Sie sind auf Seite 1von 7

Powder Technology, 72 (1992) 31-37 31

Scaling considerations for circulating fluidized bed risers


G. S. Patience*, J. Chaouki
Dkpatiement de Gt!nie Chimique, Ecole Polytechnique de Monttial, Monheal, Que., H3C 3A7 (Canada)
F. Berruti and R. Wong
Department of Chemical and Petroleum Engineering, University of Calgay, Calgary, Alta., T2N I N4 (Canada)
(Received June 10, 1991; in revised form December 27, 1991)
Abstract
The ratio between actual gas velocity to particle velocity in the hydrodynamically fully developed region of
Circulating Fluidized Bed risers (CFB) may be approximated by cp = 1 +5.6/Fr +0.47Frp =U&VP. This ratio,
termed the slip factor, is about 2 at operating conditions characteristic of industrial risers several meters in
diameter and agrees with observations of J. M. Matsen (in D. L. Keairns (ed.), Fluidization Technology, Vol. 2,
Hemisphere, 1976, p. 135). The proposed relationship between the gas and solids velocity is an adequate first
approximation to estimate gas and solids residence times, blower capacity and standpipe length.
Introduction
Circulating Fluidized Beds (CFB) are being consid-
ered as alternatives to more conventional Fluidized
Bed processes because of their apparent intrinsic ad-
vantages, including short and controllable residence
times for the gas and solids, high turn down ratios and
flexibility. Industrial scale plants for coal combustion,
aluminum oxide calcination, catalytic cracking,
Fischer-Tropsch synthesis successfully employ this tech-
nology. Contractor and Chaouki [l] and Gianetto et
al. [2] have discussed a number of potential catalytic
processes that are likely candidates for CFB technology.
The main difference between bubbling, turbulent beds
and CFB risers is gas velocity. Typical gas velocities
in CFBs range from 2-10 m s-l. At these velocities,
solids are readily entrained by the gas and are carried
to the top of the vessel. Cyclones and rough cut sep-
arators separate the solids from the gas phase. The
solids are returned to the riser bottom by a standpipe.
The longitudinal solids hold-up in the riser, discussed
by Yerushalmi et al. [3] and by Li and Kwauk [4],
exhibits a relatively dense region at the solids entry
point and a dilute phase above it. A number of models
have been proposed to characterize the riser hydro-
dynamics. Rhodes and Geldart [5] used the entrainment
model, developed for fluidized beds by Wen and Chen
[6], to describe the dilute phase and adapted a classical
*Presently at E.I. du Pont de Nemours & Co., Wilmington,
DE 19880. USA.
two phase model of a bubbling fluidized bed to the
dense phase. Kunii and Levenspiel[7] adopted a similar
approach and correlated decay constants based on a
number of experimental investigations to model the
decrease in solids hold-up along the riser. Also, mea-
surements of the internal flow structure of the riser
by Hartge et al. [ES], Bader et al. [9], and Bolton and
Davidson [lo] indicated that large radial gradients exist,
with significantly higher concentrations of solids near
the wall.
A complete description of the hydrodynamics of such
a flow structure is difficult. The gas-solid flow is typically
characterized by large relative velocities between the
two phases. Two mechanisms have been proposed to
account for the difference in velocity: Yerushalmi et
al. [3] suggested that particles agglomerate into clusters
whose void fraction approaches E,.,,~, whereas Rhodes
et al. [ll] and Berruti and Kalogerakis [12] postulated
that particles ascend in the core in a dilute phase and
descend along the wall as a dense annulus. Ishii et aE.
[13] have recently incorporated both mechanisms into
a clustering annular flow model.
