You are on page 1of 21

Chemical Engineering Science 62 (2007) 2068 2088

www.elsevier.com/locate/ces

Dilute gassolid two-phase ows in a curved 90 duct bend: CFD simulation


with experimental validation
B. Kuan , W. Yang, M.P. Schwarz
Cooperative Research Centre for Clean Power from Lignite Division of Minerals, Commonwealth Science and Industrial Research Organisation, Box 312,
Clayton South, Victoria 3169, Australia
Received 31 May 2006; received in revised form 20 October 2006; accepted 21 December 2006
Available online 13 January 2007

Abstract
Computational uid dynamics (CFD) simulations of dilute gassolid ow through a curved 90 duct bend were performed. Non-uniform
sized glass spheres with a mean diameter of 77 m were used as the dispersed phase. The curved bend is square-sectioned (150 mm 150 mm)
and has a turning radius of 1.5D (D = duct hydraulic diameter). Turbulent ow quantities for Re = 100, 000 were calculated based on a
differential Reynolds stress model. The solids mass loading considered is 0.00206 and hence justies the application of one-way coupling
between gas and particles. A Lagrangian particle-tracking algorithm which takes into account the effect of shear-slip lift (SSL) force on particles
and particle-wall interactions (PWIs) has been utilised to predict velocities of the dispersed phase. The predictions were compared against the
experimental data measured using LaserDoppler Anemometry (LDA). The study found that the predicted gas ow eld has a strong inuence
over the predicted particle velocities. PWI model considerably affects the prediction of particle velocity and distribution of particles at the
inner duct wall within the bend. Inclusion of the SSL force also helps the distribution of the particle tracks towards the duct centre in the
vertical duct downstream of the bend. Within the bend, particle velocities near the inner wall have been grossly over-predicted in the simulation,
especially at mid-bend. The present study thus highlights the importance of the predicted gas ow eld, SSL force and particle-wall collisions
to Lagrangian particle tracking.
2007 Elsevier Ltd. All rights reserved.
Keywords: Multiphase ow; Mathematical modelling; Simulation; Penumatic conveying; LDA; Turbulence

1. Introduction
Elbows and bends are commonly used in pneumatic conveying systems to change ow direction so as to transport the suspended material to the desired delivery point within a limited
space. In the case of coal-red power plants that operate
on a continuous supply of pulverised coal to furnaces, maldistribution of pulverised fuel often occurs as coal particles
are pneumatically transported from the mill through ducts
consisting of numerous bends and straight sections.
Apart from the duct geometry, the coal pulverisation process
is also a strong contributor to mal-distribution of the pulverised
fuel. Field measurements at a lignite-red power plant carried
out to support this study found that, depending on factors such
Corresponding author. Tel.: +61 3 95458687; fax: +61 3 95628919.

E-mail address: benny.kuan@csiro.au (B. Kuan).


0009-2509/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2006.12.054

as feed size and mill speed, the fuel pulverisation process produces coal particles ranging between 10 and 1000 m in size.
Fig. 1 shows a typical coal particle size distribution measured
at the mill outlet of a lignite-red power station (McIntosh and
Borthwick, 1984). The extreme non-uniformity in particle size
combined with a centrifugal effect arising from the duct bend
are believed to lead to the formation of a stratied gassolid
ow, known as a particle rope, downstream of the elbow even
at a low solids mass loading, L < 0.1. This invariably creates
difculties for the plant operators to monitor and control the
pulverised fuel supply to individual burners, and hence to maintain an optimal combustion condition inside the furnace.
There are a large number of documented studies, both numerical and experimental, on particle roping in dust conveying
systems with solids mass loading L > 0.3, but most of them
focus on particles with a size distribution that is either heavily
skewed towards the lower end of the size range (Yilmaz and

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Fig. 1. Typical coal size distribution at the mill outlet.

Levy, 1998; Akilli et al., 2001) or more evenly spread across


the entire size range (Huber and Sommereld, 1994). In both
cases, the tested particle samples contain a considerable amount
of ne particles. A comprehensive experimental investigation
was conducted by Huber and Sommereld (1994) using phaseDoppler anemometry (PDA) in various pipe elements. They
used glass particles with a mean diameter of 40 m and their
results demonstrated the signicance of wall-roughness in preventing the glass beads from settling in a horizontal pipe.
The research group at Lehigh University has conducted several physical experiments ( Yilmaz and Levy, 1998, 2001; Akilli
et al., 2001; Bilirgen and KLevy, 2001) investigating formation and dispersion of the particle ropes involving pulverised
coal particles. A closed ow loop consisting of two 6.1 m horizontal pipes, one 3.4 m vertical pipe and two 90 elbows with
adjustable turning radii was set up to facilitate measurements
of particle rope characteristics in a vertical pipe following a
horizontal-to-vertical bend. All pipe elements had an internal
diameter of 0.154 m. They identied a number of parameters
that affect the behaviour of the particle rope downstream of
a horizontal-to-vertical duct bend. These included solids mass
loading, conveying velocity, and duct bend radius.
Levy and Mason (1998) reported a similar study utilising
three different particle sizes. Their numerical simulations indicated a change in particle roping characteristics with particle
size: coarse particles form a concentrated rope which propagates more than 2.4D downstream of the bend; for the ne
particles, they only noticed a localised peak in particle concentration just downstream of the bend. There were however
no published experimental data at the time to validate their
calculations.
The effect of particle size on solid- as well as gas-phase
motions in vertical and horizontal two-phase ows has been
extensively studied in the literature. With the aid of a laserDoppler velocimeter (LDV), Tsuji et al. (1984) tested plastic
particles with diameters ranging from 200 m to 3 mm in a
vertical pipe having an inner diameter of 30.5 mm and studied
the effect of particle size on particle motion and gas turbulence.
At high solids mass loading (L > 2.0), they found the mean
air velocity prole attens across the duct cross-section with
decreasing particle size. Fan et al. (1997) performed a similar
experiment with 100 and 300 m quartz particles in a 100 mm

