Sunteți pe pagina 1din 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/258699880

Dynamics and Predictive Control of Gas Phase Propylene Polymerization in


Fluidized Bed Reactors

Data · November 2013


DOI: 10.13140/RG.2.1.4298.3128

CITATION READS

1 183

5 authors, including:

Ahmad Shamiri Mohd azlan Hussain


UCSI University University of Malaya
70 PUBLICATIONS   1,030 CITATIONS    242 PUBLICATIONS   2,807 CITATIONS   

SEE PROFILE SEE PROFILE

Farouq S. Mjalli Navid Mostoufi


Sultan Qaboos University University of Tehran
203 PUBLICATIONS   3,877 CITATIONS    265 PUBLICATIONS   2,701 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Carbon dioxide removal from contaminated air View project

Development of an advance model and optimal control system for gas-phase olefin polymerization in fluidized-bed catalytic reactor View project

All content following this page was uploaded by Ahmad Shamiri on 19 May 2014.

The user has requested enhancement of the downloaded file.


PROCESS SYSTEMS ENGINEERING AND PROCESS SAFETY
Chinese Journal of Chemical Engineering, 21(9) 1015—1029 (2013)
DOI: 10.1016/S1004-9541(13)60565-0

Dynamics and Predictive Control of Gas Phase Propylene


Polymerization in Fluidized Bed Reactors*

Ahmad Shamiri1,2,**, Mohamed azlan Hussain1,3, Farouq sabri Mjalli4, Navid Mostoufi5 and
Seyedahmad Hajimolana1
1
Department of Chemical Engineering, University of Malaya, Kuala Lumpur 50603, Malaysia
2
Training Center, Razi Petrochemical Company, Bandar Imam 161, Iran
3
UM Power Energy Dedicated Advanced Centre (UMPEDAC), University of Malaya, Kuala Lumpur 50603, Malaysia
4
Petroleum and Chemical Engineering Department, Sultan Qaboos University, Muscat 123, Oman
5
Process Design and Simulation Research Center, School of Chemical Engineering, College of Engineering, Uni-
versity of Tehran, Tehran 11155/4563, Iran

Abstract A two-phase dynamic model, describing gas phase propylene polymerization in a fluidized bed reactor,
was used to explore the dynamic behavior and process control of the polypropylene production rate and reactor
temperature. The open loop analysis revealed the nonlinear behavior of the polypropylene fluidized bed reactor, jus-
tifying the use of an advanced control algorithm for efficient control of the process variables. In this case, a central-
ized model predictive control (MPC) technique was implemented to control the polypropylene production rate and
reactor temperature by manipulating the catalyst feed rate and cooling water flow rate respectively. The corre-
sponding MPC controller was able to track changes in the setpoint smoothly for the reactor temperature and pro-
duction rate while the setpoint tracking of the conventional proportional-integral (PI) controller was oscillatory with
overshoots and obvious interaction between the reactor temperature and production rate loops. The MPC was able
to produce controller moves which not only were well within the specified input constraints for both control vari-
ables, but also non-aggressive and sufficiently smooth for practical implementations. Furthermore, the closed loop
dynamic simulations indicated that the speed of rejecting the process disturbances for the MPC controller were also
acceptable for both controlled variables.
Keywords model predictive control, fluidized bed reactor, propylene polymerization, Ziegler-Natta catalyst

1 INTRODUCTION is highly exothermic. To maintain acceptable polymer


production rate, which is an important goal for indus-
Modeling and control of polymerization process try, it is necessary to keep the bed temperature above
in fluidized bed reactors such as polypropylene pro- the dew point of the reactants to avoid gas condensation
duction are challenging issues in process control en- and below the melting point of the polymer to prevent
gineering. This is due to many factors such as the high particle melting, agglomeration and consequent reactor
non-linearity of the process dynamics involving com- shut down. For these reasons, process stabilization for
plicated reaction mechanisms, complex flow character- propylene polymerization in a fluidized bed reactor is
istics of gas and solids, various heat and mass transfer a challenging problem to be addressed through an ef-
mechanisms and the interaction between the process ficient control system design.
control loops. Many studies were reported on the mod- Most of the reactor design and control problems
eling and control of ethylene polymerization processes are associated with achieving adequate production rate
using various types of algorithms [1-6]. However, there and heat removal from the reactor. The steady-state
is little work on the modeling and control of gas phase and dynamic behavior of the reactor is influenced by
propylene polymerization in fluidized bed reactors. many process variables such as the superficial gas
The schematic of an industrial gas-phase fluid- velocity, feed gas temperature, monomer concentration,
ized bed polypropylene reactor with control design is catalyst activity and catalyst feed rate. Choi and Ray
shown in Fig. 1. The feed gas stream provides mono- [1] used a dynamic model considering bubble and
mer, hydrogen and nitrogen, and at the same time agi- emulsion phases in the bed in the first attempt to de-
tates and fluidizes the bed through the distributor and scribe the dynamics of polypropylene production. They
also removes the heat of polymerization reaction. Po- showed that a PI feedback control scheme can be used
lymerization occurs in the fluidized bed in the pres- to control the process transients, but it was limited by
ence of Ziegler-Natta catalyst and triethyl aluminum the recycle gas cooling capacity. Dadebo et al. [2]
co-catalyst. The unreacted gas exits the top of the reactor showed that for the temperature control of industrial
and is then compressed and cooled before being fed gas phase polyethylene reactors, the nonlinear error
back into the bottom of the fluidized bed. The polymer trajectory controller (ETC) exhibited significantly supe-
production rate in this system is limited by heat removal rior responses in terms of speed, damping and robust-
from the circulating gas since the polymerization reaction ness compared with an optimally-tuned proportional

Received 2012-02-17, accepted 2013-01-12.


* Supported by the Research Grants of the Research Council of Malaya.
** To whom correspondence should be addressed. E-mail: a.shamiri@um.edu.my
1016 Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013

Figure 1 Simplified schematic of an industrial gas-phase fluidized bed polypropylene reactor with a model predictive con-
trol (MPC) design
TI—temperature transmitter; FT—mass flow transmitter

