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

A characterization of constant pp-mean curvature surfaces in the Heisenberg group H1H_{1}

Hung-Lin Chiu and Hsiao-Fan Liu Department of Mathematics, National Tsing Hua University, Hsinchu, Taiwan 300, R.O.C. [email protected] Department of Mathematics, TamKang University, New Taipei City 25137, Taiwan, R.O.C. [email protected]
Abstract.

In Euclidean 33-space, it is well known that the Sine-Gordon equation was considered in the nineteenth century in the course of investigations of surfaces of constant Gaussian curvature K=1K=-1. Such a surface can be constructed from a solution to the Sine-Gordon equation, and vice versa. With this as motivation, employing the fundamental theorem of surfaces in the Heisenberg group H1H_{1}, we show in this paper that the existence of a constant pp-mean curvature surface (without singular points) is equivalent to the existence of a solution to a nonlinear second-order ODE (1.2), which is a kind of Liénard equations. Therefore, we turn to investigate this equation. It is a surprise that we give a complete set of solutions to (1.2) (or (1.5)), and hence use the types of the solution to divide constant pp-mean curvature surfaces into several classes. As a result, after a kind of normalization, we obtain a representation of constant pp-mean curvature surfaces and classify further all constant pp-mean curvature surfaces. In Section 9, we provide an approach to construct pp-minimal surfaces. It turns out that, in some sense, generic pp-minimal surfaces can be constructed via this approach. Finally, as a derivation, we recover the Bernstein-type theorem which was first shown in [3] (or see [7, 8]).

Key words and phrases:
Keywords: Heisenberg group, Pansu sphere, p-Minimal surface, Liénard equation, Bernstein theorem
1991 Mathematics Subject Classification:
1991 Mathematics Subject Classification. Primary: 53A10, 53C42, 53C22, 34A26.

1. Introduction and main results

In literature, the Heisenberg group and its sub-Laplacian are active in many fields of analysis and sub-Riemannian geometry, control theory, semiclassical analysis of quantum mechanics, etc. (cf. [18, 19, 20, 21, 22]). It also has applications in signal analysis and geometric optics [31, 32, 33]. Research on the sub-Riemannian geometry and its analytical consequences, in particular geodesics, has been studied widely and extensively in the past ten years (cf. [19, 23, 24, 25, 26, 27, 29, 28, 30]). In this paper, the Heisenberg group is studied as a pseudo-hermitian manifold. Like Euclidean geometry, it is a branch of Klein geometries, and the corresponding Cartan geometry is pseudo-hermitian geometry.

Recall that the Heisenberg group H1H_{1} is the space 3\mathbb{R}^{3} associated with the group multiplication

(x1,y1,z1)(x2,y2,z2)=(x1+x2,y1+y2,z1+z2+y1x2x1y2),(x_{1},y_{1},z_{1})\circ(x_{2},y_{2},z_{2})=(x_{1}+x_{2},y_{1}+y_{2},z_{1}+z_{2}+y_{1}x_{2}-x_{1}y_{2}),

which is a 33-dimensional Lie group. The space of all left invariant vector fields is spanned by the following three vector fields:

e̊1=x+yz,e̊2=yxz and T=z.\mathring{e}_{1}=\frac{\partial}{\partial x}+y\frac{\partial}{\partial z},~{}~{}\mathring{e}_{2}=\frac{\partial}{\partial y}-x\frac{\partial}{\partial z}~{}~{}\mbox{ and }~{}~{}T=\frac{\partial}{\partial z}.

The standard contact bundle on H1H_{1} is the subbundle ξ\xi of the tangent bundle TH1TH_{1} which is spanned by e̊1\mathring{e}_{1} and e̊2\mathring{e}_{2}. It is also defined to be the kernel of the contact form

Θ=dz+xdyydx.\Theta=dz+xdy-ydx.

The CR structure on H1H_{1} is the endomorphism J:ξξJ:\xi\to\xi defined by

J(e̊1)=e̊2 and J(e̊2)=e̊1.J(\mathring{e}_{1})=\mathring{e}_{2}~{}~{}\mbox{ and }~{}~{}J(\mathring{e}_{2})=-\mathring{e}_{1}.

One can view H1H_{1} as a pseudo-hermitian manifold with (J,Θ)(J,\Theta) as the standard pseudo-hermitian structure. There is a natural associated connection \nabla if we regard all these left invariant vector fields e̊1,e̊2\mathring{e}_{1},\mathring{e}_{2} and TT as parallel vector fields. A naturally associated metric on H1H_{1} is the adapted metric gΘg_{\Theta}, which is defined by gΘ=dΘ(,J)+Θ2g_{\Theta}=d\Theta(\cdot,J\cdot)+\Theta^{2}. It is equivalent to define the metric by regarding e̊1,e̊2\mathring{e}_{1},\mathring{e}_{2} and TT as an orthonormal frame field. We sometimes use <,><\cdot,\cdot> to denote the adapted metric. In this paper, we use the adapted metric to measure the lengths and angles of vectors, and so on.

A pseudo-hermitian transformation (or a Heisenberg rigid motion) in H1H_{1} is a diffeomorphism in H1H_{1} which preserves the standard pseudo-hermitian structure (J,Θ)(J,\Theta). We let PSH(1)PSH(1) be the group of Heisenberg rigid motions, that is, the group of all pseudo-hermitian transformations in H1H_{1}. For details of this group, we refer readers to [4, 6], in which the fundamental theorem in the Heisenberg groups has been studied. We say that two surfaces are congruent if they differ by an action of a Heisenberg rigid motion.

The concept of minimal surfaces or constant mean curvature surfaces plays an important role in differential geometry to study the basic properties of manifolds. There is an analogous concept in pseudo-hermitian manifolds, which are called pp-minimal surfaces. In this paper, we focus on studying such kinds of surfaces in the Heisenberg group H1H_{1}. Throughout this article, all objects we discuss are assumed to be CC^{\infty} smooth, unless we specify otherwise. Suppose Σ\Sigma is a surface in the Heisenberg group H1H_{1}. There is a one-form II on Σ\Sigma which is induced from the adapted metric gΘg_{\Theta}. This induced metric is defined on the whole surface Σ\Sigma and is called the first fundamental form of Σ\Sigma. The intersection TΣξT\Sigma\cap\xi is integrated to be a singular foliation on Σ\Sigma called the characteristic foliation. Each leaf is called a characteristic curve. A point pΣp\in\Sigma is called a singular point if at which the tangent plane TpΣT_{p}\Sigma coincides with the contact plane ξp\xi_{p}; otherwise, pp is called a regular (or non-singular) point. Generically, a point pΣp\in\Sigma is a regular point, and the set of all regular points is called the regular part of Σ\Sigma. On the regular part, we can choose a unit vector field e1e_{1} such that e1e_{1} defines the characteristic foliation. The vector e1e_{1} is determined up to a sign. Let e2=Je1e_{2}=Je_{1}. Then {e1,e2}\{e_{1},e_{2}\} forms an orthonormal frame field of the contact bundle ξ\xi. We usually call the vector field e2e_{2} a horizontal vector field. Then the pp-mean curvature HH of the surface Σ\Sigma is defined by

(1.1) e1e2=He1.\nabla_{e_{1}}e_{2}=-He_{1}.

The pp-mean curvature HH is only defined on the regular part of Σ\Sigma. There are two more equivalent ways to define the pp-mean curvature from the point of view of variation and a level surface (see [3, 34]). We remark that this notion of mean curvature was proposed by J.-H. Cheng, J.-F. Hwang, A. Malchiodi, and P. Yang from the geometric point of view to generalize the one introduced first by S. Pauls in H1H_{1} for graphs over the xyxy-plane [35]. Also, in [36], M. Ritoré and C. Rosales exposed another method to compute the mean curvature of a hypersurface. If H=0H=0 on the whole regular part, we call the surface a pp-minimal surface. The pp-mean curvature is the line curvature of a characteristic curve, and hence the characteristic curves are straight lines (for the detail, see [3]). There also exists a function α\alpha defined on the regular part such that αe2+T\alpha e_{2}+T is tangent to the surface Σ\Sigma. We call this function the α\alpha-function of Σ\Sigma. It is uniquely determined up to a sign, which depends on the choice of the characteristic direction e1e_{1}. Define e^1=e1\hat{e}_{1}=e_{1} and e^2=αe2+T1+α2\hat{e}_{2}=\frac{\alpha e_{2}+T}{\sqrt{1+\alpha^{2}}}, then {e^1,e^2}\{\hat{e}_{1},\hat{e}_{2}\} forms an orthonormal frame field of the tangent bundle TΣT\Sigma. Notice that e^2\hat{e}_{2} is uniquely determined and independent of the choice of the characteristic direction e1e_{1}. In [6], it was shown that these four invariants,

I,e1,α,H,I,e_{1},\alpha,H,

form a complete set of invariants for surfaces in H1H_{1}. We remark that all the results provided in [6] still hold in the C2C^{2}-category. For each regular point pp, we can choose suitable coordinates around pp to study the local geometry of such surfaces. There always exists a coordinate system (x,y)(x,y) of pp such that

e1=x.e_{1}=\frac{\partial}{\partial x}.

We call such coordinates (x,y)(x,y) a compatible coordinate system. It is determined up to a transformation in (2.19). Notice that the compatible coordinate systems are dependent on the characteristic direction e1e_{1}.

Let ΣH1\Sigma\subset H_{1} be a constant pp-mean curvature surface with H=cH=c. In terms of a compatible coordinate system (U;x,y)(U;x,y), the α\alpha-function satisfies the following equation

(1.2) αxx+6ααx+4α3+c2α=0,\alpha_{xx}+6\alpha\alpha_{x}+4\alpha^{3}+c^{2}\alpha=0,

which first appeared as a Codazzi-like equation (1.12)(1.12) in [4] with D=1/αD=-1/{\alpha}.

It is a nonlinear ordinary differential equation and is an example of the so-called Liénard equations, named after the French physicist Alfred-Marie Liénard. The Liénard equations were intensely studied as they can be used to model oscillating circuits. Conversely, in this paper, we show that if there exists a solution α(x,y)\alpha(x,y) to the Liénard equation (1.2), we are able to construct a constant pp-mean curvature surface with H=cH=c and this given solution α\alpha as its α\alpha-function. This result is summarized as Theorem 1.1. One motivation of this theorem comes from the famous Sine-Gordon Equation (SGE),

uttuxx=sin(u)cos(u),u_{tt}-u_{xx}=\sin(u)\cos(u),

which is considerably older than the Korteweg de Vries Equation (KdV). It was discovered in the late eighteenth century in the study of pseudospherical surfaces, that is, surfaces of Gaussian curvature K=1K=-1 immersed in 3\operatorname{\mathbb{R}}^{3}, and it was intensively studied for this reason. It arises from the Gauss-Codazzi equations for pseudospherical surfaces in 3\operatorname{\mathbb{R}}^{3} and is known as an integrable equation [13]. In addition, it can also be viewed as a continuum limit [14]. Its solutions and solitons have been widely discussed by the Inverse Scattering Transform and other approaches.

There is a bijective relation between solutions uu of the SGE with (u)(0,π2)\Im(u)\subset(0,\frac{\pi}{2}) and the classes of pseudospherical surfaces in 3\operatorname{\mathbb{R}}^{3} up to rigid motion. If u:2u:\operatorname{\mathbb{R}}^{2}\rightarrow\operatorname{\mathbb{R}} is a solution such that sinucosu\sin u\cos u is zero at a point u0u_{0}, then the immersed pseudoshperical surface has cusp singularities. For example, the pseudospherical surfaces corresponding to the 1-soliton solutions of SGE are the so-called Dini surfaces and have a helix of singularities.

The study of line congruences gives rise to the concept of Bäcklund transformations. A line congruence L:MML:M\rightarrow M^{*} is called a Bäcklund transformation with the constant angle θ\theta between the normal to MM at pp and the normal to MM^{*} at p=L(p)p^{*}=L(p) and the distance between pp and pp^{*} is sinθ\sin\theta for all pMp\in M. The classical Bäcklund transformation for the Sine-Gordon equation was constructed in the nineteenth century by Swedish differential geometer Albert Bäcklund by means of a geometric construction [15, 16, 17]. We then are motivated to present analogous theorems for Heisenberg groups.

Theorem 1.1.

The existence of a constant pp-mean curvature surface ((without singular points)) in H1H_{1} is equivalent to the existence of a solution to the equation (1.2). More precisely, let α(x,y)\alpha(x,y) and H(x,y)H(x,y) be two arbitrary smooth functions on a coordinate neighborhood (U;x,y)2(U;x,y)\subset\operatorname{\mathbb{R}}^{2}. If they satisfy the following integrability condition

(1.3) (h(y)Hα(e2α𝑑x)𝑑x)Hx+ek(y)Hy=(e2α𝑑x)(αxx+6ααx+4α3+αH2),\left(h(y)-\int H\alpha\left(e^{\int 2\alpha dx}\right)dx\right)H_{x}+e^{k(y)}H_{y}=\left(e^{\int 2\alpha dx}\right)(\alpha_{xx}+6\alpha\alpha_{x}+4\alpha^{3}+\alpha H^{2}),

for some functions k(y)k(y) and h(y)h(y) in the variable yy, then there exists an embedding X:UH1X:U\rightarrow H_{1} ((provided that UU is small enough)) such that the surface Σ=X(U)\Sigma=X(U) has HH and α\alpha as its pp-mean curvature and α\alpha-function, respectively, and e^1=x\hat{e}_{1}=\frac{\partial}{\partial x}e^2=a(x,y)x+b(x,y)y\hat{e}_{2}=a(x,y)\frac{\partial}{\partial x}+b(x,y)\frac{\partial}{\partial y} with aa and bb defined as (2.15) and (2.18). In addition, such embeddings are unique, up to a Heisenberg rigid motion.

In particular, when H=cH=c is a constant, the integrability condition (1.3) reads (1.2) for each pair of functions k(y)k(y) and h(y)h(y), and the resulting surface Σ=X(U)\Sigma=X(U) is a constant pp-mean curvature surface with H=cH=c, the given function α(x,y)\alpha(x,y) as its α\alpha-function, and e^1=x\hat{e}_{1}=\frac{\partial}{\partial x}.

In this article, we sometimes call the Liénard equation (1.2) as the Codazzi-like equation from the geometrical point of view [4, 6]. We would also like to specify that a graph (x,y,u(x,y))(x,y,u(x,y)) is pp-minimal if it satisfies the pp-minimal equation (see [3])

(1.4) (uy+x)2uxx2(uy+x)(uxy)uxy+(uxy)2uyy=0.(u_{y}+x)^{2}u_{xx}-2(u_{y}+x)(u_{x}-y)u_{xy}+(u_{x}-y)^{2}u_{yy}=0.

This is a degenerate hyperbolic and elliptic partial differential equation.

Theorem 1.1 is the fundamental theorem for surfaces in H1H_{1}. After we make a detailed investigation of the original version of the integrability conditions (2.1), (1.3) is more useful than the previous one in some sense (for the origin version, we refer readers to [6] or Theorem 2.1 of this paper). It also appeared as Theorem H in [4] with a slightly different formulation as the authors of [4] did not prescribe the metric.

Theorem 1.1 follows from our detailed study on the integrability condition (see (2.1)) of the fundamental theorem (Theorem 2.1) for surfaces in H1H_{1}. Actually, if we let ω^12\hat{\omega}_{1}{}^{2} be the Levi-Civita connection associated to the induced metric with respect to the orthonormal frame field {e^1,e^2}\{\hat{e}_{1},\hat{e}_{2}\}, as specified in Theorem 1.1, then (1.3) means that ω^1,2α\hat{\omega}_{1}{}^{2},\alpha, and HH satisfy the integrability condition (2.1). This is equivalent to saying that a,b,αa,b,\alpha and HH satisfy the integrability condition (2.13) (see Subsection 2.2), which is another version of (2.1). We then have Theorem 1.1.

Given a function α(x,y)\alpha(x,y) in a coordinate neighborhood (U;x,y)2(U;x,y)\subset\operatorname{\mathbb{R}}^{2} which satisfies the Codazzi-like equation (1.2), we are able to construct a family of constant pp-mean curvature surfaces. Therefore, it suggests that a good strategy is to investigate constant pp-mean curvature surfaces by means of the Codazzi-like equation (1.2); in particular, pp-minimal surfaces. In this paper, we will focus on the theory of pp-minimal surfaces. Strategically, we first study the equation (1.2) with c=0c=0, that is,

(1.5) αxx+6ααx+4α3=0.\alpha_{xx}+6\alpha\alpha_{x}+4\alpha^{3}=0.

For nonlinear ordinary differential equations, it is known that it is rarely possible to find explicit solutions in closed form, even in power series. Fortunately, we indeed obtain a complete set of solutions to (1.5) in a simple form (see Section 4), stated as follows.

Theorem 1.2.

Besides the following three special solutions to (1.5),

α(x)=0,1x+c1,12(x+c1),\alpha(x)=0,\ \frac{1}{x+c_{1}},\ \frac{1}{2(x+c_{1})},

we have the general solution to (1.5) of the form

(1.6) α(x)=x+c1(x+c1)2+c2,\alpha(x)=\frac{x+c_{1}}{(x+c_{1})^{2}+c_{2}},

which depends on two constants c1c_{1} and c2c_{2}, and c20c_{2}\neq 0.

In Subsection 5.1, we are able to use the types of the solutions in Theorem 1.2 to divide the pp-minimal surfaces into several classes, which are vertical, special type I, special type II and general type (see Definition 5.1 and 5.2). Each type of these pp-minimal surfaces is open and contains no singular points. Generically, each pp-minimal surface is a union of these types of surfaces, and ”type” can be shown to be invariant under an action of a Heisenberg rigid motion. Now for each type, whether it is special or general, if a function α\alpha is given, then Proposition 5.3, 5.4 and 5.5 express the formula for the induced metric a,ba,b (see (5.5), (5.6) and (5.7)), which is a representation of II, on the pp-minimal surfaces with this given α\alpha as α\alpha-function. From these formulae, we see that such constructed pp-minimal surfaces depend upon two functions k(y)k(y) and h(y)h(y) for each given α\alpha. Nonetheless, in Section 6, we proceed to normalize these invariants to the following normal forms in terms of an orthogonal coordinate system (x,y)(x,y), which is a coordinate system such that a=0a=0. Such a coordinate system is determined up to a translation on (x,y)(x,y), thus we call it a normal coordinate system.

Theorem 1.3.

Let ΣH1\Sigma\subset H_{1} be a pp-minimal surface. Then, in terms of a normal coordinate system (x,y)(x,y), we can normalize the α\alpha-function and the induced metric a,ba,b to be the following normal forms:

  1. (1)

    α=1x+ζ1(y)\alpha=\frac{1}{x+\zeta_{1}(y)}, and a=0,b=α21+α2a=0,b=\frac{\alpha^{2}}{\sqrt{1+\alpha^{2}}} if Σ\Sigma is of special type I,

  2. (2)

    α=12x+ζ1(y)\alpha=\frac{1}{2x+\zeta_{1}(y)}, and a=0,b=|α|1+α2a=0,b=\frac{|\alpha|}{\sqrt{1+\alpha^{2}}} if Σ\Sigma is of special type II,

  3. (3)

    α=x+ζ1(y)(x+ζ1(y))2+ζ2(y)\alpha=\frac{x+\zeta_{1}(y)}{(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)}, and a=0,b=|α||x+ζ1(y)|1+α2a=0,b=\frac{|\alpha|}{|x+\zeta_{1}(y)|\sqrt{1+\alpha^{2}}} if Σ\Sigma is of general type,

for some functions ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y), in which ζ2(y)\zeta_{2}(y) is unique up to a translation on yy, and ζ1(y)\zeta_{1}(y) is unique up to a translation on yy as well as its image.

Therefore, ζ1(y)\zeta_{1}(y) constitutes a complete set of invariants for pp-minimal surfaces of special type I (or of special type II), whereas both ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) constitute a complete set of invariants for pp-minimal surfaces of general type. We hence give a representation for pp-minimal surfaces (see Section 6). From Theorem 1.3, together with Theorem 1.1, the following version of the fundamental theorem holds immediately for pp-minimal surfaces in H1H_{1}.

Analogs for constant pp-mean curvature surfaces are also included in Subsection 5.2. We first derive the general form of α\alpha to the Codazzi-like equation (1.2) for c0c\neq 0 ( stated as Theorem 1.4), and Proposition 5.6 provides the formula for the induced metric aa and bb.

Theorem 1.4.

Besides the trivial solution α(x)=0\alpha(x)=0, the Codazzi-like equation (1.2) for nonzero cc has the special solutions

12|c|tan(|c|x+|c|K1),α(x)=12|c|tan(|c|2x+|c|2K2),K1,K2,-\frac{1}{2}|c|\tan(|c|x+|c|K_{1}),\ \alpha(x)=\frac{1}{2}|c|\tan(-\frac{|c|}{2}x+\frac{|c|}{2}K_{2}),\quad K_{1},K_{2}\in\operatorname{\mathbb{R}},

and the general solution of the form

(1.7) α(x)=λsin(2λx+c1)c2cos(2λx+c1),\alpha(x)=\lambda\frac{\sin{(2\lambda x+c_{1}})}{c_{2}-\cos{(2\lambda x+c_{1})}},

which depends on two constants c1c_{1} and c2c_{2}, and c=2λc=2\lambda.

