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

Translators of the Gauss curvature flow

Muhittin Evren Aydin Department of Mathematics, Faculty of Science. Firat University. Elazig 23200 Turkey [email protected]  and  Rafael López Departamento de Geometría y Topología. Universidad de Granada. Avenida Fuentenueva, s/n. Granada 18071 Spain [email protected]
Abstract.

A KαK^{\alpha}-translator is a surface in Euclidean space 3\mathbb{R}^{3} that moves by translations in a spatial direction and under the KαK^{\alpha}-flow, where KK is the Gauss curvature and α\alpha is a constant. We classify all KαK^{\alpha}-translators that are rotationally symmetric. In particular, we prove that for each α\alpha there is a KαK^{\alpha}-translator intersecting orthogonally the rotation axis. We also describe all KαK^{\alpha}-translators invariant by a uniparametric group of helicoidal motions and the translators obtained by separation of variables.

Key words and phrases:
KαK^{\alpha}-translator, Darboux surface, surface of revolution, separation of variables.
1991 Mathematics Subject Classification:
53C44, 53A15, 35J96

1. Introduction and results

The flow by powers of the Gauss curvature KK was initiated by Chow [8] after the articles of Firey and Tso ([11, 15]). These works were the starting point of the theory of the flow by the Gaussian curvature ([1, 16]), a topic of high activity in geometric analysis that continues to the present. Given a smooth immersion X:Σ3X:\Sigma\to\mathbb{R}^{3} of a strictly convex surface Σ\Sigma in Euclidean space 3\mathbb{R}^{3}, we consider the KαK^{\alpha}-flow as a one-parameter family of smooth immersions Xt=X(,t):Σ3X_{t}=X(\cdot,t)\colon\Sigma\to\mathbb{R}^{3}, t[0,T)t\in[0,T) such that X0=XX_{0}=X and satisfying the flow

tX(p,t)=K(p,t)αN(p,t),(p,t)Σ×[0,T),\frac{\partial}{\partial t}X(p,t)=-K(p,t)^{\alpha}N(p,t),\quad(p,t)\in\Sigma\times[0,T),

where α\alpha\in\mathbb{R} is a constant, N(p,t)N(p,t) is the unit normal of X(p,t)X(p,t) and K(p,t)K(p,t) is the Gauss curvature at X(p,t)X(p,t). An interesting problem is the evolution of a surface through the flow. As an example for the reader, if α=1\alpha=1 the surface becomes spherical ([3]).

If the surface moves under the KαK^{\alpha}-flow along a spatial direction v3\vec{v}\in\mathbb{R}^{3}, then X0X_{0} satisfies Kα=λN,vK^{\alpha}=\lambda\langle N,\vec{v}\rangle for some positive constant λ\lambda. The vector v\vec{v} is assumed to be unitary and it is called the speed of the flow. After a dilation, we can assume that λ=1\lambda=1. A surface Σ\Sigma is called a translator by the KαK^{\alpha}-flow with speed v\vec{v}, or simply, a KαK^{\alpha}-translator in case that v\vec{v} is understood, if

Kα=N,v.K^{\alpha}=\langle N,\vec{v}\rangle. (1)

The notion of translator by positive powers of the Gauss curvature appeared in [17]. See also [9, 14].

In this paper we want to find examples of KαK^{\alpha}-translators under the geometric condition that the surface is defined kinematically as the movement of a curve by a uniparametric family of rigid motions of 3\mathbb{R}^{3}. Following Darboux [10, Livre I], we consider surfaces parametrized by Ψ(s,t)=A(t)γ(s)+β(t)\Psi(s,t)=A(t)\cdot\gamma(s)+\beta(t), where γ\gamma and β\beta are two spatial curves and A(t)A(t) is an orthogonal matrix. The cases that we are interesting are: rotational surfaces (A(t)A(t) is a uniparametric group of rotations and β\beta is constant), helicoidal surfaces (A(t)A(t) is a uniparametric group of helicoidal motions), translation surfaces (A(t)A(t) is the identity) and ruled surfaces (γ\gamma is a straight-line). Since we have left to one side the study of the evolution of the surface, we do not require that KK to be positive but only that KαK^{\alpha} has sense. For example, KK may be negative if α\alpha\in\mathbb{Z}. This will be implicitly assumed in all the results.

In the flow by powers of the Gaussian curvature, it is known that the K1/4K^{1/4}-flow is special because it has an interpretation in affine differential geometry as the affine normal flow of a convex surface ([2, 6]). However, in base of our proofs, this case will appear as natural because when we express the equation (1) for each one of the above three types of surfaces and when α\alpha is exactly 1/41/4, this equation is greatly simplified due to a cancellation of terms. This makes the arguments easier than in the general case of α\alpha.

Planes are examples of KαK^{\alpha}-translators for any α>0\alpha>0 provided the plane is parallel to the speed of the flow. Throughout this paper, we will discard planes as examples of KαK^{\alpha}-translators. Other examples that we will not consider in the class of KαK^{\alpha}-translators are the surfaces whose Gauss curvature is constant, where now the equation (1) says that the unit normal vector NN makes a constant angle with a fixed direction. Similarly, we discard the case α=0\alpha=0.

It is important to point out that there is not a priori relation between the speed vector v\vec{v} and the special parametrization of each of the above Darboux surfaces. So, if one considers the study of translators one can prescribed the speed v\vec{v}, usually, the vector (0,0,1)(0,0,1) in the literature. However, if one considers one of the above types of Darboux surfaces, the parametrization has no relation with v\vec{v}. This can be clearly seen for rotational surfaces (also helicoidal surfaces). The rotation axis of the surface is, initially, independent of the vector v\vec{v}. However, as we will see, for rotational and helicoidal surfaces, the speed must be parallel to the axis (Propositions 2.1 and 3.1). Something similar occurs for translations surfaces where, if one prescribes that the surface is z=f(x)+g(y)z=f(x)+g(y), the speed may be arbitrary.

The organization of this paper is as follows. In Section 2, we obtain all rotational KαK^{\alpha}-translators. Recall that in [17], Urbas obtains these surfaces for α(0,1/2]\alpha\in(0,1/2] using the Legendre transform (see also [13] for a similar calculation). Our approach uses simple geometric arguments and it holds for any α\alpha (Theorem 2.2). In particular, for each value of α\alpha, we prove the existence of rotational examples intersecting orthogonally the rotation axis (Corollary 2.4). Section 3 is devoted to the study of KαK^{\alpha}-translators of helicoidal type. Although we are not able to obtain explicit parametrizations of these surfaces, we do a first integration of the generating curve (Theorem 3.2), also in terms of the Bour function (Theorem 3.3). In Section 4, we obtain the classification of all solutions of (1) obtained by separation of variables z=f(x)+g(y)z=f(x)+g(y), where (x,y,z)(x,y,z) are the canonical coordinate system of 3\mathbb{R}^{3}. Although these solutions depend on a particular choice of coordinates of 3\mathbb{R}^{3}, our result holds for any speed vector v\vec{v}. In particular, we prove that there are KαK^{\alpha}-translators only if α=1/4\alpha=1/4 (Theorems 4.1 and 4.3). Besides we provide new examples of K1/4K^{1/4}-translators of type z=f(x)+g(y)z=f(x)+g(y), we also give new examples of K1/4K^{1/4}-translators obtained by separation of variables of type z=f(x)g(y)z=f(x)g(y) (Example 4.4). Finally, in Section 5, we investigate the existence of ruled KαK^{\alpha}-translators, proving that there are not ruled KαK^{\alpha}-translators, except trivial cases (Theorem 5.1).

