Research Article Approximate Hamiltonian Symmetry Groups and Recursion M. Nadjafikhah

advertisement
Hindawi Publishing Corporation
Advances in Mathematical Physics
Volume 2013, Article ID 568632, 9 pages
http://dx.doi.org/10.1155/2013/568632
Research Article
Approximate Hamiltonian Symmetry Groups and Recursion
Operators for Perturbed Evolution Equations
M. Nadjafikhah1 and A. Mokhtary2
1
2
School of Mathematics, Iran University of Science and Technology, Narmak, Tehran 1684613114, Iran
Department of Complementary Education, Payame Noor University, Tehran 19395-3697, Iran
Correspondence should be addressed to A. Mokhtary; rdmokhtary@gmail.com
Received 11 October 2012; Revised 18 December 2012; Accepted 1 January 2013
Academic Editor: Rutwig Campoamor-Stursberg
Copyright © 2013 M. Nadjafikhah and A. Mokhtary. This is an open access article distributed under the Creative Commons
Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is
properly cited.
The method of approximate transformation groups, which was proposed by Baikov et al. (1988 and 1996), is extended on
Hamiltonian and bi-Hamiltonian systems of evolution equations. Indeed, as a main consequence, this extended procedure is
applied in order to compute the approximate conservation laws and approximate recursion operators corresponding to these
types of equations. In particular, as an application, a comprehensive analysis of the problem of approximate conservation laws
and approximate recursion operators associated to the Gardner equation with the small parameters is presented.
1. Introduction
The investigation of the exact solutions of nonlinear evolution
equations has a fundamental role in the nonlinear physical
phenomena. One of the significant and systematic methods for obtaining special solutions of systems of nonlinear
differential equations is the classical symmetries method,
also called Lie group analysis. This well-known approach
originated at the end of nineteenth century from the pioneering work of Lie [1]. The fact that symmetry reductions for
many PDEs cannot be determined, via the classical symmetry
method, motivated the creation of several generalizations of
the classical Lie group approach for symmetry reductions.
Consequently, several alternative reduction methods have
been proposed, going beyond Lie’s classical procedure and
providing further solutions. One of these techniques which
is extremely applied particularly for nonlinear problems is
perturbation analysis. It is worth mentioning that sometimes
differential equations which appear in mathematical modelings are presented with terms involving a parameter called
the perturbed term. Because of the instability of the Lie point
symmetries with respect to perturbation of coefficients of
differential equations, a new class of symmetries has been
created for such equations, which are known as approximate
(perturbed) symmetries. In the last century, in order to have
the utmost result from the methods, combination of Lie symmetry method and perturbations are investigated and two
different so-called approximate symmetry methods (ASMs)
have been developed. The first method is due to Baikov et al.
[2, 3]. The second procedure was proposed by Fushchich and
Shtelen [4] and later was followed by Euler et al. [5, 6]. This
method is generally based on the perturbation of dependent
variables. In [7, 8], a comprehensive comparison of these two
methods is presented.
As it is well known, Hamiltonian systems of differential
equations are one of the famous and significant concepts in
physics. These important systems appear in the various fields
of physics such as motion of rigid bodies, celestial mechanics,
quantization theory, fluid mechanics, plasma physics, and
so forth. Due to the significance of Hamiltonian structures,
in this paper, by applying the linear behavior of the Euler
operator, characteristics, prolongation, and Fréchet derivative
of vector fields, we have extended ASM on the Hamiltonian
and bi-Hamiltonian systems of evolution equations, in order
to investigate the interplay between approximate symmetry
groups, approximate conservation laws, and approximate
recursion operators.
2
Advances in Mathematical Physics
The structure of the present paper is as follows. In
Section 2, some necessary preliminaries regarding to the
Hamiltonian structures are presented. In Section 3, a comprehensive investigation of the approximate Hamiltonian
symmetries and approximate conservation laws associated
to the perturbed evolution equations is proposed. Also, as
an application of this procedure, approximate Hamiltonian
symmetry groups, approximate bi-Hamiltonian structures,
and approximate conservation laws of the Gardner equation
are computed. In Section 4, the approximate recursion operators are studied and the proposed technique is implemented
for the Gardner equation as an application. Finally, some
concluding remarks are mentioned at the end of the paper.
2. Preliminaries
In this section, we will mention some necessary preliminaries
regarding Hamiltonian structures. In order to be familiar
with the general concepts of the ASM, refer to [9]. It is
also worth mentioning that most of this paper’s definitions,
theorems and techniques regarding Hamiltonian and biHamiltonian structures are inspired from [10].
Let 𝑀 ⊂ 𝑋 × π‘ˆ denote a fixed connected open subset
of the space of independent and dependent variables π‘₯ =
(π‘₯1 , . . . , π‘₯𝑝 ) and 𝑒 = (𝑒1 , . . . , π‘’π‘ž ). The algebra of differential
functions 𝑃(π‘₯, 𝑒(𝑛) ) = 𝑝[𝑒] over 𝑀 is denoted by A.
We further define A𝑙 to be the vector space of 𝑙-tuples of
differential functions, 𝑃[𝑒] = (𝑃1 [𝑒], . . . , 𝑃𝑙 [𝑒]), where each
𝑃𝑗 ∈ A.
A generalized vector field will be a (formal) expression of
the form
𝑝
v = ∑ πœ‰π‘– [𝑒]
𝑖=1
π‘ž
πœ•
πœ•
+ ∑ πœ™ [𝑒] 𝛼
πœ•π‘₯𝑖 𝛼=1 𝛼
πœ•π‘’
(1)
in which πœ‰π‘– and πœ™π›Ό are smooth differential functions. The
Prolonged generalized vector field can be defined as follows:
(n)
pr
π‘ž
v=v+∑ ∑
𝛼=1 ♯𝐽≤𝑛
πœ™π›Όπ½
πœ•
[𝑒] 𝛼 ,
πœ•π‘’π½
(2)
whose coefficients are determined by the formula
𝑝
𝑝
𝑖=1
𝑖=1
𝛼
πœ™π›Όπ½ = 𝐷𝐽 (πœ™π›Ό − ∑ πœ‰π‘– 𝑒𝑖𝛼 ) + ∑ πœ‰π‘– 𝑒𝐽,𝑖
,
(3)
with the same notation as before. Given a generalized vector
field v, its infinite prolongation (or briefly prolongation) is the
formally infinite sum as follows:
𝑝
pr v = ∑ πœ‰π‘–
𝑖=1
π‘ž
πœ•
πœ•
+ ∑ ∑ πœ™π½
,
πœ•π‘₯𝑖 𝛼=1 𝐽 𝛼 πœ•π‘’π½π›Ό
(4)
where each πœ™π›Όπ½ is given by (3), and the sum in (4) now extends
over all multi-indices 𝐽 = (𝑗1 , . . . , π‘—π‘˜ ) for π‘˜ ≥ 0, 1 ≤ π‘—π‘˜ ≤ 𝑝.
