Sei sulla pagina 1di 31

Znd. Eng. C h e m . Res.

1994,33, 449-479

449

REVIEWS
Gas Phase Ethylene Polymerization: Production Processes, Polymer
Properties, and Reactor Modeling
Tuyu Xie, Kim B. McAuley,' James C. C. HSU,and David W. Bacon
Department of Chemical Engineering, Queen's University, Kingston, Ontario, Canada K 7 L 3N6

A review of relevant macroscopic and microscopic processes of gas phase ethylene polymerization,
both chemical and physical, is given. The commercial technology development of gas-phase ethylene
polymerization processes is illustrated through a selective survey of the patent literature. Both
advantages and disadvantages of gas phase polymerization processes are addressed, and the challenges
of laboratory studies of gas phase polymerization are also outlined. Physicochemical phenomena
of ethylene polymerization using heterogeneous catalysts are discussed, including examination of
catalyst preparation, polymer morphological development, and elementary chemical reactions.
Metallocene-based catalysts and their kinetic performance for olefin polymerizations are also
discussed. The current state of the art for reactor modeling of polymerization rate, molecular
weight development, reactor dynamics, and resin grade transition strategies is illustrated on the
basis of the most recent academic studies. Finally, relationships between resin properties and
polymer microstructures as well as characterization methods are described briefly. In particular,
temperature-rising elution fractionation technology is emphasized for characterization of ethylene
copolymers. The fundamental issues involved in gas phase ethylene polymerization and the&
interrelationships are also discussed in some detail.
Contents
1. Introduction
2. Gas Phase Polymerization Processes
2.1. Commercial Gas Phase
Polymerization Processes
2.2. Experimental Methods
3. Physicochemical Phenomena
3.1. Catalysts for Gas Phase
Polymerization
3.2. Polymer Particle Morphology
Developments
3.3. Chemical Reactions
4. Reactor Modeling
4.1. Kinetic Modeling
4.2. Dynamic Process Modeling and
Control
5. Polymer Properties and
Characterization
5.1. Physical and Mechanical
Properties
5.2. Polyethylene Characterization
6. Summary
7. References

449
450
450
455
456
456
460
461
463
463
467
469
469
470
474
474

1. Introduction

Polyethylene (PE) is the largest synthetic commodity


polymer in terms of annual production and is widely used
throughout the world due to its versatile physical and
chemical properties. The American Society for Testing
and Materials (ASTM) has classified PE into four groups:
I (low density) at 0.910-0.925 g/cm3, I1 (also low density)
at 0.926-0.940 g/cm3, I11 (high-density copolymers) at
0SSS-5SS5/94/2633-0449$04.50/0

0.941-0.959 g/cm3, and IV (high-density homopolymer)


at 0.960 g/cm3and above (Redman,1991). In the literature,
however, PE is normally classified as low density (LDPE)
at 0.910-0.930 g/cm3 and high density (HDPE) a t 0.9310.970 g/cm3. Low-densitypolyethylene is further classified
as low-densitypolyethylene (LDPE) and linear low-density
polyethylene (LLDPE) based on polymer chain microstructure and synthesis processes. According to the figures
reported in Mod. Plast. (19931, polyethylene production
in the United States alone was over 10 million tons in
1992. The annual production of PE in Europe is about
9 million tons (Redman, 1991). The current annual
worldwide capacity for PE production is over 30 million
tons. Figure 1shows the US. PE production profile over
the past-decade. Although the annual rate of increase
slowed down slightly at the end of the 1980s, the average
annual increase rate is about 8 94 for HDPE and about 5 %
for LDPE and LLDPE for the past decade. Consumption
of PE is still rising through the 1990s with development
of synthesis and processing technology. The main markets
and applications of LLDPE and HDPE are summarized
in Figure 2 after James (1986) and Foster (1991).
Polyethylene is commercially produced exclusively by
continuous processes. On the basis of polymerization
mechanisms and reactor operating conditions, PE production processes can be classified into at least five process
categories as shown in Table 1. Among them, the gas phase
polymerization process is the most recently developed and
also the most versatile. Since its emergence, this process
has been challenging other existing processes for market
share, particularly, for production of LLDPE, due to its
economic and technological advantages. Many excellent
reviews of ethylene polymerization processes have been
published (Vandenbergand Perka, 1977;Short, 1981;Choi
and Ray, 1985a; Nowlin, 1985; James, 1986; Beach and
Kissin, 1986). However, the fundamental issues involved
0 1994 American Chemical Society

450 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994


'Ooo

eI-z

zation can be traced back to the 1950s. Dye (1962),whose


L
patent was filed in 1957, was perhaps the first to adopt a

6000 -

LDPE & LLDPE

f
HDPE

vj

2ooo

.""" 80

82

86

84

88
90
TIME, year

92

94

96

Figure 1. U S . polyethylene production profile (data from Mod.


Plast. (198C-1993)).

0.97 0.96 -

- 0.95 rn

0.94

v)

z
W

0 0.93 -

SHEET

0.92 -

0.91 0.90 1

0.01

1.oo
10.0
MELT INDEX, 9/10 min

0.10

100

Figure 2. Markets and applications for low-pressure polyethylene.

in gas phase polymerization processes have not been


comprehensively discussed. With development of commercial gas phase polymerization processes, the importance of more complete understanding of gas phase
polymerization processes has been recognized by the
academic community. Significant experimental and reactor modeling work has been carried out in recent years.
In the present review, the authors focus on the main issues
involved in gas phase ethylene polymerization processes,
including commercial production technology development,
physicochemical phenomena, kinetictdynamic reactor
modeling and control; polymer properties and characterization, as well as their interrelationships.
2. Gas Phase Polymerization Processes

The distinguishing characteristic of gas phase polymerization is that the system does not involve any liquid
phase in the polymerization zone. Polymerization does
occur at the interface between the solid catalyst and the
polymer matrix, which is swollen with monomers during
polymerization. The gas phase plays a role in the supply
of monomers, mixing of polymer particles, and removal of
reaction heat. Hence, gas phase polymerization is also
called dry polymerization in some patents (Dormenval et
al., 1975; Havas and Mangin, 1976). In this section,
commercial gas phase polymerization processes and experimental studies of gas phase polymerization are discussed.
2.1. Commercial Gas Phase Polymerization Processes. The invention of gas phase ethylene polymeri-

fluidized bed reactor for gas phase olefin polymerization.


The original reactor consisted of three concentric superimposed vertical sections. Polymer particles were discharged through an extruder, which was connected to the
bottom section of the reactor. The reactor was operated
at a pressure of 30 atm and a temperature of 100 "C for
ethylene polymerization. Goins (1960)carried out ethylene
copolymerization in a countercurrent fluidized bed reactor
in the presence of inert diluent gas. In this process,
polymer particles are passed downward in the reactor and
monomer mixed with diluent gas is passed countercurrently upward in the reactor and monomer mixed with
diluent gas is passed countercurrently upward through a
series of vertical fluidized bed reaction zones. The reaction
zones can be controlled independently by taking off-gas
from the last reaction zone, cooling it, and recycling
portions of such off-gas to each of the reaction zones. Both
patents (Dye, 1962;Goins, 1960)were assigned to Phillips
Petroleum Company. Schmid et al. (1967) carried out
ethylene polymerization in a stirred fluidized bed reactor.
In this configuration, polymer particles are moved in the
direction of monomer flow by stirring, and reaction heat
is removed by cooling the walls of the reactor, by the gas
stream, and by introduction of liquified monomers. The
patent of Schmid et al. was assigned to BASF. The benefits
of gas phase polymerization were recognized by these
pioneer inventors. Although these inventions have not
been directly applied in commercial gas phase polymerization processes, the fundamental ideas demonstrated
by these inventors provided the foundation for the later
commercial gas phase process development.
The first commercial gas phase polymerization plant
using a fluidized bed reactor was constructed by Union
Carbide in 1968 at Seadrift, TX (Rasmussen, 1972;
Batleman, 1975; Burdett, 1988). This process was developed initially for HDPE production. The success of this
novel technology led to the extension of the process to
LLDPE, which was produced initially on a commercial
reactor in 1975 (Davis, 1978; Burdett, 1988). The Union
Carbide gas phase process, commonly called UNIPOL,
has been licensed worldwide with more than 25 licenses
operating in 14 different countries (Burdett, 1988). Production of LLDPE using gas phase processes is more
difficult than production of HDPE because the difference
between the melting point and polymerizationtemperature
is much narrower for LLDPE. The catalyst types and
equipment design developed for HDPE cannot be used to
produce LLDPE because of the potential for agglomeration
of polymer particles. Hence, significant engineering and
chemistry research has been required to assure the success
of gas phase LLDPE production. According to Karol
(19831,the keys to the success of the UNIPOL technology
for LLDPE production are the proprietary catalysts that
operate at low pressure and low temperature and which
are suitable for use in a gas phase fluidized bed reactor.
Union Carbide Corporation received the 1979Kirkpatrick
Chemical Engineering Achievement Award in recognition
of the innovation of the UNIPOL process (Chem. Eng.,
1979).
The UNIPOL process has been described in several US.
Patents (Miller, 1977; Levine and Karol, 1977; Karol and
Wu, 1978; Wagner et al., 1981; Jorgensen et al., 1982). A
simplified flow diagram is shown in Figure 3. The fluidized
bed reactor consists of a reaction zone and a disengagement
zone. The reaction zone has a height to diameter ratio of
about 6-7.5. The disengagement zone has a diameter to

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 451


Table 1. Polymerization Processes and Reactor Operating Conditions

reactor type

conventional
high-press. process
tubular or autoclave

high-press.
bulk process
autoclave

solution
Polymn
CSTR

reactor press., atm


temp, "C
polymn mech
loci of polymn
density, g/cm3
melt index, g/10 min
reference

1200-3000
130-350
free radical
monomer phase
0.910-0.930
0.10-100
Doak (1986)

600-800
200-300
coordination
monomer phase
0.910-0.955
0.80-100
Grunig and Luft (1986);
Villermaux et al. (1989)

100
140-200
coordination
solvent
0.910.970
0.50-105
James (1986)

GAS RECYCLE

CATALYST
FEEDER

GAS FEED
7

EXCHANGER

Figure 3. Union Carbide gas phase ethylene polymerization


process-UNIPOL (Miller, 1977;Levine and Karol, 1977;Karol and
Wu, 1978; Wagner et al., 1981; Jorgensen et al., 1982).

height ratio of about 1-2. To maintain a viable fluidized


bed, superficial flow through the bed is about 2-6 times
the minimum flow required for fluidization. It is essential
that the bed always contain polymer particles to prevent
the formation of localized "hot spots" and to entrap and
distribute the powdery catalyst. On startup, the reaction
zone is usually charged with a base of polymer particles
before gas flow is initiated.
The catalyst is stored in a catalyst feeder under a
nitrogen blanket. Catalyst is injected into the bed at a
rate equal to its consumption rate at 114 to 314 of the
height of the bed. Catalyst concentration in the fluidized
bed is essentially equal to the catalyst concentration in
the product, namely on the order of about 0.005-0.50%
of bed volume. Fluidization is achieved by a high rate of
gas recycle to and through the bed, typically on the order
of about 50 times the rate of feed of make-up gas. The
pressure drop through the bed is typically on the order of
1 psig. Make-up gas is fed to the bed at a rate equal to
the rate at which polymer particles are withdrawn. A gas
analyzer, position above the bed as shown in Figure 3,
determines the composition of the gas being recycled, and
the make-up gas composition is adjusted accordingly to
maintain an essentially steady-state gaseous composition
within the reaction zone.
The gas which has not been consumed in the bed passes
through the enlarged disengagement zone where entrained
particles drop back into the bed. Particle entrainment is
further reduced by a cyclone and a filter to avoid deposition
of polymer on heat-transfer surfaces and compressor
blades. Polymerization heat is removed by a heat exchanger before the recycle gas is compressed and returned
to the reactor. The fluidized bed can maintain itself at

slurry polymn
loop or CSTR
30-35
85-110
coordination
solid
0.930.970
<0.01-80
Choi and Ray (1985a);
Short (1981)

gas phase
polymn
fluidized or
stirred bed
30-35
80-100
coordination
solid
0.910-0.970
<0.01-200
James (1986)

an essentially constant temperature under steady-state


conditions. To increase heat removal capacity and productivity, UNIPOL reactors can be operated with an inlet
gas temperature at the bottom of the fluidized bed which
is below the dew point temperature of that gas. Recycle
gas condenses in the external cooler, and the liquid droplets
revaporize upon entry into the bed. It has been observed
that droplets pass through the distributor as a mist and
quickly wet the surface and pores of the polymer particles.
The liquid vaporizes very quickly above the distributor.
There is a dramatic change in the gas flow pattern in the
gas feed region below the distributor. However, there is
no apparent effect on bulk circulation patterns of bubbles
and resin particles above the gas distributor (Burdett,
1988).
The distribution plate at the bottom of a reactor plays
an important role in the operation of the reactor. As the
polymer particles are hot and possibly active, they must
be prevented from settling to avoid agglomeration. Maintaining sufficient recycle and make-up gas flow rate
through the distributor to achieve fluidization at the base
of the bed is very important in fluidized bed polymerization
reactor operation. The polymer particles are withdrawn
close to the distributor through sequential operation of a
pair of timed valves, defining a segregation zone as shown
in Figure 3.
The reactor is operated at a temperature below the
melting point of the polymer particles. For HDPE
production, the operating temperature is 90-110 "C. An
operating temperature of about 90 "C or lower is preferred
for production of LLDPE, which contains about 15 mol
% of one or more of the Ca to c6 &-olefins (Levine and
Karol, 1977).
The reaction zone and disengagement zone of the reactor
shown in Figure 3 are connected by a transition section
having sloped walls. During polymerization, some fine
particles from the disengagement zone fall onto the sloped
walls of the transition section. These fine particles build
up during reactor operation. Since the fine particles
contain active catalyst, they further polymerize and form
solid sheets which can grow until they block recycle gas
flow or slide off the sloped walls into the polymerization
zone. In the polymerization zone, the solid sheets block
the flow of gas and cause fusion of the polymerization
particles. Thus, large chunks of polymer can be formed
which can block the entire polymerization zone unless the
reactor is shut down and the sheets areremoved. To avoid
this problem, Union Carbide designed a modified fluidized
bed reactor shown in Figure 4 (Brown et al., 1981). In this
process, the enlarged disengagement zone, cyclone, and
filter are eliminated. Polymerization heat is removed by
an internal heat exchanger. Reactor temperature is
controlled by adjusting the temperature or flow rate of
coolant. According to the patent (Brown et al., 1981),
ethylene copolymers with densities of 0.91-0.97 gIcm3 and
melt indexes of 0.1-100 or more can be produced using

452 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

CATALYST
FEEDER
INERT GAS

GAS FEED

COMPRESSOR

l h

ACTIVATOR
FEEDER

Figure 4. Gas-phase ethylene polymerization process with internal


heat exchanger (Brown et al., 1981).

this process. Compared with the process shown in Figure


3, this modified process has the following advantages:
1. The formation of sheets on the sloped walls of the
transition zone is eliminated so that the frequency of
reactor stoppages is reduced.
2. The depth of the bed in the polymerization zone can
be varied over a wide range allowing a greatly increased
rate of reactor output with good operation.
3. A smaller initial charge of powdered material is
required to start up successfully.
4. The fabrication cost of a fluidized bed reactor is
reduced.
5. Gas velocity can be much lower using an internal
heat exchanger, thus reducing operation energy.
6. Since heat removal is independent of gas mass flow
rate, the reactor pressure can also be decreased to a limit
defined by the polymerization kinetics.
7. The internal heat exchanger also acts as a sets of
baffles, and tends to deter the migration of bubbles to the
center of the bed and to increase mixing near the reactor
walls.
8. Recycle gas is at the same temperature as the entire
fluidized bed.
Despite these advantages, there are no published reports
on the application and performance of this type reactor.
About 1964 BASF first developed a gas phase ethylene
polymerization process using Phillips catalysts (Wisseroth,
1969;Sailors and Hogan, 1981),and a commercialgas phase
HDPE plant was built in Germany in 1976 (Choi and Ray,
1985a). BASF uses a continuous stirred-bed reactor for
gas phase ethylene polymerization as shown in Figure 5
(Trieschmann et al., 1977; Choi and Ray, 1985a). The
reactor is equipped with an anchor agitator and is operated
at higher pressure and temperature (35 atm, 100-110 OC)
than those employed in UNIPOL. Polymer is discharged
through a cyclone. Unreacted monomer and the heattransfer agent, such as cyclohexane, are removed through
the top of the reactor and liquefied by a condenser. The
liquefied recycle monomer and cyclohexane are fed into
the reactor through a valve located at the bottom of the
reactor, which is controlled by a thermostat. Liquefied
monomer and cyclohexane can be fed into the reactor by
a metering pump or by provision of an adequate static
pressure difference between condenser and reactor. Fresh
monomer is also fed through the bottom of the reactor.
The temperature in the polymerization zone can be
maintained at a constant level by controlling the rate of
feed of liquefied monomer and cyclohexane. Liquefied

RECYCLE

Figure 5. BASF gas phase ethylene polymerization process


(Trieschmann et al., 1977; Choi and Ray, 1985a).
RECYCLE MONOMER-HYDROGEN-QUENCH VAPOR

CATALYG

SPYY NOZZLES

-h

MONOMERAND HYDROGEN INLETS

I!
POLYMER
DISCHARGE

Figure 6. AMOCO gas phase polymerization process (Shepard et


al.,1976;Jezl et al., 1976;Peters et al.,1976;Jezl and Peters, 1978a,b).

monomer and cyclohexane vaporize in the reaction zone


and rise through the bed at a very slow velocity of about
2 cm/s so that the polymer particles are not fluidized. The
bed is mixed by agitation with a mean circulation time of
30-100 s. Fresh monomer is added at a rate such that the
pressure in the system remains constant. The pressure
and temperature in the polymerization zone are maintained at levels such that the vaporized monomer and heattransfer agent remain in the gaseous state. Magnesiumsupported titanium halides and aluminum alkyl compounds and silica-supported modified chromium oxide
catalysts are used for HDPE production.
The AMOCO (Standard Oil Company) gas phase
polymerization process was developed in the mid-1970s
(Shepard et al., 1976; Jezl et al., 1976; Peters et al., 1976;
Jezl and Peters 1978a,b). A compartmented horizontal
stirred bed reactor is used in this process as shown in
Figure 6. The reactor is divided into several stirred
polymerization zones which are separated by weirs. Makeup and recycle monomer and hydrogen are fed into each
reaction zone along the bottom of the reactor. Catalyst
and quench liquid are introduced into each polymerization
zone through spray nozzles at the top of the reactor.
Unreacted monomer, hydrogen, and quench vapor, normally isobutane or isopentane, are removed through the
top of the reactor to the recycle and separation sections.
The weirs extend upward to somewhat above the middle
of the reactor so that the polymer bed fills about half the
volume of the reactor. As the polymer powder exceeds
the weir height, it falls into the adjacent reaction zone in
the direction of the polymer discharge. The polymer

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 463


discharge vessel contains a postpolymerization zone,
wherein a controllable amount of adiabatic polymerization
takes place, producing heat which, together with externally
added heat, melts the polymer to form easily transferable
liquid polymer. Following treatment with water and
suitable additives, the melt polymer is finally converted
to pellets.
Each polymerization zone is equipped with flat paddles
extending transversely from the shaft and making close
clearance with the inside wall of reactor to ensure adequate
bed mixing. The paddles are so constructed to minimize
any foward or backward movement of the bed. The
agitation speed is about 5-30 rpm to provide the desired
heat and mass transfer, to avoid polymer particle being
thrown up into the space above the bed, and to provide
a slow and regular turnover of the entire polymer bed
contained in the reactor. The distinguishing feature of
the AMOCO process, compared with previous processes,
is that it can be operated like a set of CSTRs in series. The
residence time distribution (RTD) of this process almost
approaches that of plug flow. Tracer studies on the
commercial reactor indicate that the RTD in the AMOCO
reactor is roughly equivalent to that of three to five wellstirred tanks in series. This unique feature of the
horizontal reactor permits production of ethylene-propylene copolymers with high impact strength and good
stiffness (Brockmeier, 1991). AMOCO and Chisso have
been collaborating for commercialization of ethylenepropylene gas phase polymerization processes (Krieger,
1992).
Soon after the Union Carbide process was introduced,
Naphtachimie developed a gas phase polymerization
process using a fluidized bed reactor (Dormenval et al.,
1975; Havas and Magin, 1976). A commercial-scale
fluidized bed plant with an initial capacity of 25 000 t/y
was commissioned in 1975 (Delgrange, 1987). The gas
phase polymerization technology developed by Naphtachimie was later transferred to BP Chemicals(Delgrange,
1987; Brockmeier, 1987). After the amalgamation of
Napthachimie into BP Chemicals, high priority was given
to further development of the novel gas phase polymerization process, as can be confirmed by recent patents
assigned to BP Chemicals (Dumain and Raufast, 1985;
Collomb-Ceccarini and Crouzet, 1986; Bailly and Speakman, 1986;Raufast, 1987;Bailly and Collomb, 1988; Chinh
and Dumain, 1990;Speakman, 1991;Matens and Moterol,
1992; Havas and Lalanne-Magne, 1992). The BP Chemicals gas phase process now has a capacity of 170 kt/y and
can produce both LLDPE and HDPE with a density range
from 0.91 to 0.96 (Delgrange, 1987; Redman, 1991; Speakman, 1991). The major distinction of the BP Chemicals
process from the UNIPOL process is that prepolymer
particles rather than catalyst particles are fed into the
fluidized bed reactor. The BP Chemicalsgas phase process
is a continuous two-stage polymerization process, as shown
in Figure 7 (Chinh and Dumain, 1990; Redman, 1991;
Matens and Moterol, 1992; Havas and Lalanne-Magne,
1992). This process uses a combination of a stirred tank
reactor and a fluidized bed reactor in series.
During conventional gas phase reactor operation, the
catalyst and the cocatalyst may be brought into contact
either prior to their introduction into the fluidized bed,
or in the interior of the reactor. Whichever method is
employed, the polymerization reaction always starts up
very abruptly and attains a maximum rate soon after the
catalyst system is introduced into the fluidized bed. I t is
therefore in the initial phase of polymerization that the
risks of hot spots forming and grains bursting into fine

MONOMER

RECYCLE

Q
9

MONOMER AND GAS FEED


P

-4'

COMPRESSOR

Figure 7. BP Chemicalstwo-stage gas phase ethylenepolymerization


process (Chinh and Dumain, 1990; Redman, 1991; Maters and
Moterol, 1992;Havas and Lalanne-Magne, 1992).

