Journal of Nonlinear Mathematical Physics

Volume 25, Issue 2, March 2018, Pages 324 - 350

Pseudo-Hermitian Reduction of a Generalized Heisenberg Ferromagnet Equation. I. Auxiliary System and Fundamental Properties

Authors
A. B. Yanovski
Department of Mathematics & Applied Mathematics, University of Cape Town, Rondebosch 7700, Cape Town, South Africa,Alexandar.Ianovsky@uct.ac.za
T. I. Valchev
Institute of Mathematics and Informatics, Bulgarian Academy of Sciences, Acad. G. Bonchev Str., 1113 Sofia, Bulgaria,tiv@math.bas.bg
Received 8 October 2017, Accepted 9 January 2018, Available Online 6 January 2021.
DOI
10.1080/14029251.2018.1452676How to use a DOI?
Keywords
Lax representation; gauge-equivalent equations; recursion operators; reductions
Abstract

We consider an auxiliary spectral problem originally introduced by Gerdjikov, Mikhailov and Valchev (GMV system) and its modification called pseudo-Hermitian reduction which is extensively studied here for the first time. We describe the integrable hierarchies of both systems in a parallel way and construct recursion operators. Using the concept of gauge equivalence, we construct expansions over the eigenfunctions of recursion operators. This permits us to obtain the expansions for both GMV systems with arbitrary constant asymptotic values of the potential functions in the auxiliary linear problems.

Copyright
© 2018 The Authors. Published by Atlantis Press and Taylor & Francis
Open Access
This is an open access article distributed under the CC BY-NC 4.0 license (http://creativecommons.org/licenses/by-nc/4.0/).

1. Introduction

We are going to study the auxiliary linear problem

L˜ψ=(ixλS)ψ=0,S=(0uvεu*00v*00),λ (1.1)
and the theory of expansions over the adjoint solutions related to it. Above ε = ± 1 and * stands for complex conjugation. The complex valued functions u and v (the potentials) are assumed to be smooth for x and they satisfy the relation:
ε|u|2+|v|2=1. (1.2)

The potentials must also satisfy some asymptotic conditions when x → ±∞ to be discussed further in text.

The system (1.1) for ε = 1 was considered by Gerdjikov, Mikhailov and Valchev [6] with the sign + before S. It is a particular case of a vector system introduced independently by Golubchik and Sokolov [12]. For reasons that will become clear later we shall prefer this equivalent form (one just needs to change (u, v) to (−u, −v)).

Let us adopt the following convention: if ε appears in formulas then it will be + 1 or − 1, if it is used as an index or label it will mean either + or −. Thus we shall denote the original Gerdjikov-Mikhailov−Valchev system by GMV+ while the general case will be referred to as GMVε system.

According to [6, 7] the GMV+ system arises naturally when one looks for integrable system having a Lax representation [L, A] = 0 with L and A subject to Mikhailov-type reduction requirements, see [1618]. This is also true for both GMV± systems. Indeed, the Mikhailov reduction group G0 acting on the fundamental solutions of the system (1.1) could be defined as generated by the two elements g1 and g2 acting in the following way:

g1(ψ)(x,λ)=[Qεψ(x,λ*)Qε]1,Qε=diag(1,ε,1),g2(ψ)(x,λ)=Hψ(x,λ)H,H=diag(1,1,1) (1.3)
where † denotes Hermitian conjugation. Since g1 g2 = g2 g1 and g12=g22=id , G0=2×2 . We shall call the reduction defined by g1, g2 for ε = 1 Hermitian while the reduction defined by g1, g2 for ε = −1 pseudo-Hermitian. The requirement that G0 is a reduction group immediately gives that the coefficients of the operators L˜ and
A˜=it+k=0nλkA˜k,A˜ksl(3,) (1.4)
must satisfy:
HSH=S,HA˜kH=(1)kA˜k,QεSQε=S,QεA˜kQε=A˜k. (1.5)

One may consider a more general form of pseudo-Hermitian reduction, i.e. one with Qε1ε2=diag(1,ε1,ε2),ε12=ε22=1 . However, it is easily checked this does not give anything new compared to the pseudo-Hermitian reduction under consideration here.

As it can be checked the matrix S has constant eigenvalues. We have g−1Sg = J0, where g is of the form:

g=12(101εu*2vεu*v*2uv*),J0=diag(1,01). (1.6)

In case ε = + 1 the matrix g belongs to the group SU(3) (g = g−1) and when ε = −1 to the group SU(2,1) (Q gQ = g−1). Further on we shall use the general notation SU(ε) referring to both cases, i.e. it implies SU(ε) ≡ S U(3) when ε = 1 and SU(ε) SU(2, 1) when ε = −1.

Since g(x) ∈ SU(ε), the values of S(x) will be in the orbit 𝒪J0(SU(ε)) of J0 with respect to SU(ε) (it is a submanifold of isu (ε)). Thus S(x)𝒪J0(SU(ε))𝔤1 where 𝔤1 is the space of the matrices X in sl (3,) such that HXH = − X, see (2.6) for the reason for this notation. Let us also note that conversely, if we assume that S(x)𝒪J0(SU(ε))𝔤1 then on the first place S has the form as in (1.1). Next, as easily checked, the eigenvalues of the matrix S are

μ1=0,μ2=μ3=ε|u|2+|v|2.

But since they coincide with 0,±1 we must have ε|u|2 + |v|2 = 1.

Our approach to the GMV± system will be based on the fact that it is gauge equivalent to a generalized Zakharov-Shabat auxiliary systems (GZS systems) on the algebra sl (3,) , see Sec. 3. Generalized Zakharov-Shabat systems, called Caudrey-Beals-Coifman (CBC) systems [2] when J is complex, are probably the best known auxiliary linear problems. In their most general form they are written on an arbitrary fixed simple Lie algebra 𝔤 in some finite dimensional irreducible representation and have the form:

Lψ=(ix+q(x)λJ)ψ=0. (1.7)

Here, q(x) and J belong to 𝔤 in some fixed representation, ψ belongs to the corresponding group. The element J must be such that the kernel of adJ (adJ(X)[J,X],X𝔤) is a Cartan subalgebra 𝔥J of 𝔤 . The potential q(x) belongs to the orthogonal complement 𝔥J of 𝔥 with respect to the Killing form:

X,Y=tr(adXadY)X,Y𝔤. (1.8)

It is assumed that q(x) is sufficiently smooth and it converges to 0 fast enough when x → ±∞. The system (1.7) is called GZS (CBC) system over 𝔤 in canonical gauge. When the algebra is understood from the context and as a representation is chosen the canonical one it is just called GZS (CBC) system. The system (1.7) is gauge equivalent to

L˜ψ˜=(ixλS(x))ψ˜=0,S(x)𝒪J. (1.9)

Usually, it is assumed that limx → ±∞S(x) = J where the convergence is sufficiently fast. Here, 𝒪J stands for the orbit of J under the adjoint action of the group G corresponding to 𝔤:𝒪J={X˜:X˜=gJg1,gG} . The concept of gauge transformation, gauge equivalent auxiliary problems and gauge equivalent soliton equations originates from the famous work [27] where it has been employed to solve an equation that is a classical analogue of equations describing waves in magnetic chains (spin 1 / 2). It has been shown that one of the nonlinear evolution equations (NLEEs) related to L˜ is the Heisenberg ferromagnet equation

St=12i[S,Sxx],Sisu(2),S2=1,limx±S(x,t)=diag(1,1)
being gauge equivalent to the famous nonlinear Schrödinger equation, see [4] for an extensive discussion on that issue. It should be mentioned that the soliton equations solvable through the auxiliary linear problem L˜ψ˜=0 in the case sl (3,) is related to a classical analog of the equation describing spin 1 particle chains dynamics, see [3].

Later, the results of [27] were generalized to the soliton equations hierarchies associated with L and L˜ , the conservation laws of those NLEEs, the hierarchies of their Hamiltonian structures etc. This was achieved by generalizing the so-called AKNS approach [1] (generating operators or recursion operators approach). Initially, this was done in the sl (2,) case, next in the case of GZS system on arbitrary semisimple Lie algebra [5, 9, 19]. Now, this theory is referred to as the gauge-covariant theory of the recursion operators related to the GZS (CBC) systems in canonical and pole gauge. For a detailed explanations of all these issues and more references (prior to 2008), see the monograph [8].

For GZS system in pole gauge most of the essential issues could be reformulated from the canonical gauge. Perhaps the main difficulty is to express explicitly all quantities depending on q and its derivative through S and its derivatives. While there is a clear procedure how to achieve that goal, in each particular case the details could be different. The procedure has been developed in detail in the PhD thesis [19]. In the case of sl (3,) in general position (with no reductions) it has been carried out in [20].

