Skip to main content
Proceedings. Mathematical, Physical, and Engineering Sciences logoLink to Proceedings. Mathematical, Physical, and Engineering Sciences
. 2022 Jul 27;478(2263):20220074. doi: 10.1098/rspa.2022.0074

Integrable nonlinear evolution equations in three spatial dimensions

A S Fokas 1,2,3,
PMCID: PMC9326974  PMID: 35909419

Abstract

There are integrable nonlinear evolution equations in two spatial variables. The solution of the initial value problem of these equations necessitated the introduction of novel mathematical formalisms. Indeed, the classical Riemann–Hilbert problem used for the solution of integrable equations in one spatial variable was replaced by a non-local Riemann–Hilbert problem or, more importantly, by the so-called d-bar formalism. The construction of integrable nonlinear evolution equations in three spatial dimensions has remained the key open problem in the area of integrability. For example, the two versions of the Kadomtsev–Petviashvili (KP) equation constitute two-dimensional generalizations of the celebrated Korteweg–de Vries equation. Are there three-dimensional generalizations of the KP equations? Here, we present such equations. Furthermore, we introduce a novel non-local d-bar formalism for solving the associated initial value problem.

Keywords: integrable, nonlinear, evolution, multi-dimensional, d-bar

1. Introduction

The celebrated Korteweg–de Vries (KdV) and nonlinear Schrödinger (NLS) equations are prototypical integrable nonlinear evolution equations in one spatial variable. Every integrable nonlinear evolution equation in one spatial dimension has several integrable versions in two spatial dimensions. Two such physically significant generalizations of the KdV are the Kadomtsev-Petviashvili I (KPI) and II (KPII) equations. Similarly, two analogous generalizations of the NLS are the Davey–Stewartson I (DSI) and II (DSII) equations [1,2]. It is important to emphasize that integrable nonlinear evolution equations arise in a variety of physical applications. For example, in the context of fluid mechanics, the KP equations arise in the weakly nonlinear, weakly dispersive, weakly two-dimensional limit of inviscid, irrotational water waves, and in the case of KPI when the surface tension is dominant [3].

The modern history of integrable equations begins with the famous works of Martin Kruskal and his colleagues on the Cauchy problem of the KdV, using what was later called the inverse scattering transform method [4,5]. The next important step was taken by Peter Lax, who established that the crucial property of an integrable equation is its formulation as the compatibility of two linear eigenvalue equations, which were later called a Lax pair [6]. The existence of a Lax pair continues to provide the defining property of integrability [7].

The inverse scattering transform method for the solution of integrable evolution equations in 1 + 1, i.e. in one spatial and one temporal dimensions, can be considered as a nonlinear Fourier transform method [8]. As shown in [9], the classical Fourier transform pair and the nonlinear generalization of this pair needed for the implementation of the inverse scattering transform method can be constructed by performing the so-called spectral analysis of the time-independent part of the Lax pair via the formulation of a Riemann–Hilbert problem. This is a specific mathematical problem in the theory of sectionally analytic functions, namely of functions that are analytic in different sectors covering the complex plane and have given jumps across the curves defining these sectors [10]. The main difference between the linear and nonlinear Fourier transform pairs is the following: the construction of the classical Fourier transform pair is based on a scalar Riemann–Hilbert problem, which can be solved in closed form, whereas in the nonlinear case, one formulates a matrix Riemann–Hilbert problem, which is equivalent to a linear integral equation of the Fredholm type. Incidentally, in certain simple cases, this integral equation can be transformed to the so-called Gelfand–Levitan–Marchenko integral equation, which was used in the original integration of the KdV equation [5].

A formal method for the solution of the Cauchy problem of equations in 2 + 1, i.e. in two spatial and one temporal dimensions, was developed by Mark Ablowitz, the author of the present paper, and their collaborators. This approach was later made rigorous in several publications; see, for example, [1117] as well as [18,19]. The extension from 1 + 1 to 2 + 1 involved the following unexpected generalization: for some equations in 2 + 1, like KPI and DSI, instead of formulating a classical Riemann–Hilbert problem, it is now required to formulate a non-local Riemann–Hilbert problem; for KPI and DSI this was achieved in [20] and [21], respectively. However, in general, it is necessary to go beyond the realm of analytic functions and to consider functions that are nowhere analytic in the complex plane. This leads to the formulation of the so-called d-bar problem [10]. Incidentally, this problem was first introduced in the field of integrability in the elegant analysis of Beals and Coifman of certain problems in 1 + 1 [22,23], although for such problems a Riemann–Hilbert formalism is not only still adequate but actually preferable to the d-bar formalism. The implementation of the d-bar method to equations in 2 + 1, where its use is indispensable, was presented in [24] for KPII and in [25] for DSII (see also [26]).

A fundamental open question in the field of integrability has been the existence of integrable evolution equations in more than two spatial dimensions. Substantial progress in this question was presented in [27], where 4 + 2 generalizations of the KP and the DS equations were presented, namely equations in four spatial and two temporal dimensions. The solution of the Cauchy problem of the 4 + 2 generalization of the KP was presented in [28]. The solution of the Cauchy problem of the 4 + 2 generalization of the DS equation was sketched in [27] and presented in detail in [29]. The solution of the Cauchy problem of the 4 + 2 generalization of the three-wave interaction equations is presented in [30]. These problems were solved using a non-local d-bar formalism, where the non-locality takes the form of a two-dimensional integral: the d-bar problem involves a two-dimensional integral over lR and lI, where l is a complex variable. Hence, the nonlinear Fourier transform (spectral function) depends on four real parameters, kR,kI,lR,lI, where kR and kI are the real and imaginary parts, respectively, of the spectral parameter k. This is consistent with the fact that q depends on four spatial variables.

