You are on page 1of 14

Chemical Engineering Science 60 (2005) 25572570

www.elsevier.com/locate/ces
Application of a three-layer modeling approach for solids transport in
horizontal and inclined channels
A. Ramadan
a,
, P. Skalle
b
, A. Saasen
c
a
Department of Petroleum Engineering, University of Tulsa, NCDB, 2450 E. Marshall St., Tulsa, OK 74110, USA
b
Department of Petroleum Engineering and Applied Geophysics, NTNU, Trondheim, Norway
c
Statoil , N-4035 Stavanger, Norway
Received 10 May 2003; received in revised form 27 August 2004; accepted 5 December 2004
Abstract
The three-layer model concept developed previously for solidliquid ow has been adapted to model solids transport in inclined channels.
The present model predicts the pressure loss and transport rate of solids in Newtonian and power-law uid suspensions by assuming
stratied ow conditions. Sets of stationary sand bed transport rate tests were performed to verify the predictions of the model. A 70-mm
ow loop was constructed to measure the average transport rates and critical ow rates, which are required to initiate the motion of solids
bed particles. The tests were carried out by eroding stationary sand beds with water and an aqueous solution of poly anionic cellulose
(PAC) in a transparent pipe. Four sand beds with different particle size ranges were used. The average transport rates of the beds were
predicted using the model. The model predictions show a satisfactory agreement with experimentally measured results when the grain
Reynolds number is between 15 and 400 and the ow rate is sufciently higher than the critical ow rate. Therefore, with some degree
of limitation, the three-layer model can be applicable for predicting the transport rates of stationary solids beds in inclined channels for
both Newtonian and power-law uids.
2005 Elsevier Ltd. All rights reserved.
Keywords: Modeling; Non-Newtonian uid; Moving bed; Particle; Slurries; Suspension; Solid transport
1. Introduction
Suspension ows are known to display substantially
non-uniform concentration distributions, which inuence
hydraulic characteristics of the ow. Many engineering
problems involve suspension ows in ducts of complex
geometry. During drilling of inclined wells, drilling uid
circulates through the annular space between the drill pipe
and the borehole. The recirculation is mainly to transport
rock cutting from the well to the surface. Often accurate
predictions of cutting transport rate and frictional pressure
loss are necessary in the design of the drilling process. Esti-
mating the cutting transport rate requires accurate hydraulic
predictions and knowledge of cutting concentration prole

Corresponding author. Tel.: +1 9186315174; fax: +1 9186315009.


E-mail address: ramadan-ahmed@utulsa.edu (A. Ramadan).
0009-2509/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2004.12.011
in the annulus. Nonetheless, obtaining a simple solution for
the concentration prole is difcult. As a result, different
modeling approaches are proposed to predict the concentra-
tion prole, transport rate and frictional pressure loss.
Since the introduction of directional wells, considerable
efforts have been made to solve the cutting transport prob-
lem in highly inclined and horizontal wells. For good cutting
transport, vertical ow is preferable because cutting fall in
the opposite direction to the velocity of the drilling uid
(mud). For an inclined well, the direction of cutting settling
is still vertical, but the uid velocity has a reduced vertical
component. This decreases the uids capability to suspend
drilled cuttings. At a high angle of inclination (when the
inclination is measured from vertical), a cutting particle that
slips through the mud has a short distance to travel before
striking the borehole wall. Once it has reached the wall,
the particle has little chance of entrainment because local
2558 A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570
Flow
direction
Pipe m
ovem
ent
Fig. 1. Sticking of drill pipe during tripping operation.
(a) (b) (c)
Fig. 2. Typical borehole geometric congurations in inclined wells.
uid velocities near the wall are very low and insufcient
to re-entrain the particle into the ow. The most common
drilling problems related to cutting transport are either insuf-
cient hole cleaning or excessive annular velocity that leads
to borehole erosion and fracturing of the formation. Usu-
ally insufcient hole cleaning results in a buildup of cutting
concentration in the annulus, consequently sticking of the
pipe, lost circulation and hindering the running of a casing.
Fig. 1 shows how cutting beds cause problems when the
drillstring is moved axially during a tripping operation (Rasi,
1994). Unlike vertical wells, the ows in inclined wells are
not annular due to the formation of cutting bed on the low
side of the borehole. As a result, borehole congurations in
inclined wells vary with cutting bed height, eccentricity of
the drill pipe and diameters of the borehole and drill pipe.
The most common types of borehole geometric congura-
tions in inclined wells are presented in Fig. 2.
During drilling of both vertical and inclined wells, non-
Newtonian uids such as clay (Bentonite), oil-based muds
and polymer-viscosied uids are often used as a drilling
uid. An aqueous solution of poly anionic cellulose (PAC)
is one of the commonly used polymer-viscosied drilling
uids. Experimental and modeling studies suggested that,
in addition to the ow rate, the rheology determines cutting
transport capacity of drilling uids. Non-Newtonian uid
properties such as shear-thinning and thixotropic behaviors
(yield stress) are normally considered in selecting a drilling
uid that has higher cutting transport ability with optimum
frictional pressure loss. As a result, in practical eld appli-
cations, non-Newtonian uid properties such as yield stress,
consistency index and Power-law exponent are routinely
controlled on the surface along with the ow rate (pump
rate). Commercially available hydraulic and cutting trans-
port models are used to predict the effects of drilling uid
properties and other drilling parameters. However, model
predictions need improvement to minimize hydraulics and
cutting transport-related problems. Often such problems can
lead to expensive and costly operations. Therefore, a cut-
ting transport model that is applicable for non-Newtonian
uid is essential to obtain reasonable hydraulic and cutting
transport predictions.
In order to model stratied solidliquid ow in pipes,
Wilson (1976) developed a one-dimensional two-layer
model that assumes ow of two layers (phases). The
solidliquid mixture in each layer is considered as a single
phase. Each layer has a separate velocity, with momentum
transfer between the layers due to shear forces at their in-
terface. This approach has been extended by a number of
workers including Gavignet and Sobey (1989) who pre-
sented the two-layer cutting transport modeling concept.
The two-layer modeling approach assumes that the lower
part of the inclined wellbore forms a bed that slides up
the annulus together with a layer of suspension. In this
approach, two mechanisms can be expected to affect the
cutting bed. One of the mechanisms is saltation that occurs
when the drag force is exerted by the liquid phase on a
particle resting at the interface. The drag force initiates the
movement of the bed particle. The other one is sliding that
takes place when the shear stress exerted by the liquid over
the rough surface and the pressure gradient result in sliding
of the entire bed by overcoming the solidsolid friction at
the wall. However, the two-layer modeling approach ignores
the most important phenomenon that occurs near the bed.
Bagnold (1954) experimentally found the existence of dis-
persive shear stress. Dispersive shear stress originates from
the interchange of momentum, which is caused by inter-
particle collision and hydrodynamic interference (FredsZe
and Deigaard, 1992; Davis, 1996). In addition to the uid
shear stress, the dispersive shear stress is responsible for the
entrainment of stationary bed particles. The two-layer mod-
eling approach neglects the presence of the dispersed layer.
Therefore, this assumption reduces the accuracy and applica-
bility of the model, because the physical phenomena occur-
ring near the bed is different from the other part of the ow.
Experimental observation indicates the existence of a highly
concentrated thin layer of suspension near the bed. The con-
cept of dispersive layer appears to have a phenomenological
background. As a result, the suspension above a solids bed
is considered to have two different uid layers with distinct
concentration proles.
The concept of dispersive layer has been employed by
Doron and Barnea (1993) to extend the two-layer modeling
approach to a three-layer scheme. Their model considered
A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570 2559
c
y
0.5
s
db
i
Fdb Fd d
Fb b

