23)
(x • e) = - I [c 2 iJ!(e) + r
(2.24)
<*[&&]'
2
^)],
= (-!)• ^ ^ ( ^ ^ - ^ [ ^ ( e J J - ' ^ - e ) * ,
by the binomial expansion. Using
from (2.15), (2.24) we obtain immediately (2.26)
^[(H^y-Hsie)]
(-iy{r2s[(H1(e)y-Hs(e)}
= +£
2<
(J)
r2(S
~ ° K » i ( e ) ) 8 " ' - ^ s - t ( e ) ] (x • e ) ' } .
Substitution of (2.23) in (2.26) and use of the binomial expansion give (2.27)
c 2s [ ( ^ ( e ) ) ' - H.(e)] = ( - l ) V 2 s [ ( ^ ( e ) ) * -
ff.(e)] +(I) + (II) + (III),
where s-2
,
(i) = (-i) E(- 1 )' ca *(j) | - a( '" t) [ fi i(*)] t •[(^(ejr'-ff.^e)],
Conformal Invariants of Submanifolds
187
s-2
(2.28)
(II) =
(-l)^(-l)*(Mr^[i7l(e)]
-iW
•[(F1(e)r<-Fs_((e)], s-2
(III) -
( - l ) s ^ ( --li)W *(')r2(s-')[(if1(e)r(-^_t(e)] t=2
^C) c2(t "" V2u ^ i(e " )]< "" [ifi(e)]tt t-1
«=1 ^ '
By using (2.26) for s = 2 and (1.8), together with its corresponding equation for Mn, we immediately see that (2.21) holds for i = 2. We shall use the induction to show (2.21) for a general i. To this end we assume that (2.21) holds for i = 2,..., s — 1, and shall prove that it also holds for i = s. By means of (1.11) and (2.21) for i — 2,... ,s - 1, from (2.28) we obtain 2a
s-2 / 's
(D=c 52Q[H1(e)]tHs-t(e) + (-l)'|;(-l)*(j)ca*ra(-«)[F1(c)]t
•'E a (*7'V i ( e ) ]' w -*-' ( e ) ' i=i
^
J
'
(2.29) (II) = (-1) V { £ ( - 1 ) * ( j ) [H1(e)]<Ws_t(e) + D - D * ( ! ) ' i f (* ; ' ) [Hi(e)]'*.-.-i(e)}, (III) = (-!)• £ ( - ! ) ' ( ! ) » * - « > ' j f ( ' ; <) [^(e)] J W._ w (e) t=2
W
j=0
V
3
'
C.-C. HSIUNG & J. J. LEVKO
188 It is easy to see that
*-2
(2 (2.30)
/
x
j-l
(ID=(-I)V^[(-I)^Q+D-I)I(;)G_|)] 3-
.[^(e)]^-^)
= (-l)'+1ra'2(*)[^i(«)]JW.-i(e). because of (2.20) and
1 1 D-tff*) =-i +(fl • ^ Similarly, by replacing £ and j respectively by t + u and j — u with new t and j , and noticing that (£) and Hc-d are zero for 6 < 0 and d > c — 2, (III) of (2.29) can be written as
u=l
j=l
t=l
\
'
/
\
J
/
\
/
•[^ 1 (e)]'[H 1 (e)] j W s _t_ j (e) (2.31) s-3
(-l)s|:(-l)'+1Qc2V2^)[F1(e-)]t •'i:a(a7')[«iw]iw.-.-i(e), 1=1
\
•>
'
since
+
' S \ ft + U\ fs — t — U \i + u) \ u
)C .")C;-".-)-G)C;*)(0-
^ ( 0 -
Substituting (I) of (2.29), (II) of (2.30) and (III) of (2.31) in (2.27) gives immediately (2.21) for i = s, and hence Lemma 2 is proved. • Now let us proceed to prove our Theorem 4. From (2.13), (2.21) it follows immediately that (2.32)
[H2i(e)]n/{2i)
(2.33)
\H2i+1(e)\n/{2i+1)
dVn = [W 2 i(e)] B/(2 ° dVn, dVn = | W 2 i + 1 ( e ) | " / ( 2 i + 1 ) dVn.
Conformal Invariants of Submanifolds
189
Integrating both sides of (2.32), (2.33) over the corresponding unit (m - 1 ) spheres 5 m _ 1 , 5 m _ 1 , and using (1.10), (1.11) we obtain (2.34) K2i=K2i, (2.35)
| / r 2 i + 1 | = |tf 2 i+i|,
where K2i and K2i+i are denned by
(2.36)
R2i =
{^2i{e)}n/(2i) dam-AdVn,
(2.37)
|^ 2 i + 1 | = | j T m JW« + i(e)| B / ( W + 1 ) (» m _i}«iV n ,
\J-
ddm-i being the volume element of 5 m _ 1 at x. If M " is compact and oriented, then by integrating both sides of (2.34) and (2.35) over Mn we have (2.38)
/ JMn
K2i=
f JM"
K2i,
f
\K2i+1\=
JM"
f
\K2i+i\.
JM"
Hence our proof of Theorem 4 is complete.
•
REFERENCES
[1] [2] [3]
[4] [5]
[6] [7] [8] [9]
W. B L A S C H K E , Vorlesungen uber Differentialgeometrie. Ill, Springer, Berlin, 1929. B. Y. C H E N , An invariant of conformal mappings, Proc. Amer. Math. Soc. 4 0 (1973), 563-564. S. S. CHERN & J. SIMONS, Some cohomology classes in principal fiber bundles and their application to Riemannian geometry, Proc. Nat. Acad. Sci. U. S. A. 6 8 (1971), 791-794. L. P. ElSENHART, Riemannian Geometry, Princeton University Press, Princeton, 1949. J. H A A N T J E S , Conformal representations of an n-dimensional Euclidean space with a non-definite fundamental form on itself, Nederl. Akad. Wetensch. Proc. Ser. A 40 (1939), 700-705. C. C. HsiUNG & L. R. M U G R I D G E , Conformal invariants of submanifolds, Proc. Amer. Math. Soc. 6 3 (1977), 316-318. C. C. HSIUNG & L. R. MUGRIDGE, Euclidean and conformal invariants of submanifolds, Geom. Dedicata 8 (1979), 31-38. H. W E Y L , On the volume of tubes, Amer. J. Math. 6 1 (1939), 461-472. J. H. W H I T E , A global invariant of conformal mappings in space, Proc. Amer. Math. Soc 3 8 (1973), 162-164.
CHUAN-CHIH HSIUNG
J O H N J. L E V K O
Department of Mathematics Lehigh University Bethlehem, Pennsylvania 18015
Department of Mathematics University of Scranton Scranton, Pennsylvania 18510
Received October 16, 1986.
590 TENSOR, N. S.
VOL. 48(1989)
A CERTAIN CLASS OF ALMOST HERMITIAN MANIFOLDS.
By Lew FRIEDLAND and Chuan-Chih HSIUNG.
Abstract. This paper deals with a certain class of almost Hermitian manifolds, which includes Kaehlerian manifolds. We obtain, among other results, an analogue of Schur's theorem concerning the holomorphic sectional curvature and extensions of some known properties of Kaehlerian manifolds. § 1. Introduction. Let M be a Riemannian manifold of dimension 2n. The manifold M is said to be an almost Hermitian manifold if it admits an almost complex structure tensor F/ and a Riemannian metric g^ such that (1.1)
gijFn'F^g^,
where, by the usual tensor convention, when a Latin letter appears in a term as a subscript and also a superscript, then this letter is summed over its range. This paper is a study of the class of almost Hermitian manifolds M satisfying (1.2)
[Dj, Dk\Fih = (DjDk-DkDj)Fih
= 0,
where D denotes the covariant differentiation with respect to the affine connection r, with components / y * in local coordinates x1, •••, x2", of the Riemannian metric g. We shall refer to such manifolds as almost L manifolds and the corresponding structures F? as almost L structures. Obviously, Kaehlerian manifolds are almost L manifolds since M is Kaehlerian if (1.3)
DiFjk = 0
for all ij, k.
Section 2 (respectively, section 3) describes the fundamental notation, definitions and well-known results on Riemannian structures (respectively, almost complex structures) which are needed for later discussion. In section 4 several necessary conditions for an almost complex structure to be an almost L structure are expressed in terms of the Riemann and Ricci curvature tensors. In section 5, by studying various sectional curvatures including constant holomorphic sectional curvature, we obtain a geometric property of an almost L manifold, conditions for certain almost L manifolds and an analogue of Schur's theorem concerning the holomorphic sectional curvature of an almost L manifold. In section 6 we give some sufficient conditions for an almost L structure and certain other almost Hermitian structures to be Kaehlerian.
Received July 30, 1988 and in the revised form December 5, 1988.
591 A certain class of almost Hermitian manifolds.
253
Throughout this paper we suppose all structures and tensors to be of class C 2 at least. § 2. Riemannian structures. Let M be a Riemannian manifold of dimension m^2 with positive definite, symmetric Riemannian metric tensor gtj, and let || g" || be the inverse matrix of || g,j ||. Then the Riemann curvature tensor, the Ricci curvature tensor and the scalar curvature of M are given, respectively, by
(2. i)
R\jk=rtj/ax" - r,v dxJ+r,w , - r,yv*,
(2.2)
Rij = R ijk,
(2.3)
R = gijRa-
The Riemann curvature tensor Rhak further satisfies the identity (2.4)
Rhijk + Rhjki + Rhkij — 0,
where Rhijk — ghlR ijk-
The following identities are known, respectively, as the Bianchi identity and the Ricci identity: (2.5) (2.6)
D,Rhtjk + DjRhikl + DkRhiu = 0, DiDjTkh-DjDiTkh=
s T V * * * , - TshR'~kji,
where Tkh is an arbitrary tensor of type (1, 1). The sectional curvature of the manifold M at a point p with respect to the twodimensional plane determined by two linearly independent tangent vectors u* and v' of M at p is given by (2 7)
K=
Rhijkuhv'uJvk
(ghkgij- ghjgik) uhviuivk'
which is the Gaussian curvature of the two-dimensional geodesic submanifold of M tangent to the plane at p. If, at every point of the manifold, the sectional curvature does not depend on the two-dimensional section through the point, then (2-8)
Rhijk =
K(ghkgij—ghjgik)-
The manifold is said to be locally Euclidean or locally flat if K = 0, i.e., if Rhijk = Q- For non-zero K, from (2.5) and (2.8) it is easy to show that (2-9)
Rij = (m-\)Kgu,
(2.10)
R = m(m-l)K,
and for m > 3, K and consequently R are absolute constants on the manifold and M is said to be of constant curvature. Further, M is an Einstein manifold as a consequence of (2.9). § 3. Almost complex structures. In this section M is a Riemannian manifold as in section 2 but with dimension m = 2n. If there exists on M a tensor F/ of type (1,1) satisfying
592
254
L. Friedland and C.-C. Hsiung. F/Fjk=-d?,
(3.1)
where Sk are the Kronecker symbols defined by
then F,' is said to define an almost complex structure on M. From (3.1) it follows that the almost complex structure Fi induces an automorphism J : TPM —• TPM of the tangent space TPM of M at each point p in M such that for each vector vkeTpM, (3.2)
J(vk) = Ft"v'.
Since J(v) is orthogonal to v and J2= —I, where / is the identity operator on TPM', the tangent space of M at each point has an induced complex structure and the manifold M is said to be almost complex. Further, if the tensor Fi satisfies (1.1) or equivalently, g{u, v) = g(J(u), J(v)), then Fi defines an almost Hermitian structure on M and the manifold M is said to be almost Hermitian. In this section and the remaining sections when we refer to an almost Hermitian structure Fi, we shall mean the structure tensor Fi and the Riemannian metric gu satisfying (1.1). By using the multiplication of matrices, from (3.1) we readily see that a necessary condition for the existence of an almost complex structure on a Riemannian manifold M is that the dimension of M be even, and therefore that an almost complex manifold is always orientable and the orientation depends only on the tensor F/. As a consequence of (3.1) and (1.1) the tensor Fu of type (0, 2) defined by FiJ = gJkFik
(3.3)
is skew-symmetric. Thus on an almost Hermitian manifold M there is a differential form (3.4)
Qj^Fijdx'Adx''.
If the differential form co is closed, that is, if (3.5)
dco = 0,
then Fi is said to be an almost Kaehlerian structure and M an almost Kaehlerian From (3.4) and (3.5) it follows that an almost Kaehlerian structure satisfies (3.6)
FH^D.Fij
+ DiFjH +
manifold.
DjF^O.
The tensor Fhij is skew-symmetric in all indices. An almost Hermitian structure Fi (respectively, manifold) satisfying (3.7)
Fi=-DjF/
= 0,
is said to be an almost semi-Kaehlerian structure (respectively, manifold). In particular, the structure Fi is Kaehlerian if (1.3) holds, and consequently the Nijenhuis torsion tensor (3.8)
Nij" = Ft (DhFf - DjFhk) - F/ (DhFik - DtFhk)
is zero, so that the integrability condition of the almost complex structure Fi is satisfied. In general, when iV;/ = 0 an almost Hermitian structure is defined to be Hermitian. If an almost Kaehlerian structure is integrable then the structure is Kaehlerian. We have
593 A certain class of almost Hermitian manifolds.
255
as well that an almost Kaehlerian structure is almost semi-Kaehlerian. Consequently Ff is harmonic. Proof. Multiplying (3.3) by Fhi and using (3.1) yield (3.9)
FijF^^-dl
By covariant differentiation of (3.9) noting that FijDhFij=Q,
(3.10)
and using the notation of (3.6) and (3.7) we have (3.11)
F^F'^lFjFi.
Thus an almost semi-Kaehlerian structure Ff satisfies (3.12)
FhiJF" = 0.
Multiplication of (3.11) by Fkh and use of (3.9) give Fk=-^-FhijF»Fkh.
(3.13)
From (3.6), (3.7) and (3.13) we conclude that an almost Kaehlerian structure or manifold is almost semi-Kaehlerian. An almost Hermitian structure Ff (respectively, manifold) satisfying (3.14)
Gi/ = DiFjk + DjFik = 0
is an almost Tachibana or nearly Kaehlerian structure (respectively, manifold). Since Ff — g'JFjj = 0, from (3.7) and (3.14) it follows that a nearly Kaehlerian manifold is almost semi-Kaehlerian. Further, it is known that a nearly Kaehlerian manifold is Kaehlerian if the structure is integrable. Let M be an almost Hermitian manifold with an almost complex structure Ff satisfying (1.1). Then the two-dimensional plane determined by an arbitrary tangent vector v' of M and the tangent vector J (v() at a point p is said to be a holomorphic plane, and the sectional curvature with respect to a holomorphic plane to be the holomorphic sectional curvature of M at p. If the holomorphic sectional curvature at a point p is independent of the holomorphic plane through p, then M is said to have constant holomorphic curvature at p. § 4. Almost L structures. In this section we shall express the defining condition of an almost L structure, given in section 1, in terms of the Riemann and Ricci curvature tensors. Lemma 4.1.
An almost Hermitian structure Ff is an almost L structure if and only
if (4.1)
FfR\k
=
FshRsijk.
The result is an immediate consequence of the Ricci identity (2.6) applied to the tensor Ff. Lemma 4.2.
On an almost Hermitian manifold condition (4.1) is equivalent to
594 256
L. Friedland and C.-C. Hsiung.
(4.2)
Rhijk — FhrFfRrsjk • Proof. Suppose (4.1) holds. Then multiplying (4.1) by Fj and using (3.1) we obtain — R rjk = Fr'Fs
RSijk
Or
Rhrjk
=
Fr'FhSRsijk-
Conversely, suppose (4.2) holds. Multiplying (4.2) by F{ we obtain Fi'Rhijk= — Fh Rruk
FrRrijk.
or Fi'R iJk =
Lemma 4.3. On an almost L manifold each of the following four conditions holds Rxt=-\F/Fk,Rhileh
(4.3)
(4.4)
7ijpkl g=i!+JF"F"i? t f M =0,
(4.5)
Rij = FihFJkRhk,
(4.6)
R=
FihFj"RJhki,
where R is given by (2.3). Proof. ( i ) By Lemmas 4.1 and 4.2, condition (4.2) holds on an almost L manifold. Multiplying (4.2) by gkk gives Rij = FkrFisRkJsr.
(4.7) By the identity (2.4) we obtain F
r
FfRkjsr=
~ F — —r
Ff(Rksrj~\~
Rkrjs)
rt Krskj— *
?i Kjskr — — r
t t Kkjsr~
r
fiKjSkr,
which with (4.7) implies (4.3). ( i i ) (4.4) is obtained by multiplying (4.3) by ghi. (iii) Using (4.7) and Lemma 4.1 we conclude that D — p s c Tnk
(iv)
— psr(D
Multiplying (4.7) by giJ gives R = FkrFjsRrsJk
=
FrkFjsRr skj
-
Remark. It is obvious that condition (4.3) is more general than (4.2), and condition (4.6) is, in turn, more general than (4.3). § 5. Sectional curvatures. the sectional curvatures.
In this section we characterize an almost L structure by
Theorem 5.1. If M is an almost L manifold with almost complex structure J, then the sectional curvature K(u, v) of M at a point p with respect to the plane determined by two arbitrary tangent vectors u and v of M at p is equal to the sectional curvature K(J(u), J (v)) with respect to the plane determined by the tangent vectors J (u) and J (v) at p. Proof. From (2.7), (3.2), (4.2), and (5.1)
FtJFkJ = gik,
595 A certain class of almost Hermitian manifolds.
257
we obtain K(U„\ K (J(U),
RhiikFPhupFqivqFriurFkvs
r(,,v> J (V))-
(ghkgtJ-ghjgik)Fp>>uPFq
RpqrsUPVqUrVs m s p a
r
„ , s ~K{U,
s V).
It should be noted that Theorem 5.1 may be considered as a geometric property of an almost L manifold. Theorem 5.2. On an almost L manifold of real dimension 2n, n>2, if the sectional curvature K(u, v) is constant for any linearly independent tangent vectors u and v of M at each point, then the curvature K is zero, and M is locally Euclidean. Proof. The theorem follows on multiplying (2.8) by ghl, substituting the resulting equation into (4.1) and multiplying by diFor a Kaehlerian manifold Theorem 5.2 reduces to the following known result. Corollary 5.1. If the sectional curvature of a Kaehlerian manifold M at each point is the same for all possible planes, then the sectional curvature is zero, and M is Euclidean. Lemma 5.1. A necessary and sufficient condition for a Riemannian In-manifold M with an almost complex structure F/ to have constant holomorphic sectional curvature K at each point with respect to the almost complex structure, where K may depend on the point of the manifold, is that the Riemann curvature tensor satisfy (5.2)
(Rrqsk + Rrksq) FprFf
+ (Rrisk 4" Rrksi) FpTFqs-\-(Rrisq~t~ Rrqsi) FpTFks
+ (RrpSk + Rrksp) FtrFqs + (Rrpsq + Rrqsp) FirFkS + (R,psi + R^p)
r Fk
Fqs
= -4K {gpigqk + giqgPk + gqPgik). Proof. To prove the necessity of condition (5.2) we suppose that M has constant holomorphic sectional curvature K. Then by using (2.7), (5.1) and Fpkupuk-Q, we obtain the holomorphic sectional curvature K with respect to an arbitrary tangent vector u' oi M : K^
RrislcFprupuiFqsuquk gpigqkupuiuquk
Thus (5.3)
RriskFprupuiFqsuquk=-Kgpiupuigqkuquk
holds for any tangent vector u' of M. By collecting all the coefficients of a general term upu'uquk on both sides of (5.3) through interchanging the indices p, i, q, k in all possible cases, i.e., interchanging a pair of them and keeping the other pair fixed, interchanging both pairs, keeping one index fixed and interchanging the other three, and interchanging all of them without keeping any one fixed, we can obtain (5.2). To prove the sufficiency of condition (5.2) we suppose that (5.2) holds. Multiplying both sides of (5.2) by ^u'u"^ for any tangent vector u' of M and summing for p, i, q, k we can see that all the terms on the left-hand side of (5.2) are then equal to each other, and the same is true of all the terms on the right-hand side. Thus (5.3) holds for any tangent vector u' of M, and M has constant holomorphic sectional curvature at each point.
596
258
L. Friedland and C.-C. Hsiung. Lemma 5.2.
If the manifold M in Lemma 5.1 satisfies
(5.4)
Rhijk =
FfFfFfFfRp,,,,
then condition (5.2) becomes (5.5)
(Rrisk
+ Rrksq) FiTFpS + (Rrpsk + Rrksp)
+ Rrksi) Fp Fq -\-(Rrqsk
FqrFiS
= —2K (gPigqk + giqgPk + gqPgik). If M is an almost L manifold, then condition (5.2) becomes RnskFprFqs + RrqSkFi"Fps + RrpskF/F? = — K {gpigqk + gtqgpk + gqpgtk)-
(5.6)
Proof. It is easy to see that (5.4) implies (5.7) (5.8)
FrhFs'Rhijk =
F/Fk'RPqrs,
J
Fq Fs Rhijk — Fi
FkrRPqrs.
Substituting (5.8) in (5.2) yields (5.5) immediately. To prove the second part of the lemma we first notice that from (4.2) it follows that an almost L manifold satisfies condition (5.4). Then by using (4.2) and the identity (2.4) we can reduce (5.5) to (5.6). Theorem 5.3. A necessary and sufficient condition for an almost L manifold M of real dimension 2n to have constant holomorphic sectional curvature K at each point, where K may depend on the point of the manifold, is that the Riemann curvature tensor be given by (5.9)
RhiJk=-^[{9hkgij
- ghjgik) + (FhkFiJ - FMFik) - 2FhiFJk\,
where „ and F/ are, respectively, the tensors of the Riemannian metric and the almost L structure of M. Further, the Ricci tensor and scalar curvature of such a manifold with constant holomorphic sectional curvature K are given, respectively, by (5.10)
R^-^Kga,
(5.11)
R=
n(n+l)K.
As a consequence of (5.10) M is an Einstein manifold. Further, if M is compact and K is positive, then the first Betti number of M is zero. Proof. For the necessity of condition (5.9) we suppose that M has constant holomorphic sectional curvature K. Then by Lemma 5.2 we have (5.6). Multiplying (5.6) by FhpFjq and using (4.2) and (5.1) yield (5.12)
RWk-RijHk-RjpskFisFhp^~K(FhiFJk-FhkFij+ghJgik).
While in consequence of (2.4) and (4.2) we have (5.13)
RjPskFisFh —RjpskFhsFi —(Rjpsk — Rjspk) FtsFh — ~ Rj/cpsFi FH = — Rhuk •
By interchanging h and i in (5.12), subtracting the resulting equation from (5.12), and using (5.13) we obtain
597 A certain class of almost Hermitian manifolds. 2Rhijk - Ruhk + RhjiK + RMjk=-K
259
[2FhiFJk - {ghkgij - ghjgik) -{FhkFij -
FhjFik)],
which implies (5.9). For the sufficiency of condition (5.9) we suppose that (5.9) holds. By using (5.9) the left side of (5.3) is reduced to -^-{FPkFqi + FqPFik-gikgqp-gpkgiq-2gPigqk)
upu'u"uk =
-Kgpigqkupuiuqu\
so that (5.3) holds for any tangent vector u' of M. Hence M has constant holomorphic sectional curvature at each point. (5.10) is obtained by multiplying (5.9) by ghk, and (5.11) by multiplying (5.10) by gu. Since for K>0, by (5.10) the Ricci curvature of M is positive definite, the last part of the theorem follows from a well-known theorem of S. Bochner (see e.g., [3] 1 ' p. 37). Remark. In the above proof of the sufficiency of condition (5.9) we require only an almost Hermitian structure rather than an almost L structure. So if (5.9) holds for an almost Hermitian manifold M, then M has constant holomorphic sectional curvature at each point. Concerning the holomorphic sectional curvature of an almost L manifold we have the following analogue of Schur's theorem. Theorem 5.4. If an almost L manifold has non-zero constant holomorphic sectional curvature K at each point, then K is an absolute constant on the manifold, and the manifold is said to be of constant holomorphic curvature. Proof. Covariantly differentiating (5.9) and similar equations, substituting these in the Bianchi identity (2.5), multiplying the resulting equation by g"ghl, and making use of F/ = 0 , FijFjk=—gik and FhiFhi — 2n we obtain (n2-l)DkK
= 0 for any A:.
Thus K is an absolute constant for n>2. For a Kaehlerian manifold Theorem 5.4 reduces to the following known result. Corollary 5.2. If a Kaehlerian manifold M has constant holomorphic sectional curvature at each point, then M is of constant holomorphic curvature. Theorem 5.5. If M is an almost L manifold with nonzero constant holomorphic sectional curvature at each point, then M is almost Kaehlerian. Proof. Substituting (5.9) and two similar equations in the Bianchi identity (2.5) gives (5.14)
D, (FucF,, - FvFik - 2FhiFJk) + Dj(FhlFik - FhkFu - 2FhiFkl) + Dk (FuFa - FhlFtj - 2FhiFu) = 0,
since, by Theorem 5.4, K is an absolute constant on M. Expanding (5.14) by differentiating covariantly, multiplying the resulting equation by Fih and using (3.10) we obtain DlFjk + DJFkl + DkFu = 0. Thus, M is almost Kaehlerian.
1) Numbers in brackets refer to the references at the end of the paper.
598 260
L. Friedland and C.-C. Hsiung.
§ 6. Kaehlerian structures. In this section we give various sufficient conditions for an almost Hermitian structure to be Kaehlerian. We give two results [1], the latter we shall need later on. Lemma 6.1.
An almost Hermitian structure Fi satisfying
(6.1)
DtFjk = DjFik
is Kaehlerian. Proof. By the skew-symmetry of the tensor Fu it follows that (6.2)
DtFjk
+DtF^O.
Taking the sum of (6.2) and the two similar equations obtained from it by cyclic permutations of the indices i,j and k and using (6.1) we obtain DiFjk +
DiFkj+DjFkt=0,
and subtracting (6.2) from this gives DjFki = 0. Lemma 6.2.
An almost Hermitian structure Fi satisfying FiJDkDkFij = 0
(6.3) is Kaehlerian, where Dk = gikDj. Proof. From (3.10) we have 0 = DkDk(FijFij)
= 2(FiJDkDkFIJ
+
DkFijDkF"),
which, together with (6.3), gives DkFijDkFii = Q and therefore DkFij-Q. The following theorem gives relations among almost L structures, almost Kaehlerian structures and nearly Kaehlerian structures. Theorem 6.1. (6.4)
If an almost semi-Kaehlerian structure Fi satisfies D.FijD'FJ" =
aDkFijDkFii,
where a is a non-zero constant, then Q defined by (6.5)
Q=
R+-^Fh'FJkRkak
is non-negative or nonpositive according as a is positive or negative. Further, the structure Fi is Kaehlerian if and only if Q = 0. Proof. From (3.1) and (3.7) it follows that Dk(FiFjk) =
F/DkFi=0,
and therefore (6.6)
D' (F/DkFi)
= DkFiDlFk
+ F/D'D^Fi = 0.
