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

Ramification and unicity theorems for Gauss maps of complete space-like stationary surfaces in four-dimensional Lorentz-Minkowski space

\fnmLi \surOu [email protected] \orgdivSchool of Mathematics, \orgnameSichuan university, \orgaddress\streetNo.24, Renmin South Road, \cityChengdu, \postcode610064, \stateSichuan Province, \countryChina
Abstract

In this paper, we investigate the value distribution properties for Gauss maps of space-like stationary surfaces in four-dimensional Lorentz-Minkowski space 3,1\mathbb{R}^{3,1}, focusing on aspects such as the number of totally ramified points and unicity properties. We not only obtain general conclusions similar to situations in four-dimensional Euclidean space, but also consider the space-like stationary surfaces with rational graphic Gauss image, which is an extension of degenerate space-like stationary surfaces.

keywords:
Gauss map, stationary surface, ramification, unicity theorem
pacs:
[

MSC Classification]53A10, 53C42, 32A22, 51B20, 30C15

1 Introduction

It is well known that value distribution properties for Gauss maps of regular minimal surfaces in n\mathbb{R}^{n} play a crucial role in the theory of minimal surfaces. For a minimal surface MM in 3\mathbb{R}^{3}, the Gauss map is defined by the unit normal vector G(p)𝕊2G(p)\in\mathbb{S}^{2}, for pMp\in M. The surface MM is canonically considered as an open Riemann surface with some conformally metric ds2ds^{2} and by the minimality of MM, the Gauss map is a meromorphic function under the sphere stereographic projection. S. S. Chern [1] introduced the generalized Gauss map GG for an oriented regular surface in n\mathbb{R}^{n}, which map pMp\in M to the point in Qn2={[(z1,,zn)]n1:z12++zn2=0}Q_{n-2}=\{[(z_{1},\cdots,z_{n})]\in\mathbb{CP}^{n-1}:z_{1}^{2}+\cdots+z_{n}^{2}=0\} corresponding to the oriented tangent plane of MM at pp. Then GG is a holomorphic map from the Riemann surface MM to n1\mathbb{CP}^{n-1}. So there are many analogous results between Gauss maps of minimal surfaces and meromorphic mappings.

One of them is the Picard theorem. Through the efforts of R. Osserman [2, 3, 4], F. Xavier [5], Mo-Osserman [6], H. Fujimoto [7] finally proved that the Gauss map of a non-flat complete minimal surface in 3\mathbb{R}^{3} can omit at most 44 points in 𝕊2\mathbb{S}^{2}. H. Fujimoto [7] also gave the estimate for the number of exceptional values for Gauss map of minimal surfaces in 4\mathbb{R}^{4}. The number ”four” in 3\mathbb{R}^{3} is the best possible upper bound since a lot of examples of complete minimal surfaces whose Gauss maps miss 44 points exist [3, 8]. And the geometry interpretation of the maximal number exceptional value is given by Y. Kawakami [9] (for 3\mathbb{R}^{3}) and R. Aiyama, K. Akutagawa, S. Imagawa and Y. Kawakami [10] (for 4\mathbb{R}^{4} ). The number of exceptional values of Gauss map of minimal surfaces in n\mathbb{R}^{n} were investigated by Fujimoto [11] and M. Ru [12].

The second one is the ramification problem. In 1992, H. Fujimoto [13] studied the ramification of Gauss maps of complete regular minimal surfaces in 3\mathbb{R}^{3}. One says that g:M¯=C{}g:M\to\bar{\mathbb{C}}=C\cup\{\mathbb{\infty}\} is ramified over a point aj¯(1jq)a_{j}\in\bar{\mathbb{C}}(1\leq j\leq q) with multiplicity at least ej(1jq)e_{j}(1\leq j\leq q) if all zeros of the function g(z)ajg(z)-a_{j} have orders at least eje_{j}. If the image of gg omits aja_{j}, one will say that gg is ramified over aja_{j} with multiplicity ej=e_{j}=\infty. Specifically, H. Fujimoto obtained the following result:

Theorem 1.1.

[13] Let M be a non-flat complete minimal surface in 3\mathbb{R}^{3}. If there are q(q>4)q(q>4) distinct points a1,,aq¯a_{1},\cdots,a_{q}\in\bar{\mathbb{C}} such that the Gauss map of M is ramified over aja_{j} with multiplicity at least eje_{j} for each jj, then

j=1q(11/ej)4.\sum_{j=1}^{q}(1-1/e_{j})\leq 4.

In particular, if the Gauss map omits five distinct points, then MM must be flat. In 1993, M. Ru [14] extended the ramification result to Gauss maps of minimal surfaces in n\mathbb{R}^{n}. Afterwards, J. Lu [15], S. J. Kao [16], P. H. Ha, L. B. Phuong, P. D. Thoan, G. Dethloff [17, 18] studied the number of exceptional values and ramification of the Gauss map of complete minimal surfaces in n\mathbb{R}^{n} on annular end.

The third one is the unicity problem. H. Fujimoto [19] also investigated the uniqueness theorem for Gauss maps of minimal surfaces in n\mathbb{R}^{n}, which is analogue to the Nevanlinna unicity theorem [20] for meromorphic functions on the complex plane \mathbb{C}: Two meromorphic functions on \mathbb{C} sharing 55 distinct values must be identically equal to each other. Here, two functions g1,g2g_{1},g_{2} share value aa means g11(a)=g21(a)g_{1}^{-1}(a)=g_{2}^{-1}(a). For the Gauss map of minimal surface, H. Fujimoto proved the following theorem:

Theorem 1.2.

[19] Let MM and M^\hat{M} be two non-flat minimal surfaces in 3\mathbb{R}^{3} with their Gauss maps gg and g^\hat{g} respectively. Suppose that there is a conformal diffeomorphism Φ\Phi from MM onto M^\hat{M} and g,g^g,\hat{g} share qq distinct points a1,a2,,aqa_{1},a_{2},\cdots,a_{q}, then the following statements hold:

  • If q7q\geq 7 and either MM or M^\hat{M} is complete, then gg^Φg\equiv\hat{g}\circ\Phi.

  • If q6q\geq 6 and both MM and M^\hat{M} are complete and have finite total curvature, then gg^Φg\equiv\hat{g}\circ\Phi.

The number 77 is the best possible since H. Fujimoto construct two mutually distinct isometric complete minimal surfaces whose Gauss maps are distinct and have the same inverse images for six points. Later, H. Fujimoto [21], J. Park and M. Ru [22] gave generalizations of the unicity theorem to minimal surfaces in n(n>3)\mathbb{R}^{n}(n>3). After that, many mathematicians studied unicity theorem for Gauss maps of minimal surfaces (for more details, see [23, 24, 25, 26]).

It is quite natural to study the value distribution problem for Gauss maps of complete space-like stationary surfaces (i.e. surfaces with zero mean curvature and positive-definite metric) in nn-dimensional Lorentz space n1,1\mathbb{R}^{n-1,1}. E. Calabi [27] showed that any complete maximal space-like surface in 2,1\mathbb{R}^{2,1} has to be affine linear, that is the Gauss map must be constant. Furthermore, the ramification problem and the unicity problem of weakly complete maxfaces in 2,1\mathbb{R}^{2,1} are investigated by Y. Kawakami in [28]. Here, maxfaces are maximal (or stationary) space-like surfaces with some admissible singularities, as introduced by Umehara and Yamada [29]. In [30], the author, C, Cheng and L. Yang [30] proved the exceptional value theorem for Gauss maps of complete space-like stationary surfaces in 3,1\mathbb{R}^{3,1}:

Theorem 1.3.

Let M be a non-flat complete space-like stationary surface in 3,1\mathbb{R}^{3,1}, (ψ1,ψ2)(\psi_{1},\psi_{2}) be the Gauss map of MM, and qiq_{i} be the number of exceptional values of ψi(i=1,2)\psi_{i}(i=1,2). If neither ψ1\psi_{1} nor ψ2\psi_{2} are constant, then min{q1,q2}3\min\{q_{1},q_{2}\}\leq 3 or q1=q2=4q_{1}=q_{2}=4.

This result cannot be further improved without additional assumptions, since every minimal surface in 3\mathbb{R}^{3} is a space-like stationary surface in 3,1\mathbb{R}^{3,1}. They also considered the situation that Gauss image lies in a graph of a rational function of degree mm, which contain the degenerate cases (m1)(m\leq 1). They obtained the following result:

Theorem 1.4.

[30] Let M be a non-flat complete space-like stationary surface in 3,1\mathbb{R}^{3,1}, whose Gauss map (ψ1,ψ2)(\psi_{1},\psi_{2}) satisfies ψ2=f(ψ1)\psi_{2}=f(\psi_{1}) with ff a rational function of degree mm, then the number q1q_{1} of the exceptional values of ψ1\psi_{1} should satisfy |Ef|q1m|Ef|+3|E_{f}|\leq q_{1}\leq m-|E_{f}|+3, where Ef={z¯,f(z)=z¯}E_{f}=\{z\in\bar{\mathbb{C}},f(z)=\bar{z}\}.

In this paper, we study the ramification and the unicity problem for Gauss maps of complete space-like stationary surfaces in 3,1\mathbb{R}^{3,1}. Since every complete space-like stationary surface in 3,1\mathbb{R}^{3,1} corresponds to a complete regular minimal surface in 4\mathbb{R}^{4} which has same Gauss map, thus we can easily get Theorem 3.1 and Theorem 4.1 for ramification problem and unicity problem respectively.

Moreover, we extend the ramification result of H. Fujimoto [13] and the unicity theorem of H. Fujimoto [19] to the complete space-like stationary surfaces in 3,1\mathbb{R}^{3,1}, whose Gauss map satisfies ψ2=f(ψ1)\psi_{2}=f(\psi_{1}) where ff is a rational function with degree mm. In the case m1m\leq 1, MM is called degenerate, which contains the minimal surfaces in 3\mathbb{R}^{3}, maximal surfaces in 2,1\mathbb{R}^{2,1}, 2-degenerate space-like stationary surface, and three types of space-like stationary graphs in 3,1\mathbb{R}^{3,1} (see more details in [30]).

We first give some basic preliminaries in section 2. And in section 3, we study the ramification problem of Gauss map for space-like stationary surface in 3,1\mathbb{R}^{3,1} and get the following theorem:

Main Theorem A.

Let M3,1M\subset\mathbb{R}^{3,1} is a non-flat complete space-like stationary surface with Gauss map G=(ψ1,ψ2)G=(\psi_{1},\psi_{2}) satisfying ψ2=f(ψ1)\psi_{2}=f(\psi_{1}), where ff is a rational function of degree mm. Set Ef={z¯,f(z)=z¯}E_{f}=\{z\in\bar{\mathbb{C}},f(z)=\bar{z}\} and assume that there are q(q>m2|Ef|+3)q(q>m-2|E_{f}|+3) distinct points a1,,aq¯Efa_{1},\cdots,a_{q}\in\bar{\mathbb{C}}\setminus E_{f}, such that the Gauss map ψ=ψ1\psi=\psi_{1} of MM is ramified over aja_{j} with multiplicity at least eje_{j} for each jj, then

γ:=|Ef|+j=1q(11ej)m|Ef|+3.\gamma:=|E_{f}|+\sum_{j=1}^{q}(1-\frac{1}{e_{j}})\leq m-|E_{f}|+3. (1)

In equation (1)(\ref{mr}), if all ej(1jq)e_{j}(1\leq j\leq q) are \infty, then we can easily get the Theorem 1.4.

