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

Constant sectional curvature surfaces with a semi-symmetric non-metric connection

Muhittin Evren Aydin Department of Mathematics, Faculty of Science, Firat University, Elazig, 23200 Turkey meaydin@firat.edu.tr Rafael López Departamento de Geometría y Topología Universidad de Granada 18071 Granada, Spain rcamino@ugr.es  and  Adela Mihai Technical University of Civil Engineering Bucharest, Department of Mathematics and Computer Science, 020396, Bucharest, Romania and Transilvania University of Braşov, Interdisciplinary Doctoral School, 500036, Braşov, Romania adela.mihai@utcb.ro, adela.mihai@unitbv.ro
Abstract.

Consider the Euclidean space 3\mathbb{R}^{3} endowed with a canonical semi-symmetric non-metric connection determined by a vector field 𝖢𝔛(3)\mathsf{C}\in\mathfrak{X}(\mathbb{R}^{3}). We study surfaces when the sectional curvature with respect to this connection is constant. In case that the surface is cylindrical, we obtain full classification when the rulings are orthogonal or parallel to 𝖢\mathsf{C}. If the surface is rotational, we prove that the rotation axis is parallel to 𝖢\mathsf{C} and we classify all conical rotational surfaces with constant sectional curvature. Finally, for the particular case 12\frac{1}{2} of the sectional curvature, the existence of rotational surfaces orthogonally intersecting the rotation axis is also obtained.

Key words and phrases:
rotational surface; sectional curvature; semi-symmetric connection; non-metric connection
1991 Mathematics Subject Classification:
53B40, 53C42, 53B20

1. Introduction

Friedmann and Schouten introduced in 1924 the notion of a semi-symmetric connection in a Riemannian manifold [4]. An affine connection ~\widetilde{\nabla} in a Riemannian manifold (M~,g~)(\widetilde{M},\tilde{g}) is said to be semi-symmetric connection if there is a non-zero vector field 𝖢𝔛(M~)\mathsf{C}\in\mathfrak{X}(\widetilde{M}) such that its torsion TT satisfies the identity

(1) T~(X,Y)=g~(𝖢,Y)Xg~(𝖢,X)Y,X,Y𝔛(M~).\widetilde{T}(X,Y)=\tilde{g}(\mathsf{C},Y)X-\tilde{g}(\mathsf{C},X)Y,\quad X,Y\in\mathfrak{X}(\widetilde{M}).

If in addition ~g~=0\widetilde{\nabla}\tilde{g}=0, the connection ~\widetilde{\nabla} is called a semi-symmetric metric connection [6]. Yano studied semi-symmetric metric connections with zero curvature and when the covariant derivative of the torsion tensor vanishes [14]. Submanifolds of Riemannian manifolds with semi-symmetric metric connections have been also investigated: without aiming a complete list, we refer to the readers to [7, 8, 9, 10, 12, 13].

If ~g~0\widetilde{\nabla}\tilde{g}\not=0, the connection is called semi-symmetric non-metric connection (snm-connection to abbreviate) [1, 2]. In this case, there is a relation between ~\widetilde{\nabla} and the Levi-Civita connection ~0\widetilde{\nabla}^{0} of (M~,g~)(\widetilde{M},\tilde{g}), namely,

(2) ~XY=~X0Y+g~(𝖢,Y)X,X,Y𝔛(M~).\widetilde{\nabla}_{X}Y=\widetilde{\nabla}^{0}_{X}Y+\tilde{g}(\mathsf{C},Y)X,\quad X,Y\in\mathfrak{X}(\widetilde{M}).

Such as it occurs for semi-symmetric metric connections, it is natural to study submanifolds of Riemannian manifolds endowed with a snm-connection ~\widetilde{\nabla}. Let MM be a submanifold of M~\widetilde{M}. Denote by \nabla (resp. 0\nabla^{0}) the induced connection on MM by ~\widetilde{\nabla} (resp. ~0\widetilde{\nabla}^{0}). The Gauss formulas are given by

~XY=XY+h(X,Y),~X0Y=X0Y+h0(X,Y),\begin{split}\widetilde{\nabla}_{X}Y&=\nabla_{X}Y+h(X,Y),\\ \widetilde{\nabla}^{0}_{X}Y&=\nabla^{0}_{X}Y+h^{0}(X,Y),\end{split}

for all X,Y𝔛(M)X,Y\in\mathfrak{X}(M), where hh is a (0,2)(0,2)-tensor field on MM and h0h^{0} is the second fundamental form of MM. It is known that h=h0h=h^{0} [2]. Hence that problems of extrinsic nature are the same one that for the Levi-Civita connection.

We consider intrinsic geometry of submanifolds. One of the main concepts in intrinsic Riemannian geometry is that of sectional curvature. It is natural to carry this concept for snm-connections. However, the sectional curvature of (M~,g~)(\widetilde{M},\tilde{g}) with respect to ~\widetilde{\nabla} cannot be defined by the usual way as the Levi-Civita connection ~0\widetilde{\nabla}^{0}. This is because if R~\widetilde{R} is the curvature tensor of ~\widetilde{\nabla}, the quantity g~(R~(e1,e2)e2,e1)\tilde{g}(\widetilde{R}(e_{1},e_{2})e_{2},e_{1}), where {e1,e2}\{e_{1},e_{2}\} is an orthonormal basis of π\pi, depends on the basis {e1,e2}\{e_{1},e_{2}\}: see Sect. 2 for details. In contrast, the third author of this paper, jointly with I. Mihai, proved that g~(R~(e1,e2)e2,e1)+g~(R~(e2,e1)e1,e2)\tilde{g}(\widetilde{R}(e_{1},e_{2})e_{2},e_{1})+\tilde{g}(\widetilde{R}(e_{2},e_{1})e_{1},e_{2}) is independent on the basis [11]. Then they introduced the following notion of sectional curvature for snm-connections.

Definition 1.1.

Let (M~,g~)(\widetilde{M},\tilde{g}) be a Riemannian manifold endowed with a snm-connection ~\widetilde{\nabla}. If π\pi is a plane in TpM~T_{p}\widetilde{M} with an orthonormal basis {e1,e2}\{e_{1},e_{2}\}, then the sectional curvature of π\pi with respect to ~\widetilde{\nabla} is defined by

(3) K~(π)=g~(R~(e1,e2)e2,e1)+g~(R~(e2,e1)e1,e2)2.\widetilde{K}(\pi)=\frac{\tilde{g}(\widetilde{R}(e_{1},e_{2})e_{2},e_{1})+\tilde{g}(\widetilde{R}(e_{2},e_{1})e_{1},e_{2})}{2}.

Once we have the notion of sectional curvature, it is natural to ask for those submanifolds with constant sectional curvature. As for the Levi-Civita connection, this question is difficult to address in all its generality.

In this paper, we consider that the ambient space is the 33-dimensional Euclidean space 3\mathbb{R}^{3} endowed with the Euclidean metric ,\langle,\rangle. The amount of snm-connections of 3\mathbb{R}^{3} is given by the vector fields 𝖢\mathsf{C} in the definition (1) of a semi-symmetric connection. One of the simplest choices of snm-connections of 3\mathbb{R}^{3} is that 𝖢\mathsf{C} is a canonical vector field. To be precise, let (x,y,z)(x,y,z) be canonical coordinates of 3\mathbb{R}^{3} and let {x,y,z}\{\partial_{x},\partial_{y},\partial_{z}\} be the corresponding basis of 𝔛(3)\mathfrak{X}(\mathbb{R}^{3}). In fact, if the vector field 𝖢\mathsf{C} is assumed to be canonical, namely 𝖢{x,y,z}\mathsf{C}\in\{\partial_{x},\partial_{y},\partial_{z}\} then, after a change of coordinates of 3\mathbb{R}^{3}, 𝖢\mathsf{C} is a unit constant vector field.

Definition 1.2.

A snm-connection ~\widetilde{\nabla} on 3\mathbb{R}^{3} is said to be canonical if 𝖢𝔛(3)\mathsf{C}\in\mathfrak{X}(\mathbb{R}^{3}) is a unit constant vector field.

