Math. Res. Lett. Volume 24, Number 2, 503–534, 2017

A characterization of Clifford hypersurfaces among embedded constant mean curvature hypersurfaces in a unit sphere Sung-Hong Min and Keomkyo Seo

Let Σ be an n(≥ 3)-dimensional compact embedded hypersurface in a unit sphere with constant mean curvature H ≥ 0 and with two distinct principal curvatures λ and μ of multiplicity n − 1 and 1, respectively. It is known that if λ > μ, there exist many compact embedded constant mean curvature hypersurfaces [26]. In this paper, we prove that if μ > λ, then Σ is congruent to a Clifford hypersurface. The proof is based on the arguments used by Brendle [10].

1

Introduction

504

2

Preliminaries

507

3

4

Simons-type identity for constant mean curvature hypersurfaces

509

First and second order derivatives of the two-point function

516

5

Proof of Main Theorem

521

6

Appendix: The case of H = 0

525

Acknowledgements

531

References

531

2010 Mathematics Subject Classification: 53C40, 53C42. Key words and phrases: Clifford hypersurface, Simons-type identity, constant mean curvature, embedded hypersurface.

503

504

S.-H. Min and K. Seo

1. Introduction Let Σ be an n-dimensional compact embedded hypersurface in an (n + 1)dimensional unit sphere Sn+1 with constant mean curvature H. In case of minimal surfaces in S3 (i.e., n = 2 and H = 0), Brendle [10] ingeniously proved the famous Lawson conjecture which states that the only embedded minimal torus in S3 is the Clifford torus from a sharp estimate for a twopoint function by using the maximum principle. It was observed that the embeddedness condition can be replaced by the weaker assumption that the minimal torus is Alexandrov-immersed in S3 [9]. The technique using the maximum principle for a two-point function was also used by Andrews-Li [5], who gave a complete classification of embedded constant mean curvature tori in S3 . More generally, the proof of Lawson conjecture was extended to a class of embedded Weingarten tori in S3 [11]. Hauswirth-Kilian-Schmidt [15] obtained that every mean-convex Alexandrov embedded constant mean curvature torus in S3 is rotationally symmetric by using integrable systems. It is interesting to find the higher-dimensional analogues of these results. One possible approach to the higher-dimensional problem is to characterize a Clifford hypersurface among embedded constant mean curvature hypersurfaces in Sn+1 . Unfortunately, even when H = 0, it is well-known that there exist infinitely many mutually noncongruent embedded minimal hypersurfaces in Sn+1 which are homeomorphic to the Clifford hypersurface [17]. Recall that an n-dimensional Clifford hypersurface in Sn+1 with constant mean curvature H has two distinct principal curvatures λ and μ of multiplicity n − k and k, respectively. Moreover it is given by

S

n−k



1 √ 1 + λ2



 ×S

k

1

 1 + μ2

 ,

where λ and μ satisfy nH = (n − k)λ + kμ and λμ + 1 = 0. In view of this observation, we restrict ourselves to consider compact embedded constant mean curvature hypersurfaces in a unit sphere with two distinct principal curvatures. Otsuki [22] proved that if the multiplicities of two distinct principal curvatures are greater than 1, then the minimal hypersurface is locally congruent to a Clifford minimal hypersurface. Later, by studying an ordinary differential equation derived from the two distinct principal curvature condition, Otsuki [23, 24] also proved that a compact embedded minimal hypersurface in Sn+1 with two distinct principal curvatures of multiplicity n − 1 and 1, respectively, is congruent to a Clifford minimal

A characterization of Clifford hypersurfaces

505

hypersurface (see also [12]). Therefore he gave the following characterization of Clifford minimal hypersurfaces in Sn+1 . Theorem 1.1 ([22–24]). Let Σ be an n(≥ 3)-dimensional compact embedded minimal hypersurface in Sn+1 with two distinct principal curvatures of multiplicity n − k and k for  1 ≤ k ≤n − 1.  Then Σ is congruent to a Clifford   minimal hypersurface Sn−k

n−k n

× Sk

k n

.

In case of constant mean curvature hypersurfaces in Sn+1 with two distinct principal curvatures, Wei [31] obtained the analogue of Otsuki’s result, provided the multiplicities of two principal curvatures are at least 2, applying a similar argument as in [22]. Theorem 1.2 ([31]). Let Σ be an n(≥ 3)-dimensional hypersurface in Sn+1 with constant mean curvature H and with two distinct principal curvatures λ and μ of multiplicities n − k and k, respectively, 2≤k≤  for  n − 2. Then Σ 1 1 n−k k √ is isometric to a Clifford hypersurface S × S √1+μ , where 2 1+λ2 λ and μ satisfy nH = (n − k)λ + kμ and λμ + 1 = 0. Therefore it suffices to consider constant mean curvature hypersurfaces with two distinct principal curvatures λ and μ, μ being simple (i.e., multipliticity 1). Perdomo [26] obtained the existence of compact embedded constant mean curvature hypersurfaces in Sn+1 other than the totally geodesic n-spheres and Clifford hypersurfaces (see also [12, 27, 33]). Indeed, he constructed such examples by analyzing an ordinary differential equation arising from the two distinct principal curvatures λ and μ satisfying that λ > μ. More precisely, he proved π Theorem 1.3 ([26]). For any integer m ≥ 2 and H between cot m and √ (m2 −2) n−1 n+1 √ , there exists a compact embedded hypersurface in S with conn m2 −1 stant mean curvature H other than the totally geodesic n-spheres and Clifford hypersurfaces.

On the other hand, in the study of n-dimensional constant mean curvature hypersurfaces in Sn+1 with two distinct principal curvatures of multiplicity n − 1 and 1, it mostly requires an additional assumption to obtain a characterization of Clifford hypersurfaces. For instance, Perdomo [25] and Wang [30] independently obtained a curvature integral inequality for minimal hypersurfaces in Sn+1 with two distinct principal curvatures, which

506

S.-H. Min and K. Seo

characterizes a Clifford minimal hypersurface. Later, Wei [32] showed that the similar curvature integral inequality holds for hypersurfaces with the vanishing m-th order mean curvature (i.e., Hm ≡ 0). More precisely, they proved Theorem ([25, 30, 32]). Let M be an n(≥ 3)-dimensional closed hypersurface in Sn+1 with Hm ≡ 0 (1 ≤ m < n) and with two distinct principal curvatures, one of them being simple. Then

n(m2 − 2m + n) |A|2 ≤ Vol(M ), m(n − m) M where equality holds if and only if M is isometric to a Clifford hypersurface  m n−m n−1 1 S × S n n . In [3], Andrews-Huang-Li obtained a uniqueness of Clifford hypersurface among compact embedded Weingarten hypersurfaces in the unit sphere with two distinct principal curvatures satisfying a linear relation between them. Very recently, the authors [21] obtained a more general sharp curvature integral inequality for hypersurfaces in Sn+1 with constant m-th order mean curvature and with two distinct principal curvatures, which generalizes Simons’ integral inequality and gives a characterization of Clifford hypersurfaces in Sn+1 . In contrast to the 2-dimensional problem for embedded constant mean curvature tori, we consider embedded constant mean curvature hypersurfaces with two distinct principal curvatures of multiplicity n − 1 and 1 without assuming any topological restriction. In this paper, we give the following characterization theorem (Theorem 5.3) of Clifford hypersurfaces: Theorem. Let Σ be an n(≥ 3)-dimensional compact embedded hypersurface in Sn+1 with constant mean curvature H ≥ 0 and with two distinct principal curvatures λ and μ, μ beingsimple. If μ >  λ, then Σ is con|λ| 1 n−1 1 √ √ gruent to a Clifford hypersurface S ×S , where λ = 1+λ2 1+λ2 √ 2 2 nH− n H +4(n−1) . 2(n−1) The key ingredients in the proof of our theorem are the following: We first define a suitable two-point function on an embedded constant mean curvature hypersurface based on the non-collapsing argument and compute the first and second order derivatives of the two-point function. This technique was pioneered by Huisken [18] and was developed by Andrews [2].

A characterization of Clifford hypersurfaces

507

Secondly, we obtain a Simons-type identity for constant mean curvature hypersurfaces with two distinct principal curvatures. Indeed, this provides a sufficient condition for constant mean curvature hypersurfaces to attain the equality in Kato’s inequality. Combining with Simons-type identity and adapting the arguments by Brendle [10] with a slight modification finally gives a characterization of Clifford hypersurfaces. We remark that every constant mean curvature torus in S3 has two distinct principal curvatures which implies that there is no umbilic point. (See [20] for minimal tori and [14, 16] for constant mean curvature tori in S3 .) Moreover, constant mean curvature tori in S3 automatically satisfy the condition that μ > λ. Hence our main theorem can be regarded as an extension of the results by Brendle [10] and Andrews-Li [5] to higher-dimensional cases.

