You are on page 1of 19

Using Simulation Technology to Improve Profitability

In the Polymer Industry

David A. Tremblay*

Aspen Technology Inc.


Ten Canal Park
Cambridge MA, 02141-2201
USA

Key words: Polyester, PET, Poly(ethylene terephthalate), process simulation, optimization

Prepared for presentation at the AIChE Annual Meeting, Houston Texas, March 14-19, 1999

Copyright  David A. Tremblay, Aspen Technology Incorporated, 1999.

AIChE shall not be responsible for statements or opinions contained in papers or printed in its publications
Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Using Simulation Technology to Improve Profitability


In the Polymer Industry

David A. Tremblay*

Aspen Technology Inc.


Ten Canal Park
Cambridge MA, 02141-2201
USA

Copyright  David A. Tremblay, 1999.

Abstract
Process simulation is a valuable tool to increase the profitability of polyester processes by reducing
costs, increasing yield, and improving product quality. Although simulation has been heavily used in
the chemical process industries for several decades, the polymer manufacturing industry has only begun
to take advantage of this technology during the past five to ten years. Recently, the commercial
development of polymer process simulation packages, such as Polymers Plus®, has made it possible to
simulate polymer processes in a simple and straightforward manner.
In this discussion, we will examine the technical challenges that must be overcome to develop and
validate a polyester process model. Several case studies will be presented. Finally, we will review how a
rigorous process model can be leveraged to support additional business needs including operator training
and process control.

Copyright © 1999 Page 1 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Process Description
Most polyethylene terephthalate (PET) is produced from purified terephthalic acid (TPA) and ethylene
glycol (EG). Although many process configurations are found in the polyester industry, they all involve
a series of three or more reactors. For demonstration purposes, we have developed a model of a typical
five-reactor process, as shown in Figure 1.
Solid terephthalic acid is mixed with ethylene glycol in carefully metered amounts. The mole ratio of
ethylene glycol to TPA is a critical process variable, so many plants use control schemes to adjust the
glycol feed rate to keep the paste density constant (the density is a good indicator of mole ratio). The
resulting paste is fed to the first reactor, which is known as a primary esterification reactor or PE.
Typically, the PE is operated at a pressure of 1-8 bar and a temperature of 255-280°C. At lower
temperatures the reactor performance is limited by the solubility of TPA in the oligomer. The behavior
of the reactor is highly non-ideal because the apparent reactor volume depends on the amount of solid
TPA. At higher temperatures the reactor performance is limited by the solid-liquid mass-transfer rate.
Under these conditions, the reactor performance depends on the TPA particle size.
Oligomer from the primary esterifier is fed to the secondary esterifier (SE). The secondary esterifier is
usually run close to atmospheric conditions with temperatures a bit higher than the PE. Frequently, the
SE is divided into several chambers to enhance the effectiveness of the reactor. In our model, we assume
the SE is divided into three equal-sized chambers. Each chamber is represented as an ideal CSTR
reactor. These assumptions are justified by sensitivity studies that indicate that vapor back mixing
between the stages has little influence on the model predictions.
Primary Secondary Spray Condenser and
Esterification Esterification Vacuum System
Column Column
Steam

Solid Steam Steam Makeup


TPA Glycol
Make-up Water Water
EG

Catalyst
Additives
Recycle
EG Bypass
Recycle Recycle Crude EG
EG EG to Recovery

Catalyst
Paste Additives

Primary Secondary
Paste Low Intermediate High
Esterification Esterification
Tank Polymerizer Polymerizer Polymerizer
Reactor Reactor

Figure 1. Process flow diagram of a typical five-reactor direct esterification PET process, including the
reactor train and vapor recovery system.

Copyright © 1999 Page 2 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

The vapor from each esterification reactor is rectified in a multi-stage distillation column. In industry,
many processes use one column to support both reactors. The columns remove water and other volatile
reaction by-products, including acetaldehyde. Excess ethylene glycol is recovered and returned to the
paste tank and the esterification reactors. The recycle rate to each esterification reactor can be
manipulated to control the local monomer ratio.
The third reactor, or low polymerizer, is typically composed of a simple CSTR. The low polymerizer
(LP) operates at a medium vacuum pressure (50-500 Torr). This stage strips off most of the excess
ethylene glycol and water remaining in the polymer. In most plants, the polymer intrinsic viscosity in the
low polymerizer is below 0.2 dl/g, and the LP behaves ideally. At higher viscosity levels, the low
polymerizer becomes increasing mass-transfer limited.
The final reactors, known as the intermediate polymerizer (IP) and high polymerizer (HP), operate at
lower vacuum pressures, often as low as one Torr. Many plants use disk-ring reactors, which contain a
number of annular disks attached to a rotating shaft. Polymer flows through holes cut into the disks. As
the disks rotate they generate a surface film, which enhances the evaporation rate. Due to the high
viscosity of the polymer, the performance of the finishing reactors is limited by the liquid-vapor mass-
transfer rate. This makes the reactor performance a function of the shaft rotation rate, as well as the
temperature, pressure, and throughput.
The vapor from the polymerization reactors is recovered using a cascade of spray condensers. The spray
condensers recover most of the ethylene glycol generated by the reactions. The ethylene glycol
recovered in the spray condensers may be further purified or recycled directly to the paste tank,
depending on the product grade.
There are several factors that must be considered to simulate the reactors in a melt polyester process.
These include non-ideal mixing, mass-transfer limitations, and uncertainty in the true operating volume.
Many industrial reactors contain baffles that divide the reactor into two or more mixing zones. These
reactors can be represented as a series of CSTRs. Recycle streams can be used to account for back
mixing in the liquid and vapor phases.
The finishing reactors can be simulated as a series of CSTR reactors, with each space between
successive disks represented as one ideal mixing stage [1]. In the model presented here, we simulate the
IP and HP as mass-transfer-limited plug-flow reactors. Simulation studies indicate that the dynamic and
steady state responses of both models are in good agreement when there are five or more ideal mixing
stages.
Component Characterization
In conventional chemical processes, component characterization is straightforward. The molecular
structure of each of the reactants is well defined. In a polymerization process, however, the polymer is
composed of a distribution of molecules. The average properties of the distribution change as the
reactions proceed.
In our model, we characterize the polymer using the segment approach, in which the reactions are
written in terms of monomers and polymer repeat units and end groups. Using this approach, we define
an end-group and repeat segment associated with each monomer. Additional segments are defined to
account for some of the side reactions that occur in this process. Figure 2 summarizes the main
components and segments in our model.

