Está en la página 1de 17

PRAMANA 

c Indian Academy of Sciences Vol. 80, No. 5


— journal of May 2013
physics pp. 739–755

The symmetries and conservation laws


of some Gordon-type equations in Milne space-time
S JAMAL1 , A H KARA1,∗ , A H BOKHARI2 and F D ZAMAN2
1 School of Mathematics, University of the Witwatersrand, Johannesburg, Private Bag 3 Wits 2050,
South Africa
2 Department of Mathematics and Statistics, King Fahd University of Petroleum and Minerals,

Dhahran, 31261, Saudi Arabia


∗ Corresponding author. E-mail: Abdul.Kara@wits.ac.za

MS received 10 January 2012; revised 3 September 2012; accepted 6 December 2012

Abstract. In this letter, the Lie point symmetries of a class of Gordon-type wave equations that
arise in the Milne space-time are presented and analysed. Using the Lie point symmetries, it is
showed how to reduce Gordon-type wave equations using the method of invariants, and to obtain
exact solutions corresponding to some boundary values. The Noether point symmetries and conser-
vation laws are obtained for the Klein–Gordon equation in one case. Finally, the existence of higher-
order variational symmetries of a projection of the Klein–Gordon equation is investigated using the
multiplier approach.

Keywords. Conservation laws; Milne space-time; Gordon-type equations.

PACS Nos 02.30.Hq; 02.30.Jr; 02.30.Xx; 02.40.Ky

1. Introduction

A vast amount of work has been published in the literature studying differential equations
(DEs) in terms of the Lie point symmetries admitted by them [1,2]. These symmetries
play an important role in finding exact analytic solutions of the nonlinear DEs. Other than
Lie point symmetries, Noether symmetries are also widely studied and are associated, in
particular, with those DEs which possess Lagrangians. These symmetries represent phys-
ical features of DEs via the conservation laws they admit. The interesting link between
symmetries and conservation laws in mathematical physics is provided in the classic work
of Noether [3] showing that for every infinitesimal transformation admitted by the action
integral of a system, there exists a conservation law. The Noether symmetries, which are
symmetries of the Euler–Lagrange systems, have interesting applications in the study of
properties of particles moving under the influence of gravitational fields.

DOI: 10.1007/s12043-013-0518-3; ePublication: 9 April 2013 739


S Jamal et al

Recently, some published results were aimed at understanding Noether symmetries of


Lagrangians that arise from certain pseudo-Riemannian metrics of interest [4,5]. More
recently, Noether symmetries of the Euler–Lagrange equations on the Milne metric [6]
were found and a discussion of the results were given by comparing Noether symmetries
on the Milne metric with those of other conventional symmetries of the same space-time
[7]. Concerning the pure wave equation (homogeneous), it is a priori clear that it will
admit a maximal Noether symmetry group on a flat manifold. In that spirit, the work of
Mahadi [7] gives limited information and needs further understanding. With this example
in mind, we extend the work of Mahadi [7] by studying Klein–Gordon [8] equation on the
Milne metric and see how Noether symmetry structures change when classical wave equa-
tions are coupled with an inhomogeneous term. For completeness, we also investigate
the existence of higher-order variational symmetries of a projection of the Klein–Gordon
equation using the multiplier approach.
The plan of the paper is as follows: In §2 we derive Lie point symmetries of some
Gordon-type wave equations and illustrate the reduction of a Gordon-type wave equation
on a Milne manifold. In §3, we determine the Noether point symmetries of the Klein–
Gordon equation and construct the associated conserved densities. Lastly, we list some
higher-order symmetries and conservation laws of a projected Klein–Gordon equation
in §4.
We present some of the definitions and notations below. Intrinsic to a Lie algebraic
treatment of differential equations is the universal space A (see [2]). The space A is
the vector space of all differential functions of all finite orders and forms an algebra.
Consider an r th-order system of partial differential equations of n independent variables
x = (x 1 , x 2 , . . . , x n ) and m dependent variables u = (u 1 , u 2 , . . . , u m )

G μ (x, u, u (1) , . . . , u (r ) ) = 0, μ = 1, . . . , m̃, (1)

where u (1) , u (2) , . . . , u (r ) denote the collections of all first-, second-, . . ., r th-order par-
tial derivatives, that is, u iα = Di (u α ), u iαj = D j Di (u α ), . . . respectively, with the total
differentiation operator with respect to x i given by

