Acessibilidade / Reportar erro

Height estimates and half-space theorems for hypersurfaces in product spaces of the type ×Mn

Abstract

We obtain height estimates and half-space theorems concerning a wide class of hypersurfaces immersed into a product space ×Mn, the so-called generalized linear Weingarten hypersurfaces, which extends that one having some constant higher order mean curvature.

Key words
Product spaces; generalized linear Weingarten hypersurfaces; height estimates; half-space theorems

1 - INTRODUCTION

The last decades have seen a steadily growing interest in the study of a priori estimates for the height function of constant mean curvature compact graphs or, more generally, compact hypersurfaces with boundary having some constant higher order mean curvature. This problem has gained special attention, being considered by several authors probably motivated by the fact that these estimates turn out to be a very useful tool in order to investigate existence and uniqueness results for complete hypersurfaces with constant higher order mean curvature, as well as to obtain information on the topology at infinity of such hypersurfaces (see, for instance, Aledo et al. 2008ALEDO JA, ESPINAR JM & GÁLVEZ JA. 2008. Height estimates for surfaces with positive constant mean curvature in 𝕄2×ℝ. Illinois J Math 52(1): 203–211. , 2010ALEDO JA, ESPINAR JM & GÁLVEZ JA. 2010. The Codazzi equation for surfaces. Adv Math 224(6): 2511-2530., Alías & Dajczer 2007ALÍAS LJ & DAJCZER M. 2007. Constant mean curvature hypersurfaces in warped product spaces. Proc Edinb Math Soc 50: 511–526. , Cheng & Rosenberg 2005, Espinar et al. 2009ESPINAR JM, GÁLVEZ JA & ROSENBERG H. 2009. Complete surfaces with positive extrinsic curvature in product spaces. Comment Math Helv 84(2): 351–386. , Heinz 1969HEINZ E. 1969. On the nonexistence of a surface of constant mean curvature with finite area and prescribed rectifiable boundary. Arch Rational Mech Anal 35: 249–252. , Hoffman et al. 2006HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2×ℝ. Trans Amer Math Soc 358: 491–507. , Korevaar et al. 1992KOREVAAR N, KUSNER R, MEEKS W & SOLOMON B. 1992. Constant Mean Curvature Surfaces in Hyperbolic Space. Amer J Math 114: 1–43. , 1989KOREVAAR N, KUSNER R & SOLOMON B. 1989. The Structure of Complete Embedded Surfaces with Constant Mean Curvature. J Differ Geom 30: 465–503. , Rosenberg 1993ROSENBERG H. 1993. Hypersurfaces of constant curvature in space forms. Bull Sci Math 117: 211–239. ).

A height estimate of compact graphs with positive constant mean curvature in the Euclidean space n+1 and boundary in a hyperplane, were first obtained by Heinz 1969HEINZ E. 1969. On the nonexistence of a surface of constant mean curvature with finite area and prescribed rectifiable boundary. Arch Rational Mech Anal 35: 249–252. . More specifically, denoting by H the mean curvature, Heinz proved that the height of such a graph can rise at most 1/H. More than twenty years after that, Korevaar et al. 1992KOREVAAR N, KUSNER R, MEEKS W & SOLOMON B. 1992. Constant Mean Curvature Surfaces in Hyperbolic Space. Amer J Math 114: 1–43. obtained a sharp bound for compact graphs and for compact embedded hypersurfaces in the hyperbolic space n+1 with nonzero constant mean curvature and boundary in a horosphere. More generally, given an arbitrary Riemannian manifold Mn, height estimates in the product space ×Mn for constant mean curvature compact embedded hypersurfaces with boundary in a slice were exhibited by Hoffman et al. 2006HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2×ℝ. Trans Amer Math Soc 358: 491–507. and Aledo et al. 2008, for n=2, and by Alías & Dajczer 2007ALÍAS LJ & DAJCZER M. 2007. Constant mean curvature hypersurfaces in warped product spaces. Proc Edinb Math Soc 50: 511–526. , for an arbitrary dimension n.

Regarding hypersurfaces having some constant higher order mean curvature, this was done firstly by Rosenberg 1993ROSENBERG H. 1993. Hypersurfaces of constant curvature in space forms. Bull Sci Math 117: 211–239. , who proved height estimates for compact embedded hypersurfaces with zero boundary values either in the Euclidean space or in the hyperbolic space, generalizing the previous estimates of Heinz and Korevaar. Later on, Cheng & Rosenberg 2005CHENG X & ROSENBERG H. 2005. Embedded positive constant r-mean curvature hypersurfaces in Mm×ℝ. An Acad Bras Cienc 77: 183–199. were able to generalize these estimates for compact graphs with some constant higher order mean curvature in the product manifold ×Mn, with boundary in a slice. As application of their height estimates, they used the Alexandrov’s reflection technique to prove that a noncompact properly embedded hypersurface having constant higher order mean curvature in ×Mn, where Mn is a compact manifold with sectional curvature bounded from below, has at least two ends or, equivalently, it cannot lie in a half-space. The same technique was used by Hoffman et al. 2006HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2×ℝ. Trans Amer Math Soc 358: 491–507. in order to obtain some information on the topology at infinity of properly embedded surfaces of constant mean curvature in ×M2. More recently, Rosenberg et al. 2013ROSENBERG H, SCHULZE F & SPRUCK J. 2013. The Half-Space Property and Entire Positive Minimal Graphs in M×ℝ. J Diff Geom 95: 321–336. showed that an entire minimal graph with nonnegative height function in a product space ×Mn, whose base Mn is a complete Riemannian manifold having non-negative Ricci curvature and with sectional curvature bounded from below, must be a slice.

