Você está na página 1de 10

4070

Ind. Eng. Chem. Res. 2010, 49, 40704079

Numerical Simulation of the Gas-Solid Flow in Fluidized-Bed Polymerization


Reactors
De-Pan Shi, Zheng-Hong Luo,* and An-Yi Guo
Department of Chemical and Biochemical Engineering, College of Chemistry and Chemical Engineering,
Xiamen UniVersity, Xiamen 361005, China

A three-dimensional computational fluid dynamics (CFD) model, using an Eulerian-Eulerian two-fluid model
which incorporates the kinetic theory of granular flow, was developed to describe the gas-solid two-phase
flow in fluidized-bed polymerization reactors. Corresponding simulations were carried out in a commercial
CFD code Fluent. The entire flow field in the reactors was calculated by the model. The predicted pressure
drop data were in agreement with the classical calculated data. In addition, the model was used to describe
the solid holdup distributions, the bubble behaviors, and the solid velocity vectors in the free and agitated
fluidized-bed polymerization reactors, respectively. The effects of the addition of an agitator on the gas-solid
flow behaviors were preliminarily investigated via the model. The simulation results showed that the addition
of an agitator can strengthen the fluidization efficiency and reduce the operation stability of the bed. However,
the simulation results also showed that the total fluidization quality of the free fluidized bed was higher than
that of the agitated fluidized bed at a superficial gas velocity of 0.5 m s-1.
1. Introduction
Polyolefins can be produced in various types of reactors, such
as autoclave, continuous stirred tank, tubular loop, or fluidized
bed (FBR). The last one is certainly the most important because
of its simple construction and excellent heat- and mass-transfer
characteristics.1 For instance, various technologies, including
Hypol, Innovene, Unipol, Spheripol, etc., are designed to
produce polypropylene. Among them, there are different reactor
arrangements in essence.2-4 FBR is one of their central reactors,
which is generally used to produce high-impact polypropylene.4
In the fluidized-bed olefin polymerization reactor, small catalyst
and/or polymer particles react with monomers to form polymer
particles in the gas phase, and the polymer particles are produced
as a solid suspension in the gas stream.2-5 Accordingly, the
reacting system is considered to be a mixture of gas and solid
phases, namely, a gas-solid two-phase system. For efficient
operation and to accomplish the desired results, it is imperative
that a good fluidization quality is achieved to ensure good
gas-solid contact, uniformity of temperature, and minimum gas
bypassing. For these reasons, computational fluid dynamics
(CFD) is becoming more and more an engineering tool to predict
flows in various types of apparatuses on the industrial scale.6-8
Furthermore, CFD is an emerging technique and holds great
potential in providing detailed information on complex fluid
dynamics.9-11
In general, two different categories of CFD models are used,
namely, the Lagrangian and Eulerian models.6-8 The Lagrangian
model solves equations of motion for each particle, taking into
account particle-particle collisions and the forces acting on the
particle, whereas the Eulerian model considers full interpenetrating continua subject to continuity and momentum equations.
Considerable attention has been devoted in recent years to the
application of CFD to gas-solid FBRs.9-17 A comprehensive
review has been published on these CFD models and experiments applied to FBRs.18 Most authors have used Eulerian
models, including continuity and momentum equations for two
interpenetrating continua, one representing the gas and the other
* To whom correspondence should be addressed. Tel.: +86-5922187190. Fax: +86-592-2187231. E-mail: luozh@xmu.edu.cn.

the solid. To achieve closure, a granular temperature model has


usually been introduced. When a turbulent flow of the gas phase
is assumed, a k- model is also incorporated. In addition,
different authors have adopted different assumptions with respect
to such aspects as boundary conditions, interphase momentum
transfer (drag) relationships, and parameters in the Eulerian
model. As a whole, these models were able to provide good
qualitative and reasonable quantitative agreement with limited
experimental findings with the help of fitting parameters.19-21
However, previous studies mainly concentrated on prediction
of their gas-solid holdup distributions and the effects of the
gas velocity on them via a CFD model along with simplification
of the flow field as a two-dimensional (2D) field.19-24 Most
authors applied CFD to the free FBR or the agitated FBR
without comparing the gas-solid flow in the free FBR to that
in the agitated FBR.19-26 In addition, less attention has been
paid to the bubble behaviors and characteristics in FBRs.27 In
practice, many features of the gas-solid FBRs, like excellent
solid mixing and heat- and mass-transfer properties, can be
correlated to the presence of bubbles, which dominate their
behaviors.28 A deeper knowledge of the FBR hydrodynamics
and how such hydrodynamics are affected by the addition of
the agitator would provide the basis for the development of a
fully predictive model.29
Recently, Lu et al.28 used a 2D Eulerian-Eulerian model
extended with the kinetic theory of granular flow (KTGF) to
simulate the bubble behaviors in a free gas-solid FBR. The
simulated values were compared to the values from the Darton
bubble-size equation and the Davidson model for isolated
bubbles.28 Busciglio et al.27 also described the bubble behaviors
in a free gas-solid FBR via a 2D Eulerian CFD model. Their
simulated data were validated with experimental data.27 Witt
etal.30 usedathree-dimensional(3D)multiphaseEulerian-Eulerian
technique to predict the transient bubble formation in a free
FBR. Unfortunately, these models were used to describe the
bubble behaviors in the free FBRs. To the best of our
knowledge, thus far, there were not any open reports regarding
the application of CFD to the fluidized-bed olefin polymerization
reactor investigating their bubble behaviors.