Incidentally, another approach towards integrability in multi-dimensions was presented in [31], where it was shown that integrable nonlinear evolution equations exist in any number of dimensions; however, these equations have the major disadvantage that they involve a non-local commutator.

The questions of reducing integrable equations from 4 + 2 to equations in three dimensions and of establishing that the initial value problems of the resulting 3 + 1 equations are well-posed, and although discussed in several papers, including [32] and [33], they remain open. Here, we finally solve this problem. In particular, we solve the initial value problem of the following equation for the complex-valued function q(x,y,z,t), where x,y,z,t, are real independent variables:

qt+αqxxx+γx1(qyyqzz+2iqyz)+βqqx=0,qC,x,y,z,tR, 1.1

where α, β, γ are arbitrary real or complex constants. In the particular case when z=0 and α, β, γ are real, equation (1.1) reduces to the KP equation

qt+αqxxx+γx1qyy+βqqx=0,qR,x,y,tR. 1.2

Similarly, the following system of the complex-valued functions Qj(x,y,z,t), j=1,2, where x, y, z, t are real independent variables, provides a three-spatial-dimensions integrable generalization of the general DS-type system

(1)j1it¯Qj+(αx2+βη¯2)Qj+γQj(δx1η¯+δ1η¯1x)Q1Q2=0,j=1,2,Q1,Q2C,η=y+iz,x,y,z,tR,η¯=12(y+iz)and(η¯1f)(y,z)=1πR2f(y,z)ηηdydz,} 1.3

where α β, γ, δ are arbitrary real or complex constants. In the particular case when z=0 and α, β, γ, δ are real, equations (1.3) reduce to the DS-type system

(1)j1itQj+(αx2+βy2)Qj+γQj(δx1y+δ1y1x)Q1Q2=0,j=1,2andQ1,Q2C,x,y,tR,} 1.4

where we have renamed β/4 as β and δ/2 as δ. In the particular case of Q1=q, Q2=σq¯, σ2=1, equations (1.4) reduce to the DS equation

iqt+(αx2+βy2)q+σγq(δx1y+δ1y1x)|q|2=0,qC,x,y,tR. 1.5

It is noted that equation (1.1) has the disadvantage that there does not appear to be a real reduction of it in 3 + 1 involving a single equation, i.e. it seems necessary to consider the system of two equations for the real and imaginary parts of q. Similarly, whereas equations (1.4) admit a reduction to the single DS equation (1.5), there does not appear to be a reduction to a single equation of the system in equations (1.3).

A 3 + 2 generalization of the KP equation is given by

qt+iqρ+αqxxx+γx1(qyyqzz+2iqyz)+βqqx=0, 1.6

where α, β, γ are arbitrary real or complex constants. This equation involves a second time-like variable, ρ. This variable could be interpreted, for example, as a specific quantity measuring the dependence of the underlying process on a particular property of the system.

What is the motivation for introducing a second time-like variable? It is shown in proposition 3.2 that the time dependence of the nonlinear Fourier transform (spectral function) associated with equation (1.6) involves the exponential exp(iTt+iRρ), where T and R are specific real functions of the spectral variables kR,kI,ξ. In the linear limit, the spectral function becomes the Fourier transform of the solution q, and kR,kI,ξ are the three variables in the Fourier space corresponding to the three variables, x,y,z in the physical space. The above time dependence shows that equation (1.6) is a dispersive PDE and hence its Cauchy problem is well-posed. On the other hand, the time dependence of the spectral function associated with equation (1.1) involves the exponential exp(iTtRt), which immediately raises the question of whether this equation is well-posed. The main achievement of this work is to show that this is indeed the case: it is shown in proposition 3.5 that there is a novel d-bar formalism that ensures R is always non-negative. Hence, the underling initial value problem is well-posed. The well-posedness of equation (1.1) is consistent with the results presented in the companion paper [34], where a multitude of explicit non-singular solutions are presented, including multi-solitons, multi-breathers, and multi-rational solutions.

The implementation of the d-bar method to the system in equations (1.3), as well as to the reduced version (to three spatial dimensions) of the 4 + 2 generalization of the three-wave interaction equations, is work in progress.

2. Three-dimensional extensions of the Kadomtsev–Petviashvili and the Davey–Stewartson-type systems

It was shown in [27] that the complexification of the independent variables x, y, t of the KP equation yield the following 4 + 2 integrable extension of the KP equation:

qt¯=14qx¯x¯x¯32qqx¯+34x¯1qy¯y¯, 2.1

where

t=t1+it2,x=x1+ix2,y=y1+iy2,t1,t2,x1,x2,y1,y2R. 2.2

Equations (2.2) imply

t¯=12(t1+it2),x¯=12(x1+ix2),y¯=12(y1+iy2)and(x¯1f)(x1,x2)=1πR2f(x1,x2)xxdx1dx2.} 2.3

It is important to emphasize that equation (2.1) is a genuine equation in 4 + 2, namely the dependent variable q(x1,x2,y1,y2,t1,t2) depends on the four spatial variables x1, x2, y1, y2 and on the two temporal variables t1, t2.

Letting

ξ=ax1+bx2,τ=a~t1+b~t2,a,b,a~,b~are real constants, 2.4

we find

x¯=Aξandt¯=A~τ,A=a+ib2,A~=a~+ib~2. 2.5

Hence, equation (2.1) becomes

A~qτA34qξξξ+32Aqqξ341Aξ1qy¯y¯=0.