integral derivative (PID) controller over a wide range and efficient process control scheme, model predictive
of operating conditions. Ali et al. [3] proposed a control (MPC) algorithm, is gradually becoming
non-linear process model based control (modified ge- popular in the process industries due to their ability to
neric model control) for the control of liquid propylene handle process interactions, constraints on manipulated
polymerization. The controller manipulated the hy- and output variables and time delays as well as their
drogen flow rate and cooling water flow to follow the ability to control nonlinear processes [25]. The unified
setpoints for cumulative melt index and reaction con- framework of MPC and its capacity to incorporate
version. They showed that the proposed controller has various practical control objectives and requirements
good setpoint tracking and disturbance rejection prop- have found extensive acceptance in the chemical
erties and it was superior to the conventional generic process industries. However, linear model predictive
model control and proportional-integral (PI) control control (LMPC) algorithms are incapable to handle
algorithms. Ibrehem et al. [4] modeled the process in the complexities encountered in nonlinear processes.
four phases namely: bubble, cloud, emulsion and solid Nonlinear model predictive control (NMPC) has good
phases with polymerization reactions occurring in the capability in improving control and operation of nonlin-
emulsion and solid phases to describe the olefin po- ear processes such as these polymerization processes.
lymerization in fluidized bed reactor. They imple- The basic principle of NMPC is the same as LMPC
mented neural-networks based predictive controller with the exception that the model describing the proc-
(NNMPC) to control the system and showed that the
ess dynamics is nonlinear [12, 14]. Successful applica-
performance of the NNMPC was superior to the PID
tions of NMPC on polymerization reactors have been
controller in terms of the setpoint tracking. Summary
of relevant studies in olefin polymerization and its reported, which is also capable of dealing with unan-
control are listed in Table 1. The summary shows that ticipated changes in the process dynamics through the
most of the control studies are related to polyethylene state estimator. Summary of relevant studies in olefin
and only three of them are related to polypropylene. polymerization and its control are listed in Table 1.
Due to nonlinearities in the process dynamics and Based on our knowledge it is the first time that a
difficulties involved in the control of the gas phase multiple-input, multiple output (MIMO) model pre-
propylene polymerization fluidized bed reactor, an dictive control (MPC) strategy as a centralized MPC
efficient process control scheme is vital for stable and technique is used to control the polypropylene produc-
efficient operation of the process. The most commonly tion rate and reactor temperature for setpoint tracking
used controllers in industrial applications are propor- and disturbance rejection. Furthermore, a two-phase
tional-integral (PI) controllers [23]. A more intelligent model with comprehensive kinetics for propylene
Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013 1017

Table 1 Summary of control studies in olefin polymerization


Reactor type Model Control algorithm Variables controlled Reference
fluidized bed polyethylene two phase PID feed concentration [7]
pressure
temperature
bed level
continues stirred bed polypropylene CSTR (well mixed) PID temperature [8]
bed level
pressure
fluidized bed polyethylene CSTR (well mixed) NMPC feed flow rates [9]
LMPC temperature
PI bleed flow and pressure
fluidized bed polyethylene CSTR (well mixed) ETC temperature [2]

fluidized bed polyethylene CSTR (well mixed) PID melt index and density [10]

fluidized bed polypropylene and two phase PI temperature [1]


polyethylene
hollow shaft liquid-pool CSTR (well mixed) GMC melt index [3]
polypropylene monomer conversion
fluidized bed polyethylene CSTR (well mixed) IMC melt index [11]
density
fluidized bed polyethylene two phase MPC (INCA) melt index [12]
density
temperature
production rate
fluidized bed polyethylene four phase NNMPC molecular weight [4]
temperature
fluidized bed polyethylene two phase and CSTR (well mixed) PID temperature [13]
pressure
bed level
polyethylene CSTR CSTR (well mixed) NMPC partial pressures of the [14]
gas phase compositions
production rate
fluidized bed polyethylene two phase PI temperature [15]
monomer concentration
fluidized bed polyethylene two phase PI temperature [16]
NMPC monomer concentration
fluidized bed polyethylene CSTR (well mixed) NMPC molecular weight distribution [17]

fluidized bed polyethylene two phase PID bed level [18]

fluidized bed polyethylene CSTR (well mixed) PI temperature [19]


pressure
bed level
density
melt flow index
production rate
fluidized bed polyethylene CSTR (well mixed) PID density and melt flow index [20]
optimal servo system
fluidized bed polyethylene CSTR (well mixed) PID temperature [21]

slurry polypropylene CSTR (well mixed) NMPC melt flow index [22]
amount of unreacted monomer
fluidized bed polyethylene two phase PI temperature [23]

fluidized bed polyethylene CSTR (well mixed) fuzzy logic temperature [24]

homo-polymerization in a fluidized bed reactor that excess gas in the emulsion phase with polymerization
considers the presence of particles in the bubbles and reaction in both phases is also considered [26-29].
1018 Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013

2 MATHEMATICAL MODEL OF PROPYLENE assumes that all particles have the same mean diame-
POLYMERIZATION ter of 500×10−6 m [1, 2, 4, 13, 26]. Polypropylene parti-
cles of such size are Geldart B, thus, the constants of
2.1 Reactor model this type of particles were chosen and shown in Table
2. It is worth noting that the same approach was
In the present study, the kinetic model of propyl- adopted in similar modeling attempts [13, 31] and pro-
ene homo-polymerization over a Ziegler-Natta cata- vided good prediction of the reactor performance.
lyst (based on the kinetic model developed by Shamiri The correlations required for estimating bubble
et al. [30]) and the dynamic two-phase flow structure volume fraction in the bed, voidage of emulsion and
proposed by Cui et al. [27, 28] were combined and im- bubble phases, gas velocities in emulsion and bubble
plemented to provide a more realistic understanding of phases and mass and heat transfer coefficients for the
the phenomena encountered in the reactor hydrody- improved two-phase model are summarized in Table 2.
namics. Fluidized bed reactors are able to produce a broad
The two-phase model presented in this work range of polypropylene.

Table 2 Correlations and equations used in the two-phase model


Parameter Formula Reference
2 1/ 2
minimum fluidization velocity Remf = [(29.5) + 0.357 Ar ] − 29.5 [32]
bubble velocity U b = U 0 − U e + ubr [33]
1/ 2
bubble rise velocity ubr = 0.711( gd b ) [34]

U0 − δU b
emulsion velocity Ue = [34]
1−δ

bubble diameter d b = d b0 [1 + 27(U 0 − U e )]1 / 3 (1 + 6.84 H ) [35]


d b0 = 0.0085 (for Geldart B)
−1
⎛ 1 1 ⎞
mass transfer coefficient K be = ⎜ + ⎟ [33]
⎝ K bc K ce ⎠
1/ 2 1/ 4
Ue Dg g
K bc = 4.5 + 5.85
db d b5 / 4

Dgε eubr
K ce = 6.77
db
−1
⎛ 1 1 ⎞
heat transfer coefficient H be = ⎜ + ⎟ [33]
H
⎝ bc H ce ⎠

( kg ρgCpg )
1/ 2
U e ρg Cpg g1 / 4
H bc = 4.5 + 5.85 5/ 4
db d b

1/ 2
⎛ ε eubr ⎞
H ce = 6.77 ( ρg Cpg kg )
1/ 2
⎜ d3 ⎟
⎝ b ⎠

⎡ ⎛ U 0 − U mf ⎞⎤
bubble phase fraction δ = 0.534 ⎢1 − exp ⎜ − ⎟⎥ [27]
⎣ ⎝ 0.413 ⎠⎦

⎛ U 0 − U mf ⎞
emulsion phase porosity ε e = ε mf + 0.2 − 0.059 exp ⎜ − ⎟ [27]
⎝ 0.429 ⎠

