Journal of Nonlinear Mathematical Physics

Volume 25, Issue 3, July 2018, Pages 442 - 461

Pseudo-Hermitian Reduction of a Generalized Heisenberg Ferromagnet Equation. II. Special Solutions

Authors
T. I. Valchev
Institute of Mathematics and Informatics, Bulgarian Academy of Sciences, Acad. G. Bonchev Str., 1113 Sofia, Bulgaria,tiv@math.bas.bg
A. B. Yanovski
Department of Mathematics & Applied Mathematics, University of Cape Town, Rondebosch 7700, Cape Town, South Africa,Alexandar.Ianovsky@uct.ac.za
Received 22 November 2017, Accepted 19 March 2018, Available Online 6 January 2021.
DOI
10.1080/14029251.2018.1494747How to use a DOI?
Keywords
generalized HF equation; dressing method; soliton solutions; quasi-rational solutions
Abstract

This paper is a continuation of our previous work in which we studied a sl(3, ℂ) Zakharov-Shabat type auxiliary linear problem with reductions of Mikhailov type and the corresponding integrable hierarchy of nonlinear evolution equations. Now, we shall demonstrate how one can construct special solutions over constant background through Zakharov-Shabat’s dressing technique. That approach will be illustrated on the example of the generalized Heisenberg ferromagnet equation related to the linear problem for sl(3, ℂ). In doing this, we shall discuss the differences between the Hermitian and pseudo-Hermitian cases.

Copyright
© 2018 The Authors. Published by Atlantis 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

In [7], the system of completely integrable equations

iut+uxx+(uux*+vvx*)ux+(uux*+vvx*)xu=0,i=1,ivt+vxx+(uux*+vvx*)vx+(uux*+vvx*)xv=0(1.1)
and the corresponding auxiliary spectral problem were introduced and studied. Above, the subscripts mean partial differentiation, * denotes complex conjugation and u, v: ℝ2 → ℂ are smooth functions subject to the condition |u|2+|v|2=1(|z|=zz*,z). System (1.1) is of particular interest since it is an integrable generalization of the classical integrable Heisenberg ferromagnet equation (HF)
St=S×Sxx.

HF describes the dynamics of a spin chain with unit spin vector S = (S1,S2,S3), see [4, 6, 21] for more details.

A vector system generalizing (1.1) was introduced quite independently by Golubchik and Sokolov [11]. Motivated by studies of another classical model from theoretical physics, the Landau-Lifshitz equation, the authors managed to find a Lax representation for the vector system, discussed the algebraic meaning of the Lax pair and developed its Hamiltonian formulation.

System (1.1) kept attracting some further attention and extensive studies. In [8, 9], the authors described the hierarchy of nonlinear evolution equations (NLEEs) associated with (1.1), the hierarchies of conservation laws as well as the hierarchies of Hamiltonian structures in the case when u(x) → eiδ± and v(x) → 0 sufficiently fast as x±∞. Moreover, the generalized Fourier transform interpretation of the inverse scattering transform for (1.1) was established and special soliton solutions to (1.1) were constructed via dressing method. A deep study of the properties of the recursion operators for (1.1) and the geometry linked to those was carried out in [22,24].

In [23], the authors of the current paper considered a slightly more general system of NLEEs having the form:

iut+uxx+(εuux*+vvx*)ux+(εuux*+vvx*)xu=0,ε2=1,ivt+vxx+(εuux*+vvx*)vx+(εuux*+vvx*)xv=0(1.2)
where u and v satisfy the constraint:
ε|u|2+|v|2=1.(1.3)

Obviously, (1.2) coincides with (1.1) when ε = 1 and, much like it, (1.2) has a zero curvature representation [L(λ), A(λ)] = 0 with Lax operators given by:

L(λ)=ixλS,λ,S=(0uvεu*00v*00),(1.4)
A(λ)=it+λA1+λ2A2,A2=(1/30002/3ε|u|2εu*v0v*u2/3|v|2),(1.5)
A1=(0abεa*00b*00),a=iuxi(εuux*+vvx*)ub=ivxi(εuux*+vvx*)v.(1.6)

The form of (1.4)(1.6) is rather similar to the form of the Lax pair of the original HF [8, 21, 23]. This is why we call (1.2) generalized Heisenberg ferromagnet equations. Following the convention in [23], the case when ε = 1 will be referred to as Hermitian reduction while the case when ε = −1 — pseudo-Hermitian. As we shall discuss in more detail at the end of this section, the pseudo-Hermitian reduction allows for different classes of solutions compared to the Hermitian one. This observation initially motivated us to extensively study the pseudo-Hermitian reduction.

In [23], we described the integrable hierarchy associated with (1.2) in terms of recursion operators and derived completeness relations of their eigenfunctions. Instead of building the theory ab initio, our analysis was based on the gauge equivalence between the auxiliary spectral problem with scattering operator in the form (1.4) and that one for a nonlinear Schrödinger equation, i.e. a generalized Zakharov-Shabat system. This allowed us to obtain all our results for arbitrary constant asymptotic values of the potential functions appearing in the auxiliary linear problems.

Another issue of fundamental importance concerns the solutions of (1.2). In the present paper, we intend to show how one can derive particular solutions to (1.2) in the simplest case of trivial (constant) background. In doing this, we shall not make use of the gauge equivalence between (1.4)(1.6) and a generalized Zakharov-Shabat’s system in canonical gauge. Our approach will be based on Zakharov-Shabat’s dressing method that seems to be suitable for the system of equations we are interested in.

