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

\UseRawInputEncoding

On triharmonic hypersurfaces in space forms

Yu Fu School of Data Science and Artificial Intelligence, Dongbei University of Finance and Economics, Dalian 116025, P. R. China [email protected]  and  Dan Yang School of Mathematics and Statistics, Liaoning University, Shenyang, 110036, China [email protected]
Abstract.

In this paper we study triharmonic hypersurfaces immersed in a space form Nn+1(c)N^{n+1}(c). We prove that any proper CMC triharmonic hypersurface in the sphere 𝕊n+1\mathbb{S}^{n+1} has constant scalar curvature; any CMC triharmonic hypersurface in the hyperbolic space n+1\mathbb{H}^{n+1} is minimal. Moreover, we show that any CMC triharmonic hypersurface in the Euclidean space n+1\mathbb{R}^{n+1} is minimal provided that the multiplicity of the principal curvature zero is at most one. In particular, we are able to prove that every CMC triharmonic hypersurface in the Euclidean space 6\mathbb{R}^{6} is minimal. These results extend some recent works due to Montaldo-Oniciuc-Ratto and Chen-Guan, and give affirmative answer to the generalized Chen’s conjecture.

Key words and phrases:
kk-harmonic maps, Triharmonic hypersurfaces, constant mean curvature, constant scalar curvature
1991 Mathematics Subject Classification:
Primary 53C40, 58E20; Secondary 53C42

1. Introduction

Let ϕ:(M,g)(N,g¯)\phi:(M,g)\rightarrow(N,{\bar{g}}) be a smooth map between Riemannian manifolds MM and NN. A kk-harmonic map ϕ\phi, proposed by Eells and Lemaire [8], is a critical point of the kk-energy functional

Ek(ϕ)=12M|(d+d)kϕ|2vg.\displaystyle E_{k}(\phi)=\frac{1}{2}\int_{M}\big{|}(d+d^{*})^{k}\phi\big{|}^{2}v_{g}.

The Euler-Lagrange equation is given by τk(ϕ)0\tau_{k}(\phi)\equiv 0 , where τk(ϕ)\tau_{k}(\phi) is the kk-tension field. The concept of kk-harmonic map is a natural generalization of harmonic map, and in particular for k=2k=2 and k=3k=3, the critical points of E2E_{2} or E3E_{3} are biharmonic or triharmonic maps, respectively (c.f.[24], [25],[11], [12]).

In recent years, biharmonic maps and biharmonic submanifolds have been widely studied, see [10, 3, 19] and the references therein. There are also a lot of results on triharmonic maps and triharmonic submanifolds (c.f.[12, 11, 13, 14, 17, 16, 1, 2]). Concerning triharmonic hypersurfaces in a space form Nn+1(c)N^{n+1}(c), Maeta [11] proved that any compact CMC triharmonic hypersurface in Nn+1(c)N^{n+1}(c) for c0c\leq 0 is minimal. Recently, Montaldo-Oniciuc-Ratto [15] gave a systematic study of CMC triharmonic hypersurface in space forms. The authors proved that

Theorem 1.1.

([15]) Let MnM^{n} be a CMC triharmonic hypersurface in Nn+1(c)N^{n+1}(c) with c0c\leq 0 and assume that the squared norm of the second fundamental form SS is constant. Then MnM^{n} is minimal.

In particular, the assumption SS being constant was removed for n=2n=2 in [15].

Theorem 1.2.

([15]) Let M2M^{2} be a CMC triharmonic surface in N3(c)N^{3}(c). Then M2M^{2} is minimal if c0c\leq 0 and M2M^{2} is an open part of the small hypersphere if c>0c>0.

Very recently, Chen-Guan investigated triharmonic CMC hypersurfaces in a space form Nn+1(c)N^{n+1}(c) under some assumptions on the number of distinct principal curvatures in [6, 7].

Theorem 1.3.

([6]) Let MnM^{n} (n3n\geq 3) be a CMC proper triharmonic hypersurface with at most three distinct principal curvatures in Nn+1(c)N^{n+1}(c). Then MnM^{n} has constant scalar curvature.

Theorem 1.4.

([7]) Let MnM^{n} (n4n\geq 4) be a CMC proper triharmonic hypersurface with four distinct principal curvatures in Nn+1(c)N^{n+1}(c). If zero is a principal curvature with multiplicity at most one, then MnM^{n} has constant scalar curvature.

We recall Chen’s conjecture in the literature of biharmonic submanifolds: any biharmonic submanifold in the Euclidean space n+1\mathbb{R}^{n+1} is minimal. There are some important progress in recent years to support the conjecture under some geometric restrictions (see, for instances [9, 19]), however the general case remains open. Taking into account Chen’s conjecture, Maeta [11] further proposed the generalized Chen’s conjecture on kk-harmonic submanifolds:

Conjecture : Any kk-harmonic submanifold in the Euclidean space n+1\mathbb{R}^{n+1} is minimal.

In this paper, we are able to determine the geometry of CMC triharmonic hypersurfaces in a space form Nn+1(c)N^{n+1}(c) without the restrictions on the number of principal curvatures. We will prove the following statements:

Theorem 1.5.

Let MnM^{n} be a CMC triharmonic hypersurface in the hyperbolic space n+1\mathbb{H}^{n+1}. Then MnM^{n} is minimal.

Moreover, restricting ourselves on the case of c>0c>0, we get

Theorem 1.6.

Let MnM^{n} be a CMC proper triharmonic hypersurface in the sphere 𝕊n+1\mathbb{S}^{n+1}. Then MnM^{n} has constant scalar curvature.

Combining our Theorem 1.6 with the results of Montaldo-Oniciuc-Ratto (Theorem 1.9, [15]) and Chen-Guan (Corollary 1.7, [7]), we provide a more general result for CMC triharmonic hypersurfaces in 𝕊n+1\mathbb{S}^{n+1}.

Corollary 1.7.

Let MnM^{n} be a CMC proper triharmonic hypersurface in the sphere 𝕊n+1\mathbb{S}^{n+1}. Then either
(1)   H2=2H^{2}=2 and MnM^{n} is an open part of Sn(1/3)S^{n}(1/\sqrt{3}), or
(2) H2(0,t0]H^{2}\in(0,t_{0}] and H2=t0H^{2}=t_{0} if and only MnM^{n} is an open part of Sn1(a)×S1(1a2)S^{n-1}(a)\times S^{1}(\sqrt{1-a^{2}}), where aa is given by

a2=2(n1)2n2H2+2n(n1)+nHn2H2+4(n1)a^{2}=\frac{2(n-1)^{2}}{n^{2}H^{2}+2n(n-1)+nH\sqrt{n^{2}H^{2}+4(n-1)}}

and t0t_{0} is the unique real root belonging to (0,2)(0,2) of the polynomial

fn=n4t32n2(n25n+5)t2(n1)(2n5)(3n5)t(n1)(n2)2.f_{n}=n^{4}t^{3}-2n^{2}(n^{2}-5n+5)t^{2}-(n-1)(2n-5)(3n-5)t-(n-1)(n-2)^{2}.

Let us recall the generalized Chern conjecture (c.f.[4, 5]), which says that: any closed hypersurface in the unit sphere 𝕊n+1\mathbb{S}^{n+1} with constant mean curvature and constant scalar curvature is isoparametric. Since the class of CMC proper triharmonic hypersurfaces in a sphere have constant scalar curvature, the next important problem is to study whether these hypersurfaces are isoparametric. The problem remains open in its full generality. The readers may refer to the recent important progress on the generalized Chern conjecture due to Tang and Yan et al. [22, 23].

Considering the case c=0c=0, we obtain a characterization under an assumption on the multiplicity of zero principal curvature.

Theorem 1.8.

Let MnM^{n} be a CMC triharmonic hypersurface in the Euclidean space n+1\mathbb{R}^{n+1}. If zero is a principal curvature with multiplicity at most one, then MnM^{n} is minimal.

In particular, we can prove

Theorem 1.9.

Any CMC triharmonic hypersurface in the Euclidean space 6\mathbb{R}^{6} is minimal.

Remark 1.10.

Note that Theorems 1.8 and 1.9 give partial affirmative answers to the generalized Chen’s Conjecture.

Remark 1.11.

The assumption that the multiplicity of the principal curvature zero is at most one was necessary in [7] for treating triharmonic hypersurfaces with four distinct principal curvatures in space forms. In our results, we only need this for c=0c=0 and n>5n>5.

At last, we point out that for a CMC proper triharmonic hypersurface in Nn+1(c)N^{n+1}(c), the two equations in (2.7) are quite similar to the equations of a proper biharmonic hypersurface in a space form, see for instance [9]. This is reasonable because the geometry property of triharmonicity is much weaker than biharmonicity. Hence, it is expected that more geometric features of triharmonic hypersurfaces could be found. Interestingly, we can achieve a complete classification of CMC triharmonic hypersurfaces in Nn+1(c)N^{n+1}(c) with c0c\neq 0. This will benefit us in studying biharmonic hypersurfaces in Nn+1(c)N^{n+1}(c).

The paper is organized as follows. In Section 2, we recall some background on the theory of triharmonic hypersurfaces in space forms and derive some useful lemmas, which are very important for us to study the geometric properties of triharmonic hypersurfaces. In Section 3, we give the proofs of Theorems 1.5 and 1.6. In Section 4, we finish the proofs of Theorems 1.8 and 1.9.