⎛ U 0 − U mf ⎞
bubble phase porosity ε b = 1 − 0.146 exp ⎜ − ⎟ [27]
⎝ 4.439 ⎠

volume of polymer phase in the emulsion phase VPe = AH (1 − ε e ) (1 − δ ) [26]


volume of polymer phase in the bubble phase VPb = AH (1 − ε b ) δ [26]
volume of the emulsion phase Ve = A(1 − δ ) H [26]
volume of the bubble phase Vb = Aδ H [26]
Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013 1019

The UNIPOL Polypropylene Technology is a the superficial gas velocity increases in a fluidized bed
simple and delicate processing system, comprising of due to more solid particles entering bubbles and more
either one or two gas-phase fluidized-bed reactors. To gas entering the emulsion phase [13, 18, 27, 28, 31].
produce homopolymers and random copolymers, a (3) The bubbles are assumed to be spherical with
single reactor is used [36]. In this work a single reactor uniform size and rise up through the bed in plug flow
is used to produce propylene homopolymers. Charac- at constant velocity.
terization of polymer properties was modeled using the (4) Mass and heat transfer resistances between
population balance and the method of moments [30, 37]. gas and solids in the emulsion and bubble phases are
Assuming that the only significant consumption negligible (i.e., low to moderate catalyst activity) [40].
of monomer is by the propagation reaction and hy- (5) Radial concentration and temperature gradi-
drogen by transfer to hydrogen, the following expres- ents in the reactor are negligible due to the agitation
sions for the consumption rate of component (mono- produced by the up-flowing gas.
mer and hydrogen) can be obtained: (6) Elutriation of solids at the top of the bed is
For monomer: negligible.
Ns (7) Constant particle size exists throughout the
Ri = ∑ [ M i ]Y (0, j )kp ( j ) , i =1 (1) bed [1, 2, 4, 13, 26].
j =1 (8) The direction of mass transfer was assumed to
For hydrogen: be from bubble to emulsion phase.
Based on these assumptions, the following dynamic
Ns material balances were derived for all the components
Ri = ∑ [ M i ]Y (0, j )kfh ( j ) , i=2 (2) in the bed.
j =1 For bubbles:
The total polymer production rate for each phase [ M i ]b,in U b Ab − [ M i ]b U b Ab − Rvε b [ M i ]b −
can be then calculated from
Ab
K be ([ M i ]b − [ M i ]e )Vb − (1 − ε b )
VPFR ∫
2 Ri ,b dz =
Rp = ∑ M w,i Ri (3)
i =1 d
in which Ri is the instantaneous rate of reaction for
(Vbε b [M i ]b ) (4)
dt
monomer and hydrogen. The reaction rate constants
were taken from literature and are listed in Table 3. For emulsion:
The assumptions made in developing the equa- [ M i ]e,in U e Ae − [ M i ]e U e Ae − Rvε e [ M i ]e +
tions of the improved two phase model are summa-
⎛ δ ⎞
rized below: K be ([ M i ]b − [ M i ]e )Ve ⎜ ⎟ − (1 − ε e ) Ri ,e =
(1)The polymerization reactions occur in both ⎝1−δ ⎠
bubble and emulsion phases. d
(2) The emulsion phase is considered to be per- (Veε e [M i ]e ) (5)
fectly mixed, and it may contain more gas at higher dt
gas velocities than at the minimum fluidization condi- The energy balances can be expressed as
tion. A better mixing between two phases occurs when For bubbles:

Table 3 Reactions rate constants


Reaction rate constant
Formation Initiation Propagation Activation energy
−1 −1 −1 −1 −1 −1 −1 −1 −1
kf(j)/s ki(j)/L·mol ·s kh(j)/L·mol ·s Khr/L·mol ·s kp(j)/L·mol ·s E/J·mol−1
site type 1 1 22.88 0.1 20 208.6 30124.8
site type 2 1 54.93 0.1 20 22.8849 30124.8
Ref. [38] [39] [38] [38] [39] [39]
Reaction rate constant
Transfer Deactivation
−1 −1 −1 −1 −1 −1 −1 −1
kfm(j)/L·mol ·s kfh(j)/L·mol ·s kfr(j)/L·mol ·s kfs(j)/L·mol ·s kds(j)/s−1
site type 1 0.0462 7.54 0.024 0.0001 0.00034
site type 2 0.2535 7.54 0.12 0.0001 0.00034
Ref. [39] [39] [38] [38] [39]
1020 Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013

m simulation of the heat removal system [2] given as


U b Ab (Tb,in − Tref ) ∑ [ M i ]b,in C pi −
dq qss − q
i =1 = (12)
m dt τ
U b Ab (Tb − Tref ) ∑ [ M i ]b C pi − where q is the dynamic heat removal rate, and qss is
i =1 the steady state heat removal rate given as
⎛ m ⎞ qss = Fg C pg (Te − Tin )
Rv (Tb − Tref ) ⎜ ∑ ε b C pi [ M i ]b + (1 − ε b ) ρ polC p ,pol ⎟ + (13)
⎝ i =1 ⎠ where Fg is the gas flow rate, Cpg is the gas heat ca-
Ab ΔH R pacity and Te is the temperature of the gas entering the
(1 − ε b ) Rpb dz + H be (Te − Tb )Vb −
VPFR ∫ heat exchanger, which is also the reactor temperature.
m
Tin is the outlet temperature of the gas from heat ex-
d changer, which is also the temperature of the gas en-
Vbε b (Tb − Tref ) ∑ C pi ([M i ]b ) =
i =1 dt tering the reactor, given as
⎛ ⎛ m ⎞⎞ d Twi [1 − exp(γ )] − T (1 − η )
⎜⎜ Vb ⎜ ε b ∑ C pi [ M i ]b + (1 − ε b ) ρ pol C p ,pol ⎟ ⎟⎟ (Tb − Tref ) Tin = (14)
η − exp(γ )
⎝ ⎝ i =1 ⎠ ⎠ dt
(6) with
For emulsion: ⎡ 1 1 ⎤
γ = AexU ex ⎢ + ⎥ (15)
m ⎣ Fg C pg Fcw C pw ⎦
U e Ae (Te,in − Tref ) ∑ [ M i ]e,in C pi −
i =1
and
m Fg C pg
U e Ae (Te − Tref ) ∑ [ M i ]e C pi − η= (16)
i =1
Fcw C pw
⎛ m ⎞ Twi is the temperature of cooling water entering the
Rv (Te − Tref ) ⎜ ∑ ε e C pi [ M i ]e + (1 − ε e ) ρ polC p ,pol ⎟ − heat exchanger, Aex is the contact area and Uex is a
⎝ i =1 ⎠ heat transfer coefficient. Fcw and Cpw are the flow rate
δ ⎞
(1 − ε e ) Rpe ΔH R − H beVe ⎛⎜
and heat capacity of the cooling water, respectively.
⎟ (Te − Tb ) −
⎝1− δ ⎠
m
d 3 MULTIPLE-INPUT, MULTIPLE OUTPUT
Veε e (Te − Tref ) ∑ C pi ([M i ]e ) = (MIMO) MODEL PREDICTIVE CONTROL
i =1 dt STRATEGY
⎛ ⎛ m ⎞⎞ d
⎜⎜ Ve ⎜ ε e ∑ C pi [ M i ]e + (1 − ε e ) ρ pol C p ,pol ⎟ ⎟⎟ (Te − Tref ) MPC is an optimization-based control strategy
⎝ ⎝ i =1 ⎠ ⎠ dt which is well suited for constrained, multivariable
(7) processes. The model predictive controller predicts the
The initial conditions for solution of the model future behavior of the actual system over a time inter-
equations are as follows: val defined by the prediction horizon. The simplified
diagram of the MPC algorithm shown in Fig. 2 indi-
[ M i ]b,t =0 = [ M i ]in (8) cates that a process model is used in MPC in order to
predict the controlled variable and in this system the
Tb (t = 0) = Tin (9) process model is used in parallel to the plant. The pre-
dicted controlled variable is fed back to the controller
[ M i ]e,t =0 = [ M i ]in (10) where it is used in an on-line optimization procedure,
which minimizes an appropriate cost function to de-
Te (t = 0) = Tin (11) termine the optimal input to the plant acting as the
manipulated variable. The optimization problem is
2.2 Heat exchanger model solved subject to constraints on input and output vari-
ables and also constraints imposed by the nonlinear
model equations. The controller output is implemented
As shown in Fig. 1, the external gas cooler is a in real time and then at the next sampling time. The
countercurrent single-pass shell and tube heat exchanger difference between the plant measurement yp and the
with the recycle gas on the tube side and cooling water process model output ym is also fed to the controller to
on the shell side. A linear first-order dynamic model eliminate steady state offset and compensate modeling
obtained by imposing first order dynamics on the heat uncertainties. The cost function usually is specified as
removal rate with a corresponding time constant, τ, to a sum of quadratic future errors between the reference
yield a linear dynamic model was considered for trajectory and predicted plant output, and the predicted
Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013 1021