b

d
Ss
Si
Sdb
Sd
Sb
Solids Bed
Suspended Layer
Dispersed Layer
(a) (b)
Fig. 3. (a) Schematic representation of shear stresses acting in the
three-layer mechanics model; and (b) assumed concentration proles in
three-layer modeling scheme.
the existence of a dispersive layer, which is sandwiched be-
tween the suspended layer and the bed as shown in Fig. 3a.
The dispersed layer was considered to have a higher concen-
tration gradient compared to the suspended layer (Fig. 3b).
No-slip condition between the solid particles and the uid
was assumed. This assumption is applicable when the ow
is in horizontal or near-horizontal congurations. The model
predictions showed satisfactory agreement with experimen-
tal data, which were obtained in horizontal pipes. Nguyen
and Rahman (1998) have adopted the three-layer model for
cutting transport predictions. They have studied the effect
of various geometric and ow parameters on the hydraulics
and transport rate of cuttings. Recently, Cho et al. (2002)
have proposed a new approach to overcome the limitations
of the existing three-layer model and adopted the model for
inclined channel ows. Nonetheless, the model neglects the
settling actions of the particles, which may signicantly af-
fect the model predictions.
Although some efforts have been expended on develop-
ing the three-layer cutting transport model, its application
is still limited to Newtonian uid ows in horizontal and
nearly horizontal channels. In the present study, the three-
layer model is extended for use in non-Newtonian inclined
channel ows. The no-slip assumption between the particles
and the uid, which was the shortcoming of the previous
modeling studies, has been eliminated. The no-slip assump-
tion is often used to simplify the model; however, it restricts
the model to horizontal and near-horizontal congurations.
2. Model development
The three-layer model presented in the study is developed
to overcome the limitations in the existing models, which
are used to predict cutting transport in inclined and horizon-
tal wells. Material balance equations of the two phases and
momentum equations of the three layers are combined to
develop the model. Additional equations are introduced to
estimate the average concentration of the suspended layer,
and thickness and velocity of the dispersed layer. The thick-
ness of the dispersed layer is modeled using the pseudo-
hydrostatic pressure gradient concept and assuming linearly
varying particle concentration in the dispersed layer. As pro-
posed by Wilson (1987), the velocity prole of the dispersed
layer is estimated based on the turbulent mixing length the-
ory. The convectiondiffusion equation is used to determine
the average concentration of the suspended layer. In order
to compare the model predictions with the measured data
(measured average transport rate), transport rate predictions
were performed at different bed heights and average trans-
port rates were determined.
2.1. Model assumptions
Like other modeling procedures, the three-layer model has
its own idealizations and assumptions that are necessary to
represent the physical model in a series of simple mathemat-
ical relationships between the ow parameters and proper-
ties of the uid and solid particles. Therefore, the following
assumptions are made to develop the present model:
(i) Distinct imaginary interface lines are to exist between
the dispersed layer and suspended layer, and between
the dispersed layer and the bed.
(ii) Uniform layers are present without signicant variation
in concentration and thickness along the length of the
channel.
(iii) The relative velocity between the particles and the uid
is negligible in the bed.
(iv) The ow is steady and turbulent.
(v) Bed shear stress variation in the lateral direction is neg-
ligible.
(vi) Stratied and well-compacted bed to resist the applied
shear.
Nguyen and Rahman (1998) mentioned that there are var-
ious transport modes depending on a given set of operating
conditions. They considered ve modes of transport in their
model. Based on visual observation of a ow loop test, Ford
et al. (1990) categorized the transport process into different
transport modes, which are presented in Fig. 4. Accordingly,
Figs. 4a and b show sub-critical conditions where the ow
velocity is not strong enough to lift the top-most particles of
the bed layer into the dispersed layer. Furthermore, at a low
angle of inclination (:) the gravitation force, which is acting
downhill, becomes high enough to overcome the pressure
gradient and the shear stress on the bed. Consequently, the
bed begins to slide down as shown in Fig. 4a. Otherwise,
the bed becomes stationary.
At higher ow rates (mean uid velocities), the intensity
of the local velocity at the bed becomes strong enough to en-
train the particles. As a result, the dispersed and suspended
layers will form with three different bed layer sliding pos-
sibilities as shown in Figs. 4ce. Accordingly, the bed layer
may slide down or slide up or be stationary during the trans-
port. If the bed layer does not slide at higher ow veloci-
ties, then it gets thinner and eventually disappears, leaving
2560 A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570
Fig. 4. Possible modes of solids transport in inclined channel.
behind the dispersed and suspended layers as shown in Fig.
4f. Finally, further increase in the ow velocity suspends the
bed particles and forms a fully suspended ow as shown in
Fig. 4g.
2.2. Material balance
With the above assumptions, the material balance equa-
tions of solid and liquid phases for stratied solidliquid
ows in inclined channels can be written as
(u
s
v
s
cos :)c
s
A
s
+(u
d
v
d
cos :)c
d
A
d
+u
b
c
b
A
b
=(u
ave
v
ave
cos :)c
ave
A
t
=T
r
(1)
and
u
s
(1 c
s
)A
s
+u
d
(1 c
d
)A
d
+u
b
(1 c
b
)A
b
=u
ave
(1 c
ave
)A
t
, (2)
respectively, where u is the mean uid ow velocity in the
layers, v the particle settling velocity in the layers, A the
cross-sectional area of the layers, c the average volumetric
concentration of solids particles in the layers, and T
r
the
transport rate of the solids. The subscripts s, d and b
refer to the suspended layer, dispersed layer and bed layer,
respectively.
2.3. Momentum balance
For the stationary bed condition, the average settling ve-
locity of the particles in the stream, v
ave
, can be estimated
by applying settling momentum balance for solids particles.
Thus, the total settling momentum equals the sum of settling
momentum of particles in each layer. Since the settling ve-
locity in the bed is negligible, the average settling velocity
equation after simplication becomes
v
ave
=
v
s
c
s
A
s
+v
d
c
d
A
d
c
s
A
s
+c
d
A
d
+c
b
A
b
. (3)
Here, it is worthwhile to give the physical meaning of
terms in Eqs. (1) and (2). The terms in the right-hand side of
Eq. (1) represent the solid transport rates in each layer, while
the terms in the right-hand side of Eq. (2) represent the uid
ow rates in each layer. These material balance equations
cannot yield the solution by themselves; therefore, we need
to include the force balance equations as well. Under steady-
state condition, the sum of all forces acting on a single
layer must be zero. Thus, the momentum equation for the
suspended layer becomes
A
s
dp
dx
t
s
S
s
t
i
S
i
j
s
A
s
g cos : =0, (4)
where : is the angle of inclination from vertical as shown
in Fig. 4a, dp/dx the pressure gradient, t
i
is the interfacial
shear stress between the suspended and the dispersed layers,
S
i
is the perimeter in contact between these layers as shown
in Fig. 3a and t
s
is the interstitial shear stress acting on the
perimeter S
s
.
The shear stress at the interface can be estimated by adopt-
ing the most widely used DarcyWeibach equation. Accord-
ingly, the interfacial shear stress is given by
t
i
=
1
8
j
s
f
i
(U
s
U
d
)
2
, (5)
where U
s
and U
d
are mean ow velocities of the suspended
and dispersed layers, respectively. The friction coefcient at
the interface, f
i
, can be obtained using the friction factor
correlations. However, it is necessary to multiply the friction
coefcient by a factor of 2, to account for the dispersive
stress at the interface (Doron and Barnea, 1993). The friction
A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570 2561
factor correlation, which is applicable for power-law uids,
is given by Szilas et al. (1981) as
1

f
=2 log
_
10
[/2
Re f
(2n)/2n
+
d
p
3.71D
h
_
, (6)
where [ is given by
[ =1.51
1/n
_
0.707
n
+2.12
_

