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

Hypersurfaces of Constant Higher Order
Mean Curvature in M×M\times\mathbb{R}

R. F. de Lima, F. Manfio and J. P. dos Santos Departamento de Matemática - Universidade Federal do Rio Grande do Norte [email protected] ICMC–Universidade de São Paulo [email protected] Departamento de Matemática - Universidade de Brasília [email protected]
Abstract.

We consider hypersurfaces of products M×M\times\mathbb{R} with constant rr-th mean curvature Hr0H_{r}\geq 0 (to be called HrH_{r}-hypersurfaces), where MM is an arbitrary Riemannian nn-manifold. We develop a general method for constructing them, and employ it to produce many examples for a variety of manifolds M,M, including all simply connected space forms and the hyperbolic spaces 𝔽m\mathbb{H}_{\mathbb{F}}^{m} (rank 11 symmetric spaces of noncompact type). We construct and classify complete rotational Hr(0)H_{r}(\geq 0)-hypersurfaces in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} and in 𝕊n×\mathbb{S}^{n}\times\mathbb{R} as well. They include spheres, Delaunay-type annuli and, in the case of 𝔽m×,\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, entire graphs. We also construct and classify complete Hr(0)H_{r}(\geq 0)-hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which are invariant by either parabolic isometries or hyperbolic translations. We establish a Jellett-Liebmann-type theorem by showing that a compact, connected and strictly convex HrH_{r}-hypersurface of n×\mathbb{H}^{n}\times\mathbb{R} or 𝕊n×\mathbb{S}^{n}\times\mathbb{R} (n3)(n\geq 3) is a rotational embedded sphere. Other uniqueness results for complete HrH_{r}-hypersurfaces of these ambient spaces are obtained.

Key words and phrases:
higher order mean curvature – rr-minimal – product space.
2010 Mathematics Subject Classification:
53B25 (primary), 53C24, 53C42 (secondary).

1. Introduction

In his pioneering work [34], H. Rosenberg initiated the study of minimal and constant mean curvature hypersurfaces of product spaces M×,M\times\mathbb{R}, where MM is an arbitrary Riemannian nn-manifold. Since then, many results on this subject have been obtained by many authors, and considerably understanding of the geometry of these hypersurfaces has been achieved, mostly in the particular case MM is a simply connected space form.

Following this path, we approach here hypersurfaces of M×M\times\mathbb{R} with constant rr-th mean curvature Hr0H_{r}\geq 0 (for some r{1,,n}r\in\{1,\dots,n\}), which we call HrH_{r}-hypersurfaces. Let us recall that the (non normalized) rr-th mean curvature HrH_{r} of a hypersurface is the rr-th elementary symmetric polynomial of its principal curvatures, so that it constitutes a natural extension of the mean curvature (r=1r=1) and the Gauss-Kronecker curvature (r=nr=n).

We focus on constructing and classifying HrH_{r}-hypersurfaces of products M×.M\times\mathbb{R}. To this end, we use a special type of graph built on families of parallel hypersurfaces of M.M. In fact, for any given constant Hr0,H_{r}\geq 0, we obtain HrH_{r}-graphs in M×M\times\mathbb{R} for those Riemannian manifolds MM which admit a local family of parallel hypersurfaces, each of them having constant principal curvatures. Following [5], such hypersurfaces are called isoparametric. We point out that many Riemannian manifolds MM admit isoparametric hypersurfaces, such as space forms, hyperbolic spaces, warped products, and 𝔼(κ,τ)\mathbb{E}(\kappa,\uptau) homogeneous spaces.

By suitably “gluing” pieces of HrH_{r}-graphs, we construct properly embedded HrH_{r}-hypersurfaces in M×M\times\mathbb{R} when MM is either the standard nn-sphere 𝕊n\mathbb{S}^{n} or one of the hyperbolic spaces 𝔽m\mathbb{H}_{\mathbb{F}}^{m} (rank 11 symmetric spaces of non compact type). In this setting, we show that there exists a rotational HrH_{r}-sphere in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} if and only if Hr>C𝔽(r),H_{r}>C_{\mathbb{F}}(r), where the constant C𝔽(r)C_{\mathbb{F}}(r) is defined as the limit of the rr-th mean curvature of a geodesic sphere of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} as its radius goes to infinity. (In particular, C𝔽(r)C_{\mathbb{F}}(r) is positive for 1r<n=dim𝔽m1\leq r<n=\dim\mathbb{H}_{\mathbb{F}}^{m} and vanishes for r=nr=n.) On the other hand, as we also show, for any r{1,,n}r\in\{1,\dots,n\} and any constant Hr>0,H_{r}>0, there exists a rotational HrH_{r}-sphere in 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}.

We remark that rotational hypersurfaces of a general product M×M\times\mathbb{R} are defined here as those which are foliated by horizontal geodesic spheres centered at an “axis” {o}×,\{o\}\times\mathbb{R}, oM.o\in M. (See Section 4 for more details.)

We provide other examples of properly embedded rotational Hr(>0)H_{r}(>0)-hypersurfaces in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} and in 𝕊n×\mathbb{S}^{n}\times\mathbb{R} as well, including Delaunay-type annuli and, in the case of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, entire graphs over 𝔽m.\mathbb{H}_{\mathbb{F}}^{m}. Then, we classify those complete connected rotational Hr(>0)H_{r}(>0)-hypersurfaces of these product spaces whose height functions are Morse-type (i.e., have isolated critical points), which include all the properly embedded rotational Hr(>0)H_{r}(>0)-hypersurfaces we obtain here.

We also construct and classify complete connected Hr(>0)H_{r}(>0)-hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} with no horizontal points (critical points of the height function) which are invariant by parabolic isometries (i.e, foliated by horospheres) or by hyperbolic translations (i.e., foliated by equidistant hypersurfaces). In the latter case, of course, only the real hyperbolic space m:=n\mathbb{H}_{\mathbb{R}}^{m}:=\mathbb{H}^{n} is considered.

Our methods work equally well for HrH_{r}-hypersurfaces with Hr=0,H_{r}=0, the so called rr-minimal hypersurfaces. By applying them, we obtain a one-parameter family of rotational, properly embedded catenoid-type rr-minimal nn-annuli in 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}. Similarly, we obtain a one-parameter family of rotational, properly embedded Delaunay-type rr-minimal nn-annuli in 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}. Then, we show that these annuli are the only complete connected rr-minimal rotational hypersurfaces of these product spaces (besides horizontal hyperplanes and, in the case r=n,r=n, cylinders over geodesic spheres).

Analogously to the case of Hr(>0)H_{r}(>0)-hypersurfaces, we construct and classify the complete connected rr-minimal hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which are invariant by either parabolic isometries or hyperbolic translations.

The study of HrH_{r}-hypersurfaces of a Riemannian manifold leads naturally to considerations on their uniqueness properties. On this matter, Montiel and Ros [28] (see also [26]) showed the following Alexandrov-type theorem:

The only compact, connected, and embedded HrH_{r}-hypersurfaces in n,\mathbb{R}^{n}, n,\mathbb{H}^{n}, or an open hemisphere of  𝕊n\mathbb{S}^{n} are geodesic spheres.

In [18], this result was extended to the context of HrH_{r}-hypersurfaces of n×,\mathbb{H}^{n}\times\mathbb{R}, where the geodesic spheres in the statement are replaced by rotational spheres.

Here, we establish uniqueness results for rotational HrH_{r}-spheres of n×\mathbb{H}^{n}\times\mathbb{R} and 𝕊n×\mathbb{S}^{n}\times\mathbb{R}, n3.n\geq 3. The case n=2n=2 was settled in [1] (for r=1r=1) and in [20] (for r=2r=2). More precisely, we show that any compact connected strictly convex HrH_{r}-hypersurface Σ\varSigma of n×\mathbb{H}^{n}\times\mathbb{R} or 𝕊n×\mathbb{S}^{n}\times\mathbb{R} (n3)(n\geq 3) is necessarily an embedded rotational sphere. Assuming Σ\varSigma complete, instead of compact, the same conclusion holds if, in addition, the height function of Σ\varSigma has a critical point and, in the case Σn×\varSigma\subset\mathbb{H}^{n}\times\mathbb{R}, the least principal curvature of Σ\varSigma is bounded away from zero. Finally, we show that, for n3,n\geq 3, any connected, properly immersed, and strictly convex Hr(>0)H_{r}(>0)-hypersurface of  𝕊n×\mathbb{S}^{n}\times\mathbb{R} is necessarily an embedded rotational HrH_{r}-sphere.

It is worth mentioning that these uniqueness results constitute applications of the main theorems in [11], which concern convexity properties of hypersurfaces in M×,M\times\mathbb{R}, MM being either a Hadamard manifold or the sphere 𝕊n.\mathbb{S}^{n}. Besides, the noncompact cases are based on height estimates we establish here for strictly convex vertical graphs in arbitrary products M×.M\times\mathbb{R}.

The paper is organized as follows. In Section 2, in addition to the usual setting of notation and basic concepts, we include a brief presentation of the hyperbolic spaces 𝔽m\mathbb{H}_{\mathbb{F}}^{m} and of the Maximum-Continuation Principle for HrH_{r}-hypersurfaces. In Section 3, we introduce graphs on parallel hypersurfaces and establish two key lemmas. In Section 4 (resp. Section 5), we construct and classify complete rotational Hr(>0)H_{r}(>0)-hypersurfaces (resp. rr-minimal hypersurfaces) in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} and 𝕊n×,\mathbb{S}^{n}\times\mathbb{R}, whereas in Section 6 (resp. Section 7) we do the same for complete translational ones (i.e, invariant by either parabolic or hyperbolic isometries). In the concluding Section 8, we prove the aforementioned uniqueness results.

2. Preliminaries

Let Σn\varSigma^{n}, n2,n\geq 2, be an oriented hypersurface of a Riemannian manifold M¯n+1.\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu^{n+1}. Set ¯\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu for the Levi-Civita connection of M¯,\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu, NN for the unit normal field of Σ,\varSigma, and AA for its shape operator with respect to N,N, so that

AX=¯XN,XTΣ,AX=-\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu_{X}N,\,\,X\in T\varSigma,

where TΣT\varSigma stand for the tangent bundle of Σ\varSigma. The principal curvatures of Σ,\varSigma, that is, the eigenvalues of the shape operator A,A, will be denoted by k1,,knk_{1}\,,\dots,k_{n}.

Given an integer r0,r\geq 0, we define the (non normalized) rr-th mean curvature HrH_{r} of the hypersurface ΣM¯\varSigma\subset\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu as:

(1) Hr:={1ifr=0.i1<<irki1kirif  1rn.0ifr>n.H_{r}:=\left\{\begin{array}[]{cl}1&{\rm if}\,\,r=0.\\[4.30554pt] \displaystyle\sum_{i_{1}<\cdots<i_{r}}k_{i_{1}}\dots k_{i_{r}}&{\rm if}\,\,1\leq r\leq n.\\[12.91663pt] 0&{\rm if}\,\,r>n.\end{array}\right.

Notice that H1H_{1} and HnH_{n} are the non normalized mean curvature and Gauss-Kronecker curvature functions of Σ,\varSigma, respectively, i.e.,

H1=traceAandHn=detA.H_{1}={\rm trace}\,A\quad\text{and}\quad H_{n}=\det A.
Definition 1.

With the above notation, given a constant Hr,H_{r}\in\mathbb{R}, we say that ΣM¯\varSigma\subset\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu is an HrH_{r}-hypersurface of M¯\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu if its rr-th mean curvature is constant and equal to Hr.H_{r}. In the case Hr=0H_{r}=0, we say that Σ\varSigma is an rr-minimal hypersurface of M¯.\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu.

Definition 2.

A hypersurface ΣM¯\varSigma\subset\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu is said to be convex at xΣx\in\varSigma if, at this point, all the nonzero principal curvatures have the same sign. If, in addition, none of these principal curvatures is zero, then Σ\varSigma is said to be strictly convex at x.x. We call Σ\varSigma convex (resp. strictly convex) if it is convex (resp. strictly convex) at all of its points.

The ambient spaces we shall consider are the products M¯n+1=Mn×\mkern 1.5mu\overline{\mkern-1.5muM\mkern-1.5mu}\mkern 1.5mu^{n+1}=M^{n}\times\mathbb{R} — where MnM^{n} is some Riemannian manifold — endowed with the standard product metric:

,=,M+dt2.\langle\,,\,\rangle=\langle\,,\,\rangle_{M}+dt^{2}.

In this setting, we denote the gradient of the projection π\pi_{\mathbb{R}} of M×M\times\mathbb{R} by t\partial_{t} , which is easily seen to be a parallel field on M×.M\times\mathbb{R}.

Let Σ\varSigma be a hypersurface of M×M\times\mathbb{R}. Its height function ξ\xi and its angle function Θ\varTheta are defined by the following identities:

ξ(x):=π|ΣandΘ(x):=N(x),t,xΣ.\xi(x):=\pi_{\scriptscriptstyle\mathbb{R}}|_{\varSigma}\quad\text{and}\quad\varTheta(x):=\langle N(x),\partial_{t}\rangle,\,\,x\in\varSigma.

A critical point of ξ\xi is called horizontal, whereas a point on which Θ\varTheta vanishes is called vertical. Notice that xΣx\in\varSigma is horizontal if and only if Θ(x)=±1.\varTheta(x)=\pm 1.

We shall denote the gradient field and the Hessian of a function ζ\zeta on Σ\varSigma by ζ\nabla\zeta and Hessζ,{\rm Hess}\,\zeta, respectively. It is easily checked that

(2) ξ=tΘNandΘ=Aξ.\nabla\xi=\partial_{t}-\varTheta N\quad\text{and}\quad\nabla\varTheta=-A\nabla\xi.

From (2), for all X,YTΣ,X,Y\in T\varSigma, one has ¯Xξ=(Θ¯XN+X(Θ)N).\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu_{X}\,\nabla\xi=-(\varTheta\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu_{X}N+X(\varTheta)N). Hence,

(3) Hessξ(X,Y)=ΘAX,YX,YTΣ.{\rm Hess}\,\xi(X,Y)=\varTheta\langle AX,Y\rangle\,\,\,\forall X,Y\in T\varSigma.

Given t,t\in\mathbb{R}, the set Pt:=M×{t}P_{t}:=M\times\{t\} is called a horizontal hyperplane of M×.M\times\mathbb{R}. Horizontal hyperplanes are all isometric to MM and totally geodesic in M×.M\times\mathbb{R}. In this context, we call a transversal intersection Σt:=ΣPt\varSigma_{t}:=\varSigma\mathrel{\text{\vbox{ \halign{#\cr\smash{$-$}\crcr$\pitchfork$\crcr} }}}P_{t} a horizontal section of Σ.\varSigma. Any horizontal section Σt\varSigma_{t} is a hypersurface of PtP_{t} . So, at any point xΣtΣ,x\in\varSigma_{t}\subset\varSigma, the tangent space TxΣT_{x}\varSigma of Σ\varSigma at xx splits as the orthogonal sum

(4) TxΣ=TxΣtSpan{ξ}.T_{x}\varSigma=T_{x}\varSigma_{t}\oplus{\rm Span}\{\nabla\xi\}.

We will adopt the notation ϵn\mathbb{Q}_{\epsilon}^{n} for the simply connected space form of constant sectional curvature ϵ{0,1,1};\epsilon\in\{0,1,-1\}; the Euclidean space n\mathbb{R}^{n} (ϵ=0\epsilon=0), the unit sphere 𝕊n\mathbb{S}^{n} (ϵ=1\epsilon=1), and the hyperbolic space n\mathbb{H}^{n} (ϵ=1\epsilon=-1).

2.1. The hyperbolic spaces 𝔽m\mathbb{H}_{\mathbb{F}}^{m}

Many of our results in this paper involve the Riemannian manifolds known as hyperbolic spaces, which include the canonical (real) nn-dimensional hyperbolic space n.\mathbb{H}^{n}. These manifolds are precisely the rank 11 symmetric spaces of non-compact type, which can be described through the four normed division algebras: \mathbb{R} (real numbers), \mathbb{C} (complex numbers), 𝕂\mathbb{K} (quaternions) and 𝕆\mathbb{O} (octonions). They are denoted by

m,m,𝕂mand𝕆2,m1,\mathbb{H}_{\mathbb{R}}^{m},\,\,\mathbb{H}_{\mathbb{C}}^{m},\,\,\mathbb{H}_{\mathbb{K}}^{m}\,\,\,\text{and}\,\,\,\mathbb{H}_{\mathbb{O}}^{2},\,\,\,\,m\geq 1,

and called real hyperbolic space, complex hyperbolic space, quaternionic hyperbolic space and Cayley hyperbolic plane, respectively.

We will adopt the unified notation 𝔽m\mathbb{H}_{\mathbb{F}}^{m} for the hyperbolic spaces, where m=2m=2 for 𝔽=𝕆.\mathbb{F}=\mathbb{O}. The real dimension of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} is n=mdim𝔽.n=m\dim\mathbb{F}. In particular, 𝕆2\mathbb{H}_{\mathbb{O}}^{2} has dimension n=16.n=16. We will keep the standard notation n\mathbb{H}^{n} for the real hyperbolic space n\mathbb{H}_{\mathbb{R}}^{n} and assume n2.n\geq 2.

Denoting by |||\,\,| the norm in 𝔽𝕆\mathbb{F}\neq\mathbb{O}, and taking in 𝔽m\mathbb{F}^{m} standard coordinates (z1,,zm),(z_{1},\dots,z_{m}), we have that 𝔽m\mathbb{H}_{\mathbb{F}}^{m} is modeled by the unit ball

B:={(z1,,zm)𝔽m;i=1m|zi|2<1}B:=\left\{(z_{1}\,,\dots,z_{m})\in\mathbb{F}^{m}\,;\,\sum_{i=1}^{m}|z_{i}|^{2}<1\right\}

equipped with the Hermitian form whose coefficients in these coordinates are

gij:=δij1i=1m|zi|2+ziz¯j(1i=1m|zi|2)2,g_{ij}:=\frac{\delta_{ij}}{1-\sum_{i=1}^{m}|z_{i}|^{2}}+\frac{z_{i}\bar{z}_{j}}{(1-\sum_{i=1}^{m}|z_{i}|^{2})^{2}}\,,

where the bar denotes conjugation.

For 𝔽=,\mathbb{F}=\mathbb{R}, the metric gg defined by gijg_{ij} correspond to the Klein model for n.\mathbb{H}^{n}. Also, for m=1m=1, gg reduces to the canonical Poincaré metric of n\mathbb{H}^{n}. In particular, 1\mathbb{H}_{\mathbb{C}}^{1} is isometric to 2\mathbb{H}^{2}, and 𝕂1\mathbb{H}_{\mathbb{K}}^{1} is isometric to 4\mathbb{H}^{4}. (See [8] for a detailed discussion on the hyperbolic spaces 𝔽m\mathbb{H}_{\mathbb{F}}^{m}, 𝔽𝕆.\mathbb{F}\neq\mathbb{O}.)

The description of a model for the Cayley hyperbolic plane 𝕆2\mathbb{H}_{\mathbb{O}}^{2} is more involved. We refer to [2] and the references therein for an account on this space.

We remark that, being symmetric, the hyperbolic spaces 𝔽m\mathbb{H}_{\mathbb{F}}^{m} are homogeneous. In addition, they are included in a distinguished class of Lie groups known as Damek-Ricci spaces (see Example 3 in the next section). In this context, it can be shown that any hyperbolic space 𝔽m\mathbb{H}_{\mathbb{F}}^{m} is a Hadamard-Einstein manifold with nonconstant (except for n\mathbb{H}^{n}) sectional curvatures pinched between 1-1 and 1/4-1/4 (cf. [5, Sections 4.1.9 and 4.2]).

2.2. The Maximum-Continuation Principle

Two major tools employed in the study of hypersurfaces of constant curvature are the Maximum Principle and the Continuation Principle for solutions of elliptic PDE’s. In the case of the rr-th mean curvature HrH_{r}, it was shown in [9, Proposition 3.2] that a hypersurface Σ\varSigma of a Riemannian manifold MM with Hr>0H_{r}>0 which is strictly convex at a point is given locally by a graph of a solution of an elliptic PDE. From the Continuation Principle (see, e.g., [23, 32]), if two such solutions are defined in a domain Ω\Omega and coincide in a subdomain ΩΩ,\Omega^{\prime}\subset\Omega, then they coincide in Ω.\Omega. These facts, together with [21, Theorem 1.1], give the following result.

Maximum-Continuation Principle.

Let Σ1,Σ2M\varSigma_{1}\,,\varSigma_{2}\subset M be complete connected Hr(>0)H_{r}(>0)-hypersurfaces of a Riemannian manifold MM which are tangent at a point xΣ1Σ2x\in\varSigma_{1}\cap\varSigma_{2} . Let N(x)TxΣiN(x)\in T_{x}\varSigma_{i}^{\perp} (i=1,2)(i=1,2) be the common unit normal and assume that Σ2\varSigma_{2} is strictly convex at one of its points. Then, if (near xx) Σ1\varSigma_{1} remains above Σ2\varSigma_{2} with respect to N(x),N(x), one has Σ1=Σ2\varSigma_{1}=\varSigma_{2} .

3. HrH_{r}-Graphs on Parallel Hypersurfaces

In this section, we give a detailed description of graphs in M×M\times\mathbb{R} which are built on families of parallel hypersurfaces of M.M. As we mentioned before, they will constitute our main tool for constructing HrH_{r}-hypersurfaces in product spaces M×.M\times\mathbb{R}.

With this purpose, consider an isometric immersion

f:M0n1Mnf:M_{0}^{n-1}\rightarrow M^{n}

between two Riemannian manifolds M0n1M_{0}^{n-1} and Mn,M^{n}, and suppose that there is a neighborhood 𝒰\mathscr{U} of M0M_{0} in TM0TM_{0}^{\perp} without focal points of f,f, that is, the restriction of the normal exponential map expM0:TM0M\exp^{\perp}_{M_{0}}:TM_{0}^{\perp}\rightarrow M to 𝒰\mathscr{U} is a diffeomorphism onto its image. In this case, denoting by η\eta the unit normal field of f,f, there is an open interval I0,I\owns 0, such that, for all pM0,p\in M_{0}, the curve

(5) γp(s)=expM(f(p),sη(p)),sI,\gamma_{p}(s)=\exp_{\scriptscriptstyle M}(f(p),s\eta(p)),\,s\in I,

is a well defined geodesic of MM without conjugate points. Thus, for all sI,s\in I,

fs:M0Mpγp(s)\begin{array}[]{cccc}f_{s}:&M_{0}&\rightarrow&M\\ &p&\mapsto&\gamma_{p}(s)\end{array}

is an immersion of M0M_{0} into M,M, which is said to be parallel to f.f. Observe that, given pM0p\in M_{0}, the tangent space fs(TpM0)f_{s_{*}}(T_{p}M_{0}) of fsf_{s} at pp is the parallel transport of f(TpM0)f_{*}(T_{p}M_{0}) along γp\gamma_{p} from 0 to s.s. We also remark that, with the induced metric, the unit normal ηs\eta_{s} of fsf_{s} at pp is given by

ηs(p)=γp(s).\eta_{s}(p)=\gamma_{p}^{\prime}(s).
Definition 3.

Let ϕ:Iϕ(I)\phi:I\rightarrow\phi(I)\subset\mathbb{R} be an increasing diffeomorphism, i.e., ϕ>0.\phi^{\prime}>0. With the above notation, we call the set

(6) Σ:={(fs(p),ϕ(s))M×;pM0,sI},\varSigma:=\{(f_{s}(p),\phi(s))\in M\times\mathbb{R}\,;\,p\in M_{0},\,s\in I\},

the graph determined by {fs;sI}\{f_{s}\,;\,s\in I\} and ϕ,\phi, or (fs,ϕ)(f_{s},\phi)-graph, for short.

Notice that, for a given (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma, and for any sIs\in I , fs(M0)f_{s}(M_{0}) is the projection on MM of the horizontal section Σϕ(s)Σ,\varSigma_{\phi(s)}\subset\varSigma, that is, these sets are the level hypersurfaces of Σ.\varSigma.

For an arbitrary point x=(fs(p),ϕ(s))x=(f_{s}(p),\phi(s)) of such a graph Σ,\varSigma, one has

TxΣ=fs(TpM0)Span{s},s=ηs+ϕ(s)t.T_{x}\varSigma=f_{s_{*}}(T_{p}M_{0})\oplus{\rm Span}\,\{\partial_{s}\},\,\,\,\partial_{s}=\eta_{s}+\phi^{\prime}(s)\partial_{t}.

So, a unit normal to Σ\varSigma is

(7) N=ϕ1+(ϕ)2ηs+11+(ϕ)2t.N=\frac{-\phi^{\prime}}{\sqrt{1+(\phi^{\prime})^{2}}}\eta_{s}+\frac{1}{\sqrt{1+(\phi^{\prime})^{2}}}\partial_{t}\,.

