This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Constant weighted mean curvature hypersurfaces in Shrinking Ricci Solitons

Igor Miranda Igor Miranda
Instituto de Matemática e Estatística, Universidade Federal Fluminense, Campus Gragoatá, Niterói, RJ
24210-200,
Brazil
[email protected]
 and  Matheus Vieira Matheus Vieira
Departamento de Matemática, Universidade Federal do Espírito Santo, Vitória, ES
29075-910
Brazil
[email protected]
Abstract.

In this paper, we study constant weighted mean curvature hypersurfaces in shrinking Ricci solitons. First, we show that a constant weighted mean curvature hypersurface with finite weighted volume cannot lie in a region determined by a special level set of the potential function, unless it is the level set. Next, we show that a compact constant weighted mean curvature hypersurface with a certain upper bound or lower bound on the mean curvature is a level set of the potential function. We can apply both results to the cylinder shrinking Ricci soliton ambient space. Finally, we show that a constant weighted mean curvature hypersurface in the Gaussian shrinking Ricci soliton (not necessarily properly immersed) with a certain assumption on the integral of the second fundamental form must be a generalized cylinder.

Key words and phrases:
level sets, constant weighted mean curvature, f-minimal, shrinking Ricci soliton, λ\lambda-hypersurfaces
2010 Mathematics Subject Classification:
MSC 53C42 and MSC 53C44
This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) Finance Code 001- from September, 2016 to March, 2021

1. Introduction

Let (M¯n+1,g¯,f)\left(\bar{M}^{n+1},\bar{g},f\right) be a smooth metric measure space and let MnM^{n} be a oriented hypersurface in M¯n+1\bar{M}^{n+1}. The weighted mean curvature vector Hf\vec{H}_{f} is defined by

Hf=H+(¯f)\vec{H}_{f}=\vec{H}+\left(\bar{\nabla}f\right)^{\perp}

and the weighted mean curvature HfH_{f} is defined by Hf=HfN\vec{H}_{f}=-H_{f}N, where \perp is the projection to the normal bundle and NN is the unit normal vector. A hypersurface MnM^{n} is said to have constant weighted mean curvature (CWMC) if HfH_{f} is a constant (see Section 2). These hypersurfaces are also known as λ\lambda-hypersurfaces where λ=Hf\lambda=H_{f}. CWMC hypersurfaces can be seen as critical points of the weighted area functional with respect to weighted volume-preserving variations (see [9], [23]). When Hf=0H_{f}=0 they are called ff-minimal hypersurfaces. For certain choices of ff, ff-minimal hypersurfaces are very important singularities of the mean curvature flow in n+1\mathbb{R}^{n+1}, namely self-shrinkers (see p.758 and p.768 in [15]), translating solitons (see p.153 and p.154 in [19]) and self-expanders (see p.9016 and p.9017 in [3]). It turns out that (n+1,g¯can,f)\left(\mathbb{R}^{n+1},\bar{g}_{\text{can}},f\right) is a Ricci soliton for each of these choices of ff (see Section 2). Thus the study of self-shrinkers, translating solitons and self-expanders in n+1\mathbb{R}^{n+1} is equivalent to the study of ff-minimal hypersurfaces in the shrinking, steady and expanding Ricci soliton (n+1,g¯can,f)\left(\mathbb{R}^{n+1},\bar{g}_{\text{can}},f\right) respectively.

There has been a great interest in CWMC hypersurfaces in the Gaussian shrinking Ricci soliton (n+1,g¯can,f)\left(\mathbb{R}^{n+1},\bar{g}_{\text{can}},f\right) with f(x)=|x|24f(x)=\frac{|x|^{2}}{4}. In [23] Mcgonagle-Ross classified stable CWMC hypersurfaces properly immersed in the Gaussian shrinking Ricci soliton. They also proved that there are no CWMC hypersurfaces properly immersed in this ambient space with index one. In [7] Q.M.Cheng-Ogata-Wei classified complete CWMC hypersurfaces in the Gaussian shrinking Ricci soliton with a certain condition on the norm of the second fundamental form and the mean curvature. In [9] Q.M.Cheng-Wei classified complete CWMC hypersurfaces embedded in the Gaussian shrinking Ricci soliton with polynomial volume growth and HHf0H-H_{f}\geq 0. In [17] Guang classified compact CWMC surfaces embedded in the Gaussian shrinking Ricci soliton with Hf0H_{f}\geq 0 and constant norm of the second fundamental form. He also classified complete CWMC hypersurfaces embedded in the Gaussian shrinking Ricci soliton with a certain condition on the norm of the second fundamental form. In [10] Q.M.Cheng-Wei generalized this result to complete CWMC surfaces and removed the condition on HfH_{f}.

There has also been a great interest in f-minimal and CWMC hypersurfaces in the cylinder shrinking Ricci soliton (n+1k×S2(k1)k,g¯,f)\left(\mathbb{R}^{n+1-k}\times S_{\sqrt{2\left(k-1\right)}}^{k},\bar{g},f\right), where g¯\bar{g} is the product metric and f(x,y)=|x|24f\left(x,y\right)=\frac{\left|x\right|^{2}}{4} (here xx is the position vector in nk+1\mathbb{R}^{n-k+1} and yy is the position vector in k+1\mathbb{R}^{k+1}). In [12] X.Cheng-Mejia-Zhou classified compact f-minimal hypersurfaces in a cylinder shrinking Ricci soliton with a certain condition on the norm of the second fundamental form. They also classified compact f-minimal hypersurfaces in this ambient space with index one. In [14] X.Cheng-Zhou generalized these results to complete hypersurfaces. In [2] Barbosa-Santana-Upadhyay obtained theorems for CWMC hypersurfaces similar to the results of Mcgonagle-Ross [23] but in the cylinder shrinking Ricci soliton ambient space. They also obtained theorems for CWMC hypersurfaces in cylinder shrinking Ricci soliton similar to the results of X.Cheng-Mejia-Zhou [12] and X.Cheng-Zhou [14] about LfL_{f}-stable (which are those that the second variation of weighted area is non-negative) ff-minimal hypersurfaces.

In this paper we study geometric properties and classification results for CWMC hypersurfaces in shrinking Ricci solitons.

Let (M¯n+1,g¯,f)\left(\bar{M}^{n+1},\bar{g},f\right) be a gradient Ricci soliton with

R¯ic+¯¯f=λg¯,\bar{R}ic+\bar{\nabla}\bar{\nabla}f=\lambda\bar{g},

where λ\lambda is a constant. We know that

|¯f|2+R¯2λf=C,\left|\bar{\nabla}f\right|^{2}+\bar{R}-2\lambda f=C,

where R¯\bar{R} is the scalar curvature of M¯n+1\bar{M}^{n+1} and CC is a constant (see p.85 in [18] and Lemma 1.1 in [4]). Let us define

D±(c1,c2)={xM¯n+1;f(x)<Γ±(x)},D^{\pm}(c_{1},c_{2})=\left\{x\in\bar{M}^{n+1};f\left(x\right)<\Gamma^{\pm}(x)\right\},

where

Γ±(x)=12λ{14(±c1+c12+4c2)2+R¯(x)C}\Gamma^{\pm}(x)=\frac{1}{2\lambda}\left\{\frac{1}{4}\left(\pm c_{1}+\sqrt{c_{1}^{2}+4c_{2}}\right)^{2}+\bar{R}(x)-C\right\}

and c1c_{1} and c2c_{2} are constants such that c12+4c20c_{1}^{2}+4c_{2}\geq 0.

Note that D+(c1,c2)D^{+}(c_{1},c_{2}) and D(c1,c2)D^{-}(c_{1},c_{2}) are open subsets of M¯n+1\bar{M}^{n+1}. In this paper we will always use the above notations and conventions.

In Theorem 3 in [27], Vieira-Zhou proved that if MnM^{n} is a complete ff-minimal hypersurface properly immersed in the cylinder shrinking Ricci soliton then: (a) MnM^{n} cannot lie inside the closed product B¯2(nk)n+1k(0)×S2(k1)k(0)\bar{B}^{n+1-k}_{\sqrt{2\left(n-k\right)}}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0), unless Mn=S2(nk)nk(0)×S2(k1)k(0)M^{n}=S^{n-k}_{\sqrt{2\left(n-k\right)}}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0) (see also the earlier work of Cavalcante-Espinar [6] for CWMC hypersurfaces in the Gaussian shrinking Ricci soliton n+1\mathbb{R}^{n+1}); (b) MnM^{n} cannot lie outside the product B2(n+1k)n+1k(0)×S2(k1)k(0)B^{n+1-k}_{\sqrt{2\left(n+1-k\right)}}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0). We generalize this result in two different ways: we extend the result to CWMC hypersurfaces and we consider an arbitrary shrinking Ricci soliton ambient space.

Theorem 1.

Let M¯n+1\bar{M}^{n+1} be a shrinking Ricci soliton and let MnM^{n} be a complete CWMC hypersurface immersed in M¯n+1\bar{M}^{n+1} with finite weighted volume.

(a) Suppose that trMn¯¯fatr_{M^{n}}\bar{\nabla}\bar{\nabla}f\geq a for some a>0a>0. If MnM^{n} lies in D(|Hf|,a)¯\overline{D^{-}(|H_{f}|,a)} then MnD(|Hf|,a)M^{n}\subseteq\partial D^{-}(|H_{f}|,a).

(b) Suppose that trMn¯¯fbtr_{M^{n}}\bar{\nabla}\bar{\nabla}f\leq b for some b>0b>0. If MnM^{n} lies outside D+(|Hf|,b)D^{+}(|H_{f}|,b) then MnD+(|Hf|,b)M^{n}\subseteq\partial D^{+}(|H_{f}|,b).

We remark that ¯¯f\bar{\nabla}\bar{\nabla}f is a (0,2)(0,2) tensor in M¯n+1\bar{M}^{n+1}, so we can consider its restriction to MnM^{n} (in this case the trace trMn¯¯ftr_{M^{n}}\bar{\nabla}\bar{\nabla}f is a function in MnM^{n}). Note that we can state Theorem 1 in a different way using the fact that

trMn¯¯f=nλR¯+R¯ic(N,N),tr_{M^{n}}\bar{\nabla}\bar{\nabla}f=n\lambda-\bar{R}+\bar{R}ic\left(N,N\right),

where NN is the unit normal vector of MnM^{n}.

Using the above result together with a classification result for the level sets of the potential function (see Theorem 17) we obtain a new result for the cylinder shrinking Ricci soliton ambient space.

Corollary 2.

Let MnM^{n} be a complete CWMC hypersurface properly immersed in the cylinder shrinking Ricci soliton n+1k×S2(k1)k(0)\mathbb{R}^{n+1-k}\times S_{\sqrt{2(k-1)}}^{k}(0), where n3n\geq 3 and 2kn12\leq k\leq n-1.