∂ ∂ ∂
Di = + u iα α + u iαj α + · · · , i = 1, . . . , n, (2)
∂x i ∂u ∂u j

where the summation convention is used whenever appropriate. A current  =


(1 , . . . , n ) is conserved if it satisfies

Di i = 0 (3)

along the solutions of (1). It can be shown that every admitted conservation law arises
from multipliers Q μ (x, u, u (1) , . . .) such that

Q μ G μ = Di i (4)

holds identically (that is, off the solution space) for some current . The conserved
vector may then be obtained by the homotopy operator (see [1,9,10]). Other works on
symmetries and conservation laws can be found in [11–13].

740 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

DEFINITION 1

The Euler operator, for each dependent variable u α , is defined by


δ ∂  ∂
α
= α
+ (−1)s Di1 · · · Dis α , α = 1, . . . , m. (5)
δu ∂u s≥1
∂u i1 ···is

Note. In most literature, a variational problem consists of finding the extrema (maxima
or minima) of a functional

L[u] = L(x, u (n) )dx


in some class of functions u = f (x) defined over , where  ⊂ X is an open, connected


subset with smooth boundary ∂ (we consider the Euclidean space with X = R n ). The
integrand L(x, u (n) ), called the Lagrangian of the variational problem L, is a smooth
function of x, u and various derivatives of u [2].

DEFINITION 2

A Lie–Bäcklund operator is given by


∂ ∂  ∂
X = ξ i i + ηα α + ζiα1 ...is α , (6)
∂x ∂u s≥1
∂u i 1 ...i s

where ξ i , ηα ∈ A and the additional coefficients are determined uniquely by the


prolongation formulae

ζiα = Di (W α ) + ξ j u iαj ,
ζiα1 ...is = Di1 . . . Dis (W α ) + ξ j u αji1 ...is , s > 1. (7)

In (7), W α is the Lie characteristic function given by


W α = ηα − ξ j u αj . (8)
A Lie symmetry generator of (1) is a one-parameter Lie group transformation that
leaves the given differential equation invariant under the transformation of all indepen-
dent and dependent variables. In this paper, we shall assume that X is a Lie point operator,
i.e., ξ and η are functions of x and u and are independent of derivatives of u. A gener-
alized operator of the form X̃ = ηα ∂/∂u α + · · · is called a canonical or evolutionary
representation of X .

DEFINITION 3

If we include point-dependent gauge terms f 1 , . . . , f n , the Noether symmetries X are


given by
X (L) + L Di (ξ i ) = Di ( f i ). (9)

Pramana – J. Phys., Vol. 80, No. 5, May 2013 741


S Jamal et al

DEFINITION 4

The Noether operator associated with a Lie–Bäcklund operator X is given by


δ  δ
Ni = ξi + Wα α + Di1 · · · Dis (W α ) α , i = 1, . . . , n, (10)
δu i s≥1
δu ii 1 ···i s

where the Euler–Lagrange operators with respect to derivatives of u α are obtained from
(5) by replacing u α by the corresponding derivatives, e.g.,
δ ∂  ∂
α = α + (−1)s D j1 · · · D js α ,
δu i ∂u i s≥1
∂u i j1 ··· js
i = 1, . . . , n, α = 1, . . . , m. (11)

Noether’s Theorem. For any Noether symmetry X corresponding to a given Lagrangian


L ∈ A, there exists a conserved current i = (1 , . . . , n ), i ∈ A, defined by

i = f i − N i (L) , i = 1, . . . , n, (12)

which is a conserved current of the Euler–Lagrange equations (δL/δu α ) = 0.

2. Lie symmetries of Gordon-type equations in Milne space-time

Consider the Milne metric [6]

ds 2 = −dt 2 + t 2 (dx 2 + e2x (dy 2 + dz 2 )) (13)

which represents an empty Universe and is of interest in relativity for being a special
case of a well-known Friedmann–Lemaître–Robertson–Walker metric [6,8]. The Klein–
Gordon equation [8] on (13) is obtained by
 
1 ∂  ij ∂
u = √ |−g|g u = k(u), (14)
|−g| ∂ x i ∂xi
and takes the form

u x x − t 2 u tt + e−2x u yy + e−2x u zz − 3tu t + 2u x − t 2 k(u) = 0. (15)