In section 4, we study the unicity problem for Gauss maps of complete stationary space-like surfaces in 3,1\mathbb{R}^{3,1} and get the following theorem:

Main Theorem B.

Let MM and M^\hat{M} be two complete non-flat space-like stationary surfaces in 3,1\mathbb{R}^{3,1} with Gauss maps G=(ψ1,ψ2)G=(\psi_{1},\psi_{2}) and G^=(ψ^1,ψ^2)\hat{G}=(\hat{\psi}_{1},\hat{\psi}_{2}) respectively. Suppose ψ2=f(ψ1)\psi_{2}=f(\psi_{1}) and ψ^2=f(ψ^1)\hat{\psi}_{2}=f(\hat{\psi}_{1}) with ff a rational function of degree mm and Φ:MM^\Phi:M\to\hat{M} is a conformal diffeomorphism. Assume there are qq points a1,a2,,aq¯a_{1},a_{2},\cdots,a_{q}\in\bar{\mathbb{C}} such that

ψ11(ai)=(ψ^1Φ)1(ai),i=1,,q.\psi_{1}^{-1}(a_{i})=(\hat{\psi}_{1}\circ\Phi)^{-1}(a_{i}),i=1,\cdots,q.

Then, we have ψ1ψ^1Φ\psi_{1}\equiv\hat{\psi}_{1}\circ\Phi if qm|Ef|+6q\geq m-|E_{f}|+6, where Ef={z¯,f(z)=z¯}E_{f}=\{z\in\bar{\mathbb{C}},f(z)=\bar{z}\}.

2 Preliminaries

2.1 The Gauss map of space-like stationary surfaces in 3,1\mathbb{R}^{3,1}

Denote 3,1\mathbb{R}^{3,1} be the 44-dimensional Minkowski space with the Minkowski inner product:

𝐮,𝐯=u1v1+u2v2+u3v3u4v4,\langle\mathbf{u},\mathbf{v}\rangle=u_{1}v_{1}+u_{2}v_{2}+u_{3}v_{3}-u_{4}v_{4},

where 𝐮=(u1,u2,u3,u4),𝐯=(v1,v2,v3,v4)3,1\mathbf{u}=(u_{1},u_{2},u_{3},u_{4}),\mathbf{v}=(v_{1},v_{2},v_{3},v_{4})\in\mathbb{R}^{3,1}. 𝐮3,1\mathbf{u}\in\mathbb{R}^{3,1} is space-like if 𝐮,𝐮>0\langle\mathbf{u},\mathbf{u}\rangle>0; 𝐮\mathbf{u} is time-like if 𝐮,𝐮<0\langle\mathbf{u},\mathbf{u}\rangle<0; 𝐮\mathbf{u} is called a null vector or a light-like vector if 𝐮,𝐮=0\langle\mathbf{u},\mathbf{u}\rangle=0.

Let 𝐱:M3,1\mathbf{x}:M\to\mathbb{R}^{3,1} be an oriented space-like stationary surface in the Minkowski space. That is the mean curvature vector field 𝐇\mathbf{H} of MM vanishes everywhere, and the pull-back metric ds2=d𝐱,d𝐱ds^{2}=\langle d\mathbf{x},d\mathbf{x}\rangle is positive-definite everywhere. MM is stationary if and only if the restriction of each coordinate function on MM is harmonic.

Let (u,v)(u,v) be local isothermal parameters in a neighborhood of pMp\in M, then the tangent space at pp is TpM=span{𝐱u,𝐱v}T_{p}M=\text{span}\{\mathbf{x}_{u},\mathbf{x}_{v}\}. The Gauss map of MM is defined by

G:pMTpM𝐆1,12,G:p\in M\mapsto T_{p}M\in\mathbf{G}_{1,1}^{2},

where 𝐆1,12\mathbf{G}_{1,1}^{2} is the Lorentz–Grassmann manifold consisting of all oriented space-like 2-plane in 3,1\mathbb{R}^{3,1}. Denote

Q1,1={[𝐳]3:z12+z22+z32z42=0},Q_{1,1}=\{[\mathbf{z}]\in\mathbb{CP}^{3}:z_{1}^{2}+z_{2}^{2}+z_{3}^{2}-z_{4}^{2}=0\},
Q1,1+={[𝐳]Q1,1:|z1|2+|z2|2+|z3|2|z4|2>0}.Q_{1,1}^{+}=\{[\mathbf{z}]\in Q_{1,1}:|z_{1}|^{2}+|z_{2}|^{2}+|z_{3}|^{2}-|z_{4}|^{2}>0\}.

Since 𝐱u,𝐱u=𝐱v,𝐱v>0\langle\mathbf{x}_{u},\mathbf{x}_{u}\rangle=\langle\mathbf{x}_{v},\mathbf{x}_{v}\rangle>0, and 𝐱u,𝐱v=0\langle\mathbf{x}_{u},\mathbf{x}_{v}\rangle=0, then [𝐱z]=[12(𝐱ui𝐱v)]Q1,1+[\mathbf{x}_{z}]=[\frac{1}{2}(\mathbf{x}_{u}-i\mathbf{x}_{v})]\in Q_{1,1}^{+}, where z=u+ivz=u+iv. We may identify 𝐆1,12\mathbf{G}_{1,1}^{2} with Q1,1+Q_{1,1}^{+} via the map

i:Π=span{𝐮,𝐯}𝐆1,12[𝐳]=[𝐮i𝐯]Q1,1+.i:\Pi=\text{span}\{\mathbf{u},\mathbf{v}\}\in\mathbf{G}_{1,1}^{2}\mapsto[\mathbf{z}]=[\mathbf{u}-i\mathbf{v}]\in Q_{1,1}^{+}.
Proposition 2.1.

[30] For Q1,1Q_{1,1} and Q1,1+Q_{1,1}^{+}, we have:

  1. (1)

    Q1,1Q_{1,1} is biholomorphic to ¯×¯=𝕊2×𝕊2\bar{\mathbb{C}}\times\bar{\mathbb{C}}=\mathbb{S}^{2}\times\mathbb{S}^{2}, with ¯={}\bar{\mathbb{C}}=\mathbb{C}\cup\{\infty\} the extended complex plane.

  2. (2)

    Q1,1+Q_{1,1}^{+} is biholomorphic to {(w1,w2)¯×¯:w2w¯1}\{(w_{1},w_{2})\in\bar{\mathbb{C}}\times\bar{\mathbb{C}}:w_{2}\neq\bar{w}_{1}\}, where ¯=\overline{\infty}=\infty.

The holomorphic map Ψ:Q1,1¯×¯\Psi:Q_{1,1}\to\bar{\mathbb{C}}\times\bar{\mathbb{C}} is defined by

  • Ψ([𝐳])=(z1+iz2z3+z4,z1iz2z3+z4)\Psi([\mathbf{z}])=(\frac{z_{1}+iz_{2}}{z_{3}+z_{4}},\frac{z_{1}-iz_{2}}{z_{3}+z_{4}}) whenever z3+z40z_{3}+z_{4}\neq 0;

  • Ψ([𝐳])=(z4z3z1iz2,)\Psi([\mathbf{z}])=(\frac{z_{4}-z_{3}}{z_{1}-iz_{2}},\infty) whenever z3+z4=z1+iz20 and z1iz20z_{3}+z_{4}=z_{1}+iz_{2}\equiv 0\text{ and }z_{1}-iz_{2}\neq 0;

  • Ψ([𝐳])=(,z4z3z1+iz2)\Psi([\mathbf{z}])=(\infty,\frac{z_{4}-z_{3}}{z_{1}+iz_{2}}) whenever z3+z4=z1iz20 and z1+iz20z_{3}+z_{4}=z_{1}-iz_{2}\equiv 0\text{ and }z_{1}+iz_{2}\neq 0;

  • Ψ([𝐳])=(,)\Psi([\mathbf{z}])=(\infty,\infty) whenever z3+z4=z1iz2=z1+iz20z_{3}+z_{4}=z_{1}-iz_{2}=z_{1}+iz_{2}\equiv 0.

The metric of Q1,1+Q_{1,1}^{+} in terms of w1w_{1} and w2w_{2} is :

g=Re[4dw¯1dw2(w¯1w2)2].g=\text{Re}\left[\frac{4d\bar{w}_{1}dw_{2}}{(\bar{w}_{1}-w_{2})^{2}}\right]. (2)

2.2 Weierstrass representation of space-like stationary Surfaces in 3,1\mathbb{R}^{3,1}

Denote

φ=(φ1,φ2,φ3,φ4):=(x1z,x2z,x3z,x4z)dz.\varphi=(\varphi_{1},\varphi_{2},\varphi_{3},\varphi_{4}):=\Big{(}\frac{\partial x_{1}}{\partial z},\frac{\partial x_{2}}{\partial z},\frac{\partial x_{3}}{\partial z},\frac{\partial x_{4}}{\partial z}\Big{)}dz.

Then, the harmonicity of xkx_{k} forces φk\varphi_{k} to be a holomorphic 1-form that can be globally defined on MM. [𝐱z]Q1,1+[\mathbf{x}_{z}]\in Q_{1,1}^{+} is equivalent to saying that

φ12+φ22+φ32φ42\displaystyle\varphi_{1}^{2}+\varphi_{2}^{2}+\varphi_{3}^{2}-\varphi_{4}^{2} =\displaystyle= 0,\displaystyle 0, (3)
|φ1|2+|φ2|2+|φ3|2|φ4|2\displaystyle|\varphi_{1}|^{2}+|\varphi_{2}|^{2}+|\varphi_{3}|^{2}-|\varphi_{4}|^{2} >\displaystyle> 0.\displaystyle 0. (4)

Let

(ψ1,ψ2)=Ψ(φ),dh=12(φ3+φ4).(\psi_{1},\psi_{2})=\Psi(\varphi),\quad dh=\frac{1}{2}(\varphi_{3}+\varphi_{4}).

Then, ψ1,ψ2\psi_{1},\psi_{2} are meromorphic functions on MM, and dhdh is a holomorphic 1-form on MM. If dh0dh\equiv 0, then (3) implies

φ1iφ20 or φ1+iφ20.\varphi_{1}-i\varphi_{2}\equiv 0\text{ or }\varphi_{1}+i\varphi_{2}\equiv 0.

Hence, φ=(φ1,±iφ1,φ3,φ3)\varphi=(\varphi_{1},\pm i\varphi_{1},\varphi_{3},-\varphi_{3}) and the zeros of φ1\varphi_{1} and φ3\varphi_{3} do not coincide. Otherwise, the zeros of dhdh are discrete.

