1 Introduction and statement of the result

A constant angle surface in the three-dimensional Euclidean space \({\mathbb {R}}^3\) is an orientable surface whose Gauss map makes a constant angle with a fixed vector direction \({{\mathbf {d}}}\) ( [1]). In particular, the projection of \({{\mathbf {d}}}\) on the tangent plane of the surface is a principal direction and the corresponding principal curvature is 0. Dillen et al. generalized, in [4], the notion of the constant angle surfaces to those surfaces where the tangent projection of \({{\mathbf {d}}}\) is a principal direction without necessarily corresponding with zero principal curvature. See also [9] for the case of hypersurfaces.

Definition 1.1

Let \({{\mathbf {d}}}\in {\mathbb {R}}^3\) be a unitary vector. We say that a surface M in \({\mathbb {R}}^3\) has a canonical principal direction with respect to \({{\mathbf {d}}}\) if the tangent part \({{\mathbf {d}}}^\top \) along M is a principal direction.

We abbreviate by saying a CPD surface where the direction \({{\mathbf {d}}}\) is understood. The study of CPD surfaces in different ambient spaces has been of interest recently: See [4, 5, 8, 17].

The purpose of this paper is to classify all CPD surfaces whose mean curvature is a linear function of the next three functions on M:

$$\begin{aligned} \text{ height } \text{ function }&f(p)=\langle p,{\mathbf {v}}\rangle ,\\ \text{ angle } \text{ function }&f(p)=\langle N(p),{\mathbf {v}}\rangle ,\\ \text{ support } \text{ function }&f(p)=\langle N(p) , p \rangle . \end{aligned}$$

Here, \({\mathbf {v}}\in {\mathbb {R}}^3\) is a fixed unit vector and N is the Gauss map of M. The interest for these three types of mean curvatures is the following:

  1. 1.

    Height function. A surface M whose mean curvature H satisfies

    $$\begin{aligned} H(p)=\lambda \langle p,{\mathbf {v}}\rangle +\mu ,\ \lambda ,\mu \in {\mathbb {R}}, \end{aligned}$$
    (1)

    is a model of a liquid drop where the direction of the gravity is indicated by the vector \({\mathbf {v}}\), the constant \(\lambda \) depends only on the physical properties of the drop and \(\mu \) is a volume constraint: See [7] as a general reference. A surface whose mean curvature satisfies (1) is called a capillary surface.

  2. 2.

    Angle function. A surface M whose mean curvature H satisfies

    $$\begin{aligned} H(p)=\lambda \langle N(p),{\mathbf {v}}\rangle +\mu ,\ \lambda ,\mu \in {\mathbb {R}}, \end{aligned}$$
    (2)

    is called a \(\lambda \)-translating soliton. This notion generalizes the translating soliton (\(\mu =0\)). Translating solitons appear in the theory of the mean curvature flow being the equation of the limit flow by a proper blowup procedure near type II singular points ( [12, 13, 20]).

  3. 3.

    Support function. A surface M whose mean curvature H satisfies

    $$\begin{aligned} H(p)=\lambda \langle N(p),p\rangle +\mu ,\ \lambda ,\mu \in {\mathbb {R}}, \end{aligned}$$
    (3)

    is called a \(\lambda \)-shrinker surface. This generalizes the concept of self-expander (\(\lambda >0, \mu =0)\) and self-shrinker (\(\lambda <0, \mu =0\)). Both types of surfaces appear in the context of self-similar solutions to the mean curvature flow as models of the flow behavior near a singularity, where M does not change the shape but is contracted (self-shrinkers) or dilated (self-expanders) by the flow: See [6, 11, 13].

We observe that if \(\lambda =0\) in the three above equations, then the mean curvature H is constant on the surface. This particular case was treated previously in [9] (see Proposition 2.2 below).