Copyright © 1999 Page 3 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Name Description Structure

H2O Water H2O


EG Ethylene-glycol HO(CH2)2OH
TPA Dissolved terephthalic acid O O
HO C C OH
TPAS Solid terephthalic acid
DEG Diethylene-glycol HO(CH2)2O(CH2)2OH
AA Acetaldehyde H(C=O)CH3
PET Polyethylene terephthalate Structure made up of segments
t-EG Ethylene-glycol end segment ~O(CH2)2OH
b-EG Ethylene-glycol repeat segment ~O(CH2)2O~
t-TPA Terephthalic acid end segment O O
C C OH

b-TPA Terephthalic acid repeat segment O O


C C

t-DEG Diethylene-glycol end segment ~O(CH2)2O(CH2)2OH


b-DEG Diethylene-glycol repeat segment ~O(CH2)2O(CH2)2O~
t-vinyl Oxyvinyl end segment ~OCH=CH2

Figure 2. Components, including polymer segments, included in the model. Note that the prefixes “t”
and “b” refer to “terminal” (end group) and “bound” (repeat) segments.

Segments provide a convenient basis for defining reaction stoichiometry and for calculating end-use
properties. The model tracks the mass and mole flow rates of the monomers, water, acetaldehyde, DEG,
and each of the segments. The mass flow rate of the polymer component is calculated from the sum of
the mass flow rate of the individual segments. The model also tracks the mole flow rate of the polymer
(the zeroth moment of the molecular weight distribution). Secondary properties, such as the acid end
group content and intrinsic viscosity, are calculated from component flow rates and polymer moments.
The physical properties of the polymer, such as density, enthalpy, and heat capacity, are calculated using
group contribution techniques based on the segmental composition. The polymer component is assumed
to be non-volatile.
End-use properties, such as end-group concentrations, intrinsic viscosity, and DEG content, are
calculated from the mole flow rates of monomers and polymer segments. The definitions of these
properties are documented at the end of this article.

Copyright © 1999 Page 4 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Phase Equilibrium
The model considers solid-liquid and liquid-vapor phase equilibrium. For TPA, the solid-liquid
equilibrium is calculated from the solubility constant, which was fit against solubility data over a wide
range of temperatures:
∆H m ∆S m
xTPA γ TPA = ln K s = −
RT R
For convenience the polymer is defined to have a fixed reference mole weight. The mole fractions and
molar activity coefficient are calculated on an apparent component basis using this reference mole
weight.
The concentration of the volatile components at the liquid-vapor interface is calculated using the activity
coefficient and vapor pressure of each component:
xi γ i Pi v = y i P
The activity coefficients are calculated using the polymer-NRTL equation [2]. Using this equation, the
activity coefficients depend on the apparent mole fractions and binary interaction parameters. Since the
polymer is the main solvent in the system, the water-polymer and ethylene glycol-polymer interaction
parameters control the phase equilibrium in the reactors. The NRTL binary parameters are critical
constants in the model because the reaction rates and mass-transfer rates depend on the liquid-phase
concentrations.
The activity coefficient of water controls the relationship between the predicted acid value and the
reactor pressure. When the activity coefficient is high, the predicted acid value is very sensitive to the
reactor pressure. When the activity coefficient is low, the predicted acid value is less sensitive to
pressure. Thus, the water-polymer interaction parameters can be fit against process data taken over a
range of operating pressures.
The activity coefficient of ethylene glycol controls the relationship between the intrinsic viscosity and
pressure in the finishing reactors. The higher the activity coefficient, the stronger the predicted trend
between the intrinsic viscosity and pressure. Thus, the ethylene glycol-polymer interaction parameter
can be fit against IV/pressure data.
Reaction Kinetics
The reaction mechanisms and kinetics of the direct esterification process are examined in many open
sources. The rate parameters used in our model are based on our interpretation of experimental work
published by Yokoyama [3,4], Otton and Ratton [5,6], and Ravindrananth and Mashelkar [7,8].
The main reactions in the process are shown in Figure 3. In the early stages of the process, esterification
reactions (1-4 in the figure) are dominant. These reactions are reversible with equilibrium constants
close to unity. Polymerization reactions (5 and 6) involve the nucleophilic attack of the terminal ester
group in one chain by the terminal alcohol group of a second chain. These reactions lead to rapid
molecular weight growth. The polymerization reactions are also reversible, with equilibrium constant
slightly below unity. Rearrangement reactions (7 and 8) involve ester groups inside the chain. These
reactions are responsible for randomizing the molecular weight distribution and redistributing segments
inside the polymer chains. Since the number of internal ester groups is much higher than the number of
terminal ester groups, the molecular weight distribution approaches the most-probable distribution.