The authors have extended the results of this paper for (spacelike and timelike) KαK^{\alpha}-translators in Lorentz-Minkowski space ([4]).

2. Rotational KαK^{\alpha}-translators

In this section, we classify the surfaces of revolution satisfying (1). A first question is if there is a relation between the rotation axis of the surface and the speed vector v\vec{v}. As it is expectable, we prove that the vector v\vec{v} must be parallel to the rotation axis.

Proposition 2.1.

Let Σ\Sigma be a KαK^{\alpha}-translator. If Σ\Sigma is a surface of revolution, then the rotation axis is parallel to the speed vector v\vec{v}.

Proof.

After a rigid motion, we can assume that the rotation axis LL is the zz-axis. The generating curve of a rotational surface is a curve included in the coordinate xzxz-plane which can be assumed that it is a graph on the xx-axis, except in the case that this curve is a straight-line parallel to the zz-axis. In this particular situation, the surface is a circular cylinder where we know that K=0K=0. Thus a circular cylinder satisfies (1) if N,v=0\langle N,\vec{v}\rangle=0, that is, the vector vv is parallel to the zz-axis, proving the result for this particular case.

Suppose now the general case that the generating curve writes as z=f(x)z=f(x) where f:I+f\colon I\subset\mathbb{R}^{+}\to\mathbb{R} is a smooth function. Then a parametrization of Σ\Sigma is

X(r,θ)=(rcosθ,rsinθ,f(r)),X(r,\theta)=(r\cos\theta,r\sin\theta,f(r)), (2)

where θ\theta\in\mathbb{R}. The unit normal vector NN is

N(r,θ)=11+f2(fcosθ,fsinθ,1)N(r,\theta)=\frac{1}{\sqrt{1+f^{\prime 2}}}(-f^{\prime}\cos\theta,-f^{\prime}\sin\theta,1)

and the principal curvatures of Σ\Sigma are

κ1=f′′(1+f2)3/2,κ2=fr(1+f2)1/2.\kappa_{1}=\frac{f^{\prime\prime}}{(1+f^{\prime 2})^{3/2}},\quad\kappa_{2}=\frac{f^{\prime}}{r(1+f^{\prime 2})^{1/2}}.

If v=(v1,v2,v3)\vec{v}=(v_{1},v_{2},v_{3}), then (1) writes as

Kα=(κ1κ2)α=(ff′′r(1+f2)2)α=11+f2(v1fcosθv2fsinθ+v3),K^{\alpha}=(\kappa_{1}\kappa_{2})^{\alpha}=\left(\frac{f^{\prime}f^{\prime\prime}}{r(1+f^{\prime 2})^{2}}\right)^{\alpha}=\frac{1}{\sqrt{1+f^{\prime 2}}}(-v_{1}f^{\prime}\cos\theta-v_{2}f^{\prime}\sin\theta+v_{3}),

or equivalently

A0+A1cosθ+A2sinθ=0,A_{0}+A_{1}\cos\theta+A_{2}\sin\theta=0,

where

A0=(ff′′r(1+f2)2)αv3(1+f2)1/2,A1=v1f(1+f2)1/2,A2=v2f(1+f2)1/2.\begin{split}A_{0}&=\left(\frac{f^{\prime}f^{\prime\prime}}{r(1+f^{\prime 2})^{2}}\right)^{\alpha}-v_{3}(1+f^{\prime 2})^{-1/2},\\ A_{1}&=v_{1}f^{\prime}(1+f^{\prime 2})^{-1/2},\quad A_{2}=v_{2}f^{\prime}(1+f^{\prime 2})^{-1/2}.\end{split}

Since the functions {1,cosθ,sinθ}\{1,\cos\theta,\sin\theta\} are linearly independent, we conclude A0=A1=A2=0A_{0}=A_{1}=A_{2}=0. If ff^{\prime} is constantly 0, then ff is constant and the surface is a plane. This case was initially discarded. Therefore from A1=A2=0A_{1}=A_{2}=0, we find that v1=v2=0v_{1}=v_{2}=0, concluding that v\vec{v} is parallel to LL. ∎

After a rigid motion, we will assume that the rotation axis is the zz-axis and consequently from this proposition, that v=(0,0,1)\vec{v}=(0,0,1) after a symmetry about the xyxy-plane if necessary. Then (1) is

(ff′′r(1+f2)2)α=1(1+f2)1/2.\left(\frac{f^{\prime}f^{\prime\prime}}{r(1+f^{\prime 2})^{2}}\right)^{\alpha}=\frac{1}{(1+f^{\prime 2})^{1/2}}.

Set g=f/(1+f2)1/2g=f^{\prime}/(1+f^{\prime 2})^{1/2}. Notice that κ1=g\kappa_{1}=g^{\prime} and rκ2=gr\kappa_{2}=g. In terms of gg, the above equation becomes

g(1g2)12αg=r.\frac{g}{(1-g^{2})^{\frac{1}{2\alpha}}}g^{\prime}=r.

A first integration gives

g2={1mer2,m>0,α=121(m2α12αr2)2α2α1,m,α12.g^{2}=\left\{\begin{array}[]{lll}1-me^{-r^{2}},m>0,&&\alpha=\frac{1}{2}\\ 1-\left(m-\frac{2\alpha-1}{2\alpha}r^{2}\right)^{\frac{2\alpha}{2\alpha-1}},m\in\mathbb{R},&&\alpha\neq\frac{1}{2}.\end{array}\right. (3)

In the case α=1/2\alpha=1/2, the condition g20g*2\geq 0 says that the domain of rr is when r2log(m)r^{2}\geq\log(m). Otherwise, we need to distinguish if 2α/(2α1)2\alpha/(2\alpha-1) is negative or positive, that is, α\alpha belongs to (0,1/2)(0,1/2) or not. If α(0,1/2)\alpha\in(0,1/2), the parenthesis in (3) must be positive, yielding r2>2α/(2α1)mr^{2}>2\alpha/(2\alpha-1)m. On the other hand, using now that g20g^{2}\geq 0, we have r22α/(2α1)(m1)0r^{2}\geq 2\alpha/(2\alpha-1)(m-1)\geq 0, so this is the restriction on rr because 2α/(2α1)<02\alpha/(2\alpha-1)<0. If α∫̸[0,1/2]\alpha\not\int[0,1/2], then 2α/(2α1)>02\alpha/(2\alpha-1)>0. Since the parenthesis in (3) must be positive because g2<1g^{2}<1, then we obtain r2<2α/(2α1)mr^{2}<2\alpha/(2\alpha-1)m and from the fact that g20g^{2}\geq 0, the restriction is r22α/(2α1)(m1)r^{2}\geq 2\alpha/(2\alpha-1)(m-1). To summarize, we have

{2α2α1(m1)r2,α(0,12)2α2α1(m1)r2<2α2α1m,α[0,12].\left\{\begin{array}[]{lll}\frac{2\alpha}{2\alpha-1}(m-1)\leq r^{2},&&\alpha\in(0,\frac{1}{2})\\ \frac{2\alpha}{2\alpha-1}(m-1)\leq r^{2}<\frac{2\alpha}{2\alpha-1}m,&&\alpha\not\in[0,\frac{1}{2}].\end{array}\right. (4)

Hence we deduce

f={±(1mer21)1/2,α=12±((m2α12αr2)2α12α1)1/2,α12.f^{\prime}=\left\{\begin{array}[]{ll}\pm\left(\frac{1}{m}e^{r^{2}}-1\right)^{1/2},&\alpha=\frac{1}{2}\\ \pm\left(\left(m-\frac{2\alpha-1}{2\alpha}r^{2}\right)^{\frac{2\alpha}{1-2\alpha}}-1\right)^{1/2},&\alpha\neq\frac{1}{2}.\end{array}\right. (5)

As conclusion, we have the classification of all KαK^{\alpha}-translators that also are surfaces of revolution.