2. Preliminaries Let F : Σn → Sn+1 (⊂ Rn+2 ) be a compact embedded constant mean curvature hypersurface in Sn+1 with two distinct principal curvatures, one of them being simple. Let ν(x) be the unit normal vector at x ∈ Σ in Sn+1 . Let h and A be the second fundamental form and the shape operator of Σ, respectively. Note that A is a self-adjoint endomorphism of the tangent space at each point x in Σ such that A(X), Y = h(X, Y ) for all X, Y ∈ Tx Σ. Since Σ has two distinct principal curvatures and one of them is simple, we may assume that λ = λ1 = · · · = λn−1 and μ = λn , where each λi denotes the principal curvature on Σ for 1 ≤ i ≤ n. The normalized mean curvature H is defined by n

1 n−1 1 1 H = tr(h) = λi = λ + μ. n n n n i=1

Since Σ is a compact embedded hypersurface, Σ divides Sn+1 into two connected components. Because the mean curvature of F (Σ) in Sn+1 is constant, we may assume that H ≥ 0 by choosing the suitable orientation of Σ. Let R be the region satisfying that ν points out of R. The mean curvature vector  satisfies that H  = −nHν(x). H 1 For a positive function Ψ on Σ, we denote by BT (x, Ψ(x) ) a ball with 1 radius Ψ(x) which touches Σ at F (x) inside the region R in Sn+1 . Note that our notation BT (x, r) is different from a ball Br (x) centered at x with 1 1 radius r > 0. Then BT (x, Ψ(x) ) is a ball of radius Ψ(x) centered at p(x) =

508 F (x) − (1)

S.-H. Min and K. Seo 1 Ψ(x) ν(x)

in Rn+2 . Define the two-point function Z : Σ × Σ → R by

Z(x, y) := Ψ(x)(1 − F (x), F (y) ) + ν(x), F (y) .

It is easy to check that for any y ∈ Σ, ⎧ 1 ⎪ ⎪ ⎨Z(x, y) > 0 if F (y) ∈ intBT (x, Ψ(x) ), 1 ), Z(x, y) = 0 if F (y) ∈ ∂BT (x, Ψ(x) ⎪ ⎪ ⎩Z(x, y) < 0 if F (y) ∈ BT (x, 1 ), Ψ(x)

since 2 Z(x, y) = |F (y) − p(x)|2 − Ψ(x)



1 Ψ(x)

2 .

We recall the definition of the interior ball curvature at x ∈ Σ, which was originally given by Andrews-Langford-McCoy [4] (see also [5]). Definition 2.1. The interior ball curvature k is a positive function on Σ defined by   1 k(x) := inf : BT (x, r) ∩ Σ = {x}, r > 0 . r Because Σ is compact and embedded in Sn+1 , one can see that the function k is a well-defined positive function on Σ. From the definition of k(x) for every point x ∈ Σ, it follows that k(x)(1 − F (x), F (y) ) + ν(x), F (y) ≥ 0 for all y ∈ Σ. Let Φ(x) := max{λ(x), μ(x)} be the maximum value of the principal curvatures of Σ in Sn+1 at F (x). Note that the two distinct principal curvature condition guarantees that Σ has no umbilic point and hence Φ(x) − H > 0. Motivated by the works of Brendle [10] and Andrews-Li [5], we introduce the constant κ as follows: k(x) − H . x∈Σ Φ(x) − H

κ := sup

For convenience, we will write ϕ(x) := Φ(x) − H.

A characterization of Clifford hypersurfaces

509

Proposition 2.2 (Uniform boundedness of κ). Let Σ be a compact embedded constant mean curvature hypersurface with two distinct principal curvatures in Sn+1 . Then there exists a constant K > 0 satisfying 1 ≤ κ < K. Proof. By definition, one sees that ϕ > 0. Because Σ is compact, ϕ is uniformly bounded and k is uniformly bounded above. From the definition of k, it immediately follows that k(x) ≥ Φ(x) for all x ∈ Σ, which gives the conclusion.  Define a positive function Ψ(x) := κϕ(x) + H = κ(Φ(x) − H) + H on Σ. Then Ψ(x) ≥ k(x). It follows that (2)

Z(x, y) = Ψ(x)(1 − F (x), F (y) ) + ν(x), F (y) ≥ 0

for all (x, y) ∈ Σ × Σ. Therefore if there exists a point (x, y) ∈ Σ × Σ satisfying that Z(x, y) = 0, then ∂Z ∂Z (x, y) = (x, y) = 0, ∂xi ∂yi since the function Z attains its global minimum at (x, y). Note that the global minimum of the function Z is attained at (x, x) ∈ Σ × Σ for all x ∈ Σ. Furthermore, one can see that there exists a point (x, y) ∈ Σ × Σ satisfying that Z(x, y) = 0 and x = y by making use of the compactness of Σ and the property of interior ball curvature derived from the embeddedness of Σ (see Lemma 5.1).

3. Simons-type identity for constant mean curvature hypersurfaces Let Σ be an n(≥ 3)-dimensional compact embedded constant mean curvature hypersurface in Sn+1 with two distinct principal curvatures. The traceless part of the second fundamental form h is defined to be a differential 2-form η on Σ with the coefficient function ηij in local coordinates as follows: ηij := hij − δij H, where δij is the Kronecker delta. The corresponding traceless shape operator ˚ is defined by A   ˚ A(X), Y = η(X, Y )

510

S.-H. Min and K. Seo

for all X, Y ∈ Tx Σ, where Tx Σ denotes the tangent space of Σ at x ∈ Σ. In 1970, Otsuki [22] observed the following interesting property of the eigenspace of principal curvatures. Theorem 3.1 ([22]). Let Σ be a hypersurface immersed in an (n + 1)dimensional Riemannian manifold of constant curvature such that the multiplicities of principal curvatures are all constant. Then we have the following: • The distribution of the space of principal vectors corresponding to each principal curvature is completely integrable. • If the multiplicity of a principal curvature is greater than 1, then this principal curvature is constant on each integral submanifold of the corresponding distribution of the space of principal vectors. Let (x1 , . . . , xn ) be the geodesic normal coordinates at x ∈ Σ (i.e. the metric tensor is given by gij = δij and the Christoffel symbol Γkij (x) at x vanishes). We may assume that hij = λi δij with λ = λ1 = · · · = λn−1 and μ = λn . We will denote the coefficient function of the covariant derivative ∇Σ h by hijk . Then ∂hij hijk (x) = (x) ∂xk at x ∈ Σ. As a consequence of Theorem 3.1, one can compute ηijk for 1 ≤ i, j, k ≤ n. Lemma 3.2. Let Σ be a constant mean curvature hypersurface in Sn+1 with two distinct principal curvatures λ and μ, μ being simple: λ = λ1 = · · · = λn−1 and μ = λn . Then for all 1 ≤ i, j, k ≤ n, we have ηijk = hijk and

⎧ ⎪ ⎨ηijk = 0 ηiik = 0 ⎪ ⎩ ηnnn = −(n − 1)ηiin

if i, j, k are all distinct, if k = n, for i = 1, . . . , n − 1.

Proof. One can easily see that ηijk = hijk for 1 ≤ i, j, k ≤ n on Σ. If i, j, k are all distinct, then at least two of them are contained in the set {1, . . . , n − 1}. Using the Codazzi equations, we may assume that i and j are in the set

A characterization of Clifford hypersurfaces {1, . . . , n − 1}. Since hijk =

∂hij ∂xk

ηijk =

511

at x, the first part of Theorem 3.1 implies ∂hij ∂λi = δij = 0 ∂xk ∂xk

at x ∈ Σ. To check the last two equalities, we let i, j ∈ {1, . . . , n − 1}. Then hiik = hjjk for any k ∈ {1, . . . , n}, and hiij = 0 are direct consequences of the second part of Theorem n−1 3.1. The constant mean curvature assumption implies that hnnk = − i=1 hiik . Hence the conclusion immediately follows.  When a constant mean curvature hypersurface has two distinct principal curvatures, we first prove the following useful identity. Proposition 3.3. Let Σ be a constant mean curvature hypersurface in Sn+1 ˚ is with two distinct principal curvatures λ and μ, μ being simple. Then |A| strictly positive and ˚2 = |∇Σ A|

(3)

n+2 Σ ˚ 2 |∇ |A|| . n

Remark 3.4. It is well-known that a constant mean curvature hypersurface Σ in space forms satisfies ˚ 2 − |∇Σ |A|| ˚ 2 ≥ 2 |∇Σ |A|| ˚ 2, |∇Σ A| n

