Hypersurfaces of Constant Higher Order
Mean Curvature in
Abstract.
We consider hypersurfaces of products with constant -th mean curvature (to be called -hypersurfaces), where is an arbitrary Riemannian -manifold. We develop a general method for constructing them, and employ it to produce many examples for a variety of manifolds including all simply connected space forms and the hyperbolic spaces (rank symmetric spaces of noncompact type). We construct and classify complete rotational -hypersurfaces in and in as well. They include spheres, Delaunay-type annuli and, in the case of entire graphs. We also construct and classify complete -hypersurfaces of 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 -hypersurface of or is a rotational embedded sphere. Other uniqueness results for complete -hypersurfaces of these ambient spaces are obtained.
Key words and phrases:
higher order mean curvature – -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 where is an arbitrary Riemannian -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 is a simply connected space form.
Following this path, we approach here hypersurfaces of with constant -th mean curvature (for some ), which we call -hypersurfaces. Let us recall that the (non normalized) -th mean curvature of a hypersurface is the -th elementary symmetric polynomial of its principal curvatures, so that it constitutes a natural extension of the mean curvature () and the Gauss-Kronecker curvature ().
We focus on constructing and classifying -hypersurfaces of products To this end, we use a special type of graph built on families of parallel hypersurfaces of In fact, for any given constant we obtain -graphs in for those Riemannian manifolds 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 admit isoparametric hypersurfaces, such as space forms, hyperbolic spaces, warped products, and homogeneous spaces.
By suitably “gluing” pieces of -graphs, we construct properly embedded -hypersurfaces in when is either the standard -sphere or one of the hyperbolic spaces (rank symmetric spaces of non compact type). In this setting, we show that there exists a rotational -sphere in if and only if where the constant is defined as the limit of the -th mean curvature of a geodesic sphere of as its radius goes to infinity. (In particular, is positive for and vanishes for .) On the other hand, as we also show, for any and any constant there exists a rotational -sphere in
We remark that rotational hypersurfaces of a general product are defined here as those which are foliated by horizontal geodesic spheres centered at an “axis” (See Section 4 for more details.)
We provide other examples of properly embedded rotational -hypersurfaces in and in as well, including Delaunay-type annuli and, in the case of , entire graphs over Then, we classify those complete connected rotational -hypersurfaces of these product spaces whose height functions are Morse-type (i.e., have isolated critical points), which include all the properly embedded rotational -hypersurfaces we obtain here.
We also construct and classify complete connected -hypersurfaces of 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 is considered.
Our methods work equally well for -hypersurfaces with the so called -minimal hypersurfaces. By applying them, we obtain a one-parameter family of rotational, properly embedded catenoid-type -minimal -annuli in Similarly, we obtain a one-parameter family of rotational, properly embedded Delaunay-type -minimal -annuli in Then, we show that these annuli are the only complete connected -minimal rotational hypersurfaces of these product spaces (besides horizontal hyperplanes and, in the case cylinders over geodesic spheres).
Analogously to the case of -hypersurfaces, we construct and classify the complete connected -minimal hypersurfaces of which are invariant by either parabolic isometries or hyperbolic translations.
The study of -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 -hypersurfaces in or an open hemisphere of are geodesic spheres.
In [18], this result was extended to the context of -hypersurfaces of where the geodesic spheres in the statement are replaced by rotational spheres.
Here, we establish uniqueness results for rotational -spheres of and , The case was settled in [1] (for ) and in [20] (for ). More precisely, we show that any compact connected strictly convex -hypersurface of or is necessarily an embedded rotational sphere. Assuming complete, instead of compact, the same conclusion holds if, in addition, the height function of has a critical point and, in the case , the least principal curvature of is bounded away from zero. Finally, we show that, for any connected, properly immersed, and strictly convex -hypersurface of is necessarily an embedded rotational -sphere.
It is worth mentioning that these uniqueness results constitute applications of the main theorems in [11], which concern convexity properties of hypersurfaces in being either a Hadamard manifold or the sphere Besides, the noncompact cases are based on height estimates we establish here for strictly convex vertical graphs in arbitrary products
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 and of the Maximum-Continuation Principle for -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 -hypersurfaces (resp. -minimal hypersurfaces) in and 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 , be an oriented hypersurface of a Riemannian manifold Set for the Levi-Civita connection of for the unit normal field of and for its shape operator with respect to so that
where stand for the tangent bundle of . The principal curvatures of that is, the eigenvalues of the shape operator will be denoted by .
Given an integer we define the (non normalized) -th mean curvature of the hypersurface as:
(1) |
Notice that and are the non normalized mean curvature and Gauss-Kronecker curvature functions of respectively, i.e.,
Definition 1.
With the above notation, given a constant we say that is an -hypersurface of if its -th mean curvature is constant and equal to In the case , we say that is an -minimal hypersurface of
Definition 2.
A hypersurface is said to be convex at if, at this point, all the nonzero principal curvatures have the same sign. If, in addition, none of these principal curvatures is zero, then is said to be strictly convex at We call 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 — where is some Riemannian manifold — endowed with the standard product metric:
In this setting, we denote the gradient of the projection of by , which is easily seen to be a parallel field on
Let be a hypersurface of . Its height function and its angle function are defined by the following identities:
A critical point of is called horizontal, whereas a point on which vanishes is called vertical. Notice that is horizontal if and only if
We shall denote the gradient field and the Hessian of a function on by and respectively. It is easily checked that
(2) |
From (2), for all one has Hence,
(3) |
Given the set is called a horizontal hyperplane of Horizontal hyperplanes are all isometric to and totally geodesic in In this context, we call a transversal intersection a horizontal section of Any horizontal section is a hypersurface of . So, at any point the tangent space of at splits as the orthogonal sum
(4) |
We will adopt the notation for the simply connected space form of constant sectional curvature the Euclidean space (), the unit sphere (), and the hyperbolic space ().
2.1. The hyperbolic spaces
Many of our results in this paper involve the Riemannian manifolds known as hyperbolic spaces, which include the canonical (real) -dimensional hyperbolic space These manifolds are precisely the rank symmetric spaces of non-compact type, which can be described through the four normed division algebras: (real numbers), (complex numbers), (quaternions) and (octonions). They are denoted by
and called real hyperbolic space, complex hyperbolic space, quaternionic hyperbolic space and Cayley hyperbolic plane, respectively.
We will adopt the unified notation for the hyperbolic spaces, where for The real dimension of is In particular, has dimension We will keep the standard notation for the real hyperbolic space and assume
Denoting by the norm in , and taking in standard coordinates we have that is modeled by the unit ball
equipped with the Hermitian form whose coefficients in these coordinates are
where the bar denotes conjugation.
For the metric defined by correspond to the Klein model for Also, for , reduces to the canonical Poincaré metric of . In particular, is isometric to , and is isometric to . (See [8] for a detailed discussion on the hyperbolic spaces , )
The description of a model for the Cayley hyperbolic plane 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 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 is a Hadamard-Einstein manifold with nonconstant (except for ) sectional curvatures pinched between and (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 -th mean curvature , it was shown in [9, Proposition 3.2] that a hypersurface of a Riemannian manifold with 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 and coincide in a subdomain then they coincide in These facts, together with [21, Theorem 1.1], give the following result.
Maximum-Continuation Principle.
Let be complete connected -hypersurfaces of a Riemannian manifold which are tangent at a point . Let be the common unit normal and assume that is strictly convex at one of its points. Then, if (near ) remains above with respect to one has .
3. -Graphs on Parallel Hypersurfaces
In this section, we give a detailed description of graphs in which are built on families of parallel hypersurfaces of As we mentioned before, they will constitute our main tool for constructing -hypersurfaces in product spaces
With this purpose, consider an isometric immersion
between two Riemannian manifolds and and suppose that there is a neighborhood of in without focal points of that is, the restriction of the normal exponential map to is a diffeomorphism onto its image. In this case, denoting by the unit normal field of there is an open interval such that, for all the curve
(5) |
is a well defined geodesic of without conjugate points. Thus, for all
is an immersion of into which is said to be parallel to Observe that, given , the tangent space of at is the parallel transport of along from to We also remark that, with the induced metric, the unit normal of at is given by
Definition 3.
Let be an increasing diffeomorphism, i.e., With the above notation, we call the set
(6) |
the graph determined by and or -graph, for short.
Notice that, for a given -graph , and for any , is the projection on of the horizontal section that is, these sets are the level hypersurfaces of
For an arbitrary point of such a graph one has
So, a unit normal to is
(7) |
In particular, its angle function is
(8) |
A key property of -graphs is that the trajectories of on them are lines of curvature, that is, is one of its principal directions. (Notice that, by (8), so never vanishes on an -graph.) More precisely (cf. [12, 36]),
(9) |
We point out that, besides being lines of curvature, the trajectories of on an -graph , when properly reparametrized, are also geodesics. This follows from the fact that , and consequently is constant along the horizontal sections of (see [36, Lemma 5]). It should also be noticed that these trajectories project on the geodesics given by (5) (Fig. 1).