Copyright © 1999 Page 5 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

# Reaction Stoichiometry Population Balance


1 EG + TPA ↔ t-EG ≈ t-TPA + H2O M + M ↔ P2 + W
2 EG + t-TPA ↔ t-EG ≈ b-TPA~ + H2O M + Pn ↔ Pn+1 + W
3 ~t-EG + TPA ↔ ~b-EG ≈ t-TPA + H2O Pn + M ↔ Pn+1 + W
4 ~t-EG + t-TPA ↔ ~b-EG ≈ b-TPA~ + H2O Pn + Pm ↔ Pn+m + W
5 ~t-EG + t-TPA ≈ t-EG ↔ ~b-EG ≈ t-TPA + EG Pn + P2 ↔ Pn+1 + M
6 ~t-EG + ~b-TPA ≈ t-EG ↔ ~b-EG ≈ b-TPA~ + EG Pn + Pm ↔ Pn+m-1 + M
7 ~t-EG + t-TPA ≈ b-EG~ ↔ ~b-EG ≈ t-TPA + t-EG~ Pn + Pm ↔ Pn+1 + Pm-1
8 ~t-EG + ~b-TPA ≈ b-EG~ ↔ ~b-EG ≈ b-TPA~ + t-EG~ Pn + Pm ↔ Pn+q + Pm-q
9 ~b-TPA ≈ t-EG + EG → ~t-TPA + DEG Pn + M → Pn’ + M’
10 ~b-TPA ≈ t-EG + ~t-EG → ~t-TPA + t-DEG Pn + Pm → Pn-1’ + Pm’
11 ~b-TPA ≈ t-EG → ~t-TPA + AA Pn → Pn-1 + AA
12 ~b-TPA ≈ b-EG~ → ~t-TPA + ~t-vinyl Pn → Pn-m +Pm
13 ~t-EG + ~b-TPA ≈ t-vinyl → ~b-TPA ≈ b-EG + AA Pn + Pm → Pn-m-1 + AA
Notes:
The double-tilde “≈”indicates that two segments are attached by a covalent single bond
A tilde “~” before or after a segment indicates that one side of the segment is attached to a chain
“Pn” indicates a linear polymer chain composed of “n” segments.
“M” refers to monomers, “W” refers to water

Figure 3. A summary of the reactions included in the model. Analogous reactions involving various
combinations of EG, DEG and their respective segments are also included in the model. Using this
technique, the smallest polymer chain has a length of two segments (this is equivalent to one PET repeat
unit).

A number of side reactions are also involved in the process. These include thermal scission and the
formation of diethylene glycol and acetaldehyde. All of the side reactions have a strong influence on the
quality of the polymer. The DEG content, for example, affects the viscosity, density, and dying
characteristics of the polyester. The scission reactions influence the acid content and yellowness of the
polymer. Scission and acetaldehyde formation reduce the polymer yield (these reactions lead to the loss
of ethylene glycol in the form of acetaldehyde). A good model must consider these side reactions in
order to identify economic opportunities related to the improvement of yield and quality.
Diethylene glycol segments are formed throughout the process, especially in the esterification reactors.
There are substantial debates about the exact mechanism by which the DEG is formed. One proposed
mechanism involves a reaction between a molecule of ethylene glycol or an ethylene glycol end group
and a cyclic transition state generated by terminal glycol ester groups. Another proposed mechanism

Copyright © 1999 Page 6 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

involves a two-step process in which the terminal glycol group in one chain undergoes a scission
reaction to form ethylene oxide, which subsequently reacts with monomeric ethylene glycol or with an
ethylene glycol end group in a second chain. Both of these mechanisms result in second-order reactions
as shown in reactions 9 and 10 in Figure 3.
Two types of thermal scission reactions occur. Thermal scission involving the terminal glycol group
generates an acid group and ethylene oxide. The ethylene oxide may react with another molecule to
form DEG as described above, or it may rearrange to form acetaldehyde (reaction 11). Scission reactions
inside the chain generate vinyl end groups and acid end groups (reaction 12). Subsequent
polycondensation reactions at the oxyvinyl ester group consume these vinyl groups and generate a vinyl
alcohol, which spontaneously rearranges to for acetaldehyde (reaction 13). These reactions destroy
alcohol groups and generate acid groups. This has a profound influence on the acid content and
molecular weight of the polymer. Since the scission reactions have higher activation energies than the
polymerization and esterification reactions, the process is very sensitive to the temperature of the
finishing reactors.
A number of other reactions are likely to occur, especially at high temperatures and in the presence of
trace amounts of oxygen. These reactions lead to the generation of color bodies in the polymer chain.
For simplicity, these reactions are not considered in our model. With sufficient data, however, these
reactions can be characterized through correlations generated by traditional methods or regressed from
the L/a/b color data and operating conditions using a neural network.
The esterification and polymerization reactions are acid-catalyzed. Various metal compounds, including
diantimony triacetate and germanium oxide, are also used to catalyze the reaction rates. These metal
compounds are inhibited by unreacted alcohol groups [9]. In the first few reactors the concentration of
acid and alcohol groups is very high, so the acid catalysis mechanism dominates. Metal catalysis
dominates in the polymerization reactors. Phosphate compounds are added to the polymer to stabilize
the catalysts, reducing the thermal scission rates. Our model accounts for both catalyst inhibition
mechanisms.
Model Validation
Even though the rate constants and other parameters in the model can be determined from literature
sources, it is still critical to fine-tune the model against process data. Each plant has its own
idiosyncrasies. Heat- and mass-transfer coefficients are equipment-specific. The operating volumes of
the reactors are usually poorly characterized. Flow meters may not be calibrated properly. Contaminants
in the raw material may influence the side reactions in the process. To account for these factors, the
model must be tuned against real process data.
Validation is also an important means of reaching organization-wide agreement on the quality and
reliability of the model. Testing the model against process data clarifies the strengths and weaknesses of
the model. Since simulation models are used to find new operating conditions, there must be convincing
proof that the model is representative of the plant.
Validation data should be collected over as wide a range of operating conditions as possible. To validate
a steady-state model, the data should reflect stable operating conditions. Ideally, the data should be
collected over several shifts and averaged. Collecting data in this fashion may allow the model
developer to evaluate the stability of the process and to weed out invalid or atypical data.
Validating dynamic models is a bit more difficult. Data needs to be collected during process upsets or
during start up, shut down, and grade transition operations. In some cases, it is necessary to artificially