In particular, its angle function is

(8) Θ=11+(ϕ)2\varTheta=\frac{1}{\sqrt{1+(\phi^{\prime})^{2}}}\,\cdot

A key property of (fs,ϕ)(f_{s},\phi)-graphs is that the trajectories of ξ\nabla\xi on them are lines of curvature, that is, ξ\nabla\xi is one of its principal directions. (Notice that, by (8), 0<Θ<1,0<\varTheta<1, so ξ\nabla\xi never vanishes on an (fs,ϕ)(f_{s},\phi)-graph.) More precisely (cf. [12, 36]),

(9) Aξ=ϕ′′(1+(ϕ)2)3ξ.A\nabla\xi=\frac{\phi^{\prime\prime}}{(\sqrt{1+(\phi^{\prime})^{2}})^{3}}\nabla\xi.

We point out that, besides being lines of curvature, the trajectories of ξ\nabla\xi on an (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma, when properly reparametrized, are also geodesics. This follows from the fact that Θ\varTheta, and consequently ξ,\|\nabla\xi\|, is constant along the horizontal sections of Σ\varSigma (see [36, Lemma 5]). It should also be noticed that these trajectories project on the geodesics γp=γp(s)\gamma_{p}=\gamma_{p}(s) given by (5) (Fig. 1).

Refer to caption
Figure 1. Trajectory of ξ\nabla\xi on an (fs,ϕ)(f_{s},\phi)-graph.

Let us compute now the principal curvatures of an (fs,ϕ)(f_{s},\phi)-graph Σ.\varSigma. For that, let {X1,,Xn}\{X_{1}\,,\dots,X_{n}\} be the orthonormal frame of principal directions of Σ\varSigma in which Xn=ξ/ξ.X_{n}=\nabla\xi/\|\nabla\xi\|. In this case, for 1in1,1\leq i\leq n-1, the fields XiX_{i} are all horizontal, that is, tangent to MM (cf. (4)). Therefore, setting

(10) ϱ:=ϕ1+(ϕ)2\varrho:=\frac{\phi^{\prime}}{\sqrt{1+(\phi^{\prime})^{2}}}

and considering (7), we have, for all i=1,,n1,i=1,\dots,n-1, that

ki=AXi,Xi=¯XiN,Xi=ϱ¯Xiηs,Xi=ϱkis,k_{i}=\langle AX_{i},X_{i}\rangle=-\langle\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu_{X_{i}}N,X_{i}\rangle=\varrho\langle\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu_{X_{i}}\eta_{s},X_{i}\rangle=-\varrho k_{i}^{s},

where kisk_{i}^{s} is the ii-th principal curvature of fs.f_{s}\,. Also, it follows from (9) that kn=ϱ.k_{n}=\varrho^{\prime}. Thus, the array of principal curvatures of the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma is

(11) ki=ϱkis(1in1)andkn=ϱ.k_{i}=-\varrho k_{i}^{s}\,\,(1\leq i\leq n-1)\quad\text{and}\quad k_{n}=\varrho^{\prime}.

Now, considering the above identities and writing, for 1rn,1\leq r\leq n,

Hr=i1<<irnki1kir+i1<<ir1ki1kir1kn,H_{r}=\sum_{i_{1}<\cdots<i_{r}\neq n}k_{i_{1}}\dots k_{i_{r}}+\sum_{i_{1}<\cdots<i_{r-1}}k_{i_{1}}\dots k_{i_{r-1}}k_{n}\,,

we have that the rr-th mean curvature of the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma is

(12) Hr=(1)rHrsϱr+(1)r1Hr1sϱr1ϱ,H_{r}=(-1)^{r}H_{r}^{s}\varrho^{r}+(-1)^{r-1}H_{r-1}^{s}\varrho^{r-1}\varrho^{\prime},

where HrsH_{r}^{s} denotes the rr-th mean curvature of fs.f_{s}.

Due to equality (12), the function defined in (10) — to be called the ϱ\varrho-function of the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma — will play a major role in the sequel. We remark that, up to a constant, the ϱ\varrho-function of Σ\varSigma determines its ϕ\phi-function. Indeed, it follows from equality (10) that

(13) ϕ(s)=s0sϱ(u)1ϱ2(u)𝑑u+ϕ(s0),s0I.\phi(s)=\int_{s_{0}}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du+\phi(s_{0}),\,\,\,s_{0}\in I.

We introduce now a special type of family of parallel hypersurfaces which will be used for constructing HrH_{r}-hypersurfaces in M×.M\times\mathbb{R}.

Definition 4.

Following [5], we call a parallel family {fs:M0M;sI}\{f_{s}:M_{0}\rightarrow M\,;\,s\in I\} isoparametric if, for each sI,s\in I, any principal curvature of fsf_{s} is constant (depending on ss). If so, each hypersurface fsf_{s} is also called isoparametric.

We should mention that there is some mismatch regarding the nomenclature for isoparametric hypersurfaces. In some contexts, isoparametric hypersurfaces are defined as those which, together with its parallel nearby hypersurfaces, have constant mean curvature. It is shown that some manifolds MM admit hypersurfaces which are isoparametric in this sense, and non isoparametric as we defined.

Let Σ\varSigma be an (fs,ϕ)(f_{s},\phi)-graph such that the parallel family {fs;sI}\{f_{s}\,;\,s\in I\} is isoparametric. Then, for any r=1,,n1,r=1,\dots,n-1, the rr-th mean curvature of fsf_{s} is a function of ss alone, which we assume to be no vanishing. In this setting, writing τ:=ϱr\uptau:=\varrho^{r}, and considering (12) with HrH_{r} constant, we obtain the following result, which turns out to be our main lemma.

Lemma 1.

Let {fs:M0M;sI}\{f_{s}:M_{0}\rightarrow M\,;\,s\in I\} be an isoparametric family of hypersurfaces whose r(<n)r(<n)-mean curvatures HrsH_{r}^{s} never vanish. Given r{1,,n}r\in\{1,\dots,n\} and Hr,H_{r}\in\mathbb{R}, let τ\uptau be a solution of the first order differential equation

(14) y=a(s)y+b(s),sI,y^{\prime}=a(s)y+b(s),\,\,\,s\in I,

where the coefficients aa and bb are the functions

(15) a(s):=rHrsHr1sandb(s):=(1)r1rHrHr1sa(s):=\frac{rH_{r}^{s}}{H_{r-1}^{s}}\quad{\rm and}\quad b(s):=\frac{(-1)^{r-1}rH_{r}}{H_{r-1}^{s}}\,\cdot

Then, if  0<τ<1,0<\uptau<1, the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma with ϱ\varrho-function τ1/r\uptau^{1/r} is an HrH_{r}-hypersurface of the product M×.M\times\mathbb{R}. Conversely, if an (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma has constant rr-th mean curvature HrH_{r} , then τ:=ϱr\uptau:=\varrho^{r} is a solution of (14).

Regarding equation (14), recall that its general solution is

(16) τ(s)=1μ(s)(τ0+s0sb(u)μ(u)𝑑u),s0,sI,τ0,\uptau(s)=\frac{1}{\mu(s)}\left(\uptau_{0}+\int_{s_{0}}^{s}{b(u)}{\mu(u)}du\right),\,\,s_{0},s\in I,\,\uptau_{0}\in\mathbb{R},

where μ\mu is the exponential function

μ(s)=exp(s0sa(u)𝑑u),sI.\mu(s)=\exp\left(-\int_{s_{0}}^{s}a(u)du\right),\,\,\,s\in I.

It follows from Lemma 1 that, as long as MM admits isoparametric hypersurfaces with non vanishing rr-th mean curvature, for any Hr,H_{r}\in\mathbb{R}, there exist HrH_{r}-graphs in M×M\times\mathbb{R}. (Notice that, the interval II and the constant τ0\uptau_{0} in (16) can be chosen in such a way that the corresponding solution τ\uptau of (14) satisfies 0<τ<10<\uptau<1.) This includes, as a trivial case, the Euclidean space n.\mathbb{R}^{n}. In the next examples, we shall consider other manifolds MM to which Lemma 1 applies.

Example 1 (sphere 𝕊n\mathbb{S}^{n}).

It is a well known fact that isoparametric hypersurfaces in 𝕊n\mathbb{S}^{n} are abundant and include all its geodesic spheres. In fact, the classification of the isoparametric hypersurfaces of 𝕊n\mathbb{S}^{n} is a long stand open problem (see, e.g., [15]).

Example 2 (warped products).

Let M=I×ωFn1M=I\times_{\omega}F^{n-1} be a warped product, where the basis II\subset\mathbb{R} is an open interval and the fiber FF is an arbitrary (n1)(n-1)-dimensional manifold. For each sI,s\in I, define fsf_{s} as the standard immersion F{s}×ωFM.F\hookrightarrow\{s\}\times_{\omega}F\subset M. It is well known that, with the induced metric, ={fs;sI}\mathscr{F}=\{f_{s}\,;\,s\in I\} is a parallel family of totally umbilical hypersurfaces of MM with constant principal curvatures ω/ω\omega^{\prime}/\omega (see, e.g., [6]). In particular, \mathscr{F} is isoparametric. Hence, if ω\omega^{\prime} never vanishes, Lemma 1 applies to M.M.

Example 3 (Damek-Ricci spaces).

Let us consider the Riemannian manifolds known as Damek-Ricci spaces. These are Lie groups endowed with a left invariant metric with especial properties (see [5, 15]). For instance, all Damek-Ricci spaces are both Hadamard and Einstein manifolds. As we have mentioned, the hyperbolic spaces 𝔽m\mathbb{H}_{\mathbb{F}}^{m} are Damek-Ricci spaces. In fact, they are the only ones which are symmetric. Their isoparametric hypersurfaces include their geodesic spheres, as well as their horospheres. Finally, we point out that, in [13], the authors obtained families of isoparametric hypersurfaces with non vanishing rr-th curvatures in Damek-Ricci harmonic spaces.

Example 4 (E(κ,τ)E(\kappa,\uptau)-spaces).

In [16], it was proved that there exist isoparametric families of parallel surfaces with nonzero constant principal curvatures in 𝔼(k,τ)\mathbb{E}(k,\uptau) spaces satisfying k4τ20.k-4\uptau^{2}\neq 0. (Those include the products 2×\mathbb{H}^{2}\times\mathbb{R} and 𝕊2×\mathbb{S}^{2}\times\mathbb{R}, the Heisenberg space Nil3{\rm Nil}_{3}, the Berger spheres, and the universal cover of the special linear group SL2(){\rm SL}_{2}(\mathbb{R})).

In the next two sections we construct properly embedded HrH_{r}-hypersurfaces in products M×M\times\mathbb{R} by suitably “gluing” HrH_{r}-graphs. To this task, the following elementary fact will be considerably helpful.

Lemma 2.

Let ϱ:[a1,a2]\varrho:[a_{1},a_{2}]\rightarrow\mathbb{R} be a differentiable function which satisfies:

0<ϱ|(a1,a2)<1.0<\varrho{|_{(a_{1},a_{2})}}<1.

Assume that one (or both) of the following hold:

  • i)

    ϱ(a2)=1\varrho(a_{2})=1 and ϱ(a2)>0.\varrho^{\prime}(a_{2})>0.

  • ii)

    ϱ(a1)=1\varrho(a_{1})=1 and ϱ(a1)<0.\varrho^{\prime}(a_{1})<0.

Under these conditions, there exists δ>0\delta>0 such that the improper integral

(17) a2δa2ϱ(s)ds1ϱ2(s)\int_{a_{2}-\delta}^{a_{2}}\frac{\varrho(s)ds}{\sqrt{1-\varrho^{2}(s)}}

is convergent if  (i) occurs. Analogously, the improper integral

a1a1+δϱ(s)ds1ϱ2(s)\int_{a_{1}}^{a_{1}+\delta}\frac{\varrho(s)ds}{\sqrt{1-\varrho^{2}(s)}}

is convergent if (ii) occurs.

Proof.

Assume that (i) occurs. In this case, there exist positive constants, δ\delta and C,C, such that ϱ(s)C>0s(a2δ,a2).\varrho^{\prime}(s)\geq C>0\,\forall s\in(a_{2}-\delta,a_{2}). Therefore, since 0<ϱ|(a1,a2)<1,0<\varrho{|_{(a_{1},a_{2})}}<1,

a2δa2ϱ(s)ds1ϱ2(s)\displaystyle\int_{a_{2}-\delta}^{a_{2}}\frac{\varrho(s)ds}{\sqrt{1-\varrho^{2}(s)}} \displaystyle\leq a2δa2ϱ(s)dsϱ(s)1ϱ2(s)1Cϱ(a2δ)1dϱ1ϱ2\displaystyle\int_{a_{2}-\delta}^{a_{2}}\frac{\varrho^{\prime}(s)ds}{\varrho^{\prime}(s)\sqrt{1-\varrho^{2}(s)}}\leq\frac{1}{C}\int_{\varrho(a_{2}-\delta)}^{1}\frac{d\varrho}{\sqrt{1-\varrho^{2}}}
=\displaystyle= 1C(π2arcsin(ϱ(a2δ)))π2C,\displaystyle\frac{1}{C}\left(\frac{\pi}{2}-\arcsin(\varrho(a_{2}-\delta))\right)\leq\frac{\pi}{2C}\,,

which proves the convergence of the integral (17). The case (ii) is analogous. ∎

4. Rotational Hr(>0)H_{r}(>0)-hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} and 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}.

Rotational hypersurfaces in simply connected space forms ϵn\mathbb{Q}^{n}_{\epsilon} or products ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R} are among the most classical hypersurfaces of these spaces. In the case of ϵn×,\mathbb{Q}^{n}_{\epsilon}\times\mathbb{R}, they are obtained by rotating (with the aid of the group of isometries of ϵn\mathbb{Q}_{\epsilon}^{n}) a plane curve about an axis {o}×,\{o\}\times\mathbb{R}, oϵno\in\mathbb{Q}_{\epsilon}^{n}. Consequently, any connected component of any horizontal section Σt\varSigma_{t} of a rotational hypersurface Σ\varSigma in ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R} lies in a geodesic sphere of ϵn×{t}\mathbb{Q}_{\epsilon}^{n}\times\{t\} with center at the axis. This fact suggests the following definition.

Definition 5.

A hypersurface ΣM×\varSigma\subset M\times\mathbb{R} is called rotational, if there exists a fixed point oMo\in M such that any connected component of any horizontal section Σt\varSigma_{t} is contained in a geodesic sphere of M×{t}M\times\{t\} with center at o×{t}.o\times\{t\}. If so, the set {o}×\{o\}\times\mathbb{R} is called the axis of Σ.\varSigma. In particular, any horizontal hyperplane Pt:=M×{t}P_{t}:=M\times\{t\} is a rotational hypersurface of M×M\times\mathbb{R} with axis at any point oM.o\in M.

In what follows, we construct and classify complete rotational Hr(>0)H_{r}(>0)-hypersurfaces in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} and 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}.

4.1. Rotational Hr(>0)H_{r}(>0)-hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}

Let us consider a family

(18) :={fs:𝕊n1𝔽m;s(0,+)}\mathscr{F}:=\{f_{s}:\mathbb{S}^{n-1}\rightarrow\mathbb{H}_{\mathbb{F}}^{m}\,;\,\,\,s\in(0,+\infty)\}

of isoparametric concentric geodesic spheres of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} indexed by their radiuses, that is, for a fixed o𝔽mo\in\mathbb{H}_{\mathbb{F}}^{m}, and each s(0,+),s\in(0,+\infty), fs(𝕊n1)f_{s}(\mathbb{S}^{n-1}) is the geodesic sphere Ss(o)S_{s}(o) of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} with center at oo and radius s.s.

We remark that any sphere fsf_{s}\in\mathscr{F} is strictly convex. Also, in accordance to the notation of Section 3, for each s(0,+),s\in(0,+\infty), we choose the outward orientation of fsf_{s} , so that any principal curvature kisk_{i}^{s} of  fsf_{s} is negative. In this setting, the function s(0,+)Hrss\in(0,+\infty)\mapsto H_{r}^{s} is positive for rr even and negative for rr odd. Hence, for any constant Hr>0,H_{r}>0, the coefficients aa and bb in (15) are given by

(19) a(s)=r|Hrs||Hr1s|andb(s)=rHr|Hr1s|a(s)=-\frac{r|H_{r}^{s}|}{|H_{r-1}^{s}|}\quad\text{and}\quad b(s)=\frac{rH_{r}}{|H_{r-1}^{s}|}\,\cdot

The principal curvatures kisk_{i}^{s} of the geodesic spheres fsf_{s}\in\mathscr{F} are (n=dim𝔽mn=\dim\mathbb{H}_{\mathbb{F}}^{m}):

(20) k1s=12coth(s/2)with multiplicitynp1k2s=coth(s)with multiplicityp,\begin{aligned} k_{1}^{s}&=-\frac{1}{2}\coth(s/2)\,\,\,\text{with multiplicity}\,\,\,n-p-1\\[6.45831pt] k_{2}^{s}&=-\coth(s)\,\,\,\text{with multiplicity}\,\,\,p\end{aligned}\,\,,

where p=n1p=n-1 for n\mathbb{H}^{n}, p=1p=1 for m\mathbb{H}_{\mathbb{C}}^{m}, p=3p=3 for 𝕂m\mathbb{H}_{\mathbb{K}}^{m}, and p=7p=7 for 𝕆2\mathbb{H}_{\mathbb{O}}^{2} (see, e.g., [7, pgs. 353, 543] and [25]).

From equalities (20), we obtain the rr-mean curvatures HrsH_{r}^{s} of the geodesic spheres fsf_{s} of 𝔽m.\mathbb{H}_{\mathbb{F}}^{m}. For instance, in n,\mathbb{H}^{n}, n2,n\geq 2, we have

(21) Hrs=(1)r(n1r)cothr(s)(1rn1),H_{r}^{s}=(-1)^{r}{{n-1}\choose{r}}\coth^{r}(s)\quad(1\leq r\leq n-1),

whereas for m,m>1,\mathbb{H}_{\mathbb{C}}^{m},\,m>1, one has

Hrs\displaystyle H_{r}^{s} =\displaystyle= (12)r(n2r)cothr(s/2)\displaystyle\left(-\frac{1}{2}\right)^{r}{{n-2}\choose{r}}\coth^{r}(s/2)
+\displaystyle+ (1)r(12)r1(n2r1)cothr1(s/2)coth(s)\displaystyle(-1)^{r}\left(\frac{1}{2}\right)^{r-1}{{n-2}\choose{r-1}}\coth^{r-1}(s/2)\coth(s)

if 1rn21\leq r\leq n-2, and

(23) Hn1s=(1)n1coth(s)cothn2(s/2)/2n2.H_{n-1}^{s}=(-1)^{n-1}\coth(s)\coth^{n-2}(s/2)/2^{n-2}.

Analogously, one obtains the rr-th mean curvature functions HrsH_{r}^{s} for the other hyperbolic spaces. A direct computation from this data yields the following

Lemma 3.

The functions aa e bb defined in (19) have the following properties:

  • i)

    aa is negative and increasing for 1rn11\leq r\leq n-1, and vanishes for r=n.r=n.

  • ii)

    bb is positive and increasing for 1<rn1<r\leq n, and b=H1b=H_{1} for r=1.r=1.

In particular, we have the inequalities

(24) a(s)0,b(s)0,anda(s)+b(s)>0s(0,+).a^{\prime}(s)\geq 0,\,\,\,\,b^{\prime}(s)\geq 0,\,\,\,\,\,\text{and}\ \,\,\,\,a^{\prime}(s)+b^{\prime}(s)>0\,\,\,\forall s\in(0,+\infty).

We point out that, in the above setting, one has (cf. [27])

(25) |Hrs|=(n1r)sr+𝒪(s2r)|H_{r}^{s}|={{{n-1}\choose{r}}}s^{-r}+\mathcal{O}(s^{2-r})

in a neighborhood of s=0.s=0. In particular,

(26) lims0|Hrs|=+.\lim_{s\rightarrow 0}|H_{r}^{s}|=+\infty.

In what follows, by means of the family ,\mathscr{F}, we will construct complete rotational HrH_{r}-hypersurfaces in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which are made of pieces of (fs,ϕ)(f_{s},\phi)-graphs. With this purpose, we will look for solutions τ(s)\uptau(s) of the equation y=ay+by^{\prime}=ay+b (with aa and bb as in (19)) satisfying suitable initial conditions. Let us recall that, in this context, the general solution of y=ay+by^{\prime}=ay+b is

(27) τ(s)=1μ(s)(τ0+s0sb(u)μ(u)𝑑u)μ(s)=exp(s0sa(u)𝑑u),s0,s(0,+),τ0.\begin{aligned} \uptau(s)&=\frac{1}{\mu(s)}\left(\uptau_{0}+\int_{s_{0}}^{s}{b(u)}{\mu(u)}du\right)\\[6.45831pt] \mu(s)&=\exp\left(-\int_{s_{0}}^{s}a(u)du\right)\end{aligned}\,\,,\quad s_{0}\,,s\in(0,+\infty),\,\,\uptau_{0}\in\mathbb{R}.

Concerning the solutions τ(s)\uptau(s), we will be also interested in those which can be extended to s=0.s=0. Notice that, in principle, neither aa nor bb are defined at s=0s=0, which makes this point a singularity. However, the function bb is easily extendable to s=0.s=0. Indeed, we can set b(0)=Hrb(0)=H_{r} if r=1r=1, and b(0)=0b(0)=0 if r>1r>1 (from (26)). As for a,a, it follows from (25) that, for 1r<n,1\leq r<n,

(28) lims0|a(s)|=+andlims0s|a(s)|<+,\lim_{s\rightarrow 0}|a(s)|=+\infty\quad\text{and}\quad\lim_{s\rightarrow 0}s|a(s)|<+\infty,

which characterizes s=0s=0 as a regular singular point of y=ay+b.y^{\prime}=ay+b. This means that, despite the fact that aa is not defined at s=0,s=0, this equation has a nonnegative solution τ\uptau defined at s=0s=0 that satisfies τ(0)=0\uptau(0)=0 (cf. [33, Theorem 3.1], [35, Lemma 4.4]). More precisely, this solution is

(29) τ(s):={1μ(s)0sb(u)μ(u)𝑑uifs(0,+)0ifs=0,\uptau(s):=\left\{\begin{array}[]{lcl}\frac{1}{\mu(s)}\int_{0}^{s}{b(u)}{\mu(u)}du&\text{if}&s\in(0,+\infty)\\[6.45831pt] 0&\text{if}&s=0\end{array}\right.\,,

where μ\mu is a solution of y+ay=0.y^{\prime}+ay=0. Notice that the function τ\uptau defined in (29) is also the solution of y=ay+by^{\prime}=ay+b in the case r=n,r=n, i.e., for a=0.a=0. (Just set μ(s)=1.\mu(s)=1.)

As we shall see, the geometry of the HrH_{r}-hypersurfaces we construct from (fs,ϕ)(f_{s},\phi)-graphs is closely related to the growth of τ\uptau as s+.s\rightarrow+\infty. Taking that into account, for a given family of parallel geodesic spheres \mathscr{F} in 𝔽m,\mathbb{H}_{\mathbb{F}}^{m}, we define

(30) C𝔽(r):=lims+|Hrs|,r=1,,n.C_{\mathbb{F}}(r):=\lim_{s\rightarrow+\infty}|H_{r}^{s}|,\,\,\,r=1,\dots,n.

In particular, C𝔽(n)=0.C_{\mathbb{F}}(n)=0. Notice that, since 𝔽m\mathbb{H}_{\mathbb{F}}^{m} is homogeneous, the constant C𝔽(r)C_{\mathbb{F}}(r) is well defined, that is, it does not depend on the family \mathscr{F} of geodesic spheres.

It follows from equalities (21)–(23) that

  • i)

    C(r)=(n1r)(1rn1).C_{\mathbb{R}}(r)={{n-1}\choose{r}}\quad(1\leq r\leq n-1).

  • ii)

    C(r)=(12)r(n2r)+(12)r1(n2r1)(1rn2).C_{\mathbb{C}}(r)=\left(\frac{1}{2}\right)^{r}{{n-2}\choose{r}}+\left(\frac{1}{2}\right)^{r-1}{{n-2}\choose{r-1}}\quad(1\leq r\leq n-2).

  • iii)

    C(n1)=12n2C_{\mathbb{C}}(n-1)=\frac{1}{2^{n-2}}\,\cdot

Similarly, one can compute the other constants C𝔽(r)C_{\mathbb{F}}(r) and easily conclude that

C𝔽(r)>0r{1,,n1}.C_{\mathbb{F}}(r)>0\,\,\,\forall r\in\{1,\dots,n-1\}.

The next proposition shows the relation between the solution τ\uptau of y=ay+by^{\prime}=ay+b and the constant C𝔽(r).C_{\mathbb{F}}(r). Notice that, for 1rn1,1\leq r\leq n-1, the identities (19) yield:

lims+b(s)a(s)=HrC𝔽(r)\lim_{s\rightarrow+\infty}\frac{-b(s)}{\phantom{-}a(s)}=\frac{H_{r}}{C_{\mathbb{F}}(r)}\,\cdot
Proposition 1.

The following assertions hold:

  • i)

    The solution τ\uptau defined in (29) is increasing, i.e., τ>0\uptau^{\prime}>0 in (0,+).(0,+\infty).

