The gl3(C) rational Gaudin model governed by 3×3 Lax matrix is applied to study the three-wave resonant interaction system (TWRI) under a constraint between the potentials and the eigenfunctions. And the TWRI system is decomposed so as to be two finite-dimensional Lie-Poisson Hamiltonian systems. Based on the generating functions of conserved integrals, it is shown that the two finite-dimensional Lie-Poisson Hamiltonian systems are completely integrable in the Liouville sense. The action-angle variables associated with non-hyperelliptic spectral curves are computed by Sklyanin's method of separation of variables, and the Jacobi inversion problems related to the resulting finite-dimensional integrable Lie-Poisson Hamiltonian systems and three-wave resonant interaction system are analyzed.
Citation: Xue Geng, Liang Guan, Dianlou Du. Action-angle variables for the Lie-Poisson Hamiltonian systems associated with the three-wave resonant interaction system[J]. AIMS Mathematics, 2022, 7(6): 9989-10008. doi: 10.3934/math.2022557
[1] | Simone Fiori . Coordinate-free Lie-group-based modeling and simulation of a submersible vehicle. AIMS Mathematics, 2024, 9(4): 10157-10184. doi: 10.3934/math.2024497 |
[2] | Hashem Althagafi, Ahmed Ghezal . Solving a system of nonlinear difference equations with bilinear dynamics. AIMS Mathematics, 2024, 9(12): 34067-34089. doi: 10.3934/math.20241624 |
[3] | Chun-Ku Kuo, Dipankar Kumar, Chieh-Ju Juan . A study of resonance Y-type multi-soliton solutions and soliton molecules for new (2+1)-dimensional nonlinear wave equations. AIMS Mathematics, 2022, 7(12): 20740-20751. doi: 10.3934/math.20221136 |
[4] | Baiying He, Siyu Gao . The nonisospectral integrable hierarchies of three generalized Lie algebras. AIMS Mathematics, 2024, 9(10): 27361-27387. doi: 10.3934/math.20241329 |
[5] | Hui Wang, Xue Wang . Bifurcations of traveling wave solutions for the mixed Korteweg-de Vries equation. AIMS Mathematics, 2024, 9(1): 1652-1663. doi: 10.3934/math.2024081 |
[6] | Ramses van der Toorn . Elementary properties of non-Linear Rossby-Haurwitz planetary waves revisited in terms of the underlying spherical symmetry. AIMS Mathematics, 2019, 4(2): 279-298. doi: 10.3934/math.2019.2.279 |
[7] | Adil Jhangeer, Ali R Ansari, Mudassar Imran, Beenish, Muhammad Bilal Riaz . Lie symmetry analysis, and traveling wave patterns arising the model of transmission lines. AIMS Mathematics, 2024, 9(7): 18013-18033. doi: 10.3934/math.2024878 |
[8] | Sixing Tao . Breather wave, resonant multi-soliton and M-breather wave solutions for a (3+1)-dimensional nonlinear evolution equation. AIMS Mathematics, 2022, 7(9): 15795-15811. doi: 10.3934/math.2022864 |
[9] | Shumoua F. Alrzqi, Fatimah A. Alrawajeh, Hany N. Hassan . An efficient numerical technique for investigating the generalized Rosenau–KDV–RLW equation by using the Fourier spectral method. AIMS Mathematics, 2024, 9(4): 8661-8688. doi: 10.3934/math.2024420 |
[10] | Amer Ahmed, Abdulghani R. Alharbi, Ishak Hashim . Exact and numerical solutions of the generalized breaking soliton system: Insights into non-linear wave dynamics. AIMS Mathematics, 2025, 10(3): 5124-5142. doi: 10.3934/math.2025235 |
The gl3(C) rational Gaudin model governed by 3×3 Lax matrix is applied to study the three-wave resonant interaction system (TWRI) under a constraint between the potentials and the eigenfunctions. And the TWRI system is decomposed so as to be two finite-dimensional Lie-Poisson Hamiltonian systems. Based on the generating functions of conserved integrals, it is shown that the two finite-dimensional Lie-Poisson Hamiltonian systems are completely integrable in the Liouville sense. The action-angle variables associated with non-hyperelliptic spectral curves are computed by Sklyanin's method of separation of variables, and the Jacobi inversion problems related to the resulting finite-dimensional integrable Lie-Poisson Hamiltonian systems and three-wave resonant interaction system are analyzed.
It is well known that Gaudin models describe completely integrable classical and quantum long-range interacting spin chains. Originally it was introduced by M. Gaudin for the simple Lie algebra su(2) [1] and later generalized to arbitrary semi-simple Lie algebras [2,3]. This model attracted considerable interest among theoretical and mathematical physicists, playing a distinguished role in the realm of integrable systems. For example, classical Hamiltonian systems associated with Lax matrices of the Gaudin-type in the context of a general group-theoretic approach [4], multi-Hamiltonian formulations (see e.g., [5]) and their integrable discretizations through B¨acklund transformations (see e.g., [6]), separation of variables of the rational Gaudin models (see e.g., [7,8,9,10]).
The separation of variables for finite-dimensional integrable systems are important for constructing action-angle variables. Much work on the separation of variables for finite dimensional integrable systems associated with hyperelliptic spectral curves was done(see e.g., [9,10,11,12,13,14,15,16]). It is worth noting that Sklyanin gave an efficient way to construct separated variables for the classical integrable SL(3) magnetic chain and sl(3) Gaudin model related to the non-hyperelliptic spectral curves [8]. After this classical work, the general cases [17,18,19] and a sufficient condition to guarantee the "separation polynomial" of Sklyanin under the corresponding r-matrix were provided in [20]. Action-angle variables of the finite-dimensional integrable systems can be constructed by using separated variables from the Liouville theorem and integration of Hamiltonian equations of motion. They were also derived with the help of the scattering data for the first-order matrix differential operator [21], or by finding real Darboux coordinates for scattering data [22,23], or by using the algebraic-geometric techniques [23,24,25]. Action-angle variables of the finite-dimensional integrable systems related to non-hyperelliptic spectral curves can also be obtained in dealing with the stationary equations of soliton hierarchies [24,25].
Based on the previous research on Gaudin models [8,10], the motivation of this paper is to use the gl3(C) rational Gaudin model governed by the Lax matrix to study the constraint systems associated with the generalized TWRI system: [26,27]
ulk,t=clkulk,x+(cml−cmk)ulmumk,1≤l,k,m≤3, | (1.1) |
where l,k,m are not equal to each other, the real constants c12,c21,c13,c31,c23,c32 are six real constants and u12,u21,u13,u31,u23,u32 are complex functions of two real independent variables x and t. Also, as a basic integrable model in mathematical physics, which has important applications in nonlinear optics, plasma physics, acoustics, fluid dynamics, solid-state physics and other fields, many studies have been done on the TWRI system. For example, Zakharov and Manakov [28,29] and Kaup [30] considered the inverse scattering transformation of the TWRI equations. It is shown that the TWRI equations can be reduced to the generic sixth Painlevé equation [31]. Some explicit solutions of the TWRI system are obtained in [32,33,34,35,36,37,38,39], including soliton solutions, rational solutions, rogue wave solutions and so on. The finite dimensional Hamiltonian system associated with the TWRI system is proved completely integrable in the Liouville sense [26]. Based on the theory of trigonal curves, explicit algebro-geometric solutions of the TWRI system have been achieved in [40] by using the asymptotic properties of the Baker-Akhiezer functions and the meromorphic functions [41,42].