2. Preliminaries

Let Nn+1(c)N^{n+1}(c) be an (n+1)(n+1)-dimensional Riemannian space form with constant sectional curvature cc. For an isometric immersion ϕ:MnNn+1(c)\phi:M^{n}\rightarrow N^{n+1}(c), we denote by \nabla the Levi-Civita connection of MnM^{n} and ~\widetilde{\nabla} the Levi-Civita connection of Nn+1(c)N^{n+1}(c). The Riemannian curvature tensors of MnM^{n} are respectively given by

R(X,Y)Z=(XYYX[X,Y])Z,\displaystyle R(X,Y)Z=(\nabla_{X}\nabla_{Y}-\nabla_{Y}\nabla_{X}-\nabla_{[X,Y]})Z,
R(X,Y,Z,W)=R(X,Y)W,Z.\displaystyle R(X,Y,Z,W)=\langle R(X,Y)W,Z\rangle.

The Gauss and Weingarten formulae are stated, respectively, as

~XY\displaystyle\widetilde{\nabla}_{X}Y =XY+h(X,Y)ξ,\displaystyle=\nabla_{X}Y+h(X,Y)\xi,
~Xξ\displaystyle\widetilde{\nabla}_{X}\xi =AX.\displaystyle=-AX.

Here X,Y,Z,WX,Y,Z,W are tangent vector fields on MM, ξ\xi is the unit normal vector field on MM, hh is the second fundamental form of MM, and AA is the shape operator.

Let us choose an orthonormal frame {ei}i=1n\{e_{i}\}_{i=1}^{n} of MM. With this frame, define eiej=kΓijkek\nabla_{e_{i}}e_{j}=\sum_{k}\Gamma_{ij}^{k}e_{k}, where Γijk\Gamma_{ij}^{k} are the connection coefficients.

Denote by

Rijkl=\displaystyle R_{ijkl}= R(ei,ej,ek,el),hij=h(ei,ej),\displaystyle R(e_{i},e_{j},e_{k},e_{l}),\quad h_{ij}=h(e_{i},e_{j}),
hijk=\displaystyle h_{ijk}= ek(hij)h(ekei,ej)h(ei,ekej),\displaystyle e_{k}(h_{ij})-h(\nabla_{e_{k}}e_{i},e_{j})-h(e_{i},\nabla_{e_{k}}e_{j}),
=\displaystyle= ek(hij)l(Γkilhlj+Γkjlhil).\displaystyle e_{k}(h_{ij})-\sum_{l}(\Gamma_{ki}^{l}h_{lj}+\Gamma_{kj}^{l}h_{il}).

From the definition of the Gauss curvature tensor we obtain

Rijkl\displaystyle R_{ijkl} =ei(Γjlk)ej(Γilk)+m(ΓjlmΓimkΓilmΓjmk(ΓijmΓjim)Γmlk).\displaystyle=e_{i}(\Gamma_{jl}^{k})-e_{j}(\Gamma_{il}^{k})+\sum_{m}\Big{(}\Gamma_{jl}^{m}\Gamma_{im}^{k}-\Gamma_{il}^{m}\Gamma_{jm}^{k}-(\Gamma_{ij}^{m}-\Gamma_{ji}^{m})\Gamma_{ml}^{k}\Big{)}. (2.1)

Moreover, the Gauss and Codazzi equations are given, respectively, by

Rijkl\displaystyle R_{ijkl} =c(δikδjlδilδjk)+(hikhjlhilhjk),\displaystyle=c(\delta_{ik}\delta_{jl}-\delta_{il}\delta_{jk})+(h_{ik}h_{jl}-h_{il}h_{jk}), (2.2)
hijk\displaystyle h_{ijk} =hikj.\displaystyle=h_{ikj}. (2.3)

The mean curvature function HH and the squared norm of the second fundamental form SS are written respectively as

H=1ni=1nhiiandS=i,j=1nhij2.\displaystyle H=\frac{1}{n}\sum_{i=1}^{n}h_{ii}\quad{\rm and}\quad S=\sum_{i,j=1}^{n}h_{ij}^{2}. (2.4)

From the Gauss equation, the scalar curvature RR is given by

R=n(n1)c+n2H2S.\displaystyle R=n(n-1)c+n^{2}H^{2}-S. (2.5)

We recall a fundamental characterization result on CMC triharmonic hypersurfaces in Nn+1(c)N^{n+1}(c).

Proposition 2.1.

(c.f.[15]) A CMC hypersurface ϕ:MnNn+1(c)\phi:M^{n}\rightarrow N^{n+1}(c) is triharmonic if the mean curvature HH and the squared norm of the second fundamental form SS on MnM^{n} satisfy

{H(ΔS+S2ncSn2cH2)=0,HAS=0.\left\{\begin{split}&H({\rm\Delta}\,S+S^{2}-ncS-n^{2}cH^{2})=0,\\ &HA\nabla S=0.\end{split}\right. (2.6)

According to (2.6), it is clear that minimal hypersurfaces are automatically triharmonic in Nn+1(c)N^{n+1}(c). A triharmonic hypersurfaces in Nn+1(c)N^{n+1}(c) is called proper if it is not minimal.

In the following, we will consider a CMC proper hypersurface MnM^{n} in a space form Nn+1(c)N^{n+1}(c). Then (2.6) becomes

{ΔS+S2ncSn2cH2=0,AS=0.\left\{\begin{split}&{\rm\Delta}\,S+S^{2}-ncS-n^{2}cH^{2}=0,\\ &A\nabla S=0.\end{split}\right. (2.7)

For a hypersurface MnM^{n} in Nn+1(c)N^{n+1}(c), we denote by λi\lambda_{i} for 1in1\leq i\leq n its principal curvatures. The number of distinct principal curvatures is locally constant and the set of all points here is an open and dense subset of MnM^{n}. Denote by MAM_{A} this set. On a non-empty connected component of MAM_{A}, which is open, the number of distinct principal curvatures is constant. On that connected component, the multiplicities of the distinct principal curvatures are constant and hence λi\lambda_{i} are always smooth and the shape operator AA is locally diagonalizable (see[18, 20, 21]).

Denote by 𝒩:={pM:S(p)0}\mathcal{N}:=\{p\in M:\nabla S(p)\neq 0\} and 𝒩MA\mathcal{N}\subset M_{A}. If SS is constant, then 𝒩\mathcal{N} is an empty set. From now on, we assume that SS is not constant, that is 𝒩\mathcal{N}\neq\emptyset. We will work in 𝒩\mathcal{N}.

Observing from the second equation of (2.7), it is known that S\nabla S is a principal direction with the corresponding principal curvature 0. Hence, we may choose an orthonormal frame {ei}i=1n\{e_{i}\}_{i=1}^{n} such that e1e_{1} is parallel to S\nabla S and the shape operator AA is diagonalizable with respect to {ei}\{e_{i}\}, i.e., hij=λiδijh_{ij}=\lambda_{i}\delta_{ij}, where λi\lambda_{i} is the principal curvature and λ1=0\lambda_{1}=0.

Suppose that MnM^{n} has dd distinct principal curvatures μ1=0,μ2,,μd\mu_{1}=0,\mu_{2},\cdots,\mu_{d} with d4d\geq 4, that is

λi=μαwheniIα,\lambda_{i}=\mu_{\alpha}\quad{\rm when}\quad i\in I_{\alpha},

where

Iα={0βα1nβ+1,,0βαnβ}I_{\alpha}=\Big{\{}\sum_{0\leq\beta\leq\alpha-1}n_{\beta}+1,\cdots,\sum_{0\leq\beta\leq\alpha}n_{\beta}\Big{\}}

with n0=0n_{0}=0 and nα+n_{\alpha}\in\mathbb{Z}_{+} satisfying 1αdnα=n\sum_{1\leq\alpha\leq d}n_{\alpha}=n, namely, nαn_{\alpha} is the multiplicity of μα\mu_{\alpha}. For convenience, we will use the range of indices 1α,β,γ,d1\leq\alpha,\beta,\gamma,\cdots\leq d except special declaration.

We collect a lemma for later use.

Lemma 2.2.

(c.f.[7]) The connection coefficients Γijk\Gamma_{ij}^{k} satisfy:
(1) Γijk=Γikj\Gamma_{ij}^{k}=-\Gamma_{ik}^{j}.
(2) Γiik=ek(λi)λiλk\Gamma_{ii}^{k}=\frac{e_{k}(\lambda_{i})}{\lambda_{i}-\lambda_{k}} for iIαi\in I_{\alpha} and kIαk\notin I_{\alpha}.
(3) Γijk=Γjik\Gamma_{ij}^{k}=\Gamma_{ji}^{k} if the indices satisfy one of the following conditions:
(3a) i,jIαi,j\in I_{\alpha} but kIαk\notin I_{\alpha};
(3b) i,j2i,j\geq 2 and k=1k=1.
(4) Γijk=0\Gamma_{ij}^{k}=0 if the indices satisfy one of the following conditions:
(4a) j=kj=k;
(4b) i=jI1i=j\in I_{1} and kI1k\notin I_{1};
(4c) i,kIα,iki,k\in I_{\alpha},i\neq k and jIαj\notin I_{\alpha};
(4d) i,j2,iIα,jIβi,j\geq 2,i\in I_{\alpha},j\in I_{\beta} with αβ\alpha\neq\beta and k=1k=1.
(5) Γjik=λjλkλiλkΓijk,\Gamma_{ji}^{k}=\frac{\lambda_{j}-\lambda_{k}}{\lambda_{i}-\lambda_{k}}\Gamma_{ij}^{k}, Γkij=λkλjλiλjΓikj\Gamma_{ki}^{j}=\frac{\lambda_{k}-\lambda_{j}}{\lambda_{i}-\lambda_{j}}\Gamma_{ik}^{j} for λi\lambda_{i}, λj\lambda_{j} and λk\lambda_{k} are mutually different.
(6) ΓijkΓjik+ΓikjΓkij+ΓjkiΓkji=0\Gamma_{ij}^{k}\Gamma_{ji}^{k}+\Gamma_{ik}^{j}\Gamma_{ki}^{j}+\Gamma_{jk}^{i}\Gamma_{kj}^{i}=0 for λi\lambda_{i}, λj\lambda_{j} and λk\lambda_{k} are mutually different.

