Sunteți pe pagina 1din 12

Available online at www.sciencedirect.

com

Powder Technology 182 (2008) 342 – 353


www.elsevier.com/locate/powtec

Drag of non-spherical solid particles of regular and irregular shape


E. Loth ⁎
Department of Aerospace Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois, United States
Received 21 November 2006; received in revised form 30 April 2007; accepted 4 June 2007
Available online 12 June 2007

Abstract

The drag of a non-spherical particle was reviewed and investigated for a variety of shapes (regular and irregular) and particle Reynolds
numbers (Rep). Point-force models for the trajectory-averaged drag were discussed for both the Stokes regime (Rep ≪ 1) and Newton regime
(Rep ≫ 1 and sub-critical with approximately constant drag coefficient) for a particular particle shape. While exact solutions were often available
for the Stokes regime, the Newton regime depended on: aspect ratio for spheroidal particles, surface area ratio for other regularly-shaped particles,
and min–med–max area for irregularly shaped particles. The combination of the Stokes and Newton regimes were well integrated using a general
method by Ganser (developed for isometric shapes and disks). In particular, a modified Clift–Gauvin expression was developed for particles with
approximately cylindrical cross-sections relative to the flow, e.g. rods, prolate spheroids, and oblate spheroids with near-unity aspect ratios.
However, particles with non-circular cross-sections exhibited a weaker dependence on Reynolds number, which is attributed to the more rapid
transition to flow separation and turbulent boundary layer conditions. Their drag coefficient behavior was better represented by a modified
Dallavalle drag model, by again integrating the Stokes and Newton regimes. This paper first discusses spherical particle drag and classification of
particle shapes, followed by the main body which discusses drag in Stokes and Newton regimes and then combines these results for the
intermediate regimes.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Drag non-spherical; Spheroid; Reynolds sphericity; Irregular risk

1. Drag of spherical particles The drag force arises from pressure and viscous stresses applied
to the particle surface and resists the relative velocity (w).
With respect to the velocities of the different phases, the Assuming that the particle and the far-field velocities are steady
particle velocity (v) is defined as the translational velocity of the and the flow is spatially uniform away from the particle,
particle center of mass (xp). The continuous-fluid velocity (u) is ∇ · u@p = 0, the magnitude of drag is primarily dictated by the
generally defined in all areas of the domain unoccupied by particle Reynolds number (Rep), defined as
particles. However, a hypothetical continuous-phase velocity
can be extrapolated to the particle centroid as u@p and termed qf jwjd
Rep u : ð2Þ
the “unhindered velocity”. The relative velocity of the particles lf
(w) is then based on the unhindered velocity, i.e., along a
particle trajectory In this expression d is the particle diameter, ρf is the
continuous-phase density, and μf is the continuous-phase
wðtÞuvðtÞ  u@p ðtÞ: ð1Þ
viscosity. Stokes [1] derived the drag force for a sphere in the
It is important to note that u@p does not include the fluid limit of negligible convection terms (Rep ≪ 1) as
dynamic effects resulting from the presence of the particle itself.
FD ¼ 3kdlf w for Rep ≪1: ð3Þ

⁎ 306 Talbot Laboratory, 104 South Wright Street, Urbana, IL 61801-2935, This is often referred to as the Stokes' drag regime owing to his
United States. Tel.: +1 217 244 5581. derivation. For increasing particle Reynolds numbers, the flow be-
E-mail address: loth@uiuc.edu. hind the particle is remains an attached laminar wake for Rep b 22,
0032-5910/$ - see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2007.06.001
E. Loth / Powder Technology 182 (2008) 342–353 343

conditions indicate a transition owing to the appearance and


growth of a wake separation bubble [2]. Very high Reynolds
numbers (ca. N 300,000) correspond to the supercritical regime,
but particles are rarely at such conditions.
Since there is no general analytical solution for intermediate
Reynolds numbers, these conditions are generally prescribed by
an empirical expression of the drag coefficient. There are a large
number of curve-fits which have been proposed for a sphere at
these intermediate conditions. For example, Clift et al. [2] give a
ten-part empirical curve set for sphere drag up to Rep of 106, i.e.
the conditions include creeping flow and up to and past the drag
crisis. However, closed-form single equation expressions are
generally preferred and three common versions are given below.
An approximate form proposed by Dallavalle [3] for the sub-
critical regime is given by
Fig. 1. Drag coefficient for a smooth solid sphere at various Reynolds numbers
for incompressible flow conditions with experimental data reported in White [4] " rffiffiffiffiffiffiffiffiffiffiffiffiffiffi#2
and various theories and fits. 0:4Rep
fRe ¼ 1þ for Rep b 2  105 : ð6Þ
24
then as a separated laminar wake up to Rep b 130, to an unsteady
transitional wake up to Rep b 1000 and then turbulent wake [2–4]. Another form with similar accuracy is given by White [4] as
At high Reynolds numbers given by 2000 b Rep b 300,000, the
24 6
boundary layer initiating at the front of the particle (θ = 0°) will be CD ¼ þ pffiffiffiffiffiffiffi þ 0:4 for Rep b 2  105 : ð7Þ
laminar and separate at θ ∼ 80° producing a fully turbulent wake Rep 1 þ Rep
behind the particle [2]. The drag coefficient (CD) written in terms
of the total drag force is defined as A generally more accurate sub-critical expression (to within
k 6% of experiments as shown in Fig. 1) is given by Clift and
FD u  d 2 qf CD ww: ð4Þ Gauvin [5] as
8
 
24 0:42
Comparing against Eq. (3), the creeping drag coefficient is CD ¼ ð1 þ 0:15Re0:687
p Þ þ for Rep b 2
simply 24/Rep. Measured drag coefficients are plotted in Fig. 1 Rep 1 þ 42;500
Re1:16 p
for a wide variety of particle Reynolds numbers. For small Rep  105 : ð8Þ
values, the Stokes drag is appropriate. Measurements in the
range of 2000 b Rep b 300,000 indicate that CD is approximately The term in square brackets is also known as the Schiller–
constant in this Rep range (Fig. 1). This nearly constant value is Naumann [6] drag expression which is quite accurate up to a
called the “critical drag coefficient” (CD,crit) with a value of moderate Reynolds number. This can be written in terms of the
about 0.4 to 0.45. This Rep range is often called the “Newton Stokes correction as
regime”, owing to his assumption of a constant drag coefficient
for ballistics. One may normalize the drag force by the creeping fRe ¼ 1 þ 0:15Re0:687
p for Rep b 800: ð9Þ
flow solution to obtain a Stokes correction
This is perhaps the most commonly used drag correction
FD ðRep Þ CD ðRep Þ expression in multiphase flows since many particles are
fRe u ¼ : ð5Þ
FD ðRep Y0Þ 24=Rep constrained to Rep values in this range. Note that the drag
given by Eqs. (9)–(11) represents the time-averaged component
This ratio is unity for Rep ≪ 1 and is proportional to Rep for for Rep N 200, since some force unsteadiness results from the
the Newton regime (ca. 3000 b Rep b 200,000). The intermediate unsteady separated wake region.

