Sei sulla pagina 1di 12

Article

pubs.acs.org/IECR

Experimental and Modeling Analysis of Propylene Polymerization in


a Pilot-Scale Fluidized Bed Reactor
Ahmad Shamiri,† M. A. Hussain,†,* Farouq Sabri Mjalli,‡ Mohammad Saleh Shafeeyan,†
and Navid Mostouޤ

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

Petroleum and Chemical Engineering Department, Sultan Qaboos University, Muscat 123, Oman
§
Process Design and Simulation Research Center, School of Chemical Engineering, College of Engineering, University of Tehran,
P.O. Box 11155/4563, Tehran, Iran

ABSTRACT: A pilot-scale fluidized bed reactor was built and used to study and gain a better understanding of the propylene
polymerization process at conditions similar to those of industrial fluidized-bed reactors. Propylene polymerization reaction was
carried out continuously in the reactor and dynamic temperatures, and propylene and hydrogen concentrations data were
collected. Simulated dynamic profiles of the two-phase model and the well-mixed model were compared with the actual plant
data. It was found that predictions of the two-phase model were in better agreement with the pilot plant data at typical industrial
operating conditions. This agreement is due to the realistic assumptions of the two-phase model as compared to the well-mixed
model. The maximum deviations between the pilot plant data and the two-phase model prediction for the propylene, hydrogen
concentrations, reactor temperature, and polypropylene production rate were about 3.4 mol %, 0.15 mol %, 2.4 °C, and 0.8 g/s,
respectively.

1. INTRODUCTION framework to represent the transient behavior of a fluidized bed


Fluidized bed reactors are widely used in the polymer industries propylene polymerization reactor. Shi et al.13 developed a
due to their capability in carrying out a variety of chemical three-dimensional computational fluid dynamics (CFD) model
reactions and their high heat and mass transfer rates with to present the gas−solid two-phase flow in fluidized bed
uniform particle mixing.1−6 Therefore, development of accurate polymerization reactors. They used an Eulerian−Eulerian two-
and rigorous models for the kinetics of heterogeneous fluid model which includes the kinetic theory of granular flow.
polymerization, mass and heat transfer, and hydrodynamic In addition, the distribution of solid holdup, the behavior of
characteristics in the fluidized bed is essential for understanding bubbles and the velocity vectors of solid particles in free and
the ongoing phenomena in the reactor and for designing more agitated fluidized bed polymerization reactors were described in
productive reactors. Mass and heat transfer restrictions, details by them. Khare et al.14 proposed a model for steady-
especially in the presence of highly reactive catalyst particles, state and dynamic propylene polymerization in a gas phase
can become considerable at the particle level.7 Model stirred-bed reactors. In their model, they considered a set of
development for investigating industrial polymerization pro- reaction kinetic equations for Ziegler−Natta catalyst as well as
cesses, reactor design, and operation need an integrated thermodynamic parameters that accurately define the polymer
approach comprising of kinetics, reactor hydrodynamics, as properties. However, all these previous studies were theoretical
well as heat and mass transfer processes. in nature and lack the necessary experimental validation.
Different modeling approaches have been proposed in the To the best of our knowledge, only one experimental study
literature to describe the hydrodynamics, mass and heat transfer has been published concerning gas-phase propylene homo-
characteristics, and chemical reactions in fluidized bed polymerization in a fluidized bed reactor in which propylene
polyolefin reactors. Choi and Ray8 proposed a dynamic two- polymerization was carried out in a batch reactor for which a
phase model in which the reactor consists of the emulsion and simplified single-phase dynamic model was proposed by Meier
bubble phases. They considered the polymerization reaction et al.15 They extended a compartment reactor model based on
only in the emulsion phase, assuming that the bubbles are solid- the small-scale reactor. This model is able to determine
free. McAuley et al.9 presented a dynamic model to describe the temperature and concentration profiles inside the reactor as
olefin polymerization in a gas-phase fluidized-bed reactor. They well as molecular-weight distribution of the polymer. They
considered the fluidized bed polyolefin reactor as a well-mixed studied vertical particle mixing and segregation in a pressurized
reactor. A mathematical model consisting of bubble, emulsion, pilot-scale fluidized bed reactor and proposed a simplified
and particulate phases with plug flow behavior was developed dynamic model to compare the measured temperature and
by Fernandez and Lona.10 Hatzantonis et al.11 considered the
reactor to consist of a perfectly mixed emulsion phase and a Received: August 21, 2013
bubble phase divided into several solid-free well-mixed Revised: April 23, 2014
compartments in series. Harshe et al. 12 developed a Accepted: May 1, 2014
comprehensive mathematical model based on a mixing cell Published: May 1, 2014

© 2014 American Chemical Society 8694 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Figure 1. Schematic of pilot scale fluidized bed propylene polymerization reactor.

