Вы находитесь на странице: 1из 13

Chemical Engineering Science 62 (2007) 3397 3409

www.elsevier.com/locate/ces

Computation of gas and solid dispersion coefcients in turbulent risers and


bubbling beds
Veeraya Jiradilok a , Dimitri Gidaspow a, , Ronald W. Breault b
a Illinois Institute of Technology, Chicago, IL, USA
b US Department of Energy, Morgantown, WV, USA

Received 1 July 2006; received in revised form 30 January 2007; accepted 30 January 2007
Available online 24 March 2007

Abstract
A literature review shows that dispersion coefcients in uidized beds differ by more than ve orders of magnitude. To understand the
phenomena, two types of hydrodynamics models that compute turbulent and bubbling behavior were used to estimate radial and axial gas
and solid dispersion coefcients. The autocorrelation technique was used to compute the dispersion coefcients from the respective computed
turbulent gas and particle velocities.
The computations show that the gas and the solid dispersion coefcients are close to each other in agreement with measurements. The
simulations show that the radial dispersion coefcients in the riser are two to three orders of magnitude lower than the axial dispersion
coefcients, but less than an order of magnitude lower for the bubbling bed at atmospheric pressure. The dispersion coefcients for the bubbling
bed at 25 atm are much higher than at atmospheric pressure due to the high bed expansion with smaller bubbles.
The computed dispersion coefcients are in reasonable agreement with the experimental measurements reported over the last half century.
2007 Elsevier Ltd. All rights reserved.
Keywords: Fluidization; Gas-particle ow; Computational uid dynamics; Reynolds stresses

1. Introduction
Traditional design of gasiers for the FutureGen project and
other reactors requires the knowledge of dispersion coefcients,
as demonstrated by Breault (2006). However, they are known to
vary by 5 orders of magnitudes (Gidaspow et al., 2004; Breault,
2006).
From experimental investigations over the last half-century,
the dispersion coefcients are known to be large for large diameter bubbling beds and small at low gas velocities. Surprisingly,
they differ by two to three orders of magnitudes at the same
gas velocity. Hence, a better understanding of the phenomena
causing such large differences is needed. This study presents a

computational method of determining the gas and solid axial and radial dispersion coefcients in bubbling beds and
risers.
The physical denition of dispersion coefcients is based
on the kinetic theory of gases (Bird et al., 2002; Chapman
and Cowling, 1970) and granular ow (Gidaspow, 1994). For
diffusion of gases or particles the diffusivity, D, is dened as
the mean free path, L, times the average velocity, C, as shown
below:
D = L C,
where the peculiar velocity, C is given by
C = c v,

Corresponding author. Department of Chemical Engineering, Illinois In-

stitute of Technology, 10 west 33rd street, Chicago, IL 60616, USA.


Tel.: +1 312 567 3045; fax: +1 312 567 8874.
E-mail address: gidaspow@iit.edu (D. Gidaspow).
0009-2509/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2007.01.084

(1)

(2)

where c is instantaneous velocity and v is the velocity averaged


over velocity space, as shown below:


1
v=
cf (c) dc with n = f (c) dc.
(3)
n

3398

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

Tartan and Gidaspow (2004). measured the instantaneous


particle velocities, c, in a riser and computed the hydrodynamic
velocities, v, as a function of time. From these two types of
velocities the particle stresses, CC, and the Reynolds stresses,
v  v  , were computed, where v  =vv and v is the time averaged
velocity. In the center of the riser the velocity, v, could be
obtained from Poiseuille ow.
The mean free path was obtained from the average velocity
and collision time, :

Table 1
Hydrodynamic viscosity model
Continuity equations

j(g g )
+ (g g vg ) = 0
jt
j(s s )
+ (s s vs ) = 0
jt
Momentum equations

L = C ,

(4)

since C and C 2 differ by only 10%.


The dispersion matrix can be dened as
DP = CC .

j(s s vs )
+ (s s vs vs ) = Ps I + s + B (vg vs )
jt

(6)

This denition is identical to that used in single phase turbulent ow (Hinze, 1959). For gasparticle ow we have two
types of dispersion coefcients, that for the gas and for the particles. In this paper we computed only the turbulent dispersion
coefcients.
Hence the dispersion coefcients in the x and y directions
are computed from the normal Reynolds stresses in the x and
y directions. For the riser, the normal Reynolds stresses in the
direction of ow are shown here to be about two orders of
magnitude larger than the radial Reynolds stresses due to the
fact that the radial velocities are small compared to the axial
velocities. This provides an explanation for the large anisotropy
of the dispersion coefcients in the riser and not in bubbling
beds, where the velocities in radial and axial directions are
similar.

Constitutive equations
(1) Denitions

g + s = 1
(2) Gas pressure
g
Pg = g RT
(3) Stress tensor (i = gas or solid)

i = 2i Di + (i 23 i ) tr(Di )I


with
Di = 21 [vi + (vi )T ]
(4) Empirical particulate phase viscosity and stress model
Ps = G(g ) s
G(g ) = 108.686g +8.577 dyne/cm2

s = 0.1651/3
s g0 poise
(5) Fluid-particulate interphase drag coefcients for  < 0.8 (based on the
Ergun equation)

2. Hydrodynamics models
The physical principles used are the conservation laws of
mass, momentum and energy for each phase, the uid phase and
the particulate phases. This approach is similar to that of Soo
(1967) for multiphase ow and of Jackson (1985) for uidization. A Newtonian or power law constitutive equation for the
surface stress of phase k will depend at least on its symmetrical gradient of velocity. The kinetic theory of granular ow
provides a physical motivation for such an approach (Gidaspow,
1994). Hence, the general balance laws of mass and momentum for each phase, with phase change, are given by Eqs. (7)
and (8) and the constitutive equation for the stress is given by
Eq. (9).
Continuity equation for phase k:
j
k.
(k k ) + (k k vk ) = m
jt


s
s
k g g
+
g
k=g,s

(5)