2069

diameter vertical tube. At L < 2.0, they observed a attening of


the mean particle velocity prole with increasing particle size.
Tsuji and Morikawa (1982) has also performed the same
LDV measurements for a horizontal two-phase ow, but only
for L > 2.0. They observed a stronger inuence of the particles on the background gas turbulence as the particle size
increased. Lan et al. (2002) experimentally studied the effect of
particle size, solids mass loading and wall roughness on a twophase horizontal channel ow. While all tests were performed
at L > 0.2, their results indicate that the ow was increasingly
sensitive to particle size distribution as L decreases.
In pneumatic transport, it is commonly accepted that movement of large particles is predominantly inuenced by particlewall interactions rather than uid turbulence. In recognition of
this, Frank et al. (1993), Kussin and Sommerfeld (2002), and
Sommerfeld and Huber (1999) conducted series of experiments
to study particle-wall interaction for a large combination of particle size, material and wall roughness in an attempt to create
a reliable physical particle-wall collision model.
Compared to the ow regime prevailing at the power plant
mill duct (i.e., L < 0.1), all of the above-mentioned studies
examined gassolid ow behaviour in the L > 0.2 range and
hence their results were not representative of the dilute mill-duct
ow considered in the present study. Further, most of the published conveying duct experiments only consider the scenario
where a substantial amount of ne particles were present in the
system. In the present study, a dispersed phase with a wider size
distribution but a smaller fraction of ne particles is considered
so as to better represent the gassolid ow system prevailing in
the mill-duct ow. Numerical simulation of such a ow using
Lagrangian particle tracking technique requires a large set of
particle trajectories to be solved. A special discretisation technique has been developed to overcome the need for an exceedingly large number of particle tracks. This paper aims to study,
both numerically and experimentally, the behaviour of a dilute
gassolid ow which uses non-uniform sized spherical glass
beads as the dispersed phase and at a low solids mass loading
L = 0.00206. Glass spheres having a volume weighted mean
diameter of 77 m were tested in purpose-designed laboratory
experiments which involve a square-sectioned horizontal-tovertical 90 bend.
Numerical simulations were performed for the gas phase and
then followed by the solution of solids motion. The solution of
particle motion is based on a Lagrangian particle tracking approach where a single particle track represents the trajectories
of a group of particles of the same size. The calculation considered various hydrodynamic forces acting on the particles, such
as drag and shear-slip lift (SSL), as well as particle wall interactions. The hydrodynamic forces were determined based on
the knowledge of the predicted gas ow. An accurate solution
of the gas ow eld is thus crucial for a realistic prediction of
the particle motion.
One main feature of the square-sectioned duct ow considered in this study is a strong curvature in the streamwise direction along the duct length. This necessitates the application of
a turbulence model that is capable of resolving the extra strains
arising from curvature in the streamlines. There exists a wide

2070

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

range of turbulence models in the published literature dealing


with streamline curvature effects. Nallasamy (1987), and Patel
and Sotiropoulos (1997) have both provided comprehensive
reviews on various approaches to model streamline curvature
effects in the turbulence model.
Gibson (1978) and Rodi and Scheuerer (1983) have proposed
a number of ad hoc modications to the standard two-equation
turbulence model to address this problem. However, the applicability of these model corrections has largely been limited to
ows with simple geometry where axis of streamline curvature
is xed relative to the principal ow direction, such as ow
over a back-ward step. Richmond and Patel (1991) have tested
one of the above models in various curved wall-bounded ows.
The model produced satisfactory mean velocity prediction on
the convex wall while under-estimating the mean velocities on
the concave wall.
Choi et al. (1989) have applied an algebraic Reynolds stress
model in the core ow and mixing-length hypothesis in the
viscous sublayer in an attempt resolve streamline curvature in
a square-sectioned U-bend. The model performs better than the
standard k. model and provides a reasonable representation
of the primary ow but it tends to over-predict gas velocity at
the convex wall and over-predict it at the concave wall.
Luo and Lakshminarayana (1996) have applied a differential
Reynolds stress model and a modied eddy dissipation rate 
transport equation proposed by Shih et al. (1995) to account
for streamline curvature effects. They tested their model in 2D
90 and 180 pipe bends. The results indicated a negligible
improvement on the predicted mean gas velocities as compared
to the case with unmodied differential Reynolds stress model.
However, the new model produced a higher level of turbulence
intensity at two stations inside and shortly downstream of the
bend.
Patel and Sotiropoulos (1997) has thus concluded that, for
ows containing streamline curvature, a differential Reynolds
stress model without any modications should provide reasonable ow prediction as compared to that based on casespecic two-equation turbulence models. The current study
adopts the differential Reynolds stress model of Speziale et al.
(1991) which is tensor-invariant and does not require the addition of wall-reection terms into its pressurestrain correlation
model.

of the converging nozzle follows Borgers contractor prole


(Borger, 1973) to ensure ow uniformity within 0.5%.
The duct geometry and ow coordinate system are shown
in Fig. 2(b). The square-sectioned (150 mm 150 mm) test
section is constructed using 10 mm thick Perspex sheets, and the
main components of the test facility include a 3.5 m horizontal
straight duct, a 90 bend with a turning radius R of 225 mm
and a 1.8 m vertical straight duct. The R/D ratio for the bend
is thus 1.5.
Glass spheres with a volume weighted mean diameter of
77 m were released into the gas ow from a uidised-bed
feeder at a rate of 2 kg h1 to give a solids mass loading L of
0.00206. A digital balance underneath the uidized-bed feeder
monitors the rate at which the particles are released so as to
ensure a dilute gassolid ow regime inside the test section.
Data on particle size distribution (based on volume fraction)
are obtained from a wet analysis using a Malvern particle size
analyser and are shown in Fig. 3.
Gas phase measurements were obtained at a bulk gas velocity Ub , of 10 m s1 in the absence of the glass spheres. A mist
of ne sugar particles generated from a jet atomiser using 5%
sugar solution have a mean diameter of 1 m and were used as
seeding particles in the single-phase measurement. Turbulence
intensity in the main stream was approximately 1% at the centre of duct cross-section 10D upstream of the 90 bend. The
Reynolds number, based on the bulk velocity, hydraulic diameter of the square-sectioned duct and kinematic viscosity of air
was subsequently 105 .
Flow measurements were performed on the duct vertical
symmetry plane. At each streamwise location along the duct
length, ow measurements, including mean gas and solids
velocities, and turbulence intensities were carried out. In nearwall areas, ow statistics were measured at 1 mm from the
duct walls. LDA measurements outside the vertical symmetry
plane were not taken because of poor access of the laser beam
to the probe locations. Secondary gas ow, which is characterised by a pair of counter-rotating vortices along the duct
length downstream of the bend, was not studied in the present
work. It is assumed to display similar characteristics to the
laboratory ow of Sudo et al. (2001) where R/D = 2.0.

