Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Department of Chemical Engineering & Energy Sustainability

Faculty of Engineering

KNC 3213 PROCESS CONTROL SYSTEM

GROUP PROJECT (GROUP 10)


POLYMERIZATION OF POLYPROPYLENE IN FLUIDIZED BED REACTOR

Jannatul Najihah Binti Mohamed Johari (63268)


Muhammad Hafiezeen Hisham Bin Mohd Khairul Hisham (63453)
Nurul Ainil Fatihah Binti Haidir (63700)
Nurul Aishah Farhani Binti Mohd Fuad (63701)
Puteri Nur Camielya Binti Kamarulzaman (63742)

Bachelor of Engineering with Honours


(Chemical Engineering)
2019
ABSTRACT

Polyolefin is one of the most famous thermoplastic polymers that derived from olefins.
It is also an important base material that can be used for various consumer products. The
example of commodity plastics from polyolefin is polypropylene (PP) and polyethylene (PE).
The most economical methods for the production of commodity polymers for olefin
polymerization is by using Fluidized Bed Reactor (FBR). In FBR, polymerization occurs by
the presence of catalyst, which is Ziegler-Natta and triethyl aluminum as its co-catalyst. For
olefin polymerization in the gas phase, two types of main mechanism arise in the process which
are free-radical mechanism and coordination mechanism. Unfortunately, there are several
problems from this polymerization process which is the worldwide plastic growth is forecast
between 5% and 5.5% per annum over the next few years, leading to a worldwide demanding
rising to around 400 million tons by 2020. In view of this high demand for polyolefin, it is
important to evaluate and improve the commodities production in Fluidized Bed Reactor
(FBR). Thus, in this process we need to observe the dynamic profile of the polymerization of
polypropylene in FBR, solve the mathematical modeling of FBR and also build and compute
the proposed Model Predictive Control (MPC). Based on the results obtained, the decreasing
concentrations in both emulsion and bubble phase are due to the high superficial gas velocity
and the increasing amount of catalyst used in the reaction. However, the effect of superficial
gas velocity on the bubble phase temperature increases after injection of the catalyst as the
polymerization reaction starts due to the progress of exothermic reaction of propylene
polymerization. Moreover, MPC controller provides the best response in the comparison of
performance between MPC and PID. MPC has been able to achieve the best set point with no
overshoot and oscillatory curve.

ii
Table of Content

Topic Page

ABSTRACT ii
1.0 INTRODUCTION 1
1.1 Background Study 1
1.2 Problem Statement 2
1.3 Objectives 2
2.0 LITERATURE REVIEW 3
3.0 METHODOLOGY 9
3.1 Mathematical Modelling 9
3.2 Model Predictive Control (MPC) 10
4.0 RESULT & DISCUSSION 12
4.1 Process Modelling & Result Analysis 12
4.2 Process Control System & Result Analysis 15
4.2.1 Proportional Integral Derivative (PID) 15
4.2.2 Model Predictive Control (MPC) 17
5.0 CONCLUSION 19
6.0 REFERENCES 20
Appendix A: Two Phase Model 24
Appendix B: Running Model 31

iii
List of Figure

Topic Page

Figure 2.1: Model predictive control system 6


Figure 3.1: The flowchart of kinetic model of propylene polymerization 9
Figure 3.2: Flowchart of the MPC diagram of polymerization process 11
Figure 4.1.1: Monomer concentration in emulsion phase 12

Figure 4.1.2: Monomer concentration in bubble phase 13

Figure 4.1.3: Concentration profile of hydrogen emulsion inside of reactor 13

Figure 4.1.4: Concentration profile of hydrogen bubble inside of reactor 14


Figure 4.1.5: Effect of superficial gas velocity on the emulsion phase temperature 14

Figure 4.1.6: Effect of superficial gas velocity on the bubble phase temperature 15

Figure 4.2.1.1: Response curve for PID controller having a unit step change 15
Figure 4.2.1.2: Block diagram of Simulink for PID controller 16
Figure 4.2.2.1: Response curve for MPC controller having unit step change 17
Figure 4.2.2.2(a): Block diagram of Simulink for MPC controller 18
Figure 4.2.2.2(b): Block diagram of Simulink for unit step change MPC controller 18

iv
1.0 INTRODUCTION

1.1 Background Study

Polyolefin is one of the most famous thermoplastic polymers that derived from olefins.
It is also an important base material that can be used for various consumer products. The
example of commodity plastics from polyolefin is polypropylene (PP) and polyethylene (PE).
Polypropylene (PP) is widely used in various applications such as household products, fibers
and fabrics, automotive, packaging and medical applications. In addition, PP is an important
and growing use in the automotive industry (Tolinski, 2010). For automotive applications, 8%
of the world’s polypropylene production was used in 2007. This is because the replacement of
metal elements with lightweight, but strong polypropylene improves the fuel efficiency. Thus,
application in the transportation sector will continue to grow strongly due to the need to
minimize the vehicle weight which contributing to an improved fuel economy. Furthermore,
the application of polypropylene is very useful in the packaging industry. The use of low-
density plastics such as polypropylene decreases weight per packaging unit, reduces shipping
charges, minimizes energy consumption and cut downs the packaging contribution to urban
solid waste (Carlson, 2009). These widespread applications are expected to have a progressive
impact on the global polyolefin market.

According to (Abbasi et al., 2016), they have been identified that the most economical
methods for the production of commodity polymers for olefin polymerization is by using
Fluidized Bed Reactor (FBR). In FBR, polymerization occurs by the presence of catalyst,
which is Ziegler-Natta and triethyl aluminum as its co-catalyst. The unreacted gas leaves the
reactor’s top and is then compressed and cooled before being pumped back into the FBR’s
bottom. The limitation of the production rate of this process is the heat removal from the
circulating gas due to the highly exothermic reaction of polymerization (Shamiri et al., 2013).
There are several advantages that can be found in the usage of FBR (Werther, 2000). First, the
properties of FBR are similar to liquid which make it easier to handle and transport the solid.
Next, the mixing of solid is intense where there are no hot spots, which it gives a uniform
temperature distribution and solid production throughout the FBR. The heat-transfer
coefficients between bed and immersed heating or cooling surfaces in FBR are high. However,
there are some disadvantages occur in the FBR which is it required an expensive solids
separations or gas purification equipment due to solids entrainment by fluidizing gas (Werther,

1
2000). Furthermore, due to agglomeration of solid in the FBR, the possibility of defluidization
to occur is high.

1.2 Problem Statement

According to Biron (2016), the worldwide plastic growth is forecast between 5% and
5.5% per annum over the next few years, leading to a worldwide demanding rising to around
400 million tons by 2020. In view of this high demand for polyolefin, it is important to evaluate
and improve the commodities production in Fluidized Bed Reactor (FBR). It can be a
challenging task due to extreme nonlinearities in the dynamics of the reactor. This is due to the
complicated reaction, heat and mass transfer, controlling variables such as temperature and
concentration, and solid characteristics that occur in the reactor.

1.3 Objective

There are two objectives in this study:

1. To observe the dynamic profile of the polymerization of polypropylene in Fluidized


Bed Reactor (FBR).
2. To build and compute the proposed Model Predictive Control (MPC).

2
2.0 LITERATURE REVIEW

In research by Atan et.al., (2019), stated that there are varies types of polyolefin formed
via the former mechanism known as fluidized bed reactor (FBR), vertical-stirred reactor, and
horizontal-stirred-bed reactor but utmost of the mathematical models was replicated by
expending a fluidized bed reactor. For olefin polymerization in the gas phase, two types of
main mechanism arise in the process which are free-radical mechanism and coordination
mechanism. To define the dynamic behavior of the olefin polymerization process in gas phase,
several mathematical models have been applied. The modelling can occur in several type of
phases through FBR depending on the assumption of the sites of the reaction and location
during the process. In the research, the single-phase is acknowledged as a well-mixed model
where the reaction were implicit to happen in the emulsion phase, for two-phase model the
reaction were presumed occurred in the emulsion and bubble phase and for the third stage
prototype, it is presumed that the reaction happened in bubble, cloud and emulsion phase. To
simplify the complexity of various stages in the FBR, the extra phases were ignored during the
model implementation. The FBR was used to extract the olefin polymerization reaction by
using several types of catalyst that suitable with the reactor such as Ziegler-Natta (Z-N)
catalyst. Based on experiment by authors, there would be slightly differed in calculated
mathematical simulations due to the assumption to simplify the intricacy of the nonlinearity
phenomena and the amount of phases in the reaction mechanism and also those models need
to be validated more to lessen the difference between model production and manufacturing
data to improve the end products to be not harmful to the environment and have high quality
standard.