particles are greatest. Hot spots may lead to formation


of agglomerates and to settling of polymer inside the
fluidized bed (Collomb-Ceccarini and Crouzet, 1986).
Furthermore, during polymerization small variations in
the feed rates of catalyst, monomer, and comonomer or in
the withdrawal rate of polymer will also cause an unexpected increase in the quantity of heat evolved by the
polymerization. If the heat cannot be removed efficiently,
these small variations can cause hot spots in the reaction
bed and formation of agglomerates by melting polymer.
Such variations can therefore make it difficult to obtain
a polymer of consistent quality, in particular, of constant
molecular weight and particle size (Chinh and Dumain,
1990; Matens and Moterol, 1992; Havas and LalanneMagne, 1992). These problems can be eliminated by
adopting a prepolymerization stage, as shown in Figure 7.
Prepolymerization gives advantages in polymer particle
size control and control of the catalyst activity in the
fluidized bed reactor. Prepolymerization can be carried
out in a liquid hydrocarbon medium or in a gas phase
stirred reactor at temperatures from 40 to 115"C. Catalyst
is introduced into the prereactor in the form of dry powder
or in suspension in a liquid hydrocarbon. Prepolymerization is carried out to a conversion wherein the prepolymer contains 0.002-10 millimol of transition metal/g
of polymer. The diameter of the prepolymer is in the
range from 200 to 250 pm (Bailly and Speakman, 1986;
Chinh and Dumain, 1990; Martens and Moterol, 1992;
Havas and Lalanne-Magne, 1992). Prepolymer is fed into
the fluidized bed reactor through a metering feed device.
The fluidized bed reactor operates at a superficial velocity
of approximately 0.5 m/s, 2-8 times the minimum fluidization velocity (Dumain and Raufast, 1985). To avoid
induction time a t startup, the prepolymer is treated with
triethylaluminum for polymerization with chromium oxide
as catalyst (Durand and Morterol, 1986a,b; Bailly and
Speakman, 1986). To create more porosity, the prepolymer
is also treated with n-hexane to remove low molecular
weight polymer (wax) (Bailly and Speakman, 1986).
Monomer, comonomer, hydrogen, and inert gas are fed
through the bottom of the fluidized bed. Some comonomer
and inert volatile hydrocarbon, such as isopentane, are
introduced into the inlet line of the heat exchanger to
avoid fine particles depositing on heat exchanger surfaces
and compressor blades. The ratio of comonomer to
monomer partial pressure is kept constant (normally 0.10.2)) in the reactor (Chinh and Dumain, 1990).
To avoid pressure fluctuations during polymer discharge,
BP Chemicals developed a continuous discharge device

4M Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

Table 2. Ethylene Gas Phaw Polymerization Prowsws and Reactor Operating Conditions
Union Carbide
BASF
AMOCO
nvletortype
fluidized bed
stirred bsd
horizontal stirred bsd
catdlyst

caialyat size,pm
press., atm
temp. O C

comonomer
MWcontrol
MWD. Mw/Mm

density, g/cmJ

palymer particle size. pm


reference

supported Ti. V and


CrCOa catalyat
30-250
20-30
75-110

supported CrCOS

~ ~ p p o r t Ti
s d and
crcos catalyst

catalyat
-35

100-110
1-butene

1-buteneor 1-hexene
Ha
Ha
4-30
narrow tu broad
0.91-0.87
500-1300
Rasmuassn (1972):
Trieaehmann et al. (1977):
N i d e t t i et al. (1988)
Choi and Ray (1985a)

consisting of continuously rotating plug valves shown as


A and B in Figure 7. These valves are connected with
each other in such a way that the two valves do not open
at the same time, rotating a t a speed of about 0.5-1.0 rpm.
The volume of the vessel between the valves is between
0.2% and 1% ofthevolumeofthefluidizedsolidcontained
in the reactor (Raufast, 1987).
One of the disadvantages of gas phase polymerization
is that the reactor operating temperature must be lower
than the melting point ofthe polymer produced. Recently,
Bailly and Collomb (1988;1990)found that agglomeration
ofthe polymer particles commencesat higher temperatures
when pulverulent inorganic particles are employed. Thus,
higher polymerization temperaturescan be used to produce
LLDPE. The pulverulent inorganic particles, suchas silica
oralumina,arechemicallyinerttocatalystsandmonomers.
The inert particles have porosity 0.2-2.5 mL/g, surface
area 20-90 mz, and diameter 0.5-20fim. The particle size
is 5(t500 times smaller than the mean diameter of
prepolymer particles. According to the patent (Bailly and
Collomb, 1988,LLDPE with density 0.860-0.910 g/cm3
can be produced with 0.005-0.2 wt % inert particles based
on polymer in the reaction zone. The inert particles are
mixed with prepolymer particles and then fed into the
reactor. Rhee et al. (1991)also used inert particles, e.g.,
carbon black, silica, and clays, in the production of
ethylene-propylene rubbers and found that the reactor
can be operated a t temperatures higher than softening
points of copolymers when amounts of about 0.3-50% by
weight of inert material are employed. These inventions
suggest that the soft polymer particles can he stabilized
using inert particles. However, the morphology development of polymer polymerized at or near melting point
has not been reported.
Typical reactor operating conditions and production
capabilities of the gas phase processes discussed above
are summarized in Table 2.
Compared with slurry and solution processes, gas phase
processes have many distinct advantages. Slurry-based
processesproducePEcoveringa broad rangeof melt index;
however, the range of densities attainable is limited. As
density decreases, resin solubility in the dilute increases.
At a density of about 0.930 g/cm3 sufficient dissolution
occurs to foul the reactor. Hence, slurry processes are not
suitable for LLDPE production. On the other hand,
solution processes can produce PE over a broad range of
densities, but only a limited range of molecular weights.
As molecular weight increases, solution viscosity increases
also. At some point, the increased viscosity limits reactor
operability and productivity. Thus, solution processes are
not suitable for making high molecular weight PE.
Production capabilities of slurry and solution processes
are shown schematically in Figure 8, after James (1986).

BP Chemic&
stirred reactor and
f l u i d d bed
supported Ti and

crco2 catalyat

-60 for prepolymn

20-40

15-25

70-110

70-115

5-17
0.95-0.98

620

propylene or 1-butene
H,and temperature

I-butene or I-hexene
Ha
0.91-0.96
300-1200

Shepard et al. (1976);


Speakman (1991):Bailly
Jezl and Petern (1978a,b)
and SpsaLman (1986)

0.970

0.980

0.950

a
-In 0 . 0
5 0.93
0
0.920
0.910
0.01

0.10

1.00

10.0

100

MELT INDEX, 9/10 min


Figure 8. Production capabilities of aolution and slurry polymerization processes.

0910

"

"1

"

A
I"

U"

I""

MELT INDEX, g/10 min


FiKure 9. Production capability of gas phase polymerization
processes (James. 1986; Speakman, 1991).

Gas phase processes,which do not involve any liquid phase


in the reaction zone, are not constrained hy solubility and
viscosity. Hence, a gas phase process can produce a
complete range of PE with densities of 0.91-0.97 g/cm$
and melt indexes of <0.01 g/10min up to 200 g/10 min,
asshowninFigure9(James, 1986,Speakman, 1991). Other
major advantages of gas phase processes over liquid phase
processes include production of high comonomer content
polyethylenes, such as high impact strength ethylene
propylene copolymers,ethylene-propylenediene rubbers.
Comonomer content is limited in the liquid phase processes
because of problems with viscosity effects from dissolved
copolymer and comonomer (Brockmeier, 1991;Rhee et
al.,1991;Burdett. 1992). Furthermore,gasphaseprocesaes
can reduce construction costs up to 30% and operating
costs up to 35% over conventional liquid phase processes
(Burdett, 1992; Krieger, 1992).

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 455


Table 3. Comparison of High-pressure Free-Radical
Process and Low-Pressure Gas Phase Process

high-pressure process
tubular or autoclave reactor

high operating pressure,


3400 atm
high reaction temperature,
-300 OC
high energy requirement
high capital cost
limited to low-density PE
no comonomer required
(homopolymer) for
LDPE production
catalyst is less sensitive to
impurities in the feedstock
low raw material cost
can produce ethylene and
vinyl acetate copolymer
polymer with long-chain
branching

low-pressure process
fluidized bed or stirred
bed reactor
low operating pressure,
-20 atm
low reaction temperature,
80-110 OC
low energy requirement
low capital cost
can produce both high- and
low-density PE
up to l0-15% comonomer
required (copolymer)
catalyst is very sensitive
to impurities
high raw material cost
currently a limited range
of comonomers
mostly linear polymer
chains with short-chain
branching

Compared with slurry and solution processes, however,


gas phase processes do have the following disadvantages:
1. Reactor operating temperature is limited to the resin
softening point. Thus the productivity of catalyst is also
limited.
2. The poor heat-transfer efficiency of the gas phase is
a disadvantage. Hence, additional inert heat transfer agent
is required to maintain stable reactor operating conditions
at high production rates.
3. There is a possibility of sintering and agglomeration
of the polymer particles due to formation of local hot spots
when a high-activity catalyst is used.
4. As polymerization progresses, fine particles will
deposit on heat-transfer surfaces, compressor blades, and
sloped walls of the reactor.
5. Sensitivity to variation in operating conditions, such
as flow rates of catalyst and monomer feed and polymer
discharge, is a disadvantage.
Gas phase processes offer many advantages over conventional high-pressure processes. The conventional highpressure free-radical process can produce only low-density
polyethylene due to the nature of its polymerization
mechanisms. Table 3 compares some features of the highpressure LDPE process and the low-pressure LLDPE gas
phase process (Imhausen et al., 1981; Karol, 1983).
Significant progress has been made in overcoming the
disadgantages of gas phase polymerization and in improving heat removal capacity, as discussed above. Overall,
gas phase polymerization processeshave had a great impact
on polyethylene production technologydevelopment.
2.2. Experimental Methods. Despite the commercial
success of gas phase ethylene polymerization technology,
the public literature contains no accounts of fundamental
scale-up studies of gas phase processes. There is a need,
therefore, for a comprehensive understanding of detailed
polymerization behavior in gas phase polymerization. A
challenge for academic researchers studying gas phase
polymerization of ethylene is how to scale down commercial
processes for experimental laboratory studies.
Kinetic studies of ethylene gas phase polymerization
were first carried out by Edgecombe (1963) and Lipman
and Norrish (1963) using conventional Ziegler-Natta
catalysts. Wisseroth (1969) reported on gas phase polymerization of ethylene using Phillips catalysts. With
the development of commercial gas phase polymerization
processes, more university-based researchers have at-

rt TO WATER BATH

/1

FROM WATER BATH

3
Figure 10. Laboratory reactors for gas phase ethylene polymerization. (1) Choi et al. (1983, 1985b); (2) Lynch et al. (1991); (3)
Dusseault (1991).

tempted to extend experimental studies of gas phase


polymerization in recent years. Choi et al. (1983,1985b)
designed a 1-L gas phase stirred bed reactor equipped
with an anchor-type agitator. Mabilon and Spitz (1985)
and Spitz et al. (1988)used a 675-cm3stainless steel reactor
stirred by vertical oscillationsfor gas phase polymerization.
However, the detailed reactor configuration used in this
investigation was not described. More recently, Jejelowo
et al. (1990) and Lynch et al. (1991) constructed a 1-L
stainless steel reactor with a specially designed paddle
agitator. The reactor temperature is controlled using a
26-L oil bath. To improve reactor mixing and heat transfer,
Dusseault (1991) designed an 800-cm3 horizontal stirred
reactor with both internal cooling coil and external cooling/
heating jacket. The agitator is designed in such a way so
that the entire polymer bed can be turned over. Weist et
al. (1989) used a 20-mm-diameter fluidized bed reactor to
study polymer morphology development at the initial
polymerization stage. However, the reactor configuration
and performance were not described. Figure 10 shows
typical reactors used for ethylene gas phase polymerization
studies reported in the literature.
Because the reactivity of the catalyst is very high, only
a very small amount of catalyst, normally in the order of
milligrams, is required for laboratory-scale polymerization
using a 0.5-1.0-L reactor. Therefore, inert particles are
usually added into the reactor to assist the initial mixing
of the catalyst. Glass beads with diameters of 500-5000
pm are often used (Keii, 1972; Doi et al., 1982; Choi et al.,
1983,1985b; Dusseault, 1991). Mabilon and Spitz (1985)
and Spitz et al. (1988) used salt (KC1) or polyethylene
powder to assist initial mixing. Lynch and Wanke (1991)
used 450-pm Teflon powder. Glass beads may further
grind the catalyst during the initial stage of polymerization.
Salt is sensitive to moisture; hence careful drying procedure
is required, but the separation of polymer from salt is very
convenient. Although polyethylene powder is effective
for initial mixing, it is not good for polymer property
analysis because newly produced polymer can not be
separated from the initial polymer powder. However, it
may be useful for polymerization rate measurements
without requiring polymer separation.
The laboratory-scale reactors shown in Figure 10 can
provide satisfactory temperature control within a restricted
operating temperature range. With high-activity catalysts,

456

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

Table 4. Main Requirements of Industrial Catalysts for


Ethylene Gas Phase Polymerization

polymerization process
requirements
high productivity
proper kinetic behavior
catalyst morphology
control of polymer morphology
good comonomer incorporation
easy feed to reactor
low cost and reproducible
catalyst preparation

polymer property control


molecular weight
molecular weight distribution
density
chain-branching distribution
particle size and bulk density
polymer chain unsaturation

the initial polymerization rate is very high, and so the


initial heat is difficult to remove with an oil or water bath
cooling system. Therefore, reactor temperature control
systems shown in Figure 10 are still not adequate for a
broad range of reactor operating conditions. To carry out
gas phase ethylene polymerization under commercial
reactor operating conditions, the reactor must be able to
provide fast heat removal to avoid hot spots in the
polymerization zone and adequate mixing conditions to
disperse a small amount of catalyst powder.
3. Physicochemical Phenomena
In light of the preceding observations, it is evident that
gas phase polymerization involves a complicated physicochemical transition from gaseous monomer to solid
polymer, although the synthesis process itself is relatively
simple compared with liquid phase processes. In gas phase
polymerization, catalyst preparation and polymer particle
morphology development play important roles in reactor
operation. Hence, it is crucial to understand the relevant
microscopic processes of polymerization and their relationships.
3.1. Catalysts for Gas-Phase Polymerization. Catalysis is the most active area of research in ethylene
polymerization with worldwide participation in both
industrial and academic laboratories. In recent years,
hundreds of publications on polyethylene synthesis and
manufacture have been recorded annually in Chemical
Abstracts. Most are related to catalysis studies. The
extent and variety of research work dealing with the same
problem reflects not only the great interest and extensive
commercial applications in this area, but also the complexity of catalysts for ethylene polymerizations. Research
progress concerning catalysts for ethylene polymerization
has been a focus of several recent international conferences
(Seymour and Cheng, 1987; Kaminsky and Sinn, 1988;
Quirk et al., 1988; and Keii and Soga, 1990; Vandenberg
and Salamone, 1992),and excellent reviews have appeared
in the literature (Zakaharov and Yermakov, 1979; Karol,
1984; Hsieh, 1984; McDaniel, 1985; Nowlin, 1985; Hsieh
et al., 1987; Barbe et al., 1987; Tait, 1989; Kryzhanovskii
and Pranchev, 1990; Dusseault and Hsu, 1993). Catalyst
development plays a vital role in the polymerization of
ethylene, particularly in successful gas phase polymerization processes (Karol, 1983). For gas phase polymerization, although high productivity of the catalyst is a
necessary condition, other features, such as copolymerization kinetic behavior and catalyst particle size and shape
also play important roles in the production of high-quality
LLDPE and HDPE. The main requirements of industrial
catalysts are summarized in Table 4 (Karol et al., 1987;
Dall'Occo et al., 1988; Burdett, 1988).
Two major catalyst families have been commonly used
in commercial ethylene production: titanium1 vanadium
based catalysts (Ziegler-Natta catalysts) and chromium
oxide based catalysts (Phillips catalysts). In the past four

Table 5. Properties of Silica Support Used for Gas Phase


Polymerization (Joraensen et al.. 1982: Nicoletti et al.. 1988)

silica
particle size, pm
surface area, m2/g
pore size, A
concn of catalyst metal,
mmol/g of support

range
10-250
at least 3
at least 80
0.054.5

preferable range
30-100
50
100
0.2-0.3

decades, research has been focused mainly on titanium


and chromium oxide based catalysts (Karol et al., 1988).
A typical industrial catalyst consists of active metal,
modifiers, and inert support. Porous silica and magnesium
halide are common supports. Table 5 shows typical
properties of silica support used for commercial gas phase
polymerization (Jorgensen et al., 1982; Nicoletti et al.,
1988). Internal pore size distribution is in the range of
80-1000 A, and with the distribution peak in the range of
100-200 A, depending on preparation (Weist et al., 1987).
A comparison of the performance of titanium-, vanadium-,
and chromium-based catalysts is summarized in Table 6
(McDaniel, 1985;Karol, 1988). The significant difference
between Ti- and Cr-based catalysts is that the molecular
weight distribution (MWD) of polymers made using Ti
catalysts in narrower than that using Cr-based catalysts.
High activity or productivity of catalyst means that the
catalyst leaves no more than 10 ppm catalyst metal and
30 ppm chloride in the polymer formed. Titanium levels
higher than this level lead to polymer color problems, and
higher chloride contents result in excessive corrosion
problems in processing equipment (Hsieh, 1980, 1984).
According to recent patents, the metal content in polymer
is very low for the catalyst used in gas phase polymerizations, e.g., <5 ppm of Ti (Jorgensen et al., 1982), <0.5
ppm of Ti (Collomb-Ceccarini and Crouzet, 1986), and
about 2 ppm of Cr (Speakman, 1991).
Several methods have been developed for Ti-based
catalyst preparation (Jorgensen et al., 1982; Karol, 1984;
Collomb-Ceccarini et al. 1986). Three steps are normally
involved: (a) formation of a precursor containing Ti, (b)
incorporation of the precursor onto the support, and ( c )
activation of the catalyst. The different methods can be
distinguished by the mode of formation of a precursor
containing titanium, as summarized in Table 7 together
with some examples (Karol, 1984). Method 1in Table 7
is often used for chromium oxide based catalyst preparation. Methods 2 and 3 are commonly used for Ti-based
industrial catalyst preparation (Jorgensen et al., 1982;
Collom-Ceccarini et al., 1986; Nicoletti et al., 1988). The
catalyst precursor prepared using method 2 in Table 7 is
then impregnated onto a porous support such as silica, as
shown in Table 5. Impregnation of the support with the
precursor is accomplished by dissolving the precursor in
the electron donbr compound and admixing the support
with dissolved precursors. The solvent is then removed
by drying at a temperature of up to 85 "C (Jorgenson et
al., 1982; Nicoletti et al., 1988). The catalyst precursor
prepared by ball-milling (method 3 in Table 7) has a
relatively small diameter, generally less than 50 pm. It is
too small to be used directly for gas phase polymerization
in a fluidized bed reactor. For this reason, the precursor
and cocatalyst are converted to prepolymer particles in a
prepolymerization stage as mentioned for BP Chemicals
processes. The concentration of titanium in the prepolymer is controlled in the range of 4-30 ppm. Prepolymer
particle size is about 70-250 pm before it is fed into the
fluidized bed reactor (Collomb-Ceccarini et al., 1986).
Hence, one of the functions of prepolymerization is to
form prepolymer in place of silica as a catalyst support for
gas phase polymerization.

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 467


Table 6. Performance of Titanium, Vanadium, and Chromium Oxide Based Catalysts
(McDaniel,
1986: Karol, 1988)
.

Mg/Ti/EDa
activator
productivity
Hz response
MWD
a-olefin incorporation
activity decay rate
polymer chain unsaturation
long-chain branching

R3A1
high
moderate
narrow, 3-6
moderate
moderate
low
no

VC13(THF)3
R&l + halocarbon promoter
high
high
intermediate-broad
high
low-moderate
very low
no

CrOs
thermal, 300-1000 OC
high
low
intermediate-broad, 6-30
high
very low
one C=C per polymer molecule
low concn of long-chain branching

ED is an electron donor.
Table 8. Approximate Dimensions of MgCll Supported
Table 7. Operational Methods to Produce High-Activity
Catalyst (Chien et al.. 1983: Chien. 1987: 1988)
Catalysts (Karol, 1984)
1.chemical anchoring to surface of support
pore
surface
(a) TiCWMg(0H)Cl
volume,
area,
(b) CrOa/SiOz
composition
diameter,A
cm3/g
m2/g
(c)
. . (CKHdoCr/SiO,
.
MgC12 (HCl treated)
-4000
0.41
2. formation of bimetallic complexes
700-1500
0.41
2.0-7.0
(a) MgClz + 2TiC4 + 8Poc13 [TiCl~~lZ[Mg(POC13)sl~+~2POC13MgCldEB
MgClz/EB/PC/AlEts
300-500
1.2
50-70
(b)2MgC1~+ Tic4 + 7THF [TiCls(THF)l-[Mg2Cl~(THF)61+
MgC12/EB/PC/AlEt3/TiC&
25-170
1.3
100-150
(c) MgC12 + Tic4 + 4CH3COzCzHs TiMgCl&H&OzCzHs)4
3. insertion into defects of support
(a) MgClz + Tic14 + ethyl p-toluate ball-milling
MgClz with an electron donor (internal Lewis base), such
(b) MgClz + TiQedioxane
ball-milling
as ethyl benzoate, can produce a disordered crystal
4. formation of high surface area sponge
structure and reduce crystalline size to 20-40 A. The
~~

I,_

(a) Tic4 + EtzAlCl

-- - - +- +

bwmyl ether

~~

surface area of the support increases to a maximum value


of ca. 6.5 m2/g. The surface area of the support further
treated solid
[TiCl~~(EtA1C12)o,03(ether)o,oJ
increases through each step of the modification. Table 8
(b) Mg(0Et)z + Tic4 [MgClz.Mg(OEt)z.Mg(TiC&)l
shows
approximate dimensions of a MgClz supported
(c) Mg(0Et)Z + Ti(O-nC,H& EtAlC12 trimetallic sponge
catalyst obtained by ball-milling (Chien et al., 1983;Chien,
5. formation of solid solutions by cocrystallization (coprecipitation)
(a) EtMgCl + Tic4 TiCls-MgClz organic fragments
1987, 1988). Although the composition of commercial
catalysts may not be the same as that shown in Table 8,
the main features of MgC12 supported catalysts are the
The patent literature usually gives only the state of the
same. From Table 8 it can be seen that MgClz supported
art for preparation of catalysts. The reasons for the choices
of different compounds are not explained. To better
catalysts have large surface areas and internal pore
understand the effect of each component used on the
volumes. Both surface area and pore volume change with
modifier composition. Incorporation of T i c 4 with such
properties of a catalyst, Chien and his co-workers have
fully characterized a typical MgC12 supported catalyst:
a support results in dispersion or isolation of the titanium
centers so that the population of active centers which are
MgC12/EB/PC/A1Et3/TiC14-A1Et3/MPT,where EB, PC,
and MPT are abbreviations for ethyl benzoate, p-cresol,
accessible for polymerization increases dramatically comand methyl p-toluate, respectively (Chien, 1987, 1988,
pared with that form unsupported conventional Ti cat1992). The major findings of academic studies on MgC12
alysts. This is the main reason why supported Ti catalysts
have productivities so much higher than those of consupported catalysts have been comprehensively reviewed
by Dusseault and Hsu (1993). The main reasons that
ventional catalysts. In addition, magnesium ions can
stabilize active titanium from deactivation, enhance chainMgCl2 is the most widely used as a catalyst support can
be summarized as follows:
transfer processes (& of PE decreases when Mg/Ti ratio
increases), and lead to narrow MWD for polyethylene
1. MgClz has crystalline forms similar to those of Tic13
production (Karol, 1984; Nowlin, 1985).
(conventional Ziegler-Natta catalyst) (Chien, 1987;Barbe,
1987; Dusseault and Hsu, 1993). This distinct feature
According to Chien (19871, both internal Lewis base
and internal alcoholincrease surface area and pore volume,
suggests that MgClz should have the ability to mimic the
but they do not compete for the same surface sites of MgC12
structure of active Tic13 and the ability to efficiently
crystal because the content of EB on the support is the
incorporate TiC14.
same after addition of internal alcohol. Therefore, they
2. MgCl2 possesses desirable morphology as a support.
may adsorb on different planes of MgClz crystal. AluMercury porosimetry showed that MgClz has a large
minum alkyls such as triethylaluminum (AlEt3) are usually
number of pores with radii > 200 nm (Chien et al., 1983).
used as cocatalysts to activate the catalyst. The activating
The structure of MgClz is sufficiently resistant to fracture
action of AlEt3 is believed to reduce Ti4+ to Ti3+ and
by operational manipulation, but weak enough so that its
alkylation. In addition to the activation requirement,
internal structure will be broken down during polymerAlEt3 also acts as a scavenger of impurities of the
ization.
polymerization system. The optimum ratio of Al/Ti is in
3. MgC12 has a lower electronegativity compared with
those of other metal halides. A metal halide with lower
the range of 10-30 (Dusseault and Hsu, 1993). AlEt3 can
also reduce MgClz support particle size and increase surface
electronegativity than Tic13 will increase productivity of
area 2-3-fold if AlEt3 is used as an internal modifier during
ethylene polymerization (Soga et al., 1987). MgClz also
support preparation (Chien, 19871, as shown in Table 8.
enhances chain-transfer reactions because M ndecreases
Compared with Ti-based catalyst preparation, chrowith an increase in Mg/Ti ratio (Karol, 1984).
mium-based catalyst preparation seems to be simpler.
4. MgC12 is inert to the chemicals used for polymeriChromium-based catalyst (Phillips catalyst) is usually
zation. It can be left in the final polymer product without
prepared by impregnating a chromium compound onto a
deashing at the levels used in commercial production.
wide-pore silica and then calcining an oxygen environment
Because of the above features of MgClz, Tic14 can be
to activate the catalyst. This leaves the chromium in the
incorporated with MgClz in different ways. Co-milling of
,9-TiC1,.xEtAlC12
Tic4

458 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

hexavalent state, monodispersed on the silica surface.