Most of the studies on the hydrodynamics of CFB
systems reported in the literature have been conducted
using laboratory scale units (i.e. relatively short and
narrow). Scale-up to industrial reactors several meters
in diameter and tens of meters in height is uncertain
at best. Experimental rigs. are not only limited by
diameter and height constraints but also by the maximum
circulation rates attainable. Matsen [14] reported that
typical industrial FCC units operate at solids fluxes
0032-5910/92/$5.00
0 1992 - Elsevier Sequoia. All rights reserved
32
between 500 and 1500 kg m- s-l. The majority of
experimental rigs employ non-mechanical devices and
Geldart group B particles, which facilitate circulation
rate control but may ultimately limit the maximum
solids fluxes attainable. Although CFB boilers generally
operate at solids circulation rates less than 100 kg m-
S -I, catalytic reactors require different operating con-
ditions.
Despite the growing body of literature, more fun-
damental information on the hydrodynamics of large
scale CFB reactors is needed to assess the potential
of this technology and to establish design criteria. Scale-
up parameters are useful for the design of industrial
CFB units. These parameters should not only estimate
the overall pressure drop for a given gas velocity and
circulation rate, necessary to size compressors and the
standpipe, but also adequately predict reactor per-
formance including gas-solids contact efficiency and
heat transfer characteristics. The internal flow structure
of small experimental units is well understood, but Dry
and La Nauze [15] suggest that the symmetry of the
radial solids distribution measured in small units may
not apply to large units. However, experiments con-
ducted in a commercial FCC riser by Saxton and Worley
[16] using a radiation attenuation technique, indicate
that a two-phase type flow pattern might adequately
describe the flow phenomenon. In addition to the
uncertainties in scaling-up the riser diameter, few studies
address the effect of height on the hydrodynamics. In
tall risers, differences in the gas velocity between the
top and bottom of 50% are conceivable at high cir-
culation rates. Grace [17] indicates that further com-
plications may arise from the exit and entrance effects,
wall intrusions or roughness and the coefficient of
restitution of the particles. Glicksman et al. [18] doc-
umented the increase in the solids void fraction by
changing the geometry of the exit from a smooth elbow
to a sharp one and presented data suggesting that
objects intruding into the riser may significantly influence
the local solids hold-up.
Glicksman et al. [18], Ishii et al. [13] and Ishii and
Murakami [19] proposed scaling laws to predict the
behaviour in large scale units. Scale-up criteria were
derived based on the principles of fluid-particle systems.
The criteria were then verified in geometrically similar
small scale lab units. Axial solids hold-up and pressure
fluctuations were generally used as the basis for com-
parison. Despite the differences in derivations, Glicks-
man et al. [18] maintain that the scaling laws proposed
in the literature are not dissimilar. They examined two
units with a 34 mm and 152 mm square cross section.
Ishii et al. [13] developed scaling parameters based on
the Clustering Annular Flow Model and validated the
theory experimentally in two geometrically similar units
200 mm and 50 mm in diameter. However, only very
low circulation rates and gas velocities were considered
(U,<2 m s-l, G,<20 kg mm2 s-l). Jazayeri [36] de-
veloped a graph, based on the data of Van Swaaij et
al. [20], that predicts the suspension density at various
gas velocities and circulation rates.
In the present work, we consider a large pool of
data, measured in laboratory and pilot scale risers, to
develop a generalized scaling criterion.
Slip factor as a scaling law
The slip factor is the ratio of the actual gas velocity
to particle velocity,
4p = U&V
P (1)
The average particle velocity, VP, is evaluated based
on the solids circulation rate,
VP =G,/ p,(l - 6)
(2)
and the void fraction, E, is calculated assuming that
the time average pressure drop is attributable only to
the hydrostatic head of the solids,
E = 1 - dpl(p,g dz)
(3)
Clearly, neglecting the particle acceleration contri-
bution to the pressure drop restricts the analysis to
the hydrodynamically fully developed region of the riser.
Moreover, eqn. (3) ignores the wall shear stress con-
tribution to the pressure drop, which may be significant
under certain operating conditions as shown by Van
Swaaij et al. [20] using y-ray absorption. Their results
indicated that at low riser density conditions the mea-
sured pressure drop was systematically higher than the
hydrostatic head of solids and at high densities it was
lower. They showed that the wall shear stress in the
fully developed region is 20-40% of the total pressure
drop at gas velocities greater than 13 m s-l and mass
fluxes greater than 180 kg me2 s-. At gas velocities
below 6 m s-l and circulation rates up to 350 kg m-
S - the contribution of the shear to the measured
pressure drop was negative and less than 25% of the
total pressure drop.