We first derive some crucial lemmas for studying CMC proper triharmonic hypersurfaces in a space form.

Lemma 2.3.

Denoting by Pα=e1(μα)μαP_{\alpha}=\frac{e_{1}(\mu_{\alpha})}{\mu_{\alpha}} for 2αd2\leq\alpha\leq d, we have

e1(Pα)=Pα2+c+mI1ΓααmΓ11m,\displaystyle e_{1}(P_{\alpha})=P_{\alpha}^{2}+c+\sum_{m\in I_{1}}\Gamma_{\alpha\alpha}^{m}\Gamma_{11}^{m}, (2.8)
ej(Pα)=ΓααjPαmI1ΓααmΓjm1forjI1andj1.\displaystyle e_{j}(P_{\alpha})=\Gamma_{\alpha\alpha}^{j}P_{\alpha}-\sum_{m\in I_{1}}\Gamma_{\alpha\alpha}^{m}\Gamma_{jm}^{1}\,\,\,{\rm for}\,\,j\in I_{1}\,\,{\rm and}\,\,j\neq 1. (2.9)
Proof.

For iIαi\in I_{\alpha} and α1\alpha\neq 1, it follows from (2.1) and the terms (1), (2) and (4) in Lemma 2.2 that

R1i1i=\displaystyle R_{1i1i}= e1(Γii1)ei(Γ1i1)+m(ΓiimΓ1m1Γ1imΓim1(Γ1imΓi1m)Γmi1)\displaystyle e_{1}(\Gamma_{ii}^{1})-e_{i}(\Gamma_{1i}^{1})+\sum_{m}\Big{(}\Gamma_{ii}^{m}\Gamma_{1m}^{1}-\Gamma_{1i}^{m}\Gamma_{im}^{1}-(\Gamma_{1i}^{m}-\Gamma_{i1}^{m})\Gamma_{mi}^{1}\Big{)}
=\displaystyle= e1(Γii1)+ei(Γ11i)m(ΓiimΓ11m+Γ1imΓim1+(Γ1im+Γim1)Γmi1)\displaystyle e_{1}(\Gamma_{ii}^{1})+e_{i}(\Gamma_{11}^{i})-\sum_{m}\Big{(}\Gamma_{ii}^{m}\Gamma_{11}^{m}+\Gamma_{1i}^{m}\Gamma_{im}^{1}+(\Gamma_{1i}^{m}+\Gamma_{im}^{1})\Gamma_{mi}^{1}\Big{)}
=\displaystyle= e1(Γii1)mI1(ΓiimΓ11m+Γ1imΓim1+(Γ1im+Γim1)Γmi1)\displaystyle e_{1}(\Gamma_{ii}^{1})-\sum_{m\in I_{1}}\Big{(}\Gamma_{ii}^{m}\Gamma_{11}^{m}+\Gamma_{1i}^{m}\Gamma_{im}^{1}+(\Gamma_{1i}^{m}+\Gamma_{im}^{1})\Gamma_{mi}^{1}\Big{)}
=\displaystyle= e1(Γii1)(Γii1)2mI1ΓiimΓ11m.\displaystyle e_{1}(\Gamma_{ii}^{1})-(\Gamma_{ii}^{1})^{2}-\sum_{m\in I_{1}}\Gamma_{ii}^{m}\Gamma_{11}^{m}.

On the other hand, from the Gauss equation (2.2) we get R1i1i=cR_{1i1i}=c. Therefore, we obtain (2.8).

For iIαi\in I_{\alpha} and α1\alpha\neq 1, jI1j\in I_{1} and j1j\neq 1, it follows from (2.1) and Lemma 2.2 that

Riji1\displaystyle R_{iji1} =ei(Γj1i)ej(Γi1i)+m(Γj1mΓimiΓi1mΓjmi(ΓijmΓjim)Γm1i)\displaystyle=e_{i}(\Gamma_{j1}^{i})-e_{j}(\Gamma_{i1}^{i})+\sum_{m}\Big{(}\Gamma_{j1}^{m}\Gamma_{im}^{i}-\Gamma_{i1}^{m}\Gamma_{jm}^{i}-(\Gamma_{ij}^{m}-\Gamma_{ji}^{m})\Gamma_{m1}^{i}\Big{)}
=ei(Γji1)+ej(Γii1)+m(Γjm1Γiim+Γim1Γjmi+(ΓijmΓjim)Γmi1)\displaystyle=-e_{i}(\Gamma_{ji}^{1})+e_{j}(\Gamma_{ii}^{1})+\sum_{m}\Big{(}\Gamma_{jm}^{1}\Gamma_{ii}^{m}+\Gamma_{im}^{1}\Gamma_{jm}^{i}+(\Gamma_{ij}^{m}-\Gamma_{ji}^{m})\Gamma_{mi}^{1}\Big{)}
=ej(Γii1)+mI1Γjm1ΓiimΓiijΓii1,\displaystyle=e_{j}(\Gamma_{ii}^{1})+\sum_{m\in I_{1}}\Gamma_{jm}^{1}\Gamma_{ii}^{m}-\Gamma_{ii}^{j}\Gamma_{ii}^{1}, (2.10)

which together with Riji1=0R_{iji1}=0 gives (2.9). Note that Riji1=0R_{iji1}=0 follows from the Gauss equation (2.2) directly. We thus complete the proof. ∎

Lemma 2.4.

For any q+q\in\mathbb{Z}_{+}, we have

f(q):=2αdnαμαPαq={0,whenqisodd;(q1)!!q!!(c)q2nH,whenqiseven.\displaystyle\begin{split}f(q):=\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}P^{q}_{\alpha}=\left\{\begin{array}[]{ll}\quad 0,&{\rm when}\,\,q\,\,{\rm is}\,\,{\rm odd};\\ \frac{(q-1)!!}{q!!}(-c)^{\frac{q}{2}}nH,&{\rm when}\,\,q\,\,{\rm is}\,\,{\rm even}.\end{array}\right.\end{split} (2.11)
Proof.

Since the case n1=1n_{1}=1 has been obtained in [7], we only need to prove it for n1>1n_{1}>1.

Taking into account the definition of HH from the first expression of (2.4), we have

2αdnαμα=nH.\displaystyle\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}=nH. (2.12)

Since HH is constant, differentiating (2.12) with respect to e1e_{1}, we obtain

2αdnαe1(μα)=0,\displaystyle\sum_{2\leq\alpha\leq d}n_{\alpha}e_{1}(\mu_{\alpha})=0,

which is equivalent to

2αdnαμαPα=0.\displaystyle\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}P_{\alpha}=0. (2.13)

Differentiating (2.12) with respect to eme_{m} for mI1m\in I_{1} and m1m\neq 1, we obtain

2αdnαem(μα)=0.\displaystyle\sum_{2\leq\alpha\leq d}n_{\alpha}e_{m}(\mu_{\alpha})=0. (2.14)

Differentiating (2.13) with respect to e1e_{1}, from (2.8), (2.12), μαΓααm=em(μα)\mu_{\alpha}\Gamma_{\alpha\alpha}^{m}=e_{m}(\mu_{\alpha}) and (2.14) we have

0\displaystyle 0 =2αdnα(e1(μα)Pα+μαe1(Pα))\displaystyle=\sum_{2\leq\alpha\leq d}n_{\alpha}\Big{(}e_{1}(\mu_{\alpha})P_{\alpha}+\mu_{\alpha}e_{1}(P_{\alpha})\Big{)}
=2αdnα(μαPα2+μα(Pα2+c+mI1ΓααmΓ11m))\displaystyle=\sum_{2\leq\alpha\leq d}n_{\alpha}\Big{(}\mu_{\alpha}P_{\alpha}^{2}+\mu_{\alpha}(P_{\alpha}^{2}+c+\sum_{m\in I_{1}}\Gamma_{\alpha\alpha}^{m}\Gamma_{11}^{m})\Big{)}
=22αdnαμαPα2+cnH+2αdmI1nαμαΓααmΓ11m\displaystyle=2\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}P_{\alpha}^{2}+cnH+\sum_{2\leq\alpha\leq d}\sum_{m\in I_{1}}n_{\alpha}\mu_{\alpha}\Gamma_{\alpha\alpha}^{m}\Gamma_{11}^{m}
=22αdnαμαPα2+cnH+mI1(2αdnαem(μα))Γ11m\displaystyle=2\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}P_{\alpha}^{2}+cnH+\sum_{m\in I_{1}}\Big{(}\sum_{2\leq\alpha\leq d}n_{\alpha}e_{m}(\mu_{\alpha})\Big{)}\Gamma_{11}^{m}
=22αdnαμαPα2+cnH,\displaystyle=2\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}P_{\alpha}^{2}+cnH,

which is equivalent to

2αdnαμαPα2=12cnH.\displaystyle\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}P_{\alpha}^{2}=-\frac{1}{2}cnH. (2.15)

Equations (2.13) and (2.15) imply that (2.11) holds for q=1,2q=1,2. Next we will prove that it holds for general qq by induction.