10.1021/ie901424g 2010 American Chemical Society


Published on Web 04/05/2010

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

4071

Figure 1. FBR configurations: (a) free FBR; (b) agitated FBR; (c) stirrer.

In this work, we develop a 3D CFD model based on the


Eulerian-Eulerian approach to describe the gas-solid twophase flow in the fluidized-bed propylene polymerization reactor.
The entire flow field in the FBR is calculated by the model.
Furthermore, the model is used to describe the bubble behaviors
in the free and agitated fluidized-bed polymerization reactors,
respectively. The effect of the agitator on the gas-solid flow
behaviors is preliminarily investigated via the model.

(R F ) + (RsFsb
V s) ) 0
t s s

(2)

The momentum balance equations for the gas and solid phases
can be written as

(R F b
V ) + (RgFgb
V gb
V g) ) -Rgp + g + Kgl(V
bs t g g g
b
V g) + RgFgg (3)

2. 3D Model for FBRs


Spheripol technology is one of the most widespread commercial methods to be used to produce polypropylene. Commonly, its key part constitutes of two liquid-phase loop reactors
and a gas-phase FBR. In this work, a pilot-plant-scale polypropylene-agitated FBR of Spheripol technology in a Chinese
chemical plant shown in Figure 1 was selected as our object.
The agitated FBR selected consists of a vertical 0.5-m-i.d.
cylinder of height 1.5 m, and there is a stirrer in the FBR. In
order to investigate the effect of the stirrer on the flow behaviors,
a free FBR with the same size and shape but no stirrer, compared
to the agitated FBR, is also selected in this work. More detailed
information regarding the FBR configurations is shown in Figure
1. Furthermore, the flow systems in FBRs are both supposed to
be mixtures of gas and solid phases.2-5,31-34
In the present study, to simulate the 3D reactors, a 3D physical
model of the reactor system must be available. Hence, the 3D
physical models and their meshes were both constructed in
Gambit 2.3.16 (Ansys Inc., Columbus, OH) first.
3. CFD Model
On the basis of KTGF, a 3D Eulerian-Eulerian two-fluid
model is used to describe the gas-solid two-phase flow in the
above FBRs.
3.1. Eulerian-Eulerian Two-Fluid Equations. This section
describes the modeling equations employed in the present
Eulerian-Eulerian two-fluid CFD model.
The continuity equations for phase n (n ) g for the gas phase
and s for the solid phases) may be written as

(R F ) + (RgFgb
V g) ) 0
t g g

(1)

g ) Rgg(V
bg + V
bgT)

(4)

(R F b
V ) + (RsFsb
V sb
V s) ) -Rsp - ps + s +
t s s s
Kls(V
bg - b
V s) + RsFsg (5)
2
s ) Rss(V
bs + V
bsT) + Rs s - s V
bsI
3

(6)

3.2. KTGF. The two-fluid model requires constitutive equations to describe the rheology of the solid phase, i.e., the
viscosity and pressure gradient of the solid phase. When the
particle motion is dominated by collision interaction, concepts
from fluid kinetic theory can be introduced to describe the
effective stresses in the solid phase resulting from particle
streaming (kinetic contribution and direct collisions) collision
contribution.9,7,35-37 Constitutive relations for the solid-phase
stress based on the kinetic theory concepts have been derived
by Lun et al.36 Moreover, their equations have been accepted
widely and are also applied in this work.
ps ) RsFss[1 + 2g0Rs(1 + es)]

4
s ) Rs2Fsdsg0(1 + es)
3

(7)

(8)

where
g0 )

1
1 - (Rs /Rs,max)/1/3

(9)

4072

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

1
s ) ss
3

(10)

In addition, a transport equation for the granular temperature is


also needed and is suggested by Ding and Gidaspow:38
3
(F R ) + (FsRsF
W ss) ) (-psI + s):F
Ws +
2 t s s s
(kss) - s + gs (11)

Table 1. Physical Properties of Gas and Solid Phases


ds/m
1 10

-3

g/pa s

910.0

21.56

1.081 10-5

CD )

15FsdsRss
12
16
(41 1 + 2(4 - 3)Rsg0 +
4(41 - 33)
5
15
33)Rsg0

Fg/kg m-3

where

where the diffusion coefficient for granular energy, ks, is given


by Syamlal et al.:39
ks )

Fs/kg m-3

3
24
R Re
1+
RgRes
20 g s

[ (

Res )

at Rg e 0.8, Ksg ) 150

12(1 - es2)g0
ds

FsRs2s1.5

(14)

gs ) -3Kgss

(15)

In this study, the granular energy was assumed to be at steady


state and dissipated locally, and the convection and diffusion
were also neglected.7,37,39 Accordingly, eq 11, which is a
complete granular temperature transport equation, can be
simplified to an algebraic equation. The simplified equation is
as follows:
0 ) (-psI + s):F
W s - s

(16)

There are many similar models for the solid-phase dynamic


viscosity. The selected model in this work is as follows:36,40,41
s ) s,col + s,kin + s,fr

Vs - b
V g|
Fgds | b
g

Rs(1 - Rg)g

(13)