(4)

which is so-called Kato’s inequality [6, 19, 28, 34]. It would be interesting to characterize the equality case. Proposition 3.3 gives a sufficient condition for Kato’s inequality (4) to attain the equality. Proof of Proposition 3.3. Since Σ has two distinct principal curvatures λ ˚ and μ, the functions λ − H and μ − H never vanish. Thus the function |A| is strictly positive. For x ∈ Σ, we choose the geodesic normal coordinates at x as above. Then we have (5)

Σ ˚2

|∇ A| =

n

i,j,k=1

2 ηijk

=

n

i=1

2 ηiii

+3

n

2 ηiik

i,k=1 i=k

+

n

i,j,k=1 i,j,k distinct

2 2 + 3(n − 1)η11n = ηnnn 2 2 = (n − 1)2 η11n + 3(n − 1)η11n 2 = (n − 1)(n + 2)η11n ,

2 ηijk

512

S.-H. Min and K. Seo

 where we used the relations ηnnn = − n−1 i=1 ηiin = −(n − 1)η11n in the sec˚ = ∇Σ |A| ˚ 2, ˚ ∇Σ |A| ond and third equality. Since 2|A|

(6)

1 ˚ 2 |2 |∇Σ |A| ˚2 4|A| n 1 = ηii ηiik ηjj ηjjk ˚2 |A| i,j,k=1

˚ 2= |∇Σ |A||

n n 1 1 = ηii ηiik ηjj ηjjk + ηii ηiin ηjj ηjjn . ˚2 ˚2 |A| |A| i,j=1 i,j,k=1 k=n

Using Lemma 3.2, one sees that the first term of the right hand side of the identity (6) vanishes. Moreover n

ηii = (n − 1)η11 + ηnn = 0.

i=1

Therefore the second term of the right hand side of the identity (6) can be written as n 1 ηii ηiin ηjj ηjjn ˚2 |A| i,j=1 n−1

n−1 2 1 ηii ηiin ηjj ηjjn + ηii ηiin ηnn ηnnn + ηnn ηnnn ηnn ηnnn 2 ˚ ˚ |A| i=1 |A|2 i,j=1 1 2 2 2 2 2 2 = (n − 1)2 η11 η11n + 2(n − 1)3 η11 η11n + (n − 1)4 η11 η11n 2 ˚ |A|

1 = ˚ |A|2

=

n2 (n − 1)2 2 2 η11 η11n . ˚2 |A|

From the fact that ˚ 2 = (n − 1)η 2 + (n − 1)2 η 2 = n(n − 1)η 2 > 0, |A| 11 11 11 we finally obtain (7)

2 ˚ 2 = n(n − 1)η11n |∇Σ |A|| .

A characterization of Clifford hypersurfaces

513

Hence combining the equations (5) and (7), 2 ˚ 2 = (n − 1)(n + 2)η11n = |∇Σ A|

n+2 Σ ˚ 2 |∇ |A|| , n



which completes the proof.

The following second order partial differential equation of the second fundamental form of a minimal hypersurface in Sn+1 was established by Simons [29]. ΔΣ |A|2 − 2|∇Σ A|2 + 2(|A|2 − n)|A|2 = 0. More generally one can obtain the analogue of the above equation by Simons for a constant mean curvature hypersurface Σ in a Riemannian manifold. The Gauss equations and the Ricci formulas state that Rijkl = (δik δjl − δil δjk ) + (hik hjl − hil hjk ), n n

hrj Rrikl + hir Rrjkl , hijkl = hijlk + r=1

r=1

where Rijkl denotes the components of the Riemann curvature tensor of Σ. The Laplacian of h can be computed by making use of the Codazzi equations as follows: ΔΣ hij = =

=

n

k=1 n

k=1 n

hijkk = hkikj + hkikj +

k=1

+

=

n

n

hkijk

k=1 n

k,r=1 n

hri Rrkjk +

n

hri (δrj δkk − δrk δkj + hrj hkk − hrk hkj )

k,r=1

hkr (δrj δik − δrk δij + hrj hik − hrk hij )

k,r=1 n

n

k=1

r=1

hkkij + nhij − hij + nH



n

=

k=1

hri hrj n

hir hrk hkj + hij − nHδij +

k,r=1

n

hkr Rrijk

k,r=1

hkkij + (n − |A|2 )hij + nH

hik hkr hrj − |A|2 hij

k,r=1

n

r=1

hri hrj − nHδij .

514

S.-H. Min and K. Seo

Therefore we have ΔΣ hij = (n − |A| )hij + nHij − nHδij + nH 2

(8)

n

hik hkj .

k=1

Note that the the above equation (8) holds for any hypersurface Σ in Sn+1 . In the following we have second-order elliptic partial differential equation on the trace-less second fundamental form, which was obtained by Al´ıasde Almeida-Brasil [1]. For completeness we give the proof which is slightly different from their proof. Proposition 3.5 ([1]). Let Σ be a constant mean curvature hypersurface in Sn+1 with two distinct principal curvatures λ and μ, μ being simple. Then Σ ˚ 2 ˚ − 2 |∇ |A|| + (|A|2 − n)|A| ˚ ΔΣ |A| ˚ n |A|

n(n − 2) ˚ 2 = 0. ˚ + sgn(λ − μ)  H|A| − 2nH 2 |A| n(n − 1)

Proof. Using the equation (8), we get n

=

ηij ΔΣ ηij =

i,j=1 n

n

ηij ΔΣ hij

i,j=1

2 (n − |A|2 )ηij + nH

i,j=1

= (n − |A| ) 2

n

i=1

n

ηij (ηik ηkj + Hηik δkj + Hδik ηkj )

i,j,k=1

ηii2

+ nH

n

ηii3

+ 2nH

i=1

2

n

ηii2 .

i=1

˚ 2 , and the right hand side is ˚ 2 − |∇Σ A| The left hand side is equal to 12 ΔΣ |A| n 2 2 2 2 ˚ ˚ equal to (n − |A| )|A| + 2nH |A| + nH i=1 ηii3 . We also see that n

3 3 ηii3 = (n − 1)η11 − (n − 1)3 η11

i=1 3 = −n(n − 1)(n − 2)η11 = −sgn(λ − μ) 

n−2 n(n − 1)

˚ 3, |A|

A characterization of Clifford hypersurfaces ˚ 3 = (|A| ˚ 2 ) 32 = sgn(λ − μ) since |A| lowing Simons-type identity:



515

3

3 . Therefore we have the foln(n − 1) η11

˚ 2 + 2(|A|2 − n)|A| ˚ 2 − 2|∇Σ A| ˚2 ΔΣ |A| 2n(n − 2) ˚ 3 = 0. ˚ 2 + sgn(λ − μ)  H|A| − 4nH 2 |A| n(n − 1) ˚ 2 = 2|A|Δ ˚ Σ |A| ˚ + 2|∇Σ |A|| ˚ 2, Since ΔΣ |A| ˚+ ΔΣ |A|

˚2 ˚ 2 |∇Σ A| |∇Σ |A|| − ˚ ˚ |A| |A|

n(n − 2) ˚ 2 = 0. ˚ − 2nH 2 |A| ˚ + sgn(λ − μ)  H|A| + (|A|2 − n)|A| n(n − 1) Therefore applying the equation (3) gives the conclusion.



Applying Proposition 3.5 to the function ϕ = Φ − H, where Φ is the maximum value of the principal curvatures, we get the following: Corollary 3.6. Let Σ be a constant mean curvature hypersurface in Sn+1 with two distinct principal curvatures λ and μ, μ being simple. Then ΔΣ ϕ −

2 |∇Σ ϕ|2 + (|A|2 − n)ϕ − 2nH 2 ϕ + sgn(λ − μ)nf (n)Hϕ2 = 0, n ϕ

where the function f (n) is defined by  f (n) :=

n−2 n−1

n−2

if Φ = μ, if Φ = λ.