Replacing q by Qq, y by y/Y¯, where Q and Y are complex constants, we find

qτ¯+αqξξξ+βqqξ+γξ1qy¯y¯=0,α,β,γC, 2.6

where

α=14A3A~,β=32AQA~andγ=34Y2AA~. 2.7a

Equation (2.6) with a~t1=2t,t2=0 is equation (1.1), whereas equation (2.6) with a~t1=2t,b~t1=2ρ is equation (1.6).

Equations (2.7a) imply

A~=A34α,Q=A2β6αandY2=A4γ3α. 2.7b

The t-independent part of the Lax pair of equation (2.1), after the change of variables in equations (2.4), becomes

[y¯+k2(Aξ+k)2+q]μ=0.

Thus, the t-independent part of the Lax pair of equation (2.6) is

[Yy¯+k2A2(ξ+kA)2+Qq]μ=0.

Replacing k with Ak and simplifying using equations (2.7b), we find

[(γ3α)(1/2)y¯ξ22kξβ6αq]μ=0. 2.8

Similarly, the t-part of the Lax pair of equation (2.1), after the change of variables in equations (2.4), yields

{τ¯4αk3+4α(ξ+k)3+βq(ξ+k)+β2[qξ+(γ3α)(1/2)ξ1qy¯]}μ=0. 2.9

Renaming τ, ξ, y1, y2 as t, x, y, z, and renaming γ/4 as γ, we find the following result: the integrable 3 + 1 equation (1.1) possesses the Lax pair

[(γ3α)(1/2)(y+iz)x22kxβ6αq]μ=0 2.10

and

{t+4α(x3+3kx2+3k2x)+βq(x+k)+β2[qx+(γ3α)(1/2)x1(qy+iqz)]}μ=0. 2.11

Similarly, equation (1.6) possesses the Lax pair formed by equation (2.10) and by the equation formed from equation (2.11) by replacing t with t+iρ.

We next consider the 2 + 1 version of the DS type system

(1)j1tQj+14(x2+η2)Qj12Qj(x1η+η1x)Q1Q2=0,j=1,2andQ1,Q2C,x,η,tR.} 2.12

The complexification of the independent variables

t=t1+it2,x=x1+ix2andη=η1+iη2,t1,t2,x1,x2,η1,η2R 2.13

yields the 4 + 2 integrable system

(1)j1t¯Qj+14(x¯2+η¯2)Qj12Qj(x¯1η¯+η¯1x¯)Q1Q2=0,j=1,2, 2.14

where

η¯=12(η1+iη2),(η¯1f)(η1,η2)=1πR2f(η1,η2)ηηdη1dη2, 2.15

and similarly for x¯, x¯1. It should be noted that there are typos for this system in [27], which are corrected in [29].

Using the change of variables in equations (2.4), proceeding as with the KP equation and finally renaming τ, ξ, η1, η2 as t, x, y, z, we find equations (1.3), where

α=iA24A~,β=iY24A~andδ=A1YA~,

and the parameter γ is generated by the scaling of Q1 and Q2.

3. The Cauchy problem for a three-dimensional extension of the Kadomtsev–Petviashvili

For simplicity of notation, we suppress the dependence on k¯, namely instead of F(k,k¯), we write F(k). Thus, F(k) means that F depends on (kR,kI) instead of kR+ikI, where kR and kI denote the real and the imaginary parts, respectively, of k.

(a) . Spectral analysis of the t-independent part of the Lax pair

In what follows, instead of analysing equation (2.10), we will analyse the equation

[y+izx22kx+q(x,y,z)]μ(x,y,z,k)=0,x,y,zR,kC. 3.1

If γ/α is real, then equation (3.1) can be mapped to equation (2.10) by replacing y and z with (γ/3α)1/2y and (γ/3α)1/2z, respectively, and by rescaling q. If γ/α is complex, then equation (3.1) can also be mapped to equation (2.10) with appropriate scaling of y and z.

We assume that q vanishes for large values of x,y,z. In what follows, we will show that a solution of equation (3.1), which tends to 1 for large values of x,y,z, satisfies the following linear integral equation:

μ(x,y,z,k)=1+R3G(xx,yy,zz,k)q(x,y,z)μ(x,y,z,k)dxdydz, 3.2

where

G(x,y,z,k)=1(2π)3R3ei(ξx+ηy+ζz)ξ22ikξ+iηζdξdηdζ. 3.3

Furthermore, μ/k¯ satisfies the linear integral equation

μk¯(x,y,z,k)=F(x,y,z,k)+R3G(xx,yy,zz,k)q(x,y,z)×μk¯(x,y,z,k)dxdydz, 3.4

where F is defined by

F(x,y,z,k)=R3Gk¯(xx,yy,zz,k)q(x,y,z)μ(x,y,z,k)dxdydz 3.5

and

Gk¯(x,y,z,k)=12iπ3Rξei[ξx+2kRξy+(ξ2+2iξkI)z]dξ. 3.6

Indeed, we define the Green’s function G by

Gy+iGzGxx2kGx=δ(x)δ(y)δ(z). 3.7

Using the identity

δ(x)δ(y)δ(z)=1(2π)3R3ei(ξx+ηy+ζz)dξdηdζ 3.8

and looking for a solution of equation (3.7) in the form

G=1(2π)3R3g(ξ,η,ζ,k)ei(ξx+ηy+ζz)dξdηdζ,

we find equation (3.3).

In order to compute G/k¯, we note that

ξ22ikξ+iηζ=2i{k[η2ξ+i2ξ(ζξ2)]}.

