You are on page 1of 49

Results. Math.

60 (2011), 53–101
c 2011 Springer Basel AG
1422-6383/11/010053-49
published online July 7, 2011
DOI 10.1007/s00025-011-0167-0 Results in Mathematics

Equations of Pseudo-Spherical Type


(After S.S. Chern and K. Tenenblat)
Enrique G. Reyes
For Prof. Keti Tenenblat on her 65th birthday

Abstract. This paper surveys some recent developments around the notion
of a scalar partial differential equation describing pseudo-spherical sur-
faces due to Chern and Tenenblat. It is shown how conservation laws,
pseudo-potentials, and linear problems arise naturally from geometric
considerations, and it is also explained how Darboux and Bäcklund trans-
formations can be constructed starting from geometric data. Classification
results for equations in this class are stated, and hierarchies of equations of
pseudo-spherical type are introduced, providing a connection between dif-
ferential geometry and the study of hierarchies of equations which are the
integrability condition of sl(2, R)-valued linear problems. Furthermore,
the existence of correspondences between any two solutions to equations
of pseudo-spherical type is reviewed, and a correspondence theorem for
hierarchies is also mentioned. As applications, an elementary immersion
result for pseudo-spherical metrics arising from the Chern–Tenenblat con-
struction is proven, and non-local symmetries of the Kaup–Kupershmidt,
Sawada–Kotera, fifth order Korteweg–de Vries and Camassa–Holm (CH)
equation with non-zero critical wave speed are considered. It is shown
that the existence of a non-local symmetry of a particular type is enough
to single the first three equations out from a whole family of equations
describing pseudo-spherical surfaces while, in the CH case, it is shown
that it admits an infinite-dimensional Lie algebra of non-local symme-
tries which includes the Virasoro algebra.

Mathematics Subject Classification (2010). Primary 53B20;


Secondary 37K10, 76M60.

Keywords. Equations of pseudo-spherical type, Bäcklund transformations,


Darboux transformations, Hierarchies of pseudo-spherical type,
integrability, nonlocal symmetries, Kaup–Kupershmidt equation,
Sawada–Kotera equation, Korteweg–de Vries equation,
Camassa–Holm equation.
54 E. G. Reyes Results. Math.

1. Introduction
The class of scalar differential equations of pseudo-spherical type (or “describ-
ing pseudo-spherical surfaces”) was introduced by Chern and Tenenblat,
see [12], motivated by the following observation (Sasaki [70]): the domains
of generic—in a sense to be made precise in Sect. 2—solutions u(x, t) of equa-
tions integrable by the Ablowitz, Kaup, Newell, and Segur (AKNS) inverse
scattering approach [1] can be equipped, whenever their associated linear prob-
lems are real, with Riemannian metrics of constant Gaussian curvature equal
to −1. Chern and Tenenblat called a scalar partial differential equation Ξ = 0
of pseudo-spherical type if generic solutions of Ξ = 0 determine Riemannian
metrics of constant Gaussian curvature −1 on open subsets of their domains
(a precise definition is given in Sect. 2). An example of such a PDE is of course
the ubiquitous sine–Gordon equation [12,22,70,78].
The importance of this structure was recognized first in the realm of inte-
grable systems: it was pointed out in [12] that some fundamental properties
of equations of pseudo-spherical type, such as the existence of conservation
laws and Bäcklund transformations can be studied by geometric means, and
also that these equations are the integrability condition of sl(2, R)-valued lin-
ear problems, so that one can try to solve them using a scattering/inverse
scattering approach [2,3]. These observations were then developed in several
papers by Prof. Tenenblat and her coworkers, see [4,10,51] and the monograph
[78], and by the present author in the later papers [31,55,56,60,61]. Also, the
authors of [12], and then several other researchers, see [20,37,38,51,54,57],
obtained characterization results for equations of pseudo-spherical type, essen-
tially “unwrapping” the zero curvature condition arising from the aforemen-
tioned fact that equations in this class admit associated sl(2, R)-valued linear
problems. These characterization theorems are quite technical and not always
of straightforward applicability, but they have proven useful in obtaining actual
explicit classifications of equations describing pseudo-spherical surfaces, see
[30,52,53], and as a tool for the comparison of different types of integrability
for evolution equations [25,54].
It has been also realized that, independently of their integrability prop-
erties, new analytic results about equations of pseudo-spherical type can be
found motivated by geometric considerations: inspired by the fact that two sur-
faces of constant Gaussian curvature equal to −1 are locally indistinguishable,
Kamran and Tenenblat [38]—and then the present author [57]—have showed
that if Ξ = 0 and Ξ  = 0 describe pseudo-spherical surfaces, there exists a
smooth mapping transforming a (generic) local solution u(x, t) of Ξ = 0 into
a (generic) local solution u (x,   = 0. In some instances one can find
t ) of Ξ
an explicit formula for u (x, 
t ) in terms of u(x, t), see [38,57,58]. Also, moti-
vated by the observation that evolution equations ut = F which are the inte-
grability condition of a non-trivial one-parameter family of linear problems
vx = X v, vt = T v (in which X and T are, say, sl(2, R)-valued functions of u
Vol. 60 (2011) Equations of Pseudo-Spherical Type 55

and a finite number of its derivatives with respect to x) are usually members of
infinite hierarchies of evolution equations uτn = Fn such that (see for instance
[19]): (a) the flows generated by the equations uτn = Fn commute, and (b)
each equation uτn = Fn is the integrability condition of a linear problem of
the form vx = X v, vτn = Tn v, hierarchies of evolution equations describ-
ing pseudo-spherical surfaces were defined in [59]; it was proven that they do
possess characteristics (a) and (b), and moreover, that there exist correspon-
dences between (suitably generic) solutions of any two hierarchies of equations
of pseudo-spherical type.
Thus, the Chern–Tenenblat structure is of interest for integrability, anal-
ysis and certainly geometry, since it is quite natural to ask which equations
(besides the classical sine–Gordon equation [22]) allow one to construct pseudo-
spherical metrics. With respect to analysis, one can trace back the relevance
of the class of equations of pseudo-spherical type to the local uniqueness of
surfaces of constant Gaussian curvature; with respect to integrability, one
would say that the relevance of the Chern–Tenenblat structure arises from
the fact that the Chern–Tenenblat construction [12] encodes in differential
geometric terms the idea of a Wahlquist–Eastbrook quadratic pseudopoten-
tial [80]. This is the reason behind the fact that one can uncover conservation
laws and Darboux/Bäcklund transformations starting from geometric consid-
erations.
In this article the developments mentioned above are reviewed, and some
new results on the application of the Chern–Tenenblat theory to integrable
systems are discussed. Specifically, it is shown how to construct non-local
symmetries of some interesting equations of mathematical physics (the Kaup–
Kupershmidt, Sawada–Kotera, 5th-order KdV and Camassa–Holm equations),
how non-local symmetries can be used to single out integrable equations,
and how the Virasoro algebra appears in the theory of non-local symme-
tries. This last part builds up on work reported in the papers [31,33,34,60–
63].

Guide: Section 2 contains basic definitions (Sect. 2.1), characterization


results (Sect. 2.2), and the definition of hierarchies of pseudo-spherical type
(Sect. 2.3). Section 3 is on associated linear problems (Sect. 3.1), conserva-
tion laws (Sect. 3.2), pseudo-potentials (Sect. 3.2), and conservation laws and
pseudo-potentials for the Camassa–Holm equation (Sect. 3.3). Section 4 is on
Darboux/Bäcklund transformations and modified equations; in particular, a
recently introduced modified Camassa–Holm equation is considered. Section 5
is on correspondence theorems for equations and hierarchies of pseudo-spher-
ical type. Finally, Sect. 6 is on non-local symmetries. The general theory is
briefly discussed in Sect. 6.1, the Kaup–Kupershmidt, Sawada–Kotera and
5th-order KdV are discussed in Sect. 6.2, and non-local symmetries of the
Camassa–Holm equation are considered in Sect. 6.3.
56 E. G. Reyes Results. Math.

2. Equations of Pseudo-Spherical Type


2.1. Basic Definitions
Definition 2.1. A two-dimensional manifold M is a pseudo-spherical surface if
there exist one-forms ω 1 , ω 2 , ω 3 on M satisfying the independence condition
ω 1 ∧ ω 2 = 0, and such that the following structure equations hold:
d ω1 = ω3 ∧ ω2 , d ω2 = ω1 ∧ ω3 , d ω3 = ω1 ∧ ω2 . (2.1)
If M is a pseudo-spherical surface, it is a Riemannian manifold equipped
with the metric ds2 = (ω 1 )2 + (ω 2 )2 , the differential form ω 3 is the corre-
sponding torsion-free metric connection one-form, and its Gaussian curvature
is K = −1 [35,38,78].
Now Definition 2.1 will be connected with (scalar) differential equations.
Independent variables will be denoted by x, t, while the dependent variable
will be denoted by u. Following standard usage [49], by a differential function
one means a smooth function which depends on x, t, u, and a finite number
of derivatives of u. Partial derivatives ∂ n+m u/∂xn ∂tm , n, m ≥ 0, will be also
denoted by uxn tm or even by ux...xt...t if deemed appropriate.
Definition 2.2. A scalar differential equation
Ξ(x, t, u, ux , . . . , uxn tm ) = 0 (2.2)
in two independent variables x, t is of pseudo-spherical type (or, it describes
pseudo-spherical surfaces) if there exist one-forms ω α , α = 1, 2, 3,
ω α = fα1 (x, t, u, . . . , uxr tp ) dx + fα2 (x, t, u, . . . , uxs tq ) dt, (2.3)
whose coefficients fαβ are differential functions, such that the one-forms ω α =
ω α (u(x, t)) satisfy the structure Eqs. (2.1) whenever u = u(x, t) is a solution
to Eq. (2.2).
Consequently, if Ξ = 0 describes pseudo-spherical surfaces with asso-
ciated one-forms ω α , and u(x, t) is a solution to Ξ = 0 such that (ω 1 ∧
ω 2 )(u(x, t)) = 0, Definition 2.1 implies that the domain of u(x, t) can be
equipped with a Riemannian metric of constant Gaussian curvature K = −1.
This is an intrinsic description. In other words, this theory does not provide one
with a prescription for constructing one-forms ω13 , ω23 such that ω 1 , ω 2 , ωij
(i, j = 1, 2, 3, ωij = −ωji , ω12 = ω 3 ) satisfy—whenever these one-forms are
pulled back to (open subsets of) R2 via local solutions—the structure equa-
tions of a two-dimensional Riemannian surface isometrically immersed in R3
(see, however, Theorem 5.5 for what can be proved along these lines).
A classical example of an equation describing pseudo-spherical surfaces
is the sine–Gordon equation, see [22,78]. Note however that, in contradistinc-
tion with this standard case, the dependent variable u of a general equation
of pseudo-spherical type does not have a geometric interpretation: geometry
appears only when one considers the associated one-forms ω α .
Vol. 60 (2011) Equations of Pseudo-Spherical Type 57

Remark 2.3. The structure Eqs. (2.1) could be also interpreted as Gauss–
Codazzi equations, as Gürses and Nutku did in [32]. This is an interesting
approach, for instance in [32] the authors use it to obtain a geometrically
motivated generalization of the Korteweg–de Vries equation. However, it would
appear that this point of view does not allow one to obtain either analytic cor-
respondences or natural geometric interpretations of quadratic pseudo-poten-
tials and conservation laws, as it does happen in the present theory. Other
applications of differential geometry to differential equations also exist in the
literature. The earliest references known to this author are [17,18,45]; a some-
what newer paper, also related to Prof. K. Tenenblat’s work, is [81]; a modern
treatise with extensive bibliography is [64]; and a recent work, also with a large
number of relevant references, is P. Olver’s [50].
The functions fαβ could, presumably, all depend only on x and t, but this
case would be uninteresting from the point of view of differential equations
and it is therefore excluded from further considerations.
Example. Burgers’ equation ut = uxx + uux is an equation of pseudo-spherical
type. Associated one-forms ω α are
     
1 1 β 1 1 2 2 λ
ω = u− dx + ux + u dt, ω = λ dx + u + β dt,
2 λ 2 2 2
and ω 3 = −ω 2 , in which λ is a nonzero parameter and β(x) is a solution to
the equation β 2 − λβx = 0.
The expression “PSS equation” will be sometimes utilized below in-
stead of the terms “equation of pseudo-spherical type” or “equation describing
pseudo-spherical surfaces”.
Definition 2.4. Let Ξ = 0 be a PSS equation with associated one-forms ω α , α =
1, 2, 3. A solution u(x, t) of Ξ = 0 is I-generic if (ω 3 ∧ ω 2 )(u(x, t)) = 0;
II-generic if (ω 1 ∧ ω 3 )(u(x, t)) = 0; and III-generic if (ω 1 ∧ ω 2 )(u(x, t)) = 0.
For instance, the traveling wave u(x, t) = 2 ex+t /(1 + ex+t ) is a III-
generic solution of Burgers equation, but it is not I-generic. Definition 2.4
allows one to refine the geometric interpretation of PSS equations given above.
In fact, the structure equations of a surface equipped with a metric of signature
(1, ),  = ±1, see for instance [35,78], imply:
Proposition 2.5. Let Ξ = 0 be a PSS equation with associated one-forms
ω α , α = 1, 2, 3; let u(x, t) be a solution of Ξ = 0, and set ω α = ω α (u(x, t)).
(a) If u(x, t) is I-generic, ω 2 and ω 3 determine a Lorentzian metric of
Gaussian curvature K = −1 on the domain of u(x, t), with connection
one-form given by ω 1 .
(b) If u(x, t) is II-generic, ω 1 and −ω 3 determine a Lorentzian metric of
Gaussian curvature K = −1 on the domain of u(x, t), with connection
one-form given by ω 2 .
58 E. G. Reyes Results. Math.

(c) If u(x, t) is III-generic, ω 1 and ω 2 determine a Riemannian metric of


Gaussian curvature K = −1 on the domain of u(x, t), with connection
one-form given by ω 3 .
The invariance properties of the structure Eqs. (2.1) are spelled out in
the following straightforward proposition:
Proposition 2.6. Let ω α , α = 1, 2, 3, be one-forms whose coefficients are dif-
ferential functions. Let u(x, t) be a smooth function, and set ω α = ω α (u(x, t)).
The structure Eqs. (2.1) are invariant under the transformations
 1 = ω 1 cos ρ + ω 2 sin ρ,
ω
 2 = −ω 1 sin ρ + ω 2 cos ρ,
ω  3 = ω 3 + dρ ;
ω (2.4)
1 1 3
 = ω cosh ρ − ω sinh ρ,
ω
 2 = ω 2 + dρ ,
ω  3 = −ω 1 sinh ρ + ω 3 cosh ρ ;
ω (2.5)
1 1 2 2 3
 = ω + dρ, ω
ω  = ω cosh ρ + ω sinh ρ,
 = ω sinh ρ + ω 3 cosh ρ ;
ω 3 2
(2.6)
in which ρ is any differential function and ρ = ρ(u(x, t)).
Transformation (2.4) corresponds to a rotation of the moving frame dual
to the coframe {ω 1 , ω 2 }; transformations (2.5) and (2.6) correspond to Lorentz
boosts.
Remark 2.7. Proposition 2.6 implies that the one-forms ω α associated to a
PSS equation Ξ = 0 are far from being unique. In terms of the geometry of
differential equations, see for instance [40], the one-forms ω α are defined on
the equation manifold S (∞) of Ξ = 0, and the definition of a PSS equation
asks that the structure Eqs. (2.1) are satisfied on S (∞) . This means that, for
instance, the form Ω3 = dω 3 − ω 1 ∧ ω 2 may depend on arbitrarily high deriv-
atives of the dependent variable u, well beyond the order of the equation, and
Ω3 = 0 only once one takes in account the equation Ξ = 0 and its differential
consequences. This observation turns out to be of importance for motivating
the characterization results to be reviewed in Sect. 2.2.
Example. In addition to the one-forms appearing in the previous example, one
can check that the Burgers equation ut = uxx +uux describes pseudo-spherical
surfaces with associated one-forms
  
1 1 1 1 1
ω1 = u − ux λ − ux 2 u + ux 2 λ + uxx dx + ux + u2
2 4 2 2 4

1 1 1 1
− ux λ u − ux 3 − ux 2 u2 + ux 2 λ u + uuxx + ux 2 + uxxx dt,
2 4 8 4
   
2 1 1 2 1 2
ω = λ + ux u − ux λ dx + λu + ux + ux u − ux λu dt,
2 2 2
Vol. 60 (2011) Equations of Pseudo-Spherical Type 59

  
3 1 2 1 2 1 1
ω = −λ + ux λ + ux u − ux λ − uxx dx + − λ u + ux λ u
4 2 2 2

1 1 1
+ ux 3 + ux 2 u2 − ux 2 λ u − uuxx − ux 2 − uxxx dt,
4 8 4
in which λ is a real parameter. Indeed, one obtains the following:
     
1 1 1 2
dω 1 − ω 3 ∧ ω 2 = Dx2 Ξ+ −λ+λux − uux Dx Ξ+ − ux Ξ dx ∧ dt
2 2 4
  
1 1
dω 2 − ω 1 ∧ ω 3 = −λ + u Dx Ξ + ux Ξ dx ∧ dt
2 2
   
1 1
dω 3 − ω 1 ∧ ω 2 = −Dx2 Ξ + λ − λux + uux Dx Ξ + u2x Ξ dx ∧ dt,
2 4
in which Ξ = −ut + uxx + uux , so that the structure Eqs. (2.1) are satisfied
only on solutions to Burgers’equation.
The following proposition appears in the classical Chern–Tenenblat pa-
per [12]. It is the key to the application of the foregoing constructions to the
theory of integrability:
 
Proposition 2.8. Given a local coframe σ 1 , σ 2 and corresponding connection
1 2
one-form σ 3 on a Riemannian surface M , there exists a new coframe {θ , θ }
3
and new connection one-form θ on M satisfying
1 2 2 1 3 2
dθ = 0, dθ = θ ∧ θ , and θ + θ = 0, (2.7)
if and only if the surface M is pseudo-spherical.
 
