Startseite The elliptic sinh-Gordon equation in a semi-strip
Artikel Open Access

The elliptic sinh-Gordon equation in a semi-strip

  • Guenbo Hwang EMAIL logo
Veröffentlicht/Copyright: 1. Juni 2017

Abstract

We study the elliptic sinh-Gordon equation posed in a semi-strip by applying the so-called Fokas method, a generalization of the inverse scattering transform for boundary value problems. Based on the spectral analysis for the Lax pair formulation, we show that the spectral functions can be characterized from the boundary values. We express the solution of the equation in terms of the unique solution of the matrix Riemann–Hilbert problem whose jump matrices are defined by the spectral functions. Moreover, we derive the global algebraic relation that involves the boundary values. In addition, it can be verified that the solution of the elliptic sinh-Gordon equation posed in the semi-strip exists if the spectral functions defined by the boundary values satisfy this global relation.

MSC 2010: 47K15; 35Q55

1 Introduction

We study the boundary problem for the elliptic sinh-Gordon equation posed in a semi-strip,

(1.1) q x x + q y y = sinh q , ( x , y ) Ω ,

where Ω = { ( x , y ) 2 : 0 < x < , 0 < y < L } . It is well known that the elliptic sinh-Gordon equation is completely integrable [1, 3] and hence equation (1.1) can be analyzed by the inverse scattering transform. Indeed, the classical inverse scattering transform was applied to solve the elliptic sinh-Gordon equation posed in the entire plane { - < x , y < } (see [3, 15]). For more complicated or general domains, the so-called Fokas method can be applied to solve boundary value problems [2, 4, 5, 6, 7] (see also the monograph [9] and references therein). It should be noted that the method can be considered as a significant extension of the inverse scattering transform for boundary value problems. Regarding the boundary value problem for the elliptic sinh-Gordon equation, the Fokas method has been applied to solve problem (1.1) posed in the half plane { - < x < ,  0 < y < } (see [13]) and the quarter plane { 0 < x , y < } (see [14]). It has been shown in [13, 14] that the solution of the equation can be expressed in terms of the unique solution of the matrix Riemann–Hilbert problem defined by the spectral functions. It also has been shown that the solution of the equation exists if the spectral functions determined by the boundary values satisfy the so-called global relation.

In this paper, we extend the results presented in [13, 14] to the semi-strip (see also [11, 10, 16] for analogous results). The rigorous analysis of the Fokas method involves the following steps:

(i) Assuming that a smooth solution q ( x , y ) exists, we express the solution q ( x , y ) in terms of the unique solution of the matrix Riemann–Hilbert problem defined by the spectral functions { a j ( k ) , b j ( k ) } , j = 1 , 2 , 3 . These spectral functions are defined by boundary values { q ( x , 0 ) , q y ( x , 0 ) } , { q ( 0 , y ) , q x ( 0 , y ) } and { q ( x , L ) , q y ( x , L ) } , respectively. It should be remarked that the spectral functions satisfy the global algebraic relation that involves all boundary values:

(1.2)

(1.2a) a 3 ( k ) = a 1 ( k ) a 2 ( - k ) + b 1 ( k ) b 2 ( - k ) , k - ,
(1.2b) b 3 ( k ) e - 2 ω 2 ( k ) L = a 2 ( k ) b 1 ( k ) - a 1 ( k ) b 2 ( k ) , k - .

(ii) Given boundary values { g 0 ( x ) , g 1 ( x ) } , { f 0 ( y ) , f 1 ( y ) } and { h 0 ( x ) , h 1 ( x ) } , where

(1.3)

(1.3a) q ( x , 0 ) = g 0 ( x ) , q y ( x , 0 ) = g 1 ( x ) ,
(1.3b) q ( 0 , y ) = f 0 ( y ) , q x ( 0 , y ) = f 1 ( y ) ,
(1.3c) q ( x , L ) = h 0 ( x ) , q y ( x , L ) = h 1 ( x ) ,

we define the spectral functions and q ( x , y ) in terms of the solution for the Riemann–Hilbert problem formulated in step (i). Assuming that the spectral functions satisfy the global relation (1.2), we prove that the function q ( x , y ) defined by the solution of the Riemann–Hilbert problem solves equation (1.1) and satisfy the boundary values (1.3).

The outline of this work is the following. In Section 2, we introduce the Lax pair for the elliptic sinh-Gordon equation and eigenfunctions that satisfy both parts of the Lax pair. In Section 3, we discuss the spectral functions defined by the boundary values and the global algebraic relation that involves all boundary values. In Section 4, we formulate the matrix Riemann–Hilbert problem whose jump matrices are uniquely defined by the spectral functions. Analyzing the Riemann–Hilbert problem, we show that the solution of the equation exists if the boundary values satisfy the global relation. We end with concluding remarks in Section 5.

2 Preliminaries

It is well known that the elliptic sinh-Gordon equation (1.1) admits the following compatibility condition of the Lax pair [3, 13, 15]:

(2.1)

(2.1a) μ x + ω 1 ( k ) [ σ 3 , μ ] = Q ( x , y , k ) μ ,
(2.1b) μ y + ω 2 ( k ) [ σ 3 , μ ] = i Q ~ ( x , y , k ) μ ,

where k is a spectral parameter, μ is a 2 × 2 matrix-valued eigenfunction and

ω 1 ( k ) = - 1 2 i ( k - 1 4 k ) , ω 2 ( k ) = - 1 2 ( k + 1 4 k ) , Q ( x , y , k ) = 1 4 ( i 2 k ( cosh q - 1 ) - ( r + sinh q 2 k ) r - sinh q 2 k - i 2 k ( cosh q - 1 ) )

with

Q ~ ( x , y , k ) = Q ( x , y , - k ) , r ( x , y ) = i q x ( x , y ) + q y ( x , y ) , σ 3 = ( 1 0 0 - 1 ) ,

and the matrix commutator defined by [ σ 3 , A ] = σ 3 A - A σ 3 . We write σ ^ 3 A = [ σ 3 , A ] for the matrix commutator. It is convenient to use the notation

e σ ^ 3 ξ A = e σ 3 ξ A e - σ 3 ξ = ( a 11 e 2 ξ a 12 e - 2 ξ a 21 a 22 ) .