Chromium trioxide (CrOa) has been impregnated most
commonly (McDaniel, 1985). Cr03chemically anchors to
the silica surface to form a new surface composition of
silica (Karol, 1984; Hsieh, 1984; McDaniel, 1985; Nowlin,
1985). However,whether the chromium atoms are isolated
as chromate, or exist in pairs as dichromates, is still debated
in chromium catalyst studies. It is believed that chromium
trioxide probably binds to the silica as chromate initially,
and during calcining some chromate may convert to
dichromate (McDaniel, 1985).
At least six methods have been described in preparation
of Cr03 catalysts for ethylene gas phase polymerization
(Baillyand Speakman, 1986). The main distinction among
them is due to incorporation of modifiers. After CrOs is
impregnated with porous silica, the catalyst is further
impregnated with a titanium compound such as titanium
tetraisopropoxide. Then, the catalyst is activated by a
heat treatment at 350-1000 "C under air for 30 min-5 h.
The final catalyst contains 0.05-10% chromium and 0.5020% titanium (Bailly and Speakman, 1986; Speakman,
1991).
Catalyst prepared following the above procedure is not
suitable for direct use in gas phase polymerization because
of its long induction time. Cr(VI)/silica catalyst develops
polymerizationactivity gradually when exposed to ethylene
monomer. It is believed that Cr(V1)is reduced by ethylene
to the Cr(I1) active site. Below 100 "C the induction time
becomes longer until at 60 "C there is almost no activity.
However, at 150 "C, the catalyst exhibits an immediate
and constant activity in solution polymerization (McDaniel,
1985). For gas phase polymerization, the normal operating
temperature is between 80 and 100 "C. The Cr-based
catalyst has a significant induction period which causes
difficulty not only in starting polymerization, but also in
maintaining constant polymerization conditions in the
reaction zone. When the polymerization is started, the
catalyst activity rapidly accelerates, which also makes
polymerization conditions difficult to control. Therefore,
the catalyst has to be prereduced to eliminate the induction
time. CO and alkyls of aluminum, boron, and zinc can be
used as reducing agents to eliminate induction time
(McDaniel, 1985). However,for gas phase polymerization,
catalyst with fully activated chromium could lead to the
formation of hot spots and cause agglomeration due to
poor heat transfer in the reaction zone. According to Bailly
and Speakman (1986) and Speakman (1991),the best way
to overcome these problems is to convert the catalyst to
prepolymer particles through a prepolymerization stage
as described above.
It should be pointed out that the activation temperature
is a very important parameter in controlling catalyst
activity and polymer molecular weight for ethylene
polymerization using Cr-based catalyst. Figure 11shows
the effects of activation temperature on the reactivity and
molecular weight of polymer (McDaniel, 1985,1988).Here,
activity is defined as the inverse of the time needed to
produce 5000 g of polymer/g of catalyst. The activity
increases with an increase in activation temperature up to
maximum at around 925 "C, and then declines as sintering
destroys the surface area and porosity of the silica base.
The weight-average molecular weight decreases with an
increase in calcining temperature up to 925 "C; then the
trend reverses as the silica begins to sinter. The melt
index shows the same type of relationship. The relationships shown in Figure 11suggestthat the chain termination
reaction is very sensitive to the activation temperature
for Cr-based catalyst systems.

_I_I

lo

200

150

600

600
700
800 900 lo00
ACTIVATION TEMPERATURE ('C)

Figure 11. Effecta of activation temperature on reactivity and


polymer properties (McDaniel, 1985, 1988). Reprinted with permission from McDaniel (1985). Copyright 1985 Academic Press.

The silica supported catalysts described above have been


successfully used for PE production. Deashing processes
have been eliminated for many years. However, silica does
remain in the final polymer product as an impurity. This
impurity causes the polyethylene to have a poor film
appearance. Hence, it is desirable to eliminate silica by
use of polymer particles as a support. Research in this
area has been reported in recent patent publications (Cana
et al., 1992). Polypropylene and polyethylene or their
copolymers have been used as supports. The catalyst
precursor is bonded on the polymer particles. However,
detailed information about how the catalyst is bonded on
the polymer particles is not given. The results using
polymer supported catalyst are sufficiently promising that
polymer supported catalysts may replace nonpolymer
supported catalysts for the manufacture of polyethylene
film product in the near future.
The term "supported catalyst" is used in a very wide
sense and includes not only systems in which the transition
metal compound is linked to the substrate by means of a
chemical bond, but also systems in which the transition
metal atom may occupy a position in a lattice structure,
or where complexation, adsorption, or even occlusion may
take place (Tait, 1989). In commercialapplications, MgC12
supported catalyst is further supported on silica or polymer
through prepolymerization. All of these supports play a
role in polymerization and in polymer morphology development.
Besides the heterogeneous catalysts discussed above,
another catalyst family that may soon be commercialized
in gas phase olefin polymerizations is metallocene-based
catalysts. The metallocene-based catalysts are also called
homogeneous Ziegler-Natta catalysts in the literature. The
first homogeneous Ziegler-Natta catalyst, CpzTiClz/AlEt~C1(Cp = cyclopentadienyl) was independently discovered
by Natta and Pino (1957) and by Breslow and Newburg
(1957). The kinetic behavior of this catalyst was studied
by Breslow and Newburg (19591, Chien (19591, and others
(Reichert, 1983). The productivity of this catalyst is very
low, less than 300 g of PE/(g of Tiehaatm) (Chien, 1959;
Chien et al., 1990). However, if cocatalysts (AlEt, or
AlEtC12) are treated with water, both productivity and
molecular weight (MW) of PE increase dramatically
(Reichert and Meyer, 1973;Andresen et al., 1976;Chihlar

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 469


et al., 1978). It was believed that aluminoxane plays an
important role in the increase in productivity (Andresen
et al., 1976;Chihlar et al., 1978;Sinnand Kaminsky, 1980).
Significant breakthroughs however occurred in the early
1980swhen Kaminsky and his co-workersfound very high
productivity for ethylene polymerization using CpzZrCl2
as catalyst and methylaluminoxane (MAO) as cocatalyst
(Sinnet al., 1980; Kaminsky et al., 1983;Kaminsky, 1983;
Herwig and Kaminsky, 1983;Kaminsky and Luker, 1984).
Furthermore, Ewen (1984) and Kaminsky et al. (1985)
demonstrated that chiral rac-Et[IndH412TiClz and racEt[Ind]zZrClz/MAO (Ind = indenyl) catalysts are capable
of stereospecific polymerization of a-olefins. Since the
mid-l980s, metallocene-based catalysts have become of
academic and commercial interest.
The level of industrial research activities in metallocene
catalysis has been tremendous. More than 50 companies
worldwide have been involved in metallocene catalyst
studies. Among them, Mitsui Petrochemical, Indemitsu,
Exxon, Mitsui Toatsu, Hoechst, Asahi Chemical, and Dow
Chemicalare the leading companies in terms of the number
of patents issued. To date, over 50 metallocene catalysts
have been described based on Ti, Zr, and Hf metallocenes
(SRI, 1993). Exxon Chemical was the first company to
apply metallocene catalysis to polyolefin production on a
commercial scale. The company built a 15 000 t/y highpressure autoclave reactor unit at the Baton Rouge site.
This plant was started up on June 1991and has produced
specialty linear ethylene-based polymers called EXACT
resins: about 20 grades with a range of densities from
0.865 to 0.935 g/cm3 (Mod. Plast., 1991; SRI, 1993;
Montagna and Floyd, 1993; Rogers, 1993a). To further
increase product range, Exxon and Mitsui Petrochemical
have been collaborating for commercialization of gas phase
process technology using metallocene catalysts (SRI,1993;
Rogers, 1993a). To further increase product range, Exxon
and Mitsui Petrochemical have been collaborating for
commercialization of gas phase process technology using
metallocene catalysts (SRI, 1993; Rogers, 1993b). The
gas phase process is expected on stream in the mid-1990s
and will be operated at temperatures of 80-90 "C and will
produce 100 000 t/y of ethylene copolymers with densities
ranging from 0.86 to 0.975 g/cm3. Dow Chemical, in
competition with Exxon, has developed a new family of
polyolefins using Constrained Geometry Catalyst Technology, This technology was commercialized under the
trade name INSITE in mid-1993 (Rotman and Wood, 1993;
Schwank, 1993; Lindsay, 1993; Mod. Plast., 1993; Story
and Knight, 1993). The distinct characteristic of INSITE
polyethylenes is improvement in physical properties
without sacrificingprocessability as a result of a significant
amount of long-chain branching formed in polymer chains
(SRI, 1993; Stevens, 1993). Thus, for the first time,
polyethylenes made through coordination polymerization
have the melt rheological advantages currently obtainable
only in free-radical-polymerized high-pressure LDPE.
INSITE resins are now produced in solution processes
(Rotman and Wood, 1993). However, Dow claims to have
successfully carried out polymerizations using its metallocene catalysts in other processes as well, including highpressure, gas phase, and slurry processes (SRI, 1993).Mobil
Chemical has also carried out ethylene and l-hexene
copolymerization using metallocene catalysts in a pilotscale fluidized bed reactor. Commercialization of a gas
phase process using a metallocene catalyst is on the way
(Furtek, 1993). In addition to the companies above, there
are others in Japan, Germany, and Belgium that appear

to be well on the way to commercialization of metallocene


technologiesfor polyolefins (SRI, 1993;Kuber, 1993;Payn,
1993).
While polyolefin companies worldwidejostle in the race
to commercialize metallocene-based technologies, the
fundamental studies of metallocene catalysts carried out
in companies are highly confidential. In academic institutes the kinetic studies of metallocene catalyzed olefin
polymerization were mainly carried out in solution processes. The following metallocene-based catalysts have
been studied extensively: CpZMClZ/MAO (M = Ti, Zr,
Hf) (Herwig and Kaminsky, 1983; Kaminsky et al., 1983;
Kaminsky and Luker, 1984;Kaminsky and Hahnsen, 1987;
Chien and Wang, 1988; Mallin et al., 1988; Ahlers and
Kaminsky, 1988; Tsutaui and Kashiwa, 1988; Tait, 1988;
Chien and Wang, 1989;Chien and Wang, 1990;Kaminaka
and Soga, 1992; Uozumi and Soga, 1992); [IndlzZrCld
MA0 (Ahlers and Kaminsky, 1988; Chien and He, 1991ad); chiral metallocenes, e.g., Et[IndIzZrClz (ethylene
bis(indeny1)zirconium dichloride), Et[IndIzZrMez, and
Et[IndH4]zZrCl2 (Ewen, 1984;Kaminsky et al., 1985,1988;
Kaminsky, 1986;Ahlers and Kaminsky, 1988;Drogemuller
et al., 1988; Ewen et al., 1988; Tait, 1988; Antberg et al.,
1990; Chien and He, 1991a; Chien and Sugimoto, 1991;
Kioka et al., 1990; 1992; Heiland and Kaminsky, 1992;
Soga and Kaminaka, 1992; Uozumi and Soga, 1992; Soga
et al., 1993; Chien and Nozaki, 1993; Chien and Gong,
1993); and other metallocene derivatives (Kaminsky et
al., 1986,1991; Ewen et al., 1988; Tait et al., 1992; Chien
et al., 1992). Compared with heterogeneous Ziegler-Natta
catalysts, metallocene-basedcatalysts have distinct kinetic
characteristics for olefin polymerizations, which are summarized as follows.
Superhigh productivity and good polymerization
rate profile. Polymerization rate reaches a maximum
within 5 min, decays slightly, and then remains almost
constant. For the Cp2MRz (M= Ti, Zr, Hf; R = C1, CH3)
catalyst family, zirconium catalysts are more active than
titanium or hafnium systems, at temperatures over 50 OC.
The cocatalyst methylaluminoxane is better suited than
ethylaluminoxane or isobutylaluminoxane (Sinn and Kaminsky, 1980; Kaminsky et al., 1983, 1986; Herwig and
Kaminsky, 1983;Kaminsky and Hahnsen, 1987;Mallin et
al., 1988; Tsutsui and Kashiwa, 1988; Chien and Razavi,
1988; Chien and Wang, 1989,1990; Chien and Sugimoto,
1991; Tait et al., 1992). For Et[Indl2MC12 (M = Zr,Hf),
Hf catalysts show higher activities (Heiland and Kaminsky,
1992). Productivity of metallocene catalysts can be up to
3 X lo7 g of PE/(g of Zr*h) (Kaminsky et al., 1986).
Very high initiation efficiency. Active center concentration is 75-100% of Zr for ethylene polymerization
using zirconocene/MAO, increases monotonically with
temperature, and is sensitive to Al/Zr ratio (Chien and
Razavi, 1988; Tait, 1988; Chien and Wang, 1989; Chien
and He, 1991b; Tait et al., 1992).
Narrow molecular weight distribution. In general,
polydispersity of PE or its copolymers produced using
metallocene catalysts is between 2 and 5; in most cases,
polydispersity is about 2 (Tsutsui and Kashiwa, 1988;
Chien and Wang, 1990; Chien and He, 1991a;Uozumi and
Soga, 1992). Polydispersity can be controlled by mixing
metallocene catalysts or by mixing metallocene with
heterogeneous Ziegler-Natta catalysts (Ahlers and Kaminsky, 1988; Kioka et al., 1990; Heiland and Kaminsky,
1992; Fries and Bowles, 1993).
Molecular weight sensitivity to hydrogen a n d
temperature. Chain transfer to hydrogen and @-hydride
elimination rate constants are 2-3 orders of magnitude
greater than the corresponding values found for MgClz

460

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

supported heterogeneous Ziegler-Natta catalysts (Kaminsky and Luker, 1984; Chien and Wang, 1990). Hydrogen also reduces polymerization rate (Kaminsky and
Luker, 1984; Kioka et al., 1992).
Uniform incorporation of comonomers. Composition of copolymer is almost the same as the monomer feed
composition, and comonomer is uniformly distributed in
copolymer chains regardless of chain length (Chien and
He, 1991a,d; Uozumi and Soga, 1992). The effect of
comonomer on ethylene polymerization rate depends on
polymerization temperature. 1-Hexene or propylene
enhances ethylene consumption rate at 30 and 50 "C
(Tsutsui and Kashiwa, 1988; Koivumaki and Seppala,
1993). However, 1-hexene reduces ethylene polymerization rate at 70 and 95 "C (Chien and Nozaki, 1993;
Koivumaki and Seppala, 1993).
Stereochemical control i n microstructures of olefin
polymers. Metallocene catalysts by manipulating the
structures of organocompoundscan polymerize olefins with
very high stereoregularity to give either isotactic or
syndiotatic polymers (Ewen, 1984;Kaminsky et al., 1985,
1986; Kaminsky, 1986; Ewen et al., 1988; Antberg et al.,
1990; Chien and Sugimoto, 1991; Antberg et al., 1991;
Uozumi and Soga, 1992).
Very high cocatalyst ratio. To obtain high polymerization rate and high active center concentration, the
Al/Zr ratio needs to be 1OL1O4orhigher (Chien and Razavi,
1988; Sinn et al., 1988; Chien and Wang, 1989). If the
AVZr ratio is small, the polymerization rate increasesslowly
and has an induction period (Chien and Sugimoto, 1991).
The reason for high cocatalyst ratio is still unclear. The
cost of cocatalyst could be more than 200-300 times the
cost of the catalyst in commercial production (SRI, 1993).
Hence, high cocatalyst ratio is one of the major barriers
to commercialization of metallocene technologies.
Metallocene catalysts must be supported on a carrier
for use in gas phase processes. Chien and He (1991~)
found
that Et[Ind12ZrClz supported on Si02 does not change
polymerization behavior. Interestingly, a smaller amount
of MA0 is required in the case of the supported catalyst
than for the homogeneous system to achieve the same
catalytic activity. Et[IndHrlzZrClz supported on A1203
or MgC12 can be activated by common trialkylaluminums;
however, the catalyst cannot be activated by common
alkylaluminums if it is supported on Si02 (Soga and
Kaminaka, 1992;Soga et al., 1993). The MgClz supported
metallocene catalysts produced a broader MWD (Soga et
al., 1993). CpzZrClz supported on A1203 or MgClz produces
atactic polypropylene, but is inactive to propylene using
Si02 as a carrier (Kaminaka and Soga, 1992). The effect
of a carrier on the performance of metallocene catalysts
is unclear at present.
For six decades, the polyolefin industry has experienced
several great revolutions. Despite the fact that there are
some technological difficulties, commercialization of metallocene-based catalyst technologies may be the greatest
revolution of all. The impact of metallocene-based
catalysts on polyolefin industry is perhaps beyond the
current imagination.
3.2. Polymer Particle Morphology Developments.
One of the most important features of ethylene polymerization using heterogeneous catalysts is replication of
polymer particles (Berger and Grieveson, 1965;Mackie et
al., 1967). This suggeststhat the morphologydevelopment
of polymer particles depends on the original morphology
of the catalysts. Polymer growth on titanium crystal
surfaces has been carefully examined using electron
microscopy (Carradine and Rase, 1971;Baker et al., 1973),

and it was found that the polymer first accumulates on


boundaries of catalyst fragments and at existing cracks.
Carradine and Rase (1971) observed that the polymer
appearing on lateral surfaces is nodular in character while
that growing on basal surfaces is fibrous. Baker et al.
(1973) observed nodular polymer growth around small
catalyst particles (-0.1 pm in diameter) during gas phase
polymerization. The shape of the polymer growth was
usually governed by that of the active catalyst particle.
Wristers (1973a,b) examined the morphology of classical
Ziegler-Natta catalyst using electron scanning microscopy,
and found that the gross catalyst particles are composed
of primary particles whose size and shape depend on the
preparation conditions of the catalyst. According to Hock
(1966) and Graff et al. (1970),the size of a primary particle
is in the range 10-1000 A in diameter. The shape and
reactivity of primary catalyst crystals are responsible for
the fiberlike structures that make up the final polymer
particles (Wristers, 1973a,b). On the basis of electron
microscopy and adsorption data, Bukatov et al. (1982)
suggested that the 10-30-pm catalyst granules consist of
three levels of particles: 200-1000-A, 1000-5000-A, and
<5-pm particles. The size of catalyst particle in the final
polymer matrix is in the range of 200 to 500 A. Mkrtchyan
et al. (1986) used microspherical mesoporous copolymers
of styrene and divinylbenzene as a catalyst support, and
found that the microspheres of support polymer also
dispersed in the polyethylene matrix.
The morphology replication nature of ethylene polymerization provides a guideline for catalyst development
(Galli et al., 1981, 1984). Polymer morphology control
can be considered in the catalyst preparation stage.
Although the reactivity and chemicalcompositionof MgClz
or silica supported catalyst have been changed compared
with those of a conventional Ziegler-Natta catalyst, the
morphological replication nature of ethylene polymerization using the supported catalyst is the same. The
morphology of a final polymer particle is governed by the
morphology of the supported catalyst. However, for
supported catalysts, the primary particle is the support
material, e.g., MgC12 crystal or silica. It has been experimentally observed that polymer grows on the edges of
MgC12 crystals treated with T i c 4 and activated by AlEt3
(Barbe et al., 1983;Galliet al., 1984). Polymer morphology
development using supported catalysts has been examined
in many laboratories. All the evidence shows that the
polymer particle size and shape replicate its parent catalyst
size and shape (Galli et al., 1981, 1984; Hosemann et al.,
1982;Munoz-Escalona et al., 1987,1988;Karol et al., 1987;
Kakugo et al., 1988;Dall'Occo et al., 1988). This suggests
that the MgClz particle and/or silica surface favors the
most isotropic distribution of Tic13 active centers. Figure
12 shows the replication relationship between silica
supported catalyst and polymer particles made with the
catalyst (Munoz-Escalona et al., 1988).
The morphology of a supported catalyst depends on the
preparation conditions and various modifiers. No detailed
morphology study on a supported catalyst has yet been
reported to date. The particle size shown in Table 8,
estimated by Chien (1987), is very small. The results in
Table 8 suggest that MgC12 supported catalysts are
composed of minute crystals with large surface area and
pore volume. As discussed in the previous section, the
catalyst particle size depends on preparation conditions.
Commercial catalysts have a particle size distribution in
the range of 10-30 pm (Galli et al., 1981, 1984; Bohm et
al., 1988; Kakugo et al., 1988; Kim and Woo, 1989). Kim
and Woo (1989)examined TiCUMgCldSiOZcatalyst using

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 461


100

8o
70

CATALYST

/
,
0-38

45-75
3845

100-250
500-1000
75-100
250-500
>loo0

PARTICLE SIZE, micron

Figure 12. Particle size replication of ethylene polymerizationwith


silica-supportedZiegler-Nattacatalyst (Munoz-Escalonaet al., 1988).

scanning electron microscopy and found that the catalyst


consists of small particles with a rough surface. However,
the internal structure of the catalyst was not shown. If
the catalyst is impregnated on silica support, the particle
size is in the range of 30-250 pm (Bailly and Speakman,
1986; Munoz-Escalona, 1988). The average polymer
particle size is about 20 times larger in diameter than the
catalyst particle size (Galli et al., 1981, 1984; Karol et al.,
1987).
Kakugo et al. (l988,1989a,b) recently studied internal
morphologydevelopment of polypropylene particles using
transmission electron microscopy. Their results indicate
that each polymer primary particle has a nucleus 50-170
A in diameter, which is believed to be the catalyst primary
particle. It is interesting to notice that the catalyst primary
particle size identified in the polymer particle is in excellent
agreement with catalyst crystal particle size estimated by
Chien (19871, as shown in Table 8. This suggests that the
minicrystal size of a catalyst achieved by ball-milling
cannot be reduced further during polymerization, or that
the Ti catalyst active center is fixed on the surface of this
mini MgClz crystal and not inside the minicrystal. The
diameter of a polymer primary particle is about 0.2 pm.
The fine polymer globules, about 1 pm in diameter, are
formed by fusion of several polymer primary particles.
The gross polymer particles are composed of these primary
particle aggregates. Weist et al. (1989) found that the
pore distribution of silica supported catalyst does not
change at the initial stage of polymerization. However,
the pore size of a catalyst increases with an increase in
polymer yield, indicating fragmentation during polymerization. The SEM photographs of polymer particles
show that the catalyst particles are completely encapsulated by polymer in the early stage (<5 min) of polymerization. Rupture of the outer layers of polymers particles
occurs as early as 5 min after the initiation of polymerization. Both spherical and fiberlike particles are formed
during ethylene polymerization (Kim and Woo, 1989).
On the basis of the available evidence in the literature,
polymer morphological development during polymerization with supported heterogeneous catalysts can be
summarized as follows:
1. An activated porous catalyst particle is composed of
primary particles of size in the range of 50-200 A. These
catalyst primary particles are held together by van der

