A Note On Polar Fluid Flow Through Porous Media With Forchheimer Effects by Hamdan

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Adv. Theor. Appl. Mech., Vol. 4, 2011, no.

3, 123 – 134

A Note on Polar Fluid Flow through Porous Media

with Forchheimer Effects

M. H. Hamdan

Department of Mathematical Sciences


University of New Brunswick
P.O. Box 5050, Saint John
New Brunswick, E2L 4L5, Canada
hamdan@unb.ca

Abstract

Equations governing the unsteady flow of an incompressible polar fluid through


isotropic porous media are derived using intrinsic volume averaging. Porous
microstructure is taken into account when modeling the Darcy resistance and the
Forchheimer effects. The model equations take into account the rotational
viscosity in the viscous shear and Forchheimer terms.

Keywords: Forchheimer Effects, Polar Fluid, Porous Media

1 Introduction
In previous work, [9, 15], equations governing the steady flow of an
incompressible polar fluid have been derived using intrinsic volume averaging.
The model equations take into account the Darcy resistance to the flow and ignore
the microscopic inertial effects that inevitably arise due to tortuosity of the flow
path in the porous sediment, and are important in the study of flow at high
Reynolds number (cf. [1,4,5,6,7,22,25,26]). In the current note we consider the
more general unsteady flow of a polar fluid through a porous sediment and
develop governing equations that take into account Forchheimer inertial effects in
addition to the Darcy resistance. It is assumed that the porosity of the medium is a
function of time and space; hence, this type of unsteady flow model might find
applications in the study of corrosion effects of porous sediments under polar fluid
flow. While most of the available models of polar and micropolar fluid flow
through porous media do not take into account the rotational viscosity in the
124 M. H. Hamdan

Darcy resistance and Forchheimer terms, (c.f. [10,16,19,20,21,22,23,24,25,26


27,28]), this work will illustrate how the rotational viscosity arises in model
formulations using the method of volume averaging. Typically, in the flow of a
polar fluid through free-space, if the rotational viscosity is zero then the model
equations reduce to those of the Navier-Stokes equations, [3]. Furthermore, the
rotational viscosity increases with a reduction in length scale, [3]. In the case of
polar fluid flow through a porous sediment, the inclusion of rotational viscosity in
the Darcy and Forchheimer terms is an imperative due to the small, pore-level
length scales associated with this type of flow domain.

2 Governing Equations
Motion of a polar fluid in free space has received considerable attention in
the literature due the the various applications it enjoys (c.f. [3,8,11,12,13,14] and
the references therein). This motion is described by the equations of continuity,
linear momentum, and angular momentum, which take the following form when
the fluid is incompressible, and body forces and body couple are absent, [3]:
r
∇ •V = 0 …(1)
r
⎡ ∂V r r⎤ r r
ρ ⎢ + ∇ • VV ⎥ = −∇p + ( μ + τ )∇ 2V + 2τ (∇ × G ) …(2)
⎣ ∂t ⎦
r
⎡ ∂G r r⎤ r r r r
ρk 2 ⎢ + ∇ • VG ⎥ = 2τ∇ × V + (α + β − γ )∇(∇ • G ) + ( β + γ )∇ 2G − 4τG …(3)
⎣ ∂t ⎦
r r
where V is the velocity vector field, G is the micro-rotation vector, p is the
pressure, t is the time variable, ρ is the fluid density, k is the radius of gyration
of a fluid element, in a given volume element, about the centroid of the volume
element, and α , β , γ , μ , and τ are viscosity coefficients ( τ is the rotational
viscosity, μ is the fluid viscosity, and α , β , γ are gradient viscosities).

In order to develop a set of field equations governing the flow of a polar fluid
through an isotropic porous medium, the equations governing the flow in free
space, i.e. equations (1), (2) and (3) are averaged over a Representative
Elementary Volume (REV), using the averaging Rules discussed in Appendix 1 of
this work. The effects of the porous microstructure on the flowing fluid will be
accounted for through the concept of a Representative Unit Cell (RUC),
introduced by Du Plessis and Masliyah, [6,7].