The slip factor is not commonly reported in the
scientific literature. However, Bolton and Davidson [lo],
Yang [22], and Kunii and Levenspiel [7] have used the
slip velocity as a parameter to model the hydrodynamics
of experimental risers,
v,, = U,lE - VP
(4)
In addition, the slip velocity, VsI , has been used as a
parameter for heat and mass transfer correlations.
Rhodes and Geldart [5] and Patience and Chaouki [21]
have assumed that Vsl is equal to the particle terminal
velocity from which the particle velocity is calculated.
33
However, work by Patience et al. [23] clearly shows
that the latter assumption is not true and better criteria
are required to evaluate the average solids velocity.
In large industrial scale FCC risers Matsen [14]
reported that the slip factor cp, is approximately equal
to 2 and hence the particle velocity equals Ug/2e.
Comparisons between large scale industrial units and
experimental units is complicated not only because of
the differences in geometry but also because of the
differences in operating conditions; high circulation
rates, high temperatures and pressures. Sternerding [24]
showed that the riser pressure drop was independent
of the transport gas properties in the range,
1.4 lo- <p < 3.7 lo- Pa-s
and
0.33 < pg <5. 0 kg rnm3.
However, Findlay and Knowlton [32] suggest that the
solids mass fraction is inversely proportional to the gas
density to the 0.6 power. More study on the effect of
gas properties on suspension density is required.
Van Swaaij et al. [20] reported data with FCC catalyst
in the developed region of an 0.18 m diameter riser
at circulation rates up to 500 kg mm2 s-l and their
experimental circulation rates are compared with the
predicted values assuming a slip factor of 2 in Fig. 1.
Reasonable agreement between the predictions and the
experimental values is evident despite the wide range
of operating conditions. The gas velocity varied between
4.3 to 15.1 m s- and the mass flux from 133 to 514
kg m- s-.
Figure 2 illustrates the longitudinal solids density at
gas velocities of 6 and 8 m s-l and various solids
circulation rates in a riser 83 mm in diameter and 5
600

E 500
\
r
-
r
400
E:
: 300
4
e
2 200
2
a
100
0
I I r I I
op=z
l p = 1+5.6/Fr+0.47Fr,041
0 100 200 300 400 500 600
Experimental Mass Flux (kg/ms)
Fig. 1. Data of Van Swaaij et al. [20] compared with predictions
assuming a slip factor of 2, and a slip factor = 1 +5.6/Fr +0.47Fr,4.
5
I I I I 1 I
V 0 l T
V v oe
4 Us= 6 m/s -
wcm 0 Gs= 102 kg/n+
l Gs= 108 kg/m s
u,= 6 m/s
2
.M
0
V G,= 26 kg/n+
!z 2
I Gs= 07 kg/m s
l
0 100 200 300 400 500 600 700
P .u,p (kdm3)
Fig. 2. Longitudinal suspension density, Patience [25].
W
U,= 6 m/s
0 G,= 102 kg/n+
l G,= 196 kg/m s
WI = 6 m/s
V Gs= 26 kg/ms
v G,= 67 kg/ms
l
, w
t
C
I I
02 10 20 30 40
Slip Factor, (p
Fig. 3. Variation of the slip factor along the riser length.
m tall as reported by Patience [25]. Sand with a mean
diameter of 275 ,um and a density of 2630 kg rnF3 was
used in all experiments. The suspension density in the
riser decays exponentially, reaches a steady value around
the middle of the column and eventually increases
toward the top. The exit configuration was an abrupt
reducing right angle. Brereton [26] attributes the in-
crease in the density at the exit to internal refluxing
of solids.