This equation together with the identity

k¯(1kk~)=πδ(kRk~R)δ(kIk~I),k=kR+ikI,k~=k~R+ik~I, 3.9

imply the formula

Gk¯=1(2π)3π(2i)Rei(ξx+ηy+ζz)|2ξ|2ξ|η=2ξkRζ=ξ2+2kIξdξ,

which is equation (3.6).

Putting equation (3.6) into equation (3.5), we find

F(x,y,z,k)=Rξq^(ξ,k)ei[ξx+2kRξy+(ξ2+2ξkI)z]dξ, 3.10

where q^(ξ,k) is defined by

q^(ξ,k)=116iπ2R3ei[ξx+2kRξy+(ξ2+2ξkI)z]q(x,y,z)μ(x,y,z,k)dxdydz. 3.11

It turns out that the functions μ and μ/k¯ satisfy the following relationship:

μ(x,y,z,k)k¯=Rξei[ξx+2kRξy+(ξ2+2ξkI)z]q^(ξ,k)μ(x,y,z,k+iξ)dξ. 3.12

Indeed, the product of the exponential and the function μ(x,y,z,k+iξ) appearing in the above integral satisfy the following equation:

ei[ξx+2kRξy+(ξ2+2ξkI)z]μ(x,y,z,k+iξ)=ei[ξx+2kRξy+(ξ2+2ξkI)z]+R3G^(xx,yy,zz,k,ξ)q(x,y,z)ei[ξx+2kRξy+(ξ2+2ξkI)z]×μ(x,y,z,k+iξ)dxdydz, 3.13

where

G^(x,y,z,k,ξ)=1(2π)3R3ei[(ξ~+ξ)x+(η~+2kRξ)y+(ζ~+ξ2+2ξkI)z]ξ~22ikξ~+iη~ζ~+2ξξ~dξ~dη~dζ~. 3.14

Letting

ξˇ=ξ~+ξ,ηˇ=η~+2kRξandζˇ=ζ~+ξ2+2ξkI,

we find

G^(x,y,z,k,ξ)=1(2π)3R3ei[ξˇx+ηˇy+ζˇz]ξˇ22ikξˇ+iηˇζˇdξˇdηˇdζˇ=G(x,y,z,k). 3.15

In equation (3.13), replacing G^ with G(xx,yy,zz,k), multiplying the resulting equation by ξq^(ξ,k), and integrating over ξ, we find an equation identical to equation (3.5).

Equation (3.12), together with the boundary condition that μ(x,y,z,t,k) tends to 1 for large values of k, define a classical d-bar problem [7]. The solution of this problem is given by

μ(x,y,z,k)=11πR3ξei[ξx+2lRξy+(ξ2+2ξlI)z]lkq^(ξ,l)μ(x,y,z,l+iξ)dξdlRdlIandkC,l=lR+ilI.} 3.16

This equation implies the expansion

μ(x,y,z,k)=1+μ1(x,y,z)k+O(1k2),k, 3.17

where μ1(x,y,z,t) can be explicitly computed in terms of the numerator of the integrand of the right-hand side of equation (3.16). On the other hand, putting the expansion of equation (3.17) into equation (3.1), we find

q(x,y,z)=2(μ1(x,y,z))x. 3.18

Hence,

q(x,y,z)=2πR3ξei[ξx+2lRξy+(ξ2+2ξlI)z]q^(ξ,l)μ(x,y,z,l+iξ)dξdlRdlI. 3.19

In summary, the spectral analysis of equation (3.1) gives rise to the following nonlinear Fourier transform pair: the direct transform is given by equation (3.11), with μ defined in terms of q by equation (3.2), whereas the inverse transform is given by equation (3.19), with μ defined in terms of q^ by equations (3.16). In the linear limit of q, μ tends to 1 and the above transform pair becomes

q^(ξ,k)=116iπ2R3ei[ξx+2kRξy+(ξ2+2ξkI)z]q(x,y,z)dxdydzandq(x,y,z)=2πR3ξei[ξx+2kRξy+(ξ2+2ξkI)z]q^(ξ,k)dξdkRdkI.} 3.20

The change of variables

ξ=k1,2kRξ=k2andξ2+2ξkI=k3

maps the transform pair in equations (3.20) to the usual three-dimensional Fourier transform pair.

The above analysis provides the basis for the following result:

Proposition 3.1. —

Let q(x,y,z)S(R3), where S denotes the space of Schwartz functions and suppose that the L1 and L2 norms of q are sufficiently small.

Let μ(x,y,z,k),(x,y,z)R3,kC, be the unique solution of equation (3.2). A nonlinear Fourier transform of q, denoted by q^(ξ,k),ξR,kC, is defined by the right-hand side of equation (3.11).

Given q^(ξ,k), let μ(x,y,z,k) be the unique solution of equation (3.16). Let q be defined by the right-hand side of equation (3.16). Then q constitutes the inverse of the nonlinear Fourier transform q^.

Proof. —

The result stated in proposition 3.1 is derived earlier. It is straightforward to make this formalism rigorous under the assumption that the L1 and L2 norms of q are sufficiently small. Indeed, in this case, both the linear integral equations (3.2) and (3.16) have unique solutions. The easiest way to obtain these results is to use the fact that for a ‘small q’, the above formalism reduces to the formulae for the direct and inverse Fourier transform pair, which are well-defined if the L1 and L2 norms of q exist.