Note that the Lax pair (2.1) can be written as the simple formulation

d [ e ( ω 1 ( k ) x + ω 2 ( k ) y ) σ ^ 3 μ ( x , y , k ) ] = e ( ω 1 ( k ) x + ω 2 ( k ) y ) σ ^ 3 W ( x , y , k ) ,

where the differential 1-form W is given by

(2.2) W ( x , y , k ) = Q ( x , y , k ) μ ( x , y , k ) d x + i Q ~ ( x , y , k ) μ ( x , y , k ) d y .

Hence, we define eigenfunctions that satisfy both parts of the Lax pair (2.1) as

(2.3) μ j ( x , y , k ) = I + ( x j , y j ) ( x , y ) e - ( ω 1 ( k ) ( x - ξ ) + ω 2 ( k ) ( y - η ) ) σ ^ 3 W j ( ξ , η , k ) ,

where W j is the differential form given in (2.2) with μ j . Since the differential 1-form W ( x , y , k ) is closed, the integration in (2.3) does not depend on paths [6, 14]. We choose three distinct points ( x j , y j ) , j = 1 , 2 , 3 (see Figure 1),

( x 1 , y 1 ) = ( 0 , L ) , ( x 2 , y 2 ) = ( 0 , 0 ) , ( x 3 , y 3 ) = ( , y ) .

More specifically, the eigenfunctions μ j ( x , y , k ) , j = 1 , 2 , 3 , satisfy the following integral equations:

μ 1 ( x , y , k ) = I + 0 x e - ω 1 ( x - ξ ) σ ^ 3 ( Q μ 1 ) ( ξ , y , k ) 𝑑 ξ - i y L e - ( ω 1 ( k ) x + ω 2 ( k ) ( y - η ) ) σ ^ 3 ( Q ~ μ 1 ) ( 0 , η , k ) 𝑑 η ,
μ 2 ( x , y , k ) = I + 0 x e - ω 1 ( k ) ( x - ξ ) σ ^ 3 ( Q μ 2 ) ( ξ , y , k ) 𝑑 ξ + i 0 y e - ( ω 1 ( k ) x + ω 2 ( k ) ( y - η ) ) σ ^ 3 ( Q ~ μ 2 ) ( 0 , η , k ) 𝑑 η ,
μ 3 ( x , y , k ) = I - x e - ω 1 ( k ) ( x - ξ ) σ ^ 3 ( Q μ 3 ) ( ξ , y , k ) 𝑑 ξ .

Figure 1 
               The three distinct points for the eigenfunctions 
                     
                        
                           
                              μ
                              j
                           
                        
                        
                        {\mu_{j}}
                     
                  , 
                     
                        
                           
                              j
                              =
                              
                                 1
                                 ,
                                 2
                                 ,
                                 3
                              
                           
                        
                        
                        {j=1,2,3}
                     
                  .
Figure 1

The three distinct points for the eigenfunctions μ j , j = 1 , 2 , 3 .

Note that the off-diagonal components of the matrix-valued eigenfunctions μ j involve the explicit exponential terms and

Re ω 1 ( k ) < 0 for  Im k > 0 , Re ω 2 ( k ) < 0 for  Re k > 0 .

Thus, let the domains D j in the complex k-plane, j = 1 , , 4 (see Figure 2), be defined by

D 1 = { k : Re ω 1 ( k ) < 0 } { k : Re ω 2 ( k ) < 0 } ,
D 2 = { k : Re ω 1 ( k ) < 0 } { k : Re ω 2 ( k ) > 0 } ,
D 3 = { k : Re ω 1 ( k ) > 0 } { k : Re ω 2 ( k ) > 0 } ,
D 4 = { k : Re ω 1 ( k ) > 0 } { k : Re ω 2 ( k ) < 0 } .

As a result, the domains of analyticity and boundedness for the eigenfunctions can be determined:

  1. μ 1 ( x , y , k ) is analytic and bounded for k ( D 2 , D 4 ) ,

  2. μ 2 ( x , y , k ) is analytic and bounded for k ( D 1 , D 3 ) ,

  3. μ 3 ( x , y , k ) is analytic and bounded for k ( D 3 D 4 , D 1 D 2 ) .

For convenience, we write each column of μ j ( x , y , k ) as follows:

μ 1 = ( μ 1 ( 2 ) , μ 1 ( 4 ) ) , μ 2 = ( μ 2 ( 1 ) , μ 2 ( 3 ) ) , μ 3 = ( μ 3 ( 34 ) , μ 3 ( 12 ) ) ,

where the superscripts indicate the analytic and bounded domains D j , j = 1 , , 4 , for the columns of the matrix-valued eigenfunctions. Using integration by parts, in the appropriate domain, we know

(2.4) μ j ( x , y , k ) = I + O ( 1 k ) as  k .

Since trace ( Q ) = trace ( Q ~ ) = 0 , equation (2.4) implies that det μ j = 1 , j = 1 , 2 , 3 . The eigenfunctions enjoy the same symmetry as Q and Q ~ :

(2.5) μ 11 ( x , y , k ) = μ 22 ( x , y , - k ) , μ 21 ( x , y , k ) = - μ 12 ( x , y , - k ) ,

where the subscripts denote the ( i , j ) -component of the matrix.

Regarding the boundary values, we assume that g j , h j H 1 ( + ) and f j H 1 ( [ 0 , L ] ) for j = 0 , 1 .

Figure 2 
               The domain 
                     
                        
                           
                              D
                              j
                           
                        
                        
                        {D_{j}}
                     
                   and the oriented contour 
                     
                        
                           
                              L
                              j
                           
                        
                        
                        {L_{j}}
                     
                  , 
                     
                        
                           
                              j
                              =
                              
                                 1
                                 ,
                                 …
                                 ,
                                 4
                              
                           
                        
                        
                        {j=1,\ldots,4}
                     
                  .
Figure 2

The domain D j and the oriented contour L j , j = 1 , , 4 .