2. Experimental test facility

In order to ensure the ow scenario tested in the laboratory


model reects the physical duct operating condition, a dynamic
similarity analysis was performed prior to the experiment. Our
analysis is based on the collective experience of Boothroyd
(1971) and Fan and Zhu (1998). In general they found the ow
systems that share similar geometric and kinematic boundary
conditions display similar dynamic responses when the dimensionless numbers Re, Rep , St, and Fr are the same

2.1. Rig conguration


All experiments were conducted in an open-circuit
horizontal-to-vertical suction wind tunnel system which was
set up for an earlier study by Yang and Kuan (2006). A
schematic sketch of the entire test facility is provided in
Fig. 2(a). It consists of an open-circuit suction wind tunnel
where the airow is drawn into the system by means of a centrifugal fan through an entry piece that consists of an elliptical
bell-mouth inlet and a honeycomb ow straightening section.
The air then passes through a converging nozzle to attain a
higher velocity before entering into the test section. Design

2.2. Dynamic similarity

Re =

 f Ub D

Rep =

 f UT d p


(1)
,

(2)

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

2071

Fig. 2. Experimental ow system: (a) schematic diagram; (b) duct geometry and ow coordinate system.

Fig. 3. Particle size distributions by volume ( wet analysis data;  modelled


distribution).

p dp2
p
; p =
;
f
18

Ub2
,
Fr =
Dg

St =

f =

D
,
Ub

(3)

(4)

whereby Re denotes duct Reynolds number; Rep is particle


Reynolds number; St is particle Stokes number as commonly
dened in the literature (Giddings et al., 2004); Fr is Froude
number for gassolid ows; UT denotes particle terminal
velocity; Ub is uid bulk velocity based on uid mass ow
rate and duct cross-section area.
Stokes number St is a ratio of particle response time to the
uid travelling time; a higher St thus implies more collisions
of the particles with the walls of the bend. The Froude number
Fr measures the dominance of inertial effects over gravity on
the airow.
Dynamic similarity parameters are compared in Table 1. The
table also lists some of the characteristic ow parameters for
both cases. The smaller duct diameter and particle sizes in the
laboratory system as compared to the industrial ow are largely
responsible for the lower Re, Rep , and St values at the upper
end of the particle size range. A lower Rep implies that the particles in the laboratory ow system are more tightly coupled to
the gas ow than in the industrial counterpart. The power plant
mill-duct ow contains a greater proportion of coarse particles
with a very high St, indicating a higher chance of particle-wall
collisions. However, motion of the larger particles is beyond the
scope of the present investigation because it is predominantly

2072

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Table 1
Flow and dynamic similarity parameters
Laboratory rig
Flow parameters
D (m)
 (kg m1 s1 )
f (kg m3 )
p (kg m3 )
L
dp (m)
UT (m s1 )
Ub (m s1 )
Re

4
(min.)
1.21 103

Dynamic similarity parameters


3.17 104
Rep
St
8.23 104
Fr
S/dp

Power station

0.15
1.8 105
1.18
2500
0.00206
77
(mean)
0.36
10.0
1.0 105
1.82
3.05
8.24
80.9

driven by particles own inertia as well as collisions with the


wall surfaces, instead of gassolids interactions.
As seen in Table 1, the current experiment covers the lower
end of the industrial particle size range which has lower Rep
and St values. One can thus expect the laboratory rig to realistically reproduce the characteristics of the power station millduct ow in the absence of very coarse particles.
Apart from dynamic similarity parameters that characterise the level of particlegas interactions in both systems,
Table 1 also lists inter-particle spacing S (with respect to dp ) as
estimated from


 1/3
S/dp =
,
(5)
6p
p = 
p

Lf

(6)
+1

which have been used by Sommerfeld (2000) to classify the


importance of interaction mechanisms in dispersed two-phase
ows. In general, effect of inter-particle collisions can be
ignored for an inter-particle space, S/dp greater than 10.
According to Table 1, both ow systems satisfy this condition
and hence inter-particle collisions were not considered in this
study.
3. Mathematical models

160
(max.)
1.1

20
(5%)
0.0155

11.5
13.2

0.0124
0.063

1.76
1.95 105
0.78
1400
0.1
80
(50%)
0.224
28.6
2.0 106
0.716
1.701
6.88
21.1

410
(95%)
2.32

38.1
44.7

notations below
jUi
= 0,
dxi

(7)

 


jUi Uj
jUj
jUi
jP
j

ui uj ,
=
+
+
dxj
jxi
jxj
jxj
jxi

(8)

jUk ui uj


dxk

j
=
jxk



2
k2
 + Cs
3


jui uj

jxl

2
+ Gij ij + ij ,
3




2
t j
j
jUk 
+
+ C1 Gkk C2
=
,
dxk
jxk

 jxl
k
k
Gij = ui uk

jUj
jUi
uj uk
,
jxk
jxk

(9)

(10)

(11)

where Ui and ui are, respectively, mean and uctuating velocities; ui uj denotes the Reynolds stress tensor; Gij is the turbulence production term; ij is the modelled pressurestrain
correlation given by Speziale et al. (1991). Model constants Cs ,
C1 , C2 , and
 , respectively, have the following values: 0.22,
1.45, 1.83, and 1.375. This approach is known as differential
Reynolds stress modelling (DRSM).

3.1. Gas ow

4. Particle tracks

Steady-state, isothermal gas ow properties and turbulence quantities are calculated numerically by solving a set
of governing partial differential equations (PDE) using a
commercial CFD software ANSYS CFX-10. The Reynoldsaveraged NavierStokes equations and the Reynolds stress
transport equations considered are written in Cartesian tensor

4.1. Hydrodynamic forces


Instantaneous positions and velocities of the dispersed phase
are solved through a Lagrangian particle tracking method. Motion of individual particles suspended in a continuous uid is
determined by numerically integrating the equations of motion

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

for the dispersed phase in a uid ow. The equation of particle


motion may be expressed as
mp

dup
= FD + Fg + Fpg + FA + Fsl ,
dt

dxp
= up ,
dt

with uf = U
r =

+ u

(15)

24
fD .
Rep

As Rep increases beyond 1000, the corresponding fD leads to


a constant drag coefcient CD of 0.44.
The force components due to gravity Fg , added mass FA and
pressure gradient Fpg are, respectively, given by

Fg = m p 1

f
p


g,

(17)

f dup
1
FA = m p
,
2
p dt

(18)

Fpg = 41 dp3 P .

