6.1 Polyethylene and Related Polymers
Polyethylene is produced commercially in very
large quantities in many parts of the world. The monomer can be
synthesized from various sources. Today, however, most of ethylene
comes from petroleum by high temperature cracking of ethane or
gasoline fractions. Other potential sources can probably be found,
depending upon the availability of raw materials.
Two main types of polyethylene are manufactured
commercially. These are low (0.92–0.93 g/cm3) and
high (0.94–0.97 g/cm3) density polymers. The
low-density material is branched while the high-density one is
mostly linear and much more crystalline. The most important
applications for the low-density polyethylene are in films, sheets,
paper, wire and cable coatings, and injection molding. The
high-density material finds use in blow molded objects and in
injection molding.
6.1.1 Preparation of Polyethylene by a Free-Radical Mechanism
Up to the late 1960s, most low-density
polyethylene was produced commercially by high-pressure
free-radical polymerization. Much of this has now been replaced by
preparation of copolymers of ethylene with α-olefins by
coordination polymerization. These preparations are discussed
further in this chapter. High-pressure polymerizations of ethylene,
however, might still practiced in some places and it is, therefore,
discussed here. The reaction requires a minimum pressure of
500 atm [1] to proceed. The
branched products contain long and short branches as well as
vinylidene groups. With an increase in pressure and temperature of
polymerization, there is a decrease in the degree of branching and
in the amount of vinylidene groups [2, 3].
Free-radical commercial polymerizations are
conducted at 1,000–3,000 atm pressure and 80–300°C. The
reaction has two peculiar characteristics: (1) a high exotherm and
(2) a critical dependence on the monomer concentration. In
addition, at these high pressures oxygen acts as an initiator. At
2,000 atm pressure and 165°C temperature, however, the maximum
safe level of oxygen is 0.075% of ethylene gas in the reaction
mixture. Any amount of oxygen beyond that level can cause explosive
decompositions. History of polyethylene manufacture contains
stories of workers being killed by explosions. Yet, the oxygen
concentration in the monomer is directly proportional to the
percent conversion of monomer to polymer, though inversely
proportional to the polymer’s molecular weight. This limits many
industrial practices to conducting the reactions below
2,000 atm and below 200°C. These reactions were done,
therefore, between 1,000 and 2,000 atm pressures. Small
quantities of oxygen, limited to 0.2% of ethylene, are accurately
metered in [4, 5]. The conversion per each pass in continuous
reactors is usually low, about 15–20%.
There is an induction period that varies
inversely with the oxygen concentration to the power of 0.23.
During this period oxygen is consumed autocatalytically. This is
not accompanied by any significant decrease in pressure. A high
concentration of ethylene is necessary for a fast rate of chain
growth, relative to the rate of termination. Also, high
temperatures are required for practical rates of initiation.
If oxygen is completely excluded and the pressure
is raised to between 3,500 and 7,750 atm, while using
relatively low temperatures of 50–80°C, linear polyethylene forms
[6]. The reactions take about
20 h. Various solvents can be used, like benzene, isooctane,
methyl, or ethyl alcohols. Higher ethyl alcohol concentrations and
low concentrations of the initiator result in higher molecular
weights. The products range from 2,000 for wax-like polymers to
4,000,000 for nearly intractable materials. Favorite free-radical
initiators for this reaction are benzoyl peroxide,
azobisisobutyronitrile, di-t-butylperoxydicarbonate,
di-t-butyl peroxide, and
dodecanoyl peroxide. Above conditions differ, however, from typical
commercial ones, because such high pressures and long reaction
times are not practical.
The actual commercial conditions vary, depending
upon location and individual technology of each company. Often,
tubular and multiple-tray autoclaves are used [7]. Good reactor design must permit dissipation
of the heat of polymerization (800–1,000 cal/g), with good
control over other parameters of the reaction. Tubular reactors are
judged as having an advantage over stirred autoclaves in offering
greater surface-to-volume ratios and better control over residence
time [7]. On the other hand, the
stirred autoclaves offer a more uniform temperature distribution
throughout the reactor.
The tubular reactors have been described as
consisting of stainless steel tubes between 0.5 and 1 in. in
internal and about 2 in. in the external diameters. The
residence time in these tubes is from 3 to 5 min, and they can
be equipped with pistons for pressure regulation. Pressure might
also be controlled by flow pulses to the reactor [8]. For the oxygen-initiated reactions, the
optimum conditions are [7]
0.03–0.1% oxygen at 190–210°C and 1,500 atm pressure. At this
pressure, the density of ethylene is 0.46 g/cm3.
This compares favorably with the critical density of ethylene that
is 0.22 g/cm3. Once the polymerization is
initiated, the liquid monomer acts as a solvent for the polymer.
Impurities, such as acetylene or hydrogen cause chain transferring
and must be carefully removed. In some processes, hindered phenols
are added in small quantities (between 10 and 1,000 ppm). This
has the effect of reducing long-chain branching and yields film
grade resins with better clarity, lower haze, and a reduced amount
of microgels. Also, diluents are used in some practices. Their main
purpose is to act as heat-exchanging mediums, but they can also
help remove the polymer from the reactor. Such diluents are water,
benzene, and ethyl or methyl alcohols. Sometimes, chain
transferring agents like carbon tetrachloride, ketones, aldehydes,
or cyclohexane might also be added to control molecular weight. The
finished product (polymer–monomer mixture) is conveyed to a
separator where almost all of the unreacted ethylene is removed
under high pressure (3,500–5,000 psi) and recycled. The
polymer is extruded and palletized. Ethylene conversion per pass is
a limiting factor on the economics. A tubular reactor is
illustrated in Fig. 6.1.

Fig.
6.1
Illustration of a tubular reactor
Polyethylene prepared in this way may have as
many as 20–30 short branches per 10,000 carbon atoms in the chain
[9] and one or two long-chain
branches per molecule, due to “backbiting” [10] (explained in Chap.
3):

The reaction results in predominantly ethyl and
butyl branches. The ratio of ethyl to butyl groups is roughly 2:1
[11, 12]. Chain transferring to the tertiary
hydrogens at the location of the short branches causes elimination
reactions and formation of vinylidene groups [13, 14]. This
mechanism also accounts for formation of low molecular weight
species.

Commercial grades of low-density polyethylene
vary widely in the number of short and long branches, average
molecular weights, and molecular weight distributions. M w/M n is between 20 and 50 for
commercial low-density materials. The short branches control the
degree of crystallinity, stiffness, and polymer density. They also
influence the flow properties of the molten material.
6.1.2 Preparation of Polyethylene by Coordination Mechanism
Low-density polyethylene can be prepared by
coordination polymerization through copolymerization of ethylene
with α-olefins. This is discussed in the section on copolymers of
ethylene. Finding catalytic systems that would allow formation of
amorphous, low-density polyethylene from the monomer alone by
low-pressure polymerization, however, is an economically worthwhile
goal. To this end, considerable research is being carried out to
develop such catalytic systems. Particular attention is given to
metallocenes and other single-site catalysts for olefin
polymerization. Originally, the metallocene catalysts were typical
metal complexes with two cyclopentadienyl or substituted
cyclopentadienyl groups. Many variations were developed since.
These materials are used in combination with methyl aluminoxane and
have the potential of forming the polymers with high precision.
Nevertheless, at this time it is probably still safe to say that
low-density polyethylene is prepared by many but perhaps not by all
of the processes and catalytic systems mentioned in this book. This
is because the material is manufactured all over the world and
different considerations govern the decisions on the processes and
catalytic systems. The same is probably true of high-density
polyethylene.
New catalysts based on palladium and nickel
complexes with bulky α-diimine ligands were developed
[17–19]. They can yield highly branched or
moderately branched polymers of ethylene, as well as propylene and
1-hexene. The polyolefins produced by such catalysts can contain a
considerable amount of branches along the backbone that are
randomly distributed throughout the molecules. In these catalysts,
the molecular weight-limiting β-hydrogen elimination process that
is common to palladium and nickel catalysts has been suppressed
through use of bulky α-diamine ligands [19]. This allows formation of high molecular
weight polymers from ethylene and α-olefins. A nickel-based
catalyst can be illustrated as follows:

It is claimed that the branching of polyethylene
can be controlled to the extent that the product can even be more
branched than conventional low-density polyethylene (1,2—300
branches/1,000 atoms) [18,
19]. The cationic Ni-diimine
catalyst shown above (R = H, CH3), with the
methylaluminoxane analog, has been found to polymerize ethylene in
toluene at room temperature at the rate of 110,000 kg/Ni/h.
This is comparable to the metallocene rates. The Pd-based catalysts
are less active than their Ni analogs [19].
When nickel catalysts are used, the extent of
branching is a function of the temperature, ethylene pressure, and
catalyst structure. Branching increases as the temperature rises.
At higher ethylene pressure less branching occurs. Brookhart et al.
illustrate the mechanism of polymerization as follows
[18]:
where X = CH3, Br; M = Pd, Ni;
R = (CH2CH2) n CH3;
Ar = 2,6-dialkylphenyl.

A patent for the polymerization process of
olefins (especially ethylene, α-olefins, cyclopentene, and some
fluorinated olefins) describes the above catalytic systems
[20]. The hindered diimines
stabilize alkyl Ni(II) or Pd(II) with cationic complexes. After
preparation, the complexes are reduced with methylaluminoxane and
then activated with Lewis acids capable of forming non-coordinating
counterions [20].
In addition, preparation of catalysts based on
iron and cobalt [21] was also
reported. These are complexes of bulky pyridine bis-imine ligands
with iron or cobalt that are also activated by methylaluminoxane:

The iron-based catalysts are reported to be
considerably more active than the cobalt analogs [21]. The yield of linear, narrow molecular
weight distribution polyethylene per gram is reported to be very
high [21].
Baugh et al. [22] synthesized and characterized a series of
nickel(II) and iron(II) complexes of the general formula
[LMX2] containing bidentate (for M = Ni) and tridentate
(for M = Fe) heterocycle-imine ligands. Activation of these
pre-catalysts with methyl aluminoxane yields active catalyst
systems for the oligomerization/polymerization of ethylene.
Compared to α-diimine nickel and bis(imino)pyridine iron catalysts,
both metal systems provide only half of the steric protection and
consequently the catalytic activities are significantly
lower.
Lower activities were attributed to reduced
stability of the active species under polymerization conditions.
The lower molecular weights of their products were explained to be
the result of increased hydrogen transfer rates. Variations within
the heterocyclic components of the ligand showed that both steric
and electronic factors influence polymerization behavior of such
catalysts.
Hanaoka, Oda, and coworkers report
[23] that single-site
polymerization catalysts are of considerable interest industrially
today, because they afford highly controllable polymerization
performances based on precise design of catalyst architecture and
their industrial applications. Among them, they point to
constrained geometry catalyst and phenoxy-induced complex, they
call phenics–Ti, that are used together with methyl aluminoxane

These are half-metallocene catalysts with an
anionic armed-pendant that have now been well developed for
industrial production of copolymers of ethylene with 1-olefins.
Modification at the cyclopentadienyl ring system has been mainly
tuned to finely control polymerization behaviors such as activity,
molecular weight, and regiochemistry. In general, minimizing
2,1-insertion is essential to obtain high molecular weight
polyolefins; otherwise, facile 6-elimination occurs, leading to
termination of chain growth. Thus, the largely open coordination
sites of half-metallocene catalyst systems possess an indispensable
problem of irregularity in propagation. Through tuning bulkiness of
substituents on the bridged-silicon unit of phenics–Ti, is claimed
to have demonstrated that 2,1-insertion of propylene can also
controlled by the bridging substituents to produce high molecular
weight polypropylene [23].
Hong and coworkers [24] concluded that it is generally desirable to
immobilize the single-site metallocene catalysts on a suitable
carrier to obtain ideal product morphology. Ultrahigh molecular
weight polyethylenes were successfully prepared by them through
titanium complexes bearing phenoxy-imine chelate ligands

They demonstrated that the catalyst can be
immobilized on silica. The product yields ultrahigh molecular
weight polyethylene. Increased polymerization temperature resulted
in higher activity, but lower molecular weight of
polyethylene.
6.1.3 Commercial High-Density Polyethylene, Properties, and Manufacture
High-density polyethylene
(0.94–0.97 g/cm3) is produced commercially with two
types of catalysts:
1.
Ziegler–Natta type catalysts
2.
Transition metal oxides on various supports
The two catalytic systems are used at different
conditions. Both types have undergone evolution from earlier
development. The original practices are summarized in
Table 6.1.
Table
6.1
Typical conditions for some preparation of
high-density polyethylene
Ziegler–Natta process
|
Cr2O3/support
|
MoO3/support
|
|
---|---|---|---|
Approx. temperature
|
75°C
|
140°C
|
234°C
|
Approx. pressure
|
60 psi
|
420 psi
|
1,000 psi
|
Usual state of the polymer in reaction
mixture
|
Suspension
|
Suspension
|
Suspension
|
The Ziegler process yields polyethylene as low as
0.94/cm3 in density, but process modifications can
result in products with a density of 0.965 g/cm3.
The transition metal oxide catalysts on support, on the other hand,
yield products in the density range of
0.960–0.970 g/cm3.
The original development by Ziegler led to what
appears to be an almost endless number of patents for various
coordination-type catalysts and processes. As described in
Chap.
4, such catalysts have been vastly improved.
Progress was made toward enhanced efficiency and selectivity. The
amount of polymer produced per gram of the transition metal has
been increased manyfold. In addition, new catalysts, based on
zirconium compounds complexed with methyl aluminoxane oligomers
(sometimes called Kominsky catalysts), were developed. They yield
very high quantities of polyethylene per gram of the catalyst. For
instance, a catalyst, bis(cyclopentadienyl)-zirconium dichloride
combined with methylaluminoxane, is claimed to yield 5,000 kg
of linear polyethylene per gram of zirconium per hour
[14].
An important factor in the catalysts activity is
the degree of oligomerization of the aluminoxane moiety. The
catalytic effect is enhanced by increase in the number of
alternating aluminum and oxygen atoms. These catalysts have long
storage life and offer such high activity that they need not be
removed from the product, because the amount present is negligible
[14, 15]. This makes the work-up of the product
simple.
The continuous solution processes are usually
carried out between 120 and 160°C at
400–500 lb/in.2 pressure. The diluents may be
cyclohexane or isooctane. In one zone reactors, the solid catalyst
is evenly dispersed throughout the reactor. In the two zone
reactors (specially constructed), the polymerizations are conducted
with stirring in the lower zone where the catalysts are present in
concentrations of 0.2–0.6% of the diluent. Purified ethylene is fed
into the bottom portions of the reactors. The polymers that form
are carried with small portions of the catalyst to the top and
removed. To compensate for the loss, additional catalysts are added
intermittently to the upper “quiescent” zones.
In suspension or slurry polymerizations, various
suspending agents, like diesel oil, lower petroleum fractions,
heptane, toluene, mineral oil, chlorobenzene, or others, are used.
The polymerization temperatures are kept between 50 and 75°C at
only slightly elevated pressures, like 25 lb/in.2.
If these are batch reactions, they last between 1 and 4 h. The
slurry reactor is illustrated in Fig. 6.2.