Proceeding with the picture described above, in this paper our aim is to obtain height estimates and half-space theorems of a wide class of hypersurfaces immersed into a product space ×Mn, which extends that one having some constant higher order mean curvature. Precisely, we consider in ×Mn generalized linear Weingarten hypersurfaces, by meaning that there exists a linear relation involving some of the corresponding higher order mean curvatures (for more details, see Section 3). We point out that our results offer improvements of those ones obtained in Alías & Dajczer 2007ALÍAS LJ & DAJCZER M. 2007. Constant mean curvature hypersurfaces in warped product spaces. Proc Edinb Math Soc 50: 511–526. , Alías et al. 2016ALÍAS LJ, MASTROLIA P & RIGOLI M. 2016. Maximum Principles and Geometric Applications. Springer: Monographs in Mathematics, 595 p., Cheng & Rosenberg 2005CHENG X & ROSENBERG H. 2005. Embedded positive constant r-mean curvature hypersurfaces in Mm×ℝ. An Acad Bras Cienc 77: 183–199. and Hoffman et al. 2006HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2×ℝ. Trans Amer Math Soc 358: 491–507. . Furthermore, we are able to prove half-space theorems for complete noncompact generalized linear Weingarten hypersurfaces in ×Mn, generalizing some results of Cheng & Rosenberg 2005CHENG X & ROSENBERG H. 2005. Embedded positive constant r-mean curvature hypersurfaces in Mm×ℝ. An Acad Bras Cienc 77: 183–199. and Hoffman et al. 2006. Recently, the authors proved similar results for the case of hypersurfaces immersed into warped product manifolds (see de Lima & de Lima 2018DE LIMA EL & DE LIMA HF. 2018. Height estimates and topology at infinity of hypersurfaces immersed in a certain class of warped products. Aequat Math 92: 737–761. ). However, as we will see, the results presented there do not contemplate those obtained here.

This manuscript is organized in the following way: In Section 2 we introduce some basic facts and notations that will appear in the proofs of our results. In particular, we recall some geometric conditions which guarantee the ellipticity of the linearized operator of the higher order mean curvature (see Lemmas 2 and 3). In Section 3, we establish our first main results concerning height estimates of compact generalized linear Weingarten hypersurfaces in ×Mn (see Theorems 1 and 2). In Section 4, as application of our height estimates, we prove half-space theorems related to noncompact generalized linear Weingarten hypersurface immersed in ×Mn, supposing that the fiber Mn is compact (see Theorems 3 and 4). Finally, when Mn is not necessarily compact, using a generalized version of the Omori-Yau maximum principle for trace type differential operators, we prove other half-space theorem, which is of independent interest by itself (see Theorem 5).

2 - PRELIMINARIES

In this section we will introduce some basic facts and notations that will appear along the paper. In this sense, along this work we will always consider Mn a (connected) n-dimensional Riemannian manifold and I an open interval in . Let us denote by M¯n+1=I×Mn the product manifold endowed with the Riemannian metric

,=πI*(dt2)+πM*(,M),
where πI and πM denote the canonical projections from I×Mn onto each factor, ,M is the Riemannian metric on the fiber Mn and I is endowed with the metric dt2. Observe that t is a unitary vector field globally defined on M¯n+1=I×Mn, which determines on M¯n+1 a codimension one foliation by totally geodesic slices {t}×M.

Throughout this paper, we will study (connected) two-sided hypersurfaces ψ:ΣnM¯n+1 immersed into the product Riemannian manifold M¯n+1=I×Mn, which means that there exists a unitary normal vector field N globally defined on Σn. As usual, we also denote by , the metric of Σn induced via ψ. In this setting, we consider two particular functions naturally attached to the two-sided hypersurface Σn, namely, the (vertical) height function h=πψ and the angle function Θ=N,t.

Let us denote by A:𝔛(Σ)𝔛(Σ) the shape operator (or Weingarten endomorphism) of Σn in M¯n+1=I×Mn with respect to N, which is given by AX=¯XN, where ¯ stands for the Levi-Civita connection of M¯n+1. A fact well known is that the curvature tensor R of the hypersurface Σn can be described in terms the shape operator A and of the curvature tensor R¯ of the ambient space M¯n+1=I×Mn by the Gauss equation given by

R(X,Y)Z=(R¯(X,Y)Z)+AX,ZAYAY,ZAX(2.1)
for every tangent vector fields X,Y,Z𝔛(Σ), where () denotes the tangential component of a vector field in 𝔛(M¯n+1) along Σn.

Associated with the shape operator A there are n algebraic invariants, which are the elementary symmetric functions Sr of its principal curvatures κ1,,κn, given by

Sr=Sr(κ1,,κn)=i1 irκi1κir,1rn.
As it is well known, the r-mean curvature Hr of the hypersurface Σn is defined by
(nr)Hr=Sr(κ1,,κn).
In particular, when r=1,
H1=1niκi=1ntr(A)=H
is just the mean curvature of Σn. When r=2, H2 defines a geometric quantity which is related to the (intrinsic) scalar curvature S of the hypersurface. For instance, when the ambient space has constant sectional curvature c, it follows from the Gauss equation that S=(n1)(c+H2). In general, it also follows from Gauss equation of the hypersurface that when r is odd Hr is extrinsic (and its sign depends on the chosen orientation), while when r is even Hr is an intrinsic geometric quantity.

It is a classical fact that the higher order mean curvatures satisfy a very useful set of inequalities, usually alluded as Newton’s inequalities. For future reference, we collect them here. A proof can be found in Hardy et al. 1989HARDY G, LITTLEWOOD JE & PÓLYA G. 1989. Inequalities. 2nd ed. Cambridge Mathematical Library. 338 p. .

Lemma 1. Let ψ:ΣnM¯n+1 be a two-sided hypersurface immersed into a product space M¯n+1=I×Mn. For each 1rn, if H1,,Hr are nonnegative on Σn, then:

  • (a) HrHr+2Hr+12;

  • (b) H1H21/2Hr1/r,

and equality holds only at umbilical points.

