1. Introduction
The geometry of the real special linear group
is very rich, and there are many important and interesting papers investigating its fundamental properties; see, e.g., [
1,
2,
3,
4]. We can define a canonical left-invariant Riemannian metric on
with the isometry group of dimension 4. It admits a structure of naturally reductive homogeneous space. On the other hand, it is possible to equip
with a left-invariant metric such that the isometry group is only three-dimensional; see, e.g., [
5]. The Iwasawa decomposition allows to make use of global coordinates on
. A contact form can be defined in a canonical way, and it can be regarded as a connection form of the principal circle bundle
over the hyperbolic plane of constant curvature
. The projection becomes a Riemannian submersion. A canonical homogeneous Sasakian structure of constant holomorphic
-sectional curvature
may be also defined; see, e.g., [
2]. For a better understanding of
geometry, several investigations must be carried out to study its submanifolds. Over the last two decades, a large number of papers have investigated the geometry of curves and surfaces in
. Among them, we mention only a few: [
5,
6,
7,
8].
The present paper is structured as follows. The next section is a detailed description of the geometry of
. We collect several fundamental properties and put them together in order to obtain a self-contained paper. In
Section 3, we are interested in some surfaces in
and recall rotational surfaces, parallel surfaces and conoids.
Section 4 is devoted to Weingarten conoids in
. As basic examples, we have minimal and flat conoids and conoids with constant Gaussian curvature, respectively, constant mean curvature conoids. All these are studied in detail. Theorem 1 gives the classification of Weingarten conoids in
. In
Section 5, we study surfaces that are invariant by the left action of the nilpotent group
given in the Iwasawa decomposition. After we study minimal (respectively flat)
-surfaces in
, we prove that every
-surface in
is a Weingarten surface (Theorem 2).
4. Weingarten Conoids
A surface M immersed in a three-dimensional Riemannian manifold is called a Weingarten surface if there exists some (smooth) relation between its mean curvature H and its Gaussian curvature K. Obviously, minimal surfaces, CMC surfaces, flat surfaces and constant Gaussian curvature surfaces are typical examples of Weingarten surfaces.
The existence of a Weingarten relation
means that curvatures
H and
K (as functions of parameters
u and
v) are functionally related, and this is equivalent to the
Jacobian condition
at any
. The Jacobian condition characterizes Weingarten surfaces, and it is used to identify them when an explicit Weingarten relation cannot be immediately found.
In this section, we study conoids in that are Weingarten surfaces, and we call them Weingarten conoids.
1. The first example is obtained from minimal conoids, when
. They are studied in [
10], and the following result is obtained:
Proposition 2. A surface of the formwhere is the only minimal conoid. Moreover, this surface is helicoidal, namely, it is invariant under any helicoidal motion in . 2. The second example is given by flat conoids; they are Weingarten surfaces with . We obtain the following result.
Proposition 3. The only flat conoids in are rotational surfaces parameterized byfor which the generating curve is a vertical geodesic in . Here, is a real constant. Proof. The Gaussian curvature
K is computed in (
39). Then, the surface is flat if and only if
This equation is equivalent to
, where
is a smooth function depending on
v. Thus, we obtain
where
.
After we take the derivative with respect to
u, we find
If
is different from zero in a point, then it is different from zero on an open set. Therefore, on that open set, we need to have
. This equation has one (real) solution, call it
. It follows that
on that open set. This can only be possible if
is a constant
, which implies
, which is false. So,
on a certain open set. We deduce that
depends only on
v. Again, we need to have
, and hence
.
For the last part of the statement, we just notice that the generating curve is parameterized by , , and this represents a vertical geodesic in . □
3. For the third example, we consider conoids with constant Gaussian curvature; that is, the Weingarten function is , where . The following statement is true:
Proposition 4. The only conoids in with constant Gaussian curvature are the flat ones.
Proof. Using the expression of the Gaussian curvature from (
39), we obtain the following partial differential equation
. Denote by
. Since
, the previous equation may be rewritten as
where we set
. The solution of this equation depends on the sign of
, so we have to distinguish three situations.