A generalized vector field v is a generalized infinitesimal
symmetry of a system of differential equations as follows:
Δ πœˆ [𝑒] = Δ πœˆ (π‘₯, 𝑒(𝑛) ) = 0,
𝜈 = 1, . . . , 𝑙,
(5)
if and only if
pr v [Δ πœˆ ] = 0,
𝜈 = 1, . . . , 𝑙,
(6)
for every smooth solution 𝑒 = 𝑓(π‘₯).
Among all the generalized vector fields, those in which
the coefficients πœ‰π‘– [𝑒] of the πœ•/πœ•π‘₯𝑖 are zero play a distinguished
role. Let 𝑄[𝑒] = (𝑄1 [𝑒], . . . , π‘„π‘ž [𝑒]) ∈ Aπ‘ž be a π‘ž-tuple of
differential functions. The generalized vector field
π‘ž
v𝑄 = ∑ 𝑄𝛼 [𝑒]
𝛼=1
πœ•
πœ•π‘’π›Ό
(7)
is called an evolutionary vector field, and 𝑄 is called its
characteristic.
A manifold 𝑀 with a Poisson bracket is called a Poisson
manifold, the bracket defining a Poisson structure on 𝑀. Let
𝑀 be a Poisson manifold and 𝐻 : 𝑀 → R be a smooth
function. The Hamiltonian vector field associated with 𝐻
is the unique smooth vector field vΜ‚H on 𝑀 satisfying the
following identity:
vΜ‚H = {𝐹, 𝐻} = − {𝐻, 𝐹}
(8)
for every smooth function 𝐹 : 𝑀 → R. The equations
governing the flow of vΜ‚H are referred to as Hamilton’s
equations for the “Hamiltonian” function 𝐻.
Let π‘₯ = (π‘₯1 , . . . , π‘₯π‘š ) be local coordinates on 𝑀 and 𝐻(π‘₯)
be a real-valued function. The following basic formula can be
obtained for the Poisson bracket:
π‘š π‘š
{𝐹, 𝐻} = ∑ ∑ {π‘₯𝑖 , π‘₯𝑗 }
𝑖=1 𝑗=1
πœ•πΉ πœ•π»
.
πœ•π‘₯𝑖 πœ•π‘₯𝑗
(9)
In other words, in order to compute the Poisson bracket of
any pair of functions in some given set of local coordinates, it
suffices to know the Poisson brackets between the coordinate
functions themselves. These basic brackets,
𝐽𝑖𝑗 (π‘₯) = {π‘₯𝑖 , π‘₯𝑗 } ,
𝑖, 𝑗 = 1, . . . , π‘š,
(10)
are called the structure functions of the Poisson manifold 𝑀
relative to the given local coordinates and serve to uniquely
determine the Poisson structure itself. For convenience, we
assemble the structure functions into a skew-symmetric π‘š ×
π‘š matrix 𝐽(π‘₯), called the structure matrix of 𝑀. Using ∇𝐻
to denote the (column) gradient vector for 𝐻, the local
coordinate form (9) for the Poisson bracket can be written
as
{𝐹, 𝐻} = ∇𝐹 ⋅ 𝐽∇𝐻.
(11)
Therefore, in the given coordinate chart, Hamilton’s
equations take the form of
𝑑π‘₯
= 𝐽 (π‘₯) ∇𝐻 (π‘₯) .
𝑑𝑑
(12)
Alternatively, using (9), we could write this in the “bracket
form” as follows:
𝑑π‘₯
= {π‘₯, 𝐻} ,
𝑑𝑑
(13)
Advances in Mathematical Physics
3
the 𝑖th component of the right-hand side being {π‘₯𝑖 , 𝐻}. A
system of first-order ordinary differential equations is said to
be a Hamiltonian system if there is a Hamiltonian function
𝐻(π‘₯) and a matrix of functions 𝐽(π‘₯) determining a Poisson
bracket (11) whereby the system takes the form (12).
If
D = ∑ 𝑃𝐽 [𝑒] 𝐷𝐽 ,
𝐽
𝑃𝐽 ∈ A
(14)
is a differential operator, its (formal) adjoint is the differential
operator D∗ which satisfies
∫ 𝑃 ⋅ D𝑄 𝑑π‘₯ = ∫ 𝑄 ⋅ D∗ 𝑃 𝑑π‘₯
Ω
Ω
(15)
for every pair of differential functions 𝑃, 𝑄 ∈ A which vanish
when 𝑒 ≡ 0. Also, for every domain Ω ⊂ R𝑝 and every
function 𝑒 = 𝑓(π‘₯) of compact support in Ω. An operator D
is self-adjoint if D∗ = D; it is skew-adjoint if D∗ = −D.
The principal innovations needed to convert a Hamiltonian system of ordinary differential equations (12) to a
Hamiltonian system of evolution equations are as follows
(refer to [10] for more details):
(i) replacing the Hamiltonian function 𝐻(π‘₯) by a Hamiltonian functional H[𝑒],
(ii) replacing the vector gradient operation ∇𝐻 by the
variational derivative 𝛿H of the Hamiltonian functional, and
(iii) replacing the skew-symmetric matrix 𝐽(π‘₯) by a skewadjoint differential operator D which may depend on
𝑒.
The resulting Hamiltonian system will take the form of
πœ•π‘’
= D ⋅ 𝛿H [𝑒] .
πœ•π‘‘
(16)
3. Approximate Hamiltonian Symmetries and
Approximate Conservation Laws
Consider a system of perturbed evolution equations:
πœ•π‘’
= 𝑃 [𝑒, πœ€]
πœ•π‘‘
(19)
in which 𝑃[𝑒, πœ€] = 𝑃(π‘₯, 𝑒(𝑛) , πœ€) ∈ Aπ‘ž , π‘₯ ∈ R𝑝 , 𝑒 ∈ Rπ‘ž and πœ€
is a parameter.
Substituting according to (19) and its derivatives, we see
that any evolutionary symmetry must be equivalent to one
whose characteristic 𝑄[𝑒, πœ€] = 𝑄(π‘₯, 𝑑, 𝑒(π‘š) , πœ€) depends only
on π‘₯, 𝑑, 𝑒, πœ€ and the π‘₯-derivatives of 𝑒. On the other hand,
(19) itself can be considered as the equations corresponding
to the flow exp(𝑑v𝑝 ) of the evolutionary vector field with
characteristic 𝑃. The symmetry criterion (6), which in this
case is
𝐷𝑑 π‘„πœˆ = pr v𝑄 (π‘ƒπœˆ ) + π‘œ (πœ€π‘ ) ,
𝜈 = 1, . . . , π‘ž,
(20)
can be readily seen to be equivalent to the following Lie
bracket condition on the two approximate generalized vector
fields. Indeed, this point generalizes the correspondence
between symmetries of systems of first-order perturbed
ordinary differential equations and the Lie bracket of the
corresponding vector fields.
Considering the above assumptions, some useful relevant
theorems and definitions could be rewritten as follows.
