02.02.2013 Views

The second variational formula for Willmore submanifolds

The second variational formula for Willmore submanifolds

The second variational formula for Willmore submanifolds

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>The</strong> <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> <strong>for</strong> <strong>Willmore</strong><br />

<strong>submanifolds</strong><br />

n ∗ in S<br />

§0. Introduction<br />

Dedicated to Professor S. S. Chern at the occasion of his 90th birthday<br />

Zhen Guo, Haizhong Li, Changping Wang<br />

Abstract: In [17] the third author presented Moebius geometry <strong>for</strong> <strong>submanifolds</strong><br />

in S n and calculated the first <strong>variational</strong> <strong><strong>for</strong>mula</strong> of the <strong>Willmore</strong><br />

functional by using Moebius invariants. In this paper we present<br />

the <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> <strong>for</strong> <strong>Willmore</strong> <strong>submanifolds</strong>. As an application<br />

of these <strong>variational</strong> <strong><strong>for</strong>mula</strong>s we give the standard examples of<br />

<strong>Willmore</strong> hypersurfaces {W m k := Sk ( � (m − k)/m)×S m−k ( � k/m), 1 ≤<br />

k ≤ m − 1} in S m+1 (which can be obtained by exchanging radii in the<br />

Clif<strong>for</strong>d tori S k ( � k/m) × S m−k ( � (m − k)/m)) and show that they are<br />

stable <strong>Willmore</strong> hypersurfaces. In case of surfaces in S 3 , the stability of<br />

the Clif<strong>for</strong>d torus S 1 ( 1<br />

√ 2 ) × S 1 ( 1<br />

√ 2 ) was proved by J. L. Weiner in [18].<br />

We give also some examples of m-dimensional <strong>Willmore</strong> <strong>submanifolds</strong><br />

in an n-dimensional unit sphere S n .<br />

Keywords: <strong>Willmore</strong> functional, <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong>, stability,<br />

<strong>Willmore</strong> conjecture<br />

2000 MR Subject Classification 53A30, 53B25<br />

Let M be an m-dimensional submanifold immersed in Sn . <strong>The</strong> <strong>Willmore</strong> functional W is<br />

defined by<br />

W (M) =<br />

� � m<br />

m 2<br />

m − 1<br />

� �<br />

m<br />

S − m�H�<br />

2� 2<br />

dM, (0.1)<br />

M<br />

∗ <strong>The</strong> <strong>second</strong> author is partially supported by the Alexander von Humboldt Research Fellowship and the<br />

US-China Cooperative Research NSF Grant INT-9906856 and the third author is partially supported by<br />

973-project, DFG466-CHV-II3/127/0, Qiushi Award and RFDP.<br />

1


which is a con<strong>for</strong>mal invariant under Moebius (or con<strong>for</strong>mal) trans<strong>for</strong>mations of S n (see [3],<br />

[17], [19]), where S is the square of the length of the <strong>second</strong> fundamental <strong>for</strong>m, H is the mean<br />

curvature vector and dM is the volume element of the induced metric on M. In case of a<br />

surface in S 3 , this functional is equivalent to the <strong>Willmore</strong> functional We(M) := 4 �<br />

M H2 dM<br />

<strong>for</strong> surfaces in R 3 . <strong>The</strong> so-called <strong>Willmore</strong> conjecture states that We(M) ≥ 8π 2 <strong>for</strong> any<br />

embedded torus in R 3 , well-studied (cf. [2], [9], [19], [20], [21]) since 1965. An equivalent<br />

version of the <strong>Willmore</strong> conjecture states that W (T 2 ) ≥ 8π 2 <strong>for</strong> any embedded torus T 2 in<br />

S 3 and the equality holds if and only if T 2 is Moebius equivalent to the Clif<strong>for</strong>d torus. Many<br />

ef<strong>for</strong>ts and partial results have been obtained <strong>for</strong> this conjecture, but up to now it is still<br />

open.<br />

From the analytic point of view, the <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> of the <strong>Willmore</strong> functional<br />

might be very important to solve the <strong>Willmore</strong> conjecture. In 1978, J. L. Weiner gave the<br />

<strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> <strong>for</strong> minimal surfaces, which is a special class of <strong>Willmore</strong> surfaces<br />

(cf. [18]). But it is rather complicated to give the <strong>second</strong> variation <strong><strong>for</strong>mula</strong> <strong>for</strong> general<br />

<strong>Willmore</strong> surfaces by using Euclidean invariants. In [12] M. Peterson derives such a <strong><strong>for</strong>mula</strong>.<br />

In [10] and [11] B. Palmer calculated the <strong>second</strong> variation of a <strong>Willmore</strong> surface in S 3 by<br />

using its con<strong>for</strong>mal Gauss map and used the <strong><strong>for</strong>mula</strong> to study the nonexistence of stable<br />

<strong>Willmore</strong> surfaces under some conditions.<br />

Since the <strong>Willmore</strong> functional (0.1) is invariant under the Moebius group (cf. [3], [17]),<br />

one can use the framework of Moebius geometry and Moebius invariants to calculate the<br />

<strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong>. It is the key point of this paper. For any submanifold M in<br />

S n we can introduce a Moebius invariant metric g on M. <strong>The</strong>n the <strong>Willmore</strong> functional<br />

is exactly the volume functional of g. <strong>The</strong> third author computed the first variation and<br />

got the Euler - Lagrange equations in [17]. Submanifolds in S n satisfying these equations<br />

are called <strong>Willmore</strong> <strong>submanifolds</strong> or Moebius minimal <strong>submanifolds</strong>. In this paper we give<br />

the <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> of the <strong>Willmore</strong> functional <strong>for</strong> <strong>submanifolds</strong> in S n by using<br />

Moebius invariants. Although this <strong><strong>for</strong>mula</strong> looks very complicated, in case of surfaces in<br />

S 3 (which is the most important case) the <strong><strong>for</strong>mula</strong> is in good <strong>for</strong>m (cf. §2, (2.43)). Using<br />

the Euler-Lagrange � equations we find the standard examples of <strong>Willmore</strong> hypersurfaces<br />

(m − k)/m) × Sm−k �<br />

( k/m), 1 ≤ k ≤ m − 1} in Sm+1 , which is (euclidean)<br />

{W m k := Sk (<br />

minimal if and only if 2k = m. In some sense, W m � k can be considered � as the dual hypersurface<br />

of the standard minimal hypersurface S k (<br />

k/m) × S m−k (<br />

that the hypersurface W m k are stable <strong>Willmore</strong> hypersurfaces.<br />

(m − k)/m) in S m+1 . We show<br />

We organize this paper as follows. In §1 we give Moebius invariants and local <strong><strong>for</strong>mula</strong>s in<br />

Moebius geometry <strong>for</strong> <strong>submanifolds</strong> in S n . In §2 we calculate the <strong>second</strong> variation <strong><strong>for</strong>mula</strong><br />

<strong>for</strong> <strong>Willmore</strong> <strong>submanifolds</strong> in S n . As an application we prove in §3 that {W m k } are stable<br />

<strong>Willmore</strong> hypersurfaces.<br />

§1. Moebius invariants and local <strong><strong>for</strong>mula</strong>s <strong>for</strong> <strong>submanifolds</strong> in S n<br />

Let x0 : M → S n be an m-dimensional compact submanifold with boundary ∂M, {e1, ..., em}<br />

be a local orthonormal basis of T M with respect to the induced metric dx0 · dx0 and<br />

{θ1, ..., θm} be its dual basis. Let {em+1, ..., en} be the local normal orthonormal vector<br />

field. We make use of the following convention on the ranges of indices:<br />

1 ≤ i, j, k, · · · ≤ m; m + 1 ≤ α, β, γ, . . . ≤ n<br />

2


and we shall agree that repeated indices are summed over the respective ranges. <strong>The</strong>n the<br />

structure equation of x0 can be written as<br />

dx0 = � θiei<br />

dei = � θijej + � h α ijθjeα − θix0<br />

deα = − � h α ijθjei + � θαβeβ.<br />

� h α iieα are the first, the <strong>second</strong><br />

<strong>The</strong> quantities I = dx0 · dx0, II = � hα ijθi ⊗ θjeα and H = 1<br />

m<br />

fundamental <strong>for</strong>m and the mean curvature vector of x0 in Sn , respectively. We define the<br />

function ρ : M → R by<br />

ρ =<br />

�<br />

m<br />

�II − HI�. (1.1)<br />

m − 1<br />

<strong>The</strong> metric g = ρ 2 dx0 · dx0 is called a Moebius metric which is invariant under Moebius<br />

trans<strong>for</strong>mations in S n (cf.[17]); it is positive definite at any non-umbilical point. <strong>The</strong>n the<br />