concentration profiles with the model results. The polymer- were continuously charged into the reactor to produce a broad
ization reactor was considered to be well-mixed and single- distribution of polymeric particles during contact with
phase in which the particulate phase remains at minimum reactants. The Ziegler−Natta catalyst contains titanium as the
fluidizing conditions. In addition, the kinetic model involved active metal and polymerization reaction occurs on the catalyst
only the propagation and deactivation of the catalyst reactions. active sites. As the reaction proceeds, catalyst portions are
They used metallocene rac-Me2Si[Ind]2ZrCl2 as a catalyst and scattered and particles grow into the final polypropylene
triisobutylaluminum (TIBA) as a cocatalyst to increase the product.16 Reactants (i.e., propylene and hydrogen) and
catalyst activity. nitrogen act as the fluidization gas as well as the heat transfer
The main focus of this work is to implement a medium. Solid particles are separated from the unreacted gases
comprehensive model for a continuous gas-phase propylene in the disengaging zone. The recycled gas is mixed with the
polymerization with heterogeneous Ziegler−Natta catalyst in a makeup gas after heat removal and is fed back to the reactor
pilot scale fluidized-bed catalytic reactor (FBR) and to validate inlet. The polymer particles were continuously withdrawn from
the two-phase model along with the comprehensive two-site a position above the distributor.
kinetic scheme through a real-time study. A unique and novel The detailed design summary of the pilot scale fluidized bed
fluidized bed pilot plant, resembling an industrial unit, was reactor is given in Table 1. The reactor (R 01) includes a
designed and built for this purpose. The pilot-scale fluidized fluidized bed and a disengagement zone. The inner diameter of
bed reactor was simulated using a two-phase model and the the reactor is 10 cm, and the height of the fluidized bed zone is
results were compared with the experimental data in terms of 150 cm. Height and diameter of the disengagement zone are
the reactor temperature, concentration profiles of propylene both 25 cm. Catalyst particles were injected into the bed at a
and hydrogen and polypropylene production rate. In addition, point 9 cm above the gas distributor. Product samples were
the experimental and calculated temperatures were compared withdrawn from three different locations, i.e., 16, 26, and 40 cm
for different polypropylene grades. above the distributor plate. The produced polymer can be
discharged semicontinuously by opening a valve connected to
2. EXPERIMENTAL STUDIES the product vessel at a point 5 cm above the gas distributor.
2.1. Experimental Setup. A pilot-scale fluidized bed The gas distributor was a stainless steel perforated plate of 100-
reactor was constructed in the Pilot Plant Laboratory at the mesh size. Temperature in the reactor was measured at 6
University of Malaya. The main objective for employing such a different positions, starting at 16 cm above the distributor.
setup was to study the catalytic polymerization of olefins at high Radial temperature gradients in the reactor can be neglected
pressure and operating conditions close to the industrial units. due to the agitation produced by the up-flowing gas and small
Figures 1−3 illustrate the schematic, picture, and detailed diameter of the reactor. Temperature uniformity is a well-
drawing of the pilot scale fluidized bed reactor, respectively. known advantage of fluidized bed reactors. In fact, the solid
Ziegler−Natta catalyst and triethyl aluminum as the cocatalyst circulation rate inside the bed is high enough that it can
8695 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Figure 2. Fluidized bed propylene polymerization reactor detailed drawing.

disperse the heat generated by the reaction instantaneously. temperature entering the reactor at start up to enable the
Therefore, the radial temperatures of the reactor were not taken reactants to reach the required reaction temperature. An air-
but only temperatures at different levels in the reactor were driven piston gas booster compressor (P101) was used to
measured. One cyclone and four filters were placed to separate compensate the pressure drop through the loop. A buffer vessel,
entrained solid particles from the exiting gas. installed downstream the compressor, was used to damp
The gas leaving the reactor was cooled by a shell and tube pressure fluctuations. Flow of gas through the reactor was
heat exchanger (E-401). A control valve (TCV101) was used to controlled by a control valve (FCV301) and was measured by a
manipulate the reactor gas inlet temperature. Heat was flow meter (FT301) that is located just before the reactor. An
removed from the reactor by cooling the gaseous recycle important requirement of a fluidized bed reactor is that the
stream outside the reactor. A heater was used to control the gas velocity of the recycle stream must be sufficient to keep the bed
8696 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Figure 3. Picture of the pilot scale fluidized bed polypropylene reactor in this study.

Table 1. Detail Design Parameters of the Pilot Scale content of the gas was measured continuously by the online
Fluidized Bed Reactor hygrometer.
A gas sample flow from the reactor was analyzed
parameters value
continuously by a Perkin-Elmer Clarus 580 gas chromatograph
catalyst type Ziegler−Natta (GC) to measure and regulate the gas composition inside the
catalyst density ρc 2370 kg/m3
reactor. Due to agitation produced by the up-flowing gas as well
average catalyst particle diameter dp 80 μm
as the small height and diameter of the pilot scale reactor,
maximum catalyst feed rate ṁ cat 3.02 g/h
which ensure uniformity of reactants in the bed ,and also
minimum fluidization velocity Umf 0.1 ms−1
because the presence of solid particles throughout the bed
maximum bubble size dBv max 0.0008 m
would choke the gas line, only one gas sample was taken from
Reactor Design
the top of the reactor and analyzed continuously by an online
reaction zone inner Di,RZ 0.1016 m
diameter Perkin-Elmer Clarus 580 gas chromatograph (GC) to measure
cross- ARZ 0.00785 m2 and regulate the gas composition inside the reactor. The three-
sectional column model Arnel 1117PPC provided a guaranteed analysis
area
of hydrogen, nitrogen, and propylene in approximately 8.5 min
height HRZ 1.5 m
using a flame ionization detector (FID) and two thermal
volume VRZ 0.011775 m3
disengagement inner Di,DZ 0.25 m
conductivity detectors (TCD/TCD). Three columns were
zone diameter installed in the GC. Column A used nitrogen as the carrier gas,
cross- ADZ 0.0490625 m2 while columns B and C used hydrogen. Column A performed a
sectional full-range hydrogen analysis on the gas samples using a TCD
area
with nitrogen carrier gas. Oxygen, nitrogen, carbon monoxide
height HDZ 0.25 m
and carbon dioxide were analyzed by the fixed gas column (B)
volume VDZ 0.0097 m3
reactor volume Vreactor 0.0215 m3
using a TCD with hydrogen carrier gas. The hydrocarbon
maximum pressure drop across Δpbed 0.0325 atm
column (C) was used to analyze propylene using a FID with
fluidized bed hydrogen carrier gas. Columns A and B were linked
Distributor Plate Design electronically and eluted to the data handling system as one
distributor plate type stainless steel perforated plate response and appear as one column. Therefore, only two
100-mesh size columns of data handling were required. All three columns
could be run simultaneously or independently.
in a fluidized state. This was ensured by monitoring and Highly pure raw materials are required for catalytic olefin
stabilizing the flow of the gaseous stream. The Easidew online polymerization to avoid poisoning the catalyst. Propylene,
hygrometer (MICHELL, EA2-MON-100) was used to measure hydrogen, and nitrogen gases were purified in separate
the moisture content of the gas within the reactor. In this purification systems (Entegris GateKeeper gas purifiers) to
experiment, two gas samples for humidity measurement were remove traces of oxygen, water vapor, and carbon monoxide.
taken from inlet and outlet of the reactor and the moisture Three mass flow meters (Brooks) were used in the fresh feed
8697 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