Regarding now the system GMV±, it was the Hermitian case, i.e. GMV+, that was mainly considered so far. The first paper in which GMV+ was studied in detail for the asymptotic conditions limx → ±∞ u = 0, limx → ±∞ v = expiΦ± was [7]. It contained discussion of the spectral properties of the GMV+ system, two operators whose product play the role of a recursion operator were presented and expansions over the so-called adjoint solutions were derived. We shall also discuss these issues, however, we want to stress on the following: a) We shall be dealing with both GMV± systems simultaneously; b) Our approach will be completely different, based on the gauge equivalence we mentioned above. Consequently, we shall be able to consider general asymptotic conditions constant limits limx → ±∞ u and limx → ±∞ v; c) Our point of view on the recursion operators when reductions are present is somewhat different from that adopted in [7].

2. Algebraic preliminaries

In order to proceed further, we shall need some information about the algebra sl (3,) and the involutions we are going to use which we introduce below. All facts from the theory of the semisimple Lie algebras we use are classical, see [13]. We also use the normalizations adopted in this monograph and most of its notation.

The Lie algebra sl (3,) is a simple Lie algebra of rank 2. We shall denote its Killing form tr (adXadY) by ⟨X, Y⟩ where as usual adX(Y) ≔ [X, Y]. It is known that tr (adXadY) = 6 tr (XY) which simplifies considerably the calculations. A Cartan subalgebra could be introduced using any regular element Xsl(3,) and constructing the space 𝔥X=keradX . As it is known, the subalgebra of the diagonal matrices is a canonical choice for Cartan subalgebra. It is also equal to 𝔥=keradJ where J is any diagonal matrix diag (λ1, λ2, λ3) with distinct λi. In that case, we shall call 𝔥 the Cartan subalgebra. For the canonical choice of the Cartan subalgebra the system of roots Δ for sl (3,) is

Δ={αi,j=εiεj,ij,i,j=1,2,3} (2.1)
where εi are functionals acting on 𝔥 in the following way: εi(diag (h1, h2, h3)) = hi. Then the set of positive roots Δ+ consists of the elements αi:
α1=ε1ε2,α2=ε2ε3,α3=ε1ε3=α1+α2. (2.2)

The corresponding root vectors Eα, α ∈ Δ together with the matrices Hα1,Hα2 written below:

Eα1=16e12,Eα2=16e23,Eα3=16e13,Eα1=16e21,Eα2=16e32,Eα3=16e31,Hα1=16(e11e22),Hα2=16(e22e33) (2.3)
form the Cartan-Weil basis of sl (3,) associated with the Cartan subalgebra 𝔥 . Here as usual eij means a matrix whose only nonzero entry equal to one is located in the intersection of the i-th row and j-th column. The matrices Hα1,Hα2 span 𝔥 and the matrices Eα span 𝔥 .

The relations (1.5) have the following Lie algebraic meaning. First, the map h: XHXH = HXH−1 is involutive automorphism of sl (3,). Next, Qε defines a complex conjugation σε of sl (3,) :

σε(X)=QεXQε. (2.4)

The complex conjugation σε defines the real form su (3) (ε = + 1) or the real form su (2,1) (ε = − 1) of sl (3,) . In order to treat both cases simultaneously we shall adopt the notation su (ε) meaning su(3) when ε = + 1 and su (2,1) when ε = − 1. Note that the complex conjugation σε commutes with the automorphism h. Now, let us introduce the spaces:

𝔤j={X:h(X)=(1)jX},j=0,1. (2.5)

Then we shall have the splittings

sl(3,)=𝔤0𝔤1su(ε)=(𝔤0su(ε))(𝔤1su(ε)). (2.6)

The space 𝔤1 consists of all off-diagonal matrices X = (xij) for which x23 = x32 = 0 and 𝔤0 consists of all traceless matrices X = (xij) for which x12 = x21 = x13 = x31 = 0. Of course, S involved in the GMV± satisfies S𝔤1 . Also, since h is automorphism, the spaces 𝔤0 and 𝔤1 are orthogonal with respect to the Killing form.

Below we shall see that h we already introduced (Xh(X) = HXH) is closely related to the automorphism Xk(X) = KXK where

K=(001010100). (2.7)

So we shall need how h and k act on the Cartan-Weil basis. Note that similar to h the automorphism k is also involutive, that is k2 = id. For h we get that

h(E±α1)=E±α1,h(E±α2)=E±α2,h(E±α3)=E±α3,h(H1)=H1,h(H2)=H2. (2.8)
while for k we have
k(E±α1)=Eα2,k(E±α3)=Eα3,k(E±α2)=Eα1,k(H1)=H2,k(H2)=H1. (2.9)

The above formulas become simpler if we introduce the action 𝒦 of k on the roots:

𝒦(±α1)=α2,𝒦(±α3)=α3,𝒦(±α2)=α1. (2.10)

We observe that 𝒦 maps positive roots into negative and vice-versa. Of course, 𝒦 just as k satisfies 𝒦2=id . There is no need to introduce the action of h on the roots since this is simply the identity. With the above notation we have

k(Eα)=E𝒦α,h(Eα)=r(α)Eα (2.11)
where r(α) = 1 if α = ± α1, ± α3 and r(α) = −1 if α = ± α2. Note that r(α)r(−α) = 1.

The invariance under the group generated by g1, g2 means that if ψ is the common G0-invariant fundamental solution of the linear problem

L˜ψ˜=(ixλS(x))ψ˜=0,S(x)𝒪J0sl(3,) (2.12)
and the linear problem of the type:
A˜ψ˜=itψ˜+k=0nλkA˜kψ˜=0,A˜ksl(3,) (2.13)
we must have S𝔤1isu(ε) and
A˜2k+1𝔤1isu(ε),A˜2k𝔤0isu(ε),k=0,1,2, (2.14)

The conditions on A˜k and S coincide with (1.5), in particular, we have that h(S) = −S and σε(S) = −S. This forces S to be in the form we introduced in (1.1). Also, as a direct consequence from the last relations we obtain that

hadS=adSh,σεadS=adSσε. (2.15)

Consequently, the spaces keradS=𝔥S (which obviously is a Cartan subalgebra) and its orthogonal complement 𝔥S are invariant under h. Each of them splits into two h-eigenspaces (for the eigenvalues +1 and −1 respectively):

𝔥S=𝔣0𝔣1,𝔥S=𝔥0𝔥1. (2.16)

Several consequences follow from the above:

  1. (1)

    adS and adS1 interchange the spaces 𝔣0 and 𝔣1 :

    adS𝔣0=𝔣1,adS𝔣1=𝔣0,adS1𝔣0=𝔣1,adS1𝔣1=𝔣0. (2.17)

  2. (2)

    Since S𝔤1,S1=S22361𝔤0 , the spaces 𝔥0,𝔥1 are 1-dimensional and are spanned by S and S1=S2231 respectively.

  3. (3)

    From the previous item follows that

    Sx𝔤1,(S1)x=(S2)x𝔤0 (2.18)
    and since Sx and (S1)x are orthogonal to 𝔥S we have
    Sx𝔣1,S1x(S1)x=(S2)x𝔣0. (2.19)

Another issue we must discuss is the relation between h from one side and adS1 and πS from the other. Here πS is the orthogonal projection on the space 𝔥S with respect to the Killing form and adS1 is defined only on the space 𝔥S . As one can show, see [7, 22], adS1 could be expressed as a polynomial p3(adS) in adS where only odd degrees of adS are present ( adL11 through adL1 in the notation of [7]). From the other side, this means that πS could be expressed as a polynomial in adS where enter only the even powers of adS. Consequently, one obtains that

adS1h=hadS1,πSh=hπS (2.20)
and in the same way
adS1σε=σεadS1,πSσε=σεπS. (2.21)

Remark 2.1.

If we understand adS1 as polynomial p3(adS) it is clear that instead of adS1πS one can simply write adS1 which is usually done without mentioning.

For convenience we formulate these facts as a proposition:

Proposition 2.1.

The projection πS commutes with h and σε.

