Acessibilidade / Reportar erro

Embedded positive constant r-mean curvature hypersurfaces in Mm × R

Abstracts

Let M be an m-dimensional Riemannian manifold with sectional curvature bounded from below. We consider hypersurfaces in the (m + 1)-dimensional product manifold M × R with positive constant r-mean curvature. We obtain height estimates of certain compact vertical graphs in M × R with boundary in M × {0}. We apply this to obtain topological obstructions for the existence of some hypersurfaces. We also discuss the rotational symmetry of some embedded complete surfaces in S² × R of positive constant 2-mean curvature.

product manifold; hypersurface; r-mean curvature


Seja M uma variedade riemanniana de dimensão m, de curvatura seccional limitada de abaixo. Consideramos as hipersuperfícies na variedade produto M × R de dimensão m+1, com curvatura r-média constante positiva. Obtemos uma estimativa para altura das alguns gráficos verticais em M × R com seus fronteiras em M × {0}. Aplicamos isto para obter as obstruções topológicas sobre existência das algumas hipersuperfícies. Também discutimos a simetria rotacional das algumas superfícies completas em S² × R de curvatura 2-média constante positiva.

variedade produto; hipersuperfície; curvatura r-média


MATHEMATICAL SCIENCES

Embedded positive constant r-mean curvature hypersurfaces in Mm × R

Xu ChengI; Harold RosenbergII,* * Member Academia Brasileira de Ciências

IInstituto de Matemática, Universidade Federal Fluminense-UFF, Centro, 24020-140 Niterói, RJ, Brasil

IIUniversité Paris 7, Institut de Mathématiques, 2 Place Jussieu, 75251 Paris cedex 05, France

Correspondence Correspondence to Harold Rosenberg E-mail: rosen@math.jussieu.fr

ABSTRACT

Let M be an m-dimensional Riemannian manifold with sectional curvature bounded from below. We consider hypersurfaces in the (m + 1)-dimensional product manifold M × R with positive constant r-mean curvature. We obtain height estimates of certain compact vertical graphs in M × R with boundary in M × {0}. We apply this to obtain topological obstructions for the existence of some hypersurfaces. We also discuss the rotational symmetry of some embedded complete surfaces in S2 × R of positive constant 2-mean curvature.

Key words: product manifold, hypersurface, r-mean curvature.

RESUMO

Seja M uma variedade riemanniana de dimensão m, de curvatura seccional limitada de abaixo. Consideramos as hipersuperfícies na variedade produto M × R de dimensão m+1, com curvatura r-média constante positiva. Obtemos uma estimativa para altura das alguns gráficos verticais em M × R com seus fronteiras em M × {0}. Aplicamos isto para obter as obstruções topológicas sobre existência das algumas hipersuperfícies. Também discutimos a simetria rotacional das algumas superfícies completas em S2 × R de curvatura 2-média constante positiva.

Palavras-chave: variedade produto, hipersuperfície, curvatura r-média.

1 INTRODUCTION

If

m+1 is an (m+1)-dimensional oriented Riemannian manifold and Sm is a hypersurface in , the r-mean curvature of S, denoted by Hr, is the weighted r'th symmetric function of the second fundamental form (see Definition 2.1). The hypersurfaces with constant r-mean curvature include those of constant mean curvature, and of constant Gauss-Kronecker curvature. In some situations, for example, = m+1, m+1, m+1(–1), a hypersurface of constant r-mean curvature is a critical point for a certain variational problem (See the related works in (Reilly 1973), (Rosenberg 1993), (Barbosa and Colares 1997) and (Elbert 2002)).

Heinz discovered that a compact graph S in

m+1 with zero boundary values, of constant mean curvature H ¹ 0, is at most a height from its boundary. A hemisphere in m+1 of radius shows this estimate is optimal. Using Alexandrov reflection techniques, it follows that a compact embedded hypersurface with constant mean curvature H ¹ 0 and boundary contained in xm+1 = 0, is at most a distance from the hyperplane xm+1 = 0.

Height estimates have been obtained by the second author (Rosenberg 1993) for compact graphs S in

m+1 with zero boundary values and with a higher order positive constant r-mean curvature. He also obtained such height estimates in space forms. For example, in 3, such a graph of positive constant Gauss curvature can rise at most from the plane containing the boundary. In general, the maximum height is . For r = 1, this result in hyperbolic space m+1(–1) was proved by Korevaar, Kusner, Meeks and Solomon (Korevaar et al. 1992). We refer the reader to (Rosenberg 1993) for applications of these height estimates.

In (Hoffman et al. 2005), Hoffman, Lira and Rosenberg obtained height estimates for some vertical graphs in M2 × of nonzero constant mean curvature, where M is a Riemannian surface.