Proposition 2. An approximate evolutionary vector field v𝑄
is a symmetry of the system of perturbed evolution equations
𝑒𝑑 = 𝑃[𝑒, πœ€] if and only if
πœ•v𝑄
+ [v𝑃 , v𝑄] = π‘œ (πœ€π‘ )
πœ•π‘‘
(21)
Clearly, for a candidate Hamiltonian operator D the correct
expression for the corresponding Poison bracket has the form
of
holds identically in (π‘₯, 𝑑, 𝑒(π‘š) , πœ€). (Here πœ•v𝑄/πœ•π‘‘ denotes the
evolutionary vector field with characteristic πœ•π‘„/πœ•π‘‘.).
{P, L} = ∫ 𝛿P ⋅ D𝛿L 𝑑π‘₯,
Any approximate conservation law of a system of perturbed evolution equations takes the form of
(17)
whenever P, L ∈ F are functionals. Off course, the Hamiltonian operator D must satisfy certain further restrictions in
order that (17) be a true Poisson bracket. A linear operator
D : Aπ‘ž → Aπ‘ž is called Hamiltonian if its Poisson bracket
(17) satisfies the conditions of skew-symmetry and the Jacobi
identity.
Proposition 1. Let D be a Hamiltonian operator with Poisson
bracket (17). To each functional H = ∫ 𝐻 𝑑π‘₯ ∈ F, there is an
evolutionary vector field pr vΜ‚H , called the Hamiltonian vector
field associated with H, which for all functionals P ∈ F
satisfies the following identity:
pr vΜ‚H (P) = {P, H} .
(18)
Indeed, vΜ‚H has characteristic D𝛿H = DE(𝐻), in which E is
Euler operator (Proposition 7.2 of [10]).
𝐷𝑑 𝑇 + Div 𝑋 = π‘œ (πœ€π‘ ) ,
(22)
in which Div denotes spatial divergence. Without loss of generality, the conserved density 𝑇(π‘₯, 𝑑, 𝑒(𝑛) , πœ€) can be assumed to
depend only on π‘₯-derivatives of 𝑒. Equivalently, for Ω ⊂ 𝑋,
the functional
T [𝑑; 𝑒, πœ€] = ∫ 𝑇 (π‘₯, 𝑑, 𝑒(𝑛) , πœ€) 𝑑π‘₯
Ω
(23)
is a constant, independent of 𝑑, for all solutions 𝑒 such that
𝑇(π‘₯, 𝑑, 𝑒(𝑛) , πœ€) → 0 as π‘₯ → πœ•Ω. Note that if 𝑇(π‘₯, 𝑑, 𝑒(𝑛) , πœ€)
is any such differential function, and 𝑒 is a solution of the
perturbed evolutionary system 𝑒𝑑 = 𝑃[𝑒, πœ€], then
𝐷𝑑 𝑇 ≈ πœ•π‘‘ 𝑇 + pr v𝑝 (𝑇) ,
(24)
4
Advances in Mathematical Physics
where πœ•π‘‘ = πœ•/πœ•π‘‘ denotes the partial 𝑑-derivative. Thus 𝑇 is the
density for a conservation law of the system if and only if its
associated functional T satisfies the following identity:
πœ•T
+ pr v𝑝 (T) = π‘œ (πœ€π‘ ) .
πœ•π‘‘
(25)
In the case that our system is of Hamiltonian form, the
bracket relation (18) immediately leads to the Noether
relation between approximate Hamiltonian symmetries and
approximate conservation laws.
Definition 3. Let D be a π‘ž × π‘ž approximate Hamiltonian
differential operator. An approximate distinguished functional
for D is a functional G ∈ F satisfying D𝛿G = π‘œ(πœ€π‘ ) for all
π‘₯, 𝑒.
In other words, the Hamiltonian system corresponding to
a distinguished functional is completely trivial: 𝑒𝑑 = 0.
Now, according to [10], the perturbed Hamiltonian version of Noether’s theorem can be presented as follows.
Theorem 4. Let 𝑒𝑑 = D𝛿H be a Hamiltonian system of
perturbed evolution equations. An approximate Hamiltonian
vector field vΜ‚P with characteristic D𝛿P, P ∈ F determines an
approximate generalized symmetry group of the system if and
Μƒ ≈ P − G differing
only if there is an equivalent functional P
only from P by a time-dependent approximate distinguished
Μƒ determines an approximate
functional G[𝑑; 𝑒, πœ€], such that P
conservation law.
(26)
can in fact be written in Hamiltonian form in two distinct
ways. Firstly, we see
2
3
𝑒𝑑 = 𝐷π‘₯ (3𝑒 + 2πœ€π‘’ − 𝑒π‘₯π‘₯ ) = D𝛿H1 ,
(27)
where D = 𝐷π‘₯ and
𝑒2
πœ€
H1 [𝑒, πœ€] = ∫ (𝑒3 + 𝑒4 + π‘₯ ) 𝑑π‘₯
2
2
(28)
is an approximate conservation law. Note that D is certainly
skew-adjoint and Hamiltonian. The Poisson bracket is
{P, L} = ∫ 𝛿P ⋅ 𝐷π‘₯ (𝛿L) 𝑑π‘₯.
(29)
The second Hamiltonian form is
𝑒𝑑 = (4𝑒𝐷π‘₯ + 2𝑒π‘₯ + 3πœ€ (𝑒𝑒π‘₯ + 𝑒2 𝐷π‘₯ ) − 𝐷π‘₯3 ) 𝑒 = E𝛿H0 ,
(30)
in which
1
H0 [𝑒, πœ€] = ∫ 𝑒2 𝑑π‘₯
2
In [11], we have comprehensively analyzed the problem
of approximate symmetries for the Gardner equation. We
have shown that the approximate symmetries of the Gardner
equation are given by the following generators:
v1 = πœ•π‘₯ ,
v2 = πœ•π‘‘ ,
v3 = 6π‘‘πœ•π‘₯ + (2πœ€π‘’ − 1) πœ•π‘’ ,
v4 = πœ€v1 ,
v5 = πœ€v2 ,
(32)
v6 = πœ€ (6π‘‘πœ•π‘₯ − πœ•π‘’ ) = πœ€v3 ,
v7 = πœ€ (π‘₯πœ•π‘₯ + 3π‘‘πœ•π‘‘ − 2π‘’πœ•π‘’ ) ,
with corresponding characteristics
𝑄1 = 𝑒π‘₯ ,
𝑄2 = 6 (𝑒 + πœ€π‘’2 ) 𝑒π‘₯ − 𝑒π‘₯π‘₯π‘₯ ,
𝑄3 = 6𝑑𝑒π‘₯ + 1 − 2πœ€π‘’,
𝑄4 = πœ€π‘„1 = πœ€π‘’π‘₯ ,
(33)
𝑄5 = πœ€π‘„2 = πœ€ (6𝑒𝑒π‘₯ − 𝑒π‘₯π‘₯π‘₯ ) ,
Example 5. The Gardner equation
𝑒𝑑 = 6 (𝑒 + πœ€π‘’2 ) 𝑒π‘₯ − 𝑒π‘₯π‘₯π‘₯ ,
E is skew-adjoint and satisfies the Jacobi identity. Therefore
it is Hamiltonian.
