Вы находитесь на странице: 1из 11

Chemical Engineering Science 66 (2011) 11891199

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Dynamic modeling of gas phase propylene homopolymerization in uidized


bed reactors
Ahmad Shamiri a,b, Mohamed Azlan Hussain a,n, Farouq Sabri Mjalli c, Navid Mostou d,
Mohammad Saleh Shafeeyan a
a

Department of Chemical Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia


Research center, Razi Petrochemical Company, P.O. Box 161, Bandar Imam, Iran
c
Petroleum & Chemical Engineering Department, Sultan Qaboos University, Muscat 123, Oman
d
Process Design and Simulation Research Center, School of Chemical Engineering, College of Engineering, University of Tehran, P.O. Box 11155/4563, Tehran, Iran
b

a r t i c l e in f o

abstract

Article history:
Received 29 June 2010
Received in revised form
3 December 2010
Accepted 16 December 2010
Available online 22 December 2010

A new model with comprehensive kinetics for propylene homopolymerization in uidized bed reactors
was developed to investigate the effect of mixing, operating conditions, kinetic and hydrodynamic
parameters on the reactor performance as well as polymer properties. Presence of the particles in the
bubbles and the excess gas in the emulsion phase was considered to improve the two-phase model, thus,
considering the polymerization reaction to take place in both the bubble and emulsion phases. It was
shown that in the practical range of supercial gas velocity and catalyst feed rate, the ratio of produced
polymer in the bubble phase to the total production rate is roughly between 10% and 13%, which is a
substantial amount and cannot be ignored. Simulation studies were carried out to compare the results of
the improved two-phase, conventional well-mixed and constant bubble size models. The improved twophase and well mixed models predicted a narrower and safer window at the same running conditions
compared with the constant bubble size model. The improved two-phase model showed close dynamic
behavior to the conventional models at the beginning of polymerization, but starts to diverge with the
evolution of time.
& 2010 Elsevier Ltd. All rights reserved.

Keywords:
Mathematical modeling
Fluidization
Propylene polymerization
Catalyst
Emulsion
Bubble

1. Introduction
Fluidized bed reactors (FBR) are used in various industrial
applications due to their advantages such as their ability to carry
out different chemical reactions, good particle mixing and high rate
of mass and heat transfer. Due to these advantages, accurate
models of the complex reactions, hydrodynamic features and mass
and heat transfer in the uidized bed reactor are vital to enable
scientists and engineers designing more efcient reactors. Fig. 1
shows schematic of an industrial gas-phase uidized bed polypropylene reactor. As shown in this gure, small particles of
ZieglerNatta catalyst and triethyl aluminum co-catalyst are
charged continuously to the reactor and react with the reactants
(propylene and hydrogen) to produce a broad distribution of
polymer particles. The catalyst particles are porous and composed
of small sub fragments containing titanium as the active metal. As
the monomer diffuses through the porous catalyst, it becomes
polymerized by reacting on the active sites of the catalyst surface.
As the polymerization continues, the catalyst fragments are

Corresponding author. Tel.: + 60 379675214; fax: + 60 379675319.


E-mail address: mohd_azlan@um.edu.my (M.A. Hussain).

0009-2509/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2010.12.030

dispersed and the particle grows into the nal polymer product
(Zacca et al., 1996). The feed gas, which consists of propylene,
hydrogen and nitrogen, provides uidization through the distributor, acts as the heat transfer medium as well as supplys reactants
for the growth of the polymer particles. The uidized particles are
released from the unreacted gases in the releasing zone. The solidfree gas is then combined with the fresh feed stream after heat
removal and recycled back to the reactor through the gas distributor. Most of the unreacted monomer is recycled to the reactor
(McAuley et al., 1990) and the monomer conversion per pass
through the bed can vary from 2% to 3% while the overall monomer
conversion can be as high as 98%. The polypropylene product is
continuously withdrawn from a point close to the bottom of the
reactor and above the gas distributor. The unreacted gas is
recovered from the product, which is sent to the nishing section
of the plant.
In heterogeneous systems, polymerization occurs in the presence of different phases with inter-phase mass and heat transfer
and chemical reaction. Phenomena such as the complex ow
characteristics of gas and solids, the kinetics of heterogeneous
polymerization and various heat and mass transfer mechanisms
must be incorporated in any realistic modeling approach. Various
methods for describing the hydrodynamics of the uidized bed

1190

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

Emulsion phase
Recycle Stream
Catalyst particle

Bubble phase
Compressor

Reactor
Polymer particle
Improved two-phase
Emulsion phase

Catalyst

Powder
PP

Bubble phase

Constant bubble size


Fresh Feed
Propylene
Hydrogen
Nitrogen

One phase

Finishing
Area

Well mixed
Fig. 1. Schematic of an industrial uidized bed polypropylene reactor.

polyolen reactors have been proposed in the literature. McAuley


et al. (1994) and Xie et al. (1994) considered the uidized bed
polyolen reactor as a well mixed reactor. They compared the
simple two-phase and the well mixed models at steady state
conditions and showed that the well mixed model does not present
signicant error in predicting the temperature and monomer
concentration in the reactor compared to the simple two-phase
model. Choi and Ray (1985) presented a simple two-phase model in
which the reactor consists of emulsion and bubble phases. They
assumed that the polymerization reaction occurs only in the
emulsion since the bubbles are solids-free. Fernandez and Lona
(2001) proposed a heterogeneous three-phase model that considers bubble, emulsion and particulate phases with plug ow
behavior. Hatzantonis et al. (2000) assumed that the reactor
comprising of a mixed emulsion and bubble phases is divided into
several solids-free well-mixed compartments in series. Alizadeh
et al. (2004) proposed a tanks-in-series model to represent the
hydrodynamics of the reactor. A comprehensive mathematical
model based on the mixing cell was developed by Harshe et al.
(2004) to simulate the transient behavior of uidized bed polypropylene reactors. This model was coupled with steady state
population balance equations. Rigorous multi-monomer, multisite
polymerization kinetics was also incorporated in the model.
Ibrehem et al. (2008) proposed that the uidized bed can be
composed of bubble, cloud, emulsion and solid phases and
considered the polymerization reactions in the emulsion and solid
phases. This model also accounts for the effect of catalyst particles
type and porosity on the rate of reaction.
In all previous models mentioned above, it was assumed that no
reaction takes place in the bubble phase. Kiashemshaki et al. (2006)