Experimental data, shown in Fig. 2, are replotted in
terms of slip factor, instead of suspension density, in
Fig. 3. At the entrance, the calculated slip factors are
greater than 20 and at the top of the riser they are
greater than 7. However, in the centre of the riser,
above the acceleration region, the flow profile is flat
and perhaps hydrodynamically fully developed. The
slip factor is close to 2 in this region, which agrees
with measurements in risers up to 1.5 m in diameter
34
that operate at elevated circulation rates with Group
A powders as reported by Matsen [14]. At the entrance,
the slip factor increases with mass flux, which may be
attributable, in part, to the overprediction of the solids
fraction because the acceleration contribution to the
pressure drop was neglected. Also, the slip factor ap-
pears to be greater at higher gas velocities at the top
of the riser.
In Fig. 4, data measured by Wong [27] in a 3 m tall
riser 50 mm in diameter is shown. The apparent slip
factor in the acceleration region is plotted together
with the actual slip factor in which the acceleration
contribution of the particles is taken into account. The
contribution of particle acceleration to pressure drop,
hence density, was calculated based on the work by
Weinstein and Li [30]. The figure indicates that, although
the actual slip factor in the acceleration zone is greater
than 2, ignoring the acceleration effect greatly over-
estimates the slip factor, hence total solids hold-up.
The slip factor calculated in the hydrodynamically
developed region, based on data reported by a number
of researchers, is plotted against the gas velocity in
Fig. 5. Table 1 summarizes the particle characteristics
and riser geometry of each study. Both Geldart A and
B powders were used in the experiments for which the
particle terminal velocities vary between 0.2 m s-l and
2 m s-l. A slip factor of 2 correlates the data reasonably
at gas velocities between 6 and 12 m s-l. This agrees
with the value reported by Matsen [14] for industrial
risers, which typically operate at velocities greater than
8 m s-l.
To account for the increase in q with decreasing gas
velocity, as shown in the figure, the following relationship
is proposed:
cp = 1 + 5.6jFr +0.47Frp4*
(5)
6-
I I I
0 U =
GE
7. 9 m/ s
= 57 kg/ m' s
5- d; = 174 / . un
0 (p experimental
0 0 0 corrected for acceleration
9. 4-
i
9 0
0 E 3-
0 0
.? 0
Ci
2 D
CJ 0 00
1 .o
Height (m)
Fig. 4. Slip factors in the hydrodynamically developing region
(acceleration) of a riser, Wong [27].
10
I I I 1 I I I
_- -
( P = 1+5~3/Fr+0.47Fr,~.~~
06
0 2 4 6 8 10 12 14 16
up b/s)
Fig. 5. Slip factors in the hydrodynamically fully developed region
at different gas velocities. Data referenced in Table 1.
Agreement between predicted and experimental cir-
culation rates of Van Swaaijs data [20] using this
correlation is good, as shown in Figs. 1 and 5. The fit
is superior compared to the single parameter estimate
of rp=2.
Equation (5) suggests that, at gas velocities much
greater than single particle terminal velocities, the solids
hold-up increases with diameter, i.e.
4(1- l ) = (1 +5.6(Dg)o.5/U, + 0.47Fr,0-41)cG,lUg
(6)
The effect of riser diameter on suspension density has
not been fully explored. Arena et al. [31] studied two
risers 41 mm and 120 mm in diameter and concluded
that the density increased with diameter at the same
operating conditions. Kato et al. [34] reported that, for
small tubes, the density increased with diameter to the
0.4 power, whereas a power of 0.2 fit data collected
by Findlay and Knowlton [38] better. Larger riser
diameters were used in the latter study. The correlation
predicts that the solids hold-up is relatively independent
of particle characteristics as long as the superficial gas
velocity is much greater than the particle terminal
velocity. Moreover, it suggests that the solids hold-up
is relatively insensitive to gas properties.
At high gas velocities, typical of pneumatic conveying,
5.6/ Fr tends to zero and cp approaches 1 +0.47Frp41.
