Sunteți pe pagina 1din 38

Polymer Science & Engineering Course

Chapter 2: Polyolefins 1
Polyolefins
1. Lecture outline
1. What are polyolefins?
2. Olefins source.
3. Types of polyolefins and their properties
4. Polymerization reactions of polyolefins
4.1 Ziegler-Natta catalysis
4.2 Metallocenes
4.3 Polymerization mechanism
4.4 Polymerization kinetics.
2. What are polyolefins
A polyolefin is a polymer produced from an olefin or alkene as a monomer. In organic
chemistry, an alkene, olefin or olefine is unsaturated chemical compound containing at
least one carbon to carbon double bond. The simplest alkene is ethylene.
A special family of these olefins is -olefines, which have a double bond at the
primary or position, Structure I. This location of a double bond enhances the reactivity
of the compound and makes it useful for a number of applications.

(I)
3. Olefins source
The most common industrial synthesis path for alkenes is cracking of petroleum.
Cracking is the process whereby complex organic molecules are broken down into
simpler molecules (e.g. light hydrocarbons) by the breaking of carbon-carbon bonds in
the precursors.
Steam cracking
It is a petrochemical process in which saturated hydrocarbons are broken down into
smaller, often unsaturated, hydrocarbons. It is the principal industrial method for
Polymer Science & Engineering Course
Chapter 2: Polyolefins 2
producing the lighter alkenes (or commonly olefins), including ethene (or ethylene) and
propene (or propylene).
In steam cracking, a gaseous or liquid hydrocarbon feed like Naphtha, LPG or Ethane
is diluted with steam and then briefly heated in a furnace (obviously without the presence
of oxygen). Typically, the reaction temperature is very hot around 850Cbut the
reaction is only allowed to take place very briefly. In modern cracking furnaces, the
residence time is even reduced to milliseconds (resulting in gas velocities reaching speeds
beyond the speed of sound) in order to improve the yield of desired products. After the
cracking temperature has been reached, the gas is quickly quenched to stop the reaction
in a transfer line exchanger.
The products produced in the reaction depend on the composition of the feed, the
hydrocarbon to steam ratio and on the cracking temperature & furnace residence time.
Light hydrocarbon feeds (such as ethane, LPGs or light naphthas) give product streams
rich in the lighter alkenes, including ethylene, propylene, and butadiene. Heavier
hydrocarbon (full range & heavy naphthas as well as other refinery products) feeds give
some of these, but also give products rich in aromatic hydrocarbons and hydrocarbons
suitable for inclusion in gasoline or fuel oil. The higher cracking temperature (also
referred to as severity) favors the production of ethene and benzene, whereas lower
severity produces relatively higher amounts of propene, C4-hydrocarbons and liquid
products.
Ethylene is synthesized by steam cracking; however, propylene is just a side product.
Propylene is synthesized by another method such as propane dehydrogenation. FCC LPG
is an important source propylene and butylene.
Types of polyolefins and their properties
1. Polyethylene
1.1 Low-density polyethylene (LDPE)
This polymer is made by free-radical polymerization. It is made using supercritical
ethylene (T
c
= 282.4 K, P
c
= 50.4 bar) under severe polymerization conditions in: (i)
autoclaves (1500-2000 atm, 180-290 C) or (ii) tubular reactors (1500-3500 atm, 140-180
C).
Reaction mechanism
As mentioned earlier, the polymerization reaction is a free-radical polymerization, with
organic peroxides providing the source of the free radicals, which are short-lived reactive
Polymer Science & Engineering Course
Chapter 2: Polyolefins 3
intermediates with unpaired electron. The reaction starts when a free radical reacts with
an ethylene molecule, forming a new molecule until the growth of long chain molecules
ends. At the high pressures involved, the polymerization step is very rapid. The
polymerization process can be described by the classic kinetic description of free-radical
(chain) polymerization.
When simplifying the free radical mechanism, the following steps can be
distinguished:

1. Initiation. Free-radical sites for polymerization are formed by reaction between
primary initiator free-radical fragments and vinyl molecules, Schemes 1and 2. Free
Radicals are generated by thermal dissociation of initiators. Free radicals are
molecules with unpaired electrons.
2. Propagation. Polymerization then proceeds through a series of additions of monomer
molecules to the growing polymer chain, with the free-radical site jumping to the end
of the growing chain after each addition, Scheme 3.
3. Termination. Active free radicals are destroyed by either combination (coupling) or,
more rarely, disproportionation, Schemes 4 and 5 respectively.

These steps are summarized in schemes 1-5 below. These steps are the part of the
mechanism that determines the polymerization rate (i.e. conversion). Peroxide type and
concentration, as well as reactor temperature control the initiation rate, equation 1. The
propagation rate increases with pressure (i.e. ethylene concentration) and temperature.

2
1
] [ 2
] [ 2 2
= +
=
M k R M M R
I fk R R I
t i
k
d d
k
i
d
(1)

Note that to deriving the kinetics model for the polymerization reaction, (quasi) steady
state assumption is made. This assumption states that the concentration of chain radicals
increases initially, but almost instantaneously reaches a constant steady state value. The
rate of change of the concentration of radicals quickly becomes and remains zero during
the course of the polymerization. This is equivalent to stating that the rates of initiation
and termination of radicals are equal, Equation 1b.
In most polymerizations, the second step (the addition of the primary radical to
monomer) is much faster than the first step. The thermal dissociation of the initiator is the
rate-determining step in the initiation sequence, and the rate of initiation is given by
Equation 1a. It can be proved that the rate of polymerization can be estimated by the
following expression (derive it):
Polymer Science & Engineering Course
Chapter 2: Polyolefins 4
2 / 1
] [
] [

=
t
d
p p
k
I fk
M k R (2)

Where [I] is initiator concentration, and f is the initiator efficiency which is defined as the
fraction of the radicals produced in dissociation reaction that initiate polymer chains. It is
value is usually less than unity. k
t
is the overall termination reaction rate constant.


Scheme 1

Scheme 2

Scheme 3

There is another class of reactions that determines the molecular weight and molecular
weight distribution of the polymer. There are

4. Chain transfer reactions. Active free radical sites at the ends of growing chains jump to
another site on the same polymer molecule, another polymer molecule, a solvent,
monomer, or modifier molecule. Chain transfer affects the size, structure, and end
Polymer Science & Engineering Course
Chapter 2: Polyolefins 5
groups of the polymers. Chain transfer agents (modifiers) are able to donate hydrogen
atoms to this reaction.