Case 1:
The solution is an affine function in
s, equivalently
, where
are two smooth functions depending on
v. Squaring and then taking the derivative with respect to
u yields
Multiply by
u and take (again) the derivative with respect to
u to obtain
Because this relation is valid for arbitrary
in an open set, we must have
and
for any
v. Then,
and
, a real constant. The result from Proposition 3 is obtained.
Case 2:
The solution of the differential equation is given by
where
are two smooth functions depending on
v.
Similarly as before, we square and take the derivative with respect to
u. We obtain
Multiply by
u and, again, take the derivative with respect to
u twice; we obtain
Adding the Equation (
46) multiplied by
with the Equation (
47), we obtain
This equation must be fulfilled for arbitrary
; hence,
. This leads to
, which is a contradiction.
Case 3:
The solution of the differential equation is given by
where
are two smooth functions depending on
v. A similar technique as above yields a nonexistent result. □
4. A fourth example consists in CMC conoids, and they are obtained when the Weingarten function
, where
. Since minimal conoids were discussed before, we consider
. The mean curvature of a conoid is given in (
43). Consider the function
. We have to solve the partial differential equation
Integrating with respect to
v, we find
, where
A is a smooth function on
u. Take the derivative with respect to
u to obtain
. As
, after some elementary computations, we obtain the following equation
Squaring, we obtain
Take the derivative with respect to
v. Recall that we excluded the minimal conoids; hence,
does not vanish on a certain open interval. Moreover the expression
cannot be identically zero. After simplifications, we find
The function
A cannot be constant, otherwise
is trivially zero. So, after we take two more derivatives with respect to
u and perform the necessary simplifications, we obtain a contradiction.
The conclusion is given by the following:
Proposition 5. Any CMC conoid in is minimal. Moreover, it is a rotational surface parameterized by (45). We have seen that all these four “classical” types of Weingarten conoids in are helicoidal surfaces (with the special case when they are also rotational surfaces). Let us study the general case to understand if the function is affine for any Weingarten conoid.
The key is to write the Jacobian condition that characterizes a Weingarten surface. With
K and
H obtained in (
39) and (
43), respectively, we develop the equation
. Therefore, the following holds:
- (i)
either ,
- (ii)
or , .
Obviously, the second situation cannot occur. We now focus on the first differential equation. The already known solution is
(with
). We are looking for other solutions. Equation (i) is equivalent to
. The general solution is
where
.
We formulate the following:
Theorem 1. Let M be a conoid in , parameterized as in (34). Then, M is a Weingarten surface if and only if either the function ϕ is affine or where , . A Weingarten surface is called linear Weingarten if there exists a linear relation W between H and K, namely, the functional relation between H and K can be written as for some with .
Corollary 1. Linear Weingarten nontrivial conoids in do not exist.
Proof. By nontrivial Weingarten conoids, we understand the conoids obtained in the second case of the previous theorem. With
, we can compute
where
.
Suppose that we have the linear relation between H and K. Starting from , after straightforward computations, we obtain a polynomial in of degree 8. The leading coefficient is . Now, analyzing the other coefficients (under the hyprthesis ), we conclude and , which leads to a contradiction. □
5. -Surfaces
An immersed surface
is called an
-surface if it is invariant under the left translations of the subgroup
of
. It follows that such a surface can be parameterized as
with
and
. See, e.g., [
12].
Let us briefly sketch the geometry of these parameterized surfaces.
The pull-back of the metric
on
M is
Consider the orthonormal frame on
M
where
. Its orthonormal coframe is given by
We have
and
; hence,
where
is the connection matrix on
M.
The second structure equation is written as
. More precisely, we have
where
.
It follows that the Ricci tensor and the scalar curvature of
M are given by
where the Gaussian curvature
K of
M is given by
We now easily compute
It follows that the unit normal to
M is
Remark that
has the same orientation as
.