(a) If MnM^{n} lies inside the closed product B¯rn+1k(0)×S2(k1)k(0)\bar{B}^{n+1-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0) with r=|Hf|+Hf2+2(nk)r=-|H_{f}|+\sqrt{H_{f}^{2}+2\left(n-k\right)}, then Hf0H_{f}\geq 0 and Mn=Srnk(0)×S2(k1)k(0)M^{n}=S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0).

(b) MnM^{n} cannot lie outside the product Brn+1k(0)×S2(k1)k(0)B^{n+1-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0) with r=|Hf|+Hf2+2(n+1k)r=|H_{f}|+\sqrt{H_{f}^{2}+2\left(n+1-k\right)}.

In Theorem 3.5 in [17], Guang proved that if MnM^{n} is a compact CWMC hypersurface in the Gaussian shrinking Ricci soliton n+1\mathbb{R}^{n+1} satisfying Hf0H_{f}\geq 0 and

|A|212+Hf(Hf+Hf2+2n)2n,|A|^{2}\leq\frac{1}{2}+\frac{H_{f}\left(H_{f}+\sqrt{H_{f}^{2}+2n}\right)}{2n},

then Mn=Srn(0)M^{n}=S^{n}_{r}(0), where r=Hf+Hf2+2nr=-H_{f}+\sqrt{H_{f}^{2}+2n}. We generalize this result in two different ways: we assume an upper bound on the mean curvature (this assumption is weaker since H2n|A|2H^{2}\leq n|A|^{2}) and we consider an arbitrary shrinking Ricci soliton ambient space. Here AA is the second fundamental form and HH is the mean curvature.

Theorem 3.

Let M¯n+1\bar{M}^{n+1} be a shrinking Ricci soliton and let MnM^{n} be a compact CWMC hypersurface immersed in M¯n+1\bar{M}^{n+1}. Suppose that trMn¯¯fatr_{M^{n}}\bar{\nabla}\bar{\nabla}f\geq a for some a>0a>0. Assume that f(p)=supMnff(p)=\sup_{M^{n}}f is a regular value of ff. If

H2Hf|Hf|+Hf2+4a2,H\leq\frac{2H_{f}-|H_{f}|+\sqrt{H_{f}^{2}+4a}}{2},

then MnD(|Hf|,a)M^{n}\subseteq\partial D^{-}(|H_{f}|,a).

Remark 4.

The theorem above generalizes Theorem 3.5 in [17]. Indeed, when the ambient space is the Gaussian shrinking Ricci soliton (see Example 13) we have trMn¯¯f=n2tr_{M^{n}}\bar{\nabla}\bar{\nabla}f=\frac{n}{2} (so a=n2a=\frac{n}{2}). Assuming that Hf0H_{f}\geq 0 and

|A|212+Hf(Hf+Hf2+2n)2n=(Hf+Hf2+2n)24n|A|^{2}\leq\frac{1}{2}+\frac{H_{f}\left(H_{f}+\sqrt{H_{f}^{2}+2n}\right)}{2n}=\frac{\left(H_{f}+\sqrt{H_{f}^{2}+2n}\right)^{2}}{4n}

and using the fact that 1nH2|A|2\frac{1}{n}H^{2}\leq|A|^{2} we have

H2Hf|Hf|+Hf2+2n2.H\leq\frac{2H_{f}-|H_{f}|+\sqrt{H_{f}^{2}+2n}}{2}.

Now using Theorem 3 we conclude that Mn=Srn(0)M^{n}=S^{n}_{r}(0).

In particular, we obtain a new result for the cylinder shrinking Ricci soliton ambient space.

Corollary 5.

Let MnM^{n} be a compact CWMC hypersurface immersed in the cylinder shrinking Ricci soliton nk+1×S2(k1)k(0)\mathbb{R}^{n-k+1}\times S^{k}_{\sqrt{2(k-1)}}(0), where n3n\geq 3 and 2kn12\leq k\leq n-1. If Hf0H_{f}\geq 0 and

HHf+Hf2+2(nk)2,H\leq\frac{H_{f}+\sqrt{H_{f}^{2}+2(n-k)}}{2},

then Mn=Srnk(0)×S2(k1)k(0)M^{n}=S^{n-k}_{r}(0)\times S^{k}_{\sqrt{2(k-1)}}(0), where r=Hf+Hf2+2(nk)r=-H_{f}+\sqrt{H_{f}^{2}+2(n-k)}.

Note that the assumption on the upper bound of HH is sharp because equality holds for Mn=Srnk(0)×S2(k1)k(0)M^{n}=S^{n-k}_{r}(0)\times S^{k}_{\sqrt{2(k-1)}}(0), where r=Hf+Hf2+2(nk)r=-H_{f}+\sqrt{H_{f}^{2}+2(n-k)}.

This is a new result even for ff-minimal hypersurfaces.

Corollary 6.

Let MnM^{n} be a compact ff-minimal hypersurface immersed in the cylinder shrinking Ricci soliton nk+1×S2(k1)k(0)\mathbb{R}^{n-k+1}\times S^{k}_{\sqrt{2(k-1)}}(0), where n3n\geq 3 and 2kn12\leq k\leq n-1. If

H2(nk)2,H\leq\frac{\sqrt{2(n-k)}}{2},

then Mn=Srnk(0)×S2(k1)k(0)M^{n}=S^{n-k}_{r}(0)\times S^{k}_{\sqrt{2(k-1)}}(0), where r=2(nk)r=\sqrt{2(n-k)}.

In Theorem 1.2 in [7], Q.M.Cheng-Ogata-Wei proved that if MnM^{n} is a complete CWMC hypersurface in the Gaussian shrinking Ricci soliton n+1\mathbb{R}^{n+1} with polynomial volume growth and satisfying

(HHf2)2Hf24+n2,\left(H-\frac{H_{f}}{2}\right)^{2}\geq\frac{H_{f}^{2}}{4}+\frac{n}{2},

then Mn=Srn(0)M^{n}=S^{n}_{r}(0), where r=Hf+Hf2+2nr=-H_{f}+\sqrt{H_{f}^{2}+2n}. We generalize this result to any ambient space which is a smooth metric measure space.

Theorem 7.

Let (M¯n+1,g¯,f)(\bar{M}^{n+1},\bar{g},f) be a smooth measure metric space and let MnM^{n} be a complete hypersurface immersed in M¯n+1\bar{M}^{n+1} with finite weighted volume. Suppose that trM¯¯fb\text{tr}_{M}\bar{\nabla}\bar{\nabla}f\leq b for some b>0b>0. If

HHf+Hf2+4b2,H\geq\frac{H_{f}+\sqrt{H_{f}^{2}+4b}}{2},

then Mnf1(γ)M^{n}\subseteq f^{-1}(\gamma) for some γ\gamma\in\mathbb{R}.

In particular, we recover the result of Q.M.Cheng-Ogata-Wei (see Corollary 26) and we obtain a new result for the cylinder shrinking Ricci soliton ambient space.

Corollary 8.

Let MnM^{n} be a complete CWMC hypersurface properly immersed in the cylinder shrinking Ricci soliton n+1k×S2(k1)k(0)\mathbb{R}^{n+1-k}\times S^{k}_{\sqrt{2(k-1)}}(0), where n3n\geq 3 and 2kn12\leq k\leq n-1. Then

infH<Hf+Hf2+2(n+1k)2.\inf H<\frac{H_{f}+\sqrt{H_{f}^{2}+2(n+1-k)}}{2}.

This is a new result even for ff-minimal hypersurfaces.

Corollary 9.

Let MnM^{n} be a complete ff-minimal hypersurface properly immersed in the cylinder shrinking Ricci soliton n+1k×S2(k1)k(0)\mathbb{R}^{n+1-k}\times S^{k}_{\sqrt{2(k-1)}}(0), where n3n\geq 3 and 2kn12\leq k\leq n-1. Then

infH<2(n+1k)2.\inf H<\frac{\sqrt{2(n+1-k)}}{2}.

In the second part of this paper, we obtain some classification theorems for the Gaussian shrinking Ricci soliton ambient space. Le-Sesum [21] proved that if MnM^{n} is a complete self-shrinker in the Gaussian shrinking Ricci soliton n+1\mathbb{R}^{n+1} with polynomial volume growth and satisfying |A|2<1/2|A|^{2}<1/2, then MnM^{n} is a hyperplane passing through the origin. Later, Cao-Li [5] extended this result to arbitrary codimension by showing that if |A|21/2|A|^{2}\leq 1/2, then MnM^{n} is a generalized cylinder. Later, Guang [17] extended this result to CWMC hypersurfaces by assuming polynomial volume growth and a condition on the norm of the second fundamental form. Many of the classification results for CWMC hypersurfaces assume that the hypersurface has polynomial volume growth in order to use integration techniques. Recently, there has been some papers trying to avoid this assumption by using the Omori-Yau maximum principles ([7], [8], [25], etc). We can replace the assumption of polynomial volume growth in Guang’s result by an assumption on the integral of the second fundamental form.

Theorem 10.

Let MnM^{n} be a complete embedded CWMC hypersurface in the Gaussian shrinking Ricci soliton n+1\mathbb{R}^{n+1}. If ALfq(M)A\in L_{f}^{q}(M) for some q1q\geq 1 and

|A|Hf2+2|Hf|2,|A|\leq\frac{\sqrt{H_{f}^{2}+2}-|H_{f}|}{2},

then either MnM^{n} is a hyperplane or it is a generalized cylinder Srk(0)×nkS^{k}_{r}(0)\times\mathbb{R}^{n-k}, where 1kn1\leq k\leq n.

We remark that for Mn=Srn(0)M^{n}=S^{n}_{r}(0) the assumption on the upper bound is not satisfied unless r=2nr=\sqrt{2n} (that is Hf=0H_{f}=0) and for Srk(0)×nkS^{k}_{r}(0)\times\mathbb{R}^{n-k} the assumption on the upper bound is not satisfied unless r=2kr=\sqrt{2k} (that is Hf=0H_{f}=0). Therefore the result is only sharp in these cases. It would be interesting to find a sharp result for Hf0H_{f}\neq 0.

This is a new result even for self-shrinkers.

Corollary 11.

Let MnM^{n} be a complete embedded self-shrinker in the Gaussian shrinking Ricci soliton n+1\mathbb{R}^{n+1}. If ALfq(M)A\in L^{q}_{f}(M) for some q1q\geq 1 and |A|21/2|A|^{2}\leq 1/2, then either MnM^{n} is a hyperplane passing through the origin or it is a generalized cylinder Srk(0)×nkS^{k}_{r}(0)\times\mathbb{R}^{n-k}, with 1kn1\leq k\leq n and r=2kr=\sqrt{2k}.

We remark that Ancari-Miranda [1] proved recently it is possible to obtain a similar result if |A|21/2|A|^{2}\leq 1/2 and HLfq(M)H\in L^{q}_{f}(M) for some even number q2q\geq 2 which is a weaker assumption when qq is even.

For self-shrinker submanifolds MnM^{n} in n+p\mathbb{R}^{n+p}, Q.M.Cheng-Peng [8] proved that if sup|A|2<1/2\sup|A|^{2}<1/2, then MnM^{n} is a hyperplane in n+1\mathbb{R}^{n+1}. For codimension 1 we generalize Q.M.Cheng-Peng’s result to CWMC hypersurfaces.