The outline is as follows. In Section 2, we review some basics about Lie-Poisson structure and coadjoint representaive theory of Lie-algebra gl3(C). In Section 3, Lax matrix Vλ is introduced to study the TWRI system (1.1) in the Lie-Poisson structure, the integrability of this restricted system and the relation between the finite dimensional Hamiltonian system and the TWRI system (1.1) are also discussed. In Section 4, the 9N dimensional Poisson manifold (gl3(C)∗)N is reduced to 6N dimensional symplectic manifold by making a restriction on 3N common level set of Casimir functions, from which 3N pairs of separated variables are constructed on the 6N dimensional symplectic manifold. In Section 5, based on the Hamilton-Jacobi theory, the generating function to construct the canonical transformation is obtained from separated variables to action-angle variables in implicit form. Further, the functional independence of conserved integrals is proved in terms of the evolution of angle type variables. In addition, the Jacobi inversion problems for the Lie-Poisson Hamiltonian systems related to the TWRI system (1.1) are established.
We start from a general definition of the Lie-Poisson structure [43]. Let G be a Lie group and g be its Lie algebra. Let also g∗ be the dual of g with the natural pairing ⟨A,B⟩ between g and g∗. Consider the algebra of smooth function F∈C∞(g∗), then its gradient ∇F∈g is defined as
⟨x,∇F(y)⟩=limε→0F(y+εx)−F(y)ε,y,x∈g∗. |
For any smooth functions F and G∈C∞(g∗), Lie-Poisson brakect is defined as
{F,G}(y)=⟨y,[∇F(y),∇G(y)]⟩ | (2.1) |
When g is the matrix Lie-algebra gl3(C), we take the trace form ⟨A,B⟩=tr(AB) as the pairing between gl3(C) and gl3(C)∗ and fix the basis of gl3(C) as
Ekl=(δmlδnk)mn,1≤l,k≤3, |
where δij are the Kronecker delta functions, which satisfy the commutative relations
[Emn,Ekl]=δnkEml−δlmEkn. |
Since the trace form ⟨A,B⟩=tr(AB) is a non-degenerate pairing, we can make an identification gl3(C)≅gl3(C)∗. In the sense of ⟨Emn,ekl⟩=δmkδnl, one can find that the dual basis of {Ekl,1≤k,l≤3} is {ekl=Elk,1≤m,n≤3}. Hereafter, we choose y=3∑l,k=1yklekl∈gl3(C)∗ and ykl are coordinates on gl3(C)∗. In these coordinates, the gradient ∇F∈gl3(C) is determined by
∇F=3∑k,l=1∂F∂yklEkl. |
The Lie-Poisson bracket (2.1) become
{F,G}(y)=⟨y,[∇F(y),∇G(y)]⟩=tr(y[∇F(y),∇G(y)]) | (2.2) |
Specially, the brackets between the coordinates are
{ylk,ymn}=⟨y,[Ekl,Enm]⟩=δlnymk−δmkyln,1≤l,k,m,n≤3. | (2.3) |
Using the cyclicity of the trace, the Hamiltonian vector field associated with (2.5) for a smooth function F is represented as
XF(y)=[∇F(y),y]. |
For any g∈GL3(C) and X∈gl3(C), the adjoint action of g∈GL3(C) on its Lie algebra gl3(C) is
AdgX=gXg−1. |
Then for any ξ∈gl3(C)∗, the coadjoint action Ad∗g of GL3(C) in the dual space gl3(C)∗ is defined as
⟨Ad∗gy,X⟨=⟨y,Adg−1X⟩=⟨y,g−1Xg⟩=tr(yg−1Xg)=tr(gyg−1X)=⟨gyg−1,X⟩. |
It follows that Ad∗gy=gyg−1=Adgy. The coadjoint orbit passing through any y∈gl3(C)∗ is
Oy={Ad∗gy|g∈GL3(C)}={y=gyg−1|g∈GL3(C)} |
According to the representative theory of Lie group, we know that the coadjoint orbits of GL3(C) are the symplectic leaves of the Lie-Poisson structure on gl3(C)∗.
A smooth function h on gl3(C)∗ is a Casimir if an only if it is invariant under the coadjoint action
h(Ad∗gy)=h(y)∀g∈GL3(C),y∈gl3(C)∗. |
Choose ˜hk(y)=tr(yk), where y∈gl3(C)∗,k=1,2,3,⋯, we have
˜hk(Ad∗gy)=tr((Ad∗gy)k)=tr((gyg−1)k)=tr(gykg−1)=tr(yk)=˜hk(y). |
Thus, we choose three Casimir functions as tr(y),12tr(y2) and 13tr(y3).
To study the Lie-Poisson Hamiltonian systems, we will use the direct product of N copies of gl3(C)∗ which is denoted by (gl3(C)∗)N. The standard Lie-Poisson structure on (gl3(C)∗)N is given by
{F,G}=N∑j=1⟨yj,[∇jF,∇jG]⟩,∇jF=3∑k,l=1∂F∂ykljElk | (2.4) |
where yj∈gl3(C)∗,j=1,2,⋯,N.
The Hamiltonian vector field associated with Lie-Poisson bracket (2.4) for a smooth function F is
XjF(yj)=[∇jF(yj),yj],j=1,…,N | (2.5) |
and the 3N Casimir functions are tr(yj),12tr(y2j),13tr(y3j),j=1,…,N.
In this section, we shall use the gl3(C) rational Gaudin model governed by the Lax matrix
Vλ=(Vlk(λ))3×3=β+N∑j=1yjλ−λj,β=diag(β1,β2,β3) | (3.1) |
to study the TWRI system (βl are different constants).
To obtain the Lie-Poisson Hamiltonian systems from the Lax matrix (3.1), we choose
F1(λ)=trVλ,F2(λ)=12tr(V2λ),F3(λ)=13tr(V3λ) | (3.2) |
as the Hamiltonians. Let tkλ be the variables of Fk(λ), according to Hamiltonian vector field (2.5), the Lie-Poisson Hamiltonian systems for Fk(λ),k=1,2,3 are
yj,tkλ=[∇jFk(λ),yj]=1λ−λj[Vk−1λ,yj],j=1,…,N. | (3.3) |
Proposition 3.1. The Lax matrix Vτ satisfies the Lax equations along the Fk(λ)-flows:
ddtkλVτ=[1λ−τVk−1λ,Vτ],k=2,3 |
with λ, τ are two different constant spectral parameters.