Fig. 2. Schematic of ellipsoidal particles.


344 E. Loth / Powder Technology 182 (2008) 342–353

2. Shapes of non-spherical solid particles

Non-spherical solid bodies can be classified as either


regularly-shaped particles (e.g., ellipsoids, cones, disks) or
irregularly-shaped particles (non-symmetric rough surfaces).
For regularly shaped particles, a common non-spherical shape is
the ellipsoid as defined in Fig. 2 with an aspect ratio
dO
E¼ : ð10Þ
d⊥

Where d‖ and d⊥ are the parallel and normal diameters, and


respectively correspond to the semi-minor and semi-major axis
lengths. Spheroids are classified as oblate (E b 1) or prolate
Fig. 3. Influence of eccentricity on spheroid drag correction based on cylinder
(E N 1). For oblate spheroids, the shape becomes a circular disk in data from Dressel [13] and Clift et al. [2], where limiting solutions are for the
the limit E → 0, while prolate spheroids approach a needle shape disk and needle case while the approximate solution is for the moderate aspect
as E → ∞. The volume of an ellipsoid is proportional to the ratio case.
product of the diameters of the three major axes, so the effective
diameter is given by
shapes that cannot be defined by a simple geometric surface. For
d ¼ d⊥ E 1=3 ¼ dO E 2=3 : ð11Þ these crushed-particle shapes, a simple eccentricity approximation
is often not reasonable and instead the degree of deviation from a
Of course, E = 1 reverts the geometry to a sphere. sphere is generally be characterized by the surface area.
Other regularly-shaped particles generally include simple cross-
sections, e.g. ellipses or circles for the case of cones, spheroids, 3. Drag of non-spherical solid particles
ellipsoids and cylinders, and while cuboids (also called “rectangular
parallelepipeds” or rectangular boxes) are made up of connecting 3.1. Spheroids and ellipsoids in creeping flow
orthogonal rectangles faces so that they are characterized by three
perpendicular lengths. These can come from natural processes such The drag of non-spherical solid particles will depend on the
as crystallization. Irregularly shaped particles can stem from degree of non-sphericity as well as their orientation to the flow
random coagulation but also can result from break-up of larger solid (since the drag will generally be anisotropic with respect to
objects or particles. In particular, grinding and crushing processes direction). As note above, the spheroid is defined by an aspect
(e.g., used to produce coal particles and sand grit) create complex ratio (E). We may define the Stokes correction for ellipsoid of
aspect ratio (fE) as the ratio of creeping solid spheroid drag force
to creeping solid sphere drag force (where both have the same
Table 1 volume equivalent diameter), we have
Stokes correction factors for spheroids
Spheriod shape fE‖ fE⊥ FD ðE; Rep Y0Þ
fE ¼ : ð12Þ
Oblate exact ð4=3ÞE1=3 ð1  E2 Þ ð8=3ÞE1=3 ðE2  1Þ 3kdlf w
(E b 1) 2 Þcos1 E 2 Þcos1 E
E þ ð12E
pffiffiffiffiffiffiffiffi
1E2
E  ð32E
pffiffiffiffiffiffiffiffi
1E2
Thus fE approaches unity as E approaches unity. Exact,
   
limiting and approximate solutions for the drag on spheroids at
Oblate approx. 4 E 1=3 3 2E 1=3 creeping flow conditions (Rep ≪ 1) were derived by Oberbeck
(0.25 b E b 1) þ E þ E
5 5 5 5 [7]. Based on Eq. (12), the Stokes correction factors based on
the drag parallel and perpendicular to the axis of symmetry ( fE‖
Disk (E b 0.25) 8 1=3 16 1=3 and fE⊥) are given in Table 1. Approximate correction factors
E E
3k 9k are also shown which are based on small departures from the
spherical shape, i.e. |E − 1|≪ 1. In contrast, limiting solutions
Prolate exact ð4=3ÞE1=3 ð1  E2 Þ ð8=3ÞE1=3 ðE2  1Þ (tending to the disk or needle geometries) are obtained for very
(E N 1) pffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi small or very large E values. The trends for the parallel drag
ð2E2 1ÞlnðEþ E 2 1Þ ð2E2 3ÞlnðEþ E2 1Þ
E pffiffiffiffiffiffiffiffi Eþ pffiffiffiffiffiffiffiffi
2E 1 2E 1 correction factor as a function of aspect ratio are shown in Fig. 3
where the approximate and limiting conditions are shown to be
   
Prolate approx. 4 E 1=3 3 2E 1=3 reasonable over a wide range of aspect ratios in comparison to
(6 N E N 1) þ E þ E
5 5 5 5 the exact values. Also shown on this graph are some results for
cylinders which correspond especially well to the limiting cases
Needle (E N 6) of disks and needles (Ecyl = Lcyl / dcyl). Note that the drag for a
ð2=3ÞE2=3 ð4=3ÞE2=3
cylinder with near unity aspect ratio (Ecyl ∼ 1) is somewhat
lnð2EÞ  1=2 lnð2EÞ  1=2
larger than that predicted for spheroids with the same aspect
E. Loth / Powder Technology 182 (2008) 342–353 345