Let us compute now the principal curvatures of an -graph For that, let be the orthonormal frame of principal directions of in which In this case, for the fields are all horizontal, that is, tangent to (cf. (4)). Therefore, setting
(10) |
and considering (7), we have, for all that
where is the -th principal curvature of Also, it follows from (9) that Thus, the array of principal curvatures of the -graph is
(11) |
Now, considering the above identities and writing, for
we have that the -th mean curvature of the -graph is
(12) |
where denotes the -th mean curvature of
Due to equality (12), the function defined in (10) — to be called the -function of the -graph — will play a major role in the sequel. We remark that, up to a constant, the -function of determines its -function. Indeed, it follows from equality (10) that
(13) |
We introduce now a special type of family of parallel hypersurfaces which will be used for constructing -hypersurfaces in
Definition 4.
Following [5], we call a parallel family isoparametric if, for each any principal curvature of is constant (depending on ). If so, each hypersurface 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 admit hypersurfaces which are isoparametric in this sense, and non isoparametric as we defined.
Let be an -graph such that the parallel family is isoparametric. Then, for any the -th mean curvature of is a function of alone, which we assume to be no vanishing. In this setting, writing , and considering (12) with constant, we obtain the following result, which turns out to be our main lemma.
Lemma 1.
Let be an isoparametric family of hypersurfaces whose -mean curvatures never vanish. Given and let be a solution of the first order differential equation
(14) |
where the coefficients and are the functions
(15) |
Then, if the -graph with -function is an -hypersurface of the product Conversely, if an -graph has constant -th mean curvature , then is a solution of (14).
It follows from Lemma 1 that, as long as admits isoparametric hypersurfaces with non vanishing -th mean curvature, for any there exist -graphs in . (Notice that, the interval and the constant in (16) can be chosen in such a way that the corresponding solution of (14) satisfies .) This includes, as a trivial case, the Euclidean space In the next examples, we shall consider other manifolds to which Lemma 1 applies.
Example 1 (sphere ).
It is a well known fact that isoparametric hypersurfaces in are abundant and include all its geodesic spheres. In fact, the classification of the isoparametric hypersurfaces of is a long stand open problem (see, e.g., [15]).
Example 2 (warped products).
Let be a warped product, where the basis is an open interval and the fiber is an arbitrary -dimensional manifold. For each define as the standard immersion It is well known that, with the induced metric, is a parallel family of totally umbilical hypersurfaces of with constant principal curvatures (see, e.g., [6]). In particular, is isoparametric. Hence, if never vanishes, Lemma 1 applies to
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 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 -th curvatures in Damek-Ricci harmonic spaces.
Example 4 (-spaces).
In [16], it was proved that there exist isoparametric families of parallel surfaces with nonzero constant principal curvatures in spaces satisfying (Those include the products and , the Heisenberg space , the Berger spheres, and the universal cover of the special linear group ).
In the next two sections we construct properly embedded -hypersurfaces in products by suitably “gluing” -graphs. To this task, the following elementary fact will be considerably helpful.
Lemma 2.
Let be a differentiable function which satisfies:
Assume that one (or both) of the following hold:
-
i)
and
-
ii)
and
Under these conditions, there exists such that the improper integral
(17) |
is convergent if (i) occurs. Analogously, the improper integral
is convergent if (ii) occurs.
Proof.
Assume that (i) occurs. In this case, there exist positive constants, and such that Therefore, since
which proves the convergence of the integral (17). The case (ii) is analogous. ∎
4. Rotational -hypersurfaces of and
Rotational hypersurfaces in simply connected space forms or products are among the most classical hypersurfaces of these spaces. In the case of they are obtained by rotating (with the aid of the group of isometries of ) a plane curve about an axis . Consequently, any connected component of any horizontal section of a rotational hypersurface in lies in a geodesic sphere of with center at the axis. This fact suggests the following definition.
Definition 5.
A hypersurface is called rotational, if there exists a fixed point such that any connected component of any horizontal section is contained in a geodesic sphere of with center at If so, the set is called the axis of In particular, any horizontal hyperplane is a rotational hypersurface of with axis at any point
In what follows, we construct and classify complete rotational -hypersurfaces in and
4.1. Rotational -hypersurfaces of
Let us consider a family
(18) |
of isoparametric concentric geodesic spheres of indexed by their radiuses, that is, for a fixed , and each is the geodesic sphere of with center at and radius
We remark that any sphere is strictly convex. Also, in accordance to the notation of Section 3, for each we choose the outward orientation of , so that any principal curvature of is negative. In this setting, the function is positive for even and negative for odd. Hence, for any constant the coefficients and in (15) are given by
(19) |
The principal curvatures of the geodesic spheres are ():
(20) |
where for , for , for , and for (see, e.g., [7, pgs. 353, 543] and [25]).
From equalities (20), we obtain the -mean curvatures of the geodesic spheres of For instance, in we have
(21) |
whereas for one has
if , and
(23) |
Analogously, one obtains the -th mean curvature functions for the other hyperbolic spaces. A direct computation from this data yields the following
Lemma 3.
The functions e defined in (19) have the following properties:
-
i)
is negative and increasing for , and vanishes for
-
ii)
is positive and increasing for , and for
In particular, we have the inequalities
(24) |
We point out that, in the above setting, one has (cf. [27])
(25) |
in a neighborhood of In particular,
(26) |
In what follows, by means of the family we will construct complete rotational -hypersurfaces in which are made of pieces of -graphs. With this purpose, we will look for solutions of the equation (with and as in (19)) satisfying suitable initial conditions. Let us recall that, in this context, the general solution of is
(27) |
Concerning the solutions , we will be also interested in those which can be extended to Notice that, in principle, neither nor are defined at , which makes this point a singularity. However, the function is easily extendable to Indeed, we can set if , and if (from (26)). As for it follows from (25) that, for
(28) |
which characterizes as a regular singular point of This means that, despite the fact that is not defined at this equation has a nonnegative solution defined at that satisfies (cf. [33, Theorem 3.1], [35, Lemma 4.4]). More precisely, this solution is
(29) |
where is a solution of Notice that the function defined in (29) is also the solution of in the case i.e., for (Just set )
As we shall see, the geometry of the -hypersurfaces we construct from -graphs is closely related to the growth of as Taking that into account, for a given family of parallel geodesic spheres in we define
(30) |
In particular, Notice that, since is homogeneous, the constant is well defined, that is, it does not depend on the family of geodesic spheres.
Similarly, one can compute the other constants and easily conclude that
The next proposition shows the relation between the solution of and the constant Notice that, for the identities (19) yield:
Proposition 1.
Proof.
To prove (i), let us observe first that, since the solution in (29) is positive in and we have that is increasing near Assume that is not increasing in In this case, has a first critical point in which is necessarily a local maximum. However, considering (24) and the equality we have
which implies that is a local minimum — a contradiction. Therefore, is increasing in which proves (i).
Now, assume We claim that, in this case, as . Indeed, for any fixed and ,
Since is decreasing, this inequality gives that in Hence,
which clearly implies the claim on
The above proposition immediately gives the following result.
Corollary 1.