Although in the literature, Eqs. (2) and (3) have been studied when \(\mu =0\), the case \(\mu \not =0\) makes sense in the context of manifolds with density [10, 16]. Indeed, consider \({\mathbb {R}}^3\) with a positive smooth density function \(e^\phi \), \(\phi \in C^\infty ({\mathbb {R}}^3)\), which serves as a weight for the volume and the surface area. For a variation \(\{M_t:t\in (-\epsilon ,\epsilon )\}\) of M, let \(A_\phi (t)\) and \(V_\phi (t)\) be the weighted area and the enclosed weighted volume of \(M_t\), respectively. Then, the first variation of \(A_\phi (t)\) and \(V_\phi (t)\) is

$$\begin{aligned} A'_\phi (0)=-2\int _M H_\phi u\, \mathrm{d}A_\phi ,\quad V'_\phi (0)=\int _M u\, \mathrm{d}A_\phi , \end{aligned}$$

where u is the normal component of the variation and \(H_\phi =H- \langle \nabla \phi ,N\rangle /2\) is called the weighted mean curvature  [19]. Then, it is immediate that M is a critical point of \(A_\phi \) for a given weighted volume if and only if \(H_\phi =\mu \) is a constant function. Equations (2) and (3) appear with suitable choices of the density function \(\phi \); namely, for Eq. (2), we take \(\phi (p)=2\lambda \langle p,{\mathbf {v}}\rangle \), and for Eq. (3), the function if \(\phi (p)=\lambda |p|^2\).

Surfaces with a canonical principal direction have suitable parametrization. It was proved in [9] that a CPD surface with respect to \({{\mathbf {d}}}\) admits the following parametrization

$$\begin{aligned} X(s,t) = \gamma (s)+ (f(t) \eta (s) + g(t) {{\mathbf {d}}}), \end{aligned}$$
(4)

where \(\gamma =\gamma (s)\), \(s\in I\subset {\mathbb {R}}\) is a curve parametrized by the arc length and contained in a plane orthogonal to \({{\mathbf {d}}}\), \(\eta =\eta (s)\) is a unit vector field along \(\gamma \) and orthogonal to both \(\gamma \) and \({{\mathbf {d}}}\). Finally, \(f=f(t)\), \(g=g(t)\), \(t\in J\subset {\mathbb {R}}\), are smooth functions that satisfy \(f'(t)^2 +g'(t)^2=1\) for all \(t\in J\). The curve \(\gamma \) is named the directrix of M, and the curve (f(t), g(t)) is the profile curve of M. As a consequence, we may take \(\eta \) to be the principal normal vector to \(\gamma \) and thus \(\gamma ''(s)=\kappa (s)\eta (s)\), where \(\kappa \) is the curvature of \(\gamma \).

Once introduced the above notation, we give the main result of this paper.

Theorem 1.1

Let M be a CPD surface parametrized as (4). If M satisfies (1), (2) or (3), then M is one of the following surfaces:

  1. 1.

    A cylindrical surface where the base curve is \(\alpha (t)=f(t)\eta +g(t){{\mathbf {d}}}\), \(\eta \) is a constant vector, the rulings are orthogonal to \({{\mathbf {d}}}\) and \(\alpha \) satisfies the one-dimensional case of (1), (2) or (3).

  2. 2.

    A surface of revolution whose generating curve is \(\alpha \) and the rotational axis is parallel to \({{\mathbf {d}}}\). The curve \(\alpha \) satisfies a specific second-order ODE depending on condition (1), (2) or (3).

  3. 3.

    A cylindrical surface where the base curve is \(\gamma _a(s)=\gamma (s)+a\eta (s)\), \(a\in {\mathbb {R}}\), \(\gamma _a\) satisfies the one-dimensional case of (1), (2) or (3) and the rulings are parallel to \({{\mathbf {d}}}\).

By a cylindrical surface, we mean a ruled surface where all the rulings are parallel. Then, the base curve of the surface is contained in a plane orthogonal to the rulings. In Theorem 1.1, items (1) and (3), the base curve of the cylindrical surface satisfies that its curvature is a linear function of the height, angle or support function depending on the case. In such a case, we will say then that the curve satisfies the one-dimensional case of (1), (2) or (3).

The proof of Theorem 1.1 is separated in each one of the three equations: See Sects. 34 and 5, respectively.

2 Preliminaries

In this section, we give an expression of the mean curvature H for a CPD surface and we classify the surfaces with constant mean curvature. Firstly, we need to give some observations about parametrization (4) in the following proposition.

Proposition 2.1

Let M be a CPD surface parametrized by (4).

  1. 1.

