• Nenhum resultado encontrado

A gap theorem for complete constant scalar curvature hypersurfaces in the de Sitter space

N/A
N/A
Protected

Academic year: 2018

Share "A gap theorem for complete constant scalar curvature hypersurfaces in the de Sitter space"

Copied!
14
0
0

Texto

(1)

A gap theorem for complete constant scalar

curvature hypersurfaces in the de Sitter space

Aldir Brasil Jr.

a

, A. Gervasio Colares

a,1

, Oscar Palmas

b,∗,2 aDepartamento de Matemática, Universidade Federal de Ceará, 60.455-760 Fortaleza CE, Brazil

bDepartamento de Matemáticas, Facultad de Ciencias, UNAM, México 04510 DF, Mexico

Received 25 May 2000

Abstract

To each immersed complete space-like hypersurfaceMwith constant normalized scalar curvature

Rin the de Sitter spaceS1n+1, we associate supH2, whereHis the mean curvature ofM. It is proved that the condition supH2Cn(R)¯ , whereR¯=(R−1) >0 andCn(R)¯ is a constant depending

only onRandn, implies that eitherMis totally umbilical orMis a hyperbolic cylinder. It is also proved the sharpness of this result by showing the existence of a class of new rotation constant scalar curvature hypersurfaces inS1n+1such that supH2 > Cn(R)¯ . © 2001 Elsevier Science B.V.

All rights reserved.

MSC:53C42; 53C40; 53A10

Subj. Class.:Differential geometry; General relativity

Keywords:de Sitter space; Hyperbolic cylinder; Rotation hypersurfaces

1. Introduction

LetRn1+2be the real vector spaceRn+2endowed with the Lorentzian metrich,igiven by hv, wi = −v0w0+v1w1+· · ·+vn+1wn+1; that is,R1n+2=Ln+2is the Lorentz–Minkowski

(n+2)-dimensional space. We define the de Sitter space as the following hyperquadric of

Rn1+2:S1n+1= {x R1n+2; |x|2=1}. The induced metrich,imakesS1n+1into a Lorentz manifold with constant sectional curvature 1.

∗Corresponding author.

E-mail addresses:aldir@mat.ufc.br (A. Brasil Jr.), gcolares@mat.ufc.br (A. Gervasio Colares), opv@hp.fciencias.unam.mx (O. Palmas).

1Research partially supported by CNPq and PRONEX.

2Research partially supported by DGAPA (Project IN119298), CONACYT (Project 27969), and CNPq. 0393-0440/01/$ – see front matter © 2001 Elsevier Science B.V. All rights reserved.

(2)

In 1977, Goddard [8] conjectured that the only complete space-like hypersurfaces (those whose induced metric is Riemannian) inSn1+1with constant mean curvature are the totally umbilical ones. Montiel [12] proved this conjecture in the compact case and exhibited a counterexample for the general case. The motivation to the study of space-like hypersurfaces in space–times comes from its relevance in general relativity. In particular, the de Sitter space is a space–time model with constant sectional curvature. On the other hand, constant mean curvature hypersurfaces are relevant for studying propagation of gravitational waves.

It is quite natural to pose a Goddard-like question for constant scalar curvature hypersur-faces inS1n+1. Partial results were obtained in [5,20,21]. For the constant scalar curvature compact Riemannian case in spheres, a rigidity theorem is given in [10]. Recently, Li [11] proved that a compact space-like hypersurface with constant normalized scalar curvature

R >1 inSn1+1must be totally umbilical (see also [14] for a similar result in a more general space–time and [1] for related results involving generalr-mean curvatures). On the other hand, Montiel [13] proved that a complete space-like hypersurface with constant mean cur-vatureH2 =4(n1)/n2inS1n+1and more than one topological end is isometric to the hyperbolic cylinderH1(1 coth2r)×Sn−1(1tanh2r).

In this paper, we substitute Montiel hypotheses for constant normalized scalar curvature

R > 1 and supH2 Cn(R)¯ , where H is the mean curvature and Cn(R)¯ is a constant depending only on the dimensionn3 andR¯ =R1. We then classify all hypersurfaces satisfying these conditions. More precisely, we prove the following theorem (which was announced in [3]).

