数学杂志  2014, Vol. 34 Issue (2): 251-258   PDF    
扩展功能
加入收藏夹
复制引文信息
加入引用管理器
Email Alert
RSS
本文作者相关文章
ZENG Fan-qi
MA Bing-qing
THE CLASSIFICATION OF GRADIENT RICCI ALMOST SOLITONS
ZENG Fan-qi, MA Bing-qing    
Department of Mathematics, Henan Normal University, Xinxiang 453007, China
Abstract: We study the classification of a gradient Ricci almost soliton. Using similar methods as in [11] for n ≥ 5, we obtain that the Weyl curvature tensor is harmonic or Einstein under the assumption that the Bach tensor is flat.
Key words: gradient Ricci almost solitons     Bach tensor     Weyl curvature tensor    
近黎奇梯度孤立子的分类
曾凡奇, 马冰清    
河南师范大学数学系, 河南 新乡 453007
摘要:本文研究黎奇梯度孤立子的分类问题.利用与文献[11]类似的方法, 在Bach张量等于零的条件下, 对于n ≥ 5, 证明了流形是Einstein的或者Weyl曲率张量是调和的.
关键词黎奇梯度孤立子    Bach张量    Weyl曲率张量    
1 Introduction

Let $(M^n,g)$ be an $n$-dimensional Riemannian manifold. If there exist two smooth functions $f$, $\lambda$ on $(M^n,g)$ such that

$\begin{equation} R_{ij}+f_{ij}=\lambda g_{ij}, \end{equation}$ (1.1)

then $(M^n,g)$ is called a gradient Ricci almost soliton which was introduced by Pigola, Rigoli, Rimoldi and Setti in [1], where $R_{ij}$ denotes the Ricci curvature of $(M^n,g)$. Clearly, the above gradient Ricci almost solitons generalize the concept of gradient Ricci solitons which play a very important role in Hamilton's Ricci flow as it corresponds to the self-similar solutions and often arises as singularity models, for a survey in this subject we refer to the work due to Cao in [2]. When $\lambda=\rho R+\mu$ in (1.1) with $\rho,\mu$ two real constants, $(M^n,g)$ is called the gradient $\rho$-Einstein soliton (see [3]) which is a special case of $(m,\rho)$-quasi-Einstein manifolds defined in [4], where $R$ is the scalar curvature of $(M^n,g)$. For the recent research on this direction, see [5-10] and the references therein.

In this paper, using a similar idea used in [11-13], we derive some formulas, and establish a link between the Cotton tensor $C_{ijk}$ and the 3-tensor $D_{ijk}$, that is, $C_{ijk}=D_{ijk}-W_{ijkl}f^l$, where $W_{ijkl}$ is the Weyl curvature tensor. By virtue of this relationship we give some classifications of gradient Ricci almost solitons.

2 Preliminaries

We use moving frames in all calculations and adopt the following index convention:

$1\leq i,j,k,\cdots\leq n,\ \ \ \ \ \ 2\leq a,b,c,\cdots\leq n$

throughout this paper.

Lemma 2.1 Let $(M^n,g)$ be a gradient Ricci almost soliton satisfying (1.1). Then we have

$\begin{align} &\Delta f=n\lambda-R, \end{align}$ (2.1)
$\begin{align} &(|\nabla f|^2)_{i}=2\lambda f_i-2R_{ij}f^j, \end{align}$ (2.2)
$\begin{align} &\frac{1}{2}R_{,i}=(n-1)\lambda_i+R_{ij}f^j, \end{align}$ (2.3)

where $f^j=g^{jk}f_k$.

Proof Equations (2.1) and (2.2) are direct consequences of (1.1) and the fact

$(|\nabla f|^2)_i=2f_jf_{ij}=2f_j(\lambda g_{ij}-R_{ij})=2\lambda f_i-2R_{ij}f_j.$

By the second Bianchi identity, we get

$\begin{align*} \frac{1}{2}R_{,i}=R_{ij,j}=&(\lambda g_{ij}-f_{ij})_{,j}\\ =&\lambda_i-f_{ijj}\\ =&\lambda_i-(\Delta f)_{i}-R_{ij}f_j, \end{align*}$

where we used the Ricci identity $f_{ijj}=(\Delta f)_{i}+R_{ij}f_j$. Insertting (2.1) and (2.2) into the above equation gives (2.3). We complete the proof of Lemma 2.1.

