Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Classication of the group invariant solutions for

contaminant transport in saturated soils under


radial uniform water ows

M.M. Potsane and R.J. Moitsheki


Center for Dierential Equations, Continuum Mechanics and Applications,
School of Computational and Applied Mathematics,
University of the Witwatersrand, Private Bag 3,
Johannesburg 2050, South Africa.

Abstract

The transport of chemicals through soils to the groundwater or precipita-


tion at the soils surfaces leads to degradation of these resources. Serious
consequences may be suered in the long run. In this paper, we consid-
er macroscopic deterministic models describing contaminant transport in
saturated soils under uniform radial water ow backgrounds. The arising
convection-dispersion equation given in terms of the stream functions is an-
alyzed using classical Lie point symmetries. A number of exotic Lie point
symmetries is admitted. Group invariant solutions are classied according
to the elements of the one dimensional optimal systems. We analyzed the
group invariant solutions which satisfy the physical boundary conditions.

1 Introduction
Prediction of water and contaminant movements in soils is important to the
theory of soil salinity and underground water pollution. Among others, the
sources of contamination are movement of agricultural and industrial con-
taminants. Some exact solutions are constructed for constraint assumptions,
as such, investigation and study of these problems is extremely dicult and
challenging [1], even when the problem is given in terms of linear partial
dierential equations [2].
Broadbridge et al. [2, 3, 4] applied the classical Lie point symmetries
to analyze the convection dispersion equations arising in solute transport
theory. A compendium of exact (group-invariant) solutions is given in [5].
Other researchers used dierent techniques to obtain exact solutions (see e.g.
[6, 7, 8]). Chen et al. [9] constructed analytical solutions for two-dimensional
advection-dispersion equation with transverse dispersivities depending lin-
early on the spatial variable. Yadav et al. [10] constructed analytical solu-
tions for solute transport in semi-innite porous domain. Symmetry reduc-
tions and group invariant solutions were constructed for models describing
motion of a polytropic gas [11].

1
Exact analytical solutions for contaminant transport in porous media
have been constructed for example in [12, 13, 14, 15]. Chrysikopoulos and
Sim [12] developed the one-dimensional stochastic model describing virus
transport in homogeneous, saturated semi-nite porous media. The ideas in
[12] are extended to three dimensional problems in [13, 14]. Experimental
investigations of acoustically enhanced colloid transport in water-saturated
packed columns [15]. In this paper we focus on theoretical macroscopic
deterministic models given in terms of partial dierential equations.
Numerical models are able to simulate complex reactive transport phe-
nomena, but can be time consuming to construct and subject to numerical
discretization errors [7]. Also, available packages have signicant disagree-
ment in their prediction of solute transport [16]. Therefore, exact solutions
are very important because they are needed both as validation tests for nu-
merical schemes and also provide insight into the water and solute transport
processes.
In this paper, we construct group-invariant solutions and classify them
according to the one-dimensional optimal systems. This is a signicant ad-
vancement of the work by [4] and [6]. We consider contaminant transport
under radial uniform water ows. The resulting convection-dispersion equa-
tion given in terms of stream function is analyzed when dispersion coecient
is proportional to the water pore velocity. It turns out the Lie point symme-
tries are admitted when the dispersion coecient is a constant in one case
and depending on water pore velocity in the other. In Section 2, we pro-
vide the mathematical models. We briey discuss the symmetry methods
in Section 3. The problem at hand is analyzed in Section 4 and the group
invariant solutions are classied according to the one-dimensional optimal
systems in Section 5 and some physically realistic solutions are discussed.
lastly, some concluding remarks are given in Section 6.

2 Mathematical model
The theoretical background in this section is obtained from Hillel [17]. The
macroscopic deterministic model, which is based on local conservation laws
is given by
(c)
= (D(v)c) (cV), (2.1)
t
where D(v) denotes the coecient of hydrodynamic dispersion and is
given in cartesian coordinates. Experimental and theoretical observations
show that the dispersion coecient can take the power law form D(v) = v p
with being a proportionality constant and 1 p 2 (see e.g. [18]). Here,
v is the pore water velocity. Given uniform water ow in saturated soils,
then the continuity equation is given by V=0. Uniform saturated water
ow implies that the soil water content = s , where s is the water content