From now on, unless otherwise specified, we denote by 𝖢\mathsf{C} a unit constant vector field on 3\mathbb{R}^{3}.

Definitively, the problem that we study is the classification of surfaces with constant sectional curvature for a given canonical snm-connection ~\widetilde{\nabla} of 3\mathbb{R}^{3}. A way to tackle this problem is to impose a certain geometric condition on the surface. A natural condition is that the surface is invariant by a one-parameter group of rigid motions. Denote KK by the sectional curvature with respect to the induced connection on the surface from ~\widetilde{\nabla}. Assuming a certain invariance of the surface, it allows us to expect that the equation K=cK=c can be expressed as an ordinary differential equation, where, under mild conditions, the existence is assured. For example, we can assume that the surface is invariant by a group of translations or that the surface is invariant by a group of rotations. In the first case, the surface is called cylindrical and in the second one, rotational surface, or surface of revolution.

The organization of this paper is according to both types of surfaces. In Sect. 2 we prove an useful formula for computing the sectional curvature KK of a surface in terms of that of 3\mathbb{R}^{3} and the Gaussian and mean curvatures of the surface. We will show some explicit examples of computations of sectional curvatures.

Section 3 is devoted to cylindrical surfaces. A cylindrical surface can be parametrized by ψ(s,t)=γ(s)+tw\psi(s,t)=\gamma(s)+t\vec{w}, sIs\in I\subset\mathbb{R}, tt\in\mathbb{R}, where w3\vec{w}\in\mathbb{R}^{3} is a unitary vector and γ:I3\gamma\colon I\to\mathbb{R}^{3} is a curve contained in a plane orthogonal to w\vec{w}. The surface is invariant by the group of translations generated by w\vec{w}. After computing the sectional curvature KK in Thm. 3.1, in Cor. 3.2, we prove that any cylindrical surface whose rulings are parallel to 𝖢\mathsf{C} has constant sectional curvature KK, being K=12K=\frac{1}{2}. Another interesting case of cylindrical surfaces is that the rulings are orthogonal to 𝖢\mathsf{C}. We obtain a full classification of these cylindrical surfaces with KK constant depending on the sign of KK (Cor. 3.3). For the particular values K=1/2K=1/2 and K=1/2K=-1/2, in Cor. 3.4 we obtain explicit parametrizations of the surfaces.

Rotational surfaces are invariant by rotations about an axis LL of 3\mathbb{R}^{3} and such surfaces with KK constant will be studied in Sect. 4. It is worth to point out that there is not a priori relation between the axis LL and the vector field 𝖢\mathsf{C} that defines the canonical snm-connection. However, we prove in Thm. 4.1 that LL and 𝖢\mathsf{C} must be parallel. In Thm. 4.3, we classify all conical rotational surfaces with KK constant proving that these surfaces are planes or circular cylinders. As a last observation, when K=12K=\frac{1}{2}, in Thm. 4.5, the existence of rotational surfaces orthogonally intersecting the rotation axis is also obtained.

2. Preliminaries

Let (M~,g~)(\widetilde{M},\tilde{g}) be a Riemannian manifold of dimension 2\geq 2 and let ~\widetilde{\nabla} be an affine connection on M~\widetilde{M}. The torsion and curvature of ~\widetilde{\nabla} are respectively a (1,2)(1,2)-tensor field T~\widetilde{T} and a (1,3)(1,3)-tensor field R~\widetilde{R} defined by

T~(X,Y)=~XY~YX[X,Y],R~(X,Y)Z=~X~YZ~Y~XZ~[X,Y]Z,\begin{split}\widetilde{T}(X,Y)&=\widetilde{\nabla}_{X}Y-\widetilde{\nabla}_{Y}X-[X,Y],\\ \widetilde{R}(X,Y)Z&=\widetilde{\nabla}_{X}\widetilde{\nabla}_{Y}Z-\widetilde{\nabla}_{Y}\widetilde{\nabla}_{X}Z-\widetilde{\nabla}_{[X,Y]}Z,\end{split}

for X,Y,Z𝔛(M~)X,Y,Z\in\mathfrak{X}(\widetilde{M}). Let ~\widetilde{\nabla} be a snm-connection on (M~,g~)(\widetilde{M},\tilde{g}) determined by a vector field 𝖢𝔛(M~)\mathsf{C}\in\mathfrak{X}(\widetilde{M}). Using (2), there is also a relation between R~\widetilde{R} and the Riemannian curvature tensor R~0\widetilde{R}^{0} of ~0\widetilde{\nabla}^{0} ([1, 11]). Indeed, for orthonormal vectors e1,e2TpM~e_{1},e_{2}\in T_{p}\widetilde{M}, pM~p\in\widetilde{M}, we have

g~(R~(e1,e2)e2,e1)=g~(R~0(e1,e2)e2,e1)e2(g~(𝖢,e2))+g~(𝖢,~e20e2)+g~(𝖢,e2)2.\tilde{g}(\widetilde{R}(e_{1},e_{2})e_{2},e_{1})=\tilde{g}(\widetilde{R}^{0}(e_{1},e_{2})e_{2},e_{1})-e_{2}(\tilde{g}(\mathsf{C},e_{2}))+\tilde{g}(\mathsf{C},\widetilde{\nabla}^{0}_{e_{2}}e_{2})+\tilde{g}(\mathsf{C},e_{2})^{2}.

Although the first term at the right hand-side is the sectional curvature of the plane section π=span{e1,e2}\pi=\text{span}\{e_{1},e_{2}\}, the term at the left hand-side depends on the choice of the basis of π\pi. Therefore, the value g~(R~(e1,e2)e2,e1)\tilde{g}(\widetilde{R}(e_{1},e_{2})e_{2},e_{1}) does not stand for a sectional curvature. The quantity (3) was proposed in [11] as the definition of sectional curvature of π\pi with respect to ~\widetilde{\nabla} because it is independent on the basis in TpM~T_{p}\widetilde{M}. In case that {e1,e2}\{e_{1},e_{2}\} is an arbitrary basis of π\pi, it is immediate to see

(4) K~(π)=g~(R~(e1,e2)e2,e1)+g~(R~(e2,e1)e1,e2)2(g~(e1,e1)g~(e2,e2)g~(e1,e2)2).\widetilde{K}(\pi)=\frac{\tilde{g}(\widetilde{R}(e_{1},e_{2})e_{2},e_{1})+\tilde{g}(\widetilde{R}(e_{2},e_{1})e_{1},e_{2})}{2(\tilde{g}(e_{1},e_{1})\tilde{g}(e_{2},e_{2})-\tilde{g}(e_{1},e_{2})^{2})}.

From now on, suppose that M~\widetilde{M} is the Euclidean space 3\mathbb{R}^{3}. We compute the sectional curvature of a plane of 3\mathbb{R}^{3}.

Proposition 2.1.

Let ~\widetilde{\nabla} be a canonical snm-connection on 3\mathbb{R}^{3}. If π\pi is a plane of 3\mathbb{R}^{3}, then its sectional curvature is

K~(π)=u,𝖢2+v,𝖢22,\widetilde{K}(\pi)=\frac{\langle\vec{u},\mathsf{C}\rangle^{2}+\langle\vec{v},\mathsf{C}\rangle^{2}}{2},

where {u,v}\{\vec{u},\vec{v}\} is an orthonormal basis of π\pi. As a consequence, K~(π)\widetilde{K}(\pi) is constant with 0K~(π)120\leq\widetilde{K}(\pi)\leq\frac{1}{2}. Furthermore, K~(π)=0\widetilde{K}(\pi)=0 (resp. K~(π)=12\widetilde{K}(\pi)=\frac{1}{2}) if and only if π\pi is perpendicular to 𝖢\mathsf{C} (resp. π\pi is parallel to 𝖢\mathsf{C}).

Proof.

