Sunteți pe pagina 1din 21

Chemical Engineering Science 62 (2007) 2068 – 2088

www.elsevier.com/locate/ces

Dilute gas–solid two-phase flows 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 fluid dynamics (CFD) simulations of dilute gas–solid flow 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 flow quantities for Re = 100, 000 were calculated based on a
differential Reynolds stress model. The solids mass loading considered is 0.00206 and hence justifies 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 Laser–Doppler Anemometry (LDA). The study found that the predicted gas flow field has a strong influence
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 flow field, SSL force and particle-wall collisions
to Lagrangian particle tracking.
䉷 2007 Elsevier Ltd. All rights reserved.

Keywords: Multiphase flow; Mathematical modelling; Simulation; Penumatic conveying; LDA; Turbulence

1. Introduction as feed size and mill speed, the fuel pulverisation process pro-
duces coal particles ranging between 10 and 1000 m in size.
Elbows and bends are commonly used in pneumatic convey- Fig. 1 shows a typical coal particle size distribution measured
ing systems to change flow direction so as to transport the sus- at the mill outlet of a lignite-fired power station (McIntosh and
pended material to the desired delivery point within a limited Borthwick, 1984). The extreme non-uniformity in particle size
space. In the case of coal-fired power plants that operate combined with a centrifugal effect arising from the duct bend
on a continuous supply of pulverised coal to furnaces, mal- are believed to lead to the formation of a stratified gas–solid
distribution of pulverised fuel often occurs as coal particles flow, known as a particle rope, downstream of the elbow even
are pneumatically transported from the mill through ducts at a low solids mass loading, L < 0.1. This invariably creates
consisting of numerous bends and straight sections. difficulties for the plant operators to monitor and control the
Apart from the duct geometry, the coal pulverisation process pulverised fuel supply to individual burners, and hence to main-
is also a strong contributor to mal-distribution of the pulverised tain an optimal combustion condition inside the furnace.
fuel. Field measurements at a lignite-fired power plant carried There are a large number of documented studies, both nu-
out to support this study found that, depending on factors such merical and experimental, on particle roping in dust conveying
systems with solids mass loading L > 0.3, but most of them
∗ Corresponding author. Tel.: +61 3 95458687; fax: +61 3 95628919. focus on particles with a size distribution that is either heavily
E-mail address: benny.kuan@csiro.au (B. Kuan). skewed towards the lower end of the size range (Yilmaz and
0009-2509/$ - see front matter 䉷 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2006.12.054
B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088 2069