Previous study by Shamiri et.al., (2014), stated that the polymerization reaction was
deliberated to form in bubble and the emulsion phase in two-phase model. The experimental
data in modelling is near to expected data of two-phase model especially at longer elapsed time.
The two-phase model provides more realistic result by seeing the occurrence of solid in bubbles
and emulsion phase presence at condition beyond the lowest fluidization. Hydrodynamics of
the reactor approach the well-mixed condition through the early stages of fluidization though
this mechanism becoming impractical in the reaction since extra formation of bubbles and the
decreasing in the degrees of mass and heat transfer. Then, rate of reaction and temperature was
increasing due to the occurrence of a greater amount of catalyst. In this research the reactions
of olefin polymerization were studied by designing a model that was authorized with data
produced in a pilot plant. The study done by researcher using well-mixed models and the pilot

3
plant data shows that the well-mixed models underrate the polypropylene concentration and
emulsion phase temperature.

Further research by Abbasi, Shamiri & Hussain (2016), found the model was accurate
in estimate the steady and stabilize production rate very close to the real data with small
deviation. The author succeeds to increase 20% of production rate in contrast with
Kiashemshaki et al. (2006) due to overview of recycled elutriated solids into the reactor.
However, when the apparent gas velocity was increased, gas through the bed become faster
and more solid particle were transport. The polymer particle and amount of catalyst available
in the reactor bed will be decreased and therefore lead to slower rate of reaction due to reduce
in reaction extent. Moreover, Author also introduced a dynamic calculating model in terms of
calculate both temperature and comonomer concentration which does not include in research
Kiashemshaki (2006).

In research by Ho et al. (2012), it is stated that the gas phase polymerization reactions
of propylene occur in bubble and emulsion phase. Hence, the reactor bed temperature must be
over the dew point purpose of the reactants to dodge the build-up of gas inside the reactor.
Simultaneously, the reactor bed temperature should also be underneath the liquefying in order
to anticipate molecule softening the polymer, accumulation and thusly reactor shut down.
Because of the excessive nonlinearities associated with the elements involved in the reactor, it
is in this way past the capacity of the customary PID controller with fixed controller settings
to accomplish incredible control of the reactor factors. Therefore, a devolved Two Inputs Two
Outputs (TITO) Adaptive Predictive Model-Based Control (APMBC) strategy was employed.
As it is being compared to the other two controllers which are the conventional PI controllers
tuned using the Internal Model Control (IMC) method’ and ‘the standard Ziegler-Nichols (Z-
N) method, it is able to track set-point properties effectively and able to synchronize the process
disturbance efficiently. It is common for the polymerization processes to have a highly
nonlinear dynamic behavior leading to an abysmal performance of controllers based on
conventional internal models to be poor or for it to need a considerable amount of effort in
controller tuning (Abbasi et al., 2018). Nonetheless justifiable that such a good performance
resulting from an advanced control scheme can be obtained at expense of increased
computational load.

The focal point of the literature in terms of modelling the operation of a catalytic olefin
polymerization in FBRs are the mixing conditions and the number of phases existing in the

4
bed. The existence of many phases that carry their own chemical reactions and interphase heat
and mass transfer in various heterogeneous polymerization systems, give out the result in an
inclusive and representative model that will account for all these complex gas and solid flows,
mechanisms of mass and heat transfers and polymerization kinetics (Abbasi et al., 2018).
McAuley et al. (1994) and Xie et al. (1994) describe fluidized-bed polyolefin reactors by
introducing a well-mixed reactor. This approach and simple two-phase models in steady state
are being compared and conclusively said that the latter shows very little errors when it predicts
the monomer concentration and reactor temperature. Choi and Ray (1985) recommended the
simple two-phase model for reactor in two phases of emulsion and bubble. They conclude on
the solid-free bubbles which shows that polymerization only takes place in the emulsion phase.
McAuley et al. (1990) recommended a much simpler well-mixed model and further predicted
that between the emulsion and bubble phases there is an unlimited heat and mass transfer.

It should be said that in dealing with the strict boundaries enforced in a few industrial
processes many conventional control algorithms are not sufficient, specifically once a first-rate
commodity is needed. This is particularly applying for polymerization processes in which fixed
properties like average molecular weight with the effect on plastic quality should be satisfied.
Therefore, an essential part for the process control structure is to adopt a model predictive
control (MPC) as a proper way, that makes use of a process tailored dynamic model. As
Campello et al. (2003) mentioned, to deal with limitations that involve the input and output
variables in procedures MPC algorithms have been used for chemical process control through
their usage simplicity and capability. The MPC is an optimization-based control approach
which is very suitable for multi-variable yet constrained processes and can also predicts the
actual system’s future behaviour over a time period defined by the prediction prospect (Abbasi
et al., 2018). Figure 2.1 shows the implication made by the simpler diagram of the MPC
algorithm that is a parallel process model to the controller is used in MPC to predict the
controlled variable.

5
Figure 2.1: Model predictive control system (Shamiri et al., 2013)

According to Calvert & Handwerk (1973), the standard expansion size for all particle
are standard at initial condition due to the very narrow dispensation’s diameter of the support
particle used to produce the catalyst and the feeding devices with specific design. The
separation device is assumed as the only exit way from the reactor for the catalyst and polymer
particles. Then, for the separation system, their efficiency was considered as 100% and had
neglect all elutriation effects (Aronson, 1983). Inside the reactor, there have separator which
only allow the catalyst and polymer particles with specific maximum size. The triangle pattern
that used for simulation function required extra computational intention to handle the sharp
irregularities at the triangle vertices. Based on result, the fluidized bed using continuous and
differentiable function model has smooth result for sample balance. By design the separation
process correctly, the models can reach the desired reactor operation condition hence will
enhance the performance of catalyst (Grosso & Chiovetta, 2005)

In dynamic 2 phase model, emulsion and bubble phase are assumed to happen in
polymerization reactions. Insulation was implemented around reactor body to reduce the heat
loss to the surrounding (Kleinhenz, et.al., 2009). The feed gas was assumed free from
contamination thus poisoning reaction can be neglect. Elutriation of solid particle from the
reactor was also neglected (Shamiri et.al., 2014) and [19-Kashiaki]. Two-phase model shows
more realistic result because the solid exist in bubble and emulsion phase at condition beyond
the minimum fluidization (Shamiri et.al., 2014). In polymerization process, the exothermic
reaction of polymerization occur hence will rise the reactor temperature (Arnold, 2015).
Shamiri, Hussain, Mjalli, Shafeeyan, & Mostoufi (2014) conclude that, by doing assumption
on operation condition, the result will not accurate in describe the dynamic of gas-solid
6
distribution on the rate of heat and mass transfer and chemical reaction in FBR. However, if a
good assumption and correlation coefficient is applied in this model, the simulation result will
close to experimental data. The Solid entrainment, inert gas and particle size distribution
movement in FBR will give varies temperature between the predicted model and industrial
model.

The equation for consumption rate of each component can be evaluate by using moment
equation by assumed the monomer will eliminate by propagation reaction (McAuley,
MacGregor & Hamielec, 1990). In modelling polymerization, the common method was used
is the moment’s method, since the developer can make prediction on the polymer’s properties.
Then, semi-dynamic also can use to foretell the polymer’s properties but fail to generate
dynamic temperature and comonomer concentration profile in the model (Kiashemshaki,
Mostoufi & Sotudeh-Gharebagh, 2006). According to Rhodes (2008), when the velocity of
fluidizing gas is higher than terminal velocity, the granular particle will slip off from the bed.
To achieve more realistic result, the solid entrainment cannot be neglect and need to inspect in
the model for mass and energy balance (Abbasi, Shamiri & Hussain, 2016).

In the beginning, design and modelling of a gas-phase polymerization reactor was built
to figure out the dynamic behavior of FBR for ethylene and propylene polymerization (Choi &
Harmon, 1985). Ziegler-Natta catalyst was used in model gas-phase polymerization FBR
(Ibrehem, Hussain & Ghasem, 2009). Meier, Weickert, Pater & Westerterp (1999), stated that
although small amount of catalyst was used in FBR for polymerization, it can give impact to
temperature gradient. Huang & Xie (2014), introduced a mathematical model that can used to
foretell the molecular weight distribution. Catalyst poison can reduce the propagation rates in
polymerization process, so alkyl aluminum was added to eliminate much of polar poison
(Abubaker & Mustafa, 2016).