Fig.
6.2
Commercial flow reactor for slurry
polymerization of ethylene with Ziegler–Natta catalysts as
illustrated in a British patent # 826 523
Polymerizations catalyzed by transition metal
oxides on support were described variously as employing
solid/liquid suspensions, fixed beds, and solid/gas-phase
operations. It appears, however, that the industrial practices are
mainly confined to use of solid/liquid suspension processes. The
polymerization is carried out at the surface of the catalyst
suspended in a hydrocarbon diluent.
In continuous slurry processes, the temperatures
are kept between 90 and 100°C and pressures between 400 and
450 lb/in.2. The catalyst concentrations range
between 0.004 and 0.03% and typical diluents are n-pentane and n-hexane. Individual catalyst particles
become imbedded in polymer granules as the reaction proceeds. The
granules are removed as slurry containing 20–40% solids.
There are variations in the individual processes.
In some procedures, the temperature is kept high enough to keep the
polymer in solution. In others, it is kept deliberately low to
maintain the polymer in slurry. The products are separated from the
monomer that is recycled. They are cooled, precipitated (if in
solution), and collected by filtration or centrifugation.
Various reactors were developed to handle
different slurry polymerization processes. The slurry is maintained
in suspension by ethylene gas. The gas rises to the top and
maintains agitation while the polymer particles settle to the
bottom where they are collected.
Several companies adopted loop reactors. These
are arranged so that the flowing reactants and diluents
continuously pass the entrance to a receiving zone. The heavier
particles gravitate from the flowing into the receiving zone while
the lighter diluents and reactants are recycled. To accommodate
that, the settling area must be large enough for the heavy polymer
particles to be collected and separated.
In addition to suspension, a gas-phase process
was developed. No diluent is used in the polymerization step.
Highly purified ethylene gas is combined continuously with a
dry-powdery catalyst and then fed into a vertical fluidized bed
reactor. The reaction is carried out at 270 psi and 85–100°C.
The circulating ethylene gas fluidizes the bed of growing granular
polymer and serves to remove the heat [15]. Formed polymer particles are removed
intermittently from the lower sections of the vertical reactor. The
product contains 5% monomer that is recovered and recycled. Control
of polymer density is achieved by copolymerization with α-olefins.
Molecular weights and molecular weight distributions are controlled
by catalyst modifications, by varying operating conditions, and/or
use of chain transferring agents [15], such as hydrogen [16]. This is illustrated in
Fig. 6.3.

Fig.
6.3
Illustration of a gas-phase process (from
Burdett, by permission of the American Chemical Society)
The reactors for the fluidized gas-phase process
are simple in design. There are no mechanical agitators and they
rely upon blowers to keep the bed fluidized and well mixed.
Catalysts and cocatalysts are fed directly to the reactor
[25].
Rieger and coworkers [26] investigated gas-phase polymerization of
ethylene with supported α-diimine nickel catalysts. The reaction of
2,5 and 2,6 and 1,4 dithiane ligands with Ni(acac)2 and
trityl tetrakis(pentafluorophenyl)borate gave the corresponding
Ni(II) complexes in high yields. These complexes were supported on
silica without a chemical tether and were used as catalysts for
ethylene polymerization reactions in the gas phase. Furthermore,
ethylene was polymerized with the unsupported 2,5-complexes in
homogeneous solution for comparison. The influence of the ligand
structure, hydrogen, and temperature on the polymerization
performance was investigated. The supported catalysts showed
moderate to high activities and produced polyethylenes ranging from
high-density polyethylene to linear low-density polyethylene,
without further addition of a α-olefin comonomer.

The weight average molecular weights of most
commercial low- and high-density polyethylenes range between 5,000
and 300,000. Very low molecular weight polyethylene waxes and very
high molecular weight materials are also available. The molecular
weight distributions for high-density polyethylene vary between 4
and 15. The product generally has fewer than three branches per
thousand carbon atoms [9].
Table 6.2 summarizes the properties of various
polyethylenes.
Table
6.2
Properties of commercial polyethylene
Properties
|
Free-radical polymerization
|
Ziegler–Natta type catalysts
|
Metal oxides on support
|
---|---|---|---|
Density
|
0.92–0.93 g/cm3
|
0.94 g/cm3
|
0.95–0.96 g/cm3
|
Melting point
|
108–110.7°C
|
129–131°C
|
136°C
|
% Amorphous
|
43.1
|
25.8
|
25.8
|
Structure
|
20–30 ethyl and butyl branches/1,000
carbons, a few long branches
|
Mainly linear 7 ethyl branches/1,000
carbons
|
Almost linear
|
Double bonds
|
0.6–2/1,000 carbons
|
0.1–1/1,000 carbons
|
Up to 3/1,000 carbons
|
Types of bonds
|
15% terminal vinyl
|
43% terminal
|
94% terminal
|
68% vinylidene
|
32% vinylidene
|
1% vinylidene
|
|
17% internal trans olefin
|
25% internal trans olefins
|
5% internal trans olefins
|
6.1.4 Materials Similar to Polyethylene
Materials that are quite similar to polyethylene
can be obtained from other starting materials. The most prominent
is formation of polymethylene and similar high molecular weight
paraffin hydrocarbons from diazoalkanes. The reaction was
originally carried out by Pechmann [27] when small quantities of a white flocculent
powder formed in an ether solution of diazomethane. Bamberger and
Tschirner [28] showed that this
white powder is polymethylene –(–CH2–) n – that melts at 128°C. The
synthesis was improved since by introduction of various catalysts.
The reaction can yield highly crystalline polymers that melt at
136.5°C [29] with the molecular
weight in millions [30]. Among the
catalysts, boron compounds are very efficient [30]. Bawn et al. [31] postulated the mechanism of catalytic
action. It consists of initial coordination of a monomer with the
initiator, BF3. This is followed by a loss of nitrogen
and a shift of a fluorine atom from boron to carbon. The successive
additions of molecules of diazoalkane follow a similar path with a
shift of the chain fragment to the electron-deficient carbon:

The resulting macromolecules are still reactive
toward additional diazoalkanes. The above step-growth
polymerization reactions can also yield block copolymers:

Formation of polymethylene by this reaction is
not practical for commercial utilization.
Colloidal gold and fine copper powder also
catalyze diazoalkane polymerizations. The reaction appears to
precede by formation of alkylidine or carbene species that are
bound to the surfaces of metals [31–33]. The
initiations are completed by additions of diazoalkanes to the bound
carbenes followed by liberation of nitrogen. Termination may take
place by chain transfer, perhaps to a monomer, or to the solvent
[31–33].
Many different diazoalkanes lend themselves to
these polymerization reactions. Polypentylidine, polyhexylidine,
polyheptilidine, and polyoctylidine form with a gold complex
catalyst, AuCl3-pyridine [34].
An entirely different route to preparation of
macroparaffins is through a high-pressure reaction between hydrogen
and carbon monoxide. Transition metals, like finely divided
ruthenium, catalyze this reaction. At pressures of about
200 atm and temperatures below 140°C, polymethylene of
molecular weight as high as 100,000 forms [35]:

6.2 Polypropylene
Propylene monomer, like ethylene, is obtained
from petroleum sources. Free-radical polymerizations of propylene
and other α-olefins are completely controlled by chain transferring
[36]. They are, therefore,
polymerized by coordination polymerization. At present, mainly
isotactic polypropylene is being used in large commercial
quantities. Also, there is some utilization of atactic
polypropylene as well. Syndiotactic polypropylene, on the other
hand, is still mainly a laboratory curiosity.
The polypropylene that was originally described
by Natta contained less than 50% of isotactic fractions. The
remainder was atactic material. Some stereoblocks composed of
isotactic and atactic polypropylenes were also formed. This type of
product forms when α-olefins are polymerized in inert hydrocarbons
with catalysts prepared by reducing high valence metal compounds,
like TiCl4, with organometallic compounds like
Al(C2H5) prepared by reducing high valence
metal compounds, like TiCl4, with organometallic
compounds like Al(C2H5)3.
Later heterogeneous highly crystalline catalysts
based on transition metals (valence 3 or less) like
TiCl2, TiCl3, ZrCl3, and
VCl3 were developed that yielded stereospecific
polypropylene. The metal halides were combined with selected metal
alkyls. Only those alkyls were picked that would not destroy the
crystalline lattice of the transition metal salts in the process of
the reaction. The resultant catalysts yielded crystalline
polypropylenes with high fractions of the isotactic material. The
products, however, also contained some low molecular weight
fractions, some amorphous and stereoblock materials, that still
required costly purification and separations to obtain relatively
pure isotactic polypropylene. The atactic polymer is a wax-like
substance that lacks toughness. Also, presence of amorphous
materials, or very low molecular weight compounds, causes tackiness
and impedes processing. Table 6.3 lists some of the
catalysts and the amounts of crystallinity in polymers that were
reported by Natta et al. [37]. To
avoid costly purification of isotactic polypropylene,
three-component catalyst systems were developed. Some of the
original ones appear to have been reported by Natta, himself, who
found that addition of Lewis bases enhances the quantity of the
crystalline material. Table 6.4 shows the effects of addition of Lewis bases
on the amount of crystallinity, reported by Natta et al.
[38].
Table
6.3
Polypropylenes prepared by Natta
[37]
Transition metal halide
|
Metal alkyl halide
|
% Crystallinity
|
---|---|---|
TiCl3 (β)
|
Al(C2H5)3
|
40–50
|
TiCl3 (α, γ, or δ)
|
Al(C2H5)3
|
96–98
|
TiCl3 (α, γ, or δ)
|
Al(C2H5)2CI
|
96–98
|
TiCl3 (α, γ, or δ)
|
Be(C2H5)2
|
94–96
|
TiCl3 (α, γ, or δ)
|
Mg(C2H5)2
|
78–85
|
TiCl3 (α, γ, or δ)
|
Zn(C2H5)2
|
30–40
|
VCl3
|
Al(C2H5)3
|
73
|
TiCl2
|
Al(C2H5)3
|
75
|
Table
6.4
Effect of addition of Lewis bases on the
amount of crystalline fraction in polypropylene
Transition metal halides
|
Aluminum alkyl
|
Lewis base
|
% Crystallinity
|
---|---|---|---|
TiCl3
|
2Al(C2H5)Br2
|
Pyridine
|
>98.5
|
TiCl3
|
2Al(C2H5)Cl2
|
N(C2H5)3
|
95
|
TiCl3
|
2Al(C2H5)Cl2
|
NH(C2H5)2
|
93
|
TiCl3
|
2Al(C2H5)Br2
|
N+(C4H9)4I−
|
>99
|
TiCl3
|
2Al(C2H5)Cl2
|
N+(C4H9)4Br−
|
96
|
Many other three-component systems were developed
since [39–43]. Also, development of more active catalysts
[44, 45] eliminates a need to remove them from the
finished product [15]. The first
improvement in catalyst productivity came from treating
TiCl3 (formed from TiCl4 and
Al(C2H5)Cl2) with aliphatic ethers
resulting in yields of 520 g of polymer for each gram of Ti
[46]. Further improvement was
achieved by supporting TiCl3 on MgCl2 or by
producing a supported catalyst by reacting TiCl4 with
Mg(OC2H5) or with other magnesium compounds.
This raised the productivity to over 3,000 g of polymer for
every gram of Ti [46]. The
products, however, contained low percentages of the isotactic
isomer (20–40%). Addition of a Lewis base like N,N,N′,N′-tetramethyl ethylenediamine in solid
component and ethyl benzoate in solution raised the isotactic
content to 93% with a productivity of 2,500 g of polymer per
gram of Ti [41]. Claims are made
today for much greater catalyst activity. It was reported, for
instance, that catalyst efficiencies of 40 kg of polymer per
1 g of Ti can be achieved. Such yields require proper choice
of catalysts and control over polymerization conditions. The
isotactic fractions in the products are reported to range from 95
to 97% [47–49].
In a catalyst system
TiCl3/MgCl2/C6H5COOC2H5/Al(C2H5)3,
the high activity was initially attributed to higher propagation
rates rather than to an increase in the concentration of the active
sites [50]. The higher activity of
these catalysts, however, was shown instead to be due to higher
numbers of active centers and only slightly higher values of
K P
[51]. Subsequent trends in
modifications of supported Ziegler–Natta catalysts consisted of
using sterically hindered amines [52–54]. For
instance, 2,2,6,6-tetramethylpiperidine might be used together with
different trialkylaluminum compounds as modifier-cocatalyst systems
for the supported catalysts:
where X represents a halogen.