<strong>Willmore</strong> functional in (0.1) is exactly the Moebius volume functional <strong>for</strong> g:<br />

�<br />

W (M) :=<br />

M<br />

ρ m dM = V olg(M), (1.2)<br />

where dM is the volume element <strong>for</strong> the metric dx0 · dx0. In this paper, it is our purpose to<br />

calculate the <strong>second</strong> variation in the framework of Moebius geometry. We need the following<br />

notation and local <strong><strong>for</strong>mula</strong>s. For more details we refer to [17].<br />

be the Lorentz space with the inner product given by<br />

Let R n+2<br />

1<br />

< X, W >= −x 0 w 0 + x 1 w 1 + · · · + x n+1 w n+1 ,<br />

where X = (x 0 , x 1 , · · · , x n+1 ), W = (w 0 , w 1 , · · · , w n+1 ) ∈ R n+2 . <strong>The</strong> half cone in R n+2<br />

1<br />

defined as<br />

C n+1<br />

+<br />

For the immersion x0 : M → S n we define<br />

= {X ∈ R n+2<br />

1 | < X, X >= 0, x 0 > 0}.<br />

Y = ρ(1, x0) : M → C n+1<br />

+ . (1.3)<br />

If ˜x0 : M → Sn is Moebius equivalent to x0, then we have ˜ Y = Y T <strong>for</strong> some T ∈ O(n + 1, 1).<br />

Thus<br />

g =< dY, dY >= ρ 2 dx0 · dx0<br />

(1.4)<br />

is a Moebius invariant. In the following we assume that x0 is an immersion without umbilical<br />

point, which implies that g is positive definite on M. Let Ei = ρ −1 ei, then {Ei} is an<br />

orthonormal basis with respect to the metric g, and its dual basis is {ωi = ρθi}. Set<br />

Yi := Ei(Y ), N := − 1 1<br />

∆Y − < ∆Y, ∆Y > Y,<br />

m 2m2 Eα := (H α , eα + H α x0), (1.5)<br />

where ∆ is the Laplacian of the metric g and H = H α eα is the mean curvature vector with<br />

respect to the induced metric dx0 · dx0.<br />

3<br />

is


Lemma 1.1([17]) {Y, N, Yi, Eα} satisfy the following conditions<br />

< Y, Y >=< N, N >= 0, < Y, N >= 1, < Yi, Yj >= δij;<br />

< Y, Yi >=< N, Yi >=< Y, Eα >=< N, Eα >= 0;<br />

< Eα, Yi >= 0, < Eα, Eβ >= δαβ.<br />

Lemma 1.1 shows that {Y, N, Yi, Eα} <strong>for</strong>ms a Moebius moving frame in R n+2<br />

1<br />

<strong>The</strong> structure equations can be written as<br />

dY = � ωiYi<br />

dN = � ψiYi + � φαEα<br />

dYi = −ψiY − ωiN + � ωijYj + � ωiαEα<br />

dEα = −φαY − � ωiαYi + � ωαβEβ.<br />

By differentiating these equations and using Cartan’s lemma, we obtain<br />

ψi = �<br />

Aijωj, Aij = Aji; ωiα = �<br />

B α ijωj, B α ij = B α ji; φα = �<br />

C α i ωi.<br />

We have the following equations:<br />

�<br />

B α ii = 0, �<br />

i<br />

α,i,j<br />

�<br />

B α �2 ij = m − 1<br />

m<br />

, �<br />

j<br />

along M.<br />

(1.6)<br />

B α ij,j = −(m − 1)C α i . (1.7)<br />

<strong>The</strong> relations between these Moebius invariants and Euclidean invariants are given by<br />

Aij = −ρ −2 (Hessij(logρ) − ei(logρ)ej(logρ) − H α h α ij)<br />

− 1<br />

2 ρ−2 (|∇logρ| 2 − 1 + �<br />

(H<br />

α<br />

α ) 2 )δij, (1.8)<br />

B α ij = ρ −1 (h α ij − H α δij), (1.9)<br />

C α i = − 1<br />

m − 1 Bα ij,j = −ρ −2 (H α ,i + (h α ij − H α δij)ej(logρ)). (1.10)<br />

§2. <strong>The</strong> <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> of the <strong>Willmore</strong> functional<br />

In this section we calculate the <strong>second</strong> variation of the <strong>Willmore</strong> functional (or the Moebius<br />

volume functional) defined by (1.2) or (0.1). Since the variation of the volume depends only<br />

on the normal component of the <strong>variational</strong> vector field (cf. [17]), we will consider the normal<br />

variation.<br />

Let x : M × R → S n be a smooth variation of x0 such that x(·, t) = x0 and dxt(T M) =<br />

dx0(T M) on ∂M <strong>for</strong> each (small) t. <strong>The</strong>se two boundary conditions disappear if ∂M = Ø.<br />

For each t we denote by{ei} a local orthonormal basis <strong>for</strong> T M with respect to dxt · dxt,<br />

by {θi} its dual basis and by {eα} a local orthonormal basis <strong>for</strong> the normal bundle of xt.<br />

be the canonical lift of xt and gt =< dY, dY > be the<br />

Let Y = ρ(1, x) : M × R → C n+1<br />

+<br />

4


Moebius metric of xt. Let {Ei := ρ−1ei} be a local orthonormal basis <strong>for</strong> gt with dual basis<br />

{ωi = ρθi}. <strong>The</strong>n the volume <strong>for</strong> gt can be written as<br />

�<br />

W (t) := V olgt(M) =<br />

From Lemma 1.1 in section 1, we can choose a moving frame<br />

in R n+2<br />

1<br />

{Y, N, Y1, · · · , Ym, Em+1, · · · , En}<br />

M<br />

ω1 ∧ · · · ∧ ωm. (2.1)<br />

along M × R, which satisfies the conditions in Lemma 1.1 <strong>for</strong> each t. Let d denote<br />

the differential operator on M × R, then we can find 1-<strong>for</strong>ms<br />

{V, Vα, Ψi, Φα, Ωi, Ωij, Ωiα, Ωαβ}<br />

on M × R with Ωij = −Ωji and Ωαβ = −Ωβα such that<br />

Taking the differentials of these equations we get<br />

dY = V Y + ΩiYi + VαEα, (2.2)<br />

dN = −V N + ΨiYi + ΦαEα, (2.3)<br />

dYi = −ΨiY − ΩiN + ΩijYj + ΩiαEα, (2.4)<br />

dEα = −Φα − VαN − ΩiαYi + ΩαβEβ. (2.5)<br />

dV = Ψi ∧ Ωi + Φα ∧ Vα, (2.6)<br />

dΩi = Ωij ∧ Ωj + V ∧ Ωi − Vα ∧ Ωiα, (2.7)<br />

dVα = Ωαβ ∧ Vβ + Ωi ∧ Ωiα + V ∧ Vα, (2.8)<br />

dΨi = Ωij ∧ Ψj − Φα ∧ Ωiα + Ψi ∧ V, (2.9)<br />

dΦα = Ωαβ ∧ Φβ + Ψi ∧ Ωiα + Φα ∧ V, (2.10)<br />

dΩij = Ωik ∧ Ωkj + Ωiα ∧ Ωαj − Ψi ∧ Ωj − Ωi ∧ Ψj, (2.11)<br />

dΩiα = Ωij ∧ Ωjα + Ωiβ ∧ Ωβα − Ψi ∧ Vα − Ωi ∧ Φα, (2.12)<br />

dΩαβ = Ωαγ ∧ Ωγβ + Ωαi ∧iβ −Φα ∧ Vβ − Vα ∧ Φβ. (2.13)<br />

Since Y = ρ(1, x), if we write the normal variation vector field of x in T S n by<br />

then by (1.5) we can find a function v : M × R → R such that<br />

∂x<br />

∂t = ρ−1 vαeα, (2.14)<br />

∂Y<br />

∂t = vY + vαEα. (2.15)<br />

From (2.2), (2.15) and the fact that d = ωiEi + dt ∂<br />

∂t on C∞ (M × R), we get<br />

V = vdt, Vα = vαdt, Ωi = ωi. (2.16)<br />

5


Since T ∗ (M × R) = T ∗ M ⊕ T ∗ R we can write<br />

Ψi = ψi + aidt, Φα = φα + bαdt, (2.17)<br />

Ωij = ωij + Pijdt, Ωiα = ωiα + Liαdt, Ωαβ = ωαβ + Qαβdt, (2.18)<br />

where {ai, bα, Pij, Qαβ} are local functions with Pij = −Pji and Qαβ = −Qβα. Let {Bα ij, Aij, Cα i }<br />

be Moebius invariants <strong>for</strong> xt defined in §1. If we denote by dM the exterior differential operator<br />