The collision dissipation of energy, s, is modeled using the


correlation by Lun et al.:36
s )

(22)

(23)

(12)

where
1
) (1 + es)
2

0.687

) ]

(17)

Rgds

Vs - b
V g|
7 RsFg | b
4
ds
(24)

3.4. CFD Modeling Strategy. As discussed earlier, the CFD


with the Eulerian-Eulerian approach has been used to study
the gas-solid interactions in this work. The RNG k- model
is used to take into account the turbulence, whereas KTGF has
been used to close the momentum balance equation for the solid
phase. The above equations are solved by the commercial CFD
code Fluent 6.3.26 (Ansys Inc., Columbus, OH) in double
precision mode. The phase-coupled SIMPLE algorithm is used
to couple the pressure and velocity, and the multiple reference
frame (MRF) model is used to simulate the agitated FBR. In
addition, as described in section 2, a commercial grid-generation
tool, Gambit 2.3.16 (Ansys Inc., Columbus, OH), is used to
generate the 3D geometries and their grids. Simple grid
sensitivity was carried out, the least cells needed to conserve
the mass of the solid phase in the dynamics modeling were
studied, and in total 107 000 and 161 000 cells were needed
for the free and agitated FBRs, respectively. Furthermore, the
simulations were executed in a Pentium 4 CPU running at 2.83
GHz with 4GB of RAM.
4. Simulation Conditions

where
s,col
s,kin )

4
) RsFsdsg0(1 + es)
5

10dsFss
4
1 + (1 + es)Rsg0
96Rs(1 + es)g0
5

s,fr )

(18)
2

ps sin
2I2D

(19)

(20)

3.3. Drag Force Model. In this work, the transfer of forces


between the gas and solid phases is described according to the
empirical drag law based on work by Gidaspow et al.40
Gidaspows model combines Wen and Yus model42 via the
Ergun equation.43 Corresponding equations are shown as
follows:
Vs - b
V g | -2.65
3 RsRgFg | b
at Rg > 0.8, Ksg ) CD
Rg
4
ds

4.1. Physical Properties of the Gas and Solid Phases.


During polymerization in FBRs, the growth rate of the polymer
particles is very slow, and their growth in diameter is mainly
determined by the residence time of the polymer particles in
FBRs. In addition, the simulated flow time in this study is very
short (about 10 s) because of the intense computational time.
In view of this, the integral of the above gas-solid system with
the propylene polymerization in FBRs is simulated. The
properties of the typical gas and solid phases at certain
polymerization times are listed in Table 1.
4.2. Boundary Conditions and Model Parameters. The
minimum fluidization velocity (Umf) and the particle terminal
velocity (UT) can be estimated by using Wen and Yus
equations:42

Umf )

(21)

g
dsFg

{[

33.72 + 0.0434

ds3Fg(Fs - Fg)g
g2

0.5

- 33.7

(25)

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

4073

Table 2. Boundary Conditions and Model Parameters


description

value

turbulence model
granular viscosity
granular bulk viscosity
frictional viscosity
angle of internal friction
granular temperature
drag law
coefficient of restitution for particle-particle collisions
inlet boundary condition
outlet boundary condition
wall boundary condition
initial bed height
initial volume fraction of the solid phase
operating pressure
inlet gas velocity
rotating speed
oulet pressure
maximum iterations
convergence criteria
time step

k- (RNG, dispersed)
Gidaspow et al.40
Lun et al.36
schaeffer
30
algebraic
Gidaspow et al.40
0.9
velocity inlet
pressure outlet
no slip for air, specularity coefficient 0 for the solid phase16,17
0.39 m
0.63
1.40 106 Pa
0.5 m s-1
10 rpm (available in the stirred model)
1.013 25 105 Pa
30
1 10-3
1 10-3 s

at Re < 0.4, UT )

at 0.4 < Re < 500, UT )

(Fs - Fg)ds2g
18g

at 500 < Re < 200000, UT )

4(Fs - Fg)2g
225Fgg

(26)

0.5

(27)

ds

3.1(Fs - Fg)gds
Fg

0.5

(28)

where
Re )

dsFgUT
g

(29)

According to the above equations and corresponding data shown


in Table 1, the values obtained for Umf and UT are 0.1126 and
1.118 m s-1, respectively. In addition, the superficial gas
velocity must be operated between the values of Umf and UT.
According to the process description, Ug is always set to 3-5
times Umf. Thus, the value of 0.5 m s-1 is used in this study.
As described in section 3.4, the CFD model was solved in
Fluent. The detailed settings in the software are list in Table 2.
5. Results and Discussion
The hydrodynamic characteristics of the entire flow field in
the free and agitated FBRs, such as the pressure drop, solid
holdup distribution, bubble behavior, and solid velocity vector
profile, are investigated via the above model, respectively. The
physical properties and model parameters are listed in Tables 1
and 2. In addition, as described in section 1, many researchers6-43
studied the gas-solid two-phase flow; a set of reference values
of these parameters can be selected. In the present study, two
important parameters, including the restitution coefficient (es)
and specularity coefficient (), were investigated. We find that
a little change of would lead to a significant change of the
pressure drop in the loop reactor (here no results are given
because of the limited space). However, the pressure drop is
not sensitive to the changes of the restitution coefficient.
Therefore, the default value of 0.9 for the restitution coefficient
in Fluent was chosen. Furthermore, our foregone sensitivity
analysis of the specularity coefficient shows that, with an
increase of , the difference between the predicted pressure drop
at the corresponding flow velocity and flow time and that