Other analogous amines, like
1,2,4-trimethylpiperazine and 2,3,4,5-tetraethylpiperidine, are
also used in preparations of titanium halide catalysts supported on
MgCl2. The amine remains as a built-in modifier in the
catalyst system [53].
Subsequent research efforts concentrated on
soluble catalytic systems, like
di-η5-cyclopentadienyldiphenyltitanium and
tetrabenzylzirconium complexed with methylaluminoxane,
(CH3)2Al– [–O–Al(CH3)–]
n
–Al(CH3)2. Such catalysts, however, yield
products that contain only about 85% isotactic polypropylene
[55–61], and only if the reactions are conducted at
low temperatures, −45°C or lower. A major breakthrough occurred
when rigid chiral metallocene initiators were developed, like
1,1-ethylene-di-η5-indenylzirconium dichloride,
complexed with methylaluminoxane. In place of zirconium, titanium
and hafnium analogs can also be used. These catalysts are highly
isospecific [62–64] when used at low temperatures. The compounds
are illustrated in Chap.
4. Typical catalysts consist of aluminum to
transition metal ratios of 103 or 104:1. Many of them yield 98–99%
isotactic fractions of the polymer. In addition, these are very
active catalysts, yielding large quantities of polymer per gram of
zirconium.
It was also reported that elastomeric
polypropylenes can be formed from the monomer with the aid of some
metallocene catalysts [62–64]. Because
rigid, chiral metallocene catalysts produce isotactic
polypropylene, while the achiral ones produce the atactic form,
Waymouth and Coates [62] prepared
a bridged metallocene catalyst with indenyl ligands that rotate
about the metal–ligand bond axis. The rotation causes the catalyst
to isomerize between chivalric and nonchiralic geometries:

Indenyl ligands, however, were found to rotate
faster that the polymerization reaction. This prevents formation of
stereoregular polymer blocks [62].
To overcome that, phenyl substituents were added to the ligands to
slow down the rotation below the speed of monomer insertion, yet
rotate faster than the time required for formation of the whole
polymeric chain. The product, a catalytic system of
bis(2-phenylindenyl)zirconium dichloride plus methylaluminoxane,
was found to yield elastomeric block copolymers of isotactic and
atactic polypropylene [62].
Vincenzo et al. [63] reported that 13C NMR
microstructural analysis of polypropylene samples produced with two
representative “oscillating” metallocene catalysts was found to be
largely different in steric hindrance. The original mechanistic
proposal of an “oscillation” between the two enantiomorphous, a
racemic-like (isotactic-selective) and a meso-like
(non-stereoselective) conformation, according to them, cannot
explain the observed polymer configuration.
They further feel that isotactic-stereoblock
nature of the polymers obtained with this catalyst proves
unambiguously that the active cation “oscillates” between the two
enantiomorphous racemic-like conformations at an average frequency
that, even at high propene concentration, is only slightly lower
than that of monomer insertion. The less hindered catalyst gives
instead a largely stereoirregular polypropylene, which is the
logical consequence of a faster ligand rotation; however, depending
on the use conditions (in particular, on the nature of the
cocatalyst and the polarity of the solvent), the polymerization
products may also contain appreciable amounts of a fairly isotactic
fraction. The peculiar microstructure of this fraction, with
isotactic blocks of the same relative configuration spanned by
short atactic ones, rules out the possibility that the latter are
due to an active species in meso-like conformation and point rather
to a conformationally “locked” racemic-like species with restricted
ring mobility. The hypothesis of a stereorigidity induced by the
proximity to a counter anion, which would play the role of the
inter-annular bridge in the racemic-bis(indenyl) ansa-metallocenes,
was tested by computer modeling and found viable.
Preparation of elastomeric polypropylenes was
also reported by Chien et al. [64]. Two metallocene catalysts of different
stereospecificities were used. The isospecific catalyst precursors
were either rac-ethylene
bis-(1-η5-indenyl)zirconium dichloride or rac-dimethylsilylene
bis(1-η5-indenyl)zirconium dichloride. The unspecific
one was ethylene bis(9-η5-fluorenyl)zirconium
dichloride. The precursors were activated with triphenyl carbenium
tetrakis(pentafluorophenyl)borate and triisobutylaluminum. The
resultant catalysts exhibit very high activity, yielding products
that range from tough plastomers to weak elastomers [64].
6.2.1 Manufacturing Techniques
The earliest commercial methods used slurry
polymerizations with liquid hydrocarbon diluents, like hexane or
heptane. These diluents carried the propylene and the catalyst.
Small amounts of hydrogen were fed into the reaction mixtures to
control molecular weights. The catalyst system consisted of a deep
purple or violet-colored TiCl3 reacted with diethyl
aluminum chloride. The TiCl3 was often prepared by
reduction of TiCl4 with an aluminum powder. These
reactions were carried out in stirred autoclaves at temperatures
below 90°C and at pressures sufficient to maintain a liquid phase.
The concentration of propylene in the reaction mixtures ranged
between 10 and 20%. The products formed in discrete particles and
were removed at 20–40% concentrations of solids. Unreacted monomer
was withdrawn from the product mixtures and reused. The catalysts
were deactivated and dissolved out of the products with alcohol
containing some HCl, or removed by steam extraction. This was
followed by extraction of the amorphous fractions with hot liquid
hydrocarbons.
Later bulk polymerization processes were
developed where liquid propylene was either used as the only
diluent in a loop reactor or permitted to boil out to remove the
heat of reaction. The second was done in stirred vessels with vapor
space at the top. More recently, gas-phase polymerizations of
propylene were introduced. The technology is similar to the
gas-phase technology in ethylene polymerizations [15] described in Sect. 6.1.
6.2.2 Syndiotactic Polypropylene
Isotactic polypropylene received most attention
because it is commercially more desirable. Nevertheless,
syndiotactic polypropylene, though less crystalline, has greater
clarity, elasticity, and impact resistance. It melts, however, at
lower temperature. This isomer was originally prepared with both,
heterogeneous, titanium-based catalysts and soluble, vanadium-based
ones. The heterogeneous catalysts gave very low yields of the
syndiotactic fractions. In fact, original samples contained only a
few percent of the desired material, almost an impurity. The yield
of syndiotactic polypropylene increased with a decrease in
polymerization temperature, but still remained low [65].
Highly syndiotactic polypropylene was prepared by
Natta et al. [38] with homogeneous
catalysts formed from VCl4 or from vanadium
tri-acetylacetonate, aluminum dialkyl halide, and anisole at −48 to
−78°C. No isotactic fractions formed. This led to development of
many effective soluble catalysts. The catalyst components and the
conditions for their preparation are quite important in maintaining
control over syndiotactic placement. For the most effective soluble
catalyst the ratio of AIR2X to the vanadium compound has
to be maintained between 3 and 10 [66]. The organic portion of the organoaluminum
compound can be either methyl, ethyl, isobutyl, neopentyl, phenyl,
or methylstyryl [9, 67]. In addition to VCl4 and to
vanadium tri-acetylacetonate [66],
various other vanadates can be used, like [VO(OR) x Cl3−x ], where x = 1, 2, or 3 [65]. The exact nature of the vanadium compound,
however, is very important to the resultant steric arrangement of
the product. For instance, VCl4 combined with
Al(C2H5)2F forms a heterogeneous
catalyst that yields the isotactic isomer [65]. Vanadium tri-acetylacetonate, on the other
hand, upon reacting with
Al(C2H5)2F forms a soluble
catalyst that yields the syndiotactic isomer [66]. Addition of certain electron donors
increases the amount of syndiotactic placement. These are anisole,
furan, diethyl ether, cycloheptanone, ethyl acetate, and thiophene
[67]. The optimum results are
obtained when an anisole to vanadium ratio is 1:1. Also, the
highest amount of syndiotactic polymer is obtained when the soluble
catalysts are prepared and used at low temperatures. Even at low
temperatures, however, like −78°C, the amount of syndiotacticity
that can be obtained with a specific catalyst decreases with time
[65, 66, 68]. This
indicates a deterioration of the syndiotactic placing sites. On the
other hand, polymerization of propylene with soluble vanadium
tri-acetylacetonate–Al(C2C5)2Cl
system was reported to be a “living” type polymerization
[69]. The product has a narrow
molecular weight distribution (M w/M n = 1.05–1.20). A kinetic
study indicates an absence of chain transferring and termination at
temperatures below −65°C.
More recent catalysts for syndiotactic
polypropylene are complexes, like i-propyl(cyclopentadienyl-1-fluorenyl)hafnium
dichloride with methyl aluminoxane [70]. Another, similar catalyst is i-propyl(η5-cyclopentadienyl-η3-fluorenyl)zirconium
dichloride with methyl aluminoxane. These catalysts yield polymers
that are high in syndiotactic material (the zirconium-based
compound yields 86% of racemic pentads) [70, 71].
Commercial production of syndiotactic polypropylene is in the early
stages. What catalytic system is used, however, is not disclosed at
this time. Some of the properties of the two isomers, isotactic and
syndiotactic polypropylenes, are compared in
Table 6.5.
Table
6.5
Comparison of isotactic and syndiotactic
polypropylenes
Typical
|
|||||
---|---|---|---|---|---|
Isomer
|
Crystal structure
|
Density at 25°C
|
MP (°C)
|
M
w
|
M
n
|
Isotactic
|
Monoclinic
|
0.92–0.43 g/cm3
|
171–186
|
220–700 K
|
38–160 K
|
Triclinic
|
0.943 g/cm3
|
||||
Hexagonal
|
|||||
Syndiotactic
|
Orthorhombic
|
0.89–0.91 g/cm3
|
138
|
6.3 Polyisobutylene
The original commercial methods for preparing
high molecular weight polyisobutylene by cationic polymerization in
good yields were reported in 1940. The reaction was carried out at
−40 to −80°C in a diluent with BF3 catalysis
[72]. This developed into current
commercial practices of polymerizing isobutylene at −80 to −100°C,
using liquid ethylene or methyl chloride as a diluent
[73, 74]. Even at these low temperatures the reaction
is quite violent. Methods were developed, therefore, to dissipate
the heat. In one of them, called “flash polymerization process” the
catalyst (a Lewis acid, like BF3 or AlCl3,
for instance) is added in solution to the cooled isobutylene
solution. The polymerization takes place very rapidly and is
complete in a few seconds with the heat of the reaction being
removed by vaporization of the diluent. Such reactions, however,
are very difficult to carry out in conventional batch reactors. Two
types of procedures were, therefore, adopted [75]. The first one is built around a moving
stainless steel belt contained inside a gas-tight reactor housing.
Isobutylene and liquid ethylene from one source and a Lewis acid in
ethylene solution (0.1–0.3% based on monomer) from another source
are fed continuously onto the moving belt where they are mixed and
moved. The movement of the belt is adjusted at such a speed that
the polymerization is complete before the polymer arrives at the
end of its travel, where it is removed with a scraper and further
processed.
In the second process, the polymerization is
carried out in multiple kneaders or mixers. These are arranged in a
series of descending steps. Here the reaction mixture is carried
from one kneader to another with the temperature being raised at
each station and completed at the last one.
All commercially important polyisobutylenes are
linear, head to tail polymers, with tertiary butyl groups at one
end of the chains and vinylidene groups at the other:

The differences lie in molecular weights. They
range from 2,000 to 20,000 for viscous liquids to between 100,000
and 400,000 for high molecular weight elastomers that resemble
unmilled crepe rubber. The polymers degrade readily from thermal
abuse. They can be stabilized effectively, however, by adding small
quantities (0.1–1.0%) of such stabilizers as aromatic amines,
phenols, or sulfur compounds. Polyisobutylenes are soluble in many
hydrocarbons and are resistant to attacks by many chemicals.
Coordination polymerizations with Ziegler–Natta
catalysts yield similar polymers that range from viscous liquids to
rubbery solids. At 0°C, a catalyst with a 1:16 Ti to Al molar ratio
yields a polymer with a molecular weight of 5,000–6,000
[76]. The molecular weight,
however, is dependent upon the reaction time. This contrasts with
polymerizations of ethylene, propylene, and 1-butene by such
catalysts, where the molecular weights of the products are
independent of the reaction time. In addition, there are some
questions about the exact molecular structures of the products
[76].
Bochmann and coworkers [77] carried out polymerizations of isobutylene
and copolymerizations with isoprene using cationic zirconocene
hydride complexes. The combination of [Cp2ZrH] with
various trityl salts of weakly coordinating anions gives binuclear
cationic hydrides
[Cp′4Zr2H(μ-H)2]+X−
which are powerful initiators for the polymerization of isobutene
and its copolymerization with isoprene. The temperature dependence
of M is indicative of a cationic mechanism. The highest molecular
weights are obtained only under scrupulously dry conditions.
High molecular weight polyisobutylene has fair
tensile strength but suffers from the disadvantage of considerable
cold flow. A copolymer of isobutylene with some isoprene for
cross-linking is, therefore, used as a commercial elastomer and
called “butyl rubber.” The isoprene is present in the copolymer in
only minor proportions (1.4–4.5%). The uncross-linked material is
very similar to polyisobutylene. Copolymers of isobutylene with
other dienes are also called butyl rubbers. They can also be
terpolymers, where the third component may be cyclopentadiene for
improved ozone resistance.
The molecular weights of the copolymers vary
inversely with the quantities of isoprene incorporated, the
polymerization temperatures, and amount of impurities present
during polymerization. Impurities like n-butene or water act as chain
transferring agents [79].
To maintain uniform molecular weights, the
conversions are usually kept from exceeding 60%.
6.4 Poly(α-olefin)s
Many α-olefins were polymerized by the
Ziegler–Natta catalysts to yield high polymers and many such
polymers were found to be stereospecific and crystalline.
Polymerizations of α-olefins of the general structure of
,
where x is 0–3 and R
denotes CH3, CH–(CH3)2,
C(CH3)3, or C6H5, can
be catalyzed by vanadium trichloride/triethyl aluminum
[80]. The conversions are fairly
high, though higher crystallinity can be obtained with
titanium-based catalysts [81].
Addition of Lewis bases, such as
(C4H9)2O,
(C4H9)3N, or
(C4H9)3P, to the catalyst system
further increases crystallinity [82].