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


the mean particle velocity profile with increasing particle size.
Tsuji and Morikawa (1982) has also performed the same
LDV measurements for a horizontal two-phase flow, but only
for L > 2.0. They observed a stronger influence of the par-
ticles on the background gas turbulence as the particle size
increased. Laín et al. (2002) experimentally studied the effect of
particle size, solids mass loading and wall roughness on a two-
phase horizontal channel flow. While all tests were performed
at L > 0.2, their results indicate that the flow was increasingly
sensitive to particle size distribution as L decreases.
In pneumatic transport, it is commonly accepted that move-
Fig. 1. Typical coal size distribution at the mill outlet. ment of large particles is predominantly influenced by particle-
wall interactions rather than fluid turbulence. In recognition of
this, Frank et al. (1993), Kussin and Sommerfeld (2002), and
Levy, 1998; Akilli et al., 2001) or more evenly spread across Sommerfeld and Huber (1999) conducted series of experiments
the entire size range (Huber and Sommerfield, 1994). In both to study particle-wall interaction for a large combination of par-
cases, the tested particle samples contain a considerable amount ticle size, material and wall roughness in an attempt to create
of fine particles. A comprehensive experimental investigation a reliable physical particle-wall collision model.
was conducted by Huber and Sommerfield (1994) using phase- Compared to the flow regime prevailing at the power plant
Doppler anemometry (PDA) in various pipe elements. They mill duct (i.e., L < 0.1), all of the above-mentioned studies
used glass particles with a mean diameter of 40 m and their examined gas–solid flow behaviour in the L > 0.2 range and
results demonstrated the significance of wall-roughness in pre- hence their results were not representative of the dilute mill-duct
venting the glass beads from settling in a horizontal pipe. flow considered in the present study. Further, most of the pub-
The research group at Lehigh University has conducted sev- lished conveying duct experiments only consider the scenario
eral physical experiments ( Yilmaz and Levy, 1998, 2001; Akilli where a substantial amount of fine particles were present in the
et al., 2001; Bilirgen and KLevy, 2001) investigating forma- system. In the present study, a dispersed phase with a wider size
tion and dispersion of the particle ropes involving pulverised distribution but a smaller fraction of fine particles is considered
coal particles. A closed flow loop consisting of two 6.1 m hor- so as to better represent the gas–solid flow system prevailing in
izontal pipes, one 3.4 m vertical pipe and two 90◦ elbows with the mill-duct flow. Numerical simulation of such a flow using
adjustable turning radii was set up to facilitate measurements Lagrangian particle tracking technique requires a large set of
of particle rope characteristics in a vertical pipe following a particle trajectories to be solved. A special discretisation tech-
horizontal-to-vertical bend. All pipe elements had an internal nique has been developed to overcome the need for an exceed-
diameter of 0.154 m. They identified a number of parameters ingly large number of particle tracks. This paper aims to study,
that affect the behaviour of the particle rope downstream of both numerically and experimentally, the behaviour of a dilute
a horizontal-to-vertical duct bend. These included solids mass gas–solid flow which uses non-uniform sized spherical glass
loading, conveying velocity, and duct bend radius. beads as the dispersed phase and at a low solids mass loading
Levy and Mason (1998) reported a similar study utilising L = 0.00206. Glass spheres having a volume weighted mean
three different particle sizes. Their numerical simulations indi- diameter of 77 m were tested in purpose-designed laboratory
cated a change in particle roping characteristics with particle experiments which involve a square-sectioned horizontal-to-
size: coarse particles form a concentrated rope which propa- vertical 90◦ bend.
gates more than 2.4D downstream of the bend; for the fine Numerical simulations were performed for the gas phase and
particles, they only noticed a localised peak in particle con- then followed by the solution of solids motion. The solution of
centration just downstream of the bend. There were however particle motion is based on a Lagrangian particle tracking ap-
no published experimental data at the time to validate their proach where a single particle track represents the trajectories
calculations. of a group of particles of the same size. The calculation consid-
The effect of particle size on solid- as well as gas-phase ered various hydrodynamic forces acting on the particles, such
motions in vertical and horizontal two-phase flows has been as drag and shear-slip lift (SSL), as well as particle wall inter-
extensively studied in the literature. With the aid of a laser- actions. The hydrodynamic forces were determined based on
Doppler velocimeter (LDV), Tsuji et al. (1984) tested plastic the knowledge of the predicted gas flow. An accurate solution
particles with diameters ranging from 200 m to 3 mm in a of the gas flow field is thus crucial for a realistic prediction of
vertical pipe having an inner diameter of 30.5 mm and studied the particle motion.
the effect of particle size on particle motion and gas turbulence. One main feature of the square-sectioned duct flow consid-
At high solids mass loading (L > 2.0), they found the mean ered in this study is a strong curvature in the streamwise direc-
air velocity profile flattens across the duct cross-section with tion along the duct length. This necessitates the application of
decreasing particle size. Fan et al. (1997) performed a similar a turbulence model that is capable of resolving the extra strains
experiment with 100 and 300 m quartz particles in a 100 mm 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 of the converging nozzle follows Borger’s contractor profile
with streamline curvature effects. Nallasamy (1987), and Patel (Borger, 1973) to ensure flow uniformity within ±0.5%.
and Sotiropoulos (1997) have both provided comprehensive The duct geometry and flow coordinate system are shown
reviews on various approaches to model streamline curvature in Fig. 2(b). The square-sectioned (150 mm × 150 mm) test
effects in the turbulence model. section is constructed using 10 mm thick Perspex sheets, and the
Gibson (1978) and Rodi and Scheuerer (1983) have proposed main components of the test facility include a 3.5 m horizontal
a number of ad hoc modifications to the standard two-equation straight duct, a 90◦ bend with a turning radius R of 225 mm
turbulence model to address this problem. However, the appli- and a 1.8 m vertical straight duct. The R/D ratio for the bend
cability of these model corrections has largely been limited to is thus 1.5.
flows with simple geometry where axis of streamline curvature Glass spheres with a volume weighted mean diameter of
is fixed relative to the principal flow direction, such as flow 77 m were released into the gas flow from a fluidised-bed
over a back-ward step. Richmond and Patel (1991) have tested feeder at a rate of 2 kg h−1 to give a solids mass loading L of
one of the above models in various curved wall-bounded flows. 0.00206. A digital balance underneath the fluidized-bed feeder
The model produced satisfactory mean velocity prediction on monitors the rate at which the particles are released so as to
the convex wall while under-estimating the mean velocities on ensure a dilute gas–solid flow regime inside the test section.
the concave wall. Data on particle size distribution (based on volume fraction)
Choi et al. (1989) have applied an algebraic Reynolds stress are obtained from a wet analysis using a Malvern particle size
model in the core flow and mixing-length hypothesis in the analyser and are shown in Fig. 3.
viscous sublayer in an attempt resolve streamline curvature in Gas phase measurements were obtained at a bulk gas veloc-
a square-sectioned U-bend. The model performs better than the ity Ub , of 10 m s−1 in the absence of the glass spheres. A mist
standard k. model and provides a reasonable representation of fine sugar particles generated from a jet atomiser using 5%
of the primary flow but it tends to over-predict gas velocity at sugar solution have a mean diameter of 1 m and were used as
the convex wall and over-predict it at the concave wall. seeding particles in the single-phase measurement. Turbulence
Luo and Lakshminarayana (1996) have applied a differential intensity in the main stream was approximately 1% at the cen-
Reynolds stress model and a modified eddy dissipation rate  tre of duct cross-section 10D upstream of the 90◦ bend. The
transport equation proposed by Shih et al. (1995) to account Reynolds number, based on the bulk velocity, hydraulic diam-
for streamline curvature effects. They tested their model in 2D eter of the square-sectioned duct and kinematic viscosity of air
90◦ and 180◦ pipe bends. The results indicated a negligible was subsequently 105 .
improvement on the predicted mean gas velocities as compared Flow measurements were performed on the duct vertical
to the case with unmodified differential Reynolds stress model. symmetry plane. At each streamwise location along the duct
However, the new model produced a higher level of turbulence length, flow measurements, including mean gas and solids
intensity at two stations inside and shortly downstream of the velocities, and turbulence intensities were carried out. In near-
bend. wall areas, flow statistics were measured at 1 mm from the
Patel and Sotiropoulos (1997) has thus concluded that, for duct walls. LDA measurements outside the vertical symmetry
flows containing streamline curvature, a differential Reynolds plane were not taken because of poor access of the laser beam
stress model without any modifications should provide rea- to the probe locations. Secondary gas flow, which is charac-
sonable flow prediction as compared to that based on case- terised by a pair of counter-rotating vortices along the duct
specific two-equation turbulence models. The current study length downstream of the bend, was not studied in the present
adopts the differential Reynolds stress model of Speziale et al. work. It is assumed to display similar characteristics to the
(1991) which is tensor-invariant and does not require the addi- laboratory flow of Sudo et al. (2001) where R/D = 2.0.
tion of wall-reflection terms into its pressure–strain correlation
model. 2.2. Dynamic similarity