The mathematical model was interpreted in the forms of differential and algebraic
equations and contain 3 main aspects which are kinetic, transport and hydrodynamic
phenomenon (Atan et al. 2019). The modelling for olefin polymerization process is complex
because of its high nonlinearity process dynamic due to complexity of the reaction mechanism,
heat transfer phenomena and physical properties of the flow behavior in solid and gas. Heat
transfer become complex because of gas phase consist of different phase such as solid,
emulsion, bubble and cloud (Ibrehem, Hussain & Ghasem, 2009). Method of moments and
population balance modeling can use to identify physicochemical properties because of their

7
capability in determining a broad range of a complicated dynamic polymerization process
(Kiparissides et al. 2004; Ramkrishna & Mahoney, 2002). Other than used co-catalyst to
activate the potential active sites, Salau, Nuemann, Trierweiler & Secchi (2008) introduces 2
further stages that can activating the site which are spontaneous site activation and site
activation using hydrogen. To control the molecular weight and the chain length, site
transformation reaction at a supplementary stage was suggested (Zacca & Debling al. 1996;
Harshe et al. 2004). According to Shamiri et al. (2011), since stiffness differential equation
used in this polymerization process, the best numerical method can be used are Backward
Euler’s and Rosenbrock methods.

8
3.0 METHODOLOGY

3.1 Mathematical Modelling

A kinetic model in a gas-phase fluidized-bed reactor for the production of


polypropylene was developed in this work. Figure 3.1 below shows the flowchart of the kinetic
model for this polymerization process. Firstly, a reaction mechanism is needed in this process
in order to help in determining the kinetic model. The reaction mechanism that involves is
activation, initiation and propagation. The kinetic model itself consists of catalyst flowrate,
reaction rate constant and energy activation at specific temperature of the process. The catalyst
feed rate plays important role as an operating parameter especially in the fluidized bed reactors
of propylene operation. Hydrogen flowrate, monomer flowrate and catalyst concentration
should be considered in the kinetic model as well.

Reaction Mechanism

Kinetic Model

Mass Balance Model

Energy Balance
Do not
converge
Compute the Model

Converge

End

Figure 3.1: The flowchart of kinetic model of propylene polymerization

9
Mass balance model is another model that is important in this polymerization process.
This model includes the composition of raw materials such as olefin itself. The mass balance
also involves the Physical Chemical properties of the raw material and the volume
specification. After the model have been produced, we will proceed to the next step which is
energy balance. This step involves the transport phenomena and hydrodynamic correlation.
Next, the model is ready to be computed. The coding should be related to the model and control
in order to make the model converge. Once it is converged, the process has come to the end of
the polymerization process. If the model does not converge, the process of kinetic modelling
should be repeated again until it is converged, and the final product is formed.

3.2 Model Predictive Control (MPC)

Model Predictive Control (MPC) itself consists of two types which is the Linear
Predictive Control (LMPC) and Nonlinear Model Predictive Control (NMPC). In this process
model, we will focus more on the NMPC as it is similar to the linear counterpart except the
process prediction and optimization of nonlinear dynamic models that is used. Figure 3.2 below
shows the flowchart of MPC diagram in this polymerization process. Firstly, the process data
should be determined before proceeding to the next step. Suitable parameters will be
considered in the process data such as bubble velocity, emulsion velocity, mass transfer
coefficient and others. Next, we need to choose a reactor type in this process and Fluidized-
bed reactor (FBR) is chosen. After that, a model is determined which is a two-phase model.

After the reactor type and the model have been chosen, we will proceed this reaction
with determination of the control algorithm. PI (Proportional-Integral) controller and NMPC
controller is chosen to be observed. Next, we proceed in choosing the controlled variables
which is the temperature and monomer concentration. The model is then ready to compute and
to see whether it is converged or not. Once it is converged, the process has come to the end of
the polymerization process. If the model does not converge, the process of MPV modelling
should be repeated again until it is converged, and the final product is formed.

10
Process Data

Reactor Type

Model

Control Algorithm

Does not Variables Controlled


converge

Compute the Model

Converge

End

Figure 3.2: Flowchart of the MPC diagram of polymerization process

11
4.0 RESULT & DISCUSSION

4.1 Process Modelling and Result Analysis

Figure 4.1.1 below shows the monomer concentration of the emulsion phase shows the
highest concentration which is about 502.59051 mol/m3 at the very first second of the reaction.
The concentration of the monomer is then started to decrease hastily down the graph at about
41 minutes (2500s) until 66 minutes (4000s) with the concentration of about 502.59016
mol/m3. The monomer concentration is then continuously remained to be at steady state
throughout the reaction. When the velocity of the superficial gas is increase, the emulsion phase
velocity will be decreasing. As a result, the concentration of polymer in emulsion phase become
reduced.

Figure 4.1.2 below shows the concentration of monomer in bubble phase decrease
rapidly down the graph at approximately 50 minutes (3000s). The concentration of monomer
seems to decrease approximately from 504 mol/m3 to 375 mol/m3 due to the starting point of
the polymerization reaction. Like the emulsion phase in Figure 4.1.1, the concentration of
polymer in bubble phase is decrease because of high superficial gas velocity. Theoretically,
when superficial gas velocity is increase, it will promote more solid and gas particle within the
bubbles. However, if the superficial gas velocity is too high, it will disturb the interaction of
the solid catalyst inside the bubble phase and resulting the lower polymer production.

Another reason of decreasing concentration in both figures is because of when the


amount of catalyst used is increase, the monomer concentration for emulsion and bubble phase
are decrease due to the increase in the reaction rate.
Concentration (mol/m3)

Time (s)

12
Figure 4.1.1: Monomer concentration in emulsion phase

Concentration (mol/m3)

Time (s)

Figure 4.1.2: Monomer concentration in bubble phase


Hydrogen gas is normally used to regulate the molecular weight and its distribution and
also regulate the product grade transition in a pilot-scale propylene polymerization plant. The
concentration of hydrogen profile versus time is shown above. From both figures, it can be
seen that the hydrogen concentration decreases at the start of polymerization reaction due to its
consumption via the mechanism of transfer to hydrogen reaction. From Figure 4.1.3, the
concentration of hydrogen in emulsion phase decrease rapidly at approximately 35 minutes
(2000s). Same goes to the Figure 4.1.4 where the concentration of hydrogen in bubble phase
also decrease rapidly at approximately 35 minutes (2000s). The differences between these two
figures is just that the concentration of hydrogen in emulsion phase decrease from
approximately 55.0294 mol/m3 to 55.0264 mol/m3, while the concentration of hydrogen in
bubble phase decrease from approximately 55 mol/m3 to 2 mol/m3.
Concentration (mol/m3)

Time (s)

13
Figure 4.1.3: Concentration profile of hydrogen emulsion inside of reactor

Concentration (mol/m3)

Time (s)

Figure 4.1.4: Concentration profile of hydrogen bubble inside of reactor


In the Figure 4.1.5, it is shown that the highest temperature is achieved at the beginning
of the reaction at approximately 317.1605 K. Then, due to the polymerization reaction the
temperature of the reactor decreases slightly from 317.1605 K to 317.1596 K approximately.
The temperature then continuously remained at steady state constantly throughout of the
reaction. The reducing temperature of emulsion phase is due to increasing of the superficial
gas velocity where convective cooling by the gas become higher. This model is working below
the industrial safety limit of 353.15 K (80°C) which the reactor does not undergo temperature
runaway. If the temperature of the reactor increases beyond the safety limit, it will cause a
particle melting, agglomeration and resulting in reactor shutdown.
Temperature (K)

Time (s)

Figure 4.1.5: Effect of superficial gas velocity on the emulsion phase temperature

14
Figure 4.1.6 below shows the graph that temperature start rising rapidly after injection
of the catalyst and reaches its steady value after about approximately 50 minutes (3000s). As
expected in this system, the reactor temperature increases from approximately 320 K to 865 K
as the polymerization reaction starts due to the progress of exothermic reaction of propylene
polymerization. The temperature is then continuously remained at the maximum temperature
of 865 K. The energy is transferred between the phases because of temperature rising.
Temperature (K)

Time (s)

Figure 4.1.6: Effect of superficial gas velocity on the bubble phase temperature
4.2 Process Control System & Result Analysis

4.2.1 Proportional Integral Derivative (PID) Controller

Figure 4.2.1.1: Response curve for PID controller having a unit step change

15
This closed loop response is set at P = 30, I = 10 and D = 10, and the set point is been
changed from 0 to 1. From Figure 4.2.1.1, the closed loop response is less sluggish as the
presence of oscillatory is at minimum. There is an overshoot happen at the early stage of set
point change. The presence of oscillatory happens to be near the maximum overshoot point.
Thus, the response is a bit aggressive because during the overshoot happening there shown
some oscillatory in the response curve. This response from Figure 4.2.1.1 shown that the
maximum deviation is in a range of +0.45. The response curve also shown that the settling time
is fast as it takes around 10 s to 15 s to reach the set point line. Thus, this closed loop response
able to achieve the satisfactory set point responses as it manage to reach the set point line in a
short period despite having an overshoot with a presence of minimum oscillatory. Figure
4.2.1.2 show the block diagram of Simulink used in order to achieve response curve shown in
Figure 4.2.1.1.