on T ∗M, then we have d = dM + dt ∧ ∂<br />

∂t on T ∗ (M × R). It follows from (2.6), (2.16)<br />

and (2.17) (comparing the terms in T ∗ (M) ∧ dt ) that<br />

ai = −v,i + vαC α i , (2.19)<br />

where v,i := Ei(v). Similarly we get from (2.8), (2.16), (2.17), (2.18) that<br />

and get from (2.7), (2.16), (2.18) that<br />

Liα = vα,i<br />

(2.20)<br />

∂ωi<br />

∂t = (Pij + vδij − vαB α ij)ωj. (2.21)<br />

By a direct calculation in a similar way, we obtain from (2.9)∼(2.13) that<br />

∂ωij<br />

∂t = (Pij,k + B α ikvα,j − B α jkvα,i − aiδkj + ajδik)ωk, (2.22)<br />

∂ψi<br />

∂t = (ai,j + PikAkj + vα,iC α j − bαB α ij − vAij)ωj (2.23)<br />

∂φα<br />

∂t = (bα,i + QαβC β<br />

i − Aijvα,j + ajB α ji − vC α i )ωi, (2.24)<br />

∂ωiα<br />

∂t = (vα,ij + PikB α kj − B β<br />

ijQβα + Aijvα + bαδij)ωj<br />

(2.25)<br />

∂ωαβ<br />

∂t = (Qαβ,i + vβ,jB α ji − vα,jB β<br />

ji + vβC α i − vαC β<br />

i )ωi, (2.26)<br />

where {vα,i} are covariant derivatives of {vα}. Since φα = Cα i ωi and ψi = Aijωj, from (2.21),<br />

(2.23) and (2.24) we have<br />

∂C α i<br />

∂t = bα,i + QαβC β<br />

i + ajB α ij − Aijvα,j + PijC α j + B β<br />

ijC α j vβ − 2vC α i , (2.27)<br />

∂Aij<br />

∂t = ai,j + PikAkj − PkjAki + vα,iC α j − B α ijbα + AikB α kjvα − 2vAij. (2.28)<br />

Since ωiα = B α ijωj, from (2.21) and (2.25) we have<br />

∂B α ij<br />

∂t = vα,ij − vB α ij + PikB α kj − PkjB α ki − B β<br />

ijQβα + vβB α ikB β<br />

kj + Aijvα + bαδij. (2.29)<br />

6


Multiplying B α ij to (2.29) and using (1.7) we get<br />

m − 1<br />

m v = Bα ijvα,ij + B α ikB β<br />

kj Bα jivβ + AijB α ijvα. (2.30)<br />

Making use of (2.21), (2.30), (1.7), Green’s <strong><strong>for</strong>mula</strong> and the conditions on ∂M, we get by<br />

noting (2.34) of [17]<br />

W ′ (t) = m2<br />

� �<br />

−(m − 1)C<br />

m − 1 M<br />

α i,i + B β<br />

ikBα kjB β<br />

ji + AijB α �<br />

ij vαdMt, (2.31)<br />

where dMt is volume element of gt.<br />

We summarize as follows<br />

Proposition 2.1. ([17]) x0 : M → S n is a <strong>Willmore</strong> submanifold (i.e., its <strong>Willmore</strong><br />

functional is stationary) if and only if<br />

−(m − 1)C α i,i + B β<br />

ik Bα kjB β<br />

ji + AijB α ij = 0. (2.32)<br />

Since ∂<br />

∂t ◦ dM − dM ◦ ∂<br />

∂t = 0 (or equivalently d2 = (dM + dt ∧ ∂<br />

∂t )2 = 0) and<br />

C α i,jωj = dMC α i + C α j ωji + C β<br />

i ωβα,<br />

we get<br />

∂C α i,j<br />

∂t = −Cα ∂ωk<br />

i,k<br />

∂t (Ej) + ( ∂Cα i<br />

∂t ),j + C α ∂ωki<br />

k<br />

∂t (Ej) + C β ∂ωβα<br />

i<br />

∂t (Ej). (2.33)<br />

By a direct calculation and use of (2.21), (2.22), (2.26) and (2.33), we get<br />

∂C α i,i<br />

∂t vα = −(C α i Pki),kvα − vC α i,ivα + vβB β<br />

kiCα �<br />

α ∂Ci i,kvα +<br />

∂t<br />

+C α k B β<br />

ki vβ,ivα + akC α k vα − makC α k vα − C β<br />

i Qαβ,ivα − B α kiC β<br />

i vβ,kvα<br />

+B β<br />

ki Cβ<br />

i vα,kvα − C β<br />

i C α i vβvα + �<br />

i,β<br />

�<br />

C β<br />

i<br />

�2 �<br />

v<br />

α<br />

2 α.<br />

By a straight<strong>for</strong>ward calculation and using (2.27), (2.19), we get from (2.34) that<br />

∂C α i,i<br />

∂t vα = −(C α �<br />

α ∂Ci i Pkivα),k +<br />

∂t vα<br />

�<br />

+ (vB<br />

,i<br />

α kivα,i),k<br />

+(m − 1)(vC α k vα),k − (C β<br />

i Qαβvα),i − (bαvα,i),i − mvC α i,ivα<br />

+C β<br />

i,iQαβvα + C α i,kB β<br />

kivαvβ + bαvα,ii − vB α kivα,ik + Aikvα,kvα,i<br />

−2C α k B β<br />

kivβvα,i + 2vC α i vα,i − m �<br />

�<br />

�<br />

C<br />

k α<br />

α �2 k vα<br />

+C β<br />

i B β<br />

ikvα,kvα + �<br />

(C<br />

i,β<br />

β<br />

�<br />

2<br />

i ) v<br />

α<br />

2 α.<br />

7<br />

�<br />

,i<br />

vα<br />

(2.34)<br />

(2.35)


From the conditions that x(., t) = x0 and dxt(T M) = dx0(T M) on ∂M <strong>for</strong> each t, we see<br />

that vα|∂M = 0 and<br />

0 = ∂<br />

∂t (dxt) = d( ∂x<br />

∂t ) = ρ−1dvαeα (2.36)<br />

at the boundary ∂M. It follows from (2.35) and Green’s <strong><strong>for</strong>mula</strong> that<br />