obtained via the classical Newitt model44,45 increases. A good


prediction of the pressure drop when equals 0 can be obtained.
Meanwhile, the value of 0 for is also that at the free slip
boundary condition for the solid phase wall boundary condition
and can be found in typical literature.16,17 Therefore, the value
of 0 for the specularity coefficient was chosen and is shown in
Table 2. Unless otherwise noted, the parameters used for the
following simulation are those in Table 2.
5.1. Pressure Drop. It is well-known that the bed pressure
drop in FBRs is an important parameter in the proper scaling
up and design of these reactors. The bed pressure drop can
always be described by the buoyant weight of the suspension:44,45
Ps ) (Fs - Fg)(1 - )gL

(30)

However, in this study, because the gas-phase density is up to


21.56 kg m-3, it is necessary to consider the effect of the gasphase weight on the pressure drop:
Pg ) FggL

(31)

Corresponding pressure drops calculated by the above equations


are depicted in Figures 2 and 3, which show the bed pressure
drop profiles as a function of the flow time in two types of
FBRs, respectively. As a whole, Figures 2 and 3 prove that the
simulated bed pressure drop data are in agreement with the
classical calculated data. Also, the slight difference shown in

Figure 2. Pressure drop versus flow time in the free FBR at an initial flow
velocity of 0.5 m s-1.

4074

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

Figure 3. Pressure drop versus flow time in the agitated FBR at an initial
flow velocity of 0.5 m s-1.

Figures 2 and 3 may result from neglect of the pressure drop


caused by friction and particle collision in the classical
calculation. Moreover, according to Figure 2, it is worth noting
that four typical regions in the free FBR can be found: the startup stage ( ) 0 s), slow drop stage (0 s < < 4.0 s), vibration
stage (4.0 s e < 6.4 s), and stable fluidization stage ( g
6.4 s). The maximum bed pressure drop at the start-up point/
stage is higher than that at the other three stages because the
interparticle locking is overcome. Since then, the bed pressure
drop decreases slowly because of formation of the gas-phase
flow field and the looseness of the solid phase following the
flow proceeding in the period of 0-4.0 s. In the period of
4.0-6.4 s, the first air bubble comes into being and develops
with time; correspondingly, the bed pressure drop fluctuates
greatly in this period. After the vibration stage, the bed pressure
drop fluctuates with time around a mean value shown in Figure
2, namely, the stable fluidization stage. In this stage, the
fluidization process in the whole free FBR is accomplished.
Therefore, the bed pressure drop is also close to a certain steadystate value. Compared with Figure 2, Figure 3 shows a similar
curve, which indicates that there are similar change trends of
the bed pressure drop in the two types of FBRs. Nevertheless,
one still notices that the whole fluctuation range of the bed
pressure drop in the free FBR is lower than that in the agitated
FBR from Figures 2 and 3. Furthermore, Figures 2 and 3
indicate that the fluctuation in the free FBR is more frequent
than that in the agitated FBR.
5.2. Solid Holdup Distribution. Solid holdup is one of the
most important parameters in FBRs. If the solid holdup is too
high somewhere, the corresponding polymerization rate may
be too high there because of the high concentration of catalyst,
which leads to a highly exothermic reaction. In addition, the
highly exothermic reaction may lead to the appearance of hot
spots if the heat of polymerization cannot be efficiently removed.
In this section, the solid holdup distributions in the free and
agitated FBRs are investigated using the above model, respectively. In addition, we also point out that the lack of consideration of other distributions (i.e., temperature distribution) and
the effects of some other operating parameters, such as
temperature, pressure, etc., on the flow hydrodynamics in FBRs
are limitations due to the limited space in this paper.
Figures 4 and 5 show the profiles of the average solid holdup
as a function of the height of the horizontal plane from the
bottom of FBRs in the free and agitated FBRs at the stable
fluidization stage, respectively. Corresponding solid holdup
distribution data at 10 s for the two reactors are shown in Figures
6 and 7.

Figure 4. Average solid volume holdup versus the height of the horizontal
plane from the bottom of the free FBR at an initial flow velocity of 0.5
m s-1 and a flow time of 10 s.

Figure 5. Average solid volume holdup versus the height of the horizontal
plane from the bottom of the agitated FBR at an initial flow velocity of 0.5
m s-1 and a flow time of 10 s.

Figure 4 shows that the average solid holdup changes very


little with a change of the height of the horizontal plane from
the bottom of the free FBR in the range of 0-0.6 m. This means
that the granule distribution in the main body of the free FBR
is homogeneous. Namely, the fluidization quality is perfect.
Hereafter, the average solid holdup descends to 0 with an
increase of the height of the horizontal plane from the bottom
of the free FBR. Accordingly, one knows that the actual height
of the free FBR is about 0.78 m according to Figure 4.
Furthermore, Figure 6 proves that the fluidization quality in the
free FBR is perfect. In practice, Figure 6 shows that the amount
of bubbles is few and the emulsion phase is the main body in
the free FBR. Although there are still granule agglomerations,
not any granules adhere to the wall in the free FBR.
Compared with Figure 4, Figure 5 shows a similar curve,
which proves that there are similar change trends of the average
solid holdup in the free and agitated FBRs. Nevertheless, one
also obtains that the actual height of the agitated FBR is about
in the range of 0.7-0.85 m according to Figure 5. Figure 7
shows that there are many bubbles generated in the agitated
FBR after fluidization for a long time (10 s). It is helpful to
have heat transfer between the gas and solid phases and leads
to an increase of the granule entrainment at the breakup of
bubble. However, the large volume and excessive amount of
bubbles generated are harmful to the stability of FBR. From
the solid phase shown in Figure 7, we can observe more obvious
granule agglomeration phenomena than that in Figure 6. The

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

4075

Figure 6. Solid volume holdup distribution in the free FBR at an initial flow velocity of 0.5 m s-1 and a flow time of 10 s.