In this paper, we will consider orientable hypersurfaces in an (m + 1)-dimensional oriented product manifold M × with positive constant r-mean curvature. Here M is an m-dimensional Riemannian manifold with sectional curvature bounded below. The typical models are when M = m, m, m(–1). We will obtain height estimates for such compact vertical graphs with boundary in M × {0}. A significant difference between this case and the previous cases we discussed is the nature of the linearized operator. It is no longer a divergence form operator and we do not have a flux formula. We prove that,

THEOREM 1.1. (Th. 4.1). Let M be an m-dimensional oriented Riemannian manifold. Let S be a compact vertical graph in the (m+1)-dimensional product manifold M × with positive constant Hr, for some 1< r < m, with boundary in M × {0}. Let h: S ® denote the height function of S.

(i) If the sectional curvature of M satisfies K > 0, then on S,

(ii) When r = 2, if the sectional curvature of M satisfies K > –t (t > 0), and H2 > t, then on S,

(iii) When r = 1, if the sectional curvature of M satisfies K > –t (t > 0), and , then on S,

We give some applications of Theorem 1.1 using the Alexandrov reflection technique. First we give the extrinsic vertical diameter estimate of compact embedded surfaces in M × with positive constant r-mean curvature (under the same curvature assumptions as in Theorem 1.1. See Th.4.2); Secondly, we prove that if M is an m-dimensional compact Riemannian manifold and S is a noncompact properly embedded hypersurface in Mm × with positive constant r-mean curvature (under the same curvature assumptions as in Theorem 1.1), then the number of ends of S is not one (Th.4.3). Thus we give a topological obstruction for the existence of such a hypersurface. For example, if S is a noncompact properly embedded hypersurface in m × of positive constant r-mean curvature for 1 < r < m, then the number of ends of S is more than one (Cor.4.1).

In this paper, we also consider the properties of symmetry of certain embedded hypersurfaces of positive constant r-mean curvature. Alexandrov (Alexandrov 1962) showed an embedded constant mean curvature hypersurface in the Euclidean space must be a standard round hypersurface. The symmetric properties of hypersurfaces (with or without boundary) in the Euclidean space and the other space forms with constant mean curvature or higher order constant r-mean curvatures have been studied in varied degrees. We discuss the rotational symmetry of a complete embedded surface in 2 × of positive constant 2-mean curvature and obtain that,

THEOREM 1.2. (Th. 5.1; Cor. 5.1). Let 2 be the unit sphere in 3 and be the open upper hemisphere in 2. Let S be a complete orientable embedded surface in 2 × with H2 = constant > 0. If S Ì × , then S has the following properties:

(i) S is topologically a sphere; and

(ii) S is a compact surface of revolution about o' × , o' Î , that is, S is foliated by round circles in × {t}, the centers of which are o' × t, where t Î (hmin,hmax) and hmin,hmax are the maximum and minimum of the height of S respectively. Moreover, there exists a horizontal level × {t0}, t0 = , such that S divides into two (upper and lower) symmetric parts; and

(iii) Let f0Î [0,) denote the polar angle between the north pole of the open upper hemisphere and o'. Then H2 > –(if f0 = 0,H2 > 0) and the generating curve of the rotational surface S can be denoted by l(u) = (f(u), t(u)) where –1< u < 1, and f denotes the polar coordinate (about o') of 2, and:

where C is a real constant.

We conjecture a compact immersed surface in

2 × of positive constant 2-mean curvature is rotationally symmetric about a vertical line. Moreover, such a surface in 2 × is rotational when it is topologically a sphere.

2 PRELIMINARIES

Let

m+1 be an (m + 1)-dimensional oriented Riemannian manifold and let Sm be an m-dimensional orientable Riemannian manifold. Suppose x : S ® is an isometric immersion. We choose a unit normal field N to S and define the shape operator A associated with the second fundamental form of S, i.e., for any p Î S,

A : TpS ® TpS, áA(X),Yñ = – áX N,Yñ , X,Y Î TpS,

where is the Riemannian connection in .

Let l1,...,lm denote the eigenvalues of A. The r-th symmetric function of l1,...,lr, denoted by Sr, is defined as

DEFINITION 2.1. With the above notations, Hr = Sr, r = 1,...,m, is called the r-mean curvature of x.

Particularly, H1 = H is the mean curvature; Hm is the Gauss-Kronecker curvature. H2 is , modulo a constant, the scalar curvature of S, when the ambient space m+1 is Einstein (Elbert 2002).

