Paper The following article is Open access

Deformations of dispersionless Lax systems*

Published 7 November 2023 © 2023 The Author(s). Published by IOP Publishing Ltd
, , Citation Wojciech Kryński 2023 Class. Quantum Grav. 40 235013 DOI 10.1088/1361-6382/ad0748

0264-9381/40/23/235013

Abstract

We study dispersionless Lax systems and present a systematic method for deriving new integrable systems from given ones. Our examples include the dispersionless Hirota equation, the generalized heavenly equation, and equations related to Veronese webs.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 license. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

In this paper we study dispersionless integrable systems that arise as integrability conditions for foliation $\mathcal{F}$ on bundle $\mathcal{B}$ with 1-dimensional fibers over manifold $\mathcal{M}$. This framework can be adapted to the heavenly equations [29, 30], the Manakov–Santini system [24] (see also [7]), the Dunajski–Tod equation [9] (see also [3]), the hyper-CR equation [6], the dispersionless Hirota equation [8] (known also as the abc-equation [31]), equations related to GL(2)-structures and web geometry [11, 17, 18, 2022] and many others [1, 2, 5, 10, 13, 15, 23, 25]. The quotient space $\mathcal{T} = \mathcal{B}/\mathcal{F}$ is usually referred to as the (real) twistor space, and it gives rise to the following double fibration picture

Our objective is to develop a method for deriving new integrable systems on $\mathcal{M}$. We shall equip the twistor space $\mathcal{T}$ with additional geometric structures and demonstrate how deformations of these structures correspond to deformations of the original integrable system. In our applications, we will focus on isospectral systems only.

The motivation behind this work is to provide a geometric justification for recent results presented in [17, 27], where the authors have obtained new integrable systems by replacing constants with specific functions. Our significant novel application is a class of new deformations of Schief's generalized heavenly equation [30]. We establish a one to one correspondence between anti-self-dual metrics of split signature satisfying the Einstein vacuum equation in four dimensions and solutions $u = u(x^1,x^2,x^3,x^4)$ to

where $u_i = \partial_{x^i}u$, $u_{ij} = \partial_{x^i}\partial_{x^j}u$, and λi 's are arbitrary differentiable functions of two independent variables. In section 3.3 we prove that the equation possesses an isospectral Lax pair for any choice of λi 's. Clearly, for $\lambda_i = \mathrm{const}$ the equation reduces to the generalized heavenly equation. On the other hand, when λi 's are non-constant functions, the equation becomes a direct generalization of deformations considered before in [16], where only one λi depends on two variables, or in [27], where λi 's depend on xi only. It is worth mentioning that in [16] an algebraic approach of deformations of symmetry algebras is employed, while in [27] normal forms of Nijenhuis operators are used (an approach previously applied in the context of the Einstein–Weyl geometry in dimension 3 [17]). A priori the two methods are unrelated, but in the present paper, we generalize results of both approaches simultaneously.

Our primary goal, however, is to present a unified framework that can be subsequently applied to other equations. This framework is elaborated upon in section 2. The approach developed in the present paper is based on twistor methods that originate from works of Penrose [28] and Hitchin [12]. All the aforementioned systems describe well known classes of geometric structures on $\mathcal{M}$. In particular, as in the case of the heavenly equations, these structures encompass anti-self-dual metrics (e.g. [1, 3, 7, 9, 29]), Einstein–Weyl structures (e.g. [7, 8, 10, 21, 24]), or higher dimensional counterparts like GL(2)-structures or Veronese and Kronecker webs (see [11, 17, 18, 20, 22]). However, establishing the generality of solutions often proves to be a subtle and nontrivial task. Furthermore, as demonstrated in [17, 26], it can be even more challenging to determine whether two equations describe the same set of structures, as these equations are often non equivalent from the viewpoint of equivalence of partial differential equations (PDEs). On the contrary, our method benefits from the twistorial framework and automatically establishes a correspondence between solutions of different equations. This idea was initially developed in [21] and applied in the specific case of the dispersionless Hirota equation. In the present paper we extend this approach. In addition to the aforementioned heavenly equation, we provide other applications in section 3. In particular we consider the hierarchy of [19] related to the Veronese webs in section 3.2 (note that in dimension 4 the structures arise in the context of GL(2)-geometry and exotic holonomy groups [4, 22]).