stream to measure the flow of propylene (FT302), hydrogen injection. After eliminating the catalyst poisons from the system
(FT304), and nitrogen (FT303). in the beginning of the experiment, a low amount of catalyst
For injecting a small amount of Ziegler−Natta catalyst into was injected into the reactor. Inlet temperature and fluidization
the reactor, a continuous feeding system was developed. A ball velocity were kept constant during the experiments. However,
valve, installed under the catalyst container, was used as the total and partial pressures were changing because injection of
dosing device. A small flow of pressurized high purity nitrogen the fresh gas was not done continuously. Temperature and
was permanently kept in the whole injection system to feed the concentration of gas components inside the reactor were
catalyst into the reactor through the ball valve and stop gas monitored continuously with the instruments mentioned in
leakage into the injection system. A kill gas injection system, section 2.1.
composed of CO2 gas, was installed for terminating the This pilot plant was equipped with a local/remote control
polymerization reaction at emergency conditions, such as system. Monitoring and data acquisition were performed
power failure, to stop or reduce the polymerization reaction, or through a PC-based software system which included various
to maintain the reactor bed temperature below the sintering capabilities such as trending, monitoring, recording and
temperature of the polymer. It was injected manually into the printing of process data as well as process control variables.
system in case these emergency conditions took place. The full SCADA is illustrated in Figure 4. Control of the heat
The system was designed to work at a maximum pressure of exchanger unit was also carried out through the use of this
30 bar and temperature of 100 °C. The reactor differential software.
pressure indicator (DPT211), thermocouples, and pressure
indicators located inside the reactor in different places were 3. MATHEMATICAL MODELING
utilized to monitor temperatures and pressures of the system. In the present study, the kinetic model of propylene
In the event of a measured value indicating a dangerous homopolymerization over Ziegler−Natta catalyst was combined
situation, the experiment was stopped. A rupture disc and a with the dynamic single-phase and two-phase hydrodynamic
pressure relief valve were installed on top of the reactor and set models. The strength of the heterogeneous Ziegler−Natta
to a relief pressure of 30 and 32 bar, respectively, to protect the catalysts to generate polymers with broad molecular weight
reactor from overpressure. distributions has long being identified. The single type of active
2.2. Experimental Procedure. The reactor was charged center kinetic model is not adequate to define the kinetic
with 2 kg of a polypropylene powder with an average particle behavior of propylene homopolymerization. Therefore, two
size of 500 μm before the startup. The system must be free of types of active sites were considered for the catalyst in the
contaminants such as moisture, oxygen, carbon monoxide, or present work. The kinetic scheme, comprising of a series of
carbon dioxide before polymerization takes place in the reactor. elementary reactions, and the rate parameters, used for the two
This was achieved by purging the system with high purity models, are listed in Table 3. The reaction rate constants used
liquefied nitrogen. The reactor was always exposed to air and in this work are given in Table 4.
moisture and consequently needed to be purged with high The model consists of mass balances on the species
purity nitrogen until the moisture level dropped to less than10 presented in the reactor which are written as a series of
ppmv prior to charging the Ziegler−Natta catalyst. Triethyla- algebraic and differential equations. Consumption rate of each
luminum (TEAL) was used to further scavenge impurities such component (monomer and hydrogen) and the polymer
as moisture and oxygen from the reactor. TEAL removes production rate were modeled using population balance and
catalyst poisons from the reactor and thereby maximizing the the method of moments. The corresponding moment
catalyst activity and productivity. TEAL diluted in hexane was equations are given in Table 5. Details of the kinetic model
then injected to the reactor until the moisture level further are reported by Shamiri et al.17
dropped to less than 1 ppmv. Based on the assumption that the only significant
The setup ran with gas inlet temperature of 60 °C, measured consumption of propylene monomer is by the propagation
by the temperature transmitter TT101 (Figure 1). When the reaction, where the consumption of hydrogen gas occurs via
system was at the required operating conditions (Table 2), the transfer to hydrogen, the following equations for the
system was regarded as being ready to start the catalyst consumption rate of monomer and hydrogen can be stated:
For monomer:
Table 2. Operating and Gas Composition Conditions and NS
Physical Properties Used in the Experimental Study Ri = ∑ [Mi]Y (0, j)k p(j), i = 1
j=1 (1)
operating conditions physical properties
Tin (K)=333.15 dp (m) = 500 × 10−6 For hydrogen:
Tref (K) = 342.15 μ (Pa·s) = 1.14 × 10−4 NS
Tw(K) = 350.15 ρg (kg/m3) = 23.45 Ri = ∑ [Mi]Y (0, j)k fh(j), i = 2
D (m) = 0.1016 ρpol (kg/m3) = 910 j=1 (2)
H (m) = 1.5 εmf = 0.45
V (m3) = 0.0215 The total polymer production rate for each phase can be
P (bar) = 22 calculated from
propylene concentration (mol %) = 59.18
2
hydrogen concentration (mol %) = 6.46
nitrogen concentration (mol %) = 33.7
Rp = ∑ MwiR i
i=1 (3)
propane concentration (mol %) = 0.47
C6+ concentration (mol %) = 0.19 where Ri is the instantaneous rate of reaction.
8698 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Figure 4. Software-based control and data acquisition system (SCADA) for the pilot scale fluidized bed polypropylene reactor.

