You are on page 1of 27

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257668931

Flow of Particulate-Fluid Suspension in a Channel with Porous Walls

Article  in  Transport in Porous Media · May 2013


DOI: 10.1007/s11242-013-0137-y

CITATIONS READS

8 218

3 authors, including:

Zhaoqin Huang
China University of Petroleum, Qingdao, China
98 PUBLICATIONS   1,985 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

An efficient numerical simulation for discrete fracture model based on mixed multi-scale finite element method View project

All content following this page was uploaded by Zhaoqin Huang on 28 April 2015.

The user has requested enhancement of the downloaded file.


Transp Porous Med (2013) 98:147–172
DOI 10.1007/s11242-013-0137-y

Flow of Particulate-Fluid Suspension in a Channel


with Porous Walls

Jun Yao · Ke Tao · Zhaoqin Huang

Received: 23 September 2012 / Accepted: 30 January 2013 / Published online: 19 February 2013
© Springer Science+Business Media Dordrecht 2013

Abstract The problem of the dispersed particulate-fluid two-phase flow in a channel with
permeable walls under the effect of the Beavers and Joseph slip boundary condition is con-
cerned in this paper. The analytical solution has been derived for the longitude pressure dif-
ference, stream functions, and the velocity distribution with the perturbation method based on
a small width to length ratio of the channel. The graphical results for pressure, velocity, and
stream function are presented and the effects of geometrical coefficients, the slip parameter
and the volume fraction density on the pressure variation, the streamline structure and the
velocity distribution are evaluated numerically and discussed. It is shown that the sinusoidal
channel, accompanied by a higher friction factor, has higher pressure drop than that of the
parallel-plate channel under fully developed flow conditions due to the wall-induced curva-
ture effect. The increment of the channel’s width to the length ratio will remarkably increase
the flow rate because of the enlargement of the flow area in the channel. At low Reynolds
number ranging from 0 to 65, the fluids move forward smoothly following the shape of the
channel. Moreover, the slip boundary condition will notably increase the fluid velocity and
the decrease of the slip parameter leads to the increment of the velocity magnitude across the
channel. The fluid-phase axial velocity decreases with the increment of the volume fraction
density.

Keywords Two-phase flow · Particulate-fluid suspension · Permeable boundary ·


Beavers and Joseph slip condition · Perturbation method

1 Introduction

Transport phenomena in a channel with permeable walls exist in numerous industrial, envi-
ronmental, and biological applications. Such applications involve the extraction of crude oil

J. Yao (B) · K. Tao · Z. Huang


Department of Reservoir Engineering, School of Petroleum Engineering,
China University of Petroleum (Huadong), Qingdao 266555, China
e-mail: RCOGFR_UPC@126.com

123
148 J. Yao et al.

from reservoirs (Arbogast and Lehr 2006; Huang et al. 2010), fuel cells (Huang et al. 2009),
flow through oil filters (Iliev and Laptev 2004), surface and groundwater flow (Furman 2008),
contaminant transport in fractured porous media (Sudicky and Frind 1982), blood flow in
a capillary (Mariamma and Majhi 2000), transfer of therapeutic agents (Baber 2009), etc.
The corresponding transport mechanism has been intensively investigated in the last decades
(Shavit 2009). However, most of the former researchers focused on the single-phase flow.
A large variety of two-phase flows exist and they can be classified into three general
categories according to the shape of the interface (Ishii 1975): separated flow, mixed flow,
and dispersed flow. In this paper, we only refer to the dispersed two-phase flow. Typical cases
of dispersed flows are the oil-water flow in the carbonate reservoir, the gas-water flow in fuel
cells and the sediment transport in river. For the circumstance of dispersed flow, one-phase
is assumed to be dilute and comprised of spherical inclusions such as solid particles, liquid
droplets, and gas bubbles carried by the other phase. The influence of the particulate objects
on the fluid-phase is represented by the source terms added to the right side of the Navier-
Stokes equation. If the particulate-phase is thought to be a continuum, then the two-phase
model consisted of two continuity and two momentum equations (Drew 1983; Kleinstreuer
2003) is acceptable.
The flow structure considered in this paper is a free flow channel bounded by surrounding
porous medium. As is well-known, the existence of permeable boundary will enlarge the
fluid velocity in the channel and the classical no-slip condition at the fluid/porous interface
is improper. Different boundary conditions at the fluid/porous interface have been derived
until now, but no general conclusion has been made. The work of Beavers and Joseph (1967)
was among the earlier attempts to find a suitable interface condition. They performed an
experimental and analytical investigation of fluid flow past a porous material and concluded
that the mass flux through the channel is much larger than that predicted by the parallel
flow with a lower impermeable wall. In order to describe the forced flow in the composite
channel, a Stokes flow is considered in the free fluid region while the momentum transport
in the homogeneous porous medium is described by Darcy’s law. Due to the different order
of the corresponding partial differential equations, an ad hoc slip boundary condition, the
Beavers and Joseph interface condition, is proposed,

du  α
= 1/2 (u i − u D ) (1)
dy  y=0+ k
Where u denotes the tangential fluid velocity, u i is the tangential fluid velocity at the interface,
u D is the Darcy velocity in the bulk; k is the permeability of the porous medium and α is the
empirical dimensionless coefficient which depends on the structure of the permeable material
within the boundary region and ranges from 0.1 to 4.
Larson and Higdon (1986, 1987) considered a simple shear flow through a semi-infinite,
periodic configuration made of cylinders, with flow either perpendicular or parallel to the
axes of the cylinders. The predetermination of the nominal interface of the medium, the plane
passing through the axes of the top row of cylinders, is needed to determine the slip velocity.
They stated that the slip velocity is very sensitive to the position of the interface and that
determining a consistent value for α is impossible because of the uncertainty of the interfacial
location. In 1992, Sahraoui and Kaviany (1992) pointed out that α is a complex function of
the porosity of the medium, the fluid Reynolds number, the channel width, deviation distance
from the interface, bulk flow direction, and surface structure of the porous medium.
Subsequently, Le Bars and Worster (2006) gave a physical interpretation for the question-
able dimensionless coefficient α for their limiting Darcy-Brinkman case: α is equal to the
square root of the porosity. Neale and Nader (1974) objected the velocity slip situation at the

123
Flow of Particulate-Fluid Suspension 149

interface and stated that Brinkman equation is mathematically and physically preferable to
Darcy’s law when considering transition region effects in porous media. They established a
relationship between the reduced viscosity and the structural parameter α:μe /μ = α 2 where
μe denotes the effective viscosity in the Brinkman equation and μ is the fluid viscosity.
However, Nield (1991, 2009) himself recommended to drop the Brinkman term and use the
Beavers and Joseph boundary condition.
Actually the length-scale constraints imposed on the macroscopic conservation equa-
tions in both homogeneous fluid and porous regions are not satisfied in the fluid-porous
inter-region (Whitaker 1999). To overcome this difficulty, the volume averaging method is
utilized to derive jump boundary conditions involving effective coefficients computed from
the solution of the corresponding closure problems (Wood et al. 2000; Valdés-Parada et al.
2006; Morales-Zárate et al. 2008; Valdés-Parada et al. 2009a,b; Aguilar-Madera et al. 2011).
At the fluid/porous interface, the jump stress condition is (Ochoa-Tapia and Whitaker 1995a)
 