Waals force. There are interstices or pores between the


primary particles, providing a large surface area.
2. At the beginning of polymerization,monomer diffuses
through the interstices and adsorbs onto the surface of
catalyst primary particles, where the active centers are
located (Bohm, 1978a).
3. The propagation reaction occurs on the surface of
the catalyst primary particles. When monomer is converted to polymer, polymer precipitates and accumulates
around the catalyst primary particles. As polymerization
proceeds, the interstices are gradually filled with polymers.
Hydraulic pressure or stress between catalyst primary
particles increases due to accumulation of polymer chains
in the pores leading to separation of catalyst primary
particles. Polymer around the catalyst primary particles
cements these primary particles. The distance between
the catalyst primary particles increases with an increase
in polymer yield. Thus polymer particles grow isotropically during polymerization. When the polymer particles
with high polymer yield are examined using electron
microscopy, catalyst primary particles are found to be
uniformly dispersed in the polymer matrix as though the
catalyst were broken up during polymerization. In fact,
this should be regarded as a separation process. Catalyst
primary particles should be broken in the preparation
stage. Interstices or pores must exist between catalyst
primary particles, where active centers are located, so that
polymerization occurs on the surfaces of the interstices,
leading to separation of catalyst primary particles.
4. The polymer around the catalyst primary particles
is swollen with monomers (ethylene and a-olefin). Since
olefin polymerization is a highly exothermic process, the
temperature within the polymer primary particle could
increase rapidly, and may exceed the softening or melting
point of the resin (Karol et al., 1987). Due to the rapid
growth of primary particles, the polymer chains between
primary particles fuse together to form polymer primary
particle aggregates of diameters in the range of 0.5-1.0
pm.
5. The final polymer particles are composed of aggregates of primary particles with average diameters in the
range of 200-500 pm (Galli et al., 1981, 1984; Kakugo,
1988; Bohm et al., 1988). The morphology of the final
particles depends on polymer yield and the balance
between annealing and distorting forces during particle
growth. The primary particles grow isotropically (Kakugo,
1988);therefore, in general, the polymer particles replicate
their parent catalyst particles in shape and size.
The mechanism described above can be expressed
schematically as in Figure 13. To avoid confusion between
catalyst primary particle and polymer primary particle,
the catalyst primary particle is called a domain in Figure
13. Polymer primary particles are formed with a catalyst
domain as a core, which is the smallest resulting catalyst
particle during polymerization.
3.3. Chemical Reactions. From the preceding observations, the chemical reactions of ethylene polymerization can be envisioned as occurring at the interface
between the solid catalyst and the polymer matrix, where
the active centers are located. From gas-state monomer
to solid-state polymer, ethylene experiences a dramatic
physicochemical transition within a very short time. The
polymerization environment changes with the composition
of catalyst, polymerization process, reactant composition,
reactor operating conditions, and extent of polymerization.
Although intensive research activity has been focused on
Ziegler-Natta catalyst systems since their discovery in
the early 1950s, no definitive, unequivocal chemical

462 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994


CATALYST
'zH4

LOW
POLYMER
YIELD

MEDIATE

10-30 micron in diameter


(30-100 micron for silica supported catalyst)
Composed of minute crystal domains
20-200 A"in diameter

Table 9. Summary of Elementary Reactions for Ethylene


and a-Olefin Copolymerization
reaction

description
Activation

Formation of polymer primary particles,


-0 2 micron in diameter
Separation of catalyst domain

Growth of primary particles


Formetion of primary particle aggregates,
-1 micron in diameter

Figure 13. Schemeof polyethylenemorphology developmentduring


gas phase polymerization.

reaction mechanisms have been developedto fully describe


the kinetic behavior of ethylene homo/copolymerization,
due to the complexity of the systems employed. Nevertheless, the key elementary reactions have been established, which include formation of active centers, insertion
of monomer into the growing polymer chains, chaintransfer reactions, and catalyst deactivation. Most of the
proposed mechanisms are based on information about
polymerization rate, molecular weight and its distribution,
polymer chain microstructure, and active center concentrations. Detailed mechanisms have been discussed in
monographs (Boor, 1979; Kissin, 1985) and in recent
reviews (Taitand Watkins, 1989;Dusseault and Hsu, 1993).
Among the proposed propagation mechanisms, the one
proposed by Cossee (Cossee, 1964; Arlman and Cossee,
1964)together with its subsequent modifications has been
widely adopted. With Cossee's mechanism in mind, Bohm
(1978a) proposed perhaps the most comprehensive elementary reactions involved in the ethylene polymerization
process.
Since commercial production of LLDPE and HDPE
consists of a copolymerization process, obviously, copolymerization mechanisms are required to understand
kinetic behavior and polymer properties. Therefore,
Bohm's reaction model has been modified and extended
to the ethylene copolymerization processes in recent
modeling studies (Villermaux et al., 1989; de Carvalho et
al., 1989,1990;McAuley et al., 1990;Lorenzini et al., 1991;
Hutchinson et al., 1992). Furthermore, to illustrate
multimodal distribution of polymer chain composition and
broad moleclar weight distribution, the multiple active
center concept has been adopted in recent modeling
studies. Therefore, a comprehensive chemical reaction
mechanism should include not only reactions proposed
by Bohm (1978) but also reactions involving multicomponents at different active centers. Whether each type of
active center has the same or different reaction mechanisms is still uncertain (multiple center formation itself
has not been proved fundamentally). If one assumes that
all of the active centers perform the same reaction
mechanisms, but with different reaction rates for each
elementary reaction, then elementary reactions which are
commonly adopted in modeling studies can be summarized
as in Table 9, where C, is the catalyst potential active

spontaneous activation
activation by aluminum
alkyl (A)
activation by electron
donor (E)
activation by hydrogen

(Hz)

activation by monomer
1 (Mi)
activation by monomer
2 (Mz)
Initiation
initiation of M 1 by normal
active center
initiation of Mz by normal
active center
initiation of MI by active
center with H
initiation of Mz by active
center with H
initiation of MI by active
center with A
initiation of Mz by active
center with A
initiation of M 1 by active
center with E
initiation of Mz by active
center with E
Propagation
KPll
propagation of chain
type 1with M 1
P m , n , l + [MI3
Pm+l,n,I
K~LZ
propagation of chain
type 1with M 2
P*m,n,l + [ M 2 1
p*m,nt1,2
K~PI
propagation of chain
type 2 with MI
P m , n , z + [MI3 -* P m + l , n , *
KPln
propagation of chain
type 2 with Mz
P m , n , z + [Mz] P m , n + l , z
Chain Transfer
Kbpl
spontaneous chain transfer
or &elimination
P*m.n.i
PO+ 4'm.n
KIN
chain transfer to hydrogen
P m n L + [HJ
PH.O
+ Q'm.n
(Hz)
KfAl
chaintransfer to aluminum
P m , n , L + [AI
P*A,O + Q'm,n
alkyl (A)
Kls,
chain transfer to electron
P m , n , i + [El -+ P*E,o
+ P'm,n
donor (E)
KMl3
chain transfer to M 1
P m , n i + [MI]
P*I,o,I+ Q'm,n
K m
chain transfer to Mz
P m , n , i + [Mz]
*o,I,,
+ q'm.n
Deactivation
KW
spontaneous deactivation

Pm,n,i

-*

'

cd

qm,n

Kd7i

Pm,n,i

+ lZ1

plm,n,i

cd

+ qm,n

d'

+ qm,n

- '
-*

KW
KdW

..+

cd

Pm.n.i

4m.n

KdHl

Pm,n,+

+ [H21

d'

+ qm,n

d'

+ 4m,n

KdJ&

Pm,n,L

rMj]

deactivation by impurities or
poison (Z)
deactivation by aluminum
alkyl (A)
deactivation by electron
donor (E)
deactivation by hydrogen (H2)
deactivation by monomers

Other Possible Reactions


formation of short-chain
branches
P I , O J + q'ra
pr+1+,1
K'pz
formation of short-chain
branches
p O , l , z + q'ra
Pr++1.2
K'Pl
formation of long-chain
branches (rare)
P*m,n,l + q f r a
~mtr,nts,l
K'PI
formation of long-chain
branches (rare)
p m , n , 2 + Q'ra
Pm+r,nta,z
K'Pl

-+

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 463


center; P*o is the active center without polymer chain;
P*m,n,iis the active center with m units of monomer 1and
n units of monomer 2, with the third subscript i denoting
the chain terminal monomer type bonded to the active
center; qm,nis a dead polymer chain with a terminal double
bond; and q,,,, is a dead polymer chain without terminal
double bond. For simplification of notation, no subscript
is shown corresponding to the type of active center. The
reactions shown in Table 9 should be considered to occur
at each type of active center. The terminal model is
assumed to be valid for binary copolymerizations. However, if the penultimate model is valid for some systems,
the mechanisms in Table 9 can be modified accordingly.
Hutchinson et al. (1992) considered transformation reactions, which are assumed to form different active centers.
If an active center formed by transformation is independent of the original polymer chain, then the effect of a
transformation reaction is on the initiation reactions which
have been shown in Table 9. If an active center formed
by transformation depends on the original polymer chain
produced by this center, then the formation of each active
center must be treated individually. The detailed mechanisms of transformation reactions have not been given
(Hutchinson, 1992). Formation of short-chain branches
was proposed by Bohm (1978). It was assumed that a
polymer chain with a terminal double bond will participate
in propagation at an active center with one monomer unit.
Formation of long-chain branching is unlikely for ZieglerNatta polymerization systems. However, for Phillips
catalyst systems, polymer chains with long-chain branching
are formed, although branching frequency is low
(McDaniel, 1985). For ethylene polymerization with a
Ziegler-Natta catalyst at high temperature and high
pressure, polymer chains dissolve in the monomer phase.
Hence, polymer chains with a terminal double bond are
able to diffuse to active centers to participate in propagation and to form long-chain branching. However,
detailed information regarding this reaction has not been
published to date. The mechanisms in Table 9 are
summarized based on Ti-based catalysts. Reaction mechanisms for chromium oxide based catalysts are a subset
of those in Table 9. It should be pointed out that secondorder deactivation of active centers is not included in Table
9.
4. Reactor Modeling

Reactor modeling is the determination of a quantitative


relationship between reactor performance and reactor
operating conditions. It requires comprehensive understanding of polymerization processes,physical phenomena,
and chemical reaction mechanisms. The importance and
benefits of reactor modeling have been widely recognized
by both industrial and academic researchers. The application capability of a model depends on scope of the
modeling effort. Ray (1986, 1991) defined a modeling
hierarchy as microscale, mesoscale, and macroscale, according to the characteristics of the polymerization reactor
systems. The relationships between modeling scale and
polymerization systems are outlined in Figure 14 after
Ray (1986,1991). The emphasis of each modeling level,
in particular for ethylene polymerization systems, can be
summarized as follows:
1. Microscale modeling: polymerization rate development, or polymer yield; reactant species distribution;
molecular weight development and ita distribution; chemical composition of polymer chains and its distribution;
microstructures of polymer chains, including chain branching, unsaturated groups, and sequence distribution.

+ 4,

MICROSCALE

MESOSCALE

1
I

M ACROSCALE

Figure 14. Levels of polymerization reactor modeling.

2. Mesoscale modeling: interphase heat and mass


transfer; intraphase heat and mass transfer; fluid mechanics and micromixing; polymer particle morphology
development; polymer particle size and distribution.
3. Macroscale modeling: macromixing and residence
time distribution; overall material and energy balance;
heat and mass transfer from the reactor; reactor dynamics
and control; polymer grade transition and control.
There are no definitive boundaries between these
modeling levels. In fact, they often overlap during
modeling studies. For instance, knowledge of some
microscale modeling is required for mesoscale modeling
studies, and macroscale modeling depends upon understanding of microscale and mesoscale phenomena. A
complete model for ethylene polymerization, including
all three levels, has not been developed in the literature.
However, significant modeling efforts with emphasis on
specific levels have been published. Microscale and
mesoscale modeling studies are normally reported as
kinetic modeling in the literature. Macroscale modeling
is often referred to as dynamic modeling. Major contributions to ethylene polymerization reactor modeling are
now discussed.
4.1. Kinetic Modeling. Kinetic modeling has focused
mainly on polymerization rate and polymer molecular
weight development, which can be expressed as follows:
2

i=l j - 1

MW f ~ ~ , ~ ~ * ~ l , ~ ~ j .-.)
l , ~ ~ (2)
~ l
where binary copolymerization is assumed. With the longchain approximation, polymerization rate is a function of
the propagation rate constants, monomer concentrations,
and active center concentration. Complexity can arise
through the expression for active center concentration.
Molecular weight development is a function of all the
components which affect polymer chain length, as shown
in Table 9. One of the important features of ethylene
polymerization is the broad molecular weight distribution
of the polymer. Two theories have been developed to
explain broad MWD of polyethylene, namely, diffusion
theory and multiple active center theory. The former
emphasizes the effect of monomer concentration on
molecular weight development. The latter is more concerned with the influence of active centers and associated
kinetic parameters. Some of the experimental results in
support of these theories are as follows.
Diffusional limitations:
1. Polymerization rate depends on stirring speed when
agitation speed is below a critical level (Berger and
Grieveson, 1965; Bohm, 1978b).

, ~ ~ l , ~

464 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

2. Molecular weight of PE formed in the initial period


of polymerization is much higher than that of polymer
formed in later stages (Crabtree et al., 1973).
3. The polymerization rate of ethylene increases significantly in the presence of small amounts of comonomers,
such as propylene and l-hexene (Finogenova et al., 1980;
Soga et al., 1989). Soga et al. (1989) found that the
crystallinity of the copolymer decreases with an increase
in comonomer concentration. The reactivity ratio of
propylene to ethylene is less than 1. Therefore, an increase
in reaction rate in the presence of a small amount of
propylene is due to a decrease in the monomer diffusion
resistance. This suggests that diffusion resistance is very
significant in highly crystalline polymers, such as HDPE
(Soga et al., 1989).
The results of Finogenova et al. (1980) and Soga et al.
(1989) are very convincing. In fact, the effect of crystallinity on monomer concentration in the polymer matrix
has not been comprehensivelystudied. However,the effect
of a-olefins on the ethylene polymerization rate depends
on the a-olefin types. It is not certain whether physical
or chemical effects dominate. Crabtrees result is in
contradiction with data reported by Bohm (1978~)and
others (Kissin, 1985). Bohm (1978~)
showed that numberaverage molecular weight increases significantly with
reaction time during the early period of polymerization
and then levels off with an increase in reaction time.
Multiple active centers: The following phenomena
cannot be explained by diffusion limitation theory (Zucchini and Cecchin, 1983; Karol, 1984).
1. Different transition metal catalysts can provide large
changes in polymer MWD.
2. Chemical modification of catalyst changes the
breadth of the MWD.
3. Heterogeneous catalyst can produce broad MWD,
even when the polymer is in solution.
4. Homogeneous soluble catalysts provide narrow
MWD, even when the polymer is insoluble in the reaction
medium.
5. High-activity catalyst does not necessarily provide
broad MWD.
6. Ti-based catalyst can provide polymer with narrow
MWD, and vanadium-based catalyst produces resins of
broad MWD. A combination of these two catalyst systems,
however, gives MWD of intermediate breadth with the
same reactor operating conditions (Nicoletti et al., 1988).
7. Compositionof LLDPE separated using temperaturerising elution fractionation had two or three distinct peaks
(Wild et al., 1982a;Usami et al., 1986;Kelusky et al., 1987;
Mirabella et al., 1987;Hosoda, 1988). The reactivityratios,
as measured by 13C NMR, were different for each peak
(Usami et al., 1986).
Although the above phenomena cannot provide conclusive proof of multiple active center formation mechanisms, they do suggestthat molecular weight development
of PE depends on the nature of the catalyst and on the
polymerization mechanisms. Many mathematical models
have been developed to quantitatively describe the kinetic
phenomena of ethylene polymerization. Most of the early
models focused on polymerization rate development (Keii,
1972). However,the emphasis has been placed on polymer
quality development in more recent modeling studies. The
progress of kinetic model development is now discussed.
Bohm (1978a) developed a model based on a comprehensive chemical reaction scheme, which is applicable to
both homogeneous and heterogeneous catalyst systems.
The model development is based on assumptions of steady
state for all active species, kinetic rate constants inde-

pendent of chain length, and slow deactivation of the


catalyst. However, the model is limited to polymerization
rate and number-average molecular weight (M,)
of homopolymerization. Higher order molecular weight and
MWD were not considered.
Dussealt (1991)and Dusseault and Hsu (1993)developed
areaction rate model for gas phase ethylene polymerization
based on the multiple active sites hypothesis. This model
is able to incorporate different distribution functions of
active centers to simulate polymerization rate behavior.
Both first-order and second-order deactivation mechanisms are considered in the model calculations (Dussealt,
1991). However, the model treatment is limited to
polymerization rate. The effect of multiple active site
distribution on polymer molecular weight development
was not considered.
de Carvalho et al. (1989) developed a comprehensive
ethylene and a-olefin copolymerization model accounting
or the formation, initiation, and deactivation of active
centers, and for the spontaneous and chain transfer to
hydrogen, monomer, and organometallics. The state of
the art to model copolymerization on multiple active sites
was described in detail. This general model was simplified
by assuming a stationary state of growing chains and
neglecting deactivation of active sites. The molecular
weight distribution was developed on the basis of these
assumptions. This model shows that the instantaneous
polydispersity for each active site is equal to 2, and the
instantaneous polydispersity for the entire amount of
polymer produced would deviate from 2 by the ratio of the
variance of the propagation rate constants to the mean
propagation rate constant. The effect of the propagation
rate constant distribution on molecular weight development was demonstrated by assuming unimodal and skewed
bimodal distributions of propagation rate constants. The
pseudo-kinetic rate constant method was used to evaluate
the kinetic rate constants for each site. However, chaintransfer reactions were not considered in the active center
fraction calculations. Effects of chain-transfer reactions
on active center fraction calculations can be very important
if chain-transfer reactions are significant (Xie and
Hamielec, 1993a-c). de Carvalho et al. (1990) recently
extended their model to calculate both the stereo and
chemical sequence distributions of copolymer. The models
developed by de Carvalho et al. (1989,1990)have not been
evaluated using experimental data.
Villermauax et al. (1989) and Lorenzini et al. (1991)
developed a model for ethylene and a-olefin copolymerization with Ziegler-Natta catalyst at high temperature
and high pressure. The model development is based on
a single active site, which is then extended to two or three
active sites using a mixing law. This model can fit
molecular weight averages and distribution data over the
range of the experimental conditions. The kinetic parameters were estimated by fitting the model to experimental data (Lorenzini et al., 1991). However, the effects
of monomer type on the reactivity of active sites are
neglected in the parameter estimation. In other words,
ethylene and butene-1 copolymerization were treated as
r1 = r2 = 1. The polydispersity of polymer obtained for
their system at higher temperature and high pressure is
between 3 and 5. It is worth noting that the polymerization
occurs in a supercritical phase, the polymer being soluble
in the monomer mixture. Therefore, there is no monomer
diffusion limitation in this system. However, the polydispersity of polymer is similar to that obtained in a lowtemperature system. This suggests that multiple active
sites play a central role in determining molecular weight

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 465

.Second-order deactivating cataIyst(I


Actlvlty :-5500 g / g - c a t hr

0
E

"i

8-

*.

Prediction

D,

1 x 10-~ c d i s

+ + Laboratory Data

7ooOo

960000

3
4 s
,%J

'

3oooo

4oooo-

*++L++**+*-'+++++

ib

io io

4b

**+**++,++++++

io ab

7b

io ~b id^ iio

Time in Hours
Figure 15. Comparison of model predictions and industrial laboratory data for gas phase ethylene copolymerization in a fluidized
bed reactor. Reprinted with permission from McAuley et al. (1990).
Copyright 1990 AIChE.

development. On the basis of this model analysis, the


realkylation, spontaneous transfer and deactivation reactions play only minor roles during high-temperature and
high-pressure polymerization.
McAuley et al. (1990)developed a comprehensive model
for gas phase ethylene copolymerization in a fluidized bed
reactor, such as the UNIPOL system shown in Figure 3.
Besides elementary reactions described by de Carvalho et
al. (19891, the model also accounts for the effect of
impurities on catalyst deactivation. This model can be
used to predict average chain lengths for multiple copolymerization. The kinetic rate constants in each active
site are expressed as pseudo-kinetic rate constants. Applying Stockmayer's bivariate distribution function (Stockmayer, 1945) to each active site, McAuley et al. (1990),for
the first time, were able to calculate molecular weight and
composition distributions for ethylene copolymers. The
model was evaluated assuming two active sites of catalyst.
The model predictions are in excellent agreement with
experimental data as shown in Figure 15.
The models discussed above focused on the effects of
chemical reactions and catalyst on polymer property
development using the hypothesis of multiple active center
distribution. Physical phenomena were simplified in these
model developments. Assuming the polymer particles to
be isothermal and spherical, Galvan and Tirrell (1986)
developed a model for molecular weight calculation of
olefin homopolymerization which accounted for two active
sites and the limitation of monomer diffusion. The
simulation results showed that the presence of different
types of active sites is responsible for the broad MWD of
olefin polymerization. Even in the case in which there are
apparent monomer diffusion limitations, multiple active
sites still play an important role in determining molecular
weight development.
Significant contributions have been made by Ray and
his co-workers over the past decade toward a better
understanding of olefin polymerizations using ZieglerNatta catalysts (Ray, 19881,including experimental studies
of olefin polymerization behavior (Taylor et al., 1981;Yuan
et al., 1982; Choi et al., 1983; Choi and Ray, 1985b; Yoon
and Ray, 1987); polymer molecular weights and polydispersity modeling (Nagel et al., 1980; Floyd et al., 1987;
1988); microparticle, macroparticle, and external film
mass- and heat-transfer phenomena (Floyd et al., 1986ac); particle ignition, extinction, monomer sorption phenomena, and particle morphology development (Hutchinson and Ray, 1987, 1990, 1991a,b; Hutchinson et al.,
1992); and reactor dynamics and control studies (Choi
and Ray, 1985c, 1988).
On the basis of the polymer morphology development

n
2

) . . .