Using (2) we compute

~uu=u,𝖢u,\displaystyle\widetilde{\nabla}_{\vec{u}}\vec{u}=\langle\vec{u},\mathsf{C}\rangle\vec{u}, ~uv=v,𝖢u,\displaystyle\widetilde{\nabla}_{\vec{u}}\vec{v}=\langle\vec{v},\mathsf{C}\rangle\vec{u},
~vu=u,𝖢v,\displaystyle\widetilde{\nabla}_{\vec{v}}\vec{u}=\langle\vec{u},\mathsf{C}\rangle\vec{v}, ~vv=v,𝖢v,\displaystyle\widetilde{\nabla}_{\vec{v}}\vec{v}=\langle\vec{v},\mathsf{C}\rangle\vec{v},

and

~u~vv=v,𝖢2u,\displaystyle\widetilde{\nabla}_{\vec{u}}\widetilde{\nabla}_{\vec{v}}\vec{v}=\langle\vec{v},\mathsf{C}\rangle^{2}\vec{u}, ~v~uv=u,𝖢v,𝖢v,\displaystyle\widetilde{\nabla}_{\vec{v}}\widetilde{\nabla}_{\vec{u}}\vec{v}=\langle\vec{u},\mathsf{C}\rangle\langle\vec{v},\mathsf{C}\rangle\vec{v},
~v~uu=u,𝖢2v,\displaystyle\widetilde{\nabla}_{\vec{v}}\widetilde{\nabla}_{\vec{u}}\vec{u}=\langle\vec{u},\mathsf{C}\rangle^{2}\vec{v}, ~u~vu=u,𝖢v,𝖢u.\displaystyle\widetilde{\nabla}_{\vec{u}}\widetilde{\nabla}_{\vec{v}}\vec{u}=\langle\vec{u},\mathsf{C}\rangle\langle\vec{v},\mathsf{C}\rangle\vec{u}.

Also it is easy to see [u,v]=0[\vec{u},\vec{v}]=0. Hence the curvature tensor R~\widetilde{R} is determined by

R~(u,v)v=v,𝖢2uu,𝖢v,𝖢v,R~(v,u)u=u,𝖢2vu,𝖢v,𝖢u.\begin{split}\widetilde{R}(\vec{u},\vec{v})\vec{v}&=\langle\vec{v},\mathsf{C}\rangle^{2}\vec{u}-\langle\vec{u},\mathsf{C}\rangle\langle\vec{v},\mathsf{C}\rangle\vec{v},\\ \widetilde{R}(\vec{v},\vec{u})\vec{u}&=\langle\vec{u},\mathsf{C}\rangle^{2}\vec{v}-\langle\vec{u},\mathsf{C}\rangle\langle\vec{v},\mathsf{C}\rangle\vec{u}.\end{split}

This gives the formula for K~(π)\widetilde{K}(\pi). The last statement is a consequence of this formula. ∎

Remark 2.2.

The notion of scalar curvature at a point p3p\in\mathbb{R}^{3} with respect to a snm-connection ~\widetilde{\nabla} can be introduced in a similar manner as for the Levi-Civita connection. Let {u,v,w}\{\vec{u},\vec{v},\vec{w}\} be an orthonormal basis of Tp3T_{p}\mathbb{R}^{3}, p3p\in\mathbb{R}^{3}. The scalar curvature ρ\rho with respect to ~\widetilde{\nabla} is defined by

ρ(p)=K~(u,v)+K~(u,w)+K~(v,w),p3.\rho(p)=\widetilde{K}(\vec{u},\vec{v})+\widetilde{K}(\vec{u},\vec{w})+\widetilde{K}(\vec{v},\vec{w}),\quad p\in\mathbb{R}^{3}.

If ~\widetilde{\nabla} is canonical, then by Prop. 2.1 the scalar curvature is constant, namely ρ(p)=1\rho(p)=1, for every p3p\in\mathbb{R}^{3}.

We conclude this section establishing a relation between the sectional curvatures KK and K~\widetilde{K} of a surface in 3\mathbb{R}^{3} in terms of the Gaussian and the mean curvatures of the surface. Let MM be an oriented surface immersed in 3\mathbb{R}^{3} and NN its unit normal vector field. Let also ~\widetilde{\nabla} be a snm-connection on 3\mathbb{R}^{3} determined by an arbitrary vector field 𝖢𝔛(3)\mathsf{C}\in\mathfrak{X}(\mathbb{R}^{3}). We have the decomposition of 𝖢\mathsf{C} in its tangential and normal components with respect to MM,

𝖢=𝖢+𝖢,NN.\mathsf{C}=\mathsf{C}^{\top}+\langle\mathsf{C},N\rangle N.

If X,Y,Z,U𝔛(M)X,Y,Z,U\in\mathfrak{X}(M), then the Gauss equation with respect to ~\widetilde{\nabla} is ([2]):

(5) R(X,Y)Z,U=R~(X,Y)Z,Uh(X,Z),h(Y,U)+h(X,U),h(Y,Z)+𝖢,N(h(X,Z),NY,Uh(Y,Z),NX,U).\begin{split}\langle R(X,Y)Z,U\rangle&=\langle\widetilde{R}(X,Y)Z,U\rangle-\langle h(X,Z),h(Y,U)\rangle+\langle h(X,U),h(Y,Z)\rangle\\ &+\langle\mathsf{C},N\rangle\left(\langle h(X,Z),N\rangle\langle Y,U\rangle-\langle h(Y,Z),N\rangle\langle X,U\rangle\right).\end{split}
Proposition 2.3.

Let MM be an oriented surface in 3\mathbb{R}^{3} and denote by GG and HH the Gaussian curvature and the mean curvature of MM, respectively, with respect to the Levi-Civita connection. Then

(6) K=K~+G𝖢,NH.K=\widetilde{K}+G-\langle\mathsf{C},N\rangle H.

Moreover, if pMp\in M then there is an orthogonal basis {e1,e2}\{e_{1},e_{2}\} of TpMT_{p}M such that

(7) R(e1,e2)e2,e1=R~(e1,e2)e2,e1+h11h22g11h22𝖢,N,R(e2,e1)e1,e2=R~(e2,e1)e1,e2+h11h22g22h11𝖢,N,\begin{split}\langle R(e_{1},e_{2})e_{2},e_{1}\rangle&=\langle\widetilde{R}(e_{1},e_{2})e_{2},e_{1}\rangle+h_{11}h_{22}-g_{11}h_{22}\langle\mathsf{C},N\rangle,\\ \langle R(e_{2},e_{1})e_{1},e_{2}\rangle&=\langle\widetilde{R}(e_{2},e_{1})e_{1},e_{2}\rangle+h_{11}h_{22}-g_{22}h_{11}\langle\mathsf{C},N\rangle,\end{split}

where gij=ei,ejg_{ij}=\langle e_{i},e_{j}\rangle and hijh_{ij} are the coefficients of the second fundamental form hh.

Proof.

Since the codimension of MM in 3\mathbb{R}^{3} is 11, it has trivially flat normal bundle. Let {e1,e2}\{e_{1},e_{2}\} be an orthogonal basis of TpMT_{p}M such that g12=0g_{12}=0 and h120=0h_{12}^{0}=0, where hij0h_{ij}^{0} are the coefficients of the second fundamental form of MM with respect to the Levi-Civita connection: see [3, Props. 3.1 and 3.2]. Therefore we have h12=0h_{12}=0 because h=h0h=h^{0}. By the Gauss equation (5) we obtain (7). With respect to this basis, we have

G=h11h22g11g22,H=g22h11+g11h222g11g22.G=\frac{h_{11}h_{22}}{g_{11}g_{22}},\quad H=\frac{g_{22}h_{11}+g_{11}h_{22}}{2g_{11}g_{22}}.

Identity (6) is a consequence of (7) and formulas (4) for KK and K~\widetilde{K}. ∎

Remark 2.4.