Proof. Assume that the orthonormal frames dual to the coframes σ 1 , σ 2 and
1 2 α
{θ , θ } possess the same orientation. Then, the one-forms σ α and θ must be
connected by means of a rotation and a gauge transformation, that is,
1 2 3
θ = σ 1 cos ρ + σ 2 sin ρ, θ = −σ 1 sin ρ + σ 2 cos ρ, θ = σ 3 + dρ. (2.8)
1 2 3
It follows that θ , θ , θ satisfying (2.7) exist if and only if the Pfaffian system
σ 3 + dρ − σ 1 sin ρ + σ 2 cos ρ = 0 (2.9)
on the space of coordinates (x, t, ρ) is completely integrable for ρ(x, t), and
this happens if and only if M is pseudo-spherical. 

If M is a pseudo-spherical surface and {v1 , v2 } is the orthonormal mov-


ing frame dual to the moving coframe {θ1 , θ2 }, then the integral curves of the
vector fields v1 and v2 are, respectively, geodesics and horocycles of M (see
[12,78]). It follows, therefore, that the function ρ satisfying Eq. (2.9) is pre-
cisely the angle connecting an arbitrary orthonormal framing of M with an
orthonormal framing along geodesics.
60 E. G. Reyes Results. Math.

Remark 2.9. The notion of a scalar differential differential equation of pseudo-


spherical type generalizes mutatis mutandis to systems of equations, see the
paper by Ding and Tenenblat [20]. For instance, it is shown in [20] that the
nonlinear Schrodinger equation
i qt + qxx − 2|q|2 q = 0,
written as a system of equations for real-valued functions u and v defined via
q(x, t) = u(x, t) + i v(x, t), describes pseudo-spherical surfaces.
2.2. Characterization Results
The term “characterization results” refers to a set of theorems on the form a
differential equation must have in order to describe pseudo-spherical surfaces.
As a rule, these theorems are quite technical (see [12] and the later papers
[20,37,38,51,54,57]) but, as stated in Sect. 1, they have proven to be very
useful in classifying PSS equations [20,30,52,53] and in comparing different
types of integrability [25,54].
The starting point for these theorems is the assumption that the differ-
ential equation one is interested in is equivalent to the structure Eqs. (2.1),
so that the phenomenon explained in Remark 2.7 does not occur. One way to
formalize this idea (for evolution equations) is due to Kamran and Tenenblat
[38]:
Given a kth order scalar differential equation ut = F (x, t, u, . . . , uk ),
consider the differential ideal IF generated by the two-forms
du ∧ dx + F (x, t, u, . . . , uxk )dx ∧ dt,
duxi ∧ dt − uxi+1 dx ∧ dt, 1 ≤ i ≤ k − 1, (2.10)
on the reduced kth order jet space with coordinates x, t, u, ux1 , . . . , uxk , so that
local solutions to the evolution equation correspond to integral submanifolds
of the exterior differential system {IF , dx ∧ dt}.
Definition 2.10. A scalar differential equation ut = F (x, t, u, . . . , uk ) is of
strictly pseudo-spherical type if there exist one-forms ω α = fα1 dx + fα2 dt
whose coefficients are differential functions fαβ depending at most on deriva-
tives of order k, such that the two-forms
Ω1 = dω 1 − ω 3 ∧ ω 2 , Ω2 = dω 2 − ω 1 ∧ ω 3 , Ω3 = dω 3 − ω 1 ∧ ω 2 , (2.11)
generate the equation ideal IF .
Thus, if ut = F (x, t, u, . . . , uxk ) is of strictly pseudo-spherical type, the
ideal IF is algebraically equivalent to a system of differential forms satisfy-
ing the pseudo-spherical structure Eqs. (2.1) whenever u(x, t) is a solution
of the equation ut = F . The differential ideal generated by the two-forms
{Ω1 , Ω2 , Ω3 } will be denoted by JF . Analogous definitions (and corresponding
characterization results) can be stated for other types of equations, see for
instance [20,37,51].
Vol. 60 (2011) Equations of Pseudo-Spherical Type 61

The characterization theorems considered here are valid only for evolu-
tion equations of strictly pseudo-spherical type, see [12,38,54,57]. They are
consequences of the fact that Definition 2.10 implies that the dependence of
its associated functions fαβ on derivatives of u is very restricted, and that
these functions satisfy several identities which allow one to “reconstruct” the
right hand side of the equation ut = F :
Lemma 2.11. Necessary and sufficient conditions for the kth order equation
ut = F to be of strictly pseudo-spherical type are the conjunction of (a) The
functions fαβ satisfy the constraints
2 2 2
fα1,uxa = 0; fα2,uxk = 0; f11,u + f21,u + f31,u = 0, (2.12)
in which a ≥ 1 and α = 1, 2, 3; and (b) The functions F and fαβ satisfy the
three identities

k−1
− f11,u F + uxi+1 f12,uxi + f21 f32 − f31 f22 + f12,x − f11,t = 0, (2.13)
i=0

k−1
−f21,u F + uxi+1 f22,uxi + f12 f31 − f11 f32 + f22,x − f21,t = 0, (2.14)
i=0

k−1
−f31,u F + uxi+1 f32,uxi + f21 f12 − f11 f22 + f32,x − f31,t = 0. (2.15)
i=0

This lemma is a direct consequence of the equality of the ideals IF and


JF , see [54]. Using Lemma 2.11 one can prove, [54], for instance, the following:
Theorem 2.12. Let fαβ , 1 ≤ α ≤ 3, 1 ≤ β ≤ 2 be differential functions sat-
isfying (2.12)–(2.15), and let f21 = λ be a non–zero real parameter. Suppose
HL = 0, in which H = f11 f11,u − f31 f31,u and L = f11 f31,u − f31 f11,u , and
set T := f11,t f31,u − f11,u f31,t . The equation ut = F (x, t, u, . . . , uxk ) is strictly
pseudo-spherical with associated one-forms ω α = fα1 dx + fα2 dt if, and only if,
(a) The function F is given in terms of fαβ by the formula

1 1  2 
k−2
k−1
2
F = uxi+1 (Dx f22 )uxi + f31 − f11 uxi+1 Ai + T
L i=1 HL i=0

f11
−f31,u f12,x + f11,u f32,x + (ux f32,u + f32,x )
L
f31 λ 1
− (ux f12,u + f12,x ) + Dx f22 − (f31,t f11 − f11,t f31 );
L H L
(b) The uxj -derivatives of f12 and f32 , 1 ≤ j ≤ k − 1, are given in terms
of f11 and f31 by the formulae fa2,uxj = −(1/L){fa1 Aj − fa1,u (Dx f22 )uxj },
a = 1, 3;
(c) The u-derivatives of f12 and f32 satisfy the equation f31,u f12,u −
f11,u f32,u + A0 = 0; and
62 E. G. Reyes Results. Math.

(d) The functions f12 and f32 satisfy the constraints



k−2
fa1 f22 1 −fa1  i
fa2 = + uxi+1 A + T + fb1,u Dx f22
λ H λ i=0
fa1
+ (f31,u f12,x − f11,u f32,x ),
λH
in which a = 1, 3; b = 3 if a = 1, and b = 1 if a = 3.
2
The functions Aj are defined recursively as follows: set M = f31,u −
2
f11,u ; P = f11,u f31,uu − f31,u f11,uu ; Q = f11,u,x f31,u − f31,u,x f11,u ; and R =
f11,u,x f31 − f31,u,x f11 . Then, Ak−1 := 0 and, for 0 ≤ j ≤ k − 2,
1
Aj = −Dx Aj+1 + (ux Lu + λH − R) Aj+1
L
1
− (ux P − λM − Q) (Dx f22 )u j+1 + f22,uxj+1 H.
L x

Thus, if one is given an evolution equation ut = F , one can compare it


with the formula appearing in (a) of Theorem 2.12 and, using (b), (c) and
(d) reconstruct the one-forms ω α associated with ut = F . If this is possible
to do, the given equation is of strictly pseudo-spherical type. It may hap-
pen, however, that for a given evolution equation ut = F items (a)–(d) of
Theorem 2.12 are inconsistent. Then, one must conclude that ut = F is not
of strictly pseudo-spherical type with one-forms satisfying the hypotheses of
this theorem. One could then check a different characterization. The results
appearing in [12,54] are exhaustive with respect to evolution equations and
associated one-forms satisfying the a priori constraint f21 = λ (a condition
dictated by inverse scattering, see [4,12]), and so one is always able to decide
whether a given evolution equation is in the class of strictly pseudo-spherical
equations or not. For instance, [54], the equation
ut = (ux + δ)3 uxxx , δ ∈ R,
one of the integrable equations appearing in the classification of integrable
evolution equations of Mikhailov et al. [47], is not of strictly pseudo-spherical
type.
The first application of the characterization results of [12,54] is to the
classification of equations of strictly pseudo-spherical type: one starts from
theorems such as Theorem 2.12 above, one fixes the order of the equations one
is interested in, and asks, for example, that the PSS equation to be constructed
be independent of the parameter λ. Then, working with (a)–(d) above one can
eventually fix completely the form the equation ut = F ! For example, the
following theorem is the main result of [53]:
Theorem 2.13. Let ut = uxxx +G(u, ux , uxx ) be a differential equation describ-
ing pseudo-spherical surfaces with associated one-forms (2.3) satisfying ω 2 =
λdx+f22 dt, in which λ is a real parameter. Then, G is independent of λ if and
Vol. 60 (2011) Equations of Pseudo-Spherical Type 63

only if, up to a change of the dependent variable, the aforementioned equation


is one of the following:

ut = uxxx + (αu3 + βu2 + γu)x ,


ut = uxxx + βuxx − 3α(uux )x + αuux (3αu − 2β),
ut = (ux − l)xx − [(δu + μ)(ux − l)]x ,
ut = uxxx + αuxx + βux + γu + η,

where α, β, δ = 0, γ, μ, η are real constants, and l is a differentiable function


of u.

Other classification results appear in [12,20,52] and in the recent (and


technically very demanding) paper by Gomes-Neto [30], in which he classifies
completely a large class of fifth order evolution equations of strictly pseudo-
spherical type making use of the characterization results obtained by Kamran
and Tenenblat [38].
Another application is to the theory of integrability for partial differential
equations: much effort has been directed toward the classification of evolution
equations which are formally integrable, that is, of equations which admit a
“formal symmetry of infinite rank” [47,49,76] or recursion operator [49]. It is
not possible to review this notion here without veering away from the main
topic of this paper, but one can state that formal integrability is the most
general notion of integrability for evolutionary partial differential equations
available today. The reader is referred to [47,49,76] and also [25] for details.
An important problem is to investigate whether formally integrable equations
possess characteristics considered to be naturally associated with integrable
equations. For instance, an integrable equation should be the integrability
condition of an overdetermined linear problem, so that, at least in principle,
the equation may be amenable of study via scattering/inverse scattering. As
it will be seen in the following section, equations of pseudo-spherical type are
the integrability condition of an sl(2, R)-valued linear problem. Thus, in order
to check whether a formally integrable equation ut = F admits an associated
linear problem, one naturally asks whether ut = F is of pseudo-spherical type,
and this can be checked using the characterization results. The following result
combines theorems proven in [47,76] and [25,54]:

Theorem 2.14. (a) Every formally integrable second order evolution equation
is equivalent, under a contact transformation of the form

t = χ(t), x = φ(t, x, u, ux ), u = ψ(t, x, u, ux ),

to one of the following:


64 E. G. Reyes Results. Math.

ut = uxx + h(x, t)u, (2.16)


ut = uxx + uux + g(x, t), (2.17)

ut = Dx ux u−2 , (2.18)

ut = Dx ux u−2 − x , (2.19)

ut = Dx ux u−2 + x2 u + xu. (2.20)
(b) Equations (2.16)–(2.20) are of strictly pseudo-spherical type.
(c) Every second order evolution equation ut = F (x, t, u, ux , uxx ) which pos-
sesses a formal symmetry of infinite rank is of pseudo-spherical type.
Other results along the same lines are possible. For instance, recall that
a Monge–Ampère equation is a second order equation of the form

Auxx + 2Buxt + Cutt + D + E uxx utt − u2xt = 0,
where A, B, C, D and E are functions of x, t, u, ux , ut .
Theorem 2.15. Any Monge–Ampère second order equation which is Darboux-
integrable is of pseudo-spherical type.
Proof. Darboux integrability is nicely explained in the textbook by Ivey and
Landsberg [35], and also in the lecture notes by Bryant et al. [7]. It is known
(see Theorem 2.1 in [7]) that any Monge–Ampère equation which is Darboux-
integrable is equivalent to the standard wave equation uxt = 0 under a contact
transformation. Theorem 2.15 follows because this equation is of pseudo-spher-
ical type: it is enough to define associated one-forms ω α as
 
ω 1 = −f 2 ux + fx + ux dx + −f 2 C + ft + C dt
ω 2 = −2ux f dx − 2Cf dt
 
ω 3 = −f 2 ux + fx − ux dx + −f 2 C + ft − C dt,
in which C = 0 and f = f (x, t) is a function depending explicitly on x, t. Note
that ω 1 ∧ ω 2 = −2f (Cfx − ux ft )dx ∧ dt, so that generically the non-triviality
condition ω 1 ∧ ω 2 ≡ 0 holds. 
It is an open problem to investigate the validity of the converse to Theorem
3 (or 4). Some further remarks on this point are in Sect. 6.2.
2.3. Hierarchies of Pseudo-Spherical Type
In this subsection a differential function will be a real-valued smooth function
depending on a finite number of independent variables x, t, τ1 , τ2 , . . ., the
variable u, and a finite number of x-derivatives of u. The independent variable
t will be also denoted by τ0 . Assume that uτi = Fi is a countable sequence of
evolution equations and set
∞ ∞
∂ ∂ ∂  j ∂
Dx = + uxj+1 and Dτi = + D Fi . (2.21)
∂x j=0 ∂uxj ∂τi j=0 x ∂uxj
Vol. 60 (2011) Equations of Pseudo-Spherical Type 65

