Sei sulla pagina 1di 5

Ind. Eng. Chem. Res.

1994,33, 1979-1983

1979

Free Settling of Nonspherical Particles


Miloslav Hartman, Otakar Trnka, and Karel Svoboda
Institute of Chemical Process Fundamentals, Academy of Sciences of the Czech Republic,
165 02 Prague 6-Suchdol, Czech Republic

New correlations have been developed to estimate the steady-state free-fall conditions of isolated,
nonspherical, isometric particles moving in Newtonian fluids. The proposed relationships cover the
Stokes, the Newton, and the transitional flow regions. Their explicit forms enable the direct
computation of the particle size corresponding to a chosen terminal velocity and the straightforward
calculation of the terminal velocity of a given nonspherical particle. The terminal velocities of
crushed particles of limestone and lime have been measured and compared to the values predicted
by the proposed formulas.

Introduction
Reliable correlations for estimating the terminal freefall velocity of particles moving through different fluids
are of great interest because of practical needs. Such
relationships are at the basis of the design and operation
of chemical process equipment such as fluidized bed
reactors, particle separators, classifiers, and crystallizers
(e.g., Yates, 1983).
The balance of forces acting on a particle which is falling
freely through an infinite fluid of constant density and
viscosity in a gravitational force field leads in the case of
steady state to

C g e ; = (4/3)Ar

(la)

In general,the drag coefficientCDis an intricate function


of the flow conditions, Ret, as well as the particle shape,
9 (Pettyjohn and Christiansen, 1948; Hartman and Yates,
1993):

C, = CD(Ret,J/)

(1b)

The particle shape cannot be readily and rigorously


described. Although some progress has been made in
attempts to characterize shape more sophisticatedly (e.g.,
Clark et al., 19891,the sphericity,9,is the most appropriate
single measure for characterizing the shape of close-toisometric nonspherical particles (Hartman and Coughlin,
1993). It is defined as
J/=
surface area of equivalent sphere
surface area of the particle

)bothof the same volume 5 1


(2)

Thus, J/ = 1for perfect spheres, and 0 C 9 C 1holds for


other bodies. The sphericity can be calculated exactly for
geometrical bodies such as spheroids, ellipsoids,and other
manufactured shapes. The sphericityof irregularlyshaped
particles can be determined photographically(Broadhurst
and Becker, 1975)or with the aid of empirically established
relations between the sphericity of particles and the
voidage of a packed bed of these particles. Viewing
particles through a microscope and comparison with the
published data (e.g., Yates, 1983; Kunii and Levenspiel,
1991; Hartman and Coughlin, 1993) will usually provide
a realistic value of 9.
As can be seen from eq 2, the sphericitymay be rigorously
defined only for purely geometric objects without any
~~

microstructure. Fractal studies (e.g., Mandelbrot, 1977;


Clark et al., 1989) suggest that the surface area of many
real particles is generally indeterminate and is influenced
by the technique with which it is measured. It appears
that a certain effectivesurfacearea may be exercisedwhich
will vary with the flow conditions. It should be noted that
the sphericity used in this work follows the packed bed/
Ergunapproach (Yates, 1983;Kunii and Levenspiel, 1991).
There are several ways of defining the size of a particle
of shape other than spherical. The most relevant size
parameter is the surface/volume diameter, d, (the diameter of a sphere having the same external surface area/
volume ratio as the particle). Similar fractal logic may be
applied to dsvas in the case of the particle sphericity.
In most practical fluid-solid systems, particles are
irregular and they are generally measured by screen
analysisprovided that they are larger than about 0.08 mm.
The screen size or sieve size, dp, is then the arithmetic
mean of the aperture of the screen which just lets the
particles pass through and the next finer screen below on
which they are retained. Unfortunately, there is no general
relationship between the screen size and the surface/
volume diameter. For isometric particles, it appears as a
feasible approach to employ the equivalent spherical
diameter based on the same volume, which can be
approximated by the screen size:
dv h = d m - dP