divided the reactor into four serial sections, each consisting of the
bubble phase as plug ow and the emulsion phase as well mixed.
They modeled the reactor at steady state conditions and considered
that polymerization reactions occur in both the emulsion and
bubble phases. To gain a better understanding of propylene
homopolymerization over a heterogeneous ZieglerNatta catalyst
in uidized bed reactors, a two-site homopolymerization kinetic
scheme has been employed in this study. Assuming that mass and
heat transfer resistances between the solid polymer particles and
the emulsion gas are negligible, extensive and comparative study
using the well-mixed, constant bubble size and improved twophase models were carried out to investigate the inuence of key
features (such as supercial gas velocity, catalyst feed rate,
monomer and hydrogen feed concentration and feed temperature)
on the predicted reactor variables and polypropylene properties.

2. Polymerization mechanism and kinetic model


In the present study, a comprehensive mechanism was considered to describe the kinetics of propylene homopolymerization over
a ZieglerNatta catalyst based on the kinetic model developed by
Shamiri et al. (2010). Characterization of the polymer properties was
accomplished using the method of moments. The set of mass balance
equations for the active sites and reacted monomers, which are
represented by a series of algebraic and differential equations,
were used separately for the emulsion phase as a well mixed
and the bubble phase as a plug ow reactors with a two-site type
catalyst, which represents the situation more commonly faced with
heterogeneous ZieglerNatta catalysts. Although the mechanism of

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

polymerization in both phases is the same, the rates of reaction in the


bubble and emulsion phases are different based on this dynamic twophase model due to different amounts of solids in each phase and
volume of polymer in the bubble phase (Vpb) and the emulsion phase
(Vpe) as described in Section 3.2. Introducing different ow rates of
catalyst inside the bubble and emulsion phases results in different
reaction conditions for the two phases and affects the production
rate, temperature and monomer conversion in both phases.
Assuming that the only signicant consumption of monomer is
by the propagation reaction and that of hydrogen by the transfer to
hydrogen reaction, the following expression for the consumption
rate of components (monomer and hydrogen) can be obtained:
For monomer:
Ri

NS
X

Mi Y0,jkp j

i1

i2

j1

For hydrogen:
Ri

NS
X

Mi Y0,jkfh j

j1

The total polymer production rate for each phase can be then
calculated from
Rp

2
X

Mwi Ri

i1

where Ri is the instantaneous rate of reaction for monomer and


hydrogen. The polymer production rates of the emulsion phase are
indicated as Rpe and the bubble phase as Rpb.
The reaction rate constants were taken from literature and are
given in Table 1. In this study, the effect of temperature (i.e.,
activation energies) on the polymerization kinetics was considered
only for the propagation reaction. It has been well established that
under various conditions, when the catalyst particles are small
enough and the catalyst activity is not high (low to moderate
polymerization rates), mass and heat transfer resistances inside the
polymer particles and between the gas and the solid polymer
particles do not play an important role and will not signicantly
affect the behavior of the reactor as well as the properties of the
polyolen (Zacca et al., 1996).

1191

models to describe this behavior. A combination of kinetics,


hydrodynamics and transport phenomena are required for modeling these reactors. Choi and Ray (1985) proposed a two-phase,
constant bubble size model consisting of a mixed emulsion phase
and a plug-ow bubble phase to describe the dynamic behavior of
the FBR. McAuley et al. (1994) proposed a simplied well-mixed
model by assuming an unrestricted mass and heat transfer
between the bubble and emulsion phases.
In the present study, a unied modeling approach was adopted
for the gassolid uidization. The two-phase ow model was
developed to describe the dynamic behavior of the polypropylene
gas phase uidized bed reactor. To calculate a better estimation of
the average bed voidage and mass and energy balances equations,
the improved two-phase model was derived based on a combination of the kinetic developments described in the previous section
and the dynamic two-phase model (Alizadeh et al., 2004;
Kiashemshaki et al., 2006; Jafari et al., 2004; Cui et al., 2000, 2001).
3.1. Simplied hydrodynamic models
The simple two-phase ow approach for the gas-phase olen
polymerization model has been used previously by many researchers (McAuley et al., 1994; Choi and Ray, 1985; Hatzantonis et al.,
2000; Fernandez and Lona, 2001; Harshe et al., 2004). This model
considers that the bubbles are solids-free while the emulsion
remains at minimum uidization conditions. However, the voidage
of the emulsion phase may vary far from that at the minimum
uidization. In addition, the bubbles may contain different portions
of solid particles (Cui et al., 2000, 2001). Based on this idea, Cui et al.
(2000, 2001) proposed the dynamic two-phase ow for the
uidized bed hydrodynamics (particle concentration in emulsion
and bubbles varies with gas velocity). Since assuming the minimum
uidization condition for the emulsion phase in the polypropylene
reactor (simple two-phase model) is not realistic (Alizadeh et al.,
2004), the dynamic two-phase ow of uidized beds, proposed by
Cui et al. (2000, 2001), was used for improving the two-phase model
in this study and to calculate a better estimation of the average bed
voidage. The correlations needed for estimating the bubble volume
fraction in the bed, the voidage of the emulsion phase and bubble
phases, the emulsion phase and bubble phases gas velocities and
mass and heat transfer coefcients for the improved two-phase and
constant bubble size models are summarized in Table 2.