�<br />

Γ1:= {<br />

M<br />

∂<br />

∂t |t=0(C α i,i)}vαdM<br />

�<br />

= {−mvC<br />

M<br />

α i,ivα + C β<br />

i,iQαβvα + C α i,kB β<br />

kivαvβ + bαvα,ii<br />

−vB α kivα,ik + Aikvα,kvα,i − 2C α k B β<br />

kivβvα,i + 2vC α i vα,i<br />

−m �<br />

�<br />

�<br />

C<br />

k α<br />

α �2 k vα + C β<br />

i B β<br />

ikvα,kvα + �<br />

(C<br />

i,β<br />

β<br />

�<br />

2<br />

i )<br />

α<br />

v 2 α}dM.<br />

(2.37)<br />

Using (2.28), (2.29), (2.19) and the facts that Pij = −Pji and �<br />

j B α ij,j = −(m − 1)C α i , we<br />

get<br />

∂ �<br />

B<br />

∂t<br />

β<br />

ikBα kjB β<br />

ji + AijB α �<br />

ij vα<br />

= (vB α ijvα,j),i − (m − 1)(vC α i vα),i + (aiB α ijvα),j<br />

−2(m − 1)vC α i,ivα + 2B β<br />

ijB α jkvβ,ikvα + (B β<br />

ikBβ kj + Aij)vα,ijvα − vB α ijvα,ij<br />

+(m − 1)C β<br />

i,iQαβvα + 2(m − 1)vC α j vα,j + C γ<br />

i B α ijvγ,jvα − C γ<br />

i B α ijvα,jvγ<br />

+ �<br />

3B β<br />

ijB α jkB γ<br />

klBβ li + 4Bα ikB γ<br />

kjAji + (m − 1)C α i C γ<br />

�<br />

i vαvγ<br />

+ � �<br />

(B<br />

i,j β<br />

β<br />

ijbβ)( �<br />

B<br />

α<br />

α ijvα) + ( �<br />

⎛<br />

bαvα) ⎝<br />

α<br />

�<br />

(B<br />

β,i,j<br />

β<br />

ij) 2 ⎞<br />

+ tr(A) ⎠<br />

+ �<br />

A<br />

i,j<br />

2 �<br />

ij v<br />

α<br />

2 α + B β β<br />

ikAkjBji( �<br />

v<br />

α<br />

2 α).<br />

Using the boundary conditions and Green’s theorem we have<br />

�<br />

Γ2:=<br />

where<br />

�<br />

=<br />

M<br />

M<br />

{ ∂ β<br />

|t=0(Bik ∂t Bα klB β<br />

lj + AijB α ij)}vαdM<br />

{−2(m − 1)vC α i,ivα + 2B β<br />

ijB α jkvβ,ikvα + (B β<br />

ikBβ kj + Aij)vα,ijvα<br />

−vB α ijvα,ij + (m − 1)C β<br />

i,iQαβvα + 2(m − 1)vC α j vα,j + C γ<br />

i B α ijvγ,jvα<br />

−C γ<br />

i B α ijvα,jvγ + �<br />

3B β<br />

ijB α jkB γ<br />

klBβ li + 4Bα ikB γ<br />

kjAji + (m − 1)C α i C γ<br />

�<br />

i vαvγ<br />

+ � �<br />

(B β<br />

ijbβ)( �<br />

B α ijvα) + ( � m − 1<br />

bαvα)( + tr(A))<br />

m<br />

ij<br />

+ �<br />

ij<br />

β<br />

A 2 ij<br />

�<br />

α<br />

α<br />

α<br />

v 2 α + B β β<br />

ikAkjBji( �<br />

α<br />

v 2 α)}dM,<br />

(2.38)<br />

(2.39)<br />

bα = − 1<br />

�<br />

∆vα + B<br />

m<br />

α ikB β<br />

kivβ + 1<br />

2m (1 + m2 �<br />

κ)vα , (2.40)<br />

8


tr(A) = 1<br />

2m<br />

m<br />

+ κ, (2.41)<br />

2<br />

where κ is the normalized scalar curvature of the metric g = ρ 2 dx0 · dx0. <strong>The</strong> function v<br />

is given by (2.15) and (2.30). Thus <strong>for</strong> a <strong>Willmore</strong> submanifold x0, we get from (2.31) and<br />

(1.7) that<br />

W ′′ (0) = −m 2<br />

�<br />

M<br />

∂C α i,i<br />

∂t |t=0vαdM<br />

+ m2<br />

� �<br />

∂ β<br />

|t=0(Bik m − 1 M ∂t Bα kjB β<br />

ji + AijB α �<br />

ij) vαdM.<br />

Thus the <strong><strong>for</strong>mula</strong> of the <strong>second</strong> variation is given by (2.37) and (2.39). We conclude that<br />

<strong>The</strong>orem 2.1. Let x : M → S n be a compact <strong>Willmore</strong> submanifold without boundary.<br />

<strong>The</strong>n the <strong>second</strong> variation <strong><strong>for</strong>mula</strong> of the Moebius volume functional is given by<br />

W ′′ �<br />

(0) =<br />

M<br />

�<br />

m 2 (m − 2)vC α i,ivα − m 2 C α i,jB β<br />

jivαvβ + m2 (m − 2)<br />

vB α ijvα,ij<br />

m − 1<br />

−m 2 Aijvα,ivα,j + m2 (2m − 1)<br />

m − 1 Cβ i B α ijvβ,jvα + m 2 (m + 1) �<br />

�<br />

�<br />

C<br />

i α<br />

α �2 i vα<br />

−m 2 C β<br />

i B β<br />

�<br />

2<br />

ijvα,jvα − m<br />

i,β<br />

�<br />

C β<br />

i<br />

� 2 �<br />

α<br />

v 2 α + 2m2<br />

m − 1 Bβ<br />

ijB α jkvβ,ikvα<br />

+ m2<br />

m − 1 (Bβ<br />

ik Bβ<br />

kj + Aij)vα,ijvα − m2<br />

m − 1 Cβ<br />

i B α ijvα,jvβ<br />

+ m2 �<br />

3B<br />

m − 1<br />

γ<br />

ijB α jkB β<br />

klBγ li + 4Bα ikB β<br />

kjAji �<br />

vαvβ + m2<br />

m − 1<br />

+ m2<br />

m − 1 Bβ<br />

�<br />

�<br />

β<br />

ikAkjBji v<br />

α<br />

2 �<br />

α<br />

+ m �<br />

(∆vα) 2 +<br />

α<br />

�<br />

A<br />

ij<br />

2 �<br />

ij v<br />

α<br />

2 α<br />

m(m − 2)<br />

m − 1 Bα ijB β<br />

ij∆vαvβ<br />

� �<br />

m(m − 2)<br />

+<br />

tr(A) − 1 vα∆vα −<br />

m − 1 m<br />

⎛<br />

�<br />

⎝<br />

m − 1 α<br />

�<br />

B<br />

ijβ<br />

α ijB β<br />

⎞2<br />

⎠<br />

ijvβ<br />

�<br />

− 1 + 2m<br />

m − 1 tr(A)<br />

� �<br />

� �<br />

B<br />

ij α<br />

α �2 ijvα − tr(A)<br />

�<br />

1 + m<br />

m − 1 tr(A)<br />

�<br />

�<br />

v<br />

α<br />

2 �<br />

α dM.<br />

(2.42)<br />

In case of the surface in S 3 , we omit all α and β because the codimension now is one,<br />

the <strong><strong>for</strong>mula</strong> is reduced to<br />

Corollary 2.1. For a <strong>Willmore</strong> surface x0 : M → S 3 , the <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> is<br />

9


given by<br />

W ′′ �<br />

(0) =<br />

M<br />

−4 �<br />

ij<br />

{2(∆f) 2 + 2f∆f + 12 �<br />

Aijfifj + (4 �<br />

ij<br />

ij<br />

A 2 ij + 4 �<br />

CiBijfjf + 4 �<br />

Aijfijf<br />

i<br />

ij<br />

C 2 i + 7<br />

8 + K − 2K2 )}dM,<br />

where K is Gaussian curvature of the Moebius metric g = ρ 2 dx0 · dx0.<br />

(2.43)<br />

Remark 2.1. <strong>The</strong> <strong>second</strong> <strong>variational</strong> <strong><strong>for</strong>mula</strong> <strong>for</strong> <strong>Willmore</strong> surfaces in S 3 might be important<br />

to solve the <strong>Willmore</strong> conjecture. As far as we know the only known stable example of a<br />

<strong>Willmore</strong> torus is the Clif<strong>for</strong>d torus. From the existence result of L. Simon in [15], we know<br />

that the <strong>Willmore</strong> conjecture is true if one can show that the only stable <strong>Willmore</strong> torus<br />

embedded in S 3 is the Clif<strong>for</strong>d torus.<br />

§3. <strong>Willmore</strong> tori in S m+1 and their stabilities<br />

In this section we present a class of important examples of <strong>Willmore</strong> hypersurfaces called<br />

<strong>Willmore</strong> tori. As an application of <strong>The</strong>orem 2.1 we show that they are stable <strong>Willmore</strong><br />

hypersurfaces.<br />

Let R m+2 be the (m+2)-dimensional Euclidean space with inner product . We write<br />

R m+2 = R k+1 × R m−k+1 , 1 ≤ k ≤ m − 1. For any vector ξ ∈ R m+2 there is a unique<br />

decomposition ξ = ξ1 + ξ2 with ξ1 ∈ R k+1 and ξ2 ∈ R m−k+1 . For another vector η = η1 + η2<br />

the inner product of them can be written as < ξ, η >=< ξ1, η1 > + < ξ2, η2 >. Let<br />

ξ1 : S k → R k+1 and ξ2 : S m−k → R m−k+1 be standard embeddings of unit spheres. Let<br />

x : S k (a1) × S m−k (a2) → S m+1 ⊂ R m+2 be the embedded hypersurface x = a1ξ1 + a2ξ2 with<br />

a 2 1 + a 2 2 = 1. It is easy to check that<br />

(i) the unit normal vector of M := S k (a1) × S m−k (a2) in S m+1 is given by<br />

em+1 = −a2ξ1 + a1ξ2;<br />

(ii) the <strong>second</strong> fundamental <strong>for</strong>m of M is given by<br />

II = − < dx, dem+1 >= a1a2(< dξ1, dξ1 > − < dξ2, dξ2 >);<br />

(iii) the induced metric of M is given by<br />

If we take {ei} and {ωi}such that<br />

I = a 2 1|dξ1| 2 + a 2 2|dξ2| 2 .<br />

k�<br />

d(a1ξ1) = ωiei, d(a2ξ2) =<br />

n�<br />

ωjej,<br />

i=1<br />

j=k+1<br />

10


then we have<br />

where<br />

m�<br />

I = ω<br />

i=1<br />

2 k� a2<br />

i , II = ω<br />

i=1<br />

a1<br />

2 m� a1<br />

i − ω<br />

j=k+1<br />

a2<br />

2 j := hijωiωj, (3.1)<br />

⎧<br />

⎪⎨<br />

hij =<br />

⎪⎩<br />

a2<br />

a1 δij, if 1 ≤ i, j ≤ k,<br />

− a1<br />

a2 δij, if k + 1 ≤ i, j ≤ n.<br />

(3.2)<br />

<strong>The</strong>orem 3.1. Let W = S k (a1) × S m−k (a2) be the hypersurface imbedded into S m+1 ,<br />

where a 2 1 + a 2 2 = 1. <strong>The</strong>n W is a <strong>Willmore</strong> hypersurface if and if<br />