In order to find the Lie point symmetries of the above Gordon-type equation we restrict
k(u) to some special cases. These cases are assumed by keeping in mind the fact that we
allow the inhomogeneous term, k(u), to be taken as sin(u) and some powers of u. The
criterion that yields the Lie point is given by the invariance condition [2]

X [u x x − t 2 u tt + e−2x u yy + e−2x u zz − 3tu t + 2u x − t 2 k(u)]|eq.(15)=0 = 0, (16)

where X is the prolonged symmetry generator in the jet space. Thus, the invariance
of differential equations (15) leads to the Lie point symmetries possessed by (15). The

742 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

procedure for finding Lie point symmetries is well known [2] and therefore will be given
without derivations. It turns out, from the symmetry study, that some special polynomial
cases of k(u) arise. Also, in line with the literature, we consider the sine-Gordon equation.
Thus, we study the four cases for k(u) in (15) given by
(i) k(u) = sin(u) (sine-Gordon)
(ii) k(u) = u (Klein–Gordon)
(iii) k(u) = u 3
(iv) k(u) = u n , n = 0, 1, 3.

Case i. Following the symmetry criterion, we find that eq. (16) in this case admits ten Lie
point symmetries given by
ex
X1 = ∂ x − e x ∂t ,
t
X 2 = ∂y ,
e−x ex y
X3 = ∂ y − ex y∂t + ∂x ,
t t
X 4 = ∂z ,
e−x ex z
X5 = ∂z − ex z∂t + ∂x ,
t t
X 6 = y∂z − z∂ y ,
X 7 = −∂x + y∂ y + z∂z ,
2e−x y 2e−x z
X8 = ∂y + ∂z + e−x (−1 − e2x (y 2 + z 2 ))∂t
t t
e−x (−1 + e2x (y 2 + z 2 ))
+ ∂x ,
t
X 9 = 2y∂x − 2yz∂z + (e−2x − y 2 + z 2 )∂ y ,
X 10 = −2yz∂ y + 2z∂x + (e−2x + y 2 − z 2 )∂z .

Case ii. When k(u) = u, we have a Klein–Gordon equation. Equation (16) in this case
admits 13 Lie point symmetries given by
X 1 = u∂u ,
X 2 = F1 (x, y, z, t)∂u ,
ex
X3 = ∂ x − e x ∂t ,
t
X 4 = ∂y ,
e−x ex y
X5 = ∂ y − ex y∂t + ∂x ,
t t

Pramana – J. Phys., Vol. 80, No. 5, May 2013 743


S Jamal et al

X 6 = y∂z − z∂ y ,
X 7 = −∂x + y∂ y + z∂z ,
X 8 = ∂z ,
e−x ex z
X9 = ∂z − ex z∂t + ∂x ,
t t
X 10 = −2∂x + 2y∂ y + 2z∂z + u∂u ,
2e−x y 2e−x z
X 11 = ∂y + ∂z + e−x (−1 − e2x (y 2 + z 2 ))∂t
t t
e−x (−1 + e2x (y 2 + z 2 ))
+ ∂x ,
t
X 12 = 2y∂x − 2yz∂z + (e−2x − y 2 + z 2 )∂ y ,
X 13 = −2yz∂ y + 2z∂x + (e−2x + y 2 − z 2 )∂z ,

where

e2x t 2 F1 (t, x, y, z) − F1,zz − F1,yy − 2e2x F1,x − e2x F1,x x


+ 3e2x tF1,t + e2x t 2 F1,tt = 0.

Case iii. When k(u) = u 3 , the (Gordon-type) eq. (16) yields a set of 15 Lie point
symmetries:

X 1 = t∂t − u∂u ,
ex (−1 + t 2 )
X 2 = −2ex tu∂u + ex (1 + t 2 )∂t + ∂x ,
t
ex (1 + t 2 )
X 3 = −2ex tu∂u + ex (−1 + t 2 )∂t + ∂x ,
t
X 4 = ∂y ,
e−x (−1 + t 2 ) ex (−1 + t 2 )y
X 5 = −2ex t yu∂u + ∂ y + ex (1 + t 2 )y∂t + ∂x ,
t t
X 6 = y∂z − z∂ y ,
X 7 = −∂x + y∂ y + z∂z ,
e−x (1 + t 2 ) ex (1 + t 2 )y
X 8 = −2ex t yu∂u + ∂ y + ex (−1 + t 2 )y∂t + ∂x ,
t t
X 9 = ∂z ,
e−x (1 + t 2 ) ex (1 + t 2 )z
X 10 = 2ex t zu∂u − ∂z − ex (−1 + t 2 )z∂t − ∂x ,
t t