˚2 = ˚ 2 = n ϕ2 and if Φ = λ, then |A| Proof. Note that if Φ = μ, then |A| n−1 2 n(n − 1)ϕ . The conclusion follows from Proposition 3.5 and the linearity of ΔΣ and ∇Σ .  For later use, we define a constant g(n) depending on the dimension n as follows:  1 if Φ = μ, g(n) = n−1 n − 1 if Φ = λ. ˚ 2 = ng(n)ϕ2 . Then one can write |A|

516

S.-H. Min and K. Seo

4. First and second order derivatives of the two-point function Let Σ be an n(≥ 3)-dimensional compact embedded constant mean curvature hypersurface in Sn+1 with two distinct principal curvatures λ and μ of the multiplicity n − 1 and 1, respectively. Consider a pair of points (x, y) ∈ Σ × Σ such that Z(x, y) = 0. Then by the equation (2) ∂Z ∂Z (x, y) = (x, y) = 0. ∂xi ∂yi Let us choose geodesic normal coordinates (x1 , . . . , xn ) at x in Σ satisfying that hij = λi δij with λ = λ1 = · · · = λn−1 and μ = λn and geodesic normal coordinates (y1 , . . . , yn ) at y in Σ. Therefore the first order derivatives of the function Z(x, y) are given by   ∂Ψ(x) ∂F (x) ∂Z (9) (x, y) = (1 − F (x), F (y) − Ψ(x) , F (y) 0= ∂xi ∂xi ∂xi   n

∂F (x) k + hi (x) , F (y) , ∂xk k=1

and (10)

    ∂Z ∂F (y) ∂F (y) 0= (x, y) = −Ψ(x) F (x), + ν(x), . ∂yi ∂xi ∂xi

In this section, using these relations in geodesic normal coordinates as above, we are able to compute the second order derivatives of the function Z. Proposition 4.1. At the point (x, y) ∈ Σ × Σ satisfying that Z(x, y) = 0 and x = y, we have  n n 2

| ∂Ψ(x) ∂2Z ∂xi | (x, y) = Δ Ψ(x) − 2 Σ ∂xi 2 Ψ(x) − λi (x) i=1 i=1  2 + |A(x)| − n Ψ(x) (1 − F (x), F (y) ) + nΨ(x) + nHΨ(x) ν(x), F (y) − nH F (x), F (y) .

A characterization of Clifford hypersurfaces Proof. Differentiating the equation (9) in the direction

∂ ∂xi

517

gives

  n n

∂2Z ∂Ψ(x) ∂F (x) (x, y) = ΔΣ Ψ(x) (1 − F (x), F (y) ) − 2 , F (y) ∂xi 2 ∂xi ∂xi i=1 i=1   n n

∂hi k (x) ∂F (x) − Ψ(x) ΔΣ F (x), F (y) + , F (y) ∂xi ∂xk +

n

k=1 i=1

hi k (x) −hik (x)ν(x) − δik F (x), F (y) .

i,k=1

Since F : Σ → Sn+1 ⊂ Rn+2 is a constant mean curvature hypersurface, ΔΣ F (x) + nF (x) = −nHν(x). By using the Codazzi equations, n n n n

∂hi k (x) Σ k = ∇ ∂ hi (x) = hiki (x) = hiik (x) = 0 ∂xi ∂xi i=1

i=1

i=1

i=1

at x. Thus

  n n

∂2Z ∂Ψ(x) ∂F (x) (x, y) = ΔΣ Ψ(x) (1 − F (x), F (y) ) − 2 , F (y) ∂xi 2 ∂xi ∂xi i=1

i=1

+ nΨ(x) F (x), F (y) + nΨ(x)H ν(x), F (y)

− |A(x)|2 ν(x), F (y) − nH F (x), F (y) . Rearranging the above formula by using the equation (1) with Z(x, y) = 0 yields n

∂2Z (11) (x, y) = ΔΣ Ψ(x) + |A(x)|2 − n Ψ(x) (1 − F (x), F (y) ) 2 ∂xi i=1   n

∂Ψ(x) ∂F (x) + nΨ(x) − 2 , F (y) ∂xi ∂xi i=1

+ nHΨ(x) ν(x), F (y) − nH F (x), F (y) . Using the formula (9) with

 (12)

∂Z ∂xi (x, y)

∂F (x) , F (y) ∂xi



=

= 0, we have

(1 − F (x), F (y) ) ∂Ψ(x) , Ψ(x) − λi (x) ∂xi

518

S.-H. Min and K. Seo

Putting the equation (12) in the equation (11), we get the conclusion.



Let wi (x, y) be the reflection of the vector ∂F∂x(x) in Rn+2 with respect to i the hyperplane orthogonal to F (x) − F (y) and passing through the origin. The vector wi (x, y) is given by ∂F (x) wi (x, y) = −2 ∂xi



∂F (x) F (x) − F (y) , ∂xi |F (x) − F (y)|



F (x) − F (y) . |F (x) − F (y)|

We remark that {w1 (x, y), . . . , wn (x, y)} is the set of mutually orthogonal unit tangent vectors in TF (y) Sn+1 . On the other hand, the following three properties hold at (x, y) for 1 ≤ i ≤ n.   ∂Z ∂F (y) • , Ψ(x)F (x) − ν(x) = − (x, y) = 0, ∂yi ∂yi   ∂F (x) , F (y) ∂xi • wi (x, y), Ψ(x)F (x) − ν(x) = Z(x, y) = 0, 1 − F (x), F (y)

• |F (y)|2 |Ψ(x)F (x) − ν(x)|2 − F (y), Ψ(x)F (x) − ν(x) 2 = (1 + Ψ(x)2 ) − Ψ(x)2 = 1 = 0.  (y) ∂F (y) Thus one sees that Span ∂F , . . . , = Span (w1 (x, y), . . . , wn (x, y)). ∂y1 ∂yn Moreover, if we choose the coordinates at y satisfying that for 1 ≤ i = j ≤ n

  ∂F (y) wi (x, y), ≥ 0 and ∂yi

  ∂F (y) wi (x, y), = 0, ∂yj

then the above three properties implies that (13)

wi (x, y) =

∂F (y) . ∂yi

Equipped with the local coordinates chosen as above, we are able to get the following second order derivatives at the global minimum points of the two-point function Z. Proposition 4.2. At the point (x, y) ∈ Σ × Σ satisfying that Z(x, y) = 0 and x = y, we have ∂2Z (x, y) = λi (x) − Ψ(x). ∂xi ∂yi

A characterization of Clifford hypersurfaces Proof. Differentiating the equation (9) in the direction

∂ ∂yi

519

gives

    ∂2Z ∂Ψ(x) ∂F (y) ∂F (x) ∂F (y) (x, y) = − F (x), − Ψ(x) , ∂xi ∂yi ∂xi ∂yi ∂xi ∂yi   n

∂F (x) ∂F (y) + hik (x) , ∂xk ∂yi k=1

=

1 1 − F (x), F (y)

     ∂F (y) ∂F (x) × (λi (x) − Ψ(x)) , F (y) F (x), ∂xi ∂yi   ∂F (x) ∂F (y) + (λi (x) − Ψ(x)) , . ∂xi ∂yi

∂Z Here the second equality follows from the equation (9) with ∂x (x, y) = 0. i Moreover it can be expressed in terms of wi (x, y) as follows:   ∂F (x) F (x) − F (y) ∂2Z (x, y) = −2 (λi (x) − Ψ(x)) , ∂xi ∂yi ∂xi |F (x) − F (y)|   F (x) − F (y) ∂F (y) × , |F (x) − F (y)| ∂yi   ∂F (x) ∂F (y) + (λi (x) − Ψ(x)) , ∂xi ∂yi   ∂F (y) = (λi (x) − Ψ(x)) wi (x, y), . ∂yi

Plugging the equation (13) into the above identity gives the conclusion.



Proposition 4.3. At the point (x, y) ∈ Σ × Σ satisfying that Z(x, y) = 0 and x = y, we have n

∂2Z i=1

∂yi 2

(x, y) = nΨ(x) + nHΨ(x) F (x), ν(y) − nH ν(x), ν(y) .

Proof. Differentiating the equation (10) in the direction n

∂2Z i=1

∂yi 2

∂ ∂yi

gives

(x, y) = −Ψ(x) F (x), ΔΣ F (y) + ν(x), ΔΣ F (y)

= nΨ(x) F (x), F (y) + nHΨ(x) F (x), ν(y)

− n ν(x), F (y) − nH ν(x), ν(y)

= nΨ(x) + nHΨ(x) F (x), ν(y) − nH ν(x), ν(y) .