Identity (6) is satisfied for any vector field 𝖢𝔛(3)\mathsf{C}\in\mathfrak{X}(\mathbb{R}^{3}). Notice also that KK is invariant by translations of 3\mathbb{R}^{3}. This is because GG and HH do no change, as well as K~\widetilde{K} because a plane π\pi is not affected by translations. However, rigid motions change the value of K~\widetilde{K} and, consequently of KK. This is because of the presence of the vector field 𝖢\mathsf{C} in (2) for computing the successive covariant derivatives.

Simple consequences of the relation (6) appear in the following result.

Corollary 2.5.
  1. (1)

    For a plane, we have K=K~K=\widetilde{K}. In particular, K0K\geq 0 and equality holds if and only if the plane is orthogonal to the vector field 𝖢\mathsf{C}.

  2. (2)

    For a cylindrical surface whose rulings are parallel to 𝖢\mathsf{C}, we have K=K~K=\widetilde{K}.

Proof.

It is immediate because for a plane we have G=H=0G=H=0, and for a cylindrical surface with rulings parallel to 𝖢\mathsf{C} we have G=0G=0 and 𝖢,N=0\langle\mathsf{C},N\rangle=0. ∎

Thanks to this corollary we see that a plane and a cylindrical cylinder satisfy the equality K=K~K=\widetilde{K}. In general, a surface satisfies K=K~K=\widetilde{K} if and only if G=𝖢,NHG=\langle\mathsf{C},N\rangle H. In case that 𝖢\mathsf{C} is a canonical vector field, we construct such a surface as follows.

Example 2.6.

Let 𝖢=z\mathsf{C}=\partial_{z}. To find a surface satisfying G=z,NHG=\langle\partial_{z},N\rangle H, we consider surfaces that are graphs of smooth functions z=u(x,y)z=u(x,y), where (x,y)Ω2(x,y)\in\Omega\subset\mathbb{R}^{2}. Then it is not difficult to find that the relation G=N,zHG=\langle N,\partial_{z}\rangle H is written by

2(uxxuyyuxy2)=(1+uy2)uxx2uxuyuxy+(1+ux2)uyy.2(u_{xx}u_{yy}-u_{xy}^{2})=(1+u_{y}^{2})u_{xx}-2u_{x}u_{y}u_{xy}+(1+u_{x}^{2})u_{yy}.

We find solutions of this equation by the technique of separation of variables. Assuming u(x,y)=f(x)+g(y)u(x,y)=f(x)+g(y), for smooth functions f=f(x)f=f(x) and g=g(y)g=g(y), xIx\in I\subset\mathbb{R}, yJy\in J\subset\mathbb{R}, the above equation becomes

(8) 2f′′g′′=f′′(1+g2)+g′′(1+f2),2f^{\prime\prime}g^{\prime\prime}=f^{\prime\prime}(1+g^{\prime 2})+g^{\prime\prime}(1+f^{\prime 2}),

for all xIx\in I, yJy\in J. Here a prime denotes the derivative with respect to each variable. A solution of Eq. (8) appears when ff and gg are linear functions, f′′=g′′=0f^{\prime\prime}=g^{\prime\prime}=0 identically. Then MM is a plane parallel to the xyxy-plane. We discard this case by assuming f′′g′′0f^{\prime\prime}g^{\prime\prime}\neq 0 on I×JI\times J. Dividing Eq. (8) by f′′g′′f^{\prime\prime}g^{\prime\prime}, we obtain

21+g2g′′=1+f2f′′.2-\frac{1+g^{\prime 2}}{g^{\prime\prime}}=\frac{1+f^{\prime 2}}{f^{\prime\prime}}.

Since the left hand-side depends only on the variable yy and the right hand-side on the variable xx, then we deduce the existence of the nonzero constant cc such that

21+g2g′′=1c=1+f2f′′.2-\frac{1+g^{\prime 2}}{g^{\prime\prime}}=\frac{1}{c}=\frac{1+f^{\prime 2}}{f^{\prime\prime}}.

Notice that if c=1/2c=1/2, then 1+g2=01+g^{\prime 2}=0, which it is not possible. By solving these equations, we obtain, up to translations of xx and yy and suitable constants,

u(x,y)=1clogcos(cx)2c1clogcos(cy2c1)).u(x,y)=-\frac{1}{c}\log\cos(cx)-\frac{2c-1}{c}\log\cos(\frac{cy}{2c-1})).

See Fig. 1 for the particular case c=1c=1.

Refer to caption
Figure 1. Graph of z=log(cos(x)cos(y))z=-\log(\cos(x)\cos(y)).

3. Cylindrical surfaces

Let MM be a cylindrical surface in 3\mathbb{R}^{3} whose rulings are parallel to w\vec{w}, where w3\vec{w}\in\mathbb{R}^{3}, |w|=1|\vec{w}|=1. If γ=γ(s)\gamma=\gamma(s) is the generating curve of MM contained in a plane orthogonal to w\vec{w}, then a parametrization of MM is

(9) ψ(s,t)=γ(s)+tw,sI,t.\psi(s,t)=\gamma(s)+t\vec{w},\quad s\in I,t\in\mathbb{R}.

Without loss of generality, we suppose that γ\gamma is parametrized by arc-length. Let 𝐧{\bf n} be the unit normal vector of γ\gamma and let κ\kappa be the Frenet curvature of γ\gamma with γ′′=κ𝐧\gamma^{\prime\prime}=\kappa{\bf n}. Since γ\gamma is contained in a plane orthogonal to w\vec{w}, consider the orientation on γ\gamma such that (γ,w,𝐧)=1(\gamma^{\prime},\vec{w},{\bf n})=1, where (a,b,c)(\vec{a},\vec{b},\vec{c}) stands for the determinant of the matrix formed by three vectors a\vec{a}, b\vec{b}, c\vec{c} of 3\mathbb{R}^{3}.

Theorem 3.1.

Let ~\widetilde{\nabla} be a canonical snm-connection on 3\mathbb{R}^{3}. If MM is a cylindrical surface parametrized by (9), then its sectional curvature KK with respect to ~\widetilde{\nabla} is

(10) K=12(w,𝖢2+γ,𝖢2κ𝐧,𝖢).K=\frac{1}{2}\left(\langle\vec{w},\mathsf{C}\rangle^{2}+\langle\gamma^{\prime},\mathsf{C}\rangle^{2}-\kappa\langle{\bf n},\mathsf{C}\rangle\right).
Proof.

The tangent plane of MM is spanned by an orthonormal basis {e1,e2}\{e_{1},e_{2}\}, where e1=ψs=γe_{1}=\psi_{s}=\gamma^{\prime} and e2=ψt=we_{2}=\psi_{t}=\vec{w}. We know that the Gaussian curvature is G=0G=0. The Gauss map and the mean curvature of MM are given by

N=γ×w,H=(γ,w,γ′′)2=κ2.N=\gamma^{\prime}\times w,\quad H=\frac{(\gamma^{\prime},w,\gamma^{\prime\prime})}{2}=\frac{\kappa}{2}.