ratio (E). This is due to the increased surface area for the include non-continuum effects which are beyond the scope of
cylinder as compared to a spheroid with the same volumetric this article), then the net correction for oblate and prolate shapes
diameter (to be discussed in the next sub-section), so that a based on Eq. (15) and Table 1 becomes:
cylinder with a unity aspect ratio has a drag which is 14.4% pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
larger than that of a sphere. It can also be seen that variations in E1=3 1  E2
E from 0.01 to 100 result in only moderate changes in the h fE i ¼ for Eb1
cos1
pEffiffiffiffiffiffiffiffiffiffiffiffiffiffi
overall drag correction (less than four-fold). This arises because E 1=3 E2  1 : ð16Þ
increases in friction drag for longer bodies are approximately h fE i ¼  pffiffiffiffiffiffiffiffiffiffiffiffiffiffi for EN1
balanced by decreases in pressure drag, while the reverse is true ln E þ E2  1
for shorter bodies. As such, drag for particles whose aspect
ratios are near unity can often be reasonably estimated by As may be expected, the minimum drag shape for a given
simply using Stokesian drag and the volume-based diameter. particle volume when averaging over all orientations is a sphere
Generally, spheroids traveling at an initial orientation with (for which fE = 1). However, the minimum drag shape for a
no rotation will continue to fall at the same orientation, though a stationary orientation is a prolate spheroid with E = 1.955
broadside orientation tends to be more stable [2]. If the initial moving parallel to the axis of symmetry, where fE‖ = 0.9555
orientation is parallel or perpendicular to the relative velocity, (Fig. 3).
no torque exists since the drag acts in the opposite direction to
the gravitational force. However, it is possible for a spheroid to 3.2. Other regularly-shaped particles in creeping flow
be at an orientation which is neither parallel nor perpendicular.
Due to the linearity of the drag in creeping flow conditions, the Regularly-shaped non-spheroidal particles such as cones,
net drag force may be obtained based on a simple combination cubes, etc. (Fig. 2) do not typically have analytical solution for the
of the individual components [2,4], i.e. drag even in the creeping flow limit. As a first approximation,
their shapes and corresponding drag corrections may be
FD ¼ 3klf ðwO fEO þ w⊥ fE⊥ Þ for Rep ≪1: ð13Þ approximated as ellipsoids by determining an effective aspect
ratio (E). As shown in Fig. 3, this approach works well for
In this expression, w‖ and w⊥ are the components of relative cylinders since their shape is quite similar to that of a spheroid.
velocity which are parallel and perpendicular to the particle's However, additional accuracy may be obtained in terms of a shape
axis of symmetry (and whose vector sum is w). If both velocity correction ( fshape), defined similar to Eq. (12). As with the
components are non-zero, FD will not be parallel to w so that spheroidal shape factor fshape is inversely proportional to the
the particle will have a glide angle with respect to g [4]. One change in terminal velocity (for d = const.) since the drag
may recast the force of Eq. (11) into a drag component in the dependence is linear in creeping flow:
relative velocity direction and then define the remaining fluid

j
dynamic force as lift. However, herein we will consider the CD;shape Wterm;sphere
force of Eq. (11) as drag and define lift to be the result of non- fshape u ¼ : ð17Þ
CD;sphere Rep ≪1 & const:vol:
Wterm;shape
zero shear associated with the surrounding fluid or non-zero
relative rotation of the particle.
To estimate the shape factor for non-spheroidal regularly-
Similarly, the drag force for ellipsoids with aspects ratios E1,
shaped particles, it is common to consider two dimensionless
E2 and E3 (and other orthotropic particles) will be based on the
area parameters: the surface and the projected area ratios. Each
Stokes correction factor in each direction and the component of
of these areas can be normalized by the surface area of a sphere
relative velocity associated with each axis direction
which has the same volume, i.e.
FD ¼ 3klf ðw1 fE1 þ w2 fE2 þ w3 fE3 Þ for Rep ≪1: ð14Þ
Asurf Aproj
A⁎surf u ; A⁎proj u 1 2 : ð18Þ
In Brownian motion, all orientations can be assumed to be kd 2 4 kd
equally possible since molecular interaction is random. In this
case, one may consider an average correction factor by integrating The surface area ratio will always be greater than one since
over all possible orientations. For particles which are regularly the spherical geometry has the minimum surface area for a
given volume, i.e., Asurf⁎ ≥ 1. The inverse of the surface area
shaped about three axes, the drag for many realizations (or long
times) is based on the average of the inverse correction factors, i.e.: ratio is more commonly defined as the “sphericity ratio” [8], the
“sphericity” [2], or the “shape factor” [9]. For a cylinder with a
3 1 1 1 length to diameter ratio (Ecyl), ratio (Ecyl), geometry relations
¼ þ þ for Rep ≪1: ð15Þ
h fE i fE1 fE2 fE3 give the surface area ratio and equivalent volume diameter as:
 
Since spheroids have two axes which are equal in length and Lcyl ⁎ 2Ecyl þ 1 3Ecyl 1=3
Ecyl u ; Asurf ¼ ; d ¼ dcyl : ð19Þ
perpendicular to the axis of symmetry, this yields fE1 = fE‖ and dcyl 2 Þ1=3
ð18Ecyl 2
fE2 = fE3 = fE⊥. If one assumes that Stokesian drag is appropriate
which is reasonable for most small particles in a liquid (small The projected area ratio will depend on the orientation of the
particles in air undergoing Brownian motion will generally particle as well as its shape. For example, a long cylinder will have
346 E. Loth / Powder Technology 182 (2008) 342–353

Table 2 Since creeping drag is not highly sensitive to details in


Projected and surface area ratios along with Stokesian drag corrections for particle shape, the spheroidal approximation is reasonable for
regular particle shapes from data of Pettyjohn and Christiansen [29], Ahmadi
[42] and predicted corrections based on Leith [10] technique
many particles, e.g., cylinders as shown in Fig. 3.
From Table 2, it can be noted that increased deviations from
Shape A⁎proj A⁎surf fshape
a sphere or a spheroid result in decreased accuracy of Eq. (20).
Predicted Measured Highly flattened or elongated particle shapes (E ≪ 1 or E ≫ 1)
Sphere 1 1 1 1 tend to be over-predicted as discussed by Ganser [11]. For
Cube octahedron ∼1 1.10 1.03 1.03 example, application of Eq. (20) to disks yields a drag which is
Octahedron ∼1 1.17 1.05 1.07
about 14% too high for both orientations when compared to the
Cube ∼1 1.24 1.08 1.08
Tetrahedron ∼1 1.49 1.15 1.19 exact solution. More complex correlations have also been
2-sphere cluster 1.26 (broadside) 1.26 1.12 1.1 offered but in general empirical measurements for a specific
3-sphere cluster 1.44 (broadside) 1.44 1.20 1.27 shape are recommended when the deviation from spherical
4-sphere cluster 1.59 (broadside) 1.59 1.26 1.32 condition is extreme.
4-sphere cluster 1.59 (broadside) 1.59 1.26 1.17

3.3. Irregularly-shaped particles in creeping flow

Many naturally occurring particles closely resemble regu-


⁎ N 1 if it falls broadside, whereas it will have Aproj
Aproj ⁎ b 1 if it larly-shaped particles (e.g. crystals which form disk or rods) so

falls vertically along its axis. Equations for Asurf and Aproj ⁎ for that the above drag corrections of Table 1 or Eq. (20) are
double-cones, cuboids, etc. are given as a function of orientation generally adequate. However, many particles are more irregular
in Clift et al. [2] for non-isometric conditions. There are also other and pose a significant challenge for a number of reasons. Firstly,
measures of particle non-sphericity but these are the two most the surface area or projected area of such particles may be
commonly used parameters for particles with symmetry and turn difficult or impossible to measure so that approximations based
out to be the most effective in correlating the drag [2,9,10,13,21]. on the above area ratios are of little use. Secondly, experiments
In general, one would expect that larger values of Aproj ⁎ or for irregular particles at creeping flow conditions are not as