Theorem 2.2.

Let Σ\Sigma be a KαK^{\alpha}-translator. If Σ\Sigma is a surface of revolution about the zz-axis, then Σ\Sigma is a circular cylinder of arbitrary radius or Σ\Sigma parametrizes as (2) where

f(r)={±r(1met21)1/2𝑑t,m>0,α=12±r((m2α12αt2)2α12α1)1/2𝑑t,m,α12.f(r)=\left\{\begin{array}[]{lll}\pm\int^{r}\left(\frac{1}{m}e^{t^{2}}-1\right)^{1/2}\,dt,m>0,&&\alpha=\frac{1}{2}\\ \pm\int^{r}\left(\left(m-\frac{2\alpha-1}{2\alpha}t^{2}\right)^{\frac{2\alpha}{1-2\alpha}}-1\right)^{1/2}\,dt,m\in\mathbb{R},&&\alpha\neq\frac{1}{2}.\end{array}\right. (6)

Furthermore, the maximal domain of the function f(r)f(r) is

  1. (1)

    [logm,)[\sqrt{\log{m}},\infty), if α=1/2\alpha=1/2.

  2. (2)

    [2α2α1m,)[\sqrt{\frac{2\alpha}{2\alpha-1}m},\infty), if α(0,1/2)\alpha\in(0,1/2).

  3. (3)

    [2α2α1(m1),2α2α1m)[\sqrt{\frac{2\alpha}{2\alpha-1}(m-1)},\sqrt{\frac{2\alpha}{2\alpha-1}m}), if α[0,1/2]\alpha\not\in[0,1/2]. In this case, we have

    limr2α2α1(m1)f(r)=0,limr2α2α1mf(r)=.\lim_{r\to\sqrt{\frac{2\alpha}{2\alpha-1}(m-1)}}f^{\prime}(r)=0,\quad\lim_{r\to\sqrt{\frac{2\alpha}{2\alpha-1}m}}f(r)=\infty.

In all these cases, we understand that if in the radicand in the left-end of the interval is negative, then the value of this end is 0.

Proof.

It remains to prove the limits in the case (3). The first limit is consequence of (4) and (5). For the second limit, note that the maximal domain of ff is a bounded interval and that limr2α2α1mf(r)=\lim_{r\to\sqrt{\frac{2\alpha}{2\alpha-1}m}}f(r)=\infty by (5). ∎

We point that it is expectable that in the case α[0,1/2]\alpha\not\in[0,1/2] the domain cannot be [0,)[0,\infty) because there are no entire graphs that are KαK^{\alpha}-translators if α>1/2\alpha>1/2 ([17, Sect. 4]) and if α<0\alpha<0 ([18, Th. 6.1]). It deserves to note the case α=1/4\alpha=1/4 because it is possible to integrate explicitly (6) (see also [14]). See Figure 1.

Corollary 2.3.

Rotational K1/4K^{1/4}-translators form a uniparametric family of surfaces parametrized by (2), where

f(r)=12(rm+r21+(m1)log(m+r21+r))+c,m,c.f(r)=\frac{1}{2}\left(r\sqrt{m+r^{2}-1}+(m-1)\log\left(\sqrt{m+r^{2}-1}+r\right)\right)+c,\quad m,c\in\mathbb{R}.

The maximal domain of ff is [1m,)[\sqrt{1-m},\infty) if m<1m<1 and [0,)[0,\infty) if m1m\geq 1. For the value m=1m=1, ff is the parabola f(r)=r2/2f(r)=r^{2}/2, the graphic of f(r)f(r) intersects orthogonally the rotation axis and the surface is a paraboloid.

Refer to caption
Figure 1. Generating curves of rotational K1/4K^{1/4}-translators for different values of the parameter mm.

By (4), we point out that the maximal domain of ff is not [0,)[0,\infty) in general. However, an interesting case to investigate is if there are generating curves that meet orthogonally the rotation axis. We prove that this occurs for all cases of α\alpha. See Figure 2.

Corollary 2.4.

For each α\alpha, there are rotational KαK^{\alpha}-translators whose generating curves intersect orthogonally the rotation axis. These surfaces are unique up to vertical translations. Furthermore,

  1. (1)

    If α(0,12]\alpha\in(0,\frac{1}{2}], the maximal domain of ff is [0,)[0,\infty), limrf(r)=\lim_{r\to\infty}f(r)=\infty and

    f(r)=(12α)(12α2α)α12αr112α+o(r112α).f(r)=(1-2\alpha)\left(\frac{1-2\alpha}{2\alpha}\right)^{\frac{\alpha}{1-2\alpha}}r^{\frac{1}{1-2\alpha}}+o(r^{\frac{1}{1-2\alpha}}).
  2. (2)

    If α[0,12]\alpha\not\in[0,\frac{1}{2}], the maximal domain is [0,2α2α1)[0,\sqrt{\frac{2\alpha}{2\alpha-1}}), with

    limr2α2α1f(r)=.\lim_{r\to\sqrt{\frac{2\alpha}{2\alpha-1}}}f(r)=\infty.
Proof.

The condition on the orthogonality with the rotation axis requires that ff is defined at r=0r=0 and f(0)=0f^{\prime}(0)=0. From (5), it is immediate that mm must be 11 and the same occurs in the particular case α=1/2\alpha=1/2. This solution is C2C^{2} at r=0r=0 because from (5), we have limr0f′′(r)=1\lim_{r\to 0}f^{\prime\prime}(r)=1. Hence ff is CC^{\infty} in its domain by regularity ([5, 12]). The uniqueness is consequence of the solvability of (6).

The behaviour of ff at infinity is consequence of (5) and the L’Hôpital rule. Indeed, if δ=1/(12α)\delta=1/(1-2\alpha), then

limrf(r)rδ=limrf(r)δrδ1=1δ(12α2α)α12α.\lim_{r\to\infty}\frac{f(r)}{r^{\delta}}=\lim_{r\to\infty}\frac{f^{\prime}(r)}{\delta r^{\delta-1}}=\frac{1}{\delta}\left(\frac{1-2\alpha}{2\alpha}\right)^{\frac{\alpha}{1-2\alpha}}.

Refer to caption
Refer to caption
Figure 2. Generating curves of rotational KαK^{\alpha}-translators intersecting orthogonally the rotation axis. Left: α=1/2\alpha=1/2 and the maximal domain is [0,)[0,\infty). Right: α=1\alpha=1 and the maximal domain is [0,2)[0,\sqrt{2}).

We point out that in some particular cases, the integrals in (6) can be explicitly solved. Here, we denote by fαf_{\alpha} to emphasize the parameter α\alpha where we also assume m=1m=1.

  1. (1)

    Case α=1\alpha=1. Then

    f(r)=r4r22r2f^{\prime}(r)=\frac{r\sqrt{4-r^{2}}}{2-r^{2}}

    and

    f1(r)=4r2±2tanh1(124r2),f_{1}(r)=\mp\sqrt{4-r^{2}}\pm\sqrt{2}\tanh^{-1}\left(\frac{1}{\sqrt{2}}\sqrt{4-r^{2}}\right),

    defined on [0,2).[0,2).