We compute the covariant derivatives as follows

~e1e1=γ′′+γ,𝖢γ,~e1e2=w,𝖢γ,~e2e1=γ,𝖢w,~e2e2=w,𝖢w.\begin{array}[]{lll}\widetilde{\nabla}_{e_{1}}e_{1}=\gamma^{\prime\prime}+\langle\gamma^{\prime},\mathsf{C}\rangle\gamma^{\prime},&\widetilde{\nabla}_{e_{1}}e_{2}=\langle\vec{w},\mathsf{C}\rangle\gamma^{\prime},\\ \widetilde{\nabla}_{e_{2}}e_{1}=\langle\gamma^{\prime},\mathsf{C}\rangle\vec{w},&\widetilde{\nabla}_{e_{2}}e_{2}=\langle\vec{w},\mathsf{C}\rangle\vec{w}.\end{array}

Because [e1,e2]=~e10e2~e20e1[e_{1},e_{2}]=\widetilde{\nabla}^{0}_{e_{1}}e_{2}-\widetilde{\nabla}^{0}_{e_{2}}e_{1}, we conclude that [e1,e2]=0[e_{1},e_{2}]=0. We also compute

~e1~e2e2=w,𝖢2γ,~e2~e1e2=w,𝖢γ,𝖢w,~e2~e1e1=(γ′′,𝖢+γ,𝖢2)w,~e1~e2e1=γ′′,𝖢w+γ,𝖢w,𝖢γ,\begin{split}\widetilde{\nabla}_{e_{1}}\widetilde{\nabla}_{e_{2}}e_{2}&=\langle\vec{w},\mathsf{C}\rangle^{2}\gamma^{\prime},\\ \widetilde{\nabla}_{e_{2}}\widetilde{\nabla}_{e_{1}}e_{2}&=\langle\vec{w},\mathsf{C}\rangle\langle\gamma^{\prime},\mathsf{C}\rangle\vec{w},\\ \widetilde{\nabla}_{e_{2}}\widetilde{\nabla}_{e_{1}}e_{1}&=(\langle\gamma^{\prime\prime},\mathsf{C}\rangle+\langle\gamma^{\prime},\mathsf{C}\rangle^{2})\vec{w},\\ \widetilde{\nabla}_{e_{1}}\widetilde{\nabla}_{e_{2}}e_{1}&=\langle\gamma^{\prime\prime},\mathsf{C}\rangle\vec{w}+\langle\gamma^{\prime},\mathsf{C}\rangle\langle\vec{w},\mathsf{C}\rangle\gamma^{\prime},\end{split}

and thus

K~=12(w,𝖢2+γ,𝖢2).\widetilde{K}=\frac{1}{2}\left(\langle\vec{w},\mathsf{C}\rangle^{2}+\langle\gamma^{\prime},\mathsf{C}\rangle^{2}\right).

By (6) we find

K=12(w,𝖢2+γ,𝖢2(γ,w,𝖢)(γ′′,γ,w)).K=\frac{1}{2}\left(\langle\vec{w},\mathsf{C}\rangle^{2}+\langle\gamma^{\prime},\mathsf{C}\rangle^{2}-(\gamma^{\prime},\vec{w},\mathsf{C})(\gamma^{\prime\prime},\gamma^{\prime},\vec{w})\right).

The result follows because γ×w=𝐧\gamma^{\prime}\times\vec{w}={\bf n} and γ′′=κ𝐧\gamma^{\prime\prime}=\kappa{\bf n}. ∎

We distinguish two particular cases, when the rulings are parallel or orthogonal to the constant vector field 𝖢\mathsf{C}.

Corollary 3.2.

Any cylindrical surface whose rulings are parallel to 𝖢\mathsf{C} has constant sectional curvature K=1/2K=1/2 with respect to a canonical snm-connection determined by 𝖢\mathsf{C}.

Suppose that the rulings are orthogonal to 𝖢\mathsf{C}. In the next result we are going to obtain explicit parametrizations of cylindrical surfaces with constant sectional curvature. Without loss of generality, we suppose that 𝖢=z\mathsf{C}=\partial_{z} and w=(0,1,0)\vec{w}=(0,1,0). Then γ\gamma is contained in the xzxz-plane, say γ(s)=(x(s),0,z(s))\gamma(s)=(x(s),0,z(s)), for smooth functions x,z:Ix,z\colon I\to\mathbb{R}. The case that MM is a plane is particular. Any plane of 3\mathbb{R}^{3} perpendicular to z\partial_{z} can be viewed as a cylindrical surface with rulings orthogonal to z\partial_{z}. By Prop. 2.1, we know that its curvature KK is constant with 0K120\leq K\leq\frac{1}{2}. We discard this case.

Corollary 3.3.

Let ~\widetilde{\nabla} be the canonical snm-connection determined by z\partial_{z} and MM be a non-planar cylindrical surface whose rulings are orthogonal to z\partial_{z}. If the sectional curvature KK with respect to ~\widetilde{\nabla} is constant, then the parametrization of the generating curve γ\gamma is

  1. (1)

    Case K>0K>0, then γ(s)=(s12Ktanh2(2Kt)𝑑t,log(cosh(2Ks)))\gamma(s)=(\int^{s}\sqrt{1-2K\tanh^{2}\left(\sqrt{2K}t\right)}\,dt,-\log(\cosh(\sqrt{2K}s))).

  2. (2)

    Case K=0K=0, then γ(s)=(±tan1(s21)s21,log(s))\gamma(s)=(\pm\tan^{-1}\left(\sqrt{s^{2}-1}\right)-\sqrt{s^{2}-1},-\log(s)).

  3. (3)

    Case K<0K<0, then γ(s)=(s12Ktan2(2Kt)dt,log(cos(2Ks))\gamma(s)=(\int^{s}\sqrt{1-2K\tan^{2}\left(\sqrt{-2K}t\right)}\,dt,-\log(\cos(\sqrt{-2K}s)).

Proof.

Since γ\gamma is parametrized by arc-length, we know x2+z2=1x^{\prime 2}+z^{\prime 2}=1 and γ=(x,0,z)\gamma^{\prime}=(x^{\prime},0,z^{\prime}). By the choice of orientation on γ\gamma given in Thm. 3.1, the normal vector is 𝐧=(z,0,x){\bf n}=(-z^{\prime},0,x^{\prime}). Identity (10) is

z′′=z22K.z^{\prime\prime}=z^{\prime 2}-2K.

The solution of this equation depends on the sign of KK. Up to an additive constant on the functions xx and zz as well as in the parameter ss, which it is only a translation of the surface (Rem. 2.4), we have

  1. (1)

    K>0K>0; then z(s)=log(cosh(2Ks))z(s)=-\log(\cosh(\sqrt{2K}s)).

  2. (2)

    K=0K=0; then z(s)=log(s)z(s)=-\log(s).

  3. (3)

    K<0K<0; then z(s)=log(cos(2Ks))z(s)=-\log(\cos(\sqrt{-2K}s)).

The result follows from the identity x2+z2=1x^{\prime 2}+z^{\prime 2}=1. ∎

In Fig. 2 we depict some graphics of the generating curves for different values of KK. Notice that the domain of γ\gamma is not \mathbb{R} in general because the root that appears in the integrals that define the xx-coordinate of γ\gamma. For example, if K=0K=0, then s[1,)s\in[1,\infty).

Refer to caption
Refer to caption
Refer to caption
Figure 2. Graphics of generating curves of Cor. 3.3: K=1K=1 (left), K=0K=0 (middle) and K=1K=-1 (right).

It is worth to consider the cases K=1/2K=1/2 and K=1/2K=-1/2. In such a case, the integrals of Cor. 3.3 can be explicitly solved.

Corollary 3.4.

Let ~\widetilde{\nabla} be the canonical snm-connection determined by z\partial_{z}, MM a non-planar cylindrical surface whose rulings are orthogonal to z\partial_{z}, and KK the sectional curvature of MM with respect to ~\widetilde{\nabla}.

  1. (1)

    If K=12K=\frac{1}{2}, then γ(s)=(tan1(sinh(s)),logcosh(s))\gamma(s)=(\tan^{-1}(\sinh(s)),-\log\cosh(s)).

  2. (2)

    If K=12K=-\frac{1}{2}, then γ(s)=(2sin1(2sin(s))cot1(cot(s)21),logcos(s))\gamma(s)=(\sqrt{2}\sin^{-1}\left(\sqrt{2}\sin(s)\right)-\cot^{-1}\left(\sqrt{\cot(s)^{2}-1}\right),-\log\cos(s)).

For K=1/2K=1/2, the curve γ\gamma in (1) is called grim reaper. The usual parametrization of the grim reaper is y(x)=log(cos(x))y(x)=-\log(\cos(x)) in the (x,y)(x,y)-plane 2\mathbb{R}^{2}. This is deduced immediately by letting x=tan1(sinh(s))x=\tan^{-1}(\sinh(s)). The grim reaper is a remarkable curve in the theory of curve-shortening flow [5].