3. Hydrodynamics
Complex mixing and contacting ow patterns, transport phenomena and polymerization reactions make uidized bed reactors
highly non-ideal and difcult to characterize. Many studies have
been attempted to model such non-ideality using various mixing

3.1.1. Well-mixed model


McAuley et al. (1990, 1994) proposed a simplied well-mixed
model by assuming that bubbles are small with unrestricted mass
and heat transfer between the bubble and emulsion phases and
that the temperature and composition are uniform in the gas phase

Table 1
Reactions rate constants.
Reaction

Rate constant

Unit
1

Formation

kf(j)

Initiation

ki(j)
kh(j)
khr(j)

l mol  1 s  1
l mol  1 s  1
l mol  1 s  1

Propagation
Activation energy

kp(j)

l mol  1 s  1
k cal mol  1

Transfer

kfm(j)
kfh(j)
kfr(j)
kfs(j)

l mol  1
l mol  1
l mol  1
l mol  1 s  1

Deactivation

kds(j)
kdI(j)

s1
l mol  1 s  1

Site type 1
1
22.88
0.1
20
208.6
7.2
0.0462
7.54
0.024
0.0001
0.00034
2000

Site type 2
1

Reference
Shamiri et al. (2010)

54.93
0.1
20

Luo et al. (2009)


Shamiri et al. (2010)
Shamiri et al. (2010)

22.8849
7.2

Luo et al. (2009)

0.2535
7.54
0.12
0.0001
0.00034
2000

Luo et al. (2009)


Luo et al. (2009)
Shamiri et al. (2010)
Shamiri et al. (2010)
Luo et al. (2009)
Shamiri et al. (2010)

1192

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

Table 2
Correlations and equations used in the two -phase and constant bubble size models.
Parameter

Formula

Reference

Minimum uidization velocity

Remf 29:52 0:357Ar1=2 29:5


Ub U0  Umf + ubr
ubr 0.711(gdb)1/2

Lucas et al. (1986)

Bubble velocity
Bubble rise velocity
Emulsion velocity

mf
Ue e 1
d
mf
db db0[1+ 27(U0  Ue)]1/3(1+ 6.84H) db0 0.0085 (for Geldart B)

1
Kbe K1bc K1ce
 1=2 
 
D g 1=4
Kbc 4:5 Ud e 5:85 g 5=4
b
db


D e u
Kce 6:77 g deb br

1
Hbe H1bc H1ce
U r C 
k r C 1=2 g 1=4
Hbc 4:5 e dg pg 5:85 g g pg5=4
b
db

1=2
Hce 6:77rg Cpg kg 1=2 eedu3br

Bubble diameter
Mass transfer coefcient

Heat transfer coefcient

Kunni and Levenspiel (1991)


Kunni and Levenspiel (1991)
Kunni and Levenspiel (1991)
Hilligardt and Werther (1986)
Kunni and Levenspiel (1991)

Kunni and Levenspiel (1991)

U0 Umf
Ub

Bubble phase fraction

Emulsion phase porosity


Volume of polymer phase in the emulsion phase

ee emf

Kunni and Levenspiel (1991)

Vp AH(1  ee)(1  d)

throughout the bed. The following assumptions were also made for
the well mixed model:
1. Mass and heat transfer rates between emulsion and bubble
phase are high or bubbles are small enough; therefore, the
polymerization reactor is considered to be a single phase
(emulsion phase), well mixed reactor (McAuley et al., 1994;
Alizadeh et al., 2004).
2. The emulsion phase remains at minimum uidization.
3. Uniform temperature and composition exists throughout the bed.
Based on the above assumptions, dynamic material and energy
balance equations can be written for the monomer and hydrogen
based on the previously mentioned assumptions (Hatzantonis
et al., 2000). The mole balance is


dMi 
V emf
U0 A Mi in Mi  Rv emf Mi 1emf Ri
dt

The energy balance is given by


"

m
X

#
Mi Cpi V emf V1emf rpol Cp,po1

i1

U0 A

m
X

"
Mi Cpi TTref Rv

i1

1emf DHR Rp

m
X

m
X
dT
Mi Cpi Tin Tref
U0 A
dt
i1
#

The assumptions of the constant bubble size model are as follows:


1. The uidized bed comprises of two phases, bubble and emulsion, and reactions occur only in the emulsion phase.
2. The emulsion phase is considered to be perfectly mixed, at
minimum uidization, interchanging heat and mass with the
bubble phase at uniform rates over the bed height.
3. The bubbles are spherical with uniform size and in plug ow at
constant velocity.
4. Mass and heat transfer resistances between the gas and the solid
polymer particles in the emulsion phase are negligible (i.e.,
small catalyst particles, low to moderate catalyst activity or
polymerization rates; Floyd et al., 1986).
Using these assumptions, steady-state mass and energy balances can be derived to describe the variation of the monomer
concentration and temperature in the bubble phase (Hatzantonis
et al., 2000). The mole balance for monomer and hydrogen can be
written as follows:
dMi b
K
be Mi e Mi b
dz
Ub

Mi Cpi emf 1emf rpo1 Cp,po1 TTref

i1

Internal energy of the monomer is considered to be negligible in


the energy balance equation. The initial conditions for solving the
model equations are as follows:
Mi t o Mi in

Tt 0 Tin

By integrating the local monomer concentration [Mi]b on the


bed height, one can calculate the mean concentration of the i-th
monomer in the bubble phase
Z


1 H
M i b
Mi b dh Mi e Mi e,in Mi e
H 0



U
K H
1exp  be
9
 b
Ub
Kbe H
The bubble phase energy balance is expressed by the following
equation:
m
X

3.1.2. Constant bubble size model


The two-phase constant bubble size model proposed by Choi
and Ray (1985) considers that the emulsion phase (or the dense
phase) is at minimum uidization conditions. This model has been
adopted in many previous studies for the gas-phase olen polymerization (McAuley et al., 1994; Choi and Ray, 1985; Hatzantonis
et al., 2000; Harshe et al., 2004).

i1

Mi b Cpi

dTb
H
be Tb Te
dz
Ub

10

The mean temperature in the bubble phase can be calculated by


integrating Eq. (10) over the total bed height
!!
Z
U Cp
1 H
H H
1exp  be
Tb
Tb dh Te Tin Te b
11
H 0
Hbe H
Ub C p