Table 3. Elementary Chemical Reactions of Propylene Table 4. Rate Parameters Used for the Two-Phase and Well-
Homo-polymerization Mixed Models (Obtained at 69 °C)
reaction description rate
reaction constant unit site type 1 site type 2
k f (j) formation of active sites
N *(j) + cocatalyst ⎯⎯⎯→ N (0, j) formation kf(j) s−1 1 1
k i(j) initiation of active sites initiation ki(j) m3 kmol−1 22.9 54.9
N (0, j) + M ⎯⎯⎯→ N (1, j) s−1
k p(j) propagation kh(j) m3 kmol−1 0.1 0.1
N (r , j) + M ⎯⎯⎯⎯→ N (r + 1, j) s−1
k fm(j) chain transfer to monomer kh m3 kmol−1 20 20
N (r , j) + M ⎯⎯⎯⎯⎯→ N (1, j) + Q (r , j) s−1
k fh(j) transfer to hydrogen propagation kp(j) m3 kmol−1 208.6 22.9
N (r , j) + H 2 ⎯⎯⎯⎯→ NH(0, j) + Q (r , j) s−1
k h(j) activation kcal kmol−1 7200 7200
NH(0, j) + M ⎯⎯⎯⎯→ N (1, j) energy
kh r(j) transfer kfm(j) m3 kmol−1 0.046 0.253
NH(0, j) + AlEt3 ⎯⎯⎯⎯⎯→ N (1, j) s−1
k fr(j) transfer to cocatalyst kfh(j) m3 kmol−1 7.54 7.54
N (r , j) + AlEt3 ⎯⎯⎯⎯→ N (1, j) + Q (r , j) s−1
k fs(j) spontaneous transfer kfr(j) m3 kmol−1 0.024 0.12
N (r , j) ⎯⎯⎯⎯→ NH(0, j) + Q (r , j) s−1
kds(j) deactivation reactions kfs(j) m3 kmol−1 0.0001 0.0001
N (r , j) ⎯⎯⎯⎯→ Nd(j) + Q (r , j) s−1
−1
kds(j) deactivation kds(j) s 0.00034 0.00034
N (0, j) ⎯⎯⎯⎯→ Nd(j)
kds(j)
NH(0, j) ⎯⎯⎯⎯→ Nd(j)
The number-average molecular weight and the weight-
average molecular weight, M̅ n and M̅ w, can be determined using
the method of moments as follows:
8699 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Table 5. Moment correlations Table 6. Hydrodynamic Correlations and Equations Used


for the Two-Phase Model
dY (0, j)
= [M]{k i(j)N (0, j) + k h(j)NH(0, j)} + NH(0, j)k hr(j)[AlEt3]
dt parameter formula ref

⎪ Rv ⎫
⎪ Minimum fluidization Remf = [(29.5)2 + 0.357Ar ]1/2 − 29.5 25
− Y (0, j)⎨

k fh( j)[H 2 ] + k fs(j) + k ds(j) + ⎬

velocity
⎩ Vp⎭
Bubble velocity Ub = U0 − Ue + ubr 26
dY (1, j) Bubble rise velocity 1/2 26
= [M]k i(j)N (0, J ) + NH(0, J ){k h(j)[M] + k hr(j)[AlEt3]} ubr = 0.711(gdb)
dt
Emulsion velocity U0 − δUb 27
+ Y (0, j){k fm(j)[M] + k fr(j)[AlEt3]} + [M]k p(j)Y (0, j) Ue =
1−δ


− Y (1, j)⎨ k (j)[M] + k fr(j)[AlEt3] + k fh(j)[H 2] + k fs(j) Bubble diameter db = db0[1 + 27(U0 − Ue)]1/3 (1 + 6.84H ) 28
⎪ fm

db0 = 0.0085 (for Geldart B)
R ⎫ mass transfer ⎛ 1 26
1 ⎞
⎪ −1
+ kds(j) + v ⎬
Vp ⎪

coefficient Kbe = ⎜ + ⎟
⎝ Kbc Kce ⎠
dY (2, j) ⎛ D 1/2g 1/4 ⎞
= [M]k i(j)N (0, j) + NH(0, j){k h(j)[M] + k hr(j)[AlEt3]} ⎛U ⎞
Kbc = 4.5⎜ e ⎟ + 5.85⎜⎜ ⎟
g
dt
5/4 ⎟
⎝ db ⎠ ⎝ db ⎠
+ Y (0, j){k fm(j)[M] + k fr(j)[AlEt3]}
+ [M]k p(j){2Y (1, j) + Y (0, j)} − Y (2, j) ⎛ Dg εeubr ⎞
Kce = 6.77⎜ ⎟


⎝ db ⎠
⎨ k (j)[M] + k fr(j)[AlEt3] + k fh(j)[H 2] + k fs(j) + kds(j)
⎪ fm heat transfer ⎛ 1 26
1 ⎞
−1

coefficient Hbe = ⎜ + ⎟
Rv ⎫

⎝ Hbc Hce ⎠
+ ⎬
Vp ⎪
⎭ ⎛ Ueρg Cpg ⎞ (kgρg Cpg)1/2 g 1/4
Hbc = 4.5⎜⎜ ⎟⎟ + 5.85
dX(n , j) ⎝ db ⎠ db5/4
= Y (n , j){k fm(j)[M] + k fr(j)[AlEt3] + k fh(j)[H 2] + k fs(j)
dt
R ⎛ ε u ⎞1/2
+ kds(j)} − X(n , j) v Hce = 6.77(ρg Cpgkg)1/2 ⎜ e 3br ⎟
Vp ⎝ db ⎠
n = 0, 1, 2 bubble phase fraction ⎡ ⎛ U − Umf ⎞⎤ 29
δ = 0.534⎢1 − exp⎜− 0 ⎟⎥
⎣ ⎝ 0.413 ⎠⎦
⎛ ∑NS (X(1, j) + Y (1, j)) ⎞ emulsion phase ⎛ U − Umf ⎞ 29
M̅ n = Mw⎜ NS ⎟
j=1
porosity εe = εmf + 0.2 − 0.059 exp⎜− 0 ⎟
⎜ ∑ (X(0, j) + Y (0, j)) ⎟ ⎝ 0.429 ⎠
⎝ j=1 ⎠ (4) bubble phase porosity ⎛ U − Umf ⎞ 29
εb = 1 − 0.146 exp⎜− 0 ⎟
⎝ 4.439 ⎠
⎛ ∑NS (X(2, j) + Y (2, j)) ⎞ volume of polymer VPe = AH(1 − εe)(1 − δ) 18
M̅ w = Mw⎜ NS ⎟
j=1 phase in the
⎜ ∑ (X(1, j) + Y (1, j)) ⎟ emulsion phase
⎝ j=1 ⎠ (5) volume of polymer VPb = AH(1 − εb)δ 18
phase in the bubble
phase
The polydispersity index (PDI) is defined by the ratio of
volume of the Ve = AH(1 − δ) 18
weight-average to number-average molecular weights: emulsion phase
volume of the bubble Vb = AδH 18
M̅ phase
PDI = w
M̅ n (6)
[Mi]b,(in) UbAb − [Mi]b UbAb − R vεb[Mi]b
3.1. Dynamic Two-Phase Model. Based on the dynamic Ab
two-phase model, polymerization reactions were considered to
− Kbe([Mi]b − [Mi]e )Vb − (1 − εb)
VPFR
∫ R i dz
b