744 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

e−x (−1 + t 2 ) ex (−1 + t 2 )z


X 11 = −2ex t zu∂u + ∂z + ex (1 + t 2 )z∂t + ∂x ,
t t
X 12 = −2y∂x + 2yz∂z + (−e−2x + y 2 − z 2 )∂ y ,
2e−x (−1 + t 2 )y 2e−x (−1 + t 2 )z
X 13 = ∂y + ∂z
t t
− 2e−x tu(1 + e2x (y 2 + z 2 ))∂u + e−x (1 + t 2 )(1 + e2x (y 2 + z 2 ))∂t
e−x (−1 + t 2 )(−1 + e2x (y 2 + z 2 ))
+ ∂x ,
t
2e−x (1 + t 2 )y 2e−x (1 + t 2 )z
X 14 = ∂y + ∂z − 2e−x tu(1 + e2x (y 2 + z 2 ))∂u
t t
+ e−x (−1 + t 2 )(1 + e2x (y 2 + z 2 ))∂t
e−x (1 + t 2 )(−1 + e2x (y 2 + z 2 ))
+ ∂x ,
t
X 15 = 2yz∂ y − 2z∂x + (−e−2x − y 2 + z 2 )∂z .

Case iv. In this case k(u) = u n , n = 0, 1, 3, the general polynomial Gordon-type equation
(16) admits 11 Lie point symmetries given by
2u t − nt
X1 = ∂u + ∂t ,
−3 + n −3 + n
ex
X2 = ∂ x − e x ∂t ,
t
X 3 = ∂y ,
e−x ex y
X4 = ∂ y − ex y∂t + ∂x ,
t t
X 5 = ∂z ,
X 6 = y∂z − z∂ y ,
X 7 = −∂x + y∂ y + z∂z ,
e−x ex z
X8 = ∂z − ex z∂t + ∂x ,
t t
2e−x y 2e−x z
X9 = ∂y + ∂z + e−x (−1 − e2x (y 2 + z 2 ))∂t
t t
e−x (−1 + e2x (y 2 + z 2 ))
+ ∂x ,
t
X 10 = −2y∂x + 2yz∂z + (−e−2x + y 2 − z 2 )∂ y ,
X 11 = 2yz∂ y − 2z∂x + (−e−2x − y 2 + z 2 )∂z .

Pramana – J. Phys., Vol. 80, No. 5, May 2013 745


S Jamal et al

2.1 Reduction of order of eq. (15) – an illustration

In this section we briefly show how the order of (1+3) Klein–Gordon equation (15) can
be reduced using its symmetries. In the first reduction, the equation with four independent
variables is reduced to a partial DE that has two independent variables. The reduced equa-
tion may then be analysed further using another Lie symmetry reduction or an appropriate
alternative method. In the second reduction (§2.1.1), we obtain exact solutions.
Since [X 6 , X 7 ] = 0, where X 6 and X 7 appear as Lie symmetries in all the above cases,
we may begin reducing with either X 6 or X 7 . Suppose we reduce (15) by X 6 = y∂z − z∂ y .
The characteristic equations are

dx dt dy dz du
= = = = .
0 0 −z y 0

Integrating yields α = y 2 + z 2 and eq. (15) is reduced to

1 2 3 2
u x x − u tt + 2 e−2x (αu αα + u α ) − u t + 2 u x − k(u) = 0 (17)
t2 t t t
with u = u(x, t, α).
If we then reduce eq. (17) by X 7 = −∂x + y∂ y + z∂z , we obtain the scaling
transformation X̄ = −∂x + 2α∂α . We now have the characteristic equations,

dt dx dα du
= = = .
0 −1 2α 0
By integrating, we obtain β = ln α + 2x and eq. (17) reduces to

2 4 3
u ββ (2 + e−β ) − u tt + 2 u β − u t − k(u) = 0 (18)
t2 t t
with u = u(t, β).
Equation (18) may be further analysed or reduced using the underlying symmetries.
It turns out that the Lie point symmetries are cumbersome and involve special functions
such as Bessel functions for k = u.
For k = u 3 , eq. (18) admits one symmetry

t∂t − u∂u .