2. General constructions

In this section, we start with a general framework of deformations and then we formulate theorem 2.2 which asserts that the solution space of a deformed equation is in a one to one correspondence with the solution space of the original equation. Once the definitions are introduced, the result is straightforward. For applications it will be crucial to determine whether two equations are mutual deformations or to construct deformations of a given equation. That, in general requires some effort, but once the work is done theorem 2.2 can be applied.

2.1. Equations

Let

be a vector bundle over a manifold $\mathcal{M}$ and let

be a rank-1 fiber bundle over $\mathcal{M}$. Denote by

the pullback bundle of U. Then, for any section u of U, there is a corresponding pullback section $\hat u$ of $\hat U$. Moreover, let $L_0,\ldots, L_s$ be differential operators acting on sections of $\hat U$ with values in the tangent bundle $T\mathcal{B}$,

Hence, for a given a section u of U, $L_0(\hat u),\ldots, L_s(\hat u)$ are vector fields on $\mathcal{B}$ depending on u and its derivatives up to certain order.

Definition. We say that u is a solution to the Lax system defined by $(L_i)_{i = 0,\ldots,s}$ if the distribution

on $\mathcal{B}$ is integrable. The solution space of the Lax system will be denoted $\mathfrak{U}_{(L_0,\ldots,L_s)}$.

The fibers of $\mathcal{B}$ typically come with a given parametrization denoted as λ, which is often referred to as the spectral parameter. Later, for simplicity, we shall write $L_i(u)$ instead of $L_i(\hat u)$. The Lax systems are usually defined by pairs $(L_0,L_1)$ of vector fields. The foliation of integral leaves of Du will be denoted $\mathcal{F}_u$. We stress that $\mathcal{F}_u$ essentially depends on a solution $u\in\mathfrak{U}_{(L_0,\ldots,L_s)}$.

For the rest of the paper we shall assume that U and $\mathcal{B}$ are trivial bundles

Moreover, we shall study $\mathcal{M}$ locally around a given point and hence we will assume for simplicity that $\mathcal{M} = \mathbb{R}^n$. The later identification gives the initial coordinates on $\mathcal{M}$.

2.2. Twistor correspondence

Let $u\in\mathfrak{U}_{(L_0,\ldots,L_s)}$ be a solution to a Lax system $(L_i)_{i = 0,\ldots,s}$ on $\mathcal{M}$. The corresponding twistor space is defined as the quotient space $\mathcal{T}_u = \mathcal{B}/\mathcal{F}_u$ of the foliation $\mathcal{F}_u$ associated to u. Let

be the quotient mapping.

We assume that fibers of $\pi_\mathcal{M}$ intersect fibers of $\pi_{\mathcal{T}_u}$ transversally, i.e. all $L_i(u)$ for $i = 0,\ldots,s$, are nowhere tangent to the fibers of $\pi_\mathcal{M}$. It follows that one can establish a correspondence between points in $\mathcal{M}$ and certain curves in $\mathcal{T}_u$, and conversely a correspondence between points in $\mathcal{T}_u$ and submanifolds in $\mathcal{M}$. Indeed, let $x\in \mathcal{M}$ and denote by $\hat\gamma_x$ the curve in $\mathcal{B}$ defined as the fiber $\pi^{-1}_\mathcal{M}(x)$. Let $\gamma_{x,u} = \pi_{\mathcal{T}_u}(\hat\gamma_x)$. Then $\gamma_{x,u}$ is a curve in $\mathcal{T}_u$. Conversely, let $p\in \mathcal{T}_u$ and denote by $\hat F_p$ the submanifold in $\mathcal{B}$ defined as the fiber $\pi^{-1}_{\mathcal{T}_u}(p)$. Let $F_p = \pi_\mathcal{M}(\hat F_p)$. Then Fp is a submanifold of $\mathcal{M}$. For a given solution u, we denote

Our aim now is to introduce a class of distinguished coordinates on $\mathcal{M}$ parameterized by solutions u of a Lax system. It is sufficient to find a map from $\Gamma_u$ to $\mathbb{R}^n$ for any solution $u\in\mathfrak{U}_{(L_0,\ldots,L_s)}$, because of the correspondence between points in $\mathcal{M}$ and $\Gamma_u$.