Figure 4.2.1.2: Block diagram of Simulink for PID controller

16
4.2.2 Model Predictive Control (MPC) Controller

Figure 4.2.2.1: Response curve for MPC controller having unit step change

This closed loop set point is been changed from 0 to 1. The closed loop response shown
in Figure 4.2.2.1 has the best set point response. There is no overshoot happen throughout the
response. Besides, the response also not shown any presence of oscillatory in the response
curve. Thus, this response has zero maximum deviation of curve. The response curve shown
the faster settling time as it takes only 0 s to 1 s to reach the set point line. Therefore, this closed
loop response able to achieve the better set point responses as it manage to reach the set point
line in a very short period compared to the response curve for PID controller shown in Figure
4.2.1.1, as there are no occurrence of the overshoot and oscillatory in the response curve. Figure
4.2.2.2(a) and Figure 4.2.2.2(b) below shown the block diagram of Simulink used for this MPC
controller in order to achieve response curve shown in Figure 4.2.2.1.

17
Figure 4.2.2.2(a): Block diagram of Simulink for MPC controller

Figure 4.2.2.2(b): Block diagram of Simulink for unit step change MPC controller

18
5.0 CONCLUSION

In conclusion, the project had successfully achieved the objectives. The dynamic profile
of the polymerization of polypropylene in Fluidized Bed Reactor (FBR) can be observed
through the graph in the section 4.1. The monomer concentration of the emulsion phase (Figure
4.1.1) shows that the concentration is decreasing and remained to be at the steady state
throughout the reaction. Similarly, the concentration of monomer in bubble phase (Figure
4.1.2) also decreases. The decreasing concentrations in both figures are due to the high
superficial gas velocity and the increasing amount of catalyst used. In addition, the
concentration of hydrogen gas has been used to regulate the molecular weight and its
distribution and to regulate the product grade transition in a pilot-scale propylene
polymerization plant. As the result, the concentration of hydrogen in emulsion phase (Figure
4.1.3) and bubble phase (Figure 4.1.4) are decreasing. Next, the temperature of emulsion phase
(Figure 4.1.5) decreases due to the increasing of the superficial gas velocity where convective
cooling by the gas becomes higher and the temperature then continuously remained at steady
state constantly throughout of the reaction. However, the effect of superficial gas velocity on
the bubble phase temperature (Figure 4.1.6) increases after injection of the catalyst as the
polymerization reaction starts due to the progress of exothermic reaction of propylene
polymerization.

Another objective that had successfully achieved in our study is to build and compute
the proposed Model Predictive Control (MPC), as shown in section 4.2. In the comparison of
performance between MPC and PID, MPC controller provides the best response. It has been
able to achieve the best set point with no overshoot and oscillatory curve. Moreover, the
response curve in MPC controller (Figure 4.2.2.1) had shown the faster settling time compare
to PID controller (Figure 4.2.1.1).

19
6.0 REFERENCES

Abbasi, M. R., Shamiri, A., & Hussain, M. A. (2016). Dynamic modeling and Molecular
Weight Distribution of ethylene copolymerization in an industrial gas-phase Fluidized-
Bed Reactor. Advanced Powder Technology, 27(4), 1526–1538.
https://doi.org/10.1016/j.apt.2016.05.014

Abbasi, M. R., Shamiri, A., & Hussain, M. A. (2019). A Review On Modelling And Control
of Olefin Polymerization In Fluidized-bed Reactors. Rev Chem Eng, 35(3), 311-333.

Abubaker, A. A., & Mustafa, M. A. (2016). Mathematical Modelling and Simulation of an


Industrial Propylene Polymerization Batch Reactor. Chemical Engineering and
Analytical Science, 1(1), 10–17.

Arnold, E. M. (2015). Temperature Control of an Emulsion Polymerization Process. Honors


Research Projects, 106.
Atan, M. F., Hussain, M. A., Abbasi, M. R., Hossain Khan, M. J., & Abdul Patah, M. F. (2019).
Advances in mathematical modeling of gas-phase olefin polymerization. Processes,
7(2), 1–33. https://doi.org/10.3390/pr7020067

Atan, M. F., Hussain, M. A., Abbasi, M. R., Hossain Khan, M. J., & Abdul Patah, M. F. (2019).
Advances in mathematical modeling of gas-phase olefin polymerization. Processes, 7(2),
1–33. https://doi.org/10.3390/pr7020067

Biron, M. (2016). Thermoplastics: Economic Overview. Practicle and Advances Information.


pp.22-111

Calvert, W. L., & Handwerk, R. H. (1973). United States Patent - 3.779.712. Particulate Solids
Injector Apparatus, 8.

Carlson A., (2009) International Polyolefins Conference 2009, Society of Plastics


Engineers, Houston,22-25.

Choi, K. Y., & Harmon Ray, W. (1985). The dynamic behaviour of fluidized bed reactors for
solid catalysed gas phase olefin polymerization. Chemical Engineering Science,
40(12),2261–2279. https://doi.org/10.1016/0009-2509(85)85128-9

Dashti, A., & Ramazani S.A., A. (2008). Modelling and Simulation of Olefin Polymerization
at Microstructure Level. Iran. J. Chem. Chem. Eng., 27(2).

20
Ghasem, N. M., Hussain, M. A., & Sata, S. A. (2006). Application of model predictive control
to batch polymerization reactor. Universitas Scientiarum, 11(1), 49-58.

Grosso, W. E., & Chiovetta, M. G. (2005). Modelling a fluidized-bed reactor for the catalytic
polymerization of ethylene: Particle size distribution effects. Latin American Applied
Research, 35(1), 67–76.

Huang, K., & Xie, R. (2014). Modeling of molecular weight distribution of propylene slurry
phase polymerization on supported metallocene catalysts. Industrial and Engineering
Chemistry, 20(1), 338–344. Retrieved from doi: 10.1016/j.jiec.2013.03.029

Harshe, Y. M., Utikar, R. P., & Ranade, V. V. (2004). A computational model for predicting
particle size distribution and performance of fluidized bed polypropylene reactor.
Chemical Engineering Science, 59(22–23), 5145–5156.
https://doi.org/10.1016/j.ces.2004.09.005

Ibrehem, A. S., Hussain, M. A., & Ghasem, N. M. (2009). Modified mathematical model for
gas phase olefin polymerization in fluidized-bed catalytic reactor. Chemical
Engineering Journal, 149(1–3), 353–362.

Kiashemshaki, A., Mostoufi, N., & Sotudeh-Gharebagh, R. (2006). Two-phase modeling of a


gas phase polyethylene fluidized bed reactor. Chemical Engineering Science.
https://doi.org/10.1016/j.ces.2006.01.042

Kiparissides, C., Alexopoulos, A., Roussos, A., Dompazis, G., & Kotoulas, C. (2004).
Population Balance Modeling of Particulate Polymerization Processes. Industrial &
Engineering Chemistry Research, 43, 7290–7302.

Kleinhenz, J., Yuan, Z., Sacksteder, K., & Caruso, J. (2009). Development of a reactor for the
extraction of oxygen and volatiles from lunar Regolith. 47th AIAA Aerospace Sciences
Meeting Including the New Horizons Forum and Aerospace Exposition, (January), 1–11.
https://doi.org/10.2514/6.2009-1203

McAuley, K. B., MacGregor, J.F., & Hamielec, A.E. (1990). A kinetic model for industrial
gas‐phase ethylene copolymerization. AIChe, 36(6).

McGregor, J. F. (1988). Control of polymerization reactors. Dynamics and Control of Chemical


Reactors and Distillation Columns, 31-35.