    We have \(1-f(t)\kappa (s)\not =0\) for all \(s\in I\), \(t\in J\).

  2. 2.

    If the function g is constant, then M is a plane orthogonal to \({{\mathbf {d}}}\).

  3. 3.

    If \(\gamma \) is a straight line, then M is a cylindrical surface whose rulings are orthogonal to \({{\mathbf {d}}}\).

  4. 4.

    If \(\gamma \) is a circle, then M is a surface of revolution whose rotational axis is parallel to \({{\mathbf {d}}}\).

Proof

  1. 1.

    Since \(\partial _sX=(1-\kappa (s)f(t))\gamma '(s)\not =0\) by regularity of M, we conclude that \(1-\kappa (s)f(t)\not =0\).

  2. 2.

    If \(g(t)=a\) for all \(s\in I\), then \(f(t)= \epsilon t + b\), with \(b\in {\mathbb {R}}\) and \(\epsilon \in \{-1,1\}\). Then, (4) gives

    $$\begin{aligned} X(s,t) = \gamma (s)+ (\epsilon t+b) \eta (s) + a {{\mathbf {d}}}= (\gamma (s)+ a {{\mathbf {d}}}) + (\epsilon t+b) \eta (s). \end{aligned}$$

    It follows that \(\gamma (s)+a{{\mathbf {d}}}\) is a planar curve and the curves \(t\mapsto (\epsilon t+b)\eta (s)\) are segments of straight lines coplanar to \(\gamma \). This proves that M is a plane orthogonal to \({{\mathbf {d}}}\).

  3. 3.

    If \(\gamma \) is a straight line, then \(\eta \) is a unit constant vector and from (4), the surface M is a cylindrical surface whose base curve is \(\alpha (t)= f(t)\eta +g(t){{\mathbf {d}}}\) and the rulings are orthogonal to \({{\mathbf {d}}}\).

  4. 4.

    After a change of coordinates, we suppose \(\gamma (s)=(r\cos (s),r\sin (s),0)\) and \({{\mathbf {d}}}=(0,0,1)\). By (4), we find

    $$\begin{aligned} X(s,t)=(r(1-f(t))\cos (s), r(1-f(t))\sin (s),g(t)), \end{aligned}$$

    proving that M is a surface of revolution about the \({{\mathbf {d}}}\)-axis.

\(\square \)

We need to have an expression of the mean curvature H of a CPD surface. The following result is proved in [9].

Proposition 2.2

Let M be a CPD surface. If M is not a plane orthogonal to \({{\mathbf {d}}}\), then the principal curvatures \(\kappa _1\) and \(\kappa _2\) with respect to the Gauss map \(N(s,t)=- g'(t) \eta (s) + f'(t) {{\mathbf {d}}}\) are

$$\begin{aligned} \kappa _1(X(s,t))= \dfrac{f''(t)}{g'(t)} , \ \ \kappa _2(X(s,t)) = \dfrac{g'(t) \kappa (s)}{1- f(t) \kappa (s)}. \end{aligned}$$

The mean curvature H with respect to N is

$$\begin{aligned} H(X(s,t))= \dfrac{1}{2} \left( \dfrac{f''(t)}{g'(t)} + \dfrac{g'(t) \kappa (s)}{1- f(t) \kappa (s)} \right) . \end{aligned}$$
(5)

In Theorem 1.1, we exclude the case \(\lambda =0\), that is, that the surface has constant mean curvature. The CPD surfaces with constant mean curvature were obtained in [9, Cor. 14]. By completeness, we give its proof.

Proposition 2.3

If M is a CPD surface with constant mean curvature H, then M is a plane, a circular cylinder with axis orthogonal to \({{\mathbf {d}}}\) or a surface of revolution whose axis is parallel to \({{\mathbf {d}}}\).