Figure 7. Solid volume holdup distribution in the free FBR at an initial flow velocity of 0.5 m s-1 and a flow time of 10 s.

Figure 8. Solid volume holdup distribution of the vertical plane across the vertical axis in the free FBR at an initial flow velocity of 0.5 m s-1 versus flow
time.

corresponding solid holdup distribution in the agitated FBR is


inhomogeneous.
5.3. Bubble Behaviors. As described in section 1, many
characteristic features of gas-solid FBRs are related to the
presence of bubbles and dominated by their behaviors. Here,
the bubble behaviors in two types of FBRs are simulated. On
the basis of Figures 2 and 3, one knows that both of the flow
fields in the free and agitated FBRs are in the stable fluidization
period when g 6.4 s. Here, we provide the simulated results
in 0-10 s of flow time, and the results are presented in Figures

8 and 9. Furthermore, Figures 8 and 9 show the bubble


formation processes with representation of the solid volume
fraction distribution of a vertical plane across the vertical axis
as a function of the flow time in the free and agitated FBRs,
respectively. Typical 3D diagrams of bubble evolution in the
free and agitated FBRs are shown in Figures 10 and 11,
respectively.
According to Figure 8, one can obtain the processes of bubble
formation, development, and breakup following the flow
proceeding in the free FBR. At 3 s in Figure 8, only the granules

4076

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

Figure 9. Solid volume holdup distribution of the vertical plane across the vertical axis in the agitated FBR at an initial flow velocity of 0.5 m s-1 versus
flow time.

Figure 10. Visual representations of bubble formation in the free FBR at an initial flow velocity of 0.5 m s-1.

Figure 11. Visual representations of bubble formation in the agitated FBR at an initial flow velocity of 0.5 m s-1.

in the bottom of the free FBR become flexible to form an


emulsion phase due to the initial quiescence state of the granule
phase and the interparticle locking; accordingly, the bed height
rises a little. The bubbles begin to form with further emulsification of the granules at 4 s. Simultaneously, as shown at 4.0,
4.2, and 4.4 s in Figure 8, the bubbles formed appear to deform
because of the interaction between granules and also develop
upward before arriving at the free space in the free FBR.
Accordingly, the film between the gas and emulsion phases
becomes thinner with development of the bubbles and breaks
up ultimately at about 4.6 s in Figure 8. This leads to the rise

of these granules in the fractured bubbles, and they also drop


back to the wall because of gravitation. In addition, at 4.6 s
when the film and first bubble break up, it represents that
FBR is basically emulsified and the inner resistance decreases.
Correspondingly, the formations of new bubbles are simple,
which leads to an increase of the bubble velocity. Therefore,
the total fluidized velocity in FBR is very fast. It leads to the
formation of the stable flow field in the FBR after a short time.
According to Figure 2, we think that the stable field will form
at 6.4 s. Namely, since 4.6 s, the solid volume fraction
distribution data of the vertical plane across the vertical axis

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

4077

Figure 12. Solid velocity vector profiles in the free FBR at an initial flow velocity of 0.5 m s-1 and a flow time of 10 s.

are similar to each other and their profiles are shown at 4.6,
4.8, 10 s in Figure 8. Figure 10 shows the typical 3D diagrams
of bubble evolution from formation, deformation, to breakup
and gives 3D visualization results. According to Figure 10, one
knows that the bubble-like cirque with narrow top and wide
bottom can be obtained at 4.0 s and its shape continues to change
because of granular actions following the flow proceeding.
Simultaneously, it is split into many small bubbles along the
axial aspect. For instance, both the volume and height of the
bubbles at 4.2 s are larger than those at 4.0 s. One also observes
many small bubbles along the axial aspect of the bubble at 4.2 s
in Figure 10. Furthermore, according to Figure 10, one can find
that the bubble continues to rise along its axial aspect and it
has been split into eight small bubbles with similar shape and
different volume at 4.4 s. In addition, the obtained small bubbles
can also rise.
Compared to the bubble behaviors in the free FBR shown in
Figures 8 and 10, the bubble behaviors shown in Figures 9 and
11 are similar. However, because of the addition of a stirrer,
the first bubble obtained in the agitated FBR is unstable and
there are many small bubbles to form, along with the formation
of the first bubble. In practice, the stirrer breaks up the early
bubbles, which leads to an increase of the charge capacity of
the granules in bubbles. Furthermore, some air whorls come
into being by the edge of the stirrer and can also leave the stirrer
to form small bubbles. Some of the small bubbles can be
incorporated to form big bubbles. Therefore, there are still many
small bubbles in the agitated FBR at 10 s in Figure 9. Although
the small bubbles can strengthen the fluidization efficiency, they
reduce the operation stability of the bed because of the addition
of the agitator.
5.4. Solid Velocity Vector. The rising motions of the rotating
torus in the free and agitated FBRs are also simulated via the
above model. The simulated velocity vectors for the solid phase
in a vertical plane across the vertical axis in the two FBRs are
shown in Figures 12 and 13, respectively.
Figure 12 shows that there is an obvious circular upflow in
the vertical plane due to the bubble motion. There are some
small dimensional circular regions (small circulations) in the
bottom of the free FBR. For the dimension of the whole FBR,
the solid phase is lifted up from its middle position and comes
back along with the breakup of the bubbles due to gravitation.