2. Experimental test facility In order to ensure the flow scenario tested in the laboratory
model reflects the physical duct operating condition, a dynamic
2.1. Rig configuration similarity analysis was performed prior to the experiment. Our
analysis is based on the collective experience of Boothroyd
All experiments were conducted in an open-circuit (1971) and Fan and Zhu (1998). In general they found the flow
horizontal-to-vertical suction wind tunnel system which was systems that share similar geometric and kinematic boundary
set up for an earlier study by Yang and Kuan (2006). A conditions display similar dynamic responses when the dimen-
schematic sketch of the entire test facility is provided in sionless numbers Re, Rep , St, and Fr are the same
Fig. 2(a). It consists of an open-circuit suction wind tunnel
where the airflow is drawn into the system by means of a cen-  f Ub D
Re = , (1)
trifugal fan through an entry piece that consists of an elliptical 
bell-mouth inlet and a honeycomb flow straightening section.
The air then passes through a converging nozzle to attain a  f UT d p
Rep = , (2)
higher velocity before entering into the test section. Design 
B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088 2071

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

whereby Re denotes duct Reynolds number; Rep is particle


Reynolds number; St is particle Stokes number as commonly
defined in the literature (Giddings et al., 2004); Fr is Froude
number for gas–solid flows; UT denotes particle terminal
velocity; Ub is fluid bulk velocity based on fluid mass flow
rate and duct cross-section area.
Stokes number St is a ratio of particle response time to the
fluid 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 airflow.
Dynamic similarity parameters are compared in Table 1. The
table also lists some of the characteristic flow parameters for
both cases. The smaller duct diameter and particle sizes in the
Fig. 3. Particle size distributions by volume (— wet analysis data;  modelled
laboratory system as compared to the industrial flow are largely
distribution).
responsible for the lower Re, Rep , and St values at the upper
end of the particle size range. A lower Rep implies that the par-
p dp2 ticles in the laboratory flow system are more tightly coupled to
p D
St = ; p = ; f = , (3) the gas flow than in the industrial counterpart. The power plant
f 18 Ub mill-duct flow contains a greater proportion of coarse particles
 with a very high St, indicating a higher chance of particle-wall
Ub2 collisions. However, motion of the larger particles is beyond the
Fr = , (4)
Dg 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 Power station

Flow parameters
D (m) 0.15 1.76
 (kg m−1 s−1 ) 1.8 × 10−5 1.95 × 10−5
f (kg m−3 ) 1.18 0.78
p (kg m−3 ) 2500 1400
L 0.00206 0.1
dp (m) 4 77 160 20 80 410
(min.) (mean) (max.) (5%) (50%) (95%)
UT (m s−1 ) 1.21 × 10−3 0.36 1.1 0.0155 0.224 2.32
Ub (m s−1 ) 10.0 28.6
Re 1.0 × 105 2.0 × 106

Dynamic similarity parameters


Rep 3.17 × 10−4 1.82 11.5 0.0124 0.716 38.1
St 8.23 × 10−4 3.05 13.2 0.063 1.701 44.7
Fr 8.24 6.88
S/dp 80.9 21.1

driven by particle’s own inertia as well as collisions with the notations below
wall surfaces, instead of gas–solids interactions.
As seen in Table 1, the current experiment covers the lower jUi
= 0, (7)
end of the industrial particle size range which has lower Rep dxi
and St values. One can thus expect the laboratory rig to realis-    
jUi Uj jP j jUi jUj
tically reproduce the characteristics of the power station mill- =− +  + − ui uj , (8)
duct flow in the absence of very coarse particles. dxj jxi jxj jxj jxi
Apart from dynamic similarity parameters that charac-  

terise the level of particle–gas interactions in both systems, jUk ui uj j 2 k2 jui uj
Table 1 also lists inter-particle spacing S (with respect to dp ) as =  + Cs
dxk jxk 3  jxl
estimated from
  2
 1/3 + Gij − ij + ij , (9)
S/dp = , (5) 3
6p   
jUk  j t j  2
1 = + + C1 Gkk − C2 , (10)
p =  (6) dxk jxk
 jxl k k
p
+1
Lf jUj jUi
Gij = ui uk − uj uk , (11)
jxk jxk
which have been used by Sommerfeld (2000) to classify the
importance of interaction mechanisms in dispersed two-phase
where Ui and ui are, respectively, mean and fluctuating veloc-
flows. In general, effect of inter-particle collisions can be
ignored for an inter-particle space, S/dp greater than 10. ities; ui uj denotes the Reynolds stress tensor; Gij is the tur-
According to Table 1, both flow systems satisfy this condition bulence production term; ij is the modelled pressure–strain
and hence inter-particle collisions were not considered in this correlation given by Speziale et al. (1991). Model constants Cs ,
study. 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. Mathematical models

3.1. Gas flow 4. Particle tracks

Steady-state, isothermal gas flow properties and turbu- 4.1. Hydrodynamic forces
lence quantities are calculated numerically by solving a set
of governing partial differential equations (PDE) using a Instantaneous positions and velocities of the dispersed phase
commercial CFD software ANSYS CFX-10. The Reynolds- are solved through a Lagrangian particle tracking method. Mo-
averaged Navier–Stokes equations and the Reynolds stress tion of individual particles suspended in a continuous fluid is
transport equations considered are written in Cartesian tensor determined by numerically integrating the equations of motion
B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088 2073

for the dispersed phase in a fluid flow. The equation of particle f dp2 |f |
motion may be expressed as Res = , (23)

dup Res
mp = FD + Fg + Fpg + FA + Fsl , (12) = 0.5 ; f = ∇ × uf . (24)
dt Rep
dxp Assuming that all force components, except the drag, are con-
= up , (13)
dt stant during a time step t, Eq. (12) can be integrated analyti-
where subscript p represents particle properties and subscripts cally to yield
D, g, pg, A and sl, respectively, denote force components arising r
up = uf + (up0 − uf )e−t/r + (1 − e−t/r )
from drag, gravity, flow pressure gradient, added mass effect mp
and slip-shear lift. × (Fg + FA + Fsl + Fpg ) (25)
The drag force is calculated from
with superscript 0 indicating the start of a time step.
uf − u p Similarly, Eq. (13) can be integrated analytically to
F D = mp (14)
r
xp = xp0 + up t. (26)
with uf = U + u and particle relaxation time r defined by
Eqs. (25) and (26) were solved within a given cell in the particle
p dp2 tracking calculations.
r = (15) In order to solve the instantaneous particle velocity and
18fD
location using the integrated equations of particle motion,
and the Schiller–Naumann drag correlation fD for a sphere Eqs. (25) and (26), for every particle track in the flow domain,
the instantaneous fluid velocity has to be specified at all par-
1 + 0.15Re0.687
p , Rep 1000, ticle locations. This is made possible through the application
fD = (16)
0.01833Rep , Rep > 1000, of a classical stochastic approach of Gosman and Ioannides
(1981), which estimates fluctuating gas velocity components
24 on the basis of isotropic turbulence. Subsequent particle track
CD = fD .
Rep integration which takes into account the turbulent dispersion
effect was thus carried out.
As Rep increases beyond 1000, the corresponding fD leads to
a constant drag coefficient CD of 0.44. 5. Particle-wall interactions
The force components due to gravity Fg , added mass FA and
pressure gradient Fpg are, respectively, given by The present simulation adopts a modified version of the PWI