Similarly for turbulent oscillations, the dispersion matrix is dened as


DT = v  v  .

j(g g vg )


+ (g g vg vg ) = P I + g B vg vs + g g
jt

(7)

 = 150

2s g
g s
+ 1.75
| g s |
g d p
2g dp2

for  > 0.8 (based on the empirical correlation)


3
4

 = Cd

g s |g s |
dp

2.65
g

where
Cd =

24
[1 + 0.15Re0.697
]
p
Rep

Cd = 0.44

Rep =

for Rep > 1000

g g dp |vg vs |
g

for Rep < 1000

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409


Table 2
Hydrodynamic kinetic theory model

3399

Table 2 (continued)
where

75
 ds 1/2
384 s

Continuity equations

dil =

j(g g )
+ (g g vg ) = 0
jt

(10) Collisional energy dissipation




4
2 2
s = 3(1 e )s s g0
vs
ds

j(s s )
+ (s s vs ) = 0
jt

(11) Gassolid drag coefcient


Momentum equations

j(g g vg )
+ (g g vg vg ) = P + g B (vg vs ) + g g
jt
j(s s vs )
+ (s s vs vs ) = Ps + s + B (vg vs )
jt
+ s (s g )g
( = 1/3 C 2 )

Fluctuating energy equation for particles


3 j
(s s ) + (s s vs ) = (Ps I + s ) : vs + (
s ) s
2 jt
Constitutive equations
(6) Denitions

B = 150

3
4

B = Cd

g s |vg vs |
dp

(),

g < 0.74

g  0.74

where

0.0214

0.5760 +

4( 0.7463)2 + 0.0044


0.0038
() =
0.0101 +

4(


0.7789)
+ 0.0040

31.8295 + 32.8295

Cd =

g + s = 1

2s g
g s |vg vs |
+ 1.75
,
g dp
2g dp2

24
[1 + 0.15Re0.697
]
p
Rep

Cd = 0.44

(0.74 <   0.82)


(0.82 <   0.97)
( > 0.97)

for Rep < 1000

for Rep > 1000

(7) Gas pressure

g g dp |vg vs |
g

g
Pg = g RT

Rep =

(8) Stress tensor (i = gas or solid)

Boundary conditions for particle phase (Johnson and Jackson, 1987)


(1) Velocity

i = 2i Di + (i 23 i ) tr(Di )I

jvs,w
6s s,max
us,w =

3 s s g0 jn

with
Di = 21 [vi + (vi )T ]

(2) Granular temperature

2
3 s s vs,slip
g0 3/2

j
+
w =
w jn
6s,max w

Ps = s s [1 + 2(1 + e)g0 s ]

s =

2sdil
(1 + e)g0

4
1 + (1 + e)g0 s
5


4

+ 2s s ds g0 (1 + e)
5

where

w =

2 )  g 3/2
3 (1 ew
s s 0
4s,max


4 2

s = s s ds g0 (1 + e)
3

where g0 is the radial distribution function and sdil is the particle phase
dilute viscosity

g0 = 1

sdil

s

1/3 1

s,max

5
=
 dp 1/2
96 p

(9) Granular conductivity of uctuating energy

2
2

6
1 + (1 + e)g0 s
dil + 22s s ds g0 (1 + e)
(1 + e)g0
5

Momentum equation for phase k:


j
(k k vk ) + (k k vk vk )
jt

= k k g + [k ] +
(vl vk ) + m
k vk .

(8)

Constitutive equations for stress (above minimum uidization):


[k ] = [Pk + k vk ][I ] + 2k [Sk ],
[Sk ] = 21 [vk + (vk )T ] 13 vk [I ].

(9)

3400

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

These equations are similar to Bowens (1976) balance laws


for multicomponent mixtures. The principle difference is the
appearance of the volume fraction of phase k denoted by k .
In the case of phases not all the space is occupied at the same
time by all phases, as it is by components due to the negligible size of molecules. As in the case of the mixture equations
for components, the mixture equations for phases show that the
sum of the phase change productions in Eq. (7) is zero and the
sum of the drag forces in Eq. (8) is zero. In convective form,
the phase change momentum in Eq. (8) is zero, insuring invariance under a change of frame of reference for translation.
Eq. (9) is the usual Newtonian expression for the stress which
arises from the assumption that the stress is a function of its
own symmetrical gradient of velocity. For the uid Pk is the
uid pressure. When this form is substituted into the momentum equation, the result is not the usual momentum balance
presented by Gidaspow (1986) and widely used in gasliquid
two-phase ow. It is a slightly modied version of the momentum balance called model B by Bouillard et al. (1989).
In this study we are applying this model to single size
particlegas system with no reaction or phase change. The
particle viscosity and the solid stress are input into the viscosity model given in Table 1. In the kinetic theory version,
Table 2, the particle viscosity and solid stress are automatically

computed. The restitution coefcient is a tting parameter. In


the simulation presented here the drag was also modied for
ow of FCC particles as described in Jiradilok et al. (2006).
The numerical scheme used in this study is the implicit continuous Eulerian (ICE) approach. The model uses donor cell
differencing. The conservation of momentum and energy equations are in mixed implicit form. It means that the momentum
equations are fully explicit. The continuity equations excluding
mass generation are in implicit form.
3. Simulations
3.1. Flow of FCC particles in the riser
3.1.1. System properties
Jiradilok et al. (2006) have already shown that the kinetic
theory model with the EMMS approach, as shown in Table 2
(Yang et al., 2004), is capable of computing turbulent uidization of FCC particles in agreement with experimental data. The
computed granular temperatures, particles viscosities, solid
pressures and oscillation frequencies agreed with experiments.
The simulations were carried out for the riser section of a
circulating uidized bed. A two-dimensional Cartesian coordinate system was used. Initially the riser column was empty and

Outlet
System Geometry and System Properties

Riser diameter

0.186 m.