Since each function of sin(x)\sin(x) and cos(x)\cos(x) differs by a sign if we replace the angle xx by x+π2λx+\frac{\pi}{2\lambda}, we can assume, without loss of generality, that c20c_{2}\geq 0. We use c2c_{2} to divide constant pp-mean curvature surfaces into two classes. One class is those surfaces with c2>1c_{2}>1, and the other is that with 0c210\leq c_{2}\leq 1. It is worth our attention that, for surfaces with c2>1c_{2}>1, the denominator of the formula for α\alpha is never zero. That means that the surfaces won’t extend to one with singular points. Moreover, if the surface is closed, it will be a closed constant pp-mean curvature surface without singularity, which means this surface is of type of torus. It indicates that it is possible to find a Wente-type torus in the class of these surfaces. We can also normalize these invariants to the normal forms. However, Proposition 5.6 implies that the normalization for constant pp-mean curvature surfaces needs to be modified since aa must be nonzero. Therefore, the transformation laws help us obtain the normal coordinates (x~,y~)(\tilde{x},\tilde{y}) in Subsection 6.5 and hence we have the complete set of invariants {ζ1(y),ζ2(y)}\{\zeta_{1}(y),\zeta_{2}(y)\} for constant pp-mean curvature surfaces (see Theorem 6.3).

Theorem 1.5.

Given two arbitrary functions ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) defined on (c,d)(c,d)\subset\operatorname{\mathbb{R}}, and ζ2(y)0\zeta_{2}(y)\neq 0 for all y(c,d)y\in(c,d) (note that (c,d)(c,d) may be the whole line \operatorname{\mathbb{R}}), then

  1. (1)

    there exist an open set U(e,f)×(c,d)(2;x,y)U\subset(e,f)\times(c,d)\subset(\operatorname{\mathbb{R}}^{2};x,y), for some (e,f)(e,f)\subset\operatorname{\mathbb{R}}, and an embedding X:UH1X:U\rightarrow H_{1} such that Σ=X(U)\Sigma=X(U) is a pp-minimal surface of special type I ((or of special type II)) with (x,y)(x,y) as a normal coordinate system and ζ1(y)\zeta_{1}(y) as its ζ1\zeta_{1}-invariant;

  2. (2)

    there exist an open set U(e,f)×(c,d)(2;x,y)U\subset(e,f)\times(c,d)\subset(\operatorname{\mathbb{R}}^{2};x,y), for some (e,f)(e,f)\subset\operatorname{\mathbb{R}}, and an embedding X:UH1X:U\rightarrow H_{1} such that Σ=X(U)\Sigma=X(U) is a pp-minimal surface of general type with (x,y)(x,y) as a normal coordinate system and ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) as its ζ1\zeta_{1}- and ζ2\zeta_{2}-invariants.

Moreover, such embeddings in (1)(1) and (2)(2) are unique, up to a Heisenberg rigid motion.

Due to Theorem 1.5, for each pair of function ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y), we define in Subsection 6.3 eight maximal pp-minimal surfaces in the sense specified in Theorem 1.6. Roughly speaking, it says that any connected pp-minimal surface with type is a part of one of these eight classes. One notices that generically, a pp-minimal surface is a union of those pp-minimal surfaces with type.

Theorem 1.6.

Given two arbitrary functions ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) defined on (c,d)(c,d)\subset\operatorname{\mathbb{R}}, and ζ2(y)0\zeta_{2}(y)\neq 0 for all y(c,d)y\in(c,d) ((note that (c,d)(c,d) may be the whole line \operatorname{\mathbb{R}})), then all the eight pp-minimal surfaces

(1.8) SI(ζ1),SI+(ζ1),SII(ζ1),SII+(ζ1);andΣI(ζ1,ζ2),ΣII(ζ1,ζ2),ΣII+(ζ1,ζ2)andΣIII(ζ1,ζ2)\begin{split}&S_{I}^{-}(\zeta_{1}),\ S_{I}^{+}(\zeta_{1}),\ S_{II}^{-}(\zeta_{1}),\ S_{II}^{+}(\zeta_{1});\ \textrm{and}\\ &\Sigma_{I}(\zeta_{1},\zeta_{2}),\ \Sigma_{II}^{-}(\zeta_{1},\zeta_{2}),\ \Sigma_{II}^{+}(\zeta_{1},\zeta_{2})\ \textrm{and}\ \Sigma_{III}(\zeta_{1},\zeta_{2})\end{split}

are immersed, in addition, they are maximal in the following sense:

  • Any connected pp-minimal surface of special type I with ζ1(y)\zeta_{1}(y) as the ζ1\zeta_{1}-invariant is a part of either SI(ζ1)S_{I}^{-}(\zeta_{1}) or SI+(ζ1)S_{I}^{+}(\zeta_{1}).

  • Any connected pp-minimal surface of special type II with ζ1(y)\zeta_{1}(y) as the ζ1\zeta_{1}-invariant is a part of either SII(ζ1)S_{II}^{-}(\zeta_{1}) or SII+(ζ1)S_{II}^{+}(\zeta_{1}).

  • Any connected pp-minimal surface of type I with ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) as the ζ1\zeta_{1}- and ζ2\zeta_{2}-invariants is a part of ΣI(ζ1,ζ2)\Sigma_{I}(\zeta_{1},\zeta_{2}).

  • Any connected pp-minimal surface of type II with ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) as the ζ1\zeta_{1}- and ζ2\zeta_{2}-invariants is a part of either ΣII(ζ1,ζ2)\Sigma_{II}^{-}(\zeta_{1},\zeta_{2}) or ΣII+(ζ1,ζ2)\Sigma_{II}^{+}(\zeta_{1},\zeta_{2}).

  • Any connected pp-minimal surface of type III with ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) as the ζ1\zeta_{1}- and ζ2\zeta_{2}-invariants is a part of ΣIII(ζ1,ζ2)\Sigma_{III}(\zeta_{1},\zeta_{2}).

As applications of this theory, in Section 8, we give a complete description of the structures of the singular sets of pp-minimal surfaces in the Heisenberg group H1H_{1}.

Theorem 1.7.

The singular set of a pp-minimal surface is either

  1. (1)

    an isolated point; or

  2. (2)

    a C1C^{1} smooth curve.

In addition, an isolated singular point only happens in the surfaces of special type I with ζ1=const.\zeta_{1}=\textrm{const.}, that is, a part of the graph u=0u=0 contains the origin as the isolated singular point.

The result in Theorem 1.7 is just a special case of Theorem 3.3 in [3]. However, we give a computable proof of this result for pp-minimal surfaces. We also have the description of how a characteristic leaf goes through a singular curve, which is called a ”go through” theorem in [3].

Theorem 1.8.

Let ΣH1\Sigma\subset H_{1} be a pp-minimal surface. Then the characteristic foliation is smooth around the singular curve in the following sense that each leaf can be extended smoothly to a point on the singular curve.

Due to Theorem 1.8, we have the following result.

Theorem 1.9.

Let Σ\Sigma be a pp-minimal surface of type II ((III)). If it can be smoothly extended through the singular curve, then the other side of the singular curve is of type III ((II)).

Theorem 1.9 plays a key point to enable us to recover the Bernstein-type theorem (see Section 8), which was first shown in the original paper [3] (or see [1, 7, 8]), and says that

u(x,y)=Ax+By+C,u(x,y)=Ax+By+C,

for some constants A,B,CA,B,C\in\operatorname{\mathbb{R}}, and

u(x,y)=ABx2+(A2B2)xy+ABy2+g(Bx+Ay),u(x,y)=-ABx^{2}+(A^{2}-B^{2})xy+ABy^{2}+g(-Bx+Ay),

where A,BA,B are constants such that A2+B2=1A^{2}+B^{2}=1 and gC()g\in C^{\infty}(\operatorname{\mathbb{R}}), are the only two classes of entire smooth solutions to the pp-minimal graph equation (1.4). In addition, in Section 7, we present some basic examples which, in particular, help us figure out the Bernstein-type theorem.

Finally, in Section 9, depending on a parametrized curve 𝒞(θ)=(x(θ),y(θ),z(θ))\mathcal{C}(\theta)=(x(\theta),y(\theta),z(\theta)) for θ\theta\in\operatorname{\mathbb{R}}, we deform the graph u=0u=0 in some way to construct pp-minimal surfaces with parametrization

(1.9) Y(r,θ)=(x(θ)+rcosθ,y(θ)+rsinθ,z(θ)+ry(θ)cosθrx(θ)sinθ),Y(r,\theta)=(x(\theta)+r\cos{\theta},y(\theta)+r\sin{\theta},z(\theta)+ry(\theta)\cos{\theta}-rx(\theta)\sin{\theta}),

for rr\in\operatorname{\mathbb{R}}. It is easy to checked that YY is an immersion if and only if either Θ(𝒞(θ))(y(θ)cosθx(θ)sinθ)20orr+(y(θ)cosθx(θ)sinθ)0\Theta(\mathcal{C}^{\prime}(\theta))-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2}\neq 0\ \textrm{or}\ r+(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})\neq 0 for all θ\theta (see Remark 9.2). In particular, the surface YY defines a pp-minimal surface of special type I if the curve 𝒞\mathcal{C} satisfies

(1.10) z(θ)+x(θ)y(θ)y(θ)x(θ)(y(θ)cosθx(θ)sinθ)2=0,z^{\prime}(\theta)+x(\theta)y^{\prime}(\theta)-y(\theta)x^{\prime}(\theta)-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2}=0,

for all θ\theta, or equivalently

(1.11) z(θ)=[(y(θ)cosθx(θ)sinθ)2+y(θ)x(θ)x(θ)y(θ)]𝑑θ.z(\theta)=\int\big{[}(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2}+y(\theta)x^{\prime}(\theta)-x(\theta)y^{\prime}(\theta)\big{]}d\theta.

In addition, the corresponding ζ1\zeta_{1}-invariant reads

(1.12) ζ1(θ)=y(θ)cosθx(θ)sinθ[x(θ)cosθ+y(θ)sinθ]𝑑θ,\zeta_{1}(\theta)=y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta}-\int\big{[}x^{\prime}(\theta)\cos{\theta}+y^{\prime}(\theta)\sin{\theta}\big{]}d\theta,

where [x(θ)cosθ+y(θ)sinθ]𝑑θ\int\big{[}x^{\prime}(\theta)\cos{\theta}+y^{\prime}(\theta)\sin{\theta}\big{]}d\theta is an anti-derivative of the function x(θ)cosθ+y(θ)sinθx^{\prime}(\theta)\cos{\theta}+y^{\prime}(\theta)\sin{\theta}.

Similarly, the surface YY defines a pp-minimal surface of general type if the curve 𝒞\mathcal{C} satisfies

(1.13) z(θ)+x(θ)y(θ)y(θ)x(θ)(y(θ)cosθx(θ)sinθ)20,z^{\prime}(\theta)+x(\theta)y^{\prime}(\theta)-y(\theta)x^{\prime}(\theta)-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2}\neq 0,

for all θ\theta. In addition, the corresponding ζ1\zeta_{1}- and ζ2\zeta_{2}-invariant read

(1.14) {ζ1(θ)=y(θ)cosθx(θ)sinθ[x(θ)cosθ+y(θ)sinθ]𝑑θζ2(θ)=z(θ)+x(θ)y(θ)y(θ)x(θ)(y(θ)cosθx(θ)sinθ)2.\left\{\begin{split}\zeta_{1}(\theta)&=y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta}-\int\big{[}x^{\prime}(\theta)\cos{\theta}+y^{\prime}(\theta)\sin{\theta}\big{]}d\theta\\ \zeta_{2}(\theta)&=z^{\prime}(\theta)+x(\theta)y^{\prime}(\theta)-y(\theta)x^{\prime}(\theta)-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2}.\end{split}\right.

From this construction, together with Theorem 7.8 which gives parametrizations for pp-minimal surfaces of special type II, we conclude that we have generically provided a parametrization for any given pp-minimal surface (see the argument in Section 9).

Acknowledgments. The first author’s research was supported in part by NCTS and in part by MOST 109-2115-M-007-004 -MY3. The second author’s research was supported in part by MOST 108-2115-M-032-008-MY2 and 110-2115-M-032-005-MY2.

2. The Fundamental Theorem for surfaces in H1H_{1}

In this section, we first review the fundamental theorem for surfaces in the Heisenberg group H1H_{1} (Theorem 2.1). For the details, we refer the reader to [6]. Next, we give another version (Theorem 1.1) of this theorem in terms of compatible coordinate systems.

2.1. The Fundamental Theorem for surfaces in H1H_{1}

Recall that there are four invariants for surfaces induced on a surface Σ\Sigma from the Heisenberg group H1H_{1}:

  1. (1)

    The first fundamental form (or the induced metric) II, which is the adapted metric gΘg_{\Theta} restricted to Σ\Sigma. This metric is actually defined on the whole surface Σ\Sigma.

  2. (2)

    The directed characteristic foliation e1e_{1}, which is a unit vector field TΣξ\in T\Sigma\cap\xi. This vector field is only defined on the regular part of Σ\Sigma.

  3. (3)

    The α\alpha-function α\alpha, which is a function defined on the regular part such that αe2+TTΣ\alpha e_{2}+T\in T\Sigma, where e2=Je1e_{2}=Je_{1}.

  4. (4)

    The pp-mean curvature HH, which is a function on the regular part defined by e1e2=He1\nabla_{e_{1}}e_{2}=-He_{1}.

These four invariants constitute a complete set of invariants for surfaces in H1H_{1}. That is, if ϕ:Σ1Σ2\phi:\Sigma_{1}\rightarrow\Sigma_{2} is a diffeomorphism between these two surfaces which preserves these four invariants, then ϕ\phi is the restriction of a Heisenberg rigid motion Φ\Phi. We have the integrability condition

(2.1) ω^1(e^1)2=Hα(1+α2)1/2,ω^1(e^2)2=2α+α(e^1α)1+α2,e^2H=e^1e^1α+6α(e^1α)+4α3+αH2(1+α2)1/2,\begin{split}\hat{\omega}_{1}{}^{2}(\hat{e}_{1})&=\frac{H\alpha}{(1+\alpha^{2})^{1/2}},\\ \hat{\omega}_{1}{}^{2}(\hat{e}_{2})&=2\alpha+\frac{\alpha(\hat{e}_{1}\alpha)}{1+\alpha^{2}},\\ \hat{e}_{2}H&=\frac{\hat{e}_{1}\hat{e}_{1}\alpha+6\alpha(\hat{e}_{1}\alpha)+4\alpha^{3}+\alpha H^{2}}{(1+\alpha^{2})^{1/2}},\end{split}

where e^2=αe2+T1+α2\hat{e}_{2}=\frac{\alpha e_{2}+T}{\sqrt{1+\alpha^{2}}}, which is nothing to do with the orientation of Σ\Sigma but a vector field and only determined by the contact form Θ\Theta. Moreover, e^1=e1\hat{e}_{1}=e_{1}, which is the characteristic direction and is determined up to a sign (if we choose e^1\hat{e}_{1} such that e^1e^2\hat{e}_{1}\wedge\hat{e}_{2} is compatible with the orientation of Σ\Sigma then e^1\hat{e}_{1} is unique). The form ω^12\hat{\omega}_{1}{}^{2} is the Levi-Civita connection form with respect to the frame {e^1,e^2}\{\hat{e}_{1},\hat{e}_{2}\}. We present the following fundamental theorem (see [6]).

Theorem 2.1 (The Fundamental theorem for surfaces in H1H_{1}).

Let (Σ,g)(\Sigma,g) be a Riemannian 22-manifold, and let α^,H^\hat{\alpha},\hat{H} be two real-valued functions on Σ\Sigma. Assume that gg, together with α^,H^\hat{\alpha},\hat{H}, satisfies the integrability condition (2.1), with α,H\alpha,H replaced by α^,H^\hat{\alpha},\hat{H}, respectively. Then for every point pΣp\in\Sigma, there exist an open neighborhood UU containing pp, and an embedding X:UH1X:U\rightarrow H_{1} such that g=X(I),α^=Xαg=X^{\ast}(I),\hat{\alpha}=X^{\ast}\alpha and H^=XH\hat{H}=X^{\ast}H. And X(e^1)X_{\ast}(\hat{e}_{1}) defines the foliation on X(U)X(U) induced from H1H_{1}. Moreover, XX is unique up to a Heisenberg rigid motion.

2.2. The new version of the integrability conditions

The goal of this subsection is to express the integrability condition (2.1) in terms of a compatible coordinate system (x,y)(x,y). We write

(2.2) e^2=a(x,y)x+b(x,y)y,\hat{e}_{2}=a(x,y)\frac{\partial}{\partial x}+b(x,y)\frac{\partial}{\partial y},

for some functions aa and b0b\neq 0. We can assume, without loss of generality, that b>0b>0, that is, both xy\frac{\partial}{\partial x}\wedge\frac{\partial}{\partial y} and e^1e^2\hat{e}_{1}\wedge\hat{e}_{2} define the same orientation on Σ\Sigma. The two functions aa and bb are a representation of the first fundamental form II. The dual co-frame {ω^1,ω^2}\{\hat{\omega}^{1},\hat{\omega}^{2}\} of {e^1,e^2}\{\hat{e}_{1},\hat{e}_{2}\} is

(2.3) ω^1=dxabdy,ω^2=1bdy.\begin{split}\hat{\omega}^{1}&=dx-\frac{a}{b}dy,\\ \hat{\omega}^{2}&=\frac{1}{b}dy.\end{split}

Then the Levi-Civita connection forms are uniquely determined by the following Riemannian structure equations

(2.4) dω^1=ω^2ω^2,1dω^2=ω^1ω^1,2\begin{split}d\hat{\omega}^{1}&=\hat{\omega}^{2}\wedge\hat{\omega}_{2}{}^{1},\\ d\hat{\omega}^{2}&=\hat{\omega}^{1}\wedge\hat{\omega}_{1}{}^{2},\end{split}

with the normalized condition

(2.5) ω^1+2ω^2=10.\hat{\omega}_{1}{}^{2}+\hat{\omega}_{2}{}^{1}=0.

A computation shows that

(2.6) dω^1=d(ab)dy=(bdaadbb2)dy=dybbdaadbb=ω^2bdaadbb.\begin{split}d\hat{\omega}^{1}&=-d\left(\frac{a}{b}\right)\wedge dy=-\left(\frac{bda-adb}{b^{2}}\right)\wedge dy\\ &=\frac{dy}{b}\wedge\frac{bda-adb}{b}=\hat{\omega}^{2}\wedge\frac{bda-adb}{b}.\end{split}

By comparing with the first equation of the Riemannian structure equations (2.4), we have

(2.7) ω^2=1bdaadbb+a11ω^2\hat{\omega}_{2}{}^{1}=\frac{bda-adb}{b}+a_{11}\hat{\omega}^{2}

for some function a11a_{11}. On one hand, the second equation of (2.4) and the normalized condition (2.5) imply

(2.8) dω^2=ω^1ω^12=(dxabdy)(bdaadbb+a11ω^2)=(ay+abbya11babax+a2b2bx)dxdy.\begin{split}d\hat{\omega}^{2}&=\hat{\omega}^{1}\wedge\hat{\omega}_{1}{}^{2}\\ &=-(dx-\frac{a}{b}dy)\wedge\left(\frac{bda-adb}{b}+a_{11}\hat{\omega}^{2}\right)\\ &=\left(-a_{y}+\frac{a}{b}b_{y}-\frac{a_{11}}{b}-\frac{a}{b}a_{x}+\frac{a^{2}}{b^{2}}b_{x}\right)dx\wedge dy.\end{split}

On the other hand, one sees

(2.9) dω^2=(d1b)dy=bxb2dxdy.d\hat{\omega}^{2}=\left(d\frac{1}{b}\right)\wedge dy=-\frac{b_{x}}{b^{2}}dx\wedge dy.

Equations (2.8) and (2.9) yield

(2.10) a11=bxbbay+abyaax+a2bbx.a_{11}=\frac{b_{x}}{b}-ba_{y}+ab_{y}-aa_{x}+\frac{a^{2}}{b}b_{x}.