Scheme 4


Scheme 5

Polymer Science & Engineering Course
Chapter 2: Polyolefins 6
Chain transfer
During the polymerization reaction, the active polymer may react with monomer,
initiator, or solvent; this will stop the growth of the active polymer and produce active
species that will start the polymerization reaction again. The radical displacement is
termed chain transfer reaction, it may be depicted as

+ + A X M XA M
n
k
n
tr
(3)

Chain transfer results in the production of a new radical, A, which then reinitiates
polymerization

+ M M A
a
k
(4)

Chain transfer reaction results in a decrease in the size of the produced polymer chains.
The effect of this reaction on polymerization rate depends on whether the rate of re-
initiation, k
a
, is comparable to that of the original propagating radical. For example, for k
p

>> k
tr
, the polymerization rate will not chang if k
a
k
p
; however, the polymerization rate
decreases if k
a
< k
p
.
The relative decrease in polymer molecular weight, which is usually defined by the
number average degree of polymerization X
n
, depends on the magnitude of the transfer
constant, k
tr
.

o
n
n
M
M
X = (5)

The degree of polymerization is defined as the polymerization rate divided by the sum of
the rates of all chain breaking reactions (termination plus transfer reactions). For the
general case of a polymerization terminated by a combination of coupling and
disproportionation and chain transfer to monomer, initiator, and the chain transfer agent,
X
n
can be written as:

CTA tr I tr M tr td tc
p
n
R R R R R
R
X
, , ,
+ + + +
= (6)

Assuming that disproportionation does not occur and replacing other reactions by the
following expressions
Polymer Science & Engineering Course
Chapter 2: Polyolefins 7

p
CTA tr
CTA
p
I tr
I
p
M tr
M
i
i p p
k
k
C
k
k
C
k
k
C
M M k R
, , ,
1
] [ ] [
= = =
=

=
(7)

Combining Equations 6 and 7 yields

] [
] [
] [ ] [
1
3 2
2
2 2
M
CTA
C
M fk k
R k
C C
M k
R k
X
CTA
d p
p t
I M
p
p t
n
+ + + = (8)

This equation is often called Mayo Equation.
This equation determines quantitatively the effect of different polymerization reaction
components as well as polymerization temperature on the molecular weight of the
product. For example, if the monomer is too active in transfer, it sets the upper limit of
the molecular weight that can be achieved as long as all other factors are optimized. The
effect of temperature is affected by the activation energies of the kinetic constants; for
example, if
CTA
E < E
p
, polymer molecular weight increases with temperature increase.
Note that this is applicable if all other conditions are not affected by temperature change.
Molecular weight distribution
This discussion is correct for high-conversion polymerization. During the
polymerization reaction, the molecular weight depends on the ratio [M]/[I]
1/2
according to
Equation 8. Note that the molecular weight depends on the ratio [M]/[I] according to the
third term in this equation. However, this term could be negligible compared to the first
term which is the case. In the usual situation [I] decreases faster than [M] and the
molecular weight of the polymer produced at any instant increases with time. The overall
molecular weight distribution for high or complete conversion polymerizations are quite
broad, with PDI being in the range 2-5.
Chain transfer to polymer
Another possible transfer reaction is chain transfer to polymer. This reaction is
negligible at low conversions because the polymer concentration is low. However, this
reaction cannot be neglected for the practical situation where polymerization is carried to
complete or high conversion. Transfer to polymer results in the formation of a radical site
on a polymer chain. The polymerization of monomer at this site leads to the production of
a branched polymer, which drastically decreases the crystallinity of a polymer.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 8
The extent of branching is greater in polymers, such as poly(vinyl acetate), PVC, and
PE, which have very reactive propagating radicals.

Long-chain branching
The extent of branching in polyethylene varies depending on the polymerization
temperature and other reaction conditions. Long chain branches increases with increasing
polymerization temperature, polymer residence time, and conversion, but decreases with
increasing pressure. The branches in PE are of two types: short branches (<7 carbons)
and long branches. Long branches due intermolecular chain transfer arising from
reactions shown in Scheme 6. These reactions lead to branches which are, on the average,
as long as the main polymer chain. This sort of branching has an observable effect on the
solution viscosity of the polymer and thus greatly influences its processing
characteristics. Polyethylene melt viscosity decreases with increasing degree of long-
chain branching at constant weight average molecular weight, Figure 1; whereas, in
poly(vinyl acetate) melt viscosity increases under the same conditions. This is attributed
to the following facts. First, the resistance to flow, i.e. viscosity, increases with increasing
crystallinity. Second, the degree of crystallinity decreases with the increase in degree of
branching.
Long-chain branching levels are usually quite low, less than 2 per 1000 repeating unit.




Scheme 6



Figure 1. The dependence of PE viscosity on chain branching
1 , ,
,
, , +
+ +
c m b n
k m
c m b n
P D D P
tr

Polymer Science & Engineering Course
Chapter 2: Polyolefins 9
The rate low for the branching reaction, Scheme 6, can be described as

=
=
=
=
1 0
, 1
1
] [ ] [
] [
n b
b n
n
n tot
tot
pol
tr
pol
tr
mD
P P
P k R

(9)

where
pol
tr
R is the rate of transfer to polymer, P
n
is polymer radical of length n, P
n,b
is
polymer radical of length n with b branch points, D
n,b
is dead polymer of length n with b
branch points. Notice that the rate expression in Equation 9 is written assuming that all
repeat units on all dead polymer chains have an equal probability of reaction, as
represented by 1, the total concentration of polymerized monomer units in the system.

Note that the reaction is proportional to the number of repeating units contained in the
chains. This has two consequences. First, the importance of transfer to polymer increases
with polymer conversion. Thus, polymerization operating at high monomer conversion
will have significantly higher branching than low-conversion systems. Second, longer
polymer chains are more likely to participate in branching reaction than short chains.

Effect of branching on MWD
Since the reactivated chains increase in length through subsequent propagation, this
leads to broadening of the molecular weight distribution reflected by an increase in the
weight average molecular weight. A PDI of 20 to 50 is considered typical.

Short-chain branching
Intramolecular H-atom removal, often called back-biting occurs via the formation of a
six-membered ring, as shown in Scheme 7. The propagating radical, A, abstracts
hydrogen from the fifth, sixth, and seventh methylene groups from the radical end, B-D.
The resulting radicals propagate with monomer to form n-hexyl, n-amyl, and n-butyl
branches, respectively. Ethylene branches arise from radical D undergoes a second
intramolecular transfer reaction after the addition of one ethylene molecule prior to
further propagation, leading to 1,3-diethyl, Scheme 8A, and 2-ethylhexyl branches,
Scheme 8b.
This mechanism is important for high-pressure LDPE production at 150-300 C, where
the number of short branches is in the range of 20-50 per 1000 ethylene repeating units.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 10
The level of short-chain branches significantly decreases Polyethylene crystallinity and
gives LDPE some of its unique properties; LDPE density is 0.92 g cm
-3
compared to 0.98
g cm
-3
for linear Polyethylene. This means that properties dependent on crystallinity such
as stiffness, chemical resistance increases with increasing density or decreasing the
amount of short-chain branching in polymers; whereas permeability to liquids and gases
and toughness decrease under the same conditions.