Thus, the Weierstrass representation of space-like stationary surfaces in 3,1\mathbb{R}^{3,1} can be stated as follows:

Theorem 2.2.

[31] Let ψ1,ψ2\psi_{1},\psi_{2} be meromorphic functions and dhdh be a holomorphic 1-form on a Riemann surface MM. If ψ1,ψ2,dh\psi_{1},\psi_{2},dh satisfy the regularity conditions (1),(2)(1),(2) and the period condition (3)(3):

  1. (1)

    ψ1ψ2¯\psi_{1}\neq\bar{\psi_{2}} on MM, and their poles do not coincide;

  2. (2)

    The zeros of dhdh coincide with the poles of ψ1\psi_{1} or ψ2\psi_{2} with the same order;

  3. (3)

    For any closed curve CC on MM:

    Cψ1dh=Cψ2𝑑h¯,ReCdh=0=ReCψ1ψ2dh.\int_{C}\psi_{1}dh=-\overline{\int_{C}\psi_{2}dh}\quad,\quad Re\int_{C}dh=0=Re\int_{C}\psi_{1}\psi_{2}dh\ .

Then, the following equation defines a space-like stationary surface 𝐱:M3,1\mathbf{x}:M\rightarrow\mathbb{R}^{3,1}:

𝐱=2Re(ψ1+ψ2,i(ψ1ψ2),1ψ1ψ2,1+ψ1ψ2)𝑑h.\mathbf{x}=2Re\int\big{(}\psi_{1}+\psi_{2},-i(\psi_{1}-\psi_{2}),1-\psi_{1}\psi_{2},1+\psi_{1}\psi_{2}\big{)}dh\ . (5)

Conversely, every space-like stationary surface 𝐱:M3,1\mathbf{x}:M\rightarrow\mathbb{R}^{3,1} can be represented as equation (5), where dhdh, ψ1\psi_{1}, ψ2\psi_{2} satisfy the conditions (1),(2)(1),(2), and (3)(3).

The induced metric is:

ds2=𝐱z,𝐱z¯=φ,φ¯=2|ψ1ψ2¯|2|dh|2.ds^{2}=\langle\mathbf{x}_{z},\mathbf{x}_{\bar{z}}\rangle=\langle\varphi,\bar{\varphi}\rangle=2|\psi_{1}-\bar{\psi_{2}}|^{2}|dh|^{2}. (6)

Actually, as shown in [30], ψ1\psi_{1} and ψ2\psi_{2} correspond to two null vectors 𝐲,𝐲\mathbf{y},\mathbf{y}^{*} respectively, which span the normal plane of MM at the considered point. Moreover, ψ2ψ¯1\psi_{2}\neq\bar{\psi}_{1} is equivalent to saying that 𝐲𝐲\mathbf{y}\neq\mathbf{y}^{*} everywhere.

Remark 2.3.

If we let ψ11ψ2\psi_{1}\equiv-\frac{1}{\psi_{2}}, (5) yields the representation for a minimal surface in 3\mathbb{R}^{3}. If ψ11ψ2\psi_{1}\equiv\frac{1}{\psi_{2}}, then we can obtain the Weierstrass representation for a maximal surface in 2,1\mathbb{R}^{2,1}. Actually, all minimal surfaces in 3\mathbb{R}^{3} and maximal surfaces in 2,1\mathbb{R}^{2,1} are space-like stationary surfaces in 3,1\mathbb{R}^{3,1}.

Remark 2.4.

The induced action on 𝕊2\mathbb{S}^{2} of the Lorentz transformation is just the Möbius transformation on 𝕊2\mathbb{S}^{2} (i.e., a fractional linear transformation on ¯\bar{\mathbb{C}}). Since (ψ1,ψ2)=Ψ([𝐱z])(\psi_{1},\psi_{2})=\Psi([\mathbf{x}_{z}]), then the Gauss maps ψ1,ψ2\psi_{1},\psi_{2} and the height differential dhdh can be transformed as below:

ψ1aψ1+bcψ1+d,ψ2a¯ψ2+b¯c¯ψ2+d¯,dh(cψ1+d)(c¯ψ2+d¯)dh,\psi_{1}\Rightarrow\frac{a\psi_{1}+b}{c\psi_{1}+d},\quad\psi_{2}\Rightarrow\frac{\bar{a}\psi_{2}+\bar{b}}{\bar{c}\psi_{2}+\bar{d}},\quad dh\Rightarrow(c\psi_{1}+d)(\bar{c}\psi_{2}+\bar{d})dh,

where S=(abcd)SL(2,)S=\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right)\in SL(2,\mathbb{C}).

Let

φk=φk,k=1,2,3,φ4=iφ4,\varphi^{*}_{k}=\varphi_{k},k=1,2,3,\quad\varphi^{*}_{4}=i\varphi_{4},

we have

k=14(φk)2=k=14φk2φ42=0,\sum_{k=1}^{4}(\varphi^{*}_{k})^{2}=\sum_{k=1}^{4}\varphi^{2}_{k}-\varphi_{4}^{2}=0,

and

k=14|φk|2=k=14|φk|2k=13|φk|2|φ4|2>0.\sum_{k=1}^{4}|\varphi^{*}_{k}|^{2}=\sum_{k=1}^{4}|\varphi_{k}|^{2}\geq\sum_{k=1}^{3}|\varphi_{k}|^{2}-|\varphi_{4}|^{2}>0. (7)

Then

𝐱=Re(φ1,φ2,φ3,φ4)𝑑z\mathbf{x}^{*}=Re\int(\varphi^{*}_{1},\varphi^{*}_{2},\varphi^{*}_{3},\varphi^{*}_{4})dz

defines a simply-connected regular minimal surface MM^{*} in 4\mathbb{R}^{4}. Conversely, given a simply-connected regular minimal surface MM^{*} in 4\mathbb{R}^{4}, the corresponding stationary surface MM may not be space-like. Thus 𝐱𝐱\mathbf{x}\leftrightarrow\mathbf{x}^{*} gives a one-to-one correspondence between all simply-connected generalized stationary surface MM in 3,1\mathbb{R}^{3,1} and all simply-connected generalized minimal surface MM^{*} in 4\mathbb{R}^{4}.

If MM is complete, MM^{*} is also complete by the equation (7). Thus, every simply-connected complete space-like stationary surface in 3,1\mathbb{R}^{3,1} can be viewed as a simply-connected complete regular minimal surface in 4\mathbb{R}^{4}.

2.3 Space-like stationary surfaces with rational graphical Gauss image

Let MM be a space-like stationary surface in 3,1\mathbb{R}^{3,1}. If the Gauss image G(M)G(M) of MM lies in a hyperplane HAQ1,1H_{A}\subset Q_{1,1} with [A]2,1[A]\in\mathbb{CP}^{2,1}, then MM is called degenerate. If MM is 1-degenerate, then the Gauss map (ψ1,ψ2)(\psi_{1},\psi_{2}) satisfies ψ2=MS(ψ1):=aψ1+bcψ1+d\psi_{2}=M_{S}(\psi_{1}):=\frac{a\psi_{1}+b}{c\psi_{1}+d}, where S:=(abcd)SL(2,)S:=\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right)\in SL(2,\mathbb{C}). ψ1ψ2¯\psi_{1}\neq\bar{\psi_{2}} implies that ψ1\psi_{1} must omit all the points in ESE_{S}, where

ES={z¯,MS(z)=z¯}.E_{S}=\{z\in\overline{\mathbb{C}},M_{S}(z)=\bar{z}\}.

Under the conjugate similarity equivalence relation (S1conjS2 if and only if S2=±T¯S1T1S_{1}\stackrel{{\scriptstyle conj}}{{\sim}}S_{2}\text{ if and only if }S_{2}=\pm\bar{T}S_{1}T^{-1}, where T=(abcd)SL(2,)T=\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right)\in SL(2,\mathbb{C}). See section 2.4 in [30] for more details ), MM is one of the following types:

  1. 1.

    MM is a maximal surface in 2,1\mathbb{R}^{2,1}. In this case, Sconj(1001)S\stackrel{{\scriptstyle conj}}{{\sim}}\left(\begin{array}[]{cc}1&0\\ 0&1\end{array}\right) and |ES|=|E_{S}|=\infty ;

  2. 2.

    MM is a degenerate space-like stationary surface of hyperbolic type, which is congruent to entire space-like stationary graph of F:21,1F:\mathbb{R}^{2}\to\mathbb{R}^{1,1}. In this case, Sconj(eu00eu)S\stackrel{{\scriptstyle conj}}{{\sim}}\left(\begin{array}[]{cc}e^{u}&0\\ 0&e^{-u}\end{array}\right) with u(0,)u\in(0,\infty) and |ES|=2|E_{S}|=2 ;

  3. 3.

    MM is a minimal surface in 3\mathbb{R}^{3} or a degenerate space-like stationary surface of elliptic type, which can be deformed by a minimal surface in 3\mathbb{R}^{3}. In this case, Sconj(0ieiαieiα0)S\stackrel{{\scriptstyle conj}}{{\sim}}\left(\begin{array}[]{cc}0&ie^{-i\alpha}\\ ie^{i\alpha}&0\end{array}\right) with α(0,π2]\alpha\in(0,\frac{\pi}{2}] and |ES|=0|E_{S}|=0;

  4. 4.

    MM is a degenerate space-like stationary surface of parabolic type. In this case, Sconj(1101)S\stackrel{{\scriptstyle conj}}{{\sim}}\left(\begin{array}[]{cc}1&1\\ 0&1\end{array}\right) and |ES|=1|E_{S}|=1.

If MM is 2-degenerate, that is the Gauss image G(M)G(M) lies in the intersection of two linear independent hyperplanes in Q1,1Q_{1,1}. Then ψ1\psi_{1} or ψ2\psi_{2} is a constant function. By Theorem 4.3 of [30], MM has to be an entire graph F:(x1,x2)2h(x1,x2)𝐲01,1F:(x_{1},x_{2})\in\mathbb{R}^{2}\to h(x_{1},x_{2})\mathbf{y}_{0}\in\mathbb{R}^{1,1}, where h(x1,x2)h(x_{1},x_{2}) is a harmonic map and 𝐲0\mathbf{y}_{0} is a null vector. Since the Gauss curvature K0K\equiv 0, the universal covering space M~\tilde{M} of MM is conformally equivalent to the whole complex plane \mathbb{C}.

More generally, if the Gauss map (ψ1,ψ2)(\psi_{1},\psi_{2}) satisfies ψ2=f(ψ1)\psi_{2}=f(\psi_{1}) with ff a rational function of degree mm, then ψ1ψ2¯\psi_{1}\neq\bar{\psi_{2}} implies that ψ1\psi_{1} must omit all the points in EfE_{f}, where

Ef={z¯,f(z)=z¯}.E_{f}=\{z\in\overline{\mathbb{C}},f(z)=\bar{z}\}.

Then we have the following results:

Proposition 2.5.