Proof. By making use of (3.3), one infers
ddtkλVτ=N∑j=1τ−λjyj,tkλ=1λ−τ[Vk−1λ,N∑j=1yjτ−λj]−1λ−τ[Vk−1λ,N∑j=1yjλ−λj]=1λ−τ[Vk−1λ,Vτ−β]−1λ−τ[Vk−1λ,Vλ−β]=[1λ−τVk−1λ,Vτ]. |
Based on Proposition 3.1, for any λ,τ, a direct calculation shows that
{Fl(τ),Fk(λ)}=ddtkλFl(τ)=1ltr(ddtkλVlτ)=1ltr([1λ−τVk−1λ,Vlτ])=0, k,l=2,3, |
Therefore, Fk(λ),k=1,2,3 can be regarded as the generating function of integrals of the Hamiltonian systems generated from it. Using (3.1), we arrive at
F1(λ)=trβ+N∑j=1h1jλ−λj:=trβ+∞∑l=0F1,lλl+1,F2(λ)=12tr(β2)+N∑j=1E1,jλ−λj+N∑j=1h2j(λ−λj)2:=12tr(β2)+∞∑l=0F2,lλl+1, | (3.4) |
where
h1j=tr(yj),F1,l=N∑j=1λljh1j,E1,j=tr(βyj)+N∑k=1k≠jtr(yjyk)λj−λk,h2j=12tr(y2j),F2,l=N∑j=1λljE1,j+lN∑j=1λl−1jh2j,l=0,1,…. |
Similarly, we have
F3(λ)=13tr(V3λ)=13tr(β3)+N∑j=1E2,jλ−λj+N∑j=1E3,j(λ−λj)2+N∑j=1h3j(λ−λj)3:=13tr(β3)+∞∑l=0F3,lλl+1, | (3.5) |
where
E2,j=tr(β2yj)+N∑k=1k≠j1λj−λk[tr(βyjyk+βykyj)+N∑i=1i≠k,jtr(yjykyi+yjyiyk)3(λk−λi)+N∑i=1i≠k,jtr(yjykyi+yjyiyk)3(λj−λi)]+N∑k=1k≠jtr(y2kyj−y2jyk)(λj−λk)2E3,j=tr(βy2j)+N∑k=1k≠jtr(y2jyk)λj−λk,h3j=13tr(y3j),F3,l=N∑j=1λljE2,j+lN∑j=1λl−1jE3,j+12l(l−1)N∑j=1λl−2jh3j,l=0,1,…. |
which implies {Elj,Ekm}=0, l,k=1,2,3,j,m=1,…,N and the following fact.
Corollary 3.1. Functions F1,l,F2,l,F3,l(l≥1) are in involution in pairs with respect to the Lie-Poisson bracket (2.4).
Now, we consider two Lie-Poisson Hamiltonian systems generated by the Hamiltonians
H=γ0F1,1+γ1(trβF1,1−F2,1+12(F1,0)2)+γ3P(β1)P(β2)+γ2[F3,1+12[(trβ)2−tr(β2)]F1,1−trβF2,1+12trβ(F1,0)2−F1,0F2,0]+γ4P(β1)P(β3)+γ5P(β2)P(β3) | (3.6) |
=α1N∑j=1λjy11j+α2N∑j=1λjy22j+α3N∑j=1λjy33j+c−112N∑j=1y12jN∑j=1y21j+c−113N∑j=1y13jN∑j=1y31j+c−123N∑j=1y23jN∑j=1y32j |
where
γ0=α1(β2−β3)β21+α2(β3−β1)β22+α3(β1−β2)β23(β1−β2)(β2−β3)(β1−β3),γ1=α1(β3−β2)β1+α2(β1−β3)β2+α3(β2−β1)β3(β1−β2)(β2−β3)(β1−β3),γ2=α1(β2−β3)+α2(β3−β1)+α3(β1−β2)(β1−β2)(β2−β3)(β1−β3),γ3=α2−α1(β1−β2)3(β2−β3)(β1−β3),γ4=α1−α3(β1−β2)(β2−β3)(β1−β3)2,γ5=α3−α2(β1−β2)(β2−β3)2(β1−β3),P(βj)=β2jF1,0−βj(trβF1,0−F2,0)+F3,0+13[(trβ)2−tr(β2)]F1,0−trβF2,0 |
and
H1=F2,1−(β2−β3)2P2(β1)−(β1−β3)2P2(β2)+(β1−β2)2P2(β3)2(β1−β2)2(β1−β3)2(β2−β3)2=β1N∑j=1λjy11j+β2N∑j=1λjy22j+β3N∑j=1λjy33j+N∑j=1y12jN∑j=1y21j+N∑j=1y13jN∑j=1y31j+N∑j=1y23jN∑j=1y32j. | (3.7) |
respectively, where α1,α2,α3 are different constants and clk=βl−βkαl−αk,1≤l,k≤3,l≠k.
The equations of motion for H and H1 are
yj,x=[∇jH,yj], j=1,…,N | (3.8) |
and
yj,t=[∇jH1,yj], j=1,…,N | (3.9) |
which are exactly the Lie-Poisson Hamiltonian systems associated with TWRI system (1.1).
In fact, the Lie-Poisson Hamiltonian systems (3.8) and (3.9) are generated respectively by the N copies of adjoint representations of the spectral problems
φx=Uφ,φ=(φ1φ2φ3),U=λA+U0=λ(α1000α2000α3)+(0u12u13u210u23u31u320) | (3.10) |
and
φt=Wφ,W=(β1λc12u12c13u13c21u21β2λc23u23c31u31c32u32β3λ) | (3.11) |
under the Bargmann constraint
ulk=c−1lkN∑j=1ylkj,(1≤l,k≤3,l≠k). | (3.12) |
Proposition 3.2. The Lie-Poisson Hamiltonian systems (3.8) and (3.9) admit the Lax representations
ddxVλ=[U,Vλ] |
and
ddtVλ=[W,Vλ], |
respectively, where
U=(α1λu12u13u21α2λu23u31u32α3λ),W=(β1λc12u12c13u13c21u21β2λc23u23c31u31c32u32β3λ) |
with
U0=(0u12u13u210u23u31u320)=f(y1,⋯,yN)=(0c−112∑Nj=1y12jc−113∑Nj=1y13jc−121∑Nj=1y21j0c−123∑Nj=1y23jc−131∑Nj=1y31jc−132∑Nj=1y32j0). |
It follows that the integrals of motion of the Lie-Poisson Hamiltonian systems (3.8) and (3.9) are provided by Fk(λ),k=1,2,3.