The paper itself is organized as follows. Next section contains our main results and is divided into four subsections. The first subsection is preliminary— its purpose is to give the reader some basic idea of Zakharov-Shabat’s dressing method and how it could be applied to linear bundles (1.4). In doing this, we are going to use dressing factors which are meromorphic functions of the spectral parameter. Our analysis depends on the poles of the dressing factor. It turns out there exist three different cases:

  • Generic case, when the poles of the dressing factor are complex numbers in general position, i.e. those are not real or imaginary numbers;

  • The case, when the poles of the dressing factor are imaginary;

  • Degenerate case, when the poles are real.

Each of these cases is discussed in a separate subsection. Most of the considerations in these subsections are general in the sense that they allow one to construct special solutions to any NLEE belonging to the integrable hierarchy of (1.4) not just to the generalized HF. In order to specify the NLEE within the hierarchy, one needs to substitute its dispersion law into the general equations we are going to derive.

The first two of the cases mentioned above lead to soliton type solutions while the degenerate one leads to quasi-rational ones. In the context of the classification of solutions, it becomes quite obvious why the pseudo-Hermitian reduction deserves to be studied. There are essential differences between the properties of the solutions obtained in the Hermitian and pseudo-Hermitian cases. For example, some soliton type solutions develop singularities in the pseudo-Hermitian case while their counterparts are non-singular in the Hermitian one. Moreover, it turns out that quasi-rational solutions are possible in the pseudo-Hermitian case only. This situation is similar to the case of the classical (scalar) nonlinear Schrödinger equation iqt +qxx +κ|q|2q = 0, κ = ±1 for q : ℝ2 → ℂ. As it is well known, the defocusing nonlinear Schrödinger equation (κ = −1) for lim|x|→∞ q = 0 does not possess soliton solutions unlike the self-focusing case (κ = 1), see [6]. On the other hand, the defocusing nonlinear Schrödinger equation admit quasi-rational solutions for the afore-mentioned asymptotic condition while the self-focusing case does not.

Last section contains some concluding remarks.

2. Special Solutions

In this section, we shall present our main results: construction of special solutions to (1.2) over constant background. More specifically, we shall assume its solutions obey the following boundary condition:

limx±u(x,t)=0,limx±v(x,t)=1(2.1)
that is easily seen to be compatible with (1.3). Our approach to obtain particular solutions will be based on Zakharov-Shabat’s dressing method [25, 26]. Since we aim at providing a self-contained exposition, we shall remind the reader its basics as applied to linear bundles in pole gauge.

2.1. Dressing method and linear bundles in pole gauge

The dressing method represents an indirect way to solve completely integrable equations, i.e. one constructs new solutions to a given NLEE from a known (seed) solution. In doing this, one essentially uses the existence and the form of Lax representation.

In order to see how the method works, let us consider an arbitrary NLEE from the integrable hierarchy related to (1.4) and denote by (u0, v0) a known solution subject to (1.3) and (2.1). Then, as it is shown in [23], the NLEEs has a Lax representation

[L0(λ),A0(λ)]=0(2.2)
where L0 and A0 are given by:
L0(λ)=ixλS(0),S(0)=(0u0v0εu0*00v0*00),(2.3)
A0(λ)=it+k=1NλkAk(0),N2,λ.(2.4)

The coefficients Ak(0), k = 1,...,N can be uniquely expressed through (u0,v0) and their x-derivatives of order up to Nk. Here we shall not specify the form of the coefficients of the second Lax operator because only their asymptotic behavior as x±∞ is essential for our further considerations. We refer the reader who is interested in this issue to our previous paper [23] where they can find more detailed explanations.

Similarly to (1.4)(1.6), the above Lax operators fulfill the following symmetry conditions:

HL0(λ)H=L0(λ),HA0(λ)H=A0(λ),H=diag(1,1,1),(2.5)
Qε(S(0))Qε=S(0),Qε(Ak(0))Qε=Ak(0),Qε=diag(1,ε,1)(2.6)
where † stands for Hermitian conjugation and ε = ±1.

Let ψ0 be an arbitrary fundamental solution to the auxiliary linear problem:

L0(λ)ψ0(x,t,λ)=0.(2.7)

Due to (2.2), ψ0 also fulfills the linear problem:

A0(λ)ψ0(x,t,λ)=ψ0(x,t,λ)f(λ)(2.8)
where
f(λ):=limx±k=1Nλk[g0(x,t)]1Ak(0)(x,t)g0(x,t)(2.9)
is the dispersion law of the NLEE. Above,
g0:=12(101εu0*2v0εu0*v0*2u0v0*)
is the gauge transform diagonalizing S(0), see [23] for more explanations. The dispersion law of a NLEE is an essential feature encoding the time dependence of its solutions and that way it labels the NLEE within the integrable hierarchy. The dispersion law of (1.2) is
f(λ)=λ2diag(1,2,1)/3.(2.10)

Linear problems (2.7) and (2.8) will be referred to as bare (seed) linear problems and their fundamental solutions will be called bare (seed) fundamental solutions. We shall denote the set of all bare fundamental solutions by 0.

As discussed in [23], (2.5) and (2.6) are due to the action of the reduction group 𝕑2 × 𝕑2 on the set of bare fundamental solutions, see [13, 14] for more explanations. Indeed, in our case the following 𝕑2 × 𝕑2-action

ψ0(x,t,λ)Hψ0(x,t,λ)H,(2.11)
ψ0(x,t,λ)Qε[ψ0(x,t,λ*)]1Qε(2.12)
leads to (2.5) and (2.6) respectively.