As it is well-known, the inverse scattering transform method, as well as its d-bar generalization, are based on the spectral analysis of the t-independent part of the Lax pair. Different time evolutions impose different time dependence on the associated nonlinear Fourier transform. In our case, equations (1.1) and (1.6) impose different time dependence on q^(ξ,k). We first consider equation (1.6).

(b) . Solution of the Cauchy problem of equation (1.6)

Proposition 3.2. —

Given the complex-valued function q0(x,y,z)S(R3), whose L1 and L2 norms are sufficiently small, let μ0(x,y,z,k) denote the solution of equation (3.2), where q is replaced by q0. Let q^0(ξ,k) be defined by the right-hand side of equation (3.11), with q and μ replaced by q0 and μ0, respectively. Let μ(x,y,z,t,ρ,k) be the solution of equation (3.16), with q^(ξ,k) replaced by

q^(ξ,k,t,ρ)=q^0(ξ,k)ei(Tt+Rρ),T=ξ3+3ξ(kR2kI2)3ξ2kIandR=3ξkR(ξ+2kI).} 3.21

Let q(x,y,z,t,ρ) be defined by the right-hand side of equation (3.19), with q^(ξ,k) and μ(x,y,z,k) replaced by q^(ξ,k,t,ρ) and μ(x,y,z,t,ρ,k), respectively. Then q(x,y,z,t,ρ) satisfies

qt+iqρ14qxxx+34x1(qzzqyy2iqyz)+32qxq=0, 3.22

with

q(x,y,z,0,0)=q0(x,y,z). 3.23

Proof. —

The equation satisfied by μ(x,y,z,t,ρ,k) is equivalent to the d-bar problem

μ(k)k¯=RE(k,ξ)μ(k+iξ)ξq^0(ξ,k)dξ 3.24a

and

μ(k)=1+μ1k+O(1k2),|k|, 3.24b

where

E(k,ξ)=eiξx+i(k+k¯)ξy+[iξ2+ξ(kk¯)]z+iTt+iRρ, 3.25

and for convenience of notation, we have suppressed the x,y,z,t,ρ dependence of μ and E.

In what follows, using the dressing method, we will show directly that if μ is defined by equations (3.24) and if q is defined in terms of μ via

q(x,y,z,t,ρ)=2(μ1(x,y,z,t,ρ))x, 3.26

or equivalently via equation (3.19), with q^(ξ,k) and μ(x,y,z,k) replaced by q^(ξ,k,t,ρ) and μ(x,y,z,t,ρ,k), respectively, then q solves equation (3.22). For this purpose, we introduce the operators

Dx=x+kandD=y+iz+k2. 3.27

The crucial property of these operators is that they ‘commute’ with the d-bar operator defined in equation (3.24a). In particular,

k¯(Dxμ(k))=RE(k,ξ)Dxμ(k+iξ)dν,dν=ξq^0(ξ,k)dξ. 3.28

Indeed, differentiating equation (3.24a) with respect to x, we find

k¯μx(k)=RE(k,ξ)[μx(k+iξ)+iξμ(k+iξ)]dν.

Also,

k¯(kμ)(k)=RE(k,ξ)kμ(k+iξ)dν.

Adding these two equations, we find equation (3.28). Similar considerations are valid for D. Also,

k¯(fμ)(k)=RE(k,ξ)(qμ)(k+iξ)dν.

Hence, the function L1(k), defined by

L1(k)=(DDx2+q)μ, 3.29

where again we have suppressed the x,y,z,t,ρ dependence of L1(k), satisfies equation (3.24a). Hence, L1 vanishes provided that L1(k)=O(1/k), |k|. This is indeed the case provided that q is defined by equation (3.26).

The equation L1(k)=0 is equation (3.1) with μ and q replaced by μ(x,y,z,t,ρ,k) and q(x,y,t,ρ), respectively. This equation provides the (t,ρ)-independent part of the Lax pair of equation (3.22). In order to define the (t,ρ)-dependent part, we introduce the operator

Dτ=t+iρ+k3. 3.30

It is straightforward to verify that Dτ also commutes with the d-bar operator defined in equation (3.24a). This fact is a direct consequence of the identity

iTR+k3=(k+iξ)3.

Since DτDx3 is of the order k as |k|, we define L2(k) by

L2(k)=(DτDx3+Q1Dx+Q2)μ,

where Q1 and Q2 are functions of x,y,z,t,ρ, chosen by the requirement that L2(k)=O(1/k) as |k|. Hence, the (t,ρ)-dependent part of the Lax pair is

μt+iμρμxxx3k2μx3kμxx+Q1(μx+kμ)+Q2μ=0, 3.31a

where

Q1=3μ1x=32qandQ2=3μ2x+3μ1xxQ1μ1. 3.32a

The terms of O(1/k) of the equation L1(k)=0 imply

μ1y+iμ1zμ1xx2μ2x+qμ1=0. 3.31b

Thus, the second of equations (3.32a) yields

Q2=32(μ1xx+μ1y+iμ1z)=34(qx+x1qy+ix1qz). 3.32b

Either the compatibility of the Lax pair, or the terms of order 1/k of equation (3.31a), yield equation (3.22).

Remark 3.3. —

A simple computation shows that, as expected, the exponential exp[iξx+2ikRξy+i(ξ2+2ξkI)z+iTt+iRρ], with T and R defined in equations (3.21), satisfies the linear version of equation (3.22) (this equation can be solved numerically by letting q=u+iv, and then solving the system of the two resulting equations).