Figure 2 Model predictive control system

Table 4 Operating conditions and physical parameters considered in this work for
modeling fluidized bed polypropylene reactors
Operating conditions Physical parameters
Propylene Hydrogen Catalyst
H Dt V Tref Tin P
concentration concentration feed rate μ/Pa·s ρg/kg·m−3 ρs/kg·m−3 dp/m εmf
/m /m /m3 /K /K /MPa /mol·L−1 /mol·L−1 /g·s−1
7.08 3 50 353.15 328.15 2.5 0.9 0.015 0.3 1.14×10−4 23.45 910 500×10−6 0.45

control effort along the prediction horizon.


ymax (k ) ≥ y (k + i − 1) ≥ ymin (k ) (20)
Due to its ability and simplicity the Model Pre-
dictive Control (MPC) algorithm is implemented to Depending on the problem formulation, parame-
control the gas phase propylene polymerization process ters such as the control horizon, prediction horizon
in this fluidized bed reactor. The MIMO based MPC and weighting matrices in the optimization formula-
controller used in this work takes into account both tion has to be finalized to obtain the best performance
the output and manipulated variables and minimizes for the predicted output.
the following cost functions V(k) as given by [41, 42] In this study, a MIMO system is controlled by the
P ny centralized MPC techniques. This system takes the
{ }
2
V (k ) = ∑∑ w yj ⎡⎣ yi (k + i ) − rj (k + i ) ⎤⎦ + reactor temperature (Te) and production rate (Rp) as
i =1 j =1 controlled variables, with the catalyst feed rate (Fcat)
M nmv and cooling water flow rate (Fcw) as manipulated
∑∑ {wΔj u Δu j (k + i − 1)}
2
(17) variables , and the hydrogen concentration (CH) , pro-
i =1 j =1 pylene concentration (CM) and superficial gas velocity
(U0) as measured disturbances. The improved two
where k is the current sampling interval, k + i is a phase model is written in MATLAB and implemented
future sampling interval (within the prediction horizon), in S-functions in order to be accessed by Simulink.
P is the prediction horizon, ny is the number of plant
outputs, w yj is the weight for output j, [ y j (k + i ) − 4 RESULTS AND DISCUSSION
rj (k + i )] is the predicted deviation at future instant
4.1 Open loop analysis of the propylene polym-
k + i , M is the control horizon, nmv is the number of
erization fluidized bed reactor
manipulated variables (inputs), Δu j (k + i − 1) is the
predicted adjustment (i.e., move) in manipulated To demonstrate the non-linearity of the reactor,
variable j at future (or current) sampling interval the process was simulated at the operating conditions
(k + i − 1) , and wΔj u is the weight for input j. given in Table 4 to explore the effect of step change in
The cost function is also subjected to inequality the superficial gas velocity, catalyst feed rate, cooling
constraints of the manipulated and output variables water flow rate and hydrogen and propylene concen-
given below: trations on the emulsion phase temperature and poly-
Manipulated variable constraint is given as: propylene production rate. The results of the open-loop
umax (k ) ≥ u (k + i − 1) ≥ umin (k ) (18) simulations are shown in Figs. 3 to 7.
As shown in Fig. 3 (a) and 3 (b), the superficial
Manipulated variable rate constraint is given as gas velocity was changed from the nominal value (0.4
m·s−1) using increments of 0.05 m·s−1 in both positive
Δumax (k ) ≥ u (k + i − 1) ≥ Δumin (k ) (19)
and negative directions. The superficial gas velocity has
Output variable constraint is given as: a significant effect on the emulsion phase temperature
1022 Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013

(a) (b)
Figure 3 Effect of step change in U0 on (a) the emulsion phase temperature and (b) production rate (Fcat = 0. 3 g·s−1, CM =
0.9 mol·L−1 and CH = 0.015 mol·L−1)
U0/m·s−1: a—0.4; b—0.25; c—0.3; d—0.35; e—0.45; f—0.5; g—0.55

(a) (b)
Figure 4 Effect of step change in the catalyst feed rate on (a) the emulsion phase temperature and (b) production rate
(U0 = 0.45 m·s−1, CM = 0.9 mol·L−1 and CH = 0.015 mol·L−1)
Fcat/g·s−1: a—0.3; b—0.6; c—0.5; d—0.4; e—0.2; f—0.1