Asurf would correspond to larger drag values, and indeed this is plentiful as that for regular particles. Thirdly, the shape char-
the case. Leith [10] suggested the following correlation of these acterization of the particles is often not documented (e.g. a study
two area ratios for the Stokes shape correction factor may mention simply that sand particle of an effective diameter
was used) and is generally difficult to measure. Finally, within
1 qffiffiffiffiffiffiffiffiffi 2 pffiffiffiffiffiffiffiffiffi any mean shape characterization for such particles, irregularity
fshape ¼ A⁎proj þ A⁎surf for Rep ≪1: ð20Þ
3 3 generally leads to random protuberances orientations so that the
statistical variation of drag is often large.
This relationship was based on the fact that one-third of the Given these challenges, it can be expected that only
drag of a sphere is form drag (related to the projected area) while approximate and probabilistic predictions are possible for
two-thirds is friction drag (related to the surface area), and that the highly irregular particles. A wide variety of shape-characteriz-
form and skin friction drags are proportional to the particle ing parameters have been put forth for irregular particles, but the
diameter. This ratio also approximately holds for other non- most common and most successful is the Corey shape function
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
spherical particles with small deviations from a sphere, though it (CSF). This function is given by dmax = dmax dmed which
does not hold for very high or very low aspect ratios [2]. A review employs the three lengths of a particle in mutually-perpendic-
by Ganser [11] indicated that Eq. (20) gives reasonable results for ular directions: the particle's longest dimension (dmax), the
many well-defined shapes with moderate aspect ratios. In shortest dimension (dmin), and the intermediate or medium
particular, isometric particles with equal sides (cubes, octahedron, direction (dmed). A convenient version of CSF used herein is the
tetrahedron, etc.) have a projected area which is approximately the “max–med–min area”:
same for all orientations so that the projected areas for these
shapes can be estimated as unity. Table 2 lists some regular dmax dmed
A⁎mmm u 2
: ð22Þ
particle shapes along with measured and predicted Stokes dmin
correction factors. If the particle can be reasonably approximated
as having a surface area ratio close to that of a spheroid, one can This form is chosen for its consistency with Eq. (18) so that
use the following relationships for a given aspect ratio: increasing area ratios correspond to increasing drag corrections.
Since particles will generally lie on a plate on their broadside, the
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi! maximum and medium dimensions can be obtained through
E2=3 E4=3 1 þ 1  E2 visualization from above, while the minimum dimension
A⁎surf ¼ þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ln pffiffiffiffiffiffiffiffiffiffiffiffiffiffi for EV1
2 4 1  E2 1  1  E2 (thickness) can be obtained from visualization from a side. This
:
1 E 1=3 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi must then be repeated for a statistically large number of samples to
1
A⁎surf ¼ 2=3 þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sin ð 1  E 2 Þ for Ez1 obtain an average max–med–min area. While cumbersome, this
2E 2 1  E2 shape characterization is much more convenient than the
ð21Þ sphericity because the surface area of irregular particles is often
E. Loth / Powder Technology 182 (2008) 342–353 347

However, particles with strong indentations or pores could


radically alter the inscribed to circumscribed ratio. Unfortunately
there is insufficient quantitative data to develop models for such
shapes.

3.4. Non-spherical particles in the Newton-drag regime

Consistent with sphere drag, it is helpful to consider the drag


at high Reynolds numbers before proceeding to intermediate
values. Fortunately, non-spherical solid particles tend to have
drag coefficients which are approximately independent of
Reynolds number (ca. over a range 104 b Rep b 105) so that one
may reasonably define an approximately constant critical drag
coefficient in a Newton-drag regime. For convenience, this drag
coefficient can be normalized by that of a sphere with the same
volume (similar to the definition of fshape):
Fig. 4. Stokesian drag correction for irregular particles in terms of max–med–
min area with Rep b 0.5 data from Dressel [13], Baba and Komar [33] and sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Jimenez and Madsen [12].
Cshape u
CD;shape;crit
CD;sphere;crit j const:vol:
¼
Wterm;sphere
Wterm;shape
: ð24Þ
impractical to measure, and the same is true for other measures
like not-roundedness, anisometry and bulkiness ratio. Further-
more, the max–med–min area has been found to be generally Thus, Cshape can be thought of as a Newton-drag correction;
better at correlating drag modifications [2,12,13,34]. Note that a though Thompson and Clark [16] referred to this value as
projection area is difficult to characterize for an irregular particle “scruple”. Consistent with the Clift–Gauvin expression, Eq. (24)
since its asymmetry generally leads to torque imbalances and thus will use CD,sphere,crit ≈ 0.42, which is the approximate average for
complex tumbling. However, the free-fall orientation for irregular a sphere over 104 b Rep b 105 (Fig. 1). Measured Cshape values
particles is not uniformly random over all possible angles as they based on free-fall trajectories of different particle geometries are
tend to fall mostly in the so-called broadside orientation so that given in Table 3 and further geometries are given by Lasso and
dmin is roughly parallel to the average fall direction. Weidmann [17]. For regularly-shaped particle without a measured
Some approximate values for Ammm ⁎ for irregular non- Cshape value, one must resort to correlations based on particle
spherical particle shapes are given by Jimenez and Madsen [12], shape parameters.
e.g. 1.2 for smooth well-rounded shapes, 2 for shapes with some Similar to the creeping flow conditions, area ratios are ex-
edges but naturally rounded, and 6 for crushed sediment with pected to dominate the drag corrections. While drag at high
sharp edges. The correlation of this area ratio with the Stokesian Reynolds numbers is normally defined by the projected area, it is
drag correction for three well-characterized experiments of difficult to determine Aproj⁎ for some particles since the trajec-
particles in free-fall is given in Fig. 4. As expected for irregular tories will generally include secondary motion so that they are not
shapes, there is significant scatter but a strong trend is observed always falling in a broadside orientation. Therefore, several
with increasing area ratio correlated with increasing drag. A authors (Clift et al. [2]; Thompson and Clark [16]; Ganser [11])
general curve fit (averaged over many orientations and particle recommend only using sphericity, i.e. relating Cshape to Asurf ⁎ .
samples) can be given as The surface area ratios for spheroids and cylinders are given in

h fshape i ¼ ðA⁎mmm Þ0:09 for Rep ≪1: ð23Þ

This is nearly identical to the fit proposed by Dressel [13]


except that a 0.1 exponent was used. Other shape factors
“perimeter”, “circularity”, “shape entropy”, “polygonal harmo-
nics”, etc. have also been put forth to characterize the shape
irregularity, but the impact of the max–med–min area tends to be
strongest [2,14] which may be attributed to the fact that creeping
flow is not sensitive to sharp edges. However, it should be kept in
mind, that Eq. (23) is only a rough estimate, as evidenced by the
large experimental data spread shown in Fig. 4. For particles, for
which even Ammm ⁎ is not readily available, it is useful to recall
a theorem of Hill and Power [15] which states that the drag
coefficient in creeping flow is at most the drag of a circumscribing Fig. 5. An irregularly-shaped particle with illustration of its circumscribed
sphere and at least the drag of an inscribing sphere (Fig. 5). sphere (thick black line) and inscribed sphere (dashed line).
348 E. Loth / Powder Technology 182 (2008) 342–353