Scheme 7

Scheme 8
Polymer Science & Engineering Course
Chapter 2: Polyolefins 11
1.2 High-density polyethylene (HDPE)
This polyethylene is made with Ziegler-Natta catalysts. It has a totally different
structure from that obtained by radical polymerization in having a much lower degree of
branching (0.5-3 vs. 15-30 side chains per 500 monomer units). Note that chain transfer
to polymer is not possible in coordination polymerization. LDPE and HDPE are referred
to branched and linear polyethylenes respectively. The low degree of branching results in
high crystallinity (70-90%) compared to (40-60%) in the case of LDPE. This increases
polymer density (0.94-0.96 g/mL vs. 0.91-0.93 g/mL) and crystalline melting temperature
(133-138 vs. 105-115 C). Compared to LDPE, HDPE has increased:

Tensile strength
Stiffness
Chemical resistance
Upper used temperature
combined with decreased:
o Elongation
o Permeability
o Resistance to stress crack
o Low-temperature brittleness
In summary, HDPE has much greater rigidity than LDPE and can be used in structural
applications.
Average molecular weight of HDPE
Most of HDPEs have number average molecular weights of 5010
3
-25010
3
. Various
specialty HDPEs are produced by polymerization to higher molecular weights. Increased
molecular weights results in increased tensile strength, elongation, and low temperature
impact resistance. High-molecular weight HDPE (0.25-1.5 10
6
g/mol) are also produced
as well as Ultra high-molecular weight HDPE (>1.5 10
6
g/mol), which has the highest
abrasion resistance and impact strength.
1.3 Linear low-density polyethylene (LLDPE)
It is a copolymer of ethylene and -olefins (generally small amounts of 1-butene, 1-
hexene, or 1-octene). LLDPE has controlled amounts of ethyl, n-butyl and, n-hexyl
branches, that makes it equivalent to LDPE. Thus, it is properties are very similar to
LDPE properties. Thus, LDPE is now being supplemented by LLDPE. However, not all
LDPE properties are achievable by LLDPE. LLDPE is produced in a gas-phase reactor
(P ~ 20 atm and ~ 80 C); this process can be used to produce HDPE and PP as well.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 12
While all the properties of LDPE are not achieved by the new material, the economics of
the low pressure, solvent-free process limit the need to build new high-pressure plants.
Note that PE products with even lower densities, 0.88 g/cm
3
, are sometimes called very
low-density polyethylene (VLDPE) but are chemically identical to LLDPE.

Table 1. Molecular structure and properties of various grades of polyethylene.

2. Polypropylene
Polypropylene is the second most important commercial polyolefin. Isotactic PP has
the lowest density (0.90-0.91 g/mL) of the major plastics. It has a high crystalline melting
point of (165-175 C). Polymerization of propylene is carried out with coordination
catalysts. Ethylene, propylene and other -olefins can be polymerized in the same
equipment with very little modifications leading to highly flexible operation. Before
proceeding, tacticity concept will be reviewed.
Tacticity
What is tacticity?
As un illustration, let us consider polystyrene. It is often drawn as a flat picture, Figure 2.
But polymers aren't really flat like that. The carbon atoms aren't really in a straight line
like that, nor are the hydrogens and phenyl groups all placed at perfect right angles. The
carbon chain is more of a zigzag with the phenol (pendant) group tends to be point away
from the chain as shown in Figure 3.
Tacticity is simply the way pendant groups are arranged along the backbone chain of a
polymer. Normally, some of the atoms in the polymer chain will have small chains of
atoms attached to them. These small chains are called pendant groups. The chains of
Polymer Science & Engineering Course
Chapter 2: Polyolefins 13
pendant groups are much smaller than the backbone chain. Pendant chains have just a
few atoms, but the backbone chain usually has hundreds of thousands of atoms, Figure 4.

Figure 2. Typical polystyrene representation


Figure 3. Dimensional representation of polystyrene




Figure 4. Pendant groups.

The regularity in configurations of successive stereocenters determines the overall
order (tacticity) of the polymer chain. A stereocenter, or stereogenic centre, is any atom
in a molecule bearing groups such that an interchanging of any two groups leads to a
stereoisomer. Stereoisomers are isomeric molecules whose atomic connectivity is the
same but whose atomic arrangement in space is different.
If the R groups on successive stereocenters are randomly distributed on the two sides of
the planar zigzag polymer chain, the polymer does not have order and is termed atactic.
Two types of ordered or tactic structures can occur: isotactic and syndiotactic. An
isotactic structure occurs when the stereocenter in each repeating unit in the polymer
chain has the same configuration. All the R groups will be located on one side of the
plane of the carbon-carbon polymer chain either all above or all below the plane of the
chain. A syndiotactic polymer structure occurs when the configurations of the
stereocenters alternate from one repeating unit to the next with the R groups located

Polymer Science & Engineering Course
Chapter 2: Polyolefins 14
alternately on the opposite sides of the plane of the polymer. These structures are shown
in Figures 5-6 for a dummy polymer and polypropylene respectively.


Figure 5. schematic illustrations of atactic (a), syndiotactic (b), and isotactic (c) polymer.



Figure 6. Schematic illustrations of PP
Factors control the stereo arrangement of the propylene polymer
1. The degree of branching.
2. The pendant methyl sequence: is the monomer unit always added in the head-to-tail
manner or it is sometimes inserted head-to-head or tail-to-tail, Figure 7.
3. Right or left hand: is it stereospecific, Figure 8.
Significance of stereoregularity
The regularity or lack of regularity in polymers affects their properties by way of large
differences in their abilities to crystallize. Atactic polymers are amorphous
(noncrystalline), soft materials with lower physical strength. The corresponding isotactic

(a)
(b)
(c)
Polymer Science & Engineering Course
Chapter 2: Polyolefins 15


Figure 7. Head-to-tail (A) and tail-to
tail (B) addition of propylene to the
growing PP chain.
Figure 8. Addition of propylene to the
growing PP chain to the same (A) and
opposite (B) hand.