However, using (2.6) and (3.7) we obtain (6.7)
D'DkFi = D{DkFij = FihR\ki
which, together with (2.4), (6.5) and FjkFhJ= -ghk,
+
F^R^,
implies that
599 A certain class of almost Hermitian manifolds. FjkD'DkF/ = FjkFihRjhki
(6.8)
261
-R=-Q.
Combining (6.4), (6.6) and (6.8) we then have (6.9)
aDkFuDkFij,
Q=
which is non-negative for a > 0 and non-positive for a < 0 . In particular, Q — 0 if and only if DkFjj = 0, that is, if and only if the structure is Kaehlerian. The following two corollaries [2] are immediate consequences of Theorem 6.1. Corollary 6.1.
If an almost Hermitian structure F/ is nearly Kaehlerian, then
(6.10)
g^O,
where equality holds if and only if the structure is Kaehlerian. Proof. From (3.14) it follows immediately that a—I. Corollary 6.2.
If an almost Hermitian structure F/ is almost Kaehlerian, then
(6.11)
Q<0,
where equality holds if and only if the structure is Kaehlerian. Proof. From (3.6) it follows that a= —«- since DtFijD'F* = - DkFij (DkFiJ + DJFki) = - DkFijD kFiJ - DkFHD 'F kJ = -DkFiJDkFiJ-DkFiJDiFJk. By Theorem 5.5, Corollary 6.2, and (4.4), Theorem 5.5 may be extended as follows. Corollary 6.3. If M is an almost L manifold of real dimension In with non-zero constant holomorphic sectional curvature at each point, then M is Kaehlerian. Using Corollary 6.3 and a well-known theorem of S. Bochner for Kaehlerian manifolds (see e.g. [3] p. 161) we immediately obtain Corollary 6.4. The Betti numbers bt of a compact almost L manifold of real dimension In with positive constant holomorphic sectional curvature at each point are given by (6.12)
hP=l, fep+i = 0,
(0^2/><2/>+1^2H).
Corollaries 6.1, 6.2 and Lemma 4.3 imply that an almost L structure Ff is Kaehlerian if it is either a nearly Kaehlerian structure or an almost Kaehlerian structure. The hypothesis of this result may be weakened slightly as in the following lemma: Lemma 6.3. either (6.13)
If an almost L structure Fi is almost semi-Kaehlerian and satisfies DkGhij = 0,
or (6.14)
DkFhli = Q,
where GhiJ and Fhij are defined in (3.14) and (3.6) respectively, then the structure is
600 262
L. Friedland and C.-C. Hsiung.
Kaehlerian. Proof. From (1.2) it follows that (6.15)
DkDhF/ = DhDkF/.
Multiplying (6.15) by g'k and using (3.7) give (6.16) ( i)
D
(6.17)
D'DiFn^O.
Thus the structure Ff is Kaehlerian by Lemma 6.2. ( i i ) Suppose (6.14) holds. Similarly as in ( i ) we obtain (6.17) by multiplying (6.14) by 81 Lemma 6.4. If M is a compact almost Hermitian manifold with almost Hermitian structure Ff and f (DiFihDkFiJ
(6.18)
+ FhkF"RhiJk-R)
dV = 0,
JM
where dV is the volume element of M, then M is almost semi-Kaehlerian. Proof Let ^i = F3kDkFiJ. Then by Green's theorem we have [Di£idV
(6.19)
= 0.
JM
Using the Ricci identity (2.6), condition (6.18) and the relation FjhFik — ghk, we compute Dk (FJkDkFtj), so that (6.19) becomes f FJkDkDtF/d
(6.20)
V=0.
JM
By Green's theorem we obtain as well (6.21)
(Dk {FJkDiF/) dV={ JM
(DkF*DiFf
+ FJkDkD(Ff)
dV = 0,
JM
which, together with (6.20) gives (6.22)
fDkFJkDiF/dV
= 0.
JM
Since DkFikDtF} is non-negative, from (6.22) it follows that DtF/ = 0, so that M is almost semi-Kaehlerian. Corollary 6.5. If M is an almost Hermitian manifold of real dimension In, and the Riemann curvature tensor satisfies (5.9), then M is Kaehlerian with constant holomorphic sectional curvature. Proof. By the remark following Theorem 5.3, M has constant holomorphic sectional curvature. Further, M is an almost L manifold since, given (5.9), condition (4.2) holds. The result then follows from Corollary 6.3.
601 A certain class of almost Hermitian manifolds. _ „ Department of Mathematics .r . . „, T „, . State University of New York Geneseo, New York 14454 U
' S'
A
'
263
Department of Mathematics _, . _, TT „ „, . Christmas-Saucon Hall #14 „ ,,, ™» , o«,,r Bethlehem, PA. 18015 U. S. A.
REFERENCES.
[1] C. C. Hsiung and J. J. Levko : Complex Laplacians on almost-Hermitian manifolds, J. Differential Geometry, 5 (1971), 383-403. [2] S. Koto : Some theorems on almost Kaehlerian spaces, / Math. Soc. Japan, 12 (1960), 422-433. [3] K. Yano and S. Bochner : Curvature and Betti numbers, Annals of Math. Studies, No. 32, Princeton University Press, Princeton, 1953.
602
Annali di Matematica pura ed applicata (IV), Vol. CLXVIII (1995), pp. 133-149
A New Class of Almost Complex Structures (*). CHUAN-CHIH HSIUNG - BONNIE XIONG
Abstract. - The purpose of this paper is to introduce a new class of almost complex structures J on a Riemannian manifold M by using a certain identity for the relationship between the tensor FJ of J and the Riemann curvature tensor Rhijk ofM. This class contains the Kahlerian structures, and its relationship with some known classes of almost Hermitian structures defined by similar identities is discussed. For convenience we call each structure of this new class an almost C-structure, and a manifold with an almost C-structure an almost C-manifold. We obtain an analogue of F. Schur's theorem concerning the holomorphic sectional curvature of an almost Hermitian C-manifold, and some sufficient conditions for an almost Hermitian C-manifold to be Kahlerian. We show that these results are also true for a manifold with a complex structure.
1. - Introduction. Let M be a Riemannian 2w-manifold, and g^, Ftj and Rhijk be respectively the components of a Riemannian metric tensor g, the tesnor of an almost complex structure, and the Riemann curvature tensor, with respect to g, of M. Throughout this paper all Latin indices take the values 1 , . . . ,2n unless stated otherwise. By using the following identities for the relationship between Fj and Rhijk, we can define three wellknown classes of almost complex structures on the manifold M: (1-1)
Rhijk =
d-2)
Rhijk = FhFiSRrsjk
(1-3)
Rhijk
=
F^F^R^ic, + FhFfRrUk
+ FhFkRrijs
>
FhpFiqFjrFk&RPQrs,
where the repeated indices imply summation. F o r the class JE of almost complex structures (or manifolds), let £t denote the sub-
(*) Entrata in Redazione il 18 marzo 1993. Indirizzo degli AA.: CHUAN-CHIH HSIUNG: Department of Mathematics, Lehigh University, Bethlehem, Pennsylvania 18015; BONNIE XIONG: Department of Mathematics, University of Scranton, Scranton, Pennsylvania 18510.
603 134
C.-C.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
class of almost complex structures (or manifolds) satisfying (li), then we have (1.4)
£1c£2c£3c£.
Since the class X of the Kahlerian structures (or manifolds) belong to J£I , as i decreases the structures (or manifolds) in J£J resemble Kahlerian structures (or manifolds) more closely. The structures (or manifolds) of class £ x were studied first by E. CARTAN [1] as a direct generalization of the notion of symmetric spaces, and the by A. LICHNERROWICZ [6] and many other authors, and were said to be para-Kahler by R. B. RIZZA[7], and semi-symmetric by Z. I. SZAB6[9], and were called i^-structures (or F-manifolds) by S. SAWAKI and K. SEKIGAWA [8] and almost L-structures (or L-manifolds) by C. C. HSIUNG and L. FRIEDLAND [5]. The structures (or manifolds) of class £ 3 were called .Rif-strucutres (or /ZX-manifolds) by L. VANHECKE [10], A. GRAY and L. M. HERVELLA [2] have also used identities, but expressed in terms of covariant derivatives, to define the sixteen classes of all almost Hermitian manifolds. On the other hand, C. C. HSIUNG [3], [4], has recently established the following theorem. THEOREM 1.1. - If F/ are the components of a complex structure on a Riemannian 2/n-manifold M without a constant sectional curvature, which is zero or nonzero, then
(1-5)
FiiFijRijisk + Fi?Fi!Rijhk + FhFiiRiji2k
=0
for all ii,i2, is and k. In particular, ifM is of constant sectional curvature K with respect to g, then FJ satisfy (1.5) or not according as K is zero or not. It should be remarked that when 2n < 4, at least two of the four indices i\, ii, h, k will be the same, and therefore (1.5) will hold automatically. Now we use condition (1.5) to define a new class of almost complex structure, and call an almost complex structure Fj satisfying this condition (1.5) an almost C-structure, and the coresponding manifold an almost C-manifold. The purpose of this paper is to study this new class, denoted by Q, of almost C-structures (or C-manifolds). Paragraph 2 (respectively, § 3) contains the fundamental notation, definitions and well-known results on Riemannian structures (respectively, almost complex structures) which are needed for the later discussions. In § 4 we first prove the formula (1.4), which is known for almost Hermitian structures, and then use the almost C-structures to introduce a new classification of all almost complex structures on a Riemannian manifold, together with a relationship between this classification and the old one. Finally some necessary conditions for an almost complex structure to be an almost C-structure are expressed in terms of Riemann curvature and Ricci curvature tensors. Paragraph 5 deals with the almost Hermitian C-structures by means of the holomorphic sectional curvatures. At first a necessary and sufficient condition is given for
604 C.-C. HsiUNG - B.
XIONG:
A new class of almost complex structures
135
an almost Hermitian C-structure on a Riemannian manifold M to have constant holomorphic sectional curvature K with respect to an almost Hermitian metric at each point of the manifold M. Then we use the condition to obtain an analogue of F. Schur's theorem, which states that if if is nonzero, then K is an absolute constant. Finally we find a sufficient condition for such a manifold M with nonzero K to be Kahlerian. Paragraph 6 is an application of § 5 to a manifold with a Hermitian complex structure.
2. - Riemannian structures. Let M be a Riemannian manifold of dimension m 3= 2 with positive definite, symmetric Riemannian metric tensor g^, and let ||flr*'|| be the inverse matrix of \\gy\\. Then the Riemann curvature tensor, the Ricci curvature tensor and the scalar curvature of M are given, respectively, by
(2.1)
R% = rfj/dx" - rA/dx* + r,\A$ - ri)rik,
(2.2)
Ry = R ijk,
(2.3)
R = 9ijRij,
where r fj are the components, in local coordinates x1, ..., xm, of the affine connection
rofg. The Riemann curvature tensor Rkijk further satisfies the identity (2.4)
Rhijk + Rigid + Rhkij ~ 0 >
where Rhijk = ghiRlijk- In geneal, we shall follow the usual tensor convention that indices can be raised and lowered by using gij and g$ respectively. The following identities are known, respectively, as the Bianchi identity and the Ricci identity: (2.5)
^iRhijk + VjRhikl + VkRhUj = 0 .
(2.6)
V, V,- Tkh - Vj V, Tkh = Tk'Rh# -
TshRskji,
where V denotes the covariant derivation with respect to T, and Tk is an arbitrary tensor of type (1.1). The sectional curvature of the manifold M at a point P with respect to the two-dimensional plane determined by two linearly independent tangent vectors ul and vl of M at P is given by RkiikUhViUiVk
(2-7)
K=
^ , k i i k(9hk9ij-9hj9ik)u'lvlu:>vlc
605 136
C.-C.
HSIUNG
- B.
A new class of almost complex structures
XIONG:
which is the Gaussian curvature of the two-dimensional geodesic submanifold of M tangent to the plane at P. If, at any point of the manifold, the sectional curvature does not depend on the two-dimensional section through the point, then (2.8)
Rhijk = K( 9hk 9ij ~ 9hj 9ik) •
The manifold is said to be locally Euclidean or locally flat if K = 0, i.e., if Rhijk = 0. For nonzero K, from (2.5) and (2.8) it is easy to show that (2.9)
Rij =
(2.10)
R =
m(m-l)Kgij, m(m-1)K,
and for m S 3, K and consequently R are absolute constants on the manifold M, and M is said to be of constant curvature. Further, M is an Einstein manifold as a consequence of (2.9). 3. - Almost complex structures. In this section M is a Riemannian manifold as in § 2 but with dimension m = 2n. If there exists on M a tensor Fj of type (1,1) satisfying FJFjk=-6\,
(3.1)
where Sk are the Kronechker deltas defined by °i
— •
1, 0,
% = k, i *k,
then FJ is said to define an almost complex structure on M. From (3.1) it follows that the almost complex structure Fj induces an automorphism /:
TpM-+TpM
of the tangent space TpM of M at each point P in M such that for each vector vk e TPM, (3.2)
J(vk) = Fikvi.
Since J(v) is orthogonal to v and J2 = - I where / is the identity operator on TpM, the tangent space of M at each point P has an induced complex structure and the manifold M is said to be almost complex. Further, if the tensor Fj satisfies (3.3)
QijF^F^gu,
or equivalently, g(u, v) = g(J(u), J(v)), then Fj is said to define an almost Hermitian structure on M and the manifold is said to be almost Hermitian. In this section and the remaining sections when we refer to an almost Hermitian structure Fj we shall mean that the structure tensor Fj and the Riemannian metric g^ satisfy (3.3).
606 C.-C.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
137
By using the multiplication of matrices, from (3.1) we readily see that a necessary condition for the existence of an almost complex structure on a Riemannian manifold M is that the dimension of M be even. It is also well-known that an almost complex manifold is always orientable and the orientation depends only on the tensor As a consequence of (3.1) and (3.3) the tensor Fy of type (0,2) defined by Fii = gjkFik
(3.4)
is skew-symmetric. Thus on an almost Hermitian manifold M there is a differential form OJ = Fijdx1 A dxj,
(3.5)
where A is the exterior multiplication. If the differential of w is closed, that is, if (3.6)
dco = 0, j
then Ft is said to be an almost Kahlerian structure, and M an almost Kahlerian manifold. From (3.5) and (3.6) it follows that an almost Kahlerian structure satisfies (3.7)
Fm = lh Fij + Vi Fih + V, Fhi = 0.
The tensor Fhij is skew-symmetric in all indices. An almost Hermitian structure F/ (respetively, manifold) satisfying (3.8)
F{= - VjFi = 0
is said to be an almost semi-Kahlerian structure (respectively, manifold). In particular, the structure F? is Kahlerian if V{Fk = 0 holds, and consequently the torsion tensor (3.9)
N/
= F%h(V»^* - VjFkk) - Fjk(VhF{k - V ^ * )
is zero, so that the integrability condition of the almost complex structure Fj is satisfied. In general, when Nyk = 0, an almost Hermitian structure is define to be Hermitian. If an almost Kahlerian structure is integrable, then the structure is Kahlerian. We have as well that an almost Kahlerian structure is almost semi-Kahlerian. Consequently Ftj is harmonic. PROOF.
(3.10)
- Multiplying (3.4) by Fhi and using (3.1) yield FijFhi=
-Sf.
By covarient differentiation of (3.10), noting that (3.11)
F*VhFv = 0,
607 138
C.-C.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
and using the notation of (3.7) and (3.8) we have FhijFiJ = 2FhiFi.
(3.12)
Thus an almost semi-Kahlerian structure F£ satisfies (3.13)
* V * = 0h
Multiplication of (3.12) by Fk and use of (3.10) give
(3.14)
**=-|*VW.
From (3.7), (3.8) and (3.14) we conclude that an almost Kahlerian structure or manifold is almost semi-Kahlerin. An almost Hermitian structure F^ (respectively, manifold) satisfying ViFf + VjFik = 0
(3.15)
is an almost Tachibana or nearly Kahlerian structure (respectively, manifold). Since Fi = gijFy = 0, from (3.8) and (3.15) it follow that a nearly Kahleerian manifold is almost semi-Kahlerian. Further, it is known that a nearly Kahlerian manifold is Kahlerian if the structure is integrable. Let M be an almost Hermitian manifold with an almost complex structure Ftj satisfying (3.3). Then the two-dimensional plane determined by an arbitrary tangent vector v% of M and the tangent vector J(vl) at a point P is said to be a holomorphic plane, and the sectional curvature with respect to a holomorphic plane to be the holomorphic sectional curvature of M at P. If the holomorphic sectional curvature at a point P is independent of the holomorphic plane through P, then M is said to have constant holomorphic curvature at P.
4. - Almost C-structures. The relation (1.4) for almost Hermitian structures is well known; for completeness we include its proof here. To prove <£iC<£2, i.e., (1.1) implies (1.3), we suppose that (1.1) holds. From (1.1) and (3.1) it follows that FhrFksRrijs
= FkrFksRjsri
= FhrFkeFfFs"Rpqri
= -
Fh'F/Rpkri.
Then by substituting this for the third term on the right-hand side of (1.2), we obtain (1.2). For proving £ 2 c £3 we assume that (1.2) holds. Substituting (1.2) for each term on its right-hand side and using (3.1) we can easily obtain Rhijk = FfFfRhsqk + Ft"FkgRhsjq + FjSFkqRhisq . Thus by substituting this in the right-hand side of (1.3) we arrive at (1.3).
608 C.-C.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
139
For similar properties of almost C-structures we have THEOREM 4.1.
(4.1)
jdcecJE,,
(4.2)
£2 n e = £x.
PROOF. - To prove £i c Q, suppose that (1.1) holds. Substituting (1.1) for each term on the left-hand side of (1.5) and using (2.4) we then obtain (1.5). To show C c £ 3 , for convenience we use {i^i^k) to denote the left-hand side of (1.5), and set
' Ax = F^F^R^,
A2 =
F^FJRnk,
i
(4.3)
• At = Fk Fk'R^litt
A^F^Fi/Rij^,
_ A5 = Fi^F^Rij^k, Then using (i3i2iiA:) and {izki2i\)
A6 =
Fi3lFkJRijili2.
we obtain, from (1.5), respectively
A2=A1+A6,
A2=AZ+A6,
and therefore (4.4)
A1+A5=A3+A6.
Similarly, from (fci^is) and ( f c ^ i ^ ) it follows that (4.5)
A 4 = A 1 + A 3 = A B + A6,
which, together with (4.4), implies that A3=AS, (4.6)
F^Fi/R^ k
=
i.e., F^FJR^.
h
Multiplying (4.6) by Fb Fa and using (3.1) we obtain (1.3). Combining the above two results thus gives (4.1). To prove (4.2) we need LEMMA 4.1. - (1.5) implies
FCRijhh + FhRwh
(4.7) PROOF.
(4.8)
+ KiRvkk
= °•
- Suppose that (1.5) holds. Then substituting (4.6) in (1.5) yields FJFJ^ k
+ F^F^R^
+ Fj Fj Rmk
= 0.
Multiplying (4.8) by Fh and using (3.1) we thus obtain (4.7).
q.e.d.
For convenience we denote by 5C the class of the almost complex structures Fj satisfying (4.7). Then by Lemma 4.1, Cc5C, so if £2C\X=£X, then £2f\e = £x.
609
140
C.-C. HSIUNG - B. XIONG: A new class of almost
complex
structures
Now suppose that Ftj e X. Then multiplying (4.7) by Fk%s and using (3.1) we get
On the other hand, from the identity (2.4) we have (4.10)
— tii^ij
— iti3ijiz ~ Ki3jii2>
(4.11)
—Kisi1ji=
— /tjjjiji + Itisiii] •
Substituting (4.10) and (4.11) in (4.9) gives (4.12)
+ FhRikki">
Ricn^ = ~ Fk^iFi^Riiiajh
~ FkHFklRim2
- FkHF^Rhjili
• • - F,'SF- lR
= F.^F^R-
=
••• - F,isF- 'R ••
where the last step is a consequence of (4.7). F o r £2 H X, F{* satisfies (1.2) so that (4.13)
Rkjhh = Fk^FjRihhk
+ Fkl3FiilRisjii2 +
F 3
k Fi2lRi3ii1i-
Combining (4.12) and (4.13) gives the following condition for Fj e £2: (4 14)
FJ3F- lR ••• + F, *3F- {R- •• • = 0
Multiplying (4.14) by FrkFsh we thus obtain (1.1), and hence £2r\Xc£1. £x c £ 2 and JEJ C e c X, we have £2 C\ X = J ^ . q.e.d.
Since
By combining (1.4), (4.1) and (4.2) we thus have ttcJE^
c ^2 c c e c
\£3c£.
From now on we shall study almost Hermitian C-structures F / with respect to the metric g. LEMMA 4.2. - For an almost Hermitian hodls:
±-FJFuR,hjkl)
(4.15)
Rhi = -
(4.16)
Q = R + j-FhiFuRm
(4-17)
Rij =
FihFikRhk,
(4.18)
R =
FihFjkRjm,
where R is given by (2.3).
C-structure
=0,
Fj, each of the following
610 C.-C. PROOF.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
141
- Using (2.2), (4.1), (1.3) and noticing FhiFjk=-ghk,
(4.19)
we can easily obtain (4.17). Again from (1.3) it follows that, in consequence of (1.5) and (3.1), Rhijk = FirFksRshjr +
(4.20)
FfFfR^r.
Multiplying (4.20) by ghk and using (4.17) give (4.15) immediately. (4.16) is obtained by multiplying (4.15) by gh\ Substituting the identity (2.4) in (4.15) we obtain Rhi = ±Fi (FklRmj
(4.21)
+ FuRm)
= | * y (F*Rm
(4.18) is obtained by multiplying (4.21) by gh\
+ F^R^)
FiF&Rm.
=
q.e.d.
Lemma 4.2 is known for an almost Hermitian //-structure (see, for instance, [5]).
5. - Holomorphic sectional curvature. In this section we characterize an almost Hermitian C-structure by the holomorphic sectional curvatures. LEMMA 5.1. - A necessary and sufficient condition for a Riemannian in-manifold M with an almost complex structure Fj to have constant holomorphic sectional curvature K at each point with respect to the almost complex structure, where K may depend on the point of the manifold, is that the Riemann curvature tensor satisfy (5.1)
(Rrqsk + Rrksq)FpFi + (Brisk + ^rksi) Fpr FqS + + (Rrisq + R^F/FS
+ (Rrpsk + Rrksp)FirFqs
+ (RrpSi + RHsp)FkrFqs This lemma is due to C. C.
HSIUNG
and L.
+ ( f l w + R^)
FtrFks +
= - AK{gpigqk + giqgpk +
g^g*).
FRIEDLAND [5].
LEMMA 5.2. - For an almost C-structure i*V condition (5.1) becomes (5.2)
RriskF/Fg3 + Rr&FfFp* + R^FfF?
PROOF.
- Due to ec£^
= - K(gpigqk + giqgpk +
girpgik).
of (4.1), substituting (1.3) for Rhijk in the expression
611 142
C.-C. HsiUNG - B. XlONG: A new class of almost complex structures
FqhFjRHjk
and using (3.1) we obtain FqhFjRhijk
(5-3)
FivFkrRvqn.
-
Substituting (5.3) in (5.1) yields immediately (5.4)
{R-risk + Rrksi ) Fpr FqS + (Rrgsk + Rrksq ) Fi Fp" +
+ (Rrpsk + Rrksp)FqrFts = - 2K(gpigqk + g^g^ + gqpgnc)By the identity (2.4) we have FprFqsRrisk=
- FprFqsRrski
(5-6)
FirFpsRrqsk=
— FirFpsRrskq
+ FirFpsRr1csq,
(5-7)
FqrFisRrpsk=
— FqrFisRrskp
+F^F^Rricgp.
(5.5)
+FprFqsRrksi.
Similarly,
Adding together (5.5), (5.6), (5.7) and making use of (1.5), we obtain (5.8) FpFqRrisk
= FprFqsRrsik + FiFpRrsqk
+ FfFp'R^sk + FqF'Ry^
+ FprFqsRrksi + F* Fps Rrksq + Fqr Fts Rrksp = FprFq$Rrksi + Ff Fps R^ Substituting (5.8) in (5.4) gives (5.2) immediately. q.e.d. C. C. HSIUNG and L. almost L-structure F{*.
FRIEDLAND
+ FJF^R^
+
+ Fq Fts Rrksp.
[5] have shown that Lemma 5.2 is also true for an
THEOREM 5.1. - Let M be a Riemannian 2n-manifold with an almost C-structure Fj3 which is almost Hermitian with respect to a Riemannian metric g^ ofM. Then a necessary and sufficient condition for M to have constant holomorphic sectional curvature K with respect to F/ at each point, where K may depend on the point of the manifold, is that the Riemann curvature tensor of M is given by
(5.9)
Rijkh - FirFksRshjr
= - ^K(ghk9ij
+ ghjgik - F^F, -
FhjFik).
Further, the Ricci tensor and scalar curvature of such a manifold with constant holomorphic sectional curvature K are given respectively by (5-10)
R,=
^ K g
(5.11)
R = n(n + 1)K.
i h
As a consequences of (5.10), M is an Einstein manifold. Further, ifM is compact and K is positive, then the first Betti number of M is zero. PROOF.
- For the necessity of condition (5.9) we suppose that M has constant holo-
612
C.-C. HSIUNG - B. XIONG: A new class of almost complex structures
143
morphic sectional curvature K. Then by Lemma 5.2 we have (5.2). Multiplying (5.2) by FhpFjq and using (3.1) and (5.12)
FijF^g*,
we obtain (5.13)
Rhijk - RrqhkFjqFir
- RjpskFisFhp
= - K(FhiFjk
- F^F^
+
g^g^).
On the other hand, a use of (4.7) and (3.1) yields (5.14)
F^F^Rpjs/c
= - F^iF/RpjM
+ F/R^)
= Rijkh -
F^F/R^.
- FisFkpRPjhs
= - K{FhiFjk
- FhkFij +
g^ga)-
- FisFjpRpkhs
= - K{FuFkj
- FhjFik +
g^g^).
Substituting (5.14) in (5.13) we find (5.15)
Rhijk + Rhkji - F{rFjqR^M
Interchanging j and k in (5.15) gives (5.16)
Rhikj + Rhjid - FfF^R^j
For convenience we rewrite (4.20) as follows: Rhijk = FtsFkpRphjS
(5.17)
+ Fir Fjq Rhq/cr.