520

S.-H. Min and K. Seo

Proposition 4.4. For (x, y) ∈ Σ × Σ satisfying that Z(x, y) = 0 and x = y, n n n

∂2Z ∂2Z ∂2Z + 2 + ∂xi 2 ∂xi ∂yi ∂yi 2 i=1

i=1

i=1

= (1 − F (x), F (y) )   n 2

| ∂Ψ(x) ∂xi | 2 2 × ΔΣ Ψ(x) − 2 + |A(x)| − n Ψ(x) − nHΨ(x) + nH . Ψ(x) − λi (x) i=1

Proof. Applying Proposition 4.1, Proposition 4.2, and Proposition 4.3, we have (14)

n n n

∂2Z ∂2Z ∂2Z + 2 + ∂xi 2 ∂xi ∂yi ∂yi 2 i=1

i=1

i=1

= (1 − F (x), F (y) )  n

× ΔΣ Ψ(x) − 2 i=1

2 | ∂Ψ(x) ∂xi |

Ψ(x) − λi (x)







+ |A(x)|2 − n Ψ(x)

+ 2nH + nHΨ(x) ν(x), F (y) − nH F (x), F (y)

+ nHΨ(x) F (x), ν(y) − nH ν(x), ν(y) . In order to get the conclusion, we need the following computations: • ν(x), F (y) = −Ψ(x) (1 − F (x), F (y) ), 1 Since BT (x, Ψ(x) ) touches Σ at F (x) and F (y) simultaneously, the center 1 p(x) of the geodesic ball BT (x, Ψ(x) ) is given by

p(x) = F (x) −

1 1 ν(x) = F (y) − ν(y), Ψ(x) Ψ(x)

which gives ν(y) = ν(x) + Ψ(x)(F (y) − F (x)). Thus • F (x), ν(y) = F (x), ν(x) + Ψ(x)(F (y) − F (x))

= −Ψ(x) (1 − F (x), F (y) ). Moreover • ν(x), ν(y) = ν(x), ν(x) + Ψ(x)(F (y) − F (x))

= 1 + Ψ(x) ν(x), F (y) = 1 − Ψ(x)2 (1 − F (x), F (y) ), • F (x), F (y) = 1 − (1 − F (x), F (y) ).

A characterization of Clifford hypersurfaces

521

Combining these computations with the equation (14), we get the conclu sion. Since Φ(x) ≤ k(x) ≤ Ψ(x), one sees that for 1 ≤ i ≤ n ⎛ ⎞

Ψ(x) − λi = Ψ(x) − ⎝nH − λj ⎠ = Ψ(x) +

j=i

λj − nH ≤ n(Ψ(x) − H).

j=i

We remark that Ψ(x) − λj < n(Ψ(x) − H) for some 1 ≤ j ≤ n because F (Σ) has two distinct principal curvatures. As a consequence of Proposition 4.4, we have the following: Corollary 4.5. For (x, y) ∈ Σ × Σ satisfying that Z(x, y) = 0 and x = y, n n n

∂2Z ∂2Z ∂2Z + 2 + ∂xi 2 ∂xi ∂yi ∂yi 2 i=1

i=1

i=1

≤ (1 − F (x), F (y) )   |∇Σ Ψ(x)|2 2 2 × ΔΣ Ψ(x) − + |A(x)| − n Ψ(x) − nHΨ(x) + nH . Ψ(x) − H Moreover, if κ > 1, equality holds only when ∇Σ Ψ(x) = 0.

5. Proof of Main Theorem We begin with showing the existence of a global minimum point (x, y) ∈ Σ × Σ of the function Z which is not contained in the diagonal D = {(x, x) : x ∈ Σ} ⊂ Σ × Σ when κ > 1. Lemma 5.1. If κ > 1, then there exists a point (x, y) ∈ Σ × Σ \ D such that Z(x, y) = 0, where D is the diagonal. Proof. Since Σ is compact, κ is attained at some point x ∈ Σ. Thus (15)

Ψ(x) = κϕ(x) + H = k(x).

By the definition of the interior ball curvature k(x) at x ∈ Σ, there exists a 1 point y ∈ Σ satisfying that y ∈ BT (x, k(x) ) ∩ Σ \ {x}. This is equivalent to that there exists a point y ∈ Σ \ {x} such that Z(x, y) = 0, which follows from the definition of the function Z and the equation (15). 

522

S.-H. Min and K. Seo

By definition of the interior ball curvature k(x), it holds Φ(x) ≤ k(x) for every x ∈ Σ, in general. However the following proposition shows that if Σ has a constant mean curvature H > 0 and two distinct principal curvatures of multiplicity n − 1 and 1, then k(x) = Φ(x) for every x ∈ Σ. Proposition 5.2. Let Σ be an n(≥ 3)-dimensional compact embedded hypersurface in Sn+1 with constant mean curvature H with two distinct principal curvatures, one of them being simple. If H > 0. Then the interior ball curvature k(x) is the same as the maximum principal curvature Φ(x) for all x ∈ Σ. Proof. Suppose that κ > 1. By Lemma 5.1, there exists a point (x, y) ∈ Σ × Σ with x = y satisfying that Z(x, y) = 0. Using Corollary 3.6 and Corollary 4.5 together with Ψ(x) = κϕ(x) + H, we get  n  n n

∂2Z

∂2Z ∂2Z 1 (16) +2 + (1 − F (x), F (y) ) ∂xi 2 ∂xi ∂yi ∂yi 2 i=1

i=1

i=1

2κ |∇Σ ϕ(x)|2 ≤ κΔΣ ϕ(x) − + |A(x)|2 − n (κϕ(x) + H) n ϕ(x) − nH(κϕ(x) + H)2 + nH = H|A(x)|2 − κ2 nHϕ(x)2 − nH 3 − sgn(λ − μ)κnf (n)Hϕ(x)2 , where f (n) =

n−2 n−1

if Φ = μ, and f (n) = n − 2 if Φ = λ. Using the relation ˚ 2 = |A|2 − nH 2 = ng(n)ϕ2 , |A|

we get 1 (1 − F (x), F (y) )



n n n

∂2Z ∂2Z ∂2Z + 2 + ∂xi 2 ∂xi ∂yi ∂yi 2 i=1

i=1



i=1

≤ −nHϕ(x)2 (κ2 + sgn(λ − μ)f (n)κ − g(n)) < −nHϕ(x)2 (1 + sgn(λ − μ)f (n) − g(n)) ≤0 where we used the identity 1 + sgn(λ − μ)f (n) − g(n) = 0. However, since the point (x, y) ∈ Σ × Σ \ D is a global minimum point of the function Z, we see n n n

∂2Z ∂2Z ∂2Z 0≤ + 2 + , ∂xi 2 ∂xi ∂yi ∂yi 2 i=1

i=1

i=1

A characterization of Clifford hypersurfaces

523

which is a contradiction. From Proposition 2.2, it follows that k(x) = Φ(x) = Ψ(x)



for all x ∈ Σ. We are now ready to prove our main theorem.

Theorem 5.3. Let Σ be an n(≥ 3)-dimensional compact embedded hypersurface in Sn+1 with constant mean curvature H ≥ 0 and with two distinct principal curvatures λ and μ, μ beingsimple. If μ >  λ, then Σ is con|λ| 1 n−1 1 √ √ gruent to a Clifford hypersurface S ×S , where λ = 1+λ2 1+λ2 √ 2 2 nH− n H +4(n−1) . 2(n−1) Proof. If H = 0, then Σ is congruent to a Clifford minimal hypersurfaces from Theorem 1.1 by Otsuki. Thus it suffices to consider the case of H > 0. Since μ > λ, we have Φ = μ. From Proposition 5.2, we have Φ(x)(1 − F (x), F (y) ) + ν(x), F (y) ≥ 0, for all x, y ∈ Σ. Fix x ∈ Σ and choose an orthonormal frame {e1 , . . . , en } in a neighborhood of x such that h(en , en ) = Φ. Let γ(t) be a geodesic on Σ such that γ(0) = F (x) and γ  (0) = en . For simplicity, let us identify the hypersurface Σ with its image under the embedding F , so that F (x) = x. Define a function f : R → R by f (t) := Z(F (x), γ(t)) = Φ(x)(1 − F (x), γ(t) ) + ν(x), γ(t) . Then, by definition, f (t) ≥ 0 and f (0) = 0. A simple computation shows

  f  (t) = − Φ(x)F (x) − ν(x), γ  (t) ,   f  (t) = Φ(x)F (x) − ν(x), γ(t) + h(γ  (t), γ  (t))ν(γ(t)) ,    f  (t) = Φ(x)F (x) − ν(x), γ  (t) + (∇Σ γ  (t) h)(γ (t), γ (t))ν(γ(t))  + h(γ  (t), γ  (t))∇γ  (t) ν(γ(t)) , where ∇ is the covariant derivative of Rn+2 . In particular, it follows that f (0) = f  (0) = 0,

524

S.-H. Min and K. Seo f  (0) = Φ(x)F (x) − ν(x), F (x) + Φ(x)ν(x) = 0.