and production rate even for small changes in the su- rate is depicted in Fig. 4 (a) and 4 (b). Once the emul-
perficial gas velocity which indicates that it is an im- sion phase temperature and polypropylene production
portant parameter for the heat removal and conse- rate reached steady state conditions at catalyst feed
quently the control of reactor temperature and produc- rate of 0.3 g·s−1, the catalyst feed rate was changed to
tion rate. Although one can manipulate the fluidizing show the effect of manipulating this variable on the
gas velocity to control the reactor temperature and process dynamics. Small step changes in the catalyst
production rate but changes in fluidization character- feed rate result in significant changes in the reactor
istics (e.g. solids entrainment) may occur when the temperature and polymer production rates. The slightly
gas velocity is varied [1] which is also difficult to con- symmetrical nature of these profiles due to the sys-
trol in practice and thus the fluidizing gas velocity is tematic positive and negative variations in the catalyst
not used as the manipulated variable. flow rate indicates the slightly nonlinear relationships.
This test clearly shows that the negative steps in In this case, using a conventional controller can be
superficial gas velocity have more significant effect acceptable provided that the effect of disturbances,
than the corresponding positive steps for both output measurement noise, modeling uncertainty and process
variables (emulsion phase temperature and production interaction are not pronounced.
rate) and produce non-symmetrical responses. The In the industry, the properties of polypropylene
variation of the emulsion phase temperature and poly- are controlled by varying the monomer and hydrogen
propylene production rate as the superficial gas veloc- feed rates. Increasing the hydrogen to propylene ratio
ity change deviates from the linear behavior. For such in the reactor leads to increased chain transfer and to
nonlinear relationships, the use of conventional con- lower molecular weight. Changes in monomer and
trollers tends to produce poor control of the process hydrogen feed rates also act as disturbances to the
variables, thus justifying the use of a more efficient reactor temperature and polymer production rate. To
control scheme to properly regulate the effect of super- explore these effects, positive and negative steps were
ficial gas velocity on the process variables. introduced in the hydrogen and propylene feed con-
The effect of step changes in the catalyst feed centrations by 0.005 mol·L−1 and 0.1 mol·L−1 respec-
rate on the emulsion phase temperature and production tively. The effect of a step change in the hydrogen
Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013 1023

(a) (b)
Figure 5 Effect of step change in the hydrogen feed concentration on (a) emulsion phase temperature and (b) production
rate (U0 = 0.45 m·s−1, Fcat = 0. 3g·s−1 and CM = 0.9 mol·L−1)
CH/mol·L−1: a—0.015; b—0; c—0.005; d—0.01; e—0.02; f—0.025; g—0.3

(a) (b)
Figure 6 Effect of step change in the propylene feed concentration on (a) the emulsion phase temperature and (b) produc-
tion rate (Fcat = 0. 3 g·s−1, U0 = 0.45 m·s−1 and CH = 0.015 mol·L−1)
CM/mol·L−1: a—0.7; b—1; c—0.9; d—0.8; e—0.6; f—0.5; g—0.4

concentration in the feed on the emulsion phase tem- in terms of the control algorithm used.
perature and production rate predicted by the model is Figure 7 illustrates the effect of step changes in
shown in Fig. 5 (a) and 5 (b). The initial hydrogen the cooling water flow rate on the emulsion phase
feed concentration was set at 0.015 mol·L−1 and step temperature. By changing the cooling water flow rate
changes was introduced from the steady state condi- from an initial value of 3×105 g·s−1, the emulsion
tions. As can be seen, hydrogen concentration has highly
nonlinear relationship with the reactor temperature
and polymer production rate. Increasing the hydrogen
concentration leads to enhancing the chain transfer
reaction rate and consequently lowers the degree of
polymerization due to the shorter polymer chain length.
The effect of varying the feed propylene concen-
tration on the emulsion phase temperature and produc-
tion rate is shown in Fig. 6 (a) and 6 (b). The propyl-
ene concentration was changed from the nominal
value of 0.7 mol·L−1. By increasing the feed propylene
concentration, the emulsion phase temperature and
production rate increases. Increasing the propylene
concentration leads to enhancing the reaction rate and
consequently the degree of polymerization. This ob- Figure 7 Effect of step change in the cooling water flow
servation, in addition to the ones discussed previously, rate on the emulsion phase temperature (Fcat = 0. 3 g·s−1,
U0 = 0.45 m·s−1, CM = 0.9 mol·L−1 and CH = 0.015 mol·L−1)
show nonlinear behavior as inferred from the produced
dynamic profiles. Such observations indicate that the Fcw/g·s−1: a—3×105; b—1.5×105; c—2×105; d—2.5×105; e—
process model under investigation needs special attention 3.5×105; f—4×105; g—4.5×105
1024 Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013

Figure 8 A simplified schematic of the process sub-models and control system of the gas phase propylene polymerization
fluidized bed reactor

phase temperature changes significantly and shows and Cooper [43] were used as the starting values but
strong dependency. This indicates that it has an im- the exact values used in this work were the result of
portant role on the heat removal and on the control of further fine tuning based on actual control perform-
the fluidized bed polymerization rector temperature. ance. In addition to the selection of controller tuning
parameters, appropriate values were chosen for the
4.2 Closed loop performance of the model predic- input/output variables constraints and imposed on the
tive controller scheme system based on practical experience. The values of
the upper and lower limits of the constraints for the
Fcat − Rp profile are: 0≤Fcat≤100%, −20%≤ΔFcat≤
A simplified schematic of the process sub-models 20% and for the Fcw − Te profile are: 0≤Fcw≤100%,
and control system of the gas phase propylene polym- −20%≤ΔFcw≤20%. The control system has a sam-
erization fluidized bed reactor is shown in Fig. 8. pling time of 15 s, a prediction horizon of 20, a con-
trol horizon of 2, an output variables weight of 0.005
4.2.1 MPC controller design for Rp and 0.0005 for Te. The manipulated variables
The results presented in the previous section weights selected were 0.1 for Fcat and 0.1 for Fcw.
show the nonlinear behavior of the system as inferred
from the produced dynamic profiles. Such observa- 4.2.2 Setpoint tracking
tions imply that the process model under investigation It is interesting to see how the nonlinear poly-
needs special attention in terms of the control algo- propylene fluidized bed reactor behaves under advanced
rithm to be used for this process. Despite the fact that and conventional control algorithms. Hence the closed
the superficial gas velocity, catalyst feed rate and hy- loop performance of the MPC scheme in tracking a
drogen and propylene concentrations have significant series of setpoint changes for emulsion phase tempera-
effect on the output variables, based on actual process ture and production rate in the polymerization reactor
experience, not all of them are suitable to be selected was evaluated. For this purpose, a series of setpoint
as manipulated variables. In the industrial fluid- changes in opposite directions were introduced for the
ized-bed olefin polymerization reactor, the tempera- controlled variables. The magnitude of the setpoints
ture is usually controlled by manipulating the cooling introduced for these variables were typical of the re-
water flow rate of an upstream heat exchanger which spective nominal operating ranges. The commonly
is used to remove the heat of the exothermic polym- used conventional controllers in industrial applications
erization reaction from the recycle gas stream [2], and are proportional integral (PI) controllers due to their
changes to the catalyst feed rate can be used to control relative simple design, simple control structure and
the polymer production rate [11]. Based on this approach, moderate cost [23]. Therefore, for comparison purposes,
Fcat and Fcw were selected as manipulated variables a conventional PI controller, tuned using the standard
and Rp and Te were used as control variables to design Ziegler-Nichols (Z-N) [44] open loop tuning method,
the centralized MPC in this work (see Fig. 1). In this was included in this simulation work. The PI control-
case, superficial gas velocity (U0), propylene concen- ler tuning parameters were calculated based on an
tration (CM), and hydrogen concentration (CH) were analysis of the open loop process reaction curve for a
classified as the disturbance variables. particular operational region of the process, where the
To ensure good performance of the MPC con- open loop response was approximated by a first order
troller, the tuning parameters must be appropriately plus dead time (FOPDT) dynamic model [K = −0.3
tuned. The tuned parameters suggested by Shridhar (process gain), τ = 890 s (process time constant) and
Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013 1025