Definition. A collection of corank-1 submanifolds $(T_1, T_2,\ldots,T_m)$ of $\mathcal{T}_u$ is called a transversal system for a family $\Gamma_u$ if any $\gamma\in\Gamma_u$ and Ti , $i = 1,\ldots,m$, intersect transversally exactly at one point.

Definition. Let $T_u = (T_1,\ldots,T_m)$ be a transversal system of submanifolds of $\mathcal{T}_u$ for a family $\Gamma_u$ and assume that any Tl possess a fixed local coordinate system $(x_l^1,\ldots,x_l^k)\colon T_l\to\mathbb{R}^k$. Any function $x_l^j$ defines a function on $\Gamma_u$ by the formula

where the right hand side is the value of $x_l^j$ at the intersection point of γ and Tl (unique by definition), see figure 1. We say that maps $(x_l^j)$ induce coordinates on $\Gamma_u$ if there is a subset $\{x^1,\ldots,x^n\}\subset\{x_l^j\ |\ l = 1,\ldots,m,\ j = 1,\ldots,k\}$ defining local coordinates on $\Gamma_u$. If this is the case then $\phi_u = (x^1,\ldots,x^n)\colon \Gamma_u\to \mathbb{R}^n$ are called induced coordinates on $\mathcal{M}$.

Figure 1.

Figure 1. Transversal system $T_u = (T_1,T_2,T_3,T_4)$ and induced coordinates $(x^i)$ of a point x represented by a curve γx in the twistor space $\mathcal{T}_u$.

Standard image High-resolution image

The coordinate functions obtained from the construction are very special. Indeed, we have the following.

Proposition 2.1. If $(x^i)_{i = 1,\ldots,n}$, is a coordinate system on $\mathcal{M}$ induced by a transversal system on Tu , then all foliations $x^i = \mathrm{const}$, $i = 1,\ldots,n$, are tangent to the projections from $\mathcal{B}$ to $\mathcal{M}$ of certain integral leaves of $\mathcal{F}_u$.

Proof. Indeed, a condition $x^i = c$ for some $c\in\mathbb{R}$ fixes a submanifold $S\subset T_s$, where Ts is one of the submanifolds from the transversal system Tu . Points $p\in S$ correspond to submanifolds $F_p\subset\mathcal{M}$ that consist of points represented by curves from $\Gamma_u$ intersecting S. On the other hand these submanifolds are, by definition, projections to $\mathcal{M}$ of integral submanifolds of Du . □

2.3. Deformations of Lax systems

Let $(L_i)_{i = 0,\ldots,s}$ be a Lax system on $\mathcal{M}$. Our goal is to consider two different transversal systems, $T_u^1$ and $T_u^2$, on the twistor space $\mathcal{T}_u$, for any solution $u\in\mathfrak{U}_{(L_0,\ldots,L_s)}$ (see figure 2). The two transversal systems give rise to two induced coordinate systems $\phi_u^1\colon\mathcal{M}\to\mathbb{R}^n$ and $\phi_u^2\colon\mathcal{M}\to\mathbb{R}^n$, respectively.

Figure 2.

Figure 2. Two transversal systems on the twistor space $\mathcal{T}_u$, giving two different coordinate systems on $\mathcal{M}$.