2
at saturation. By Darcys law, which states that water ux is proportional
to the gradient of the hydraulic pressure head, that is V = ks , where
is the total hydraulic pressure head and ks is the hydraulic conductivity at
saturation (see e.g. [17]), as such we obtain the Laplaces equation 2 = 0.
Under these assumptions, Equation (2.1) then becomes

c
= (D(v)c) + k c, (2.2)
t
where k = ks /s and v = |k|. Note that this problem becomes extremely
dicult to solve exactly when v must be the modulus of the potential ow
velocity eld for an incompressible uid (see e.g. [2]). However, one may
transform Equation (2.1) from cartesian coordinates (x, y) to the streamline
coordinates (, ) using Laplace preserving or conformal transformations
[19], see also [2]. The resulting solute transport Equation is given by

c c
= v 2 [D(v)c] + v 2 , (2.3)
t
( )
where =
, . The velocity potential is and is a conjugate
harmonic stream function such that x = y and y = x . Also, the
functions and satisfy the Laplace equation. In radial water ows, the
velocity potential is given by = (Q/s ) logr and the stream function
= (Q/s ) arctan(y/x), where Q is the source strength (pumping rate).
Introducing the normalized concentration, velocity potential and time given
by C = c/cs , = logR and = t/ts respectively with cs , ts and R being
the concentration, time at soil saturation and the distance or radius from
the point source, we may write Equation (2.3) as

C C
= v 2 [D(v)C] + v 2 . (2.4)

We refer to Equation (2.4) as the governing equation. Equation (2.4) is


susceptible to symmetry analysis (see e.g. [3, 4]).

3 Algebraic techniques for symmetry reduction


In brief, a symmetry of a dierential equation is an invertible transformation
of the dependent and independent variables that does not change the original
dierential equation. Symmetries depend continuously on a parameter and
form a group; the one-parameter group of transformations. This group can
be determined algorithmically by hand or by computer software programs
such as YaLie [20], Reduce [21], Dimsym [22] and so on. The theory of
application of Lie groups to dierential equations may be found in texts such
as those of [23, 25, 26]. Given a second order partial dierential equation

3
such as Equation (2.4) describing contaminant transport under radial water
ows, we seek transformations of the form

= + 1 (, , C) + O(2 )
= + 2 (, , C) + O(2 ) , (3.1)

C = C + (, , C) + O(2 )

generated by the vector eld



X = 1 (, , C) + 2 (, , C) + (, , C) . (3.2)
C
Note that the transformations in (3.1) are equivalent to the one-parameter
Lie group of transformations that leaves the equation in question invariant.
Since Equation (2.4) is second order, then one may prolong the symmetry
generator (3.2) accordingly. The invariance criterion is then given by

X [2] (Equation(2.4))|Equation(2.4) = 0,

where X [2] is the second prolongation given by



X 2 = 1 + 2 + + + + .
C C C C

Here , and are the extended or prolonged innitesimals (see


e.g. [25]). The invariance criterion results in the overdetermined system of
linear homogeneous partial dierential equations known as the determining
equations, which may be solved even by interactive programs such as YaLie
[20] and Reduce [21].

4 Lie point symmetry analysis of Equation (2.4)


We consider contaminant transport under steady radial water ow in sat-
urated soils. In this case, the relevant normalized point source, the water
velocity and the dispersion coecient are given by = log R, v = e
and D(v) = ep , respectively (see e.g. [3]). One may simply omit the
dependence of contaminant concentration on the clockwise polar angle co-
ordinate. In the initial symmetry analysis, Equation (2.4) with arbitrary p

and , admits the time translation and the scaling of C, that is, X1 =


and X2 = C , and the innite symmetry generator
C

X = (, ) ,
C
where is any solution of Equation (2.4). Extra symmetries are admitted
only when p = 0 and p = 2.

4
4.1 Constant dispersion coecient
Given p = 0, then the dispersion coecient becomes a proportionality con-
stant, . Given dispersion coecient as an arbitrary constant = 0, Equa-
tion (2.4) admits nite two extra symmetries





X3 = 2 ,


( ) (4.1)
2

1 e

X4 = 2 + C .
2 4 C
Further, symmetries may be admitted when = 1 and = 1. Note that
= 1 implies negative dispersion coecient which is not physically real-
istic, and as such we focus on the case = 1. Given p = 0 and = 1, then
Equation (2.4) admits nite four extra symmetries given by



X3 = e ,











X4 = 2 ,


( 2 ) (4.2)


e

X5 = + C ,

2 C



( )

2 e 2

X6 = + C .