du  du  μ
μe −μ = β 1/2 u i (2)
dy  y=0− dy  y=0+ k
Here β is the stress jump coefficient and it may be positive or negative but must be of
the order of 1 (Ochoa-Tapia and Whitaker 1995b). Valdés-Parada et al. (2007a) related this
parameter to a mixed stress tensor combining the global and the Brinkman stresses at the
inter-region. The model consisting of macroscopic averaging equations coupled with the cor-
responding interface conditions is referred as the two-domain approach (TDA). Goyeau et
al. (2003) introduced one-domain approach (ODA), which considers the porous medium as a
pseudo fluid and employs one single equation to describe the fluid flow both in the free fluid
region and porous medium, to obtain an explicit expression for the jump parameter. They
concluded that β is related to the continuous spatial variations of the porous structure near
the interface. Recently, Chandesris and Jamet (2006) established a stress jump condition by
solving the ODA with matched asymptotic expansions. The variations of the porosity and
permeability in the transition region were expressed through two excess quantities. Further-
more, Jamet et al. (2007, 2009) analyzed the physical meaning based on two-step up-scaling
method and they showed that jump parameters can be interpreted as surface-excess quantities
whose value depend linearly on the position of the discontinuous interface.
Some authors have considered the influence of slip effects on the flow. Ali et al. (2008)
have studied the slip effects on the peristaltic transport of MHD fluid with variable viscosity.
Hayat et al. (2008) considered the effect of partial slip on the peristaltic flow in a porous
medium. Mishra and Rao (2005) study the peristaltic transport of two-layered system. The
Brinkman equation is applied in the porous medium and the shear stress jump boundary
condition derived by Ochoa-Tapia and Whitaker is used at the interface. RaviKumar et al.
(2010) studied the peristaltic pumping in a finite lengths tube with permeable wall and
they used the Beavers and Joseph slip condition. In this paper, the slip boundary condition,
i.e., the Beavers and Joseph boundary condition, is used and the jump boundary condition
will be covered in future work.
However, most of the former studies focused on the single-phase flow. To the best of
the authors’ knowledge, no investigation has been made yet to analyze the influence of
slip on the dispersed two- phase flow, which is the purpose of this article. In Sect. 2, the
mathematical formulation is discussed and the non-dimensional dispersed two-phase model
using continuum approach is built. Section 3 uses the perturbation technique to acquire the
analytical solution for the two-phase model. In Sect. 4, the dimensionless pressure difference
between the inlet and the outlet of the channel is calculated by integrating the pressure gradient
along the center of the channel. The verification of the perturbation solution is conducted in

123
150 J. Yao et al.

Sect. 5; the numerical results and discussion are presented in Sect. 6. The conclusions have
been summarized in Sect. 7.

2 Mathematical Model

Consider a symmetric two-dimensional channel with length L and width 2h occupied by


a mixture of particulates and incompressible Newtonian viscous fluid in Fig. 1. The upper
interface between the channel and the porous medium is expressed as

H ∗ (x ∗ ) = h + h ∗1 (x ∗ ) (3)

h ∗1 (x ∗ ) is assigned different expressions according to the shape of the channel walls. In


the case of flow in fractured media, rock fractures are usually described by the parallel plate
model with smooth surfaces and an aperture 2h (Weerakone et al. 2011) as shown in Fig. 1a.
However, real fractures are rough and in partial contact (Brown et al. 1995; Weerakone et al.
2011). Taking into account the surface roughness, we construct the two-dimensional rough
channels with sinusoidal fluctuation on each wall in Fig. 1(b).
As shown in Fig. 1, the free flow region f filled with the dispersed particulate-fluid
two-phase is sandwiched by the surrounding homogeneous porous medium region p where
the fluid-phase seeps. These two regions are separated by the intermediate sharp interfaces.
Here, we assume that the particulate size should be large enough so that the particulates in
the channel cannot penetrate into the pores of the porous medium. To ensure the correctness
of this assumption, the characteristic length scales of the systems must satisfy

lm << ld < lf < r0 << L (4)

where lm , ld and lf are the length scales associated to the micro-pores, particulate-phase, and
fluid-phase respectively. r0 denotes the size of the averaging volume.
In the porous media, the fluid flow can be described by the Darcy’s law
k d p∗
u ∗D = − (5)
μf dx ∗

(a) (b)
y* y*
Porous Medium
Porous Medium

A Interface
Interface
h h
Free Flow Region
0 x* 0
Free Flow Region
x*

Interface

Interface
L

Porous Medium L
Porous Medium

Fig. 1 The schematic diagrams of the problem with a h ∗1 (x ∗ ) = 0 and b h ∗1 (x ∗ ) = A sin[2π(x ∗ /L − 1/4)]
respectively

123
Flow of Particulate-Fluid Suspension 151

where u ∗D denotes the Darcy velocity, k is the permeability of the porous medium and the fluid
viscosity μf . The negative sign signifies the velocity direction is opposite to the direction of
the pressure gradient. An important assumption for the Darcy model is that the pore-scale
Reynolds number for the porous matrix Rep is sufficiently small. The inertial effects in the
porous media can be neglected if Rep is lower than 0.1 (Vafai and Kim 1990).

ρf u ∗D k
Rep = (6)
μf
In the channels, the Mass and momentum conservation equations for the fluid-phase and
particulate-phase acquired using the continuum method are expressed as follows (Drew 1983;
Srivastava and Srivastava 1997; Mekheimer et al. 1998; Jiménez-Lozano et al. 2011).

Fluid-Phase
 ∗
∂u ∗f ∗
∗ ∂u f ∗ ∂u f ∂ p∗
(1 − C)ρf + u +v = −(1 − C) +(1 − C)μs ∇ 2 u ∗f + C S(u ∗p − u ∗f )
∂t ∗ f
∂x∗ f
∂ y∗ ∂x∗
(7)
 ∗ ∗ ∗ ∗
∂vf ∂v ∂v ∂ p
(1 − C)ρf + u ∗f f∗ + vf∗ f∗ = −(1 − C) ∗ +(1−C)μs ∇ 2 vf∗ +C S(vp∗ −vf∗ )
∂t ∗ ∂x ∂y ∂y
(8)
∂   ∂  
(1 − C)u ∗f + ∗ (1 − C)vf∗ = 0 (9)
∂x∗ ∂y
Particulate-Phase:
 
∂u ∗p ∂u ∗p ∂u ∗p ∂ p∗
Cρp + u ∗p + vp∗ = −C + C S(u ∗f − u ∗p ) (10)
∂t ∗ ∂x∗ ∂ y∗ ∂x∗
 
∂vp∗ ∂vp∗ ∂vp∗
∂ p∗
Cρp + u ∗p + vp∗ + C S(vf∗ − vp∗ )
= −C (11)
∂t ∗ ∂x∗ ∂ y∗
∂ y∗
∂  ∗ ∂  ∗
Cu + Cvp = 0 (12)
∂x∗ p
∂ y∗
In the above Eqs. 7–12, x ∗ and y ∗ are Cartesian coordinates, (u ∗f , vf∗ ) denotes the fluid-phase
velocity components corresponding to x coordinate and y coordinate respectively; (u ∗, ∗
p vp )
denotes the particulate-phase velocities. C is the volume fraction density of the particulates,
ρf , ρp are the actual densities of the fluid-phase and particulate-phase, and Cρp, (1 − C)ρf
are the densities of the particulate-phase and the fluid-phase in the mixture. p ∗ represents the
pressure, μs is the mixture viscosity, and S denotes the drag coefficient of interaction for the
force exerted by one phase on the other.
Here the volume fraction density C is chosen to be constant which is acceptable for small
concentration of particulates. And we assume the small concentration of the particulates
leads to the ignorance of the interaction among particulates and the interaction between the
particulates and the porous medium. The expression for the drag coefficient S is selected as
(Srivastava and Srivastava 1997)
9 μf
S= λ (C)
2 r2
1
4 + 3(8C − 3C 2 ) 2 + 3C
λ (C) = (13)
(2 − 3C)2

123
152 J. Yao et al.