For each 0rn, one defines the r-th Newton transformation Pr:𝔛(Σ)𝔛(Σ) of the hypersurface Σn by setting P0=I (the identity operator) and, for 1rn, via the recurrence relation

Pr=(nr)HrIAPr1,
Equivalently,
Pr=j=0r(nj)(1)rjHjArj,
so that the Cayley-Hamilton theorem gives Pn=0. Observe also that when r is even, the definition of Pr does not depend on the chosen unitary normal vector field N, but when r is odd there is a change of sign in the definition of Pr. Moreover, it is easy to see that each Pr is a self-adjoint operator which commutes with shape operator A, that is, if a local orthonormal frame on Σn diagonalizes A, then it also diagonalizes each Pr. More specifically, if {E1,,En} is such a local orthonormal frame with A(Ei)=κiEi, then
Pr(Ei)=μi,rEi,
where
μi,r=i1 ir,ijiκi1κir.
It follows from here that for each 0rn1, we have
tr(Pr)=crHr,withcr=(nr)(nr)=(r+1)(nr+1).

Let stand for the Levi-Civita connection of the two-sided hypersurface Σn. Associated to each Newton transformation Pr, one has the second order linear differential operator Lr:C(Σ)C(Σ) for r=0,1,,n1, defined by

Lru=tr(Prhess u),
where hess u:𝔛(Σ)𝔛(Σ) denotes the self-adjoint linear operator metrically equivalent to the Hessian of u, Hess u, which are given by
hess u(X)=XuandHess (X,Y)=hess u(X),Y,
respectively, for all X,Y𝔛(Σ). In particular, L0=Δ, the Laplacian of Σn, which is always an elliptic operator in divergence form. More generally, it is well known that the operator Lr is elliptic if and only if Pr is positive definite.

For our applications, it will be useful to have some geometric conditions which guarantee the ellipticity of Lr when r1. For r=1, the next lemma assures the ellipticity of L1 (see Lemma 3.10 of Elbert 2002ELBERT MF. 2002. Constant positive 2-mean curvature hypersurfaces. Illinois J Math 46: 247–267. ).

Lemma 2. Let ψ:ΣnM¯n+1 be a two-sided hypersurface immersed into a product space M¯n+1=I×Mn. If H20 on Σ, then L1 is elliptic or, equivalently, P1 is positive definite (for an appropriate choice of the orientation N).

When r2, the following lemma give us sufficient conditions to guarantee the ellipticity of Lr. The proof is given in Proposition 3.2 of Cheng & Rosenberg 2005CHENG X & ROSENBERG H. 2005. Embedded positive constant r-mean curvature hypersurfaces in Mm×ℝ. An Acad Bras Cienc 77: 183–199. (see also Proposition 3.2 of Barbosa & Colares 1997BARBOSA JLM & COLARES AG. 1997. Stability of hypersurfaces with constant r-mean curvature. Ann Global Anal Geom 15: 277–297. ).

Lemma 3. Let ψ:ΣnM¯n+1 be a two-sided hypersurface (with or without boundary) immersed into a product space M¯n+1=I×Mn with Hr+10 on Σn, for some 2rn1. If there exists an interior point p of Σn such that all the principal curvatures at p are nonnegative, then for all 1kr the operator Lk is elliptic and the (k+1)-mean curvature Hk+1 is positive.

Next, we close this section with the following formulas, which will be essential for the proofs of our main results (for more details see Proposition 6 and Lemma 27 of Alías et al. 2013ALÍAS LJ, IMPERA D & RIGOLI M. 2013. Hypersurfaces of constant higher order mean curvature in warped products. Trans Am Math Soc 365(2): 591-621. ).

Proposition 1. Let ψ:ΣnM¯n+1 be a two-sided hypersurface immersed into a product space M¯n+1=I×Mn. For every r=0,,n1:

  • (a) The height function satisfies

    Hess h(X,Y)=ΘAX,Yand Lrh=crΘHr+1,
    wherecr:=(nr)(nr)=(r+1)(nr+1).

  • (b) The angle function satisfies

    LrΘ=crr+1Hr+1,hcrΘr+1(nH1Hr+1(nr1)Hr+2)Θi=1nμi,rKM(N*,Ei*)|N*Ei*|2,
    where{E1,,En}is an orthonormal frame onΣndiagonalizingA, KMdenotes the sectional curvature of the fiberMn, μi,rstands for the eigenvalues ofPkand, for every vector fieldX𝔛(M¯), X*is the orthogonal projection onTM.

3 - HEIGHT ESTIMATES OF GENERALIZED LINEAR WEINGARTEN HYPERSURFARCES

This section is devoted to establish our results concerning to estimates of the height function h of a wide class of two-sided hypersurfaces immersed into a product Riemannian manifold M¯n+1=×Mn, which extends that one having some constant higher order mean curvature. Specifically, let us consider ψ:ΣnM¯n+1 a two-sided hypersurface immersed into a product space M¯n+1=×Mn. We say that Σn is (r,s)-linear Weingarten, for some 0rsn1, if there exist nonnegative real numbers br,,bs (at least one of them nonzero) such that the following linear relation holds on Σn:

k=rsbkHk+1=d.(3.1)

Thus, naturally attached to a (r,s)-linear Weingarten two-sided hypersurface we have the constant d given by (3.1). We note that the (r,r)-linear Weingarten two-sided hypersurfaces are exactly the two-sided hypersurfaces having d=Hr+1 constant. On the other hand, if the ambient space has zero sectional curvature and taking into account that in this case S¯=H2, where S¯ stands for the normalized scalar curvature of Σn, we observe that the (0,1)-linear Weingarten two-sided hypersurfaces are called simply linear Weingarten two-sided hypersurfaces. Throughout this paper, we will always denote by d the constant given by equation (3.1).