21
Ram, S. S., D. Dinesh Kumar, & B. Meenakshipriya. (2016). Designing of PID Controllers for
pH Neutraliztion Process. Indian Journal of Science and Technology, Vol 9(12).

Ramkrishna, D., & Mahoney, A. W. (2002). Population balance modeling. Promise for the
future. Chemical Engineering Science, 57(4), 595–606. https://doi.org/10.1016/S0009-
2509 (01)00386-4

Rhodes, Martin, J. (2008). Introduction to particle technology (2nd Edition). New York, United
States: John Wiley & Sons Inc

Robert, G., Aronson, S., Charleston, W., & Va. (1985). U. S. Patent Nov. 26, 1985. Fluidized
Bed Discharge Process, (19).

S. Moritomi, T. Watanabe, S. Kanzaki, Polypropylene Compounds for Automotive


Applications and reference therein. Retrieve from
http://www.sumitomochem.co.jp/english/rd/report/theses/docs/20100100_a2g.pdf

Salau, N, P, G., Nuemann, G, A., Trierweiler, J, O., & Secchi, A, R. (2008). Dynamic
Behaviour and Control in an Industrial Fluidized-Bed Polymerization Reactor. Industrial
& Engineering Chemistry Research, 47(16), 6058–6069.

Shamiri, A., Hussain, M. A., Mjalli, F. S., & Mostoufi, N. (2010). Kinetic modelling of
propylene homopolymerization in a gas-phase fluidized-bed reactor. Chemical
Engineering Journal, 240-249.

Shamiri, A., Azlan Hussain, M., Sabri Mjalli, F., Mostoufi, N., & Saleh Shafeeyan, M. (2011)
. Dynamic modeling of gas phase propylene homopolymerization in fluidized bed
reactors. Chemical Engineering Science, 66(6), 1189–1199.
https://doi.org/10.1016/j.ces.2010.12.030

Shamiri, A., Hussain, M. A., Mjalli, F. S., Mostoufi, N., & Hajimolana, S. (2013). Dynamic
and predictive control of gas phase propylene polymerization in fluidized bed reactors.
Chinese Journal of Chemical Engineering, 21(9), 1015-1029.

Shamiri, A., Hussain, M. A., Mjalli, F. S., Shafeeyan, M. S., & Mostoufi, N. (2014).
Experimental and modeling analysis of propylene polymerization in a pilot-scale

22
fluidized bed reactor. Industrial and Engineering Chemistry Research, 53(21), 8694–
8705. https://doi.org/10.1021/ie501155h

Tolinski, M. (2010), Plastics Engineering, vol. 66


Varshouee, G. H., Heydarinasab, A., Vaziri, A., & Zarand, S. G. (2019). A mathematical model
for investigating the effect of reaction temperature and hydrogen amount on the catalyst
yield during propylene polymerization. Kem. Ind, 68, 269-280.

Varshouee, H. G., Heydarinasab, A., Vaziri, A., & Roozbahani, B. (2018). Hydrogen Effect
Modelling On Ziegler-Natta Catalyst And Final Product Properties In Propylene
Polymerization. Bull. Chem. Soc. Ethiop., 32(2), 371-386

Weickert, G, B., Meier, G, B., Pater, J, T, M., & Westerterp, K, R. (1999). The particle as
microreactor: catalytic propylene polymerizations with supported metallocenes and
Ziegler-Natta catalysts. Chemical Engineering Science, 54(15–16), 3291–3296. Retrieved
from 10.1016/S0009-2509(98)00441-2

Werther, J. (2000). Fluidized-bed reactors. Ullmann’s encyclopedia of industrial chemistry.

Zacca, J. J., Debling, J. A., & Ray, W. H. (1996). Reactor residence time distribution effects
on the multistage polymerization of olefins - I. Basic principles and illustrative examples,
polypropylene. Chemical Engineering Science, 51(21), 4859–4886.
https://doi.org/10.1016/0009-2509 (96)00258-8

Zhang, Z., Chuang, Y.-Y., & Chen, J. (2014). Methodology of data reconciliation and
parameter estimation for process systems with multi-operating conditions
. Chemometrics and Intelligient Laboratory Systems, 110-119.

23
APPENDIX

Matlab Coding

A. Two Phase Model

function
dfdt=TwoPhaseModel(t,f,Mono,H2,Im,Fcat,U0,Tin,Tref,Press,yC3H6,yH2,yN2,
yC2H,yC6star)
%where t=independent variable, f=vector, Input=Monomer,H2,Catalyst flow
rate,velocity,inlet T,reference T,P,Coefficient
dfdt=zeros(44,1);

%Press=22; unit in bar


%T=72+273.15; unit in Kelvin K
R=8.314e-5; %unit in m3bar/molK
Z=0.9; %assume-non ideal gas
% Cg=Press/(t*Z*R); %unit in mol/m3
% N2=yN2*Cg;

% Concentration of co-catalyst(convert unit from kg/s to mol/m3)


Vreactor=0.0215; % Reactor volume
mw_Ti=47.9/1000; %unit in kg/mol
Al_ratio=1.01; %ratio of Al in Ti
Fcat_mol1=Fcat/mw_Ti; %molar flow rate of co-catalyst in mol/s
Fcat_mol2=Fcat_mol1*3600; %molar flow rate of co-catalyst in mol/h
Vgas=U0*3600; %Total gas injection in m3/h
Fcat=Fcat_mol2/Vgas; %unit in mol/m3
AlEt3=Fcat*Al_ratio; % Concentration of AlEt3 in mol/m3
AlEt3_e=0.88*AlEt3; % Concentration of AlEt3 in emulsion phase, in
mol/m3
AlEt3_b=0.12*AlEt3; % Concentration of AlEt3 in bubble phase, in mol/m3

% Umf, minimum fluidization velocity in m/s


% rho_g=23.45;
% rho_g depends on temperature hence need to determine the desired
rho_g
rho_s=910; % solid density in kg/m3
g=9.81; % gravitational accelertion in m/s2
mw_Mono=42.08e-3; %unit in kg/mol
mw_H2=2e-3; %unit in kg/mol
mw_N2=28.014e-3; %unit in kg/mol
mw_C2H=25.02934e-3; %unit in kg/mol
mw_C6star=92e-3; %unit in kg/mol
mw_avg=yC3H6*mw_Mono+yH2*mw_H2+yN2*mw_N2+yC2H*mw_C2H+yC6star*mw_C6star;
rho_g0=(Press*mw_avg)/(Z*Tin*R); % density of gas at Tinlet in kg/m3
T=72+273.15; % T=Treaction in K
rho_g=(rho_g0*Tin)/T; % desired density of gas at T in kg/m3
dp=500e-6; % density of particle in m
miug=1.14e-4; % viscosity dynamic in Pa.s
Ar=(rho_g*(rho_s-rho_g)*g*dp^3)/miug^2; % Archimedes number, no unit
Remf=sqrt((29.5)^2+0.357*Ar)-29.5; % No unit
Umf=(Remf*miug)/(dp*rho_g); % in m/s

delta=0.534*(1-exp(-(U0-Umf)/0.413));
eps_mf=0.5; % void fraction at the bed at minimum fluidization
eps_e=eps_mf+0.2-0.059*exp(-(U0-Umf)/0.429);

24
eps_b=1-0.149*exp(-(U0-Umf)/4.439);

% kbe, mass transfer coefficient


h=1.5; %unit in m
eps_mf=0.5; % void fraction at the bed at minimum fluidization
dbO=0.0085; % for Geldart B
g=9.81; % gravitational acceleration in m/s2
Dg=6e-7; % gas diffusion coefficient in unit m2/s found in Ibrehem
(2009)
Ub=((-delta*U0)+((1-delta)*0.711*(g*dbO)^(0.5))/(1-2*delta)); % Bubble
velocity in m/s
Ue=Umf/eps_mf*(1-delta); % Elmusion velocity in m/s
db=dbO*((1+27*(U0-Ue))^(1/3))*(1+6.84*h);
Ubr=0.711*((g*db)^(1/2)); % Bubble rise velocity
kbc=4.5*(Ue/db)+5.85*((Dg^(1/2))*(g^(1/4))/(db^(5/4)));
kce=6.77*((Dg*eps_e*Ubr)/db);
kbe=1/((1/kbc)+(1/kce)); % Bubble to emulsion mass transfer
coefficient, kbe in unit 1/s