(a) (b)
MPC; setpoint PI; setpoint
Figure 9 Comparison of the performance between the MPC and the PI controller (Kc = 7, τI = 46.6666, τD = 0) in tracking
series of setpoint changes in the Rp loop

(a) (b)
MPC; setpoint PI; setpoint
Figure 10 Comparison of the performance between the MPC and the PI controller (Kc = −34, τI = 271.5654, τD = 0) in track-
ing series of setpoint changes in the Te loop

α = 80 (process dead time) for the Te profile, and respectively. It can be seen that the PI controller
K = 42.5 (process gain), τ = 988 s (process time con- moves are vigorous for the Rp profile. Although the
stant) and α = 5 (process dead time) for the Rp profile]. controller moves observed for the Te profile (Fig. 12)
The parameters tuned using Ziegler and Nichols are less vigorous and physically realizable, such con-
[44] method were used as starting values but the exact troller moves typically are not desired during practical
values used in this work were the result of further fine implementations due to possible ‘wear and tear’ which
tuning. can severely shorten the life span of the actuator. In
Figures 9 and 10 show the performance of the general, the MPC is able to not only produce control-
MPC as compared to the PI controller in tracking the ler moves which are well within the specified input
series of setpoint changes for Rp and Te profiles, re- constraints, but also the controller moves produced are
spectively. The MPC controller is able to track the non-aggressive and smooth for practical implementa-
changes in the setpoints for both profiles. The quality tions. The high quality control performance of the
of the setpoint tracking as demonstrated by the MPC MPC is attributed to its ability to handle constraints in
in terms of its ability to attain minimal overshoot and the inputs, of which the conventional PI controller was
minimize the effect of process variables interaction is unable to achieve due to the high degree of interaction
acceptable. In contrast, the PI controller produced between these process variables. To summarize, Table 5
noticeable overshoots for both loops, in addition to the shows the integral absolute error (IAE) for the two
obvious effect of process variables interaction for the controllers in tracking the series of setpoint changes
Te loop. This illustrates the failure of the decentralized for both controlled variables and the performance of
PI controller to track the setpoints for both loops due the MPC is also superior.
to loop interactions as well as the process nonlinearity.
The controller moves for MPC and PI controllers 4.2.3 Disturbance rejection
in tracking these series of setpoint changes for the Rp The results thus far have demonstrated the good
and the Te profiles are shown in Figs. 11 and 12, performance of the MPC in handling the servo process
1026 Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013

(a) (b)
Figure 11 Comparison of the corresponding controller moves between the MPC and the PI controller for the Rp profile
MPC; PI

(a) (b)
Figure 12 Comparison of the corresponding controller moves between the MPC and the PI controller for the Te profile
MPC; PI

Table 5 Integral absolute error (IAE) for the MPC and mol·L−1) at the time of 13000 s. Figs. 13 to 15 show
the PI controller in tracking series of setpoint changes for the performance of the MPC in rejecting the effect of
the production rate (Rp) and the emulsion phase such changes on the Rp and Te profiles. Generally, the
temperature (Te) profiles MPC controller was able to reject these disturbances
Controller IAE for Rp profile IAE for Te profile in an efficient manner. As a general observation, the
4 superficial gas velocity had the largest effect on both
MPC 4.31×10 2.22×103
the Rp and Te profiles. On the contrary, hydrogen and
5
PI controller 1.84×10 4.3×103 propylene concentrations have negligible effect on
both of the profiles.

control problem for the polymerization reactor as 5 CONCLUSIONS


compared to the conventional controller. However, for
a particular controller to be practically useful in the A two-phase dynamic model describing the gas
process industry, the controller must additionally be phase propylene polymerization in a fluidized bed
able to handle regulatory problems effectively. In this reactor was used to investigate the dynamic behavior
case, three variables i.e., hydrogen concentration (CH), and control on the polypropylene production rate and
superficial gas velocity (U0) and propylene concentra- reactor temperature. The hydrodynamics of the fluid-
tion (CM) were identified as the measured disturbance ized bed reactor of polypropylene production was based
variables. The disturbance variables were introduced on the dynamic two-phase concept of fluidization.
one at a time to observe their individual effect on the Results of the open loop analysis revealed that the
Rp and Te profiles. Each disturbance variable was in- catalyst feed rate and gas superficial velocity have
creased by 15% of its nominal value (superficial gas significant effect on the transient behavior of the
velocity from 0.4 m·s−1 to 0.46 m·s−1, hydrogen con- polypropylene fluidized bed reactor and the process
centration from 0.01 mol·L−1 to 0.0115 mol·L−1 and shows nonlinear behavior, especially by the variation
propylene concentration from 0.8 mol·L−1 to 0.92 of the superficial gas velocity.
Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013 1027

(a) (b)
Figure 13 Performance of the MPC controller in rejecting the effect of superficial gas velocity (U0) on the Rp and Te profiles
(a 15% increase in the superficial gas velocity was introduced at time = 13000)

(a) (b)
Figure 14 Performance of the MPC controller in rejecting the effect of hydrogen concentration (CH) on the Rp and Te pro-
files (a 15% increase in hydrogen concentration was introduced at time = 13000 s)

(a) (b)
Figure 15 Performance of the MPC controller in rejecting the effect of propylene concentration (CM) on the Rp and Te pro-
files (a 15% increase in the propylene concentration was introduced at time = 13000 s)

Two control algorithms namely, the conventional loops while the MPC moves were shown to be
PI and MPC were tested for the stabilization of the non-aggressive and sufficiently smooth for practical
process during setpoint and disturbances changes. The implementations. By introducing changes in the dis-
closed-loop simulations for setpoint changes revealed turbance variables, where each disturbance variable
that the MPC is able to track the changes in the setpoint was increased by 15% of its nominal value, the closed
effectively with minimal overshoot while the PI con- loop dynamic simulations indicated that the speed of
troller produces oscillatory response for the reactor disturbance rejection for the MPC controller was ac-
temperature and production profiles. The PI controller ceptable for both the Rp and Te profiles. In conclusion,
produces overshoots for both Rp and Te loops, in addi- the results demonstrated excellent performance of the
tion to the obvious effect of loop interaction for both MPC in handling servo and regulatory process control
1028 Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013