Theorem 1.1. Let Mn be an n-dimensional(n 3) complete space-like hypersurface immersed inS1n+1 with constant normalized scalar curvature R and= R1 > 0.

Suppose also thatsupH2Cn(R)¯ ,where Cn(R)¯ =

1

n2

(n1)2nR¯−2

n2 +2(n−1)+

n2

nR¯2

. (1)

Then either

1. supH2= ¯Rand M is totally umbilical; or

2. supH2=Cn(R)¯ and M is isometric to the hyperbolic cylinder H1(1−coth2r)×Sn−1(1− tanh2r).

Theorem 1.1 has the following consequences:

1. The only complete constant normalized scalar curvature space-like hypersurfaces with ¯

R=supH2are the umbilical ones.

2. There are no complete constant normalized scalar curvature space-like hypersurfaces such thatR <¯ supH2< Cn(R)¯ .

3. The only complete constant normalized scalar curvature space-like hypersurfaces with supH2=Cn(R)¯ are the hyperbolic cylinders.

(3)

hyper-surface distinct from the above ones and such that supH2=C. These examples constitute a class of new rotation hypersurfaces with constant scalar curvature inS1n+1. In fact, we prove the following theorem.

Theorem 1.2. Fix an integern3and let(2/n,+∞).Then, for eachC Cn(R)¯ ,

there exists a complete immersed space-like hypersurface inS1n+1with normalized scalar curvatureR = ¯R+1and mean curvature H satisfyingsupH2 =C.Moreover, ifC = Cn(R)¯ ,then the hypersurface is a hyperbolic cylinder.

Chern et al. [4] associated to each compact hypersurface in the Euclidean sphereSn+1

withH = 0, the square of the norm of its second fundamental form (or equivalently, its scalar curvature) and asked if the image of such a function is discrete. A Lorentzian version dual of this question could be to associate to each complete space-like hypersurface inS1n+1

with constant scalar curvature the value supH2and ask if the image of such a function is discrete.

The results we obtain give, for eachn3, a detailed picture of the setGnof the complete constant normalized scalar curvature space-like hypersurfaces inS1n+1. In fact, our results can be interpreted as follows: givenn3, we associate to each such a complete constant normalized scalar curvature space-like hypersurfaceMn the coordinate pair(R,¯ supH2)

in the first quadrant of a 2-plane, thus obtaining Fig. 1. In this figure, the bisector ray

Fig. 1. The plane(R,¯ supH2), where the curve shown is the graph ofCn(R)¯ corresponding to the hyperbolic

(4)

corresponds to the umbilical hypersurfaces and the curve shown there corresponds to the hyperbolic cylinders. We see that for eachR¯6=1, there is no complete hypersurface of the kind studied here and such that the corresponding point(R,¯ supH2)lies above the bisector ray and below the curve, thus yielding a gap (see Section 4). Our results also show that fornandR¯ fixed, the image of the functionGn →Rgiven byMn 7→ supH2is the set { ¯R} ∪[C(R),¯ ), which obviously is not discrete, thus giving an answer to a Lorentzian version dual to the Chern–do Carmo–Kobayashi’s question.

2. Preliminaries

LetMnbe ann-dimensional complete orientable manifold and letf :Mn S1n+1

Rn1+2be a space-like immersion ofMninto the(n+1)-dimensional de Sitter spaceS1n+1. Choose a unit normalηalongf and denote byA: TpM → TpMthe linear map of the tangent spaceTpMat the pointp ∈M, associated to the second fundamental form off alongη,

hAX, Yi = −h∇XY, ηi,

whereX and Y are tangent vector fields on M and is the connection on S1n+1. Let {e1, . . . , en}be an orthonormal basis which diagonalizesAwith eigenvalueski, i.e.,Aei = kiei,i = 1, . . . , n.|A|2 = Pk2i is the square of the norm of the second fundamental form,H =(1/n)P

kiis the mean curvature off andR = ¯R+1 is the normalized scalar curvature, where

¯

R=X

i6=j kikj.