Typical conditions on the velocity and spin vectors are the no-slip and no-spin
assumptions on the solid matrix. These are implemented in this work and translate
r r r
to: V = G = 0 on the stationary solid matrix.
Polar fluid flow through porous media 125

3 The Averaged Governing Equations


The following intrinsically averaged equations are obtained by a direct
application of the averaging rules of Appendix 1 to equations (1), (2) and (3),
respectively:

r 1 r
∇ • ϕv + ∫ V • ndS = 0 …(4)
VS
r
⎡ ∂ϕv r r⎤ ρ r r r r 1
ρ⎢ + ∇ • ϕv v ⎥ + ∫ VV • ndS + ρ∇ • ϕ < V oV o > ϕ = − ϕ∇P − ∫ p o ndS +
⎣ ∂t ⎦ VS VS
r μ +τ r μ +τ r r 2τ
( μ + τ )∇ 2ϕv + ∫ V • ndS + ∫ ∇V • ndS + 2τ∇ × ϕg + ∫ G × ndS …(5)
V S V S V S
r
∂ϕg rr r r ρk 2 r r r
ρk 2
+ ρ k 2 ∇ • ϕv g + ρ k 2 ∇ • ϕ < V o G o > ϕ + ∫ VG • ndS = 2τ ∇ × ϕv
∂t V S
2τ r r (α + β − γ ) r o
− ∫ V × ndS . + (α + β − γ )ϕ∇ < ∇ • G > ϕ + ∫ (∇ • G ) ndS +
V S V S
r β +γ r β +γ r r
( β + γ )∇ 2ϕg + ∫ G • ndS + ∫ ∇G • ndS . − 4τ ϕg . …(6)
V S V S
The surface integrals and deviation terms in equations (4), (5) and (6) contain the
necessary information on the interactions between the flowing fluid and the
porous structure. These are analyzed in what follows.
For the incompressible flow at hand, continuity of mass flow translates into
r r
vanishing normal component of velocity, and the surface integral, ∫ V • ndS
S
vanishes. Furthermore, the no-slip and no-spin conditions on solid walls imply
that the terms involving velocity and microrotation vectors explicitly, vanish. The
rr r rr r
surface integrals ∫ VV • ndS , ∫ V • ndS , ∫ G × ndS , ∫ VG • ndS , ∫ V × ndS ,
S S S S S
r
∫ G • ndS are thus ignored in equations (4), (5) and (6).
S
r r r r
The volume filters ∇ • ϕ < V oV o >ϕ , appearing in (5), and ∇ • ϕ < V oG o >ϕ ,
appearing in (6), represent the hydrodynamic dispersion of the average velocity
and average micro-rotation, respectively. Du Plessis and Masliyah, [6,7], analyzed
the velocity deviation term and concluded that it is negligible in the absence of
high porosity and velocity gradients. Du Plessis and Diedericks, [5], justified
ignoring terms involving the divergence of the product of porosity and velocity
deviations on the bases of uniform average flow through porous media of uniform
porosities. Accordingly, the above volume filters are ignored in equations (5) and
(6). Equations (4), (5), and (6) thus reduce to the following forms, respectively:
126 M. H. Hamdan

r
∇ • ϕv = 0 . …(7)
r
⎡ ∂ϕv r r⎤ 1 r μ +τ r
ρ⎢ + ∇ • ϕv v ⎥ = − ϕ∇P − ∫ p o ndS + ( μ + τ )∇ 2ϕv + ∫ ∇V • ndS
⎣ ∂t ⎦ VS V S
r
+ 2τ∇ × ϕg …(8)
r
2 ⎡ ∂ϕg r r⎤ r 1 r r
ρk ⎢ + ∇ • ϕv g ⎥ = 2τ ∇ × ϕv + (α + β − γ )ϕ∇( ∇ • ϕg ) + ( β + γ )∇ 2ϕg
⎣ ∂t ⎦ ϕ
r (α + 2 β ) r o
− 4τ ϕg + ∫ (∇ • G ) ndS …(9)
V S
where, in arriving at equation (9), we made use of the following substitutions in
equation (6), obtained using averaging Rules (v) and (ii):
r
r < ∇•G > 1 r 1 r r
< ∇ • G >ϕ = = ∇ • ϕ < G >ϕ + ∫ G • n dS =
ϕ ϕ ϕV S
. …(10)
1 r 1 r
∇ • ϕ < G >ϕ = ∇ • ϕg
ϕ ϕ