Table 3 projected area (vs. area associated with volumetric diameter)


Newton-drag corrections for various regular particle shapes based on data from was nearly equal to unity for cylinders for Ecyl N 10. This
the Pettyjohn and Christiansen [29], Stringham et al. [18], Clarke et al. [35],
⁎ . Data collected and
corresponds to a shape factor of 2.4 Asurf
Gogus et al. [32], and predictions

reviewed by Christiansen and Barker [21] for cylinders of
Particle shape Asurf CD,crit Cshape
Ecyl = 1.75 and Ecyl = 2.5 indicate secondary motion is sensitive
Predicted Measured to ρ⁎ and that affects the mean drag coefficients (up to ± 20%),
Sphere 1.00 0.42 1.0 1.0 but in general cylinders and prolate ellipsoids can be
Spheroid (E = 0.5) 1.095 0.84 2.1 2.0 approximately represented for a wide variety of density ratios
Cube octahedron 1.10 0.83 2.1 2.0
by the following expression:
Octahedron 1.17 1.14 2.8 2.7
Cube 1.24 1.40 3.3 3.3
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Cylinder (Ecyl = 1) 1.31 1.50 4.0 3.6 hCshape ic1 þ 0:7 A⁎surf  1 þ 2:4ðA⁎surf  1Þ for EN1:
Equilateral prism 1.43 2.20 4.9 5.2
Tetrahedron 1.49 1.90 5.3 4.5 ð26Þ

Interestingly, the result for long cylinders (Ecyl ≫ 1) is


Eqs. (21) and (19), while Table 3 gives area ratios for other shapes. approximately consistent with that based on a fixed finite-length
This table also shows that the Newton-drag corrections tend to cylinder in the broadside orientation, i.e. 〈Cshape〉 ≈ Cshape⊥.
⁎ with greater sensitivity than at creeping flow
increase with Asurf However, this is not generally true for other non-spherical
conditions, i.e. Cshape values of Table 3 are larger than fshape particle shapes because secondary motion tends to yield higher
values of Table 2 for the same shape. This may be attributed to the drag than that of a fixed broadside motion. As such, averaged
increased sensitivity of particle boundary layer separation to free-fall data is important to obtain reasonable results of the
increased shape convolutions at high Rep conditions. This mean drag.
sensitivity influences the extent of separated wake flow and the To account for secondary motion, a potentially more accu-
pressure drag component (in contrast, creeping conditions lead rate but complex technique is to determine the time-evolving
to attached flow over the particle regardless of any shape particle orientation during its trajectory based on applied
convolutions. torque and particle dynamics, and then apply a Cshape based on
The drag corrections for shapes in Table 3 which have this instantaneous local orientation. While more demanding
aspect ratios of order unity or less as well as and data for disks numerically, this has the advantage of employing the more
(0.012 ≤ Ecyl ≤ 0.515) and a spheroid (E = 0.5) free-falling in detailed drag data associated with various orientations (which
liquids are plotted as a function of the surface area ratio in can be obtained from fixed wind or water tunnel studies) and
Fig. 6. Despite the wide variety of particle shapes and surface furthermore allows inclusion of the lift and side forces to
area ratios, the results indicate a good correlation given by: simulate lateral motion. An example of this was conducted
by List et al. [22] for an ellipsoid of E = 0.5, whereby the
hCshape i ¼ 1 þ 1:5ðA⁎surf  1Þ1=2 þ 6:7 ð25Þ oscillatory motion was investigated and at least qualitative

ðA⁎surf  1Þ for EV1 and q⁎ f1:

However, this fit does not take into account inertia effects
which can be significant at low aspect ratios and high density
ratios. In particular, disks at high Reynolds numbers can
undergo significant pitching and/or tumbling depending on a
dimensionless inertia parameter⁎
(I⁎), which in turn is propor-
⁎ kq dO
tional to density ratio I u : As a result, measured drag
64d⊥
coefficients for disks falling in air [43] indicate a somewhat
reduced drag for a given area ratio.
The data of Stringham et al. [18] indicate that inertial effects
become important especially once the Reynolds number based
on the disk diameter becomes larger than 2000. As such, the
above drag correction expression is only approximate.
For cylinders and prolate spheroids (E N 1), secondary
motion has also been found to be significant only at extreme
aspect ratios. For example, Jayaweera and Mason [19] found
that unsteady motion such as fluttering was present in liquids at
high Reynolds numbers up to Ecyl b 25, whereas Gorial and
O'Callaghan [20] found that cylinders in air were generally Fig. 6. Newton-drag correction for regularly-shaped particles based on data of
stable for Ecyl N 4. However, both studies (and those they Table 3, Pettyjohn and Christiansen [29], Stringham et al. [18], and Haider and
reviewed) indicated that a drag coefficient based on broadside Levenspiel [23].
E. Loth / Powder Technology 182 (2008) 342–353 349

is functionally similar for all particle shapes and that the


differences are simply the corrections at the two extremes (given
by fshape and Cshape). Ganser argued this dependency based on
dimensional analysis, and it can be expressed via CD⁎ = f (Re⁎p )
by normalizing the drag coefficient and Reynolds number as:

CD
CD⁎ u
Cshape
: ð28Þ
⁎ Cshape Rep
Rep u
fshape

Evidence of this dependency can be seen in Fig. 8 where a free-


fall data for a large variety of non-spherical particles from
Jayaweera and Mason [19], Haider and Levenspiel [23], Madhav
and Chhabra [26], Tran-Cong et al. [14] are found to
Fig. 7. Newton-drag correction for irregularly-shaped particles in terms of max– approximately collapse for these normalized parameters. In this
med–min area based on data of Albertson [30], Alger and Simons [31], Gogus et al. conversion, a broadside orientation was employed for the Stokes
[32], and Smith and Chueng [34]. Additional studies have suggested Cshape ∼ 3 for correction, i.e. oblate particles are related to fE‖ while prolate
sand particles (Cheng [25]) and Cshape ∼ 4–4.5 for rocks (Clarke et al. [35]).
particle are related to fE⊥. Note that, the cluster of spheres data
from Tran-Cong et al. [14] corresponds to an fshape of 1.17 and a
agreement was obtained with respect to modeling secondary Cshape of 1.9. Unfortunately, there is little or no detailed data
motion. which can be used to assess Cshape and fshape of coal particles,
The case of irregular particles is further complicated by the though limited settling data at high concentrations and interme-
randomness of the shapes and dynamics. However, a review of diate conditions has been measured by Turian et al. [44].
experimental studies for mean drag of irregularly-shaped Upon close inspection of Fig. 8, one notes that there tends to
particles suggests that it is again reasonable to use the max– be two different curves along which the data collapse which are
med–min area ratio (as was used for the Stokes correction related to the shapes and orientations which dictate the cross-
factor). A qualitative dependency of the mean Newton shape section relative to the flow, or “C/S” for short. In particular,
factor for available well-characterized data is shown in Fig. 7 particles which can have a “circular C/S” such as spheres and
which can be reasonably correlated as needles (which tend to fall broadside) are fitted by the solid line,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi and particles which will have a “non-circular C/S” such as
hCshape i ¼ 6A⁎mmm  1: ð27Þ cubes, cones, disks (which tend to fall broadside) are fitted by
the dashed lines. Further data for agricultural particles are given
This expression, unlike Eq. (23) leads to a drag correction by Gorial and O'Callaghan [20] and Xie and Zhang [27] and
greater than unity as A⁎mmm approaches unity. This trend may generally follow the trends given above. This segregation may
be attributed to the increased drag associated with tumbling be related to the fluid dynamics over the particle surface. In
and flow separation for surface roughness at high Rep (whereas particular, particles which present a circular C/S to the flow will
creeping flow drag is relatively insensitive to roughness). How- have a surface separation point that will vary substantially with
ever, as with the Stokes correction for irregular particles, there is Rep (Fig. 1) while particles which present a non-circular C/S to
substantial uncertainty associated with Eq. (28) such that it is only the flow tend to have a separation point at the surface edges
a rough guide to the expected drag correction sampled over many which initiates at a low Rep (ca. 10 or less) and stays fixed
particles. regardless of Rep changes beyond that level.
Another example of the normalized drag curve is shown in
3.5. Non-spherical particles at intermediate Reynolds numbers Fig. 9 for particle shapes that correspond to various forms
of atmospheric precipitation including: spherical drops
Based on the above, non-spherical particles at intermediate (corresponding to clouds and small rain drops), various forms of
Rep can also be expected to produce a significant increase in snow in the form of dendrites, needles and disks, as well as hail
drag coefficient as compared to that for spheres. Such a result is shapes in the forms of a cone-hemisphere and a spheroid with
shown in the inset of Fig. 8 for the regularly-shaped data of E = 0.5. The limiting shape corrections for the spheroid and the
Haider and Levenspiel [23]. Determining a robust dependency needles are well represented by Table 1 and Eqs. (25) and (27). The
at intermediate Rep for non-spherical particles has challenged data again falls along the two curves seen in Fig. 8 corresponding to
many researchers and there have been more than a dozen forms whether the particle cross-section is circular or non-circular. One
of correlations (see review by Chhabra et al. [24]). The most would expect all prolate spheroids to yield a drag-curve with a
successful approaches are those by Ganser [11] and Cheng [25] “circular cross-section” but an interesting issue is whether this is
which use a combination of the Stokes drag correction and the also true for some oblate spheroids. The spheroid data in Fig. 9 and
Newton-drag correction. These approaches assume that the de- the E = 0.5 resolved-surface simulations of Comer and Kleinstreuer
pendence which governs the variation from Rep ≪ 1 to Rep,crit [28] indicate that such a drag-curve holds for E ≥ 0.5, while the disk
350 E. Loth / Powder Technology 182 (2008) 342–353

Fig. 8. Free-fall experimental data for various particle shapes in terms of normalized drag coefficient and Reynolds number with fits for both approximately circular
and non-circular cross-sections. The inset plot shows the conventional relationship between CD and Rep without normalization.

data (with Ecyl ≤ 0.05) indicate “non-circular cross-section” cylinders and even oblate spheroids with E N 0.5. As such, Eq.
characteristics. Other simulations by Comer and Kleinstreuer for (30) can be referred to as the “circular C/S Clift–Gauvin fit”.
Rep ranging from 40 to 120 indicate that spheroids with E =0.2 lie For moderate particle Reynolds numbers, a normalized
approximately in the middle of the circular and non- circular cross- Schiller–Naumann expression may be similarly defined by
section curves. Note that, the data from List and Schemanauer [22] modifying Eq. (11), i.e.
indicate that the cone-hemisphere shape has an fshape of 1.5 and a
Cshape of 2, while the dendritic snow flake (with a thickness to outer f ¼ fshape ½1 þ 0:15ðRe⁎p Þ0:687  for Re⁎p b800 &c circular C=S:
diameter ratio of 4%) has an fshape of 3.1 and a Cshape of 25. ð30Þ
Use of a dimensionless Clift–Gauvin expression based on
(10) yields Other expressions for spherical particles could also be
normalized in this fashion. For the second group of particles
24 h i whose relative cross-section is non-circular, a modified version
CD⁎ ¼ ⁎ 1 þ 0:15ðRep Þ
⁎ 0:687
of the Clift–Gauvin formula similar to that proposed by Ganser
Rep
0:42 can be obtained based on the best fit to the data as
þ for c circular C=S: ð29Þ
1 þ ðRe
42;500
⁎ 1:16
Þp
CD⁎ ¼
24
Re⁎p
½
1 þ 0:035ðRe⁎p Þ0:74  ð31Þ

As shown in Figs. 8 and 9, this gives good correlation for


0:42
particles for a wide range of Reynolds numbers whose relative þ for non  circular C=S:
cross-section (C/S) is approximately circular, e.g. spheres, 1 þ ðRe33⁎ Þ0:5
p
E. Loth / Powder Technology 182 (2008) 342–353 351

of irregularly-shaped particles is generally high or unknown


since they exhibit little or no drag crisis [30–32]. This is
because such particles tend to have a consistent bluff-body
separation and tumbling motion for a large range of Rep.
With respect to irregular particles, the results are generally
consistent with Eq. (33) but data are not available over the entire
range of Reynolds numbers (from Stokes to Newton behavior) for
a single shape set in order to establish this correlation as definitive.
Even less is known about the effect of large-scale contortions
for such conditions, but some experiments have investigated the
effect of small-scale roughness as discussed in the following
section.