Standard image High-resolution image

The coordinate systems $\phi_u^1$ and $\phi_u^2$ are parameterized by solutions $u\in\mathfrak{U}_{(L_0,\ldots,L_s)}$ and, a priori, take values in different copies of $\mathbb{R}^n$. However, we can as well assume that all maps $\phi^1_u$, $u\in\mathfrak{U}_{(L_0,\ldots,L_s)}$, take values in a one fixed copy of $\mathbb{R}^n$, denoted $\mathcal{M}_1$, and similarly all maps $\phi^2_u$, $u\in\mathfrak{U}_{(L_0,\ldots,L_s)}$, take values in another fixed copy of $\mathbb{R}^n$, denoted $\mathcal{M}_2$. If it is the case, we can consider $L^1_{i,u} = \phi^1_{u*}L_i$ and $L^2_{i,u} = \phi^2_{u*}L_i$, defined on $\mathcal{M}_1$ and $\mathcal{M}_2$, respectively, and we look for new Lax systems: $(L^1_0,\ldots,L^1_s)$ on $\mathcal{M}_1$ and $(L^2_0,\ldots,L^2_s)$ on $\mathcal{M}_2$ such that

The two Lax system, $(L^1_0,\ldots,L^1_s)$ on $\mathcal{M}_1$ and $(L^2_0,\ldots,L^2_s)$ on $\mathcal{M}_2$, obtained in this manner will be referred to as deformations of the original system. In examples, the procedure follows these steps:

  • Step 1.  
    Consider a Lax system in a specific coordinate system and prove that it can be interpreted as an induced coordinate system defined by a transversal system for each solution function u;
  • Step 2.  
    Deform all transversal systems and find new coordinates for each solution function u;
  • Step 3.  
    Derive a new Lax system based on the formulas in new coordinates.

We introduce the following formal definition.

Definition. Let $(L_i)_{i = 0,\ldots,s}$ be a Lax system on $\mathcal{M}$ and let $\mathfrak{U} = \mathfrak{U}_{(L_0,\ldots,L_s)}$ be its solution space. A Lax system $(\tilde L_i)_{i = 0,\ldots,s}$ on $\tilde{\mathcal{M}}$ is a weak deformation defined by a family of transversal systems $(T_u)_{u\in\mathfrak{U}}$ and induced coordinates $(\phi_u)_{u\in\mathfrak{U}}$ if for any solution $u\in \mathfrak{U}$ the pullback

to $\tilde{\mathcal{M}}$ via the corresponding φu is a solutions to $(\tilde L_i)_{i = 0,\ldots,s}$.

Let us point out that in the definition above there is a different transformation φu for each solution function u. On the other hand, we assume that they all map $\mathcal{M}$ to $\tilde{\mathcal{M}}$, which is a fixed manifold, and therefore we can consider a new Lax system on $\tilde{\mathcal{M}}$

Definition. A Lax system $(\tilde L_i)_{i = 0,\ldots,s}$ on $\tilde{\mathcal{M}}$ is a deformation of a Lax system $(L_i)_{i = 0,\ldots,s}$ on $\mathcal{M}$ if the two systems are mutual weak deformations of each other and

holds.

We get the following tautological result.

Theorem 2.2. If a Lax system $(\tilde L_i)_{i = 0,\ldots,s}$ on $\tilde{\mathcal{M}}$ is a deformation of a Lax system $(L_i)_{i = 0,\ldots,s}$ on $\mathcal{M}$ then there is a one to one correspondence between the two corresponding solution spaces $\mathfrak{U}_{(L_0,\ldots,L_s)}$ and $\mathfrak{U}_{(\tilde L_0,\ldots,\tilde L_s)}$.

Remark. An important aspect of paper [27] is the realizability of vacuum anti-self-dual Einstein metrics as solutions to the equations that arise from the normal forms of Nijenhuis tensors. One can obtain an alternative proof of this result by demonstrating that the equations in [27] are deformations of the heavenly equation (we achieve this in section 3.3), and then apply theorem 2.2. This observation is general: the geometric structures described by deformed Lax systems exhibit the same level of generality as those described by the original ones.