Proof. From (3.1) we see that<br />

S :=<br />

H := 1<br />

m<br />

a1 =<br />

�<br />

m − k<br />

m , a2<br />

�<br />

k<br />

=<br />

m .<br />

m�<br />

h<br />

i,j<br />

2 � �<br />

a2<br />

2<br />

� �<br />

a1<br />

2<br />

ij = k + (m − k) , (3.3)<br />

a1<br />

a2<br />

m�<br />

i=1<br />

hii = 1<br />

�<br />

k<br />

m<br />

a2<br />

a1<br />

− (m − k) a1<br />

a2<br />

�<br />

, (3.4)<br />

ρ 2 = m<br />

m − 1 (S − mH2 ). (3.5)<br />

Substituting (3.2) and (3.5) into (1.8)-(1.10) we know that the <strong>Willmore</strong> condition<br />

is equivalent to the equation<br />

(m − k)<br />

� �<br />

a2<br />

6<br />

a1<br />

−(m − 1)Ci,i + BikBkjBji + AijBij = 0<br />

+ (2m − 3k)<br />

� �<br />

a2<br />

4<br />

a1<br />

+ (m − 3k)<br />

� �<br />

a2<br />

2<br />

From the equations (3.6) and a2 1 + a2 �<br />

m−k<br />

2 = 1 we get a1 = m and a2<br />

�<br />

k = m .<br />

We call<br />

W m k = S k<br />

⎛�<br />

⎞<br />

m − k<br />

⎝ ⎠ × S<br />

m<br />

m−k<br />

⎛�<br />

⎞<br />

k<br />

⎝ ⎠ , 1 ≤ k ≤ m − 1,<br />

m<br />

a <strong>Willmore</strong> torus.<br />

a1<br />

− k = 0. (3.6)<br />

Remark 3.1. It is remarkable that the <strong>Willmore</strong> torus W m k is (Euclidean) minimal if<br />

and only if 2k = m. We note that W m �<br />

k can be obtained by exchanging the radii k/m and<br />

�<br />

m − k/m in the Clif<strong>for</strong>d minimal torus Sk �� �<br />

k × S m<br />

m−k �� �<br />

m−k . m<br />

Remark 3.2. It is known that any minimal surface in S n is a <strong>Willmore</strong> surface (see Weiner<br />

[18]) (also see Pinkall’s examples of non-minimal <strong>Willmore</strong> surfaces [13]). <strong>The</strong>orem 3.1<br />

shows that m-dimensional (m ≥ 3) minimal hypersurfaces are not <strong>Willmore</strong> hypersurfaces<br />

in general.<br />

11


Remark 3.3. In [6], the <strong>second</strong> author proved an integral inequality of Simons’ type <strong>for</strong><br />

m-dimensional compact <strong>Willmore</strong> hypersurfaces in S m+1 and gave a characterization of <strong>Willmore</strong><br />

tori W m k by use of his integral inequality.<br />

From now on we study the stability of <strong>Willmore</strong> tori W m k defined in <strong>The</strong>orem 3.1. For<br />

W m k we get from (3.2), (3.3), (3.4) and (3.5) that<br />

⎧ �<br />

⎨ k<br />

m−k<br />

hij =<br />

⎩<br />

δij, if 1 ≤ i, j ≤ k,<br />

�<br />

m−k − k δij, if k + 1 ≤ i, j ≤ m,<br />

S = k3 + (m − k) 3<br />

k(m − k)<br />

From (1.8) and (1.9) we have<br />

⎧<br />

⎨<br />

=<br />

⎩<br />

Bij = ρ −1 ⎧<br />

⎨<br />

(hij − Hδij) =<br />

⎩<br />

(3.7)<br />

m − 2k<br />

, H = −� k(m − k) , ρ2 = m2<br />

. (3.8)<br />

m − 1<br />

Aij = ρ −2 Hhij + 1<br />

2 ρ−2 (1 − H 2 )δij<br />

ρ −2 3km−k2 −m 2<br />

2k(m−k) δij, if 1 ≤ i, j ≤ k,<br />

ρ −2 m2 −k 2 −km<br />

2k(m−k) δij, if k + 1 ≤ i, j ≤ m,<br />

(3.9)<br />

tr(A) = (m − 1)(m2 − 3km + 3k2 )<br />

, (3.10)<br />

2km(m − k)<br />

ρ −1� m−k<br />

k δij, if 1 ≤ i, j ≤ k<br />

−ρ −1� k<br />

m−k δij, if k + 1 ≤ i, j ≤ m,<br />

(3.11)<br />

and � B 2 ij = m−1<br />

m . From the last equation in (1.7) we obtain Ci = 0. We are going to<br />

calculate<br />

W ′′ (0) = −m 2 Γ1 + m2<br />

m − 1 Γ2,<br />

{ ∂<br />

∂t |t=0(BikBklBli + AijBji)}fdM and f ∈ C ∞ (M) is<br />

where Γ1 = � ∂Ci,i<br />

M |t=0fdM, Γ2 = ∂t �<br />

M<br />

the normal component of the <strong>variational</strong> vector field. From (2.40) we have<br />

b = − 1<br />

�<br />

∆f +<br />

m<br />

(m − 1)(m2 + k2 �<br />

− km)<br />

f . (3.12)<br />

2km(m − k)<br />

Substituting (3.9), (3.11) and (3.12) into (2.37), we get<br />

−m2 �<br />

Γ1 =<br />

M<br />

�<br />

m(∆f) 2 + (m − 1)(m2 + k2 �<br />

− km)<br />

f∆f<br />

2k(m − k)<br />

� �<br />

2 2 (m − 1)(3km − k − m )<br />

−<br />

|∇1f|<br />

M 2k(m − k)<br />

2 + (m − 1)(m2 − km − k2 )<br />

|∇2f|<br />

2k(m − k)<br />

2<br />

�<br />

dM<br />

+m 2<br />

�<br />

M<br />

vBijfijdM.<br />

12<br />

dM<br />

(3.13)


Substituting (3.9), (3.11) and (3.12) into (2.39), by a straight<strong>for</strong>ward calculation we get<br />

m2 m − 1 Γ2 = − m2 − k2 �<br />

+ km<br />

f∆fdM +<br />

2k(m − k) M<br />

5m2 + 5k2 �<br />

− 9km<br />

f∆1fdM<br />

2k(m − k) M<br />

+ 5k2 + m 2 − km<br />

2k(m − k)<br />

�<br />

M<br />

f∆2fdM − m2<br />

�<br />

vBijfijdM<br />

m − 1 M<br />

+ �<br />

2km(m − k)(m 2 + 7km − 7k 2 ) − m(m 2 − k 2 + km)(k 2 + m 2 − km)<br />

+ k(3km − m 2 − k 2 ) 2 + (m − k)(m 2 − km − k 2 ) 2�<br />

m − 1<br />

4k2m2 (m − k) 2<br />

�<br />

f<br />

M<br />

2 dM.<br />

(3.14)<br />

In (3.13) and (3.14) we denote by ∆ the Laplace operator on W m k with respect to the<br />

Moebius metric g = ρ2dx · dx. � We write g = g1 ⊕ g2 �<br />

according to the decomposition<br />

M m k := M1×M2 := S k �� m−k<br />

m<br />

×S m−k �� k<br />

m<br />

. We denote by {∆1, ∆2} the Laplace operators<br />

of {g1, g2} and by {∇1, ∇2} the gradient operators on M1 and M2, respectively. From (2.30),<br />

(3.9) and (3.11), we have<br />

From(3.13), (3.14) and (3.15) we get<br />

W ′′ �<br />

(0) =<br />

vBijfij = 1<br />

⎛�<br />

m − k<br />

⎝<br />

m k ∆1f<br />

�<br />

k<br />

−<br />

m − k ∆2f<br />

⎞2<br />

⎠ . (3.15)<br />

M<br />

�<br />

m(∆f) 2 + m(k2 + m2 − km − 2m)<br />

f∆f<br />

2k(m − k)<br />

+ m(m − 2)<br />

⎛�<br />

m − k<br />

⎝<br />

m − 1 k ∆1f<br />

�<br />

k<br />

−<br />

m − k ∆2f<br />

⎞2<br />

⎠<br />

+ m(3km − k2 − m2 ) + 6(m − k) 2<br />

f∆1f<br />

2k(m − k)<br />

+ m(m2 − km − k2 ) + 6k2 f∆2f +<br />

2k(m − k)<br />

�<br />

2(m − 1) 2<br />

f dM.<br />

m<br />

(3.16)<br />

We denote the Laplace operators with respect to the Euclidean metric dx · dx by ∆M, ∆M1<br />

and ∆M2 on W m k , M1 and M2, respectively. Since ρ = constant and the Moebius metric<br />

g = ρ 2 dx · dx, we have ∆M = ρ 2 ∆, ∆M1 = ρ 2 ∆1, ∆M2 = ρ 2 ∆2 and ∆M = ∆M1 + ∆M2. From<br />