r 1 r (α + β − γ ) r o
(α + β − γ )ϕ < ∇(∇ • G ) >ϕ = (α + β − γ )ϕ∇( ∇ • ϕg ) + ∫ (∇ • G ) ndS
ϕ V S
.
…(11)

4 Surface Integral Appearing in the Angular Momentum


Equation
The surface integral term appearing in (9), namely,
(α + 2β ) r o r
∫ (∇ • G ) ndS , arises in averaging the term containing ∇(∇ • G ) in the
V S
r
angular momentum equation (3) using Rule (ii), in which F = ∇ • G . This
r
surface integral involves the deviation of ∇ • G from its true value. Since
r r
∇ • G ≠ 0 , the deviation (∇ • G ) o is not necessarily zero for all porous structures,
and its quantification depends in part on an accurate description of the porous
microstructure. Let us first re-write Rule (ii) in the following superficial average
of F:
1
< ∇F >= ∇ < F > + ∫ F ndS . …(12)
VS
r
Now, taking F = ∇ • G in (12) we obtain:
r r 1 r
< ∇ (∇ • G ) > = ∇ < ∇ • G > + ∫ (∇ • G )ndS . …(13)
VS
Polar fluid flow through porous media 127

Clearly, the problem has been transformed into that of evaluating a surface
r r
integral of ∇ • G . Again, since ∇ • G ≠ 0 the above surface integral is not
necessarily zero for all porous microstructure. Using Green’s theorem, we can
r
replace the surface integral in (13) by the line integral ∫ G • nds , where C is any
C
piecewise smooth and closed boundary curve, bounding a solid surface within the
REV. This line integral vanishes, since it involves the micro-rotation vector
explicitly. The angular momentum equation (9) thus takes the following intrinsic
averaged form:
r
2 ⎡ ∂ϕg r r⎤ r 1 r r
ρk ⎢ + ∇ • ϕv g ⎥ = 2τ ∇ × ϕv + (α + β − γ )ϕ∇( ∇ • ϕg ) + ( β + γ )∇ 2ϕg
⎣ ∂t ⎦ ϕ
r
− 4τ ϕg . …(14)

5 Surface Integral Appearing in the Linear Momentum Equation


Effects of the porous matrix on the flowing fluid occur through the portion
of the surface area of the solid matrix that is in contact with the polar fluid. The
surface integral in (8) involves the pressure deviation term and the velocity
gradient, and contains the necessary information to quantify the pressure and
friction forces exerted by the porous matrix on the fluid. A similar surface integral
arises in the case of averaging the Navier-Stokes equations and, in fact, if τ = 0
then the surface integral in (8) is the same as the integral obtained in Whitaker,
[29], and referred to by Whitaker as a surface filter, and the integral obtained in
Du Plessis and Masliyah, [6,7]. By comparison with the averaged Navier-Stokes
momentum equations, the right-hand-side of (8) is identified with the force that
gives rise to Darcy resistance and the Forchheimer inertial terms.

As discussed in Whitaker, [29,30], Ma and Ruth, [17], Ruth and Ma, [18], and Du
Plessis and Diedericks, [5], the above surface integral can be decomposed into
two parts: one is a shear force integral (which accounts for the viscous drag
effects that predominate in the Darcy regime, that is, for small Reynolds number
flow), and the other is an inertial force integral (which accounts for inertial drag
effects that predominate in the Forchheimer regime, that is, for high Reynolds
number flow).