Riser inlet diameter

0.093 m.

Riser height

8 m.

Particle size

54 m

Particle density

1398 kg/m3

Restitution coefficient, e

0.9

Wall restitution coefficient, ew

0.6

Specularity coefficient, 

0.6

Grid size, (x y)

0.465 cm 2.68 cm

Grid number

42 (radial) 300 (axial)

Time step

5 10-5

Inlet
Fig. 1. System geometry for simulations based on Wei et al. (1998a) experiments.

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

3401

800
0.4
8

700

0.35
7

600

0.3

500

0.25

0.2

300

0.15

200

0.1

Height, m.

400

5
4
3
2
1

0.05

100

0
0.0001
0
5

10

0.001
0.01
0.11
Laminar Granular Temperature m2/s2

15

Fig. 2. (a) Snapshot of solid volume fraction, (b) axial laminar granular temperature prole for Ws = 132 kg/m2 s and Ug = 4.57 m/s.

the velocities of both phases were assumed to be zero. At the


outlet, atmospheric pressure was prescribed. Particularly important is also the specication of appropriate boundary conditions at the wall. For the gas a non-slip boundary condition was
used. For the granular temperature wall boundary condition the
Johnson and Jackson (1987) boundary condition was used. It
was obtained by equating the granular ux to collisional dissipation with a correction for slip.
Fig. 1 gives the system geometry and the system properties
for describing the experiments of Wei et al. (1998a). The main
uidizing gas is air. The solid phase consisted of FCC particles.
A restitution coefcient of 0.9 was used (Jiradilok et al., 2006).

Table 3
Equations for obtaining the averaged velocity and stresses

3.1.2. Flow structure


We have shown that the standard kinetic theory-based CFD
model with a modied drag as suggested by Li group (Yang
et al., 2004) is capable of correctly describing the coexistence
of the dense and dilute regimes for ow of FCC particles in a
riser in the turbulent regime (Jiradilok et al., 2006).
Fig. 2(a) displays the snap shot at 7.2 s for the solid ux of
132 kg/m2 s and the gas velocity of 4.25 m/s. The top part of
the riser is dilute and the bottom part is dense. The structure
at the bottom part is core-annular. There is a low concentration
of solid at the center and a high solid volume fraction near the
wall, which approximately agrees with the experimental data.
Using the kinetic theory model, the laminar granular temperature was computed. The axial prole of laminar granular

3.1.3. Reynolds stresses


Reynolds stresses are produced due to the random velocity
uctuations of the hydrodynamic velocity. This is the principal
characteristic of turbulent ow. The Reynolds stresses are used
to estimate the turbulent part of granular temperature and the
dispersion coefcients. The Reynolds stress can be calculated
as a function of hydrodynamic velocity and mean velocity, as
shown in Table 3.
The time-average values of normal Reynolds stresses per
unit bulk density in the axial direction of gas and solid phases
with various heights are shown as Fig. 3. The values vary as a
function of the radial position for both phases. The oscillations
on the top, 2 m, and bottom, 6 m, parts occur due to the effects
of the inlet and outlet.

The mean velocity


particle

v i (r) =

The normal Reynolds


stress

vi vi =

m
1 
vik (r, t)
m k=1

m
1 
(vik (r, t) vi (r))(vik (r, t) vi (r))
m k=1

i represents x or y directions, m is the total number of data over a given time


period.

temperature is shown as Fig. 2(b). The laminar granular temperature increases with increasing bed height due to the oscillation of individual particles.

3402

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

10

0.25

9
8

0.2
vx vx m2/s2

vyvy m2/s2

7
6
5
4

0.15
0.1

3
2

0.05

1
0
-1

-0.8

-0.6

-0.4

-0.2

-1

-0.8

-0.6

-0.4

-0.2

-1

-0.8

-0.6

-0.4

-0.2

-1

-0.8

-0.6

-0.4

-0.2

0.25

10
9
8

0.2
vx vx m2/s2

vyvy m2/s2

7
6
5
4
3
2

0.15
0.1
0.05

1
0
-1

-0.8

-0.6

-0.4

-0.2

10

0.25

9
0.2

8
vx vx m2/s2

vyvy m2/s2

7
6
5
4
3

0.15
0.1
0.05

2
1

0
-1

-0.8

-0.6

-0.4

-0.2

r/R
Gas phase
Solid phase

r/R
Gas phase

Solid phase

Fig. 3. Axial normal Reynolds stress per bulk density of gas and solid phases
at (a) 2 m, (b) 4 m and (c) 6 m.

Fig. 4. Radial normal Reynolds stress per bulk density of gas and solid phases
at (a) 2 m, (b) 4 m and (c) 6 m.

