Paper The following article is Open access

Affine spheres and finite gap solutions of Tzitzèica equation

and

Published 21 November 2018 © 2018 The Author(s). Published by IOP Publishing Ltd
, , Citation J Inoguchi and S Udagawa 2018 J. Phys. Commun. 2 115020 DOI 10.1088/2399-6528/aaeaa0

2399-6528/2/11/115020

Abstract

The purpose of the present paper is to give an explicit form of the finite gap solutions to the Tzitzèica equation (2D Toda equation of type ${{\rm{A}}}_{2}^{(2)}$) in terms of Riemann theta function. We give explicit expressions of proper affiene spheres derived from finite gap solutions to the Tzitzèica equation.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

This paper concerns with the following 2 dimensional Toda lattice (2DTL) of type ${{\rm{A}}}_{2}^{(2)}$:

Equation (1.1)

Nowadays, this 2DTL is called the Tzitzèica equation named after a Romanian mathematician G. Tzitzèica.

In the context of soliton theory, Tzitzèica equation was discovered by Bullough and Dodd [1], Z̆iber and S̆abat [2] independently (see also Mikhailov [3], p. 74).

In addition, Mikhailov [3] pointed out that Tzitzèica equation is obtained as a reduction of periodic 2DTL

of period 3 under the condition ${u}_{3}=0$, ${u}_{1}=-{u}_{2}=u$. This reduction corresponds to the affine root system ${{\rm{A}}}_{2}^{(2)}$ (or the reduced root system ${\mathrm{BC}}_{1}$) [4]. Note that this reduction was already known in projective differential geometry, see [5].

Early year in the twenty century, Tzitzèica [6] studied non-degenerate surfaces, that is, surfaces with non-degenerate second fundamental form in the Euclidean 3-space E3 with the property that $K/{\bar{d}}^{4}$ is constant over the surface, where K is the Euclidean Gaussian curvature function and $\overline{d}$ is the Euclidean distance function from the origin of E3 to each point of the surface. Tzitzèica noticed that $K/{\bar{d}}^{4}$ is invariant under equiaffine transformations of E3.

Tzitzèica's observation initiated affine differential geometry of surfaces. The surfaces discovered by Tzitzèica are now refereed as proper affine sphere due to the fact that a proper affine sphere is the set of points where the affine distance from the origin is non-zero constant. In fact the affine distance function d from the origin satisfies $K/{\bar{d}}^{4}=1/{d}^{4}$.

On every proper affine sphere, there exists a unique (up to sign) semi-Riemannian metric invariant under equiaffine transformations of E3. Such a metric is called the Blashcke metric. Note that the Blaschke metric of a proper affine sphere is conformal to the Euclidean second fundamental form. There are two classes of proper affine spheres and they are said to be indefinite or definite according as the non-degenerate affine metric is indefinite or definite, respectively. In this paper, we study indefinite proper affine spheres.

The Gauss-Codazzi equations for indefinite proper affine spheres are described by the Tzitzèica equation, where $h={e}^{u}\ {dx}\ {dt}$ is Blaschke metric on M and the affine mean curvature is normalized to −1. Tzitzèica equation arises not only in affine differential geometry but also in many other realms of mathematical physics and differential geometry. For instance elliptic versions of Tzitzèica equation are integrability conditions of the following three kinds of surfaces: (1) Lagrangian minimal surfaces in complex projective plane (cf. [7]), (2) Lagrangian minimal surfaces in complex hyperbolic plane [8] and (3) affine spheres with positive definite Blaschke metric [9]. These elliptic versions are closely related to certain Kähler-Einstein metrics [10].

For the geometry of surfaces in 3-dimensional spaces using the soliton theory we refer to [11] and [12]. More generally, for the close relation between minimal surfaces (nonlinear sigma models) in symmetric spaces and 2DTL, for example, see [13, 14] and [15].

Some explicit solutions of Tzitzèica equation (1.1) are known. For example, the hexenhut and the Jonas Kelch are realized using the solution of $u=0$ and $u=\mathrm{log}(1-3/(2{\cosh }^{2}(\sqrt{3}(x+t)/2))$, respectively. Moreover, the finite gap solutions of (1.1) may be described by Cherdantsev and Sharipov in [16] as follows :

Equation (1.2)

where θ is the Prym-theta function for some compact Riemann surface of genus $2g$ and the constant C is determined by some spetral data. They also gave the Baker-Akhiezer function in terms of the spectral data (see [16]). On the other hand, we may expect the existence of the solution of (1.1) in terms of the Jacobi elliptic functions because the elliptic version of (1.1) is solved in terms of the Jacobi elliptic function by Castro-Urbano (see [17]).

In this paper, first of all, we construct the solution of (1.1) in terms of the Jacobi elliptic functions. However, it is not clear what kind of the spectral data produces the solution in terms of the Jacobi elliptic functions as a special case of the finite gap solutions. Therefore, we explicitly give the spectral data for the solution in terms of the Jacobi elliptic functions. Moreover, we can describe the finite gap solutions more explicitly in terms of Riemann theta function. Precisely speaking, we can give the constant C in (1.2) explicitly (see (5.8), (5.9) and theorem 5.5). Moreover, we can give Abelian differentials of second kind on the spectral curve explicitly (see (5.7)) and consequently we obtain a real frame for indefinite proper affine spheres in terms of the Baker-Akhiezer function.

For our purpose stated above, first of all, in section 4, we deduce the solution of the Tzitzèica equation and gives a Blaschke immersion of indefinite proper affine spheres in terms of the Jacobi elliptic functions. Therefore, the well known elliptic function theory hides in the background. We employ the elliptic function theory and reconstruct the whole theory for solutions of the Tzitzèica equation in terms of the Riemann theta function (see (4.17)), and for Blaschke immersions of indefinite proper affine spheres in terms of the meromorphic function $\hat{{\rm{\Psi }}}$ which is a solution of the Schrödinger equation ${\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}={e}^{u}\ \hat{{\rm{\Psi }}}$ (see theorem 4.9). These reconstructions give a nice story for a general case of the corresponding problems. Since the elliptic curve used in section 4 is the Prym variety of the spectral curve of genus 2, we must consider the spectral curve of genus $2g$ and the g-dimensional Prym variety for the general case. The standard argument in the integrable systems and the result in section 4 help us to reconstruct the solutions of the problems for the general case. In section 5, we get the solutions of the Tzitzèica equation and give expressions of the immersions of indefinite proper affine spheres of finite type from the points of views of the algebro-geometric approach to the integrable systems.

2. Indefinite proper affine spheres

2.1.  Fundamental equations

We start with description of affine spheres in equiaffine differential geometry. For more details we refer to the textbook [18] by Nomizu and Sasaki.

Let ${{\bf{R}}}^{3}$ be a Cartesian 3-space. We denote by $\det :{{\bf{R}}}^{3}\times {{\bf{R}}}^{3}\times {{\bf{R}}}^{3}\longrightarrow {\bf{R}}$ the determinant function, which defines a volume element of ${{\bf{R}}}^{3}$. Then the triplet ${{\bf{A}}}^{3}=({{\bf{R}}}^{3},D,\det )$ is an equiaffine 3-space, i.e., it satisfies $D\ \det \,=\,0$, where D is the canonical covariant differentiation for ${{\bf{R}}}^{3}$. An immersion $\psi :M\longrightarrow {{\bf{A}}}^{3}$ of an oriented 2-manifold is said to be an affine immersion if $(M,\psi )$ admits a complementary subbundle ${ \mathcal N }$ of the tangent bundle TM of M in the pull-back bundle ${\psi }^{* }(T{{\bf{A}}}^{3})$, that is, if we have ${\psi }_{* }(T{{\bf{A}}}^{3})={TM}\oplus { \mathcal N }$. A local section ξ of ${ \mathcal N }$ is called a transversal vector field of $(M,\psi )$. In the following, we assume that there is a global transversal vector field ξ to $(M,\psi )$ so that ${ \mathcal N }={\cup }_{p\in M}{\bf{R}}{\xi }_{p}$. We then call a triplet $(M,\psi ,\xi )$ an affine immersion.

For any vector fields X and Y on M we have the Gauss formula

We easily verify that ${\rm{\nabla }}$ defines a torsion free linear connection on M and h is a symmetric tensor field on M. The symmetric tensor field h is called the affine fundamental form derived from ξ. We also have the Weingarten formula

The endomorphism field S of TM is called the affine shape operator with respect to ξ. The 1-form ${ \mathcal T }$ is called the transversal connection form. For example, an immersion $\psi :M\longrightarrow {{\bf{A}}}^{3}\setminus \{0\}$ with the property that the position vector field ψ is transversal to M is an affine immersion and in particular it is called centro-affine immersion. For an affine immersion $(M,\psi ,\xi )$, we may define a volume element ϑ on M by $\vartheta (X,Y)=\det (X,Y,\xi )$ for any vector fields X and Y on M. We address here some known facts about affine immersions.

  • For an affine immersion, the rank of the affine fundamental form is independent of the choice of the transversal vector field.
  • The triplet $(M,{\rm{\nabla }},\vartheta )$ is equiaffine, i.e., ${\rm{\nabla }}\vartheta =0$ if and only if ${ \mathcal T }=0$.
  • Let $\psi :M\longrightarrow {{\bf{A}}}^{3}$ be an affine immersion with non-degenerate affine fundamental form h. Then there is a unique (up to sign) transversal vector field ξ such that ${ \mathcal T }=0$ and the volume element of the semi-Riemannian surface $(M,h)$ coincides with ϑ. This ξ is called the Blaschke normal vector filed and the pair $(\psi ,\xi )$ is called a Blaschke immersion. The affine fundamental form h with respect to the Blaschke normal vector field is traditionally called the Blaschke metric of $(M,\psi )$. For a Blaschke immersion $(M,\psi ,\xi )$ we have the following fundamental equations.

Here R is the curvature tensor field of the connection ${\rm{\nabla }}$. The Codazzi equation implies that $C:= {\rm{\nabla }}h$ is a symmetric covariant tensor field. We call the C cubic form of $(M,\psi ,\xi )$.

2.2.  Affine spheres

Now let $(M,\psi ,\xi )$ be a Blaschke immersion with indefinite Blaschke metric h. Then we can take an asymptotic coordinate system $(x,t)$ on a simply connected domain ${\bf{D}}\subset M$ with respect to the Lorentz conformal structure determined by h. Namely, the Blaschke metric h is represented as $h={e}^{u}\ {dx}\ {dt}$ with respect to $(x,t)$. It then follows from Fact 2 and Fact 3 that the cubic form C is represented as $C={{Adx}}^{3}+{{Bdt}}^{3}$.

Definition 2.1. A Blaschke immersion $(M,\psi ,\xi )$ is said to be an affine sphere if $S=k\ I$ for some constant k. Moreover, a affine sphere $(M,\psi ,\xi )$ is said to be proper if $k\ne 0$ and improper if $k=0$.

  • Let $(M,\psi ,\xi )$ be a proper affine sphere. Then all the Blaschke normals meet in one point.

2.3.  Tzitz`eica equation

In the following, we consider proper affine spheres with indefinite Blaschke metric h. In particular, for such proper affine spheres we have $\xi =-H\psi $, where $H=\tfrac{1}{2}\mathrm{trace}\ S$ is the affine mean curvature and it is a negative constant. By scaling the affine metric, without loss of generality, we may assume that $H=-1$, hence we may take $\xi =\psi $. Note that $(M,\psi ,\xi )$ is centro-affine.

Set ${\partial }_{x}=\tfrac{\partial }{\partial x}$, ${\partial }_{t}=\tfrac{\partial }{\partial t}$, then the induced connection ${\rm{\nabla }}$ is given by

Set ${\psi }_{t}={\psi }_{* }({\partial }_{t})$ and ${\psi }_{x}={\psi }_{* }({\partial }_{x})$ then the Gauss-Weingarten equations are given by

Equation (2.1)

Here we introduce a matrix valued function (called a framing) ${ \mathcal F }=({\psi }_{t}\ {\psi }_{x}\ \psi )$. Since ψ is a Blaschke immersion, we see that $\det { \mathcal F }={e}^{u}\ne 0$. We rewrite (2.1) in the following way [19, 20]:

Equation (2.2)

The compatibility condition of the system (2.2) is given by

Equation (2.3)

The first equation is the Gauss equation and the last two equations are Coddazi equations. Throughout this paper we assume that indefinite proper affine spheres are weakly regular, that is, ${AB}\ne 0$ [20]. Then the Coddazi equation means that $A=A(x)$ and $B=B(t)$. Therefore, choosing the coordinate system $(x,t)$ we may assume that the cubic form is represented as $C={{dx}}^{3}+{{dt}}^{3}$ from the first step. In this case, we have $A=B=1$ and (2.3) is called the Tzitzèica equation [6].

The compatibility condition (2.3) is invariant under the replacements

Equation (2.4)

respectively. We denote by ${ \mathcal F }(\nu )$ the solution to the system

Using λ with ${\lambda }^{3}=\nu $ and setting

we obtain a new framing $F(\lambda )$ with

where

Equation (2.5)

Thus for any indefinite proper afine sphere $\psi :M\to {{\bf{A}}}^{3}$, there exists a map $F(\lambda ):{\bf{D}}\times {{\bf{R}}}^{* }\to {SL}(3,{\bf{R}})$ satisfying (2.5). We call $F(\lambda )$ a coordinate extended framing of an indefinite proper affine sphere $(M,\psi )$. Coordinate extended framings are examples of 'extended framings' which will be introduced in later section (definition 3.2).

Before closing this section, we state the assumption on u throughout the paper. We assume that $u(x,t)=u(t,x)$ with respect to the coordinate system $(x,t)$ so that the cubic form is given by $C={{dx}}^{3}+{{dt}}^{3}$. Of course, this assumption is independent of introducing the non-zero parameter ν.