Now we know that there are 3N Casimir functions of the Lie-Poisson structure (2.4): tr(yj),12tr(y2j),13tr(y3j),j=1,…,N and 3N involutive first integrals E1,j,E2,j,E3,j,j=1,⋯,N. Hence, they are integrable.
The first two typical members of TWRI vector fields {Xm} are [26]
X0=(0(β1−β2)u12(β1−β3)u13(β2−β1)u210(β2−β3)u23(β3−β1)u31(β3−β2)u320), | (3.13) |
X1=(0c12u12x+(c13−c23)u13u32c13u13x+(c12−c23)u12u23c21u21x+(c23−c13)u23u310c23u23x+(c12−c13)u21u13c31u31x+(c23−c12)u21u32c32u32x+(c13−c12)u12u310). | (3.14) |
The zero-curvature equation Ut=Wx−[U,W] leads to the TWRI system (1.1)
(U0)t=X1. |
From Corollary 3.1, it is not difficult to verify the involutivity {H,H1}=0, which implies the commutativity of Hamiltonian vector fields. The importance of Hamiltonians H,H1 are that the differential f∗ maps [∇jH,yj], [∇jH1,yj] exactly into the TWRI vector fields X0,X1
f∗([∇1H,y1],⋯,[∇NH,yN])=X0,f∗([∇1H1,y1],⋯,[∇NH1,yN])=X1, |
respectively. Thus solutions of TWRI system (1.1) can be obtained by solving two compatible Lie-Poisson Hamiltonian systems with ordinary differential equations:
Proposition 3.3. Let yj be a compatible solution of the Lie-Poisson Hamiltonian systems (3.8) and (3.9). Then
ulk=c−1lkN∑j=1ylkj,1≤l,k≤3,l≠k |
solves TWRI system (1.1).
In this section, we will give the separated canonical coordinates on the common level set of the Casimir functions to deal with the Lie-Poisson Hamiltonian systems.
Remark 1. Let (C1j,C2j,C3j) be regular values of the map defined by the Casimirs: yj→(tr(yj),12tr(y2j),13tr(y3j)). Then restricted on the common level set of Casimir functions
{y1,⋯,yj,⋯,yN|tr(yj)=C1j,12tr(y2j)=C2j,13tr(y3j)=C3j,j=1,…,N}, | (4.1) |
the 9N dimensional Poisson manifold (gl3(C)∗)N is naturally reduced to a 6N dimensional symplectic manifold, by which 3N pairs of canonical variables can be introduced.
In the following, we give the first 3N−2 pairs of separated variables μi,νi by Sklyanin's method [8].
In fact, the characteristic polynomial of Lax matrix Vλ for the TWRI system (1.1) is a constant independent of variables x and t with the expansion
det(zI−Vλ)=z3−F1(λ)z2+(12F21(λ)−F2(λ))z−(F3(λ)−F2(λ)F1(λ)+16F31(λ))=0, | (4.2) |
which defines a non-hyperelliptic algebraic curve by introducing variable ζ=a(λ)z:
ζ3−a(λ)F1(λ)ζ2+a2(λ)(12F21(λ)−F2(λ))ζ−a3(λ)(F3(λ)−F2(λ)F1(λ)+16F31(λ))=0 |
with
a(λ)=N∏j=1(λ−λj). |
Follow the method in [8], the canonical separated variables μi(i=1,…,3N−2) are defined by zeros of some polynomial B(λ) and the corresponding conjugate coordinates νi(i=1,…,3N−2) related to μi by the equation
ν3i−F1(μi)ν2k+(12F21(μi)−F2(μi))νi−(F3(μi)−F2(μi)F1(μi)+16F31(μi))=0, | (4.3) |
It follows from (2.3) that νi are eigenvalues of the matrix Vμi. Therefore, there must exists a similarity transformation Vμi→˜Vμi=PVμiP−1 for each i such that the matrix ˜Vμi is block-triangular
˜V21(μi)=˜V31(μi)=0 | (4.4) |
and νi is the eigenvalue of Vμi splitted from the upper block,
νi=˜V11(μi). | (4.5) |
Thus, the problem is reduced to determining the matrix P and polynomial B(λ). Let P be of the form:
P=(100p10001). |
Note that the matrix
˜Vλ=PVλP−1=(V11(λ)−pV12(λ)V12(λ)V13(λ)V21(λ)+pV11(λ)−p(pV12(λ)+V22(λ))V22(λ)+pV12(λ)V23+pV13(λ)V31(λ)−pV32(λ)V32(λ)V33(λ)) |
depends on two parameters λ and p. Hence, we can consider the condition (4.4) as the set of two algebraic equations
{˜V21(λ)=V21(λ)+pV11(λ)−p(pV12(λ)+V22(λ))=0,˜V31(λ)=V31(λ)−pV32(λ)=0 | (4.6) |
for two variable λ and p. Eliminating p from (4.6), one obtains the polynomial equation for λ:
B(λ)=V32(λ)V31(λ)[V11(λ)−V22(λ)]+V232(λ)V21(λ)−V231(λ)V12(λ)=0. | (4.7) |
Based on (3.1) and (4.7), the polynomial B(λ) of degree 3N−2 can be denote as
B(λ):=(β1−β2)N∑j=1y31jN∑j=1y32jn(λ)a3(λ), | (4.8) |
where
n(λ)=3N−2∏i=1(λ−μi),a(λ)=N∏j=1(λ−λj):=N∑j=0ajλN−j,(a0=1). | (4.9) |
Expressing p from ˜V31(λ)=0 as p=−V31(λ)/V32(λ) and substituting it into Eq (4.5) of νi, we arrive at
νi=˜V11(μi)=V11(μi)−V12(μi)V31(μi)V32(μi),i=1,…,3N−2, | (4.10) |
which produces 3N−2 pairs of variables μi,νi. It is easy to see from (2.3) and the above expressions that
{Vlk(τ),Vmn(λ)}=1λ−τ[(Vmk(τ)−Vmk(λ))δln−(Vln(τ)−Vln(λ))δmk],1≤l,k,m,n≤3 | (4.11) |
with λ,τ are two different constant spectral parameters. We have from (4.11) that
A(λ)=V11(λ)−V12(λ)V31(λ)V32(λ), | (4.12) |
which, together with B(λ) defined by (4.7), B(λ) and A(τ) satisfy
{{A(τ),A(λ)}=0,{B(τ),B(λ)}=0,{A(τ),B(λ)}=1λ−τ(V232(λ)V232(τ)B(τ)−B(λ)). | (4.13) |
Proposition 4.1. The canonical separated variables μj and νj constructed from B(λ) in (4.8) and A(μj) in (4.10) satisfy the relations
{μi,μj}=0,{νi,νj}=0,{νi,μj}=δij1≤i,j≤3N−2. |
Proof. The commutativity of B in (4.13) implies the commutativity of μj (zeros of B(λ)). The Poisson brackets of νj can be calculated by using the definition of μj. From B(μj)=0 for j=1,…,3N−2, it follows that
0={F,B(μj)}={F,B(λ)}|λ=μj+B′(μj){F,μj} |
that is
{F,μj}=−{F,B(λ)}|λ=μjB′(μj), | (4.14) |
for any function F. In the same way we have
{νi,F}={A(μi),F}={A(τ),F}|τ=μi+A′(μi){μi,F}. |
Now we turn to prove {νi,μj}=δij. Noting that
{νi,μj}={A(τ),μj}|τ=μi+A′(μi){μi,μj}={A(τ),μj}|τ=μi, |
we obtain by using (4.14) and the third equation in (4.13) that
{νi,μj}=−{A(τ),B(λ)}|τ=μiλ=μjB′(μj)=1μj−μi(V232(μj)V232(μi)B(μi)−B(μj))1B′(μj), | (4.15) |
which vanishes for μi≠μj due to B(μi)=B(μj)=0 and is evaluated via L'Hôpital rule for μi=μj to produce the proclaimed result. Similarly, the commutativity of νi(1≤i≤3N−2) can be shown by the first equation of (4.13).