Due to the turbulences in the direction of ow, the normal


Reynolds stresses per unit bulk density in the axial direction
are higher than those in the radial direction for both phases,
as shown in Fig. 4. We see that the Reynolds stresses per unit
bulk density for the gas and the solid are close to each other.

the viscosity model, as given in Table 1. The system geometry and the system properties are summarized in Fig. 5. For
boundary conditions, the non-slip boundary condition was used
for the gas phase at the wall. For the solid phase, the free-slip
boundary condition was chosen. Initially, the bed was lled with
particles at a solid volume fraction of 0.58. In this study we
simulated bubbling beds under two different pressures, a high
pressure, 25 atm, and a low pressure, atmospheric pressure.

3.2. Bubbling commercial size uidized beds


3.2.1. System properties
To show that the model predicts the effects of pressure on
dispersion coefcients, two bubbling uidizations were simulated. Simulations of bubbling commercial size uidized beds
from a two-dimensional rectangular bed were performed using

3.2.2. Flow structure


Instantaneous snapshots of the solid volume fraction proles
of the bubbling bed are shown in Fig. 6. Figs. 6(a) and (b) were
obtained with atmospheric pressure and 25 atm, respectively.

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

3403

Geometry of the reactors


Height
12.19 m.
Reactor Diameter (OD)
1.22 m.
Reactor Diameter (ID)
1.14 m.
Angle
75
4
Number of inlet cell
Position of outlet right hand side
Outer diameter of the outlet
0.16 m.
There is one cell of outlet on right hand side.
Solid properties
Density
2500 kg/m3
Diameter
500 m
Fluid properties
Air
Grid size
Radial
x = 4.064 cm

y = 16.475 cm

Axial

Number of Grid

Radial Axial

30 74 cells

(Including boundary wall cells)


Operating Condition

Gas velocity

Initial solid volume fraction

Initial bed height

Temperature

Pressure
Case 1
Case 2

0.45 m/s
0.58
4 m.
300 K
Atmospheric pressure, 1 atm
25 atm

Fig. 5. System geometry and operating conditions for bubbling bed simulations.

Table 4
A comparison of expanded height ratio and equivalent bubble diameter at
two pressures
P (atm)

Expanded height ratio (H/Ho)

Equivalent bubble diameter (m)

1
25

1.21
2.40

0.70 0.14
0.58 0.18

bubble diameter can be estimated as a function of the distance


above the distributor in a bubbling uidized bed, as follows:

(10)
DB = 0.54(U0 Umf )0.4 (h + 4 A0 )0.8 /g 0.2 ,

Fig. 6. Instantaneous plots of solid volume fraction eld in bubbling bed


simulations: (a) 1 atm, (b) 25 atm.

The bed expansion ratio is calculated based on the uidized


bed height and the initial bed height,  = Hf /H0 . Using Davidsons bubble-growth model (Darton et al., 1977), the equivalent

where DB is the equivalent bubble diameter, U0 is the superthe minimum uidization velocity, h
cial gas velocity, Umf is
is the initial bed height, 4 A0 is 0.03 m for a porous-plate gas
distributor.
In the high-pressure system there is a high expansion and
small bubbles, as shown in Table 4. It is well known in the
uidization community (Sobreiro and Monteiro, 1982; Rowe
et al., 1983; Piepers et al., 1984; Gidaspow, 1994) that under
high pressure, Geldarts Type B powders undergo considerable
expansion before bubbling.
At atmospheric pressure, the equivalent bubble diameter obtained from the above equation is 0.58 m. This shows a reasonable agreement with bubble size in Table 4.

3404

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

100

Power Spectral Magnitude

Axial Solid Velocity, m/s

90
4
3
2
1

Axial

80

Radial

70
60
50
40
30
20
10

20

25

30

35
Time, sec

40

45

50

0.8

0.025

0.4

Power Spectral Magnitude

Radial Velocity, m/s

2
3
Frequency (Hz)

0.030

0.6

0.2
0
-0.2
-0.4
-0.6
-0.8

20

25

30

35
Time, sec

40

45

50

0.015
0.010
0.005
0.000

Frequency (Hz)

Fig. 7. Typical time series of axial (a) and radial (b) hydrodynamic velocities
(v) for particles in the center region at a bed height of 4.

3.2.3. Fluctuations
Fig. 7 shows typical time series of axial and radial hydrodynamic velocities for particles at 25 atm. The hydrodynamic
velocities are obtained directly from the code. Their frequency
distributions of hydrodynamic velocities are shown in Fig. 8(a).
The main frequency for the axial direction is 0.233 Hz, and
the main frequency for the radial direction is 0.3 Hz. Fig. 8(b)
shows the power spectrum density of bed void at 25 atm. The
dominant frequency (f ) of porosity oscillations in the bubbling bed can be estimated by an analytical solution, as follows
(Gidaspow et al., 2001; Jung et al., 2005):


1  g 1/2 (3s /g + 2)s 1/2
f=
,
2 H
s0

0.020

(11)

where s0 and H0 are some initial solid volume fraction and
initial bed height.
The time averaged solid volume fraction at the center of
column is approximately 0.80. The initial bed height is 4 m. The
calculated main frequency for porosity oscillations obtained
from the above equation is 0.24 Hz. This shows a reasonably
good agreement with the main frequency in Fig. 8.

Fig. 8. Frequency and power spectral magnitude of (a) hydrodynamic velocities and (b) bed void. Main frequency: axial velocity = 0.233 Hz; radial
velocity = 0.3 Hz; bed void = 0.3 Hz.