Proof

It is immediate that a plane orthogonal to \({{\mathbf {d}}}\) is a CPD surface with respect to \({{\mathbf {d}}}\) and with zero mean curvature. Suppose now that M is not the above surface, in particular, the function g(t) in (4) is not constant by item 2 of Proposition 2.1. By Proposition 2.2, we have

$$\begin{aligned} \dfrac{f''(t)}{g'(t)} + \dfrac{g'(t) \kappa (s)}{1- f(t) \kappa (s)}=2H, \end{aligned}$$

where the right-hand side is a constant. Differentiating with respect to s, we obtain

$$\begin{aligned} \frac{g'(t)\kappa '(s)}{(1-f(t)\kappa (s))^2}=0. \end{aligned}$$

Since \(g'(t)\not =0\), we conclude \(\kappa '(s)=0\) for every \(s\in I\). This means that \(\gamma \) is a straight line or a circle.

  1. 1.

    Case \(\gamma \) is a straight line. By Proposition 2.1, the surface M is a cylindrical surface whose rulings are parallel to \(\gamma \). Since H is constant, then M is a plane parallel to \(\gamma \) (\(H=0\)) or M is a circular cylinder (\(H\not =0\)) whose axis is the straight line, being \(\gamma \) orthogonal to \({{\mathbf {d}}}\).

  2. 2.

    Case \(\gamma \) is a circle. By Proposition 2.1 again, M is a surface of revolution whose axis is parallel to \({{\mathbf {d}}}\).

\(\square \)

3 CPD capillary surfaces

In this section, we prove Theorem 1.1 for capillary surfaces. By the expression of H in (5), Eq. (1) is written as

$$\begin{aligned} \dfrac{1}{2} \left( \dfrac{f''(t)}{g'(t)} + \dfrac{g'(t) \kappa (s)}{1- f(t) \kappa (s)} \right) =\lambda \left( \langle \gamma (s),{\mathbf {v}}\rangle +(f(t)\langle \eta (s),{\mathbf {v}}\rangle +g(t)\langle {{\mathbf {d}}},{\mathbf {v}}\rangle ) \right) +\mu , \end{aligned}$$

for all \(s\in I\), \(t\in J\). We differentiate with respect to s, obtaining

$$\begin{aligned} \frac{g'(t)\kappa '(s)}{(1-f(t)\kappa (s))^2}=2\lambda (1-f(t)\kappa (s))\langle \gamma '(s),{\mathbf {v}}\rangle , \end{aligned}$$
(6)

where we have used \(\gamma ''(s)=\kappa (s)\eta (s)\). We discuss according to \(\kappa \).

  1. 1.

    Case \(\kappa '(s)=0\) for every \(s\in I\). Then, the curvature function \(\kappa \) is constant. Moreover, Eq. (6) implies \(\langle \gamma '(s),{\mathbf {v}}\rangle = 0\), that is, \(\langle \gamma (s),{\mathbf {v}}\rangle \) is a constant function. From Proposition 2.1, we know that M is a cylindrical surface whose rulings are orthogonal to \({{\mathbf {d}}}\) or M is a surface of revolution whose axis is parallel to \({{\mathbf {d}}}\).

    1. (a)

      Case \(\kappa =0\). Then, \(\eta \) is a constant vector and Eq. (1) becomes

      $$\begin{aligned} \dfrac{1}{2} \dfrac{f''(t)}{g'(t)} = \lambda \langle f(t)\eta +g(t){{\mathbf {d}}},{\mathbf {v}}\rangle +\lambda \langle \gamma (s),{\mathbf {v}}\rangle +\mu . \end{aligned}$$

      Define \(\alpha (t)=f(t)\eta +g(t){{\mathbf {d}}}\) and \({\tilde{\mu }}= \lambda \langle \gamma (s),{\mathbf {v}}\rangle +\mu \). Let us observe that \({\tilde{\mu }}\) is constant. On the other hand, \(\alpha \) is parametrized by the arc length and its curvature \({\tilde{\kappa }}\) satisfies

      $$\begin{aligned} \frac{{\tilde{\kappa }}(t)}{2}=\lambda \langle \alpha (t) , {\mathbf {v}}\rangle + {\tilde{\mu }}. \end{aligned}$$

      Consequently, the planar curve \(\alpha \) satisfies the one-dimensional case of Eq. (1) as a linear combination of its height function.