Definition 2.16. Let uτi = Fi be a countable number of evolutionary equa-


tions. They are a hierarchy of equations describing pseudo-spherical surfaces
(or, uτi = Fi is a hierarchy of pseudo-spherical type) if there exist differential
[n]
functions fαβ and hαj such that for each n ≥ 0, the one-forms Θα given by

n
[n]
Θα = fα1 d x + fα2 d t + hαk d τk (2.22)
k=1

satisfy the equations


[n] [n] [n] [n] [n] [n] [n] [n] [n]
dH Θ1 = Θ3 ∧ Θ2 , dH Θ2 = Θ1 ∧ Θ3 , dH Θ3 = Θ1 ∧ Θ2 , (2.23)
[n]
in which d
H Θα is computed by means of dH (dx) = dH (dτi ) = 0 and dH g =
n
Dx gdx + k=0 Dτk g dτk for any differential function g.
Definition 2.17. A local smooth solution of a hierarchy of pseudo-spherical
type uτi = Fi is a sequence {u[n] (x, τ0 , . . . , τn ) : n ≥ 0} of smooth functions
u[n] : V [n] ⊂ Rn+2 → R such that for each n ≥ 0 the following two conditions
hold:
(a) u[n] is a local smooth solution of the equations uτi = Fi , i = 0, . . . , n,
(b) u[n+1] |V [n] = u[n] .
A geometric content for Definitions 2.16 and 2.17 is provided by the
following theorem [59]:
Theorem 2.18. Let uτi = Fi be a hierarchy of pseudo-spherical type with asso-
ciated one-forms

n
[n]
Θα = fα1 dx + fα2 dt + hαk dτk , n ≥ 0, (2.24)
k=1

and let the sequence of real-valued smooth functions {u[n] : n ≥ 0} be an arbi-


trary solution of the hierarchy uτi = Fi . For each n ≥ 0, let V [n] ⊆ Rn+2 be
(an open subset of) the domain of the function u[n] .
[0] [0]
1. Assume that n = 0. If Θ1 ∧ Θ2 (u[0] (x, t)) = 0, the tensor
   
[0] [0]
ds2 = Θ1 u[0] (x, t) ⊗ Θ1 u[0] (x, t)
   
[0] [0]
+Θ2 u[0] (x, t) ⊗ Θ2 u[0] (x, t)
defines a Riemannian metric of constant Gaussian curvature K = −1 on
[0] 
V [0] , and the one-form Θ3 u[0] (x, t) is the corresponding Levi–Civita
connection one-form.
2. Assume that n ≥ 1. Equip V [n] with a flat pseudo-Riemannian metric of
index s, and let ι : D ⊆ R2 → V [n] be a smooth function from an open
[n] 
set D ⊆ R2 into V [n] . Suppose that the forms Θα u[n] (x, τ0 , . . . , τn )
satisfy the conditions
66 E. G. Reyes Results. Math.

  
ι∗ Θα[n]
u[n] (x, τ0 , . . . , τn ) = 0;
   (2.25)
[n] [n]
ι∗ Θ1 ∧ Θ2 u[n] (x, τ0 , . . . , τn ) = 0.

(a) The set D can be equipped with the structure of a pseudo-spherical


surface: the pair of one-forms
     
[n] [n]
ι∗ Θ1 u[n] (x, τ0 , . . . , τn ) and ι∗ Θ2 u[n] (x, τ0 , . . . , τn )
 
[n] 
are a moving coframe on D, and ι∗ Θ3 u[n] (x, τ0 , . . . , τn ) is the
corresponding Levi–Civita connection one-form.
(b) If the index s is equal to n, there exists an isometric immersion of
the pseudo-spherical surface (D, ds2 ) constructed in (a), in which
  2   2
[n] [n]
ds2 = ι∗ Θ1 u[n] (x, τ0 , . . . , τn ) +ι∗ Θ2 u[n] (x, τ0 , . . . , τn ) ,

into the flat pseudo-Riemannian manifold V [n] . Moreover, the


normal bundle of the immersion is flat.
Example. It follows from the seminal paper by Chern and Peng [11] that the
Korteweg–de Vries hierarchy is of pseudo-spherical type. Indeed, associated
functions are
f11 = 1 − u; f12 = λux − uxx − 2u2 + 2u − λ2 u + λ2 ; (2.26)
3
f21 = λ; f22 = λ + 2λu − 2ux ; (2.27)
2 2 2
f31 = −1 − u; f32 = λux − uxx − λ u − 2u − λ − 2u; (2.28)
and
1 1 (i+1)
h1i = λBx(i+1) − Bxx − u B (i+1) + B (i+1) ; (2.29)
2 2
h2i = λB (i+1) − Bx(i+1) ; (2.30)
1 1 (i+1)
h3i = λBx(i+1) − Bxx − u B (i+1) − B (i+1) ; (2.31)
2 2
in which

i
B (i) = Bj λ2(i−j),
j=0

and the functions Bj are defined recursively by means of


B0,x = 0, (2.32)
Bj+1,x = Bj,xxx + 4uBj,x + 2ux Bj , j ≥ 0. (2.33)
The functions Fi , i ≥ 0, are given by
1 1
Fi = Bi+1,xxx + ux Bi+1 + 2 u Bi+1,x = Bi+2,x . (2.34)
2 2
Vol. 60 (2011) Equations of Pseudo-Spherical Type 67

For instance, if B0 = 1 and all integration constants are set to zero, the equa-
tion uτ0 = F0 is the standard KdV equation ut = uxxx + 6uux , while uτ1 = F1
and uτ2 = F2 are, respectively,
uτ1 = uxxxxx + 20 ux uxx + 10 uuxxx + 30 u2 ux ;
uτ2 = uxxxxxxx + 70 uxx uxxx + 42 ux uxxxx + 14 u uxxxxx + 70 u3x
+280 u ux uxx + 70 u2 uxxx + 140 u3 ux .
An elementary solution to the KdV hierarchy is the sequence of functions
{u[n] : n ≥ 0} given by

u[n] (x, t, τ1 , . . . , τn ) = 2 sech2 x + 4t + 16τ1 + 64τ2 + · · · + 4n+1 τn .
(2.35)
Now, there are three facts one needs to state and prove for a hierarchy of
pseudo-spherical type, as advanced in Sect. 1: Each equation uτi = Fi of the
hierarchy should describe pseudo-spherical surfaces; each such equation should
be the compatibility condition of linear problems of the form vx = X v, vτn =
Tn v; each pair of equations uτi = Fi , uτj = Fj , i = j, should determine
commuting flows. The first two properties are quite easy to establish [59]:
Proposition 2.19. Let uτi = Fi be a hierarchy of pseudo-spherical type with
associated one-forms

n
[n]
Θα = fα1 dx + fα2 dt + hαi dτi , n ≥ 0.
i=1

The equations uτi = Fi describe pseudo-spherical surfaces with associated one-


forms
ω0α = fα1 dx + fα2 dt (if i = 0) and ωiα = fα1 dx + hαi dτi (if i ≥ 1),
(2.36)
and moreover, the one-forms σjα = fα2 dt + hαj dτj (j fixed) and σij α
= hαi dτi +
hαj dτj (i, j fixed, i = j) satisfy the structure Eqs. (2.1) of a pseudo-spherical
surface.
Proposition 2.20. Let uτi = Fi be a hierarchy of pseudo-spherical type. There
exist sl(2, R)-valued functions X and Ti such that for each i ≥ 0, uτi = Fi is
the integrability condition of the linear problem
vx = X v, vτi = Ti v.
This last proposition is an immediate corollary of Proposition 2.19, as
it can be seen from Eq. (3.2) in Sect. 3. To show that hierarchies of pseudo-
spherical type generate an infinite number of pairwise commuting flows is
more involved. Motivated by soliton theory as it appears for instance in Prof.
L. Dickey’s treatise [19], one first defines hierarchies of strictly pseudo-spherical
68 E. G. Reyes Results. Math.

type, in analogy with the single equation case reviewed in the previous sub-
section. Then one can prove a theorem on commuting flows, see Theorem 2.25
below:
Consider a countable number of evolution equations uτi = Fi , and assume
that the ith equation is of order ki . For each n ≥ 0, let J (n) be a manifold with
coordinates (x, t, τ1 , . . . , τn , u, ux , . . . , uxM (n) ), where M (n) is the maximum of
the orders kj , 0 ≤ j ≤ n, and let I (n) be the differential ideal generated by
the two-forms

n
du ∧ dx + F0 dx ∧ dt + Fi dx ∧ dτi ; (2.37)
i=1

n
duxk ∧ dt − uxk+1 dx ∧ dt − Dxk Fj dτj ∧ dt; 0 ≤ k ≤ k0 − 1; (2.38)
j=1

n
duxk ∧ dτj − uxk+1 dx ∧ dτj − Dxk F0 dt ∧ dτj − Dxk Fl dτl ∧ dτj , (2.39)
l=1

in which for each j = 1, . . . , n the index k takes the values k = 0, . . . , kj − 1.

Definition 2.21. A countable collection of evolution equations uτi = Fi is a


hierarchy of strictly pseudo-spherical type if there exist differential functions
[n] [n]
fαβ and hαj such that for every n ≥ 0, the two-forms Ωi , given by Ω1 =
[n] [n] [n] [n] [n] [n] [n] [n] [n] [n] [n]
dΘ1 − Θ3 ∧ Θ2 , Ω2 = dΘ2 − Θ1 ∧ Θ3 , Ω3 = dΘ3 − Θ1 ∧ Θ2 , and

n
[n]
Θα = fα1 dx + fα2 dt + hαj dτj ,
j=1

generate the ideal I (n) .

Instead of Lemma 2.11, one has the following result, see [59]:

Proposition 2.22. Consider a sequence of equations uτi = Fi of order ki and


for each n ≥ 0 let M (n) be the maximum of the orders kj , 0 ≤ j ≤ n. Neces-
sary and sufficient conditions for the equations uτi = Fi to define a hierarchy
of strictly pseudo-spherical type with associated functions fαβ and hαj are the
conjunction of:
2 2 2
1. The functions fα1 satisfy f11,u + f21,u + f31,u = 0 ;
2. For each n ≥ 0, the functions fαβ and hαj satisfy fα1,uxk = 0, 1 ≤ k ≤
M (n); fα2,uxk = 0, k0 ≤ k ≤ M (n); hαi,uxk = 0, ki ≤ k ≤ M (n);
3. The functions Fi , fαβ , and hαj satisfy, for all i, j ≥ 0,

⎨ −(f11,t + f11,u F0 ) + Dx f12 = f31 f22 − f32 f21
−(f21,t + f21,u F0 ) + Dx f22 = f11 f32 − f12 f31 (2.40)

−(f31,t + f31,u F0 ) + Dx f32 = f11 f22 − f12 f21 ;
Vol. 60 (2011) Equations of Pseudo-Spherical Type 69

⎨ −(f11,τi + f11,u Fi ) + Dx h1i = f31 h2i − h3i f21
−(f21,τi + f21,u Fi ) + Dx h2i = f11 h3i − h1i f31 (2.41)

−(f31,τi + f31,u Fi ) + Dx h3i = f11 h2i − h1i f21 ;

⎨ −Dτi f12 + Dt h1i = f32 h2i − h3i f22
−Dτi f22 + Dt h2i = f12 h3i − h1i f32 (2.42)

−Dτi f32 + Dt h3i = f12 h2i − h1i f22 ;

⎨ −Dτj h1i + Dτi h1j = h3i h2j − h3j h2i
−Dτj h2i + Dτi h2j = h1i h3j − h1j h3i (2.43)

−Dτj h3i + Dτi h3j = h1i h2j − h1j h2i .
Corollary 2.23. Let uτi = Fi be a hierarchy of strictly pseudo-spherical type
with associated functions fαβ and hαj . Then, the equation ut = F0 is strictly
pseudo-spherical with associated one-forms ω0α = fα1 dx + fα2 dt, and uτj = Fj
is strictly pseudo-spherical with associated one forms ωjα = fα1 dx + hαj dτj .

Now one shows that a hierarchy uτi = Fi of strictly pseudo-spherical type


generates a family of pairwise commuting flows or, equivalently, see [19,40,49],
that the function Fp is a generalized symmetry of the equation uτq = Fq for
all p, q ≥ 0 . The notion of a generalized symmetry is classical, so it will be
recalled here only to fix notation. For any smooth function f (x, t, u, . . . , uxk )
define the operator f∗ (the formal linearization of f ) by means of [47,49]

k
∂f
f∗ = Dxi , (2.44)
i=0
∂u xi

and consider Dt defined as in (2.21). A function G(x, t, u, . . . , uxk ) is a general-


ized symmetry of ut = F if and only if the equation Dt G = F∗ G is identically
satisfied. Equivalently, [49], an evolution equation uτ = G is a generalized
symmetry of ut = F if the equation
∂G
+ F∗ G − G∗ F = 0 (2.45)
∂t
holds whenever u(x, t) is a solution of ut = F . If Ft = Gt = 0, Eq. (2.45)
means that, at least formally, the flows of ut = F and uτ = G commute.

Lemma 2.24. Let ut = F (x, t, u, . . . , uxm ) be an evolution equation of strictly


pseudo-spherical type with associated one-forms ω α = fα1 dx + fα2 dt. Let G be
a smooth function depending on x, t, u and a finite number of derivatives of u,
and let u(x, t) be a local solution of ut = F . Consider the deformed one-forms
ω α (u(x, t)) + τ Λα (u(x, t)), in which Λα (u(x, t)) is given by

Λα (u(x, t)) = fα1,u (u(x, t)) G(u(x, t)) dx



m−1
∂ i G(u(x, t))
+ fα2,uxi (u(x, t)) dt. (2.46)
i=0
∂xi
70 E. G. Reyes Results. Math.