[30] Let MM be a non-flat complete space-like stationary surface in 3,1\mathbb{R}^{3,1}. If the Gauss map (ψ1,ψ2)(\psi_{1},\psi_{2}) satisfies ψ2=f(ψ1)\psi_{2}=f(\psi_{1}) with ff a rational function of degree mm, then

  • m5;m\leq 5;

  • If m2m\geq 2, then m1|Ef|m+32m-1\leq|E_{f}|\leq\frac{m+3}{2}.

  • The number of exceptional values of ψ1\psi_{1} in ¯\bar{\mathbb{C}} cannot exceed m|Ef|+3m-|E_{f}|+3.

2.4 Metrics with negative curvature

Let Ω\Omega be a domain in \mathbb{C} of hyperbolic type which can be endowed with a Poincare´\acute{e} metric denoted by

ds2=λΩ(z)2|dz|2,ds^{2}=\lambda_{\Omega}(z)^{2}|dz|^{2},

where λΩ(z)\lambda_{\Omega}(z) is a positive C2C^{2}-function on Ω\Omega satisfying the condition ΔlogλΩ=λΩ2.\Delta log\lambda_{\Omega}=\lambda_{\Omega}^{2}. In particular, for a disc 𝔻(R)={z,|z|<R}\mathbb{D}(R)=\{z\in\mathbb{C},|z|<R\}, we have

λ𝔻(R)(z)=2RR2|z|2.\lambda_{\mathbb{D}(R)}(z)=\frac{2R}{R^{2}-|z|^{2}}.
Theorem 2.6.

[32] Let Ω\Omega be a domain in \mathbb{C} and λ\lambda be a positive C2C^{2}-function on Ω\Omega satisfying the condition Δlogλλ2\Delta log\lambda\geq\lambda^{2}. Then, for every holomorphic map f:𝔻(R)Ω,f:\mathbb{D}(R)\to\Omega,

|f(z)|λ(f(z))2RR2|z|2.|f^{\prime}(z)|\lambda(f(z))\leq\frac{2R}{R^{2}-|z|^{2}}. (8)

Especially, if f(z)=zf(z)=z, then we get

λ(z)2RR2|z|2.\lambda(z)\leq\frac{2R}{R^{2}-|z|^{2}}.

3 Ramification problem

Let MM be a space-like stationary surfaces in 3,1\mathbb{R}^{3,1} with Gauss map G=(ψ1,ψ2)G=(\psi_{1},\psi_{2}), and the induced metric (6)(\ref{phi5}) satisfies

ds2=2|ψ1ψ2¯|2|dh|22(1+|ψ1|2)(1+|ψ2|2)|dh|2=ds~2.ds^{2}=2|\psi_{1}-\bar{\psi_{2}}|^{2}|dh|^{2}\leq 2(1+|\psi_{1}|^{2})(1+|\psi_{2}|^{2})|dh|^{2}=d\tilde{s}^{2}.

Then (M,ds~2)(M,d\tilde{s}^{2}) is a complete Riemann surface if ds2ds^{2} is complete. Thus the following ramification result for Gauss map of stationary surfaces in 3,1\mathbb{R}^{3,1} can be derived from Theorem III in [33] :

Theorem 3.1.

Let M3,1M\subset\mathbb{R}^{3,1} is a non-flat complete space-like stationary surface with Gauss map G=(ψ1,ψ2)G=(\psi_{1},\psi_{2}). Then if ψ1,ψ2\psi_{1},\psi_{2} are not constant, and ψi(i=1,2)\psi_{i}(i=1,2) is ramified over aij(1jqi)a^{ij}(1\leq j\leq q_{i}) with multiplicity at least mijm_{ij}, then min{γ1,γ2}3\min\{\gamma_{1},\gamma_{2}\}\leq 3, or γ1=γ2=4\gamma_{1}=\gamma_{2}=4, where

γ1:=Σj=1q1(11m1j),γ2:=Σj=1q2(11m2j).\gamma_{1}:=\Sigma_{j=1}^{q_{1}}(1-\frac{1}{m_{1j}}),\quad\gamma_{2}:=\Sigma_{j=1}^{q_{2}}(1-\frac{1}{m_{2j}}).
Remark 3.2.

If all mij=(i=1,2,j=1,,qi)m_{ij}=\infty(i=1,2,j=1,\cdots,q_{i}), that is aija^{ij} are exceptional values of ψi\psi_{i}. Then γi=qi(i=1,2)\gamma_{i}=q_{i}(i=1,2) are the number of exceptional values, and the above theorem is a generalization of Theorem 1.21.2 in [30].

Next, we will discuss the ramification problem with condition ψ2=f(ψ1)\psi_{2}=f(\psi_{1}) where f(z)=P(z)Q(z)f(z)=\frac{P(z)}{Q(z)} is a rational function of degree mm. Throughout the proof process, we may assume MM is simply connected, otherwise we consider its universal covering space. By Koebe’s uniformization theorem, MM is conformally equivalent to the whole complex plane \mathbb{C} or to the unit disc 𝔻\mathbb{D}. For the case M=M=\mathbb{C}, γ\gamma defined in (1) satisfies γ2\gamma\leq 2 by the following theorem:

Theorem 3.3.

[34] Let ff be a non-constant meromorphic function over \mathbb{C}, if there are qq distinct points a1,a2,,aq¯a_{1},a_{2},\cdots,a_{q}\in\overline{\mathbb{C}} such that all the roots of the equation f(z)=aif(z)=a_{i} have multiplicity at least eie_{i}(ei=e_{i}=\infty if f(z)=aif(z)=a_{i} has no root). Then

i=1q(11ei)2.\sum_{i=1}^{q}(1-\frac{1}{e_{i}})\leq 2. (9)
Proposition 3.4.
  • If MM is 2-degenerate (m=0)(m=0) or 1-degenerate of hyperbolic type(m=1,|Ef|=2m=1,|E_{f}|=2), MM is an entire graph over 2\mathbb{R}^{2} and the universal covering space M~\tilde{M} of MM is conformally equivalent to the whole complex plane. Then γ2\gamma\leq 2 by Theorem 3.3.

  • If MM is 1-degenerate of elliptic type(m=1,|Ef|=0m=1,|E_{f}|=0), MM can be deformed by a minimal surface in 3\mathbb{R}^{3}, thus γ4\gamma\leq 4 by Theorem 1.1.

Thus, we just need to consider the 1-degenerate of parabolic type (m=1,|Ef|=1m=1,|E_{f}|=1) and the case 2m52\leq m\leq 5. Since EfE_{f}\neq\emptyset, without loss of generality, we can assume Ef\infty\in E_{f} under a suitable Möbius transformation. This means f()=f(\infty)=\infty, hence

m=degreeP>degreeQ=n.m=degreeP>degreeQ=n.

Set

Ef={c1,c2,,cl,}E_{f}=\{c_{1},c_{2},\cdots,c_{l},\infty\}

and let b1,,bsb_{1},\cdots,b_{s} be the zeros of Q(z)Q(z) with orders r1,,rsr_{1},\cdots,r_{s} respectively. Then

|f(z)z¯|C(Πi=1l|zci|)(Πj=1s|zbj|rj)(1+|z|2)ml2.|f(z)-\bar{z}|\leq C\left(\Pi_{i=1}^{l}|z-c_{i}|\right)\left(\Pi_{j=1}^{s}|z-b_{j}|^{-r_{j}}\right)(1+|z|^{2})^{\frac{m-l}{2}}.

Let ψ:=ψ1\psi:=\psi_{1}, the induced metric on MM is

ds2=|f(ψ)ψ¯|2|dh|2(Πi=1l|ψci|2)(Πj=1s|ψbj|2rj)(1+|ψ|2)ml|dh|2=(Πi=1l|ψci|2)(1+|ψ|2)ml|ω|2=ds~2,\begin{split}ds^{2}&=|f(\psi)-\bar{\psi}|^{2}|dh|^{2}\\ &\leq\left(\Pi_{i=1}^{l}|\psi-c_{i}|^{2}\right)\left(\Pi_{j=1}^{s}|\psi-b_{j}|^{-2r_{j}}\right)(1+|\psi|^{2})^{m-l}|dh|^{2}\\ &=\left(\Pi_{i=1}^{l}|\psi-c_{i}|^{2}\right)(1+|\psi|^{2})^{m-l}|\omega|^{2}=d\tilde{s}^{2},\end{split} (10)

then ds~2d\tilde{s}^{2} is also a complete metric on MM, where

ω=dhΠj=1s|ψbj|rj\omega=\frac{dh}{\Pi_{j=1}^{s}|\psi-b_{j}|^{r_{j}}}

is a holomorphic 1-form on MM with no zero. Since

m1|Ef|=l+1m+32,m-1\leq|E_{f}|=l+1\leq\frac{m+3}{2},

we have 0<ml20<m-l\leq 2.

For a non-zero meromorphic function ff, the divisor νf\nu_{f} of ff is defined as follows:

νf(a)={k if f has a zero of order k at ap if f has a pole of order p at a0 otherwise.\nu_{f}(a)=\begin{cases}k&\text{ if $f$ has a zero of order $k$ at $a$, }\\ -p&\text{ if $f$ has a pole of order $p$ at $a$, }\\ 0&\text{ otherwise.}\end{cases}
Lemma 3.5.

Assume that there are qq distinct points a1,,aq¯Efa_{1},\cdots,a_{q}\in\bar{\mathbb{C}}\setminus E_{f}, such that the meromorphic function ψ\psi on 𝔻(R)\mathbb{D}(R) is ramified over aja_{j} with multiplicity at least eje_{j} for each j. If l+qj=1q1ej1>ϵ>(l+q)ϵ>0l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-1>\epsilon>(l+q)\epsilon^{\prime}>0, where 0<ϵ<10<\epsilon^{\prime}<1. Set

v=(1+ψ2)(ml)p2ψΠi=1l(ψci)1ϵΠj=1q(ψaj)11ejϵv=\frac{(1+\psi^{2})^{\frac{(m-l)p}{2}}\psi^{\prime}}{\Pi_{i=1}^{l}(\psi-c_{i})^{1-\epsilon^{\prime}}\Pi_{j=1}^{q}(\psi-a_{j})^{1-\frac{1}{e}_{j}-\epsilon^{\prime}}} (11)

on 𝔻(R){z,ψ(z)=aj for some j}\mathbb{D}(R)\setminus\{z,\psi(z)=a_{j}\text{ for some j}\} and v=0v=0 on 𝔻(R){z,ψ(z)=aj for some j}\mathbb{D}(R)\cap\{z,\psi(z)=a_{j}\text{ for some j}\}. Set

p=l+qj=1q1ej1ϵml.p=\frac{l+q-\sum_{j=1}^{q}\frac{1}{e}_{j}-1-\epsilon}{m-l}.

Then vv is a continuous function in 𝔻(R)\mathbb{D}(R), and there exists a positive constant BB such that

|v|B2RR2|z|2.|v|\leq B\frac{2R}{R^{2}-|z|^{2}}. (12)
Proof.

First, we prove the continuity of vv. Obviously, vv is continuous on

𝔻(R){z,ψ(z)=aj for some 0jq}.\mathbb{D}(R)\setminus\{z,\psi(z)=a_{j}\text{ for some }0\leq j\leq q\}.