Now, we are in position to establish our first existence result.
Theorem 1.
Given and a constant the following hold:
-
i)
If there exists an embedded strictly convex rotational -sphere in which is symmetric with respect to a horizontal hyperplane.
-
ii)
If there exists an entire strictly convex rotational -graph in which is tangent to at a single point, and whose height function is unbounded above. Consequently, there are no compact -hypersurfaces of for such values of .
Proof.
Let be an arbitrary family of parallel geodesic spheres of as in (18). Consider the functions and defined in (19) and let be the solution (29) of the ODE
If we have from Corollary 1 that there exists satisfying
Hence, by Lemma 1, the -graph with -function is a rotational -graph of over the open ball such that
(32) |
From Proposition 1-(i), one has which implies that Hence, by Lemma 2, extends to , i.e.,
is well defined. In particular,

Notice that is an isolated minimum of the height function of Thus, is strictly convex at In addition, by the identities (11), at any point of all the principal curvatures are positive. Therefore, is strictly convex.
As we know, the angle function of is given by
(33) |
Since , we have from (32) that as . This, together with (33), implies that the tangent spaces of along are vertical. Hence, the trajectories of all emanate from and meet orthogonally (Fig. 3). Recall that these trajectories are geodesics which foliate
Now, set for the reflection of with respect to and define
that is, is the “gluing” of and along the -sphere which is -differentiable. Since the trajectories of are also for being geodesics, the resulting hypersurface is -differentiable with vertical tangent spaces along Therefore, is a rotational strictly convex -hypersurface of which is homeomorphic to and is symmetric with respect to This proves (i).
Under the hypotheses in (ii), it follows from Corollary 1 that satisfies:
so that the -graph with -function is an entire rotational -graph of over Since and for any is contained in the closed half-space and is tangent to at Also, the height function of is unbounded above. Indeed, from Proposition 1-(i), , and so , is increasing. Thus, for a fixed and any one has
which implies that is unbounded above. Also, arguing as for the graph in the preceding paragraphs, we conclude that is strictly convex.
Observe that the mean curvature vector of “points upwards”, that is, its mean convex side is the connected component of which contains the axis In particular, is foliated by the balls

Let us suppose that there exists a compact -hypersurface such that Considering the fact that is “horizontally and vertically unbounded”, it is easily seen that, after a suitable vertical translation, we can assume (Fig. 4). Now, translate downward until it has a first contact with Since is strictly convex, the Maximum-Continuation Principle applies and gives that and coincide, which is clearly impossible. This shows that such a cannot exist and finishes the proof of (ii). ∎
Remark 1.
Let be an entire -graph as in Theorem 1-(ii). Since , we must have Also, the associated function is positive, bounded and increasing, so that
which implies that as This, together with (11), gives that the principal curvature goes to zero as In particular, the least principal curvature function of is not bounded away from zero.
Next, we apply the method of -graphs to produce one-parameter families of -annuli in For that, fix and consider the solution of given in (27), which satisfies the initial condition From (19), we have
(34) |
Therefore, given it follows from (34) that if and only if Thus, considering (26), we can define as the largest positive constant with the following property:
(35) |
Observing that is a positive function for any we distinguish the cases:


Theorem 2.
Given and let be as in (35). Then there exists a one-parameter family
of properly embedded rotational -hypersurfaces in which are all homeomorphic to the -annulus In addition, the following assertions hold:
-
i)
If each 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 , each hypersurface is symmetric with respect to and has unbounded height function.
Proof.
Fix and recall that is decreasing near Thus, if it follows from (31) that there exists such that
Let us observe that a critical point of is necessarily a minimum, since Therefore, must have a unique local minimum at a point between and In particular, (Fig. 5a).