3. The GMV± system and its gauge equivalent

It has been pointed out [7] that the GMV+ system is gauge equivalent to a GZS type system. For the system GMV± the situation is the same. After the gauge transformation L˜L=g1(x)L˜g(x) with g(x) given in (1.6) we obtain the system

Lψ=(ix+qλJ0)ψ=0 (3.1)
where q′ = ig−1 gx. After expressing g with u and v, we obtain
q=i2(εuux*+vvx*2(uvxvux)εuux*+vvx*2ε(v*ux*u*vx*)2(εu*ux+v*vx)2ε(v*ux*u*vx*)εuux*+vvx*2(uvxvux)εuux*+vvx*). (3.2)

As it should be for GZS system J0 is a real diagonal matrix and its diagonal elements are ordered to decrease going down along the main diagonal. This is in fact our motivation when we choose the sign in front of S. As one could see, in the above q′ is not off-diagonal as GZS requires and in [7] authors just stopped here. It turns out, however, that the issue could be easily addressed making one more gauge transformation. Indeed, due to the condition ε|u|2 + |v|2 = 1 the diagonal part of q′ equals i2(εuux*+vvx*)J=b(x)J where J′ = diag (1, −2, 1). Moreover, the diagonal part is real, that is b(x)=i2(εuux*+vvx*) is real. Then the gauge transformation

LL=(exp(iJxb(y)dy))L(exp(iJxb(y)dy)) (3.3)
maps the spectral problem into
Lψ=(ix+qλJ0)ψ=0 (3.4)
where
q(x)=exp[iJxb(y)dy]q(x)exp[iJxb(y)dy].

One can easily see that q is off-diagonal and since the entries of q′ decay when x → ±∞ the entries of q also decay. Now, since the diagonal elements of J0 are distinct we conclude that GMV± is a problem of GZS type in pole gauge on sl (3,) . We shall exploit this fact but we want to address first some other issues. The first one is whether we can be more precise about the potential q in the GZS system (3.4) to which GMV± is equivalent. With direct calculation we are able to establish that the potential q in (3.4) has special properties, namely QεqQε = q and KqK = q where K has been introduced in (2.7). We also notice that KJ0K = − J0 and QεJ0Qε=J0 .

The next issue is about the asymptotic conditions S must satisfy. Passing from GZS system in canonical gauge (3.4) to GZS in pole gauge (1.9) usually involves the Jost solution ψ0 of (3.4) such that limxψ0(x)=1 at λ = 0. The gauge transformation is then ψψ˜=ψ01ψ . In that case we obtain S=ψ01J0ψ0 and limxS(x)=J0. However, this is not our case and such a condition is not compatible with the reductions. Natural for the GMV± system are conditions when u, v tend to some constant values. In this work we shall assume that limx±u(x)=u±,limx±v(x)=v± . In that case the limit of g when x → ±∞ is equal to

g±=12(101εu±*2v±εu±*v±*2u±v±*). (3.5)

Now we notice the following. If we come back to how q was constructed, we shall see that the function

ψ0=eiJxb(y)dyg1 (3.6)
(for the definitions of g and b see the explanations after (3.2)) is as a matter of fact a solution to the system (3.4) for λ = 0. It satisfies:
limxψ0=g1,limx+ψ0=exp(iΦJ)g+1 (3.7)
where
Φ=+b(y)dy(mod2π). (3.8)

Remark 3.1.

One can easily show that b(x) is a conservation law density for all the NLEEs associated to the GMV± system. Thus Φ is a conservation law and does not depend on time. Moreover, as it is well-known the conservation laws densities are local, this shows

B(x,t)=xb(y,t)dy
is a local expression (depends on x, t only through u, v and its derivatives).

So we can make a gauge transformation where instead of ψ0 we use ψ0. Since J0 and J′ commute, we obtain that S(x)=ψ01J0ψ0 satisfies

limxS(x)=gJ0g1=S,limx+S(x)=g+J0g+1=S+ (3.9)
S±=(0u±v±εu±*00v±*00), (3.10)
i.e. we have the correct asymptotic for S(x). The last remark, which is in fact quite important, is that taking the explicit form of ψ0, see (3.6), one gets the following important formula:
Kψ0=ψ0H,ψ01K=Hψ01 (3.11)
where H was introduced earlier in relation to the GMV± system and K is given in (2.7). One could also obtain the same result in another way. Suppose one starts with the GZS system and assume it has the property that if ψ0 is a fundamental solution, then 0K also is. Then naturally 0 is also a fundamental solution. So we must have 0 = ψ0P where P is some non-degenerate matrix. Taking x → −∞ we get that Kg1=g1P . But, since Kg1=g1H we have P = H. Finally, we formulate the results of the above discussions as a theorem:

Theorem 3.1.

The GMV± system is gauge equivalent to a canonical GZS linear problem on sl (3,)

Lψ=(ix+qλJ0)ψ=0 (3.12)
subject to a Mikhailov reduction group generated by the two elements h1 and h2. These elements act on the fundamental solutions ψ of the system (3.12) in the following way:
h1(ψ)(x,λ)=[Qεψ(x,λ*)Qε]1,Qε=diag(1,ε,1),ε=±1,h2(ψ)(x,λ)=Kψ(x,λ)K. (3.13)

Since h12=h22=id and h1 h2 = h2 h1 we have again a 2×2 reduction.

The properties of the fundamental solutions of the auxiliary systems play a paramount role in the spectral theory of such systems and for the theory of integration of the NLEEs associated with them. Since for the GZS systems with p reductions these questions have been studied in detail, see [11] (whether or not K is Coxeter automorphism does not make any difference here) the above theorem opens the possibility to study the GMV± system using its gauge-equivalent GZS system with 2×2 reduction. In [11], it was considered much more complicated case of a CBC system and the automorphism has order p > 1 (in our case p = 2 since K2 = 1). Then the complex plane is divided into p sectors by straight lines through the origin and in each sector there is a fundamental analytic solution to the corresponding CBC system. In our case the things are much simpler: we have the real line dividing into upper and lower half-plane. Then the system (3.12) possesses fundamental analytic solutions (FAS) χ±(x, λ) of the form m±(x, λ) exp (−i λxJ0). The functions m+ (x, λ) (m(x, λ)) are meromorphic in λ in the upper (lower) half plane +() . Naturally, χ±(x, λ) have the same analytic properties in λ as m±(x, λ). The points of that belong to the discrete spectrum are those at which m±(x, λ) and their inverse have singularities with respect to λ in +() . It is usually considered the case when these singularities are poles, see for example [10] in the general case and [11] for the case of reductions. There is no difficulty to include the discrete spectrum for the GMV± in our considerations but this will make all the formulas much more complicated so we shall do it elsewhere. Here we intend to explain mainly how our approach works and most of the issues are algebraic. Thus, in what follows we shall assume that there is no discrete spectrum for the GZS system gauge equivalent GMV± and consequently no discrete spectrum for GMV±.

Continuing with the properties of m±(x, λ), if the potential q(x) has integrable derivatives up to the n-th order then

m±(x,λ)=1+i=1nai(x)λi+o(λn) (3.14)
when |λ| → ∞ and λ remains in + ( respectively). The asymptotic is uniform in x , and the coefficients ai(x) could be calculated through q and its x-derivatives. In particular, for absolutely integrable q we have lim|λ|m±(x,λ)=1 . The functions m± allow extensions by continuity to the real line, these extensions will be denoted by the same letters. In addition, m± satisfy:
Km±(x,λ)K=m(x,λ),limxm±(x,λ)=1. (3.15)

Since KJ0K = −J0, the solutions χ± also allow extensions by continuity to the real line (denoted again by the same letters) and satisfy:

Kχ±(x,λ)K=χ(x,λ). (3.16)

It is simple to see that in our case χ± will also satisfy:

Qε(χ±)(x,λ)Qε=(χ(x,λ*))1. (3.17)

We build from χ±(x, λ) fundamental analytic solutions χ˜±(x,λ) of the system (2.12) by setting:

χ˜±(x,λ)=ψ01χ±(x,λ). (3.18)

Now we have:

Theorem 3.2.

The solutions χ˜±(x,λ) satisfy a:

Qε(χ˜±(x,λ*))Qε=(χ˜(x,λ))1,Hχ˜±(x,λ)H=χ˜(x,λ)KH. (3.19)

Proof. The proof of the first statement is straightforward, one needs to see only that QεJ0Qε=J0 . The second one is obtained if one takes into account (3.11):

Hχ˜±(x,λ)=Hψ01χ±(x,λ)=ψ01Kχ±(x,λ)=ψ01Kχ±(x,λ)K2=ψ01χ(x,λ)K=χ˜±(x,λ)K.

Finally, we note that we have the same type of asymptotic behavior both for GMV+ and GMV which allowed to treat both cases simultaneously. Of course, the solutions we speak of are different for different choices of ε but we denote them by the same letter since it will not cause ambiguities. As it is well known, the asymptotic of χ±(x, λ) when x → ±∞ are of the type a±±exp(iλJ0x) . So the asymptotic of χ˜±(x,λ) are of the type ga±exp(iλJ0x) when x → −∞ and g+expi(ΦJ)a+±exp(iλJ0x) when x → +∞.

4. Recursion operators for the GMV± system

The recursion operators (Λ-operators) are theoretical tools that permit to describe [8]:

  • The hierarchies of the NLEEs related to the auxiliary linear problem.