In Eq. 13, μf is the fluid viscosity and r is radius of the particulates. λ(C) is a function
obtained by Tam (1969) to modify the classical Stokes drag to account for the finite particulate
fractional volume. For the present problem, the empirical relation acquired by Charm and
Kurland (1974) is used to calculate the suspension viscosity μs .
1
μs = μf
1 − λq C
 
1107
λq = 0.07 exp 2.49C + exp(−1.69C) (14)
T
where T is the absolute temperature.
For this coupled flow problem, the boundary conditions are:
∂u ∗f ∂u ∗p
= = 0 , vf∗ = vp∗ = 0 at y ∗ = 0
∂ y∗ ∂ y∗

k ∂u ∗f
u ∗f = + u ∗D at y ∗ = H ∗ (x ∗ ) (15)
α ∂ y∗
The first condition in Eq. 15 implies that both the fluid phase x-velocity and the particulate
phase x-velocity reach its maximum at the center of the channel and its corresponding
y-velocities remain zero. The second condition is the recomposition of the Beavers and
Joseph slip boundary condition.
Introduce stream functions as follows:
∂ f∗ ∂ p∗ ∂ f∗ ∂ p∗
u ∗f = , u ∗
= , v ∗
= − , v ∗
= − (16)
∂ y∗ p
∂ y∗ f
∂x∗ p
∂x∗
and the dimensionless variables
x∗ y∗ u ∗f u ∗p vf∗
x= L , y= h , uf = u ∗D , up = u ∗D , vf = u ∗D ,
vp∗ f∗ p∗ p∗
vp = u ∗D , f = hu ∗D , p = hu ∗D , p= ρf u ∗2
, ε= h
L (17)
D

The suspension parameters are defined as follows:


u ∗D hρf Sh 2 Sh 2 ρf
R= , M= , N= (18)
(1 − C)μs (1 − C)μs (1 − C)ρp μs
With the above dimensionless variables, the dimensionless stream functions can be written
as
∂ f ∂ p ∂ f ∂ p
uf = , up = , vf = −ε , vp = −ε (19)
∂y ∂y ∂x ∂x
The continuity equations are satisfied automatically.
Eliminating pressure terms by cross differentiation, the dimensionless steady-state equa-
tions of motion in terms of stream function are given by
 
(1 − C)ε R f y ∇˜ 2 f x − f x ∇˜ 2 f y = ∇˜ 2 ∇˜ 2 f + C M(∇˜ 2 p − ∇˜ 2 f )
  (20)
Cε R py ∇˜ 2 px − px ∇˜ 2 py = C N (∇˜ 2 f − ∇˜ 2 p )

where
∂2 ∂2
∇˜ 2 = ε 2 2 + 2
∂x ∂y

123
Flow of Particulate-Fluid Suspension 153

The dimensionless formulation for the upper surrounding permeable wall is



1 0 ≤ x ≤ 1 for parallel channel
H (x) = (21)
1 + ϕ sin [2π(x − 1/4)] 0 ≤ x ≤ 1 for rough channel

where ϕ=A/h is the dimensionless amplitude ratio of the sinusoidal channel. The fluid-phase
volume flow rate in the half of the channel is


(x)
H

Q ∗f = (1 − C)u ∗D h u f dy (22)
0

Define the dimensionless flow rate


Q ∗f
F= (23)
(1 − C)u ∗D h

Then, from Eq. 19


(x)
H
∂ f
F= dy = f (H ) − f (0) (24)
∂y
0

If we choose the zero value of the streamline at y = 0, then at the fluid/porous interface we
have

F = f (H ) (25)

So the dimensionless boundary conditions for this problem are

f = p = 0, f yy = pyy = 0 at y = 0
f = F, f y = β f yy + 1 at y = H (26)

Here β = −k 1/2 / hα is a dimensionless parameter, named slip parameter here.


The fluid-phase Reynolds number is defined as

ρf Ui lv ρf Q ∗f
Ref = = (27)
μf μf

where lv = h is the dimensional characteristic length and Ui = Q ∗f / h is the dimensional


average velocity in the channel. The friction factor f is calculated according to its standard
definition (Niceno and Nobile 2001)

Gh Gh 3
f = = (28)
2ρa Ui2 2ρa Q ∗2
f

where G = −d p/dx is the negative pressure gradient and ρa is the average density calculated
by the following equation

ρa = Cρp + (1 − C)ρf (29)

123
154 J. Yao et al.

3 Perturbation Solution

The perturbation method is employed to solve the problem based on the small channel width
to length ratio ε. Expand the flow quantities in a power series of the small parameter ε
f = f0 + ε f1 + ε 2 f2 + · · ·
p = p0 + ε p1 + ε 2 p2 + · · ·
u = u 0 + εu 1 + ε 2 u 2 + · · ·
v = v0 + εv1 + ε 2 v2 + · · · (30)

F = F0 + ε F1 + ε 2 F2 + · · ·
∂P ∂ P0 ∂ P1 ∂ P2
= +ε + ε2 + ···
∂x ∂x ∂x ∂x
Substituting Eq. 30 into Eqs. 20 and 26 and collecting terms of equal powers of ε, and
equating identical powers of ε on both sides of the equations, the equations are reduced to a
series of linear equations.

3.1 The Zeroth-Order System

The zero-order equations are



f0yyyy + C M p0yy − f0yy = 0
(31)
C N f0yy − p0yy = 0
With the boundary conditions
f0 = p0 = 0 , f0yy = p0yy = 0
(32)
f0 = F0 , f0y = β f0yy + 1

3.2 The First-Order System

The first-order equations are


 
(1 − C)R f0y f0x yy − f0x f0yyy = f1yyyy + C M p1yy − f1yy
  (33)
C R p0y p0x yy − p0x p0yyy = C N f1yy − p1yy
With boundary conditions
f1 = p1 = 0 , f1y y = p1yy = 0
(34)
f1 = F1 , f1y = β f1yy
The systems of Eqs. 31–34 are solved order by order as shown in 7.
The expressions for the stream function f (x, y) and p (x, y) up to first-order are
(1)
f = f0 + ε f1
(35)
p(1) = p0 + ε p1

Express the calculation result up to first order of the dimensionless flow rate F (1)
F (1) = F0 + ε F1 (36)

123
Flow of Particulate-Fluid Suspension 155

Then substituting F0 = F (1) − ε F1 into the assembled solutions and neglecting terms
greater than O(ε), the closed-form solution in terms of the flow rate are obtained. The corre-
sponding dimensionless velocities can be achieved with relation expressed in Eq. 16.

4 Pressure Difference

Substituting Eq. 30 into the dimensionless motion equations and equating the coefficients
of equal powers of ε on both sides, we get the zero- and first-order terms for the pressure
gradient
∂ p0
= g1 f0yyy + g2 p0y − f0y
∂x (37)
∂ p1 L
= f0x f0yy − f0y f0yx + g1 f1yyy + g2 p1y − f1y
∂x h
where
μs L C SL
g1 = , g2 = (38)
h 2 ρf u D (1 − C)ρf u D
The dimensionless pressure difference in the longitude direction of the channel is defined
as

1
dp
P = dx (39)
dx
0

Since the pressure difference in the longitude direction is independent of y, the integral
in Eq. 39 can be evaluated along the axis at y = 0. Substituting Eq. 37 into Eq. 39 and
integrating from 0 to 1, we have the expression for the pressure drop using the zero- and
first-order terms for the pressure gradient,

1
(1) ∂ p0 ∂ p1
P = ( +ε ) dx (40)
∂x ∂x
0

Similarly, substituting F0 = F (1) − ε F1 into the above Eq. 40 and neglecting terms
greater than O(ε), the dimensionless pressure gradient in terms of the flow rate is obtained
as shown in 7.

5 Verification of the Solution

To verify the solution, the problem of the incompressible Newton single-phase flow in the
parallel plate model is tested because it is the case whose analytical solution is known.
A fully developed flow between two parallel plates similar to the channel in Fig. 1a, 2h apart
and L long, is considered. Both the top plate and the bottom one are fixed. Besides, a pressure
gradient drives the flow in the x direction with the boundary conditions: u = 0 at y = 1,
du/dy = 0 at y = 0, where the dimensionless height is y = y ∗ / h, the dimensionless
velocity u = μu ∗ /Gh 2 in which G = −d p/dx is the negative pressure gradient and
μ is the viscosity of the fluid. The exact velocity distribution can be easily obtained from the
dimensionless Stokes equation,

123
156 J. Yao et al.

Fig. 2 Comparison between velocity profiles obtained by using the exact analytical solution and the
perturbation solution from this work

1 1
u = − y2 + (41)
2 2

The velocity profile in the parallel plates is also obtained using the solution in Sect. 3 and 4.
In this calculation, the volume fraction density C is zero and the slip parameter is also set to
zero to ensure the non-slip boundary condition at y = 0. Moreover, The Reynolds number
Ref is 1,000, the flow rate Q ∗f is 0.001 m2 /s, and the negative pressure gradient G = 3 Pa/m.
Fig. 2 compares the exact analytical velocity profile and the velocity distribution obtained
using the perturbation solution in Sect. 3 and 4 in the transverse direction. It is clearly shown
that the results agree well.

