• Nie Znaleziono Wyników

CHARACTERIZATIONS OF COMPLEX SPACE FORMS BY MEANS OF GEODESIC SPHERES AND TUBES

N/A
N/A
Protected

Academic year: 2021

Share "CHARACTERIZATIONS OF COMPLEX SPACE FORMS BY MEANS OF GEODESIC SPHERES AND TUBES"

Copied!
10
0
0

Pełen tekst

(1)

C O L L O Q U I U M M A T H E M A T I C U M

VOL. 71 1996 NO. 2

CHARACTERIZATIONS OF COMPLEX SPACE FORMS BY MEANS OF GEODESIC SPHERES AND TUBES

BY

J. G I L L A R D (LEUVEN)

We prove that a connected complex space form (M n , g, J ) with n ≥ 4 can be characterized by the Ricci-semi-symmetry condition e R XY · % = 0 and by e the semi-parallel condition e R XY · σ = 0, considering special choices of tan- gent vectors X, Y to small geodesic spheres or geodesic tubes (that is, tubes about geodesics), where e R, % and σ denote the Riemann curvature tensor, e the corresponding Ricci tensor of type (0, 2) and the second fundamental form of the spheres or tubes and where e R XY acts as a derivation.

1. Introduction. In a previous article [1] the following question was stated: which are the Riemannian manifolds all of whose small geodesic spheres or geodesic tubes are semi-symmetric? In fact, one investigated the weaker Ricci-semi-symmetry condition e R XY · % = 0 and also the semi-parallel e condition e R XY · σ = 0 for these hypersurfaces, in view of the strong similari- ties shown in [2], [4] between the intrinsic geometry determined by the Ricci tensor % and the extrinsic properties related to the second fundamental form e σ of the geodesic sphere or tube. The main result was that a connected Rie- mannian manifold (M n , g) with n ≥ 4 is a real space form if and only if its small geodesic spheres are Ricci-semi-symmetric or semi-parallel, where for small geodesic tubes it was sufficient that these conditions are satisfied for the so-called horizontal tangent vectors X, Y to the tube. As a consequence, these properties cannot hold for complex space forms, except when they are flat.

In this paper we look for a special class of tangent vectors X, Y to the tubes or spheres which makes each of the two conditions e R XY · % = 0 and e R e XY · σ = 0 characteristic for complex space forms. It will turn out that the appropriate tangent vectors are the horizontal ones (in the sense of Section 3 and 4), where in the case of geodesic tubes one has additionally to restrict to special points (see Section 2).

1991 Mathematics Subject Classification: 53C20, 53C25, 53C40, 53C55.

[253]

(2)

The author wishes to express his gratitude to Dr. E. Boeckx and Prof.

L. Vanhecke for their support and many useful suggestions.

2. Preliminaries. Let (M, g) be an n-dimensional, connected, smooth Riemannian manifold, with n ≥ 4. Denote by ∇ the Levi-Civita connection and by R and % the corresponding Riemannian curvature tensor and Ricci tensor, respectively. We use the sign convention

R XY = ∇ [X,Y ] − [∇ X , ∇ Y ] for tangent vector fields X, Y on M .

Next, we treat some general aspects of complex space forms. Suppose that (M, g, J ) is a K¨ ahler manifold , that is, J is a (1, 1)-tensor field on M such that

(1) J 2 = −I, g(J X, J Y ) = g(X, Y ), ∇J = 0

for all tangent vector fields X, Y on M . The holomorphic sectional curvature H(u) for a unit tangent vector u ∈ T x M, x ∈ M is the sectional curvature of the plane spanned by {u, J u}. So, H(u) = R uJ uuJ u = g(R uJ u u, J u).

If H(u) is independent of u then it is independent of x, i.e., H(u) = c, c ∈ R and then (M, g, J ) is called a space of constant holomorphic sectional cur- vature c or a complex space form. Further, a K¨ ahler manifold of constant holomorphic sectional curvature c is characterized by the following curvature tensor:

R XY Z = c

4 {g(X, Z)Y − g(Y, Z)X (2)

+ g(J X, Z)J Y − g(J Y, Z)J X + 2g(J X, Y )J Z}.