Typically, FCC risers operate at slip factors near two;
Govier and Aziz [37] suggest that in pneumatic conveying
1 < cp < 2, which agrees with the proposed correlation.
Brereton [26] reports large differences in the slip
factor between a smooth exit and an abrupt geometry.
The slip varies between 1.88 and 2.32 for sand particles
(open squares in Fig. 5) in a CFB with a smooth exit.
The slip factor for sand in the same unit with an abrupt
exit geometry (filled squares) varies from 8.2 at low
gas velocities to 3.6. Most of the data reported in the
35
TABLE 1. References and experimental conditions for the data in Fig. 5
Key
Study Riser geometry Particle properties Remarks
V Arena et al. [LB]
0 Brereton [26]
n
A
0 Patience [25]
+ Rhodes and Geldart [5]
v Stemerding [24]
0 Van Swaaij et al. [20]
A Wong [27]
0
Estimates.
D Height Type
dP
Cm) Cm)
(pm) Zg mm3)
(m SC)
0.041 6.4 Glass 88 2600 0.46 Smooth exit
0.152 9.3 Sand 148 2650 0.99 Smooth exit
0.152 9.3 Sand 148 2650 0.99 Abrupt exit
0.152 9.3 Alumina 65 3500 0.36 Abrupt exit
0.083 5 Sand 275 2630 1.9 Abrupt exit, 20 < T< 250
0.152 5 FCC 64 1800 0.2 Abrupt exit
0.051 2-10 FCC 65 1600 0.18 , Developed region, 15 < T < 500
0.18 FCC 65 1400 0.16 Developed region
0.05 3 Sand 93 2500 0.48 Abrupt exit
0.05 3 Sand 174 2500 1.2 Abrupt exit
CFB with the abrupt exit geometry varies between 4
and 6. The slip factor with the alumina particles and
an abrupt exit geometry (open triangles) lies between
2.1 and 4.7. The large slip velocities reported by Brereton
[36] may be a result of the contribution of two different
phenomena. Firstly, the abrupt exit configuration, in
this unit, might affect the solids behaviour significantly.
Secondly, secondary gas was supplied to the column
which extends the acceleration region of the riser. The
combination of the abrupt exit and extended accel-
eration zone may prevent the establishment of a fully
developed flow region. However, experiments by Pa-
tience [25] and Wong [27], conducted in short risers
with abrupt right angle exits, for which the acceleration
zone would presumably affect the solids loading more
than in taller units, gave much lower slip factors.
Whereas values of rp greater than those predicted
by the proposed correlation are evident in the work
of Brereton [26], smaller values of cp are calculated
from data obtained in the experimental unit of Arena
et al. [28] as reported by Louge and Chang [29] and
Yang [22]. In this case, the slip velocity approaches
the single particle terminal velocity. At gas velocities
of 5 m s-l and circulation rates ranging from 80 to
390 kg rnp2 s-l the slip varies between 1.15 and 1.055.
Inaccuracies in the measurement of the circulation rate
could explain the differences in slip factor as reported
by Brereton [26] and Arena et al. [28]. The slip factor
is inversely proportional to the mass flux; hence, under-
estimating the circulation rate will result in large cal-
culated slip factors and overestimating the rate gives
low values of q. Patience and Chaouki [33] show that
using the particle velocity along the downcomer wall,
the technique used by Brereton [26] in the experiments
with sand, can underestimate the circulation rate by
up to 40%. Brereton [26] used a butterfly valve technique
in the experiments with alumina and eqn. (5) fits these
data well compared to the experiments with sand where
the solids circulation rates were made by tracking the
wall velocity in the downcomer.
The general agreement between slip factors reported
for industrial FCC reactors and experimental risers
suggests that pressure drop predictions may be possible
without having to develop large and costly pilot plants.
That is, given the desired residence time of the gas
and solids, the blower requirements and height and
diameter of the reactor may be calculated. The insistence
of geometrical similarity, as suggested by Ishii et al.