In our case it is convenient to define a linear mapφ:TpM→TpMby

hφX, Yi = hAX, Yi −HhX, Yi.

It is easily checked that tr(φ)=0 and that

|2= 1

2n

X

(ki−kj)2,

so that|φ|2 =0 if and only ifMnis totally umbilical. Note that the eigenvalues ofφare given byµi =ki−H, so that|φ|2=Piµ2i = |A|2−nH2and

X

i

ki3=nH3+3HX i

µ2i X i

µ3i. (2)

We have the Gauss equation relatingR¯,Hand|A|2:

n(n1)R¯=n2H2− |A|2. (3)

The standard examples of space-like umbilical hypersurfaces with constant mean curvature in the de Sitter space are given by

(5)

wherea Rn1+2,|a|2=ρ =1,0,1 andτ2> ρ. The corresponding mean curvatureH

of such hypersurfaces satisfies

H2= τ

2

τ2ρ

(see [12], for instance) andMnis isometric to a hyperbolic space, an Euclidean space or a sphere ifρ=1,0,1, respectively.

On the other hand, the hyperbolic cylinders are the hypersurfaces given by

Mn= {pS1n+1; −p02+p21+ · · · +pk2= −sinh2r}

withr Rand 1≤k n1. Such hyperbolic cylinders have constant mean curvature

nH=kcothr+(nk)tanhr. Thus we have

H2 4(n−1)

n2 ,

and the equality is attained fork = 1 and coth2r = (n1). Their normalized scalar curvature isR =1+(1/n)(2+(n2)tanh2r). We point out that these examples have only two distinct constant principal curvatures at each point and one of them has multiplicity one. Moreover, they are isometric to the Riemannian productH1(1coth2r)×Sn−1(1 tanh2r).

Also, we observe that the cylindersHn−1(1coth2r)×S1(1tanh2r)have normalized scalar curvatureR =1+(1/n)(2+(n2)coth2r).

The Laplacian1acts on anyC2-functionf defined onMas(1f )ij=1fij =Pkfijkk and the Laplacian of the second fundamental formhijis given by

1hij=

X

k hijkk.

We may write

1hij=

X

k

(hijkk−hikjk)+

X

k

(hikjk−hikkj)+

X

k

(hikkj−hkkij),

so that

1hij=nHij+

X

m,k

himRmkjk+

X

m,k

hkmRmijk,

whereHijdenotes the second covariant derivative ofH. LetT =P

i,jTijωi⊗ωjbe the symmetric tensor defined by

Tij=nHδij−hij.

Following Cheng and Yau [6], we introduce the operator L1 associated toT acting on

C2-functionsf onMnby

L1f =

X

i,j

Tijfij =

X

i,j

(6)

Around a given pointp Mwe choose an orthonormal frame field{e1, . . . , en}with dual frame field{w1, . . . , wn}so thathij=kiδijatp. Using (3) and (4), we have

L1(nH)=nH1(nH)−

X

i

ki(nH)ii= 1 21(nH)

2 −X

i

(nH)2i X i

ki(nH)ii

=1

2n(n−1)1R+ 1 21|A|

2

−n2|∇H|2X i

ki(nH)ii.

On the other hand, Simons formula (see [17, p. 320, (3.5)]) implies 1

21|A| 2

= |∇A|2+nX i

kiHii+ 1 2

X

i,j

Rijij(ki −kj)2. From the last two equations, we have

L1(nH)= 1

2n(n−1)1R+ |∇A| 2

−n2|∇H|2+1

2

X

i,j

Rijij(ki−kj)2. (5)

3. Classification of complete constant scalar curvature hypersurfaces with

¯

RRR¯¯supsupsupHHH222

Here we prove Theorem 1.1; for this we will need the following three lemmas.