3 Spectral analysis

3.1 Spectral functions

Note that the matrix eigenfunctions μ 1 , μ 2 and μ 3 are fundamental solutions of the Lax pair (2.1) and μ 1 ( 0 , L , k ) = I and μ 2 ( 0 , 0 , k ) = I . Hence, the eigenfunctions are related by the so-called spectral functions, also known as the scattering matrices, S 1 ( k ) , S 2 ( k ) and S 3 ( k ) :

(3.1)

(3.1a) μ 3 ( x , y , k ) = μ 2 ( x , y , k ) e - ( ω 1 ( k ) x + ω 2 ( k ) y ) σ ^ 3 S 1 ( k ) , k ( + , - ) ,
(3.1b) μ 1 ( x , y , k ) = μ 2 ( x , y , k ) e - ( ω 1 ( k ) x + ω 2 ( k ) y ) σ ^ 3 S 2 ( k ) , k ( i + , i - ) ,
(3.1c) μ 3 ( x , y , k ) = μ 1 ( x , y , k ) e - ( ω 1 ( k ) x + ω 2 ( k ) ( y - L ) ) σ ^ 3 S 3 ( k ) , k ( - , + ) .

Let x = 0 , y = 0 in (3.1a) and (3.1b) and x = 0 , y = L in (3.1c). Then the spectral functions are given by

S 1 ( k ) = μ 3 ( 0 , 0 , k ) , S 2 ( k ) = μ 1 ( 0 , 0 , k ) , S 3 ( k ) = μ 3 ( 0 , L , k ) .

From the symmetry (2.5) of the eigenfunctions, we write the spectral functions S 1 ( k ) , S 2 ( k ) and S 3 ( k ) as

S j ( k ) = ( a j ( k ) - b j ( - k ) b j ( k ) a j ( - k ) ) , j = 1 , 2 , 3 .

Since det S j ( k ) = 1 , j = 1 , 2 , 3 , we find the following identities:

a j ( k ) a j ( - k ) + b j ( k ) b j ( - k ) = 1 , j = 1 , 2 , 3 .

Motivated by the above arguments, we define

Φ 1 ( x , k ) = μ 3 ( x , 0 , k ) , Φ 2 ( y , k ) = μ 1 ( 0 , y , k ) , Φ 3 ( x , k ) = μ 3 ( x , L , k ) .

More specifically, the functions Φ j , j = 1 , 2 , 3 , satisfy the integral equations

Φ 1 ( x , k ) = I - x e - ω 1 ( k ) ( x - ξ ) σ ^ 3 ( Q 0 Φ 1 ) ( ξ , k ) 𝑑 ξ , k ( - , + ) ,  0 x < ,
Φ 2 ( y , k ) = I - i y L e - ω 2 ( k ) ( y - η ) σ ^ 3 ( Q ~ 0 Φ 2 ) ( η , k ) 𝑑 η , k ,  0 y L ,
Φ 3 ( x , k ) = I - x e - ω 1 ( k ) ( x - ξ ) σ ^ 3 ( Q L Φ 3 ) ( ξ , k ) 𝑑 ξ , k ( - , + ) ,  0 x < ,

where Q 0 ( x , k ) = Q ( x , 0 , k ) , Q ~ 0 ( y , k ) = Q ( 0 , y , - k ) and Q L ( x , k ) = Q ( x , L , k ) . Note that

S j ( k ) = Φ j ( 0 , k ) , j = 1 , 2 , 3 ,

which immediately imply that the spectral functions a j ( k ) and b j ( k ) , j = 1 , 3 , are analytic and bounded for Im k < 0 , while the spectral functions a 2 ( k ) and b 2 ( k ) are analytic and bounded for k except for the essential singularities k = 0 and . Moreover, due to the symmetry (2.5), the functions Φ j , j = 1 , 2 , 3 , can also be written as

Φ j ( x , k ) = ( A j ( x , k ) - B j ( x , - k ) B j ( x , k ) A j ( x , - k ) ) , j = 1 , 3 , Φ 2 ( y , k ) = ( A 2 ( y , k ) - B 2 ( y , - k ) B 2 ( y , k ) A 2 ( y , - k ) ) .

In summary, we will define the integral representations for the spectral functions below.

Definition 3.1.

Given q ( x , 0 ) = g 0 ( x ) and q y ( x , 0 ) = g 1 ( x ) , the map

{ g 0 ( x ) , g 1 ( x ) } { a 1 ( k ) , b 1 ( k ) }

is defined by

(3.2)

a 1 ( k ) = 1 - 1 4 0 { i 2 k ( cosh g 0 ( ξ ) - 1 ) A 1 ( ξ , k )
(3.2a) - ( i g ˙ 0 ( ξ ) + g 1 ( ξ ) + 1 2 k sinh g 0 ( ξ ) ) B 1 ( ξ , k ) } d ξ , Im k < 0 ,
b 1 ( k ) = - 1 4 0 e - 2 ω 1 ( k ) ξ { ( i g ˙ 0 ( ξ ) + g 1 ( ξ ) - 1 2 k sinh g 0 ( ξ ) ) A 1 ( ξ , k )
(3.2b) - i 2 k ( cosh g 0 ( ξ ) - 1 ) B 1 ( ξ , k ) } d ξ , Im k < 0 ,

where the functions A 1 and B 1 are the solutions of the x-part of the Lax pair (2.1a) with y = 0 , i.e., A 1 and B 1 solve the following system of ordinary differential equations:

A 1 x = 1 4 [ i 2 k ( cosh g 0 ( x ) - 1 ) A 1 - ( i g ˙ 0 ( x ) + g 1 ( x ) + 1 2 k sinh g 0 ( x ) ) B 1 ] ,
B 1 x - 2 ω 1 ( k ) B 1 = 1 4 [ ( i g ˙ 0 ( x ) + g 1 ( x ) - 1 2 k sinh g 0 ( x ) ) A 1 - i 2 k ( cosh g 0 ( x ) - 1 ) B 1 ]

with lim x ( A 1 , B 1 ) = ( 1 , 0 ) .

Definition 3.2.