4.015
n
1.057, (7)
where d
p
denotes the mean diameter of the particles. The
hydraulic diameter of the layer is D
h
and n is the ow be-
havior index. The density of the suspended layer, j
s
, is esti-
mated as j
s
=c
s
j
p
+(1 c
s
)j
L
, where j
p
and j
L
are the
density of solids particles and the uid, respectively. Similar
suspension density formulas can be used for the other lay-
ers. Eq. (6) is applicable for both smooth and rough walls.
The generalized Reynolds number for power-law uids is
calculated as
Re =
8j
s
U
2n
s
D
n
h
K
s
_
6n +2
n
_
n
, (8)
where K
s
is the suspended layer consistency index, which
is roughly estimated by K
s
=K(1+2.5c
s
) (Einstein, 1906).
Likewise, the interstitial uid shear stress, t
s
, in Eq. (4) can
be estimated as t
s
=f
s
j
s
u
2
s
/8, where f
s
is the friction factor
between the uid and the channel, which can be determined
numerically from Eq. (6).
Furthermore, the momentum balance of the dispersed
layer can be written as
A
d
dp
dx
+t
i
S
i
(t
d
S
d
+F
d
)
(t
db
S
db
+F
db
) j
d
A
d
g cos : =0, (9)
where t
d
is the uid shear stress between the suspension
layer and the channel acting on the perimeter, S
d
. The uid
shear stress between the dispersion layer and the bed, t
db
,
acts on the perimeter, S
db
. The dispersive force, F
db
, in-
cludes the inter-particle collision and hydrodynamic inter-
ference forces that act on the interface, S
db
. For the sake of
simplicity, dispersive force, F
db
, and other forces in the mo-
mentum equations are considered in unit length basis. Based
on the above denition, an equivalent dispersive shear stress
can be estimated as t
dis
=F
db
/S
db
.
The dry friction force, F
d
, is acting on the boundaries
of S
d
. This force is estimated using a pseudohydrostatic
pressure distribution. Thus, the approximate equation for the
dry friction force is given by
F
d
gj
d
sin :(j
p
j
L
)c
d
S
d
h
d
cos
_
0
b
+0
d
2
_
, (10)
where j
d
is the dry dynamic friction coefcient between the
particles and the wall of the channel and g is the gravitational
acceleration. The thickness of the dispersed layer is h
d
. The
angular bed thicknesses (Fig. 3a) of the bed and dispersed
layer are 0
b
and 0
d
, respectively.
The uid shear stress at the interface between the bed and
the dispersed layer is estimated by
t
db
=
1
8
j
d
f
db
(U
d
U
b
)
2
, (11)
where f
db
is the interfacial uid friction factor that can be es-
timated using Eq. (6). However, in this case, we do not need
to multiply the factor by two, as suggested by Televantos
et al. (1979), because the dispersive force, F
db
, takes care
of the dispersive shear stress. The interstitial shear stress,
t
d
, can be evaluated in a similar way to the interstitial shear
stress, t
s
. However, the hydraulic diameter of the dispersed
layer can be approximated by the hydraulic diameter of the
suspended layer. Direct calculation of the hydraulic diameter
from the ow area and the perimeter of the dispersed layer
does not give a reasonable friction factor, because this layer
is at least in contact with one moving layer, sharing almost
50% of its perimeter. As a result, the use of directly calcu-
lated hydraulic diameter overestimates the pressure loss and
the interfacial shear stress.
Similarly, the momentum equations of the bed can be
written as
A
b
dp
dx
+(t
db
S
db
+F
db
) (t
b
s
b
+F
b
)
j
b
A
b
g cos : =0, (12)
where t
b
is the interstitial shear stress between the bed layer
and the channel, and S
b
is the perimeter in contact with the
channel. The dry friction force, F
b
, acts on the boundary
S
b
. The interstitial uid shear stress between the bed layer
and the channel, t
b
, can be evaluated in a similar way to the
interstitial uid shear, t
s
.
Theoretically upward sliding of the bed occurs from the
stationary condition when F
b
>F
st
, where F
st
is the static
friction force acting on the surface of the channel. Similarly,
if F
b
<F
st
, then the bed will slide downward as shown in
Fig. 4c. During upward sliding of the bed, the magnitude of
the dry friction force becomes that of the dynamic friction
force, F
dyn
. Similarly for downward sliding F
b
= F
dyn
.
The static friction force and the dynamic friction force can
be evaluated by F
st
=F
n
j
s
and F
dyn
=F
n
j
d
, respectively,
where F
n
is the normal force and j
s
the static friction factor
between the bed and the channel. The normal force can
be estimated by integrating the pseudohydrostatic pressure
distribution over the boundary of the bed in contact with the
channel. Thus, the local normal force acting on a differential
element on the surface of the channel can be calculated as
dF
n
=g(j
p
j
L
)c
b
_
D
2
(cos 0 1) +h
_

_
D
2
d0 cos 0
_
sin :, (13)
where h is the bed height and 0 the angular displacement
(Fig. 5). The total normal force can be obtained by integrat-
ing Eq. (13) from 0
b
to 0
b
. Subsequently, the simplied
form of the total normal force acting on the surface of the
2562 A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570

b
d
h-(r-rcos())
r-rcos()
h

Fig. 5. Geometric considerations in the determination of the pseudohy-


drostatic pressure distribution in the bed layer.
channel becomes
F
n
gDc
bu
(j
p
j
L
)(0.5D(0.50
bu
+0.25 sin 20
bu
0.5 sin 0
bu
) +h sin 0
bu
) sin :. (14)
2.4. Thickness and velocity of dispersed layer
In order to complete the three-layer modeling, additional
model equations are necessary for determining the following
unknowns: (i) average concentration of the suspended layer;
(ii) thickness of dispersed layer; and (iii) velocity of the
dispersed layer. Often the thickness of the dispersed layer
is determined using the pseudohydrostatic pressure gradient
concept and assuming linearly varying particle concentration
distribution in the dispersed layer (FredsZe and Deigaard,
1992). Thus, the average concentration of dispersed layer is
approximately given by
c
d

c
s
+c
b
2
. (15a)
In drilling operations, average suspended layer concentra-
tion, c
s
, is very small compared to the average concentra-
tion of dispersed layer. Therefore, for the iterative numerical
procedure, the value of c
d
can be initialized as 0.5c
b.
After adopting for the three-layer model approach, the
formula for dispersed layer thickness can be estimated as
h
d
=
t
dis
g sin : c
b
(j
P
j
L
) tan [
D
, (15b)
where tan([
D
) is an equivalent dynamic friction factor.
Bagnold (1954) experimentally showed that the value of
tan([
D
) is around 0.75 near the bed. In naturally formed
beds of solids the concentration of solids is usually taken to
be 0.5. Thus, Eq. (15b) can be rewritten in a simple form as
h
d
=
2.67t
dis
g sin :(j
P
j
L
)
. (16)
Based on the turbulent mixing length theory, Wilson
(1987) proposed the following relation to determine the
velocity prole of the dispersed layer relative to the bed
(FredsZe and Deigaard, 1992):
U
d,r
(y) =
U
t

2
_
2y
h
d
. (17)
The von Karman constant, , is 0.4 and the distance from
the mean bed level is y. The friction velocity, U
t
, is given by
(t
i
/j
s
)
0.5
. An expression for the mean relative velocity can
be obtained by averaging the value U
d,r
over the thickness
h
d
. Thus, the nal expression for absolute mean velocity of
the dispersed layer is
U
d
=
4