  • The hierarchies of conservation laws for these NLEEs.

  • The hierarchies of compatible Hamiltonian structures of these NLEEs.

The issue about the Hamiltonian structures is closely related to a beautiful interpretation of the recursion operators originating from the work [15]. In fact, the adjoint of the recursion operators could be interpreted as Nijenhuis tensors on the infinite dimensional manifold of ‘potentials’ (the functions S(x)), see [8].

Recursion operators arise naturally when one tries to find the hierarchy of Lax pairs related to a particular auxiliary linear problem [28]. So suppose we want to find the NLEEs having Lax representation [L˜,A˜]=0 with L˜=ixλS as in (2.12) and

A˜=it+k=0nλkA˜k,A˜n𝔥S. (4.1)

The first thing we notice is that we must have ixA˜n𝔥S and A˜0 = const in x. Using a gauge transformation depending only on t one can achieve A˜0=0 . Then one can show that the coefficients A˜k for k = 1, 2,…n−1 are calculated recursively. In order to see it we first recall that the elements S and S1=S2231 span 𝔥S=keradS and that

S,S=12,S1,S1=4,S,S1=0. (4.2)

As before, denote by πS the orthogonal projection (with respect to the Killing form) onto the space 𝔥S . Then the orthogonal projection onto 𝔥S will be 1πS and for Xsl(3,) we shall have

(1πS)X=S12S,X+S14S1,X. (4.3)

Putting

A˜sd=(1πS)A˜s,A˜sa=πSA˜s, (4.4)
we obtain that the coefficients A˜k for k = 1, 2,…n − 1 are obtained recursively in the following way:
A˜k1a=Λ˜±A˜ka, (4.5)
A˜kd=i{S12±xA˜ka,Sydy+S14±xA˜ka,S1ydy} (4.6)
where
Λ˜±(Z˜)=iadS1πS(xZ˜+Sx12±xZ˜,Sydy+S1x4±xZ˜,S1ydy) (4.7)
and for the sake of brevity we write here and below S1x instead of (S1)x .

The operators Λ˜± are recursion (generating) operators for GMV± system. As one can see, they do not depend on the second reduction, that is on the choice of the real form. In fact, the recursion operators for the sl (3,) -GZS system in general position were already known both in canonical and pole gauge [20, 21] and one can obtain Λ˜ for GMV from those as it was described in [22]. The expressions for the recursion operators for sl (n,) -CBC in pole gauge is also known [23]. From it one also could easily find Λ˜± . Let us note that in [7] we already mentioned what is called recursion operator is in fact Λ˜±2 and it has been introduced as a product of two operators which differ from Λ˜±. This complicates the picture (we shall discuss that a little further).

Let us continue with the calculation of the coefficients A˜k. For the coefficient A˜n1 one has that ixA˜n=[S,A˜n1] and since A˜n𝔥S there are scalar functions α, β such that A˜n=αS+βS1 . This gives

iαxS+iβxS1+iαSx+iβ(S1)x𝔥S
and therefore α and β are constants. Thus
A˜n1=iαadS1Sx+iβadS1S1x
and the hierarchy of NLEEs related to L˜ in general position is
adS1tS+(Λ˜±)n(adS1(αSx+β(S1x)=0. (4.8)

In the case of GMVε system A˜n1isu(ε) so α and β must be real. Next, if A˜n𝔥1 and n = 1(mod2) because of the (4.5) and (4.6) we shall have automatically

A˜2k(𝔥1𝔣1)isu(ε),A˜2k1(𝔥0𝔣0)isu(ε). (4.9)

If A˜n𝔥0 and n = 0(mod2) we have again (4.9). Thus the Mikhailov type reductions are compatible with the action of recursion operator and the general form of the equations related to the GMVε system is:

adS1tS=k=0ra2k(Λ˜±)2kadS1(Sx)+k=1ma2k1(Λ˜±)2k1(adS1(S1x) (4.10)
where ai are some real constants. This is the hierarchy found in [7] although it was presented there in a different form.

Let us denote the space of rapidly decreasing functions Z˜(x) with values in 𝔣0(𝔣1) by 𝔣0[x](𝔣1[x]) . Since 𝔣0 and 𝔣1 are orthogonal with respect to the Killing form and taking into account Remark 2.1 we have

Z˜𝔣0[x],Λ˜±(Z˜)=iadS1(xZ˜+S1x4±xZ˜,S1ydy)𝔣1[x], (4.11)
Z˜𝔣1[x],Λ˜±(Z˜)=iadS1(xZ˜+Sx12±xZ˜,Sydy)𝔣0[x]. (4.12)

Now, let us do the following:

  • Put in the above expressions S = −L1 (in order to express everything through L1).

  • Take into account that what is called Killing form in [7] and denoted ⟨X, Y⟩ is not the canonical definition of the Killing form, it is in fact tr (XY) and as we know in our notation ⟨X, Y⟩ = 6 tr (XY).

  • Take into account that because Z˜ is orthogonal to 𝔥S we have Z˜(x),Sx=Z˜x,S(x) and Z˜(x),S1x=Z˜x,S1(x) .

Then one could convince himself that the operators Λ1±,Λ2± introduced in [7] are defined by the expressions standing in the right hand sides of (4.11) and (4.12), see [22]. In other words, Λ1±,Λ2± are restrictions of the operators Λ˜± on the spaces 𝔣0[x] and 𝔣1[x] respectively. If the hierarchy of the soliton equations (4.10) is written in terms of Λ1±,Λ2± it acquires the form that was presented in [7].

5. Completeness relations for the GMV± system

We have seen that the consideration of the NLEEs hierarchy (4.10) shows that the operator ‘moving’ the equations along the hierarchy is Λ˜±2 and it does not depend on the real form. The geometric interpretation of the hierarchies and their conservation laws which is developed for the case of GMV+ [24] also suggests that the appropriate operator we must consider is Λ˜±2 . It remains to consider the third important aspect of the recursion operators – their relations to the so-called expansions over the adjoint solutions. In that approach one considers these expansions and then finds operators, for which the adjoint solutions entering into them are eigenfunctions. The importance of the adjoint solutions is related to the fact that they form complete systems, so roughly speaking, one can expand S and its variation over them and then have the famous interpretation of the inverse scattering method as a generalized Fourier transform, see [8, 14]. Recently, the theory of the expansions over adjoint solutions has been extended to the case when there are reductions. In [11] the theory for the CBC systems with p reductions was presented in full taking into account the discrete spectrum while in [25] is discussed also the gauge-covariant formulation of the expansions with p reductions. The geometric aspects of the NLEEs associated with CBC systems in canonical and pole gauge with reductions were presented recently in [26].

Our approach will be the following. We start from the spectral theory of the generating operators for (3.12) (a GZS system in canonical gauge) which is very well known and from it obtain the theory of the generating operators for GMV± (a GZS system in pole gauge). So let us sketch the theory for GZS in canonical gauge, for all the details see [8, 20]. The adjoint solutions for the system GZS system (3.12) are defined as follows:

eα±π0(χ±Eαχ^±),αΔ (5.1)
where π0 is the orthogonal projector on 𝔥 where 𝔥 is the Cartan subalgebra of sl(3,) and χ^ stands for the matrix inverse to χSL(3,) . The generating operators Λ± then have the form:
Λ±(Y(x))=adJ01[ixY+π0adqY(x)+adq+x(1π0)adqY(y)dy]. (5.2)

As mentioned, the above formula together with its derivation could be found in many sources, see for example [5, 8, 10, 11]. From the asymptotic behaviour of the solutions χ± follows that for α > 0 we have

Λ(eα+(x,λ))=λeα+(x,λ),Λ(eα(x,λ))=λeα(x,λ)Λ+(eα+(x,λ))=λeα+(x,λ),Λ+(eα(x,λ))=λeα(x,λ). (5.3)

(It is of no importance whether or not we have some reductions). Let us define now the adjoint solutions for the GMV± in the following way:

e˜α±πS(χ˜±Eαχ˜ˆ±),αΔ. (5.4)

Let us recall now that we have a reduction defined by the automorphism h. Taking into account that h commutes with πS (see Proposition (2.1)) and the properties of ψ0 (see (3.11)) we see that h(e˜α±)(x,λ) equals

πS(Hχ˜±Eαχ˜^±H)(x,λ)=πS(χ˜KEαKχ˜^)(x,λ)=πS(χ˜E𝒦αχ˜^)(x,λ)
so we have:
h(e˜α±)(x,λ)=e˜𝒦(x,λ). (5.5)

One proves easily that πS=Ad(ψ01)π0Ad(ψ0) , and then one sees that for α ∈ Δ we have that e˜α±(x,λ)=Ad(ψ01)eα±(x,λ) . Here Ad (r) denotes the adjoint action of the group on its algebra, that is Ad (r)X = rXr−1. All the theory then develops in a similar way as in the case when we make a gauge transformation with a Jost solution which is very well known [10]. For example, one has that

Λ˜±=Ad(ψ01)Λ±Ad(ψ0). (5.6)

As a consequence, for α ∈ Δ+ we have

Λ˜(e˜α+(x,λ))=λe˜α+(x,λ),Λ˜(e˜α(x,λ))=λe˜α(x,λ),Λ˜+(e˜α+(x,λ))=λe˜α+(x,λ),Λ˜+(e˜α(x,λ))=λe˜α(x,λ). (5.7)

We are now ready to discuss the completeness of the adjoint solutions for the GZS system in pole gauge (which under reductions becomes our GMV± problem). In order to simplify the formulas we shall have further, let us adopt the following notation: for the functions X(x),Y(x):sl(3,) we put

X,Y+X(x),Y(x)dx. (5.8)

First, we write the completeness relations for the ‘adjoint ‘ solutions of the canonical gauge GZS system (details can be found in [5, 8]). We also remind that we disregard the discrete spectrum. Then for the GZS system in canonical gauge (3.12) the following theorem holds:

Theorem 5.1.

