• Nenhum resultado encontrado

A new characterization of the Euclidean sphere

N/A
N/A
Protected

Academic year: 2019

Share "A new characterization of the Euclidean sphere"

Copied!
8
0
0

Texto

(1)

(Annals of the Brazilian Academy of Sciences)

Printed version ISSN 0001-3765 / Online version ISSN 1678-2690 http://dx.doi.org/10.1590/0001-3765201820170689

www.scielo.br/aabc | www.fb.com/aabcjournal

A new characterization of the Euclidean sphere

ABDÊNAGO A. BARROS1, CÍCERO P. AQUINO2and JOSÉ N.V. GOMES3 1Departamento de Matemática, Universidade Federal do Ceará, Campus do Pici,

Bloco 914, 60455-760 Fortaleza, CE, Brazil

2Departamento de Matemática, Universidade Federal do Piauí, Campus Universitário Ministro Petrônio Portella, 64049-550 Teresina, PI, Brazil

3Departamento de Matemática, Universidade Federal do Amazonas, Av. General Rodrigo Octávio, 6200, 69080-900 Manaus, AM, Brazil

Manuscript received on September 6, 2017; accepted for publication on April 4, 2018

ABSTRACT

In this paper, we obtain a new characterization of the Euclidean sphere as a compact Riemannian manifold with constant scalar curvature carrying a nontrivial conformal vector field which is also conformal Ricci vector field.

Key words:Conformal vector field, Scalar curvature, Euclidean sphere, Einstein manifold.

INTRODUCTION

In the middle of the last century many geometers tried to prove a conjecture concerning the Euclidean sphere as the unique compact orientable Riemannian manifold(Mn,g)admitting a metric of constant scalar curvatureS and carrying a nontrivial conformal vector fieldX. Among them, we cite Bochner and Yano (1953), Goldberg and Kobayashi (1962), Lichnerowicz (1955), Nagano and Yano (1959), Obata (1962), Obata and Yano (1965, 1970) and Tashiro (1965). The attempts to prove this conjecture resulted into the rich literature which has currently been attracting a lot of attention in the mathematical community. We address the reader to the book of Yano (1970) for a summary of those results. Ejiri gave a counter example to this conjecture building metrics of constant scalar curvature on the warped productS1×fFn−1, where

Fn−1 is a compact Riemannian manifold of constant scalar curvature, while S1 stands for the Euclidean circle. In his example the conformal vector field isX= f(d/dt), whered/dtis a unit vector field onS1and

f satisfies a certain ordinary differential equation, see Ejiri (1981) for details.

The primary concept involved in the study of this subject is of Lie derivatives. After all, what is the geometric meaning of the Lie derivative of a tensorT (or of a vector fieldY) with respect to a vector field

(2)

X?This is a method that uses the flow ofX to push values ofT back to pand then differentiate. The result is called the Lie derivativeLXTofTwith respect toX. A vector fieldXon Riemannian manifold(Mn,g)is called conformal ifLXgis a multiple ofg. There are important applications of Lie derivatives in the study of how geometric objects such as Riemannian metrics, volume forms, and symplectic forms behave under flows. For instance, it is well-known that the Lie derivative of a vector fieldY with respect toX is zero if and only ifY is invariant under the flow ofX. It turns out that the Lie derivative of a tensor has exactly the same interpretation. For more details see the book of Lee (2003).

We highlight that Nagano and Yano (1959) have proved that the aforementioned conjecture is true when (Mn,g)is an Einstein manifold, i.e., the Ricci tensor of metricgsatisfiesRic=S

ng. In this case,Sis constant for dimensionsn≥3. Thus ifX is the conformal vector field with conformal factorρ,that is,LXg=2ρg, we deduce LXRic=2ρRic.With this setting we define aconformal Ricci vector fieldon a Riemannian manifold(Mn,g)as a vector fieldX satisfying

LXRic=2βRic, (1)

for some smooth functionβ:MnR.In particular, on Einstein manifolds this concept is equivalent to the classical conformal vector field. With this additional condition the aforementioned conjecture is true. More precisely, we have the following theorem.

Theorem 1. Let(Mn,g),n3,be a compact orientable Riemannian manifold with constant scalar

curva-ture carrying a nontrivial conformal vector fieldXwhich is also a conformal Ricci vector field. ThenMnis

isometric to a Euclidean sphereSn(r). Moreover, up to a rescaling, the conformal factorρis given by

ρ=τ−hv n

andX is the gradient of the Hodge-de Rham function which in this case, up to a constant, is the function

1

nhv, wherehv is a height function on a unitary sphereS

nandτis an appropriate constant.