We use the identity (2.4) to obtain: - FirFjq R„,hk — FisFjpRpkhs
(5.18) (5.19)
F^F^Rpjte
+ F^F^R^j
(5-20)
Rhijk + Rhjki = Rijkh •
=
FfFfR^u,
=
FisFkpRphjS,
Substituting (5.18), (5.19) in the sum of (5.15), (5.16), (5.17) and making use of (5.20) we can easily arrive at (5.9). To prove the sufficiency of condition (5.9) we need LEMMA 5.3. - (5.9)
(5.21)
implies
Rhijk + FjrFksRrshi
= -^K(ghk9ij ~ QhjSik + FhkFij ~ FhjFik -
2FhiFjk).
PROOF. - The identity (2.4) readily gives Ff Fks Rnhi
=
—
FksFjrRrhis
+
F^FfRrifo.
Then we obtain (5.21) by substituting (5.9) for each term on the right-hand side of the above equation. q.e.d. Now we are in a position to prove the sufficiency of condition (5.9). F o r this purpose we suppose that (5.9) holds. Since the holomorphic sectional curvature K with respect to Fj and an arbitrary
613 144
C.-C.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
tangent vector u% of M at a point P is the sectional curvature of M at P with respect to the holomorphic plane determined by the tangent vector ul and its holomorphic tangent vector Fjluj, by using (2.7), (5.12) and FyUluj = 0 we obtain RriskFpru"uiFqsu''uk
gvigqku'puiu'iuk and therefore (5.22)
R*AFvrv?uiF*v.*v.t=
-
Kgpigqku'>ulu«uk.
Substituting (5.21) for R^ on the left-hand side of (5.22), and (5.9) for FirFksRskjr in the resulting equation, and making use of (3.1), (5.12), Rpikqupuluquk = 0, and gpqgikupuluquk = gqkgpiupulu9uk, we can see that the left-hand side of (5.22) becomes automatically the right-hand side of (5.22), so that (5.22) holds for any tangent vector M1 of M. Hence M has constant holomorphic sectional curvarue at each point. (5.10) is obtained by multiplying (5.9) by gjh and using (4.17), (5.12), and (5.11) is obtained by multiplying (5.10) by gij. Since for K > 0, by (5.10) the Ricci curvature of M is positive definite, the last part of the theorem follows from a well-known theorem of S. BOCHNER (see, e.g., [11, p. 37]). q.e.d. Concerning the holomorphic sectional curvature of an almost Hermitian C-manifold we have the following analogue of Schur's theorem. THEOREM 5.2. - / / an almost Hermitian C-manifold M of real dimension 2n ^ 4 has nonzero constant holomorphic sectional curvature K at each point, then K is an absolute constant on the manifold, and therefore the manifold is of constant holomorphic curvature. PROOF. - By Theorem 5.1 the C-manifold M is Einstein, and it is well known that using (5.10) we can easily show that K is an absolute constant for n ^ 2. Theorem 5.2 is well known for a Kahlerian manifold and an almost Hermitian L-manifold (for the latter case see, for instance, [5]). THEOREM 5.3. -IfMis an almost Hermitian C-manifold of real dimension In ^ > 4 with nonzero constant holomorphic sectional curvature K at each point and satisfies
(5-23)
Rhijk VtFih + RhWVjFih
+ RmVkF*
= 0,
where FJ is the almost Hermitian C-structure, Rhijk is the Riemann curvature tensor, and V is the covariant derivation, with respect to the almost Hermitian metric g{j of Fj, then M is almost Kahlerian.
614 C.-C. PROOF.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
145
- We shall prove our theorem by using the Bianchi identity: gijgMWiRhyk
(5.24)
+ VJRKM + v ^ - ) = o.
Substituting (4.6) in (5.21) gives (5.25)
Rhijk + F^FfR^
= —Kigali
~ dhjgtk + FhkFij ~ FhjFik ~
2FhiFjk).
Denote by (5.25)x (or (5.25)2) the equation obtained from (5.25) by changing the indices j , k to k, I (or I, j) respectively. Take the covariant derivative V; on the both sides of (5.25) and denote the resulting equation by V; (5.25). Consider the equation (5.26)
i^[V ; (5.25) + 7,(5.25)! + V,(5.25)2].
Using (4.19) we can easily obtain the left-hand side of F^V;(5.25): (5.27)
+ FhrFisR^k)
F^dR^
= 2FihVlRm
+
2RmVlFih.
Similarly, the left-hand sides of ^ 7 , ( 5 . 2 5 ) ! and FihVk(5.25)2 are respectively (5.28)
FihVj(Rhm
(5.29)
Fih Vk (Rkaj + FhrFisRrdj)
+ F^FfRrsu)
= 2FAVjRhikl
+
2RhmVjFih,
= 2Fih VkRm + 2Rm VkFih.
Thus from (5.27), (5.28), (5.29) and the Bianchi identity (2.5) it follows that the lefthand side of the equation (5.26) is equal to 2(RmVlFih
(5.30)
+ RhiklVjFik
+
RmVkFih),
which is zero by condition (5.23). Since K is an absolute constant by Theorem 5.2, the right-hand side of the equation (5.26) is (5.31)
±KFih [V,(FuFij - FyF* - 2FhiFjk) + + VjiFuF* - FhkFa - 2FhiFkl) + Vk(FhjFa - FuFq -
2FhiFlj)1.
From (3.1) we have FijFM = - **, The last of which gives Fih^iFih (5.31) to
FijFih = Sf,
V,(FihFih) = 0,
= 0. Using these equations we readily reduce
2(» + DKCJiFjt + VjFu + VkFij).
615 146
C.-C.
HSIUNG
- B. XIONG: A new class of almost complex structures
By comparing the both sides of the equation (5.26) we thus obtain VtFjk + VjFkl + VkFtj = 0, which shwos that M is almost Kahlerian. For an almost L-manifold, Theorem 5.3 becomes the following known result (see, for instance, [5, Theorem 5.5]). COROLLAEY 5.1. - If M is an almost Hermitian L-manifold with non-zero constant holomorphic sectional curvature at each point, then M is almost Kahlerian. PROOF.
- From (1.1) and (5.27) it follows that Rhijk^lF1
=0.
Similarly, using (5.28) and (5.29) instead of (5.27) we have RkiklVjFih = 0,
RhiljVkFih
= 0.
Thus condition (5.23) is automatically satisfied for an almost Hermitian L-manifold. Since by Theorem 4.1 an almost L-manifold is also an almost C-manifold, Corollary 5.1 follows immediately from Theorem 5.3. We need the following known relations [5, Theorem 6.1 and Corollary 6.2] among almost Kahlerian structures, almost semi-Kahlerian structures and Kahlerian structures. THEOREM
5.4. - If an almost semi-Kahlerian structure Fj satisfies Vk Fij V* F * = aVk Ftj VkFij,
(5.32)
where V is the covariant derivation with respect to the Hermitian metric g^ ofFj, and a is a nonzero constant, then Q defined by (5.33)
Q=R +
±FhiF*Rhijk
is nonnegative or nonpositive according as a is positive or negative. Further, the structure Fj is Kahlerian if and anly if Q = 0. PROOF.
- From (3.1) and (3.8) it follows that Vk(FlJFjk) = FjkVkFJ
= 0,
and therefore that (5.34)
V H F / V ^ / ) = ^FjkVkFJ
+ FjkT^kFJ
= 0.
Using (3.8) and the Ricci identity (2.6) we obtain (5.35)
YVkFJ
= ViVi**' - V ^ F * = F^R\ki
+ Fh^Rhk,
616 C.-C.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
147
which, together with (2.3), (2.4), (4.19) and (5.33), implies that (5.36)
= FikFihRm
Fj^V.FJ
- R.
On the other hand, we have, by (2.4), F*FihRjUd
(5.37)
= - F*Fih (Rm
+
Rm),
and also F*FARm
(5.38)
= F*FuRm
=
F*FihRm,
both of which together give F*FihRm
(5.39)
±F*FhiRm.
= -
Substituting (5.39) in (5.36) and using (5.33) we obtain (5.40)
FfWtFj
= - Q.
From (5.34), (5.32) and (5.40) it thus follows that FfVW
= - YFfVkFJ
= - aV^^F*
=-Q,
or Q = aVkFijVkFi>.
(5.41)
Since VkFijVkFij =s 0, if a > 0 then Q ^ 0, and if a < 0 then Q s= 0. Furthermore, Q = = 0 if and only if
i.e., if and only if VkFy = 0, or Fj is Kahlerian. COROLLARY
5.2. - / / an almost Hermitian structure FJ is almost Kahlerian,
then (5.42)
Q « 0,
where the equality holds if and only if the structure F / is Kahlerian. PROOF. jk
VkF^F
- From (3.7) it follows that = - VfcF#(V*F*' + V'f **) = = - V ^ V ^ * ' - VtFfiVF*
= - VuFtfV'F* -
V^VF^.
Since an almost Kahlerian structure is also an almost semi-Kahlerian structure, by (5.49) we have a = - 1 / 2 , and the last part of Corollary 5.2 follows readily from Theorem 5.4.
617 148
C.-C.
HSIUNG
- B.
XIONG:
A new class of almost complex structures
By Theorem 5.3, Corollary 5.2 and equation (4.16), Theorem 5.3 can be extended as follows. THEOREM 5.5. -IfMis an almost Hermitian C-manifold of real dimension 2n 2* & 4 with nonzero constant holomorphic sectional curvature at each point and satisfies condition (5.23), then M is Kahlerian.
Using Theorem 5.5 and a well-known theorem of S. Bochner for Kahlerian manifold (see, for instance, [11, p. 161]) we immediately obtain THEOREM 5.6. - The Betti numbers 6, of a compact almost Hermitian C-manifold of real dimension In > 4 with positive constant holomorphic sectional curvature at each point and satisfying condition (5.23) are given by
(5.43)
&2p = l ,
b2p + 1 = 0,
(0^2p<2p
+
1^2n).
6. - Complex structures. By Theorem 1.1 a complex structure on a Riemannian 2w-manifold without a constant sectional curvature, which is zero or nonzero, is an almost C-structure. In order to apply some of our above discussions to complex structures on such manifolds we need the following lemma. LEMMA 6.1. -On a Riemannian manifold M of dimension 2n with an almost complex structure F^ there is a Riemannian metric with respect to which F/ is almost Hermitian. PROOF.
(6.D
- Let h^ be any Riemannian metric on M. Then kv =
FikFilhu
is also a Riemannian metric on M, since by (3.1), d e t a i l = (2w)2redet||fey||, where det||fci;|| denotes the determinant of the square matrix ||ft^-||. Define a tensor g^ by (6.2)
2gij = hij + kij.
Using (6.2), (6.1) and (3.1) we can easily show that Fi Fj gu - gtj,
so that F/ is almost Hermitian with respect to the Riemannian metric g^ on M. q.e.d. Using Lemma 6.1, from Theorems 5.2, 5.5 and 5.6 we thus obtain the following theorem.
618
C.-C. HSIUNG - B. XIONG: A new class of almost complex structures
149
THEOREM 6.1. - Suppose that a Riemannian manifold M of dimension 2n 3= 4 without a constant sectional curvature, which is zero or nonzero, has a complex structure Fj and nonzero constant holomorphic sectional curvature K with respect to an almost Hermitian metric g^ of Fi at each point of M. (i) Then K is an absolute constant
on the manifold
M.
(ii) If the complex structure Fj satisfies condition (5.23), then M is Kahlerian. Further, if M is compact and K > 0, then the Betti numbers bi of M are given by (6.3)
b
2 p
=l,
b 2 p + 1 = 0,
(0^2p<2p
+
1^2n).
REFERENCES [1] E. CARTAN, Lecons sur la geometric des espaces de Riemann, 2nd ed., Paris (1946). [2] A. GRAY - L. M. HERVELLA, The sixteen classes of all Hermitian manifolds and their linear invariants, Ann. Mat. Pura Appl., 123 (1980), pp. 35-58. [3] C. C. HSIUNG, Nonexistence of a complex structure on the six-sphere, Bull. Inst. Math. Acad. Sinica, Taiwan, 14 (1986), pp. 231-247. [4] C. C. HSIUNG, Some conditions for a complex structure, to be published. [5] C. C. HSIUNG - L. FRIEDLAND, A certain class of almost Hermitian manifolds, Tensor, 48 (1989), pp. 252-263. [6] A. LICHNEROWICZ, Courbure, nombres de Betti et espaces symmetriques, Proc. Internat. Congress Math. (Cambridge, Mass., 1950), Amer. Math. Soc, Vol. II (1952), pp. 216-223. [7] G. B. RIZZA, Varieta parakdhleriane, Ann. Math. Pura Appl., 98 (1974), pp. 47-61. [8] S. SAWAKI - S. K. SEKIGAWA, Almost Hermitian manifolds with constant holomorphic sectional curvature, J. Diff. Geometry, 9 (1974), pp. 123-134. [9] Z. I. SZABO, Structure theorems on Riemannian spaces satisfying R(X, Y)-R = 0 - I, J. Diff. Geometry, 17 (1982), pp. 531-582. [10] L. VANHECKE, Almost Hermitian manifolds with J-invariant Riemann curvature, Rend. Sem. Matem. Torino, 34 (1975-76), pp. 487-498. [11] K. YANO - S. BOCHNER, Curvature and Betti numbers, Annals of Math. Studies, No. 32, Princeton University Press, Princeton (1953).
BULLETIN OF THE INSTITUTE OF MATHEMATICS ACADEMIA SINICA Volume 23, Number 3, September 1995
T H E SPECTRAL GEOMETRY OF ALMOST L MANIFOLDS BY CHUAN-CHIH HSIUNG (M^a)
AND CHUAN-XI WU ( J ^ H )
1. Introduction. Let (M,g) be an m-dimensional compact Riemannian manifold M with a Riemannian metric g. Throughout this paper all manifolds are supposed to be C°° and connected.
By A we denote the
Laplacian with respect to the metric g, acting on p-forms, 0 < p < m, on M. If a p-form a satisfies AQ
= Xa,
where A G R, the set of real numbers, then a is called a p-eigenform, and A the eigenvalue associated with a. Furthermore, the set of the eigenvalues associated with all the p-eigenforms is called the spectrum of A on p-forms on M, which will be denoted by Specp(M,g). Specp{M,g)
Thus
= {0 > A1>p > A2,p > . . . > - o o } ,
where each eigenvalue A;iP is repeated as many times as its multiplicity, which is finite and the spectrum Specp(M,g)
is discrete since A is an elliptic
operator. It is an interesting problem to investigate how the spectra reflect the geometry of M. J. Milnor [10] has shown that there may exist two nonisoReceived by the editors March 5, 1994. AMS 1980 Subject Classification: Primary 53C15, 53C55 58G25. T h e work of the second author was partially supported by the National Natural Science Foundation of the People's Republic of China and the C. C. Hsiung Fund at Lehigh University. 229
230
THE SPECTRAL GEOMETRY OF ALMOST L MANIFOLDS
[September
metric Riemannian structures on a compact manifold M such that for each p the spectra with respect to the Riemannian metrics are the same. Since then, there are many such examples; see, for instance, [6] and the references there. Thus the spectra do not determine the manifold M up to an isometry. However, the relationship between the geometry of a Riemannian manifold and its spectra has been extensively studied. One of the problems in this aspect is as follows: Let (M,g) and (M',g') p
manifolds with Spec (M,g)
p
= Spec (M',g')
be compact Riemannian
for a fixed p. Then is it true
that (M,g) is of constant sectional curvature if and only if (M',g')
is also
so? The answer to the problem is yes in the following cases: 1. p = 0 and m < 5, [1], [12]; 2. p = 1 and m = 2,3 or 16 < m < 93, [13]; 3. p = 2 and m = 2,3,6, 7,14 or 17 < m < 178, [14]. The above results are not true in general without restrictions on the dimension of M. V. K. Patodi [11] has given a counterexample in the case p — 0 and m > 5, but also proved the following result: Let (M,g) and p
p
be compact Riemannian manifolds with Spec (M,g)
— Spec (M',g')
(M',g') for
p = 0 and 1. Then (M,g) is of constant sectional curvature if and only if (M',g')
is also so.
Let (M,g,J)
be a compact Kahler manifold with complex structure J
and Kahler metric g, and let the real dimension of M be denoted by m = 2n. A natural problem is the following : Let (M,g,J) p
compact Kahler manifolds with Spec (M,g)
and (M',g',J')
p
= Spec (M',g')
be
for a fixed p.
Then is it true that (M, g, J) is of constant holomorphic sectional curvature if and only if (M',g', J') is also so? The answer to the problem is yes in the following cases: 1. p = 0 and m < 10, [12]; 2. p = 1 and 16 < m < 102, [13]; 3. p = 2 and m = 2,6,8,14 or 18 < m < 188, [14]. For p = 2, B. Y. Chen [4] and S. I. Goldberg [5] proved that if (M,g, J) is a compact Kahler manifold with Spec2(M,g)
= Spec2(CPn,g0)
where CPn is
1995]
CHUAN-CHIH HSIUNG AND CHUAN-XI WU
231
the complex projective space of complex dimension n, and go is the FubiniStudy metric, then (M,g,J)
is holomorphically isometric to
(CPn,go,Jo),
where Jo is the standard complex structure of CP r a . On the other hand, there are several well-known classes of almost' Hermitian manifolds. One of them is so called the class of almost L manifolds by L. Friedland and C. C. Hsiung [5], which were studied first by E. Cartan [3], and then by A. Lichnerowicz [8] and many other authors. An almost Hermitian manifold M satisfying [ V ^ V f c ] / ^ (V,V fc - V f c V , ) i f = 0 is called an almost L manifold (see [5]), where J = (F/ 1 ) is the almost complex structure of M, and V denotes the covariant differentiation with respect to the affine connection of the Hermitian metric. It is obvious that a Kahler manifold is an almost L manifold. An example of a non-Kahler almost L manifold has been given by H. Yanamoto [15] by constructing a certain hypersurface of a 7-dimensional Euclidean space. The purpose of this paper is to study the spectral geometry of almost L manifolds. The main results of this paper are listed in the following theorem. Theorem. Let (M,g,J) p
folds with Spec (M,g) dimension
of M. n
tive space CP
and (M',g',J') p
— Spec (M',g') n
Let (
where p = 0,1,2, and m be the real be the complex n-dimensional
projec-
with the Fubini-Study metric go and the standard complex
structure JQ. Consider the following (*) (M,g,J)
be compact almost L mani-
statements:
is of constant holomorphic sectional curvature H if and only
if (M',g',J')
is of constant holomorphic sectional curvature H', and
H = H'; (**) (M,g,J)
is Kahlerian and holomorphically isometric to
Then we have the following: (i) (*) is true for p = 0, and m < 10. (ii) (**) is true if Spec°(M,g)
= Spec°{CPn,g0)
and m < 10.
(iii) (*) is true for p — 1 and m = 2 or 16 < m < 102.
(CPn,g0,Jo)-
232
THE SPECTRAL GEOMETRY OF ALMOST L MANIFOLDS
(iv) (**) is true if Specl(M,g)
= Spec1 (CPn,g0)
[September
and TO = 2 or 16 < m <
102. (v) (*) is true for p = 2 and m = 2,6,8,14 or 18 < m < 188. (vi) (**) is true if Spec2(M,g) 18 <
TO
= Spec2 (CPn, g0) and m = 2,6,8,14 or
< 188.
(vii) (*) is true for p = 0 and 1. (viii) (**) is true if Specp(M,g)
= Specp(CPn,g0)
Remarks. 1. When (M,g,J)
for p = 0 and 1.
and (M',g',J')
are Kahler manifolds,
parts (i)-(vi) of the above theorem are reduced to the known results in ([12], [13], [14]) mentioned before. 2. In parts (vii) and (viii) of the above theorem, there are no restrictions on the dimension of M, and the results are new even if the manifolds are Kahlerian. Hence we have the following. Corollary 1. (*) is true for compact Kahler manifolds (M,g,J) p
(M',g',J')
p
with Spec (M,g)
= Spec (M',g')
corollary 2. (**) is true if (M,g,J) p
and Spec (M,g)
p
n
= Spec (CP ,g0)
and
for p = 0 and 1. is a compact Kahler
manifold,
for p = 0 and 1.
2. P r e l i m i n a r i e s . Let (M,g) be a Riemannian manifold of dimension TO > 2 with positive definite symmetric Riemannian metric g = (gij), and let (
an
d S we denote the Riemann
curvature tensor, the Ricci curvature tensor and the scalar curvature g13 Rij of M, respectively. The Riemann curvature tensor Rhijk satisfies the identity (2.1)
Rhijk + Rhjki + Rhkij = 0,
where Rhijk = guR ijfe, and the Einstein tensor denoted by G — (Gij) is given by
1995]
CHUAN-CHIH HSIUNG AND CHUAN-XI WU
233
5 (2.2)
Gij = Rij
gij.
The following Ricci identity is known: V ^ T f e " - VjVlTkh
(2.3)
= TksRhsji
-
T,hR'kji,
where Tk is an arbitrary tensor of type (1.1). We denote the square of the norm of a tensor D by \D\2. Then, from (2.2) it follows that \G\2 = IRA2 - — and that IGI2 = 0 holds on M if and only if (M, g) is an Einstein space. The Minakshisundaram-Pleijel-Gaffney's formula for Specp(M,g)
is
given by oo
oo
£ V * - ' * tj.0 ~ (47rt)-m/2^aiiPf,
(2.4)
fc=0
where (2.5)
«o, P —
(2.6)
ffll.p —
DL^ 1 /m\
J
/ m - 2\
3\p) ~ U - i J .
SdM,
JM 2
«2,p = / [ci(m,p)5 + c 2 ( m , p ) | ^ i | 2 + c 3 (rn,p)|i?| 2 ]dM,
(2.7) and ,„„. (2 8)
m
P)
'
^ >
.„ „, (2 9) ' ,„ „x
,
(2.10)
C2(m P)
'
1 fm\
1 / m - 2\
= T2{P)-Ap-l) 1 /m\ = - M ( p ) 1 /m\
+
+
1 /m - 4
2\p-2
1 /TO - 2 \ /TO - 4 2 ( P - l ) - 2 ( P - 2 1 /TO - 2 \ 1 /TO - 4 \
«3(m)P) = - y - - ( p _ J + - ( p _ 2 ) ,
dM being the volume element of M, and ( m ) a binomial coefficient.
The
coefficients aDiP, a,i p and a,2>p have been calculated for p — 0 by many authors (see [1], [9]), and determined for all p by V. K. Patodi [11]. R e m a r k s . 1. Let (M,g) and (M',g') p
folds. If Spec {M,g) we have
p
= Spec (M',g')
be compact Riemannian mani-
for some p, then from (2.4) and (2.5)
234
THE SPECTRAL GEOMETRY OF ALMOST L MANIFOLDS
[September
(1) m = dim M = dim M' = m'\ (2) V o l M = V o l M ' . 2. For a geometric quantity A on (M,g), we will denote the corresponding quatity on (M',g')
by A'.
3. Almost L manifolds. Now let (M,g,J)
be an almost L manifold
with a Hermitian metric g^ and an almost Hermitian structure J whose components are F{ J. Then we have F^F3k
(3-D
= -6l
guF^
= ghk
and (3.2)
[V,-, Vk}F,
h
= (V.-Vfc - VfcV,-)^ h = 0.
L. Priedland and C. C. Hsiung have given a thorough study on almost L manifolds in [5], where we can find the following Lemmas 1, 3, 4. L e m m a 1. An almost Hermitian manifold is an almost L manifold if and only if Fi SR
(3-3)
sjk
= Fs
Rsijk,
which is equaivalent to yo.Q)
*LijkiTr
rs
iLijrS)
implying RijFrlFsj
(3.5)
= Rrs.
L e m m a 2. On an almost L manifold each of the following holds: yo.Dj
ttijksrr
ii,ijrkrs
(3.7)
RirFjT
(3.8)
RijklFkl
=
2FjTRir,
(3.9)
2RijklFjl
=
RikjlFjl,
where Fi} = gjkF{h,
i™ = F f c J g i k .
— —RjrFir,
,
identities
1995]
CHUAN-CHIH HSIUNG AND CHUAN-XI WU
235
Proof. (3.6) and (3.7) follow readily from (3.4), (3.5) and (3.1). Multiplying (3.4) by glsFpr
and using (3.1) we obtain
RjrFpr
(3.10)
~RijPiFil.
=
Prom identity (2.1) it follows that jr
p
'
V-^ipO * ^iljp) "•iljpj
—
fi-lpij
i r
—•*
fi'ijpl T r
ti'iljp rijpii,
which together with (3.10) implies (3.8). Similarly, by identity (2.1) we have
2RzjklF>1
= F>lRijkl
+ FljRilkj
= Fjl(Rijkl
+
Riljk)
which is (3.9). L e m m a 3. A necessary and sufficient condition for an almost L manifold M of real dimension m to have constant holomorphic sectional curvature H at each point is that the Riemann curvature tensor be given by TT
(3.11)
Rhijk
= —[(ghkgij - ghjgik) + (FhkFij - FhjFik)
-
2FhiFjk}.
Further, the Ricci tensor and scalar curvature of such a manifold with constant holomorphic sectional curvature H are given, respectively, by (3.12)
Rl3 =
(3.13)
S =
^±lH9ij,
m ( m +
2)
g.
Lemma 4. / / M is an almost L manifold of real dimension > 4 and nonzero constant holomorphic sectional curvature, then M is Kdhlerian. For an almost Hermitian manifold of real dimension m, we define its Bochner curvature tensor B = (Bijki) as follows:
236
THE SPECTRAL GEOMETRY OF ALMOST L MANIFOLDS Bijki = Rijki
~~i{Rjkgn
(3.14)
- Rjigik + gjkRu
+ FjkFi
r
Rri
- F^F,
— FikFjrRri
[September
— gjiRik
r
Rrk
+ FuFf
— 2FirRrjFki
Rrk 2FijFkrRrl)
-
g ( m + 2 ) ( m + 4)
{gjkgu - gjigik + FjkFu
L e m m a 5. Let M be an almost L manifold
~ Fj[Fik
-
2FijFki).
of real dimension
m.
Then
\B\2 = \R\2--^—\Rl\2 + m +4 ' (m + 2 ) ( m + 4)
(3.15)
Proof. We can prove this l e m m a by using (3.1) and L e m m a s 1 a n d 2. A l t h o u g h t h e c o m p u t a t i o n is elementary, it is long, so we sketch it here. A t first, we p u t •A-ijkl =Rjk9il
— Rjigik + gjkRil
+ FjkFtTRri 3 16
( -
— gjiRik + FuF-rRrk
— FjiF^Rrk
-
FikF.TRrl
) - 2Fi TRrjFu Bijki =gjk9u
2FijFkrRri,
-
- g3igik + FjkFu
— FjiFik
-
2FijFki.