4. Rotational surfaces

In this section we study rotational surfaces with constant sectional curvature. A first problem is the relation between the axis LL of the surface and the vector field 𝖢\mathsf{C} that defines the canonical snm-connection. As we said in the Introduction, there is no a priori a relation between both. However, we prove that they must be parallel.

Theorem 4.1.

Let ~\widetilde{\nabla} be a canonical snm-connection on 3\mathbb{R}^{3} determined by the vector field 𝖢\mathsf{C} and MM be a rotational surface in 3\mathbb{R}^{3} about an axis LL. If MM has constant sectional curvature KK, then either MM is any plane and K0K\geq 0 or LL is parallel to 𝖢\mathsf{C}.

Proof.

After a change of coordinates in 3\mathbb{R}^{3}, we can suppose that the axis LL of MM is the zz-axis. Let 𝖢=ax+by+cz\mathsf{C}=a\partial_{x}+b\partial_{y}+c\partial_{z}, for a,b,ca,b,c\in\mathbb{R}. Let also γ:I3\gamma\colon I\subset\mathbb{R}\to\mathbb{R}^{3} be the generating curve of MM which we can assume that it is contained in the xzxz-plane, namely,

γ(s)=(x(s),0,z(s)),sI.\gamma(s)=(x(s),0,z(s)),\quad s\in I\subset\mathbb{R}.

We also assume that γ\gamma is parametrized by arc-length, that is, x2+z2=1x^{\prime 2}+z^{\prime 2}=1. Let κ=xz′′zx′′\kappa=x^{\prime}z^{\prime\prime}-z^{\prime}x^{\prime\prime} be its Frenet curvature with respect to the induced Levi-Civita connection ~0\widetilde{\nabla}^{0}. A parametrization of MM is

ψ(s,t)=(x(s)cost,x(s)sint,z(s)),sI,t.\psi(s,t)=(x(s)\cos t,x(s)\sin t,z(s)),\quad s\in I,t\in\mathbb{R}.

For the computation of KK, we calculate all terms of (6). The tangent plane of MM is spanned by {e1,e2}={ψs,ψt}\{e_{1},e_{2}\}=\{\psi_{s},\psi_{t}\}, where

e1=(xcost,xsint,z),e2=(xsint,xcost,0).\begin{split}e_{1}&=(x^{\prime}\cos t,x^{\prime}\sin t,z^{\prime}),\\ e_{2}&=(-x\sin t,x\cos t,0).\end{split}

The coefficients of the first fundamental form are g11=1g_{11}=1, g12=0g_{12}=0 and g22=x2g_{22}=x^{2}. The unit normal vector of MM is

N=(zcost,zsint,x).N=(-z^{\prime}\cos t,-z^{\prime}\sin t,x^{\prime}).

Then it is immediate

(11) 𝖢,N=azcostbzsint+cx,H=12x(z+xκ),G=zκx.\begin{split}\langle\mathsf{C},N\rangle&=-az^{\prime}\cos t-bz^{\prime}\sin t+cx^{\prime},\\ H&=\frac{1}{2x}(z^{\prime}+x\kappa),\\ G&=\frac{z^{\prime}\kappa}{x}.\end{split}

We now calculate K~\widetilde{K}. For this we employ the definition (3) taking into account that now the denominator is 2(g11g22g122)=2x22(g_{11}g_{22}-g_{12}^{2})=2x^{2}. We begin computing the covariant derivatives ~eiej\widetilde{\nabla}_{e_{i}}e_{j}, 1i,j21\leq i,j\leq 2. From (3), we have

~e1e1\displaystyle\widetilde{\nabla}_{e_{1}}e_{1} =\displaystyle= ψss+𝖢,e1e1,\displaystyle\psi_{ss}+\langle\mathsf{C},e_{1}\rangle e_{1},
~e1e2\displaystyle\widetilde{\nabla}_{e_{1}}e_{2} =\displaystyle= ψst+𝖢,e2e1,\displaystyle\psi_{st}+\langle\mathsf{C},e_{2}\rangle e_{1},
~e2e1\displaystyle\widetilde{\nabla}_{e_{2}}e_{1} =\displaystyle= ψst+𝖢,e1e2,\displaystyle\psi_{st}+\langle\mathsf{C},e_{1}\rangle e_{2},
~e2e2\displaystyle\widetilde{\nabla}_{e_{2}}e_{2} =\displaystyle= ψtt+𝖢,e2e2.\displaystyle\psi_{tt}+\langle\mathsf{C},e_{2}\rangle e_{2}.

Similarly, the covariant derivatives of second order are calculated. We obtain

~e1~e2e2\displaystyle\widetilde{\nabla}_{e_{1}}\widetilde{\nabla}_{e_{2}}e_{2} =\displaystyle= (~e2e2)s+𝖢,~e2e2e1,\displaystyle(\widetilde{\nabla}_{e_{2}}e_{2})_{s}+\langle\mathsf{C},\widetilde{\nabla}_{e_{2}}e_{2}\rangle e_{1},
~e2~e1e2\displaystyle\widetilde{\nabla}_{e_{2}}\widetilde{\nabla}_{e_{1}}e_{2} =\displaystyle= (~e1e2)t+𝖢,~e1e2e2,\displaystyle(\widetilde{\nabla}_{e_{1}}e_{2})_{t}+\langle\mathsf{C},\widetilde{\nabla}_{e_{1}}e_{2}\rangle e_{2},
~e2~e1e1\displaystyle\widetilde{\nabla}_{e_{2}}\widetilde{\nabla}_{e_{1}}e_{1} =\displaystyle= (~e1e1)t+𝖢,~e1e1e2,\displaystyle(\widetilde{\nabla}_{e_{1}}e_{1})_{t}+\langle\mathsf{C},\widetilde{\nabla}_{e_{1}}e_{1}\rangle e_{2},
~e1~e2e1\displaystyle\widetilde{\nabla}_{e_{1}}\widetilde{\nabla}_{e_{2}}e_{1} =\displaystyle= (~e2e1)s+𝖢,~e2e1e1.\displaystyle(\widetilde{\nabla}_{e_{2}}e_{1})_{s}+\langle\mathsf{C},\widetilde{\nabla}_{e_{2}}e_{1}\rangle e_{1}.

Obviously, [e1,e2]=0[e_{1},e_{2}]=0. The curvature R~\widetilde{R} is

R~(e1,e2,e2,e1)=(~e2e2)s(~e1e2)t,e1+𝖢,~e2e2=x2(bcostasint)2,R~(e2,e1,e1,e2)=(~e1e1)t(~e2e1)s,e2+x2𝖢,~e1e1=x2(x(acost+bsint)+cz)2.\begin{split}\widetilde{R}(e_{1},e_{2},e_{2},e_{1})&=\langle(\widetilde{\nabla}_{e_{2}}e_{2})_{s}-(\widetilde{\nabla}_{e_{1}}e_{2})_{t},e_{1}\rangle+\langle\mathsf{C},\widetilde{\nabla}_{e_{2}}e_{2}\rangle\\ &=x^{2}(b\cos t-a\sin t)^{2},\\ \widetilde{R}(e_{2},e_{1},e_{1},e_{2})&=\langle(\widetilde{\nabla}_{e_{1}}e_{1})_{t}-(\widetilde{\nabla}_{e_{2}}e_{1})_{s},e_{2}\rangle+x^{2}\langle\mathsf{C},\widetilde{\nabla}_{e_{1}}e_{1}\rangle\\ &=x^{2}\left(x^{\prime}(a\cos t+b\sin t)+cz^{\prime}\right)^{2}.\end{split}

This gives

K~=R~(e1,e2,e2,e1)+R~(e2,e1,e1,e2)2x2=12((bcostasint)2+(x(acost+bsint)+cz)2).\begin{split}\widetilde{K}&=\frac{\widetilde{R}(e_{1},e_{2},e_{2},e_{1})+\widetilde{R}(e_{2},e_{1},e_{1},e_{2})}{2x^{2}}\\ =&\frac{1}{2}\left((b\cos t-a\sin t)^{2}+(x^{\prime}(a\cos t+b\sin t)+cz^{\prime})^{2}\right).\end{split}

Finally, using (6), we obtain

K=12((bcostasint)2+(x(acost+bsint)+cz)2)+G(cxz(acost+bsint))H.\begin{split}K&=\frac{1}{2}\left((b\cos t-a\sin t)^{2}+(x^{\prime}(a\cos t+b\sin t)+cz^{\prime})^{2}\right)\\ &+G-(cx^{\prime}-z^{\prime}(a\cos t+b\sin t))H.\end{split}

The above expression can be written as a polynomial equation of type

n=02(An(s)cos(nt)+Bn(s)sin(nt))=0.\sum_{n=0}^{2}(A_{n}(s)\cos(nt)+B_{n}(s)\sin(nt))=0.

