Journal of Nonlinear Mathematical Physics

Volume 25, Issue 2, March 2018, Pages 309 - 323

Bilinear Identities and Hirota’s Bilinear Forms for the (γn, σk)-KP Hierarchy

Authors
Yuqin Yao
Departmentof Applied Mathematics, China Agricultural University, Beijing, 100083, People’s Republic of China,yyqinw@126.com
Juhui Zhang
Departmentof Applied Mathematics, China Agricultural University, Beijing, 100083, People’s Republic of China
Runliang Lin
Department of Mathematical Sciences, Tsinghua University, Beijing, 100084, People’s Republic of China
Xiaojun Liu
Departmentof Applied Mathematics, China Agricultural University, Beijing, 100083, People’s Republic of China
Yehui Huang
School of Mathematics and Physics, North China Electric Power University, Beijing, 102206, People’s Republic of China
Received 31 August 2017, Accepted 8 January 2018, Available Online 6 January 2021.
DOI
10.1080/14029251.2018.1452675How to use a DOI?
Keywords
(γn, σk)-KP hierarchy; bilinear identity; τ-function; Hirota's bilinear form
Abstract

In this paper, we discuss how to construct the bilinear identities for the wave functions of the (γn, σk)-KP hierarchy and its Hirota’s bilinear forms. First, based on the corresponding squared eigenfunction symmetry of the KP hierarchy, we prove that the wave functions of the (γn, σk)-KP hierarchy are equal to the bilinear identities given in Sec.3 by introducing N auxiliary parameters zi, i = 1, 2, …, N. Next, we derived the bilinear equations for the tau-function of the (γn, σk)-KP hierarchy. Then, we obtain the bilinear equations for the taufunction of the mixed type of KP equation with self-consistent sources (KPESCS), which includes both the first and the second type of KPESCS as special cases by setting n = 2 and k = 3. Finally, using the relation between the Hirota bilinear derivatives and the usual partial derivatives, we show the procedure of translating the Hirota’s bilinear equations into the mixed type of KPESCS.

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

Sato theory has important applications in the theory of integrable systems. It reveals the infinite dimensional Grassmannian structure of space of tau-functions, where the tau-function are solutions for the Hirota’s bilinear form of KP hierarchy. The KP hierarchy can be expressed in terms of pseudo-differential operator and has the bilinear identities [3, 4].

Soliton equations with self-consistent sources (SESCS) are important integrable models in many fields of physics, such as hydrodynamics, state physics, plasma physics. For example, the KdV equation with self-consistent sources describes the interaction of long and short capillary-gravity waves. The nonlinear Schrödinger equation with self-consistent sources represents the nonlinear interaction of an electrostatic high frequency wave with the ion acoustic wave in a two component homogeneous plasma. The KP equation with self-consistent sources describes the interaction of a long wave with a short wave packet propagating on the x-y plane at some angle to each other.

As an infinite dimensional integrable system, it has been generalized to large sets of integrable hierarchies by introducing new flows [7, 16]. In [8], Liu and his collaborators construct an extended KP hierarchy by introducing a new vector field τk . This new extended KP hierarchy can be reduced to the k-constrained KP hierarchy, the Gelfand-Dickey hierarchy with self-consistent sources, the first type of KP equation with self-consistent sources(KPESCS) and the second type of KPESCS. In [18], Yao and her collaborators propose a new (γn, σk)-KP hierarchy with two new time series γn and σk. This new (γn, σk)-KP hierarchy can be regarded as a generalization of the extended KP hierarchy, which consists of a γn-flow, a σk-flow as well as a mixed γn-and σk-evolution equations of the eigenfunctions [8]. The (γn, σk)-KP hierarchy contains the mixed type of KP equation with self-consistent sources (KPESCS), which can also be reduced to both the first type and the second type of KPESCS as special cases. Also, the constrained flows of the (γn, σk)-KP hierarchy can be regarded as a generalization of the Gelfand-Dickey hierarchy (GDH), which contains the first, the second as well as the mixed type of GDH with self-consistent sources.