Corollary 12.

Let MnM^{n} be a complete embedded CWMC hypersurface in n+1\mathbb{R}^{n+1}. If

sup|A|<Hf2+2|Hf|2,\sup|A|<\frac{\sqrt{H_{f}^{2}+2}-|H_{f}|}{2},

then MnM^{n} is a hyperplane.

This paper is organized as follows. In section 2, we describe our notations and conventions and prove some basic results. In section 3, we prove Theorem 1, Theorem 3, Theorem 7 and related results. In Section 4, we prove Theorem 10 and related results.

2. Preliminaries and basics results

In this section we describe our notations and conventions and prove some basic results.

Smooth metric measure spaces. Let (M,g)\left(M,g\right) be a Riemannian manifold and let ff be a smooth function on MM. The triple (M,g,f)\left(M,g,f\right) is called a smooth metric measure space. The measure efdvole^{-f}dvol is called weighted volume. If uu and vv are functions on MnM^{n} the Lf2L_{f}^{2} inner product uu and vv is defined by

u,vLf2(M)=Muvef\left\langle u,v\right\rangle_{L_{f}^{2}\left(M\right)}=\int_{M}uve^{-f}

and the LfpL_{f}^{p} norm of uu is defined by

|u|Lfp(M)=(M|u|pef)1p,\left|u\right|_{L_{f}^{p}\left(M\right)}=\left(\int_{M}\left|u\right|^{p}e^{-f}\right)^{\frac{1}{p}},

where p1p\geq 1. The operator

Δf=Δf,\Delta_{f}=\Delta-\left\langle\nabla f,\nabla\cdot\right\rangle

is called weighted Laplacian. It is well known that the weighted Laplacian is a densely defined self-adjoint operator in Lf2L_{f}^{2}, that is, if uu and vv are smooth functions on MM with compact support we have

M(Δfu)vef=Mu,vef.\int_{M}\left(\Delta_{f}u\right)ve^{-f}=-\int_{M}\left\langle\nabla u,\nabla v\right\rangle e^{-f}.

The Bakry-Émery-Ricci curvature is defined by

Ricf=Ric+f.Ric_{f}=Ric+\nabla\nabla f.

The triple (M,g,f)\left(M,g,f\right) is called a gradient Ricci soliton if

Ricf=λg,Ric_{f}=\lambda g,

where λ\lambda is a constant. If λ\lambda is positive, zero or negative it is called shrinking, steady or expanding, respectively.

Hypersurfaces in smooth metric measure spaces. Let (M¯n+1,g¯,f)\left(\bar{M}^{n+1},\bar{g},f\right) be a smooth metric measure space and let MnM^{n} be a oriented hypersurface of M¯n+1\bar{M}^{n+1}. The second fundamental form is defined by

A(u,v)=¯uv,N,A\left(u,v\right)=\langle\bar{\nabla}_{u}v,N\rangle,

where uu and vv are vector fields on MnM^{n}, ¯\bar{\nabla} is the Riemannian connection of M¯n+1\bar{M}^{n+1} and NN is the unit normal vector. The mean curvature vector is defined by

H=(trMnA)N,\vec{H}=(tr_{M^{n}}A)N,

The weighted mean curvature vector is defined by

Hf=H+(¯f).\vec{H}_{f}=\vec{H}+\left(\bar{\nabla}f\right)^{\perp}.

We remark that

Hf=i=1n¯eiei,NN+¯f,NN.\vec{H}_{f}=\sum_{i=1}^{n}\langle\bar{\nabla}_{e_{i}}e_{i},N\rangle N+\langle\bar{\nabla}f,N\rangle N.

We see that Hf\vec{H}_{f} does not depend on the choice of the normal vector. The weighted mean curvature is defined by

Hf=HfN.\vec{H}_{f}=-H_{f}N.

A hypersurface is said to have constant weighted mean curvature (CWMC) if the weighted mean curvature is constant. Note that (Mn,g,f)\left(M^{n},g,f\right) is also a smooth metric measure space with weighted measure efdvolMe^{-f}dvol_{M} and weighted Laplacian

Δf=Δf,,\Delta_{f}=\Delta-\left\langle\nabla f,\nabla\cdot\right\rangle,

where g=g¯|Mg=\bar{g}|M, \nabla is the Riemannian connection of MnM^{n} and Δ\Delta is the Laplacian of MnM^{n}.

Example 13.

Let M¯n+1=n+1\bar{M}^{n+1}=\mathbb{R}^{n+1}, g¯=g¯can\bar{g}=\bar{g}_{can} and f(x)=|x|24f\left(x\right)=\frac{\left|x\right|^{2}}{4}. The triple (n+1,g¯,f)\left(\mathbb{R}^{n+1},\bar{g},f\right) is a gradient shrinking Ricci soliton with Bakry-Émery-Ricci curvature

R¯icf=12g¯.\bar{R}ic_{f}=\frac{1}{2}\bar{g}.

A hypersurface MnM^{n} of n+1\mathbb{R}^{n+1} is a CWMC hypersurface if and only if

H=x,N2+Hf.H=\frac{\langle x,N\rangle}{2}+H_{f}.

Note that Hf=0H_{f}=0 if and only if MnM^{n} is a self-shrinker. The weighted volume of MnM^{n} is e|x|24dvolMe^{-\frac{|x|^{2}}{4}}dvol_{M} and the weighted Laplacian of MnM^{n} is given by

Δf=Δ12x,.\Delta_{f}=\Delta-\frac{1}{2}\left\langle x,\nabla\cdot\right\rangle.

Note that Δf\Delta_{f} is the operator \mathcal{L} introduced by Colding and Minicozzi [15]. We say that (n+1,g¯,f)\left(\mathbb{R}^{n+1},\bar{g},f\right) is the Gaussian shrinking Ricci soliton.

Example 14.

Let M¯n+1=n+1\bar{M}^{n+1}=\mathbb{R}^{n+1}, g¯=g¯can\bar{g}=\bar{g}_{can} and f(x)=|x|24f\left(x\right)=-\frac{\left|x\right|^{2}}{4}. The triple (n+1,g¯,f)\left(\mathbb{R}^{n+1},\bar{g},f\right) is a gradient expanding Ricci soliton with Bakry-Émery-Ricci curvature

R¯icf=12g¯.\bar{R}ic_{f}=-\frac{1}{2}\bar{g}.

A hypersurface MnM^{n} of n+1\mathbb{R}^{n+1} is a CWMC hypersurface if and only if

H=x,N2+Hf.H=-\frac{\langle x,N\rangle}{2}+H_{f}.

Note that Hf=0H_{f}=0 if and only if MnM^{n} is a self-expander. The weighted volume of MnM^{n} is e|x|24dvolMe^{\frac{|x|^{2}}{4}}dvol_{M} and the weighted Laplacian of MnM^{n} is given by

Δf=Δ+12x,.\Delta_{f}=\Delta+\frac{1}{2}\left\langle x,\nabla\cdot\right\rangle.
Example 15.

Let M¯n+1=n+1\bar{M}^{n+1}=\mathbb{R}^{n+1}, g¯=g¯can\bar{g}=\bar{g}_{can} and f(x)=a,xf\left(x\right)=\langle a,x\rangle, where an+1a\in\mathbb{R}^{n+1}. The triple (n+1,g¯,f)\left(\mathbb{R}^{n+1},\bar{g},f\right) is a gradient steady Ricci soliton with Bakry-Émery-Ricci curvature

R¯icf=0.\bar{R}ic_{f}=0.

A hypersurface MnM^{n} of n+1\mathbb{R}^{n+1} is a CWMC hypersurface if and only if

H=a,N+Hf.H=\langle a,N\rangle+H_{f}.

Note that Hf=0H_{f}=0 if and only if MnM^{n} is a translating soliton. The weighted volume of MnM^{n} is ea,xdvolMe^{-\langle a,x\rangle}dvol_{M} and the weighted Laplacian of MnM^{n} is given by

Δf=Δa,.\Delta_{f}=\Delta-\left\langle a,\nabla\cdot\right\rangle.

In [22] Lopez classified CWMC surfaces in the steady Ricci soliton 3\mathbb{R}^{3} that are invariant by rotations and translations.

Example 16.

Let M¯n+1=n+1k×S2(k1)k\bar{M}^{n+1}=\mathbb{R}^{n+1-k}\times S_{\sqrt{2\left(k-1\right)}}^{k} with product metric g¯\bar{g} and potential function f(x,y)=|x|24f\left(x,y\right)=\frac{\left|x\right|^{2}}{4}. Here xx is the position vector in nk+1\mathbb{R}^{n-k+1} and yy is the position vector in k+1\mathbb{R}^{k+1}. The triple (n+1k×S2(k1)k,g¯,f)\left(\mathbb{R}^{n+1-k}\times S_{\sqrt{2\left(k-1\right)}}^{k},\bar{g},f\right) is a gradient shrinking Ricci soliton with Bakry-Émery-Ricci curvature

R¯icf\displaystyle\bar{R}ic_{f} =¯¯f+R¯ic\displaystyle=\bar{\nabla}\bar{\nabla}f+\bar{R}ic
=12gn+1k+12gS2(k1)k\displaystyle=\frac{1}{2}g_{\mathbb{R}^{n+1-k}}+\frac{1}{2}g_{S_{\sqrt{2\left(k-1\right)}}^{k}}
=12g¯.\displaystyle=\frac{1}{2}\bar{g}.

A hypersurface MnM^{n} of n+1k×S2(k1)k\mathbb{R}^{n+1-k}\times S_{\sqrt{2\left(k-1\right)}}^{k} is a CWMC hypersurface if and only if

H=x,N2+Hf.H=\frac{\langle x,N\rangle}{2}+H_{f}.

The weighted volume of MnM^{n} is e|x|24dvolMe^{-\frac{|x|^{2}}{4}}dvol_{M} and the weighted Laplacian of MnM^{n} is given by

Δf=Δ12x,.\Delta_{f}=\Delta-\frac{1}{2}\left\langle x,\nabla\cdot\right\rangle.

We say that (n+1k×S2(k1)k,g¯,f)\left(\mathbb{R}^{n+1-k}\times S_{\sqrt{2\left(k-1\right)}}^{k},\bar{g},f\right) is the cylinder shrinking Ricci soliton.

In the rest of the section we prove some results for the level set of the potential function of a shrinking Ricci soliton.

Theorem 17.

Let M¯n+1\bar{M}^{n+1} be a complete gradient shrinking Ricci soliton with constant scalar curvature. Suppose that

Mn={xM¯n+1:f(x)=γ},M^{n}=\left\{x\in\bar{M}^{n+1}:f\left(x\right)=\gamma\right\},

where γ\gamma is a regular value of ff. Then MnM^{n} is a CWMC hypersurface with

Hf=nλ2λγC2λγR¯+C,H_{f}=\frac{n\lambda-2\lambda\gamma-C}{\sqrt{2\lambda\gamma-\bar{R}+C}},

and

γ=12λ{14(Hf+Hf2+4(nλR¯))2+R¯C}.\gamma=\frac{1}{2\lambda}\left\{\frac{1}{4}\left(-H_{f}+\sqrt{H_{f}^{2}+4\left(n\lambda-\bar{R}\right)}\right)^{2}+\bar{R}-C\right\}.