3. Twisted loop algebras and twisted loop groups

3.1.  6-symmetric spaces

Let $G={SL}(3,{\bf{R}})$ be a real special linear group of degree 3 and ${\mathfrak{g}}={\mathfrak{sl}}(3,{\bf{R}})$ its Lie algebra. We denote by ${{\mathfrak{g}}}^{{\bf{C}}}={\mathfrak{sl}}(3,{\bf{C}})$ the complexification of ${\mathfrak{g}}$. Define two automorphisms σ and γ of ${{\mathfrak{g}}}^{{\bf{C}}}$ by

where $\omega =\exp (2\pi \sqrt{-1}/3)$. We then see that $\tau =\sigma \gamma =\gamma \sigma $ is an automorphism of order 6 [20]. Let $\epsilon =-{\omega }^{2}$ be the sixth root of unity. Denote by ${{\mathfrak{g}}}_{j}^{{\bf{C}}}$ be the ${\epsilon }^{j}$-eigenspace of τ, where $j=0$, 1, 2, 3, 4, 5. Set ${{\mathfrak{g}}}_{j}={{\mathfrak{g}}}_{j}^{{\bf{C}}}\cap {\mathfrak{g}}$. For example, we have ${{\mathfrak{g}}}_{0}$ consists of the marices of the form

which is a Lie subalgebra of ${\mathfrak{g}}$ and is isomorphic to ${\mathfrak{o}}(1,1)$ in ${\mathfrak{sl}}(3,{\bf{R}})$. Let K be a subgroup of G, which consists of the matrices of the form

and is isomorphic to ${SO}(1,1)\subset {SL}(3,{\bf{R}})$. We denote by ${{SO}}^{0}(1,1)$ the identity component of ${SO}(1,1)$. Then ${{SO}}^{0}(1,1)$ is isomorphic to $\exp ({{\mathfrak{g}}}_{0})=: {K}_{+}$. We denote by ${SO}(2,1)$ the Lorentzian group with Lie algebra ${{\mathfrak{g}}}_{0}\oplus {{\mathfrak{g}}}_{2}\oplus {{\mathfrak{g}}}_{4}$ and its identity component by ${{SO}}^{0}(2,1)$. We then easily see the following fact.

Proposition 3.1. The automorphism $\tau $ of order 6 gives a semi-Riemannian 6-symmetric space structure on ${SL}(3,{\bf{R}})/{{SO}}^{0}(1,1)$. The involution $\sigma $ gives ${SL}(3,{\bf{R}})/{{SO}}^{0}(2,1)$ the outer semi-Riemannian symmetric space structure. Moreoevr, we have the homogeneous projection $\pi :{SL}(3,{\bf{R}})/{{SO}}^{0}(1,1)\longrightarrow {SL}(3,{\bf{R}})/{{SO}}^{0}(2,1)$.

3.2.  Twisted loop algebras

Let $\hat{\sigma }$ be an involution on ${{\mathfrak{g}}}^{{\bf{C}}}$ defined by

We then have $\hat{\sigma }({{\mathfrak{g}}}_{j})={{\mathfrak{g}}}_{-j}$. We consider a twisted loop algebra

where $| | \xi | | ={\sum }_{n\in {\bf{Z}}}| | {\xi }_{n}| | $ for $\xi (\lambda )={\sum }_{n\in {\bf{Z}}}{\xi }_{n}{\lambda }^{n}$ and $| | {\xi }_{n}| | ={\max }_{j\in \{\mathrm{1,2,3}\}}{\sum }_{i=1}^{3}| {({\xi }_{n})}_{{ij}}| $. Then, ${\rm{\Lambda }}{{\mathfrak{g}}}_{\tau }^{{\bf{C}}}$ is a Banach Lie algebra (see [21, 22]). Set

where $[j]\in {\bf{Z}}/6{\bf{Z}}$.

We decompose each element $\xi \in {\rm{\Lambda }}{{\mathfrak{g}}}_{\tau }$ as $\xi ={\xi }_{+}+{\xi }_{0}+{\xi }_{-}$, where ${\xi }_{+}(\lambda )={\sum }_{j\gt 0}{\xi }_{j}{\lambda }^{j},{\xi }_{-}(\lambda )={\sum }_{j\lt 0}{\xi }_{j}{\lambda }^{j}$. Next, we extend the involution $\hat{\sigma }$ to ${\rm{\Lambda }}{{\mathfrak{g}}}_{\tau }$ by

We may write ξ as

Equation (3.1)

We now define the following Banach Lie subalgebras of ${\rm{\Lambda }}{{\mathfrak{g}}}_{\tau }$ :

Equation (3.2)

It then follows from (3.1) that

Equation (3.3)

which is a vector space decomposition into Banach Lie subalgebras.

For a positive integer $d\equiv 1(\mathrm{mod}\ 6)$, we define the vector subspace ${{\rm{\Lambda }}}_{d}^{\hat{\sigma }}$ of ${{\rm{\Lambda }}}^{\hat{\sigma }}{{\mathfrak{g}}}_{\tau }$ by

Equation (3.4)

3.3.  Twisted loop groups

Let ${\rm{\Lambda }}{G}_{\tau }$ be a twisted loop group defined by

whose Lie algebra is ${\rm{\Lambda }}{{\mathfrak{g}}}_{\tau }$. Here we denote the Lie group automorphism corresponding to τ by the same letter. Define subgroup ${{\rm{\Lambda }}}^{\hat{\sigma }}{G}_{\tau }$ of ${\rm{\Lambda }}{G}_{\tau }$ by

Consider a map $g:{\bf{D}}\subset {{\bf{R}}}^{2}\ni (x,t)\,\longmapsto \,g(x,t)\in {\rm{\Lambda }}{G}_{\tau }$. We extend the involution $\hat{\sigma }$ to the above map by the rule

Equation (3.5)

Define the mapping spaces ${\rm{\Lambda }}{G}_{\tau }({\bf{D}})$ and ${{\rm{\Lambda }}}^{{\hat{\sigma }}^{* }}{G}_{\tau }({\bf{D}})$ by

Equation (3.6)

Equation (3.7)

3.4.  Affines spheres of finite type

Now we return to indefinite proper affine spheres. Let $\psi :{\bf{D}}\subset {{\bf{R}}}^{2}\to {{\bf{A}}}^{3}$ be an indefinite proper affine sphere parametrized by global asymptotic coordinates $(x,t)$ as before. The coordinate extended framing $F(\lambda )$ can be extended analytically on ${{\bf{C}}}^{* }$. The (extended) map $F(\lambda )$ is uniquely determined by the values on ${S}^{1}\subset {{\bf{C}}}^{* }$ (cf. [20], p. 234). Hence $F(\lambda )$ is regarded as a map into ${\rm{\Lambda }}{G}_{\tau }$. In addition, since we assumed that $u(x,t)=u(t,x)$, it follows from (2.5) that ${\hat{\sigma }}^{* }(U(\lambda ))=V(\lambda )$. Therefore, the coordinate extended framing $F(\lambda )$ can be considered as a map $F\in {{\rm{\Lambda }}}^{{\hat{\sigma }}^{* }}{G}_{\tau }({\bf{D}})$. Moreover, since

we see that the 3rd column vector of the coordinate extended framing $F(\lambda ){| }_{\lambda =1}$ gives a Blaschke immersion $\psi :{\bf{D}}\longrightarrow {{\bf{A}}}^{3}$. Here we give the following definition (compare with [20, 23]).

Definition 3.2. An element $F\in {{\rm{\Lambda }}}^{{\hat{\sigma }}^{* }}{G}_{\tau }({\bf{D}})$ which satisfies ${F}^{-1}{dF}=\lambda {\alpha }_{1}{dt}+{\alpha }_{0}+{\lambda }^{-1}{\alpha }_{-1}{dx}$ with ${\alpha }_{-1}={\hat{\sigma }}^{* }({\alpha }_{1})$ is said to be an extended framing for indefinite proper affine sphere.

Remark 3.3. One can check that every extended framing $F\in {{\rm{\Lambda }}}^{{\hat{\sigma }}^{* }}{G}_{\tau }(D)$ induces a harmonic map (nonlinear σ-model) $F\cdot {K}_{+}:{\bf{D}}\to G/{K}_{+}$ (see proposition 3.1, [24]).

Next we introduce the following notion for affine spheres.

Definition 3.4. An indefinite proper affine sphere is said to be of finite type if its extended framing a is obtained from the following differential equation under certain initial condition:

where $d\equiv 1\ ({\rm{mod}}\ 6)$.

One can see that α satisfies Maurer-Cartan equation and hence there is a solution $a\in {{\rm{\Lambda }}}^{{\hat{\sigma }}^{* }}{G}_{\tau }({\bf{D}})$ to the equation ${a}^{-1}{da}=\alpha $. Moreover this a is an extended framing for affine spheres.

(hexenhut).

Example 3.5 We present a trivial solution $u=0$ of Tzitzèica equation and corresponding affine sphere. Take a matrix $A=\left(\begin{array}{ccc}0 & 0 & 1\\ 1 & 0 & 0\\ 0 & 1 & 0\end{array}\right)$. Note that $\hat{\sigma }(A)={A}^{-1}$. Then we easily see that $a(x,t)=\exp (t\lambda A+x{\lambda }^{-1}{A}^{-1})$ is an extended framing of an indefinite proper affine sphere ${\psi }_{0}$ of finite type. After some elementary calculation we obtain

Hence the resulting surface ${\psi }_{0}$ is affine congruent to the hexenhut $Z({X}^{2}+{Y}^{2})=1$ in ${{\bf{A}}}^{3}(X,Y,Z)$ (see Figure 1). The hexenhut ${\psi }_{0}$ corresponds to the trivial solution $u=0$ to the Tzitzèica equation [25].

Figure 1.

Figure 1. Hexenhut.

Standard image High-resolution image

Remark 3.6. One can establish the Symes method (also called AKS-scheme) for constructing indefinite proper affine sphere of finite type. This scheme was established in a separate publication [26].

4. Construction of Blaschke immersions of affine spheres

4.1. Blaschke immersions in terms of elliptic functions

In this section, we give a solution of Tzitzèica equation (1.1) in terms of elliptic functions and represent affine spheres in terms of elliptic functions. Although these formulas have been known in some literature, our purpose here is to represent them in terms of Riemann theta function. This achievement gives a nice story for the description of the solutions in terms of the spectral curves of higher genus and Prym-theta functions.

Introducing new coordinates $\hat{x}$, $\hat{y}$ by

we have $4\tfrac{\partial }{\partial x}\tfrac{\partial }{\partial t}=\tfrac{{\partial }^{2}}{\partial {\hat{x}}^{2}}-\tfrac{{\partial }^{2}}{\partial {\hat{y}}^{2}}$.

We now assume that a solution u depends only on the parameter $\hat{y}$ and write $u=u(\hat{y})$. We suppose that an initial condition ${e}^{u(0)}=\tfrac{\ \alpha \ }{2},{u}_{\hat{y}}(0)=0$, where α is a real number with $\alpha \gt 2$. If we set $Y(\hat{y})={e}^{u(\hat{y})}$, then the equation (1.1) may rewritten as

where

If we set $Y={\zeta }_{2}-({\zeta }_{2}-{\zeta }_{3}){X}^{2}$ and ${p}^{2}=\tfrac{{\zeta }_{3}-{\zeta }_{2}}{{\zeta }_{1}-{\zeta }_{2}}$ we obtain

hence, setting $X(\hat{y})=\widetilde{X}(\sqrt{2}\sqrt{{\zeta }_{2}-{\zeta }_{1}}\ \hat{y})$ and $\hat{z}=\sqrt{2}\sqrt{{\zeta }_{2}-{\zeta }_{1}}\ \hat{y}$, we have

Then, we see that $\widetilde{X}(\hat{z})=\mathrm{sn}\ (\hat{z},p)$, where $\mathrm{sn}\ (\hat{z},p)$ is the Jacobi $\mathrm{sn}$-function with modulus p. Therefore, we obtain a solution of (1.1) as follows.

Equation (4.1)

The function $Y(\hat{y})$ can be extended as a function of complex variables. Extending $\hat{y}$ as a complex valued function, we see that $Y(\hat{y})$ is a doubly-periodic function with the periods $2{\omega }^{0},2{\omega }^{{\prime} }{}^{0}$, where

Equation (4.2)

and $K(p)$ is the complete elliptic integral of 1st kind, ${K}^{{\prime} }(p)=K({p}^{{\prime} })$ and ${p}^{{\prime} }=\sqrt{1-{p}^{2}}$. The Y can be essentially represented by Weierstrass $\wp $-function. In fact, if we set $P(z)=P(\sqrt{2}\sqrt{-1}\hat{y})=Y(\hat{y})-\tfrac{\ a\ }{3}$ then we have

where ${\eta }_{j}={\zeta }_{j}-\tfrac{\ a\ }{3}$, $j=1$, 2, 3 with the property ${\eta }_{1}+{\eta }_{2}+{\eta }_{3}=0$. The solution of this equation is given by Weierstrass $\wp $-function $\wp (z)$, which is related to the Jacobi $\mathrm{sn}$-function by

where ${k}^{2}=\tfrac{{\eta }_{1}-{\eta }_{3}}{{\eta }_{1}-{\eta }_{2}}=\tfrac{{\zeta }_{1}-{\zeta }_{3}}{{\zeta }_{1}-{\zeta }_{2}}=1-{p}^{2}={({p}^{{\prime} })}^{2}$, hence we have $k={p}^{{\prime} }$ and ${k}^{{\prime} }=p$. As it is well-known, $\wp $-function is doubly-periodic with periods $2\omega (k),2{\omega }^{{\prime} }(k)$, where

Since we know that $k={p}^{{\prime} }$ and ${k}^{{\prime} }=p$, the above equation yields the following.