% Hbe, Heat transfer coefficient


Cp_Mono=1500*mw_Mono; %unit in J/mol K
CpH2=14320*mw_H2; %unit in J/mol K
CpN2=1041*mw_N2; %unit in J/mol K
Cp_pol=4016.64; %unit in J/mol K found in Harshe et al (2004)
Cpg=1778.2; % Specific heat capacity of gaseous stream in J/kg K found
in Harshe et al (2004)
kg=2.55e-2; % Gas thermal conductivity in W/mK found in Choi et al
(1985)
g=9.81; % gravitational acceleration in m/s2
Hbc=4.5*((Ue*rho_g*Cpg)/db)+5.85*((kg*rho_g*Cpg^(1/2))*(g^(1/4))/(db^(5
/4)));%unit in 1/s
Hce=6.77*((rho_g*Cpg*kg)^(1/2))*((eps_e*Ubr)/db^3); %unit in 1/s
Hbe=1/((1/Hbc)+(1/Hce)); % Bubble to emulsion mass transfer
coefficient, kbe in unit 1/s

% Heat of reaction
Hr_change=2.343e6; %unit in J/kg found in Choi et al

VPFR=Vreactor*eps_b; % VPFR Volume of plug flow reactor

rho_pol=rho_s;

% Assumption: 60% Site 1 and 40% Site 2


Fin1=0.6*Fcat;
Fin2=0.4*Fcat;

% Assumption: 88% Emulsion phase and 12% Bubble phase


Finstar_1e=0.88*Fin1;
Finstar_2e=0.88*Fin2;
Finstar_1b=0.12*Fin1;
Finstar_2b=0.12*Fin2;

% Te_in=Tin;
% Tb_in=Tin;

%L/mol s-> m3/mol.s (Conv. Factor = /1000)


% Rate constant for formation, kf(j) in unit m3/s
kf1=1/1000;

25
kf2=2/1000;

% Rate constant for initiation, ki(j) in unit m3/mol.s


ki1=22.88/1000;
ki2=54.93/1000;

% Rate constant for reinitiating by monomer M, kh(j)in unit m3/mol.s


kh1=0.1/1000;
kh2=0.1/1000;

% Rate constant for reinitiating by co-catalyst, khr(j) in unit


m3/mol.s
khr1=20/1000;
khr2=20/1000;

% Propagation rate constant, kp(j) in unit m3/mol.s


kp11=208.6/1000;
kp22=22.8849/1000;

% Activation energy, EA for the propagation reaction in unit k.cal/mol


Rg=1.98722; % general gas constant (cal/mol.K)
T=72+273.15;
Ea1=7200;%
Ea2=7200;
kp1=kp11*exp(-Ea1*((1/T)-(1/Tref))/Rg);
kp2=kp22*exp(-Ea2*((1/T)-(1/Tref))/Rg);

% Transfer rate constant, kfm(j) in unit 1/mol


kfm1=0.0462;
kfm2=0.2535;

% Transfer rate constant, kfh(j) in unit 1/mol


kfh1=7.54;
kfh2=7.54;

% Transfer rate constant,kfr(j) in unit 1/mol


kfr1=0.024;
kfr2=0.12;

% Spontaneous transfer rate constant, kfs(j) in unit 1/mol.s


kfs1=0.0001;
kfs2=0.0001;

% Spontaneous deactivation rate constant, kds(j)in unit 1/s


kds1=0.00034;
kds2=0.00034;

% Deactivation by impurities rate constant, kdl(j) in unit 1/mol.s


kdl1=2000;
kdl2=2000;

% Ri= Sum([Mi]*Y(0,j)*kp(j) i=1,2 j=1,2


R_e=f(39).*f(13)*kp1+f(39).*f(14)*kp2; % Comsumption rate of monomer at
emulsion phase
R_b=f(40).*f(15)*kp1+f(40).*f(16)*kp2; % Comsumption rate of monomer at
bubble phase
Rh_e=f(41).*f(13)*kfh1+f(41).*f(14)*kfh2; % Consumption rate H2 at
emulsion phase

26
Rh_b=f(42).*f(15)*kfh1+f(42).*f(16)*kfh2; % Consumption rate H2 at
bubble phase

% Rp=Sum(mwi*Ri)
Rpe=mw_Mono*R_e+mw_H2*Rh_e; % Total polymer production rate at emulsion
phase
Rpb=mw_Mono*R_b+mw_H2*Rh_b; % Total polymer production rate at bubble
phase

% Parameters for correlations and equations in the two phase


% d=0.1016; % Diameter in m
% pi=3.142; % pi value
h=1.5; % Height in unit m
A=0.00785; % Cross-sectional area of reactor in m3

Vpe=A*h*(1-eps_e)*(1-delta); % Volume of polymer at emulsion phase


Vpb=A*h*(1-eps_b)*(delta); % Volume of polymer at bubble phase

% Rv=volumetric polymer phase outflow rate from the reactor in unit


m3/s
Rve=((mw_Mono*R_e+mw_H2*Rh_e)/rho_s); % Emulsion phase
Rvb=((mw_Mono*R_b+mw_H2*Rh_b)/rho_s); % Bubble phase

Ve=A*(1-delta)*h; % Volume of emulsion phase


Vb=A*(delta)*h; % Volume of bubble phase

% Im=impurity such as carbon monoxide in kmol/m3


% Im=0;

% Kinetic Model
% dN*(j)/dt=Fin*(j)-kf(j)N*(j)-N*(j)(Rv/Vp) j=1,2
dfdt(1)=Finstar_1e-kf1*f(1)-f(1)*(Rpe/Vpe); % Emulsion phase site 1
mol/s
dfdt(2)=Finstar_2e-kf2*f(2)-f(2)*(Rpe/Vpe); % Emulsion phase site 2
mol/s
dfdt(3)=Finstar_1b-kf1*f(3)-f(3)*(Rpb/Vpb); % Bubble phase site 1 mol/s
dfdt(4)=Finstar_2b-kf2*f(4)-f(4)*(Rpb/Vpb); % Bubble phase site 2 mol/s

% dN(0,j)/dt=kf(j)N*(j)-N(0,j)(ki(j)[M]+kds(j)+kdl(j)[Im]+(Rv/Vp))
dfdt(5)=kf1*f(1)-f(5)*(ki1*f(39)+kds1+kdl1*Im+(Rve/Vpe)); % Emulsion
phase site 1 mol/s
dfdt(6)=kf2*f(2)-f(6)*(ki2*f(39)+kds2+kdl2*Im+(Rve/Vpe)); % Emulsion
phase site 2 mol/s
dfdt(7)=kf1*f(3)-f(7)*(ki1*f(40)+kds1+kdl1*Im+(Rvb/Vpb)); % Bubble
phase site 1 mol/s
dfdt(8)=kf2*f(4)-f(8)*(ki2*f(40)+kds2+kdl2*Im+(Rvb/Vpb)); % Bubble
phase site 2 mol/s

% dNH(0,j)/dt=Y(0,j)(kfh(j)[H2]+kfs(j)-
NH(0,j))(kh(j)[M]+khr(j)*AlEt3+kds(j)+kdl(j)[Im]+(Rv/Vp))
dfdt(9)=f(13)*(kfh1*f(41)+kfs1-
f(9))*(kh1*f(39)+khr1*AlEt3_e+kds1+kdl1*Im+(Rve/Vpe)); % Emulsion phase
site 1
dfdt(10)=f(14)*(kfh2*f(41)+kfs2-
f(10))*(kh2*f(39)+khr2*AlEt3_e+kds2+kdl2*Im+(Rve/Vpe)); % Emulsion
phase site 2

27
dfdt(11)=f(15)*(kfh1*f(42)+kfs1-
f(11))*(kh1*f(40)+khr1*AlEt3_b+kds1+kdl1*Im+(Rvb/Vpb)); % Bubble phase
site 1
dfdt(12)=f(16)*(kfh2*f(42)+kfs2-
f(12))*(kh2*f(40)+khr2*AlEt3_b+kds2+kdl2*Im+(Rvb/Vpb)); % Bubble phase
site 2