(See for example [11].) We also have another useful characterization:

Theorem 2.1 [8]. Let (M n , g, J ) be a connected K¨ ahler manifold with dimension n ≥ 4. Then M is a complex space form if and only if R XJ X X is proportional to J X for any vector X tangent to M .

Now, let m be a point in an arbitrary Riemannian manifold M and γ a geodesic parametrized by arc length such that γ(0) = m. Denote u = γ 0 (0).

Next, let {E 1 , . . . , E n } be the parallel orthonormal frame field along γ with

E 1 (0) = u. Let G m (r) denote the geodesic sphere centered at m and with

radius r < i(m), the injectivity radius at m. For a point p = γ(r) =

exp m (ru) ∈ G m (r) we have the following expansions for the curvature tensor

R, the Ricci tensor e % and the second fundamental form σ of G e m (r) with

respect to {E 1 , . . . , E n }:

(3)

(3) R e abcd (p) = 1

r 2 (δ ac δ bd − δ ad δ bc ) +



R abcd − 1

3 (R ubud δ ac + R uauc δ bd − R ubuc δ ad − R uaud δ bc )



(m) + O(r), (4) % e ab (p) = n − 2

r 2 δ ab +



% ab − 1

3 % uu δ ab − n 3 R uaub

 (m)

+ r



∇ u % ab − 1

4 ∇ u % uu δ ab − n + 1

4 ∇ u R uaub

 (m)

+ r 2  1

2 ∇ 2 uu % ab − 1

10 ∇ 2 uu % uu δ ab − n + 2

10 ∇ 2 uu R uaub

+ 1

9 R uaub % uu − 1 45

n

X

λ,µ=2

R 2 uλuµ δ ab

− n + 2 45

n

X

λ=2

R uauλ R ubuλ



(m) + O(r 3 ),

(5) σ ab (p) = 1

r δ ab − r

3 R uaub (m) + O(r 2 )

for a, b, c, d = 2, . . . , n, where R abcd = g(R E

a

E

b

E c , E d ) and similarly for the other tensors. We refer to [2], [5], [6], [9] for more details.

Since we are working in a K¨ ahler manifold we can make a specific choice for E 2 by means of the initial condition E 2 (0) = J u = J γ 0 (0). Hence, E 2 = J E 1 = J γ 0 . When (M n , g, J ) is a space of constant holomorphic sectional curvature c, we can write down complete formulas for e R, % and σ. e Using the technique of Jacobi vector fields [9] we find

(6) σ = λg + µη ⊗ η.

This together with (2) and the Gauss equation yields R e XYZW =  c

4 + λ 2



{g(X, Z)g(Y, W ) − g(X, W )g(Y, Z)}

(7)

+ c

4 {g(J X, Z)g(J Y, W ) − g(J Y, Z)g(J X, W ) + 2g(J X, Y )g(J Z, W )}

+ µλ{g(X, Z)η(Y )η(W ) + g(Y, W )η(X)η(Z)

− g(X, W )η(Y )η(Z) − g(Y, Z)η(X)η(W )}.

By contraction we then obtain (8) % = e



(n − 2)λ 2 + (n + 1) c 4 + µλ

 g +



(n − 3)µλ − 3c 4



η ⊗ η,

(4)

where g denotes the induced metric and λ =

√ c 2 cot

√ c

2 r, µ+λ = √ c cot √

c r for c > 0, η(X) = g(X, E 2 (r)) and X, Y, Z, W are tangent vectors to G m (r).

When c < 0 one has to replace cot by coth and the formulas for c = 0 are obtained by taking the limit as c → 0.

Now, we will consider geodesic tubes, that is, tubes about a geodesic curve. We refer to [4], [5], [7], [9], [10] for more details. Let σ : [a, b] → M be a smooth embedded geodesic curve and let P r denote the tube of radius r about σ, where we suppose r to be smaller than the distance from σ to its nearest focal point. In that case, P r is a hypersurface of M . Let σ be parametrized by the arc length and denote by {e 1 , . . . , e n } an orthonormal basis of T σ(a) M such that e 1 = ˙σ(a). Further, let E 1 , . . . , E n be the vector fields along σ obtained by parallel translation of e 1 , . . . , e n . Then E 1 = ˙σ and {E 1 , . . . , E n } is a parallel orthonormal frame field along the geodesic σ. Next, let p ∈ P r and denote by γ the geodesic through p which cuts σ orthogonally at m = σ(t). We parametrize γ by arc length such that γ(0) = m and take (E 2 , . . . , E n ) such that E 2 (t) = γ 0 (0) = u. Finally, let {F 1 , . . . , F n } be the orthonormal frame field along γ obtained by parallel translation of {E 1 (t), . . . , E n (t)} along γ.