These forms satisfy the structure equations of a pseudo-spherical surface up to


terms of order τ 2 if and only if the function G is a generalized symmetry of
the equation ut = F .
Lemma 2.24 is proven in [56]. Motivated by the above discussion on gen-
eralized symmetries, it will be assumed that neither the functions Fj , nor the
associated functions fαβ and hαj , depend explicitly on the independent “time”
variables τ0 , τ1 , τ2 , . . ..
Theorem 2.25. Let uτi = Fi be a hierarchy of strictly pseudo-spherical type
with associated one-forms

n
[n]
Θα = fα1 dx + fα2 dt + hαk dτk , n ≥ 0. (2.47)
k=1

The function Fj is a generalized symmetry of the equation uτi = Fi for all


i, j ≥ 0.
Proof. (Sketch) Corollary 2.23 says that the equations ut = F and uτj = Fj
are strictly pseudo-spherical with associated one forms ω0α = fα1 dx + fα2 dt
and ωjα = fα1 dx + hαj dτj respectively, and so one can apply the last lemma.
For instance, the equations uτi = Fi i > 0, determine generalized symme-
tries of the equation ut = F : Consider the deformations of the one-forms ω0α
induced by u → u + τi Fi , and set Λα0 = fα1,τi dx + fα2,τi dt, in which fαβ,τi =
Dτi fαβ . These one-forms are of the type (2.46) because of Proposition 2.22.
Now, it can be checked that the deformed one-forms ω0α +τi Λα 0 describe pseudo-
spherical surfaces to first order in τi if and only if the equations
d Λ 1 = ω 3 ∧ Λ 2 + Λ3 ∧ ω 2 , d Λ 2 = ω 1 ∧ Λ 3 + Λ1 ∧ ω 3 ,
d Λ 3 = ω 1 ∧ Λ 2 + Λ1 ∧ ω 2 , (2.48)
hold whenever u(x, t) is a solution of the equation ut = F , in which the exte-
rior derivative is taken in (x, t) space. The first equation of (2.48) holds if and
only if
− f11,τi t + f12,τi x = f31 f22,τi − f32 f21,τi + f31,τi f22 − f21 f32,τi . (2.49)
But, because of Eqs. (2.41) and (2.42), Eq. (2.49) is equivalent to
−h2i (−f31,t + f32,x ) + h3i (−f21,t + f22,x )
= −h2i (f22 f11 − f21 f12 ) + h3i (f32 f11 − f31 f12 ),
and this equation holds whenever u(x, t) is a solution of ut = F , since ut = F
describes pseudo-spherical surfaces with associated one-forms ω0α . The other
two equations of (2.48) are treated in the same fashion, and analogous com-
putations show that the equations uτj = Fj determine generalized symmetries
of the equations uτi = Fi , i = j. 
Vol. 60 (2011) Equations of Pseudo-Spherical Type 71

3. Conservation Laws, Pseudo-Potentials and Associated Linear


Problems
As pointed out in Sect. 1, the existence of associated linear problems,
conservation laws, and classical Bäcklund/Darboux transformations of PSS
equations, can be studied via the geometry of pseudo-spherical surfaces [4,12,
54–56,70,78]:

3.1. Associated Linear Problems


Perhaps the most basic observation connecting the theory of PSS equations
with integrable systems is the fact that if Ξ(x, t, u, . . . ) = 0 describes pseudo-
spherical surfaces with associated one-forms ω α , it is the integrability condition
of a sl(2, R)-valued linear problem. Indeed, the sl(2, R)-valued linear problem
∂v ∂v
= Xv, = T v, (3.1)
∂x ∂t
in which X and T are determined by the sl(2, R)-valued one-form
 
1 ω2 ω1 − ω3
Ω = Xdx + T dt = , (3.2)
2 ω1 + ω3 −ω 2
is integrable whenever u(x, t) is a solution to Ξ = 0. In other words, the pseudo-
spherical structure Eqs. (2.1) imply that the “zero curvature condition”
∂X ∂T
− + [X, T ] = 0 (3.3)
∂t ∂x
is identically satisfied whenever u(x, t) is a solution of Ξ = 0. As already
observed, this fact means that one may hope to study PSS equations via
scattering/inverse scattering techniques [2,3].
An interesting example of the foregoing comments is provided by the
equation
{ut − [αg(u) + β]ux }x = g  (u), (3.4)
in which g(u) satisfies the constraint
g  + μg = −θ
and μ, θ, α, β are real numbers. Rabelo [51] proved that Eq. (3.4) is of pseudo-
spherical type with associated one-forms
ω 1 = ζux dx + ζ(αg + β)ux dt, (3.5)
 2 
ζ g−θ
ω 2 = λ dx + + βλ dt, (3.6)
λ
ω 3 = (ζg  /λ) dt, (3.7)
2 2
in which ζ = αλ − μ and λ is a real parameter. Beals et al. [4] used the linear
problem (3.1) determined by (3.2) and (3.5)–(3.7) to solve (3.4), by adapting
to this case the rigorous scattering/inverse scattering method developed by
72 E. G. Reyes Results. Math.

Beals and Coifman [2,3]. Much more recently, it has been realized that a par-
ticular instance of Eq. (3.4) describes the propagation of ultra-short optical
pulses in nonlinear media! In fact, the short pulse equation
1
uxt = u + uu2x + u2 uxx , (3.8)
2
derived from physical principles by Schäfer and Wayne [73], corresponds to
Rabelo’s equation (3.4) with g(u) = (1/2) u2 , α = 1, β = 0, μ = 0, and
θ = −1. Some integrability properties of Eq. (3.8) have been also discussed by
Sakovich and Sakovich [66–68] independently of [4,51].
Remark 3.1. It may happen that the one-forms ω α associated to a given PSS
equation Ξ = 0 give rise to an associated linear problem which cannot be used
for obtaining solutions via scattering/inverse scattering. This is the case with
the third equation appearing in the Rabelo–Tenenblat classification reviewed
in Theorem 2.13, see [53]. In a related vein, it may happen that the associated
linear problem (3.2) is “trivial”, in some sense, even if the one-forms ω α deter-
mine non-trivial pseudo-spherical metrics. To decide when a linear problem is
trivial is a difficult problem which has been considered by Marvan and Sako-
vich, see [44,65], and references therein. A related discussion appears also in
the recent paper [31].
Remark 3.2. One can interpret Eqs. (3.1)–(3.3) in terms of connections [18,70]:
Let u(x, t) be a solution to Ξ = 0, let π : U × SL(2, R) → U , in which U is (an
open subset of) the domain of the solution u(x, t), be a principal fiber bundle,
and consider the sl(2, R)-valued one-form Ω(u(x, t)) determined by (3.2). This
differential form can be thought of as a flat connection one-form on the bun-
dle π, and the solution v(x, t) to (3.1) as a covariantly constant section on a
vector bundle U × V → U associated to π, in which V is some two-dimensional
vector space. This interpretation has been of use in the study of transforma-
tions of solutions of PSS equations, see [57], and it has been further refined by
Shchepetilov [75].
3.2. Pseudo-potentials and Conservation Laws
Definition 3.3. Let Ξ = 0 be a scalar partial differential equation in two inde-
pendent variables. A function Γ is a pseudo-potential for Ξ = 0 if there exist
smooth functions f and g depending on x, t, u, a finite number of derivatives
of u and Γ, such that the one-form
ΩΓ = dΓ − (f dx + gdt)
satisfies
dΩΓ = 0 mod ΩΓ
whenever u(x, t) is a solution to Ξ = 0.
In other words, Γ is a pseudo-potential for Ξ = 0 if and only if ft = gx
whenever Γx = f, Γt = g and u(x, t) is a solution to Ξ = 0. Of course, if
Vol. 60 (2011) Equations of Pseudo-Spherical Type 73

the functions f and g above do not depend on Γ at all, ΩΓ is a local con-


servation law for Ξ = 0 with potential Γ. Definition 3.3 is motivated by the
seminal work of Wahlquist and Eastbrook [80] on quadratic pseudo-potentials
of the KdV equation. Analogous definitions have been given by Sasaki [71]
and Molino [48]. Moreover, it has been recently realized by the author that a
more general definition of pseudo-potential has been put forward by Bryant
and Griffiths, see [6] and the textbook [35].
Now, let Ξ = 0 be an equation of pseudo-spherical type and consider the
Pfaffian system
ω 3 + dρ − ω 1 sin ρ + ω 2 cos ρ = 0 (3.9)
arising from Eq. (2.9) of Proposition 2.8. This proposition implies that (3.9)
is completely integrable for ρ whenever u(x, t) is a solution to the equation
Ξ = 0, and therefore the function ρ is a pseudo-potential for Ξ = 0.
This pseudo-potential has a clear geometric interpretation, as stated in
Proposition 2.8. Moreover, Eqs. (2.7) and (2.8) imply that for each solution
u(x, t) and corresponding solution ρ(x, t) of (3.9), the one-form θ1 = ω 1 cos ρ +
ω 2 sin ρ is closed. Thus, as observed by Chern and Tenenblat in [12], if the
functions fαβ can be expanded as power series in a parameter, so can ρ(x, t)
and θ1 , and therefore, in principle, equations describing (one-parameter fam-
ilies of) pseudo-spherical surfaces possess an infinite number of conservation
laws! The reader is referred to [10], and also to the later paper [82], for compu-
tations of conservation laws using precisely this idea. The following elementary
proposition (see [55]) connects these remarks with standard facts in the theory
of integrable systems:
Proposition 3.4. Let Ξ = 0 be a differential equation describing pseudo-spherical
surfaces with associated one-forms ω α . Under the changes of variables Γ =
tan(ρ/2) and Γ  = cot(ρ/2), Eq. (3.9) and the closed one-form θ1 = ω 1 cos ρ +
ω 2 sin ρ become, respectively,
 
− 2 d Γ = ω 3 + ω 2 − 2 Γ ω 1 + Γ2 ω 3 − ω 2 ; Θ = ω 1 − Γ ω 3 − ω 2 ;
(3.10)
and
 
2 d γ = ω3 − ω2 − 2 γ ω1 + γ 2 ω3 + ω2 ;  = −ω 1 + γ ω 3 + ω 2 ;
Θ
(3.11)
 have been further simplified by adding an
in which the one-forms Θ and Θ
exact one-form.
The functions Γ and γ appearing in (3.10) and (3.11) are quadratic
pseudo-potentials for the PSS equation Ξ = 0. An easy computation using
these equations (see for instance [63]) shows that, conversely, if a differential
equation admits a quadratic pseudo-potential then it is of pseudo-spherical
type.
74 E. G. Reyes Results. Math.

As will be explained in Sect. 4, Eqs. (3.9)–(3.11) can be also used to obtain


Darboux and/or Bäcklund transformations for PSS equations. In this section
focus is shifted to the analysis of a particular example, the Camassa–Holm
equation with non-zero critical wave speed:
3.3. The Camassa–Holm Equation
The Camassa–Holm (CH) equation, [8], reads
− 2 ux uxx + 3 ux u + 2 ux κ − uuxxx + ut − uxxt = 0. (3.12)
As shown in [8], this equation describes shallow water waves; the parameter
κ is the critical shallow water wave speed. More recently, derivations of (3.12)
as a model for shallow water waves of moderate amplitude have been obtained
by Johnson [36] and Constantin and Lannes [15]. CH models breaking waves,
that is, it admits solutions which develop singularities in finite time (see [14]
and also [5]). It also admits peaked traveling waves that interact like solitons
[8,5,42] and moreover, these peaked waves are orbitally stable: their shapes
are stable under small perturbations and therefore they can be recognized
physically [16,41].
The integrability properties Eq. (3.12)—its bi-Hamiltonian character, the
existence of a Lax pair, the existence of a recursion operator—have been
discussed by Camassa and Holm [8], Camassa et al. [9], and by Fokas and
Fuchssteiner [23,26–28] (see also [24]); well-posedness, weak solutions and the
existence of wave breaking have been analyzed by Li and Olver [43] and also
by Constantin and Escher [13,14]; and, it has been proven that the Camas-
sa–Holm equation (3.12) can be interpreted as a geodesic flow on the Virasoro
group, see Misiolek [46]. Finally, the case κ = 0 has been analyzed from the
geometric point of view advocated in this paper in [33,34,60,63].
Theorem 3.5. The Camassa–Holm equation with non-zero critical wave speed,
ut − uxxt = 2 ux uxx − 3 ux u − 2 κ ux + uuxxx , (3.13)
is of pseudo-spherical type.
Proof. Indeed, it is enough to define one-forms ω α , α = 1, 2, 3, as
 
β 1
ω 1 = uxx − u − 2 κ − β + 2 − 2 dx
η η
 
ux − ux β 1−β 2
+ + + u − 1 + u β − u u xx + 2 κ u dt (3.14)
η η2
β + u η 2 − 1 − ux η
ω 2 = ηdx − dt (3.15)
η
ω 3 = (uxx − u + 1) dx
 
uβ − β 2 1 − u ux − ux β
+ + u + + − u u xx − u dt, (3.16)
η2 η2 η
Vol. 60 (2011) Equations of Pseudo-Spherical Type 75

in which the parameters β and η are constrained by



−η 4 + 2 κ + 1 − β 2 − 2 β κ η 2 + 1 − 2 β + β 2 = 0.

The constraint on β and η appearing in the above proof is interesting
because it says that the CH equation describes one-parameter families of
pseudo-spherical surfaces with parameter moving along an elliptic curve,
instead of just along (open subsets of) the real line. One simplifies the asso-
−1
ciated one-forms (3.14)–(3.16) writing β = (s − κ − 1) − s − κ and η =

2 + κ2 + 2 κ − s2 . If ω = fα1 dx + fα2 dt, one finds that the functions fαβ
α

read
f11 = uxx − u − κ + s,
f12 = u2 − uuxx
 2 √
s − 1 − s − 2 sκ + κ2 + κ u + ux 2 + κ2 + 2 κ − s2 − (κ − s)
− ,
s−κ−1

f21 = 2 + κ2 + 2 κ − s2 ,

 2 + κ2 + 2 κ − s2
2 2
f22 = −u 2 + κ + 2 κ − s + ux − ,
s−κ−1
f31 = uxx − u + 1,

2 (−κ − 2 + s) u + ux 2 + κ2 + 2 κ − s2 + 1
f32 = u − uuxx − .
s−κ−1
Now rotate these one-forms and construct a quadratic pseudo-potential γ1
using Eq. (3.11). The resulting expressions are quite involved, so they are not
reproduced here. Change variables from γ1 to γ via

2 + κ2 + 2 κ − s2
γ1 = γ + ,
−s + κ + 1
and define a new parameter λ via s = κ + 1 + λ−1 . One obtains the quadratic
pseudo-potential equations
γ2 1
γx = −u + uxx − −κ+ λ (3.17)
2λ 2
1 uγ 2 1 1
γt = γ 2 + − ux γ − λ2 + κ u − u uxx + λ κ + u2 + λ u (3.18)
2 2λ 2 2
and the parameter dependent conservation law
∂γ ∂γ ∂γ
= −λ + λ uxx − u − ux γ. (3.19)
∂t ∂x ∂x
Equations (3.17) and (3.18) also allow one to obtain a simple linear problem
associated to the Camassa–Holm equation (3.12). Indeed, they imply that the
CH equation (3.13) is the integrability condition of Ψx = XΨ, Ψt = T Ψ, in
which
76 E. G. Reyes Results. Math.

 
0 − 21λ
X=
u − uxx + κ − (1/2) λ 0
and
 1 u

(1/2) ux 2 + 2λ
T = .
(1/2) λ2 + uuxx − u2 − λ κ − κ u − (1/2) λ u −(1/2) ux
In order to obtain conservation laws from this set-up, one expands the
pseudo-potential γ in powers of λ. Specifically, one takes either

 ∞

γ= γn λn/2 or γ =λ+ γn λ−n .
n=1 n=0

Detailed calculations of conservation laws using these expansions appear (in


the κ = 0 case) in [31,60]. Here, as an example, only the first expansion will
be briefly considered. Substituting the first expansion for γ into (3.17), and
writing m = uxx − u − κ, one finds the conserved densities

γ1 = 2 m,
mx
γ2 = −(1/2) ,
m  
1 mx 1  mx  2 1
γ3 = √ − + ,
2m m x 2 m 2
and further densities are found via the recursion relation
1
n+2
γn,x = − γk γn−k+2 n ≥ 3.
2
k=1

4. Darboux/Bäcklund Transforms and Modified Equations


In this paper a Darboux transformation is any transformation connecting two
solutions of a given equation (or of a pair of equations); a Bäcklund trans-
formation is any transformation connecting derivatives of two solutions of a
given equation (or of a pair of equations). Classical examples are provided by
the Korteweg–de Vries and sine–Gordon equations respectively. They are now
recalled within the context of PSS equations:
Example. The KdV equation ut = uxxx + 6uux describes pseudo-spherical
surfaces with associated one-forms

ω 1 = (1 − u)dx + −uxx + λux − λ2 u − 2u2 + λ2 + 2u dt, (4.1)
2
 3
ω = λdx + λ + 2λu − 2ux dt, (4.2)
3
 2 2 2

ω = (−1 − u)dx + −uxx + λux − λ u − 2u − λ − 2u dt, (4.3)
Vol. 60 (2011) Equations of Pseudo-Spherical Type 77

see [12,70]. After rotating the coframe {ω 1 , ω 2 } and changing Γ to −Γ, the
Pfaffian system appearing in (3.10) can be written as