Moreover the fact that f ≥ 0 implies that f  (0) = 0. Hence 0 = f  (0) = Φ(x)F (x) − ν(x), en + hnnn (x)ν(x) = −hnnn (x), since ∇γ  (t) ν(γ(t)) is tangent to Σ. Therefore we get en λ = h11n = 1 hnnn = 0. Combining this with Lemma 3.2, one sees that λ and μ are − n−1 constant on Σ, which implies that Σ is an isoparametric hypersurface in Sn+1 with two distinct principal curvatures. From the classification of isoparametric hypersurfaces with two principal curvatures due to  Cartan [13],it follows 1 n−1 1 √ 1 √ × S , that Σ is congruent to the Riemannian product S 2 2 1+μ 1+λ where λ and μ satisfy nH = (n − 1)λ + μ. We now claim that λμ + 1 = 0 on Σ. To see this, let {e1 , . . . , en , en+1 } be a local orthonormal frame around p ∈ Σ such that en+1 is normal to Σ and hij = λi δij at p. Let {ω1 , . . . , ωn , ωn+1 } be a dual coframe. We use the following convention on the ranges of indices: 1 ≤ A, B, C, . . . ≤ n + 1 and 1 ≤ i, j, k, . . . ≤ n. Then the structure equations of a unit sphere Sn+1 are given by dωA = −

(17)

dωAB = − ΩAB =

1 2

n+1

B=1 n+1

ωAB ∧ ωB ,

ωAB + ωBA = 0,

ωAC ∧ ωCB + ΩAB ,

C=1 n+1

KABCD ωC ∧ ωD ,

C,D=1

KABCD = δAC δBD − δAD δBC . We restrict these forms to Σ. Then we have ωn+1 = 0 on Σ. Moreover 0 = dωn+1 = −

n

ωn+1,i ∧ ωi

i=1

Recall that hijk is defined by

and

ωn+1,i =

n

j=1

hij ωj =

n

i=1

λi ω i .

A characterization of Clifford hypersurfaces

n

hijk ωk = dhij −

k=1

n

hik ωkj −

k=1

n

525

hkj ωki .

k=1

Let θij := (λi − λj )ωij = θji . Then we have n

hijk ωk = δij dλj − (λi − λj )ωij = δij dλj − θij .

k=1

Since each λi is constant on Σ, Lemma 3.2 shows that θin = δin dλn −

n

hink ωk = −

k=1

n−1

k=1

for 1 ≤ i ≤ n − 1. This implies that ωin = n − 1, using the equation (17) gives 0 = dωin = −

n

hink ωk − hinn ωn = 0

θin λ−μ

= 0. Therefore, for 1 ≤ i ≤

ωik ∧ ωkn − ωi,n+1 ∧ ωn+1,n + ωi ∧ ωn = (λμ + 1) ωi ∧ ωn ,

k=1

which shows thatλμ + 1 = 0 onΣ. Therefore Σ is congruent √ to a Clifford hy persurface Sn−1 μ > λ.

√ 1 1+λ2

× S1



|λ| 1+λ2

, where λ =

nH−

n2 H 2 +4(n−1) 2(n−1)

since 

6. Appendix: The case of H = 0 Our proof of Theorem 5.3 still works for the case of H = 0. Although Otsuki gave a classification theorem for embedded minimal hypersurfaces with two distinct principal curvatures in Theorem 1.1, we here give another proof of Theorem 1.1. If H = 0, then μ = −(n − 1)λ. Therefore, by choosing a suitable orientation, we have μ(x) > λ(x) for every x ∈ Σ. The proof is divided into two cases: κ = 1 and κ > 1. The proof in case of κ = 1 is similar to that of Theorem 5.3 with Φ = μ. For this reason, it suffices to consider the case of κ > 1. The proof uses basically Brendle’s argument [10]. Proposition 6.1. Let (x, y) ∈ Σ × Σ \ D such that Z(x, y) = 0. Then ∇Σ Φ(x) = 0.

526

S.-H. Min and K. Seo

Proof. Since the function Z attains its global minimum at (x, y), it follows from the inequality (16) and H = 0 that 0≤

n n n

∂2Z ∂2Z ∂2Z + 2 + ∂xi 2 ∂xi ∂yi ∂yi 2 i=1

i=1 2

i=1

= H|A(x)| − κ nHϕ(x) − nH 3 − sgn(λ − μ)κnf (n)Hϕ(x)2 2

2

= 0. Therefore equality holds in Corollary 4.5, which implies that ∇Σ Ψ(x) = 0. Hence we obtain that ∇Σ Φ(x) = 0.  For a point (x, y) such that Z(x, y) = 0, we choose open neighborhoods U1 and U2 of x and y, respectively, such that U1 × U2 ∩ D = ∅. Then there exist a constant Λ1 > 0 depending on x and y such that sup {|∇Σ Ψ|, |∇Σ F |, |A|2 } < Λ1 ,

(18)

U1 ×U2

and (19)

inf {Ψ − λ, Ψ − μ, 1 − F (x), F (y) } >

U1 ×U2

1 . Λ1

For a sufficiently small ε > 0, there exist open neighborhoods N1 ⊂ U1 and N2 ⊂ U2 of x and y such that |Z(x, y)| < ε and |dZ(x, y)| < ε for (x, y) ∈ N1 × N2 . Obviously, the neighborhood N1 × N2 is disjoint from D. In order to compute the second order derivatives for an arbitrary point (x, y) ∈ N1 × N2 , let us choose geodesic normal coordinates (x1 , . . . , xn ) at x satisfying that hij = λi δij at x. We recall that the vector wi (x, y) is defined by the reflection of ∂F∂x(x) in i n+2 R with respect to the hyperplane orthogonal to F (x) − F (y) and passing through the origin. Thus the vector wi (x, y) is given by   ∂F (x) F (x) − F (y) F (x) − F (y) ∂F (x) −2 , wi (x, y) = . ∂xi ∂xi |F (x) − F (y)| |F (x) − F (y)| For any point (x, y) ∈ N1 × N2 , we obtain the following estimates:   ∂F ∂Z (20) (y), Ψ(x)F (x) − ν(x) = − (x, y) < ε, ∂yi ∂yi

A characterization of Clifford hypersurfaces

 (21)

| wi (x, y), Ψ(x)F (x) − ν(x) | =

∂F ∂xi (x), F (y)

527



1 − F (x), F (y)

Z(x, y) < Λ21 ε,

and (22)

|F (y)|2 |Ψ(x)F (x) − ν(x)|2 − F (y), Ψ(x)F (x) − ν(x) 2 = (1 + Ψ(x)2 ) − (Ψ(x) − Z(x, y))2 = 1 + 2Ψ(x)Z(x, y) − Z(x, y)2 > 1 − ε2 = 0.

Let V (x, y) be the orthogonal projection of Ψ(x)F (x) − ν(x) onto ∂F (y)}ni=1 and ν(y). For a suitably chosen TF (y) Sn+1 , which is spanned by { ∂y i small ε, we can conclude that • F (y) and Ψ(x)F (x) − ν(x) are linearly independent, • |V (x, y)|2 = 1 + 2Ψ(x)Z(x, y) − Z(x, y)2 from the formula (22). The inequality (20) implies that ∂F ∂F • The set { ∂y , . . . , ∂y , V } is a basis of TF (y) Sn+1 . Moreover, we can 1 n make the angle between V and ν(y) arbitrarily close to 0 in N1 × N2 for a sufficiently small ε > 0.

Note that the vectors wi (x, y) and wj (x, y) for 1 ≤ i, j ≤ n and i = j are mutually orthogonal unit vectors in TF (y) Sn+1 . Finally the inequality (21) implies that • The set {w1 , . . . , wn , V } is also a basis of TF (y) Sn+1 . Moreover, we can make the angle between wi (1 ≤ i ≤ n) and V arbitrarily close to π2 in N1 × N2 for a sufficiently small ε > 0. Therefore we choose geodesic normal coordinates at y satisfying that for 1≤i 0,     ∂F ∂F • wi (x, y), ∂y (y) ≥ 0 and w (x, y), (y) = 0. i ∂yj i

Moreover, the magnitude of the difference vector between wi (x, y) and can be controlled by Z(x, y) and |dZ(x, y)| as follows:

∂F ∂yi (y)

528

S.-H. Min and K. Seo

Lemma 6.2. Let (x, y) ∈ Σ × Σ \ D such that Z(x, y) = 0. Then there exist open neighborhoods N1 of x and N2 of y in Σ satisfying that (N1 × N2 ) ∩ D = ∅ and there exists a constant Λ2 > 0 depending only on x, y such that wi (x, y) −

∂F (y) ∂yi

2

 ≤ Λ2 Z(x, y) + |dZ(x, y)|

for any point (x, y) ∈ N1 × N2 . Proof. Let us choose the open neighborhoods N1 and N2 of x and y as above. Furthermore, for an arbitrary point (x, y) ∈ N1 × N2 , we choose the ∂F geodesic normal coordinates at x and y as above. If wi (x, y) = ∂y (y), then i ∂F it is trivial. Thus we may assume that wi (x, y) = ∂yi (y). Since we can make ∂F the angle between ∂y (y) and wi (x, y) sufficiently small, we may assume that i ∂F the angle between the vectors wi (x, y) − ∂y (y) and V is less than π4 . It is i easy to see that 2  2 ∂F ∂F V wi (x, y) − (y) < 2 wi (x, y) − (y), . ∂yi ∂yi |V | Therefore wi (x, y) −

∂F (y) ∂yi

2

<

2 |V |2

2 = |V |2 

 wi (x, y) −



∂F (y), V ∂yi



2

∂F wi (x, y) − (y), Ψ(x)F (x) − ν(x) ∂yi  2 ∂F (x), F (y) ∂xi ∂Z =4 (x, y) Z(x, y) + 1 − F (x), F (y)

∂yi



2

≤ 4Λ4 εZ(x, y) + 4ε(2Λ21 + 1) |dZ(x, y)|  ≤ Λ2 Z(x, y) + |dZ(x, y)| , for any point (x, y) ∈ N1 × N2 .