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

where the average heat capacity of the reacting components is


given as
Cp

Nm h
X

Mi

i
b

i1

CpMi

12

The dynamic molar balance for the i-th component in the


emulsion phase can be written as
Ve emf

"

dMi e
Ue Ae emf Mi e,in Mi e
dt
Ve dKbe
M i b Mi e Rv emf Mi e 1emf Ri

1d

13

The dynamic emulsion phase energy balance can be written as


#
dTe
Ve emf Mi e Cpi Ve 1emf rPo1 Cp,po1
dt
i1
m
X

m
X

Ve emf Cpi

i1

Ue Ae emf
Ue Ae emf

m
X

dMi e
Te Tref
dt

Mi e,in Cpi Te,in Tref

i1
m
X

Ve dHbe
Te T b
1d
!
m
X
Mi e Cpi Te Tref

Rv 1emf rpo1 Cp,po1 emf

i1

1emf DHR Rp

up through the bed at a constant velocity and the particles present a


downward ow, growing in size and weight as they ow downwards. These facts make the assumption of the plug ow valid for
such a regime in the bubble phase.
The assumptions made in developing the equations of the
improved model are summarized below:
1. The polymerization reactions occur in both bubble and emulsion phases.
2. The emulsion phase is considered to be perfectly mixed and
does not remain at the minimum uidization condition.
3. The bubbles are assumed to be spherical with uniform size and
travel up through the bed in plug ow at constant velocity.
4. Mass and heat transfer resistances between gas and solids in the
emulsion and bubble phases are negligible (i.e., low to moderate
catalyst activity; Floyd et al., 1986).
5. Radial concentration and temperature gradients in the reactor
are negligible due to the agitation produced by the upowing gas.
6. Elutriation of solids at the top of the bed is negligible.
7. Constant particle size is assumed throughout the bed.
Based on these assumptions, the following dynamic material
balances were written for all the components in the bed.
For bubbles:

Mi e Cpi Te Tref 

i1

1193

14

The boundary and initial conditions for solving the model


equations are as follows:
Mi b,z 0 Mi in

15

Tb z 0 Tin

16

Mi e,t 0 Mi in

17

Te t 0 Tin

18

3.2. Improved hydrodynamic model


The conventional constant bubble size and well mixed models
assume that the emulsion is at minimum uidization (ee emf) and
that bubbles are solids-free (eb 1). By these assumptions, it is not
possible to predict the effect of gassolid distribution on the
apparent reaction and heat/mass transfer rates in the uidized
beds properly at velocities higher than the minimum uidization.
However, the presence of solids in the bubbles has been shown
experimentally (Aoyagi and Kunii, 1974) and theoretically
(Gilbertson and Yates, 1996). The emulsion phase also does not
remain at minimum uidization conditions and it may contain
more gas at higher gas velocities (Abrahamson and Geldart, 1980).
A better mixing of the two phases, which results in more solid
particles entering the bubbles and more gas entering the emulsion
phase, occurs while the supercial gas velocity increases in a
uidized bed. In the present study, the dynamic two-phase model,
proposed by Cui et al. (2000, 2001), was employed to provide a
more realistic understanding of the hydrodynamics phenomena
and improve the quantitative understanding of the actual process.
In the bubbling uidized bed reactor, the upward motion of the
gas bubbles causes enough mixing of solid particles in the emulsion
phase or dense phase. Consequently, the concentrations of various
species and temperature are nearly uniform in the emulsion phase.
Therefore, a pseudo-homogeneous well mixed model can be
applied to this phase (McAuley et al., 1990). The bubbles travel

Mi b,in Ub Ab Mi b Ub Ab Rv eb Mi b Kbe Mi b Mi e Vb


Z
A
d
V e M 
1eb b
Rib dz
dt b b i b
VPFR

19

For emulsion:
Mi e,in Ue Ae Mi e Ue Ae Rv ee Mi e Kbe Mi b Mi e Ve

1d

d
Ve ee Mi e
1ee Rie
dt

20

The direction of mass transfer was assumed to be from the


bubble to emulsion phase.
The energy balances can be expressed as
For bubbles:
m
X

Ub Ab Tb,in Tref

Mi b,in Cpi Ub Ab Tb Tref

i1

Mi b Cpi

i1

m
X

Rv Tb Tref

m
X

eb Cpi Mi b 1eb rpol Cp,pol

i1

A DHR
1eb b
VPFR

Rpb dz

m
X

d
Mi b
dt
i1
!
m
X
d
T T
Cpi Mi b 1eb rpol Cp,pol
Vb eb
dt b ref
i1

Hbe Te Tb Vb Vb eb Tb Tref

Cpi

21

For emulsion:
m
X

Ue Ae Te,in Tref

Mi e,in Cpi Ue Ae Te Tref

i1
m
X

Rv Te Tref
Hbe Ve

Mi e Cpi

i1

ee Cpi Mi e 1ee rpol Cp,pol 1ee Rpe DHR

i1

1d

Ve ee

m
X

m
X
i1

Te Tb Ve ee Te Tref

m
X
i1

Cpi

!!

Cpi Mi e 1ee rpol Cp,pol

d
Mi e
dt
d
Te Tref
dt

22

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

where
U0 dUb
Ue
1d

23

Ub U0 Ue ubr

24




U0 Umf
d 0:534 1exp 
0:413

25

ee emf 0:20:059exp 


eb 10:146exp 

U0 Umf
0:429

U0 Umf
4:439


26


27

VPe AH1ee 1d

28

VPb AH1eb d

29

Ve A1dH

30

Vb AdH

31

The initial conditions for solving the model equations are as


follows:
Mi b,t 0 Mi in

32

Tb t 0 Tin

33

Mi e,t 0 Mi in

34

Te t 0 Tin

35

4. Results and discussion


Dynamic modeling and simulation studies of the gas phase
propylene polymerization in the uidized bed reactor was carried
out using the improved two-phase, conventional well mixed and
constant bubble size models along a comprehensive two-site kinetic
scheme to verify the effects of different hydrodynamic sub-models,
mixing conditions and model assumptions on the dynamic response
and grade transition of the process. Comprehensive and comparative simulations were conducted to evaluate the effect of key
features such as catalyst feed rate, supercial gas velocity, monomer, hydrogen and impurities feed concentrations on the polymer
production rate, emulsion and bubble phase temperature and the
properties of the polypropylene product. Simulations were carried
out at the operating conditions shown in Table 3.
Based on the advantages of the improved model over the previous
ones, it can be expected that the improved model provides a more
realistic result compared to the well mixed and constant bubble size
models. Even though the model has not been validated experimentally, the recommendations of previous experimental ndings were
considered in the improved model. It is worth mentioning that the
same modeling approach was adopted for polyethylene production
Table 3
Operating conditions and physical parameters considered in this work for modeling
uidized bed polypropylene reactors.
Operating conditions

Physical parameters

V (m3) 50
Tref (K)353.15
Tin (K) 317.15
P (bar) 25
Propylene concentration (mol/l) 1
Hydrogen concentration (mol/l) 0.015
Catalyst feed rate (g/s) 0.2

m (Pa.s) 1.14  10  4
rg (kg/m3) 23.45
rs (kg/m3) 910
dp (m) 500  10
emf 0.45

6

and compared with the steady state industrial data for the molecular
weight distribution (Kiashemshaki et al., 2006) and the results
showed that the improved two-phase model for this system is in
good agreement with the industrial data.