(3a)

For a long period of time, considerableeffort was directed


to the determination of the free-fall velocity of spheres in
various fluids. Results of such experiments were interpreted and correlated in terms of the drag coefficient of
spheres related to the Reynolds number (Clift and Grace,
1978; Flemmer and Banks, 1986; Turton and Levenspiel,
1986). Comparison of these correlations indicates that
the differences in the drag coefficientare practically always
less than 4% (Hartman et al., 1990; Hartman and Yates,
1993). These correlations are equivalent from the standpoint of accuracy. It can easily be shown that a deviation
of 4% in CD leads to an error of less than 2% in Ret for
the Reynolds numbers widely ranging from 1to 2 X lo5.
In contrast to the spheres, relatively little experimental
work was done with nonspherical particles (e.g., Pettyjohn
and Christiansen, 1948; Becker, 1959; Christiansen and
Barker, 1965). Based on their data, Pettyjohn and
Christiansen (1948) introduced simple corrections to the
spherical particle equations for the Stokes region

* Author to whom correspondence should be addressed.


osse-~ss~19~12s33-i9~g~o4.5o10
0 1994 American Chemical Society

1980 Ind. Eng. Chem. Res., Vol. 33, No. 8, 1994

where
K = 0.843 10g,~($/0.065), Re, C 0.05, 0.67 < < 1 (3c)

and the Newton region

CD= 5.31 - 4.88$, 2

lo3 C Re, C 2 X lo6,


0.67 < $ < 1 (3d)

Under flow conditions in the intermediate (transition)


region (0.05 C Ret C 20001, the authors recommend that
the experimental curves CD = CD(Ret,$) be used in trial
and error estimations of the free-fall terminal velocity of
nonspherical particles. In many applications, particularly
in the case of fluidized beds, the particles of interest have
terminal velocities within the intermediate regime and
convenient methods of calculation are necessary.
Recently, Haider and Levenspiel (1989) provided empirical correlations CDvs Ret and # for Newtonian fluids.
Since Ret is present in botheq 1and in tedious expressions
for CD= CD(Ret,$),an appropriate trial and error procedure
is required to predict the terminal velocity. It appears
that the correlation of Haider and Levenspiel (1989) tends
to overestimate CD for spheres at Ret L 100, and the
expressions of Turton and Levenspiel (1986) or Flemmer
and Banks (1986) should be preferred in such events
(Hartman et al., 1989).
In practice, a situation frequently occurs when the size
of particles, which are just entrained under the operation
conditions of interest, is to be determined. Successive
approximations are needed for calculating the size of a
particle which has a given terminal velocity when we
employ the above implicit relationships 1and lb. To avoid
the "double cycled" trial and error computations, the
particle size must be eliminated from eq 1. As it holds
that
(4)
eq 1 can be rewritten as

Re, = CD(Re,,$)/Y

(5)

where Y is the dimensionless group defined by

Table 1. Physical Properties of the Gas-Solid System and


Terminal Velocity Conditions Determined by Experiment.
d , x 104 (m)
Ar
Ut(m/a)
Ret
Y
CD
a. Crushed Particles of Limestone
Density, p s = 2600 kg/m3; Particle sphericity, $ = 0.55
1.12
134.2
0.410
3.044 6.344
19.31
2.25
1088.0
0.794
11.84
0.8740
10.35
2.82
2142.5
1.059
19.80 0.3680
7.286
4.50
8704.7
1.927
57.49
0.06108
3.511
2.570
96.27
0.02575
2.479
5.65
17232.0
b. Lime Particles
Density, pa = 1600 kg/m3;Particle Sphericity, $ = 0.78
1.12
80.70
0.407
3.022 3.899
11.78
2.25
654.3
0.631
9.413 1.046
9.846
2.82
1288.0
0.776
14.51
0.5622
8.158
1.510
45.05
0.07634
3.439
4.50
5235.0
2.370
88.78 0.01974
1.753
5.65
10362.0
a Properties of air at 20 OC: density, pf = 1.2kg/m3; viscosity, 1.81
x 1od Pa s.