Differentiating (2.11) with respect to e1e_{1} yields

0\displaystyle 0 =2αdnα(e1(μα)Pαq+qμαPαq1e1(Pα))\displaystyle=\sum_{2\leq\alpha\leq d}n_{\alpha}\Big{(}e_{1}(\mu_{\alpha})P_{\alpha}^{q}+q\mu_{\alpha}P_{\alpha}^{q-1}e_{1}(P_{\alpha})\Big{)}
=2αdnα(μαPαq+1+qμαPαq1(Pα2+c+mI1ΓααmΓ11m))\displaystyle=\sum_{2\leq\alpha\leq d}n_{\alpha}\Big{(}\mu_{\alpha}P_{\alpha}^{q+1}+q\mu_{\alpha}P_{\alpha}^{q-1}(P_{\alpha}^{2}+c+\sum_{m\in I_{1}}\Gamma_{\alpha\alpha}^{m}\Gamma_{11}^{m})\Big{)}
=2αd((1+q)nαμαPαq+1+cqnαμαPαq1)+q2αdnαPαq1mI1em(μα)Γ11m\displaystyle=\sum_{2\leq\alpha\leq d}\Big{(}(1+q)n_{\alpha}\mu_{\alpha}P_{\alpha}^{q+1}+cqn_{\alpha}\mu_{\alpha}P_{\alpha}^{q-1}\Big{)}+q\sum_{2\leq\alpha\leq d}n_{\alpha}P_{\alpha}^{q-1}\sum_{m\in I_{1}}e_{m}(\mu_{\alpha})\Gamma_{11}^{m}
=(1+q)f(q+1)+cqf(q1)+qmI1{2αdnαem(μα)Pαq1}Γ11m.\displaystyle=(1+q)f(q+1)+cqf(q-1)+q\sum_{m\in I_{1}}\Big{\{}\sum_{2\leq\alpha\leq d}n_{\alpha}e_{m}(\mu_{\alpha})P_{\alpha}^{q-1}\Big{\}}\Gamma_{11}^{m}. (2.16)

On the other hand, since (2.11) holds for f(q1)f(q-1), we differentiate f(q1)=2αdnαμαPαq1f(q-1)=\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}P^{q-1}_{\alpha} with respect to eje_{j} for jI1j\in I_{1} and j1j\neq 1. It follows from (2.9) that

0\displaystyle 0 =2αd(nαej(μα)Pαq1+(q1)nαμαPαq2ej(Pα))\displaystyle=\sum_{2\leq\alpha\leq d}\Big{(}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{q-1}+(q-1)n_{\alpha}\mu_{\alpha}P_{\alpha}^{q-2}e_{j}(P_{\alpha})\Big{)}
=2αd(nαej(μα)Pαq1+(q1)nαμαPαq2(ΓααjPαmI1ΓααmΓjm1))\displaystyle=\sum_{2\leq\alpha\leq d}\Big{(}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{q-1}+(q-1)n_{\alpha}\mu_{\alpha}P_{\alpha}^{q-2}\big{(}\Gamma_{\alpha\alpha}^{j}P_{\alpha}-\sum_{m\in I_{1}}\Gamma_{\alpha\alpha}^{m}\Gamma_{jm}^{1})\Big{)}
=2αd(nαej(μα)Pαq1+(q1)nαej(μα)Pαq1(q1)mI1nαPαq2em(μα)Γjm1)\displaystyle=\sum_{2\leq\alpha\leq d}\Big{(}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{q-1}+(q-1)n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{q-1}-(q-1)\sum_{m\in I_{1}}n_{\alpha}P_{\alpha}^{q-2}e_{m}(\mu_{\alpha})\Gamma_{jm}^{1}\Big{)}
=2αd(qnαej(μα)Pαq1(q1)mI1nαPαq2em(μα)Γjm1).\displaystyle=\sum_{2\leq\alpha\leq d}\Big{(}qn_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{q-1}-(q-1)\sum_{m\in I_{1}}n_{\alpha}P_{\alpha}^{q-2}e_{m}(\mu_{\alpha})\Gamma_{jm}^{1}\Big{)}.

Hence the following relation holds for any jI1j\in I_{1} and j1j\neq 1

q2αdnαej(μα)Pαq1=mI1{(q1)2αdnαem(μα)Pαq2}Γjm1.\displaystyle q\sum_{2\leq\alpha\leq d}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{q-1}=\sum_{m\in I_{1}}\Big{\{}(q-1)\sum_{2\leq\alpha\leq d}n_{\alpha}e_{m}(\mu_{\alpha})P_{\alpha}^{q-2}\Big{\}}\Gamma_{jm}^{1}. (2.17)

Since (2.17) holds for any qq, letting q=2q=2, (2.17) reduces to

22αdnαej(μα)Pα=mI1{2αdnαem(μα)}Γjm1,\displaystyle 2\sum_{2\leq\alpha\leq d}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}=\sum_{m\in I_{1}}\Big{\{}\sum_{2\leq\alpha\leq d}n_{\alpha}e_{m}(\mu_{\alpha})\Big{\}}\Gamma_{jm}^{1},

which together with (2.14) yields

22αdnαej(μα)Pα=0.\displaystyle 2\sum_{2\leq\alpha\leq d}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}=0. (2.18)

Letting q=3q=3 and using (2.18), (2.17) reduces to

32αdnαej(μα)Pα2=mI1{22αdnαem(μα)Pα}Γjm1=0.\displaystyle 3\sum_{2\leq\alpha\leq d}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{2}=\sum_{m\in I_{1}}\Big{\{}2\sum_{2\leq\alpha\leq d}n_{\alpha}e_{m}(\mu_{\alpha})P_{\alpha}\Big{\}}\Gamma_{jm}^{1}=0.

Similarly, we can gradually show

q2αdnαej(μα)Pαq1=0.\displaystyle q\sum_{2\leq\alpha\leq d}n_{\alpha}e_{j}(\mu_{\alpha})P_{\alpha}^{q-1}=0. (2.19)

Hence, combing (2) with (2.19) gives

(q+1)f(q+1)+cqf(q1)=0.\displaystyle(q+1)f(q+1)+cqf(q-1)=0. (2.20)

When qq is even, both of q1q-1 and q+1q+1 are odd. From (2.20), f(q1)=0f(q-1)=0 can yield f(q+1)=0f(q+1)=0 as well.

When qq is odd, both of q1q-1 and q+1q+1 are even. We conclude from (2.20) that

f(q+1)\displaystyle f(q+1) =cqq+1f(q1)\displaystyle=-\frac{cq}{q+1}f(q-1)
=cqq+1×(q2)!!(q1)!!(c)(q1)/2nH\displaystyle=-\frac{cq}{q+1}\times\frac{(q-2)!!}{(q-1)!!}(-c)^{(q-1)/2}nH
=q!!(q+1)!!(c)(q+1)/2nH,\displaystyle=\frac{q!!}{(q+1)!!}(-c)^{(q+1)/2}nH, (2.21)

which completes the proof of Lemma 2.4. ∎

Remark 2.5.

While assuming an additional condition that the multiplicity of zero principal curvature is at most one, special cases of Lemmas 2.3 and 2.4 were derived in [7].

Lemma 2.6.

The equation ΔS+S2ncSn2cH2=0{\rm\Delta}\,S+S^{2}-ncS-n^{2}cH^{2}=0 is equivalent to

6α=2dnαμα2Pα2+2(α=2dnαμα2Pα)(α=2dnαPα+mI1Γmm1)+(α=2dnαμα2)2\displaystyle-6\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}^{2}+2\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}\Big{)}\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}P_{\alpha}+\sum_{m\in I_{1}}\Gamma_{mm}^{1}\Big{)}+\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}^{2}
(n+2)c(α=2dnαμα2)c(α=2dnαμα)2=0.\displaystyle-(n+2)c\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}-c\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}\Big{)}^{2}=0.
Proof.

Since ei(S)=0e_{i}(S)=0 for 2in2\leq i\leq n, it follows from Lemma 2.2 that

ΔS\displaystyle\Delta S =i=1n(eieiSeieiS)\displaystyle=-\sum_{i=1}^{n}(\nabla_{e_{i}}\nabla_{e_{i}}S-\nabla_{\nabla_{e_{i}e_{i}}}S)
=e1e1S+e1(S)i=2nΓii1\displaystyle=-e_{1}e_{1}S+e_{1}(S)\sum_{i=2}^{n}\Gamma_{ii}^{1}
=e1e1S+e1(S)(α=2dnαPα+mI1Γmm1).\displaystyle=-e_{1}e_{1}S+e_{1}(S)\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}P_{\alpha}+\sum_{m\in I_{1}}\Gamma_{mm}^{1}\Big{)}. (2.22)

Noting S=α=2dnαμα2S=\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}, it follows from (2.8) that

e1(S)\displaystyle e_{1}(S) =2α=2dnαμαe1(μα)=2α=2dnαμα2Pα,\displaystyle=2\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}e_{1}(\mu_{\alpha})=2\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}, (2.23)
e1e1(S)\displaystyle e_{1}e_{1}(S) =4α=2dnαμα2Pα2+2α=2dnαμα2Pα2+2cα=2dnαμα2+2α=2dmI1nαμα2ΓααmΓ11m\displaystyle=4\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}^{2}+2\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}^{2}+2c\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}+2\sum_{\alpha=2}^{d}\sum_{m\in I_{1}}n_{\alpha}\mu_{\alpha}^{2}\Gamma_{\alpha\alpha}^{m}\Gamma_{11}^{m}
=6α=2dnαμα2Pα2+2cα=2dnαμα2+2mI1{α=2dnαμα2Γααm}Γ11m.\displaystyle=6\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}^{2}+2c\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}+2\sum_{m\in I_{1}}\Big{\{}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Gamma_{\alpha\alpha}^{m}\Big{\}}\Gamma_{11}^{m}. (2.24)