(31)
𝑄6 = πœ€π‘„3 = πœ€ (6𝑑𝑒π‘₯ + 1) ,
𝑄7 = πœ€ (2𝑒 + π‘₯𝑒π‘₯ + 3𝑑 (6𝑒𝑒π‘₯ − 𝑒π‘₯π‘₯π‘₯ )) ,
(up to sign).
For the first Hamiltonian operator D = 𝐷π‘₯ , there is one
independent nontrivial approximate distinguished functional, the mass P0 = M = ∫ 𝑒𝑑π‘₯ which is consequently
approximately conserved.
For the above seven characteristics, we have
𝑄𝑖 ≈ 𝐷π‘₯ 𝛿P𝑖 ,
𝑖 = 1, 2, 4, 5, 6,
(34)
with the following approximately conserved functionals:
1
P1 = H0 [𝑒, πœ€] = ∫ 𝑒2 𝑑π‘₯,
2
πœ€
1
P2 = H1 [𝑒, πœ€] = ∫ (𝑒3 + 𝑒4 + 𝑒π‘₯2 ) 𝑑π‘₯,
2
2
πœ€
P4 = πœ€P1 = ∫ 𝑒2 𝑑π‘₯,
2
1
P5 = πœ€P2 = πœ€ ∫ (𝑒3 + 𝑒π‘₯2 ) 𝑑π‘₯,
2
P6 = πœ€ ∫ (3𝑑𝑒2 + π‘₯𝑒) 𝑑π‘₯.
(35)
Advances in Mathematical Physics
5
For the second Hamiltonian operator E = 4𝑒𝐷π‘₯ + 2𝑒π‘₯ +
3πœ€(𝑒𝑒π‘₯ + 𝑒2 𝐷π‘₯ ) − 𝐷π‘₯3 ,
̃𝑖 ,
𝑄𝑖 ≈ E𝛿P
𝑖 = 2, 4, 5, 7,
(36)
the following approximately conserved functionals are the
corresponding approximate conservation laws:
Μƒ 5 = πœ€P
Μƒ 2 = πœ€ ∫ 1 𝑒2 𝑑π‘₯,
P
2
(37)
In this case, nothing new is obtained. Note that the other
approximate conservation law P5 did not arise from one
of the geometrical symmetries. According to Theorem 4,
however, there is an approximate Hamiltonian symmetry
which gives rises to it, namely vΜ‚P5 . The characteristic of this
approximate generalized symmetry is
≈ πœ€ (𝑒π‘₯π‘₯π‘₯π‘₯π‘₯ − 10𝑒𝑒π‘₯π‘₯π‘₯ − 20𝑒π‘₯ 𝑒π‘₯π‘₯ + 30𝑒2 𝑒π‘₯ ) .
(38)
Note that 𝑄5 happens to satisfy the Hamiltonian condition
(34) for D with the following functional:
P5 =
πœ€
2
− 5𝑒2 𝑒π‘₯π‘₯ + 5𝑒4 ) 𝑑π‘₯.
∫ (𝑒π‘₯π‘₯
2
𝑑 󡄨󡄨󡄨󡄨
󡄨 𝑃 [𝑒 + πœ–π‘„ [𝑒, πœ€]]
π‘‘πœ– σ΅„¨σ΅„¨σ΅„¨πœ–=0
(40)
Proposition 8. If 𝑃 ∈ Aπ‘Ÿ and 𝑄 ∈ Aπ‘ž then
Μƒ 7 = 1 P6 = πœ€ ∫ (3𝑑𝑒2 + π‘₯𝑒) 𝑑π‘₯.
P
2
2
𝑄5 ≈ E𝛿P5 = Eπœ€ (3𝑒2 − 𝑒π‘₯π‘₯ )
D𝑃 (𝑄) =
for any 𝑄 ∈ Aπ‘ž .
Μƒ 2 = P1 = ∫ 1 𝑒2 𝑑π‘₯,
P
2
Μƒ 4 = πœ€ P0 = πœ€ ∫ 𝑒 𝑑π‘₯,
P
2
2
perturbed differential operator D𝑃 : Aπ‘ž → Aπ‘Ÿ defined so
that
D𝑃 (𝑄) ≈ pr v𝑄 (𝑃) .
(41)
Theorem 9. Suppose that Δ[𝑒, πœ€] = 0 be a system of perturbed
differential equations. If R : Aπ‘ž → Aπ‘ž is a linear operator
such that for all solutions 𝑒 of Δ,
Μƒ ⋅ DΔ ,
DΔ ⋅ R ≈ R
(42)
where R : Aπ‘ž → Aπ‘ž is a linear differential operator, then R
is an approximate recursion operator for the system.
Suppose that Δ[𝑒, πœ€] = 𝑒𝑑 − 𝐾[𝑒, πœ€] is a perturbed evolution equation. Then DΔ = 𝐷𝑑 − D𝐾 . If R is an approximate
recursion operator, then it is not hard to observe that the
Μƒ in (42) must be the same as R. Therefore,
operator R
the condition (42) in this case reduces to the commutator
condition
(39)
Consequently, another approximate conservation law is provided for the Gardner equation.
Keeping on this procedure recursively, further approximate conservation laws could be generated. But, this procedure will be done in the next section by applying approximate
recursion operators.
4. Approximate Recursion Operators
Definition 6. Let Δ be a system of perturbed differential equations. An approximate recursion operator for Δ is a linear operator R : Aπ‘ž → Aπ‘ž in the space of π‘ž-tuples of differential
functions with the property that whenever v𝑄 is an approxiΜƒ ≈ R𝑄.
mate evolutionary symmetry of Δ, so v𝑄̃ is with 𝑄
For nonlinear perturbed systems, there is an analogous
criterion for a differential operator to be an approximate
recursion operator, but to state it we need to introduce the
notion of the (formal) Fréchet derivative of a differential
function.
Definition 7. Let 𝑃[𝑒, πœ€] = 𝑃(π‘₯, 𝑒(𝑛) , πœ€) ∈ Aπ‘Ÿ be an π‘Ÿ-tuple
of differential functions. The Fréchet derivative of 𝑃 is the
R𝑑 ≈ [D𝐾 , R]
for an approximate recursion operator of a perturbed evolution equation.
From (4), we can conclude that if R is an approximate
recursion operator, then for all 𝑙 ≥ 1 in which πœ€π‘™ R =ΜΈ 0, πœ€π‘™ R is
an approximate recursion operator as follows:
(πœ€π‘™ R)𝑑 = πœ€π‘™ R𝑑 ≈ πœ€π‘™ [D𝐾 , R] ≈ [D𝐾 , πœ€π‘™ R] .
In order to illustrate the significance of the above theorem, we
discuss a couple of examples, including the potential Burgers’
equation and the Gardner equation. In the first example, we
apply some technical methods, used in Examples 5.8 and 5.30
of [10].