This type of integral has received extensive analysis, and its quantification gives
closure to the problem of flow through a porous structure, [5]. Quantification of
this surface integral depends on evaluating the surface integral of pressure
deviations, namely, ∫ p o ndS , and the surface integral of the directional derivative
S
r
r ∂V
in the direction of the normal vector, namely, ∫ ∇V • ndS = ∫ r dS . Clearly,
S S ∂n
128 M. H. Hamdan

these surface integrals are dependent on the flow velocity but independent of the
type of flowing fluid; hence, we will rely on what is already established in the
literature to evaluate the surface integral in equation (8).

To accomplish this, we let f1 be the velocity-independent viscous shear geometric


factor that depends on the geometry of the porous medium and gives rise to the
Darcy resistance, and f 2 the velocity-dependent inertial geometric factor that
gives rise to the Forchheimer inertial term. Following Du Plessis and Diedericks,
[5], the Churchill-Usagi total frictional effects, f, of the porous matrix on the fluid
may be expressed as:
f m = f1 + f 2
m m
…(15)
where m is a shifting factor that Du Plessis’ results, [4], have shown to produce
reasonable correlation when its value is unity. Furthermore, in terms of the factor
f1 , hydrodynamic permeability, η , is given by (cf. [5]):
ϕ
η= . …(16)
f1

The surface integral in (8) may then be written in terms of the superficial velocity
r r
average (namely, < V >= ϕ < V > ϕ = ϕv , as per Whitaker, [29,30]), as:
1 r r r
− ∫ [ − p o n + ( μ + τ )∇V • n]dS = − ( μ + τ ) fϕv = −( μ + τ )( f1 + f 2 )ϕv …(17)
V S

and equation (8) takes the form:

r
⎡ ∂ϕv r r⎤ r r r
ρ⎢ + ∇ • ϕv v ⎥ = − ϕ∇P + ( μ + τ )∇ 2ϕv + 2τ∇ × ϕg − ( μ + τ )( f1 + f 2 )ϕv . …(18)
⎣ ∂t ⎦

Equation (17) clearly demonstrates how the rotational viscosity, τ , enters into the
Darcy and Forchheimer terms. This viscosity is significant in the flow of polar
fluid through porous media due to the small length scales associated with the flow
domain, and the significant increase in viscosity with decreasing length scale.
Expressions for f1 and f 2 require a mathematical description of the porous matrix
and its microstructure. Du Plessis and Diedericks, [5], carried out extensive
analysis on evaluating these geometric factors for isotropic porous media, based
on Du Plessis and Masliyah’s, [6,7], concept of a Representative Unit Cell (RUC),
which they defined as the minimal REV in which the average properties of the
porous medium are embedded. The following porosity functions for granular and
sponge-like isotropic porous media, as given in Du Plessis and Diedericks, [5], are
adopted in this work:
Polar fluid flow through porous media 129

Material Geometric Factor Geometric Factor


Type f1 f2
(Forchheimer drag
coefficient is C d )
r
Granular 36(1 − ϕ ) 2 / 3 ρd ϕv Cd (1 − ϕ ) 2 / 3
d 2 [1 − (1 − ϕ )1 / 3 ][1 − (1 − ϕ ) 2 / 3 ] d 2 μ[1 − (1 − ϕ ) 2 / 3 ]2
r
Consolidated 36T (T − 1) ρd ϕv C d T (T − 1)
(Sponges and ϕd 2 d 2 μϕ (3 − T )
Metallic Foams)

Material Hydrodynamic Permeability


Type ϕ
η=
f1
Granular d ϕ[1 − (1 − ϕ )1 / 3 ][1 − (1 − ϕ ) 2 / 3 ]
2

36(1 − ϕ ) 2 / 3
Consolidated ϕ 2d 2
(Sponges and
36T (T − 1)
Metallic Foams)

Table 1. Geometric Factors and Hydrodynamic Permeability for Two Types of


Porous Media

An expression for the tortuosity component, T, appearing in Table 1 is given in


Du Plessis and Diedericks, [5], by the following form:

1 3 9 − 8ϕ ⎡ 4π 1 ⎧ 8ϕ 2 − 36ϕ + 27 ⎫⎤
= + cos ⎢ + cos −1 ⎨ ⎬⎥ . …(19)
T 4ϕ 2ϕ ⎢⎣ 3 3 ⎩ (9 − 8ϕ ) 3/ 2
⎭⎥⎦

Using the values of Table 1 for the geometric factors in equation (18) brings
closure to the intrinsic averaged linear momentum equations.