Differentiating S=α=2dnαμα2S=\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2} with respect to eme_{m} for mI1m\in I_{1} and m1m\neq 1, we get

0=em(S)=2α=2dnαμαem(μα)=2α=2dnαμα2Γααm.\displaystyle 0=e_{m}(S)=2\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}e_{m}(\mu_{\alpha})=2\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Gamma_{\alpha\alpha}^{m}. (2.25)

Combining (2.24) with (2.25) gives

e1e1(S)=6α=2dnαμα2Pα2+2cα=2dnαμα2.\displaystyle e_{1}e_{1}(S)=6\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}^{2}+2c\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}. (2.26)

Substituting (2.23) and (2.26) into (2), we have

ΔS=\displaystyle\Delta S= 6α=2dnαμα2Pα2+2(α=2dnαμα2Pα)(α=2dnαPα+mI1Γmm1)\displaystyle-6\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}^{2}+2\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}\Big{)}\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}P_{\alpha}+\sum_{m\in I_{1}}\Gamma_{mm}^{1}\Big{)}
2cα=2dnαμα2.\displaystyle-2c\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}.

Hence, the proof has been completed. ∎

3. Proof of Theorems 1.5 and 1.6

The proof of Theorems 1.5 and 1.6: We will prove Theorems 1.5 and 1.6 by deriving a contradiction from the assumption that 𝒩={pMn:S(p)0}\mathcal{N}=\{p\in M^{n}:\nabla S(p)\neq 0\}\neq\emptyset.

Taking q=1,3,5,,2d3q=1,3,5,\cdots,2d-3 in Lemma 2.4, we have

{n2μ2P2+n3μ3P3++ndμdPd=0,n2μ2P23+n3μ3P33++ndμdPd3=0,n2μ2P22d3+n3μ3P32d3++ndμdPd2d3=0,\displaystyle\begin{split}\left\{\begin{array}[]{ll}n_{2}\mu_{2}P_{2}+n_{3}\mu_{3}P_{3}+\cdots+n_{d}\mu_{d}P_{d}=0,\\ n_{2}\mu_{2}P_{2}^{3}+n_{3}\mu_{3}P_{3}^{3}+\cdots+n_{d}\mu_{d}P_{d}^{3}=0,\\ \quad\,\vdots\\ n_{2}\mu_{2}P_{2}^{2d-3}+n_{3}\mu_{3}P_{3}^{2d-3}+\cdots+n_{d}\mu_{d}P_{d}^{2d-3}=0,\end{array}\right.\end{split} (3.1)

which is a (d1)(d-1)-th order equation system with a non-zero solution. Hence on 𝒩\mathcal{N} we have

|P2P3PdP23P33Pd3P22d3P32d3Pd2d3|=P2Pd2αβd(Pα2Pβ2)=0.\displaystyle\begin{split}\left|\begin{array}[]{llll}P_{2}&P_{3}&\cdots&P_{d}\\ P_{2}^{3}&P_{3}^{3}&\cdots&P_{d}^{3}\\ \,\,\vdots&\,\,\vdots&\cdots&\,\,\vdots\\ P_{2}^{2d-3}&P_{3}^{2d-3}&\cdots&P_{d}^{2d-3}\end{array}\right|=P_{2}\cdots P_{d}\prod_{2\leq\alpha\leq\beta\leq d}(P_{\alpha}^{2}-P_{\beta}^{2})=0.\end{split} (3.2)

Taking q=2,4,6,,2d2q=2,4,6,\cdots,2d-2 in Lemma 2.4, we obtain

{n2μ2P22+n3μ3P32++ndμdPd2=12ncH,n2μ2P24+n3μ3P34++ndμdPd4=38nc2H,n2μ2P22d2+n3μ3P32d2++ndμdPd2d2=(2d3)!!(2d2)!!(c)d1nH.\displaystyle\begin{split}\left\{\begin{array}[]{ll}n_{2}\mu_{2}P_{2}^{2}+n_{3}\mu_{3}P_{3}^{2}+\cdots+n_{d}\mu_{d}P_{d}^{2}=-\frac{1}{2}ncH,\\ n_{2}\mu_{2}P_{2}^{4}+n_{3}\mu_{3}P_{3}^{4}+\cdots+n_{d}\mu_{d}P_{d}^{4}=\frac{3}{8}nc^{2}H,\\ \quad\,\vdots\\ n_{2}\mu_{2}P_{2}^{2d-2}+n_{3}\mu_{3}P_{3}^{2d-2}+\cdots+n_{d}\mu_{d}P_{d}^{2d-2}=\frac{(2d-3)!!}{(2d-2)!!}(-c)^{d-1}nH.\end{array}\right.\end{split} (3.3)

Next we consider all possible cases.

Case 1. P2P3Pd0P_{2}P_{3}\cdots P_{d}\neq 0 at some p𝒩p\in\mathcal{N}. Then from (3.2) we have that 2αβd(Pα2Pβ2)=0\prod_{2\leq\alpha\leq\beta\leq d}(P_{\alpha}^{2}-P_{\beta}^{2})=0 at pp. Without loss of generality, we assume P22P32=0P_{2}^{2}-P_{3}^{2}=0, i.e., P2=±P3P_{2}=\pm P_{3}. Now the former d2d-2 equations in the system of equations (3.1) determine a new system of equations as follows:

{(n2μ2±n3μ3)P3++ndμdPd=0,(n2μ2±n3μ3)P33++ndμdPd3=0,(n2μ2±n3μ3)P32d5++ndμdPd2d5=0,\displaystyle\begin{split}\left\{\begin{array}[]{ll}(n_{2}\mu_{2}\pm n_{3}\mu_{3})P_{3}+\cdots+n_{d}\mu_{d}P_{d}=0,\\ (n_{2}\mu_{2}\pm n_{3}\mu_{3})P_{3}^{3}+\cdots+n_{d}\mu_{d}P_{d}^{3}=0,\\ \quad\quad\quad\vdots\\ (n_{2}\mu_{2}\pm n_{3}\mu_{3})P_{3}^{2d-5}+\cdots+n_{d}\mu_{d}P_{d}^{2d-5}=0,\end{array}\right.\end{split} (3.4)

which has a non-zero solution. Then we have

|P3P4PdP33P43Pd3P32d5P42d5Pd2d5|=P3Pd3αβd(Pα2Pβ2)=0.\displaystyle\begin{split}\left|\begin{array}[]{llll}P_{3}&P_{4}&\cdots&P_{d}\\ P_{3}^{3}&P_{4}^{3}&\cdots&P_{d}^{3}\\ \,\,\vdots&\,\,\vdots&\cdots&\,\,\vdots\\ P_{3}^{2d-5}&P_{4}^{2d-5}&\cdots&P_{d}^{2d-5}\end{array}\right|=P_{3}\cdots P_{d}\prod_{3\leq\alpha\leq\beta\leq d}(P_{\alpha}^{2}-P_{\beta}^{2})=0.\end{split}

Since P3P4Pd0P_{3}P_{4}\cdots P_{d}\neq 0, we have that 3αβd(Pα2Pβ2)=0\prod_{3\leq\alpha\leq\beta\leq d}(P_{\alpha}^{2}-P_{\beta}^{2})=0. Without loss of generality, we assume that P32=P42P_{3}^{2}=P_{4}^{2}. Proceeding in this way, we obtain that P22=P32==Pd2:=P2P_{2}^{2}=P_{3}^{2}=\cdots=P_{d}^{2}:=P^{2} at pp. Now (3.3) becomes

{12ncH=n2μ2P22+n3μ3P32++ndμdPd2=nHP2,38nc2H=n2μ2P24+n3μ3P34++ndμdPd4=nHP4,(2d3)!!(2d2)!!(c)d1nH=n2μ2P22d2+n3μ3P32d2++ndμdPd2d2=nHP2d2.\displaystyle\begin{split}\left\{\begin{array}[]{ll}-\frac{1}{2}ncH=n_{2}\mu_{2}P_{2}^{2}+n_{3}\mu_{3}P_{3}^{2}+\cdots+n_{d}\mu_{d}P_{d}^{2}=nHP^{2},\\ \frac{3}{8}nc^{2}H=n_{2}\mu_{2}P_{2}^{4}+n_{3}\mu_{3}P_{3}^{4}+\cdots+n_{d}\mu_{d}P_{d}^{4}=nHP^{4},\\ \quad\,\,\vdots\\ \frac{(2d-3)!!}{(2d-2)!!}(-c)^{d-1}nH=n_{2}\mu_{2}P_{2}^{2d-2}+n_{3}\mu_{3}P_{3}^{2d-2}+\cdots+n_{d}\mu_{d}P_{d}^{2d-2}=nHP^{2d-2}.\end{array}\right.\end{split}

Since nH0nH\neq 0, the first two equations of the above system imply that

P2=12candP4=38c2,P^{2}=-\frac{1}{2}c\quad{\rm and}\quad P^{4}=\frac{3}{8}c^{2},

and hence c=P=0c=P=0. It is a contradiction, so this case is ruled out.