13


(3.8) and (3.16), we have<br />

W ′′ (0) =<br />

m − 1<br />

2k(m − k)m3 � �<br />

2k(m − k)(m − 1)(∆Mf)<br />

M<br />

2<br />

+m 2 (k 2 + m 2 − km − 2m)f∆Mf<br />

+2(m − 2) ((m − k)∆M1f − k∆M2f) 2<br />

+m �<br />

m(3km − k 2 − m 2 ) + 6(m − k) 2�<br />

f∆M1f<br />

+m �<br />

m(m 2 − km − k 2 ) + 6k 2�<br />

f∆M2f<br />

+4k(m − k)m 2 f 2 ) dM.<br />

(3.17)<br />

Let λi, λ ′ i and µij be the eigenvalues of Laplace operators ∆M1, ∆M2 and ∆M, respectively,<br />

then we have<br />

λi = i(k + i − 1) n<br />

n − k , λ′ j = j(n − k + j − 1) n<br />

k , µij = λi + λ ′ j,<br />

where i and j are nonnegative integers. Let fi, f ′ i be eigenfunctions corresponding to λi and<br />

λ ′ i, respectively, then gij(p, q) = fi(p)f ′ j(q) ((p, q) ∈ M1×M2) is an eigenfunction corresponding<br />

to µij. For any f ∈ C ∞ (M m k ) we have the decomposition of f into eigenfuctions<br />

f = �<br />

i+j�=0<br />

where cij and c0 are constants. Thus we have<br />

∆M1f = − �<br />

λicijgij, ∆M2f = − �<br />

i+j�=0<br />

i+j�=0<br />

Substituting (3.18) and (3.19) into <strong><strong>for</strong>mula</strong> (3.17), we get<br />

W ′′ (0) ≥ m−1<br />

2k(m−k)m3 � �<br />

M i+j�=0<br />

cijgij + c0, (3.18)<br />

λ ′ jcijgij, ∆Mf = − �<br />

�<br />

2k(m − k)(m − 1)µ 2 ij<br />

i+j�=0<br />

−m2 (k2 + m2 − km − 2m)µij + 2(m − 2) �<br />

(m − k)λi − kλ ′ �2 j<br />

−m (m(3km − k 2 − m 2 ) + 6(m − k) 2 ) λi<br />

−m (m(m 2 − km − k 2 ) + 6k 2 ) λ ′ j + 4k(m − k)k 2 ) c 2 ijg 2 ijdM<br />

=<br />

m−1<br />

2k(m−k)m3 � �<br />

i+j�=0 M {2(m − k)(m2 − 2m + k)λ2 i<br />

−2(2m 3 + km 3 − 6km 2 + 3k 2 m)λi<br />

+2k(m 2 − m − k)(λ ′ j) 2 − 2m(m 3 − m 2 − km 2 + 3k 2 )λ ′ j<br />

+4k(m − k)λiλ ′ j + 4k(m − k)m 2 }c 2 ijg 2 ijdM.<br />

14<br />

µijcijgij. (3.19)<br />

(3.20)<br />

(3.21)


If we set<br />

A(i, j) = 2(m − k)(m 2 − 2m + k)λ 2 i − 2(2m 3 + km 3 − 6km 2 + 3k 2 m)λi<br />

then one can easily verify that<br />

+2k(m 2 − m − k)(λ ′ j) 2 − 2m(m 3 − m 2 − km 2 + 3k 2 )λ ′ j<br />

+4k(m − k)λiλ ′ j + 4k(m − k)m 2 ,<br />

A(i, j) = � km<br />

(m−k) 2 (m2 − 2m + k)i(k + i − 1) + ( 2km j(m − k + j − 1) − 2km)�<br />

m−k<br />

· � m(m−k)<br />

k 2<br />

(m2 − m − k)j(m − k + j − 1) + ( 2(m−k)m<br />

i(k + i − 1) − 2(m − k)m) k<br />

�<br />

− m2 (m2−2m+k)(m2−m−k) ij(k + i − 1)(m − k + j − 1).<br />

k(m−k)<br />

(3.22)<br />

From (3.22) it is not difficult to see that A(i, j) ≥ 0 and A(i, j) = 0 if and only if (i, j) =<br />

(1, 0), (0, 1) or (1, 1). Thus we have proved the main result in this section:<br />

<strong>The</strong>orem 3.2. All <strong>Willmore</strong> tori Sk �� �<br />

m−k × S m<br />

m−k �� �<br />

k → S m<br />

m+1 , 1 ≤ k ≤ m − 1, are<br />

stable.<br />

Now we would like to propose the following generalized <strong>Willmore</strong> conjecture in S m+1 (cf.<br />

Pinkall [14] and Kobayashi [5]):<br />

Generalized <strong>Willmore</strong> Conjecture: Let Gk be a m-dimensional manifold which is diffeomorphic<br />

to S k × S m−k and x : Gk → S m+1 be an imbedding, where 1 ≤ k ≤ m − 1.<br />

Set<br />

τk(x) =<br />

� � m<br />

m 2<br />

m − 1<br />

� �<br />

m<br />

S − mH<br />

2� 2<br />

dM, (3.23)<br />

Gk<br />

where dM is the volume element of Gk with respect to the induced metric dx · dx, S the<br />

square of the length of the <strong>second</strong> fundamental <strong>for</strong>m and H the mean curvature of x. <strong>The</strong>n<br />

τk(x) ≥<br />

4π m+2<br />

2 (m − k) k<br />

2 k m−k<br />

2<br />

m m<br />

2 −2 (m − 1)Γ( k+1<br />

2<br />

m−k+1 )Γ( ) 2<br />

(3.24)<br />

and equality holds if and only if x(M) is Moebius equivalent to W m k . Here Γ is the gamma<br />

function and the term on the right hand side of (3.24) is the Moebius volume of W m k .<br />

Remark 3.4. When m = 2, the above generalized <strong>Willmore</strong> conjecture reduces to the<br />

<strong>Willmore</strong> conjecture.<br />

§4. Examples of <strong>Willmore</strong> Submanifolds in S n<br />

In this section, we will give some examples of <strong>Willmore</strong> <strong>submanifolds</strong> in a unit sphere S n .<br />

We first prove<br />

15


<strong>The</strong>orem 4.1. Let x : M → S n be a m-dimensional submanifold in an n-dimensional unit<br />

sphere. <strong>The</strong>n x(M) is a m-dimensional <strong>Willmore</strong> submanifold in S n if and only if, <strong>for</strong> any<br />

α with m + 1 ≤ α ≤ n,<br />

−ρm−2 [SH α + Hβh β<br />

ijhα ij − hα ijh β<br />

ikhβkj − m| � H| 2H α ]<br />

+(m − 1)∆(ρm−2H α ) − (ρm−2 ),ij(mHαδij − hα ij) = 0,<br />

(4.1)<br />

where ∆ and (.),ij are the Laplacian and covariant derivatives with respect to the induced<br />

metric dx · dx, respectively.<br />

Proof. From [17], we have the following relations between the connections of the Moebius<br />

metric ρ 2 dx · dx and the induced metric dx · dx, respectively,<br />

By use of (1.10), we get<br />

thus<br />

ωij = θij + (lnρ)iθj − (lnρ)jθi, ωαβ = θαβ. (4.2)<br />

ρC α i,jθj = C α i,jωj<br />

= dC α i + C α j ωji + C β<br />

i ωβα<br />

= dC α i + C α j θji + C β<br />

i θβα + [(lnρ)kθi − (lnρ)iθk] · C α k ,<br />

ρC α i,j = 2ρ −3 ρj[H α ,i + (h α ik − H α δik)(lnρ)k] − ρ −2 [H α ,ij + (h α ik,j − H α ,jδik)(lnρ)k]<br />

−ρ −2 (h α ik − H α δik)(lnρ)kj + C α k (lnρ)kδij − (lnρ)iC α j .<br />

Letting i = j and making summation over i in (4.3), we have<br />

�<br />

C<br />

i<br />

α i,i = −ρ−3∆H α − ρ−3 (hα ik − Hαδik)(lnρ)ki −2(m − 2)ρ−3H α ,i (lnρ)i − (m − 3)ρ−3 (lnρ)i(lnρ)j(hα ij − Hαδij). On the other hand, we have from (1.8) and (1.9)<br />