3.6. Effect of roughness and turbulence

Small-scale roughness and turbulence is well known to


reduce Rep,crit, since both can cause the particle's boundary
layer to transition sooner to turbulent flow than for a smooth
sphere in steady flow. As a result, the drag crisis occurs at lower
Fig. 9. Normalized drag relationships for various forms of precipitation shapes Reynolds numbers in the presence of increased turbulence or
based on data of Jayaweera and Mason [19], Beard and Pruppacher [36], particle roughness. To quantify the impact on Rep,crit (defined
Stringham et al. [18], and List and Schemenauer [22]. when CD drops below 0.3), the particle surface variations (drms ′ )

This correlation is referred to herein as the “non-circular


Ganser fit” and, as shown in Fig. 8, gives reasonable agreement
for a wide variety of shapes. One may also consider the form
proposed by Cheng for sand particles, which is similar to
normalizing the Dallevalle drag expression given by Eq. (9)
except that Cheng used an exponent of 1.5 (vs. 2). However,
this was based on limited irregular particle data (in terms of both
shape and Reynolds number range) and neither of these
exponent values gives good agreement with the more
comprehensive non-spherical data presented in Figs. 8 and 9.
Fortunately, a simple but successful variant is to employ an
exponent of unity:

24
CD⁎ ¼ þ 0:4 for non  circular C=S: ð32Þ
Re⁎p

This is referred to as the “non-circular C/S Cheng fit” and is


found to give a remarkably good correlation as shown in Figs. 8
and 9, though it does not allow for potential dips which may
occur at about Re⁎p of about 1000 (prior to reaching the critical
drag coefficient).
The maximum Re⁎p used for the normalized drag correlations
is not well known for most shapes. Particles that are nearly
spherical but with small-scale rough surfaces (ca. 1% or less)
or spheroids with moderate aspect ratios (e.g. for 1/2 b E b 2)
can yield a decreased Rep,crit but generally display a nearly
constant Newton-like drag coefficient for Re⁎p up to 40,000
[29].
Fig. 10. Critical Reynolds number dependence on: a) surface roughness at low
However, the Newton regime for disks tends to be limited to turbulence, and b) turbulence intensity for various roughness values, based on
Re⁎p N 1000 since inertia (I⁎) effects can become important data from Torobin and Gauvin [37], Achenbach [38,39], and Neve and Jaafar
thereafter. In contrast, the maximum Rep for the Newton regime [40].
352 E. Loth / Powder Technology 182 (2008) 342–353

However, this relationship belies a complex influence of


turbulence for urms′⁎ N 0.1. In particular, high turbulence gives
rise to a “trans-critical” regime occurs for which the drag
coefficient twice drops below 0.3 as Rep increases. This regime
is shown in Fig. 11 whereby the drag coefficient undergoes an
initial drop, followed by a rise, followed by a second drop. This
figure also indicates that the drag can be quite sensitive at high
turbulence levels, e.g. at turbulence levels of 35% CD can
change by an order of magnitude for only a factor of three
change in Rep. As such, significant caution is needed when
employing the sub-critical CD expressions in turbulence levels
beyond a few %.

Fig. 11. Drag coefficient dependence on turbulence levels including data from 4. Conclusions
Gauvin and Clamen [41].
This study is a comprehensive review of particle drag associated
are normalized by its diameter and the free-stream turbulence with non-spherical shapes. The focus was on trajectory-averaged
′ ) are normalized by the relative velocity:
levels (urms drag coefficients based on particles in free-fall. Drag coefficients
for regular and irregular shapes have been collected herein for all
V
V⁎ urms available experimental and numerical data, including recent results
urms u ð33aÞ
w which are important to establishing robust models. For both very
high and very low Reynolds numbers, it was found that aspect ratio
V
V⁎ drms (E) was the best shape parameter for spheroids, surface area ratio
drms u : ð33bÞ
d (Asurf ⁎) was the best for regular particles, and max–med–min area
ratio (Ammm⁎) was the best for irregular particles.
The roughness effect is shown in Fig. 10a for a relatively low As suggested by Ganser, the key to determining the drag
′⁎ up to 1.5%, which leads to a four-fold
turbulence level for drms coefficient for non-spherical particles at intermediate Reynolds
reduction in Rep,crit. Conditions with very high surface irregu- numbers is to employ a normalized expression for the drag
′⁎ on the order of 15% or higher) have been studied
larities (drms coefficient based on the Stokes and Newton regimes. The
in the context of irregular particles as discussed in previous resulting drag coefficient models can be obtained from the
section. While very high Rep studies for such conditions are collected equations, figures and tables for the generally broadside
scare, there is no evidence of any drag coefficients below 0.3 for condition in terms of fshape and Cshape. However, it was found that
such particles. Furthermore, such particles are not likely to have the overall behavior also depends on the cross-sectional shape
a drag-crisis at all since the highly irregular surfaces will cause relative to the fall velocity. It should be kept in mind that the
the particles to have turbulent flow over the particle at modest correlations are based on mean free-fall characteristics and that
Rep values for which the measured CD is approximately con- the area ratios give only a gross generalization of the shape
stant (e.g. non-circular cross-section particles in Fig. 8). Unfor- characteristics, e.g. regularly-shaped particles of different shapes
tunately, there is insufficient quantitative data at intermediate but the same sphericity ratio can have significantly different
surface roughness levels (0.015 b drms ′⁎ b 0.15) to determine the Newton-based drag coefficients (that are not predicted by the
behavior of the transition. above functions). Thus, it is always best to obtain CD based on
The effect of turbulence is qualitatively equivalent to that of experiments with that same particle shape and Reynolds number,
small roughness levels on the particle surface. This effect is shown especially for shapes that are substantially non-spherical.
′⁎ b 10− 4)
in Fig. 10b for particles that are effectively smooth (drms However, the theoretical models and empirical correlations put
for which Rep,crit reduces to values as low as 20,000 at turbulence forth all have limits of for which they may be considered
levels of 20%. On the other hand, turbulence levels less than 1% reasonable. Furthermore, additional data is required to improve
have little effect on critical Reynolds number. The combined their accuracy and increase their ranges of applicability.
effect of roughness and turbulence intensity is also shown in
Fig. 10b, where it can seen that turbulence has a more profound Acknowledgement
effect and that the combination gives even larger reductions in
Rep,crit. Correspondingly, roughness becomes less important at The author would like to especially thank Mr. Vladimir
high turbulence intensities. An empirical expression which ap- Kudinok for creating the figures as well as assisting in the data
proximately correlates the combined effects of the data in Fig. 10a collection and analysis.
and b is given by
References
V⁎ 0:03
log10 Rep;crit ¼ 4:4ðdrms þ 0:002Þ
[1] G.G. Stokes, On the effect of the inertial friction of fluids on the motion of
 V⁎ 0:55
1:9ðurms V⁎
Þ ðdrms þ 0:002Þ0:07 : ð34Þ pendulums, Trans. Camb. Phil. Soc. 9 (part II) (1951) 8–106.
E. Loth / Powder Technology 182 (2008) 342–353 353