Let 𝒮 be the space of sufficiently smooth functions defined on the real axis with values in 𝔥 tending fast enough to zero when x → ±∞. Then for every Z(x)𝒮 the following expansion formulas hold (η = ±):

Z(x)=12παΔ+eαη[(x,λ)eαη,[J0,Z]eαη(x,λ)eαη,[J0,Z]]dλ. (5.9)

Let us make now a gauge transformation in the above completeness relations for the GZS system in canonical gauge, one will easily see that they transform to completeness relations for the GZS system in pole gauge (which becomes GMV± if we impose reductions):

Theorem 5.2.

Let 𝒮˜ be the space of sufficiently smooth functions defined on the real axis with values in 𝔥S(x) tending fast enough to zero when x → ±∞. Then for every Z˜(x)𝒮˜ the following expansion formulas hold (η = ±):

Z˜(x)=12π[αΔ+e˜αη(x,λ)e˜αη,[S,Z˜]e˜αη(x,λ)e˜αη,[S,Z˜]]dλ. (5.10)

Note that we have two expansions here − for η = + and η = −. One is over the eigenfunctions of Λ˜, the other is over the eigenfunctions of Λ˜+. In fact the above is the spectral theorem for these operators.

Now we assume that we have the reductions defined by h and σε and we are going to transform these expansions so that it will be easier to see the action of h and σε on these expansions when we expand functions that have values in the eigenspaces 𝔣0,𝔣1 respectively or are ‘real’ or ‘imaginary’ with respect to the complex conjugation σε.

Let us start with the reduction defined by h. We have seen that h acts on the adjoint solutions according to (5.5). Then using the invariance of the Killing form with respect to the action of h we get

+e˜αη(x,λ),[S,Z˜]dx=+h(e˜αη(x,λ)),h([S,Z˜])dx=+e˜𝒦αη(x,λ),H[S,Z˜]Hdx=+e˜𝒦αη(x,λ),[S,h(Z˜)]dx.

Therefore

e˜αη(x,λ)e˜αη(x,λ),[S,Z˜]=h(e˜𝒦αη(x,λ))+e˜𝒦αη(x,λ),[S,h(Z˜)]dx.

Suppose for example h(Z˜)=Z˜ . Then the above is simply written as:

e˜αη(x,λ)e˜αη(x,λ),[S,Z˜]=h(e˜𝒦αη(x,λ))e˜𝒦αη(x,λ),[S,Z˜].

Since 𝒦 maps positive roots into negative and vice versa, we shall have:

αΔ+e˜αη(x,λ)e˜αη(x,λ),[S,Z˜]=αΔ+h(e˜αη(x,λ))e˜αη(x,λ),[S,Z˜],αΔ+e˜αη(x,λ)e˜αη(x,λ),[S,Z˜]=αΔ+h(e˜αη(x,λ))e˜αη(x,λ),[S,Z˜].

Now, let us make the change of variables λ ↦ −λ in the integral standing in the right hand side of (5.10) and use the above relations. We obtain:

Z˜(x)=12π[αΔ+h(e˜αη(x,λ))e˜αη,[S,Z˜]h(e˜αη(x,λ))e˜αη,[S,Z˜]]dλ. (5.11)

But then combining the original expansion with this one we immediately get:

Z˜(x)=12π[αΔ+s˜αη(x,λ)e˜αη,[S,Z˜]s˜αη(x,λ))e˜αη,[S,Z˜]]dλ (5.12)
where
s˜±αη(x,λ)=12(e˜±αη(x,λ)+h(e˜±αη(x,λ))). (5.13)

In view of the formulas we shall write a little further, we also note, that one has

e˜αη,[S,Z˜]=h(e˜αη),h([S,Z˜])=h(e˜αη),[S,Z˜]
so one can write
e˜αη,[S,Z˜]=a˜αη,[S,Z˜]
where
a˜±αη(x,λ)=12(e˜±αη(x,λ)h(e˜±αη(x,λ))). (5.14)

Thus finally we obtain:

Z˜(x)=12π[αΔ+s˜αη(x,λ)a˜αη,[S,Z˜]s˜αη(x,λ)a˜αη,[S,Z˜])]dλ. (5.15)

If instead of h(Z˜)=Z˜ we assume that h(Z˜)=Z˜ then in the same manner we shall obtain

Z˜(x)=12π[αΔ+a˜αη(x,λ)e˜αη,[S,Z˜]a˜αη(x,λ)e˜αη,[S,Z˜]]dλ=12π[αΔ+a˜αη(x,λ)s˜αη,[S,Z˜]a˜αη(x,λ)s˜αη,[S,Z˜])]dλ. (5.16)

Naturally,

h(s˜±αη(x,λ))=s˜±αη(x,λ),h(a˜±αη(x,λ))=a˜±αη(x,λ). (5.17)

Thus in case h(Z˜)=Z˜ or h(Z˜)=Z˜ the expansions could be written in terms of new sets of adjoint solutions, s˜+αη(x,λ) in the first case and a˜±αη(x,λ) in the second, that reflect the symmetry of Z˜ .

Returning to the properties of the new families we introduced, we easily see that s±αη(x,λ) and a±αη(x,λ) are not eigenfunctions of Λ˜ or Λ˜+. In order to see what happens, we look how Λ˜± is related to h and σε .

Lemma 5.1.

The following relations hold:

hΛ˜±=Λ˜±h,σεΛ˜±=Λ˜±σε. (5.18)

Proof. The proof of the first relation is straightforward, let us prove the second one. First, assume that σε(Z˜)=Z˜ (the case when σε(Z˜)=Z˜ is treated in a similar way). First, we remark that σε(Sx) = − Sx and that σε(S1) = −S1, consequently σε(S1x) = −S1x. Next, we have

Z˜,Sx*=σε(Z˜),σε(Sx)=Z˜,Sx
and similarly
Z˜,S1x*=σε(Z˜),σε(S1x)=Z˜,S1x.

So in case σε(Z˜)=Z˜ (we call Z˜ real in this case) the integrands in the formula for recursion operator are purely imaginary while for σε(Z˜)=Z˜ (we then say that Z˜ is purely imaginary) these expressions will be real. Then, taking into account (2.21) we get

σεΛ˜±(Z˜)=iadS1πS(x(σεZ˜)+σεSx12±xZ˜,Sy* dy+(σεS1x)4±xZ˜,S1y* dy)=iadS1πS(x(Z˜)+Sx12±xZ˜,Sydy+S1x4±xZ˜,S1ydy)=Λ˜±(Z˜)=Λ˜±(σεZ˜).

As mentioned, one can show in a similar way that when σε(Z˜)=Z˜ we have the same relation. Finally, as each Z˜ could be written in a unique way as Z˜=Z˜1+Z˜2 where σε(Z˜1)=Z˜1,σε(Z˜2)=Z˜2 then the second relation in (5.18) is proved for arbitrary Z˜.

Because of the relations (5.7) and the above lemma, for α ∈ Δ+ one has

Λ˜(s˜α+(x,λ))=λa˜α+(x,λ),Λ˜(s˜α(x,λ))=λa˜α(x,λ),Λ˜(a˜α+(x,λ))=λs˜α+(x,λ),Λ˜(a˜α(x,λ))=λs˜α(x,λ), (5.19)
Λ˜+(s˜α+(x,λ))=λa˜α+(x,λ),Λ˜+(s˜α(x,λ))=λa˜α(x,λ),Λ˜+(a˜α+(x,λ))=λs˜α+(x,λ),Λ˜+(a˜α(x,λ))=λs˜α(x,λ) (5.20)
and one sees that the functions in the expansions when we have some symmetry with respect to h are eigenfunctions for Λ˜2(Λ˜+2) with eigenvalue λ2. This together with the fact that when recursively finding the coefficients for the pencil (4.1) one effectively uses Λ˜+2 has led to the interpretation that in case we have 2 reduction defined by h the role of the Generating Operator is played by Λ˜±2 .