2.4. Higher order deformations

Higher order deformation can be defined by replacing some of Ti in a transversal system with their tangent bundles (or higher order tangent bundles). This can be visualized as a limit, where two (or more) submanifolds Ti in a transversal system become arbitrary close. Then, one can introduce coordinates of a curve γx as a first jet (or a higher order jet) at the intersection point of γx and Ti . This approach was applied in the case of the Hirota equation in [21]. We shall not explain the details here, but postpone it to separate studies (equations that can be treated in this way include the hyper-CR equation, the Manakov–Santini system and the system that governs the half-integrable Cayley structures introduced in [23]).

3. Examples

This section contains several examples illustrating the general constructions from the previous section. A common feature among all of them is that the twistor space is fibered over a projective space. This is manifested in the Lax pairs by the absence of a term in the direction of $\partial_\lambda$, i.e. the Lax systems are isospectral. It follows that the spectral coordinate λ on the bundle $\mathcal{B}$ descends to a well defined function on $\mathcal{T}_u$, for any solution $u\in\mathfrak{U}_{(L_i)}$.

3.1. Dispersionless Hirota equation

It is proved in [8] that the solutions to the dispersionless Hirota equation are in a one to one correspondence with the 3-dimensional hyper-CR Einstein–Weyl structures. The equation, written on manifold $\mathcal{M} = \mathbb{R}^3$ with coordinates $x^1,x^2,x^3$, reads

Equation (1)

where $a,b,c\in\mathbb{R}$ are non-zero constants such that $a+b+c = 0$, and $u_i = \partial_iu$. The corresponding Lax pair is of the following form (see [8])

Equation (2)

where λ is an additional affine coordinate on the rank-1 bundle $\mathcal{B} = \mathcal{M}\times\mathbb{R} P^1$. Note that for any value λ, the two vector fields L0 and L1 span an integrable distribution on $\mathcal{M}$. Indeed, the Hirota equation is the integrability condition. The corresponding conformal metric $[g]$ and the Weyl connection can be found in [8] (formula (4)).

Let $\lambda_1, \lambda_2,\lambda_3$ be such that $a = \lambda_2-\lambda_3$, $b = \lambda_3-\lambda_1$ and $c = \lambda_1-\lambda_2$. Performing a Möbius transformation $\lambda\mapsto\frac{1}{\lambda_1-\lambda}$ one puts the Lax pair (2) in the form

Equation (3)

which clearly gives the same Lax equation for u. One checks that for $\lambda = \lambda_i$ the distribution spanned by L0 and L1 is tangent to the foliation $x^i = \mathrm{const}$.

It follows that coordinates $(x^1,x^2,x^3)$ can be interpreted as induced coordinates from a transversal system. Indeed, $\mathcal{T}_u = \mathcal{B}/D_u$ is the space of all integral manifolds of $D_u = \operatorname{span}\{L_0,L_1\}$. Moreover, $\mathcal{T}_u$ is two-dimensional and, as mentioned before, λ is a well defined function on $\mathcal{T}$ because it is constant on leaves of Du . The transversal system is defined by submanifolds

Function xi is a diffeomorphism from Ti to $\mathbb{R}$, i.e. it parameterizes leaves of the aforementioned foliation $x^i = \mathrm{const}$ on $\mathcal{M}$.

Let us replace now Ti by general submanifolds of $\mathcal{T}_u$ of the form

for some functions $f_i\colon\mathcal{T}_u\to\mathbb{R}$ on the twistor space (with 0 being a regular value). It is proved in [21] that any choice of coordinates xi on $\tilde T_i$ transforms (1) into

Equation (4)

where now each $\lambda_i(x^i)$, $i = 1,2,3$, is a function of one variable. Precisely, λi is defined as the restriction of function λ from $\mathcal{T}_u$ to submanifold $\tilde T_i$, and xi is a coordinate function on $\tilde T_i$, which consequently becomes an induced coordinate on $\mathcal{M}$. In the previous case this restriction of λ to Ti is just a constant function. The Lax pair for (4) is given by

Equation (5)

Equation (4) was derived for the first time in [17] where the authors analyzed the so-called Nijenhuis operators associated to Veronese webs. The general assumption in their approach is that the Nijenhuis operators have no singular points which is reflected in the condition $\partial_{x^i}\lambda_i(x^i)\neq 0$. In this case a simple coordinate change (a point transformation) gives $\lambda_i(x^i) = x^i$. In our approach $\lambda_i(x^i)$ can be arbitrary smooth function of one variable. In particular we admit $\partial_{x^i}\lambda_i(x^i) = 0$ for certain values of xi .