(3) To examine the agreement between the measured


results and predictions of the proposed correlation.
Experimental Section
The terminal velocities of very narrow fractions of
limestone and lime particles determined from the dependence of pressure drop versus gas velocity for a shallow
bed of such particles were measured in a fluidization
column (Kunii and Levenspiel, 1991).
Materials. The present study was conducted with a
high-grade limestone and its calcine. The particles
comprised five very narrow, carefully sieved fractions:
0.100 - 0.125 mm (2, = 0.112 mm), 0.20 - 0.25 mm (a, =
0.225 mm), 0.25 - 0.315 mm
= 0.282 mm), 0.40 - 0.50
mm
= 0.450 mm), and 0.5 - 0.63 mm
= 0.565 mm).
The Czech Standard series of sieves was employed, in which
the apertures of the adjacent sieves are arranged in
multiples of 1.25. Examinations with an optical microscope indicated that the collected particles had approximately equal axes at right angles to each other. In
other words, their near-isometric form can be assumed.
While sharp edges could be seen on the limestone material,
the lime particles appeared quite round. Particle sphericity was determined by a procedure based on pressure drop
measurements of the fixed (static) bed (Svoboda and
Hartman, 1981). According to this method, the particle
sphericity can be expressed from the Ergun equation
(Hartman and Coughlin, 1993)
hp

&2

One can note that the particle size does not occur on the
right-hand side of eq 6. Similarly as in the case of Ut (eqs
l a and lb), the estimation of the particle size corresponding
to a given terminal velocity and physical properties of a
system also requires an iterative ("single cycled") solution
of eqs l b and 5.
This article is a sequel to recent studies of ours on the
predictive relationships for steady-state free fall of particles
in a Newtonian fluid. The primary objectiveof this project
was to propose an explicit formula that would make it
possible to estimate the size of a nonspherical particle
directly from its terminal velocity under different conditions of operation.
The work was divided into three parts: (1)To measure
the terminal velocities of nonspherical limestone and lime
particles. (2) To develop an explicit relationship Ret =
Ret( Y,$) by treatment of the voluminous experimental
data on free-settling rates of nonspherical isometric
particles amassed by Pettyjohn and Christiansen (1948).

(ap

(ap

- 1.75(1-

c)pfU2

e3d,

(ap

U
- 150(1- E ) ~ / +- 0 (7a)

Bed voidage was determined by weighing a mass of


particles slowly poured into a glass cylinder:

The values of E slightly decline with increasing particle


size and amount to 0.52 (limestone) and 0.42 (lime).
Repeated measurements provided average values of the
sphericity, $ = 0.55 for limestone and $ = 0.78 for lime.
Dried air was used at 20 OC as an operation medium.
Physical properties of both phases are given in Table 1.
Apparatus. An experimentalsetup with a glass column
of inside diameter 8.5 cm and height 2 m was used. The
column was equipped with an interchangeable gas plate
distributor and fluidizing grid. Their free area was varied
to avoid excessive pressure drop. The distance between

Ind. Eng.Chem. Res., Vol. 33, No. 8,1994 1981


Table 2. Physical Properties of the ZiquidLSolid' System
and Terminal Velocity Conditions Determined by
Exwriment
dpX108(m) p.(kg/m9
Ar
&Cis)
Ret
Y
CD
7.004
5.189 1.350
1.119
2703.7
141.42 0.0516
6.287
1.201
2703.7
6.088 1.033
174.75 0.0564
2.046
2532.3
766.71 0.1046 19.234 0.1437 2.764
2.504
2589.8 1464.3 0.1357 30.538 0.06857 2.094
3.220
2529.6 2981.7 0.1738 50.295 0.03126 1.572
4.080
2453.5 5720.0 0.2134 78.249 0.01592 1.246
~