6 Results and Discussions

Through the perturbation method based on the small channel width to length ratio ε, we
have procured the analytical solutions for the particulate-fluid flow in channels with perme-
able boundaries. The behavior of this dispersed two-phase flow is affected by the follow-
ing parameters: the channel’s amplitude ratio ϕ (parallel or rough), the channel’s width to
length ratio ε, the slip parameter β, the volume fraction density C, and the fluid Reynolds
number Ref .
In this section, our aim is to analyze the effects of the above interesting coefficients
on: (1) the pressure rise in the longitude direction of the channel P(1) ; (2) streamlines
and velocity profiles. The parameters related to the fluid flow in porous medium are shown
below: permeability kp = 1 × 10−6 m2 , fluid viscosity μf = 0.001 Pa s, constant pressure
gradient d p/dx = −0.1 Pa/m. According to the Darcy’s law, the seepage velocity in the

123
Flow of Particulate-Fluid Suspension 157

Fig. 3 Variation of dimensionless pressure difference P(1) with the dimensionless flow rate F(1) at:
ε = 0.01, β = −0.1, C = 0.1, ϕ = 0, 0.1, 0.2, 0.3, 0.4

outside porous medium is 1 × 10−4 m/s. And the dimensionless pressure difference across
the channel is −1 × 104 .

6.1 Pressure Difference

First, the influence of the channel’s amplitude ratio on the variation of dimensionless pressure
difference P(1) is presented in Fig. 3. P(1) varies linearly with the dimensionless flow rate
F(1) and all the lines cross at one point, which corresponds to the flow rate with P(1) = 0. F(1)
increases with the pressure gradient. These observations are also illustrated in the forthcoming
Figs. 4, 5, 6 and will not be mentioned further.
Figure. 3 shows that the flow rate F(1) will be larger in the parallel-plate channel (ϕ = 0)
compared with that in the sine-shaped channel if P(1) and other parameters are held fixed.
This indicates that the sinusoidal channel has higher pressure drop than that of the parallel-
plate channel under fully developed flow conditions because of the larger resistance exerted
on the flow by the coarse surface. For the sine-shaped channels, the rougher the channel is,
i.e., the larger the value ϕ is, the smaller the flow rate will be. This means the increment of
the amplitude increases the effect of the curvature.
For the flow in fractured media, the rock fractures vary widely in size. Therefore, it is
important to study the effect of the parameter ε on the flow properties. One can observe from
Fig. 4 that the increase of ε will remarkably increase the flow rate F(1) under the same pressure
gradient. In this calculation, the length of the channel is kept constant and the increase of ε
implies the increase of the width, i.e., the enlargement of the flow area in the channel.
Figure 5 indicates the pressure drop along the channel for different values of the slip para-
meter β. Here β = 0 corresponds to the non-slip boundary condition and the dimensionless

123
158 J. Yao et al.

Fig. 4 Variation of dimensionless pressure difference P(1) with the dimensionless flow rate F(1) at:
ϕ = 0.1, α = 1, C = 0.1, ε = 0.0025, 0.005, 0.0075, 0.01

Fig. 5 Variation of dimensionless pressure difference P(1) with the dimensionless flow rate F(1) at:
ϕ = 0.1, ε = 0.01, C = 0.1, β = 0, −0.025, −0.1, −0.5, −1

123
Flow of Particulate-Fluid Suspension 159

Fig. 6 Variation of dimensionless pressure difference P(1) with the dimensionless flow rate F(1) at:
ϕ = 0.1, ε = 0.01, β = −0.1, C = 0, 0.05, 0.1, 0.2, 0.4

slip coefficient α introduced in the Beavers and Joseph boundary condition ranges from 0.1
to 4. Under the non-slip condition, the flow rate F(1) is smaller than that with slip condition if
the pressure gradient and all other parameters are kept constant. With the decrease of β from 0
to −1, F(1) increases largely. This indicates that the permeable boundary will notably increase
the fluid flow rate in the channel and the use of the non-slip condition will underestimate the
fluid velocity and the flow rate if the permeability of the surrounding porous medium is large
enough to reinforce the flow in the channel.
The existence of particulates in the fluid-phase will inevitably affect the fluid flow in
the channel. In Fig. 6, the variation of P(1) versus F(1) at different volume fraction C are
drawn. Here C = 0 represents the particle-free fluid flow. It is observed that the presence
of the particulates diminishes the flow rate and that an increase of C results in a decrease
of the dimensionless flow rate F(1) if the pressure gradient and all other parameters are kept
constant.

6.2 Streamline and Velocity Distribution

To discuss the behavior of the streamlines and velocity distributions in the sinusoidal channel,
numerical calculations for several values of the channel’s amplitude ratio (ϕ), the channel’s
width to length ratio (ε), the slip parameter (β), and the volume fraction density (C) are
carried out. In the following calculation, the pressure gradient in the channel is exactly the
same as that in the porous medium.

123
160 J. Yao et al.

(1)
Fig. 7 Streamlines f in sinusoidal channels with ε = 0.01, β = −0.1, C = 0.1 at different amplitude
to width ratio and Reynolds number: a ϕ = 0.1, Ref = 33.91805, b ϕ = 0.2, Ref = 31.39353, c ϕ = 0.3,
Ref = 27.49343, and d ϕ = 0.4, Ref = 22.63375

Table 1 The flow rate Q ∗f and the friction factor f for varying amplitude ratio ϕ with ε = 0.01,
β = −0.1, C = 0.1

ϕ=0 ϕ = 0.1 ϕ = 0.2 ϕ = 0.3 ϕ = 0.4

Q ∗f (m2 /s) 3.48 × 10−5 3.39 × 10−5 3.14 × 10−5 2.75 × 10−5 2.26 × 10−5
f 0.042 0.044 0.052 0.067 0.100

6.2.1 Effects of the Geometrical Parameters

The streamline patterns in sinusoidal channels with different amplitude ratio ϕ are plotted in
Fig. 7. Since the pressure drop is kept constant in all the cases, the calculated Reynolds number
does not turn out to be an integer. At these low Reynolds number, the flow moves forward
smoothly following the shape of the channel and no separation occurs. As ϕ increases, the
film thickness decreases and the streamlines are more compact at the trough. On the other
hand, the wall-induced curvature effect increases, which would exert more resistance to the
flow in the channel.
The fluid-phase Reynolds number Ref decreases with the increment of ϕ owing to the
increasing resistance of the channel and the increasing frictional energy loss. The friction
factor f is calculated using Eq. 28 to quantitatively study the resistance of the channel geom-
etry on the fluid flow. As shown in Table 1, the friction factor increases with the increment
of the amplitude ratio ϕ and the friction factors corresponding to the sinusoidal channels are

123
Flow of Particulate-Fluid Suspension 161

(1)
Fig. 8 Velocity distributions uf across sinusoidal channels at x = 0 (left) and at x = 1/2 (right) for different
values of ϕ with α = 1, C = 0.1, ε = 0.005 (top) and ε = 0.01 (bottom)

always higher than that of the parallel-plate channel. The value for the channel with ϕ = 0.4
is more than two times larger than that for the parallel-plate channel.
(1)
Figure 8 shows the velocity distributions uf at x = 0 (minimum section) and at
x = 1/2 (maximum section) for different values of ϕ with the channel’s width to length ratio

123
162 J. Yao et al.

Fig. 8 continued

ε = 0.005 and ε = 0.01. The velocity profiles are parabolic and reach its maximum
at the centerline y = 0. In this figure, the horizontal lines are added to denote the
upper boundaries for each case and the intersecting points of the horizontal lines and
the parabolic profiles signify the slip velocities at the fluid/porous interface. At x = 0,
the increase of ϕ will not influence the shape of the velocity profile, but the slip velocity

123
Flow of Particulate-Fluid Suspension 163

becomes higher for higher values of ϕ. Moreover, the maximum velocities at y = 0


in the sinusoidal channels are larger than that in the parallel-plate channel owing to
the decrease of the film thickness. On the contrary, at x = 1/2, the increase of ϕ
decreases the slip velocity and the maximum velocities at y = 0 in the sinusoidal
channels are smaller than that in the parallel-plate channel to satisfy the continuity of
mass.
Two different values of the channel’s width to length ratio ε are selected in the calculation:
ε = 0.005 (top) and ε = 0.01 (bottom). Here the channel length is kept constant; the increase
of ε means the augmentation of the channel width. The change of ε will not affect the trend
and the shape of the velocity profiles. However, the increase of ε, i.e., the augmentation of
the channel width, will enlarge the velocity values.