Given q ( 0 , y ) = f 0 ( y ) and q x ( 0 , y ) = f 1 ( y ) , the map

{ f 0 ( y ) , f 1 ( y ) } { a 2 ( k ) , b 2 ( k ) }

is defined by

(3.3)

a 2 ( k ) = 1 + i 4 0 L { i 2 k ( cosh f 0 ( η ) - 1 ) A 2 ( η , k )
(3.3a) + ( i f 1 ( η ) + f ˙ 0 ( η ) - 1 2 k sinh f 0 ( η ) ) B 2 ( η , k ) } d η , k ,
b 2 ( k ) = - i 4 0 L e - 2 ω 2 ( k ) η { ( i f 1 ( η ) + f ˙ 0 ( η ) + 1 2 k sinh f 0 ( η ) ) A 2 ( η , k )
(3.3b) + i 2 k ( cosh f 0 ( η ) - 1 ) B 2 ( η , k ) } d η , k ,

where the functions A 2 and B 2 are the solutions of the y-part of the Lax pair (2.1b) with x = 0 , i.e., A 2 and B 2 solve the following system of ordinary differential equations:

A 2 y = - i 4 [ i 2 k ( cosh f 0 ( y ) - 1 ) A 2 + ( i f 0 ( y ) + f ˙ 1 ( y ) - 1 2 k sinh f 0 ( y ) ) B 2 ] ,
B 2 y - 2 ω 2 ( k ) B 2 = i 4 [ ( i f 0 ( y ) + f ˙ 1 ( y ) + 1 2 k sinh f 0 ( y ) ) A 2 + i 2 k ( cosh f 0 ( y ) - 1 ) B 2 ]

with lim y ( A 2 , B 2 ) = ( 1 , 0 ) .

Definition 3.3.

Given q ( x , L ) = h 0 ( x ) and q y ( x , L ) = h 1 ( x ) , the map

{ h 0 ( x ) , h 1 ( x ) } { a 3 ( k ) , b 3 ( k ) }

is defined by

(3.4)

a 3 ( k ) = 1 - 1 4 0 { i 2 k ( cosh h 0 ( ξ ) - 1 ) A 3 ( ξ , k )
(3.4a) - ( i h ˙ 0 ( ξ ) + h 1 ( ξ ) + 1 2 k sinh h 0 ( ξ ) ) B 3 ( ξ , k ) } d ξ , Im k < 0 ,
b 3 ( k ) = - 1 4 0 e - 2 ω 1 ( k ) ξ { ( i h ˙ 0 ( ξ ) + h 1 ( ξ ) - 1 2 k sinh h 0 ( ξ ) ) A 3 ( ξ , k )
(3.4b) - i 2 k ( cosh h 0 ( ξ ) - 1 ) B 3 ( ξ , k ) } d ξ , Im k < 0 ,

where the functions A 3 and B 3 are the solutions of the x-part of the Lax pair (2.1a) with y = L , i.e., A 3 and B 3 solve the following system of ordinary differential equations:

A 3 x = 1 4 [ i 2 k ( cosh h 0 ( x ) - 1 ) A 3 - ( i h ˙ 0 ( x ) + h 1 ( x ) + 1 2 k sinh h 0 ( x ) ) B 3 ] ,
B 3 x - 2 ω 1 ( k ) B 3 = 1 4 [ ( i h ˙ 0 ( x ) + h 1 ( x ) - 1 2 k sinh h 0 ( x ) ) A 3 - i 2 k ( cosh h 0 ( x ) - 1 ) B 3 ]

with lim x ( A 3 , B 3 ) = ( 1 , 0 ) .

In what follows, we derive the global relation that involves all boundary values. It should be noted that the global relation plays a crucial role in the implementation of the Fokas method for boundary value problems. Evaluating equations (3.1a) and (3.1b) at x = 0 and y = L , we find

S 3 ( k ) = μ 2 ( 0 , L , k ) e - ω 2 ( k ) L σ ^ 3 S 1 ( k ) ,
I = μ 2 ( 0 , L , k ) e - ω 2 ( k ) L σ ^ 3 S 2 ( k ) .

We then obtain the following global relation in terms of the scattering matrices:

(3.5) e ω 2 ( k ) L σ ^ 3 S 3 ( k ) = S 2 - 1 ( k ) S 1 ( k ) .

In particular, the global relation (3.5) can be written as

(3.6)

(3.6a) a 3 ( k ) = a 1 ( k ) a 2 ( - k ) + b 1 ( k ) b 2 ( - k ) , k - ,
(3.6b) b 3 ( k ) e - 2 ω 2 ( k ) L = a 2 ( k ) b 1 ( k ) - a 1 ( k ) b 2 ( k ) , k - .

3.2 Spectral analysis at boundary values

In this section we characterize the boundary values in terms of the spectral functions. This can be done by the spectral analysis for the Lax pair at x = 0 , y = 0 and y = L , respectively.

Proposition 3.4.

The inverse map

{ a 1 ( k ) , b 1 ( k ) } { q ( x , 0 ) , q y ( x , 0 ) }

to the map defined in Definition 3.1 is given by

cosh q ( x , 0 ) = 1 - 8 i lim k k M 11 x ( x ) - 8 lim k ( k M 21 ( x ) ) 2 ,
i q x ( x , 0 ) + q y ( x , 0 ) = - 4 i lim k k M 21 ( x ) ,

where M ( x ) is the solution of the matrix Riemann–Hilbert problem

(3.7) M - ( x ) ( x , k ) = M + ( x ) ( x , k ) J ( x ) ( x , k ) , k

with the jump matrix J ( x ) given by

(3.8) J ( x ) ( x , k ) = ( 1 b 1 ( - k ) a 1 ( k ) e - 2 ω 1 ( k ) x b 1 ( k ) a 1 ( - k ) e 2 ω 1 ( k ) x 1 a 1 ( k ) a 1 ( - k ) ) , k .

Proof.

Evaluating equation (3.1a) at y = 0 , we find

μ 3 ( x , 0 , k ) = μ 2 ( x , 0 , k ) e - ω 1 ( k ) x σ ^ 3 S 1 ( k ) , k ( + , - ) ,  0 x < .