Mb
1

Mn2

0.0

0.2

0.4

4 mOl/f
1

'

30,000

0.8

0.8

1.0

WElaHT FRACTION
Figure 17. Polydispersityresulting from combinationof two polymer
fractions, each with polydispersity of 2. Reprinted with permission
from Floyd et al. (1987). Copyright 1987 John Wiley & Sons.

discussed above, as shown in Figure 13,a multigrain model


of particle growth was used to study heat and mass
transport phenomena. According to Floyd et al. (1986a),
intraparticle temperature gradients are negligible for both
the microparticles and the macroparticles in gas or slurry
polymerization reactors. Exceptions would be for large
highly active catalyst particles early in the lifetime of the
polymer particles in gas phase reactors. This conclusion
is in agreement with commercial observations (Bailly and
Speakman, 1986). Intraparticle monomer concentration
gradients in the microparticles can be significant for highactivity-catalyst systems having large primary crystals of
catalyst in a gas phase polymerization system. However,
concentration gradients in the macroparticles will normally
be negligible for gas phase polymerization.
Floyd et al. (1987, 1988) further studied the effects of
intraparticle and external boundary layer transport resistance on kinetic behavior and polymer property development based on the multigrain model. It was found
that diffusion resistance affects the polymerization rate
more strongly than the molecular weights. Their simulation results show that even in a highly diffusion-limited
situation, the polydispersity drops below 4 within 30 min
after polymerization begins. A typical simulation result
is shown in Figure 16 (Floyd et al., 1987). Therefore,
monomer concentration gradients in the macroparticle

466 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994


10-1

u * 2 cm/s

fi inslonloneous breakup

Table 10. Heat- and Mass-Transfer Resistance in Gas


Phase Polymerization (Flosed et al., 1986a.b. 1987. 1988)
microparticle
T
negligible, <1 O C
gradients
[MI
can be significant for high-activity
catalyst and for those yielding large
primary crystallites after breakup
macroparticle
T
negligible except for large particles
gradients
of high-activity catalyst early
in particle lifetime
negligible except for large particles
[MI
of high-cavity catalyst early
in particle lifetime
externalfilm
T
often very significant during
gradients
initial stage of particle growth
negligible, <1%
[MI
~~

00
IO2

3 r

0.5

1.0

1.5

2.0

2.5

3.0

hslonloneous brrohuv

00

05

I.o

1.5

20

25

3.0

TIME (sec)
Figure 18. External film heat- and mass-transfer resistances in gas
phase polymerization of polyethylene under stirred bed conditions
with high activity catalyst. Reprinted with permission from Floyd
et al. (1986~).Copyright 1986 John Wiley & Sons.

alone cannot explain the breadth of the MWD under


normal reactor operating conditions. However, reactivities
of active sites affect polydispersity significantly. Figure
17 showsthe effects of molecular weight and weight fraction
produced in each site on polydispersity assuming two active
sites (Floyd et al., 1987). This figure shows that the
molecular weights produced by the two sites must differ
by at least an order of magnitude to achieve broad
polydispersities (PD > 5). If polydispersity is the same
for each site, the overall polydispersity is always at a
maximum when each site contributes 50% of the total
weight of polymer. A bimodal MWD will result if the
polymer with polydispersity greater than 6.0 is produced
from two types of active site. To produce unimodal MWD
with high polydispersity, a large number of different types
of active site (>4) is required. However, an increase in
the number of active site types may not necessarily increase
the polydispersity. Polydispersity and MWD are not
sensitive to catalyst particle size distribution.
Floyd et al. (1986b,c) calculated external film heattransfer resistances in gas phase ethylene polymerization
in a stirred bed reactor using the Ranz-Marshall correlation. They found that operating a gas phase reactor
with either a low- or high-activity catalyst would lead to
a significant particle temperature rise that would influence
observed kinetics early in the polymerization, and would
lead to agglomeration and sticking of polymer particles
for high-activity catalyst above 30-60 pm in diameter.
Figure 18shows simulations of external film temperatures
and monomer concentration gradients of polymer particles
in an early stage of polymerization (Floyd et al., 1986~).
It can be seen that the concentration gradient is less than
1%; however, temperature rise on the particle surface is
very significant for large catalyst particles. I t should be
pointed out that this simulation was based on a stirred
bed reactor with a very slow gas flow velocity (2 cm/s). For
fluidizedbed reactors, the temperature rise should be much
less than that shown in Figure 18. The main results of
heat and mass transport for gas phase polymerization are
summarized in Table 10 (Floyd et al., 1986a,b, 1987,1988).
Hutchinson and Ray (1991a)recently studied the effect
of prepolymerization on polymer property development
and found that prepolymerization does not affect molecular
weight development, but increases polymer bulk density.
Bulk density increases with an increase in reactor operating
temperature. If prepolymerization is carried out at high

temperature and at low pressure, reactor fouling is more


serious. An attempt to model morphology development
of polymer particles has also been made in a recent
publication (Hutchinson et al., 1992).
Significant modeling studies of heat and mass transfer
during olefin polymerization from the gas phase have also
been carried out by Laurence and Chiovetta (1983). Their
model development was based on the following
assumptions: microparticle and macroparticle are spherical; diffusivity is independent of temperature; and microparticle sizes are equal. Their simulations showed that
there is no significant monomer diffusion effect during
the progress of the reaction. The monomer concentration
profile becomes almost uniform inside the microparticles
within 3 min. The temperature gradient is less than 1"C.
However, the monomer gradient in the initial stage is
significant. The resultsof Floyd et al. (1986a,b, 1987,1988)
discussed above are in general agreement with model
predictions by Laurence and Chiovetta (1983). Webb et
al. (1989a) determined sorption of ethylene onto the
catalyst and silica support using solid-gas chromatography
and gravimetry and found that ethylene sorbs on both the
silica support and the catalyst. However, sorption on
polymerization catalyst is about 20% higher than on its
support. The sorption of ethylene in the high yield polymer
particles is 50 times less than the sorption of ethylene by
the bare catalyst. Webb et al. (1989b) measured internal
porosity and intraparticle diffusivity using solid-gas
chromatography. It was observed that internal void space
and transport rate decrease significantly at a yield of 0.1
g of polymer/g of catalyst. Polymerization to such low
yields fills the macropores of the catalyst nearest the
external particle surface with polymer. However,the rate
of adsorption of monomer is very fast and is not rate
limiting. Webb et al. (1991) further discussed morphological influences in the initial stage of gas phase polymerization and proposed a kinetic model accounting for
nonideal morphological influences.
On the basis of the above studies, it is safe to state that
multiple active sites play an important role in determining
PE molecular weight development, whereas the effect of
diffusion limitations on molecular weight development
plays only a minor role. However, diffusion limitation
may be significant in controlling the polymerization rate
for a system with a high-activity catalyst and highly
crystalline polymer matrix. Even in a diffusionally limited
situation, MWD is still controlled by the nature of the
catalyst system and the ethylene polymerization mechanisms. Therefore, in the future, polymer property
modeling should focus on multiple active site effects and
exploit the mechanisms of formation of different active
centers. The current state of the modeling of molecular
weight development published in the literature is summarized in Table 11. Accounting for multiple active sites,
detailed elementary chemical reactions, and mole fraction

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 467


Table 11. Approaches to Polymer Property Modeling
model structure
P, ,P polydispersity

R,, P, ,P chain length


distribution, polymer
composition distribution,
stereo and chemical sequence
length distribution

R,, M,,a, Ma,MWD, short chain


branches, double bonds

R,, P,,

p,, bivariate molecular


weight and composition
distribution of copolymer

M,,a,, polydispersity

basic assumption
isothermal, spherical
particles, two active sites,
effect of diffusion control
1imitation
multiple active sites,
concentration of reactant
is independent of active
sites, deactivation depends
on site type, steady state
for growing chains
two or three active sites,
steady state for living
chains, first-order
deactivation
multiple active sites,
monomer concentration
in the active sites is
proportional to gas phase
pressure and independent
of sites, steady state for
growing chains, uniform gas
composition and temperature
multiple active sites,
consider effect of diffusion
limitation, steady state
for growing chains, multiple
grain particles

of each active site, Xie et al. (1993) recently developed a


comprehensive model for molecular weight calculations
of ethylene and a-olefin copolymers. The model can
predict molecular weight with broad polydispersity ranging
from 2 to 40 or higher depending on polymerization
conditions. The model can be used for copolymer molecular weight calculations of not only heterogeneous
Ziegler-Natta catalyst but also metallocene-based catalyst
systems.
4.2. Dynamic Process Modeling and Reactor Control. Dynamic process modeling considers the behavior
of the entire reactor system, as shown in Figure 14.
Therefore, dynamic modeling should incorporate kinetics
of polymerization, heat and mass transfer, and reactor
characteristics in arealistic manner. Choi and Ray ( 1 9 8 5 ~
1988)were the first to model the dynamic behavior of gas
phase polymerization in a fluidized bed reactor and in a
stirred bed reactor. Their model development for a
fluidized bed reactor is based on the followingassumptions:
1. The emulsion phase is perfectly backmixed and is at
incipient fluidizing conditions with constant voidage.
2. The flow of gas in excess of minimum fluidization
passes through the bed in the form of bubbles. The bubbles
are spherical, uniform, and in plug flow. The bubble phase
is always at a quasi-steady state.
3. Polymerization occurs only in the emulsion phase.
4. Mass and heat transfer occur at uniform rates over
the bed height.
5. There is negligible mass- and heat-transfer resistance
between the solid polymer particles and the emulsion
phase.
6. No elutriation of solids occurs.
7. There is no catalyst deactivation.
8. The fluidization bed height (reaction zone height) is
constant.
The model simulations show both steady-state multiplicity and bifurcation phenomena in a fluidized bed
reactor. Temperature runaway in the reactor can occur
in open-loop operation if parameters are set beyond the
Holf bifurcation point or if disturbances carry the reactor
beyond the domain of attraction of the lower steady state.
Under feedback temperature control the reactor can be

reactor and process

reference

batch, slurry system,


homopolymerization

Galvan and Tirrell, (1986)

batch or CSTR, slurry


system, copolymerization

de Carvalho et al. (1989,1990)

CSTR, high-temperature
and high-pressure
copolymerization

Villermaux et al. (1989);


Lorenzini et al. (1991)

fluidized bed reactor or


CSTR, multicomponent
gas phase polymerization

McAuley et al. (1990)

batch, slurry or gas phase


homo- or copolymerization

Nagel et al. (1980);


Floyd et al. (1987);
Hutchinson et al. (1992)

stabilized, provided that the recycle heat exchanger has


sufficient cooling capacity. However, even under closedloop control, temperature runaway can occur in some
situations when the rate of heat removal by recirculating
gas becomes lower than the rate of heat generation in the
reactor. Hence, additional coolant is required to maintain
stable reactor operation conditions for high reaction rate
systems.
Assuming that the stirred bed reactor is adiabatic and
heat removal depends on evaporation of liquid monomer,
Choi and Ray (1988) found that a stirred bed reactor was
open-loop-unstable for most operating conditions with or
without a bed level controller. The reactor instability was
due to the limited heat removal capability of the system.
Although the reactor is open-loop-unstable, the dynamics
are so sluggish that application of feedback controllers on
bed level, temperature, and pressure can easily stabilize
the system.
Under the assumptions listed above, Choi and Ray
(1985~)did not consider the existence of a maximum
bubble size, nor the effects of a varying bed height or inert
gas on the fluid dynamics of the reactor. Talbot (1990)
and McAuley et al. (1993) relaxed those assumptions and
considered the bubbles to grow only to their maximum
stable size under certain reactor operating conditions, and
assumed further than the bubbles are of constant size at
all levels within the bed and travel through the bed in
plug flow. The simulation results show that bubble size
critically influences the rate of heat and mass transfer
within the bed, and when the bubbles are as small as those
predicted by the maximum stable size correlations, there
is little or no resistance to the transfer of monomer and
heat between the phases. McAuley et al. (1993) further
simplified the two-phase model to the well-mixed model
by assuming unrestricted mass and energy transfer between the bubble and emulsion phases. The simulations
show that only minor deviations exist between the
predictions of the two models for superficial velocities in
the range of 3-6 times the minimum fluidization velocity
and for the small bubble size predicted by the maximum
stable size correlations. Therefore, the well-mixed gas
assumption is valid for modeling gas phase polymerization.

'

468 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

30u

SPECIFICATION LIMITS

COMPLETED

O.DOO

n.nn2

0.n08

0.004

0.008

o.010

Partlcle Slze (cm)

Figure 19. Particle size distribution of catalyst (Talbot, 1990).

.in

Dlstrlbutlon

,,,,,,,,, ,,,,,,,,, .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

o. aa

a. as

a.

io

0.

IS

a. 20

Partlcle Size (cm)

Figure 20. Effect of catalyst particle size distribution on polymer


particle size distribution (Talbot, 1990).

The dynamic model developed by Talbot (1990) also


accounts for the effects of a varying bed height and inert
gas on the fluid dynamics. These modifications permit
the model to simulate reactor startup behavior. It was
found that when the bed height is allowed to increase to
a desired level before product removal begins, the temperature rise in the bed is more rapid than when the bed
height is kept constant at a desired level. Hence, a
temperature overshoot may occur when a varying bed
height startup procedure is employed. Using a population
balance approach, Talbot (1990) also developed a model
for polymer particle size calculations. This model can be
used to predict the effects of catalyst particle size and
distribution on polymer particle size development. For
instance, a distribution of catalyst particle size can produce
particle size distribution of polymer product which is broad
and skewed, as shown in Figures 19 and 20 (Talbot, 1990).
Understanding the dynamic behavior of a reactor is
important not only for reactor design and process control
under normal operating conditions, but also for product
quality control, particularly during the resin grade transition period. Although gas phase polymerization is a
versatile process, as shown in Figure 9, specific reactor
operating conditions produce a specific grade of polyethylene. To meet the market demand, a polyethylene process
is required to produce a number of different product grades
in the same reactor. For example, the Phillips plant
produces as many as 50 different grades of polymer
(McDaniel, 1988). Therefore, all polyethylene manufacturers regularly face grade changeover operations, and

2
TIME, hr

Figure 21. Grade changeover scheme for ethylene polymerization


(Sinclair, 1987).

detailed procedures for grade changeover are important


knowledge for an individual manufacturer. Very little
information on this topic is available in the open literature.
Sinclair (1987) has outlined the considerable economic
importance and difficulties of grade changeoveroperation.
Figure 21 shows an example of lost production time and
off-specificationproduct during changeover from one grade
to another (Sinclair, 1987). In the case shown in Figure
21, the polymerization rate must be reduced to 88% of the
nameplate capacity at the beginning of the grade change
point a to maintain safe reactor conditions. The polymerization rate is held at the reduced rate until point b
while the reactor conditions change toward those required
for the new target grade. At point c, the polymerization
rate is gradually increased within the reactor safety
limitations. At point d, the polymerization rate reaches
the lower specification limit for the product, and steadystate conditions for the new grade are finally reached at
point e. Off-specification product is produced between
the beginning of the grade change (point a) and the
attainment of specifications for the new grade (point d).
The amount of this product is the area under the
polymerization rate curve from point a to point d. The
lost production time can be estimated as the difference
between what the reactor could have produced during the
grade changeover and the actual production during the
change. Therefore, knowledge of how to minimize offspecificationproduct and maximize production rate during
grade changeovers is crucial in plant operation.
Ramanathan and Ray (1991) have simulated grade
transition strategies for two fluidized bed polyolefin
reactors in series and found that catalyst feed rate,
hydrogen feed rate, and gas bleed rate are important
variables for accomplishing rapid grade transitions while
maintaining the polymer production rate. More recently,
optimal open-loop policies for accomplishing grade
changeovers of polyethylene production at commercial
operating conditions were studied by McAuley and
MacGregor (1992) using dynamic model-based optimization. McAuley and MacGregor (1991)related the multiple
active center model with appropriate simplifications to
the melt index of polyethylene so that the model can be
used to predict melt index which can be measured in the
industrial laboratory environment. Adjustable parameters
in the model are updated on-line by a recursive prediction
error method as measurement data become available.
Therefore, the model can be used for on-line prediction
based on laboratory measurements and for grade transition
policy studies. A desirable grade transition policy must
satisfy the following conditions:

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 469


Oo4

0
0

0.4
0.8
1.2
1.6
SCALED MOLE FRACTION OF BUTENE IN POLYMER

2.0

Figure 22. Steady-state copolymer compositiondistributionsof two


grades of polyethylene (McAuley and MacGregor, 1992).

a. Safety and stability of reactor performance during


the grade changeover must be ensured. Excursions above
the sticking temperature of the polymer and disruption
of fluidization, heat transfer, and mixing in the bed cannot
be tolerated.
b. Minimum quantity of off-specification polymer and
short transition time are desired.
c. All reactor operating variables should settle at their
prespecified final values when grade transition is complete.
Production of consistent on-specification product under
new reactor operating conditions is required.
With these conditions as criteria, McAuley and MacGregor (1992) used the following integral quadratic objective
functions to obtain optimal transition policies.

a,(ln MI, - In
Vpdes)'

+ a,(pi - P,,)~I dt + J~[w5(VP- R p d e ~ ) ~dt] + wS(Tf8p - Tde8)2 +


w T ( B d 8 p - Bwde8)z (4)

where Fl(u) and Fz(u)are objective functions, MI and p


are the melt index and density of polymer, respectively,
Tfsp and Tdes are the final steady-state temperature and
desired temperature, Vp and Vpdes are the gas bleed valve
position and desired valve settling position, R, is the
polymerization rate, B, is the bed level, t o and tf are the
grade transition beginning time and the time at which the
reactor reaches steady state beyond the end of the
changeover, and wi (i = 1-8) and ai (j = 1-4) are weighting
factors related to the relative costs of deviations of each
variable. Subscripts c and sp represent the accumulated
value and the set point, respectively. Three policies of
grade changeover were demonstrated using eqs 3 and 4:
McAuley and MacGregor (1992) pointed out that the
molecular weight and composition distributions may not
reach the steady-state specification even though the
average properties MI, and p, of the polymer in the reactor
are within their specification limits, as shown in Figures
22 and 23 for a changeover from grade B to grade A. Figure
22 shows the steady-state composition distributions for
both grades A and B. At the end of the changeover, the
composition of grade A is as shown in figure 23, which is
very different from the steady-state grade A shown in
Figure 22. This is because some of the old polymer

0.4

0.8

1.2

1.6

2.0

SCALED MOLE FRACTION OF BUTENE IN POLYMER

Figure 23. Composition distribution of copolymer at end of grade


transition for polyethylene (McAuley and MacGregor, 1992).

produced before the start of the grade changeover remains


in the reactor after the changeover, having been mixed
with the new polymer produced during the changeover.
Figures 22 and 23 suggest that a longer time is required
for a complete grade changeover of the composition
distribution, compared with the times for average properties.
Except the bed level, all of the operating variables
discussed above only affect the properties of polymer
produced after the beginning of the grade changeover,
because the properties of dead polymer are determined
by the kinetics of polymerization and reactor operating
conditions at the time at which the polymer was produced.
Therefore, reactor operating variables only affect newly
born polymer in the reactor. Thus, to achieve fast grade
changeover, properties of the new polymer should be
brought as close as possible to their targets. On the basis
of this philosophy, McAuley and MacGregor (1993)
designed a nonlinear model-inverse based feedback controller, choosing instantaneous MIi and pi as controlled
variables, instead of accumulated values. An on-line
polymer property inference scheme (McAuleyand MacGregor, 1991) and open-loop optimal policies (McAuley and
MacGregor, 1992) are implemented in the designed
controller. The control action also accounts for impurities
and catalyst composition disturbance during grade
changeovers. Simulation results show that the controller
is able to control the properties of many grades of
polyethylene and to accomplish nearly optimal changeover
between these grades.
5. Polymer Properties and Characterization

The objective of process development and kinetic/


dynamic modeling studies is to produce a quality polymer
with high productivity. Hence, it is very important to
understand the relationships among polymer end-use
properties, polymer characteristics, and reactor operating
conditions.
5.1. Physical and Mechanical Properties. Physical
and mechanical properties of polyethylene material depend
on density, crystallinity, molecular weight and its distribution, chemical composition, and other characteristics.
These variables, in turn, are controlled by polymerization
conditions and are not mutually independent. The
essential variable, which affects all the physical properties,
is the chain microstructure of polyethylene. Backbone
linearity contributes to improve tensile and tear properties,
while branching increases toughness of the material. The
combination of the two in LLDPE resins gives physical
properties that are generally superior to those of conven-

470 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

* STIFFNESS

'TENSILE STRENGTH
* HEAT RESISTANCE
* HARDNESS
* PERMEATION RESISTANCE
'SHRINMGE

* DISTORTION RESISTANCE
* OPTICALS 7

;
v)

* STRESS CRACK RESISTANCE


* ELONGATION
*TENSILE STRENGTH
* MELT STRENGTH
* ORIENTATION
* ELASTlClY

* STRESS CRACK RESISTANCE

* FLEXIBILITY
* IMPACT STRENGTH

/
)O

\
I

DENSITY, g/ml

0 97

Figure 24. Relationships between polymer properties and resin


density (Foster, 1991).

tional LDPE (Sinclair, 1983). Commercial resins are


mainly specified by density and molecular weight (melt
index) and distribution (melt flow ratio). In other words,
the processing and end-use properties are mainly governed
by these variables.
Density of PE depends on the following factors: the
number of branches; branch chain length; branch frequency distribution; chemical composition distribution;
molecular weight and distribution.
In general, the density of PE decreases with an increase
in branch numbers; the more branching, the lower the
density (Shirayama et al., 1972; Sinclair, 1983). The
crystallinity of PE decreases significantly with an increase
in branch frequency and size. Hence, any physical
properties related to crystallinity, such as stiffness and
yield stress, will be affected by branching or chemical
composition. Density decreases with an increase in
molecular weight. This is because of the inhibition of
crystallization by longer molecular chains. In commercial
copolymers made using heterogeneous Ziegler-Natta catalysts, high molecular weight fractions generally contain
less comonomer units and fewer branch chains, while low
molecular fractions contain 2-4 times the comonomer
content of the high molecular weight fractions. This
suggests that comonomer units are more sensitive to chain
transfer and other termination reactions during polymerization. This is also one of the reasons why the density
of polymer depends on MWD. The basic relationships
between polymer properties and density can be summarized in Figure 24 (Foster, 1991).
The effects of molecular weight and distribution on
mechanical properties of polymers are very complex
(Martin et al., 1972; Nunes et al., 1982). Stiffness and
hardness of PE decrease with an increase in molecular
weight for molecular weights < 106. For molecular weights
> lo6, the flexural stiffness is independent of molecular
weight (Margolies, 1970). The stiffness of a resin is a
function of crystallinity. As molecular weight of P E
increases, the percent crystallinity decreases so that the
stiffness decreases (Shirayama et al., 1972; Nunes et al.,
1982).
As the molecular weight of PE increases, the tensile
impact and the Izod impact increase and reach a maximum
at a molecular weight of about 106, and then decrease
gradually with an increase in molecular weight (Margolies,
1970).
Figure 25 summarizes general relationships between
polymer properties and molecular weight (Foster, 1991).
The significance of these relationships also depends on

0.1

MELT INDEX

100

-MOLECULAR WEIGHT

Figure 25. Relationships between polymer properties and resin


molecular weight (Foster, 1991).

0.93-0.97
HDPE
giml

I,
LLDPE
0.91-0.93 giml

LDPE
0.91-0 93 glml

Figure 26. Schematic of molecular structures for commercial


polyethylene.

MWD. Resins with a broad MWD show enhanced melt


shear thinning compared to that of narrow MWD resins,
and therefore tend to flow more easily at the high shear
rates prevailing in resin processing. Hence, broad MWD
resins are widely used where extrusion cross-sectionalareas
are relatively large and where a broad MWD results in
easier processing. In blow molding of large containers,
the combination of high molecular weight and broad MWD
gives high melt strength, low parison sag and easy
processing; the same applies for pipe extrusion (Sinclair,
1983).
5.2. Polyethylene Characterization. Commercial
polyethylenes are essentially branched polymers, LDPE
with long-chain branching, and both LLDPE and HDPE
with short-chain branching, as illustrated Figure 26.
Hence, polyethylene characterization requires analysis of
not only molecular size but also branching and composition.
In commercial practice, polyethylene resin grade specification is generally quoted in terms of melt index and
density (ASTM, 1990) rather than molecular weight and
composition. Polydispersity is indicated using a melt flow
ratio (Spitz, 1987) rather than the ratio of M,/i@,,. The
melt index and melt flow ratio give onlyrelative properties
of a polymer and cannot provide detailed information
about the polymer structure. It is difficult to relate melt
index to polymerization conditions. However, melt index
measurement is the most convenient way to determine
relative molecular size in an industrial environment, which
is why melt index is generally used for PE quality control.

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 471


40

32

A b

ia

25

16 .