(19)

The present calculation adopts a SSL model formulated by Mei


(1992). In vector form, the model is written as
Fsl =

 3
d  Csl ((uf up ) f ),
8 p f

(23)

Assuming that all force components, except the drag, are constant during a time step t, Eq. (12) can be integrated analytically to yield

and the SchillerNaumann drag correlation fD for a sphere



1 + 0.15Re0.687
, Rep 1000,
p
(16)
fD =
0.01833Rep ,
Rep > 1000,
CD =

Res
;
Rep

(13)

and particle relaxation time r dened by

18fD

= 0.5

(14)

p dp2

f dp2 |f |

(12)

where subscript p represents particle properties and subscripts


D, g, pg, A and sl, respectively, denote force components arising
from drag, gravity, ow pressure gradient, added mass effect
and slip-shear lift.
The drag force is calculated from
uf u p
F D = mp
r

Res =

2073

f = uf .

up = uf + (up0 uf )et/r +

(24)

r
(1 et/r )
mp

(Fg + FA + Fsl + Fpg )

(25)

with superscript 0 indicating the start of a time step.


Similarly, Eq. (13) can be integrated analytically to
xp = xp0 + up t.

(26)

Eqs. (25) and (26) were solved within a given cell in the particle
tracking calculations.
In order to solve the instantaneous particle velocity and
location using the integrated equations of particle motion,
Eqs. (25) and (26), for every particle track in the ow domain,
the instantaneous uid velocity has to be specied at all particle locations. This is made possible through the application
of a classical stochastic approach of Gosman and Ioannides
(1981), which estimates uctuating gas velocity components
on the basis of isotropic turbulence. Subsequent particle track
integration which takes into account the turbulent dispersion
effect was thus carried out.
5. Particle-wall interactions
The present simulation adopts a modied version of the PWI
model of Matsumoto and Saito (1970). The base-case model
allows the particles to either slide along the wall surface when
the angle of incidence is small, or rebound away from the wall
after impact. However, it is based on the assumption of a constant restitution coefcient and dynamic friction, both of which
are sensitive to a range of parameters, such as incidence angle
and wall material, as found in published experimental investigations of Frank et al. (1993), Sommerfeld and Huber (1999).
For the present numerical calculation that involves collisions
between the glass spheres and the Perspex duct walls, particle
velocity components as well as particle angle of incidence are
dened in Fig. 4. Impact test data for glass beads on Plexiglass

(20)

where
4.1126
f (Rep , Res ),
Csl =
Res

(21)

1/2
(1 0.3314 )
e(0.1Rep ) + 0.3314 1/2 , Rep 40,
f (Rep , Res ) =

Rep 40,
0.0524( Rep )1/2 ,
(22)

Fig. 4. Denition of velocities before and after impact and particle angle of
incidence.

2074

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

plates are utilized to characterize the PWI (Sommerfeld and


Huber, 1999). The modied PWI model is given by
up2 = Et up1 ,
vp2 = En vp1 ,
En = max(1.0 0.0151 , 0.73),
d = max(0.4 0.009261 , 0.15)
ucrit = 3.5d (1 + En )|vp1 |

1.0 d (1 + En )|vp1 /up1 |,

up1 ucrit (sliding collision),


Et =

5/7,

up1 < ucrit (non-sliding collision).

(27)

The effect of wall roughness is introduced through a semiempirical approach that modies the smooth-wall incident
angle 1 with a random component characterising the presence
of a rough wall
1 = 1 +  ,

(28)

where  represents a random component sampled from a


Gaussian distribution function. is a Gaussian random number
with zero mean and standard deviation of unity, and  is the
standard deviation of the wall roughness angle.  is approximately 3.8 for a Plexiglass plate.
In our simulations, we have assumed the particles are nonrotating and the effect of particle rotation on particle velocity
after the collision is negligible.
6. Numerical procedure
The PDE (8)(10), are discretised following a nite volume approach. The advection terms were approximated using
a scheme developed by Barth and Jesperson (1989) which
is more than rst-order accurate at mesh discontinuities and
provides higher-order accuracy for smoothly varying meshes.

All calculations were performed on a 80 80 135 (y z s)


body-tted grid and the computed ow domain was constructed
using hexahedral cells. The grid has a minimum wall spacing
of y + < 16 and was found to be sufciently rened to produce
a grid-independent solution as shown in an earlier study by
Kuan (2005).
The current simulations utilise a set of fully turbulent inow
conditions generated from a separate calculation in a straight
horizontal pipe of the same cross-section. The solid particles
were released into the duct at the same velocity as the gas
phase, i.e., zero slip, and from random locations on the inlet
plane.
In a dilute two-phase ow where particle volume fraction is
less than 106 (i.e., L < 0.00211), it is generally accepted that
the transfer of particle momentum to the carrier-phase is negligible (Sommerfeld, 2000). Thus, at L=0.00206, the gas motion
could be considered independent of the solid-phase (i.e., oneway coupling). When particles are introduced into a turbulent
gas ow, they can either enhance or reduce the gas turbulence.
This is known as turbulence modulation which is assumed to
be insignicant in our simulation.
In the general dilute gassolid ows, the quality of the
dispersed-phase prediction based on Lagrangian particle tracking is directly dependent on the number of particle tracks solved
in the simulation. In cases where the particle size distribution
is wide, one has to proportionally increase the total number
of the considered particle tracks so as to adequately resolve
the smallest size fraction. This often leads to an exceedingly
large set of particle tracks which are very computationally
demanding to solve and analyse.
In the present study, we have developed a methodology
which allows one to overcome this limitation. The particle size
distribution (by volume) as measured in a wet analysis by the
Malvern size analyser (Fig. 3) was rst discretised into 12 characteristic particle size fractions (Table 2). One can calculate
the associated particle volume ow rate for each size fraction
to give a total solids volume ow rate of 2.207E 7 m3 s1
such that L = 0.00206. A xed number of particle tracks were

Table 2
Modelled particle size distribution and a discrete representation of the particulate ow by particle tracks
i

dp,i (m)

Vol.%

Particle vol. ow
rate (m3 /s)

Ni

Particle vol. ow
rate per track (m3 /s)

i (1/s)