Now, let us apply the gauge (dressing) transform

𝒢:01=𝒢0:={𝒢ψ0|ψ00}
where 𝒢 (x, t, λ) is a 3 × 3-matrix with unit determinant. Since the Lax operators are transformed through:
L0L1:𝒢L0𝒢1,A0A1:𝒢A0𝒢1,(2.13)
the operators L1 and A1 commute as well. We shall require that the auxiliary linear problems remain covariant under the dressing transform, i.e. we have
L1(λ)ψ1(x,t,λ)=0,A1(λ)ψ1(x,t,λ)=ψ1(x,t,λ)f(λ),ψ11(2.14)
for L1 and A1 being of the same form as L0 and A0. Thus we can write
L1(λ)=ixλS(1),S(1)=(0u1v1εu1*00v1*00),(2.15)
A1(λ)=it+k=1NλkAk(1)(2.16)
where u1 and v1 are some new (unknown) functions solving the same NLEE and subject to (1.3) and (2.1). The coefficients Ak(1), k = 1,...,N depend on (u1, v1) and their x-derivatives of order up to Nk in the same way as Ak(0) depend on (u0, v0) and their x-derivatives. Therefore, the dressed Lax pair has the same asymptotic behavior as the bare one.

Finding (u1, v1) means that we have obtained a new solution to the same NLEE and this is the main idea underlying the dressing method. Symbolically, it could be presented as the following sequence of steps:

(u0,v0)(L0,A0)ψ0𝒢ψ1(L1,A1)(u1,v1).

Though the covariance of the linear problems is a strong condition, it does not determine the dressing factor 𝒢 uniquely. Indeed, after comparing (2.7) and (2.8) with (2.14), we see that 𝒢 must solve the system of linear partial differential equations:

ix𝒢λ(S(1)𝒢𝒢S(0))=0,(2.17)
it𝒢+k=1Nλk(Ak(1)𝒢𝒢Ak(0))=0.(2.18)

Equations (2.17) and (2.18) tell us nothing about the λ-dependence of 𝒢 . This is why we need to make a few assumptions about its behavior with respect to λ in order to obtain more specific results. Let us assume the dressing factor and its derivatives in x and t are defined in the neighbourhood of λ = 0. After setting λ = 0 in (2.17) and (2.18), we immediately see that

x𝒢|λ=0=t𝒢|λ=0=0.

These relations imply that 𝒢 should depend non-trivially on λ otherwise it will be merely a constant. We shall pick up 𝒢|λ=0 = 𝟙 since it does not lead to any loss of generality. In fact, this normalization corresponds to the normalization of the fundamental analytic solutions to the auxiliary linear problems at λ = 0, see [23].

Equation (2.17) allows one to find u1 and v1 in terms of the seed solution and the dressing factor. At this point, we require that 𝒢 as well as its derivatives in x and t are regular when |λ| → ∞. After dividing (2.17) by λ and setting |λ| → ∞, we derive the following interrelation:

S(1)=𝒢S(0)𝒢1,𝒢(x,t):=lim|λ|𝒢(x,t,λ).(2.19)

Since S(0) is determined by (u0,v0) and S(1) is determined by (u1,v1), the above relation allows us to construct another solution of our system starting from a known one.

The form of the dressed operators and the zero curvature condition imply L1 and A1 obey (2.5) and (2.6) too. Therefore, the set of the dressed fundamental solutions is subject to the 𝕑2 ×𝕑2-action

ψ1(x,t,λ)Hψ1(x,t,λ)H,ψ11,(2.20)
ψ1(x,t,λ)Qε[ψ1(x,t,λ)]1Qε.(2.21)

The 𝕑2 ×𝕑2 action on 0 and 1 implies that the dressing factor is not entirely arbitrary but obeys the symmetry relations:

H𝒢(x,t,λ)H=𝒢(x,t,λ),(2.22)
Qε𝒢(x,t,λ*)Qε=[𝒢(x,t,λ)]1.(2.23)

A simple ansatz for dressing factor meeting all the requirements discussed so far is given by:

𝒢(x,t,λ)=𝟙+λj[Bj(x,t)μj(λμj)+HBj(x,t)Hμj(λ+μj)],μj\{0}.(2.24)

Generally speaking, the dressing procedure does not preserve the spectrum of scattering operator, see [6]. Indeed, let us assume for simplicity that L0 has no discrete eigenvalues, i.e. the spectrum of the bare scattering operator coincides with the real line, see [23]. Denote the resolvent of L0 by R0(λ). According to (2.13), the dressing transform maps the bare resolvent onto

R1(λ)=𝒢R0(λ)𝒢1.(2.25)

Equation (2.25) implies that the singularities of the dressing factor and its inverse “produce” singularities of the dressed resolvent operator R1(λ). This way the dressing procedure adds new discrete eigenvalues to the bare scattering operator.

In order to find L1 and A1 through (2.19), we need to know Bj(x,t) = Res(𝒢 (x,t,λ); μj). The algorithm to find the residues of (2.24) consists in two steps. In the first step, one considers the identity 𝒢 𝒢 −1 = 𝟙 which gives rise to a set of algebraic relations for Bj. The form of those relations crucially depends on the location of μj— whether the poles of the dressing factor are arbitrary complex numbers with nonzero real and imaginary parts (generic case) or the poles are either imaginary or real numbers. If the poles are complex numbers in generic position or purely imaginary then we obtain soliton type solutions. For poles lying on the real line, we have degeneracy in the spectrum of the scattering operator. As a result, we obtain quasi-rational solutions. Due to all these differences, we shall consider the three cases in separate subsections.

Though the algebraic relations may be different, they always imply that the residues of 𝒢 are some singular (degenerate) matrices which could be decomposed as follows:

Bj(x,t)=Xj(x,t)FjT(x,t)(2.26)
where Xj(x, t) and Fj(x, t) are rectangular matrices of certain rank and the superscript T stands for matrix transposition. After substituting (2.26) into the algebraic relations, we are able to express Xj through Fj.