2 4 C

4.2 Velocity-dependent dispersion coecient


The case p = 2, is in agreement with solute transport theory. This implies
that the dispersion coecient is now given in terms of the water pore veloc-
ity. In this case, Equation (2.4) with arbitrary proportionality constant ,
admits extra four nite Lie point symmetries given by


X3 = e2 ,





( 4 )

2

1 e e

X4 = C + ,

4 2 C

( ) (4.3)
1 e2

1

X5 = + C ,

4 8 4 C



( 2 )

4

1 e e 2
X6 = 2
+ C .

2 4 16 2 4 C

5
Note that the admitted symmetry structure and number is not aected by
the constant , that is we obtain the same symmetries up to the specied
value.

5 Classication of group-invariant solutions


In general it is possible to reduce the number of independent variables by
one using any linear combination of the admitted base vectors such as those
in Equations (4.2) and (4.3). In other words, for each l -parameter subgroup
(or equivalently l -dimensional subalgebra) of the full symmetry group (or
symmetry algebra) there is a corresponding family of group-invariant solu-
tions, which may be innite [23]. Thus one needs a systematic means to
classify these solutions such that none can be derived from the other. In
order to classify these solutions one needs to construct a set of optimal sys-
tems.
Denition: [23] An optimal system of lparameter group-invariant so-
lutions to a dierential equation (or system of dierential equations) is a
collections of solutions with the properties
(i) Each solution in the list is invariant under some lparameter symme-
try group of the dierential equation (or system of dierential equa-
tions).

(ii) If there exists another solution which is invariant under lparameter


symmetry group, then there is a further symmetry generator admitted
by the equation (or system) which maps this old solution to the new
one.
Clearly, an optimal system is a set of elements which lead to symmetry
reductions that are not equivalent by any transformation. Since we aim to
classify the group invariant solutions of Equation (2.4), we construct the
optimal system for the Lie point symmetries in Equations (4.2) and (4.3).

5.1 Construction of the optimal system of subalgebras


In this section we adopt the method in [23] to construct the one-dimensional
system of subalgebras of the algebra spanned by the base vectors in array of
Equations (4.2). To construct the optimal system we rst need to determine
the commutators of the admitted symmetries.
Denition: [24] Given the generators

X1 = 1i (x, u) i
+ 1 (x, u)
x u
and

X2 = 2i (x, u) + 2 (x, u) ,
xi u

6
admitted by a k th-order partial dierential equation
( )
u(k) = F x, u, u(1) , ..., u(k1) ,

then the commutator of X1 and X2 is dened by


( ( ) ( ))
[X1 , X2 ] = X1 2i X2 1i i
+ (X1 (2 ) X2 (1 )) .
x u
One may list the commutators of symmetries in Equation (4.2) together
with the time translation and concentration scaling in Table 1.

Table 1: Commutators of the admitted symmetries.

[Xi , Xj ] X1 X2 X3 X4 X5 X6
X1 0 0 0 2X1 2X3 4X4 2X2
X2 0 0 0 0 0 0
X3 0 0 0 X3 X2 2X5
X4 2X1 0 X3 0 X5 2X6
X5 2X3 0 X2 X5 0 0
X6 2X2 4X4 0 2X5 2X6 0 0

Furthermore we construct a set of one-dimensional subalgebras which


are equivalent to a unique element of the set under some element of the
adjoint representation given by
n

( ) 2
Ad e Xi
Xj = (AdXi )n Xj = Xj [Xi , Xj ] +
[Xi , [Xi , Xj ]] ,
n! 2!
n=0
(5.1)
where the commutator of Xi and Xj is dened above. The adjoint repre-
sentation of base vectors in Equation (4.2) is given in Table 2. Let
X = a1 X1 + a2 X2 + a3 X3 + a4 X4 + a5 X5 + a6 X6 . (5.2)
It remains to use the adjoint table to simplify as much as possible the con-
stants in Equation (5.2). The key here, is the recognition of the function
(X) = a24 4a1 a6 , which is invariant under the full adjoint action [23]. To
begin the simplication we concentrate on the constants a1 , a4 , a6 . If X is
given as in (5.2), then

6
e=
X aei Xi = Ad(exp(X1 )) Ad(exp(X6 ))X (5.3)
i=1
has the coecients

ae1 = a1 2a4 + 42 a6 ,
ae4 = a4 4a6 + 4(a1 2a4 + 42 a6 ), (5.4)

ae6 = a6 + 2(a4 4a6 ) + 4 2 (a1 2a4 + 42 a6 ).