We also introduce endomorphisms of T(S), the Newton transformations, defined by

T0 = I,

Tr = SrI – Sr–1A + ... + (–1)r Ar, r = 1, ... , m.

It is obvious that Tr, r = 0,1,...,m are symmetric linear operators and Tr = SrI – ATr–1.

Let e1,...,em be the principal directions corresponding respectively to the principal curvatures l1,..., lm. For i = 1,...,m, let Ai denote the restriction of the transformation A to the (m – 1)-dimensional subspace normal to ei, and let Sr(Ai) denote the r-symmetric function associated to Ai.

One has the following properties of Tr and Sr.

PROPOSITION 2.1. For 0< r < m, 1< i < m,

(i) Tr(ei) = ei = Sr(Ai);

(ii) (m – r)Sr = trace(Tr) = Sr(Ai);

(iii) (r + 1)Sr+1 = trace(ATr) = liSr(Ai);

(iv)

> H2.

Proposition 2.1 can been verified directly.

PROPOSITION 2.2. For each r (1< r < m), if H1, H2,...,Hr are nonnegative, then

(i) H1> > > ... > ;

(ii) Hr–1 Hr+1< ;

(iii) H1Hr > Hr+1, where, H0 = 1, Hm+1 = 0.

See the proof of (i) and (ii) of Proposition 2.2 in (Hardy et al. 1989. p.52). The inequality in (iii) can be obtained from (ii).

Given a function f in C2(S) for p Î S, the linear operator Hessian of f is defined as Hess f(X) = ÑXf), X Î TpS, where Ñ is the induced connection on S.

With Tr and Hess, we can define a differential operator Lr as follows:

DEFINITION 2.2. Given f Î C2(S), 0< r < m,

Given a local coordinate frame of S at p, by direct computation, we have locally the expression of Lr,

where

are the connection coefficients of Ñ.

From the above local expression, we know that the linear operator Lr is elliptic if and only if Tr is positive definitive. Clearly, L0 = tr(Hess f) = div(Ñ f) is elliptic.

In this paper, the ambient space which we study is an (m + 1)-dimensional product manifold Mm × , where M is an m-dimensional Riemannian manifold. S will be a hypersurface in M × with positive constant r-mean curvature, i.e, Hr = constant > 0.

We use t to denote the last coordinate in M × .

The height function, denoted by h, of S in M × is defined as the restriction of the projection t : M × ® to S , i.e., if p Î S,x(p) Î M × {t}, then h(p) = t. We have = h.

3 THE ELLIPTIC PROPERTIES OF THE OPERATOR Lr

The following Prop. 3.1 is known (cf. Elbert 2002. Lemma 3.10). For completeness, we give its proof.

PROPOSITION 3.1. Let

m+1 be an (m + 1)-dimensional oriented Riemannian manifold and let Sm be a connected m–dimensional orientable Riemannian manifold. Suppose x : S ® is an isometric immersion. If H2 > 0, then the operator L1 is elliptic.

PROOF. Since L1 is elliptic if and only if T1 is positive definite, we will prove that T1 is positive definite.

By H2 > 0 and > H2, H1 is nonzero and has the same sign on S. We may choose the normal N such that H1 is positive. So S1 = mH1 > 0.

By = å + 2S2 > , we have S1 > li. Then S1(Ai) = S1 – li > 0. Since T1ei = S1(Ai)ei (Prop. 2.1 (i)), T1 is positive definite.

In Prop. 3.2, we will prove the ellipticity of Lr when the hypersurface S with positive r-mean curvature has an elliptic or parabolic point.

PROPOSITION 3.2. Let

m+1 be an (m + 1)-dimensional oriented Riemannian manifold and let Sm be a connected m–dimensional orientable Riemannian manifold (with or without boundary). Suppose x : S ® is an isometric immersion with Hr > 0 for some 1< r < m. If there exists an interior point p of S such that all the principle curvatures at p are nonnegative, then for all 1< j < r – 1, the operator Lj is elliptic, and the j-mean curvature Hj is positive.

PROOF. Since Lj is elliptic if and only if Tj is positive definite, it is sufficient to prove that the eigenvalues of Tj are positive, that is Sj(Ai) > 0 on S, for all 1 < j < r – 1, 1 < i < m.

In the following proof, the ranges of j and i are 1 < j < r – 1, 1 < i < m respectively.

Since Sr > 0 and the principal curvatures of x at p are nonnegative, we have that at p, at least r principal curvatures are positive. For simplicity, we suppose that l1, l2,..., lr are positive.

Then, by direct verification, at p, Sj(Ai) > 0.

By continuity, there exists an open intrinsic ball B(p) Ì S with center p such that the functions Sj(Ai) > 0 on B(p).