problems for this polymerization reactor. Remf Reynolds number at minimum fluidization velocity, Remf =
U mf ρg d p / μg
Rp production rate, kg·s−1
NOMENCLATURE
Rpb bubble phase production rate, kg·s−1
Rpe emulsion phase production rate, kg·s−1
A cross sectional area of the reactor, m2 Rv volumetric polymer phase outflow rate from the reactor, m3·s−1
Aex heat transfer area of the heat exchanger, m2 t time, s
Ar Archimedes number, d p3 ρg ( ρs − ρg ) g / μ 2 T temperature of the gas entering the exchanger, K
CH hydrogen concentration, kmol·m−3 Tin temperature of the inlet gaseous stream, K
CM propylene concentration, kmol·m−3 Twi temperature of the cooling water entering the heat exchanger, K
Cpi specific heat capacity of component i, J·kg−1·K−1 U0 superficial gas velocity, m·s−1
Cpg specific heat capacity of gaseous stream, J·kg−1·K−1 Ub bubble velocity, m·s−1
Cp,pol specific heat capacity of solid product, J·kg−1·K−1 Ue emulsion gas velocity, m·s−1
Cpw specific heat capacity of the cooling water, J·kg−1·K−1 Uex overall heat transfer coefficient of the heat exchanger, J·K−1·s−1·m−2
Dg gas diffusion coefficient, m2·s−1 Umf minimum fluidization velocity, m·s−1
Dt reactor diameter, m V reactor volume, m3
db bubble diameter, m Vb volume of the bubble phase, m3
dp particle diameter, m Ve volume of the emulsion phase, m3
Fg gas flow rate, kg·s−1 Vp volume of polymer phase in the reactor, m3
Fcat catalyst feed rate, kg·s−1 Vpb volume of polymer phase in the bubble phase, m3
Fcw cooling water flow rate, kg·s−1 Vpe volume of polymer phase in the emulsion phase, m3
H height of the reactor, m VPFR volume of PFR, m3
Hbe bubble to emulsion heat transfer coefficient, W·m−3·K−1 Y(n,j) nth moment of chain length distribution for living polymer pro-
Hbc bubble to cloud heat transfer coefficient, W·m3·K−1 duced at a site of type j
Hce cloud to emulsion heat transfer coefficient, W·m3·K−1 γ dimensionless variable defined in Eq. (14)
ΔHR heat of reaction, J·kg−1 δ volume fraction of bubbles in the bed
j active site type εb void fraction of bubble for Geldart B particles
kds(j) spontaneous deactivation rate constant for a site of type j εe void fraction of emulsion for Geldart B particles
kf(j) formation rate constant for a site of type j εmf void fraction of the bed at minimum fluidization
kfh(j) transfer rate constant for a site of type j with terminal monomer η dimensionless variable defined in Eq. (14)
M reacting with hydrogen μ gas viscosity, Pa·s
kfm(j) transfer rate constant for a site of type j with terminal monomer ρ density, kg·m−3
M reacting with monomer M ρg gas density, kg·m−3
kfr(j) transfer rate constant for a site of type j with terminal monomer ρpol polymer density, kg·m−3
M reacting with AlEt3 Subscripts
kfs(j) spontaneous transfer rate constant for a site of type j with ter- 1 propylene
minal monomer M 2 hydrogen
kh(j) rate constant for reinitiating of a site of type j by monomer M b bubble phase
khr(j) rate constant for reinitiating of a site of type j by cocatalyst e emulsion phase
ki(j) rate constant for initiation of a site of type j by monomer M g gas mixture property
kp(j) propagation rate constant for a site of type j with terminal i component type number
monomer M reacting with monomer M in inlet
kg gas thermal conductivity, W·m−1·K−1 j active site type number
kp(j) propagation rate constant for a site of type j with terminal mf minimum fluidization
monomer M reacting with monomer M pol polymer
Kbc bubble to cloud mass transfer coefficient, s−1 PP polypropylene
Kbe bubble to emulsion mass transfer coefficient, s−1 ref reference condition
Kce cloud to emulsion mass transfer coefficient, s−1
M monomer (propylene)
[Mi] concentration of component i in the reactor, kmol·m−3
REFERENCES
[Mi]in concentration of component i in the inlet gaseous stream, kmol·m−3
Mw monomer molecular weight, kg·kmol−1 1 Choi, K.Y., Ray, W.H., “The dynamic behaviour of fluidized bed re-
NS number of active site types actors for solid catalysed gas phase olefin polymerization”, Chem.
P pressure, Pa Eng. Sci., 40 (12), 2261-2279 (1985).
2 Dadebo, S.A., Bell, M.L., McLellan, P.J., McAuley, K.B., “Tem-
q dynamic heat removal rate, J·s−1
perature control of industrial gas phase polyethylene reactors”, J.
qss steady state heat removal rate, J·s−1
Process Contr., 7 (2), 83-95 (1997).
r number of units in polymer chain
3 Ali, M.A.H., Betlem, B., Weickert, G., Roffel, B. “Non-linear model
R instantaneous consumption rate of monomer, kmol·s−1 based control of a propylene polymerization reactor”, Chem. Eng.
Ri instantaneous rate of reaction for monomer i, kmol·s−1 Process., 46 (6), 554-564 (2007).
R(j) rate at which monomer M is consumed by propagation reactions 4 Ibrehem, A.S., Hussain, M.A., Ghasem, N.M., “Mathematical model
at sites of type j and advanced control for gas-phase olefin polymerization in
Chin. J. Chem. Eng., Vol. 21, No. 9, September 2013 1029