Let us consider now the other reduction that we have in the GMV± system, the one defined by the complex conjugation σε. In accordance with Theorem 3.2 the FAS fulfill Qε(χ˜(x,λ))Qε=(χ˜±(x,λ*))1 so on the real axis Qε(χ˜(x,λ))Qε=(χ˜±(x,λ*))1 . Then taking into account Proposition (2.1) for α ∈ Δ

Qεe˜αη(x,λ)Qε=πSχ˜η(x,λ)QεEαQεχ˜^η(x,λ*)
and we need to take into account how Xσε(X)=QεXQε acts on the Cartan-Weil basis. One easily sees that
σ+E±α1=Eα1,σ+E±α2=Eα2,σ+E±α3=Eα3σE±α1=Eα1,σE±α2=Eα2,σE±α3=Eα3σεHαj=Hαj,j=1,2. (5.21)

In order to write down these relations in a concise way, we introduce the symbol qε(α) which is equal to −1 if ε = 1 for arbitrary α ∈ Δ but if ε = −1 then qε(α) = 1 for α = ±α1, ±α2 and qεα3) = −1. Then qε(α)qε(−α) = 1 and we have

σεEα=qε(α)Eα,σεHαi=Hαi,i=1,2. (5.22)

With the above notation Qεe˜αη(x,λ)Qε=qε(α)e˜αη(x,λ) or if one prefers expression through the complex conjugation

σε(e˜αη(x,λ))=qε(α)e˜αη(x,λ*). (5.23)

Naturally, we have

σε(e˜αη(x,λ))=qε(α)e˜αη(x,λ),λ. (5.24)

Consider now the coefficients e˜αη,[S,Z˜] introduced in the above and and let us assume that σεZ˜=Z˜. We have

e˜αη,[S,Z˜]=σεe˜αη,σε[S,Z˜]*=qε(α)e˜αη,[S,Z˜]*.

Consequently:

σε(e˜αη(λ),[S,Z˜]e˜αη(x,λ))=e˜αη(λ),[S,Z˜]e˜αη(x,λ). (5.25)

In case σεZ˜=Z˜ we obtain in a similar way

σε(e˜αη(λ),[S,Z˜]e˜αη(x,λ))=e˜αη(λ),[S,Z˜]e˜αη(x,λ). (5.26)

One could also see that in case σεZ˜=Z˜ we have

e˜αη(λ),[S,Z˜]*=σεe˜αη(λ),σε[S,Z˜]=σεe˜αη(λ),[S,Z˜].

Thus when σεZ˜=Z˜ (respectively σεZ˜=Z˜ ) the expansions (5.10) acquire the form:

Z˜(x)=12π[αΔ+(1+σε)(e˜αη(x,λ)e˜αη(λ),[S,Z˜])]dλ (5.27)
and
Z˜(x)=12π[αΔ+(1σε)(e˜αη(x,λ)e˜αη(λ),[S,Z˜])]dλ (5.28)
respectively.

In order to obtain expansions when both reductions are present, we notice that since σε and h commute, in case σεZ˜=Z˜ one has:

σε(e˜αη,[S,Z˜]s˜αη(x,λ))=e˜αη,[S,Z˜]s˜αη(x,λ),σε(e˜αη,[S,Z˜]a˜αη(x,λ))=e˜αη,[S,Z˜]a˜αη(x,λ) (5.29)
and in case σεZ˜=Z˜ we have
σε(e˜αη,[S,Z˜]s˜αη(x,λ))=e˜αη,[S,Z˜]s˜αη(x,λ),σε(e˜αη,[S,Z˜]a˜αη(x,λ))=e˜αη,[S,Z˜]a˜αη(x,λ). (5.30)

Then we obtain:

Theorem 5.3.

The following expansions hold.

  • In case σεZ˜=Z˜,hZ˜=Z˜

    Z˜(x)=14π[αΔ+(1+σε)(s˜αη(x,λ)e˜αη,[S,Z˜](λ))]dλ. (5.31)

  • In case σεZ˜=Z˜,hZ˜=Z˜

    Z˜(x)=14π[αΔ+(1+σε)(a˜αη(x,λ)e˜αη,[S,Z˜](λ))]dλ. (5.32)

  • In case σεZ˜=Z˜,hZ˜=Z˜

    Z˜(x)=14π[αΔ+(1σε)(s˜αη(x,λ)e˜αη,[S,Z˜](λ))]dλ. (5.33)

  • In case σεZ˜=Z˜,hZ˜=Z˜

    Z˜(x)=14π[αΔ+(1σε)(a˜αη(x,λ)e˜αη,[S,Z˜](λ))]dλ. (5.34)

Corollary 5.1.

In view of Lemma 5.1, we see that the functions that stay in the integrands of the above expressions, that is (1±σε)s˜αη(x,λ) and (1±σε)a˜αη(x,λ) remain eigenfunctions of Λ˜±2 with eigenvalues λ2, that is, the reduction defined by a real form does not change the recursion operators.

Let us make some final comments about the expansions we obtained. In the case GMV+ the families s˜±αη(x,λ) and a˜±αη(x,λ) as well as the relations (5.19), (5.20) were introduced in [7]. However, they were written in terms of the restrictions of Λ˜± on the spaces 𝔣0[x](𝔣1[x]) , namely through the operators Λ1±,Λ2± we mentioned already. This complicates their form and obscures their meaning. We believe that the form in which we cast them now and in relation to Lemma 5.1 is are much easier to understand. As about the expansions (5.15), (5.16) and about expansions in Theorem 5.3 they were not presented until now in their general form but only for the specific cases of expansions of the functions Sx and (S1)x. Moreover, since in [7] the theory of the GMV+ have been developed not in parallel (though by analogy) to the corresponding system in canonical gauge, the knowledge about the GZS system in canonical gauge could not be exploited in full and everything ought to be developed from the beginning. In particular, there have been considered more restricted boundary conditions than necessary, namely the case u± = 0, v± = exp (iΦ±). Consequently, all the results were obtained in less generality and with more effort.

The above expansions have big theoretical interest. They show that similar to the general position case (without reductions) the ’potential’ and its variation could be expanded over a family of functions that are eigenfunctions of the operator that appears when one recursively solves the equations for the coefficients A˜k . Of course, it is important also to know that when we have reductions the expansions could be written in a different form. Practically however, it is easier to adopt the following attitude. We could simply ’forget’ that we have reductions. In other words, the whole Gauge Covariant Theory of the Generating Operators will proceed in the usual way, with or without reductions. Basically, all the formulas will remain true, one must simply take into account the following:

  • Some additional symmetries of the scattering data (see the next Section) which lead to symmetries in the corresponding Riemann-Hilbert problem when one uses it for finding exact solutions.

  • The operators that permit to ‘move’ along the hierarchies are Λ˜±2 .

  • Some of the conservation laws and Poisson structures in the corresponding hierarchies trivialize.

For the first of these items one can see [7], as for the second and the third, according to the general theory (see [8] or [20] specifically for the case of sl(3,)) we have:

  • The NLEEs (4.10) have the following series of conservation laws:

    DB(s)=[ySy,Λ˜±s(adS1B˜y)dy]dx,B𝔥,B=const,B˜=gBg1 (5.35)
    where s = 1,2,… Now, if k(B) = − B, then adS1B˜x𝔥0 and if s is even Λ˜±s(B˜x)𝔥0[x] . Since Sy𝔥1[x] we have DB(s)=0 . This of course does not happen if s is odd. Analogously, DB(s)=0 if s is odd and k(B) = B. Thus indeed some of the conservation laws trivialize.

  • The NLEEs (4.10) are Hamiltonian with respect to the hierarchy of symplectic forms:

    Ω˜(p)(δS1,δS2)=δS1,Λ˜padS1δS2dx,p=0,1,2, (5.36)
    where δ1S, δ2S are some variations of S and
    Λ˜=12(Λ˜++Λ˜). (5.37)

    Now, since ⟨S, S⟩ = 12, δ1S, δ2S are orthogonal to S and since h(S) = −S we have h(δ1S) = −δ1S, h(δ1S) = −δ1S so finally δ1S,δ2S𝔣1[x] . Further, adS1δ2S𝔣0[x] and therefore, if p is even Ω˜(p) is identically zero. The simplest case is of course the case p = 0 that corresponds to the Kirillov Poisson structure adS (restricted to the manifold of the potentials) which obviously trivializes. On the contrary, for p odd, these structures are not trivial. The symplectic structure for p = 1 corresponds to the Poisson tensor i∂x (restricted to the manifold of the potentials) and does not trivialize, see [24] for the details.