4.1. Effect of catalyst feed rate


Catalyst feed rate is an important operating parameter in the
operation of polypropylene uidized bed reactors. Simplied
hydrodynamic models ignore the existence of catalyst in the
bubbles and assume that the polymer is produced only in the
emulsion. However, using the improved hydrodynamic model, it
was possible to show that the emulsion phase contains about 88%
of the catalyst while the bubbles carry about 12% of the catalyst
introduced into the reactor. Therefore, the portion of reaction
occurring in the bubbles is signicant and must be considered
(Kiashemshaki et al., 2006; Shamiri et al., 2010).
Fig. 2 shows the effect of catalyst feed rate on the polypropylene
production rate predicted by the three models and the production
distribution rate in emulsion and bubble phases predicted by the
improved two-phase model. It is clear that the polymer production
rate is directly proportional to the catalyst feed rate. As the catalyst
feed rate increases, the polymer production rate also increases due
to the increase in the available active sites. The polymer production
rate predicted by the improved two-phase model is lower than
the conventional models. Consideration of the excess gas in the
emulsion phase increases the void fraction, which leads to a
decrease in the polymer production rate compared to the conventional models that assume an emulsion phase at minimum
uidization. The improved two-phase model shows that the rate
of changes of production of the emulsion phase is higher than that
of the bubble phase. This is mainly due to the higher amount of
catalyst in the emulsion phase compared to that in the bubble
phase. Nevertheless, it is noticeable that about 12% of the polymer
is produced in the bubble phase.
Fig. 3a and b shows the effect of catalyst feed rate on
temperature and monomer concentration, respectively, in the
emulsion and bubble phases predicted by the three models. As
the catalyst feed rate increases, the temperature of emulsion phase
and the bubble phase increases, while the monomer concentration
decreases due to the increase in the reaction rate. The improved
two-phase model temperature prediction clearly lies between the
responses gained by the other two models and it is closer to the
well-mixed model response. The constant bubble size model shows

Production rate (g/s)

1194

3250
3000
2750
2500
2250
2000
1750
1500
1250
1000
750
500
250
0
0.1

Improved two phase


Well mixed
Constant bubble size
Rpe
Rpb

0.2

0.3
0.4
Catalyst feed rate (g/s)

0.5

Fig. 2. Effect of catalyst feed rate on the polymer production rate predicted by the
three models and the production rate distribution in the emulsion and bubble
phases predicted by the improved two-phase model (U0 0.35 m/s).

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

365
360

Temperature (K)

355
350

Improved two-phase, emulsion


Well mixed, emulsion
Constant bubble size, emulsion
Improved two-phase, bubble

345
340
335
330
325
320
315

Propylene concentration (mol/lit)

1.000
0.995
0.990
0.985
0.980
0.975
0.970
0.965
0.1

Improved two-phase, emulsion


Well mixed, emulsion
Constant bubble size, emulsion
Improved two-phase, Bubble

0.2

0.3
0.4
Catalyst feed rate (g/s)

0.5

Fig. 3. Effect of catalyst feed rate on (a) the temperature and (b) propylene
concentration in the emulsion phase and the bubble phase predicted by the three
models at U0 0.35 m/s.

higher emulsion phase temperature and lower monomer concentration compared to the other models. In this model, it was
assumed that the energy is transferred between the phases due
to existence of the temperature gradient between the bubbles and
the emulsion as well as by the diffusing gas. On the other hand, the
only mechanism by which mass can be transferred from the
bubbles to the emulsion phase is by diffusion through the bubble
clouds. Due to the short bubble mean residence time, the model
uses the mean temperature and concentration in the bubble phase
over the bed height (Choi and Ray, 1985; Hatzantonis et al., 2000).
Thus, the temperature evaluated by the model is over-predicted
while the monomer concentration is under-predicted in the
emulsion phase. It can be seen in Fig. 3a and b that the temperature
of the emulsion phase is higher and the monomer concentration is
lower than that of the bubble phase due to higher amount of
catalyst in the emulsion phase compared to the bubble phase.
Comparing the emulsion phase and the bubble phase temperature
proles with the monomer concentration proles reveals the
reverse relationship between these two variables.

4.2. Effect of supercial gas velocity


Direct relation of the supercial gas velocity to the monomer
residence time in the reactor, heat removal rate from the reactor,
particle mixing and uidization conditions makes it an important
process parameter. Thus, it is important to study its effect on these
process conditions. The effect of supercial gas velocity on the