For the hypersurface P r one then has the following expansions with re- spect to this parallel frame field [4], [10]:

R e 1abc (p) =



R 1abc − 1

2 R 1ubu δ ac + 1

2 R 1ucu δ ab

 (m) (9)

+ r



∇ u R 1abc − 1

3 ∇ u R 1ubu δ ac + 1

3 ∇ u R 1ucu δ ab

 (m) + r 2  1

2 ∇ 2 uu R 1abc + 1

6 R 1ubu R aucu − 1

6 R 1ucu R aubu

− 1

8 ∇ 2 uu R 1ubu δ ac + 1

8 ∇ 2 uu R 1ucu δ ab

− 1

8 R 1u1u R 1ubu δ ac + 1

8 R 1u1u R 1ucu δ ab

− 1 24

n

X

λ=3

R 1uλu R buλu δ ac

+ 1 24

n

X

λ=3

R 1uλu R cuλu δ ab



(m) + O(r 3 ),

R e abcd (p) = 1

r 2 (δ ac δ bd − δ ad δ bc ) + R abcd (m) (10)

− 1

3 (R budu δ ac − R bucu δ ad + R aucu δ bd − R audu δ bc )(m)

+ O(r),

(5)

% e 11 (p) = % 11 (m) − (n − 1)R 1u1u (m) + O(r), (11)

% e 1a (p) = % 1a (m) − n − 1

2 R 1uau (m) (12)

+ r



u % 1a − n

3 ∇ u R 1uau

 (m) + r 2  1

2 ∇ 2 uu % 1a − n + 1

8 ∇ 2 uu R 1uau + 1

6 % uu R 1uau

− 3n − 5

24 R 1u1u R 1uau − n + 1 24

n

X

λ=3

R 1uλu R auλu

 (m) + O(r 3 ),

% e ab (p) = n − 3 r 2 δ ab +



% ab − n − 1 3 R aubu

(13)

− 1

3 % uu δ ab − 2

3 R 1u1u δ ab



(m) + O(r), σ 11 (p) = O(r),

(14)

σ 1a (p) = − r

2 R 1uau (m) + O(r 2 ), (15)

σ ab (p) = 1

r δ ab + O(r) (16)

for a, b, c, d ∈ {3, . . . , n}.

Now, suppose that (M n , g, J ) is a K¨ ahler manifold. Then, a point p = exp m (ru) on the geodesic tube P r will be called a special point when u = J ˙σ(t), that is, F 2 = J F 1 . For complex space forms of holomorphic sectional curvature c, computing the second fundamental form of P r by means of the technique of Jacobi vector fields at such a special point yields [7]

(17) σ(p) = λ g + µ η ⊗ η,

where g denotes the induced metric and λ =

√ c 2 cot

√ c

2 r, µ+λ = − √ c tan √

c r for c > 0. The values for c < 0 are obtained as usual by replacing the trigonometric functions by the corresponding hyperbolic functions and for c = 0 one has to take the limit c → 0. The tensor η in this case is determined by η(X)(p) = g(X, F 1 (r)) for tangent vectors X to P r at the special point p. Since σ has the same form as in (6), proceeding in the same way results in formally the same expressions for e R and % as in (7) and (8), respectively. e One only has to keep in mind that in the case of geodesic tubes, these formulas are only valid for the special points.

3. Horizontally Ricci-semi-symmetric and horizontally semi-

parallel geodesic spheres. A vector X ∈ T p G m (r) is called horizontal if

(6)

X is orthogonal to J γ 0 |p , where γ denotes the unit speed geodesic connecting m and p. This means that η(X) = 0. Moreover, the space of horizontal tangent vectors to G m (r) at p is spanned by E 3 (r), . . . , E n (r).