and syndiotactic polymers are usually obtained as highly crystalline materials. The
ordered structures are capable of packing into a crystal lattice, while the unordered
structures are not.
The chain regularity of isotactic polypropylene is responsible for its high melting
temperature and crystallinity, making it ideal for injection molding and extrusion
applications due to its excellent rigidity, toughness, and temperature resistance.
Copolymers and terpolymers
Beyond homopolymers, there is a wide range of copolymers and terpolymers, usually
of ethylene and butene, of two types: random and impact.
Random copolymers
Random copolymers typically contain up to 6 % (by weight) of ethylene or other
comonomers inserted at random within the chain.
Similar to homopolymers, the structural parameters of M
w
and MWD play similar role
for the properties of the random copolymers, while, the stereospecificity concept changes
its meaning. In fact, the introduction of a comonomer into polymeric chain determines a
discontinuity that deeply affects the molecules crystallization behavior. Crystallization
speed slows down, forcing lower total crystallinity and a reduction of the melting
temperature related to the less perfect structure of the crystals.
Random copolymers are used where clarity, lower melting point, or lower modulus is
desirable.
Impact copolymers
Impact copolymers, also known as heterophasic copolymers, usually contain up to
about 40 % ethylene-propylene rubber (EPR), regularly distributed inside the
Polymer Science & Engineering Course
Chapter 2: Polyolefins 16
semicrystalline PP homopolymer matrix. Those copolymers are used where impact
strength is important, especially at low temperatures.
The improvement of the impact properties while retaining sufficient elastic modulus
performance, (good stiffness impact strength balance) when coupled with the basic
thermal, physical and chemical properties of the PP, has opened wide field of
applications. Key issues are found mainly in the production of structural items in the
automotive and durable goods, or also in packing industries.
The correlation between the structure and the properties of PP heterophasic
copolymers shows a more complex situation compared to what is observed for the
structurally simpler homopolymers or random copolymers. This is a direct result of and
depends on the mutual interactions of the complex structures and interactions between
these various phases. Detailed discussion of this issue is beyond the scope of this course.
Catalysts for olefin polymerization
1. Ziegler-Natta catalysts
A Z-N catalyst can be defined as a transition metal compound bearing a metal-carbon
bond able to carry out a repeated insertion of olefin units. Usually, though not
necessarily, the catalyst consists of two components: (i) a transition metal salt, as TiCl
4
,
and (ii) a main-group metal alkyl (activator) which serves the purpose of generating the
active metal-carbon bond.
Highly active heterogeneous Z-N catalyst, which is used extensively nowadays, has
active sites, TiCl
4
, for polymerization. These active sites are located at the surface and
edges of the crystalline structure of titanium chloride, MgCl
2
. This support has high
surface area and pore volume, thus; it increases catalyst activity by many times.
The cocatalysts (activators) used with MgCl
2
supported catalysts are invariably Al-
trialkyls, triethyl-Aluminum (TEA) and triiso-butyl-Aluminum (TIBA) being the most
preferred ones. For these catalyst systems, active sites are generated through the
interaction of the transition metal atoms on the catalyst surface and the organoaluminum
compound (cocatalyst). Because the cocatalysts work as a Lewis acid (electron acceptor),
it also used to scavenge polar impurities from the reactor. These impurities are electron
donor, such as oxygen, sulfur and nitrogen compounds, and moisture that poison the
cationic active sites.
A major advance in heterogeneous Z-N catalysts was the use of internal and external
donors which improve the properties of the produced polymer as well as increase catalyst
activity.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 17
Electron Donor (Lewis base)
It was found that the use of the catalyst-cocatalyst system alone would produce
polyolefins, especially polypropylene, with low stereoselectivity (II < 50%). This
problem was overcome by the addition of appropriate Lewis bases, which made it
possible to obtain highly active and stereoselective catalysts. The catalyst system consists
of: (i) catalyst, TiCl
4
, support, MgCl
2
, and a Lewis base, usually referred to as internal
donor (Di), and (ii) cocatalyst, Al-trialkyl, and a second Lewis base, usually called
external donor (D
e
).
The effect of the donors depends on the donors themselves and the other two
components in the system. The effect of these donors is usually in: (i) catalyst activity,
(ii) catalyst stereoselectivity, and (iii) polymer molecular weight distribution.
Various mechanisms have been proposed for the action of the donors in improving
stereoselectivity:

1. Lewis base coordinates to a Mg ion neighboring a less isoselective Ti site. It
influences the active site via both steric and electronic effects.
2. By the addition of the donor, the isotacticity of the highly isotactic site is improved.
3. Part of a less isotactic sites are transformed to a highly isotactic sites.
4. The activity of less isotacitic sites decreases.

Note that, depending on catalyst generation, the external donor is used because internal
donor dissolved in the monomer after reacting with the cocatalyst. This reduces its
concentration in the catalyst system. External donor is added to compensate for the loss;
otherwise, there is a lose of stereoselectivity.
Catalyst generations
There are 6 generations of Ziegler-Natta catalysts. The difference in these generations is
in: (i) catalyst, (ii) support, and (iii) donors. A summary of different generations is in
Table 2. Different families of donors are shown in Figure 8.


Figure 9. General formulas for internal and external donors.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 18
Table 2. Performance of different Ziegler-Natta catalysts.


I.I (isotactic index) is a measure of the isotactic content of a polymer. It is the percentage
of polymer sample insoluble in a hydrocarbon solvent such as boiling n-heptane.
mmmm measures the isotactic sequence in a polymer chain; in this case, five
stereocenters are considered, Figure 10.


Figure 10. Pentad.

Table 2 shows an updated comparison of the performance achievable with the different
catalyst generations. The comparison based on data obtained under bulk polymerization
conditions at 70 C and for a two-hours run. For the first two generations, the
polymerization rate was very low as well as the isotacticity. Thus, two successive steps,
after polymerization, were required: (i) polymer purification from catalyst traces, and (ii)
atactic polymer removal. For the third generation, the polymerization reaction became
more steroeselective; however, the activity was not really high. However, this changes a
lot when Lewis base, benzoate, is used. Both catalyst activity and polymerization
stereoselectivity increases and PDI decreases while we are moving from the 3
rd
to the 5
th

generation. However, this is not the case for succinate-based systems.
1st
2nd
3rd
4th
5th
Polymer Science & Engineering Course
Chapter 2: Polyolefins 19
Therefore, as a function of the donor structure, it is possible to modulate both the catalyst
performance and polymer structure.
2. Philips catalysts
Phillips catalysts are based on Cr(IV) supported on silica and alumina. The true
structure of the Phillips catalyst is not well understood. A mixture of chromium oxide and
silicon oxide (CrO
3
/SiO
2
) is used to create the active sites. The catalyst does not require
addition of chemical activators before the polymerization, since the active site is
produced prior to the polymerization by thermal activation at high temperatures (600C,
for instance). Phillips catalysts are used in both gas-phase and slurry processes.
Polyethylene made with Phillips catalysts has very broad molecular weight distribution,
with polydispersities ranging from 12 to 24. Interestingly, hydrogen is not an effective
chain-transfer agent and generally decreases catalyst activity.
3. Metallocene catalysts
What are metallocenes?
Metallocene catalysts have structures similar to ferrocene (Cp
2
Fe), Figure 11, which is
a widely known complex with a sandwich structure, where the transition metal lies
between two cyclopentadienyl (Cp) rings. Cp
2
ZrCl
2
(Figure 12) is a well studied example
of a metallocene catalyst for olefin polymerization. Metallocenes are soluble and
therefore can be used as homogeneous catalysts, but they can also be supported on inert
carriers such as silica, alumina, and magnesium chloride, among other inorganic and
organic supports, and used as heterogeneous catalysts



Figure 11. Ferrocene. Figure 12. Structure of typical metallocene
catalyst, Cp
2
ZrCl
2
.