6 The Final Forms of the Field Equations


In light of the analyses above, the intrinsic volume averaged set of
equations governing the flow of a polar fluid through a variable-porosity, isotropic
porous structure, corresponding to equations (1), (2), and (3), are equations (7),
(18) and (14), respectively. The averaged equations take the following final form
r
in terms of the specific discharge, q , and the superficial average of the micro-
r r r r r
rotation vector, defined by: q = ϕ < V > ϕ = ϕv ; N = ϕ < G > ϕ = ϕg :
Conservation of Mass:
130 M. H. Hamdan

r
∇ • q = 0. …(20)

Linear Momentum Equation:


r r r
⎡ ∂q r q ⎤ r r
ρ ⎢ + (q • ∇)( )⎥ = − ϕ∇P + ( μ + τ )∇ 2 q + 2τ∇ × N − ( μ + τ )( f1 + f 2 )q …(21)
⎣ ∂t ϕ ⎦

Angular Momentum Equation:


r r
2 ⎡ ∂N N ⎤
r r 1 r r
ρk ⎢ + (q • ∇)( )⎥ = 2τ ∇ × q + (α + β − γ )ϕ∇( ∇ • N ) + ( β + γ )∇ 2 N
⎣ ∂t ϕ ⎦ ϕ
r
− 4τ N . …(22)

For constant-porosity media, the averaged equations (7), (18) and (14) take the
following forms, respectively, in terms of the intrinsic averaged velocity and
microrotation vector, obtained by dividing each of the equations throughout by ϕ :
r
∇•v = 0 …(23)
r
⎡ ∂v r r⎤ r r r
ρ⎢ + (v • ∇)v ⎥ = − ∇P + ( μ + τ )∇ 2 v + 2τ∇ × g − ( μ + τ )( f1 + f 2 )v …(24)
⎣ ∂t ⎦
r
2 ⎡ ∂g r r⎤ r r r r
ρk ⎢ + (v • ∇) g ⎥ = 2τ ∇ × v + (α + β − γ )∇ (∇ • g ) + ( β + γ )∇ 2 g − 4τg . …(25)
⎣ ∂t ⎦

Conclusions

In this work, we developed a set of equations governing the flow of a polar


fluid through an isotropic porous using the method of intrinsic volume averaging.
The equations are valid for variable-porosity media, hence cast in terms of the
specific discharge and the microrotation superficial average. For constant-porosity
media, the equations are written in their final form in terms of intrinsic averages.
The derived equations are valid for both low and high Reynolds number flows.
Expression for the Darcy resistance and Forchheimer drag are obtained based on
descriptions of the porous microstructure available in the porous media literature.
Both terms include the effects of rotational viscosity, which is significant at the
length scales dealt with in porous media; hence, ignoring the rotational viscosity
cannot be justified.

APPENDIX 1

Averaging Rules

Following Bachmat and Bear (1986), a Representative Elementary Volume, REV,


is a control volume that contains fluid and porous matrix in the same proportion as
Polar fluid flow through porous media 131

the whole porous medium. In other words, it is a control volume whose porosity is
the same as that of the whole porous medium. The porosity, ϕ , is defined as the
ratio of the pore volume to the bulk volume of the medium. In terms of the REV,
porosity is:
V
ϕ= ϕ …(1.1)
V
where Vϕ is the pore volume within the REV, which contains the fluid, and V is
the bulk volume of the REV. In terms of microscopic and macroscopic length
scales, l and L respectively, the REV is chosen such that

l 3 << V << L3 . …(1.2)

The volumetric phase average of a quantity F is defined as:


1
< F > = ∫ FdV …(1.3)
V Vφ
and the intrinsic phase average (that is, the volumetric average of F over the
effective pore space, Vφ ) is defined as:
1
< F >ϕ = ∫ FdV . …(1.4)
Vϕ Vφ

The relationship between the volumetric phase average and the intrinsic phase
average is obtained from equations (1.1), (1.3) and (1.4), and takes the form:

< F >= ϕ < F > ϕ . …(1.5)

Averaging theorems are then written in the following forms, [15]. Let F and H be
r
volumetrically additive scalar quantities, F a vector quantity, and c a constant
(whose average is itself), then:

(i)… < cF >= c < F > =c ϕ < F > ϕ .


1
(ii)… < ∇F >= ϕ∇ < F >ϕ +
V S∫ F o ndS

where S is the surface area of the solid matrix in the REV that is in contact with
r
the fluid, and n is the unit normal vector pointing into the solid. The quantity
F o = F − < F > is the deviation of the averaged quantity from its true
(microscopic) value.

(iii)… < F m H >=< F > m < H > = ϕ < F m H >ϕ = ϕ < F >ϕ mϕ < H >ϕ
(iv)… < FH >= ϕ < FH >ϕ = ϕ < F >ϕ < H >ϕ +ϕ < F o H o >ϕ
132 M. H. Hamdan

r r 1
(v) < ∇ • F >= ∇ • ϕ < F > ϕ + ∫ F • ndS .
V S
1
(vi)… < ∇ × F >= ∇ × ϕ < F > ϕ − ∫ F × ndS .
V S
(vii)… Due to the no-slip condition, a surface integral is zero if it contains the
fluid velocity vector explicitly.

(viii)… Due to the no-spin condition, a surface integral is zero if it contains the
spin vector explicitly. r
r
∂F ∂ r ∂ϕ < F > ϕ
(ix)… < >= < F >= .
∂t ∂t ∂t

Porosity, ϕ , may or may not be independent of time. It is time-dependent in cases


of flow of a corrosive material through a porous sediment. In either case, we
consider here the general case of porosity being a function of both position and
time. We have adopted the following notation in the current work for the phase-
averaged velocity vector and the averaged micro-rotation vector, respectively:
r r r r
< V >ϕ = v , < G > ϕ = g , and the following notation for intrinsic-averaged
pressure: < p > ϕ = P .

References
[1] B. Alazmi and K. Vafai, Analysis of variants within the porous media
transport models, J. Heat Transfer, 122 (2000), 303-326.

[2] Y. Bachmat and J. Bear, Macroscopic modeling of transport phenomena in


porous media, I: The Continuum Approach, Transp. in Porous Media, 1(1986),
213-240.

[3] S.C. Cowin, The theory of polar fluids, Adv. Appl. Mech., 14(1974), 279–347.

[4] J.P. Du Plessis, Analytical quantification of coefficients in the Ergun equation


for fluid friction in a packed bed, Transp. in Porous Media, 16(1994), 189-207.

[5] J.P. Du Plessis and G.P.J. Diedericks, Pore-scale modeling of interstitial


phenomena. In “Fluid Transport in Porous Media”, J.P. Du Plessis, ed., 61-104,
Computational Mechanics Publications, 1997.

[6] J.P. Du Plessis and J.H. Masliyah, Mathematical modeling of flow through
consolidated isotropic porous media, Transp. in Porous Media, 3(1988), 145-161.
Polar fluid flow through porous media 133

[7] J.P. Du Plessis and J.H. Masliyah, Flow through isotropic granular porous
media, Transp. in Porous Media, 6(1991), 207-221.

[8] C.F. Chan Man Fong, M.N. Farah, M.H. Hamdan and M.T. Kamel, Polar fluid
flow between two eccentric cylinders: inertial effects, Int. J. Pure and Applied
Mathematics, 23#2(2005), 267-283.

[9] M.H. Hamdan and J.D. Rehkopf, Micropolar fluid flow through porous media,
Engineering Mech. Symposium, CSCE, 72-77, Winnipeg, Manitoba, Canada,
1994.