Equation (4.3)

We see that

Equation (4.4)

where we used the well-known relations

Remark 4.1.  $Y(\hat{y})={\zeta }_{2}+\displaystyle \frac{{\zeta }_{1}-{\zeta }_{2}}{{\mathrm{sn}}^{2}\ (\sqrt{2}\sqrt{{\zeta }_{2}-{\zeta }_{1}}\ \hat{y},p)}$ is also a solution of (1.1). The Weierstrass $\wp (z)$-function is defined by

The sum of the right hand side is absolutely and uniformly convergent on a compact subset of ${\bf{C}}$. It is an elliptic function.

On the other hand, rewriting (2.1) by the new coordinates $\hat{x}$, $\hat{y}$ and differentiating the rewritten equation one more time with respect to $\hat{x}$ we have ${\psi }_{\hat{x}\hat{x}\hat{x}}=2a{\psi }_{\hat{x}}+2\psi $. Let ${\mu }_{1}$, ${\mu }_{2}$ and ${\mu }_{3}$ be the solutions of the characteristic equation ${\mu }^{3}-2a\mu -2=0$, that is,

Equation (4.5)

We then may write $\psi (\hat{x},\hat{y})=({e}^{{\mu }_{1}\hat{x}}{C}_{1}(\hat{y}),\ {e}^{{\mu }_{2}\hat{x}}{C}_{2}(\hat{y}),\ {e}^{{\mu }_{3}\hat{x}}{C}_{3}(\hat{y}))$. Substituting this expression into the rewritten equation of (2.1), in particular, into ${\psi }_{\hat{x}\hat{y}}=\tfrac{1}{2}{u}_{\hat{y}}{\psi }_{\hat{y}}-{e}^{u}{\psi }_{\hat{y}}$, we obtain $(2{e}^{u}+{\mu }_{j}^{2}-2a){C}_{j}^{{\prime} }(\hat{y})={u}_{\hat{y}}{e}^{u}{C}_{j}(\hat{y})$. Using the relations

which, together with $\det { \mathcal F }={e}^{u}$, yield the explicit parametrization

Equation (4.6)

where c1, c2 and c3 are non-zero real numbers with

This is a rotational surface as an orbit of the group ${SO}(1,1)$ of hyperbolic rotations (Lorentz boosts) up to affine transformations. When α tends to 2, the solution converges to a trivial solution $u=0$.

Remark 4.2. Analogously, we may get a solution of (1.1) which depends only on the variable $\hat{x}$ as follows.

The corresponding affine sphere is obtained as follows.

Therefore, this is a rotational surface as a orbit of the rotational group SO(2) of elliptic type up to affine transformations. When α tends to 2, the above solution converges to the 1-soliton solution

of Tzitzèica equation. The rotational surface corresponding to the 1-soliton is called the Jonas Kelch (Figure 2, see also [25]).

Figure 2.

Figure 2. Jonas Kelch.

Standard image High-resolution image

4.2. 

In the rest of the section 4, we represent the formulas (4.1) and (4.6) in terms of the Riemann theta function.

We consider an elliptic curve ${ \mathcal C }$ defined by ${B}^{2}=4(\eta -{\eta }_{1})(\eta -{\eta }_{2})(\eta -{\eta }_{3})$. The curve is parametrized by $(B,\eta )=({\wp }^{{\prime} }(z),\wp (z))$. We may choose a cycle $\{{a}_{1},{b}_{1}\}$ of ${ \mathcal C }$ so that the following holds :

Equation (4.7)

If we take ${w}_{1}=\displaystyle \frac{\pi \sqrt{-1}}{\omega (k)}\displaystyle \frac{d\eta }{B}$, then we easily see that

Then, the elliptic curve has a Riemann period matrix $(2\pi \sqrt{-1}\ {\rm{\Pi }}(k))$. Setting $\widetilde{B}=\sqrt{2}\sqrt{-1}B,\eta =\zeta -\displaystyle \frac{\ a\ }{3}$, we obtain ${\widetilde{B}}^{2}=-8(\zeta -{\zeta }_{1})(\zeta -{\zeta }_{2})(\zeta -{\zeta }_{3})$, which is denoted by ${{ \mathcal C }}^{0}$. We may choose a cycle $\{{a}_{1}^{0},{b}_{1}^{0}\}$ of ${{ \mathcal C }}^{0}$ by ${a}_{1}^{0}={b}_{1}$ and ${b}_{1}^{0}=-{a}_{1}$, so that the following relations hold:

Equation (4.8)

If we take ${w}_{1}^{0}=\tfrac{\pi \sqrt{-1}}{{\omega }^{0}}\tfrac{d\zeta }{\widetilde{B}}$, then we see that

Equation (4.9)

Thus, the curve ${{ \mathcal C }}^{0}$ has a Riemann period matrix $(2\pi \sqrt{-1}\ {\rm{\Pi }}(p))$. Let ${\zeta }_{w}$ be the Weierstrass ζ-function defined by

It is a meromorphic function on ${\bf{C}}$ and a odd function, but not an elliptic function. In fact, we have the following properties :

Equation (4.10)

where $\omega =\omega (k)$ and ${\omega }^{{\prime} }={\omega }^{{\prime} }(k)$. In the sequel, we keep on using this notation when there is no confusion. The last equation in (4.10) follows from the 2nd and 3rd equations in (4.10) and from the fact ${\int }_{{\rm{\Gamma }}}{\zeta }_{w}(z)\ {dz}=2\pi \sqrt{-1}$, where Γ is a fundamental parallelogram which contains a pole of ${\zeta }_{w}(z)$ inside. Let ${{\rm{\Omega }}}_{\infty }^{0}$ be an Abelian differential of 2nd kind defined by

Equation (4.11)

Set ${{\bf{U}}}^{0}={\int }_{{b}_{1}^{0}}{{\rm{\Omega }}}_{\infty }^{0}$. It follows from (4.10) and ${b}_{1}^{0}=-{a}_{1}$ that

hence we obtain

Equation (4.12)

Similarly, we see that ${\int }_{{a}_{1}^{0}}{{\rm{\Omega }}}_{\infty }^{0}=0$. Therefore, ${{\rm{\Omega }}}_{\infty }^{0}$ is the normalized Abelian differential of 2nd kind.

Remark 4.3. Alternatively, since we have ${{\rm{\Omega }}}_{\infty }^{0}=\left(-\tfrac{\ {z}^{-2}\ }{\sqrt{2}\sqrt{-1}}+O(1)\right)\ {dz}$, it follows from the reciprocity laws for differentials of 1st and 2nd kind that

4.3. Riemann theta function and Jacobi theta functions

For ${\rm{\Pi }}={\rm{\Pi }}(p)$, $z\in {\bf{C}}$, set $\tau =\tfrac{{\rm{\Pi }}}{2\pi \sqrt{-1}},v=\tfrac{z}{2\pi \sqrt{-1}}$. The Riemann theta function $\theta (z;{\rm{\Pi }})$ for elliptic curve with the period matrix $(2\pi \sqrt{-1}\ {\rm{\Pi }})$ is defined by

which absolutely and uniformly converges on a compact subset of ${\bf{C}}$. The Jacobi theta functions are described in terms of the Riemann theta function as follows (see [27] and [28]).

Equation (4.13)

We then see that ${\theta }_{3}(v+\tfrac{1}{2})={\theta }_{0}(v),{\theta }_{2}(v+\tfrac{1}{2})=-{\theta }_{1}(v)$ (see [28]). Although the Jacobi theta functions are not elliptic functions, there are some relations between the Jacobi theta functions and the Jacobi elliptic functions as follows.

Equation (4.14)

Moreover, the following beautiful formula is known

Equation (4.15)

where $E(p)={\int }_{0}^{K(p)}{\mathrm{dn}}^{2}(u,p)\ {du}$ is the complete elliptic integral of 2nd kind. We now calculate

Equation (4.16)

where we have used (4.2) and ${p}^{2}=\tfrac{{\zeta }_{3}-{\zeta }_{2}}{{\zeta }_{1}-{\zeta }_{2}}$. If we set $C={\zeta }_{1}+\tfrac{E(p)}{K(p)}({\zeta }_{2}-{\zeta }_{1})$ then we have

which is a solution of (1.1) (see (4.1)). On the other hand, for ${\bf{e}}=\pi \sqrt{-1}$ we see that

Thus, we obtain the following formula.

Equation (4.17)

where $C={\zeta }_{1}+\tfrac{\ E(p)\ }{K(p)}({\zeta }_{2}-{\zeta }_{1})$ and ${\bf{e}}=\pi \sqrt{-1}$.

Lemma 4.4. The $C$ in (4.17) is given by

Equation (4.18)

Proof. First of all, we have from the definition of $E(p)$ that $C={\zeta }_{2}+\tfrac{({\zeta }_{3}-{\zeta }_{2})}{K(p)}{\int }_{0}^{K(p)}{\mathrm{sn}}^{2}\ u\ {du}$. Secondly, we have from (4.4) that

We calculate

where we used (4.2), (4.3) and (4.10). □

4.4. Expression of Blaschke immersions in terms of Riemann theta function

We gave a formula of Blaschke immersion in terms of elliptic functions in (4.6). In this section, we rewrite the formula (4.6) in terms of Riemann theta function. For this purpose, we consider the elliptic curve ${{ \mathcal C }}^{0}$ as a Prym variety of a compact Riemann surface $\hat{{ \mathcal C }}$ of genus 2. When $u=u(\hat{y})$ is given by (4.1), the spectral curve is defined by the equation $\det \ (U(\lambda )+V(\lambda )-\mu I)=0$, where $U(\lambda )$ and $V(\lambda )$ are as in (2.5). Since the spectral curve is independent of the parameter $\hat{y}$, if we take $\hat{y}=0$ then the initial condition of $u=u(\hat{y})$ implies that the defining equation of the spectral curve is ${\mu }^{3}-2a\mu ={\lambda }^{3}+{\lambda }^{-3}$. The unramified covering map $\lambda \to \nu ={\lambda }^{3}$ yields ${\mu }^{3}-2a\mu =\nu +{\nu }^{-1}$. Projectivizing this affine plane curve, we obtain a compact Riemann surface $\hat{{ \mathcal C }}$ of genus 2. Since ${(\nu -{\nu }^{-1})}^{2}={(\nu +{\nu }^{-1})}^{2}-4={({\mu }^{3}-2a\mu )}^{2}-4$, setting $\widetilde{\nu }=\nu -{\nu }^{-1}$, we see that $\hat{{ \mathcal C }}$ may be regarded as a hyperelliptic curve ${({\mu }^{3}-2a\mu )}^{2}-4={\widetilde{\nu }}^{2}$. We may observe that $\hat{{ \mathcal C }}$ admits holomorphic involution σ and an anti-holomorphic involution ρ defined by $\sigma (\mu ,\nu )=(-\mu ,-\nu )$ and $\rho (\mu ,\nu )=(\overline{\mu },{\overline{\nu }}^{-1})$. If we transform the parameters by

then we obtain the elliptic curve ${\hat{{ \mathcal C }}}^{0}:{\widetilde{B}}^{2}=-8(\zeta -{\zeta }_{1})(\zeta -{\zeta }_{2})(\zeta -{\zeta }_{3})$. We may use ${\hat{{ \mathcal C }}}^{0}$ as a Prym variety of $\hat{{ \mathcal C }}$. Let $\varphi :\hat{{ \mathcal C }}\longrightarrow {\hat{{ \mathcal C }}}^{0}$ be a covering map defined as above. We may choose the canonical homology basis $\{{\hat{a}}_{1},{\hat{a}}_{2},{\hat{b}}_{1},{\hat{b}}_{2}\}$ of $\hat{{ \mathcal C }}$ so that

hold (see Figure 3).

Figure 3.

Figure 3. Canonical homology basis.