For any q Î S, since S is connected, there exists a path g(t) (t Î [0,1]) in S, joining p to q with g(0) = p and g(1) = q.

Define J = {t Î [0,1]|Sj(Ai) > 0 on g|[0,t]}. Let t0 = sup J.

Note Sj(Ai) > 0 on B(p) so t0 > 0. By continuity, at t0, Sj(Ai) > 0.

We will prove that Sj(Ai) > 0 at t0 and hence t0Î J.

We first show that Sr–1(Ai) > 0 at t0. Otherwise, there exists 1 < i < r – 1 such that Sr–1(Ai) = 0 at t0. For this i, by Sr = liSr–1(Ai) + Sr(Ai), we have Sr(Ai) = Sr > 0 at t0. So, at t0, Sj(Ai) > 0, 1 < j < r – 1 and Sr(Ai) > 0. By (ii) in Prop. 2.2, at t0, the following inequality holds,

H1(Ai) >H2> ... > Hr–1> Hr.

Then at t0, Hr–1(Ai) > 0, i.e., Sr–1(Ai) > 0, which is a contradiction. Thus, we have Sr–1(Ai) > 0 at t0.

Next, by Sr–1(Ai) > 0 and H1(Ai) > H2

> ... > Hr–1 at t0, we have that Sj(Ai) > 0 at t0. Hence t0Î J.

If t0 < 1, by continuity, there exists an open intrinsic ball B(g(t0)) of center g(t0) such that Sj(Ai) > 0 on B(g(t0)), which contradicts our choice of t0 = sup J. Hence, t0 = 1.

So we obtain that at q, Sj(Ai) > 0. Hence the Lj, for all 1 < j < r – 1 are elliptic.

By Prop. 2.1 (ii), Hj (for all 1 < j < r – 1) are positive.

4 THE HEIGHT ESTIMATES OF THE VERTICAL GRAPH

In this section, we will consider hypersurfaces in an (m + 1)-dimensional product manifold Mm × of positive constant r-mean curvature.

LEMMA 4.1. Let M be an m–dimensional oriented Riemannian manifold and let S be an immersed orientable hypersurface in M × (with or without boundary). Then

where 1 < r < m + 1, h denotes the height function of S, and n = áN,ñ.

PROOF. Fix p Î S. Let {ei} be a geodesic orthonormal frame of S at p. So at p, ej(p) = 0. We can assume {ei} are the principal directions at p, that is, Aei(p) = liei, where li are the eigenvalues (principal curvatures) of A at p (this frame can been obtained by rotating {ei}).

Observe that the vertical translation in M × is an isometry. Hence h = is a Killing vector field, and

h = 0, for i = 1,...,m.

We have

where the T denotes the tangent part of the vector of T(M × ) to S.

Then

Hence

Since rSrn is independent of the choice of the frame, we obtain that, on S

In order to prove the following Lemma 4.2, we first recall some known results.

Let S Ì

m+1 be an oriented hypersurface. Let D be a compact domain of S. Consider a variation of D, denoted by f. f:(–,) × ® M × , > 0, such that for each s Î (–,) , the map fs : {s} × ® M × , fs(p) = f(s, p) is an immersion, and f0 = .

Let At(p) be the shape operator of S(t) at p and for 0 < r < m, let Sr(t)(p) be the r-th symmetric function of the eigenvalues of At(p). We are interested in the first variation of Sr. To calculate this one may differentiate the equation rSr = tr(Tr–1A). We refer the reader to (Rosenberg 1993) for details of this calculation in space forms. In general ambient spaces, Elbert (Elbert 2002. Proposition 3.2) proved the following result.

Set Es(p) = (p, s) and fs = < Es, Ns > , where Ns is the unit normal to fs(D). We have

where

N is defined as N(X) = (N, X)N, the curvature operator of , and denotes the tangent part of Es.

LEMMA 4.2. With M and S as in Lemma 4.1, assume Hr of S is constant, for some r, 1< r < m + 1. Then, on S,

PROOF. We may choose the variation fs : (–, ) × D Ì (–, ) × S ® M × given by vertical translation of M × , (x, t0) ® (x, t0 + s), (x, t0) Î M × R.

Under this variation, Sr(s) is constant, i.e., Sr(s) = 0. Also by the hypothesis Sr = constant on S, we have (Sr) = 0, for s = 0. Hence, when s = 0, we have f0 = áE0, N0ñ = á, Nñ = n and

Thus (4.6) holds on S.

We will express tr(T1

N) using the sectional curvature of M in order to prove Theorem 4.1.