2
3
_
t
i
j
P
+U
b
. (18)
2.5. Average concentration of suspended layer
At this stage, the average concentration of the suspended
layer is unknown. Wilson and Pugh (1988) suggested that
the concentration in this layer is negligible. However,
Doron (1994) has used a convectiondiffusion model ap-
proach to determine the local concentration (c
sl
) prole in
this layer. Consequently, a one-dimensional time depen-
dent, convectiondiffusion equation for an inclined channel
becomes
jc
sl
jt
=
j
jy
_
I
jc
sl
jy
+v
s
c
sl
sin :
_
, (19)
where I is the solid particles diffusion coefcient. The rst
term in the bracket is the diffusive term that represents the
upward ux of solid particles by the action of turbulent ed-
dies or other hydrodynamic interactions and the second one
is the connective term, which represents the sedimentation
ux of the particles due to the gravitational force. In order
to fulll the steady-state condition, the sum of terms in the
bracket must be zero, yielding the steady-state convection
diffusion equation as
I
jc
sl
jy
=v
S
c
sl
sin :. (20)
For a rectangular duct, assuming constant settling velocity
and diffusivity coefcient, the average concentration of the
suspended layer can be obtained from Eq. (20). Thus, the av-
erage concentration of the suspended layer is approximately
estimated by
c
s
c
d
I
1 e
(v
s
/I) sin : h
s
v
s
sin : h
s
, (21)
where h
s
is the thickness of the suspended layer. Obviously,
Eq. (21) is not an exact equation of the average concentra-
tion, but it gives a reliable approximate value. At this point,
it is important to note that similar analysis may not be appli-
cable for the dispersed layer, because in the dispersed layer,
A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570 2563
the normal component of the settling velocity, v
d
, is bal-
anced not only by the hydrodynamic diffusion but also by
the particle collision. As a result, the use of Eq. (20) for the
dispersed layer can be inappropriate. Based on experimental
investigation, FredsZe and Deigaard (1992) have suggested
the use of linearly varying particle concentration prole for
the dispersed layer.
Theoretical and experimental studies on particle disper-
sion in turbulent ows indicate that the diffusivity of solid
particles is roughly equal to or greater than the diffusiv-
ity of the uid. This assumption is supported by theoretical
results of homogeneous isotropic turbulence (Mohs et al.,
2000). Moreover, the Reynolds analogy between mass and
momentum transport suggests that uid diffusivity is the
same as eddy viscosity (Young and Leeming, 1997). There-
fore, the solid particle turbulent diffusivity coefcient in the
suspended layer can be estimated as I =U
t
(1 y/h
s
)y.
This means that the diffusivity coefcient has a parabolic
prole with zero value at the bottom of the suspended layer.
In order to avoid zero diffusivity value, the coefcient is usu-
ally evaluated a small distance from the interface between
the suspended and dispersed layers. FredsZe and Deigaard
(1992) have recommended this distance to be twice the mean
particle diameter of a solids bed. In the present study, this
distance is assumed to be h
d
/100. The laboratory experi-
ments indicated that the thickness of the dispersed layer and
the diffusivity of the suspended layer increase with the ow
velocity.
Determination of the settling velocity of the particles is
necessary in Eqs. (1), (3) and (21). Settling has a negative
impact on solid transport and it depends on the properties of
the uid and suspended particles. Generally, a solid particle
falling in a uid under the action of gravity accelerates until
the buoyancy and drag force just balance the gravitational
force, and then it continues to fall at constant velocity, which
is given by (Dedegil, 1987)
v =
_
4gd
p
(j
p
j
f
)
3j
L
C
D
, (22)
where C
D
is the particle drag coefcient that is determined
by (White, 1974):
C
D
=
24
Re
p
+
6
1 +Re
0.5
p
+0.4, (23)
where Re
p
is the particle Reynolds number. The above equa-
tion can be valid for Newtonian and non-Newtonian uids
if the denition of the particle Reynolds number is the same
in both cases. Hence, it is necessary to dene the particle
Reynolds number in a more general form as (Dedegil, 1987)
Re
p
=
j
L
v
2
s
t
, (24)
where t is the shear stress, which is determined by the rheo-
logical model of the uid at representative shear rate v
s
/d
p
.
The effect of concentration on settling velocity needs to be
considered when calculating the free terminal settling ve-
locity of a single particle, because its effect is signicant at
high concentrations. Therefore, settling velocity obtained by
Eq. (22) has to be modied by the hindered settling factor
to account for hydrodynamic interference and particle col-
lision. For solids volume fractions between 0.001 and 0.4,
this factor is given as f
S
=e
5.9c
(Govier and Aziz, 1972).
2.6. Overall momentum and material balance
Finally, the solution for a given mode of transport is de-
termined by solving these sets of equations simultaneously.
However, the numerical calculation procedure can be sim-
pler for a sliding bed case, if the overall momentum and
mass balance equations are expressed as
U
s
A
s
+U
d
A
d
+U
b
A
b
=U
ave
A
t
, (25)
A
t
dp
dx
(t
s
s
s
) (t
b
s
b
+F
dyn
)
(j
s
A
s
+j
b
A
b
)g cos : =0, (26)
where U
s
, U
d
, and U
b
are the mean ow velocity of the
layers. The average velocity of the stream, U
ave
, is the ratio
of the total ow rate of the suspension to the total cross-
section area of the channel. A general formula for the mean
ow velocities of the layers can be written as
U
i
=
Q
i
A
i
=u
i
v
i
c
i
cos :, (27)
where Q
i
and A
i
are the ow rate and ow cross-sectional
area of a given layer, respectively. The subscript i can be
any of the subscripts (s, d, b or ave) used in Eqs. (1) and (2).
At high inclination angles (i.e., cos : 0), the following
assumptions can be made: U
s
=u
s
; U
d
=u
d
; and U
b
=u
b
. For
incompressible ow, the total ow rate across the channel
is given by
Q=Q
s
+Q
d
+Q
b
. (28)
When a stationary bed slides the relative velocity between
the bed and the suspended layer approaches the critical ve-
locity and results in the formation of a very thin dispersed
layer. Hence, the dry friction forces and the uid shear stress
forces acting between the channel wall and the dispersed
layer can be neglected in the derivation of Eq. (26).
2.7. Average transport rate
To compare the model prediction with the measured data,
we need to predict the average transport rate of solids.
Eq. (1) only determines the transport rate at a given bed
height or sand volume. The average transport rate is esti-
mated by taking the weighted average as
T
ave
=
1
V
s
_
0
V
s
T
r
dV, (29)
2564 A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570
where V denotes the volume of sand during the test that
ranges from the initial sand volume, V
s
to zero.
3. Computation procedures and modes of transport
A system of 11 equations ((2), (4), (9), (12), (15a), (16),
(18), (21) and (27) for each layer) with 11 unknowns (c
s
, c
d
,
U
s
, U
d
, U
b
, u
s
, u
d
, u
b
, h
d
, t
dis
, and dp/dx) can be estab-
lished, when all the constitutive relations are substituted into
the conservation equations. This system is non-linear and
requires an iterative procedure to get a numerical solution.
Determining the mode of cutting transport and obtaining a
stable numerical procedure are the most important parts of
the three-layer modeling. In order to determine the mode of
transport, the numerical procedure needs some criteria for
the range of ow parameters. This requires in-depth under-
standing of the model equations and their physical interpre-
tations. Thus, it is essential to see the inuence of modes of
transport on the key ow parameters such as pressure-drop
and ow rate.
Fig. 4 presents possible modes of solids transport in in-
clined channels. At low ow rates, the ow velocity is not
sufcient enough to initiate the movement of the bed parti-
cles; and a two-layer ow pattern shown in Fig. 4a will be
formed. The two-layer ow pattern has a stationary bed of
solids and clear liquid layer. However, as the ow velocity
increases, the intensities of hydrodynamic forces grow in
strength and eventually reach a stage at which these forces
are strong enough to lift the topmost particles of the bed.
At this stage, the three-layer ow pattern with a dispersed
layer establishes as shown in Fig. 4ce. Therefore, in the
model, the thickness of a dispersed layer (h
d
) can be used
to determine the presence of the dispersed layer or the tran-
sition from a two-layer to three-layer ow pattern. As the
ow velocity further increases, the dispersed layer will get
thicker. In this condition, the intensity of the hydrodynamic
and dispersive shear stresses at the interface between the
bed and the dispersed layer may become strong enough to
slide the stationary bed. The sliding of the bed occurs when
the dry friction force (F
b
) is greater than the static friction
force (F
st
). If the bed does not slide at a higher ow veloc-
ity, then the bed layer will get thinner and eventually disap-
pear. The ow pattern changes to a two-layer ow pattern
(suspended and dispersed layers) as shown in Fig. 4f. Fur-
ther increase in the ow velocity will suspend all the bed
particles and form a fully suspended ow pattern (Fig. 4g).
However, in its present form, the model cannot predict the
transition from a three-layer ow pattern to a two-layer or a
fully suspended ow pattern.
In addition to the dispersed layer thickness (h
d
) and dry
friction force (F
b
), the pressure loss can give some infor-
mation about the transition from stationary bed to sliding
bed ow condition. In order to see the effect of sliding on
the hydraulics, the frictional pressure loss is estimated us-
ing the classical two-layer model approach for sliding bed
Flowrate
P
r
e
s
s
u
r
e