In the proof of Proposition 4.1, Proposition 4.2, and Proposition 4.3, we used the property Z(x, y) =

∂Z ∂Z (x, y) = (x, y) = 0 ∂xi ∂yi

at a global minimum point (x, y). However it is no longer valid in general. Using the geodesic normal coordinates we picked in the above, we have the

A characterization of Clifford hypersurfaces

529

following second order derivatives in general. For any point (x, y) ∈ N1 × N2 , n

∂2Z (x, y) = (23) ∂xi 2

 ΔΣ Ψ(x) − 2

i=1

n

i=1

2 | ∂Ψ(x) ∂xi |

Ψ(x) − λi (x) 

+ |A(x)|2 − n Ψ(x) (1 − F (x), F (y) ) + nΨ(x) + 2

n

i=1

∂Ψ(x) ∂xi

∂Z − |A(x)|2 Z(x, y), Ψ(x) − λi (x) ∂xi

1 ∂F (y) 2 ∂2Z (24) (x, y) = λi (x) − Ψ(x) − wi (x, y) − ∂xi ∂yi 2 ∂yi   1 ∂F (y) ∂Z − F (x), (x, y), 1 − F (x), F (y)

∂yi ∂xi n

∂2Z (25) (x, y) = nΨ(x) − nZ(x, y). ∂yi 2 i=1

Lemma 6.3. Let (x, y) ∈ Σ × Σ \ D such that Z(x, y) = 0. Then there exist open neighborhoods N1 of x and N2 of y in Σ satisfying that (N1 × N2 ) ∩ D = ∅ and there exists a constant Λ > 0 depending only on x, y such that (26)

n n n

∂2Z ∂2Z ∂2Z (x, y) + 2 (x, y) + (x, y) ∂xi 2 ∂xi ∂yi ∂yi 2 i=1 i=1 i=1 ≤ Λ Z(x, y) + |dZ(x, y)|

for all (x, y) ∈ N1 × N2 . Proof. Applying the estimates (18), (19), the equalities (23), (24), (25), and Lemma 6.2, we have n n n

∂2Z ∂2Z ∂2Z + 2 + ∂xi 2 ∂xi ∂yi ∂yi 2 i=1 i=1 i=1   2 |∇Σ Ψ(x)|2 2 ≤ (1 − F (x), F (y) ) ΔΣ Ψ(x) − + |A(x)| − n Ψ(x) n Ψ(x) + 2nΛ21 |dZ| + Λ1 Z + n(Λ2 Z + Λ2 |dZ|) + 2nΛ21 |dZ| + nZ ≤ Λ Z + |dZ| .



530

S.-H. Min and K. Seo

We define the set Ω by Ω := {x ∈ Σ : there exists a point y ∈ Σ \ {x} such that Z(x, y) = 0}. Lemma 5.1 shows that the set Ω is nonempty. Moreover we can prove the following: Theorem 6.4. The set Ω is an open subset of Σ. We need the following strict maximum principle for a degenerate second order elliptic partial differential equation. Theorem 6.5 ([7, 8]). Let Ω be an open subset of an n-dimensional Riemannian manifold, and let X1 , . . . , Xm be smooth vector fields on Ω. Assume that ϕ : Ω → R is a nonnegative smooth function satisfying m

j=1

(D2 ϕ)(Xj , Xj ) ≤ −L inf (D2 ϕ)(ξ, ξ) + L|dϕ| + Lϕ, |ξ|≤1

where L is a positive constant. Let F = {x ∈ Ω : ϕ(x) = 0} be the zero set of the function ϕ. Moreover,  suppose that γ : [0, 1] → Ω is a smooth path such that γ(0) ∈ F and γ  (s) = m j=1 fj (s)Xj (γ(s)) for suitable smooth functions f1 , . . . , fm : [0, 1] → R. Then γ(s) ∈ F for all s ∈ [0, 1]. Proof of Theorem 6.4. Take an open subset Ω of 2n-dimensional manifold Σ × Σ as an open neighborhood N1 × N2 of (x, y) in Lemma 6.3 with the ∂ usual product topology. Let Xi = ∂x + ∂y∂ i for i = 1, . . . , n. Then Lemma 6.3 i shows that the condition of Theorem 6.5 is satisfied. Applying Theorem 6.5  with L = Λ, we have the conclusion. We now complete our proof of Theorem 1.1. By Proposition 6.1 and Theorem 6.4, we see that ΔΣ Ψ(x) = 0 for all x ∈ Ω. Thus Corollary 3.6 implies that |A(x)|2 = n for all x ∈ Ω. Furthermore, since |A|2 = n(n − 1)λ2 by minimality and two distinct principal curvatures condition, we conclude that λ and μ are constant on Ω, which implies that Ψ is constant on Ω. By analytic continuation for solutions of elliptic partial differential equations, we see that Ψ is constant on Σ. Hence λ and μ are constant on Σ, which shows that Σ is an isoparametric minimal hypersurface with two distinct principal curvatures. From the Cartan’s classification of isoparametric hypersurfaces and analysis of the structure equations as before, it follows that

A characterization of Clifford hypersurfaces

531

Σ is congruent to a Clifford minimal hypersurface. However, this is a contradiction to the fact that κ = 1 on any Clifford minimal hypersurface, which completes the proof.

Acknowledgements The authors would like to thank Professor Simon Brendle for his valuable comments on this work. The first author was supported in part by the MOST research grant 102-2115-M-002-013-MY3 in Taiwan. The second author was supported in part by Samsung Science and Technology Foundation under Project Number SSTF-BA1501-04.

References [1] L. J. Al´ıas, S. C. de Almeida, and A. Brasil, Jr., Hypersurfaces with constant mean curvature and two principal curvatures in S n+1 , An. Acad. Brasil. Ciˆenc. 76 (2004), no. 3, 489–497. [2] B. Andrews, Noncollapsing in mean-convex mean curvature flow, Geom. Topol. 16 (2012), no. 3, 1413–1418. [3] B. Andrews, Z. Huang, and H. Li, Uniqueness for a class of embedded Weingarten hypersurfaces in S n+1 , in Introduction to modern mathematics, Vol. 33 of Adv. Lect. Math. (ALM), 95–107, Int. Press, Somerville, MA (2015). [4] B. Andrews, M. Langford, and J. McCoy, Non-collapsing in fully nonlinear curvature flows, Ann. Inst. H. Poincar´e Anal. Non Lin´eaire 30 (2013), no. 1, 23–32. [5] B. Andrews and H. Li, Embedded constant mean curvature tori in the three-sphere, J. Differential Geom. 99 (2015), no. 2, 169–189. [6] P. B´erard, Remarques sur l’´equation de J. Simons, in Differential geometry, Vol. 52 of Pitman Monogr. Surveys Pure Appl. Math., 47–57, Longman Sci. Tech., Harlow (1991). [7] J.-M. Bony, Principe du maximum, in´egalite de Harnack et unicit´e du probl`eme de Cauchy pour les op´erateurs elliptiques d´eg´en´er´es, Ann. Inst. Fourier (Grenoble) 19 (1969), no. fasc. 1, 277–304 xii. [8] S. Brendle, Ricci flow and the sphere theorem, Vol. 111 of Graduate Studies in Mathematics, American Mathematical Society, Providence, RI (2010), ISBN 978-0-8218-4938-5.