The KP hierarchy can be expressed in bilinear form using Hirota’s bilinear operators [6]. In this formalism, solutions to the KP equation can be obtained without knowing its Lax pair. Researchers have paid much attention on the subject of bilinear identities because of its importance in Sato theory. By using the bilinear identities of soliton hierarchies [24, 9, 15], we can derive the Hirota bilinear forms for all the equations in the hierarchies. Recently, Lin and his collaborators give the bilinear identities for the wave functions of the KP hierarchy with a squared eigenfunction symmetry in [11]. Considering the squared eigenfunction symmetry as an auxiliary flow, they also give the bilinear identities for the extended KP hierarchy. They obtain the generating functions of the Hirota bilinear forms for the extended KP hierarchy by constructing the τ-function for the extended KP hierarchy.

This paper is organized as follows. In Section 2, we briefly recall the KP hierarchy and the (γn, σk)-KP hierarchy. In Section 3, the bilinear identities of the (γn, σk)-KP hierarchy are constructed. In Section 4, the τ-function of the (γn, σk)-KP hierarchy is introduced. The generation functions for the Hirota bilinear form of the (γn, σk)-KP hierarchy are obtained. In Section 5, we show the procedure of translating the Hirota bilinear forms into nonlinear partial differential equations. Conclusions are given in the last section.

2. The KP hierarchy and (γn, σk)-KP hierarchy

Let

L=+u11+u22+
be a pseudo-differential operator whose coefficients are considered as generators of a differential algebra 𝒜[4] .

The well-known KP hierarchy

Ltn=[Bn,L],n (2.1)
can be constructed from the compatibility condition of the following linear systems [3, 4]
Lψ=λψ, (2.2a)
ψtn=Bnψ,Bn=(Ln)+,n, (2.2b)
where {tn} are the time variables with t1 = x and Bn stands for the differential part of Ln. The compatibility of tn − flow and tm − flow of the KP hierarchy (2.1) leads to the following zero-curvature equations
(Bm)tn(Bn)tm=[Bn,Bm],m,n. (2.3)

Supposing that W = 1 + ω1− 1 + ω2− 2 + ⋯ is a dressing operator satisfying

tnW=(WnW1)W,n, (2.4)
then the operator L defined by
L=WW1 (2.5)
is a solution to the KP hierarchy (2.1).

Let the wave functions and the adjoint wave functions be

ψ(t,λ)=Weη(t,λ), (2.6a)
ψ*(t,λ)=(W*)1eη(t,λ),η(t,λ)=i1tiλi, (2.6b)
where W* is the formal adjoint of W defined by (iaii)*i()iai , we find that the wave function (2.6a) satisfies the KP hierarchy (2.2) while the adjoint wave function satisfies
L*ψ*=λψ*, (2.7a)
ψ*tn=B n*ψ*,B n*=[(L*)n]+,n. (2.7b)

Similarly, we can get the following hierarchy

L tn*=[L*,B n*],n (2.8)
from the linear systems (2.7a). If the operator W* is a solution of
tn[(W*)1]=[(W*)1W*]1(W*)1,n, (2.9)
then the adjoint operator L*=(W*)1W* is a solution to the hierarchy (2.8).

For any fixed k , by defining a new variable τk whose vector field is given by

τk=tki=1Ns0ς is1ts,

Liu and his collaborators introduce a new extended KP hierarchy [8]

Ltn=[Bn,L],(n,nk), (2.10a)
Lτk=[Bk+i=1Nqi1ri,L], (2.10b)
qi,tn=Bn(qi), (2.10c)
ri,tn=B n*(ri), (2.10d)
qi,τk=Bk(qi), (2.10e)
ri,τk=B k*(ri),i=1,,N. (2.10f)

The compatibility of tn − flow and τk − flow of (2.10) gives rise to the following zero-curvature equations

Bn,τk(Bk+i=1Nqi1ri)tn+[Bn,Bk+i=1Nqi1ri]=0.

