Você está na página 1de 6

Back to Basics

Making Plastics:
From Monomer to Polymer
Pete Sharpe
Emerson Process
Management

Versatility, ease of manufacture, and relatively


low cost make plastics some of the most useful
materials for a wide range of applications.
This article explains the chemistry and production
processes behind two of the most popular plastics
polyethylene and polypropylene.

lastics are some of the worlds most diverse and useful


manufacturing materials. While plastics encompass a
large number of materials, polyethylene and polypropylene are two of the major plastic types found in many
consumer goods, from car parts to shopping bags to water
pipes.
Several types of reactors can make polymers, including
fluidized-bed, loop, autoclave, and tubular reactors. Different
grades of polyethylene and polypropylene offer a wide range
of physical properties, such as density, stiffness, flexibility,
opacity, melting point, texture, and strength. Manipulating
variables in the reactor, such as monomer, comonomer, catalyst, and cooling media flowrates, can control key quality
parameters. Additives and colorings can modify the appearance of the polymer.
Polymer plants are semi-continuous processes. Raw
material is fed into a reactor continuously at the front end,
while polymer powder and pellets are packaged in batches.
Most sites operate multiple lines with many bins and silos
for storage and blending. The plastics are eventually delivered to customers using barges, trucks, or rail cars.
This article describes the different polyolefin production processes, their key operating parameters, and ways
to use automation for improved quality control and higher
throughput.

The chemistry of making plastic


Its helpful to understand some of the chemistry behind
the polymerization reaction to appreciate how the process
works and the complexity involved in making plastic. A
polymerization reaction starts with a primary ingredient
(monomer), such as ethylene or propylene.
24

www.aiche.org/cep September 2015 CEP

Ethylene (C2H4) is a stable molecule with two carbon


atoms and a double bond. Polyethylene (PE) is a made by
the reaction of multiple ethylene molecules in the presence
of catalyst to break the double bond and connect the carbon
atoms into a chain (Figure 1). The longer the chain, the
higher the molecular weight. Polymers can have molecular
weights in the millions.
Similarly, polypropylene (PP) is made by breaking the
double bond in a propylene (C3H6) molecule, in the presence
of a catalyst, to form long chains of three-carbon-atom molecules (Figure 2). The third carbon atom adds a complexity:
On which side of the chain will the methyl (CH3) groups
fall? They could all be on one side of the chains centerline,
or backbone (isotactic), they could appear alternately on
opposite sides of the backbone (syndiotactic), or their positions could be random (atactic). These arrangements have
different physical properties.
Polymerization reactions will also consume hydrogen, which is required to quench the reaction (i.e., end
the chains), and some will involve a secondary ingredient
(known as a comonomer). Since the concentrations of these
components in the reactor affect the probabilities that speH

H
C

H
H

Ethylene

Polyethylene

p Figure 1. Ethylene is a stable molecule with two carbon atoms


connected by a double bond. Polyethylene is made by the reaction of
multiple ethylene molecules in the presence of catalyst.
Copyright 2015 American Institute of Chemical Engineers (AIChE)

cific reactions will take place, the composition in the reactor


effectively sets the amount of branching and the length of
the chain.
The polymer industry employs many catalysts, and new
catalysts are developed every year. Different catalysts are
used to create polymers with particular properties, even in
the same reactor. Each PE or PP process licensor incorporates proprietary catalyst formulations in its reactor designs.
Depending on the type of reactor, the catalysts can be solid
particles or suspended in hydrocarbon or solvent.
Polymerization is a highly exothermic reaction, and it
requires continuous cooling to prevent runaway reactions.
Most reactor systems include an emergency quench that
quickly shuts down the reactor if the temperature reaches
a predetermined setpoint. Before the implementation of
redundant control mechanisms, a runaway reaction could
cause the reactor to become completely plugged with plastic.
Process designs have since changed and safety systems have
been implemented to prevent such incidents.