Moreover R¯{0,2λ,,nλ}\bar{R}\in\{0,2\lambda,...,n\lambda\}.

Proof.

Using the fact that N=¯f|¯f|N=\frac{\bar{\nabla}f}{\left|\bar{\nabla}f\right|} and ¯¯f=R¯ic+λg¯\bar{\nabla}\bar{\nabla}f=-\bar{R}ic+\lambda\bar{g} we have

A(v,w)\displaystyle A\left(v,w\right) =¯vN,w\displaystyle=-\left\langle\bar{\nabla}_{v}N,w\right\rangle
=¯v¯f|¯f|,w\displaystyle=-\left\langle\bar{\nabla}_{v}\frac{\bar{\nabla}f}{\left|\bar{\nabla}f\right|},w\right\rangle
=1|¯f|¯¯f(v,w)\displaystyle=-\frac{1}{\left|\bar{\nabla}f\right|}\bar{\nabla}\bar{\nabla}f\left(v,w\right)
=1|¯f|(R¯ic(v,w)λv,w).\displaystyle=\frac{1}{\left|\bar{\nabla}f\right|}\left(\bar{R}ic\left(v,w\right)-\lambda\left\langle v,w\right\rangle\right).

Then

Hf\displaystyle H_{f} =H¯f,N\displaystyle=H-\left\langle\bar{\nabla}f,N\right\rangle
=1|¯f|(nλtrMnR¯ic)¯f,¯f|¯f|\displaystyle=\frac{1}{\left|\bar{\nabla}f\right|}\left(n\lambda-tr_{M^{n}}\bar{R}ic\right)-\left\langle\bar{\nabla}f,\frac{\bar{\nabla}f}{\left|\bar{\nabla}f\right|}\right\rangle
=1|¯f|(nλR¯+R¯ic(N,N)|¯f|2).\displaystyle=\frac{1}{\left|\bar{\nabla}f\right|}\left(n\lambda-\bar{R}+\bar{R}ic\left(N,N\right)-\left|\bar{\nabla}f\right|^{2}\right).

It is well known that if M¯n+1\bar{M}^{n+1} is a gradient Ricci soliton then R¯ic(¯f)=12¯R¯\bar{R}ic\left(\bar{\nabla}f\right)=\frac{1}{2}\bar{\nabla}\bar{R} (see Equation 1.7 in [4]). Using this and the fact that R¯\bar{R} is constant and N=¯f|¯f|N=\frac{\bar{\nabla}f}{\left|\bar{\nabla}f\right|} we have R¯ic(N,N)=0\bar{R}ic\left(N,N\right)=0. We find that

Hf=1|¯f|(nλR¯|¯f|2).H_{f}=\frac{1}{\left|\bar{\nabla}f\right|}\left(n\lambda-\bar{R}-\left|\bar{\nabla}f\right|^{2}\right).

Using the fact that R¯+|¯f|2=2λf+C\bar{R}+\left|\bar{\nabla}f\right|^{2}=2\lambda f+C we get the first part of the result.

Now we prove the second part of the result. Multiplying by |¯f|\left|\bar{\nabla}f\right| we obtain

|¯f|2+Hf|¯f|(nλR¯)=0.\left|\bar{\nabla}f\right|^{2}+H_{f}\left|\bar{\nabla}f\right|-\left(n\lambda-\bar{R}\right)=0.

Solving the quadratic equation we get

|¯f|=12(Hf+Hf2+4(nλR¯)),\left|\bar{\nabla}f\right|=\frac{1}{2}\left(-H_{f}+\sqrt{H_{f}^{2}+4\left(n\lambda-\bar{R}\right)}\right),

or

|¯f|=12(HfHf2+4(nλR¯)).\left|\bar{\nabla}f\right|=\frac{1}{2}\left(-H_{f}-\sqrt{H_{f}^{2}+4\left(n\lambda-\bar{R}\right)}\right).

Claim: The second case does not happen. We will prove the claim later. Assuming the claim we have

|¯f|2=14(Hf+Hf2+4(nλR¯))2.\left|\bar{\nabla}f\right|^{2}=\frac{1}{4}\left(-H_{f}+\sqrt{H_{f}^{2}+4\left(n\lambda-\bar{R}\right)}\right)^{2}.

Using the fact that |¯f|2=2λfR¯+C\left|\bar{\nabla}f\right|^{2}=2\lambda f-\bar{R}+C we get

f=12λ{14(Hf+Hf2+4(nλR¯))2+R¯C}.f=\frac{1}{2\lambda}\left\{\frac{1}{4}\left(-H_{f}+\sqrt{H_{f}^{2}+4\left(n\lambda-\bar{R}\right)}\right)^{2}+\bar{R}-C\right\}.

This proves the second part of the result.

Now we prove the claim. We only need to show that

|¯f|=12(HfHf2+4(nλR¯))0.\left|\bar{\nabla}f\right|=\frac{1}{2}\left(-H_{f}-\sqrt{H_{f}^{2}+4\left(n\lambda-\bar{R}\right)}\right)\leq 0.

Since R¯\bar{R} is constant on M¯n+1\bar{M}^{n+1} we know that R¯{0,2λ,,nλ,(n+1)λ}\bar{R}\in\left\{0,2\lambda,...,n\lambda,\left(n+1\right)\lambda\right\} (see Theorem 1 in [16]). We will show that R¯{0,2λ,,nλ}\bar{R}\in\left\{0,2\lambda,...,n\lambda\right\}. Suppose that R¯=(n+1)λ\bar{R}=\left(n+1\right)\lambda. In this case by Proposition 3.3 in [24] we see that M¯n+1\bar{M}^{n+1} is Einstein, which implies that

R¯ic=R¯n+1g¯=λg¯.\bar{R}ic=\frac{\bar{R}}{n+1}\bar{g}=\lambda\bar{g}.

Since ¯¯f+R¯ic=λg¯\bar{\nabla}\bar{\nabla}f+\bar{R}ic=\lambda\bar{g} we have ¯¯f=0\bar{\nabla}\bar{\nabla}f=0. This implies that |¯f|\left|\bar{\nabla}f\right| is constant on M¯n+1\bar{M}^{n+1}. Since 2λf=|¯f|2R¯+C2\lambda f=-\left|\bar{\nabla}f\right|^{2}-\bar{R}+C we see that ff is constant on M¯n+1\bar{M}^{n+1} and ¯f=0\bar{\nabla}f=0. This contradicts the fact that MnM^{n} is a level set of the potential function. Therefore R¯{0,2λ,,nλ}\bar{R}\in\left\{0,2\lambda,...,n\lambda\right\}. Since nλR¯0n\lambda-\bar{R}\geq 0 we have

HfHf2+4(nλR¯)0.-H_{f}-\sqrt{H_{f}^{2}+4\left(n\lambda-\bar{R}\right)}\leq 0.

This proves the claim. ∎

Remark 18.

Note that when the ambient is a shrinking Ricci soliton with scalar curvature R¯=nλ\bar{R}=n\lambda, there is no level set which is ff-minimal. Indeed,

Hf=1|¯f|(nλR¯|¯f|2)H_{f}=\frac{1}{\left|\bar{\nabla}f\right|}\left(n\lambda-\bar{R}-\left|\bar{\nabla}f\right|^{2}\right)

implies |¯f|=0\left|\bar{\nabla}f\right|=0, a contradiction.

Using the normalization λ=1/2\lambda=1/2 and C=R¯C=\bar{R} (which is used in many papers, see for example pag 2 in [4]) we have:

Corollary 19.

Let M¯n+1\bar{M}^{n+1} be a complete gradient shrinking Ricci soliton with constant scalar curvature. Suppose that

Mn={xM¯n+1:f(x)=γ},M^{n}=\left\{x\in\bar{M}^{n+1}:f\left(x\right)=\gamma\right\},

where γ\gamma is a regular value of ff. Then MnM^{n} is a CWMC hypersurface with

Hf=n2γ2R¯2γ,H_{f}=\frac{n-2\gamma-2\bar{R}}{2\sqrt{\gamma}},

and

γ=14(Hf+Hf2+4(n2R¯))2.\gamma=\frac{1}{4}\left(-H_{f}+\sqrt{H_{f}^{2}+4\left(\frac{n}{2}-\bar{R}\right)}\right)^{2}.

Moreover R¯{0,1,,n2}\bar{R}\in\{0,1,...,\frac{n}{2}\}.

For 3-dimensional shrinking Ricci solitons we have:

Corollary 20.

Let M¯3\bar{M}^{3} be a complete gradient shrinking Ricci soliton with constant scalar curvature. Suppose that

M2={xM¯3:f(x)=γ},M^{2}=\left\{x\in\bar{M}^{3}:f\left(x\right)=\gamma\right\},

where γ\gamma is a regular value of ff. Then M2M^{2} is a CWMC hypersurface and R¯{0,1}\bar{R}\in\{0,1\}. Moreover

Hf=1γγin the caseR¯=0,H_{f}=\frac{1-\gamma}{\sqrt{\gamma}}\ \ \text{in the case}\ \ \bar{R}=0,

and

Hf=γin the caseR¯=1.H_{f}=-\sqrt{\gamma}\ \ \text{in the case}\ \ \bar{R}=1.

For 4-dimensional shrinking Ricci solitons we have:

Corollary 21.

Let M¯4\bar{M}^{4} be a complete gradient shrinking Ricci soliton with constant scalar curvature. Suppose that

M3={xM¯4:f(x)=γ},M^{3}=\left\{x\in\bar{M}^{4}:f\left(x\right)=\gamma\right\},

where γ\gamma is a regular value of ff. Then M3M^{3} is a CWMC hypersurface and R¯{0,1,32}\bar{R}\in\{0,1,\frac{3}{2}\}. Moreover

Hf=32γ2γ,in the caseR¯=0,H_{f}=\frac{3-2\gamma}{2\sqrt{\gamma}},\ \ \text{in the case}\ \ \bar{R}=0,
Hf=12γ2γ,in the caseR¯=1,H_{f}=\frac{1-2\gamma}{2\sqrt{\gamma}},\ \ \text{in the case}\ \ \bar{R}=1,

and

Hf=γ,in the caseR¯=32.H_{f}=-\sqrt{\gamma},\ \ \text{in the case}\ \ \bar{R}=\frac{3}{2}.

Another consequence of Theorem 17 is the following.

Corollary 22.

Let r>0r>0. Then Srnk(0)×S2(k1)k(0)S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}(0) is a CWMC hypersurface in the cylinder shrinking Ricci soliton n+1k×S2(k1)k(0)\mathbb{R}^{n+1-k}\times S_{\sqrt{2\left(k-1\right)}}^{k}(0) and r=Hf+Hf2+2(nk)r=-H_{f}+\sqrt{H_{f}^{2}+2\left(n-k\right)}.