6.4.1 Properties of Poly(α-olefin)s
Many poly(α-olefin)s reported in the literature
are not used commercially for various reasons.
Table 6.6
lists some of the olefins polymerized by the Ziegler–Natta
catalysts [72, 83].
Table
6.6
Properties of poly(α-olefin)s
Monomer
|
State
|
MP (°C)
|
---|---|---|
![]() |
Crystalline
|
136
|
![]() |
Crystalline
|
165–168
|
![]() |
Crystalline
|
124–130
|
![]() |
Crystalline
|
75
|
![]() |
Rubber, amorphous
|
–
|
![]() |
Some crystallinity
|
45
|
![]() |
Crystallinity in pendant groups
|
70; 100
|
![]() |
Crystalline, hard
|
240–285
|
![]() |
Crystalline, hard
|
200–240
|
![]() |
Crystalline, hard
|
300–350
|
![]() |
Crystalline, hard
|
350
|
![]() |
Crystalline, hard
|
160
|
![]() |
Rubber, amorphous
|
–
|
![]() |
Crystalline, slightly rubbery
|
158
|
![]() |
Crystalline, intractable
|
360
|
![]() |
Rubber, amorphous
|
–
|
6.4.2 Poly(butene-1)
Isotactic poly(butene-1) is produced commercially
with three-component coordination-type catalysts. It is
manufactured by a continuous process with simultaneous additions to
the reaction vessel of the monomer solution, a suspension of
TiCl2–AlCl3, and a solution of diethyl
aluminum chloride [84]. The
effluent containing the suspension of the product is continually
removed from the reactor. Molecular weight control is achieved
through regulating the reaction temperature. The effluent contains
approximately 5–8% of atactic polybutene that is dissolved in the
liquid carrier. The suspended isotactic fractions (92–98%) are
isolated after catalyst decomposition and removal. The product has
a density of 0.92 g/cm3 and melts at
124–130°C.
Isotactic polybutene crystallizes into three
different forms. When it cools from the melt, it originally
crystallizes into a metastable crystalline one. After several days,
however, it transforms into a different form. Noticeable changes in
melting point, density, flexural modulus, yield, and hardness
accompany this transformation. The third crystalline form results
from crystallization from solution. The polymer exhibits good
impact and tear resistance. It is also resistant to environmental
stress-cracking.
6.4.3 Poly(4-methyl pentene-1)
Another commercially produced polyolefin is
isotactic poly(4-methyl pentene-1). The polymer carries a trade
name of TPX. This material is known for high transparency, good
electrical properties, and heat resistance. Poly(4-methyl
pentene-1) has a density of 0.83 g/cm3. This
polyolefin exhibits poor load-bearing properties and is susceptible
to UV degradation. It is also a poor barrier to moisture and gases
and scratches readily. This limits its use in many
applications.
Poly(4-methyl pentene) is produced by the same
process and equipment as polypropylene. A post finishing de-ashing
step, however, is required. In addition, aseptic conditions are
maintained during manufacture to prevent contamination that may
affect clarity.
A number of similar polyolefins with pendant side
groups are known. These include poly(3-methyl butene-1),
poly(4,4-dimethyl pentene-1), and poly(vinyl cyclohexane). Due to
their increased cohesive energy, ability to pack into tight
structures, and the effect of increasing stiffness of the pendant
groups, some of these polymers have a high melting point. This can
be seen from Table 6.9. Many of these polymers, however, tend to
undergo complex morphological changes on standing. This can result
in fissures and planes of weakness in the structure.
6.5 Copolymers of Ethylene and Propylene
Many monomers have been copolymerized with
ethylene by a variety of polymerization methods. When ethylene is
copolymerized with other olefins, the resultant hydrocarbon
polymers have reduced regularity and lower density, lower softening
point, and lower brittle point.
Copolymers of ethylene and propylene are a
commercially important family of materials. They vary from
elastomers that can contain 80% ethylene and 20% propylene to
polypropylene that is modified with small amounts of ethylene to
improve impact resistance.
Metallocene catalysts can produce both random and
alternating copolymers of ethylene and propylene [85]. At present there does not appear to be any
commercial utilization of alternating copolymers. They were
reported to form in polymerizations catalyzed by bridged fluorenyl
catalysts [85].
6.5.1 Ethylene and Propylene Elastomers
The commercial ethylene-propylene rubbers typically
range in propylene content from 30 to 60%, depending upon intended
use. Such copolymers are prepared with Ziegler–Natta type
catalysts. Soluble catalysts and true solution processes are
preferred. The common catalyst systems are based on
VCl4, VOCl3, V(Acac)3,
VO(OR)3, VOCl(OR)2, VOCl2(OR),
etc. with various organoaluminum derivatives. The products are
predominantly amorphous. Polymerization reactions are usually
carried out at 40°C in solvents like chlorobenzene or pentane. The
resultant random copolymers are recovered by alcohol precipitation.
Because these elastomers are almost completely saturated,
cross-linking is difficult. A third monomer, a diene is, therefore,
included in the preparation of these rubbers that carry the trade
names EPTR or EPDM. Inclusion of third monomers presents some
problems in copolymerization reactions. For instance, it is
important to maintain constant feed mixtures of monomers to obtain
constant compositions. Yet, two of the three monomers are gaseous
and the third one is a liquid. Natta [86] developed a technique that depends upon
maintaining violent agitation of the solvent while gaseous monomers
were bubbled through the liquid phase. This was referred to as
“semi-flow technique.” The process allows the compositions of
gaseous and liquid phases to be in equilibrium with each other and
to be more or less constant [87].
Other techniques evolved since. All are designed to maintain
constant polymerization mixtures.
The vanadium-based catalyst systems deteriorate
with time and decrease in the number of catalytic centers as the
polymerizations progress. The rate of decay is affected by
conditions used for catalyst preparation, compositions of the
catalysts, temperature, solvents, and Lewis bases. It is also
affected by the type and concentration of the third monomer
[88–90]. Additions of chlorinated compounds to the
deactivated catalysts, however, help restore activity
[91, 92]. Catalyst decay can also be overcome by
continually feeding catalyst components into the polymerization
medium [93].
While third monomer can be a common diene, like
isoprene, more often it is a bridged ring structure with at least
one double bond in the ring. In typical terpolymer rubbers with
60–40 ratios of ethylene to propylene the diene components usually
comprise about 3% of the total. Some specialty rubbers, however,
may contain 10% of the diene or even more. Reaction conditions are
always chosen to obtain 1,2 placement of the diene. Dienes in
common use are ethylidine norbornene, methylene norbornene,
1,4-hexadiene, dicyclopentadiene, and cycloocatadiene:

In addition to the above, the patent literature
describes many other dienes. An idealized picture of a segment of
an uncross-linked gum stock might be shown as having the following
structure:

6.5.2 Copolymers of Ethylene with α-Olefins and Ethylene with Carbon Monoxide
Many copolymers of ethylene with α-olefins are
prepared commercially. Thus ethylene is copolymerized with
butene-1, where a comonomer is included to lower the regularity and
the density of the polymer. Many copolymers are prepared with
transition metal oxide catalysts on support. The comonomer is
usually present in approximately 5% quantities. This is sufficient
to lower the crystallinity and to markedly improve the impact
strength and resistance to environmental stress-cracking.
Copolymers of ethylene with hexene-1, where the hexene-1 content is
less than 5%, are also produced for the same reason.
In most cases, the monomers that homopolymerize
by Ziegler–Natta coordination catalysts also copolymerize by them
[94]. In addition, some monomers
that do not homopolymerize may still copolymerize to form
alternating copolymers. Because the lifetime of a growing polymer
molecule is relatively long (can be as long as several minutes),
block copolymerization is possible through changes in the monomer
feeds. Also, the nature of the transition metal compound influences
the reactivity ratios of the monomers in copolymerizations. On the
other hand, the nature of the organometallic compound has no such
effect [95]. It also appears that
changes in the reaction temperature between 0 and 75°C have no
effect on the r values.
Copolymers can be formed using either soluble or heterogeneous
Ziegler–Natta. One problem encountered with the heterogeneous
catalysts is the tendency by the formed polymers to coat the active
sites. This forces the monomers to diffuse to the sites and may
cause starvation of the more active monomer if both diffuse at
equal rates.
Many different block copolymers of olefins, like
ethylene with propylene and ethylene with butene-1, are
manufactured. Use of the anionic coordination catalysts enables
variations in the molecular structures of the products. It is
possible to vary the length and stereoregularity of the blocks.
This is accomplished by feeding alternately different monomers into
the reactor. When it is necessary that the blocks consist of pure
homopolymers, then after each addition the reaction is allowed to
subside. If any residual monomer remains, it is removed
[96]. This requires a long
lifetime for the growing chains and an insignificant amount of
termination. The stability of the anion depends upon the catalyst
system. One technique for catalyst preparation is to form
TiCl3 by reducing TiCl4 with diethyl aluminum
chloride followed by careful washing of the product of reduction to
remove the by-product Al(C2H5)Cl. Other
reports describe using the α-form of TiCl3 or heat
treating it to form the β or γ-forms that yields more
stereospecific products.
The transition metal oxide catalysts on support,
such as the CrO3/silica–alumina (Phillips) and
MoO3/Al2O3 (Standard Oil), are
used to copolymerize minor quantities of α-olefins with ethylene.
Such copolymerizations introduce short pendant groups into polymer
backbones.
Ethylene and other olefins can also be
copolymerized with carbon monoxide to form polymers of aliphatic
ketones, using transition metal catalysts, like palladium(II)
coupled with non-coordinating anions. There are numerous reports of
such catalysts in the literature. One example is a compound
composed of bidentate diarylphosphinopropane ligand and two
acetonitrile molecules coordinating Pd2+ coupled with
BF3 counterions. This compound,
bis(acetonitrile)palladium(II)-1,3-bis(diphenyl-phosphino)propane-(tetrafluoborate),
can be illustrated as follows [97]:

When the tetrafluoroborate is replaced with a
perchlorate, the compound is a very active catalyst [97].
One copolymer of ethylene and carbon monoxide are
available commercially. The material offered under the trade name
of Carilon is actually a terpolymer, because it contains a small
quantity of propylene. It is reported [98] that use of a palladium catalyst permits
formation of perfectly alternating interpolymer.08 The
product is reported to be a tough, chemical resistant
material.
Hustad et al. [99] developed a technique to make polydisperse
polyethylene diblock copolymers with 1-octene with a distribution
of block lengths. When melted and compressed into films, the
distinct polymeric segments self-assemble into layered patterns of
semi-crystalline and hard and amorphous phases. Because each phase
has a different refractive index, the block copolymer, shown below,
can function as a photonic crystal and scatter visible light.

Nomura and coworkers [100] studied copolymerization of ethylene with
various pentenes:
where n = 8, 12. The
polymerizations were carried out with titanium catalysts
illustrated below. Titanium compound were combined with methyl
aluminoxane:


Their results show that the monomer reactivities
are influenced not only by substituents on the olefins but also by
the nature of the catalytically active species.
Derlin and Kaminsky [101] reported copolymerizations of ethylene and
propylene with a sterically hindered monomer, 3-methyl-1-butene,
using titanium and zirconium metallocenes with methyl aluminoxane
cocatalyst.
Tritto and coworkers [102] reported that the complex
[Pd(k2-P,O-{2-(2-MeOC6H4)2P}C6H4SO3)Me(DMSO)]
was investigated as a single-component catalyst for the
copolymerization of ethylene with norbornene. The catalyst was
illustrated as follows:

The copolymers were obtained in very good yields
and molar masses were significantly higher than those of
polyethylene. Three copolymers were formed:

Determination of microstructure and reactivity
ratios revealed a strong inherent tendency to form alternating
copolymers.
Jordan [103]
described copolymerization of ethylene with vinyl ethers and with
vinyl fluoride. The catalyst used was (ortho-phospheno-arenesulfonate)PdMe(pyridine).
The reaction was illustrated as follows:

6.5.3 Copolymers of Propylene with Dienes
Although presently lacking industrial importance,
alternating copolymers can be made from propylene and butadiene
[104] and also from propylene and
isoprene [105]. Copolymers of
propylene and butadiene form with vanadium- or titanium-based
catalysts combined with aluminum alkyls. The catalysts have to be
prepared at very low temperature (−70°C). Also, it was found that a
presence of halogen atoms in the catalyst is essential
[75]. Carbonyl compounds, such as
ketones, esters, and others, are very effective additives. A
reaction mechanism based on alternating coordination of propylene
and butadiene with the transition metal was proposed by Furukawa
[104].
6.5.4 Copolymers of Ethylene with Vinyl Acetate
Various copolymers of ethylene with vinyl acetate
are prepared by free-radical mechanism in emulsion polymerizations.
Both reactivity ratios are close to 1.0 [106]. The degree of branching in these
copolymers is strongly temperature-dependent [107]. These materials find wide use in such
areas as paper coatings and adhesives. In addition, some are
hydrolyzed to form copolymers of ethylene with vinyl alcohol. Such
resins are available commercially in various ratios of polyethylene
to poly(vinyl alcohol), can range from 30% poly(vinyl alcohol) to
as high as 70%.
Vinyl acetate residues in ethylene–vinyl acetate
copolymers reduce regularity of polyethylene. This reduces
crystallinity in the polymer. Materials containing 45% vinyl
acetate are elastomers and can be cross-linked with
peroxides.
6.5.5 Ionomers
Another group of commercial copolymers of
ethylene is those formed with acrylic and methacrylic acids, where
ethylene is the major component. The copolymerizations are carried
out under high pressures. These materials range in comonomer
content from 3 to 20%. Typical values are 10%. A large proportion
of the carboxylic acid groups (40–50%) are prereacted with metal
ions like sodium or zinc. The copolymer salts are called ionomers
with a trade name like Syrlin. The materials tend to behave
similarly to cross-linked polymers at ambient temperature by being
stiff and tough. Yet they can be processed at elevated
temperatures, because aggregation of the ionic segments from
different polymeric molecules is destroyed. The material becomes
mobile but after cooling the aggregates reform. Ionomers exhibit
good low temperature flexibility. They are tough,
abrasion-resistant resins that adhere well to metal surfaces.
6.6 Homopolymers of Conjugated Dienes
Many different polymers of conjugated dienes are
prepared commercially by a variety of processes, depending upon the
need. They are formed by free-radical, ionic, and coordinated
anionic polymerizations. In addition, various molecular weights
homopolymers and copolymers, ranging from a few thousand for liquid
polymers to high molecular weight ones for synthetic rubbers, are
on the market.
6.6.1 Polybutadiene
1,3-Butadiene, the simplest of the conjugated
dienes, is produced commercially by thermal cracking of petroleum
fractions and catalytic dehydrogenation of butane and butene.
Polymerization of butadiene can potentially lead to three
poly(1,2-butadiene)s, atactic, isotactic, and syndiotactic and two
cis and trans forms of poly(1,4-butadiene).
This is discussed in Chaps.
3 and 4.
Free-radical polymerizations of 1,3-butadiene
usually result in polymers with 78–82% of 1,4-type placement and
18–22% of 1,2-adducts. The ratio of 1,4 to 1,2 adducts is
independent of the temperature of polymerization. Moreover, this
ratio is obtained in polymerizations that are carried out in bulk
and in emulsion. The ratio of trans-1,4 to cis-1,4 tends to decrease, however, as
the temperature of the reaction decreases. Polybutadiene polymers
formed by free-radical mechanism are branched because the residual
unsaturations in the polymeric chains are subjects to free-radical
attacks:

Should branching become excessive, infinite
networks can form. The products become cross-linked, insoluble, and
infusible. Such materials are called popcorn polymers. This phenomenon is
more common in bulk polymerizations. The cross-linked polymers form
nodules that occupy much more volume than the monomers from which
they formed and often clog up the polymerization equipment,
sometimes even rupturing it.
High molecular weight homopolymers of
1,3-butadiene formed by free-radical mechanism lack the type of
elastomeric properties that are needed from commercial rubbers.
Copolymers of butadiene, however, with styrene or acrylonitrile are
more useful and are prepared on a large scale. This is discussed in
another section.
6.6.1.1 Liquid Polybutadiene
Low molecular weight liquid homopolymers of 1,3
butadiene, also some liquid copolymers, find industrial uses in
many applications. These materials can range in molecular weights
from 500 to 5,000 depending upon the mode of polymerization. Liquid
polybutadienes formed by cationic polymerizations are high
trans-1,4 content. Such
materials find applications in industrial coatings. They are
usually prepared with Lewis acids in chlorinated solvents. When the
reactions are catalyzed by AlCl3 at −78°C, two types of
polymers form [108]. One is
soluble and the other is insoluble, depending upon the extent of
conversion. AlCl3, AlBr3, and
BF3–Et2O produce polymers with the same
ratios of trans-1,4 to 1,2
adducts. These range from 4 to 5. Polymerizations carried out in
ethylene chloride [108] catalyzed
by TiCl4 yield products with lower ratios of
trans-1,4 to 1,2 adducts.
The ratios of the two placements are affected by the solvents. They
are also affected by additions of complexing agents, such as
nitroethane and nitrobenzene [108]. The changes, however, are small.
Hydroxyl-terminated liquid polybutadienes are
prepared for reactions with diisocyanates to form elastomeric
polyurethanes (see Chap. 6). Such materials can be
prepared by anionic polymerizations as “living” polymers and then
quenched at the appropriate molecular weight. These polybutadienes
can also be formed by free-radical mechanism. The microstructures
of the two products differ, however, and this may affect the
properties of the finished products. To form hydroxyl-terminated
polymers by free-radical mechanism, the polymerization reactions
may be initiated by hydroxyl radicals from hydrogen peroxide.
A new approach to preparation of
hydroxyl-terminated liquid polybutadiene is to use a cyclic
monomer, 1,5-cyclooctadiene, a ruthenium metathesis catalyst (see
Chap.
5, Grubbs catalysts in section on metathesis ring
opening polymerization) and an acetate functionalized chain
transfer agent [109,
110]. The acetate-functionalized
chain transfer agent is cis-2-butene-1,4-diacetate. The
reaction can be carried out without a solvent and proceeds at 50°C
over 6 h under an inert gas purge [111]. The acetate protecting groups provide
compatibility with the ruthenium catalyst. Subsequent to
polymerization the acetate groups can be converted to hydroxyl end
groups with the aid of a base, like sodium methoxide.
Liquid polybutadienes that are high in 1,2
placement are also available commercially. These range from
reactive polymers containing approximately 70% of vinyl groups to
very reactive ones containing more than 90% of 1,2 units. The
materials are formed by anionic polymerization with either sodium
naphthalene, or with sodium dispersions, or with organolithium
initiators in polar solvents. Carboxyl group terminated liquid
polybutadienes are predominantly used as modifiers for epoxy resins
(Chap.
7). They are formed by anionic mechanisms in
solution with organolithium catalysts like diphenylethanedilithium,
butanedilithium, isoprenelithium, or lithium methylnaphthalene
complexes. Cyclohexane is the choice solvent. The reaction is
quenched with carbon dioxide to introduce the terminal carboxyl
groups.
6.6.1.2 High Molecular Weight Polybutadiene
High molecular weight polybutadiene homopolymers
are prepared commercially with anionic catalysts and with
coordination ones. Polybutadiene formed with sodium dispersions was
prepared industrially in the former USSR, and perhaps might still
be produced in that area today. This sodium-catalyzed polybutadiene
contains 65% of 1,2-adducts [112]. Many of the preparations by others,
however, utilize either alkyllithium or Ziegler–Natta type
catalysts prepared with titanium tetra iodide or preferably
containing cobalt.
Because high molecular weight polybutadiene can
be prepared by different catalytic systems, the choice of catalyst
is usually governed by the desired microstructure of the product.
Alfin catalysts yield very high molecular weight polymers with a
large amount of trans-1,4
structures. Both the molecular weight and microstructure can be
affected significantly, however, by variations in the Alfin
catalyst components. These can be the alkyl groups of the
organometallic compounds or alkoxide portions.
When butadiene is polymerized with lithium metal
or with alkyllithium catalysts, inert solvents like hexane or
heptane must be used to obtain high cis-1,4 placement (see Chap.
4). Based on 13C NMR spectra,
1,4-polybutadiene formed with n-butyllithium consists of blocks of
cis-1,4 units and
trans-1,4 units that are
separated by isolated vinyl structures [113]:

The quantity of such units in the above
polybutadienes is approximately 48–58% trans-1,4, 33–45% cis-1,4, and 7–10% 1,2 units
[112]. There is little effect of
the reaction temperatures upon this composition. As described in
Chap.
3, however, addition of Lewis bases has a
profound effect. Reactions in tetrahydrofuran solvent result in 1,2
placement that can be as high as 87%.
The microstructures of polybutadienes prepared
with Ziegler–Natta catalysts vary with catalyst composition. It is
possible to form polymers that are high either in 1,2 placement or
in 1,4 units. The catalysts and the type of placement are
summarized in Table 6.7.
Table
6.7
Microstructures of polybutadienes prepared
with some coordination catalysts
Microstructure (%)
|
|||
---|---|---|---|
Catalyst
|
Cis-1,4
|
Trans-1,4
|
1,2
|
TiI4/Al(C2H5)3
|
95
|
2
|
3
|
TiBr4/Al(C4H9)3
|
88
|
3
|
9
|
β-TiCl3/Al(C2H5)3
|
80
|
12
|
8
|
Ti(OC4H9)4/Al(2H5)3
|
–
|
–
|
99–100
|
Ti(OC6H5)4/R3Al
|
90–100
|
–
|
–
|
(π-Cycloocatadiene)2Ni,CF3CO2H
|
100
|
–
|
–
|
Bis(π-Crotyl NiCl)
|
92
|
–
|
–
|
Bis(π-Crotyl NiI)
|
–
|
94
|
–
|
CoCl2/Al(C2H5)2Cl
|
96–97
|
2.5
|
1–1.5
|
CoCl2/Al(C2H5)3
|
94
|
3
|
3
|
CoCl2/Al(C2H5)3/pyridine
|
90–97
|
–
|
–
|
Co-stearate/AlR2Cl
|
98
|
1
|
1
|
VCl3/AlR3
|
–
|
99
|
1
|
VOCl3/Al(C2H5)3
|
–
|
97–9
|
2–3
|
VCl4/AlCl3
|
–
|
95
|
–
|
Cr(C6H5CN)6/Al(C2H5)3
|
100
|
–
|
–
|
MoO2(OR)3/Al(C2H5)3
|
–
|
–
|
75
|
Butadiene can be polymerized with chromium oxide
catalyst on support to form solid homopolymers. The products,
however, tend to coat the catalyst within a few hours after the
start of the reaction and interfere with further polymerization.
Polybutadiene can also be prepared in the presence of molybdenum
catalyst promoted by calcium hydride. The product contains 80% of
1,4 units and 20% of 1,2 units. Of the 1,4 units, 62.5% are
cis and 37.5% are
trans [110].
Cobalt oxide on silica–alumina in the presence of
alkyl aluminum also yields high cis-1,4 structure polymers. An all 1,2
polybutadiene can be prepared with n-butyllithium modified with
bis-piperidino ethane. The atactic polymer can be formed in hexane
at −5 to +20°C temperature [111].
The 100% 1,2 placement was postulated to proceed according to the
following scheme [111]. First a
complex forms between the base and butyllithium:

The above complex reacts with butadiene to form a
new complex:

This is followed by insertion of butadiene into
the carbon–lithium bond:

Annunziata et al. [114] reported that Group 4 metals complexes
bearing anilidomethylpyridine ligands were prepared by them. After
activation by AlBu2H and methylalumoxane, the catalysts
were tested in 1,3-butadiene and α-olefin polymerization. The
zirconium complexes showed higher activity than the titanium
analogous. Polymerization of ethylene resulted in all cases in the
production of high molecular weight linear polyethylene. On the
other hand, propylene polymerization tests provided substantially
atactic polypropylene. 1,3-Butadiene polymerizations produced
cis-1,4 polybutadiene. Use
of zirconium complexes produced polymers with a content of
cis-1,4 units higher than
99.9% were claimed.
6.6.2 Polyisoprene
Polyisoprenes occur in nature. They are also
prepared synthetically. Most commercial processes try to duplicate
the naturally occurring material.
6.6.2.1 Natural Polyisoprenes
Rubber hydrocarbon is the principle component of
raw rubber. The subject is discussed in greater detail in
Chap.
7. Natural rubber is 97% cis-1,4 polyisoprene. It is obtained by
tapping the bark of rubber trees (Hevea brasiliensis) and collecting the
exudates, a latex consisting of about 32–35% rubber. A similar
material can also be found in the sap of many other plants and
shrubs. The structure of natural rubber has been investigated over
100 years, but it was only after 1920, however, that the chemical
structure was elucidated. It was shown to be a linear polymer
consisting of head to tail links of isoprene units, 98% bonded
1,4.
6.6.2.2 Synthetic Polyisoprenes
In following natural rubber, the synthetic
efforts are devoted to obtaining very high cis-1,4 polyisoprene and to forming a
synthetic “natural” rubber. Two types of polymerizations yield
products that approach this. One is through use of Ziegler–Natta
type catalysts and the other through anionic polymerization with
alkyllithium compounds in hydrocarbon solvents. One commercial
process, for instance, uses reaction products of TiCl4
with triisobutylaluminum at an Al/Ti ratio of 0.9–1.1 as the
catalyst. Diphenyl ether or other Lewis bases are sometimes added
as catalyst modifiers [113–116]. The
process results in an approximately 95% cis-1,4 polyisoprene product.
Typically, such reactions are carried out on continuous basis,
usually in hexane and take 2–4 h. Polymerizations are often
done in two reaction lines, each consisting of four kettles
arranged in series. The heat of the reaction is partially absorbed
by precooling the feed streams. The remaining heat is absorbed on
cooled surfaces. When the stream exits, the conversion is about
80%. Addition of a shortstop solution stabilizes the product.
Alkyllithium-initiated polymerizations of
isoprene yield polymers with 92–93% cis-1,4 content. One industrial process
uses butyllithium in a continuous reaction in two lines each
consisting of four reaction kettles. The heat of the reaction is
removed by vaporization of the solvent and the monomer. The
catalyst solution is added to the solvent stream just before it is
intensively mixed with the isoprene monomer stream and fed to the
first reactor. After the stream leaves each reactor, small
quantities of methanol are injected between stages into the
reaction mixture. This limits the molecular weight by stopping the
reaction. Fresh butyllithium catalyst is added again at the next
stage in the next reactor to initiate new polymer growth
[117–119].
As is described in Chaps.
3 and 4, the monomer placement into the
polyisoprene chain can occur potentially in nine different ways.
These are the three tactic forms of the 1,2 adducts, two 1,4
adducts, cis and
trans, and three tactic
forms of 3,4-adducts. In addition, there is some possibility of
head to head and tail to tail insertion, though the common addition
is head to tail. Table 6.8 presents the various microstructures that
can be obtained in polymerizations of isoprene with different
catalysts.
Table
6.8
Polymerization products of isoprene
Mode of polymerization
|
Solvent
|
Approximate
|
|||
---|---|---|---|---|---|
% Cis-1,4
|
% Trans-1,4
|
% 1,2
|
% 3,4
|
||
Free radical
|
Emulsion in water
|
32
|
65
|
6
|
7
|
Cationic
|
–
|
37
|
51
|
4
|
9
|
Chloroform (30°C)
|
–
|
90
|
4
|
6
|
|
Anionic
|
|||||
Lithium
|
Pentane
|
94
|
0
|
6
|
|
Ethyllithium
|
Pentene
|
94
|
0
|
6
|
|
Butyllithium
|
Pentene
|
93
|
0
|
7
|
|
Sodium
|
Pentene
|
0
|
43
|
6
|
51
|
Ethylsodium
|
Pentene
|
6
|
42
|
7
|
45
|
Butylsodium
|
Pentane
|
4
|
35
|
7
|
54
|
Potassium
|
Pentane
|
0
|
52
|
8
|
40
|
Ethylpotassium
|
Pentane
|
24
|
39
|
6
|
31
|
Butylpotassium
|
Pentane
|
20
|
41
|
6
|
34
|
Rubidium
|
Pentane
|
5
|
47
|
8
|
39
|
Cesium
|
Pentene
|
4
|
51
|
8
|
37
|
Ethyllithium
|
Ethyl ether
|
6
|
29
|
5
|
60
|
Ethylsodium
|
Ethyl ether
|
0
|
14
|
10
|
76
|
Lithium
|
Ethyl ether
|
4
|
27
|
5–7
|
63–65
|
Alfin
|
Pentane
|
27
|
52
|
5
|
16
|
Coordination catalysis
|
|||||
α-TiCl3/AlR3
|
91
|
||||
VCl3/Al(C2H5)3
|
99
|
||||
TiCl4/Al(C2H5)3
|
95–96
|
||||
Ti
l4/AlR3 + amine
|
100
|
–
|
–
|
||
CoCl2/AlR3 + pyridine
|
96
|
||||
V(acetylacetonate)3/AlR3
|
90
|
||||
Ti(OR)4/Al(C2H5)3
|
95
|
Cationic polymerizations of isoprene proceed more
readily than those of butadiene, though both yield low molecular
weight liquid polymers. AlCl3 and stannic chloride can
be used in chlorinated solvents at temperatures below 0°C. Without
chlorinated solvents, however, polymerizations of isoprene require
temperatures above 0°C. At high conversions, cationic
polymerizations of isoprene result in formations of some
cross-linked material [120]. The
soluble portions of the polymers are high in trans-1,4 structures. Alfin catalysts
yield polymers that are higher in trans-1,4 structures than free-radical
emulsion polymerizations [121].
Chromium oxide catalysts on support polymerize
isoprene-like butadiene to solid polymers. Here too, however,
during the polymerization process, polymer particles cover the
catalyst completely within a few hours from the start of the
reaction and retard or stop further polymer formation. The
polymerization conditions are the same as those used for butadiene.
The reactions can be carried out over fixed bed catalysts
containing 3% chromium oxide on
SiO2–Al2O3. Conditions are 88°C
and 42 kg/cm2 pressure with the charge containing
20% of isoprene and 80% isobutane [122]. The mixed molybdenum–alumina catalyst
with calcium hydride also yields polyisoprene.
Lithium metal dispersions form polymers of
isoprene that are high in cis-1,4 contents as shown in
Table 6.8. These polymers form in hydrocarbon
solvents. This is done industrially and the products are called
Coral rubbers. They contain only a small percentage of
3,4-structures and no trans-1,4 or 1,2 units. The materials
strongly resemble Hevea
rubber.
Use of Ziegler–Natta catalysts, as seen from
Table 6.8, can yield an almost all cis-1,4-polyisoprene or an almost all
trans-1,4-polyisoprene. The
microstructure depends upon the ratio of titanium to aluminum.
Ratios of Ti:Al between 0.5:1 and 1.5:1 yield the cis isomer. A 1:1 ratio is the optimum.
Ratios of Ti:Al between 1.5:1 and 3:1 yield the trans structures [123]. The titanium to aluminum ratios also
affect the yields of the polymers as well as the microstructures.
There also is an influence on the molecular weight of the product
[124]. Variations in catalyst
compositions, however, do not affect the relative amounts of 1,4 to
3,4 or to 1,2 placements. Only cis and trans arrangements are affected. In
addition, the molecular weights of the polymers and the
microstructures are relatively insensitive to the catalyst
concentrations. The temperatures of the reactions, however, do
affect the rates, the molecular weights, and the
microstructures.
Use of Ziegler–Natta catalysts, as seen from
Table 6.8, can yield an almost all cis-1,4-polyisoprene or an almost all
trans-1,4-polyisoprene. The
microstructure depends upon the ratio of titanium to aluminum.
Variations in catalyst compositions, however, do not affect the
relative amounts of 1,4 to 3,4 or to 1,2 placements. Only
cis and trans arrangements are affected. In
addition, the molecular weights of the polymers and the
microstructures are relatively insensitive to the catalyst
concentrations. The temperatures of the reactions, however, do
affect the rates, the molecular weights, and the
microstructures.
6.7 Methyl Rubber, Poly(2,3-dimethylbutadiene)
Early attempts at preparations of synthetic
rubbers resulted in developments of elastomers from
2,3-dimethylbutadiene. The material, called “methyl rubber,” was
claimed to yield better elastomeric properties than polybutadiene.
Methyl rubber was produced in Germany during World War I where the
monomer was prepared from acetone. The polymerizations were carried
out by free-radical mechanism and anionically, using sodium metal
dispersions for initiation. Later, it was demonstrated that
2,3-dimethyl polybutadiene can be polymerized to very high
cis-1,4 polymer with
Ziegler–Natta catalysts [125,
126].
6.8 Chloroprene Rubber, Poly(2-chloro-1,3-butadiene)
2-Chloro-1,3-butadiene (chloroprene) was
originally synthesized in 1930. The material can polymerize
spontaneously to an elastomer that has good resistance to oil and
weathering. Commercial production of chloroprene rubber started in
1932. Since then, many types of polymers and copolymers were
developed with the trivial generic name of neoprene.
The monomer can be prepared from acetylene:

It can also be formed from butadiene.
Chloroprene is polymerized commercially by
free-radical emulsion polymerization. The reaction is carried out
at 40°C to a 90% conversion. A typical recipe for such an emulsion
polymerization is as follows [127]:
Material
|
Parts
|
Water
|
150
|
Chloroprene
|
100
|
Rosin
|
4 (stabilizer)
|
NaOH
|
0.8 (stabilizer)
|
K2S2O8
|
|
Sulfur
|
|
Methylene-bis-(Na-naphthalenesulfonic
acid)
|
0.7
|
When the polymerization reaches 90% conversion,
the reaction mixture is cooled to 20°C and tetraethylthiuram
disulfide is added. This is done to prevent the pendant
unsaturation in the polychloroprene backbones from cross-linking or
forming branches. An unmodified polymer is difficult to process
even at a 70% conversion. To overcome this, a
sulfur–tetraethylthiuram modification is carried out.
When the product is treated with the thiuram, an
exchange reaction takes place to yield a stable, thiuram-modified
polymer of reduced molecular weight. It is believed that the
reaction takes place through cleavage of the sulfur links, formed
during polymerization in the presence of sulfur, and formation of
free radicals [128]

After the reaction with tetraethylthiuram
disulfide is completed, the latex is acidified with acetic acid,
short of coagulation. The rubber is then recovered at a low
temperature (about −15°C) in the form of sheets by deposition of
the latex on cooled rotating drums [127, 128].
The polymer, formed by this technique, consists
of about 85% of trans-1,4
units, 10% of cis-1,4
units, 1.5% of 1,2 units, and 1.0% of 3,4 units. The polymer is
essentially linear with a molecular weight equal to approximately
100,000. The sulfur-modified polychloroprenes are sold under a
trade name of Neoprene-G. An unmodified version prepared with
mercaptan chain transferring agents (Neoprene W) is a polymer with
a molecular weight of about 200,000 [128, 130].
Table 6.9 lists the structures of polychloroprenes
that form by free-radical polymerization at different temperatures.
Chloroprene polymerizes by cationic polymerization with the aid of
Lewis acids in chlorinated solvents. When aluminum chloride is used
in a mixture of ethyl chloride–methylene chloride solvent mixture
at −80°C, the polymer has 50% 1,4 units [131, 132]. If
it is polymerized with boron trifluoride, the product consists of a
50–70% 1,4-adducts. A very high trans-1,4 poly(2-chloro-1,3-butadiene)
forms by X-ray radiation polymerization of large crystals of
chloroprene at −130 to −180°C. It is 97.8% trans-1,4 [130]. Presumably the mechanism of
polymerization is free radical.
Table
6.9
Structures of polychloroprenes formed by
free-radical polymerization
Polymerization temperature (°C)
|
% 1,4
|
References
|
|||
---|---|---|---|---|---|
Cis
|
Trans
|
% 1,2
|
% 3,4
|
||
−40
|
5
|
94
|
0.9
|
0.3
|
[112]
|
−20
|
6
|
91.5
|
0.7
|
0.5
|
[113]
|
−10
|
7
|
–
|
–
|
–
|
[112]
|
10
|
9
|
84
|
1.1
|
1.0
|
[112]
|
40
|
10
|
86, 81
|
1.6
|
1.0
|
[112]
|
40
|
13
|
88.9
|
0.9
|
0.3
|
[113]
|
100
|
13
|
71
|
2.4
|
2.4
|
[112]
|
6.9 Special Polymers from Dienes
There are many reports in the literature of
preparations of polymers from various other substituted dienes.
Most have no commercial significance. Some are, however,
interesting materials. An example is a polymer of 2-t-butyl-1,3-butadiene formed with
TiCl4 and either alkylaluminum or aluminum hydride
catalysts [132]. The polymer is
crystalline and melts at 106°C. It can be dissolved in common
solvents. Based on X-ray data, the monomer placement is high
cis-1,4.
Poly(carboxybutadiene)s also forms with
coordination catalysts [133–135]:
where R = CH3; R′ = CH3,
C2H5, C4H9, or
C6H5.

Based on the mode of packing of the chains in the
crystalline regions and from the encumbrance of the side groups in
relationship to the main chain, an erythro configuration can be assigned
[134]. The polymers, therefore,
are trans-erythro-isotactic.
Polymerization of 1,3-pentadiene can potentially
result in five different insertions of the monomers. These are
1,4-cis, 1,4-trans, 1,2-cis, 1,2-trans, and 3,4. In addition, there are
potentially 3-cis-1,4 and 3
trans-1,4 structures
(isotactic, syndiotactic, and atactic). Formations of trans-1,4 isotactic, cis-1,4 isotactic, and cis-1,4 syndiotactic polymers are
possible with Ziegler–Natta catalysts [136–138].
Amorphous polymers also form that are predominantly cis-1,4 or trans-1,4, but lack tactic order.
Stereospecificity in poly(1,3-pentadiene) is strongly dependent
upon the solvent used during the polymerization. Thus, cis-1,4 syndiotactic polymers form in
aromatic solvents and trans-1,2 in aliphatic ones. The
preparations require cobalt halide/aluminum alkyl dichloride(or
dialkyl chloride) catalysts in combinations with Lewis bases. To
form a trans-1,4 structure,
a catalyst containing aluminum to titanium ratio close to 5 must be
used [139].
6.10 Cyclopolymerization of Conjugated Dienes
Conjugated dienes like isoprene, butadiene, and
chloroprene cyclopolymerize with catalysts consisting of aluminum
alkyls, like ethylaluminum dichloride, and titanium tetrachloride
[141]. The ladder polymers that
form contain fused cyclic structures. The products prepared in
hexane are generally insoluble powders, while those prepared in
aromatic solvents are soluble even when the molecular weights are
high [142]. A high ratio of the
transition metal halide to that of the aluminum alkyl must be used.
Such a ratio might, conceivably, mean that the mechanism of
polymerization is cationic. Also, conventional cationic initiators
can be used to yield similar products. The cyclization occurs
during propagation. Unsaturation in the products can vary from none
to as high as 80%, depending upon the initiator used
[142].
Different mechanisms were offered to explain the
cyclization of 1,3 dienes [142,
143]. The cyclization might
conceivably occur by a sequential process:
or, perhaps from attacks by the propagating carbon cation on
trans-1,4 double bonds:
where R+ can represent either a propagating carbon
cation or an initiating species. The extensive cyclization may be a
result of a sequential process [142, 143].


Cyclopolymerizations typically result in low
conversions and dormant reaction mixtures. When additional monomer
is added, the dormant mixtures reinitiate polymerizations that
again proceed to some limited conversions. If the original dormant
mixtures are allowed to stand for a long time the unreacted
monomers are slowly consumed [142].
Polymerization of 2,3-dimethylbutadiene-1,3 with
Ziegler–Natta catalysts consisting of Al(i-C4H9)3–TiCl4
yields cis-1,4-polydimethylbutadiene as
described earlier. This, however, takes place when the aluminum
alkyl is in excess. If, on the other hand, the ratio of Al to Ti is
1 or less, cyclic polymer forms instead. The product has reduced
unsaturation and some trans-1,4 units in the chain
[144]. A complex catalyst,
consisting of Al(i-C4H9)3–CoCl2,
yields polymers that are predominantly cis-1,4 with about 20% of 1,2 units. On
the other hand, acid catalysts, like
Al(C2H5)Cl2, yield cyclic polymers
[143, 144]. A polymer formed with the aid of X-ray
radiation at low temperatures also contains cyclic units and some
trans-1,4 [145]. Butadiene and isoprene also form this
type of polymer at the same conditions [145].
6.11 Copolymers of Dienes
Several different elastomers, copolymers of
butadiene, are produced commercially. The major ones are copolymers
of butadiene with styrene and butadiene with acrylonitrile. Some
terpolymers, where the third component is an unsaturated carboxylic
acid, are also manufactured. Block copolymers of isoprene with
styrene and butadiene with styrene are important commercial
elastomers.
6.11.1 GR-S Rubber
Copolymerization of butadiene with styrene by
free-radical mechanism has been explored very thoroughly
[146]. The original efforts
started during World War I in Germany. Subsequent work during the
1930s was followed by a particularly strong impetus in the United
States during World War II. This led to a development of GR-S
rubber in the United States and Buna-S rubber in Germany. After
World War II further refinements were introduced into the
preparatory procedures and “cold” rubber was developed.
Industrially, the copolymer is prepared by emulsion
copolymerization of butadiene and styrene at low temperatures in a
continuous process. A typical product is a random distribution
copolymer, with the butadiene content ranging from 70 to 75%. The
diene monomer placement is roughly 18% cis-1,4; 65% trans-1,4; and 17% 1,2. M n of these copolymers is
about 100,000.
A “redox” initiator is used in the cold process,
but not in the “hot” one. Also, the “hot” process is carried out at
about 50°C for 12 h to approximately 72% conversion. The
“cold” process is also carried for 12 h, but at about 5°C to a
60% conversion. The two recipes for preparation of GR-S rubbers are
shown in Table 6.10 for comparison of the “hot” and “cold”
processes.
Table
6.10
Typical recipes for preparation of
butadiene–styrene rubbers by emulsion polymerization
Material
|
“Hot” process
|
“Cold” process
|
||
---|---|---|---|---|
Parts
|
Purpose
|
Parts
|
Purpose
|
|
Butadiene
|
75
|
Comonomer
|
72
|
Comonomer
|
Styrene
|
25
|
Comonomer
|
28
|
Comonomer
|
Water
|
180
|
Carrier
|
180
|
Carrier
|
Fatty acid soap
|
5.0
|
Emulsifier
|
4.5
|
Emulsifier
|
n-Dodecyl mercaptan
|
0.5
|
Chain transferring agent
|
–
|
–
|
t-Dodecyl mercaptan
|
–
|
–
|
0.2
|
Chain transferring agent
|
Potassium persulfate
|
0.3
|
Initiator
|
–
|
–
|
Auxiliary surfactant
|
–
|
–
|
0.3
|
Stabilizer
|
Potassium chloride
|
–
|
–
|
0.3
|
Stabilizer
|
p-Menthane hydroperoxide
|
–
|
–
|
0.06
|
Initiator system
|
Ferrous sulfate
|
–
|
–
|
0.01
|
|
Ethylenediamine tetraacetic acid sodium
salt
|
–
|
–
|
0.05
|
|
Sodiumformaldehyde sulfoxylate
|
0.05
|
In both polymerizations, the unreacted monomer
has to be removed. In the “hot” one the reaction is often quenched
by addition of hydroquinone, and in the “cold” one by addition of
N,N-diethyldithiocarbamate. After the
monomers are steam stripped in both processes, an antioxidant like
N-phenyl-2-naphthylamine is
added. The latex is usually coagulated by addition of a sodium
chloride–sulfuric acid solution. The “cold” process yields polymers
with less branching than the “hot” one, slightly higher
trans to cis ratios.
During the middle 1960s a series of
butadiene–styrene and isoprene–styrene block-copolymer-elastomers were
developed. These materials possess typical rubber-like properties
at ambient temperatures, but act like thermoplastic resins at
elevated ones. The copolymers vary from diblock structures of
styrene and butadiene
to triblock ones, like styrene–butadiene–styrene:


A typical triblock copolymer may consist of about
150 styrene units at each end of the macromolecule and some 1,000
butadiene units in the center. The special physical properties of
these block copolymers are due to inherent incompatibility of
polystyrene with polybutadiene or polyisoprene blocks. Within the
bulk material, there are separations and aggregations of the
domains. The polystyrene domains are dispersed in continuous
matrixes of the polydienes that are the major components. At
ambient temperature, below the T g of the polystyrene,
these domains are rigid and immobilize the ends of the polydiene
segments. In effect they serve both as filler particles and as
cross-links. Above T
g of polystyrene, however, the domains are easily
disrupted and the material can be processed as a thermoplastic
polymer. The separation into domains is illustrated in
Fig. 6.4.