For $n\geq3$, the Weyl curvature tensor and the Cotton tensor are defined by

$\begin{aligned} W_{ijkl}=&R_{ijkl}-\frac{1}{n-2}(A_{ik}g_{jl}-A_{il}g_{jk} +A_{jl}g_{ik}-A_{jk}g_{il})\\ =&R_{ijkl}-\frac{1}{n-2}(R_{ik}g_{jl}-R_{il}g_{jk} +R_{jl}g_{ik}-R_{jk}g_{il})\\ &+\frac{R}{(n-1)(n-2)}(g_{ik}g_{jl}-g_{il}g_{jk}) \end{aligned}$ (2.4)

and

$\begin{equation}C_{ijk}=A_{kj,i}-A_{ki,j},\end{equation}$ (2.5)

where $A_{ij}$ is called the Schouten tensor given by

$A_{ij}=R_{ij}-\frac{R}{2(n-1)}g_{ij}.$

From the definition of the Cotton tensor, we have that $C_{ijk}$ is skew-symmetric in the first two indices and trace-free in any two indices:

$C_{ijk}=-C_{jik},\qquad g^{ij}C_{ijk}=g^{ik}C_{ijk}=0.$

The divergence of the Weyl curvature tensor is related to the Cotton tensor by

$\begin{equation} -\frac{n-3}{n-2}C_{ijk}=W_{ijkl,}{}^{l}. \end{equation}$ (2.6)

For $n\geq4$, the Bach tensor is defined by

$\begin{equation} B_{ij}=\frac{1}{n-3}W_{ikjl,}{}^{lk}+\frac{1}{n-2}W_{ikjl}R^{kl}. \end{equation}$ (2.7)

Using (2.6), we may extend the definition of Bach tensor in dimensions including 3 as follows:

$\begin{equation} B_{ij}=\frac{1}{n-2}(C_{kij,}{}^{k}+W_{ikjl}R^{kl}). \end{equation}$ (2.8)

As in [11], see also [8, 12, 13], we define the following 3-tensor $D$ by

$\begin{aligned} D_{ijk}=&\frac{1}{n-2}(R_{kj}f_i-R_{ki}f_j)+\frac{1}{(n-1)(n-2)} (R_{il}g_{jk}f^l-R_{jl}g_{ik}f^l)\\ &-\frac{R}{(n-1)(n-2)}(g_{kj}f_i-g_{ki} f_j). \end{aligned}$ (2.9)

Then we have that $D_{ijk}$ is skew-symmetric in the first two indices and trace-free in any two indices:

$D_{ijk}=-D_{jik},\qquad g^{ij}D_{ijk}=g^{ik}D_{ijk}=0.$

Lemma 2.2 Let $(M^n,g)$ be a gradient Ricci almost soliton satisfying (1.1). Then the Cotton tensor, D-tensor and the Weyl curvature tensor are related by

$\begin{equation} C_{ijk}=D_{ijk}-W_{ijkl}f^l. \end{equation}$ (2.10)

proof Using formula (1.1), we have

$\begin{aligned} R_{kj,i}-R_{ki,j} =&(\lambda g_{kj}-f_{kj})_{,i} -(\lambda g_{ki}-f_{ki})_{,j}\\ =&\lambda_{i}g_{kj}-\lambda_{j}g_{ki}+f_{kij}-f_{kji}\\ =&\lambda_{i}g_{kj}-\lambda_{j}g_{ki}-R_{ijkl}f^l. \end{aligned}$

Therefore,

$\begin{aligned} C_{ijk}=&A_{kj,i}-A_{ki,j}\\ =&\lambda_{i}g_{kj}-\lambda_{j}g_{ki}-R_{ijkl}f^l\\ &-\frac{1}{2(n-1)}(R_{,i}g_{jk}-R_{,j}g_{ik})\\ =&-\frac{1}{(n-1)}(R_{il}g_{kj}f_l-R_{jl}g_{ki}f_l)-R_{ijkl}f^l\\ =&D_{ijk}-W_{ijkl}f^l, \end{aligned}$

where the third equality used equation (2.3). It completes the proof of Lemma 2.2.