The factors Fj are determined in the second step. For this to be done, we consider (2.17) and (2.18). As for the algebraic relations discussed above, the calculations here depend on the location of the poles of 𝒢 . However, (2.17) always leads to the following result:

FjT(x,t)=Fj,0T(t)[ψ0(x,t,μj)]1
allowing one to construct Fj through a bare fundamental solution defined in the neighbourhood of μj and some arbitrary x-independent matrices Fj,0. The matrices Fj,0 depend on t in a way governed by (2.18). Regardless of the location of μj, we derive exponential t-dependence. Thus in order to recover the time dependence in all formulas, we may use the rule below:
Fj,0T(t)Fj,0T(t)eif(μj)t
where f(λ) is the dispersion law of the NLEE, see (2.9). We shall demonstrate in more detail that procedure in the subsections to follow.

2.2. Soliton type solutions I. Generic case

Let us start with the case when the poles of (2.24) are complex numbers in generic position, i.e. μj2 for all j. From the symmetry condition (2.23), we immediately deduce that

[𝒢(x,t,λ)]1=𝟙+λi[QεBi(x,t)Qεμj*(λμi*)+QεHBi(x,t)HQεμi*(λ+μi*)].(2.27)

Thus the dressing factor and its inverse have poles located at different points.

Let us consider the identity 𝒢 (x, t, λ)[𝒢 (x, t, λ)]−1 = 𝟙. Since it holds identically in λ, the residues of 𝒢 𝒢 −1 should vanish. After evaluating the residue of 𝒢 𝒢 −1 at μi* we easily obtain the following algebraic relation:

[𝟙+μi*j(Bjμj(μi*μj)+HBjHμj(μi*+μj))]QεBiQε=0.(2.28)

Evaluation of the residues at ±μi and μi* leads to equations that can easily be reduced to (2.28), thus giving us no new constraints.

It is seen from (2.28) that each Bi should be a degenerate matrix, hence it can be factored

Bi(x,t)=Xi(x,t)FiT(x,t)(2.29)
where Xi and Fi are two rectangular matrices. After substituting (2.29) into (2.28), one obtains the linear system
QεFi*=μi*j(XjFjTQεFi*μj(μjμi*)HXjFjTHQεFi*μj(μi*+μj))(2.30)
for the factors Xj. Solving it, allows one to express Xj through Fj. This is easier when the dressing factor has just a single pair of poles: μ and −μ. Assuming X and F are column-vectors, we get:
X=(μ*FTQεF*μ(μμ*)μ*FTHQεF*μ(μ+μ*)H)1QεF*.(2.31)

Let us return now to the general case. The factors Fj can be found from (2.17). For that purpose we rewrite (2.17) in the following way

λS(1)=ix𝒢𝒢1+λ𝒢S(0)𝒢1
and calculate the residues of its right-hand side at λ = μi. After taking into account (2.30), we obtain the linear differential equation
ixFiT+μiFiTS(0)=0(2.32)
which is immediately solved to give
FiT(x,t)=Fi,0T(t)[ψ0(x,t,μi)]1.(2.33)

Above, Fi,0 are “integration constant” matrices which depend on t however. Evaluation of the residues at −μi or ±μi* gives equations that are easily reduced to (2.32) after taking into account the symmetries of S1(0).

The t-dependence of Fi,0 is determined from (2.18) after it is rewritten as follows:

k=1NλkAk(1)=it𝒢𝒢1+𝒢k=1NλkAk(0)𝒢1.(2.34)

Evaluation of the residues of both hand sides of (2.34) for λ = μi gives

itFiTFiTk=1NμikAk(0)=0.(2.35)

Yet again, we do not need to consider the residues for λ=±μi* and λ = −μi due to the symmetries of the coefficients of A(λ).

After substituting (2.33) into (2.35), we derive a linear differential equation for Fi,0, namely

itFi,0TFi,0Tf(μi)=0(2.36)
where f(λ) is the dispersion law of the NLEE. Equation (2.36) is easily solved and allows us to state the rule: in order to recover the t-dependence in all formulas, one has to make the following substitution
Fi,0TFi,0Teif(μi)t.(2.37)

Now let us apply the general results obtained above to some seed solution. An obvious choice for seed solution satisfying (1.3) and (2.1) is

S(0)=(001000100).(2.38)

One can prove that the bare scattering operator in this case has no discrete eigenvalues.

For a bare fundamental solution we shall choose

ψ0(x,λ)=(cos(λx)0isin(λx)010isin(λx)0cos(λx)).(2.39)

This fundamental solution is invariant with respect to (2.11) and (2.12) which makes it rather convenient for the calculations to follow.

From this point on, we shall focus on the simplest case when the dressing factor has a single pair of poles and X and F are column-vectors. After substituting (2.29) and (2.31) into (2.24) and (2.27), equation (2.19) can be written in components as follows:

u1=[μ|F1|2+μ*(ε|F2|2+|F3|2)](μ2(μ*)2F2(F3)*μ*[μ*|F1|2+μ(ε|F2|2+|F3|2)]2,(2.40)
v1=[μ|F1|2+μ*(ε|F2|2+|F3|2)]μ*[μ*|F1|2+μ(ε|F2|2+|F3|2)]2×{μ*[μ|F1|2+μ*(ε|F2|2+|F3|2)]+(μ2(μ*)2)ε|F2|2}(2.41)
where FT = (F1,F2,F3). It is not hard to check that (2.40) and (2.41) satisfy (1.3) for any F.