  • ii)

    Both the solutions τ\uptau in (27) and (29) satisfy the following equality:

    (31) lims+τ(s)={+ifr=n.Hr/C𝔽(r)if1rn1.\lim_{s\rightarrow+\infty}\uptau(s)=\left\{\begin{array}[]{lcl}+\infty&\text{if}&r=n.\\[6.45831pt] H_{r}/C_{\mathbb{F}}(r)&\text{if}&1\leq r\leq n-1.\end{array}\right.
Proof.

To prove (i), let us observe first that, since the solution τ\uptau in (29) is positive in (0,+)(0,+\infty) and τ(0)=0,\uptau(0)=0, we have that τ\uptau is increasing near 0.0. Assume that τ\uptau is not increasing in (0,+).(0,+\infty). In this case, τ\uptau has a first critical point s0s_{0} in (0,+)(0,+\infty) which is necessarily a local maximum. However, considering (24) and the equality τ=aτ+b,\uptau^{\prime}=a\uptau+b, we have

τ′′(s0)=a(s0)τ(s0)+a(s0)τ(s0)+b(s0)=a(s0)τ(s0)+b(s0)>0,\uptau^{\prime\prime}(s_{0})=a^{\prime}(s_{0})\uptau(s_{0})+a(s_{0})\uptau^{\prime}(s_{0})+b^{\prime}(s_{0})=a^{\prime}(s_{0})\uptau(s_{0})+b^{\prime}(s_{0})>0,

which implies that s0s_{0} is a local minimum — a contradiction. Therefore, τ\uptau is increasing in (0,+),(0,+\infty), which proves (i).

Suppose that τ\uptau is as in (27). If r=n,r=n, since bb is increasing, one has

τ(s)=τ0+s0sb(u)𝑑uτ0+b(s0)(ss0),\uptau(s)=\uptau_{0}+\int_{s_{0}}^{s}b(u)du\geq\uptau_{0}+b(s_{0})(s-s_{0}),

which implies that τ(s)+\uptau(s)\rightarrow+\infty as s+.s\rightarrow+\infty.

Now, assume 1rn1.1\leq r\leq n-1. We claim that, in this case, μ(s)+\mu(s)\rightarrow+\infty as s+s\rightarrow+\infty. Indeed, for any fixed s0>0,s_{0}>0, and s>s0s>s_{0} ,

|a(s)|=r|Hrs||Hr1s|>rC𝔽(r)|Hr1s0|>0.|a(s)|=\frac{r|H_{r}^{s}|}{|H_{r-1}^{s}|}>\frac{rC_{\mathbb{F}}(r)}{|H_{r-1}^{s_{0}}|}>0.

Since |a||a| is decreasing, this inequality gives that inf|a|>0\inf|a|>0 in (0,+).(0,+\infty). Hence,

μ(s)=es0sa(u)𝑑u=es0s|a(u)|𝑑u>e(inf|a|)(ss0),\mu(s)=e^{-\int_{s_{0}}^{s}a(u)du}=e^{\int_{s_{0}}^{s}|a(u)|du}>e^{(\inf|a|)(s-s_{0})},

which clearly implies the claim on μ.\mu.

From the expression of τ,\uptau, we have (apply l’Hôpital to the second summand):

lims+τ(s)=lims+τ0μ(s)+lims+b(s)μ(s)μ(s)=lims+b(s)μ(s)a(s)μ(s)=HrC𝔽(r),\lim_{s\rightarrow+\infty}\uptau(s)=\lim_{s\rightarrow+\infty}\frac{\uptau_{0}}{\mu(s)}+\lim_{s\rightarrow+\infty}\frac{b(s)\mu(s)}{\mu^{\prime}(s)}=\lim_{s\rightarrow+\infty}\frac{b(s)\mu(s)}{-a(s)\mu(s)}=\frac{H_{r}}{C_{\mathbb{F}}(r)}\,,

which finishes the proof of (31) when τ\uptau is defined as in (27). The proof for the solution τ\uptau in (29) is completely analogous. ∎

The above proposition immediately gives the following result.

Corollary 1.

Let τ\uptau be as in (29). Then, there exists s0(0,+)s_{0}\in(0,+\infty) satisfying

0<τ|(0,s0)<1andτ(s0)=10<\uptau|_{(0,s_{0})}<1\,\,\,\,\text{and}\,\,\,\,\,\uptau(s_{0})=1

if and only if Hr>C𝔽(r)H_{r}>C_{\mathbb{F}}(r) (Fig. 2a). Consequently, for 1rn11\leq r\leq n-1 and any constant Hr(0,C𝔽(r)),H_{r}\in(0,C_{\mathbb{F}}(r)), the following inequality holds (Fig. 2b):

0<τ(s)<1s(0,+).0<\uptau(s)<1\,\,\,\forall s\in(0,+\infty).
Refer to caption
a Hr>C𝔽(r)H_{r}>C_{\mathbb{F}}(r)
Refer to caption
b 0<HrC𝔽(r)0<H_{r}\leq C_{\mathbb{F}}(r)
Figure 2. Graphs of τ\uptau (as in (29)) according to the sign of HrC𝔽(r).H_{r}-C_{\mathbb{F}}(r).

Now, we are in position to establish our first existence result.

Theorem 1.

Given r{1,,n}r\in\{1,\dots,n\} and a constant Hr>0,H_{r}>0, the following hold:

  • i)

    If  Hr>C𝔽(r),H_{r}>C_{\mathbb{F}}(r), there exists an embedded strictly convex rotational HrH_{r}-sphere in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which is symmetric with respect to a horizontal hyperplane.

  • ii)

    If  0<HrC𝔽(r),0<H_{r}\leq C_{\mathbb{F}}(r), there exists an entire strictly convex rotational HrH_{r}-graph in 𝔽m×[0,+)\mathbb{H}_{\mathbb{F}}^{m}\times[0,+\infty) which is tangent to 𝔽m×{0}\mathbb{H}_{\mathbb{F}}^{m}\times\{0\} at a single point, and whose height function is unbounded above. Consequently, there are no compact HrH_{r}-hypersurfaces of  𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} for such values of HrH_{r} .

Proof.

Let \mathscr{F} be an arbitrary family of parallel geodesic spheres of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} as in (18). Consider the functions aa and bb defined in (19) and let τ\uptau be the solution (29) of the ODE y=ay+b.y^{\prime}=ay+b.

If Hr>C𝔽(r),H_{r}>C_{\mathbb{F}}(r), we have from Corollary 1 that there exists s0(0,+)s_{0}\in(0,+\infty) satisfying

τ(0)=0<τ|(0,s0)<1=τ(s0).\uptau(0)=0<\uptau|_{(0,s_{0})}<1=\uptau(s_{0}).

Hence, by Lemma 1, the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma^{\prime} with ϱ\varrho-function ϱ:=τ|[0,s0)r\varrho:=\sqrt[r]{\uptau|_{[0,s_{0})}} is a rotational HrH_{r}-graph of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} over the open ball Bs0(o)𝔽mB_{s_{0}}(o)\subseteq\mathbb{H}_{\mathbb{F}}^{m} such that

(32) ϕ(s)=0sϱ(u)1ϱ2(u)𝑑u,s[0,s0).\phi(s)=\int_{0}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du,\,\,\,\,s\in[0,s_{0}).

From Proposition 1-(i), one has τ(s0)>0,\uptau^{\prime}(s_{0})>0, which implies that ϱ(s0)>0.\varrho^{\prime}(s_{0})>0. Hence, by Lemma 2, ϕ\phi extends to s0s_{0} , i.e.,

ϕ(s0):=limss0ϕ(s)\phi(s_{0}):=\lim_{s\rightarrow s_{0}}\phi(s)

is well defined. In particular, Σ=Ss0×{ϕ(s0)}.\partial\varSigma^{\prime}=S_{s_{0}}\times\{\phi(s_{0})\}.

Refer to caption
Figure 3. The trajectories of ξ\nabla\xi on Σ\varSigma^{\prime} emanate from oo and meet Σ\partial\varSigma^{\prime} orthogonally.

Notice that oΣo\in\varSigma^{\prime} is an isolated minimum of the height function ξ\xi of Σ.\varSigma^{\prime}. Thus, Σ\varSigma^{\prime} is strictly convex at o.o. In addition, by the identities (11), at any point of Σ{o},\varSigma^{\prime}-\{o\}, all the principal curvatures are positive. Therefore, Σ\varSigma^{\prime} is strictly convex.

As we know, the angle function Θ\varTheta of Σ\varSigma^{\prime} is given by

(33) Θ=11+(ϕ)2\varTheta=\frac{1}{\sqrt{1+(\phi^{\prime})^{2}}}\,\cdot

Since ϱ(s0)=1\varrho(s_{0})=1, we have from (32) that ϕ(s)+\phi^{\prime}(s)\rightarrow+\infty as ss0s\rightarrow s_{0} . This, together with (33), implies that the tangent spaces of Σ\varSigma^{\prime} along Σ\partial\varSigma^{\prime} are vertical. Hence, the trajectories of ξ\nabla\xi all emanate from oo and meet Σ\partial\varSigma^{\prime} orthogonally (Fig. 3). Recall that these trajectories are geodesics which foliate Σ.\varSigma^{\prime}.

Now, set Σ′′\varSigma^{\prime\prime} for the reflection of Σ\varSigma^{\prime} with respect to 𝔽m×{ϕ(s0)}\mathbb{H}_{\mathbb{F}}^{m}\times\{\phi(s_{0})\} and define

Σ:=closure(Σ)closure(Σ′′),\varSigma:={\rm closure}\,(\varSigma^{\prime})\cup{\rm closure}\,(\varSigma^{\prime\prime}),

that is, Σ\varSigma is the “gluing” of Σ\varSigma^{\prime} and Σ′′\varSigma^{\prime\prime} along the (n1)(n-1)-sphere Ss0(o)×{ϕ(s0)},S_{s_{0}}(o)\times\{\phi(s_{0})\}, which is CC^{\infty}-differentiable. Since the trajectories of ξ\nabla\xi are also C,C^{\infty}, for being geodesics, the resulting hypersurface Σ\varSigma is CC^{\infty}-differentiable with vertical tangent spaces along Ss0(o)×{ϕ(s0)}.S_{s_{0}}(o)\times\{\phi(s_{0})\}. Therefore, Σ\varSigma is a rotational strictly convex HrH_{r}-hypersurface of  𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which is homeomorphic to 𝕊n\mathbb{S}^{n} and is symmetric with respect to 𝔽m×{ϕ(s0)}.\mathbb{H}_{\mathbb{F}}^{m}\times\{\phi(s_{0})\}. This proves (i).

Under the hypotheses in (ii), it follows from Corollary 1 that τ\uptau satisfies:

τ(0)=0<τ|(0,+)<1,\uptau(0)=0<\uptau|_{(0,+\infty)}<1,

so that the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma with ϱ\varrho-function ϱ:=τ|[0,+)r\varrho:=\sqrt[r]{\uptau|_{[0,+\infty)}} is an entire rotational HrH_{r}-graph of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} over 𝔽m×{0}.\mathbb{H}_{\mathbb{F}}^{m}\times\{0\}. Since ϕ(0)=0\phi(0)=0 and ϕ(s)>0\phi(s)>0 for any s>0,s>0, Σ\varSigma is contained in the closed half-space 𝔽m×[0,+)\mathbb{H}_{\mathbb{F}}^{m}\times[0,+\infty) and is tangent to 𝔽m×{0}\mathbb{H}_{\mathbb{F}}^{m}\times\{0\} at o.o. Also, the height function of Σ\varSigma is unbounded above. Indeed, from Proposition 1-(i), τ\uptau, and so ϱ\varrho, is increasing. Thus, for a fixed δ>0,\delta>0, and any s>δ,s>\delta, one has

ϕ(s)=0sϱ(u)1ϱ2(u)𝑑uδsϱ(u)𝑑uϱ(δ)(sδ),\phi(s)=\int_{0}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du\geq\int_{\delta}^{s}\varrho(u)du\geq\varrho(\delta)(s-\delta),

which implies that ϕ\phi is unbounded above. Also, arguing as for the graph Σ\varSigma^{\prime} in the preceding paragraphs, we conclude that Σ\varSigma is strictly convex.

Observe that the mean curvature vector of Σ\varSigma “points upwards”, that is, its mean convex side Λ\Lambda is the connected component of (𝔽m×)Σ(\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R})-\varSigma which contains the axis {o}×.\{o\}\times\mathbb{R}. In particular, Λ\Lambda is foliated by the balls Bs(o)×{ϕ(s)},s(0,+).B_{s}(o)\times\{\phi(s)\},\,\,s\in(0,+\infty).

Refer to caption
Figure 4. After a downward translation of Σ~,\tilde{\varSigma}, it has a contact with Σ.\varSigma.

Let us suppose that there exists a compact HrH_{r}-hypersurface Σ~\widetilde{\varSigma} such that 0<HrC𝔽(r).0<H_{r}\leq C_{\mathbb{F}}(r). Considering the fact that Λ\Lambda is “horizontally and vertically unbounded”, it is easily seen that, after a suitable vertical translation, we can assume Σ~Λ\widetilde{\varSigma}\subset\Lambda (Fig. 4). Now, translate Σ~\widetilde{\varSigma} downward until it has a first contact with Σ.\varSigma. Since Σ\varSigma is strictly convex, the Maximum-Continuation Principle applies and gives that Σ\varSigma and Σ~\widetilde{\varSigma} coincide, which is clearly impossible. This shows that such a Σ~\widetilde{\varSigma} cannot exist and finishes the proof of (ii). ∎

Remark 1.

Let Σ𝔽m×\varSigma\subset\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} be an entire HrH_{r}-graph as in Theorem 1-(ii). Since C𝔽(n)=0C_{\mathbb{F}}(n)=0, we must have r<n.r<n. Also, the associated function τ:[0,+)\uptau:[0,+\infty)\rightarrow\mathbb{R} is positive, bounded and increasing, so that

lims+τ(s)=0,\lim_{s\rightarrow+\infty}\uptau^{\prime}(s)=0,

which implies that ϱ(s)0\varrho^{\prime}(s)\rightarrow 0 as s+.s\rightarrow+\infty. This, together with (11), gives that the principal curvature kn=ϱ(s)k_{n}=\varrho^{\prime}(s) goes to zero as s+.s\rightarrow+\infty. In particular, the least principal curvature function of Σ\varSigma is not bounded away from zero.

Next, we apply the method of (fs,ϕ)(f_{s},\phi)-graphs to produce one-parameter families of Hr(>0)H_{r}(>0)-annuli in 𝔽m×,\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, 1r<n.1\leq r<n. For that, fix s0=λ>0s_{0}=\lambda>0 and consider the solution τ(s)\uptau(s) of y=ay+by^{\prime}=ay+b given in (27), which satisfies the initial condition τ0=τ(λ)=1.\uptau_{0}=\uptau(\lambda)=1. From (19), we have

(34) τ(λ)=a(λ)+b(λ)=r(Hr|Hrλ|)|Hr1λ|\uptau^{\prime}(\lambda)=a(\lambda)+b(\lambda)=\frac{r(H_{r}-|H_{r}^{\lambda}|)}{|H_{r-1}^{\lambda}|}\,\cdot

Therefore, given Hr>0,H_{r}>0, r{1,,n1},r\in\{1,\dots,n-1\}, it follows from (34) that τ(λ)<0\uptau^{\prime}(\lambda)<0 if and only if Hr<|Hrλ|.H_{r}<|H_{r}^{\lambda}|. Thus, considering (26), we can define δHr\delta_{H_{r}} as the largest positive constant with the following property:

(35) λ(0,δHr)τ(λ)=1andτ(λ)<0.\lambda\in(0,\delta_{H_{r}})\quad\Leftrightarrow\quad\uptau(\lambda)=1\,\,\,\text{and}\,\,\,\uptau^{\prime}(\lambda)<0.

Observing that τ|[λ,+)\uptau|_{[\lambda,+\infty)} is a positive function for any λ(0,δHr),\lambda\in(0,\delta_{H_{r}}), we distinguish the cases:

  • i)

    There exists λ¯(λ,+)\mkern 1.5mu\overline{\mkern-1.5mu\lambda\mkern-1.5mu}\mkern 1.5mu\in(\lambda,+\infty) such that τ(λ¯)=1\uptau(\mkern 1.5mu\overline{\mkern-1.5mu\lambda\mkern-1.5mu}\mkern 1.5mu)=1  (Fig. 5a).

  • ii)

    τ(s)<1\uptau(s)<1 for all s(λ,+)s\in(\lambda,+\infty)  (Fig. 5b).

Refer to caption
a
Refer to caption
b
Figure 5. The two types of solutions τ\uptau, as in (27), satisfying τ(λ)=1\uptau(\lambda)=1.
Theorem 2.

Given r{1,,n1}r\in\{1,\dots,n-1\} and Hr>0,H_{r}>0, let δHr>0\delta_{H_{r}}>0 be as in (35). Then there exists a one-parameter family

𝒮={Σ(λ);λ(0,δHr)}\mathscr{S}\>=\{\varSigma(\lambda)\,;\,\lambda\in(0,\delta_{H_{r}})\}

of properly embedded rotational HrH_{r}-hypersurfaces in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which are all homeomorphic to the nn-annulus 𝕊n1×.\mathbb{S}^{n-1}\times\mathbb{R}. In addition, the following assertions hold:

  • i)

    If  Hr>C𝔽(r),H_{r}>C_{\mathbb{F}}(r), each Σ(λ)𝒮\varSigma(\lambda)\in\mathscr{S} is Delaunay-type, i.e., it is periodic in the vertical direction, and has unduloids as the trajectories of the gradient of its height function.

  • ii)

    If 0<HrC𝔽(r)0<H_{r}\leq C_{\mathbb{F}}(r), each hypersurface Σ(λ)𝒮\varSigma(\lambda)\in\mathscr{S} is symmetric with respect to 𝔽m×{0}\mathbb{H}_{\mathbb{F}}^{m}\times\{0\} and has unbounded height function.

Proof.

Fix λ(0,δHr)\lambda\in(0,\delta_{H_{r}}) and recall that τ\uptau is decreasing near λ.\lambda. Thus, if Hr>C𝔽(r),H_{r}>C_{\mathbb{F}}(r), it follows from (31) that there exists λ¯(λ,+)\bar{\lambda}\in(\lambda,+\infty) such that

0<τ|(λ,λ¯)<1=τ(λ¯)=τ(λ).0<\uptau|_{(\lambda,\bar{\lambda})}<1=\uptau(\bar{\lambda})=\uptau(\lambda).

Let us observe that a critical point s1s_{1} of τ\uptau is necessarily a minimum, since τ′′(s1)=a(s1)τ(s1)+b(s1)>0.\uptau^{\prime\prime}(s_{1})=a^{\prime}(s_{1})\uptau(s_{1})+b^{\prime}(s_{1})>0. Therefore, τ\uptau must have a unique local minimum at a point between λ\lambda and λ¯.\mkern 1.5mu\overline{\mkern-1.5mu\lambda\mkern-1.5mu}\mkern 1.5mu. In particular, τ(λ¯)>0\uptau^{\prime}(\bar{\lambda})>0 (Fig. 5a).

Refer to caption
Figure 6. The “block” of a Delaunay-type HrH_{r}-hypersurface in 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}.

Setting τλ:=τ|(λ,λ¯)\uptau_{\lambda}:=\uptau|_{(\lambda,\bar{\lambda})}, it follows from the above considerations and Lemmas 1 and 2 that the (fs,ϕ)(f_{s},\phi)-graph Σ(λ)\varSigma^{\prime}(\lambda) with ϱ\varrho-function ϱ(s)=τλr\varrho(s)=\sqrt[r]{\uptau_{\lambda}} is a bounded HrH_{r}-hypersurface of 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}. Moreover, Σ(λ)\varSigma^{\prime}(\lambda) is homeomorphic to 𝕊n1×(λ,λ¯)\mathbb{S}^{n-1}\times(\lambda,\bar{\lambda}) and has boundary (see Fig. 6):

Σ(λ)=(Sλ(o)×{0})(Sλ¯(o)×{ϕ(λ¯)}).\partial\varSigma^{\prime}(\lambda)=(S_{\lambda}(o)\times\{0\})\cup(S_{\bar{\lambda}}(o)\times\{\phi(\bar{\lambda})\}).

We also have that the tangent spaces of Σ(λ)\varSigma^{\prime}(\lambda) are vertical along its boundary Σ(λ)\partial\varSigma^{\prime}(\lambda), for ϱ(λ)=ϱ(λ¯)=1.\varrho(\lambda)=\varrho(\bar{\lambda})=1. Therefore, we obtain a properly embedded rotational HrH_{r}-hypersurface Σ(λ)\varSigma(\lambda) from Σ(λ)\varSigma^{\prime}(\lambda) by continuously reflecting it with respect to the horizontal hyperplanes 𝔽m×{kϕ(λ¯)},k.\mathbb{H}_{\mathbb{F}}^{m}\times\{k\phi(\bar{\lambda})\},\,k\in\mathbb{Z}. This proves (i).

Now, let us suppose that 0<HrC𝔽(r).0<H_{r}\leq C_{\mathbb{F}}(r). In this case, (31) gives that

0<τ|(λ,+)<1,0<\uptau|_{(\lambda,+\infty)}<1,

so that the (fs,ϕ)(f_{s},\phi)-graph Σ(λ)\varSigma^{\prime}(\lambda) determined by ϱ=τ|(λ,+)1/r\varrho=\uptau|_{(\lambda,+\infty)}^{1/r} is an HrH_{r}-hypersurface of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} with boundary Σ(λ)=Sλ(o)×{0}\partial\varSigma^{\prime}(\lambda)=S_{\lambda}(o)\times\{0\} (Fig. 7). By reflecting Σ(λ)\varSigma^{\prime}(\lambda) with respect to 𝔽m×{0},\mathbb{H}_{\mathbb{F}}^{m}\times\{0\}, as we did before, we obtain the embedded HrH_{r}-hypersurface Σ(λ)\varSigma(\lambda) as stated.

It remains to show that the height function of Σ(λ)\varSigma(\lambda) is unbounded. For that, we have just to observe that the infimum of τ\uptau in [λ,+)[\lambda,+\infty) is positive, since τ\uptau itself is positive in this interval, and its limit as s+s\rightarrow+\infty is Hr/C𝔽(r)>0.H_{r}/C_{\mathbb{F}}(r)>0. So, the same is true for ϱ=τ1/r.\varrho=\uptau^{1/r}. Therefore,

ϕ(s)=λsϱ(u)1ϱ2(u)𝑑u>λsϱ(u)𝑑u>infϱ|[λ,+)(sλ),\phi(s)=\int_{\lambda}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du>\int_{\lambda}^{s}\varrho(u)du>\inf\varrho|_{[\lambda,+\infty)}(s-\lambda),

from which we conclude that ϕ\phi is unbounded. ∎

Remark 2.

The case 𝔽=\mathbb{F}=\mathbb{R} of Theorem 1 was previously established in [18], whereas the case 𝔽=\mathbb{F}=\mathbb{R} and r=1r=1 of Theorem 2 was considered in [3]. Nevertheless, the methods employed in these works are different from ours, and are not applicable to the products 𝔽m×,\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, 𝔽.\mathbb{F}\neq\mathbb{R}.

Refer to caption
Figure 7. The (fs,ϕ)(f_{s},\phi)-graph Σ(λ),\varSigma^{\prime}(\lambda), on which all the trajectories of ξ\nabla\xi emanate from Σ(λ)\partial\varSigma^{\prime}(\lambda) orthogonally.

We proceed now to the classification of the complete rotational HrH_{r}-hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} whose height functions are Morse-type, i.e., have isolated critical points (if any). As we shall see, besides cylinders over geodesic spheres, these hypersurfaces are precisely the ones we obtained in Theorems 1 and 2. In particular, any of them is embedded. We point out that, in [3], it was shown that, for any H1>0,H_{1}>0, there exist complete rotational H1H_{1}-hypersurfaces in n×\mathbb{H}^{n}\times\mathbb{R} which are not embedded. In accordance with our results, the height function of none of these H1H_{1}-hypersurfaces is Morse-type.

Firstly, let us recall that the HrH_{r}-hypersurfaces in Theorems 1 and 2 were constructed from a single (fs,ϕ)(f_{s},\phi)-graph whose associated τ\uptau-function is a solution of the ODE y=ay+b,y^{\prime}=ay+b, where aa and bb are as in (19). For such a τ\uptau, there is a maximal interval (s0,s1)(s_{0},s_{1}), 0s0<s1+,0\leq s_{0}<s_{1}\leq+\infty, such that 0<τ|(s0,s1)<1.0<\uptau|_{(s_{0},s_{1})}<1.

Notice that each choice of HrH_{r} determines the function bb and, so, the equation y=ay+b.y^{\prime}=ay+b. The corresponding graph, then, is determined by the ordering of the constants HrH_{r} and C𝔽(r),C_{\mathbb{F}}(r), as well as by the values of s0s_{0} and τ(s0).\uptau(s_{0}).

Below, we list all the occurrences of s0s_{0} and τ(s0)\uptau(s_{0}) in Theorems 1 and 2 with respect to the ordering of HrH_{r} and C(r)C_{\mathscr{F}}(r):

  • C1)

    s0=0s_{0}=0, τ(s0)=0\uptau(s_{0})=0, Hr>C𝔽(r)H_{r}>C_{\mathbb{F}}(r).