    2. (b)

      Case \(\kappa \not =0\). We know that M is a rotational surface whose axis is parallel to \({{\mathbf {d}}}\). Since \(\langle \gamma '(s),{\mathbf {v}}\rangle = 0\), a differentiation with respect to s yields \( \kappa (s) \langle \eta (s), {\mathbf {v}}\rangle = 0\). We deduce that \(\langle \eta (s), {\mathbf {v}}\rangle = 0\), and it follows that \({\mathbf {v}}\) is orthogonal to both \(\gamma '\) and \(\eta \). This says that \({{\mathbf {d}}}\) is parallel to \({\mathbf {v}}\). By Eq. (1), the generating curve \(\alpha (t)=f(t)\eta +g(t){{\mathbf {d}}}\) satisfies the ODE

      $$\begin{aligned} \dfrac{1}{2} \left( \dfrac{f''(t)}{g'(t)} + \dfrac{g'(t) \kappa }{1- f(t) \kappa } \right) = \epsilon \lambda g(t) + b \end{aligned}$$

      where \(\epsilon = \langle {{\mathbf {d}}}, {\mathbf {v}}\rangle = \pm 1\) and \(b=\lambda \langle \gamma (s),{\mathbf {v}}\rangle + \mu \) is a constant.

  2. 2.

    Case \(\kappa '(s_0)\not =0\) at some \(s_0\in I\). In particular, \(\kappa '(s)\not =0\) in a subinterval \(I'\) of I around \(s=s_0\). By simplicity, we assume \(I'=I\). We now write (6) as

    $$\begin{aligned} 2 \frac{\lambda \langle \gamma '(s),{\mathbf {v}}\rangle }{\kappa '(s)}= \frac{g'(t)}{(1-f(t)\kappa (s))^3}. \end{aligned}$$

    By differentiating with respect to t, we have

    $$\begin{aligned} g''(t) (1 - f(t) \kappa (s)) + 3 g'(t) f'(t) \kappa (s) = 0. \end{aligned}$$
    (7)

    We distinguish two possibilities:

    1. (a)

      Subcase \(g''(t) \ne 0\) in some \(t_0\in J\). Then, \(g''(t)\not =0\) in an interval \(K\subset J\) around \(t=t_0\). In \(I\times K\), we can write (7) as

      $$\begin{aligned} \frac{1}{\kappa (s)}= f(t)- 3 \frac{g'(t) f'(t)}{g''(t)}. \end{aligned}$$

      Since the left-hand side depends only on s and the right-hand side on t, then they are constant functions; in particular, \(1/\kappa (s)\) is constant, a contradiction because \(\kappa '(s)\not =0\).

    2. (b)

      Subcase \(g''(t) = 0\) for every t. From (7), we have \(f'(t)\kappa (s)=0\) for every \(s\in I\), \(t\in J\). Since \(\kappa '(s)\not =0\), then \(\kappa (s)\not =0\) for some \(s\in I\), and thus \(f'(t)=0\) for every \(t\in J\). Hence, \(g'(t)= \epsilon \in \{-1,1\}\), f is a constant function \(f(t)=a\), \(a\in {\mathbb {R}}\), and \(g(t)=\epsilon t + b\), \(b\in {\mathbb {R}}\). Parametrization (4) of M becomes

      $$\begin{aligned} X(s,t) = \gamma (s)+ a \eta (s) + (\epsilon t + b) {{\mathbf {d}}}. \end{aligned}$$

      This proves that M is a cylindrical surface whose base curve is the planar curve \(\gamma _a(s) = \gamma (s)+a\eta (s)\) and the rulings are parallel to \({{\mathbf {d}}}\). Under this situation, Eq. (1) becomes

      $$\begin{aligned} \dfrac{\epsilon \kappa (s)}{2(1- a \kappa (s))} = \lambda \left( \langle \gamma _a(s) ,{\mathbf {v}}\rangle +(\epsilon t + b)\langle {{\mathbf {d}}},{\mathbf {v}}\rangle \right) +\mu . \end{aligned}$$

      The derivative with respect to t gives \( \lambda \epsilon \langle {{\mathbf {d}}},{\mathbf {v}}\rangle =0\), and we deduce that \({{\mathbf {d}}}\) is orthogonal to \({\mathbf {v}}\). Eq. (1) simplifies into

      $$\begin{aligned} \dfrac{ \kappa (s)}{1- a \kappa (s)} = 2 \epsilon \lambda \langle \gamma _a(s),{\mathbf {v}}\rangle + 2 \epsilon \mu . \end{aligned}$$
      (8)

      Finally, a simple computation gives that the curvature of \(\gamma _a\) is \(\kappa _{\gamma _a}=\dfrac{\epsilon \kappa (s)}{1- a \kappa (s)} \). Thus, from (8), we deduce that \(\gamma _a\) satisfies the one-dimensional case of (1).