6.2.2 Effects of the Slip Parameter

When the surrounding walls are permeable, the slip condition instead of the classical non-
slip condition should be considered at the clear/porous interface. The influence of the slip
parameter β on the velocity distributions in the sinusoidal channels with ε = 0.01, ϕ = 0.1,
C = 0.1 at x = 0 (minimum section) and x = 1/2 (maximum section) is examined
in Fig. 9. The corresponding fluid-phase Reynolds numbers at different slip parameters
(β = 0, −0.025, −0.1, −0.2, −0.5) are 26.18372, 28.11888, 33.91805, 41.64100, and
64.78288. The decrease of β would increase the Reynolds number. One can observe from
Fig. 9 that the shape of the velocity distribution is nearly the same. But at x = 0, the
velocity magnitude is much larger than that at x = 1/2 to satisfy the continuity of
mass.
In Fig. 9, β = 0 corresponds to the non-slip condition. In this case, the surrounding walls
are treated as tight solid boundary with no effect on the fluid flow in the channel. For the
slip cases (β  = 0), the decrease of β leads to the increment of the velocity magnitude across
the channel. We can conclude that the slip boundary condition would largely increase the
velocity value in the channel compared with the non-slip condition. So the slip parameter
is an important factor used to accurately describe the flow in the channel surrounding by
permeable porous medium.

6.2.3 Effects of the Volume Fraction Density

For the particulate-fluid suspension flow, the existence of particulates in the fluid-phase
will inevitably affect the fluid flow in the channel. Figure 10 shows the effect of the
volume fraction density C on the fluid-phase velocity distributions in the channels with
ε = 0.01, ϕ = 0.1, β = −0.1. The corresponding fluid-phase Reynolds numbers at
different volume fraction density (C = 0, 0.025, 0.05, 0.1, 0.2) are 43.21841, 40.2570,
37.78778, 33.91805, and 30.14938. The increase of C would decrease the Reynolds
number.
One can observe from Fig. 10 that the velocity distribution is parabolic and the velocity
magnitude at x = 0 is much larger than that at x = 1/2 to satisfy the continuity of mass.
Besides, the slip velocity decreases with the increase of the volume fraction density. In Fig. 10,
C = 0 corresponds to the particle-free fluid flow. The velocity magnitude of the particle-free
fluid flow is larger than that of the particulate suspension flow. This is because the presence of
the particulates reduces the inertia of the fluid and this causes a slightly smaller fluid velocity.

123
164 J. Yao et al.

(1)
Fig. 9 Velocity distributions uf across sinusoidal channels with ϕ = 0.1, ε = 0.01, C = 0.1, and
β = 0, −0.025, −0.1, −0.2, −0.5 at different locations: a at x = 0, b at x = 1/2

For the particulate suspension cases (C  = 0), the fluid-phase axial velocity decreases with
increment in the particle concentration. The volume fraction density C appears to be a strong
parameter influencing the flow.

123
Flow of Particulate-Fluid Suspension 165

(1)
Fig. 10 Velocity distributions uf across sinusoidal channels with ϕ = 0.1, ε = 0.01, β = −0.1, and
C = 0, 0.025, 0.05, 0.1, 0.2 at different locations: a at x = 0, b at x = 1/2

7 Conclusions

In this paper, we have studied the problem of the dispersed particulate-fluid two-phase
flow in a channel with porous walls under the effect of the Beavers and Joseph slip

123
166 J. Yao et al.

boundary condition. The analytical solutions of longitude pressure difference, stream-


lines and its corresponding velocities at distinct situations are given using the perturbation
method. Numerical calculations for several values of the channel’s width to length ratio
ε, the channel’s amplitude ratio ϕ, the slip parameter β, and the volume fraction density
C are carried out to analyze the effects of the above interesting coefficients on the pres-
sure rise in the longitude direction of the channel P(1) and the streamlines and velocity
profiles.
At low Reynolds number, the flow moves forward smoothly following the shape of
the channel and no separation occurs. It is found that the sinusoidal channel has higher
pressure drop than that of the parallel-plate channel under fully developed flow condi-
tions. The increment of the amplitude ratio ϕ increases the effect of the curvature and
decreases the flow rate in the channel. The friction factors corresponding to the sinu-
soidal channels are always higher than that of the parallel-plate channel. The incre-
ment of the channel’s width to length ratio ε will remarkably increase the flow rate
under the same pressure gradient because of the enlargement of the flow area in the
channel.
The slip boundary condition will notably increase the fluid velocity and the decrease
of the slip parameter β leads to the increment of the velocity magnitude across the chan-
nel. The slip parameter is an important factor used to accurately describe the flow in
the channel surrounding by permeable porous medium. The velocity magnitude of the
particle-free fluid flow is larger than that of the particulate suspension flow because the
presence of the particulates reduces the inertia of the fluid. For the particulate suspension
cases, the fluid-phase axial velocity decreases with the increment in the particle concen-
tration. The volume fraction density C appears to be a strong parameter influencing the
flow.

Acknowledgments This study was supported by the National Basic Research Program of China (”973”
Program) (Grant No. 2011CB201004), the National Natural Science Foundation of China (Grant No.
51234007), and the Fundamental Research Funds for the Central Universities (Grant No. 11CX06026A).

Appendix 1 Zeroth-, and First-Order Solutions

The solution at the zeroth order is

(F0 − H ) (H 3 − 3H 2 F0 + 6β F0 H )
f0 = − y 3
− y (42)
2H 2 (H − 3β) 2H 2 (H − 3β)
(F0 − H ) (H 3 − 3H 2 F0 + 6β F0 H ) 3(F0 − H )
p0 =− y 3
− y+ y (43)
2H 2 (H − 3β) 2H 2 (H − 3β) M H 2 (H − 3β)

The solution at the first order is


f1 = a1 F02 + a2 F0 + a3 y 7 + a4 F02 + a5 F0 + a6 y 5

+ a7 F02 + a8 F0 + a9 y 3 + a10 F02 + a11 F0 + a12 y (44)

p1 = a1 F02 + a2 F0 + a3 y 7 + a13 F02 + a14 F0 + a15 y 5

+ a16 F02 + a17 F0 + a18 y 3 + a19 F02 + a20 F0 + a21 y (45)

123
Flow of Particulate-Fluid Suspension 167

The closed-form solutions in terms of the flow rate are written as follows
 
(1) F (1) − H −6β F (1) − H 2 +3H F (1)
f = − y 3
+ y +ε a1 F (1)2 +a2 F (1) +a3 y 7
2H (H −3β)
2 2H (H − 3β)
 
(1)2
+ a4 F + a5 F + a6 y 5 + a7 F (1)2 + a8 F (1) + b1 y 3
(1)

 
+ a10 F (1)2 + a11 F (1) + b2 y (46)

F (1) − H −6β F (1) M H − 6H − M H 3 + 3M H 2 F (1) + 6F (1)
p(1) = − y 3
+ y
2H 2 (H − 3β) 2M H 2 (H − 3β)
  
+ε a1 F (1)2 + a2 F (1) + a3 y 7 + a13 F (1)2 + a14 F (1) + a15 y 5
  
+ a16 F (1)2 + a17 F (1) + b3 y 3 + a19 F (1)2 + a20 F (1) + b4 y (47)

where the expressions a1 ∼ a20 and b1 ∼ b4 are given below.