6. Application. Scattering data

As an application we are going to discuss the implications of the reduction over the scattering data related to the sl (3,) -GZS system in canonical and in pole gauge. First of all, as well-known from the theory of the CBC (GZS) systems, there are essentially two types of scattering data: a) associated with the asymptotic of FAS b) associated with the coefficients in the expansions over the adjoint solutions. All these sets of scattering data are equivalent, but we are not going to enter into this issue here.

6.1. Scattering data related to the asymptotics

As it follows from the general theory of CBC (GZS) systems, the functions m±(x, λ) through which we introduced the FAS for (3.12) in case of the real λ have asymptotics, see for example [5] or [20]:

m+(x,λ)=eiλJ0xχ+(x,λ)S+(λ),x,m+(x,λ)=eiλJ0xχ+(x,λ)T(λ)D+(λ),x+,m(x,λ)=eiλJ0xχ(x,λ)S(λ),x,m(x,λ)=eiλJ0xχ(x,λ)T+(λ)D(λ),x+. (6.1)

The matrices S+, T+ are upper triangular with with diagonal elements equal to 1, the matrices S, T are lower triangular with with diagonal elements equal to 1 and the matrices D± are diagonal with determinant 1. Besides, D+ (λ) could be defined in the upper half-plane + while D (λ) could be defined in the lower half-plane . They are meromorphic and the poles correspond to the discrete spectrum. As agreed, we shall assume that there is no discrete spectrum, so D+ (λ) (D(λ)) are analytic in +() respectively.

Since on the real line both χ±(x, λ) exist, there is a non−degenerate matrix R(λ) such that χ+ (x, λ) = χ (x, λ)R(λ) and we get that

R(λ)=S^(λ)S+(λ)=D^(λ)T^+(λ)T(λ)D+(λ),λ.

One could cast this in an equivalent form, introducing a matrix T(λ)

T(λ)=T(λ)D+(λ)S^+(λ)=T+(λ)D(λ)S^(λ),λ. (6.2)

The matrices S+, T+ are upper-triangular with diagonal elements equal to 1, the matrices S, T are lower triangular with diagonal elements equal to 1 and the matrices D± are diagonal. The above shows that the matrices we just introduced are factors of two Gauss decompositions of the matrix T(λ), which is called the transition matrix. We have the following representations of the Gauss factors:

S+(λ)=expαΔ+sα+(λ)Eα,S(λ)=expαΔ+sα(λ)Eα,D+(λ)=expi=12di+(λ)Hi,D(λ)=expi=12di(λ)Hi,T+(λ)=expαΔ+tα+(λ)Eα,T(λ)=expαΔ+tα(λ)Eα. (6.3)

As it is known, see [5, 8, 10] each one of the families of functions

s={sα+(λ),sα(λ),λ,αΔ+},t={tα+(λ),tα(λ),λ,αΔ+} (6.4)
could be taken as a scattering data, from which one could reconstruct the potential q (and hence the potential function S). We have reductions and they impose some restrictions on the scattering data. To start with, since Kχ±(x,λ)K=χ(x,λ), we get that
KS+(λ)K=S(λ),KT(λ)K=T+(λ),KD+(λ)K=D(λ). (6.5)

Using the quantities introduced in (6.3), the above relations could be written in an equivalent form: for α ∈ Δ+

sα(λ)=s𝒦α+(λ),tα(λ)=t𝒦α+(λ),sα+(λ)=s𝒦α(λ),tα+(λ)=t𝒦α(λ),d1+(λ)=d2(λ),d2+(λ)=d1(λ) (6.6)
where 𝒦 is the action of the automorphism K on the roots. The second involution leads to the relations:
Qε(S+(λ))Qε=(S(λ))1,Qε(T+(λ))Qε=(T(λ))1,(D+(λ))=(D(λ))1, (6.7)
(sα+(λ))*=qε(α)sα(λ),(tα+(λ))*=qε(α)tα(λ),αΔ+(di+(λ))*=di(λ),i=1,2 (6.8)
where in all the above relations λ is real.

Now it is not hard to find the corresponding relations for the GMV± system. Indeed, because of (3.18) we get

eiλJ0xg1χ˜+(x,λ)S+(λ),x,eiλJ0xg+1χ˜+(x,λ)eiΦJT(λ)D+(λ),x+,eiλJ0xg1χ˜(x,λ)S(λ),x,eiλJ0xg+1χ˜(x,λ)eiΦJT+(λ)D(λ),x+. (6.9)

Since J′ = H1H2 we easily get that

eiΦJT(λ)D+(λ)=T˜(λ)D˜+(λ),eiΦJT+(λ)D(λ)=T˜+(λ)D˜(λ) (6.10)
where
D˜+(λ)=eiΦJD+(λ),D˜(λ)=eiΦJD(λ),T˜±(λ)=eiΦJT±(λ)eiΦJ. (6.11)

We also put S˜±(λ)=S±(λ) . Since KJK = J the factors D˜±(λ),S˜±(λ),T˜+(λ) obey the same relations (6.5) as the factors D±(λ),S±(λ),T+(λ) . Also, we can write

D˜+(λ)=expi=12d˜i+(λ)Hi,D˜(λ)=expi=12d˜i(λ)HiT˜±(λ)=expαΔ+t˜±α±(λ)E±α,S˜±(λ)=expαΔ+s˜±α±(λ)E±α (6.12)
where s˜±α=s±α and
d˜1±=d1±+iΦ,d˜2±=d2±iΦ,t˜±α±=t±α±exp±iα(J)Φ. (6.13)

So setting S˜±(λ)=S±(λ) and s˜±α±=s±α± one could put (6.9) in a more symmetric form

eiλJ0xg1χ˜+(x,λ)S˜+(λ),x,eiλJ0xg+1χ˜+(x,λ)T˜(λ)D˜+(λ),x+,eiλJ0xg1χ˜(x,λ)S˜(λ),x,eiλJ0xg+1χ˜(x,λ)T˜+(λ)D˜(λ),x+. (6.14)

One can see that again for α ∈ Δ+

s˜α(λ)=s˜𝒦α+(λ),t˜α(λ)=t˜𝒦α+(λ),s˜α+(λ)=s˜𝒦α(λ),t˜α+(λ)=t˜𝒦α(λ),d˜1+(λ)=d˜2(λ),d˜2+(λ)=d˜1(λ) (6.15)
and we get that the families
˜s={s˜α+(λ),s˜α(λ),λ,αΔ+},˜t={t˜α+(λ),t˜α(λ),λ,αΔ+} (6.16)
could be considered as scattering data for the GMV± system. Taking into account that Φ is integral of motion for the NLEEs associated to the GMV± system (see Remark 3.1) we notice that passing from quantities without tilde to quantities with tilde is essentially the same thing, that is S could be reconstructed using the same scattering data as the scattering data for q, something that of course could be expected.

6.2. Scattering data related to the expansions over the adjoint solutions

Now let us turn our attention to the scattering data associated with the expansions over the adjoint solutions of the GMV±. First we make the following observation. Suppose B=const𝔥 . Then the function B˜=gBg1 obviously satisfies [S,B˜]=0, so B˜ takes values in 𝔥S . Moreover, we have that B˜,S=B,J0 = const and B˜,S1=B,J1 = const where J1=J02231 . But J0 and J1 span 𝔥, so B is of the form B = a0J0 + a1J1 where a0,a1 are some numbers. Then B˜=a0S+a1S1. From the other side, we have seen that Sx and S1,x are orthogonal to 𝔥S, so we see that B˜x is also orthogonal to 𝔥S , that is, belongs to 𝔥S. Take now any fundamental solution χ˜ and a any root vector Eα. We have

B˜x,χ˜Eαχ˜1=B˜x,πS(χ˜Eαχ˜1)=[S,adS1B˜x],πS(χ˜Eαχ˜1).

From the other side, since ix(χ˜Eαχ˜1)=λ[S,χ˜Eαχ˜1] , we have that

x(B˜,χ˜Eαχ˜1)=B˜x,χ˜Eαχ˜1.

Combining these relations we obtain that

[S,adS1B˜x],πS(χ˜Eαχ˜1)dx=B˜,χ˜Eαχ˜1|,λ,αΔ (6.17)
provided the integral exists, that is the right hand exists. Putting in the above formulas χ˜=χ˜± we see that we can calculate the coefficients in the expansions from Theorem 5.2 for the functions adS1B˜x. We present the result of these calculations as a theorem.