~~

~~~

s = [E(log(Re,), - l o g ( f t e ~ ~ ) ~ / r n ~ ' / ~(8)


P I

Properties ofthe dimethyl ester of phthalic acid at 25 O C : density,


1.323 X 1W2Pa 8. Glass spheres ($
= 1) of different size and density.
0

pf = 1189 kg/ms; viscosity, pf =

these perforated plates was 3 cm. A ring with a height of


1.5 cm was situated just above the fluidizing grid for
measurements of the pressure drop across the bed. The
elutriated particles were separated in a cyclone, and the
dust was removed in a fabric filter. The flow rate of air
was gradually increased, and the pressure drop of the
fluidized bed vs the velocity of air was recorded. The
terminal velocities given in this work are the arithmetic
mean of the velocity at which the decline in pressure drop
becomes noticeable and that when practically all the
particles are entrained out of the column. Such values
are related to the mean sieve size of the particles.
The difference between the end and the beginning of
entrainment is a function of the width of a sieved fraction
(0.0264.13 mm) and varies from about 0.05 for the smallest
particles to approximately 0.25 m/s for the largest particles.
Replicate measurements of the terminal velocity showed
reproducibility better than 5%. The measured terminal
velocities are also given in Table 1. Their values vary
from0.410 to 2.57 m/s for the limestone particles and from
0.407 to 2.37 m/s for the lime particles.

Various forms of the correlations were attempted and


tested. An approach making use of a departure in the
behavior of a nonspherical particle from that of the sphere
proved to be productive. The correlation of Turton and
Levenspiel (1986) for the drag coefficient was employed
to describe the free-fall conditions for spheres. Using the
simplex minimization method, a reasonable order polynomial was fittedto the data for the nonspherical particles:
log Re,(Y,$) = log R e t ( Y , l )+ P ( Y , $ )

(9)

where
log R e t ( Y , l ) = 0.77481 - 0.56032 log Y +
0.024246(10g

n2- O.O038056(10g Y)'

(10)

P(Y,$) = -0.10118(1- $) log Y +


0.092944(1- $)(log
- 0.0098356(1- $)(log 0.12666(1- $)2 log Y (11)

n2

v3

for 0.01 < Ret < 16 OOO and 0.67 < $ < 1 and with s =
0.0470. One should note that eq 10 holds for spheres ($
= 1).
An effort was also made to derive a simpler formula
than eq 9. An entirely different approach, similar to that
of Churchill and Usagi (1972), was applied. This method
employs the limiting solutions for large and small values
of the independent variable. The use of a simple equation
originated from its canonical form is proposed for interpolation between the asymptotic solutions

Free-Fall Velocity Measurements with Glass


Spheres. Experimentswere also conducted in a0.0507-m
glass column with a glass spheres-dimethyl ester of
phthalic acid system. The detailed description of a the
experimental apparatus as well as that of the procedure
can be found elsewhere (Hartman et al., 1992). The
measured terminal velocities and physical properties of
the system are given in Table 2.

Development of the Relationship Ret = &( Y,$)


from the Data of Pettyjohn and Christiansen
(1948)
These authors amassed a considerable volume of
experimental data on the free fall of well-defined bodies
such as spheres ($ = 1, Ret E (0.0359,22640)), cube
octahedrons ($ = 0.906, Ret E (0.00 592,17 410)), octahedrons ($ = 0.846, Ret E (0.00 837,17 350)), cubes ($ =
0.806,Ret E (0.0076,16 OBO)), andtetrahedrons ($ = 0.67,
Ret E (0.00691,13 310)) in different liquids. More than
400 data points are more or less uniformly distributed