1
2
3
4
5
6
7
8
9
10
11
12

5
18
30
45
57
65
76
89
103
125
140
152

2.001E 5
5.202E 5
3.112E 3
2.905E 1
3.260
10.35
38.10
34.03
9.982
2.500
1.000
4.887E 1

4.416E 14
1.148E 13
6.868E 12
6.411E 10
7.196E 9
2.284E 8
8.409E 8
7.510E 8
2.203E 8
5.518E 9
2.207E 9
1.079E 9

8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000

5.520E 18
1.435E 17
8.585E 16
8.014E 14
8.994E 13
2.855E 12
1.051E 11
9.387E 12
2.754E 12
6.897E 13
2.759E 13
1.348E 13

8.434E 2
4.700E 3
6.073E 2
1.680
9.276
19.85
45.73
25.43
4.813
6.744E 1
1.920E 1
7.332E 2

100.0

2.207E 7

96 000

Total

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

2075

Fig. 5. Mean longitudinal gas velocity proles within the bend ( prediction;  data).

then allocated to each size fraction and we have used a total


of 96,000 particle tracks in the present study. Individual particle tracks in each size fraction thus carry a uniform portion
of the solids volume ow rate as shown in Table 2. One
can then determine a particle number ow rate i represented by each track and subsequently apply this to convert
number of particle tracks into number of real particles during the statistical averaging process as discussed in the next
section.

7. Collection and averaging of predicted particle statistics


Similar to the data acquisition process performed during the
physical experiment, the predicted particle statistics are collected at each measurement station on the duct centre-plane.
We have adopted the technique by Uijttewall and Oliemans
(1996) for estimating the mean solids ow quantities on the
measurement plane. This involves setting up 27 equally spaced
bins along the duct diameter to collect and store particle track

2076

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Fig. 6. Mean longitudinal gas velocity proles in the vertical duct ( prediction;  data).

information. The particle statistics stored inside each bin are


then averaged based on Table 2 and Eq. (29) to produce a representative value for the entire bin.

i=1,12 up ni i
Up = 
.
(29)
i=1,12 ni i
In Eq. (29), ni is the number of particle tracks for size fraction
i and i is the corresponding particle number ow rate per
particle track.

8. Results and discussions


8.1. Gas ow predictions and validation
Predicted and measured proles of mean longitudinal gas
velocities are compared in Figs. 5 and 6. Overall, the numerical
prediction provides a good representation of the measured ow
eld except in the outer-wall region (r < 0.4) within the bend.
The laboratory data indicates a growing layer of slow gas stream
next to the outer duct wall. This is presumably due to an adverse

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Fig. 7. Mean longitudinal gas velocity proles along the outer wall, 1 mm
from the surface.

pressure gradient on the concave face (outer wall) within the


rst half of the bend. In the second half of the bend, the gas
at the outer wall is found to accelerate towards the bend exit.
The gas ow near the inner wall, however, displays an opposite
trend: it accelerates steadily between = 0 and 45 , and then
slows down abruptly from = 45 . This is consistent with the
ndings of a similar experiment by Sudo et al. (2001) which
attributed this behaviour to changes in local pressure gradients
at the convex face as well as uid transport by the secondary
ow.
The DRSM adopted is unable to correctly resolve such a
pressurevelocity interaction and hence over-predicts the mean
longitudinal gas velocity by as much as 18% at = 60 in
the outer-wall region within the bend. This is shown in Fig. 7
which compares the predicted and measured near-wall longitudinal gas velocities at 1 mm above the outer wall. This possibly
points to a deciency in the pressurestrain correlation model
of Speziale et al. (1991) or even the modelled turbulent dissipation rate transport equation, Eq. (10), for resolving streamline
curvature effects as discussed in Gibson et al. (1981) and Luo
and Lakshminarayana (1996). In the inner-wall region, however, the DRSM captures the ow acceleration and deceleration
process satisfactorily.
Between = 90 and s/D = 0.5, the numerical prediction
indicates the formation of a recirculation zone at the inner wall
(Figs. 5 and 6). Although the measured proles at these two
stations provide no direct evidence as to the existence of this
recirculation zone, it is highly probable that ow reversal might
have taken place over a very short distance between =90 and
s/D = 0.5 in the physical experiment as a result of an adverse
pressure gradient.
Downstream of the bend, the maximum peak in the measured
prole has moved from r = 0.36 at = 90 to r = 0.3 at
s/D = 0.5, indicating a movement of the core gas ow towards
the outer wall under the inuence of centrifugal effect. At s/D=
1.0, a smaller peak in the measured velocity prole at r = 0.86
suggests that part of the core ow has been carried to the innerwall region (Fig. 6). According to Sudo et al. (2001), this is
attributable to the existence of a secondary ow which carries
the core gas ow from the outer wall to the inner wall along

2077

the duct circumference. The DRSM is found to reproduce the


core ow movement correctly in the simulation.
Sudo et al. (2001) also found that the secondary ow improves the mixing between the fast and the slow gas streams
which, respectively, originate from the outer and the inner walls.
This helps the mean gas ow stabilise and develop towards a
fully developed structure downstream of s/D = 1.0. While the
current set of centre-plane data is insufcient for a direct validation of the ability of the numerical model to capture secondary
ow motion, one can still apply it to assess the predicted core
ow movement which strongly affects secondary ow motion
inside the duct. A comparison of transverse gas velocity proles
in Fig. 8 shows that the core of the ow is initially drawn to the
upper inner duct wall at the bend entrance as evidenced from a
positive transverse velocity at =0 . Further into the bend, negative transverse velocities imply that the core ow is gradually
moving towards the outer wall. As the core ow travels from
the inner wall to the outer wall, it entrains the surrounding uid
on either side of it towards the outer wall and then around the
side walls. A pair of counter-rotating vortices (i.e., secondary
motion) is thus formed inside the bend. The speed at which the
core ow approaches the outer wall is therefore a good indicator
of the strength of the induced secondary motion. As seen from
Fig. 8, the DRSM predicts an excessive movement of the core
ow towards the outer wall from as early as 60 . The same
trend persists downstream of the bend (Fig. 9) which indicates
an over-prediction of the transverse velocity and hence of the
secondary ow motion by up to 50%. One would thus expect
the secondary motion to disperse more slowly in the simulation
than in the experiment.
Figs. 10 and 11compare the measured and predicted turbulence intensity u u (normalised by the bulk velocity Ub )
inside the duct. On the outer wall, the data indicates a rise in
near-wall turbulence intensity between =0 and 30 , followed
by a drop to 0.1 at s/D = 0.5. By contrast, the turbulence intensity at the inner wall rises considerably in the second half
of the bend to a maximum of 1.5 at s/D = 0.5. This is directly
due to a strong reduction of the local gas ow in the streamwise
direction. The numerical solution provides a qualitative representation of these trends. Flow uctuations as represented by
the turbulence intensities, however, have been severely underpredicted in these regions. This is mainly due to the application
of the eddy dissipation transport equation (10), which ignores
the contribution arising from the extra strain associated with
streamline curvature.
Turbulence intensity falls rapidly after s/d = 0.5 and gradually recovers back to the same level as that at the bend entrance,
buy symmetry in the turbulence structure is not completely
restored even at s/D = 9.0.
8.2. Validation of the predicted particle tracks
Based on the predicted gas ow eld, Lagrangian particle
tracking was rst performed without the implementation of
the SSL force and PWI models. The calculation utilised a
uniform restitution coefcient E = 0.73. The predicted mean