  2. (2)

    Case α=1/3\alpha=1/3. Now

    f(r)=±12r4+r2f^{\prime}(r)=\pm\frac{1}{2}r\sqrt{4+r^{2}}

    and the solution is

    f1/3(r)=±16(4+r2)3/2,f_{1/3}(r)=\pm\frac{1}{6}\left(4+r^{2}\right)^{3/2},

    defined on [0,).[0,\infty).

  3. (3)

    For α=1/6\alpha=1/6, we have

    f(r)=±(2r2+11)1/2,f^{\prime}(r)=\pm(\sqrt{2r^{2}+1}-1)^{1/2},

    and

    f1/6(r)=±2r2+11(2r22r2+11)3r,f_{1/6}(r)=\pm\frac{\sqrt{\sqrt{2r^{2}+1}-1}\left(2r^{2}-\sqrt{2r^{2}+1}-1\right)}{3r},

    and defined on [0,)[0,\infty).

3. Helicoidal KαK^{\alpha}-translators

Consider a helicoidal surface Σ\Sigma in 3\mathbb{R}^{3} with axis zz whose generating curve γ\gamma is included in the xzxz-plane and pitch hh. Without loss of generality, we can assume that γ(r)=(r,0,f(r))\gamma(r)=(r,0,f(r)) where f:I+f\colon I\subset\mathbb{R}^{+}\to\mathbb{R} is a smooth function. Then Σ\Sigma parametrizes by

X(r,θ)=(rcosθ,rsinθ,f(r)+hθ),rI,θ.X(r,\theta)=(r\cos\theta,r\sin\theta,f(r)+h\theta),\quad r\in I,\theta\in\mathbb{R}. (7)

If D=r2(1+f2)+h2D=r^{2}(1+f^{\prime 2})+h^{2}, then the unit normal vector NN is

N=1D(hsinθrfcosθ,hcosθrfsinθ,r)N=\frac{1}{\sqrt{D}}(h\sin\theta-rf^{\prime}\cos\theta,-h\cos\theta-rf^{\prime}\sin\theta,r)

and the Gauss curvature is

K=r3ff′′h2D2.K=\frac{r^{3}f^{\prime}f^{\prime\prime}-h^{2}}{D^{2}}.

Then Σ\Sigma is a KαK^{\alpha}-translator if

(r3ff′′h2D2)α=D1/2(v1(hsinθrfcosθ)v2(hcosθ+rfsinθ)+v3r).\left(\frac{r^{3}f^{\prime}f^{\prime\prime}-h^{2}}{D^{2}}\right)^{\alpha}=D^{-1/2}\left(v_{1}(h\sin\theta-rf^{\prime}\cos\theta)-v_{2}(h\cos\theta+rf^{\prime}\sin\theta)+v_{3}r\right). (8)

As a first conclusion, we prove that v\vec{v} must be parallel to the zz-axis. The proof is similar of Proposition 2.1.

Proposition 3.1.

Let Σ\Sigma be a KαK^{\alpha}-translator. If Σ\Sigma is a helicoidal surface, then the axis is parallel to the speed vector v\vec{v}.

Proof.

Equation (8) can be written as A0+A1cosθ+A2sinθ=0A_{0}+A_{1}\cos\theta+A_{2}\sin\theta=0. From A1=A2=0A_{1}=A_{2}=0, we obtain

v1hv2rf=0,v1rf+v2h=0.v_{1}h-v_{2}rf^{\prime}=0,\quad v_{1}rf^{\prime}+v_{2}h=0.

Combining both equations, we have v2(h2+r2f2)=0v_{2}(h^{2}+r^{2}f^{\prime 2})=0. Thus v2=0v_{2}=0 and hence, v1=0v_{1}=0. ∎

From this proposition, we can assume that v=(0,0,1)\vec{v}=(0,0,1). Then (8) is

(r3ff′′h2D2)α=rD.\left(\frac{r^{3}f^{\prime}f^{\prime\prime}-h^{2}}{D^{2}}\right)^{\alpha}=\frac{r}{\sqrt{D}}. (9)

We will obtain a first integration of this equation. For this, let

g(r)=r2(1+f(r)2)+h2.g(r)=r^{2}(1+f^{\prime}(r)^{2})+h^{2}.

Then r2g=g+r3ff′′h2\frac{r}{2}g^{\prime}=g+r^{3}f^{\prime}f^{\prime\prime}-h^{2} and thus (9) is equivalent to

rg=2g+2r1αg4α12α.rg^{\prime}=2g+2r^{\frac{1}{\alpha}}g^{\frac{4\alpha-1}{2\alpha}}.

If α=1/2\alpha=1/2, the solution is g(r)=mr2er2g(r)=mr^{2}e^{r^{2}}, with mm\in\mathbb{R}. If α1/2\alpha\not=1/2, the solution of this equation is

g(r)=r2(12α2αr2+m)2α12α,m.g(r)=r^{2}\left(\frac{1-2\alpha}{2\alpha}r^{2}+m\right)^{\frac{2\alpha}{1-2\alpha}},\quad m\in\mathbb{R}.

In terms of the function ff, we have

1+f2=(12α2αr2+m)2α12αh2r2.1+f^{\prime 2}=\left(\frac{1-2\alpha}{2\alpha}r^{2}+m\right)^{\frac{2\alpha}{1-2\alpha}}-\frac{h^{2}}{r^{2}}.

In particular, this gives a restriction on the domain of f(r)f(r).

Theorem 3.2.

Let Σ\Sigma be a KαK^{\alpha}-translator. If Σ\Sigma is a helicoidal surface about the zz-axis and pitch hh, then Σ\Sigma parametrizes as (7), where

f(r)={±r(met2h2t21)1/2𝑑t,m>0,α=12±r((12α2αt2+m)2α12αh2t21)1/2𝑑t,m,α12.f(r)=\left\{\begin{array}[]{lll}\pm\int^{r}\left(me^{t^{2}}-\frac{h^{2}}{t^{2}}-1\right)^{1/2}\,dt,m>0,&&\alpha=\frac{1}{2}\\ \pm\int^{r}\left(\left(\frac{1-2\alpha}{2\alpha}t^{2}+m\right)^{\frac{2\alpha}{1-2\alpha}}-\frac{h^{2}}{t^{2}}-1\right)^{1/2}\,dt,m\in\mathbb{R},&&\alpha\neq\frac{1}{2}.\end{array}\right. (10)

We now show examples of ff^{\prime} for some choices of α\alpha. In all cases, the pitch is h=1h=1. See Figures 3 and 4

  1. (1)

    Case α=1/2\alpha=1/2. We take m=1m=1 in (10). The function ff is defined provided r2log((1+r2)/r2)r^{2}\geq\log((1+r^{2})/r^{2}), that is, [r0,)[r_{0},\infty), where r00.898r_{0}\approx 0.898.

  2. (2)

    Case α=1/4\alpha=1/4. Let m=0m=0. Then 1+f2=1/r2+r21+f^{\prime 2}=-1/r^{2}+r^{2}. Now the restriction on rr is r4r210r^{4}-r^{2}-1\geq 0, that is, the domain is [r0,)[r_{0},\infty) with r01.272r_{0}\approx 1.272. This case appeared in [14].

  3. (3)

    Case α=1\alpha=1. Let m=0m=0 in (10). Then 1+f2=4/r41/r21+f^{\prime 2}=4/r^{4}-1/r^{2} and the domain of ff is [0,r0][0,r_{0}] with r01.249r_{0}\approx 1.249. Here limr0f(r)=limr0f(r)=\lim_{r\to 0}f(r)=\lim_{r\to 0}f^{\prime}(r)=\infty.