over the respective sphericities.
What is presented below are the results of efforts to
develop a convenient mathematical approximation to the
above experimental results. The proposed correlation is
then confronted with the experimental measurements of
the authors. The results of Pettyjohn and Christiansen
(1948) plotted in any form exhibit a considerable scatter.
More than 400 experimental data points were fitted by
minimizing the standard deviation between the experimental values and the values estimated from a proposed
relationship:

The quantities CD and K contained in eq 12 are given


by eq 3d (for small Y, large Ret) and eq 3c (for large Y,
small Ret), respectively.
The power ''n" occurring in the general canonical
expression of Churchill and Usagi (1972) was evaluated
for all the respective bodies of Pettyjohn and Christiansen
(1948). It moderately fluctuated between 1.2 and 1.4
around the mean value 1.351for the nonspherical particles.
However, the measured results for the spheres led to an
appreciably different value of 0.79. With this result, the
correlation for spheres takes the form

Re,(Y) = m [ 1 + 7.38Y"995]1.266,$ = 1
(13)
Y
Comparison with the Measured Data for the
Particles of Limestone, Particles of Lime, and
Glass Spheres
The proposed eqs 9 and 12 were subjected to testing
with the use of the experimental data measured for the
particles of limestone and lime, The measured terminal
velocities as well as related quantities are presented in
Table 1. Figures 1 and 2 illustrate reasonably good
agreement between the predicted Reynolds numbers at
the terminal velocity and the values measured at room
temperature. The proposed correlation 9 fib the experimental data points for limestone and lime with an accuracy
of 3-22 5%. As also can be seen in Figures 1 and 2, the
predictions of eqs 9 and 12 that are represented by the
curves 1 and 2, respectively, are not far apart. The
differences do not exceed about 105% , Considerably lower

1982 Ind. Eng. Chem. Res., Vol. 33, No. 8, 1994

Figure

1. Comparison of measured and predicted values of the


Reynolds number at terminal velocity: fluid, air; temperature, 20
"C; bodies, crushed particles of limestone; sphericity, = 0.55; (0)
experimental data points; Ut = 0.41-2.57 m/s; curve 1, predictions
of eq 9; curve 2, predictions of eq 12; curve 3, predictions of eq 10
for spheres (+ = 1).

Figure 3. Comparison of measured and predicted values of the


Reynolds number at terminal velocity: fluid, the dimethyl ester of
phthalic acid; temperature, 25 OC; bodies, glass spheres, # = 1;( 0 )
experimental data points; Ut= 0.0516-0.2134 m/s. Predictions of
eqs 10 and 13 coincide in the solid line shown in this figure.
velocity as well as the overall behavior of such particles
should be determined by experiment.
In order to enable the solving of engineering problems
related to the motion of nonspherical, isometric particles
in fluids from different angles,other expressionshave been
derived. Thus, the correlations Ret = Ret(Ar,$) and CD
= CD(Ret,$) are given in the Appendix.

Conclusions

4
1 1Y'"

Figure 2. Comparison of measured and predicted values of the


Reynolds number at terminal velocity: fluid, air; temperature, 20
O C ; bodies, sieved particles of lime; sphericity, # = 0.78; (0)
experimental data points; Ut= 0.40-2.37 m/s; curve 1, predictions
of eq 9; curve 2, predictions of eq 12.

values of Ret shown by curve 3 for the spheres in Figure


1 show that the particle shape has to be accounted for in