1195

polymer production rate predicted by various models is illustrated


in Fig. 4. The polymer production rate decreases as the supercial
gas velocity is increased. An increase in the supercial gas velocity
decreases the monomer mean residence time, which leads to a
decrease in the monomer conversion and polymer production rate.
The production rate predicted by the improved model is less than
the conventional models due to considering emulsion phase being
at conditions beyond the minimum uidization velocity of gas,
which results in a higher void fraction and lower reaction rate of the
emulsion phase.
Effect of supercial gas velocity on the polymer production rate
of emulsion and bubble phases at various catalyst feed rates
predicted by the improved two-phase model is illustrated in
Fig. 5. By increasing the supercial gas velocity, polymer production rate in the emulsion and bubble phases decreases since it
decreases the monomer mean residence time, which leads to lower
monomer conversion and polymer production rate.
Fig. 6 shows the effect of supercial gas velocity on the ratio of
the bubble phase production rate to the total production rate
predicted by the improved two-phase model. As shown in this
gure, the increase in supercial gas velocity increases the ratio of
polymer production in the bubble phase to the total production
rate (Rpb/Rp). By increasing the supercial gas velocity, more gas
and solids enter the bubbles that results an increase in the bubble
contribution to the reaction rate. The bubble contribution to the
total production is roughly 1013%, which is a signicant amount
that should be considered for a more reliable model prediction
(Kiashemshaki et al., 2006; Shamiri et al., 2010).
Fig. 7 displays the effect of supercial gas velocity on the
emulsion phase temperature predicted by various models. As
shown in this gure, the increase in the supercial gas velocity
results in lower temperature for the emulsion phase due to higher
convective cooling by the gas. The results show that the temperature predicted by the constant bubble size model is higher than
those predicted by the improved two-phase and well mixed models
at the same operating conditions. Furthermore, the constant
bubble size model predicts a temperature above the accepted
industrial safety limit of 80 1C (Meier et al., 2001). Working above
this critical temperature may lead to particle melting, agglomeration and resulting in reactor shutdown. High gas velocities are
required to reduce these risks. However, high gas velocities reduce
the monomer conversion per pass through the reactor bed and can
lead to greater elutriation of small particles from the bed. Under the
same conditions, the improved and well mixed models predict an
emulsion phase temperature below the maximum safe bed operation temperature in which the bed does not undergo temperature
runaway. As shown in Figs. 3a and 7, the safe range of variation in
the main operating parameters such as the catalyst feed rate and
supercial gas velocity calculated by the improved model and the
well mixed model are wider than that achieved by the constant
bubble size model.
Fig. 8 shows the effect of supercial gas velocity on the emulsion
and bubble phase temperatures at different catalyst feed rate
predicted by the improved hydrodynamic model. This gure
illustrates that increase in supercial gas velocity results in a
decrease in temperature of the emulsion and bubble phases due to
lower conversion of the monomer and higher convective cooling
capacity of the gas. Small values of supercial gas velocity lead to
higher sensitivity of the emulsion and bubble phases temperatures
to changes in the catalyst feed rate.

4.3. Effect of feed composition


Catalyst activity is sensitive to existing impurities such as
carbon monoxide. Effect of carbon monoxide concentration on

1196

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

2400

Improved two-phase
Well mixed
Constant bubble size

Production rate (g/s)

2200
2000
1800
1600
1400
1200
1000
800

0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65
U0 (m/s)
Fig. 4. Effect of supercial gas velocity on the product rate (Fcat 0.2 g/s).

Production rate (g/s)

4000

Fcat=0.2 g/s
Fcat=0.4 g/s
Fcat=0.6 g/s

3500
3000
2500
2000
1500
1000
500

370
Emulsion phase temperature (K)

4500

Improved two-phase
Well mixed
Constant bubble size

365
360
355

Safety limit

350
345
340
335
330
325

0
0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65
U0 (m/s)
Fig. 5. Effect of supercial gas velocity on the production rate of bubble (solid) and
emulsion (dashed) phases at various catalyst feed rates predicted by the improved
two-phase model.

320
0.2

11.5

Fcat=0.2 g/s
Fcat=0.4 g/s
Fcat=0.6 g/s

11.0
10.5
10.0
0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65
U0 (m/s)

Fig. 6. Effect of U0 on the production rate ratio (Rpb/Rp) at different catalyst feed rate
predicted by the improved two-phase model.

Temperature (K)

Production rate ratio (Rpb/Rp (%))

13.0

12.0

0.4
U0 (m/s)

0.5

0.6

Fig. 7. Effect of supercial gas velocity on the emulsion phase temperature


(Fcat 0.2 g/s).

13.5

12.5

0.3

390
385
380
375
370
365
360
355
350
345
340
335
330
325
320

Fcat=0.2 g/s
Fcat=0.4 g/s
Fcat=0.6 g/s

0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65
U0 (m/s)
Fig. 8. Effect of U0 on the emulsion phase (dashed) and the bubble (solid) phase
temperature predicted by the improved two-phase model at different catalyst
feed rate.

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

1350

Production rate (g/s)

Emulsion phase temperature (K)

Rp
Rpe
Rpb

1200
1050
900
750
600
450
300
150
0
0.0

0.5
1.0
1.5
2.0
Carbon monoxide concentration (ppm)

342
340
338
336
334
332
330
328
326
324
322
320
318
316

Fig. 9. Effect of carbon monoxide concentration in the feed on the production rate
(U0 0.35 m/s and Fcat 0.2 g/s).

4.4. Dynamic response of the reactor


Comparison of the three models for their dynamic responses was
also carried out in the present work. In all cases, the reactor started
with a bed of inert polymer pellets and circulating gas when the
catalyst feed ow starts (at time zero). Comparison between the
dynamic responses of the well mixed, constant bubble size and
improved two-phase models for temperature and monomer concentration is made in Fig. 10a and b. These gures show that there are
similar trends in the evolution of temperature and concentration at
the start-up conditions of the reactor. However, at the steady state
conditions (after 1 h) when the emulsion phase temperature and
concentration reach their nal values, it is clear that the models reach
different values. As shown in Fig. 10a and b, the constant bubble size
model shows the largest deviation compared to the other two
models. As previously described, this is mainly due to the assumptions of the constant bubble size model in energy and mass balance
equations. The improved two-phase model shows closer behavior to
the well mixed model compared to the constant bubble size model.
The emulsion phase temperature and concentration proles
predicted by the improved model lie between those gained by the
other two conventional models. This is mainly due to the assumption of the emulsion phase not being at minimum uidization
condition, considering of the catalyst distribution between the
emulsion and bubble phases and polymerization reaction in both
the bubble and emulsion phases. It is well known that since the gas
velocity the bubbling uidized bed reactor needs to be high enough
to keep the particles in uidized state, the residence time of
monomers is low and the monomer conversion per pass through
the bed is quite low (less than 3%). Therefore, a large amount of
unreacted gas should be recycled to the reactor, typically on the
order of about 40 times the rate of make-up gas for more reaction in
commercial uidized bed reactors (McAuley et al., 1990; Alizadeh
et al., 2004). Thus, the variation in the monomer concentration in
each pass of the gas through the reactor predicted by the three
models are low compared to the initial value of monomer
concentration (1 mol/l), which also show the conversion being
less than 1.5% in all cases. In summary, the improved two-phase

2000

4000
6000
Time (s)

8000

10000

1.000
Emulsion phase propylene
concentration (mol/lit)

the emulsion and the bubble phase production rates, predicted by


the improved model is shown in Fig. 9. Higher carbon monoxide
concentration results in lowering the production rate in both
emulsion and bubble phases. Adsorption of carbon monoxide on
the catalyst sites deactivates the sites and results in an instantaneous drop in the propagation rate.