Copyright © 1999 Page 7 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

disturb the process by making a step change in one or more of the operating conditions. These data make
it possible to determine the various time constants in the system. For example, by making a step change
in the temperature set point of a reactor, one can determine the thermal inertia of the reactor.
Each data set should include all of the operating conditions and feed conditions in the plant. Ideally, the
oligomer should be sampled at the outlet of each reactor. Each sample should be analyzed for the end
group content, intrinsic viscosity, and DEG content. These data can be used to fine-tune the various
reaction rate constants.
In developing and validating a model, it is critical to remember what a model really is – a simplified
representation of the plant. Simulation models are subject to the law of diminishing returns, unnecessary
details and complications should be avoided. No model is perfect, and no model can reproduce plant
data with perfect fidelity. There are several sources of differences between the model predictions and
process data, including analytical error, imperfect calibration of flow meters and liquid levels, deviations
from steady state conditions, and simplifying assumptions built into the simulation.
Case Study #1: Using the Model to Evaluate the Influence of Feed Mole Ratio
The ratio of EG to TPA in the feed is a key process parameter. Optimizing this ratio can result in
improved quality and higher production rates for a process. In this example we carried out a sensitivity
study to evaluate the influence of the feed mole ratio on the process, holding the reactor operating
conditions constant. The results of this study are summarized in Figure 4, below.
As shown, there is an optimal mole ratio at which the intrinsic viscosity is a maximum. Although the
production rate was held constant in this example, the maximum intrinsic viscosity can be translated into
a maximum production rate at constant IV.
The optimal mole ratio depends on the operating conditions of the reactors, so there is no universal
“best” mole ratio at which to run a given plant. The optimal mole ratio is influenced by the thermal
scission and acetaldehyde formation reactions, which destroy glycol end groups.
At very low mole ratios, the final intrinsic viscosity drops of quickly. Under these conditions, all of the
polymer end groups are terminated by acid ends, which cannot react with each other. In effect, the
polymer becomes “capped” and the growth stops. At very high glycol/TPA ratios, the polymer has
mostly alcohol ends. The alcohol end groups undergo polymerization reactions, so there is substantial
growth. This reaction, however, is not favored (the equilibrium constant is less than one).
At intermediate mole ratios, the polymer chains grow by esterification and polymerization. The
esterification reaction is favored (the equilibrium constant is greater than one), so the reaction is less
limited by chemical equilibrium. Even more importantly, the esterification reaction generates water
instead of ethylene glycol. Water is far more volatile than glycol, so it is easier to remove from the
polymer melt. This also favors the forward reaction.

Copyright © 1999 Page 8 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

0.70 160
0.65 140

Acid Content, mmol/kg


Intrinsic Viscosity, dl/g

0.60 Intrinsic Viscosity


120
0.55
0.50 100
0.45 80
0.40 60
0.35
Acid Content 40
0.30
0.25 20
0.20 0
1.00 1.05 1.10 1.15 1.20
Feed Mole Ratio EG/TPA

Figure 4. Product acid content and intrinsic viscosity as a function of the feed mole ratio at constant
throughput. All reactor operating conditions are held constant for this study. These results demonstrate
that there is an optimal mole ratio for the process. The optimal mole ratio depends on the throughput and
operating conditions.

Case Study #2: Increasing Throughput


In the previous study we examined the influence of the feed mole ratio on the product acid content and
intrinsic viscosity. The results are intriguing because they show that there is an optimal mole ratio.
However, a simple sensitivity study leaves many questions unanswered. For example, what are the best
operating conditions for each reactor? What production rate can be obtained if the intrinsic viscosity is
held constant?
Process optimization studies can answer these questions. In an optimization study, one or more process
variables are manipulated to minimize or maximize an objective function while meeting constraints.
Manipulated variables may include operating conditions, such as temperatures and pressures, feed flow
rates, catalyst and additive levels, and other factors that can be controlled in the process. The upper and
lower limit of each variable is specified to ensure the optimization finds a feasible set of operating
conditions.
Product quality is usually the key constraint. There are tight quality specifications for any given product
grade. Typically, these specifications include the DEG content, acid end-group concentration, and
intrinsic viscosity. In addition, there may be physical constraints on the equipment, for example, a
reactor may be limited by the available duty or hot oil supply.
In an optimization study, these constraints are defined as equality or inequality constraints. Equality
constraints set the target and tolerance of a given model prediction, for example, IV=0.610±0.005.
Inequality constraints represent lower or upper limits in the predicted variables.