The cocatalysts commonly used with Ziegler-Natta catalysts (AIEt
2
Cl, AIEt
3
) are not
able to activate metallocene catalysts very well. Even though metallocenes can be
activated with these cocatalysts, they generally deactivate very fast, and consequently
have no commercial interest. The use of bulky activators, such as methylaluminoxane
(MAO), is necessary to produce highly active metallocene catalysts. In fact, it can be said
Polymer Science & Engineering Course
Chapter 2: Polyolefins 20
that it was the discovery of MAO as a good cocatalyst for metallocenes that made
possible the metallocene revolution that we are currently experiencing in polyolefin
manufacture. Other bulky Lewis acids such as trispentafluorophenylborane (B(C
6
F
5
)
3
) are
also very good cocatalysts for metallocenes.
The effect of catalyst system on polymerization reaction
A large variety of metallocene catalysts can be obtained by altering the simple
structure of Cp
2
ZrCl
2
. The nature of the transition metal and the structure of the ligand
have a large effect on catalyst behavior. The shape, geometry, and chemical structure of
the ligand can affect the activity and selectivity of the catalyst. The symmetry imposed by
ligands around the active site determines the geometry for monomer coordination and
insertion, and consequently the relative orientation of catalyst and growing polymer
chain. As an example, consider (2-phenylindenyl)zirconium dichloride. This metallocene,
as shown in Figure 13, has no bridge between the two indenyl rings compared to the
catalyst shown in Figure 14. This means that the two rings can spin around freely.
Sometimes the rings will be pointed in opposite directions, called rac form. Other times
the rings will be pointing in the same direction, called meso form. The compound spends
some time in the rac form, then flips around, and becomes the meso form. After awhile it
will flip back again. This happens over and over again. So how this will affect on the
polymerization reaction?

Figure 13 rac and meso forms for (2-phenylindenyl)zirconium.

Figure 14 Bridged metallocene catalyst.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 21
When the zirconocene is in the rac form, propylene monomer can only approach in an
orientation which will give isotactic polypropylene, Scheme 9a. Note the monomer in the
left side could not approach the active site due to the repulsion forces between the methyl
groups. But when the zirconocene flips and becomes the meso form, propylene monomer
can approach in any orientation. This will give atactic polypropylene, Scheme 9b.
Note that the zirconocene is constantly flipping back and forth between the two forms. It
even does this while polymerization is taking place. This means that the same polymer
chain will end up having blocks that are atactic and blocks that are isotactic, Figure 15.
This kind of polypropylene is called elastomeric polypropylene. Elastomers can be
stretched to many times their original length, and can bounce back into their original
shape without permanent deformation. By the proper choice of polymerization conditions
it is possible to produce polypropylene that have chains varying from mainly atactic
through block atactic-isotactic to mainly isotactic configurations.



Scheme 9. Propylene reaction with active metallocene catalyst with rac form (a), and meso
form (b).

Figure 15 Elastomeric propylene.

There is a difference between polyolefins made by Z-N catalysts and those made by
metallocenes. Metallocene catalysts offer a narrow molecular weight distribution and a
(a)
(b)
Polymer Science & Engineering Course
Chapter 2: Polyolefins 22
significantly reduced low molecular weight fraction. Metallocenes are more sensitive to
hydrogen than Z-N catalysts, which results in the ability to produce very high melt flow
rate polymers directly in the reactor. Metallocenes offer better stiffness and clarity.
However, yields from metallocenes have been low.
It is widely expected that, Ziegler-Natta catalysts are and will be for years to come, the
leading industrial catalyst systems for the production of polyolefins. Figure 15 show the
forecasts for PE and PP demand up to 2010.



Figure 16 Worldwide forecast PE demand (a), and PP demand (b).
b
a
Polymer Science & Engineering Course
Chapter 2: Polyolefins 23
Mechanism of polyolefin polymerization
In this part, the focus will be on Z-N catalyst. The mechanism is poorly understood
because it takes place on the surface of an insoluble particle, a difficult situation to probe
experimentally. Structure II is generally considered as the active species formed from
titanium chloride and alkyl-aluminum components. The empty square represents the
vacant site of the titanium complex. The cocatalyst shares one of its methyl groups (R)
group with the titanium.

(II)

Whereas different opinions are reported in the literature about the nature of the
polymerization centers, it is assumed that polymerization of olefins with Z-N catalysts
involves a stepwise insertion of the monomer into a transition metal-carbon bond as
follows:
R C C Mt C C n R Mt
n
= + ) ( (10)

Polymerization with coordination catalysts proceeds via two main steps: monomer
coordination to the active site, and monomer insertion into the growing polymer chain.
Before insertion, the double bond in the olefin monomer coordinates to the coordination
vacancy of the transition metal. After the olefin is inserted into the growing polymer
chain, another one can coordinate to the vacant site; thus the process of insertion is
repeated to increase the size of the polymer chain by one monomer unit at a time until
chain transfer takes place. In the case of copolymerization, there is a competition between
comonomers to coordinate to the active sites and to be inserted into the growing polymer
chains. Different rates of coordination and insertion of comonomers determine the final
chemical composition of the copolymer chain.

Definition. Monomer coordination is orienting the monomer with respect to the growing
chain end and brings about stereoselective addition
Insertion mode and regioselectivity
The insertion of an -olefin in the metal-carbon bond may take place in two different
ways:
Polymer Science & Engineering Course
Chapter 2: Polyolefins 24
Mt-P+CH
2
=CH-CH
3
Mt-CH
2
-CH(CH
3
)-P (1,2 or primary insertion)
Mt-CH(CH
3
)-CH
2
-P (2,1 or secondary insertion)

(11)
It is proved that the 1,2 insertion mode is operative in isospecific polymerization of
olefins. Secondary insertion results in a temporarily inactive catalyst site, called usually
dormant site. This site is activated by hydrogen.
Chain termination
Z-N polymerization has the characteristic of living polymerizations with regard to
active sites but not individual chains. The life time of propagating chains is very short
compared to life time of active sites, seconds-minutes vs. hours.
Definition. Living polymerization is a form of addition polymerization where the ability
of a growing polymer chain to terminate has been removed. This can be accomplished in
a variety of ways. Chain termination and chain transfer reactions are absent and the rate
of chain initiation is also much larger than the rate of chain propagation. The result is that
the polymer chains grow at a more constant rate than seen in traditional chain
polymerization and their lengths remain very similar (i.e. they have a very low
polydispersity index). Living polymerization is a popular method for synthesizing block
copolymers since the polymer can be synthesized in stages, each stage containing a
different monomer. Additional advantages are predetermined molar mass and control
over end-groups. Living polymerization in the literature is often called "living"
polymerization or controlled polymerization.
Several chain-transfer mechanisms are operative in coordination polymerization: -
hydride elimination, -methyl elimination transfer to monomer, transfer to cocatalyst, and
transfer to chain transfer agent, commonly hydrogen. The type of termination reaction
determines the chemical group bounded to the active site and the terminal chemical group
in the polymer chain. The first three types produce unsaturated chain ends, while the last
two types produce saturated chain ends. Equations 12-16 summarize these reactions:

Transfer with hydrogen
P CH CH CH H Mt H P CH CH CH Mt
tH
k
+ +
3 3 2 3 2
) ( ) (
2
(12)
Polymer Science & Engineering Course
Chapter 2: Polyolefins 25
Transfer with monomer
P CH C CH
CH CH CH Mt CH CH CH P CH CH CH Mt
tM
k
= +
= +
3 2
3 2 2 3 2 3 2
) (
) (
(13)
Transfer with cocatalyst
P CH CH CH Al R R Mt AlR P CH CH CH Mt
tMR
k
+ +
3 2 2 3 3 2
) ( ) ( (14)
-hydride elimination
P CH CH CH H Mt P CH CH CH Mt
hyd t
k
= +
3 2 3 2
) ( ) (