[13] and Ishii and Murakami [19], appears to be too
restrictive. In the present work, the proposed slip factor
model is shown to provide a reasonable estimate of
the average gas and solids residence times. Particle
characteristics do not affect the slip factor as long as
the operating gas velocity is significantly greater than
the particle terminal velocity (in the fully developed
flow regime of the riser). Van Swaaij et al. [20] and
Stemerding [24] used Geldart Group A powders,
whereas Patience [25], Brereton [26] and Wong [27]
used Geldart Group B particles. Zhang et al. [35] suggest
that particle density and particle size distribution of
Geldart A materials do not affect the radial voidage
profile when comparing systems operating at the same
suspension density.
Conclusions
Correlations available in the literature do not seem
to predict the relationship between the gas and solids
velocity adequately. An examination of a large pool of
data from both experimental laboratory scale CFBs and
industrial units indicates that the ratio of interstitial
36
gas velocity to particle velocity, or slip factor, is ap-
proximately equal to 2 in the hydrodynamically fully
developed flow regime of risers at superficial gas ve-
locities greater than 6 m s-. However, an improved
relationship is also proposed,
which better describes the slip factor dependence on
gas velocity, riser diameter, and particle properties.
The suspension density increases with diameter and
decreases with increasing gas velocity. At high gas
velocities cp approaches 1 + 0.47Frto.41. This relationship
applies to the region, above the acceleration zone at
the entrance and below the deceleration zone at the
exit where values of C+J are greater. Therefore, to estimate
gas and solids residence times, blower capacity, and
standpipe length requirements the entrance and exit
effects must be considered. However, entrance and exit
lengths are typically much shorter than the developed
region, so errors in ignoring the pressure drop con-
tribution in a 40 m tall riser would be less than 10%.
The slip factor, cp, is reported to be independent of
gas properties and particle characteristics (at gas ve-
locities much greater than the single particle terminal
velocity).
List of symbols
D riser diameter
d,
particle diameter
dPld.z
pressure gradient
Fr Froude number, U&D).
Fr,
Froude number, V&gD)~5
GS
solids flux in riser
g
gravitational constant
P time-averaged pressure
T temperature
utx
superficial gas velocity
v,
particle velocity
KI
slip velocity
v,
particle terminal velocity
z vertical co-ordinate
Greek letters
E void fraction
E
mf
void fraction at minimum fluidization
Ps
gas density
4
particle density
cp
slip factor
References
1 R. M. Contractor and J. Chaouki, in P. Basu, M. Horio and
M. Hasatani (eds.), Circulating Fluidized Bed Technology I I I ,
Pergamon, Oxford, 1991, p. 39.
A. Gianetto, S. Pagliolico, G. Rover0 and B. Ruggeri, Chem
Eng. Sk., 45 (8) (1990) 2219.
J. Yerushalmi, M. Cankurt, D. Geldart and B. Liss, AI ChE
Symp. Ser., 74 (76) (1978) 1.
Y. Li and M. Kwauk, in J. R. Grace and J. Matsen (eds.),
Fluidization ZZJ Plenum, New York, 1980, p. 537.
M. J. Rhodes and D. Geldart, Chem. Eng. Res. Des., 67 (1989)
20.
10
11
12
13
14
15
16
17
18
C. Y. Wen and L. H. Chen, AZChE J ., 28 (1982) 117.
K. Kunii and 0. Levenspiel, Powder Technol., 61 (1990) 193.
E.-U. Hartge, Y. Li and J. Werther, in P. Basu (ed.), Circulating
Fluidized Bed Technology, Pergamon, Toronto, 1986, p. 153.
R. Bader, J. Findlay and T. M. Knowlton, in P. Basu and
J. F. Large (eds.), Circulating Fluidized Bed Technology I I ,
Pergamon, Oxford, 1988, p. 123.
L. W. Bolton and J. F. Davidson, in P. Basu and J. F. Large
(eds.), Circulating Fluidized Bed Technology I I , Pergamon,
Oxford, 1988, p. 139.