Lemma 3.1. LetMnbe an immersed space-like hypersurface inS1n+1with constant nor-malized scalar curvature. Then

L1(nH)= |∇A|2n2|∇H|2+ |φ|2(nnH2+ |φ|2)nHX

i

µ3i. (6)

Proof. SinceRis constant, (3) and (5) imply

L1(nH)= |∇A|2−n2|∇H|2+ 1 2

X

i,j

(1+kikj)(ki−kj)2= |∇A|2−n2|∇H|2

+1 2n

X

i

k2i +1

2n

X

j

kj2X i,j

kjki+ 1 2

X

i,j ki3kj+

1 2

X

i,j kikj3−

X

i,j k2ik2j.

Makingi=j, we obtain

L1(nH)= |∇A|2−n2|∇H|2+n|A|2−n2H2+ |A|4−nH

X

i k3i.

Using (2) in the expression above, we have

L1(nH)= |∇A|2n2|∇H|2+n|φ|2+ |A|4nHX

i

µ3i 3nH2|φ|2n2H4.

Then

L1(nH)= |∇A|2−n2|∇H|2+n|φ|2+ |φ|4+n2H4 +2nH2|φ|23nH2|φ|2n2H4nHX

(7)

This yields

L1(nH)= |∇A|2−n2|∇H|2+ |φ|2(n−nH2+ |φ|2)−nH

X

i µ3i,

which proves the lemma.

Lemma 3.2 (Okumura [17, p. 210]). Let µi,i = 1, . . . , n be real numbers such that

Pµ

i =0andPµ2i =β ≥0,withβconstant. Then −√n−2

n(n1)β

3

≤Xµ3i n−2 n(n1)β

3,

and equality holds on the right-hand side, if and only if

µ1= · · · =µn−1= −

s

1

n(n1)β, µn=

r

n1

n β.

Lemma 3.3. LetMn be an n-dimensional complete(n 3)space-like hypersurface im-mersed into the de Sitter spaceS1n+1 with constant normalized scalar curvature R and

¯

R=R1.Then the following inequalities are equivalent: 1. |Z|2n/((n2)(nR¯2))[n(n1)R¯2+4(n1)R¯+n],

2. (n2)/np(n(n1)R¯+ |Z|2)(|Z|2nR)¯ n2(n1)R¯+(n2)/n|Z|2,

where|Z|2=sup|A|2.Equivalence also holds in case of equality.

Proof. We will prove that (2) implies (1). Squaring (2) and simplifying, we get

n

−2

n

[(n2)2R¯2(n2(n1)R)¯ ]|Z|2

≤(n2(n1)R)¯ 2+(n2)2(n1)R¯2,

hence,

(n2)(nR¯2)|Z|2n[n(n1)R¯24(n1)R¯+n],

which gives (1). The other implication is immediate.

We will also use the maximum principle at infinity for complete manifolds due to Omori [16] and Yau [19]:

LetMnbe an n-dimensional complete Riemannian manifold whose Ricci curvature is bounded from below. Let f be aC2-function bounded from below on Mn.Then for each ε >0there exists a pointpε∈Msuch that

|∇f|(pε) < ε, 1f (pε) >−ε, inff ≤f (pε) <inff +ε.

Proof of Theorem 1.1. Applying Lemma 3.2 to the eigenvalues ofφ, we obtain

X

µ3i

n2 √

n(n1)(|φ|

(8)

Using this fact and Lemma 3.1, we get

L1(nH)≥ |∇A|2−n2|∇H|2+ |φ|2PH(|φ|), (8)

wherePHis the polynomial in|φ|given by

PH(|φ|)=n−nH2+ |φ|2−

n(n2)

n(n1)|H||φ|. (9)

SinceR¯is constant and positive, Lemma 4.1 of [2] implies

|∇A|2n2|∇H|20.

Substituting this fact in (8), we obtain

L1(nH)≥ |φ|2PH(|φ|). (10)

From (3),

|2= |A|2nH2= n−1

n (|A|

2

−nR).¯ (11)

Thus we may seePH(|φ|)asPR¯(|A|), where

PR¯(|A|)=n2(n1)R¯+n−2 n |A|

2

−n−n 2

q

(n(n1)R¯+ |A|2)(|A|2nR).¯ (12)

Hence, we may write (10) as

L1(nH)≥

n1

n (|A|

2

−nR)P¯ R¯(|A|). (13)

We claim thatPR¯(psup|A|2)0. By hypothesis (see (1)), supH2Cn(R)¯ =

1

n2

(n1)2nR¯−2

n2 +2(n−1)+

n2

nR¯2

.