4. Dispersion coefcients
There are two kinds of mixing in uidization: that due to individual particle oscillations and that due to cluster oscillations.
An order of magnitude estimate of the dispersion coefcient
due to individual particles oscillations can be obtained from
the laminar granular temperature (Jiradilok et al., 2006). Turbulent dispersion coefcients can be obtained as a function of
normal Reynolds stress corresponding the Lagrangian integral
time scale as described below.
4.1. Turbulent dispersion coefcients calculation
The dispersion coefcients in the radial and axial directions
are expressed as in Hinze (1959), as follows:
DT (a) = v  (a)2 TL ,

(12)

where v  (a)2 is the mean square uctuating velocity corresponding to normal Reynolds stress and TL is the Lagrangian

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

1
Axial correlation coefficient

0.8
0.6
0.4
0.2
0

-0.2
RL (y,t)
-0.4

0.1

0.2

0.3

0.4
0.4

0.6

0.5

0.2

0.6

4.2. Characteristic lengths

Time, sec

Fig. 10 shows the snapshots of solid volume fractions of the


computed clusters at 6.5, 7.5 and 8.5 s. The length and width
of clusters can be approximated from characteristic lengths estimated from the relation between the dispersion coefcients
and the oscillating velocity as

Radial corelation Coefficient

1
0.8
0.6
0.4

Dispersion coefcients (D)


= characteristic length oscillating velocity.

0.2
0
0

0.1

0.2

-0.4
RL (x,t)
-0.6

0.3

0.4
0.6

0.4

0.5
0.2

0.6
0

Time, sec
Fig. 9. Typical autocorrelation functions of solid phase (a) axial; (b) radial
for Ws = 98.8 kg/m2 s and Ug = 3.25 m/s.

integral time scale of the particle and gas motion, dened by



 
v (t)v  (t + t  ) 
TL =
RL (
a , t  ) dt  =
dt ,
(13)
0
0
v2
where v  here is Lagrangian velocity uctuations and the autocorrelation function, given by
RL (a, t  ) =

v  (t)v  (t + t  )
v2

(14)

Eulerian turbulence characteristics can be obtained from


Lagrangian turbulence characteristic (Hinze, 1959). The relationship between the Eulerian and the Lagrangian turbulence
characteristics has been given by Hay and Pasquil as
TL = TE ,

(15)

where  is the coefcient, TE is the Eulerian integral time scale


of the particle and gas motion.
In order to estimate the order of magnitude of the dispersion coefcient, the Eulerian integral time scale approximately
equals Lagrangian integral time scale (Hinze, 1959):
TL TE .

Eq. (12) is a special case of the second-order tensor dispersion matrix, Eq. (6).
Fig. 9 shows typical autocorrelation functions of solid phase
in radial and axial directions for Ws = 98.8 kg/m2 s and Ug =
3.25 m/s. The autocorrelation function decays with time from
the maximum value of one, and goes to zero. For the radial
autocorrelation, the prole dips below zero, then oscillates to a
stationary value of zero due to the wall limitation of x direction.
For the direction of ow, the autocorrelation coefcient simply
decayed exponentially, corresponding to Roy et al. (2005) in
a liquidsolid system and Godfroy et al. (1999) in a gassolid
riser.

-0.6

-0.2

3405

(16)

(17)

The oscillating velocities are obtained from the square root


of normal Reynolds stress. Fig. 11 shows the radial distribution
of characteristic lengths in the axial and the radial directions.
The length and width of clusters depend on the position corresponding to Fig. 10. The lengths and widths of clusters are
approximately 10100 and 0.54 cm, respectively.
4.3. Turbulent dispersion coefcients
The computed radial and axial particle and gas dispersion
coefcients varied as a function of positions. The values of
dispersion coefcients are reported by averaging over the cross
section of a riser or bubbling beds. The standard deviations are
based on these lateral position variations.
For ow of FCC particles in the riser, the experimental solid
ux of 98.8 kg/m2 s and the gas velocity of 3.25 m/s (Wei
et al., 1998a) was used for the computation of the gas and solid
dispersion coefcients. The computation of gas dispersion coefcients was studied based on Jiradilok et al. (2006) study. A
comparison of radial and axial solid and gas dispersion coefcients at various heights is summarized in Table 5. The solid
and gas dispersions are of same order of magnitudes, because
the Reynolds stresses per unit bulk density do not differ from
each other. Substantially both dispersion coefcients vary from
top to bottom of the riser, as expected. The simulations show
that the radial dispersion coefcients in the riser are two to
three orders of magnitude lower than the axial dispersion coefcients.
For bubbling commercial size uidized bed simulations, the
effects of pressure on dispersion coefcients were studied. A
comparison of radial and axial solid and gas dispersion coefcients at 25 atm and at an atmospheric pressure is given in
Table 6. They were calculated at 4 m above the distributor. The

3406

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

Fig. 10. Snapshots of solid volume fraction at 6.5, 7.5 and 8 s for Ws = 98.8 kg/m2 s and Ug = 3.25 m/s.

dispersion coefcients for the bubbling bed at 25 atm are much


higher than at atmospheric pressure due to the high bed expansion with smaller bubbles.
Fig. 12 shows that solid dispersion coefcients increase with
the bed diameter because the bubble diameter increases with
the bed diameter. The computed axial solid dispersion coefcients of bubbling bed at atmospheric pressure agree with the
measured data. The differences between the simulations and
the experiments are in part due to different denitions of the
dispersion coefcients.
Figs. 13 and 14 show the comparisons of computed axial
and radial gas dispersion coefcients with the literature survey
by Breault (2006). The computed dispersion coefcients are in
the range of the literature data.
The axial solid dispersion coefcients for 850 m cork particle were generated at the National Energy Technology Laboratory (NETL) and measured using the autocorrelation method.
This is the same method used in this study. The data are reported in Fig. 15. A numerical simulation of the NETL riser is
in progress.
Comparisons between computed solid dispersion coefcients
and the literature survey for both directions, axial and radial,
are shown in Figs. 15 and 16, respectively. The computations
show that the gas and the solid dispersion coefcients are close

to each other in agreement with measurements. The simulations


show that the radial dispersion coefcients in the riser are two
to three orders of magnitude lower than the axial dispersion
coefcients, but less than an order of magnitude lower for the
bubbling bed at atmospheric pressure.