Refer to caption
Refer to caption
Refer to caption
Figure 3. Generating curves of helicoidal KαK^{\alpha}-translators: α=1/2\alpha=1/2 (left), α=1/4\alpha=1/4 (middle) and α=1\alpha=1 (right).
Refer to caption
Refer to caption
Refer to caption
Figure 4. Helicoidal KαK^{\alpha}-translators: α=1/2\alpha=1/2 (left), α=1/4\alpha=1/4 (middle) and α=1\alpha=1 (right).

Following Lee [14] we may also use the approach of Bour ([7]). More explicitly, the Bour coordinates are defined by rsr\mapsto s and θt,\theta\mapsto t, t=θ+Θt=\theta+\Theta, where

ds=(1+r2r2+h2f2)1/2dr,dΘ=hr2+h2fdr.ds=\left(1+\frac{r^{2}}{r^{2}+h^{2}}f^{\prime 2}\right)^{1/2}dr,\quad d\Theta=\frac{h}{r^{2}+h^{2}}f^{\prime}dr.

The first fundamental form is now I=ds2+U2dt2,I=ds^{2}+U^{2}dt^{2}, where the so-called Bour function is introduced by the relation U2=r2+h2.U^{2}=r^{2}+h^{2}. Using the Bour function UU, the terms rr, ff and Θ\Theta can be determined by

{r=U2h2,df2=U2(U2h2)2(U2(1(dUds)2)h2)ds2,dΘ=hU2df.\left\{\begin{array}[]{l}r=\sqrt{U^{2}-h^{2}},\\ df^{2}=\frac{U^{2}}{(U^{2}-h^{2})^{2}}\left(U^{2}\left(1-(\frac{dU}{ds})^{2}\right)-h^{2}\right)ds^{2},\\ d\Theta=\frac{h}{U^{2}}df.\end{array}\right. (11)

With the above discussion, the Gauss curvature is now K=(d2U/ds2)/UK=-(d^{2}U/ds^{2})/U and N,(0,0,1)=dU/ds.\left\langle N,(0,0,1)\right\rangle=dU/ds. Then (1) is now

1Ud2Uds2=(dUds)1α-\frac{1}{U}\frac{d^{2}U}{ds^{2}}=\left(\frac{dU}{ds}\right)^{\frac{1}{\alpha}}

or equivalently

1dU/dsd2Uds2=U(dUds)1αα.\frac{1}{dU/ds}\frac{d^{2}U}{ds^{2}}=-U\left(\frac{dU}{ds}\right)^{\frac{1-\alpha}{\alpha}}.

Setting P=dU/dsP=dU/ds and dPdU=(ds/dU)(d2U/ds2)\frac{dP}{dU}=(ds/dU)(d^{2}U/ds^{2}), we have

dPdU=UP1αα.\frac{dP}{dU}=-UP^{\frac{1-\alpha}{\alpha}}.

A first integration is

ds={m1eU22dU,α=1/2(m2α12αU2)α12αdU,α1/2,ds=\left\{\begin{array}[]{lll}m^{-1}e^{\frac{U^{2}}{2}}dU,&&\alpha=1/2\\ (m-\frac{2\alpha-1}{2\alpha}U^{2})^{\frac{\alpha}{1-2\alpha}}dU,&&\alpha\neq 1/2,\end{array}\right. (12)

where mm\in\mathbb{R} is a integration constant with m>0m>0 if α=1/2\alpha=1/2. Because ss can be viewed a function of UU, we may interchange their roles. Therefore, we obtain again a classification of helicoidal KαK^{\alpha}-translators in terms of the Bour function UU.

Theorem 3.3.

Let Σ\Sigma be a KαK^{\alpha}-translator. If Σ\Sigma is a helicoidal surface about the zz-axis and pitch hh, then Σ\Sigma parametrizes as

X(U,t)=(U2h2cos(tΘ(U)),U2h2sin(tΘ(U)),f(U)+h(tΘ(U))),X(U,t)=(\sqrt{U^{2}-h^{2}}\cos(t-\Theta(U)),\sqrt{U^{2}-h^{2}}\sin(t-\Theta(U)),f(U)+h(t-\Theta(U))),

where UU is the Bour function, dΘ=hU2dfd\Theta=hU^{-2}df and dfdf is

{±m1UU2h2(U2(1m2eU2)h2)12eU22dU,α=1/2±UU2h2(U2(1(m2α12αU2)2α2α1)h2)12(m2α12αU2)α12αdU,α1/2,\left\{\begin{array}[]{lll}\frac{\pm m^{-1}U}{U^{2}-h^{2}}\left(U^{2}\left(1-m^{2}e^{-U^{2}}\right)-h^{2}\right)^{\frac{1}{2}}e^{\frac{U^{2}}{2}}dU,&&\alpha=1/2\\ \frac{\pm U}{U^{2}-h^{2}}\left(U^{2}\left(1-(m-\frac{2\alpha-1}{2\alpha}U^{2})^{\frac{2\alpha}{2\alpha-1}}\right)-h^{2}\right)^{\frac{1}{2}}(m-\frac{2\alpha-1}{2\alpha}U^{2})^{\frac{\alpha}{1-2\alpha}}dU,&&\alpha\neq 1/2,\end{array}\right. (13)

with mm\in\mathbb{R} (m>0m>0 if α=1/2\alpha=1/2).

Proof.

Because we see ss as a function of UU in (12), UU can be considered as a new variable. Then we have the first equality in (11), where rr depends on this new variable UU and as do ff and Θ.\Theta. Considering this, together θ=tΘ\theta=t-\Theta in (7), we have the parametrization of the helicoidal KαK^{\alpha}-translator. Up to α=1/2\alpha=1/2 or not, from (12) we complete the proof. ∎

Again, for some particular values of α\alpha, (12) can be explicitly integrated. We present the cases α=1/4\alpha=1/4 ([14]), α=1/3\alpha=1/3 and α=1\alpha=1.

  1. (1)

    Case α=1/4.\alpha=1/4. The solution is

    s={12(Um+U2+mcosh1(U)),U{m,1}when m<0,12U2,U>0when m=0,12(Um+U2+msinh1(U)),U>0when m>0.s=\left\{\begin{array}[]{lll}\frac{1}{2}\left(U\sqrt{m+U^{2}}+m\cosh^{-1}(U)\right),&U\geq\{\sqrt{-m},1\}&\mbox{when }m<0,\\ \frac{1}{2}U^{2},&U>0&\mbox{when }m=0,\\ \frac{1}{2}\left(U\sqrt{m+U^{2}}+m\sinh^{-1}(U)\right),&U>0&\mbox{when }m>0.\end{array}\right.
  2. (2)

    Case α=1/3\alpha=1/3. The solution is

    s=mU+U36,s=mU+\frac{U^{3}}{6},

    where U(0,).U\in(0,\infty).

  3. (3)

    Case α=1.\alpha=1. Then,

    s={2mtanh1(U2m),U2mwhen m>02U,U>0when m=02mtan1(U2m),U2mwhen m<0.s=\left\{\begin{array}[]{lll}\sqrt{\frac{2}{m}}\tanh^{-1}\left(\frac{U}{\sqrt{2m}}\right),&U\leq\sqrt{2m}&\mbox{when }m>0\\ \frac{2}{U},&U>0&\mbox{when }m=0\\ -\sqrt{\frac{2}{-m}}\tan^{-1}\left(\frac{U}{\sqrt{-2m}}\right),&U\leq\sqrt{-2m}&\mbox{when }m<0.\end{array}\right.