T h e n (3.14) becomes (3.17)
1 S —r-Aijki + T —^77 —rzBijkim + 4 (m + 2 ) ( m + 4)
B^ki — Rijki
For simplifying the c o m p u t a t i o n we choose an o r t h o n o r m a l local coord i n a t e s y s t e m so t h a t (3.18)
gij = 8ij.
where % are Kronecker deltas. Using ( 3 . 4 ) , . . . , (3.9), we obtain (3.19)
RijkiAi>kl
=
(3.20)
RijkiBljkl
= 85,
ijkl
16\R1\2,
(3.21)
AijkiA
= 16[5 2 + (m + 4 ) | i ? 1 | 2 ] ,
(3.22)
A3kiBljkl
= 16(m + 2 ) 5 ,
(3.23)
BijklBijkl
= 8 m ( m + 2),
1995]
CHUAN-CHIH HSIUNG AND CHUAN-XI WU
237
Substituting (3.19),..., (3.23) in \B\2 expressed by (3.17) thus gives (3.15). L e m m a 6. An almost L manifold with real dimension m > 4 is of constant holomorphic sectional curvature if and only if B = 0 and G = 0. Proof. Suppose that an almost L manifold M of real dimension m is of constant holomorphic sectional curvature. Then (3.11), (3.12) and (3.13) hold by Lemma 3. Substituting (3.11), (3.12) and (3.13) in (3.14) and using (3.1), we obtain B = 0. G = 0 follows from (2.2), (3.12) and (3.13). Conversely, suppose that B = 0 and G = 0. From G = 0 it follows that M is an Einstein space, so that S is constant for TO > 4. Substituting R-ij = m9{j
m
(3-14) and using (3.1) and B — 0, we obtain
g Rijki = —7—~^:[9jk9ii — 9ji9ik + FjkFu — FikFji Tit I lit
l - Li I
2FijFki}.
Hence by Lemma 3, M has constant holomorphic sectional curvature H = 45 m(m+2) '
4. Proof of the Theorem, (i) From (2.5), • • •, (2.10), we have (4.1)
a 0 , 0 = [ dM = Vol M, JM
(4.2)
a M = \ I SdM, 6 JM
(4.3)
a 2 , 0 = - ^ f [2\R\2 - 2 | ^ | 2 + 5 S2]dM. The case m = 2. Since |i?| 2 = S2 and g is always an Einstein metric,
we have |-Ri|2 = S 2 / 2 and therefore a2,o = ^ JM S2dM. of (M,g,J)
and (M',g',J')
Since the roles
in the theorem are the same, we need only to
prove the "if part, and the "only if " part can be proved in the same way just by interchanging the roles of (M,g,J) that (M',g',J')
and (M',g',J').
So we assume
has constant holomorphic sectional curvature H'.
5 ' = 2H'. From ao,o = a'00, a^o = a\ 0 and c^.o = a' 20 , it follows that SdM = 2H'Vol M,
/ JA
/
S2dM = 4iJ' 2 Vol M, M
Then
238
THE SPECTRAL GEOMETRY OF ALMOST L MANIFOLDS
[September
• l)dM)2
so that the equality holds in the Schwarz inequality (fM(S 2
(JM S dM)
<
• (JM UM). Thus 5 is constant, and S = S' = 2H'.
The case 4 < m < 10. Substituting (3.15) and (2.2) for \R\2 and \RX\2 respectively in (4.3), we obtain (4.4)
a2.0
360 JM
2\B\2 +
2(12
-m)
G\2 +
m +4
5m 2 + 8m + 12 m(m + 2)
dM.
Thus a2,o — 0-2 o c a n t>e written as 2\B\Z +
/ (4.5)
JM
f JM'
2(12
m
),
G
|2
5 +
8m + 12_ 2 ' dM 7n(rra + 2)
™2
m +4
5m 2 + 8 m + 12 m(m + 2)
2\B'\2+2+-^^\G>\2 m +4
?'2
dM'.
As in the former case we need only to assume that (M',g',J') 2
constant holomorphic sectional curvature H'. Then \B'\ Lemma 6. Since S' is constant, ao,o = fto,o S'2dM'.
an(
is of
2
= \G'\ — 0 by
a
i i,o = a'i o imply / M S2dM >
In fact, using the Schwarz inequality we have / dM J ( / S2dM) > ( f SdM) = JM J \JM J \JM J = 5' 2 (Vol M ' ) 2
S'dM' M'
= 5 /2 Vol M • Vol M' dM'.
= Vol M JM'
Thus (4.5) gives B = 0, G = 0 and / M 5 2 d M = / M , S'2dM', and hence by Lemma 6, (M,g, J) is of constant holomorphic sectional curvature H = H'. (ii) Since (CPn,go,Jo) c > 0, (M,g,J)
has constant holomorphic sectional curvature
has constant holomorphic sectional curvature c > 0 by
Theorem 1. From Lemma 4 it follows that (M,g,J) (M,g,J)
n
is holomorphically isometric to ( C P , go, Jo)•
(iii) From (2.5), • • •, (2.10), we have (4.6)
is Kahlerian.
Thus
1995]
CHUAN-CHIH HSIUNG AND CHUAN-XI WU
(4.7)
aM = ^-=^ /
6
239
S dM,
JM
(4.8) o 2 ! = — / [2(m - 15)|#| 2 + 2(90 - m ) ^ 360 J M
2
+ 5(m - 12)5 2 ]dM.
The Proof for the case m = 2 is similar to that for the corresponding case in (i), and therefore is omitted here. For a general m, as in the proof of (i), by (3.15) and (4.8) we obtain 02,1 = ^ 360
M CA 1
j
- 15)|5| 2 + —^— [m(102 - m) + 120]|G| 2 m +4 1 1 + — — [m 2 (5m - 52) + 72m + 120152 \dM.
/ U(m iM I
m{m + 2)
J
For 16 < m < 102, all coefficients of \B\2, \G\2 and S 2 are positive. Also as before we may assume that (M',g', J') is of constant holomorphic sectional curvature H', so that \B'\2 = \G'\ = 0. As in the proof of Theorem 1, we have JM S2dM > JM, S'2dM'. Furthermore, a 2 ,i = a'2
1
can be written as
I J 2(m - 15)|B|2 + —^— [m(102 - m) + 120] |G|2 + (4.10)
. 1 „. fm 2 (5m - 52) + 72m + 1201J S2 \dM m(m + 2) L J
= / ( 2 ( m - 1 5 ) | B ' | 2 + — — L[ m ( 1 0 2 - m ) + 120l|G'| 2 JM' I m + 4 + —. — [m 2 (5m - 52) + 72m + 1201S'2 \dM', m(m + 2) L v ' J which implies that \B\2 = |G| 2 = 0. Hence (M,g, J) is of constant holomorphic sectional curvature H = H'. (iv) follows from (iii) and Lemma 4. (v) From §2, for p = 2, the coefficients are given by ,.,., (4.11)
m(m-l) a0,2 = *
/, ,o\ (4.12)
ai, 2 =
f m(m-l)^T. ,, - / dM = —^— -Vol M. JM 2
m 2 - 1 3 m + 24 f — / iZ
JM
OJ!1J
SdM.
240
THE SPECTRAL GEOMETRY OF ALMOST L MANIFOLDS
[September
(4.13)
«2,2 = =^r / [5(m 2 - 25m + 120)S 2 - 2(m 2 - 181m + 1080)|i?!\' 7^0 JM + 2(m 2 - 31m + 240)|i?| 2 ] dM.
The proof for the case m = 2 is similar to that in (i) or (iii). For a general m, as in the proofs of (i) and (iii), by using (3.15) the coefficient a2v2 can be written as
(4.14)
a2>2 = ^
J (Ax\B\2 + ,42|G|2 + A3S2)dM,
where Ax = 2(m 2 - 3 1 m + 240), -(m 3 - 193m2 + 852m + 480), m + 4v ' 4 3 2 A3 = — - ( 5 m - 117m + 724m - 732m - 480). m{m + 2) A2 =
Thus for m = 6,8 or 14, or 18 < m < 180, we have Ax > 0, A2 > 0, A3 > 0. The remaining part of the proof is completely similar to that in (i) or (iii). (vi) follows from (v) and Lemma 4. (vii) This for the case m = 2 follows readily from (i) or (iii). So we may assum m > 4. Multiplying (4.5) by (m — 15) and subtracting the resulting equation from (4.10), we obtain f [150|C| 2 + — (m + (4.15)
= / JM'
Assume that (M',g',J') 2
Then \B'\
10)S2]dm
m
JM JM
[150|G'| 2 + — ( m + 10)5' 2 ]rfM'. m
is of constant holomorphic sectional curvature W.
2
= \G'\ = 0. As before we have JM S2dM
> JM, S'2dM' which
together with (4.15) implies that |G| 2 = 0. Thus from (4.5) it follows \B\2 = 0. Hence (M,g, J) is of constant holomorphic sectional curvature H = H'. (viii) follows from (vii) and Lemma 4.
1995]
CHUAN-CHIH HSIUNG AND CHUAN-XI WU
241
References 1 M. Berger, Le spectre des varietes riemanniennes, Rev. Roumaine Math. Pures Appl., 1 3 (1969), 915-931. 2. M Berger, P. Gauduchon and E. Mazet, Le spectre d'une variete riemannienne, Lecture Notes in Math., Vol. 194, Springer, Berlin, 1971. 3. E. Cartan, Lecons sur La geometrie des espaces de Riemann, 2nd ed., Paris, 1946. 4. B. Y. Chen and L. Vanhecke, The spectrum of the Laplacian of Kdhler manifolds, Proc. Amer. Math. S o c , 79 (1980), 82-86. 5. L. Friedland and C. C. Hsiung, A certain class of almost Hermitian manifolds, Tensor, 4 8 (1989), 252-263. 6. S. I. Goldberg, A characterization of complex projective space, C. R. Math. Rep. Acad. Sci. Canada, 6 (1984), 193-198. 7. C. Gordon, Isospectral closed Riemannian manifolds which are not locally isometric, J. Differential Geometry, 3 7 (1993), 639-649. 8. A. Lichnerowicz, Courbure, nombres de Betti et espaces symmetriqv.es, Proc. Internat. Congress Math. (Cambridge, Mass. 1950), Amer. Math. S o c , Vol. I I , (1952), 216-223. 9. H. F. McKean and I. M. Singer, Curvature and the eigenvalues of the Laplacion, J. Differential Geometry, 1 (1967), 43-69. 10. J. Milnor, Eigenvalues of the Laplace operator on certain manifolds, Proc. Nat. Acad. Sci. U. S. A., 5 1 (1964), 542. 11. V. K. Patodi, Curvature and the fundomental solution of the heat operator, J. Indian Math. S o c , 3 4 (1970), 269-285. 12. S. Tanno, Eigenvalues of the Laplacian of Riemannian manifolds, Tohoku Math. J., 2 5 (1973), 391-403. 13. S. Tanno, The spectrum of the Laplacian for 1-forms, Proc. Amer. Math. S o c , 4 5 (1974), 125-129. 14. GR. Tsagas and C. Kockinos, The geometry and the Laplace operator on the exterior 2-forms on a compact Riemannian manifold, Proc. Amer. Math. S o c , 7 3 (1979), 109-116. 15. H. Yanamoto, On orientable hypersurface of R7 satisfying R(X,Y) • F = 0, Sci. Rep. Nagaoka Tech. College, 8 (1972), 9-14.
Department of Mathematics, Lehigh University, 14E. Packer Avenue, Bethlehem, PA. 18015-3174, U.S.A. Hubei University, Wuhan, CHINA
SUT Journal of Mathematics Vol. 31, No. 2 (1995), 133-154
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES OF ALMOST HERMITIAN MANIFOLDS Chuan-Chih HSIUNG, Wenmao Y A N G and Lew F R I E D L A N D (Received July 10, 1995) A b s t r a c t . Friedland and Hsiung [1] proved an analogue of F. Schur's theorem concerning the holomorphic sectional curvature of some almost Hermitian manifolds called almost Hermitian L-manifolds, of which Kahlerian manifolds are special ones. Recently, Hsiung and Xiong [3] gave a classification of almost Hermitian manifolds and extended the above work of Friedland and Hsiung to a new class of almost Hermitian manifolds called the class of almost C Hermitian manifolds. In this paper we shall further extend the above work of Hsiung and Xiong by studying the general sectional, the holomorphic sectional and the holomorphic bisectional curvatures of almost Hermitian manifolds of all classes, together with some relationship among the three types of sectional curvatures.
§1. Introduction Let M be a Riemannian 2n-manifold, and gij, J / and Rhijk the components of a Riemannian metric tensor, and an almost complex structure J, and the curvature tensor, of M respectively. Throughout this paper, all Latin indices take the values 1, ...,2n unless stated otherwise. By using the following identities. Hsiung and Xiong [3] have defined the following four classes of almost complex structures on the Riemannian manifold M: (1-1)
(1.2)
(l.oj
Rhijk =
Rhijk
=
Jh "i Rrsjk
J^J^Rrajk,
+ Jh "j Rriak + Jh ^k
-Hhijk — ^h
i
j
k
Rrijst
^pqrsj
T h e work of the second author was partially supported by the National Natural Science Foundation of the People's Republic of China and the C.C. Hsiung Fund at Lehigh University. 133
633 134
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
(1.4)
JJJi'Rrsisk + J^J^Rrshk + J^JifRr^k
= 0,
where the repeated indices imply summation. Let £ and K denote the classes of almost complex structures (or manifolds) and the Kahlerican structures (or manifolds), respectively. Let C\, £2, £3 and C denote the classes of almost complex structures (or manifolds) satisfying (1.1),...,(1.4) respectively. Hsiung and Xiong [3] have showed the following inclusion relation: (1.5)
Kc£iC
£
2 c
c£3c£.
Thus for 1 = 1,2,3 as i decreases, the structures (or manifolds) in £j resemble Kahlerian structures (or manifolds) more closely. If J\ and gij satisfy (1-6)
9ijJhJk
=9hk,
then the almost complex structure J and the manifold M are called an almost Hermitian structure and an almost Hermitian manifold, respectively, and g^ is called an almost Hermitian metric. For simplicity, throughout this paper, unless stated otherwise, by an almost Hermitian manifold M we shall always mean a manifold with an almost Hermitian structure J and an almost Hermitian metric g^. Friedland and Hsiung [1] called an almost Hermitian structure J (or manifold M) an almost L structure (or manifold) if it satisfies (1-7)
[V;, V f c ]J> = (VjVfc - VkVjVS
= 0,
where V denotes the Levi-Civita connection of g^. Obviously, Kahlerian manifolds are almost L manifolds since M is Kahlerian if and only if (1.8)
ViJ/=0
for all
i,j,k.
Friedland and Hsiung [1] have obtained a necessary and sufficient condition for an almost L manifold to have constant holomorphic sectional curvature H at each point and showed that H is an absolute constant for such a manifold. Hsiung and Xiong [3] have proved that an almost L manifold is an almost Hermitian £1 manifold and extended the above result of Friedland and Hsiung to an almost Hermitian C manifold. The purpose of this paper is to extend further the above results of Hsiung and Xiong to an almost Hermitian manifold of each class with respect to a general sectional curvature or holomorphic sectional curvature, or holomorphic bisectional curvature, and to discuss the relationship among the three types of sectional curvatures for each of these manifolds.
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
135
In §2 (resp. §3) we recall some fundamental notation, definitions and wellknown results on Riemannian structures (resp. almost complex structures) which are needed for the later discussions. In §§ 4,5 and 6, we give a necessary and sufficient condition for an almost Hermitian manifold of each class to be of constant general sectional curvature, or constant holomorphic sectional curvature or constant holomorphic bisectional curvature at each point of the Riemannian manifold, respectively. Some relationship among the three types of sectional curvatures for an almost Hermitian manifold of each class are derived in §7. For simplicity we shall denote an almost Hermitian £* manifold by AHi for i = 1, 2, 3, and a Kahlerian manifold, an almost Hermitian C manifold and an almost Hermitian manifold respectively by K,AHC and AH. From (1.5) we thus obtain the following inclusion relations AH
(1.9)
K c AHi C 2HC
C AHZ
C
AH
-
Now we introduce the new notion of AH[ manifold which denotes an almost Hermitian manifold satisfying (1.10)
Rhijk =
-JhJ^Rrsjk-
It should be noted that the difference between (1.1) and (1-10) is only a sign, and therefore that AH[ c AHC C AH3, and the intersection of the two classes AH 1 and AH[ is the class of locally Euclidean spaces, that is, the class of spaces with Rhijk — 0 §2. Riemannian structures Let M be a Riemannian manifold of dimension m > 2 with Riemannian metric tensor g^, and let (g%:>) be the inverse matrix of (gij). We shall follow the usual tensor convention that indices can be raised and lowered by using g%i and g^ respectively. Let Rhijk, Rij,R denote the Riemannian curvature tensor, the Ricci curvature tensor and the scalar curvature of M, respectively. The following identities are known, the last two of which are called the Bianchi identity and the Ricci identity respectively: (2.1)
Rhijk + Rhjki + Rhkij = 0,
(2.2)
VeRhijk + VjRhM + VkRhitj = 0,
(2.3)
ViVj-Tfc* - VjViTkh
= Tk°Rhsji
-
TahRakji,
635 136
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
where V denotes the Levi-Civita connection of M, and Tkh is an arbitrary tensor of type (1,1). The sectional curvature with respect to the two-dimensional plane (u, v) determined by two linearly independent tangent vectors u and v of M at a point p is given by RhijkUhViUjVk h• l , k.—r-r
TS w \ K = K(U, V) = -.
{9hk9ij-9hj9ik)u v u^v
(2 4) where (2.5) (2.6)
R{u,v,u,v) [g(u,v)}2 g(u,u)g(v,v)' R(u, v, u, v) = g(u,u) = giju%u\
RhijkUkvlujvk,
g(u,v) = giju'v3,
g{v,v) = Sy-uV
Note that K(u, v) is the Gaussian curvature of the two-dimensional geodesic submanifold of M tangent to the plane (u, v) at P . If the sectional curvature at any point of the Riemannian manifold does not depend on the two-dimensional plane at the point, then (2.7) Rhijk = K(ghkgij - ghj9ik)The Riemannian manifold is said to be locally Euclidean or locally flat if K — 0, i.e., if Rhijk — 0. For nonzero function K on M, from (2.2) and (2.7) it is easy to show that (2.8)
^
(2.9)
=
(771-1)^,-,
R = m(m-l)K,
and for m > 3, K and therefore R are absolute constants on the manifold M and M is said to be of constant curvature. Furthermore, M is a n Einstein manifold as a consequence of (2.8). Now we want to define an angle between 2-planes through a point p in the tangent space TP(M) of the Riemannian m-manifold M at p. Let II = (a, b) and II' = (c, d) be two 2-planes determined respectively by orthonormal tangent vectors a, b and c, d at the point p. Then the determinant
(2.10)
(11,11')
g(a,c) g(b, c)
g(a,d) g(b, d)
is called the inner product of II and II'. When II and II' coincide, since o and b are not parallel, we have (2.11)
(n, n ) = g(a, a)g(b, b) - [g(a, b)}2 > 0.
So we can define the angle (II, II') between II and II' such that
(2.12)
cos(n, n') -
(n'n)
0
< / n n'> < TT/2.
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
137
§3. Almost complex structures In this section M is a Riemannian manifold as in § 2 but with dimension m — 2n. If a tensor J J of type (1.1) on M satisfies
(3.1)
V J / = -6l
where 6* are the Kronecker deltas defined by
then J •?' is called an almost complex structure on M, and M is called an almost complex manifold. If J is Hermitian, then as a consequence of (3.1) and (1.6) the tensor Jj,- of type (0,2) defined by (3-2)
Jij=9jkJik
is skew-symmetric. If the differential form Jtj is closed, then J^ is called an almost Kahlerian structure, and M an almost Kahlerian manifold. It is clear that an almost Kahlerian structure satisfies (3.3)
Jhij ~ V/j Jij + ViJjh + VjJhi = 0.
The tensor Jhij is skew-symmetric in all indices. An almost Hermitian structure J^ satisfying
(3.4)
Ji»-Vj.V=0
is called an almost semi-Kahlerian structure. An almost Kahlerian structure is almost semi-Kahlerian. An almost Hermitian structure J J satisfying (3.5)
V i J J -* + V,-Jifc = 0
is called a nearly Kahlerian structure. Since J^ = g^Jij = 0, from (3.4) and (3.5) it follows that a nearly Kahlerian manifold is almost semi-Kahlerian. Let M be an almost Hermitian manifold with an almost complex structure JJ. Then the two-dimensional plane (u, Ju) determined by an arbitrary tangent vector u of M and the tangent vector Ju at a point p is called a holomorphic plane, and the sectional curvature with respect to the holomorphic plane at p is called the holomorphic sectional curvature at p. If the holomorphic sectional curvature at p is independent of the holomorphic plane at p,
637 138
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
then M is said to be of constant holomorphic sectional curvature at p. We easily obtain H(u) = K(u, Ju) = (3.6)
R
hpjq Jj PJkuH
^
UJ
2
^
[a(u,u)} R(u, Ju, u, Ju) \g(u,u)}2
Let v be another tangent vector of M at p. Then the holomorphic bisectional curvature B(u,v) of M at p with respect to the vectors u and v is denned as follows (see Goldberg and Kobayashi [2]): (3.7)
B = B(u,v) =
- R ^ J ^ J v l g{u,u)g(v,v)
It is clear that B(u,u) = H(u). So the holomorphic bisectional curvature is a generalization of the holomorphic sectional curvature. Now let u and v be two unit tangent vectors of M at p, and let
gijulvj,
cosO — g(Ju,v) = JijUlv3, (3.8)
cosO' = g{Jv,u) =
-cos6.
Furthermore, for two holomorphic planes II = (u,Ju) have, in consequence of (2.10),
and II' = (v,Jv),
we
(n,n) = (n'n') = i, (3.9)
(n,II')=cos 2 (£ + cos 2 0,
which together with (2.12) imply (3.10)
cos(n, IT') = cos2 <> / + cos2 6.
Thus (3.11)
0 < cos2 (/> + cos2 0 < 1.
In particular, when II and II' are orthogonal,
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
139
For t h e later developments, using (3.1), we can easily show t h a t t h e following identities (3,12),(3,13),...,(3,16) are equivalent respectively t o identities (2.1), (1.1) (1.4): (3-12)
Jj Jk (Rhipq + Rhpqi + Rhqip) = 0,
(3.13)
J/Rhipk JhPRpijk
(3.14) (3.15)
+ JiPRhpjk
+ JkpRhijP
= 0,
+ JjPRhipk
+ J/fRhijp
= 0,
Jh Ji Rpqjk — Jj Jk Rhipq — 0, JhpRkpij
(3.16)
+ ^ pRkpjh
+ J/RkPhi
= 0.
§4. G e n e r a l s e c t i o n a l c u r v a t u r e s In this section we shall discuss general sectional curvatures of almost Hermitian manifolds. At first we have T h e o r e m 4 . 1 . If an almost Hermitian 2n-manifold M2n is of constant sectional curvature K at each point, then M2n is an AH3 manifold.
general
Proof. T h e identity (2.7) yields "h
i
j
k Rpqrs
=
Jh Ji Jj Jk ^{9ps9qr
— **• \JhsJirJj
= K{9hk9ij
Jk
~
9pr9qs)
~ JhrJisJj
~ 9hj9ik) =
Jk )
Rhijk,
which is t h e defining equation (1.3) of an AH3 manifold.
•
Now we want t o prove the following theorem for some smaller classes of almost Hermitian manifolds. T h e o r e m 4 . 2 . If an AH\ or AH{ 2n-manifold M2n for n > 1 h a s general sectional curvature K, then M2n is locally Hat.
constant
Proof. Suppose t h a t M2n is an AH\ manifold of constant general sectional curvature K in (1.1), we can easily obtain (4.1)
K(9hk9ij
- 9hj9ik + JhjJik ~ JhkJij)
= 0.
Multiplying (4.1) by gij we have (n — l)ghk K = 0, which implies K = 0. Hence, by (2.7), Rhijk = 0 is deduced. T h u s M2n is locally flat. If M2n is an AH[, manifold of constant general sectional curvature K, from (1.9), we have (4.2)
K{ghk9%j - 9hj9ik - JhjJik + JhkJij)
which yields nghkK
= 0. Hence M2n is also locally
= 0, flat.
•
140
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
§5. Holomorphic sectional curvatures In this section we discuss holomorphic sectional curvatures of AH manifolds. For an AH manifold, Friedland & Hsiung [1] have established Theorem 5.1. A necessary and sufficient condition for an almost Hermitian 2n-manifold M2n with an almost complex structure J^ and a Riemannian metric gij to be of constant holomorphic sectional curvature H at each point is that the Riemann curvature tensor Rhijk ofM2n with respect to g^ satisfies: (.tijrak "r tCkrsj)Jh
^i
' \K-irsk "T •K-krsi/'Jh ^ j
+ (Rirsj
+ RjrsijJk
+ (Rhrsj
+ Rjrsh)Ji
(5.1)
^k
~^~ {Rhrsk + RkTshl^i
Jj
+ {•"•hrsi + Rirsh)Jk
^j
Jk
= 4H(gkigjk + gij9hk + ghjgik)-
In this section, we use Theorem 5.1 to deduce a necessary condition for an AH 2n-manifold M2n of each special class to have constant holomorphic sectional curvature H at each point. At first we have Theorem 5.2. A necessary condition for an AH 2n-manifold M2n to be of constant holomorphic sectional curvature H at each point is that the Riemann curvature tensor Rhijk of M2n satisfies r> hijk
rL
_ T > TP J ( _ Jrij hipq<Jj Jk
p pqrsJ h "i
Jx
jPjQjrjs Jj Jk
1 1 4 + -Phijk + i;Qhijk + -^HGhijk,
(5-2) where
*hijk
=
{^•hpqk^'i
^ipqk^h
)Jj
+ (Rjpqk ~ ttjqpk)<Jh Ji
(5.3)
+ {RjqpiJh ~ RjqphJi )«^fc >
Whijk — (5-4)
(5.5)
^tipqjh^h + \RjipqJh
^i ~
< {^-hkpqJi ttjhpqJi
RikpqJ h J^j
Q
Hfc i
Ghijk — ghkgij — ghjgik + JhkJij ~ JhjJik ~ ^JhiJjk-
Furthermore, the Ricci tensor and scalar curvature of such a manifold are given respectively by (5.6)
Rij = -Rpq J." J." + 3Rpiqr J™ J/ + ZRpj^JwJS
+ 4(n +
l)H9ij,
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
R = 3Rp3rqJpqJrs
(5.7)
Proof. Multiplying (5.1) by Jh^J^ tively, we obtain (5.8)
+ 4n(n +
l)H.
and changing hi,ii
Rhijk + Rhkji = Thijk — ^H(JjiJjk
141
back to h,i
+ ghj9ik +
respec-
JhkJji,
where Thijk = (Rhqpk + Rhkpq)JiPJj
q
+ (Rjqpk + Rjkpq)Ji Jh \o.&)
\-Ttspqr ~r ftsrqpjJfi
+ (Rhqpi + Rhipq) - V Jj ' + (Rjqpi +
Rjipq)JkPJh9
^i Jj **k •
Interchange of h and i in (5.8) yields (5.10)
Rihjk + Rikjh = Tihjk - ^H{JihJjk
+ 9ij9hk +
JikJjh)-
Subtracting (5.10) from (5.8) and using (2.1) and (5.5) we easily have (5.11)
SRhijk = T^jk
— T^jk
+ 4 i / Ghijk-
On t h e other hand, from (5.9), (2.1), (5.3), (5.4) it follows t h a t J-hijk
*-ihjk
(5.12)
=
~otlhipqJj
Jfc ~ dUpqrsJh
i
j
fc
+ Phijk + Qhijk-
Substitution of (5.12) in (5.11) gives immediately (5.2). Multiplying (5.2) by ghk and using (3.1), (3.2) we can obtain (5.6). Moreover, (5.7) follows similarly by multiplying (5.6) by g%K T h e following theorem is a consequence of Theorem 5.2. T h e o r e m 5 . 3 . A necessary condition for an AH3 2n-manifold M2n constant holomorphic sectional curvature H at each point is that the curvature tensor Rhijk of M2n satisfies 2 Rhijk — 7;{RpqkjJh J% J<~RpkqiJh Jj ~ z(RpgJkJh (5.13)
+
-HGhijk-
Ji
+ Rpjiq^h
+ RpiqkJh Jj
^k )
+ RpiJqJh ^k )
to be of Riemann
641
142
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
Furthermore, the Ricci tensor and scalar curvature given respectively by Rij = 3RPiqrJpqJjr
(5.14)
+ 2(n +
R = 3RpSrqJpqJrs+4:n(n
(5.15)
of such a manifold
l)H9ij,
+
l)H.