Standard image High-resolution image

Let $\{{w}_{1},{w}_{2}\}$ be the basis of ${{\rm{H}}}^{1}(\hat{{ \mathcal C }},{\bf{C}})$ such that $(2\pi \sqrt{-1}\ {\rm{I}}\ \hat{T})$ is the Riemann period matrix for $\hat{{ \mathcal C }}$, where $\hat{T}=({\hat{T}}_{{ij}})$ and ${\hat{T}}_{{ij}}={\int }_{{\hat{b}}_{i}}{w}_{j}$ for i, $j=1$, 2. It is well known that the Jacobian variety $J(\hat{{ \mathcal C }})$ of $\hat{{ \mathcal C }}$ is a complex torus ${{\bf{C}}}^{2}/{\rm{\Lambda }}$, where ${\rm{\Lambda }}={\mathrm{Span}}_{{\bf{Z}}}\{2\pi \sqrt{-1}{\rm{I}},\hat{T}\}$. The basis $\{{w}_{1},{w}_{2}\}$ can be constructed as follows. Set ${u}_{1}=\displaystyle \frac{d\mu }{\widetilde{\nu }},{u}_{2}=\displaystyle \frac{\mu d\mu }{\widetilde{\nu }}$. We then see that ${\varphi }^{* }{w}_{1}^{0}=-\tfrac{\pi \sqrt{-1}}{{\omega }^{0}}\ {u}_{2}$. We want to look for ${w}_{1},{w}_{2}$ which satisfy ${w}_{1}+{w}_{2}=-\tfrac{\pi \sqrt{-1}}{{\omega }^{0}}\ {u}_{2}$. Note that $\{{u}_{1},{u}_{2}\}$ is a basis of ${{\rm{H}}}^{1}(\hat{{ \mathcal C }},{\bf{C}})$ and ${\sigma }^{* }{u}_{1}={u}_{1}$ and ${\sigma }^{* }{u}_{2}=-{u}_{2}$. If we set ${C}_{{ij}}={\int }_{{\hat{a}}_{i}}{u}_{j}$ for i, $j=1$, 2 then we have ${C}_{11}=-{C}_{21}$ and ${C}_{12}={C}_{22}=-2{\omega }^{0}$. The last equation follows from ${\int }_{{\hat{a}}_{j}}{\varphi }^{* }{w}_{1}^{0}={\int }_{\varphi ({\hat{a}}_{j})}{w}_{1}^{0}={\int }_{{a}_{1}^{0}}{w}_{1}^{0}=2\pi \sqrt{-1}$ for $j=1$, 2. The Riemann bilinear relation means that ${C}_{11}\ne 0$. Thus, we find ${w}_{1},{w}_{2}$ :

with the properties ${\varphi }^{* }{w}_{1}^{0}={w}_{1}+{w}_{2}$ and ${\sigma }^{* }{w}_{1}=-{w}_{2}$. If we set ${T}_{j1}={\int }_{{\hat{b}}_{j}}({w}_{1}-{w}_{2})$ then we have ${T}_{11}=-{T}_{21}=: T$. We then know that

Equation (4.19)

for ${\rm{\Pi }}={\rm{\Pi }}(p)$, which is as in (4.9). Moreover, since $\overline{{\rho }^{* }{u}_{j}}=-{u}_{j}$ for $j=1$, 2 we have that

Equation (4.20)

In fact, since $\hat{{ \mathcal C }}$ is an M-curve (see [27]), there are 3 real ovals, which are connected components of fixed points of ρ, and we may take ${\hat{b}}_{j}\ (j=1,2)$ as these ovals. In this case, the number C11 above is real and $\overline{{\rho }^{* }{w}_{j}}={w}_{j}$ for $j=1,2$ hold, which is compatible with the reality of Π. We define a map ${ \mathcal B }:\hat{{ \mathcal C }}\longrightarrow \mathrm{Prym}(\hat{{ \mathcal C }})$ by ${ \mathcal B }(\hat{P})={\int }_{{\hat{P}}_{0}}^{\hat{P}}({w}_{1}+{w}_{2})$. If we change $\hat{P}\to \hat{P}+{m}_{j}{\hat{a}}_{j}+{n}_{j}{\hat{b}}_{j}$, where ${m}_{j},{n}_{j}\in {\bf{Z}}$ and $j=1$, 2, then ${ \mathcal B }(\hat{P})$ changes into ${ \mathcal B }(\hat{P})+2\pi \sqrt{-1}{m}_{j}\,+{\rm{\Pi }}{n}_{j}$, hence it is a well-defined map into $\mathrm{Prym}(\hat{{ \mathcal C }})\cong {{ \mathcal C }}^{0}={\bf{C}}/{\rm{\Gamma }}$, where ${\rm{\Gamma }}={\mathrm{Span}}_{{\bf{Z}}}\{2\pi \sqrt{-1},{\rm{\Pi }}\}$. We have the relations as follows.

Equation (4.21)

We may consider the function f on $\hat{{ \mathcal C }}$ defined by $f(\hat{P})=\theta ({ \mathcal B }(\hat{P})-{\bf{e}})$ for ${\bf{e}}\in {\bf{C}}$, where θ is the Riemann theta function $\theta (z)=\theta (z;{\rm{\Pi }}(p))$ as in section 4.3. Let ${\hat{{ \mathcal C }}}_{0}$ be the simply-connected domain obtained by removing all $\hat{a}$- and $\hat{b}$-cycles from $\hat{{ \mathcal C }}$. The boundary of ${\hat{{ \mathcal C }}}_{0}$ is denoted by $\partial {\hat{{ \mathcal C }}}_{0}={\hat{a}}_{1}{\hat{b}}_{1}{\hat{a}}_{1}^{-1}{\hat{b}}_{1}^{-1}{\hat{a}}_{2}{\hat{b}}_{2}{\hat{a}}_{2}^{-1}{\hat{b}}_{2}^{-1}$. We then know that f is single-valued on ${\hat{{ \mathcal C }}}_{0}$. However, since the theta function satisfies the relations $\theta (z+2\pi \sqrt{-1})=\theta (z),\theta (z+{\rm{\Pi }})\,=\exp \left(-\tfrac{1}{2}{\rm{\Pi }}-z\right)\theta (z)$, it is no problem to consider f on $\hat{{ \mathcal C }}$ when we search for the zeros of the function f. With respect to this problem, there is a study of J. Fay.

(cf [29]).

Lemma 4.5 If $f(\hat{P})=\theta ({ \mathcal B }(\hat{P})-{\bf{e}})\not\equiv 0$ then the zeros of f is a degree 2 divisor $\hat{{ \mathcal D }}$. Moreover, we have ${ \mathcal B }(\hat{{ \mathcal D }})\equiv {\bf{K}}+2{\bf{e}}\ (\mathrm{mod}\ {\rm{\Gamma }})$, where ${\bf{K}}$ is given by

and ${\hat{b}}_{j}(0)$ is the initial point of the path ${\hat{b}}_{j}$ in the boundary $\partial {\hat{{ \mathcal C }}}_{0}$.

This is, of course, a special one of the higher genus case, which is stated in the next section. Fay states the property of the divisor $\hat{{ \mathcal D }}$ in terms of the Abelian map ${ \mathcal A }:\hat{{ \mathcal C }}\longrightarrow J(\hat{{ \mathcal C }})$. Therefore, we here address the outline of the proof. When $\hat{P}\in {\hat{a}}_{j}({\rm{resp.}}\ {\hat{b}}_{j})$, we denote by ${\hat{P}}^{-}$ the corresponding point of ${\hat{a}}_{j}^{-1}({\rm{resp.}}\ {\hat{b}}_{j}^{-1})$. Set ${f}^{-}(\hat{P})=f({\hat{P}}^{-})$. If f is not identically zero, then the number n of the zeros of f is given by

But, since we may observe that ${ \mathcal B }({\hat{P}}^{-})={ \mathcal B }(\hat{P})+{\rm{\Pi }}$ when $\hat{P}\in {\hat{a}}_{j}$ and ${ \mathcal B }({\hat{P}}^{-})={ \mathcal B }(\hat{P})-2\pi \sqrt{-1}$ when $\hat{P}\in {\hat{b}}_{j}$, the property of θ yields $n=\tfrac{1}{2\pi \sqrt{-1}}{\sum }_{j=1}^{2}{\int }_{{\hat{a}}_{j}}d{ \mathcal B }=2$. Thus, the divisor $\hat{{ \mathcal D }}$ is of degree 2 and expressed as $\hat{{ \mathcal D }}=\{{\hat{p}}_{1},{\hat{p}}_{2}\}$. By the residue theorem, we have ${ \mathcal B }(\hat{{ \mathcal D }})=\displaystyle \frac{1}{2\pi \sqrt{-1}}{\int }_{\partial {\hat{{ \mathcal C }}}_{0}}{ \mathcal B }(\hat{P})d\mathrm{log}f$. Using the same calculation as in above, we obtain ${ \mathcal B }(\hat{{ \mathcal D }})\equiv \displaystyle \frac{1}{2\pi \sqrt{-1}}{\sum }_{j=1}^{2}{\int }_{{\hat{a}}_{j}}{ \mathcal B }({w}_{1}+{w}_{2})+{\sum }_{j=1}^{2}{\int }_{{\hat{b}}_{j}}d\mathrm{log}f$, $(\mathrm{mod}\ {\rm{\Gamma }})$. Finally, if we denote by ${\hat{b}}_{j}(1)$ the end point of the path ${\hat{b}}_{j}$ in the boundary $\partial {\hat{{ \mathcal C }}}_{0}$, then the property of θ yields $f({\hat{b}}_{j}(1))=\exp (-\tfrac{1}{2}{\rm{\Pi }}-{ \mathcal B }({\hat{b}}_{j}(0))+{\bf{e}})f({\hat{b}}_{j}(0))$, which implies the formula of ${ \mathcal B }(\hat{{ \mathcal D }})$ stated in lemma 4.5.

In our case, take ${\bf{e}}=\pi \sqrt{-1}$. Then, since the zeros of $\theta (z)$ are given by $z=\pi \sqrt{-1}\pm \tfrac{1}{\ 2\ }{\rm{\Pi }}\ (\mathrm{mod}\ {\rm{\Gamma }})$, it follows from (4.32) below that $\hat{{ \mathcal D }}=\{{\hat{P}}_{2},{\hat{P}}_{-2}\}$, where ${\hat{P}}_{2}=({\mu }_{2},1),{\hat{P}}_{-2}=(-{\mu }_{2},-1)$ in terms of affine coordinate system $(\mu ,\nu )$. Moreover, we see that ${ \mathcal B }(\hat{{ \mathcal D }})\equiv 0\ (\mathrm{mod}\ {\rm{\Gamma }})$.

Next, we look for the solution $\hat{{\rm{\Psi }}}=\hat{{\rm{\Psi }}}(x,t,\hat{\nu })$ of the Schrödinger equation ${\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}={e}^{u}\hat{{\rm{\Psi }}}$ on $\hat{{ \mathcal C }}$ of which the integrability condition ensures that ${e}^{u}$ coincides with the one in (4.17). Moreover, the evaluation of $\hat{{\rm{\Psi }}}$ at $\nu =1$ gives the Blaschke immersion $\psi (\hat{x},\hat{y})$ in (4.6). For this purpose, we want to describe $\hat{{\rm{\Psi }}}(x,t,\hat{\nu })$ in terms of the theta function θ by the following conditions : Denote by ${\hat{P}}_{0}$ and ${\hat{P}}_{\infty }$ the points on $\hat{{ \mathcal C }}$ which corresponds to the values $\hat{\nu }=0$ and $\hat{\nu }=\infty $, respectively, where $\hat{\nu }$ and ${\hat{\nu }}^{-1}$ are the local coordinates around ${\hat{P}}_{0}$ and ${\hat{P}}_{\infty }$ of $\hat{{ \mathcal C }}$, respectively.

  • (1)  
    $\hat{{\rm{\Psi }}}$ is a meromorphic function on $\hat{{ \mathcal C }}\setminus \{{\hat{P}}_{0},{\hat{P}}_{\infty }\}$ and the divisor of the poles is nonspecial and given by ${\hat{{ \mathcal D }}}_{\infty }=\{{\hat{p}}_{1},{\hat{p}}_{2}\}$ which is independent of the parameters x and t,
  • (2)  
    $\hat{{\rm{\Psi }}}$ has the following asymptotic expansions.

Consider the Abelian differential ${\omega }_{\hat{{\rm{\Psi }}}}$ defined by ${\omega }_{\hat{{\rm{\Psi }}}}=d\mathrm{log}\hat{{\rm{\Psi }}}$. Then, the principal parts of ${\omega }_{\hat{{\rm{\Psi }}}}$ around $\hat{\nu }=0$ and $\hat{\nu }=\infty $ are given by $-x{\hat{\nu }}^{-2}d\hat{\nu }$ and $t{\hat{\nu }}^{2}d{\hat{\nu }}^{-1}$, respectively. Since the residue of ${\omega }_{\hat{{\rm{\Psi }}}}$ is zero, it follows from the condition (1) that there are two points of zeros, which are denoted by ${\hat{q}}_{1}(x,t)$ and ${\hat{q}}_{2}(x,t)$. Therefore, ${\omega }_{\hat{{\rm{\Psi }}}}$ may be described as follows.

Equation (4.22)

where $\omega ({\hat{q}}_{j},{\hat{p}}_{j})$ is the normalized Abelian differential of 3rd kind with the principal part ${(\hat{\nu }-{\hat{q}}_{j})}^{-1}d\hat{\nu }$ and $-{(\hat{\nu }-{\hat{p}}_{j})}^{-1}d\hat{\nu }$, and ${\hat{{\rm{\Omega }}}}_{\infty }$ and ${\hat{{\rm{\Omega }}}}_{0}$ are the normalized Abelian differentials of second kind of the forms ${\hat{{\rm{\Omega }}}}_{\infty }=(-{\hat{\nu }}^{-2}+O(1))d\hat{\nu }\quad ({\rm{near}}\quad \ \hat{\nu }=0)$ and ${\hat{{\rm{\Omega }}}}_{0}=({\hat{\nu }}^{2}+O(1))d{\hat{\nu }}^{-1}\quad ({\rm{near}}\quad \ \hat{\nu }=\infty )$, respectively. We may define ${\hat{{\rm{\Omega }}}}_{\infty }$ and ${\hat{{\rm{\Omega }}}}_{0}$ as follows.