We point out that the one-dimensional case of capillary Eq. (1) and the rotational capillary surfaces are well known. For the first one, we refer [18, pp. 1130–40], also [14]. A detailed description of the rotational capillary surfaces lies in [7].

4 CPD \(\lambda \)-translating solitons

We now give the proof of Theorem 1.1 for CPD \(\lambda \)-translating solitons. In order to compute (2), we need to know \(\langle N(p),{\mathbf {v}}\rangle \). By Proposition 2.2, we find

$$\begin{aligned} \langle N(s,t), {\mathbf {v}}\rangle = - g'(t) \langle \eta (s) , {\mathbf {v}}\rangle + f'(t) \langle {{\mathbf {d}}}, {\mathbf {v}}\rangle . \end{aligned}$$

By combining (2) and (5), we obtain

$$\begin{aligned} \frac{f''(t)}{g'(t)} + \frac{g'(t) \kappa (s)}{1- f(t) \kappa (s)} = - 2 \lambda g'(t) \langle \eta (s) , {\mathbf {v}}\rangle + 2 \lambda f'(t) \langle {{\mathbf {d}}}, {\mathbf {v}}\rangle +2 \mu , \end{aligned}$$

for all \(s\in I\), \(t\in J\). By taking the derivative with respect to s, and using that \(g'(t)\not =0\), we get

$$\begin{aligned} \frac{\kappa '(s)}{(1- f(t) \kappa (s) )^2} = - 2 \lambda \langle \eta '(s) , {\mathbf {v}}\rangle = 2 \lambda \kappa (s) \langle \gamma '(s) , {\mathbf {v}}\rangle . \end{aligned}$$

We differentiate this expression with respect to t, obtaining now

$$\begin{aligned} \frac{2 \kappa (s) \kappa '(s)f'(t)}{(1- f(t) \kappa (s) )^3} =0. \end{aligned}$$

Then, \((\kappa ^2(s))'f'(t)=0\). We discuss two cases.

  1. 1.

    Suppose there exists \(t_0\in J\) such that \(f'(t_0)\not =0\). Then, \(\kappa \) is a constant function. In such a case, by Proposition 2.1 we conclude that M is either a cylindrical surface whose rulings are orthogonal to \({{\mathbf {d}}}\) or M is a surface of revolution with axis parallel to \({{\mathbf {d}}}\).

    1. (a)

      Case \(\kappa =0\). Now \(\eta \) is a constant vector. Equation (2) is now

      $$\begin{aligned} \frac{f''(t)}{g'(t)} = - 2 \lambda g'(t) \langle \eta (s) , {\mathbf {v}}\rangle + 2 \lambda f'(t) \langle {{\mathbf {d}}}, {\mathbf {v}}\rangle +2 \mu . \end{aligned}$$
      (9)

      Define the planar curve \(\alpha (t)= f(t) \eta + g(t) {{\mathbf {d}}}\) is a planar curve. A computation of its principal normal vector \(N_\alpha \) gives \(N_\alpha (s)= - g'(t) \eta + f'(t) {{\mathbf {d}}}\) and its curvature is \(\kappa _\alpha (t)= \frac{f''(t)}{g'(t)} \). Thus, (9) is written as

      $$\begin{aligned} \kappa _\alpha (t) = 2 \lambda \langle N_\alpha (t),{\mathbf {v}}\rangle + 2 \mu , \end{aligned}$$

      proving that \(\alpha \) satisfies the one-dimensional case of (2).