We point out that Obata and Yano (1965) have obtained the same conclusion of the preceding theorem under the hypothesisLXRic=αg,for some smooth functionα defined inMn.Moreover, whenMn is an Einstein compact Riemannian manifold the result of Nagano and Yano (1959) is a consequence of Theorem 1. We also observe that the compactness ofMnin our result is an essential hypothesis. In fact, let us consider the hyperbolic spaceHn as a hyperquadric of the Lorentz-Minkowski space Ln+1 andv a nonzero fixed vector inLn+1. After a straightforward computation, it is easy to verify that the orthogonal projectionv⊤of

vonto the tangent bundleTHnprovides a nontrivial conformal vector field onHnfor an appropriate choice ofv. Consequently, sinceHnis Einstein, it follows thatv⊤is also a nontrivial conformal Ricci vector field onHn.

PRELIMINARIES AND AUXILIARY RESULTS

To start with, we consider the Hilbert-Schmidt norm for tensors on a Riemannian manifold(Mn,g), i.e., the inner producthT,Si=tr T S.It is important to notice that for an orthonormal basis{e1, . . . ,en}, we can use the natural identification of(0,2)-tensors with(1,1)-tensors,T(ei,ej) =g(Tei,ej), to write

hT,Si=

i,j

Ti jSi j=

i

(3)

We recall that the divergence of a(1,r)-tensorT onMnis the(0,r)-tensor given by

(divT)(v1, . . . ,vr)(p) =tr w7→(∇wT)(v1, . . . ,vr)(p)

,

wherep∈Mnand(v1, . . . ,vr)∈TpM×. . .×TpM.

LetX be a smooth vector field onMn, and letϕbe its flow. For anyp∈Mn, iftis sufficiently close to zero,ϕt is a diffeomorphism from a neighborhood ofpto a neighborhood ofϕt(p), soϕt∗pulls back tensors atϕt(p)to ones at p.

Given a(0,r)-tensorT onMn, the Lie derivative ofT with respect toX is the(0,r)-tensorL

XT given by

(LXT)p= d dt

t=0(ϕ

tT)p=lim t→0 1

t

ϕt∗(Tϕ(t,p))−Tp

.

Fortunately, there is a simple formula for computing the Lie derivative without explicitly finding the flow. For any(0,r)-tensorT onMn,

(LXT)(Y1, . . . ,Yr) = X(T(Y1, . . . ,Yr))−T([X,Y1],Y2, . . . ,Yr)−. . .

−T(Y1, . . . ,Yr−1,[X,Yr]),

whereY1, . . . ,Yrare any smooth vector fields onMn, and[X,Yi]stands for the Lie bracket ofX andYi. In what follows we prove some lemmas and integral formulas which will be required later.

Lemma 1. For any symmetric(0,2)-tensorT on a Riemannian manifold(Mn,g)andXX(M), holds

LX|T|2=2hLXT,Ti −2hT2,LXgi.

In particular, ifXis a conformal vector field with conformal factorρ,then we haveLX|T|2=2hLXT,Ti −

4ρ|T|2.

Proof.Let{e1, . . . ,en}be a geodesic orthonormal frame at p∈Mn. Then we have

LX|T|2 = LX

i,j

Ti jTi j

=2

i,j

Ti jX(Ti j)

= 2

i,j

Ti j{(LXT)i j+T([X,ei],ej) +T(ei,[X,ej])}

= 2

i,j

Ti j(LXT)i j2

i,j

Ti j{T(∇eiX,ej) +T(ei,∇ejX)}.

Whence, we use thatT is symmetric to deduce

LX|T|2 = 2hT,LXTi −2

i,j

Ti j{g(∇eiX,Tej) +g(Tei,∇ejX)}

= 2hT,LXTi −

j

2g(Te

jX,Tej)−

i

2g(Tei,Te iX)

= 2hT,LXTi −2

i

(LXg)(Tei,Tei)

= 2hT,LXTi −2

i,j

Ti j2(LXg)i j,

(4)

Corollary 1. Under the assumptions of Lemma 1 we have LX|T|2=0,provided that LXg=2ρg and

LXT =2ρT.

Another useful result is given by the following lemma.

Lemma 2. Let(Mn,g)be a Riemannian manifold endowed with a symmetric(0,2)-tensorT.Then it holds

XhT,gi=hLXT,gi − hT,LXgi.

In particular, ifLXT =2βT,then

2βtr(T) =h(tr(T)),Xi+hT,LXgi.