We use (3) to write this expression as

|Z|2 n

(n2)(nR¯2)[n(n−1)R¯

2

+4(n1)R¯+n],

where again|Z|2=sup|A|2. By Lemma 3.3, this is equivalent to

n2

n

q

(n(n1)R¯+ |Z|2)(|Z|2nR)¯ n2(n1)R¯+n−2 n |Z|

2.

In view of (12), this last inequality shows thatPR¯(psup|A|2) 0, so that our claim is proved.

On the other hand,

L1(nH)=

X

i,j

(nHδijnhij)(nH)ij =

X

i

(nHnhii)(nH)ii

=nX i

H (nH)ii−n

X

i

(9)

where sup|H|is the sup of the absolute value of the mean curvatureHinMandC=minki the minimum of the principal curvatures inM.

From (13) and (14), we have

n1

n (|A|

2

−nR)P¯ R¯(|A|)L1(nH)≤n(sup|H| −C)1(nH). (15) Since R¯ is the second symmetric function of the eigenvalues of A, we have from the Newton inequalities thatR¯H2, and henceR¯supH2. Gauss equation impliesRicM ≥ (n1)14nH2, so that the Ricci curvature is bounded from below. Thus we may apply Omori and Yau’s result to the function

f =p 1

1+(nH)2,

which is a positive smooth function onM. Then,

|∇f|2= 1

4

|∇(nH)2|2

(1+(nH))2, (16)

1f = −1

2

1(nH)

(1+(nH))3/2+

3|∇(nH)|2

4(1+nH)5/2. (17)

Let{pk},k ∈ N, be a sequence of points inMgiven by Omori and Yau’s result, such that

lim

k→∞f (pk)=inff, 1f (pk) >−

1

k, |∇f|

2(p k) <

1

k2. (18)

Using (16)–(18) and the fact that

lim

k→∞(nH)(pk)=supp∈M(nH)(p),

we have

−1

k <−

1 2

1(nH)

(1+(nH))3/2(pk)+ 3

k2(1+nH(pk)) 3/2. Hence,

1(nH)

(1+nH)3/2(pk) < 2

k

1

1+(nH)(pk)+ 3

k

. (19)

Evaluating (15) at the pointspk, makingk→ ∞and using (19), we have

0n−1

n (sup|A|

2

−nR)P¯ R¯

q

sup|A|2

≤L1(nsup|H|) ≤n(sup|H| −C)1(nH) < n(sup|H| −C)(1+nH)3/2(pk)

2

k

×

1

1+(nH)(pk)+ 3

k

(10)

Ifk→ ∞, we have that the last expression goes to zero, so either

(sup|A|2nR)¯ =0, or PR¯

q

sup|A|2

=0.

In the first case, Gauss equation implies that

nsup|φ|2=(n1)(sup|A|2nR)¯ =0,

so that|φ|2=0 andMnis totally umbilical.

In the second case, i.e., ifPR¯psup|A|2=0, then Lemma 3.3 implies that sup|A|2= n

(n2)(nR¯2)[n(n−1)R¯

2

+4(n1)R¯+n]. (20)

Using (11), we have

supH2= 1 n2sup|A|

2

−n−1

n R.¯ (21)

Substituting (20) in (21), we obtain supH2=Cn(R)¯ (defined in (1)) and so we have the following equality:

L1(nsupH )= n−1 n (sup|A|

2

−nR)P¯ R¯

q

sup|A|2

.

Thus, equality also holds on the right-hand side of Okumura’s lemma (Lemma 3.2). After re-enumeration, we havek1 =k2 = · · · =kn−1,k1 6= kn, wherek1 =tanhr andkn = cothr. ThenMnis isometric toH1(1coth2r)×Sn−1(1tanh2r), finishing the proof

of the theorem.