7
Three cases arise.
Case 1. If (X) > 0, then we choose to be a real root of a1 2a 2
4 +4 a6 =
0 and = a6 /(8a6 2a4 ). This implies ae1 = ae6 = 0 and ae4 = (X) = 0.
Thus, X is equivalent to the vector X e = X4 + ae2 X2 + ae3 X3 + ae5 X5 . Acting on
e
X by adjoint maps generated by X5 and X3 , namely Ad (exp (ae5 X5 )) and
Ad (exp (ae3 X3 )) , as such X5 and X3 in X e vanish. No further simplications
are possible, therefore X is equivalent to a multiple of X4 + aX2 for some
a R provided (X) > 0.
Case 2. If (X) < 0, we set = 0 and = a4 /4a1 so that ae4 = 0. One
may then assume both the coecients of X1 and X6 to be unity. Thus, X
is equivalent to the vector X e = X1 + X6 + ae2 X2 + ae3 X3 + ae5 X5 . Acting on
e
X by adjoint maps generated by X5 and X3 , namely Ad (exp ((ae3 /2) X5 ))
and Ad (exp ((ae5 /2) X3 )) , as such X5 and X3 in X e vanish. No further
simplications are possible, as such X is equivalent to X1 + X6 + aX2 , a R
given (X) < 0.
Case 3. (X) = 0. Two subcases arise. If not all the coecients a1 , a4 , a6
are zero, then we are free to choose and such that ae1 = 0 and ae4 =
ae6 = 0. In this case X is equivalent to a multiple of X e = X1 + ae2 X2 +
ae3 X3 + ae5 X5 . Acting on Xe by adjoint maps generated by X1 and X3 , namely
Ad (exp ((ae3 /2ae5 ) X5 )) and Ad (exp ((ae2 /ae5 ) X3 )) , ae5 = 0, as such X2
and X3 in X e vanish. Therefore X is equivalent to a multiple of X1 X5 . If
ae5 = 0, then acting on X e by Ad (exp ( (ae3 /2) X5 )) and so X is equivalent
to a multiple of X1 + aX2 , a R. If all the coecients a1 , a4 , a6 are zero,
and assuming a3 = 0 (say a3 = 1), then acting on X by Ad (exp (ae2 X5 ))
and Ad (exp ((ae5 /2) X5 )) yield X3 , a multiple of X. If a3 = 0 but a5 = 0,
acting by any group generated by X1 , namely, Ad (exp (X1 )) gives a nonzero
coecient in front X3 implying that this case is similar to the case when
a3 = 0. Thus, the only remaining vectors are multiples of X2 . The set of
one-dimensional optimal system is given by
{ }
X4 + aX2 , X1 + X6 + aX2 , X1 X5 , X1 + aX2 , X2 , X3 , a R. (5.5)

Repeating these calculations, we construct the one-dimensional optimal


system for the Lie point symmetries in Equation (4.3) together with the
time translation and the scaling of C and obtain
{ }
X5 + aX2 , X1 + X6 + aX2 , X1 X4 , X1 + aX2 , X2 , X3 , a R. (5.6)

5.2 Construction of group-invariant solutions


The group-invariant solutions are constructed in this section. We further
classify the group-invariant solutions according to the elements of the one-
dimensional optimal systems. The reductions and group invariant solutions
are listed in Tables 3 and 4.
Denition:[25] u = H(x) is a group-invariant solution of a kth order PDE

8
corresponding to the appropriate admitted symmetry generator if and only
if u = H(x) satises

X(u H(x)) = 0, when u = H(x),


that is,
H
i (x, H(x)) = (x, H(x)). (5.7)
xi
The solution of Equation (5.7) is obtained from the characteristics equation

d 1 d 2 d
1
= 2
= ... = .
dx dx u
Classication of the group-invariant by the elements of the optimal system
in (5.5) and (5.6) are listed in Tables 3 and 4, respectively. Wherever they
appear k1 and k2 are arbitrary constants, 1 F1 (b, c; z) and U (b, c; z) are the
conuent hypergeometric functions, Ai(z) and Bi(z) are the Airy functions,
while Jn (z) and (z) represent the Bessel function of rst kind and Euler
gamma function respectively and erf(z) represents an error function. A
well documented review of such functions is presented by Abramowitz and
Stegun [27].