Proof.

By Example 16 we have λ=12\lambda=\frac{1}{2}, R¯=k2\bar{R}=\frac{k}{2} and C=R¯C=\bar{R} (since CR¯=|¯f|22λfC-\bar{R}=\left|\bar{\nabla}f\right|^{2}-2\lambda f). We have

Srnk(0)×S2(k1)k(0)={(x,y)n+1k×S2(k1)k(0);f(x,y)=γ},S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}(0)=\left\{(x,y)\in\mathbb{R}^{n+1-k}\times S_{\sqrt{2\left(k-1\right)}}^{k}(0);f\left(x,y\right)=\gamma\right\},

where γ=r24\gamma=\frac{r^{2}}{4}. By Theorem 17 we have

γ=14(Hf+Hf2+2(nk))2.\gamma=\frac{1}{4}\left(-H_{f}+\sqrt{H_{f}^{2}+2\left(n-k\right)}\right)^{2}.

This proves the result. ∎

Using similar computations as in the proof of Theorem 17 we show that the level sets cannot be ff-minimal when the ambient space is a steady or expanding Ricci soliton with constant scalar curvature.

Proposition 23.

Let M¯n+1\bar{M}^{n+1} be a complete gradient Ricci soliton with constant scalar curvature. If there exists a regular level set of the potential function f which is f-minimal, then M¯n+1\bar{M}^{n+1} is a gradient shrinking Ricci soliton.

Proof.

Let Mn=f1(γ)M^{n}=f^{-1}(\gamma) be a regular level set of ff which is ff-minimal. By the proof of Theorem 17 we have

Hf=1|¯f|(nλR¯|¯f|2).H_{f}=\frac{1}{\left|\bar{\nabla}f\right|}\left(n\lambda-\bar{R}-\left|\bar{\nabla}f\right|^{2}\right).

From the assumption that MnM^{n} is f-minimal we have

nλR¯|¯f|2=0.n\lambda-\bar{R}-|\bar{\nabla}f|^{2}=0.

First suppose that M¯n+1\bar{M}^{n+1} is a steady soliton (λ=0\lambda=0). By Theorem 1.3 in [29] we have R¯0\bar{R}\geq 0. From the equation above we conclude that |¯f|=0|\bar{\nabla}f|=0, which is a contradiction with the fact that MnM^{n} is a level of the potential function.
Now suppose that M¯n+1\bar{M}^{n+1} is an expanding Ricci soliton (λ<0\lambda<0). By Theorem 1 in [16] we have R¯{(n+1)λ,nλ,,2λ,0}\bar{R}\in\{(n+1)\lambda,n\lambda,...,2\lambda,0\}. Using the same argument as in the proof of Lemma 17 we have R¯nλ\bar{R}\geq n\lambda. From the equation above we conclude that |¯f|=0|\bar{\nabla}f|=0, which is a contradiction with the fact that MnM^{n} is a level of the potential function. ∎

In Proposition 5.3 in [5], Cao-Li showed that there are no compact self-expanders in the expanding Ricci soliton (n+1,g¯can,f=|x|2/4)(\mathbb{R}^{n+1},\bar{g}_{can},f=-|x|^{2}/4). Thus, the above result was expected for expanding Ricci solitons with a proper potential function.

3. Geometric results in shrinking Ricci solitons

In this section we prove Theorem 1, Theorem 3, Theorem 7 and related results.

Proof of Theorem 1.

Proof.

Fact (a). Let uu be a function on MnM^{n}. If uu is bounded from above and Δfu0\Delta_{f}u\geq 0 then uu is constant.

Fact (b). Let uu be a function on MnM^{n}. If uu is bounded from below and Δfu0\Delta_{f}u\leq 0 then uu is constant.

Since MnM^{n} has finite weighted volume both facts follow from Corollary 1 in [11]. We remark that the extension of these results to smooth metric measure spaces is straightforward. See also Theorem 25 and Remark 26 in [26].

We have

f(u,v)=¯¯f(u,v)+¯f,(¯uv).\nabla\nabla f\left(u,v\right)=\bar{\nabla}\bar{\nabla}f\left(u,v\right)+\left\langle\bar{\nabla}f,\left(\bar{\nabla}_{u}v\right)^{\perp}\right\rangle.

We see that

Δff\displaystyle\Delta_{f}f =trMn¯¯f+¯f,Hf,f\displaystyle=tr_{M^{n}}\bar{\nabla}\bar{\nabla}f+\left\langle\bar{\nabla}f,\vec{H}\right\rangle-\left\langle\nabla f,\nabla f\right\rangle
=trMn¯¯f+¯f,H+(¯f)¯f,¯f\displaystyle=tr_{M^{n}}\bar{\nabla}\bar{\nabla}f+\left\langle\bar{\nabla}f,\vec{H}+\left(\bar{\nabla}f\right)^{\perp}\right\rangle-\left\langle\bar{\nabla}f,\bar{\nabla}f\right\rangle
=trMn¯¯f+¯f,Hf|¯f|2.\displaystyle=tr_{M^{n}}\bar{\nabla}\bar{\nabla}f+\left\langle\bar{\nabla}f,\vec{H}_{f}\right\rangle-\left|\bar{\nabla}f\right|^{2}.

Proof of Item (a). Since trMn¯¯fatr_{M^{n}}\bar{\nabla}\bar{\nabla}f\geq a we have

Δffa|Hf||¯f||¯f|2.\Delta_{f}f\geq a-\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}.

Claim: We have

a|Hf||¯f||¯f|20onMn,a-\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\geq 0\,\,\,on\,\,\,M^{n},

and equality holds if and only if MnD(|Hf|,a)M^{n}\subseteq\partial D^{-}(|H_{f}|,a). We will prove this claim later. By the claim we have

Δffa|Hf||¯f||¯f|20.\Delta_{f}f\geq a-\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\geq 0.

Since fΓf\leq\Gamma^{-} on MnM^{n}, by Fact (a) we see that ff is constant on MnM^{n}. We conclude that

a|Hf||¯f||¯f|2=0onMn.a-\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}=0\,\,\,on\,\,\,M^{n}.

By the claim we have MnD(|Hf|,a)M^{n}\subseteq\partial D^{-}(|H_{f}|,a). Now we prove the claim. We have

a|Hf||¯f||¯f|20onMn\displaystyle a-\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\geq 0\,\,\,on\,\,\,M^{n}
(|¯f|+|Hf|2)2Hf24+aonMn\displaystyle\iff\left(\left|\bar{\nabla}f\right|+\frac{\left|H_{f}\right|}{2}\right)^{2}\leq\frac{H_{f}^{2}}{4}+a\,\,\,on\,\,\,M^{n}
|¯f|12(|Hf|+Hf2+4a)onMn.\displaystyle\iff\left|\bar{\nabla}f\right|\leq\frac{1}{2}\left(-\left|H_{f}\right|+\sqrt{H_{f}^{2}+4a}\right)\,\,\,on\,\,\,M^{n}.

Since |¯f|2=2λfR¯+C\left|\bar{\nabla}f\right|^{2}=2\lambda f-\bar{R}+C and

2λΓR¯+C=14(|Hf|+Hf2+4a)2,2\lambda\Gamma^{-}-\bar{R}+C=\frac{1}{4}\left(-\left|H_{f}\right|+\sqrt{H_{f}^{2}+4a}\right)^{2},

we have

MnD(|Hf|,a)¯\displaystyle M^{n}\subset\overline{D^{-}(|H_{f}|,a)} fΓonMn\displaystyle\iff f\leq\Gamma^{-}\,\,\,on\,\,\,M^{n}
2λfR¯+C2λΓR¯+ConMn\displaystyle\iff 2\lambda f-\bar{R}+C\leq 2\lambda\Gamma^{-}-\bar{R}+C\,\,\,on\,\,\,M^{n}
|¯f|214(|Hf|+Hf2+4a)2onMn\displaystyle\iff\left|\bar{\nabla}f\right|^{2}\leq\frac{1}{4}\left(-\left|H_{f}\right|+\sqrt{H_{f}^{2}+4a}\right)^{2}\,\,\,on\,\,\,M^{n}
|¯f|12(|Hf|+Hf2+4a)onMn.\displaystyle\iff\left|\bar{\nabla}f\right|\leq\frac{1}{2}\left(-\left|H_{f}\right|+\sqrt{H_{f}^{2}+4a}\right)\,\,\,on\,\,\,M^{n}.

In the last line we used the fact that a0a\geq 0. We conclude that

a|Hf||¯f||¯f|20onMnMnD(|Hf|,a)¯.a-\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\geq 0\,\,\,on\,\,\,M^{n}\iff M^{n}\subset\overline{D^{-}(|H_{f}|,a)}.

Moreover equality holds in the left hand side if and only if MnD(|Hf|,a)M^{n}\subseteq\partial D^{-}(|H_{f}|,a). This proves the claim.

Proof of Item (b). Since trMn¯¯fbtr_{M^{n}}\bar{\nabla}\bar{\nabla}f\leq b we have

Δffb+|Hf||¯f||¯f|2.\Delta_{f}f\leq b+\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}.

Claim. We have

b+|Hf||¯f||¯f|20onMn,b+\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\leq 0\,\,\,on\,\,\,M^{n},

and equality holds if and only if MnD+(|Hf|,b)M^{n}\subseteq\partial D^{+}(|H_{f}|,b). We will prove this claim later. By the claim we have

Δffb+|Hf||¯f||¯f|20.\Delta_{f}f\leq b+\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\leq 0.

Since fΓ+f\geq\Gamma^{+} on MnM^{n}, by Fact (b) we see that ff is constant on MnM^{n}. We conclude that

b+|Hf||¯f||¯f|2=0onMn.b+\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}=0\,\,\,on\,\,\,M^{n}.

By the claim we have MnD+(|Hf|,b)M^{n}\subseteq\partial D^{+}(|H_{f}|,b). Now we prove the claim. We have

b+|Hf||¯f||¯f|20onMn\displaystyle b+\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\leq 0\,\,\,on\,\,\,M^{n}
(|¯f||Hf|2)2Hf24+bonMn\displaystyle\iff\left(\left|\bar{\nabla}f\right|-\frac{\left|H_{f}\right|}{2}\right)^{2}\geq\frac{H_{f}^{2}}{4}+b\,\,\,on\,\,\,M^{n}
|¯f|12(|Hf|+Hf2+4b)onMn.\displaystyle\iff\left|\bar{\nabla}f\right|\geq\frac{1}{2}\left(\left|H_{f}\right|+\sqrt{H_{f}^{2}+4b}\right)\,\,\,on\,\,\,M^{n}.