CH3

CH3

Polypropylene
(Isotactic)

H
C
H

H
n

C
H

Propylene

CH3

CH3

Polypropylene
(Syndiotactic)

H
n

le M
Sing

al
Bimo
d

Probability

oda
l

p Figure 2. Polypropylene is made by breaking the double bond in a


propylene molecule in the presence of a catalyst. The resulting polymer
may be isotactic, with all of the methyl groups on the same side of the
polymer backbone, syndiotactic, with alternate methyl groups on opposite
sides of the backbone, or atactic (not shown), with the methyl groups
arranged randomly.

Molecular Weight

p Figure 3. The molecular weight distribution of a bimodal polymer has


more than one peak.

Copyright 2015 American Institute of Chemical Engineers (AIChE)

Key quality specifications


Although polymer properties can be modified somewhat
in the blending and extruding steps, the conditions in the
reactor set the product grade(s) as defined by a few key
quality measures that can be made. In general, poly
ethylene is categorized based on its density, the linearity of
the molecules (i.e., degree of branching), and its molecular
weight (length of the chains). Other polymer qualities,
including the molten flowing property, known as the meltflow index, are a function of the crystalline structure and are
also primarily determined by the polymerization reaction.
The melt-flow index determines the polymers behavior in
downstream operations, such as extrusion, blow molding, or
film production.
The most common polyethylene grades are:
high-density polyethylene (HDPE). This polymer has
a density greater than or equal to 0.941 g/cm3. It has a low
degree of branching, with mostly linear molecules, has high
tensile strength, is resistant to many chemicals, and is used
in such products as bottles, jugs, water pipes, and toys.
medium-density polyethylene (MDPE). With a density
range between 0.925 and 0.940 g/cm3, MDPE has better
shock and stress crack resistance than HDPE and is commonly used for gas pipes, plastic bags, and packaging films.
low-density polyethylene (LDPE). This grade has a
density range of 0.9100.925 g/cm3 and a high degree of
short branching. Its loosely packed crystal structure gives it
a lower tensile strength, but desirable molten flow properties. LDPE can be used for both rigid containers and flexible films.
linear low-density polyethylene (LLDPE). This highly
linear polymer has a very large number of short branches. It
has higher tensile strength than LDPE, and can be made into
thinner films with high puncture and stress crack resistance.
LLDPE is commonly used in packaging films and bubble
wrap, as well as cable sheathing, toys, and buckets.
Each of these major grades of polyethylene includes hundreds of different formulations. The grades and formulations
can be blended to create even more polymers each with a
unique set of properties. Bimodal polymers, whose molecular weight distributions have more than one peak (Figure 3),
can also be produced, further expanding the range of plastics
properties.
Similarly, polypropylene has a few main product grades
with many subgrades. The major polypropylene grades are:
homopolymer consists of only propylene, with no
comonomer
random copolymer contains comonomer (typically
ethylene) randomly arranged in the polymer chain
blocked copolymer contains one or more comonomer
units alternating with monomer units.
Polypropylenes melt-flow index is a measure of its
CEP September 2015 www.aiche.org/cep

25

Back to Basics

26

www.aiche.org/cep September 2015 CEP

Reactor
Section

Intermediate
Storage

Extrusion

Catalyst
Preparation
Intermediate
Storage

Bagging,
Bulk Loading

p Figure 4. A polymer production process typically involves these general


steps.
Quench Water

Feed Hopper

Die

Pelletizer
Blade Drive

Main Drive
Barrel Heating
Units

Water and
Pellets to
Classifier

p Figure 5. In an extruder, a screw conveys polymer powder from a feed


hopper to a die. Along the way, the powder moves through heated units,
which cause it to melt. At the end of the extruder, the die shapes the
polymer to the desired width, and a blade cuts the polymer into pellets of
the desired length.
Compressor