On the other hand, (2.33) leads to the following expression for F:

F(x)=(F01cos(μx)+iF03sin(μx)F02F03cos(μx)+iF01sin(μx)),F0=(F01F02F03).(2.42)

In order to evaluate X, we shall need the following quadratic expressions as well:

FTQεF*=(|F01|2+|F03|2)cosh(2κx)2|F01F03|cosφsinh(2κx)+ε|F02|2,(2.43)
FTQεHF*=[(|F03|2|F01|2)cos(2ωx)2|F01F03|sinφsin(2ωx)+ε|F02|2](2.44)
where ω = Re μ > 0, κ = Im μ > 0 and φ=argF01argF03. Taking into account the structure of (2.39)(2.42), it is natural to consider in detail the following three elementary cases, see [9].
  1. (1)

    First, let us assume F02=0 and F01±F03. Then for F we have

    F(x)=(F01cos(μx)+iF03sin(μx)0F03cos(μx)+iF01sin(μx))(2.45)
    and (2.43), (2.44) now could be simplified to
    FTQεF*=Acosh(2κx+ξ0),(2.46)
    FTQεHF*=Acos(2ωx+δ0)(2.47)
    where
    coshξ0=|F01|2+|F03|2A,sinhξ0=2|F01F03|cosφA,(2.48)
    A=(|F01|2+|F03|2)24|F01F03|cos2φ,(2.49)
    cosδ0=|F01|2|F03|2A,sinδ0=2|F01F03|sinφA.(2.50)

    After substituting (2.45)(2.50) into (2.40) and (2.41), we obtain the dressed solution to (1.2) at a fixed moment of time:

    u1(x)=0,(2.51)
    v1(x)=exp[4iarctanκcos(2ωx+δ0)ωcosh(2κx+ξ0)].(2.52)

    In order to recover the time evolution in (2.51) and (2.52), one needs to make the substitution:

    φφ,ξ0ξ0,δ0δ0,AAexp(4ωκt3)(2.53)
    that follows directly from (2.10) and (2.37). It is seen that (2.51) and (2.52) remain invariant under transformation (2.53), so the dressed solution is stationary and non-singular. The solution just derived coincides with that found in [9] (see formula (3.26)) for the generalized HF with Hermitian reduction.

  2. (2)

    Let us assume F020. Without any loss of generality we could just set F02=1. There exist two elementary options: F01=F03 and F01=F03. Let us consider the former one. After recovering the time evolution, the vector F(x) is given by:

    F(x,t)=(F01eiμx+iμ2t3e2iμ2t3F01eiμx+iμ2t3).

    The solution in this case reads:

    u1(x,t)=4ωκ[2ωeϑ(x,t)+ε(ωiκ)eϑ(x,t)]ei[ωx+(ω2κ2)t+δ](ωiκ)[2ωeϑ(x,t)+ε(ω+iκ)eϑ(x,t)]2,(2.54)
    v1(x,t)=18εωκ2(ωiκ)[2ωeϑ(x,t)+ε(ω+iκ)eϑ(x,t)]2(2.55)
    where
    ϑ(x,t)=κ(x+2ωt)+ln|F0,1|,δ=(π2+argF0,1).

    Solution (2.54), (2.55) has no singularities and goes into the one obtained in [9] (see equation (3.27)) when ε = 1.

  3. (3)

    Assume now F02=1 and F01=F03. In this case F(x,t) is given by:

    F(x,t)=(F01eiμx+iμ2t3e2iμ2t3F01eiμx+iμ2t3).

    The corresponding dressed solution looks as follows:

    u1(x,t)=4ωκ[2ωeϑ˜(x,t)+ε(ωiκ)eϑ˜(x,t)]ei[ωx+(κ2ω2)t+δ˜](ωiκ)[2ωeϑ˜(x,t)+ε(ω+iκ)eϑ˜(x,t)]2,(2.56)
    v1(x,t)=18εωκ2(ωiκ)[2ωeϑ˜(x,t)+ε(ω+iκ)eϑ˜(x,t)]2(2.57)
    where
    ϑ˜(x,t)=κ(x2ωt)+ln|F0,1|,δ˜=π2argF0,1.

Formally, this solution could be derived from (2.54), (2.55) by applying the transform x → −x and F0,1F0,3.

All the solutions constructed explicitly in this subsection are connected with four distinct poles (quadruples) of 𝒢 and 𝒢−1, i.e. four discrete eigenvalues of the dressed scattering operator. This is why we call such solutions “quadruplet” solutions.

2.3. Soliton type solutions II. Imaginary poles

Let us assume now that all the poles of 𝒢 are purely imaginary, i.e. we have μj = iκj for some real numbers κj ≠ 0. Thus we could write

𝒢(x,t,λ)=𝟙+λi[Bi(x,t)iκi(λiκi)+HBi(x,t)Hiκi(λ+iκi)](2.58)
for the dressing factor and
[𝒢(x,t,λ)]1=𝟙+λi[QεBi(x,t)Qεiκi(λ+iκi)+QεHBi(x,t)HQεiκi(λiκi)].(2.59)
for its inverse. Since 𝒢 and 𝒢 −1 have the same poles, the identity 𝒢 𝒢 −1 = 𝟙 gives rise to algebraic relations that do not follow from (2.28). It is easily checked that the following equations hold true:
limλiκi(λiκi)2(𝒢𝒢1=𝟙)BiQεHBi=0,(2.60)
limλiκiλ(λiκi)2(𝒢𝒢1=𝟙)BiQεHΩiΩiQεHBi=0(2.61)
where
Ωi=𝟙+Biiκi+jiκiBjiκj(κiκj)+jκiHBjHiκj(κi+κj).

There is no need to perform similar calculations for λ = −iκi since all algebraic equations can be reduced to (2.60) and (2.61).