Take a point z0z_{0} with ψ(z0)=aj\psi(z_{0})=a_{j} for some jj. Now, we write

ψ(z)aj=(zz0)r0h(z),\psi(z)-a_{j}=(z-z_{0})^{r_{0}}h(z),

where r0=νψ(z0)ejr_{0}=\nu_{\psi}(z_{0})\geq e_{j} and h(z)h(z) is a meromorphic function with h(z0)0h(z_{0})\neq 0. Then

ψ=(zz0)r01(r0h(z)+(zz0)h(z)),\psi^{\prime}=(z-z_{0})^{r_{0}-1}(r_{0}h(z)+(z-z_{0})h^{\prime}(z)), (13)

that is νψ(z0)=νψ(z0)1=r01\nu_{\psi^{\prime}}(z_{0})=\nu_{\psi}(z_{0})-1=r_{0}-1. Since

νv(z0)\displaystyle\nu_{v}(z_{0}) =νψ(z0)(11ejϵ)νψ(z0)\displaystyle=\nu_{\psi^{\prime}}(z_{0})-(1-\frac{1}{e}_{j}-\epsilon^{\prime})\nu_{\psi}(z_{0})
=r01r0(11eiϵ)\displaystyle=r_{0}-1-r_{0}(1-\frac{1}{e}_{i}-\epsilon^{\prime})
=r0ei1+r0ϵ>0,\displaystyle=\frac{r_{0}}{e_{i}}-1+r_{0}\epsilon^{\prime}>0,

then limzz0v=0\lim_{z\to z_{0}}v=0. This implies that vv is continuous on 𝔻(R)\mathbb{D}(R).

Obviously, (12) holds if z𝔻(R){z,ψ(z)=aj for some j}z\in\mathbb{D}(R)\cap\{z,\psi(z)=a_{j}\text{ for some j}\}. If z𝔻(R){z,ψ(z)=aj for some j}z\in\mathbb{D}(R)\setminus\{z,\psi(z)=a_{j}\text{ for some j}\}, let Ω={a1,,aq,c1,,cl}\Omega=\mathbb{C}\setminus\{a_{1},\dots,a_{q},c_{1},\dots,c_{l}\} and λΩ\lambda_{\Omega} be the Poincare´\acute{e} metric of Ω\Omega, that is λΩ\lambda_{\Omega} satisfies ΔlogλΩ=λΩ2,\Delta log\lambda_{\Omega}=\lambda_{\Omega}^{2}, then (8) implies that

|ψ(z)|λΩ(ψ(z))2RR2|z|2.|\psi^{\prime}(z)|\lambda_{\Omega}(\psi(z))\leq\frac{2R}{R^{2}-|z|^{2}}. (14)

Moreover, λΩ\lambda_{\Omega} satisfies(see e.g. p.250 of [35])

λΩ(ψ)1|ψai|log|ψai|1 near ai,\lambda_{\Omega}(\psi)\sim\frac{1}{|\psi-a_{i}|\log|\psi-a_{i}|^{-1}}\quad\text{ near }a_{i},
λΩ(ψ)1|ψcj|log|ψcj|1 near cj,\lambda_{\Omega}(\psi)\sim\frac{1}{|\psi-c_{j}|\log|\psi-c_{j}|^{-1}}\quad\text{ near }c_{j},
λΩ(ψ)1|ψ|log|ψ| near .\lambda_{\Omega}(\psi)\sim\frac{1}{|\psi|\log|\psi|}\quad\text{ near }\infty.

Denote

u=(1+|ψ|2)(ml)p2Πi=1l|ψci|1ϵΠj=1q|ψaj|11ejϵλΩ(ψ),u=\frac{(1+|\psi|^{2})^{\frac{(m-l)p}{2}}}{\Pi_{i=1}^{l}|\psi-c_{i}|^{1-\epsilon^{\prime}}\Pi_{j=1}^{q}|\psi-a_{j}|^{1-\frac{1}{e}_{j}-\epsilon^{\prime}}\lambda_{\Omega}(\psi)},

then by a direct calculation, we get uu is bounded by a constant BB .

Thus by Theorem 2.6, we get

v=u|ψ|λΩ(ψ)B2RR2|z|2.v=u|\psi^{\prime}|\lambda_{\Omega}(\psi)\leq B\frac{2R}{R^{2}-|z|^{2}}.

Now, we begin to proof the Main Theorem A.

Proof.

If the equation (1) is wrong, then

l+1+j=1q(11ej)>ml+2,l+1+\sum_{j=1}^{q}(1-\frac{1}{e_{j}})>m-l+2,

that is

2l+qj=1q1ejm1>0.2l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-m-1>0. (15)

Take ϵ\epsilon^{\prime} with

2l+qj=1q1ejm1l+q>ϵ>2l+qj=1q1ejm1q+m,\frac{2l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-m-1}{l+q}>\epsilon^{\prime}>\frac{2l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-m-1}{q+m},

and set

p~=1p=mll+qj=1q1ej1ϵ.\tilde{p}=\frac{1}{p}=\frac{m-l}{l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-1-\epsilon}.

Then

0<p~=mll+qj=1q1ej1ϵ<mlml+2ϵ<1,0<\tilde{p}=\frac{m-l}{l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-1-\epsilon}<\frac{m-l}{m-l+2-\epsilon}<1,

and

p~1p~>ϵp~1p~\displaystyle\frac{\tilde{p}}{1-\tilde{p}}>\frac{\epsilon^{\prime}\tilde{p}}{1-\tilde{p}} =(ml)ϵl+qj=1q1ej1ϵm+l\displaystyle=\frac{(m-l)\epsilon^{\prime}}{l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-1-\epsilon-m+l}
=(ml)ϵ2l+qj=1q1ej1mϵ\displaystyle=\frac{(m-l)\epsilon^{\prime}}{2l+q-\sum_{j=1}^{q}\frac{1}{e_{j}}-1-m-\epsilon}
>(ml)ϵ(m+q)ϵϵ\displaystyle>\frac{(m-l)\epsilon^{\prime}}{(m+q)\epsilon^{\prime}-\epsilon}
=(ml)ϵ(ml)ϵ+(l+q)ϵϵ>1.\displaystyle=\frac{(m-l)\epsilon^{\prime}}{(m-l)\epsilon^{\prime}+(l+q)\epsilon^{\prime}-\epsilon}>1.

Now, we consider M=𝔻M=\mathbb{D} and

E={z𝔻,ψ(z)=0}.E=\{z\in\mathbb{D},\psi^{\prime}(z)=0\}.

Non-flatness of MM implies E𝔻E\neq\mathbb{D}.

Let ω=gdz\omega=gdz, where gg is a holomorphic function on 𝔻\mathbb{D} with no zero, and consider the following many-valued function:

η=g11p~Πi=1l(ψci)p~(1+pϵ)1p~Πj=1q(ψaj)p~(11ejϵ)1p~ψp~1p~\eta=\frac{g^{\frac{1}{1-\tilde{p}}}\Pi_{i=1}^{l}(\psi-c_{i})^{\frac{\tilde{p}(1+p-\epsilon^{\prime})}{1-\tilde{p}}}\Pi_{j=1}^{q}\left(\psi-a_{j}\right)^{\frac{\tilde{p}(1-\frac{1}{e_{j}}-\epsilon^{\prime})}{1-\tilde{p}}}}{\psi^{\prime\frac{\tilde{p}}{1-\tilde{p}}}}

on 𝔻=𝔻E\mathbb{D}^{\prime}=\mathbb{D}\setminus E.

Take an arbitrary single-value branch of η\eta, still denoted by η\eta for the matter of convenience. Let

w=F(z)=η𝑑zw=F(z)=\int\eta dz

be a holomorphic mapping from 𝔻\mathbb{D}^{\prime} to \mathbb{C}, satisfying F(0)=0,F(z)=η(z)0F(0)=0,F^{\prime}(z)=\eta(z)\neq 0. Hence there exists a holomorphic inverse mapping z=H(w)z=H(w) on a neighborhood of 0. Let 𝔻(R)={w:|w|R}\mathbb{D}(R)=\{w:|w|\leq R\} is the largest ball that HH can be defined, then R<+R<+\infty (otherwise, HH is a non-constant bounded entire function, which contracts to the Liouville theorem) and there exists a point aa on the boundary of 𝔻(R)\mathbb{D}(R), such that HH cannot be extended beyond a neighborhood of aa. Let la:={ta:0t<1}l_{a}:=\{ta:0\leq t<1\} be the straight line segment starting from 0, then H(la)H(l_{a}) must be a divergent curve in 𝔻\mathbb{D}^{\prime} as tt tends to 11.

Lemma 3.6.

There exists a point with |a|=R|a|=R such that H(la)H(l_{a}) is a divergent curve in 𝔻\mathbb{D}.

Proof.

If H(la)H(l_{a}) tends to a point z0z_{0} such that ψ(z0)=aj\psi^{\prime}(z_{0})=a_{j} for some 0jq0\leq j\leq q, then we have two cases:

Case 1: ψ(z0)aj\psi(z_{0})\neq a_{j} for all 1jq1\leq j\leq q, then νη(z0)=kp~1p~\nu_{\eta}(z_{0})=-\frac{k\tilde{p}}{1-\tilde{p}}. Since

kp~1p~>p~1p~>1,\frac{k\tilde{p}}{1-\tilde{p}}>\frac{\tilde{p}}{1-\tilde{p}}>1,

then

R\displaystyle R =la|dw|=H(la)|dwdz||dz|\displaystyle=\int_{l_{a}}|dw|=\int_{H(l_{a})}\left|\frac{dw}{dz}\right||dz|
=H(la)|η||dz|=,\displaystyle=\int_{H(l_{a})}|\eta||dz|=\infty,

which contradicts with R<R<\infty.

Case 2: ψ(z0)=aj\psi(z_{0})=a_{j} for some j{1,2,,q}j\in\{1,2,\cdots,q\} . Then νψ(z0)ej\nu_{\psi}(z_{0})\geq e_{j}, and νψ(z0)=νψ(z0)1\nu_{\psi^{\prime}}(z_{0})=\nu_{\psi}(z_{0})-1.

νη(z0)\displaystyle\nu_{\eta}(z_{0}) =p~1p~(νψ(z0)(11ejϵ)νψ(z0)+1)\displaystyle=\frac{\tilde{p}}{1-\tilde{p}}(\nu_{\psi}(z_{0})(1-\frac{1}{e_{j}}-\epsilon^{\prime})-\nu_{\psi}(z_{0})+1)
=p~1p~(1νψ(z0)(1ej+ϵ))\displaystyle=\frac{\tilde{p}}{1-\tilde{p}}(1-\nu_{\psi}(z_{0})(\frac{1}{e_{j}}+\epsilon^{\prime}))
p~1p~(1ej(1ej+ϵ))\displaystyle\leq\frac{\tilde{p}}{1-\tilde{p}}(1-e_{j}(\frac{1}{e_{j}}+\epsilon^{\prime}))
=ejϵp~1p~<1\displaystyle=-\frac{e_{j}\epsilon^{\prime}\tilde{p}}{1-\tilde{p}}<-1

then H(la)|η||dz|=+\int_{H(l_{a})}|\eta||dz|=+\infty and forces a contradiction.