    In particular, we can express UU in terms of ss,

    U={m2tanh(2ms),s(2m)1/2when m>02s,s>0when m=0m2tan(2ms),s(2m)1/2when m<0.U=\left\{\begin{array}[]{lll}\sqrt{\frac{m}{2}}\tanh\left(\sqrt{2m}s\right),&\mid s\mid\leq(2m)^{-1/2}&\mbox{when }m>0\\ \frac{2}{s},&s>0&\mbox{when }m=0\\ -\sqrt{\frac{-m}{2}}\tan\left(\sqrt{-2m}s\right),&\mid s\mid\leq(-2m)^{-1/2}&\mbox{when }m<0.\end{array}\right.

4. KαK^{\alpha}-translators of translation type

By a translation surface of 3\mathbb{R}^{3} we mean a surface given by the sum of two curves contained in two coordinate planes. This is a particular case of a Darboux surface where A(t)A(t) is the identity. After a rigid motion, the surface is the sum of the curves γ(x)=(x,0,f(x))\gamma(x)=(x,0,f(x)) and β(y)=(0,y,g(y))\beta(y)=(0,y,g(y)), where f:If\colon I\subset\mathbb{R}\to\mathbb{R} and g:Jg\colon J\subset\mathbb{R}\to\mathbb{R} are smooth functions in one variable. Thus the surface parametrizes by

X(x,y)=(x,y,f(x)+g(y)),xI,yJ.X(x,y)=(x,y,f(x)+g(y)),\ x\in I,y\in J. (14)

If we see the surface Σ\Sigma as the graph of z=f(x)+g(y)z=f(x)+g(y), then the problem of finding all translation surfaces that are KαK^{\alpha}-translator is equivalent to ask which are the solutions of (1) obtained by separation of variables z=f(x)+g(y)z=f(x)+g(y). In this section we classify all KαK^{\alpha}-translators of translation type.

We calculate all terms of equation (1). Let us observe that once we have the parametrization (14), we cannot prescribed the speed v\vec{v} because the parametrization (14) was previously fixed after a rigid motion. Thus in order to calculate all terms of (1), the vector v=(v1,v2,v3)\vec{v}=(v_{1},v_{2},v_{3}) is assumed in all its generality. The computation of (1) gives

(f′′g′′)α(1+f2+g2)2α=v3v1fv2g(1+f2+g2)1/2,\frac{(f^{\prime\prime}g^{\prime\prime})^{\alpha}}{(1+f^{\prime 2}+g^{\prime 2})^{2\alpha}}=\frac{v_{3}-v_{1}f^{\prime}-v_{2}g^{\prime}}{(1+f^{\prime 2}+g^{\prime 2})^{1/2}}, (15)

where the prime denotes the derivatives with respect to the corresponding variables xx or yy in each case.

In order to clarify the arguments, we separate the case α=1/4\alpha=1/4. This case appears because the denominators in (15) are cancelled.

Theorem 4.1.

Let Σ\Sigma be a K1/4K^{1/4}-translator with speed v\vec{v} of translation type parametrized by (14). Up to a change of the roles of ff and gg, the function ff is f(x)=x2/m+ax+bf(x)=x^{2}/m+ax+b, a,b,ma,b,m\in\mathbb{R}, m0m\not=0 and gg is one of the following functions depending on the speed v\vec{v}:

  1. (1)

    If the speed is v=(0,0,1)\vec{v}=(0,0,1), then g(y)=my2/4+cy+dg(y)=my^{2}/4+cy+d, c,dc,d\in\mathbb{R}.

  2. (2)

    If the speed is v=(0,1,v3)\vec{v}=(0,1,v_{3}), then

    g(y)=v3(2c+my)m(32)2/3(2c+my)2/3m+d,g(y)=\frac{v_{3}(2c+my)}{m}-\frac{\left(\frac{3}{2}\right)^{2/3}(2c+my)^{2/3}}{m}+d, (16)

    where c,dc,d\in\mathbb{R}.

Proof.

Now (15) is

(f′′g′′)1/4=v3v1fv2g.(f^{\prime\prime}g^{\prime\prime})^{1/4}=v_{3}-v_{1}f^{\prime}-v_{2}g^{\prime}. (17)

Let us observe the symmetry of the roles of ff and gg, so it is enough to distinguish cases according to the function ff. Notice that f′′f^{\prime\prime} is not constantly 0 because the case K=0K=0 was initially discarded.

  1. (1)

    Case f′′=2/m0f^{\prime\prime}=2/m\not=0 is a non-zero constant, m0m\not=0. In particular, f(x)=x2/m+ax+bf(x)=x^{2}/m+ax+b, a,ba,b\in\mathbb{R}. From (17), v1=0v_{1}=0 and (2g′′/m)1/4=v3v2g(2g^{\prime\prime}/m)^{1/4}=v_{3}-v_{2}g^{\prime}. Since g′′0g^{\prime\prime}\not=0 (otherwise, K=0K=0), then g′′=m2(v3v2g)4g^{\prime\prime}=\frac{m}{2}(v_{3}-v_{2}g^{\prime})^{4}. If v2=0v_{2}=0, we assume that v3=1v_{3}=1 and the solution is g(y)=my2/4+cy+dg(y)=my^{2}/4+cy+d, c,dc,d\in\mathbb{R}. If v20v_{2}\not=0, then the solution of this equation is (16).

  2. (2)

    Suppose that f′′f^{\prime\prime} is not constant. Differentiating successively (17) with respect to xx and next with respect to yy, we obtain f′′′(x)g′′′(y)=0f^{\prime\prime\prime}(x)g^{\prime\prime\prime}(y)=0 for all x,yx,y. If at some xx, f′′′(x)0f^{\prime\prime\prime}(x)\not=0, then g′′g^{\prime\prime} is constant in some interval and we are in the previous case (1) interchanging the roles of ff and gg. Thus f′′′=0f^{\prime\prime\prime}=0 in its domain, which it is a contradiction because f′′f^{\prime\prime} is not a constant function.

Let us observe that the surfaces of the case (1) of Theorem 4.1 are affinities of the paraboloid z=x2+y2z=x^{2}+y^{2} and that the speed v=(0,0,1)\vec{v}=(0,0,1). Other consequence of this result is that we find K1/4K^{1/4}-translators where the speed v\vec{v} is parallel to the xyxy-plane by choosing v=(0,1,0)\vec{v}=(0,1,0). If we assume that the speed v\vec{v} is (0,0,1)(0,0,1), as usually is taken as a convention in the literature (e.g. [9, 17]), then changing the roles of yy and zz, we can provide examples of translation surfaces X(x,z)=(x,f(x)+g(z),z)X(x,z)=(x,f(x)+g(z),z) whose speed is (0,0,1)(0,0,1).

Example 4.2.

Let α=1/4\alpha=1/4 and v=(0,0,1)\vec{v}=(0,0,1). Suppose that a K1/4K^{1/4}-translator parametrizes as y=f(x)+g(z)y=f(x)+g(z). In this case, (1) is

(f′′g′′)1/4=g.(f^{\prime\prime}g^{\prime\prime})^{1/4}=g^{\prime}.