For k = u 4 , the symmetries of (18) are

3 1
X∗ = t∂t − u∂u ,
4 2
X ∗∗ = 4t 2 (1 + 2eβ )∂β + t 3 (1 + 4eβ )∂t − 2t 2 u(1 + 4eβ )∂u .

Using X ∗∗ , eq. (18) reduces to the ordinary differential equation (ODE)

γ 4 Fγ γ + γ Fγ + 4F − 8γ 2 F 4 = 0, (19)

746 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

where
te−β/4
γ = and F = ueβ/2 (1 + 2eβ )1/2 .
(1 + 2eβ )1/4
It turns out that eq. (19) admits the Lie point symmetry

3
G = − β∂β + F∂ F
2
which leads us to the first-order ODE
2q + 24q 2/3 p 4/3 − 12q 2/3 p 1/3
qp = , (20)
2 p + 3q 1/3 p 2/3

where p = γ 2 F 3 and q = γ 5 F 3 .
There are no symmetries for k = u n , n = 0, 1, 3, 4 and k = sin u in eq. (18).
Also, one may consider reduction by studying the underlying conservation laws. This
would require methods other than the variational one, i.e., Noether’s theorem, since
eq. (18) is not variational.
For the Klein–Gordon case k(u) = u in (18), it can be shown, for e.g., that a conserved
vector of (18) is (β , t ), where

β = 2(1 + 2eβ ) BesselJ(1, t)u β ,


1 β
t = e t[(t BesselJ(0, t) − 2 BesselJ(1, t)
2
− t BesselJ(2, t))u − 2t BesselJ(1, t)u t ],

such that Dβ β + Dt t = 0 along the solutions of (18) and where BesselJ is the Bessel
function of the first kind ((β is the conserved flow and t , the conserved density).
For k(u) = u 3 in (18), the components of the conserved vector are:

β = (1 + 2eβ )t 2 [u t u β − uu βt ],
1
t = − t 2 [eβ t 2 u 4 + 2eβ t 2 u t 2 + 4u(eβ tu t − 2eβ u β − (1 + 2eβ )u ββ )].
4

Similarly, for k(u) = u 4 in (18), the components of the conserved vector are

1
β = − (1 + 2eβ )t 3 (40eβ u 2 + 4eβ t 2 u 5 − 5u β ((1 + 4eβ )tu t
5
+ 4(1 + 2eβ )u β ) + 5tu(10eβ u t + 2eβ tu tt + (1 + 4eβ )u βt )),
1 1
t = −t 4 (−2eβ (1 + 4eβ )u 2 + eβ (1 + 4eβ )t 2 u 5 + eβ tu t ((1 + 4eβ )tu t
5 2
+ 4(1 + 2eβ )u β ) − u(2eβ (3 + 8eβ )u β
+ (1 + 2eβ )(2eβ tu βt + (1 + 4eβ )u ββ ))).

Pramana – J. Phys., Vol. 80, No. 5, May 2013 747


S Jamal et al

The cases k(u) = sin u and k(u) = u n , n = 1, 3, 4 in eq. (18) do not yield any
conserved vectors.

2.1.1 Exact solutions/boundary conditions. We reduce eq. (15) by first using the
symmetry X = ∂ y to obtain
1 1 3 2
u x x − u tt + 2 e−2x u zz − u t + 2 u x − k(u) = 0,
t2 t t t
u = u(x, z, t), u(x, 0, t) = 0, u(x, 1, t) = 1. (21)
Reduce (21) further using the symmetry X = (e /t)∂x − e ∂t . The characteristic
x x

equations are
tdx dt dz du
= = = .
e x −e x 0 0
Integrating yields t˜ = (1/t)e−x and eq. (21) is reduced to
1
u zz − k(u) = 0, (22)
t˜2
where u(z, t˜) and boundary conditions transform to
u(0, t˜) = 0, u(1, t˜) = 1.
If k(u) = u, as in the Klein–Gordon case, then the general solution to eq. (22) is
u(z, t˜) = et˜z C1 (t˜) + e−t˜z C2 (t˜). (23)
We plot the function t˜ over different ranges: {x, −50, 50}, {t, 1, 10} (figure 1) and
{x, −10, 10}, {t, 1, 10} (figure 2). When t → ±∞, t˜ → 0, and when t → 0, t˜ becomes
large.