Case 2. Pα=0P_{\alpha}=0 for all α=2,,d\alpha=2,\cdots,d at some p𝒩p\in\mathcal{N}. In this case, we have

e1S=e12αdnαμα2=22αdnαμαe1(μα)=22αdnαμα2Pα=0.e_{1}S=e_{1}\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}^{2}=2\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}e_{1}(\mu_{\alpha})=2\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}=0.

This contradicts S0\nabla S\neq 0 at pp and this case is also ruled out.

Case 3. For any given point p𝒩p\in\mathcal{N}, some terms of PαP_{\alpha} are zero and the others are not zero. In this case, without loss of generality, assume Pα=0P_{\alpha}=0 for α=2,,r\alpha=2,\cdots,r and Pα0P_{\alpha}\neq 0 for α=r+1,,d\alpha=r+1,\cdots,d. Then the first drd-r equations in (3.1) form a new system of equations

{nr+1μr+1Pr+1+nr+2μr+2Pr+2++ndμdPd=0,nr+1μr+1Pr+13+nr+2μr+2Pr+23++ndμdPd3=0,nr+1μr+1Pr+12(dr)1+nr+2μr+2Pr+22(dr)1++ndμdPd2(dr)1=0,\displaystyle\begin{split}\left\{\begin{array}[]{ll}n_{r+1}\mu_{r+1}P_{r+1}+n_{r+2}\mu_{r+2}P_{r+2}+\cdots+n_{d}\mu_{d}P_{d}=0,\\ n_{r+1}\mu_{r+1}P_{r+1}^{3}+n_{r+2}\mu_{r+2}P_{r+2}^{3}+\cdots+n_{d}\mu_{d}P_{d}^{3}=0,\\ \quad\quad\quad\vdots\\ n_{r+1}\mu_{r+1}P_{r+1}^{2(d-r)-1}+n_{r+2}\mu_{r+2}P_{r+2}^{2(d-r)-1}+\cdots+n_{d}\mu_{d}P_{d}^{2(d-r)-1}=0,\end{array}\right.\end{split} (3.5)

which is a (dr)(d-r)-th order equation system with non-zero solutions. So the coefficient determinant is zero, that is r+1αβd(Pα2Pβ2)=0\prod_{r+1\leq\alpha\leq\beta\leq d}(P_{\alpha}^{2}-P_{\beta}^{2})=0. Without loss of generality, we assume that Pr+12=Pr+22P_{r+1}^{2}=P_{r+2}^{2}. Proceeding in this way, we can show that Pr+12==Pd20P_{r+1}^{2}=\cdots=P_{d}^{2}\neq 0. Denote by P2:=Pr+12==Pd2P^{2}:=P_{r+1}^{2}=\cdots=P_{d}^{2}. Then (3.3) becomes

{(nr+1μr+1++ndμd)P2=12ncH,(nr+1μr+1++ndμd)P4=38nc2H,(nr+1μr+1++ndμd)P6=516nc3H,(nr+1μr+1++ndμd)P2d2=(2d3)!!(2d2)!!(c)d1nH.\displaystyle\begin{split}\left\{\begin{array}[]{ll}(n_{r+1}\mu_{r+1}+\cdots+n_{d}\mu_{d})P^{2}=-\frac{1}{2}ncH,\\ (n_{r+1}\mu_{r+1}+\cdots+n_{d}\mu_{d})P^{4}=\frac{3}{8}nc^{2}H,\\ (n_{r+1}\mu_{r+1}+\cdots+n_{d}\mu_{d})P^{6}=-\frac{5}{16}nc^{3}H,\\ \quad\quad\quad\vdots\\ (n_{r+1}\mu_{r+1}+\cdots+n_{d}\mu_{d})P^{2d-2}=\frac{(2d-3)!!}{(2d-2)!!}(-c)^{d-1}nH.\end{array}\right.\end{split} (3.6)

The above system of equations means that nr+1μr+1++ndμd0n_{r+1}\mu_{r+1}+\cdots+n_{d}\mu_{d}\neq 0 since c0c\neq 0 and H0H\neq 0. The first two equations of (3.6) force that P2=34cP^{2}=-\frac{3}{4}c, and the second and the third equation of (3.6) force that P2=56cP^{2}=-\frac{5}{6}c. Hence we have P=c=0P=c=0, a contradiction.

In conclusion, we have that 𝒩={pM:S(p)0}\mathcal{N}=\{p\in M:\nabla S(p)\neq 0\} is empty and hence SS has to be a constant. From (2.5), we conclude that the scalar curvature RR of MnM^{n} is constant as well. But for c<0c<0, the first equation of (2.7) means that S2ncSn2cH2=0S^{2}-ncS-n^{2}cH^{2}=0, a contradiction. This completes the proof of Theorems 1.5 and 1.6. \hfill\square

4. Proofs of Theorems 1.8 and 1.9

In this section, we mainly concern CMC triharmonic hypersurfaces in the Euclidean space n+1\mathbb{R}^{n+1}. For any dimension nn, we need another assumption that the multiplicity of the zero principal curvature is at most one as discussed in [7] for four distinct principal curvatures.

The proof of Theorem 1.8: Assume that 𝒩={pMn:S(p)0}\mathcal{N}=\{p\in M^{n}:\nabla S(p)\neq 0\}\neq\emptyset. We will prove Theorems 1.8 by deriving a contradiction.

Taking q=1,2,3,,d1q=1,2,3,\cdots,d-1 in Lemma 2.4, we have

{n2μ2P2+n3μ3P3++ndμdPd=0,n2μ2P22+n3μ3P32++ndμdPd2=0,n2μ2P2d1+n3μ3P3d1++ndμdPdd1=0,\displaystyle\begin{split}\left\{\begin{array}[]{ll}n_{2}\mu_{2}P_{2}+n_{3}\mu_{3}P_{3}+\cdots+n_{d}\mu_{d}P_{d}=0,\\ n_{2}\mu_{2}P_{2}^{2}+n_{3}\mu_{3}P_{3}^{2}+\cdots+n_{d}\mu_{d}P_{d}^{2}=0,\\ \quad\,\vdots\\ n_{2}\mu_{2}P_{2}^{d-1}+n_{3}\mu_{3}P_{3}^{d-1}+\cdots+n_{d}\mu_{d}P_{d}^{d-1}=0,\end{array}\right.\end{split} (4.1)

which is a (d1)(d-1)-th order equation system with a non-zero solution. Hence on 𝒩\mathcal{N} we have

|P2P3PdP22P32Pd2P2d1P3d1Pdd1|=P2Pd2αβd(PαPβ)=0.\displaystyle\begin{split}\left|\begin{array}[]{llll}P_{2}&P_{3}&\cdots&P_{d}\\ P_{2}^{2}&P_{3}^{2}&\cdots&P_{d}^{2}\\ \,\,\vdots&\,\,\vdots&\cdots&\,\,\vdots\\ P_{2}^{d-1}&P_{3}^{d-1}&\cdots&P_{d}^{d-1}\end{array}\right|=P_{2}\cdots P_{d}\prod_{2\leq\alpha\leq\beta\leq d}(P_{\alpha}-P_{\beta})=0.\end{split} (4.2)

Next we consider three possible cases.

Case 1. P2P3Pd0P_{2}P_{3}\cdots P_{d}\neq 0 at some p𝒩p\in\mathcal{N}. Then from (4.2) we have that 2αβd(PαPβ)=0\prod_{2\leq\alpha\leq\beta\leq d}(P_{\alpha}-P_{\beta})=0 at pp. Without loss of generality, we assume P2P3=0P_{2}-P_{3}=0. Now the former d2d-2 equations in the system of equations (4.1) determine a new system of equations as follows:

{(n2μ2+n3μ3)P3++ndμdPd=0,(n2μ2+n3μ3)P32++ndμdPd2=0,(n2μ2+n3μ3)P3d2++ndμdPdd2=0,\displaystyle\begin{split}\left\{\begin{array}[]{ll}(n_{2}\mu_{2}+n_{3}\mu_{3})P_{3}+\cdots+n_{d}\mu_{d}P_{d}=0,\\ (n_{2}\mu_{2}+n_{3}\mu_{3})P_{3}^{2}+\cdots+n_{d}\mu_{d}P_{d}^{2}=0,\\ \quad\quad\quad\vdots\\ (n_{2}\mu_{2}+n_{3}\mu_{3})P_{3}^{d-2}+\cdots+n_{d}\mu_{d}P_{d}^{d-2}=0,\end{array}\right.\end{split} (4.3)

which has a non-zero solution. Then we have

|P3P4PdP32P42Pd2P3d2P4d2Pdd2|=P3Pd3αβd(PαPβ)=0.\displaystyle\begin{split}\left|\begin{array}[]{llll}P_{3}&P_{4}&\cdots&P_{d}\\ P_{3}^{2}&P_{4}^{2}&\cdots&P_{d}^{2}\\ \,\,\vdots&\,\,\vdots&\cdots&\,\,\vdots\\ P_{3}^{d-2}&P_{4}^{d-2}&\cdots&P_{d}^{d-2}\end{array}\right|=P_{3}\cdots P_{d}\prod_{3\leq\alpha\leq\beta\leq d}(P_{\alpha}-P_{\beta})=0.\end{split}

Since P3P4Pd0P_{3}P_{4}\cdots P_{d}\neq 0, we have that 3αβd(PαPβ)=0\prod_{3\leq\alpha\leq\beta\leq d}(P_{\alpha}-P_{\beta})=0. Without loss of generality, we assume that P3=P4P_{3}=P_{4}. Similarly, we obtain that P3=P4==Pd:=PP_{3}=P_{4}=\cdots=P_{d}:=P at pp. Since nH0nH\neq 0, the first equation of the above system (4.3) implies that

(2αdnαμα)P=nHP=0\Big{(}\sum_{2\leq\alpha\leq d}{n_{\alpha}\mu_{\alpha}}\Big{)}P=nHP=0

and hence P=0P=0. It is a contradiction.