Now, we are in position to state and prove our first main result. More precisely, we will establish an estimate for the height function concerning (r,s)-linear Weingarten two-sided hypersurface in a product space ×Mn.

Theorem 1. Let M¯n+1=×Mn be a product space whose the fiber Mn has nonnegative sectional curvature KM. Let ψ:ΣnM¯n+1 be a compact (r,s)-linear Weingarten two-sided hypersurface with (s+1)-mean curvature Hs+10 on Σn, for some 0sn1, and boundary Σn contained into the slice {t0}×Mn for some t0. Suppose that the angle function Θ does not change sign on Σn. Then,

  • (a) Either maxht0 and

    Σ n [ t 0 , t 0 + 1 min | H 1 | ] × M n ,

  • (b) or minht0 and

    Σ n [ t 0 1 min | H 1 | , t 0 ] × M n .

Proof. First of all it is clear from our hypothesis on the (s+1)-mean curvature that either maxht0 or minht0. So, we begin by assuming that maxht0 and let us choose an interior point p0 of Σn such that the height function reaches its maximum and the orientation so that Θ0. Then, Proposition 1 yields

0Hess h(p0)(v,v)=Θ(p0)Av,v(p0),vTp0Σ,
that is, at p0 all the principal curvatures are nonnegative. Since we are assume that Hs+10 on Σn, we must have Hs+10 on Σn. In particular, we can apply Lemma 3 (or Lemma 2 if s=1) to guarantee the ellipticity of the operator Lk for every k=r,,s and Hk+1 is positive on Σn for every 0ks. So, for instance, we have
Lsh=csΘHs+10
and, consequently, by the weak maximum principle we obtain that ht0 on Σn.

Now let us consider on Σn the smooth function φ=ch+Θ, where c is a positive constant to be chosen in an appropriate way. Then Proposition 1 gives

Lkφ=ckk+1Hk+1,hΘ(nk+1)(nH1Hk+1(nk1)Hk+2(k+1)cHk+1)Θi=1nμi,kKM(N*,Ei*)|N*Ei*|2,(3.2)
where {E1,,En} is an orthonormal frame on Σn diagonalizing A with PkEi=μi,kEi, for every i=1,,n and k=r,,s, and X* denotes the orthogonal projection on TM.

Since Hk+1 is positive for every k=0,,s, from Lemma 1 we get

H1Hk+1Hk+2H1Hk+1Hk+12Hk1=Hk+1Hk(H1HkHk+1).
By using once more Lemma 1 it follows from here that
H1Hk+1Hk+2Hk+1Hk(H1HkHk(k+1)/k)=Hk+1(H1Hk1/k)0.
Then, the previous inequality implies that
nH1Hk+1(nk1)Hk+2(k+1)cHk+1=(k+1)Hk+1(H1c)+(nk1)(H1Hk+1Hk+2)(k+1)Hk+1(H1c)0,(3.3)
provided that c:=minH1.

On the other hand, since the operator Lk is elliptic for every k=r,,s or, equivalently, Pk is positive definite, we get that its eigenvalues μi,k are all positive on Σn. Then, by our assumption on the sectional curvature KM of the fiber Mn we have

i=1nμi,kKM(N*,Ei*)|N*Ei*|20,
Hence, by using (3.3) and taking into account that Θ0, we infer from (3.2) and the previous inequality that
Lkφckk+1Hk+1,h.(3.4)

Proceeding, we introduce the following second order linear differential operator L:C(Σ)C(Σ) defined by

L=k=rs(k+1)ck1bkLk=tr(Phess),
where the tensor P:𝔛(Σ)𝔛(Σ) is given by
P=k=rs(k+1)ck1bkPk.
Since (k+1)ck1bk0 for every k=r,,s and each operator Lk is elliptic (equivalently, each Pk is positive definite) we see that the operator P is positive definite and, consequently, the operator L is elliptic too. So, equation (3.4) and the fact that Σn is (r,s)-linear Weingarten imply that
Lφ0.
By using once more the weak maximum principle for the elliptic operator L we get
φmaxΣφct0,
that is,
c(ht0)1.
Therefore, we conclude that
ht0+1minH1.
This proves (a).

In the case minht0, we choose an interior point q0 of Σn satisfying minh=h(q0) and the orientation so that Θ0. Then,

0Hess h(q0)(v,v)=Θ(q0)Av,v(q0),vTq0Σ,
that is, at q0 all the principal curvatures must be nonnegative. So, reasoning as in the previous case we see that each operator Lk is elliptic for every k=r,,s, Hk+1 is positive on Σn for every 0ks and ht0 on Σn.

Besides, keeping the notation of case (a), it follows that φ=ch+Θ satisfies, by equations (3.2), (3.3) and our assumption on KM,

Lkφckk+1Hk+1,h,
which implies that Lφ0. Therefore, by weak maximum principle we conclude that
φminΣφct0,
that is,
ht01minH1.
This finishes the proof of the theorem. ◻

Remark 1. We observe that the estimate given in Theorem 1 is sharp in the sense that it is reached by the hemisphere Σ+={x𝕊n;x10} of the standard sphere 𝕊n in n+1. Indeed, it follows easily that Σ+ is a totally umbilical hypersurface (in fact, it is a vertical graph) with H1=1, boundary {0}×𝕊n1{0}×n and has the maximum height 1.

We observe that Theorem 1 above does not follow from Theorem 1 of de Lima & de Lima 2018, because there the authors just assume that Hs+10 and, with this constraint, it is not possible to obtain item (b) above.

Let us also point out that when s=r, that is, the hypersurface has constant (r+1)-mean curvature Hr+1, Theorem 1 improves the estimate obtained in Theorem 4.1(i) of Cheng & Rosenberg 2005. Indeed, it is easy to see that the inequality

1minH11Hr+11/(r+1)
holds for every r=0,,n1. Moreover, this result is also an extension of Theorem 3.5 of Alías & Dajczer 2007 (case α=0) and Proposition 1 of Hoffman et al. 2006HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2×ℝ. Trans Amer Math Soc 358: 491–507. (case τ=0).