occur in the bubble phase and the emulsion phase. The heat d
= (Vbεb[Mi]b )
loss through the reactor wall to the surroundings due to the dt (7)
absence of insulation was also considered in the model. For emulsion:
Poisoning reactions were neglected because it was assumed that
[Mi]e,(in) UeAe − [Mi]e UeAe − R vεe[Mi]e
the feed gas was free of impurities. Elutriation of solid particles
from the reactor was also neglected. The required correlations ⎛ δ ⎞
+ Kbe([Mi]b − [Mi]e )Ve⎜ ⎟ − (1 − ε )R
⎝1 − δ ⎠ e ie
for the dynamic two phase model are given in Table 6.
Based on the above assumptions, the following dynamic d
= (Veεe[Mi]e )
dt (8)
material balances were derived:18−23
For bubbles: Moreover, the energy balances can be written as
8700 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

For bubbles: • Homogeneous composition and temperature throughout


m the bed were considered.
UbAb(Tb,(in) − Tref ) ∑ [Mi]b,(in) Cpi − UbAb(Tb − Tr ef ) • The feed gas is free of impurities, thus, poisoning
i=1 reactions were neglected.
m
• No elutriation of solid particles occurs from the reactor.
∑ [Mi]b Cpi
i=1
Based on the above assumptions, the following dynamic
m
material and energy balance equations can be derived:19
The molar balance is given by
−R v(Tb − Tref )(∑ εbCpi[Mi]b + (1 − εb)ρpol Cp,pol)
i=1 d[Mi]
A ΔH (Vεmf ) = U0A([Mi]in − [Mi]) − R vεmf [Mi]
+ (1 − εb) b R
VPFR
∫ R pbdz dt
− (1 − εmf )R i (15)
m
d
+Hbe(Te − Tb)Vb − Vbεb(Tb − Tr ef ) ∑ Cpi ([Mi]b ) The energy balance is written as
i=1
d t m
dT
m
d [∑ [Mi]CpiVεmf + V (1 − εmf )ρpol Cp,pol]
=(Vb(εb ∑ Cpi[Mi]b + (1 − εb)ρpol Cp,pol)) (Tb − Tref ) i=1 dt
dt m
i=1
= U0A ∑ [Mi]Cpi(Tin − Tref )
(9)
i=1
For emulsion: m
m −U0A ∑ [Mi]Cpi(T − Tr ef )
UeAe(Te,(in) − Tref ) ∑ [Mi]e,(in) Cpi − UeAe(Te − Tref ) i=1
m
i=1
m − R v[∑ [Mi]Cpiεmf + (1 − εmf )ρpol Cp,pol](T − Tref )
∑ [Mi]e Cpi i=1
i=1
m
+(1 − εmf )ΔHR R p − πDHhw (Te − Tw )
−R v(Te − Tref )(∑ εeCpi[Mi]e + (1 − εe)ρpol Cp,pol) (16)
i=1 The initial conditions are as follows:
− (1 − εe)R peΔHR
[Mi]t = o = [Mi]in (17)
⎛ δ ⎞
−HbeVe⎜ ⎟(T − T ) − V ε (T − T ) T (t = 0) = Tin (18)
⎝1 − δ ⎠ e b e e e ref
m
∑ Cpi d ([Mi]e ) − πDHhw(1 − δ)(Te − Tw)= 4. RESULTS AND DISCUSSION
i=1 dt
m
In the present study, the two-phase model along with the
d comprehensive two-site kinetic scheme, discussed in section 3
(Ve(εe ∑ Cpi([Mi]e ) + (1 − εe)ρpol Cp,pol)) (Te − Tref )
dt was validated with the actual pilot plant data. The operating
i=1
conditions and gas composition for producing polypropylene in
(10)
this work are listed in Table 2.
The initial conditions are given by A comparative study with the actual dynamic pilot plant data
[Mi]b, t = 0 = [Mi]in in terms of reactor temperature, propylene gas concentration
(11)
and polypropylene production rate was conducted to validate
[Mi]e, t = 0 = [Mi]in the two-phase model for propylene homopolymerization in the
(12)
gas phase fluidized bed reactor. Comparison between the
Tb(t = 0) = Tin (13) results of two-phase and conventional well-mixed models with
the pilot plant data in terms of the propylene concentration
Te(t = 0) = Tin (14) inside the reactor is illustrated in Figure 5. As can be seen in
Figure 5, predicted data of the two-phase model are in good
Solving the above set of equations, changes in concentration agreement with the experimental data, especially at longer
and temperature inside the reactor can be obtained. elapsed times. Based on considering the presence of solids in
3.2. Well-Mixed Model. For the sake of comparison, the bubbles and the emulsion phase being at conditions beyond
formulation of the well-mixed model is discussed briefly in this the minimum fluidization the two-phase model provides more
section. In the simplified well-mixed model proposed by realistic results. However, the two-phase model under-predicts
McAuley et al.,9,24 the following simplifying assumptions were the experimental data at shorter elapsed time. This is mainly
made: due to very high rate of mass and heat transfer between the
• Due to the absence of insulation, the heat loss through bubbles and the emulsion phase at the very beginning of the
the reactor wall to the surroundings was considered. The process startup at which the concentration difference between
polymerization reactor is considered to be a well-mixed the two phases is maximum. This situation during the initial
reactor due to high mass and heat transfer rates between stages of fluidization makes the hydrodynamics of the reactor
bubble and emulsion phases.4,19,21 approach the well-mixed condition. However, this mechanism
• The emulsion phase remains at minimum fluidization. is definitely unrealistic later in the reaction as more bubbles are
8701 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Figure 7. Comparison between experimental data and calculated


Figure 5. Comparison between experimental data and calculated temperatures.
propylene concentration.