Relation (2.60) implies that Bi(x,t) is degenerate, so (2.29) holds again for some rectangular matrices Xi and Fi (we shall assume that these matrices are simply column vectors). Due to (2.60) the factors Fi satisfy the quadratic relation

FiTQεHFi*=0.(2.62)

After substituting (2.29) into (2.61), we see that there exist functions αi such that:

ΩiQεHFi*=Xiαi,αi*=αi.(2.63)

These relations allow one to find all vectors Xi in terms of Fi and αi. In the simplest case when the dressing factor has a single pair of poles ±iκ (2.63) is easily solvable to give:

X=(α+FTQεF*2iκH)1QεHF*.(2.64)

After taking into account (2.29), (2.58), (2.59) and (2.64), equation (2.19) could be written down in components as follows:

u1(x,t)=2(|F1(x,t)|2+iκα(x,t))F2(x,t)(F3(x,t))*(|F1(x,t)|2=iκα(x,t))2,(2.65)
v1(x,t)=1+2(2iκα(x,t)ε|F2(x,t)|2)|F1(x,t)|2iκα(x,t)+4iκα(x,t)(iκα(x,t)ε|F2(x,t)|2)(|F1(x,t)|2iκα(x,t))2.(2.66)

It is easy to check that (2.65) and (2.66) are interrelated through |v|2 +ε|u|2 = 1 for any real-valued α and F obeying (2.62).

Like in the generic case (see the previous subsection) we can find Fi and αi by analyzing (2.17) and (2.18). Considerations similar to those in the generic case lead to the following linear differential equations:

limλiκi(λiκi)2(ix𝒢𝒢1+λ𝒢S(0)𝒢1=λS(1))ixFiT+μiFiTS(0)=0,(2.67)
limλiκi(λiκi)2(ix𝒢𝒢1+λ𝒢S(0)𝒢1=λS(1))ixαi=FiTS(0)QεHFi*.(2.68)

Equation (2.67) shows that (2.33) holds true for the vectors Fi while (2.68) implies that the scalar functions αi can be expressed in terms of the seed fundamental solution as given by:

αi(x)=αi,0+Fi,0T[ψ0(x,iκi)]1ψ0(x,iκi)λKψ0QεHFi,0*.(2.69)

Above, αi,0 is an integration constant and Kψ(λ) = [ψ(λ)]−1HQε[ψ(−λ*)]−1 QεH measures the “deviation” of the solution ψ from invariant solutions, i.e. if ψ is invariant with respect to (2.11) and (2.12), then Kψ = 𝟙.

Finally, we need to recover the time dependence in all formulas. For this to be done we consider (2.18) as we did in the previous subsection. As a result, we derive a set of equations for Fi and αi, namely:

limλiκi(λiκi)2[it𝒢𝒢1+𝒢kλkAk(0)𝒢1]=0itFiTFjTk(iκi)kAk(0)=0,(2.70)
limλiκiλ{(λiκi)2[it𝒢𝒢1+𝒢kλkAk(0)𝒢1]}=0itαi+FiTkk(iκi)k1Ak(0)QεHFi*=0.(2.71)

Equation (2.70) coincides with (2.35). Thus after substituting (2.33) into (2.70), we derive (2.36) and (2.37) holds again. On the other hand, (2.71) is reduced to

itαi,0+Fi,0Tdf(iκi)dλKψ0HQεFi,0*=0.(2.72)

If the bare fundamental solution is invariant with respect to both reductions, i.e. Kψ0 = 𝟙, then (2.72) is simplified to

itαi,0+Fi,0Tdf(iκi)dλHQεFi,0*=0.(2.73)

Let us consider the case when the seed solution S(0) is picked up as in (2.38) and the corresponding fundamental solution is given by (2.39). As we discussed in the previous subsection, for (2.39) we have Kψ0 = 𝟙. From this point on, we shall assume that 𝒢 has a single pair of imaginary poles.

In this case the 3-vector F reads:

F(x)=(cosh(κx)F01sinh(κx)F03F02cosh(κx)F03sinh(κx)F01),F0=(F01F02F03).(2.74)

Our further considerations depend on whether or not F02 is equal to 0.

  1. (1)

    Let us first assume F02=0. Then we may set |F01|=|F03|=1 and (2.74) and (2.69) now give

    |F1(x)|2=cosh(2κx)sinh(2κx)cos(φ),2φ=argF01argF03(2.75)
    α(x)=α0+2xsin(2φ)(2.76)
    where α0 is t-independent in this case. After substituting (2.75) and (2.76) into (2.65) and (2.66), then taking into account (2.74), we get the following stationary solution to (1.2):
    u1(x)=0,(2.77)
    v1(x)=[cosh(2κx)sinh(2κx)cos(2φ)+iκ(α0+2xsin(2φ))cosh(2κx)sinh(2κx)cos(2φ)iκ(α0+2xsin(2φ))]2.(2.78)

  2. The dressed solution just derived does not “feel” the presence of ε, i.e. whether the reduction is Hermitian or pseudo-Hermitian, hence it formally coincides with the one obtained in [9]. The right-hand side of (2.78) could be rewritten as:

    v1(x)=exp[4iarctan(κ(α0+2xsin(2φ))cosh(2κx)sinh(2κx)cos(2φ))].

    When we have F0,1 = ±F0,3, it simplifies to

    v1(x)=[e2κx+iκα0e2κxiκα0]2.

    It is seen that v1 is nontrivial only if α0 ≠ 0.

  3. (2)

    Assume now that F020. Thus we may simply set F02=1. Our further analysis depends on the value of ε. Let us pick up ε = 1 first. Due to (2.62) and the invariance of (2.39) under the reductions, we have

    F0TQεHF0*=0|F01|2|F03|2=1.(2.79)

    A natural parametrization for F01 and F03 is the following one:

    F01=cosh(θ)ei(δ+φ),F03=sinh(θ)ei(δφ),θ>0.(2.80)

Then using (2.80), we have

|F1(x,t)|2=cosh2(κxθ)cos2φ+cosh2(κx+θ)sin2φ,α(x,t)=xsinh(2θ)sin(2θ)2κt.