It is still worth pointing out that the same argument done in the proof of Theorem 4.2 of Cheng & Rosenberg 2005CHENG X & ROSENBERG H. 2005. Embedded positive constant r-mean curvature hypersurfaces in Mm×ℝ. An Acad Bras Cienc 77: 183–199. by using the Alexandrov reflection technique, enable us to get the following consequence of Theorem 1 concerning compact embedded (r,s)-linear Weingarten hypersurfaces:

Corollary 1. Let M¯n+1=×Mn be a product space whose the fiber Mn has nonnegative sectional curvature KM. Let ψ:ΣnM¯n+1 be a compact embedded (r,s)-linear Weingarten two-sided hypersurface with (s+1)-mean curvature Hs+10 on Σn, for some 0sn1. Suppose that the angle function Θ does not change sign on Σn. Then Σn is symmetric about some slice {t0}×Mn, t0, and the extrinsic vertical diameter of Σn is no more than 2min|H1|.

Proceeding, we are able to relax the assumption on the sectional curvature KM of the fiber Mn letting it be bounded from below by a negative constant. For this, we will assume that the mean curvature satisfies a certain condition, which holds automatically when the sectional curvature of the fiber is nonnegative. In what follows, we will denote by c=minH1. So, we get the following result.

Theorem 2. Let M¯n+1=×Mn be a product space whose the fiber Mn has sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a compact (r,s)-linear Weingarten two-sided hypersurface with (s+1)-mean curvature Hs+10 on Σn, for some 0sn1, and boundary Σn contained into the slice {t0}×Mn for some t0. Suppose that the angle function Θ does not change sign on Σ and c(r+1)minHk+1α(s+1)maxHk for every k=r,,s. Then,

  • (a) Either maxht0 and

    Σn[t0,t0+(r+1)d(r+1)dc(s+1)αβ]×Mn,
    wheredis given by (3.1) and β=k=rsbkmaxHk.

  • (b) or minht0 and

    Σn[t0(r+1)d(r+1)dc(s+1)αβ,t0]×Mn,
    whered is given by (3.1) andβ=k=rsbkmaxHk.

Remark 2. We note that in the case of hypersurfaces having constant (r+1)-mean curvature Hr+1, our assumption on c in Theorem 2 becomes cHr+1αmaxHr. In particular, it is weaker than assumption (7.77) of Theorem 7.19 of Alías et al. 2016ALÍAS LJ, MASTROLIA P & RIGOLI M. 2016. Maximum Principles and Geometric Applications. Springer: Monographs in Mathematics, 595 p.. Moreover, the constant (r+1)d(r+1)dc(s+1)αβ is just given by Hr+1cHr+1αmaxHr. Furthermore, by Lemma 1 we have cHr+11/(r+1), which implies that

Hr+1cHr+1αmaxHrHr+1Hr+1(r+2)/(r+1)αmaxHr.
In this setting, our estimate improves that one given in Theorem 7.19 of Alías et al. 2016ALÍAS LJ, MASTROLIA P & RIGOLI M. 2016. Maximum Principles and Geometric Applications. Springer: Monographs in Mathematics, 595 p. for the compact case.

On the other hand, we also observe that, since KM can be negative in Theorem 2, it does not follow from Theorem 1 of de Lima & de Lima 2018.

Proof of Theorem 2. In what follows, we keep the notations established in Theorem 1. Let us suppose maxht0 first. Then, as in the previous result, taking the angle function Θ nonpositive, it is easy to see that the operator Lk is elliptic for every k=r,,s, the (k+1)-mean curvature Hk+1 is positive for every 0ks and ht0. Besides, by equations (3.4) and (3.3) we get that the function φ defined in Theorem 1 satisfies

Lkφckk+1Hk+1,hΘi=1nμi,kKM(N*,Ei*)|N*Ei*|2.(3.5)
Since the eigenvalues μi,k are all positive on Σn and using our assumption on KM we have
μi,kKM(N*,Ei*)|N*Ei*|2μi,kα|N*Ei*|2,(3.6)
for every i=1,,n and k=r,,s. With a straightforward computation, we can show that
|N*Ei*|2=|N*|2|Ei*|2N*,Ei*2=|h|2Ei,h21,
which jointly with (3.6) imply that
i=1nμi,kKM(N*,Ei*)|N*Ei*|2αtr Pk=αckHkαckmaxHk.
From here and (3.5) we infer that
Lφk=rs(k+1)αΘbkmaxHk(s+1)αβΘ,(3.7)
where β=k=rsbkmaxHk. On the other hand, by using Proposition 1 we get that
Lh=k=rs(k+1)ck1bkLkh=k=rs(k+1)ΘbkHk+1(r+1)dΘ.(3.8)

So, let us consider on Σn the smooth function given by

φ̃=φ(s+1)αβ(r+1)dh=(r+1)dc(s+1)αβ(r+1)dh+Θ.
Then, equations (3.7) and (3.8) yield
Lφ̃=Lφ(s+1)αβ(r+1)dLh(s+1)αβΘ(s+1)αβΘ=0.
Hence, we can apply once more the weak maximum principle to conclude that
φ̃maxΣφ̃(r+1)dc(s+1)αβ(r+1)dt0,
that is,
(r+1)dc(s+1)αβ(r+1)d(ht0)1.(3.9)
Now, the assumption on c gives
(r+1)dc(s+1)αβ=(r+1)ck=rsbkHk+1(s+1)αk=rsbkmaxHk=k=rsbk((r+1)cHk+1(s+1)αmaxHk)0.
Therefore, from equation (3.9) we arrive to
ht0+(r+1)d(r+1)dc(s+1)αβ
as desired.