5.3 Some physical examples


5.3.1 Given constant dispersion coecient
Example (a)
Suppose a concentration C0 of a solute is supplied to a single point, in an
instant of time. We require to determine the subsequent concentration of
the pollution at various distances from where it was released (see e.g. [28]).
We would expect concentration to vanish at large distance, that is

C(, ) 0, as . (5.8)

The generator X4 in Equation (4.2) (also, an element of the one di-


mensional optimal systems given a = 0) leads to an invariant solution in
functional form given by
C = G(), (5.9)

where = e and G satises the equation

2 3 G + (4 2 1)G = 0,

and hence ( )
1
G = c1 + c2 erf .
2

9
5
=1
4.5 =5
=10
4 =20

3.5

C
2.5

1.5

0.5

0
0 5 10 15
R

Figure 1: Contaminant concentration prole along the radius. Here C0 = 5.

z
2
e d is the error function [27]. In terms of the original
2
erf(z) =
0
variable and subject to the boundary conditions, we obtain
( )
e
C = C0 erfc . (5.10)
2

Here, the erfc(.) is the complement error function dened by (1-erf(.)).


Solution (5.10) is depicted in Figure 1.
Total ux across R = Ra is given by
( )
1 C C0 eRa /4
2
Ra
C+ = C erfc . (5.11)
R R R=Ra
0
2 Ra

The contaminant ux (5.11) is depicted in Figure 2. Total ux across Ra =


4, increases and attens at large time.

Example (b)
The X6 -invariant solution is given in functional form as
( 2 )
1 e
C = exp G(), (5.12)
4

where

= e and G satises the ODE G + G = 0.

We impose the boundary conditions

C 0, and C = ( ), .

10
8

J
3

0
10 20 30 40 50 60 70 80 90 100

Figure 2: Solute ux across a xed radius. Here Ra = 4 and C0 = 10.

Innite concentration at the origin implies that there is a high supply of


contaminants at this point. Furthermore, contaminant concentration van-
ishes when time evolves. In terms of the original variables we obtain the
exact (group-invariant) solution given by
( 2 )
1 e
C = exp . (5.13)
4
Solution (5.13) is depicted in Figures 3 and 4. In Figure 3, a sharp peak
of concentration is observed shortly after = 0 and decreases at later stage.
This may be interpreted as an injection of contaminants at a single point,
that is, the concentration at a single point increases but due to diusion at
larger time it smoothes out. Note that here we have restricted our analysis
using symmetry generator X6 . This symmetry generator leads to simpler
and realistic exact solution. In Figure 4, we observe that concentration at
the origin decreases with time. Furthermore, this concentration vanishes at
large distances.

5.3.2 Given velocity-dependent dispersion coecient


Example (a) Steady state solution
The time translation leads to the analysis of the steady state contaminant
transport. Steady state solutions may be constructed subject to the follow-
ing imposed boundary conditions;

C = C0 , R=0 (5.14)

and
1 dC
C+ = 0, R = Ra . (5.15)
R dR

11
0.4

0.35

0.3

0.25
C
0.2

0.15

0.1

0.05

0
0 5 10 15 20 25 30 35 40

Figure 3: Contaminant concentration prole as time evolves.

1
=1
0.9 =2
=5
0.8

0.7

0.6
C

0.5

0.4

0.3

0.2

0.1

0
0 0.5 1 1.5 2 2.5 3 3.5 4
R

Figure 4: Contaminant concentration prole at various xed times.

12
1

0.9

0.8

0.7

0.6

C
0.5

0.4

0.3

0.2 p=1
p=1.5
0.1 p=2

0
0 1 2 3 4 5
R

Figure 5: Steady contaminant prole for concentration given in Equation


(5.16). Here = 1.

The boundary condition (5.14) implies that pollutants are supplied at the
origin, and boundary condition (5.15) correspond to the assumption that
pollutants are not carried though at some distance Ra , rather it accumulates
here. We obtain the exact solution
{( ) ( p )}
1 1 R
C = C0 1 + exp , (5.16)
p
where is given by
( ) ( )
Rap2 Rap
=1 1+ exp .
p

The solution (5.16) is depicted in Figures 5, 6 and 7. We observe in Figure


5 that concentration starts decreasing and converge to some value at large
distance for p = 2 than for lower values of p, whereas has an opposite
eect as shown in 6.