Remark 3.4. —

If the initial condition q0 satisfies the relation

q0(x,y,z)¯=q0(x,y,z), 3.33a

then, the solution of equation (3.22) satisfies the relation

q(x,y,z,t,ρ)¯=q(x,y,z,t,ρ). 3.33b

Indeed, using the transformations ξξ, ηη in equation (3.3), we find

G(x,y,z,k)¯=G(x,y,z,k¯).

Then, equation (3.2) together with the assumption on q0 imply

μ(x,y,z,k)¯=μ(x,y,z,k¯).

Then, the definition of q^(ξ,k) implies

q^(ξ,k)¯=q^(ξ,k¯).

The definition of μ, namely equation (3.24), implies

μ(x,y,z,t,ρ,k)¯=μ(x,y,z,t,ρ,k¯)

and then the definition of q implies that q satisfies equation (3.33b).

(c) . Solving equation (1.1)

Following an analysis similar to that presented in proposition 3.2, it follows that now, instead of equation (3.21), we have

q^(ξ,k,t)=q^0(ξ,k)eiTt4kRξ(ξ+2kI)t, 3.34

where T is defined in equation (3.21). This means that the integral equation defining μ(x,y,z,t,k,ξ) is well-defined only if we can ensure that

kRξ(ξ+2kI)0. 3.35

For this purpose, since we need to control the sign of ξ+2kI, we write the double integral over dξdkI in the form

0dkI2kIdξ+0dkI2kIdξ+0dkI2kIdξ+0dkI2kIdξ.

Furthermore, since we also need to control the sign of ξ, we split the first and fourth integrals as follows:

0dkI0dξ+0dkI02kIdξ+0dkI2kIdξ+0dkI2kIdξ+0dkI2kI0dξ+0dkI0dξ. 3.36

In the first three integrals in equation (3.36), ξ+2kI0, thus in these integrals, we need kRξ0; hence, in the first integral, kR0, in the second, kR0, and in the third, kR0. Similarly, in the last three integrals in equation (3.36), ξ+2kI0, thus in these integrals, we need kRξ0; hence, in the fourth integral, kR0, in the fifth, kR0, and in the sixth, kR0. Hence, if we define the operator LkR,kI,ξ by

LkR,kI,ξ[]=0dkR[0dkI(0dξ+2kIdξ)+0dkI(2kIdξ+0dξ)]+0dkR(0dkI02kIdξ+0dkI2kI0dξ)[], 3.37

then this operator has the distinctive property that the inequality in equation (3.35) is satisfied in the domain of integration.

Using LkR,kI,ξ and following similar steps to those used in proposition 3.2, we can derive the result below.

Proposition 3.5. —

Given the complex-valued function q0(x,y,z), whose L1 and L2 norms are sufficiently small, let μ0(x,y,z,k) denote the solution of equation (3.2), where q is replaced by q0. Let q^0(ξ,k) be defined by the right-hand side of equation (3.11), with q and μ replaced by q0 and μ0, respectively. Let μ(x,y,z,t,k) be the solution of the linear integral equation

μ(x,y,z,t,k)=11πLlR,lI,ξ[Eξq^0(ξ,k)μ(x,y,z,t,l+iξ)lk], 3.38a

where

E:=E(x,y,z,t,lR,lI,ξ)=exp[iTt3lRξ(ξ+2lI)t+iξx+2ilRξy+iξ(ξ+2lI)z]. 3.38b

Let q(x,y,z,t) be defined by

q(x,y,z,t)=2xμ1(x,y,z,t), 3.39

where μ1 is the coefficient of 1/k in the large k expansion of μ. Then, q solves the 3 + 1 equation

qt14qxxx34x1(qyyqzz+2iqyz)+32qqx=0, 3.40a

with

q(x,y,z,0)=q0(x,y,z). 3.40b

Remark 3.6. —

In the linear limit μ1, the above result gives rise to the following novel transform pair:

q^(ξ,k)=116iπ2R3ei[ξx+2kRξy+ξ(ξ+2kI)z]q(x,y,z)dxdydz 3.41a

and

q(x,y,z)=2πLlR,lI,ξ[ξei[ξx+2kRξy+ξ(ξ+2kI)z]q^(ξ,k)]. 3.41b

This pair is the proper pair for solving the linear version of equation (3.40a), namely the equation obtained by neglecting the term qqx.

It is interesting that equations (3.41) imply that q^(ξ,k) automatically satisfies the condition

q^(ξ,k)=0,forkRξ(ξ+2kI)<0. 3.42

The easiest way to prove equations (3.41), as well as to derive equation (3.42), is to use the change of variables k1=ξ,k2=2kRξ,k3=ξ(ξ+2kI). Then, defining q(x,y,z) by the right-hand side of equation (3.41b), it immediately follows that q^(ξ,k) is given by the right-hand side of equation (3.41a). Equation (3.42) is a direct consequence of the Fourier transform pair bijection, where the role of the usual direct and inverse transforms is now reversed.

Remark 3.7. —

It is important to note that a given evolution PDE often imposes an appropriate constraint that needs to be satisfied by the initial datum. A specific such constraint for KPI is analysed in [35]. Such constraints are, of course, ubiquitous for initial-boundary value problems reflecting the compatibility of the solution and its derivatives at the intersection of the spatial and temporal domains. Interestingly, for the case of the evolution PDE (1.1), the proper constraint is given by equation (3.42). What happens if the given initial datum violates this constraint? In this case, it is expected that there will be a singularity at t=0 (for a specific example involving initial-boundary value problems, see [36]). The analysis of this singularity is work in progress.