Finally, the case minht0 follows as above and this finishes the proof of the theorem. ◻

As consequence of Theorem 2, the analogue of Corollary 1 also holds in this situation:

Corollary 2. Let M¯n+1=×Mn be a product space whose the fiber Mn has sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a compact embedded (r,s)-linear Weingarten two-sided hypersurface with (s+1)-mean curvature Hs+10 on Σn, for some 0sn1. Suppose that the angle function Θ does not change sign on Σ and c(r+1)minHk+1α(s+1)maxHk for every k=r,,s. Then Σn is symmetric about some slice {t0}×Mn, t0, and the extrinsic vertical diameter of Σn is no more than 2(r+1)d(r+1)dc(s+1)αβ.

4 - HALF-SPACE THEOREMS

The aim of this section is to give nonexistence results, in the form of half-space theorems, concerning complete two-sided hypersurfaces in the product Riemannian manifold ×Mn. We point out that our results do not assume that some higher order mean curvature of the hypersurface is constant. In this setting, when the fiber Mn is compact, our results generalize those one obtained by Cheng & Rosenberg 2005 and Hoffman et al. 2006HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2×ℝ. Trans Amer Math Soc 358: 491–507. for the case in which the mean curvature or some higher order mean curvature is constant (see Theorems 3 and 4 below). Moreover, in the case in which Mn is not necessarily compact, by using a generalized version of the Omori-Yau maximum principle for trace type differential operators, we prove other interesting half-space theorem (see Theorem 5 below).

According to Hoffman et al. 2006HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2×ℝ. Trans Amer Math Soc 358: 491–507. , we say that a two-sided hypersurface in a product space ×Mn lies in an upper or lower half-space if it is, respectively, contained in a region of ×Mn of the form

[a,+)×Mnor(,a]×Mn,
for some real number a.

As an application of Theorem 1 we get the following result:

Theorem 3. Let M¯n+1=×Mn be a product space whose fiber Mn is compact and has nonnegative sectional curvature. Let ψ:ΣnM¯n+1 be a noncompact (r,s)-linear Weingarten two-sided properly immersed hypersurface with (s+1)-mean curvature bounded away from zero, for some 0sn1, and such that its angle function does not change sign. Then, Σn cannot lie in a half-space.

Proof. Let us assume by contradiction that Σn lies in an upper half-space, that is, Σn[a,+)×Mn, for some a. For any number t0a, we denote by Σt0 the hypersurface

Σt0={(t,p)Σn;tt0}.
Then, Σt0 is a compact (r,s)-linear Weingarten two-sided hypersurface with boundary contained into the slice {t0}×M and minht0, because Mn is compact and the immersion is proper. Hence, by Theorem 1 we must have Hs+10 on Σt0 and Σt0[t01c(t0),t0]×Mn, where c(t0)=minΣt0H10, that is,
t0a1c(t0).
Because Hs+1 is bounded away from zero we get infHs+10, which implies infH10. Thus
t0a1c(t0)1infH1.
Then choosing t0 large enough we reached a contradiction.

Finally, if Σn is contained into a lower half-space, we may apply the same argument above to arrive at a contradiction. ◻

Similarly, we can reason as in Theorem 3 to obtain as consequence of Theorem 2 the following result, where we keep the notation c=minH1.

Theorem 4. Let M¯n+1=×Mn be a product space whose fiber Mn is compact with sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a noncompact (r,s)-linear Weingarten two-sided properly immersed hypersurface with bounded away from zero (s+1)-mean curvature, for some 0sn1, and such that its angle function does not change sign. Suppose that c(r+1)minHk+1α(s+1)maxHk for every k=r,s. Then, Σn cannot lie in a half-space.

In order to treat the case in which the fiber is not compact, we will make use of a generalized version of the Omori-Yau maximum principle for trace type differential operators proved in Alías et al. 2016. Let Σn be a Riemannian manifold and let =tr(𝒫hess) be a semi-elliptic operator, where 𝒫:𝔛(Σ)𝔛(Σ) is a positive semi-definite symmetric tensor. Following the terminology introduced by Pigola et al. 2005PIGOLA S, RIGOLI M & SETTI AG. 2005. Maximum principles on Riemannian manifolds and applications. Mem American Math Soc 174(822): 109. , we say that the Omori-Yau maximum principle holds on Σn for the operator if, for any function uC2(Σ) with u*=supu +, there exists a sequence of points (pj)Σn satisfying

u(pj)u*1j,|u(pj)| 1jandu(pj) 1j
for every j. Equivalently, for any smooth function uC2(Σ) with u*=infu there exists a sequence of points (pj)Σn satisfying
u(pj) u*+1j,|u(pj)| 1jandu(pj)1j
for every j.

We quote a suitable version of the Omori-Yau maximum principle for trace type differential operators on a complete noncompact Riemannian manifold (see Theorem 6.13 of Alías et al. 2016ALÍAS LJ, MASTROLIA P & RIGOLI M. 2016. Maximum Principles and Geometric Applications. Springer: Monographs in Mathematics, 595 p.).

Lemma 4. Let Σn be a complete noncompact Riemannian manifold; let oΣn be a reference point and denote by ro the Riemannian distance function from o. Assume that the sectional curvature of Σn satisfies

KΣG2(ro),(4.1)
with GC1([0,+)) satisfying
G(0)0,G(t)0and 1G(t)L1(+).(4.2)
Let 𝒫 be a positive semi-definite symmetric tensor on Σn. If suptr(𝒫) +, then the Omori-Yau maximum principle holds on Σn for the semi-elliptic operator =tr(𝒫hess).

In particular, Lemma 4 remains true if we replace condition (4.1) by the stronger condition of Σn having sectional curvature bounded from below by a constant.