Observing that

dwdz=gΠi=1l(ψci)p~(1+pϵ)Πj=1q(ψaj)p~(11ejϵ)(ψ)p~(dwdz)p~,\frac{dw}{dz}=\frac{g\Pi_{i=1}^{l}(\psi-c_{i})^{\tilde{p}(1+p-\epsilon^{\prime})}\Pi_{j=1}^{q}\left({\psi-a_{j}}\right)^{\tilde{p}(1-\frac{1}{e_{j}}-\epsilon^{\prime})}}{(\psi^{\prime})^{\tilde{p}}}\left(\frac{dw}{dz}\right)^{\tilde{p}},

then

|dzdw|2=|(ψH)|2p~|dzdw|2p~|gH|2Πi=1l|ψHci|2p~(1+pϵ)Πj=1q|ψHaj|2p~(11ejϵ),\left|\frac{dz}{dw}\right|^{2}=\frac{|(\psi\circ H)^{\prime}|^{2\tilde{p}}\left|\frac{dz}{dw}\right|^{2\tilde{p}}}{|g\circ H|^{2}\Pi_{i=1}^{l}|\psi\circ H-c_{i}|^{2\tilde{p}(1+p-\epsilon^{\prime})}\Pi_{j=1}^{q}|\psi\circ H-a_{j}|^{2\tilde{p}(1-\frac{1}{e_{j}}-\epsilon^{\prime})}},

and the pull back metric on 𝔻(R)\mathbb{D}(R) is

Hds~2\displaystyle H^{*}d\tilde{s}^{2} =((Πi=1l|ψci|2)(1+|ψ|2)ml|g|2)H|dzdw|2|dw|2\displaystyle=\bigg{(}\left(\Pi_{i=1}^{l}|\psi-c_{i}|^{2}\right)(1+|\psi|^{2})^{m-l}|g|^{2}\bigg{)}\circ H\left|\frac{dz}{dw}\right|^{2}|dw|^{2}
=(Πi=1l|ψ(w)ci|2)(1+|ψ(w)|2)ml|g(w)|2|ψ(w)|2p~|g(w)|2Πi=1l|ψ(w)ci|2p~(1+pϵ)Πj=1q|ψ(w)aj|2p~(11ejϵ)|dw|2\displaystyle=\frac{\left(\Pi_{i=1}^{l}|\psi(w)-c_{i}|^{2}\right)(1+|\psi(w)|^{2})^{m-l}|g(w)|^{2}|\psi^{\prime}(w)|^{2\tilde{p}}}{|g(w)|^{2}\Pi_{i=1}^{l}|\psi(w)-c_{i}|^{2\tilde{p}(1+p-\epsilon^{\prime})}\Pi_{j=1}^{q}|\psi(w)-a_{j}|^{2\tilde{p}(1-\frac{1}{e_{j}}-\epsilon^{\prime})}}|dw|^{2}
=((1+|ψ|2)(ml)p2ψΠi=1l(ψci)1ϵΠj=1q(ψaj)11ejϵ)2p~|dw|2\displaystyle=\left(\frac{(1+|\psi|^{2})^{\frac{(m-l)p}{2}}\psi^{\prime}}{\Pi_{i=1}^{l}(\psi-c_{i})^{1-\epsilon^{\prime}}\Pi_{j=1}^{q}(\psi-a_{j})^{1-\frac{1}{e}_{j}-\epsilon^{\prime}}}\right)^{2\tilde{p}}|dw|^{2}
(B2RR2|w|2)2p~.\displaystyle\leq\left(B\frac{2R}{R^{2}-|w|^{2}}\right)^{2\tilde{p}}.

Thus

H(la)𝑑s~=laH𝑑s~0R(2RR2|r|2)p~𝑑r.\int_{H(l_{a})}d\tilde{s}=\int_{l_{a}}H^{*}d\tilde{s}\leq\int_{0}^{R}\left(\frac{2R}{R^{2}-|r|^{2}}\right)^{\tilde{p}}dr. (16)

The integral (16) is convergent since 0<p~<10<\tilde{p}<1, which contradicts to the completeness of (M,ds~2)(M,d\tilde{s}^{2}). ∎

Remark 3.7.

In equation (1)(\ref{mr}),

  • If m=0,|Ef|=1m=0,|E_{f}|=1(MM is 2-degenerate), then γ2\gamma\leq 2;

  • If m=1,|Ef|=2m=1,|E_{f}|=2(MM is 1-degenerate of hyperbolic type), then γ2\gamma\leq 2;

  • If m=1,|Ef|=0m=1,|E_{f}|=0(MM is 1-degenerate of elliptic type), then γ4\gamma\leq 4.

These conclusions are agree with the Proposition 3.4.

Thus, by combining the above proof process and Proposition 3.4, we get Main Theorem A.

4 Unicity problem

Similarly as in Section 3, we first consider Riemann surface MM and metric

ds~2=2(1+|ψ1|2)(1+|ψ2|2)|dh|2.d\tilde{s}^{2}=2(1+|\psi_{1}|^{2})(1+|\psi_{2}|^{2})|dh|^{2}.

Thus we can easily get the following theorem by Theorem 1.21.2 in [23].

Theorem 4.1.

Let MM and M^\hat{M} be two complete non-flat space-like stationary surfaces in 3,1\mathbb{R}^{3,1} with Gauss maps G=(ψ1,ψ2)G=(\psi_{1},\psi_{2}) and G^=(ψ^1,ψ^2)\hat{G}=(\hat{\psi}_{1},\hat{\psi}_{2}) respectively. Assume Φ:MM^\Phi:M\to\hat{M} is a conformal diffeomorphism and ψ1,ψ2,ψ^1,ψ^2\psi_{1},\psi_{2},\hat{\psi}_{1},\hat{\psi}_{2} are not constant. If for each i(i=1,2)i(i=1,2), ψi\psi_{i} and ψ^i\hat{\psi}_{i} share pi>4p_{i}>4 distinct values and ψiψ^iΦ\psi_{i}\neq\hat{\psi}_{i}\circ\Phi, then we have min{p1,p2}6.\min\{p_{1},p_{2}\}\leq 6. In particular, if p17p_{1}\geq 7 and p27p_{2}\geq 7, then either ψ1ψ^1Φ\psi_{1}\equiv\hat{\psi}_{1}\circ\Phi or ψ2ψ^2Φ\psi_{2}\equiv\hat{\psi}_{2}\circ\Phi, or both hold.

Remark 4.2.

Since every minimal surface in 3\mathbb{R}^{3} is also a space-like stationary surface in 3,1\mathbb{R}^{3,1} with Gauss map G=(ψ1,ψ2)G=(\psi_{1},\psi_{2}) and ψ2=1ψ1\psi_{2}=-\frac{1}{\psi_{1}}, then theorem 1.2 and the optimality in 3\mathbb{R}^{3} tell us the above result is also optimal.

Next, we will consider the space-like stationary surface with Gauss map G=(ψ1,ψ2)G=(\psi_{1},\psi_{2}) satisfying ψ2=f(ψ1)\psi_{2}=f(\psi_{1}), where ff is a rational function with degree mm. Now, assume Φ:MM^\Phi:M\to\hat{M} be the conformal diffeomorphism, and (ψ1,ψ2)(\psi_{1},\psi_{2}) and (ψ1^,ψ2^)(\hat{\psi_{1}},\hat{\psi_{2}}) be the Gauss map of MM and M^\hat{M} respectively, where ψ2=f(ψ1)\psi_{2}=f(\psi_{1}), ψ2^=f(ψ1^)\hat{\psi_{2}}=f(\hat{\psi_{1}}). For brevity, we denote ψ1,ψ^1Φ\psi_{1},\hat{\psi}_{1}\circ\Phi by ψ,ψ^\psi,\hat{\psi}.

If M,M^M,\hat{M} are 2-degenerate (m=0)(m=0) or 1-degenerate of hyperbolic type(m=1,|Ef|=2m=1,|E_{f}|=2), then they are entire graphs over 2\mathbb{R}^{2} and their universal covering spaces are conformally equivalent to the whole complex plane. Then the Main Theorem B is followed by the Nevanlinna unicity theorem [20] for meromorphic functions on the complex plane \mathbb{C}.

If M,M^M,\hat{M} are 1-degenerate of elliptic type(m=1,|Ef|=0m=1,|E_{f}|=0), then they can be deformed by minimal surfaces in 3\mathbb{R}^{3}, and the Main Theorem B is followed by Theorem 1.2.

Thus, we only need to consider the proof when m2m\geq 2 and M,M^M,\hat{M} are 1-degenerate of parabolic type(m=1,|Ef|=1m=1,|E_{f}|=1). For each α,β¯\alpha,\beta\in\bar{\mathbb{C}}, the chordal distance[13] between α,β\alpha,\beta is

|α,β|=|αβ|1+|α|21+|β|2,|\alpha,\beta|=\frac{|\alpha-\beta|}{\sqrt{1+|\alpha|^{2}}\sqrt{1+|\beta|^{2}}},

if α,β\alpha\neq\infty,\beta\neq\infty, and |α,β|=|β,α|=11+|α|2|\alpha,\beta|=|\beta,\alpha|=\frac{1}{\sqrt{1+|\alpha|^{2}}}, if β=\beta=\infty.

Lemma 4.3.

[19] Let ff and f^\hat{f} be two mutually distinct non-constant meromorphic functions on a Riemann surface MM and qq distinct points a1,a2,,aq(q>4)a_{1},a_{2},\cdots,a_{q}(q>4). Assume that f1(aj)=f^1(aj)(1jq).f^{-1}(a_{j})=\hat{f}^{-1}(a_{j})(1\leq j\leq q). For a0>0a_{0}>0 and ϵ\epsilon with q4>qϵ>0q-4>q\epsilon>0, set

λ:=(Πi=1q|f,ai|log(a0|f,ai|2))1+ϵ,\lambda:=\left(\Pi_{i=1}^{q}|f,a_{i}|\log\left(\frac{a_{0}}{|f,a_{i}|^{2}}\right)\right)^{-1+\epsilon},
λ^:=(Πi=1q|f^,ai|log(a0|f^,ai|2))1+ϵ,\hat{\lambda}:=\left(\Pi_{i=1}^{q}|\hat{f},a_{i}|\log\left(\frac{a_{0}}{|\hat{f},a_{i}|^{2}}\right)\right)^{-1+\epsilon},
dτ2:=|f,f^|2λλ^f1+|f|2f^1+|f^|2d\tau^{2}:=|f,\hat{f}|^{2}\lambda\hat{\lambda}\frac{f^{\prime}}{1+|f|^{2}}\frac{\hat{f}^{\prime}}{1+|\hat{f}|^{2}} (17)

outside the E:=i=1qf1(ai)E:=\cup_{i=1}^{q}f^{-1}(a_{i}) and dτ2=0d\tau^{2}=0 on EE. Then, for a suitably chosen a0a_{0}, dτ2d\tau^{2} is continuous on MM, and has strictly negative curvature on the set {dτ20}\{d\tau^{2}\neq 0\}.