2078

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Fig. 8. Mean transverse gas velocity proles within the bend ( prediction;  data).

particle velocity proles were obtained from the data collection process as outlined previously and they are compared
against the measured proles in Figs. 12 and 13. This case
will be referred to as the baseline simulation in the following
discussions.
Compared to the measured proles, the baseline simulation
has over-predicted longitudinal particle velocities at the outer
wall within the bend (Fig. 12) and up to 3D downstream of the
bend (Fig. 13). This is partly attributable to an over-prediction

of the near-wall gas velocities (see Figs. 5 and 6). Particle-wall


collisions also strongly affect particle velocity prediction in
this area and this is illustrated in Fig. 14 which plots predicted
particle velocities (un-averaged) for three particle sizes at =
75 . One can see from the gure that the predicted velocity
proles for ne (5 m) and coarse (> 76 m) particles display
distinctively different characteristics. The ne particles show a
strong coupling with the gas phase while the coarse particles
are moving much more slowly within r < 0.4 of the outer wall.

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

2079

Fig. 9. Mean transverse gas velocity proles in the vertical duct ( prediction;  data).

This has led us to believe that the coarse particles have already
separated from the carrier uid and they are moving under the
inuence of their own inertia as well as particle-wall collisions
in the simulation. This subsequently implies that particle size
distribution also affects the prediction of particle velocity in the
outer wall region of the duct bend.
Away from the outer wall, the predicted values follow the
measured velocity proles with a maximum error of 20%

towards the inner wall, except at = 75 where large discrepancies between the measured and predicted particle velocities
are found at r > 0.8 (Fig. 12). With reference to Fig. 14, this
arises from a small group of ne particles possessing negative
longitudinal velocities and is a direct result of a recirculation
zone in the predicted gas ow eld at = 90 (see Fig. 5).
The ne particles that have been entrained by the recirculating gas ow at = 90 are thrown back towards = 75 .

2080

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088


Fig. 10. Turbulence intensity ( u u /Ub ) proles within the bend ( prediction;  data).

Velocity statistics pertaining to particles travelling upstream


thus cancel out those travelling downstream in the statistical
averaging process. Hence, one can conclude that the predicted
particle motion in the inner-wall region of the duct bend is
very sensitive to the background gas ow eld.
Mean transverse velocity proles for the particles are presented in Figs. 15 and 16. The baseline calculation is able to

correctly capture the transverse particle motion on the vertical centre-plane of the duct bend (Fig. 15). Both the experimental measurement and the calculation suggest a thin layer
of particles having very small transverse velocity components
at r < 0.2 between = 45 and 75 . This phenomenon arises
as some of the particles are moving towards the outer wall and
some are moving (rebounding) away from the wall within the

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

2081


Fig. 11. Turbulence intensity ( u u /Ub ) proles in the vertical duct ( prediction;  data).

LDA probe volume. As a result, the two types of motion cancel


each other out during the averaging process. This observation
is consistent with the ndings of a number of published studies
on solid saltation, e.g. Tanire et al. (1997) and Ciccone et al.
(1990).
In the vertical duct, the particle tracks tend to possess a larger
transverse velocity component than that in the physical ow

(Fig. 16). This directly relates to a considerable discrepancy


between the measured and predicted gas-phase velocities as
seen in Fig. 9. Hence, one should expect the predicted transverse
velocity prole to approach the measured distribution as the
gas-phase solution improves.
Following the baseline calculation, Lagrangian particle
tracking was repeated with the inclusion of SSL force and

2082

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Fig. 12. Mean longitudinal particle velocity proles within the bend ( baseline prediction;  prediction considering SSL and PWI models;  data).

sliding/non-sliding PWI in the numerical model. Figs. 12,


13, 15 and 16 indicate that they have no major effect on the
predicted mean particle velocities. However, they do contribute
considerably to the distribution of particles in the duct as illustrated in Fig. 17. The gure plots the normalised distribution
of particle tracks in the inner- and outer-wall regions along the
duct, and it is based on the number of particle tracks passing
through a 5.0 mm layer next to the duct walls. At each station,

particle tracks in the near-wall layer are counted and then


the sum is normalised by the total number of particle tracks
passing through the symmetry axis of that station. According
to the gure, the application of SSL force and PWI models
has helped distribute a larger portion of the particle tracks towards the outer duct wall within the bend as compared to the
baseline result. In this region, the baseline model only allows
the particles to rebound away from the wall after particle-wall

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

2083

Fig. 13. Mean longitudinal particle velocity proles in the vertical duct ( baseline prediction;  prediction considering SSL and PWI models;  data).

collisions. By comparison, the PWI model offers the particles


a second option which is to slide along the wall if their angle
of incidence is below a critical limit. This directly contributes
to a stronger presence of particle tracks at the outer wall of the
duct bend.
In the vertical duct, slip velocities (i.e., uf up ) are small
but positive for a large majority of the particle tracks in the
inner-wall layer. The SSL force, Eq. (20), thus distributes more

particle tracks away from the inner wall, resulting in a weaker


presence of particle tracks at the duct inner wall between s/D=
3.0 and 9.0 as compared to the baseline case.
Streamwise variation of the mean longitudinal velocity for
the particles within a 1.5 mm layer next to the outer wall is
shown in Fig. 18. In the early part of the bend, the data show
a rather rapid decrease in particle velocity as the majority
of the particles collide with the outer wall. The wall particle

2084

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Fig. 14. Predicted particle velocities by particle size (un-averaged) at = 75 .