%X 2 O
y

- c

P4:
m

16

$ 13

5 12

10

2 7.8 -

g 6.3
b>5.0

"3;

1
-

"1
Y

a
2

a 10
3

a 6
6

2.5

2.0

4
0 01

01
10
MELT INDEX, 9/10min

10

100

10

12
14
16
MELT FLOW RATIO

ia

20

Figure 27. Relationship between weight-averagemolecular weight


and melt index of polyethylene. (m) Sinclair (1983);(+) Spitz (1987);
(0)LLDPE (1-octene as comonomer); (*) LLDPE (1-butene as
comonomer); (A)HDPE (Bremner and Rudin, 1990).

Figure 28. Relationship between polydispersity and melt flow ratio


of polyethylene (Spitz, 1987).

One way to relate the melt index to polymerization


conditions is to correlate melt index with molecular
weights. Unfortunately, only limited data are available
in the literature for correlation of melt index and molecular
weight. Figure 27 shows the relationship between weightaverage molecular weight and melt index (Sinclair, 1983;
Spitz, 1987; Bremner et al., 1990). It can be seen that
molecular weight increases with a decrease in melt index;
however, there is no simple relationship between them
over a broad range of molecular weights. According to
measurements by Bremner and Rudin (1990),Mwand the
melt index relationship depend on the density or polymer
composition. The data for LLDPE with 1-butene as a
comonomer are very scattered. The data for LLDPE with
1-octene comonomer are in agreement with the data
reported by Sinclair (1983) and Spitz (1987). These data
can be correlated with the solid line shown in Figure 27.
A linear relationship between MWand melt index can only
be used over a limited range of molecular weights. The
results of Bremner and Rudin show that, at the same melt
index, HDPE has a higher molecular weight than LLDPE.
Poor correlation between molecular weight and melt
index can be attributed to both measurement error and
the nature of PE structure. Commercial low-pressure PES
are copolymers. Polymers with the same molecular weight
may have different compositions, i.e., short branches. The
melt flow property depends on both molecular weight and
branching. Hence, polymers with the same molecular
weight may have different melt index values. In other
words, melt index is a function of the whole distribution
of molecular weights and compositions and not just a single
molecular weight average value (Bemner et al., 1990;Pang
and Rudin, 1992). Furthermore, melt indexes and molecular weights published in the literature are often
measured using commercial polymers, for which the
synthesis histories of the samples axe often unknown.
Therefore, caution must be taken when using such data.
Figure 28 shows the relationship between polydispersity
and melt flow ratio, which is used to indicate broadness
of molecular weight distribution (Spitz, 1987). Although
the data are scattered, polydispersity seems to be linearly
related to flow ratio.
Although melt index and melt flow ratio can provide
relevant information about molecular weight and distribution, the true molecular weight and its distribution are
required to better understand polymer properties and their

relationships with polymerization mechanisms and reactor


operating conditions. Size exclusion chromatography
(SEC and GPC) can, in principle, provide both molecular
weight averages and distribution. The Waters Associates
150C ALC/GPC (high-temperature GPC) is commonly
used in both academic and industrial research for PE
molecular weight measurement. o-Dichlorobenzene or
1,2,44richlorobenzene is used as solvent (Kissin, 1985).
Virtually all polyethylenes are insoluble at room temperature; hence, the GPC is normally operated at 135-150
OC.
A GPC consists of two essential sections: a separation
section and a detector section. A GPC column fractionates
according to the hydrodynamic volume of polymer solute
in the mobile phase. Since the hydrodynamic volume
depends on both molecular weight and composition
(branch frequency), the polymer eluted in the detector
cell will not be uniform either in molecular weight or in
composition (Vela Estrada and Hamielec, 1992-93).
Hence, strictly speaking, GPC can only be valid for
molecular weight measurement of homopolyethylene.
Current commercially available detectors for GPC
measurements include (i) a differential refractive index
detector (DRI) to monitor polymer concentration in the
eluant as a function of retention time; (ii) continuous
measurement of the weight-average molecular weight of
the eluting species using a low-angle laser light scattering
photometer (LALLS); (iii) continuous measurement of
the molecular weight using a continuous viscometer (CV).
The LALLS signal is proportional to CMand CV responds
according to CMa, while DRI is proportional to C, where
C is concentration and M is molecular weight (Pang and
Rudin, 1992). DRI and CV are insensitive to high
molecular weight and low concentration fraction. On the
other hand, LALLS is very sensitive to high molecular
weight fractions but not to low molecular weight. Therefore, DRI and CV measurements may lose information on
high molecular weight fractions and LALLS may lose the
information on low molecular weight fractions, so that
three different molecular weights and distributions may
be obtained for the same PE sample using three different
detectors. Pang and Rudin (1992) concluded that no
current GPC technique provides a picture of the true
molecular weight distribution of broad MWD polyethylene.
For copolymer, such as LLDPE, GPC can provide only
approximate values of molecular weight and its distribution

472

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994

if the entire polymer sample is used. Since copolymer has


a bivariate distribution of molecular weight and chemical
composition, it is desirable to measure both molecular
weight and chemical composition of the polymer. This
can be achieved by a combination of temperature-rising
elution fractionation and GPC techniques.
The fractionation of polyethylene copolymer on the basis
of molecular weight alone can occur only if a solvent
gradient elution technique is performed at a temperature
above the melting point, while the fractionation takes place
mainly on the basis of the degree of short-chain branching
if the temperature-rising elution technique is employed.
The term temperature-rising elution fractionation
(TREF) was first coined by Shirayama et al. (1965) to
describe a method used to fractionate low-density polyethylene according to the degree of short-chain branching.
The actual technique had been described earlier by
Desreux and Spiegels, who first recognized the potential
of using elution at different temperatures to achieve a
crystallization separation (Hawkins and Smith, 1958).
Although the principle of TREF has been recognized for
over four decades, significant development of TREF
analytical techniques occurred only in the past decade. It
is interesting to note that the development of TREF
technology and application has taken place almost exclusively within industrial research laboratories (Hawkins
and Smith, 1958; Shirayama et al., 1965,1972; Wild and
Ryle, 1977;Bergstrom and Avela, 1979;Nakano and Goto,
1981; Wild et al., 1982a,b; Knobeloch and Wild, 1984;
Kimura et al., 1984; Usami et al., 1986; Kelusky et al.,
1987; Cady, 1987; Hosoda, 1988; Mirabella et al., 1987,
1992; Wild, 1989; Wilfong and Knight, 1990). Only very
recently has this technology captured the attention of the
academic community (Schouterden et al., 1987; Kulin et
al., 1988; Hamielec, 1989; Glockner, 1990; Vela Estrada
and Hamielec, 1992-93). No TREF instrument has been
commercialized to date. Hence, TREF instruments are
currently constructed by individual laboratories. Few
university-based polymer institutes have built TREF
equipment in recent years, e.g., McMaster University (Vela
Estrada and Hamielec, 1992-93), University of Waterloo,
and the University of Alberta. The development of TREF
is clearly driven by the need to understand the nature of
polyethylene copolymerswhich exhibit behavior indicative
of considerable structural heterogeneity and to estimate
the kinetic parameters associated with detailed kinetic
models.
The fundamental basis of TREF technology is the
relationship between crystallizability and branching of
polymer. Crystallizability of a polymer decreases with an
increase in branch frequency in polymer chains. In other
words, polymer chains with different branch concentrations crystallize at different temperatures. The higher
the degree of crystallinity, the higher is the dissolution
temperature of polymer chains; hence, polymer chains with
fewer branches dissolve at higher temperatures. To
separate polymer chains with different branch concentrations effectively in a TREF process, the polymer is first
coated on a spherical support material, normally a column
packing, by slowly cooling a polymer solution within the
support material. Thus, polymer chains with fewer
branches deposit on the support first, and polymer chains
with more branches deposit later as the temperature
decreases. To separate the polymer fractions, the polymer
precipitated on the support is stripped off by extraction
with a solvent at increasing temperatures. Therefore,
polymer eluted at lower temperature has a higher frequency of branches and polymer eluted at higher tem-

BRANCH FREQUENCY

Figure 29. Schematic representation of temperature rising elution


fractionation.

perature has a lower branch concentration. One may


envisage polymer chains with different branch concentrations depositing on the support in different layers, as
shown in Figure 29. In a real TREF process, the thickness
of polymer on a support may be very thin, depending on
the surface area of support and amount of polymer used.
Detailed information regarding polymer deposition on the
support has not been reported.
Two types of TREF mode have been developed, namely,
preparative TREF and analytical TREF. In the preparative mode, the polymer sample is fractionated and
subsequently each fraction is analyzed by additional
methods such as GPC and NMR. Thus, relatively large
samples (1-log) are separated (Wild, 1990,1993;Glockner,
1990; Vela Estrader and Hamielec, 1992-93). Typical
preparative TREFs have been designed by Bergstrom and
Avela (1979) and Usami et al. (1986). The main elements
include an oven or a bath with temperature control, a
packed column, and solvent and solution flow systems. A
TREF column is normally packed with glass beads
(Nakano et al., 1981),Chromosorb P (Bergstromand Avela,
1979; Wild et al., 1982; Kulin et al., 1988), or silanated
silica gel (Kelusky et al., 1987). Cooling and heating rates
are in the range 0.5-6.5 C/h. Analytical TREF refers to
fractionation systems in which the separated polymer
eluate is continuously monitored as a function of elution
temperature. This type of operation usually fractionates
only small quantities (<0.5g) of polymer (Wild, 1990,1993).
Typical analytical TREF systems have been developed
by Wild et al. (1982b), Usami et al. (1986), Kelusky et al.
(19871, and Hazlitt (1990). Perhaps the most advanced
analytical TREF developed to date is that by Hazlitt
(19901, shown in Figure 30. The instrumentation used in
analytical TREF is derived directly from the already welldeveloped GPC with necessary modifications of a temperature rise control system (Wild, 1990). Analytical
TREF is a quick method to obtain the short-chainbranching distribution, but the amount of sample used is
too small to measure the MWD of separate fractions by
GPC.
The main results for polyethylene homo and copolymers
using TREF measurement are summarized as follows.
The number of branches in fractions (methyl groups/
100 C) decreases with an increase in elution temperature
(Shirayama et al., 1972; Wild et al., 1982; Usami et al.,
1986; Schouterden et al., 1987; Kelusky et al., 1987;

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 473

,r m

40

40

Figure 30. Schematic diagram for an autoanalytical TREF instrument. Reprinted with permission from Hazlitt (1990). Copyright
1990 John Wiley & Sons.

$70

Figure 32. Bivariate distribution of composition and molecular


weight for LLDPE and LDPE (Wild et al., 1982a; Wild, 1991).
Reprinted with permission from Wild (1991). Copyright 1991
Springer-Verlag.

40
V

-.30

0
0

e,,,,,,
E 20

w7

L L
O

-M, 10X I 0

50

100

Figure 31. Branching frequency dependence on molecular weight


(Schouterden et al., 1987).

Mirabella et al., 1987;Wilfong and Knight, 1990). Branch


frequency is generally a linear function of elution temperature; however, a straight-line relationship is not
universal. Nakano and Goto (1981)showed that polymers
with the same branch concentration, CH3/1000C (carbon),
may not necessarily have the same elution temperature
because the crystallizability of polyethylene depends on
the type of branching. The expression CH3/1000 C
includes all the chain ends (Wilfong, 1990), so the real
branch number or comonomer composition is less than
CH3/1000 C. In fact, comonomer content in polymer
chains is not a linear function of elution temperature
(Kimura et al., 1984).
The branching number of the eluate fraction decreases
with an increase in molecular weight (Mirabella et al.,
1987; Schouterden et al., 1987; Hosoda, 1988; Wilfong et
al., 1990). Figure 31 shows the relationship between the
relative value of branching concentration and molecular
weight (Schouterden et al., 1987). This relationship
suggeststhat the growing chain ends with comonomer units
are more sensitive to chain-transfer and termination
reactions.

IO

0
IO1

IO'

IO'

IO'

HOLECULRR WEJCHT
Figure 33. Contour map of distributionof compositionandmolecular
weight for LLDPE. Reprinted with permission from Hosoda (1988).
Copyright 1988 The Society of Polymer Science, Japan.

Perhaps the most significant finding using TREF is the


shape of polymer composition distribution. Homopolyethylene (LDPE) always shows a unimodal distribution
of CH3/1000C (Wild et al., 1982a,b). Ethylenecopolymers
generally show a bimodal distribution (Wild et al., 1982a,b;
Usami et al., 1986; Cady, 1987; Kelusky et al., 1987) or a
trimodal distribution (mirabella, 1987;Hosoda, 1988).This
evidence strongly suggests that the polymer chains are
formed at different active sites. Figure 32 shows bivariate
distributions of molecular weights and branching for
LLDPE and LDPE (Wild et al., 1982a;Wild, 1990). Figure
33 shows a contour map of a trimodal distribution from
a LLDPE sample (Hosoda, 1988).
Most TREF measurement results published in the
literature are based on commercial polyethylenes. Un-

474 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994


fortunately, the polymer synthesis histories are usually
unknown. Hence, these results cannot provide detailed
information relating polymer microstructure to polymerization conditions. Therefore, systematic studies are
necessary to understand the composition development of
polymers at known polymerization conditions.

of gas phase ethylene polymerization behavior over


commercial reactor operating conditions, off-line characterization of polymer properties, more detailed kinetic/
dynamic modeling of polymerization processes accounting
for commercial reactor operating conditions, and reliable
estimation of model parameters.

6. Summary

7. Literature Cited

In view of the preceding observations, it can be seen


that gas phase ethylene polymerization technology has
been well-established using both fluidized bed reactors
and stirred bed reactors (vertical and horizontal reactors).
Gas phase processes, in contrast with other established
processes, can produce polyethylenes with a broad range
of densities and molecular weights by controlling the
content of comonomerswithout the limitation of the liquid
phase. However, the reactor operating temperature for
gas phase polymerization is limited by the resin softening/
melting point. Because of poor heat removal efficiency in
gas phase polymerization and the use of high-activity
catalyst, preventing polymer particle agglomerationduring
the gas phase polymerization reactor operation is a major
challenge. In commercial practice, particle agglomeration
is avoided by eliminating the possibility of hot-spot
formation, enhancing heat-transfer capacity by introducing
an inert heat-transfer agent, and controlling catalyst
properties through proper catalyst morphology and activity
design or by prepolymerization. In contrast to the
significant industrial activity, relatively little experimental
work in gas phase polymerization has been conducted in
academic laboratories. In general, gas phase polymerization on a laboratory scale is carried out in a stirred bed
reactor in semibatch mode. Reactor mixing and heat
removal are major problems in a laboratory reactor
operation.
Significant progress in reactor modeling for ethylene
polymerization has been made during the past decade,
particularly in understanding molecular weight development and heat and mass transport phenomena. It has
been recognized that the catalyst nature and polymerization mechanisms play important roles in determining
molecular weight development of polyethylene, and diffusion limitations may only play a minor role in controlling
molecular weight distribution. Therefore, the multiple
active sites theory has been adopted in modeling studies
of polymer molecular weight and composition development. Multiple active sites theory can be incorporated
with diffusion limitations when mass-transfer resistance
is operative. However, very few of the models in the
literature have been evaluated using experimental data.
Mechanisms for formation of multiple active centers have
not been fundamentally explained even though the multiple sites theory is supported by significant experimental
evidence, such as the multimodal composition distribution
of LLDPE. Moreover, a comprehensive data base for gas
phase ethylene polymerization has not yet been established
and few kineticJdynamic model parameters have been
estimated. Besides the heterogeneous catalysts, a new
catalyst family, metallocene-based catalysts, may soon be
commercialized for gas phase processes. However, little
work has been published for olefin gas phase polymerizations using metallocene-based catalysts. A general
morphology development model is not yet available, and
more experimental data are needed for better understanding of both catalyst morphology and the polymer
particle morphology development. Therefore, systematic
studies are required for better understanding of gas phase
polymerization processes, including on-line measurement

Ahlers, A.; Kaminsky, W. Variation of Molecular weight Distribution


of Polyethylenes Obtained with Homogeneous Ziegler-Natta
Catalysts. Makromol. Chem. Rapid Commun. 1988,9,457.
Andresen, A.; Cordes, H. G.; Herwig, J.; Kaminsky, W.; Merck, A.;
Mottweiler, R.; Pein, J.; Sinn, H.; Vollmer, H. J. Halogen-Free
Soluble Ziegler Catalysts for the Polymerization of Ethylene.
Control of Molecular weight by Choice of Temperature. Angew.
Chem., Znt. Ed. Engl. 1976,15,630.
Antberg, M.; Dolle, V.; Klein, R.; Rohrmann, J.; Spaleck, W.; Winter,
A. Catalytic Olefin Polymerization; Keii, T., Soga, K., Eds.;
Elsevier: Tokyo, 1990; p 501.
Antberg, M.; Dolle,V.; Haftka, S.;Rohrmann, J.;Spaleck, W.; Winter,
A.; Zimmermann, H. J. Stereospecific Polymerizations with
metallocene catalysts: Products and Technology Aspects. Makromol. Chem., Macromol. Symp. 1991,48149, 333.
Arlman, E. J.; Cossee, P. Ziegler-Natta Catalysis 111. Stereospecific
Polymerization of Propene with the Catalyst System TiCls-AlEt3.
J. Catal. 1964, 3, 99.
ASTM. American Society for Testing Materials 1990Annual Book
of ASTM Standards; Philadelphia 1990;Vol. 8.01, Plastics (1)pp
395, 458.
Bailly, J. C.; Speakman, J. G. Process for the Polymerization of
Ethylene or the Copolymerization of ethylene and a-olefins in a
Fluidized Bed in the Presence of a Chromium Based Catalyst.
Eur. Patent Appl. EP0175532 Al, March 26, 1986.
Bailly, J. C.; Collomb, J. Process for Polymerizing One or More
1-Olefin in the Gas Phase in the Presence of Pulverulent Inorganic
Material. PCT Int. Appl. W08802379, A l , April 7, 1988.
Bailly, J. C.; Collomb, J. Process for Polymerizing One or More
1-Olefin in the Gas Phase in the Presence of Pulverulent Inorganic
Material. US. Patent 4970279, Nov 13, 1990.
Baker, R. T. K.; Harris, P. S.; Waite, R. J.; Roper, A. N. Continuous
Electron Microscopic Observation of the behaviour of ZieglerNatta Catalysts. J.Polym. Sci. Part E,Polym. Lett. 1973,11,45.
Barbe, P. C.; Noristi, L.; Baruzzi, G.; Marchetti, E. Microscopic
Analysis of Polyolefin Initial formation on TiCWMgCl2Supported
Catalysts Makromol. Chem., Rap. Commun. 1983, 4, 249.
Barbe, P. C.; Cecchin, G.; Noristi, L. The Catalytic System TiComplex/MgCl2. Adu. Polym. Sci. 1987,81, l.
Batleman, H. L. Gas-Phase Polymerization-a New Route to HDPE.
Plast. Eng. 1975, April, 73.
Beach, D. L.; Kissin, Y. V. Encylcopedia of Polymer Science and
Engineering; Kroschwitz, J. I. Ed.; John Wiley & Sons: New York,
1986; Vol. 6, p 454.
Berger, M. N.; Grieveson, B. M. Kinetics of the Polymerization of
Ethylene with a Ziegler-Natta Catalyst I. Principal Kinetic
Features. Makromol. Chem. 1965,83, 80.
Bergstrom, C.; Avela, E. Investigation of the Composite Molecular
Structure of LDPE by Using Temperature Rising Elution Fractionation. J. Appl. Polym. Sci. 1979, 23, 163.
Bohm, L. L. Reaction Model for Ziegler-Natta Polymerization.
Polymer 1978a, 19, 545.
Bohm, L. L. Ethylene Polymerization Process with a Highly Active
Ziegler-Natta Catalyst 1. Kinetics. Polymer 197813, 29, 533.
Bohm, L. L. Ethylene Polymerization with a Highly Active ZieglerNatta Catalyst: 2. Molecular weight Regulation. Polymer 1978c,
19, 562.
B o b , L. L.; Franke, R.; Thum, G. Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky,
W., Sinn, H., Eds.; Springer-Verlag: Berlin Heidelberg, 1988; p
391.
Boor, J. Jr. Ziegler-Natta Catalysts and Polymerizations;Academic
Press: New York, 1979.
Bremner, T.;Rudin, A. Melt Flow Index Values and Molecular Weight
Distributions of Commercial Thermoplastics. J. Appl. Polym.
Sci. 1990, 41, 1617.
Breslow, D. S.; Newburg, N. R. Bis(cyclopentadieny1) Titanium
Dichloride Alkylaluminum Complexes as Catalysts for the Polymerization of Ethylene. J. Am. Chem. SOC.1957, 79, 5072.

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 475


Breslow, D. S.; Newburg, N. R. Bis(cyclopentadieny1)Titanium
Dichloride Alkylaluminum Complexes as Soluble Catalysts for
the Polymerization of Ethylene. J.Am. Chem. SOC.
1959,81,81.
Brockmeier, N. F. Encyclopedia of Polymer Scienceand Engineering;
Kroschwitz, J. I., Ed.; John Wiley & Sons: New York, 1987; Vol.
7, p 480.
Brockmeier, N. F. The AMOCO/CHISSOGas Phase Polypropylene
Copolymer Technology. Presented at Society of Plastics Engineers, International Conference, Houston, TX, February 1991.
Brown, G. L.; Warner, D. F.; Byon, J. H. Exothermic Polymerization
in a Vertical Fluid Bed Reactor System Containing CoolingMeans
Therein and Apparatus Therefore. U.S. Patent 4255542, March
10, 1981.
Bukatov, G. D.; Zaikovskii, V. I.; Zakharov, V. A.; Kryukava, G. N.;
Fenelonov, V. B.; Zagrafskaya,R. V. The Morphology of Polypropyelene Granules and its Link with the Titanium Trichloride
Texture. Polym. Sci. U.S.S.R. 1982,24, 599.
Burdett, I. D. The Union Carbide UNIPOL Process: Polymerization
of Olefinsin a Gas-Phase Fluidized Bed. Presented at the AIChE
Annual Meeting, Washington, DC, Nov 27-Dec 2, 1988.
Burdett, I. D. A Continuing Success: The Unipol Process.
CHEMTECH 1992, Oct, 616.
Cady, L. D. LLDPE Properties Tied to Branch Distribution. Plast.
Eng. 1987, 43, 25.
Cana, K. J.; Brady, R. C.; Gau, Y. Process for the Production of
Polyethylene. US. Patent 5102841, April 7, 1992.
Carradine, W. R., Jr.; Rase, H. F. Effect of Surface Characteristics
on Polymer Growth on Titanium Trichloride Catalyst. J. Appl.
Polym. Sci. 1971, 15, 889.
Chem. Eng. 1979,Dec 3,80. New Route to Low-DensityPolyethylene.
Chien, J. C. W. Kinetics of Ethylene Polymerization Catalyzed by
Bis(cyclopentadieny1)Titanium Dichloride Dimethylaluminium
Chloride. J. Am. Chem. SOC.
1959, 81, 86.
Chien, J. C. W. Advances in Polyolefins; Seymeur, R. B., Cheng, T.,
Eds.; Plenum Press: New York, 1987; p 255.
Chien, J. C. W. TransitionMetal CatalyzedPolymerizations;Quirk,
R. P., Hoff, R. E., Klingensmith, G. B., Tait, B. L., Goodall, B. L.,
Eds.; Cambridge University Press: Cambridge, 1988; p 55.
Chien, J. C. W. Catalysis in Polymer Synthesis; Vanderberg, E. J.,
Salamone, J. C., Eds.; American Chemical Society: Washington,
DC, 1992; p 27.
Chien, J. C. W.; Wang, B. P. Metallocene-Methylaluminoxane
Catalysts for Polymerization I. Trimethylaluminum as Cocativator. J. Polym. Sci. A: Polym. Chem. 1988,26, 3089.
Chien,J. C. W.; Razavi,A. Mettallocene-MethylaluminoxaneCatalyst
for Olefin Polymerization 11. bis-q64Neomenthylcyclopentaldienyl) Zirconium Dichloride. J.Polym. Sci. A: Polym. Chem. 1988,
26, 2369.
Chien, J. C. W.; Wang, B. P. Mettallocene-Methylaluminoxane
catalysts for Olefin PolymerizationIV. Active Site Determinations
and Limitation of W O Radiolabeling Technique. J.Polym. Sci.
A: Polym. Chem. 1989,27,1539.
Chien, J. C. W.; Wang, B. P. Mettallocene-Methylaluminoxane
Catalysts for Polymerization V. Comparison of CpzZrClz and
CpZrCls. J. Polym. Sci. A: Polym. Chem. 1990,28, 15.
Chien, J. C. W.; He, D. Olefin Copolymerization with Metallocene
Catalysts I. Comparison of Catalysts. J. Polym. Sci. A: Polym.
Chem. 1991a, 29, 1585.
Chien, J. C. W.; He, D. Olefin Copolymerization with Metallocene
Catalysts 11. Kinetics, Cocatalysts, and Addivities. J. Polym.
Sci. A: Polym. Chem. 1991b, 29, 1595.
Chien, J. C. W.; He, D. Olefin Copolymerization with Metallocene
Catalysts 111. Supported Metallocene/Methylaluminoxane Catalyst for Olefin Copolymerization. J.Polym. Sci. A: Polym. Chem.
1991c, 29, 1603.
Chien, J. C. W.; He, D. Olefin Copolymerization with Metallocene
Catalysts IV. Metallocene/MethylaluminoxaneCatalyzed Olefin
Terpolymerization. J. Polym. Sci. A: Polym. Chem. 1991d, 29,
1609.
Chien, J. C. W.; Sugimoto, R. Kinetics and Stereochemical Control
of Propylene Polymerization Initiated by Ethylene Bis(4,5,6,7Tetrahydro-1-indenyl)Zirconium Dichloride/Methylaluminoxane
Catalyst. J. Polym. Sci. A: Polym. Chem. 1991,29, 459.
Chien, J. C. W.; Nozaki, T. Ethylene-Hexene Copolymerization by
Heterogeneousand HomogeneousZiegler-Natta Catalysts ans the
Comonomer effect. J.Polym. Sci. A: Polym. Chem. 1993,31,227.
Chien, J. C. W.; Gong, B. M. Hexene-1 Polymerization by Homogeneous Zirconocene and Heterogeneous Supported Tic13 Catalysts. J. Polym. Sci. A: Polym. Chem. 1993,31, 1747.