Let us write the expressions of the Levi-Civita connection ∇ on
M and the (scalar) second fundamental form
h of the immersion
, respectively:
Notice, as it is also pointed out in [
12], that the surface
M has no geodesic points.
The mean curvature of
M is
Weingarten -Surfaces
The first problems we wish to investigate are the minimal -surfaces, respectively, the flat -surfaces.
Proposition 6. An -surface defined by Equation (49) is minimal if and only if, under the initial conditions and , the function ϕ is Proof. The formula (
57) is equivalent to that found in [
7] for
. The minimality condition
leads to the differential equation
With the initial condition
, we obtain
. Then, the conclusion holds immediately.
Remark that the minimality condition is also equivalently to . □
Remark 2. In [12], the authors provide examples of constant mean curvature -surfaces by considering a certain height function to be constant. Proposition 7. An -surface defined by Equation (49) is flat if and only if, under the initial conditions and , the function ϕ iswhere is is the inverse function of the function , and . Proof. The flatness condition
is equivalent to the differential equation
Remark that
for any
v in an open interval where all the expressions make sense.
Consider the function , , which satisfies the following properties:
p is a strictly decreasing function, hence it is invertible.
p is an odd function, i.e., , .
, .
We have seen that
; hence,
, where the integration constant is obtained from the initial conditions:
. It follows that
Since
, the function
is increasing, and because
(due to the initial conditions), the function
is positive (for
). It follows that
. So, the definition domain of
v restricts to
.
For the sake of simplicity, let us write
. We compute
At this point, with
given by (
58), straightforward computations yield
. □
Remark 3. The flatness equation can be written in terms of the function ϕ as follows: We conclude this section with the following statement:
Theorem 2. Every -surface is a Weingarten surface.
Proof. The proof is elementary, because we have concrete expressions for the mean curvature H and for the Gaussian curvature K. We remark that they only depend on the variable v, so, the Jacobi condition is automatically satisfied. □
Remark 4. It is well known that with the metric (see Appendix A) is identified with the anti-de Sitter space . The Lie group acts on by the Ad-action. See, e.g., [7]. The -orbit of a space-like vector (in endowed with the pseudoscalar product) is the hyperbolic plane , the -orbit of a time-like vector is the Lorentz sphere and the -orbit of a null vector is the light-cone. Inoguchi noticed in [7] that rotational surfaces, as well as conoids, are Hopf cylinders over the hyperbolic plane and over Lorentz sphere, respecitvely. Of course, different metrics are considered on . In the same article, Inoguchi studied surfaces obtained over the -orbit of a null vector, i.e., the inverse image in of a curve in the light-cone of . This surface is precisely the -surface defined in this section. 6. Conclusions
A Weingarten surface is a surface for which the mean curvature
H is connected with its Gaussian curvature
K by a functional relation. These surfaces were introduced by J. Weingarten [
13,
14] in the context of the problem of finding all surfaces isometric to a given surface of revolution. Several geometric problems that involve Weingarten surfaces have been formulated and solved, especially in the Euclidean 3-space. See, e.g., [
15,
16,
17,
18]. Later on, generalized Weingarten surfaces in the Euclidean space of dimension 3 were studied. See, e.g., [
19,
20,
21]. The study of Weingarten surfaces may be extended to other ambient spaces (e.g., [
22,
23]). In this paper, we consider surfaces in the real special linear group
. This group is one of the most important Lie groups, not only because it consists in all linear transformations of
that preserve the oriented area but also because its universal cover is one of the eight Thurston geometries.
Along this article, we briefly describe some aspects of the geometry of the real special linear group and study some of its surfaces. We classify Weingarten conoids and show that there is no linear Weingarten nontrivial conoids in . We also prove that the only conoids in with constant Gaussian curvature are the flat ones. Finally, we show that any surface that is invariant under left translations of the subgroup (and we call them -surfaces, is a Weingarten surface.
There are several types of surfaces in to be studied, and we plan to investigate some of them in future works.