Since the functions {cos(nt),sin(nt)}\{\cos(nt),\sin(nt)\}, 0n20\leq n\leq 2, are linearly independent, then all coefficients AnA_{n} must vanish identically. The computation of these coefficients yields

A2=(b2a2)z24,B2=12abz2,A1=az(H+cx),B1=bz(H+cx),A0=14(a2+b2+2c2)z2+12(a2+b2)x2cHxK+G.\begin{split}A_{2}&=\frac{(b^{2}-a^{2})z^{\prime 2}}{4},\\ B_{2}&=-\frac{1}{2}abz^{\prime 2},\\ A_{1}&=az^{\prime}(H+cx^{\prime}),\\ B_{1}&=bz^{\prime}(H+cx^{\prime}),\\ A_{0}&=\frac{1}{4}\left(a^{2}+b^{2}+2c^{2}\right)z^{\prime 2}+\frac{1}{2}\left(a^{2}+b^{2}\right)x^{\prime 2}-cHx^{\prime}-K+G.\end{split}

From A2=0A_{2}=0 and B2=0B_{2}=0, we have the following discussion of cases.

  1. (1)

    Case z=0z^{\prime}=0 identically. Then zz is a constant function and this implies that MM is a horizontal plane. In particular, x2=1x^{\prime 2}=1. Without loss of generality, we suppose x(s)=sx(s)=s. Since G=H=0G=H=0, equation A0=0A_{0}=0 is simply

    K=12(a2+b2).K=\frac{1}{2}(a^{2}+b^{2}).

    This proves the result in this case.

  2. (2)

    Case that z(s)0z^{\prime}(s)\not=0 at some value ss. Then z0z^{\prime}\not=0 around ss and A2=B2=0A_{2}=B_{2}=0 implies a=b=0a=b=0. Thus 𝖢=±z\mathsf{C}=\pm\partial_{z} and this proves that 𝖢\mathsf{C} is parallel to the zz-axis, which it is the rotation axis of MM.

Once proved Thm. 4.1, we can suppose that the vector field 𝖢\mathsf{C} is z\partial_{z} and MM is a rotational surface about the zz-axis. Following the proof of that theorem, all coefficients AnA_{n} and BnB_{n}, 1n21\leq n\leq 2 are trivially 0 except A0A_{0} which it is

K=z22+GxH.K=\frac{z^{\prime 2}}{2}+G-x^{\prime}H.

Using the value of GG and HH given in (11), the above equation gives us the expression of KK of a rotational surface in terms of its generating curve, namely,

(12) K=12x((2zxx)κ+z(xzx)).K=\frac{1}{2x}((2z^{\prime}-xx^{\prime})\kappa+z^{\prime}(xz^{\prime}-x^{\prime})).

We study when the parenthesis of the right hand-side of (12) are 0 identically.

Proposition 4.2.

If KK is constant in (12), then the functions 2zxx2z^{\prime}-xx^{\prime} and xzxxz^{\prime}-x^{\prime} cannot vanish identically in II.

Proof.
  1. (1)

    Case 2zxx=02z^{\prime}-xx^{\prime}=0. Then neither xx^{\prime} nor zz^{\prime} can vanish identically. From (12), we have

    x2=8Kx22.x^{\prime 2}=\frac{8K}{x^{2}-2}.

    Since x2+z2=1x^{\prime 2}+z^{\prime 2}=1, then

    x2=44+x2.x^{\prime 2}=\frac{4}{4+x^{2}}.

    Combining both equations, we get 8K+(2K1)x2+2=08K+(2K-1)x^{2}+2=0, then xx is a constant function, which it is a contradiction.

  2. (2)

    Case xzx=0xz^{\prime}-x^{\prime}=0. Since z=x/xz^{\prime}=x^{\prime}/x, then κ=x3x2\kappa=-\frac{x^{\prime 3}}{x^{2}}. Thus (12) is

    2Kx2=x4(x2).2Kx^{2}=x^{\prime 4}(x-2).

    On the other hand, it follows x2+z2=1x^{\prime 2}+z^{\prime 2}=1 that

    x2=x21+x2.x^{\prime 2}=\frac{x^{2}}{1+x^{2}}.

    Combining both equations we obtain that x=x(s)x=x(s) is a constant function. From xzx=0xz^{\prime}-x^{\prime}=0, we have zz constant too, which it is a contradiction by regularity of γ\gamma.

In the following two results we study the case when the generating curve γ\gamma of MM has constant curvature κ\kappa, that is, γ\gamma is a straight-line and a circle. First, suppose that γ\gamma is a straight-line. This implies that MM is a conical rotational surface.

Theorem 4.3.

Let ~\widetilde{\nabla} be a canonical snm-connection and MM be a rotational surface about the zz-axis. Assume that the sectional curvature KK of MM with respect to ~\widetilde{\nabla} is constant. If the generating curve of MM is a straight-line, then either MM is a circular cylinder and K=1/2K=1/2, or MM is a horizontal plane and K=0K=0.

Proof.

We follow the notation of Thm. 4.1. Since γ\gamma is parametrized by arc-length, then there is a real number θ\theta\in\mathbb{R} such that γ\gamma can be written as

γ(s)=(c1,c2)+(cosθ,sinθ)s,c1,c2.\gamma(s)=(c_{1},c_{2})+(\cos\theta,\sin\theta)s,\quad c_{1},c_{2}\in\mathbb{R}.

Equation (12) is now

2K(scosθ+c1)sin2θ(scosθ+c1)+sinθcosθ=0.2K(s\cos\theta+c_{1})-\sin^{2}\theta(s\cos\theta+c_{1})+\sin\theta\cos\theta=0.

This is a polynomial equation on ss, so all coefficients must vanish. Therefore

(2Ksin2θ)cosθ\displaystyle(2K-\sin^{2}\theta)\cos\theta =\displaystyle= 0,\displaystyle 0,
(2Ksin2θ)c1+sinθcosθ\displaystyle(2K-\sin^{2}\theta)c_{1}+\sin\theta\cos\theta =\displaystyle= 0.\displaystyle 0.
  1. (1)

    Case cosθ=0\cos\theta=0. Then γ(s)=(c1,±s+c2)\gamma(s)=(c_{1},\pm s+c_{2}). In particular c1>0c_{1}>0. This implies that MM is a circular cylinder of radius c1\sqrt{c_{1}}. The second equation gives K=1/2K=1/2.

  2. (2)

    Case cosθ0\cos\theta\neq 0. Then 2Ksin2θ=02K-\sin^{2}\theta=0 and the second equation gives sinθ=0\sin\theta=0. Thus γ(s)=(±s+c1,c2)\gamma(s)=(\pm s+c_{1},c_{2}) and MM is a horizontal plane of equation z=c2z=c_{2}. Here K=0K=0.

Finally, we suppose that γ\gamma is a circle. This implies that MM is torus of revolution or a rotational ovaloid.

Theorem 4.4.

Let ~\widetilde{\nabla} be a canonical snm-connection and MM be a rotational surface about the zz-axis. Assume that the sectional curvature KK of MM with respect to ~\widetilde{\nabla} is constant. Then the generating curve of MM cannot be a circle.

Proof.