The fallen granules can flow down along the wall. Therefore,
the above results can also lead to the formation of big circular
upflows, namely, big circulations. The combined action of the
small and big circulations leads to a good mixing result in the
free FBR. However, a vortex appears between the small and
big circulations, namely, in the middle of the free FBR. It is
harmful to the matter and heat transfers between the quiescent
regions and the bottom of the FBR.
Figure 13 shows that there is no vortex to be formed in the
middle of the agitated FBR because of the addition of the stirrer.
One knows that the motion of the granules in the agitated FBR
is mainly influenced by two actions. The granules can be lifted
up because of the action of the gas phase along the vertical
axis and can also be rotated in the agitated FBR because of the
action of the stirrer. Besides, other actions including granule
gravitation can also influence their motion. Accordingly, as
shown in Figure 13, most granules in the agitated FBR lift up
along the right side of the vertical plane across the vertical axis
and collide with the fallen granules to form a vortex. Some
granules flow back to the bottom of the agitated FBR. In
addition, some granules inside the fractured bubbles also fall
back to the bottom of the agitated FBR.
6. Conclusions
In this study, a 3D CFD model was developed to describe
the gas-solid two-phase flow in fluidized-bed polymerization
reactors.ThemodelincorporatedKTGFwiththeEulerian-Eulerian
approach. The pressure drop data calculated according to the
classical equation were employed to verify the model. The
predicted pressure drop data were found to agree well with the
classical calculated data. Furthermore, the hydrodynamic characteristics of the entire flow field in the free and agitated FBRs,
such as the pressure drop, solid holdup distribution, bubble
behavior, and solid velocity vector profile, were investigated
via the above model, respectively. Particular attention was paid
to the effect of the addition of the agitator on the gas-solid
flow behavior.
The simulated results show that both of the flow fields in the
free and agitated FBRs as a function of the flow time can be
divided into four periods. Namely, the start-up period ( ) 0 s),
slow drop period (0 s < < 4.0 s), vibration period (4.0 s e

4078

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

Figure 13. Solid velocity vector profiles in the agitated FBR at an initial flow velocity of 0.5 m s-1 and a flow time of 10 s.

< 6.4 s), and steady-state fluidization period ( g 6.4 s).


Furthermore, the typical profiles of the average solid hold-up
and solid hold-up distribution in the free and agitated FBRs at
the steady-state fluidization period are also obtained. In addition,
the simulated results show that the addition of the agitator can
strengthen the fluidization efficiency and yet reduce the operation stability of the bed. The simulation results also showed
that the fluidization quality of the free fluidized bed is higher
than that of the agitated fluidized bed at a superficial gas velocity
of 0.5 m s-1.
Because we know that the main function of the agitator in
the agitated FBR is to prevent the granule from adhering to the
wall of FBR, the modeling results show that granules will not
adhere to the wall in the free FBR, so it is not necessary to add
the agitator to FBR; that is why the newly developed fluidizedbed polymerization reactors are always free FBR. Further studies
on the 3D CFD model for the gas-solid two-phase flow in FBR
are in progress in our group.
Acknowledgment
The authors thank the National Natural Science Foundation
of China (Grant 20406016) and China National Petroleum Corp.
for supporting this work. We also thank Dr. Z. Yao (Department
of Chemical Engineering and Biochemical Engineering, Zhejiang University) for his valuable discussion in this work. The
authors also thank the anonymous reviewers for comments on
this manuscript. The simulation work is implemented by
advanced software tools (Fluent 6.3.26 and Gambit 2.3.16)
provided by the China National Petroleum Corp. and its
subsidiary company.
Appendix
Nomenclature
Cd ) drag coefficient

ds ) particle diameter, m
D ) pipe diameter, m
es ) particle-particle restitution coefficient
ew ) particle-wall restitution coefficient
g ) gravitational acceleration, m s-2
jjI ) identity matrix
I2D ) second invariant of the deviatoric stress tensor
Kgs ) interphase exchange coefficient, kg m2 s-1
p ) pressure, Pa
ps ) particulate phase pressure, Pa
Res ) particles Reynolds number
t ) flow time, s
Umf ) minimum fluidization velocity
Ut ) particle terminal velocity
Vg ) gas velocity, m s-1
Vs ) solid velocity, m s-1
Vs,w ) solid velocity at the wall, m s-1
Rg ) volume fraction of the gas phase
Rs ) volume fraction of the solid phase
Rs,m ) maximum volume fraction of the solid phase
) voidage
) specularity factor
g ) viscosity of the gas phase, Pa s
s ) solids shear viscosity, Pa s
s,col ) solids collisional viscosity, Pa s
s,kin ) solids kinetic viscosity, Pa s
s,fr ) solids frictional viscosity, Pa s
) angle of internal friction, deg
s ) granular temperature, m2 s-2
s ) collisional dissipation of energy, m2 s-2
jjg ) shear stress of the gas phase, N m-2
jjs ) shear stress of the solid phase, N m-2
s ) solid bulk viscosity, Pa s
Fg ) gas density, kg m-3
Fs ) solid density, kg m-3