Therefore,

(2.11) ω^21=bdaadbb+a11ω^2=(baxabx)dxb+(bayaby)dyb+(bxbbay+abyaax+a2bbx)dyb=(baxabxb)dx+(bxb2aaxb+a2bxb2)dy.\begin{split}\hat{\omega}_{2}{}^{1}&=\frac{bda-adb}{b}+a_{11}\hat{\omega}^{2}\\ &=(ba_{x}-ab_{x})\frac{dx}{b}+(ba_{y}-ab_{y})\frac{dy}{b}+\left(\frac{b_{x}}{b}-ba_{y}+ab_{y}-aa_{x}+\frac{a^{2}}{b}b_{x}\right)\frac{dy}{b}\\ &=\left(\frac{ba_{x}-ab_{x}}{b}\right)dx+\left(\frac{b_{x}}{b^{2}}-\frac{aa_{x}}{b}+\frac{a^{2}b_{x}}{b^{2}}\right)dy.\end{split}

From the connection form formula (2.11), we have

(2.12) ω^1(e^1)2=abxbaxb=ax+abxb,ω^1(e^2)2=a(abxbax)b+b(aaxbbxb2a2bxb2)=bxb.\begin{split}\hat{\omega}_{1}{}^{2}(\hat{e}_{1})&=\frac{ab_{x}-ba_{x}}{b}=-a_{x}+a\frac{b_{x}}{b},\\ \hat{\omega}_{1}{}^{2}(\hat{e}_{2})&=\frac{a(ab_{x}-ba_{x})}{b}+b\left(\frac{aa_{x}}{b}-\frac{b_{x}}{b^{2}}-\frac{a^{2}b_{x}}{b^{2}}\right)=-\frac{b_{x}}{b}.\end{split}

Therefore, in terms of a compatible coordinate system (U;x,y)(U;x,y), the integrability condition (2.1) is equivalent to

(2.13) ax+abxb=Hα(1+α2)1/2,bxb=2α+ααx1+α2,aHx+bHy=αxx+6ααx+4α3+αH2(1+α2)1/2.\begin{split}-a_{x}+a\frac{b_{x}}{b}&=\frac{H\alpha}{(1+\alpha^{2})^{1/2}},\\ -\frac{b_{x}}{b}&=2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}},\\ aH_{x}+bH_{y}&=\frac{\alpha_{xx}+6\alpha\alpha_{x}+4\alpha^{3}+\alpha H^{2}}{(1+\alpha^{2})^{1/2}}.\end{split}

2.3. The computation of the first fundamental form II

We would like to solve the first two equations of (2.13), which are part of the integrability condition. From the second equation of (2.13), it is easy to see that

(2.14) ln|b|=(2α+ααx1+α2)dx+k(y)\ln{|b|}=\int-\left(2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}}\right)dx+k(y)

for some function k(y)k(y) in the variable yy, that is

(2.15) |b|=ek(y)e(2α+ααx1+α2)dx,=ek(y)e2α𝑑x(1+α2)12,\begin{split}|b|&=e^{k(y)}e^{\int-\left(2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}}\right)dx},\\ &=e^{k(y)}\frac{e^{-\int 2\alpha dx}}{(1+\alpha^{2})^{\frac{1}{2}}},\end{split}

where 2α𝑑x\int 2\alpha dx is an anti-derivative of 2α2\alpha with respect to xx. Throughout this paper, we always assume, without loss of generality, that b>0b>0. For aa, we substitute the second equation of (2.13) into the first one to obtain the first-order linear ODE

(2.16) ax+a(2α+ααx1+α2)+Hα(1+α2)1/2=0.a_{x}+a\left(2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}}\right)+\frac{H\alpha}{(1+\alpha^{2})^{1/2}}=0.

To solve aa, we choose the integrating factor u=e(2α+ααx1+α2)𝑑xu=e^{\int\left(2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}}\right)dx} such that the one-form

(2.17) u([a(2α+ααx1+α2)+Hα(1+α2)1/2]dx+da)u\left(\left[a\left(2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}}\right)+\frac{H\alpha}{(1+\alpha^{2})^{1/2}}\right]dx+da\right)

is an exact form. Therefore, using the standard method of ODE, one sees that

(2.18) a=e(2α+ααx1+α2)dx(h(y)Hα(1+α2)1/2e(2α+ααx1+α2)𝑑x𝑑x)=e2α𝑑x(1+α2)12(h(y)Hα(e2α𝑑x)𝑑x),\begin{split}a&=e^{\int-\left(2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}}\right)dx}\left(h(y)-\int\frac{H\alpha}{(1+\alpha^{2})^{1/2}}e^{\int\left(2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}}\right)dx}dx\right)\\ &=\frac{e^{-\int 2\alpha dx}}{(1+\alpha^{2})^{\frac{1}{2}}}\left(h(y)-\int H\alpha\left(e^{\int 2\alpha dx}\right)dx\right),\end{split}

for some function h(y)h(y) in yy and Hα(e2α𝑑x)𝑑x\int H\alpha\left(e^{\int 2\alpha dx}\right)dx is an anti-derivative of Hα(e2α𝑑x)H\alpha\left(e^{\int 2\alpha dx}\right) with respect to xx. From (2.18) and (2.15), we conclude that the first fundamental form II (or aa and bb) is determined by α\alpha and HH, up to two functions k(y)k(y) and h(y)h(y). We are thus ready to prove a more useful version of the fundamental theorem for surfaces (see Theorem 1.1).

2.4. The proof of Theorem 1.1 with arbitrary α\alpha and HH

We define a Riemannian metric on UU by regarding {x,a(x,y)x+b(x,y)y}\{\frac{\partial}{\partial x},a(x,y)\frac{\partial}{\partial x}+b(x,y)\frac{\partial}{\partial y}\} as an orthonormal frame field, where aa and bb are specified as (2.15) and (2.18) with h(y)h(y) and k(y)k(y) given in (1.3). Then it is easy to see that α\alpha and HH, together with aa and bb, satisfy the integrability condition (2.13), and hence, by the fundamental theorem for surfaces (see Theorem 2.1), UU can be embedded uniquely as a surface with HH and α\alpha as its pp-mean curvature and α\alpha-function, respectively. In addition, the characteristic direction e^1=x\hat{e}_{1}=\frac{\partial}{\partial x} and e^2=a(x,y)x+b(x,y)y\hat{e}_{2}=a(x,y)\frac{\partial}{\partial x}+b(x,y)\frac{\partial}{\partial y} define the induced metric on the embedded surface, as desired (see Theorem 2.1 for the original version).

2.5. The Transformation law of invariants

First of all, we compute the transformation law of compatible coordinate systems. Let (x,y)(x,y) and (x~,y~)(\tilde{x},\tilde{y}) be two compatible coordinate systems and ϕ\phi a coordinate transformation, i.e., (x~,y~)=ϕ(x,y)(\tilde{x},\tilde{y})=\phi(x,y). Then we have ϕx=x~\phi_{*}\frac{\partial}{\partial x}=\frac{\partial}{\partial\tilde{x}}, which means that

[ϕ](10)=(x~xx~yy~xy~y)(10)=(10),[\phi_{*}]\left(\begin{array}[]{c}1\\ 0\end{array}\right)=\left(\begin{array}[]{cc}\tilde{x}_{x}&\tilde{x}_{y}\\ \tilde{y}_{x}&\tilde{y}_{y}\end{array}\right)\left(\begin{array}[]{c}1\\ 0\end{array}\right)=\left(\begin{array}[]{c}1\\ 0\end{array}\right),

where [ϕ][\phi_{*}] is the matrix representation of ϕ\phi_{*} with respect to these two coordinate systems. We then have the coordinates transformation

(2.19) x~=x+Γ(y),y~=Ψ(y),\tilde{x}=x+\Gamma(y),\ \ \ \tilde{y}=\Psi(y),

for some functions Γ(y)\Gamma(y) and Ψ(y)\Psi(y). Since det[ϕ]=Ψ(y)\det{[\phi_{*}]}=\Psi^{{}^{\prime}}(y), we immediately have that Ψ(y)0\Psi^{{}^{\prime}}(y)\neq 0 for all yy. Next, we compute the transformation law of representations of the induced metrics. Suppose that the representations of the induced metric are, respectively, given by a,ba,b and a~,b~\tilde{a},\tilde{b}, that is, e^2=ax+by=a~x~+b~y~\hat{e}_{2}=a\frac{\partial}{\partial x}+b\frac{\partial}{\partial y}=\tilde{a}\frac{\partial}{\partial\tilde{x}}+\tilde{b}\frac{\partial}{\partial\tilde{y}}. By (2.19), we have the following transformation law of the induced metric as

(2.20) a~=a+bΓ(y),b~=bΨ(y).\tilde{a}=a+b\Gamma^{\prime}(y),\ \ \ \tilde{b}=b\Psi^{\prime}(y).

Notice that we have omitted the sign of pull-back ϕ\phi^{*} on a~\tilde{a} and b~\tilde{b}. Since the pp-mean curvature and α\alpha-function are function-type invariants, they transform by pull-back.

3. Constant pp-mean curvature surfaces

In this section, we aim to prove Theorem 1.1 with constant HH and then provide a new tool to study the constant pp-mean curvature surfaces. More precisely, we will indicate that one can convert the investigation of the constant pp-mean curvature surfaces into the study of the so-called Codazzi-like equation.

3.1. The proof of Theorem 1.1 with constant HH

Let ΣH1\Sigma\subset H_{1} be a constant pp-mean curvature surface with H=cH=c. Then, in terms of a compatible coordinate system (U;x,y)(U;x,y), the integrability condition (2.13) is reduced to

(3.1) ax+abxb=cα(1+α2)1/2,bxb=2α+ααx1+α2,αxx+6ααx+4α3+c2α=0.\begin{split}-a_{x}+a\frac{b_{x}}{b}&=\frac{c\alpha}{(1+\alpha^{2})^{1/2}},\\ -\frac{b_{x}}{b}&=2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}},\\ \alpha_{xx}+6\alpha\alpha_{x}&+4\alpha^{3}+c^{2}\alpha=0.\end{split}

In other words, there exists the α\alpha satisfying the Codazzi-like equation

(3.2) αxx+6ααx+4α3+c2α=0,\alpha_{xx}+6\alpha\alpha_{x}+4\alpha^{3}+c^{2}\alpha=0,

which is a nonlinear ordinary differential equation. Conversely, given an arbitrary function α(x,y)\alpha(x,y) on a coordinate neighborhood (U,x,y)2(U,x,y)\subset\operatorname{\mathbb{R}}^{2} which satisfies the Codazzi-like equation (3.2) (or (1.2)). It is easy to see that equation (3.2) is just the equation (1.3) in the constant pp-mean curvature cases. Namely, that α\alpha satisfies equation (3.2) means it satisfies (1.3) for arbitrary functions h(y)h(y) and k(y)k(y). Therefore, the embedding Σ=X(U)\Sigma=X(U) in Theorem 1.1 is a constant pp-mean curvature surface with H=cH=c, the given function α(x,y)\alpha(x,y) as its α\alpha-function, and e^1=x\hat{e}_{1}=\frac{\partial}{\partial x}. Notice that, given (2.15) and (2.18), the constant pp-mean curvature surface determined by the given function α\alpha usually is not unique. They depend on two functions h(y)h(y) and k(y)k(y) in yy. The assertion is completed.

Throughout the rest of this paper, we will apply the tool to the subject of the pp-minimal surfaces.

4. Solutions to the Liénard equation

Since we bring in a strategy to study pp-minimal surfaces using the understanding of the Codazzi-like equation (4.1), we focus on studying this equation in this section. First of all, we suppose that α\alpha is regarded as a function in xx and we want to discuss the solutions to the Codazzi-like equation

(4.1) αxx+6ααx+4α3=0.\alpha_{xx}+6\alpha\alpha_{x}+4\alpha^{3}=0.

This is actually one kind of the so-called Liénard equation [2]. Derivation of explicit solutions (see Theorem 1.2) to the equation (4.1) is given below.

4.1. The proof of Theorem 1.2

Using v=αv=\alpha^{\prime} to see that (4.1) becomes

(4.2) vdvdα+6αv+4α3=0,v\frac{dv}{d\alpha}+6\alpha v+4\alpha^{3}=0,

which is the second kind of the Abel equation (cf. [9, 10]). Apparently, given v0v\neq 0, equation (4.2) can be written as

(4.3) dvdα=2α(3v+2α2)v.\frac{dv}{d\alpha}=-\frac{2\alpha(3v+2\alpha^{2})}{v}.

Denote u=1vu=\frac{1}{v}. Then (4.2) or (4.3) becomes

(4.4) dudα=6αu2+4α3u3.\frac{du}{d\alpha}=6\alpha u^{2}+4\alpha^{3}u^{3}.

We apply Chiellini’s integrability condition (stated in [11, 12]) for the Abel equation to (4.4), which is exactly integrated with k=290k=\frac{2}{9}\neq 0. It can be checked that

ddα(4α36α)=ddα(23α2)=43α=29(6α).\frac{d}{d\alpha}\left(\frac{4\alpha^{3}}{6\alpha}\right)=\frac{d}{d\alpha}\left(\frac{2}{3}\alpha^{2}\right)=\frac{4}{3}\alpha=\frac{2}{9}(6\alpha).

Let u=6α4α3ωu=\frac{6\alpha}{4\alpha^{3}}\omega, i.e., ω=23α2u\omega=\frac{2}{3}\alpha^{2}u and ω0\omega\neq 0. The equation (4.4) turns out to be

(4.5) dωdα=ωα(2+9ω+9ω2),\frac{d\omega}{d\alpha}=\frac{\omega}{\alpha}(2+9\omega+9\omega^{2}),

which is a separable first-order ODE, i.e.,

(4.6) dωω(2+9ω+9ω2)=dαα.\frac{d\omega}{\omega(2+9\omega+9\omega^{2})}=\frac{d\alpha}{\alpha}.

Using the method of partial fractions to integrate the left-hand side, we have

(4.7) (12ω33ω+1+32(3ω+2))𝑑ω=dαα,\int\left(\frac{1}{2\omega}-\frac{3}{3\omega+1}+\frac{3}{2(3\omega+2)}\right)d\omega=\int\frac{d\alpha}{\alpha},

which gives

(4.8) 12(ln|ω(3ω+2)||3ω+1|2)=ln|α|+const.\frac{1}{2}\left(\ln\frac{|\omega(3\omega+2)|}{|3\omega+1|^{2}}\right)=\ln|\alpha|+const.

The implicit solution to (4.5) is then expressed as

(4.9) ω(3ω+2)(3ω+1)2=Cα2,\frac{\omega(3\omega+2)}{(3\omega+1)^{2}}=C\alpha^{2},

provided that ω0,13,23\omega\neq 0,-\frac{1}{3},-\frac{2}{3}, and CC is an arbitrary nonzero constant. Hence, with the assumption α0\alpha\neq 0, we have

(4.10) 4α2(3Cα21)u2+4(3Cα21)u+3C=0,4\alpha^{2}(3C\alpha^{2}-1)u^{2}+4(3C\alpha^{2}-1)u+3C=0,

or equivalently,

(4.11) 3C(α)2+4(3Cα21)α+4α2(3Cα21)=0.3C(\alpha^{\prime})^{2}+4(3C\alpha^{2}-1)\alpha^{\prime}+4\alpha^{2}(3C\alpha^{2}-1)=0.

This yields

(4.12) α=2(13Cα2)±213Cα23C=(12Cα2)±12Cα2C,\begin{split}\alpha^{\prime}&=\frac{2(1-3C\alpha^{2})\pm 2\sqrt{1-3C\alpha^{2}}}{3C}\\ &=\frac{(1-2C\alpha^{2})\pm\sqrt{1-2C\alpha^{2}}}{C},\end{split}

for a nonzero constant CC. Since we have assumed that v=αv=\alpha^{\prime} is not zero, neither is 12Cα21-2C\alpha^{2}. To obtain the general solutions, we proceed to solve equation (4.12) by means of the variable separable method. We rewrite (4.12) as

(4.13) dxC=dα(12Cα2)±12Cα2=(12Cα2)12Cα2[(12Cα2)±12Cα2][(12Cα2)12Cα2]dα=(12Cα2)12Cα22Cα2(12Cα2)dα=(12Cα2±12Cα212Cα2)dα.\begin{split}\frac{dx}{C}&=\frac{d\alpha}{(1-2C\alpha^{2})\pm\sqrt{1-2C\alpha^{2}}}\\ &=\frac{(1-2C\alpha^{2})\mp\sqrt{1-2C\alpha^{2}}}{[(1-2C\alpha^{2})\pm\sqrt{1-2C\alpha^{2}}][(1-2C\alpha^{2})\mp\sqrt{1-2C\alpha^{2}}]}d\alpha\\ &=\frac{(1-2C\alpha^{2})\mp\sqrt{1-2C\alpha^{2}}}{-2C\alpha^{2}(1-2C\alpha^{2})}d\alpha\\ &=\left(\frac{-1}{2C\alpha^{2}}\pm\frac{1}{2C\alpha^{2}\sqrt{1-2C\alpha^{2}}}\right)d\alpha.\end{split}

Case I. If C<0C<0, we use the trigonometric substitution 2|C|α=tanθ,π2<θ<π2\sqrt{2|C|}\alpha=\tan{\theta},\ -\frac{\pi}{2}<\theta<\frac{\pi}{2}, to get

(4.14) dα2Cα212Cα2=12Cα22Cα+c1,\int\frac{d\alpha}{2C\alpha^{2}\sqrt{1-2C\alpha^{2}}}=-\frac{\sqrt{1-2C\alpha^{2}}}{2C\alpha}+c_{1},

for some c1c_{1}\in\mathbb{R}. Substituting (4.14) into (4.13), we obtain

x+c1C=112Cα22Cα,\frac{x+c_{1}}{C}=\frac{1\mp\sqrt{1-2C\alpha^{2}}}{2C\alpha},

that is,

(4.15) 2α(x+c1)1=12Cα2,2\alpha(x+c_{1})-1=\mp\sqrt{1-2C\alpha^{2}},

for some c1c_{1}\in\mathbb{R}. Taking the square of both sides and noticing that α0\alpha\neq 0, we obtain

(4.16) α(x)=x+c1(x+c1)2+c2,\alpha(x)=\frac{x+c_{1}}{(x+c_{1})^{2}+c_{2}},

for some c1,c2c_{1},\ c_{2}\in\mathbb{R} and c2<0c_{2}<0. If α\alpha satisfies 2α(x+c1)1=+12Cα22\alpha(x+c_{1})-1=+\sqrt{1-2C\alpha^{2}} in (4.15), then we have α(x+c1)>0\alpha(x+c_{1})>0, and hence (4.16) implies x+c1<|c2|x+c_{1}<-\sqrt{|c_{2}|} or |c2|<x+c1\sqrt{|c_{2}|}<x+c_{1}. On the other hand, if α\alpha satisfies 2α(x+c1)1=12Cα22\alpha(x+c_{1})-1=-\sqrt{1-2C\alpha^{2}}, then we have α(x+c1)<0\alpha(x+c_{1})<0, and we then obtain that |c2|<x+c1<|c2|-\sqrt{|c_{2}|}<x+c_{1}<\sqrt{|c_{2}|} by (4.16).
Case II. If C>0C>0, we use the trigonometric substitution 2Cα=sinθ,π2<θ<π2\sqrt{2C}\alpha=\sin{\theta},\ -\frac{\pi}{2}<\theta<\frac{\pi}{2}, to get

(4.17) dα2Cα212Cα2=12Cα22Cα+c1,\int\frac{d\alpha}{2C\alpha^{2}\sqrt{1-2C\alpha^{2}}}=-\frac{\sqrt{1-2C\alpha^{2}}}{2C\alpha}+c_{1},

for some c1c_{1}\in\mathbb{R}. Substituting (4.17) into (4.13), we obtain

x+c1C=112Cα22Cα,\frac{x+c_{1}}{C}=\frac{1\mp\sqrt{1-2C\alpha^{2}}}{2C\alpha},

that is,

(4.18) 2α(x+c1)1=12Cα2,2\alpha(x+c_{1})-1=\mp\sqrt{1-2C\alpha^{2}},

for some c1c_{1}\in\mathbb{R}. Taking the square of both sides and noticing that α0\alpha\neq 0, we obtain

(4.19) α(x)=x+c1(x+c1)2+c2,\alpha(x)=\frac{x+c_{1}}{(x+c_{1})^{2}+c_{2}},

for some c1,c2c_{1},\ c_{2}\in\mathbb{R} and c2>0c_{2}>0.

As above, while we try to get the general solutions to (4.1), we have assumed that α0,α0,ω0,ω13\alpha^{\prime}\neq 0,\alpha\neq 0,\omega\neq 0,\omega\neq-\frac{1}{3} and ω23\omega\neq-\frac{2}{3}. Now suppose α=0\alpha^{\prime}=0 on an open interval, then (4.1) immediately implies α=0\alpha=0 on that interval. It is easy to see that ω=0\omega=0 is equivalent to α=0\alpha=0. Finally, since ω=23α2u\omega=\frac{2}{3}\alpha^{2}u, we see that

ω=13α2u=12α=2α2α(x)=12(x+c1).\omega=-\frac{1}{3}\Leftrightarrow\alpha^{2}u=\frac{-1}{2}\Leftrightarrow\alpha^{\prime}=-2\alpha^{2}\Leftrightarrow\alpha(x)=\frac{1}{2(x+c_{1})}.

Similarly, we have

ω=23α(x)=1(x+c1),\omega=-\frac{2}{3}\Leftrightarrow\alpha(x)=\frac{1}{(x+c_{1})},

for some c1Rc_{1}\in R. We hence complete the proof of Theorem 1.2. A similar argument establishes the following proof.