AijB α ij + B β<br />

ikBβ kjBα ij = −ρ−3 (lnρ)ij(hα ij − Hαδij) + ρ−3 (lnρ)i(lnρ)j(hα ij − Hαδij) +ρ−3 [h β<br />

ikhβkj hαij − Hβh β<br />

ijhα ij − SH α + m| � H| 2H α ].<br />

Putting (4.5) into (2.32), we get<br />

m−1<br />

ρ m+1 {− ρm−2<br />

m−1 [SHα + H β h β<br />

ijhα ij − hα ijh β<br />

ikhβkj − m| � H| 2H α ]<br />

+ρm−2∆H α + m−2<br />

m−1ρm−2 (lnρ),ij(H αδij − hα ij) + 2(m − 2)ρm−2 (lnρ)iH α ,i<br />

+ (m−2)2<br />

m−1 ρm−2 (lnρ)i(lnρ)j(hα ij − Hαδij)} = 0.<br />

By a direct computation we can check the following identity<br />

− 1<br />

m−1 (ρm−2 )ij(mHαδij − hα ij) + ∆(ρm−2H α )<br />

= (m−2)2<br />

m−1 ρm−2 (lnρ)i(lnρ)j(hα ij − Hαδij) + m−2<br />

m−1ρm−2 (lnρ)ij(hα ij − Hαδij) +2(m − 2)ρm−2 (lnρ)iH α ,i + ρm−2∆H α .<br />

(4.3)<br />

(4.4)<br />

(4.5)<br />

(4.6)<br />

(4.7)<br />

Thus (4.6) is equivalent to (4.1) by use of (4.7). We complete the proof of <strong>The</strong>orem 4.1. <strong>The</strong><br />

16


Gauss equations of an m-dimensional submanifold x : M → S n give<br />

Rij = (m − 1)δij + mH β h β<br />

ij − h β<br />

ikhβkj , (4.8)<br />

R = m(m − 1) + m 2 | � H| 2 − S, (4.9)<br />

where Rij and R are the Ricci curvature and scalar curvature of M, respectively, with respect<br />

to the induced metric dx · dx. By use of (4.8) and (4.9) we have,<br />

SH α + H β h β<br />

ijh α ij − h α ijh β<br />

ik hβ<br />

kj − n| � H| 2 H α = (Rij − (m − 1)H β h β<br />

ij)(h α ij − H α δij). (4.10)<br />

Combining (4.1) and (4.10), we prove Corollary 4.1. x : M → S n is an m-dimensional<br />

<strong>Willmore</strong> submanifold in an n-dimensional unit sphere if and only if, <strong>for</strong> any α with m+1 ≤<br />

α ≤ n,<br />

−ρ m−2 (Rij − (m − 1)H β h β<br />

ij)(h α ij − H α δij)<br />

+(m − 1)∆(ρ m−2 H α ) − (ρ m−2 ),ij(mH α δij − h α ij) = 0.<br />

(4.11)<br />

Remark 4.1. When m = 2, Corollary 4.1 implies that x : M 2 → S n is a <strong>Willmore</strong> surface<br />

if and only if<br />

∆H α + h α ijh β<br />

ijH β − 2| � H| 2 H α = 0, 3 ≤ α ≤ n,<br />

which was first proved by Weiner in [18]. Thus all minimal surfaces in S n are <strong>Willmore</strong><br />

surfaces. In particular, the Veronese surfaces in S n are <strong>Willmore</strong> surfaces. In [8], Li-Wang-<br />

Wu gave a Moebius characterization of Veronese surfaces in S n .<br />

Remark 4.2. Fix the index α with m + 1 ≤ α ≤ n, define ✷ α : M → R by<br />

✷ α f = (mH α δij − h α ij)fij,<br />

where f is any smooth function on M. We know that ✷ α is a self-adjoint operator (cf.<br />

Cheng-Yau [4] and Li [7]). This operator naturally appears in the <strong>Willmore</strong> equations (4.1)<br />

or (4.11).<br />

<strong>The</strong>orem 4.2. Let x : M → S n be a m-dimensional minimal submanifold in an ndimensional<br />

unit sphere. If M is an Einstein manifold, then x : M → S n is a <strong>Willmore</strong><br />

submanifold.<br />

Proof. From our assumption<br />

we have, by use of (4.9),<br />

H α = 0, m + 1 ≤ α ≤ n, Rij = R<br />

m δij, (4.12)<br />

ρ 2 = S − m| � H| 2 = S = m(m − 1) − R = constant. (4.13)<br />

Thus from (4.12) and (4.13) we know that (4.11) holds. We conclude that M is a <strong>Willmore</strong><br />

submanifold.<br />

17


Example 4.1. Sm �<br />

2(m+1)<br />

(<br />

S m (<br />

�<br />

m<br />

) → Sm+p with p = 1(m<br />

− 1)(m + 2). Let<br />

2<br />

2(m + 1)<br />

m<br />

) = {(x0, x1, · · · , xm) ∈ R m+1 ; � x 2 i =<br />

2(m + 1)<br />

m<br />

}.<br />

Let E be the space of (m + 1) × (m + 1) symmetric matrices (uij), (i, j = 1, · · · , m), such<br />

that � uii = 0; it is a vector space of dimension 1m(m<br />

+ 3). We define a norm in E by<br />

2<br />

||(uij)|| 2 = � u2 ij. Let Sm+p with p = 1(m<br />

− 1)(m + 2) be the unit hypersphere in E. <strong>The</strong><br />

2<br />

mapping of Sm �<br />

2(m+1)<br />

) into Sm+p , defined by<br />

m<br />

uij = 1<br />

�<br />

m<br />

2 m + 1 (xixj − 2<br />

m δij),<br />

is an isometric minimal immersion. (Actually, this gives an imbedding of the real projective<br />

space of m-dimension into Sm+p ). We know that Sm �<br />

2(m+1)<br />

( ) is an Einstein manifold, thus<br />

m<br />

from <strong>The</strong>orem 4.2 this immersion is a <strong>Willmore</strong> submanifold.<br />

Example 4.2. CP m 2m/(m+1) → Sm(m+2)−1 . We can define a minimal immersion of the<br />

with holomorphic sectional curvature<br />

m-dimensional complex projective space CP m 2m/(m+1)<br />

2m/(m+1) into a unit sphere Sm(m+2)−1 such that the usual coordinate functions of Rm(m+2) are all independent hermitian harmonic functions of degree 1 on CP m 2m/(m+1) (see Wallach<br />

[16]). We also know that CP m 2m/(m+1) is an Einstein manifold with Ricci curvature m. <strong>The</strong>re-<br />

<strong>for</strong>e from <strong>The</strong>orem 4.2 we conclude that this immersion is a <strong>Willmore</strong> submanifold.<br />

Proposition 4.1. Let M = Sm1 (a1) × · · · × Smr (ar) be the submanifold imbedded into<br />

Sm+r−1 , where m1 + · · · + mr = m. <strong>The</strong>n M is a <strong>Willmore</strong> submanifold if and only if<br />

Proof. Consider<br />

ai =<br />

�<br />

m − mi<br />

, i = 1, · · · , r. (4.14)<br />

m(r − 1)<br />

R m+r = R m1+1<br />

r�<br />

mr+1<br />

× · · · × R , m =<br />

i=1<br />

S m1 (a1) × · · · × S mr (ar) = {(a1x1, · · · , arxr) ∈ R m+r : |xi| = 1, i = 1, · · · , r},<br />

where Smi (ai) ⊂ Rmi+1 r�<br />

, (i = 1, · · · , r),<br />

a<br />

i=1<br />

2 i = 1, is a m-dimensional submanifold in Sm+r−1 .<br />

Writing x = (a1x1, · · · , arxr) : M = Sm1 (a1) × · · · × Smr (ar) → Sm+r−1 , x1 · x1 = · · · =<br />

xr · xr = 1. <strong>The</strong>n r − 1 orthogonal normal vector fields of M are<br />

mi,<br />

en+λ = (aλ1x1, · · · , aλrxr), λ = 1, · · · , r − 1<br />

where (aλ1, · · · , aλr) satisfies that (r × r) matrix<br />

⎛<br />

⎜<br />

A = ⎜<br />

⎝<br />

a1<br />

a11<br />

.<br />

· · ·<br />

· · ·<br />

· · ·<br />

ar<br />

a1r<br />

.<br />

ar−11 · · · ar−1r<br />

18<br />

⎞<br />

⎟<br />


is an orthogonal matrix, thus<br />

r−1<br />

�<br />

λ=1<br />

aλiaλj = δij − aiaj. (4.15)<br />

<strong>The</strong>re<strong>for</strong>e the first fundamental <strong>for</strong>m and the <strong>second</strong> fundamental <strong>for</strong>m of M are given by<br />