model of Matsumoto and Saito (1970). The base-case model
f allows the particles to either slide along the wall surface when
Fg = m p 1 − g, (17)
p the angle of incidence is small, or rebound away from the wall
after impact. However, it is based on the assumption of a con-
1 f dup stant restitution coefficient and dynamic friction, both of which
FA = − m p , (18)
2 p dt are sensitive to a range of parameters, such as incidence angle
and wall material, as found in published experimental investi-
Fpg = − 41 dp3 ∇P . (19) gations of Frank et al. (1993), Sommerfeld and Huber (1999).
For the present numerical calculation that involves collisions
The present calculation adopts a SSL model formulated by Mei between the glass spheres and the Perspex duct walls, particle
(1992). In vector form, the model is written as velocity components as well as particle angle of incidence are
defined in Fig. 4. Impact test data for glass beads on Plexiglass
 3
Fsl = d  Csl ((uf − up ) × f ), (20)
8 p f
where
4.1126
Csl = √ f (Rep , Res ), (21)
Res

⎨ (1 − 0.3314 )
1/2

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



0.0524( Rep )1/2 , Rep 40, Fig. 4. Definition of velocities before and after impact and particle angle of
(22) incidence.
2074 B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088

plates are utilized to characterize the PWI (Sommerfeld and All calculations were performed on a 80 × 80 × 135 (y × z × s)
Huber, 1999). The modified PWI model is given by body-fitted grid and the computed flow domain was constructed
using hexahedral cells. The grid has a minimum wall spacing
up2 = Et up1 ,
of y + < 16 and was found to be sufficiently refined to produce
vp2 = −En vp1 , a grid-independent solution as shown in an earlier study by
Kuan (2005).
En = max(1.0 − 0.0151 , 0.73),
The current simulations utilise a set of fully turbulent inflow
d = max(0.4 − 0.009261 , 0.15) conditions generated from a separate calculation in a straight
horizontal pipe of the same cross-section. The solid particles
ucrit = 3.5d (1 + En )|vp1 | were released into the duct at the same velocity as the gas


⎪ 1.0 − d (1 + En )|vp1 /up1 |, phase, i.e., zero slip, and from random locations on the inlet

⎪ plane.
⎨ up1 ucrit (sliding collision),
Et = (27) In a dilute two-phase flow where particle volume fraction is

⎪ 5/7, less than 10−6 (i.e., L < 0.00211), it is generally accepted that


⎩ the transfer of particle momentum to the carrier-phase is negli-
up1 < ucrit (non-sliding collision).
gible (Sommerfeld, 2000). Thus, at L=0.00206, the gas motion
The effect of wall roughness is introduced through a semi- could be considered independent of the solid-phase (i.e., one-
empirical approach that modifies the “smooth-wall” incident way coupling). When particles are introduced into a turbulent
angle 1 with a random component characterising the presence gas flow, they can either enhance or reduce the gas turbulence.
of a rough wall This is known as turbulence modulation which is assumed to
1 = 1 +  , (28) be insignificant in our simulation.
In the general dilute gas–solid flows, the quality of the
where  represents a random component sampled from a dispersed-phase prediction based on Lagrangian particle track-
Gaussian distribution function. is a Gaussian random number ing is directly dependent on the number of particle tracks solved
with zero mean and standard deviation of unity, and  is the in the simulation. In cases where the particle size distribution
standard deviation of the wall roughness angle.  is approxi- is wide, one has to proportionally increase the total number
mately 3.8◦ for a Plexiglass plate. of the considered particle tracks so as to adequately resolve
In our simulations, we have assumed the particles are non- the smallest size fraction. This often leads to an exceedingly
rotating and the effect of particle rotation on particle velocity large set of particle tracks which are very computationally
after the collision is negligible. demanding to solve and analyse.
In the present study, we have developed a methodology
6. Numerical procedure which allows one to overcome this limitation. The particle size
distribution (by volume) as measured in a wet analysis by the
The PDE (8)–(10), are discretised following a finite vol- Malvern size analyser (Fig. 3) was first discretised into 12 char-
ume approach. The advection terms were approximated using acteristic particle size fractions (Table 2). One can calculate
a scheme developed by Barth and Jesperson (1989) which the associated particle volume flow rate for each size fraction
is more than first-order accurate at mesh discontinuities and to give a total solids volume flow rate of 2.207E − 7 m3 s−1
provides higher-order accuracy for smoothly varying meshes. such that L = 0.00206. A fixed number of particle tracks were

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

i dp,i (m) Vol.% Particle vol. flow Ni Particle vol. flow i (1/s)
rate (m3 /s) rate per track (m3 /s)

1 5 2.001E − 5 4.416E − 14 8000 5.520E − 18 8.434E − 2


2 18 5.202E − 5 1.148E − 13 8000 1.435E − 17 4.700E − 3
3 30 3.112E − 3 6.868E − 12 8000 8.585E − 16 6.073E − 2
4 45 2.905E − 1 6.411E − 10 8000 8.014E − 14 1.680
5 57 3.260 7.196E − 9 8000 8.994E − 13 9.276
6 65 10.35 2.284E − 8 8000 2.855E − 12 19.85
7 76 38.10 8.409E − 8 8000 1.051E − 11 45.73
8 89 34.03 7.510E − 8 8000 9.387E − 12 25.43
9 103 9.982 2.203E − 8 8000 2.754E − 12 4.813
10 125 2.500 5.518E − 9 8000 6.897E − 13 6.744E − 1
11 140 1.000 2.207E − 9 8000 2.759E − 13 1.920E − 1
12 152 4.887E − 1 1.079E − 9 8000 1.348E − 13 7.332E − 2

Total 100.0 2.207E − 7 96 000


B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088 2075

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

then allocated to each size fraction and we have used a total 7. Collection and averaging of predicted particle statistics
of 96,000 particle tracks in the present study. Individual par-
ticle tracks in each size fraction thus carry a uniform portion Similar to the data acquisition process performed during the
of the solids volume flow rate as shown in Table 2. One physical experiment, the predicted particle statistics are col-
can then determine a particle number flow rate i repre- lected at each measurement station on the duct centre-plane.
sented by each track and subsequently apply this to convert We have adopted the technique by Uijttewall and Oliemans
number of particle tracks into number of “real” particles dur- (1996) for estimating the mean solids flow quantities on the
ing the statistical averaging process as discussed in the next measurement plane. This involves setting up 27 equally spaced
section. “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 profiles in the vertical duct (— prediction;  data).

information. The particle statistics stored inside each bin are 8. Results and discussions
then averaged based on Table 2 and Eq. (29) to produce a rep-
resentative value for the entire bin. 8.1. Gas flow predictions and validation

i=1,12 up ni i Predicted and measured profiles of mean longitudinal gas
Up =  . (29)
i=1,12 ni i
velocities are compared in Figs. 5 and 6. Overall, the numerical
prediction provides a good representation of the measured flow
In Eq. (29), ni is the number of particle tracks for size fraction field except in the outer-wall region (r ∗ < 0.4) within the bend.
i and i is the corresponding particle number flow rate per The laboratory data indicates a growing layer of slow gas stream
particle track. next to the outer duct wall. This is presumably due to an adverse
B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088 2077

the duct circumference. The DRSM is found to reproduce the


core flow movement correctly in the simulation.
Sudo et al. (2001) also found that the secondary flow im-
proves 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 flow stabilise and develop towards a
fully developed structure downstream of s/D = 1.0. While the
current set of centre-plane data is insufficient for a direct valida-
tion of the ability of the numerical model to capture secondary
flow motion, one can still apply it to assess the predicted core
flow movement which strongly affects secondary flow motion
inside the duct. A comparison of transverse gas velocity profiles
in Fig. 8 shows that the core of the flow is initially drawn to the
Fig. 7. Mean longitudinal gas velocity profiles along the outer wall, 1 mm upper inner duct wall at the bend entrance as evidenced from a
from the surface.
positive transverse velocity at =0◦ . Further into the bend, neg-
ative transverse velocities imply that the core flow is gradually
moving towards the outer wall. As the core flow travels from
pressure gradient on the concave face (outer wall) within the the inner wall to the outer wall, it entrains the surrounding fluid
first half of the bend. In the second half of the bend, the gas on either side of it towards the outer wall and then around the
at the outer wall is found to accelerate towards the bend exit. side walls. A pair of counter-rotating vortices (i.e., secondary
The gas flow near the inner wall, however, displays an opposite motion) is thus formed inside the bend. The speed at which the
trend: it accelerates steadily between = 0◦ and 45◦ , and then core flow approaches the outer wall is therefore a good indicator
slows down abruptly from = 45◦ . This is consistent with the of the strength of the induced secondary motion. As seen from
findings of a similar experiment by Sudo et al. (2001) which Fig. 8, the DRSM predicts an excessive movement of the core
attributed this behaviour to changes in local pressure gradients flow towards the outer wall from as early as 60◦ . The same
at the convex face as well as fluid transport by the secondary trend persists downstream of the bend (Fig. 9) which indicates
flow. an over-prediction of the transverse velocity and hence of the
The DRSM adopted is unable to correctly resolve such a secondary flow motion by up to 50%. One would thus expect
pressure–velocity interaction and hence over-predicts the mean the secondary motion to disperse more slowly in the simulation
longitudinal gas velocity by as much as 18% at = 60◦ in than in the experiment.
the outer-wall region within the bend. This is shown in Fig. 7 Figs. 10 and 11compare the measured and predicted tur-
which compares the predicted and measured near-wall longitu- bulence intensity u u (normalised by the bulk velocity Ub )
dinal gas velocities at 1 mm above the outer wall. This possibly inside the duct. On the outer wall, the data indicates a rise in
points to a deficiency in the pressure–strain correlation model near-wall turbulence intensity between =0◦ and 30◦ , followed
of Speziale et al. (1991) or even the modelled turbulent dissipa- by a drop to 0.1 at s/D = 0.5. By contrast, the turbulence in-
tion rate transport equation, Eq. (10), for resolving streamline tensity at the inner wall rises considerably in the second half
curvature effects as discussed in Gibson et al. (1981) and Luo of the bend to a maximum of 1.5 at s/D = 0.5. This is directly
and Lakshminarayana (1996). In the inner-wall region, how- due to a strong reduction of the local gas flow in the streamwise
ever, the DRSM captures the flow acceleration and deceleration direction. The numerical solution provides a qualitative repre-
process satisfactorily. sentation of these trends. Flow fluctuations as represented by
Between = 90◦ and s/D = 0.5, the numerical prediction the turbulence intensities, however, have been severely under-
indicates the formation of a recirculation zone at the inner wall predicted in these regions. This is mainly due to the application
(Figs. 5 and 6). Although the measured profiles at these two of the eddy dissipation transport equation (10), which ignores
stations provide no direct evidence as to the existence of this the contribution arising from the extra strain associated with
recirculation zone, it is highly probable that flow reversal might streamline curvature.
have taken place over a very short distance between =90◦ and Turbulence intensity falls rapidly after s/d = 0.5 and gradu-
s/D = 0.5 in the physical experiment as a result of an adverse ally recovers back to the same level as that at the bend entrance,
pressure gradient. buy symmetry in the turbulence structure is not completely
Downstream of the bend, the maximum peak in the measured restored even at s/D = 9.0.
profile has moved from r ∗ = 0.36 at = 90◦ to r ∗ = 0.3 at
s/D = 0.5, indicating a movement of the core gas flow towards
the outer wall under the influence of centrifugal effect. At s/D= 8.2. Validation of the predicted particle tracks
1.0, a smaller peak in the measured velocity profile at r ∗ = 0.86
suggests that part of the core flow has been carried to the inner- Based on the predicted gas flow field, Lagrangian particle
wall region (Fig. 6). According to Sudo et al. (2001), this is tracking was first performed without the implementation of
attributable to the existence of a secondary flow which carries the SSL force and PWI models. The calculation utilised a
the core gas flow from the outer wall to the inner wall along uniform restitution coefficient E = 0.73. The predicted mean
2078 B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088

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

particle velocity profiles were obtained from the data collec- of the near-wall gas velocities (see Figs. 5 and 6). Particle-wall
tion process as outlined previously and they are compared collisions also strongly affect particle velocity prediction in
against the measured profiles in Figs. 12 and 13. This case this area and this is illustrated in Fig. 14 which plots predicted
will be referred to as the baseline simulation in the following particle velocities (un-averaged) for three particle sizes at =
discussions. 75◦ . One can see from the figure that the predicted velocity
Compared to the measured profiles, the baseline simulation profiles for fine (5 m) and coarse (> 76 m) particles display
has over-predicted longitudinal particle velocities at the outer distinctively different characteristics. The fine particles show a
wall within the bend (Fig. 12) and up to 3D downstream of the strong coupling with the gas phase while the coarse particles
bend (Fig. 13). This is partly attributable to an over-prediction 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 profiles in the vertical duct (— prediction;  data).

This has led us to believe that the coarse particles have already towards the inner wall, except at = 75◦ where large discrep-
separated from the carrier fluid and they are moving under the ancies between the measured and predicted particle velocities
influence of their own inertia as well as particle-wall collisions are found at r ∗ > 0.8 (Fig. 12). With reference to Fig. 14, this
in the simulation. This subsequently implies that particle size arises from a small group of fine particles possessing negative
distribution also affects the prediction of particle velocity in the longitudinal velocities and is a direct result of a recirculation
outer wall region of the duct bend. zone in the predicted gas flow field at = 90◦ (see Fig. 5).
Away from the outer wall, the predicted values follow the The fine particles that have been entrained by the recirculat-
measured velocity profiles with a maximum error of 20% ing gas flow 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 ) profiles within the bend (— prediction;  data).

Velocity statistics pertaining to particles travelling upstream correctly capture the transverse particle motion on the verti-
thus cancel out those travelling downstream in the statistical cal centre-plane of the duct bend (Fig. 15). Both the experi-
averaging process. Hence, one can conclude that the predicted mental measurement and the calculation suggest a thin layer
particle motion in the inner-wall region of the duct bend is of particles having very small transverse velocity components
very sensitive to the background gas flow field. at r ∗ < 0.2 between = 45◦ and 75◦ . This phenomenon arises
Mean transverse velocity profiles for the particles are pre- as some of the particles are moving towards the outer wall and
sented in Figs. 15 and 16. The baseline calculation is able to 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 ) profiles in the vertical duct (— prediction;  data).

LDA probe volume. As a result, the two types of motion cancel (Fig. 16). This directly relates to a considerable discrepancy
each other out during the averaging process. This observation between the measured and predicted gas-phase velocities as
is consistent with the findings of a number of published studies seen in Fig. 9. Hence, one should expect the predicted transverse
on solid saltation, e.g. Tanière et al. (1997) and Ciccone et al. velocity profile to approach the measured distribution as the
(1990). gas-phase solution improves.
In the vertical duct, the particle tracks tend to possess a larger Following the baseline calculation, Lagrangian particle
transverse velocity component than that in the physical flow 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 profiles within the bend (— baseline prediction;  prediction considering SSL and PWI models;  data).

sliding/non-sliding PWI in the numerical model. Figs. 12, particle tracks in the near-wall layer are counted and then
13, 15 and 16 indicate that they have no major effect on the the sum is normalised by the total number of particle tracks
predicted mean particle velocities. However, they do contribute passing through the symmetry axis of that station. According
considerably to the distribution of particles in the duct as illus- to the figure, the application of SSL force and PWI models
trated in Fig. 17. The figure plots the normalised distribution has helped distribute a larger portion of the particle tracks to-
of particle tracks in the inner- and outer-wall regions along the wards the outer duct wall within the bend as compared to the
duct, and it is based on the number of particle tracks passing baseline result. In this region, the baseline model only allows
through a 5.0 mm layer next to the duct walls. At each station, 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 profiles in the vertical duct (— baseline prediction;  prediction considering SSL and PWI models;  data).

collisions. By comparison, the PWI model offers the particles particle tracks away from the inner wall, resulting in a weaker
a second option which is to slide along the wall if their angle presence of particle tracks at the duct inner wall between s/D=
of incidence is below a critical limit. This directly contributes 3.0 and 9.0 as compared to the baseline case.
to a stronger presence of particle tracks at the outer wall of the Streamwise variation of the mean longitudinal velocity for
duct bend. the particles within a 1.5 mm layer next to the outer wall is
In the vertical duct, slip velocities (i.e., uf – up ) are small shown in Fig. 18. In the early part of the bend, the data show
but positive for a large majority of the particle tracks in the a rather rapid decrease in particle velocity as the majority
inner-wall layer. The SSL force, Eq. (20), thus distributes more 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 profiles within the bend (— baseline prediction;  prediction considering SSL and PWI models;  data).

velocity then gradually recovers to about 0.56Ub by 90◦ with Compared to the measured profile, our predictions based
the help of the gas flow entraining and resuspending the on the SSL and particle-wall collision models are only able
rebounded particles. 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 profiles in the vertical duct (— baseline prediction;  prediction considering SSL and PWI models;  data).

process inside the bend. Universally applying a lower resti- particle lose much of their momentum as a result of roping.
tution coefficient to replace the particle-wall collision model Within the rope, the particles are strongly affected by inter-
did not cause a further reduction in particle velocity between particle collisions. One can therefore further improve particle
= 15◦ and 45◦ where particle-wall collisions are expected to velocity predictions in this region by considering inter-particle
be the most frequent. Therefore, particle-wall collision is not collisions in the numerical simulation.
the dominating mechanism through which the particles lose
their momentum. 9. Conclusion
Although we are not able to directly verify the location and
size of particle rope in the duct with the measured data, a higher Numerical simulations have been performed for a curved 90◦
concentration of particle tracks in the outer wall layer between duct bend. The gas-phase simulation was based on a DRSM
= 0◦ and s/D = 1.0 (Fig. 17) does suggest the presence of approach and the solids flow was solved by Lagrangian parti-
a particle rope inside the duct. It is thus very likely that the cle 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 PWI model can also considerably affect the prediction of
and a particle-wall interaction (PWI) model, but ignored the particle velocities and distribution of particles near wall.
effects of particle rotation, turbulence modulation and inter- • Implementation of the shear-slip lift (SSL) force model
particle collisions. The CFD solutions have then been validated did not directly contribute to a more accurate prediction of
against the experimental data in an effort to identify areas where solids motion. However, it acts to redistribute particle tracks
further work is necessary to improve the accuracy of numerical across the flow depending on local gas velocity gradients
prediction for flows with dilute solid suspension. The present and slip velocity, particularly near the inner duct wall down-
study demonstrated the importance of a range of issues that one stream of the bend. This makes it an important hydrody-
needs to consider when performing a numerical simulation of namic force component to consider in shear or curved flow
gas–solid flow inside a power station mill duct, including: calculations.
• The PWI model adopted in this study does not allow near-
• Turbulent gas flow solution based on the current DRSM can wall particles to lose as much momentum as the data suggests
provide a reasonable representation of the mean gas flow at at the outer wall, though it was able to reproduce a similar
the duct vertical centre-plane, except in regions next to the trend. We thus expect the particles to attain a much lower
concave wall and more than 3D downstream of the bend. The near-wall velocity through mechanisms other than PWI in
former is a deficiency, common to all differential Reynolds the physical flow.
stress models, which has not yet been fully resolved by the
fluids modelling community. The turbulence model has con-
siderably under-predicted the decay of secondary motion in Notation
the vertical duct.
• In the inner-wall region and also more than 3D downstream C model constants
of the bend, the predicted particle motion critically depends CD drag coefficient
on the quality of the gas flow solution. Near the outer wall dp particle diameter, m
where the particles are more likely to collide with the wall, D hydraulic diameter of the duct, m
B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088 2087

E restitution coefficient f fluid


En restitution coefficient normal to a surface g gravity
Et restitution coefficient tangential to a surface i size fraction index
F force vector, N i, j, k tensor index
Fr Froude number p partic
fD Schiller–Naumann drag correlation pg pressure gradient
G turbulence production, kg m−2 s−2 sl slip-shear lift
g gravity vector, m s−2 1 pre-impact state
k turbulence kinetic energy, m2 s−2 2 post-impact state
L solids mass loading = ṁp /ṁf
m mass, kg Acknowledgements
ṁ mass flow rate, kg s−1
Ni number of particle tracks allocated to size frac- The authors gratefully acknowledge the financial and other
tion i support received for this research from the Cooperative Re-
n number of particle tracks search Centre (CRC) for Clean Power from Lignite, which is
R duct turning radius established and supported under the Australian Government’s
Re duct Reynolds number Cooperative Research Centres program.
Res particle Reynolds number based on fluid rotation
S inter-particle distance, m References
St particle Stokes number
s, r ∗ curvilinear coordinate system on the duct Akilli, H., Levy, E.K., Sahin, B., 2001. Gas–solid flow behaviour in a
plane of symmetry; r ∗ = 0 at horizontal pipe after vertical-to-horizontal elbow. Powder Technology 116,
43–52.
outer wall; r ∗ = 1 at inner wall, m Barth, T.J., Jesperson, D.C., 1989. The design and application of upwind
u instantaneous velocity vector, m s−1 schemes on unstructured meshes. AIAA Paper 89-0366.
u fluctuating velocity vector, m s−1 Bilirgen, H., Levy, E.K., 2001. Mixing and dispersion of particle ropes in
u , v  fluctuating longitudinal andtransverse lean phase pneumatic conveying. Powder Technology 119, 134–152.
velocity components, m s−1 Boothroyd, R.G., 1971. Flowing Gas–solids Suspensions. Chapman & Hall,
London.
U mean gas velocity vector, m s−1 Borger, G.G., 1973. The optimisation of wind tunnel contractions for the
U, V mean longitudinal and transverse subsonic range. Ph.D. Thesis, Ruhr University, Germany.
velocity components, m s−1 Choi, Y.D., Iacovides, H., Launder, B.E., 1989. Numerical computation of
Ub bulk gas velocity, m s−1 turbulent flow in a square-sectioned 180 deg bend. Journal of Fluids
Engineering 111, 59–68.
UT particle terminal velocity, m s−1 Ciccone, A.D., Kawall, J.G., Keffer, J.F., 1990. Flow visualization/digital
image analysis of saltating particle motion. Experiments in Fluids 9,
Greek letters 65–73.
Fan, L.S., Zhu, C., 1998. Principles of Gas–Solid Flows. Cambridge University
p particle volume fraction Press, Cambridge, pp. 87–110.
1 smooth wall incidence angle, deg Fan, J., Zhang, X., Cheng, L., Cen, K., 1997. Numerical simulation and
ij Kronecker delta experimental study of two-phase flow in a vertical pipe. Aerosol Science
 standard deviation of the wall and Technology 27, 281–292.
roughness angle;  = 3.8 for a Frank, TH., Schade, F.-P., Petrak, D., 1993. Numerical simulation and
experimental investigation of a gas–solid two-phase flow in a horizontal
plexiglass plate, deg channel. International Journal of Multiphase Flow 19 (1), 187–198.
 eddy dissipation rate, m2 s−3 Gibson, M.M., 1978. An algebraic stress and heat-flux model for turbulent
duct turning angle, deg shear flow with streamline curvature. International Journal of Heat and
 gas dynamic viscosity, kg m−1 s−1 Mass Transfer 21, 1609–1617.
0 static friction coefficient Gibson, M.M., Jones, W.P., Younis, B.A., 1981. Calculation of turbulent
boundary layers on curved surfaces. Physics of Fluids 24 (3), 386–395.
d dynamic friction coefficient Giddings, D., Aroussi, A., Pickering, S.J., Mozaffari, E., 2004. A 1/4 scale test
 gas kinematic viscosity, m2 s−1 facility for PF transport in power station pipelines. Fuel 83, 2195–2204.
Gaussian random number with zero Gosman, A.D., Ioannides, E., 1981. Aspects of computer simulation of liquid-
mean andstandard deviation of 1 fuelled combustors. AIAA Paper No. 81-0323.
 density, kg m−3 Huber, N., Sommerfield, M., 1994. Characterization of the cross-sectional
particle concentration distribution in pneumatic conveying systems. Powder
 time scale, s Technology 79, 191–210.
particle rate, s−1 Kuan, B., 2005. CFD simulation of dilute gas–solid two-phase flows with
f fluid rotation vector,s−1 different solid size distributions in a curved 90◦ duct bend. Australia
& New Zealand Industrial and Applied Mathematics Journal 46 (E),
Subscripts C744–C763.
Kussin, J., Sommerfeld, M., 2002. Experimental studies on particle behaviour
A added mass and turbulence modification in horizontal channel flow with different wall
D drag roughness. Experiments in Fluids 33, 143–159.
2088 B. Kuan et al. / Chemical Engineering Science 62 (2007) 2068 – 2088