4.2. The proof of Theorem 1.4

Using a similar transformation between v,uv,u and ω\omega, the ODE (1.2) is converted to be

(4.20) dωω(2+9ω+9ω2)=4α4α2+c2dα(ω=4α3+c2α6αu),\frac{d\omega}{\omega(2+9\omega+9\omega^{2})}=\frac{4\alpha}{4\alpha^{2}+c^{2}}d\alpha\quad(\omega=\frac{4\alpha^{3}+c^{2}\alpha}{6\alpha}u),

which has the implicit solution expressed as

(4.21) ω(3ω+2)(3ω+1)2=K(4α2+c2),\frac{\omega(3\omega+2)}{(3\omega+1)^{2}}=K(4\alpha^{2}+c^{2}),

provided that ω0,13,23\omega\neq 0,-\frac{1}{3},-\frac{2}{3}, and K=econstK=e^{const} is arbitrary nonzero constant. If ω=0,13\omega=0,-\frac{1}{3} or 23-\frac{2}{3}, we have the trivial solution or the special solutions derived. Hence, with the assumption α0\alpha\neq 0, we have

(4.22) α=13K(4α2+c2)±13K(4α2+c2)6K,\alpha^{\prime}=\frac{1-3K(4\alpha^{2}+c^{2})\pm\sqrt{1-3K(4\alpha^{2}+c^{2})}}{6K},

for a nonzero constant KK. Since we have assumed that v=α0v=\alpha^{\prime}\neq 0, neither is 13K(4α2+c2)1-3K(4\alpha^{2}+c^{2}). In order to obtain the general solutions, we rationalize (4.22) as

(4.23) dx6K=dα(13K(4α2+c2))±13K(4α2+c2)=(13K(4α2+c2)±13K(4α2+c2)13K(4α2+c2))dα.\begin{split}\frac{dx}{6K}&=\frac{d\alpha}{(1-3K(4\alpha^{2}+c^{2}))\pm\sqrt{1-3K(4\alpha^{2}+c^{2})}}\\ &=\left(\frac{-1}{3K(4\alpha^{2}+c^{2})}\pm\frac{1}{3K(4\alpha^{2}+c^{2})\sqrt{1-3K(4\alpha^{2}+c^{2})}}\right)d\alpha.\end{split}

Next, we will deal with integrations of both sides. Note that

(4.24) 13K(4α2+c2)=(13Kc2)12Kα20,1-3K(4\alpha^{2}+c^{2})=(1-3Kc^{2})-12K\alpha^{2}\geq 0,

which implies

13Kc212Kα20.1-3Kc^{2}\geq 12K\alpha^{2}\geq 0.

Moreover, that K>0K>0 implies 03K(4α2+c2)0\leq 3K(4\alpha^{2}+c^{2}). We further assume that 3K(4α2+c2)13K(4\alpha^{2}+c^{2})\leq 1.

  • If 13Kc2=01-3Kc^{2}=0, then 3Kc2=13Kc^{2}=1, which yields 3K(4α2+c2)=12Kα2+1>13K(4\alpha^{2}+c^{2})=12K\alpha^{2}+1>1, a contradiction.

  • If 13Kc2>01-3Kc^{2}>0, then we let

    (4.25) 12Kα=13Kc2sinθ.\sqrt{12K}\alpha=\sqrt{1-3Kc^{2}}\sin\theta.

    This gives

    (4.26) dα=13Kc212Kcosθdθ,13K(4α2+c2)=(13Kc2)cos2θ.d\alpha=\sqrt{\frac{1-3Kc^{2}}{12K}}\cos\theta d\theta,\quad 1-3K(4\alpha^{2}+c^{2})=(1-3Kc^{2})\cos^{2}\theta.

    It turns out that the integration of 13K(4α2+c2)13K(4α2+c2)dα\frac{1}{3K(4\alpha^{2}+c^{2})\sqrt{1-3K(4\alpha^{2}+c^{2})}}d\alpha becomes

    (4.27) 112Kdθ1(13Kc2)cos2θ=112K(tan1(tanθ3Kc2)3Kc2)+const.\frac{1}{\sqrt{12K}}\int\frac{d\theta}{1-(1-3Kc^{2})\cos^{2}\theta}=\frac{1}{\sqrt{12K}}\left(-\frac{\tan^{-1}\left(\frac{\tan\theta}{\sqrt{3Kc^{2}}}\right)}{\sqrt{3Kc^{2}}}\right)+const.

    By (4.25), we have

    (4.28) 112K(tan1(2α|c|13K(4α2+c2))3Kc2)+const..\frac{1}{\sqrt{12K}}\left(\frac{\tan^{-1}\left(\frac{2\alpha}{|c|\sqrt{1-3K(4\alpha^{2}+c^{2})}}\right)}{\sqrt{3Kc^{2}}}\right)+const..

It is easy to see that

dα3K(4α2+c2)=13Ktan1(2αc)2c+const..\int\frac{d\alpha}{3K(4\alpha^{2}+c^{2})}=\frac{1}{3K}\frac{\tan^{-1}\left(\frac{2\alpha}{c}\right)}{2c}+const..

Therefore, (4.23) shows

(4.29) x6K=13Ktan1(2αc)2c±16K|c|(tan1(2α|c|13K(4α2+c2)))+C,\frac{x}{6K}=-\frac{1}{3K}\frac{\tan^{-1}\left(\frac{2\alpha}{c}\right)}{2c}\pm\frac{1}{6K|c|}\left(\tan^{-1}\left(\frac{2\alpha}{|c|\sqrt{1-3K(4\alpha^{2}+c^{2})}}\right)\right)+C,

where CC is a constant. Note that the sum/difference identity of tan(x±y)\tan(x\pm y) implies

(4.30) tan1A±tan1B=tan1(A±B1AB)+nπ,n.\tan^{-1}A\pm\tan^{-1}B=\tan^{-1}\left(\frac{A\pm B}{1\mp AB}\right)+n\pi,\quad n\in\mathbb{Z}.

Assuming c>0c>0, (4.29) becomes

cx+C1+nπ=tan1(2cα13K(4α2+c2)2cαc213K(4α2+c2)±4α2),-cx+C_{1}+n\pi=\tan^{-1}\left(\frac{2c\alpha\sqrt{1-3K(4\alpha^{2}+c^{2})}\mp 2c\alpha}{c^{2}\sqrt{1-3K(4\alpha^{2}+c^{2})}\pm 4\alpha^{2}}\right),

which is

tan(cxc1)=2cα13K(4α2+c2)2cαc213K(4α2+c2)±4α2.\tan(-cx-c_{1})=\frac{2c\alpha\sqrt{1-3K(4\alpha^{2}+c^{2})}\mp 2c\alpha}{c^{2}\sqrt{1-3K(4\alpha^{2}+c^{2})}\pm 4\alpha^{2}}.

After a direct computation, the square sum of the denominator and numerator is equal to

(4α2+c2)2(13c2K),(4\alpha^{2}+c^{2})^{2}(1-3c^{2}K),

and hence we have

sin(cxc1)=2αc13K(4α2+c2)2αc(4α2+c2)(13c2K),and cos(cxc1)=c213K(4α2+c2)±4α2(4α2+c2)(13c2K).\begin{split}\sin{(-cx-c_{1})}&=\frac{2\alpha c\sqrt{1-3K(4\alpha^{2}+c^{2})}\mp 2\alpha c}{(4\alpha^{2}+c^{2})\sqrt{(1-3c^{2}K)}},\\ \mbox{and }\cos{(-cx-c_{1})}&=\frac{c^{2}\sqrt{1-3K(4\alpha^{2}+c^{2})}\pm 4\alpha^{2}}{(4\alpha^{2}+c^{2})\sqrt{(1-3c^{2}K)}}.\end{split}

If we let

c2=±113c2K,c_{2}=\pm\frac{1}{\sqrt{1-3c^{2}K}},

and notice that c=2λc=2\lambda, we compute

λsin(cxc1)c2cos(cxc1)=α2c213K(4α2+c2)±2c22c2(4α2+c2)13c2K2c213K(4α2+c2)8α2=α\begin{split}&-\lambda\frac{\sin{(-cx-c_{1})}}{c_{2}-\cos{(-cx-c_{1})}}\\ =&\alpha\frac{-2c^{2}\sqrt{1-3K(4\alpha^{2}+c^{2})}\pm 2c^{2}}{2c_{2}(4\alpha^{2}+c^{2})\sqrt{1-3c^{2}K}-2c^{2}\sqrt{1-3K(4\alpha^{2}+c^{2})}\mp 8\alpha^{2}}\\ =&\alpha\end{split}

If K<0K<0, then 13K(4α2+c2)01-3K(4\alpha^{2}+c^{2})\geq 0 automatically. It is easy to see that either 13K(4α2+c2)=01-3K(4\alpha^{2}+c^{2})=0 or 13K(4α2+c2)>01-3K(4\alpha^{2}+c^{2})>0. If 13K(4α2+c2)=01-3K(4\alpha^{2}+c^{2})=0, then (13Kc2)+(12Kα2)=0(1-3Kc^{2})+(-12K\alpha^{2})=0, which yields 13Kc2=01-3Kc^{2}=0 since both (13Kc2)(1-3Kc^{2}) and 12Kα2-12K\alpha^{2} are non-negative. We use (4.22) to have α=0\alpha^{\prime}=0, i.e., α\alpha is a constant w.r.t. xx, which gives α=0\alpha=0. For the later case, 13K(4α2+c2)>01-3K(4\alpha^{2}+c^{2})>0, and we further assume that 13Kc2>01-3Kc^{2}>0, otherwise, we get α=0\alpha=0. Direction calculations give the same equation (4.29).

4.3. The phase plane

We remark that when c=0c=0 (the case of pp-minimal surfaces), (3.2) is one of the so-called Liénard equation [2]

(4.31) αxx=f(α,αx)=(6ααx+4α3).\alpha_{xx}=f(\alpha,\alpha_{x})=-(6\alpha\alpha_{x}+4\alpha^{3}).

If we imagine a simple dynamical system consisting of a particle of unit mass moving on the α\alpha-axis, and if f(α,αx)f(\alpha,\alpha_{x}) is the force acting on it, then (4.31) is the equation of motion. The values of α\alpha (position) and αx\alpha_{x} (velocity), which at each instant characterize the state of the system, are called its phases, and the plane of the variables α\alpha and αx\alpha_{x} is called the phase plane. Using v=αxv=\alpha_{x} to see that (4.31) can be replaced by the equivalent system

(4.32) {dαdx=v,dvdx=(6αv+4α3).\left\{\begin{split}\frac{d\alpha}{dx}&=v,\\ \frac{dv}{dx}&=-(6\alpha v+4\alpha^{3}).\end{split}\right.

In general, a solution of (4.32) is a pair of functions α(x)\alpha(x) and v(x)v(x) defining a curve on the phase plane. It follows from the standard theory of ODE that if x0x_{0} is any number and (α0,v0)(\alpha_{0},v_{0}) is any point in the phase plane, then there exists a unique solution (α(x),v(x))(\alpha(x),v(x)) of (4.32) such that α(x0)=α0\alpha(x_{0})=\alpha_{0} and v(x0)=v0v(x_{0})=v_{0}. If this solution (α(x),v(x))(\alpha(x),v(x)) is not a constant, then it defines a curve on the phase plane called a path of the system; otherwise, it defines a critical point. All paths together with critical points form a directed singular foliation on the phase plane with critical points as singular points of the foliation, and each path lies in a leaf of the foliation. It is easy to see that this foliation is defined by the vector field V=(v,(6αv+4α3))V=(v,-(6\alpha v+4\alpha^{3})) and (0,0)(0,0) is the only critical point. We express the direction field (or the directed singular foliation) as Figure 4.2 (Figure 4.2 for c=1.5c=1.5) as follows.

Refer to caption
Figure 4.1. direction field VV
Refer to caption
Figure 4.2. direction field for c=32c=\frac{3}{2}

5. The classification of constant pp-mean curvature surfaces

5.1. The classification of pp-minimal surfaces

Theorem 1.2 suggests we divide (locally) the pp-minimal surfaces into several classes. In terms of compatible coordinates (x,y)(x,y), the function α(x,y)\alpha(x,y) is a solution to the Codazzi-like equation (4.1) for any given yy. By Theorem 1.2, the function α(x,y)\alpha(x,y) hence has one of the following forms of special types

0,1x+c1(y),12x+c1(y),0,\ \frac{1}{x+c_{1}(y)},\ \frac{1}{2x+c_{1}(y)},

and general types

x+c1(y)(x+c1(y))2+c2(y),\frac{x+c_{1}(y)}{(x+c_{1}(y))^{2}+c_{2}(y)},

where, instead of constants, both c1(y)c_{1}(y) and c2(y)c_{2}(y) are now functions of yy. Notice that c2(y)0c_{2}(y)\neq 0 for all yy. We now use the types of the function α(x,y)\alpha(x,y) to define the types of pp-minimal surface as follows.

Definition 5.1.

Locally, we say that a pp-minimal surface is

  1. (1)

    vertical if α\alpha vanishes (i.e., α(x,y)=0\alpha(x,y)=0 for all x,yx,y).

  2. (2)

    of special type I if α=1x+c1(y)\alpha=\frac{1}{x+c_{1}(y)}.

  3. (3)

    of special type II if α=12x+c1(y)\alpha=\frac{1}{2x+c_{1}(y)}.

  4. (4)

    of general type if α=x+c1(y)(x+c1(y))2+c2(y)\alpha=\frac{x+c_{1}(y)}{(x+c_{1}(y))^{2}+c_{2}(y)} with c2(y)0c_{2}(y)\neq 0 for all yy.

We further divide pp-minimal surfaces of general type into three classes as follows:

Definition 5.2.

We say that a pp-minimal surface of general type is

  1. (1)

    of type I if c2(y)>0c_{2}(y)>0 for all yy.

  2. (2)

    of type II if c2(y)<0c_{2}(y)<0 for all yy, and either x<c1(y)c2(y)x<-c_{1}(y)-\sqrt{-c_{2}(y)} or x>c1(y)+c2(y)x>-c_{1}(y)+\sqrt{-c_{2}(y)}.

  3. (3)

    of type III if c2(y)<0c_{2}(y)<0 for all yy, and c1(y)c2(y)<x<c1(y)+c2(y)-c_{1}(y)-\sqrt{-c_{2}(y)}<x<-c_{1}(y)+\sqrt{-c_{2}(y)}.

We notice that the type is invariant under an action of a Heisenberg rigid motion and the regular part of a pp-minimal surface ΣH1\Sigma\subset H_{1} is a union of these types of surfaces. The corresponding paths of each type of α\alpha are marked on the phase plane (Figure 4.2). We express some basic facts about pp-minimal surfaces with type as follows.

  • If α\alpha vanishes, then it is part of a vertical plane.

  • The two concave downward parabolas represent α=1x+c1,12x+c1\alpha=\frac{1}{x+c_{1}},\frac{1}{2x+c_{1}} respectively. The one for α=1x+c1\alpha=\frac{1}{x+c_{1}} is on top of the one for α=12x+c1\alpha=\frac{1}{2x+c_{1}}. For surfaces of special type I, we have that

    (5.1) α{,ifxc1from the right,,ifxc1from the left;\begin{split}\alpha\rightarrow\left\{\begin{array}[]{rl}\infty,&\ \textrm{if}\ x\rightarrow-c_{1}\ \ \textrm{from the right},\\ -\infty,&\ \textrm{if}\ x\rightarrow-c_{1}\ \ \textrm{from the left};\end{array}\right.\end{split}

    and, for surfaces of special type II, we have that

    (5.2) α{,ifxc12from the right,,ifxc12from the left;\begin{split}\alpha\rightarrow\left\{\begin{array}[]{rl}\infty,&\ \textrm{if}\ x\rightarrow\frac{-c_{1}}{2}\ \ \textrm{from the right},\\ -\infty,&\ \textrm{if}\ x\rightarrow\frac{-c_{1}}{2}\ \ \textrm{from the left};\end{array}\right.\end{split}
  • The closed curves with the origin removed correspond to the family of solutions

    α(x)=x+c1(x+c1)2+c2,\alpha(x)=\frac{x+c_{1}}{(x+c_{1})^{2}+c_{2}},

    where c1,c2c_{1},c_{2} are constants and c2>0c_{2}>0, which are of type I. There exists a zero for α\alpha-function at x=c1x=-c_{1}. For surfaces of type I, we have |α|12c2|\alpha|\leq\frac{1}{2\sqrt{c_{2}}}, so α\alpha is a bounded function for each fixed yy, that is, along each path on the phase plane, α\alpha is bounded. Therefore, there are no singular points for surfaces of type I.

  • The curves in between the two concave downward parabolas are of type II. The α\alpha-function of type II does not have any zeros. For surfaces of type II, it can be checked that

    (5.3) α{,ifxc1+c2from the right,,ifxc1c2from the left.\begin{split}\alpha\rightarrow\left\{\begin{array}[]{rl}\infty,&\ \textrm{if}\ x\rightarrow-c_{1}+\sqrt{-c_{2}}\ \ \textrm{from the right},\\ -\infty,&\ \textrm{if}\ x\rightarrow-c_{1}-\sqrt{-c_{2}}\ \ \textrm{from the left}.\end{array}\right.\end{split}
  • The curves beneath the lower concave downward parabolas are of type III. There exists a zero for α\alpha-function at x=c1x=-c_{1}. For surfaces of type III, we have

    (5.4) α{,ifxc1+c2from the left,,ifxc1c2from the right.\begin{split}\alpha\rightarrow\left\{\begin{array}[]{rl}-\infty,&\ \textrm{if}\ x\rightarrow-c_{1}+\sqrt{-c_{2}}\ \ \textrm{from the left},\\ \infty,&\ \textrm{if}\ x\rightarrow-c_{1}-\sqrt{-c_{2}}\ \ \textrm{from the right}.\end{array}\right.\end{split}
Proposition 5.3.

Suppose α(x,y)=x+c1(y)(x+c1(y))2+c2(y)\alpha(x,y)=\frac{x+c_{1}(y)}{(x+c_{1}(y))^{2}+c_{2}(y)}, which is of general type. Then the explicit formula for the induced metric on a pp-minimal surface with this α\alpha as its α\alpha-function is given by

(5.5) a=|α|h(y)|x+c1(y)|1+α2,b=|α|ek(y)|x+c1(y)|1+α2,a=\frac{|\alpha|h(y)}{|x+c_{1}(y)|\sqrt{1+\alpha^{2}}},\ \ b=\frac{|\alpha|e^{k(y)}}{|x+c_{1}(y)|\sqrt{1+\alpha^{2}}},

for some functions h(y)h(y) and k(y)k(y).

Proof.

If α=x+c1(y)(x+c1(y))2+c2(y)\alpha=\frac{x+c_{1}(y)}{(x+c_{1}(y))^{2}+c_{2}(y)}, we choose ln|(x+c1(y))2+c2(y)|\ln{|(x+c_{1}(y))^{2}+c_{2}(y)|} as an anti-derivative of 2α2\alpha with respect to xx. Simple computations imply

e2α𝑑x(1+α2)12=1(x+c1(y))2+((x+c1(y))2+c2(y))2=|α||x+c1(y)|1+α2.\begin{split}\frac{e^{-\int 2\alpha dx}}{(1+\alpha^{2})^{\frac{1}{2}}}&=\frac{1}{\sqrt{(x+c_{1}(y))^{2}+((x+c_{1}(y))^{2}+c_{2}(y))^{2}}}\\ &=\frac{|\alpha|}{|x+c_{1}(y)|\sqrt{1+\alpha^{2}}}.\end{split}

Substituting the above formula into (2.15) and (2.18), equation (5.5) follows. ∎

Similarly, we have

Proposition 5.4.

Suppose α(x,y)=1x+c1(y)\alpha(x,y)=\frac{1}{x+c_{1}(y)}, which is of special type I. Then the explicit formula for the induced metric on a pp-minimal surface with this α\alpha as its α\alpha-function is given by

(5.6) a=α2h(y)1+α2,b=α2ek(y)1+α2.a=\frac{\alpha^{2}h(y)}{\sqrt{1+\alpha^{2}}},\ \ b=\frac{\alpha^{2}e^{k(y)}}{\sqrt{1+\alpha^{2}}}.
Proof.

To obtain (5.6), we choose 2ln|x+c1(y)|2\ln{|x+c_{1}(y)|} as an anti-derivative of 2α2\alpha with respect to xx. ∎

Proposition 5.5.

Suppose α(x,y)=12x+c1(y)\alpha(x,y)=\frac{1}{2x+c_{1}(y)}, which is of special type II. Then the explicit formula for the induced metric on a pp-minimal surface with this α\alpha as its α\alpha-function is given by

(5.7) a=|α|h(y)1+α2,b=|α|ek(y)1+α2.a=\frac{|\alpha|h(y)}{\sqrt{1+\alpha^{2}}},\ \ b=\frac{|\alpha|e^{k(y)}}{\sqrt{1+\alpha^{2}}}.
Proof.

To have (5.7), we choose ln|2x+c1(y)|\ln{|2x+c_{1}(y)|} as an anti-derivative of 2α2\alpha with respect to xx. ∎

5.2. Analogs for constant pp-mean curvature surfaces

If H=c0H=c\neq 0 (we assume that c=2λ>0c=2\lambda>0), then the Codazzi-like equation (1.2) has the trivial solution α(x)=0\alpha(x)=0 and the general solution of the form

(5.8) α(x)=λsin(2λx+c1)c2cos(2λx+c1),\alpha(x)=\lambda\frac{\sin{(2\lambda x+c_{1}})}{c_{2}-\cos{(2\lambda x+c_{1})}},

which depends on two constants c1c_{1} and c2c_{2}.

Proposition 5.6.