    2. (b)

      Case \(\kappa \not =0\). Then, \(\langle \gamma '(s) , {\mathbf {v}}\rangle =0\), that is, \(\langle \gamma (s) , {\mathbf {v}}\rangle \) is a constant function. Moreover, taking another derivative with respect to s, we find \( \kappa \langle \eta (s) , {\mathbf {v}}\rangle = 0 \), so \({\mathbf {v}}\) is orthogonal to both \(\gamma '(s)\) and \(\eta (s)\) which proves that \({\mathbf {v}}\) is parallel to \({{\mathbf {d}}}\). Now Eq. (2) becomes

      $$\begin{aligned} \frac{f''(t)}{g'(t)} + \frac{g'(t) \kappa }{1- f(t) \kappa } = 2 \epsilon \lambda f'(t) +2 \mu , \end{aligned}$$

      with \( \epsilon = \langle {{\mathbf {d}}}, {\mathbf {v}}\rangle = \pm 1 \).

  2. 2.

    Case \(f'(t)=0\) for every \(t\in J\). Then, f is a constant function \(f(t)=a\), \(a\in {\mathbb {R}}\). From \(f'(t)^2+g'(t)^2=1\), we deduce that \(g(t)= \epsilon t +b\), \(\epsilon \in \{-1,1\}\), \(b\in {\mathbb {R}}\). The parametrization X in (4) is now

    $$\begin{aligned} X(s,t) = (\gamma (s)+ a \eta (s)) + (\epsilon t +b) {{\mathbf {d}}}. \end{aligned}$$

    This proves that M is a cylindrical surface over the planar curve \(\gamma _a(s)=\gamma (s)+a\eta (s)\) and the rulings are parallel to \({{\mathbf {d}}}\). Now Eq. (2) becomes

    $$\begin{aligned} \frac{ \kappa (s)}{1- a \kappa (s)} = - 2 \lambda \langle \eta (s) , {\mathbf {v}}\rangle +2 \mu \epsilon . \end{aligned}$$
    (10)

    Again the left-hand side of (10) is the curvature \(\kappa _{\gamma _a}\) of \(\gamma _a\), and consequently (10) asserts that \(\gamma _a\) satisfies the one-dimensional case of (2).

We point out that the rotational surfaces that are \(\lambda \)-translating solitons are classified in [3] if \(\mu =0\) and in [15] when \(\mu \not =0\).

5 CPD \(\lambda \)-shrinker surfaces

We now give the proof of Theorem 1.1 for CPD \(\lambda \)-shrinker surface. From the expression of N in Proposition 2.2, we have

$$\begin{aligned} \begin{array}{ccl} \langle N(p),p \rangle &{}=&{} \langle - g'(t) \eta (s) + f'(t) d , \gamma (s)+ (f(t) \eta (s) + g(t) d) \rangle \\ &{}=&{}-g'(t) \langle \eta (s) , \gamma (s) \rangle - f(t) g'(t) + f'(t) g(t). \end{array} \end{aligned}$$

By the expression of H in (5), Eq. (3) is written as

$$\begin{aligned} \frac{f''(t)}{g'(t)} + \frac{g'(t) \kappa (s)}{1- f(t) \kappa (s)} = -2 \lambda g'(t) \langle \eta (s) , \gamma (s) \rangle -2 \lambda f(t) g'(t) + 2 \lambda f'(t) g(t) + 2\mu , \end{aligned}$$

for all \(s\in I\), \(t\in J\). Since \(g'(t)\not =0\), we divide the above expression by \(g'(t)\), obtaining

$$\begin{aligned} \frac{f''(t)}{g'(t)^2} + \frac{ \kappa (s)}{1- f(t) \kappa (s)} = -2 \lambda \langle \eta (s) , \gamma (s) \rangle -2 \lambda f(t) + 2 \lambda \frac{f'(t) g(t)}{g'(t)} + 2\frac{\mu }{g'(t)}. \end{aligned}$$

We observe that the right-hand side in the above equation is a sum of a function depending only on s and other function depending only on t. Thus, if we differentiate with respect to s and then with respect to t, we obtain 0. Therefore, looking at the left-hand side, we conclude

$$\begin{aligned} 0=\frac{2\kappa (s)\kappa '(s)f(t)}{(1-f(t)\kappa (s))^3}= \frac{(\kappa ^2(s))' f(t)}{(1-f(t)\kappa (s))^3}. \end{aligned}$$

Then, \( (\kappa ^2(s))' f'(t) =0 \). Now the discussion follows the next two cases.

  1. 1.