1 −x
Figure 1. t˜ = e .
t

748 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

1 −x
Figure 2. t˜ = e .
t

Keeping this in mind, we plot u(z, t˜) from (23) in figure 3: {t˜, 0, 10}, {z, −5, 5} and
figure 4: {t˜, 0, 10}, {z, −5, 5}, choosing particular C1 s and C2 s.
If we impose the boundary conditions on eq. (23), the solution of u(z, t˜) may be
expressed in terms of the hyperbolic sine function, namely,
1 1
u(z, t˜) = et˜z − e−t˜z .
˜
2 sinh(t ) 2 sinh(t˜)
The graph of this solution, for the range {t˜, 0, 10}, {z, −5, 5}, is given in figure 5.

Figure 3. C1 = 100t˜ 2 , C2 = −2 sin(t˜).

Pramana – J. Phys., Vol. 80, No. 5, May 2013 749


S Jamal et al

Figure 4. C1 = 100, C2 = −2.

Remark. For the sine-Gordon case, with k(u) = sin(u) in eq. (22), we obtain

u(z, t˜) = ±2 Jacobi Amplitude


 
1 4t˜2
× −(2t˜2 − D1 (t˜))(z + D2 (t˜))2 , − ,
2 −2t˜2 + D1 (t˜)

where Jacobi amplitude [v, m] refers to the amplitude am(v  m) for Jacobi elliptic
functions.

Figure 5. u(z, t˜) = (1/2 sinh(t˜))et˜z − (1/2 sinh(t˜))e−t˜z .

750 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

3. Noether symmetries of Klein–Gordon equation for the Milne space-time

Consider the wave eq. (15) with k(u) = u (Klein–Gordon), which has the Lagrangian,

1 3 2x 2 1 2x 2 1 2 1 2 1 3 2x 2
L= t e u + te u x + tu y + tu z − t e u t . (24)
2 2 2 2 2
We assume that

X = ξ(t, x, y, z, u)∂x + τ (t, x, y, z, u)∂t + η(t, x, y, z, u)∂ y

+γ (t, x, y, z, u)∂z + φ(t, x, y, z, u)∂u

is a Noether point operator that satisfies (9) with gauge vector f i (i = 1, 2, 3, 4), depen-
dent on (t, x, y, z, u). Then the Noether symmetry criterion (9) for the Lagrangian given
by eq. (24) takes the form,

X L + L[Dt τ + Dx ξ + D y η + Dz γ ] = Dt f 1 + Dx f 2 + D y f 3 + Dz f 4 .

Separation by derivatives of u yields the following overdetermined system:

u 3t : τu = 0,
u 3x : ξu = 0,
u 3y : ηu = 0,
u 3z : γu = 0,
3 1 1 1
u 2t : −e2x t 3 ξ − e2x t 2 τ − e2x t 3 γz − e2x t 3 η y − e2x t 3 ξx
2 2 2 2
1
+ e2x t 3 τt − e2x t 3 φu = 0,
2
1 1 1 1 1
u 2x : e2x tξ + e2x τ + e2x tγz + e2x tη y − e2x tξx + e2x tτt
2 2 2 2 2
+ e2x tφu = 0,
1 1 1 1 1
u 2y : τ + tγz − tη y + tξx + tτt + tφu = 0,
2 2 2 2 2
1 1 1 1 1
u 2z : τ − tγz + tη y + tξx + tτt + tφu = 0,
2 2 2 2 2
u t u z : e2x t 3 γt − tτz = 0,
u x u z : −e2x tγx − tξz = 0,
u y u z : −tγ y − tηz = 0,
u t u y : e2x t 3 ηt − tτ y = 0,
u y u x : −e2x tηx − tξ y = 0,

Pramana – J. Phys., Vol. 80, No. 5, May 2013 751


S Jamal et al

u t u x : e2x t 3 ξt − e2x tτx = 0,


u t : − f 1,u − e2x t 3 φt = 0,
u x : − f 2,u + e2x tφx = 0,
u y : − f 3,u + tφ y = 0,
u z : − f 4,u + tφz = 0,
3 1