In [17] it is also proved that equations with constant coefficients λi are contactly non-equivalent to the one with non-constant λi (one can check that the corresponding symmetry groups are different). Note that the coordinate change between coordinate systems induced by different transversal systems does not establish a contact equivalence of the Lax systems. The reason is that the coordinate change depends on a given solution u (it is a subtle point in the definition of deformations which makes the procedure nontrivial).

On the other hand there is a Bäcklund transformation between solutions to (1) and (4). It was found in [17] and is given by

where u and $\tilde u$ are solutions to (1) and (4), respectively, where in (4) $\lambda_i(x^i)$ are replaced by $\tilde\lambda_i(x^i)$ for clarity.

3.2. Higher dimensional Veronese webs

Veronese webs are higher dimensional counterparts of the hyper-CR Einstein–Weyl structures described by equation (1). We refer to [8] for the proof of the 3-dimensional correspondence between the Veronese webs and the Einstein–Weyl structures. The general case was studied in [19] where it is additionally proved that the webs underlie very particular integrable paraconformal structures (so called totally geodesic GL(2)-geometries).

The Veronese webs on $\mathcal{M} = \mathbb{R}^n$ are 1-parameter families of corank-one foliations [14, 18, 31] such that the annihilating 1-forms ωλ give rise to a rational normal curves $\lambda\mapsto\mathbb{R}\omega_\lambda(x)\in P(T^*_x\mathcal{M})$ at each point $x\in\mathcal{M}$. The curves replace cones of null directions of $[g]$ in the 3-dimensional case. It is proved in [19] that the Veronese webs are in a one to one correspondence with solutions to the following system

Equation (6)

where $i,j,k = 1,\ldots,n$. Indeed, the equations are equivalent to the integrability condition

where ωλ is given by

Equation (7)

System (6) appeared in [8] for the first time and turned out to be a Lax system with

Equation (8)

The corresponding twistor space is two dimensional and the most natural transversal system can be defined as before

where now $i = 1,\ldots,n$. Note that, as in the 3 dimensional case, λ is constant along vector fields Li independently of u, and therefore can be treated as a function on the twistor space $\mathcal{T}_u$.

Deforming Ti as in the 3-dimensional case we get that the constants λi can be replaced by arbitrary (smooth) functions of one variable $\lambda_i = \lambda_i(x_i)$. Indeed we have

Equation (9)

$i,j,k = 1,\ldots,n$, as an integrability condition for ωλ given by (7) with constants λi replaced by functions. Note that, if $\partial_{x^i}\lambda_i\neq 0$ then coordinate change (a point transformation) reduces $\lambda_i(x^i) = x^i$.

Finally, as in the case of the single Hirota equation we find that (6) and (9) are related by a Bäcklund transformation

where $\tilde\lambda_i = \tilde\lambda_i(x^i)$ are functions.

3.3. Heavenly equations

In a recent paper [15] an interesting approach to the heavenly equations is presented. In this context $\mathcal{M} = \mathbb{R}^4$ and one looks for split-signature self-dual Ricci flat metrics. It is well known that metrics of this type are in a one to one correspondence with solutions to the (first) Plebański equation

Equation (10)

which has a Lax pair on $\mathcal{B} = \mathcal{M}\times\mathbb{R} P^1$ of the following form

The authors of [15] utilize the eigenfunctions as coordinates, where eigenfunctions are understood as solutions to linear system

Equation (11)

for unknown function ψ, where $\lambda^*$ is a fixed value of λ. It follows that any solution ψ is a function on a submanifold of $\mathcal{T}_u$ defined as

where as in the previous examples λ is treated here as a function on $\mathcal{T}_u$. Conversely, any function on $T_{\lambda^*}$ is a solution to (11).