[10] P.S. Hiremath and P.M. Patil, Free convection effects on the oscillatory flow
of a couple stress fluid through a porous medium, Acta Mechanica, 98(1993),
143-158.

[11] M.T. Kamel and C.F. Chan Man Fong, Micropolar fluid flow between two
eccentric coaxially rotating spheres, Acta Mechanica, 99(1993), 155-171.

[12] M.T. Kamel and M.H. Hamdan, Aspects of thin-film polar fluid lubrication,
J. Applied Mathematics and Computation, 80#1(1996), 33-42.

[13] M.T. Kamel and P.N. Kaloni, Cosserat fluid flow through a curved pipe,
ZAMP, 28(1977), 551-576.

[14] M.T. Kamel, P.N. Kaloni and E.M. Tory, Two-dimensional internal flows of
polar fluids, J. Rheol., 23(1979), 141-150.

[15] M.T. Kamel, D. Roach and M.H. Hamdan, On the Micropolar Fluid Flow
through Porous Media, In: Mathematical Methods, System Theory and Control,
Proceedings of the 11th MAMECTIS ’09, ISBN 978-960-474-094-9, 190-197,
WSEAS Press, 2009.

[16] Y.J. Kim, Heat and mass transfer in MHD micropolar flow over a vertical
moving porous plate in a porous medium, Transp. in Porous Media, 56(2004), 17-
37.

[17] H. Ma and D.W. Ruth, The microscopic analysis of high Forchheimer


number flow in porous media, Transp. in Porous Media, 13(1993), 139-160.

[18] D.W. Ruth and H. Ma, Numerical analysis of viscous, incompressible flow in
a diverging-converging RUC, Transp. in Porous Media, 13(1993), 161-177.

[19] P.M. Patil, Effects of free convection on the oscillatory flow of a polar fluid
through a porous medium in the presence of variable wall heat flux, J.
Engineering Physics and Thermophysics, 81#5(2008), 905-922.
134 M. H. Hamdan

[20] P.M. Patil and P.S. Hiremath, A note on the effects of couple stresses on the
flow through a porous medium, Rheol. Acta, 31#2(1992), 206–207.

[21] A. Raptis, Effects of couple stresses on the flow through a porous medium,
Rheol. Acta, 21(1982), 736–737.

[22] A. Raptis, Unsteady free convective flow through a porous medium, Int. J.
Engineering Sciences, 21(1983), 345-348.

[23] A. Raptis, Boundary layer flow of a micropolar fluid through a porous


medium, Journal of Porous Media, 3#1(2000), 95–97.

[24] A. Raptis, N. Kafousias and C. Massalas, Free convection and mass transfer
flow through a porous medium bounded by an infinite vertical porous plate with
constant heat flux, ZAMM, 62(1982), 489-491.

[25] A. Raptis and C.R. Perdikis, Oscillatory flow through a porous medium by
the presence of free Convective flow, Int. J. Engineering Sciences, 23(1985), 51-
55.

[26] A. Raptis and H.S. Takhar, Polar fluid through a porous medium, Acta
Mechanica, 135(1999), 91–93.

[27] A. Raptis, G. Tzivanidis and N. Kafousias, Free convection and mass transfer
flow through a porous medium bounded by an infinite vertical limiting surface
with constant suction, Letters Heat Mass Transfer, 8(1981), 417-424.

[28] P.G. Siddeshwar and C.V. Sri Krishna, Linear and non-linear analyses of
convection in a micropolar fluid occupying a porous medium, Int. J. Non-linear
Mechanics, 38(2003), 1561–1579.

[29] S. Whitaker, Volume averaging of transport equations, In “Fluid Transport


in Porous Media”, J.P. Du Plessis, ed., 1-60, Computational Mechanics
Publications, 1997.

[30] S. Whitaker, The Method of Volume Averaging. Kluwer Academic


Publishers, Dordrecht, 1999.

Received: June. 2010

You might also like