Setting , it follows from the above considerations and Lemmas 1 and 2 that the -graph with -function is a bounded -hypersurface of Moreover, is homeomorphic to and has boundary (see Fig. 6):
We also have that the tangent spaces of are vertical along its boundary , for Therefore, we obtain a properly embedded rotational -hypersurface from by continuously reflecting it with respect to the horizontal hyperplanes This proves (i).
Now, let us suppose that In this case, (31) gives that
so that the -graph determined by is an -hypersurface of with boundary (Fig. 7). By reflecting with respect to as we did before, we obtain the embedded -hypersurface as stated.
It remains to show that the height function of is unbounded. For that, we have just to observe that the infimum of in is positive, since itself is positive in this interval, and its limit as is So, the same is true for Therefore,
from which we conclude that is unbounded. ∎
Remark 2.

We proceed now to the classification of the complete rotational -hypersurfaces of 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 there exist complete rotational -hypersurfaces in which are not embedded. In accordance with our results, the height function of none of these -hypersurfaces is Morse-type.
Firstly, let us recall that the -hypersurfaces in Theorems 1 and 2 were constructed from a single -graph whose associated -function is a solution of the ODE where and are as in (19). For such a , there is a maximal interval , such that
Notice that each choice of determines the function and, so, the equation The corresponding graph, then, is determined by the ordering of the constants and as well as by the values of and
Below, we list all the occurrences of and in Theorems 1 and 2 with respect to the ordering of and :
-
C1)
, , .
-
C2)
, , .
-
C3)
, , .
-
C4)
, , .
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 in cases C1 and C3, with and that in cases C2 and C4.
Let be a hypersurface of a Riemannian manifold It is easily seen that is a hypersurface of whose tangent spaces are all vertical, so that is a principal direction of with vanishing principal curvature. In particular, on Also, for all the -th mean curvatures of and at and coincide. In particular, is an -hypersurface of if and only if is an -hypersurface of We call the cylinder over
Theorem 3.
Proof.
Suppose that is not a cylinder. In this case, we have that the open set on which never vanishes is nonempty. Since contains no vertical points, for a given , there is an open neighborhood of in which is a graph over an open set of . Thus, since is rotational and contains no horizontal points, after possibly a reflection with respect to a horizontal hyperplane, we can assume that is an -graph over (Recall that, in our setting, the -function of an -graph is required to be radially increasing.)
By Lemma 1, the function associated to is a solution of with and as in (19). In addition, since is complete, there exists a maximal interval , such that In particular, we have the following two possibilities:
Suppose that After a vertical translation, we can assume that
In particular, If these equalities imply that the sphere of is contained in and that vanishes at all of its points. This, however, contradicts that the height function of is Morse-type. Hence, so that the -function of satisfies the initial condition
If by the uniqueness of solutions of linear ODE’s satisfying an initial condition, the function such that coincides with the one in case C1 above. Thus, the corresponding -functions also coincide, which clearly implies that is an open set of the (strictly convex) -sphere obtained in Theorem 1-(i). Therefore, by the Maximum-Continuation Principle, coincides with this -sphere. If then coincides with the solution of case C2. Analogously, we conclude that is an entire graph as in Theorem 1-(ii).
4.2. Rotational -hypersurfaces of
In this section, we apply the method of -graphs to construct and classify rotational -hypersurfaces in
As we did before, let us fix a point and consider a family
(36) |
of parallel geodesic spheres of with radius and center As is well known, each is totally umbilical, having principal curvatures all equal to with respect to the outward orientation. In particular, is isoparametric.
From a direct computation, we get that the coefficients and of the ODE determined by and any given are
(37) |
and that the corresponding general solution is:
(38) |
where and
Also, it is easily checked that
(39) |
is a well defined solution of satisfying
Given an integer it will be convenient to introduce the following constant:
(40) |
Proposition 2.
Let be the solution (39). Then, the following hold:
-
i)
in
-
ii)
-
iii)
if
Proof.
Since the functions and 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 In this case, since and the integral is positive, we have that satisfies (ii) for
If , for a fixed and any one has
which implies that the first integral goes to infinity as since the same is true for the integral . It follows from this fact that as if which proves (ii).
The identity in (iii) follows directly from the definitions of (for ) and (as in (40)). ∎
From the above proposition, we get the following existence result for -hypersurfaces of
Theorem 4.
Given and a constant there exists an -sphere in which is symmetric with respect to a horizontal hyperplane. Furthermore:
-
i)
is strictly convex if or and
-
ii)
is convex if and
-
iii)
is non convex if and
Proof.
Let be an arbitrary family of parallel geodesic spheres of as given in (36). Consider the functions and defined in (37) and let be the solution (39) of the ODE
From Proposition 2-(ii), there exists a positive such that
so that determines an -graph over Since and (by Proposition 2-(i)), we can proceed just as in the proof of Theorem 1-(i) to obtain from the embedded -sphere of which is symmetric with respect to
If or and we have from Proposition 2, items (ii) and (iii), that Hence, for all spheres have negative principal curvatures, which, together with equalities (11), gives that is strictly convex. This proves (i).
If and Proposition 2-(iii) yields However, is totally geodesic in which implies that, except for the principal curvatures of vanish at all points of the horizontal section Therefore, is convex on and strictly convex on .
Finally, assuming and we have from Proposition 2-(iii) that Observing that, for has positive principal curvatures, we conclude, as in the last paragraph, that is strictly convex (resp. convex, non convex) on if (resp. ). In particular, is non convex. This shows (iii) and concludes our proof. ∎
Remark 3.
In our next theorem we show the existence of one-parameter families of rotational Delaunay-type -annuli in This result, then, generalizes the analogous one obtained in [31] for
First, let us introduce the constant
and observe that, for the positive constants , (as in (37)), and satisfy the following relation:
(41) |
In this setting, if we define
(42) |
then a solution of such that , satisfies:
(43) |
Theorem 5.
Given , and there exists a one-parameter family of properly embedded Delaunay-type rotational -hypersurfaces in .
Proof.
Given consider the solution as in (38) such that and From (43), we have that is decreasing in a neighborhood of
Observe that is positive in Also, setting
we have that on So, for and
which implies that as
If since and is positive, we have that as
It follows from the above considerations that there exists such that
(44) |
From this point on, the proof is entirely analogous to that of Theorem 2-(i). ∎
A classification result for rotational -hypersurfaces of can be achieved in the same way we did for their congeners in To see this, assume that is a complete connected rotational -hypersurface of whose height function is Morse-type. Assuming that is non cylindrical, we have, as before, that there exists an open set which is an -graph, The corresponding -function, restricted to a maximal interval satisfies:
which yields or
If then (Otherwise, the height function of would not be Morse-type.) In this case, coincides with the -function of the -sphere of Theorem 4, and then itself coincides with this sphere. (Notice that any of the spheres obtained in Theorem 4 is strictly convex on an open set.)
If then is decreasing in a neighborhood of . Thus, In particular, as so that Analogously, this gives that coincides with the -annulus of Theorem 5,
Summarizing, we have the following result.
Theorem 6.
5. Rotational -minimal Hypersurfaces of and
In this section, we shall see that the method of -graphs can be used for construction and classification of rotational -minimal hypersurfaces of and A major distinction from the case of -hypersurfaces is that the Maximum-Continuation Principle is no longer available.
Theorem 7.
Given there exists a one-parameter family
of complete rotational -minimal -annuli in with the following properties:
-
i)
If is a cylinder over a geodesic sphere of of radius
-
ii)
If , is catenoid-type. More precisely, it is symmetric with respect to and is the geodesic sphere of of radius centered at the point of the axis. In addition, each of the parts of above and below is a rotational graph over (Fig. 8).
Furthermore, up to ambient isometries, any complete connected rotational -minimal hypersurface of is either a member of or a horizontal hyperplane.
Proof.
Given is immediate that a cylinder over a geodesic sphere of of radius is an -minimal rotational annulus of which yields (i).
Assume that and let be the parallel family of geodesic spheres of centered at the axis point The ODE determined by and is given by
(45) |
Since given the function
is clearly a solution of (45) which satisfies
In addition, So, setting it follows from Lemma 2 that
is well defined. Therefore, by Lemma 1, the -graph is an -minimal hypersurface of Notice that is a graph over with boundary (Fig. 8).