Equation (4.23)

If we set $\mu ={\hat{\nu }}^{-1}$, we then see that

near $\hat{\nu }=0$, where C is as that in (4.18). It then follows from (4.11) that ${\sigma }^{* }{\hat{{\rm{\Omega }}}}_{\infty }=-{\hat{{\rm{\Omega }}}}_{\infty },{\sigma }^{* }{\hat{{\rm{\Omega }}}}_{0}=-{\hat{{\rm{\Omega }}}}_{0}$ and $\overline{{\rho }^{* }{\hat{{\rm{\Omega }}}}_{\infty }}={\hat{{\rm{\Omega }}}}_{0}$. Now, since ${\int }_{{\hat{a}}_{j}}{\omega }_{\hat{{\rm{\Psi }}}}=2\pi \sqrt{-1}{m}_{j}$, the integration of (4.22) over the cycle ${\hat{a}}_{j}$ yields that ${m}_{j}\in {\bf{Z}}$. Set ${\bf{U}}=({U}_{1},{U}_{2})=\left({\int }_{{\hat{b}}_{1}}{\hat{{\rm{\Omega }}}}_{\infty },{\int }_{{\hat{b}}_{2}}{\hat{{\rm{\Omega }}}}_{\infty }\right)$ and ${\bf{V}}=({V}_{1},{V}_{2})=\left({\int }_{{\hat{b}}_{1}}{\hat{{\rm{\Omega }}}}_{0},{\int }_{{\hat{b}}_{2}}{\hat{{\rm{\Omega }}}}_{0}\right)$. Then, the various properties of ${\hat{{\rm{\Omega }}}}_{\infty },{\hat{{\rm{\Omega }}}}_{0}$ under the involutions σ, ρ and ${\int }_{{\hat{b}}_{j}}{\hat{{\rm{\Omega }}}}_{\infty }=-{\int }_{\varphi ({\hat{b}}_{j})}{{\rm{\Omega }}}_{\infty }^{0}=-{\int }_{{b}_{1}^{0}}{{\rm{\Omega }}}_{\infty }^{0}=-{{\bf{U}}}^{0}$ imply that

Equation (4.24)

Remark 4.6. Since ${w}_{1}+{w}_{2}=\displaystyle \frac{\pi \sqrt{-1}}{{\omega }^{0}}(1+O({\hat{\nu }}^{2}))d\hat{\nu }$, we have

which is nothing but the consequence of the reciprocity law for differentials of 1st and 2nd kinds.

The integration of (4.22) over the cycle ${\hat{b}}_{j}$ and the reciprocity law gives

Equation (4.25)

Define $F(z)$ by

Let $\{{z}_{1},{z}_{2},{z}_{3}\}$ be the set of points where ${{\mathfrak{P}}}^{{\prime} }(z)=0$. We then assume that ${z}_{1}={\omega }^{{\prime} },{z}_{2}=\omega ,{z}_{3}=\omega +{\omega }^{{\prime} }$. Note that ${\mathfrak{P}}({z}_{1})={\eta }_{1}$, ${\mathfrak{P}}({z}_{2})={\eta }_{2}$, ${\mathfrak{P}}({z}_{3})={\eta }_{3}$. We denote by ${\hat{P}}_{1}$, ${\hat{P}}_{2}$, ${\hat{P}}_{3}$ the points on $\hat{{ \mathcal C }}$ corresponding to ${\eta }_{1}$, ${\eta }_{2}$, ${\eta }_{3}$,respectively. Precisely, ${\hat{P}}_{1}=({\mu }_{1},1)$, ${\hat{P}}_{2}=({\mu }_{2},1)$, ${\hat{P}}_{3}=({\mu }_{3},1)$ by the affine coordinate $\hat{P}=(\mu ,\nu )$. We choose a path γ from ${\hat{P}}_{0}$ to ${\hat{P}}_{\infty }$ as in Figure 3 and we fix a path from ${\hat{P}}_{0}$ to ${\hat{P}}_{1}$ in the following. Since we have ${\int }_{\gamma }({w}_{1}+{w}_{2})\,\equiv 0\ (\mathrm{mod}\quad {\rm{\Gamma }})$ and we also have ${\int }_{\gamma }({w}_{1}+{w}_{2})$ is purely imaginary by $\rho (\gamma )=-\gamma $ and $\overline{{\rho }^{* }{w}_{j}}={w}_{j}$, we may fix a path from ${\hat{P}}_{1}$ to ${\hat{P}}_{-1}=\sigma ({\hat{P}}_{1})$ which passes through only $\hat{a}$-cycles. It then follows from

that

In fact, this is verified directly in (4.32) below. We have

We have ${\int }_{{\hat{P}}_{1}}^{\hat{P}}{\hat{{\rm{\Omega }}}}_{\infty }+\tfrac{1}{\ 2\ }{\mu }_{1}={\hat{\nu }}^{-1}+O(\hat{\nu })$ and ${\int }_{{\hat{P}}_{1}}^{\hat{P}}{\hat{{\rm{\Omega }}}}_{0}+\displaystyle \frac{1}{\ 2\ }{\mu }_{1}=C\hat{\nu }+O({\hat{\nu }}^{3})\ ({\rm{near}}\quad \hat{\nu }=0)$. We put

Equation (4.26)

where ${\bf{e}}\in {\bf{C}}$ is chosen so that $f(\hat{P})\not\equiv 0$ and the divisor of the poles of ${{\rm{\Phi }}}_{\theta }$ is $\hat{{ \mathcal D }}=\{{\hat{p}}_{1},{\hat{p}}_{2}\}$. We observe that ${{\rm{\Phi }}}_{\theta }{{\rm{\Phi }}}_{e}$ is invariant under the translation $\hat{P}\longrightarrow \hat{P}+{m}_{j}{\hat{a}}_{j}+{n}_{j}{\hat{b}}_{j}$ by the property of θ and (4.24). Therefore, it is a meromorphic function on $\hat{{ \mathcal C }}$. It then follows from lemma 4.5 and (4.25) that $\hat{{\rm{\Psi }}}{{\rm{\Phi }}}_{\theta }^{-1}{{\rm{\Phi }}}_{e}^{-1}$ is a holomorphic function on $\hat{{ \mathcal C }}$,hence it is a constant. Evaluating it at $\hat{\nu }=0$ we see that the constant is equal to $\displaystyle \frac{\theta ({\bf{e}})}{\theta ((x-t){{\bf{U}}}^{0}+{\bf{e}})}$. We thus obtain the following.

Lemma 4.7.  $\hat{{\rm{\Psi }}}(x,t,\hat{P},{\bf{e}})$ is given by

where ${{\rm{\Phi }}}_{e}(x,t,\hat{P})$ is as that in (4.26), and we fix the path from ${\hat{P}}_{0}$ to ${\hat{P}}_{1}$.

Moreover, using the expression of $\hat{{\rm{\Psi }}}$ we see that the following reality condition holds:

Equation (4.27)

As we see later, we may prove that $\hat{{\rm{\Psi }}}$ satisfies the Schrödinger equation ${\partial }_{t}{\partial }_{x}\hat{{\rm{\Psi }}}={e}^{u}\hat{{\rm{\Psi }}}$. For $U(\lambda )$, $V(\lambda )$ given in (2.5), let ${F}_{\lambda }:{{\bf{R}}}^{2}\longrightarrow {{\bf{C}}}^{3}$ be a solution of the differential equation

To simplify this equation, define ${\hat{F}}_{\nu }$ by

We then obtain

where

Equation (4.28)

If we express ${\hat{F}}_{\nu }$ as ${\hat{F}}_{\nu }=({\psi }_{0}\ {\psi }_{1}\ {\psi }_{2})$ then we write down the system of differential equations which we must solve as follows.

Equation (4.29)

Lemma 4.8.  $\{{\psi }_{0}={\nu }^{-1}{\partial }_{t}\hat{{\rm{\Psi }}},{\psi }_{1}={e}^{-u}{\partial }_{x}\hat{{\rm{\Psi }}},{\psi }_{2}=\hat{{\rm{\Psi }}}\}$ is a solution of the system of differential equations (4.29) above.

Set ${\hat{\psi }}_{0}={\hat{\nu }}^{6}{e}^{-\tfrac{u}{2}}{\psi }_{0},{\hat{\psi }}_{1}={\hat{\nu }}^{3}{e}^{\tfrac{u}{2}}{\psi }_{1},{\hat{\psi }}_{2}=\psi $ with $\nu ={\hat{\nu }}^{3}$. We define $\hat{W}(x,t,\hat{P})$ by

Equation (4.30)

It follows from (4.29) that $\hat{W}$ is independent of the parameters x and t. Thus, we may write $\hat{W}(x,t,\hat{P})=\hat{W}(\hat{P})$.

We will arrive at the following conclusion.

Theorem 4.9. For $\hat{{\rm{\Psi }}}$ in lemma 4.7, the Blaschke immersion $\psi (\hat{x},\hat{y})$ given in (4.6) can be described in terms of Riemann theta functions as

which is a solution of (2.1), where ${\hat{c}}_{1}{\hat{c}}_{2}{\hat{c}}_{3}=1$.

Moreover, the equation ${\partial }_{x}\ {\partial }_{t}\ \hat{{\rm{\Psi }}}={e}^{u}\ \hat{{\rm{\Psi }}}$ holds, where ${e}^{u}$ coincides with one given by (4.17), that is, a solution of the Tzitzèica equation (1.1).

Proof. Proof. Since $\hat{{\rm{\Psi }}}$ is single-valued, we carry out the calculations of integrations using the paths ${\gamma }_{+},{\hat{b}}_{1+},{\hat{a}}_{1+}$ by the parts of the disjoint unions $\gamma ={\gamma }_{+}\cup {\gamma }_{-},{\hat{b}}_{1}={\hat{b}}_{1+}\cup {\hat{b}}_{1-},{\hat{a}}_{1}={\hat{a}}_{1+}\cup {\hat{a}}_{1-}$

Let $\{{\hat{P}}_{1},{\hat{P}}_{2},{\hat{P}}_{3}\}$ be the points of $\hat{{ \mathcal C }}$ as above, which are also points of $\nu =1$. It follows from (4.10) and (4.11) that

where $j=1,2,3$. Therefore, we obtain

Equation (4.31)

On the other hand, since ${\varphi }^{* }\displaystyle \frac{d\zeta }{\widetilde{B}}={\varphi }^{* }\displaystyle \frac{{dz}}{\sqrt{2}\sqrt{-1}}$ we obtain

Equation (4.32)

We here remark that if $u=\sqrt{2}\sqrt{{\zeta }_{2}-{\zeta }_{1}}\ \hat{y}$ then $v=\displaystyle \frac{\ u\ }{2K(p)}=\displaystyle \frac{\ {{\bf{U}}}^{0}(x-t)\ }{2\pi \sqrt{-1}}$ by (4.2) and (4.12). We now set $u=\sqrt{2}\sqrt{{\zeta }_{2}-{\zeta }_{1}}\ \hat{y}$. It then follows from (4.13), (4.14), (4.31) and (4.32) that

Take a function $C(\hat{P})$ on $\hat{{ \mathcal C }}$ and define $\widetilde{{\rm{\Psi }}}$ by

where $C({\hat{P}}_{1})=C({\hat{P}}_{3})=1,C({\hat{P}}_{2})={(\hat{{\rm{\Psi }}}({\omega }^{0},-{\omega }^{0},{\hat{P}}_{2},\pi \sqrt{-1}))}^{-1}=0$. Let $\widetilde{P}$ be the point near ${\hat{P}}_{2}$ along ${\hat{a}}_{1+}$. We may write ${ \mathcal B }(\widetilde{P})=-\displaystyle \frac{1}{2}{\rm{\Pi }}-(a-\pi )\sqrt{-1}$, where $a\in {\bf{R}}$. We then have ${ \mathcal B }({\hat{P}}_{2})={\mathrm{lim}}_{a\to \pi }{ \mathcal B }(\widetilde{P})$. We define $\widetilde{W}$ using $\widetilde{{\rm{\Psi }}}$ as well as the way we defined $\hat{W}$ using $\hat{{\rm{\Psi }}}$. We then see that

Here, we note that

Therefore, we see that

Since we may obtain the explicit expression of $\widetilde{{\rm{\Psi }}}({\hat{P}}_{j}),(j=1,2,3)$ using the Jacobi elliptic functions, we obtain the following.

which implies that $\sqrt{\widetilde{W}({\hat{P}}_{1})\widetilde{W}({\hat{P}}_{2})\widetilde{W}({\hat{P}}_{3})}=\sqrt{2}\sqrt{{\zeta }_{2}-{\zeta }_{1}}({\zeta }_{3}-{\zeta }_{1})$.

Therefore, $\psi (\hat{x},\hat{y})$ in (4.6) can be described as those forms stated in the theorem. Next, we show that the $\hat{{\rm{\Psi }}}$ satisfies the Schrödinger equation ${\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}={e}^{u}\hat{{\rm{\Psi }}}$. Near ${\hat{P}}_{0}$, $\hat{{\rm{\Psi }}}(x,t,\hat{\nu })=\exp (x{\hat{\nu }}^{-1})(1+{\sum }_{j=1}^{\infty }{\hat{\xi }}_{j}{\hat{\nu }}^{j})$. We set ${e}^{u}={\partial }_{t}{\hat{\xi }}_{1}$. We then see that

Since the poles of $\hat{{\rm{\Psi }}}$ are independent of the parameters $x,t$, we see that the poles of $({\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}-{e}^{u}\hat{{\rm{\Psi }}})$ coincides with the zeros of ${\hat{{\rm{\Psi }}}}^{-1}$. But, since the zeros of $({\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}-{e}^{u}\hat{{\rm{\Psi }}})$ changes with $x,t$, if $\hat{{\rm{\Psi }}}\not\equiv 0$ then it follows from lemma 4.5 that the zeros of $\hat{{\rm{\Psi }}}$ are the degree 2 divisor $\{{\hat{q}}_{1}(x,t),{\hat{q}}_{2}(x,t)\}$. We then have $\hat{{\rm{\Phi }}}\in {{\rm{H}}}^{0}(\hat{{ \mathcal C }},{{ \mathcal O }}_{\hat{{ \mathcal C }}}({\hat{q}}_{1}(x,t)+{\hat{q}}_{2}(x,t)))$. Since $\{{\hat{q}}_{1}(x,t),{\hat{q}}_{2}(x,t)\}$ is a non-special divisor at the origin $(x,t)=(0,0)$, it remains non-special near the origin. Thus, $\hat{{\rm{\Phi }}}$ must be a constant on $\hat{{ \mathcal C }}$. Evaluating it at $\hat{\nu }=0$ we obtain $\hat{{\rm{\Phi }}}\equiv 0$, hence we have ${\partial }_{x}\ {\partial }_{t}\ \hat{{\rm{\Psi }}}-{e}^{u}\hat{{\rm{\Psi }}}\equiv 0$. Finally, we show that the ${e}^{u}$ above coincides with one in (4.17). Near ${\hat{P}}_{0}$, we may write ${\int }_{{\hat{P}}_{1}}^{\hat{P}}{\hat{{\rm{\Omega }}}}_{0}+\displaystyle \frac{1}{\ 2\ }{\mu }_{1}\,=C\hat{\nu }+O({\hat{\nu }}^{3})$, where $C=\tfrac{a}{3}-\tfrac{{\zeta }_{w}({\omega }^{{\prime} })}{{\omega }^{{\prime} }}$ as in (4.18). Differentiating $\mathrm{log}\hat{{\rm{\Psi }}}$ by $\hat{\nu }$ and setting $\hat{\nu }=0$, we obtain from lemma 4.7.