Note that the eigenfunction μ 2 ( 1 ) ( x , 0 , k ) is analytic and bounded for k D 1 D 2 and μ 2 ( 3 ) ( x , 0 , k ) is analytic and bounded for k D 3 D 4 . Thus, we formulate the matrix Riemann–Hilbert problem (3.7) with the jump matrix J ( x ) ( x , k ) given by (3.8), where the sectionally meromorphic functions M ± ( x ) are defined by

(3.9)

M + ( x ) ( x , k ) = ( μ 2 ( 1 ) ( x , 0 , k ) a 1 ( - k ) , μ 3 ( 12 ) ( x , 0 , k ) ) , Im k > 0 ,
M - ( x ) ( x , k ) = ( μ 3 ( 34 ) ( x , 0 , k ) , μ 2 ( 3 ) ( x , 0 , k ) a 1 ( k ) ) , Im k < 0

with det M ± ( x ) = 1 and M ± ( x ) = I + O ( 1 k ) as k . The remaining proof follows arguments similar to arguments in [14]. ∎

Proposition 3.5.

The inverse map

{ a 2 ( k ) , b 2 ( k ) } { q ( 0 , y ) , q x ( 0 , y ) }

to the map defined in Definition 3.2 is given by

cosh q ( 0 , y ) = 1 + 8 lim k k M 11 y ( y ) + 8 lim k ( k M 21 ( y ) ) 2 ,
i q x ( 0 , y ) + q y ( 0 , y ) = - 4 i lim k k M 21 ( y ) ,

where M ( y ) is the solution of the matrix Riemann–Hilbert problem

M - ( y ) ( y , k ) = M + ( y ) ( y , k ) J ( y ) ( y , k ) , k i

with the jump matrix J ( y ) given by

(3.10) J ( y ) ( y , k ) = ( 1 b 2 ( - k ) a 2 ( k ) e - 2 ω 2 ( k ) y b 2 ( k ) a 2 ( - k ) e 2 ω 2 ( k ) y 1 a 2 ( k ) a 2 ( - k ) ) , k i .

Proof.

The proof is similar to the arguments discussed in Proposition 3.4. From the spectral relation (3.1b) with x = 0 , we find the jump matrix J ( y ) ( y , k ) given in (3.10) and the sectionally meromorphic functions M ± ( y ) given by

(3.11)

M + ( y ) ( y , k ) = ( μ 2 ( 1 ) ( 0 , y , k ) a 2 ( - k ) , μ 1 ( 4 ) ( 0 , y , k ) ) , Re k > 0 ,
M - ( y ) ( y , k ) = ( μ 1 ( 2 ) ( 0 , y , k ) , μ 2 ( 3 ) ( 0 , y , k ) a 2 ( k ) ) , Re k < 0

with det M ± ( y ) = 1 and M ± ( y ) = I + O ( 1 k ) as k . Note that the eigenfunction μ 2 ( 1 ) ( 0 , y , k ) is analytic and bounded for k D 1 D 4 and the function μ 2 ( 3 ) ( 0 , y , k ) is analytic and bounded for k D 2 D 3 . ∎

The spectral analysis at the boundary y = L is similar to Proposition 3.4.

Proposition 3.6.

The inverse map

{ a 3 ( k ) , b 3 ( k ) } { q ( x , L ) , q y ( x , L ) }

to the map defined in Definition 3.3 is given by

cosh q ( x , L ) = 1 - 8 i lim k k M 11 x ( L ) - 8 lim k ( k M 21 ( L ) ) 2 ,
i q x ( x , L ) + q y ( x , L ) = - 4 i lim k k M 21 ( L ) ,

where M ( L ) is the solution of the matrix Riemann–Hilbert problem

(3.12) M - ( L ) ( x , k ) = M + ( L ) ( x , k ) J ( L ) ( x , k ) , k

with the jump matrix J ( L ) given by

(3.13) J ( L ) ( x , k ) = ( 1 b 3 ( - k ) a 3 ( k ) e - 2 ω 1 ( k ) x b 3 ( k ) a 3 ( - k ) e 2 ω 1 ( k ) x 1 a 3 ( k ) a 3 ( - k ) ) , k .

4 Riemann–Hilbert problem

In this section, we first formulate the matrix Riemann–Hilbert problem under the assumption that the solution q ( x , y ) of (1.1) exists and then we discuss the existence of the solution q ( x , y ) satisfying the boundary conditions. Using equations (3.1) and the global relation (3.6), after tedious but straightforward calculations, we formulate the following matrix Riemann–Hilbert problem:

(4.1) M - ( x , y , k ) = M + ( x , y , k ) J ( x , y , k ) , k ,

where the oriented contours = L 1 L 2 L 3 L 4 are given by (cf. Figure 2)

(4.2)

(4.2a) L 1 = D 1 D 2 , L 2 = D 2 D 3 ,
(4.2b) L 3 = D 3 D 4 , L 4 = D 4 D 1 ,

and the jump matrices are defined by

(4.3)

(4.3a) J 1 ( x , y , k ) = ( 1 0 b 2 ( k ) a 1 ( - k ) a 3 ( - k ) e 2 θ ( x , y , k ) 1 ) , k L 1 ,
(4.3b) J 2 ( x , y , k ) = ( a 2 ( k ) a 1 ( k ) a 3 ( - k ) - b 1 ( - k ) a 1 ( k ) e - 2 θ ( x , y , k ) - b 3 ( k ) a 3 ( - k ) e 2 ( θ ( x , y , k ) - ω 2 ( k ) L ) 1 ) , k L 2 ,
(4.3c) J 3 ( x , y , k ) = ( 1 - b 2 ( - k ) a 1 ( k ) a 3 ( k ) e - 2 θ ( x , y , k ) 0 1 ) , k L 3 ,
(4.3d) J 4 ( x , y , k ) = J 1 J 2 - 1 J 3 = ( 1 b 3 ( - k ) a 3 ( k ) e - 2 ( θ ( x , y , k ) - ω 2 ( k ) L ) b 1 ( k ) a 1 ( - k ) e 2 θ ( x , y , k ) a 2 ( - k ) a 1 ( - k ) a 3 ( k ) ) , k L 4

with θ ( x , y , k ) = ω 1 ( k ) x + ω 2 ( k ) y . The matrix-valued functions M ± are sectionally meromorphic and defined as follows:

M + ( x , y , k ) = { ( μ 2 ( 1 ) a 1 ( - k ) , μ 3 ( 12 ) ) for  k D 1 , ( μ 3 ( 34 ) , μ 2 ( 3 ) a 1 ( k ) ) for  k D 3 ,
M - ( x , y , k ) = { ( μ 1 ( 2 ) a 3 ( - k ) , μ 3 ( 12 ) ) for  k D 2 , ( μ 3 ( 34 ) , μ 1 ( 4 ) a 3 ( k ) ) for  k D 4 .