Since |¯f|2=2λfR¯+C\left|\bar{\nabla}f\right|^{2}=2\lambda f-\bar{R}+C and

2λΓ+R¯+C=14(|Hf|+Hf2+4b)2,2\lambda\Gamma^{+}-\bar{R}+C=\frac{1}{4}\left(\left|H_{f}\right|+\sqrt{H_{f}^{2}+4b}\right)^{2},

we have

MnM¯n+1D+(|Hf|,b)\displaystyle M^{n}\subset\bar{M}^{n+1}\setminus D^{+}(|H_{f}|,b) fΓ+onMn\displaystyle\iff f\geq\Gamma^{+}\,\,\,on\,\,\,M^{n}
2λfR¯+C2λΓ+R¯+ConMn\displaystyle\iff 2\lambda f-\bar{R}+C\geq 2\lambda\Gamma^{+}-\bar{R}+C\,\,\,on\,\,\,M^{n}
|¯f|214(|Hf|+Hf2+4b)2onMn\displaystyle\iff\left|\bar{\nabla}f\right|^{2}\geq\frac{1}{4}\left(\left|H_{f}\right|+\sqrt{H_{f}^{2}+4b}\right)^{2}\,\,\,on\,\,\,M^{n}
|¯f|12(|Hf|+Hf2+4b)onMn.\displaystyle\iff\left|\bar{\nabla}f\right|\geq\frac{1}{2}\left(\left|H_{f}\right|+\sqrt{H_{f}^{2}+4b}\right)\,\,\,on\,\,\,M^{n}.

In the last line we used the fact that b0b\geq 0. We conclude that

b+|Hf||¯f||¯f|20onMnMnM¯n+1D+(|Hf|,b).b+\left|H_{f}\right|\left|\bar{\nabla}f\right|-\left|\bar{\nabla}f\right|^{2}\leq 0\,\,\,on\,\,\,M^{n}\iff M^{n}\subset\bar{M}^{n+1}\setminus D^{+}(|H_{f}|,b).

Moreover equality holds in the left hand side if and only if MnD+(|Hf|,b)M^{n}\subseteq\partial D^{+}(|H_{f}|,b). ∎

The following lemma will be important throughout this work in order to guarantee the equivalence between properness and finite weighted volume.

Lemma 24.

Let (M¯n+1,g¯,f)\bar{M}^{n+1},\bar{g},f) be a complete gradient shrinking Ricci soliton with constant scalar curvature and let MnM^{n} be a complete CWMC hypersurface in M¯n+1\bar{M}^{n+1}. Suppose that ¯¯f(N,N)k1\bar{\nabla}\bar{\nabla}f(N,N)\geq k_{1} or trMn¯¯fk2\text{tr}_{M^{n}}\bar{\nabla}\bar{\nabla}f\leq k_{2} where k1k_{1} and k2k_{2} are constants. Then MnM^{n} is properly immersed if and only if MnM^{n} has finite weighted volume.

Proof.

When ¯¯f(N,N)k1\bar{\nabla}\bar{\nabla}f(N,N)\geq k_{1}, the result follows from Theorem 1.3 in [13].

Now suppose that trMn¯¯fk2\text{tr}_{M^{n}}\bar{\nabla}\bar{\nabla}f\leq k_{2}. Using the fact that R¯ic+¯¯f=λg¯\bar{R}ic+\bar{\nabla}\bar{\nabla}f=\lambda\bar{g} we have

(n+1)λ\displaystyle(n+1)\lambda =\displaystyle= R¯+Δ¯f\displaystyle\bar{R}+\bar{\Delta}f
=\displaystyle= R¯+trMn¯¯f+¯¯f(N,N).\displaystyle\bar{R}+\text{tr}_{M^{n}}\bar{\nabla}\bar{\nabla}f+\bar{\nabla}\bar{\nabla}f(N,N).

We see that

¯¯f(N,N)(n+1)λR¯k2.\bar{\nabla}\bar{\nabla}f(N,N)\geq(n+1)\lambda-\bar{R}-k_{2}.

Using Theorem 1.3 in [13] we get the result. ∎

Proof of Corollary 2.

Proof.

By Example 16 we have λ=12\lambda=\frac{1}{2} and C=R¯C=\bar{R} (since CR¯=|¯f|22λfC-\bar{R}=\left|\bar{\nabla}f\right|^{2}-2\lambda f). We have ¯¯f=12gRn+1k\bar{\nabla}\bar{\nabla}f=\frac{1}{2}g_{R^{n+1-k}} and

trMn¯¯f\displaystyle tr_{M^{n}}\bar{\nabla}\bar{\nabla}f =Δ¯f¯¯f(N,N)\displaystyle=\bar{\Delta}f-\bar{\nabla}\bar{\nabla}f\left(N,N\right)
=n+1k2¯¯f(N,N).\displaystyle=\frac{n+1-k}{2}-\bar{\nabla}\bar{\nabla}f\left(N,N\right).

In particular

nk2trMn¯¯fn+1k2.\frac{n-k}{2}\leq tr_{M^{n}}\bar{\nabla}\bar{\nabla}f\leq\frac{n+1-k}{2}.

For D(|Hf|,a)D^{-}(|H_{f}|,a) with a=nk2a=\frac{n-k}{2} we have

Γ=14(|Hf|+Hf2+2(nk))2,\Gamma^{-}=\frac{1}{4}\left(-\left|H_{f}\right|+\sqrt{H_{f}^{2}+2\left(n-k\right)}\right)^{2},

and for D+(|Hf|,b)D^{+}(|H_{f}|,b) with b=nk+12b=\frac{n-k+1}{2} we have

Γ+=14(|Hf|+Hf2+2(n+1k))2.\Gamma^{+}=\frac{1}{4}\left(\left|H_{f}\right|+\sqrt{H_{f}^{2}+2\left(n+1-k\right)}\right)^{2}.

Since MnM^{n} is properly immersed we know that MnM^{n} has finite weighted volume (see Lemma 24).

Proof of Item (a). Using the fact that f(x,y)=|x|24f(x,y)=\frac{|x|^{2}}{4} we have D(|Hf|,a)=Brn+1k(0)×S2(k1)k(0)D^{-}(|H_{f}|,a)=B^{n+1-k}_{r}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}(0) and D(|Hf|,a)=Srnk(0)×S2(k1)k(0)\partial D^{-}(|H_{f}|,a)=S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}(0) where r=|Hf|+Hf2+2(nk)r=-|H_{f}|+\sqrt{H_{f}^{2}+2\left(n-k\right)}. Using the fact that MnM^{n} lies in B¯rn+1k(0)×S2(k1)k(0)\bar{B}^{n+1-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0) and Item (a) of Theorem 1 we have MnSrnk(0)×S2(k1)k(0)M^{n}\subseteq S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}(0). Since MnM^{n} is complete we have Mn=Srnk(0)×S2(k1)k(0)M^{n}=S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}(0). On the other hand, by Corollary 22 we have r=Hf+Hf2+2(nk)r=-H_{f}+\sqrt{H_{f}^{2}+2\left(n-k\right)}, so Hf0H_{f}\geq 0.

Proof of Item (b). Using the fact that f(x,y)=|x|24f(x,y)=\frac{|x|^{2}}{4} we have D+(|Hf|,b)=Brnk+1(0)×S2(k1)k(0)D^{+}(|H_{f}|,b)=B^{n-k+1}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0) and D+(|Hf|,b)=Srnk(0)×S2(k1)k(0)\partial D^{+}(|H_{f}|,b)=S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0) where r=|Hf|+Hf2+2(n+1k)r=|H_{f}|+\sqrt{H_{f}^{2}+2\left(n+1-k\right)}. Using the fact that MnM^{n} lies outside Brn+1k(0)×S2(k1)k(0)B^{n+1-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0) and Item (b) of Theorem 1 we have MnSrnk(0)×S2(k1)k(0)M^{n}\subseteq S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0). Since MnM^{n} is complete we have Mn=Srnk(0)×S2(k1)k(0)M^{n}=S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2(k-1)}}^{k}(0). On the other hand, by Corollary 22 we have r=Hf+Hf2+2(nk)r=-H_{f}+\sqrt{H_{f}^{2}+2\left(n-k\right)}. Therefore

Hf+Hf2+2(nk)\displaystyle-H_{f}+\sqrt{H_{f}^{2}+2(n-k)} =\displaystyle= |Hf|+Hf2+2(n+1k)\displaystyle|H_{f}|+\sqrt{H_{f}^{2}+2(n+1-k)}
>\displaystyle> |Hf|+Hf2+2(nk).\displaystyle|H_{f}|+\sqrt{H_{f}^{2}+2(n-k)}.

Hence

Hf>|Hf|,-H_{f}>|H_{f}|,

which is a contradiction. ∎

For completeness we include an application of Theorem 1 to the Gaussian shrinking soliton ambient space. We omit the proof (it is similar to the proof of Corollary 2).

Corollary 25.

Let MnM^{n} be a complete CWMC hypersurface properly immersed in the Gaussian shrinking Ricci soliton n+1\mathbb{R}^{n+1}.

(a) If MnM^{n} lies inside B¯rn+1(0)\bar{B}^{n+1}_{r}(0) with r=|Hf|+Hf2+2nr=-|H_{f}|+\sqrt{H_{f}^{2}+2n}, then Hf0H_{f}\geq 0 and Mn=Srn(0)M^{n}=S^{n}_{r}(0).

(b) If MnM^{n} lies outside Brn+1(0)B^{n+1}_{r}(0) with r=|Hf|+Hf2+2nr=|H_{f}|+\sqrt{H_{f}^{2}+2n}, then Hf0H_{f}\leq 0 and Mn=Srn(0)M^{n}=S^{n}_{r}(0).

Proof of Theorem 3.

Proof.

Since MnM^{n} is a compact hypersurface, the potential function achieves its maximum in some pMp\in M. From the fact that the gradient of ff points in the direction of greatest increase and f(p)=0\nabla f(p)=0, we see that ¯f(p)\bar{\nabla}f(p) and N(p)N(p) have the same direction. Therefore

Hf+|¯f|(p)\displaystyle H_{f}+|\bar{\nabla}f|(p) =\displaystyle= Hf+¯f(p),N(p)\displaystyle H_{f}+\langle\bar{\nabla}f(p),N(p)\rangle
=\displaystyle= H(p).\displaystyle H(p).

Using the fact that |¯f|2+R¯2λf=C\left|\bar{\nabla}f\right|^{2}+\bar{R}-2\lambda f=C and the hypothesis on the mean curvature, we get

2λf(p)+CR¯=|¯f|(p)\displaystyle\sqrt{2\lambda f(p)+C-\bar{R}}=|\bar{\nabla}f|(p) \displaystyle\leq 2Hf|Hf|+Hf2+4a2Hf\displaystyle\frac{2H_{f}-|H_{f}|+\sqrt{H_{f}^{2}+4a}}{2}-H_{f}
=\displaystyle= |Hf|+Hf2+4a2.\displaystyle\frac{-|H_{f}|+\sqrt{H_{f}^{2}+4a}}{2}.