d
r
o
p
Stationary bed
Sliding bed
D
A
q
min
C
B
E
q
max
F
q
s
Fig. 6. Flow characteristic curves of stationary and sliding beds.
ow with a low-friction factor and stationary bed ow. The
frictional pressure loss is presented in Fig. 6 as a function
of the ow rate for both stationary and moving bed cases.
The pressure loss for the stationary bed ow is higher than
the sliding bed ow, because the stationary bed ow has rel-
atively lower hydraulic diameter and higher ow velocity.
An important feature of the sliding bed pressure loss curve
is the presence of a minimum ow rate below which the
bed becomes stationary. Hence, the sliding bed does not ex-
ist below this ow rate. In order to get the overall picture,
let us begin the cleaning at point A as shown in Fig. 6 and
increase the ow rate from this point on. If we increase the
ow rate so rapidly that the transport rate of the bed, which
increases the hydraulic diameter, is negligible, then we will
reach point B where the bed starts to slide over the chan-
nel wall. Consequently, the ow mode becomes the slid-
ing bed ow and the pressure loss has to follow the lower
curve. As a result, the pressure loss reduces to the value that
corresponds to point C. Previous experimental and simula-
tion studies (Doron and Barnea, 1993; Nguyen and Rahman,
1998) conrmed a similar pressure loss reduction at the be-
ginning of sliding. A further increase in the ow rate does
not change the ow pattern. The pressure loss continuously
increases and follows the curve CD for higher ow rates.
However, if the ow rate decreases after reaching point C,
the curve follows the path CE, because the dynamic friction
factor is lower than the static friction factor. Further decrease
in ow rate stops the sliding of the bed and increases the
pressure loss to the corresponding value of point F. The up-
per curve begins to govern the course of the pressure loss.
As indicated in the gure, the pressure curve may show a
hysteresis depending on the static and dynamic friction fac-
tors between the bed and the channel wall. When the ow
rate reaches q
max
, the stationary bed is at the onset of slid-
ing. This condition can be extremely unstable. The stability
of a bed mainly depends on the friction factor of the chan-
nel, and roughness and thickness of the bed. When a con-
stant pressure is maintained at the beginning of sliding, then
A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570 2565
Table 1
Common computational data for the model predictions
Particle density 2600 kg/m
3
Channel diameter 70 mm
Density of water 1000 kg/m
3
Static friction factor 0.5
Density of PAC solution 1000 kg/m
3
Dynamic friction factor 0.25
Viscosity of water 0.001 Pa s
the ow rate abruptly increases from q
max
to q
s
. Therefore,
the critical sliding condition can be one of the causes of the
ow instability we have in stratied solidliquid ows.
The model simulations were performed assuming the
same condition as the test. The average particle size was
used as the particle size. Common computational data for
the model predictions are listed in Table 1 .
4. Experimental equipment and procedures
The experiments were conducted in a ow loop that has
a transparent test section with uid re-circulation facilities.
The ow loop is designed so that the ow is fully developed
and free of any entrance effect at the test section. A simpli-
ed schematic diagram of the loop is presented in Fig. 7.
As shown from the gure, the loop consisted of the follow-
ing items: (i) test section (a 4-m long transparent pipe with
internal diameter of 70 mm); (ii) overhead tank to maintain
constant pressure head at the inlet of the channel; (iii) cir-
culation tank for preparing and re-circulating the test uids
(water and PAC solutions); (iv) hydrocyclone to separate the
solids downstream of the channel; and (v) centrifugal pump
to recycle the uid. The loop is equipped with a control
mechanism to maintain a more or less constant ow rate of
uid during a test run. Details of the ow loop and equip-
ment that were used in the experiment are available in pre-
viously published studies (Ramadan, 2001; Ramadan et al.,
2003).
The uid level in the overhead tank was kept relatively
constant in order to maintain steady-state ow condition in
the test section. Minimumand maximumlevel switches were
used to regulate the pump in accordance with the uid level
in the overhead tank. A magnetic ow meter was placed
Test Section Flow
Meter
Differencial Pressure
Transmitter
Hydrocyclone
Sand
Overhead
Tank
Control Valve
Centrifugal
Pump
Mixing &
Recirculation Tank
DP
FR
Fig. 7. Schematic diagram of the ow loop.
upstream of the channel to measure the ow rate. The pres-
sure loss across the test section was continuously measured
using a differential pressure transmitter. Both the ow me-
ter and differential pressure transmitter were connected to a
personal computer for on-line display and recording. Tem-
perature measurements of the uid were made at the circu-
lation tank. A manually operated valve, placed downstream
of the hydrocyclone, was used to regulate the ow rate.
The average transport rate tests were performed by placing
the sand sample uniformly in the test section and measuring
the time of complete removal of the sample, while keeping
the ow rate constant. Critical ow rate (critical velocity)
tests were performed before every transport rate test. The
critical ow rate was measured by visual observation of
the movement of the bed particles (Ramadan et al., 2003).
Rheologies of the PACsolutions were measured before every
test and maintained at K=0.050 Pa s
0.7
and n=0.7. During
the test run, the temperature of the test uid was maintained
at 20