Then a small geodesic sphere G m (r) is said to be horizontally Ricci-semi- symmetric if e R XY · % = 0 for all horizontal tangent vectors on G e m (r).

The notion of horizontally semi-parallel geodesic spheres is defined in a similar way by means of the condition e R XY · σ = 0.

First, we prove the following result for complex space forms.

Theorem 3.1. Let (M n , g, J ), n ≥ 4, be a complex space form. Then the small geodesic spheres in M are horizontally Ricci-semi-symmetric and horizontally semi-parallel.

P r o o f. Using (8) it is easy to see that

−( e R XY · %)(W, W ) = 2µ e 2 η( e R XY W )η(W ),

where µ 2 = (n − 3)µλ − 3c/4. But η( e R XY W ) = −g( e R XY E 2 , W ). So, we have to show that

(18) R e XY E 2 = 0

for horizontal tangent vectors to G m (r).

Using (6) we see in the same way that (18) implies e R XY · σ = 0.

By means of (7) it is easy to verify that (18) is indeed satisfied for horizontal tangent vectors.

Next, we prove the converse theorems.

Theorem 3.2. Let (M n , g, J ), n ≥ 4, be a K¨ ahler manifold such that its small geodesic spheres are horizontally semi-parallel. Then (M, g, J ) is a complex space form.

P r o o f. Using (3) and (5) and considering the coefficient of r −1 in the power series expansion of

( e R ab · σ) cd = 0 for a, b = 3, . . . , n and c, d = 2, . . . , n yields

−δ ac R dubu + δ bc R duau − δ ad R cubu + δ bd R cuau = 0.

Next, take a = d 6= b and c = J u (that is, c = 2). Then we also have a 6= c, b 6= c since a, b ≥ 3, and we get R J uubu = 0 for b ≥ 3. This implies that R uJ uux = 0 for x orthogonal to J u. Hence, Theorem 2.1 yields that (M, g, J ) is a complex space form.

Theorem 3.3. Let (M n , g, J ), n ≥ 4, be a K¨ ahler manifold such that its

small geodesic spheres are horizontally Ricci-semi-symmetric. Then (M, g, J )

is a complex space form.

(7)

P r o o f. The assumption in the theorem yields ( e R ab · %) e cd = 0 for a, b = 3, . . . , n and c, d = 2, . . . , n. Using the power series expansions (3) and (4) and considering the coefficient of r −2 , r −1 and r 0 gives three conditions in which we make the choice b = d 6= a and c = J u (that is, c = 2). This leads to the following conditions:

% aJ u = n

3 R auJ uu , (19)

(∇ u %) aJ u = n + 1

4 (∇ u R) auJ uu , (20)

0 = 1

2 (∇ 2 uu %) aJ u − n + 2

10 (∇ 2 uu R) auJ uu

(21)

+ 1

9 R auJ uu % uu − n + 2 45

n

X

λ=2

R λuJ uu R λuau

for a orthogonal to span{u, J u}.

These three conditions are exactly those needed in the proof of Theo- rem 12 of [3, pp. 198–201]. Applying the same method (polarization and summation procedures) therefore leads to the required result.

4. Horizontally Ricci-semi-symmetric and horizontally semi- parallel geodesic tubes. In [1] a tangent vector X to a small geodesic tube P r is said to be horizontal if X is orthogonal to F 1 , the parallel translate of ˙σ along γ.

Now, if (M n , g, J ) is a K¨ ahler manifold, for special points p ∈ P r we see that X ∈ T p P r is horizontal if X is orthogonal to J γ 0 |p . Hence, a horizontal vector X at a special point p is determined by the condition η(X) = 0 and the spaces of horizontal vectors at p are spanned by F 3 , . . . , F n at p.

Next, a small geodesic tube P r will be called horizontally Ricci-semi- symmetric for special points if e R XY · % = 0 for all horizontal tangent vectors e X, Y at special points, and similarly P r is said to be horizontally semi- parallel for special points if e R XY · σ = 0 for the same choice of vectors X, Y .

We then have

Theorem 4.1. Let (M n , g, J ), n ≥ 4, be a complex space form. Then the small geodesic tubes in M are horizontally Ricci-semi-symmetric and horizontally semi-parallel for special points.