Note that det M ± = 1 and M ± = I + O ( 1 k ) as k . Indeed, using the global relation (3.6), we find

a 2 ( k ) - b 1 ( - k ) b 3 ( k ) e - 2 ω 2 ( k ) L = a 1 ( k ) a 3 ( - k ) ,

which implies that det ( J 2 ) = det ( J 4 ) = 1 .

The Riemann–Hilbert problem (4.1) can be solved by a Cauchy-type integral equation. Let J ~ = I - J . Then equation (4.1) becomes

M + ( x , y , k ) - M - ( x , y , k ) = M + ( x , y , k ) J ~ ( x , y , k ) .

Applying the Plemelj formula [9], the solution M of the Riemann–Hilbert problem (4.1) can be expressed as

M ( x , y , k ) = I + 1 2 i π M + ( x , y , k ) J ~ ( x , y , k ) d k k - k .

Expanding the integrand for k, we find

M ( x , y , k ) = I - 1 2 i π k M + ( x , y , k ) J ~ ( x , y , k ) 𝑑 k + O ( 1 k 2 ) , k .

Thus, we expand the solution M of the Riemann–Hilbert problem (4.1) as

(4.4) M ( x , y , k ) = I + M ( 1 ) ( x , y ) k + M ( 2 ) ( x , y ) k 2 + O ( 1 k 3 ) , k .

If we substitute the expansion (4.4) into the x-part of the Lax pair (2.1a), the ( 2 , 1 ) -component at O ( 1 ) yields

(4.5) i q x ( x , y ) + q y ( x , y ) = - 4 i M 21 ( 1 ) ( x , y )

and the ( 1 , 1 ) -component at O ( 1 k ) implies

M 11 x ( 1 ) ( x , y ) = - 1 8 i ( cosh q ( x , y ) - 1 ) - 1 4 ( i q x ( x , y ) + q y ( x , y ) ) M 21 ( 1 ) ( x , y ) .

Simplifying the above equation with (4.5), we obtain the reconstruction formula for q ( x , y ) in terms of the solution of the Riemann–Hilbert problem

cosh q ( x , y ) = 1 - 8 i M 11 x ( 1 ) ( x , y ) - 8 ( M 21 ( 1 ) ) 2 .

Similarly, if we substitute the expansion (4.4) into the y-part of the Lax pair, the solution q ( x , y ) can be written equivalently as

cosh q ( x , y ) = 1 + 8 M 11 y ( 1 ) ( x , y ) + 8 ( M 21 ( 1 ) ) 2 .

Let us now state the existence theorem for the elliptic sinh-Gordon equation in the semi-strip.

Theorem 4.1.

Let the functions g j ( x ) , h j ( x ) H 1 ( R + ) and f j ( y ) H 1 ( [ 0 , L ] ) , j = 0 , 1 , be given with the sufficiently small H 1 norms. Let the functions a j ( k ) , b j ( k ) , j = 1 , 2 , 3 , be given by (3.2), (3.3) and (3.4) in Definitions 3.1, 3.2 and 3.3, respectively. Suppose that the spectral functions a j and b j , j = 1 , 2 , 3 , satisfy the global relation (3.6). Let M ( x , y , k ) be the solution of the following matrix Riemann–Hilbert (RH) problem

(4.6) M - ( x , y , k ) = M + ( x , y , k ) J ( x , y , k ) , k ,

where det ( M ± ) = 1 , M ± = I + O ( 1 k ) as k , the oriented contours L are defined in (4.2) and the jump matrices J are given in (4.3).

Then the Riemann–Hilbert problem is uniquely solvable and the function q ( x , y ) defined by

(4.7) i q x + q y = - 4 i lim k k M 21 , cosh q ( x , y ) = 1 - 8 i lim k k M 11 x - 8 lim k ( k M 21 ) 2

solves the elliptic sinh-Gordon equation (1.1) satisfying the boundary conditions

(4.8)

(4.8a) q ( x , 0 ) = g 0 ( x ) , q y ( x , 0 ) = g 1 ( x ) ,
(4.8b) q ( 0 , y ) = f 0 ( y ) , q x ( 0 , y ) = f 1 ( y ) ,
(4.8c) q ( x , L ) = h 0 ( x ) , q y ( x , L ) = h 1 ( x ) .

Proof.

Using the dressing method and the vanishing lemma presented in [7, 9], we can prove the unique solvability of the Riemann–Hilbert problem (4.6) and then we can also verify that the q ( x , y ) defined in (4.7) solves the elliptic sinh-Gordon equation (1.1).

The proof that q ( x , y ) given in (4.7) satisfies the boundary values is similar to the argument presented in [14]. Here, we only state the proof of (4.8c). We first define

(4.9) M ( L ) ( x , k ) = { M ( x , L , k ) J 1 ( x , L , k ) for  k D 1 , M ( x , L , k ) for  k D 2 , M ( x , L , k ) F ( x , k ) for  k D 3 , M ( x , L , k ) J 3 - 1 ( x , L , k ) F ( x , k ) for  k D 4 ,

where

F ( x , k ) = ( 1 - b 2 ( - k ) a 1 ( k ) a 3 ( k ) e - 2 ( ω 1 ( k ) x + ω 2 ( k ) L ) 0 1 ) .

It is convenient to denote M ( x , L , k ) and M ( L ) ( x , k ) for k D j , j = 1 , , 4 , by M j ( x , L , k ) and M j ( L ) ( x , k ) , respectively. Then, equations (4.6) and (4.9) can be written as

(4.10)

(4.10a) M 2 ( x , L , k ) = M 1 J 1 ( x , L , k ) , M 2 ( x , L , k ) = M 3 J 2 ( x , L , k ) ,
(4.10b) M 4 ( x , L , k ) = M 3 J 3 ( x , L , k ) , M 4 ( x , L , k ) = M 1 J 4 ( x , L , k ) ,

and

(4.11)

(4.11a) M 1 ( L ) ( x , k ) = M 1 ( x , L , k ) J 1 ( x , L , k ) , M 2 ( L ) ( x , k ) = M 2 ( x , L , k ) ,
(4.11b) M 3 ( L ) ( x , k ) = M 3 ( x , L , k ) F ( x , k ) , M 4 ( L ) ( x , k ) = M 4 ( x , L , k ) J 3 - 1 ( x , L , k ) F ( x , k ) .