C by placing a heating element inside the circulation


tank.
5. Result and discussion
The conditions for the experiments were selected to cover
the stationary bed condition at different ow rates (ow ve-
locities). Physical characteristics of the sand beds and mea-
sured critical ow rates are presented in Table 2 (Ramadan
et al., 2003). Figs. 8 ad present measured average transport
rate and model predictions for water. Both the measured and
model predicted transport rates show a generally increasing
trend with the ow rate. The model predictions show a sat-
isfactory agreement with the measured data for the coarse
sand beds (S3 and S4) regardless of the bed thickness and
angle of inclination. As the ow rate approaches the criti-
cal ow rate, the model prediction becomes inadequate for
the ne sand beds (S1 and S2). During the ne sand bed
tests, we have also observed the formation of dunes and rip-
ples. Henderson (1966) suggested that the transport mech-
anisms of solids depend greatly on the difference between
the critical velocity and the mean ow velocity, and the par-
ticle size. As the difference between these two ow rates
decreases, there is a great possibility of dune and ripple for-
mation. This phenomenon predominantly occurs when the
bed particles are less than the viscous sublayer. The forma-
tion of dunes and ripples considerably affects the intensity
and modes of solids and cutting transport. As a result, the
three-layer model, which assumes a uniform bed thickness,
fails to predict the transport rate near the critical ow rate.
The grain Reynolds number is often used to describe the
possibilities of dune and ripple formation in sediment trans-
port (FredsZe and Deigaard, 1992). After simplication, a
generalized form of the grain Reynolds number is written as
Re
G
=
j
L
U
2
t
t
o
, (30)
2566 A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570
Table 2
Average critical ow rates of water and PAC solution with 1-litter sand bed in a horizontal channel
Particle Particle size Mean particle Critical velocity (m/s) Range of Re (10
4
) Range of Re
G
Critical ow rate (10
3
m
3
/s)
size code range (mm) size (mm)
PAC Water PAC Water PAC Water PAC Water
S1 0.1250.50 0.38 0.51 0.26 0.430.70 1.803.47 34 917 1.83 0.92
S2 0.501.20 0.85 0.68 0.32 0.430.66 1.953.32 47 3542 2.43 1.14
S3 2.003.50 2.75 0.54 0.50 0.330.66 2.943.40 917 130230 1.95 1.81
S4 4.505.50 5.00 0.54 0.56 0.270.43 3.235.60 1325 230405 1.93 2.02
Flow Rate [10
-3
m
3
/s]
1.5 2.0 2.5 3.0
T
r
a
n
s
p
o
r
t

R
a
t
e

[
K
g
/
m
i
n
]
0.1
1
S1 Model
S2 Model
S3 Model
S1, Re
G
=9-14
S2, Re
G
=22-35
S3, Re
G
=131-155
Flow Rate [10
-3
m
3
/s]
1.5 2.0 2.5 3.0
T
r
a
n
s
p
o
r
t

R
a
t
e

[
K
g
/
m
i
n
]
0.1
1
S1 Model
S2 Model
S3 Model
S1, Re
G
=10-13
S2, Re
G
=23-44
S3, Re
G
=150-178
(a)
(c) (d)
(b)
Flow Rate [10
-3
m
3
/s]
1.5 2.0 2.5 3.0 3.5 4.0
T
r
a
n
s
p
o
r
t

R
a
t
e

[
K
g
/
m
i
n
]
0.01
0.1
1
S1 Model
S2 Model
S3 Model
S4 Model
S1, Re
G
=9-17
S2, Re
G
=35-42
S3, Re
G
=130-230
S4, Re
G
=230-405
Flow Rate [10
-3
m
3
/s]
1.5 2.0 2.5 3.0
T
r
a
n
s
p
o
r
t

R
a
t
e

[
K
g
/
m
i
n
]
0.1
1
S1 Model
S2 Model
S3 Model
S1, Re
G
=10-12
S2, Re
G
=24-37
S3, Re
G
=140-166
Fig. 8. Measured and predicted average transport rates as a function of total ow rate for water: (a) 1-litter sand bed and : =90

; (b) 2.5-litter sand bed


and : =90

; (c) 4-litter sand bed and : =90

; (d) 1-litter sand bed and : =78

.
where t
o
is the shear stress, which is determined by the
rheological model of the uid at the representative shear
rate, U
t
/d
p
. The ranges of grain Reynolds numbers for the
sand beds are presented in Table 2. The grain Reynolds
number increases with increasing bed particle size. Model
predictions are reasonable at high grain Reynolds numbers.
Figs. 9ad present the frictional pressure losses measured
at the beginning of the transport rate test and the model
predictions for water test. During the test, the ow was fully
turbulent with the Reynolds number ranging from 18,000 to
56,000. In general, the frictional pressure loss predictions
show a satisfactory agreement with the measured data.
In order to determine the effect of rheology, transport rate
measurements and simulations were performed using PAC
solutions. The tests were performed at low Reynolds num-
bers that were ranging from 2700 to 7000. Model predicted
average transport rates are presented in Fig. 10a together
with the measured ones. For the ne sand beds (S1 and S2),
the gures indicated that the model predictions are consider-
ably higher than the measured data. For PAC solutions, the
thickness of viscous sublayer is greater than that of water
because of its high viscosity. Hence, the viscous sublayer
completely buries protruding particles of the ne beds (S1
and S2) forming dunes and ripples. Dunes and ripples were
formed during the experiment with the ne sand beds. As a
result, the model prediction becomes unsatisfactory for the
ne sand beds at low ow rates. The discrepancies between
the predicted and measured transport rates increase as the
ow rate and grain Reynolds number decrease. For coarse
sand beds (S3 and S4), the model predictions are accept-
able, although the accuracy the model reduces as the grain
Reynolds number decreases.
Fig. 10b presents pressure loss measurements along with
the model predictions for the PAC solution. For the ne
sand beds (S1 and S2), the model predictions are consid-
erably higher than the measured values. Model predictions
for coarse sand beds (S3 and S4) are relatively acceptable,
although they have some discrepancies with the measured
A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570 2567
Flow Rate [10
-3
m
3
/s]
1.5 2.0 2.5 3.0
F
r
i
c
t
i
o
n
a
l

P
r
e
s
s
u
r
e

L
o
s
s

[
P
a
/
m
]
10
100
S1 Model
S2 Model
S3 Model
S1 Measured
S2 Measured
S3 Measured Re=2.8 - 5.0 E+04
Flow Rate [10
-3
m
3
/s]
1.5 2.0 2.5 3.0
F
r
i
c
t
i
o
n
a
l

P
r
e
s
s
u
r
e

L
o
s
s

[
P
a
/
m
]
10
100
S1 Model
S2 Model
S3 Model
S1 Measured
S2 Measured
S3 Measured
Re=3.0 - 4.5 E+04
1.5 2.0 2.5 3.0
F
r
i
c
t
i
o
n
a
l

P
r
e
s
s
u
r
e

L
o
s
s

[
P
a
/
m
]
10
100
S1 Model
S2 Model
S3 Model
S1 Measured
S2 Measured
S3 Measured Re=2.8 - 4.8 E+04
1.5 2.0 2.5 3.0 3.5 4.0
F
r
i
c
t
i
o
n
a
l

P
r
e
s
s
u
r
e

L
o
s
s

[
P
a
/
m
]
10
100
S1 Model
S2 Model
S3 Model
S4 Model
S1 Measured
S2 Measured
S3 Measured
S4 Measured Re=1.8 - 5.6 E+04
(d) (c)
Flow Rate [10
-3
m
3
/s] Flow Rate [10
-3
m
3
/s] (b) (a)
Fig. 9. Measured and predicted pressure gradient as a function of total ow rate for water: (a) 1-litter sand bed and : =90

; (b) 2.5-litter sand bed and


: =90

; (c) 4-litter sand bed and : =90

; (d) 1-litter sand bed and : =78

.
Flow Rate [10
-3
m
3
/s]
2.0 2.5 3.0 3.5 4.0
T
r
a
n
s
p
o
r
t

R
a
t
e

[
K
g
/
m
i
n
]
0.01
0.1
1
S1 Model
S2 Model
S3 Model
S4 Model
S1, Re
G
=3-4
S2, Re
G
=4-7
S3, Re
G
=9-17
S4, Re
G
=13-25
(a)
(b) Flow Rate [10
-3
m
3
/s]
2.0 2.5 3.0 3.5 4.0
F
r
i
c
t
i
o
n
a
l

P
r
e
s
s
u
r
e

L
o
s
s

[
P
a
/
m
]
100
1000
S1 Model
S2 Model
S3 Model
S4 Model
S1 Measured
S2 Measured
S3 Measured
S4 Measured Re=2.7 - 7.0 E+03
Fig. 10. Measured and predicted average transport rates and pressure
gradient as a function of total ow rate for PAC with 1-litter sand bed
and : =90