Case 2. Pα=0P_{\alpha}=0 for all α=2,,d\alpha=2,\cdots,d at some p𝒩p\in\mathcal{N}. Then

e1S=e12αdnαμα2=22αdnαμαe1(μα)=22αdnαμα2Pα=0,e_{1}S=e_{1}\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}^{2}=2\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}e_{1}(\mu_{\alpha})=2\sum_{2\leq\alpha\leq d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}=0,

which contradicts S0\nabla S\neq 0 at pp.

Case 3. For any given point p𝒩p\in\mathcal{N}, some terms of PαP_{\alpha} are zero and the others are not zero. In this case, without loss of generality, assume Pα=0P_{\alpha}=0 for α=2,,r\alpha=2,\cdots,r and Pα0P_{\alpha}\neq 0 for α=r+1,,d\alpha=r+1,\cdots,d. Then the first drd-r equations in (4.1) form a new system of equations

{nr+1μr+1Pr+1+nr+2μr+2Pr+2++ndμdPd=0,nr+1μr+1Pr+12+nr+2μr+2Pr+22++ndμdPd2=0,nr+1μr+1Pr+1dr+nr+2μr+2Pr+2dr++ndμdPddr=0,\displaystyle\begin{split}\left\{\begin{array}[]{ll}n_{r+1}\mu_{r+1}P_{r+1}+n_{r+2}\mu_{r+2}P_{r+2}+\cdots+n_{d}\mu_{d}P_{d}=0,\\ n_{r+1}\mu_{r+1}P_{r+1}^{2}+n_{r+2}\mu_{r+2}P_{r+2}^{2}+\cdots+n_{d}\mu_{d}P_{d}^{2}=0,\\ \quad\quad\quad\vdots\\ n_{r+1}\mu_{r+1}P_{r+1}^{d-r}+n_{r+2}\mu_{r+2}P_{r+2}^{d-r}+\cdots+n_{d}\mu_{d}P_{d}^{d-r}=0,\end{array}\right.\end{split} (4.4)

which is a (dr)(d-r)-th order equation system with non-zero solutions. Thus the coefficient determinant is zero, that is r+1αβd(PαPβ)=0\prod_{r+1\leq\alpha\leq\beta\leq d}(P_{\alpha}-P_{\beta})=0. Without loss of generality, we assume that Pr+1=Pr+2P_{r+1}=P_{r+2}. Similar discussion as the above yields Pr+1==Pd0P_{r+1}=\cdots=P_{d}\neq 0. Denote by P:=Pr+1==PdP:=P_{r+1}=\cdots=P_{d}.

It follows from (4.4) that nr+1μr+1++ndμd=0n_{r+1}\mu_{r+1}+\cdots+n_{d}\mu_{d}=0, because of P0P\neq 0. In this case, we have n2μ2+nrμr=nHn_{2}\mu_{2}+\cdots n_{r}\mu_{r}=nH. Therefore we deduce from Lemma 2.6 that

6α=2dnαμα2Pα2+2(α=2dnαμα2Pα)(α=2dnαPα+mI1Γmm1)+(α=2dnαμα2)2=0,\displaystyle-6\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}^{2}+2\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}P_{\alpha}\Big{)}\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}P_{\alpha}+\sum_{m\in I_{1}}\Gamma_{mm}^{1}\Big{)}+\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}^{2}=0,

that is

2(α=r+1dnα3)P2α=r+1dnαμα2+2(α=r+1dnαμα2)PmI1Γmm1\displaystyle 2\Big{(}\sum_{\alpha=r+1}^{d}n_{\alpha}-3\Big{)}P^{2}\sum_{\alpha=r+1}^{d}n_{\alpha}\mu_{\alpha}^{2}+2\Big{(}\sum_{\alpha=r+1}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}P\sum_{m\in I_{1}}\Gamma_{mm}^{1}
+(α=2dnαμα2)2=0,\displaystyle+\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}^{2}=0, (4.5)

where we have used Pα=0P_{\alpha}=0 for α=2,,r\alpha=2,\ldots,r.

Because we have assumed that the multiplicity of the zero principal curvature is one, i.e. I1={1}I_{1}=\{1\}, (4) becomes

2(α=r+1dnα3)P2α=r+1dnαμα2+(α=2dnαμα2)2=0.\displaystyle 2\Big{(}\sum_{\alpha=r+1}^{d}n_{\alpha}-3\Big{)}P^{2}\sum_{\alpha=r+1}^{d}n_{\alpha}\mu_{\alpha}^{2}+\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}^{2}=0. (4.6)

Differentiating with respect to e1e_{1} on both sides of equation (4.6), it follows from (2.8) that

8(α=r+1dnα3)P3α=r+1dnαμα2+4(α=2dnαμα2)(α=r+1dnαμα2)P=0.\displaystyle 8\Big{(}\sum_{\alpha=r+1}^{d}n_{\alpha}-3\Big{)}P^{3}\sum_{\alpha=r+1}^{d}n_{\alpha}\mu_{\alpha}^{2}+4\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}\Big{(}\sum_{\alpha=r+1}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}P=0. (4.7)

Dividing (4.7) by 4P4P, and then subtracting (4.6), we have

(α=2dnαμα2)(α=2rnαμα2)=0.\displaystyle\Big{(}\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\Big{)}\Big{(}\sum_{\alpha=2}^{r}n_{\alpha}\mu_{\alpha}^{2}\Big{)}=0. (4.8)

Since S=α=2dnαμα20S=\sum_{\alpha=2}^{d}n_{\alpha}\mu_{\alpha}^{2}\neq 0, it follows from (4.8) that α=2rnαμα2=0\sum_{\alpha=2}^{r}n_{\alpha}\mu_{\alpha}^{2}=0, and hence μα=0\mu_{\alpha}=0 for α=2,,r\alpha=2,\cdots,r. This is a contradiction since α=2rnαμα=nH0\sum_{\alpha=2}^{r}n_{\alpha}\mu_{\alpha}=nH\neq 0. \hfill\square

The proof of Theorem 1.9: Let us restrict the case of a CMC triharmonic hypersurface M5M^{5} in 6\mathbb{R}^{6}. We will still work on 𝒩={pM5:S(p)0}\mathcal{N}=\{p\in M^{5}:\nabla S(p)\neq 0\}\neq\emptyset and prove Theorem 1.9 with a contradiction.

When the multiplicity of zero principal curvature is one, we can derive a contradiction from Theorem 1.8.

By Theorem 1.3, we only need to deal with the case that the multiplicity of zero principal curvature is two, i.e. I1={1,2}I_{1}=\{1,2\}. In this case, the principle curvatures on M5M^{5} are respectively 0,0,λ3,λ4,λ50,0,\lambda_{3},\lambda_{4},\lambda_{5} with λ3+λ4+λ5=5H\lambda_{3}+\lambda_{4}+\lambda_{5}=5H.

According to the proof of Theorem 1.8, we can deduce a contradiction when Pα0P_{\alpha}\neq 0 or Pα=0P_{\alpha}=0 for all PαP_{\alpha}. Let us consider the remaining case that P3=0P_{3}=0, P4=P50P_{4}=P_{5}\neq 0. It follows from (4.4) that λ4+λ5=0\lambda_{4}+\lambda_{5}=0 and λ3=5H\lambda_{3}=5H. For simplicity, we denote λ4=λ5=:μ\lambda_{4}=-\lambda_{5}=:\mu and P4=P5=:PP_{4}=P_{5}=:P. Then the squared norm of the second fundamental form SS is given by S=2μ2+25H2S=2\mu^{2}+25H^{2}. Since e2(S)=0e_{2}(S)=0, we have e2(μ)=0e_{2}(\mu)=0 and hence Γ442=Γ552=0\Gamma_{44}^{2}=\Gamma_{55}^{2}=0. In this case, we deduce from (2.9) that e2(P)=Γ442PΓ442Γ221=0e_{2}(P)=\Gamma_{44}^{2}P-\Gamma_{44}^{2}\Gamma_{22}^{1}=0. Moreover, e1(S)=4μe1(μ)=4μ2Pe_{1}(S)=4\mu e_{1}(\mu)=4\mu^{2}P and hence e2e1(S)=0e_{2}e_{1}(S)=0. Since e1(S)0e_{1}(S)\neq 0, ei(S)=0e_{i}(S)=0 for i2i\geq 2 and Γ211=0\Gamma_{21}^{1}=0, we find

0=e1e2(S)e2e1(S)=[e1,e2](S)=(e1e2e2e1)S=Γ121e1(S),\displaystyle 0=e_{1}e_{2}(S)-e_{2}e_{1}(S)=[e_{1},e_{2}](S)=(\nabla_{e_{1}}e_{2}-\nabla_{e_{2}}e_{1})S=\Gamma_{12}^{1}e_{1}(S),

which means that Γ121=Γ112=0\Gamma_{12}^{1}=-\Gamma_{11}^{2}=0. Then (2.8) turns into

e1(P)=P2.\displaystyle e_{1}(P)=P^{2}. (4.9)