For any α(x,y)=λsin(2λx+c1(y))c2(y)cos(2λx+c1(y))\alpha(x,y)=\lambda\frac{\sin{(2\lambda x+c_{1}(y)})}{c_{2}(y)-\cos{(2\lambda x+c_{1}(y))}}, the explicit formula for the induced metric on a constant pp-mean curvature surface with 2λ2\lambda as its pp-mean curvature and this α\alpha as its α\alpha-function is given by

(5.9) a=c2(1+α2)1/2+c2c2(c2cos(cx+c1))(1+α2)1/2+h(y)|c2cos(cx+c1)|(1+α2)1/2,a=-\frac{\frac{c}{2}}{(1+\alpha^{2})^{1/2}}+\frac{\frac{c}{2}c_{2}}{(c_{2}-\cos{(cx+c_{1})})(1+\alpha^{2})^{1/2}}+\frac{h(y)}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}},

and

(5.10) b=ek(y)|c2cos(cx+c1)|(1+α2)1/2,b=\frac{e^{k(y)}}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}},

for some functions h(y)h(y) and k(y)k(y).

Proof.

Let f(x)=ln|c2cos(cx+c1|,c=2λf(x)=\ln{|c_{2}-\cos{(cx+c_{1}}|},c=2\lambda. Then f(x)=2α(x)f^{\prime}(x)=2\alpha(x). If we choose 2α𝑑x=f(x)\int 2\alpha dx=f(x), it is easy to see

e2α𝑑x=|c2cos(cx+c1)|,e^{\int 2\alpha dx}=|c_{2}-\cos{(cx+c_{1})|},

and hence,

a=e2α𝑑x(1+α2)1/2(h(y)cα(e2α𝑑x)𝑑x)=1|c2cos(cx+c1)|(1+α2)1/2(h(y)cα|c2cos(cx+c1)|𝑑x),\begin{split}a&=\frac{e^{-\int 2\alpha dx}}{(1+\alpha^{2})^{1/2}}\left(h(y)-\int c\alpha(e^{\int 2\alpha dx})dx\right)\\ &=\frac{1}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}}\left(h(y)-\int c\alpha|c_{2}-\cos{(cx+c_{1})}|dx\right),\end{split}

where

cα|c2cos(cx+c1)|𝑑x=c22sin(cx+c1)c2cos(cx+c1)|c2cos(cx+c1)|𝑑x={c2cos(cx+c1),c2cos(cx+c1)>0c2cos(cx+c1),c2cos(cx+c1)<0.\begin{split}\int c\alpha|c_{2}-\cos{(cx+c_{1})}|dx&=\int\frac{c^{2}}{2}\frac{\sin{(cx+c_{1}})}{c_{2}-\cos{(cx+c_{1})}}|c_{2}-\cos{(cx+c_{1})}|dx\\ &=\left\{\begin{array}[]{rl}-\frac{c}{2}\cos{(cx+c_{1})},&c_{2}-\cos{(cx+c_{1})}>0\\ \frac{c}{2}\cos{(cx+c_{1})},&c_{2}-\cos{(cx+c_{1})}<0\end{array}\right..\end{split}

Direct computations imply

(5.11) a=h(y)|c2cos(cx+c1)|(1+α2)1/2+c2(c2cos(cx+c1))+c2c2(c2cos(cx+c1))(1+α2)1/2=c2(1+α2)1/2+c2c2(c2cos(cx+c1))(1+α2)1/2+h(y)|c2cos(cx+c1)|(1+α2)1/2\begin{split}a&=\frac{h(y)}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}}+\frac{-\frac{c}{2}(c_{2}-\cos{(cx+c_{1})})+\frac{c}{2}c_{2}}{(c_{2}-\cos{(cx+c_{1})})(1+\alpha^{2})^{1/2}}\\ &=-\frac{\frac{c}{2}}{(1+\alpha^{2})^{1/2}}+\frac{\frac{c}{2}c_{2}}{(c_{2}-\cos{(cx+c_{1})})(1+\alpha^{2})^{1/2}}+\frac{h(y)}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}}\end{split}

and

(5.12) b=e2α𝑑x(1+α2)1/2ek(y)=ek(y)|c2cos(cx+c1)|(1+α2)1/2.b=\frac{e^{-\int 2\alpha dx}}{(1+\alpha^{2})^{1/2}}e^{k(y)}=\frac{e^{k(y)}}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}}.

Example 5.7 (Pansu Sphere).

The Pansu sphere in [5] is the union of graphs of ff and f-f given by

f(z)=12λ2(λ|z|1λ2|z|2+cos1(λ|z|)),|z|1λ,f(z)=\frac{1}{2\lambda^{2}}\left(\lambda|z|\sqrt{1-\lambda^{2}|z|^{2}}+\cos^{-1}(\lambda|z|)\right),|z|\leq\frac{1}{\lambda},

which has a constant pp-mean curvature c=2λc=2\lambda. Its α\alpha-function is of the following form,

α=λsin(2λs)(1cos(2λs))=λtan(λs+π2),\alpha=\frac{\lambda\sin(2\lambda s)}{(1-\cos(2\lambda s))}=-\lambda\tan(\lambda s+\frac{\pi}{2}),

and the metric is given by

(5.13) a=λ1+α2 and b=2λ21+α2(1cos2λs).a=\frac{-\lambda}{\sqrt{1+\alpha^{2}}}\mbox{ and }b=\frac{2\lambda^{2}}{\sqrt{1+\alpha^{2}}(1-\cos 2\lambda s)}.

6. A representation of constant pp-mean curvature surfaces

Let ΣH1\Sigma\subset H_{1} be a pp-minimal surface. We define an orthogonal coordinate system (x,y)(x,y) to be a compatible coordinate system such that a=0a=0, that is, e^2=by\hat{e}_{2}=b\frac{\partial}{\partial y}.

Proposition 6.1.

There always exists an orthogonal coordinate system around any regular point of a pp-minimal surface Σ\Sigma.

Proof.

Suppose that pΣp\in\Sigma is a regular point and (x,y)(x,y) is an arbitrary compatible coordinate system around pp. Since H=0H=0, equations (2.15) and (2.18) imply that the ratio ab=h(y)eg(y)\frac{-a}{b}=\frac{-h(y)}{e^{g(y)}} is just a function of yy. Now we define another compatible coordinates (x~,y~)(\tilde{x},\tilde{y}) by

(x~,y~)=(x+Γ(y),Ψ(y)),(\tilde{x},\tilde{y})=(x+\Gamma(y),\Psi(y)),

for some functions Γ(y)\Gamma(y) and Ψ(y)\Psi(y) such that Γ(y)=ab\Gamma^{\prime}(y)=\frac{-a}{b}. By the transformation law (2.20) of the representation of the induced metric, we have a~=0\tilde{a}=0. This means that (x~,y~)(\tilde{x},\tilde{y}) are orthogonal coordinates around pp. ∎

6.1. The proof of Theorem 1.3

It will be suitable to choose an orthogonal coordinate system (U;x,y)(U;x,y) to study a pp-minimal surface. One sees easily from (2.19) and (2.20) that any two orthogonal coordinate systems (x,y)(x,y) and (x~,y~)(\tilde{x},\tilde{y}) are transformed by

(6.1) x~=x+C,y~=Ψ(y),\tilde{x}=x+C,\ \ \ \tilde{y}=\Psi(y),

for a constant CC and a function Ψ(y)\Psi(y). That is, the orthogonal coordinate systems are determined, up to a constant CC on the coordinate xx and scaling on the coordinate yy. The transformation law of the representation of the induced metric hence reduces to

(6.2) a~=a=0,b~=bΨ(y).\tilde{a}=a=0,\ \tilde{b}=b\Psi^{\prime}(y).

In terms of orthogonal coordinate systems, the integrability condition hence reads

(6.3) bxb=2α+ααx1+α2,αxx+6ααx+4α3=0.\begin{split}-\frac{b_{x}}{b}&=2\alpha+\frac{\alpha\alpha_{x}}{1+\alpha^{2}},\\ \alpha_{xx}&+6\alpha\alpha_{x}+4\alpha^{3}=0.\end{split}

Then the α\alpha-function determines the metric representation bb, and hence a pp-minimal surface, up to a positive function ek(y)e^{k(y)} as (2.15) specified (or see (5.5),(5.6) and (5.7) ). Therefore, from the transformation law of the induced metric (6.2), we are able to choose another orthogonal coordinate system (x~,y~)=ϕ(x,y)(\tilde{x},\tilde{y})=\phi(x,y) with Ψ\Psi satisfying ek(y)Ψ=1e^{k(y)}\Psi^{{}^{\prime}}=1. That is, we can further normalize bb such that k(y~)=0k(\tilde{y})=0 for each type, no matter it is special or general. Here k(y~)k(\tilde{y}) is the function kk in the numerator of b~\tilde{b} (see (5.5),(5.6) and (5.7)) with α,x,y\alpha,x,y replaced by α~,x~,y~\tilde{\alpha},\tilde{x},\tilde{y}. In fact, for a general type (the special types are similar), it is possible to choose another orthogonal coordinate system (x~,y~)=ϕ(x,y)(\tilde{x},\tilde{y})=\phi(x,y) such that

ϕb~=|α||x+c1(y)|1+α2.\phi^{*}\tilde{b}=\frac{|\alpha|}{|x+c_{1}(y)|\sqrt{1+\alpha^{2}}}.

One sees that such orthogonal coordinate systems are unique up to a translation on the two variables xx and yy. In other words, there are constants C1C_{1} and C2C_{2} such that

(6.4) (x~,y~)=ϕ(x,y)=(x+C1,y+C2).(\tilde{x},\tilde{y})=\phi(x,y)=(x+C_{1},y+C_{2}).

We call such an orthogonal coordinate system a normal coordinate system. Indeed, since ϕα~=α\phi^{*}\tilde{\alpha}=\alpha, Definition 5.1 indicates the following transformation law for c1(y)c_{1}(y) and c2(y)c_{2}(y) functions

(6.5) c~1(y~)=c1(y~C2)C1, for special type I,c~1(y~)=c1(y~C2)2C1, for special type II, andc~1(y~)=c1(y~C2)C1,c~2(y~)=c2(y~C2), for general type,\begin{split}\tilde{c}_{1}(\tilde{y})&=c_{1}(\tilde{y}-C_{2})-C_{1},\textrm{ for special type I},\\ \tilde{c}_{1}(\tilde{y})&=c_{1}(\tilde{y}-C_{2})-2C_{1},\textrm{ for special type II, and}\\ \tilde{c}_{1}(\tilde{y})&=c_{1}(\tilde{y}-C_{2})-C_{1},\ \tilde{c}_{2}(\tilde{y})=c_{2}(\tilde{y}-C_{2}),\textrm{ for general type},\\ \end{split}

where c~1(y~)\tilde{c}_{1}(\tilde{y}) and c~2(y~)\tilde{c}_{2}(\tilde{y}) are with respect to α~\tilde{\alpha}. Namely, c2(y)c_{2}(y) is unique up to a translation on yy, and c1(y)c_{1}(y) is unique up to a translation on yy and its image as well. We denote these two unique functions c1(y)c_{1}(y) and c2(y)c_{2}(y) by ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y), respectively. We then complete the proof of Theorem 1.3.

Both two functions ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y) are invariants under a Heisenberg rigid motion. Therefore we call them ζ1\zeta_{1}- and ζ2\zeta_{2}-invariants, respectively. In terms of ζ1\zeta_{1} and ζ2\zeta_{2}, we thus have the version of the fundamental theorem for pp-minimal surfaces in H1H_{1} (Theorem 1.5).

6.2. The proof of Theorem 1.5

Given ζ1(y)\zeta_{1}(y), for (1) in Theorem 1.5, we define α,a,b\alpha,a,b on UU by

α=1x+ζ1(y),a=0andb=α21+α2.\alpha=\frac{1}{x+\zeta_{1}(y)},a=0\ \textrm{and}\ b=\frac{\alpha^{2}}{\sqrt{1+\alpha^{2}}}.

Notice that (e,f)(e,f) needs to be chosen so that (e,f)×(c,d)(e,f)\times(c,d) does not contain the zero set of x+ζ1(y)x+\zeta_{1}(y). Then they satisfy the integrability condition (3.1) with c=0c=0, and hence UU together with α,a,b\alpha,a,b can be embedded into H1H_{1} to be a pp-minimal surface with α\alpha as its α\alpha-function, and the induced metric a,ba,b. Moreover the characteristic direction e1=xe_{1}=\frac{\partial}{\partial x}. From the type of α\alpha, this minimal pp-surface is of special type I. In view of a=0a=0 and b=α21+α2b=\frac{\alpha^{2}}{\sqrt{1+\alpha^{2}}}, we see that the coordinates (x,y)(x,y) are a normal coordinate system. Therefore ζ1(y)\zeta_{1}(y) is the corresponding ζ1\zeta_{1}-invariant. The uniqueness follows from the fundamental theorem for surfaces in H1H_{1} or theorem 1.3. This completes the proof of (1) for the special type I. Both proofs of (1) for the special type II and of (3) are similar with (e,f)(e,f) chosen according to their types. For the special type II, note that (e,f)(e,f) needs be chosen so that (e,f)×(c,d)(e,f)\times(c,d) does not contain the zero set of 2x+ζ1(y)2x+\zeta_{1}(y), and we define α,a,b\alpha,a,b on UU by

α=12x+ζ1(y),a=0andb=|α|1+α2.\alpha=\frac{1}{2x+\zeta_{1}(y)},a=0\ \textrm{and}\ b=\frac{|\alpha|}{\sqrt{1+\alpha^{2}}}.

For (3), we see that (e,f)(e,f) needs be chosen so that (e,f)×(c,d)(e,f)\times(c,d) contains no the zero set of (x+ζ1(y))2+ζ2(y)(x+\zeta_{1}(y))^{2}+\zeta_{2}(y), and we define α,a,b\alpha,a,b on UU by

α=x+ζ1(y)(x+ζ1(y))2+ζ2(y),a=0andb=|α||x+ζ1(y)|1+α2.\alpha=\frac{x+\zeta_{1}(y)}{(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)},a=0\ \textrm{and}\ b=\frac{|\alpha|}{|x+\zeta_{1}(y)|\sqrt{1+\alpha^{2}}}.

Thus we complete the proof of Theorem 1.5.

We remark that, in terms of normal coordinates (x,y)(x,y), the co-frame formula (2.3) reads

(6.6) ω^1=dxabdy=dx,ω^2=1bdy,\begin{split}\hat{\omega}^{1}&=dx-\frac{a}{b}dy=dx,\\ \hat{\omega}^{2}&=\frac{1}{b}dy,\end{split}

and hence the induced metric II (the first fundamental form) reads

(6.7) I=ω^1ω^1+ω^2ω^2=dxdx+1b2dydy,={dxdx+[(x+ζ1(y))2+(x+ζ1(y))4]dydy,for special type I,dxdx+[1+(2x+ζ1(y))2]dydy,for special type II,dxdx+[(x+ζ1(y))2+[(x+ζ1(y))2+ζ2(y)]2]dydy,for general type.\begin{split}I&=\hat{\omega}^{1}\otimes\hat{\omega}^{1}+\hat{\omega}^{2}\otimes\hat{\omega}^{2}=dx\otimes dx+\frac{1}{b^{2}}dy\otimes dy,\\ &=\left\{\begin{array}[]{ll}dx\otimes dx+\big{[}(x+\zeta_{1}(y))^{2}+(x+\zeta_{1}(y))^{4}\big{]}dy\otimes dy,&\textrm{for {\bf special type I}},\\ &\\ dx\otimes dx+\big{[}1+(2x+\zeta_{1}(y))^{2}\big{]}dy\otimes dy,&\textrm{for {\bf special type II}},\\ &\\ dx\otimes dx+\big{[}(x+\zeta_{1}(y))^{2}+[(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)]^{2}\big{]}dy\otimes dy,&\textrm{for {\bf general type}}.\\ \end{array}\right.\end{split}

From (6.7), we see immediately that the induced metric II degenerates on the singular set {(x,y)|x+ζ1(y)=0}\{(x,y)\ |\ x+\zeta_{1}(y)=0\} for surfaces of special type I. Therefore, it cannot extend smoothly through the singular set. On the other hand, this phenomenon does not happen for both special type II and general type.

6.3. The maximal pp-minimal surfaces and the proof of Theorem 1.6

From the proof of Theorem 1.5, it is clear to see that

  • for the special type I, the open rectangle UU in (1) of Theorem 1.5 can be extended to be either

    (6.8) UI={(x,y)2|y(c,d),x+ζ1(y)<0},orUI+={(x,y)2|y(c,d),x+ζ1(y>0},\begin{split}U_{I}^{-}&=\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),\ x+\zeta_{1}(y)<0\},\ \textrm{or}\\ U_{I}^{+}&=\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),\ x+\zeta_{1}(y>0\},\end{split}

    which depends on that UU is originally contained in UIU_{I}^{-} or UI+U_{I}^{+}. Notice that the embedding XX might be just extended to be an immersion. Since both UIU_{I}^{-} and UI+U_{I}^{+} are connected and simply connected, the immersion XX is unique, up to a Heisenberg rigid motion. We denote these two pp-minimal surfaces of special type I by SI(ζ1)=X(UI)S_{I}^{-}(\zeta_{1})=X(U_{I}^{-}) and SI+(ζ1)=X(UI+)S_{I}^{+}(\zeta_{1})=X(U_{I}^{+}). From (6.7), we see that the induced metric II degenerates on the singular set {(x,y)2|y(c,d),x+ζ1(y)=0}\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),\ x+\zeta_{1}(y)=0\}.

  • for the special type II, the open rectangle UU in (2) of Theorem 1.5 can be extended to be either

    (6.9) UII={(x,y)2|y(c,d), 2x+ζ1(y)<0},orUII+={(x,y)2|y(c,d), 2x+ζ1(y>0},\begin{split}U_{II}^{-}&=\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),\ 2x+\zeta_{1}(y)<0\},\ \textrm{or}\\ U_{II}^{+}&=\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),\ 2x+\zeta_{1}(y>0\},\end{split}

    which depends on the fact that UU is originally contained in UIIU_{II}^{-} or UII+U_{II}^{+}. The embedding XX might be extended to be an immersion. Since both UIIU_{II}^{-} and UII+U_{II}^{+} are connected and simply connected, the immersion XX is unique, up to a Heisenberg rigid motion. We denote these two pp-minimal surfaces of special type II by SII(ζ1)=X(UII)S_{II}^{-}(\zeta_{1})=X(U_{II}^{-}) and SII+(ζ1)=X(UII+)S_{II}^{+}(\zeta_{1})=X(U_{II}^{+}).

  • when ζ2(y)>0\zeta_{2}(y)>0 for all y(c,d)y\in(c,d), since there exist no zeros of (x+ζ1(y))2+ζ2(y)=0(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)=0, the open rectangle UU in (3) can be extended to be the product space

    VI=×(c,d).V_{I}=\operatorname{\mathbb{R}}\times(c,d).

    Since the extended immersion XX is unique, up to a Heisenberg rigid motion, we denote the pp-minimal surface of type I by ΣI(ζ1,ζ2)=X(VI)\Sigma_{I}(\zeta_{1},\zeta_{2})=X(V_{I}).

  • when ζ2(y)<0\zeta_{2}(y)<0 for all y(c,d)y\in(c,d), since the zero set (x+ζ1(y))2+ζ2(y)=0(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)=0 consists of two separated curves defined by x+ζ1(y)+ζ2(y)=0x+\zeta_{1}(y)+\sqrt{-\zeta_{2}(y)}=0 and x+ζ1(y)ζ2(y)=0x+\zeta_{1}(y)-\sqrt{-\zeta_{2}(y)}=0, respectively, the open rectangle UU in (3) can be extended to be one of the following three domains:

    (6.10) VII={(x,y)2|y(c,d),x<ζ1(y)ζ2(y)},VII+={(x,y)2|y(c,d),x>ζ1(y)+ζ2(y)},andVIII={(x,y)2|y(c,d),ζ1(y)ζ2(y)<x<ζ1(y)+ζ2(y)},\begin{split}V_{II}^{-}&=\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),x<-\zeta_{1}(y)-\sqrt{-\zeta_{2}(y)}\},\\ V_{II}^{+}&=\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),x>-\zeta_{1}(y)+\sqrt{-\zeta_{2}(y)}\},\ \ \textrm{and}\\ V_{III}&=\{(x,y)\in\operatorname{\mathbb{R}}^{2}\ |\ y\in(c,d),-\zeta_{1}(y)-\sqrt{-\zeta_{2}(y)}<x<-\zeta_{1}(y)+\sqrt{-\zeta_{2}(y)}\},\end{split}

    Since the extended immersion XX is unique, up to a Heisenberg rigid motion, we denote these two pp-minimal surfaces of type II by ΣII(ζ1,ζ2)=X(VII)\Sigma_{II}^{-}(\zeta_{1},\zeta_{2})=X(V_{II}^{-}) and ΣII+(ζ1,ζ2)=X(VII+)\Sigma_{II}^{+}(\zeta_{1},\zeta_{2})=X(V_{II}^{+}), and the pp-minimal surface of type III by ΣIII(ζ1,ζ2)=X(VIII)\Sigma_{III}(\zeta_{1},\zeta_{2})=X(V_{III}).