  • C2)

    s0=0s_{0}=0, τ(s0)=0\uptau(s_{0})=0, HrC𝔽(r)H_{r}\leq C_{\mathbb{F}}(r).

  • C3)

    s0>0s_{0}>0, τ(s0)=1\uptau(s_{0})=1, Hr>C𝔽(r)H_{r}>C_{\mathbb{F}}(r).

  • C4)

    s0>0s_{0}>0, τ(s0)=1\uptau(s_{0})=1, HrC𝔽(r)H_{r}\leq C_{\mathbb{F}}(r).

The cases C1 and C2 correspond to Theorem 1-(i) and Theorem 1-(ii), respectively, whereas C3 and C4 correspond to Theorem 2-(i) and Theorem 2-(ii). We also remark that s1<+s_{1}<+\infty in cases C1 and C3, with τ(s1)=1,\uptau(s_{1})=1, and that s1=+s_{1}=+\infty in cases C2 and C4.

Let M0M_{0} be a hypersurface of a Riemannian manifold M.M. It is easily seen that Σ:=M0×\varSigma:=M_{0}\times\mathbb{R} is a hypersurface of M×M\times\mathbb{R} whose tangent spaces are all vertical, so that t\partial_{t} is a principal direction of Σ\varSigma with vanishing principal curvature. In particular, Hn=0H_{n}=0 on Σ.\varSigma. Also, for all r{1,,n1},r\in\{1,\dots,n-1\}, the rr-th mean curvatures of M0M_{0} and Σ\varSigma at xM0x\in M_{0} and (x,t)Σ(x,t)\in\varSigma coincide. In particular, M0M_{0} is an Hr(<n)H_{r(<n)}-hypersurface of MM if and only if Σ\varSigma is an Hr(<n)H_{r(<n)}-hypersurface of M×.M\times\mathbb{R}. We call Σ:=M0×\varSigma:=M_{0}\times\mathbb{R} the cylinder over M0.M_{0}.

Theorem 3.

Let Σ\varSigma be a connected complete rotational Hr(>0)H_{r}(>0)-hypersurface of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} whose height function is Morse-type. Then, Σ\varSigma is either a cylinder over a geodesic sphere of  𝔽m\mathbb{H}_{\mathbb{F}}^{m} or one of the embedded HrH_{r}-hypersurfaces of Theorems 12.

Proof.

Suppose that Σ\varSigma is not a cylinder. In this case, we have that the open set Σ0Σ\varSigma_{0}\subset\varSigma on which Θξ\varTheta\nabla\xi never vanishes is nonempty. Since Σ0\varSigma_{0} contains no vertical points, for a given x0Σ0x_{0}\in\varSigma_{0} , there is an open neighborhood Σ\varSigma^{\prime} of x0x_{0} in Σ0\varSigma_{0} which is a graph over an open set Ω\Omega of 𝔽m\mathbb{H}_{\mathbb{F}}^{m}. Thus, since Σ\varSigma is rotational and Σ0\varSigma_{0} contains no horizontal points, after possibly a reflection with respect to a horizontal hyperplane, we can assume that Σ\varSigma^{\prime} is an (fs,ϕ)(f_{s},\phi)-graph over Ω.\Omega. (Recall that, in our setting, the ϕ\phi-function of an (fs,ϕ)(f_{s},\phi)-graph is required to be radially increasing.)

By Lemma 1, the function τ=ϱr\uptau=\varrho^{r} associated to Σ\varSigma^{\prime} is a solution of y=ay+b,y^{\prime}=ay+b, with aa and bb as in (19). In addition, since Σ\varSigma is complete, there exists a maximal interval (s0,s1)(s_{0},s_{1}), 0s0<s1+,0\leq s_{0}<s_{1}\leq+\infty, such that 0<τ|(s0,s1)<1.0<\uptau|_{(s_{0},s_{1})}<1. In particular, we have the following two possibilities:

τ(s0)=0orτ(s0)=1.\uptau(s_{0})=0\quad\text{or}\quad\uptau(s_{0})=1.

Suppose that τ(s0)=0.\uptau(s_{0})=0. After a vertical translation, we can assume that

ϕ(s)=s0sϱ(u)1ϱ2(u)𝑑u.\phi(s)=\int_{s_{0}}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du.

In particular, ϕ(s0)=ϕ(s0)=0.\phi(s_{0})=\phi^{\prime}(s_{0})=0. If s0>0,s_{0}>0, these equalities imply that the sphere Ss0(o)×{0}S_{s_{0}}(o)\times\{0\} of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} is contained in Σ,\partial\varSigma^{\prime}, and that ξ\nabla\xi vanishes at all of its points. This, however, contradicts that the height function of Σ\varSigma is Morse-type. Hence, s0=0,s_{0}=0, so that the τ\uptau-function of Σ\varSigma^{\prime} satisfies the initial condition τ(0)=0.\uptau(0)=0.

If Hr>C𝔽(r),H_{r}>C_{\mathbb{F}}(r), by the uniqueness of solutions of linear ODE’s satisfying an initial condition, the function τ\uptau such that τ(0)=0\uptau(0)=0 coincides with the one in case C1 above. Thus, the corresponding ϕ\phi-functions also coincide, which clearly implies that Σ\varSigma^{\prime} is an open set of the (strictly convex) HrH_{r}-sphere obtained in Theorem 1-(i). Therefore, by the Maximum-Continuation Principle, Σ\varSigma coincides with this HrH_{r}-sphere. If HrC𝔽(r),H_{r}\leq C_{\mathbb{F}}(r), then τ\uptau coincides with the solution of case C2. Analogously, we conclude that Σ\varSigma is an entire graph as in Theorem 1-(ii).

Let us suppose now that τ(s0)=1.\uptau(s_{0})=1. Since 0<τ<10<\uptau<1 in (s0,s1),(s_{0},s_{1}), τ\uptau is decreasing near s0s_{0} , which implies that r<n.r<n. (Indeed, for r=n,r=n, we have τ=b>0.\uptau^{\prime}=b>0.) In this case, as we have discussed, |a(s)|+|a(s)|\rightarrow+\infty as s0s\rightarrow 0 (cf. (28)), and b(0)b(0) is 0 or H1H_{1}. In particular, any solution τ\uptau of y=ay+by^{\prime}=ay+b at s=0s=0 must satisfy τ(0)=0,\uptau(0)=0, so that s00.s_{0}\neq 0. Hence, s0(0,δHr),s_{0}\in(0,\delta_{H_{r}}), where δHr\delta_{H_{r}} is the positive constant defined in (35).

Setting s0=λs_{0}=\lambda and observing that any of the hypersurfaces obtained in Theorem 2 is strictly convex at some of its points, we can argue as in the second from the last paragraph and conclude that Σ\varSigma is the HrH_{r}-hypersurface Σ(λ)\varSigma(\lambda) of Theorem 2-(i) or Theorem 2-(ii), according to whether Hr>C𝔽(r)H_{r}>C_{\mathbb{F}}(r) or HrC𝔽(r).H_{r}\leq C_{\mathbb{F}}(r). This finishes the proof. ∎

4.2. Rotational Hr(>0)H_{r}(>0)-hypersurfaces of 𝕊n×\mathbb{S}^{n}\times\mathbb{R}

In this section, we apply the method of (fs,ϕ)(f_{s},\phi)-graphs to construct and classify rotational Hr(>0)H_{r}(>0)-hypersurfaces in 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}.

As we did before, let us fix a point o𝕊no\in\mathbb{S}^{n} and consider a family

(36) :={fs:𝕊n1𝕊n;s(0,π)}\mathscr{F}:=\{f_{s}:\mathbb{S}^{n-1}\rightarrow\mathbb{S}^{n}\,;\,s\in(0,\pi)\}

of parallel geodesic spheres fsf_{s} of 𝕊n\mathbb{S}^{n} with radius ss and center o.o. As is well known, each fsf_{s} is totally umbilical, having principal curvatures all equal to cots-\cot s with respect to the outward orientation. In particular, \mathscr{F} is isoparametric.

From a direct computation, we get that the coefficients aa and bb of the ODE y=ay+by^{\prime}=ay+b determined by \mathscr{F} and any given Hr>0H_{r}>0 are

(37) a(s)=(nr)cotsandb(s)=brtanr1(s),br=rHr(n1r1)1,a(s)=-(n-r)\cot s\quad\text{and}\quad b(s)=b_{r}\tan^{r-1}(s)\,,\,\,\,b_{r}=rH_{r}{{{n-1}\choose{r-1}}}^{-1},

and that the corresponding general solution is:

(38) τ(s):=(sins0sins)nr(τ0+brsinnr(s0)s0ssinn1(u)cosr1(u)𝑑u),s0,s(0,),\uptau(s):=\left(\frac{\sin s_{0}}{\sin s}\right)^{n-r}\left(\uptau_{0}+\frac{b_{r}}{\sin^{n-r}(s_{0})}\int_{s_{0}}^{s}\frac{\sin^{n-1}(u)}{\cos^{r-1}(u)}du\right),\,\,s_{0}\,,s\in(0,\mathcal{R}),

where τ0=τ(s0)\uptau_{0}=\uptau(s_{0})\in\mathbb{R} and

:={π/2ifr>1πifr=1.\mathcal{R}:=\left\{\begin{array}[]{ccc}\pi/2&\text{if}&r>1\\ \pi&\text{if}&r=1\end{array}\right..

Also, it is easily checked that

(39) τ(s):={brsinnr(s)0ssinn1(u)cosr1(u)𝑑uifs(0,)0ifs=0\uptau(s):=\left\{\begin{array}[]{lcl}\frac{b_{r}}{\sin^{n-r}(s)}\int_{0}^{s}\frac{\sin^{n-1}(u)}{\cos^{r-1}(u)}du&\text{if}&s\in(0,\mathcal{R})\\[6.45831pt] 0&\text{if}&s=0\end{array}\right.

is a well defined solution of y=ay+by^{\prime}=ay+b satisfying y(0)=0.y(0)=0.

Given an integer n2,n\geq 2, it will be convenient to introduce the following constant:

(40) S(n)=0π/2sinn1(s)𝑑s,n2.S(n)=\int_{0}^{\pi/2}\sin^{n-1}(s)ds,\,\,n\geq 2.
Proposition 2.

Let τ\uptau be the solution (39). Then, the following hold:

  • i)

    τ>0\uptau^{\prime}>0 in (0,).(0,\mathcal{R}).

  • ii)

    limsτ(s)=+.\displaystyle\lim_{s\rightarrow\mathcal{R}}\uptau(s)=+\infty.

  • iii)

    limsπ2τ(s)=H1S(n)\displaystyle\lim_{s\rightarrow\frac{\pi}{2}}\uptau(s)=H_{1}S(n) if r=1.r=1.

Proof.

Since the functions aa and bb in (37) are both increasing (when nonconstant), the proof of (i) is entirely analogous to the one given in Proposition 1-(i).

To prove (ii), let us first assume r=1.r=1. In this case, since sinπ=0\sin\pi=0 and the integral 0πsinn1(u)𝑑u\int_{0}^{\pi}\sin^{n-1}(u)du is positive, we have that τ\uptau satisfies (ii) for =π.\mathcal{R}=\pi.

If r>1r>1, for a fixed δ(0,π/2)\delta\in(0,\pi/2) and any s(δ,π/2),s\in(\delta,\pi/2), one has

0ssinn1(u)cosr1(u)𝑑uδstanr1(u)sinnr(u)𝑑usinnr(δ)δstanr1(u)𝑑u,\int_{0}^{s}\frac{\sin^{n-1}(u)}{\cos^{r-1}(u)}du\geq\int_{\delta}^{s}\tan^{r-1}(u){\sin^{n-r}(u)}du\geq\sin^{n-r}(\delta)\int_{\delta}^{s}\tan^{r-1}(u)du,

which implies that the first integral goes to infinity as sπ/2,s\rightarrow\pi/2, since the same is true for the integral δstanr1(u)𝑑u\int_{\delta}^{s}\tan^{r-1}(u)du. It follows from this fact that τ(s)+\uptau(s)\rightarrow+\infty as sπ/2s\rightarrow\pi/2 if r>1,r>1, which proves (ii).

The identity in (iii) follows directly from the definitions of τ\uptau (for r=1r=1) and S(n)S(n) (as in (40)). ∎

From the above proposition, we get the following existence result for Hr(>0)H_{r}(>0)-hypersurfaces of 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}.

Theorem 4.

Given r{1,,n}r\in\{1,\dots,n\} and a constant Hr>0,H_{r}>0, there exists an HrH_{r}-sphere Σ\varSigma in 𝕊n×\mathbb{S}^{n}\times\mathbb{R} which is symmetric with respect to a horizontal hyperplane. Furthermore:

  • i)

    Σ\varSigma is strictly convex if r>1r>1 or r=1r=1 and H1>1/S(n).H_{1}>1/S(n).

  • ii)

    Σ\varSigma is convex if r=1r=1 and H1=1/S(n).H_{1}=1/S(n).

  • iii)

    Σ\varSigma is non convex if r=1r=1 and 0<H1<1/S(n).0<H_{1}<1/S(n).

Proof.

Let \mathscr{F} be an arbitrary family of parallel geodesic spheres of 𝕊n\mathbb{S}^{n} as given in (36). Consider the functions aa and bb defined in (37) and let τ\uptau be the solution (39) of the ODE y=ay+b.y^{\prime}=ay+b.

From Proposition 2-(ii), there exists a positive s0<s_{0}<\mathcal{R} such that

0=τ(0)<τ|(0,s0)<1=τ(s0),0=\uptau(0)<\uptau|_{(0,s_{0})}<1=\uptau(s_{0}),

so that τ|[0,s0)\uptau|_{[0,s_{0})} determines an (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma^{\prime} over Bs0(0)𝕊n.B_{s_{0}}(0)\subset\mathbb{S}^{n}. Since τ(s0)=1\uptau(s_{0})=1 and τ(s0)>0\uptau^{\prime}(s_{0})>0 (by Proposition 2-(i)), we can proceed just as in the proof of Theorem 1-(i) to obtain from Σ\varSigma^{\prime} the embedded HrH_{r}-sphere Σ\varSigma of 𝕊n×\mathbb{S}^{n}\times\mathbb{R} which is symmetric with respect to Pϕ(s0):=𝕊n×{ϕ(s0)}.P_{\phi(s_{0})}:=\mathbb{S}^{n}\times\{\phi(s_{0})\}.

If r>1r>1 or r=1r=1 and H1>1/S(n),H_{1}>1/S(n), we have from Proposition 2, items (ii) and (iii), that 0<s0<π/2.0<s_{0}<\pi/2. Hence, for s(0,s0),s\in(0,s_{0}), all spheres fsf_{s} have negative principal curvatures, which, together with equalities (11), gives that Σ\varSigma is strictly convex. This proves (i).

If r=1r=1 and H1=1/S(n),H_{1}=1/S(n), Proposition 2-(iii) yields s0=π/2.s_{0}=\pi/2. However, fπ/2f_{\pi/2} is totally geodesic in 𝕊n,\mathbb{S}^{n}, which implies that, except for kn=H1>0,k_{n}=H_{1}>0, the principal curvatures of Σ\varSigma vanish at all points of the horizontal section Σϕ(π/2)=ΣPϕ(π/2).\varSigma_{\phi(\pi/2)}=\varSigma\cap P_{\phi(\pi/2)}. Therefore, Σ\varSigma is convex on Σϕ(π/2)\varSigma_{\phi(\pi/2)} and strictly convex on ΣΣϕ(π/2)\varSigma-\varSigma_{\phi(\pi/2)} .

Finally, assuming r=1r=1 and 0<H1<1/S(n),0<H_{1}<1/S(n), we have from Proposition 2-(iii) that s0>π/2.s_{0}>\pi/2. Observing that, for π/2<s<π,\pi/2<s<\pi, fsf_{s} has positive principal curvatures, we conclude, as in the last paragraph, that Σ\varSigma is strictly convex (resp. convex, non convex) on Σϕ(s)\varSigma_{\phi(s)} if s<π/2s<\pi/2 (resp. s=π/2,s=\pi/2, s>π/2s>\pi/2). In particular, Σ\varSigma is non convex. This shows (iii) and concludes our proof. ∎

Remark 3.

Except for the assumptions on the convexity of Σ,\varSigma, the case r=1r=1 of Theorem 4 was proved in [30]. The case r=n=2r=n=2 was considered in [9]. It should also be mentioned that, for n=2n=2 and r=1,r=1, the non convexity of Σ\varSigma as stated in (iii) was pointed out in [1, Remark 2.8].

In our next theorem we show the existence of one-parameter families of rotational Delaunay-type Hr(>0)H_{r}(>0)-annuli in 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}. This result, then, generalizes the analogous one obtained in [31] for r=1.r=1.

First, let us introduce the constant

Cr:=nrn(nr)C_{r}:=\frac{n-r}{n}{{n}\choose{r}}

and observe that, for 1r<n,1\leq r<n, the positive constants HrH_{r} , brb_{r} (as in (37)), and CrC_{r} satisfy the following relation:

(41) nrbr=CrHr\frac{n-r}{b_{r}}=\frac{C_{r}}{H_{r}}\,\cdot

In this setting, if we define

(42) δHr:=arctan(Cr/Hr)1/r(0,π/2),\delta_{H_{r}}:={\rm arctan}\,(C_{r}/H_{r})^{1/r}\,\in(0,\pi/2),

then a solution τ\uptau of y=ay+by^{\prime}=ay+b such that τ(s0)=1\uptau(s_{0})=1, s0(0,π/2),s_{0}\in(0,\pi/2), satisfies:

(43) τ(s0)<00<s0<δHr.\uptau^{\prime}(s_{0})<0\quad\Leftrightarrow\quad 0<s_{0}<\delta_{H_{r}}.
Theorem 5.

Given n2,n\geq 2, r{1,,n1}r\in\{1,\dots,n-1\}, and Hr>0,H_{r}>0, there exists a one-parameter family  𝒮={Σ(λ); 0<λ<δHr}\mathscr{S}=\{\varSigma(\lambda)\,;\,0<\lambda<\delta_{H_{r}}\} of properly embedded Delaunay-type rotational HrH_{r}-hypersurfaces in 𝕊n×\mathbb{S}^{n}\times\mathbb{R}.

Proof.

Given λ(0,δHr),\lambda\in(0,\delta_{H_{r}}), consider the solution τ\uptau as in (38) such that s0=λs_{0}=\lambda and τ0=τ(λ)=1.\uptau_{0}=\uptau(\lambda)=1. From (43), we have that τ\uptau is decreasing in a neighborhood of λ.\lambda.

Observe that τ\uptau is positive in (0,).(0,\mathcal{R}). Also, setting

μ(s)=(sinssinλ)nr,\mu(s)=\left(\frac{\sin s}{\sin\lambda}\right)^{n-r}\,,

we have that μ>1\mu>1 on (λ,π/2).(\lambda,\pi/2). So, for r>1r>1 and s>λ,s>\lambda,

τ(s)>1μ(s)(1+λsb(u)𝑑u)=1μ(s)(1+brλstanr1(u)𝑑u),\uptau(s)>\frac{1}{\mu(s)}\left(1+\int_{\lambda}^{s}b(u)du\right)=\frac{1}{\mu(s)}\left(1+b_{r}\int_{\lambda}^{s}\tan^{r-1}(u)du\right),

which implies that τ(s)+\uptau(s)\rightarrow+\infty as sπ/2.s\rightarrow\pi/2.

If r=1,r=1, since sinπ=0\sin\pi=0 and λπμ(s)𝑑s\int_{\lambda}^{\pi}\mu(s)ds is positive, we have that τ(s)+\uptau(s)\rightarrow+\infty as sπ.s\rightarrow\pi.

It follows from the above considerations that there exists λ¯(0,)\mkern 1.5mu\overline{\mkern-1.5mu\lambda\mkern-1.5mu}\mkern 1.5mu\in(0,\mathcal{R}) such that

(44) τ(λ)=τ(λ¯)=1andτ(λ)<0<τ(λ¯).\uptau(\lambda)=\uptau(\bar{\lambda})=1\,\,\,\text{and}\,\,\,\uptau^{\prime}(\lambda)<0<\uptau^{\prime}(\bar{\lambda}).

From this point on, the proof is entirely analogous to that of Theorem 2-(i). ∎

A classification result for rotational HrH_{r}-hypersurfaces of 𝕊n×\mathbb{S}^{n}\times\mathbb{R} can be achieved in the same way we did for their congeners in 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}. To see this, assume that Σ\varSigma is a complete connected rotational Hr(>0)H_{r}(>0)-hypersurface of 𝕊n×\mathbb{S}^{n}\times\mathbb{R} whose height function is Morse-type. Assuming that Σ\varSigma is non cylindrical, we have, as before, that there exists an open set ΣΣ\varSigma^{\prime}\subset\varSigma which is an (fs,ϕ)(f_{s},\phi)-graph, fs.f_{s}\in\mathscr{F}. The corresponding τ\uptau-function, restricted to a maximal interval (s0,s1),(s_{0},s_{1}), satisfies:

0<τ|(s0,s1)<1,   0s0<s1,0<\uptau|_{(s_{0}\,,s_{1})}<1,\,\,\,0\leq s_{0}<s_{1}\leq\mathcal{R},

which yields τ(s0)=0\uptau(s_{0})=0 or τ(s0)=1.\uptau(s_{0})=1.

If τ(s0)=0,\uptau(s_{0})=0, then s0=0.s_{0}=0. (Otherwise, the height function of Σ\varSigma would not be Morse-type.) In this case, τ\uptau coincides with the τ\uptau-function of the HrH_{r}-sphere of Theorem 4, and then Σ\varSigma itself coincides with this sphere. (Notice that any of the spheres obtained in Theorem 4 is strictly convex on an open set.)

If τ(s0)=1,\uptau(s_{0})=1, then τ\uptau is decreasing in a neighborhood of s0s_{0} . Thus, r<n.r<n. In particular, |a(s)|+|a(s)|\rightarrow+\infty as s0,s\rightarrow 0, so that s0(0,δHr).s_{0}\in(0,\delta_{H_{r}}). Analogously, this gives that Σ\varSigma coincides with the HrH_{r}-annulus Σ(λ)\varSigma(\lambda) of Theorem 5, λ=s0.\lambda=s_{0}.

Summarizing, we have the following result.

Theorem 6.

Let Σ\varSigma be a connected complete rotational Hr(>0)H_{r}(>0)-hypersurface of 𝕊n×\mathbb{S}^{n}\times\mathbb{R} whose height function is Morse-type. Then, Σ\varSigma is either a cylinder over a strictly convex geodesic sphere of  𝕊n\mathbb{S}^{n} or one of the embedded HrH_{r}-hypersurfaces of Theorems 45.

Remark 4.

Regarding the hypothesis on the height function of Σ\varSigma in Theorem 6, we point out that a rotational embedded H1(>0)H_{1}(>0)-torus in 𝕊n×\mathbb{S}^{n}\times\mathbb{R} whose height function is non Morse-type was obtained in [30].

5. Rotational rr-minimal Hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} and 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}.

In this section, we shall see that the method of (fs,ϕ)(f_{s},\phi)-graphs can be used for construction and classification of rotational rr-minimal hypersurfaces of  𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} and 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}. A major distinction from the case of Hr(>0)H_{r}(>0)-hypersurfaces is that the Maximum-Continuation Principle is no longer available.

Theorem 7.

Given r{1,,n},r\in\{1,\dots,n\}, there exists a one-parameter family

𝒮={Σ(λ);λ>0}\mathscr{S}=\{\varSigma(\lambda)\,;\,\lambda>0\}

of complete rotational rr-minimal nn-annuli in  𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} with the following properties:

  • i)

    If r=n,r=n, Σ(λ)\varSigma(\lambda) is a cylinder over a geodesic sphere of  𝔽m\mathbb{H}_{\mathbb{F}}^{m} of radius λ>0.\lambda>0.

  • ii)

    If r<nr<n, Σ(λ)\varSigma(\lambda) is catenoid-type. More precisely, it is symmetric with respect to P0=𝔽m×{0},P_{0}=\mathbb{H}_{\mathbb{F}}^{m}\times\{0\}, and P0Σ(λ)P_{0}\cap\varSigma(\lambda) is the geodesic sphere of  𝔽m\mathbb{H}_{\mathbb{F}}^{m} of radius λ\lambda centered at the point o𝔽mo\in\mathbb{H}_{\mathbb{F}}^{m} of the axis. In addition, each of the parts of Σ(λ)\varSigma(\lambda) above and below P0P_{0} is a rotational graph over 𝔽mBλ(o)\mathbb{H}_{\mathbb{F}}^{m}-B_{\lambda}(o) (Fig. 8).

Furthermore, up to ambient isometries, any complete connected rotational rr-minimal hypersurface of  𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} is either a member of 𝒮\mathscr{S} or a horizontal hyperplane.

Proof.

Given λ>0,\lambda>0, is immediate that a cylinder over a geodesic sphere of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} of radius λ\lambda is an nn-minimal rotational annulus of 𝔽m×,\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, which yields (i).