: (a) average transport rates as a function of total ow rate;


and (b) pressure gradient as a function of total ow rate.
data. Model predictions for the PAC solution are not as good
as that of water. The reasons for this can be the rheology
of the uids, Reynolds number and ow characteristics of
the polymer. According to Eq. (4), frictional pressure loss is
a function of the wall shear stress and the interfacial shear
stress. These two shear stresses mainly depend on the ow
velocity and friction factor. For water test, the ow was most
likely in the turbulent regime with Reynolds number greater
than 10,000. The PAC solution tests were conducted in the
transition regime. The accuracy of the friction factor cor-
relations, Eq. (6), is relatively low in the transition regime,
even though it is applicable for both turbulent and transition
regimes. In addition, the PAC solution to some extent ex-
hibits drag-reducing behavior, which may affect the velocity
prole, friction factor and the ow regime.
To show the effect of angle inclination on the transport
rate, theoretical simulations were made for 1-litter sand beds
with different particle sizes (S1 and S4) at different inclina-
tion angles. Results of the simulations are presented in Figs.
11a and b for water and PAC, respectively. The transport
rate decreases as the inclination angle increases. The gures
show the presence of two distinct parts in the transport ver-
sus inclination graph. At lower inclination angles, there is a
high-gradient part that suggests strong inuence of the angle
of inclination. The second part is a low-gradient part, which
is shown at higher inclination angles. In this part, the effect
of inclination on the transport rate can be neglected. There-
fore, the no-slip assumption, used by Nguyen and Rahman
(1998), which neglects the effect of inclination angle, is ac-
ceptable in the low-gradient part.
2568 A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570
Angle of Inclination [degrees]
30 40 50 60 70 80 90
T
r
a
n
s
p
o
r
t

R
a
t
e

[
K
g
/
m
i
n
]
1
2
3
4
5
Q=0.0017 m
3
/s S4
Q=0.0025 m
3
/s S4
Q=0.0033 m
3
/s S4
Q=0.0017 m
3
/s S1
Q=0.0025 m
3
/s S1
Q=0.0033 m
3
/s S1
Angle of Inclination [degrees]
30 40 50 60 70 80 90
T
r
a
n
s
p
o
r
t

R
a
t
e

[
K
g
/
m
i
n
]
1
2
3
4
5
(a)
(b)
Q=0.0017 m
3
/s S4
Q=0.0025 m
3
/s S4
Q=0.0033 m
3
/s S4
Q=0.0017 m
3
/s S1
Q=0.0025 m
3
/s S1
Q=0.0033 m
3
/s S1
Fig. 11. Model predictions of the transport rate as a function of angle of
inclination for different sand beds: (a) water and (b) PAC.
The effect of particle size on the transport rate as the in-
clination angle and ow rate change is also shown in Figs.
11a and b. For both uids, the inuence of particle size on
the transport rate is strong at high ow rates and low incli-
nation angles, and the coarse sand bed has a higher transport
rate than the ne bed. Cutting transport with water appears
more sensitive to the particle size. It is also worthwhile to
compare model-predicted cutting transport capacities of wa-
ter and PAC. The model predictions indicated that, at a given
ow rate, PAC has better cutting transport capacity than wa-
ter for this particular simulation condition. One possible ex-
planation for this could be the rheology of the PAC solution
with high consistency index (i.e., 0.052 Pa s
n
). Even though
the PAC solution has a shear thinning behavior, it is more
viscous than water even at higher shear rates. As a result,
the hydrodynamic force (drag force) acting on particles at
a give local uid velocity becomes relatively high for the
PAC solution. In addition to this, higher consistency index
reduces particle settling velocity of the suspended bed par-
ticles. The reduction in the settling velocity has a positive
effect in the cutting transport process. Therefore, because of
these two phenomena, the PAC solution shows better cutting
transport capacity than water. However, the PAC solution
requires higher critical velocity (Table 2) than water for the
ne beds, because the viscous sublayer tends to be thicker
for the PAC solution.
Flow Rate [10
-3
m
3
/s]
0
2
4
6
8
10
12
S1
S2
S3
S4
Flow Rate [10
-3
m
3
/s]
0
2
4
6
8
10
12
S1
S2
S3
S4
2.0 2.4 2.8 3.2 3.6 4.0
2.0 2.4 2.8 3.2 3.6 4.0