Compressor

Reactor 2

Polymer reactor systems


The heart of a PE or PP plant is the reactor system. There
are many different licensors and reactor configurations, each
of which has competitive advantages depending on the types
of polymers to be made. This brief overview of the major
reactor types is not intended to be a comprehensive review
of all reactor configurations. Additional information is available from the licensors and in Ref. 1.
Fluidized-bed reactors. Union Carbide introduced one
of the first gas-phase polymer process technologies in the
1970s under the tradename Unipol. This process, now
licensed by Dow, uses a low-pressure, fluidized-bed reactor
to make polymer powder. As shown in Figure 6, monomer,
comonomer, and catalyst are injected into a recycle gas loop

Feed
Preparation

Reactor 1

The basic manufacturing process


Polyethylene and polylpropylene production processes
involve similar operations (Figure 4).
PE and PP processes require very pure monomer and
comonomer(s) to minimize undesirable side reactions that
might affect the polymer structure and properties. Most
ethylene and propylene producers make both chemical and
polymer grades. Typical polymer-grade ethylene and propylene are at least 99.95% pure, with limits on certain impurities (e.g., acetylene and propadiene) in the parts-per-million
range. Depending on the feed source, some polymer plants
require one or more distillation columns at the front end to
purify the raw material or to recycle the unreacted monomers to the reactors.
In the catalyst preparation area, solid or liquid catalysts
are mixed, activated, and loaded into the catalyst feeding
system. The catalyst is fed, together with the raw materials,
to the reactor section. Solvent and unreacted monomers are
separated and recycled to the process, while the polymer
powder is sent to intermediate storage or directly to extruder
feed bins.
An extruder (Figure 5) is a large screw-type device
with a heated barrel that compresses and melts the polymer
powder and forces it through a die with many small holes.
Additives and coloring may be added to the polymer in the
extruder. On the other side of the die, blades cut the extruding plastic into pellets; water quenches the pellets and carries
them to a classifier, which separates the normal-sized pellets
from the clumps. The pellets are conveyed (typically pneumatically) to storage silos for further classification, blending,
and shipping. The final step in the process is either packaging via bagging or loading into bulk shipping containers,
such as rail cars or trucks.

and fed into the bottom of the polymer bed. Polymer product
is withdrawn from the bed and sent to a powder-disengaging
drum, where the gases are purged, and the powder is routed
to a product silo. Unreacted reactor gas is compressed,
cooled, and recycled to the bottom of the reactor.
In some cases, two reactors are operated in series. Polymer made in the first reactor is fed to a second reactor where
a comonomer can be introduced. This configuration uses
the resin from Reactor 1 as the building blocks to produce
blocked copolymers in Reactor 2.

Reactor

average molecular weight. Polypropylene resin has a wide


range of applications, from blow-molded containers to films
to fibers, and more.

Polymer

Polymer
Powder
Monomer
Comonomer
Catalyst

Comonomer
Resin Transfer
Monomer
Catalyst

p Figure 6. In a gas-phase fluidized-bed polymerization process (blue),


monomer, comonomer, and catalyst are injected into the bottom of the
fluidized-bed reactor and the polymer is produced as a powder. In a dualreactor system (green), two fluidized-bed reactors are operated in series,
and comonomer is introduced in the second reactor.
Copyright 2015 American Institute of Chemical Engineers (AIChE)

In addition to Dow, INEOS Technologies and LyondellBassell license gas-phase, fluidized-bed reactor systems.
INEOS and Japan Polypropylene Corp. license gas-phase,
horizontal, stirred-tank reactor systems. CB&I also offers
gas-phase process technology reactors, but with an agitated
polymer bed (1).
Loop reactors. Another common scheme is the highpressure loop reactor (Figure 7). Monomer, comonomer,
and catalyst are injected into the lower part of the loop; the
polymer slurry is pumped around the loop and the product
is removed as a slurry from the opposite side of the loop. A
cooling water jacket on the reactor removes the reaction heat
to maintain a constant temperature. The product slurry goes
through a series of flash drums, where the solvent and hydrocarbons are removed for recycling to the process. The powder is transported, typically pneumatically, to a pelletizer and
product storage bins for further processing. In some cases,
a second loop reactor is operated in series to make bimodal
polymers. INEOS, LyondellBassell, and Chevron Phillips
Chemical (among others) license loop reactor designs.
Autoclave and tubular reactors. Some processes employ
high-pressure, stirred autoclave reactors that operate at pressures in the range of 2,0003,000 bar (29,00043,000 psi).
ExxonMobil Chemical, Eni Versalis, and LyondellBassell
(among others) license autoclave reactor systems. Jacketed,
tubular reactors are another option. Autoclave and tubular
reactors are much less common than fluidized-bed and loop
reactors, and are not discussed further here.