Lemma 4.4.

[19] Let ff and f^\hat{f} be two mutually distinct non-constant meromorphic functions on a Riemann surface MM satisfies the same assumption as in Lemma 4.3, then for the metric dτ2d\tau^{2} defined by (17), there is a constant C>0C>0 such that

dτ2C4R2(R2z2)2|dz|2.d\tau^{2}\leq C\frac{4R^{2}}{(R^{2}-z^{2})^{2}}|dz|^{2}.

Combine the completeness of MM and inequality (10), the conformal metric ds2ds^{2} can be stated as follows:

ds2=Πi=1l|ψci|(1+|ψ|2)ml|h|2|dz|2,ds^{2}=\Pi_{i=1}^{l}|\psi-c_{i}|(1+|\psi|^{2})^{m-l}|h|^{2}|dz|^{2},

where hh is a holomorphic function with no zeros on a simply connected open set UU.

Since there exists a local non-zero holomorphic function ζ\zeta on UU such that ds2=|ζ|2Φ(ds^2)ds^{2}=|\zeta|^{2}\Phi^{*}(d\hat{s}^{2}), thus

Πi=1l|ψci|2(1+|ψ|2)ml|h|2|dz|2=|ζ|2Πi=1l|ψ^ci|2(1+|ψ^|2)ml|h^|2|dz|2,\Pi_{i=1}^{l}|\psi-c_{i}|^{2}(1+|\psi|^{2})^{m-l}|h|^{2}|dz|^{2}=|\zeta|^{2}\Pi_{i=1}^{l}|\hat{\psi}-c_{i}|^{2}(1+|\hat{\psi}|^{2})^{m-l}|\hat{h}|^{2}|dz|^{2},
Πi=1l|ψci|(1+|ψ|2)ml2|h|=|ζ|Πi=1l|ψ^ci|(1+|ψ^|2)ml2|h^|.\Rightarrow\Pi_{i=1}^{l}|\psi-c_{i}|(1+|\psi|^{2})^{\frac{m-l}{2}}|h|=|\zeta|\Pi_{i=1}^{l}|\hat{\psi}-c_{i}|(1+|\hat{\psi}|^{2})^{\frac{m-l}{2}}|\hat{h}|.

Let kk be a non-zero holomorphic function on UU satisfies k2=hh^ζk^{2}=h\hat{h}\zeta, then

ds2=|k|2Πi=1l(|ψci||ψ^ci|)(1+|ψ|2)ml2(1+|ψ^|2)ml2|dz|2.ds^{2}=|k|^{2}\Pi_{i=1}^{l}(|\psi-c_{i}||\hat{\psi}-c_{i}|)(1+|\psi|^{2})^{\frac{m-l}{2}}(1+|\hat{\psi}|^{2})^{\frac{m-l}{2}}|dz|^{2}.

Assume there are qq distinct points a1,,aq¯a_{1},\cdots,a_{q}\in\bar{\mathbb{C}} such that ψ1(ai)=ψ^1(ai)\psi^{-1}(a_{i})=\hat{\psi}^{-1}(a_{i}) for all 1iq1\leq i\leq q. Since ψ1(Ef)=ψ^1(Ef)=\psi^{-1}(E_{f})=\hat{\psi}^{-1}(E_{f})=\emptyset, then Ef{a1,,aq}E_{f}\subset\{a_{1},\cdots,a_{q}\}. We may assume aq=Efa_{q}=\infty\in E_{f}.

Since qm|Ef|+6=ml+5q\geq m-|E_{f}|+6=m-l+5, then qm+l4>0q-m+l-4>0. Take a positive real number δ\delta with

qm+l4q>δ>min{qm+l4ml+q,q42m+2lq}\frac{q-m+l-4}{q}>\delta>\min\left\{\frac{q-m+l-4}{m-l+q},\frac{q-4-2m+2l}{q}\right\}

and set

p=mlq4qδ<1.p=\frac{m-l}{q-4-q\delta}<1.

Then

p1p>1,δp1p>1.\frac{p}{1-p}>1,\quad\frac{\delta p}{1-p}>1.

If ψψ^\psi\not\equiv\hat{\psi}, consider the function

η~=k11pΠi=1l((ψci)(ψ^ci))12(1p)Πj=1q1((ψaj)(ψ^aj))p(1δ)2(1p)((ψψ^)2ψψ^Πj=1q1(1+|aj|2)1δ)p2(1p)\tilde{\eta}=k^{\frac{1}{1-p}}\frac{\Pi_{i=1}^{l}\left((\psi-c_{i})(\hat{\psi}-c_{i})\right)^{\frac{1}{2(1-p)}}\Pi_{j=1}^{q-1}\left((\psi-a_{j})(\hat{\psi}-a_{j})\right)^{\frac{p(1-\delta)}{2(1-p)}}}{\left((\psi-\hat{\psi})^{2}\psi^{\prime}\hat{\psi}^{\prime}\Pi_{j=1}^{q-1}(1+|a_{j}|^{2})^{1-\delta}\right)^{\frac{p}{2(1-p)}}} (18)

on M=MEM^{\prime}=M\setminus E, where E={zM,ψ(z)ψ^(z)=0, or ψ(z)=ψ^(z)}.E=\{z\in M,\psi^{\prime}(z)\hat{\psi}^{\prime}(z)=0,\text{ or }\psi(z)=\hat{\psi}(z)\}.

Denote

w=F~(z)=η~𝑑z,w=\tilde{F}(z)=\int\tilde{\eta}dz, (19)

by similar method in the proof of Main Theorem A, we can find a neighborhood UU of point 0 and a positive number R, such that the inverse map H~=(F~|U)1:𝔻(R)UM\tilde{H}=(\tilde{F}|_{U})^{-1}:\mathbb{D}(R)\to U\subset M^{\prime} is a holomorphic map. Since dσ2=|η~|2|dz|2d\sigma^{2}=|\tilde{\eta}|^{2}|dz|^{2} has strictly negative curvature on MM^{\prime} by Lemma 4.3, then R<R<\infty. And there exists a point aa with |a|=R|a|=R, such that for the line segment

la={ta,0t<1},l_{a}=\{ta,0\leq t<1\},

the image γ~=H~(la)\tilde{\gamma}=\tilde{H}(l_{a}) is a divergent curve in MM^{\prime} as tt tends to 11.

Lemma 4.5.

There exists a point with |a|=R|a|=R such that H~(la)\tilde{H}(l_{a}) is a divergent curve in MM.

Proof.

It is suffices to consider that H~(la)\tilde{H}(l_{a}) tends to a point z0Ez_{0}\in E.

Case 1: If ψ(z0)=ψ^(z0)\psi(z_{0})=\hat{\psi}(z_{0}) and there exists some aj,a_{j}, such that ψ(z0)=ψ(z0)^=aj\psi(z_{0})=\hat{\psi(z_{0})}=a_{j}, then

νη~(z0)=p1p((1δ)(νψ(z0)+νψ^(z0))2min{νψ(z0),νψ^(z0)}+νψ(z0)+νψ^(z0)2)δp1p(νψ(z0)+νψ^(z0))<δp1p.\begin{split}\nu_{\tilde{\eta}}(z_{0})&=\frac{p}{1-p}\left((1-\delta)(\nu_{\psi}(z_{0})+\nu_{\hat{\psi}}(z_{0}))-2\min\{\nu_{\psi}(z_{0}),\nu_{\hat{\psi}}(z_{0})\}+\nu_{\psi}(z_{0})+\nu_{\hat{\psi}}(z_{0})-2\right)\\ &\leq-\frac{\delta p}{1-p}(\nu_{\psi}(z_{0})+\nu_{\hat{\psi}}(z_{0}))\\ &<-\frac{\delta p}{1-p}.\end{split}

Otherwise,

νη~(z0)=p1p(2min{νψ(z0),νψ^(z0)})2δp1p.\nu_{\tilde{\eta}}(z_{0})=-\frac{p}{1-p}(2\min\{\nu_{\psi}(z_{0}),\nu_{\hat{\psi}}(z_{0})\})\leq-\frac{2\delta p}{1-p}.

Since δp1p>1\frac{\delta p}{1-p}>1, then

R\displaystyle R =la|dw|=γ~|dwdz||dz|\displaystyle=\int_{l_{a}}|dw|=\int_{\tilde{\gamma}}\left|\frac{dw}{dz}\right||dz|
=γ~|η~||dz|=,\displaystyle=\int_{\tilde{\gamma}}|\tilde{\eta}||dz|=\infty,

which contradicts with R<R<\infty.

Case 2: H~(la)\tilde{H}(l_{a}) tends to a point z0z_{0} such thatψ(z0)ψ^(z0)=0\psi^{\prime}(z_{0})\hat{\psi^{\prime}}(z_{0})=0, we can easily get νη~(z0)<p1p<1\nu_{\tilde{\eta}}(z_{0})<-\frac{p}{1-p}<-1. This also contradicts that RR is finite. ∎

Combine with (18) and (19), we obtain

|dwdz|=|k|Πi=1l(|ψci||ψ^ci|)12Πj=1q1(|ψaj||ψ^aj|)p(1δ)2(|ψψ^|2|ψ||ψ^|Πj=1q1(1+|aj|2)1δ)p2|dwdz|p.\left|\frac{dw}{dz}\right|=\frac{|k|\Pi_{i=1}^{l}\left(|\psi-c_{i}||\hat{\psi}-c_{i}|\right)^{\frac{1}{2}}\Pi_{j=1}^{q-1}\left(|\psi-a_{j}||\hat{\psi}-a_{j}|\right)^{\frac{p(1-\delta)}{2}}}{\left(|\psi-\hat{\psi}|^{2}|\psi^{\prime}||\hat{\psi}^{\prime}|\Pi_{j=1}^{q-1}(1+|a_{j}|^{2})^{1-\delta}\right)^{\frac{p}{2}}}\left|\frac{dw}{dz}\right|^{p}.