d
i
s
/

d
b

d
i
s
/

d
b
(a)
(b)
Fig. 12. Model predictions of the ratio of equivalent dispersive shear stress
(t
dis
) to the uid shear stress (t
db
) as a function of total ow rate in a
horizontal channel and for different sand beds: (a) water and (b) PAC.
Comparisons are made between the equivalent dispersive
shear stress (t
dis
) and the uid shear stress (t
db
) using the
simulation results. Figs. 12a and b present the ratio of these
two shear stresses as a function of the ow rate for water and
PAC, respectively. The gures reveal that the contribution of
the dispersive shear stress in the momentum exchange be-
tween the dispersed layer and the bed is considerably higher
than that of the uid shear stress. The ratio decreases as the
particle size increases due to the increase in the friction fac-
tor that directly inuences the uid shear stress. Especially
for water, the ratio is strongly inuenced by the particle size
that determines the bed roughness. The effect of bed rough-
ness on the friction factor is not as signicant for PAC solu-
tion as for water. This is partly due to the fact that the PAC
solution ow lies in the low Reynolds number range, and
partly due to the fact that the viscous sublayer tends to be
thicker for the PAC solution. In turbulent ows, the nature
of the ow is affected by the bed particle size (wall rough-
ness). The inuence of bed particle size can be effective as
long as the viscous sublayer near the bed is thin compared to
the particle size. This mostly happens at increased Reynolds
number. In practical applications, turbulent ows of non-
Newtonian uids usually fall in the low Reynolds number
range. As a result, for non-Newtonian uids such as poly-
mer solutions (PAC solutions), the inuence of the particle
size on the cutting transport rate is relatively less.
A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570 2569
Fig. 12a indicates that the inuence of the ow rate on
the stress ratio is insignicant for water. This means that
the dispersive shear stress and the uid shear stress increase
in the same magnitudes as ow rate increases, resulting in
approximately constant stress ratio. However, for the PAC
solution, the ratio increases signicantly as the ow rate
increases. In this case, the relative change in uid shear stress
with the increase in ow rate is less, because the ow is
in the transition regime and the friction factor signicantly
decreases as ow rate increases.
6. Conclusion
A three-layer solid transport model is developed and ap-
plied to simulate solid particle transport in inclined chan-
nels. The results from a series of experiments are com-
pared with the model predictions. The data obtained clearly
demonstrated the limitations and application ranges of the
model. Consequently, the present formulation of the three-
layer model can be applicable to predict the transport rate
of stationary beds for both Newtonian and power-law uids
if the ow rate is not close to the critical ow rate and the
grain Reynolds number is between 15 and 400. This means
that when the eroding uid velocity is relatively high over
stationary beds of coarse particles, the model is applicable
for both Newtonian and power-law uids. The model predic-
tions for different angles of inclination indicate that the no-
slip assumption in the formulation of the three-layer model
is acceptable at high inclination angles (horizontal and near-
horizontal).
Notation
A
b
cross-section area of the bed
A
d
cross-section area of the dispersed layer
A
s
cross-section area of the suspended layer
A
t
ow cross-section area of the channel
c concentration
c
b
mean volumetric concentration of solids in the
bed
c
d
mean volumetric concentration of solids in dis-
persed layer
c
s
mean volumetric concentration of solids in the
suspended layer
c
sl
local volumetric concentration of solids in the
suspended layer
c
ave
average concentration of solids
d
p
mean diameter of bed particles
dp/dx pressure gradient
D pipe diameter
D
h
hydraulic diameter of a layer
f
db
interfacial uid friction factor between the bed
and dispersed layer
f
s
friction factor between the channel and sus-
pended layer
F
b
dry friction force between the bed and channel
wall
F
d
dry friction forces on the boundaries of the chan-
nel wall and dispersed layer
F
db
dispersive force
F
dyn
dynamic friction force between the bed and
channel wall
F
st
static friction force between the bed and channel
wall
g gravitational acceleration
h bed height
h
d
thickness of the dispersed layer
h
s
thickness of the suspended layer
K uid consistency index
K
s
consistency index of the suspended layer
n ow behavior index
Q total ow rate of suspension through the channel
Q
b
ow rate of the bed
Q
d
ow rate of the dispersed layer
Q
s
ow rate of the suspended layer
Re Generalized Reynolds number
Re
p
particleReynolds number
Re
G
grain Reynolds number
S
b
perimeter of the bed in contact with the channel
wall
S
d
perimeter of contact surface between the channel
and dispersion layer
S
db
perimeter of contact surface between the bed and
dispersion layer
S
i
perimeter in contact between suspended and dis-
persed layers
S
s
channel perimeter in contact with the suspended
layer
T
r
transport rate of solids or volume ow rate of
solids
T
ave
average transport rate of solids or average vol-
ume ow rate of solids
u
ave
mean uid phase velocity in the mean ow
u
b
mean ow velocity of the bed
u
d
mean uid phase velocity in the dispersed layer
u
s
mean uid phase velocity in the suspended layer
U
ave
mean ow velocity, which is dened as the total
ow rate divided by the cross-sectional area of
the channel.
U
c
critical velocity
U
d
mean ow velocity of the dispersed layer
U
d,r
local velocity of the dispersed layer relative to
the bed
U
s
mean ow velocity of the suspended layer
U
t
friction velocity
v
d
average settling velocity of solid particles in the
dispersed layer
v
s
average settling velocity of solid particles in the
suspended layer
2570 A. Ramadan et al. / Chemical Engineering Science 60 (2005) 25572570
v
ave
average settling velocity of solid particles in the
total ow
V volume of sand
V
s
initial sand volume
y vertical distance from the mean bed level
Greek letters
: angle of inclination from vertical
I diffusion coefcient of solid particles
0 angular displacement
0
b
angular bed thickness of the bed
0
d
angular bed thickness for dispersed layer
von Karman constant
j
d
dry dynamic friction coefcient between the bed
and the channel wall
j
L
uid viscosity
j
s
static friction coefcient between the bed and
channel wall
j
p
density of solid particles
j
s
density of suspended layer
j
d
density of dispersed layer
j
L
uid density
t shear stress acting on settling particles at repre-
sentative shear rate
t
b
interstitial uid shear stress between the bed and
channel wall
t
d
uid shear stress between the channel wall and
suspended layer
t
db
uid shear stress between the bed and dispersion
layer
t
dis
equivalent dispersive shear stress between the
bed and dispersed layer
t
i
interfacial shear stress between the suspended
layer and dispersed layer
t
s
interstitial shear stress between the channel and
suspended layer
[
D
equivalent dynamic friction angle
Acknowledgements
The authors express their appreciation to the staff in the
workshops and laboratories at the Department of Petroleum
Engineering and Applied Geophysics, NTNU for their assis-
tance in building the ow loop and Statoil for the nancial
support of this project.
References
Bagnold, R.A., 1954. Experiments on a gravity-free dispersion of large
spheres in a Newtonian uid under shear. Proceedings of the Royal
Society of London (A) 225, 4963.
Cho, H., Shah, S.N., Osisanya, S.O., 2002. A three-segment hydraulic
model for cuttings transport in coiled tubing horizontal and deviated
drilling. Journal of Canadian Petroleum Technology 41 (6), 3239.
Davis, R.H., 1996. Hydrodynamic diffusion of suspended particles: a
symposium. Journal of Fluid Mechanics 301, 325335.
Dedegil, M.Y., 1987. Particle drag coefcient and settling velocity of
particles in non-Newtonian suspensions. Journal of Fluids Engineering-
Transactions, ASME 109 (3), 319323.
Doron, P., Barnea, D., 1993. A three-layer model for solidliquid ow
in horizontal pipes. International Journal of Multiphase Flow 19,
10291043.
Doron, P., 1994. Investigation of hydrodynamic phenomena in the ow of
solidliquid mixtures in pipes. Ph.D. Dissertation, Tel-Aviv University,
Tel-Aviv.
Einstein, A., 1906. Investigation on the theory of the Brownian movement.
Annals of Physics 19, 371.
Ford, J.T., Peden, J.M., Oyeneyin, M.B., Gao, E., Zarrough, R., 1990.
Experimental investigation of drilled cuttings transport in inclined
boreholes, SPE paper 20421, Presented at the 1990 SPE Annual
Technical Conference and Exhibition, New Orleans, 2326 September.
FredsZe, D., Deigaard, R., 1992. Mechanics of Coastal Sediment
Transport. World Scientic Publishing Co., Singapore, pp. 194226.
Gavignet, A.A., Sobey, I.J., 1989. Model aids cuttings transport prediction.
Journal of Petroleum Technology 41, 916921.
Govier, G.W., Aziz, K., 1972. The Flow of Complex Mixtures in Pipes.
Robert E. Krieger Pub. Co., Malabar, FL, pp. 1322.
Henderson, F.M., 1966. Open Channel Flow. Macmillan Series, NewYork.
Mohs, B., Mittendorff, I., Oliemanns, R.V.A., 2000. Results from a two-
dimensional turbulent-model for dispersion and deposition of droplets
in horizontal annular dispersed gas/liquid ow. International Journal
of Multiphase Flow 26, 949975.
Nguyen, D., Rahman, S.S., 1998. A three-layer hydraulic program for
effective cuttings transport and hole cleaning in highly deviated and
horizontal wells. SPE Drilling & Completion 13 (3), 182189.
Ramadan, A., 2001. Mathematical modeling and experimental
investigation of solids and cuttings transport. Ph.D. Dissertation,
Norwegian University of Science & Technology, Trondheim, Norway.
Ramadan, A., Skalle, P., Johansen, S.T., 2003. A mechanistic model to
determine the critical ow velocity required to initiate the movement
of spherical bed particles in inclined channels. Chemical Engineering
Science 58 (10), 21532163.
Rasi, M., 1994. Hole cleaning in large, high-angle wellbores. SPE Paper
27464, Presented at the IADC/SPE Drilling Conference, Dallas, 1518
February.
Szilas, A.P., Bobok, E., Navratil, L., 1981. Determination of turbulent
pressure loss of non-Newtonian oil ow in rough pipes. Rheologica
Acta 20 (5), 487496.
Televantos, Y., Shook, C., Carleton, A., Streat, M., 1979. Flow of slurries
of coarse particles at high solids concentrations. Canadian Journal of
Chemical Engineering 57, 255262.
Wilson, K.C., 1976. A unied physical-based analysis of solid-liquid
pipeline ow. In Proc. 4th Int. Conf. on the Hydraulic Transport of
Solids in pipes, Banff, Alberta, 116.
Wilson, K.C., 1987. Analysis of bed load motion at high shear stresses.
Journal of Hydrologic Engineering ASCE 113 (1), 97103.
Wilson, K.C., Pugh, F.J., 1988. Dispersive-force modeling of turbulent
suspension in heterogeneous slurry ow. Canadian Journal of Chemical
Engineering 66 (5), 721727.
White, F.M., 1974. Viscous Fluid Flow. McGraw-Hill, New York.
Young, J., Leeming, A., 1997. A theory of particle deposition in turbulent
pipe ow. Journal of Fluid Mechanics 340, 129159.

You might also like