3 R H  (C M H − C N H + N H − 2MCβ − 2β N + 2Cβ N )
a1 = −
280 N H 5 (H − 3β)3

1 R H (5N H − 5C N H + 5C M H − 9MCβ − 9β N + 9Cβ N )
a2 =
280 N H 4 (H − 3β)3

1 R H (2N H − 2C N H + 2C M H − 3MCβ − 3β N + 3Cβ N )
a3 = −
280 N H 3 (H − 3β)3

3 R H (N H − C N H + C M H − MCβ − β N + Cβ N )
a4 =
40 N H 4 (H − 3β)2
3 R H  (C M − C N + N )
a5 = −
40 N H 2 (H − 3β)2

1 R H (C M − C N + N )
a6 =
40 N H (H − 3β)2
3 RH
a7 = − (−210β 3 N − 210MCβ 3 − 11H 3 C N + 280β 2 H N
280 N H 2 (H − 3β)4
−99MCβ H 2 + 99Cβ H 2 N + 280MCβ 2 H − 99β H 2 N + 11H 3 MC
+11H 3 N + 210C Nβ 3 − 280C H Nβ 2 )
3 RH
a8 = (−147C Nβ 2 + 147MCβ 2 + 147Nβ 2 + 9C M H 2
280 N (H − 3β)4
−9C N H 2 + 9N H 2 − 68MC Hβ − 68N Hβ + 68C N Hβ)
1 RH H
a9 = − (−147C Nβ 2 + 147C Mβ 2 + 8C M H 2 + 61C N Hβ + 8N H 2
280 N (H − 3β)4
1 F1
−8C N H 2 + 147Nβ 2 − 61N Hβ − 61MC Hβ) −
2 H 2 (H − 3β)
3 RH
a10 = (5C M H 3 − 181C H Nβ 2 + 181MC Hβ 2 + 55Cβ N H 2 − 147Nβ 3
280 N (H − 3β)4
−55MCβ H 2 + 147N Cβ 3 − 5C N H 3 + 5N H 3 + 181H Nβ 2 − 55Nβ H 2 − 147MCβ 3 )
1 RHH2
a11 = − (11N H 2 − 102β N H − 279C Nβ 2 + 102C N Hβ + 279MCβ 2
280 N (H − 3β)4
−11C N H 2 + 11C M H 2 + 279Nβ 2 − 102MC Hβ)

123
168 J. Yao et al.

1 RHH3
a12 = (−28H Nβ − 93C Nβ 2 − 3C N H 2 + 93Nβ 2 + 3MC H 2 + 3N H 2
280 N (H − 3β)4
3 (H − 2β)F1
+28H C Nβ + 93MCβ 2 − 28MC Hβ) +
2 (H − 3β)H
3 RH
a13 = (4Cβ N H 2 − 3C H Nβ 2 + MC H 3 + N H 3 − 12β
40 N H (H − 3β)3
5

+6H − 4β N H 2 + 3H Nβ 2 − 4MCβ H 2 + 3MC Hβ 2 − C N H 3 )


3 RH
a14 =− (10H − 18β − C N H 3 + 3Cβ N H 2 + N H 3 − 3β N H 2
40 N H 4 (H − 3β)3
+MC H 3 − 3MCβ H 2 )
1 RH
a15 = (12H − C N H 3 + 3C Nβ H 2 + N H 3 − 3β N H 2 + MC H 3
40 N H (H − 3β)3
3

−3MCβ H 2 − 18β)
3 RH
a16 =− (−99β N H 4 − 11C N H 5 + 280MCβ 2 H 3 + 140H 3
280 N H 4 (H − 3β)4
−980β H 2 + 2100Hβ 2 − 210N H 2 β 3 + 99C Nβ H 4 − 210C M H 2 β 3 + 280Nβ 2 H 3
−99C Mβ H 4 + 11C M H 5 + 210C N H 2 β 3 + 11N H 5 − 1260β 3 − 280C Nβ 2 H 3 )
3 RH
a17 = (140H 2 − 840β H + 1260β 2 − 147C N H 2 β 2 + 9N H 4
280 N H (H − 3β)4
2

+9C M H 4 +147N H 2 β 2 −9C N H 4 −68Nβ H 3 −68MCβ H 3 +147MCβ 2 H 2 +68Cβ N H 3 )


1 RH
a18 =− (8N H 4 + 147N H 2 β 2 − 61Nβ H 3 − 147C N H 2 β 2
280 N H (H − 3β)4
−8C N H 4 + 140H 2 − 840β H + 1260β 2 + 147MC H 2 β 2 − 61MCβ H 3 + 61N Cβ H 3
1 F1
+8C M H 4 ) −
2 H 2 (H − 3β)
3 RH
a19 = (−12600β H + 12600Cβ H − 11760β M H 3 + 27720Mβ 2 H 2
280 N M H 5 (H − 3β)4
2

−22680M Hβ 3 + 6720M 2 β 2 H 4 − 11340M 2 β 3 H 3 + 7560M 2 β 4 H 2 − 1890M 2 β H 5


+5N M 2 H 8 − 144C M 2 H 6 − 144M N H 6 + 5C M 3 H 8 + 11760MCβ H 3 − 27720MCβ 2 H 2
+210M 2 H 6 + 1680M H 4 + 10080C M 2 β 3 H 3 − 5040M Nβ 2 H 4 + 181C M 3 β 2 H 6
−55C M 3 β H 7 − 5040C M 2 β 2 H 4 − 147C M 3 β 3 H 5 − 7560C M 2 β 4 H 2 + 22680C M Hβ 3
−147N M 2 β 3 H 5 + 10080N Mβ 3 H 3 + 1296C M 2 β H 5 + 144C M N H 6 + 1296β M N H 5
+181N M 2 β 2 H 6 − 55N M 2 β H 7 − 5C N M 2 H 8 − 7560M Nβ 4 H 2 − 15120Cβ 2
−181C N M 2 β 2 H 6 +55C N M 2 β H 7 −1296C M Nβ H 5 +147C N M 2 β 3 H 5 −10080C M Nβ 3 H 3
+5040C M Nβ 2 H 4 + 7560C M Nβ 4 H 2 − 2520C H 2 + 15120β 2 + 2520H 2 − 1680C M H 4 )
1 RH
a20 =− (−60480Hβ + 60480Cβ H − 30240Mβ H 3
280 N M H 4 (H − 3β)4
2

+60480Mβ 2 H 2 − 45360M Hβ 3 + 630M 2 β 2 H 4 − 840M 2 β H 5 + 11N M 2 H 8 − 48C M 2 H 6


−48N M H 6 + 11C M 3 H 8 + 30240C Mβ H 3 − 60480C Mβ 2 H 2 + 210M 2 H 6 + 5040M H 4
+2016M Nβ 2 H 4 + 279C M 3 β 2 H 6 − 102C M 3 β 2 H 6 + 2016C M 2 β 2 H 4 + 45360C Mβ 3 H
−384C M 2 β H 5 + 48C M N H 6 − 384M Nβ H 5 + 279N M 2 β 2 H 6 − 102N M 2 β H 7
−11C N M 2 H 8 − 68040Cβ 2 − 279C N M 2 β 2 H 6 + 102C N M 2 β H 7
+384C M Nβ H 5 − 2016C N Mβ 2 H 4 − 12600C H 2 + 68040β 2 + 12600H 2 − 5040C M H 4 )

123
Flow of Particulate-Fluid Suspension 169

1 RH
a21 = (−22680β H + 22680Cβ H − 2520Mβ H 3 − 630M 2 β 2 H 4
280 N M H 3 (H − 3β)4
2

+210M 2 β H 5 + 3N M 2 H 8 + 48C M 2 H 6 + 48M N H 6 + 3C M 3 H 8 + 2520C Mβ H 3


+840M H 4 + 1512M Nβ 2 H 4 + 93C M 3 β 2 H 6 − 28C M 3 β H 7 + 1512C M 2 β 2 H 4
−576C M 2 β H 5 − 48C M N H 6 −576β M N H 5 +93N M 2 β 2 H 6 − 28N M 2 β H 7 −3C N M 2 H 8
−22680Cβ 2 − 93C N M 2 β 2 H 6 + 28C N M 2 β H 7 + 576C M Nβ H 5 − 1512C M Nβ 2 H 4
−5040C H 2 + 22680β 2 + 5040H 2 − 840C M H 4 )
3F1 3β F1 3F1
+ − +
M H 2 (H − 3β) H (H − 3β) 2(H − 3β)
1 RH H
b1 = − (−147C Nβ 2 + 147C Mβ 2 + 8C M H 2 + 61C N Hβ + 8N H 2
280 N (H − 3β)4
−8C N H 2 + 147Nβ 2 − 61N Hβ − 61MC Hβ)
1 RHH3
b2 = (−28H Nβ − 93C Nβ 2 − 3C N H 2 + 93Nβ 2 + 3MC H 2 + 3N H 2
280 N (H − 3β)4
+28H C Nβ + 93MCβ 2 − 28MC Hβ)
1 RH
b3 = − (8N H 4 + 147N H 2 β 2 − 61Nβ H 3 − 147C N H 2 β 2
280 N H (H − 3β)4
−8C N H 4 +140H 2 − 840β H +1260β 2 +147MC H 2 β 2 −61MCβ H 3 +61N Cβ H 3 +8C M H 4 )
1 RH
b4 = (−22680β H + 22680Cβ H − 2520Mβ H 3 − 630M 2 β 2 H 4
280 N M H 3 (H − 3β)4
2