Therefore

ff(p)12λ{(|Hf|+Hf2+4a2)2+R¯C},\displaystyle f\leq f(p)\leq\frac{1}{2\lambda}\left\{\left(\frac{-|H_{f}|+\sqrt{H_{f}^{2}+4a}}{2}\right)^{2}+\bar{R}-C\right\},

which implies that MnM^{n} lies in D(|Hf|,a)¯\overline{D^{-}(|H_{f}|,a)}. By Theorem 1, we conclude that MnD(|Hf|,a)M^{n}\subseteq\partial D^{-}(|H_{f}|,a). ∎

Proof of Corollary 5.

Proof.

Take a=nk2a=\frac{n-k}{2}. As in the proof of Corollary 2 Item (a) we have trMn¯¯fatr_{M^{n}}\bar{\nabla}\bar{\nabla}f\geq a and D(Hf,a)=Srnk(0)×S2(k1)k(0)\partial D^{-}(H_{f},a)=S^{n-k}_{r}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}(0) where r=Hf+Hf2+2(nk)r=-H_{f}+\sqrt{H_{f}^{2}+2\left(n-k\right)}. The result follows from Theorem 3. ∎

Proof of Theorem 7.

Proof.

We compute

Δf\displaystyle\Delta f =\displaystyle= trM¯¯f¯f,NH\displaystyle\text{tr}_{M}\bar{\nabla}\bar{\nabla}f-\langle\bar{\nabla}f,N\rangle H
=\displaystyle= trM¯¯f+(HfH)H\displaystyle\text{tr}_{M}\bar{\nabla}\bar{\nabla}f+(H_{f}-H)H
\displaystyle\leq b+(HfH)H\displaystyle b+(H_{f}-H)H
=\displaystyle= Hf2+4b4(HHf2)20.\displaystyle\frac{H_{f}^{2}+4b}{4}-\left(H-\frac{H_{f}}{2}\right)^{2}\leq 0.

In the third line we used the assumption on ¯¯f\bar{\nabla}\bar{\nabla}f and in the fourth line we used the assumption on HH. Let φCc(Mn)\varphi\in C_{c}^{\infty}(M^{n}). Integrating by parts, we have

Mφ2|f|2ef\displaystyle\int_{M}\varphi^{2}|\nabla f|^{2}e^{-f} \displaystyle\leq Mφ2(|f|2Δf)ef\displaystyle\int_{M}\varphi^{2}(|\nabla f|^{2}-\Delta f)e^{-f}
=\displaystyle= Mφ2Δffef\displaystyle-\int_{M}\varphi^{2}\Delta_{f}fe^{-f}
=\displaystyle= Mφ2,fef\displaystyle\int_{M}\langle\nabla\varphi^{2},\nabla f\rangle e^{-f}
=\displaystyle= 2Mφ,φfef\displaystyle 2\int_{M}\langle\nabla\varphi,\varphi\nabla f\rangle e^{-f}
\displaystyle\leq M[2|φ|2+12φ2|f|2]ef.\displaystyle\int_{M}\left[2|\nabla\varphi|^{2}+\frac{1}{2}\varphi^{2}|\nabla f|^{2}\right]e^{-f}.

Therefore

12Mφ2|f|2ef2M|φ|2ef.\frac{1}{2}\int_{M}\varphi^{2}|\nabla f|^{2}e^{-f}\leq 2\int_{M}|\nabla\varphi|^{2}e^{-f}.

Fix p0Mnp_{0}\in M^{n} and consider a sequence φjC0(Mn)\varphi_{j}\in C_{0}^{\infty}(M^{n}) such that φj=1\varphi_{j}=1 on BjM(p0)B^{M}_{j}(p_{0}), φj=0\varphi_{j}=0 on MnB2jM(p0)M^{n}\setminus B^{M}_{2j}(p_{0}) and |φj|1j|\nabla\varphi_{j}|\leq\frac{1}{j}. Using the monotone convergence theorem and the fact that the weighted volume is finite we have

M|f|2ef=0,\int_{M}|\nabla f|^{2}e^{-f}=0,

which implies that ff is constant on MnM^{n}. ∎

Using Theorem 7, Lemma 24 and Theorem 17 we obtain a new result for ambient spaces which are shrinking Ricci solitons with constant scalar curvature. Note that taking b=n2b=\frac{n}{2}, λ=12\lambda=\frac{1}{2}, R¯=0\bar{R}=0 and C=0C=0 in the next result we recover Theorem 1.2 in [7].

Corollary 26.

Let M¯n+1\bar{M}^{n+1} be a shrinking Ricci soliton with constant scalar curvature and let MnM^{n} be a complete CWMC hypersurface properly immersed in M¯n+1\bar{M}^{n+1}. Suppose that trM¯¯fb\text{tr}_{M}\bar{\nabla}\bar{\nabla}f\leq b for some b>0b>0. If

HHf+Hf2+4b2,H\geq\frac{H_{f}+\sqrt{H_{f}^{2}+4b}}{2},

then Mnf1(γ)M^{n}\subseteq f^{-1}(\gamma) for some γ>0\gamma>0. Moreover, if f1(γ)f^{-1}(\gamma) is connected and complete then Mn=f1(γ)M^{n}=f^{-1}(\gamma) where

γ=12λ{14(Hf+Hf2+4(nλR¯))2+R¯C}.\gamma=\frac{1}{2\lambda}\left\{\frac{1}{4}\left(-H_{f}+\sqrt{H_{f}^{2}+4(n\lambda-\bar{R})}\right)^{2}+\bar{R}-C\right\}.
Proof.

Since MnM^{n} is properly immersed we know that MnM^{n} has finite weighted volume (see Lemma 24). Using Theorem 7 and Lemma 24 we have Mnf1(γ)M^{n}\subseteq f^{-1}(\gamma). Suppose that f1(γ)f^{-1}(\gamma) is connected and complete. Then Mn=f1(γ)M^{n}=f^{-1}(\gamma). Using Theorem 17 we get the conclusion of γ\gamma. ∎

Proof of Corollary 8.

Proof.

Assuming that the conclusion of the result is false we will show that this leads to a contradiction. Take b=n+1k2b=\frac{n+1-k}{2}. If the conclusion is false we have

HHf+Hf2+4b2.H\geq\frac{H_{f}+\sqrt{H_{f}^{2}+4b}}{2}.

As in the proof Corollary 2 we have trM¯¯fb\text{tr}_{M}\bar{\nabla}\bar{\nabla}f\leq b, λ=12\lambda=\frac{1}{2} and C=R¯=k2C=\bar{R}=\frac{k}{2}. Using Corollary 26 we have Mn=f1(γ)M^{n}=f^{-1}(\gamma) where

γ=14(Hf+Hf2+2(nk))2.\gamma=\frac{1}{4}\left(-H_{f}+\sqrt{H_{f}^{2}+2\left(n-k\right)}\right)^{2}.

Using the fact that f(x,y)=|x|24f(x,y)=\frac{\left|x\right|^{2}}{4} we see that Mn=Srnk(0)×S2(k1)k(0)M^{n}=S_{r}^{n-k}\left(0\right)\times S_{\sqrt{2\left(k-1\right)}}^{k}\left(0\right) with r=Hf+Hf2+2(nk)r=-H_{f}+\sqrt{H_{f}^{2}+2\left(n-k\right)}. However for MnM^{n} we have (see Corollary 5)

H=Hf+Hf2+2(nk)2<Hf+Hf2+2(n+1k)2=Hf+Hf2+4b2,H=\frac{H_{f}+\sqrt{H_{f}^{2}+2(n-k)}}{2}<\frac{H_{f}+\sqrt{H_{f}^{2}+2(n+1-k)}}{2}=\frac{H_{f}+\sqrt{H_{f}^{2}+4b}}{2},

a contradiction. ∎

4. Constant weighted mean curvature hypersurfaces in n+1\mathbb{R}^{n+1}

In this section we prove Theorem 10 and Corollary 12.

Proof of Theorem 10.

Proof.

By [17] we have

Δf|A|2=2(12|A|2)|A|22HftrA3+2|A|2.\Delta_{f}|A|^{2}=2\left(\frac{1}{2}-|A|^{2}\right)|A|^{2}-2H_{f}\text{tr}A^{3}+2|\nabla A|^{2}.

Considering q1q\geq 1 as in the hypothesis, we have

|A|qΔf|A|2\displaystyle|A|^{q}\Delta_{f}|A|^{2} =\displaystyle= 2(12|A|2)|A|q+22|A|qHftrA3+2|A|q|A|2\displaystyle 2\left(\frac{1}{2}-|A|^{2}\right)|A|^{q+2}-2|A|^{q}H_{f}\text{tr}A^{3}+2|A|^{q}|\nabla A|^{2}
\displaystyle\geq 2(12|A|2)|A|q+22|Hf||A|q+3+2|A|q|A|2\displaystyle 2\left(\frac{1}{2}-|A|^{2}\right)|A|^{q+2}-2|H_{f}||A|^{q+3}+2|A|^{q}|\nabla A|^{2}
=\displaystyle= |A|q+2(12|A|22|Hf||A|)+2|A|q|A|2.\displaystyle|A|^{q+2}(1-2|A|^{2}-2|H_{f}||A|)+2|A|^{q}|\nabla A|^{2}.

Note that since

|A|Hf2+2|Hf|2,|A|\leq\frac{\sqrt{H_{f}^{2}+2}-|H_{f}|}{2},

we have that 12|A|22|Hf||A|01-2|A|^{2}-2|H_{f}||A|\geq 0. Therefore

|A|qΔf|A|22|A|q|A|2.|A|^{q}\Delta_{f}|A|^{2}\geq 2|A|^{q}|\nabla A|^{2}.

Let φCc(Mn)\varphi\in C_{c}^{\infty}(M^{n}). Integrating by parts we have

2Mφ2|A|q|A|2ef\displaystyle 2\int_{M}\varphi^{2}|A|^{q}|\nabla A|^{2}e^{-f} \displaystyle\leq Mφ2|A|qΔf|A|2ef\displaystyle\int_{M}\varphi^{2}|A|^{q}\Delta_{f}|A|^{2}e^{-f}
=\displaystyle= Mφ2|A|q,|A|2ef\displaystyle-\int_{M}\langle\nabla\varphi^{2}|A|^{q},\nabla|A|^{2}\rangle e^{-f}
=\displaystyle= 2Mq|A|qφ2||A||2ef4M|A|q2+1φ,φ|A|q2|A|ef\displaystyle-2\int_{M}q|A|^{q}\varphi^{2}|\nabla|A||^{2}e^{-f}-4\int_{M}\langle|A|^{\frac{q}{2}+1}\nabla\varphi,\varphi|A|^{\frac{q}{2}}\nabla|A|\rangle e^{-f}
\displaystyle\leq 2Mq|A|qφ2||A||2ef+4M|A|q2+1|φ|φ|A|q2||A||ef.\displaystyle-2\int_{M}q|A|^{q}\varphi^{2}|\nabla|A||^{2}e^{-f}+4\int_{M}|A|^{\frac{q}{2}+1}|\nabla\varphi|\varphi|A|^{\frac{q}{2}}|\nabla|A||e^{-f}.