In the present case $\mathcal{T}_u$ is 3-dimensional, and consequently $T_{\lambda^*}$ is a 2-dimensional surface in $\mathcal{T}$. It follows that there are two functionally independent functions on each $T_{\lambda^*}$. Konopelchenko et al [15] uses the functions as coordinates on $\mathcal{M}$. In our terminology they are induced coordinates for a transversal system. For the Plebański equation it is enough to take two distinct values of $\lambda^*$, say λ1 and λ2, and the corresponding transversal system $T_u = (T_1,T_2)$. We get four functions $x^i_l$, $i,l = 1,2$ in this way, as needed.

However, one can take $T_u = (T_1,T_2,T_3, T_4)$ for four different values of $\lambda^* = \lambda_l$, $l = 1,2,3,4$. Then coordinates on Tl give 8 functions $x^i_l$ in total. We pick 4 of them, one for each λl . It is proved in [15] (theorem 4.1) that as a result one gets the general heavenly equation (see also [30])

Equation (12)

The corresponding Lax pair has the following form

Now, the transversal system Tu can be deformed to a quadruple $\tilde T_u = (\tilde T_1,\tilde T_2,\tilde T_3, \tilde T_4)$ of arbitrary surfaces (transversal with respect to the family $\Gamma_u$) in $\mathcal{T}_u$. Similarly to the Hirota equation such a deformation replaces each λi by a smooth function which in general can be written in the form

Equation (13)

where $(x_i^1,x_i^2)$ are two coordinate functions on $\tilde T_i$. One of them, say $x_i^1$, becomes a coordinate on $\mathcal{M}$ (denote it by xi ). Denote by φi the pullback of the second coordinate function (i.e. $x_i^2$) from $\tilde T_i$ to $\mathcal{M}$. Then $\phi^i = \phi^i(x^1,x^2,x^3,x^4)$ is a general function satisfying (11) for $\lambda^* = \lambda_i(x^1_i,x^2_i)$.

The Lax pair in the new coordinate system reads

where $\lambda_i = \lambda_i(x_i^1,x_i^2)$ and $\tilde U_{ij}$ are certain expressions involving second order derivatives of the original u and of the corresponding transformation of coordinates. However, it turns out that $\tilde U_{ij}$ can be computed explicitly as

where on the right hand side there are second order derivatives with respect to the new coordinates. The formula holds provided that new coordinates, as well as functions φi , are defined as functions on $\tilde T_i$, or equivalently they are solutions to

Equation (14)

in new coordinates. The integrability condition for the new Lax pair gives (12) (with λi upgraded to functions). Notice, that there is one obvious solution to (14) which is $\phi^i = \partial_i\tilde u = \tilde u_i$ (written in new coordinates). Summarizing, if we replace for simplicity $u: = \tilde u$, we get the following equation

In line with the previous examples, one can provide a Bäcklund transformation for the generalized heavenly equation with different values of λi , regardless they are constant or not. In the current case, the transformation is expressed via second-order equations:

Note that equations are linear with respect to u, provided that the corresponding λi 's are constant.

As mentioned in the Introduction, specific cases of this particular deformation appeared independently in [16, 27]. In [16] authors consider deformations of the symmetry algebra and their algebraic procedure give our deformed equation but only with respect to λ4 and coordinate x4 (it is case (I) of [16]). Function Q in [16] is given by

On the other hand authors of [27] apply approach of the Nijenhuis operators and get $\lambda_i = \lambda_i(x^i)$. This corresponds to the case when the foliation $x^1_i = \mathrm{const}$ on $\tilde T_i$ coincide with the foliation $\lambda = \mathrm{const}$ on $\mathcal{T}_u$ restricted to $\tilde T_i$. As in the previous examples our method covers singularities of the Nijenhuis operators—corresponding to points where $\partial_{x^i}\lambda_i(x^i) = 0$.

Acknowledgments

I am grateful to Boris Kruglikov for helpful conversations and comments.

Data availability statement

No new data were created or analyzed in this study.

Please wait… references are loading.
10.1088/1361-6382/ad0748