4. Examples of constant scalar curvature hypersurfaces satisfyingsupsupsupHHH222> C> C> Cnnn(((R)R)R)¯¯¯ To show that the restriction supH2Cn(R)¯ in Theorem 1.1 is sharp, we will give here for fixedn 3,R >¯ 2/nand every numberC > Cn(R)¯ , a complete hypersurface with constant scalar curvatureR¯such that supH2=C. This is the content of Theorem 1.2 which we will prove here. First we recall the notion of a rotation hypersurface in the context of the de Sitter space [15] (see [7] for the Riemannian case).

Definition 4.1. Denote byPk ak-dimensional linear subspace of Rn1+2 and by O(P2)

the identity component of the Lorentz group O(1, n+1)which leavesP2pointwise fixed. ChooseP2andP3such thatP2P3, and letCbe a space-like curve inS1n+1(P3P2). Then the orbit ofCunder O(P2)is a hypersurfaceMnS1n+1called the rotation hypersur-face generated byC, andCis called the profile curve ofM.Mis spherical (resp. parabolic, hyperbolic) if the restrictionh,i|P2 is a Lorentzian metric (resp. a Riemannian metric, a degenerate quadratic form).

(11)

LetP3S1n+1be the subspace{(y0, y1, y2,0, . . . ,0)∈R1n+2;yi ∈R}, and the profile curveCbe given by the functionsyi =yi(s),i=0,1,2, parameterized by arc-length, so that the following conditions hold (from now on, we omit the variablesfor convenience):

−y02+y12+y22=1, y0′2+y1′2+y2′2=1. (22) We may describey0, y1in terms ofy2, as follows. Set

y0=

q

y221 coshϕ, y1=

q

y221 sinhϕ

for a function ϕ to be defined. Differentiating these equations and substituting into the second equation in (22), we get

1=(y221)ϕ′2 y

2 2y2′2

y221 +y 2 2, so that

ϕ=

Z q

y2′2+y221

y221 ds.

Hence, the coordinates of the profile curveCmay be given in terms ofy2and its derivatives. In this setting, if we rotateC with respect toP2 = {(y0, y1,0, . . . ,0)∈ Rn1+2;yi ∈ R}, we may calculate the principal curvatures of our rotation hypersurface (see [7,15]), which are given by

κi =

q

y2′2+y22δ y2

, (23)

κn=

y2′′+y2

q

y2′2+y22δ

, (24)

wherei=1, . . . , n1 andδ= −1,0,1, when the hypersurface is spherical, parabolic or hyperbolic, respectively.

It is worth noting that the hyperbolic cylindersH1(1 coth2r)×Sn−1(1 tanh2r),

r >0 already mentioned in this paper may be given as rotation hypersurfaces as follows: if 2/n <R <¯ 1, we sety2=coshr, where

tanh2r= nR¯−2 n2 ,

while ifR >¯ 1, we usey2=sinhr, where coth2r =nR¯−2

n2 .

(12)

curvatures to obtain the following formulas forHandR¯:

nH=(n1)

q

y2′2+y22δ

y2 +

y2′′+y2

q

y2′2+y22δ

, (25)

nR¯=(n2)y ′2

2 +y22−δ

y22 +2

y2′′+y2

y2

. (26)

Now we use the following result (see [9,18]):

Ifis constant, the last equation has the first integral given by

G(y2, y2′)=y2n−2(y2′2+y22−1− ¯Ry22). (27)

We translate the study of the level curves ofGin the(y2, y2′)-plane into information about the profile curves of our rotation hypersurfaces; for example, the critical points(yc,0)of

G, satisfying

n(1− ¯R)yc2(n2)δ=0,

or, equivalently,

yc2= (n−2)δ

n(1− ¯R), (28)

correspond exactly to the hyperbolic cylinders

H1(1 coth2r)×Sn−1(1 tanh2r)

with principal curvatureski =tanhrandkn=cothr. We use (25), (26) and (28) to write H2in terms ofR¯, obtaining

H2= 1 n2

(n1)2nR¯−2

n2 +2(n−1)+

n2

nR¯2

,

which is the value ofCn(R)¯ defined in (1) and corresponding to the hyperbolic cylinders, thus defining a functionR¯7→H2. Fig. 1 (see Introduction) shows the graph of this function on a(R,¯ supH2)-plane. We remark that in the vertical axis we plot supH2and notH2. Also, if we take an umbilical hypersurface, we have supH2 = H2 = ¯R, so that these hypersurfaces are represented by the first quadrant bisector ray.