Remark 3. As it is well known, especially significant examples of functions G satisfying the condition (4.2) in Lemma 4 are given by (see, for instance Alías et al. 2016ALÍAS LJ, MASTROLIA P & RIGOLI M. 2016. Maximum Principles and Geometric Applications. Springer: Monographs in Mathematics, 595 p. and Pigola et al. 2005PIGOLA S, RIGOLI M & SETTI AG. 2005. Maximum principles on Riemannian manifolds and applications. Mem American Math Soc 174(822): 109. )

G(t)=tj=1Nlogj(t),t1,
where logj stands for the j-th iterated logarithm.

Now, we are in ready to state and prove our last half-space theorem.

Theorem 5. Let M¯n+1=×Mn be a product space whose fiber Mn has sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a complete noncompact (r,s)-linear Weingarten two-sided hypersurface with positive (s+1)-mean curvature, for some 1rsn1. Suppose that sup|Hr| + and, if s2, there exists an elliptic point in Σn. Assume further that the shape operator satisfies |A|G(ro), where GC1([0,+)) satisfies (4.2) and ro is the distance function from a reference point of Σn. The following holds:

  • (a) either supΘ0 or Σn cannot lie in an upper half-space;

  • (b) either infΘ 0 or Σn cannot lie in a lower half-space.

Proof. We begin stating that the sectional curvature KΣ of Σ satisfies the assumption (4.1) of Lemma 4. Indeed, denoting by K¯ the sectional curvature of the ambient space, it follows from Gauss equation (2.1) that if {X,Y} is an orthonormal basis for an arbitrary plane tangent to Σn, then

KΣ(X,Y)=K¯(X,Y)+AX,XAY,YAX,Y2K¯(X,Y)|AX||AY||AX|2K¯(X,Y)2|A|2,(4.3)
where the last inequality follows from the fact that
|AX|2tr(A2)|X|2=|A|2
for every unitary vector X tangent to Σn. Taking into account that
K¯(X,Y)=KM(X*,Y*)|X*Y*|2
we obtain from our hypothesis on KM that K¯(X,Y)α, because |X*Y*|2|XY|21. Hence, since the shape operator satisfies |A|G(ro), equation (4.3) yields
KΣα2G2(ro),
which concludes the claim.

We prove part (a) first. To do this, we assume that Θ0 and argue by contradiction, that is, we suppose that Σn lies in an upper half-space. Equivalently, the height function of Σn satisfies h*=infh.

We set the second order linear differential operator :C(Σ)C(Σ) by

=k=rsck1bkLk=tr (𝒫hess),
where the tensor 𝒫:𝔛(Σ)𝔛(Σ) is given by
𝒫=k=rsck1bkPk.
Since Σn has an elliptic point (that is, all the principal curvatures are positive in such a point), the operator Lk is elliptic for every k=r,,s or, equivalently, Pk is positive definite. Then, 𝒫 is a positive linear combination of the Pk’s, so that it is positive definite. Thus, is a trace type elliptic operator. Besides, by using the identity tr(Pk)=ckHk we obtain from Lemma 1 that
tr(𝒫)=k=rsbkHkk=rsbkHrk/r,
which implies that suptr(𝒫) +. Hence, we are ready to apply Lemma 4 to guarantee that the Omori-Yau maximum principle holds on Σn for the operator . Then, there exists a sequence of points (pj)Σn having the following properties:
limh(pj)=h,|h(pj)|1jandLh(pj)1j.
In particular, by Proposition 1 we get
1j h(pj)=k=rsΘ(pj)bkHk+1(pj)=dΘ(pj).
Since |h|2=1Θ2, we see that Θ(pj)1. So, taking limits we conclude that d0, which gives a contradiction.

In case (b), we reason again by contradiction, that is, by assuming that Θ0 and Σn is contained into a lower half-space so that the height function satisfies h*=suph +. Then, reasoning as in part (a), it is not difficult to see once more that d0, characterizing a contradiction. This concludes the proof of the theorem. ◻

Let us observe that the proof of Theorem 5 remains true with the stronger assumption that KΣ is bounded from below by a constant, which implies the validity of the Omori-Yau’s maximum principle. For instance, reasoning as in the proof of Theorem 5 we see that KΣ is bounded from below since sup|A|2 +. On the other hand, the hypothesis on Hr in Theorem 5, sup|Hr| , can be replaced by supH1 , because of Lemma 1. In this case, taking into account the relation

|A|2=n2H12n(n1)H2,
it follows that the condition sup|A|2 + is equivalent to supH1 +. This proves the following result:

Corollary 3. Let M¯n+1=×Mn be a product space whose fiber Mn has sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a complete noncompact (r,s)-linear Weingarten two-sided hypersurface with positive (s+1)-mean curvature, for some 1rsn1. Suppose that sup|H1| + and, if s2, there exists an elliptic point in Σn. The following holds:

  • (a) either supΘ0 or Σn cannot lie in an upper half-space;

  • (b) either infΘ 0 or Σn cannot lie in a lower half-space.

In the case of hypersurfaces having constant mean curvature the assumption of existence of an elliptic point can be dropped as follows.

Corollary 4. Let M¯n+1=×Mn be a product space whose fiber Mn has sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a complete noncompact two-sided hypersurface with positive constant mean curvature and such that infH2. The following holds:

  • (a) either supΘ0 or Σn cannot lie in an upper half-space;

  • (b) either infΘ 0 or Σn cannot lie in a lower half-space.

In other words we have:

  • (a’) There is no complete noncompact two-sided hypersurface having positive constant mean curvature, angle function nonpositive and contained into an upper half-space;

  • (b’) There is no complete noncompact two-sided hypersurface having positive constant mean curvature, angle function nonnegative and contained into a lower half-space.