The next lemma links the norm of $D_{ijk}$ to the geometry of the level surfaces of the function $f$ on $(M^n,g)$. The proof can be found in [15, Proposition 2.3] and [11, Proposition 3.1].

Lemma 2.3 Let $(M^n,g)$ be a Riemannian manifold and let $\Sigma_c=\{x|f(x)=c\}$ be the level surface with respect to regular value $c$ of $f$. Choose local orthonormal frame $\{e_1, e_2, \cdots, e_n\}$ on $(M^n,g)$ such that $e_1=\nabla f/|\nabla f|$ and $\{e_2, \cdots, e_n\}$ tangent to $\Sigma_c$. Denote by $|D_{ijk}|$ the norm of the $3$-tensor $D$, and by $g_{ab}$ the induced metric on $\Sigma_c$. We have

$\begin{equation*} |D_{ijk}|^2=\frac{2|\nabla f|^2}{(n-1)(n-2)^2}\biggl((n-2)\sum\limits_{a=2}^nR_{1a}^2 +(n-1)\Big|R_{ab}-\frac{R-R_{11}}{n-1}g_{ab}\Big|^2\biggr), \end{equation*}$

where $R_{ij}={\rm Ric}(e_i,e_j)$ are the components of the Ricci curvature on $(M^n,g)$, $R$ is the scalar curvature of $(M^n,g)$. Note that the indices $2\leq a,b,c,\cdots\leq n$, then $R_{ab}$ denotes the Ricci tensor of $(M^n,g)$ restricted to the tangent space of $\Sigma_c$ and $g^{ab}R_{ab}=R-R_{11}$.

3 Some Results

With the help of Lemma 2.3, we can obtain the following result.

Proposition 3.1 Let $(M^n,g)$ be a gradient Ricci almost soliton satisfying (1.1) with $D_{ijk}=0$. Let $\Sigma_c=\{x|f(x)=c\}$ be the level surface with respect to regular value $c$ of $f$. Then for any local orthonormal frame $\{e_1, e_2, \cdots, e_n\}$ with $e_1=\nabla f/|\nabla f|$ and $\{e_2, \cdots, e_n\}$ tangent to $\Sigma_c$, we have

(1) $|\nabla f|$, $\Delta f$, $\lambda$ and the scalar curvature $R$ of $(M^n, g)$ are all constant on $\Sigma_c$;

(2) $R_{1a}=0$ and $e_1=\nabla f /|\nabla f |$ is an eigenvector of the Ricci operator;

(3) the second fundamental form $h_{ab}$ of $\Sigma_c$ is of the form $h_{ab}=\frac{H}{n-1} g_{ab}$;

(4) the mean curvature $H=\frac{(n-1)\lambda-(R-R_{11})}{|\nabla f|}$ is constant on $\Sigma_c$;

(5) on $\Sigma_c$ the Ricci tensor of $(M^n, g)$ either has a unique eigenvalue $\nu$, or has two distinct eigenvalues $\nu$ and $\sigma$ of multiplicity $1$ and $n-1$ respectively. In either case, $e_1=\nabla f /|\nabla f |$ is an eigenvector of $\nu$. Moreover, both $\nu$ and $\sigma$ are constant on $\Sigma_c$.

Proof Under this chosen orthonormal frame, we have $f_1=|\nabla f|$ and $f_2=f_3=\cdots=f_n=0$. When $D_{ijk}=0$, we have from Lemma 2.3 that

$\begin{equation} R_{1a}=0 \end{equation}$ (3.1)

and

$\begin{equation} R_{ab}=\frac{R-R_{11}}{n-1}g_{ab}. \end{equation}$ (3.2)

Therefore, we obtain from (2.2) and (2.3)

$(|\nabla f|^2)_{a}=0,\ \ \ \ \forall\ a,$

which show that $|\nabla f|$ is constant on $\Sigma_c$. We derive form (2.2) and (2.3)