formed and rates of mass and heat transfer are reduced. The reactor occurs between the temperature transmitter TT102 at
maximum difference between the experimental data and the bottom and TT107 at the top of the reactor at about 7 °C.
predictions of the two-phase and well-mixed models are The maximum temperature observed in this experiment was
about 3.4 and 5 mol %, respectively. These differences are about 87.5 °C sensed by TT102. Since it was the closest sensor
further due to the effect of inert gas on the hydrodynamic to the catalyst injection point, the presence of a higher amount
behavior of the fluidized bed reactor. of catalyst causes a higher reaction rate and consequently higher
Concentration profiles of inert gases versus time are shown temperature. Nevertheless, the average bed temperature was
in Figure 6. When the polymerization reaction starts, propylene used for comparison with predictions of the theoretical models.
The figure also shows good agreement between the calculated
and experimental average bed temperature profiles. As time
evolves, predictions of the two-phase model become closer to
the average bed temperature than does the well-mixed model. It
should be noted that considering more realistic assumptions of
the two-phase model, results obtained by this model are closer
to real values compared to the well mixed model. At the start of
reaction, the well-mixed model shows a close dynamic response
to the average bed temperature. However, upon reaching a
temperature of about 81 °C, this model begins to deviate
considerably with time due the fact that it was assumed that the
bubbles are solid-free and the emulsion stays at minimum
fluidization, interchanging heat and mass with the bubble phase
at uniform rates at all operating conditions. It is obvious that
due to these simplified assumptions, it is not capable to
accurately describe the dynamics of gas−solid distribution on
the chemical reaction and rate of heat and mass transfer in the
Figure 6. Concentration profile of inert gas versus time inside of fluidized bed reactors by the well-mixed model. The maximum
reactor. difference between the pilot plant temperature data and that of
the two-phase and well-mixed models were about 2.4 and 3.5
concentration is reduced due to its conversion to polypropy- °C, respectively.
lene. By this decrease in the propylene concentration during the A comparison between the results of two-phase model with
polymerization reaction, an expected increase in the concen- the pilot plant data in terms of the polypropylene production
tration of inert components (nitrogen, propane, ethane, and rate is illustrated in Figure 8. As shown in Figure 8, the two-
C6+) in the system was observed. This is similar to what phase model predictions are in good agreement with the
happens in an industrial scale polymerization reactor. experimental data. However, the two-phase model overpredicts
Experimental reactor temperatures obtained at different the experimental values within the initial reaction time range of
levels of the fluidized bed and those obtained based on the 0−2000 s. This is mainly due to very high rate of
two-phase and well-mixed models are presented in Figure 7, polymerization reaction as well as mass and heat transfer at
demonstrates that the temperature starts rising rapidly after the beginning of the process startup. The maximum difference
injection of the catalyst and reaches its steady value after about between the experimental data and predictions of the two-
2 h. As expected in this system, the reactor temperature phase model is about 0.8 g/s.
increases as the polymerization reaction starts due to progress Hydrogen gas is used to regulate the molecular weight and its
of the exothermic reaction of propylene polymerization. During distribution as well as grade transition of product in a pilot-scale
a single experiment, the largest temperature gradient inside the propylene polymerization plant. Hydrogen concentration
8702 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Figure 10. Evolution of number and weight-average molecular weights


and PDI with the time in the reactor (propylene = 60 mol %,
Figure 8. Comparison between experimental data and calculated hydrogen = 0.5 mol %, and nitrogen = 39.5 mol %).
production rate using the two-phase model.
indicates that the weight-average molecular weight, number-
profile versus time is shown in Figure 9. It is obvious that the average molecular weight and PDI of the polymer increase
hydrogen concentration decreases at the start of the polymer- rapidly at the beginning of the polymerization and eventually
stabilize within less than half an hour of the process startup.
The final value of the number-average molecular weight,
weight-average molecular weight and PDI, under the conditions
of this modeling, are close to 28900, 145000, and 5,
respectively. This indicates that the produced polypropylene
attains a broad molecular weight distribution.
A comparison between actual reactor temperatures and those
calculated based on the model presented in this work is shown
in Figure 11. The gas composition conditions for producing

Figure 9. Concentration profile of hydrogen versus time inside of


reactor.

ization reaction due to its consumption via the mechanism of


transfer to hydrogen reaction. The simulation results based on
the two-phase model were found to be in a close agreement
with the experimental data in terms of hydrogen concentration
profile. As discussed before, this is mainly due to the sensible
assumption and correlation coefficient considered in the two-
Figure 11. Comparison between experimental data and calculated
phase model. The maximum difference between the pilot plant temperatures using the two-phase model at different polypropylene
data and the two phase model prediction for the hydrogen grades.
concentration was as little as 0.15 mol %. The over prediction
of the two-phase model can be attributed to the effect of inert
gas on the process hydrodynamics which is neglected in this
work. different polypropylene grades employed in this study are listed
The polymer molecular weight and its distribution affect in Table 7. As can be seen in Figure 11, there is a good
most of the essential properties of the polymer, such as tensile agreement between the calculated and pilot plant temperatures.
strength, thermal stability, stiffness, hardness, softening point The maximum difference between the industrial data and the
and impact strength significantly. While the PDI represents the model prediction is 1.5 K for polypropylene grade B type. The
distribution of individual molecular masses in a batch of differences between predicted and industrial temperatures are
polymers and is used as a measure of the width of the molecular due to the effect of inert gas, particle size distribution, and solid
weight distribution. Figure 10 shows the polypropylene entrainments on the hydrodynamic behavior of the fluidized
molecular weight distribution as a function of time. This figure bed reactor.
8703 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Table 7. Gas Composition Conditions for Producing Different Grades of Polypropylene


gas unit A B C D E F
propylene concn mol % 64.945 59.18 54.9 49.95 44.96 40.04
hydrogen concn mol % 4.85 6.46 4.83 4.1 3.9 3.8
nitrogen concn mol % 29.95 33.71 39.96 45.71 50.9 55.93