P r o o f. In the same way as in Theorem 3.1 we find that e R XY F 1 = 0 implies e R XY · % = 0 and e e R XY · σ = 0 for X, Y tangent to P r . So, we have to show that

R e XY F 1 = 0

(22)

(8)

for horizontal tangent vectors at special points. But at special points e R has the same form as in (7). Using the horizontality of X, Y , it is easy to see that (22) holds.

Finally, we consider the converse theorems.

Theorem 4.2. Let (M n , g, J ), n ≥ 4, be a K¨ ahler manifold all of whose geodesic tubes are horizontally semi-parallel for special points. Then (M, g, J ) is a complex space form.

P r o o f. The assumption yields ( e R ab · σ) 1c = 0 for a, b, c = 3, . . . , n.

Using the power series expansions (9), (10), (14)–(16) and considering the coefficient of r −1 yields R 1cab = 0. Now, take b = c = J a. Then, since F 1 (0) = −J u, we get R J uJ aaJ a = 0 and hence R uaJ aa = 0, for a orthogonal to the plane (u, J u). Since this must hold for all tubes, the result follows from Theorem 2.1.

Theorem 4.3. Let (M n , g, J ), n ≥ 4, be a K¨ ahler manifold all of whose geodesic tubes are horizontally Ricci-semi-symmetric for special points. Then (M, g, J ) is a complex space form.

P r o o f. Using (9)–(13) we can write down the power series expansion for ( e R ab · %) e 1a = 0, a, b = 3, . . . , n.

Considering the coefficient of r −2 and taking b = J a results in %(u, a) = R aJ uuJ u + (n − 3)R uJ aaJ a for any unit tangent vectors a, u on M , with a orthogonal to u and J u. Switching a and u and subtracting the equations obtained yields, for n 6= 4 and a, u as above, that %(u, a) = (n − 4)R uJ aaJ a

and hence %(a, J u) = (n − 4)R auJ uu . Although the coefficient of R auJ uu in this expression is different from the one in (19), using a similar polarization and summation procedure as in the first part of the proof of Theorem 12 in [3, p. 198] gives the result for n 6= 4. (We omit the details.)

For n = 4 we consider the coefficient of r 0 . In this expression we regroup equal terms and use the identity ∇ 2 uu % 1a = ∇ 2 uu R 1uau + ∇ 2 uu R 1bab . Finally, taking b = J a results in

0 = (2R uaJ aa + R auJ uu )(R J uuJ uu − R J uaJ ua ) + R auJ au (2R uJ aaJ a − R aJ uuJ u )

for a, u unit tangent vectors on M , with a orthogonal to u and J u.

First, we replace a and u by a/kak and u/kuk respectively. Then we obtain a homogeneous expression which is also valid for non-unit vectors a and u.

Next, we polarize this expression, replacing a by αa + βu, which we may do, since αa + βu is orthogonal to u, J u if a is orthogonal to u, J u. Writing down the coefficient of α 3 β 2 and β 5 yields

(23) AB + DC = 0, DB − AC = 0,

(9)

where

A = 2R uaJ aa + R auJ uuu , B = R J uuJ uu − R auau ,

C = R auJ au ,

D = R aJ uuJ u − 2R uJ aaJ a .

Since (23) is a homogeneous system of linear equations with determinant different from zero if A 6= 0, we always get AB = 0. Explicitly, this means (24) (2R uaJ aa + R auJ uu )(R J uuJ uu − R auau ) = 0

for unit tangent vectors a, u on M , with a orthogonal to u, J u.

Again, we homogenize (24) and polarize, replacing a by αa + βu and u by βa − αu. Writing down the coefficients of the polynomial obtained by this procedure gives

(25)

 

 

 

 

 

 

 

 

 

 

 

 

(2H + G)X = 0,

2(2H + G)E + 3KX = 0,

(2H + G)Z + 24KE − (G − H)X = 0,

2(2H + G)F + 3K(Z + X) − 2(G − H)E = 0,

(2H + G)Y + 24K(F + E) − (G − H)Z − (2G + H)X = 0, 2(2G + H)E − 3K(Z + Y ) + 2(G − H)F = 0,

(2G + H)Z − 24KF + (G − H)Y = 0, 2(2G + H)F − 3KY = 0,

(2G + H)Y = 0, where

X = R J uuJ uu − R auau , Y = R J aaJ aa − R auau ,

Z = 2R J uuJ aa + 4R J uaJ ua − 2R auau , E = R aJ uuJ u , F = R uJ aaJ a ,

G = R auJ uu , H = R uaJ aa , K = R auJ au .