Example (b) Transient state solution


It is quite dicult to construct exact solutions for transient contaminant
transport subject to these boundary conditions (5.14) and (5.15). However,
if one assumes that at an initial time, say = 1, the concentration at the
point source is given by a constant and that this concentration vanishes at
large distances and prolonged periods, then using the symmetry combination
X1 + aX2 from Table 4, the group invariant solution is given by
{( ) }
1 1 4a
a
C = C0 e exp R2 , a < 0 and > 0. (5.17)
4

13
1

0.9

0.8

0.7

0.6

C 0.5

0.4

0.3

0.2 =1
=2
0.1
=100
0
0 1 2 3 4 5
R

Figure 6: Steady contaminant prole for concentration given in Equation


(5.16). Here R = Ra = 2.

0.8

0.6
C

0.4

0.2

0
6
5
4 4
3
2 2
1
0 0
p

Figure 7: Eects of and p on concentration prole for solution given in


Equation (5.16). Here R = Ra = 2.

14
1
=0.1
0.9 =0.5
=1
0.8

0.7

0.6

C 0.5

0.4

0.3

0.2

0.1

0
0 0.5 1 1.5 2 2.5 3 3.5 4
R

Figure 8: Concentration prole for solution given in Equation (5.17). Here


R = Ra = 2, C0 = 1 and = 1

0.14
=1
=1.5
0.12 =2

0.1

0.08
C

0.06

0.04

0.02

0
0 1 2 3 4 5
R

Figure 9: Eects of on the concentration prole for solution given in


Equation (5.17). Here R = Ra = 2, C0 = 1 and = 2.

15
Solution (5.17) is depicted in Figures 8 and 9.
In his work, Philip (1994) considered the instantaneous point source for
contaminant dispersion during radial water ow in porous media. Exact
solutions were constructed for the two- and three- dimensional models with
dispersion coecient depending on Peclet number. Here, we consider mod-
els in stream functions coordinates. Exact close-form (similarity) solutions
are constructed using the elements of the one dimensional optimal systems.
These new solutions may be viewed as representing the continual supply of
contaminant at a point (source) which are dispersed radially.

6 Some concluding remarks


In this paper, we have focused only on the two dimensional solute concentra-
tion eld within water from a single injection well. The considered problem
is a signicant improvement in the study of solute transport under radial
water background since we analyze the convection dispersion equation in
stream functions. We have observed that extra Lie point symmetries are
admitted when the dispersion coecient is a constant or when it is given
as a power law function of velocity, with exponent being given by two. We
have classied the group invariant solutions by the elements of the optimal
systems. In fact, new exact solutions are constructed. The symmetry (in-
variant) solution are obtainable when dispersion coecient is a constant or
is given by the Taylors theory of mixing in soils.

Acknowledgements
RJM wishes to thank the National Research Foundation of South Africa
under REDIBA program, for the continued generous nancial support.

Conict of interests
The authors declare that there is no conict of interest regarding the pub-
lication of this article.

References
[1] F. Catania, M. Massab`o and O. Paladino, Estimation of transport
and kinetic parameters using analytical solutions of the 2D advection-
dispersion-reaction model, Environmetrics, 17: 199-216, 2006.

[2] P. Broadbridge, J.M. Hill and J. Goard, Symmetry reductions of equa-


tions for solute transport in soil, Nonlinear Dynamics, 22(1): 15-27,
2000.

16
[3] P. Broadbridge, R.J. Moitsheki and M.P. Edwards, Analytical solutions
for two dimensional solute transport with velocity-dependent disper-
sion, Environmental Mechanics: Water, Mass and Energy transfer in
the biosphere, 129: 145-153, 2002.

[4] R.J. Moitsheki, P. Broadbridge and M.P. Edwards, Group invariant so-
lutions for two dimensional solute transport under realistic water ows,
Quaestiones Mathematicae, 29(1), 73-83, 2006.

[5] R.J. Moitsheki, Invariant solutions for transient solute transport in sat-
urated and unsaturated soil, PhD thesis, University of Wollongong,
2004.

[6] J.R. Philip, Some exact solutions of convection-diusion and diusion


equations, Water Resoursces research, 30(12): 3545-3551, 1994

[7] J.R. Craig and T. Heidlauf, Coordinate mapping of analytical contam-


inant transport solutions to non-uniform ow elds, Advances in water
Resources, 32: 353-360, 2009.

[8] J.E. Houseworth and J. Leem, A quasilinear model for solute transport
under unsaturated ow, Vadose Zone Journal, 8(4): 1031-1037, 2009.

[9] J-S Chen, C-F Ni and C-P Liang, Analytical power series solution-
s to the two-dimensional advection-dispersion equation with distance-
dependent dispersivities, Hydrological Processes, 22: 4670-4678, 2008.