+210M 2 β H 5 + 3N M 2 H 8 + 48C M 2 H 6 + 48M N H 6 + 3C M 3 H 8 + 2520C Mβ H 3


+840M H 4 + 1512M Nβ 2 H 4 + 93C M 3 β 2 H 6 − 28C M 3 β H 7 + 1512C M 2 β 2 H 4
−576C M 2 β H 5 −48C M N H 6 − 576β M N H 5 +93N M 2 β 2 H 6 −28N M 2 β H 7 −3C N M 2 H 8
−22680Cβ 2 − 93C N M 2 β 2 H 6 + 28C N M 2 β H 7 + 576C M Nβ H 5 − 1512C M Nβ 2 H 4
−5040C H 2 + 22680β 2 + 5040H 2 − 840C M H 4 )

Appendix 2 Expressions for the Pressure Gradient

The dimensionless pressure gradient in terms of the flow rate is written as


 
∂ p0 ∂ p1 9 ε H  F (1)2
+ε =− (S1 g1 + S2 g2 + S7 )
∂x ∂x 140 N h M H 5 (H − 3β)4
2

3 F (1)
+ [ε H  (S3 g1 + S4 g2 + S8 ) + S9 ]
140 N h M H 4 (H − 3β)4
2

3 1
− [ε H  (S5 g1 + S6 g2 + S10 ) + S11 ] (48)
140 N h M 2 H 3 (H − 3β)4
where the expressions S1 ∼ S11 are given below.

S1 = −99C Rhβ M 3 H 5 − 280RC N hβ 2 M 2 H 4 − 99Rβ N h M 2 H 5 + 280Rβ 2 N h M 2 H 4


+99RCβ N h M 2 H 5 + 210RC N hβ 3 M 2 H 3 + 280RChβ 2 M 3 H 4 − 210RChβ 3 M 3 H 3
−11RC N h M 2 H 6 + 11R N h M 2 H 6 + 11RCh M 3 H 6 − 210R N h M 2 β 3 H 3
S2 = −1960RCβ Mh H 3 + 4620RC Mhβ 2 H 2 − 3780RC Mhβ 3 H + 1260R M N hβ 4 H 2
+1260RCh M 2 β 4 H 2 − 216R M N hβ H 5 − 1680R M N hβ 3 H 3 − 24RC M N h H 6
−1680RC M 2 hβ 3 H 3 + 840RCh M 2 β 2 H 4 − 216RChβ M 2 H 5 + 840R M N hβ 2 H 4

123
170 J. Yao et al.

−2100RCh Hβ + 315Rhβ M 2 H 5 + 280RC Mh H 4 − 1120Rhβ 2 M 2 H 4 + 1890Rhβ 3 M 2 H 3


−1260Rhβ 4 M 2 H 2 + 1960Rhβ M H 3 − 4620Rhβ 2 M H 2 + 3780Rhβ 3 M H + 24Rh N M H 6
+24RCh M 2 H 6 − 840RC N hβ 2 M H 4 + 216RC N hβ M H 5 + 1680RC N hβ 3 M H 3
−1260RC N hβ 4 M H 2 − 420Rh H 2 − 2520Rhβ 2 + 2520RChβ 2 − 35Rhβ 2 H 6
−280Rh M H 4 + 2100Rh Hβ + 420RCh H 2
S3 = 204RC N hβ M 2 H 5 + 441RChβ 2 M 3 H 4 + 27RCh M 3 H 6 + 441R N hβ 2 M 2 H 4
+27R N h M 2 H 6 − 204R N hβ M 2 H 5 − 441RC N hβ 2 M 2 H 4 − 204RChβ M 3 H 5
−27RCh N M 2 H 6
S4 = −11340Rhβ 2 − 2100Rh H 2 − 336R N hβ 2 M H 4 + 64R N hβ M H 5 − 8RC N h M H 6
+64RChβ M 2 H 5 − 336RChβ 2 M 2 H 4 − 5040RChβ M H 3 + 10080RChβ 2 M H 2
−7560RChβ 3 M H + 5040Rhβ M H 3 − 10080Rhβ 2 M H 2 − 10080RChβ H + 8R N h M H 6
8RCh M 2 H 6 + 7560Rhβ 3 M H + 840RCh M H 4 − 105Rhβ 2 M 2 H 4 + 140Rhβ M 2 H 5
+336RC N hβ 2 M H 4 − 64RC N hβ M H 5 + 11340RChβ 2 − 35Rh M 2 H 6 − 840Rh M H 4
+2100RCh H 2 + 10080Rhβ H
S5 = 8R N h M 2 H 6 − 8RC N h M 2 H 6 + 147R N hβ 2 M 2 H 4 − 61R N hβ M 2 H 5 + 8RCh M 3 H 6
−61RChβ M 3 H 5 − 147RC N hβ 2 M 2 H 4 + 61C R N hβ M 2 H 5 + 147RChβ 2 M 3 H 4
S6 = 420Rhβ M H 3 + 840RCh H 2 − 140Rh M H 4 + 140RCh M H 4 + 3780Rhβ H − 8RCh M 2 H 6
−3780RChβ H − 35Rhβ M 2 H 5 + 96R N hβ M H 5 + 105Rhβ 2 M 2 H 4 − 252RChβ 2 M 2 H 4
+96RChβ M 2 H 5 + 8RC N h M H 6 − 252R N hβ 2 M H 4 − 840Rh H 2
−8R N h M H 6 − 420RChβ M H 3 + 252RC N hβ 2 M H 4
−96RC N hβ M H 5 − 3780Rhβ 2 + 3780RChβ 2
S7 = −1120L N M 2 β 2 H 4 − 1260L N M 2 β 4 H 2 + 1890L N M 2 β 3 H 3 − 35L N M 2 H 6
+315L Nβ M 2 H 5
S8 = −105L Nβ 2 M 2 H 4 − 35L N M 2 H 6 + 140L Nβ M 2 H 5
S9 = 140h M N g2 H 5 − 140g1 h N M 2 H 5 − 3780g2 h M N H 2 β 3 + 1260g1 h Nβ M 2 H 4
−1260g2 h Nβ M H 4 + 3780g2 h Nβ 2 M H 3 − 3780g1 h Nβ 2 M 2 H 3
+3780g1 h Nβ 3 M 2 H 2
S10 = −35L Nβ M 2 H 5 + 105L Nβ 2 M 2 H 4
S11 = −140g1 h N M 2 H 5 + 140g2 h N M H 5 + 1260g1 h Nβ M 2 H 4 − 3780g1 h Nβ 2 M 2 H 3
+3780g1 h Nβ 3 M 2 H 2 − 1260g2 h Nβ M H 4 + 3780g2 h Nβ 2 M H 3
−3780g2 h Nβ 3 M H 2

References

Arbogast, T., Lehr, L.H.: Homogenization of a Darcy-Stokes system modeling vuggy porous media. Comput.
Geosci. 10, 291–302 (2006)
Ali, N., Hussain, Q., Hayat, T., Asghar, S.: Slip effects on the peristaltic transport of MHD fluid with variable
viscosity. Phys. Lett. A 372, 1477–1489 (2008)
Aguilar-Madera, C.G., Valdés-Parada, F.J., Goyeau, B., Ochoa-Tapia, J.A.: One-domain approach for heat
transfer between a porous medium and a fluid. Int. J. Heat Mass Transf. 54, 2089–2099 (2011)
Beavers, G., Joseph, D.D.: Boundary conditions at a naturally permeable wall. J. Fluid Mech. 30, 197–207
(1967)
Brown, S.R., Stockman, H.W., Reeves, S.J.: Applicability of the Reynolds equation for modeling fluid flow
between rough surfaces. Geophys. Res. Lett. 22, 2537–2540 (1995)