Proof. Multiplying the defining equation (1.3) of an Aff 3 -manifold by and JalJbk, we obtain respectively (5.16)
RrsjkJh
J% — RhirsJj
(5.17)
RhrskJi
Jj
are
JahJbl
Jk >
— RirsjJh
Jk
Moreover, multiplying (5.17) by ghk and glk gives respectively (5.18)
RrsJ{
^O.iy^
Jhpiqru
Jj
Jj
= Rij:
— ihpjqrJ
c/j
Using (5.17) and (2.1), we can reduce (5.3) and (5.4) respectively to (5.20)
(5.21)
Phijk
Qhijk
= 2{RhmkJiq
- RipgkJ^J/
+ 2(RhkpqJip
= -2RhipgJ/Jk"
RhipqJjPJkq,
-
RikpqJ»)J.q.
-
Substitution of (1.3), (5.20), (5.21), (5.16) in (5.2) yields *jilhijk
(5-22)
~
dMpqjkJh
+ RhkpqJi
Ji
Jj
i ri.hpqk'Jj
J^ 9
— RikpqJh'Jj
RipqkJj
Jh
+2HGhijk-
By means of (5.16), (5.17), (2.1) the first five terms on the right-hand side of (5.22) can be reduced to — 6RpqjkJh
(5.23)
Ji
+ RpjiqJh
+ { — RipqjJh
Jk
— (—RipqkJj
Jjq +
Jk
— RipqkJj
+ ftipqjJk Jfi ) RipqkJjPJh9).
Substituting (5.23) in (5.22), we readily arrive at (5.13).
Jfr
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
143
(5.14) can be obtained by multiplying (5.13) by ghk and using (2.1), (5.19) or by applying (5.18), (5.19) to (5.6). Equation (5.15) follows immediately by multiplying (5.14) by g%i • Now suppose that M2n is an AH$ Einstein manifold of constant holomorphic sectional curvature H. Since M 2 n is an Einstein manifold, we have
(5-24)
Rij = J ^ y ,
which together with (5.14) implies (5.25)
Rij = SRpigrJVJf
+ 2(n + l)H9ij
=
—9ij, 2n~
so that Rij = aRPiqrJvqJjr = (IHgij, and therefore R = aRpiqrJpqJlr = 2n(3H. On an AH\ manifold of constant holomorphic sectional curvature H, /3= ^ [ 1 ] , that is (5.26)
H =
R n(n+ 1)
holds. We investigate an AH3 manifold satisfying (5.26) Lemma 5.4. If M2n is an AH3 Einstein 2n-manifold of constant holomorphic sectional curvature H with H = ni^+\)' ^nen (5.27)
yo.ZiO)
(5.29)
Rij =
•H'ij —
Q-=R
nonzero
n+1 -r—Hgij,
-itpiqrJ
Jj
+ RPiqrJpgJir
i
= 0.
Proof. Equation (5.27) follows from (5.24) and (5.26). Substituting (5.26) in the right-hand side of (5.25), we obtain (5.30)
(n + l)H9ij
=
^RngrJnjf.
Substitution of (5.30) in the left part of (5.25) gives (5.28) immediately. (5.29) is obtained by multiplying (5.28) by glj. D
144
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
Corollary 5.5. On a compact AHz Einstein 2n-manifold of constant positive holomorphic sectional curvature H satisfying (5.26), the Grst Betti number of M2n is zero. Proof. The result follows from (5.27) because the Ricci curvature tensor is positive definite [5]. It is known (see, for instance, [1]) that Q < 0 on an almost Kahlerian manifold, and Q > 0 on a nearly Kahlerian manifold. Furthermore, Q = 0 if and only if the manifold is Kahlerian. Thus the following corollary is obvious due to (5.29). Corollary 5.6. An almost Kahlerian or a nearly Kahlelrian Einstein 2nmanifold of constant nonzero holomorphic sectional curvature H satisfying (5.26) is Kahlerian. In the following, we shall discuss Theorem 5.3 for three special AH3 manifolds. Theorem 5.7. A necessary condition for an AH2 2n-manifold M2n to be of constant holomorphic sectional curvature H at each point is that the Riemann curvature tensor Rhijk satisfies Rhijk = ~^\RpqkjJh •A + RpkqiJfr Jj ^Rpjiq^h (5.31)
+
^k '
^HGhijk.
Furthermore, The Ricci tensor and scalar curvature of such a manifold are given respectively by (5.32)
Rij = ZRpiqrJPoj/
(5.33)
R = 3RpSrqJpqJrs
+ 2(n +
l)Hgijt
+ 4n(n + \)H.
Proof. Substituting the defining equation (1.2) of an AH2 manifold in the second term on the right-hand side of (5.13) we obtain (5.31) immediately. Multiplying (5.31) by ghk and using (5.19), (5.23) is deduced. Furthermore, (5.33) follows by multiplying (5.32) by gij. O Lemma 5.8. An almost Kahlerian or a nearly Kahlerian AHC manifold is Kahlerian.
M2n
Proof. The defining equation (1.4) of an AHC manifold is equivalent to (5-34)
Rhijk = RqhjpJi Jk + RhqkpJi Jj •
Multiplying (5.34) by ghkgli and using (2.1), we can easily obtain (5.29). Thus the lemma is proved in the same way as Corollary 5.6 was proved. •
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
145
T h e o r e m 5 . 9 . [3]. A necessary condition for an AHC 2n-manifo Id M2n to be of constant holomorphic sectional curvature H at each point is that the curvature tensor Rhijk satisfies (5-35)
Rhijk = RpqkjJhJiq
+
yHGhijk-
Furthermore, the Ricci tensor and scalar curvature given respectively by (5.36)
Rij =
(5.37)
R=
of such a manifold are
——Hgij, n{n+l)H.
As a consequence of (5.36), M2n is an Einstein manifold. Furthermore, if M2n is compact and H is positive, then the first Betti number of M2n is zero. Proof. Since a n AHC manifold is AH3, the curvature tensor Rhijk (5.13), which can be rewritten as Rhijk ~ Ttl+Rpqkj J h ^i + *h>qjkJj Ji + \RpikqJh + \RpiqjJh
^k + "fjiq^h
*A + RpkqiJh
^j )\RpkqiJ h / j + ^vjiq^h
satisfies
Jj ) ^k )
(5.38) + HGhijk] • By (2.1), (5.38) is reduced t o Rhijk = 'Zl'+RpqkjJh "A + \RpqjkJh
"'i + RpqkiJ h •* j
^^vqij^h
^k )
(5.39) + {RpkqiJh Jj + Rpjiq^h
^k ) + 2HGhijk],
which together with (3.21) and (5.34) gives (5.35). By multiplying (5.34) by ghk and using (5.18) we can easily obtain (5.40)
R^ =
\RqhjPJhqJiP-
On t h e other hand, from (3.8) it follows t h a t (5.41)
RpqkjJh
^i ~ ~RpkjqJi
Jh + RpjkqJi
Jh -
Multiplying (5.41) by ghk, we have (5.42)
RpqkjJ^Ji " =
-iRpkjqJ^Ji".
Multiplying (5.35) by ghk using (5.42) and (5.40), we thus arrive a t (5.36). E q u a t i o n (5.37) is obvious, and the last p a r t of this theorem follows from t h e same argument as in the proof of Corollary 5.5. •
146
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
Theorem 5.10 [1]. A necessary and sufficient condition for an AH\ 2nmanifold M2n to be of constant holomorphic sectional curvature H at each points is that the curvature tensor Rhijk satisfies (5.43)
Rhijk = jHGhijk-
Furthermore, the Ricci tensor and scalar curvature of such a manifold are respectively given by (5.36), (5.37). As a consequence of (5.36), M2n is an Einstein manifold. Furthermore, if M2n is compact and H is positive, then the first Betti number of M2n is zero. Proof. (5.43) is a consequence of (5.35), (1.1), (1.5). For the sufficiency of the theorem, we notice that from (3.6) it follows that M2n has constant holomorphic sectional curvature H is and only if (5.44)
RriskJhruhJjsujuk
-Hghiuhuigjkujuk
=
holds for any tangent vector ul of M2n. If (5.43) holds, then by substituting (5.43) in the left-hand side of (5.44), we can easily show that the left-hand side of (5.44) becomes automatically the right-hand side of (5.44). The other part of the theorem follows from the same argument as in the proof of Corollary 5.5. • §6. Holomorphic bisectional curvatures This section is devoted to a study of the holomorphic bisectional curvatures of AH manifolds. At first we have Theorem 6.1. A necessary and sufficient condition for an AH 2n-manifold M2n to be of constant holomorphic bisectional curvature B at each point is that the Riemann curvature tensor Rhijk satisfies
(6.1)
K-hpjqJi Jk J<~RipjqJh ^k + "-hpkqJi Jj + RipkqJhJj" = -ABghigjk.
Proof. To prove the necessity of condition (6.1) we assume that M2n is of constant holomorphic bisectional curvature B. Then, from (3.7) it follows that (6.2)
RhpiqJipJkquhuivjvk
=
-Bghi9jkuhuivjvk
for any tangent vectors u and v of M2n. By collecting all the coefficients of a general term uhulv^vk on the left-hand side of (6.2) by interchanging the
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
147
indices h,i,j,k in all possible cases, i.e., interchanging h, i and keeping j,k, interchanging j , k and keeping h, i and interchanging h, i and interchanging j , k at the same time, we can easily obtain the left-hand side of (6.1). In the same way we can show that all the coefficients of the general term uhulv^vk on the right-hand side of (6.2) is the right-hand side of (6.1). To prove the sufficiency of condition (6.1) we suppose that (6.1) holds. Multiplying both sides of (6.1) by uhulv:jvk for any tangent vectors u and v of M2n and summing for h, i, j , k we can see that all the terms on the left-hand side of the resulting equation are equal to each other. Thus (6.2) holds for any tangent vectors u and v of M2n, that is, M2n is of constant holomorphic bisectional curvature at each point. Hence the proof of this theorem is complete. Theorem 6.2. A necessary and sufficient condition for an AH 2n-manifold M2n to be of constant holomorphic bisectional curvature B with respect to gij at each point is that the Riemann curvature tensor Rhijk satisfies ~Rhijk (6.3)
=
RpqrsJh Ji Jj Jk ~t~ ^VQJk^h *A + RhipgJ/J^ + 4BJhiJjk.
Furthermore the Ricci tensor and the scalar curvature of such a manifold are given respectively by
(6.5)
R = RprqsJpqJrs
+ AnB.
Proof. To prove the necessity of condition (6.3), we suppose that M2n is of constant holomorphic bisectional curvature B. Then, by Theorem 6.1, we have (6.1). Multiplying (6.1) by Jr%Jsk, (6.3) is immediately deduced. Also, we obtain (6.4) by multiplying (6.3) by ghk, and obtain (6.5) by multiplying (6.4) by fl«. To prove the sufficiency of condition (6.3) we suppose that (6.3) holds. Multiplying (6.3) by JjJbk, (6.1) is readily obtained. Thus M2n is of constant holomorphic bisectional curvature at each point by Theorem 6.1. D The following corollary is an obvious consequence of (6.1) and (6.3). Corollary 6.3. A necessary and sufficient condition for an AH 2n-manifold M2n to be of zero holomorphic bisectional curvature at each point is that the Riemann curvature tensor Rhijk satisfies RhpjqJi Jk + -RipjqJh ^h + H-hpkqJi Jj (6.6)
+RiPk9JhpJjq
= 0,
148
C. C. HSIUNG, W. YANG AND L. FRIEDLAND or — Rhijk = RpqrsJh <J% Jj ^k "+" **-pqjkJ h *** (6.7)
+ RhipqJj
Jk •
R e m a r k . From (1.9) and (6.7) it follows immediately t h a t an AH[ has zero holomorphic bisectional curvature at each point. T h e o r e m 6 . 4 . A necessary condition for an AH3 2n-manifold of constant holomorphic bisectional curvature B at each point Riemann curvature tensor Rhijk satisfies (6.8)
—Rhijk = RhipqJjPJkq
Furthermore, the Ricci curvature are given, respectively, by (6.9)
(6.10)
+
R = Rhiqp
+
M2n to be is that the
^BJhiJjk-
and the scalar curvature
Rij = RhiQPJjPJhq
manifold
of such a
manifold
2B9ij,
Jip Jhq + 4nB.
Proof. By means of (1.3), we can see t h a t on the right-hand side of (6.3) t h e first t e r m is Rhijk, and the second and third terms are the same, so t h a t (6.3) becomes (6.8). Multiplying (6.8) by ghk, (6.9) is deduced and, furthermore, (6.10) is obtained by multiplying (6.9) by g%i. • C o r o l l a r y 6.5. zero holomorphic
A necessary condition for an AH3 manifold bisectional curvature is that M2n is an AH[
Proof. This follows immediately from (6.8) and (1.9)
M2n to be of manifold.
•
T h e o r e m 6.6. A necessary condition for an AKi 2n-manifold of constant holomorphic bisectional curvature B at each point Riemann curvature tensor Rhijk satisfies (6-11)
Rhijk — RpiqkJhVJjq
Furthermore, the Ricci curvature are given, respectively, by
+ B(ghigjk
-
JhiJjk)-
and the scalar curvature
(6.12)
R^ = RpiqkJWJj1*
(6.13)
R = RpiqkJT">Jik
+
2B9ij,
+ 4nB.
M2n to be is that the
of such a
manifold
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
149
Proof. Since M2n is an AH3 manifold by (1.9), (6.8) holds. By multiplying (6.8) by Jek, we obtain RpthiJj
— RjphiJe
= 2BJhigj(,
which becomes, after some changes of indices, RpijkJh — RhPjkJiP = ZBJjkghi-
(6-14) Similarly, we have (6-15)
RpkhiJjP — RjphiJkP = 2BJhi9jk-
Subtracting (6.14) and (6.15) from (3.14), which holds for an AH2 manifold, we have (6.16)
RhpjkJiP + RjphiJkP = -B(Jik9hi
+ Jhi9jk)-
By multiplying (6.16) by J ' , we arrive at (6.11). Multiplying (6.11) by ghk, (6.12) is deduced, and, furthermore, (6.13) is obtained by multiplying (6.12) by gli. D Theorem 6.7. If an AHC 2n-manifold M2n is of constant holomorphic bisectional curvature B, then B must be zero and M2n is an AH[ manifold. Proof. Since M2n is an AH3 manifold by (1.9), (6.14) holds. By changing the subscripts i,j, k cyclically from (6.14), we obtain two more equations. On the left-hand side of these three equations, the sum of the three first terms is zero by (2.1), and the sum of the three second terms is zero by (3.16), so that we obtain 2B(Jikghi + JkiQhj + Jij9hk) = 0, which implies B = 0, and hence M2n is an AH[ by Corollary 6.5.
•
§7. The relationship among the three types of sectional curvatures In this section we shall assume, unless stated otherwise, that M is an almost Hermitian 2n-manifold with an almost Hermitian structure J and an almost Hermitian metric g whose respective components are J J and gij. Moreover, let u and v be two unit tangent vectors of M at a point p, and let (f>, 9,0' be the angles between u and v, Ju and v,u and Jv, respectively. Then, from (2.4) and (3.8) it follows that the sectional curvature of M with respect to the two-dimensional plane (u, v) determined by two linearly independent unit tangent vectors u and v at p is given by (7.1)
K(u, v) = -R(u, v, u, v) s i n - 2 4>.
150
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
Theorem 7.1. IfM is an AHi manifold, then the sectional curvature K(u, v) and the holomorphic bisectional curvature B(u,v) satisfy B(u, v) = K(u, v) sin2 cf> + K{u, Jv) sin2 6. Proof. Prom (1.1), (2.1), (3.7) and (7.1) we obtain -RhijkuhJpiupvjJqkvq
B(u,v) =
Rhkij)uhJpiupvjJqkv'1
= (Rhjki
+
=
-Rhjikuhvij;vPJqkv« +
RhkijuhJqkv«Jpiv?vj -{RhjPqJiPJkq)uhvjuivk
=
(RhkPqJipJjq)uhJrkvruiJsjvs
-
= -RhjikUflvjuivk (7.3)
-
= K(u,v) sin2
RhkijUhJrkvruiJajva Jv)sm26.
Corollary 7.2. Assume that M is an AHi-manifold. If the two holomorphic planes II = (u, Ju) and II' = (u, Jv) are ortJiogonai, then (7.4)
B(u,v) =
K(u,v)+K(u,Jv).
Proof. This result is immediately deduced from the previous theorem and (3.10). • For a Kahlerian manifold, Goldberg and Kobayashi obtained (7.4) in [2], but missed the orthogonality condition of the two holomorphic planes (u, Ju) and (v, Jv). Corollary 7.3. IfM is an AH[ manifold, then the sectional curvature K(u, v) and the holomorphic bisectional curvature B(u,v) satisfy (7.5)
-B(u, v) = K{u, v) sin2 4> + K(u, Jv) sin 9.
In particular, if (u, Ju) and (v, Jv) are orthogonal, then (7.6)
-B{u,v)
= K{u,v) +
K(u,Jv).
Proof. This corollary is deduced by imitating the proof of Theorem 7.1.
•
If an almost Hermitian manifold M has constant general sectional curvature, then from the definition M must also have constant holomorphic sectional curvature, but the following theorem shows that M does not necessarily have constant holomorphic bisectional curvature.
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
151
Theorem 7.4. If an almost Hermitian manifold M has nonzero constant general sectional curvature K, then the holomorphic bisectional curvature B(u, v) of M satisfies (7.7)
B(u,v) = K(cos2<j> + cos26) =
tfcos(II,n'),
where (II, II') is the angle between the two planes II := (u, Ju) and U' := {v, Jv) defined by (2.12). Proof. By (3.7) and (2.7), we have B(u,v) = -K(ghkgij
-
ghjgik)uh-JJvPv3Jqkvq.
Substitution of (3.8) and (3.10) in the above equation gives immediately (7.7).
•
The following two corollaries are immediate consequences of Theorem 7.4 and (3.11). Corollary 7.5. For an almost Hermitian manifold M with nonzero constant general sectional curvature K, the holomorphic bisectional curvature B(u, v) has the same sign as K and
(7.8)
0
for
K > 0,
K < B(u, v)<0
for
K < 0.
Moreover, the absolute value \B(u,v)\ reaches zero, the minimum; when the two holomorphic planes U — (u, Ju) and II' = (v, Jv) are orthogonal, and reaches \K\ when U and TV coincide. Corollary 7.5. A necessary and sufficient condition for two holomorphic planes H = (u, Ju) and W = (v, Jv) of an almost Hermitian manifold M with nonzero constant general sectional curvature to be orthogonal is that the holomorphic bisectional curvature of M determined by II and II' is zero. Theorem 7.7. Let M be an AH\ manifold with nonzero constant holomorphic sectional curvature H. Then the general sectional curvature K(u, v) and the holomorphic bisectional curvature B(u,v) are
(7.10)
B{u,v) = Hcos2 M ^
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
152
respectively. Proof. By (7.1), (5.43), (5.5) and (3.8) we obtain -K(u, v) sin2 4> =
Rhijkuhviujvk ^HGhijkuhviu^vk
=
=- — (sin2./. + 3 cos2 6),
(7.11)
which implies (7.9). Similarly, we have, in consequence of (3.7) and (3.10), -B(u,v)
RhijkuhJpiupvjJqkv'1
=
±HGhiikuhJpWJqkv« = H cos2
(7.12) which implies (7.10).
,
•
The following corollary is an immediate consequence of (7.9). Corollary 7.8. Using the same notation as in Theorem 7.7 for orthogonal vectors u and v we have ^
for
H>0,
TI
(7.13)
—>K>H for H < 0. 4 In Corollary 7.8, f = K occurs when Ju and v are orthogonal, and H — K occurs when Ju and v coincide. Since a Kahlerian manifold is an AH\ manifold, the following corollary is obvious. Corollary 7.9. Both Theorem 7.7and Corollary 7.8 are true for a Kahlerian manifold. For (7.9) and Corollary 7.8 for a Kahlerian manifold see Kon and Yano [5, pp. 76-77]. Corollary 7.10. Any AHi or Kahlerian 2n-manifold M for n > 2 of constant holomorphic bisectional curvature is locally Hat. Proof. Since M is of constant holomorphic bisectional curvature B, every holomorphic sectional curvature H(u, Ju) is constant. Suppose H ^ 0. Then from (7.12), cos 2 ((II, II')/2) is constant, which is impossible for n > 2. Thus H = 0 which implies that Rhijk = 0 by (5.43). Hence M is locally flat. • For AHC manifolds, which are more general than AH\ manifolds, of nonzero constant holomorphic sectional curvature, we have the following theorem.
HOLOMORPHIC SECTIONAL AND BISECTIONAL CURVATURES
153
Theorem 7.11. If M is an AHC manifold of nonzero constant holomorphic sectional curvature H, then the general sectional curvature K and the holomorphic bisectional curvature B of M satisfy (7.14)
2
A _L + Ifl*. 0 ^ sin 2 n 2_6 -TT H r K{u, v) sin2
„2
(7.15)
Proof. It is known [3] that the Riemann curvature tensor Rhijk of M satisfies (7.16)
Rhijk = RhpqjJi Jk ~ 2^^hjSik
+ 9hi9jk + JhjJik + JfiiJjk)-
Substituting (7.16) in (7.1) and using (3.8), (3.10) and (7.1) for K(u,Jv), obtain K(u, v) sin 2 <j) - RhrjaJi
we
r
Jkauhviuivk
+ 2H(9hj9ik + 9hi9jk + JhjJik +
JhiJjk)uhvlujvk
RhijkuhJpivi>ujJqkv'>
=
+ -H(l + cos2
JSJ^Rhrjs^JjvPvijfv* + -H(ghjgik RhrjsUhUrVjVa
2 tf(i
which implies (7.15)
+ 9hi9jk + JhjJik + JhiJjk)uhJplupv3
Jqkvq
+ -H(1 + COS2
+ cos(n,n')), •
Corollary 7.12. If an AHC manifold M is of constant holomorphic sectional curvature H, then the sectional curvature K and the holomorphic bisectional curvature B satisfy (7.17)
B(u, v) = K{u, v) sin2 <£ + K(u, Jv) sin2 0.
Proof. The corollary follows immediately from (7.14) and (7.15).
•
Remark. Theorem 7.1 shows that (7.17) also holds for an AH\ manifold M, but there is no constant holomorphic sectional curvature condition on M.
154
C. C. HSIUNG, W. YANG AND L. FRIEDLAND
Corollary 7.13. If M is an AHC manifold of nonzero constant holomorphic sectional curvature H, then the holomorphic bisectional curvature B of M satisfies ^
y,
for
H>0,
for
H<0.
In Corollary 7.13, y = B(u,v) occurs when II = (u, Ju) and II'(v, Jv) are orthogonal, and H = B(u,v) occurs when II and II' coincide. References 1. L. Friedland At C.C. Hsiung, A certain class of almost Hermitian manifolds, Tensor 48 (1989), 252-263.. 2. S.I. Goldberg &z S. Kobayashi, Holomorphic bisectional curvature, J. Differential Geometry 1 (1967), 225-233.. 3. C.C. Hsiung & B. Xiong, A new class of almost complex structures (to appear)in Ann. Mat. Pura. Appl.. 4. K. Yano, Differential Geometry on Complex and Almost Complex Spaces, Macmillan, New York, (1965). 5. K. Yano & S. Bochner, Curvature and Betti numbers, Annals of Math. Studies 32 (1953), Princeton University Press, Princeton. Chuan-Chih Hsiung Lehigh University Wenmao Yang Wuhan University Lew Friedland State University of New York, Geneseo
SUT Journal of Mathematics Vol. 32, No. 2 (1996), 163-178
THE SPECTRAL GEOMETRY OF SOME ALMOST HERMITIAN MANIFOLDS Chuan-Chih Hsiung, Wenmao Yang and Bonnie Xiong (Received October 3, 1996) A b s t r a c t . Let (Mj,gi) be a certain almost Hermitian 2n-manifold Mi with a Hermitian metric gi for t = 1,2, which is more general than an almost L manifold (a Kahlerian manifold is known to be a special almost L manifold). Let Specp(Mi,gi) denote the spectrum of the real Laplacian on p-forms on M%. The purpose of this paper is to show that for some special values of p and n, if Specp(M\,gi) = Specp(M^,g2), then ( M i , g i ) is of constant holomorphic sectional curvature H\ if and only if (M2,$72) is of constant holomorphic sectional curvature H?, and H2 — H\ • The corresponding results on almost L manifolds were obtained by C. C. Hsiung and C. X. Wu (The spectral geometry of almost L manifolds, Bull. Inst. Math. Acad. Sinica, 23 (1995), 229-241). AMS 1991 Mathematics
Subject Classification.
Primary 53C15, 53C35, 58G25.
Key words and phrases. Spectrum, Laplacian, almost complex structures, almost Hermitian structures, holomorphic sectional curvature, Bochner curvature tensor.