I = dx · dx = a 2 1dx1 · dx1 + · · · + a 2 rdxr · dxr,<br />

r−1 �<br />

II = − [a1aλ1dx1 · dx1 + · · · + araλrdxr · dxr]em+λ.<br />

λ=1<br />

We get the components of the <strong>second</strong> fundamental <strong>for</strong>m<br />

⎛<br />

(h m+λ<br />

⎜<br />

ij ) = ⎜<br />

⎝<br />

− aλ1<br />

a1 E1<br />

. ..<br />

aλr<br />

ar Er<br />

⎞<br />

⎟<br />

⎠ , λ = 1, · · · , r − 1 (4.16)<br />

where Ei is a (mi × mi) unit matrix. It follows that ρ 2 = m(S − mH 2 )/(m − 1) is constant<br />

and ∇ ⊥ eα = 0, ∆H α = 0, α = m + λ.<br />

From (4.1), we know that M is a <strong>Willmore</strong> submanifold if and only if<br />

SH α + H β h β<br />

ijh α ij − h α ijh β<br />

ik hβ<br />

kj − m| � H| 2 H α = 0, n + 1 ≤ α = m + λ ≤ n + r − 1. (4.17)<br />

Putting<br />

H m+λ = 1<br />

r� aλi<br />

mi,<br />

m i=1<br />

ai<br />

1 ≤ λ ≤ r − 1, S = � 1 − a<br />

mi<br />

i=1<br />

2 i<br />

a2 ,<br />

i<br />

− �<br />

h<br />

µ,i,j,k<br />

m+λ<br />

ij h m+µ<br />

ik hm+µ<br />

� mi<br />

kj =<br />

i a3 aλi(1 − a<br />

i<br />

2 i ), �<br />

h<br />

i,j<br />

m+λ<br />

ij h m+µ<br />

ij = � aλiaµi<br />

mi<br />

i a2 i<br />

into (4.17), by use of (4.15) and (4.16) we can easily check that (4.17) can be reduced to<br />

�<br />

aλi{−<br />

i<br />

mi �<br />

mj(1 −<br />

mai j<br />

1<br />

a2 ) −<br />

j<br />

1 � δij − aiaj<br />

mimj<br />

m j a2 i aj<br />

mi<br />

a3 (1 − a<br />

i<br />

2 i ) + 1<br />

m2 � δjk − ajak<br />

mimjmk } = 0,<br />

j,k<br />

aiajak<br />

where 1 ≤ λ ≤ r − 1. (4.18) is equivalent to<br />

<strong>The</strong>re<strong>for</strong>e (4.19) becomes<br />

−2m +<br />

�<br />

mj(1 −<br />

j<br />

1<br />

a2 ) −<br />

j<br />

1 � δij − aiaj<br />

mimj<br />

m j a2 i aj<br />

mi<br />

a3 (1 − a<br />

i<br />

2 i ) + 1<br />

m2 � δjk − ajak<br />

mimjmk = 0.<br />

j,k<br />

aiajak<br />

− mi<br />

mai<br />

m − mi<br />

a 2 i<br />

+ �<br />

j<br />

mj<br />

a 2 j<br />

(4.18)<br />

(4.19)<br />

(1 + mj<br />

) = 0, i = 1, · · · , r. (4.20)<br />

m<br />

19


We solve (4.20) to get<br />

a 2 i =<br />

We complete the proof of Proposition 4.1.<br />

m − mi<br />

, i = 1, · · · , r.<br />

m(r − 1)<br />

Remark 4.3. When r = 2, Proposition 4.1 reduces to <strong>The</strong>orem 3.1.<br />

Remark 4.4. We also know that S m1 (a1) × · · · × S mr (ar) in S m+r−1 is a m-dimensional<br />

minimal submanifold if and only if<br />

ai =<br />

� mi<br />

m ,<br />

�<br />

mi = m, i = 1, · · · , r.<br />

i<br />

Thus it follows that S m1 (a1) × · · · × S mr (ar) in S m+r−1 is an m-dimensional minimal and<br />

<strong>Willmore</strong> submanifold if and only if<br />

m1 = m2 = · · · = mr = m<br />

r , ai<br />

�<br />

=<br />

1<br />

r .<br />

Acknowledgements. <strong>The</strong> third author would like to thank Professor U. Simon <strong>for</strong> his<br />

hospitality during his research stay at the TU Berlin. We would like to thank the referee <strong>for</strong><br />

his useful suggestions and comments.<br />

References<br />

[1] Blaschke, W.:Vorlesungen uber Differentialgeometrie. Vol.3, Springer Berlin,1929.<br />

[2] Bryant, R.: A duality theorem <strong>for</strong> <strong>Willmore</strong> surfaces, J. Differential Geom. 20(1984),<br />

23-53.<br />

[3] Chen, B.Y.: Some con<strong>for</strong>mal invariants of <strong>submanifolds</strong> and their applications, Bol. Un.<br />

Math. Ital. (4)10 (1974), 380-385.<br />

[4] Cheng, S.Y. and Yau, S.-T.: Hypersurfaces with constant scalar curvature, Math. Ann.<br />

225(1977), 195-204.<br />

[5] Kobayashi, O.: A <strong>Willmore</strong> type problem <strong>for</strong> S 2 × S 2 , Differential geometry and differential<br />

equations (Shanghai, 1985), 67-72, Lecture Notes in Math., 1255, Springer, Berlin,<br />

1987.<br />

[6] Li, H., <strong>Willmore</strong> hypersurfaces in a sphere, To appear in Asian J. of Math., 5(2001).<br />

[7] Li, H., Global rigidity theorems of hypersurfaces, Ark. Mat. 35(1997), 327-351.<br />

[8] Li, H., Wang, C. and Wu, F., A Moebius characterization of Veronese surfaces in S n ,<br />

Math. Ann. 319(2001), 707-714.<br />

20


[9] Li, P. and Yau, S.-T.: A new con<strong>for</strong>mal invariant and its application to <strong>Willmore</strong> conjecture<br />

and the first eigenvalue of compact surface, Invent. Math. 69(1982), 269-291.<br />

[10] Palmer, B.: <strong>The</strong> con<strong>for</strong>mal Gauss map and the stability of <strong>Willmore</strong> surfaces, Ann.<br />

Global Anal. Geom. Vol. 9, No. 3 (1991), 305-317.<br />

[11] Palmer, B.: Second <strong>variational</strong> <strong><strong>for</strong>mula</strong>s <strong>for</strong> <strong>Willmore</strong> surfaces, <strong>The</strong> problem of Plateau,<br />

221-228, World Sci. Publishing, River Edge, NJ, 1992.<br />

[12] Peterson, M.A.: Geometrical methods <strong>for</strong> the elasticity theory of membranes, J. Math.<br />

Physics, 26(4), 1985, 711-717.<br />

[13] Pinkall, U.: Hopf tori in S 3 , Invent. Math. 81(1985), 379-386.<br />

[14] Pinkall, U.: Inequalities of <strong>Willmore</strong> type <strong>for</strong> <strong>submanifolds</strong>, Math. Z. 193 (1986), 241-<br />

246.<br />

[15] Simon, L: Existence of surfaces minimizing the <strong>Willmore</strong> energy, Comm. Analysis<br />

Geometry, vol. 1, n.2 (1993), 281-326.<br />

[16] Wallach, N.R.: Minimal immersions of symmetric spaces into spheres, ”Symmetric<br />

Space”, Ed. Boothby and Weiss, Dekker, New York, 1972, 1-40.<br />

[17] Wang, C.P.: Moebius geometry of <strong>submanifolds</strong> in S n , Manuscripta Math. 96(1998),<br />

517-534.<br />

[18] Weiner, J.L.: On a problem of Chen, <strong>Willmore</strong>,et al., Indiana University Journal,<br />

27(1978), 19-35.<br />

[19] <strong>Willmore</strong>, T.J.: Total curvature in Riemannian geometry. Ellis Horwood Limitd, 1982.<br />

[20] <strong>Willmore</strong>, T.J.: Note on embedded surfaces, An. Sti. Univ. ”Al. I. Cuza” Iasi Sect. I a<br />

Mat., 11(1965), 493–496.<br />

[21] <strong>Willmore</strong>, T.J.: Mean curvature of Riemannian immersions, J. London Math. Soc.,<br />

3(1971), 307–310.<br />

Zhen Guo Haizhong Li<br />

Department of Mathematics Department of Mathematical Sciences<br />

Yunnan Normal University Tsinghua University<br />

Kunming, 650092, P. R. of China Beijing, 100084, P. R. of China<br />

e-mail:guominzhi@ynmail.com e-mail:hli@math.tsinghua.edu.cn<br />

Changping Wang<br />

Department of Mathematics<br />

Peking University<br />

Beijing,100871, P. R.of China<br />

e-mail:wangcp@pku.edu.cn<br />

Eingegangen am 11. April 2001<br />

21

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!