For any fixed n,k , Yao and her collaborators propose the (γn, σk)-KP hierarchy with two generalized time series γn and σk in [18]

Lts=[Bs,L],(n,sn,sk), (2.11a)
Lγn=[Bn+αni=1Nqi1ri,L],(nk), (2.11b)
Lσk=[Bk+βki=1Nqi1ri,L], (2.11c)
qi,ts=Bs(qi), (2.11d)
ri,ts=B s*(ri), (2.11e)
αn(qi,σkBk(qi))βk(qi,γnBn(qi))=0, (2.11f)
αn(ri,σk+B k*(ri))βk(ri,γn+B n*(ri))=0,i=1,,N, (2.11g)
where αn and βk are constants, qi and ri(i = 1, 2, ⋯, N) are generalized eigenfunctions and adjoint eigenfunctions. It’s easy to see that the KP hierarchy can be derived from (2.11) by setting αn = 0 and βk = 0. The commutativity of (2.11b) and (2.11c) under (2.11f) and (2.11g) gives rise to the following zero-curvature equations
Bn,σkBk,γn+[Bn,Bk]+βk[Bn,i=1Nqi1ri]++αn[i=1Nqi1ri,Bk]+=0. (2.12)

Supposing that the operator W in (2.5) satisfies the following evolution equations

tsW=(WsW1)W,(sn,sk) (2.13a)
Wγn=(WnW1)W+αni=1Nqi1riW,(nk) (2.13b)
Wσk=(WkW1)W+βki=1Nqi1riW, (2.13c)
we can prove that the operator L defined by (2.5) satisfies (2.11a) (2.11b) and (2.11c) (see[18] for the proof).

When we take n = 2, k = 3 and set γ2 = y, σ3 = t, u1 = u, the mixed KPESCS

4ut31uyy12uuxuxxx3α2i=1N(qiri)y+4β3i=1N(qiri)x+3α2i=1N(qiri,xxqi,xxri)=0, (2.14a)
α2(qi,tqi,xxx3uqi,x32qi1uy32qiux32α2qij=1Nqjrj)β3(qi,yqi,xx2uqi)=0, (2.14b)
α2(ri,tri,xxx3uri,x+32ri1uy32riux+32α2rij=1Nqjrj)β3(ri,y+ri,x+2uri)=0,i=1,2,,N, (2.14c)
can be obtained from (2.12), (2.11f) and (2.11g).

In particular, if α2 = β3 = 0 (resp., α2 = 0, β3 = 1 or α2 = 1, β3 = 0 or α2 = 1, β3 = 1), the nonlinear equations (2.14) will be reduced to the KP equation [3, 4](resp., the first type [12, 13, 17], or the second type [5, 8, 12], or the mixed type of KP equation with self-consistent sources [18]). The KP equation with self-consistent sources has important applications in physics [10, 13].

3. Bilinear Identities for the (γn, σk)-KP hierarchy

We introduce zi -flows (i = 1, 2, ⋯, N) as

ziL=[qi1ri,L],i=1,2,,N, (3.1)
where qi and ri are the eigenfunctions and their adjoint ones, respectively to construct the bilinear identities for (2.11). According to the results given in [1], the relation between the operator W and the auxiliary parameters zi(i = 1, 2, ⋯, N) satisfies
Wzi=qi1riW,i=1,2,,N. (3.2)

Let ξ(t,λ)=in,ktiλi+γnλn+σkλk, the action of pseudo-differential operator on ξ(t, λ) is defined by

mξ(t,λ)=λm,meξ(t,λ)=λmeξ(t,λ)
for any integer m.

Denoting z=(z1,z2,,zN) , t = (t1,…, tn − 1, γn, tn + 1,…, tk − 1, σk, tk + 1, ⋯) or t = (t1,…, tn − 1, γn, tn + 1,…, tk − 1, σk, tk + 1, ⋯), the wave function and the adjoint wave function with auxiliary parameters zi(i = 1, 2, ⋯, N) can be defined as

ω(z,t,λ)=Weξ(t,λ), (3.3a)
ω*(z,t,λ)=(W*)1eξ(t,λ). (3.3b)

Before giving the bilinear identities for (2.11), let’s recall a useful lemma [3]:

Lemma 1.

Let P and Q be two pseudo-differential operators, Q* is the formal adjoint of Q, then

ResPQ*=ResλP(eξ(t,λ))Q(eξ(t,λ)), (3.4)
where Res(iaii)=a1 and Resλ(iaiλi)=a1 .

Now we have the following theorems:

Theorem 1.