We see that ζ1\zeta_{1}-invariant is the only invariant for pp-minimal surfaces of special types, and ζ1\zeta_{1}- and ζ2\zeta_{2}-invariants are the only two invariants for general type. From the above, Theorem 1.6 immediately follows.

6.4. Symmetric pp-minimal surfaces

A pp-minimal surface is called symmetric if ζ1\zeta_{1}-invariant is a constant for the special types; both ζ1\zeta_{1}- and ζ2\zeta_{2}-invariants are constants for the general types. Since ζ1\zeta_{1}, up to a translation on its image, is an invariant, we presently have

Theorem 6.2.

All symmetric pp-minimal surfaces of the same special type are locally congruent to one another, whereas for the general type, locally there is a family of symmetric pp-minimal surfaces, depending on a parameter on \mathbb{R}.

6.5. The normalization of constant pp-mean curvature surfaces

Unlike pp-minimal surfaces, Proposition 5.6 indicates that normalizing aa to be zero is not applicable. Therefore, we would like to normalize aa and bb so that they look like the induced metric of the Pansu sphere given in (5.13) as possible as we can. Let (x~,y~)=(x+Γ(y),Ψ(y))(\tilde{x},\tilde{y})=(x+\Gamma(y),\Psi(y)) be another compatible coordinates. The transformation laws imply

(6.11) a~=a+bΓ(y)=c2(1+α2)1/2+±(λc2)+h(y)+ek(y)Γ(y)|c2cos(cx+c1)|(1+α2)1/2,\begin{split}\tilde{a}&=a+b\Gamma^{\prime}(y)\\ &=-\frac{\frac{c}{2}}{(1+\alpha^{2})^{1/2}}+\frac{\pm(\lambda c_{2})+h(y)+e^{k(y)}\Gamma^{\prime}(y)}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}},\end{split}

where we have the sign ”-” if c2cos(cx+c1)<0c_{2}-\cos{(cx+c_{1})}<0. Therefore, if we can choose Γ(y)\Gamma(y) such that ±(λc2)+h(y)+ek(y)Γ(y)=0\pm(\lambda c_{2})+h(y)+e^{k(y)}\Gamma^{\prime}(y)=0, we have

(6.12) a~=λ(1+α2)1/2.\tilde{a}=-\frac{\lambda}{(1+\alpha^{2})^{1/2}}.

Similarly,

(6.13) b~=bΨ(y)=ek(y)Ψ(y)|c2cos(cx+c1)|(1+α2)1/2,\tilde{b}=b\Psi^{\prime}(y)=\frac{e^{k(y)}\Psi^{\prime}(y)}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}},

which implies that if Ψ(y)\Psi(y) is chosen to be ek(y)Ψ(y)=2λ2e^{k(y)}\Psi^{\prime}(y)=2\lambda^{2}, we obtain

(6.14) b~=2λ2|c2cos(cx+c1)|(1+α2)1/2,\tilde{b}=\frac{2\lambda^{2}}{|c_{2}-\cos{(cx+c_{1})}|(1+\alpha^{2})^{1/2}},

It is easy to see that such a coordinate system (x~,y~)(\tilde{x},\tilde{y}) is unique up to a translation, i.e., (x~,y~)=(x+C1,y+C2)(\tilde{x},\tilde{y})=(x+C_{1},y+C_{2}) for some constants C1,C2C_{1},C_{2}. We call it the normal coordinate system.

Theorem 6.3.

In normal coordinates (x,y)(x,y), the functions c1(y)c_{1}(y) and c2(y)c_{2}(y) in the expression α(x)=λsin(2λx+c1)c2cos(2λx+c1)\alpha(x)=\lambda\frac{\sin{(2\lambda x+c_{1}})}{c_{2}-\cos{(2\lambda x+c_{1})}} are unique in the following sense: up to a translation on yy, c2(y)c_{2}(y) is unique up to a sign; and c1(y)c_{1}(y) is unique up to a constant. We denote these two unique functions by

ζ1(y)=c1(y),ζ2(y)=c2(y).\zeta_{1}(y)=c_{1}(y),\ \zeta_{2}(y)=c_{2}(y).

Therefore, {ζ1(y),ζ2(y)}\{\zeta_{1}(y),\zeta_{2}(y)\} is a complete set of invariants for those surfaces (α\alpha not vanishing).

7. Examples of pp-minimal surfaces

7.1. Examples of special type I

The following is a family of pp-minimal surfaces. They are defined by the graphs of

(7.1) u=Ax+By+C,u=Ax+By+C,

for some real constants A,BA,B and CC. It is easy to see that (B,A,C)(-B,A,C) or (x,y)=(B,A)(x,y)=(-B,A) is the only singular point of the graph of u=Ax+By+Cu=Ax+By+C.

Lemma 7.1.

The graph defined by (7.1) is congruent to the graph of u=0u=0.

Proof.

After the action of the left translation by (B,A,C)(B,-A,-C), we have

(B,A,C)(x,y,u)=(x+B,yA,uCAxBy)=(x+B,yA,0).\begin{split}(B,-A,-C)(x,y,u)&=(x+B,y-A,u-C-Ax-By)\\ &=(x+B,y-A,0).\end{split}

This completes the proof. ∎

Example 7.2.

The p-minimal surface defined by the graph of u=0u=0 corresponds to α=1r\alpha=\frac{1}{r}, where r=x2+y2r=\sqrt{x^{2}+y^{2}}. Indeed, let us consider a surface defined by

X:(x,y)(x,y,0).X:(x,y)\rightarrow(x,y,0).

The horizontal normal can be calculated as

(7.2) e2=(uxy)re1+(uy+x)re2=yx2+y2e1+xx2+y2e2,e_{2}=\frac{(u_{x}-y)}{r}\overset{\circ}{e_{1}}+\frac{(u_{y}+x)}{r}\overset{\circ}{e_{2}}=\frac{-y}{\sqrt{x^{2}+y^{2}}}\overset{\circ}{e_{1}}+\frac{x}{\sqrt{x^{2}+y^{2}}}\overset{\circ}{e_{2}},

and then

(7.3) e1=xx2+y2e1+yx2+y2e2=xx2+y2x+yx2+y2y.e_{1}=\frac{x}{\sqrt{x^{2}+y^{2}}}\overset{\circ}{e_{1}}+\frac{y}{\sqrt{x^{2}+y^{2}}}\overset{\circ}{e_{2}}=\frac{x}{\sqrt{x^{2}+y^{2}}}\frac{\partial}{\partial x}+\frac{y}{\sqrt{x^{2}+y^{2}}}\frac{\partial}{\partial y}.

For the α\alpha-function, making use of

αe2+T=α(yx2+y2,xx2+y2,(x2+y2)x2+y2)+(0,0,1).\alpha e_{2}+T=\alpha(\frac{-y}{\sqrt{x^{2}+y^{2}}},\frac{x}{\sqrt{x^{2}+y^{2}}},\frac{-(x^{2}+y^{2})}{\sqrt{x^{2}+y^{2}}})+(0,0,1).

to derive α=1x2+y2\alpha=\frac{1}{\sqrt{x^{2}+y^{2}}}, and hence (0,0)(0,0) is the only singular point. Notice that (x,y)(x,y) is not a compatible coordinate system. In terms of the polar coordinates (r,θ)(r,\theta) with the coordinates transformation x=rcosθx=r\cos{\theta} and y=rsinθy=r\sin{\theta}, that is, we consider the re-parametrization

X:(r,θ)(rcosθ,rsinθ,0).X:(r,\theta)\rightarrow(r\cos{\theta},r\sin{\theta},0).

It represents

(7.4) Xr=(cosθ,sinθ,0),Xθ=(rsinθ,rcosθ,0).X_{r}=(\cos{\theta},\sin{\theta},0),\ X_{\theta}=(-r\sin{\theta},r\cos{\theta},0).

From (7.3), it is easy to see that e1=Xr=re_{1}=X_{r}=\frac{\partial}{\partial r}, and thus (r,θ)(r,\theta) is a compatible coordinate system. For the α\alpha-function and the induced metric aa and bb, we solve the equation

αe2+T1+α2=aXr+bXθ\frac{\alpha e_{2}+T}{\sqrt{1+\alpha^{2}}}=aX_{r}+bX_{\theta}

to get e2=(sinθ,cosθ,r)e_{2}=(-\sin{\theta},\cos{\theta},-r) from (7.2), and to obtain

α=1r,a=0,b=α21+α2,\alpha=\frac{1}{r},a=0,b=\frac{\alpha^{2}}{\sqrt{1+\alpha^{2}}},

which, from the formula of bb, implies that the polar coordinates (r,θ)(r,\theta) is a normal coordinate system. Since α=1r\alpha=\frac{1}{r}, the surface XX is of special type I with the ζ1(θ)=0\zeta_{1}(\theta)=0 as ζ1\zeta_{1}-invariant, and hence XX is a symmetric pp-minimal surface.

Refer to caption
Figure 7.1. The characteristic direction field of XX
Refer to caption
Figure 7.2. e1=xe_{1}=\frac{\partial}{\partial x} for x>12g(y)x>-\frac{1}{2}g^{\prime}(y)

Given Theorem 6.2, we immediately have the following theorem.

Theorem 7.3.

A symmetric pp-minimal surface of special type I is locally congruent to the graph of u=0u=0.

Proof.

This is because that ζ1\zeta_{1} is constant for a symmetric pp-minimal surface of special type I and the function ζ1\zeta_{1}, up to a constant, is a complete invariant. ∎

Theorem 7.4.

In terms of normal coordinates (x,y)(x,y), the induced metric (the first fundamental form) on a pp-minimal surface of special type I degenerates on the singular set {(x,y)|x=ζ1(y)}\{(x,y)\ |\ x=-\zeta_{1}(y)\}. Therefore if it is not symmetric, then it will never smoothly extend through the singular set.

Proof.

In terms of normal coordinates, a pp-minimal surface of special type I is given by a function ζ1(y)\zeta_{1}(y), in which we have

α=1x+ζ1(y),a=0,b=α21+α2.\alpha=\frac{1}{x+\zeta_{1}(y)},\ a=0,\ b=\frac{\alpha^{2}}{\sqrt{1+\alpha^{2}}}.

Therefore,

e^2=ax+by=α21+α2y.\hat{e}_{2}=a\frac{\partial}{\partial x}+b\frac{\partial}{\partial y}=\frac{\alpha^{2}}{\sqrt{1+\alpha^{2}}}\frac{\partial}{\partial y}.

This is equivalent to that the induced metric II is degenerate on the singular set on which α\alpha blows up. In fact, we have

I=ω^1ω^1+ω^2ω^2=dxdx+(1+α2α4)dydy,I=\hat{\omega}^{1}\otimes\hat{\omega}^{1}+\hat{\omega}^{2}\otimes\hat{\omega}^{2}=dx\otimes dx+\left(\frac{1+\alpha^{2}}{\alpha^{4}}\right)dy\otimes dy,

where (ω^1,ω^2)(\hat{\omega}^{1},\hat{\omega}^{2}) is the dual co-frame of (e^1,e^2)(\hat{e}_{1},\hat{e}_{2}). If it is not symmetric, and suppose that it can be smoothly extended beyond the singular set, then the singular set is actually a singular curve and the induced metric must be non-degenerate. This completes the proof. ∎

7.2. Examples of special type II

The other family of pp-minimal surfaces is defined by the graph of

(7.5) u=ABx2+(A2B2)xy+ABy2+g(Bx+Ay),u=-ABx^{2}+(A^{2}-B^{2})xy+ABy^{2}+g(-Bx+Ay),

for some real constants AA and BB such that A2+B2=1A^{2}+B^{2}=1 and gC()g\in C^{\infty}(\operatorname{\mathbb{R}}).

Lemma 7.5.

The graph defined by (7.5) is congruent to the graph of u=xy+g(y)u=xy+g(y).

Proof.

Since A2+B2=1A^{2}+B^{2}=1, the matrix

(ABBA)\left(\begin{array}[]{rl}A&B\\ -B&A\end{array}\right)

defines a rotation on 2\mathbb{R}^{2}. Let

(XY)=(ABBA)(xy),\left(\begin{array}[]{c}X\\ Y\end{array}\right)=\left(\begin{array}[]{rl}A&B\\ -B&A\end{array}\right)\left(\begin{array}[]{c}x\\ y\end{array}\right),

we have

XY=(Ax+By)(Bx+Ay)=ABx2+(A2B2)xy+ABy2,\begin{split}XY&=(Ax+By)(-Bx+Ay)\\ &=-ABx^{2}+(A^{2}-B^{2})xy+ABy^{2},\end{split}

which implies

u=XY+g(Y).u=XY+g(Y).

This completes the proof. ∎

Example 7.6.

We now study the example of the graph of u=xy+g(y)u=xy+g(y). We consider a parametrization of the graph defined by

X:(x,y)(x,y,xy+g(y)),X:(x,y)\rightarrow(x,y,xy+g(y)),

then we have

(7.6) Xx=(1,0,y)=e̊1,Xy=(0,1,x+g(y)).X_{x}=(1,0,y)=\mathring{e}_{1},\ \ X_{y}=(0,1,x+g^{\prime}(y)).

The horizontal normal e2e_{2} is taken to be

e2={(uxy)De̊1+(uy+x)De̊2(uxy)De̊1(uy+x)De̊2={2x+g(y)De̊2=e̊2,if 2x+g(y)>02x+g(y)De̊2=e̊2,if 2x+g(y)<0,e_{2}=\left\{\begin{array}[]{l}\frac{(u_{x}-y)}{D}\mathring{e}_{1}+\frac{(u_{y}+x)}{D}\mathring{e}_{2}\\ \\ -\frac{(u_{x}-y)}{D}\mathring{e}_{1}-\frac{(u_{y}+x)}{D}\mathring{e}_{2}\end{array}\right.\\ =\left\{\begin{array}[]{ll}\frac{2x+g^{\prime}(y)}{D}\mathring{e}_{2}=\mathring{e}_{2}&,\ \textrm{if}\ 2x+g^{\prime}(y)>0\\ \\ -\frac{2x+g^{\prime}(y)}{D}\mathring{e}_{2}=\mathring{e}_{2}&,\ \textrm{if}\ 2x+g^{\prime}(y)<0\end{array},\right.

where D=|2x+g(y)|D=|2x+g^{\prime}(y)|. Combining with (7.6), one sees

(7.7) e1=Je2=e̊1=Xx=x.e_{1}=-Je_{2}=\mathring{e}_{1}=X_{x}=\frac{\partial}{\partial x}.

We proceed to compute the α\alpha-function, aa and bb in terms of (x,y)(x,y), which is a compatible coordinate system. These are obtained from αe2+T1+α2=aXx+bXy\frac{\alpha e_{2}+T}{\sqrt{1+\alpha^{2}}}=aX_{x}+bX_{y} immediately as follows.

(7.8) a=0,b=α1+α2,andα=12x+g(y).a=0,\ b=\frac{\alpha}{\sqrt{1+\alpha^{2}}},\ \textrm{and}\ \ \alpha=\frac{1}{2x+g^{\prime}(y)}.

(i) Thus, from the formula of bb, the coordinates (x,y)(x,y) is a normal coordinate system on the part where 2x+g(y)>02x+g^{\prime}(y)>0, and we have ζ1(y)=g(y)\zeta_{1}(y)=g^{\prime}(y). It is easy to see that the graph of u=xy+g(y)u=xy+g(y), for gC()g\in C^{\infty}(\operatorname{\mathbb{R}}), is just the maximal surface SII+(g(y))S_{II}^{+}(g^{\prime}(y)) when we restrict to the domain {(x,y)|y, 2x+g(y)>0}\{(x,y)\ |\ y\in\operatorname{\mathbb{R}},\ 2x+g^{\prime}(y)>0\}.

(ii) For the other part with 2x+g(y)<02x+g^{\prime}(y)<0, we have b<0b<0. Therefore, instead of (x,y)(x,y), the new coordinate system (x~,y~)=(x,y)(\tilde{x},\tilde{y})=(x,-y) is a normal coordinate system (notice that the compatible coordinates are chosen so that b>0b>0). The invariants α,a\alpha,a and bb read

(7.9) a=0,b=α1+α2>0,andα=12x+g(y~)<0,a=0,\ b=-\frac{\alpha}{\sqrt{1+\alpha^{2}}}>0,\ \textrm{and}\ \ \alpha=\frac{1}{2x+g^{\prime}(-\tilde{y})}<0,

and hence ζ1(y~)=g(y~)\zeta_{1}(\tilde{y})=g^{\prime}(-\tilde{y}). Here means the derivative with respect to yy.

(iii) For the other part with 2x+g(y)<02x+g^{\prime}(y)<0, we can say something more. If, instead of x\frac{\partial}{\partial x}, we choose x-\frac{\partial}{\partial x} as the characteristic direction, that is, e1=xe_{1}=-\frac{\partial}{\partial x}, then the coordinates (x~,y~)=(x,y)(\tilde{x},\tilde{y})=(-x,-y) lead to the normal coordinate system for the part with 2x+g(y)<02x+g^{\prime}(y)<0. As a result, we consider the re-parametrization of the surface

X:(x~,y~)(x~,y~,x~y~+g(y~))X:(\tilde{x},\tilde{y})\rightarrow(-\tilde{x},-\tilde{y},\tilde{x}\tilde{y}+g(-\tilde{y}))

such that e1=x~=Xx~e_{1}=\frac{\partial}{\partial\tilde{x}}=X_{\tilde{x}}. Similarly, from αe2+T1+α2=aXx~+bXy~\frac{\alpha e_{2}+T}{\sqrt{1+\alpha^{2}}}=aX_{\tilde{x}}+bX_{\tilde{y}}, we have

(7.10) a=0,b=α1+α2>0,andα=12x~+g~y~(y~)>0,a=0,\ b=\frac{\alpha}{\sqrt{1+\alpha^{2}}}>0,\ \textrm{and}\ \ \alpha=\frac{1}{2\tilde{x}+\frac{\partial\tilde{g}}{\partial\tilde{y}}(\tilde{y})}>0,

where g~(y~)\tilde{g}(\tilde{y}) is defined by g~(y~)=g(y~)\tilde{g}(\tilde{y})=g(-\tilde{y}). Thus (x~,y~)(\tilde{x},\tilde{y}) is the normal coordinate system and we have ζ1(y~)=g~y~(y~)\zeta_{1}(\tilde{y})=\frac{\partial\tilde{g}}{\partial\tilde{y}}(\tilde{y}).

Let R(g(y))R(g(y)) and L(g(y))L(g(y)) be the part of the surface with 2x+g(y)>02x+g^{\prime}(y)>0 and 2x+g(y)<02x+g^{\prime}(y)<0, respectively. In terms of the notations defined in Subsection 6.3, we see that R(g(y))=SII+(g(y))R(g(y))=S_{II}^{+}(g^{\prime}(y)) and L(g(y))=SII(g(y))L(g(y))=S_{II}^{-}(g^{\prime}(y)). Then, comparing with (7.8) and (7.10), we immediately have the following proposition, due to Theorem 1.5.

Proposition 7.7.

The surface L(g(y))(orSII(g(y)))L(g(-y))\left(\textrm{or}\ S_{II}^{-}(-g^{\prime}(y))\right) and R(g(y))(orSII+(g(y)))R(g(y))\left(\textrm{or}\ S_{II}^{+}(g^{\prime}(y))\right) are congruent to each other. They in fact differ by an action of the Heisenberg rigid motion (x,y,t)(x,y,t)(x,y,t)\rightarrow(-x,-y,t).

Theorem 7.8.

Any pp-minimal surface of special type II is locally a part of the surface defined by u=xy+g(y)u=xy+g(y) for some gC()g\in C^{\infty}(\mathbb{R}), up to a Heisenberg rigid motion. In addition, it is symmetric if and only if g(y)g(y) is linear in the variable yy. Therefore, any symmetric pp-minimal surface of special type II is locally a part of the surface defined by the graph of u=xyu=xy, up to a Heisenberg rigid motion.

Proof.

Any pp-minimal surface of special type II locally has the following normal representation

(7.11) a=0,b=|α|1+α2,α=12x+ζ1(y),a=0,\ b=\frac{|\alpha|}{\sqrt{1+\alpha^{2}}},\ \ \alpha=\frac{1}{2x+\zeta_{1}(y)},

in terms of normal coordinates (x,y)(x,y). Therefore, comparing with (7.8), the proof is finished if we choose gg such that g(y)=ζ1(y)g^{\prime}(y)=\zeta_{1}(y). Moreover, it is symmetric if and only if ζ1(y)=g(y)=\zeta_{1}(y)=g^{\prime}(y)= constant, i.e., gg is linear in yy. ∎

7.3. Examples of types I, II and III

We consider the surface ΣH1\Sigma\in H_{1} defined on 2\operatorname{\mathbb{R}}^{2} by

(7.12) X:(s,t)(x,y,z)=(scosθ(t),ssinθ(t),t).X:(s,t)\rightarrow(x,y,z)=(s\cos\theta(t),s\sin\theta(t),t).

Then it can be calculated that

Xs=(cosθ(t),sinθ(t),0),Xt=(sθsinθ(t),sθcosθ(t),1).X_{s}=(\cos\theta(t),\sin\theta(t),0),\quad X_{t}=(-s\theta^{\prime}\sin\theta(t),s\theta^{\prime}\cos\theta(t),1).