Example 10. Consider the potential Burgers’ equation
𝑒𝑑 = 𝑒π‘₯π‘₯ + πœ€π‘’π‘₯2 .
(43)
6
Advances in Mathematical Physics
As mentioned in [7], approximate symmetries of the potential
Burgers’ equation are given by the following twelve vector
fields
v1 = πœ•π‘₯ ,
v2 = πœ•π‘‘ ,
v3 = π‘₯πœ•π‘₯ + 2π‘‘πœ•π‘‘ ,
v4 = 2π‘‘πœ•π‘₯ − (π‘₯𝑒 − πœ€π‘‘
v5 = (𝑒 − πœ€π‘‘
𝐷𝐾 = 𝐷π‘₯2 + 2πœ€π‘’π‘₯ 𝐷π‘₯ .
𝑒2
) πœ•π‘’ ,
2
2
𝑒
) πœ•π‘’ ,
2
(R1 )𝑑 = (𝐷π‘₯ + πœ€π‘’π‘₯ )𝑑 = πœ€π‘’π‘₯𝑑 = πœ€ (𝑒π‘₯π‘₯π‘₯ + 2πœ€π‘’π‘₯ 𝑒π‘₯π‘₯ ) = πœ€π‘’π‘₯π‘₯π‘₯ .
(49)
(44)
v7 = πœ€v1 ,
v8 = πœ€v2 ,
v11 = πœ€π‘’πœ•π‘’ = πœ€v5 ,
v12 = πœ€ (4π‘₯π‘‘πœ•π‘₯ + 4𝑑2 πœ•π‘‘ − (π‘₯2 + 2𝑑) π‘’πœ•π‘’ ) = πœ€v6 ,
plus the infinite family of vector fields
(45)
where 𝑓, 𝑔 are arbitrary solutions of the heat equation 𝑒𝑑 =
𝑒π‘₯π‘₯ .
R1 𝑄12 = πœ€ ((π‘₯2 + 6𝑑) 𝑒π‘₯ + 2π‘₯ (𝑒 + 2𝑑𝑒π‘₯π‘₯ ) + 4𝑑2 𝑒π‘₯π‘₯π‘₯ ) . (51)
To obtain the characteristics depending on π‘₯ and 𝑑, we
require a second approximate recursion operator, which by
inspection, we guess to be
R2 = 𝑑R1 +
The corresponding characteristics of the first twelve
approximate symmetries are
(R2 )𝑑 = 𝑑(R1 )𝑑 + R1 = 𝑑 [𝐷𝐾 , R1 ] + R1 ,
𝑄2 = 𝑒π‘₯π‘₯ + πœ€π‘’π‘₯2 ,
(46)
(52)
(53)
whereas
1
[𝐷𝐾 , R2 ] = 𝑑 [𝐷𝐾 , R1 ] + [𝐷π‘₯2 + 2πœ€π‘’π‘₯ 𝐷π‘₯ , π‘₯]
2
𝑒2
𝑄4 = π‘₯𝑒 + 2𝑑𝑒π‘₯ − πœ€π‘‘ ,
2
= 𝑑 [𝐷𝐾 , R1 ] + (𝐷π‘₯ + πœ€π‘’π‘₯ ) = 𝑑 [𝐷𝐾 , R1 ] + R1 ,
(54)
𝑒2
𝑄5 = 𝑒 − πœ€π‘‘ ,
2
𝑒2
) + 4π‘₯𝑑𝑒π‘₯ + 4𝑑2 (𝑒π‘₯π‘₯ + πœ€π‘’π‘₯2 ) ,
2
𝑄7 = πœ€π‘„1 = πœ€π‘’π‘₯ ,
𝑄8 = πœ€π‘„2 = πœ€π‘’π‘₯π‘₯ ,
proving that R2 is also an approximate recursion operator.
There is thus a doubly infinite hierarchy of approximate
generalized symmetries of potential Burgers’ equation, with
characteristics R𝑙2 Rπ‘˜1 𝑄1 , π‘˜, 𝑙 ≥ 0. For instance, 𝑄2 = R1 𝑄1 ,
𝑄3 = 2R2 R1 𝑄1 and so on.
Example 11. Consider the Gardner equation, which was
shown to have two Hamiltonian structures with
𝑄9 = πœ€π‘„3 = πœ€ (π‘₯𝑒π‘₯ + 2𝑑𝑒π‘₯π‘₯ ) ,
𝑄10 = πœ€π‘„4 = πœ€ (π‘₯𝑒 + 2𝑑𝑒π‘₯ ) ,
D = 𝐷π‘₯ ,
𝑄11 = πœ€π‘„5 = πœ€π‘’,
𝑄12 = πœ€π‘„6 = πœ€ ((π‘₯2 + 2𝑑) 𝑒 + 4π‘₯𝑑𝑒π‘₯ + 4𝑑2 𝑒π‘₯π‘₯ ) ,
(47)
(up to sign).
π‘₯
.
2
Using the fact that R1 satisfies (4), we readily find
𝑄1 = 𝑒π‘₯ ,
𝑄3 = π‘₯𝑒π‘₯ + 2𝑑 (𝑒π‘₯π‘₯ + πœ€π‘’π‘₯2 ) ,
(50)
Comparing these two verifies (4) and proves that R1 is an
approximate recursion operator for the potential Burgers’
equation.
There is thus an infinite hierarchy of approximate symmetries, with characteristics Rπ‘˜1 𝑄1 , π‘˜ = 0, 1, 2, . . . For example,
the next characteristic after 𝑄12 in the sequence is
v10 = πœ€ (2π‘‘πœ•π‘₯ − π‘₯π‘’πœ•π‘’ ) = πœ€v4 ,
v𝑓,𝑔 = (𝑓 (π‘₯, 𝑑) (1 − πœ€π‘’) + πœ€π‘” (π‘₯, 𝑑)) πœ•π‘’ ,
On the other hand, the commutator is computed using
Leibniz’ rule for differential operators:
[𝐷𝐾 , R1 ] = πœ€π‘’π‘₯π‘₯π‘₯ .
v9 = πœ€ (π‘₯πœ•π‘₯ + 2π‘‘πœ•π‘‘ ) = πœ€v3 ,
𝑄6 = (π‘₯2 + 2𝑑) (𝑒 − πœ€π‘‘
(48)
We must verify (4). The time derivative of the first approximate recursion operator R1 on the solutions of the potential
Burgers’ equation is the multiplication operator as follows:
𝑒2
) πœ•π‘’ ,
2
v6 = 4π‘₯π‘‘πœ•π‘₯ + 4𝑑2 πœ•π‘‘ − (π‘₯2 + 2𝑑) (𝑒 − πœ€π‘‘
Inspection of 𝑄1 , 𝑄2 , 𝑄7 , 𝑄8 leads us to the conjecture
that R1 = 𝐷π‘₯ + πœ€π‘’π‘₯ is an approximate recursion operator,
since 𝑄3 = R1 𝑄1 , 𝑄8 = R1 𝑄7 , and so forth. To prove this,
we note that the Fréchet derivative for the right-hand side of
potential Burgers’ equation is
E = 4𝑒𝐷π‘₯ + 2𝑒π‘₯ + 3πœ€ (𝑒𝑒π‘₯ + 𝑒2 𝐷π‘₯ ) − 𝐷π‘₯3 .