where ${C}_{0}(x)$ is a function of the parameter x only. It then follows from ${e}^{u}={\partial }_{t}{\hat{\xi }}_{1}$ that ${e}^{u}=C\,+{\partial }_{t}{\left.\displaystyle \frac{d}{d\hat{\nu }}\right|}_{\hat{\nu }=0}\mathrm{log}\theta ({ \mathcal B }(\hat{P})-(x-t){{\bf{U}}}^{0}-{\bf{e}})$. It follows from the reciprocity law for 1st and 2nd kinds that

which, together with ${e}^{u}={\partial }_{t}\ {\hat{\xi }}_{1}$, yields

5. Blaschke immersions of finite type in terms of Prym-theta functions

Let $d\equiv 1\ \mathrm{mod}\ 6$ and $\xi (x,t,\lambda )\in {{\rm{\Lambda }}}_{d}^{{\hat{\sigma }}^{* }}({\bf{D}})$ be a solution of $d\xi =[\xi ,{a}^{-1}{da}]$. For λ, $\mu \in {\bf{C}}$, the spectral curve is defined by the equation $\det (\xi (x,t,\lambda )-\mu I)=0$. However, since $\xi =\mathrm{Ad}\ ({a}^{-1})\xi (\lambda )$ for some initial data $\xi (\lambda )\in {{\rm{\Lambda }}}_{d}^{\hat{\sigma }}$, we see that the spectral curve is independent of the parameters x, t and given by the equation $\det (\xi (\lambda )-\mu I)=0$, which becomes ${\mu }^{3}-\tfrac{1}{2}(\mathrm{trace}\ \xi {(\lambda )}^{2})\mu =\det \xi (\lambda )$. Since the involution σ given in section 3 has a property that $\sigma (\xi (\lambda ))=\xi (-\lambda )$ and hence

holds, the spectral curve has a holomorphic involution σ defined by $\sigma (\mu ,\lambda )=(-\mu ,-\lambda )$, where $Q=\left(\begin{array}{ccc}0 & 1 & 0\\ 1 & 0 & 0\\ 0 & 0 & 1\end{array}\right)$. Moreover, since $\xi {(\lambda )}^{* }=\xi ({\overline{\lambda }}^{-1})$, the spectral curve also admits an anti-holomorphic involution ρ defined by $\rho (\mu ,\lambda )=(\overline{\mu },{\overline{\lambda }}^{-1})$. Set $P(\lambda )=\displaystyle \frac{1}{2}\mathrm{trace}\ \xi {(\lambda )}^{2},\quad Q(\lambda )=\det \xi (\lambda )$. The existence of the involutions σ, ρ and the special forms of elements of each ${{\mathfrak{g}}}_{j}\ (j=0,1,2,3,4,5)$ implies that $P(\lambda )$ and $Q(\lambda )$ have the expansions of the forms

where we have expressed d as $d=6k+1$. Therefore, we see that the spectral curve has $12d$ branch points. Projectivizing this affine plane curve, we have a compact Riemann surface ${ \mathcal C }$, whose genus is given by $(6d-2)$. A three-fold covering map given by ${ \mathcal C }\longrightarrow \hat{{ \mathcal C }},(\mu ,\lambda )\longrightarrow (\mu ,\nu ),\nu ={\lambda }^{3}$, is unramified, hence the genus $\hat{g}$ of $\hat{{ \mathcal C }}$ is $\hat{g}=2d$. The curve $\hat{{ \mathcal C }}$ also admits the involutions σ and ρ. We assume that $\hat{{ \mathcal C }}$ has no branch points over $\nu \in {S}^{1}$. We denote by $P(\nu )$ and $Q(\nu )$ the corresponding Laurent polynomials obtained by setting $\nu ={\lambda }^{3}$ in $P(\lambda )$ and $Q(\lambda )$, respectively. Let $\{{\nu }_{1},{\nu }_{2},\,\cdots ,\,{\nu }_{4k+1}\}$ be the set of $(4k+1)$ points of ${S}^{1}$ which are distinct each other. For example, we take ${\nu }_{j}=\exp (2\pi \sqrt{-1}j/(4k+1))$. For each j, the equation ${({\mu }^{3}-P({\nu }_{j})\mu )}^{2}-Q{({\nu }_{j})}^{2}=0$ has solutions $\mu =\pm {\mu }_{j1},\pm {\mu }_{j2},\pm {\mu }_{j3}$ with ${\mu }_{j1}\lt {\mu }_{j2}\lt {\mu }_{j3}$ and ${\mu }_{j1}+{\mu }_{j2}+{\mu }_{j3}=0$ because there is no branch points on ${S}^{1}$. Therefore, we may assume that our spectral curve is a hyperelliptic curve defined by

Note that both of $P({\nu }_{j})$ and $Q({\nu }_{j})$ are real for each $j=1$, $2,\,\cdots ,\,4k+1$.

Keeping these in mind, we consider a compact Riemann surface $\hat{{ \mathcal C }}$ of genus $\hat{g}$ obtained by projectivizing hyperelliptic curve defined by

This is an M-curve (see [27]). We see that the involutions σ and ρ act on $\hat{{ \mathcal C }}$ by $\sigma (\mu ,\widetilde{\nu })=(-\mu ,-\widetilde{\nu })$ and $\rho (\mu ,\widetilde{\nu })=(\overline{\mu },-\overline{\widetilde{\nu }})$.

We may choose the canonical homology basis $\{{\hat{a}}_{1},\,\cdots ,\,{\hat{a}}_{\hat{g}},{\hat{b}}_{1},\,\cdots ,\,{\hat{b}}_{\hat{g}}\}$ of $\hat{{ \mathcal C }}$ so that $\sigma ({\hat{a}}_{j})=-{\hat{a}}_{d+j},\sigma ({\hat{b}}_{j})\,=-{\hat{b}}_{d+j}$ hold for $j=1,2,\,\cdots ,\,d$ (see Figure 4).

Figure 4.

Figure 4. Canonical homology basis.

Standard image High-resolution image

Let $\{{u}_{1},\,\cdots ,\,{u}_{\hat{g}}\}$ be the basis of ${{\rm{H}}}^{1}(\hat{{ \mathcal C }},{\bf{C}})$ given by

We find the properties ${\sigma }^{* }{u}_{j}={u}_{j},{\sigma }^{* }{u}_{d+j}=-{u}_{d+j}$ for $j=1$, $2,\,\cdots ,\,d$. Define two $d\times d$-matrices $K=({K}_{{ij}})$ and $\hat{K}=({\hat{K}}_{{ij}})$ by ${K}_{{ij}}={\int }_{{\hat{a}}_{i}}{u}_{j}$ and ${\hat{K}}_{{ij}}={\int }_{{\hat{a}}_{i}}{u}_{d+j}$. Note that ${\int }_{{\hat{a}}_{d+i}}{u}_{j}=-{\int }_{{\hat{a}}_{i}}{u}_{j}$ and ${\int }_{{\hat{a}}_{d+i}}{u}_{d+j}={\int }_{{\hat{a}}_{i}}{u}_{d+j}$. Now, we may construct a new basis $\{{w}_{1},\,\cdots ,\,{w}_{\hat{g}}\}$ of ${{\rm{H}}}^{1}(\hat{{ \mathcal C }},{\bf{C}})$ as follows. If we write ${\bf{w}}={}^{t}({w}_{1},\,\cdots ,\,{w}_{d})$, $\hat{{\bf{w}}}={}^{t}({w}_{d+1},\,\cdots ,\,{w}_{\hat{g}})$, ${\bf{u}}={}^{t}({u}_{1},\,\cdots ,\,{u}_{d})$ and $\hat{{\bf{u}}}={}^{t}({u}_{d+1},\,\cdots ,\,{u}_{\hat{g}})$, we put

We then see that $\{{w}_{1},\,\cdots ,\,{w}_{\hat{g}}\}$ is the normalized basis of ${{\rm{H}}}^{1}(\hat{{ \mathcal C }},{\bf{C}})$ with the property ${\sigma }^{* }{\bf{w}}=-\hat{{\bf{w}}}$. Define a matrix $\hat{T}=({\hat{T}}_{\alpha \beta })$ by ${\hat{T}}_{\alpha \beta }={\int }_{{\hat{b}}_{\alpha }}{w}_{\beta }$, where α, $\beta =1$, $2,\,\cdots ,\,\hat{g}$. Moreover, define two matrices ${\rm{\Pi }}=({{\rm{\Pi }}}_{{ij}})$ and $T=({T}_{{ij}})$ by ${{\rm{\Pi }}}_{{ij}}={\int }_{{\hat{b}}_{i}}({w}_{j}+{w}_{d+j})$ and ${T}_{{ij}}={\int }_{{\hat{b}}_{i}}({w}_{j}-{w}_{d+j})$,respectively. We then see that

where ${i}^{{\prime} }=d+i$, ${j}^{{\prime} }=d+j$. From this, we may find a compact Riemann surface ${{ \mathcal C }}^{0}$ of genus d with the Riemann period matrix $(2\pi \sqrt{-1}\ {\rm{I}}\ {\rm{\Pi }})$ by Torelli's theorem. The Jacobian variety of ${{ \mathcal C }}^{0}$ is nothing but the Prym variety $\mathrm{Prym}(\hat{{ \mathcal C }})\cong {{\bf{C}}}^{d}/{\rm{\Gamma }}$ of $\hat{{ \mathcal C }}$, where ${\rm{\Gamma }}={\mathrm{Span}}_{{\mathbb{Z}}}\{2\pi \sqrt{-1}{\rm{I}},{\rm{\Pi }}\}$. A map ${ \mathcal B }:\hat{{ \mathcal C }}\longrightarrow \mathrm{Prym}(\hat{{ \mathcal C }})$ is defined by ${ \mathcal B }(\hat{P})={}^{t}({{ \mathcal B }}_{1}(\hat{P}),{{ \mathcal B }}_{2}(\hat{P}),\,\cdots ,\,{{ \mathcal B }}_{d}(\hat{P}))$ and ${{ \mathcal B }}_{j}(\hat{P})={\int }_{{\hat{P}}_{0}}^{\hat{P}}({w}_{j}+{w}_{d+j})$ for $j=1$, $2,\,\cdots ,\,d$. Since we have ${w}_{j}+{w}_{d+j}=2\pi \sqrt{-1}{\sum }_{i=1}^{d}{({\hat{K}}^{-1})}_{{ij}}{u}_{d+i}$ by the above expression, using the coordinate $\hat{\nu }={\mu }^{-1}$ around $\mu =\infty $, we have

Equation (5.1)

Consider the Riemann theta function θ on $\mathrm{Prym}(\hat{{ \mathcal C }})$, which is defined by

where $z\in {{\bf{C}}}^{d}$ and $\langle \cdot ,\cdot \rangle $ is the standard inner product of ${{\bf{C}}}^{d}$. The theta function has the quasi-periodic properties:

where ${{\bf{e}}}_{1},\,\cdots ,\,{{\bf{e}}}_{d}$ is a standard basis of ${{\bf{C}}}^{d}$ and $z={}^{t}({z}_{1},\,\cdots ,\,{z}_{d})$. Since $\hat{{ \mathcal C }}$ is an M-curve, we may take ${\hat{b}}_{j}^{{\prime} }$ s $(j=1,2,\,\cdots ,\,\hat{g})$ as real ovals. Therefore, we have $\rho ({\hat{a}}_{j})=-{\hat{a}}_{j},\rho ({\hat{b}}_{j})={\hat{b}}_{j}$ for $j=1,2,\,\cdots ,\,\hat{g}$, and thus we have $\overline{{\rho }^{* }{w}_{\alpha }}={w}_{\alpha }$ for $\alpha =1,2,\,\cdots ,\,\hat{g}$. From this, we see that Π is real and $\overline{\theta (z)}=\theta (\overline{z})$. Moreover, since $\overline{{\rho }^{* }{u}_{d+j}}=-{u}_{d+j}$ we see that $\hat{K}$ is a real matrix. Note that ${\int }_{{\hat{P}}_{0}}^{{\hat{P}}_{\infty }}({w}_{j}+{w}_{d+j})\equiv 0(\mathrm{mod}\ {\rm{\Gamma }})$ for $j=1,\,\cdots ,\,d$ because ${\sigma }^{* }{w}_{j}=-{w}_{d+j}$. Therefore, we obtain

Equation (5.2)

Consider a function f on $\hat{{ \mathcal C }}$ defined by $f(\hat{P})=\theta ({ \mathcal B }(\hat{P})-{\bf{e}})$ for ${\bf{e}}\in {{\bf{C}}}^{d}$. The method similar to those in section 4.4 yield the following.

(cf. [29]).