5. Conclusions
We have shown how to compute radial and axial particle and
gas dispersion coefcients in the turbulent regime of a riser
with ow of FCC particles and in bubbling commercial size
uidized beds at low and high pressures.
The dispersion coefcients were computed from the turbulent
velocity oscillations of the gas and the particles obtained
by direct numerical solutions of the coupled NavierStokes
equations for gasparticle ow in the two uid model.
The computed dispersion coefcients are in reasonable agreement with the experimental measurements reported over the
last half century. The CFD computations suggest that the reported differences in the dispersion coefcients may be due
to geometrical effects of the risers and the bubbling beds,
since the geometry strongly affects the local gas and particle
velocities from which the dispersion coefcients are derived.

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

100
80
60
40

0.2

0.4

0.6

0.8

r/R

0.01
A
0.001
B

0.2

0.4

0.6

0.8

10

400 cm

1
0.1
Bed Diameter (m)

0.1

Li , 1989
Wei, 2001

0.01
0.001
0.0001

600 cm

Fig. 11. Radial distributions of characteristic lengths (a) axial; (b) radial for
Ws = 98.8 kg/m2 s and Ug = 3.25 m/s.

10

CFD-Riser-FCC particles
4 m.
Dry, 1989
6 m.
2 m.
CFD,1 atm
Kim, 1998 & 1999
(Bubbling)

r/R
200 cm

Axial,
Bubbling, 1 atm
Computation
May (1959)
Thiel and Potter (1978)
Avidan et al (1985)
Morooka et al (1972)
Du et al (2002)
Jung et al (2005)
Lewis et al (1962)
de Groot (1967)
Liu and Gidaspow (1981)
Lee and Kim (1990)
Mostoufi et al (2001)
Jung et al (2005)
This study (Computation)

Fig. 12. Effect of the bed diameter on experimental and computed solid
dispersion coefcients for bubbling and turbulent uidized beds for Geldart
A and B particles (Avidan and Yerushalmi, 1985; Du et al., 2002; de Groot,
1967; Jung et al., 2005; Lee and Kim, 1990; Lewis et al., 1962; Liu and
Gidaspow, 1981; May, 1959; Morooka et al., 1972; Mostou and Chaouki,
2001; Thiel and Potter, 1978).

Axial Gas Dispersion Coefficient (m2/s)

0.1

0.0001
0.01

20
0

Radial Charteristic Length (cm)

1
Solids Dispersion Coefficients (m2/s)

Axial Charecteristic Length (cm)

The computed dispersion coefcients and the normal stresses


allow the computation of characteristic lengths of clusters.
The lengths and widths agree with snapshots of volume fraction of solid.

3407

4
5
6
7
Gas Velocity (m/s)

Fig. 13. Effect of the gas velocity on experimental and computed axial gas
dispersion coefcients (Dry and White, 1989; Kim and Namkung, 1998,
1999; Li and Weinstein, 1989; Wei et al., 2001).

Table 5
A comparison of computed radial and axial solid and gas dispersion coefcients at various heights for FCC particles in a riser

s

Height (m)

2
4
6

0.25
0.14
0.04

Solid dispersion coefcient (m2 /s)

Gas dispersion coefcient (m2 /s)

Axial

Radial

Axial

Radial

0.347 0.120
1.221 0.289
1.331 0.757

0.002 0.001
0.001 0.001
0.009 0.005

0.614 0.323
2.032 0.927
1.134 0.821

0.004 0.004
0.002 0.002
0.004 0.003

Table 6
A comparison of computed radial and axial solid and gas dispersion coefcients at two pressures for the bubbling beds
P (atm)

1
25

s

0.41
0.20

10

Solids dispersion coefcient (m2 /s)

Gas dispersion coefcient (m2 /s)

Axial

Radial

Axial

Radial

0.069 0.042
0.791 0.373

0.012 0.005
0.030 0.009

0.06 0.027
0.891 0.464

0.019 0.011
0.040 0.018

Radial Gas Dispersion Coefficient (m2/s)

3408

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409

The dispersion coefcients for the bubbling bed at 25 atm are


much higher than at atmospheric pressure due to the high bed
expansion with smaller bubbles.
The computations show that the gas and the solid dispersion
coefcients are close to each other in agreement with measurements. The simulations show that the radial dispersion
coefcients in the riser are two to three orders of magnitude
lower than the axial dispersion coefcients, but less than an
order of magnitude lower for the bubbling bed at atmospheric
pressure.

10
1
Leckner,
(hot) 2000

0.1

Leckner,
(cold) 2000

Wei, 2001

CFD, 1atm
(Bubbling)
Rhodes,
1993

0.01
6 m. CFD-riser Werther, 1992
2 m. (FCC particles)
4 m.

0.001
Leckner, 2002

0.0001

Adanez, 1997

Notation
0

10

Gas Velocity (m/s)


Fig. 14. Effect of the gas velocity on experimental and computed radial
gas dispersion coefcients (Adanez et al., 1997; Leckner et al., 2000, 2002;
Rhodes et al., 1993; Wei et al., 2001; Werther et al., 1992).