predicting the free-fall conditions, particularly at higher
Reynolds numbers.
The comparison of the measured terminal velocities of
the spheres with the predictions of eqs 10 and 13 in terms
of Ret and Y is presented in Figure 3. The proposed eqs
10 and 13 tend to underpredict Ret slightly with an error
of 2-4% and 3-7 % ,respectively. Understandably, this is
a better agreement between the estimates and experiment
than in the case of the nonspherical particles or limestone
and lime.
Of course,the empirical correlations developed here have
the usual limitations, and they should be applied with
caution outside the experimental conditions from which
they were educed.
For example, as the particles become less isometric, the
moment of inertia is more important. Solids of widely
nonisometric shapes do not fall vertically at a constant
velocity but follow a zig-zag path or rotate. The free-fall

We have developed and verified by experiment a


correlation for the direct computation of the size of single,
isometric, nonspherical particle which corresponds to a
given terminal velocity under different operation conditions. The proposed relationship covers the Stokes, the
Newton, and the transitional flow regions. The data
measured on the nonspherical particles of limestone and
lime fit to the proposed formula with an accuracy better
than 20%.
An important part of the correlation developed here for
the nonspherical particles is an independent expression
for spheres whose maximum deviation from experiment
amounts to 4 % .

Acknowledgment
This work was supported in part by the Grant Agency
of the Academy of Sciences of the Czech Republic under
Grant No. 472113.

Nomenclature
Ar = Archimedes number = dp3gpt(pll- pf)/pf2
CD = drag coefficient of a particle

dp = particle size, m

d, = mean particle size, m


d,,, = aperture of screen, m
dsph = diameter of a sphere, m
F = cross-sectional area of vessel, m2
g = acceleration due to gravity, m/s2 (9.807)
H = height of bed, m
AH = increment of bed height, m
K = Stokes law shape factor defined in eq 3c
m = number of experimental data points
P(Y,+) = quantity defined in eqs 11, A-3, and A-6
AP = pressure drop, Pa

Ind. Eng. Chem. Res., Vol. 33, No. 8,1994 1983


Ret = Reynolds number at terminal velocity of a particle =
UtdpPtlClt

(Re& = experimental value of Ret


(&e& = computed value of Ret
s = sum of squared errors defined in eq 8
U = superficial velocity of fluid, m/s
Ut = terminal velocity of an isolated particle in an infinite
fluid
W = mass of particles, kg
Y = dimensionless group = (4/3)Cg(~l - pr)~/(Ut~pH)l=
(4/3)(Ar/Ret3)= C d R e t
Greek Symbols
t = bed voidage
pf = fluid viscosity, Pa s, kg/(m s)
pf = fluid density, kg/m3
pa = particle density, kg/m3
= particle sphericity, shape factor
Other Symbol
log denotes base-ten or Briggsian logarithm

P(Re,,$) = -0.03874(1- $) log Re,


0.09238(1- $)(log Ret)2+ 0.06003(1- $)(log ReS3
0.01005(1- $)(log ReJ4 - 0.003571(1- $)(log Re,) 0.005697(1- $)2(10gRet) (A-6)
for 0.01 < Ret < 16 000 and 0.67 < $ < 1 and with s =
0.0476. Equation A-5 proposed by Turton and Levenspiel
(1986) predicts reliably the drag coefficient for spheres
(Hartman and Yates, 1993). An iterative solution of eq
1 together with eq A-6 can serve as a check on the
predictions provided by the explicit formula A-1.

Literature Cited
Becker, H. A. The Effects of Shape and Reynolds Number on Drag
in the Motion of a Freely Oriented Body in an Infinite Fluid. Can.
J. Chem. Eng. 1969,37,85.
Broadhurst, T. E.; Becker, H. A. Onset of Fluidization and Slugging