Proof.Let{e1, . . . ,en}be a geodesic orthonormal frame at a pointp∈Mn.By the symmetry ofT, we have

hLXT,gi=tr(LXT) =

i

X(Tii) +2T(e iX,ei)

= XhT,gi+2hT,Xi

= XhT,gi+hT,LXgi,

that finishes the proof of the lemma.

Now we remind the reader that the traceless tensor of a symmetric(0,2)-tensorT on a Riemannian manifold Mn,gis given by

˚

T =T−tr(T)

n g. (2)

With this setting we prove the next corollary.

Corollary 2. Let(Mn,g)be a Riemannian manifold endowed with a symmetric (0,2)-tensorT such that LXT =2βT, withLXg=2ρg.Then we have:

1. 2 βρ

tr(T) =h(tr(T)),Xi.

2. Iftr(T)is a non null constant, thenβ =ρ andLXT˚ =2ρT˚.

Proof.SinceLXg=2ρgwe have immediately the first item. Now, iftr(T)is a non zero constant, then we haveβ =ρ,which impliesLXT˚ =LXTtr(T)

n LXg=2ρT˚,finishing the proof of the corollary.

Next, we apply the previous results to the Ricci tensor. First of all, given a conformal vector fieldXon a Riemannian manifoldMnsuch thatLXg=2ρg,we have the next well-known formulae, see e.g. Obata and Yano (1970).

LXRic=−(n2)2ρ(∆ρ)g, (3)

LXS=2(n1)∆ρ2Sρ, (4) and

LXG=−(n2)

∇2ρ−1 n(∆ρ)g

, (5)

whereG=Ric−S

ng=Ric˚ .

We claim that ifMnis compact with constant scalar curvatureSandρ is not constant, then equation

(4) allows us to infer thatSis positive. Indeed, sinceρ is not constant S

(5)

Lemma 3. Let (Mn,g) be a compact Riemannian manifold with constant scalar curvature such that LXRic=2βRicandLXg=2ρg. Then we haveβ=ρandLX|G|2=0.

Proof.SinceLXRic=2βRicandS=tr(Ric)is a positive constant we get by Corollary 2 thatβ =ρ and

LXG=2ρG.Therefore, applying Corollary 1 we have LX|G|2

=0, which completes the proof of the lemma.

Taking into account the second contracted Bianchi identity: div(Ric) = 1

2dS and using the identity div(S

ng) =

1

ndSwe obtain the following relation

div(G) =n−2 2n dS. Therefore, we can write

ρdiv(G)(∇ρ) =n−2

4n h∇S,∇ρ 2

i.

The next equation is well-known. For details of a more general case see for example Lemma1 in Barros and Gomes (2013).

div(ρG(∇ρ)) =ρdiv(G)(∇ρ) +ρh2ρ,Gi+G(∇ρ,∇ρ). (6)

SincehG,gi=0, from (5) we have

hLXG,Gi=−(n2)h2ρ,Gi.

Applying Lemma 1 to this identity we obtain

h∇2ρ,Gi=− 1 n−2

1

2LX|G| 2

+2ρ|G|2. (7)

Comparing (6) and (7) we infer

div(ρG(∇ρ)) = n−2

4n h∇S,∇ρ 2

i − 1 n−2

ρ

2LX|G| 2

+2ρ2|G|2

+G(∇ρ,∇ρ). (8)

In what follows we assume that(M,g)is an orientable Riemannian manifold. IfMis not orientable, we take the orientable double coveringM˜ ofMand induce, in the natural manner, the Riemannian metricg˜on

˜

M. Then(M,g)and(M˜,g)˜ have the same local geometry.

As a direct consequence from (8) and Stokes’ Theorem we obtain the following lemma.

Lemma 4. Let (Mn,g) be a compact orientable Riemannian manifold endowed with a conformal vector

fieldX of conformal factorρ, then

Z

M

G(∇ρ,∇ρ)dM= 1 n−2

Z

M

ρ

2LX|G| 2

+2ρ2|G|2dM+n−2 4n

Z

M

ρ2∆SdM.

Before stating our next result we note that|∇2

ρ−∆ρn g|2

=|∇2

ρ|2

−1n(∆ρ)2. Accordingly, the Bochner formula becomes

1

2∆|∇ρ| 2

= G(∇ρ,∇ρ) +S n|∇ρ|

2

+|∇2ρ−∆ρ n g|

2

+1 n(∆ρ)

2

−h∇ρ,∇ Sρ

n−1+

LXS

2(n1)

(6)

where in the last term we use equation (4). Then,

1

2∆|∇ρ| 2

= G(∇ρ,∇ρ) +|∇2ρ−∆ρ n g|

2

−S |∇ρ|

2

+ρ∆ρ

n(n−1)

− 1

2n(n1)

(LXS)∆ρ+nhS,∇ρ2i+nh∇ρ,(LXS)i.