(55)
Hence, the operator connecting our hierarchy of approximate
Hamiltonian symmetries is
R = E ⋅ D−1 = 4𝑒 + 3πœ€π‘’2 + (2 + 3πœ€π‘’) 𝑒π‘₯ 𝐷π‘₯−1 − 𝐷π‘₯2 .
(56)
Advances in Mathematical Physics
7
Therefore, our results on approximate bi-Hamiltonian systems will provide ready-made proofs of the existence of
infinitely many approximate conservation laws and approximate symmetries for the Gardner equation.
Note that the Fréchet derivative for the right-hand side of
Gardner’s equation is
Definition 13. A pair of skew-adjoint π‘ž × π‘ž matrix of
differential operators D and E is said to form an approximately Hamiltonian pair if every linear combination π‘ŽD +
𝑏E, π‘Ž, 𝑏 ∈ R, is an approximate Hamiltonian operator. A
system of perturbed evolution equations is an approximate biHamiltonian system if it can be written in the form of
D𝐾 = 6 (1 + 2πœ€π‘’) 𝑒π‘₯ + 6 (𝑒 + πœ€π‘’2 ) 𝐷π‘₯ − 𝐷π‘₯3 ,
R𝑑 = (4 + 6πœ€π‘’) 𝑒𝑑 + (2𝑒π‘₯𝑑 + 3πœ€π‘’π‘‘ 𝑒π‘₯ +
πœ•π‘’
= 𝐾1 [𝑒, πœ€] ≈ D𝛿H1 ≈ E𝛿H0 ,
πœ•π‘‘
3πœ€π‘’π‘’π‘₯𝑑 ) 𝐷π‘₯−1
= 12𝑒𝑒π‘₯ (2 + 5πœ€π‘’) − (4 + 6πœ€π‘’) 𝑒π‘₯π‘₯π‘₯
(57)
(63)
where D, E form an approximately Hamiltonian pair, and H0
and H1 are appropriate Hamiltonian functionals.
Lemma 14. If D, E are skew-adjoint operators, then they form
an approximately Hamiltonian pair if and only if D, E and
D + E are all approximate Hamiltonian operators.
+ (6𝑒𝑒π‘₯π‘₯ (2 + 5πœ€π‘’) + 12𝑒π‘₯2 (1 + 5πœ€π‘’)
−𝑒π‘₯π‘₯π‘₯π‘₯ (2 + 3πœ€π‘’) − 3πœ€π‘’π‘₯ 𝑒π‘₯π‘₯π‘₯ ) 𝐷π‘₯−1 .
Theorem 12. Let 𝑄0 = πœ€π‘’π‘₯ . For each π‘˜ ≥ 0, the differential
polynomial π‘„π‘˜ = Rπ‘˜ 𝑄0 is a total π‘₯-derivative, π‘„π‘˜ = 𝐷π‘₯ π‘…π‘˜ ,
and hence we can recursively define π‘„π‘˜+1 = Rπ‘„π‘˜ . Each π‘„π‘˜ is
the characteristic of an approximate symmetry of the Gardner
equation.
Corollary 15. Let D and E be Hamiltonian differential
operators. Then D, E form an approximately Hamiltonian pair
if and only if
pr vDπœƒ (ΘE ) + pr vEπœƒ (ΘD ) = π‘œ (πœ€π‘ ) ,
(64)
where
Proof. To prove this theorem, we apply the similar method
applied in Theorem 5.31 of [10].
We proceed by induction on π‘˜, so suppose that π‘„π‘˜ =
Rπ‘˜ 𝑄0 for some π‘…π‘˜ ∈ A. From the form of the approximate
recursion operator,
π‘„π‘˜+1 = πœ€ (4π‘’π‘„π‘˜ + 2𝑒π‘₯ 𝐷π‘₯−1 π‘„π‘˜ − 𝐷π‘₯2 π‘„π‘˜ )
= πœ€π·π‘₯ (2𝑒𝐷π‘₯−1 π‘„π‘˜ + 2𝐷π‘₯−1 (π‘’π‘„π‘˜ ) − 𝐷π‘₯ π‘„π‘˜ ) .
(58)
If we can prove that for some differential polynomial π‘†π‘˜ ∈ A,
π‘’π‘„π‘˜ = 𝐷π‘₯ π‘†π‘˜ , we will indeed have proved that π‘„π‘˜+1 = 𝐷π‘₯ π‘…π‘˜+1 ,
where π‘…π‘˜+1 is the above expression in brackets. Consequently,
the induction step will be completed.
To prove this fact, note that the formal adjoint of the
approximate recursion operator πœ€R is
∗
πœ€R = πœ€ (4𝑒 −
2𝐷π‘₯−1
⋅ 𝑒π‘₯ −
𝐷π‘₯2 )
=
𝐷π‘₯−1 πœ€R𝐷π‘₯ .
(59)
We apply this in order to integrate the expression π‘’π‘„π‘˜ , by
parts, so
π‘˜
π‘„π‘˜ = 𝑒Rπ‘˜ [πœ€π‘’π‘₯ ] = 𝑒π‘₯ ⋅ (πœ€R∗ ) [𝑒] + 𝐷π‘₯ 𝐴 π‘˜
(60)
for some differential function 𝐴 π‘˜ ∈ A. On the other hand,
using a further integration by parts, for some π΅π‘˜ ∈ A the
following identity holds:
∗ π‘˜
𝑒π‘₯ ⋅ (πœ€R ) [𝑒] = 𝑒π‘₯ ⋅
𝐷π‘₯−1
πœ€R [𝑒π‘₯ ]
= 𝑒π‘₯ ⋅ 𝐷π‘₯−1 π‘„π‘˜
(61)
which proves our claim.
1
(𝐴 + π΅π‘˜ ) ,
2 π‘˜
ΘE =
1
∫ {πœƒ ∧ Eπœƒ} 𝑑π‘₯ (65)
2
are the functional bi-vectors representing the respective Poisson
brackets.