Proof. It is enough to prove part (a). We suppose that Θ0 and let us reason by contradiction, that is, infh=h*. As in the proof of Theorem 5 and by remark above, we might see that the Omori-Yau maximum principle holds on Σn for the Laplacian. Then, there is a sequence of points (pj)Σn satisfying

limh(pj)=h,|h|1jandΔ h(pj)1j.
By applying Proposition 1 we find
1j Δh(pj)=nΘ(pj)H1
Since the angle function is nonpositive, taking limits here we conclude that H10, which gives a contradiction. ◻

More generally, for hypersurfaces having some constant higher order mean curvature we get the following result:

Corollary 5. Let M¯n+1=×Mn be a product space whose fiber Mn has sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a complete noncompact two-sided hypersurface with positive constant (r+1)-mean curvature, for some 1rn1. Suppose that supH1 + and, if r2, there exists an elliptic point in Σn. The following holds:

  • (a) either supΘ0 or Σn cannot lie in an upper half-space;

  • (b) either infΘ 0 or Σn cannot lie in a lower half-space.

In other words we have:

  • (a’) there is no complete noncompact two-sided hypersurface having Hr+10, an elliptic point, with supH1 +, angle function nonpositive and contained into an upper half-space;

  • (b’) there is no complete noncompact two-sided hypersurface having Hr+10, an elliptic point, with supH1 +, angle function nonnegative and contained into a lower half-space.

Finally we collect (a) and (b) in the previous corollaries in order to obtain the following result.

Corollary 6. Let M¯n+1=×Mn be a product space whose fiber Mn has sectional curvature satisfying KMα, for some positive constant α. Let ψ:ΣnM¯n+1 be a complete noncompact two-sided hypersurface with positive constant (r+1)-mean curvature, for some 0rn1. Suppose that supH1 + and, if r2, there exists an elliptic point in Σn. In addition, if r=0 assume that infH2. Then, either Θ does not vanishes identically or Σn cannot lie in a half-space. In other words, there is no complete noncompact two-sided hypersurface having Hr+10, an elliptic point, with supH1 +, angle function vanishes identically and contained into a half-space.

ACKNOWLEDGMENTS

The second author is partially supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Brazil, grant 301970/2019-0. The authors would like to thank the referee for reading the manuscript in great detail and for his/her valuable suggestions and useful comments which improved the paper.

  • ALEDO JA, ESPINAR JM & GÁLVEZ JA. 2008. Height estimates for surfaces with positive constant mean curvature in 𝕄2× Illinois J Math 52(1): 203–211.
  • ALEDO JA, ESPINAR JM & GÁLVEZ JA. 2010. The Codazzi equation for surfaces. Adv Math 224(6): 2511-2530.
  • ALÍAS LJ & DAJCZER M. 2007. Constant mean curvature hypersurfaces in warped product spaces. Proc Edinb Math Soc 50: 511–526.
  • ALÍAS LJ, IMPERA D & RIGOLI M. 2013. Hypersurfaces of constant higher order mean curvature in warped products. Trans Am Math Soc 365(2): 591-621.
  • ALÍAS LJ, MASTROLIA P & RIGOLI M. 2016. Maximum Principles and Geometric Applications. Springer: Monographs in Mathematics, 595 p.
  • BARBOSA JLM & COLARES AG. 1997. Stability of hypersurfaces with constant r-mean curvature. Ann Global Anal Geom 15: 277–297.
  • CHENG X & ROSENBERG H. 2005. Embedded positive constant r-mean curvature hypersurfaces in Mm× An Acad Bras Cienc 77: 183–199.
  • DE LIMA EL & DE LIMA HF. 2018. Height estimates and topology at infinity of hypersurfaces immersed in a certain class of warped products. Aequat Math 92: 737–761.
  • ELBERT MF. 2002. Constant positive 2-mean curvature hypersurfaces. Illinois J Math 46: 247–267.
  • ESPINAR JM, GÁLVEZ JA & ROSENBERG H. 2009. Complete surfaces with positive extrinsic curvature in product spaces. Comment Math Helv 84(2): 351–386.
  • HARDY G, LITTLEWOOD JE & PÓLYA G. 1989. Inequalities. 2nd ed. Cambridge Mathematical Library. 338 p.
  • HEINZ E. 1969. On the nonexistence of a surface of constant mean curvature with finite area and prescribed rectifiable boundary. Arch Rational Mech Anal 35: 249–252.
  • HOFFMAN D, LIRA J & ROSENBERG H. 2006. Constant mean curvature surfaces in M2× Trans Amer Math Soc 358: 491–507.
  • KOREVAAR N, KUSNER R, MEEKS W & SOLOMON B. 1992. Constant Mean Curvature Surfaces in Hyperbolic Space. Amer J Math 114: 1–43.
  • KOREVAAR N, KUSNER R & SOLOMON B. 1989. The Structure of Complete Embedded Surfaces with Constant Mean Curvature. J Differ Geom 30: 465–503.
  • PIGOLA S, RIGOLI M & SETTI AG. 2005. Maximum principles on Riemannian manifolds and applications. Mem American Math Soc 174(822): 109.
  • ROSENBERG H. 1993. Hypersurfaces of constant curvature in space forms. Bull Sci Math 117: 211–239.
  • ROSENBERG H, SCHULZE F & SPRUCK J. 2013. The Half-Space Property and Entire Positive Minimal Graphs in M× J Diff Geom 95: 321–336.

Publication Dates

  • Publication in this collection
    03 Dec 2021
  • Date of issue
    2021

History

  • Received
    20 May 2019
  • Accepted
    8 Feb 2020
Academia Brasileira de Ciências Rua Anfilófio de Carvalho, 29, 3º andar, 20030-060 Rio de Janeiro RJ Brasil, Tel: +55 21 3907-8100, CLOCKSS system has permission to ingest, preserve, and serve this Archival Unit - Rio de Janeiro - RJ - Brazil
E-mail: aabc@abc.org.br