Fig. 15. Mean transverse particle velocity proles within the bend ( baseline prediction;  prediction considering SSL and PWI models;  data).

velocity then gradually recovers to about 0.56Ub by 90 with


the help of the gas ow entraining and resuspending the
rebounded particles.

Compared to the measured prole, our predictions based


on the SSL and particle-wall collision models are only able
to quantitatively capture the particle deceleration/acceleration

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

2085

Fig. 16. Mean transverse particle velocity proles in the vertical duct ( baseline prediction;  prediction considering SSL and PWI models;  data).

process inside the bend. Universally applying a lower restitution coefcient to replace the particle-wall collision model
did not cause a further reduction in particle velocity between
= 15 and 45 where particle-wall collisions are expected to
be the most frequent. Therefore, particle-wall collision is not
the dominating mechanism through which the particles lose
their momentum.
Although we are not able to directly verify the location and
size of particle rope in the duct with the measured data, a higher
concentration of particle tracks in the outer wall layer between
= 0 and s/D = 1.0 (Fig. 17) does suggest the presence of
a particle rope inside the duct. It is thus very likely that the

particle lose much of their momentum as a result of roping.


Within the rope, the particles are strongly affected by interparticle collisions. One can therefore further improve particle
velocity predictions in this region by considering inter-particle
collisions in the numerical simulation.
9. Conclusion
Numerical simulations have been performed for a curved 90
duct bend. The gas-phase simulation was based on a DRSM
approach and the solids ow was solved by Lagrangian particle tracking. Our particle tracking calculations involved 96,000

2086

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Fig. 17. Normalised distribution of particle tracks within a 5.0 mm layer next to the walls ( baseline prediction;  prediction considering SSL and PWI
models).

Fig. 18. Streamwise variation of mean longitudinal particle velocities along the outer wall, 1.5 mm from the surface.

particle tracks and considered a range of hydrodynamic forces


and a particle-wall interaction (PWI) model, but ignored the
effects of particle rotation, turbulence modulation and interparticle collisions. The CFD solutions have then been validated
against the experimental data in an effort to identify areas where
further work is necessary to improve the accuracy of numerical
prediction for ows with dilute solid suspension. The present
study demonstrated the importance of a range of issues that one
needs to consider when performing a numerical simulation of
gassolid ow inside a power station mill duct, including:
Turbulent gas ow solution based on the current DRSM can
provide a reasonable representation of the mean gas ow at
the duct vertical centre-plane, except in regions next to the
concave wall and more than 3D downstream of the bend. The
former is a deciency, common to all differential Reynolds
stress models, which has not yet been fully resolved by the
uids modelling community. The turbulence model has considerably under-predicted the decay of secondary motion in
the vertical duct.
In the inner-wall region and also more than 3D downstream
of the bend, the predicted particle motion critically depends
on the quality of the gas ow solution. Near the outer wall
where the particles are more likely to collide with the wall,

PWI model can also considerably affect the prediction of


particle velocities and distribution of particles near wall.
Implementation of the shear-slip lift (SSL) force model
did not directly contribute to a more accurate prediction of
solids motion. However, it acts to redistribute particle tracks
across the ow depending on local gas velocity gradients
and slip velocity, particularly near the inner duct wall downstream of the bend. This makes it an important hydrodynamic force component to consider in shear or curved ow
calculations.
The PWI model adopted in this study does not allow nearwall particles to lose as much momentum as the data suggests
at the outer wall, though it was able to reproduce a similar
trend. We thus expect the particles to attain a much lower
near-wall velocity through mechanisms other than PWI in
the physical ow.
Notation
C
CD
dp
D

model constants
drag coefcient
particle diameter, m
hydraulic diameter of the duct, m

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

E
En
Et
F
Fr
fD
G
g
k
L
m
m

Ni
n
R
Re
Res
S
St
s, r

u
u
u , v 
U
U, V
Ub
UT

restitution coefcient
restitution coefcient normal to a surface
restitution coefcient tangential to a surface
force vector, N
Froude number
SchillerNaumann drag correlation
turbulence production, kg m2 s2
gravity vector, m s2
turbulence kinetic energy, m2 s2
solids mass loading = m
p /m
f
mass, kg
mass ow rate, kg s1
number of particle tracks allocated to size fraction i
number of particle tracks
duct turning radius
duct Reynolds number
particle Reynolds number based on uid rotation
inter-particle distance, m
particle Stokes number
curvilinear coordinate system on the duct
plane of symmetry; r = 0 at
outer wall; r = 1 at inner wall, m
instantaneous velocity vector, m s1
uctuating velocity vector, m s1
uctuating longitudinal andtransverse
velocity components, m s1
mean gas velocity vector, m s1
mean longitudinal and transverse
velocity components, m s1
bulk gas velocity, m s1
particle terminal velocity, m s1

Greek letters
p
1
ij




0
d





f

particle volume fraction


smooth wall incidence angle, deg
Kronecker delta
standard deviation of the wall
roughness angle;  = 3.8 for a
plexiglass plate, deg
eddy dissipation rate, m2 s3
duct turning angle, deg
gas dynamic viscosity, kg m1 s1
static friction coefcient
dynamic friction coefcient
gas kinematic viscosity, m2 s1
Gaussian random number with zero
mean andstandard deviation of 1
density, kg m3
time scale, s
particle rate, s1
uid rotation vector,s1

Subscripts
A
D

added mass
drag

f
g
i
i, j, k
p
pg
sl
1
2

2087

uid
gravity
size fraction index
tensor index
partic
pressure gradient
slip-shear lift
pre-impact state
post-impact state