[2] R. Clift, J.R. Grace, M.E. Weber, Bubbles, Drops and Particles, Academic [24] R.P. Chhabra, L. Agarwal, N.K. Sinha, Drag on non-spherical particles: an
Press, New York, 1978. evaluation of available methods, Powder Technol. 101 (1989) 288–295.
[3] J.M. Dallavalle, Micrometrics, Pitman Publishing, New York, 1948. [25] N.S. Cheng, Simplified settling velocity formula for sediment particle,
[4] F.M. White, Fluid Mechanics, McGraw-Hill, New York, 1991. J. Hydraul. Eng. 123 (1997) 149–152.
[5] R. Clift, W.H. Gauvin, Proc. CHEMECA '70, vol. 1, Butterworth, [26] G.V. Madhav, R.P. Chhabra, Drag on non-spherical particles in viscous
Melbourne, 1970, pp. 14–28. fluids, Int. J. Miner. Process. 43 (1995) 15–29.
[6] L. Schiller, A.Z. Naumann, Über die grundlegenden Berechungen bei der [27] H.-Y. Xie, D.-W. Zhang, Stokes shape factor and its application in the
Schwerkraftaufbereitung, Ver. Deut. Ing. 77 (1933) 318–320. measurement of sphericity of non-spherical particles, Powder Technol. 114
[7] A. Oberbeck, Crelles J. 81 (1876) 62. (2001) 102–105.
[8] H. Wadell, The coefficient of resistance as a function of Reynolds number [28] J.K. Comer, C. Kleinsteuer, A numerical investigation of laminar flow past
for solids of various shapes, J. Franklin Inst. 41 (1934) 310–331. non-spherical solids and droplets, ASME J. Fluids Eng. 117 (1995)
[9] C.T. Crowe, M. Sommerfeld, Y. Tsuji, Multiphase Flows with Droplets and 170–175.
Particles, CRC Press, Boca Raton, FL, 1998. [29] E.S. Pettyjohn, E.B. Christiansen, Chem. Eng. Prog. 4 (1948) 157–172.
[10] D. Leith, Drag on non-spherical objects, Aerosol. Sci. Tech. 6 (1987) [30] L. Albertson, Effect of shape on the fall velocity of gravel particles, Proc.
153–161. 5th Hydr. Conf., Studies in Engineering, University of Iowa, Iowa City,
[11] G.H. Ganser, A rational approach to drag prediction of spherical and non- Iowa, 1952.
spherical particles, Powder Technol. 77 (1983) 143. [31] G.R. Alger, Simons, J. Hyd. Eng. Div. ASCE 94 (1968) 721.
[12] J.A. Jimenez, O.S. Madsen, A simple formula to estimate settling velocity [32] M. Gogus, O.N. Ipecki, M.A. Kokpinar, Effect of particle shape on fall
of natural sediments, J. Waterw. Port Coast. Ocean Eng. (2003) 70–78. velocity of angular particles, J. Hydraul. Eng. 127 (10) (October 2001)
[13] M. Dressel, Dynamic shape factors for particle shape characterization, 860.
Part. Charact. 2 (1985) 62–66. [33] J. Baba, P.D. Komar, Measurements and analysis of settling velocities of
[14] S. Tran-Cong, M. Gay, E.E. Michaelides, Drag coefficients of irregularly natural quartz sand grains, J. Sediment. Petrol. 51 (1981) 631–640.
shaped particles, Powder Technol. 139 (2004) 21–32. [34] D.A. Smith, K.F. Chueng, Settling characteristics of calcareous sand,
[15] R. Hill, G. Power, Extremeum principles for slow viscous flow and the J. Hydraul. Eng. (June 2003) 479–483.
approximate calculation of drag, Q. J. Mech. Appl. Math. 9 (1956) [35] N.N. Clarke, P. Gabriele, S. Shuker, R. Turton, Powder Technol. 59 (1989)
313–319. 69.
[16] T.L. Thompson, N.N. Clark, A holistic approach to particle drag [36] K.V. Beard, H.R. Pruppacher, A determination of the terminal velocity and
prediction, Powder Technol. 67 (1991) 57–66. drag of small water drops by means of a wind tunnel, J. Atmos. Sci. 26
[17] I.A. Lasso, P.D. Weidman, Stokes drag on hollow cylinders and (1969) 1066–1071.
conglomerates, Phys. Fluids 29 (1986) 3921–3934. [37] L.B. Torobin, W.H. Gauvin, Can. J. Chem. Eng. 38 (1960) 142–153.
[18] G.E. Stringham, D.B. Simons, H.P. Guy, The behavior of large parti- [38] E. Achenbach, Experiments on the flow past spheres at very high Reynolds
cles falling in quiescent liquids, Prof. Pap. US Geol. Surv., vol. 562-C, numbers, J. Fluid Mech. 54 (Part. 3) (1972) 565–575.
1969. [39] E. Achenbach, The effects of surface roughness and tunnel blockage on the
[19] K.O.L.F. Jayaweera, B.J. Mason, The behavior of freely falling cylinders flow past spheres, J. Fluid Mech. 65 (Part 1) (1974) 113–125.
and cones in viscous fluid, J. Fluid Mech. 22 (1965) 709–720. [40] R.S. Neve, F.B. Jaafar, The effect of turbulence and surface roughness on
[20] B.Y. Gorial, J.R. O'Callaghan, Aerodynamic properties of grain/straw the drag of spheres in thin jets, Aeronaut. J. 86 (1982) 331–336.
materials, J. Agric. Eng. Res. 46 (1990) 275–290. [41] A. Gauvin, W.H. Clamen, Effect of turbulence on the drag coefficients of
[21] E.B. Christiansen, D.H. Barker, The effect of shape and density on the free spheres in a supercritical flow regime, AIChE J. 15 (1969) 184.
settling of particles at high Reynolds number, AIChE J. 11 (1965) [42] Ahmadi G., Clarkson University, personal communication, 2005.
145–151. [43] W.W. Willmarth, N.E. Hawk, R.L. Harvey, Steady and unsteady motions
[22] R. List, Schemenauer, Free-fall behavior of planar snow crystals, conical and wakes of freely falling disks, Phys. Fluids 7 (1964) 197–208.
graupel and small hail, J. Atmos. Sci. 28 (1971) 110–115. [44] R.M. Turian, F.-L. Hsu, K.S. Avrarnidis, D.-J. Sung, R.K. Allendorfer,
[23] A. Haider, O. Levenspiel, Drag coefficient and terminal velocity of Settling and rheology of suspensions of narrow-sized coal particles, AIChE
spherical and non-spherical particles, Powder Technol. 58 (1989) 63–70. J. 38 (1992) 969–987.

S-ar putea să vă placă și