123
Flow of Particulate-Fluid Suspension 171

Baber, K.: Modeling the transfer of therapeutic agents from the vascular space to the tissue compartment
(a continuum approach). Universitat Stuttgart, Master thesis (2009)
Charm, S.E., Kurland, G.S.: Blood Flow and Microcirculation. Wiley, New York (1974)
Chandesris, M., Jamet, D.: Boundary conditions at a planar fluid-porous interface for a Poiseuille flow. Int. J.
Heat Mass Transf. 49, 2137–2150 (2006)
Drew, D.: Mathematical modeling of two-phase flow. Annu. Rev. Fluid Mech. 15, 261–291 (1983)
Furman, A.: Modeling coupled surface-subsurface flow processes: a review. Vadose Zone J. 7, 741–756 (2008)
Goyeau, B., Lhuillier, D., Gobin, D.: Momentum transport at a fluid-porous interface. Int. J. Heat Mass Transf.
46, 4071–4081 (2003)
Hayat, T., Hussain, Q., Ali, N.: Influence of partial slip on the peristaltic flow in a porous medium. Phys. Lett.
A 387, 3399–3409 (2008)
Huang, C., Shy, S., Chien, C., Lee, C.: Parametric study of anodic microstructures to cell performance of
planar solid oxide fuel cell using measured porous transport properties. J. Power Sources 195, 2260–2265
(2009)
Huang, Z., Yao, J., Li, Y.: Permeability analysis of fractured vuggy porous media based on homogenization
theory. Sci. China Tech. Sci. 53, 839–847 (2010)
Ishii, M.: Thermo-Fluid Dynamic Theory of Two-phase flow. Volume 22 of Direction des etudes et recherches
d’électricité de France. Eyrolles, Paris (1975)
Iliev, O., Laptev, V.: On numerical simulation of flow through oil filters. Comput. Vis. Sci. 6, 139–146 (2004)
Jamet, D., Chandesris, M.: Boundary conditions at a fluid-porous interface: an a priori estimation of the stress
jump coefficients. Int. J. Heat Mass Transf. 50, 3422–3436 (2007)
Jamet, D., Chandesris, M., Goyeau, B.: On the equivalence of the discontinuous one- and two-domain
approaches for modeling of transport phenomena at a fluid-porous interface. Transp. Porous Media 78,
403–418 (2009)
Jiménez-Lozano, J., Sen, M., Corona, E.: Analysis of peristaltic two-phase flow with application to ureteral
biomechanics. Acta Mech. 219, 91–109 (2011)
Kleinstreuer, C.: Two-Phase Flow: Theory and Applications. Taylor and Francis, London (2003)
Larson, R.E., Higdon, J.J.L.: Microscopic flow near the surface of two-dimensional porous-media: 1. Axial-
flow. J. Fluid Mech. 166, 449–472 (1986)
Larson, R.E., Higdon, J.J.L.: Microscopic flow near the surface of two-dimensional porous-media: 2. Trans-
verse flow. J. Fluid Mech. 178, 119–136 (1987)
Le Bars, M., Worster, M.G.: Interfacial conditions between a pure fluid and a porous medium: implications
for binary alloy solidification. J. Fluid Mech. 550, 149–173 (2006)
Mekheimer, K.S., El Shehawey, E.F., Elaw, A.M.: Peristaltic motion of a particle-fluid suspension in a planar
channel. Int. J. Theor. Phys. 37, 2895–2920 (1998)
Mariamma, N.K., Majhi, S.N.: Flow of a Newtonian fluid in blood vessel with permeable wall–a theoretical
model. Comput. Math. Appl. 40, 1419–1432 (2000)
Mishra, M., Rao, A.R.: Peristaltic transport in a channel with a porous peripheral layer: model of a flow in
gastrointestinal tract. J Biomech. 38, 779–789 (2005)
Morales-Zárate, E., Valdés-Parada, F.J., Goyeau, B., Ochoa-Tapia, J.A.: Diffusion and reaction in three-phase
systems:Average transport equations and jump boundary conditions. Chem. Eng. J. 138, 307–332 (2008)
Neale, G., Nader, W.: Practical significance of Brinkman’s extension of Darcy’s law : coupled parallel flows
within a channel and a bounding porous medium. Can. J. Chem. Eng. 52, 415–478 (1974)
Nield, D.A.: The limitations of the Brinkman-Forchheimer equation in modeling flow in a saturated porous
medium and at an interface. Int. J. Heat Fluid Flow 12, 269–272 (1991)
Niceno, B., Nobile, E.: Numerical analysis of fluid flow and heat transfer in periodic wavy channels. Int.
J. Heat Fluid Flow 22, 156–167 (2001)
Nield, D.A.: The Beavers-Joseph boundary condition and related matters: a historical and critical note. Transp.
Porous Media 78, 537–540 (2009)
Ochoa-Tapia, J.A., Whitaker, S.: Momentum-transfer at the boundary between a porous medium and a homo-
geneous fluid: 1. Theoretical development. Int. J. Heat Mass Transf. 38, 2635–2646 (1995a)
Ochoa-Tapia, J.A., Whitaker, S.: Momentum transfer at the boundary between a porous medium and a homo-
geneous fluid. II. Comparison with experiment. Int. J. Heat Mass Transf. 38, 2635–2656 (1995b)
RaviKumar, Y.V.K., KrishnaKumari, S.V.H.N., Raman Murthy, M.V., Sreenadh, S.: Unsteady peristaltic
pumping in a finite length tube with permeable wall. Trans. ASME J. Fluids Eng. 32, 1012011–1012014
(2010)
Sudicky, E.A., Frind, E.O.: Contaminant transport in fractured porous media: analytical solutions for a system
of parallel fractures. Water Resources Res. 18, 1634–1642 (1982)
Sahraoui, M., Kaviany, M.: Slip and no-slip velocity boundary conditions at interface of porous, plain media.
Int. J. Heat Mass Transf. 35, 927–943 (1992)

123
172 J. Yao et al.

Srivastava, V.P., Srivastava, L.M.: Influence of wall elasticity and Poiseuille flow on peristaltic induced flow
of a particle-fluid mixture. Int. J. Eng. Sci. 35, 1359–1386 (1997)
Shavit, U.: Special issue on “transport phenomena at the interface between fluid and porous domains” a preface.
Transp. Porous Media 78, 327–330 (2009)
Tam, C.K.W.: The drag on a cloud of spherical particles in low Reynolds number flow. J. Fluid Mech. 38,
537–546 (1969)
Vafai, K., Kim, S.: Fluid mechanics of the interface region between a porous medium and a fluid layer-an
exact solution. Int. J. Heat Fluid Flow 11, 254–256 (1990)
Valdés-Parada, F.J., Goyeau, B., Ochoa-Tapia, J.A.: Diffusive mass transfer between a microporous medium
and an homogeneous fluid: Jump boundary conditions. Chem. Eng. Sci. 61, 1692–1704 (2006)
Valdés-Parada, F.J., Goyeau, B., Ochoa-Tapia, J.A.: Jump momentum boundary condition at a fluid-porous
dividing surface: derivation of the closure problem. Chem. Eng. Sci. 62, 4025–4039 (2007)
Valdés-Parada, F.J., Alvarez-Ramirez, J., Goyeau, B., Ochoa-Tapia, J.A.: Computation of jump coefficients
for momentum transfer between a porous medium and a fluid using a closed generalized transfer equation.
Transp. Porous Media 78, 439–457 (2009a)
Valdés-Parada, F.J., Alvarez-Ramirez, J., Goyeau, B., Ochoa-Tapia, J.A.: Jump condition for diffusive and
convective mass transfer between a porous medium and a fluid involving adsorption and chemical reaction.
Transp. Porous Media 78, 459–476 (2009b)
Whitaker, S.: The Method of Volume Averaging. Kluwer, Dordrecht (1999)
Wood, B.D., Quintard, M., Whitaker, S.: Jump condition at non-uniform boundaries: The catalytic surface.
Chem. Eng. Sci. 55, 5231–5245 (2000)
Weerakone, W.M.S.B., Wong, R.C.K., Mehrotra, A.K.: Single-phase (brine) and two-phase (DNAPL-brine)
flows in induced fractures. Transp. Porous Media 89, 75–95 (2011)

123
View publication stats

You might also like