Acknowledgements
The authors gratefully acknowledge the nancial and other
support received for this research from the Cooperative Research Centre (CRC) for Clean Power from Lignite, which is
established and supported under the Australian Governments
Cooperative Research Centres program.
References
Akilli, H., Levy, E.K., Sahin, B., 2001. Gassolid ow behaviour in a
horizontal pipe after vertical-to-horizontal elbow. Powder Technology 116,
4352.
Barth, T.J., Jesperson, D.C., 1989. The design and application of upwind
schemes on unstructured meshes. AIAA Paper 89-0366.
Bilirgen, H., Levy, E.K., 2001. Mixing and dispersion of particle ropes in
lean phase pneumatic conveying. Powder Technology 119, 134152.
Boothroyd, R.G., 1971. Flowing Gassolids Suspensions. Chapman & Hall,
London.
Borger, G.G., 1973. The optimisation of wind tunnel contractions for the
subsonic range. Ph.D. Thesis, Ruhr University, Germany.
Choi, Y.D., Iacovides, H., Launder, B.E., 1989. Numerical computation of
turbulent ow in a square-sectioned 180 deg bend. Journal of Fluids
Engineering 111, 5968.
Ciccone, A.D., Kawall, J.G., Keffer, J.F., 1990. Flow visualization/digital
image analysis of saltating particle motion. Experiments in Fluids 9,
6573.
Fan, L.S., Zhu, C., 1998. Principles of GasSolid Flows. Cambridge University
Press, Cambridge, pp. 87110.
Fan, J., Zhang, X., Cheng, L., Cen, K., 1997. Numerical simulation and
experimental study of two-phase ow in a vertical pipe. Aerosol Science
and Technology 27, 281292.
Frank, TH., Schade, F.-P., Petrak, D., 1993. Numerical simulation and
experimental investigation of a gassolid two-phase ow in a horizontal
channel. International Journal of Multiphase Flow 19 (1), 187198.
Gibson, M.M., 1978. An algebraic stress and heat-ux model for turbulent
shear ow with streamline curvature. International Journal of Heat and
Mass Transfer 21, 16091617.
Gibson, M.M., Jones, W.P., Younis, B.A., 1981. Calculation of turbulent
boundary layers on curved surfaces. Physics of Fluids 24 (3), 386395.
Giddings, D., Aroussi, A., Pickering, S.J., Mozaffari, E., 2004. A 1/4 scale test
facility for PF transport in power station pipelines. Fuel 83, 21952204.
Gosman, A.D., Ioannides, E., 1981. Aspects of computer simulation of liquidfuelled combustors. AIAA Paper No. 81-0323.
Huber, N., Sommereld, M., 1994. Characterization of the cross-sectional
particle concentration distribution in pneumatic conveying systems. Powder
Technology 79, 191210.
Kuan, B., 2005. CFD simulation of dilute gassolid two-phase ows with
different solid size distributions in a curved 90 duct bend. Australia
& New Zealand Industrial and Applied Mathematics Journal 46 (E),
C744C763.
Kussin, J., Sommerfeld, M., 2002. Experimental studies on particle behaviour
and turbulence modication in horizontal channel ow with different wall
roughness. Experiments in Fluids 33, 143159.

2088

B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 2088

Lan, S., Sommerfeld, M., Kussin, J., 2002. Experimental studies and
modelling of four-way coupling in particle-laden horizontal channel ow.
International Journal of Heat and Fluid Flow 23, 647656.
Levy, A., Mason, D.J., 1998. The effect of a bend on the particle cross-section
concentration and segregation in pneumatic conveying systems. Powder
Technology 98, 95103.
Luo, J., Lakshminarayana, B., 1996. Analysis of streamline curvature effects
on wall-bounded turbulent ows. Third International Symposium on
Engineering Turbulence Modelling and Measurements, Crete, Greece, May
1996.
Matsumoto, S., Saito, S., 1970. Monte Carlo simulation of horizontal
pneumatic conveying based on the rough wall model. Journal of Chemical
Engineering of Japan 3 (1), 223230.
McIntosh, M.J., Borthwick, I.R., 1984. Investigation of the grinding process
in an EVT millDevelopment of a process model, Report No. ND/84/003,
State Electricity Commission of Victoria, Australia.
Mei, R., 1992. An approximate expression for the shear lift force on a spherical
particle at nite Reynolds number. International Journal of Multiphase
Flow 18 (1), 145147.
Nallasamy, M., 1987. Turbulence models and their applications to the
prediction of internal ows: a review. Computers & Fluids 15 (2),
151194.
Patel, V.C., Sotiropoulos, F., 1997. Longitudinal curvature effects in turbulent
boundary layers. Progress in Aerospace Science 33, 170.
Richmond, M.C., Patel, V.C., 1991. Convex and concave surface curvature
effects in wall-bounded turbulent ows. AIAA Journal 29 (6), 895902.
Rodi, W., Scheuerer, G., 1983. Calculation of curved shear layers with twoequation turbulence models. Physics of Fluids 26, 1422.
Shih, T.H., Liou, W.W., Shabbir, A., Yang, Z., Zhu, J., 1995. A new k . eddy
viscosity model for high Reynolds number turbulent ows. Computers
Fluids 24 (3), 227238.

Sommerfeld, M., 2000. Theoretical and experimental modelling of particulate


ows. Lecture Series 2000-06, von Karman Institute for Fluid Dynamics.
Sommerfeld, M., Huber, N., 1999. Experimental analysis and modelling
of particle-wall collisions. International Journal of Multiphase Flow 25,
14571489.
Speziale, C.G., Sarkar, S., Gatski, T.B., 1991. Modelling the pressurestrain
correlation of turbulence: an invariant dynamical systems approach. Journal
of Fluid Mechanics 227, 245272.
Sudo, K., Sumida, M., Hibara, H., 2001. Experimental investigation on
turbulent ow in a square-sectioned 90-degree bend. Experiments in Fluids
30, 246252.
Tanire, A., Oestrle, B., Monnier, J.C., 1997. On the behaviour of solid
particles in a horizontal boundary layer with turbulence and saltation
effects. Experiments in Fluids 23, 463471.
Tsuji, Y., Morikawa, Y., 1982. LDV measurements of an airsolid two-phase
ow in a horizontal pipe. Journal of Fluid Mechanics 120, 385409.
Tsuji, Y., Morikawa, Y., Shiomi, H., 1984. LDV measurements of an airsolid
two-phase ow in a vertical pipe. Journal of Fluid Mechanics 139,
417434.
Uijttewall, W.S.J., Oliemans, R.V.A., 1996. Particle dispersion and deposition
in direct numerical and large eddy simulations of vertical pipe ows.
Physics of Fluids 8 (10), 25902604.
Yang, W., Kuan, B., 2006. Experimental investigation of dilute turbulent
particulate ow inside a curved 90 bend. Chemical Engineering Science
61, 35933601.
Yilmaz, A., Levy, E.K., 1998. Roping phenomena in pulverized coal conveying
lines. Powder Technology 95, 4348.
Yilmaz, A., Levy, E.K., 2001. Formation and dispersion of ropes in pneumatic
conveying. Powder Technology 114, 165185.

You might also like