Copyright © 1999 Page 9 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

We must select an objective function to optimize. The objective function can be any continuous
numerical quantity, such as the production rate, mole ratio, and so on. The optimization tool can
minimize or maximize the objective function.
In this example we maximized production rate by manipulating several process variables including the
reactor temperatures and pressures, the catalyst and stabilizer levels, and the feed mole ratio. We set
constraints on the intrinsic viscosity, the concentration of acid and vinyl end groups, and the DEG
content. The simulation results are summarized in Table 1.

Variable Type Variable Lower Upper Initial Optimized


Bound Bound Value Value
Objective Function PET production rate, kg/hr - - 7147 9626
Constraints Intrinsic viscosity, dl/g 0.595 0.610 0.606 0.605
Acid content, mmol/kg - 40.0 23.0 39.1
DEG content, weight % 1.15 1.35 0.52 1.25
Vinyl content, mmol/kg - 3.00 1.96 3.00
Manipulated Variables TPA feed mass flow, kg/hr 6140 12000 6150 8263
Feed mole ratio EG/TPA 1.05 1.25 1.10 1.19
Sb2O3 concentration, PPM 100 300 250 300
H3PO4 concentration, PPM 50 150 50 50
Fraction of EG recycled to PE 0.70 0.99 0.90 0.99
Fraction of EG recycled to SE 0.50 0.99 0.90 0.50
PE temperature, deg C 250.0 270.0 260.0 270.0
PE pressure, Torr 1500 4000 1839 4000
SE temperature, deg C 250.0 280.0 260.0 280.0
LP temperature, deg C 260.0 300.0 275.0 300.0
LP pressure, Torr 50.0 500.0 50.0 50.0
IP temperature, deg C 270.0 305.0 280.0 305.0
IP pressure, torr 1.0 3.0 2.0 1.0
HP temperature deg C 270.0 305.0 285.0 287.7
HP pressure, torr 0.8 2.0 1.0 0.8
Table 1. Results of a process optimization study for the five-reactor process shown in Figure 1. In most
optimization studies a number of variables reach their bounds. In this case, the production rate is limited
by the constraints on the acid content and vinyl content.

The results of the optimization run are very interesting. The model indicates that it is possible to achieve
higher production rates by increasing the polymerization temperatures. Note that the quality of the
polymer is not sacrificed. The acid value and vinyl content can be maintained at higher temperatures and
throughputs because the reactor residence times are reduced as the throughput increases.
The model predicts that the finishing reactors should operate at the lowest possible pressure. This is not
surprising because these reactors are mass-transfer limited. By reducing the reactor pressure, we
increase the concentration gradient between the bulk liquid phase and the liquid-vapor interface. Lower
pressures lead to higher evaporation rates, improving the effectiveness of the reactors.

Copyright © 1999 Page 10 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Predicting the best pressure and temperature of the low polymerizer is difficult. If the pressure is too low
or the temperature is too high, too much of the residual ethylene glycol is lost. This reduces the mole
ratio, which may shift the process away from the optimal mole ratio.
Table 1, above, lists the initial and final values of the objective function, constraints, and manipulated
variables. Note that several of the manipulated variables reach their upper and lower bounds. The
variables that do not reach the bounds may be at optimal conditions. In some cases, the manipulated
variables are limited by the constraints. For example, in this process the temperature of the high
polymerizer cannot be higher than 287.7°C without violating the acid and vinyl end-group constraints.
Table 2, below, summarizes the polymer streams at the base case and after optimization. Note that the
final acid content is much higher in the optimal case. This accelerates the reaction rates in the
polymerization reactors. When the process is operated at such high temperatures, however, more of the
ethylene glycol is lost to side reactions, so the feed mole ratio must be higher. Although many product
grades have tighter quality constraints on the acid content, high acid content may be desirable when the
polymer is used as the feedstock for a solid-state polymerization process.
In this example, ethylene glycol is recycled from the distillation column to the primary and secondary
esterification reactors. A portion of the ethylene glycol is removed from the secondary esterification
stage. The recycle rates to the primary and secondary esterification reactors are included as manipulated
variables in this optimization study. This allows the model to find the optimal mole ratio in each
esterification stage.

Antimony and phosphate levels were also included in this study. As expected, the model drives the
catalyst level to the upper bound and the phosphate level to the lower bound. At higher production rates,
higher catalyst levels are offset by lower residence times.

The optimization study identified thirty five percent additional capacity over the base-case conditions.
These results are very high because the base-case conditions reported here reflect industry averages. In
most plants, years of fine-tuning the process have led to operation far above the design capacity. Even in
very mature plants, however, optimization studies may identify an additional five to ten percent
capacity.