Also, since the tangent spaces of along are all vertical. Thus, considering the reflection of with respect to as before, we have that is the desired -minimal hypersurface.
Suppose now that is a complete connected rotational -minimal hypersurface of , and set
Notice that is either a horizontal hyperplane or a cylinder if and only if So, we can assume We can also assume, without loss of generality, that and all members of share the same axis
As we argued in previous proofs, under the above hypotheses, there exists an -graph and a maximal open interval such that the -function of satisfies
For we have that and so , is constant. Hence, up to a vertical translation, one has for some constant However, which implies that the closure of in meets the rotation axis non orthogonally, i.e., is not smooth at — a contradiction. So, if
For we have that is a solution of (45). In particular, is decreasing, which implies that As before, this yields Thus, setting we have , which implies that, up to a vertical translation, and, then, coincides with the half-catenoid
We conclude from the above that is the union of open half-catenoids , where In addition, has empty interior in Otherwise, there would exist a nonempty open set of which would be either horizontal or vertical, and whose boundary in would be contained in some half-catenoid. Since catenoids have no horizontal points, should be vertical and, so, part of a vertical rotational cylinder. However, rotational cylinders in are -minimal if and only if Therefore, has empty interior, which implies that is open and dense in Clearly, the intersection of two distinct members of is always transversal. This, together with the connectedness of and the density of in gives that coincides with some , which concludes our proof. ∎
Theorem 8.
Given there exists a one-parameter family
of complete rotational -minimal -annuli in with the following properties:
-
i)
If then and is a cylinder over a geodesic sphere of of radius
-
ii)
If , then and is Delaunay-type.
Furthermore, up to ambient isometries, any complete connected rotational -minimal hypersurface of is either a member of or a horizontal hyperplane.
Proof.
Statement (i) is trivial. So, assume and let be the family of parallel geodesic spheres of centered at some point In this setting, the ODE determined by and is
(46) |
Henceforth, the reasoning in the proof of Theorem 5 applies and leads to the construction of the Delaunay-type -minimal hypersurface as stated in (ii).
Now, suppose that is a complete connected rotational -minimal hypersurface of In this setting, define
As in the preceding proof, if Thus, in this case, is either a horizontal hyperplane or a cylinder over a geodesic sphere of
Suppose that and that the axis of is If then is neither a horizontal hyperplane nor a cylinder. In addition, there exists an -graph and a maximal interval , such that the -function of satisfies . So, or
The formula of the general solution of the ODE (46) gives that is positive, not defined at , and bounded away from zero. In particular, and so that is given by
It is clear from this last equality and the considerations preceding it that which implies that coincides with , . Therefore, coincides with the “block” that generates (see Fig. 6), so that is a union of open sets of members of
Let be an open set of If is nonempty, it cannot be horizontal, for no member of has horizontal points. If is vertical, then it is part of the totally geodesic cylinder In this case, a boundary point of is vertical and lies on a geodesic sphere centered at the axis and of radius However, such a boundary point also lies on some which contradicts the fact that any vertical point of lies on a geodesic sphere of radius different from (In fact, these vertical points are on geodesic spheres of radiuses and )
We conclude from the above that is open and dense in Since is connected and two distinct members of are never tangent, it follows that, for some coincides with ∎
Remark 5.
6. Translational -hypersurfaces of
Given a Hadamard manifold recall that the Busemann function of corresponding to an arclength geodesic is defined as
The level sets of a Busemann function are called horospheres of In this setting, as is well known, is a parallel family which foliates Furthermore, any horosphere is homeomorphic to , and any geodesic of which is asymptotic to — i.e., with the same point on the asymptotic boundary of — is orthogonal to each horosphere . In this case, we say that the horospheres are centered at .

Therefore, in what concerns its horospheres, a Hadamard manifold can be pictured just as the Poincaré ball model of hyperbolic space where the horospheres centered at a point are the Euclidean -spheres in which are tangent to at (Fig. 9).
In the real hyperbolic space any horosphere is totally umbilical with constant principal curvatures equal to As shown in [5, Proposition-(vi), pg. 88], any horosphere of has principal curvatures and with multiplicities and respectively. Therefore, any family of parallel horospheres of is isoparametric and its members are pairwise congruent. In addition, for any integer all horospheres of have the same (positive) -th mean curvature, which we denote by
Theorem 9.
Let be a family of parallel horospheres in hyperbolic space Then, for any even integer and any constant there exists a properly embedded, everywhere non convex -hypersurface in which is homeomorphic to Furthermore, is foliated by horospheres, is symmetric with respect to the horizontal hyperplane , and its height function is unbounded above and below.
Proof.
For each , consider the isometric immersion such that . Since all the principal curvatures of are constant and independent of the coefficients e of the ODE associated to this family are constants. Also, since is even and , we have
In this setting, consider the solution of :
(47) |
and observe that it satisfies:
(48) |