Given p = (x, t) Î M × . For X Î TpM × , let Xh denote the horizontal component of X. If {ei} denotes a geodesic frame of S at p and N denotes the unit normal to S at p, we have, by direct calculation,

(ei, N, ei, N) = (, Nh, , Nh) = K(, Nh) × |Ù Nh|2,

where K is the sectional curvature of M.

If ei are also the principal directions of A. We have Trei = Sr(Ai)ei. Then

THEOREM 4.1. Let M be an m-dimensional oriented Riemannian manifold. Let S be a compact vertical graph in the product manifold M × with boundary in M × {0} and with positive constant Hr, for some 1< r < m. Let h : S ® denote the height function of S.

(i) If the sectional curvature of M satisfies K > 0, then on S,

(ii) When r = 2, if the sectional curvature of M satisfies K > –t (t > 0), and H2 > t, then on S,

(iii) When r = 1, if the sectional curvature of M satisfies K > –t (t > 0), and > t, then on S,

PROOF. At a highest point, all the principal curvatures have the same sign. Since we assume that Hr > 0, we know that at this point, all the principal curvatures are nonnegative and the unit normal N to S is downward pointing. Since S is a vertical graph, we may choose the smooth unit normal field N to S to be downward pointing (i.e.,n = áN,ñ < 0 on S). Hence we can apply Prop. 3.1 and Prop. 3.2 to obtain that, Lr–1 is elliptic, and Hi(1 < i < r – 1) are positive.

Define j = ch + n on S, where c is a positive constant to be determined. On ¶S, j = n < 0.

Since Lr–1 is an elliptic operator, we have, by the maximum principle, that if Lr–1j > 0, then j < 0 on S. Then h < –< .

Now we will choose c such that Lr–1j > 0. By Lemma 4.1 and 4.2,

Lr–1(j) = crSrn – (S1Sr – (r + 1)Sr+1 + tr(Tr–1

N))n.

So Lr–1(j) > 0 is equivalent to

–rSrc + S1Sr – (r + 1)Sr+1 + tr(Tr–1

N) > 0.

If K > a (a < 0), we have

So we may choose c such that

–rSrc + S1Sr – (r + 1)Sr+1 + (m – r + 1)aSr–1> 0.

When H1,...,Hr are nonnegative, H1Hr> Hr+1, that is, S1Sr> (r + 1)Sr+1. So it is sufficient to choose c such that

–rSrc + S1Sr + (m – r + 1)aSr–1> 0.

Then, .

(i) Take a = 0 and choose c < . Hence h < .

(ii) Take a = –t(t > 0). Then c < H1 – t · .

When r = 2, c < H1 – t · = H1(1 – ).

By H1> , we may choose c < (1 – ) = . Hence h < .

(iii) When K > a (a < 0) and r = 1, we have a better estimate. Note

Similar to the above, we may choose c < . Take a = –t(t > 0). We have h < .

REMARK 4.1. The height estimate in Theorem 4.1 is sharp. Consider a hemisphere of the unitary round sphere

m in m+1. It is a vertical graph on m = m × {0} of Hr = 1 with boundary m–1 × {0} and has the maximum height 1.

REMARK 4.2. In the case m = 2 and r = 1, (i) and (iii) of Theorem 4.1 was proved in (Hoffman et al. 2005).

THEOREM 4.2. Let M be an m-dimensional Riemannian manifold and let S be a compact orientable embedded hypersurface in M × with Hr = constant > 0, for some 1< r < m. Then S is symmetric about some horizontal surface M × {t0}, t0Î . Moreover,

(i) If the sectional curvature of M satisfies K > 0, then the extrinsic vertical diameter of S is no more than 2;

(ii) When r = 2, if the sectional curvature of M satisfies K > –t and H2 > t(t > 0), then the extrinsic vertical diameter of S is no more than .

(iii) When r = 1, if the sectional curvature of M satisfies K > –t(t > 0), and > t, then the extrinsic vertical diameter of S is no more than .

PROOF. We will prove that S has a horizontal surface of symmetry M(t0) = M × {t0}, for some t0Î .

Note that vertical translation is an isometry, as well as reflection through each horizontal M(t). Hence we can use the Alexandrov reflection technique. One comes down from above S with the horizontal surfaces M(t). For t slightly smaller than the highest value of the height of S, the part of S above M(t) is a vertical graph of bounded gradient over a domain in M(t). The symmetry through M(t) of this part of S, is below M(t), contained in the domain bounded by S, and meets S only along the boundary. One continues to do these reflections through the M(t) translated downwards, until a first accident occurs. This plane is a plane of symmetry of S, and the part of S above this plane is a graph with zero boundary values. This proves the Theorem.