Apart from the 3N−2 pairs of separated variables above, we should add 2 pairs of conjugate variables to prove the canonical structure on the common level set of Casimir functions (4.1).
The last 2 pairs of conjugate variables can be defined as follows by direct calculation.
Proposition 4.2. Assume that the 2 pairs of additional canonical separated variables are defined by
μ3N−1=lnN∑j=1y31j,ν3N−1=(β2+β3)F2,0−F3,0(β2−β1)(β3−β1),μ3N=lnN∑j=1y32j,ν3N=G0−(β1+β3)F2,0(β2−β1)(β3−β2). | (4.16) |
Then on the common level set of Casimir functions (4.1), we have
{μi,μj}=0,{νi,νj}=0,{μi,νj}=δij,i,j=1,…,3N. | (4.17) |
It is shown that μj,νj,j=1,⋯,N are 3N pairs of conjugate variables.
In this section, the action-angle variables will be introduced by resorting to Hamilton-Jacobi theory. As a by-product, the functional independence of conserved integrals for the Liouville integrability of Lie-Poisson Hamiltonian systems (3.8) and (3.9) will be proved. Further, the Jacobi inversion problems for systems (3.8), (3.9) and TWRI Eq (1.1) will be built by using the canonical transformation from the separated variables to the action-angle variables
Let
12trβ2+N∑j=1E1,jλ−λj:=b2(λ)a(λ):=12trβ2+∞∑l=0flλl+1,13trβ3+N∑j=1E2,jλ−λj+N∑j=1E3,j(λ−λj)2:=b3(λ)a2(λ):=13trβ3+∞∑l=0glλl+1, |
where
b2(λ)=12trβ2λN+I0λN−1+I1λN−2+⋯+IN−3λ2+IN−2λ+IN−1,b3(λ)=13trβ3λ2N+I˜0λ2N−1+INλ2N−2+⋯+I3N−3λ+I3N−2, | (5.1) |
from which we can rewrite the generating functions F1(λ),F2(λ),F3(λ) as
F1(λ)=trβ+N∑j=1C1jλ−λj:=R1(λ)a(λ),F2(λ)=b2(λ)a(λ)+N∑j=1C2j(λ−λj)2=12trβ2+∞∑l=0flλl+1+N∑j=1C2j(λ−λj)2:=R2(λ)a2(λ),F3(λ)=b3(λ)a2(λ)+N∑j=1C3j(λ−λj)3=13trβ3+∞∑l=0glλl+1+N∑j=1C3j(λ−λj)3 | (5.2) |
with
R1(λ)=a(λ)(trβ+N∑j=1C1jλ−λj),R2(λ)=a(λ)b2(λ)+a2(λ)N∑j=1C2j(λ−λj)2, |
The comparison of the coefficients of λl,l=0,…,N−1 in equation
b2(λ)=a(λ)(12trβ2+∞∑l=0flλl+1) |
and the comparison of the coefficients of λl,l=0,1,…,2N−1 in equation
b3(λ)=a2(λ)(13trβ3+∞∑l=0glλl+1), |
respectively, yield
I0=12a1trβ2+f0=12a1trβ2+(β3−β1)ν3N−1+(β3−β2)ν3N,I˜0=23a1trβ3+g0=23a1trβ3+(β23−β21)ν3N−1+(β23−β22)ν3N,Ij=j∑i=0aifj−i+12aj+1trβ2,j=1,…,N−1,IN+k=k+1∑l=0(∑i,j≥0i+j=laiaj)gk+1−l+13∑i,j≥0i+j=k+2aiajtrβ3,k=0,…,2N−2. |
Let
νi=∂S∂μi,i=1,…,3N−2. |
We obtain from (4.3) that the completely separated Hamilton-Jacobi equations:
(∂S∂μi)3−R1(μi)a(μi)(∂S∂μi)2+(R21(μi)2a2(μi)−b2(μi)a(μi)−N∑j=1C2j(μi−λj)2)∂S∂μi−(b3(μi)a2(μi)+N∑j=1C3j(μi−λj)3−(b2(μi)a(μi)+N∑j=1C2j(μi−λj)2)R1(μi)a(μi)+R31(μi)6a3(μi))=0, |
with i=1,…,3N−2, from which we get an implicit complete integral of Hamilton-Jacobi equations for the generating functions F2(λ) and F3(λ):
S=3N−2∑j=1Sj(μj)=S(μ1,⋯,μ3N−2;I1,⋯,I3N−2)=3N−2∑j=1∫μj0zdλ, | (5.3) |
where z satisfies (4.2).