By Lemma 1, the -graph with is an -hypersurface of The function in this case, is given by
Notice that, by (48), one has
so that Thus, by Lemma 2, is well defined. Also, is negative on and is unbounded. Indeed, for all
which implies that is unbounded, since as
Denoting by the horoball of with boundary , it follows from the above considerations that is an -graph over which is unbounded and has boundary (Fig. 10). In particular, is homeomorphic to . Furthermore, is everywhere non convex, since, from the identities (11), one has
where is the (positive constant) -th principal curvature of .
Finally, since as in the previous theorems, we have that any trajectory of on meets orthogonally. Consequently, setting for the reflection of with respect to and defining
we have that is a properly embedded -hypersurface of which is foliated by horospheres and is homeomorphic to (Fig. 11), as we wished to prove. ∎

Our next result establishes that the conditions on the parity of and on the sign of in Theorem 9 are necessary to the conclusion.
Theorem 10.
Let be a family of parallel horospheres in Assume that, for some is a complete connected -hypersurface of with no horizontal points, and that each connected component of any horizontal section is a (vertically translated) horosphere of Under these conditions, one has Assume, in addition, that either of the following assertions holds:
-
i)
is even and
-
ii)
is odd.
Then, for some In particular,
Proof.
Let be the open set of points satisfying Our aim is to prove that is empty. Assuming otherwise, choose . Since has no horizontal points, we can suppose (after possibly a reflection about a horizontal hyperplane) that there is an open neighborhood of which is an -graph,
The -function associated to satisfies where and are the (constant) functions (15) determined by and . Also, there is a maximal interval , such that
Let us suppose that In this case, we have which gives that is, In particular, and , and is increasing (if ), or decreasing (if ) in So, vanishes in or . Assuming the former, we have that is defined at and Thus, for any the point is horizontal, contrary to our assumption. Therefore, if then which implies that for some But this contradicts the assumed positiveness of . Hence, we must have
Let us assume now that (i) holds. Then, we have Also, on
that is, is decreasing and concave in which clearly implies that and As in the preceding paragraph, this leads to the existence of a horizontal point of Therefore, if (i) holds, which implies that for some
Finally, let us assume that (ii) holds. In this case, one has which gives that is increasing and convex. From this point, we get easily to the conclusion by reasoning just as in the last paragraph. ∎
An isometry of which fixes only one point is called parabolic. Such isometries have the following fundamental property: The horospheres of centered at are invariant by parabolic isometries of that fix (cf. [17, Proposition 7.8]).
We point out that any isometry of has a natural extension to an isometry of Namely,
We call parabolic if is parabolic. More specifically, if
is the family of parallel horospheres which are invariant by we say that and are -parabolic isometries.
In the upper half-space model of , 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 of parallel horospheres in if a hypersurface of is invariant by -parabolic isometries of , then any connected component of any horizontal section is contained in a (vertically translated) horosphere of
Now, we are in position to classify all complete connected -hypersurfaces of with no horizontal points which are invariant by parabolic isometries.
Theorem 11.
Let be a family of parallel horospheres of Assume that is a complete connected -hypersurface of , with no horizontal points, which is invariant by -parabolic isometries. Then, up to ambient isometries, is either a cylinder over a horosphere of or the embedded -hypersurface obtained in Theorem 9.
Proof.
Assume that is not a cylinder over a horosphere of By Theorem 10, we have that is even and In this case, the open set
is dense in Indeed, the -th mean curvature of any nonempty open -invariant vertical set would be Thus, given there exists an -graph in with In addition, the associated -function, restricted to a maximal interval satisfies
We have that is a solution of the ODE determined by and The conditions on the parity of and the sign of , as in the proof of Theorem 9, give that is increasing and convex, which implies that and that
Consider the solution (47) of and denote it by Since and the coefficients and are constants, by the uniqueness of solutions satisfying initial conditions, we have that This, together with the homogeneity of the horospheres of implies that coincides with the -graph determined by From this fact and the density of in we conclude that coincides with the embedded -hypersurface obtained in Theorem 9, as we wished to prove. ∎
Given a totally geodesic hyperplane of let us recall that there exists a family of parallel hypersurfaces of such that the distance of any point of to is The family foliates and each member of , which is called an equidistant hypersurface, is properly embedded and homeomorphic to (Fig. 12).

We shall also write as a family of immersions:
that is, for each is the equidistant to .
Given a geodesic orthogonal to the members of any equidistant hypersurface is totally umbilical with constant principal curvatures all equal to
with respect to the unit normal (see Section 3). In particular, is isoparametric. Also, given a constant , the coefficients and of the differential equation associated to and are:
(49) |
It will be convenient to reconsider the constant and recall that, for the following identity holds:
(50) |
Our next result establishes that, for , there exists a one-parameter family of properly embedded -hypersurfaces in which are foliated by (vertical translations of) parallel equidistant hypersurfaces of In this setting, we have so we can define:
(51) |
Theorem 12.
Given , let Then, there exists a one parameter family of properly embedded and everywhere non convex -hypersurfaces of Each member of is homeomorphic to and is foliated by equidistant hypersurfaces. Moreover, is symmetric with respect to , and its height function is unbounded.
Proof.
Let be a family of parallel equidistant hypersurfaces of Since , it follows from the relation (50) that Thus, we can choose such that
(52) |
In particular,
Let be the solution of satisfying where and are the functions in (49). Then, from (52), we have
so that is decreasing near .
We claim that is decreasing on the whole interval . To show that, it suffices to prove that has no critical points in Assuming otherwise, consider satisfying Since we have that We also have and Thus,
which implies that is necessarily a local maximum for This proves the claim, for is decreasing near so that a local maximum for should be preceded by a local minimum.
We also have that is positive in Indeed, if we had for some it would give and then would be increasing near
It follows from the above considerations that
Furthermore, since is decreasing and positive, we have that as This, together with the equalities and , gives
(53) |
Therefore, the graph associated to (see Fig. 13) is an -hypersurface of whose -function is

As in the preceding proofs, we obtain a properly embedded -hypersurface of by reflecting with respect to since and . It is also clear from equalities (11) that, except for , its principal curvatures are all positive, so that is nowhere convex.
Finally, considering (53) and the fact that is decreasing, we have
which clearly implies that the height function of is unbounded. ∎
Theorem 13.
Let be a family of equidistant hypersurfaces in Given and let be the constant defined in (51). Then, there exists a complete everywhere non convex -hypersurface in which is an -graph, Furthermore, the height function of is unbounded above and below, and is asymptotic to
Proof.
Let be the solution of the differential equation associated to and (i.e., with and as in (49)) which satisfies the initial condition
From its definition, we have that satisfies In addition,
so that is a local maximum of Reasoning as in the preceding proof, we get that , and so , is positive and decreasing in From this, we conclude analogously that as
Now, for a fixed , define
and let be the corresponding -graph. As before, we have that is nowhere convex. Denoting by the convex connected component of we also have that is a graph over (Fig. 14).
For , we have
which gives that , and so the height function of is unbounded above.
Finally, given a constant there exists such that
for Thus, for such values of one has
which implies that as , since is bounded away from zero. Therefore, the height function of is unbounded below, and is asymptotic to in as we wished to prove. ∎