% dY(0,j)/dt=[M](Ki(j)N(0,j)+kh(j)NH(0,j))+khr(j)NH(0,j)*AlEt3-
Y(0,j)(kfh(j)[H2]+kfs(j)+kds(j)+kdl(j)[Im]+(Rv/Vp)
dfdt(13)=f(39)*(ki1*f(5)+kh1*f(9))+khr1*f(9)*AlEt3_e-
(f(13)*(kfh1*f(41)+kfs1+kds1+kdl1*Im+(Rve/Vpe))); % Emulsion phase site
1
dfdt(14)=f(39)*(ki2*f(6)+kh2*f(10))+khr2*f(10)*AlEt3_e-
(f(14)*(kfh2*f(41)+kfs2+kds2+kdl2*Im+(Rve/Vpe))); % Emulsion phase site
2
dfdt(15)=f(40)*(ki1*f(7)+kh1*f(11))+khr1*f(11)*AlEt3_b-
(f(15)*(kfh1*f(42)+kfs1+kds1+kdl1*Im+(Rvb/Vpb))); % Bubble phase site 1
dfdt(16)=f(40)*(ki2*f(8)+kh2*f(12))+khr2*f(12)*AlEt3_b-
(f(16)*(kfh2*f(42)+kfs2+kds2+kdl2*Im+(Rvb/Vpb))); % Bubble phase site 2

%
dY(1,j)dt=[M]ki(j)N(0,j)+NH(0,j)(kh(j)[M]+khr(j)*AlEt3)+Y(0,j)(kfm(j)[M
]+kfr(j)*AlEt3)+[M]kp(j)Y(0,j)-
Y(1,j)(kfm(j)[M]+kfr(j)*AlEt3+kfh(j)[H2]+kfs(j)+kds(j)+kdI(j)[Im]+(Rv/V
p))
dfdt(17)=f(39)*ki1*f(5)+f(9)*(kh1*f(9)+khr1*AlEt3_e)+f(13)*(kfm1*f(39)+
kfr1*AlEt3_e)+f(39)*kp1*f(13)-
f(17)*(kfm1*f(39)+kfr1*AlEt3_e+kfh1*f(41)+kfs1+kds1+kdl1*Im+(Rve/Vpe));
% Emulsion phase site 1
dfdt(18)=f(39)*ki2*f(6)+f(10)*(kh2*f(10)+khr2*AlEt3_e)+f(14)*(kfm2*f(39
)+kfr1*AlEt3_e)+f(39)*kp2*f(14)-
f(18)*(kfm1*f(39)+kfr2*AlEt3_e+kfh2*f(41)+kfs2+kds2+kdl2*Im+(Rve/Vpe));
% Emulsion phase site 2
dfdt(18)=f(40)*ki1*f(7)+f(11)*(kh1*f(11)+khr1*AlEt3_b)+f(15)*(kfm1*f(40
)+kfr1*AlEt3_b)+f(40)*kp1*f(15)-
f(19)*(kfm1*f(40)+kfr1*AlEt3_b+kfh1*f(42)+kfs1+kds1+kdl1*Im+(Rvb/Vpb));
% Bubble phase site 1
dfdt(20)=f(40)*ki2*f(8)+f(12)*(kh2*f(12)+khr2*AlEt3_b)+f(16)*(kfm2*f(40
)+kfr1*AlEt3_b)+f(40)*kp2*f(16)-
f(20)*(kfm1*f(40)+kfr2*AlEt3_b+kfh2*f(42)+kfs2+kds2+kdl2*Im+(Rvb/Vpb));
% Bubble phase site 2

%
dY(2,j)dt=[M]ki(j)N(0,j)+NH(0,j)(kh(j)[M]+khr(j)*AlEt3)+Y(0,j)(kfm(j)[M
]+kfr(j)*AlEt3)+[M]kp(j)(2Y(1,j)+Y(0,j))-
Y(2,j)(kfm(j)[M]+kfr(j)*AlEt3+kfh(j)[H2]+kfs(j)+kds(j)+kdI(j)[Im]+(Rv/V
p))
dfdt(21)=f(39)*ki1*f(5)+f(9)*(kh1*f(39)+khr1*AlEt3_e)+f(13)*(kfm1*f(39)
+kfr1*AlEt3_e)+f(39)*kp1*(2*f(17)+f(13))-
f(21)*(kfm1*f(39)+kfr1*AlEt3_e+kfh1*f(41)+kfs1+kds1+kdl1*Im+(Rve/Vpe));
% Emulsion phase site 1
dfdt(22)=f(39)*ki2*f(6)+f(10)*(kh2*f(39)+khr2*AlEt3_e)+f(14)*(kfm2*f(39
)+kfr2*AlEt3_e)+f(39)*kp2*(2*f(18)+f(14))-
f(22)*(kfm2*f(39)+kfr2*AlEt3_e+kfh2*f(41)+kfs2+kds2+kdl2*Im+(Rve/Vpe));
% Emulsion phase site 2
dfdt(23)=f(40)*ki1*f(7)+f(11)*(kh1*f(40)+khr1*AlEt3_b)+f(15)*(kfm1*f(40
)+kfr1*AlEt3_b)+f(40)*kp1*(2*f(19)+f(15))-

28
f(23)*(kfm1*f(40)+kfr1*AlEt3_b+kfh1*f(42)+kfs1+kds1+kdl1*Im+(Rvb/Vpb));
% Bubble phase site 1
dfdt(24)=f(40)*ki2*f(8)+f(12)*(kh2*f(39)+khr2*AlEt3_b)+f(16)*(kfm2*f(40
)+kfr2*AlEt3_b)+f(40)*kp2*(2*f(20)+f(16))-
f(24)*(kfm2*f(40)+kfr2*AlEt3_b+kfh2*f(42)+kfs2+kds2+kdl2*Im+(Rvb/Vpb));
% Bubble phase site 2

% dX(0,j)/dt=Y(0,j)(kfm(j)[M]+kfr(j)*AlEt3+kfh(j)[H2]+kfs(j)+kds(j))-
X(0,j)(Rv/Vp)
dfdt(25)=f(13)*(kfm1*f(39)+kfr1*AlEt3_e+kfh1*f(41)+kfs1+kds1)-
f(25)*(Rve/Vpe); % Emulsion phase site 1
dfdt(26)=f(14)*(kfm2*f(39)+kfr2*AlEt3_e+kfh2*f(41)+kfs2+kds2)-
f(26)*(Rve/Vpe); % Emulsion phase site 2
dfdt(27)=f(15)*(kfm1*f(40)+kfr1*AlEt3_b+kfh1*f(42)+kfs1+kds1)-
f(27)*(Rvb/Vpb); % Bubble phase site 1
dfdt(28)=f(16)*(kfm2*f(40)+kfr2*AlEt3_b+kfh2*f(42)+kfs2+kds2)-
f(28)*(Rvb/Vpb); % Bubble phase site 2

% dX(1,j)/dt=Y(1,j)(kfm(j)[M]+kfr(j)*AlEt3+kfh(j)[H2]+kfs(j)+kds(j))-
X(1,j)(Rv/Vp)
dfdt(29)=f(17)*(kfm1*f(39)+kfr1*AlEt3_e+kfh1*f(41)+kfs1+kds1)-
f(29)*(Rve/Vpe); % Emulsion phase site 1
dfdt(30)=f(18)*(kfm2*f(39)+kfr2*AlEt3_e+kfh2*f(41)+kfs2+kds2)-
f(30)*(Rve/Vpe); % Emulsion phase site 2
dfdt(31)=f(19)*(kfm1*f(40)+kfr1*AlEt3_b+kfh1*f(42)+kfs1+kds1)-
f(31)*(Rvb/Vpb); % Bubble phase site 1
dfdt(32)=f(20)*(kfm2*f(40)+kfr2*AlEt3_b+kfh2*f(42)+kfs2+kds2)-
f(32)*(Rvb/Vpb); % Bubble phase site 2

% dX(2,j)/dt=Y(2,j)(kfm(j)[M]+kfr(j)*AlEt3+kfh(j)[H2]+kfs(j)+kds(j))-
X(2,j)(Rv/Vp)
dfdt(33)=f(21)*(kfm1*f(39)+kfr1*AlEt3_e+kfh1*f(41)+kfs1+kds1)-
f(33)*(Rve/Vpe); % Emulsion phase site 1
dfdt(34)=f(22)*(kfm2*f(39)+kfr2*AlEt3_e+kfh2*f(41)+kfs2+kds2)-
f(34)*(Rve/Vpe); % Emulsion phase site 2
dfdt(35)=f(23)*(kfm1*f(40)+kfr1*AlEt3_b+kfh1*f(42)+kfs1+kds1)-
f(35)*(Rvb/Vpb); % Bubble phase site 1
dfdt(36)=f(24)*(kfm2*f(40)+kfr2*AlEt3_b+kfh2*f(42)+kfs2+kds2)-
f(36)*(Rvb/Vpb); % Bubble phase site 2