532

S.-H. Min and K. Seo

[9] S. Brendle, Alexandrov immersed minimal tori in S 3 , Math. Res. Lett. 20 (2013), no. 3, 459–464. [10] S. Brendle, Embedded minimal tori in S 3 and the Lawson conjecture, Acta Math. 211 (2013), no. 2, 177–190. [11] S. Brendle, Embedded Weingarten tori in S 3 , Adv. Math. 257 (2014), 462–475. [12] F. Brito and M. L. Leite, A remark on rotational hypersurfaces of S n , Bull. Soc. Math. Belg. S´er. B 42 (1990), no. 3, 303–318. ´ Cartan, Familles de surfaces isoparam´etriques dans les espaces ` [13] E. a courbure constante, Ann. Mat. Pura Appl. 17 (1938), no. 1, 177–191. [14] S. S. Chern, On surfaces of constant mean curvature in a threedimensional space of constant curvature, in Geometric dynamics (Rio de Janeiro, 1981), Vol. 1007 of Lecture Notes in Math., 104–108, Springer, Berlin (1983). [15] L. Hauswirth, M. Kilian, and M. U. Schmidt, Mean-convex Alexandrov embedded constant mean curvature tori in the 3-sphere, Proc. Lond. Math. Soc. (3) 112 (2016), no. 3, 588–622. ¨ [16] H. Hopf, Uber Fl¨ achen mit einer Relation zwischen den Hauptkr¨ ummungen, Math. Nachr. 4 (1951), 232–249. [17] W.-Y. Hsiang, On the construction of infinitely many congruence classes of imbedded closed minimal hypersurfaces in S n (1) for all n ≥ 3, Duke Math. J. 55 (1987), no. 2, 361–367. [18] G. Huisken, A distance comparison principle for evolving curves, Asian J. Math. 2 (1998), no. 1, 127–133. [19] S. Ilias, B. Nelli, and M. Soret, Caccioppoli’s inequalities on constant mean curvature hypersurfaces in Riemannian manifolds, Ann. Global Anal. Geom. 42 (2012), no. 4, 443–471. [20] H. B. Lawson, Jr., Complete minimal surfaces in S 3 , Ann. of Math. (2) 92 (1970), 335–374. [21] S.-H. Min and K. Seo, Characterizations of a Clifford hypersurface in a unit sphere via Simons’ integral inequalities, Monatsh. Math. 181 (2016), no. 2, 437–450. ˆ [22] T. Otsuki, Minimal hypersurfaces in a Riemannian manifold of constant curvature., Amer. J. Math. 92 (1970), 145–173.

A characterization of Clifford hypersurfaces

533

ˆ [23] T. Otsuki, On integral inequalities related with a certain nonlinear differential equation, Proc. Japan Acad. 48 (1972), 9–12. ˆ [24] T. Otsuki, On a bound for periods of solutions of a certain nonlinear differential equation. I, J. Math. Soc. Japan 26 (1974), 206–233. [25] O. Perdomo, Ridigity of minimal hypersurfaces of spheres with two principal curvatures, Arch. Math. (Basel) 82 (2004), no. 2, 180–184. [26] O. M. Perdomo, Embedded constant mean curvature hypersurfaces on spheres, Asian J. Math. 14 (2010), no. 1, 73–108. [27] J. B. Ripoll, Superf´ıcies invariantes de curvatura m´edia constante (1986). Tese de Doutorado, IMPA, Rio de Janeiro. [28] R. Schoen, L. Simon, and S. T. Yau, Curvature estimates for minimal hypersurfaces, Acta Math. 134 (1975), no. 3-4, 275–288. [29] J. Simons, Minimal varieties in riemannian manifolds, Ann. of Math. (2) 88 (1968), 62–105. [30] Q. Wang, Rigidity of Clifford minimal hypersurfaces, Monatsh. Math. 140 (2003), no. 2, 163–167. [31] G. Wei, Complete hypersurfaces with constant mean curvature in a unit sphere, Monatsh. Math. 149 (2006), no. 3, 251–258. [32] G. Wei, Rigidity theorem for hypersurfaces in a unit sphere, Monatsh. Math. 149 (2006), no. 4, 343–350. [33] G. Wei, Q.-M. Cheng, and H. Li, Embedded hypersurfaces with constant mth mean curvature in a unit sphere, Commun. Contemp. Math. 12 (2010), no. 6, 997–1013. [34] Y. L. Xin, Berstein type theorems without graphic condition, Asian J. Math. 9 (2005), no. 1, 31–44.

534

S.-H. Min and K. Seo

Department of Mathematics, Chungnam National University Daehak-ro 99, Yuseong-gu, Daejeon, 305-764, Korea E-mail address: [email protected] Department of Mathematics, Sookmyung Women’s University Hyochangwongil 52, Yongsan-ku, Seoul, 140-742, Korea E-mail address: [email protected] Received June 8, 2015

A characterization of Clifford hypersurfaces among ...

two distinct principal curvatures λ and μ of multiplicity n − 1 and. 1, respectively. It is known that if λ>μ, there exist many com- pact embedded constant mean curvature hypersurfaces [26]. In this paper, we prove that if μ>λ, then Σ is congruent to a Clifford hypersurface. The proof is based on the arguments used by Brendle.

224KB Sizes 0 Downloads 173 Views

Recommend Documents

Clifford-Takes-A-Trip-Clifford-8x8.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item.

SINGULARITIES OF SYMMETRIC HYPERSURFACES ...
National Council of Research and Technology (CONICET). Buenos Aires ... coordinates. This allows us to obtain critical information concerning the singu- .... by a polynomial f ∈ Fq[T] of degree k + d < q − 1 is a deep hole of the standard.

Clifford-The-Big-Red-Dog-Clifford-8x8.pdf
Page 1. Whoops! There was a problem loading more pages. Clifford-The-Big-Red-Dog-Clifford-8x8.pdf. Clifford-The-Big-Red-Dog-Clifford-8x8.pdf. Open. Extract.

Stable minimal hypersurfaces in a Riemannian manifold with pinched ...
DOI : 10.1007/s10455-011-9293-x. Cite this article as: Dung, N.T. & Seo, K. Ann Glob Anal Geom (2012) 41: 447. doi:10.1007/s10455-011-9293-x. 6 Citations ...

Clifford Organizers.pdf
than one graphic organizer there are times where I will fill in. responses from my students for one of the organizers without. having them copy it...in this instance, I use it more as a reading. tool rather than a writing tool for my students. Happy

Characterization of conjugate powers of a matrix
Since the characteristic roots of Ak are kth power of characteristic roots ai, for i = 1···n of matrix A and if Ak is a conjugate then from above lemma 1 it follows that. Ak has same roots as that of A and since there is a permutation σ ∈ Sn su

Fundamental tone of minimal hypersurfaces with finite ...
Seo Journal of Inequalities and Applications (2016) 2016:127 .... mark that the finite index condition can be omitted, since the finiteness of the L norm.

Enumeration of singular hypersurfaces on arbitrary ...
Let q ∈ X and v1,...vm be a basis for TXq. Then there exists sections s1,...sm ∈ H0(X, L) such that for all i, j ∈ {1,2...m} si(q)=0,. ∇si|q(vi) = 0 and. ∇si|q(vj) = 0.

Singularities of symmetric hypersurfaces and Reed-Solomon codes
Finite fields, Reed–Solomon codes, deep holes, symmetric polynomials, singular hypersurfaces ..... The dimension dim V of a K-variety V is the length r of the.

Clifford search strategies.pdf
Exalead: claims to have over 8 billion searchable pages. www.exalead.com/search/. • If you don't find what you are looking for at one site, go to another. 3.

A characterization of the Extended Serial ...
Oct 27, 2011 - 3On a restricted domain of preference profiles, however, Heo ..... Let R ∈ RN and π ∈ Π. For each σ ∈ ΣN , each agent does his best during.

Characterization of a trp RNA-binding Attenuation ...
the leader region of read-through trp mRNAs induces formation of an RNA ..... tides in all positions except G3 of each repeat, binds WT TRAP with similar affinity ...

Identification and characterization of a PHGPX gene ...
vine serum albumin (BSA) as a standard. SDS-PAGE analysis. An equal amount of each protein sample (35 µg) was loaded on 12–. 15 % gradient SDS-PAGE ...

A characterization of the Nash bargaining solution - Springer Link
2 Department of Economics, 260 Heady Hall, Iowa State University, Ames,. Iowa 50011, USA (e-mail address: [email protected]; http://volij.co.il). 3 Department of ...

Characterization of Activation Products in a Medical ...
photon beam was produced by impinging electrons on a tungsten target. The electron ... The mean energy of the electron beam was 18.3 MeV with a Gaussian.

A Characterization of Farsightedly Stable Networks
Jul 22, 2010 - Abstract: We study the stability of social and economic networks when .... g is simply a list of which pairs of individuals are linked to each other.

Molecular Characterization of Human Group A ...
predominant strain [Kim et al., 1999; Kim, 1993; Seo and Sim, 2000]. These distinct changes in ..... Becton, Dickinson & Company Sparks, MD, USA) for 30 min at. 37oC and ...... The best hope for its prevention is the development of an effective ...

Towards a characterization and understanding of ...
technology; (b) new pedagogy which abandons an “information transfer” ..... Keitel and K. Ruthven (Eds.), Learning from Computers: Mathematics Education and.

Visuomotor characterization of eye movements in a ...
tive analysis of the data collected in drawing, it was clear that all subjects ...... PhD dissertation, Berkeley, University of California, Computer Science. Division.