Axial Solids Dispersion (m2/sec)

100
10
1
0.1

Duet al. (2002)


Thiel and Potter (1978)
Aviden and Yerushalmi (1985)
Weiet al. (1998)
Weiet al. (1995)
Gidaspow et al. (2004), IITRiser
NETL unit,850m Cork particles
Jiradiloket al. (2006), FCC particles
2 m.
4 m.
6 m.
This study (Bubbling Bed, 500 m)

Experiment
FCCparticles-Cluster
Computation
Computation

NETL ,
Cork particles

Bubble
Computation, 1 atm

0.01
0.001
0.01

Single particle
oscillation

0.1
1
Gas Velocity (m/sec)

10

Fig. 15. Effect of the gas velocity on experimental and computed radial solid
dispersion coefcients (Du et al., 2002; Thiel and Potter, 1978; Avidan and
Yerushalmi, 1985; Wei et al., 1995, 1998a, b; Gidaspow et al., 2004; Jiradilok
et al., 2006).

Radial Solids Dispersion (m2/sec)

0.1

0.01

Du et al. (2002)
Koenigdorff and Werther (1995)
Wei et al. (1998)
We i et al. (1995)
Jiradilok et al. (2006), FCCparticles
2 m.
4 m.
6 m.
This study (Bubbling Bed, 500 m)

Experiment

a
c
Cd
dk
DP
DT
e
g
g0
L
P
Pk
vi
vi
vi vj

x or y directions
instantaneous velocity
drag coefcient
characteristic particulate phase diameter
particle dispersion coefcients
turbulent dispersion coefcients, due to cluster oscillations
coefcient of restitution
gravity
radial distribution function at contact
mean free path
continuous phase pressure
dispersed (particulate) phase pressure
hydrodynamic velocity in i direction
mean particle velocity in i direction
Reynolds stress (i = j normal Reynolds stress; i =
j shear Reynolds stress)

Greek letters
B

k




k
k
k


interphase momentum transfer coefcient


energy dissipation due to inelastic particle collision
volume fraction of phase k
granular temperature
granular conductivity
bulk viscosity of phase k
shear viscosity of phase k
density of phase k
stress of phase k
specularity coefcient

Computation

Bubble
Computation, 1 atm

Acknowledgment
This study was supported by the US Department of Energy
Grant (DE-FG26-06NT42736).

FCC particles-cluster
Computation

References
0.001
0.1
1
Gas Velocity (m/sec)

10

Fig. 16. Effect of the gas velocity on experimental and computed radial solid
dispersion coefcients (Du et al., 2002; Koenigsdorff and Werther, 1995; Wei
et al., 1995, 1998a, b; Jiradilok et al., 2006).

Adanez, J., Gayan, P., de Diego, L.F., 1997. Radial gas mixing in fast uidized
bed. Powder Technology 94, 163.
Avidan, A., Yerushalmi, J., 1985. Solids mixing in an expanded top uid
bed. A.I.Ch.E. Journal 31, 835.
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2002. Transport Phenomena.
second ed. Wiley, New York.
Bouillard, J.X., Lyczkowski, R.W., Gidaspow, D., 1989. Porosity distributions
in a uidized bed with an immersed obstacle. A.I.Ch.E. Journal 35, 908.

V. Jiradilok et al. / Chemical Engineering Science 62 (2007) 3397 3409


Bowen, R.M., 1976. Theory of mixtures. In: Continuum Physics, vol. III.
Academic Press, New York.
Breault, R.W., 2006. A review of gassolid dispersion and mass transfer
coefcient correlations in circulating uidized beds. Powder Technology
163 (12), 917.
Chapman, S., Cowling, T.G., 1970. The Mathematical Theory of Non-Uniform
Gases. Cambridge University Press, Cambridge.
Darton, R.C., LaNauze, R.D., Davidson, J.F., Harrison, D., 1977. Bubble
growth due to coalescence in uidised beds. Transactions of the Institution
of Chemical Engineers 55, 274.
Dry, R.J., White, C.C., 1989. Gas residence-time characteristics in a highvelocity circulating uidized bed of FCC catalyst. Powder Technology 58,
1723.
Du, B., Fan, L.S., Wei, F., Warsito, W., 2002. Gas and solids mixing in a
turbulent uidized bed. A.I.Ch.E. Journal 48, 1896.
Gidaspow, D., 1986. Hydrodynamics of uidization and heat transfer:
supercomputer modeling. Applied Mechanics Reviews 39, 1.
Gidaspow, D., 1994. Multiphase Flow and Fluidization: Continuum and
Kinetic Theory Descriptions. Academic Press, New York.
Gidaspow, D., Huilin, L., Mosto, R., 2001. Large scale oscillations or gravity
waves in risers and bubbling beds. In: Kwauk, M., Li, J., Yang, W.C.
(Eds.), Fluidization X. United Engineering Foundation, New York, p. 317.
Gidaspow, D., Jonghwun, J., Singh, R.K., 2004. Hydrodynamics of uidization
using kinetic theory: an emerging paradigm 2002 Flour-Daniel lecture.
Powder Technology 148, 123141.
Godfroy, L., Larachi, F., Chaouki, J., 1999. Position and velocity of a large
particle in a gas/solid riser using the radioactive particle tracking technique.
Canadian Journal of Chemical Engineering 77, 253261.
de Groot, J.H., 1967. Scaling up of uidized bed reactors. In: Drinkenburg,
A.A.H. (Ed.), Proceedings of the International Symposium on Fluidization,
vol. 348. Netherlands University Press, Eindhoven, The Netherlands.
Hinze, H.O., 1959. Turbulence. McGraw-Hill, New York.
Jackson, R., 1985. Hydrodynamic stability of uidparticle systems. In:
Fluidization. Academic Press, New York, p. 47.
Jiradilok, V., Gidaspow, D., Damronglerd, S., Koves, W.J., Mosto, R., 2006.
Kinetic theory based CFD simulation turbulent uidization of FCC particles
in a riser. Chemical Engineering Science 61, 52445259.
Johnson, P.C., Jackson, R., 1987. Frictionalcollisional constitutive relations
for granular materials, with application to plane shearing. Journal of Fluid
Mechanics 176, 6793.
Jung, J., Gidaspow, D., Gamwo, I.K., 2005. Measurement of two kinds of
granular temperatures, stresses, and dispersion in bubbling beds. Industrial
and Engineering Chemistry Research 44, 13291341.
Kim, S.D., Namkung, W., 1998. Gas backmixing in a circulating uidized
bed. Powder Technology 99, 7078.
Kim, S.D., Namkung, W., 1999. Gas backmixing in the dense region of a
circulating uidized bed. Korean Journal of Chemical Engineering 16 (4),
456461.
Koenigsdorff, R., Werther, J., 1995. Gas and solid mixing in and ow structure
modeling of the upper dilute zone of a circulating uidized bed. Powder
Technology 82, 317329.
Leckner, B., Sterneus, J., Johnsson, F., 2000. Gas mixing in circulating
uidized bed risers. Chemical Engineering Science 55, 129.
Leckner, B., Sterneus, J., Johnsson, F., 2002. Characteristics of gas mixing
in circulating uidized bed. Powder Technology 126, 28.