Now let us consider a canonical transformation from the variables μi,νi,(i=1,…,3N−2) to variables ϕi and Ii,(i=1,…,3N−2), generated by the generating function S:
3N−2∑i=1νidμi+3N−2∑i=1ϕidIi=dS, |
that satisfies
νi=∂S∂μi,ϕi=∂S∂Ii,i=1,⋯,3N−2. | (5.4) |
Resorting to Eqs (5.3), (5.4), (4.2) and (5.2), we arrive at
ϕi=∂S∂Ii=3N−2∑j=1∫μj0∂z∂Iidλ={3N−2∑j=1∫μj0(a(λ)z−R1(λ))λN−i−1R(λ)dλ,i=1,…,N−1,3N−2∑j=1∫μj0λ3N−i−2R(λ)dλ,i=N,…,3N−2, | (5.5) |
where R(λ)=3a2(λ)z2−2a(λ)R1(λ)z+12R21(λ)−R2(λ). From Eqs (5.1) and (5.2), the generating functions of integrals can be rewritten as
F2(λ)=N∑j=1C2j(λ−λj)2+12trβ2λN+I0λN−1+I1λN−2+⋯+IN−1a(λ):=K2(I1,⋯,IN−1,λ), |
F3(λ)=N∑j=1C3j(λ−λj)3+13trβ3λ2N+I˜0λ2N−1+INλ2N−2+⋯+I3N−2a2(λ):=K3(IN,…,I3N−2,λ). |
Functions I1,⋯,I3N−2 and ϕ1,…,ϕ3N−2 are variables of action type and the corresponding variables of angles, respectively. In the following, we will use these action-angle variables to discuss the equations of motion for the Lie-Poisson Hamiltonian systems generated by the Lax matrix (3.1). The Hamiltonian canonical equations for the generating functions F2(λ) and F3(λ) in terms of action-angle variables Ij and ϕj, j=1,…,3N−2, are as follows
ϕj,t2λ={∂K2(λ)∂Ij=λN−j−1a(λ),1≤j≤N−1∂K2(λ)∂Ij=0,N≤j≤3N−2, | (5.6) |
Ij,t2λ=−∂K2(λ)∂ϕj=0,1≤j≤3N−2, | (5.7) |
ϕj,t3λ={∂K3(λ)∂Ij=0,1≤j≤N−1∂K3(λ)∂Ij=λ3N−j−2a2(λ),N≤j≤3N−2, | (5.8) |
Ij,t3λ=−∂K3(λ)∂ϕj=0,1≤j≤3N−2. | (5.9) |
Let t2,l and t3,l represent the variables of F2,l-flow and F3,l-flow, respectively. According to the definition of the Lie-Poisson bracket, one infers
Ij,t2λ=∞∑l=01λl+1{Ij,F2,l}=∞∑l=01λl+1dIjdt2,lIj,t3λ=∞∑l=01λl+1{Ij,F3,l}=∞∑l=01λl+1dIjdt3,lϕj,t2λ=∞∑l=01λl+1{ϕj,F2,l}=∞∑l=01λl+1dϕjdt2,lϕj,t3λ=∞∑l=01λl+1{ϕj,F3,l}=∞∑l=01λl+1dϕjdt3,l,j=1,…,3N−2. | (5.10) |
Proposition 5.1.
(dϕdt2,1,⋯,dϕdt2,N−1,dϕdt3,1,⋯,dϕdt3,2N−1)=(Q1100Q22), | (5.11) |
with
Q11=(1A1A2⋯AN−21A1⋯AN−3⋱⋱⋮1A11),Q22=(1B1B2⋯B2N−21B1⋯B2N−3⋱⋱⋮1B11) |
where Ak's are the coefficients in the expansion
λNa(λ)=∞∑k=0Akλk, |
which could be represented through the power sums of λj, that is
A0=1,A1=s1,A2=12(s2+s21) |
with the recursive formula
Ak=1k(sk+∑m,n≥1m+n=ksmAn),sk=N∑j=1λkj, |
and Br's are obtained by comparing the coefficients of λr,r=0,1,…, in equation
λ2Na2(λ)=(∞∑k=0Akλk)2=∞∑r=0Brλr, |
where B0=A20=1,B1=2A1,⋯,Br=∑m,n≥0m+n=rAmAn with the supplementary definition A−k = B−k=0, k=1,2,….
Proof. Using (5.6), (5.8) and (5.10), it is easy to see that
∞∑l=01λl+1{Ij,F2,l}=∞∑l=01λl+1{Ij,F3,l}=0,j=1,…,3N−2,∞∑l=01λl+1{ϕj,F2,l}=λN−j−1a(λ)=∞∑k=0Akλk+j+1,j=1,…,N−1,∞∑l=01λl+1{ϕj,F3,l}=0,j=1,…,N−1,∞∑l=01λl+1{ϕj,F2,l}=0,j=N,…,3N−2,∞∑l=01λl+1{ϕj,F3,l}=λ3N−j−2a2(λ)=∞∑k=0Bkλk+j+2−N,j=N…,3N−2. | (5.12) |
By comparing the coefficients of λ−l−1 in (5.12), we deduce the Poisson brackets
{Ij,F2,l}=0,{Ij,F3,l}=0,j=1,…,3N−2,{ϕj,F2,l}=0,{ϕj,F3,l}=0,j=1,…,3N−2,{ϕj,F2,l}=Al−j,{ϕj,F3,l}=0,j=1,…,N−1,{ϕj,F2,l}=0,{ϕj,F3,l}=Bl+N−j−1,j=N,…,3N−2. | (5.13) |
Thus, the non-degenerate matrix takes on the form (5.11).
Proposition 5.2. F2,1,⋯,F2,N−1,F3,1,⋯,F3,2N−1 given in (3.4) and (3.5) are functionally independent.
Proof. We need only prove the linear independence of the gradients: ∇F2,1,⋯,∇F2,N−1, ∇F3,1,⋯∇F3,2N−1. Suppose
N−1∑k=1ck∇F2,k+2N−1∑m=1cN+m−1∇F3,m=0. |
It is easy to calculate that
0=N−1∑k=1ck{ϕj,F2,k}+2N−1∑m=1cN+m−1{ϕj,F3,m}=N−1∑k=1ckdϕjdt2,k+2N−1∑m=1cN+m−1dϕjdt3,m, |
which implies that c1=c2=⋯=c3N−2=0 because the coefficient determinant is equal to 1 by (5.11).
Remark 2. From Corollary 3.1 and the above Proposition, it is proved that the Lie-Poisson Hmailtonian systems (3.8) and (3.9) with the Hamiltonians (3.6) and (3.7) are complete integrable in the Liouville sense because their 3N−2 integrals F2,1,⋯,F2,N−1, F3,1,⋯,F3,2N−1 are involutive in pairs and functionally independent.