Proof of Theorem 1.2. Regarding Fig. 1, we will show that every point above the graph ofCn(R)¯ has associated a rotation space-like hypersurface such that(R,¯ supH2)are the coordinates of the given point. As before, we work out the details only in the spherical case. The level curves of the first integral given in (27) give rise to space-like hypersurfaces with constant scalar curvature. In particular, we obtain complete space-like hypersurfaces whenever the level curve stays inside the region{y2>1}. If 2/n <R <¯ 1, the critical point

(13)

These nearby closed level curves have itsy2-coordinate bounded between, say,y∗andy∗

(these values, of course, depend on the level curve we are considering) and so the profile curve is contained in a strip given byyy2y∗, containing the critical pointyc.

We claim that, if we consider all closed level curves which stay inside the region{y2>1}, then, for the correspondingcompletespace-like hypersurfaces (withR¯fixed), the value of supH2varies in [Cn(R),¯ ∞).

To prove the claim, we first writeH2in terms ofR¯,y2andy2′; from (26),

y2′′+y22

q

y′22+y22δ

= y2

2qy2′2+y22δ

nR¯(n2)y ′2

2 +y22−δ

y22

!

=nR¯ 2

y2

q

y2′2+y22δ

−(n2)

q

y2′2+y22δ

y2 ,

so that

H =1

2

 q

y2′2+y22δ

y2 + ¯R

y2

q

y2′2+y22δ

. (29)

Fixing the value of the first integralG(y2, y2′)=k, so that

y2n−2(y2′2+y22δ− ¯Ry22)=k,

we get

q

y′22+y22δ

y2 =

s

¯

Ry2n+k y2n .

Using this last expression in (29), we get

H =1

2

 s

¯

Ry2n+k y2n + ¯R

s

y2n

¯

Ry2n+k

.

It is easy to show thatH, as a function ofy2, is differentiable, defined in an interval of the form(1, yR¯], where

G(yR¯,0)=G(1,0)= − ¯R,

and thatHisdecreasingin its domain. These properties imply that for a closed level curve

G(y2, y2′)=kofG, withy2-extreme valuesy∗andy∗,

supH2=H2(y),

(14)

If we makey1+, the valuekofGtends toG(1,0)= − ¯Rand we have

lim y→1+supH

2 = lim

y→1+ 1 4

 s

¯

Ryn

∗ +k yn

+ ¯R

s

yn

¯

Ryn

∗+k 

2 = ∞.

On the other hand,

lim y→yc−

supH2= lim y→y−c

H2(y)=H2(yc).

But asyccorresponds to a hyperbolic cylinder,H2(yc)=Cn(R)¯ , and so the claim and the

theorem are proved.

References

[1] J.A. Aledo, L.J. Al´ıas, A. Romero, Integral formulas for compact space-like hypersurfaces in de Sitter space: applications to the case of constant higher order mean curvature, J. Geom. Phys. 31 (1999) 195–208. [2] H. Alencar, M. do Carmo, A.G. Colares, Stable hypersurfaces with constant scalar curvature, Math. Z. 213

(1993) 117–131.

[3] A. Brasil, A.G. Colares, On space-like hypersurfaces with constant scalar curvature in the de Sitter space, An. Acad. Bras. Ciencias 72 (4) (2000), in press.