Notice that e̊1|(0,0,z)=x\mathring{e}_{1}|_{(0,0,z)}=\frac{\partial}{\partial x}, e̊2|(0,0,z)=y\mathring{e}_{2}|_{(0,0,z)}=\frac{\partial}{\partial y} and θ(t)=θ(z)\theta(t)=\theta(z), i.e., θ\theta is independent of xx and yy. We rewrite XsX_{s} as

(7.13) Xs=cosθ(t)x+sinθ(t)y=cosθ(t)e̊1(X)+sinθ(t)e̊2(X)ξ,X_{s}=\cos\theta(t)\frac{\partial}{\partial x}+\sin\theta(t)\frac{\partial}{\partial y}=\cos\theta(t)\mathring{e}_{1}(X)+\sin\theta(t)\mathring{e}_{2}(X)\in\xi,

which is a vector tangent to the contact plane. Then we choose e1=Xse_{1}=X_{s}, and hence e2=Je1=sinθ(t)e̊1(X)+cosθ(t)e̊2(X)e_{2}=Je_{1}=-\sin\theta(t)\mathring{e}_{1}(X)+\cos\theta(t)\mathring{e}_{2}(X), which yields

(7.14) e1e2=(e1θ(t))(cosθ(t)e̊1(X)sinθ(t)e̊2(X))=0,(e1θ(t)=dθ(t)ds=0).\begin{split}\bigtriangledown_{e_{1}}e_{2}&=-(e_{1}\theta(t))(\cos\theta(t)\mathring{e}_{1}(X)-\sin\theta(t)\mathring{e}_{2}(X))\\ &=0,\ \left(\because e_{1}\theta(t)=\frac{d\theta(t)}{ds}=0\right).\end{split}

This implies that such surface defined by (7.12) has pp-mean curvature H=0H=0. We proceed to work out the α\alpha-function α\alpha, aa and bb. By definition, it is a function satisfying αe2+TTΣ\alpha e_{2}+T\in T\Sigma, that is,

(7.15) α(sinθe̊1+cosθe̊2)+T=EXs+FXt,\alpha(-\sin\theta\mathring{e}_{1}+\cos\theta\mathring{e}_{2})+T=EX_{s}+FX_{t},

for some functions E,FE,F. Similarly, we rewrite XsX_{s} as a linear combination of e̊1,e̊2\mathring{e}_{1},\mathring{e}_{2} and z\frac{\partial}{\partial z}, and we can express XtX_{t} as

(7.16) Xt=(sθsinθ)e̊1+(sθcosθ)e̊2+(s2θ+1)z.X_{t}=(-s\theta^{\prime}\sin{\theta})\mathring{e}_{1}+(s\theta^{\prime}\cos{\theta})\mathring{e}_{2}+(s^{2}\theta^{\prime}+1)\frac{\partial}{\partial z}.

Combining (7.15) and (7.16) and notice that Xs=e1X_{s}=e_{1}, we obtain that E=0,F=1s2θ(t)+1E=0,\ F=\frac{1}{s^{2}\theta^{\prime}(t)+1} and hence

α=sθ(t)s2θ(t)+1,a=0,b=F1+α2.\alpha=\frac{s\theta^{\prime}(t)}{s^{2}\theta^{\prime}(t)+1},\ a=0,\ b=\frac{F}{\sqrt{1+\alpha^{2}}}.

If θ(t)=0\theta^{\prime}(t)=0, then we have α=0\alpha=0. However, if θ(t)0\theta^{\prime}(t)\neq 0, then α\alpha, aa and bb read

(7.17) α=ss2+1θ(t),a=0,b=αs1+α21θ(t),\alpha=\frac{s}{s^{2}+\frac{1}{\theta^{\prime}(t)}},\ a=0,\ b=\frac{\alpha}{s\sqrt{1+\alpha^{2}}}\frac{1}{\theta^{\prime}(t)},

which means that (s,t)(s,t) is an orthogonal coordinate system, but not normal. In particular, if θ(t)>0\theta^{\prime}(t)>0 for all tt, then one sees that the pp-minimal surface has no singularities. From (7.17), we conclude that this surface is of type I if θ(t)>0\theta^{{}^{\prime}}(t)>0 for all tt. On the other hand, if θ(t)<0\theta^{{}^{\prime}}(t)<0 for all tt, then it is either of type II on which s>1θ(t)(t)s>\sqrt{-\frac{1}{\theta^{\prime}(t)}(t)} or s<1θ(t)(t)s<-\sqrt{-\frac{1}{\theta^{\prime}(t)}(t)}; or type III on which 1θ(t)(t)<s<1θ(t)(t)-\sqrt{-\frac{1}{\theta^{\prime}(t)}(t)}<s<\sqrt{-\frac{1}{\theta^{\prime}(t)}(t)}.
Finally, we can further take the coordinates (s~,t~)=(sC,θ(t))(\tilde{s},\tilde{t})=(s-C,\theta(t)) to normalize bb such that it only depends on ss and α\alpha, then we have

(7.18) α=s~+C(s~+C)2+1θ(θ1(t~)),a=0,b=α(s~+C)1+α2,\alpha=\frac{\tilde{s}+C}{(\tilde{s}+C)^{2}+\frac{1}{\theta^{\prime}(\theta^{-1}(\tilde{t}))}},\ a=0,\ b=\frac{\alpha}{(\tilde{s}+C)\sqrt{1+\alpha^{2}}},

for some constant CC, and hence

(7.19) ζ1(t~)=C,ζ2(t~)=1θ(θ1(t~)).\zeta_{1}(\tilde{t})=C,\ \ \zeta_{2}(\tilde{t})=\frac{1}{\theta^{\prime}(\theta^{-1}(\tilde{t}))}.
Refer to caption
Figure 7.3. Helicoid
Refer to caption
Figure 7.4. Conicoid

8. Structures of singular sets of pp-minimal surfaces

In this section, we assume that ΣH1\Sigma\subset H_{1} is a pp-minimal surface.

Proposition 8.1.

Let pp be a singular point of a pp-minimal surface Σ\Sigma. Then there must be a characteristic line approaching this point pp.

Proof.

Suppose no characteristic line approaches this point pp. We would like to find a contradiction. Firstly, by [3], there exists a small neighborhood of pp whose intersection with the singular set is contained in a smooth curve Γp\Gamma_{p}. If the neighborhood is small enough, then, on one side of the curve Γp\Gamma_{p}, we can find a compatible coordinate system (U;x,y)(U;x,y) such that pp is contained on the boundary of UU. Notice that, by our assumption, pp does not lie at the end of any leaf of the foliation defined by e1=xe_{1}=\frac{\partial}{\partial x}. Thus the image of the map defined on UU by (x,y)(α,αx)(x,y)\rightarrow(\alpha,\alpha_{x}) is bounded on the phase plane (see Figure 4.2). Therefore, we have that lim(x,y)pα(x,y)\lim_{(x,y)\rightarrow p}\alpha(x,y) is finite, which is a contradiction since pp is a singular point. ∎

8.1. The proof of Theorem 1.7

Due to Proposition 8.1, it suffices to show this theorem for a pp-minimal surface of some type. For general type (notice that there are no singular points for type I), we choose a normal coordinate system (x,y)(x,y) such that α\alpha, and a,ba,b read

(8.1) α(x,y)=x+ζ1(y)(x+ζ1(y))2c2(y),anda=0,b=|α||x+ζ1(y)|1+α2,\alpha(x,y)=\frac{x+\zeta_{1}(y)}{(x+\zeta_{1}(y))^{2}-c^{2}(y)},\ \textrm{and}\ a=0,\ b=\frac{|\alpha|}{|x+\zeta_{1}(y)|\sqrt{1+\alpha^{2}}},

where c(y)c(y) is a positive function of the variable yy such that ζ2(y)=c2(y)\zeta_{2}(y)=-c^{2}(y). Then the singular set is the graphs of the functions

(8.2) x=ζ1(y)±c(y).x=-\zeta_{1}(y)\pm c(y).

By (6.7), the induced metric II (or the first fundamental form) on the regular part reads

(8.3) I=dxdx+[(x+ζ1(y))2+[(x+ζ1(y))2+ζ2(y)]2]dydy.I=dx\otimes dx+\big{[}(x+\zeta_{1}(y))^{2}+[(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)]^{2}\big{]}dy\otimes dy.

Now we use the metric to compute the length of the singular set {(ζ1(y)±c(y),y)}\{(-\zeta_{1}(y)\pm c(y),y)\}, where yy belongs to some open interval. Let γ±(y)=(ζ1(y)±c(y),y)\gamma_{\pm}(y)=(-\zeta_{1}(y)\pm c(y),y), which is a parametrization of the singular set. Then the square of the velocity at yy is

(8.4) |γ±(y)|2=|ζ1(y)±c(y)|2+c2(y)>0for ally,|\gamma^{\prime}_{\pm}(y)|^{2}=|-\zeta_{1}^{\prime}(y)\pm c^{\prime}(y)|^{2}+c^{2}(y)>0\ \textrm{for all}\ y,

where we have used the fact that (x+ζ1(y))2=c2(y)(x+\zeta_{1}(y))^{2}=c^{2}(y) on the singular set. Formula (8.4) shows that the parametrized curve γ±(y)\gamma_{\pm}(y) of the singular set has a positive length. We omit similar proof for special type II.

Finally, for special type I, in terms of a normal coordinate system (x,y)(x,y), we have

(8.5) α=1x+ζ1(y),anda=0,b=α21+α2.\alpha=\frac{1}{x+\zeta_{1}(y)},\ \textrm{and}\ a=0,\ b=\frac{\alpha^{2}}{\sqrt{1+\alpha^{2}}}.

If γ(y)=(ζ1(y),y)\gamma(y)=(-\zeta_{1}(y),y) is a parametrization of the singular set {(ζ1(y),y)}\{(-\zeta_{1}(y),y)\} for yy inside some open interval, then (6.7) indicates that the induced metric II on the regular part reads

(8.6) dxdx+[(x+ζ1(y))2+(x+ζ1(y))4]dydydx\otimes dx+\big{[}(x+\zeta_{1}(y))^{2}+(x+\zeta_{1}(y))^{4}\big{]}dy\otimes dy

and the square of the velocity at yy is

(8.7) |γ(y)|2=(ζ1(y))2+[(x+ζ1(y))2+(x+ζ1(y))4]=(ζ1(y))2,|\gamma^{\prime}(y)|^{2}=(\zeta_{1}^{\prime}(y))^{2}+\big{[}(x+\zeta_{1}(y))^{2}+(x+\zeta_{1}(y))^{4}\big{]}=(\zeta_{1}^{\prime}(y))^{2},

where we have used the fact that x=ζ1(y)x=-\zeta_{1}(y) on the singular set. From formula (8.7), we see that the length of γ(y)\gamma(y) depends on whether the value ζ1(y)\zeta_{1}^{\prime}(y) is zero or not, which implies that the singular set is either an isolated point or a smooth curve of positive length. In addition, the singular set as an isolated point happens if and only if ζ1\zeta_{1} is a constant, that is, the surface is part of a plane. We thus complete the proof of this theorem 1.7.

8.2. The proof of Theorem 1.8

Around the singular point pp, we may assume that the surface is represented by a graph z=u(x,y)z=u(x,y). Let XX be a parametrization of the pp-minimal surface around pp defined by X(x,y)=(x,y,u(x,y))X(x,y)=(x,y,u(x,y)). Then

(8.8) Xx=(1,0,ux)=x+uxt=e̊1+(uxy)t;Xy=(0,1,uy)=y+uyt=e̊2+(uy+x)t,\begin{split}X_{x}&=(1,0,u_{x})=\frac{\partial}{\partial x}+u_{x}\frac{\partial}{\partial t}=\mathring{e}_{1}+(u_{x}-y)\frac{\partial}{\partial t};\\ X_{y}&=(0,1,u_{y})=\frac{\partial}{\partial y}+u_{y}\frac{\partial}{\partial t}=\mathring{e}_{2}+(u_{y}+x)\frac{\partial}{\partial t},\end{split}

which yields

I(Xx,Xx)=1+(uxy)2,I(Xy,Xy)=1+(uy+x)2,I(Xx,Xy)=(uxy)(uy+x),I(X_{x},X_{x})=1+(u_{x}-y)^{2},\ \ I(X_{y},X_{y})=1+(u_{y}+x)^{2},\ \ I(X_{x},X_{y})=(u_{x}-y)(u_{y}+x),

where II is the induced metric (first fundamental form) on the surface. Now we choose a horizontal normal as follows

e2=(uxy)e̊1+(uy+x)e̊2D,e_{2}=-\frac{(u_{x}-y)\mathring{e}_{1}+(u_{y}+x)\mathring{e}_{2}}{D},

where D=((uxy)2+(uy+x)2)1/2D=\left((u_{x}-y)^{2}+(u_{y}+x)^{2}\right)^{1/2}. Then

(8.9) e1=(uy+x)e̊1(uxy)e̊2De_{1}=\frac{(u_{y}+x)\mathring{e}_{1}-(u_{x}-y)\mathring{e}_{2}}{D}

is tangent to the characteristic curves.

We first claim that either uxx(p)0u_{xx}(p)\neq 0 or (uxy+1)(p)0(u_{xy}+1)(p)\neq 0. Let f(x,y)=uxyf(x,y)=u_{x}-y and let (x(s),y(s))(x(s),y(s)) be a parametrization of the singular curve passing through pp. Notice that we may assume, w.l.o.g., that the xx-axis past pp is transverse to the singular curve, i.e., y0y^{\prime}\neq 0. Since f(x(s),y(s))=0f(x(s),y(s))=0, taking derivative with respect to ss gives uxxx+(uxy1)y=0u_{xx}x^{\prime}+(u_{xy}-1)y^{\prime}=0. Therefore, (uxy1)(p)=0(u_{xy}-1)(p)=0 if uxx(p)=0u_{xx}(p)=0, and hence (uxy+1)(p)=2(u_{xy}+1)(p)=2.

If uxx(p)0u_{xx}(p)\neq 0, we turn to compute the angle ζ\zeta between e1e_{1} and XxX_{x}. First, from (8.8), we have

I(e1,Xx)=|e1||Xx|cosζ=(1+(uxy)2)1/2cosζ.I(e_{1},X_{x})=|e_{1}||X_{x}|\cos{\zeta}=(1+(u_{x}-y)^{2})^{1/2}\cos{\zeta}.

On the other hand, using (8.9) to get

I(e1,Xx)=(uy+x)D.I(e_{1},X_{x})=\frac{(u_{y}+x)}{D}.

Combining the above two formulae, we obtain

(8.10) cosζ=uy+xD1+(uxy)2=uy+xuxyDuxy1+(uxy)2=±uy+xuxy1+(uy+xuxy)21+(uxy)2,\begin{split}\cos{\zeta}&=\frac{u_{y}+x}{D\sqrt{1+(u_{x}-y)^{2}}}=\frac{\frac{u_{y}+x}{u_{x}-y}}{\frac{D}{u_{x}-y}\sqrt{1+(u_{x}-y)^{2}}}\\ &=\pm\frac{\frac{u_{y}+x}{u_{x}-y}}{\sqrt{1+(\frac{u_{y}+x}{u_{x}-y})^{2}}\sqrt{1+(u_{x}-y)^{2}}},\end{split}

where the sign ±\pm depends on that the sign of uxyu_{x}-y is positive or not. By the mean value theorem, it is easy to see (or see [3]) that

(8.11) limqp+uy+xuxy=uxy+1uxx(p)=limqpuy+xuxy,\lim_{q\rightarrow p^{+}}\frac{u_{y}+x}{u_{x}-y}=\frac{u_{xy}+1}{u_{xx}}(p)=\lim_{q\rightarrow p^{-}}\frac{u_{y}+x}{u_{x}-y},

and thus

(8.12) limqp+cosζ=uxy+1uxx(p)1+(uxy+1uxx(p))2=limqpcosζ,\lim_{q\rightarrow p^{+}}\cos{\zeta}=\frac{\frac{u_{xy}+1}{u_{xx}}(p)}{\sqrt{1+\left(\frac{u_{xy}+1}{u_{xx}}(p)\right)^{2}}}=-\lim_{q\rightarrow p^{-}}\cos{\zeta},

where limqp+(limqp)\lim_{q\rightarrow p^{+}}(\lim_{q\rightarrow p^{-}}) means that qpq\rightarrow p from the side in which uxyu_{x}-y is positive (negative).

If (uxy+1)(p)0(u_{xy}+1)(p)\neq 0, similar computations give the angle η\eta between e1e_{1} and XyX_{y} by

(8.13) cosη=(uxyuy+x)±1+(uxyuy+x)21+(uy+x)2,\cos{\eta}=\frac{-(\frac{u_{x}-y}{u_{y}+x})}{\pm\sqrt{1+(\frac{u_{x}-y}{u_{y}+x})^{2}}\sqrt{1+(u_{y}+x)^{2}}},

thus

(8.14) limqp+cosη=uxxuxy+1(p)1+(uxxuxy+1(p))2=limqpcosη,\lim_{q\rightarrow p^{+}}\cos{\eta}=-\frac{\frac{u_{xx}}{u_{xy}+1}(p)}{\sqrt{1+\left(\frac{u_{xx}}{u_{xy}+1}(p)\right)^{2}}}=-\lim_{q\rightarrow p^{-}}\cos{\eta},

where limqp+(limqp)\lim_{q\rightarrow p^{+}}(\lim_{q\rightarrow p^{-}}) means that qpq\rightarrow p from the side in which uy+xu_{y}+x is positive (negative).

From (8.11), it is easy to see that both uxyu_{x}-y and uy+xu_{y}+x differ by a sign on the different side of the singular curve which is defined by uxy=0u_{x}-y=0 and uy+x=0u_{y}+x=0. Therefore, from the formula of e1e_{1} (see (8.9)), together with (8.12) and (8.14), we conclude that the characteristic vector field e1e_{1} differs by a sign on the different side of the singular curve when approaching the singular point pp. This completes the proof of theorem 1.8.

8.3. The proof of Theorem 1.9

In terms of normal coordinates (x,y)(x,y), the surface Σ\Sigma is represented by two functions ζ1(y)\zeta_{1}(y) and ζ2(y)\zeta_{2}(y). Since it is of type II, we have ζ2(y)<0\zeta_{2}(y)<0 and

(8.15) α=x+ζ1(y)(x+ζ1(y))2+ζ2(y),a=0,b=|α||x+ζ1(y)|1+α2,\alpha=\frac{x+\zeta_{1}(y)}{(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)},\ a=0,\ b=\frac{|\alpha|}{|x+\zeta_{1}(y)|\sqrt{1+\alpha^{2}}},

on which either x+ζ1(y)>ζ2(y)x+\zeta_{1}(y)>\sqrt{-\zeta_{2}(y)} or x+ζ1(y)<ζ2(y)x+\zeta_{1}(y)<-\sqrt{-\zeta_{2}(y)}. The induced metric is

(8.16) I=dxdx+1b2dydy.I=dx\otimes dx+\frac{1}{b^{2}}dy\otimes dy.

We assume that Σ\Sigma lies on the part x+ζ1(y)>ζ2(y)x+\zeta_{1}(y)>\sqrt{-\zeta_{2}(y)} (the proof for the case that Σ\Sigma lies on the part x+ζ1(y)<ζ2(y)x+\zeta_{1}(y)<-\sqrt{-\zeta_{2}(y)} is similar). Suppose, in addition, that Σ\Sigma can be smoothly extended beyond the singular curve x+ζ1(y)ζ2(y)=0x+\zeta_{1}(y)-\sqrt{-\zeta_{2}(y)}=0. By theorem 1.8, the coordinates (x,y)(x,y) can be extended beyond the singular curve to be compatible coordinates. Then the α\alpha-function on the other side of the singular curve must be one of the following

  1. (1)

    1x+ζ1(y)ζ2(y)\frac{1}{x+\zeta_{1}(y)-\sqrt{-\zeta_{2}(y)}}, which is of special type I;

  2. (2)

    12x+2(ζ1(y)ζ2(y))\frac{1}{2x+2(\zeta_{1}(y)-\sqrt{-\zeta_{2}(y)})}, which is of special type II;

  3. (3)

    α=x+ζ1(y)(x+ζ1(y))2+ζ2(y)\alpha=\frac{x+\zeta_{1}(y)}{(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)}, which is of general type,

for x+ζ1(y)<ζ2(y)x+\zeta_{1}(y)<\sqrt{-\zeta_{2}(y)}. The induced metric on this other part is

(8.17) I=dxdxabdxdyabdydx+(1+a2)b2dydy.I=dx\otimes dx-\frac{a}{b}dx\otimes dy-\frac{a}{b}dy\otimes dx+\frac{(1+a^{2})}{b^{2}}dy\otimes dy.

Comparing (8.16) and (8.17), and noting that II is smooth around the singular curve, we have

a=0,b=|α||x+ζ1(y)|1+α2,a=0,\ b=\frac{|\alpha|}{|x+\zeta_{1}(y)|\sqrt{1+\alpha^{2}}},

with α=x+ζ1(y)(x+ζ1(y))2+ζ2(y)\alpha=\frac{x+\zeta_{1}(y)}{(x+\zeta_{1}(y))^{2}+\zeta_{2}(y)}. That is, cases (1) and (2) for α\alpha do not happen. Therefore, the extended coordinates beyond the singular curve are also normal coordinates. The formula of α\alpha shows that the part on the other side of the singular curve is of type III. This completes the proof of Theorem 1.9.