For given values of the 3N Casimir functions in (4.1), F0,l=∑Nj=1λljC1j are constants, which means that {ϕj,F0,l}=0. Based on (5.13) and (3.6), the solution of the Lie-Poisson Hmailtonian system (3.8) in terms of action-angle variables ϕj and Ij is
Ij(x)=Ij(0),ϕj(x)={ϕj(0)−(γ2trβ+γ1)A1−jx,j=1,…,N−1,ϕj(0)+γ2BN−jx,j=N,…,3N−2. | (5.14) |
Thus, combining (5.5) with (5.14) give rise to the Jacobi inversion problem
{ϕj(0)−(γ2trβ+γ1)A1−jx=3N−2∑k=1∫μk0(a(λ)z−R1(λ))λN−j−1R(λ)dλ,j=1,…,N−1,ϕj(0)+γ2BN−jx=3N−2∑k=1∫μk0λ3N−j−2R(λ)dλ,j=N,…,3N−2. |
For the Lie-Poisson Hamiltonian system (3.9) with respect to Lie-Poisson bracket (5.13), we obtain by using (3.7) that
Ij(t)=Ij(0),ϕj(t)={ϕj(0)+A1−jt,j=1,…,N−1,ϕj(0),j=N,…,3N−2. | (5.15) |
With the help of (5.5) and (5.15), we deduce the Jacobi inversion problem
{ϕj(0)+A1−jt=3N−2∑k=1∫μk0(a(λ)z−R1(λ))λN−j−1R(λ)dλ,j=1,…,N−1,ϕj(0)=3N−2∑k=1∫μk0λ3N−j−2R(λ)dλ,j=N,…,3N−2. |
The compatible solution of Lie-Poisson Hamiltonian systems (3.8) and (3.9) in terms of action-angle variables ϕj and Ij is
Ij(x,t)=Ij(0,0),ϕj(x,t)={ϕj(0,0)−(γ2trβ+γ1)A1−jx+A1−jt,j=1,…,N−1,ϕj(0,0)+γ2BN−jx,j=N,…,3N−2. | (5.16) |
Thus, by making use of (5.5) and (5.16), we arrive at the Jacobi inversion problem for the TWRI Eq (1.1)
{ϕj(0,0)−(γ2trβ+γ1)A1−jx+A1−jt=3N−2∑k=1∫μk0(a(λ)z−R1(λ))λN−j−1R(λ)dλ,j=1,…,N−1,ϕj(0,0)+γ2BN−jx=3N−2∑k=1∫μk0λ3N−j−2R(λ)dλ,j=N,…,3N−2. |
In this paper, two finite dimensional Lie-Poisson Hamiltonian systems associated with a 3×3 spectral problem related to three-wave resonant interaction system are presented with the help of nonlinearization method. In the framework of Lie-Poisson structure, it is easier to prove the integrability for these finite-dimensional Lie-Poisson Hamiltonian systems in Liouville sense. 3N pairs of separation of variables for these integrable systems with non-hyperelliptic spectral curves are constructed and 3N−2 pairs of them are proposed by using Sklyanin's method. In addition, apart from the variables μk,νk(k=1,⋯,3N−2), we add two pairs of conjugate variables. 3N−2 pairs of action-angle variables are introduced with the help of Hamilton-Jacobi theory. The Jacobi inversion problems for the these Lie-Poisson Hamiltonian systems and three-wave resonant interaction system are discussed. Furthermore, based upon the Jacobi inversion problems, we may use the algebro-geometric method to get the multi-variable sigma-function solutions, which will be left to a future publication. The methods in this paper can be applied to other systems of soliton hierarchies with 3×3 matrix spectral problems.
This work was supported by the National Natural Science Foundation of China (Grant Nos.12001013, 11626140 and 11271337).
The authors declare that they have no competing interests.
[1] |
M. Gaudin, Diagonalisation d′une classe d′Hamiltoniens de spin, J. Phys. France., 37 (1976), 1089–1098. https://doi.org/10.1051/jphys:0197600370100108700 doi: 10.1051/jphys:0197600370100108700
![]() |
[2] | M. Gaudin, The bethe wavefunction (Français), Paris: Masson, 1983. https://doi.org/10.1017/CBO9781107053885 |
[3] |
B. Jurˇco, Classical Yang-Baxter equations and quantum integrable systems, J. Math. Phys., 30 (1989), 1289. https://doi.org/10.1063/1.528305 doi: 10.1063/1.528305
![]() |
[4] | A. Reyman, M. Semenov-Tian-Shansky, Group-theoretical methods in the theory of finite-dimensional integrable systems, In: Dynamical systems vii, Berlin: Springer, 1994,116–225. https://doi.org/10.1007/978-3-662-06796-3_7 |
[5] |
B. Feigin, E. Frenkel, N. Reshetikhin, Gaudin model, Bethe ansatz and critical level, Commun. Math. Phys., 166 (1994), 27–62. https://doi.org/10.1007/BF02099300 doi: 10.1007/BF02099300
![]() |
[6] |
A. Hone, V. Kuznetsov, O. Ragnisco, B¨acklund transformations for the sl(2) Gaudin magnet, J. Phys. A: Math. Gen., 34 (2001), 2477. https://doi.org/10.1088/0305-4470/34/11/336 doi: 10.1088/0305-4470/34/11/336
![]() |
[7] |
E. Sklyanin, Separation of variables in the Gaudin model, J. Soviet. Math., 47 (1989), 2473–2488. https://doi.org/10.1007/BF01840429 doi: 10.1007/BF01840429
![]() |
[8] |
E. Sklyanin, Separation of variables in the classical integrable SL(3) magnetic chain, Commun. Math. Phys., 150 (1992), 181–191. https://doi.org/10.1007/BF02096572 doi: 10.1007/BF02096572
![]() |
[9] |
V. Kuznetsov, Quadrics on real Riemannian spaces of constant curvature: separation of variables and connection with Gaudin magnet, J. Math. Phys., 33 (1992), 3240. https://doi.org/10.1063/1.529542 doi: 10.1063/1.529542
![]() |
[10] |
D. Du, X. Geng, On the relationship between the classical Dicke-Jaynes-Cummings-Gaudin model and the nonlinear Schrödinger equation, J. Math. Phys., 54 (2013), 053510. https://doi.org/10.1063/1.4804943 doi: 10.1063/1.4804943
![]() |
[11] |
E. Kalnins, V. Kuznetsov, W. Miller, Quadrics on complex Riemannian spaces of constant curvature, separation of variables, and the Gaudin magnet, J. Math. Phys., 35 (1994), 1710. https://doi.org/10.1063/1.530566 doi: 10.1063/1.530566
![]() |
[12] |
J. Eilbeck, V. Enol'skii, V. Kuznetsov, A. Tsiganov, Linear r-matrix algebra for classical separable systems, J. Phys. A: Math. Gen., 27 (1994), 567. https://doi.org/10.1088/0305-4470/27/2/038 doi: 10.1088/0305-4470/27/2/038
![]() |
[13] |
J. Harnad, P. Winternitz, Classical and quantum integrable systems in 263-1263-1263-1 and separation of variables, Commun. Math. Phys., 172 (1995), 263–285. https://doi.org/10.1007/BF02099428 doi: 10.1007/BF02099428
![]() |
[14] |
C. Cao, Y. Wu, X. Geng, Relation between the Kadometsev-Petviashvili equation and the confocal involutive system, J. Math. Phys., 40 (1999), 3948. https://doi.org/10.1063/1.532936 doi: 10.1063/1.532936
![]() |
[15] |
X. Yang, D. Du, From nonlinear Schr¨odinger hierarchy to some (2+1)-dimensional nonlinear pseudodifferentiale quations, J. Math. Phys., 51 (2010), 083505. https://doi.org/10.1063/1.3453389 doi: 10.1063/1.3453389