Assume that 1r<n1\leq r<n and let ={fs;s(0,+)}\mathscr{F}=\{f_{s}\,;\,s\in(0,+\infty)\} be the parallel family of geodesic spheres of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} centered at the axis point o𝔽m.o\in\mathbb{H}_{\mathbb{F}}^{m}. The ODE determined by \mathscr{F} and Hr=0H_{r}=0 is given by

(45) y=ay,a(s)=r|Hrs||Hrs1|,s(0,+).y^{\prime}=ay,\quad a(s)=-\frac{r|H_{r}^{s}|}{|H_{r}^{s-1}|}\,,\,\,\,s\in(0,+\infty).

Since a<0,a<0, given λ>0,\lambda>0, the function

τλ(s)=exp(λsa(u)𝑑u),s[λ,+),\uptau_{\lambda}(s)=\exp\left(\int_{\lambda}^{s}a(u)du\right),\,\,\,s\in[\lambda,+\infty),

is clearly a solution of (45) which satisfies

0<τλ(s)τλ(λ)=1s[λ,+).0<\uptau_{\lambda}(s)\leq\uptau_{\lambda}(\lambda)=1\,\,\,\forall s\in[\lambda,+\infty).

In addition, τλ(λ)=a(λ)<0.\uptau_{\lambda}^{\prime}(\lambda)=a(\lambda)<0. So, setting ϱλ=τλ1/r,\varrho_{\lambda}=\uptau_{\lambda}^{1/r}, it follows from Lemma 2 that

ϕλ(s):=λsϱλ(u)1ϱλ2(u)𝑑u,s[λ,+),\phi_{\lambda}(s):=\int_{\lambda}^{s}\frac{\varrho_{\lambda}(u)}{\sqrt{1-\varrho_{\lambda}^{2}(u)}}du,\,\,\,s\in[\lambda,+\infty),

is well defined. Therefore, by Lemma 1, the (fs,ϕλ)(f_{s},\phi_{\lambda})-graph Σ(λ)\varSigma^{\prime}(\lambda) is an rr-minimal hypersurface of 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}. Notice that Σ(λ)\varSigma^{\prime}(\lambda) is a graph over 𝔽mBλ(o)¯\mathbb{H}_{\mathbb{F}}^{m}-\mkern 1.5mu\overline{\mkern-1.5muB_{\lambda}(o)\mkern-1.5mu}\mkern 1.5mu with boundary Σ(λ)=Sλ(o)\partial\varSigma^{\prime}(\lambda)=S_{\lambda}(o) (Fig. 8).

Refer to caption
Figure 8. The half rr-minimal catenoid Σ(λ),\varSigma^{\prime}(\lambda), on which all the trajectories of ξ\nabla\xi emanate from Σ(λ)\partial\varSigma^{\prime}(\lambda) orthogonally.

Also, since ϱλ(λ)=1,\varrho_{\lambda}(\lambda)=1, the tangent spaces of Σ(λ)\varSigma^{\prime}(\lambda) along Σ(λ)\partial\varSigma^{\prime}(\lambda) are all vertical. Thus, considering the reflection Σ′′(λ)\varSigma^{\prime\prime}(\lambda) of Σ(λ)\varSigma^{\prime}(\lambda) with respect to 𝔽m×{0},\mathbb{H}_{\mathbb{F}}^{m}\times\{0\}, as before, we have that Σ(λ):=closure(Σ(λ))closure(Σ′′(λ))\varSigma(\lambda):={\rm closure}\,(\varSigma^{\prime}(\lambda))\cup{\rm closure}\,(\varSigma^{\prime\prime}(\lambda)) is the desired rr-minimal hypersurface.

Suppose now that Σ\varSigma is a complete connected rotational rr-minimal hypersurface of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, r{1,,n},r\in\{1,\dots,n\}, and set

Σ0:={xΣ;Θ(x)ξ(x)0}.\varSigma_{0}:=\{x\in\varSigma\,;\,\varTheta(x)\nabla\xi(x)\neq 0\}.

Notice that Σ\varSigma is either a horizontal hyperplane or a cylinder if and only if Σ0=.\varSigma_{0}=\emptyset. So, we can assume Σ0.\varSigma_{0}\neq\emptyset. We can also assume, without loss of generality, that Σ\varSigma and all members of 𝒮\mathscr{S} share the same axis {o}×.\{o\}\times\mathbb{R}.

As we argued in previous proofs, under the above hypotheses, there exists an (fs,ϕ)(f_{s},\phi)-graph ΣΣ0\varSigma^{\prime}\subset\varSigma_{0} and a maximal open interval (s0,s1),(s_{0},s_{1}), 0s0<s1+,0\leq s_{0}<s_{1}\leq+\infty, such that the τ\uptau-function of Σ\varSigma^{\prime} satisfies 0<τ|(s0,s1)<1.0<\uptau|_{(s_{0},s_{1})}<1.

For r=n,r=n, we have that τ,\uptau, and so ϱ\varrho, is constant. Hence, up to a vertical translation, one has ϕ(s)=cs,s>0,\phi(s)=cs,\,\,s>0, for some constant c>0.c>0. However, ϕ(0)=c>0,\phi^{\prime}(0)=c>0, which implies that the closure of Σ\varSigma^{\prime} in Σ\varSigma meets the rotation axis non orthogonally, i.e., Σ\varSigma is not smooth at Σ\partial\varSigma^{\prime} — a contradiction. So, Σ0=\varSigma_{0}=\emptyset if r=n.r=n.

For r<n,r<n, we have that τ\uptau is a solution of (45). In particular, τ\uptau is decreasing, which implies that τ(s0)=1.\uptau(s_{0})=1. As before, this yields s0>0.s_{0}>0. Thus, setting s0=λ,s_{0}=\lambda, we have τ=τλ\uptau=\uptau_{\lambda}, which implies that, up to a vertical translation, ϕ=ϕλ\phi=\phi_{\lambda} and, then, Σ\varSigma^{\prime} coincides with the half-catenoid Σ(λ).\varSigma^{\prime}(\lambda).

We conclude from the above that Σ0\varSigma_{0} is the union of open half-catenoids Σ(λ)\varSigma^{\prime}(\lambda), where Σ(λ)𝒮.\varSigma(\lambda)\in\mathscr{S}. In addition, Σ1:=ΣΣ0\varSigma_{1}:=\varSigma-\varSigma_{0} has empty interior in Σ.\varSigma. Otherwise, there would exist a nonempty open set UΣ1U\subset\varSigma_{1} of Σ\varSigma which would be either horizontal or vertical, and whose boundary in Σ\varSigma would be contained in some half-catenoid. Since catenoids have no horizontal points, UU should be vertical and, so, part of a vertical rotational cylinder. However, rotational cylinders in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} are rr-minimal if and only if r=n.r=n. Therefore, Σ1\varSigma_{1} has empty interior, which implies that Σ0\varSigma_{0} is open and dense in Σ.\varSigma. Clearly, the intersection of two distinct members of 𝒮\mathscr{S} is always transversal. This, together with the connectedness of Σ\varSigma and the density of Σ0\varSigma_{0}  in Σ,\varSigma, gives that Σ\varSigma coincides with some Σ(λ)𝒮\varSigma(\lambda)\in\mathscr{S}, which concludes our proof. ∎

Theorem 8.

Given r{1,,n},r\in\{1,\dots,n\}, there exists a one-parameter family

𝒮={Σ(λ); 0<λ<0}\mathscr{S}=\{\varSigma(\lambda)\,;\,0<\lambda<\mathcal{R}_{0}\}

of complete rotational rr-minimal nn-annuli in  𝕊n×\mathbb{S}^{n}\times\mathbb{R} with the following properties:

  • i)

    If r=n,r=n, then 0=π\mathcal{R}_{0}=\pi and Σ(λ)\varSigma(\lambda) is a cylinder over a geodesic sphere of  𝕊n\mathbb{S}^{n} of radius λ.\lambda.

  • ii)

    If r<nr<n, then 0=π/2\mathcal{R}_{0}=\pi/2 and Σ(λ)\varSigma(\lambda) is Delaunay-type.

Furthermore, up to ambient isometries, any complete connected rotational rr-minimal hypersurface of  𝕊n×\mathbb{S}^{n}\times\mathbb{R} is either a member of 𝒮\mathscr{S} or a horizontal hyperplane.

Proof.

Statement (i) is trivial. So, assume r<nr<n and let ={fs;s(0,π)}\mathscr{F}=\{f_{s}\,;\,s\in(0,\pi)\} be the family of parallel geodesic spheres of 𝕊n\mathbb{S}^{n} centered at some point o𝕊n.o\in\mathbb{S}^{n}. In this setting, the ODE determined by \mathscr{F} and Hr=0H_{r}=0 is

(46) y=ay,a(s)=(nr)cots,s(0,π).y^{\prime}=ay,\quad a(s)=-(n-r)\cot s,\quad s\in(0,\pi).

Given λ(0,π/2),\lambda\in(0,\pi/2), the function

τλ(s)=(sinλsins)nr,s(0,π),\uptau_{\lambda}(s)=\left(\frac{\sin\lambda}{\sin s}\right)^{n-r}\,,\,\,s\in(0,\pi),

is easily seen to be the solution of (46) satisfying

0<τλ|(λ,πλ)<1=τ(λ)=τ(πλ).0<\uptau_{\lambda}|_{(\lambda,\pi-\lambda)}<1=\uptau(\lambda)=\uptau(\pi-\lambda).

Henceforth, the reasoning in the proof of Theorem 5 applies and leads to the construction of the Delaunay-type rr-minimal hypersurface Σ(λ)\varSigma(\lambda) as stated in (ii).

Now, suppose that Σ\varSigma is a complete connected rotational rr-minimal hypersurface of 𝕊n×.\mathbb{S}^{n}\times\mathbb{R}. In this setting, define

Σ0:={xΣ;Θ(x)ξ(x)0}.\varSigma_{0}:=\{x\in\varSigma\,;\,\varTheta(x)\nabla\xi(x)\neq 0\}.

As in the preceding proof, Σ0=\varSigma_{0}=\emptyset if r=n.r=n. Thus, in this case, Σ\varSigma is either a horizontal hyperplane or a cylinder over a geodesic sphere of 𝕊n.\mathbb{S}^{n}.

Suppose that r<nr<n and that the axis of Σ\varSigma is {o}×.\{o\}\times\mathbb{R}. If Σ0,\varSigma_{0}\neq\emptyset, then Σ\varSigma is neither a horizontal hyperplane nor a cylinder. In addition, there exists an (fs,ϕ)(f_{s},\phi)-graph ΣΣ0\varSigma^{\prime}\subset\varSigma_{0} and a maximal interval (s0,s1),(s_{0},s_{1}), 0s0<s1π0\leq s_{0}<s_{1}\leq\pi, such that the τ\uptau-function of Σ\varSigma^{\prime} satisfies 0<τ|(s0,s1)<10<\uptau|_{(s_{0},s_{1})}<1. So, τ(s0)=0\uptau(s_{0})=0 or τ(s0)=1.\uptau(s_{0})=1.

The formula of the general solution of the ODE (46) gives that τ\uptau is positive, not defined at s=0,πs=0,\pi, and bounded away from zero. In particular, s00,s1πs_{0}\neq 0,s_{1}\neq\pi and τ(s0)=τ(s1)=1,\uptau(s_{0})=\uptau(s_{1})=1, so that τ\uptau is given by

τ(s)=(sins0sins)nr,s0,s(0,π).\uptau(s)=\left(\frac{\sin s_{0}}{\sin s}\right)^{n-r}\,,\,\,s_{0}\,,s\in(0,\pi).

It is clear from this last equality and the considerations preceding it that s0<π/2<s1<π,s_{0}<\pi/2<s_{1}<\pi, which implies that τ\uptau coincides with τλ\uptau_{\lambda} , λ=s0\lambda=s_{0} . Therefore, Σ\varSigma^{\prime} coincides with the “block” Σ(λ)\varSigma^{\prime}(\lambda) that generates Σ(λ)\varSigma(\lambda) (see Fig. 6), so that Σ0\varSigma_{0} is a union of open sets of members of 𝒮.\mathscr{S}.

Let UΣΣ0U\subset\varSigma-\varSigma_{0} be an open set of Σ.\varSigma. If UU is nonempty, it cannot be horizontal, for no member of 𝒮\mathscr{S} has horizontal points. If UU is vertical, then it is part of the totally geodesic cylinder Sπ/2×.S_{\pi/2}\times\mathbb{R}. In this case, a boundary point of UU is vertical and lies on a geodesic sphere centered at the axis and of radius π/2.\pi/2. However, such a boundary point also lies on some Σ(λ)𝒮,\varSigma(\lambda)\in\mathscr{S}, which contradicts the fact that any vertical point of Σ(λ)\varSigma(\lambda) lies on a geodesic sphere of radius different from π/2.\pi/2. (In fact, these vertical points are on geodesic spheres of radiuses λ<π/2\lambda<\pi/2 and πλ>π/2.\pi-\lambda>\pi/2.)

We conclude from the above that Σ0\varSigma_{0} is open and dense in Σ.\varSigma. Since Σ\varSigma is connected and two distinct members of 𝒮\mathscr{S} are never tangent, it follows that, for some λ(0,π/2),\lambda\in(0,\pi/2), Σ\varSigma coincides with Σ(λ)𝒮.\varSigma(\lambda)\in\mathscr{S}.

Remark 5.

The case 𝔽=\mathbb{F}=\mathbb{R} of Theorem 7 was considered in [19], whereas Theorem 8 was proved in [31] for r=1.r=1. Again, the methods employed in these works is different from ours and cannot be applied to general products M×M\times\mathbb{R}, since they all rely on the Euclidean and Lorentzian geometries of the underlying spaces of 𝕊n×\mathbb{S}^{n}\times\mathbb{R} and n×.\mathbb{H}^{n}\times\mathbb{R}.

6. Translational Hr(>0)H_{r}(>0)-hypersurfaces of 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}.

Given a Hadamard manifold M,M, recall that the Busemann function 𝔟γ\mathfrak{b}_{\gamma} of MM corresponding to an arclength geodesic γ:(,+)M\gamma\colon(-\infty,+\infty)\rightarrow M is defined as

𝔟γ(p):=lims+(distM(p,γ(s))s),pM.\mathfrak{b}_{\gamma}(p):=\lim_{s\rightarrow+\infty}({\rm dist}_{M}(p,\gamma(s))-s),\,\,\,p\in M.

The level sets s:=𝔟γ1(s)\mathscr{H}_{s}:=\mathfrak{b}_{\gamma}^{-1}(s) of a Busemann function 𝔟γ\mathfrak{b}_{\gamma} are called horospheres of M.M. In this setting, as is well known, {s;s(,+)}\{\mathscr{H}_{s}\,;\,s\in(-\infty,+\infty)\} is a parallel family which foliates M.M. Furthermore, any horosphere s\mathscr{H}_{s} is homeomorphic to n1\mathbb{R}^{n-1}, and any geodesic of MM which is asymptotic to γ\gamma — i.e., with the same point pp_{\infty} on the asymptotic boundary M()M(\infty) of MM — is orthogonal to each horosphere s\mathscr{H}_{s} . In this case, we say that the horospheres s\mathscr{H}_{s} are centered at pp_{\infty} .

Refer to caption
Figure 9. A “ball model” for a Hadamard manifold.

Therefore, in what concerns its horospheres, a Hadamard manifold can be pictured just as the Poincaré ball model of hyperbolic space n,\mathbb{H}^{n}, where the horospheres centered at a point pn()p_{\infty}\in\mathbb{H}^{n}(\infty) are the Euclidean (n1)(n-1)-spheres in n\mathbb{H}^{n} which are tangent to n()\mathbb{H}^{n}(\infty) at pp_{\infty} (Fig. 9).

In the real hyperbolic space n,\mathbb{H}^{n}, any horosphere is totally umbilical with constant principal curvatures equal to 1.1. As shown in [5, Proposition-(vi), pg. 88], any horosphere of 𝔽m,\mathbb{H}_{\mathbb{F}}^{m}, 𝔽,\mathbb{F}\neq\mathbb{R}, has principal curvatures 11 and 1/21/2 with multiplicities 11 and n2,n-2, respectively. Therefore, any family \mathscr{F} of parallel horospheres of  𝔽m\mathbb{H}_{\mathbb{F}}^{m} is isoparametric and its members are pairwise congruent. In addition, for any integer r{1,,n1},r\in\{1,\dots,n-1\}, all horospheres of  𝔽m\mathbb{H}_{\mathbb{F}}^{m} have the same (positive) rr-th mean curvature, which we denote by Hr0.H_{r}^{0}.

Theorem 9.

Let :={s;s(,+)}\mathscr{F}:=\{\mathscr{H}_{s}\,;\,s\in(-\infty,+\infty)\} be a family of parallel horospheres in hyperbolic space 𝔽m.\mathbb{H}_{\mathbb{F}}^{m}. Then, for any even integer  r{2,,n1},r\in\{2,\dots,n-1\}, and any constant Hr(0,Hr0),H_{r}\in(0,H_{r}^{0}), there exists a properly embedded, everywhere non convex HrH_{r}-hypersurface Σ\varSigma in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which is homeomorphic to  n.\mathbb{R}^{n}. Furthermore, Σ\varSigma is foliated by horospheres, is symmetric with respect to the horizontal hyperplane 𝔽m×{0}\mathbb{H}_{\mathbb{F}}^{m}\times\{0\}, and its height function is unbounded above and below.

Proof.

For each s(,)s\in(-\infty,\infty), consider the isometric immersion fs:n1𝔽mf_{s}:\mathbb{R}^{n-1}\rightarrow\mathbb{H}_{\mathbb{F}}^{m} such that fs(n1)=sf_{s}(\mathbb{R}^{n-1})=\mathscr{H}_{s} . Since all the principal curvatures of fsf_{s} are constant and independent of s,s, the coefficients aa e bb of the ODE y=ay+by^{\prime}=ay+b associated to this family are constants. Also, since rr is even and 0<Hr<Hr00<H_{r}<H_{r}^{0} , we have

b<0<aand0<ba<1.b<0<a\quad\text{and}\quad 0<-\frac{b}{a}<1.

In this setting, consider the solution τ:(,0]\uptau:(-\infty,0]\rightarrow\mathbb{R} of y=ay+by^{\prime}=ay+b:

(47) τ(s)=ea(ss0)ba,s0=log(1+b/a)1/a,\uptau(s)=e^{a(s-s_{0})}-\frac{b}{a}\,,\,\,\,s_{0}=\log(1+b/a)^{-1/a},

and observe that it satisfies:

(48) 0<ba<τ(s)1=τ(0)s(,0].0<-\frac{b}{a}<\uptau(s)\leq 1=\uptau(0)\,\,\,\forall s\in(-\infty,0].
Refer to caption
Figure 10. A piece of the graph Σ\varSigma^{\prime}, on which all the trajectories of ξ\nabla\xi meet 0×{0}\mathscr{H}_{0}\times\{0\} orthogonally.

By Lemma 1, the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma^{\prime} with ϱ=τr\varrho=\sqrt[r]{\uptau} is an HrH_{r}-hypersurface of 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}. The function ϕ,\phi, in this case, is given by

ϕ(s):=s0ϱ(u)1ϱ2(u)𝑑u,s(,0).\phi(s):=-\int_{s}^{0}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du,\,\,\,s\in(-\infty,0).

Notice that, by (48), one has

τ(s)=aτ(s)+b>aba+b=0,\uptau^{\prime}(s)=a\uptau(s)+b>a\frac{-b}{\phantom{-}a}+b=0,

so that τ(0)>0.\uptau^{\prime}(0)>0. Thus, by Lemma 2, ϕ\phi is well defined. Also, ϕ\phi is negative on (,0)(-\infty,0) and is unbounded. Indeed, for all s(,0),s\in(-\infty,0),

ϕ(s)=s0ϱ(u)1ϱ2(u)𝑑us0ϱ(u)𝑑uinfϱ|[s,0]s=sϱ(s),-\phi(s)=\int_{s}^{0}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du\geq\int_{s}^{0}\varrho(u)du\geq-\inf\varrho|_{[s,0]}s=-s\varrho(s)\,,

which implies that ϕ\phi is unbounded, since ϱ(s)(b/a)1/r>0\varrho(s)\rightarrow(-b/a)^{1/r}>0 as s.s\rightarrow-\infty.

Denoting by B0B_{0} the horoball of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} with boundary 0\mathscr{H}_{0} , it follows from the above considerations that Σ\varSigma^{\prime} is an HrH_{r}-graph over 𝔽mB0\mathbb{H}_{\mathbb{F}}^{m}-B_{0} which is unbounded and has boundary Σ=0×{0}\partial\varSigma^{\prime}=\mathscr{H}_{0}\times\{0\} (Fig. 10). In particular, Σ\varSigma^{\prime} is homeomorphic to n\mathbb{R}^{n}. Furthermore, Σ\varSigma^{\prime} is everywhere non convex, since, from the identities (11), one has

ki=λiϱ<0(1in1)andkn=ϱ>0,k_{i}=-\lambda_{i}\varrho<0\,\,\,(1\leq i\leq n-1)\,\,\,\,\text{and}\,\,\,k_{n}=\varrho^{\prime}>0,

where λi\lambda_{i} is the (positive constant) ii-th principal curvature of fsf_{s} .

Finally, since ϱ(0)=1,\varrho(0)=1, as in the previous theorems, we have that any trajectory of ξ\nabla\xi on Σ\varSigma^{\prime} meets Σ\partial\varSigma^{\prime} orthogonally. Consequently, setting Σ′′\varSigma^{\prime\prime} for the reflection of Σ\varSigma^{\prime} with respect to 𝔽m×{0}\mathbb{H}_{\mathbb{F}}^{m}\times\{0\} and defining

Σ:=closure(Σ)closure(Σ′′),\varSigma:={\rm closure}\,(\varSigma^{\prime})\cup{\rm closure}\,(\varSigma^{\prime\prime}),

we have that Σ\varSigma is a properly embedded HrH_{r}-hypersurface of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which is foliated by horospheres and is homeomorphic to n\mathbb{R}^{n} (Fig. 11), as we wished to prove. ∎

Refer to caption
Figure 11. A piece of a properly embedded everywhere non convex Hr(>0)H_{r}(>0)-hypersurface of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which is foliated by horospheres.

Our next result establishes that the conditions on the parity of rr and on the sign of HrHr0H_{r}-H_{r}^{0} in Theorem 9 are necessary to the conclusion.

Theorem 10.

Let \mathscr{F} be a family of parallel horospheres in 𝔽m.\mathbb{H}_{\mathbb{F}}^{m}. Assume that, for some r{1,,n},r\in\{1,\dots,n\}, Σ\varSigma is a complete connected Hr(>0)H_{r}(>0)-hypersurface of  𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} with no horizontal points, and that each connected component of any horizontal section ΣtΣ\varSigma_{t}\subset\varSigma is a (vertically translated) horosphere of .\mathscr{F}. Under these conditions, one has r<n.r<n. Assume, in addition, that either of the following assertions holds:

  • i)

    rr is even and HrHr0.H_{r}\geq H_{r}^{0}.

  • ii)

    rr is odd.

Then, Σ=s×\varSigma=\mathscr{H}_{s}\times\mathbb{R} for some s.\mathscr{H}_{s}\in\mathscr{F}. In particular, Hr=Hr0.H_{r}=H_{r}^{0}.

Proof.

Let Σ0Σ\varSigma_{0}\subset\varSigma be the open set of points xΣx\in\varSigma satisfying Θ(x)0.\varTheta(x)\neq 0. Our aim is to prove that Σ0\varSigma_{0} is empty. Assuming otherwise, choose x0Σ0x_{0}\in\varSigma_{0} . Since Σ\varSigma has no horizontal points, we can suppose (after possibly a reflection about a horizontal hyperplane) that there is an open neighborhood ΣΣ0\varSigma^{\prime}\subset\varSigma_{0} of x0x_{0} which is an (fs,ϕ)(f_{s},\phi)-graph, fs.f_{s}\in\mathscr{F}.

The τ\uptau-function associated to Σ\varSigma^{\prime} satisfies τ=aτ+b,\uptau^{\prime}=a\uptau+b, where aa and b0b\neq 0 are the (constant) functions (15) determined by \mathscr{F} and HrH_{r} . Also, there is a maximal interval I=(s1,s2),I=(s_{1},s_{2})\subset\mathbb{R}, s1<s2+-\infty\leq s_{1}<s_{2}\leq+\infty, such that τ(I)(0,1).\uptau(I)\subset(0,1).

Let us suppose that r=n.r=n. In this case, we have a=0,a=0, which gives τ(s)=b0,\uptau^{\prime}(s)=b\neq 0, that is, τ(s)=bs+c,\uptau(s)=bs+c, c.c\in\mathbb{R}. In particular, s1>s_{1}>-\infty and s2<+s_{2}<+\infty, and τ\uptau is increasing (if b>0b>0), or decreasing (if b<0b<0) in (s1,s2).(s_{1},s_{2}). So, τ\uptau vanishes in s1s_{1} or s2s_{2} . Assuming the former, we have that ϕ\phi is defined at s1s_{1} and ϕ(s1)=0.\phi^{\prime}(s_{1})=0. Thus, for any pn1,p\in\mathbb{R}^{n-1}, the point x=(fs1(p),ϕ(s1))Σx=(f_{s_{1}}(p),\phi(s_{1}))\in\varSigma^{\prime} is horizontal, contrary to our assumption. Therefore, if r=n,r=n, then Σ0=,\varSigma_{0}=\emptyset, which implies that Σ=s×\varSigma=\mathscr{H}_{s}\times\mathbb{R} for some s.s\in\mathbb{R}. But this contradicts the assumed positiveness of HnH_{n} . Hence, we must have r<n.r<n.