3409

Lee, G.S., Kim, S.D., 1990. Axial mixing of solids in turbulent uidized
beds. Chemical Engineering Journal 44, 1.
Lewis, W.K., Gilliland, E.R., Girouard, H., 1962. Heat transfer and solids
mixing in a bed of uidized solids. Chemical Engineering Progress
Symposium Series 58, 87.
Li, J., Weinstein, H., 1989. An experimental comparison of gas backmixing in
uidized beds across the regime spectrum. Chemical Engineering Science
44 (8), 16971705.
Liu, Y., Gidaspow, D., 1981. Solids mixing in uidized bedsa hydrodynamic
approach. Chemical Engineering Science 36, 539.
May, W.G., 1959. Fluidized-bed reactor studies. Chemical Engineering
Progress 55, 49.
Morooka, S., Kato, Y., Miyauchi, T., 1972. Holdup of gas bubbles and
longitudinal dispersion coefcient of solid particles in uidized contactors
for gassolid systems. Journal of Chemical Engineering of Japan 5, 161.
Mostou, N., Chaouki, J., 2001. Local solid mixing in gassolid uidized
beds. Powder Technology 114, 23.
Piepers, H.W., Cottaar, E.J.E., Verkooijen, A.H.M., Rietema, K., 1984. Effects
of pressure and type of gas on particleparticle interaction and the
consequences for gassolid uidization behaviour. Powder Technology 37
(1), 5570.
Rhodes, M.J., Amos, G., Mineo, H., 1993. Gas mixing in gassolids risers.
Chemical Engineering Science 48 (5), 943.
Rowe, P.N., Foscolo, P.U., Hoffmann, A.C., Yates, J.G., 1983. X-ray
observation of gas uidized beds under pressure. In: Kunii, D., Joei, R.
(Eds.), Proceedings of the 4th International Conference on Fluidization,
Japan, 1983. A.I.Ch.E., New York.
Roy, S., Kemoun, A., Al-Dahhan, M.H., Dudukovic, M.P., 2005. Experimental
investigation of hydrodynamics in a liquidsolid riser. A.I.Ch.E. Journal
51, 802835.
Sobreiro, L.E.L., Monteiro, J.L.F., 1982. The effect of pressure on uidized
bed behaviour. Powder Technology 33 (1), 95100.
Soo, S.L., 1967. Fluid Dynamics of Multiphase Systems. Blaisdell Publishing.
Tartan, M., Gidaspow, D., 2004. Measurement of granular temperature and
stresses in risers. A.I.Ch.E. Journal 50, 17601775.
Thiel, W.J., Potter, O.E., 1978. The mixing of solids in slugging gas uidized
beds. A.I.Ch.E. Journal 24, 561.
Wei, F., Jin, Y., Yu, Z., Chen, W., Mori, S., 1995. Lateral and axial mixing
of the dispersed particles in CFB. Journal of Chemical Engineering of
Japan 28, 506.
Wei, F., Lin, H., Cheng, Y., Wang, Z., Jin, Y., 1998a. Proles of particle
velocity and solids fraction in a high-density riser. Powder Technology
100, 183189.
Wei, F., Cheng, Y., Jin, Y., Yu, Z., 1998b. Axial and lateral solids dispersion
in a binary-solid riser. Canadian Journal of Chemical Engineering
76 (1), 19.
Wei, F., Yang, Y., Jia, X., Yong, J., 2001. Hydrodynamics and lateral gas
dispersion in a high density circulating uidized bed reactor with bluff
internals. Chinese Journal of Chemical Engineering 9 (3), 291296.
Werther, J., Hartge, E.-U., Kruse, M., 1992. Radial gas mixing in the upper
dilute core of a circulating uidized bed. Powder Technology 70, 293.
Yang, N., Wang, W., Ge, W., Wang, L., Li, J., 2004. Simulation of
heterogeneous structure in a circulating uidized-bed riser by combining
the two-uid model with EMMS approach. Industrial and Engineering
Chemistry Research 43, 55485561.

Вам также может понравиться