[4] S.S. Chern, M. do Carmo, S. Kobayashi, Minimal submanifolds of the sphere with second fundamental form of constant length, Functional Analysis and Related Fields, Browder, Springer, Berlin, 1970, pp. 57–75. [5] Q.M. Cheng, S. Ishikawa, Space-like hypersurface with constant scalar curvature, Manuscripta Math. 95 (4)

(1998) 499–505.

[6] S.Y. Cheng, S.T. Yau, Hypersurfaces with constant scalar curvature, Math. Ann. 225 (1977) 195–204. [7] M.P. do Carmo, M. Dajczer, Rotation hypersurfaces in spaces of constant curvature, Trans. AMS 277 (2)

(1983) 685–709.

[8] A.J. Goddard, Some remarks on the existence of space-like hypersurfaces of constant mean curvature, Math. Proc. Cambridge Phil. Soc. 82 (1977) 489–495.

[9] M.L. Leite, Rotational hypersurfaces of space forms with constant scalar curvature, Manuscripta Math. 67 (1990) 285–304.

[10] H. Li, Hypersurfaces with constant scalar curvature in space forms, Math. Ann. 305 (1996) 665–672. [11] H. Li, Global rigidity for hypersurfaces, Ark. Mat. 35 (1997) 327–351.

[12] S. Montiel, An integral inequality for compact space-like hypersurfaces in the de Sitter space and applications to the case of constant mean curvature, Indiana Univ. Math. J. 37 (1988) 909–917.

[13] S. Montiel, A characterization of hyperbolic cylinders in the de Sitter space, Tôhoku Math. J. 48 (1996) 23–31.

[14] S. Montiel, Uniqueness of space-like hypersurfaces of constant mean curvature in foliated space–times, Math. Ann. 314 (1999) 529–553.

[15] H. Mori, Rotational surfaces in a pseudo-Riemannian 3-sphere, Tôhoku Math. J. 88 (1986) 29–36. [16] H. Omori, Isometric immersions of Riemannian manifolds, J. Math. Soc. Jpn. 19 (1967) 205–214. [17] M. Okumura, Hypersurfaces and a pinching problem on the second fundamental tensor, Am. J. Math. 96

(1974) 207–213.

[18] O. Palmas, Complete rotation hypersurfaces withHkconstant in space forms, Bol. Soc. Bras. Mat. 30 (2)

(1999) 139–161.

[19] S.T. Yau, Harmonic functions on complete Riemannian manifolds, Commun. Pure Appl. Math. 28 (1975) 201–228.

[20] Y. Zheng, On space-like hypersurfaces in the de Sitter space, Ann. Global Anal. Geom. 13 (1995) 317–332. [21] Y. Zheng, Space-like hypersurfaces with constant scalar curvature in the de Sitter space, Diff. Geom. Appl.

Imagem

Fig. 1. The plane ( R, ¯ supH 2 ), where the curve shown is the graph of C n ( R) ¯ corresponding to the hyperbolic cylinders H 1 × S n − 1

Referências

Documentos relacionados

In this paper we give the precise index growth for the embedded hypersur- faces of revolution with constant mean curvature (cmc) 1 in IR n (Delaunay unduloids).. When n = 3, using

Sharp upper bounds for the first positive eigenvalue of the lin- earized operator of the higher order mean curvature of closed hypersurfaces into a Riemannian space form.. Stability

Let M be a compact Riemannian manifold with nonnegative Ricci curvature and o~, a codimension-one C3-foliation of M whose leaves have constant mean curvature..

Erratum to “A gap theorem for complete constant scalar curvature hypersurfaces in the de Sitter

From Remark 6.4 we can see that any two-umbilical immersion has no more that two principal curvatures, and Theorem 4.1 says that the maximum number of principal curvatures of

Alencar, H., do Carmo, M.P.: Hypersurfaces with constant mean curvature in spheres.. Cheng, S.Y., Yau, S.T.: Hypersurfaces with constant

Leite proved the existence of compact embedded rotation hypersurfaces with constant scalar curvature not isometric to products of spheres.. More

By using the nodal domains of some natural function arising in the study of hypersurfaces with constant mean curvature we obtain some Bernstein-type theorems.. Key words: