You are on page 1of 18

Computers & Fluids 38 (2009) 572–589

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

Lattice Boltzmann simulations to determine drag, lift and torque acting on


non-spherical particles
Andreas Hölzer, Martin Sommerfeld *
Mechanische Verfahrenstechnik, Zentrum für Ingenieurwissenschaften, Martin-Luther-Universität Halle-Wittenberg, 06099 Halle (Saale), Germany

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

Article history: The drag, lift and moment coefficients of differently shaped single particles have been determined as a
Received 27 July 2007 function of the angle of incidence at particle Reynolds numbers between Re = 0.3 and 240 under different
Received in revised form 15 February 2008 conditions. For this purpose simulations of the flow around these particles have been performed using the
Accepted 2 June 2008
three-dimensional Lattice Boltzmann method. In the first case studied a particle is fixed in a uniform flow,
Available online 8 June 2008
in the second case the particle is rotating in a uniform flow to determine, among others, the Magnus lift
force and in the third case the particle is fixed in a linear shear flow. In the first case six particle shapes are
considered, i.e. a sphere, a spheroid, a cube, a cuboid and two cylinders with an axis ratio of 1 and 1.5,
respectively. In the second and third case the sphere and the spheroid are considered. At the higher Re
considered, the drag depends strongly on particle shape, the angle of incidence and particle rotation.
The lift and the torque of both the sphere and the spheroid are strongly affected by particle rotation
and fluid shear. For approximately Re 6 1, the shear induced lift for unbounded flow could not be simu-
lated as the top and bottom wall have a significant influence in the current flow configuration. The shear
induced lift of the sphere changes direction at approximately Re = 50 and the mean (over the orientation)
shear induced lift of the spheroid changes direction at approximately Re = 90.
Ó 2008 Elsevier Ltd. All rights reserved.

1. Introduction The characteristic dimension of a particle (d) has been defined


as the diameter of the volume-equivalent sphere. The particle Rey-
The motion of particles plays an important role in both techni- nolds number is based on the relative velocity of the fluid with re-
cal and natural processes. Examples are combustion of pulverised spect to the particle. In our case the particle velocity is zero, thus
coal, fibre suspension flow in paper forming, pneumatic conveying the particle Reynolds number is
of granular materials, pollutant transport in the atmosphere and
jujd
transport of sediment grains in a river. Modelling of these pro- Re ¼ ; ð1Þ
cesses relies mostly on the assumption of non-rotating spherical t
particles in a uniform flow for the drag and on the assumption of with u being the undisturbed fluid velocity at the center of the par-
rotating spherical particles in a linear shear flow at low Re for ticle and t being the kinematic viscosity. The particle spin number
the lift and the torque. However, in practice, particles are not sub- (or dimensionless particle angular velocity) is defined as the ratio
ject to such ideal conditions and the particle motion is affected, between the peripheral velocity of the volume-equivalent sphere
among others, by fluid shear and particle rotation at high Re, par- and the undisturbed fluid velocity, i.e.
ticle shape and for non-spherical particles by particle orientation.
For simulation of the motion of such particles detailed information jxPa j 2d
SPa ¼ ; ð2Þ
on the forces acting on these particles are necessary, however, only juj
averaged correlations of the drag coefficient for different particle
shapes are available [1,2]. This work is, to the authors’ knowledge, where xPa is the particle angular velocity. Analogous to the particle
the first comprehensive three-dimensional study of the drag, lift spin number, the fluid spin number (or dimensionless fluid angular
and moment coefficients of non-spherical particles as a function velocity) is defined as
of the angle of incidence, allowing for particle rotation and linear jxFl j 2d
fluid shear for a sphere and a spheroid. SFl ¼ ; ð3Þ
juj
where xFl = 0.5r  u is the fluid angular velocity, i.e. half the fluid
velocity gradient in a linear shear flow. The drag force FD and the lift
* Corresponding author. Tel.: +49 3461 462879; fax: +49 3461 462878. force FL are the components of the fluid force acting parallel and
E-mail address: martin.sommerfeld@iw.uni-halle.de (M. Sommerfeld). perpendicular to the direction of the undisturbed fluid velocity. M

0045-7930/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compfluid.2008.06.001
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 573

is the torque acting on the particle. The drag, lift and moment coef- for the shear induced lift force in an inviscid fluid for very low SFl.
ficients are defined by Again, the lift acts towards higher velocities. In the zero Re limit,
the equation for torque on a rotating sphere in a uniform flow
FD
cD ¼ ; ð4Þ was determined by Kirchhoff [12] as
1
2
qu2 p4 d2
FL 3 SPa
cL ¼ ; ð5Þ M ¼ pd qtxPa or cM ¼ 16 : ð13Þ
1
qu2 p d2 Re
2 4
The torque acting on a stationary sphere in a shear flow was calcu-
and
lated by Faxén [13] and reads
M
cM ¼ ; ð6Þ 3 SFl
1
qu2 p4 d3 M ¼ pd qtxFl or cM ¼ 16 ; ð14Þ
2 Re
with q being the fluid density. i.e. at equilibrium xPa = xFl. All analytical equations mentioned in
The first analytical solution of the flow around a particle was this section are valid for unbounded flows only. Ladenburg [14] as
derived by Stokes [3] for the flow around a stationary sphere in a well as Faxén [13] determined theoretically and e.g. Fidleris and
uniform flow in the zero Re limit. For the drag force, Stokes Whitmore [15] and Sutterby [16] found experimentally that the
obtained influence of walls is to increase the drag.
There are many experimental studies of the drag of stationary
24
F D ¼ 3pdqvu or cD ¼ : ð7Þ particles in a uniform flow, mostly for spherical particles. From
Re
the beginning of the 20th century most of the experimental drag
Oberbeck [4] derived integral equations for the flow around a values were obtained by particle settling experiments at low Re
spheroid and calculated the drag of a prolate spheroid with the and by wind tunnel experiments at high Re. Numerical simulations
symmetry axis parallel to the flow. Oseen [5] calculated the drag of the flow around a stationary sphere in a uniform flow are exten-
of both prolate and oblate spheroids with arbitrary orientation, sively discussed in Section 3. Only very few three-dimensional
including the circular disk as a special case. The prolate spheroid numerical studies on the drag of stationary non-spherical particles
with the symmetry axis parallel to the flow and with an axis ratio in a uniform flow exist. Masliyah and Epstein [17], Pitter et al. [18]
of c/a = 1.95 has the overall smallest drag coefficient. Batchelor [6] and Comer and Kleinstreuer [19] investigated numerically the flow
derived an approximate expression for the drag of cylinders with past spheroids with the symmetry axis parallel to the flow for low
the symmetry axis both parallel and perpendicular to the flow, and moderate Re up to 120. Dwyer and Dandy [20] simulated the
which is based on slender-body theory and thus valid for long cyl- flow around prolate spheroids with c/a = 1.5, 2 and 3 if the symme-
inders only. try axis is at an angle of 0°, 45° and 90° with respect to the flow at
An additional lift force and a torque appear if the particle ro- 10 6 Re 6 100. For Stokes flow, Youngren and Acrivos [21] deter-
tates or if it is placed in a shear flow. In the case of a rotating par- mined numerically the drag of cylinders with 0.5 6 c/a 6 100 with
ticle in a uniform flow, the lift force is the well-known Magnus the symmetry axis parallel to the flow.
force. For low Re, Rubinow and Keller [7] obtained analytically Experimental and numerical cD values of spheres, disks and
the expression plates, lengthwise spheroids and streamline bodies, isometric par-
p 3 ticles such as cuboids, cylinders, tetrahedrons and octahedrons and
FL ¼ d qu  xPa or cL ¼ 2SPa ; ð8Þ irregularly shaped particles such as minerals are summarised in
8
Fig. 1. One can see the strong influence of Re, particle shape and
for the Magnus force on a sphere, where the expression for cL is only
particle orientation. Crosswise disks and plates have a large drag
valid if xPa and u are perpendicular. For low Re and Re  SFl, the
coefficient over the whole range of Re. Lengthwise disks and plates
analytical solution for the shear induced lift force on a sphere was
have a large drag coefficient at low Re, but a very small drag coef-
derived by Saffman [8] and is
ficient at high Re. Isometric bodies and minerals have a small drag
rffiffiffiffiffiffiffiffi
6:46 2 pffiffiffi dux coefficient at low Re and a small or an intermediate drag coefficient
F Lz ¼ d q t ux ; ð9Þ at high Re.
4 dz
For the case of a rotating sphere in a uniform flow, measure-
if u only has a component in the x-direction and a gradient in the z-
ments of the drag and lift force were conducted by Maccoll [47],
direction. The force acts towards higher relative velocities. More
Davies [48], Tani [49] and Tanaka et al. [50] in a wind tunnel at
generally, Saffman’s equation can be written as [9]
Re  105. Barkla and Auchterlonie [51], Tsuji et al. [52] and Oesterlé
sffiffiffiffiffiffiffiffiffiffi
and Dinh [53] determined the Magnus force from particle trajecto-
6:46 pffiffiffi 2 pffiffiffi 1 4 pffiffiffiffiffiffiffiffiffiffiffiffiffi
FL ¼ 2d q t u  xFl or cL ¼ 6:46 SFl =Re: ð10Þ ries at Re between 10 and 3000. The literature indicates that cL
4 jxFl j p
obeys Rubinow and Keller’s prediction (Eq. (8)) at low Re and is
McLaughlin [10] extended Saffman’s equation for arbitrary SFl/Re ra- about 0.5 at approximately SPa > 1.5 and Re > 500, i.e. it then does
tios p introducing the additional factor 9/(6.46p)J(e) with
byffiffiffiffiffiffiffiffiffiffiffiffiffi not depend or depends only weakly on Re and SPa. Sawatzki [54]
e ¼ 2 SFl =Re in Saffman’s equation carried out measurements and Dennis et al. [55] carried out simu-
rffiffiffiffiffiffiffiffi lations to determine cM of a rotating sphere in a quiescent fluid
9 2 pffiffiffi dux depending on Re. For Re < 10, cM agreed with Kirchhoff’s prediction
F Lz ¼ JðeÞd q t ux : ð11Þ
4p dz (Eq. (13)) and became larger than Kirchhoff’s prediction with
increasing Re. The lift force of a sphere in a linear shear flow at
J(e) is an integral which must be evaluated numerically. The values
Re < 1 was determined by Cherukat et al. [56] using a shear flow
of J(e) can be found in a table as well as in a diagram in [10]. The lift
apparatus. The lift force was consistent with McLaughlin’s predic-
in McLaughlin’s equation decreases more strongly than in Saffman’s
tion (Eq. (11)).
prediction with decreasing e up to negative values at small e. Auton
Numerical simulations of the flow around a rotating sphere in a
[11] derived the analytical solution
uniform flow, around a stationary sphere in a linear shear flow and
p 3 8 around a rotating sphere in a linear shear flow were performed by
FL ¼ d qu  xFl or cL ¼ SFl ð12Þ
6 3 Salem and Oesterle [57] and Kurose and Komori [58] at Re up to 40
574 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

5
10

4
cross- and length-
10 wise disks and plates
3
10 crosswise
disks and plates
2
10
cD

1
isometric
10 bodies and
minerals sphere
0
10
spheroids
-1
10 lengthwise plates, spheroids,
and streamline bodies
-2
10
10-3 10-1 101 103 105
Re
Fig. 1. Drag coefficient of differently shaped particles from the literature as a function of the particle Reynolds number. (- - - -) Stokes; (–) sphere [22–30]; (}) disks and plates
[25,28,30–36]; (X) isometric bodies [37–40]; (j) minerals [41–43]; (s) spheroids and streamline bodies [17,20,44–46].

and 500, respectively. Cherukat et al. [59] computed the lift force behaviour of fluids at the molecular level. Advantages of the Lattice
and the angular velocity and Bagchi and Balachandar [60] in addi- Boltzmann method are its numerical efficiency due to the simplic-
tion the drag force of a sphere in a linear shear flow at Re 6 2 and ity and the local character of the model equations and the direct
0.5 6 Re 6 200, respectively, where the sphere was either freely calculation of the forces acting on walls, obstacles or particles via
rotating or non-rotating. All four references mentioned above the momentum exchange of the molecules during impacts on a so-
found the shear induced lift force in good agreement with lid surface.
McLaughlin’s prediction (Eq. (11)) for low Re. At Re higher than The Lattice Boltzmann equation is a special finite difference
approximately 60, both Kurose and Komori [58] and Bagchi and form of the Boltzmann equation [63]. The phase space is discre-
Balachandar [60] obtained a negative shear induced lift force. The tised into cubic cells and 19 discrete velocities (D3Q19 model,
first numerical simulations of the flow around a stationary sphere see [64]), i.e. one velocity vector for molecules at rest (r = 0), six
in a linear shear flow were carried out by Dandy and Dwyer [61] at vertical or horizontal velocity vectors (r = 1) and 12 diagonal
Re 6 100. Their lift compared better with Saffman’s prediction (Eq. velocity vectors (r = 2). The velocity vectors are illustrated in
(9)) than with McLaughlin’s one (Eq. (11)) at low Re and a negative Fig. 2 and read
shear induced lift force was not observed. Cherukat et al. [59] as 8
< ð0; 0; 0Þ;
> r ¼ 0; i ¼ 1;
well as Bagchi and Balachandar [60] obtained a good agreement
of the angular velocity with Faxén’s prediction (Eq. (14)) at low eri ¼ ð1; 0; 0Þc; ð0; 1; 0Þc; ð0; 0; 1Þc; r ¼ 1; i ¼ 1 . . . 6;
>
:
Re and smaller angular velocities compared to Faxén’s equation ð1; 1; 0Þc; ð1; 0; 1Þc; ð0; 1; 1Þc; r ¼ 2; i ¼ 1 . . . 12
at higher Re. Nirschl [62] determined numerically the angular
velocity of a spheroid in a linear shear flow and also obtained smal- with c being the lattice constant which is the ratio of one time to
ler mean angular velocities with increasing Re. one space step (c = Dx/Dt). The Lattice Boltzmann equation with
In the present study, cD, cL and cM were calculated for various single time (BGK) relaxation [64,65] used in the present work reads
non-spherical particles using the Lattice Boltzmann method. The
article is organised as follows. The numerical method and setup
is described in Section 2. The results of the different flow regimes
around a sphere at Re 6 400 are presented in Section 3. In Section 2
4, cD, cL and cM of a stationary sphere, cube and cylinder with 1
c/a = 1 and of a stationary spheroid, cuboid and cylinder with 2 2
c/a = 1.5 in a uniform flow are given as a function of a and Re. In 2
Section 5, cD, cL and cM of a rotating sphere and spheroid in a
uniform flow are presented and in Section 6 cD, cL, and cM of a sta-
2
tionary sphere and spheroid in a linear shear flow are presented 1 1
depending on a, Re and SPa and SFl, respectively. Conclusions are 0
drawn in Section 7. 1 2
2
2. Numerical method

The fluid flow is simulated by the Lattice Boltzmann method


2 1 2
which is an alternative approach to conventional methods for
numerically solving the conservation equations. Whereas conven- 2
tional models are based on the conservation laws formulated at the
macroscopic level, the Boltzmann equation describes the Fig. 2. Discrete velocity vectors in the D3Q19 model.
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 575

Dt the side walls, which would yield a uniform flow along the compu-
fri ðx þ eri Dt; t þ DtÞ  fri ðx; tÞ ¼  ðfri ðx; tÞ  freqi ðx; tÞÞ; ð15Þ
s tational domain without a particle. The periodic boundary condi-
where fri(x,t) is the discrete distribution function representing the tion can in general not be applied at the side walls, since with
probability of finding a number of molecules with the velocity eri periodic boundary conditions a non-axisymmetric flow would turn
at the position x and the time t, freqi ðx; tÞ is the discrete equilibrium upwards or downwards directly after the inlet. In the case of the
distribution function and s is the relaxation time. This equation can linear shear inflow the moving no-slip wall boundary condition
be split into two steps, namely the relaxation step is imposed at the top and bottom wall, yielding a linear shear flow
along the computational domain without a particle. The axially
Dt  ð0Þ

stress-free boundary condition, i.e. oux/ox = 0, is applied at the out-
fri ðx; t þ Þ ¼ fri ðx; tÞ  fri ðx; tÞ  fri ðx; tÞ ð16Þ
s let [66].
and the propagation step The curved no-slip boundary condition introduced by Bouzidi
et al. [67] is imposed on the particle surface. This boundary condi-
fri ðx þ eri Dt; t þ DtÞ ¼ fri ðx; tþ Þ: ð17Þ
tion considers the exact position of the particle surface within a
In the D3Q19 model the discrete equilibrium distribution function cell through the parameters qri. The parameters qri are illustrated
is in Fig. 4. They indicate the relative distance between the particle
! surface and the next fluid node in ri-direction. The fluid force act-
3eri  uðx; tÞ 9ðeri  uðx; tÞÞ2 3u2 ðx; tÞ ing on the particle is directly evaluated from the momentum ex-
freqi ðx; tÞ ¼ xr qðx; tÞ 1 þ þ 
c2 2c4 2c2 change between the fluid and the particle surface [67,68]. The
8 momentum exchange at one node at time t + 0.5 Dt is
< 1=3;
> r ¼ 0; i ¼ 1;
xr ¼ 1=18; r ¼ 1; i ¼ 1 . . . 6; pri0 ðx; t þ 0:5 DtÞ ¼ fri ðx; t þ DtÞeri  fri0 ðx; t þ Þeri0 ;
>
: ð22Þ
1=36; r ¼ 2; i ¼ 1 . . . 12: pri0 ðx; t þ 0:5 DtÞ ¼ ðfri ðx; t þ DtÞ þ fri0 ðx; tþ ÞÞeri0
ð18Þ
and the resulting counterforce is
Density and momentum are obtained through the equations DV
XX F ri0 ðx; t þ 0:5 DtÞ ¼ ðfri ðx; t þ DtÞ þ fri0 ðx; tþ ÞÞeri0 ; ð23Þ
qðx; tÞ ¼ fri ðx; tÞ ð19Þ Dt
r i
with eri0 pointing to the particle surface and eri pointing in the
and opposite direction of eri0 .
XX The domain size is 14.5l  5l  6.2l cells, where l is the smallest
qðx; tÞuðx; tÞ ¼ eri fri ðx; tÞ: ð20Þ dimension of the particle. The value of l is 12 cells for Re 6 240 and
r i
greater than 12 cells for Re > 240 up to 24 cells at Re = 480, which
The relaxation time determines the viscosity via the relation allows for a reasonable resolution of the flow field around the par-
ticle. The coefficients converge to 99% of the terminal value at
1 2
v¼ c ð2s  DtÞ: ð21Þ Re = 0.3 after 10,000 time steps and at Re = 240 after 4000 time
6
steps if a steady solution exists.
The BGK Lattice Boltzmann method is limited to low dimensionless The influence of the boundaries of the computational domain
velocities of about u < 0.1 and high dimensionless viscosities of and of the particle resolution on the drag, lift and torque were cal-
about m > 1/200. With Eq. (1) one obtains Re < 20d, i.e. with a diam- culated for different particle shapes depending on Re; for more de-
eter of 12 cells a maximum Reynolds number of 240 can be tails see [69]. These calculations were used to correct the drag, lift
simulated. and torque to obtain the values for unbounded flow. Fig. 5 exempl-
In the present simulation, a particle with an axis ratio c/a is arily shows cD of a fixed sphere in a uniform flow as a function of
either fixed at a certain angle of incidence a or allowed to rotate the number of cells for discretizing the diameter of the particle at
with a constant angular velocity xPa in the centre of the cross-sec-
tional area of the domain, see Fig. 3. The angle of incidence is the
angle between the undisturbed flow direction and the symmetry
axis of the particle. In all cases the symmetry axis of the particle e21
is in the x–z plane; thus only position variations in the x–z plane
are considered. As inflow (x = 0) a uniform flow or a linear shear e11
flow with the angular velocity xFl is assumed. In the case of the
uniform inflow, the symmetry boundary condition is imposed at e22
21
symmetry boundary condition (uniform inflow)
moving no-slip wall boundary condition (linear shear inflow)
FL q11
c
uniform
flow ωFl ωPa M stress-free
22

or boundary fluid solid


α FD condition
linear shear
flow a
u
z

y
x

Fig. 3. Flow configuration. Fig. 4. Parameters qri of the curved boundary condition.
576 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

1.5 1.6
Present simulation Theoretical (Vasseur
1.2 Re=0.012
1.4 Re=0.3 and Cox [70])
Re=30 Re=0.12
0.8
Normalized cD

Re=90 Re=0.3
1.3 Re=240 Re=1.2
0.4

cL
1.2 0.0

-0.4
1.1
-0.8
1.0
5 10 15 20 25 -1.2
0 2 4 6 8 10 12 14 16 18 20
Diameter in cells
Distance between walls L/d
Fig. 5. Normalised drag coefficient of a sphere as a function of the resolution.
Fig. 7. Lift coefficient of a sphere in the center between two walls in a linear shear
flow as a function of the distance between the walls at SFl = 0.025.

different Re. The influence of particle resolution on the error in cD


increases with increasing Re. In case the particle is resolved by where L is the distance between the two walls. Vasseur and Cox’s
12 cells (as done here for Re 6 240) the maximum error in cD for prediction is also plotted in Fig. 7 and agrees very well with the
Re = 240 is about 11%. The strongest influence is that of the dis- present results at Re = 0.12 and 0.3. However, the curve for
tance between the particle and the side walls on the lift force. Re = 0.012 runs below that of the prediction in the L/d range plotted,
Fig. 6 shows cL of a rotating sphere in a uniform flow as a function but seems to fit the prediction at very high L/d. The reason why the
of the normalised distance between the side walls at different Re. graph of Re = 0.012 has the highest discrepancy to Vasseur and Cox’
The influence of the boundaries decreases with increasing Re. prediction, although it has the lowest Re and therefore is expected
In all cases, except that of the lift force in a linear shear flow as a to agree better, is not clear and requires further investigations,
function of the distance between the side walls, the coefficients ap- which however are beyond the scope of this paper. For Re P 1.2,
proach an asymptotic value with increasing computational domain the functions are asymptotic. Thus, for Re P 1.2, also the side walls
as well as with increasing particle resolution at every Re, see Figs. 5 with the moving no-slip wall boundary condition have no influence
and 6 In the case of the linear shear flow, i.e. for the moving no-slip on the lift force at large distances between the side walls.
wall boundary condition at the top and bottom wall, the influence
of the distance between the top and bottom wall on the lift coeffi- 3. Flow around a sphere
cient is plotted in Fig. 7 for different Re. At Re = 0.012, 0.12 and 0.3
the lift increases almost linearly with increasing distance. Thus, at Since the stability computations of Natarajan and Acrivos [71] it
low Re, the influence of the top and bottom wall is significant for is generally accepted that the flow around a sphere undergoes a
very large distances also. This behaviour was predicted theoreti- steady non-axisymmetric regime between the steady axisymmet-
cally by Vasseur and Cox [70] for a sphere suspended in a linear ric and the time-dependent regime. The different flow regimes
shear flow between two walls. If the sphere is in the centre be- are in detail as follows. At low Re, the flow around a sphere is stea-
tween the two walls, they obtained m01 =ðRe U m Þ ¼ 0:036 (Fig. 4 dy, axisymmetric and separation does not occur (region I). The first
in [70]) or change in the flow regime appears at Re1, where the flow becomes
cL ¼ 48  0:036  SFl L=d; ð24Þ separated, but remains steady and axisymmetric (region II). With
increasing Re, the separations length and angle grow. The second
transition occurs at Re2, where the flow changes from axisymmet-
ric to planar symmetric, but still exhibits steadiness (region III). At
the third transition at Re3, the flow becomes temporally periodic
9 without vortex shedding. The flow remains planar symmetric (re-
8 Re=80 gion IV). The fourth transition appears at Re4, where temporal peri-
Re=12 odic vortex shedding arises. The flow still exhibits planar
7 Re=1.2 symmetry (region V). At the fifth transition at Re5, irregularly
Normalized cL

Re=0.024 shaped vortices appear and thus time periodicity and planar sym-
6
Re=0.003 metry are lost (region VI). There are two more flow regimes which
5 Re=0.0012 should also be mentioned. At Re  800, the wake becomes turbu-
lent, i.e. smaller scales of fluid motion arise [72,73]. At the well
4 known critical Reynolds number of approximately 2.5  105, the
3 drag drops abruptly to less than half [25,26,29,74] due to the
change of the boundary layer flow from laminar to turbulent [75].
2 Saha [76] observed numerically that the flow regimes and the
1 corresponding transition Reynolds numbers of the flow past a cube
are very similar to that past a sphere, i.e. it also undergoes a steady
2 4 6 8 10 12 14 16 18
symmetric, a steady planar symmetric and an unsteady planar
Distance between side walls L/d
symmetric regime. Some authors observed periodic motion of
Fig. 6. Normalised lift coefficient of a rotating sphere as a function of the distance freely falling disks [28,36,77] and long cylinders [78] from approx-
between the side walls. imately Re = 100 and chaotic motion at higher Re indicating a
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 577

regime of periodic and chaotic flow, respectively, for these particle 2


10
shapes also.
In the present numerical simulation the sphere is fixed in a uni- Present simulation
form flow at Re between 200 and 400, where Reynolds numbers Experimental data
are considered with an increment of 40. Therefore, the exact values Stokes
of the transition Reynolds numbers are not computed and only a 10
1

range within 40 Reynolds numbers can be stated. Experimental

cD
and numerical values of the transition Reynolds numbers Re1,
Re2, . . . , Re5, including the present values, are summarised in Table
1. The ranges of the present transition Reynolds numbers agree
excellently with the literature. The different flow regimes for 10
0

Re 6 400, experimental cD values and the present results are plot-


ted in Fig. 8 versus Re. The experimental values and the present
results of cD agree very well with each other. 0 1 2 3
The temporal behaviours of cD, cL and cM at Re = 200, 240, 280, 10 10 10 10
320, 360 and 400 including each an image of the flow field around Re
the sphere are plotted in Fig. 9. The initial condition is a uniform I II III IV V VI
flow along the computational domain. Thus, cD is very large and Re1 Re2 Re3 Re4 Re5
cL and cM are zero at t = 0. The temporal behaviour agrees perfectly
with the literature mentioned above for the different Reynolds I Steady, non-separated, axisymmetric
numbers. At Re = 200 (region II), the flow is steady and axisymmet- II Steady, separated, axisymmetric
ric with a detached vortex ring behind the sphere. Lift and torque III Steady, separated, planar symmetric
are zero. At Re = 240 (region III), the flow has a transition point at IV Time-periodic, no vortex shedding, planar symmetric
the dimensionless time tu/d  1000 where a lift and torque appear V Time-periodic vortex shedding, planar symmetric
and the drag abruptly increases. The flow is then not axisymmetric, VI Time-chaotic vortex shedding, non-symmetric
but symmetric in the y = z plane, which lies diagonally in the chan- Fig. 8. Comparison between experimental data [22,23,27,28,30] for the drag
nel. At Re = 280 (region IV), the flow is the same as at Re = 240 ex- coefficient of a sphere and present simulation as a function of the particle Reynolds
cept that small periodic oscillations appear. At Re = 320 (region V), number including the different flow regimes.
vortex shedding occurs and, therefore, the coefficients oscillate
periodically. The flow remains almost planar symmetric in the
y = z plane. At Re = 360 and 400 (region VI), the flow is irregular and the torque has only a component in y-direction. The lift force
in time and space. All calculations are axisymmetric and steady is positive in the z-direction and the torque is positive in the y-
in the beginning; calculations with a non-axisymmetric or an un- direction, see Fig. 3.
steady solution show a sharp transition to the other stage at a cer-
tain time. This transition develops automatically due to numerical 4. Stationary particle in a uniform flow
noise. It should be pointed out that all data of cD, cL and cM which
will be presented in Sections 4–6 were obtained before reaching The values of cD, cL and cM of six differently shaped particles, i.e.
the possible existing transition time because of the long computa- the sphere, the cube, the cylinder with c/a = 1 and the spheroid, the
tion time until a transition, if any, occurs. Furthermore, all flows cuboid and the cylinder with c/a = 1.5 are computed at Re = 0.3, 30,
considered in Sections 4–6 are symmetric with respect to the x– 90, 240 and for certain orientations also at Re = 480. In Fig. 10, cD is
z-plane. It should be noted that this is not due to any numerical plotted as a function of a at Re = 0.3, 30, 90 and 240. The theoretical
limitations. The lift force has only a component in z-direction curve for the sphere and the spheroid for Stokes flow are added to

Table 1
Experimental (upper table) and numerical (lower table) values of the particle Reynolds numbers at the transition between the different flow regimes past a sphere

Author Re1 Re2 Re3 Re4 Re5


Experimental transition Reynolds numbers
Taneda [79] 24
Magarvey and Bishop [80] 210 270 290
Wu and Faeth [81] 200 280
Provansal and Ormières [82] 180 280 310 360
Goldburg and Florsheim [83] 270
Sakamoto and Haniu [73] 280 300 420
Schouveiler and Provansal [84] 270
Numerical transition Reynolds numbers
Hamielec et al. [85] 22
Pruppacher et al. [86] 20
Magnaudet et al. [87] 20
Johnson and Patel [88] 20 211 270. . .280
Dennis and Walker [89] 20.5
Shirayama [90] 19
Natarajan and Acrivos [71] 210 277.5
Thompson et al. [91] 212 272
Tomboulides and Orszag [92] 212 270. . .285 285. . .300 <500
Ghidersa and Dušek [93] 272.3
Mittal [94] 350. . .375
Present Results 200. . .240 240. . .280 280. . .320 320. . .360
578 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

0.760 0.695
Re=200 Re=240 cD 0.032
cD
cLy
0.755 0.690 0.030

cL
cLz 0.028

cL, cM
0.750 0.685
cD

cD
cMy
0.0005
0.745 0.680

cM
cL, cM
0.0 cMz 0.0000
0.740 0.675 -0.0005
0 500 1000 1500 2000 0 400 800 1200 1600
t u/d t u/d

0.65 0.63
Re=280 cD cD
0.64 0.060 Re=320 0.08
0.60 cLy
cLy
0.63 0.06
0.045

cL
0.57

cL
0.62 0.04
cLz 0.030
cD

0.54

cD
0.61 cLz
cMy cMy
0.60 0.0004 0.51 0.0004

cM
cM 0.0000
0.59 cMz 0.0000 0.48
cMz -0.0004
0.58 -0.0004 0.45
0 150 300 450 600 0 80 160 240 320
t u/d t u/d

0.63 0.60
Re=360 cD cD
0.12 Re=400 0.12
cLy 0.54
0.56 cLy
0.08 0.08
cL

0.48

cL
0.04 0.04
0.49 0.42
cD

cD

cMy cLz cMy


0.0005 0.36 cLz 0.0006
cM

cM
0.42
0.0000 0.30 0.0000
cMz -0.0005 cMz
0.35 -0.0006
0.24
0 50 100 150 200 250 0 60 120 180 240
t u/d t u/d

Fig. 9. Temporal behaviour of drag, lift and moment coefficient of a sphere at Re = 200, 240, 280, 320, 360 and 400.

the diagram for Re = 0.3. For building the theoretical curve of cD with Re is the increasing influence of the form drag with increasing
versus a, the theoretical solution of Oseen [5] for lengthwise Re due to the increasing pressure change around the body. As con-
(a = 0°) and crosswise (a = 90°) spheroids and the Stokes assump- clusion, the streamlining has a greater influence on the drag at low
tion that the equations are linear in u were used. Both theoretical rather than at high Re and the size of the projected area has a great-
curves agree very well with the present results for the sphere er influence at high rather than at low Re. The lengthwise spheroid
and the spheroid; the present values are only slightly larger, prob- has the smallest drag of all particles at every Re considered. More-
ably because the Reynolds number of 0.3 is still slightly too high over, at Re P 90, all lengthwise particles with c/a = 1.5 have a
for the Stokes assumption. According to the Stokes assumption, smaller drag than the sphere because of their small projected
for the cube, cD is constant over a. However, the present results areas. The cuboid or the cube have the maximum drag at every
show that the cube has small oscillations in cD at Re = 0.3. Re considered due to their sharp edges. So far the only available
The cube has a local minimum in cD at a = 45° and the cuboid data from literature of the drag of a three-dimensional non-spher-
and both cylinders have an inflection point around 45° at ical body as a function of a are the numerical results from Dwyer
Re = 0.3. The reason for this is that these bodies have the best and Dandy [20] for a spheroid at a = 0°, 45° and 90° and
streamline shape at a = 45°. However, the variations in cD are rela- 10 6 Re 6 100. Their cD values show an almost linear slope be-
tively small at Re = 0.3. It should be pointed out that the area in the tween a = 0°, 45° and 90°, which agrees very well with the present
definition of cD (Eq. (4)) is the area of the volume-equivalent results for the spheroid.
sphere and therefore independent of a. For Re P 30, the local min- In Fig. 11, cD of the particles in lengthwise and crosswise posi-
imum and the inflection points of the cube, the cuboid and the tion is plotted versus Re. One can see that the scatter in the cD val-
cylinders at a = 45° have disappeared. Moreover, at Re P 30, all ues increases strongly with increasing Re, i.e. the drag is more
non-spherical particles have local minima in cD at orientations influenced by particle shape and orientation at high than at low Re.
where the projected area reaches local minima, i.e. at a = 0° and, Generally, for non-spherical particles a lift and a torque arise as
except for the spheroid, at 90°. In addition, at Re P 30, the non- the flow and pressure distribution around the body is not axial-
spherical particles have the maximum drag at approximately that symmetric, which is demonstrated in Fig. 12 for the cuboid at
orientation where the projected area reaches the greatest value, a = 45° and Re = 90. More precisely, the high pressure on the up-
e.g. the cube at a = 45°. The reason for the change of the behaviour stream large side face and the low pressure on the downstream
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 579

105 3.0
Re=0.3 Re=30
100 2.8

2.6
95
cD 2.4

cD
90
2.2
85 2.0

80 1.8
0 15 30 45 60 75 90 0 15 30 45 60 75 90
α [˚] α [˚]

1.6 1.2
Re=90 Re=240
1.5 1.1
1.4 1.0
1.3 0.9

cD
cD

1.2 0.8
1.1 0.7
1.0 0.6
0.9 0.5
0 15 30 45 60 75 90 0 15 30 45 60 75 90
α [˚] α [˚]

Fig. 10. Drag coefficient of stationary particles of different shape in a uniform flow as a function of the angle of incidence at Re = 0.3, 30, 90 and 240. (—) Sphere; (5) spheroid
c/a = 1.5; (h) cube; (s) cuboid c/a = 1.5; (}) cylinder c/a = 1; (/) cylinder c/a = 1.5; (- - -) theoretical sphere and spheroid for Stokes flow using Oseen’s solution [5] for 0° and 90°.

103
Sphere Cube 0˚
Spheroid c/a=1.5 0˚ Cuboid c/a=1.5 0˚
Spheroid c/a=1.5 90˚ Cuboid c/a=1.5 90˚
102
Cylinder c/a=1 0˚
Cylinder c/a=1 90˚
101 Cylinder c/a=1.5 0˚
cD

Cylinder c/a=1.5 90˚

100
Stokes sphere

10-1
10-1 100 101 102 103
Re
Fig. 11. Drag coefficient of stationary particles of different shape in a uniform flow
as a function of the particle Reynolds number.
Fig. 12. Velocity and pressure field around a stationary cuboid in a uniform flow at
large side face lead to a lift in downward direction in Fig. 12. Fig. 13 a = 45° and Re = 90.
shows cL depending on a at Re = 0.3, 30, 90 and 240. The cube expe-
riences no lift according to the Stokes assumption. The present present result for the spheroid. Furthermore, all curves for the par-
simulation shows that the cube has a very small lift of maximum ticles with c/a = 1.5 show an almost parabolic shape with the min-
0.1% compared to the drag at Re = 0.3. However, cL of the cube imum at about a = 45° at Re = 0.3. The graph of the spheroid
reaches values up to 3% of cD at Re = 30, 8%, at Re = 90 and 9% at remains nearly parabolic also at Re = 30 and 90. With increasing
Re = 240. The reason for this increase of lift is that the lift of the Re, the shape of the cL curves of the cuboid and the cylinder with
cube is almost only due to the pressure and the pressure change c/a = 1.5 becomes asymmetric and the minimum shifts towards
around a body increases with increasing Re. The cylinder with smaller a. The reason is the change of the flow and the pressure
c/a = 1 behaves similar to the cube since its shape is similar to that distribution around the bodies, which are almost symmetric
of the cube. around the particle center at every a at Re = 0.3, but in general
The theoretical curve for the spheroid, derived again from non-symmetric at Re P 30.
Oseen’s solution for lengthwise and crosswise spheroids and from At Re = 240, periodic flow separations behind the non-spherical
the Stokes assumption, is added to the diagram for Re = 0.3. The particles appear already from the beginning of the computations at
theoretical curve has a parabolic shape and agrees well with the orientations where the projected area reaches the respective great-
580 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

1 0.1
Re=0.3 Re=30
0
0.0
-1

-2 -0.1
cL

cL
-3
-0.2
-4

-5 -0.3
0 15 30 45 60 75 90 0 15 30 45 60 75 90
α [˚] α [˚]

0.2 0.2
Re=90 Re=240
0.1 0.1 periodic flow
separations
0.0 0.0
cL

cL
-0.1 -0.1

-0.2 -0.2

-0.3 -0.3
0 15 30 45 60 75 90 0 15 30 45 60 75 90
α [˚] α [˚]

Fig. 13. Lift coefficient of stationary particles of different shape in a uniform flow as a function of the angle of incidence at Re = 0.3, 30, 90 and 240. Symbols as in Fig. 10.

est values, yielding an increase of cL. As mentioned in Section 3, it son to the drag force with increasing Re; e.g. the maximum cL of
was not waited for the possible occurrence of a flow transition at a the spheroid reaches almost 30% of cD at Re = 240.
later point of time, where e.g. flow oscillations could arise for other The torque on the non-spherical particles with c/a = 1.5 acts,
orientations also. with few exceptions, to turn the particle in the crosswise position,
The maximum cL values are plotted versus Re in Fig. 14. It i.e. in the anticlockwise direction at 0° 6 a 6 90°. The reason for
should be noted that the mean values (over a) of cL, as later in this is the high pressure at the front stagnation point (see
the text also the mean values of cM, which perhaps initially appear Fig. 12). The behaviour of cM with respect to a plotted in Fig. 15
to be the more meaningful values, are not plotted because they are is very similar to that of cL except that the minima of the cuboid
zero if averaged over 180°. The maximum cL values of the particles and the cylinder with c/a = 1.5 shift towards larger a with increas-
with c/a = 1.5 decrease strongly between Re = 0.3 and Re = 30. Since ing Re and that the periodic flow separations at Re = 240 have no
the cube and the cylinder with c/a = 1 have a very small lift at observable influence.
Re = 0.3, they only experience a small change in cL between A particle is in a stable position if cM is zero and if the slope of
Re = 0.3 and Re = 30. For Re P 30, all cL values remain nearly con- the curve of cM versus a is positive. The particles with c/a = 1.5 are
stant. A comparison between Figs. 14 and 11 shows that for every in a stable position at the crosswise position (a = 90°) and the cu-
particle shape the influence of the lift force increases in compari- boid at Re = 240 also at the lengthwise position (a = 0°). For a
spheroid in an inviscid fluid, the crosswise position was also pre-
dicted theoretically as the only stable orientation [95]. For the
1 cube, stable positions are the equivalent positions at a = 0° and
10
90°. The cylinder with c/a = 1 is in a stable position at a = 90° and
Spheroid c/a=1.5 at Re = 90 and 240 it is also stable at 0°. This behaviour of the par-
Cube ticles agrees with the observation of many authors [28,36–
Cuboid c/a=1.5 38,77,78,96] that for steady motion and approximately Re > 1 sym-
Cylinder c/a=1 metrical particles fall with the maximum principal cross-sectional
0
Cylinder c/a=1.5 area normal to the direction of motion (crosswise position).
10
-cL

Fig. 16 shows the maximum cM values of the different particles


versus Re. According to the Stokes assumption, no torque acts on a
particle with three perpendicular symmetry planes [97], i.e. on all
particles considered here. This behaviour was confirmed in settling
experiments at low Re, where non-spherical particles retained the
-1 orientation in which they were released [28,37,38,96]. According
10 to the present simulation, the ratio between cM and cD is less than
-1 0 1 2 3 0.4% at Re = 0.3, i.e. the rotation is negligible compared to the trans-
10 10 10 10 10
lation. The maximum cM values of the particles with c/a = 1.5 de-
Re
crease consistently with increasing Re as shown in Fig. 16, but
Fig. 14. Maximum lift coefficient of stationary particles of different shape in a the slope is much smaller than for cD (Fig. 11). Therefore, the influ-
uniform flow as a function of the particle Reynolds number. ence of cM increases significantly with increasing Re; e.g. cM of the
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 581

0.1 0.05
Re=0.3 Re=30
0. 0.00

-0.1 -0.05

cM
cM
-0.2 -0.10

-0.3 -0.15

-0.4 -0.20
0 15 30 45 60 75 90 0 15 30 45 60 75 90
α [˚] α [˚]

0.05 0.08
Re=90 Re=240
0.04
0.00
0.00
-0.05

cM
cM

-0.04
-0.10
-0.08

-0.15 -0.12
0 15 30 45 60 75 90 0 15 30 45 60 75 90
α [˚] α [˚]

Fig. 15. Moment coefficient of stationary particles of different shape in a uniform flow as a function of the angle of incidence at Re = 0.3, 30, 90 and 240. Symbols as in Fig. 10.

0 pressure field around the rotating sphere with SPa = 1 at Re = 90.


10
The streamlines from the bottom of the sphere move along with
the sphere upwards which results in an upward deflection of all
-1 streamlines behind the rotating sphere. At Re 6 240, no wake
10 structure appears behind the sphere except for Re = 240 and
SPa = 2. More precisely, in this case vortex shedding, which starts
from the top of the sphere, occurs downstream of the sphere. Thus,
-2
10 the particle spin number of 2 can be seen as a kind of ‘‘resonance”
-cM

angular velocity for Re = 240. Furthermore, there is only one stag-


nation point around the sphere, which is no longer situated on
10
-3 Spheroid c/a=1.5 the particle surface, since the surface has a velocity. The stagnation
Cube Cuboid c/a=1.5 point is situated above the top of the sphere at Re = 0.3 and moves
Cylinder c/a=1 Cylinder c/a=1.5
-4
10 -1 0 1 2 3
10 10 10 10 10
Re
Fig. 16. Maximum moment coefficient of stationary particles of different shape in a
uniform flow as a function of the particle Reynolds number.

spheroid reaches 16% of cD at Re = 240. For the same reason as ex-


plained above for cL, cM of the cube and of the cylinder with c/a = 1
are negligible at Re = 0.3, but reach values up to 5% of cD at
Re = 240.

5. Rotating sphere and spheroid in a uniform flow

The sphere and the spheroid with c/a = 1.5 rotate in each case
with four different angular velocities (SPa = 0.5, 1, 2 and 3) in a uni-
form flow at Re = 0.3, 30, 90 and 240. For the sphere these four
angular velocities also apply at Re = 0.003 and 480. Due to particle
rotation the flow around a sphere is no longer symmetric in the
plane of rotation which results in an additional transverse lift,
the so-called Magnus force, and in an additional torque. This is Fig. 17. Streamlines and pressure field around a rotating sphere with SPa = 1 in a
demonstrated in Fig. 17 by means of the streamlines and the uniform flow at Re = 90.
582 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

upstream with increasing Re. The maximum pressure point is also although more accentuated, the same behaviour as cD of the
located at the top hemisphere, but no longer coincides with the sphere.
stagnation point. As one can see in Fig. 17, an area of low pressure Fig. 19 shows cD of the sphere and the spheroid as a function of
is situated at the bottom of the sphere causing the Magnus lift act- a at Re = 0.3, 30, 90 and 240. The curves of cD of the spheroid versus
ing towards the bottom. In addition, the lift due to viscous stress a change with particle rotation, since the rotating spheroid perma-
also acts towards the bottom at every Re considered, but is appre- nently changes its position and thus areas continuously switch be-
ciable smaller than the lift due to pressure. tween fluid and solid. Therefore, with particle rotation the areas of
The cD values of the sphere, plotted versus Re in Fig. 18, increase low and high pressure also continuously change their location. The
with increasing SPa up to SPa = 2 and become smaller again for particle rotation has major effects on the variation of the drag of
SPa = 3 at every Re considered. However, at Re = 0.3, the difference the spheroid with respect to the orientation, which increases
in sphere drag for the different SPa is negligible, but becomes larger strongly with increasing SPa and with increasing Re, i.e. it is maxi-
with increasing Re. The mean values of cD of the spheroid show, mal for Re = 240 and SPa = 3. The reason for this is the increase of
the pressure change around the spheroid with increasing Re as well
as with increasing SPa. At Re = 90 and 240, the spheroid with SPa = 3
experiences a negative drag at approximately 120° < a < 190° since
3
10 an area of high pressure exists behind the spheroid. In addition, the
SPa=0 particle rotation affects, although less strong, the shape of the cD
curves. Without rotation, the spheroid experiences the minimum
2 SPa=0.5 in cD at a = 0° and the maximum at a = 90°. With rotation, both
10
SPa=1 minimum and maximum shift towards smaller a.
SPa=2 Fig. 20 shows cL of the sphere, i.e. the Magnus lift coefficient,
1 and the mean values of cL of the spheroid as a function of Re. An
10 SPa=3
increase of SPa causes an increase of the absolute lift value of the
cD

sphere at every Re considered. The cL of the sphere fits the theoret-


ical solution equation (8) very well at low Re, i.e. for constant SPa, cL
0
10 remains constant over Re. Beyond Re of about 0.3, the calculated cL
becomes smaller than the theoretical solution for low Re. The cL
Stokes sphere values of SPa = 1, 2 and 3 are almost identical at the value of
-1 approximately 0.7 at Re = 240 and at the value of 0.6 at Re = 480,
10
10-1 100 101 102 103 which agrees with the literature [47–51,53]. The reason for this
Re behaviour is the occurrence of an additional area of low pressure
for Re P 240 and high SPa at the top of the sphere. Therefore, at
Fig. 18. Drag coefficient of a rotating sphere with SPa = 0, 0.5, 1, 2 and 3 in a uniform Re P 240, the lift does not increase infinitely with increasing SPa
flow as a function of the particle Reynolds number. but approaches an asymptotic value. As one can also see from

93 3.5
SPa Re=0.3 SPa Re=30
3.0
90
2.5
87 2.0
cD
cD

84 1.5
1.0 SPa
81
0.5
SPa
78 0.0
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

3 3
SPa Re=90 SPa Re=240
2 2

1
1
cD
cD

0
0
SPa -1
SPa
-1 -2
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

Fig. 19. Drag coefficient of a rotating sphere and spheroid (c/a = 1.5) with SPa = 0, 0.5, 1, 2 and 3 in a uniform flow as a function of the angle of incidence at Re = 0.3, 30, 90 and
240. (- - -) Sphere; (—) spheroid.
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 583

101 10 5
Sphere SPa=0.5 SPa=1
10 4
SPa=2 SPa=3
10 3 Mean value spheroid c/a=1.5
SPa=0.5 SPa=1
10 2 SPa=2 SPa=3
100

cM
-cL

Mean value 10 1
spheroid c/a=1.5 Sphere
SPa=0.5 SPa=0.5
10 0
SPa=1 SPa=1 Theoretical sphere
Theoretical sphere
SPa=2 SPa=2 (Rubinow and 10 -1 (Kirchhoff [12])
SPa=3 SPa=3 Keller [7])
10-1 10 -2
10-3 10-1 101 103 10
-3
10
-2
10
-1
10
0
10
1
10
2
10
3

Re Re
Fig. 20. Lift coefficient of a rotating sphere and mean lift coefficient of a rotating Fig. 22. Moment coefficient of a rotating sphere and mean moment coefficient of a
spheroid (c/a = 1.5) with SPa = 0.5, 1, 2 and 3 in a uniform flow as a function of the rotating spheroid (c/a = 1.5) with SPa = 0.5, 1, 2 and 3 in a uniform flow as a function
particle Reynolds number. of the particle Reynolds number.

Fig. 20, the mean values of cL of the spheroid behave similar to cL of


the sphere. cM of the sphere fits Kirchhoff’s theoretical prediction (Eq. (13))
The variations of cL of the sphere and the spheroid depending on very well. With increasing Re, cM becomes larger than the theoret-
a are plotted in Fig. 21. At Re = 0.3, the curves for the spheroid are ical solution. Sawatzki [54] and Dennis et al. [55] also observed a
almost parallel shifted to smaller cL and smaller a with increasing larger cM for high angular velocities of a sphere in a quiescent fluid
SPa. Similarly as for cD, the relative variation of cL over a increases compared to the theoretical values. The torque on a sphere is only
drastically with increasing SPa at Re = 30, 90 and 240. Again, the caused by the viscous wall shear stress on the particle surface. The
reason is the rise of the pressure change around the spheroid with velocity gradients and thus the viscous shear stresses are larger at
increasing SPa. The minimum lift at a  30° is attributed to the high Re than at low Re, resulting in a larger cM of the sphere at high
occurrence of an area of high pressure at the top of the spheroid Re compared to the corresponding value for low Re. The mean val-
and the maximum lift at a  120° to the occurrence of an area of ues of the spheroid are significantly larger than the values of the
low pressure at the top of the spheroid. sphere, since for non-spherical particles the pressure distribution
The behaviour of cM of the sphere and of the mean values of cM around the body also has a contribution in cM and acts in the same
of the spheroid are plotted in Fig. 22 as a function of Re. At low Re, direction as the viscous wall shear stress, i.e. in the direction oppo-

8 1
Re=0.3 Re=30
4 0

0 -1

-4 -2 SPa
cL

cL

SPa
-8 SPa -3
SPa
-12 -4

-16 -5
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

1 2
Re=90 Re=240
1
0

0
cL

-1
cL

SPa -1
SPa
SPa
-2
-2 SPa

-3 -3
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

Fig. 21. Lift coefficient of a rotating sphere and spheroid (c/a = 1.5) with SPa = 0, 0.5, 1, 2 and 3 in a uniform flow as a function of the angle of incidence at Re = 0.3, 30, 90 and
240. (- - -) Sphere; (—) spheroid.
584 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

site to the particle rotation. Since the pressure change around a


body becomes more important with increasing Re, the difference
in cM between the sphere and the spheroid also increases with
increasing Re; more precisely from 19% at Re = 0.3 to 149% at
Re = 240 and SPa = 3.
The dependence of cM on a is plotted in Fig. 23. An increase of
SPa causes an increase of cM for the sphere and the spheroid at
every Re and a considered. One can see the same trend as for cD
and cL, namely that the variation in cM of the spheroid becomes lar-
ger with increasing SPa and with increasing Re due to the increase
of the pressure change in each case. In contrast to cD and cL, the
dependency of cM on a is small compared to the dependency on
SPa, since the pressure stresses act in the same direction at every
a, i.e. in that opposite to the particle rotation.

6. Stationary sphere and spheroid in a linear shear flow

The sphere and the spheroid with c/a = 1.5 are placed in linear
shear flows with shear rates of SFl = 0.04 and SFl = 0.08 at Re = 0.3,
30, 90 and 240. For the sphere Re = 3 and 480 are also considered.
The streamlines and the pressure field around a sphere at SFl = 0.08 Fig. 24. Streamlines and pressure field around a stationary sphere in a linear shear
flow with SFl = 0.08 at Re = 90.
and Re = 90 are shown in Fig. 24. As one can see, the flow is not
symmetric in the shear plane. The streamlines coming from the
bottom (low velocity side) do an S-bend behind the particle to- by the shear flow at Re = 0.3, 30 and 90. At Re = 240, the drag of the
wards the top. If flow separation occurs, only a small eddy at the sphere and the maximum drag of the spheroid, which appears at
low velocity side exists and the size of the eddy decreases with a = 90°, increase slightly with increasing SFl, because in these cases
increasing SFl. At Re = 0.3, the front stagnation point is nearly in and only in these cases, the pressure change around the body in-
front of the sphere as in the uniform flow and the maximum pres- creases appreciably with increasing SFl.
sure point has shifted towards the high velocity side. For Re P 30, As mentioned in Section 2, the influence of the top and bottom
the front stagnation point as well as the maximum pressure point wall is significant at Re = 0.3 in agreement with the result of Vas-
have moved slightly towards the high velocity side. seur and Cox (Eq. (24)) and cL increases approximately linearly
The computed cD of the sphere is plotted as a function of Re in with the distance between the top and bottom wall. Thus, the pres-
Fig. 25. Fig. 26 shows cD of the sphere and the spheroid as a func- ent results at Re = 0.3 cannot be compared with Saffman’s and
tion of a at Re = 0.3, 30, 90 and 240. The drag is almost unaffected McLaughlin’s solutions (Eqs. (9) and (11)) valid for unbounded

250 6
Re=0.3 SPa Re=30
200 5
SPa
4
150
3 SPa SPa
cM

100
cM

2
50
1
0 0
-50 -1
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

3.0 2.5
Re=90 Re=240
2.5 2.0
2.0 SPa 1.5
1.5 SPa
cM

1.0
cM

1.0 SPa
0.5 SPa
0.5
0.0 0.0

-0.5 -0.5
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

Fig. 23. Moment coefficient of a rotating sphere and spheroid (c/a = 1.5) with SPa = 0, 0.5, 1, 2 and 3 in a uniform flow as a function of the angle of incidence at Re = 0.3, 30, 90
and 240. (- - -) Sphere; (—) spheroid.
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 585

1
10 3 10 0.15
Mean value
McLaughlin [10]
spheroid c/a=1.5
SFl=0 Saffman [8] SFl=0.04 0.10
10 2 SFl=0.04 0
SFl=0.08
10 Sphere
SFl=0.08
SFl=0.04 0.05

cL
cL
SFl=0.08
10 1
cD

-1 Vasseur 0.00
10 and Cox [70]

10 0 -0.05
Stokes sphere
-2
10 -0.10
-1 -1 0 1 1 2 3
10 10 10 10 10 10 10
-1 0 1 2 3
10 10 10 10 10 Re
Re
Fig. 27. Lift coefficient of a stationary sphere and mean lift coefficient of a
Fig. 25. Drag coefficient of a stationary sphere in a linear shear flow with SFl = 0, stationary spheroid (c/a = 1.5) in a linear shear flow with SFl = 0.04 and 0.08 as a
0.04 and 0.08 as a function of the particle Reynolds number. function of the particle Reynolds number.

flows. Fig. 27 shows cL of the sphere and the mean cL of the spher- shear stress on the low velocity side is larger than on the high
oid versus Re, where cL is plotted logarithmic in the left diagram velocity side; thus on both sides the viscous wall shear stress lead
and linear in the right diagram. The values of the sphere approach to a lift towards the high velocity side. The lift due to pressure acts
Vasseur and Cox’s prediction at Re = 0.3. As shown in the right dia- also towards the high velocity side, but is considerably smaller
gram in Fig. 27, the lift of the sphere changes the direction at than that due to viscous stress. In contrast, at Re = 30, 90, 240
approximately Re = 50 and the mean lift of the spheroid changes and 480, the lift is determined by both the pressure distribution
the direction at approximately Re = 90 for both SFl. The absolute around the body and the viscous stress. On the upstream side,
values of cL of the sphere and of the mean cL of the spheroid are lar- the viscous wall shear stress on the high velocity side is smaller
ger at SFl = 0.08 than at SFl = 0.04 both before and after the change than on the low velocity side, causing a lift towards the low veloc-
of direction. ity side. However, at Re = 30, a large area of low pressure is situated
At Re = 0.3 and 3, the shear induced lift of the sphere is for the at the top of the sphere (high velocity side) leading to a total lift
most part due to the viscous stress. On the upstream side the vis- towards the high velocity side. As one can see from Fig. 24, the
cous wall shear stress on the high velocity side is larger than on the pressure at the bottom of the sphere becomes similarly low as
low velocity side and on the downstream side the viscous wall the pressure at the top at Re P 90. Since on the very upstream side,

92 2.4
Re=0.3 Re=30
90
2.2
88

86 2.0
cD

cD

84
1.8
82

80 1.6
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

1.3 0.9
Re=90 Re=240
SFl
1.2 0.8
SFl

1.1 0.7
cD
cD

1.0 0.6

0.9 0.5
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

Fig. 26. Drag coefficient of a stationary sphere and spheroid (c/a = 1.5) in a linear shear flow with SFl = 0, 0.04 and 0.08 as a function of the angle of incidence at Re = 0.3, 30, 90
and 240. (- - -) Sphere; (—) spheroid; (M) SFl = 0; (h) SFl = 0.04; (s) SFl = 0.08.
586 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

the pressure on the top hemisphere (high velocity side) is larger 10


1

than on the bottom hemisphere (low velocity side), see Fig. 24,
the resulting lift due to pressure also acts towards the low velocity Theoretical sphere
0
side at Re P 90. 10 (Faxén [13])
The change of direction of the lift for a sphere occurred at
Re  60 in Kurose and Komori’s work [58] and at 50 < Re < 100 in -1
Bagchi and Balachandar’s work [60]. In contrast, Dandy and Dwyer 10

cM
[61] did not observe a negative shear induced lift force. Based on
Dandy and Dwyer’s work, Mei [98] proposed a widely used approx- -2
10
imation for the shear induced lift on a sphere at finite Re. Therefore, Sphere
the existence of the negative shear induced lift force is not widely SFl=0.04 SFl=0.08
-3
accepted up to now. 10 Mean value spheroid c/a=1.5
The curves of cL versus a are plotted in Fig. 28. At Re = 0.3, cL of SFl=0.04 SFl=0.08
the spheroid is determined, among others, by the top and bottom -4
wall and the curve of cL versus a in a shear flow is similar to that 10
10 -1 10 0 10 1 10 2 10 3
in the uniform form, however, the variations are larger in a shear
Re
flow, since the high and the low pressure on the high velocity side
reach larger and smaller values, respectively. As for the sphere, at Fig. 29. Moment coefficient of a stationary sphere and mean moment coefficient of
Re = 30, the shear induced lift is determined by a region of low a stationary spheroid (c/a = 1.5) in a linear shear flow with SFl = 0.04 and 0.08 as a
pressure at the top of the spheroid. Thus, the shear induced lift function of the particle Reynolds number.

of the spheroid is positive at every a and the curves of cL are almost


parallel shifted to larger cL with increasing SFl. The shear induced Due to the viscous wall shear stress in a shear flow a torque acts
lift of the spheroid not only changes the direction with Re, but at on the sphere in the direction of the fluid rotation. As shown in
Re = 90 also with a; e.g. it is positive at a = 0° and negative at Fig. 29, cM of the sphere fits Faxén’s theoretical prediction (Eq.
a = 90°. Except for a = 0°, the shear induced lift is consistently neg- (14)) very well at low Re. Between Re = 90 and 240, a drastic drop
ative at Re = 240. Furthermore, the difference in cL of both the of cM occurs, which is consistent with Bagchi and Balachandar’s
sphere and the spheroid between SFl = 0.04 and 0.08 is smaller than simulations [60], and cM remains almost constant beyond it. The
between SFl = 0 and 0.04. Thus, at Re = 240, cL of the sphere and the mean values of cM of the spheroid, also plotted in Fig. 29, almost
mean cL of the spheroid seem to approach the asymptotic value of follow a straight line. They are 15% larger than cM of the sphere
approximately 0.1 with increasing SFl. It should be mentioned that for both SFl at Re = 0.3 and become much larger than cM of the
the Magnus lift coefficient also showed such a asymptotic behav- sphere at higher Re; they are 478% larger for SFl = 0.04 and 664%
iour with increasing SPa. for SFl = 0.08 at Re = 240. This trend is the same as for the Magnus

12 0.5
Re=0.3 0.4 Re=30
8
0.3
SFl
4 0.2 SFl
SFl
0.1 SFl
cL

cL

0 0.0
-0.1
-4
SFl -0.2
-8 -0.3
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

0.3 0.2
Re=90 Re=240
0.2 SFl 0.1
0.1 SFl
SFl 0.0
0.0
cL
cL

SFl -0.1 SFl


-0.1 SFl

-0.2 -0.2

-0.3 -0.3
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

Fig. 28. Lift coefficient of a stationary sphere and spheroid (c/a = 1.5) in a linear shear flow with SFl = 0, 0.04 and 0.08 as a function of the angle of incidence at Re = 0.3, 30, 90
and 240. Symbols as in Fig. 26.
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 587

8 0.4
Re=0.3 SFl=0.08 Re=30
6 0.3

0.2
4 SFl
cM 0.1 SFl

cM
2
0.0
SFl=0.04
0 -0.1
SFl=0
-2 -0.2
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

0.2 0.15
Re=90 Re=240
0.10
0.1 SFl SFl
0.05
SFl
0.0 0.00

cM
cM

-0.05
-0.1
-0.10

-0.2 -0.15
0 30 60 90 120 150 180 0 30 60 90 120 150 180
α [˚] α [˚]

Fig. 30. Moment coefficient of a stationary sphere and spheroid (c/a = 1.5) in a linear shear flow with SFl = 0, 0.04 and 0.08 as a function of the angle of incidence at Re = 0.3,
30, 90 and 240. Symbols as in Fig. 26.

torque and the reason once again is that cM of the sphere is only stage, the flow undergoes a transition to the non-axisymmetric
caused by the viscous wall shear stress, but for the spheroid the stage at a certain time.
pressure has an additional influence. Apart from Re, the drag acting on a particle depends strongly on
Fig. 30 shows cM as a function of a. At Re = 0.3, the maximum the particle shape and the angle of incidence. The influence of the
torque in a linear shear flow acts on the spheroid at the crosswise particle orientation and the particle shape on the drag increases
position (a = 90°) and the minimum torque at the lengthwise posi- with increasing Re. The particle streamlining has a greater influ-
tion (a = 0°) which was predicted theoretically for the angular ence on the drag at low rather than at high Re and the size of the
velocity by Jeffery [99] and computed numerically by Nirschl projected area has a greater influence at high rather than at low
[62]. Moreover, the curves of the spheroid for SFl = 0.04 and 0.08 Re. The effect of particle rotation on the drag of a sphere and a
differ almost only by the factor 2. The ratio between cM at a = 90° spheroid is negligible at Re 6 0.3 but significant at Re P 30. In de-
and 0° is 2.31 for both SFl, which is close to Jeffery’s value of 2.24 tail, the particle rotation causes an increase of the sphere drag and
for the angular velocity. At Re = 30, 90 and 240, the curves are sim- the mean (over the orientation) spheroid drag as well as a stronger
ilar to that in a uniform flow and become more and more similar to variation of the drag of the spheroid with respect to the orienta-
them with increasing Re, i.e. the influence of fluid shear on the tor- tion. In contrast, the linear shear flow has almost no influence on
que decreases compared to the influence of the particle profile the drag for the shear rates considered within this work.
with increasing Re. At Re = 30, a very small torque and at Re = 90 The curves of the profile lift and profile torque, i.e. the lift and
and 240 a negative torque acts on the spheroid at a = 0° since the torque of a stationary non-spherical particle in a uniform flow,
the area of highest pressure is situated on the upstream high veloc- versus the angle of incidence change their shapes with Re. The rea-
ity side, as for the sphere, and thus produces a negative torque for son is the change of the flow and the pressure distribution around
a = 0°. the bodies from almost symmetric around the particle center at
every orientation at Re = 0.3 to generally non-symmetric at higher
7. Conclusions Re. In addition, the pressure change around the body increases
with increasing Re. The influence of both the profile lift coefficient
The simulations of the flow past a sphere at Re 6 400 confirm and the profile moment coefficient relative to the drag coefficient
recent results from the literature, i.e. the flow undergoes a steady increases with increasing Re.
planar symmetric regime and an unsteady planar symmetric re- An additional lift and torque act on a particle if it is rotating or
gime between the steady axisymmetric flow at low and the unstea- placed in a linear shear flow. The additional lift on a rotating par-
dy non-symmetric flow at high Re. The transition between steady ticle in a uniform flow is called Magnus force. For these both cases
axisymmetric flow and steady planar symmetric flow occurs at a sphere and a spheroid were considered. The Magnus force and
200 < Re < 240, between steady planar symmetric flow and unstea- the shear induced lift force can be of importance both at low and
dy planar symmetric flow at Re  280 and between unsteady pla- at high Re. Both the mean Magnus lift coefficient and the mean
nar symmetric flow and unsteady non-symmetric flow at shear induced lift coefficient of the spheroid behave similarly to
320 < Re < 360. In all cases the flow past a sphere is axisymmetric the corresponding values of the sphere. At high Re and high fre-
at the initial stage. If the flow is non-axisymmetric in the final quencies of particle revolution, the Magnus lift coefficient of the
588 A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589

sphere is approximately 0.6, independent of the frequency of rev- [16] Sutterby JL. Falling sphere viscometry. I. Wall and inertial corrections to
Stokes’ law in long tubes. Trans Soc Rheol 1973;17:559–73;
olution. As for the drag, at Re P 30, the particle rotation leads to
Sutterby JL. Falling sphere viscometry. II. End effects in short tubes. Trans Soc
strong variations of the lift of the spheroid with respect to the ori- Rheol 1973;17:575–85.
entation. At Re = 0.3, the lift of a particle in a linear shear flow is [17] Masliyah JH, Epstein N. Numerical study of steady flow past spheroids. J Fluid
among other factors determined by the top and bottom wall in Mech 1970;44:493–512.
[18] Pitter RL, Pruppacher HR, Hamielec AE. A numerical study of viscous flow past
the current flow configuration and cannot be compared to the well a thin oblate spheroid at low and intermediate reynolds numbers. J Atmos Sci
known theoretical solutions valid for unbounded flows [8,10]. Nev- 1973;30:125–34.
ertheless, our results for the sphere agree with the theoretical pre- [19] Comer JK, Kleinstreuer C. A numerical investigation of laminar flow past
nonspherical solids and droplets. ASME J Fluids Eng 1995;117:170–5.
diction of Vasseur and Cox [70] calculated for the same flow [20] Dwyer HA, Dandy DS. Some influences of particle shape on drag and heat
configuration. At Re = 0.3, the shear induced lift of the sphere and transfer. Phys Fluids A 1990;2:2110–8.
the mean shear induced lift of the spheroid act towards the larger [21] Youngren GK, Acrivos A. Stokes flow past a particle of arbitrary shape: a
numerical method of solution. J Fluid Mech 1975;69:377–403.
relative fluid velocity. The shear induced lift of the sphere changes [22] Allen HS. The motion of a sphere in a viscous fluid. Philos Mag Ser 5
direction at approximately Re = 50, which confirms the results of 1900;50:323–39.
Kurose and Komori [58] and Bagchi and Balachandar [60], and [23] Arnold HD. Limitations imposed by slip and inertia terms upon Stokes’s law
for the motion of spheres through liquids. Philos Mag Ser 6 1911;22:755–75.
the mean shear induced lift of the spheroid changes direction at [24] Eiffel G. Sur la résistance des sphères dans l’air en mouvement. C R Hebd
approximately Re = 90. Séances Acad Sci 1912;155:1597–9.
At Re = 0.3, the moment coefficients due to particle rotation and [25] Wieselsberger C. Versuche über den Luftwiderstand gerundeter und kantiger
Körper. Ergebnisse der Aerodynamischen Versuchsanstalt zu Göttingen
linear shear flow can be several orders of magnitude larger than
1923;2:22–35.
the moment coefficient due to particle profile, depending on the [26] Bacon DL, Reid EG. The resistance of spheres in wind tunnels and in air. NACA
frequency of revolution and on the shear rate, respectively. The Report No. 185; 1924.
variation of the moment coefficient of the spheroid due to particle [27] Liebster H. Über den Widerstand von Kugeln. Ann Phys 1927;82:541–62.
[28] Schmiedel J. Experimentelle Untersuchungen über die Fallbewegung von
rotation over the orientation increases with increasing Re. In con- Kugeln und Scheiben in reibenden Flüssigkeiten. Physik Zeitschr
trast, at Re = 240, the torque due to linear shear flow is irrelevant. 1928;29:593–610.
In both cases of particle rotation and linear shear flow the mean [29] Jacobs EN. Sphere drag tests in the variable density wind tunnel. NACA
Technical Notes No. 312; 1929.
moment coefficient of the spheroid is larger than the correspond- [30] Roos FW, Willmarth WW. Some experimental results on sphere and disk
ing moment coefficient of the sphere since the pressure distribu- drag. AIAA J 1971;9:285–91.
tion has a share in the torque of the spheroid but not in the [31] Gebers F. Ein Beitrag zur experimentellen Ermittlung des Wasserwiderstandes
gegen bewegte Körper. Schiffbau 1908;9:435–52.
torque of the sphere. This difference increases with increasing Re. [32] Föppl O. Windkräfte an ebenen und gewölbten Platten. Dissertation. Berlin:
In [100] a new correlation formula for the drag coefficient of Springer; 1911.
arbitrary shaped particles accounting for their orientation was [33] Wieselsberger C. Untersuchungen über den Reibungswiderstand von
stoffbespannten Flächen. Ergebnisse der Aerodynamischen Versuchsanstalt
established using, among others, the present numerical results. zu Göttingen 1923;1:120–6.
In the same way, correlation formulas for the lift and moment coef- [34] Squires L, Squires W. The sedimentation of thin discs. Trans Am Inst Chem
ficient can be formulated, which account for particle shape, particle Eng 1937;33:1–12.
[35] Jones AM, Knudsen JG. Drag coefficients at low reynolds numbers for flow
orientation, particle rotation and shear flow. These correlations can
past immersed bodies. AIChE J 1961;7:20–5.
be used as a basis for further development of Lagrangian point-par- [36] Willmarth WW, Hawk NE, Harvey RL. Steady and unsteady motions and
ticle simulations in order to describe the behaviour of non-spheri- wakes of freely falling disks. Phys Fluids 1964;7:197–208.
cal particles. This would require the solution of additional ordinary [37] Pettyjohn ES, Christiansen EB. Effect of particle shape on free-settling rates of
isometric particles. Chem Eng Prog 1948;44:157–72.
differential equations for the particle orientation vector. [38] McNown JS, Malaika J. Effects of particle shape on settling velocity at low
Reynolds numbers. Trans AGU 1950;31:74–82.
References [39] Heiss JF, Coull J. The effect of orientation and shape on the settling velocity of
non-isometric particles in a viscous medium. Chem Eng Prog
1952;48:133–40.
[1] Haider A, Levenspiel O. Drag coefficient and terminal velocity of spherical and
[40] Chowdhury KCR, Fritz W. Sinkversuche mit isometrischen Teilchen in
nonspherical particles. Powder Technol 1989;58:63–70.
Flüssigkeiten. Chem Eng Sci 1959;11:92–8.
[2] Ganser GH. A rational approach to drag prediction of spherical and
[41] Heywood H. Measurement of the fineness of powdered materials. Proc Inst
nonspherical particles. Powder Technol 1993;77:143–52.
Mech Eng 1938;140:257–308.
[3] Stokes GG. On the effect of the internal friction of fluids on the motion of
[42] Needham LW, Hill NW. The settling of mineral particles in water. Fuel Sci
pendulums. Cambridge Philos Trans 1851;9:8–106.
Pract 1947;26:101–15.
[4] Oberbeck A. Ueber stationäre Flüssigkeitsbewegungen mit Berücksichtigung
[43] Hottovy JD, Sylvester ND. Drag coefficients for irregularly shaped particles.
der inneren Reibung. J Reine Angew Math 1876;81:62–80.
Ind Eng Chem Process Des Dev 1979;18:433–6.
[5] Oseen CW. Neuere Methoden und Ergebnisse in der Hydrodynamik.
[44] Wieselsberger C. Der Luftwiderstand von Kugeln. Zeitschrift für Flugtechnik
Leipzig: Akademische Verlagsgesellschaft; 1927.
und Motorluftschiffahrt 1914;5:140–5.
[6] Batchelor GK. Slender-body theory for particles of arbitrary cross-section in
[45] Dryden HL, Kuethe AM. Effect of turbulence in wind tunnel measurements.
Stokes flow. J Fluid Mech 1970;44:419–40.
NACA Technical Report No. 342; 1931.
[7] Rubinow SI, Keller JB. The transverse force on a spinning sphere moving in a
[46] Abbott IH. The drag of two streamline bodies as affected by protuberances
viscous fluid. J Fluid Mech 1961;11:447–59.
and appendages. NACA Report No. 451; 1934.
[8] Saffman PG. The lift on a small sphere in a slow shear flow. J Fluid Mech
[47] Maccoll JW. Aerodynamics of a spinning sphere. J Roy Aeron Soc
1965;22:385–400;
1928;32:777–98.
Saffman PG. Corrigendum. J Fluid Mech 1968;31:624.
[48] Davies JM. The aerodynamics of golf balls. J Appl Phys 1949;20:821–8.
[9] Sommerfeld M. Bewegung fester Partikel in Gasen und Flüssigkeiten. In: VDI-
[49] Tani I. Baseball’s curved balls. Kagaku 1950;20:405–9. in Japanese.
Wärmeatlas. Berlin: Springer; 2002. Lca1.
[50] Tanaka T, Yamagata K, Tsuji Y. Experiment of fluid forces on a rotating sphere
[10] McLaughlin JB. Inertial migration of a small sphere in linear shear flows. J and spheroid. The Second KSME-JSME Fluids Engineering Conference, Seoul
Fluid Mech 1991;224:261–74. 1990:366–9.
[11] Auton TR. The lift force on a spherical body in a rotational flow. J Fluid Mech [51] Barkla HM, Auchterlonie LJ. The Magnus or Robins effect on rotating spheres. J
1987;183:199–218.
Fluid Mech 1971;47:437–47.
[12] Kirchhoff G. Vorlesungen über Mathematische Physik. Leipzig: Teubner; [52] Tsuji Y, Morikawa Y, Mizuno O. Experimental measurements of the Magnus
1876.
force on a rotating sphere at low Reynolds numbers. ASME J Fluids Eng
[13] Faxén H. Der Widerstand gegen die Bewegung einer starren Kugel in einer 1985;107:484–8.
zähen Flüssigkeit, die zwischen zwei parallelen ebenen Wänden [53] Oesterlé B, Dinh TB. Experiments on the lift of a spinning sphere in a range of
eingeschlossen ist. Ann Phys 1922;68:89–119.
intermediate Reynolds numbers. Exp Fluids 1998;25:16–22.
[14] Ladenburg R. Über den Einfluß von Wänden auf die Bewegung einer Kugel in [54] Sawatzki O. Das Strömungsfeld um eine rotierende Kugel. Acta Mech
einer reibenden Flüssigkeit. Ann Phys 1907;23:447–58.
1970;9:159–214.
[15] Fidleris V, Whitmore RL. Experimental determination of the wall effect for [55] Dennis SCR, Singh SN, Ingham DB. The steady flow due to a rotating sphere at
spheres falling axially in cylindrical vessels. Brit J Appl Phys 1961;12:490–4. low and moderate Reynolds numbers. J Fluid Mech 1980;101:257–79.
A. Hölzer, M. Sommerfeld / Computers & Fluids 38 (2009) 572–589 589

[56] Cherukat P, McLaughlin JB, Graham AL. The inertial lift on a rigid sphere [77] Field SB, Klaus M, Moore MG, Nori F. Chaotic dynamics of falling disks. Nature
translating in a linear shear flow field. Int J Multiphase Flow 1994;20:339–53. 1997;388:252–4.
[57] Salem MB, Oesterle B. A shear flow around a spinning sphere: numerical [78] Marchildon EK, Clamen A, Gauvin WH. Drag and oscillatory motion of freely
study at moderate Reynolds numbers. Int J Multiphase Flow 1998;24:563–85. falling cylindrical particles. Can J Chem Eng 1964;42:178–82.
[58] Kurose R, Komori S. Drag and lift forces on a rotating sphere in a linear shear [79] Taneda S. Experimental investigation of the wake behind a sphere at low
flow. J Fluid Mech 1999;384:183–206. Reynolds numbers. J Phys Soc Jpn 1956;11:1104–8.
[59] Cherukat P, McLaughlin JB, Dandy DS. A computational study of the inertial [80] Magarvey RH, Bishop RL. Transition ranges for three-dimensional wakes. Can
lift on a sphere in linear shear flow field. Int J Multiphase Flow J Phys 1961;39:1418–22.
1999;25:15–33. [81] Wu J-S, Faeth GM. Sphere wakes in still surroundings at intermediate
[60] Bagchi P, Balachandar S. Effect of free rotation on the motion of a solid sphere Reynolds numbers. AIAA J 1993;31:1448–55.
in linear shear flow at moderate Re. Phys Fluids 2002;14:2719–37. [82] Provansal M, Ormières D. Étude expérimentale de l’instabilité du sillage d’une
[61] Dandy DS, Dwyer HA. A sphere in shear flow at finite Reynolds number: effect sphère. C R Acad Sci II b 1998;326:489–94;
of shear on particle lift, drag and heat transfer. J Fluid Mech Ormières D, Provansal M. Transition to turbulence in the wake of a sphere.
1990;216:381–410. Phys Rev Lett 1999;83:80–3.
[62] Nirschl H. Partikelbewegungen in Scherströmungen ohne und mit [83] Goldburg A, Florsheim BH. Transition and Strouhal number for the
Berücksichtigung des Einflusses angrenzender Wände. Düsseldorf: VDI incompressible wake of various bodies. Phys Fluids 1966;9:45–50.
Verlag; 1997. [84] Schouveiler L, Provansal M. Periodic wakes of low aspect ratio cylinders with
[63] He X, Luo L-S. A priori derivation of the lattice Boltzmann equation. Phys Rev free hemispherical ends. J Fluids Struct 2001;15:565–73.
E 1997;55:6333–6; [85] Hamielec AE, Hoffman TW, Ross LL. Numerical solution of the Navier–Stokes
He X, Luo L-S. Theory of the lattice Boltzmann method: From the Boltzmann equation for flow past spheres. AIChE J 1967;13:212–24.
equation to the lattice Boltzmann equation. Phys Rev E 1997;56:6811–7. [86] Pruppacher HR, Le Clair BP, Hamielec AE. Some relations between drag and
[64] Qian YH, d’Humières D, Lallemand P. Lattice BGK models for Navier–Stokes flow pattern of viscous flow past a sphere and a cylinder at low and
equation. Europhys Lett 1992;17:479–84. intermediate Reynolds numbers. J Fluid Mech 1970;44:781–90.
[65] Chen H, Chen S, Matthaeus WH. Recovery of the Navier–Stokes equations [87] Magnaudet J, Rivero M, Fabre J. Accelerated flows past a rigid sphere or a
using a lattice-gas Boltzmann method. Phys Rev A 1992;45:5339–41. spherical particle. Part 1. Steady straining flow. J Fluid Mech
[66] Aidun CK, Lu Y, Ding E-J. Direct analysis of particulate suspensions with 1995;284:97–135.
inertia using the discrete Boltzmann equation. J Fluid Mech [88] Johnson TA, Patel VC. Flow past a sphere up to a Reynolds number of 300. J
1998;373:287–311. Fluid Mech 1999;378:19–70.
[67] Bouzidi M, Firdaouss M, Lallemand P. Momentum transfer of a Boltzmann- [89] Dennis SCR, Walker JDA. Calculation of the steady flow past a sphere at low
lattice fluid with boundaries. Phys Fluids 2001;13:3452–9. and moderate Reynolds numbers. J Fluid Mech 1971;48:771–89.
[68] Mei R, Yu D, Shyy W, Luo L-S. Force evaluation in the lattice Boltzmann [90] Shirayama S. Flow past a sphere: topological transitions of the vorticity field.
method involving curved geometry. Phys Rev E 2002;65:041023. AIAA J 1992;30:349–58.
[69] Hölzer A. Bestimmung des Widerstandes, Auftriebs und Drehmoments und [91] Thompson MC, Leweke T, Provansal M. Kinematics and dynamics of sphere
Simulation der Bewegung nichtsphärischer Partikel in laminaren und wake transition. J Fluids Struct 2001;15:575–85.
turbulenten Strömungen mit dem Lattice-Boltzmann-Verfahren. Dissertation [92] Tomboulides AG, Orszag SA. Numerical investigation of transitional and weak
Thesis. Martin-Luther-University Halle-Wittenberg; 2007. Available from: turbulent flow past a sphere. J Fluid Mech 2000;416:45–73.
http://sundoc.bibliothek.uni-halle.de/diss-online/07/08H015/prom.pdf. [93] Ghidersa B, Dušek J. Breaking of axisymmetry and onset of unsteadiness in
[70] Vasseur P, Cox RG. The lateral migration of a spherical particle in two- the wake of a sphere. J Fluid Mech 2000;423:33–69.
dimensional shear flows. J Fluid Mech 1976;78:385–413. [94] Mittal R. Planar symmetry in the unsteady wake of a sphere. AIAA J
[71] Natarajan R, Acrivos A. The instability of the steady flow past spheres and 1999;37:388–90.
disks. J Fluid Mech 1993;254:323–44. [95] Lamb H. Hydrodynamics. Cambridge: Cambridge University Press; 1932.
[72] Kim HJ, Durbin PA. Observations of the frequencies in a sphere wake and of [96] Jayaweera KOLF, Mason BJ. The behaviour of freely falling cylinders and cones
drag increase by acoustic excitation. Phys Fluids 1988;31:3260–5. in a viscous fluid. J Fluid Mech 1965;22:709–20.
[73] Sakamoto H, Haniu H. The formation mechanism and shedding frequency of [97] Gans R. Wie fallen Stäbe und Scheiben in einer reibenden Flüssigkeit.
vortices from a sphere in uniform shear flow. J Fluid Mech 1995;287:151–71. Sitzungsb math-phys Kl Bayer Akad Wiss 1911;41:191–203.
[74] Föppl O. Ergebnisse der aerodynamischen Versuchsanstalt von Eiffel, [98] Mei R. An approximate expression for the shear lift force on a spherical
verglichen mit den Göttinger Resultaten. Zeitschrift für Flugtechnik und particle at finite reynolds number. Int J Multiphase Flow 1992;18:
Motorluftschiffahrt 1912;3:118–21. 145–7.
[75] Prandtl L. Der Luftwiderstand von Kugeln. Nachr Kgl Ges Wiss Göttingen, [99] Jeffery GB. Experimental investigation of the wake behind a sphere at low
Math-phys Kl 1914:177–90. Reynolds numbers. Proc Roy Soc Lond Ser A 1922;102:161–79.
[76] Saha AK. Three-dimensional numerical simulation of the transition of flow [100] Hölzer A, Sommerfeld M. New simple correlation formula for the drag
past a cube. Phys Fluids 2004;16:1630–46. coefficient of non-spherical particles. Powder Technol 2008;184:
361–5.

You might also like