By contradiction, suppose that γ\gamma is a circle of radius r>0r>0. A parametrization of γ\gamma is

γ(s)=(c1,c2)+r(cos(s/r),sin(s/r)).\gamma(s)=(c_{1},c_{2})+r\left(\cos(s/r),\sin(s/r)\right).

Substituting into (12), we obtain

2K(c1+rcos(s/r))1r(2cos(s/r)+(c1+rcos(s/r))sin(s/r))cos(s/r)((c1+rcos(s/r))cos(s/r)+sin(s/r))=0.\begin{split}2K(c_{1}+r\cos(s/r))&-\frac{1}{r}\left(2\cos(s/r)+(c_{1}+r\cos(s/r))\sin(s/r)\right)\\ &-\cos(s/r)\left((c_{1}+r\cos(s/r))\cos(s/r)+\sin(s/r)\right)=0.\end{split}

This equation writes as

n=03(Ancos(s/r)+Bnsin(s/r))=0,\sum_{n=0}^{3}(A_{n}\cos(s/r)+B_{n}\sin(s/r))=0,

where AnA_{n} and BnB_{n} are real constants. Since all AnA_{n} and BnB_{n} must 0, a computation gives A3=r3A_{3}=-\frac{r}{3}, obtaining a contradiction. ∎

The study of solutions of (12) is difficult to do in all its generality and Thms. 4.3 and 4.4 are the first results. An interesting value for KK is 1/21/2 because this is the curvature of a circular cylinder (for any radius) and that of a plane parallel to z\partial_{z}. If K=1/2K=1/2, then Eq. (12) is

(13) κ=x(x+z)2zxx.\kappa=\frac{x^{\prime}(x+z^{\prime})}{2z^{\prime}-xx^{\prime}}.

An interesting question is if this equation has a solution for curves starting orthogonally from the rotation axis. If s=0s=0 is the time where γ\gamma intersects the zz-axis, then we need x(0)=0x(0)=0 and z(0)=0z^{\prime}(0)=0. However, the left hand-side of (13) is not defined at s=0s=0. This implies that existence of such solutions is not assured. We prove that these solutions, indeed, exist.

Theorem 4.5.

There exist rotational surfaces with constant sectional curvature K=1/2K=1/2 intersecting orthogonally the rotation axis.

Proof.

For our convenience, we work assuming that γ\gamma is locally a graph z=z(x)z=z(x). Then

x(s)=11+z(x)2,z(s)=z(x)1+z(x)2,κ=z′′(x)(1+z(x)2)3/2.x^{\prime}(s)=\frac{1}{\sqrt{1+z^{\prime}(x)^{2}}},\quad z^{\prime}(s)=\frac{z^{\prime}(x)}{\sqrt{1+z^{\prime}(x)^{2}}},\quad\kappa=\frac{z^{\prime\prime}(x)}{(1+z^{\prime}(x)^{2})^{3/2}}.

Then (13) becomes

z′′(1+z2)3/2=x1+z2+z(2zx)1+z2,\frac{z^{\prime\prime}}{(1+z^{\prime 2})^{3/2}}=\frac{x\sqrt{1+z^{\prime 2}}+z^{\prime}}{(2z^{\prime}-x)\sqrt{1+z^{\prime 2}}},

or equivalently,

(2zx)z′′(1+z2)3/2=x1+z2+z1+z2.(2z^{\prime}-x)\frac{z^{\prime\prime}}{(1+z^{\prime 2})^{3/2}}=\frac{x\sqrt{1+z^{\prime 2}}+z^{\prime}}{\sqrt{1+z^{\prime 2}}}.

This equation also writes as

ddx((2zx)z1+z2)=ddx(21+z2+x22).\frac{d}{dx}\left((2z^{\prime}-x)\frac{z^{\prime}}{\sqrt{1+z^{\prime 2}}}\right)=\frac{d}{dx}(2\sqrt{1+z^{\prime 2}}+\frac{x^{2}}{2}).

Thus there is an integration constant cc\in\mathbb{R} such that

(2zx)z1+z2=21+z2+x22+c.(2z^{\prime}-x)\frac{z^{\prime}}{\sqrt{1+z^{\prime 2}}}=2\sqrt{1+z^{\prime 2}}+\frac{x^{2}}{2}+c.

If γ\gamma intersects orthogonally the zz-axis, then we have z(0)=0z^{\prime}(0)=0. This gives c=2c=-2, obtaining a first integration of (12), namely,

(2zx)z1+z2=2(1+z21)+x22.(2z^{\prime}-x)\frac{z^{\prime}}{\sqrt{1+z^{\prime 2}}}=2(\sqrt{1+z^{\prime 2}}-1)+\frac{x^{2}}{2}.

Squaring both sides of this equation, we get

(4x2(x24)2)z2+16xz+16(x24)2=0.(4x^{2}-(x^{2}-4)^{2})z^{\prime 2}+16xz^{\prime}+16-(x^{2}-4)^{2}=0.

By standard theory of existence of ODE, this equation has a solution with initial value z(0)=0z^{\prime}(0)=0, proving the result. ∎

Acknowledgements

Rafael López is a member of the IMAG and of the Research Group “Problemas variacionales en geometría”, Junta de Andalucía (FQM 325). This research has been partially supported by MINECO/MICINN/FEDER grant no. PID2020-117868GB-I00, and by the “María de Maeztu” Excellence Unit IMAG, reference CEX2020-001105-M, funded by MCINN/AEI/10.13039/501100011033/ CEX2020-001105-M.

References

  • [1] N. S. Agashe, A semi-symmetric non-metric connection on a Riemannian manifold. Indian J. Pure Appl. Math. 23 (1992), 399–409.
  • [2] N. S. Agashe, M. R. Chafle, On submanifolds of a Riemannian manifold with a semi-symmetric non-metric connection. Tensor 55 (1994), 120–130.
  • [3] B.-Y. Chen, Total mean curvature and submanifolds of finite type. Second edition. With a foreword by Leopold Verstraelen. Series in Pure Mathematics, 27. World Scientific Publishing Co. Pte. Ltd., Hackensack (2015).
  • [4] A. Friedmann, J. A. Schouten, Über die Geometrie der halbsymmetrischen Übertragungen. Math. Z. 21 (1924), 211–223.
  • [5] H. P. Halldorsson, Self-similar solutions to the curve shortening flow. Trans. Amer. Math. Soc. 364 (2012), 5285–5309.
  • [6] H. Hayden, Subspaces of a space with torsion. Proc. London Math. Soc. 34 (1932), 27–50.
  • [7] T. Imai, Hypersurfaces of a Riemannian manifold with semi-symmetric metric connection. Tensor (N. S.) 23 (1972), 300–306.
  • [8] C. W. Lee, D. W. Yoon, J. W. Lee, Optimal inequalities for the Casorati curvatures of submanifolds of real space forms endowed with semi-symmetric metric connections. J. Inequal. Appl. 2014, 327 (2014).
  • [9] A. Mihai, C. Özgür, Chen inequalities for submanifolds of real space forms with a semi-symmetric metric connection. Taiwan. J. Math. 4 (2010), 1465–1477.
  • [10] A. Mihai and C. Özgür, Chen inequalities for submanifolds of complex space forms and Sasakian space forms endowed with semi-symmetric metric connections. Rocky Mountain J. Math. 41 (2011), 1653–1673.
  • [11] A. Mihai, I. Mihai, A note on a well-defined sectional curvature of a semi-symmetric non-metric connection. Int. Electron. J. Geom. 17(1) (2024), 15-23.
  • [12] Z. Nakao, Submanifolds of a Riemannian manifold with semisymmetric metric connections. Proc. Amer. Math. Soc. 54 (1976), 261–266.
  • [13] Y. Wang, Minimal translation surfaces with respect to semi-symmetric connections in R3R^{3} and R13R^{3}_{1}. Bull. Korean Math. Soc. 58 (2021), 959–972.
  • [14] K. Yano, On semi symmetric metric connection. Rev. Roum. Math. Pures Appl. 15 (1970), 1579–1591.