Combining (4.11) and (4.10), we find the jump conditions as

M 2 ( L ) ( x , k ) = M 1 ( L ) ( x , k ) , M 3 ( L ) ( x , k ) = M 2 ( L ) ( x , k ) J 2 - 1 ( x , L , k ) F ( x , k ) ,
M 4 ( L ) ( x , k ) = M 3 ( L ) ( x , k ) , M 4 ( L ) ( x , k ) = M 1 ( L ) ( x , k ) J 2 - 1 ( x , L , k ) F ( x , k ) .

Note that no jumps occur along the contours L 1 and L 3 and that J 2 - 1 ( x , L , k ) F ( x , k ) = J ( L ) ( x , k ) , where J ( L ) ( x , k ) is given in (3.13). Thus, we define

M ( L ) ( x , k ) = M + ( L ) ( x , k ) , k D 1 D 2 ,
M ( L ) ( x , k ) = M - ( L ) ( x , k ) , k D 3 D 4

and then equations (4.10) are equivalent to the Riemann–Hilbert problem (3.12) with the jump matrix J ( L ) ( x , k ) . Thus, evaluating (4.7) at y = L , we prove that the function q ( x , y ) defined in (4.7) satisfies the boundary values (4.8c).

In a similar way, equations (4.8a) and (4.8b) can be proved. ∎

5 Concluding remarks

In conclusion, we have studied the boundary value problem for the elliptic sinh-Gordon equation formulated in the semi-strip by the Fokas method, a generalization of the inverse scattering transform for boundary value problems. In particular, we have characterized the spectral functions in terms of the boundary values and we have derived the global algebraic relation that involves the boundary values. Furthermore, we have shown that the solution of the elliptic sinh-Gordon equation posed in the semi-strip exists provided that the spectral function defined by the boundary values satisfy the global relation. This solution can be expressed in terms of the unique solution of the matrix Riemann–Hilbert problem whose jump matrices are uniquely defined by the spectral functions. These spectral functions denoted by { a j ( k ) , b j ( k ) } , j = 1 , 2 , 3 , can be determined by the boundary values { q ( x , 0 ) , q y ( x , 0 ) } , { q ( 0 , y ) , q x ( 0 , y ) } and { q ( x , L ) , q y ( x , L ) } . However, for a well-posed boundary value problem, it is required that a subset of the boundary values should be prescribed. In this respect, it is necessary to characterize unknown boundary values, called the Dirichlet-to-Neumann map [8]. This characterization can be done by analyzing the global relation as discussed in [11, 10, 12] and we will address this issue in the near future.

Funding source: Daegu University

Award Identifier / Grant number: Research Grant

Award Identifier / Grant number: 2014

Funding statement: The work is supported by the Daegu University Research Grant, 2014.

References

[1] M. J. Ablowitz and P. A. Clarkson, Solitons, Nonlinear Evolution Equations and Inverse Scattering, Cambridge, Cambridge University Press, 1991. 10.1017/CBO9780511623998Suche in Google Scholar

[2] G. Biondini and G. Hwang, Initial-boundary value problems for discrete evolution equations: Discrete linear Schrödinger and integrable discrete nonlinear Schrödinger equations, Inverse Problems 24 (2008), Article ID 065011. 10.1088/0266-5611/24/6/065011Suche in Google Scholar

[3] M. Boiti, J.-P. Leon and F. Pempinelli, Integrable two-dimensional generalisation of the sine- and sinh-Gordon equations, Inverse Problems 3 (1987), 37–49. 10.1088/0266-5611/3/1/009Suche in Google Scholar

[4] A. S. Fokas, A unified transform method for solving linear and certain nonlinear PDEs, Proc. Roy. Soc. Lond. Ser. A 453 (1997), 1411–1443. 10.1098/rspa.1997.0077Suche in Google Scholar

[5] A. S. Fokas, On the integrability of certain linear and nonlinear partial differential equations, J. Math. Phys. 41 (2000), 4188–4237. 10.1063/1.533339Suche in Google Scholar

[6] A. S. Fokas, Two-dimensional linear PDEs in a convex polygon, Proc. Roy. Soc. Lond. Ser. A 457 (2001), 371–393. 10.1098/rspa.2000.0671Suche in Google Scholar

[7] A. S. Fokas, Integrable nonlinear evolution equations on the half-line, Comm. Math. Phys. 230 (2002), 1–39. 10.1007/s00220-002-0681-8Suche in Google Scholar

[8] A. S. Fokas, The generalized Dirichlet-to-Neumann map for certain nonlinear evolution PDEs, Comm. Pure Appl. Math. 58 (2005), 639–670. 10.1002/cpa.20076Suche in Google Scholar

[9] A. S. Fokas, A Unified Approach to Boundary Value Problems, CBMS-NSF Regional Conf. Ser. in Appl. Math. 78, SIAM, Philadelphia, 2008. 10.1137/1.9780898717068Suche in Google Scholar

[10] A. S. Fokas, J. Lenells and B. Pelloni, Boundary value problems for the elliptic sine-Gordon equation in a semi-strip, J. Nonlinear Sci. 23 (2013), 241–282. 10.1007/s00332-012-9150-5Suche in Google Scholar

[11] A. S. Fokas and B. Pelloni, The Dirichlet-to-Neumann map for the elliptic sine-Gordon equation, Nonlinearity 25 (2012), 1011–1031. 10.1088/0951-7715/25/4/1011Suche in Google Scholar

[12] G. Hwang, The Fokas method: The Dirichlet to Neumann map for the sine-Gordon equation, Stud. Appl. Math. 132 (2014), 381–406. 10.1111/sapm.12035Suche in Google Scholar

[13] G. Hwang, The elliptic sinh-Gordon equation in the half plane, J. Nonlinear Sci. Appl. 8 (2015), 163–173. 10.22436/jnsa.008.02.08Suche in Google Scholar