Set g(w)=ψ(H~(w)),g^(w)=ψ^(H~(w))g(w)=\psi(\tilde{H}(w)),\hat{g}(w)=\hat{\psi}(\tilde{H}(w)), since g=ψdzdw,g^=ψ^dzdwg^{\prime}=\psi^{\prime}\frac{dz}{dw},\hat{g}^{\prime}=\hat{\psi}^{\prime}\frac{dz}{dw}, then

|dzdw|2=(|gg^|2|g||g^|Πj=1q1(1+|aj|2)1δ)p|kH~|2Πi=1l(|gci||g^ci|)Πj=1q1(|gaj||g^aj|)p(1δ)\left|\frac{dz}{dw}\right|^{2}=\frac{\left(|g-\hat{g}|^{2}|g^{\prime}||\hat{g}^{\prime}|\Pi_{j=1}^{q-1}(1+|a_{j}|^{2})^{1-\delta}\right)^{p}}{|k\circ\tilde{H}|^{2}\Pi_{i=1}^{l}\left(|g-c_{i}||\hat{g}-c_{i}|\right)\Pi_{j=1}^{q-1}\left(|g-a_{j}||\hat{g}-a_{j}|\right)^{p(1-\delta)}}

Therefore, the induced metric of 𝔻(R)\mathbb{D}(R) from MM by H~\tilde{H} is

H~ds2\displaystyle\tilde{H}^{*}ds^{2} =|kH~|2Πi=1l(|gci||g^ci|)(1+|g|2)ml2(1+|g^|2)ml2|dzdw|2|dw|2\displaystyle=|k\circ\tilde{H}|^{2}\Pi_{i=1}^{l}(|g-c_{i}||\hat{g}-c_{i}|)(1+|g|^{2})^{\frac{m-l}{2}}(1+|\hat{g}|^{2})^{\frac{m-l}{2}}\left|\frac{dz}{dw}\right|^{2}|dw|^{2}
=((1+|g|2)ml2p(1+|g^|2)ml2p|gg^|2|g||g^|Πj=1q1(1+|aj|2)1δΠj=1q1|gaj|1δ|g^aj|1δ)p|dw|2\displaystyle=\left(\frac{(1+|g|^{2})^{\frac{m-l}{2p}}(1+|\hat{g}|^{2})^{\frac{m-l}{2p}}|g-\hat{g}|^{2}|g^{\prime}||\hat{g}^{\prime}|\Pi_{j=1}^{q-1}(1+|a_{j}|^{2})^{1-\delta}}{\Pi_{j=1}^{q-1}|g-a_{j}|^{1-\delta}|\hat{g}-a_{j}|^{1-\delta}}\right)^{p}|dw|^{2}
=(μ2Πi=1q(|g,ai||g^,ai|)ϵ(Πi=1qloga0|g,ai|2loga0|g^,ai|2)1ϵ)p|dw|2,\displaystyle=\left(\mu^{2}\Pi_{i=1}^{q}(|g,a_{i}||\hat{g},a_{i}|)^{\epsilon}\left(\Pi_{i=1}^{q}\log\frac{a_{0}}{|g,a_{i}|^{2}}\log\frac{a_{0}}{|\hat{g},a_{i}|^{2}}\right)^{1-\epsilon}\right)^{p}|dw|^{2},

where μ\mu is the function with dτ2=μ2|dw|2d\tau^{2}=\mu^{2}|dw|^{2}. On the other hand, for a given ϵ\epsilon, it holds that

limx0xϵlog1ϵa0x2<.\lim_{x\to 0}x^{\epsilon}\log^{1-\epsilon}\frac{a_{0}}{x^{2}}<\infty.

Thus there exists a constant C1C_{1} such that

H~ds2C1μ2p|dw|2.\tilde{H}^{*}ds^{2}\leq C_{1}\mu^{2p}|dw|^{2}.

By Lemma 4.4,

γ~𝑑s=laH~𝑑s<Cla(RR2|w|2)p|dw|<,\int_{\tilde{\gamma}}ds=\int_{{l_{a}}}\tilde{H}^{*}ds<C\int_{l_{a}}\left(\frac{R}{R^{2}-|w|^{2}}\right)^{p}|dw|<\infty,

which contradicts with the completeness of MM. We have necessarily ψ=ψ^\psi=\hat{\psi}, and the proof of Main Theorem B is completed.

Remark 4.6.

In this context, we cannot uniformly discuss whether the results are optimal or not, because values of mm and |Ef||E_{f}| vary. Although for each fixed mm, the value of |Ef||E_{f}| may also have several situations.

References

  • \bibcommenthead
  • Chern [1965] Chern Ss. Minimal surfaces in an Euclidean space of NN dimensions. In: Differential and Combinatorial Topology (A Symposium in Honor of Marston Morse). Princeton Univ. Press, Princeton, NJ; 1965. p. 187–198.
  • Osserman [1959] Osserman R. Proof of a conjecture of Nirenberg. Comm Pure Appl Math. 1959;12:229–232. 10.1002/cpa.3160120203.
  • Osserman [1961] Osserman R. Minimal surfaces in the large. Comment Math Helv. 1961;35:65–76. 10.1007/BF02567006.
  • Osserman [1964] Osserman R. Global properties of minimal surfaces in E3E^{3} and EnE^{n}. Ann of Math (2). 1964;80:340–364. 10.2307/1970396.
  • Xavier [1981] Xavier F. The Gauss map of a complete nonflat minimal surface cannot omit 77 points of the sphere. Ann of Math (2). 1981;113(1):211–214. 10.2307/1971139.
  • Mo and Osserman [1990] Mo X, Osserman R. On the Gauss map and total curvature of complete minimal surfaces and an extension of Fujimoto’s theorem. J Differential Geom. 1990;31(2):343–355.
  • Fujimoto [1988] Fujimoto H. On the number of exceptional values of the Gauss maps of minimal surfaces. J Math Soc Japan. 1988;40(2):235–247. 10.2969/jmsj/04020235.
  • Osserman [1986] Osserman R. A survey of minimal surfaces. 2nd ed. Dover Publications, Inc., New York; 1986.
  • Kawakami [2013] Kawakami Y. On the maximal number of exceptional values of Gauss maps for various classes of surfaces. Math Z. 2013;274(3-4):1249–1260. 10.1007/s00209-012-1115-8.
  • Aiyama et al. [2017] Aiyama R, Akutagawa K, Imagawa S, Kawakami Y. Remarks on the Gauss images of complete minimal surfaces in Euclidean four-space. Ann Mat Pura Appl (4). 2017;196(5):1863–1875. 10.1007/s10231-017-0643-6.
  • Fujimoto [1990] Fujimoto H. Modified defect relations for the Gauss map of minimal surfaces. II. J Differential Geom. 1990;31(2):365–385.
  • Ru [1991] Ru M. On the Gauss map of minimal surfaces immersed in 𝐑n{\bf R}^{n}. J Differential Geom. 1991;34(2):411–423.
  • Fujimoto [1992] Fujimoto H. On the Gauss curvature of minimal surfaces. J Math Soc Japan. 1992;44(3):427–439. 10.2969/jmsj/04430427.
  • Ru [1993] Ru M. Gauss map of minimal surfaces with ramification. Trans Amer Math Soc. 1993;339(2):751–764. 10.2307/2154296.
  • Jin and Ru [2007] Jin L, Ru M. Values of Gauss maps of complete minimal surfaces in m\mathbb{R}^{m} on annular ends. Trans Amer Math Soc. 2007;359(4):1547–1553. 10.1090/S0002-9947-06-04126-2.
  • Kao [1991] Kao SJ. On values of Gauss maps of complete minimal surfaces on annular ends. Math Ann. 1991;291(2):315–318. 10.1007/BF01445210.
  • Dethloff and Ha [2014] Dethloff G, Ha PH. Ramification of the Gauss map of complete minimal surfaces in 3\mathbb{R}^{3} and 4\mathbb{R}^{4} on annular ends. Ann Fac Sci Toulouse Math (6). 2014;23(4):829–846. 10.5802/afst.1426.
  • Ha et al. [2016] Ha PH, Phuong LB, Thoan PD. Ramification of the Gauss map and the total curvature of a complete minimal surface. Topology Appl. 2016;199:32–48. 10.1016/j.topol.2015.11.010.
  • Fujimoto [1993] Fujimoto H. Unicity theorems for the Gauss maps of complete minimal surfaces. J Math Soc Japan. 1993;45(3):481–487. 10.2969/jmsj/04530481.
  • Nevanlinna [1926] Nevanlinna R. Einige Eindeutigkeitssätze in der Theorie der Meromorphen Funktionen. Acta Math. 1926;48(3-4):367–391. 10.1007/BF02565342.
  • Fujimoto [1993] Fujimoto H. Unicity theorems for the Gauss maps of complete minimal surfaces. II. Kodai Math J. 1993;16(3):335–354. 10.2996/kmj/1138039844.
  • Park and Ru [2017] Park J, Ru M. Unicity results for Gauss maps of minimal surfaces immersed in m\mathbb{R}^{m}. J Geom. 2017;108(2):481–499. 10.1007/s00022-016-0353-z.
  • Ha and Kawakami [2018] Ha PH, Kawakami Y. A note on a unicity theorem for the Gauss maps of complete minimal surfaces in Euclidean four-space. Canad Math Bull. 2018;61(2):292–300. 10.4153/CMB-2017-015-0.
  • Ru and Ugur [2017] Ru M, Ugur G. Uniqueness results for algebraic and holomorphic curves into n()\mathbb{P}^{n}(\mathbb{C}). Internat J Math. 2017;28(9):1740003, 27. 10.1142/S0129167X17400031.
  • Jin and Ru [2007] Jin L, Ru M. Algebraic curves and the Gauss map of algebraic minimal surfaces. Differential Geom Appl. 2007;25(6):701–712. 10.1016/j.difgeo.2007.06.014.
  • Ha [2020] Ha PH. Gaussian curvature and unicity problem of Gauss maps of various classes of surfaces. Nagoya Math J. 2020;240:275–297. 10.1017/nmj.2019.5.
  • Calabi [1970] Calabi E. Examples of Bernstein problems for some nonlinear equations. In: Global Analysis (Proc. Sympos. Pure Math., Vol. XV, Berkeley, Calif., 1968). Amer. Math. Soc., Providence, R.I.; 1970. p. 223–230.
  • Kawakami [2015] Kawakami Y. Function-theoretic properties for the Gauss maps of various classes of surfaces. Canad J Math. 2015;67(6):1411–1434. 10.4153/CJM-2015-008-5.
  • Umehara and Yamada [2006] Umehara M, Yamada K. Maximal surfaces with singularities in Minkowski space. Hokkaido Math J. 2006;35(1):13–40. 10.14492/hokmj/1285766302.
  • Ou et al. [2024] Ou L, Cheng C, Yang L. On complete space-like stationary surfaces in 4-dimensional Minkowski space with graphical Gauss image. Math Z. 2024;306(1):Paper No. 16, 40. 10.1007/s00209-023-03408-1.
  • Ma et al. [2013] Ma X, Wang C, Wang P. Global geometry and topology of spacelike stationary surfaces in the 4-dimensional Lorentz space. Adv Math. 2013;249:311–347. 10.1016/j.aim.2013.09.013.
  • Ahlfors [1973] Ahlfors LV. Conformal invariants: topics in geometric function theory. McGraw-Hill Series in Higher Mathematics. McGraw-Hill Book Co., New York-Düsseldorf-Johannesburg; 1973.
  • Fujimoto [1989] Fujimoto H. Modified defect relations for the Gauss map of minimal surfaces. J Differential Geom. 1989;29(2):245–262.
  • Robinson [1939] Robinson RM. A generalization of Picard’s and related theorems. Duke Math J. 1939;5(1):118–132. 10.1215/S0012-7094-39-00513-2.
  • Nevanlinna [1970] Nevanlinna R. Analytic functions. Die Grundlehren der mathematischen Wissenschaften, Band 162. Springer-Verlag, New York-Berlin; 1970. Translated from the second German edition by Phillip Emig.