8.4. The proof of the Bernstein-type theorem

In this subsection, we will show that (7.1) and (7.5) are the only entire smooth pp-minimal graphs. Suppose that ΣH1\Sigma\subset H_{1} is an entire pp-minimal graph. First of all, since it is a graph, we notice that there is nowhere at which α\alpha is zero. Next, we claim the following lemma.

Lemma 8.2.

The induced singular characteristic foliation of Σ\Sigma does not contain a leaf along which α\alpha is of general type, that is, in terms of normal coordinates around the leaf, α\alpha is a general solution of the Codazzi-like equation ((see (4.1))).

Proof.

Suppose not. We assume that the induced singular characteristic foliation of Σ\Sigma contains such a leaf. Then there will be a piece of the surface (a neighborhood) around the leaf such that this piece is of general type. Suppose that this piece is of type I or of type III, then the entireness and the phase plane (Figure 4.2) indicate that the α\alpha-function must be extended so that it has a zero somewhere. This is a contradiction. Therefore this piece (of general type) must be of type II. Again, since it is entire, this piece can be smoothly extended through the singular curve. By Theorem 1.9, it contains a piece of type III, which lies on the other side of the singular curve. This is also a contradiction, as we argue above. We hence complete the proof of Lemma 8.2. ∎

From Lemma 8.2, we know that an entire pp-minimal graph is either of special type I or of special type II. If it is of special type II, Theorem 7.8 and Lemma 7.5 ensure that Σ\Sigma is one of the graphs in (7.5). If it contains a piece of special type I, then this piece must be symmetric by Theorem 7.4. Therefore, by Theorem 7.3 and Lemma 7.1, the surface Σ\Sigma must be one of the graphs in (7.1). We hence complete the proof of the Bernstein-type theorem. We also remark that the Bernstein-type theorem still holds for C3C^{3} surfaces in H1H_{1}.

Remark 8.3.

We point out that in [3], the Bernstein-type theorem had been proved in C2C^{2}-graphs.

9. An approach to construct pp-minimal surfaces

In this section, we provide an approach to constructing pp-minimal surfaces. It turns out that, in some sense, generic pp-minimal surfaces can be constructed by this approach, particularly, other than those pp-minimal surfaces of special type I. This approach is to perturb the surface u=0u=0 in some way. Recall we choose the parametrization of u=0u=0 by

X:(r,θ)(rcosθ,rsinθ,0),r>0,X:(r,\theta)\rightarrow(r\cos{\theta},r\sin{\theta},0),\ \ r>0,

where each half-ray lθ:r(rcosθ,rsinθ,0)l_{\theta}:r\rightarrow(r\cos{\theta},r\sin{\theta},0) with a fixed angle θ\theta is a Legendrian straight line. Therefore, the image of the action of each Heisenberg rigid motion on lθl_{\theta} is also a Legendrian straight line. Let 𝒞\mathcal{C} be an arbitrary curve 𝒞:θ(x(θ),y(θ),z(θ)),θ\mathcal{C}:\theta\rightarrow(x(\theta),y(\theta),z(\theta)),\ \theta\in\operatorname{\mathbb{R}}. Then for each fixed θ\theta and r>0r>0, the curve defined by

L𝒞(θ)(lθ):r(x(θ)+rcosθ,y(θ)+rsinθ,z(θ)+ry(θ)cosθrx(θ)sinθ)L_{\mathcal{C}(\theta)}(l_{\theta}):r\rightarrow(x(\theta)+r\cos{\theta},y(\theta)+r\sin{\theta},z(\theta)+ry(\theta)\cos{\theta}-rx(\theta)\sin{\theta})

is a Legendrian straight line. Here L𝒞(θ)L_{\mathcal{C}(\theta)} is the left translation by 𝒞(θ)\mathcal{C}(\theta). Therefore, the union of these lines constitutes a pp-minimal surface with a parametrization YY given by

(9.1) Y(r,θ)=(x(θ)+rcosθ,y(θ)+rsinθ,z(θ)+ry(θ)cosθrx(θ)sinθ).Y(r,\theta)=(x(\theta)+r\cos{\theta},y(\theta)+r\sin{\theta},z(\theta)+ry(\theta)\cos{\theta}-rx(\theta)\sin{\theta}).

This surface depends on the curve 𝒞(θ)=(x(θ),y(θ),z(θ))\mathcal{C}(\theta)=(x(\theta),y(\theta),z(\theta)). We have the following proposition about the surface YY.

Proposition 9.1.

The coordinates (r,θ)(r,\theta) are compatible coordinates for YY. In terms of this coordinate system, the α\alpha-invariant and the induced metric read

(9.2) a=(x(θ)cosθ+y(θ)sinθ)α[r+(y(θ)cosθx(θ)sinθ)]1+α2b=α[r+(y(θ)cosθx(θ)sinθ)]1+α2,\begin{split}a&=\frac{-(x^{\prime}(\theta)\cos{\theta}+y^{\prime}(\theta)\sin{\theta})\alpha}{[r+(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})]\sqrt{1+\alpha^{2}}}\\ &\\ b&=\frac{\alpha}{[r+(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})]\sqrt{1+\alpha^{2}}},\end{split}

and

(9.3) α=r+(y(θ)cosθx(θ)sinθ)[r+(y(θ)cosθx(θ)sinθ)]2+Θ(𝒞(θ))(y(θ)cosθx(θ)sinθ)2,\alpha=\frac{r+(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})}{[r+(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})]^{2}+\Theta(\mathcal{C}^{\prime}(\theta))-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2}},

where Θ(𝒞(θ))=z(θ)+x(θ)y(θ)y(θ)x(θ)\Theta(\mathcal{C}^{\prime}(\theta))=z^{\prime}(\theta)+x(\theta)y^{\prime}(\theta)-y(\theta)x^{\prime}(\theta).

Proof.

We make a straightforward computation for the invariants α,a\alpha,a and bb. Firstly, we have

(9.4) Yr=(cosθ,sinθ,y(θ)cosθx(θ)sinθ)=cosθe̊1(Y(r,θ))+sinθe̊2(Y(r,θ)).\begin{split}Y_{r}&=(\cos{\theta},\sin{\theta},y(\theta)\cos{\theta}-x(\theta)\sin{\theta})\\ &=\cos{\theta}\ \mathring{e}_{1}(Y(r,\theta))+\sin{\theta}\ \mathring{e}_{2}(Y(r,\theta)).\end{split}

From the construction of YY, we have e1=Yre_{1}=Y_{r}. Thus

(9.5) e2=Je1=sinθe̊1(Y(r,θ))+cosθe̊2(Y(r,θ)),e_{2}=Je_{1}=-\sin{\theta}\ \mathring{e}_{1}(Y(r,\theta))+\cos{\theta}\ \mathring{e}_{2}(Y(r,\theta)),

whereas we have

Yθ=(x(θ)rsinθ,y(θ)+rcosθ,z(θ)+r(y(θ)cosθy(θ)sinθx(θ)sinθx(θ)cosθ)).Y_{\theta}=(x^{\prime}(\theta)-r\sin{\theta},y^{\prime}(\theta)+r\cos{\theta},z^{\prime}(\theta)+r(y^{\prime}(\theta)\cos{\theta}-y(\theta)\sin{\theta}-x^{\prime}(\theta)\sin{\theta}-x(\theta)\cos{\theta})).

If we let

(9.6) Yθ=Ae̊1(Y(r,θ))+Be̊2(Y(r,θ))+Cz,Y_{\theta}=A\ \mathring{e}_{1}(Y(r,\theta))+B\ \mathring{e}_{2}(Y(r,\theta))+C\ \frac{\partial}{\partial z},

for some functions A,BA,B and CC. Then straightforward computations show that

(9.7) A=x(θ)rsinθ,B=y(θ)+rcosθ,C=z(θ)x(θ)y(θ)+y(θ)x(θ)+2r(y(θ)cosθx(θ)sinθ)+r2.\begin{split}A&=x^{\prime}(\theta)-r\sin{\theta},\ \ B=y^{\prime}(\theta)+r\cos{\theta},\\ C&=z^{\prime}(\theta)-x^{\prime}(\theta)y(\theta)+y^{\prime}(\theta)x(\theta)+2r(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})+r^{2}.\end{split}

We recall that the three invariants α,a\alpha,a and bb are related by

(9.8) αe2+T1+α2=aYr+bYθ.\frac{\alpha e_{2}+T}{\sqrt{1+\alpha^{2}}}=aY_{r}+bY_{\theta}.

If we substitute (9.4), (9.5) and (9.6) into (9.8), and compare the corresponding coefficients, we then obtain (9.2) and (9.3). ∎

Remark 9.2.

Let D=y(θ)cosθx(θ)sinθD=y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta}. By (9.4),(9.6) and (9.7), we have

0=YrYθBcosθAsinθ=0,Ccosθ=0,Csinθ=0r+D=0,C=0r+D=0,Θ(𝒞(θ))+2rD+r2=0,by(9.7),r+D=0,r=D±D2Θ(𝒞(θ))r+D=0,Θ(𝒞(θ))D2=0.\begin{split}0=Y_{r}\wedge Y_{\theta}&\Leftrightarrow B\cos{\theta}-A\sin{\theta}=0,\ C\cos{\theta}=0,\ C\sin{\theta}=0\\ &\Leftrightarrow r+D=0,\ C=0\\ &\Leftrightarrow r+D=0,\ \Theta(\mathcal{C}^{\prime}(\theta))+2rD+r^{2}=0,\ \ \textrm{by}\ \eqref{7031},\\ &\Leftrightarrow r+D=0,\ r=-D\pm\sqrt{D^{2}-\Theta(\mathcal{C}^{\prime}(\theta))}\\ &\Leftrightarrow r+D=0,\ \Theta(\mathcal{C}^{\prime}(\theta))-D^{2}=0.\end{split}

We conclude that YY is an immersion if and only if either Θ(𝒞(θ))D20orr+D0\Theta(\mathcal{C}^{\prime}(\theta))-D^{2}\neq 0\ \textrm{or}\ r+D\neq 0 for all θ\theta, where Θ(𝒞(θ))=z(θ)x(θ)y(θ)+y(θ)x(θ)\Theta(\mathcal{C}^{\prime}(\theta))=z^{\prime}(\theta)-x^{\prime}(\theta)y(\theta)+y^{\prime}(\theta)x(\theta).

Formula (9.3) suggests the following: That YY defines a pp-minimal surface of special type depends on whether Θ(𝒞(θ))(y(θ)cosθx(θ)sinθ)2\Theta(\mathcal{C}^{\prime}(\theta))-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2} vanishes or not.

Comparing equations (1.12) and (1.14), it is convenient to regard surfaces of special type I as surfaces of general type with ζ2\zeta_{2}-invariant vanishing. Now given two arbitrary functions ζ1\zeta_{1} and ζ2\zeta_{2}, we solve equation system (1.14) for a smooth curve 𝒞(θ)=(x(θ),y(θ),z(θ))\mathcal{C}(\theta)=(x(\theta),y(\theta),z(\theta)). Since system (1.14) is equivalent to the following system

(9.9) {ζ1(θ)=y′′(θ)cosθx′′(θ)sinθ2(x(θ)cosθ+y(θ)sinθ),ζ2(θ)=z(θ)+x(θ)y(θ)y(θ)x(θ)(y(θ)cosθx(θ)sinθ)2,\left\{\begin{split}\zeta^{\prime}_{1}(\theta)&=y^{\prime\prime}(\theta)\cos{\theta}-x^{\prime\prime}(\theta)\sin{\theta}-2\big{(}x^{\prime}(\theta)\cos{\theta}+y^{\prime}(\theta)\sin{\theta}\big{)},\\ \zeta_{2}(\theta)&=z^{\prime}(\theta)+x(\theta)y^{\prime}(\theta)-y(\theta)x^{\prime}(\theta)-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2},\end{split}\right.

which is underdetermined. Therefore, the solutions always exist. For example, we can solve the first equation of (9.9) for (x(θ),y(θ))(x(\theta),y(\theta)) and then solve for z(θ)z(\theta) from the second one. It turns out that we can find a smooth curve 𝒞\mathcal{C} such that the corresponding pp-minimal surface YY has the two given functions ζ1\zeta_{1} and ζ2\zeta_{2} as its ζ1\zeta_{1}-and ζ2\zeta_{2}-invariants. If ζ2=0\zeta_{2}=0, then YY is of special type I. We thus conclude, together with Theorem 7.8 in which states parametrizations for pp-minimal surfaces of special type II, that we generically have provided a parametrization for any given pp-minimal surface with type. In particular, we give a parametrization presentation for the eight classes of maximal pp-minimal surfaces constructed in Subsection 6.3.

Finally, we point out that these pp-minimal surfaces constructed by curves defined by (1.10) and (1.13) are all immersed surfaces at least (in some cases, they are embedded). This is because that b~0\tilde{b}\neq 0 for all points. In particular, formula (1.14) says that if α~0\tilde{\alpha}\rightarrow 0 then

b~1|Θ(𝒞(θ))(y(θ)cosθx(θ)sinθ)2|,\tilde{b}\rightarrow\frac{1}{|\Theta(\mathcal{C}^{\prime}(\theta))-(y^{\prime}(\theta)\cos{\theta}-x^{\prime}(\theta)\sin{\theta})^{2}|},

which is not zero.

Example 9.3.

If we take 𝒞\mathcal{C} to be the curve 𝒞(θ)=(0,0,z(θ))\mathcal{C}(\theta)=(0,0,z(\theta)) with z(θ)0z^{\prime}(\theta)\neq 0, then

Y(r,θ)=(rcosθ,rsinθ,z(θ)).Y(r,\theta)=(r\cos{\theta},r\sin{\theta},z(\theta)).

Taking the new coordinates (s,t)=(r,z(θ))(s,t)=(r,z(\theta)), we recover the surface of general type in Subsection 7.3 (see Figure 7.4 for the case z(θ)=θz(\theta)=\theta).

Example 9.4.

If we take 𝒞\mathcal{C} to be the curve 𝒞(θ)=(sinθ,cosθ,θ)\mathcal{C}(\theta)=(-\sin{\theta},\cos{\theta},\theta), then

Y(r,θ)=(sinθ+rcosθ,cosθ+rsinθ,r).Y(r,\theta)=(-\sin{\theta}+r\cos{\theta},\cos{\theta}+r\sin{\theta},r).

The surface of type I in Subsection LABEL:exat1 (see Figure 7.4) can be recovered by taking a rotation by π2\frac{\pi}{2} about the zz-axis.

References

  • [1] V. Barone Adesi, F. Serra Cassano, and D. Vittone, The Bernstein problem for intrinsic graphs in Heisenberg groups and calibrations, Preprint.
  • [2] T. A. Burton, The Nonlinear Wave Equation as a Liénard Equation, Funkcialaj Ekvacioj, 34 (1991) 529-545.
  • [3] Cheng, J.-H.; Hwang, J.-F.; Malchiodi, A., and Yang, P., Minimal surfaces in Pseudohermitian geometry, Annali della Scuola Normale Superiore di Pisa Classe di Scienze V , 4 (1), 129-177, 2005.
  • [4] Cheng, J.-H.; Hwang, J.-F.; Malchiodi, A., and Yang, P., A Codazzi-like equation and the singular set for C1C^{1} smooth surfaces in the Heisenberg group, J. Reine angew. Math. 671 (2012), 131-198.
  • [5] Cheng, J.-H.; Chiu, H.-L.; Hwang, J.-F., and Yang, P.,Umbilicity and characterization of Pansu spheres in the Heisenberg group, Journal fur die reine und angewandte Mathematik (738), 203-235, 2018.
  • [6] Chiu, H.-L. and Lai, S.-H., The fundamental theorem for hypersurfaces in Heisenberg groups, Calc. Var. Partial Diffrential Equations, 54 (2015), no. 1, 1091-1118.
  • [7] D. Daniekki, N. Garofalo, D. M. Nhieu and S. D. Pauls, Instability of graphical strips and a positive answer to the Bernstein problem in the Heisenberg group H1H^{1}, J. Differential Geometry, 81, pp 251–295, 2009.
  • [8] D. Danielli, N. Garofalo, DM Nhieu and S., Pauls, The Bernstein problem for Embedded surfaces in the Heisenberg group H1H_{1}, Indiana University mathematics journal, pp 563-594, 2010.
  • [9] Zaitsev, V. F. and Polyanin, A. D., Discrete-Group Methods for Integrating Equations of Nonlinear Mechanics, CRC Press, Boca Raton, 1994.
  • [10] Polyanin, A.D. and Zaitsev,V.F., Handbook of Exact Solutions for Ordinary Differential Equations, 2nd Edition, Chapman and Hall/CRC, Boca Raton, 2003.
  • [11] Liénard, A., Etude des oscillations entretenues, Revue Générale de l’Electricité 23 (1928) 901-912 & 946-954.
  • [12] A. Chiellini, Sull’integrazione dell’equazione differenziale y+Py2+Qy3=0y^{\prime}+Py^{2}+Qy^{3}=0, Bollettino dell’Unione Matematica Italiana, 10, 301-307 (1931).
  • [13] M. J. Ablowitz, D. J. Kaup, A. C. Newell and H. Segur, Method for solving the Sine-Gordon equation, Phy. Rev. Lett. 30 (1973),1262–1264
  • [14] M. Remoissenet, Waves Called Solitons: Concepts and Experiments, Springer, 1994.
  • [15] A. Bäcklund, Zur Theorie der Flächentransformationen, Math. Ann. XIX 387–422 (1882).
  • [16] A. Bäcklund, On ytor med konstant negativ krökning, Lunds Univ. Å\mathring{A}rsskr XIX (1883).
  • [17] A. Bäcklund, Einiges über Kugelkomplexe, Annali di Matematica Ser III XX 65–107 (1913).
  • [18] O. Calin, Geodesics on a certain step 2 sub-Riemannian manifold, Number 22. Annals of Global Analysis and Geometry, pp. 317-339, 2002.
  • [19] O. Calin, D. C. Chang, and P. C. Greiner, On a Step 2(k + 1) sub-Riemannian manifold, volume 14. Journal of Geometric Analysis, pp. 1-18, 2004.
  • [20] U. Hamenstiidt, Some regularity theorems for Carnot-Caratheodory metrics, volume 32. J. Differential Geom., pp. 819-850, 1990.
  • [21] P. Hartman, Ordinary Differential equations, Wiley, 1984.
  • [22] W. Liu and H. J. Sussmann, Shortest paths for sub-Riemannian metrics on rank two distri- butions, volume 118. Mem. Amer. Math. Soc., 1995.
  • [23] R. Beals, B. Gaveau, and P. C. Greiner, Hamilton-Jacobi theory and the heat kernel on Heisenberg groups, Number 79. J. Math. Pure Appl., pp. 633-689, 2000.
  • [24] R. Beals, B. Gaveau, and P.C. Greiner, On a Geometric Formula for the Fundamental Solu- tion of Subelliptic Laplacians, Number 181. Math. Nachr., pp. 81-163, 1996.
  • [25] O. Calin, D. C. Chang, and P. C. Greiner, Geometric mechanics on the Heisenberg group, Bulletin of the Institute of Mathematics, Academia Sinica, 2005.
  • [26] O. Calin and V. Mangione, Variational calculus on sub-Riemannian manifolds, volume 8, Balcan Journal of Geometry and Applications, 2003.
  • [27] B. Gaveau, Principe de moindre action, propagation de la chaleur et estimees sous-elliptiques sur certains groupes nilpotents, volume 139. Acta Math, 1977, pp. 95-153.
  • [28] A. Koranyi, Geometric properties of Heisenberg groups, Advances in Math., pp. 28-38, 1985.
  • [29] A. Koranyi, Geometric aspects of analysis on the Heisenberg group, Topics in Modern Harmonic Analysis, pp. 209-258, May-June 1982.
  • [30] A. Koranyi and H. M. Riemann, Quasiconformal mappings in the Heisenberg group, Invent. Math., pp. 309-338, 1985.
  • [31] Gerald B. Folland, Fourier Analysis and its Applications, Brooks / Cole, Pacific Grove, 1992.
  • [32] Karl-Heinz Gröchenig, Foundations of Time-Frequency Analysis, Birkhäuser, Boston, 2000.
  • [33] Ernst Binz and Sonja Pods, The Geometry of Heisenberg Groups With Applications in Signal Theory, Optics, Quantization, and Field Quantization, AMS 2008.
  • [34] Malchiodi, A., Minimal surfaces in three dimensional pseudo-Hermitian geometry, Lecture Notes of Seminario Interdisciplinare di Matematica, Vol. 6, pp. 195-207, 2007.
  • [35] S. D. Pauls, Minimal surfaces in the Heisenberg group, Geometriae Dedicata,104, pp. 201-231, 2004.
  • [36] M. Ritoré and C. Rosales, Rotationally invariant Hypersurfaces with constant mean curvature in the Heisenberg group HnH^{n}, The Journal of Geometric Analysis, v.16, n.4 pp. 703-720 (2006).