    Suppose there exists \(t_0\in J\) such that \(f'(t_0)\not =0\). Then, \(\kappa \) is a constant function. In such a case, by Proposition 2.1, we conclude that M is either a cylindrical surface whose rulings are orthogonal to \({{\mathbf {d}}}\) or M is a surface of revolution with axis parallel to \({{\mathbf {d}}}\). By the above equations, \(\kappa \) constant implies that the function \( \langle \gamma (s) , \eta (s) \rangle \) is constant.

    1. (a)

      Case \(\kappa =0\). In particular, \(\gamma '\) and \(\eta \) are constant coplanar vectors and mutually orthogonal. Define the curve \(\alpha (t)=f(t)\eta +g(t){{\mathbf {d}}}\). The principal normal vector of \(\alpha \) is \(N_\alpha (t)=-g'(t)\eta +f'(t){{\mathbf {d}}}\). Equation (3) is written as

      $$\begin{aligned} \frac{f''(t)}{g'(t)} = 2 \lambda \langle N_\alpha (t) , \gamma (s)+ \alpha (t) \rangle +2 \mu . \end{aligned}$$

      Let us observe that \( \langle N_\alpha (t) , \gamma (s) \rangle \) does not depend on s and \(\kappa _\alpha (t)= f''(t)/g'(t)\) is the curvature of \(\alpha \). Consequently, \(\alpha \) satisfies the one-dimensional case of (3).

    2. (b)

      Case \(\kappa \ne 0\). Then, \(\gamma \) is a circle. Furthermore, we have

      $$\begin{aligned} - \kappa \langle \gamma (s) , \gamma '(s) \rangle = \langle \gamma (s) , \eta '(s) \rangle = \langle \gamma (s) , \eta (s) \rangle ' = 0 , \end{aligned}$$

      which implies that \(\langle \gamma (s) , \gamma '(s) \rangle = 0\). This shows that \(\gamma \) is a circle with center in the origin. Taking another derivative, we find

      $$\begin{aligned} \langle \gamma (s) , \gamma '(s) \rangle ' = \langle \gamma (s) , \gamma ''(s) \rangle + \langle \gamma '(s) , \gamma '(s) \rangle = \kappa \langle \gamma (s) , \eta (s) \rangle +1=0 . \end{aligned}$$

      Therefore, \(\langle \gamma (s) , \eta (s) \rangle = -1/\kappa \). Thus, M is a rotational surface whose axis of revolution passes through the origin. Moreover, the generating curve satisfies

      $$\begin{aligned} \frac{f''(t)}{g'(t)} + \frac{g'(t) \kappa }{1- f(t) \kappa } = (2/ \kappa ) \lambda g'(t) -2 \lambda f(t) g'(t) + 2 \lambda f'(t) g(t) + 2\mu . \end{aligned}$$
  2. 2.

    Case \(f'(t)=0\) for every \(t\in J\). Then, f is a constant function \(f(t)=a\), \(a\in {\mathbb {R}}\). From \(f'(t)^2+g'(t)^2=1\), we deduce that \(g(t)= \epsilon t +b\), \(\epsilon \in \{-1,1\}\), \(b\in {\mathbb {R}}\). From (4), the parametrization X is written as

    $$\begin{aligned} X(s,t) = (\gamma (s)+ a \eta (s)) + (\epsilon t +b) {{\mathbf {d}}}. \end{aligned}$$

    This proves that M is a cylindrical surface over the planar curve \(\gamma _a = \gamma (s)+a\eta (s)\) and the rulings are parallel to \({{\mathbf {d}}}\). Finally, Eq. (3) becomes

    $$\begin{aligned} \frac{ \kappa (s)}{1- a \kappa (s)} = -2 \lambda \langle \eta (s) , \gamma (s) \rangle -2 \lambda a + 2 \epsilon \mu = 2 \lambda \langle \eta (s) , \gamma _a(s) \rangle + 2 \epsilon \mu . \end{aligned}$$

    Because the curvature of \(\gamma _a\) is \(\kappa _{\gamma _a}= \kappa (s)/(1- a \kappa (s))\), then \(\gamma _a\) satisfies the one-dimensional case of (3).

The cylindrical shrinker surfaces are classified in [2].