(15)
-methyl elimination
P CH CH CH CH Mt P CH CH CH Mt
tmethyl
k
= +
3 2 3 3 2
) ( ) ( (16)

The extents to which the various chain transfer reactions occur depend on the
monomer, concentration of the initiator components, temperature, and other reaction
conditions. -hydride transfer reactions are usually the predominant chain-breaking
reaction in the absence of H
2
or other active hydrogen compound. H
2
is a highly effective
chain-transfer agent and is used for molecular weight control in commercial
polymerizations of ethylene and propylene.
There are considerable differences in the extent of transfer to group IIII metal
components. TEA is much less active transfer agent compared to dialkylzinc (ZnEt
2
).
Thus, chain transfer with cocatalyst plays a relatively major role only if dialkylzinc is
used. This is due to the fact that dissociation constant of dialkylzinc is high. This
produces large concentration of alkyl groups, which reacts with the active sites bearing
the polymer chains.
Chain transfer to molecular hydrogen not only affects polymer molecular weight, but
unlike other transfer agents, also affects polymerization rate. Hydrogen often decreases
the rate of ethylene polymerization, but increases the rate of propene polymerization. The
rate depression for ethylene is attributed to the high stability of the Ti-ethyl group,
formed when transfer is followed by the addition of a monomer molecule. The reactivity
of a propagating chain in propene polymerization decreases whenever there is 2,1-
insertion (secondary insertion). The resulting 2,1-chain ends are preferentially terminated
by transfer to hydrogen; the overall result is replacement of 2,1-chain ends by more
reactive 1,2-chain ends on re-initiation.
It should be emphasized that there are many proposed models to at least partially
explain the action of the Ziegler-Natta systems, but it is only an approximation of the
more complex process that actually occurs. Scheme 10 shows one of these proposed
models.

Polymer Science & Engineering Course
Chapter 2: Polyolefins 26


Scheme 10 The bimetallic polymerization mechanism.
Catalyst deactivation
The reason for catalyst decay is not well understood. There are many hypotheses:

1. One of the most widely supported hypotheses for catalyst activity decay is the
over-reduction of the active centers, Ti
3+
, into an inactive form, Ti
2+
, by the
cocatalyst.
2. The formation of dormant sites.
3. Reaction of active sites with polar impurities.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 27
4. Physical deactivation. It is based on the formation of a polymer shell, surrounding
the catalyst, during the polymerization reaction. This shell introduces mass-
transfer limitations of the monomer molecules that lead to the observed catalyst
decay.
5. Two active sites form stable complex that is inactive for monomer
polymerization. This type of bimolecular intermediate is favored at high catalyst
concentrations and is reversible.
Kinetics of polyolefin polymerization
Observed rate behavior
It is well known that the overall kinetic behavior of any catalytic polymerization
process is the resultant of the overlapping of: the time dependence of the overall
polymerization reaction and the time dependence of any diffusion processes that may also
take place. The kinetic behavior of heterogeneous Ziegler-Natta catalysts is complicated
by adsorption reactions of monomer, cocatalyst, donors and by-products formed in
reactions leading to the formation of active centers.
A number of different types of kinetic rate-time profiles are encountered in the studies
of Ziegler-Natta catalysts, typical examples are shown in Figure 17. These profiles
resulted from the variation of different polymerization reaction factors as catalyst
preparation, cocatalyst and monomer types, and the polymerization media. Generally,
these profiles are kinetic finger-prints for a given catalyst-monomer system.
The behavior described by plot (a) is usually observed when the particle size of the
transition metal component is relatively large. The particles of the transition metal
component consist of aggregates of smaller crystals, Figure 18. The mechanical pressure
of the growing polymer chains cleaves these aggregates with the result that the initiator
surface area, number of active sites, and polymerization rate increase with time. After this
initial period, referred to as a buildup or settling period, a steady-state rate is reached,
which corresponds to the splitting of the initial particles to the smallest-sized particles.
Type (b) behavior is often observed when trialkylaluminum compounds are used as
cocatalysts. The use of a trialkylaluminum compound usually produces a more active but
less stable catalyst. In type (b) behavior the rate of polymerization increases to reach a
maximum and then decreases with increase in polymerization time.
Many high activity supported catalyst systems show either (c) or (d) type behavior, in
which the rate may either start at a maximum value or rise very rapidly to a maximum
value and then decrease rapidly with time, e.g. MgCl
2
/EB/TiCl
4
-AlEt
3
catalyst systems
when used for either ethylene or propylene polymerization (EB = ethyl benzoate). Note
Polymer Science & Engineering Course
Chapter 2: Polyolefins 28
(a)
(c)
(e)
Time
R
a
t
e
(b)
(f)
(d)

Figure 17 Typical kinetic rate-time profiles.

Figure 18 Particle fragmentation during polymerization reaction.

that this behavior indicates the presence of active sites of differing activities with some
of the active sites decaying with time.
Type (c) behavior was observed in the early polymerization of styrene by catalysts
derived from TiCl
4
and AlEt
3
Such behavior is also shown by many homogeneous
catalyst systems, e.g. CP
2
TiEtCI-AlEtCI
2
in ethylene polymerization, and by modified
catalyst systems, such as VOCI
2
2THF-AIR
3
, in vinyl chloride polymerization.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 29
Type (e) behavior is shown by Solvay & Cie ether-treated (second generation)
catalysts in propylene polymerization. Breakdown of the porous catalyst particles is
practically instantaneous on treatment with a dialkylaluminum compound and the settling
period is more or less eliminated. Thereafter the polymerization rate decreases very
gradually with time and the catalyst system shows good stability.
Type (f) behavior, for which the polymerization system shows no settling or
adjustment period but whose rate remains strictly constant with time, is rare and so is not
often encountered. Nevertheless such behavior is shown by MgCl
2
-supported catalysts
containing phthalate esters when used for the polymerization of 4-methyl-1-pentene (4-
MP-1). Here no settling period is observed and the polymerization rate remains almost
completely constant with time. Catalyst systems CP
2
TiCl
2
-aluminoxane also shown
constant polymerization rates apart from a short settling period.
Kinetic models
Kinetic investigations of olefin polymerization with heterogeneous Z-N catalysts are
very useful in the industrial development of the reaction but have made limited
contribution to the understanding of the mechanism of polymerization, due to:

1. the presence of different types of active sites on its surface
2. the activity decay during the polymerization
3. the chemical modification of the components of the catalyst during the time
4. the influence of the procedure of contact of catalyst components on the rate of
stereospecificity of the polymerization.
Comparison between coordination and free-radical polymerization
In free-radical polymerization, the initiator fragment moves away from the growing
chain and therefore can influence chain growth and termination during only the first few
monomer insertion steps. Consequently, the values of the polymerization kinetic
parameters depend mainly on the monomer type.
In contrast, in coordination polymerization chain growth and termination take place by
insertion of the monomer or chain transfer into metal-carbon bond. Consequently,
electrical and steric effects around the active site affect polymerization kinetics as much
as monomer type. The mechanisms of free-radical and coordination polymerization are
contrasted in Figure 19.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 30

Figure 19 Comparison of the monomer propagation step in free-radical and
coordination polymerization.