The (γn, σk)-KP hierarchy (2.11) is equivalent to the following bilinear identities with N auxiliary variables zi, i = 1, 2, 3, …, N,

Resλω(T,t,λ)ω*(T,t,λ)=0, (3.5a)
Resλωzi(T,t,λ)ω*(T,t,λ)=qi(T,t)ri(T,t), (3.5b)
Resλω(T,t,λ)[1qi(T,t)ω*(T,t,λ)]=qi(T,t), (3.5c)
Resλ[1ri(T,t)ω(T,t,λ)]ω*(T,t,λ)=ri(T,t),i=1,2,,N, (3.5d)
where
t=(t1,,tn1,γn,tn+1,,tk1,σk,tk+1,),t=(t1,,tn1,γn,tn+1,,tk1,σk,tk+1,),T=(z1αnγnβkσk,z2αnγnβkσk,,zNαnγnβkσk),T=(z1αnγnβkσk,z2αnγnβkσk,,zNαnγnβkσk),
and
f(T,t)=g1g2f(T,t),
g1=(t1t1)i1(tn1tn1)in1(tn1tn+1)in+1(tk1tk1)ik1(tk+1tk+1)ik+1,g2= 1i1 n1in1 n+1in+1 k1ik1 k+1ik+1i1!in1!in+1!ik1!ik+1!(1)i0i(βkγnβkγn)i(αnσkαnσk)i0i ni ki0ii!(i0i)!.

The action of ∂ 1 on the (adjoint) wave function is taken as pseudo-differential operator acting on the exponential part of the function, e.g., 1(rω)=(1rW)(eξ(t,λ)) .

Proof. Let’s prove the following observations first

(βkddγnαnddσk)m0 t1m1 tn1mn1 tn+1mn+1 tk1mk1 tk+1mk+1 tlml[ω*(T,t,λ)],=Pm0m1mn1mn+1mk1mk+1mlω*(T,t,λ) (3.6a)
(βkddγnαnddσk)m0 t1m1 tn1mn1 tn+1mn+1 tk1mk1 tk+1mk+1 tlml[ri(T,t)],=Pm0m1mn1mn+1mk1mk+1mlri(T,t) (3.6b)
where Pm0m1mn1mn+1mk1mk+1m1 is a differential operator in ∂ since the actions of the partial derivatives ti (for in,k) and ddγn,ddσk can all be written as the actions of differential operators.

Indeed, applying ts,ddγn,ddσk to ω*(T,t,λ), the following expression can be constructed

ts[ω*(T,t,λ)]=B s*[ω*(T,t,λ)], (3.7a)
ddγn[ω*(T,t,λ)]=ω γn*(z,t,λ)|z=Tαnj=1Nω zj*(T,t,λ), (3.7b)
ddσk[ω*(T,t,λ)]=ω σk*(z,t,λ)|z=Tβkj=1Nω zj*(T,t,λ), (3.7c)
which can be reduced to
(βkddγnαnddσk)[ω*(T,t,λ)]=[αnB k*βkB n*][ω*(T,t,λ)]. (3.8)

Similarly, applying ts,ddγn,ddσk to ri(T, t) and taking (2.11e) (2.11g) into consideration, we have

ts[ri(T,t)]=B s*[ri(T,t)],i=1,,N, (3.9a)
αn[ri,σk(z,t)|z=T+B k*ri(T,t)]βk[ri,γn(z,t)|z=T+B n*r(T,t)i]=0, (3.9b)
ddγn[ri(T,t)]=ri,γn(z,t)|z=Tαnj=1Nri,zj(T,t), (3.9c)
ddσk[ri(T,t)]=ri,σk(z,t)|z=Tβkj=1Nri,zj(T,t). (3.9d)

Substituting (3.9c) and (3.9d) into (3.9b), we obtain

(βkddγnαnddσk)ri(T,t)=(αnB k*βkB n*)ri(T,t). (3.10)

So the observations (3.6) can be obtained with the help of (3.7a), (3.8) and (3.9a), (3.10).