In the end we derive the following dressed solution:

u1(x,t)=2Δp*Δp2ei(κ2tδ)[sinh(κxθ)cosφ+isinh(κx+θ)sinφ],
v1(x,t)=1+2(2iκα1)Δp+4iκα(iκα1)Δp2,
Δp=cosh2(κxθ)cos2φ+cosh2(κx+θ)sin2φiκ(xsinh2θsin2φ2κt).

The result we obtained coincides with the one found in [9].

Let us consider now the case when ε = −1. Relation (2.79) is rewritten as

|F03|2|F01|2=1
hinting at the following parametrization:
F01=sinh(θ)ei(δ+φ),F03=cosh(θ)ei(δφ),θ>0.

Then we have

|F1(x,t)|2=sinh2(κxθ)cos2φ+sinh(κx+θ)sin2φ,α(x,t)=xsinh(2θ)sin(2θ)+2κt
and substituting into (2.65) and (2.66) we obtain:
u1(x,t)=2Δm*Δm2ei(κ2tδ)[cosh(κxθ)cosφ+icosh(κx+θ)sinφ],(2.81)
v1(x,t)=1+2(2iκα1)Δp+4iκα(iκα1)Δp2,(2.82)
Δm=sinh2(κxθ)cos2φ+sinh2(κx+θ)sin2φiκ[xsinh(2θ)sin2φ2κt].

It is not hard to see that Δm could vanish for particular values of the parameters and thus we obtain singular solutions.

Unlike the quadruplet solutions in the previous subsection, the soliton-like solutions we derived here are related to a pair of discrete eigenvalues (doublet) of the scattering operator. This is the reason we call such solutions “doublet” solutions.

2.4. Degenerate case

Hereafter, we shall consider the case when the poles of the dressing factor are all real. Like in doublet case (see the previous subsection), the dressing factor

𝒢(x,t,λ)=𝟙+λi=1[Bi(x,t)μi(λμi)+HBi(x,t)Hμi(λ+μi)],μi\{0}
and its inverse
[𝒢(x,t,λ)]1=𝟙+λi=1[QεBi(x,t)Qεμi(λμi)+QεHBi(x,t)HQεμi(λ+μi)]
have the same poles. Then the identity 𝒢 (λ)[𝒢 (λ)]−1 = 𝟙 gives rise to the following algebraic relations:
B1QεBi=0,(2.83)
Ω˜iQεBi=BiQεΩ˜i=0(2.84)
where
Ω˜i=𝟙Biμi+jiμiBjμj(μiμj)+jμiHBjHμj(μi+μj).(2.85)

Like in the previous case, relation (2.83) could be reduced to

FiTQεFi*=0(2.86)
for the vector Fi involved in (2.29). It is clear that for Qε = 𝟙 (2.86) leads to Fi = 0 and 𝒢 becomes equal to 𝟙. Therefore a nontrivial result is possible only for pseudo-Hermitian reduction, i.e. Qε = diag (1, −1,1).

From this point on, we shall be interested in the case when Xi and Fi are just 3-vectors. Relation (2.84) implies there exists a scalar function βi such that

Ω˜iQεFi*=Xiβi,βi*=βi.(2.87)

Relation (2.87) is a linear system of equations for Xi. This system could easily be solved when 𝒢 has a single pair of real poles. In this case for X we have:

X=(βFTHQεF*2μH)1QεF*.(2.88)

In order to obtain Fi and βi, one considers the differential equations satisfied by the dressing factor, see (2.17) and (2.18). Analysis rather similar to that in the previous subsections shows that Fi depend on bare fundamental solution according to (2.33) while βi is determined by:

β1=βi,0Fi,0T[ψ0(x,μi)]1λψ0(x,μi)K˜ψ0(μi)QεFi,0*
where K˜ψ(λ)=[ψ(λ)]1Qε[ψ(λ*)]1Qε measures the “deviation from invariance” of the solution ψ and βi,0 is an imaginary x-independent scalar function. Like in the previous cases, (2.18) leads to (2.37) recovering the time dependence of Fi. The time dependence of βi is obtained through the following substitution rule:
βi,0βi,0iFi,0Tdf(λ)dλ|λ=μiK˜ψ0(μi)QεFi,0*t(2.89)
where f(λ) is the dispersion law.

We shall focus now on the simplest case when the dressing factor has one pair of real poles. Assume the seed solution S(0) and the corresponding fundamental solution ψ0 are given by (2.38) and (2.39) respectively. Due to (2.86), for the components of the 3-vector F0=(F01,F02,F03) we have

|F01|2+|F03|2=|F02|2.(2.90)

Without any loss of generality we could set F02=1. Then the form of (2.90) suggests the parametrization:

F01(t)=cos(θ)ei(μ2t3+δ+φ),F03(t)=sin(θ)ei(μ2t3+δφ),θ[0,π/2].

Then F could be written down as:

F(x,t)=(ei(μ2t3+δ)[eiφcos(μx)cosθ+ieiφsin(μx)sinθ]e2iμ2t3ei(μ2t3+δ)[eiφcos(μx)sinθ+ieiφsin(μx)cosθ]).(2.91)

A relatively simple computation shows that β depends linearly on x and t as follows:

β=i[2μt+xsin(2θ)cos(2φ)].(2.92)

Thus taking into account (2.19), the dressed solution to (1.2) acquires the form:

u1(x,t)=2(μβ+|F1|2)F2(F3)*(μβ|F1|2)2,(2.93)
v1(x,t)=(μβ+|F1|2)(μβ1|F3|2)(μβ|F1|2)2.(2.94)

It could easily be checked that (2.93) and (2.94) satisfy (1.3) provided β and F obey (2.86) but are otherwise arbitrary.