A family of equidistant hypersurfaces in determines a group of translational isometries which we shall call -hyperbolic. In the upper half-space model of taking as a Euclidean half vertical hyperplane orthogonal to through the “origin” we have that the -hyperbolic isometries are the Euclidean homotheties from It should be noticed that the equidistant hypersurfaces of are all invariant by -hyperbolic isometries.
The natural extension of an -hyperbolic isometry of to will also be called -hyperbolic. If is a hypersurface of which is invariant by -hyperbolic isometries, it is clear that any connected component of any horizontal section of is contained in for some
Next, we classify -hypersurfaces of (without horizontal points or totally geodesic horizontal sections) which are invariant by hyperbolic translations.
Theorem 14.
Let be a family of parallel equidistant hypersurfaces in Assume that, for some is a complete connected -hypersurface of which is invariant by -hyperbolic translations. Assume further that has no horizontal points, and that no horizontal section of is totally geodesic in (i.e., Under these conditions, the following assertions hold:
-
i)
.
-
ii)
.
- iii)
Proof.
Set and assume Given , as in previous proofs, we can assume there is an open set which is an -graph containing . Its function satisfies , where and are the functions given in (49). Also, is defined in a maximal interval such that Since no horizontal section of is totally geodesic, we can assume
The maximality of gives that or In the former case, we have Then, is well defined (by Lemma 2) and so that is a horizontal point of , contrary to our hypothesis. Then, we must have In particular, near , is decreasing in which implies that Indeed, for
Assume now that , Then, we have
which contradicts that is decreasing near .
It follows from the above considerations that, if then and . Furthermore, a direct computation gives that if and only if . If , then coincides with the function of the -graph associated to the hypersurface of Theorem 12. From this, arguing as in preceding proofs, we conclude that By the same token, if then is the complete graph obtained in Theorem 13.
Let us suppose now that In this case, we must have where is an equidistant hypersurface with -th mean curvature , so that . Therefore, we have (since we are assuming ) and
which concludes our proof. ∎
7. Translational -minimal Hypersurfaces of
In this section, we construct and classify -minimal hypersurfaces in which are invariant by translational isometries. It will be convenient to consider first the case of hyperbolic isometries of
Theorem 15.
Let be a family of parallel equidistant hypersurfaces to a totally geodesic hyperplane of Then, for each there exists a one-parameter family of properly embedded -minimal hypersurfaces of which are all homeomorphic to and invariant by -hyperbolic translations. Each member has the following additional properties:
-
i)
For is a constant angle entire -minimal graph over whose height function is unbounded above and below.
For we distinguish the following cases:
-
ii)
is symmetric with respect to the horizontal hyperplane , and is contained in a slab
-
iii)
is an -graph () which is bounded above, unbounded below, and asymptotic to
-
iv)
is an entire graph over which is symmetric with respect to , and is contained is a slab
Furthermore, except for the cylinders (in the case ), and up to ambient isometries, the members of are the only complete non totally geodesic -minimal hypersurfaces of which are invariant by hyperbolic translations.
Proof.
The equation (14) determined by and is:
(54) |
For its solution is constant. So, given defining
we have that the corresponding -graph is an entire -minimal graph whose level hypersurfaces are the leaves of Clearly, the height function of is unbounded above and below. Moreover, it follows from (8) that is a constant angle hypersurface. This proves (i).
Assume that Then, defining one has
Hence, setting
(55) |
we have that the -graph is a well defined -minimal hypersurface, for Also, since the closure of intersects orthogonally. Thus, we obtain an -minimal hypersurface by reflecting about
As for the boundedness of , we first observe that, from the equalities and we have In addition, the function is bounded above by in Hence,
which finishes the proof of (ii).
Assuming now let us fix and define
Since we can mimic the final part of the proof of Theorem 13 and conclude that is unbounded below, and that the corresponding -graph is asymptotic to Also, proceeding as in (7), we can show that is bounded above. This proves (iii).
Given we have that in So, we can define as in (55), replacing by . Analogously, we have that is bounded above, and that the boundary of the -graph is .
Notice that is well defined and positive, since is neither nor Thus, we obtain a complete properly embedded -minimal hypersurface from by reflecting it with respect to , and then with respect to the totally geodesic vertical hyperplane (Fig. 15). This shows (iv).
Assume now that is a complete non totally geodesic -minimal hypersurface of which is invariant by -hyperbolic translations. Set
and suppose that Then, Otherwise, would be either a horizontal hyperplane or a cylinder over the hyperplane of In both cases, would be totally geodesic, which is contrary to our assumption.
Therefore, if for each , there is an -graph which contains , and whose -function is a solution of (54). More precisely, for some one has
Now, recall that in the cases (ii)–(iv) above, the corresponding function satisfies Thus, for the function coincides with which implies that In addition, no has horizontal or totally geodesic points, which gives that is open and dense in Therefore, coincides with
Finally, let us suppose that If then for some If then there exists an -graph whose -function is constant. In particular, up to a vertical translation, we have for some so that is the entire graph given in (i). ∎