Now we prove the bilinear identities (3.5) from (2.11) (2.13) (3.2):

To prove the bilinear identity (3.5a), it is sufficient to consider the following case

Resλω(T,t,λ)(βkddγnαnddσk)m0 t1m1 tn1mn1 tn+1mn+1 tk1mk1 tk+1mk+1 tlmlω*(T,t,λ)=0
for every mj ≥ 0.

By using Lemma 1 and observation (3.6a), we have

Resλω(T,t,λ)(βkddγnαnddσk)m0 t1m1 tn1mn1 tn+1mn+1 tk1mk1 tk+1mk+1 tlmlω*(T,t,λ)=ResλWeξ(t,λ)Pm0m1mn1mn+1mk1mk+1ml(W*)1eξ(t,λ)=ResW(W)1P m0m1mn1mn+1mk1mk+1ml*=0,
so the bilinear identity (3.5a) holds.

Notice that Wzi = qi− 1riW, i = 1, 2, …, N, we get

Resλωzi(T,t,λ)(βkddγnαnddσk)m0 t1m1 tn1mn1 tn+1mn+1 tk1mk1 tk+1mk+1 tlmlω*(T,t,λ)=Resλωzi(T,t,λ)Pm0m1mn1mn+1mk1mk+1mlω*(T,t,λ)=Resλqi(T,t)1ri(T,t)Weξ(t,λ)Pm0m1mn1mn+1mk1mk+1ml(W*)1eξ(t,λ)=Resqi(T,t)1ri(T,t)P m1m2ml*=qi(T,t)Pm0m1mn1mn+1mk1mk+1mlri(T,t),
so the bilinear identity (3.5b) is proved.

Similarly, we have the following bilinear identity

Resλω(T,t,λ)ω zi*(T,t,λ)=qi(T,t)ri(T,t). (3.11)

By substituting ω zi*=ri1qiω* into (3.11) and ωzi=qi1riω into (3.5b) respectively, the bilinear identities (3.5c) and (3.5d) can be proved.

Theorem 2.

If qi(T,t), ri(T,t) (i = 1, 2, ⋯, N),

ω(T,t,λ)=(1+i1ωiλi)eξ(t,λ),
and
ω*(T,t,λ)=(1+i1ω i*λi)eξ(t,λ)
satisfy the bilinear identities (3.5), then the pseudo-differential operators L=WW1(W=1+iwii) and functions qi and ri are solutions to the (γn, σk)-KP hierarchy (2.11).

Proof. For any m ≥ 1, denoting W˜=1+i1ω i*i and taking (3.5a) and Lemma 1 into account, we have

ResWW˜*m=ResλWeξ(t,λ)()mW˜eξ(t,λ)=Resλω(T,t,λ)()mω*(T,t,λ)=0,
which implies that the negative part of (WW˜*) is 0. Noticing that the non-negative part of (WW˜*)+ is 1, we have W˜=(W*)1 .

For any m > 0, the following computation

ResWziW1()m=ResλWzieξ(t,λ)m(W1)*eξ(t,λ)=Resλωzi(T,t,λ)mω*(T,t,λ)=qi(T,t)mri(T,t),
leads to
(WziW1)=qi(T,γ)1ri(T,γ).

Since the non-negative part of (WziW1)+ is 0, then (3.2) holds.

From the definition of W, we know that (Wti+L iW)+=0.

For m > 0 and sn, sk, we have the following computation

Res(WtsW1+L s)m=Res(WtsW1+(WsW1))m=ResλWtseξ(t,λ)()m(W*)1eξ(t,λ)+Res(WsW1(WsW1)+)m=ResλWtseξ(t,λ)()m(W*)1eξ(t,λ)+ResλWseξ(t,λ)()mW*1eξ(t,λ)=Resλωts(T,t,λ)()mω*(T,t,λ)=0,
which means that (2.13a) holds.

Similarly, we can get

ddγn[W(T,t)]=L nW, (3.12a)
ddσk[W(T,t)]=L kW, (3.12b)
i.e.,
Wγn(z,t,)|z=Tαni=1NqiriW=L nW,
Wσk(z,t)|z=Tβki=1NqiriW=L kW.

So (2.13b) and (2.13c) are proved.

Applying both sides of (3.12) to eξ(t, λ) respectively, we can get the following equalities

ddγnω(T,t,λ)=Bnω(T,t,λ),ddσkω(T,t,λ)=Bkω(T,t,λ).