1. Introduction Let (M,g) be an m-dimensional compact Riemannian manifold M with a Riemannian metric g. Throughout this paper all manifolds are supposed to be C°° and connected. The set of the eigenvalues associated with all the peigenforms, 0 < p < m, with respect to real Laplacian A and the metric g of M is called the spectrum of A on p-forms on M, which will be denoted by SpecP(M,g).Thus (1.1)
Specp{M, g) = {0 > A liP > A2,p >
> -00},
where each eigenvalue AjiP, i = 1,2,..., is repeated as many times as its multiplicity; which is finite, and the spectrum Specp(M,g) is discrete since A is an elliptic operator. It is well known that there are various examples ([13], [7] and the references there) of a pair of nonisometric manifolds with the same spectrum. Thus the The work of the second author was partially supported by the National Natural Science Foundation of the People's Republic of China and the C.C.Hsiung Fund at Lehigh University. 163
655 164
C- C. HSIUNG, W. YANG AND B. XIONG
spectra do not determine a manifold up to an isometry. However, the relationship between the geometry of a Riemannian or Kahlerian manifold and its spectra has been extensively studied. Let (M,g) and (M',g') be compact Riemannian (respectively, Kahlerian) manifolds M, M' with Riemannian (respectively, Hermitian) metrics g and g' and Specp(M,g) = Specp(M',g') for a fixed p. Various authors [1], [4], [6], [14], . . . , [17] have shown that for some spacial values of p and m, (M, g) is of constant sectional (respectively, holomorphic sectional) curvature H if and only if (M',g') is of constant sectional (respectively holomorphic sectional) curvature H' and H = H'. Recently Hsiung and Wu [8] have generalized the results on Kahlerian manifolds to almost L manifolds of which Kahlerian manifolds are special ones. The purpose of this paper is to further generalize the results of Hsiung and Wu to more general almost Hermitian manifolds. In §2 there is a classification of all almost complex structures on a Riemannian manifold, together with inclusion relations among all classes, by means of the Riemann curvature tensor and the tensor of an almost complex structure. In §3 we define various almost Hermitian structures and manifolds, and give necessary and sufficient or just necessary conditions for some classes of almost complex manifolds defined in §2 with an Hermitian structure to have constant holomorphic sectional curvature at each point. §4 contains some fundamenzal formulas for a Riemannian structure and the well-known Minakshisundaram-Pleijel-Gaffney's formula for the spectra of a Riemannian manifold. §5 deals with the spectral geometry of some almost Hermitian manifolds which are more general than L manifolds. 2. Almost complex structures Let M be a Riemannian 2n-manifold, and let gij,Rhijk,Rij,R, and J^ denote, respectively, the Riemannian metric tensor, the Riemann curvature tensor, the Ricci curvature tensor, the scalar curvature and the tensor of an almost complex structure J of M. Let {gli) be the inverse matrix of the matrix (gij). Throughout this paper all Latin indices take values l,--- ,2n unless stated otherwise. We shall follow the usual tensor convention that indices of tensors can be raised and lowered by using gt:> and g^ respectively, and that repeated indices imply summation. Moreover, if we multiply, for example, the components a,ij of a tensor of type (0,2) by the components b^k of a tensor of type(2,0), it will be always understood that j is to be summed. By using the following identities for the relationship between jj and Rhijk i Hsiung and Xiong [9] have denned the following four classes of almost complex structures on the manifold M:
(2.1)
•tthijk — Jh ^i
-^rsjky
ALMOST HERMITIAN MANIFOLDS
[^•^)
K-hijk
\4-o)
\*"^l
=
Jh Ji Rrsjk
ri-hijk
"%i Jii
^-rsi3k
=
+ Jh Jj Rrisk
Jh Ji Jj
T Ji? Ji3
Jk
+ Jh Jk
165
Rrijs-,
Rpqrsi
-"Vsiifc T J%3 Ji\
i^-rsiik
=
"•
Let £ and A denote respectively the classes of almost complex structures (or manifolds) and the Kahlerian structure (or manifolds). Let £i,£2,£3 and € denote the classes of almost complex structures (or manifolds) satisfying (2.1),- • •, (2.4) respectively. Hsiung and Xiong [9] have showed the following inclusion relations. C ^2r (2.5)
#C£i
£3C£.
Thus for i = 1,2,3, as i decreases the structures (or manifolds) in £, resemble Kahlerian structures (or manifolds) more closely. For simplicity, throughout this paper if an almost complex manifolds M admits a certain special almost complex structure, then M is also called by the same name as the structure's. If JiJ and gij satisfy (2.6)
gijJh%Jkj = 9hk,
then the almost complex structure J is called an almost Hermitian structure, and g^ is called an Hermitian metric. For simplicity, throughout this paper, unless stated otherwise, by an almost Hermitian manifold M we shall always mean a manifold with an almost Hermitian structure J and an Hermitian metric g^. Friedland and Hsiung [5] called an almost Hermitian structure J an almost L structure if it satisfies
(2.7)
[Vj, Vk]Jih = (V.Vfc - VfcV^J^ = 0,
where V denotes the covariant derivation with respect to g^. Obviously, Kahlerian structures are almost L structures since an almost Hermitian structure Ji3 is Kahlerian if
(2.8)
ViJjk = 0
for all i, j , k.
For simplicity we shall denote an almost Hermitian £j structure by AH{ for i = 1,2,3, and a Kahlerian structure, an almost Hermitian <£ structure
C- C. HSIUNG, W. YANG AND B. XIONG
166
and an almost Hermitian structure respectively by K, AHC, and AH. From (2.5) we thus obtain the following inclusion relations among almost Hermitian structures: CAH2
(2.9)
KcAHi
) AH3 c AH.
Q
AHC°
We [10] also have defined an AH[ manifold to be an almost Hermitian manifold satisfying (2.10)
Rhijk — —Jh J% Rrajk-
Since the difference between (2.1) and (2.10) is only a sign, AH[ C AHC C AHz and the intersection of the two classes AH\ and AH[ is the class of locally Euclidean spaces, that is, the classe of spaces with Rhijk = 0. Now we introduce one new classe of AH\ manifolds which are almost Hermitian manifolds satisfying (2.11)
2Rhijk = Jh "i Rrsjk + Jh Jj Rrkia + Jh Jk H-rjai'
3. Almost Hermitian structures In this section M i s a Riemannian manifold as in §2. If there exists on M a tensor J J of type (1,1) satisfying JSJjh
(3.1)
= -**,
where 6* are the Kronecker deltas defined by
(3.2)
tf=p-
:=*• 10,
i ? k,
then Ji° is said to define an almost complex structure on M, and M is called an almost complex manifold. If an almost complex J£ is almost Hermitian, then as a consequence of (3.1) and (2.6) the tensor J,j of type (0,2) defined by (3-3)
Jij = gjkJi
is skew-symmetric. Thus (3.4)
JyJkj=9ik,
JjiJjk=9ik,
and for any tangent vector v% of M, (3.5)
9ij
Jfc W
= 0,
ALMOST HERMITIAN MANIFOLDS
167
which shows that vl is orthogonal to its transform JJVK Furthermore, on an almost Hermitian manifold M, there is a differential form u = Jijdx1 A dx\
(3.6)
where x 1 , . . . ,x2n are local coordinates on M, and the wedge A denotes the exterior product. If the differential form w is closed, that is, if (3.7)
du = 0,
then J^ is called an almost Kahlerian structure. From (3.6) and (3.7) it follows that an almost Kahlerian structure satisfies (3-8)
Jhij = VhJi:j + Vi Jjh + VjJhi = 0.
The tensor J^ij is skew-symmetric in all indices. Lemma 3.1. An almost Hermitian 2n-manifold M with an Hermitian metric gij and an almost Hermitian structure J^ is Kahlerian if it satisfies (3.9)
R+^JijJklRijkl=0,
Q=
(3.10)
Rhijk =
jHGhijk,
where H is a nonzero constant, and (3.11)
Ghijk = 9hk9ij — 9hj9ik + JhkJij ~ JhjJik ~ %JhiJjk-
Proof. Substituting (3.10) and (3.11) and four similar equations in Bianchi identity (4.2) gives (3.12) ^l{JhkJ%j — JhjJik — ZJhiJjk) + Vj(JftjJjfc - JhkJil ~ 2JftiJfc;) + Vfc(J/jjJu - JhlJij - 2JhiJlj) = 0. Expanding (3.12) by differentiating covariantly, multiplying the resulting equation by Jlk and using J^VhJij = 0 we obtain (3.8) which shows that M is almost Kahlerian. Hence M is Kahlerian since it is known that an almost Kahlerian manifold with condition (3.9) is Kahlerian (see for instance, [5,
p.261]).
•
Let M be an almost Hermitian manifold with an almost Hermitian structure •V satisfying (2.6). Then the two-dimensional plane determined by an arbitrary tangent vector ul of M and the tangent vector Jfv? at a point p of M is called a holomorphic plane (u, Ju), and the sectional curvature with respect to the plane (u, Ju) is called the holomorphic sectional curvature H(u, Ju) of M at p. If the holomorphic sectional curvature at a point p is independent of the holomorphic plane through p, then M is said to have constant holomorphic sectional curvature at p. Concerning constant holomorphic sectional curvature we have the following theorems:
168
C- C. HSIUNG, W. YANG AND B. XIONG
T h e o r e m 3 . 1 [5]. A necessary and sufficient condition for an almost L Inmanifold M to be of constant holomorphic sectional curvature H at each point is that the Riemann curvature tensor Rhijk with respect to the metric gij satisfy (3.10). Furthermore, the Ricci tensor and scalar curvature of such a manifold are given respectively by (3.13)
Rij =
(3.14)
^j-Hgij,
R = n(n + 1)H.
As a consequence
of (3.13), M is an Einstein
manifold.
T h e o r e m 3.2[9]. A necessary and sufficient condition for an AHC 2n-manifold M to be of constant holomorphic sectional curvature H a t each point is that the Riemann curvature tensor Rhijk with respect to the metric g^ satisfy (3.15)
Rhijk = JhrJj"Rskir
~ jH{gkjghi
+ gki9hj — JkjJhi ~
JkiJhj)-
Furthermore, the Ricci tensor and scalar curvature of such a manifold are given respectively by (3.13) and (3.14). As a consequence of (3.13), M is an Einstein manifold. T h e o r e m 3.3[10]. A necessary condition for an AHi 2n-manifold M to be of constant holomorphic sectional curvature H at each point is that the Riemann curvature tensor Rhijk v/ith respect to the metric g^ satisfy (3.16)
Rhijk — y(RpqkjJhPJiq
+ RpkqiJh9JjQ
+ RpjiqJh9 Jk*) +
Furthermore, the Ricci tensor and scalar curvature given respectively by (3.17)
(3.18)
R^ = mviqrJpqJjT R = ZRpSrqJp
^HGhijk-
of such a manifold
+ 2(n +
l)Hgij,
+ 4n(n +
l)H.
are
4. S p e c t r a of R i e m a n n i a n m a n i f o l d s Let (M, g) be a Riemannian manifold of dimension m > 2 with a Riemannian mertic g = (gij). We shall use all the notation with 2n = m, in §2, and also t h e following identities: (4-1)
Rhijk + Rhjki + Rhkij = 0,
ALMOST HERMITIAN MANIFOLDS
and (4.2)
VlRhijk + ^jRhikl + ^kRhilj = 0.
(4.2) is called the Bianchi identity. In 1919 Einstein suggested the following equation, called the Einstein equation, R
(4.3)
±ij — Kij
rn^oi
where R is the scalar curvature: (4.4)
gijRij,
R =
and Tij is called the stress-energy tensor. When Ty = 0, (M, g) is called an Einstein manifold and gij an Einstein mertic. Assume that M is compact. To study Specp(M,g) given by (1.1) we need the following Minakshisundaram-Pleijel-Gaffney's formula: oo
*
(4.5)
oo
no Airt^r • ^
where (4.6)
(4.7)
a
i,p
1 fm 6 IP
m-2 P-l
)]/•
tfdM,
/ J JM
(4.8)
a 2|P = / [c 1 (m,p)i? 2 + c 2 (m,p)|iZi j | 2 + c 3 (m,p)|J , ? /lijfc | 2 ]dM, ./A/
and (4.9)
.
1/
m
\ 1 (m-2\
, 1 fm-4
i(::i2)-2(;-;
(4.10)
e « =-i(;)
+
(4.1D
C3KP) = ^ ( ™
m-2\ 12 Vp-1 J
1 fm-4 2 VP-2
169
661 170
C - C. HSIUNG, W. YANG AND B. XIONG
dM being the volume element of M, I
l a binomial coefficient, and \Rij\
and |i?jitjk|2 the square of the lengths of the Ricci and Riemann curvature tensors respectively given by (4.12)
\Rhijk\2
IRitf^&Rij,
=
RhiikRhiik.
The coefficients a 0 , p ,ai,p and tt2,p have been calculated for p = 0 by many authors (see [1], [12]), and determined for all p by V.K.Patodi [14]. Remarks. 1. Let (M,g) and (M',g') be compact Riemannian manifolds. If Specp(M,g) = Specp(M',g') for some p, then from (4.5) and (4.6) we have (i) m = dim M = dim M' = m', (ii) Vol.M = Vol.M'. 2. For a geometric quantity A on (M, g), we shall denote the corresponding quantity on (M',g') by A'. 5. AH2,i manifolds The purpose of this section is to study the spectral geometry of the almost Hermitian manifold M satisfying (2.2) and (2.11), which we call an AH2,4manifold. At first we need Lemma 5.1. An AHi-manifold is an AH2,4-manifold. Proof. Substituting (2.1) for the second and the third terms on the right-hand side of (2.11) and using (4.1) we can easily see that the right-hand side of (2.11) becomes automatically the left-hand side of (2.11). So an Ai^-manifold is an AH4-maniiold and therefore an ^4if2,4-manifold, since from (2.9) an AH\manifold is also an ylfi^-mamfold. • Theorem 5.1. A necessary and sufficient condition for an AH?.^ In-manifold M to be of constant holomorphic sectional curvature H at each point is that the Riemann curvature tensor Rhijk with respect to the metric g^ satisfy (3.10). Furthermore, the Ricci tensor and scalar curvature of such a manifold are given respectively by (3.13) and (3.14). As a consequence of (3.13), M is an Einstein manifold. Proof. Suppose that M is of constant holomorphic sectional curvature H at each point. Since M is an AH2 manifold, (3.16) holds. Since M is also an AH4 manifold, (2.11) also holds. Substituting (2.11) in the right-hand side of (3.16) gives condition (3.10) immediately. For the proof of the sufficiency of condition (3.10) one may see [5] as Theorem 3.1 also has this condition. • Now we want to derive some relations among the tensors Jf, Rhijk and Rij of an AH4 2n-manifold M. At first, from (4.1) follows \0.4J
J
rtfkis
—
J
ftrisk
J
rXksri — ^"J
^riks^
662 ALMOST HERMITIAN MANIFOLDS
171
Multiplying (2.11) by ghk and using (3.4), (5.4), we obtain (5.5)
ZRij
= J
Ji Rrajk
+ 2J
Jj
Rriks-
l
Similarly, multiplication of (2.11) by g i and use of (5.4) give (5.6)
2Rhk
+ Jh JkSRrs
3JhrJtSRikrs-
=
Substituting the right-hand side of (5.6) for each term on the right-hand side of (5.5) yields readily which together with (5.6) implies that (5-8)
Rhk — Jh J**'Rikra =
J^J^Rihrs,
where the last equality is due to the symmetry of h and k. Multiplying (5.8) by ghk gives JisJhrRihrs
(5.9)
= R.
l
Multiplying (5.4) by Jj and using (5.8) we obtain J rJj Rkrsi = —2Rjs.
(5.10)
Multiplication of (5.10) by gls gives JkrJsiRkrsi
(5.11)
= -2R.
%
Multiplying (5.7) by Jk yields (5.12)
JkRij
= —JjSRka,
which together with (3.5) shows that On the other hand, since we have \0.iQ)
J
•ft'rpks
:=:
*J -L^rkps'
From (5.8) and (5.14) it follows that (5.15)
Rhk = JhrJ"Rikra
= ~Jk
J%SRirks-
Multiplying (5.1) by Jjh gives (5.16)
JlSRikjs = JkrJj
J1SRirha-
Moreover, using (5.8), (3.4), (4.12) we readily obtain (5.17)
\Rij\ = \Jl8Rikra\ •
The following lemma is an immediate consequence of (5.11), Theorem 5.1 and Lemma 3.1.
663 172
C - C. HSIUNG, W. YANG AND B. XIONG
L e m m a 5 . 2 . If an AH2li 2n-manifold M is of constant holomorphic curvature H at each point, then M is Kahlerian.
sectional
T h e Bochner curvature tensor B = (Bhijk) of an almost Hermitian mmanifold M with an Hermitian mertic g and an almost Hermitian structure Ji3 is denned as follows: (5.18)
Bhijk
1 - ^—^Ahijk
= Rhijk
R +
4 ( n + 1 ) ( n +
2 )
<W,
where the components of the tensors G — (Ghijk) and A = (Ahijk) respectively by (3.11) and (5.19)
are given
Ahijk = o-hijk + bhijk — "Zchijk,
with
(5.20)
yO.ZL)
ahijk = ghkRij - 9hjRik + gijRhk -
Ohijk
=
"ijJh
**>rk
Jik^h
J*
Chijk = JjkJhTRri
(5-22)
gikRhj,
+ JhiJj
^rj
Jhj"i
-^rfc>
Rrk-
Now we have the following crucial lemma. L e m m a 5 . 3 . For an AH4 2n-manifold
M,
2
\ R ^ +
(5.23)
|IW|
= \Rhljk?
- ^ -
2
2 (n+1)
(n +
2)
*2-
Proof. From (5.18) it follows t h a t \Bhijk\
=\Rhijk\
—-^RhijkA
lJ
r>
(5 24)
'
+
2(n+l)(n
-1
Rh kGhiJk
+ 2) «
~ 4(n + l ) ( n + 2)*AhijkGh13"
+
l^W^^ +
16(n + l ) 2 ( n + 2) 2 | G f c y f c | '
By (3.1), (3.4) and other equations of this section some elementary b u t complicated computations give the following: \ahijk\2 =8(n-l)\Rij\2
+ 4R2,
ALMOST HERMITIAN MANIFOLDS
ahijkbhiik ahijkchijk
(5.25)
= SlRijl2
by (5.13),
= -8|i?y|2
by (5.13),
2
|^ijfc| = 8(n-l)|JRij| bhijkchiik
2
= SlRij|2 2
= 4\Rij\2,
Rhijkchijk
=-AlRijf
+ 4i? 2 , by (5.13),
2
+ 2R2;
\chijk\ =4n\Rij\ Rhijkahijk
173
Rhijkbhijk
= 4|i? i ,| 2
by (5.15),
(5.26)
(5.27)
RhijkGhijk
= 8R
hijk
(5.28)
by (5.10); by (5.9), (5.11);
ahijka
= 8(n - 1 ) | ^ | 2 + 4R2,
ahijkbhijk
= 8\Rij\2
ahijkchiik
by (5.13),
=-8\Rtj\2
2
\bhijk\
bhijkchiik
by (5.13), 2
+ 4R2,
= 8(n-l)\Rij\ =-8\Rij|2
by (5.13),
2
| c W j i t | = 4 n | i ? i / + 2J?2; a«j f c G^ f c = 8(n + 1)R, (5.29)
6 W j f c G^ f c = 8(n + l)fl, ChijkGhijk
= - 8 ( n + l)i?;
|Gw^|2=32n(n+l),
(5.30)
From (5.19), (5.25), • • •, (5.29) we obtain RhijkAhijk \Ahijk\2
l6\Rij\2,
=
= Z2{n + 2)\Rij\2 + 16R2,
AhijkGhijk
= 32(n + 1)R,
\Ghijk\ = 16n(n + 2). Substituting (5.27) and (5.31) in (5.24) yields (5.23) immediately.
•
Assume that M is compact. Now we can express the coefficient 02,p of formula (4.5) in terms of \Bhijk\ and (5.32)
m/Hflyl
which follows from (4.3) readily.
2
-^*
2
.
C- C. HSIUNG, W. YANG AND B. XIONG
174
Lemma 5.4. For a compact AH4 2n-manifold M, a 2 , p = [ Mn,p)\Bhijk\2
(5.33)
+ b 2 (n,p)|T i; ,| 2 +
b3(n,p)R2]dM,
JM
where bi(n,p) = c 3 (m,p), (5.34)
b2(n,p) = c2(m,p) + ——c 3 (m,p), n +1 1 2 h(n,p) = ci(m,p) + —c2(m,p) + ———c3(m,p), In n{n + i) and m = 2n. Proof. The lemma is an immediate consequence by substituting (5.23) and (5.32) in (4.8). • Lemma 5.5. An AH2,4 2n-manifold M for n>2 is of constant holomorphic sectional curvature if and only if the tensors B and T = (Tij) are zero. Proof. Suppose M to be of constant holomorphic sectional curvature H. Then (3.10), (3.11), (3.13) and (3.14) hold by Theorem 5.1. Substituting (3.10), (3.11), (3.13) and (3.14) in (5.18) shows readily that Bhijk = 0. T{j = 0 follows from (4.3), (3.13), (3.14). Conversely, suppose that B = 0 and T = 0 which implies that M is an Einstein space, so that R is constant for 2n > 4. Substituting Rij = —gu in 2n (5.18) and using (3.1) and (2.6), we obtain Rhijk
An(n + l)GhijkHence, by Theorem 5.1. M is of constant holomorphic sectional curvature H = —*
=
D
n(n+ 1) 6. The main theorem The main results of this paper are listed in the following theorem. Theorem 6.1. Let (M,g, J) and (M',g', J') be compact AH2t4 2n-manifolds with almost Hermitian structures J and J', Hermitian mertics g and g'. Let (CPn,go,Jo) be the complex n-dimensional projective space CPn with the Fubini-Study metric go and the standard complex structure Jo- Consider the following statements: (1) (*) (M,g, J) is of constant holomorphic sectional curvature H if and only if(M',g', J') is of constant holomorphic sectional curvature H', and H = H'; (2) (**) (M,g,J) is Kahlerian and holomorphically isometric to (CPn, 9o,Jo)-
ALMOST HERMITIAN MANIFOLDS
175
Then we have the following: (i) (*) is true ifSpec°(M,g) = Spec°(M',g') and 2n < 10. (ii) (**) is true ifSpec°{M,g) = Spec0(CPn, g0) and In < 10. (iii) (*) is true if Spec1(M,g) = Spec1(M',g') and 2n = 2 or 16 < 2n < 102. (iv) (**) is true if Spec1 (M, g) = Spec1 (CPn,gQ) and 2n = 2 or 16 < 2n < 102. (v) (*) is true if Spec2{M,g) = Spec2(M',g') and 2n = 2,6,8,14 or 18 < 2n < 188. (vi) (**) is true if Spec2{M,g) = Spec2(CPn,g0) and 2n = 2,6,8,14 or 18 < 2n < 188. (vii) (*) is true ifSpecp(M, g) = Specp(M', g') for p = 0 and 1. (viii) (**) is true ifSpecp{M,g) = SpecP(CPn,g0) for p = 0 and 1. Remark. For almost L manifolds M and M' Theorem 6.1 is due to Hsiung and Wu [8]. When (M,g, J) and (M',g', J') are Kahlerian manifolds, parts (i)-(vi) of Theorem 6.1 are reduced to the known results in ([14],[15],[16]) mentioned before. Proof of Theorem 6.1. It is clear that (*) and (**) hold in the case of n = 1. So we assume n > 2. (i) From (4.6), (4.7), (4.9), . . . , (4.11) for m = 2n, and (5.33), (5.34) we obtain ao 0 = f dM = Vol M,
(6.1)
JM
(6.2)
(6.3)
o i >°
a2,0
91B 360 JM
=
fi /
RdM,
,2 , 2 ( 6 - n ) n+2
'hijk I
2
5n2+4n + 3 n(n+l)
2
dM.
Since the roles of (M,g, J) and (M',g', J') in the theorem are the same, we need only to prove the "if" part, and the "only if" part can be proved in the same way as by interchanging the roles of (M, g, J) and (M',g',J'). So we assume that (M',g',J') has constant holomorphic sectional curvature H'. Then R' — 2H' by Theorem 5.1, and therefore R' is constant. Thus from ao,o = a'o,o and ai,o = a'i,o it follows that JM R2dM > JM, R'2dM'. In fact, using the Schwarz inequality we
667 176
C- C. HSIUNG, W. YANG AND B. XIONG
have
dM) ( [ R2dM) > ( [ R CIM) = ( [ R' dM'] IM
/ \JM
J
\JM
= R'2(Vol
(6.4)
J M'f
\JM'
J
= R'2Vol M • Vol M' = VolM-
[
R'2dM.
JM'
On the other hand, applying Lemma 5.5 to (M',g', J') gives | B ' ^ f c | 2 = |T y | 2 = 0.
(6.5)
Thus from 02,0 = a'2,0, (6.4), (6.3) and its corresponding equation (6,3)' for {M',g',J') and In < 10, it follows that
(6.6)
\Bhm\2
= \Ttj\2 = 0,
/ R2dM - I JM
R'2dM'.
JM'
Hence by Lemma 5.5, (M, g, J) is of constant holomorphic sectional curvature H = H'. (ii) Since ( C P n , go, Jo) is of constant holomorphic sectional curvature c > 0, (M, g, J) is of constant holomorphic sectional curvature c > 0 by part (i). From Lemma 5.2 it follows that (M, g, J) is Kahlerian. Hence (M,g, J) is holomorphically isomertic to (CPn,go, Jo). (iii) As in part (i) we have a01=2n ao,i =2n
(6.7)
1dM = In Vol M, JM
(6.8)
aM = ! L * •J
02,i = ^ (6.9)
j
{2{2n - lh)\Bhijk\2
/ RdM, JA M
+ ^ [ n ( 5 1 - n) + 30]|Ttf |
2 + — — fn2(5n - 26) + 18n + 15}R2\dM. n(n+l) For 8 < n < 51, all coefficients of \Bhijk\ , |T»j| and R2 in (6.9) are positive. Also as before we may assume that (M',g', J') is of constant holomorphic sectional curvature H' so that R! is constant and (6.5)
ALMOST HERMITIAN MANIFOLDS
177
holds. We also have (6.4). Thus from a2,i = a'2,1 and (6.9), (6.9)' follows (6.6). Hence (M,g,J) is of constant holomorphic sectional curvature H = H''. (iv) follows from (iii) and Lemma 5.2. (v) As before we have (6.10)
a0,2 = n(2n - 1) /
dM = n(2n - l)Vol M,
JM
ai, 2 = \{2n2 - 13n +12) f
(6.11)
(6.12)
R dM.