Laín, S., Sommerfeld, M., Kussin, J., 2002. Experimental studies and Sommerfeld, M., 2000. Theoretical and experimental modelling of particulate
modelling of four-way coupling in particle-laden horizontal channel flow. flows. Lecture Series 2000-06, von Karman Institute for Fluid Dynamics.
International Journal of Heat and Fluid Flow 23, 647–656. Sommerfeld, M., Huber, N., 1999. Experimental analysis and modelling
Levy, A., Mason, D.J., 1998. The effect of a bend on the particle cross-section of particle-wall collisions. International Journal of Multiphase Flow 25,
concentration and segregation in pneumatic conveying systems. Powder 1457–1489.
Technology 98, 95–103. Speziale, C.G., Sarkar, S., Gatski, T.B., 1991. Modelling the pressure–strain
Luo, J., Lakshminarayana, B., 1996. Analysis of streamline curvature effects correlation of turbulence: an invariant dynamical systems approach. Journal
on wall-bounded turbulent flows. Third International Symposium on of Fluid Mechanics 227, 245–272.
Engineering Turbulence Modelling and Measurements, Crete, Greece, May Sudo, K., Sumida, M., Hibara, H., 2001. Experimental investigation on
1996. turbulent flow in a square-sectioned 90-degree bend. Experiments in Fluids
Matsumoto, S., Saito, S., 1970. Monte Carlo simulation of horizontal 30, 246–252.
pneumatic conveying based on the rough wall model. Journal of Chemical Tanière, A., Oestérle, B., Monnier, J.C., 1997. On the behaviour of solid
Engineering of Japan 3 (1), 223–230. particles in a horizontal boundary layer with turbulence and saltation
McIntosh, M.J., Borthwick, I.R., 1984. Investigation of the grinding process effects. Experiments in Fluids 23, 463–471.
in an EVT mill—Development of a process model, Report No. ND/84/003, Tsuji, Y., Morikawa, Y., 1982. LDV measurements of an air–solid two-phase
State Electricity Commission of Victoria, Australia. flow in a horizontal pipe. Journal of Fluid Mechanics 120, 385–409.
Mei, R., 1992. An approximate expression for the shear lift force on a spherical Tsuji, Y., Morikawa, Y., Shiomi, H., 1984. LDV measurements of an air–solid
particle at finite Reynolds number. International Journal of Multiphase two-phase flow in a vertical pipe. Journal of Fluid Mechanics 139,
Flow 18 (1), 145–147. 417–434.
Nallasamy, M., 1987. Turbulence models and their applications to the Uijttewall, W.S.J., Oliemans, R.V.A., 1996. Particle dispersion and deposition
prediction of internal flows: a review. Computers & Fluids 15 (2), in direct numerical and large eddy simulations of vertical pipe flows.
151–194. Physics of Fluids 8 (10), 2590–2604.
Patel, V.C., Sotiropoulos, F., 1997. Longitudinal curvature effects in turbulent Yang, W., Kuan, B., 2006. Experimental investigation of dilute turbulent
boundary layers. Progress in Aerospace Science 33, 1–70. particulate flow inside a curved 90◦ bend. Chemical Engineering Science
Richmond, M.C., Patel, V.C., 1991. Convex and concave surface curvature 61, 3593–3601.
effects in wall-bounded turbulent flows. AIAA Journal 29 (6), 895–902. Yilmaz, A., Levy, E.K., 1998. Roping phenomena in pulverized coal conveying
Rodi, W., Scheuerer, G., 1983. Calculation of curved shear layers with two- lines. Powder Technology 95, 43–48.
equation turbulence models. Physics of Fluids 26, 1422. Yilmaz, A., Levy, E.K., 2001. Formation and dispersion of ropes in pneumatic
Shih, T.H., Liou, W.W., Shabbir, A., Yang, Z., Zhu, J., 1995. A new k . eddy conveying. Powder Technology 114, 165–185.
viscosity model for high Reynolds number turbulent flows. Computers
Fluids 24 (3), 227–238.

S-ar putea să vă placă și