1 : e2x t 3 u 2 ξ + e2x t 2 u 2 τ + e2x t 3 uφ + t 3 e2x u 2 τt + ξx + η y + γz

2 2
− f 1,t + f 2,x + f 3,y + f 4,z = 0. (25)
Solving the above system, one finds that the Noether point symmetries of the Klein–
Gordon equation for the Milne metric are given by
 x 
2x 3 e
X1 = e t ∂x − e ∂t , f i = 0,
x
t
X 2 = e2x t 3 ∂ y , f i = 0,
 −x 
2x 3 e ex y
X3 = e t ∂ y − e y∂t +
x
∂x , f i = 0,
t t
X 4 = e2x t 3 ∂z , f i = 0,
 −x 
e ex z
X 5 = e2x t 3 ∂z − ex z∂t + ∂x , f i = 0,
t t
X 6 = e2x t 3 (y∂z − z∂ y ), f i = 0,
X 7 = e t (−∂x + y∂ y + z∂z ), f i = 0,
2x 3
 −x
2e y 2e−x z
X 8 = e2x t 3 ∂y + ∂z + e−x (−1 − e2x (y 2 + z 2 ))∂t
t t

e−x (−1 + e2x (y 2 + z 2 ))
+ ∂x , f i = 0,
t
X 9 = e2x t 3 (2y∂x − 2yz∂z + (e−2x − y 2 + z 2 )∂ y ), f i = 0,
X 10 = e2x t 3 (−2yz∂ y + 2z∂x + (e−2x + y 2 − z 2 )∂z ), f i = 0.
We may then obtain the corresponding conserved flows for each X i (i = 1, . . . , 10).
For example, the symmetry X 6 yields the conserved flow,
1
t6 = − e2x t 3 (yu z u t − zu y u t + u(−yu t z + zu t y ))
2
1 2x
6 = e t (yu z u x − zu y u x + u(−yu x z + zu x y ))
x
2
y 1
6 = t (e2x t 2 zu 2 + u y (yu z − zu y ) − u(u z + zu zz + yu yz − 3e2x t zu t
2
− e2x t 2 zu tt + 2e2x zu x + e2x zu x x ))
1
6z = − t (e2x t 2 yu 2 + u z (−yu z + zu y ) − u(u y + zu yz
2
+ y(u yy + e2x (−3tu t − t 2 u tt + 2u x + u x x )))).

752 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

4. Klein–Gordon equations and higher-order variational symmetries and


conservation laws in Milne space-time

Consider the Klein–Gordon wave equation in Milne space-time with dependent variable
u as a function of x, t and y only, i.e., we have removed the spatial variable z from the
original wave equation (15) – the calculations are extremely cumbersome producing no
final outcomes. We consider the multiplier method for eq. (15), by choosing k(u) = u.
That is, since the Euler–Lagrange operator annihilates total divergences, we get
  
δ 1 1 3 2
Q 2 u x x − u tt + 2 e−2x u yy − u t + 2 u x − u = 0, (26)
δu t t t t

where Q = Q(x, y, t, u x , u x , u x x , u x y , u x x x , u x x y , u x yy , u yyy ). Although not pursued


here, the calculations may include derivatives of u with respect to t. Then

 
1 1 3 2
Q 2
u x x − u tt + 2 e−2x u yy − u t + 2 u x − u
t t t t
= D t t + D x  x + D y  y ,

where (x ,  y , t ) is the conserved flow (t being the conserved density). We obtain
the set of multipliers Qi , namely,


1 1 1
Q1 = t e 3 2x
− u x x x − yu y + yu x x y + y 2 u yy
3 6 2

1 1 1
+ y 3 u yyy − y 2 u x yy − u x x + u ,
3 2 12
Q2 = t 3 e2x (−2yu x yy + u x x y + yu yy + y 2 u yyy ),
1 3 −2x
Q3 = t e u yyy + t 3 u x x y
2

1
+ t e 2y 3 u x yy − y 3 u yy − y(u) − 3y 2 u x x y
3 2x
2

1 2 1 4
+ y u y + 3yu x x + 2yu x x x − y u yyy ,
2 2
1 3 2x
Q4 = t e (2u x yy − u yy − 2yu yyy ),
2
Q5 = t 3 e2x u yyy ,
 