$R_{,i}=2(n-1)\lambda_{i}+2\lambda f_{i}-(|\nabla f|^2)_{i}$

which means that

$\begin{equation} dR=2(n-1)d\lambda+2\lambda df-d(|\nabla f|^2). \end{equation}$ (3.3)

Taking exterior differential of the both sides of (3.3), we obtain $d\lambda\wedge df=0$. Therefore, according to the well-known Cartan's lemma, there exists a smooth function $\varphi$ such that

$d\lambda=\varphi\, df,$

which shows that $\lambda$ is also constant on $\Sigma_c$. Hence, (1) is proved.

In particular, (2) can be obtained from (3.1) directly.

By the definition of $h_{ab}$, we have

$\begin{equation} h_{ab}=\langle\nabla_{e_a}\Big(\frac{\nabla f}{|\nabla f|}\Big),e_b\rangle =\frac{1}{|\nabla f|}f_{ab}=\frac{1}{|\nabla f|}\Big(\lambda-\frac{R-R_{11}}{n-1}\Big)g_{ab},\end{equation}$ (3.4)

where the last equality used (3.2). Hence,

$\begin{equation} H=g^{ab}h_{ab} =\frac{(n-1)\lambda-(R-R_{11})}{|\nabla f|}\end{equation}$ (3.5)

and (3) is proved.

By the Codazzi equation

$R_{1cab}=\nabla_a^{\Sigma_c}h_{bc}-\nabla_b^{\Sigma_c}h_{ac},$

we get from tracing over $b$ and $c$

$R_{1a}=\nabla_a^{\Sigma_c}H-\nabla_b^{\Sigma_c}h_{ab}=\frac{n-2}{n-1}H_{,a}$

and (4) follows form $R_{1a}=0$.

Since $H$ is constant on $\Sigma_c$, we have from (3.5)

$R_{11,a}=0.$

Applying

$R_{11,a}=e_a(R_{11})-2R(\nabla_{e_a}e_1,e_1) =e_a(R_{11})-2h_{ab}R_{1b}=e_a(R_{11})$

yields $e_a(R_{11})=0$, which shows that $\nu=R_{11}$ is constant on $\Sigma_c$. By (3.2) we know that for distinct $a$, the eigenvalues of $R_{aa}$ are the same. Hence, we have the eigenvalue $\sigma$ is also constant. We obtain (5) and complete the proof of Proposition 3.1.

Theorem 3.2 Let $(M^n,g)$ be a gradient Ricci almost soliton satisfying (1.1). Then

$\begin{equation}(n-2)B_{ij}=D_{kij,}{}^{k} +\frac{n-3}{n-2}C_{kji}f^k.\end{equation}$ (3.6)

If $(M^n,g)$ is compact, then for $p\geq0$,

$\begin{equation} \int\limits_{M^n}f^pB_{ij}f^if^j\,dv_g=-\frac{1}{2}\int\limits_{M^n}f^p|D|^2\,dv_g. \end{equation}$ (3.7)

In particular, if $B_{ij}=0$, we obtain from (3.7) the 3-tensor $D_{ijk}=0$.

Proof By virtue of (2.8) and (2.10), we have

$\begin{aligned} (n-2)B_{ij}=&C_{kij,k}+W_{ikjl}R_{kl}\\ =&(D_{kij}-W_{kijl}f_l)_{,k}+W_{ikjl}R_{kl}\\ =&D_{kij,k}-W_{kijl,k}f_l-W_{kijl}f_{kl}+W_{ikjl}R_{kl}\\ =&D_{kij,k}-W_{kijl,k}f_l\\ =&D_{kij,k}+\frac{n-3}{n-2}C_{lji}f_l. \end{aligned}$

If $(M^n,g)$ is compact, we obtain using integrating by parts

$\begin{aligned} &(n-2)\int\limits_{M^n}f^pB_{ij}f_if_j\,dv_g\\ =&\int\limits_{M^n}f^p\biggl(D_{kij,}{}^{k} +\frac{n-3}{n-2}C_{kji}f^k\biggr)f_if_j\,dv_g\\ =&\int\limits_{M^n}f^pD_{kij,}{}^{k}f_if_j\,dv_g\\ =&-\int\limits_{M^n}f^pD_{kij}f^if^{kj}\,dv_g\\ =&-\frac{n-2}{2}\int\limits_{M^n}f^p|D|^2\,dv_g. \end{aligned}$

Therefore, we obtain (3.7) and complete the proof of Theorem 3.2.

