Sei sulla pagina 1di 12

Computers & Fluids 113 (2015) 53–64

Contents lists available at ScienceDirect

Computers & Fluids


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o m p fl u i d

Drag, lift and torque coefficients for ellipsoidal particles: From low to
moderate particle Reynolds numbers
Rafik Ouchene a, Mohammed Khalij a, Anne Tanière a,⇑, Boris Arcen b
a
CNRS, LEMTA, UMR 7563, Université de Lorraine – ESSTIN, 2 rue Jean Lamour, 54500 Vandoeuvre-les-Nancy, France
b
CNRS, LRGP, UMR 7274, Université de Lorraine, Nancy F-54000, France

a r t i c l e i n f o a b s t r a c t

Article history: An accurate prediction of the translational and rotational motion of ellipsoidal particles can be only given
Received 3 December 2013 if a complete set of correlations of the drag, lift and pitching torque coefficients is known. The present
Received in revised form 18 November 2014 study is thus devoted to the assessment of the available correlations in the literature through a
Accepted 4 December 2014
comparison with numerical results of the forces acting on a particle given by a full body-fitted direct
Available online 19 December 2014
numerical simulation (DNS) in the case of a uniform flow, for three different ellipsoidal particles, and
for Reynolds number ranging from 0.1 to 290. The comparison between the computed force (or hydrody-
Keywords:
namic coefficients) and the literature correlations shows clearly that certain precautions must be taken.
Non-spherical particles
Drag
The mean deviation of the most accurate correlation considered in the present study from our full DNS
Lift results can be of the order of 20% for the drag coefficient. For the lift and pitching torque coefficients,
Torque the deviations can increase up to roughly 25%. This comparison shows that further works are definitively
Direct numerical simulation necessary to develop a complete set of correlations for ellipsoidal particles outside Stokes regime.
Body-fitted method Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction particles were previously proposed and summarized in Clift et al.


[7], Allen [8], Mandø [9], and Li et al. [10].
The assumption that particles are spherical has been exten- Even if the particle shapes are idealized, a complete description
sively used in the modeling of particle transport in fluid flows for of the behavior of these particles requires the modeling of their
the past 20 years (Marchioli et al. [1], Arcen et al. [2]). However, translational and rotational motion taking its orientation into
in natural processes as in a variety of engineering applications such account, in contrast to the spherical particle case.
as dispersion of pollutant, pulverized coal combustion, pneumatic Several studies have been performed over the past 20 years, but
transport, or in all aspects of the pulp and paper industry, many they are mainly devoted to the drag force acting on non-spherical
different shapes are possible. Owing to the significant effect of particles. Chhabra et al. [13] were the first to compile and classify
the particle shapes on their motion and their practical importance the most used correlations. According to their study they showed
in hydrodynamic applications, the non-sphericity is more and that the correlation derived by Ganser [14] is the most accurate.
more considered in the modeling and simulations of particles They distinguished two types of approach that were used in later
transport in fluid flows (Mandø and Rosendahl [3], Lain and studies. The first one is to define a single correlation for all shapes
Sommerfeld [4], Njobuenwu and Fairweather [5], van Wachem and orientations (Tran-Cong et al. [15], Yow et al. [16], Holzer and
et al. [6]). Unfortunately, it is not possible to consider each shape Sommerfeld [17]). Indeed, Tran-Cong et al. [15] proposed a corre-
in the implementation of numerical models, because of the nonex- lation from their experimental data obtained from agglomerates
istence of a single approach describing accurately the sizes and of ordered packed spheres. Yow et al. [16] used data available in
shapes of non-spherical particles. Therefore, irregular particle the literature to extract a new correlation which unfortunately,
shapes are commonly idealized to some regular shapes such as does not take the orientation into account. This parameter was
spheroids, cuboids or cylinders (see Fig. 1). In addition, several introduced in the correlation of Holzer and Sommerfeld [17],
methods of size measurement, shape categorization, and shape through two shape parameters which are the sphericity (/) and
parameters used to describe regular and irregular non-spherical the crosswise sphericity. The sphericity is defined by Wadell [18]
as the ratio between the surface area of the volume equivalent
⇑ Corresponding author. sphere and the surface area of the considered particle (S). The
E-mail address: anne.taniere@univ-lorraine.fr (A. Tanière). crosswise sphericity, /? , is defined by Holzer and Sommerfeld

http://dx.doi.org/10.1016/j.compfluid.2014.12.005
0045-7930/Ó 2014 Elsevier Ltd. All rights reserved.
54 R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64

[17] as the ratio between the cross-sectional area of the volume can be applied for any particle aspect ratio. Concerning the pitching
equivalent sphere and the projected cross-sectional area of the torque, it is generally accepted that it is intimately linked to the
considered particle perpendicular to the flow (S? ). The second drag and lift forces which act at the pressure center, its location
approach consists in obtaining a drag coefficient expression for a being different from the gravity center, this induces the pitching
fixed shape and any orientations. Rosendahl [19] suggested to torque. The only formula for the pitching torque coefficient was
determine the drag coefficient at two incidence angles (a), i.e. 0° given by Zastawny et al. [20] for ellipsoidal particles and for specific
and 90° ðC D;a¼0 ; C D;a¼90 Þ, from the best experimental results or values of the aspect ratio.
from existing correlations. Then, a function linking these two lim- Up to now, there is no set of correlations of the drag, lift and
iting cases, with an inflection point at 45°, is used to represent the pitching torque coefficients for ellipsoidal particles outside the
whole range of incidence angles for a non-spherical particle. Stokes regime which takes into account the incidence angle,
Recently, Zastawny et al. [20] proposed new correlations to predict aspect ratio, and Reynolds number. Therefore, when the motion
the drag, lift, and torque (pitching and rotational) coefficients for of ellipsoidal particles has to be predicted, some choices have to
non-spherical particles from data given by a direct numerical sim- be made among available correlations even if they were derived
ulation carried out with an immersed boundary method. Their cor- under strong assumptions. For instance, several authors studied
relation for the drag force is derived following a functional form the motion of inertial ellipsoidal particles using the theoretical
similar to that proposed by Rosendahl [19], nonetheless C D;a¼0 expression for the force and torque derived under the assumption
and C D;a¼90 are given from formulas which are function of the par- of low Reynolds number (Zhang et al. [24], Mortensen [25],
ticle Reynolds number and particle-specific fit parameters. In their Marchioli et al. [26], Tian and Ahmadi [27], Marchioli and Soldati
study, Zastawny et al. [20] compared their drag coefficient for [28], Zhao et al. [29]). The torque was obtained from the expres-
ellipsoidal particles with the theoretical expression provided by sion derived by Jeffery [30] while the force was computed from
Happel and Brenner [21] at low Reynolds numbers. At higher the expression given in Happel and Brenner [21]. However, when
Reynolds numbers, the comparison is conducted with the formula the particle inertia is large, the low Reynolds number assumption
proposed by Rosendahl [19], the drag coefficient at 0° and 90° is not valid anymore. Due to the lack of correlations outside the
being given by the correlation proposed by Holzer and Sommerfeld Stokes regime, researchers are therefore generally forced to study
[17]. At low and higher Reynolds numbers, some significant differ- the non-spherical particle motion under the creeping flow
ences were noted. Currently, given this uncertainty on the accuracy assumption. An accurate prediction of the complete motion of
of the exiting drag coefficient correlations for ellipsoidal particles, non-spherical particles in a turbulent flow by a Lagrangian
it is difficult to make a choice about the drag correlation to use. approach and outside the Stokes regime would be possible if an
This choice is all the more difficult for the lift or torque coefficients accurate set of hydrodynamic coefficient correlations was avail-
since there are few studies devoted to these coefficients in the lit- able. The aim of the present work is thus twofold. Firstly, to guide
erature. In some studies, a proportionality of the lift coefficient the reader through the choice of correlations of the drag, lift and
with the drag one is assumed (Yin et al. [22], Hoerner [23]). torque coefficients outside the Stokes regime. Secondly, to pro-
Nevertheless, Zastawny et al. [20] showed the weakness of this vide data on the drag, lift, and pitching torque coefficients for dif-
assumption through their DNS results and provided a new correla- ferent ellipsoidal particles by a full direct numerical simulation
tion valid for ellipsoidal particles and for two aspects ratios. For using a body-fitted approach. It is hoped that this database will
instance, they proposed the following relation for the lift coeffi- help to develop models of these coefficients.
cient (Table 1 includes all the correlations proposed by Zastawny The document is organized in three sections. The first one is
et al. [20]): devoted to the description of the model of the motion of non-
spherical particles and to the main available correlations. In the
!
b1 b3 b7 b10 second section, the numerical simulation technique and some
CL ¼  sin ðaÞb5 þb6 Rep cos ðaÞb8 þb9 Rep ; ð1Þ the validation cases are presented. Then, the results obtained
Rebp2 Rebp4
for the non-dimensional hydrodynamical parameters (drag, lift,
and pitching torque coefficients) are given, for three different par-
where the coefficients, bi, are specific to the aspect ratio. At the ticle shapes, and discussed in the third section before the
present time, there is no correlation of the lift coefficient which conclusion.

Table 1
Drag, lift, and torque correlations.

Authors Correlations
Holzer and Sommerfeld [17] C D ¼ Re8p p1ffiffiffiffiffi þ Re16 p 1ffiffiffi
þ p3ffiffiffiffiffiffi /3=4
1
0:2
þ 0:42  100:4ð log /Þ /1 / is the sphericity and /? the crosswise sphericity
/? p / Rep ?
8
Rosendahl [19] 3
>
> C D ¼ C D; a ¼90  þ ðC
D; a ¼90   C
D; a ¼0  Þ sin a
>
>
>
< Where : 0:2
C D;a¼0 ¼ Re8p p1ffiffiffiffiffi þ Re 16 p 1ffiffiffi
þ p3ffiffiffiffiffiffi /3=4
1
þ 0:42  100:4ð log /Þ /1
>
> /? p / Rep ?
>
> 0:2
>
: C D;a¼90 ¼ Re8p p1ffiffiffiffiffi þ Re 16 p 1ffiffiffi
þ p3ffiffiffiffiffiffi /3=4
1
þ 0:42  100:4ð log /Þ /1?
/? p / Rep
8 a
Zastawny et al. [20] >
> C ¼ C D;a¼90 þ ðC D;a¼90  C D;a¼0 Þ sin 0 a
> D
>
< where
C D;a¼0 ¼ a1a2  a3a4
>
> Rep Rep
>
>C a5 a7
: D;a¼90 ¼ a  a
Rep6 Rep4
  b7 b10
CL ¼ b1
b  b3
b sin ðaÞb5 þb6 Rep cos ðaÞb8 þb9 Rep
Rep2 Rep4
  c7 c10
CT ¼ c1
c  c3
c sin ðaÞc5 þc6 Rep cos ðaÞc8 þc9 Rep
Rep2 Rep4

Hoerner [23] CL
¼ sin2 a cos a
CD
R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64 55

Fig. 1. Example of spheroid coal–particle shapes with different aspect ratios. Picture was taken from Brix et al. [11] and Powers [12].

Fig. 2. 3d Flow configuration in the inertial frame (x, y, z) (left). 3d Flow configuration in, respectively, co-moving (x0 , y0 , z0 ) and co-rotating (x00 , y00 , z00 ) frames references (right).
The angle of incidence a is defined as the angle between the direction of the fluid velocity and the longest axis of the particle. a and b are respectively the longest and smallest
semiaxes.

2. Motion of a non-spherical particle Table 2


Particle shapes.
There are few models that describe the motion of non-spherical Ellipsoid w = 5/1 Ellipsoid w = 5/2 Ellipsoid w = 5/4
particles suspended in a fluid flow. Indeed, this type of particle is
not only subjected to translation forces but also to a rotational
motion around itself. Let ðx; y; zÞ, ðx0 ; y0 ; z0 Þ, ðx00 ; y00 ; z00 Þ be respec-
tively the inertial, co-moving and co-rotating frames of reference
(see Fig. 2). The translational motion of a particle is described in
the inertial frame of reference by the conservation of momentum.
In addition, for ‘‘heavy’’ solid particles in a gas flow, the drag ~ F D , lift
~
F L and volume forces control the particle motion [24]:
d~
up ~
mp ¼ FD þ ~
F L þ V p ðqp  qf Þ~
g; ð2Þ where Ix00 , Iy00 , Iz00 represent the moment of inertia; xx00 , xy00 , xz00 are
dt
the angular velocities according to the co-rotating frame, and T x00 ,
where mp , ~ up , V p , qp , represent respectively the particle mass, T y00 , T z00 are the torque relative to the three directions.
velocity, volume, and density, qf is the fluid density, and ~ g is the Usually, the forces acting on a non-spherical particle are
gravitational acceleration. characterized by the dimensionless drag CD, lift CL and torque CT
The rotational motion for non-spherical particles is induced by coefficients which depend on the particle Reynolds number
the different location of the center of mass and the center of pres- Rep ¼ k~ uR kdp =m, where ~
uR ¼ ~
up  ~
uf is the relative velocity between
sure. Therefore, an additional torque is added and the following the particle and the fluid, ~ uf represents the velocity of the fluid at
equations have to be solved:
8 dx 00
>
> I 00 x  xy00 xz00 ðIy00  Iz00 Þ ¼ T x00 ;
> x dt
< Table 3
dxy00 Parameters of the simulations.
Iy00  xz00 xx00 ðIz00  Ix00 Þ ¼ T y00 ; ; ð3Þ
>
> dt
>
: dxz00
Rep Particle Ratio w = a/b Orientation a
Iz00 dt
 xx00 xy00 ðIx00  Iy00 Þ ¼ T z00 ;
0.1, 10, 60, 240, 290 Ellipsoids 5/1, 5/2, 5/4 0–90°
56 R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64

the particle location, dp is the volume equivalent spherical diame- term in the Navier–Stokes equations has been neglected. The final
ter, and m is the fluid kinematic viscosity. Thus, these coefficients form of governing equations reads as:
are obtained using the following formulations:
r ~
u ¼ 0;
k~
FDk k~
FLk k~
Tk 1 ð5Þ
CD ¼ ; CL ¼ ; CT ¼ : ð4Þ u  rÞ~
ð~ u¼ rp þ mD~
u;
1
2
qf u2R p4 d2p 1
2
qf u2R p4 d2p 1
2
qf u2R p8 d3p qf
For a particle which is not symmetric (unlike the sphere), the where p is the pressure. In this study, the flow was thus predicted
drag and lift coefficients depend not only on particle Reynolds by solving steady-state equations although it is known that the flow
number but also on its incidence angle a with respect to the flow regime can be unsteady for the highest particle Reynolds number
(see Fig. 2). This effect plays an important role in the motion of considered (Rep = 290). The unsteadiness of the flow past a sphere
non-spherical particles [23] through the drag and lift forces. Hence, occurs for Rep 2 ½270  280 [32], while for an ellipsoidal particle
for an accurate description of non-spherical particles motion, the with w = 2 aligned or perpendicular to the flow, the flow remains
drag, lift and torque coefficients have to be precisely predicted. steady for Rep < 250 [35]. Nevertheless, as can be noted from the
Table 1 reports the correlations that will be compared to our data given in Johnson and Patel [32] for a sphere at Rep = 300, the
numerical results in order to assess their accuracy. Concerning fluctuations of the drag coefficient around its time-averaged value
the drag coefficient three correlations have been chosen for the are weak (roughly 0.5% of the time-averaged drag coefficient). From
comparison. They were developed by Rosendahl [19], Holzer and a preliminary steady computation of flow past a sphere at Rep = 300,
Sommerfeld [17] and Zastawny et al. [20]. The lack of lift and it was shown that the difference between the present drag coeffi-
torque correlations has conducted us to use the correlations pre- cient and that of Johnson and Patel [32] is of 3.7%. It is believed that
sented in the introduction. For the lift, the expressions by Hoerner the unsteadiness of the flow has little effect on the drag, lift, and
[23] and Zastawny et al. [20] will be assessed, and for the torque, torque coefficients for the particle Reynolds number range consid-
we will consider the correlation by Zastawny et al. [20]. ered in the present study. The drag and lift are the components of
the fluid forces (pressure and friction) acting on the considered par-
3. Numerical overview ticle in the three principal directions. These forces are obtained by
integrating the total stress acting on the particle surface area. Thus,
3.1. Numerical technique both forces are:
Z Z
We want to examine the hydrodynamic forces acting on ellip- k~
F D k ¼ F D;pressure þ F D;friction ¼  p~ x ~
ndS þ ~ n  s ~
xdS
soids of revolution (Table 2), i.e. the minors semi-axes b and c Z S ZS ð6Þ
 
are equal (b = c). The range of aspect ratios w ¼ length
width
¼ ba consid- k~
F L k ¼ F L;pressure þ F L;friction ¼  p~ y ~
ndS þ ~ n  s ~
ydS;
S S
ered in this study is w 2 ½1  5 (see Table 2). This is the typical
range encountered in many technical processes such as coal com- where s and ~ n are respectively the viscous stress tensor and the
bustion (see Fig. 1). In addition, the particle Reynolds number in normal vector to the particle surface. The pitching torque is the
many technical processes is, generally, lower than 300. Thus, in torque induced by the drag and lift forces acting on the center of
this study, we studied the hydrodynamic forces acting on ellipsoi- pressure ~
xcp :
dal particles for Rep 2 ½0:1  300½. In Table 3, the aspect ratios and ~
T ¼ ð~
FD þ ~
FLÞ  ~
xcp : ð7Þ
particle Reynolds numbers computed are summarized.
Besides, it is known that the flow past such particles becomes Simulations are performed using the CFD code Ansys FluentÓ. Cen-
non-axisymmetric for Rep > 200 [20,31–35]. Hence, it was required tered scheme are used for discretizing all the numerical fluxes in
to consider simulation in a complete 3D configuration. Thus, the the Navier–Stokes equations. The problem of the coupling between
particle is supposed fixed and immersed in a tridimensional uni- velocity and pressure is tackled with a Semi-Implicit Method for
form flow. Pressure-Linked Equations (SIMPLE) [31]. The geometrical configu-
ration (Fig. 3) is a parallelepiped fluid box surrounding the particle.
3.2. Governing equation and numerical details The dimensions of the box were determined to prevent any distur-
bance of the flow due to boundary conditions, particularly in the
In the present study, it is assumed that the flow is isothermal, boundary layer. From a practical point of view, these dimensions
incompressible, the fluid is considered as Newtonian. The unsteady were imposed by the case Rep = 0.1 resulting in a greater boundary
layer where the particle is located transversely to the flow. In addi-
tion, the length of the computational box was set to have an ade-
quate distance for the development of the wake. The mesh of the

Fig. 3. Geometrical configuration and boundary conditions. Fig. 4. Computational mesh for flow past ellipsoidal particle.
R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64 57

fluid domain consists of tetrahedral elements. It was verified that comparison with the recent DNS (based on a mirroring immersed
the mesh surface of the particle is sufficiently fine to give an accu- boundary method) results of Zastawny et al. [20], it is noted that
rate integration of the forces acting on the particle surface (Fig. 4). they obtained a similar trend. Nonetheless, their prediction of the
The influence of the number of mesh cells on the hydrodynamic drag coefficient is less accurate to that given by the present DNS
coefficients for an ellipsoid (w = 5/1) and an incidence angle equals whatever the value of the Reynolds number. The mean relative
to 90° at Rep = 290 is shown in Table 4. As it can be seen, the drag deviation from the present DNS results with the drag coefficient
coefficient can be considered mesh independent with a number of correlation is 1.5% whereas from the DNS results of Zastawny
cells equals to 650 000. The difference with the finest grid result et al. [20], it is about 17%.
is less than 1%. For this aspect ratio and incidence angle, the com-
puted values of the lift and torque coefficients are close to zero  The second case is performed with three different ellipsoids
since the wake is axisymmetric. Thus, the number of cells chosen (w = 5/1, w = 5/2 and w = 5/4) in a Stokes flow. The drag and lift
for all the simulation is always greater than 500 000 (this number coefficients results are compared to the analytical solution
varies with w and a). It has to be noted that the boundary layer is given in Happel and Brenner [21] in Fig. 6. The deviations from
always represented by at least 20 nodes. the theoretical results are shown in Table 5. We also reported
The boundary conditions are presented in Fig. 3. A constant the DNS numerical results of Zastawny et al. [20] (drag and lift
velocity inflow condition is imposed at the inlet boundary whereas coefficients) in Fig. 6 and Table 5 in order to compare two differ-
homogeneous Neumann conditions are chosen at the outlet. On ent DNS approaches (body-fitted and immersed boundary
lateral boundaries, no-slip Dirichlet condition is specified. methods). Note that the case w = 5/1 was not considered by
Zastawny et al. [20].
3.3. Validation cases
From the second validation test case results, it can be noted that
Before examining the results obtained for the cases given in the DNS results for the drag and lift coefficients are qualitatively in
Table 3, the computational methodology is validated by comparing good accordance with the theoretical results of Happel and Bren-
the results obtained from the present DNS results for a sphere and ner [21]. However, a slight overestimation of the present numerical
different ellipsoidal particles with a well-known drag correlation, results is observed for the drag and lift results whatever the value
another simulation at particle scale, and theoretical results. The of the aspect ratio. The DNS, using an immersed boundary method,
two validation cases are described below. by Zastawny et al. [20] underestimates the drag coefficient while
their lift coefficient is always greater than the theoretical results.
 In the first one, we computed the drag coefficient of a sphere for As shown in Table 5, the mean relative deviation of the present
Rep 2 ½0:1  280, the present results are compared to the well- results from the theory [21] does not exceed 5% for the drag coef-
known correlation by Schiller and Naumann [36], ficient. This deviation is less than 10% for the lift coefficient what-
24Re1 0:687 ever the value of the aspect ratio. It is noted that the mean
p ð1 þ 0:15 Rep Þ, in Fig. 5. The results given by the
recent DNS of Zastawny et al. [20] are also reported. deviation increases when w ? 1 but it never exceeds 10%. In addi-
tion the maximum deviations are of the same order as the mean
For this first validation test case, the present direct numerical deviation. The increase of the mean relative deviation as the aspect
simulation applied at the scale of a spherical particle gives very ratio tends to unity is also noted for the results given by Zastawny
good results with respect to the well-known drag correlation of et al. [20]. However, the mean relative error is at least two times
Schiller and Naumann [36] for a large particle Reynolds range. In higher for a shape close to a sphere than the one obtained from
the present DNS results.
Zastawny et al. [20] also observed this deviation from the theo-
Table 4 retical expression by Happel and Brenner [21], they explained it by
Influence of the number of mesh cells on the drag, lift on torque coefficient for w = 5/1 the use of a finite domain. Nevertheless, the reasons for these dis-
at Rep = 290.
crepancies at low Reynolds number and for large aspect ratio are
Number of mesh cells CD CL CT not clear since no study of the effect of the computational domain
344,134 0.853 8.32  105 5.87  104 size, or of the number of mesh cells was provided in their study. It
664,911 0.830 1.53  105 8.56  105 is surprising that the results of Zastawny et al. [20] are better for
978,559 0.828 3.90  105 1.14  104 w = 5/2 than for w = 5/4. It is also surprising that the lift coefficient
is better predicted than the drag coefficient for w = 5/2.
In order to understand these particular trends, we present in
Fig. 7, the pressure and viscous contributions to the drag and lift
coefficients computed from Eq. (6) with the present numerical
results for the three different ellipsoidal particles at Rep = 0.1. In
this figure, it is noticed that the viscous stress is the main contribu-
tion to the drag coefficient whatever the angle of incidence and
aspect ratio. This trend is completely different for the lift coeffi-
cient which is mainly driven by the pressure. Therefore, it could
be assumed that the DNS of Zastawny et al. [20] did not accurately
capture the viscous effect for w = 5/2 since better results are
obtained for the lift coefficient than for the drag coefficient. How-
ever, for w = 5/4, this explanation does not hold since their drag
coefficient results, which are mainly governed by the viscous
effects, are in better accordance with the theory than their lift coef-
ficient. Finally, it can be only concluded that their results are less
accurate than the present body-fitted computation.
Fig. 5. Drag coefficient as a function of Rep. Present results: ( ); Schiller and To sum up, the present DNS based on a body-fitted approach
Naumann [36]: (–). Results of Zastawny et al. [20] are also reported ( ). gives very satisfactory results of the forces acting on a sphere over
58 R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64

Fig. 6. Drag (left) and lift (right) coefficients obtained for different ellipsoids at Rep = 0.1. First row: w = 5/1, second row: w = 5/2, third row: w = 5/4. Present results: ( );
theoretical results of Happel and Brenner [21]: (–); results of Zastawny et al. [20] are also reported ( ).

Table 5
Deviation of the computed drag and lift coefficients from the theoretical results by Happel and Brenner [21] in Stokes flow regime.

Drag coefficient Lift coefficient


Mean relative deviation (%) Maximum relative deviation (%) Mean relative deviation (%) Maximum relative deviation (%)
w = 5/1 Present simulation 2.14 2.45 4.2 4.29
w = 5/2 Present simulation 2.89 3.07 4.13 4.21
Zastawny et al. [20] 13.34 17.75 2.67 4.07
w = 5/4 Present simulation 4.04 4.13 8.04 8.65
Zastawny et al. [20] 11.31 13.11 23.68 54.57

a large Reynolds number range, as well as for ellipsoidal particles Brenner [21] for ellipsoidal particles is acceptable (never exceeds
at low Reynolds number. The mean relative deviation from the 10%). The present DNS results have also been compared to DNS
correlation of a widely used sphere drag correlation (Schiller and data obtained using a different numerical approach (mirroring
Naumann [36]) or from the theoretical results by Happel and immersed boundary method). This comparison shows that similar
R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64 59

Fig. 7. Pressure (dashed lines) and friction (continuous lines) drag (left) and lift (right) coefficients as a function of the angle of incidence a in a Stokes regime. ( ): w = 5/1;
(–): w = 5/2; ( ): w = 5/4.

Fig. 8. Comparison of the computed drag coefficient with the existing correlations for different particles Reynolds numbers. (–): Rosendahl [19]; ( ): present results;
( ): Zastawny et al. [20]; ( ): Holzer and Sommerfeld [17]. The dashed lines are referred to Rep = 10 and the continuous lines are referred to Rep = 290.

tendencies are given by both methods, nevertheless, the results angles of incidence and for different particle Reynolds numbers.
from the mirroring immersed boundary method DNS are less Several aspect ratios were computed (see Table 3).
accurate. The present computational methodology (domain size,
number of mesh points, Navier–Stokes solver, and computation 4.1. Drag coefficient
of the hydrodynamic force) used to compute the hydrodynamic
forces acting on spherical and non-spherical particles can be thus In Fig. 8, the computed drag coefficients are presented and
considered to be valid. In the next section, results obtained outside compared with the correlations of Holzer and Sommerfeld [17],
Stokes flow regime are shown for the ellipsoidal particles Rosendahl [19] and Zastawny et al. [20] given in Table 1. Concern-
described in Table 2. ing the CD results, it has to be noted that the value of CD,a=0°, and
CD,a=90°, used for the correlation of Rosendahl [19] were computed
4. Results and discussion from the correlation of Holzer and Sommerfeld [17]. The drag
coefficient is presented as a function of the angle of incidence.
Simulations were performed to compute the drag CD, lift CL and Zastawny et al. [20] results are not presented for the case w = 5/1
pitching torque CT coefficients for ellipsoidal particles at different since they did not consider this aspect ratio in their study.

Table 6
Deviation of the computed drag coefficients from the present results at Rep = 10.

Mean relative deviation (%) Maximum relative deviation (%)


Holzer and Sommerfeld [17] Rosendahl [19] Zastawny et al. [20] Holzer and Sommerfeld [17] Rosendahl [19] Zastawny et al. [20]
w = 5/1 6.73 3.41 12.49 7.3
w = 5/2 3.18 4.7 11.69 6.82 6.98 15.86
w = 5/4 15.35 15.73 21.32 16.78 16.78 21.54
60 R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64

Table 7
Deviation of the computed drag coefficients from the present results at Rep = 290.

Mean relative deviation (%) Maximum relative deviation (%)


Holzer and Sommerfeld [17] Rosendahl [19] Zastawny et al. [20] Holzer and Sommerfeld [17] Rosendahl [19] Zastawny et al. [20]
w = 5/1 116.1 83.73 140.7 99.22
w = 5/2 71.7 59.01 22.08 84.59 79.46 37.56
w = 5/4 18.99 17.50 7.19 25.02 25.02 10.07

Qualitatively, the present results are in good agreement with and CD,a=90° used for the correlation of Rosendahl [19] were com-
those issuing from the literature since a similar trend is observed. puted from the correlation of Holzer and Sommerfeld [17]. It is
The first main observation lies in the decrease of the drag values therefore not surprising to have similar deviations.
when the Reynolds number increases whatever the aspect ratio. At higher Reynolds number (Table 7), we note that the devia-
The second one is linked to the evolution of the drag value as a tion previously noted at low Reynolds number is lower in the case
function of incidence angle for a given Reynolds number. The drag where the shape is close to a sphere. However, when the aspect
coefficient increases for increasing incidence angle whatever the ratio increases, the deviation does not decrease in contrast to the
values of aspect ratio. An inflection point is noted around a = 45°. low Reynolds number case. The deviation is much larger for
Note that this inflection point is not captured by the correlation w = 5/2 at Rep = 290 while this was noticed for w = 5/4 at
of Holzer and Sommerfeld [17]. For a shape near to a sphere Rep = 10. Moreover, important deviations from the present results
(w = 5/4), the drag coefficient increase is attenuated whatever the are remarked when the correlation of Holzer and Sommerfeld
Reynolds value. In other words, for this aspect ratio near to a [17] or of Rosendahl [19] is used. Again, similar mean and maxi-
sphere, the drag coefficient does not significantly depend on the mum deviations are obtained. The deviation can be found to be
incidence angle as expected. more than 100% using the correlation from Holzer and Sommerfeld
Concerning the quantitative comparison of the results given, it [17] whereas the deviation is slightly smaller than 100% (but
is noticed that there is a great disparity in the results plotted. remains around 100%) when the correlation of Rosendahl [19] is
The difference between our drag coefficient values and those used for the highest aspect ratio (w = 5/1). If the correlation of
obtained from the correlations by Holzer and Sommerfeld [17], Holzer and Sommerfeld [17] does not provide an accurate predic-
Rosendahl [19], and Zastawny et al. [20] are reported in Tables 6 tion of the drag coefficient, the prediction by the correlation of
(Rep = 10) and 7 (Rep = 290) for each aspect ratio. These deviations Rosendahl [19] will be also inaccurate. These inconsistencies with
were averaged on the whole range of angles of incidence. At low respect to our results are certainly due to the fact that their
Reynolds number (Rep = 10, Table 6), we clearly note that the most correlation was developed for arbitrary shaped particles from
important deviation from our results is obtained for a shape close experimental and numerical data obtained for various forms like
to a sphere (w = 5/4) whatever the correlations used. It is noticed isometric particles, cuboids, cylinders or disks. This correlation is
that the mean relative deviation is 20% for the results of Zastawny not dedicated to ellipsoidal particles, a lack of accuracy is thus
et al. [20] and around 15% with the results of Holzer and expected. Holzer and Sommerfeld [17] found that the accuracy of
Sommerfeld [17] and of Rosendahl [19]. A similar trend is noted their correlation depends on the shape of the particles (the mean
for the calculated maximum relative deviation. The similarities relative deviation between the correlation and experimental data
noted for the correlations proposed by Holzer and Sommerfeld is about 14%, we have found a similar deviation at Rep = 10). In
[17] and Rosendahl [19] were expected since the values of CD,a=0°, comparison to the results from the correlations of Holzer and

Fig. 9. Normalized drag coefficient, CD/CD,a=90°, as a function of the angle of incidence a. ( ): Rep = 10; ( ); Rep = 60; ( ): Rep = 240; ( ): Rep = 290.
R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64 61

Fig. 10. Pressure (dashed lines) and friction (continuous lines) drag coefficients normalized by CD,a=90° as a function of the angle of incidence a for different aspect ratios.
( ): w = 5/1; ( ); w = 5/2; ( ): w = 5/4. Left: Rep = 10. Right: Rep = 290.

Sommerfeld [17] or Rosendhal [19], the accuracy of the correlation pressure and friction drag forces) normalized by CD,a=90° are given
provided by Zastawny et al. [20] is better at Rep = 290. At Rep = 10, in Fig. 10 for two values of Rep (10 and 290). The pressure drag
the mean deviation given by this latter correlation is higher than coefficient is seen to be significantly influenced by the angle of
those provided by the two other correlations. Moreover, the mag- incidence while the friction drag coefficient does not greatly vary
nitude of the deviation depends on the particle aspect ratio. More as a function of the particle orientation. Therefore, the evolution
accurate results are obtained for w = 5/2 at Rep = 10 while a better of total drag coefficient as a function of a is mainly due to the pres-
prediction is found for w = 5/4 at Rep = 290. However, the mean sure drag coefficient contribution. This trend is observed at both
relative error remains high. Unfortunately, we have no clear expla- particle Reynolds numbers. It should be noted, that the friction is
nation for these particular trends. the main contribution to the drag at Rep = 10, and whatever the
In order to better characterize the evolution of drag coefficient angle of incidence. At Rep = 290, this contribution is higher than
as a function of the Reynolds number, aspect ratio, and angle of that of pressure at low angle of incidence, then the pressure contri-
incidence, we have plotted in Fig. 9 the computed normalized drag bution dominates. Finally, from this figure, it can be seen that the
coefficients CD/CD,a=90° as a function of a for different aspect ratio friction contribution is less influenced by the particle aspect ratio
and for several particle Reynolds numbers. than the pressure contribution. For instance, at Rep = 290, the pres-
In Fig. 9, it is clearly seen that CD decreases when the Reynolds sure contribution for w = 5/4 is 4 times higher than that obtained
number increases as observed before. It is also interesting to note for w = 5/2 while the friction contributions are almost identical.
that when Rep is higher than 240 the ratio CD/CD,a=90° does not sig- To sum up, it has been shown that the present computation can
nificantly evolves as a function of the Reynolds number. As already accurately predict the drag coefficient of a spherical particle from
noticed, the incidence angle has a less important influence when Stokes flow up to Rep = 290. A very good accordance between the
the aspect ratio decreases. present DNS computations and the theoretical results in Stokes
To more thoroughly examine the drag coefficient, the pressure flow regime has been also observed for the drag coefficient of ellip-
and viscous drag contributions (the total drag force is the sum of soidal particles. For higher Reynolds numbers, the comparison

Fig. 11. Comparison of the computed lift coefficient with the existing correlations for different particle Reynolds numbers. ( ): present results; ( ): Zastawny et al.
[20]; ( ): Hoerner [23]. The dashed lines are referred to Rep = 10 and the continuous lines are referred to Rep = 290.
62 R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64

Table 8
Deviation of the computed lift coefficients from the present results at Rep = 10.

Mean relative deviation (%) Maximum relative deviation (%)


Hoerner [23] Zastawny et al. [20] Hoerner [23] Zastawny et al. [20]
w = 5/1 84.90 173.23
w = 5/2 134.86 8.70 263.23 11.72
w = 5/4 601.16 19.94 978.4 23.59

with existing drag coefficient correlations (derived from numerical In order to better understand the trend observed for the lift
and/or experimental data) shows that they do not accurately coefficient, we computed the contribution to the lift coefficient
predict the evolution of the drag coefficient as a function of the induced by the pressure and the viscous stress. The pressure and
Reynolds, aspect ratio, and angle of incidence. From a physical viscous contributions are plotted in Fig. 12 for Rep = 10 and
point of view, it has been shown the friction part is not greatly Rep = 290. It can be first noticed that the incidence angle has a weak
influenced by the angle of incidence in contrary to the pressure effect on the pressure and viscous stress lift coefficients. In
contribution. addition, the pressure is the main contribution to the lift force
whatever the particle Reynolds number and aspect ratio. Nonethe-
4.2. Lift coefficient less, there is an important difference between these two particle
Reynolds number cases. At Rep = 10, the viscous contribution is
The lift coefficient CL at Rep = 10 and Rep = 290 is shown in always negative while this contribution is positive at Rep = 290
Fig. 11 as a function of the incidence angle. When w = 5/2 or 5/1, and very close to zero. The pressure plays definitely a major role
the lift coefficient is higher for Rep = 10 than for Rep = 290, as in the lift coefficient. It is interesting to note that Cherukat et al.
already noticed for CD. This trend is also noted for w = 5/4 when [37] also found, from numerical simulations, that the viscous stress
a < 70°, but for higher angle of incidence the lift coefficient at contribution becomes negative for a spherical solid particle
Rep = 290 does not tend toward zero. The lift coefficient is thus embedded in a linear shear flow when the particle Reynolds num-
greater than that at Rep = 10 for this range of incidence angle. This ber is greater than unity. A similar trend was observed by Legendre
nonzero value of the lift force at Rep = 290 is induced by the non- and Magnaudet [38] for a spherical bubble in a linear shear flow
axisymmetry of the wake. Johnson and Patel [32] also observed when the bubble Reynolds number is greater than one. In the pres-
this effect for a sphere (w = 1) when Rep > 210. In addition, except ent study, this negative contribution is induced by the particle
for w = 5/4 at Rep = 290 and a > 70°, the computed lift coefficient shape since the ellipsoidal particles considered are embedded in
is seen to decrease as w tends toward unity (sphere). This was a uniform flow.
expected since in the present study, the lift force is only induced
by the particle shape. Finally, it has to be noted that, whatever 4.3. Pitching torque coefficient
the aspect ratio and particle Reynolds number, the lift coefficient
is maximum near an incident angle of a = 45°. In Fig. 13, the computed pitching torque coefficient is plotted as
These trends are similar to that obtained by Zastawny et al. [20]. a function of a for two aspect ratios and two particle Reynolds
In Fig. 11, the prediction of the lift coefficients given by the corre- numbers (Rep = 10 and Rep = 290). It is noticed that the pitching
lations derived by Hoerner [23] and Zastawny et al. [20] are also torque decreases as the particle Reynolds number increases. As
plotted. For sake of clarity, the curve predicted by Hoerner [23] is expected, this torque also decreases when w tends toward unity.
cut for w = 5/4. From this comparison, it is seen that the evolution These data are compared to the results given by the correlation
of CL predicted by Hoerner [23] does not coincide with the present derived by Zastawny et al. [20]. It is noted that a better accordance
results nor with those obtained by Zastawny et al. [20]. Moreover, is obtained for w = 5/2 than for w = 5/4. The deviation between the
the maximum lift coefficient predicted by Hoerner [23] is located prediction by the correlation given by Zastawny et al. [20] and the
at a = 60° while Zastawny et al. [20] and the present results show present results are shown in Table 10. It should be kept in mind
that it is located near a = 45°. It can be definitively concluded that that the pitching torque is obtained by the combination of drag
the lift coefficient cannot be only a function of incidence angle a and lift forces acting on the center of pressure. Therefore, as
and drag coefficient CD as suggested by Hoerner [23]. A better expected, the differences between the present results and those
prediction of the lift coefficient is obtained with the correlation given by the correlation of Zastawny et al. [20] are of the same
presented by Zastawny et al. [20] as can be seen in Fig. 11. In Tables order to those already noticed for the drag and lift coefficients.
8 and 9, the deviation between the prediction by Hoerner [23] or For the pitching torque, the mean relative deviation is always
Zastawny et al. [20] and the present results are given. It confirms lower than 25%. The deviation increases with increasing aspects
that the correlation by Hoerner [23] is not able to predict the lift ratios. It also increases with increasing particle Reynolds numbers
force acting on ellipsoidal particles since the mean deviation for w = 5/2 ellipsoids. On contrary, it is seen that the mean devia-
ranges from 35% to 600%. The maximum mean relative deviation tion does vary as a function of the particle Reynolds number for
for the results of Zastawny et al. [20] is about 27%. It is noted that w = 5/4. Again, these trends are different from those noticed for
this deviation error increases with increasing particle Reynolds the drag and lift coefficients and cannot be easily explained.
numbers and/or decreasing aspect ratio in contrast to the results
for the drag coefficients. It is also observed that the maximum
deviation is obtained for a shape near a sphere whatever the value 5. Conclusion
of the Reynolds number. It has to be noted that in their simula-
tions, the number of mesh cells depends on the Reynolds number The first objective of the present study was to provide, using a
considered. This number was varied from 8 to 12 cells on the diam- body-fitted DNS, the drag, lift and pitching torque coefficients for
eter of the volume equivalent sphere, it can be thus assumed that three different ellipsoidal particles and for Reynolds number rang-
this number is not enough high to capture the boundary layer at ing from 0.1 to 290, and to compare these results to some existing
high Reynolds number (keeping in mind that the boundary layer correlations (Holzer and Sommerfeld [17], Rosendahl [19] and
thickness varies approximately as Rep1/2). Zastawny et al. [20]). From this comparison, it was shown that
R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64 63

Table 9
Deviation of the computed lift coefficients from the present results at Rep = 290.

Mean relative deviation (%) Maximum relative deviation (%)


Hoerner [23] Zastawny et al. [20] Hoerner [23] Zastawny et al. [20]
w = 5/1 38.83 86.21
w = 5/2 36.64 21.34 85.15 30.25
w = 5/4 54 27.25 89.50 67.82

Fig. 12. Pressure (dashed lines) and friction (continuous lines) lift coefficients, normalized by the total lift coefficient, as a function of the angle of incidence a for different
aspect ratios. ( ): w = 5/1; (–); w = 5/2; ( ): w = 5/4. Left: Rep = 10. Right: Rep = 290.

Fig. 13. Comparison of the torque coefficients for different particles Reynolds numbers. Present results: ( ); results of Zastawny et al. [20]: ( ). The dashed lines are
referred to Rep = 10 and the continuous lines are referred to Rep = 290.

Table 10
as already noted by Zastawny et al. [20], that the lift coefficient
Deviation of the computed drag, lift, and pitching torque coefficients given by
Zastawny et al. [20] from the present results at Rep = 10 and 290. cannot be modeled as a function of only the angle of incidence
and the drag coefficient. The correlation by Zastawny et al. [20]
Mean relative deviation (%) Maximum relative deviation (%)
gives more accurate results, nevertheless, it was observed that
Rep = 10 Rep = 290 Rep = 10 Rep = 290 the mean relative deviation from our DNS results increases with
w = 5/2 2.24 15.17 4.35 20.28 increasing Reynolds number and ranges from 8% to 27%. As for
w = 5/4 24.75 23.43 26.54 30.51 the drag coefficient, they provide a correlation which is only valid
for two ellipsoid aspect ratios. The lift coefficient was also divided
into a pressure part and a viscous part. From this decomposition, it
these correlations cannot accurately predict the drag coefficient for has been noticed that the pressure is the main contribution to the
the three aspect ratios and over the entire particle Reynolds lift force whatever the aspect ratio, angle of incidence, and Rey-
number range considered. The correlation by Zastawny et al. [20] nolds number. Moreover, as also observed for a spherical particle
provides the best prediction, nevertheless, the deviation from our in a linear shear flow, the viscous part can be negative (i.e. in the
full DNS results can be of the order of 20% and their correlation opposite direction to the total lift force). Finally, we have presented
for ellipsoidal particles is valid for only two aspect ratios. From the pitching torque coefficient and compared it to the correlation
the DNS results, it was noted the viscous stress and pressure con- derived by Zastawny et al. [20]. As for the drag and lift coefficients,
tributions to the drag force vary with the Reynolds number, aspect the mean deviation from our DNS results ranges from 2% to 25%.
ratio, and angle of incidence. Generally, the viscous part is the main Therefore, the study by Zastawny et al. [20] is the only one to pro-
contribution. In addition, it was noted that the functional form of vide a complete set of correlations for ellipsoidal particles. How-
the model proposed by Rosendahl [19] to express the drag coeffi- ever, these correlations are valid for only two aspect ratios. This
cient as a function of the angle of incidence is acceptable. Concern- is one of the major drawbacks of their models. The second one is
ing the lift force, the comparison with the correlations by Hoerner that we have found that the DNS used by Zastawny et al. [20] to
[23] or Zastawny et al. [20] have shown that the correlation by derive their models does not provide an accurate prediction of
Hoerner [23] is not able to predict the lift coefficient. This means, the hydrodynamic force acting on ellipsoidal particle at low
64 R. Ouchene et al. / Computers & Fluids 113 (2015) 53–64

Reynolds number. In addition, some significant deviations have [14] Ganser GH. A rational approach to drag prediction of spherical and non-
spherical particles. Powder Technol 1993;77(2):143–52.
been also noted between the correlations derived by Zastawny
[15] Tran-Cong, Gay M, Michaelides E. Drag coefficients of irregularly shaped
et al. [20] and our DNS results outside Stokes flow regime. For all particles. Powder Technol 2004;139:21–32.
the reasons given above, we deeply believe that further efforts [16] Yow HN, Salman A, Pitt MJ. Drag correlations for particles of regular shape.
are definitively necessary to develop a complete set of correlations Powder Technol 2005;16(4):363–72.
[17] Holzer A, Sommerfeld M. New simple correlation formula for the drag
for ellipsoidal particles outside Stokes flow regime. These coefficient of non spherical particles. Powder Technol 2008;184:361–5.
correlations should be function of the angle of incidence, particle [18] Wadell H. The coefficient of resistance as a function of Reynolds number for
Reynolds number, and aspect ratio of particles in order to solids of various shapes. J Franklin Inst 1934;217:459–90.
[19] Rosendahl L. Using a multi-parameter particle shape description to predict the
accurately model the forces and torque acting on non-spherical motion of non-spherical particle shapes in swirling flow. Appl Math Model
particles. We hope that the present existing database on drag, lift 2000;24:11–25.
and torque coefficients will help to the future developments of [20] Zastawny M, Mallouppas G, Zhao F, van Wachem B. Derivation of drag and lift
force and torque coefficients for non-spherical particles in flows. Int J
such correlations. Multiphase Flow 2012;101:288–95.
[21] Happel J, Brenner H. Low Reynolds number hydrodynamics. Englewood
Acknowledgments Cliffs: Prentice-Hall; 1965.
[22] Yin CG, Rosendahl L, Kaer SK, Sorensen H. Modelling the motion of cylindrical
particles in a non uniform flow. Chem Eng Sci 2003;58(15):3489–98.
This work received funding by ANR-PLAYER; their support is [23] Hoerner SF. Fluid dynamic drag. Hoerner Fluid Dynamics; 1965.
gratefully acknowledged. We would like to thank our partners [24] Zhang H, Ahmadi G, Fan F-G, McLaughlin JB. Ellipsoidal particles transport and
deposition in turbulent channel flows. Int J Multiphase Flow
IMFT and I2M laboratories.
2001;27(6):971–1009.
[25] Mortensen PH, Andersson HI, Gillissen JJJ, Boersma BJ. Dynamics of prolate
References ellipsoidal particles in a turbulent channel flow. Phys Fluids 2008;20:093302.
[26] Marchioli C, Fantoni M, Soldati A. Orientation, distribution and deposition of
[1] Marchioli C, Soldati A, Kuerten JGM, Arcen B, Taniere A, Goldensoph G, et al. elongated, inertial fibers in turbulent channel flow. Phys Fluids
Statistics of particle dispersion in direct numerical simulations of wall- 2010;22:033301.
bounded turbulence. Results of an international collaborative benchmark test. [27] Tian L, Ahmadi G. Fiber transport and deposition in human upper
Int J Multiphase Flow 2008;34:879–93. tracheobronchial airways. J Aerosol Sci 2013;60:1–20.
[2] Arcen B, Taniere A, Zaichik L. Assessment of a statistical model for the [28] Marchioli C, Soldati A. Rotation statistics of fibers in wall shear turbulence.
transport of discrete particles in a turbulent channel flow. Int J Multiphase Acta Mech 2013;224:2311–29.
Flow 2008;34(4):419–26. [29] Zhao L, Marchioli C, Andersson HI. Slip velocity of rigid fibers in turbulent
[3] Mandø M, Rosendahl L. On the motion of non-spherical particles at high channel flow. Phys Fluids 2014;26:063302.
Reynolds number. Powder Technol 2010;202(25):1–13. [30] Jeffery GB. The motion of ellipsoidal particles immersed in a viscous fluid. Proc
[4] Lain S, Sommerfeld M. Kinetic simulations of non-spherical particle response R Soc 1922;102:161–79.
behavior. In: 8th International conference on multiphase flow; 2013. [31] Lázaro BJ, Lasheras JC. Particle dispersion in a turbulent, plane, free shear layer.
[5] Njobuenwu DO, Fairweather M. Effect of shape on inertial particle dynamics in Phys Fluids 1989;A1(6):1035–44.
a channel flow. Flow, Turbul Combust 2014;92:83–101. [32] Johnson TA, Patel VC. Flow past a sphere up to a Reynolds number of 300. J
[6] van Wachem B, Zastawny M, Zhao F, Mallouppas G. Modelling of gas–solid Fluid Mech 1999;378:19–70.
turbulent channel flow with non-spherical particles with large Stokes [33] Leonard BP. A stable and accurate convective modeling procedure based on
numbers. Int J Multiphase Flow 2015;68:80–92. quadratic upstream interpolation. Comput Meth Appl Mech Eng
[7] Clift R, Grace JR, Weber ME. Bubbles, drops and particles. New York: Academic 1979;19:59–98.
Press; 1978. [34] Hölzer A, Sommerfeld M. Lattice Boltzmann simulations to determine drag, lift
[8] Allen T. Particle size measurement. Chapman & Hall; 1981. and torque acting on non-spherical particles. Comput Fluids 2009;38:572–89.
[9] Mandø M. Turbulence modulation by non-spherical particles. Aalborg [35] Richter A, Nikrityuk PA. Drag forces and heat transfer coefficients for spherical,
University; 2009. cuboidal and ellipsoidal particles in cross flow at sub-critical Reynolds
[10] Li T, Li S, Zhao J, Lu P, Meng L. Sphericities of non-spherical objects. numbers. Int J Heat Mass Transfer 2012;55:1343–54.
Particuology 2012;10(1):97–104. [36] Schiller L, Naumann A. Fundamental calculations in gravitational processing. Z
[11] Brix J, Jensen PA, Jensen AD. Modeling char conversion under suspension fired Vereines Deutscher Ingenieure 1933;77:318–20.
conditions in O2/N2 and O2/CO2 atmospheres. Fuel 2011;90:2224–39. [37] Cherukat P, McLaughlin JB, Dandy DS. A computational study of the inertial lift
[12] Powers MC. A new roundness scale for sedimentary particles. J Sediment on a sphere in a linear shear flow field. Int J Multiphase Flow
Petrol 1953;23:117–9. 1999;25(1):15–33.
[13] Chhabra R, Agarwal L, Sinha N. Drag on non-spherical particles: an evaluation [38] Legendre D, Magnaudet J. The lift force on a spherical bubble in a viscous linear
of available methods. Powder Technol 1999;101:288–95. shear flow. J Fluid Mech 1998;368:81–126.

Potrebbero piacerti anche