(a) Γx = −u − λΓ − Γ2 , (b) Γt = Γxx − 3Γ2 λ − 2Γ3 x . (4.4)
Take a solution u of KdV and compute Γ from (a). Equation (b) is invariant
under the transformation (Γ → −Γ, λ → −λ), so that u = Γx − Γ λ − Γ2 is
also a solution of KdV. In other words, if u is a solution to KdV, so is
u(x, t, λ) = u(x, t) + 2Γx (x, t, λ), (4.5)
in which Γ is a solution to Equation (b) above. This is the standard Darboux
transformation for KdV (see for instance [64]).
Example. The sine–Gordon equation
uxt = sin u
describes pseudo-spherical surfaces with associated one-forms ([12,70,78])
1 1
ω1 = sin u dt, ω 2 = λdx + cos u dt; ω 3 = ux dx.
λ λ
The Pfaffian system (3.9) can be written as

u = φ + arccos(η φt ); φxt − 1 − η 2 φ2t cos φ = 0. (4.6)
The second equation is invariant under the discrete symmetry (φ, η) →
(π−φ, −η), and so
ū = π − φ + arccos(η φt )
is a new solution to the sine–Gordon equation. It follows that φ = (1/2)(u −
ū + π), and substituting into (4.6) one obtains the system
   
u − ū 2 u + ū
(u + ū)x = 2η sin ; (u − ū)t = sin .
2 η 2
This is the classical Bäcklund transform for the sine–Gordon equation [22,78].
The idea of using discrete symmetries of Pfaffian systems determined by
PSS equations to study their (Bäcklund, Darboux) transformations is due to
Chern and Tenenblat, see [4,12,78]. The sine–Gordon example first appeared
in [12]. Bäcklund transformations of hierarchies of pseudo-spherical type can
be also considered, see [59]. A newer example of a Darboux transformation is
furnished by the associated Camassa–Holm equation [33,34]:
Example. Consider the transformation

p = 2m, d y = p d x − p u d t, and d T = d t. (4.7)
Replacing (4.7) into the CH equation
ut − uxxt = 2 ux uxx − 3 ux u + uuxxx (4.8)
78 E. G. Reyes Results. Math.

one finds the system


 
−p2 pT
pT = −p2 uy , u= − p. (4.9)
2 p y

This is Schiff’s associated Camassa–Holm (ACH) equation, see [74]. Schiff


observed in [74] that (4.9) admits a loop group interpretation, and that this
point of view can be used to derive a Darboux transformation for ACH. In
fact, one can proceed as in the previous examples:
Equations (3.17)–(3.18) for the CH equation imply that the associated
Camassa–Holm equation admits a pseudo-potential γ determined by the
compatible equations
1 2 p λ γ2 pT 1
γy = − γ + + , γT = + γ + λu − λ2 . (4.10)
2λp 2 2p 2 p 2
Computing p(y, T ) from the first equation in (4.10) one obtains
 2 2
∂ ∂ γ (y, T )
p(y, T ) = γ (y, T ) + γ (y, T ) + − λ, (4.11)
∂y ∂y λ
and replacing (4.11) into the equation for γT , one finds an equation for γ which
is invariant under the change
√ √
y → −y, λ → − λ.
Equation (4.11) is not invariant under this change. Instead, one obtains that
 2 2
∂ ∂ γ (y, T )
p(y, T ) = − γ (y, T ) + γ (y, T ) + −λ (4.12)
∂y ∂y λ
is also a solution to ACH. Subtracting (4.11) and (4.12) one concludes that if
γ(y, T ) is determined by Eqs. (4.10), and p(y, T ) is a solution to the associated
Camassa–Holm equation (4.9), then so is
∂γ
p(y, T ) = p(y, T ) − 2 (y, T ).
∂y
This is the transformation Schiff found in [74]. Yet another Darboux transfor-
mation, and a non-linear superposition rule for ACH, appear in [33,34].
With respect to Bäcklund transformations, two further instances are
provided by the following linearizing transforms:
Example. Burgers’ equation ut = uxx + uux is a PSS equation with associated
one-forms
 
1 1 1 1 2 λ
ω = udx + ux + u dt, ω 2 = λ dx + u dt; ω 3 = −ω 2 ,
2 2 4 2
in which λ is a real parameter. The Pfaffian system (3.10) can be written as
Γx
u=2 − 2λΓ, Γt = Γxx − 2λΓΓx , (4.13)
Γ
Vol. 60 (2011) Equations of Pseudo-Spherical Type 79

so that the Cole–Hopf transformation connecting the Burgers and heat equa-
tions is recovered if λ = 0.
Example. The Ibragimov–Shabat equation
ut = uxxx + 3u2 uxx + 9uu2x + 3u4 ux (4.14)
is a PSS equation with associated one-forms
u  u 
ω1 = + u2 dx + + u6 + 8u2x + 5uuxx + 9u3 ux dt, (4.15)
x xxx
u u u 
2 xx 4
ω = λ dx + λ + u + 4uux dt, (4.16)
uu 
ω 3 = −λ dx − λ + u4 + 4uux dt,
xx
(4.17)
u
in which λ is a real parameter. The Pfaffian system (3.10) can be written as
u  
+ u2 Γ + λΓ2 , Γt = Γxxx − 3λ Γxx Γ + Γ2x + 3λ2 Γx Γ2 ,
x
Γx =
u
(4.18)
 −1 2
and therefore the linearizing transformation Γ = u exp Dx u is recovered if
λ = 0.
The Ibragimov–Shabat example is interesting not only because it shows
how a linearizing transformation arises from geometric considerations, but also
because it is an example of a PSS equation of the form ut = uxxx +G(u, ux , uxx )
whose associated one-forms are not covered by the Rabelo–Tenenblat classifi-
cation theorem (Theorem 2.13). Indeed, by Lemma 2.11, if ω 1 = f11 dx+f12 dt,
then f11 must depend at most on u, and not on derivatives of u, and this is
not satisfied by the one-form (4.15). The reason why this is not a contradic-
tion is because Lemma 2.11 was proven using the equality of two differential
ideals, IF and JF , in which the maximal integral submanifolds of {IF , dx∧dt}
are in one-to-one correspondence with local solutions of the equation at hand,
and JF is determined by two-forms satisfying the pseudo-spherical structure
equations if pulled-back by solutions (see Definition 2.10 and the subsequent
discussion). If IF is changed, so will be Lemma 2.11.
Now consider again Eq. (4.4). It hints at the fact that the quadratic
pseudo-potential arising from Proposition 2.8 also encodes “modified” versions
of PSS equations. In fact:
Example. Consider the Pfaffian system (4.4) determined by the KdV equation,
namely,

(a) Γx = −u − λΓ − Γ2 , (b) Γt = Γxx − 3Γ2 λ − 2Γ3 x . (4.19)
Since KdV is of strictly pseudo-spherical type, if Γ is a solution of (b) then u
given by (a) is a solution of KdV. This means that setting λ = 0 one recovers
the Miura transformation and the modified KdV equation.
80 E. G. Reyes Results. Math.

Example. Start from the system of Eqs. (3.17), (3.19) with κ = 0, namely,
1 2 1
m = γx + γ − λ, (4.20)
 2λ 2 
1
γt = λ ux − γ − uγ , (4.21)
λ x

in which λ = 0 is a real parameter and m = uxx − u. Set G = ∂x2 − 1, so that


u = G−1 m, ux = G−1 mx , and uxx = G−1 mxx . Expanding (4.21) and using
(4.20), one finds
  
2 −1 γ2 λ
Gγt = γx + γγxx + λγx − G∂x γG γx + − . (4.22)
2λ 2
Equation (4.22) is the modified Camassa–Holm equation introduced in [31].
It can be further simplified by applying the Galilei transformation x = X +
(λ/2)T, t = T, u = U + λ/2. The CH equation (4.8) becomes

2 ux uxx + u uxxx − λux = ut − uxxt + 3 ux u,

while the pseudo-potential Eqs. (4.20) and (4.21) now read


1 2
m = γx + γ , (4.23)
 2λ 
1
γt = λ ux − γ − uγ . (4.24)
λ x

Equation (4.23) is a perfect analog of the Miura transformation of KdV theory.


Using (4.23) and (4.24) one can write the mCH equation (4.22) in the form
γ2
γt = + λG−1 m − ∂x (γG−1 m). (4.25)
2
This equation can be transformed into a system of partial differential equa-
tions, simply setting G−1 m = v. Using Eq. (4.23) one obtains the system
γ2
γt = + λv − ∂x (γv), (4.26)
2
γ2
vxx − v = γx + . (4.27)

Note that this system is simply an interesting way of writing the nonlocal
equation (4.25) for γ. However, if v satisfies (4.26) and (4.27), then v also sat-
isfies (formally) the CH equation! Thus, (4.26) and (4.27) can be understood
as a Bäcklund transformation for the CH equation.
It is proven in [31] that mCH describes pseudo-spherical surfaces. It is
also proven in that reference that if q ∈ ( 12 , 1] ∪ ( 32 , ∞) and γ0 belongs to the
Sobolev space H q (R), the problem
Vol. 60 (2011) Equations of Pseudo-Spherical Type 81

γ2
γt = + λv − ∂x (γv), γ(0) = γ0 , (4.28)
2
γ2
vxx − v = γx + (4.29)


admits a weak solution in the space W 1,∞ (0, T ; H q−1 ) ∩ L∞ (0, T ; H q ) ×
L∞ (0, T ; H q+1 ) such that γ(0) = γ0 , and that this solution is unique for q = 1
or q > 3/2.

5. Correspondence Theorems
A correspondence theorem is a theorem relating two solutions of a pair of PSS
equations. The correspondence itself depends on the particular solution one
starts with, and it uses an explicit change of independent variables. Thus,
these theorems are quite different from the transformations of the last section.
The key observation behind these results is that two (pseudo-)Riemannian sur-
faces of constant Gaussian curvature equal to −1 are locally indistinguishable.
As will be discussed here, it follows that given two PSS equations, and two
III-generic (II-generic, I-generic) solutions u(x, t) and u x, 
( t ) of them, one
(
can relate u(x, t) and u 
x, t ) by integrating first order systems of equations,
and moreover, one can obtain a formula for u (x, 
t ) in terms of u(x, t) and
some functions determined solely by u(x, t). The first theorem of this kind was
proven by Kamran and Tenenblat, see [38], who treated the Riemannian case.
The pseudo-Riemannian case was considered in [57].
Lemma 5.1. Let Ξ(x, t, u, . . . ) = 0 be a differential equation describing pseudo-
spherical surfaces. One can choose one-forms ω α = fα1 dx + fα2 dt associated
to Ξ = 0 such that ω α = udx+βdt for at least one index α, 1 ≤ α ≤ 3. In fact,
one can always choose associated one-forms ω α such that ω 1 = udx + βdt.
Proof. Let ω α = fα1 dx + fα2 dt, be one-forms associated with the PSS equa-
tion Ξ = 0. One can always assume that f11 = 0, for, if f21 = 0, one scan
imply rotate the coframe using Transformation (2.4) with ρ = π/2 to obtain
associated one-forms ω  i = fi1 dx + fi2 dt satisfying f11 = 0, while if f21 = 0,
one can apply Transformation (2.4) with ρ = arctan(−f11 /f21 ) to find again
f11 = 0.
Assume then that ω α = fα1 dx + fα2 dt, α = 1, 2, 3, satisfy f11 = 0.
There are three cases. First, if f31 = 0, one applies Transformation (2.5) with
ρ = arcsinh(−u/f31 ) to find new associated one-forms with f11 = u. Second, if
f31 = 0 but f21 = 0, one uses Transformation (2.6) with ρ = arcsinh(−u/f21 )
to obtain f31 = u. Third, if f31 = f21 = 0, one uses (2.5) to find new associated
one-forms with f11 = f31 = 0 and f21 = 0, and proceed as in the second case.
The last assertion of the Lemma is trivial if the construction above yields
either f11 = u or f21 = u. Suppose then that the equation Ξ = 0 has associated
82 E. G. Reyes Results. Math.

one-forms ω α satisfying ω 3 = udx + βdt. Reasoning as in the first paragraph


 i satisfying either
of this proof, one gets new associated one-forms ω
f11 = 0 and f31 = u, or
  
  f11
f11 = 0 and f31 = u + Dx arctan − .
f21
In any of these two cases f31 = 0, and therefore, reasoning as in the second
paragraph, one can find new one-forms, ω new
i
say, satisfying f11 new = u, as
claimed. 
This lemma is proven in detail because it is not too well-known and also
because it can, presumably, be used to simplify the characterization results
reviewed in Sect. 2.2.
Theorem 5.2. Let Ξ(x, t, u, . . . ) = 0 and Ξ( x,  , . . . ) = 0 be two scalar PSS
t, u
equations with associated one-forms ω α = fα1 dx + fα2 dt and ω  α = fα1 d
x+
 
fα2 dt respectively. Any III-generic solution u(x, t) of Ξ(x, t, u, . . . ) = 0 deter-
mines a III-generic solution u (x,   x, 
t ) of Ξ( , . . . ) = 0 via
t, u
 
1 ∂ψ ∂ψ
◦Φ =
u (cos θf11 + sin θf21 ) − (cos θf12 + sin θf22 ) , (5.1)
J ∂t ∂x
in which Φ(x, t) = (ϕ(x, t), ψ(x, t)) is a local diffeomorphism with Jacobian J,
and the functions θ(x, t), ϕ(x, t), and ψ(x, t) are smooth solutions of the sys-
tem of equations
  ∂ϕ   ∂ψ
Φ∗ f21 + Φ∗ f22 = − sin θ f11 + cos θ f21 (5.2)
∂x ∂x
  ∂ϕ   ∂ψ
Φ∗ f21 + Φ∗ f22 = − sin θ f12 + cos θ f22 (5.3)
∂t ∂t
  ∂ϕ   ∂ψ ∂θ
Φ∗ f31 + Φ∗ f32 = f31 + (5.4)
∂x ∂x ∂x
  ∂ϕ   ∂ψ ∂θ
Φ∗ f31 + Φ∗ f32 = f32 + (5.5)
∂t ∂t ∂t
  ∂ϕ
− Φ∗ f12 J = (cos θf11 + sin θf21 )
∂t
∂ϕ
− (cos θ f12 + sin θ f22 ) (5.6)
∂x
where the pull-backs of u  and its derivatives appearing in (Φ∗ fαβ ), have been
evaluated by means of (5.1) and its derivatives.
This is the Kamran–Tenenblat theorem, [38], as it appears in [57]. As an
example, consider a non-linear superposition of sine–Gordon solutions:
Example. The sine–Gordon equation
∂2θ ∂2θ
− = sin θ cos θ (5.7)
∂u2 ∂v 2
Vol. 60 (2011) Equations of Pseudo-Spherical Type 83

describes pseudo-spherical surfaces with associated functions


f11 = 0, f12 = sin θ, f21 = − cos θ, f22 = 0, f31 = θv , f32 = θu . (5.8)
Let θ(u, v) be a III-generic solution of Eq. (5.7). Theorem 5.2 allows one to
obtain the traveling wave solution of Burgers equation from (5.7) and (5.8).
Consider a function ω(u, v) determined by the completely integrable Pfaffian
system
θv + ωu = − sin ω cos θ, and θu + ωv = cos ω sin θ, (5.9)
and introduce potentials ξ(u, v) and ζ(u, v) thus:
dξ = −(cos ω cos θ du + sin ω sin θ dv), and (5.10)
dζ = exp(ξ)(cos θ sin ω du − sin θ cos ω dv). (5.11)
Now consider Burgers’ equation u xx − 2
t = u uux with associated functions
f11 = u
, f12 = − x , f21 = 1,
u2 + u
f22 = −u, f31 = 1, f32 = − u, (5.12)
and define the transformation
 = ϕ(u, v) = 1 − ζ(u, v) − exp(ξ(u, v)),
x (5.13)

t = ψ(u, v) = ln |1 − exp(ξ(u, v))| − 1 + ζ(u, v) + exp(ξ(u, v)). (5.14)
One can check that, if θ̃ = ω(u, v) − π/2, the map Φ : (u, v) → ( x, 
t ) given
by (5.13) and (5.14) is a local diffeomorphism which satisfies the system of
Eqs. (5.2)–(5.6) (with “θ” replaced by “θ̃”). Substituting into Eq. (5.1) (again,
with “θ” replaced by “θ̃”) one finds that the pull–back of u x, 
( t ) by Φ is
1 − exp(ξ(u, v))
◦Φ =
u ,
exp(ξ(u, v))
and therefore one obtains
x+
exp( t)
x, 
(
u t) = , (5.15)
x+
1 − exp( t)
the traveling wave solution of Burgers’s equation.
As remarked above, the Kamran and Tenenblat’s result [38] covers
III-generic solutions, while the I- and II-generic cases are considered in [57]. For
illustration purposes, some details of the proofs appearing in [57] are provided
below.
Theorem 5.3. Let Ξ(x, t, u, . . . ) = 0 and Ξ( x,  , . . . ) = 0 be two PSS equa-
t, u
α
tions with associated one-forms ω = fα1 dx + fα2 dt and ω  α = fα1 dx + fα2 d
t
respectively. For any II-generic solutions u(x, t) of Ξ = 0 and u (   = 0,
x, t ) of Ξ
there exist a local diffeomorphism Υ : V → V , in which V and V are open sub-
sets of the domains of u(x, t) and u x, 
( t ) respectively, and a smooth function
μ : V → R, such that ω α (u(x, t)) and ω  α ( x, 
u( t )) satisfy
84 E. G. Reyes Results. Math.