Logistics complexities
By their very nature, polymer plants present significant
logistical challenges that require systems and experienced
personnel to operate efficiently.
The primary feedstocks ethylene and propylene
are gaseous at ambient conditions. Thus, they cannot simply
be stored in tanks, and are delivered via pipeline. Some producers operate large salt-dome caverns that can store liquid
ethylene and propylene under high pressure to accommodate
swings in demand. Any changes in the reactor consumption
must be communicated and coordinated with the pipeline
and suppliers. Typically, once a reactor train is started, it is
kept running as long as possible. Production transitions from
one recipe to another on the fly, with the objective to minimize or eliminate the amount of off-specification polymer
produced during the transition.
Most plants produce a large number of different polymer grades in each reactor, so planning and scheduling how
much of which products to make in which reactors and
where to store those products is a dynamic undertaking.
Planning starts with understanding customer demand,
evaluating the available inventories, and determining which
grades should be produced in which order in each reactor.
Copyright 2015 American Institute of Chemical Engineers (AIChE)

Production orders are passed to operations as a target


recipe with the quality specifications for A-grade (target)
material. The recipe may contain information on the catalyst
preparation, reactor conditions, target monomer and comonomer ratios, additives, and pelletizing instructions.
Off-spec polymer produced during grade transitions
must also be considered when scheduling a reactor. The
product made during the transition between similar grades
is typically within specifications, but switching between
dramatically different recipes can involve several hours of
off-spec production. Plants either sell off-grade product at a
lower price or blend it into an A-grade batch.
The production of a batch starts with the catalyst preparation. Depending on the type of reactor, catalyst may be
stored in tanks, drums, or bags, and may require mixing,
weighing, and/or sampling. Catalyst may need to be combined with an activator before it is loaded into the catalyst
charging system. Since catalyst performance has a significant impact on the polymers produced, many companies try
to track each catalyst batch and the product(s) it is used to
make. This requires a product genealogy system, which
can track the catalyst batch, as well as the reactor conditions, extruder conditions, additives, and lab qualities of the
powder and pellets throughout the production process, down
to a particular customer shipment.
Tracking genealogy is extremely difficult, because the
hold-up silos that store powder and pellets throughout the
production process are not always completely emptied
before new material is transferred into them. Because powders tend to build up on the sides of the silos, they are not
always well-mixed, and a lab sample taken at the bottom
Circulation
Pump
Solvent to
Recycle

Loop
Reactor

Cooling
Water
Jacket

Flash
Drum

Polymer
Product

Catalyst
Monomer
Comonomer
Solvent

p Figure 7. In a loop reactor system, monomer, comonomer, and catalyst


enter the reactor at the bottom of the loop, and the product, in the form of
a slurry, is withdrawn from the opposite side of the loop. The product slurry
is sent to a series of flash drums where the solvent and hydrocarbons are
removed for recycling to the process.

CEP September 2015 www.aiche.org/cep

27

Back to Basics

of the silo is not necessarily representative of the silos entire


inventory.
At the end of the production process, the polymer (as
pellets or powder) is loaded into bulk shipping containers, or
bagged and put on pallets. The final grade is then determined,
and composite samples (i.e., small samples captured over the
loading period) are analyzed for quality-control purposes.