JM
D
a2>2 = 3 ^ / ( 4 i | £ W i f c | 2 + ^2|T«| 2 +
A3R2)dM,
where Ax = 2(2n 2 - 31n + 120)
(6.13)
A2 =
- ( 2 n 3 - 1 9 3 n 2 + 426n + 120), n+2
A3 = — rr(10n 4 - 117n3 + 362n2 - 183n - 60). n(n + 1) Thus for n = 3,4 or 7, or 9 < n < 94, A\, A2, and A3 are positive. The remaining part of the proof is completely similar to that in (i) or (iii). (vi) follows from (v) and Lemma 5.2. (vii) Multiplying (6.3) by (2n - 15) and subtracting the resulting equation from (6.9) we obtain (6.14)
02,1 - (2n - 15)a2,0 = 15 / JM
10|rij| 2 + - ( n + 5)i? 2 dM. . n
Assume that (M',g',J') is of constant holomorphic curvature H'. Then we have (6.5) and (6.4) which together with 02,1 — (2n— 15)a2,o = a'2,1 ~ ( 2 n - 15)a'2,o, (6.14) and (6.14)' implies that |T^| 2 = 0. Thus from 02,0 = a'2,0) (6.3) and (6.3)' it follows that \Bhijk\ — 0. Hence (M,g, J) is of constant holomorphic sectional curvature H = H''. (viii) follows from (vii) and Lemma 5.2. •
669 178
C - C. HSIUNG, W. YANG AND B. XIONG
References [I] M.Berger, Le spectre des varietes riemaniennes, Rev. Roumaine Math. Pures Appl. 1 3 (1969), 915-931. [2] M.Berger, P.Gauduchon and E.Mazet, Le spectre d'une variete riemanienne, Lecture Notes in Math., vol. 194, Springer, Berlin, 1971. [3] E.Cartan, Lecons sur la geometrie des espaces de Riemann, 2nd ed., Paris (1946). [4] B.Y.Chen and L.Vanhecke, The spectrum of the Laplacian of Kahler manifolds, Proc. Amer. Math. Soc. 79 (1980), 82-86. [5] L.Friedland and C.C.Hsiung, A certain class of almost Hermitian manifolds, Tensor 48 (1989), 252-263. [6] S.I.Goldberg, A characterization of complex projective spaces, C. R. Math. Rep. Acad. Sci. Canada 6 (1984), 193-198. [7] C.Gordon, Isospectral closed Riemannian manifolds which are not locally isomertic, J.Differential Geometry 3 7 (1993), 639-649. [8] C.C.Hsiung and C.X.Wu, The spectral geomerty of almost L manifolds, Bull.Inst. Math. Acad. Sinica. Taiwan 2 3 (1995), 229-241. [9] C.C.Hsiung and B.Xiong, A new class of almost complex structures, Ann. Mat. Pura. Appl. 168 (1995), 133-149. [10] C.C.Hsiung, W.Yang and L.Friedland, Holomorphic sectional and bisectional curvatures of almost Hermitian manifolds, SUT J. Math 31 (1995), 133-154. [II] A.Lichnerowicz, Courbure, nombres de Betti et espaces symmertiques, Proc. Internat. Congress Math. (Cambridge, Mass. 1950), Amer. Math. Soc. I I (1952), 216-223. [12] H.F.McKean and I.M.Singer, Curvature and the eigenvalues of the Laplacian, J. Differential Geomerty 1 (1967), 43-69. [13] J.Milnor, Eigenvalues of the Laplace operator on certain manifolds, Proc. Nat. Acad. Sci. U.S.A. 5 1 (1964), 542. [14] V.K.Patodi, Curvature and the fundamental solution of the heat operator, J. Indian Math. Soc. 34 (1970), 269-285. [15] S.Tanno, Eigenvalues of the Laplacian of Riemannian manifolds, Tohoku Math. J. 25 (1973), 391-403. [16] , The spectrum of the Laplacian for 1-forms, Proc. Amer. Math. Soc. 4 5 (1974), 125-129. [17] GR.Tsagas and C.Kockinos, The geometry and the Laplace operator on the exterior 2-forms on a compact Riemannian manifold, Proc. Amer. Math. Soc. 7 3 (1979), 109116. Chuan-Chih Hsiung Department of Mathematics, Lehigh University Bethlehem, PA 18015, USA Wenmao Yang Department of Mathematics, Wuhan University Wuchang, People's Republic of China Bonnie Xiong Department of Mathematics, University of Scranton Scranton, PA 18510, USA
BULLETIN OF THE INSTITUTE OF MATHEMATICS ACADEMIA SINICA Volume 26, Number 1, March 1998
S O M E C O N D I T I O N S FOR A C O M P L E X S T R U C T U R E BY
CHUAN-CHIH HSIUNG ( J t B ^ j ) The purpose of this note is to give a sufficient condition and therefrom a necessary condition for an almost complex structure to be a complex structure. The results are contained in the following theorem. Theorem. Let J be an almost complex structure on a Riemannian 2n
2n-
3
manifold M
(n > 2). Let Ji and Rhijk denote respectively the components
of the tensor of J and the Riemann curvature tensor of M2n to a Riemannian
with respect
metric gij and a local coordinate system, where all indices
take the values 1 , . . . , 2n, If Ji3 satisfy
(i)
vn.v'-vI2v = o
for all ii, %2, j , where V denotes the covariant derivation with respect to gij, then J is a complex structure and
for all i\, ii, i-s, k, where the repeated indices imply Proof. Let xl,...,x2n M
2n
summation.
be a local coordinate system on the manifold
. Then with respect to this system the torsion tensor T^ of the almost
complex structure Ji3 is given by {6)
l
a-
J
i
[dxs
dxi
)
J
* \dxs
dxj
J>
Received by the editors June 12, 1995 and in revised form April 23, 1996. 1980 Mathematics Subject Classification (1985) Revision. Primary 53B20, 53C15, 32C10. 39
40
CHUAN-CHIH HSIUNG
[March
which can be written, in term of the covariant derivation V with respect to the Riemannian metric gij, as (4)
~ V f j / ) ~ JiB(V.Jjh
2 * - J^iVsJi*
~ V, Jsh).
From (1) it follows that T^- = 0, so that the almost complex structure J is integrable, and the Newlander and Nirenberg's theorem [1] shows that J is a complex structure. On the other hand, using the Ricci identity we obtain (5)
(o)
{V i1 V i2 — V i 2 V il)Ji3
~{vi3vi2
{' )
— V i 2 V i3)Jil
\V i3v
= J{3
H ji2ii
= ~Jix
^
ii ~ V j j V j 3 j J i 2
Ji2
it
~ J 3 "• »3»2*i>
ji2i3
**• l l i 2 J 3 '
~"~ ^ j
jiji3 ~ J j
it
i2iii3 •
Adding (5), (6), (7) together gives Vll(Vl2Jl3fc-Vl3J,2fc) + Vt2(Vl3Jllfc-V11Jl3fc) + Vl3(VllJl2fc-Vt2JMfc) (
8
)
•
u
•
u
•
u
J. J Rfe .. . 4 . / . 3 P fc .. . 4 . T. J P « .. .
— _ J
i L
*l
.7*2*3
'
J
- , L .7*1*3 ~
*2
U
7k(Rj-
4 T
Jj
\il
l
l3
l
• • *i*2*3
J*2*l
RJ
J l
*2*1*3
Nj
u
By (1) and RJ'. J L
. . 4. p i . . . 4. p i . . . — 0
21*2*3
'
J L
*2*3*1
'
J t
*3*1*2
—
U
'
(8) is reduced to (9") V1^/
/•
j
^*1
Rk •• • 4 - /•
J
t
.7*2*3 ^
°T-2
j
Ftk • • • 4 - 7- J R f c • • • — n
1 L
.7*3*1
'
J
*3
J L
,7*1*2
—
U
'
Multiplying (9) by gki and summing for k we have V-LUj
Ji1
ftiji2i3
~t~ t / i 2 ltiji3ii
~*~ ^*3 -^*j*i*2 — ^ '
Multiplying (10) by J i / and summing for i yield
• • ) *3*2*1/'
1998]
SOME CONDITIONS FOR A COMPLEX STRUCTURE
41
By changing ix, k, i2 to i2, i\, k; k, i2, i$ to i 2 , 13, k; and iu k, i3 to fc, i 3 , ii, respectively, from (11) we obtain
J %2 ^i\
H-ijkis
1 Jk
"k
^-iji2ii
' "12
Ji$
Ji\
^iji^i2
i-3 **
~r ^23 Ji\
1 Jii
Ji$
^*-iji2k
=
^ijki2
— ^'
~ U,
Adding the above three equations together and making use of (1) we immediately arrive at (2), and the proof of the theorem is complete. Referee's remark.
Condition (1) is by no means necessary for an
almost complex structure J to be integable; in fact, for n = 2 if the metric g = (gij) and the almost complex structure J are compatible, i.e., if g(u,v) — g(Ju,Jv)
for any two tangent vectors u and v, then condition (1) implies
that the metric g is flat. References 1. A. Newlander & L. Nirenberg, Complex analytic manifolds, Ann. of Math. 6 5 (1957), 391-404.
coordinates
in almost
Department of Mathematics, Lehigh University, Bethlehem, P a 18015, U.S.A.
complex
673 N O N E X I S T E N C E OF A COMPLEX
STRUCTURE
O N T H E S I X - S P H E R E . II
CHUAN-CHIH
HSIUNG
ABSTRACT. The purpose of this paper is to revise my solution to the long standing unsolved problem "does there exist a complex structure on the six- sphere?"
1. I n t r o d u c t i o n T h e notion of a complex manifold is a n a t u r a l outgrowth of t h a t of a diiferentiable manifold. Its importance lies t o a large extent in t h e fact t h a t t h e complex manifolds include t h e complex algebraic varieties and t h e Riemann surfaces as special cases, and furnish t h e geometric basis for functions of several complex variables. Its development h a s led to clarifications of classical algebraic geometry and to newresults a n d problems. A fundamental problem in the theory of complex manifolds is t o characterize the orientable manifolds of even dimension 2n which can be given a complex structure. In order t o solve this fundamental problem, it is n a t u r a l to first check all t h e spheres of even dimensions. By using characteristic classes and cohomology operations on the classes, in particular, t h e Steenrod squaring a n d reduced power operations, various necessary conditions are known for a manifold to be almost complex a n d therefore to be complex since a complex s t r u c t u r e is naturally an almost complex structure. These conditions suffice to show t h a t among the even-dimensional spheres only t h e 2sphere S 2 a n d the 6-sphere S 6 are almost complex. In fact, t h e absence of an almost complex structure on Sik for k > 1 a n d S 2 n for n > 4 was proved by W u [12] a n d jointly by Borel and'Serre [2] respectively. Kirchoff [7], [8] has shown t h a t if Sn a d m i t s an almost complex structure, t h e Sn+l admits an absolute parallelism, and A d a m s [1] t h a t Sn+l admits an absolute parallelism only for n + 1 = 1,3 a n d 7. T h e result of Adams combined with t h a t of Kirchhoff implies t h e results of W u , Borel a n d Serre. T h u s "does there exist a complex structure on S 6 " has been a long standing unsolved problem for at least forty years. In 1986 I solved this problem [5] b u t missed some conditions. For completeness I a m revising my former work with an addition of the missing conditions. Now our results can be expressed as t h e following theorem and corollaries. T h e o r e m 1. If an almost complex structure J on a Riemannian 2n-manifoId M 2 n (n > 2) without a flat metric and a metric of nonzero constant sectional curvature is a complex structure, then (i)
JiiJilRin*
+ JiiJjRiHik
+ JdJitRiji,*
= o
674
for all ii,i2,i3,k, where J^ and Rhijk are respectively the components of J and the Riemann curvature tensor of M2n, all indices take the values 1, ...,2n, and the repeated indices imply summation. C o r o l l a r y 1. There does not exist a complex manifold M2n (n > 2) with a constant nonzero a flat metric. C o r o l l a r y 2. There does not exist a complex n>2.
structure sectional
structure
on a Riemarmian 2ncurvature but without
on the 2n-sphere
S2n
for
R e m a r k . It might be t r u e t h a t there are various necessary conditions for an almost complex structure J on a Riemannian manifold M to be a complex structure. Equation (1) is a special condition which involves the metrics of M. Since a complex structure is independent of the metrics, (1) cannot serve as such a condition unless it can hold for all the metrics on M. So it is n a t u r a l to expect to impose some metric conditions on M. 2. A l m o s t C o m p l e x S t r u c t u r e s Let Mm be a differentiate m-manifold of class C°° with a R i e m a n n metric gij. On Mm if there exists a tensor J ; J of type (1,1) satisfying (2)
j^j*
= -e\
(,-,_,•,*,= l , . . . , m ) ,
where the repeated index j implies summation, and £ ? =
fl, [ 0,
i=k, _ otherwise
t h e n according to C. E h r e s m a n n [4] J^ defines an almost complex structure on Mm, and Mm is called an almost complex manifold. From (1) it follows t h a t the tensor J^ induces an automorphism J of the tangent space M™ of Mm at each point x with J 2 = —J, / being the identity operator: J : M™ -> M™, such that for each tangent vector v ,
J(vk)=
^V.
By using the multiplication of matrices, from (1) we readily see t h a t a necessary condition for the existence of an almost complex structure on a manifold Mm is t h a t the dimension m of Mm b e even. Let M2n be a 2n-manifold with an almost complex structure Ji (x), where the point x has real local coordinates xh,i,h = 1,..., n, 1,..., n. If there exists a system of complex coordinate neighborhood (z K ) given by
(3)
z K =xK
+T/~^IXK,
675 with respect to which the tensor J,- h has the numerical components
W
(jk.fV^lel
4
0
J }
()
^« - V 0
- v ^ ef
where K , K , A r u n over the range l , . . . , n , then we say t h a t the almost complex s t r u c t u r e Ji h is induced from the complex structure defined by t h e existence of such a system of complex coordinate neighborhoods. Moreover, if a n almost complex s t r u c t u r e is induced from a complex structure, then the almost complex s t r u c t u r e is called integrable. T h e following two theorems give two important characterizations of integrable almost complex structures. T h e o r e m 2 . An almost
complex
structure
is integrable
d2 = 0 or, equivalently
(5)
where d and d are two conjugate structure.
if and only if
o
d — 0,
complex operators
defined by an almost
complex
For the definitions of d a n d d see, for instance [11, Chapter IX], a n d for the proof of T h e o r e m 2 see, for instance, [10, Chapter 4, p . 12] or [3, p p . 18, 19]. T h e o r e m 3 . (Newlander and Nirenberg ture is a complex structure.
[9]). An integrable
almost complex
struc-
3. Proofs (i) Theorem 1. O u r m e t h o d is to apply Theorems 2 a n d 3. (5) for any almost complex structure J holding for any form gives a necessary conditions for J to be a complex structure. It is interesting to know t h a t if we apply (5) to any 0-form, then (5) easily becomes (see for instance, [6, p p . 255-256])
(6) W
T-k = J > t f _ dJl) _ JMMI ,]
~J>
{
dx*
dxt
]
'
[
cb*
dxi
-°l£)-0
]
~
'
where ( x 1 , . . . , x 2 n ) are real local coordinates on the manifold M 2 n , and T,-* is the well-known torsion tensor. However, if we apply d2 to the 1-form dxh, where k is arbitrary, after an extremely lengthy calculation in which several hundred or even u p to a thousand t e r m s are involved, we can prove (see, for instance, [6, p p . 255-265]) d2 = 0 to be (1). To prove the necessity of the metric conditions on the manifold M2n, let (7)
J'j=g'kJk3.
Ji3=gjkJi\
Then •Jij T1
•*jit
676
since otherwise, by (2), - 2 n = JijJji
= J{jJij
a contradiction. T h u s on the manifold M2n tensor field of type (0,2): (8)
there is a nonvanishing skew symmetric
Hij = Jij -
Now suppose t h a t the manifold M2n respect to t h e metric gij. T h e n (9)
> 0,
Jji.
is of constant sectional curvature K with
Rhijk = K(ghkgij
— ghj9ik),
which, together with (7) a n d (8), reduces (1) to (10)
K(Jilk^i3
Multiplying (10) by Jk'1
+ Ji,k{iail
+ Ji3kCm2)
= 0.
a n d using (2) we obtain (n - l)KZi,it
= 0,
which holds for n > 2 if a n d only if K = 0 since £; 2 i 3 ^ 0. T h u s J{3 satisfies (1) or not according as K is zero or not. This is a contradiction if the manifold M2n a d m i t s the two types of metrics. But this contradiction will not occur, if the manifold M2n admits only one or neither of t h e two types of metrics. Hence the proof of the theorem is completed. (ii) Corollary 1. From the last part of the above proof of Theorem 1, an almost s t r u c t u r e J^ on M2n will not satisfy (1), and therefore cannot be a complex structure. (iii) Corollary 2. A 2n-sphere S2n does not admit aflat metric since otherwise its Euler-Poincare characteristic is zero by the Gauss-Bonnet formula. Moreover, S2n is of constant positive sectional curvature by the s t a n d a r d imbedding in a Euclidean (2n + l)-space E2n+1. Hence Corollary 2 follows from Corollary 1 immediately. REFERENCES
[1] J . F . Adams, On the non-existence of elements of Hopf invariant one, Ann. of Math. 72 (1960), 20-104. [2] A. Borel and J.P. Serre, Groupes de Lie et puissances reduites de Steenrod, Amer. J. Math. 75 (1953), 409-448. [3] S.S. Chern, Complex manifolds without potential theory, Van Nostrand, Princeton, 1967. [4] C. Ehresmann, Sur la theorie des espaces fibres, Coll. Int. C.N.R.S. Topologie Algebrique, Paris (1947), 3-35. [5] C.C. Hsiung,, Nonexistence of a complex structure on the six-wphere, Bull. Inst. Math. Acad. Sinica, Taiwain 14 (1986), 231-247. [6] , Almost complex and complex structures, World Scientific, Singapore, 1995. [7] A. Kirchhoff,, Sur Vexistence de certains champs tensoriels sur les spheres a dimensions, C.R. Acad. Sci. Paris 225 (1947), 1253-1260. [8] , Beitrage zur topologischen linearen algebra, Compositio Math. 11 (1953), 1-36.
677 [9] A. Newlander and L. Nirenberg, Complex analystic coordinates in almost complex manifolds, Ann. of Math. 65 (1957), 391-404. [10] H.K. Nikerson and D.C. Spencer, Notes on differeniiable manifolds and sheaves, Mimeographed notes, Princeton Univ., Princeton, 1953-1955. [11] M. Schiffer and D.C. Spencer, Functionals of finite Riemann surfaces, Princeton Univ. Press, Princeton, 1954. [12] W . T . Wu, Sur les classes caracteristiques des structures fibrees spheriques, No. 1183, Actualites Sci. Indust., Herman, Paris, 1952. L E H I G H U N I V E R S I T Y , B E T H L E H E M , PENNSYLVANIA 18015, U.S.A.
E-mail:
cch2Qlehigh.edu
679
REMARKS ON SOME SELECTED PAPERS
1. *[4] solves the three cases k = 2,3,4 of an open problem mentioned in E. P. Lane's book (Projective Differential Geometry of Curves and Surfaces, University of Chicago Press, 1932, Ex. 19, pp. 30-31). 2. To my knowledge our theorem in [8] is the first pure mathematical theorem of a tangram (see, for instance, Martin, Gardner, Mathematical Puzzle & Divisions, Simon and Shuster, New York, 1961, p. 214). 3. In [10], [12], [13], [15], [16], [18], [19], [27], projective invariants of pairs of curves and surfaces at an ordinary common point in spaces of dimension n > 2 are given with geometric interpretations in terms of cross ratios. 4. [23] is of the same nature as those above, but there are triplets of plane curvilinear elements with a common singular point. 5. [33] gives a set of integral formulas, for an orientable hypersurface Vn with a closed boundary Vn~1 of dimension n — 1 in a Euclidean space En+1 of n + 1 > 3 dimensions, and deduces conditions for Vn to be a hypersphere. 6. [37] is an extension of H. Hopf and K. Voss, Ein Satz aus der Flachentheorie im Grossen, Archiv der Math. 3 (1952), 187-192. 7. [39], [45], [47] extend the vanishing theorems of Bochner and Lichnerowicz to manifolds with boundary. 8. Many years ago the following conjecture was popular: A compact Riemannian manifold with constant scalar curvature admitting a one-parameter group of nonisometric conformal transformations is isometric to a sphere. [51], [55], [57], [59], [65], [69], [74] and many other people's papers failed to prove the conjecture. In 1981, N. Ejiri showed that the conjecture without additional conditions is not true, (N. Ejiri, A negative answer to a conjecture of conformal transformations of Riemannian Manifolds, J. Math. Soc. Japan, 33 (1981) 261-266). 9. [68], published in 1972, showed that Hirzebruch signature theorem is also true for a manifold with a boundary which admits a free orientationreversing involution, such boundary is called a reflecting boundary. In 1975 and 1976, M. F. Atiyah, V. M. Patodi and I. M. Singer proved this for a general boundary in their joined papers (Spectral, symmetry and Riemannian geometry, I, II, III, Math. Proc. Cambridge Philos. Soc, 77 (1975) 43-69, 78 (1975) no. 3, 405-432, 79 (1976) no. 1, 71-99).
680
10. [75] shows that if two compact oriented manifolds with reflecting boundaries are cobordant, then they have equal Pontrjagin numbers; this becomes Thorn's theorem for manifolds without boundary. 11. [70] is concerned with some extensions of the theorem of W. Frenhel, J. Milnor and I. Fary on the total absolute curvature of closed curves in Euclidean space. We obtain a theorem for closed curves in a complete simply connected Riemannian n-manifold with nonpositive sectional curvature, and this leads to more precise results when the curvature is constant. 12. Let Mm, M be two m-dimensional compact oriented hypersurfaces of class C 3 immersed in a Riemannian manifold Rm+1 of constant sectional curvature. Suppose that Rm+1 admits a one-parameter continuous group G of conformal transformations satisfying a certain condition (which holds automatically when G is a group of isometric transformations). Suppose further that there is a 1 — 1 transformation TT £ G between Mm and M™ such that P = TT(P)P for each P € M and each P G M . If the r-th mean curvature for any r, 1 < r < m, of Mm at each point P 6 Mm is equal to that of M at the corresponding point P = TT(P)P, together with other conditions, in [73] we then prove that Mm and M are congruent, mod G. This is a generalization of a joint theorem of H. Hopf and Y. Katsurada in which G is a group of isometric transformations. 13. Let Mm be a submanifold of dimension m(> 2) immersed in a Riemannian n-manifold Nn(n > m). Suppose that Nn admits a continuous infinitesimal conformal vector field £, and let e be the unit normal vector field over Mm parallel in the normal bundle of Mm. Then in [77] we derive a set of integral formulas (which are extensions of those in [33]) for a closed oriented Mm in Nn with respect to £ and e, and obtain various conditions for Mm to be unbilical with respect to e. 14. [83], [91], [92] together prove the nonexistence of complex structure on the six-sphere. 15. The following long standing conjecture of H. Hopf is well known. Let M be a compact orientable Riemannian manifold of even dimension n > 2. If M has nonnegative sectional curvature, then the Euler-Poincare characteristic x(M) is nonnegative. If M has nonpositive sectional curvature, then x(M) is nonnegative or nonpositive according as n = 0 or 2 or mod 4. This conjecture for n = 4 was proved by J. W. Milnor and then by S. S. Chern by a different method. In [84] we prove this conjecture for a general n under an extra condition on higher order sectional curvature, which holds automatically for n = 4. Similar results are obtained for Kahler manifolds by using holomorphic sectional curvature,
681 and F. Schur's theorem about the constancy of sectional curvature on a Riemannian manifold is extended. 16. Let (Mj,gi) be an almost 2n-manifold Mj with a Hermitian metric
683
CURRICULUM VITAE
Birth Place: Shin-Gien, Jiangsi, China Birth Date: February 15, 1916 Parents: Father's Name: Mu Hun Hsiung Mother's Name: Tu Shih Hsiung Marriage: Wife: Wenchin Yu Child: Nancy Hsiung
Marriage Date: July 10, 1942
Education: B.S., National Chekiang University, Hangchow, China, 1932-1936 Ph.D., Michigan State University, East Lansing, MI, 1948 Career History: Instructor, University of Wisconsin, Madison, WI, 1948-1950 Lecturer, Northwestern University, Evanston, IL Fall, 1950 Research Fellow, Harvard University, Cambridge, MA, 1951-1952 Assistant Professor, Lehigh University, Bethlehem, PA, 1952-1955 Associate Professor, Lehigh University, Bethlehem, PA, 1955-1960 Visiting Associate Professor, Mathematics Research Center, U.S. Army at University of Wisconsin, Madison, WI, 1959-1960 Professor, Lehigh University, Bethlehem, PA, 1960-1984, Emeritus Visiting Specialist, University of California, Berkeley, CA, Spring, 1962 Visiting lecturer, Fudan University and Wuhan University, People's Republic of China, Spring, 1980 Visiting Professor, University of Granada, Granada, Spain, JanuaryMay, 1986 Professional Activities: Founder and editor-in-chief of the Journal of Differential Geometry, the unique international journal in the field, 1967Editor of the Bulletin of the Institute of Mathematics, Academia Sinica, Taiwan, 1972The invited organizer of a special session on differential geometry at the meeting of the American Mathematical Society on April 17-18, 1980 in Philadelphia. An editorial advisor and editor-in-chief of the Series in Pure Mathematics of the World Scientific Publishing Co., Singapore, 1982 An invited speaker at many international major conferences including:
i) The Summer Institute, sponsored by the American Mathematical Society and the National Science Foundation (NSF), on Differential Geometry in the Large, at the University of Washington (Seattle) in 1956; (ii) The Summer Institute on Relativity and Differential Geometry, at the University of California (Santa Barbara) in 1962; (iii) The Conference on Differential Geometry in the Large, sponsored by the National Mathematics Research Institute of West Germany, at Oberwolfach, West Germany, in 1964 and 1972; (iv) The International Congresses of Mathematicians in Stockholm, Sweden, in 1962, in Moscow in 1966, and in Vancouver, Canada, in 1974; (v) The Regional Conference on Differential Geometry sponsored by NSF at Michigan State University in 1970; (vi) The 13th Biennial Seminar of the Canadian Mathematical Congress at Delhousie University, Halifax, Nova Scotia, in 1971; (vii) The International Global Analysis Symposium held at the University of Warwick, England, Spring, 1972; viii) The Summer Institute on Differential Geometry sponsored by the American Mathematical Society and NSF at Stanford University in 1973.
685
B I B L I O G R A P H Y OF T H E P U B L I C A T I O N S OF C H U A N - C H I H H S I U N G
I.