Example 16. Consider the approximate Hamiltonian operators D, E associated with the Gardner equation. We have
pr vDπœƒ = ∑ 𝐷𝐽 (Dπœƒ)𝛼
𝛼,𝐽
πœ•
πœ•π‘’π½π›Ό
π‘ž
πœ•
= ∑ 𝐷𝐽 ( ∑ D𝛼𝛽 πœƒπ›½ ) 𝛼
πœ•π‘’π½
𝛼,𝐽
𝛽=1
(66)
in the case of the second approximate Hamiltonian operator
for the Gardner equation, we have
pr vEπœƒ (𝑒2 ) = 2𝑒Eπœƒ,
pr vEπœƒ (𝑒) = Eπœƒ,
pr vEπœƒ (ΘD ) = pr vEπœƒ ∫
1
{πœƒ ∧ πœƒπ‘₯ } 𝑑π‘₯ = π‘œ (πœ€π‘ )
2
(67)
trivially, by the properties of the wedge product, it is deduced
that
pr vDπœƒ (ΘE )
1
3πœ€ 2
𝑒 ) πœƒ ∧ πœƒπ‘₯ + πœƒπ‘₯ ∧ πœƒ} 𝑑π‘₯ (68)
2
2
≈ ∫ {(2 + 3πœ€π‘’) πœƒπ‘₯ ∧ πœƒ ∧ πœƒπ‘₯ } = π‘œ (πœ€π‘ ) .
Thus D and E form an approximately Hamiltonian pair.
Substituting into the previous identity, we conclude
where π‘†π‘˜ =
1
∫ {πœƒ ∧ Dπœƒ} 𝑑π‘₯,
2
= pr vDπœƒ ∫ {(2𝑒 +
= −π‘’π‘„π‘˜ + 𝐷π‘₯ π΅π‘˜ .
π‘’π‘„π‘˜ = 𝐷π‘₯ π‘†π‘˜ ,
ΘD =
(62)
Definition 17. A differential operator D : Aπ‘Ÿ → A𝑠 is
approximately degenerate if there is a nonzero differential
Μƒ : A𝑠 → A such that D
Μƒ ⋅ D ≡ π‘œ(πœ€π‘ ).
operator D
8
Advances in Mathematical Physics
𝑝
Now, according to [10], we are in a situation to state the
main theorem on approximate bi-Hamiltonian systems.
Judging from R𝑙 = π‘œ(πœ€π‘ ), when 𝑙 =ΜΈ 0, this type of approximate recursion operators have less significance than R0 .
Theorem 18. Let
Example 21. The approximate recursion operators of the
Gardner equation are
𝑒𝑑 = 𝐾1 [𝑒, πœ€] ≈ D𝛿H1 ≈ E𝛿H0
(69)
be an approximate bi-Hamiltonian system of perturbed evolution equations. Assume that the operator D of the approximately Hamiltonian pair is approximate nondegenerate. Let
R = E ⋅ D−1 be the corresponding approximate recursion
operator, and let 𝐾0 ≈ D𝛿H0 . Assume that for each 𝑛 =
1, 2, . . . one can recursively define
𝐾𝑛 ≈ R𝐾𝑛−1 ,
𝑛 ≥ 1,
(70)
meaning that for each 𝑛, 𝐾𝑛−1 lies in the image of D. Then there
exists a sequence of functionals H0 , H1 , H2 , . . . such that
R0 = E ⋅ D−1 = 4𝑒 + 2𝑒π‘₯ 𝐷π‘₯−1 + 3πœ€ (𝑒𝑒π‘₯ 𝐷π‘₯−1 + 𝑒2 ) − 𝐷π‘₯2 ,
R1 = πœ€ (4𝑒 + 2𝑒π‘₯ 𝐷π‘₯−1 − 𝐷π‘₯2 )
(75)
and we can apply R0 to the right-hand side of the Gardner
equation to obtain the approximate symmetries. The first step
in this recursion is the flow
𝑒𝑑 ≈ E𝛿H1 ≈ D𝛿H2
≈ 𝑒π‘₯π‘₯π‘₯π‘₯π‘₯ − 10𝑒𝑒π‘₯π‘₯π‘₯ − 20𝑒π‘₯ 𝑒π‘₯π‘₯ + 30𝑒2 𝑒π‘₯
(i) for each 𝑛 ≥ 1 the perturbed evolution equation
𝑒𝑑 = 𝐾𝑛 [𝑒, πœ€] ≈ D𝛿H𝑛 ≈ E𝛿H𝑛−1
(ii) the corresponding approximate evolutionary vector
fields v𝑛 = v𝐾𝑛 all mutually commute
[v𝑛 , vπ‘š ] = π‘œ (πœ€π‘ ) ,
𝑛, π‘š ≥ 0;
(72)
(iii) the approximate Hamiltonian functionals H𝑛 are all in
involution with respect to either Poisson bracket:
{H𝑛 , H𝑛 }D = π‘œ (πœ€π‘ ) = {H𝑛 , H𝑛 }E ,
𝑛, π‘š ≥ 0,
+ πœ€ (55𝑒3 𝑒π‘₯ − 39𝑒𝑒π‘₯ 𝑒π‘₯π‘₯ − 9𝑒2 𝑒π‘₯π‘₯π‘₯ − 12𝑒π‘₯3 ) ,
(71)
is an approximate bi-Hamiltonian system;
(73)
and hence provide an infinite collection of approximate
conservation laws for each of the approximate biHamiltonian systems (63).
We have seen that given an approximate bi-Hamiltonian
system, the operator R = E ⋅ D−1 , when applied successively
to the initial equation 𝐾0 = D𝛿H0 , produces an infinite
sequence of approximate generalized symmetries of the original system (subject to the technical assumptions contained in
Theorem 18). It is still not clear that R is a true approximate
recursion operator for the system, in the sense that whenever
v𝑄 is an approximate generalized symmetry, so is vR𝑄. So far,
we only know it for approximate symmetries with 𝑄 = 𝐾𝑛 for
some 𝑛. In order to establish this more general result, we need
a formula for the infinitesimal change of the approximate
Hamiltonian operator itself under a Hamiltonian flow.
which is not approximately total derivative, so we cannot
reapply the approximate recursion operator to get a meaningful approximate generalized symmetry.
But if we set
𝐾1 [𝑒, πœ€] = 𝑄5 = πœ€πΎ1 [𝑒, πœ€] = πœ€ (6𝑒𝑒π‘₯ − 𝑒π‘₯π‘₯π‘₯ ) ,
Μƒ 5 = πœ€H0 ,
H0 = P
(74)
Theorem 20. Let 𝑒𝑑 = 𝐾 ≈ D𝛿H1 ≈ E𝛿H0 be an
approximate bi-Hamiltonian system of perturbed evolution
equations. Then the operators R𝑙 = πœ€π‘™ E ⋅ D−1 , 0 ≤ 𝑙 ≤ 𝑝,
are approximate recursion operators for the system.
H1 = P5 = πœ€H1 ,
(77)
then we can apply R0 successively to 𝐾1 in order to obtain
the approximate symmetries. The first phase become
𝑒𝑑 ≈ E𝛿H1 ≈ D𝛿H2 ≈ R0 𝐾1
≈ πœ€ (𝑒π‘₯π‘₯π‘₯π‘₯π‘₯ − 10𝑒𝑒π‘₯π‘₯π‘₯ − 20𝑒π‘₯ 𝑒π‘₯π‘₯ + 30𝑒2 𝑒π‘₯ )
(78)
in which
H2 = P5 =
πœ€
2
− 5𝑒2 𝑒π‘₯π‘₯ + 5𝑒4 ) 𝑑π‘₯
∫ (𝑒π‘₯π‘₯
2
(79)
is another approximate conservation law.