Since the nature of the active sites is rather complex, as discussed above, this makes
the determination of generalized polymerization rate constant and activation energy is
virtually impossible. On the other hand, this phenomenon is responsible to for the
remarkable flexibility of coordination catalysts since polymers with completely different
properties can be made with only a few monomer types by varying the way these
monomers are inserted into the polymer chain via active-site design.
Polymerization kinetics with multiple-site catalysts
All heterogeneous Ziegler-Natta and Phillips catalysts have two or more active-site
types and many soluble Ziegler-Natta and metallocene catalysts may also show multiple-
site behavior. In addition, several metallocene catalysts, when supported on organic and
inorganic carriers, may behave like multiple-site catalysts even if they behaved as single-
site catalysts in solution polymerization. Therefore, several of the catalysts used
industrially for polyolefin manufacturing have in fact two or more active-site types.
The kinetics of polymerization with multiple-site catalysts is generally considered to
be the same as with single-site catalysts, Equations 17-30.

*
C Al C
Al
k
+
(17)
Polymer Science & Engineering Course
Chapter 2: Polyolefins 31
After the catalyst is activated, it is ready to polymerize with monomer, M, forming living
polymer chain P
1
of length 1.

1
*
P M C
i
k
+
(18)

A subsequent monomer propagation reaction with a growing polymer of chain r, P
r
,
increases its length to r+1.

1 +
+
r
k
r
P M P
p

(19)

Generally, the first monomer insertion step is considered to have a different reaction rate
constant, k
i
, form the subsequent monomer propagation steps, k
p
. Since, practically, these
constant are very difficult, if not impossible, to estimate independently, most kinetic
models assume k
i
~k
p
, or even assume that the initiation reaction proceeds very quickly.
The most important transfer reactions in coordination polymerization are: (i) -hydride
elimination, (ii) transfer to chain transfer agent, (iii) transfer to monomer, and (vi)
transfer to cocatalyst. The corresponding reactions are summarized below respectively.

=
+
r H
k
r
D C P
t *

(20)
1
*
P M C
iH
k
H
+
(21)
=
+

r CH
k
r
D C P
CH t *
3
3


(22)
r H
k
r
D C H P
tH
+ +
*
2

(23)
=
+ +
r
k
r
D P M P
tM
1

(24)
Al r CH
k
r
D C Al P
tAl
,
*
3
+ +
(25)

where
*
H
C is a metal hydride center,
*
3
CH
C and
=
r
D is a dead polymer chain containing a
terminal vinyl unsaturation. Note that it is not possible to estimate k
i
and k
iH
, thus, the
simplifying assumption k
i
~ k
iH
is adopted. In Equation (25), the cocatalyst considered is
TEA.
Reactions with polar impurities, I, leading to the formation of a deactivated site, C
d
,
and a dead polymer chain, have been modeled in a simple way

r d
k
r
D C I P
dI
+ +
(26)
d
k
C I C
dI
+
*

(27)
Polymer Science & Engineering Course
Chapter 2: Polyolefins 32

Similarly, monomolecular and bimolecular deactivation reactions have been described in
a simple manner as described by the following set of equations

d r
k
r
C D P
d
+
(28)
d
k
C C
d

*

(29)
d
k
C C
d
2 2
*

(30)

These deactivation equations are not easy to prove, but has been successful in describing
olefin polymerization processes with Ziegler-Natta and metallocene catalysts.
In case of propylene polymerization, there are two mechanisms for monomer insertion,
either primary insertion or secondary insertion. This can be described by the following set
of equations:

Dormant site formation
1 +
+
r
k
r
S M P
dm

(31)
Dormant sites reactivation by H
2

r
k
r
D C H S
reh
+ +
*
2

(32)
Reactivation of dormant sites by monomer
r
k M
r
P S
rem

,

(33)
Polymerization rate equations
The polymerization rate of polyolefins can be estimated using different forms of rate
equations. The adopted form depends on weather monomer and cocatalyst adsorption is
significant or not. If adsorption does not have significant effect, then the following
formula can be used:

] ][ [
*
M C k R
p P
= (34)
with [C
*
] is the concentration of active sites, and [M] is monomer concentration.
Adsorption phenomena are important in most ZieglerNatta polymerizations, and this
requires treatment by heterogeneous kinetics. The exact form of the resulting kinetic
expressions differs depending on the specific adsorption phenomena that are important in
the particular reaction system. Consider a LangmuirHinschelwood model where reaction
occurs only after monomer is adsorbed from solution onto the transition metal active
sites. Further, we assume that the group IIII metal component is present in solution and
competes with monomer for the same sites; that is, there is excess group IIII metal
component over and above the amount needed to activate (reduce/alkylate) the transition
Polymer Science & Engineering Course
Chapter 2: Polyolefins 33
metal sites. The fraction
A
and
M
of the transition metal sites covered with the group
IIII metal component and monomer, respectively, are given by

] [ ] [ 1
] [
] [ ] [ 1
] [
M K A K
M K
M K A K
A K
M A
M
M
M A
A
A
+ +
=
+ +
=
(35)

where [A] and [M] are the concentrations of the group IIII metal component and
monomer in solution, respectively, and K
A
and K
M
are the respective equilibrium
constants for their adsorption. Propagation occurs by reaction of adsorbed monomer at
the active sites at a rate given by

M p P
C k R = ] [
*
(36)
Combination the above equations give

] [ ] [ 1
] ][ [
*
M K A K
M C K k
R
M A
M p
P
+ +
= (37)
Termination probability
Number-average degree of polymerization, X
n
is defined, as stated at the beginning of
this section, as ratio of propagation reaction to the summation of all chain transfer
reactions. For a polymerization reaction that can be described by Equations 19, 23, 24,
31-34, it can be proved that X
n
can be described by the following equation

rem
reh
p
dm
p
tH
p
tm
n
k
k
k
X k
X k
k
k
X
k
k
k
k
X
q
=

+
+ + = =
1
1
1
1
1
(38)
with X is hydrogen to monomer molar ration, and q, which is the inverse of X
n
, is
termination probability. This parameter is dependent on the type of catalyst and the
process conditions, such as temperature, pressure and hydrogen concentration.