Theorem 6.1.

The following expansion formulas hold:

 a) adS1B˜x=12π[αΔ+(ρ˜B,α+e˜α+ρ˜B,αe˜α)]dλ b) adS1B˜x=12π[αΔ+(σ˜B,α+e˜α+σ˜B,αe˜α)]dλ. (6.18)

The coefficients ρ˜ and σ˜ (of course we have αΔ+ and λ ) are equal to

ρ˜B,α±=(S˜±)1BS˜±,Eα=(S±)1BS±,Eα,σ˜B,±α±=(D˜±)1(T˜±)1BT˜±D˜±,E±α=(D±)1(T±)1BT±D±,E±α. (6.19)

The functions

ρB,α±=(S±)1BS±,Eα,σB,±α±=(D±)1(T±)1BT±D±,E±α (6.20)
are exactly the coefficients one obtains expanding iadJ01[B,q] over the adjoint solutions of (3.12) see for example [5, 20]. Thus we have
ρ˜B,α±=ρB,α±,σ˜B,±α±=σB,±α±. (6.21)

It is known that provided that B is a regular element, the families

ρ={ρB,α+(λ),ρB,α(λ):λ,αΔ+},σ={σB,α+(λ),σB,α(λ):λ,αΔ+} (6.22)
could also be taken as sets of scattering data for GZS (3.12), that is using it one can construct the potential q. Then the families
˜ρ={ρ˜B,α+(λ),ρ˜B,α(λ):λ,αΔ+},˜σ={σ˜B,α+(λ),σ˜B,α(λ):λ,αΔ+} (6.23)
could be treated as scattering data for the GMV± system. In particular, if one chooses B = J0 and B = J1 one will obtain expansions for the functions adS1Sx and adS1S1x. However, depending on the properties of B the scattering data have different symmetries. Indeed, suppose that h(B˜)=ηB˜ , where η=±1. This as easily seen is equivalent to k(B) = ηB. Then of course we shall have also h(B˜x)=ηB˜x and
[S,adS1B˜x],πSχ˜±(x,λ)Eα(χ˜±)1(x,λ)=η[S,adS1B˜x],πSχ˜E𝒦α(χ˜)1(x,λ).

Consequently, for α ∈ Δ+ we have

ρ˜B,α+(λ)=ηρ˜B,𝒦α(λ),ρ˜B,α(λ)=ηρ˜B,𝒦α+(λ),σ˜B,α+(λ)=ησ˜B,𝒦α(λ),σ˜B,α(λ)=ησ˜B,𝒦α+(λ). (6.24)

Since k(J0) = −J0 and k(J1) = J1, taking for example B = J0, one obtains

ρ˜J0,α+(λ)=ρ˜J0,𝒦α(λ),ρ˜J0,α(λ)=ρ˜J0,𝒦α+(λ),σ˜B,α+(λ)=σ˜J0,𝒦α(λ),σ˜J0,α(λ)=σ˜J0,𝒦α+(λ) (6.25)
and one sees that
 a) adS1S˜x=12π[αΔ+(ρ˜J0,α+s˜α+ρ˜J0,αs˜α)]dλ, b) adS1S˜x=12π[αΔ+(σ˜J0,α+s˜α+σ˜J0,αs˜α)]dλ (6.26)
as it should be according to the theory we developed. If we expand adS1S1x then the expansion could be reformulated as expansion over the functions a˜α± .

7. Conclusions

In the present work we have been able to do the following:

  • We have showed that the results for the GMV = GMV+ system about the expansions over the adjoint solutions could be generalized in two directions: i) To apply for the system GMV-corresponding to a pseudo-Hermitian reduction of a certain sl (3,) -GZS system in pole gauge subject to 2×2 reduction of Mikhailov type. ii) Both for GMV+ and GMV we have been able to develop the theory of expansions for arbitrary constant asymptotic conditions of the potentials.

  • We have showed how the expansions over adjoint solutions should be modified if one takes into account possible symmetries of the functions we expand.

The above was achieved using the gauge equivalence between the GZS system in pole gauge and the one in canonical gauge, see Theorem 3.1. This has permitted to avoid repetition of cumbersome proofs and frequently to reduce the things to purely algebraic arguments.

Since the expansions over the adjoint solutions of an auxiliary linear problem are the main tool of the so-called recursion operator method to the soliton equations, it would be natural to develop the theory of these equations, their conservation laws, hierarchies of compatible Poisson structures etc., for the GZS system in pole gauge and in canonical gauge, making all the approach gauge-covariant. As mentioned in the introduction this task that has been successfully achieved in the case of the Heisenberg Ferromagnet hierarchy and nonlinear Schrödinger hierarchy and later generalized for systems on arbitrary simple Lie algebra with no reductions.

Also, one must consider the discrete spectrum which we have not done here for the sake of brevity.

Another direction in which we would like to proceed is construct special exact solutions for the first equations in the hierarchies since they are likely to have some physical applications.

Including all these issues in one article is, of course, not possible, so we intend to address them in future works.

Acknowledgments

The work has been supported by the NRF incentive grant of South Africa and grant DN 02-5 of Bulgarian Fund ”Scientific Research”.

Footnotes

a

In [6, 7] FAS have mistakenly been claimed to satisfy Hχ˜±(x,λ)H=χ˜(x,λ) .

References

[3]AE Borovik and V.Yu Popkov, Completely integrable spin - 1 chains, Sov. Phys. JETPH, Vol. 71, 1990, pp. 177-85.
[4]LD Faddeev and LA Takhtadjan, Hamiltonian Method in the Theory of Solitons, Springer Verlag, Berlin, 1987.
[13]M Goto and F Grosshans, Semisimple Lie Algebras, M. Dekker Inc., New-York & Basel, 1978. Lecture Notes in Pure and Applied Mathematics 38
[14]ID Iliev, EKh Khristov, and KP Kirchev, Spectral Methods in Soliton Equations, John Wiley & Sons Inc., New York, 1994. Pitman Monographs and Surveys in Pure and Appl. Mathematics 73
[16]AV Mikhailov, Reduction in the integrable systems. Reduction groups, Lett. JETF, Vol. 32, 1979, pp. 187-192.
[19]AB Yanovski, Gauge-covariant approach to the theory of the generating operators for soliton equations, PhD thesis: 5-87-222, Joint Institute for Nuclear Research (JINR), 1987.
[20]AB Yanovski, Generating operators for the generalized Zakharov-Shabat system and its gauge equivalent system in sl (3,ℂ) case, 1993. Preprint: Universität Leipzig, Naturwissenchaftlich Theoretisches Zentrum Report N20 http://cdsweb.cern.ch/record/256804/files/P00019754.pdf
[24]AB Yanovski and G Vilasi, Geometry of the recursion operators for the GMV system, Jour. Nonl. Math. Phys., Vol. 19, No. N3, 2012. 1250023-1/18.
[25]AB Yanovski, Gauge-covariant theory of the generating operators associated with linear problems of Caudrey-Beals-Coifman type in canonical and in pole gauge with and without reductions, Proc. Int. Conf. BGSIAM14 (Sofia), 2015, pp. 2-43.
[26]AB Yanovski and G Vilasi, Recursion operators for CBC system with reductions. Geometric theory, Il Nuovo Cimento, Vol. 38, No. C, 2015, pp. 172.
[27]VE Zakharov and LA Takhtadjan, Equivalence between nonlinear Schrödinger equation and Heisenberg ferromagnet equation, TMF, Vol. 38, 1979, pp. 26-35.
Journal
Journal of Nonlinear Mathematical Physics
Volume-Issue
25 - 2
Pages
324 - 350
Publication Date
2021/01/06
ISSN (Online)
1776-0852
ISSN (Print)
1402-9251
DOI
10.1080/14029251.2018.1452676How to use a DOI?
Copyright
© 2018 The Authors. Published by Atlantis Press and Taylor & Francis
Open Access
This is an open access article distributed under the CC BY-NC 4.0 license (http://creativecommons.org/licenses/by-nc/4.0/).

Cite this article

TY  - JOUR
AU  - A. B. Yanovski
AU  - T. I. Valchev
PY  - 2021
DA  - 2021/01/06
TI  - Pseudo-Hermitian Reduction of a Generalized Heisenberg Ferromagnet Equation. I. Auxiliary System and Fundamental Properties
JO  - Journal of Nonlinear Mathematical Physics
SP  - 324
EP  - 350
VL  - 25
IS  - 2
SN  - 1776-0852
UR  - https://doi.org/10.1080/14029251.2018.1452676
DO  - 10.1080/14029251.2018.1452676
ID  - Yanovski2021
ER  -