Proposition 3.3 Let $(M^n,g)$ be a compact gradient Ricci almost soliton satisfying (1.1) with $B_{ij}=0$. If $n\geq4$, then the Cotton tensor $C_{ijk}=0$ at all points where $\nabla f\neq0$.

Proof From Lemma 2.2 and Theorem 3.2, we conclude that $C_{ijk}=-W_{ijkl}f_l$. Under the orthonormal frame as in Lemma 2.3, we have

$\begin{equation}C_{ijk}=-W_{ijk1}|\nabla f|.\end{equation}$ (3.8)

In particular, we obtain from (3.8)

$\begin{equation}C_{ij1}=0.\end{equation}$ (3.9)

From Theorem 3.2, we get

$\frac{n-3}{n-2}C_{1ji}|\nabla f|=0.$

Hence, If $n\geq4$, then

$\begin{equation}C_{1ji}=C_{j1i}=0.\end{equation}$ (3.10)

Moreover, from (3.8) we also have that $C_{abc}=-W_{abc1}|\nabla f|$. Using (2.4) and Proposition 3.1, we obtain

$W_{abc1}=R_{abc1}=R_{1cba}=\nabla^\Sigma_{e_b}h_{ac}-\nabla^\Sigma_{e_a}h_{bc}=0. $ (3.11)

Therefore, we obtain

$\begin{equation}C_{abc}=0.\end{equation}$

Combining (3.9) with (3.10) and (3.11), we arrive at the conclusion of Proposition 3.3.

Proposition 3.4 Let $(M^4,g)$ be a compact gradient Ricci almost soliton satisfying (1.1). If $B_{ij}=0$, then the Weyl curvature tensor $W_{ijkl}=0$ at all points where $\nabla f\neq0$.

Proof Since $B_{ij}=0$, we have $D_{ijk}=C_{ijk}=0$. Hence, Lemma 2.2 shows that $W_{ijk1}=0$ for $1\leq i,j,k\leq 4$. It remains to show that $W_{abcd}=0$ for $2\leq a,b,c,d\leq 4$. This essentially reduces to show the Weyl curvature tensor is equal to zero in 3 dimensions (see [14, p.276-277] or [11, p.13]). Therefore, we have $W_{ijkl}=0$.

Theorem 3.5 Let $(M^n,g)$ be a compact gradient Ricci almost soliton satisfying (1.1) with $B_{ij}=0$.

(1) If $n\geq5$, then the Weyl curvature tensor is harmonic or Einstein.

(2) If $n=4$ and it has positive sectional curvature, then $(M^4,g)$ is rotational symmetric or Einstein.

Proof (1) If $(M^n,g)$ is not Einstein, then from the set $\{p|\nabla f(p)=0\}$ is of measure zero we have $C_{ijk}=0$ on $\Omega=\{x|\nabla f\neq0\}$ everywhere according to Proposition 3.3 and the continuity. Hence, the Weyl curvature tensor is harmonic.

(2) Under the assumption of Theorem 3.1, Proposition 3.4 shows that $(M^4,g)$ has vanishing Weyl curvature tensor at all points where $\nabla f\neq 0$. So if the set $\Omega=\{x|\nabla f\neq0\}$ is dense, by continuity of the Weyl curvature tensor we have $W_{ijkl}=0$ everywhere and $(M^4,g)$ is locally conformally flat. Recall that in any neighborhood of the level surface $\Sigma_c$, where $\nabla f\neq 0$, we can express the metric $ds^2$ by

$\begin{equation} ds^2=\frac{1}{|\nabla f|^2}(f,\theta)df^2+g_{ab}(f,\theta)d\theta^a\theta^b, \end{equation}$ (3.12)

where $\theta=(\theta^2,\cdots,\theta^n)$ denote the intrinsic coordinates on $\Sigma_c$. Since $(M^4,g)$ has vanishing Weyl curvature tensor and positive sectional curvature, the Gauss equation