Improved two-phase
Well mixed
Constant bubble size

2.5

1197

Improved two-phase
Well mixed
Constant bubble size

0.998
0.996
0.994
0.992
0.990
0.988
0.986
0.984
0

2000

4000
6000
Time (s)

8000

10000

Fig. 10. Evolution of the emulsion phase: (a) temperature and (b) propylene
concentration over the time in the reactor (Fcat 0.2 g/s and U0 0.35 m/s).

model shows the same dynamic behavior to the constant bubble


size and well mixed models at the beginning of polymerization but
starts to diverge as time evolves.

5. Conclusions
An improved two-phase approach was adopted for modeling
polypropylene production in a gas-phase uidized bed reactor.
The hydrodynamics of the uidized bed reactor of polypropylene
production was based on the dynamic two-phase concept of
uidization. This hydrodynamic model was combined with a kinetic
model to provide a better understanding of the reactor performance.
Comparative simulation studies were carried out using the
improved two-phase versus the conventional well mixed and
constant bubble size models to investigate the effect of mixing,
operating conditions, kinetic and hydrodynamic parameters on the
reactor performance as well as on the polymer properties. In the
improved two-phase model, polymerization reaction was considered to take place in both the bubble and emulsion phases to
provide a more realistic understanding of the phenomena occurring in the bed. The constant bubble size model was found to overpredict the emulsion phase temperature and propylene conversion
under typical operating conditions, while the improved two-phase
and well mixed models were in better agreement and predicted a
narrower and safe operation window at the same operating
conditions compared to the constant bubble size model.
The improved two-phase model showed that by considering the
practical range of the supercial gas velocity from 0.2 to 0.65 m/s

1198

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

and catalyst feed rate from 0.2 to 0.6 g/s, the ratio of polymer
production in the bubble phase to the total production rate (Rpb/Rp)
is roughly 1013%, which is a signicant amount that should be
considered in the model. Besides, it was shown that the supercial
gas velocity and catalyst feed rate have a strong effect on the
hydrodynamics and the reaction rate, which results in a greater
variation in the total production rate ratio. The improved twophase model shows close dynamic behavior to the constant bubble
size and well mixed models at the beginning of polymerization but
starts to differ at longer times. The presence of carbon monoxide in
the feed gas results in a decrease in the polypropylene production
rate due to catalyst deactivation and ignoring the presence of the
carbon monoxide in the feed gas would result in signicant error in
predicting the correct reactor performance.

Nomenclature
A
Ar
Cpi
Cpg
Cp,pol
CpMi
db
dp
Dg
Dt
Fcat
H
Hbe
Hbc
Hce
j
kdI(j)
kds(j)
kf (j)
kfh(j)
kfm(j)
kfr(j)
kfs(j)
kg
kh(j)
khr(j)
ki(j)
kp(j)
Kbc
Kbe
Kce
m
[Mi]
[Mi]in
Mi
Mw
NS
P
PP
r
R
Ri

cross sectional area of the reactor (m2)


Archimedes number
specic heat capacity of component i (J/kg K)
specic heat capacity of gaseous stream (J/kg K)
specic heat capacity of solid product (J/kg K)
specic heat of component i (J/kmol K)
bubble diameter (m)
particle diameter (m)
gas diffusion coefcient (m2/s)
reactor diameter (m)
catalyst feed rate (kg/s)
height of the reactor (m)
bubble to emulsion heat transfer coefcient (W/m3 K)
bubble to cloud heat transfer coefcient (W/m3 K)
cloud to emulsion heat transfer coefcient (W/m3 K)
active site type
deactivation by impurities rate constant for a site of type j
spontaneous deactivation rate constant for a site of type j
formation rate constant for a site of type j
transfer rate constant for a site of type j with terminal
monomer M reacting with hydrogen
transfer rate constant for a site of type j with terminal
monomer M reacting with monomer M
transfer rate constant for a site of type j with terminal
monomer M reacting with AlEt3
spontaneous transfer rate constant for a site of type j with
terminal monomer M
gas thermal conductivity (W/m K)
rate constant for reinitiating of a site of type j by monomer M
rate constant for reinitiating of a site of type j by cocatalyst
rate constant for initiation of a site of type j by monomer M
propagation rate constant for a site of type j with terminal
monomer M reacting with monomer M
bubble to cloud mass transfer coefcient (s  1)
bubble to emulsion mass transfer coefcient (s  1)
cloud to emulsion mass transfer coefcient (s  1)
total number of components
concentration of component i in the reactor (kmol/m3)
concentration of component i in the inlet gaseous stream
i-th component
monomer molecular weight (kg/kmol)
number of active site types
pressure (Pa)
polypropylene
number of units in polymer chain
instantaneous consumption rate of monomer (kmol/s)
instantaneous rate of reaction for monomer i (kmol/s)

R(j)
Remf
Rp
Rpb
Rpe
Rv
t
T
Tin
U0
Ub
Ue
Umf
V
Vb
Ve
Vp
Vpb
Vpe
VPFR
Y(n,j)

rate at which monomer M is consumed by propagation


reactions at sites of type j
Reynolds number of particles at minimum uidization
condition
production rate (kg/s)
bubble phase production rate (kg/s)
emulsion phase production rate (kg/s)
volumetric polymer phase outow rate from the reactor
(m3/s)
time (s)
temperature (K)
temperature of the inlet gaseous stream (K)
supercial gas velocity (m/s)
bubble velocity (m/s)
emulsion gas velocity (m/s)
minimum uidization velocity (m/s)
reactor volume (m3)
volume of the bubble phase
volume of the emulsion phase
volume of polymer phase in the reactor (m3)
volume of polymer phase in the bubble phase (m3)
volume of polymer phase in the emulsion phase (m3)
volume of plug ow reactor (m3)
n-th moment of chain length distribution for living
polymer produced at a site of type j