Books, Lecture Notes and Related Publications
1. A First Course in Differential Geometry, Wiley-Interscience Series of Texts, Monographs and Tracts, Wiley, New York, 1981, 360 pp; republished, International Press, Cambridge, MA, 1997, 360 pp. 2. Almost Complex and Complex Structures, World Scientific, Singapore, 1995, 310 pp. 3. Topics on Differential Geometry, Lecture Notes, Academia Sinica, National Taiwan Univ. & National Tsinghua Univ., 1969, 87 pp. 4. Mathematical Essays (in Honor of Professor Su Buchin's 80th birthday), Editor, World Scientific, Singapore, 1983, 278 pp. 5. (with James Foster) Differential geometry, Handbook of Applicable Math., Wiley, New York, Vol. V, 1985, Chapter 12, 423-467. 6. Convex sets, Encyclopedia Phys. Sci. Tech., Academic Press, New York, 1987, 675-689. 7. My associations with S. S. Chern, S. S. Chern, A great geometor of the Twentieth century (expanded edition), International Press, MA, 1998, 320-325.
II.
Papers 1937
[1] On the related conies, Sci. Reports, University of Chekiang, 2 (1937) 171-178. 1940 [2] On the plane sections of the tangent surface of a space curve, J. Chinese Math. Soc. 2 (1940) 239-245. [3] Note on the intersection of two space curves, Tohoku Math. J. 47 (1940) 201-209. [4] Sopra il contatto di due curve piane, Boll. Un. Nat. Ital. (2), 2 (1940) 443-451. [5] On the curvature form and the projective curvature of curves in space of four dimensions, Rev. Mat. Fis. Teor. Univ. Nac. Tucuman, (A), 1 (1940) 159-171.
686
1941 [6] A graphical construction of the sphere osculating a space curve, Tohoku Math. J. 48 (1941) 272-276. [7] The canonical lines, Duke Math. J. 8 (1941) 738-742. 1942 [8] (with F. T. Wang) A theorem on the tangram, Amer. Math. Monthly 49 (1942) 596-599. 1943 [9] Asymptotic ruled surfaces, Duke Math. J. 10 (1943) 217-237. [10] Projective differential geometry of a pair of plane curves, Duke Math. J. 10 (1943) 539-546. [11] Theory of intersection of two plane curves, Bull. Amer. Math. Soc. 49 (1943) 786-792. [12] An invariant of intersection of two surfaces, Bull. Amer. Math. Soc. 49 (1943) 877-880. [13] Projective invariants of a pair of surfaces, Duke Math. J. 10 (1943) 717-720. 1944 [14] Plane sections of certain ruled surfaces associated with a curved surface, Duke Math. J. 11 (1944) 59-64. [15] Projective invariants of intersection of certain pairs of surfaces, Bull. Amer. Math. Soc. 50 (1944) 437-441. 1945 [16] New geometrical characterizations of some special conjugate nets, Duke Math. J. 12 (1945) 249-253. [17] Some invariants of certain pairs of hypersurfaces, Bull. Amer. Math. Soc. 51 (1945) 572-582. [18] A projective invariant of a certain pair of surfaces, Duke Math. J. 12 (1945) 441-443. 1946 [19] Projective invariants of contact of two curves in space of n dimensions, Quart. J. Math., Oxford Ser. 17 (1946) 39-45.
687
1947 [20] Plane sections of the tangent surfaces of two space curves, Duke 'Math. J. 14 (1947) 151-158. [21] A study on the theory of conjugate nets, Amer. J. Math. 69 (1947) 379-390. [22] Projective theory of surfaces and conjugate nets in four-dimensional space, Amer. J. Math. 69 (1947) 607-621. [23] On triplets of plane curvilinear elements with a common singular point, Quart. J. Math., Oxford Ser. 18 (1947) 129-132. 1948 [24] Differential geometry of a surface at a parabolic point, Amer. J. Math. 70 (1948) 333-344. 1949 [25] Invariants of intersection of certain pairs of space curves, Bull. Amer. Math. Soc. 55 (1949) 623-628. [26] Rectilinear congruences, Trans. Amer. Math. Soc. 66 (1949) 419-439. [27] Invariants of intersection of certain pairs of curves in n-dimensional space, Amer. J. Math. 71 (1949) 678-686. [28] Affine invariants of a pair of hypersurfaces, Amer. J. Math. 71 (1949) 879-882. 1950 [29] A note of correction, Proc. Amer. Math. Soc. 1 (1950) 824-825. 1951 [30] Conjugate nets in three- and four-dimensional spaces, Duke Math. J. 18 (1951) 487-499. [31] A general theory of conjugate nets in projective hyperspace, Trans. Amer. Math. Soc. 70 (1951) 312-322. 1952 [32] Some curves in Riemannian space, Acad. Roy. Belg. Bull. CI. Sci. ser. 5, 38 (1952) 816-823. 1954 [33] Some integral formulas for closed hypersurfaces, Math. Scand. 2 (1954) 286-294.
688
1956 [34] A theorem on surfaces with a closed boundary, Math. Z. 64 (1956) 41-46. [35] On differential geometry of hypersurfaces in the large, Trans. Amer. Math. Soc. 81 (1956) 243-252. [36] Some integral formulas for closed hypersurfaces in Riemannian space, Pacific J. Math. 6 (1956) 291-299. 1957 [37] Some global theorems on hypersurfaces, Canad. J. Math. 9 (1957) 5-14. 1958 [38] A uniqueness theorem for Minkowski's problem for convex surfaces with boundary, Illinois J. Math. 2 (1958) 71-75. [39] Curvature and Betti numbers of compact Riemannian manifolds with boundary, Univ. Politec. Torino. Rend. Sem. Math. 17 (1958) 95-131. [40] A uniqueness theorem on two-dimensional Riemannian manifolds with boundary, Michigan Math. J. 5 (1958) 25-30. 1959 [41] (with G. F. Feeman) Characterizations of Riemann n-spheres, Amer. J. Math. 81 (1958) 691-708. 1960 [42] (with S. S. Chern and J. Hano) A uniqueness theorem on closed convex hypersurfaces in Euclidean space, J. Math. Mech. 9 (1960) 85-88. [43] Some uniqueness theorems on Riemannian manifolds with boundary, Illinois J. Math. 4 (1960) 526-540. 1961 [44] Isoperimetric inequalities for two-dimensional Riemannian manifolds with boundary, Ann. of Math. 73 (1961) 213-220. 1962 [45] A note of correction, Univ. e. Politec. Torino, Rend. Sem. Mat. (1962) 127-129.
21
1963 [46] (with S. S. Chern) On the isometry of compact submanifolds in Euclidean space, Math. Ann. 149 (1963) 278-285.
689 [47] Curvature and homology of Riemannian manifolds with boundary, Math. Z. 82 (1963) 67-81. 1964 [48] Vector fields and infinitesimal transformations on Riemannian manifolds with boundary, Bull. Soc. Math. France 92 (1964) 411-434. [49] On the congruence of hypersurfaces, Atti Accad. Naz. Lincei. Rend. CI. Sci. Fis. Mat. Nat, ser. 8, 37 (1964) 259-266. 1965 [50] (with A. L. Hilt) Vector fields and infinitesimal transformations on almost-Hermitian manifolds with boundary, Canad. J. Math. 17 (1965) 213-238. [51] On the group of conformal transformations of a compact Riemannian manifold, Proc. Nat. Acad. Sci. U.S.A. 54 (1965) 1509-1513. 1966 [52] Structures and operators on almost-Hermitian manifolds, Trans. Araer. Math. Soc. 122 (1966) 136-152. [53] (with J. I. Nassar) Parallel and central transformations of Riemannian manifolds, Rend. Circ. Mat. Palermo, ser. II, 5 (1966) 283-309. 1967 [54] (with J. K. Shahin) Affine differential geometry of closed hypersurfaces, Proc. London Math. Soc, ser. 3, 7 (1967) 715-735. [55] On the group of conformal transformations of a compact Riemannian manifold. II, Duke Math. J. 34 (1967) 337-341. [56] (with Y. K. Cheung) Curvature and characteristic classes of compact Riemannian manifolds, J. Differential Geom. 1 (1967) 89-97. 1968 [57] (with J. D. Liu) The group of conformal transformations of a compact Riemannian manifold, Math. Z. 105 (1968) 307-312. [58] (with B. H. Rhodes) Isometries of compact submanifolds of a Riemannian manifold, J. Differential Geom. 2 (1968) 9-24. [59] On the group of conformal transformations of a compact Riemannian manifold. Ill, J. Differential Geom. 2 (1968) 185-190. 1970 [60] (with S. Braidi) Submanifolds of spheres, Math. Z. 115 (1970) 235-251.
690 [61] Minimal immersions in Riemannian spheres, Studies and Essays, (Presented to Y. W. Chen on his Sixtieth Birthday), Math. Res. Center, Nat. Taiwan Univ., Taipei (1970) 223-229. 1971 [62] (with L. R. Mugridge) Conformal changes of metrics on a Riemannian manifold, Math. Z. 119 (1971) 179-187. [63] (with J. J. Levko III) Curvature and characteristic classes of compact pseudo-Riemannian manifolds, Rocky Mountain J. Math. 1 (1971) 523536. [64] (with J. J. Levko III) Complex Laplacians on almost-Hermitian manifolds, J. Differential Geom. 5 (1971) 383-403. 1972 [65] (with L. W. Stern) Conformality and isometry of Riemannian manifolds to spheres, Trans. Amer. Math. Soc. 163 (1972) 65-73. [66] (with S. S. Mittra) Isometries of compact hypersurfaces with boundary in a Riemannian space, J. Differential Geometry in honor of K. Yano, Kinokuniya, Tokyo, 1972, 145-161. [67] (with L. R. Mugridge) Riemannian manifolds admitting certain conformal changes of metric, Colloq. Mat. 26 (1972) 135-143. [68] The signature and G-signature of manifolds with boundary, J. Differential Geom. 6 (1972) 595-598. 1974 [69] (with L. L. Ackler) Isometry of Riemannian manifolds to spheres, Ann. Mat. Pura Appi, ser. 4, 99 (1974) 53-64. [70] (with F. Brickell) The total absolute curvature of closed curves in Riemannian manifolds, J. Differential Geom. 9 (1974) 177-193. 1975 [71] A remark on pinched manifolds with boundary, Ann. Mat. Pura Appi, ser. 4, 102 (1975) 103-107. [72] The generalized Poincare conjecture on higher dimensional manifolds with boundary, Bull. Inst. Math., Acad. Sinica, Taiwan, 3 (1975) 177181. 1976 [73] (with T. P. Lo) Congruence theorems for compact hypersurfaces of a Riemannian manifold, Ann. Mat. Pura Appi. 109 (1976) 289-304.
691 [74] (with N. H. Ackerman) Isometry of Riemannian manifolds to spheres. II, Canad. J. Math. 28 (1976) 63-72. [75] A remark on cobordism of manifolds with boundary, Arch. Math. Basel 27 (1976) 551-555. 1977 [76] (with L. R. Mugridge) Conformal invariants of submanifolds, Proc. Amer. Math. Soc. 62 (1977) 316-318. [77] (with J. D. Liu and S. S. Mittra) Integral formulas for closed submanifolds of a Riemannian manifold, J. Differential Geom. 12 (1977) 133151. [78] (with J. D. Liu) A generalization of the rigidity theorem of Cohn-Vossen, J. London Math. Soc. 15 (1977) 557-565. 1979 [79] (with L. R. Mugridge) Euclidean and conformal invariants of submanifolds, Geom. Dedicata. 8 (1979) 31-38. 1981 [80] (with Ann D. Bingham) A generalization of Hilbert's theorem for the nonexistence of isometric imbeddings, 1981, unpublished. [81] Isometry of compact Riemannian manifolds to spheres, Proc. Conf. Differential Equations in Analysis and Geometry, Supplement Amer. J. Math. (1981) 163-167. 1982 [82] (with K. S. park) Some uniqueness theorems on two-dimensional Riemannian manifolds immersed in a general Euclidean space, Geom. Dedicata, 12 (1982) 35-51. 1986 [83] Nonexistence of a complex structure on the six-sphere, Bull. Math., Acad. Sinica, Taiwan, 14 (1986) 231-247.
Inst.
1988 [84] (with K. M. Shiskowski) Euler-Poincare characteristic and higher order sectional curvature. I, Trans. Amer. Math. Soc. 305 (1988) 113-128. [85] (with J. J. Levko) Conformal invariants of submanifold. II, Indiana Univ. Math. J. 37 (1988) 181-189.
692 1989 [86] (with L. Friedland) A certain class of almost Hermitian manifolds, Tensor 48 (1989) 252-263. 1995 [87] (with B. Xiong) A new class of almost complex structures, Ann. Mat. Pura Appd. 168 (1995) 133-149. [88] (with C. X. Wu) The spectral geometry of almost L manifolds, Bull. Inst. Math. Acad. Sinica, Taiwan, 23 (1995) 229-241. [89] (with W. Yang and L. Friedland) Holomorphic sectional and bisectional curvatures of almost Hermitian manifolds, SUT J. Math. 31 (1995) 138-154. 1996 [90] (with W. Yang and B. Xiong) The spectral geometry of some almost Hermitian manifolds, SUT J. Math. 32 (1996) 163-178. 1998 [91] Some conditions for a complex structure, Bull. Inst. Math. Acad. Sinica, Taiwan, 26 (1998) 39-41. [92] Nonexistence of a complex structure on the six-sphere. II, to be published.
693
LIST OF P H . D . T H E S E S W R I T T E N U N D E R THE SUPERVISION OF CHUAN-CHIH HSIUNG
1. George F. Feeman, Characterizations of Riemann n-sphere, October, 1958. 2. Hanna, Id Nassar, Parallel and central transformations of Riemannian manifolds, June, 1961. 3. Arthur L. Hilt, Vector fields and infinitesimal transformations of almostHermitian manifolds with boundary, June, 1961. 4. Jamal K. Shahin, Affine differential geometry of closed hypersurfaces, October, 1965. 5. Burgess H. Rhodes, Isometries of compact submanifolds of a Riemannian manifold, June, 1967. 6. Yuk K. Cheung, Curvature and characteristic classes of compact Riemannian manifolds, June, 1967. 7. Jong D. Liu, The group of conformal transformations of a compact Riemannian manifold, June, 1967. 8. Larry R. Mugridge, Conformal changes of metrics on a Riemannian manifold, June, 1968. 9. Siham Braidi, Submanifolds of spheres, June, 1969. 10. John J. Levko III, Complex Laplacians on almost-Hermitian manifolds, October, 1969. 11. Louis W. Stern, Conformality and isometry of Riemannian manifolds to spheres, June, 1970. 12. Sitansu S. Mittra, Isometries of compact hypersurfaces with boundary in a Riemannian space, June, 1971. 13. Lynn L. Ackler, Isometry of Riemannian manifolds to spheres, October, 1971. 14. Timothy P. Lo, Congruence theorems for compact hypersurfaces of a Riemannian manifold, October, 1973. 15. Neill H. Ackerman, Isometry of Riemannian manifolds to spheres. II, October, 1974. 16. Ann Bingham, A generalization of Hilbert's theorem for the nonexistence of isometric imbeddings, June, 1979. 17. Kyu S. Park, Some uniqueness theorems on two-dimensional Riemannian manifolds immersed in a general Euclidean space, October, 1979. 18. Kenneth M. Shiskowski, Euler-Poincare characteristic and higher order sectional curvature. I, June, 1983. 19. Lew Friedland, A certain class of almost Hermitian manifolds, June, 1984. 20. Zhong C. Xiong, A new class of almost complex structures, June, 1989.
695 Permission [4] Sopra il contatto di due curve piane, reprinted from Boll. Un. Nat. Ital. (2), 2 (1940) 443-451. [8] (with F. T. Wang) A theorem on the tangram, reprinted with permission from Amer. Math. Monthly 49 (1942) 596-599. © 1949 Mathematical Association of America. 10] Projective differential geometry of a pair of plane curves, reprinted with permission from Duke Math. J. 10 (1943) 539-546. © 1943 Duke University Press. 12] An invariant of intersection of two surfaces, reprinted with permission from Bull. Amer. Math. Soc. 49 (1943) 877-880. © 1943 American Mathematical Society. 13] Projective invariants of a pair of surfaces, reprinted with permission from Duke Math. J. 10 (1943) 717-720. © 1943 Duke University Press. 15] Projective invariants of intersection of certain pairs of surfaces, reprinted with permission from Bull. Amer. Math. Soc. 50 (1944) 437-441. © 1944 American Mathematical Society. 16] New geometrical characterizations of some special conjugate nets, reprinted with permission from Duke Math. J. 12 (1945) 249-253. © 1945 Duke University Press. 18] A projective invariant of a certain pair of surfaces, reprinted with permission from Duke Math. J. 12 (1945) 441-443. © 1945 Duke University Press. 19] Projective invariants of contact of two curves in space of n dimensions, reprinted with permission from Quart. J. Math., Oxford Ser. 17 (1946) 39-45. © 1946 Oxford University Press. 23] On triplets of plane curvilinear elements with a common singular point, reprinted with permission from Quart. J. Math., Oxford Ser. 18 (1947) 129-132. © 1947 Oxford University Press. 27] Invariants of intersection of certain pairs of curves in n-dimensional space, reprinted with permission from Amer. J. Math. 71 (1949) 678-686. © 1949 Johns Hopkins University Press. 28] Affine invariants of a pair of hypersurfaces, reprinted with permission from Amer. J. Math. 71 (1949) 879-882. © 1949 Johns Hopkins University Press. 33] Some integral formulas for closed hypersurfaces, reprinted with permission from Math. Scand. 2 (1954) 286-294.
696 [34] A theorem on surfaces with a closed boundary, reprinted with permission from Math. Z. 64 (1956) 41-46. © 1956 Springer-Verlag. [35] On differential geometry of hypersurfaces in the large, reprinted with permission from Trans. Amer. Math. Soc. 81 (1956) 243-252. © 1956 American Mathematical Society. [37] Some global theorems on hypersurfaces, reprinted with permission from Canad. J. Math. 9 (1957) 5-14. © 1957 Canadian Mathematical Society. [38] A uniqueness theorem for Minkowski's problem for convex surfaces with boundary, reprinted with permission from Illinois J. Math. 2 (1958) 71-75. [39] Curvature and Betti numbers of compact Riemannian manifolds with boundary, reprinted with permission from Univ. Politec. Torino. Rend. Sem. Math. 17 (1958) 95-131. [40] A uniqueness theorem on two-dimensional Riemannian manifolds with boundary, reprinted with permission from Michigan Math. J. 5 (1958) 25-30. [42] (with S. S. Chern and J. Hano) A uniqueness theorem on closed convex hypersurfaces in Euclidean space, reprinted with permission from J. Math. Mech. 9 (1960) 85-88. © 1960 The Graduate Institute for Mathematics and Mechanics, Indiana University. [43] Some uniqueness theorems on Riemannian manifolds with boundary, reprinted with permission from Illinois J. Math. 4 (1960) 526-540. [44] Isoperimetric inequalities for two-dimensional Riemannian manifolds with boundary, reprinted with permission from Ann. of Math. 73 (1961) 213-220. [45] A note of correction, reprinted with permission from Univ. e. Politec. Torino, Rend. Sem. Mat. 21 (1962) 127-129. [46] (with S. S. Chern) On the isometry of compact submanifolds in Euclidean space, reprinted with permission from Math. Ann. 149 (1963) 278-285. © 1963 Springer-Verlag. [47] Curvature and homology of Riemannian manifolds with boundary, reprinted with permission from Math. Z. 82 (1963) 67-81. © 1963 Springer-Verlag. [48] Vector fields and infinitesimal transformations on Riemannian manifolds with boundary, reprinted with permission from Bull. Soc. Math. France 92 (1964) 411-434. © 1964 Kluwer Academic Publisher. [49] On the congruence of hypersurfaces, reprinted with permission from Atti Accad. Naz. Lincei. Rend. CI. Sci. Fis. Mat. Nat, ser. 8, 37 (1964) 259-266.
697
[51] On the group of conformal transformations of a compact Riemannian manifold, reprinted with permission from Proc. Nat. Acad. Sci. U.S.A. 54 (1965) 1509-1513. [52] Structures and operators on almost-Hermitian manifolds, reprinted with permission from Trans. Amer. Math. Soc. 122 (1966) 136-152. © 1966 American Mathematical Society. [54] (with J. K. Shahin) Affine differential geometry of closed hypersurfaces, reprinted with permission from Proc. London Math. Soc, ser. 3, 7 (1967) 715-735. © 1967 London Mathematical Society. [55] On the group of conformal transformations of a compact Riemannian manifold. II, reprinted with permission from Duke Math. J. 34 (1967) 337-341. © 1967 Duke University Press. [56] (with Y. K. Cheung) Curvature and characteristic classes of compact Riemannian manifolds, reprinted with permission from J. Differential Geom. 1 (1967) 89-97. [57] (with J. D. Liu) The group of conformal transformations of a compact Riemannian manifold, reprinted with permission from Math. Z. 105 (1968) 307-312. © 1968 Springer-Verlag. [58] (with B. H. Rhodes) Isometries of compact submanifolds of a Riemannian manifold, reprinted with permission from J. Differential Geom. 2 (1968) 9-24. [59] On the group of conformal transformations of a compact Riemannian manifold. Ill, reprinted with permission from J. Differential Geom. 2 (1968) 185-190. [60] (with S. Braidi) Submanifolds of spheres, reprinted with permission from Math. Z. 115 (1970) 235-251. © 1970 Springer-Verlag. [61] Minimal immersions in Riemannian spheres, reprinted with permission from Studies and Essays, (Presented to Y. W. Chen on his Sixtieth Birthday), Math. Res. Center, Nat. Taiwan Univ., Taipei (1970) 223-229. [63] (with J. J. Levko III) Curvature and characteristic classes of compact pseudo-Riemannian manifolds, Rocky Mountain J. Math. 1 (1971) 523-536. © 1971 Rocky Mountain Mathematics Consortium. [64] (with J. J. Levko III) Complex Laplacians on almost-Hermitian manifolds, reprinted with permission from J. Differential Geom. 5 (1971) 383-403. [65] (with L. W. Stern) Conformality and isometry of Riemannian manifolds to spheres, reprinted with permission from Trans. Amer. Math. Soc. 163 (1972) 65-73. © 1972 American Mathematical Society. [66] (with S. S. Mittra) Isometries of compact hypersurfaces with boundary in a Riemannian space, reprinted with permission from J. Differential Geometry in honor of K. Yano, Kinokuniya, Tokyo, 1972, 145-161.
698
[67] (with L. R. Mugridge) Riemannian manifolds admitting certain conformal changes of metric, reprinted with permission from Colloq. Mat. 26 (1972) 135-143. [68] The signature and G-signature of manifolds with boundary, reprinted with permission from J. Differential Geom. 6 (1972) 595-598. [69] (with L. L. Ackler) Isometry of Riemannian manifolds to spheres, reprinted with permission from Ann. Mat. Pura Appl, ser. 4, 99 (1974) 53-64. [70] (with F. Brickell) The total absolute curvature of closed curves in Riemannian manifolds, reprinted with permission from J. Differential Geom. 9 (1974) 177-193. [71] A remark on pinched manifolds with boundary, reprinted with permission from Ann. Mat. Pura Appl, ser. 4, 102 (1975) 103-107. [72] The generalized Poincare conjecture on higher dimensional manifolds with boundary, reprinted with permission from Bull. Inst. Math., Acad. Sinica, Taiwan, 3 (1975) 177-181. [73] (with T. P. Lo) Congruence theorems for compact hypersurfaces of a Riemannian manifold, reprinted with permission from Ann. Mat. Pura Appl. 109 (1976) 289-304. [74] (with N. H. Ackerman) Isometry of Riemannian manifolds to spheres. II, reprinted with permission from Canad. J. Math. 28 (1976) 63-72. © 1976 Canadian Mathematical Society. [75] A remark on cobordism of manifolds with boundary, reprinted with permission from Arch. Math. Basel 27 (1976) 551-555. [77] (with J. D. Liu and S. S. Mittra) Integral formulas for closed submanifolds of a Riemannian manifold, reprinted with permission from J. Differential Geom. 12 (1977) 133-151. [78] (with J. D. Liu) A generalization of the rigidity theorem of CohnVossen, reprinted with permission from J. London Math. Soc. 15 (1977) 557-565. © 1977 London Mathematical Society. [79] (with L. R. Mugridge) Euclidean and conformal invariants of submanifolds, reprinted with permission from Geom. Dedicata. 8 (1979) 31-38. © 1979 by D. Reidel Publishing Co. [82] (with K. S. park) Some uniqueness theorems on two-dimensional Riemannian manifolds immersed in a general Euclidean space, reprinted with permission from Geom. Dedicata, 12 (1982) 35-51. © 1982 by D. Reidel Publishing Co. [83] Nonexistence of a complex structure on the six-sphere, reprinted with permission from Bull. Inst. Math., Acad. Sinica, Taiwan, 14 (1986) 231-247.
699 [84] (with K. M. Shiskowski) Euler-Poincare characteristic and higher order sectional curvature. I, reprinted with permission from Trans. Amer. Math. Soc. 305 (1988) 113-128. © 1988 American Mathematical Society. [85] (with J. J. Levko) Conformal invariants of submanifold. II, reprinted with permission from Indiana Univ. Math. J. 37 (1988) 181-189. [86] (with L. Friedland) A certain class of almost Hermitian manifolds, reprinted from the Tensor (New Series) 48 (1989) 252-263. [87] (with B. Xiong) A new class of almost complex structures, reprinted with permission from Ann. Mat. Pura Appd. 168 (1995) 133-149. [88] (with C. X. Wu) The spectral geometry of almost L manifolds, reprinted with permission from Bull. Inst. Math. Acad. Sinica, Taiwan, 23 (1995) 229-241. [89] (with W. Yang and L. Friedland) Holomorphic sectional and bisectional curvatures of almost Hermitian manifolds, reprinted with permission from SUT J. Math. 31 (1995) 138-154. [90] (with W. Yang and B. Xiong) The spectral geometry of some almost Hermitian manifolds, reprinted with permission from SUT J. Math. 32 (1996) 163-178. [91] Some conditions for a complex structure, reprinted with permission from Bull. Inst. Math. Acad. Sinica, Taiwan, 26 (1998) 39-41.
SELECTED PAPERS OF CHUAN CHIH HSIUNG This invaluable book contains selected papers of Prof Chuan-d Hsiung, renowned mathematician in differential geometry and founder and editor-in-chief of a unique international journal in this field, the Journal of Differential Geometry. During the period of 1935-1943, Prof Hsiung was in China working on projective differential geometry under Prof Buchin Su. In 1946, he went to the United States, where he gradually shifted to global problems. Altogether Prof Hsiung has published about 100 research papers, from which he has selected 64 (in chronological order) for this volume.
ISBN 981-02-4323-5
www. worldscientific. com 4437 he
9 ll789810ll243234"