Process economics are summarized in Table 3. The prices of TPA, ethylene glycol, and PET are based
on the commodity market reports for January 1999. The utility prices are estimated based on the process
heat duties and current energy prices in the United States. We assume that the increase in throughput has
no influence on the process labor costs. These figures demonstrate that even with today’s narrow
margins, increasing the capacity of an existing plant can be very profitable. For this 170 ton per day
plant, a five- percent increase in capacity represents over one million US dollars per annum.

Copyright © 1999 Page 11 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Stream Conditions Before Optimization


Reactor PE SE LP IP HP
Total Flow kg/hr 7407.8 7515.5 7255.1 7166.1 7147.4
Temperature, deg C 260.0 260.0 275.0 280.0 285.8
Absolute pressure, torr 1839 765 50 2.0 1.0
Antimony Oxide, wt PPM 0 237 245 248 249
Phosphoric Acid, wt PPM 0 0 49 50 50
Degree of Esterification 89.43 97.66 99.47 99.79 99.78
Acid content, mmol/kg 1056.7 230.9 53.8 22.2 23.0
Hydroxyl content, mmol/kg 1614.8 1732.4 572.9 164.6 80.2
Vinyl content, mmol/kg 0.00 0.02 0.22 0.92 2.16
Saponification, mmol/kg 9994 9851 10205 10332 10359
Intrinsic viscosity, dl/g 0.049 0.057 0.140 0.376 0.606
DEG content, weight % 0.349 0.452 0.500 0.515 0.522
Number average DP 4.79 5.78 17.00 55.79 99.01
Number average Mn 920 1110 3266 10721 19026
Stream Conditions After Optimization
Reactor PE SE LP IP HP
Total Flow kg/hr 10334.2 10192.3 9745.3 9644.7 9626.3
Temperature, deg C 270.0 260.0 300.0 305.0 288.3
Absolute pressure, torr 4000 765 50 1.0 0.8
Antimony Oxide, wt PPM 0 281 294 297 298
Phosphoric Acid, wt PPM 0 0 49 50 50
Degree of Esterification 90.31 98.20 99.67 99.57 99.62
Acid content, mmol/kg 933.2 175.5 34.1 44.7 39.1
Hydroxyl content, mmol/kg 2563.2 1904.7 457.9 127.1 63.5
Vinyl content, mmol/kg 0.00 0.03 0.65 2.65 3.00
Saponification, mmol/kg 9626 9760 10208 10314 10334
Intrinsic viscosity, dl/g 0.039 0.055 0.171 0.399 0.605
DEG content, weight % 0.931 1.108 1.214 1.244 1.250
Number average DP 3.58 5.55 21.53 59.95 98.85
Number average Mn 689 1066 4137 11520 18997
Table 2. Stream conditions for the base-case conditions and for conditions optmized to obtain maximum
production rates.

Copyright © 1999 Page 12 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Optimized production rate 9.626 metric ton / hour


Original production rate - 7.147 metric ton / hour
Additional throughput 2.479 metric ton / hour
x 24 hours / day = 59.493 metric ton / day
x 365 days / year = 21715.0 metric ton / year
Selling price of polyester x $ 948 $ / metric ton
Additional revenue $ 20,585,775 $ / year

TPA feed rate at optimum 8.263 metric ton / hour


Original TPA feed rate - 6.150 metric ton / hour
Additional TPA flow rate 2.113 metric ton/hour
x 24 hours / day = 50.720 metric ton / day
x 365 days / year = 18512.9 metric ton / year
Price of fiber-grade TPA x $ 507 $ / metric ton
Additional TPA costs $ 9,386,039 $ / year

EG feed rate at optimum 3.715 metric ton / hour


Original EG feed rate - 2.676 metric ton / hour
Additional TPA flow rate 1.039 metric ton / hour
x 24 hours / day = 24.936 metric ton / day
x 365 days / year = 9101.6 metric ton / year
Price of fiber-grade EG x $ 310 $ / metric ton
Additional EG costs $ 2,821,508 $ / year

Utility costs $ 30 $ / metric ton


Additional throughput x 21715 metric ton / year
$ 651,449 $ / year

Revenue from additional sales $ 20,585,775 $ / year


Additional raw material costs $ (12,207,548) $ / year
Additional utility costs $ (651,449) $ / year
Net return $ 7,726,779 $ / year

Price indices based on January 1999 spot market


Table 3. Process economic calculations for the throughput optimization study. The monomer prices
reported here are based on the spot-market prices for the second week of January 1999. Utility prices are
estimated. We assume a constant utility price per mass of polymer, independent of the operating
condtions.

Case Study #3: Reducing Costs


Polyester is projected to be in oversupply until the middle of the next decade. Under these
circumstances, cost cutting becomes essential to survival. From a process engineering point of view, the

Copyright © 1999 Page 13 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

most promising ways to cut costs in an existing plant are to improve yield, use less raw material, and
reduce energy costs. Steady-state process models can address all of these issues.
Energy and raw material costs are linked to the feed mole ratio. At lower mole ratios, raw material costs
are reduced because less ethylene glycol is consumed. The recycle rate between the esterification
reactors and the distillation columns increases with the mole ratio. At lower mole ratios, the reactor duty
and column condenser duties are reduced.
In this example we applied multivariable optimization to reduce the process costs at constant
throughput. Constraints are defined to keep the intrinsic viscosity, end group concentrations, and DEG
within quality specifications. The objective of the optimization study was to minimize the ethylene
glycol feed rate. In this case, there is much less room for improvement because there is a minimum
feasible mole ratio.
The model identified conditions to achieve a reduction of two percent in the glycol feed flow rate at
constant throughput, which corresponds to a reduction in the feed mole ratio from 1.10 to 1.08. Further,
the model indicates that the major variable influencing the ethylene glycol loss rate is the pressure of the
low polymerizer. The optimization results are summarized in Tables 4-6.