$R^{\Sigma_c}_{abcd}=R_{abcd}+h_{aa}h_{bb}-h_{ab}^2$

and Proposition 3.1 tells us that $(\Sigma_c,g_{ab})$ is a space form with constant positive sectional curvature and $\frac{1}{|\nabla f|}(f,\theta)=\frac{1}{|\nabla f|}(f)$. Hence on $\Omega$ we have

$\begin{equation} ds^2=\frac{1}{|\nabla f|^2}(f)df^2+\varphi^2(f)g_{_{\mathbb{S}^{n-1}}}, \end{equation}$ (3.13)

where $g_{\mathbb{S}^{n-1}}$ denotes the standard metric on unit sphere $\mathbb{S}^{n-1}$. We conclude that $(M^4,g)$ is rotationally symmetric.

References
[1] Pigola S, Rigoli M, Rimoldi M, Setti A. Ricci almost solitons[J]. Ann. Sc. Norm. Super. Pisa Cl. Sci., 2011, 10(4): 757–799.
[2] Cao Huaidong. Recent progress on Ricci soliton[J]. Adv. Lect. Math. (ALM), 2009, 11: 1–38.
[3] Catino G, Mazzieri L. Gradient Einstein solitons[J]. arXiv: 1201. 6620.
[4] Huang Guangyue, Wei Yong. The classification of $(m,\rho)$-quasi-Einstein manifolds[J]. Ann. Glob. Anal. Geom., 2013, 44(3): 269–282. DOI:10.1007/s10455-013-9366-0
[5] Barros A, Ribeiro Jr E. Some characterizations for compact almost Ricci solitons[J]. Proc. Amer. Math. Soc., 2012, 140(3): 1033–1040. DOI:10.1090/S0002-9939-2011-11029-3
[6] Barros A, Gomes J. A compact gradient generalized quasi-Einstein metric with constant scalar curvature[J]. J. Math. Anal. Appl., 2013, 401(2): 702–705. DOI:10.1016/j.jmaa.2012.12.068
[7] Case J, Shu Y J, Wei G. Rigidity of quasi-Einstein metrics[J]. Diff. Geom. Appl., 2011, 29(1): 93–100. DOI:10.1016/j.difgeo.2010.11.003
[8] Chen Qiang, He Chenxu. On Bach flat warped product Einstein manifolds[J]. Pacific J. Math., 2013, 265(2): 313–326. DOI:10.2140/pjm
[9] He Chenxu, Petersen P, Wylie W. On the classification of warped product Einstein metrics[J]. Comm. Anal. Geom., 2012, 20: 271–312. DOI:10.4310/CAG.2012.v20.n2.a3
[10] Ma Bingqing, Yau Suxia. Yamabe solitons on compact Riemannian manifolds[J]. J. Henan Norm. Univ. Nat. Sci., 2012, 40(6): 12–13.
[11] Cao Huaidong, Chen Qiang. On Bach flat gradient shrinking Ricci solitons[J]. Duke Math. J., 2013, 162(6): 1149–1169. DOI:10.1215/00127094-2147649
[12] Cao Huaidong, Catino G, Chen Qiang. Mantegazza C, Mazzieri L. Bach-flat gradient Ricci solitons[J]. arXiv: 1107. 4591, to appear in Calc. Var. PDE. DOI10. 1007/s00526-012-0575-3
[13] Cao Huaidong, Chen Qiang. On locally conformally flat gradient steady Ricci solitons[J]. Trans. Amer. Math. Soc., 2012, 364(5): 2377–2391. DOI:10.1090/S0002-9947-2011-05446-2
[14] Hamilton R S. Three manifolds with positive Ricci curvature[J]. J. Diff. Geom., 17, 1982, 17(2): 255–306. DOI:10.4310/jdg/1214436922
[15] Huang Guangyue, Li Haizhong. On a classification of the quasi Yamabe gradient solitons[J]. arXiv: 1108. 6177, to appear in Methods and Applications of Analysis.