[10] R.R. Yadav, D.K. jaiswal, H.K. Yadav and Gulrana, Analytical solu-
tions for temporally dependent dispersion through homogeneous porous
media, International Journal of Hydrology Science and Technology,
2(1): 101-115, 2012.

[11] E.M.E Zayed and H.A. Zedan, Autonomous forms and exact solutions
of equations of motion of polytropic gas, International Journal of The-
oratical Physics, 40(6): 1183-1196, 2001

[12] C.V Chrysikopoulos and Y. Sim, One-dimensional virus transport in ho-


mogeneous porous media with time-dependent distribution coecient,
Journal of Hydrology, 185: 199-219, 1996.

[13] Y. Sim and C.V Chrysikopoulos, Three-dimensional analytical models


for virus transport in saturated porous media, Transport in Porous
Media, 30: 87-112, 1998.

[14] Y. Sim and C.V Chrysikopoulos, Analytical solutions for solute trans-
port in saturated porous media media with semi-innite or nite thick-
ness, Advances in Water Resources, 22(5): 507-519, 1999.

17
[15] J.M. Thomas and C.V Chrysikopoulos, Experimental investigation
of acoustically enhanced colloid transport in water-saturated packed
columns, Journal of Colloid and Interface Science, 308: 200-207, 2007.

[16] J. Woods, C.T. Simmons and K.A. Narayan, Verication of black box
groundwater models, in EMAC98: Proceedings of the Third Biennial
Engineering Mathematics and Applications Conference, Institution of
Engineers Australia, Adelaide, 523-526, 1985.

[17] D. Hillel, Environmental soil physics, Academic Press, New York, 1998.

[18] J. Bear and A. Verruijt, Modeling groundwater ow and pollution, K-


luwer Academic Publishers, Norwell, MA, 1987.

[19] G. Segol, Classical groundwater simulation: proving and improving nu-


merical models, Englewood Clis, New Jersey: Prentice-Hall, 1994.

[20] J.M. Daz, YaLie users manual, Departamento de Matematicas, Uni-


versidad de Cadiz, 2000.

[21] A.C. Hearn, Reduce users manual version 3.4, Santa Monica, Califor-
nia: Rand Publication CP78, The Rand Corporation, 1985

[22] J. Sherring, Dimsym: symmetry determination and linear dierential


equation package, Latrobe University Mathematics Dept., Melbourne,
1993.

[23] P.J. Olver, Applications of Lie groups to dierential equations, New


York: Springer-verlag, (1972).

[24] N.H. Ibragimov, Elementary Lie group analysis and ordinary dieren-
tial equations, Wiley, New York, 1999.

[25] G.W. Bluman and S.C. Anco, Symmetry and integration methods for
dierential equations, New York: Springer-verlag, (2002).

[26] G.W. Bluman, A.F. Cheviakov and S.C. Anco, Applications of sym-
metry methods to partial dierential equations, New York: Springer-
verlag, (2010)

[27] M. Abramowitz and I.A. Stegun, Handbook of mathematicsl functions,


Dover Publication, INC, New York, (1972).

[28] G.R. Fulford and P Broadbridge, Industrial mathematics: case stud-


ies in the diusion of heat and matter, Cambridge university press,
Cambridge UK, 2002.

18
Table 2: Adjoint representation for the base vectors.

Ad X1 X2 X3 X4 X5 X6
X1 X1 X2 X3 2X1 + X4 2X3 + X5 42 X1 + 2X2 4X4 + X6

19
X2 X1 X2 X3 X4 X5 X6
X3 X1 X2 X3 X3 + X4 X2 + X5 2 X2 + X6 2X5
X4 2
e X1 X2 e X 3 X4 e X5 e2 X6
X5 X1 + 2X3 2 X2 X2 X2 + X3 X4 + X5 X5 X6
X6 X1 2X2 + 4X4 + 42 X6 X2 X3 + 2X5 X4 + 2X6 X5 X6
Table 3: Group-invariant solutions for Equation (2.4) with constant D.