1
Q7 = t e u x + u − yu y ,
3 2x
2
Q8 = t 3 e2x u y .

Pramana – J. Phys., Vol. 80, No. 5, May 2013 753


S Jamal et al

These multipliers yield a set of eight conserved flows, for example, the corresponding
components of the conserved vector for Q4 are
1
4x = t (2e2x t 2 u y 2 + 2u yy 2 − 6e2x tu yy u t − 2e2x t 2 u yy u tt
12
+ e2x u yy u x − 6e2x yu yyy u x + 6e2x u x u x yy + 2e2x u yy u x x
− 2u y (u yyy + e2x (−3tu t y − t 2 u tt y + 2u x y + u x x y ))
+ u(−4e2x t 2 u yy + 2u yyyy + e2x (−6tu t yy − 2t 2 u tt yy
+ 7u x yy + 6yu x yyy − 4u x x yy ))),
1 2x 3
t4 = e t (u yy u t + 2yu yyy u t − uu t yy − 2yuu t yyy − 2u t u x yy + 2uu xt yy ),
4
y 1
4 = t (−6e2x t 2 yu y 2 − 2(3yu yy 2 + u yyy u x
12
+ u yy (−9e2x t yu t − 3e2x t 2 yu tt + 6e2x yu x − 2u x y + 3e2x yu x x )
+ e2x (−3tu t y u x − t 2 u tt y u x + 6tu t u x y + 2t 2 u tt u x y
− 2u x u x y − 2u x y u x x + u x u x x y ))
+ u y (3e2x tu t − 18e2x t yu t y + e2x t 2 u tt
− 6e2x t 2 yu tt y − 2e2x u x + 4e2x t 2 u x + 12e2x yu x y + 4u x yy
+ 6e2x tu xt + 2e2x t 2 u xtt − 5e2x u x x + 6e2x yu x x y − 2e2x u x x x )
+ u(2e2x t 2 u y + 12e2x t 2 yu yy + 3e2x tu t y
+ 18e2x t yu t yy + e2x t 2 u tt y + 6e2x t 2 yu tt yy − 2e2x u x y
− 8e2x t 2 u x y − 12e2x yu x yy
− 2u x yyy − 12e2x tu xt y − 4e2x t 2 u xtt y + 7e2x u x x y
− 6e2x yu x x yy + 4e2x u x x x y )).

5. Concluding remarks

This paper investigates a class of wave and Gordon-type equations in Milne space-time.
In particular, we conducted a Lie and Noether symmetry analysis of a Klein–Gordon
equation on this manifold. We have given some symmetry reductions to show how the
(1+3)-dimensional wave equation can be reduced to an ordinary differential equation
using the method of invariants, and obtained some exact solutions. A conserved density
of the Klein–Gordon equation is constructed. Finally, some higher-order symmetries for
the projected equation and associated conservation laws are presented. It is hoped that an
analysis of the nonlinear Klein–Gordon equation in a genuinely curved space-time will
provide interesting insight from the point of view of conserved quantities.

754 Pramana – J. Phys., Vol. 80, No. 5, May 2013


Symmetries of Gordon-type equations in Milne space-time

References

[1] S Anco and G Bluman, Eur. J. Appl. Math. 13, 545 (2002)
[2] P Olver, Application of Lie groups to differential equations (Springer, New York, 1993)
[3] E Noether, Mathematisch-Physikalische Klasse 2, 235 (1918), English translation in Transport
Theory and Statistical Physics 1(3), 186 (1971)
[4] A H Bokhari and A H Kara, Gen. Relativ. Gravit. 39, 2053 (2007)
[5] A H Bokhari, A H Kara, A R Kashif and F D Zaman, Int. J. Theor. Phys. 45(6), 1029 (2006)
[6] A D Rendall, Applications of the theory of evolution equations to general relativity, Gen-
eral Relativity – Proceedings of the 16th International Conference edited by N T Bishop and
S D Maharaj (World Scientific, Singapore, 2002)
[7] Ahmad M Mugbil, On Noether symmetry and conservation laws, Ph.D. Thesis (King Fahd
University of Petroleum and Minerals, Dhahran, Saudi Arabia, 2011).
[8] C W Misner, K S Thorne and J A Wheeler, Gravitation (W H Freeman, San Francisco, 1973)
[9] G Bluman and S Kumei, Symmetries and differential equations (Springer-Verlag, New York,
1989)
[10] W Hereman, Int. J. Quantum Chem. 106, 278 (2006)
[11] A H Kara and F M Mahomed, Int. J. Theor. Phys. 39(1), 23 (2000)
[12] A H Kara and F M Mahomed, J. Nonlinear Math. Phys. 9, 60 (2002)
[13] U Göktas and W Hereman, Physica D 123, 425 (1998)

Pramana – J. Phys., Vol. 80, No. 5, May 2013 755

También podría gustarte