Υ∗ ω
 1 = ω 1 cosh μ − ω 3 sinh μ, (5.16)
Υ∗ ω
 2 = ω 2 + dμ, (5.17)
Υ∗ ω
 3 = −ω 1 sinh μ + ω 3 cosh μ. (5.18)
Proof. Define one-forms σ α on an open subset W  of V by
1 1 1
σ1 = d x, σ 2 = d t, and σ 3 = d t. (5.19)

x 
x 
x
The pseudo-spherical structure equations imply that ( σ 1 )2 − (σ 2 )2 is a Lorentz

metric of constant Gaussian curvature K = −1 on W , and that σ 3 is the con-
1 2
nection one-form corresponding to the moving coframe { σ ,σ  }. There exists
a function Γ : V → W  , Γ(x, t) = (α(x, t), β(x, t)), and a real–valued function
θ(x, t) on V , such that
Γ∗ σ
1 = ω 1 cosh θ − ω 3 sinh θ, (5.20)
Γ∗ σ
2 = ω 1 sinh θ − ω 3 cosh θ, (5.21)
Γ∗ σ
3 = ω 2 + dθ. (5.22)
Indeed, Eqs. (5.20)–(5.22) are equivalent to the Pfaffian system
αx = α (f11 cosh θ − f31 sinh θ), αt = α (f12 cosh θ − f32 sinh θ), (5.23)
βx = α (f11 sinh θ − f31 cosh θ), βt = α (f12 sinh θ − f32 cosh θ), (5.24)
θx = f11 sinh θ − f31 cosh θ − f21 , θt = f12 sinh θ − f32 cosh θ − f22 , (5.25)
and this system is completely integrable since the one-forms ω α (u(x, t)) satisfy
the structure equations of a pseudo-spherical surface. Moreover, Γ = (α, β) is a
local diffeomorphism, αx βt − αt βx = 0, because u(x, t) is a II-generic solution
of Ξ = 0. An analogous argument can be used to find a local diffeomorphism
Γ̃ : V → W
 and a function θ̃ : V → R such that
Γ̃∗ σ
1 = ω
 1 cosh θ̃− ω
 3 sinh θ̃, Γ̃∗ σ
2 = ω
 1 sinh θ̃− ω
 3 cosh θ̃, Γ̃∗ σ
3 = ω
 2 + dθ̃.
It is then straightforward to check that Eqs. (5.16)–(5.18) are satisfied if one
defines Υ = Γ̃−1 ◦ Γ, and μ = θ − θ̃ ◦ Υ. 
In geometric terms, the function Υ is a local isometry and μ determines
a Lorentz boost. Theorem 5.3 can be used to construct a transformation from
x, 
(
a solution u(x, t) of Ξ = 0 to a solution u  = 0:
t ) of Ξ
Theorem 5.4. Let Ξ(x, t, u, . . . ) = 0 and Ξ(  x,  , . . . ) = 0 be two PSS equa-
t, u
tions, assume that Ξ = 0 has associated one-forms ω α = fα1 dx + fα2 dt, and
that there exist one-forms ω 0α associated to Ξ  = 0 such that this equation
is necessary and sufficient for the pseudo-spherical structure Eqs. (2.1) to be
satisfied. Any II-generic solution u(x, t) of Ξ(x, t, u, . . . ) = 0 determines a
II-generic solution u x, 
(  x, 
t ) of Ξ( , . . . ) = 0 by
t, u
1
u ◦ Υ = [ψt (f21 + μx ) − ψx (f22 + μt )] , (5.26)
J
Vol. 60 (2011) Equations of Pseudo-Spherical Type 85

where Υ(x, t) = (φ(x, t), ψ(x, t)) is a local diffeomorphism, and φ, ψ, μ are
solutions to the system of equations
   
Υ∗ f11 φx + Υ∗ f12 ψx = f11 cosh μ − f31 sinh μ, (5.27)
   
Υ∗ f11 φt + Υ∗ f12 ψt = f12 cosh μ − f32 sinh μ, (5.28)
 
J Υ∗ f22 = − [φt (f21 + μx ) − φx (f22 + μt )] , (5.29)
   
Υ∗ f31 φx + Υ∗ f32 ψx = −f11 sinh μ + f31 cosh μ, (5.30)
   
Υ∗ f31 φt + Υ∗ f32 ψt = −f12 sinh μ + f32 cosh μ, (5.31)

in which the pull-backs of u  and its derivatives with respect to x, 


t appearing
in the functions (Υ∗ fαβ )(x, t) have been evaluated using (5.26).
Proof. Consider the one-forms ω 0α . Applying Lemma 5.1, one obtains one-forms
 = fα1 d
ω α
x + fα2 d  = 0 such that f21 = u
t associated with the equation Ξ . Fix
a II-generic solution u(x, t) of equation Ξ(x, t, u, . . . ) = 0, and consider local
solutions φ, ψ and μ to the system of Eqs. (5.27)–(5.31) such that Υ = (φ, ψ)
is a local diffeomorphism. (That such functions do exist is a technical lemma
appearing in [57]). Define u  ◦ Υ by Eq. (5.26). Then it can be checked that
(5.26), (5.27)–(5.31) is a system of six equations equivalent to
Υ∗ ω
 1 = ω 1 cosh μ − ω 3 sinh μ, (5.32)
Υ∗ ω
 2 = ω 2 + dμ, (5.33)
Υ∗ ω
 3 = −ω 1 sinh μ + ω 3 cosh μ. (5.34)
Since Υ is a local diffeomorphism, and the one-forms ω α satisfy the struc-
 α . It fol-
ture Eqs. (2.1) of a pseudo-spherical surface, so do the one-forms ω


lows that (5.26) determines a solution of Ξ = 0, as claimed. That u  x , 
t
is  = 0 follows from the identity
 a II-generic
 solution
  of the equation
 Ξ

   
f11 ◦ Υ f32 ◦ Υ − f12 ◦ Υ f31 ◦ Υ [φx ψt − φt ψx ] = f11 f32 − f12 f31 .


Examples of transformations obtained using this result appear in [57].


In that reference a third correspondence result for I-generic solutions to PSS
equations is also stated.
Now, it is interesting to note that the results considered here can be used
to obtain (local) isometric immersions of the pseudo-spherical surfaces deter-
mined by PSS equations. Even though this is an elementary observation, it is
stated here as a theorem:
Theorem 5.5. Let Ξ = 0 be a PSS equation with associated one-forms ω α =
fα1 dx + fα2 dt, let u(x, t) be a III-generic solution, and let V be an open sub-
set of the domain of u(x, t). There exists a local isometric immersion of the
86 E. G. Reyes Results. Math.

pseudo-spherical surface V with metric ω 1 (u(x, t)) ⊗ ω 1 (u(x, t)) + ω 2 (u(x, t)) ⊗
ω 2 (u(x, t)) into R3 .
Proof. First of all, recall that the sine–Gordon equation uxt = sin u describes
pseudo-spherical surfaces [12,70,78]. More than that, it is a classical fact [22]
that the sine–Gordon equation describes immersed pseudo-spherical surfaces
via the associated one-forms
1
 1 = sin ud
ω t (5.35)
η
1
ω x + cos ud
 2 = ηd t (5.36)
η
12 = ux d
ω x (5.37)
1
13 = sin ud
ω t (5.38)
η
1
23 = ηd
ω x + cos udt, (5.39)
η
in which the classical notation ω  3 . Now, it follows
12 has been used instead of ω
from the Kamran–Tenenblat correspondence theorem [38] that there exists a
local diffeomorphism Φ from V to an open subset V of the domain of a III-
generic solution u(x, 
t) to the sine–Gordon equation, and a smooth function
θ : V → R such that
Φ∗ ω
 1 = cos θ ω 1 + sin θ ω 2 (5.40)
∗ 2 1 2
 = − sin θ ω + cos θ ω
Φ ω (5.41)
∗ 3 3
 = ω + dθ.
Φ ω (5.42)

The simple observation to be made now is that the one-forms Φ ω  and Φ∗ ω
α
αβ ,
in which ω  and ω
α
αβ are given by (5.35)–(5.39), satisfy the structure equa-
tions of a surface locally immersed in R3 . But the metric of this immersed
surface is precisely ω 1 (u(x, t)) ⊗ ω 1 (u(x, t)) + ω 2 (u(x, t)) ⊗ ω 2 (u(x, t)) because
of Eqs. (5.40) and (5.41). 
Finally, note that correspondence theorems also exist for hierarchies of
pseudo-spherical type. Since proofs (and statements) become more technical,
only a result analogous to Theorem 5.3 will be recalled in this paper. Details
and examples are in [59].
τi = Fi (
Theorem 5.6. Let uτi = Fi (x, t, u, . . . ) and u x,  , . . . ) be two hierar-
t, u
chies of pseudo-spherical type with associated one-forms given by

n
[n]
Θα = fα1 dx + fα2 dt + hαi dτi and
i=1
n
 [n] = fα1 d
Θ x + fα2 d
t+ 
hαi d
τi . (5.43)
α
i=1
Vol. 60 (2011) Equations of Pseudo-Spherical Type 87

Let {u[n] } and {


u[n] } be solutions of uτi = Fi and u τi = Fi respec-
[0]
tively, and assume that u (x, t) and u [0]
 ( 
x, t ) are III-generic. For each n ≥ 0
there exist a local diffeomorphism Υ[n] : V [n] → V [n] —in which V [n] and
V [n] are open subsets of the domains of u[n] and u [n] —and a smooth function
[n] [n]
μ : V → R such that
 = Θ cos μ[n] + Θ sin μ[n] ,
Υ[n]∗ Θ
[n] [n] [n]
(5.44)
1 1 2
 [n] = −Θ[n] sin μ[n] + Θ[n] cos μ[n] ,
Υ[n]∗ Θ (5.45)
2 1 2
[n]∗  [n] [n] [n]
Υ Θ = Θ + dμ .
3 3 (5.46)
[n] [n]
Moreover, the maps Υ and μ can be chosen so that for each n ≥ 0,
Υ[n+1] |V [n] = Υ[n] and μ[n+1] |V [n] = μ[n] . (5.47)
Remark 5.7. It is important to stress that, as it happens in the single equa-
tion case, the correspondence depends on the particular solution {u[n] } of the
hierarchy uτi = Fi one starts with, and that the proof of its existence relies
on Frobenius theorem (see [59]). Since the natural arena for the study of the
formal geometry of differential equations is the theory of infinite dimensional
jet spaces (see for instance [40] and references therein) and at this level there
is no Frobenius theorem available [40,77], the results of this subsection do not
imply that one can “transfer” information on for instance, conservation laws
or symmetries, from one hierarchy to another.

6. Nonlocal Symmetries and Particular Solutions


In this final section recent developments on the (nonlocal) symmetry struc-
ture of some equations of pseudo-spherical type are reviewed. A non-local
symmetry is, roughly, a symmetry depending on (for instance) integrals of
dependent variables, so that, from the point of view of the geometric theory of
differential equations [40], these symmetries cannot be considered as objects
defined on a universal infinite jet space. A geometric theory of nonlocal symme-
tries has been put forward by Vinogradov, Krasil’shchik and their co-workers,
see [40,79].

6.1. On Nonlocal Symmetries


The following example appears in the seminal paper [79], and it is already a
tradition to begin discussions on nonlocal symmetries quoting it:
Example. Consider Burgers’ equation
ut = uxx + uux . (6.1)
The expression
1

G = (2Sx − uS) e− 2 u dx
, (6.2)
88 E. G. Reyes Results. Math.

in which S is any function such that St = Sxx , formally satisfies the linearized
Burgers equation Gt = Gxx + G ux + u Gx whenever u is a solution of Equation
(6.1), so that G should determine a symmetry of (6.1), in analogy with the
local theory [40,49]. To make this observation rigorous, one considers an extra
dependent variable γ 1 such that γx1 = u and γt1 = ux + (1/2) u2 , so that (6.2)
becomes G = (2Sx − uS) exp(− 12 γ 1 ). Thus, G is now “local” and it could
be considered as the characteristic of a local symmetry for the “augmented”
system
ut = uxx + uux ; γx1 = u; γt1 = ux + (1/2) u2 . (6.3)
But then, as observed in [79], in order to formalize this idea, one also needs to
consider the infinitesimal variation of γ 1 as u changes to u + τ G !
This example motivates the following two definitions:
Definition 6.1. Let N be a non-zero integer. An N -dimensional covering π of
a (system of) partial differential equation(s) Ξa = 0, a = 1, . . . , k, is a pair

π = {γ b : b = 1, . . . , N }; {Xib : b = 1, . . . , N ; i = 1, . . . , n} (6.4)
of variables γ b and smooth functions Xib depending on xi , uα , γ b and a finite
number of partial derivatives of uα , such that the equations
∂γ b
= Xib (6.5)
∂xi
are compatible whenever uα (xi ) is a solution to Ξa = 0.
One usually writes π = (γ b ; Xib ) instead of (6.4). The variables γb are
considered as new dependent variables, the “nonlocal variables” of the theory.
Eqs. (6.5) relate them to the original variables uα .
Definition 6.2. Let Ξa = 0, a = 1 . . . , k, be a system of partial differential
equations, and let π = (γ b ; Xib ) be a covering of Ξa = 0. A nonlocal π–
symmetry of Ξa = 0 is a generalized symmetry
 ∂  ∂  ∂
X= ξi i + φα α + ϕb b (6.6)
i
∂x α
∂u ∂γ
b
of the augmented system
∂γ b
Ξa = 0; = Xib . (6.7)
∂xi
This definition of nonlocal symmetries is essentially in [40, pp. 249–250].
It follows that in order to find nonlocal π-symmetries, one proceeds exactly as
in the local case considered for instance in P. Olver’s treatise [49, Chapter 5].
In particular, [49, p. 291], it is enough to consider “evolutionary” vector fields
X of the form
m
∂ 
N

X= Gα α + Hb b . (6.8)
α=1
∂u ∂γ
b=1
Vol. 60 (2011) Equations of Pseudo-Spherical Type 89