4. Conclusion

Equation (1.1), where q(x,y,z,t) is a complex-valued function and x,y,z,t are real independent variables, provides an integrable generalization to three-spatial dimensions of both the KP equations. Indeed, if y=0 or z=0, the t-independent part of the Lax pair of equation (1.1), namely equation (3.31b), reduces to the corresponding t-independent part of the Lax pair of KPI or KPII, respectively. Equation (1.6) provides a 3 + 2 generalization of the KP equations. Propositions 3.2 and 3.5 summarize the main results of this paper, namely they present a non-local d-bar formalism for solving the initial-value problem of these multi-dimensional equations. The unexpected achievement of the results of proposition 3.3 is the derivation of a formalism capable of dealing with a time dependence of the spectral function whose exponential contains a non-vanishing real part. Remarkably, this has implications beyond the area of nonlinear integrable PDEs: it introduces a new powerful transform for solving a large class of linear PDEs. It is worth noting that earlier modifications of the Fourier transform, such as those used by the author for the solution of boundary value problems on the half-line or the finite interval [37], involved the straightforward restriction of the direct Fourier transform in the appropriate physical domain. Here, the new transform involves a highly non-trivial modification in the Fourier space, thus, the new transform is indeed novel.

It is important to emphasize that the new formalism introduced in this paper for integrating equation (1.1) can also be applied to other multi-dimensional nonlinear PDEs, such as equations (2.3). This will be presented in future publications.

Several questions remain open. The most important among them is whether the 3 + 1 or 3 + 2 nonlinear PDEs introduced in this paper can be used to model physical or biological phenomena. In addition, in the formalism presented in both propositions 3.2 and 3.5, it was assumed that certain norms are sufficiently small. This excludes soliton-type solutions. The question of using such a solution for the scheme presented here remains open. Finally, the analysis of the singularity occurring at t=0 in the case that the initial datum does not satisfy the constraint in equation (3.42) is work in progress.

Data accessibility

This article has no additional data.

Conflict of interest declaration

I declare I have no competing interests.

Funding

I received no funding for this study.