Let us assume now that (i) holds. Then, we have b<0<a.b<0<a. Also, on (s1,s2),(s_{1},s_{2}),

τ=aτ+b<a+b=r(Hr0Hr)Hr100andτ′′=aτ+b<0,\uptau^{\prime}=a\uptau+b<a+b=\frac{r(H_{r}^{0}-H_{r})}{H_{r-1}^{0}}\leq 0\quad\text{and}\quad\uptau^{\prime\prime}=a\uptau^{\prime}+b<0,

that is, τ\uptau is decreasing and concave in (s1,s2),(s_{1},s_{2}), which clearly implies that s2<+s_{2}<+\infty and τ(s2)=0.\uptau(s_{2})=0. As in the preceding paragraph, this leads to the existence of a horizontal point of Σ.\varSigma. Therefore, Σ0=\varSigma_{0}=\emptyset if (i) holds, which implies that Σ=s×\varSigma=\mathscr{H}_{s}\times\mathbb{R} for some s.s\in\mathbb{R}.

Finally, let us assume that (ii) holds. In this case, one has a,b>0,a,b>0, which gives that τ\uptau is increasing and convex. From this point, we get easily to the conclusion by reasoning just as in the last paragraph. ∎

An isometry φ\varphi of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} which fixes only one point p𝔽m()p_{\infty}\in\mathbb{H}_{\mathbb{F}}^{m}(\infty) is called parabolic. Such isometries have the following fundamental property: The horospheres of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} centered at p𝔽m()p_{\infty}\in\mathbb{H}_{\mathbb{F}}^{m}(\infty) are invariant by parabolic isometries of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} that fix pp_{\infty} (cf. [17, Proposition 7.8]).

We point out that any isometry φ\varphi of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} has a natural extension to an isometry Φ\Phi of 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}. Namely,

Φ(p,t)=(φ(p),t),(p,t)𝔽m×.\Phi(p,t)=(\varphi(p),t),\,\,\,(p,t)\in\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}.

We call Φ\Phi parabolic if φ\varphi is parabolic. More specifically, if

={fs:n1𝔽m;s(,+)},fs(n1)=s,\mathscr{F}=\{f_{s}:\mathbb{R}^{n-1}\rightarrow\mathbb{H}_{\mathbb{F}}^{m}\,;\,s\in(-\infty,+\infty)\},\,\,\,f_{s}(\mathbb{R}^{n-1})=\mathscr{H}_{s}\,,

is the family of parallel horospheres which are invariant by φ,\varphi, we say that φ\varphi and Φ\Phi are \mathscr{F}-parabolic isometries.

In the upper half-space model of n\mathbb{H}_{\mathbb{R}}^{n}, Euclidean horizontal translations in a fixed direction are parabolic. As for the other hyperbolic spaces, the parabolic isometries are more involved (see, e.g., [24]). Nevertheless, inspired by the real case, we say that parabolic isometries are translational.

Finally, let us remark that, given a family \mathscr{F} of parallel horospheres in 𝔽m,\mathbb{H}_{\mathbb{F}}^{m}, if a hypersurface Σ\varSigma of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} is invariant by \mathscr{F}-parabolic isometries of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, then any connected component of any horizontal section ΣtΣ\varSigma_{t}\subset\varSigma is contained in a (vertically translated) horosphere of .\mathscr{F}.

Now, we are in position to classify all complete connected Hr(>0)H_{r}(>0)-hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} with no horizontal points which are invariant by parabolic isometries.

Theorem 11.

Let \mathscr{F} be a family of parallel horospheres of 𝔽m.\mathbb{H}_{\mathbb{F}}^{m}. Assume that Σ\varSigma is a complete connected Hr(>0)H_{r}(>0)-hypersurface of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}, r{1,,n},r\in\{1,\dots,n\}, with no horizontal points, which is invariant by \mathscr{F}-parabolic isometries. Then, up to ambient isometries, Σ\varSigma is either a cylinder over a horosphere of  𝔽m\mathbb{H}_{\mathbb{F}}^{m} or the embedded HrH_{r}-hypersurface obtained in Theorem 9.

Proof.

Assume that Σ\varSigma is not a cylinder over a horosphere of 𝔽m.\mathbb{H}_{\mathbb{F}}^{m}. By Theorem 10, we have that r(<n)r(<n) is even and 0<Hr<Hr0.0<H_{r}<H_{r}^{0}. In this case, the open set

Σ0:={xΣ;Θ(x)0}\varSigma_{0}:=\{x\in\varSigma\,;\,\varTheta(x)\neq 0\}

is dense in Σ.\varSigma. Indeed, the r(<n)r(<n)-th mean curvature of any nonempty open \mathscr{F}-invariant vertical set UΣU\subset\varSigma would be Hr0>Hr.H_{r}^{0}>H_{r}. Thus, given x0Σ0,x_{0}\in\varSigma_{0}\,, there exists an (fs,ϕ)(f_{s},\phi)-graph Σx0\varSigma^{\prime}\owns x_{0} in Σ0\varSigma_{0} with fs.f_{s}\in\mathscr{F}. In addition, the associated τ\uptau-function, restricted to a maximal interval (s0,s1),(s_{0},s_{1}), s0<s1+,-\infty\leq s_{0}<s_{1}\leq+\infty, satisfies 0<τ<1.0<\uptau<1.

We have that τ\uptau is a solution of the ODE y=ay+by^{\prime}=ay+b determined by \mathscr{F} and Hr.H_{r}\,. The conditions on the parity of rr and the sign of HrHr0H_{r}-H_{r}^{0}, as in the proof of Theorem 9, give that τ\uptau is increasing and convex, which implies that s1<+s_{1}<+\infty and that τ(s1)=1.\uptau(s_{1})=1.

Consider the solution (47) of y=ay+by^{\prime}=ay+b and denote it by τ~.\tilde{\uptau}. Since τ~(0)=τ(s1)=1,\tilde{\uptau}(0)=\uptau(s_{1})=1, and the coefficients aa and bb are constants, by the uniqueness of solutions satisfying initial conditions, we have that τ(s)=τ~(ss1).\uptau(s)=\tilde{\uptau}(s-s_{1}). This, together with the homogeneity of the horospheres of 𝔽m,\mathbb{H}_{\mathbb{F}}^{m}, implies that Σ\varSigma^{\prime} coincides with the (fs,ϕ)(f_{s},\phi)-graph determined by τ~.\tilde{\uptau}. From this fact and the density of Σ0\varSigma_{0} in Σ,\varSigma, we conclude that Σ\varSigma coincides with the embedded HrH_{r}-hypersurface obtained in Theorem 9, as we wished to prove. ∎

Given a totally geodesic hyperplane 0\mathscr{E}_{0} of n,\mathbb{H}^{n}, let us recall that there exists a family :={s;s(,+)}\mathscr{F}:=\{\mathscr{E}_{s}\,;\,s\in(-\infty,+\infty)\} of parallel hypersurfaces of n\mathbb{H}^{n} such that the distance of any point of s\mathscr{E}_{s} to 0\mathscr{E}_{0} is |s|.|s|. The family \mathscr{F} foliates n,\mathbb{H}^{n}, and each member s\mathscr{E}_{s} of \mathscr{F}, which is called an equidistant hypersurface, is properly embedded and homeomorphic to n1\mathbb{R}^{n-1} (Fig. 12).

Refer to caption
Figure 12. Equidistant hypersurfaces in the Poincaré ball model of n.\mathbb{H}^{n}.

We shall also write \mathscr{F} as a family of immersions:

={fs:n1n;s(,+)},\mathscr{F}=\{f_{s}:\mathbb{R}^{n-1}\rightarrow\mathbb{H}^{n}\,;\,s\in(-\infty,+\infty)\},

that is, for each s(,+),s\in(-\infty,+\infty), fs(n1)f_{s}(\mathbb{R}^{n-1}) is the equidistant s\mathscr{E}_{s} to 0=f0(n1)\mathscr{E}_{0}=f_{0}(\mathbb{R}^{n-1}) .

Given a geodesic γp\gamma_{p} orthogonal to the members of ,\mathscr{F}, p0,p\in\mathscr{E}_{0}, any equidistant hypersurface s\mathscr{E}_{s} is totally umbilical with constant principal curvatures all equal to

ks=tanh(s)k^{s}=-\tanh(s)

with respect to the unit normal ηs=γp\eta_{s}=\gamma_{p}^{\prime} (see Section 3). In particular, \mathscr{F} is isoparametric. Also, given a constant HrH_{r} , the coefficients aa and bb of the differential equation y=ay+by^{\prime}=ay+b associated to \mathscr{F} and HrH_{r} are:

(49) a(s)=(nr)tanh(s)andb(s)=brtanh1r(s),br=rHr(n1r1)1.a(s)=-(n-r)\tanh(s)\quad\text{and}\quad b(s)=b_{r}\tanh^{1-r}(s),\,\,\,b_{r}=rH_{r}{{n-1}\choose{r-1}}^{-1}.

It will be convenient to reconsider the constant Cr:=nrn(nr)C_{r}:=\frac{n-r}{n}{{n}\choose{r}} and recall that, for 1r<n,1\leq r<n, the following identity holds:

(50) brnr=HrCr\frac{b_{r}}{n-r}=\frac{H_{r}}{C_{r}}\,\cdot

Our next result establishes that, for Hr(0,Cr)H_{r}\in(0,C_{r}), 1r<n,1\leq r<n, there exists a one-parameter family of properly embedded HrH_{r}-hypersurfaces in n×\mathbb{H}^{n}\times\mathbb{R} which are foliated by (vertical translations of) parallel equidistant hypersurfaces of n.\mathbb{H}^{n}. In this setting, we have 0<Hr/Cr<1,0<H_{r}/C_{r}<1, so we can define:

(51) sr:=arctanh(Hr/Cr)1/r.s_{r}:={\rm arctanh}\,(H_{r}/C_{r})^{1/r}.
Theorem 12.

Given r{1,,n1}r\in\{1,\dots,n-1\}, let Hr(0,Cr).H_{r}\in(0,C_{r}). Then, there exists a one parameter family 𝒮:={Σ(λ);λ(sr,+)}\mathscr{S}:=\{\varSigma(\lambda)\,;\,\lambda\in(s_{r},+\infty)\} of properly embedded and everywhere non convex HrH_{r}-hypersurfaces of  n×.\mathbb{H}^{n}\times\mathbb{R}. Each member Σ(λ)\varSigma(\lambda) of 𝒮\mathscr{S} is homeomorphic to n\mathbb{R}^{n} and is foliated by equidistant hypersurfaces. Moreover, Σ(λ)\varSigma(\lambda) is symmetric with respect to n×{0}\mathbb{H}^{n}\times\{0\}, and its height function is unbounded.

Proof.

Let :={fs;s(0,+)}\mathscr{F}:=\{f_{s}\,;\,s\in(0,+\infty)\} be a family of parallel equidistant hypersurfaces of n.\mathbb{H}^{n}. Since 0<Hr<Cr0<H_{r}<C_{r} , it follows from the relation (50) that 0<br/(nr)<1.0<b_{r}/(n-r)<1. Thus, we can choose λ>0\lambda>0 such that

(52) tanhr(λ)>brnr\tanh^{r}(\lambda)>\frac{b_{r}}{n-r}\,\cdot

In particular, λ(sr,+).\lambda\in(s_{r},+\infty).

Let τλ\uptau_{\lambda} be the solution of y=ay+by^{\prime}=ay+b satisfying y(λ)=1,y(\lambda)=1, where aa and bb are the functions in (49). Then, from (52), we have

τλ(λ)=(nr)tanh(λ)+brtanh1r(λ)<0,\uptau_{\lambda}^{\prime}(\lambda)=-(n-r)\tanh(\lambda)+b_{r}\tanh^{1-r}(\lambda)<0,

so that τλ\uptau_{\lambda} is decreasing near λ\lambda .

We claim that τλ\uptau_{\lambda} is decreasing on the whole interval [λ,+)[\lambda,+\infty). To show that, it suffices to prove that τλ\uptau_{\lambda} has no critical points in (λ,+).(\lambda,+\infty). Assuming otherwise, consider s1>λs_{1}>\lambda satisfying τλ(s1)=0.\uptau_{\lambda}^{\prime}(s_{1})=0. Since τλ(s1)=a(s1)τλ(s1)+b(s1),\uptau_{\lambda}^{\prime}(s_{1})=a(s_{1})\uptau_{\lambda}(s_{1})+b(s_{1}), we have that τλ(s1)=b(s1)/a(s1)>0.\uptau_{\lambda}(s_{1})=-b(s_{1})/a(s_{1})>0. We also have a<0a^{\prime}<0 and b0.b^{\prime}\leq 0. Thus,

τλ′′(s1)=a(s1)τλ(s1)+b(s1)<0,\uptau_{\lambda}^{\prime\prime}(s_{1})=a^{\prime}(s_{1})\uptau_{\lambda}(s_{1})+b^{\prime}(s_{1})<0,

which implies that s1s_{1} is necessarily a local maximum for τλ.\uptau_{\lambda}. This proves the claim, for τλ\uptau_{\lambda} is decreasing near λ,\lambda, so that a local maximum s1>λs_{1}>\lambda for τλ\uptau_{\lambda} should be preceded by a local minimum.

We also have that τλ\uptau_{\lambda} is positive in [λ,+).[\lambda,+\infty). Indeed, if we had τλ(s)0\uptau_{\lambda}(s)\leq 0 for some s>λ,s>\lambda, it would give τλ(s)=a(s)τ(s)+b(s)>0,\uptau_{\lambda}^{\prime}(s)=a(s)\uptau(s)+b(s)>0, and then τλ\uptau_{\lambda} would be increasing near s.s.

It follows from the above considerations that

0<τλ(s)1=τλ(λ)s[λ,+).0<\uptau_{\lambda}(s)\leq 1=\uptau_{\lambda}(\lambda)\,\,\,\forall s\in[\lambda,+\infty).

Furthermore, since τλ\uptau_{\lambda} is decreasing and positive, we have that τλ(s)0\uptau_{\lambda}^{\prime}(s)\rightarrow 0 as s+.s\rightarrow+\infty. This, together with the equalities τλ=aτλ+b\uptau_{\lambda}^{\prime}=a\uptau_{\lambda}+b and ϱλr=τλ\varrho_{\lambda}^{r}=\uptau_{\lambda}, gives

(53) lims+ϱλ(s)=(Hr/Cr)1/r>0.\lim_{s\rightarrow+\infty}\varrho_{\lambda}(s)=\left({H_{r}}/{C_{r}}\right)^{1/r}>0.

Therefore, the (fs,ϕ)(f_{s},\phi) graph Σ(λ)\varSigma^{\prime}(\lambda) associated to ϱλ\varrho_{\lambda} (see Fig. 13) is an HrH_{r}-hypersurface of n×\mathbb{H}^{n}\times\mathbb{R} whose ϕ\phi-function is

ϕλ(s)=λsϱλ(u)1ϱλ2(u)𝑑u,s[λ,+).\phi_{\lambda}(s)=\int_{\lambda}^{s}\frac{\varrho_{\lambda}(u)}{\sqrt{1-\varrho_{\lambda}^{2}(u)}}du,\,\,\,s\in[\lambda,+\infty).
Refer to caption
Figure 13. A piece of the graph Σ\varSigma^{\prime}, on which all trajectories of ξ\nabla\xi emanate from λ×{0}\mathscr{E}_{\lambda}\times\{0\} orthogonally.

As in the preceding proofs, we obtain a properly embedded HrH_{r}-hypersurface Σ(λ)\varSigma(\lambda) of n×\mathbb{H}^{n}\times\mathbb{R} by reflecting Σ(λ)\varSigma^{\prime}(\lambda) with respect to n×{0},\mathbb{H}^{n}\times\{0\}, since ϱλ(λ)=1\varrho_{\lambda}(\lambda)=1 and ϱλ(λ)<0\varrho_{\lambda}^{\prime}(\lambda)<0. It is also clear from equalities (11) that, except for kn=ϱk_{n}=\varrho^{\prime}, its principal curvatures kik_{i} are all positive, so that Σ(λ)\varSigma(\lambda) is nowhere convex.

Finally, considering (53) and the fact that ϱλ\varrho_{\lambda} is decreasing, we have

ϕλ(s)=λsϱλ(u)1ϱλ2(u)𝑑uλsϱλ(u)𝑑u(Hr/Cr)1/r(sλ),\phi_{\lambda}(s)=\int_{\lambda}^{s}\frac{\varrho_{\lambda}(u)}{\sqrt{1-\varrho_{\lambda}^{2}(u)}}du\geq\int_{\lambda}^{s}\varrho_{\lambda}(u)du\geq\left({H_{r}}/{C_{r}}\right)^{1/r}(s-\lambda),

which clearly implies that the height function of Σ(λ)\varSigma(\lambda) is unbounded. ∎

Theorem 13.

Let :={fs;s(0,+)}\mathscr{F}:=\{f_{s}\,;\,s\in(0,+\infty)\} be a family of equidistant hypersurfaces in n.\mathbb{H}^{n}. Given r{1,,n1}r\in\{1,\dots,n-1\} and Hr(0,Cr),H_{r}\in(0,C_{r}), let sr>0s_{r}>0 be the constant defined in (51). Then, there exists a complete everywhere non convex HrH_{r}-hypersurface Σ\varSigma in n×\mathbb{H}^{n}\times\mathbb{R} which is an (fs,ϕ)(f_{s},\phi)-graph, s(sr,+).s\in(s_{r},+\infty). Furthermore, the height function of Σ\varSigma is unbounded above and below, and Σ\varSigma is asymptotic to sr×(,0).\mathscr{E}_{s_{r}}\times(-\infty,0).

Proof.

Let τ\uptau be the solution of the differential equation y=ay+by^{\prime}=ay+b associated to HrH_{r} and \mathscr{F} (i.e., with aa and bb as in (49)) which satisfies the initial condition τ(sr)=1.\uptau(s_{r})=1.

From its definition, we have that srs_{r} satisfies τ(sr)=0.\uptau^{\prime}(s_{r})=0. In addition,

τ′′(sr)=a(sr)+b(sr)<0,\uptau^{\prime\prime}(s_{r})=a^{\prime}(s_{r})+b^{\prime}(s_{r})<0,

so that srs_{r} is a local maximum of τ.\uptau. Reasoning as in the preceding proof, we get that τ\uptau, and so ϱ=τ1/r\varrho=\uptau^{1/r}, is positive and decreasing in (sr,+).(s_{r},+\infty). From this, we conclude analogously that ϱ(s)(Hr/Cr)1/r\varrho(s)\rightarrow(H_{r}/C_{r})^{1/r} as s+.s\rightarrow+\infty.

Now, for a fixed s0>srs_{0}>s_{r} , define

ϕ(s):=s0sϱ(u)1ϱ2(u)𝑑u,s(sr,+),\phi(s):=\int_{s_{0}}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du,\,\,\,s\in(s_{r},+\infty),

and let Σ\varSigma be the corresponding (fs,ϕ)(f_{s},\phi)-graph. As before, we have that Σ\varSigma is nowhere convex. Denoting by Ωsr\Omega_{s_{r}} the convex connected component of nsr,\mathbb{H}^{n}-\mathscr{E}_{s_{r}}, we also have that Σ\varSigma is a graph over nΩsr\mathbb{H}^{n}-\Omega_{s_{r}} (Fig. 14).

For s>s0s>s_{0} , we have

ϕ(s)=s0sϱ(u)1ϱ2(u)𝑑us0sϱ(u)𝑑u(Hr/Cr)1/r(ss0),\phi(s)=\int_{s_{0}}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du\geq\int_{s_{0}}^{s}\varrho(u)du\geq(H_{r}/C_{r})^{1/r}(s-s_{0}),

which gives that ϕ\phi, and so the height function of Σ,\varSigma, is unbounded above.

Finally, given a constant C>0,C>0, there exists s¯(sr,s0)\bar{s}\in(s_{r},s_{0}) such that

1ϱ(s)<Cs(sr,s¯),\frac{1}{\varrho^{\prime}(s)}<-C\,\,\,\forall s\in(s_{r},\bar{s}),

for ϱ(sr)=0.\varrho^{\prime}(s_{r})=0. Thus, for such values of s,s, one has

ϕ(s)\displaystyle\phi(s) =\displaystyle= s0sϱ(u)ϱ(u)ϱ(u)1ϱ2(u)𝑑uCϱ(s¯)s0sϱ(u)1ϱ2(u)𝑑u\displaystyle\int_{s_{0}}^{s}\frac{\varrho^{\prime}(u)\varrho(u)}{\varrho^{\prime}(u)\sqrt{1-\varrho^{2}(u)}}du\leq-C\varrho(\bar{s})\int_{s_{0}}^{s}\frac{\varrho^{\prime}(u)}{\sqrt{1-\varrho^{2}(u)}}du
=\displaystyle= Cϱ(s¯)ϱ(s0)ϱ(s)dϱ1ϱ2=Cϱ(s¯)(arcsinϱ(s)arcsinϱ(s0))\displaystyle-C\varrho(\bar{s})\int_{\varrho(s_{0})}^{\varrho(s)}\frac{d\varrho}{\sqrt{1-\varrho^{2}}}=-C\varrho(\bar{s})(\arcsin\varrho(s)-\arcsin\varrho(s_{0}))
\displaystyle\leq Cϱ(s¯)(arcsinϱ(s¯)arcsinϱ(s0)),\displaystyle-C\varrho(\bar{s})(\arcsin\varrho(\bar{s})-\arcsin\varrho(s_{0})),

which implies that ϕ(s)\phi(s)\rightarrow-\infty as ssrs\rightarrow s_{r} , since ϱ(s¯)\varrho(\bar{s}) is bounded away from zero. Therefore, the height function of Σ\varSigma is unbounded below, and Σ\varSigma is asymptotic to sr×(,0)\mathscr{E}_{s_{r}}\times(-\infty,0) in n×(,0),\mathbb{H}^{n}\times(-\infty,0), as we wished to prove. ∎

Refer to caption
Figure 14. A piece of the (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma which is above n×{0}.\mathbb{H}^{n}\times\{0\}. As ss\rightarrow-\infty, the trajectories of ξ\nabla\xi converge asymptotically to sr×(0,).\mathscr{E}_{s_{r}}\times(0,-\infty).

A family ={s;s(,+)\mathscr{F}=\{\mathscr{E}_{s}\,;\,s\in(-\infty,+\infty) of equidistant hypersurfaces in n\mathbb{H}^{n} determines a group of translational isometries which we shall call \mathscr{F}-hyperbolic. In the upper half-space model of n,\mathbb{H}^{n}, taking 0\mathscr{E}_{0} as a Euclidean half vertical hyperplane orthogonal to n\partial_{\infty}\mathbb{H}^{n} through the “origin” o,o, we have that the \mathscr{F}-hyperbolic isometries are the Euclidean homotheties from o.o. It should be noticed that the equidistant hypersurfaces of \mathscr{F} are all invariant by \mathscr{F}-hyperbolic isometries.

The natural extension of an \mathscr{F}-hyperbolic isometry of n\mathbb{H}^{n} to n×\mathbb{H}^{n}\times\mathbb{R} will also be called \mathscr{F}-hyperbolic. If Σ\varSigma is a hypersurface of n×\mathbb{H}^{n}\times\mathbb{R} which is invariant by \mathscr{F}-hyperbolic isometries, it is clear that any connected component of any horizontal section Σt\varSigma_{t} of Σ\varSigma is contained in s×{t}\mathscr{E}_{s}\times\{t\} for some s(,+).s\in(-\infty,+\infty).

Next, we classify Hr(>0)H_{r}(>0)-hypersurfaces of n×\mathbb{H}^{n}\times\mathbb{R} (without horizontal points or totally geodesic horizontal sections) which are invariant by hyperbolic translations.

Theorem 14.

Let ={fs;s(,+)}\mathscr{F}=\{f_{s}\,;\,s\in(-\infty,+\infty)\} be a family of parallel equidistant hypersurfaces in n.\mathbb{H}^{n}. Assume that, for some r{1,,n},r\in\{1,\dots,n\}, Σ\varSigma is a complete connected Hr(>0)H_{r}(>0)-hypersurface of  n×\mathbb{H}^{n}\times\mathbb{R} which is invariant by \mathscr{F}-hyperbolic translations. Assume further that Σ\varSigma has no horizontal points, and that no horizontal section Σt\varSigma_{t} of Σ\varSigma is totally geodesic in n×{t}\mathbb{H}^{n}\times\{t\} (i.e., Σt0=f0(n)).\varSigma_{t}\not\subset\mathscr{E}_{0}=f_{0}(\mathbb{R}^{n})). Under these conditions, the following assertions hold:

  • i)

    r<nr<n.