By integration,

Z

M

G(∇ρ,∇ρ) +|∇2ρ−∆ρ n g|

2

+ 1

2n(ρ 2

∆S+ (LXS)∆ρ

dM=0. (9)

In the notation of (2) we have|∇˚2ρ|2

=|∇2

ρ−∆ρn g|2. Comparing Lemma 4 with equation (9) we get the lemma.

Lemma 5. Let(Mn,g) be a compact orientable Riemannian manifold endowed with a conformal vector

fieldXof conformal factorρ, then

1. R M ρ n−2 1 2LX|G|

2

+2ρ|G|2

+|∇˚2ρ|2

+S

2div(ρ∇ρ) +

LXS

2n ∆ρ

dM=0.

2. R M

ρ

n−2hLXG,Gi+| ˚

∇2ρ|2

+S

2div(ρ∇ρ) +

LXS

2n ∆ρ

dM=0.

Proof.First assertion is a direct combination of Lemma 4 and equation (9), while the second one follows from Lemma 1 and the first assertion. Indeed, from this latter lemma we have 1

2LX|G| 2

+2ρ|G|2=hLXG,Gi, which completes our proof.

We are in the right position to prove our main result.

PROOF OF THEOREM 1

Firstly, we observe that we are supposing that there exists a vector fieldXonMnsuch thatLXg=2ρgand

LXRic=2βRic, for some smooth functionsρ andβ onMn, whereρ is non-constant. Consequently, from Lemma 3 we obtainβ =ρ andLX|G|2=0. Secondly, from item (1) of Lemma 5 and equation (4) we get

Z

M

2

n−2ρ 2

|G|2+|∇2ρ+ Sρ n(n−1)g|

2

dM=0.

Taking into account that ρ is non-constant the preceding identity allows us to achieve G = 0 and

∇2

ρ =−n(n−S1)ρg.Therefore, we can apply a classical result due to Obata (1962), for instance, to con-clude thatMnis isometric to a Euclidean sphereSn(r). Moreover,2ρ=∆ρ

n gand∆(∆ρ) + S

n−1∆ρ=0(see equation (4)). Rescaling the metric we can assume thatS=n(n−1). Then we conclude that∆ρ is the first eigenvalue of the unitary sphereSn.Whence, there exists a fixed vectorvSnsuch that∆ρ=hv=1

n∆hv. Thus we have∆(ρ+1nhv) =0,which givesρ=τ1

nhv. Settingu=−ρwe obtain

Lug=22u=22ρ=2ρg=LXg.

It is also true that LuRic=2ρRic. Besides, by Hodge-de Rham decomposition theorem we can write

(7)

A MORE GENERAL CASE We notice thatLXG=−(n2)˚2ρandL

XG=LXRic−1nLX(Sg)give

LXRic = 1

nLX(Sg)−(n−2)

˚

∇2ρ

= 2ρ S

ng

+1

n(LXS)g−(n−2)

˚

∇2ρ

= 2ρ RicG

+1

n(LXS)g−(n−2)

˚

∇2ρ

= 2ρRic+1

n(LXS)g+T, whereT =−2ρG(n2)˚2ρ.Therefore, we deduce

LXRic=2ρRic+1

n(LXS)g+T, wheretr(T) =0.

Let us suppose thatLXRic=2βRic+T andLXg=2ρg, whereT is a(0,2)-tensor onMn. By using (3) and (4) we deduce

tr(T) =2(n1)∆ρ2βS and

LXS=tr(T) +2(βρ)S. In particular, ifSis a non null constant andρ6=0, we have

tr(T) =0 if and only if β =ρ. (10)

On the other hand,LXG=LXRic1

nLX(Sg)gives

LXG=2βG+Ttr(T)

n g (11)

and

LX|G|2=4(βρ)|G|2+2hT,Gi. (12) Lemma 6. Let(Mn,g),n3, be a compact orientable Riemannian manifold endowed with a conformal

vector fieldX whose conformal factor is ρ.If LXRic=2βRic+T, then the following integral formula holds:

Z

M

1

n−2 2β ρ|G| 2

+ρhT,Gi

+|∇2ρ−∆ρ n g|

2

+S

2div(ρ∇ρ) +

LXS

2n ∆ρ

dM=0.

Proof.First we notice that from (12) we obtain

ρ

2LX|G| 2

+2ρ2|G|2=2β ρ|G|2+ρhT,Gi. (13)

Therefore, using (13) in the first assertion of Lemma 5, we have Z

M

1

n−2 2β ρ|G| 2

+ρhT,Gi

+|∇2ρ−∆ρ n g|

2

+S

2div(ρ∇ρ) +

LXS

2n ∆ρ

dM=0

which completes the proof of the lemma.