Symmetry Reductions aR Group-invariant solutions

a
[ ( )
3 1
( a 1 1 )]
X4 + aX2 C = 2 G(), where = e2 a<0 C = a/2 21 1e k21 F1 a+1 , ;
2 4 e 2 + k 1 1 F 1 2 , 2 ; 4 e2 .
[ ( 2 )]
and G satises a=0 C = k2 2 k1 erf 1 2
[ 2 e
a/2 1 1
( 1a 3 1 ) ( a 1 1 )]
3 G + 14 (6 1)G 81 aG = 0. a>0 C= 2 e k 2 1 F1 2 , ;
2 4 e 2 + k 1 1 F 1 2 , 2 ; 4 e2 .
[ ]
X1 + aX2 C = ea G(), where = a<0 C = ea k1 cos ae k2 sin ae
and G satises a=0 C = e k1 +[ k2 . ]

G + G ae2 G = 0 a>0 C(, ) = ea k1 cosh ( ae ) ik2 sinh ( ae ) .

2 3
X1 + X5 C = e( e + 3 ) G(), where ( )( )
= 2 + e and G satises C = exp e + 32 3 Ai( 2 + e )k1 + Bi( 2 + e )k2
G G = 0.

20
[ [ ]
1 2 a
X1 + X6 C = exp 2(1+4 2 )
a<0 C = exp e2 (1+4 2 ) 4 ln (1 + 4 ) + 2 tan
1 (2 )
e [ ( )
] 2 2

a 1 3+ia 3 i
+aX2 14 ln (1 + 4 2 ) + 2 tan1 (2 ) G(), ei/2e (1+4 ) 2e2 (1+4 2) k F
2 1 1 4 , ;
2 e2 (1+4 2 )
( )]
i/2 e 2 (1+4 2 ) 1 3+ia 3 i
where = e 1 + 4 2 +e 2e2 (1+4 2 ) 1
k U 4 , 2 ; e2 (1+4 2 ) .
[ ]
1 2
and G satises a=0 C = exp e2 (1+4 2 ) 4 ln (1 + 4 )
[ ( ) ( )
1
6 G + 2 5 G + (1 a 2 )G = 0, 1 J 1 2 2 k1 34
2e 1+4 2 4 2e (1+4 )
( ) ]
1
(5)
+ 2e 11+ 2 J 1 2e2 (1+4 2) k 2 4 .
4
[ ]
1
a>0 C = exp e2 (1+4 2) 4 ln (1 + 4 2 ) + a arctan (2 )
2
[
2 2 1
( 3ia 3 i
)
ei/2e (1+4 ) 2e2 (1+4 2 ) k2 1 F1 4 , 2 ; e2 (1+4 2 )
]
2 2 1
( 3ia 3 i
)
+ei/2e (1+4 ) 2e2 (1+4 2) k 1 U 4 , ;
2 e2 (1+4 2 ) .
Table 4: Group-invariant solutions for Equation (2.4) given D(v) = e2 .

Symmetry Reductions aR Group-invariant solutions


[ 2 ] ( )
a 2
X5 + aX2 C = exp e4 4 + 8 ln ( ) G(), a<0 C = exp e4 4 + a ln8( )
[ ( ) ( ))]
k2
where = e 4 e2 F 4+a , 3 ; e4 + k F a , 1 ; e4
4 1 1 8 2 16 1 1 1 8 2 16 .
( )[ ( )]
2
2
and G satises a=0 C = exp e4 4 k2 k1 erf 4e .
( )[
e 2 a e2
2 6 G 41 (1 12 4 )G a>0 C = exp 4 4 + 8 ln ( ) k42
( ) ( )]
4a 3 e 4 a 1 e 4
a8 G = 0. 1 F1 8 , 2 ; 16 + k1 1 F1 8 , 2 ; 16 .

[ ] ( )
8 3 3
X1 + X4 C = exp 3 + e2 2 G(), C = exp 8
3 + e 2 2
{ 2 2 [ ( )
4 +e 1+4(4 2 +e2 )

21
where = 4 2 + e2 e 4 k1 Ai (4)4/3
( )]}
1+4(4 2 +e2 )
and G satises +k2 Bi (4)4/3
.
1
G 2 G 4 G = 0.
[ ( )
1
2
X1 + aX2 C= ea G(), a<0 C(, ) = ea
exp 1 4a)e
4 (1 k1
( ) ]
1

where = + exp 4 (1 + 1 4a)e2 k2
[ ( ) ]
e 2
and G satises a=0 C = k2 + exp 2 k1 .
( ) [( )
2 1

e
G + 2+ G + a e4 G = 0, a>0 a
C = e exp 2 k1
4 (1 1 + 4a)e
( ) ]
1
2
+ exp 4 (1 + 1 + 4a)e k2 .

X1 + X6 + aX2 leads to reductions which are unsolvable

You might also like