5. CONCLUSIONS Hce cloud to emulsion heat transfer coefficient (W/m3·K)


A two-phase model was used for simulating the propylene spontaneous deactivation rate constant for a site of
polymerization in a pilot-scale fluidized bed reactor. The model kds (j) type j
was validated with data generated in a pilot plant designed to kf (j) formation rate constant for a site of type j
study olefin polymerization reactions. To achieve a better kfh (j) transfer rate constant for a site of type j with terminal
insight on the reactor performance, the hydrodynamics of the monomer M reacting with hydrogen
fluidized bed reactor of propylene polymerization, based on the kfm (j) transfer rate constant for a site of type j with terminal
dynamic two-phase flow structure, was coupled with the kinetic monomer M reacting with monomer M
model. Moreover, the polymerization reaction was considered kfr (j) transfer rate constant for a site of type j with terminal
to take place in both the bubble phase and the emulsion phases monomer M reacting with AlEt3
in the two-phase model. kfs (j) spontaneous transfer rate constant for a site of type j
Comparative studies were performed using two-phase and with terminal monomer M
well-mixed models versus the pilot plant data. A good kh (j) rate constant for reinitiating of a site of type j by
agreement was observed between the prediction of the two- monomer M
phase model and that of the pilot plant. The maximum kh (j) rate constant for reinitiating of a site of type j by
differences between the pilot plant data and the two phase cocatalyst
model prediction for the propylene and hydrogen concen- ki (j) rate constant for initiation of a site of type j by
trations, reactor temperature and polypropylene production monomer M
rate were about 3.4 mol %, 0.15 mol %, 2.4 °C, and 0.8 g/s, kp (j) propagation rate constant for a site of type j with
respectively. The well-mixed model was found to underestimate terminal monomer M reacting with monomer M
the propylene concentration and emulsion phase temperature. Kbc bubble to cloud mass transfer coefficient (s−1)
Furthermore, a comparison between actual reactor temper- Kbe bubble to emulsion mass transfer coefficient (s−1)
atures and those calculated based on the two-phase model using Kce cloud to emulsion mass transfer coefficient (s−1)
different gas composition or polypropylene grades was carried M monomer (propylene)
out and good agreement was also achieved with an average [Mi] concentration of component i in the reactor (kmol/
error of 0.22%. m3)


[Mi]in concentration of component iin the inlet gaseous
AUTHOR INFORMATION stream
Mw monomer molecular weight (kg/kmol)
Corresponding Author N(0, j) uninitiated site of type j produced by formation
*E-mail: mohd_azlan@um.edu.my; a.shamiri@um.edu.my. Tel: reaction
+60-379677657. Fax: +60-379675319. N(r,j) living polymer molecule of length r, growing at an
Notes active site of type j, with terminal monomer M
The authors declare no competing financial interest. NH(0,j) uninitiated site of type j produced by transfer to

■ ACKNOWLEDGMENTS
The authors are grateful to the University of Malaya and the
r
Ri
hydrogen reaction
number of units in polymer chain
instantaneous rate of reaction for monomer i (kmol/
Ministry of Higher Education of Malaysia (MOHE) for s)
supporting this research project via the research grant UM.C/ Remf Reynolds number of particles at minimum fluidization
625/1/HIR/MOHE/ENG/25 which made possible the condition
publication of this paper. Rp production rate (kg/s)


Rpb bubble phase production rate (kg/s)
NOMENCLATURE Rpe emulsion phase production rate (kg/s)
Rv volumetric polymer phase outflow rate from the
A cross-sectional area of the reactor (m2)
reactor (m3/s)
AlEt3 triethyl aluminum
Tin temperature of the inlet gaseous stream (K)
Ar Archimedes number
Tw wall temperature (K)
Cpi specific heat capacity of component i (J/kg·K)
U0 superficial gas velocity (m/s)
Cp,pol specific heat capacity of solid product (J/kg·K)
Umf minimum fluidization velocity (m/s)
db bubble diameter (m)
V reactor volume (m3)
dp polymer particle diameter (m)
Vp volume of polymer phase in the reactor (m3)
D bed diameter (m) VPFR volume of PFR (m3)
hw wall heat transfer coefficient (W/m2·K) X(n,j) nth moment of chain length distribution for dead
H bed height (m) polymer produced at a site of type j
Hbe bubble to emulsion heat transfer coefficient (W/m3· Y(n,j) nth moment of chain length distribution for living
K) polymer produced at a site of type j
Hbc bubble to cloud heat transfer coefficient (W/m3·K)
8704 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705
Industrial & Engineering Chemistry Research Article

Greekletters Homopolymerization in Fluidized Bed Reactors. Chem. Eng. Sci. 2011,


ΔHR heat of reaction (J/kg) 66 (6), 1189−1199.
εmf void fraction of the bed at minimum fluidization (19) Shamiri, A.; Azlan Hussain, M.; Sabri Mjalli, F.; Mostoufi, N.
Improved Single Phase Modeling of Propylene Polymerization in a
μ gas viscosity (Pa.s)
Fluidized Bed Reactor. Comput. Chem. Eng. 2012, 36, 35−47.
ρg gas density (kg/m3) (20) Ho, Y. K.; Shamiri, A.; Mjalli, F. S.; Hussain, M. A. Control of
ρpol polymer density (kg/m3) Industrial Gas Phase Propylene Polymerization in Fluidized Bed