Chien, J. C. W.; Wu, J. C.; Kuo, C. I. Magnesium Chloride Supported


High-Mileage Catalysts for Olefin Polymerization V. BET,
Prosimetry and X-Ray Diffraction Studies. J.Polym. Sci. Polym.
Chem. Ed. 1983,21,737.
Chien, J. C. W.; Rieger, B.; Sugimoto, R.; Mallin, D. T.; Rausch, M.
D. Catalytic Olefin Polymerization; Keii, T., Soga, K., Eds.;
Elsevier: Tokyo, 1990, p 535.
Chien, J. C. W.; Llinas, G. H.; Rusch, M. D.; Lin, L. G.; Winter, H.
H.; Atwood, J. L.; Bott, S. G. Metallocene Catalysts for Olefin
PolymerizationXXIV. A Two-State Propagations. J.Polym. Sci.
A: Polym. Chem. 1992, 30, 2601.
Chinh, J. C.; Dumain, A. Process and Apparatus for the Gas Phase
Polymerization of Olefinsin a Fluidised Bed Reactor. Eur. Patent
Appl. EP351068 Al, Jan 17, 1990.
Choi, K. Y.; Ray, W. H. Recent Developments in Transition Metal
Catalyzed Olefin Polymerization-A survey I. Ethylene Polymerization. J.Macromol. Sci. Rev. Macromol. Chem. Phys. 1985a,
C25, 1.
Choi, K. Y.; Ray, W. H. Polymerization of Olefins Through
Heterogeneous Catalysts 11. Kinetics of Gas Phase Propylene
Polymerizationwith Ziegler-Natta Catalysts. J.Appl. Polym. Sci.
1985b,30, 1065.
Choi, K. Y.; Ray, W. H. The Dynamic behaviour of Fluidized Bed
Reactors for Solid Catalyzed Gas Phase Olefin Polymerization.
Chem. Eng. Sci. 1985c, 40, 2261.
Choi,K.Y.;Ray, W. H.TheDynamicBehaviourof ContinuousStirred
Bed Reactors for Solid Catalyzed Gas Phase Polymerization of
Propylene. Chem. Eng. Sci. 1988,43, 2587.
Choi, K. Y.; Taylor,T. W.;Ray, W. H. PolymerReaction Engineering;
Reichert, K. H., Geiseler, W., Eds.; Hanser Publisher: Munich,
1983; p 313.
Cihlar, J.; Mejzlik, J.; Hamrik, 0. Influence of Water on Ethylene
Polymerization Catalyzed by Titanocene Systems. Makromol.
Chem. Rapid Commun. 1978,179,2553.
Collomb-Ceccarini,J.; Crouzet, P. Process for the Polymerization or
Copolymerizationof a-Olefins in a Fluidised Bed in the presence
of a Ziegler-Natta Catalyst System. PCT Int. Appl. W08600314
A l , Jan 16, 1986.
Cossee, P. Ziegler-Natta Catalysts 1. Mechanism of Polymerization
of a-Olefins with Ziegler-Natta Catalysts. J. Catal. 1964, 3, 80.
Crabtree, J. R.; Grimsby, F. N.; Nummelin, A. J.; Sketchley, J. M.
The Role of Diffusion in the Ziegler Polymerization of Ethylene.
J. Appl. Polym. Sci. 1973, 17, 959.
Dall'Occo, T.; Zucchini, U.; Cuffiani, I. Transition Metals and
Organometallics as Catalysts for Olefin Polymerizations; Kaminsky, W., Sinn, H., Eds.; Springer-Verlag: Berlin Heidelberg,
1988; p 209.
Davis, J. C. LDPE Goes Low Pressure. Chem. Eng. 1978, Jan 2,25.
de Carvalho, A. B. M.; Gloor, P. E.; Hamielec, A. E. A Kinetic
Mathematical Model for Heterogeneous Ziegler-Natta Copolymerization. Polymer 1989, 30, 280.
de Carvalho, A. B. M.; Gloor, P. E.; Hamielec, A. E. A Kinetic Model
for Heterogeneous Ziegler-Natta Copolymerization. Part 2: Stereochemical Sequences Length Distribution. Polymer 1990, 31,
1294.
Delgrange, J. P. B.P. Chemie Gas Phase Fluid-Bed Process for Linear
Polyethylene. Chem. Age India 1987, 38, 129.
Doak, K. W. Encyclopedia of Polymer Science and Engineering;
Kroschwitz, J. I., Ed.; John Wiley & Sons: New York, 1986; Vol.
6, p 386.
Doi, Y.; Murata, M.; Yano, K. Gas Phase Polymerization of Propene
with the Supported Ziegler Catalyst: TiCWMgCld/CeH&OOCzHd
Al(CzH& Ind. Eng. Chem. Prod. Res. Dev. 1982, 21, 580.
Dormenval,R.; Havaa, L.; Mangin, P. Process for Dry Polymerization
of Olefins. U S . Patent 3922322, Nov 25, 1975.
Drogemuller, H.; Heiland, K.; Kaminsky, W. Transition Metals and
Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.; Springer-Verlag: Berlin, 1988; p 303.
Dumain, A.; Raufast, C. Device and Process for Introducing a Powder
with Catalytic Activity into a Fluidized Bed Polymerization
Reactor. Eur. Patent Appl. EP0157584 A3, Oct 9, 1985.
Durand, D. C.; Morterol, F. R. M. M.; Sandis, S. Process for the
Starting up the Polymerization of Ethylene or Copolymerization
of Ethylene and at Least one other a-Olefin in the Gas Phase in
the presence of a Catalyst Based on Chromium Oxide. Eur. Patent
Appl. EP179666 A l , April 30,1986a.
Durand, D. C.; Morterol, F. R. M. M.; Sandis, S.Process for the Start
up of Polymerization or Copolymerization in the Gas Phase of
a-Olefins in the presence of a Ziegler-Natta Catalyst System. Eur.
Patent Appl. EP0180420, May 7, 1986b.

476 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994


Dusseault, J. J. A. The Gas Phase Polymerization Kinetics of Ethylene
over MgClz-Supported Ziegler-Natta Catalyst. Ph.D. Thesis,
Queen's University, 1991.
Dusseault, J. J. A,; Hsu, C. C. MgC12 Supported Ziegler-Natta
Catalysts for Olefin Polymerization: Basic Structure, Mechanism
and Kinetic Behaviour. J.Macromol. Sci. Rev., Macromol. Chem.
Phys. 1993, C33, 103.
Dye, R. F. Polymerization Process, US. Patent 3023203, February
27, 1962.
Edgecombe, F. H. C. The Vapor Phase Polymerization of Ethylene
on an Organotitanium Catalytic Surface. Can. J. Chem. 1963,41,
1265.
Ewen, J. A. Mechanisms of Stereochemical Control in Propylene
Polymerizations with Soluble Group 4B Metallocene/Methylaluminoxane Catalysts. J. Am. Chem. SOC.
1984, 106,6355.
Ewen, J. A.; Haspeslahg, L.; Elder, M. J.; Atwood, J. L.; Zhang, H.;
Cheng, H. N. Transition Metals and Organometallicsas Catalysts
for Olefin Polymerizations; Kaminsky, W., Sinn, H., Eds.;
Springer-Verlag: Berlin, 1988a; p 281.
Ewen, J. A.; Jones, R. L.; Razavi, A.; Ferrara, J. D. Syndiospecific
Propylene Polymerizations with Group 4B Metallocenes, J. Am.
Chem. SOC.
198813, 110,6255.
Ewen, J. A,; Elder, M. J.; Jones, R. L.; Curtis, S.; Cheng, H. N.
Catalytic Olefin Polymerization; Keii, T., Soga, K., Eds.; Elsevier: Tbkyo, 1990; p 439.
Finogenova, L. T.; Zakharov, V. A.; Buniyat-Zade, A. A.; Bukatov,
G. D.: Plaksunov. T. K. Study of Copolymerization of Ethylene
with Hexene-1 on Applied Catalysts.- Polym. Sci. U.S.S.R. 1980,
22, 448.
Floyd, S.; Choi, K. Y.; Taylor, T. W.; Ray, W. H. Polymerization of
Olefins through Heterogeneous Catalysts 111. Polymer Particle
Modelling with an Analysis of Intraparticle Heat and Mass
Transfer Effects. J. Appl. Polym. Sci. 1986a, 32, 2935.
Floyd, S.; Hutchinson, R. A.; Ray, W. H. Polymerization of Olefins
through Heterogeneous Catalysis V. Gas-Liquid Mass Transfer
Limitations in Liquid Slurry Reactors. J.Appl. Polym. Sci. 1986b,
32, 5451.
Floyd, S.; Choi, K. Y.; Taylor, T. W.; Ray, W. H. Polymerization of
Olefins through Heterogeneous Catalysis IV. Modelling of Heat
and Mass Transfer Resistance in the Polymer Particle Boundary
Layer. J. Appl. Polym. Sci. 1986c, 31, 2231.
Floyd, S.; Heiskanen, T.; Taylor, T. W.; Mann, G. E.; Ray, W. H.
Polymerization of Olefins through Heterogeneous Catalysis VI.
Effect of Particle Heat and Mass Transfer on Polymerization
Behaviour and Polymer Properties. J. Appl. Polym. Sci. 1987,
33, 1021.
Floyd, S.; Heiskanen, T.; Ray, W. H. Solid Catalyzed Olefin
Polymerization. Chem. Eng. Prog. 1988, Nov, 56.
Foster, G. Polymer Reaction Engineering Course; McMaster
University: Hamilton, ON, 1991.
Fries, R. W.; Bowles, W. A. Organometallic Modified Polyolefin
Catalysts for Enhanced Molecular Properties. Metcon'93, Catalyst
Consultants Inc.: Houston, TX, 1993.
Furtek, A. B. Ultra Stength Polyethylene Resins Produced in a FluidBed Process Utilizing Metallocene-Based Catalysts. Metcon'93,
Catalyst Consultants Inc.: Houston, TX, 1993.
Galli, P.; Luciani, L.; Cechin, G. Advances in the Polymerization of
Polyolefins with Coordination Catalysts. Angew. Makromol.
Chem. 1981,94, 63.
Galli, P.; Barbe, P. C.; Noristi, L. High Yield Catalysts in Olefin
Polymerization. Angew. Makromol. Chem. 1984, 120, 73.
Galvan, R.; Tirrel, M. Molecular Weight Distribution Predictions
for Heterogeneous Ziegler-Natta Polymerization Using a TwoSite Model. Chem. Eng. Sci. 1986, 41, 2385.
Glockner, G. Temperature Rising Elution Fractionation: A Review.
J. Appl. Chem. Sci. Appl. Polym. Symp. 1990,45, 1.
Goins, R. R. Olefin Polymerization. U.S. Patent 2936303, May 10,
1960.
Graff, R. J. L.; Kortleve, G.; Vonk, C. G. On the Size of the Primary
Particles in Zieeler
- Catalvsts. J.Polym. Sci. Part B, Polym. Lett.
1970, 8, 735.
Grunie. H.: Luft. G. Polymer Reaction Engineering: Reichert, K. H.,
GeGeler; W., 'Eds.; Huthig & Wepf Virlag: Basel, Heidelberg,
1986; p 293.
Hamielec, A. E. A Brief Introduction to Some Difficult Polymer
Analysis and Characterization Problems-Heterogeneous Branched
and Crosslinked Copolymers. J. Appl. Polym. Sci. Appl. Polym.
Symp. 1989, 43, 1.
Havas, L. J.; Mangin, P. M. Method of Producing Solid Polymers.
US. Patent 3954909, May 4, 1976.

Havas, L.; Lalanne-Magne, C. Gas Phase Polymerization Process.


Eur. Patent Appl. EP475603 Al, March 18, 1992.
Hawkins, S. W.; Smith, H. The Fractionation of Polyethylene. J.
Polym. Sci. 1958, 28, 341.
Hazlitt, L. G. Determination of Short-Chain Branching Distribution
of Ethylene Copolymers by Automated Analytical Temperature
Rising Elution Fractionation. J. Appl.
.. Polym. Sci. Appl.
~. Polym.
Symi. 1990, 45, 25.
Heiland. K.: Kaminskv. W. ComDarison of Zirconocene and Hafnocene
C a d y s k for the" Polymerization of Ethylene and 1-Butene.
Makromol. Chem. 1992,193,601.
Herwig, J.; Kaminsky, W. Halogen-Free Soluble Ziegler Catalysts
with Methylaluminoxane as Catalyst. Polym. Bull. 1983,9,464.
Hock, C. W. How Tic13 Catalysts Control the Texture of asPolymerized Polypropylene. J.Polym. Sci. Part A-1 1966,4,3055.
Hoseman, R.; Hentchel, M.; Feracini, E.; Ferrero, A.; Martelli, S.;
Riva, F.; Antisari, M. V. Globules of Microparacrystals in Nascent
Isotactic Polypropylene. Polymer 1982, 23,979.
Hosoda, S. Structural Distribution of Linear Low-Density Polyethylenes. Polym. J. 1988, 20, 383.
Hsieh, H. L. Recent Development on Transition-Metal Catalysts for
Olefin Polymerizations. Polym. J . 1980, 12, 597.
Hsieh, H. L. Olefin Polymerization Catalysis Technology. Catal.
Reu.-Sci. Eng. 1984, 26, 631.
Hsieh, H. L.; McDaniel; Martin, J. L.; Smith, D. R.; Fahey, D. R.
Advances in Polyolefins; Seymour, R. B., Cheng, T., Eds.; Plenum
Press: New York, 1987; p 153.
Hutchinson, R. A.; Ray, W. H. Polymerization of Olefins Through
Heterogeneous Catalysis VII. Particle Ignition and Extinction
Phenomena. J. Appl. Polym. Sci. 1987, 34, 657.
Hutchinson, R. A,; Ray, W. H. Polymerization of Olefins Through
Heterogeneous Catalysis VIII. Monomer Sorption Effects. J.
Appl. Polym. Sci. 1990,41, 51.
Hutchinson, R. A.; Ray, W. H. Polymerization of Olefins Through
Heterogeneous Catalysis IX. Experimental Study of Propylene
Polymerization over a High Activity MgClz-Supported Catalyst.
J. Appl. Polym. Sci. 1991a, 43, 1271.
Hutchinson, R. A.; Ray, W. H. Polymerization of Olefins Through
Heterogeneous Catalysis-The Effect of Condensation Cooling
on Particle Ignition. J. Appl. Polym. Sci. 1991b, 43, 1387.
Hutchinson, R. A.; Chen, C. M.; Ray, W. H. Polymerizationof Olefins
Through Heterogeneous Catalysis X. Modellingof Particle Growth
and Morphology. J. Appl. Polym. Sci. 1992,44, 1389.
Imhausen, H. C. K. H.; Hippenstiel-Im-Hausen, J.; Newman, Ph.;
Berndt, R.; Schoffel, F.; Zink, J. A Comparison of a Modern High
pressure Tubular Low-Density Polyethylene Process with Gas
Phase Linear Low Density Polyethylene Technology. J. Appl.
Polym. Sci. Appl. Polym. Symp. 1981, 36, 1.
James, D. E. Encyclopedia of Polymer Science and Engineering;
Kroschwitz, J. I., Ed.; John Wiley & Sons: New York, 1986; Vol.
6, p 429.
Jejelowo, M. 0.;
Lynch, N. BU. D. T.; Wanke, S. E. Catalytic Olefin
Polymerization; Keii, T., Soga, K., Eds.; Kodansha: Tokyo, 1990;
p 39.
Jezl, J. L.; Peters, E. F. Horizontal Reactor for the Gas Phase
Polymerization of Monomers. US. Patent 4101289, July 18,1978a.
Jezl, J. L.; Peters, E. F. Horizontal Reactor for the Gas Phase
Polymerization of Monomers. U.S. Patent 4129701, December
12, 1978b.
Jezl, J. L.; Peters, E. F.; Hall, R. D.; Shepard, J. W. Process for the
Vapour Polymerization of Monomers in a Horizontal, QuenchCooled, Stirred-Bed Reactor Using Essentially Total off-Gas
Recycle and Melt Finishing. US. Patent 3965083, June 22,1976.
Jorgensen, R. J.; Goeke, G. L.; Karol, F. J. Catalyst Composition for
Copolymerization Ethylene. US. Patent 4349648, September 14,
1982.
Kakugo, M.; Sadatoshi, H.; Yokoyama, M.; Kojima, K. Transition
Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.; Springer-Verlag: Berlin
Heidelberg, 1988; p 433.
Kakugo, M.; Sadatoshi, H.; Yokoyama, M.; Kojima, K. Transmition
Electron Mocroscopic Observation of Nascent Polypropylene
Particles Using a New Staining Method. Macromolecules 1989a,
22, 547.
Kakugo, M.; Sadatoshi, H., Sakai, J.; Yokoyama, M. Growth of
Polypropylene Particles in Heterogeneous Ziegler-Natta Polymerization. Macromolecules 1989b, 22, 3172.
Kaminaka, M.; Soga, K. Polymerization of Propene with Catalyst
Systems Composed of A1203or MgClz Supported Zirconocene and
Al(CH&. Polymer 1992, 33, 1105.

Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994 477


Kaminsky, W. Transition Metal Catalyzed Polymerizations; Quirk,
R. P. Ed.; Harwood Academic Publishers: New York, 1983;Vol.
4,Part A, p 225.
Kaminsky, W. Stereoselective Polymerization of Olefin using Homogeneous Chiral Ziegler-Natta Catalysts. Angew. Makromol.
Chem. 1986,145/146,149.
Kaminsky, W.; Luker, H. Influence of Hydrogen on the Polymerization of Ethylene with the Homogeneous Ziegler System Bis(Cyclopentadienyl) Zirconium Dichloride/Aluminoxane. Makromol. Chem. Rapid Commun. 1984,5,225.
Kaminsky, W.; Hahnsen, H. Advances in Polyolefins; Segmour, R.
B., Cheng, T., Eds.; Plenum Press: New York, 1987;p 361.
Kaminsky, W.; Sinn, H. Transition Metals and Organometallics as
Catalysts for Olefin Polymerization; Springer-Verlag: Berlin
Heidelberg, 1988.
Kaminsky, W.; Miri, M.; Sinn, H.; Woldt, R. Bis(cyclopentadieny1)
Zirconium Dichloride and Aluminoxane as Catalyst for Olefin
Copolymerization. Makromol. Chem. Rapid Commun. 1983,4,
417.
Kaminsky, W.; Kulper, K.; Brintzinger, H. H.; Wild, F. R. W. P.
Polymerization of Propene and Butene with a Chiral Zirconocene
and Methylaluminoxane as Cocatalyst. Angew. Chem., Int. Ed.
Engl. 1985,24,507.
Kaminsky, W.; Kulper, K.; Niedoba, S. Olefin Polymerization with
High Active Soluble Zirconium Compounds Using Aluminoxane
as Cocatalyst. Makromol. Chem. Macromol. Symp. 1986,3,377.
Kaminsky, W.; Bark, A.; Spiehl, R.; Moller-Lindenhol, N.; Niedoba,
S. Transition Metals and Organometallics as Catalysts for Olefin
Polymerization; Kaminsky, W., Sinn, H., Eds.; Springer-Verlag:
Berlin, 1988;p 291.
Kaminsky, W.; Bark, A.; Arndt, M. New Polymers by Homogeneous
Zirconocene/AluminoxaneCatalysts. Makromol. Chem. Macromol. Symp. 1991,47,83.
Karol, F. J. The Polyethylene Revolution. CHEMTECH 1983,April,
222.
Karol, F. J. Studies with High Activity Catalysts for Olefin Polymerization. Catal. Rev.-Sci. Eng. 1984,26, 557.
Karol, F. J.;Wu, C. S. Ethylene Polymerization with Silane Modified
Catalyst. US. Patent 4086408,April 25, 1978.
Karol, F. J.; Wagner, B. E.; Levine, I. J.; Goeke, G. L.; Noshay, A.
Advances in Polyolefins; Seymour,R. B., Cheng, T., Eds.; Plenum
Press: New York, 1987;p 337.
Karol, F. L.; Cann, K. J.; Wagner, B. E. Transition Metals and
Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.; Springer-Verlag: Berlin Heidelberg, 1988;
p 149.
Keii,T. Kinetics of Ziegler-Natta Polymerization;Kodansha: Tokyo,
1972.
Keii,T.; Soga, K. Catalytic OlfeinPolymerization; Kodansha: Tokyo,
1990.
Kelusky, E. C.; Elston, C. T.; Murray, R. E. Charactering Polyethylene-Based Blends with Temperature Elution Frationation
(TREF) Techniques. Polym. Eng. Sci. 1987,27, 1562.
Kim, 11; Woo, S. I. MorphologicalStudy of HDPE Prepared with the
Highly Active Silica Supported TiCUMgC12 Catalyst. Polym. J.
1989,21,697.
Kimura, K.; Shigemura, T.; Yuasa, S. Characterization of Ethylene1-Butene Copolymer by Differential Scanning Calorimetry and
W N M R Spectroscopy. J. Appl. Polym. Sci. 1984,29,3161.
Kioka, M.; Tsutsui, T.; Ueda, T.; Kashiwa, N. Catalytic Olefin
Polymerization; Keii, T., Soga, K., Eds.; Elsevier: Tokyo, 1990;
p 483.
Kioka, M.; Mizuno,A.; Tsutsui, T.; Kashiwa,N. Catalysis in Polymer
Synthesis; Vandenberg, E. J., Salamone, J. C., Eds.; American
Chemical Society: Washington, DC, 1992;p 74.
Kissin, Y. V. Isospecific Polymerization of Olefins with Heterogeneous Ziegler-Natta Catalysis;Springer-Verlag: New York, 1985.
Knobeloch,D. C.; Wild, L. Polyolefins IV: Innovations inProcesses,
Products, Processing and Additives; S P E Westchase, 1984;p
427.
Koivumaki,J.; Seppala, J. V. Observationson the Rate Enhancement
Effect with MgCl*/TiC4 and CpzZrClz Catalyst Systems upon
1-Hexene Addition. Macromolecules 1993,26,5535.
Krieger, J. AMOCO to Launch New Technology for Ethylene/
Propylene Polymerization. Chem. Eng. News 1992,March 30,17.
Kryzhanovskii, A. V.; Pranchev, S. S. Synthesisof Linear Polyethylene
on SuDported Ziegler-Natta catalysis. Review. Polym. sci.
u.s.s.~~.
i990,32,1312.
Kuber, F. Dreams of the Perfect Plastic. New Sci. 1993,Aug 14,28.