Greek letters

DHR
d

eb
ee
emf
m
r
rg
rpol

heat of reaction (J/kg)


volume fraction of bubbles in the bed
void fraction of bubble for Geldart B particles
void fraction of emulsion for Geldart B particles
void fraction of the bed at minimum uidization
gas viscosity (Pa.s)
density (kg/m3)
gas density (kg/m3)
polymer density (kg/m3)

Subscripts and superscripts


1
2
b
e
g
i
in
j
mf
pol
ref

propylene
hydrogen
bubble phase
emulsion phase
gas mixture property
component type number
inlet
active site type number
minimum uidization
polymer
reference condition

Acknowledgment
The authors thank the Research Council of University of Malaya for
its support to this research under research grant (No. RG054/09AET).
References
Abrahamson, A.R., Geldart, D., 1980. Behaviour of gas-uidized beds of ne
powders: part II, voidage of the dense phase in bubbling beds. Powder Technol.
26, 4755.
Alizadeh, M., Mostou, N., Pourmahdian, S., Sotudeh-Gharebagh, R., 2004. Modeling
of uidized bed reactor of ethylene polymerization. Chem. Eng. J. 97, 2735.

A. Shamiri et al. / Chemical Engineering Science 66 (2011) 11891199

Aoyagi, M., Kunii, D., 1974. Importance of dispersed solids in bubbles for exothermic
reactions in uidized beds. Chem. Eng. Commun. 1, 191197.
Choi, K.Y., Ray, W.H., 1985. The dynamic behavior of uidized bed reactors for solid
catalyzed gas phase olen polymerization. Chem. Eng. Sci. 40, 22612278.
Cui, H.P., Mostou, N., Chaouki, J., 2000. Characterization of dynamic gassolid
distribution in the uidized beds. Chem. Eng. J. 79, 135143.
Cui, H.P., Mostou, N., Chaouki, J., 2001. Gas and solids between dynamic bubble and
emulsion in gas-uidized beds. Powder Technol. 120, 1220.
Fernandes, F.A.N., Lona, L.M.F., 2001. Heterogeneous modeling for uidized-bed
polymerization reactor. Chem. Eng. Sci. 56, 963969.
Floyd, S., Choi, K.Y., Taylor, T.W., Ray, W.H., 1986. Polymerization of olens through
heterogeneous catalysts. III. Polymer particle modeling with an analysis of
intraparticle heat and mass transfer effects. J. Appl. Polym. Sci. 32, 29352960.
Gilbertson, M.A., Yates, J.G., 1996. The motion of particles near a bubble in a gasuidized bed. J. Fluid Mech. 323, 377385.
Harshe, Y.M., Utikar, R.P., Ranade, V.V., 2004. A computational model for predicting
particle size distribution and performance of uidized bed polypropylene
reactor. Chem. Eng. Sci. 59, 51455156.
Hatzantonis, H., Yiannoulakis, H., Yiagopoulos, A., Kiparissides, C., 2000. Recent
developments in modeling gas-phase catalyzed olen polymerization uidizedbed reactors: the effect of bubble size variation on the reactors performance.
Chem. Eng. Sci. 55, 32373259.
Hilligardt, K., Werther, J., 1986. Local bubble gas hold-up and expansion of gas/solid
uidized beds. Ger. Chem. Eng. 92, 15221.
Ibrehem, A.S., Hussain, M.A., Ghasem, N.M., 2008. Modied mathematical model for
gas phase olen polymerization in uidized-bed catalytic reactor. Chem. Eng. J.
149, 353362.
Jafari, R., Sotudeh-Gharebagh, R., Mostou, N., 2004. Performance of the wideranging models for uidized bed reactors. Adv. Powder Technol. 15, 533548.

1199

Kiashemshaki, A., Mostou, N., Sotudeh-Gharebagh, R., 2006. Two-phase modeling of the gas phase polyethylene uidized bed reactor. Chem. Eng. Sci. 61,
39974006.
Kunni, D.E., Levenspiel, O., 1991. Fluidization Engineering, second ed. ButterworthHeinmann, Boston, MA.
Lucas, A., Arnaldos, J., Casal, J., Puigjaner, L., 1986. Improved equation for the
calculation of minimum uidization velocity. Ind. Eng. Chem. Process. Des. Dev.
25, 426429.
Luo, Z.H., Su, P.L., Shi, D.P., Zheng, Z.W., 2009. Steady-state and dynamic modeling of
commercial bulk polypropylene process of Hypol technology. Chem. Eng. J. 149,
370382.
McAuley, K.B., MacGregor, J.F., Hamilec, A.E., 1990. A kinetic model for industrial
gas-phase ethylene copolymerization. AIChE J. 36, 837850.
McAuley, K.B., Talbot, J.P., Harris, T.J., 1994. A comparison of two phase and
well-mixed models for uidized bed polyethylene reactors. Chem. Eng. Sci.
49, 20352045.
Meier, G.B., Weickert, G., Van Swaaij, W.P.M., 2001. Gas-phase polymerization of
propylene: reaction kinetics and molecular weight distribution. J. Polym. Sci. A:
Polym. Chem. 39, 500513.
Shamiri, A., Hussain, M.A., Farouq, S.M., Mostou, N., 2010. Kinetic modeling of
propylene homopolymerization in a gas-phase uidized-bed reactor. Chem.
Eng. J. 161, 240249.
Xie, T., McAuley, K.B., Hsu, J.C.C., Bacon, D.W., 1994. Gas phase ethylene polymerization: preparation processes, polymer properties and reactor modeling.
Ind. Eng. Chem. Res. 33, 449479.
Zacca, J.J., Debling, J.A., Ray, W.H., 1996. Reactor residence time distribution effects
on the multistage polymerization of olens-I. Basic principles and illustrative
examples, polypropylene. Chem. Eng. Sci. 51, 48594886.

Вам также может понравиться