[14] G. Hwang, The elliptic sinh-Gordon equation in the quarter plane, J. Nonlinear Math. Phys. 23 (2016), 127–140. 10.1080/14029251.2016.1135646Suche in Google Scholar

[15] M. Jaworski and D. Kaup, Direct and inverse scattering problem associated with the elliptic sinh-Gordon equation, Inverse Problems 6 (1990), 543–556. 10.1088/0266-5611/6/4/006Suche in Google Scholar

[16] B. Pelloni and D. A. Pinotsis, The elliptic sine-Gordon equation in a half plane, Nonlinearity 23 (2010), 77–88. 10.1088/0951-7715/23/1/004Suche in Google Scholar

Received: 2016-09-19
Accepted: 2017-04-09
Published Online: 2017-06-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Artikel in diesem Heft

  1. Frontmatter
  2. Asymptotic behavior of evolution systems in arbitrary Banach spaces using general almost periodic splittings
  3. Solvability of a product-type system of difference equations with six parameters
  4. On Dirichlet problem for fractional p-Laplacian with singular non-linearity
  5. Absence of Lavrentiev gap for non-autonomous functionals with (p,q)-growth
  6. On a class of fully nonlinear parabolic equations
  7. On sign-changing solutions for (p,q)-Laplace equations with two parameters
  8. Weighted Caffarelli–Kohn–Nirenberg type inequalities related to Grushin type operators
  9. On the fractional p-Laplacian equations with weight and general datum
  10. An elliptic equation with an indefinite sublinear boundary condition
  11. Liouville-type theorems for elliptic equations in half-space with mixed boundary value conditions
  12. Well/ill-posedness for the dissipative Navier–Stokes system in generalized Carleson measure spaces
  13. Hypercontractivity, supercontractivity, ultraboundedness and stability in semilinear problems
  14. Theoretical analysis of a water wave model with a nonlocal viscous dispersive term using the diffusive approach
  15. A multiplicity result for asymptotically linear Kirchhoff equations
  16. Higher-order anisotropic models in phase separation
  17. Well-posedness and maximum principles for lattice reaction-diffusion equations
  18. Existence of a bound state solution for quasilinear Schrödinger equations
  19. Existence and concentration behavior of solutions for a class of quasilinear elliptic equations with critical growth
  20. Homoclinics for strongly indefinite almost periodic second order Hamiltonian systems
  21. A new method for converting boundary value problems for impulsive fractional differential equations to integral equations and its applications
  22. Diffusive logistic equations with harvesting and heterogeneity under strong growth rate
  23. On viscosity and weak solutions for non-homogeneous p-Laplace equations
  24. Periodic impulsive fractional differential equations
  25. A result of uniqueness of solutions of the Shigesada–Kawasaki–Teramoto equations
  26. Solutions of vectorial Hamilton–Jacobi equations are rank-one absolute minimisers in L
  27. Large solutions to non-divergence structure semilinear elliptic equations with inhomogeneous term
  28. The elliptic sinh-Gordon equation in a semi-strip
  29. The Gelfand problem for the 1-homogeneous p-Laplacian
  30. Boundary layers to a singularly perturbed Klein–Gordon–Maxwell–Proca system on a compact Riemannian manifold with boundary
  31. Subharmonic solutions of Hamiltonian systems displaying some kind of sublinear growth
  32. Multiple solutions for an elliptic system with indefinite Robin boundary conditions
  33. New solutions for critical Neumann problems in ℝ2
  34. A fractional Kirchhoff problem involving a singular term and a critical nonlinearity
  35. Existence and non-existence of solutions to a Hamiltonian strongly degenerate elliptic system
  36. Characterizing the strange term in critical size homogenization: Quasilinear equations with a general microscopic boundary condition
  37. Nonlocal perturbations of the fractional Choquard equation
  38. A pathological example in nonlinear spectral theory
  39. Infinitely many solutions for cubic nonlinear Schrödinger equations in dimension four
  40. On Cauchy–Liouville-type theorems
  41. Maximal Lp -Lq regularity to the Stokes problem with Navier boundary conditions
  42. Besov regularity for solutions of p-harmonic equations
  43. The classical theory of calculus of variations for generalized functions
  44. On the Cauchy problem of a degenerate parabolic-hyperbolic PDE with Lévy noise
  45. Hölder gradient estimates for a class of singular or degenerate parabolic equations
  46. Critical and subcritical fractional Trudinger–Moser-type inequalities on
  47. Multiple nonradial solutions for a nonlinear elliptic problem with singular and decaying radial potential
  48. Quantization of energy and weakly turbulent profiles of solutions to some damped second-order evolution equations
  49. An elliptic system with logarithmic nonlinearity
  50. The Caccioppoli ultrafunctions
  51. Equilibrium of a production economy with non-compact attainable allocations set
  52. Exact behavior around isolated singularity for semilinear elliptic equations with a log-type nonlinearity
  53. The higher integrability of weak solutions of porous medium systems
  54. Classification of stable solutions for boundary value problems with nonlinear boundary conditions on Riemannian manifolds with nonnegative Ricci curvature
  55. Regularity results for p-Laplacians in pre-fractal domains
  56. Carleman estimates and null controllability of a class of singular parabolic equations
  57. Limit profiles and uniqueness of ground states to the nonlinear Choquard equations
  58. On a measure of noncompactness in the space of regulated functions and its applications
  59. p-fractional Hardy–Schrödinger–Kirchhoff systems with critical nonlinearities
  60. On the well-posedness of a multiscale mathematical model for Lithium-ion batteries
  61. Global existence of a radiative Euler system coupled to an electromagnetic field
  62. On the existence of a weak solution for some singular p ( x ) -biharmonic equation with Navier boundary conditions
  63. Choquard-type equations with Hardy–Littlewood–Sobolev upper-critical growth
  64. Clustered solutions for supercritical elliptic equations on Riemannian manifolds
  65. Ground state solutions for the Hénon prescribed mean curvature equation
  66. Quasilinear equations with indefinite nonlinearity
  67. Concentrating solutions for a planar elliptic problem with large nonlinear exponent and Robin boundary condition
  68. Retraction of: Concentrating solutions for a planar elliptic problem with large nonlinear exponent and Robin boundary condition
Heruntergeladen am 23.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/anona-2016-0206/html
Button zum nach oben scrollen