Then according to (3.5c), we find

ddγn[qi(T,t)]=Resλddγn[ω(T,t,λ)]1(qi(T,t)ω*(T,t,λ))=L +n(Resλω(T,t,λ)1(qi(T,t)w*(T,t,λ)))=L +n(qi(T,t)),
which leads to
αnj=1Nqi,zj(T,t)+qi,γn(z,t)|z=TL +nqi(T,t)=0. (3.13)

Similarly, we have

βkj=1Nqi,zj(T,t)+qi,σk(z,t)|z=TL +kqi(T,t)=0. (3.14)

So equation (2.11f) can be proved by combining (3.13) and (3.14).

Equation (2.11g) can also be proved in the similar way from (3.5d).

4. Tau-Function for (γn, σk)-KP hierarchy

Since the (adjoint) wave functions of (γn, σk)-KP hierarchy satisfy the same bilinear equation (3.5a) just as the (adjoint)wave functions of the original KP hierarchy do (if one considers zi, i = 1, 2, ⋯, N as auxiliary parameters), it is reasonable to assume the existence of tau-function and make the following assumptions

ω(T,t,λ)=τ(z,t[λ])˜τ(z,t)˜eξ(t,λ), (4.1a)
ω*(T,t,λ)=τ(z,t[λ])˜τ(z,t)˜eξ(t,λ), (4.1b)
where the ~ over a function f(z, t) is defined as
f(z,t)˜f(z1αnγnβkσk,,zNαnγnβkσk,t) (4.2)
with [λ]=(1λ,12λ2,) , z = (z1, z2,…,zN), t = (t1,…, tn−1, γn, tn + 1,…, tk−1, σk, tk + 1, ⋯). For example, according to the definition (4.2), we have
τ(z,t)˜=τ(T,t),τ(z,t[λ]˜)=τ(R,t[λ]),
where
R=(R1,R2,,RN),Ri=ziαn(γn1nλn)βk(σk1kλk),i=1,2,,N.

According to [2, 11], assume

qi(z,t)=κi(z,t)τ(z,t),ri(z,t)=ρi(z,t)τ(z,t),i=1,2,,N. (4.3)
we can get the following results
1(ri(T,t)ω(T,t,λ))=ρi(z,t˜[λ])λτ(T,t)eξ(t,λ),i=1,2,,N, (4.4a)
1(qi(T,t)ω*(T,t,λ))=κi(z,t+[˜λ])λτ(T,t)eξ(t,λ),i=1,2,,N. (4.4b)

Substituting (4.1) (4.3) and (4.4) into (3.5), we have

Resλτ(z,t[λ]˜)τ(z,t+[˜λ])eξ(tt,λ)=0, (4.5a)
Resλτzi(z,t[˜λ])τ(z,t+[˜λ])eξ(tt,λ) (4.5b)
Resλτ(z,t[˜λ])(zilogτ(z,t)˜)τ(z,t+[˜λ])eξ(tt,λ)=κi(z,t)˜ρi(z,t)˜, (4.5c)
Resλλ1τ(z,t[˜λ])κi(z,t+[˜λ])eξ(tt,λ)=κi(z,t)˜τ(z,t)˜, (4.5d)
Resλλ1ρi(z,t[˜λ])τ(z,t+[˜λ])eξ(γt,λ)=ρi(z,t)˜τ(z,t)˜. (4.5e)

Denoting y = (y1, y2, ⋯) and setting t and t′ as t + y and ty respectively, we can write the equalities (4.5) as the following systems with Hirota bilinear derivatives D˜ and Di’s:

i=0pi(2y)pi+1(D˜)exp(i=1yiDi)τ(T,t)τ(T,t)=0, (4.6a)
i=0pi(2y)[zilogτ(T,t+y)]pi+1(D˜)τ(T,t+y)τ(T,ty)+i=0pi(2y)pi+1(D˜)exp(i=1yiDi)τzi(T,t)τ(T,t)=exp(i=1yiDi)κi(T,t)ρi(T,t), (4.6b)
i=0pi(2y)pi(D˜)exp(i=1yiDi)τ(T,t)κi(T,t)=exp(i=1yiDi)κi(T,t)τ(T,t), (4.6c)
i=0pi(2y)pi(D˜)exp(i=1yiDi)ρi(T,t)τ(T,t)=exp(i=1yiDi)τ(T,t)ρi(T,t). (4.6d)
where D˜=(D1,12D2,13D3,) , Di is the Hirota bilinear derivative defined by Difg=ftigfgti and pi(y) is the i-th Schur polynomial given by expi=1yiλi=i=0pi(y)λi .