Lemma 5.1 If $f(\hat{P})=\theta ({ \mathcal B }(\hat{P})-{\bf{e}})\not\equiv 0$ for some ${\bf{e}}\in {{\bf{C}}}^{d}$ then the zeros of f is a degree $\hat{g}(=2d)$ divisor $\hat{{ \mathcal D }}$. Moreover, we have ${ \mathcal B }(\hat{{ \mathcal D }})\equiv {\bf{K}}+2{\bf{e}}\ (\mathrm{mod}\ {\rm{\Gamma }})$, where ${\bf{K}}={}^{t}({{\bf{K}}}_{1},\,\cdots ,\,{{\bf{K}}}_{d})\in {{\bf{C}}}^{d}$ and each ${{\bf{K}}}_{i}$ is given by

where $[j]=[2d+j]=j$, $[d+j]=d+j$ for $j=1,2,\,\cdots ,\,d$ and ${\hat{b}}_{i}(0)\ ({\rm{resp.}}\ {\hat{b}}_{d+i}(0))$ is the initial point of the path ${\hat{b}}_{i}\ ({\rm{resp.}}\ {\hat{b}}_{d+i})$ in the boundary $\partial {\hat{{ \mathcal C }}}_{0}$.

The description of ${{\bf{K}}}_{i}$ is proved by the method similar to that of lemma 4.5.

For $U(\lambda )$, $V(\lambda )$ given in (2.5), let ${F}_{\lambda }:{{\bf{R}}}^{2}\longrightarrow {{\bf{C}}}^{3}$ be a solution of the differential equation

To simplify this equation, define ${\hat{F}}_{\nu }$ by

We then obtain

where

Equation (5.3)

If we express ${\hat{F}}_{\nu }$ as ${\hat{F}}_{\nu }=({\psi }_{0}\ {\psi }_{1}\ {\psi }_{2})$ then we write down the system of differential equations which we must solve as follows.

Equation (5.4)

Since ${\partial }_{x}{\partial }_{t}{\psi }_{2}={e}^{u}{\psi }_{2}$, we want to describe ${\psi }_{2}$ in terms of the Riemann theta function as in theorem 4.9. We look for the solution $\hat{{\rm{\Psi }}}=\hat{{\rm{\Psi }}}(x,t,\hat{\nu })$ of the Schrödinger equation ${\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}={e}^{u}\hat{{\rm{\Psi }}}$ on $\hat{{ \mathcal C }}$ of which the integrability condition ensures that ${e}^{u}$ is a solution of the Tzitzéica equation. Moreover, the description of $\hat{{\rm{\Psi }}}$ in terms of the Riemann theta function yields the formula of the solution of Tzitzéica equation in terms of the Riemann theta function. For this purpose, we look for $\hat{{\rm{\Psi }}}$ by the following conditions : Denote by ${\hat{P}}_{0}$ and ${\hat{P}}_{\infty }$ the points on $\hat{{ \mathcal C }}$ which corresponds to the values $\hat{\nu }=0$ and $\hat{\nu }=\infty $, respectively, where $\hat{\nu }$ and ${\hat{\nu }}^{-1}$ are the local coordinates around ${\hat{P}}_{0}$ and ${\hat{P}}_{\infty }$ of $\hat{{ \mathcal C }}$, respectively.

  • (1)  
    $\hat{{\rm{\Psi }}}$ is e meromorphic function on $\hat{{ \mathcal C }}\setminus \{{\hat{P}}_{0},{\hat{P}}_{\infty }\}$ and the divisor of the poles is given by ${\hat{{ \mathcal D }}}_{\infty }=\{{\hat{p}}_{1},\,\cdots \,\cdots ,\,{\hat{p}}_{\hat{g}}\}$ which is independent of the parameters x and t,
  • (2)  
    $\hat{{\rm{\Psi }}}$ has the following asymptotic expansions.

Consider the Abelian differential ${\omega }_{\hat{{\rm{\Psi }}}}$ defined by ${\omega }_{\hat{{\rm{\Psi }}}}=d\mathrm{log}\hat{{\rm{\Psi }}}$. It follows from the condition (1) that there are $\hat{g}$ points of zeros, which are denoted by $\{{\hat{q}}_{1}(x,t),{\hat{q}}_{2}(x,t),\,\cdots ,\,{\hat{q}}_{\hat{g}}(x,t)\}$. Therefore, ${\omega }_{\hat{{\rm{\Psi }}}}$ may be described as follows.

Equation (5.5)

where $\omega ({\hat{q}}_{j},{\hat{p}}_{j})$ is the normalized Abelian differential of third kind with the principal part ${(\hat{\nu }-{\hat{q}}_{j})}^{-1}d\hat{\nu }$ and $-{(\hat{\nu }-{\hat{p}}_{j})}^{-1}d\hat{\nu }$, and ${\hat{{\rm{\Omega }}}}_{\infty }$ and ${\hat{{\rm{\Omega }}}}_{0}$ are the normalized Abelian differential of second kind with the properties ${\sigma }^{* }{\hat{{\rm{\Omega }}}}_{\infty }=-{\hat{{\rm{\Omega }}}}_{\infty },{\sigma }^{* }{\hat{{\rm{\Omega }}}}_{0}=-{\hat{{\rm{\Omega }}}}_{0}$ and $\overline{{\rho }^{* }{\hat{{\rm{\Omega }}}}_{\infty }}={\hat{{\rm{\Omega }}}}_{0}$ as those in section 4.4. Set $\hat{{\bf{U}}}={}^{t}({U}_{1},{U}_{2},\,\cdots ,\,{U}_{\hat{g}})$ and $\hat{{\bf{V}}}={}^{t}({V}_{1},{V}_{2},\,\cdots ,\,{V}_{\hat{g}})$, where ${U}_{j}={\int }_{{\hat{b}}_{j}}{\hat{{\rm{\Omega }}}}_{\infty }$ and ${V}_{j}={\int }_{{\hat{b}}_{j}}{\hat{{\rm{\Omega }}}}_{0}$ for $j=1,2,\,\cdots ,\,\hat{g}$. It follows from $\rho ({\hat{b}}_{j})={\hat{b}}_{j}$ that $\hat{{\bf{V}}}=\overline{\hat{{\bf{U}}}}$. The integration of (5.5) over the cycle ${\hat{b}}_{j}$ and the reciprocity law gives

Since ${\sigma }^{* }{\hat{{\rm{\Omega }}}}_{\infty }=-{\hat{{\rm{\Omega }}}}_{\infty }$ and $\sigma ({\hat{b}}_{j})=-{\hat{b}}_{d+j}$ we have ${U}_{d+j}={U}_{j}$ for $j=1,2,\,\cdots ,\,d$. Hence, it follows from ${\left.\displaystyle \frac{d}{d\hat{\nu }}\right|}_{\hat{\nu }=0}{{ \mathcal B }}_{j}(\hat{P})=-({U}_{j}+{U}_{d+j})=-2{U}_{j}$, the reality of the matrix $\hat{K}$ and (5.1) that ${U}_{j}={U}_{d+j}$ is purely imaginary for $j=1,2,\,\cdots ,\,d$. Therefore, we obtain

Equation (5.6)

We may give ${\hat{{\rm{\Omega }}}}_{\infty }$, ${\hat{{\rm{\Omega }}}}_{0}$ explicitly as follows.

Lemma 5.2. Define ${\hat{{\rm{\Omega }}}}_{\infty }$, ${\hat{{\rm{\Omega }}}}_{0}$ by

Equation (5.7)

where ${c}_{j}\ (j=1,2,\,\cdots ,\,d)$ is given by

Equation (5.8)

We then have the following.

  • (1)  
    cj is a real number for $j=1,2,\,\cdots ,\,d$,
  • (2)  
    ${\hat{{\rm{\Omega }}}}_{\infty }$ and ${\hat{{\rm{\Omega }}}}_{0}$ are the normalized Abelian differentials of 2nd kinds with the properties ${\sigma }^{* }{\hat{{\rm{\Omega }}}}_{\infty }\,=-{\hat{{\rm{\Omega }}}}_{\infty },{\sigma }^{* }{\hat{{\rm{\Omega }}}}_{0}=-{\hat{{\rm{\Omega }}}}_{0},\overline{{\rho }^{* }{\hat{{\rm{\Omega }}}}_{\infty }}={\hat{{\rm{\Omega }}}}_{0}$ and ${\hat{{\rm{\Omega }}}}_{\infty }=(-{\hat{\nu }}^{-2}+O(1))d\hat{\nu }$, ${\hat{{\rm{\Omega }}}}_{0}=(C+O({\hat{\nu }}^{2}))d\hat{\nu }$ near $\hat{\nu }(={\mu }^{-1})=0$, where
    Equation (5.9)

Proof. 

  • (1)  
    Since $\rho ({\hat{a}}_{i})=-{\hat{a}}_{i}$ for $i=1,2,\,\cdots ,\,d$, we have
    which, together with the reality of the matrix $\hat{K}$ and (5.8), implies that each cj is a real number.
  • (2)  
    The properties ${\sigma }^{* }{\hat{{\rm{\Omega }}}}_{\infty }=-{\hat{{\rm{\Omega }}}}_{\infty }$ and ${\sigma }^{* }{\hat{{\rm{\Omega }}}}_{0}=-{\hat{{\rm{\Omega }}}}_{0}$ are clear, and the property $\overline{{\rho }^{* }{\hat{{\rm{\Omega }}}}_{\infty }}={\hat{{\rm{\Omega }}}}_{0}$ follows from the reality of cj. The choice of cj by (5.8) implies that all the ${\hat{a}}_{i}$- and ${\hat{a}}_{d+i}-$ cycles of ${\hat{{\rm{\Omega }}}}_{\infty },{\hat{{\rm{\Omega }}}}_{0}$ are zero. Finally, for $\hat{\nu }={\mu }^{-1}$ we have ${\widetilde{\nu }}^{-1}={\hat{\nu }}^{\hat{g}+1}{(1-{\sum }_{j=1}^{\hat{g}+1}{\mu }_{j}^{2}{\hat{\nu }}^{2}+O({\hat{\nu }}^{4}))}^{-\tfrac{1}{2}}={\hat{\nu }}^{\hat{g}+1}(1+\displaystyle \frac{1}{2}{\sum }_{j=1}^{\hat{g}+1}{\mu }_{j}^{2}\ {\hat{\nu }}^{2}+O({\hat{\nu }}^{4}))$, which, together with the definition of ${\hat{{\rm{\Omega }}}}_{\infty },{\hat{{\rm{\Omega }}}}_{0}$ in (5.7), yields
    near $\hat{\nu }=0$.

Set ${\bf{U}}={}^{t}({U}_{1},{U}_{2},\,\cdots ,\,{U}_{d})$. We choose three points $\{{\hat{P}}_{1},{\hat{P}}_{2},{\hat{P}}_{3}\}$ on $\hat{{ \mathcal C }}$, which are expressed as ${\hat{P}}_{1}=({\mu }_{1},0),{\hat{P}}_{2}=(-{\mu }_{2},0),{\hat{P}}_{3}=(-{\mu }_{3},0)$ using the coordinate function $(\mu ,\widetilde{\nu })$. It follows from lemma 5.2 that

We put

Equation (5.10)

where ${\bf{e}}\in {\bf{C}}$ is chosen so that $f(\hat{P})\not\equiv 0$ and the divisor of the poles of ${{\rm{\Phi }}}_{\theta }$ is $\hat{{ \mathcal D }}=\{{\hat{p}}_{1},{\hat{p}}_{2},\,\cdots ,\,{\hat{p}}_{\hat{g}}\}$. We observe that ${{\rm{\Phi }}}_{\theta }{{\rm{\Phi }}}_{e}$ is invariant under the translation $\hat{P}\longrightarrow \hat{P}+{m}_{j}{\hat{a}}_{j}+{n}_{j}{\hat{b}}_{j}$ by the property of θ. Therefore, it is a meromorphic function on $\hat{{ \mathcal C }}$. It follows from lemma 5.1 and (5.6) that $\hat{{\rm{\Psi }}}{{\rm{\Phi }}}_{\theta }^{-1}{{\rm{\Phi }}}_{e}^{-1}$ is a holomorphic function on $\hat{{ \mathcal C }}$, hence a constant. Evaluating it at $\hat{\nu }=0$ we see that the constant is equal to $\displaystyle \frac{\theta ({\bf{e}})}{\theta ((x-t){\bf{U}}-{\bf{e}})}$. We thus obtain the following.

Lemma 5.3.  $\hat{{\rm{\Psi }}}(x,t,\hat{P},{\bf{e}})$ is given by

where ${{\rm{\Phi }}}_{e}(x,t,\hat{P})$ is as that in (5.10), and we fix the path from ${\hat{P}}_{0}$ to ${\hat{P}}_{1}$.

For our chosen canonical homology basis of $\hat{{ \mathcal C }}$ in Figure 4, we decompose ${\hat{a}}_{j}$ and ${\hat{b}}_{j}$ into the disjoint unions ${\hat{a}}_{j}={\hat{a}}_{j+}\cup {\hat{a}}_{j-}$ and ${\hat{b}}_{j}={\hat{b}}_{j+}\cup {\hat{b}}_{j-}$, where ${\hat{a}}_{j+}$ is the part of ${\hat{a}}_{j}$ lying on $\mathrm{Im}\ (\mu )\geqq 0$ and ${\hat{b}}_{j+}$ is the part of ${\hat{b}}_{j}$ lying on the upper sheet over $\widetilde{\nu }=0$. We also decompose γ, which is the fixed path from ${\hat{P}}_{0}$ to ${\hat{P}}_{\infty }$, into the disjoint union $\gamma ={\gamma }_{+}\cup {\gamma }_{-}$, where ${\gamma }_{+}$ is the part of γ lying on the upper sheet over $\widetilde{\nu }=0$ and it is nothing but the fixed path from ${\hat{P}}_{0}$ to ${\hat{P}}_{1}$.

We here prepare the following lemma.