One the other hand, taking into account the Gauss equation and (2.2), from Lemma 2.2 we obtain

0=R1212=\displaystyle 0=R_{1212}= e1(Γ221)e2(Γ121)+m(Γ22mΓ1m1Γ12mΓ2m1(Γ12mΓ21m)Γm21)\displaystyle e_{1}(\Gamma_{22}^{1})-e_{2}(\Gamma_{12}^{1})+\sum_{m}\Big{(}\Gamma_{22}^{m}\Gamma_{1m}^{1}-\Gamma_{12}^{m}\Gamma_{2m}^{1}-(\Gamma_{12}^{m}-\Gamma_{21}^{m})\Gamma_{m2}^{1}\Big{)}
=\displaystyle= e1(Γ221)(Γ221)2.\displaystyle e_{1}(\Gamma_{22}^{1})-(\Gamma_{22}^{1})^{2}. (4.10)

Based on the above discussion, (4) becomes

4μ2P2+4μ2PΓ221+(2μ2+25H2)2=0,\displaystyle-4\mu^{2}P^{2}+4\mu^{2}P\Gamma_{22}^{1}+(2\mu^{2}+25H^{2})^{2}=0, (4.11)

and hence

Γ221=4μ2P2(2μ2+25H2)24μ2P.\displaystyle\Gamma_{22}^{1}=\frac{4\mu^{2}P^{2}-(2\mu^{2}+25H^{2})^{2}}{4\mu^{2}P}. (4.12)

Noting e1(μ)=μPe_{1}(\mu)=\mu P and differentiating (4.11) with respect to e1e_{1}, from (4.9) and (4.10) we have

4P2+3PΓ221+(Γ221)2+2(2μ2+25H2)=0.\displaystyle-4P^{2}+3P\Gamma_{22}^{1}+(\Gamma_{22}^{1})^{2}+2(2\mu^{2}+25H^{2})=0. (4.13)

Substituting (4.12) into (4.13) and eliminating the terms of Γ221\Gamma_{22}^{1} one gets

16μ4P2(2(2μ2+25H2)4P2)+12μ2P2(4μ2P2(2μ2+25H2)2)\displaystyle 16\mu^{4}P^{2}\Big{(}2(2\mu^{2}+25H^{2})-4P^{2}\Big{)}+12\mu^{2}P^{2}\Big{(}4\mu^{2}P^{2}-(2\mu^{2}+25H^{2})^{2}\Big{)}
+(4μ2P2(2μ2+25H2)2)2=0,\displaystyle+\Big{(}4\mu^{2}P^{2}-(2\mu^{2}+25H^{2})^{2}\Big{)}^{2}=0,

that is

32μ4P2(2μ2+25H2)20μ2P2(2μ2+25H2)2+(2μ2+25H2)4=0.\displaystyle 32\mu^{4}P^{2}(2\mu^{2}+25H^{2})-20\mu^{2}P^{2}(2\mu^{2}+25H^{2})^{2}+(2\mu^{2}+25H^{2})^{4}=0. (4.14)

Because 2μ2+25H202\mu^{2}+25H^{2}\neq 0, (4.14) becomes

32μ4P220μ2P2(2μ2+25H2)+(2μ2+25H2)3=0.\displaystyle 32\mu^{4}P^{2}-20\mu^{2}P^{2}(2\mu^{2}+25H^{2})+(2\mu^{2}+25H^{2})^{3}=0. (4.15)

Differentiating with respect to e1e_{1} on (4.15), from (4.9) and (4.10) we have

12μ2P2500H2P2+3(2μ2+25H2)2=0.\displaystyle-12\mu^{2}P^{2}-500H^{2}P^{2}+3(2\mu^{2}+25H^{2})^{2}=0. (4.16)

Differentiating with respect to e1e_{1} on (4.16) yields

12μ2P2250H2P2+6μ2(2μ2+25H2)=0.\displaystyle-12\mu^{2}P^{2}-250H^{2}P^{2}+6\mu^{2}(2\mu^{2}+25H^{2})=0. (4.17)

Eliminating the terms of P2P^{2} between (4.16) and (4.17) gives

2μ2(12μ2+500H2)=(2μ2+25H2)(12μ2+250H2),\displaystyle 2\mu^{2}(12\mu^{2}+500H^{2})=(2\mu^{2}+25H^{2})(12\mu^{2}+250H^{2}),

that is

4μ2H2125H4=0,\displaystyle 4\mu^{2}H^{2}-125H^{4}=0, (4.18)

which implies that μ\mu is constant and S=2μ2+25H2S=2\mu^{2}+25H^{2} is constant as well. This contradicts S(p)0\nabla S(p)\neq 0.

In summary, we conclude that 𝒩={pM:S(p)0}=\mathcal{N}=\{p\in M:\nabla S(p)\neq 0\}=\emptyset and according to (2.6) we know that MnM^{n} is minimal. This completes the proof of Theorem 1.9. \hfill\square

Acknowledgement: The authors are supported by the NSFC (No.11801246) and Liaoning Provincial Education Department Project (No.LJKMZ20221561)

References

  • [1] V. Branding, S. Montaldo, C. Oniciuc and A. Ratto, Higher order energy functionals, Adv. Math. 370 (2020), Article no. 107236.
  • [2] V. Branding, S. Montaldo, C. Oniciuc and A. Ratto, Polyharmonic hypersurfaces into pseudo-Riemannian space forms, Ann. Mat. Pura Appl. (2022). https://doi.org/10.1007/s10231-022-01263-1
  • [3] B. Y. Chen, Total mean curvature and submanifolds of finite type, Series in Pure Mathematics, vol. 27, 2nd edn. World Scientific Publishing Co. Pte. Ltd., Hackensack (2015).
  • [4] S. P. Chang, A closed hypersurface with constant scalar and mean curvatures in 𝕊4\mathbb{S}^{4} is isoparametric. Comm. Anal. Geom. 1(1) (1993), 71 C-100.
  • [5] S. P. Chang, On closed hypersurfaces of constant scalar curvatures and mean curvatures in 𝕊n+1\mathbb{S}^{n+1}. Pacific J. Math. 165(1) (1994), 67–76.
  • [6] H. Chen and Z. Guan, Triharmonic CMC hypersurfaces in space forms with at most 3 distinct principal curvatures, Results Math. 77 (2022) No 4, Paper No. 155, 17 pp.
  • [7] H. Chen and Z. Guan, Triharmonic CMC hypersurfaces in space Forms, Manuscript Math. (2022). https://doi.org/10.1007/s00229-022-01409-8
  • [8] J. Eells and L. Lemaire, Selected topics in harmonic maps, In proceedings of the CBMS regional conference series in Mathematics, Providence, RI, USA, 31 (1983).
  • [9] Y. Fu, D. Yang and X. Zhan, Recent progress of biharmonic hypersurfaces in space forms, Contemp. Math. 777 (2022), 91–101.
  • [10] G. Y. Jiang, 2-harmonic maps and their first and second variational formulas, Chin. Ann. Math. Ser. A 7(4) (1986), 389–402.
  • [11] S. Maeta, kk-harmonic maps into a Riemannian manifold with constant sectional curvature, Proc. Am. Math. Soc. 140(5) (2012), 1835–1847.
  • [12] S. Maeta, The second variational formula of the kk-energy and kk-harmonic curves, Osaka J. Math. 49(4) (2012), 1035–1063.
  • [13] S. Maeta, Construction of triharmonic maps, Houston J. Math. 41(2) (2015), 433–444.
  • [14] S. Maeta, N. Nakauchi and H. Urakawa, Triharmonic isometric immersions into a manifold of non-positively constant curvature, Monatsh. Math. 177(4) (2015), 551–567.
  • [15] S. Montaldo, C. Oniciuc and A. Ratto, Polyharmonic hypersurfaces into space forms, Isr. J. Math. 249(1) (2022), 343–374.
  • [16] S. Montaldo and A. Ratto, New examples of rr-harmonic immersions into the sphere, J. Math. Anal. Appl. 458(1) (2018), 849–859.
  • [17] N. Nakauchi and H. Urakawa, Polyharmonic maps into the Euclidean space, Note Mat. 38(1) (2018), 89–100.
  • [18] K. Nomizu, Characteristic roots and vectors of a differentiable family of symmetric matrices, Linear and Multilinear Algebra 1 (1973), no. 2, 159–162.
  • [19] Y. L. Ou and B. Y. Chen, Biharmonic submanifolds and biharmonic maps in Riemannian geometry, World Scientific Publishing, Hackensack, NJ, 2020.
  • [20] P. J. Ryan, Homogeneity and some curvature conditions for hypersurfaces, Tohoku Math. J. (2) 21 (1969), 363–388.
  • [21] P. J. Ryan, Hypersurfaces with parallel Ricci tensor, Osaka Math. J. 8 (1971), 251–259.
  • [22] Z. Z. Tang, D. Y. Wei and W. J. Yan, A suffcient condition for a hypersurface to be isoparametric, Tohoku Math. J. (2) 72 (2020) no. 4, 493–505.
  • [23] Z. Z. Tang and W. J. Yan, On the Chern conjecture for isoparametric hypersurfaces, Sci. China Math. (2023) 66, 143–162.
  • [24] S. B. Wang, The first variation formula for kk-harmonic mapping, J. Nanchang Univ. 13(1) (1989).
  • [25] S. B. Wang, Some results on stability of 3-harmonic mappings, Chin. Ann. Math. Ser. A 12(4) (1991), 459–467.