(8)

Theorem 2. Let(Mn,g),n3, be a compact orientable Riemannian manifold with constant scalar

cur-vatureS. Suppose that there exists a nontrivial conformal vector fieldX onMnsuch thatLXg=2ρgand

LXRic=2βRic+T. Iftr(T) =0andR M

2ρ2

|G|2

+ρhG,TidM≥0, thenMnis isometric to a Euclidean

sphere.

Moreover, up to a rescaling, the conformal factorρ is given byρ =τ−hv

n andX is the gradient of

the Hodge-de Rham function which in this case, up to a constant, is the function 1

nhv, wherehvis a height

function on a unitary sphereSnandτis an appropriate constant.

Proof.It follows from (10) thatβ=ρ. Now we may use Lemma 6 to deduce that|∇2

ρ−∆ρn g|2

=0.But, from (4) we have∆ρ=− S

n−1ρ.Therefore, we deduce∇ 2

ρ=− S

n(n−1)ρg, from which we may apply Obata’s Theorem, consult Obata (1962), to conclude thatMnis isometric to a Euclidean sphere. Moreover, from (11) we have thatT is null tensor. So, the latter claim is proved following the same steps given in the proof of Theorem 1.

Remark 1. Notice that if T =λ ρG, with (2+λ) 0, then the conditions of the previous theorem are verified. In particular, for λ =−2, we have LXRic=2ρS

ng, which allows us to obtain the result due to

Obata and Yano (1965).

ACKNOWLEDGMENTS

The authors would like to thank the referees for their careful reading and useful comments which improved the paper. José N.V. Gomes would like to thank the Department of Mathematics at Lehigh University, where part of this work was carried out. He is grateful to Huai-Dong Cao and Mary Ann for their warm hospitality and constant encouragement. The authors are partially supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), do Ministério da Ciência, Tecnologia e Inovação (MCTI) do Brasil.

REFERENCES

BARROS A AND GOMES JN. 2013. A Compact gradient generalized quasi-Einstein metric with constant scalar curvature. J Math Anal Appl 401: 702-705.

BOCHNER S AND YANO K. 1953. Curvature and Betti Numbers. Ann Math Study. Princeton University Press, Princeton, New Jersey, US.

EJIRI N. 1981. A negative answer to a conjecture of conformal transformations of Riemannian manifolds. J Math Soc Japan 33: 261-266.

GOLDBERG SI AND KOBAYASHI S. 1962. The conformal transformation group of a compact Riemannian manifold. Amer J Math 84: 170-174.

LEE JM. 2003. Introduction to smooth Manifolds. Springer-Verlag, New York, US.

LICHNEROWICZ A. 1955. Transformations infinitésimales conformes de certaines variétés riemannienne compacte. CR Acad Sci Paris 241: 726-729.

NAGANO T AND YANO K. 1959. Einstein spaces admitting a one-parameter group of conformal transformations. Ann Math 69: 451-461.

OBATA M. 1962. Certain conditions for a Riemannian manifold to be isometric with a sphere. J Math Soc Japan 14: 333-340. OBATA M AND YANO K. 1965. Sur le groupe de transformations conformes d’une variété de Riemann dont le scalaire de courbure

est constant. CR Acad Sci Paris 260: 2698-2700.

Referências

Documentos relacionados

i) A condutividade da matriz vítrea diminui com o aumento do tempo de tratamento térmico (Fig.. 241 pequena quantidade de cristais existentes na amostra já provoca um efeito

Como salientado no início da dissertação, o objectivo do presente trabalho era o de efectuar o levantamento do estado em que se encontrava o reactor MPCVD do Centro de Tecnologia

Ousasse apontar algumas hipóteses para a solução desse problema público a partir do exposto dos autores usados como base para fundamentação teórica, da análise dos dados

didático e resolva as ​listas de exercícios (disponíveis no ​Classroom​) referentes às obras de Carlos Drummond de Andrade, João Guimarães Rosa, Machado de Assis,

[r]

O soro dos animais vacinados com lipossomo, EBS e proteolipossomos foram coletados semanalmente antes e após a infecção experimental para a detecção da produção de anticorpos IgG,

Despercebido: não visto, não notado, não observado, ignorado.. Não me passou despercebido

gulbenkian música mecenas estágios gulbenkian para orquestra mecenas música de câmara mecenas concertos de domingo mecenas ciclo piano mecenas coro gulbenkian. FUNDAÇÃO