Advanced automation opportunities


Over the past decade, improvements in digital instrumentation and advanced control and information systems
have enabled plastics producers to improve A-grade production, quality control, safety, reliability, and customer service.
Some of the technologies that have enabled this transition
are highlighted in the following sections.
Measurement and control technologies
Polymer plants face some unique measurement challenges. The primary manipulated variables for the reactor are
the catalyst-addition rate, the flowrate of the reactor cooling medium, and the monomer, comonomer, and hydrogen
flowrates. Precise measurement of the reactor feed streams
and tight control to their stoichiometric targets is critical to
maintain the desired composition inside the reactor and, ultimately, the polymers properties. Coriolis meters are a best
practice for feed stream and slurry measurements because of
their accuracy, high turndown ratio, and ability to simultaneously measure mass flow and density of multiphase streams.
The catalyst-addition rate and the reactor temperature
set the reaction rate. Because the potential for runaway
reactions exists, robust reactor temperature control is critical, so most reactors have redundant temperature measurements. In addition, fast-acting, nonlinear control algorithms
are used to ensure an aggressive response to any increase in
temperature (2). Fuzzy logic controls are a good option for
nonlinear control response.
Impulse lines that connect pressure, flow, and level

instruments to the process are prone to plugging with polymer powder, so they are typically purged with nitrogen as a
precaution. Advanced diagnostics in some new pressure and
differential-pressure transmitters use the noise signature of
the impulse line to detect the onset of plugging in a pressure
tap (Figure 8).
It can be a challenge for operators to accurately measure
the level of resin inside a fluidized-bed or stirred reactor
system and in storage silos. Noncontacting radar level detectors are recommended for internal reactor level measurements, particularly if the reactor contains agitators. Some
sites employ load cells in silos for mass measurements, but
because level is also a function of product density, overflow
can still occur.
The best practice for measuring level in most powder
and pellet silos is guided-wave radar (Figure 9), because
it can handle products with very low dielectric properties
and it is responsive to rapid level changes, independent of
density. Three-dimensional acoustic sensors and solid-state
level sensing technology, which use sound waves with three
separate antennas to measure the surface contours, provide a
3D image of the silos inside surface. However, 3D sensors
are limited by a minimum diameter-to-height ratio and can
only be used in large vessels (e.g., the minimum diameter for
a silo-type structure is approximately 9 ft).

Advanced process control


Control of the polymer composition in the reactor is
a multivariable problem, and the manipulated variables
(monomer, comonomer, catalyst, and cooling media
flowrates) interact with each other and influence the final
product qualities (density, melt-flow index, and molecular
weight distribution).
Advanced, multivariable control techniques have been
used successfully to control polymer reactor systems for
many years, although certain aspects of the polymerization process make it a difficult application (3). Some of the
Guided Wave Radar

3D Acoustic Sensor

Impulse Line
Plugging
View From
Transmitter

Inferred
Volume

Measured
Volume

View from
Distributed
Control System

p Figure 8. Plugging of an impulse line can be identified by detecting a


change in the noise in a signal from the pressure tap. As the tap begins to
plug, the noise is dampened.

28

www.aiche.org/cep September 2015 CEP

p Figure 9. Guided-wave radar can measure the level of pellets or powder


in a silo and sense rapid level changes, independent of the density of the
material. Three-dimensional acoustic sensors measure level by mapping
the surface topography of the material in the silo.

Copyright 2015 American Institute of Chemical Engineers (AIChE)

primary product qualities (e.g., density, molecular weight


distribution) cannot be readily measured online, so inferential models are often used to predict resin qualities based on
laboratory analysis of periodic samples. Traditional inferential modeling techniques, such as partial least squares (PLS)
models, linear quadratic estimation (i.e., Kalman filters), and
neural nets, have been used in the past (3). Online analyzers,
combined with heat-and-material balance calculations, allow
estimates of reaction rate and consumption of each component. These calculations can be used for feedforward control
to adjust monomer and comonomer feed rates, reactor cooling medium flowrate, and catalyst-addition rates (4).
Modeling polymer reactors is complex, because the
process is inherently nonlinear. Historically, gain scheduling
or switching between multiple controllers have been used
with linear model predictive control (MPC) tools to accommodate this process nonlinearity. More recently, nonlinear
controllers have been applied to provide control through
grade transitions, reportedly reducing off-grade production
by 2550% (2). However, there is little to no literature that
documents the robustness of these solutions and the level of
technical support that is required to keep them running.