Ind. Eng. Chem. Res., Vol. 49, No. 9, 2010

Literature Cited
(1) Ahmadzadeh, A.; Arastoopour, H.; Teymour, F.; Strumendo, M.
Population balance equations application in rotating fluidized bed polymerization reactor. Chem. Eng. Res. Des. 2008, 86, 329.
(2) Khare, N. P.; Luca, B.; Seavey, K. C.; Liu, Y. A. Steady-state and
dynamic modeling of gas-phase polypropylene processes using stirred-bed
reactors. Ind. Eng. Chem. Res. 2004, 43, 884.
(3) Khare, N. P.; Seavey, K. C.; Liu, Y. A.; Ramanathan, S.; Lingard,
S.; Chen, C. C. Steady-state and dynamic modeling of commercial slurry
high-density polyethylene (HDPE) processes. Ind. Eng. Chem. Res. 2002,
41, 5601.
(4) Yiannoulakis, H.; Yiagopoulos, A.; Kiparissides, C. Recent developments in the particle size distribution modeling of fluidized-bed olefin
polymerization reactors. Chem. Eng. Sci. 2001, 56, 917.
(5) Luo, Z. H.; Su, P. L.; You, X. Z.; Shi, D. P.; Wu, J. C. Steady-state
particle size distribution modeling of polypropylene produced in tubular
loop reactors. Chem. Eng. J. 2009, 146, 61.
(6) Wang, T. F.; Wang, J. F.; Jin, Y. A CFD-PBM coupled model for
gas-liquid flows. AIChE J. 2006, 52, 125.
(7) Ahuja, G. N.; Patwardhan, A. W. CFD and experimental studies of
solids hold-up distribution and circulation patterns in gas-solid fluidized
beds. Chem. Eng. J. 2008, 143, 147.
(8) Roy, S.; Dhotre, M. T.; Joshi, J. B. CFD simulation of flow and
axial dispersion in external loop airlift reactor. Chem. Eng. Res. Des. 2006,
84, 677.
(9) Darelius, A.; Rasmuson, A.; van Wachem, B.; Bjorn, I. N.; Folestad,
S. CFD simulation of the high shear mixing process using kinetic theory
of granular flow and frictional stress models. Chem. Eng. Sci. 2008, 63,
2188.
(10) Papadikis, K.; Gu, S.; Bridgwater, A. V. CFD modelling of the
fast pyrolysis of biomass in fluidised bed reactors: modelling the impact of
biomass shrinkage. Chem. Eng. J. 2009, 149, 417.
(11) Doroodchi, E.; Galvin, K. P.; Fletcher, D. F. The influence of
inclined plates on expansion behaviour of solid suspensions in a liquid
fluidised bedsa computational fluid dynamics study. Powder Technol. 2005,
156, 1.
(12) Kobayashi, N.; Yamazaki, R.; Mori, S. A study on the behavior of
bubbles and solids in bubbling fluidized beds. Powder Technol. 2000, 113,
327.
(13) Nieuwland, J. J.; Veneendaal, M. L.; Kuipers, J. A. M.; van Swaaij,
W. P. M. Bubble formation at single orifice in gas-fluidised beds. Chem.
Eng. Sci. 1996, 51, 4087.
(14) Zhang, Y.; Reese, J. M. Continuum modeling of granular particle
flow with inelastic inter-particle collisions. T. Inst. Chem. Eng. 2003, 81,
483.
(15) Taghipour, F.; Ellis, N.; Wong, C. Experimental and computational
study of gas-solid fluidized bed hydrodynamics. Chem. Eng. Sci. 2005,
51, 6857.
(16) Wachem, B.; Schouten, J. C.; Bleek, C.; Krishna, R.; Sinclair, J. L.
Comparative analysis of CFD models of dense gas-solid systems. AIChE
J. 2001, 47, 1035.
(17) Sasic, S.; Johnsson, F.; Leckner, B. Inlet boundary conditions for
the simulation of fluid dynamics in gas-solid fluidized beds. Chem. Eng.
Sci. 2006, 61, 5183.
(18) Dudukovic, M. P. Opaque multiphase flow: experiments and
modeling. Exp. Therm. Fluid Sci. 2002, 26, 747.
(19) Cornelissen, J. T.; Taghipour, F.; Escudie, R.; Escudie, R.; Grace,
J. R. CFD modelling of a liquid-solid fluidized bed. Chem. Eng. Sci. 2007,
62, 6334.
(20) Sinclair, J. L. Hydrodynamic modeling in circulating fluidized beds;
Blackie Academic and Professionals: London, 1997.
(21) Gidaspow, D. Multiphase flow and fluidization: Continuum and
kinetic theory descriptions; Academic Press: Boston, 1994.
(22) de Broqueville, A.; De Wilde, J. Numerical investigation of gassolid heat transfer in rotating fluidized beds in a static geometry. Chem.
Eng. Sci. 2009, 64, 1232.
(23) Gao, J. S.; Chang, J.; Lu, C. X.; Xu, C. M. Experimental and
computational studies on flow behavior of gas-solid fluidized bed with
disparately sized binary particles. Particuology 2008, 6, 59.