Following [40], G = (G1 , . . . , Gm ) is called the π-shadow of the nonlocal sym-


metry (6.8). A fundamental result appearing in [40] states that if G is a
π-shadow, and N (the number of new dependent variables) is allowed to be
infinite, it is always possible to find a further covering π̂ and a bona fide
m N
nonlocal π̂-symmetry X of the form X = α=1 Gα ∂u∂α + b=1 H b ∂γ∂ b .
From the general theory of generalized symmetries (see [40], [49, Chap-
ter 5]) one concludes that nonlocal symmetries can be used to generate solu-
tions:
Corollary 6.3. If uα i b i
0 (x ) and γ0 (x ) are solutions to the augmented system
(6.7), the solution to the Cauchy problem
∂uα ∂γ b
= Gα , = H b ; uα (xi , 0) = uα i
0 (x ), γb (xi , 0) = γ0b (xi );
∂τ ∂τ
is a one-parameter family of solutions to the augmented system (6.7). In partic-
ular, nonlocal π–symmetries send solutions to the system Ξa = 0 to solutions
of the same system.
6.2. The Kaup–Kupershmidt, Sawada–Kotera and 5th Order Korteweg–de
Vries Equations
This subsection is on nonlocal symmetries of the Kaup–Kupershmidt (KK)
25
qt = qxxxxx + 5 q qxxx + qx qxx + 5 q 2 qx , (6.9)
2
Sawada–Kotera (SK)
qt = qxxxxx + 5 q qxxx + 5 qx qxx + 5 q 2 qx , (6.10)
and fifth order Korteweg–de Vries (KdV5)
5√ 10 √
qt = qxxxxx + 6 q qxxx + 6 qx qxx + 5 q 2 qx (6.11)
3 3
equations. The usual fifth order KdV equation, see for instance [73], qt =
qxxxxx + 53 q qxxx + 103 qx qxx +
5 2
q qx , is obtained from (6.11) simply by mak-
√6
ing the replacement q → (1/ 6) q. The choice of coefficients in (6.11) will be
clear momentarily. There are at least two reasons why these three equations
are interesting to study from the point of view of nonlocal symmetries. First,
see [69], Eqs. (6.9), (6.10) and (6.11) are the only three fifth order polyno-
mial and 2-homogeneous evolution equations which admit an infinite number
of generalized symmetries (an equation ut = F is λ-homogeneous of weight
μ if it admits the one-parameter group of scaling symmetries (x, t, u) →
(a−1 x, a−μ t, aλ u), a > 0). Second, it is quite surprising that, in spite of their
similitude (note the the difference between (6.9) and (6.10) is just one coeffi-
cient) the KK, SK and KdV5 equations possess very different characteristics:
Kaup [39] introduced Eq. (6.9) in 1980 and found a solitary wave solution
for it using scattering/inverse scattering techniques [2,3], but other interest-
ing solutions, and transformations relating them, have been found explicitly
90 E. G. Reyes Results. Math.

only in the last 6 years, see [62] and references therein. On the other hand,
the Sawada–Kotera equation (6.10) was discovered in 1974, [72], its N -soliton
solutions were readily computed, and its main algebraic properties (e.g. Backl-
und transformation, zero curvature representation) were already known by
1983 [21,72]. And of course, the fifth order KdV equation (6.11) is a standard
soliton equation, see for instance [1,11,19].
These three equations describe pseudo-spherical surfaces. Since the
explicit expressions for associated one-forms are not used in what follows, only
linear problems will be presented:
Proposition 6.4. The family of equations
   
1 3
qt = qxxxxx − 4y + q qxxx + 5 q 2 qx − 2y + qx qxx , (6.12)
y y
in which y is a non-zero real parameter, is the integrability condition of the
sl(2, R)-valued linear problem Xψ = ψx , T ψ = ψt where
 
0 −y/η 2
X= (6.13)
−η 2 q 0
and
 
yqxxx − qqx 2y 2 qxx /η 2 − yq 2 /η 2
T = 2 .
η (−qxxxx + (2y + 1/y)qqxx + qx2 /y − q 3 ) −yqxxx + qqx
(6.14)
Note that Eq. (6.12) is 2-homogeneous of weight 5 for arbitrary non-zero
values of the real parameter y. The Kaup–Kupershmidt equation corresponds
to (6.12) with y = −1/4, Sawada–Kotera
√ is (6.12) with y = −1, and the fifth
order KdV is (6.12) with y = −1/ 6. Explicit associated one-forms for these
three equations have been found recently by Gomes-Neto [30] using the char-
acterization results of Kamran and Tenenblat [38]. From Proposition 6.4 one
finds a quadratic pseudo-potential for Eq. (6.12):
Lemma 6.5. Equation (6.12) admits the quadratic pseudo-potential
y
αx = −η 2 q + 2 α2 (6.15)
η
 
η2 η2 2
αt = −η 2 qxxxx + 2 η 2 y + q qxx + q − η 2 q 3 + (2 q qx − 2y qxxx )α
y y x
 
y 2 2y 2
+ q − q xx α2 , (6.16)
η2 η2
that is, the system (6.15) and (6.16) is completely integrable for α(x, t) when-
ever q(x, t) is a solution to Eq. (6.12).
Now one can use (6.15) to write Eq. (6.16) as a conservation law and
define a potential δ. One finds that δ is determined by the following two com-
patible equations:
Vol. 60 (2011) Equations of Pseudo-Spherical Type 91

2y
δx = α (6.17)
η2
 
4y 2 2y
δt = −2y qxxx + 2 qqx + − 2 qxx + 2 q 2 α. (6.18)
η η
One would like to obtain a shadow of a nonlocal symmetry of the form
G = α exp(−Lδ), for some constant L, following previous work by Galas [29]
and the present author [60–62]. It turns out that the existence of this shadow
singles out exactly the KK, SK and KdV5 equations:
Theorem 6.6. Consider the function
G = α exp(−L δ) (6.19)
on the covering π of Eq. (6.12) determined by the nonlocal variables α and δ
satisfying Eqs. (6.15)–(6.18). Then, G is the shadow of a nonlocal symmetry
for Eq. (6.12) if and only if
4y 2 + 1
L(y) = , (6.20)
10y 2
and the parameter y satisfies the equation

− 125y 2 96y 6 − 118y 4 − 1 + 23y 2 = 0. (6.21)
Theorem 6.6 can be checked with the help of a symbolic computation
software such as MAPLE. Since y cannot be equal to zero, Eq. (6.21) gives
one exactly six values for which (6.19) is a π-shadow:
1 1 1 1
y = 1, −1 ; , − ; √ , − √ .
4 4 6 6
The corresponding
√ values
√ of L are L(1) = L(−1) = 1/2, L(1/4) = L(−1/4) = 2,
and L(1/ 6) = L(−1/ 6) = 1. The equations obtained by replacing these val-
ues of y into (6.12) are, respectively, the Sawada–Kotera, Kaup–Kupershmidt,
and fifth order KdV equations! (one moves from (6.12) with y = 1 to (6.12)
with y = −1, for instance, via the obvious change of variables q → −q). Thus,
the shadow (6.19) recognizes precisely the only 2-homogeneous polynomial evo-
lution equations which possess an infinite number of generalized symmetries,
out of a whole family of equations which one could, a priori, have thought “spe-
cial”, in the sense that they are the integrability condition of an overdetermined
linear problem (although it can be checked that the “spectral parameter” η
can be eliminated via a gauge transformation), and they do admit quadratic
pseudo-potentials and conservation laws. It is still an open problem whether it
is true in general that shadows such as (6.19) single integrable equations out
of the class of equations of pseudo-spherical type.
Now one extends the shadow (6.19) to a nonlocal symmetry for (6.9)–
(6.11). Set
 
4y 2 + 1
qτ = α exp − δ (6.22)
10y 2
92 E. G. Reyes Results. Math.

and find the variations of α and δ with respect to the deformation parame-
ter τ . In order to obtain the variation of α, take derivatives with respect to τ
in Eq. (6.15) and solve the resulting equation for ατ using (6.17). One finds,
setting an integration constant equal to zero,
 
−η 4 −10y 2 4y 2 + 1 5yη 4 4y 2 + 1
ατ = exp − δ = exp(− δ). (6.23)
2y 14y 2 + 1 10y 2 14y 2 + 1 10y 2
Now consider the variation of δ with respect to the parameter τ :
 
2y 10η 2 y 2 4y 2 + 1
δxτ = δτ x = 2 ατ = exp − δ .
η 14y 2 + 1 10y 2
Since the right hand side of this equation is not a total x-derivative, one simply
defines a further potential β by means of
 
10η 2 y 2 4y 2 + 1
βx = exp − δ (6.24)
14y 2 + 1 10y 2
and sets
δτ = β. (6.25)
It remains to compute βt = δτ t = δtτ and βτ . Using (6.18), (6.22)–(6.25) one
gets
 
(4 y2 +1)δ
2 exp − 10y2

βt = −1764 y 8 + 2016 y 6 − 239 y 4 − 14 y 2 + 1 α4
125y 2 η 6 (14 y 2 + 1)

+ 85 y 3 qη 4 − 280 y 7 η 4 q + 5 qy η 4 + 190 y 5 qη 4 α2

+ 1400 y 6 η 6 qx + 450 qx y 4 η 6 + 25 qx y 2 η 6 α + (−25 y 4 η 8 q 2
+ 500 y 5 η 8 qxx − 50 q 2 y 2 η 8 + 125 y 3 η 8 qxx + 700 y 6 η 8 q 2 ) (6.26)
and
4y 2 + 1 2
βτ = − β .
20y 2
Equations (6.24) and (6.26) are not compatible for arbitrary non-zero values
of y, in contradistinction with the equations for α and δ. This happens because
the pseudo-potential α and the potential δ depend only on the realization of
the nonlinear equation (6.12) as a zero curvature condition, while β depends
explicitly on the shadow (6.19).
Theorem 6.7. Consider the pseudopotential α determined by Eqs. (6.15) and
(6.16), and the potential δ given by (6.17) and (6.18). The system of Eqs. (6.24)
and (6.26) determines a further potential β via (6.24) and (6.26) if and only if

− 125y 2 96y 6 − 118y 4 − 1 + 23y 2 = 0. (6.27)
Vol. 60 (2011) Equations of Pseudo-Spherical Type 93

If this condition is satisfied, the vector field


   
4y 2 + 1 ∂ 5yη 4 4y 2 + 1 ∂ ∂
W = α exp − δ + exp − δ +β
10y 2 ∂q 14y 2 + 1 10y 2 ∂α ∂δ
4y 2 + 1 2 ∂
− β (6.28)
20y 2 ∂β
is a nonlocal symmetry for Eq. (6.12).
Hereafter assume (6.27). The flow of the vector field (6.28) is found by
solving the system of equations
   
∂q 4y 2 + 1 ∂α 5yη 4 4y 2 + 1
= α exp − δ ; = exp − δ ; (6.29)
∂τ 10y 2 ∂τ 14y 2 + 1 10y 2
∂δ ∂β 4y 2 + 1 2
= β; =− β ; (6.30)
∂τ ∂τ 20y 2
with initial conditions
q(x, t, 0) = q0 ; α(x, t, 0) = α0 ; δ(x, t, 0) = δ0 ; β(x, t, 0) = β0 ,
(6.31)
in which q0 , α0 , δ0 and β0 are arbitrary particular solutions to the compati-
ble system of Eqs. (6.9), (6.15)–(6.18), (6.24) and (6.26). The solution to this
initial value problem is
 2

−2000y 5 η 4 exp − 4y10y+12 δ
α(τ ) =
[β0 (4y 2 + 1)τ + 20y 2 ]β0 (4y 2 + 1)(14y 2 + 1)
 2

100y 3 η 4 exp − 4y10y+1 2 δ
+ + α0 ; (6.32)
β0 (4y 2 + 1)(14y 2 + 1)
! !
20y 2 ! β0 (4y 2 + 1)τ + 20y 2 !
δ(τ ) = 2 ln ! ! + δ0 ; (6.33)
4y + 1 ! 20y 2 !

20y 2 β0
β(τ ) = . (6.34)
β0 τ (4y 2 + 1) + 20y 2
The corresponding general formula for q(x, t, τ ) can be obtained
√ from the
first equation in (6.29). In the relevant cases y = −1/4, −1, −1/ 6 one finds:
Corollary 6.8. Assume that the functions q0 , α0 , δ0 and β0 are arbitrary solu-
tions to the compatible√system of Eqs. (6.9), (6.15)–(6.18), (6.24) and (6.26)
for y = −1/4, −1, −1/ 6. Then, respectively,

τ 2 η 4 e−4 δ0 τ α0 e−2 δ0
qKK (x, t, τ ) = q0 − 2 + (6.35)
3 (τ β0 + 1) τ β0 + 1
94 E. G. Reyes Results. Math.

is a family of solutions to the Kaup–Kupershmidt equation;

8η 4 e−δ0 τ 2 4τ
qSK (x, t, τ ) = − 2 + α0 e−(1/2) δ0 + q0 (6.36)
3 (β0 τ + 4) (β0 τ + 4)

is a family of solutions to the Sawada–Kotera equation; and


√ 2
6τ −2 δ0 η 4 + 2τ e−δ0 α + q (6.37)
qKdV5 (x, t, τ ) = − 2 e 0 0
2 (β0 τ + 2) β0 τ + 2

is a family of solutions to the fifth order KdV equation.

Explicit examples of solutions (to the KK equation) obtained with the


help of Corollary 6.8 appear in [62].

6.3. Nonlocal Symmetries of the Camassa–Holm Equation


The final part of the paper is on nonlocal symmetries of the Camassa–Holm
equation with non-zero critical wave speed, Eq. (3.13), which is now written as

− m + uxx − u − κ = 0, −mt − mx u − 2 mux = 0. (6.38)

First of all, recall from Sect. 3 that the CH equation (6.38) admits a quadratic
pseudo-potential γ and a potential δ determined by
γ2
−γx + m − (1/2) + (1/2) λ = 0,
λ (6.39)
−γt − ux γ − γx λ − γx u + uxx λ = 0,

and

− δx + γ = 0, −δt + ux λ − γλ − u γ = 0 (6.40)

respectively. Furthermore, define a “nonlocal potential” β via


 
δ
−βx + m exp =0 (6.41)
λ
   
δ 2 δ
−βt − (1/2) exp γ + (1/2) exp λ2
λ λ
 
δ
−exp u m = 0, (6.42)
λ
following the reasoning of the last subsection, and consider also the equations

λx = 0 λt = 0. (6.43)

stating that λ is simply a parameter.


Vol. 60 (2011) Equations of Pseudo-Spherical Type 95

Theorem 6.9. The first order generalized symmetries of the augmented CH


system (6.38)–(6.43) are

V1 = (6.44)
∂t

V2 = (6.45)
∂x
∂ β ∂
V3 = + (6.46)
∂δ λ ∂β

V4 = (6.47)
∂β
 
δ/λ 2mγ ∂ ∂ ∂ ∂
V5 =e + mx + eδ/λ γ + eδ/λ m +β
λ ∂m ∂u ∂γ ∂δ
 
2δ/λ β2 ∂
+ e m+ (6.48)
2λ ∂β

∂ ∂ ∂ ∂
V6 =A m + (u + κ) +γ +δ
∂m ∂u ∂γ ∂δ

∂ ∂ ∂ ∂
+β +λ + κt −t (6.49)
∂β ∂λ ∂x ∂t
where A is an arbitrary function of λ. Consequently, these vector fields are
nonlocal symmetries of the CH equation (6.38).

Proposition 6.10. Assume that A and B arbitrary functions of λ and set


V6∗ = (B/A)V6 . The commutation table of the six nonlocal symmetries (6.44)–
(6.49) of the CH equation is

V1 V2 V3 V4 V5 V6
V1 A κV2 − A V1
V2
V3 − λ1 V4 1
λ V5 A V3
1
V4 λ V4 V3 A V4
V5 − λ1 V5 −V3
−λ
V6∗ −BκV2 + A V1 − B V3 − B V4 A (AB

− BA )V6

whenever u, m, γ, δ, β, λ are solutions of the augmented CH system (6.38)–


(6.43).

Theorem 6.9 has been found via extensive symbolic computations carried
out with the help of Mathematica software written by R. Hernández Heredero
(Universidad Politécnica de Madrid). It is based on earlier classification results
appearing in [33,34]. The function A(λ) appearing in (6.49) is of importance
96 E. G. Reyes Results. Math.

for the Lie algebra structure of nonlocal π-symmetries: Proposition 6.10 im-
plies that the nonlocal π-symmetries (6.44)–(6.48) generate a five-dimensional
Lie algebra G5 . More precisely, if instead of V3 , V4 , V5 we use
√ √
e = − 2λ V4 , f = 2λ V5 , h = −2λV3 ,
we find the commutators [h, e] = 2e, [h, f ] = −2f, [e, f ] = h, that is, e, f, h
generate a copy of sl(2), and therefore G5 is isomorphic to the direct sum of
sl(2) and the commutative Lie algebra generated by V1 and V2 . Now, if in
addition to V1 , . . . , V5 one considers also V6 , one can show that the Lie algebra
generated by the nonlocal symmetries V1 , . . . , V6 contains a semidirect sum of
the loop algebra over sl(2, R) and the Virasoro algebra! The paper finishes
with this observation:
Proposition 6.11. Set A(λ) = 1 and write L instead of V6 . Define the vector
fields
√ √
Tn1 = −2λn+1 V3 , Tn2 = − 2λ λn V4 , Tn3 = 2λ λn V5 ,
and
W0 = −L + κV2 , Wn = −λn L,
in which n ∈ Z and L = V6 with A(λ) = 1. Then, the following commutation
relations hold:
 1 2 2
 1 3 3
 2 3 1
Tm , Tn = 2 Tm+n , Tm , Tn = −2 Tm+n , Tm , Tn = Tm+n ;
 
 1 1
 2 1 2
Tm , Wn = m Tm+n , Tm , Wn = m − Tm+n ,
2
 
 3 1 3
Tm , Wn = m + Tm+n ;
2
[Wm , Wn ] = (m − n) Wm+n − 2κ m δm+n,0 V2 .