in Beds of Uniform Particles. AIChE J. 1975,21, 238.
Clark, N. N.; Gabriele, P.; Shuker, S. Drag Coefficient of Irregular
Particles in Newtons Settling Regime. Powder Technol. 1989,
59, 69.

Appendix
In this appendix, we provide other correlations based
on the data of Pettyjohn and Christiansen (1948) for the
steady-state free fall of isometric, nonspherical particles
in Newtonian fluids. The Reynolds number a t the terminal
velocity can be directly estimated as a function of the
Archimedes number and the sphericity from
log Re,(Ar,$) = log Re,(Ar,l)

+ P(Ar,$)

(A-1)

Clift, R.;Grace, J. R.Bubbles, Drops and Particles; Academic Press:


New York, 1978.
Christianaen, E. B.; Barker, D. H. The Effect of Shape and Density
on the Free Settling of Particles at High Reynolds Numbers.
AIChE J. 1966,11,145.
Churchill, S. W.; Usagi, R. A General Expression for the Correlation
of Rates of Transfer and Other Phenomena. AIChE J. 1972,18,
1121.

Flemmer, R.L. C.; Banks, C. L. On the Drag Coefficient of a Sphere.


Powder Technol. 1986,48,217.
Haider, A.; Levenspiel, 0. Drag Coefficient and Terminal Velocity
of Spherical and Nonspherical Particles. Powder Technol. 1989,
58,63.

where
log Re,(Ar,l) = -1.27380 + 1.04185 log Ar O.O60409(log ArI2 + O.O020226(log ArI3 (A-2)

P(Ar,$)= -0.071876(1- $1 log Ar -

0.023093(1- $)(log Ar)2 + 0.0011615(1 -$)(log Ar)3 +


0.075772(1- $)2 log Ar (A-3)

for 0.01 < Ret < 16 OOO and 0.67 < $ < 1 and with s =
0.0331. As can easily be seen, the right-hand sides of eqs
A-2 and A-3do not contain the terminal velocity. Equation
A-1 enables, therefore, the direct computation of the
terminal velocity of a given particle under different
conditions of operation. Thus, eqs 9 and A-1 are complementary to one another and cover the problems which
occur very frequently in engineering practice.
An equation for the drag coefficient of nonspherical
particles was also developed as a function of Ret and $:

where
24
CD(Ret,l)= ~ (+ 0.173
1
t

0.413
1 16300Re:.0e

(A-5)

Hartman, M.; Coughlin, R.W. On the Incipient Fluidized State of


Solid Particles. Collect Czech. Chem. Commun. 1993,58,1213.
Hartman, M.; Yates, J. G. Free-fall of Solid Particles through Fluids.
Collect. Czech. Chem. Commun. 1993,58, 961.
Hartman, M.; H a v h , V.; Tmka, 0.;chk$,M. Predicting the FreeFall Velocities of Spheres. Chem. Eng. Sci. 1989,44,1743.
Hartman, M.; Vesely, V.; Svoboda, K.; H a v h , V. Explicit Relationships for the Terminal Velocity of Spherical Particles. Collect.
Czech. Chem. Commun. 1990,55,403.
Hartman, M.; Trnka, 0.;Havlh, V. A Relationship to Estimate the
Porosity in Liquid-Solid Fluidized Beds. Chem. Eng. Sci. 1992,
47, 3162.
Kunii, D.; Levenspiel, 0. Fluidization Engineering, 2nd ed.;
Butterworth-Heinemann: Boston, 1991.
Mandelbrot, B. B. Fractab: Form, Chance, and Dimension; Freeman: San Francisco, 1977.
Pettyjohn, E. S.; Christiansen, E. B. Effect of Particle Shape on
Free-Settling Rates of Isometric Particles. Chem.Eng. Prog. 1948,
44 (2), 157.

Svoboda, K.; Hartman, M. Influence of Temperature on Incipient


Fluidization of Limestone, Lime, Coal Ash, and Corundum. Id.
Eng. Chem. Process Des. Dev. 1981,20, 319.
Turton, R.;Levenspiel, 0. A Short Note on the Drag Correlation for
Spheres. Powder Technol. 1986,47,83.
Yates, J. G. Fundamentals of Fluidized-Bed Processes; Butterworths: London, 1983.
Received for review November 17, 1993
Revised manuscript received April 11, 1994
Accepted May 24, 1994O
Abstract published in Advance ACS Abstracts, July 1,1994.

Potrebbero piacerti anche