4079

(24) Mohammad, A. D.; Shahrokh, S.; Hashemabadi, S. H.; Ghafelebashi, S. M. CFD based evaluation of polymer particles heat transfer
coefficient in gas phase polymerization reactors. Int. Commun. Heat Mass
Transfer 2008, 35, 1375.
(25) Panneerselvam, R.; Savithri, S.; Surender, G. D. CFD simulation
of hydrodynamics of gas-liquid-solid fluidised bed reactor. Chem. Eng.
Sci. 2009, 64, 1119.
(26) Zhang, K.; Brandani, S.; Bi, J. C.; Jiang, J. C. CFD simulation of
fluidization quality in the three-dimensional fluidized bed. Prog. Nat. Sci.
2008, 18, 729.
(27) Busciglio, A.; Vella, G.; Micale, G.; Rizzuti, L. Analysis of the
bubbling behaviour of 2D gas solid fluidized beds. Part II. Comparison
between experiments and numerical simulations via digital image analysis
technique. Chem. Eng. J. 2009, 148, 145.
(28) Lu, H. L.; Liu, W. T.; Li, F.; Zhao, G. B.; He, Y. R. Eulerian
simulation of bubble behaviour in a two-dimensional gas-solid bubbling
fluidized bed. Int. J. Energy Res. 2002, 26, 1285.
(29) Darelius, A.; Vella, G.; Micale, G.; Rizzuti, L. Analysis of the
bubbling behaviour of 2D gas solid fluidized beds. Part II. Comparison
between experiments and numerical simulations via digital image analysis
technique. Chem. Eng. J. 2009, 148, 145.
(30) Witt, P. J.; Perry, J. H.; Schwarz, M. P. A numerical model for
predicting bubble formation in a 3D fluidized bed. Appl. Math. Model. 1998,
22, 1071.
(31) Choi, K. Y.; Ray, W. H. The dynamic behavior of continuous
stirred-bed reactors for the solid catalyzed gas phase polymerization of
propylene. Chem. Eng. Sci. 1988, 43, 2587.
(32) Urdampilleta, I.; Gonzalez, A.; Lruin, J. J.; de la Cal, J. C.; Asua,
J. M. Origins of product heterogeneity in the Spheripol high impact
polypropylene process. Ind. Eng. Chem. Res. 2006, 45, 4178.
(33) Kosek, J.; Grof, Z.; Novak, A.; Stepanek, F.; Marek, M. Dynamics
of particle growth and overheating in gas-phase polymerization reactors.
Chem. Eng. Sci. 2001, 56, 3951.
(34) Kim, J. Y.; Choi, K. Y. Modeling of particle segregation phenomena
in a gas phase fluidized bed olefin polymerization reactor. Chem. Eng. Sci.
2001, 56, 4069.
(35) Chapman, S.; Cowling, T. G. The mathematical theory of nonuniform gases; Cambridge University Press: Cambridge, U.K., 1970.
(36) Lun, C. K. K.; Savage, S. B.; Jeffrey, D. J.; Chepurniy, N. Kinetic
theories for granular flow-inelastic particles in Couette-flow and slightly
inelastic particles in a general flow field. J. Fluid Mech. 1984, 140, 223.
(37) Boemer, A.; Qi, H.; Renz, U. Eulerian simulation of bubble
formation at a jet in a two-dimensional fluidized bed. Int. J. Multiphase
Flow 1997, 23, 927.
(38) Ding, J.; Gidaspow, D. A bubbling fluidization model using kinetictheory of granular flow. AIChE J. 1990, 36, 523.
(39) Syamlal, M.; Rogers, W.; OBrien, T. J. MFIX Documentation,
Theory Guide; National Technical Information Service: Springfield, MA,
1993; Vol. 1.
(40) Gidaspow, D.; Bezburuah, R.; Ding, J. Hydrodynamics of circulating fluidized beds: Kinetic theory approach. In: Fluidization VII, Proceedings
of the 7th Engineering Foundation Conference on Fluidization; 1992; pp
75-82.
(41) Schaeffer, D. G. Instability in the evolution equations describing
incompressible granular flow. J. Differ. Equations 1987, 66, 19.
(42) Wen, C. Y.; Yu, Y. H. Mechanics of fluidization. Chem. Eng. Prog.
Symp. Ser. 1966, 62, 100.
(43) Ergun, S. Fluid flow through packed columns. Chem. Eng. Prog.
1952, 48, 89.
(44) Lettieri, P.; Felice, R. D.; Pacciani, R.; Owoyemi, O. CFD modelling
of liquid fluidized beds in slugging mode. Powder Technol. 2006, 167, 94.
(45) Chai, C. J.; Zhang, G. L. Chemical Engineering Fluid Flows and
Heat Transport; Chemical Engineering Press (Chinese): Beijing, 2004.

ReceiVed for reView September 11, 2009


ReVised manuscript receiVed March 13, 2010
Accepted March 25, 2010
IE901424G

Você também pode gostar