Acknowledgements
Support from the Fondo Nacional de Desarrollo Cientı́fico y Tecnológico
(FONDECYT) through Grants #1070191 and #1111042 is gratefully acknowl-
edged. Thanks are also due to Denis and Stepan for letting the author work
during the (southern hemisphere) summertime.

References
[1] Ablowitz, M., Kaup, D.J., Newell, A., Segur, H.: The inverse scattering
transform–Fourier analysis for nonlinear problems. Stud. Appl. Math. 53, 249–
315 (1974)
[2] Beals, R., Coifman, R.: Scattering and inverse scattering for first order systems.
Commun. Pure Appl. Math. 37, 39–90 (1984)
Vol. 60 (2011) Equations of Pseudo-Spherical Type 97

[3] Beals, R., Coifman, R.: Scattering and inverse scattering for first order systems,
II. Inverse Prob. 3, 577–593 (1987)
[4] Beals, R., Rabelo, M., Tenenblat, K.: Bäcklund transformations and inverse
scattering solutions for some pseudospherical surface equations. Stud. Appl.
Math. 81(2), 125–151 (1989)
[5] Beals, R., Sattinger, D.H., Szmigielski, J.: Multipeakons and the classical
moment problem. Adv. Math. 154, 229–257 (2000)
[6] Bryant, R., Griffiths, P.: Characteristic cohomology of differential systems, II:
Conservation laws for a class of parabolic equations. Duke Math. J. 78, 531–676
(1995)
[7] Bryant, R., Griffiths, P., Grossmann, D.: Exterior differential systems and
Euler–Lagrange partial differential equations. University of Chicago Press,
Chicago (2003)
[8] Camassa, R., Holm, D.D.: An integrable shallow water equation with peaked
solitons. Phys. Rev. Lett. 71(11), 1661–1664 (1993)
[9] Camassa, R., Holm, D.D., Hyman, J.M.: A new integrable shallow water equa-
tion. Adv. Appl. Mech. 31, 1–33 (1994)
[10] Cavalcante, J.A., Tenenblat, K.: Conservation laws for nonlinear evolution equa-
tions. J. Math. Phys. 29(4), 1044–1049 (1988)
[11] Chern, S.S., Peng, C.K.: Lie groups and KdV equations. Manuscripta
Math. 28(13), 207–217 (1979)
[12] Chern, S.S., Tenenblat, K.: Pseudo-spherical surfaces and evolution equa-
tions. Stud. Appl. Math. 74, 55–83 (1986)
[13] Constantin, A., Escher, J.: Global weak solutions for a shallow water equa-
tion. Indiana U. Math. J. 47, 1527–1545 (1998)
[14] Constantin, A., Escher, J.: Wave breaking for nonlinear nonlocal shallow water
equations. Acta Math. 181, 229–243 (1998)
[15] Constantin, A., Lannes, D.: The hydrodynamical relevance of the Camassa–
Holm and Degasperis–Procesi equations. Arch. Ration. Mech. Anal. 192, 165–186
(2009)
[16] Constantin, A., Strauss, W.: Stability of peakons. Commun. Pure Appl.
Math. 53, 603–610 (2000)
[17] Crampin, M.: Solitons and SL(2R). Phys. Lett. A 66, 170–172 (1978)
[18] Crampin, M., Pirani, F.A.E., Robinson, D.C.: The soliton connection. Lett.
Math. Phys. 2, 15–19 (1977)
[19] Dickey, L.A.: Soliton equations and Hamiltonian systems, 2nd edn. Advanced
Series in Mathematical Physics, vol. 26. World Scientific, New Jersey (2003)
[20] Ding, Q., Tenenblat, K.: On differential systems describing surfaces of constant
curvature. J. Diff. Equ. 184, 185–214 (2002)
[21] Dodd, R.K., Fordy, A.: The prolongation structures of quasi-polynomial
flows. Proc. R. Soc. Lond. 385, 389–429 (1983)
[22] Eisenhart, L.P.: A treatise on the differential geometry of curves and surfaces.
Ginn and Company, Boston (1909)
98 E. G. Reyes Results. Math.

[23] Fokas, A.S.: On a class of physically important integrable equations. Phys. D


87, 145–150 (1995)
[24] Fokas, A.S., Olver, P.J., Rosenau, P.: A plethora of integrable bi-Hamiltonian
equations. In: Fokas, A.S., Gel’fand, I.M. (eds.) Algebraic Aspects of Integra-
ble Systems: In Memory of Irene Dorfman. Progress in Nonlinear Differential
Equations, vol. 26, pp. 93–101. Birkhauser, Boston (1996)
[25] Foursov, M.V., Olver, P.J., Reyes, E.G.: On formal integrability of evolution
equations and local geometry of surfaces. Diff. Geom. Appl. 15(2), 183–199
(2001)
[26] Fuchssteiner, B., Fokas, A.S.: Symplectic structures, their Bäcklund transforma-
tions and hereditary symmetries. Phys. D 4, 47–66 (1981)
[27] Fuchssteiner, B.: The Lie algebra structure of degenerate hamiltonian and
bi-hamiltonian systems. Prog. Theor. Phys. 68, 1082–1104 (1982)
[28] Fuchssteiner, B.: Some tricks from the symmetry-toolbox for nonlinear equa-
tions: generalizations of the Camassa–Holm equation. Phys. D 95, 229–243
(1996)
[29] Galas, F.: New nonlocal symmetries with pseudopotentials. J. Phys. A
25(15), L981–L986 (1992)
[30] Gomes Neto, V.P.: Fifth-order evolution equations describing pseudospherical
surfaces. J. Diff. Equ. 249, 2822–2865 (2010)
[31] Górka, P., Reyes, E.G.: The modified Camassa–Holm equation. Int. Math. Res.
Notices (2010). doi:10.1093/imrn/rnq163
[32] Gürses, M., Nutku, Y.: New nonlinear equations from surface theory. J. Math.
Phys. 22, 1393–1398 (1981)
[33] Hernandez Heredero, R., Reyes, E.G.: Nonlocal symmetries and a Darboux
transformation for the Camassa–Holm equation. J. Phys. A Math. Theor. 42,
182002 (2009) (Fast Track Communication, 9 pp)
[34] Hernandez Heredero, R., Reyes, E.G.: Geometric integrability of the Camassa–
Holm equation. II. Int. Math. Res. Notices (2011). doi:10.1093/imrn/RNR120
[35] Ivey, T.A., Landsberg, J.M.: Cartan for Beginners: Differential Geometry via
Moving Frames and Exterior Differential Systems. AMS, Providence (2003)
[36] Johnson, R.S.: Camassa–Holm, Korteweg–de Vries and related models for water
waves. J. Fluid Mech. 455, 63–82 (2002)
[37] Jorge, L., Tenenblat, K.: Linear problems associated to evolution equations of
type utt = F (u, ux , uxx , ut ). Stud. Appl. Math. 77, 103–117 (1987)
[38] Kamran, N., Tenenblat, K.: On differential equations describing pseudospherical
surfaces. J. Diff. Equ. 115(1), 75–98 (1995)
[39] Kaup, D.J.: On the inverse scattering problem for cubic eigenvalue problems of
the class ψxxx + 6qψx + 6rψ = λψ. Stud. Appl. Math. 62, 189–216 (1980)
[40] Krasil’shchik, I.S., Vinogradov, A.M. (eds.) Symmetries and Conservation
Laws for Differential Equations of Mathematical Physics. Translations of
Mathematical Monographs, vol. 182. AMS, Providence (1999)
[41] Lenells, J.: Stability of periodic peakons. Int. Math. Res. Notices 10, 485–499
(2004)
Vol. 60 (2011) Equations of Pseudo-Spherical Type 99

[42] Lenells, J.: Traveling wave solutions of the Camassa–Holm equation. J. Diff.
Equ. 217, 393–430 (2005)
[43] Li, Y.A., Olver, P.J.: Well-posedness and blow-up solutions for an integrable
nonlinearly dispersive model wave equation. J. Diff. Equ. 162(1), 27–63 (2000)
[44] Marvan, M.: On zero-curvature representations of partial differential equations.
In: Kowalski, O., Krupka, D. (eds.) Differential Geometry and its Applications,
pp. 103–122. Silesian University, Opava (1993). (ELibEMS, http://www.emis.
de/proceedings/5ICDGA)
[45] McIntosh, C.B.G.: Equations with soliton solutions and the pseudosphere. Lett.
Math. Phys. 5, 105–112 (1981)
[46] Misiolek, G.: A shallow water equation as a geodesic flow on the Bott–Virasoro
group. J. Geom. Phys. 24, 203–208 (1998)
[47] Mikhailov, A.V., Shabat, A.B., Sokolov, V.V.: The symmetry approach to clas-
sification of integrable equations. In: Zakharov, V.E. (ed.) What is Integrability?
pp. 115–184. Springer, New York (1991)
[48] Molino, P.: Systèmes différentiels extérieurs et équations d’évolution. In: Elective
Course in Differential Geometry, 1980–1981, pp. 70–218. University of Science
and Technology, Languedoc, Montpellier (1981)
[49] Olver, P.J.: Applications of Lie Groups to Differential Equations, 2nd edn.
Springer, New York (1993)
[50] Olver, P.J.: Invariant submanifold flows. J. Phys. A 41, 344017 (2008)
[51] Rabelo, M.: On equations which describe pseudospherical surfaces. Stud. Appl.
Math. 81, 221–248 (1989)
[52] Rabelo, M., Tenenblat, K.: On equations of type uxt = F (u, ux ) which describe
pseudospherical surfaces. J. Math. Phys. 31, 1400–1407 (1990)
[53] Rabelo, M., Tenenblat, K.: A classification of pseudospherical surface equations
of type ut = uxxx + G(u, ux , uxx ). J. Math. Phys. 33, 537–549 (1992)
[54] Reyes, E.G.: Pseudo-spherical surfaces and integrability of evolution equations.
J. Diff. Equ. 147(1), 195–230 (1998). Erratum: J. Diff. Equ. 153(1), 223–224
(1999)
[55] Reyes, E.G.: Conservation laws and Calapso–Guichard deformations of equa-
tions describing pseudo-spherical surfaces. J. Math. Phys. 41, 2968–2989 (2000)
[56] Reyes, E.G.: Some geometric aspects of integrability of differential equations in
two independent variables. Acta Appl. Math. 64(2/3), 75–109 (2000)
[57] Reyes, E.G.: On generalized Bäcklund transformations for equations describing
pseudo-spherical surfaces. J. Geom. Phys. 45(3–4), 368–392 (2003)
[58] Reyes, E.G.: Transformations of solutions for equations and hierarchies of
pseudo-spherical type. J. Phys. A Math. Gen. 36(8), L125–L132 (2003)
[59] Reyes, E.G.: Correspondence theorems for hierarchies of equations of pseudo-
spherical type. J. Diff. Equ. 225, 26–56 (2006)
[60] Reyes, E.G.: Geometric integrability of the Camassa–Holm equation. Lett. Math.
Phys. 59(2), 117–131 (2002)
[61] Reyes, E.G.: The soliton content of the Camassa–Holm and Hunter–Saxton
equations. In: Nikitin, A.G., Boyko, V.M., Popovych, R.O. (eds.) Proceedings of
100 E. G. Reyes Results. Math.

the Fourth International Conference on Symmetry in Nonlinear Mathematical


Physics, 2001, vol. 43, pp. 201–208. Proceedings of the Institute of Mathematics
of the NAS of Ukraine, Ukraine (2002)
[62] Reyes, E.G.: Nonlocal symmetries and the Kaup–Kupershmidt equation.
J. Math. Phys. 46(7), 073507 (2005)
[63] Reyes, E.G.: Pseudo-potentials, nonlocal symmetries, and integrability of some
shallow water equations. Select. Math. (New Series) 12, 241–270 (2006)
[64] Rogers, C., Schief, W.K.: Bäcklund and Darboux transformations. Geome-
try and Modern Applications in Soliton Theory. Cambridge University Press,
Cambridge (2002)
[65] Sakovich, S.Yu.: Cyclic bases of zero-curvature representations: five illustrations
to one concept. Acta Appl. Math. 83, 69–83 (2004)
[66] Sakovich, A., Sakovich, S.: The short pulse equation is integrable. J. Phys. Soc.
Japan 74, 239–241 (2005)
[67] Sakovich, A., Sakovich, S.: Solitary wave solutions of the short pulse equation.
J. Phys. A Math. Gen. 39, L361–L367 (2006)
[68] Sakovich, A., Sakovich, S.: On transformations of the Rabelo equations. SIGMA
3, 086, 8 (2007)
[69] Sanders, J.A., Wang, J.-P.: On the integrability of homogeneous scalar evolution
equations. J. Diff. Equ. 147, 410–434 (1998)
[70] Sasaki, R.: Soliton equations and pseudospherical surfaces. Phys. B 154, 343–357
(1979)
[71] Sasaki, R.: Pseudopotentials for the general AKNS system. Phys. Lett. A 73, 77–
80 (1979)
[72] Sawada, K., Kotera, T.: A method for finding N -soliton solutions of the KdV
equation and KdV-like equation. Prog. Theor. Phys. 51, 1355–1367 (1974)
[73] Schäfer, T., Wayne, C.E.: Propagation of ultra-short optical pulses in cubic
nonlinear media. Phys. D 196, 90–105 (2004)
[74] Schiff, J.: The Camassa–Holm equation: a loop group approach. Phys. D 121, 24–
43 (1998)
[75] Shchepetilov, A.V.: The geometric sense of the Sasaki connection. J. Phys. A
Math. Gen. 36, 3893–3898 (2003)
[76] Svinolupov, S.I., Sokolov, V.V.: Weak nonlocalities in evolution equations. Math.
Notes 48, 1234–1239 (1991)
[77] Takens, F.: A global version of the inverse problem in the calculus of variations.
J. Diff. Geom. 14, 543–562 (1979)
[78] Tenenblat, K.: Transformations of Manifolds and Applications to Differential
Equations. Addison Wesley Longman, England (1998)
[79] Vinogradov, A.M., Krasil’shchik, I.S.: A method of calculating higher symme-
tries of nonlinear evolutionary equations, and nonlocal symmetries. Dokl. Akad.
Nauk SSSR 253(6), 1289–1293 (1980)
[80] Wahlquist, H.D., Eastbrook, F.B.: Prolongation structures of nonlinear equa-
tions. J. Math. Phys. 16, 1–7 (1975)
Vol. 60 (2011) Equations of Pseudo-Spherical Type 101

[81] Wu, H.: Foliations on constant curvature surfaces and nonlinear partial differ-
ential equations. Houston J. Math. 24, 65–84 (1998)
[82] Wyller, J.: Nonlocal conservation laws of the DNLS-equation. Phys. Scripta 40,
717–720 (1989)
Enrique G. Reyes
Departamento de Matemática y Ciencia de la Computación
Universidad de Santiago de Chile
Casilla 307 Correo 2
Santiago, Chile
e-mail: enrique.reyes@usach.cl;
ereyes@fermat.usach.cl

Received: February 1, 2011.


Revised: May 23, 2011.
Accepted: May 24, 2011.

You might also like