Next, we obtain all complete non totally geodesic -minimal hypersurfaces of which are invariant by parabolic isometries.
Theorem 16.
Let be a family of parallel horospheres in Then, for any there exists a properly embedded -minimal hypersurface of which is invariant by -parabolic isometries. In addition, is homeomorphic to and has the following properties:
-
i)
For is a constant angle entire graph over whose height function is unbounded above and below.
-
ii)
For is symmetric with respect to and is contained in a slab
Furthermore, except for the cylinders (in the case ), and up to ambient isometries, is the only complete non totally geodesic -minimal hypersurface of which is invariant by parabolic isometries.
Proof.
The proof of the existence of as in (i) is analogous to the one given in the preceding theorem. So, let us assume In this case, the equation (14) determined by and takes the form
(57) |
and its positive solutions are
(58) |
It is easily checked that, since the horospheres of are pairwise congruent, an graph with -function does not depend on More precisely, two such graphs obtained from functions and , are isometric. Therefore, we can assume and set . Then, we have
Since writing , we have
is well defined, and so is the corresponding -graph Also, so that the tangent spaces of along its boundary are all vertical. Therefore, we obtain the stated -minimal hypersurface by reflecting about
Observe that, for all one has
Hence, setting we have that as which proves that is contained in the slab
As for the uniqueness of , notice that the following hold:
-
•
The -function of any -minimal -graph, is a positive solution of (57) (if ) or is a positive constant (if ).
-
•
has no horizontal points.
-
•
A vertical -invariant hypersurface of is -minimal if and only if
-
•
The graph in (i) has no vertical points.
These facts allow us to argue as in preceding proofs, and then show the uniqueness of as asserted. ∎
8. Uniqueness of Rotational -spheres of
In this concluding section, we concern the uniqueness of the rotational -spheres we constructed in Section 4. We restrict ourselves to , with and As we mentioned before, the case was considered in [1, 20].
We obtain a Jellett–Liebmann type theorem by showing that a compact, connected and strictly convex -hypersurface of 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 which we establish in the next proposition. First, we compute the Laplacian of both the height function and the angle function of an arbitrary hypersurface of a general product
Given a smooth function on let us denote its Laplacian by i.e.,
In particular, from equation (3), the Laplacian of is given by
(59) |
Recall that, for the Codazzi equation reads as
where is the curvature tensor of denotes the tangent component of the tangent bundle of and, by definition,
Now, let us fix and an orthonormal frame in a neighborhood of in which is geodesic at that is
Writing we have Therefore
which implies that, at the chosen point
Since is arbitrary, we get from this last equality and (60) that, on
(61) |
where denotes the Ricci curvature tensor of and
Remark 7.
For the next results, except for Theorem 19, we order the principal curvatures of a hypersurface of as
Proposition 3.
Consider an arbitrary Riemannian manifold and let be a compact vertical graph of a nonnegative function defined on a domain Assume strictly convex up to Under these conditions, the following height estimate holds:
(62) |
Proof.
Consider in the “inward” orientation, so that its angle function is non positive. Choose satisfying and define on the function
We claim that has no interior maximum. Indeed, assuming otherwise, let be a maximum point of . In this case, from (2), we have
Hence, if we had then would be an eigenvalue of at which is impossible, by our choice of Thus, is a critical point of Since is strictly convex, is necessarily its highest point. In particular, This, together with identities (59) and (61), gives that, at ,
(63) |
However, from our choice of and the strict convexity of , we have
which contradicts (63). Therefore, attains its maximum on which implies that on for Hence,
The result, then, follows from this last inequality, since it holds for any positive ∎
Remark 8.
Proposition 3 has its own importance, since it establishes height estimates for vertical graphs in making no assumptions on 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 was considered.
Theorem 17 (Jellett–Liebmann-type theorem).
Let be a compact connected strictly convex -hypersurface of (). Then, is an embedded rotational -sphere.
Proof.
Since is compact, its height function has a maximal point This, together with the strict convexity of , allows us to apply [11, Theorems 1 and 2] and conclude that is embedded and homeomorphic to . Thus, for the result follows from [18, Theorem 7.6], the Alexandrov-type theorem we mentioned in the introduction.
For we can perform Alexandrov reflections on with respect to horizontal hyperplanes coming down from above For some the reflection of the part of above will have a first contact with Then, by the Maximum-Continuation Principle, is symmetric with respect to . Therefore, assuming and identifying with we conclude that is a “bigraph” over its projection to . As a consequence, is the boundary of in .
By [11, Lemma 1], the second fundamental form of , as a hypersurface of , is positive definite. In particular, is non totally geodesic in . Thus, by [14, Theorem 1], is contained in an open hemisphere of , which implies that the same is true for that is, In this setting, we can apply Alexandrov reflections on “vertical hyperplanes” where is a totally geodesic -sphere of , and conclude that 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 In the case , we also have to impose a condition on the second fundamental form of which turns out to be a necessary hypothesis (see Remark 10, below).
Theorem 18.
Let be a complete connected strictly convex -hypersurface of () whose height function has a local extreme point. For , assume further that the least principal curvature of is bounded away from zero. Then, is an embedded rotational sphere.
Proof.
As in the previous theorem, fulfills the hypotheses of [11, Theorems 1 and 2], which implies that is properly embedded and homeomorphic to either or Furthermore, in the latter case, the height function of is unbounded and has a single extreme point which we assume to be a maximum.
For the height estimates obtained in [9, Theorem 4.1-(i)] forbid to be unbounded. Thus, in this case, is homeomorphic to and the result follows from Theorem 17.
Let us consider now the case Assume, by contradiction, that is homeomorphic to so that is unbounded below. Hence, given a horizontal hyperplane with the part of which lies above must be a vertical graph with boundary in . If not, for some between and would be orthogonal to at one of its points. Then, the Alexandrov reflection method would give that is symmetric with respect to which is impossible, since we are assuming unbounded, and the closure of in is compact.
Remark 9.
In Theorems 17 and 18, the hypothesis of strict convexity of is automatically satisfied for so it can be dropped in this case. Indeed, in both theorems, the height function has a critical point , which can be assumed to be a maximum. Then, taking the inward orientation on we have that which, together with equality (3), yields
However, on . Thus, at and then on all of , the second fundamental form is positive definite, that is, is strictly convex.
Remark 10.
Next, we consider the dual case of Theorem 18, assuming now that the height function of the hypersurface has no critical points. First, we recall that a hypersurface is said to be cylindrically bounded, if there exists a closed geodesic ball such that
Theorem 19.
Let be a proper, convex, connected -hypersurface of () with no horizontal points. For assume further that is cylindrically bounded. Then, is a cylinder over a geodesic sphere of . In particular,
Proof.
From the hypothesis and [11, Theorem 3], where is an embedded convex topological sphere of . Moreover, in the case is contained in an open hemisphere of
Since a cylinder is nowhere strictly convex, it follows from the above theorem that, for a connected, proper, and strictly convex -hypersurface of must have a horizontal point. This fact, together with Theorem 18, gives our last result:
Corollary 2.
For any connected, properly immersed, and strictly convex -hypersurface of is necessarily an embedded rotational -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 and 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 -hypersurfaces in and geometric applications. Matemática Contemporânea, Sociedade Brasileira de Matemática. 34, 19–51 (2008).
- [4] Bérard, P., Sa Earp, R.: Minimal hypersurfaces in 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 An. Acad. Brasil. Ciênc. 77, 183–199 (2005).
- [10] Daniel, B.: Isometric immersions into and 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 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 -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 -hypersurfaces, barriers and Alexandrov theorem in . Ann. Mat. Pura Appl. 194, 1809–1834 (2015).
- [19] Elbert, M. F., Nelli, B., Santos, W.: Hypersurfaces with in . 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 . 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 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 and Bull. Braz. Math. Soc., New Series 41 (2), 199–209 (2010).