![]() |
[16] |
D. Du, X. Yang, An alternative approach to solve the mixed AKNS equations, J. Math. Anal. Appl., 414 (2014), 850–870. https://doi.org/10.1016/j.jmaa.2014.01.041 doi: 10.1016/j.jmaa.2014.01.041
![]() |
[17] |
D. Scott, Classical functional Bethe ansatz for SL(N): separation of variables for the magnetic chain, J. Math. Phys., 35 (1994), 5831. https://doi.org/10.1063/1.530712 doi: 10.1063/1.530712
![]() |
[18] |
M. Gekhtman, Separation of variables in the classical SL(N) magnetic chain, Commun. Math. Phys., 167 (1995), 593–605. https://doi.org/10.1007/BF02101537 doi: 10.1007/BF02101537
![]() |
[19] |
E. Sklyanin, Separation of variables: new trends, Prog. Theor. Phys. Supp., 118 (1995), 35–60. https://doi.org/10.1143/PTPS.118.35 doi: 10.1143/PTPS.118.35
![]() |
[20] |
B. Dubrovin, T. Skrypnyk, Separation of variables for linear Lax algebras and classical r-matrices, J. Math. Phys., 59 (2018), 091405. https://doi.org/10.1063/1.5031769 doi: 10.1063/1.5031769
![]() |
[21] |
S. Manakov, Example of a completely integrable nonlinear wave field with nontrivial dynamics (lee model), Theor. Math. Phys., 28 (1976), 709–714. https://doi.org/10.1007/BF01029027 doi: 10.1007/BF01029027
![]() |
[22] |
R. Beals, D. Sattinger, On the complete integrability of completely integrable systems, Commun. Math. Phys., 138 (1991), 409–436. https://doi.org/10.1007/BF02102035 doi: 10.1007/BF02102035
![]() |
[23] |
M. Adams, J. Harnad, J. Hurtubise, Darboux coordinates and Liouville-Arnold integration in loop algebras, Commun. Math. Phys., 155 (1993), 385–413. https://doi.org/10.1007/bf02097398 doi: 10.1007/bf02097398
![]() |
[24] |
L. Dickey, Integrable nonlinear equations and Liouville's theorem, I, Commun. Math. Phys., 82 (1981), 345–360. https://doi.org/10.1007/BF01237043 doi: 10.1007/BF01237043
![]() |
[25] | L. Dickey, Soliton equations and Hamiltonian systems, Singapore: World Scientific Publishing, 2003. https://doi.org/10.1142/5108 |
[26] |
Y. Wu, X. Geng, A finite-dimensional integrable system associated with the three-wave interaction equations, J. Math. Phys., 40 (1999), 3409. https://doi.org/10.1063/1.532896 doi: 10.1063/1.532896
![]() |
[27] |
F. Calogero, A. Degasperis, Novel solution of the system describing the resonant interaction of three waves, Physica D, 200 (2005), 242–256. https://doi.org/10.1016/j.physd.2004.11.007 doi: 10.1016/j.physd.2004.11.007
![]() |
[28] | V. Zakharov, S. Manakov, Resonant interaction of wave packets in nonlinear media, JETP Lett., 18 (1973), 243–245. |
[29] | V. Zakharov, S. Manakov, The theory of resonant interaction of wave packets in nonlinear media, Sov. Phys. JETP Lett., 42 (1975), 842–850. |
[30] |
D. Kaup, The three-wave interaction-a nondispersive phenomenon, Stud. Appl. Math., 55 (1976), 9–44. https://doi.org/10.1002/sapm19765519 doi: 10.1002/sapm19765519
![]() |
[31] |
R. Conte, A. Grundland, M. Musette, A reduction of the resonant three-wave interaction to the generic sixth Painlevé equation, J. Phys. A: Math. Gen., 39 (2006), 12115. https://doi.org/10.1088/0305-4470/39/39/S07 doi: 10.1088/0305-4470/39/39/S07
![]() |
[32] |
R. Buckingham, R. Jenkins, P. Miller, Semiclassical soliton ensembles for the three-wave resonant interaction equations, Commun. Math. Phys., 354 (2017), 1015–1100. https://doi.org/10.1007/s00220-017-2897-7 doi: 10.1007/s00220-017-2897-7
![]() |
[33] |
G. Huang, Exact solitary wave solutions of three-wave interaction equations with dispersion, J. Phys. A: Math. Gen., 33 (2000), 8477. https://doi.org/10.1088/0305-4470/33/47/310 doi: 10.1088/0305-4470/33/47/310
![]() |
[34] | G. Valiulis, K. Staliˉunas, On the subject of the integrability and soliton solutions of three-wave interaction equations, Lith. J. Phys., 31 (1991), 61. |
[35] |
F. Baronio, M. Conforti, A. Degasperis, S. Lombardo, Rogue waves emerging from the resonant interaction of three waves, Phys. Rev. Lett., 111 (2013), 114101. https://doi.org/10.1103/PhysRevLett.111.114101 doi: 10.1103/PhysRevLett.111.114101
![]() |
[36] |
A. Degasperis, S. Lombardo, Rational solitons of wave resonant-interaction models, Phys. Rev. E, 88 (2013), 052914. https://doi.org/10.1103/PhysRevE.88.052914 doi: 10.1103/PhysRevE.88.052914
![]() |
[37] |
X. Wang, J. Cao, Y. Chen, Higher-order rogue wave solutions of the three-wave resonant interaction equation via the generalized Darboux transformation, Phys. Scr., 90 (2015), 105201. https://doi.org/10.1088/0031-8949/90/10/105201 doi: 10.1088/0031-8949/90/10/105201
![]() |
[38] |
B. Yang, J. Yang, General rogue waves in the three-wave resonant interaction systems, IMA J. Appl. Math., 86 (2021), 378–425. https://doi.org/10.1093/imamat/hxab005 doi: 10.1093/imamat/hxab005
![]() |
[39] |
X. Geng, Y. Li, B. Xue, A second-order three-wave interaction system and its rogue wave solutions, Nonlinear. Dyn., 105 (2021), 2575–2593. https://doi.org/10.1007/s11071-021-06727-2 doi: 10.1007/s11071-021-06727-2
![]() |
[40] |
G. He, X. Geng, L. Wu, Algebro-geometric quasiperiodic solutions to the three-wave resonant interaction hierarchy, SIAM J. Math. Anal., 46 (2014), 1348–1384. https://doi.org/10.1137/130918794 doi: 10.1137/130918794
![]() |
[41] |
X. Geng, Y. Zhai, H. Dai, Algebro-geometric solutions of the coupled modified Korteweg-de Vries hierarchy, Adv. Math., 263 (2014), 123–153. https://doi.org/10.1016/j.aim.2014.06.013 doi: 10.1016/j.aim.2014.06.013
![]() |
[42] |
J. Wei, X. Geng, X. Zeng, The Riemann theta function solutions for the hierarchy of Bogoyavlensky lattices, T. Am. Math. Soc., 371 (2019), 1483–1507. https://doi.org/10.1090/tran/7349 doi: 10.1090/tran/7349
![]() |
[43] | G. Arutyunov, Elements of classical and quantum integrable systems, Switzerland: Springer, 2019. https://doi.org/10.1007/978-3-030-24198-8 |