Lemma 5.4. Choose ${\bf{e}}=\pi \sqrt{-1}{\rm{\Delta }}$, where ${\rm{\Delta }}={}^{t}(1,1,\,\cdots ,\,1)\in {{\bf{R}}}^{d}$. When we choose the paths

we have

Equation (5.11)

$(\mathrm{mod}\ 2\pi \sqrt{-1}{{\bf{Z}}}^{d})$ and

Equation (5.12)

Moreover, each $\hat{{\rm{\Psi }}}(x,t,{\hat{P}}_{j},{\bf{e}})$ is real for $j=1,2,3$.

Proof. Set ${a}_{j}^{0}:= \varphi ({\hat{a}}_{j})=\varphi ({\hat{a}}_{d+j})$ and ${b}_{j}^{0}:= \varphi ({\hat{b}}_{j})=\varphi ({\hat{b}}_{d+j})$ for $j=1,2,\,\cdots ,\,d$, where $\varphi :\hat{{ \mathcal C }}\longrightarrow \mathrm{Prym}(\hat{{ \mathcal C }})$ is a double covering map. Moreover, set ${P}_{0}:= \varphi ({\hat{P}}_{0})$ and ${P}_{\alpha }:= \varphi ({\hat{P}}_{\alpha })=\varphi ({\hat{P}}_{-\alpha })$ for $\alpha =1,2,3$, where ${\hat{P}}_{-\alpha }=\sigma ({\hat{P}}_{\alpha })$. There exists a normalized basis ${{\bf{w}}}^{0}$ of ${{\rm{H}}}^{1}(\mathrm{Prym}(\hat{{ \mathcal C }}),{\bf{C}})$ such that ${\varphi }^{* }{{\bf{w}}}^{0}=({\bf{w}}+\hat{{\bf{w}}})$. It follows from $\sigma ({\hat{P}}_{0})={\hat{P}}_{0},\sigma ({\hat{P}}_{\infty })={\hat{P}}_{\infty },\rho ({\hat{P}}_{0})={\hat{P}}_{\infty },{\sigma }^{* }({\bf{w}}+\hat{{\bf{w}}})=-({\bf{w}}+\hat{{\bf{w}}})$ and $\overline{{\rho }^{* }({\bf{w}}+\hat{{\bf{w}}})}=({\bf{w}}+\hat{{\bf{w}}})$ that ${\int }_{{\hat{P}}_{0}}^{{\hat{P}}_{\infty }}({\bf{w}}+\hat{{\bf{w}}})\equiv 0\quad \mathrm{mod}\ 2\pi \sqrt{-1}{{\bf{Z}}}^{d}$. Hence, we may fix a path from ${\hat{P}}_{1}$ to ${\hat{P}}_{-1}=\sigma ({\hat{P}}_{1})$ which passes through only $\hat{a}$-cycles. It then follows from

that ${\int }_{{\hat{P}}_{0}}^{{\hat{P}}_{1}}({\bf{w}}+\hat{{\bf{w}}})=-\pi \sqrt{-1}d{\rm{\Delta }}\equiv \pi \sqrt{-1}{\rm{\Delta }}={\bf{e}}\quad \mathrm{mod}\ 2\pi \sqrt{-1}{{\bf{Z}}}^{d}$. Next, we also decompose ${a}_{d}^{0}$ and ${b}_{d}^{0}$ into the disjoint unions ${a}_{d}^{0}={a}_{d+}^{0}\cup {a}_{d-}^{0}$ and ${b}_{d}^{0}={b}_{d+}^{0}\cup {b}_{d-}^{0}$ as in the case of cycles in $\hat{{ \mathcal C }}$. It then follows from

that ${\int }_{{\hat{P}}_{1}}^{{\hat{P}}_{2}}({\bf{w}}+\hat{{\bf{w}}})=-\displaystyle \frac{1}{2}{{\rm{\Pi }}}_{d}$. Similarly, it follows from

that ${\int }_{{\hat{P}}_{2}}^{{\hat{P}}_{3}}({\bf{w}}+\hat{{\bf{w}}})=\pi \sqrt{-1}{{\bf{e}}}_{d}$. Thus we obtain (5.11). Next, we prove (5.12). First of all, we see that there is an Abelian differential of 2nd kind ${{\rm{\Omega }}}_{\infty }^{0}$ on $\mathrm{Prym}(\hat{{ \mathcal C }})$ such that ${\hat{{\rm{\Omega }}}}_{\infty }={\varphi }^{* }{{\rm{\Omega }}}_{\infty }^{0}+\displaystyle \frac{1}{2}d\mu $. Since ${\sigma }^{* }{\hat{{\rm{\Omega }}}}_{\infty }=-{\hat{{\rm{\Omega }}}}_{\infty }$, it then follows from

that

Similarly, we obtain ${\int }_{{\hat{P}}_{1}}^{{\hat{P}}_{3}}{\hat{{\rm{\Omega }}}}_{\infty }=-\displaystyle \frac{1}{2}{U}_{d}+\displaystyle \frac{1}{2}({\mu }_{3}-{\mu }_{1})$. □

For any function $\hat{\psi }=\hat{\psi }(x,t,\hat{P})$ on ${\bf{D}}\times \hat{{ \mathcal C }}$, we define an action ${\hat{\sigma }}^{* }$ by

where σ is the involution of $\hat{{ \mathcal C }}$.

We may verify that the following reality condition of $\hat{{\rm{\Psi }}}$ holds :

We define $\hat{W}(x,t,\hat{P})$ as in (4.30). We then see from (5.4) that $\hat{W}(x,t,\hat{P})$ is independent of the parameters x and t. Therefore, we may write it as $\hat{W}(\hat{P})$.

We now obtain the following.

Theorem 5.5. For ${\bf{e}}=\pi \sqrt{-1}{\rm{\Delta }}$, the $\hat{{\rm{\Psi }}}$ in lemma 5.3 satisfies ${\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}={e}^{u}\hat{{\rm{\Psi }}}$ and ${e}^{u}$ is a solution of the Tzitzèica equation and may be described as

where C is as in (5.9). A Blaschke immersion $\psi $ of an indefinite proper affine sphere with $\det { \mathcal F }={e}^{u}$ may be described as

where ${\hat{c}}_{1}{\hat{c}}_{2}{\hat{c}}_{3}=1$.

Proof. We choose a function ${e}^{u}$ so that the following estimate holds :

Equation (5.13)

Then, ${e}^{u}$ must be given as ${e}^{u}={\partial }_{t}\ {\hat{\xi }}_{1}=-{\partial }_{x}\ {\hat{\eta }}_{1}$. We then see that

Since the poles of $\hat{{\rm{\Psi }}}$ are independent of the parameters x, t, we see that the poles of $({\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}-{e}^{u}\hat{{\rm{\Psi }}})$ coincides with the zeros of ${\hat{{\rm{\Psi }}}}^{-1}$. However, since the zeros of $({\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}-{e}^{u}\hat{{\rm{\Psi }}})$ changes with $x,t$, if $\hat{{\rm{\Psi }}}\not\equiv 0$ then it follows from lemma 5.1 and lemma 5.3 that the zeros of $\hat{{\rm{\Psi }}}$ are the degree $\hat{g}$ divisor $\hat{{ \mathcal D }}$ given by $\hat{{ \mathcal D }}=\{{\hat{q}}_{1}(x,t),{\hat{q}}_{2}(x,t),\,\cdots ,\,{\hat{q}}_{\hat{g}}(x,t)\}$. We then have $\hat{{\rm{\Phi }}}\in {H}^{0}(\hat{{ \mathcal C }},{{ \mathcal O }}_{\hat{{ \mathcal C }}}(\hat{{ \mathcal D }}))$. Since $\hat{{ \mathcal D }}$ is non-special at the origin $(x,t)=(0,0)$, it remains non-special near the origin. Thus, $\hat{{\rm{\Phi }}}$ must be a constant on $\hat{{ \mathcal C }}$. Evaluating it at $\hat{\nu }=0$ we obtain $\hat{{\rm{\Phi }}}\equiv 0$, which implies that ${\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}\,-{e}^{u}\hat{{\rm{\Psi }}}\equiv 0$. We show ${e}^{u}$ is a solution of the Tzitzèica equation. Since ${\int }_{{\hat{P}}_{1}}^{\hat{P}}{\hat{{\rm{\Omega }}}}_{0}+\displaystyle \frac{1}{2}{\mu }_{1}=C\hat{\nu }+O({\hat{\nu }}^{3})$ and

the same method as those in the proof of theorem 4.9 yields that

Set ${\psi }_{2}=\hat{{\rm{\Psi }}}$ and define ${\psi }_{0}$, ${\psi }_{1}$ by ${\psi }_{0}={\nu }^{-1}{\partial }_{t}{\psi }_{2}$ and ${\psi }_{1}={e}^{-u}{\partial }_{x}{\psi }_{2}$, where ν is the coordinate function on $\hat{{ \mathcal C }}$. We show these satisfies all the equations in (5.4). First of all, it follows from ${\partial }_{x}{\partial }_{t}{\psi }_{2}={e}^{u}{\psi }_{2}$ that ${\partial }_{x}{\psi }_{0}={\nu }^{-1}{e}^{u}{\psi }_{2}$ and ${\partial }_{t}{\psi }_{1}=-{u}_{t}{\psi }_{1}+{\psi }_{2}$. Next, note that we may write $\nu ={\hat{\nu }}^{3}$ locally near ${\hat{P}}_{\infty }$ or ${\hat{P}}_{0}$. This, together with the definitions of ${\psi }_{0}$ and ${\psi }_{1}$, implies that

Thus, we have $({\partial }_{t}{\psi }_{0}-{u}_{t}{\psi }_{0}-{\psi }_{1}){\psi }_{2}^{-1}\longrightarrow 0$ as $\hat{\nu }\to \infty $, hence we have

Equation (5.14)

as above. Next, we show ${\partial }_{x}\ {\psi }_{1}={e}^{-2u}{\psi }_{0}$ holds using (5.10) and (5.14). It follows from (5.2) and (5.10) that $\overline{{\psi }_{2}(\rho (\hat{P}))}=({\hat{\sigma }}^{* }{\psi }_{2})(\sigma (\hat{P}))$. We calculate

hence we obtain

Equation (5.15)

Differentiating the first equation in (5.15) by the parameter t we have

Now, since

and $\overline{{\partial }_{t}{\psi }_{0}(\rho (\hat{P}))}={u}_{t}\overline{{\psi }_{0}(\rho (\hat{P}))}+\overline{{\psi }_{1}(\rho (\hat{P}))}$, where the last equation follows from (5.14), we have

by the last equation in (5.15). Therefore we obtain ${\partial }_{x}{\psi }_{1}={e}^{-2u}{\psi }_{0}$. Thus, we have proved (5.4). The integrability condition for (5.4) means that ${e}^{u}$ is the solution of the Tzitzèica equation.

On the other hand, we find that $\hat{W}(\hat{P})$ may be expressed as

We here prove

Equation (5.16)

For this, we calculate

which implies that ${\int }_{{\hat{P}}_{1}}^{\sigma ({\hat{P}}_{j})}{\hat{{\rm{\Omega }}}}_{\infty }+\displaystyle \frac{1}{2}{\mu }_{1}=-\left({\int }_{{\hat{P}}_{1}}^{{\hat{P}}_{j}}{\hat{{\rm{\Omega }}}}_{\infty }+\displaystyle \frac{1}{2}{\mu }_{1}\right)$. Therefore we have ${{\rm{\Phi }}}_{e}(x,t,{\hat{P}}_{j}){{\rm{\Phi }}}_{e}(x,t,\sigma ({\hat{P}}_{j}))=1$, from which and (5.2) and (5.11), we obtain (5.16). Set $V(x,t,{\hat{P}}_{j})=\hat{{\rm{\Psi }}}(x,t,{\hat{P}}_{j},{\bf{e}})\cdot \hat{{\rm{\Psi }}}(x,t,\sigma ({\hat{P}}_{j}),{\bf{e}})$,which is non-zero and real for $j=1$, 2, 3 by (5.16). Calculating ${\partial }_{x}{\partial }_{t}V(x,t,{\hat{P}}_{j})$ and using ${\partial }_{x}{\partial }_{t}\hat{{\rm{\Psi }}}={e}^{u}\hat{{\rm{\Psi }}}$ we have

which implies that each $\hat{W}({\hat{P}}_{j})$ is real. We now assume that ${\hat{P}}_{j}(j=1,2,3)$ is a zero of $f(\hat{P})=\theta ({ \mathcal B }(\hat{P})-{\bf{e}})$. Then ${\hat{P}}_{j}$ is also a zero of $f(\sigma (\hat{P}))$ by (5.2) and the properties of the theta function. Therefore, ${\hat{P}}_{j}$ is a pole of $\sqrt{| \hat{W}(\hat{P})| }$. Thus, the pole of $\hat{{\rm{\Psi }}}(x,t,\hat{P},{\bf{e}})$ cancel with the pole of $\sqrt{| \hat{W}(\hat{P})| }$ each other.

Therefore, as in the proof of theorem 4.9, introducing $\widetilde{{\rm{\Psi }}}(x,t,\hat{P})$ and $\widetilde{W}(\hat{P})$ with $C({\hat{P}}_{1})=C({\hat{P}}_{2})=1$ and $C({\hat{P}}_{3})={\left(\hat{{\rm{\Psi }}}\left(-\displaystyle \frac{\pi \sqrt{-1}}{2{U}_{d}},\displaystyle \frac{\pi \sqrt{-1}}{2{U}_{d}},{\hat{P}}_{3}\right)\right)}^{-1}$, which is zero by lemma 5.4, we obtain $\psi (\hat{x},\hat{y})$ as those forms stated in the theorem.□

Please wait… references are loading.
10.1088/2399-6528/aaeaa0