Remark 4.3. Consider the case m = 2 and r = 2 in Theorem 4.1 or Theorem 4.2. From the Gauss equation, we obtain that the intrinsic sectional curvature of S has a positive lower bound H2. So Bonnet-Myers theorem yields the intrinsic diameter estimate of S and hence the vertical height or the extrinsic vertical diameter estimate. That is, (a) when S is a vertical graph with zero boundary values, a geodesic on S from the highest point to ¶S (and hence the vertical height) is at most ; (b) when S is a compact embedded surface, a geodesic on S from the highest point to the lowest point (and hence extrinsic vertical diameter) is at most . But our estimates (a) and (b) are better. Moreover our estimate works for any dimension m and any r, for which Bonnet-Myers theorem indeed doesn't apply.

THEOREM 4.3. Let M be an m-dimensional compact Riemannian manifold and let S be a noncompact properly embedded hypersurface in × with Hr = constant > 0, for some 1< r < m. If the sectional curvature of M and the r-mean curvature of S satisfy the conditions in (i), (ii) or (iii) of Theorem 4.1, then the number of ends of S is not one.

PROOF. Suppose on the contrary, that S has exactly one end E. Since S is properly embedded, E must go up or down, but not both. Assume E goes down. Then S has a highest point so we can do Alexandrov reflection coming down from above S with horizontal surfaces M(t).

Since S is not invariant by symmetry in any M(t), there is no first point of contact of the symmetry of the part of S above M(t), with the part of S below M(t). But by the Alexandrov reflection technique, the part of S above each M(t) is always a vertical graph. This contradicts the height estimate for such vertical graphs.

COROLLARY 4.1. Let S be a noncompact properly embedded hypersurface in m × with Hr = constant > 0, for some 1< r < m. Then the number of ends of S is not one.

5 ROTATIONAL SYMMETRY OF SOME SURFACES IN

2 × OF CONSTANT 2-MEAN CURVATURE

If M is the unit sphere 2 in 3, an embedded surface in 2 × of positive constant 2-mean curvature has another symmetry under some restriction. If we demand the vertical projection of such a surface on 2 × {0} is contained in an open hemisphere, then it is also rotational about a vertical line parallel to the vertical -axis.

We now give some definitions which are modifications of the related concepts in

3 (cf. Hopf 1983. p.147-148).

Let be the open hemisphere in 2 and ¶ be the boundary of . Let {p,–p} denote a pair of antipodal points on ¶, and g denote a semi-circle on joining p and –p. We call the surface P = g × in 2 × a vertical geodesic strip.

DEFINITION 5.1. For a point q Î × , a point q' Î × is called a point of symmetry of q about a vertical geodesic strip P = g × if, for a geodesic l passing through q and perpendicular to P, we have q' Î l, lies on the opposite part of l divided by P, and dist(q, P) = dist(q', P). This means q' is the image of q by the isometry of S × which is reflection of each S(t) through g × {t}.

A vertical geodesic strip P = g × in 2 × is called a vertical strip of symmetry for a set W Ì 2 × if, for every point q Î W, its point of symmetry q' is also in W.

We will obtain the following:

Theorem 5.1. Let

2 be the unit sphere in 3 and be an open hemisphere in 2. Let S be a complete oriented embedded surface in 2× with H2 = constant > 0. If S Ì × , then S has the following properties:

(i) S is topologically a sphere; and

(ii) S is a surface of revolution about a vertical line parallel to the vertical -axis, that is, S is foliated by round circles in × {t}, the centers of which are the same point on modulo t, where t Î (hmin,hmax ) Ì , and hmin, hmax are the maximum and minimum of vertical height of S respectively; and

(iii) there exists t0 = such that × {t0} divides S into two (upper and lower) symmetric parts.

PROOF. By the Gauss equation, the sectional curvature of S is bounded below by a positive constant, hence S is compact by the Bonnet-Myers theorem. By the Gauss-Bonnet theorem, S is topologically a sphere. By Corollary 4.2, we know there exists t0 = such that the horizontal level × {t0} divides S into two (upper and lower) symmetric parts.

We prove S is rotational in two steps 1) and 2).

1) S has a vertical strip of symmetry about every pair of antipodal points.

Fix a pair of antipodal points {p,–p} on ¶. Then ¶ is divided into two semi-circles joining p and –p. Let q denote the rotational angle from one semi-circle to the other (0 < q < p), and g(q) denote the semi-circle on joining p and –p, with the rotational angle q.

Since S Ì × and S is compact, the vertical strip P(q) = g(q) × is disjoint from S for sufficiently small q. Note the rotation on 2 is an isometry. We may do the Alexander reflection through P(q), as q moves from 0 to p, to obtain a vertical strip of symmetry P(q0) such that S is divided into two graphs of symmetry over a domain of P(q0).