Let y = 0, the equation (4.6b) can be converted into the following forms by using the Hirota bilinear operator

κi(T,t)ρi(T,t)+Dxτzi(T,t)τ(T,t)=κi(T,t)ρi(T,t)+Dziτx(T,t)τ(T,t)=κi(T,t)ρi(T,t)+12DxDziτ(T,t)τ(T,t)=0. (4.7)

By setting n = 2, k = 3 and comparing the coefficient of y3 in equation (4.6a) and the coefficient of y2, y3 in equation (4.6c) (4.6d), we can obtain

κi(z,t)ρi(z,t)+12D1Dziτ(z,t)τ(z,t)=0, (4.8a)
[D 14+3(D2α2i=1NDzi)24D1(D3β3i=1NDzi)]τ(z,t)τ(z,t)=0, (4.8b)
[4(D3β3i=1NDzi)+3D1(D2α2i=1NDzi)D 13]τ(z,t)κi(z,t)=0, (4.8c)
[4(D3β3i=1NDzi)+3D1(D2α2i=1NDzi)D 13]ρi(z,t)τ(z,t)=0, (4.8d)
[D 12+(D2α2i=1NDzi)]τ(z,t)κi(z,t)=0, (4.8e)
[D 12+(D2α2i=1NDzi)]ρi(z,t)τ(z,t)=0,i=1,2,,N. (4.8f)

The bilinear equations (4.8) correspond to the mixed type of KP equation with self-consistent sources [5], which can be reduced to the first type or the second type of KP equation with self-consistent sources by setting α2 = 0 or β3 = 0 respectively. The KP equation with self-consistent sources describes the interaction of a long wave with a short-wave packet propagating on the x, y plane at an angle to each other [13].

5. The Procedure of Getting Nonlinear Equation from Hirota’s Bilinear Equation

At the beginning of this section, we recall two identities for arbitrary functions τ(t) and κ(t), which is proved in [11]

exp(iδiDi)κτ=e2cosh(iδii)logτeδii(κ/τ), (5.1a)
cosh(iδiDi)ττ=e2cosh(iδii)logτ. (5.1b)

By defining u= x2logτ(xt1),q=κτ,r=ρτ , we can get [11]

1τ2n=0(iδiDi)nn!κτ=exp[2n=1(iδii)2n(2n)!2u]eδiiq, (5.2a)
1τ2n=1(iδiDi)2n(2n)!ττ=exp[2n=1(iδii)2n(2n)!2u], (5.2b)
and similarly, we have
1τ2n=0(iδiDi)nn!ρτ=exp[2n=0(iδii)2n2n!2u]eδiir. (5.2c)

Comparing the coefficient of (δi)j,j0 , we can get the relations between the Hirota bilinear derivatives and the usual partial derivatives. Here are some of them

{D 14τττ2=2u1,1+12u2,D 12τττ2=2u,D1D3τττ2=21u3D 13ρττ2=r1,1,1+6ur1,D1D2ρττ2=r1,2+2r1u2D2ρττ2=r2,D3ρττ2=r3,D 12ρττ2=r1,1+2urD 13κττ2=q1,1,1+6uq1,D1D2κττ2=q1,2+2q1u2D2κττ2=q2,D3κττ2=q3,D 12κττ2=q1,1+2uq (5.3)
where the subscripts i, j, ⋯ of ui,j,…, ri,j,…, qi,j,… denote the derivatives with respect to the variables ti, tj, ⋯.