Hence g>0g^{\prime}>0 and

g′′g4=1f′′=m\frac{g^{\prime\prime}}{g^{\prime 4}}=\frac{1}{f^{\prime\prime}}=m

for some constant m0m\not=0. Integrating,

f(x)=12mx2+ax+b,g(z)=12m(3mz+c)2/3+d,f(x)=\frac{1}{2m}x^{2}+ax+b,\quad g(z)=-\frac{1}{2m}(-3mz+c)^{2/3}+d,

with 3mz+c>0-3mz+c>0, c,dc,d\in\mathbb{R}. For example, choose m=1m=1 and a=b=c=d=0a=b=c=d=0. Then the surface is

X(x,z)=(x,12x212(3z)2/3,z).X(x,z)=(x,\frac{1}{2}x^{2}-\frac{1}{2}(-3z)^{2/3},z).

If we see this surface as the graph on the xyxy-plane, and after a symmetry about the zz-plane, we have

z=13(x22y)3/2.z=\frac{1}{3}(x^{2}-2y)^{3/2}.

We now consider the general case α1/4\alpha\not=1/4.

Theorem 4.3.

If α1/4\alpha\not=1/4, there are not KαK^{\alpha}-translators that are surfaces of translation.

Proof.

Again by the symmetry of the roles of ff and gg, we discuss according to the function ff. Recall that f′′f^{\prime\prime} (and consequently, g′′g^{\prime\prime}) cannot be constantly 0 because then KK would be 0. Then (15) is

(g′′)α=(1+f2+g2)4α12(v3v1fv2g)(f′′)α.(g^{\prime\prime})^{\alpha}=(1+f^{\prime 2}+g^{\prime 2})^{\frac{4\alpha-1}{2}}(v_{3}-v_{1}f^{\prime}-v_{2}g^{\prime})(f^{\prime\prime})^{-\alpha}. (18)

We differentiate (18) with respect to xx, and after some manipulations, we arrive to

(4α1)ff′′P(1+f2+g2)(v1(f′′)1α+α(f′′)α1f′′′P)=0,(4\alpha-1)f^{\prime}f^{\prime\prime}P-(1+f^{\prime 2}+g^{\prime 2})\left(v_{1}(f^{\prime\prime})^{1-\alpha}+\alpha(f^{\prime\prime})^{-\alpha-1}f^{\prime\prime\prime}P\right)=0,

where P=v3v1fv2gP=v_{3}-v_{1}f^{\prime}-v_{2}g^{\prime}. The above equation is a polynomial equation on g=g(y)g^{\prime}=g^{\prime}(y) of degree 33, which we write as

A0+A1g+A2g2+A3g3=0,A_{0}+A_{1}g^{\prime}+A_{2}g^{\prime 2}+A_{3}g^{\prime 3}=0,

where all coefficients AiA_{i} are functions on the variable xx. Therefore they must vanish because g0g^{\prime}\not=0. We have

A1=v2(4α1)ff′′+αv2(1+f2)(f′′)α1f′′′,A3=α(f′′)α1f′′′v2.\begin{split}A_{1}&=-v_{2}(4\alpha-1)f^{\prime}f^{\prime\prime}+\alpha v_{2}(1+f^{\prime 2})(f^{\prime\prime})^{-\alpha-1}f^{\prime\prime\prime},\\ A_{3}&=\alpha(f^{\prime\prime})^{-\alpha-1}f^{\prime\prime\prime}v_{2}.\end{split}

Since 4α104\alpha-1\not=0, we obtain v2=0v_{2}=0. Now

A0=(4α1)(v3v1f)ff′′(1+f2)(v1(f′′)1α+α(f′′)α1f′′′(v3v1f)),A2=v1(f′′)1α+α(f′′)α1f′′′(v3v1f).\begin{split}A_{0}&=(4\alpha-1)(v_{3}-v_{1}f^{\prime})f^{\prime}f^{\prime\prime}-(1+f^{\prime 2})(v_{1}(f^{\prime\prime})^{1-\alpha}+\alpha(f^{\prime\prime})^{-\alpha-1}f^{\prime\prime\prime}(v_{3}-v_{1}f^{\prime})),\\ A_{2}&=v_{1}(f^{\prime\prime})^{1-\alpha}+\alpha(f^{\prime\prime})^{-\alpha-1}f^{\prime\prime\prime}(v_{3}-v_{1}f^{\prime}).\end{split}

Using A2=0A_{2}=0 into A1A_{1}, we arrive to (4α1)(v3v1f)ff′′=0(4\alpha-1)(v_{3}-v_{1}f^{\prime})f^{\prime}f^{\prime\prime}=0, obtaining a contradiction. This completes the proof. ∎

Translations surfaces appear as surfaces obtained by the translation of a curve along another curve. As we have seen, in case that the two curves are included in coordinate planes, then the surface can be written as z=f(x)+g(y)z=f(x)+g(y). As we said, the problem to classify all translation surfaces that are KαK^{\alpha}-translators is equivalent to solve equation (1) by separation of variables. Other way of separation of variables is assuming that z=f(x)g(y)z=f(x)g(y), for two smooth functions ff and gg. However, the equation (1) is difficult to solve in all its generality. We only show an example where we can obtain non-trivial examples if α=1/4\alpha=1/4.

Example 4.4.

. Assume α=1/4\alpha=1/4 and the speed is v=(0,0,1)\vec{v}=(0,0,1). Instead to assume z=f(x)g(y)z=f(x)g(y), we suppose x=f(z)g(y)x=f(z)g(y). Then the parametrization of the surface is X(z,y)=(f(z)g(y),y,z)X(z,y)=(f(z)g(y),y,z), zIz\in I, yJy\in J and (1) is

fgf′′g′′f2g2=f4g4.fgf^{\prime\prime}g^{\prime\prime}-f^{\prime 2}g^{\prime 2}=f^{\prime 4}g^{4}. (19)

Because α=1/4\alpha=1/4, then K>0K>0 and so, f′′(x)g′′(y)0f^{\prime\prime}(x)g^{\prime\prime}(y)\not=0. We divide (19) with f2gg′′,f^{\prime 2}gg^{\prime\prime}, obtaining

ff′′f2g2gg′′=f2g3g′′.\frac{ff^{\prime\prime}}{f^{\prime 2}}-\frac{g^{\prime 2}}{gg^{\prime\prime}}=f^{\prime 2}\frac{g^{3}}{g^{\prime\prime}}. (20)

Differentiating successively (20) with respect to zz and yy, we have

2ff′′(g3g′′)=0.2f^{\prime}f^{\prime\prime}\left(\frac{g^{3}}{g^{\prime\prime}}\right)^{\prime}=0.

Because ff′′0f^{\prime}f^{\prime\prime}\not=0, then g′′=ag3g^{\prime\prime}=ag^{3}, where a,a\in\mathbb{R}, a0.a\neq 0. Now (20) is

ff′′f2f2a=g2ag4=b,\frac{ff^{\prime\prime}}{f^{\prime 2}}-\frac{f^{\prime 2}}{a}=\frac{g^{\prime 2}}{ag^{4}}=b,

for some nonzero constant bb. Then g2=abg4g^{\prime 2}=abg^{4} and differentiating and using that g′′=ag3g^{\prime\prime}=ag^{3}, we deduce that b=1/2b=1/2. For the function gg, we have

g2=a2g4,g^{\prime 2}=\frac{a}{2}g^{4},

in particular, a>0a>0. The solution of this equation is g(y)=±2ay+c,g(y)=\pm\frac{2}{ay+c}, cc\in\mathbb{R}. We come back to (19), obtaining

ff′′12f2=f4a.ff^{\prime\prime}-\frac{1}{2}f^{\prime 2}=\frac{f^{\prime 4}}{a}.