■ REFERENCES
(1) Farhangiyan Kashani, A.; Abedini, H.; Kalaee, M. R. Simulation of
Reactors. J. Process Control 2012, 22 (6), 947−958.
(21) Shamiri, A.; Azlan Hussain, M.; Sabri Mjalli, F.; Mostoufi, N.
Comparative Simulation Study of Gas-Phase Propylene Polymer-
ization in Fluidized Bed Reactors Using Aspen Polymers and Two
an Industrial Linear Low Density Polyethylene Plant. Chem. Prod. Phase Models. Chem. Ind. Chem. Eng. Q. 2013, 19 (1), 13−24.
Process Model. 2011, 6 (1), No. 34. (22) Shamiri, A.; Azlan Hussain, M.; Sabri Mjalli, F.; Mostoufi, N.;
(2) Ibrehem, A. S. Hybrid Modeling of Modified Mathematical Hajimolana, S. A. Dynamics and Predictive Control of Gas Phase
Model for Gas Phase Olefin in Fluidized Bed Catalyst Reactors Using Propylene Polymerization in Fluidized Bed Reactors. Chin. J. Chem.
Artificial Neural Networks. Chem. Prod. Process Model. 2009, 4 (1), Eng. 2013, 21 (9), 1015−1029.
No. 46. (23) Shamiri, A.; Azlan Hussain, M.; Sabri Mjalli, F. Two Phase
(3) Ghasem, N.; Ang, W.; Hussain, M. Dynamic Model for Dynamic Model for Gas Phase Propylene Copolymerization in
Polyethylene Production in a Multizone Circulating Reactor. Chem. Fluidized Bed Reactor. Defect Diffusion Forum 2011, 312−315,
Prod. Process Model. 2008, 3 (1), No. 1. 1079−1084.
(4) Alizadeh, M.; Mostoufi, N.; Pourmahdian, S.; Sotudeh- (24) McAuley, K. B.; Talbot, J. P.; Harris, T. J. A Comparison of
Gharebagh, R. Modeling of Fluidized Bed Reactor of Ethylene Two-Phase and Well-Mixed Models for Fluidized-Bed Polyethylene
Polymerization. Chem. Eng. J. 2004, 97 (1), 27−35. Reactors. Chem. Eng. Sci. 1994, 49 (13), 2035−2045.
(5) Kiashemshaki, A.; Mostoufi, N.; Sotudeh-Gharebagh, R. Two- (25) Lucas, A.; Arnaldos, J.; Casal, J.; Puigjaner, L. Improved
Phase Modeling of a Gas Phase Polyethylene Fluidized Bed Reactor. Equation for the Calculation of Minimum Fluidization Velocity. Ind.
Chem. Eng. Sci. 2006, 61 (12), 3997−4006. Eng. Chem. Process Des. Dev. 1986, 25 (2), 426−429.
(6) Luo, Z.-H.; Su, P.-L.; Shi, D.-P.; Zheng, Z.-W. Steady-State and (26) Kunii, D.; Levenspiel, O. Fluidization Engineering, 2nd ed.;
Dynamic Modeling of Commercial Bulk Polypropylene Process of Butterworth-Heinmann: Boston, MA, 1991.
Hypol Technology. Chem. Eng. J. 2009, 149 (1−3), 370−382. (27) Mostoufi, N.; Cui, H.; Chaouki, J. A Comparison of Two- and
(7) Floyd, S.; Choi, K. Y.; Taylor, T. W.; Ray, W. H. Polymerization Single-Phase Models for Fluidized-Bed Reactors. Ind. Eng. Chem. Res.
of Olefins through Heterogeneous Catalysis. Iii. Polymer Particle 2001, 40 (23), 5526−5532.
Modelling with an Analysis of Intraparticle Heat and Mass Transfer (28) Hilligardt, K.; Werther, J. Local Bubble Gas Hold-up and
Effects. J. Appl. Polym. Sci. 1986, 32 (1), 2935−2960. Expansion of Gas/Solid Fluidized Beds. Ger. Chem. Eng. 1986, 9 (4),
(8) Choi, K. Y.; Ray, W. H. The Dynamic Behaviour of Fluidized Bed 215−221.
Reactors for Solid Catalysed Gas Phase Olefin Polymerization. Chem. (29) Cui, H.; Mostoufi, N.; Chaouki, J. Characterization of Dynamic
Eng. Sci. 1985, 40 (12), 2261−2279. Gas-Solid Distribution in Fluidized Beds. Chem. Eng. J. 2000, 79 (2),
(9) McAuley, K. B.; MacGregor, J. F.; Hamielec, A. E. A Kinetic 133−143.
Model for Industrial Gas-Phase Ethylene Copolymerization. AIChE J.
1990, 36 (6), 837−850.
(10) Fernandes, F. A. N.; Lona, L. M. F. Heterogeneous Modeling
for Fluidized-Bed Polymerization Reactor. Chem. Eng. Sci. 2001, 56
(3), 963−969.
(11) Hatzantonis, H.; Yiannoulakis, H.; Yiagopoulos, A.; Kiparissides,
C. Recent Developments in Modeling Gas-Phase Catalyzed Olefin
Polymerization Fluidized-Bed Reactors: The Effect of Bubble Size
Variation on the Reactor’S Performance. Chem. Eng. Sci. 2000, 55
(16), 3237−3259.
(12) Harshe, Y. M.; Utikar, R. P.; Ranade, V. V. A Computational
Model for Predicting Particle Size Distribution and Performance of
Fluidized Bed Polypropylene Reactor. Chem. Eng. Sci. 2004, 59 (22−
23), 5145−5156.
(13) Shi, D.-P.; Luo, Z.-H.; Guo, A.-Y. Numerical Simulation of the
Gas-Solid Flow in Fluidized-Bed Polymerization Reactors. Ind. Eng.
Chem. Res. 2010, 49 (9), 4070−4079.
(14) Khare, N. P.; Lucas, B.; Seavey, K. C.; Liu, Y. A.; Sirohi, A.;
Ramanathan, S.; Lingard, S.; Song, Y.; Chen, C.-C. Steady-State and
Dynamic Modeling of Gas-Phase Polypropylene Processes Using
Stirred-Bed Reactors. Ind. Eng. Chem. Res. 2004, 43 (4), 884−900.
(15) Meier, G. B.; Weickert, G.; van Swaaij, W. P. M. FBR for
Catalytic Propylene Polymerization: Controlled Mixing and Reactor
Modeling. AIChE J. 2002, 48 (6), 1268−1283.
(16) Zacca, J. J.; Debling, J. A.; Ray, W. H. Reactor Residence Time
Distribution Effects on the Multistage Polymerization of Olefins–I.
Basic Principles and Illustrative Examples, Polypropylene. Chem. Eng.
Sci. 1996, 51 (21), 4859−4886.
(17) Shamiri, A.; Hussain, M. A.; Mjalli, F. S.; Mostoufi, N. Kinetic
Modeling of Propylene Homopolymerization in a Gas-Phase Fluid-
ized-Bed Reactor. Chem. Eng. J. 2010, 161 (1−2), 240−249.
(18) Shamiri, A.; Azlan Hussain, M.; Sabri Mjalli, F.; Mostoufi, N.;
Saleh Shafeeyan, M. Dynamic Modeling of Gas Phase Propylene

8705 dx.doi.org/10.1021/ie501155h | Ind. Eng. Chem. Res. 2014, 53, 8694−8705

Potrebbero piacerti anche