Variable Type Variable Lower Upper Initial Optimized


Bound Bound Value Value
Objective Function EG Mass Flow Rate, kg/hr - - 7147 9626
Constraints Intrinsic viscosity, dl/g 0.595 0.610 0.606 0.605
Acid content, mmol/kg - 40.0 23.0 31.1
DEG content, weight % 1.15 1.35 0.52 1.25
Vinyl content, mmol/kg - 3.00 1.96 2.17
Manipulated Variables Feed mole ratio EG/TPA 1.05 1.25 1.10 1.08
Fraction of EG recycled to PE 0.70 0.99 0.90 0.99
Fraction of EG recycled to SE 0.50 0.99 0.90 0.99
PE temperature, deg C 250.0 270.0 260.0 270.0
PE pressure, Torr 1500 4000 1839 4000
SE temperature, deg C 250.0 280.0 260.0 280.0
LP temperature, deg C 260.0 300.0 275.0 300.0
LP pressure, Torr 50.0 500.0 50.0 500.0
IP temperature, deg C 270.0 305.0 280.0 305.0
IP pressure, torr 1.0 3.0 2.0 2.6
HP temperature deg C 270.0 305.0 285.0 280.5
HP pressure, torr 0.8 2.0 1.0 1.2
Table 4. Case 3 optimization results. The objective is to reduce costs by reducing the excess mole ratio.
Catalyst and additive levels and reactor volumes are held constant.
The ethylene glycol reduction would have been higher if the base-case conditions met the DEG content
constraint. The optimization study increases the pressure and temperature in the primary esterifier to
meet this constraint. Since the DEG content is very sensitive to mole ratio, these variables reach their
upper bounds.

Copyright © 1999 Page 14 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

The stream conditions are summarized in Table 5. These can be compared to the base case conditions
shown above in Table 2.
Reactor PE SE LP IP HP
Antimony Oxide, wt PPM 0 237 241 248 248
Phosphoric Acid, wt PPM 0 0 48 49 50
Degree of Esterification 88.46 97.95 99.13 99.59 99.70
Acid content, mmol/kg 1143.5 203.0 87.6 42.2 31.1
Hydroxyl content, mmol/kg 1788.1 1551.2 1025.5 145.9 72.4
Vinyl content, mmol/kg 0.00 0.04 0.19 2.30 2.17
Saponification, mmol/kg 9911 9886 10032 10308 10331
Intrinsic viscosity, dl/g 0.043 0.063 0.089 0.372 0.605
DEG content, weight % 0.746 0.907 1.184 1.243 1.250
Number average DP 4.13 6.42 9.79 54.97 98.79
Number average Mn 795 1234 1882 10564 18984
Table 5. Stream summary for case-2. Note that the final DEG content meets the specified constraint,
even though the base case conditions did not meet this constraint.

The economic benefits of reducing the ethylene glycol feed flow rate are summarized in Table 6. As
shown, the benefits are relatively small. This reflects today’s low ethylene glycol prices and the fact that
the base case condition was close to the minimum mole ratio. In addition to the 41 k$/year savings in
raw material, there are secondary benefits including reduced heat duty and cooling duty in the
esterification section. The primary esterifier heat duty is decreased by three percent, and the column
condenser duty is reduced by nearly ten percent. There are also additional savings in crude glycol
recovery costs.

Original EG feed rate 2.676 metric ton / hour


Optimized EG feed rate - 2.661 metric ton / hour
Reduction in EG flow rate 0.015 metric ton / hour
x 24 hours / day = 0.360 metric ton / day
x 365 days / year = 131.4 metric ton / year
Price of fiber-grade EG x $ 310 $ / metric ton
Savings in EG costs $ 40,734 $ / year
Table 6. Direct cost savings associated with the reduction in mole ratio at constant throughput. Not
shown are secondary savings including decreased heat and cooling duties and a small reduction in the
amount of crude EG to be purified.

Copyright © 1999 Page 15 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Other Applications of Simulation Technology


Dynamic process models are key tools for improving yield and reducing process variability. Using a
dynamic process model, one can compare alternate control schemes to find the fastest and most stable
control strategy. A dynamic process model can also be used to evaluate and improve grade transition
strategies.
A sufficiently detailed dynamic process model can be adapted for use in operator training systems. For
models to be applied in these applications, they must be able to run in real time. Further, they must
faithfully reproduce the process and control system dynamics. This requires a very detailed and robust
model. Ideally, such models should emulate or communicate through the person-machine interface used
by the real control system.
Rigorous models can support efforts to develop advanced control and on-line optimization systems.
Typically, the complexity and computational performance of first-principles rigorous models prevent
such models from being used directly in on-line applications. These models can be used, however, to
help generate a gain matrix. A gain matrix is a database that represents the response of each process
variable with respect to all other process variables. This database is required to train a fast-running
linearized or neural-net model that can be implemented in open- or closed-loop control.
Conclusions
When used properly, process simulation is a powerful tool to improve process profitability. The melt
polyester process involves many process variables, and the interactions among these variables are
complex. A simulation model captures these interactions and makes them easier to visualize and
understand.
Process optimization is the most powerful off-line application of these models. By simultaneously
adjusting a number of process variables, it is possible to take advantage of tradeoffs between
temperature and residence time, temperature and pressure, and mole ratio and other process variables.
This allows the model to identify improved operating conditions, even in very mature plants. For a
typical polyester process, the economic benefits of applying a model to improve existing process lines
may range into millions of dollars per year.
Dynamic models can be used on-line or off-line to improve the process stability. This results in higher
yields of first quality products, reduced downtime, and shorter grade transitions. These models can be
further leveraged for operator training.
To effectively apply simulation, the models must be tuned and validated against process data. This is
important for two reasons. First, there are a number of parameters, such as heat transfer coefficients and
level calibration curves, which are unique to a particular process line. Second, the models must be well
proven before they gain the political acceptance required to support major engineering decisions in an
operating plant.
The models described herein represents the efforts of many of my associates in Aspen Technology as
well as the input of a number of engineers and managers in the process industries. Further, the models
and data are based on many works published in the scientific literature, especially those listed below.