References

  • 1.Ablowitz MJ, Clarkson PA. 1991. Solitons, nonlinear evolution equations and inverse scattering. London Mathematical Society Lecture Note Series. Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 2.Fokas AS, Zakharov VE (eds). 1992. Important developments in soliton theory. Berlin, Germany: Springer. [Google Scholar]
  • 3.Ablowitz MJ, Fokas AS, Musslimani ZH. 2006. On a new nonlocal formulation of water waves. J. Fluid Mech. 562, 313-344. ( 10.1017/S0022112006001091) [DOI] [Google Scholar]
  • 4.Zabusky NJ, Kruskal MD. 1965. Interaction of ‘solitons’ in a collisionless plasma and the recurrence of initial states. Phys. Rev. Lett. 15, 240-243. ( 10.1103/PhysRevLett.15.240) [DOI] [Google Scholar]
  • 5.Gardner CS, Greene JM, Kruskal MD, Miura RM. 1967. Method for solving the Korteweg-deVries equation. Phys. Rev. Lett. 19, 1095-1097. ( 10.1103/PhysRevLett.19.1095) [DOI] [Google Scholar]
  • 6.Lax PD. 1968. Integrals of nonlinear equations of evolution and solitary waves. Commun. Pur. Appl. Math. 21, 467-490. ( 10.1002/cpa.3160210503) [DOI] [Google Scholar]
  • 7.Fokas AS. 2009. Lax pairs: a novel type of separability. Inverse Prob. 25, 123007. ( 10.1088/0266-5611/25/12/123007) [DOI] [Google Scholar]
  • 8.Ablowitz MJ, Kaup DJ, Newell AC, Segur H. 1974. The inverse scattering transform-Fourier analysis for nonlinear problems. Stud. Appl. Math. 53, 249-315. ( 10.1002/sapm1974534249) [DOI] [Google Scholar]
  • 9.Fokas AS, Gelfand IM. 1994. Integrability of linear and nonlinear evolution equations and the associated nonlinear Fourier transforms. Lett. Math. Phys. 32, 189-210. ( 10.1007/BF00750662) [DOI] [Google Scholar]
  • 10.Ablowitz MJ, Fokas AS. 2003. Complex variables: introduction and applications. 2nd edn. Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 11.Fokas AS, Sung L-Y. 1992. On the solvability of the N-wave, Davey-Stewartson and Kadomtsev-Petviashvili equations. Inverse Prob. 8, 673-708. ( 10.1088/0266-5611/8/5/002) [DOI] [Google Scholar]
  • 12.Zhou X. 1990. Inverse scattering transform for the time dependent Schrödinger equation with application to the KPI equation. Commun. Math. Phys. 128, 551-564. ( 10.1007/BF02096873) [DOI] [Google Scholar]
  • 13.Boiti M, Leon JJ-P, Pempinelli F. 1989. Spectral transform and orthogonality relations for the Kadomtsev–Petviashvili equation. Phys. Lett. A 141, 96-100. ( 10.1016/0375-9601(89)90766-4) [DOI] [Google Scholar]
  • 14.Boiti M, Pempinelli F, Pogrebkov AK. 1994. Solutions of the KPI equation with smooth initial data. Inv. Prob. 10, 505-519. ( 10.1088/0266-5611/10/3/001) [DOI] [Google Scholar]
  • 15.Boiti M, Pempinelli F, Pogrebkov AK. 1994. Properties of solutions of the KPI equation. J. Math. Phys. 35, 4683-4718. ( 10.1063/1.530808) [DOI] [Google Scholar]
  • 16.Fokas AS, Sung L-Y. 1999. The inverse spectral method for KPI without the zero mass constraint. Math. Proc. Camb. Phil. Soc. 125, 113-138. ( 10.1017/S0305004198002850) [DOI] [Google Scholar]
  • 17.Boiti M, Pempinelli F, Pogrebkov AK, Prinari B. 2001. Towards an inverse scattering theory for nondecaying potentials of the heat equation. Inv. Prob. 17, 937-957. ( 10.1088/0266-5611/17/4/324) [DOI] [Google Scholar]
  • 18.Molinet L, Saut J-C, Tzvetkov N. 2002. Well-posedness and Ill-posedness results for the Kadomtsev–Petviashvili equation. Duke Math. J. 115, 353-384. ( 10.1215/S0012-7094-02-11525-7) [DOI] [Google Scholar]
  • 19.Molinet L, Saut J-C, Tzvetkov N. 2002. Global well-posedness for the KPI equation. Math. Ann. 324, 255-275. ( 10.1007/s00208-002-0338-0) [DOI] [Google Scholar]
  • 20.Fokas AS, Ablowitz MJ. 1983. On the inverse scattering of the time-dependent Schrödinger equation and the associated Kadomtsev-Petviashvili (I) equation. Stud. Appl. Math. 69, 211-228. ( 10.1002/sapm1983693211) [DOI] [Google Scholar]
  • 21.Fokas AS, Santini PM. 1990. Dromions and a boundary value problem for the Davey-Stewartson 1 equation. Physica D 44, 99-130. ( 10.1016/0167-2789(90)90050-Y) [DOI] [Google Scholar]
  • 22.Beals R, Coifman R. Scattering, transformations spectrales et équations d’évolution non linéaires. Séminaire Équations aux dérivées partielles (Polytechnique), exp. no. 22:1–9, 1980–1981.
  • 23.Beals R, Coifman R. Scattering, transformations spectrales et équations d’évolution non linéaires II. Séminaire Équations aux dérivées partielles, Exp. no. 21:1–8, 1981–1982.
  • 24.Ablowitz MJ, Bar Yaacov D, Fokas AS. 1983. On the inverse scattering transform for the Kadomtsev-Petviashvili equation. Stud. Appl. Math. 69, 135-143. ( 10.1002/sapm1983692135) [DOI] [Google Scholar]
  • 25.Fokas AS. 1983. Inverse scattering of first-order systems in the plane related to nonlinear multidimensional equations. Phys. Rev. Lett. 51, 3-6. ( 10.1103/PhysRevLett.51.3) [DOI] [Google Scholar]
  • 26.Fokas AS, Ablowitz MJ. 1984. On the inverse scattering transform of multidimensional nonlinear equations related to first-order systems in the plane. J. Math. Phys. 25, 2494-2505. ( 10.1063/1.526471) [DOI] [Google Scholar]
  • 27.Fokas AS. 2006. Integrable nonlinear evolution partial differential equations in 4+2 and 3+1 dimensions. Phys. Rev. Lett. 96, 190201. ( 10.1103/PhysRevLett.96.190201) [DOI] [PubMed] [Google Scholar]
  • 28.Fokas AS. 2009. Kadomtsev-Petviashvili equation revisited and integrability in 4 + 2 and 3 + 1. Stud. Appl. Math. 122, 347-359. ( 10.1111/j.1467-9590.2009.00437.x) [DOI] [Google Scholar]
  • 29.Fokas AS, van der Weele MC. 2018. Complexification and integrability in multidimensions.J. Math. Phys. 59, 091413. ( 10.1063/1.5032110) [DOI] [Google Scholar]
  • 30.van der Weele MC, Fokas AS. 2021. Solving the initial value problem for the 3-wave interaction equations in multidimensions. Computat. Methods Funct. Theory 21, 9-39. ( 10.1007/s40315-020-00357-2) [DOI] [Google Scholar]
  • 31.Fokas AS. 2007. Nonlinear Fourier transforms, integrability and nonlocality in multidimensions. Nonlinearity 20, 2093-2113. ( 10.1088/0951-7715/20/9/005) [DOI] [Google Scholar]
  • 32.Fokas AS. 2008. Soliton multidimensional equations and integrable evolutions preserving Laplace’s equation. Phys. Lett. A 372, 1277-1279. ( 10.1016/j.physleta.2007.09.037) [DOI] [Google Scholar]
  • 33.Fokas AS. 2008. The D-bar method, inversion of certain integrals and integrability in 4+2 and 3+1 dimensions. J. Phys. A 41, 344006. ( 10.1088/1751-8113/41/34/344006) [DOI] [Google Scholar]
  • 34.Fokas AS, Cao Y, He J. Multi-solitons, multi-breathers and multi-rational solutions of integrable extensions of the Kadomtsev-Petviashvili equation in three-dimensions. (In preparation).
  • 35.Fokas AS, Sung L-Y. 1999. The Cauchy problem for the Kadomtsev-Petviashvili—I equation without the zero mass constraint. In Mathematical Proc. of the Cambridge Philosophical Society, vol. 125, pp. 113–138. Cambridge, UK: Cambridge University Press.
  • 36.Lenells J, Fokas AS. 2015. The nonlinear Schrödinger equation with t-periodic data: I. exact results. Proc. R. Soc. A 471, 20140925. ( 10.1098/rspa.2014.0925) [DOI] [Google Scholar]
  • 37.Fokas AS, Kaxiras E. 2022. Modern mathematical methods for computational sciences and engineering. Singapore: World Scientific. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This article has no additional data.


Articles from Proceedings. Mathematical, Physical, and Engineering Sciences are provided here courtesy of The Royal Society

RESOURCES