Here in the last line we use the Cauchy-Schwarz inequality. From the identity 2abεa2+1εb22ab\leq\varepsilon a^{2}+\frac{1}{\varepsilon}b^{2}, choosing a=|A|q2+1|φ|a=|A|^{\frac{q}{2}+1}|\nabla\varphi| and b=φ|A|q2||A||b=\varphi|A|^{\frac{q}{2}}|\nabla|A|| we have

2Mφ2|A|q|A|2ef2εM|A|q+2|φ|2ef+(2ε2q)Mφ2|A|q||A||2ef.2\int_{M}\varphi^{2}|A|^{q}|\nabla A|^{2}e^{-f}\leq 2\varepsilon\int_{M}|A|^{q+2}|\nabla\varphi|^{2}e^{-f}+\left(\frac{2}{\varepsilon}-2q\right)\int_{M}\varphi^{2}|A|^{q}|\nabla|A||^{2}e^{-f}.

Choosing ε=1q\varepsilon=\frac{1}{q}, we get

Mφ2|A|q|A|2ef1qM|A|q+2|φ|2ef.\displaystyle\int_{M}\varphi^{2}|A|^{q}|\nabla A|^{2}e^{-f}\leq\frac{1}{q}\int_{M}|A|^{q+2}|\nabla\varphi|^{2}e^{-f}.

Since |A|2C|A|^{2}\leq C, we have

Mφ2|A|q|A|2efCqM|A|q|φ|2ef.\int_{M}\varphi^{2}|A|^{q}|\nabla A|^{2}e^{-f}\leq\frac{C}{q}\int_{M}|A|^{q}|\nabla\varphi|^{2}e^{-f}.

Let us fix p0Mp_{0}\in M and consider the sequence φj\varphi_{j} such that φj=1\varphi_{j}=1 on BjM(p0)B^{M}_{j}(p_{0}), φj=0\varphi_{j}=0 on MB2jM(p0)M\setminus B^{M}_{2j}(p_{0}) and |φj|1j|\nabla\varphi_{j}|\leq\frac{1}{j}. Using the monotone convergence theorem and the assumption that ALfq(M)A\in L^{q}_{f}(M), we conclude that

|A||A|=0.|A||\nabla A|=0.

Let 𝒜={xM;|A|(x)=0}\mathcal{A}=\{x\in M;|A|(x)=0\}. The set M𝒜M\setminus\mathcal{A} is open and since

|A|||A|||A||A|=0,|A||\nabla|A||\leq|A||\nabla A|=0,

we see that |A||A| is constant on M𝒜M\setminus\mathcal{A}. If 𝒜\mathcal{A}\neq\emptyset, then using continuity we conclude that |A|=0|A|=0, which implies that MnM^{n} is a hyperplane. If 𝒜=\mathcal{A}=\emptyset, then |A|=0|\nabla A|=0. In this case using Theorem 4 in Lawson [20] and the fact that MnM^{n} is complete we conclude that MnM^{n} is a generalized cylinder. ∎

Proof of Corollary 12.

Proof.

We estimate the Bakry-Emery-Ricci tensor RicfRic_{f} of a CWMC hypersurface. Let pMp\in M and choose a orthonormal basis of TpMT_{p}M such that ejei(p)=0\nabla_{e_{j}}e_{i}(p)=0. Let hij=A(ei,ej)h_{ij}=A(e_{i},e_{j}). We have

(Ricf)ij\displaystyle(Ric_{f})_{ij} =\displaystyle= Rij+eiejf\displaystyle R_{ij}+\nabla_{e_{i}}\nabla_{e_{j}}f
=\displaystyle= Rij+¯ei¯ejf+¯f,A(ei,ej)N\displaystyle R_{ij}+\bar{\nabla}_{e_{i}}\bar{\nabla}_{e_{j}}f+\langle\bar{\nabla}f,A(e_{i},e_{j})N\rangle
=\displaystyle= Rij+12δij+x,N2hij\displaystyle R_{ij}+\frac{1}{2}\delta_{ij}+\frac{\langle x,N\rangle}{2}h_{ij}
=\displaystyle= Hhijl=1nhilhlj+x,N2hij+12δij\displaystyle-Hh_{ij}-\sum_{l=1}^{n}h_{il}h_{lj}+\frac{\langle x,N\rangle}{2}h_{ij}+\frac{1}{2}\delta_{ij}
=\displaystyle= Hfhijl=1nhilhlj+12δij\displaystyle-H_{f}h_{ij}-\sum_{l=1}^{n}h_{il}h_{lj}+\frac{1}{2}\delta_{ij}
\displaystyle\geq |Hf||hij|l=1nhilhlj+12δij\displaystyle-|H_{f}||h_{ij}|-\sum_{l=1}^{n}h_{il}h_{lj}+\frac{1}{2}\delta_{ij}
\displaystyle\geq |Hf||A|l=1nhilhlj+12δij.\displaystyle-|H_{f}||A|-\sum_{l=1}^{n}h_{il}h_{lj}+\frac{1}{2}\delta_{ij}.

In the fourth line we used the Gauss equation. Since sup|A|<Hf2+2|Hf|2\sup|A|<\frac{\sqrt{H_{f}^{2}+2}-|H_{f}|}{2} we have

Ricf|A|2|Hf||A|+12(sup|A|)2|Hf|sup|A|+12>0.Ric_{f}\geq-|A|^{2}-|H_{f}||A|+\frac{1}{2}\geq-(\sup|A|)^{2}-|H_{f}|\sup|A|+\frac{1}{2}>0.

By Theorem 4.1 in [28], the condition above implies that MnM^{n} has finite weighted volume. Since |A|q|A|^{q} is bounded and MnM^{n} has finite weighted volume, we see that |A|Lfq(M)|A|\in L_{f}^{q}(M). Applying Theorem 10 we conclude that MnM^{n} is a hyperplane. ∎

References

  • [1] S. Ancari and I. Miranda. Rigidity theorems for complete λ\lambda-hypersurfaces. Archiv der Mathematik, pages 1–16, 2021.
  • [2] E. Barbosa, F. Santana, and A. Upadhyay. λ\lambda-hypersurfaces on shrinking gradient ricci solitons. Journal of Mathematical Analysis and Applications, 492(2):124470, 2020.
  • [3] J. Bernstein and L. Wang. Smooth compactness for spaces of asymptotically conical self-expanders of mean curvature flow. International Mathematics Research Notices, 2021(12):9016–9044, 2021.
  • [4] H.-D. Cao. Recent progress on ricci solitons. arXiv preprint arXiv:0908.2006, 2009.
  • [5] H.-D. Cao and H. Li. A gap theorem for self-shrinkers of the mean curvature flow in arbitrary codimension. Calculus of Variations and Partial Differential Equations, 46(3-4):879–889, 2013.
  • [6] M. P. Cavalcante and J. M. Espinar. Halfspace type theorems for self-shrinkers. Bulletin of the London Mathematical Society, 48(2):242–250, 2016.
  • [7] Q.-M. Cheng, S. Ogata, and G. Wei. Rigidity theorems of λ\lambda-hypersurfaces. Communications in Analysis and Geometry, 24(1):45–58, 2016.
  • [8] Q.-M. Cheng and Y. Peng. Complete self-shrinkers of the mean curvature flow. Calculus of Variations and Partial Differential Equations, 52(3-4):497–506, 2015.
  • [9] Q.-M. Cheng and G. Wei. Complete λ\lambda-hypersurfaces of weighted volume-preserving mean curvature flow. Calculus of Variations and Partial Differential Equations, 57(2):32, 2018.
  • [10] Q.-M. Cheng and G. Wei. Complete λ\lambda-surfaces in 𝟛\mathbb{R^{3}}. Calculus of Variations and Partial Differential Equations, 60(1):1–19, 2021.
  • [11] S. Y. Cheng and S.-T. Yau. Differential equations on riemannian manifolds and their geometric applications. Communications on Pure and Applied Mathematics, 28(3):333–354, 1975.
  • [12] X. Cheng, T. Mejia, and D. Zhou. Simons-type equation for f-minimal hypersurfaces and applications. The Journal of Geometric Analysis, 25(4):2667–2686, 2015.
  • [13] X. Cheng, M. Vieira, and D. Zhou. Volume growth of complete submanifolds in gradient ricci solitons with bounded weighted mean curvature. International Mathematics Research Notices, 2019.
  • [14] X. Cheng and D. Zhou. Stability properties and gap theorem for complete f-minimal hypersurfaces. Bulletin of the Brazilian Mathematical Society, New Series, 46(2):251–274, 2015.
  • [15] T. H. Colding and W. P. Minicozzi. Generic mean curvature flow i; generic singularities. Annals of mathematics, pages 755–833, 2012.
  • [16] M. Fernández-López and E. García-Río. On gradient ricci solitons with constant scalar curvature. Proceedings of the American Mathematical Society, 144(1):369–378, 2016.
  • [17] Q. Guang. Gap and rigidity theorems of λ\lambda-hypersurfaces. Proceedings of the American Mathematical Society, 146(10):4459–4471, 2018.
  • [18] R. Hamilton. The formations of singularities in the ricci flow. Surveys in differential geometry, 2(1):7–136, 1993.
  • [19] D. Hoffman, T. Ilmanen, F. Martín, and B. White. Notes on translating solitons for mean curvature flow. In Minimal Surfaces: Integrable Systems and Visualisation, pages 147–168. Springer, 2017.
  • [20] H. B. Lawson. Local rigidity theorems for minimal hypersurfaces. Annals of Mathematics, pages 187–197, 1969.
  • [21] N. Q. Le and N. Sesum. Blow-up rate of the mean curvature during the mean curvature flow and a gap theorem for self-shrinkers. Communications in Analysis and Geometry, 19(4):633–659, 2011.
  • [22] R. López. Invariant surfaces in euclidean space with a log-linear density. Advances in Mathematics, 339:285–309, 2018.
  • [23] M. McGonagle and J. Ross. The hyperplane is the only stable, smooth solution to the isoperimetric problem in gaussian space. Geometriae Dedicata, 178(1):277–296, 2015.
  • [24] P. Petersen and W. Wylie. Rigidity of gradient ricci solitons. Pacific journal of mathematics, 241(2):329–345, 2009.
  • [25] S. Pigola and M. Rimoldi. Complete self-shrinkers confined into some regions of the space. Annals of Global Analysis and Geometry, 45(1):47–65, 2014.
  • [26] S. Pigola, M. Rimoldi, and A. G. Setti. Remarks on non-compact gradient ricci solitons. Mathematische Zeitschrift, 268(3):777–790, 2011.
  • [27] M. Vieira and D. Zhou. Geometric properties of self-shrinkers in cylinder shrinking ricci solitons. The Journal of Geometric Analysis, 28(1):170–189, 2018.
  • [28] G. Wei and W. Wylie. Comparison geometry for the bakry-emery ricci tensor. Journal of differential geometry, 83(2):377–406, 2009.
  • [29] Z.-H. Zhang. On the completeness of gradient ricci solitons. Proceedings of the American Mathematical Society, 137(8):2755–2759, 2009.