fluidized-bed catalytic reactors”, Chin. J. Chem. Eng., 16 (1), 84-89 temperature in industrial polyethylene fluidized bed reactor”, Chem.
(2008). Eng. Res. Des., 84 (2), 97-106 (2006).
5 Zheng, Z.W., Shi, D.P., Su, P.L., Luo, Z.H., Li, X.J., ‘‘Steady-state 25 Gu, B., Gupta, Y.P., ‘‘Control of nonlinear processes by using linear
and dynamic modeling of the base multireactor olefin polymeriza- model predictive control algorithms’’, ISA T., 47 (2), 211-216 (2008).
tion process’’, Ind. Eng. Chem. Res., 50 (1), 322-331(2011). 26 Shamiri, A., Hussain, M.A., Mjalli, F.S., Mostoufi, N., Shafeeyan
6 Luo, Z., Cao, Z., Su, Y., “Monte carlo simulation of propylene po- M.S., “Dynamic modeling of gas phase propylene homopolymeriza-
lymerization (i) effects of impurity on propylene polymerization”, tion in fluidized bed reactors”, Chem. Eng. Sci., 66 (6), 1189-1199
Chin. J. Chem. Eng., 14 (2), 194-199 (2006). (2011).
7 Vahidi, O.S.M., Mirzaei, A., “Control of a fluidized bed polyethyl- 27 Cui, H., Mostoufi, N., Chaouki, J., “Characterization of dynamic
ene reactor”, Iran. J. Chem. Chem. Eng., 27 (3), 87-101 (2008). gas-solid distribution in fluidized beds”, Chem. Eng. J., 79 (2),
8 Choi, K.Y., Ray, W.H., “The dynamic behavior of continuous 133-143 (2000).
stirred-bed reactors for the solid catalyzed gas phase polymerization 28 Cui, H., Mostoufi, N., Chaouki, J., “Gas and solids between dynamic
of propylene”, Chem. Eng. Sci., 43 (10), 2587-2604 (1988). bubble and emulsion in gas-fluidized beds”, Powder Technol., 120
9 Ali, E., Al-Humaizi, K., Ajbar, A., “Multivariable control of a simu- (1-2), 12-20 (2001).
lated industrial gas-phase polyethylene reactor”, Ind. Eng. Chem. 29 Shamiri, A., Hussain, M.A., Mjalli, F.S., “Two phase dynamic model
Res., 42 (11), 2349-2364 (2003). for gas phase propylene copolymerization in fluidized bed reactor”,
10 McAuley, K.B., MacGregor, J.F., “Optimal grade transitions in a gas Defect. Diffus. Forum, 312-315, 1079-1084 (2011).
phase polyethylene reactor”, AIChE J., 38 (10), 1564-1576 (1992). 30 Shamiri, A., Hussain, M.A., Mjalli, F.S., Mostoufi, N., “Kinetic
11 McAuley, K.B., Macgregor, J.F., “Nonlinear product property con- modeling of propylene homopolymerization in a gas-phase fluid-
trol in industrial gas-phase polyethylene reactors”, AIChE J., 39 (5), ized-bed reactor”, Chem. Eng. J., 161 (1-2), 240-249 (2010).
855-866 (1993). 31 Kiashemshaki, A., Mostoufi, N., Pourmahdian, S., Sotudeh-Gharebagh,
12 Van Brempt, W., Backx, T., Ludlage, J., Van Overschee, P., De Moor, S., “Two phase modeling of a gas phase polyethylene fluidized bed
B., Tousain, R., “A high performance model predictive controller: reactor”, Chem. Eng. Sci., 61 (12), 3997-4006 (2006).
Application on a polyethylene gas phase reactor”, Control Eng. 32 Lucas, A., Arnaldos, J., Casal, J., Puigjaner, L., “Improved equation
Pract., 9 (8), 829-835 (2001). for the calculation of minimum fluidization velocity”, Ind. Eng.
13 Sarvaramini, A., Mostoufi, N., Sotudeh-Gharebagh, R., “Influence of Chem. Proc. Dev. Des., 25 (2), 426-429 (1986).
hydrodynamic models on dynamic response of the fluidized bed 33 Kunii, D., Levenspiel, O., Fluidization Engineering, 2nd edition,
polyethylene reactor”, Int. J. Chem. React. Eng., 6 (1), 1542-1580 Boston, MA, Butterworth-Heinmann (1991).
(2008). 34 Mostoufi, N., Cui, H., Chaouki, J., “A comparison of two- and sin-
14 Seki, H., Ogawa, M., Ooyama, S., Akamatsu, K., Ohshima, M., Yang, gle-phase models for fluidized-bed reactors”, Ind. Eng. Chem. Res.,
W., ‘‘Industrial application of a nonlinear model predictive control to 40 (23), 5526-5532 (2001).
polymerization reactors’’, Control Eng. Pract., 9 (8), 819-828 35 Hilligardt, K., Werther, J., “Local bubble gas hold-up and expansion
(2001). of gas/solid fluidized beds”, German Chem. Eng., 9 (4), 215-221
15 Ali, E.M., Abasaeed, A.E., Al-Zahrani, S.M., “Improved regulatory (1986).
control of industrial gas-phase ethylene polymerization reactors”, 36 Cai, P., Chen, L., van Egmond, J., Tilston, M., “Some recent ad-
Ind. Eng. Chem. Res., 38 (6), 2383-2390 (1999). vances in fluidized-bed polymerization technology”, Particuology, 8,
16 Ali, E.M., Abasaeed, A.E., Al-Zahrani, S.M., “Optimization and (6) 578-581 (2010).
control of industrial gas-phase ethylene polymerization reactors”, 37 Shamiri, A., Hussain, M.A., Mjalli, F.S., Mostoufi, N., “Improved
Ind. Eng. Chem. Res., 37 (8), 3414-3423 (1998). single phase modeling of propylene polymerization in a fluidized
17 Ali, E.M., Ali, M.A.H., “Broadening the polyethylene molecular bed reactor”, Comput. Chem. Eng., 36 (6), 35-47 (2012).
weight distribution by controlling the hydrogen concentration and 38 McAuley, K.B., MacGregor, J.F., Hamielec, A.E., “A kinetic model
catalyst feed rates”, ISA T., 49 (1), 177-187 (2010). for industrial gas-phase ethylene copolymerization”, AIChE J., 36
18 Hassimi, A., Mostoufi, N., Sotudeh-Gharebagh, R., “Unsteady-state (6), 837-850 (1990).
modeling of the fluidized bed polyethylene reactor”, Iran. J. Chem. 39 Luo, Z.H., Su, P.L., Shi, D.P., Zheng, Z.W., “Steady-state and dy-
Eng., 6 (1), 23-39 (2009). namic modeling of commercial bulk polypropylene process of hypol
19 Chatzidoukas, C., Perkins, J.D., Pistikopoulos, E.N., Kiparissides, C., technology”, Chem. Eng. J., 149 (1-3), 370-382 (2009).
“Optimal grade transition and selection of closed-loop controllers in 40 Floyd, S., Choi, K.Y., Taylor, T.W., Ray, W.H., “Polymerization of
a gas-phase olefin polymerization fluidized bed reactor”, Chem. Eng. olefins through heterogeneous catalysis. III. Polymer particle model-
Sci., 58 (16), 3643-3658 (2003). ling with an analysis of intraparticle heat and mass transfer effects”,
20 Sato, C., Ohtani, T., Nishitani, H., “Modeling, simulation and J. Appl. Polym. Sci., 32 (1), 2935-2960 (1986).
nonlinear control of a gas-phase polymerization process”, Comput. 41 Mjalli, F.S., Al-Asheh, S., “Neural-networks-based feedback lin-
Chem. Eng., 24 (2-7), 945-951 (2000). earization versus model predictive control of continuous alcoholic
21 Salau, N.P.G., Neumann, G.A., Trierweiler, J.O., Secchi, A.R., “Dy- fermentation process”, Chem. Eng. Technol., 28 (10), 1191-1200
namic behavior and control in an industrial fluidized-bed polymeri- (2005).
zation reactor”, Ind. Eng. Chem. Res., 47 (16), 6058-6069 (2008). 42 Seborg, D.E., Edgar, T.F., Mellichamp, D.A., Process Dynamics and
22 Bolsoni, A., Lima, E.L., Pinto, J.C., “Advanced control of propylene Control, 2nd edition, John Wiley & Sons, New Jersey (2004).
polymerizations in slurry reactors”, Braz. J. Chem. Eng., 17 (4-7), 43 Shridhar, R., Cooper, D.J., “A tuning strategy for unconstrained
565-574 (2000). SISO model predictive control”, Ind. Eng. Chem. Res., 36 (3),
23 Ghasem, N.M., “Dynamic behavior of industrial gas phase fluidized 729-746 (1997).
bed polyethylene reactors under PI control”, Chem. Eng. Technol., 44 Ziegler, J.G., Nichols, N.B., Rochester, N.Y., “Optimum settings for
23 (2), 133–145 (2000). automatic controllers”, Trans. ASME, 64 (8), 759-768 (1942).
24 Ghasem, N.M., “Design of a fuzzy logic controller for regulating the

View publication stats

S-ar putea să vă placă și