By the arbitrariness of p on ¶, we have proved 1).

2) Since any two such vertical strips of symmetry intersect in a line parallel to the vertical -axis, it is sufficient to prove all of these lines coincide.

Choose a horizontal surface (t) = × {t} whose intersection with S has at least two points.

Fix this t. Let S(t) = (t)Ç S. Since S is embedded and compact, S(t) is a simple closed curve on (t). Any vertical strip of symmetry P = g × of S corresponds to a semi-circle of symmetry g(t) = g × {t} of S(t) on (t). By 1), we know that every pair of antipodal points {p, –p} on ¶(t) determines a semi-circle of symmetry g of S(t).

We will look into the relation of all the semi-circles of symmetry {g(t)}. For simplification of notation, we omit t in g(t). Fix a semi-circle of symmetry g0 on (t). Let g1 be any other one, and b denote the angle between g0 and g1 (we may choose the direction of the angle such that, if the rotation from the oriented g0 to the oriented g1 is anti-clockwise, b is positive). Then the reflection of S(t) by g1, followed by a reflection of S(t) by g0, corresponds to a rotation of S(t) through an angle 2b about the intersection of g0 and g1, and leaves S(t) invariant. By this property of rotation and the arbitrariness of p Î ¶, we know all of the rotations of g are about the same point. Hence S(t) contains a circle and thus must be this circle. Thus we have proved that all of g intersect at the same point, and S(t) is rotational about this point on (t).

Therefore we have proved that S is a surface of revolution (rotational about a vertical line parallel to the -axis).

In the following, we discuss complete (hence compact) smooth surfaces of revolution, about a vertical line, in

2 × of positive constant 2-mean curvature. We will give the parameterized equation of such surfaces.

Given o' Î 2. Then {o'} × is a vertical line. Let S denote a complete surface of revolution about {o'} × , in 2 × (that is to say that S is rotational about the vertical line {o'} × ).

Let p Î 2 × , and let (f,q,t) (0 < f < p, 0 < q < 2p) denote the local coordinate of p where t denotes the -coordinate of and (f,q) denotes the spherical coordinate of 2 about o', that is, f and q are the polar coordinate and azimuthal coordinate respectively. Then p = (sin f cos q, sin f sin q,cos f, t).

The generating curve l(s) of the rotational surface S is l(s) = (f(s),t(s)), where s is the arc-length parameter.

S can be denoted by (sin f(s) cos q,sin f(s) sin q,cos f(s), t(s)).

Take the unit normal N of S as N = (t' cos f cos q, t' cos f sin q, –t' sin f, –f'). Then the two principle curvatures of S are

and the 2-mean curvature of S is

Since s is the arc-length parameter, t'2 + f'2 = 1. Then t't" + f'f" = 0. So

Suppose y is the lowest point of S. We may choose s positive with y = l(0). We have t'(0) = 0. From l2 = t' cot f, we know that f = 0, or p at y. Without loss of generality, we assume that f = 0 at y. This implies y = o' × {ty} for some ty Î .

Since H2 > 0, f ¹ . Hence the domain of f is contained either in [0,) or in (,p]. Since we assume f(0) = 0, we have f Î [0,). This means that S stays in × , where is the open hemisphere of 2 of the center o'.

Since y = l(0) is the lowest point, we have t'(s) > 0 and hence l2(s) > 0 for sufficiently small positive s. But l2 must have the same sign, so t'(s) > 0 for all s. Hence t increases as s increases and S must be embedded.

Suppose z = l(s1) is the highest point S. Also we have f = 0 at z (i.e., z = o' × {tz} for some tz Î .

Hence the domain of s is [0,s1], and f(0) = f(s1) = 0; t'(0) = t'(s1) = 0; f'(0) = 1, f'(s1) = –1.

Since H2 = –f"(cot f), we have f" < 0. This implies f' decreases from f(0) = 1 to f'(s1) = –1. So as s increases from 0 to s1, f first increases from 0 to fmax then decreases from fmax to 0.

By equation (5.3),

where C1 is a constant to be determined by the boundary conditions.

By f(0) = 0 and f'(0) = 1, we have C1 = 1. So

We have fmin = 0, fmax = f|f' = 0 = cos–1exp(–).

Let u = f'(|u| < 1). We will give the equation of the generating curve l.

By equation (5.5), we have

Since u = f', = f" = –, we have

Then

where C is a real constant.

We have

Therefore, we obtain the equations (5.6) and (5.7) of f and t. As u decreases from 1 to –1, t increases from tmin to tmax, and f first increases from 0 to fmax then decreases from fmax to 0. Also S is symmetric about × {t0}.