Copyright © 1999 Page 16 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

References

[1] Cheong, Seong Ill, and Kyu Yong Choi, Modeling of a Continuous Rotating Disk
Polycondensation Reactor in the Synthesis of Thermoplastic Polyesters, J. Applied Polymer
Science, Vol. 61, 763-774 (1996).

[2] Chen, Chau-Chyun, A Segment-Based Local Composition Model for the Gibbs Energy of
Polymer Solutions, Fluid Phase Equilibria, Vol. 83, 301-312 (1993).

[3] Yokoyama, H, T. Sano, T. Chijiwa, and R. Kajya, A Simulation Method for Ethylene Glycol
Terephthalate Polycondensation Process, J. Japan Petroleum Institute, Vol. 21, No. 4, 19-24
(1978).

[4] Yokoyama, H, T. Sano, T. Chijiwa, and R. Kajya, Influence of Catalyst, Stabilizer, and
Temperature on the Polymerization of Ethylene Glycol Terephthalate, J. Japan Petroleum
Institute, Vol. 21, No. 3, 208-210 (1978).

[5] Otton, Jean and Serge Ratton, Investigation of the Formation of Poly(ethylene Terephthalate)
with Model Molecules: Kinetics and Mechanism of the Catalytic Esterification and Alcoholysis
Reactions I. Carboxylic Acid Catalysis (Monofunctional Reactants), J. Applied Polymer
Science: Part A: Polymer Chemistry, Vol. 26, 2183-2197 (1988).

[6] Otton, Jean and Serge Ratton, Investigation of the Formation of Poly(ethylene Terephthalate)
with Model Molecules: Kinetics and Mechanism of the Catalytic Esterification and Alcoholysis
Reactions II. Catalysis by Metallic Derivatives (Monofunctional Reactants), J. Applied Polymer
Science: Part A: Polymer Chemistry, Vol. 26, 2199-2224 (1988).

[7] Ravindrananth, K. and R.A. Mashelkar, Polyethylene Terephthalate I. Chemistry,


Thermodynamics, and Transport Properties, Chemical Engineering Science, Vol. 41, No. 9,
2197-2214, (1986).

[8] Ravindrananth, K. and R.A. Mashelkar, Polyethylene Terephthalate II. Engineering Analysis,
Chemical Engineering Science, Vol. 41, No. 12, 2969-2987, (1986).

[9] Hovenkamp, S.G., Kinetic Aspects of Catalyzed Reactions in the Formation of Poly(ethylene
Terephthalate), J. of Polymer Science Part A-1, Vol. 9 (1971).

Copyright © 1999 Page 17 of 19


Presented at AIChE Spring 1999 Meeting, Houston Texas Session: Polyester Manufacturing Processes I

Polymer Property Definitions

Acid Content Equivalent concentration of acid end groups in polymer, mmole acid / kg liquid:
 2 f TPA + 2 f TPA ( s ) + f t −TPA  6
 10
 liquid mass flow rate, kg/hr 
DEG Content Equivalent weight percent of DEG in the polymer:
 ( f DEG + f t − DEG + f b − DEG )M DEG  2
 10
 liquid mass flow rate, kg/hr 
Esterification Degree of esterification (percent conversion of acid end groups):
( 1 – acid content / saponification )102
IV Intrinsic Viscosity, dl/g:
1.07 x10 −4 (number average molecular weight )
0.83

Saponification Equivalent concentration of ester groups in polymer, mmole ester / kg liquid:


 2( fTPA + f TPA ( s ) + f t −TPA + f b −TPA )  6
 10
 liquid mass flow rate, kg/hr 
Vinyl Content Concentration of oxyvinyl end groups in polymer, mmol / kg:
 2 f t −vinyl  6
 10
 liquid mass flow rate, kg/hr 

Definition of Symbols
Ks Solubility constant of solid TPA
f Molar flow rate, kmole/hr
γ Activity coefficient (apparent mole basis)
i Component index
∆H m Heat of melting of solid TPA
Mi Molecular weight of component “i”
P Total pressure
Pvi Vapor pressure of component “i”
R Ideal gas constant
∆S m Entropy of melting of solid TPA
T Absolute temperature, deg K
xi Apparent mole fraction of component “i” in the liquid phase
yi Apparent mole fraction component “i” in the vapor phase

Copyright © 1999 Page 18 of 19

You might also like