Now, for 𝐾2 = R0 𝐾1 we have
𝑒𝑑 ≈ E𝛿H2 ≈ D𝛿H3 ≈ R0 𝐾2
≈ πœ€ (−𝑒π‘₯π‘₯π‘₯π‘₯π‘₯π‘₯π‘₯ + 14𝑒𝑒π‘₯π‘₯π‘₯π‘₯π‘₯ + 42𝑒π‘₯ 𝑒π‘₯π‘₯π‘₯π‘₯ )
+ 70πœ€ (𝑒π‘₯π‘₯ 𝑒π‘₯π‘₯π‘₯ − 𝑒2 𝑒π‘₯π‘₯π‘₯ + 2𝑒3 𝑒π‘₯ − 4𝑒𝑒π‘₯ 𝑒π‘₯π‘₯ − 𝑒π‘₯3 ) ,
(80)
Lemma 19. Let 𝑒𝑑 = 𝐾 ≈ D𝛿H be an approximate
Hamiltonian system of perturbed evolution equations with
corresponding vector field v𝐾 = vΜ‚H . Then
pr vΜ‚H (D) ≈ D𝐾 ⋅ D + D ⋅ D∗𝐾 .
(76)
where
H3 = 7πœ€ ∫ (
2
𝑒π‘₯π‘₯π‘₯
2
+ 5𝑒2 𝑒π‘₯2 + 𝑒5 ) 𝑑π‘₯
+ 𝑒𝑒π‘₯π‘₯
14
is a further approximate conservation law.
(81)
Advances in Mathematical Physics
5. Concluding Remarks
Sometimes, differential equations appearing in mathematical
modelings are written with terms involving a small parameter
which is known as the perturbed term. Taking into account
the instability of the Lie point symmetries with respect
to perturbation of coefficients of differential equations, the
approximate (perturbed) symmetries for such equations are
obtained. Different methods for computing the approximate
symmetries of a system of differential equations are available
in the literature [2–4].
The approximate symmetry method proposed by
Fushchich and Shtelen [4] is based on a perturbation of
dependent variables. This method has so many advantages
such as producing more approximate group-invariant
solutions, consistence with the perturbation theory, solving
singular perturbation problems [7, 8], and close relationship
with approximate homotopy symmetry method [12]. But
despite the above-mentioned benefits, this procedure
converts a perturbed evolution equation to an equivalent
perturbed evolutionary system. In his case, obtaining the
corresponding Hamiltonian formulation will be hard. Due
to the increase of the dimensions of Hamiltonian operators
D, E, computation of the approximate recursion operator
R = E ⋅ D−1 is difficult.
Since prolongation and Fréchet derivative of vector fields
are linear, both of the approximate symmetry methods can
be extended on the Hamiltonian structures. But due to
the significance of vector fields in Hamiltonian and biHamiltonian systems, the approximate symmetry method
proposed by Baikov et al. [2, 3] seems to be more consistent.
Acknowledgments
It is a pleasure to thank the anonymous referees for their
constructive suggestions and helpful comments which have
improved the presentation of the paper. The authors wish to
express their sincere gratitude to Fatemeh Ahangari for her
useful advice and suggestions.
References
[1] S. Lie, “On integration of a class of linear partial differential
equations by means of definite integrals,” Archiv der Mathematik, vol. 6, pp. 328–368, 1881.
[2] V. A. Baikov, R. K. Gazizov, and N. Kh. Ibragimov, “Approximate
symmetries of equations with a small parameter,” MatematicheskiΔ±Μ† Sbornik, vol. 136, no. 4, pp. 435–450, 1988.
[3] V. A. Baikov, R. K. Gazizov, and N. H. Ibragimov, “Approximate
transformation groups and deformations of symmetry Lie
algebras,” in CRC Handbook of Lie Group Analysis of Differential
Equations, N. H. Ibragimov, Ed., vol. 3, chapter 2, CRC Press,
Boca Raton, Fla, USA, 1996.
[4] W. I. Fushchich and W. M. Shtelen, “On approximate symmetry
and approximate solutions of the nonlinear wave equation with
a small parameter,” Journal of Physics A, vol. 22, no. 18, pp. L887–
L890, 1989.
9
[5] N. Euler, M. W. Shul’ga, and W.-H. Steeb, “Approximate symmetries and approximate solutions for a multidimensional LandauGinzburg equation,” Journal of Physics A, vol. 25, no. 18, pp.
L1095–L1103, 1992.
[6] M. Euler, N. Euler, and A. Köhler, “On the construction of
approximate solutions for a multidimensional nonlinear heat
equation,” Journal of Physics A, vol. 27, no. 6, pp. 2083–2092,
1994.
[7] M. Pakdemirli, M. Yürüsoy, and IΜ‡. T. DolapcΜ§i, “Comparison
of approximate symmetry methods for differential equations,”
Acta Applicandae Mathematicae, vol. 80, no. 3, pp. 243–271,
2004.
[8] R. Wiltshire, “Two approaches to the calculation of approximate
symmetry exemplified using a system of advection-diffusion
equations,” Journal of Computational and Applied Mathematics,
vol. 197, no. 2, pp. 287–301, 2006.
[9] N. H. Ibragimov and V. F. Kovalev, Approximate and Renormgroup Symmetries, Nonlinear Physical Science, Springer, Berlin,
Germany, 2009.
[10] P. J. Olver, Applications of Lie Groups to Differential Equations,
vol. 107 of Graduate Texts in Mathematics, Springer, New York,
NY, USA, 2nd edition, 1993.
[11] M. Nadjafikhah and A. Mokhtary, “Approximate symmetry
analysis of Gardner equation,” http://arxiv.org/abs/1212.3604 .
[12] Z. Zhang, “Approximate homotopy series solutions of perturbed PDEs via approximate symmetrymethod,” http://128.84.
158.119/abs/1112.4225v2.
Advances in
Hindawi Publishing Corporation
http://www.hindawi.com
Algebra
Advances in
Operations
Research
Decision
Sciences
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Journal of
Probability
and
Statistics
Mathematical Problems
in Engineering
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
The Scientific
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International Journal of
Differential Equations
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Volume 2014
Submit your manuscripts at
http://www.hindawi.com
International Journal of
Advances in
Combinatorics
Mathematical Physics
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Journal of
Complex Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International
Journal of
Mathematics and
Mathematical
Sciences
Hindawi Publishing Corporation
http://www.hindawi.com
Journal of
Hindawi Publishing Corporation
http://www.hindawi.com
Stochastic Analysis
Abstract and
Applied Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
International Journal of
Mathematics
Volume 2014
Volume 2014
Journal of
Volume 2014
Discrete Dynamics in
Nature and Society
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Journal of
Applied Mathematics
Optimization
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Discrete Mathematics
Journal of Function Spaces
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Volume 2014
Volume 2014
Download