Plant information and operations systems


Many parts of the organization sales, production planning, plant operations, customer service, logistics, quality
control, maintenance, plant management the list goes
on need access to plant data. Therefore, most sites have
systems and at least a certain level of integration to facilitate
the flow of information throughout the production cycle.
Figure 10 illustrates a typical plant architecture for a
polymer processing facility. The extent to which the plant
production systems are integrated into the full enterprise
supply chain varies from company to company. Several large
plants have implemented operations management systems
that coordinate the planning, scheduling, and execution functions required between the entry of customer orders into the
Business Systems (ERP, Dispatch)

Realtime
Database

Batch/
Planning
Recipe
Genealogy
and
Management
Tracking
Scheduling

Inventory
and
Logistics

Operations Management Systems

Laboratory
Information
System

Control Systems (DCS, PLC)

Instruments, Analyzers, Valves

p Figure 10. This polymer manufacturing plant architecture integrates


enterprise resource planning software and operations management
systems.

Copyright 2015 American Institute of Chemical Engineers (AIChE)

enterprise resource planning (ERP) software to the notification that the products have been shipped to fulfill the orders.
Integrated operations management systems can provide
benefits throughout the organization, including (4):
more-efficient sales and customer service
enhanced productivity
improved product quality
reduced inventory
more-efficient planning and scheduling
improved safety.

Looking ahead
According to PlasticsEurope, worldwide plastics
demand has been growing at a compound annual growth
rate (CAGR) of 6.7% since 2009, and reached almost
300 million m.t. in 2013 (5). With hundreds of different
polymer formulations available and new ones being developed every year, engineered polymers, made-to-order for
specific customers with customized properties, are becoming
reality. New business models will motivate polymer producers to apply modern automation and information technology
to measure and control the production process, plan and
manage grade transitions, and integrate functions of sales,
planning, operations, and logistics systems. Opportunities
CEP
abound for those who do it successfully.

PETE SHARPE is the Director, Industry Solutions Development for Emerson


Process Management. He holds a BS in chemical engineering from the
Univ. of Colorado and an MBA from the Univ. of Houston. He has over
36 years of experience in the process control industry, in both technical
and management roles, and specializes in advanced process control.
He has helped refiners, petrochemical manufacturers, utilities, and
polymer producers around the world design, justify, and implement
major automation projects and control system upgrades. He is currently responsible for managing the engineering and development of
Emersons industry-specific applications for monitoring and control of
plant assets.

Literature Cited
1. Hydrocarbon Processing, 2014 Petrochemical Processes Handbook, Gulf Publishing Co., Houston, TX (May 2014).
2. Naidoo, K., et. al., Experience with Nonlinear MPC in Polymer
Manufacturing, in R. Findeisen, et al., eds., Assessment and
Future Directions of Nonlinear Model Predictive Control,
Springer-Verlag, Berlin, Germany, pp. 383398 (2007).
3. White, D. C., Polyolefin Plants: Profitability Improvements
from Advanced Control and Information Systems, Chemical
Engineering, 105 (2), pp. 102108 (Feb. 1998).
4. Richards, J. R., and J. P. Congalidis, Measurement and
Control of Polymerization Reactors, Computers and Chemical
Engineering, 30 (1012), pp. 14471463 (Sept. 2006).
5. PlasticsEurope, Plastics the Facts 2013, Association of
Plastics Manufacturers, www.plasticseurope.org/documents/
document/20131014095824-final_plastics_the_facts_2013_published_october2013.pdf (2013).

CEP September 2015 www.aiche.org/cep

29

Você também pode gostar