  • ii)

    0<Hr<Cr0<H_{r}<C_{r} .

  • iii)

    Σ\varSigma is either the cylinder over the equidistant  sr\mathscr{E}_{s_{r}} or, up to an ambient isometry, one of the embedded hypersurfaces obtained in Theorems 1213.

Proof.

Set Σ0:={xΣ;Θ(x)0}\varSigma_{0}:=\{x\in\varSigma\,;\,\varTheta(x)\neq 0\} and assume Σ0.\varSigma_{0}\neq\emptyset. Given x0Σ0x_{0}\in\varSigma_{0} , as in previous proofs, we can assume there is an open set ΣΣ0\varSigma^{\prime}\subset\varSigma_{0} which is an (fs,ϕ)(f_{s},\phi)-graph containing x0x_{0} . Its τ\uptau function satisfies τ=aτ+b\uptau^{\prime}=a\uptau+b, where aa and bb are the functions given in (49). Also, τ\uptau is defined in a maximal interval (s0,s1)(s_{0}\,,s_{1})\subset\mathbb{R} such that 0<τ|(s0,s1)<1.0<\uptau|_{(s_{0},s_{1})}<1. Since no horizontal section of Σ\varSigma is totally geodesic, we can assume 0<s0<s1+.0<s_{0}<s_{1}\leq+\infty.

The maximality of (s0,s1)(s_{0},s_{1}) gives that τ(s0)=0\uptau(s_{0})=0 or τ(s0)=1.\uptau(s_{0})=1. In the former case, we have τ(s0)=b(s0)>0.\uptau^{\prime}(s_{0})=b(s_{0})>0. Then, ϕ(s0)\phi(s_{0}) is well defined (by Lemma 2) and ϕ(s0)=0,\phi^{\prime}(s_{0})=0, so that x=(fs0(p),ϕ(s0)),x=(f_{s_{0}}(p),\phi(s_{0})), pn1,p\in\mathbb{R}^{n-1}, is a horizontal point of Σ\varSigma, contrary to our hypothesis. Then, we must have τ(s0)=1.\uptau(s_{0})=1. In particular, near s0s_{0} , τ\uptau is decreasing in (s0,s1),(s_{0},s_{1}), which implies that r<n.r<n. Indeed, for r=n,r=n, τ=b>0.\uptau^{\prime}=b>0.

Assume now that HrCrH_{r}\geq C_{r}, r<n.r<n. Then, we have

τ(s0)\displaystyle\uptau^{\prime}(s_{0}) =\displaystyle= a(s0)+b(s0)=(nr)tanh(s0)+brtanh1r(s0)\displaystyle a(s_{0})+b(s_{0})=-(n-r)\tanh(s_{0})+b_{r}\tanh^{1-r}(s_{0})
=\displaystyle= (nr)((Hr/Cr)tanh1r(s0)tanh(s0))\displaystyle(n-r)((H_{r}/C_{r})\tanh^{1-r}(s_{0})-\tanh(s_{0}))
\displaystyle\geq (nr)(tanh1r(s0)tanh(s0))>0,\displaystyle(n-r)(\tanh^{1-r}(s_{0})-\tanh(s_{0}))>0,

which contradicts that τ\uptau is decreasing near s0s_{0} .

It follows from the above considerations that, if Σ0,\varSigma_{0}\neq\emptyset, then r<nr<n and Hr<CrH_{r}<C_{r} . Furthermore, a direct computation gives that τ(s0)0\uptau^{\prime}(s_{0})\leq 0 if and only if s0srs_{0}\geq s_{r} . If s0=λ>srs_{0}=\lambda>s_{r} , then τ\uptau coincides with the function τλ\uptau_{\lambda} of the (fs,ϕ)(f_{s},\phi)-graph associated to the hypersurface Σ(λ)\varSigma(\lambda) of Theorem 12. From this, arguing as in preceding proofs, we conclude that Σ=Σ(λ).\varSigma=\varSigma(\lambda). By the same token, if s0=sr,s_{0}=s_{r}, then Σ=Σ\varSigma=\varSigma^{\prime} is the complete graph obtained in Theorem 13.

Let us suppose now that Σ0=.\varSigma_{0}=\emptyset. In this case, we must have Σ=s×,\varSigma=\mathscr{E}_{s}\times\mathbb{R}, where s=fs(n1)\mathscr{E}_{s}=f_{s}(\mathbb{R}^{n-1}) is an equidistant hypersurface with rr-th mean curvature Hrs=HrH_{r}^{s}=H_{r} , so that s=srs=s_{r} . Therefore, we have r<nr<n (since we are assuming Hr>0H_{r}>0) and

Hr=|Hrsr|=(n1r)tanhr(sr)=Crtanhr(sr)<Cr,H_{r}=|H_{r}^{s_{r}}|={{n-1}\choose{r}}\tanh^{r}(s_{r})=C_{r}\tanh^{r}(s_{r})<C_{r}\,,

which concludes our proof. ∎

7. Translational rr-minimal Hypersurfaces of 𝔽m×.\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R}.

In this section, we construct and classify rr-minimal hypersurfaces in 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which are invariant by translational isometries. It will be convenient to consider first the case of hyperbolic isometries of n.\mathbb{H}^{n}.

Theorem 15.

Let ={fs;s(,+)}\mathscr{F}=\{f_{s}\,;\,s\in(-\infty,+\infty)\} be a family of parallel equidistant hypersurfaces to a totally geodesic hyperplane 0=f0(n1)\mathscr{E}_{0}=f_{0}(\mathbb{R}^{n-1}) of  n.\mathbb{H}^{n}. Then, for each r{1,,n},r\in\{1,\dots,n\}, there exists a one-parameter family 𝒮={Σ(λ);λ>0}\mathscr{S}=\{\varSigma(\lambda)\,;\,\lambda>0\} of properly embedded rr-minimal hypersurfaces of  n×\mathbb{H}^{n}\times\mathbb{R} which are all homeomorphic to n\mathbb{R}^{n} and invariant by \mathscr{F}-hyperbolic translations. Each member Σ(λ)𝒮\varSigma(\lambda)\in\mathscr{S} has the following additional properties:

  • i)

    For r=n,r=n, Σ(λ)\varSigma(\lambda) is a constant angle entire rr-minimal graph over n\mathbb{H}^{n} whose height function is unbounded above and below.

For r<n,r<n, we distinguish the following cases:

  • ii)

    λ>1:\lambda>1: Σ(λ)\varSigma(\lambda) is symmetric with respect to the horizontal hyperplane P0=n×{0}P_{0}=\mathbb{H}^{n}\times\{0\}, and is contained in a slab n×(α,α),\mathbb{H}^{n}\times(-\alpha,\alpha), α>0.\alpha>0.

  • iii)

    λ=1:\lambda=1: Σ(λ)\varSigma(\lambda) is an (fs,ϕ)(f_{s},\phi)-graph (s>0s>0) which is bounded above, unbounded below, and asymptotic to 0×(0,).\mathscr{E}_{0}\times(0,-\infty).

  • iv)

    λ<1:\lambda<1: Σ(λ)\varSigma(\lambda) is an entire graph over n\mathbb{H}^{n} which is symmetric with respect to 0\mathscr{E}_{0}, and is contained is a slab n×(α,α),\mathbb{H}^{n}\times(-\alpha,\alpha), α>0.\alpha>0.

Furthermore, except for the cylinders s×,s0\mathscr{E}_{s}\times\mathbb{R},\,s\neq 0 (in the case r=nr=n), and up to ambient isometries, the members of 𝒮\mathscr{S} are the only complete non totally geodesic rr-minimal hypersurfaces of  n×\mathbb{H}^{n}\times\mathbb{R} which are invariant by hyperbolic translations.

Proof.

The equation (14) determined by \mathscr{F} and Hr=0H_{r}=0 is:

(54) y=a(s)y,a(s)=(nr)tanh(s).y^{\prime}=a(s)y,\,\,\,a(s)=-(n-r)\tanh(s).

For r=n,r=n, its solution τ\uptau is constant. So, given λ>0,\lambda>0, defining

ϕ(s)=λs,s(,+),\phi(s)=\lambda s,\,\,s\in(-\infty,+\infty),

we have that the corresponding (fs,ϕ)(f_{s},\phi)-graph Σ(λ)\varSigma(\lambda) is an entire nn-minimal graph whose level hypersurfaces are the leaves of .\mathscr{F}. Clearly, the height function of Σ(λ)\varSigma(\lambda) is unbounded above and below. Moreover, it follows from (8) that Σ(λ)\varSigma(\lambda) is a constant angle hypersurface. This proves (i).

Let us suppose now that 1r<n.1\leq r<n. Given λ>0,\lambda>0, set

τλ(s):=λ(1coshs)nr,s(,+).\uptau_{\lambda}(s):=\lambda\left(\frac{1}{\cosh s}\right)^{n-r},\,\,s\in(-\infty,+\infty).

It is easily checked that τλ\uptau_{\lambda} is the solution of (54) satisfying τλ(0)=λ.\uptau_{\lambda}(0)=\lambda.

Assume that λ>1.\lambda>1. Then, defining sλ:=arccosh(λ1/(nr)),s_{\lambda}:={\rm arccosh}\,(\lambda^{1/(n-r)}), one has

0<τλ(s)1=τλ(sλ)s[sλ,+).0<\uptau_{\lambda}(s)\leq 1=\uptau_{\lambda}(s_{\lambda})\,\,\,\forall s\in[s_{\lambda},+\infty).

Hence, setting

(55) ϕλ(s):=sλsϱλ(u)1ϱλ2(u)𝑑u,ϱλ=τλ1/r,s(sλ,+),\phi_{\lambda}(s):=\int_{s_{\lambda}}^{s}\frac{\varrho_{\lambda}(u)}{\sqrt{1-\varrho_{\lambda}^{2}(u)}}du,\quad\varrho_{\lambda}=\uptau_{\lambda}^{1/r},\quad s\in(s_{\lambda},+\infty),

we have that the (fs,ϕλ)(f_{s},\phi_{\lambda})-graph Σ(λ)\varSigma^{\prime}(\lambda) is a well defined rr-minimal hypersurface, for τ(sλ)<0.\uptau^{\prime}(s_{\lambda})<0. Also, since τλ(sλ)=1,\uptau_{\lambda}(s_{\lambda})=1, the closure of Σ(λ)\varSigma^{\prime}(\lambda) intersects P0P_{0} orthogonally. Thus, we obtain an rr-minimal hypersurface Σ(λ)\varSigma(\lambda) by reflecting Σ(λ)\varSigma^{\prime}(\lambda) about P0.P_{0}.

As for the boundedness of ϕλ\phi_{\lambda}, we first observe that, from the equalities τλ=ϱλr\uptau_{\lambda}=\varrho_{\lambda}^{r} and τλ=aτλ,\uptau_{\lambda}^{\prime}=a\uptau_{\lambda}, we have ϱλ=(r/a)ϱλ.\varrho_{\lambda}=(r/a)\varrho_{\lambda}^{\prime}. In addition, the function 1/a1/a is bounded above by 1/(nr)-1/(n-r) in (0,+).(0,+\infty). Hence,

ϕλ(s)\displaystyle\phi_{\lambda}(s) =\displaystyle= sλsrϱλ(u)a(u)1ϱλ2(u)𝑑urnrϱλ(sλ)ϱλ(s)dϱλ1ϱλ2\displaystyle\int_{s_{\lambda}}^{s}\frac{r\varrho_{\lambda}^{\prime}(u)}{a(u)\sqrt{1-\varrho_{\lambda}^{2}(u)}}du\leq-\frac{r}{n-r}\int_{\varrho_{\lambda}(s_{\lambda})}^{\varrho_{\lambda}(s)}\frac{d\varrho_{\lambda}}{\sqrt{1-\varrho_{\lambda}^{2}}}
=\displaystyle= rnr(arcsinϱλ(sλ)arcsinϱλ(s))πr2(nr),\displaystyle\frac{r}{n-r}(\arcsin\varrho_{\lambda}(s_{\lambda})-\arcsin\varrho_{\lambda}(s))\leq\frac{\pi r}{2(n-r)}\,,

which finishes the proof of (ii).

Assuming now λ=1,\lambda=1, let us fix s0>0s_{0}>0 and define

ϕ(s)=s0sϱ(u)1ϱ2(u)𝑑u,ϱ=τ11/r,s(0,+).\phi(s)=\int_{s_{0}}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du,\quad\varrho=\uptau_{1}^{1/r},\quad s\in(0,+\infty).

Since τ(0)=0,\uptau^{\prime}(0)=0, we can mimic the final part of the proof of Theorem 13 and conclude that ϕ\phi is unbounded below, and that the corresponding (fs,ϕ)(f_{s},\phi)-graph Σ\varSigma is asymptotic to 0×(,0).\mathscr{E}_{0}\times(-\infty,0). Also, proceeding as in (7), we can show that ϕ\phi is bounded above. This proves (iii).

Given 0<λ<1,0<\lambda<1, we have that 0<τλ<10<\uptau_{\lambda}<1 in (0,+).(0,+\infty). So, we can define ϕλ\phi_{\lambda} as in (55), replacing sλs_{\lambda} by 0. Analogously, we have that ϕλ\phi_{\lambda} is bounded above, and that the boundary of the (fs,ϕλ)(f_{s},\phi_{\lambda})-graph Σ(λ)\varSigma^{\prime}(\lambda) is 0×{0}P0\mathscr{E}_{0}\times\{0\}\subset P_{0} .

Notice that ϕλ(0)=ϱλ(0)/1ϱλ2(0)\phi_{\lambda}^{\prime}(0)=\varrho_{\lambda}(0)/\sqrt{1-\varrho_{\lambda}^{2}(0)} is well defined and positive, since ϱλ(0)\varrho_{\lambda}(0) is neither 0 nor 1.1. Thus, we obtain a complete properly embedded rr-minimal hypersurface Σ(λ)\varSigma(\lambda) from Σ(λ)\varSigma^{\prime}(\lambda) by reflecting it with respect to P0P_{0} , and then with respect to the totally geodesic vertical hyperplane 0×\mathscr{E}_{0}\times\mathbb{R} (Fig. 15). This shows (iv).

Assume now that Σ\varSigma is a complete non totally geodesic rr-minimal hypersurface of n×\mathbb{H}^{n}\times\mathbb{R} which is invariant by \mathscr{F}-hyperbolic translations. Set

Σ0:={xΣ;Θ(x)ξ(x)0}\varSigma_{0}:=\{x\in\varSigma\,;\,\varTheta(x)\nabla\xi(x)\neq 0\}

and suppose that 1r<n.1\leq r<n. Then, Σ0.\varSigma_{0}\neq\emptyset. Otherwise, Σ\varSigma would be either a horizontal hyperplane or a cylinder over the hyperplane 0\mathscr{E}_{0} of n.\mathbb{H}^{n}. In both cases, Σ\varSigma would be totally geodesic, which is contrary to our assumption.

Therefore, if 1r<n,1\leq r<n, for each x0Σ0x_{0}\in\varSigma_{0} , there is an (fs,ϕ)(f_{s},\phi)-graph ΣΣ0\varSigma^{\prime}\subset\varSigma_{0} which contains x0x_{0} , and whose τ\uptau-function is a solution of (54). More precisely, for some c>0,c>0, one has

τ(s)=c(coshs0coshs)nr,s0,s(,+).\uptau(s)=c\left(\frac{\cosh s_{0}}{\cosh s}\right)^{n-r},\,\,s_{0}\,,s\in(-\infty,+\infty).

Now, recall that in the cases (ii)–(iv) above, the corresponding function τλ\uptau_{\lambda} satisfies τλ(0)=λ.\uptau_{\lambda}(0)=\lambda. Thus, for λ:=τ(0)=ccoshnr(s0),\lambda:=\uptau(0)=c\cosh^{n-r}(s_{0}), the function τ\uptau coincides with τλ,\uptau_{\lambda}, which implies that ΣΣ(λ).\varSigma^{\prime}\subset\varSigma(\lambda). In addition, no Σ(λ)𝒮\varSigma(\lambda)\in\mathscr{S} has horizontal or totally geodesic points, which gives that Σ0\varSigma_{0} is open and dense in Σ.\varSigma. Therefore, Σ\varSigma coincides with Σ(λ).\varSigma(\lambda).

Finally, let us suppose that r=n.r=n. If Σ0=,\varSigma_{0}=\emptyset, then Σ=s\varSigma=\mathscr{E}_{s} for some s0.s\neq 0. If Σ0,\varSigma_{0}\neq\emptyset, then there exists an (fs,ϕ)(f_{s},\phi)-graph ΣΣ0\varSigma^{\prime}\subset\varSigma_{0} whose τ\uptau-function is constant. In particular, up to a vertical translation, we have ϕ(s)=λs\phi(s)=\lambda s for some λ>0,\lambda>0, so that Σ=Σ\varSigma^{\prime}=\varSigma is the entire graph Σ(λ)\varSigma(\lambda) given in (i). ∎

Refer to caption
Figure 15. The figure shows half of the (fs,ϕ)(f_{s},\phi)-graph Σ(λ)\varSigma^{\prime}(\lambda) (above n×{0}\mathbb{H}^{n}\times\{0\}) and half of its reflection with respect to 0\mathscr{E}_{0} (below n×{0}\mathbb{H}^{n}\times\{0\}).
Remark 6.

It should be mentioned that the particular case r=1r=1 of Theorem 15 was considered in [4].

Next, we obtain all complete non totally geodesic rr-minimal hypersurfaces of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which are invariant by parabolic isometries.

Theorem 16.

Let ={fs;s(,+)}\mathscr{F}=\{f_{s}\,;\,s\in(-\infty,+\infty)\} be a family of parallel horospheres in 𝔽m.\mathbb{H}_{\mathbb{F}}^{m}. Then, for any r{1,,n},r\in\{1,\dots,n\}, there exists a properly embedded rr-minimal hypersurface Σ\varSigma of  𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} which is invariant by \mathscr{F}-parabolic isometries. In addition, Σ\varSigma is homeomorphic to n\mathbb{R}^{n} and has the following properties:

  • i)

    For r=n,r=n, Σ\varSigma is a constant angle entire graph over 𝔽m\mathbb{H}_{\mathbb{F}}^{m} whose height function is unbounded above and below.

  • ii)

    For r<n,r<n, Σ\varSigma is symmetric with respect to 𝔽m×{0}\mathbb{H}_{\mathbb{F}}^{m}\times\{0\} and is contained in a slab 𝔽m×(α,α).\mathbb{H}_{\mathbb{F}}^{m}\times(-\alpha,\alpha).

Furthermore, except for the cylinders s×\mathscr{H}_{s}\times\mathbb{R} (in the case r=nr=n), and up to ambient isometries, Σ\varSigma is the only complete non totally geodesic rr-minimal hypersurface of  n×\mathbb{H}^{n}\times\mathbb{R} which is invariant by parabolic isometries.

Proof.

The proof of the existence of Σ\varSigma as in (i) is analogous to the one given in the preceding theorem. So, let us assume r<n.r<n. In this case, the equation (14) determined by \mathscr{F} and Hr=0H_{r}=0 takes the form

(57) y=ay,a=rHr0Hr10>0,y^{\prime}=ay,\,\,\,a=\frac{rH_{r}^{0}}{H_{r-1}^{0}}>0\,,

and its positive solutions are

(58) τλ(s)=λeas,λ>0,s(,+).\uptau_{\lambda}(s)=\lambda e^{as},\,\,\,\,\lambda>0,\,\,s\in(-\infty,+\infty).

It is easily checked that, since the horospheres of 𝔽m\mathbb{H}_{\mathbb{F}}^{m} are pairwise congruent, an (fs,ϕ)(f_{s},\phi) graph with τ\uptau-function τλ\uptau_{\lambda} does not depend on λ.\lambda. More precisely, two such graphs obtained from functions τλ1\uptau_{\lambda_{1}} and τλ2\uptau_{\lambda_{2}} , λ1λ2,\lambda_{1}\neq\lambda_{2}, are isometric. Therefore, we can assume λ=1\lambda=1 and set τ:=τ1\uptau:=\uptau_{1} . Then, we have

0<τ(s)<1=τ(0)s(,0).0<\uptau(s)<1=\uptau(0)\,\,\,\forall s\in(-\infty,0).

Since τ(0)=aτ(0)=a>0,\uptau^{\prime}(0)=a\uptau(0)=a>0, writing ϱr=τ|(,0)\varrho^{r}=\uptau|_{(-\infty,0)}, we have

ϕ(s)=0sϱ(u)1ϱ2(u)𝑑u,s(,0),\phi(s)=\int_{0}^{s}\frac{\varrho(u)}{\sqrt{1-\varrho^{2}(u)}}du,\,\,\,s\in(-\infty,0),

is well defined, and so is the corresponding (fs,ϕ)(f_{s},\phi)-graph Σ.\varSigma^{\prime}. Also, τ(0)=1,\uptau(0)=1, so that the tangent spaces of Σ\varSigma^{\prime} along its boundary are all vertical. Therefore, we obtain the stated rr-minimal hypersurface Σ\varSigma by reflecting Σ\varSigma^{\prime} about P0=𝔽m×{0}.P_{0}=\mathbb{H}_{\mathbb{F}}^{m}\times\{0\}.

Observe that, for all s(,0),s\in(-\infty,0), one has

ϕ(s)=s0eau/r1e2au/r𝑑u=ra(π/2arcsin(eas/r)).-\phi(s)=\int_{s}^{0}\frac{e^{au/r}}{\sqrt{1-e^{2au/r}}}du=\frac{r}{a}(\pi/2-\arcsin(e^{as/r})).

Hence, setting α:=πr/2a>0,\alpha:={\pi r}/{2a}>0, we have that ϕ(s)α\phi(s)\rightarrow-\alpha as s,s\rightarrow-\infty, which proves that Σ\varSigma is contained in the slab 𝔽m×(α,α).\mathbb{H}_{\mathbb{F}}^{m}\times(-\alpha,\alpha).

As for the uniqueness of Σ\varSigma, notice that the following hold:

  • The τ\uptau-function of any rr-minimal (fs,ϕ)(f_{s},\phi)-graph, fs,f_{s}\in\mathscr{F}, is a positive solution of (57) (if r<nr<n) or is a positive constant (if r=nr=n).

  • Σ\varSigma has no horizontal points.

  • A vertical \mathscr{F}-invariant hypersurface of 𝔽m×\mathbb{H}_{\mathbb{F}}^{m}\times\mathbb{R} is rr-minimal if and only if r=n.r=n.

  • The graph in (i) has no vertical points.

These facts allow us to argue as in preceding proofs, and then show the uniqueness of Σ\varSigma as asserted. ∎

8. Uniqueness of Rotational HrH_{r}-spheres of ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R}

In this concluding section, we concern the uniqueness of the rotational HrH_{r}-spheres we constructed in Section 4. We restrict ourselves to ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R}, with ϵ{1,1}\epsilon\in\{-1,1\} and n3.n\geq 3. As we mentioned before, the case n=2n=2 was considered in [1, 20].

We obtain a Jellett–Liebmann type theorem by showing that a compact, connected and strictly convex HrH_{r}-hypersurface of ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R} is a rotational embedded sphere (cf. Theorem 17). We also show the uniqueness of these spheres under completeness or properness assumptions, instead of compactness (cf. Theorem 18 and Corollary 2).

For the proof of Theorem 18, we make use of a height estimate for convex graphs in M×M\times\mathbb{R} which we establish in the next proposition. First, we compute the Laplacian of both the height function ξ\xi and the angle function Θ\varTheta of an arbitrary hypersurface Σ\varSigma of a general product M×.M\times\mathbb{R}.

Given a smooth function ζ\zeta on Σ,\varSigma, let us denote its Laplacian by Δζ,\Delta\zeta, i.e.,

Δζ:=trace(Hessζ).\Delta\zeta:={\rm trace}({\rm Hess}\,\zeta).

In particular, from equation (3), the Laplacian of ξ\xi is given by

(59) Δξ=ΘH,H=H1.\Delta\xi=\varTheta H,\,\,\,H=H_{1}\,.

Recall that, for X,YTΣ,X,Y\in T\varSigma, the Codazzi equation reads as

(R¯(X,Y)N)=(YA)X(XA)Y,(\mkern 1.5mu\overline{\mkern-1.5muR\mkern-1.5mu}\mkern 1.5mu(X,Y)N)^{\top}=\left(\nabla_{Y}A\right)X-\left(\nabla_{X}A\right)Y,

where R¯\mkern 1.5mu\overline{\mkern-1.5muR\mkern-1.5mu}\mkern 1.5mu is the curvature tensor of M×,M\times\mathbb{R}, \top denotes the tangent component of the tangent bundle TΣT\varSigma of Σ,\varSigma, and, by definition,

(YA)X:=YAXAYX.\left(\nabla_{Y}A\right)X:=\nabla_{Y}AX-A\nabla_{Y}X.