We see ff^{\prime} as a function of ff, f=p(f)f^{\prime}=p(f). Then p=dpdf=f′′fp^{\prime}=\frac{dp}{df}=\frac{f^{\prime\prime}}{f^{\prime}}, so

pp2f=p3af,p^{\prime}-\frac{p}{2f}=\frac{p^{3}}{af},

which is an ODE of Bernoulli type. The solution is

f(z)=±amf12m2f.f^{\prime}(z)=\pm\frac{\sqrt{a}m\sqrt{f}}{\sqrt{1-2m^{2}f}}.

The solution f(z)f(z) of this equation together the above function g(y)g(y) provide an example of a K1/4K^{1/4}-translator given by x=f(z)g(y)x=f(z)g(y).

5. Ruled KαK^{\alpha}-translators

In this section we study the solutions of (1) when the surface is ruled. A parametrization of a ruled surface Σ\Sigma is

X(s,t)=γ(s)+tw(s),sI,t,X(s,t)=\gamma(s)+tw(s),\quad s\in I\subset\mathbb{R},t\in\mathbb{R},

where w=w(s)w=w(s) is a smooth function with |w(s)|=1|w(s)|=1 for all sIs\in I and γ\gamma is a curve parametrized by arc-length.

A case to discard is that Σ\Sigma is a cylindrical surface because K=0K=0. Thus, w=w(s)w=w(s) is a non-constant function. In such a case, we can choose γ\gamma to be the striction curve, that is, γ(s),w(s)=0\langle\gamma^{\prime}(s),w^{\prime}(s)\rangle=0 for all sIs\in I. Then

K=λ2(λ2+t2)2,λ=(γ,w,w)|w|2,K=-\frac{\lambda^{2}}{(\lambda^{2}+t^{2})^{2}},\quad\lambda=\frac{(\gamma^{\prime},w,w^{\prime})}{|w^{\prime}|^{2}},

where the parenthesis (,,)(\cdot,\cdot,\cdot) denotes the determinant of the three vectors. In particular, we are assuming that λ\lambda is not constantly 0. On the other hand, the unit normal vector NN is

N=λw+tw×w|w|λ2+t2.N=\frac{\lambda w^{\prime}+tw^{\prime}\times w}{|w^{\prime}|\sqrt{\lambda^{2}+t^{2}}}.

In particular, KK is negative, so we are assuming that α\alpha is an integer. Then (1) writes as

(1)αλ2α(λ2+t2)2α=1|w|λ2+t2(λw,v+t(w,w,v)).(-1)^{\alpha}\frac{\lambda^{2\alpha}}{(\lambda^{2}+t^{2})^{2\alpha}}=\frac{1}{|w^{\prime}|\sqrt{\lambda^{2}+t^{2}}}(\lambda\langle w^{\prime},\vec{v}\rangle+t(w^{\prime},w,\vec{v})).

or equivalently,

(1)αλ2α|w|+λw,v(λ2+t2)4α12+(w,w,v)t(λ2+t2)4α12=0.-(-1)^{\alpha}\lambda^{2\alpha}|w^{\prime}|+\lambda\langle w^{\prime},\vec{v}\rangle(\lambda^{2}+t^{2})^{\frac{4\alpha-1}{2}}+(w^{\prime},w,\vec{v})t(\lambda^{2}+t^{2})^{\frac{4\alpha-1}{2}}=0. (21)

The Wronskian of the functions {1,(λ2+t2)4α12,t(λ2+t2)4α12}\left\{1,(\lambda^{2}+t^{2})^{\frac{4\alpha-1}{2}},t(\lambda^{2}+t^{2})^{\frac{4\alpha-1}{2}}\right\} is

(4α1)(λ2+t2)4α3(4αt2λ2).(4\alpha-1)(\lambda^{2}+t^{2})^{4\alpha-3}\left(4\alpha t^{2}-\lambda^{2}\right).

Since α1/4\alpha\neq 1/4, the Wronskian if not 0, proving these three functions are linearly independent. This yields a contradiction with (21). As a conclusion, we have the following result.

Theorem 5.1.

There are not KαK^{\alpha}-translators that are ruled surfaces.

Let us notice that this result of non-existence has discarded the case that the surface is cylindrical (the rulings are parallel to v\vec{v}) or that λ=0\lambda=0, that is, the surface is developable. The condition λ=0\lambda=0 together (w,w,v)=0(w,w^{\prime},\vec{v})=0 is equivalent to the case that the base curve is a planar curve parallel to v\vec{v}, obtaining that the surface is part of a plane.

Acknowledgments

This publication is part of the Project I+D+i PID2020-117868GB-I00, supported by MCIN/ AEI/10.13039/501100011033/

References

  • [1] B. Andrews, Contraction of convex hypersurfaces in Euclidean space. Calc. Var. Partial Differ. Eq. 2 (1994), 151–171.
  • [2] B. Andrews, Contraction of convex hypersurfaces by their affine normal. J. Differential Geom. 43 (1996), 207–230.
  • [3] B. Andrews, Gauss curvature flow: the fate of the rolling stones. Invent. Math. 138 (1999), 151–161.
  • [4] M. E. Aydin, R. López, Rotational KαK^{\alpha}-Translators in Minkowski space. Preprint 2022.
  • [5] L. Caffarelli, Interior W2,pW^{2,p} estimates for solutions of the Monge-Ampère equation. Ann. of Math. 131 (1990), 135–150.
  • [6] E. Calabi, Hypersurfaces with maximal affinely invariant area. Amer. J. Math. 104 (1982), 91–126.
  • [7] M. P. do Carmo, M. Dajczer, Helicoidal surfaces with constant mean curvature. Tohoku Math. J. 34 (1982), 425–425.
  • [8] B. Chow, Deforming convex hypersurfaces by the nth root of the Gaussian curvature. J. Differential Geom. 22 (1985). 117–138.
  • [9] K. Choi, P. Daskalopoulos, K.-A. Lee, Translating solutions to the Gauss curvature flow with flat sides, Analysis &\& PDE 14 (2021), 595–616.
  • [10] G. Darboux, Leçons sur la Théorie Générale des Surfaces et ses Applications Géométriques du Calcul Infinitésimal, vol. 1–4. Chelsea Publ. Co, reprint, 1972.
  • [11] W. J. Firey, Shapes of worn stones. Mathematika 21 (1974), 1–11.
  • [12] D. Gilbarg, N. Trudinger, Second Order Elliptic Partial Differential Equations. Second edition, Springer-Verlag, 1983.
  • [13] H. Ju, J. Bao, H. Jian, Existence for translating solutions of Gauss curvature flow on exterior domains. Nonlinear Anal. 75 (2012), 3629–3640.
  • [14] H. Lee, Isometric deformations of the K1/4K^{1/4}-flow translators in R3R^{3} with helicoidal symmetry. C. R. Math. Acad. Sci. Paris 351 (2013), 477–482.
  • [15] K. Tso, Deforming a hypersurface by its Gauss-Kronecker curvature. Comm. Pure Appl. Math. 38 (1985), 867–882.
  • [16] J. Urbas, An expansion of convex hypersurfaces. J. Differential Geom. 33 (1991), 91–125.
  • [17] J. Urbas, Complete noncompact self-similar solutions of Gauss curvature flows, I: Positive powers. Math. Ann. 311 (1998), 251–274.
  • [18] J. Urbas, Complete noncompact self-similar solutions of Gauss curvature flows, II: Negative powers. Adv. Differ. Equ. 4(3) (1999), 323–346.