After substituting (2.91) and (2.92) into (2.93) and (2.94) and making some further simplifications, we derive

u1(x,t)=2Δd*Δd2e(iμ2t+δ+π4)[cos(φ+π4)sin(μx+θ)isin(φ+π4)sin(μxθ)],(2.95)
v1(x,t)=Δd*(Δd*2)Δd2(2.96)
where
Δd=cos2(μx+θ)cos2(φ+π4)+cos2(μx+θ)sin2(φ+π4)iμ[2μt+xsin(2θ)cos(2φ)].

Solutions (2.95) and (2.96) contain terms that are linear in x and t as well as terms that are bounded oscillating functions. This is why we call such solutions quasi-rational. It is easy to see that the denominator Δd has zeros for particular values of the parameters. For example, when θ = π/2 it is zero at x = 0 and t = 0.

3. Conclusion

In this paper, we have shown how Zakharov-Shabat’s dressing method can be applied to linear bundles in pole gauge whose potential functions are subject to (2.1). Following the general algorithm described in Section 2.1, we have constructed special solutions to a generalized Heisenberg ferromagnet equation, see (1.2). The simplest class of such solutions corresponds to a dressing factor with simple poles, see (2.24). We have seen that the location of the poles in λ-plane affects the form of the solutions. There exist three “pure” cases: the poles are complex numbers in generic position, the poles are imaginary and the poles are real. The first two options lead to soliton-like solutions of quadruplet and doublet type respectively, see formulas (2.51), (2.52), (2.54), (2.55), (2.56), (2.57), (2.77), (2.78), (2.81) and (2.82). All the quadruplet solutions we have derived are non-singular while the doublet solutions could have singularities in the pseudo-Hermitian case.

The case of real poles differs much from the others. It introduces certain degeneracy in the spectrum of the scattering operator and leads to quasi-rational solutions, see (2.95) and (2.96). We have seen such degeneracy is possible only if we have a pseudo-Hermitian reduction. This is a rather essential difference between the Hermitian and pseudo-Hermitian reductions. The quasi-rational solutions (2.95) and (2.96) could be non-singular for particular values of the parameters.

Apart from the “pure” cases we have discussed in the main text, there exists a situation when some poles of the dressing factor are generic complex numbers while the rest are either real or imaginary numbers. Clearly, the analysis of this “mixed” case can be reduced to the considerations we have already demonstrated.

Though we have discussed in detail special solutions to (1.2), similar procedures could be used to derive explicit solutions to any NLEE belonging to the integrable hierarchy of (1.4). The solutions of such a NLEE will have the same x-dependence as those of (1.2) but a different t-dependence.

In the present paper, we have focused on the simplest class of solutions obeying asymptotic behavior (2.1). A possible way of extending our results is by looking for solutions having more complicated behavior, e.g. non-trivial background solutions or periodic solutions. In the last few decades, nontrivial background solutions have become a topic of increased interest due to the connection to phenomena like rogue (freak) waves, see [1, 3, 17, 18]. Such solutions were obtained for classical integrable equations like the (scalar) nonlinear Schrödinger equation [1, 2, 12, 18], the 3-wave equation [3, 5] and for the scalar derivative nonlinear Schrödinger equation [19, 20]. This is why it is interesting to explicitly construct such type of solutions for the generalized Heisenberg equation and find out how the properties of (1.2) and its spectral problem will change if u and v have a different behavior.

Another direction to extend our results is by studying auxiliary spectral problems and the corresponding integrable hierarchies of NLEEs for other symmetric spaces and/or other reductions. This includes the study of multicomponent versions of (1.2) connected to the symmetric spaces SU(m + n)/S(U(m) ×U(n)) and rational bundles like the following one:

L(λ)=ixλS11λS1,λ\{0},
where
S±1=(0u±v±0)
for u± and v± being some vector functions. The above rational bundle can be viewed as a deformation of the linear bundle we have considered here and it has much more complicated properties, see [7, 10] for some discussion of the Hermitian case. The dressing method has already been successfully applied to such complicated spectral problems, see [15,16] for some interesting examples. We intend to discuss all these issues in more detail elsewhere.

Acknowledgments

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

References

[4]A.E. Borovik and V. Yu. Popkov, Completely integrable spin-1 chains, Sov. Phys. JETPH, Vol. 71, 1990, pp. 177-85.
[13]A.V. Mikhailov, Reduction in the integrable systems. Reduction groups, Lett. JETF, Vol. 32, 1979, pp. 187-192.
[21]L. Takhtadjan and L. Faddeev, The Hamiltonian Approach to Soliton Theory, Springer-Verlag, Berlin, 1987.
[23]A.B. Yanovski and T.I. Valchev. Pseudo-Hermitian reduction of a generalized Heisenberg ferromagnet equation. I. Auxiliary system and fundamental properties, ArXiv: 1709.09266 [nlin.SI]
Journal
Journal of Nonlinear Mathematical Physics
Volume-Issue
25 - 3
Pages
442 - 461
Publication Date
2021/01/06
ISSN (Online)
1776-0852
ISSN (Print)
1402-9251
DOI
10.1080/14029251.2018.1494747How to use a DOI?
Copyright
© 2018 The Authors. Published by Atlantis 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  - T. I. Valchev
AU  - A. B. Yanovski
PY  - 2021
DA  - 2021/01/06
TI  - Pseudo-Hermitian Reduction of a Generalized Heisenberg Ferromagnet Equation. II. Special Solutions
JO  - Journal of Nonlinear Mathematical Physics
SP  - 442
EP  - 461
VL  - 25
IS  - 3
SN  - 1776-0852
UR  - https://doi.org/10.1080/14029251.2018.1494747
DO  - 10.1080/14029251.2018.1494747
ID  - Valchev2021
ER  -