Observing that

Xξ=(¯Xξ)=(¯XΘN)=ΘAX,\nabla_{X}\nabla\xi=\left(\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu_{X}\nabla\xi\right)^{\top}=-\left(\mkern 1.5mu\overline{\mkern-1.5mu\nabla\mkern-1.5mu}\mkern 1.5mu_{X}\varTheta N\right)^{\top}=\varTheta AX,

we have from (2) that

XΘ=XAξ=(XA)ξ+AXξ=(XA)ξ+ΘA2X,-\nabla_{X}\nabla\varTheta=\nabla_{X}A\nabla\xi=\left(\nabla_{X}A\right)\nabla\xi+A\nabla_{X}\nabla\xi=\left(\nabla_{X}A\right)\nabla\xi+\varTheta A^{2}X,

which yields

(60) XΘ=(R¯(ξ,X)N)(ξA)XΘA2X.\nabla_{X}\nabla\varTheta=-(\mkern 1.5mu\overline{\mkern-1.5muR\mkern-1.5mu}\mkern 1.5mu(\nabla\xi,X)N)^{\top}-\left(\nabla_{\nabla\xi}A\right)X-\varTheta A^{2}X.

Now, let us fix xΣx\in\varSigma and an orthonormal frame {X1,,Xn}\{X_{1}\,,\dots,X_{n}\} in a neighborhood of xx in Σ,\varSigma, which is geodesic at x,x, that is

XiXj(x)=0i,j=1,,n.\nabla_{X_{i}}X_{j}\,(x)=0\,\,\,\forall i,j=1,\dots,n.

Writing ξj=Xj(ξ),\xi_{j}=X_{j}(\xi), we have ξ=jξjXj.\nabla\xi=\sum_{j}\xi_{j}X_{j}. Therefore

i=1n(ξA)Xi,Xi\displaystyle\sum_{i=1}^{n}\langle\left(\nabla_{\nabla\xi}A\right)X_{i}\,,\,X_{i}\rangle =\displaystyle= i=1n(ξAXi,XiAξXi,Xi)\displaystyle\sum_{i=1}^{n}\left(\langle\nabla_{\nabla\xi}AX_{i}\,,\,X_{i}\rangle-\langle A\nabla_{\nabla\xi}X_{i}\,,\,X_{i}\rangle\right)
=\displaystyle= i,j=1nξj(XjAXi,XiAXjXi,Xi)\displaystyle\sum_{i,j=1}^{n}\xi_{j}(\langle\nabla_{X_{j}}AX_{i}\,,\,X_{i}\rangle-\langle A\nabla_{X_{j}}X_{i}\,,\,X_{i}\rangle)
=\displaystyle= i,j=1nξj(XjAXi,XiAXi,XjXiAXjXi,Xi)\displaystyle\sum_{i,j=1}^{n}\xi_{j}(X_{j}\langle AX_{i},X_{i}\rangle-\langle AX_{i},\nabla_{X_{j}}X_{i}\rangle-\langle A\nabla_{X_{j}}X_{i}\,,\,X_{i}\rangle)
=\displaystyle= ξ,Hi,j=1nξj(AXi,XjXiAXjXi,Xi),\displaystyle\langle\nabla\xi,\nabla H\rangle-\sum_{i,j=1}^{n}\xi_{j}(\langle AX_{i},\nabla_{X_{j}}X_{i}\rangle-\langle A\nabla_{X_{j}}X_{i}\,,\,X_{i}\rangle),

which implies that, at the chosen point xΣ,x\in\varSigma,

i=1n(ξA)Xi,Xi=ξ,H.\sum_{i=1}^{n}\langle\left(\nabla_{\nabla\xi}A\right)X_{i}\,,\,X_{i}\rangle=\langle\nabla\xi,\nabla H\rangle.

Since xx is arbitrary, we get from this last equality and (60) that, on Σ,\varSigma,

(61) ΔΘ=Ric¯(ξ,N)ξ,HΘA2,\Delta\varTheta=\mkern 1.5mu\overline{\mkern-1.5mu\rm Ric\mkern-1.5mu}\mkern 1.5mu(\nabla\xi,N)-\langle\nabla\xi,\nabla H\rangle-\varTheta\|A\|^{2},

where Ric¯\mkern 1.5mu\overline{\mkern-1.5mu\rm Ric\mkern-1.5mu}\mkern 1.5mu denotes the Ricci curvature tensor of M×M\times\mathbb{R} and A2:=traceA2.\|A\|^{2}:={\rm trace}\,A^{2}.

Remark 7.

For the next results, except for Theorem 19, we order the principal curvatures of a hypersurface Σ\varSigma of M×M\times\mathbb{R} as

k1k2kn1kn.k_{1}\leq k_{2}\leq\cdots\leq k_{n-1}\leq k_{n}\,.
Proposition 3.

Consider an arbitrary Riemannian manifold  M,M, and let ΣM×\varSigma\subset M\times\mathbb{R} be a compact vertical graph of a nonnegative function defined on a domain ΩM×{0}.\Omega\subset M\times\{0\}. Assume Σ\varSigma strictly convex up to ΣM×{0}.\partial\varSigma\subset M\times\{0\}. Under these conditions, the following height estimate holds:

(62) ξ(x)1infΣk1xΣ.\xi(x)\leq\frac{1}{\inf_{\varSigma}k_{1}}\,\,\,\forall x\in\varSigma.
Proof.

Consider in Σ\varSigma the “inward” orientation, so that its angle function Θ\varTheta is non positive. Choose δ>0\delta>0 satisfying 1/δ<infΣk11/\delta<\inf_{\varSigma}k_{1} and define on Σ\varSigma the function

φ=ξ+δΘ.\varphi=\xi+\delta\varTheta.

We claim that φ\varphi has no interior maximum. Indeed, assuming otherwise, let xΣΣx\in\varSigma-\partial\varSigma be a maximum point of φ\varphi. In this case, from (2), we have

0=φ(x)=ξ(x)+δΘ(x)=ξ(x)δAξ(x).0=\nabla\varphi(x)=\nabla\xi(x)+\delta\nabla\varTheta(x)=\nabla\xi(x)-\delta A\nabla\xi(x).

Hence, if we had ξ(x)0,\nabla\xi(x)\neq 0, then 1/δ1/\delta would be an eigenvalue of AA at x,x, which is impossible, by our choice of δ.\delta. Thus, xx is a critical point of ξ.\xi. Since Σ\varSigma is strictly convex, xx is necessarily its highest point. In particular, Θ(x)=1.\varTheta(x)=-1. This, together with identities (59) and (61), gives that, at xx,

(63) 0Δφ=H+δA2.0\geq\Delta\varphi=-H+\delta\|A\|^{2}.

However, from our choice of δ\delta and the strict convexity of Σ\varSigma, we have

Hδ<k1H=k1(k1++kn)k12++kn2=A2,\frac{H}{\delta}<k_{1}H=k_{1}(k_{1}+\cdots+k_{n})\leq k_{1}^{2}+\cdots+k_{n}^{2}=\|A\|^{2},

which contradicts (63). Therefore, φ\varphi attains its maximum on Σ,\partial\varSigma, which implies that φ0\varphi\leq 0 on Σ,\varSigma, for φ|Σ=δΘ0.\varphi|_{\partial\varSigma}=\delta\varTheta\leq 0. Hence,

ξ(x)δΘ(x)δxΣ.\xi(x)\leq-\delta\varTheta(x)\leq\delta\,\,\,\forall x\in\varSigma.

The result, then, follows from this last inequality, since it holds for any positive δ>1/infΣk1.\delta>1/\inf_{\varSigma}k_{1}.

Remark 8.

Proposition 3 has its own importance, since it establishes height estimates for vertical graphs in M×M\times\mathbb{R} making no assumptions on M.M. In addition, no curvature of such a graph is assumed to be constant.

In the next two theorems, we apply the Alexandrov reflection technique. Since the arguments are standard, the proofs will be somewhat sketchy on this matter (see, e.g., [9, Theorems 4.2 and 5.1] and [29, Theorem 1.1]). We add that the proof of Theorem 17 is, essentially, the one for [11, Corollary 1], in which the case r=1r=1 was considered.

Theorem 17 (Jellett–Liebmann-type theorem).

Let Σ\varSigma be a compact connected strictly convex Hr(>0)H_{r}(>0)-hypersurface of  ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R} (n3n\geq 3). Then, Σ\varSigma is an embedded rotational HrH_{r}-sphere.

Proof.

Since Σ\varSigma is compact, its height function ξ\xi has a maximal point x.x. This, together with the strict convexity of Σ\varSigma, allows us to apply [11, Theorems 1 and 2] and conclude that Σ\varSigma is embedded and homeomorphic to 𝕊n\mathbb{S}^{n}. Thus, for ϵ=1,\epsilon=-1, the result follows from [18, Theorem 7.6], the Alexandrov-type theorem we mentioned in the introduction.

For ϵ=1,\epsilon=1, we can perform Alexandrov reflections on Σ\varSigma with respect to horizontal hyperplanes Pt:=𝕊n×{t}P_{t}:=\mathbb{S}^{n}\times\{t\} coming down from above Σ.\varSigma. For some t0<ξ(x),t_{0}<\xi(x), the reflection of the part of Σ\varSigma above Pt0P_{t_{0}} will have a first contact with Σ.\varSigma. Then, by the Maximum-Continuation Principle, Σ\varSigma is symmetric with respect to Pt0P_{t_{0}}. Therefore, assuming t0=0t_{0}=0 and identifying 𝕊n×{0}\mathbb{S}^{n}\times\{0\} with 𝕊n,\mathbb{S}^{n}, we conclude that Σ\varSigma is a “bigraph” over its projection π(Σ)\pi(\varSigma) to 𝕊n\mathbb{S}^{n}. As a consequence, Σ0:=Σ𝕊n\varSigma_{0}:=\varSigma\cap\mathbb{S}^{n} is the boundary of π(Σ)\pi(\varSigma) in 𝕊n\mathbb{S}^{n}.

By [11, Lemma 1], the second fundamental form of Σ0\varSigma_{0}, as a hypersurface of  𝕊n\mathbb{S}^{n}, is positive definite. In particular, Σ0\varSigma_{0} is non totally geodesic in  𝕊n\mathbb{S}^{n}. Thus, by [14, Theorem 1], Σ0\varSigma_{0} is contained in an open hemisphere 𝕊+n\mathbb{S}_{+}^{n} of 𝕊n\mathbb{S}^{n}, which implies that the same is true for π(Σ),\pi(\varSigma), that is, Σ𝕊+n×.\varSigma\subset\mathbb{S}_{+}^{n}\times\mathbb{R}. In this setting, we can apply Alexandrov reflections on “vertical hyperplanes” (𝕊n1𝕊+n)×,(\mathbb{S}^{n-1}\cap\mathbb{S}_{+}^{n})\times\mathbb{R}, where 𝕊n1𝕊n\mathbb{S}^{n-1}\subset\mathbb{S}^{n} is a totally geodesic (n1)(n-1)-sphere of 𝕊n\mathbb{S}^{n} , and conclude that Σ\varSigma is rotational. ∎

Let us show now that, regarding Theorem 17, the compactness hypothesis can be replaced by completeness if we add a one point condition on the height function of Σ.\varSigma. In the case ϵ=1\epsilon=-1, we also have to impose a condition on the second fundamental form of Σ,\varSigma, which turns out to be a necessary hypothesis (see Remark 10, below).

Theorem 18.

Let Σ\varSigma be a complete connected strictly convex Hr(>0)H_{r}(>0)-hypersurface of ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R} (n3n\geq 3) whose height function ξ\xi has a local extreme point. For ϵ=1\epsilon=-1, assume further that the least principal curvature k1k_{1} of Σ\varSigma is bounded away from zero. Then, Σ\varSigma is an embedded rotational sphere.

Proof.

As in the previous theorem, Σ\varSigma fulfills the hypotheses of [11, Theorems 1 and 2], which implies that Σ\varSigma is properly embedded and homeomorphic to either 𝕊n\mathbb{S}^{n} or n.\mathbb{R}^{n}. Furthermore, in the latter case, the height function of Σ\varSigma is unbounded and has a single extreme point x,x, which we assume to be a maximum.

For ϵ=1,\epsilon=1, the height estimates obtained in [9, Theorem 4.1-(i)] forbid ξ\xi to be unbounded. Thus, in this case, Σ\varSigma is homeomorphic to 𝕊n\mathbb{S}^{n} and the result follows from Theorem 17.

Let us consider now the case ϵ=1.\epsilon=-1. Assume, by contradiction, that Σ\varSigma is homeomorphic to n,\mathbb{R}^{n}, so that ξ\xi is unbounded below. Hence, given a horizontal hyperplane Pt=M×{t}P_{t}=M\times\{t\} with t<ξ(x),t<\xi(x), the part Σt+\varSigma_{t}^{+} of Σ\varSigma which lies above PtP_{t} must be a vertical graph with boundary in PtP_{t}. If not, for some tt^{\prime} between tt and ξ(x),\xi(x), PtP_{t^{\prime}} would be orthogonal to Σ\varSigma at one of its points. Then, the Alexandrov reflection method would give that Σ\varSigma is symmetric with respect to Pt,P_{t^{\prime}}, which is impossible, since we are assuming ξ\xi unbounded, and the closure of Σt+\varSigma_{t^{\prime}}^{+} in Σ\varSigma is compact.

It follows from the above that, for |t||t| sufficiently large, one has

ξ(x)t>1infΣk11infΣt+k1,\xi(x)-t>\frac{1}{\inf_{\varSigma}k_{1}}\geq\frac{1}{\inf_{\varSigma_{t}^{+}}k_{1}}\,,

which clearly contradicts Proposition 3. Therefore, Σ\varSigma is homeomorphic to 𝕊n\mathbb{S}^{n} and, again, the result follows from Theorem 17. ∎

Remark 9.

In Theorems 17 and 18, the hypothesis of strict convexity of Σ\varSigma is automatically satisfied for r=n,r=n, so it can be dropped in this case. Indeed, in both theorems, the height function ξ\xi has a critical point xΣx\in\varSigma, which can be assumed to be a maximum. Then, taking the inward orientation on Σ,\varSigma, we have that Θ(x)=1,\varTheta(x)=-1, which, together with equality (3), yields

AX,X=Hessξ(X,X)0XTxΣ.\langle AX,X\rangle=-{\rm Hess}\,\xi(X,X)\geq 0\,\,\,\forall X\in T_{x}\varSigma.

However, Hn=detA>0H_{n}=\det A>0 on Σ\varSigma. Thus, at x,x, and then on all of Σ\varSigma, the second fundamental form is positive definite, that is, Σ\varSigma is strictly convex.

Remark 10.

It follows from the considerations of Remark 1 that, for r<n,r<n, the hypothesis on the least principal curvature of Σ\varSigma in Theorem 18 is necessary for the conclusion. As shown by Theorem 5, the same is true for the hypothesis on the height function ξ\xi in the case ϵ=1\epsilon=1 and r<n.r<n.

Next, we consider the dual case of Theorem 18, assuming now that the height function of the hypersurface Σ\varSigma has no critical points. First, we recall that a hypersurface Σn×\varSigma\subset\mathbb{H}^{n}\times\mathbb{R} is said to be cylindrically bounded, if there exists a closed geodesic ball BnB\subset\mathbb{H}^{n} such that ΣB×.\varSigma\subset B\times\mathbb{R}.

Theorem 19.

Let Σ\varSigma be a proper, convex, connected Hr(>0)H_{r}(>0)-hypersurface of  ϵn×\mathbb{Q}_{\epsilon}^{n}\times\mathbb{R} (n3n\geq 3) with no horizontal points. For ϵ=1,\epsilon=-1, assume further that Σ\varSigma is cylindrically bounded. Then, Σ\varSigma is a cylinder over a geodesic sphere of  ϵn\mathbb{Q}_{\epsilon}^{n}. In particular, r<n.r<n.

Proof.

From the hypothesis and [11, Theorem 3], Σ=Σ0×,\varSigma=\varSigma_{0}\times\mathbb{R}, where Σ0\varSigma_{0} is an embedded convex topological sphere of ϵn\mathbb{Q}_{\epsilon}^{n} . Moreover, in the case ϵ=1,\epsilon=1, Σ0\varSigma_{0} is contained in an open hemisphere of 𝕊n.\mathbb{S}^{n}.

At a given point xΣ,x\in\varSigma, the principal curvatures are k1,,kn1,0,k_{1},\,\dots,k_{n-1},0, where k1,,kn1k_{1}\,,\dots,k_{n-1} are the principal curvatures of Σ0ϵn\varSigma_{0}\subset\mathbb{Q}_{\epsilon}^{n} at πϵn(x)Σ0\pi_{\mathbb{Q}^{n}_{\epsilon}}(x)\in\varSigma_{0} . In particular, Σ0\varSigma_{0} has constant rr-th mean curvature HrH_{r} if r<n,r<n, which implies that, in this case, Σ0\varSigma_{0} is a geodesic sphere of ϵn\mathbb{Q}_{\epsilon}^{n} (see [26, 28]). Also, Hn=0H_{n}=0 on Σ\varSigma, so we must have r<n,r<n, since we are assuming Hr>0.H_{r}>0.

Since a cylinder Σ0×𝕊n×\varSigma_{0}\times\mathbb{R}\subset\mathbb{S}^{n}\times\mathbb{R} is nowhere strictly convex, it follows from the above theorem that, for n3,n\geq 3, a connected, proper, and strictly convex HrH_{r}-hypersurface of 𝕊n×\mathbb{S}^{n}\times\mathbb{R} must have a horizontal point. This fact, together with Theorem 18, gives our last result:

Corollary 2.

For n3,n\geq 3, any connected, properly immersed, and strictly convex Hr(>0)H_{r}(>0)-hypersurface of  𝕊n×\mathbb{S}^{n}\times\mathbb{R} is necessarily an embedded rotational HrH_{r}-sphere.

9. Acknowledgments

We are indebt to Antonio Martinez, Pablo Mira, and Miguel Domínguez-Vázquez for their valuable suggestions. Fernando Manfio is supported by Fapesp, grant 2016/23746-6. João Paulo dos Santos is supported by FAPDF, grant 0193.001346/
2016.

References

  • [1] Abresch, U., Rosenberg, H.: A Hopf differential for constant mean curvature surfaces in 𝕊2×\mathbb{S}^{2}\times\mathbb{R} and 2×.\mathbb{H}^{2}\times\mathbb{R}. Acta Math. 193, 141–174 (2004).
  • [2] Aravinda, C. S., Farrel, F. T.: Exotic negatively curved structures on Cayley hyperbolic manifolds. J. Diff. Geom. 63, 41–62 (2003).
  • [3] Bérard, P., Sa Earp, R.: Examples of HH-hypersurfaces in n×\mathbb{H}^{n}\times\mathbb{R} and geometric applications. Matemática Contemporânea, Sociedade Brasileira de Matemática. 34, 19–51 (2008).
  • [4] Bérard, P., Sa Earp, R.: Minimal hypersurfaces in n×,\mathbb{H}^{n}\times\mathbb{R}, total curvature and index. Boll. Unione Mat. Ital. 9, 341–362 (2016).
  • [5] Berndt, J., Tricerri, F., Vanhecke, L.: Generalized Heisenberg groups and Damek-Ricci harmonic spaces. Lecture Notes in Mathematics 1598. Springer Verlag (1995).
  • [6] Bishop, R. L., O’Neill, B.: Manifolds of negative curvature. Trans. Amer. Math. Soc. 145, 1–49 (1969).
  • [7] Cecil, T., Ryan, P.: Geometry of hypersurfaces. Springer Verlag (2015).
  • [8] Chen, S.S., Greenberg, L.: Hyperbolic spaces. In: Ahlfors, L.V., Kra, I., Maskit, B., Nirenberg, L. (eds.) Contributions to Analysis, pp. 49–87. Academic Press, New York (1974).
  • [9] Cheng, X., Rosenberg, H.: Embedded positive constant r-mean curvature hypersurfaces in Mm×.M^{m}\times\mathbb{R}. An. Acad. Brasil. Ciênc. 77, 183–199 (2005).
  • [10] Daniel, B.: Isometric immersions into  𝕊n×\mathbb{S}^{n}\times\mathbb{R}  and  n×\mathbb{H}^{n}\times\mathbb{R}  and applications to minimal surfaces. Trans. Amer. Math. Soc. 361, Number 12, 6255–-6282 (2009).
  • [11] de Lima, R. F.: Embeddedness, convexity, and rigidity of hypersurfaces in product spaces. Preprint. To appear in Ann. Global Anal. Geom. (available at: https://arxiv.org/abs/1806.01509).
  • [12] de Lima, R.F., Roitman, P.: Helicoids and catenoids in M×.M\times\mathbb{R}. Preprint. To appear in Ann. Mat. Pura Appl. (available at: https://arxiv.org/abs/1901.07936).
  • [13] Díaz-Ramos, J. C., Domínguez-Vázquez, M.: Isoparametric hypersurfaces in Damek–Ricci spaces. Adv. in Math. 239, 1–17 (2013).
  • [14] do Carmo, M., Warner, F.: Rigidity and convexity of hypersurfaces in spheres, J. Diff. Geom. 4, 133–144 (1970).
  • [15] M. Domínguez-Vázquez: An introduction to isoparametric foliations. Preprint (2018) (avaiable at: http://xtsunxet.usc.es/miguel/teaching/jae2018.html).
  • [16] Domínguez-Vázquez, M., Manzano, J. M.: Isoparametric surfaces in 𝔼(k,τ)\mathbb{E}(k,\uptau)-spaces. Preprint. To appear in Ann. Sc. Norm. Super. Pisa Cl. Sci. (available at https://arxiv.org/abs/1803.06154).
  • [17] Eberlein, P., O’Neill, B.: Visibility manifolds. Pacific Jour. Math. 46, 45–109 (1973).
  • [18] Elbert, M. F., Sá Earp, R.: Constructions of HrH_{r}-hypersurfaces, barriers and Alexandrov theorem in n×\mathbb{H}^{n}\times\mathbb{R}. Ann. Mat. Pura Appl. 194, 1809–1834 (2015).
  • [19] Elbert, M. F., Nelli, B., Santos, W.: Hypersurfaces with Hr+1=0H_{r+1}=0 in n×\mathbb{H}^{n}\times\mathbb{R}. Manuscripta math. 149 , 507–521 (2016).
  • [20] Espinar, J., Gálvez, A.: Rosenberg, H.: Complete surfaces with positive extrinsic curvature in product spaces, Comment. Math. Helv. 84, 351–386 (2009)
  • [21] Fontenele, F., Silva, S.: A tangency principle and applications. Illinois J. of Math. 54, 213–228 (2001).
  • [22] Hauswirth, L.: Minimal surfaces of Riemann type in three-dimensional product manifolds, Pacific J. Math. 224, no. 1, 91–117 (2006).
  • [23] Kazdan, J.: Unique continuation in geometry. Comm. Pure Appl. Math., Vol. XLI 667–681 (1988).
  • [24] Kim, Y.: Quasiconformal conjugacy classes of parabolic isometries of complex hyperbolic space. Pacific J. Math., 270 129–149 (2014).
  • [25] Kim, S., Nikolayevsky, Y., Park, J.: Einstein hypersurfaces of the Cayley projective plane. Differential Geom. Appl. 69, 1–6 (2020).
  • [26] Korevaar, N.: Sphere theorems via Alexandrov for constant Weingarten curvature hypersurfaces - appendix to a note of A. Ros. J. Diff. Geom. 27, 221–223 (1988)
  • [27] Mahmoudi, F.: Constant k-curvature hypersurfaces in Riemannian manifolds. Differential Geom. Appl. 28, 1–11 (2010).
  • [28] Montiel, S., Ros, A.: Compact hypersurfaces: the Alexandrov theorem for higher order mean curvatures, Differential geometry Pitman Monogr. Surveys Pure Appl. Math., vol. 52, Longman Sci. Tech., Harlow, 279–296 (1991).
  • [29] Nelli, B., Rosenberg, H.: Simply connected constant mean curvature surfaces in 2×\mathbb{H}^{2}\times\mathbb{R}. Michigan Math. J. 54, 537–543 (2006).
  • [30] Pedrosa, R.: The isoperimetric problem in spherical cylinders, Ann. Glob. Anal. and Geom. 26, 333-–354 (2004).
  • [31] Pedrosa, R., Ritoré, M.: Isoperimetric domains in the Riemannian product of a circle with a simply connected space form and applications to free boundary problems. Indiana Univ. Math. J., 48 (4), 1357–1394 (1999).
  • [32] Protter, M.: Unique continuation for elliptic equations, Trans. Amer. Math. Soc., 95, 81–91 (1960).
  • [33] Reignier, J.: Singularities of ordinary linear differential equations and integrability. In: Conte R. (eds) The Painlevé Property. CRM Series in Mathematical Physics. Springer, New York, NY (1999).
  • [34] Rosenberg, H.: Minimal surfaces in M2×.M^{2}\times\mathbb{R}. Illinois J. Math. 46, 1177–1195. (2002).
  • [35] Teschl, G.: Ordinary differential equations and dynamical systems. AMS (2012).
  • [36] Tojeiro, R.: On a class of hypersurfaces in 𝕊n×\mathbb{S}^{n}\times\mathbb{R} and n×.\mathbb{H}^{n}\times\mathbb{R}. Bull. Braz. Math. Soc., New Series 41 (2), 199–209 (2010).