Kulin, L. I.; Meijerink, N. L.; Stack, P. Long and Short Chain


Branching Frequency in Low Density Polyethylene (LDPE). Pure
Appl. Chem. 1988,60,1403.
Laurence, R. L.; Chiovetta, M. G. Polymer Reaction Engineering;
Reichert, K. H., Geiseler, W., Eds.; Hanser Publishers: Munich,
1983;p 73.
Levine, I. J.; Karol, F. J. Preparation of Low and Medium Density
Ethylene Polymer in Fluid Bed Reactor. U S . Patent 4011382,
March 8, 1977.
Lindsav. K. F. Dow Plastics Unveils Sinde-Site Catalvst Resins for
Paciaging. Mod. Plast. 1993,Oct, 85.
Lorenzini, P.; Bertrand, P.; Vilermaux, J. Modelling Ethylene and
a-Olefin CopolymerizationUsing Ziegler-Natta Catalyst. Can. J.
Chem. Eng. 1991,69,682.
Lynch, T.; Wanke, S. E. Reactor Desogn and Operation for Gas Phase
Ethylene Polymerization Using Ziegler-Natta Catalyst. Can J.
Chem. Eng. 1991,69,332.
Lipman, R. D. A.; Norrish, R. G. M. Kinetic Studies with a Ziegler
Catalyst System in the Gaseous Phase. Proc. R. SOC.(London) A
1963,275,310.
Mabilon, G.; Spitz, R. Ethylene and Propylene Copolymerization
Using Chromium Oxide Catalyst. Eur. Polym. J. 1985,21,245.
Mackie, P.; Berger, M. N.; Lawson, D.; Grieveson,B. M. Replication
in Ziegler Polymerization. Polym. Sci. Part B, Polym. Lett. 1967,
5,493.
Mallin, P. T.; Rausch, M. D.; Chien, J. C. W. A Comparison of Cp2HfClz Versus CpzZrCls Methylaluminoxane Catalysts. Polym.
Bull. 1988,20,421.
Margolies,A. F. Effect of Molecular Weight on thePhysicalProperties
of Linear, High-DensityPolyethylene Homopolymers. Tech. Pap.
Reg. Tech. Conf.,SOC.Plast. Eng., Baltimore-Washington Sect.,
1970.
Martens, A.; Morterol,F. R. M. M. Gas Phase PolymerizationProcess.
Eur. Patent Appl. EP0476835 A l , March 25, 1992.
Martin, J. R.; Johnson, J. F.; Cooper, A. R. Mechanical Properties
of Polymers: The Influence of Molecular Weight and Molecular
Weight Distribution. J. Macromol. Sci. Rev. Macromol. Chem.
1972,C8, 57.
McAuley, K. B.; MacGregor, J. F. On-line Inference of Polymer
Properties in an Industrial Polyethylene Reactor. AIChE J. 1991,
37,825.
McAuley, K. B.; MacGregor,J. F. Optimal Grade Transition in a Gas
Phase Polymerization Reactor. AIChE J. 1992,38, 1564.
McAuley, K. B.; MacGregor,J.F. Nonlinear Product Property Control
in Industrial Gas Phase Polyethylene Reactors. AIChE J. 1993,
39,855.
McAuley, K. B.; MacGregor, J. F.; Hamielec, A. E. A Kinetic Model
for Industrial Gas-Phase Ethylene Copolymerization. AIChE J.
1990,36, 837.
McAuley, K. B.; Talbot, J. P.; Harris, T. J. A Comparison of TwoPhase and Well-Mixed Models for Fluidized Bed Polyethylene
Reactors. Submitted to Chem. Eng. Sci. 1993.
McDaniel, M. P. Supported Chromium Catalysts for Ethylene
Polymerization. Adv. Catal. 1985,33,47.
McDaniel, M. P. Controlling Polymer Properties with the Phillips
Chromium Catalysts. Ind. Eng. Chem. Res. 1988,27,1559.
Miller, A. R. Fluidized Bed Reactor. U.S. Patent 4003712,Jan 18,
1977.
Mirabella, F. M., Jr. Microstructural Characterization of ImpactResistant Polypropylene Copolymers. J.Appl. Polym. Sci. Appl.
Polym. Symp. 1992,51,117.
Mirabella, F. M., Jr.; Ford, E. A. Characterization of Linear Low
Density Polyethylene Cross-Fraction According to Copolymer
Composition and Molecular Weight. J. Polym. Sci. B: Polym.
Phys. 1987,25,777.
Mkrtchyan, S. A.; Uvarov, B. A.; Tsvetkova, V. I.; Tovmasyan, Yu.
M.; Chistyakov, S. 0.;Rachinskii, G. F.; Dyachkovskii, F. B.
Formation and Growth of Polypropylene and Polyethylene
Particles During the Polymerization of Olefins on Deposited
Catalysts. Polym. Sci. U.S.S.R.1986,28,2343.
Mod. Plast. 1991,July, 61. Exxon Cites Breakthrough in Olefins
Polymerization.
Mod. Plast. 1993, Sept, 17. Dow Sees Broad Applications for
Performance Polyolefins.
Montagna, A.; Floyd, J. C. Exxpolm Single-Site Catalysts: Leading
the Revolution to New Polyolefin Products and Processes.
Metcon9S; Catalyst Consultants Inc.: Houston, TX; 1993.
Munoz-Escalona. A.: Hernandez, J. G.; Gallardo, J. A.: Sustic, A.
Advances in Polyolefins; Seymour, R; B., Cheng, T., Eds.; Plenum
Press: New York, 1987;p 179.

478 Ind. Eng. Chem. Res., Vol. 33, No. 3, 1994


Munoz-Escalona, A.; Alarcon, C.; Albornoz, L. A.; Fuentes, A.;
Sequera, J. A. Transition Metals and Organometallics as Catalysts
for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.;
Springer-Verlag: Berlin Heidelberg, 1988; p 417.
Nagel, E. J.; Kirillov, V. A.; Ray, W. H. Prediction of Molecular
Weight Distribution for High-Density Polyolefins. Ind. Eng.
Chem. Prod. Res. Dev. 1980,19,372.
Nakano, S.; Goto, Y. Development of Automatic Cross Fractionation: Combination of Crystallizability Fractionation and Molecular Weight Fractionation. J.Appl. Polym. Sci. 1981,26,4217.
Natta, G.; Pino, P. A. Crystallizable Organometallic Complex
Containing Titanium and Aluminum. J . Am. Chem. SOC.1957,
79, 2975.
Nicoletti, J. W.; Cann, K. J.; Karol, F. J. Process for the Production
of Polyethylene Having an Intermediate Molecular Weight Distribution. Eur. Patent Appl. EP0286001 Al, Oct 12,1988.
Nowlin, T. E. Low Pressure Manufacture of Polyethylene. Prog.
Polym. Sci. 1985,11, 29.
Nunes, R. W.; Martin, J. R.; Johnson, J. F. Influence of Molecular
Weight and Molecular Weight Distribution on Mechanical Properties of Polymers. Polym. Eng. Sci. 1982,22,205.
Pang, S.;Rudin, A. Characterization of Polyolefins by Size Exclusion
Chromatography withlow-Angle Light Scattering and Continuous
Viscometer Detector. Polymer 1992,33,1949.
Payn, C. F. Perspectives on the Commercial Direction of Metallocene
Technology to 2000. Metcon93; Catalyst Consultants Inc.:
Houston, TX; 1993.
Peters, E. F.; Spangler, M. J.; Michaels, G. 0.; Jezl, J. L. Vapour
Phase Reactor off-Gas Recycle System for Use in the Vapour State
Polymerization of Monomers. U S . Patent 3971768,July 27,1976.
Quirk, R. P.; Hoff, R. E.; Klingensmith, P. J.; Tait, T.; Goodall, B.
L. Transition Metal Catalyzed Polymerization; Cambridge University Press: Cambridge, 1988.
Ramanathan, S.; Ray, W. H. A Study of the Dynamic Behaviour of
Polymerization Process Flowsheets Using the CAD Package,
POLYRED. Engineering Foundation Conference on Polymer
Reaction Engineering, Santa Barbara, CA, 1991.
Rasmussen, D. M. High Density Polyethylene Polymerized in Gas
Phase. Chem. Eng. 1972,Sept. 18,104.
Raufast, C. Fluidiaed Bed Discharge Device. Eur. Patent EP0250169
A2, Dec 23,1987.
Ray, W. H. Modelling of Polymerization Phenomena. Ber. BunsenGes. Phys. Chem. 1986,90,947.
Ray, W. H. Transition Metal Catalyzed Polymerizations-ZieglerNatta and Metathesis Polymerizations; Quirk, R. P., Ed.;
Cambridge University Press: Cambridge, 1988;p 417.
Ray, W. H. Modelling of Addition Polymerization Processes-Free
Radical, Ionic, Group Transfer, and Ziegler-Natta Kinetics. Can.
J. Chem. Eng. 1991,69,626.
Redman, J. Polyethylene. Chem. Eng. 1991,No. 504,26.
Reichert, K. H. TransitionMetal Catalyzed Polymerizations;Quirk,
R. P., Ed.; Harwood Academic Publishers: New York, 1983;Vol.
4,Part B, p 465.
Reichert, K. H.; Meyer, K. R. Ethylene Polymerization Using Soluble
Ziegler Catalysts. Makromol. Chem. 1973,169,163.
Rhee, S.J.; Baker, E. C.; Edwards, D. N.; Lee, K. H.; Moorhouse, J.
H.; Scarola, L. S.; Karol, F. J. Process for Producing Sticky
Polymers. US. Patent 4994534,Feb 19,1991.
Rogers, J. K. Exxon: Full Speed Ahead on Gas Phase Metallocene
PE. Mod. Plast. 19938, 15.
Rogers, J. K. Exxon Expands Line of Metallocene Products. Mod.
Plast. 1993b,11.
Rotman, D.; Wood, A. Dow and Exxon Jostle in Race to Bring
Metallocene Products to Market. Chem. Week 1993,Sept 15.
Sailors, H. R.; Hogan, J. P. History of Polyolefins. J. Macromol.
Sci.-Chem. 1981,A15, 1377.
Schmid, K.; Stedefeder, J.; John, G.; Haeberte, M.; Lautenschlager,
H.; Trieschmann, H. G. Process and Apparatus for the Polymerization of Monoolefins. U.S. Patent 3300457,Jan 24,1967.
Schouterden, P.; Groeninckx, G.; Van der Heijden, B.; Jansen, F.
Polymer 1987,28,2099.
Schwank, D. Single-Site Metallocene Catalysts Yield Tailor-Made
Polyolefin Resins. Mod. Plast. 1993,Aug, 49.
Seymour, R. B.; Cheng, T. Advances in Polyolefins; Plenum Press:
New York, 1987.
Shepard, J. W.; Jezl, J. L.; Peters, E. F.; Hall, R. D. Divided Horizontal
Reactor for the Vapor Phase Polymerization of Monomers a t
Different Hydrogen Levels. US.Patent 3957448,May 18,1976.

Shirayama, K.; Okada, T.; Kita, S. Distribution of Short Chain


Branching in Low Density Polyethylene. J. Polym. Sci. Part A ,
1965,3,907.
Shirayama, K.; Kita, S.; Watabe, H. Effects of Branching on Some
Properties of Ethylene/a-Olefin Copolymers. Makromol. Chem.
1972,151,97.
Short, J. N. Transition Metal Catalyzed Polymerizations Alkenes
and Dienes; Quirk, R. P., Ed.; Harwood Academic Publishers:
New York, 1981;p 651.
Sinclair, K. B. Characteristics of Linear LDPE and Description of
UCC Gas Phase Process. Process Economics Report; SRI International, Menlo Park, CA, 1983.
Sinclair, K. B. Grade Change Flexibility-Defined, Determined,
Compared. Proceedings of the Fifth International SPE Conference; 1987.
Sinn, H.; Kaminsky, W. Ziegler-Natta Catalysis. Adv. Organomet.
Chem. 1980,18,99.
Sinn, H.; Kaminsky, W.; Vollmer, H. J.; Woldt, R. Living Polymers
on Polymerization with Extremely Productive Ziegler Catalysts.
Angew. Chem., Ind. Ed. Engl. 1980,19,390.
Sinn, H.;Bliemeister, J.; Clausnitzer, D.; Tikwe, L.; Winter, H.;
Zarncke, 0. Transition Metals and Organometallics as Catalysts
for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.;
Springer-Verlag: Berlin, 1988;p 417.
Soga, K.; Kaminaka, M. Polmerization of Propene with the Heterogeneous Catalyst System Et[Ind&] zZrClz/MAO/SiOzCombined
withTrialkylaluminum. Makromol. Chem.Rapid Commun. 1992,
13, 221.
Soga, KO;Chen, S.; Doi, Y.; Shiono, T. Aduances in Polyolefins;
Seymour, R. B., Cheng, T., Eds.; Plenum Press: New York, 1987;
p 143.
Soga, K.; Yanagihara, H.; Lee, D. Effect of Monomer Diffusion in the
Polymerization of Olefins over Ziegler-Natta Catalysts. Makromol.
Chem. 1989,190,995.
Soga, K.; Kaminaka, M.; Shiono, T. Heterogeneous Metallocene
Catalysts Activated by Ordinary Alkylaluminums. Metcon93;
Catalyst Consultants Inc.: Houston, TX, 1993.
Speakman, J. G. Process for the Gas Phase (Co)polymerization of
Ethylene. Eur. Patent Appl. EP435515 Al, July 3,1991.
Spitz, R. Recent Advances in Mechanistic and Synthetic Aspects
of Polymerization, NATO AS1 Series, Series C, 215;Fontanille,
M., Guyot, A.,Eds.; Reidel Publishing Company: Dordrecht, 1987;
p 485.
Spitz, R.; Pasquet, V.; Guyot, A. Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W.,
Sinn, H., Eds.; Springer-Verlag: Berlin Heidelberg, 1988;p 405.
SRI International. Metallocenes: Catalysts for the New Polyolefin
Generation; 1993;Vol. 1.
Stevens, J. C. INSITETMCatalyst Structure/Activity Relationships
for Olefin Polymerization. Metcon93; Catalyst Consultants Inc.:
Houston, TX, 1993.
Stockmayer, W. H. Distribution of Chain Length and Composition
in Copolymers. J. Chem. Phys. 1945,13, 199.
Story, B. A.; Knight, G. W. The New Family of Polyolefins from
Insite* Technology. Metcon93; Catalyst Consultants Inc.: Houston, TX, 1993.
Tait, P. J. T. Transition Metals and Organometallics as Catalysts
f o r Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.;
Springer-Verlag: Berlin Heidelberg, 1988; p 309.
Tait, P. J. T.; Watkins, N. D. Comprehensive Polymer Science;
Eastmond, G. C., Ledwith, A., Russo, S., Sigwalt, P., Eds.;
Pergamon Press: New York, 1989;Vol. 4,p 533.
Tait, J. T.; Booth, B. L.; Jejelowo, M. 0. Catalysis in Polymer
Synthesis; Vanderberg, E. J., Salamone, J. C., Eds.; American
Chemical Society: Washington, DC, 1992;p 78.
Talbot, J. The Dynamic Modelling and Particle Effects on a Fluidised
Bed Polyethylene Reactor, Ph.D. Thesis, QueensUniversity, 1990.
Taylor, T. W.; Choi, K. Y.; Yuan, H.; Ray, W. H. Transition Metal
Catalyzed Polymerizations Alkenes and Dienes; Quirk, R. P.,
Ed.; Harwood Academic Publishers: New York, 1981;Part A, p
191.
Trieschmann, H. G.; Ambil, K. H.; Rau, W.; Wisseroth, K. Method
of Removing Heat from Polymerization Reactions of Monomers
in the Gas Phase. U.S.Patent 4012573,March 15,1977.
Tsutsui, T.; Kashiwa, N. Kinetic Study on Ethylene Polymerization
with CpzZrCll Methylaluminoxane Catalyst System. Polym.
Commun. 1988,29, 180.
Uozumi, T.; Soga, K. Copolymerization of Olefins with KaminskySinn-Type Catalysts. Makromol. Chem. 1992,193, 823.

Ind. Eng.Chem. Res., Vol. 33, No. 3, 1994 479


Usami, T.; Gotoh, Y., Takayama, S. Generation Mechanism of ShortChain BranchingDistribution in Linear Low-DensityPolyethylene.
Macromolecules 1986,19,2722.
Vandenberg, E. J.; Perka, B. C. Polymerization Processes; Schildknecht, C. E.; Skeist, I., Eds.; John Wiley & Sons: New York,
1977;p 337.
Vandenberg, E. J.; Salamone, J. C. Catalysis in Polymer Synthesis;
American Chemical Society: Washington, DC, 1992.
Vela Estrada, J. M.; Hamielec, A. E. Measurement of the Bivariate
Distribution of Composition and Molecular Weight for Binary
Copolymers-A Review. Polym. React. Eng. 1992-93, 1, 171.
Villermaux, J.; Lorenzini, P.; Bertrand, P.; Greffe, J. L. Polymer
Reaction Engineering; Reichert, K. H.; Geiseler, W., Eds.; VCH
Publishers: New York, 1989; p 350.
Wagner, B. E.; Goeke, G. L.; Karol, F. J. Process for the Preparation
of High Density Ethylene Polymers in Fluid Bed Reactor. U S .
Patent 4303771,Dec 1, 1981.
Webb, S. W.; Conner, W. C.; Laurence, R. L. Polymer Reaction
Engineering; Reichert, K. H., Geiseler,W., Eds.; VCH Publishers:
New York, 1989a;p 381.
Webb, S. W.; Conner, W. C.; Laurance, R. L. Monomer Transport
Influences in the Nascent Polymerization of Ethylene by SilicaSupported Chromium Oxide Catalyst. Macromolecules 1989b,
22,2885.
Webb,S. W.; Weist,E.L.;Chiovetta,M.G.;Laurence,R.L.;Conner,
W. C. MorphologicalInfluences in the Gas Phase Polymerization
of Ethylene by Silica Supported Chromium Oxide Catalysts. Can.
J. Chem. Eng. 1991,69,665.
Weist, E. L.; Ali, A. H.; Conner, W. C. Morphological Study of
Supported Chromium Polymerization Catalysts 1. Activation.
Macromolecules 1987,20,689.
Weist, E. L.; Ali, A. H.; Naik, B. G.; Conner, W. C. Morphological
Study of Supported Chromium PolymerizationCatalysta 2. Initial
Stages of Polymerization. Macromolecules 1989,22,3244.
Wild, L. International Conference on Polyolefins: Sixth; SPE:
Houston, 1989;p 413.
Wild, L. Temperature Rising Elution Fractionation. Adu. Polym.
Sci. 1991,98,1.
Wild, L. CompositionDistributions in Polyethylene. A Job for TREF.
Trends Polym. Sci. 1993,1, 50.
Wild, L.; Ryle, T. R. Crystallizability Distributions in Polymers: a
New Analytic Technique. Polym. Prepr. 1977,18,182.
Wild, L.; Ryle, T. R.; Knobeloch, D. C. Branching Distribution in
Linear Low Density Polyethylene. Polym. Prepr. 1982a,23,133.
Wild, L.; Ryle, T. R.; Knobeloch, D. C.; Reat, I. R. Determination
of Branching Distribution in Polyethylene and Ethylene Copolymers. J. Polym. Sci. Polym. Phys. Ed. 1982b,20,441.

Wild, L.; Ryle, T. R.; Knobeloch, D. C. Branching Distribution in


Linear Low-DensityPolyethylene. Polym. P r e p . 1982c,23,133.
Wilfong, D. L.; Knight, G. W. CrystallizationMechanismsfor LLDPE
and its Fractions. J. Polym. Sci. B: Polym. Phys. 1990,2$,861.
Wisseroth, V. K. Ethylene Gas Phase Polymerization. Angew.
Macromol. Chem. 1969,8,41.
Wristers, J. Nascent Polypropylene Morphology: Polymer Fiber. J.
Polym. Sci. Polym. Phys. Ed. 1973a,11, 1601.
Wristers, J. Direct Examination of Polymerization Catalyst by
Electron Microscopy. J.Polym. Sci. Polym. Phys. Ed. 1973b,11,
1619.
Xie, T. Y.; Hamielec, A. E. ModellingFree Radical Copolymerization
Kinetics: Evaluation of the Pseudo-Kinetic Rate Constant Method, Part I Molecular Weight Calculationsfor Linear Copolymers.
Makromol. Chem. Theory Simul. 1993a,2,421.
Xie, T. Y.; Hamielec,A. E. ModellingFree Radical Copolymerization
Kinetics: Evaluation of the Pseudo-Kinetic Rate Constant Method, Part 11: Molecular Weight Calculations for Copolymers with
Long Chain Branching. Makromol. Chem. Theory Simul. 1993b,
2,455.
Xie, T. Y.; Hamielec, A. E. ModellingFree Radical Copolymerization
Kinetics: Part I11 Molecular Weight Calculationsfor Copolymers
with Crosslinking. Makromol. Chem. Theory Simul. 1993c,2,
777.
Xie, T. Y.; McAuley, K. B.; Hsu, C. C.; Bacon, D. W. Gas Phase
Ethylene Copolymerization: Mechanisms, Kinetics, and Reactor
Modelling. Presented at the 43rd Canadian ChemicalEngineering
Conference, Ottawa, October 1993.
Yoon, J. S.;Ray, W. H. Simple Mechanistic Model for the Kinetics
and Catalyst Activity Decay of Propylene Polymerization Over
Tic&Catalyst with DEAC Cocatalyst. Znd. Eng. Chem. Res. 1987,
26,415.
Yuan, H. G.; Taylor, T. W.; Choi, K. Y.; Ray, W. H. Polymerization
of Olefin Through Heterogeneous Catalysis I. Low Pressure
Propylene Polymerization in Slurry with Ziegler-Natta Catalyst.
J.Appl. Polym. Sci. 1982,27,1691.
Zakharov, V. A.; Yermakov, Yu. I. Supported Organometallic
Catalysts for Olefin Polymerization. Catal. Rev. Sci. Eng. 1979,
19,67.
Zucchini, U.; Cechin, G. Adv. Polym. Sci. 1983,51,557.
Received for review July 6, 1993
Revised manuscript received November 5, 1993
Accepted November 16, 19938
Abstract published in Advance ACS Abstracts, January 15,
1994.

Potrebbero piacerti anche