M. 3. Rhodes, P. Laussmann, F. Villain and G. Geldart, in
P. Basu and J. F. Large (eds.), Circulating Fiuidized Bed
Technology ZZ, Pergamon, Oxford, 1988, p. 20.
F. Berruti and N. Kalogerakis, Can. J. Chem. Eng., 67 (1989)
1010.
H. Ishii, T. Nakajima and M. Horio, J. Chem. Eng. J pn., 22
(5) (1989) 484.
J. M. Matsen, in D. L. Keairns (ed.), Fluidization Technology,
Vol. 2, Hemisphere, 1976, p. 135.
R. J. Dry and R. D. La Nauze, Chem. Eng. Prog., 86 (7)
(1990) 31.
A. L. Sazton and A. C. Worley, Oil Gas J ., 68 (20) (1970)
84.
19
20
21
22
23
24
25
26
27
28
J. R. Grace, Chem. Eng. Sci., 45 (8) (1990) 1953.
L. R. Glicksman, D. Westphalen, C. Brereton and J. Grace,
in P. Basu, M. Horio and M. Hasatani (eds.), Circulating
Fluidized Bed Technology I I Z, Pergamon, Oxford, 1991, p. 119.
H. Ishii and I. Murakami, in P. Basu, M. Horio and M.
Hasatani (eds.), Circulating Fluidized Bed Technology I I I , Per-
gamon, Oxford, 1991, p. 125.
W. P. M. Van Swaaij, C. Buurman and J. W. Van Breugel,
Chem. Eng. Sci., 25 (1970) 1818.
G. S. Patience and J. Chaouki, Chem. Eng. Res. Des., 68 (A)
(1990) 301.
W.-C. Yang, in P. Basu and J. F. Large (eds.), Circulating
Fluidized Bed Technology I I , Pergamon, Oxford, 1988, p. 181.
G. S. Patience, J. Chaouki and G. Kennedy, in P. Basu, M.
Horio and M. Hasatani (eds.), CircuZating FZuidized Bed Tech-
nology I I I , Pergamon, Oxford, 1991, p. 599.
S. Sternerding, Chem. Eng. Sci., 17 (1962) 599.
G. S. Patience, Ph.D. Dissertation, Ecole Polytechnique de
Montreal, Canada, 1990.
C. M. H. Brereton, Ph.D. Dissertation, University of British
Columbia, Vancouver, Canada, 1987.
R. Wong, M.Sc. Thesis, University of Calgary, Canada, in
preparation.
U. Arena, A. Cammarota and L. Pistone, in P. Basu (ed.),
Circulating Flutdized Bed Technology, Pergamon, Toronto, 1986,
p. 119.
29 M. Louge and H. Chang, Powder TechnoZ., 60 (1990) 197.
30 H. Weinstein and J. Li, Powder Technoi., 57 (1989) 77.
37
31 U. Arena, A. Cammarota, L. Massimilla and D. Pirozzi, in
P. Basu and J. F. Large (eds.), Circulating Fluidized Bed
Technology I I , Pergamon, Oxford, 1988, p. 223.
32 J. Findlay and T. M. Knowlton, Final Report: Pipeline Gas
from Coal (I GT Hydrogasification Process), (1980) 306.
33 G. S. Patience and J. Chaouki, in P. Basu, M. Horio and
M. Hasatani (eds.), Circulating Fluidized Bed Technology I I I ,
Pergamon, Oxford, 1991, p. 627.
34 K. Kato, Y. Ozawa and H. Endo, in K. Ostergaard and A.
Sorenson (eds.), Fluidization, Engineering Foundation, New
York, 1986.
35 W. Zhang, Y. Tung and F. Johnsson, Chem. Eng. Sci., 46
(12) (1991) 3045.
36 B. Jazayeri, Hydrocarbon Process., 70 (5) (1991) 93.
37 G. W. Govier and K. A&, Z7ze Flow of Complex Mixtzues in
Pipes, Van Nostrand Reinhold, Toronto, 1972.

Das könnte Ihnen auch gefallen