% dB/dt=R-(B*(Rv/Vp))
dfdt(37)=R_e-(f(37)*(Rve/Vpe)); % Emulsion phase
dfdt(38)=R_b-(f(38)*(Rvb/Vpb)); % Bubble phase

% dfdt(39) Monomer emulsion


% M_e_in=Mono;
% M_b_in=Mono;
% Mh_e_in=H2;
Ae= Mono*Cp_Mono+H2*CpH2;
dfdt(39)=(Mono*Ue*Ae-f(39)*Ue*Ae-Rve*eps_e*f(39)+kbe*(f(40)-
f(39))*Ve*(delta/(1-delta))-(1-eps_e)*R_e)/(Ve*eps_e);

% dfdt(40) Monomer bubble


% Mh_b_in=H2;
Ab=Mono*Cp_Mono+H2*CpH2;
dfdt(40)=(Mono*Ub*Ab-f(40)*Ub*Ab-Rvb*eps_b*f(40)-kbe*(f(40)-f(39))*Vb-
(1-eps_b)*(Ab/VPFR)*R_b*h)/(Vb*eps_b);

29
% dfdt(41) Hydrogen emulsion
% Mh_e_in=H2;
% Mh_b_in=H2;
dfdt(41)=(H2*Ue*Ae-f(41)*Ue*Ae-Rve*eps_e*f(41)+kbe*(f(42)-
f(41))*Ve*(delta/(1-delta))-(1-eps_e)*Rh_e)/(Ve*eps_e); %mol/m3.s

% dfdt(42) Hydrogen bubble


dfdt(42)=(H2*Ub*Ab-f(42)*Ub*Ab-Rvb*eps_b*f(42)-kbe*(f(42)-f(41))*Vb-(1-
eps_b)*(Ab/VPFR)*Rh_b*h)/(Vb*eps_b);

% dfdt(43) T emulsion
Ae=Mono*Cp_Mono+H2*CpH2;
Be=f(39)*Cp_Mono+f(41)*CpH2+yN2*CpN2;
Ce=eps_e*Cp_Mono*f(39)+eps_e*CpH2*f(41)+eps_e*yN2*CpN2+(1-
eps_e)*rho_pol*Cp_pol;
De=Cp_Mono*dfdt(39)+CpH2*dfdt(41);

dfdt(43)=(Ue*Ae*(Tin-Tref)*Ae-Ue*Ae*(f(43)-Tref)*Be-Rve*(f(43)-
Tref)*Ce-(1-eps_e)*Rpe*Hr_change-Hbe*Ve*(delta/(1-delta))*(f(43)-
f(44))-Ve*eps_e*(f(43)-Tref)*De)/(Ve*Ce);

% dfdt(44) T bubble
Ab=Mono*Cp_Mono+H2*CpH2;
Bb=f(40)*Cp_Mono+f(42)*CpH2+yN2*CpN2;
Cb=eps_b*Cp_Mono*f(40)+eps_b*CpH2*f(42)+eps_b*yN2*CpN2+(1-
eps_b)*rho_pol*Cp_pol;
Db=Cp_Mono*dfdt(41)+CpH2*dfdt(42);

dfdt(44)=(Ub*Ab*(Tin-Tref)*Ab-Ub*Ab*(f(44)-Tref)*Bb-Rvb*(f(44)-
Tref)*Cb+(1-eps_b)*((Ab*Hr_change)/VPFR)*Rpb*h+Hbe*(f(43)-f(44))*Vb-
Vb*eps_b*(f(44)-Tref)*Db)/(Vb*Cb); %K/s

% dfdt(45) until dfdt(48) for N(1,j) j=1,2


%
dfdt(45)=ki1*f(5).*f(39)+f(9)*(kh1*f(39)+khr1*AlEt3_e)+f(13)*(kfm1*f(39
)+kfr1*AlEt3_e)-
f(45)*(kp1*f(39)+kfm1*f(39)+kfh1*f(41)+kfr1*AlEt3_e+kfs1+kds1+kdl1*Im+(
Rve/Vpe)); % Emulsion phase site 1
%
dfdt(46)=ki2*f(6).*f(39)+f(10)*(kh2*f(39)+khr2*AlEt3_e)+f(14)*(kfm2*f(3
9)+kfr2*AlEt3_e)-
f(46)*(kp2*f(39)+kfm2*f(39)+kfh2*f(41)+kfr2*AlEt3_e+kfs2+kds2+kdl2*Im+(
Rve/Vpe)); % Emulsion phase site 2
%
dfdt(47)=ki1*f(7).*f(40+f(11)*(kh1*f(40)+khr1*AlEt3_b)+f(15)*(kfm1*f(39
)+kfr1*AlEt3_b)-
f(47)*(kp1*f(40)+kfm1*f(40)+kfh1*f(42)+kfr1*AlEt3_b+kfs1+kds1+kdl1*Im+(
Rvb/Vpb))); % Bubble phase site 1
%
dfdt(48)=ki2*f(8).*f(40)+f(12)*(kh2*f(40)+khr2*AlEt3_b)+f(16)*(kfm2*f(3
9)+kfr2*AlEt3_b)-
f(48)*(kp2*f(40)+kfm2*f(40)+kfh2*f(42)+kfr2*AlEt3_b+kfs2+kds2+kdl2*Im+(
Rvb/Vpb)); % Bubble phase site 2

end

30
B. Running Model

yC3H6=0.59;
yH2=0.0646;
yN2=0.337;
yC2H=4.7e-3;
yC6star=0.19e-2;

% Total gas inlet


Press=22; % in bar
t=72+273.15;
R=8.314e-5; % in m3 bar mol-1 K-1 (unit)
Z=0.9; % non-ideal gas
Cg=Press/(t*Z*R); % in mol/m3
Mono=yC3H6*Cg;
H2=yH2*Cg;
Im=0;
Fcat1=3.02; % in kg/h
Fcat=3.02/3600; % in kg/s
U0=0.35; % in m3/s
Tin=317.15; % in K
Tref=353.15; % in K

mw_Mono=42.08; % in g/mol
mw_H2=2; % in g/mol
% Propagation rate constant, kp(j) in unit m3/mol.s
kp11=208.6/1000;
kp22=22.8849/1000;
kfh1=7.54;
kfh2=7.54;

% Activation energy, EA for the propagation reaction in unit k.cal/mol


Rg=1.98722; % general gas constant (cal/mol.K)
T=72+273.15;
Ea1=7200;%
Ea2=7200;
kp1=kp11*exp(-Ea1*((1/T)-(1/Tref))/Rg);
kp2=kp22*exp(-Ea2*((1/T)-(1/Tref))/Rg);

tspan=[0 10000];
f0=[0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 Mono Mono H2 H2 Tin Tin];
[t,f]=ode15s(@(t,f)
TwoPhaseModel(t,f,Mono,H2,Im,Fcat,U0,Tin,Tref,Press,yC3H6,yH2,yN2,yC2H,
yC6star), tspan, f0);
Temul=real(f(:,43));
Tbub=real(f(:,44));
% Tavg=(Temul(end)+Tbub(end))/2
% plot(t,f(:,43))
% Ri= Sum([Mi]*Y(0,j)*kp(j) i=1,2 j=1,2
R_e=f(:,39).*f(:,13)*kp1+f(:,39).*f(:,14)*kp2; % Comsumption rate of
monomer at emulsion phase
R_b=f(:,40).*f(:,15)*kp1+f(:,40).*f(:,16)*kp2; % Comsumption rate of
monomer at bubble phase
Rh_e=f(:,41).*f(:,13)*kfh1+f(:,41).*f(:,14)*kfh2; % Consumption rate H2
at emulsion phase
Rh_b=f(:,42).*f(:,15)*kfh1+f(:,42).*f(:,16)*kfh2; % Consumption rate H2
at bubble phase

31
% Rp=Sum(mwi*Ri)
Rpe=mw_Mono*R_e+mw_H2*Rh_e; % Total polymer production rate at emulsion
phase
Rpb=mw_Mono*R_b+mw_H2*Rh_b; % Total polymer production rate at bubble
phase
Rp=Rpe+Rpb;
Rpf=real(Rp);
plot(t,Rpf)
plot(t,f(:,44))
%plot(t,f(:,39))
% plot(t,f(:,41))
% plot(t,f(:,42))
% plot(t,f(:,43))
% plot(t,f(:,44))

32

You might also like