Polymer Science & Engineering Course
Chapter 2: Polyolefins 34
Molecular weight distribution
In Ziegler-Natta polymerization, at any given moment during the growth of a
macromolecule, many possibilities for its evolution are encountered. The chain end can
add a new monomeric unit, or may terminate. This termination can be occured
spontaneously, or by transfer reaction, or because the active centers, upon which polymer
grows, has been deactivated. Therefore, every polymer is generally composed of
macromolecules with a different degree of polymerization, i.e. length of chains.
These lengths are distributed according to a statistical function which depends on the
polymerization mechanism, and the resulting polydispersity of the polymer, with regard
to its macromolecules, i.e. molecular weight distribution, can be described by a curve in
which the number fraction, or weight fraction is plotted versus a parameter representing
the molecular size such as chain length or molecular weight.
Most industrial catalysts for olefin polymerization are heterogeneous ZieglerNatta or
chromium oxide systems (approximately 43% and 20% of all polyethylene products,
respectively, with almost all polypropylene resins produced with ZieglerNatta catalysts).
Although there are many different types of heterogeneous Ziegler-Natta catalysts, most of
them have a common characteristic: they yield polymer with a broad MWD and, in the
case of copolymerization, broad chemical composition distribution (CCD).
Different explanations for this phenomenon have been proposed, but most of them can
be classified into two categories. One proposes that a role could be played by physical
phenomena such as: a) variation in the lifetime of the growing polymer chain depending
on the degree of active centers encapsulation in the polymeric matrix, and b) limitations
in the heat and mass transfer from the bulk to the polymer. Another explanation is based
on the presence of multiplicity of active centers energetically, structurally, and
chemically different from one another and therefore having different kinetic constants.
Recently, it is widely accepted that, under most polymerization conditions, the effect of
multiple site types on polymer properties is more important than mass and heat transfer
limitations.
Flory distribution
The molecular weight distribution of the produced polymers depends on the nature of
the chain termination reactions. For polyolefins produced with homogenous Ziegler-Natta
catalysts, this distribution can be represented by a mechanistic theory that has been
developed by Flory. This theory is applicable to polymerization processes in which:

1. One type of active sites is considered.
Polymer Science & Engineering Course
Chapter 2: Polyolefins 35
2. The ratio of chain termination reactions to the chain propagation reactions does not
depend on the chain length.
3. Each chain termination reaction produces one polymer chain.
4. There are no cross-linking or other side reactions.
5. A quasi steady-state assumption is applicable.

The theory states that the instantaneous weight distribution, y
j
, of the polymer chains
produced in such a process can be given by:

) ( 2 q j
j
e q j y

= (39)
where j is the number of repeating units.
Polyolefins produced using heterogeneous Ziegler-Natta catalysts have a broad
molecular weight distribution with polydispersity indexes that range from 3 to 10. The
presence of several sites is the main reason why polyolefins made with these catalysts
have broad and sometimes multimodal molecular weight distribution. This distribution
can be considered as an average of that produced by the individual site types, weighted
by the weight fraction of polymer produced by each site type.




Figure 20 Schematic representation of multiple active site types on the surface of a catalyst
fragment. Each active type produces polymer with chains described by Florys distribution.

Polymer Science & Engineering Course
Chapter 2: Polyolefins 36
Effect of operating conditions on kinetics and MWD
1. Kinetics
1.1 Hydrogen effect
1.1.1 Propylene
Hydrogen effect depends on two factors: (i) nature of catalyst system and (ii) monomer
type. For propylene polymerization, a remarkable activating effect of hydrogen was
observed for almost all Ziegler-Natta catalyst types. Different hypotheses have been
proposed to account for the enhancement of the polymerization rate caused by hydrogen.
Among these theories, we will concentrate on the following ones as they are the most
widely used by different authors:

Increase in the number of active sites.
The change in oxidation states.
Dormant sites theory.

Note that the third theory is the most acceptable one.
1.1.2 Ethylene
For ethylene polymerization, the effect of hydrogen on the polymerization rate is
highly dependent on catalyst type. Ziegler-Natta catalysts usually show a decrease in
polymerization rate; while, different metallocene catalysts show either an increase or an
increase followed by a decrease in the rate. These are the general dependencies, finding
different effects is always possible. The decrease in catalytic activity results from
different reasons:

1. It is well known that the main termination mechanism in the absence of hydrogen is
b-hydrogen elimination, yielding intermediate metal hydride complexes. These
intermediate hydride complexes can be reactivated for propagation by insertion of a
monomer unit. Hydrogen may coordinate to the metal atom and make more metal
hydride complexes resulting in a lower propagation rate.

2. Hydrogen can cause reduction of ethylene concentration around active sites because
of its characteristics (hydrogenation of olefinic monomer).


Polymer Science & Engineering Course
Chapter 2: Polyolefins 37
1.2 Polymerization temperature
1.2.1 Propylene
For propylene, a different dependence of polymerization rate on temperature was noticed
for the various catalytic systems. For TiCl
3
catalysts the rate does increase with
increasing temperature. For slurry polymerization of propylene with a -TiCl
3
3 1 AlCl
3
-
AlEt
2
Cl catalyst, both polymerization yield and average reaction rate increase in the
temperature range 30 C to 90 C.
For MgCl
2
-supported catalysts, the polymerization rate shows a distinct maximum in
the range 60 C to 70 C and then decreases with increasing temperature. This decreasing
rate at high temperatures points to catalyst deactivation either by overreduction of the
catalyst sites or through sorption theory. The rate of reaction is proportional to the
monomer concentration sorped by the amorphous regions of the polymer. The monomer
concentration in amorphous polymer decreases by temperature and, consequently, the
reaction rate decreases.
1.2.2 Ethylene
Similarly, temperature effect depends on catalyst type. It was found that, for some
metallocene catalysts, there is an increase in catalytic activity with temperature followed
by a decrease when the temperature exceeds certain value. The value of this temperature
depends on metallocene catalyst type.
1.3 Other variables
Many variables have an effect on catalytic activity, such as cocatalyst type and
concentration, donors and their concentration and polymerization time. Studying the
effect of these variables is out of the scope of this course.
2. MWD
2.1 Hydrogen effect
MWD of the produced polymer is described by: (i) number/weight average molecular
weight, (ii) width of the distribution, which is described usually by polydispersity index
(PDI). The following discussion will consider these two parameters separately.
2.1.1 Propylene
Commercially, the most important method of controlling molecular weight is by the
use of hydrogen as a chain transfer agent. Nowadays, it is widely accepted that the
Polymer Science & Engineering Course
Chapter 2: Polyolefins 38
number average molecular weight of the produced polypropylene depends on hydrogen
concentration as described by Equation (40)

2
H 2 1
n
P K K
1
M
+
= (40)

However, a lot of research work developed different correlations. Thus, it could be
stated that, hydrogen effect on polymer molecular weight depends on: (i) polymerization
temperature, (ii) catalyst system used, and (iii) the concentration of hydrogen in the
system.
Increasing levels of hydrogen result in a somewhat larger polydispersity index,
because of the activation of site types that are dormant in the absence of hydrogen.
2.1.1 Ethylene
As a chain transfer agent, hydrogen decreases polyethylene molecular weight.
Hydrogen effect can be described by Equation (40). The relationship between these
variables can be described by other correlations as well.
The effect of hydrogen on the molecular weight distributions of the produced
polyethylenes is rather complex and depend on catalyst type.
2.2 Temperature
Polydispersity decreases with increasing polymerization temperature. Indeed some
authors reported this phenomenon both for polyethylene and polypropylene using
heterogeneous Ziegler-Natta catalyst systems. This could be explained by the increase in
activation energy with isospecificity of the active center. This leads to a greater increase
in the productivity of the highly isospecific sites compared to the weakly isospecific ones
as the polymerization temperature is increased.
Most of the works, for ethylene and propylene, show a decrease in polymers molecular
weight with increasing polymerization temperature. The higher activation energy of the
chain transfer reactions is the reason adopted by researches.

S-ar putea să vă placă și