From the above analysis, we obtain that

PROPOSITION 5.1. Let S be a complete immersed surface of revolution (about a vertical line o' × , o' Î 2) in 2 × with H2 = constant > 0. Then S has the following characters:

(i) S is topologically a sphere and embedded; and

(ii) S stays in × , where denotes the open hemisphere of 2 of the center o'; and

(iii) the generating curve of S can be denoted by l(u) = (f(u), t(u)) with parameter u (–1< u < 1), where t denotes the -coordinate of and f denotes the polar coordinate (about o') of 2, as follows:

where C is a real constant.

The equation of l implies that there exists a horizontal level × {t0}, t0Î such that S divides into two (upper and lower) symmetric parts.

By Theorem 5.1 and Proposition 5.1, we can determine the generating curvature of S satisfying the conditions of Th. 5.1 as follows:

COROLLARY 5.1. Let

2 be the unit sphere in 3 and be the open upper hemisphere in 2. Let S be a complete orientable embedded surface in 2 × with H2 = constant > 0. If S Ì × , then S has the following properties:

The 2-mean curvature H2 satisfies

(where f0Î [0,) denotes the polar angle between the north pole of the open upper hemisphere and o').

Moreover, the generating curve of the rotational surface S can be denoted by l(u) = (f(u), t(u)) where –1< u < 1, and f denotes the polar coordinate (about o') of 2, and:

where C is a real constant.

PROOF. We just need to prove H2 > –. Since S Ì 2 × , f0 + fmax < . Here fmax denotes the maximum of f. Of course, we have fmax = f|f' = 0 = cos–1 exp(–). By it, H2 > –.

REMARK 5.1. We have proved that if S is a complete embedded surface in

2 × Ì 2 × of positive constant 2-mean curvature, S is rotationally symmetric about a vertical line. We conjecture this is true for immersed such S in 2 × (also in 2 × ). This is true if one assumes that the mean curvature is constant instead of positive constant H2 under the condition that the genus of S is zero (Abresch and Rosenberg 2005).

ACKNOWLEDGMENTS

This work was done when the first author was visiting Université Paris 7 (Institut de Mathématiques de Jussieu), supported by a post-doctoral fellowship by Centre Nationale de la Recherche Scientifique (CNRS), France. She would like to thank Université Paris 7 for support and hospitality.

HOFFMAN D, LIRA J AND ROSENBERG H. 2005. Constant mean curvature surfaces in M × R. Trans Amer Math Soc. (In press).

Manuscript received on September 20, 2004; accepted for publication on January 14, 2005; contributed by HAROLD ROSENBERG*

  • ABRESCH U AND ROSENBERG H. 2005. A Hopf differential for constant mean curvature surfaces in S × R and H × R Acta Math. (In press).
  • ALEXANDROV AD. 1962. Uniqueness theorems for surfaces in the large, V. (Russian) Vestnik Leningrad. Univ. 13 1958 19: 5-8; English translation: Amer Math Soc Transl 21: 412-416.
  • BARBOSA JL AND COLARES AG. 1997. Stability of hypersurfaces with constant r-mean curvature. Ann Global Anal Geom 15: 277-297.
  • ELBERT MF. 2002. Constant positive 2-mean curvature hypersurfaces. Illinois. J Math 46: 247-267.
  • HARDY G, LITTLEWOOD J AND POLYA G. 1989. Inequalities, 2nd. Ed., Cambridge, Univ. Press.
  • HOPF H. 1983. Differential geometry in the large, Lecture Notes in Mathematics, 1000. Springer-Verlag, Berlin.
  • KOREVAAR N, KUSNER R, MEEKS W AND SOLOMON B. 1992. Constant mean curvature surfaces in hyperbolic space. Amer J Math 114: 1-43.
  • REILLY R. 1973. Variational properties of functions of the mean curvatures for hypersurfaces in space forms. J Differ Geom 8: 465-477.
  • ROSENBERG H. 1993. Hypersurfaces of constant curvature in space forms. Bull Sci Math 117: 211-239.
  • Correspondence to
    Harold Rosenberg
    E-mail:
  • *
    Member Academia Brasileira de Ciências
  • Publication Dates

    • Publication in this collection
      09 May 2005
    • Date of issue
      June 2005

    History

    • Accepted
      14 Jan 2005
    • Received
      20 Sept 2004
    Academia Brasileira de Ciências Rua Anfilófio de Carvalho, 29, 3º andar, 20030-060 Rio de Janeiro RJ Brasil, Tel: +55 21 3907-8100 - Rio de Janeiro - RJ - Brazil
    E-mail: aabc@abc.org.br