We also have the following expressions

{D1Dziτττ2=21uzi,D2Dziτττ2=21u2,zi,D zi2τττ2=21uzi,ziDziρττ2=rzi,D1Dziρττ2=r1,zi+2r1uziDziκττ2=qzi,D1Dziκττ2=q1,zi+2q1uzi (5.4)
where the subscripts zi (i = 1, 2, 3, …N) of u denote the derivatives with respect to the variables zi (i = 1, 2, 3, …, N).

By using (5.3) and (5.4), we can write (4.8) as the following nonlinear partial differential equations

1ziu+riqi=0,(uxxx+12uux+4β3k=1Nuzk4ut)x+3(yα2k=1Nzk)2u=0,4qi,t+4β3qi,zk+3qi,xy+6qi(yα2k=1Nzk)1u3α2k=1Nqi,xzk+qi,xxx+6uqi,x=0,4ri,t4β3ri,zk+3ri,xy+6ri(yα2k=1Nzk)1u3α2k=1Nri,xzkri,xxx6uri,x=0,qi,xx+2uqiqi,y+α2k=1Nqi,zk=0,ri,xx+2uri+ri,yα2k=1Nri,zk=0,i=1,2,3,,N. (5.5)

Equations (2.14) can be obtained by eliminating the auxiliary variables zi in (5.5). So we can see that the Hirota bilinear equations (4.8) correspond to the mixed type of KP equation with self-consistent sources (2.14).

6. Conclusions

The bilinear identities for the (γn, σk)-KP hierarchy [18] are constructed in this paper, which could be seen as the generating functions of all the Hirota’s bilinear equations for the zero-curvature forms in the (γn, σk)-KP Hierarchy. Many integrable 2 + 1 dimensional equations with self-consistent sources are included as special cases of this hierarchy. We have shown that the Hirota’s bilinear forms (4.8) correspond to the mixed type of KPESCS, which can be reduced to the first and the second type of KPESCS.

With the help of N auxiliary flows ( zi – flow), we obtain the bilinear identities of the whole (γn, σk)-KP hierarchy, which have many important applications. For example, taking the intimate relation between quasi-periodic solutions and bilinear identity into account, we investigate the quasi-periodic solutions for the (γn, σk)-KP Hierarchy. Under proper constraints, the (γn, σk)-KP hierarchy can be reduced to Gelfand-Dickey hierarchy (GDH), KdV equation, Bonssinesq equation and many other equations with self-consistent sources. So the bilinear identities of the (γn, σk)-KP hierarchy can help us learn the relation among these equations’ bilinear identities. We will investigate these problems in future.

Acknowledgments

This work is supported by National Natural Science Foundation of China (Grant No. 11471182, 11201477, 11301179), Beijing Natural Science Foundation (1182009) and China Scholarship Council.

References

[2]Y Cheng and YJ Zhang, Solutions for the vector k-constrained KP hierarchy, Inverse Probl., Vol. 11, 1994, pp. 5869-5884.
[3]E Date, M Kashiwara, M Jimbo, and T Miwa, Transformation groups for soliton equations, M Jimbo and T Miwa (editors), Nonlinear Integrable Systems, Classical Theory and Quantum Theory (Kyoto, 1981), World Scientific, Singapore, 1983, pp. 39-119.
[4]LA Dickey, Soliton equations and Hamiltonian systems, 2nd ed., World Scientific, Singapore, 2003.
[6]R Hirota, Direct methods in soliton theory, Cambridge University Press, Cambridge, 2004.
Journal
Journal of Nonlinear Mathematical Physics
Volume-Issue
25 - 2
Pages
309 - 323
Publication Date
2021/01/06
ISSN (Online)
1776-0852
ISSN (Print)
1402-9251
DOI
10.1080/14029251.2018.1452675How 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  - Yuqin Yao
AU  - Juhui Zhang
AU  - Runliang Lin
AU  - Xiaojun Liu
AU  - Yehui Huang
PY  - 2021
DA  - 2021/01/06
TI  - Bilinear Identities and Hirota’s Bilinear Forms for the (γn, σk)-KP Hierarchy
JO  - Journal of Nonlinear Mathematical Physics
SP  - 309
EP  - 323
VL  - 25
IS  - 2
SN  - 1776-0852
UR  - https://doi.org/10.1080/14029251.2018.1452675
DO  - 10.1080/14029251.2018.1452675
ID  - Yao2021
ER  -