First we suppose that 2H + G = 0. The last two equations in (25) then yield that HF = 0. On the contrary, if 2H + G 6= 0, we can use the first four equations to derive that F = 0. So, in both cases we obtain HF = 0, which means that R uaJ aa R uJ aaJ a = 0 for all a, u tangent to M with a orthogonal to u, J u. Replacing u by u + J u in this condition eventually leads to R uaJ aa = 0. Then the result for n = 4 follows by Theorem 2.1.

REFERENCES

[1] E. B o e c k x, J. G i l l a r d and L. V a n h e c k e, Semi-symmetric and semi-parallel geodesic spheres and tubes, Indian J. Pure Appl. Math., to appear.

[2] B. Y. C h e n and L. V a n h e c k e, Differential geometry of geodesic spheres, J. Reine

Angew. Math. 325 (1981), 28–67.

(10)

[3] M. D j o r i ´ c and L. V a n h e c k e, Almost Hermitian geometry , geodesic spheres and symmetries, Math. J. Okayama Univ. 32 (1990), 187–206.

[4] L. G h e y s e n s and L. V a n h e c k e, Total scalar curvature of tubes about curves, Math. Nachr. 103 (1981), 177–197.

[5] A. G r a y, Tubes, Addison-Wesley, Reading, 1989.

[6] A. G r a y and L. V a n h e c k e, Riemannian geometry as determined by the volumes of small geodesic balls, Acta Math. 142 (1979), 157–198.

[7] —, —, The volumes of tubes about curves in a Riemannian manifold , Proc. London Math. Soc. 44 (1982), 215–243.

[8] S. T a n n o, Constancy of holomorphic sectional curvature in almost Hermitian man- ifolds, K¯ odai Math. Sem. Rep. 25 (1973), 190–201.

[9] L. V a n h e c k e, Geometry in normal and tubular neighborhoods, Rend. Sem. Fac.

Sci. Univ. Cagliari, Supplemento al Vol. 58 (1988), 73–176.

[10] L. V a n h e c k e and T. J. W i l l m o r e, Interaction of tubes and spheres, Math. Ann.

263 (1983), 31–42.

[11] K. Y a n o and M. K o n, Structures on Manifolds, Ser. in Pure Math. 3, World Sci., Singapore, 1984.

Department of Mathematics Katholieke Universiteit Leuven Celestijnenlaan 200B

B-3001 Leuven, Belgium

E-mail: jurgen.gillard@wis.kuleuven.ac.be

Received 26 September 1995

Cytaty

Powiązane dokumenty

So, the following theorem extends the Krengel–Lin decomposition which is discussed in [KL] only for compact groups..

ANNALES SOCIETATIS MATHEMATICAE POLONAE Series I: COMMENT ATI ONES MATHEMATICAE X IX (1976) ROCZNIKI POLSKIEGO TOWARZYSTWA MATEMATYCZNEGO.. Séria I: PRACE MATE

While studying the representation theory of the trivial extension T (A) of an artin algebra A by its minimal injective cogenerator bimodule DA, Tachikawa [12] and Yamagata [13]

The ‘only if direction’ follows from the well-known result that every induced subgraph of a strongly chordal graph has a simple vertex , meaning a vertex v such that the

[r]

Berndtsson’s estimate (5) is closely related to the Ohsawa-Takegoshi extension theorem [12] but the latter cannot be deduced from it directly (it could be if (5) were true for ı D

She is pleased to thank the Department of Mathematics of Wesleyan University for generous hospitality during the spring semester of 1992.. The second author is pleased to thank

In fact, since any graph can be identified with a symmetric digraph (where each edge is replaced by a double arc) the direct product of graphs is a special case of the direct product