Fig.
6.4
Illustration of polystyrene and
polybutadiene domains
These thermoplastic elastomers are prepared by
anionic solution polymerization with organometallic catalysts. A
typical example of such preparation is polymerization of a 75/25
mixture of butadiene/styrene in the presence of sec-butyllithium in a hydrocarbon–ether
solvent blend. At these reaction conditions butadiene blocks form
first and when all the butadiene is consumed, styrene blocks form.
In other preparations, monomers are added sequentially, taking
advantage of the “living” nature of these anionic
polymerizations.
These block copolymers have very narrow molecular
weight distributions. Also, the sizes of the blocks are restricted
to narrow ranges to maintain optimum elastomeric properties.
6.11.2 GR-N Rubber
Butadiene–acrylonitrile rubbers are another group
of useful synthetic elastomers. These copolymers were originally
developed in Germany where they were found superior in oil
resistance to the butadiene–styrene rubbers. Commercially, these
materials are produced by free-radical emulsion polymerization very
similarly to the butadiene–styrene copolymers. Similarly, “hot” and
“cold” processes are employed. “Low,” “medium,” and “high” grades
of solvent-resistant copolymers are formed, depending upon the
amount of acrylonitrile in the copolymer that can range from 25 to
40%. The butadiene placement in these copolymers is approximately
77.5% trans-1,4, 12.5%
cis-1,4, and 10% of
1,2-units. Also, the polymers formed by the “cold” process are less
branched and have a narrower molecular weight distribution than
those formed by the “hot” process.
An interesting alternating copolymer of butadiene
and acrylonitrile was developed in Japan [147]. The copolymer is formed with coordination
catalysts consisting of AlR3, AlCl3, and
VOCl3 in a suspension polymerization process. The
product is more than 94% alternate and is reported to have very
good mechanical properties and good oil resistance.
6.12 Polystyrene and Polystyrene-Like Polymers
Styrene is produced in the United States from
benzene and ethylene by a Friedel–Craft reaction that is followed
by dehydrogenation over alumina at 600°C. Polystyrene was first
prepared in 1839, though the material was confused for an oxidation
product of the styrene monomer [148]. Today polystyrene is produced in very
large quantities and much is known about this material.
6.12.1 Preparation of Polystyrene by Free-Radical Mechanism
Styrene is one of those monomers that lends
itself to polymerization by free-radical, cationic, anionic and
coordination mechanisms. This is due to several reasons. One is
resonance stabilization of the reactive polystyryl species in the
transition state that lowers the activation energy of the
propagation reaction. Another is the low polarity of the monomer.
This facilitates attack by free-radicals, differently charged ions,
and metal complexes. In addition, no side reactions that occur in
ionic polymerizations of monomers with functional groups are
possible. Styrene polymerizes in the dark by free-radical mechanism
more slowly than it does in the presence of light [149]. Also, styrene formed in the dark is
reported to have greater amount of syndiotactic placement
[150]. The amount of branching in
the polymer prepared by free-radical mechanism increases with
temperature [136]. This also
depends upon the initiator used [151].
The following information has evolved about the
free-radical polymerization of styrene:
2.
Oxygen retards polymerizations of styrene. At
higher temperatures, however, the rate is accelerated due to
peroxide formation [156].
3.
The rate of styrene polymerizations in bulk is
initially, at low conversions, first order with respect to monomer
concentrations. In solution, however, it is a second order with
respect to monomer [157].
Polystyrene that is manufactured by free-radical
polymerization is atactic. Isotactic polystyrene formed with
Ziegler–Natta catalysts was introduced commercially in the 1960s,
but failed to gain acceptance. Syndiotactic polystyrene is now
being produced commercially.
Industrially, free-radical styrene
polymerizations are carried out in bulk, in emulsion, in solution,
and in suspension. The clear plastic is generally prepared by mass
polymerization. Because polystyrene is soluble in the monomer, mass
polymerization, when carried out to completion, results in a
tremendous increase in melt viscosity. To avoid this, when styrene
is polymerized in bulk in an agitated kettle, the reaction is only
carried out to 30–40% conversion. After that, the viscous syrup is
transferred to another type of reactor for the completion of the
reaction. According to one early German patent, polymerization is
completed in a plate and frame filter press [157]. Water circulating through the press
removes the heat of the reaction, and the solid polymer is formed
inside the frames. This process is still used in some places
[158].
Another approach is to use adiabatic towers.
Styrene is first partially polymerized in two agitated reaction
kettles at 80–100°C. The syrup solution of the polymer in the
monomer is then fed continually into the towers from the top. The
temperatures in the towers are gradually increased from 100–110°C
at the top to 180–200°C at the bottom. By the time the material
reaches the bottom, in about 3 h, the polymerization is 92–98%
complete [159]. The unreacted
monomer is removed and recycled. A modification of the process is
to remove the monomer vapor at the top of the tower for
reuse.
An improvement in the above procedure is the use
of agitated towers [160]. To
avoid channeling inside the towers and for better heat transfer,
three towers are arranged in series. They are equipped with slow
agitators and with grids of pipes for cooling and heating
[160]. Polymeric melt is heated
from 95 to 225°C to reduce viscosity and help heat transfer. A
solvent like ethyl benzene may be added. A vacuum devolatalizer
removes both monomer and solvent from the product
(Fig. 6.5).

Fig.
6.5
Adiabatic tower for mass polymerization of
styrene
6.12.2 Polystyrene Prepared by Ionic Chain-Growth Polymerization
Much research was devoted to both cationic and
anionic polymerizations. An investigation of cationic
polymerization of styrene with
Al(C2H5)2Cl/RCl (R = alkyl or
aryl) catalyst/cocatalyst system was reported by Kennedy
[161, 162]. The efficiency (polymerization
initiation) is determined by the relative stability and/or
concentration of the initiating carbocations that are provided by
the cocatalyst RCl. N-butyl, isopropyl, and sec-butyl chlorides exhibit low
cocatalytic efficiencies because of low tendency for ion formation.
Triphenylmethyl chloride is also a poor cocatalyst because the
triphenylmethyl ion that forms is more stable than the propagating
styryl ion. Initiation of styrene polymerizations by carbocations
is now well established [163].
Anionic polymerization of styrene
with amyl sodium yields an isotactic polymer [164]. Polymerizations catalyzed by
triphenylmethylpotassium also yield the same stereospecific
polystyrene [165]. The same is
true of organolithium compounds [166, 167].
In butyllithium initiated polymerizations of
styrene in benzene termination were claimed to occur by association
between the propagating anions and the lithium counterions with
another butyllithium molecule [168]. This was contradicted by claims that
terminations result from association of two propagating chains
[169]. Alfin catalysts polymerize
styrene to yield stereospecific products [170].
Coordination catalysts based on
aluminum alkyls and titanium halides yield isotactic polystyrene
[171–174]. The polymer matches isotactic polystyrene
formed with amylsodium. It is composed of head-to-tail sequences
with the main chain fold being helical. There are three monomer
units per each helical fold [171,
172]. The catalyst composition,
however, has a strong bearing on the microstructure of the
resultant polymer [175,
176].
Ishihara et al. reported in 1986 that
syndiotactic polystyrene can be prepared with the aid of organic or
inorganic titanium compounds activated with methylaluminoxane
[177]. There is much greater
incentive to commercialize syndiotactic polystyrene than the
isotactic one. This is because isotactic polystyrene crystallizes
at a slow rate. That makes it impractical for many industrial
applications. Syndiotactic polystyrene, on the other hand,
crystallizes at a fast rate, has a melting point of 275°C, compared
to 240°C for the isotactic one, and is suitable for use as a strong
structural material.
Many catalysts were investigated for the
preparation of syndiotactic polystyrene. High activity is claimed
for half-sandwich titanocenes of the type CpTiCl3,
IndTiCl3, and substituted IndTiCl3 with
methylaluminoxane as cocatalyst [178]. It was also reported that
BMe(C6F5)3 and other borates can
be used as precursors instead of methyl-aluminoxane [178]. In addition, it was disclosed that
fluorinated catalysts exhibit very high activities and produce
polymers with higher molecular weight polymers [178].
The mechanism of polymerization was investigated
by different groups. Grassi and Zambelli claim that the
syndiotactic styrene polymerization proceeds through a secondary
insertion of styrene into Ti–alkyl (or growing polymer chain) bond.
In the half titanocene catalyst, the polymer chain appears to be
η6-coordinated to the metal of the active species
[179]:

Maron and coworkers [180] reported that theoretical methods were
used to investigate the syndiospecificity of the styrene
polymerization catalyzed by single-site, single-component allyl
ansa-lanthanidocenes:

Two limiting chain end stereocontrol mechanisms
were studied by them, namely, migratory insertion through a site
epimerization and site stereoconfiguration independent of backside
insertion on a “stationary” polymer chain. Four consecutive
insertions of styrene were computed to reveal that (i) backside
insertions are more favorable than, or at least as favorable as,
frontside insertions. The formation of a syndiotactic polymer is
controlled by the thermodynamics. Moreover, the odd (first and
third) insertions are of 2,1-down-si-type and are kinetically favored
over the 2,1-up-re-ones.
This control is the conjunction of two effects: minimization of
styrene–styrene and styrene (phenyl ring)–fluorenyl repulsions. The
steric hindrance of the polymer chain induces a fourth insertion by
an exocyclic coordination of the fluorenyl ligand that is
compensated by the η6 coordination of one of the phenyl
ring in the growing chain.
Syndiotactic polystyrene is available
commercially under the trade name of Questra. This material is
produced with the aid of a metallocene catalyst and is sold in
several grades [181].
There is a small interest in forming isotactic
polystyrenes with vary narrow molecular weight distributions,
because of some very limited practical applications, and from
purely academic interests. Several preparations of virtually
monodisperse polystyrenes of M w/M n = 1.06 by anionic
polymerizations were developed. The materials are available
commercially [181–186], small quantities for use as standards for
GPC.
6.12.3 Polymers from Substituted Styrenes
Many derivatives of styrene can be readily
synthesized. Some are commercially available. One of them is
α-methyl styrene. It is formed from propylene and benzene by a
process that is very similar to styrene preparation:

Due to the allylic nature of α-methyl styrene it
cannot be polymerized by free-radical mechanism. It polymerizes
readily, however, by an ionic one. Resins based on copolymers of
α-methyl styrene are available commercially. Other styrene
derivatives that can be obtained commercially are:
1.
Alkyl or aryl substituted styrenes,

where R1 and R2 are alkyl
or aryl groups.
2.
Halogen derivatives,

where X1 and X2 = F, Cl,
Br, or I.
3.
Polar-substituted styrenes,
where R = CN, CHO, COOH, OCOCH3, OH, OCH3,
NO2, NH2, and SO3H.

Vinyl toluene polymerizes readily by free-radical
mechanism at 100°C. The absolute rate at that temperature is
greater than for styrene. The activation energy for vinyl toluene
polymerization is 17–19 kcal per mole while that for styrene
is 21 kcal mole. This monomer can also be polymerized by ionic
and coordination mechanisms. Earlier attempts at polymerization of
α-methyl styrene with Ziegler–Natta catalysts were not successful
[187, 188]. Later, however, it was shown that
polymerization does take place with
TiCl4/Al(C2H5)3 at
−78°C. The activity of the catalyst and the DP depend on the ratio
of aluminum to titanium, the nature of the solvent, and on the
aging of the catalyst [189]. The
optimum ratio of the aluminum alkyl to titanium chloride is
1.0–1.2. Mixing and aging of the catalyst must be done below room
temperature, and the valence of titanium must be maintained between
3 and 4 [189]. This led Sakurada
to suggest that the reaction actually proceeds via a cationic
rather than a coordinated anionic mechanism [189].
Various reports in the literature describe
cationic polymerizations of α-methyl styrene with Lewis acids
[190–192]. The products are mostly low molecular
weight polymers, some containing unsaturation with pendant
phenylindane groups. A high molecular weight polymer can be
prepared from α-methyl styrene by cationic polymerization at −90 to
−130°C with AlCl3 in ethyl chloride or in carbon
disulfide [193]. The product has
a narrow molecular weight distribution. Some T g and T m values of
polystyrene-like materials, including isotactic polystyrene, are
presented in Table 6.11.
Table
6.11
Transition temperatures of poly(phenylene
alkene)s
Polymer
|
T
g (°C)
|
References
|
T
m (°C)
|
References
|
---|---|---|---|---|
![]() |
80
|
[178]
|
(i) 240; (s) 272
|
[178]
|
![]() |
73.81
|
[178]
|
Amorphous
|
[176]
|
![]() |
77
|
[178]
|
219
|
[176]
|
![]() |
92
|
[178]
|
310
|
[176]
|
![]() |
123
|
[178]
|
330
|
[176]
|
![]() |
91
|
[178]
|
240
|
[176]
|
![]() |
84
|
[177]
|
290
|
[176]
|
![]() |
60
|
[177]
|
207–208
|
[176]
|
![]() |
80
|
[177]
|
2
|
[177]
|
![]() |
34–40
|
[177]
|
180
|
[177]
|
![]() |
60–65
|
[177]
|
240
|
[177]
|
![]() |
10
|
[177]
|
162–168
|
[177]
|
6.13 Copolymers of Styrene
Many copolymers of styrene are manufactured on a
large commercial scale. Because styrene copolymerizes readily with
many other monomers, it is possible to obtain a wide distribution
of properties. Random copolymers form quite readily by free-radical
mechanism [185, 194]. Some can also be formed by ionic
mechanism. In addition, graft and block copolymers of styrene are
also among commercially important materials.
Most comonomers differ from styrene in polarity
and reactivity. A desired copolymer composition can be achieved,
however, through utilization of copolymerization parameters based
on kinetic data and on quantum-chemical considerations. This is
done industrially in preparations of styrene–acrylonitrile,
styrene–methyl methacrylate, and styrene–maleic anhydride
copolymers of different compositions.
6.13.1 High-Impact Polystyrene
For many applications, the homopolymer of styrene
is too brittle. To overcome that, many different approaches were
originally tried. These included use of high molecular weight
polymers, use of plasticizers, fillers (glass fiber, wood flour,
etc.), deliberate orientation of the polymeric chains,
copolymerization and addition of rubbery substances. Effect of
plasticizers is too severe for practical use, and use of high
molecular weight polymers exhibits only marginal improvement. Use
of fillers, though beneficial, is mostly confined to United States.
Orientation is limited to sheets and filaments, and
copolymerization usually lowers the softening point too much.
Addition of rubbery materials, however, does
improve the impact resistance of polystyrene. This is done,
therefore, extensively. The most common rubbers used for this
purpose are butadiene–styrene copolymers. Some butadiene
homopolymers are also used but to a lesser extent. The high-impact
polystyrene is presently prepared by dissolving the rubber in a
styrene monomer and then polymerizing the styrene. This
polymerization is either done in bulk or in suspension. The product
contains styrene–butadiene rubber, styrene homopolymer, and a
considerable portion of styrene-graft copolymer that forms when
polystyrene radicals attack the rubber molecules. The product has
very enhanced impact resistance.
Past practices, however, consisted simply in
blending a mixture of polystyrene and rubber on a two-roll mill, or
in a high shear internal mixer, or passing through an extruder. The
impact strength of the product was only moderately better than that
of the unmodified polymer. Another procedure was to blend
polystyrene emulsion latex with a styrene–butadiene rubber emulsion
latex and then to coagulate the two together. The product is also
only marginally better in impact strength than styrene homopolymer.
This practice, however, may still be in existence in some
places.
In high-impact polystyrene, the rubber exists in
discrete droplets, less than 50 μm in diameter. In effect the
polymerization serves to form an oil in oil emulsion
[199] where the polystyrene is in
the continuous phase and the rubber is in a dispersed phase. The
graft copolymer that forms serves to “emulsify” this heterogeneous
polymer solution [200].
Commercial high-impact polystyrene usually
contains 5–20% styrene–butadiene rubber. The particle size ranges
from 1 to 10 μm. High-impact polystyrene may have as much as
seven times the impact strength of polystyrene, but it has only
half its tensile strength, lower hardness, and lower softening
point.
6.13.2 ABS Resins
Styrene–acrylonitrile copolymers are produced
commercially for use as structural plastics. The typical
acrylonitrile content in such resins is between 20 and 30%. These
materials have better solvent and oil resistance than polystyrene
and a higher softening point. In addition, they exhibit better
resistance to cracking and crazing and an enhanced impact strength.
Although the acrylonitrile copolymers have enhanced properties over
polystyrene, they are still inadequate for many applications.
Acrylonitrile–butadiene–styrene polymers, known as ABS resins, were
therefore developed.
Although ABS resins can potentially be produced
in a variety of ways, there are only two main processes. In one of
them acrylonitrile–styrene copolymer is blended with a
butadiene–acrylonitrile rubber. In the other one, interpolymers are
formed of polybutadiene with styrene and acrylonitrile.
In the first one, the two materials are blended
on a rubber mill or in an internal mixer. Blending of the two
materials can also be achieved by combining emulsion latexes of the
two materials together and then coagulating the mixture. Peroxide
must be added to the blends in order to achieve some cross-linking
of the elastomer to attain optimum properties. A wide range of
blends are made by this technique with various properties
[201]. Most common commercial
blends of ABS resins may contain 70 parts of styrene–acrylonitrile
copolymer (70/30) and 40 parts of butadiene–nitrile rubber
(65/35).
In the second process, styrene and acrylonitrile
are copolymerized in the presence of polybutadiene latex. The
product is a mixture of butadiene homopolymer and a graft
copolymer.
6.13.3 Copolymers of Styrene with Maleic Anhydride
Many styrene–maleic anhydride copolymers are
produced commercially for special uses. These are formed by
free-radical copolymerization and many commercial grades are
partially esterified. Molecular weights of such polymers may range
from 1,500 to 50,000, depending upon the source. The melting points
of these copolymers can vary from 110 to 220°C, depending upon
molecular weight, degree of hydrolysis and esterification, and also
the ratio of styrene to maleic anhydride.
Only a small percent of styrene copolymers
reported in the literature achieved industrial importance. Some of
the interesting copolymers of styrene that were reported but not
utilized commercially are copolymers with various unsaturated
nitriles. This includes vinylidene cyanide, fumaronitrile,
malononitrile, methacrylonitrile, acrylonitrile, and cinnamonitrile
[292]. Often, copolymerization of
styrene with nitriles yields copolymers with higher heat distortion
temperature, higher tensiles, better craze resistance, and higher
percent elongation.
Styrene was also copolymerized with many acrylic
and methacrylic esters. Products with better weathering properties
often form. Copolymerization with some acrylates lowers the value
of T g
[203].
6.14 Polymers of Acrylic and Methacrylic Esters
There are many synthetic procedures for preparing
acrylic acid and its esters. One way, used early, is to make
acrylic esters from ethylene oxide:

Another route is through oxidation of propylene
over cobalt molybdenum catalyst at 400–500°C:

Many industrial preparations start with
acetylene, carbon monoxide, and alcohol or water:

One route to α-methyl acrylic acid (or
methacrylic acid) and its esters is by a cyanohydrin reaction of
acetone:

6.14.1 Polymerizations of Acrylic and Methacrylic Esters
Free-radical bulk polymerizations of acrylate
esters exhibit rapid rate accelerations at low conversions. This
often results in formation of some very high molecular weight
polymer and some cross-linked material. The cross-linking is a
result of chain transferring by abstractions of labile tertiary
hydrogens from already formed “dead” polymeric chains
[204]. Eventually, termination by
combination of the branched radicals leads to cross-linked
structures. Addition of chain transferring agents, like mercaptans
(that reduces the length of the primary chains), helps prevent gel
formation. There are no labile tertiary hydrogens in methacrylic
esters. The growing methacrylate radicals are still capable of
abstracting hydrogens from the α-methyl groups. Such abstractions,
however, require more energy and are not an important problem in
polymerizations of methacrylic esters [205]. Nevertheless, occasional formation of
cross-linked poly(alkyl methacrylate)s does occur. This is due to
chain transferring to the alcohol moiety [206, 210].
The termination reaction in free-radical
polymerizations of the esters of acrylic and methacrylic acids
takes place by recombination and by disproportionation
[206, 207]. Methyl methacrylate polymerizations,
however, terminate at 25°C predominantly by disproportionation
[205].
Oxygen inhibits free-radical polymerization of
α-methyl methacrylate [208]. The
reaction with oxygen results in formation of low molecular weight
polymeric peroxides that subsequently decompose to formaldehyde and
methyl pyruvate [210]:

Oxygen is less effective in inhibiting
polymerizations of acrylic esters. It reacts 400 times faster with
the methacrylic radicals than with the acrylic ones. Nevertheless,
even small quantities of oxygen affect polymerization rates of
acrylic esters [216]. This
includes photopolymerizations of gaseous ethyl acrylate that are
affected by oxygen and by moisture [217].
Acrylic and methacrylic esters polymerize by
free-radical mechanism to atactic polymers. The sizes of the
alcohol portions of the esters determine the T g values of the resultant
polymers. They also determine the solubility of the resultant
polymers in hydrocarbon solvents and in oils.
Solvents influence the rate of free-radical
homopolymerization acrylic acid and its copolymerization with other
monomers. Hydrogen bonding solvents slow down the reaction rates
[219]. Due to electron
withdrawing nature of the ester groups, acrylic and methacrylic
ester polymerize by anionic but not by cationic mechanisms. Lithium
alkyls are very effective initiators of α-methyl methacrylate
polymerization yielding stereospecific polymers [213]. Isotactic poly(methyl methacrylate) forms
in hydrocarbon solvents [214].
Block copolymers of isotactic and syndiotactic poly(methyl
methacrylate) form in solvents of medium polarity. Syndiotactic
polymers form in polar solvents, like ethylene glycol dimethyl
ether, or pyridine. This solvent influence is related to Lewis
basicity [215] in the following
order:

Furthermore, polymerizations in solvating media,
like ethylene glycol dimethyl ether, tetrahydrofuran, or pyridine,
using biphenylsodium or biphenyllithium yield virtually
monodisperse syndiotactic poly(methyl methacrylate) [216].
The nature of the counterion in anionic
polymerizations of methyl methacrylate in liquid ammonia with
alkali metal amide or alkali earth metal amide catalysts is an
important variable [217]. Lithium
and calcium amides yield high molecular weight polymers, though the
reactions tend to be slow. Sodium amide, on the other hand, yields
rapid polymerizations but low molecular weight polymers. Polymers
formed with sodium amide, however, have a narrower molecular weight
distribution than those obtained with lithium and calcium amides.
Calcium amide also yields high molecular weight polymers from ethyl
acrylate and methyl methacrylate monomers in aromatic and aliphatic
solvents at temperatures from −8 to 110°C. When, however,
tetrahydrofuran or acetonitrile is used as solvents much lower
molecular weight products form [218].
Products from anionic polymerizations of methyl
methacrylate catalyzed by Grignard reagents (RMgX) vary with the
nature of the R and X groups, the reaction temperature, and the
nature of the solvent [219–221].
Secondary alkyl Grignard reagents give the highest yields and the
fastest rates of the reactions. Isotacticity of the products
increases with the temperature. When anion-radicals from alkali
metal ketyls of benzophenone initiate polymerizations of methyl
methacrylate, amorphous polymers form at temperatures from −78 to
+65°C [222].
Sodium dispersions in hexane yield syndiotactic
poly(methyl methacrylate) [223].
A 60–65% conversion is obtained over a 24-h period at a reaction
temperature of 20–25°C. Lithium dispersions [224], butyllithium [203], and Grignard reagents [225, 226]
yield crystalline isotactic poly(t-butyl acrylate). The reactions take
place in bulk and in hydrocarbon solvents. Isotactic poly(isopropyl
acrylate) forms with Grignard reagents [226, 227].
Coordination polymerizations of methyl
methacrylate with diethyliron–bipyridyl complex in nonpolar
solvents like benzene or toluene yield stereoblock polymers. In
polar solvents, however, like dimethylformamide or acetonitrile,
the products are rich in isotactic placement [229].
There are many reports in the literature on
polymerizations of acrylic and methacrylic esters with
Ziegler–Natta catalysts [230–233]. The
molecular weights of the products, the microstructures, and rates
of the polymerizations depend upon the metal alkyl and the
transition metal salt used. The ratios of the catalyst components
to each other are also important [234, 235].
In 1992 Yasuda et al. [236, 237]
reported that organolanthanide complexes of the type
Cp*2Sm-R (where Cp* is pentamethyl cyclopentadienyl, and
R is either an alkyl, alkylaluminum or a hydride) initiate highly
syndiotactic, living polymerizations of methacrylates. It was also
reported that lanthanide complexes such as
Cp*2Yb(THF)1–3,
Cp*2Sm(THF)2, and
(indenyl)2Yb(THF)2 can also initiate
polymerizations of methylmethacrylate [238]. Although very low initiator efficiencies
were observed, these were living polymerizations. The polymers that
formed had the dispersity of 1.1 and were high in syndiotactic
sequences.
Novak and Boffa, in studying lanthanide
complexes, observed an unusual facile organometallic electron
transfer process takes place that generates in situ bimetallic
lanthanide(III) initiators for polymerizations of methacrylates
[239]. They concluded that methyl
methacrylate polymerizations initiated by the Cp*2Sm
complexes occur through reductive dimerizations of methyl
methacrylate molecules to form “bisinitiators” that consists of two
samarium(III) enolates joined through their double bond terminally
[239]. Their conclusion is based
on the tendency of Cp*2Sm complexes to reductively
couple unsaturated molecules:

Montei and coworkers [240] reported that Nickel complexes
[(X,O)NiR(PPh3)] (X = N or P), designed for the
polymerization of ethylene, are effective for home- and
copolymerization of butyl acrylate, methyl methacrylate, and
styrene. Their role as radical initiators was demonstrated from the
calculation of the copolymerization reactivity ratios. It was shown
that the efficiency of the radical initiation is improved by the
addition of PPh3 to the nickel complexes as well as by
increasing the temperature. The dual role of nickel complex as
radical initiators and catalysts was exploited to succeed in the
copolymerization of ethylene with butyl acrylate and methyl
methacrylate.
Multiblock copolymers containing sequences of
both ethylene and polar monomers were thus prepared.
6.14.2 Acrylic Elastomers
Polymers of lower n-alkyl acrylates are used commercially
to only a limited extent. Ethyl and butyl acrylates are, however,
major components of acrylic
elastomers. The polymers are usually formed by free-radical
emulsion polymerization. Because acrylate esters are sensitive to
hydrolysis under basic conditions, the polymerizations are usually
conducted at neutral or acidic pH. The acrylic rubbers, like other
elastomers must be cross-linked or vulcanized to obtain optimum
properties. Cross-linking can be accomplished by reactions with
peroxides through abstractions of tertiary hydrogens with free
radicals:

Another way to cross-link acrylic elastomers is
through a Claisen condensation:

The above illustrated cross-linking reactions of
homopolymers, however, form elastomers with poor aging properties.
Commercial acrylic rubbers are, therefore, copolymers of ethyl or
butyl acrylate with small quantities of comonomers that carry
special functional groups for cross-linking. Such comonomers are
2-chloroethylvinyl ether or vinyl chloroacetate, used in small
quantities (about 5%). These copolymers cross-link through
reactions with polyamines.
6.14.3 Thermoplastic and Thermoset Acrylic Resins
Among methacrylic ester polymers, poly(methyl
methacrylate) is the most important one industrially. Most of it is
prepared by free-radical polymerizations of the monomer and a great
deal of these polymerizations are carried out in bulk. Typical
methods of preparation of clear sheets and rods consist of initial
partial polymerizations in reaction kettles at about 90°C with
peroxide initiators. This is done by heating and stirring for about
10 min to form syrups. The products are cooled to room
temperature and various additives may be added. The syrups are
solutions of about 20% polymer dissolved in the monomer. They are
poured into casting cells where the polymerizations are completed.
The final polymers are high in molecular weight, about
1,000,000.
Poly(methyl methacrylate) intended for surface
coatings is prepared by solution polymerization. The molecular
weights of the polymers are about 90,000 and the reaction products
that are 40–60% solutions are often used directly in
coatings.
A certain amount of poly(methyl methacrylate) is
also prepared by suspension polymerization. The molecular weights
of these polymers are about 60,000 and they are used in injection
molding and extrusion.
Thermosetting acrylic resins are used widely in
surface coatings. Both acrylic and methacrylic esters are utilized
and the term is applied to both of them. Often such resins are
terpolymers or even tetra polymers where each monomer is chosen for
a special function [214]. One is
selected for rigidity, surface hardness, and scratch resistance;
another for the ability to flexibilize the film, and the third one
for cross-linking it. In addition, not all comonomers are
necessarily acrylic or methacrylic esters or acids. For instance,
among the monomers that may be chosen for rigidity may be methyl
methacrylate. On the other hand, it may be styrene instead, or
vinyl toluene, etc. The same is true of the other components.
Table 6.12 illustrates some common components that
can be found in thermoset acrylic resins.
Table
6.12
Typical components of thermoset acrylic
resins
Monomers that contribute rigidity
|
Flexibilizing monomers
|
Monomers used for cross-linking
|
---|---|---|
Methyl methacrylate
|
Ethyl acrylate
|
Acrylic acid
|
Ethyl methacrylate
|
Isopropyl acrylate
|
Methacrylic acid
|
Styrene
|
Butyl acrylate
|
Hydroxyethyl acrylate
|
Vinyl toluene
|
i-Octyl acrylate
|
Hydroxypropyl acrylate
|
Acrylonitrile
|
Decyl acrylate
|
Glycidyl acrylate
|
Methacrylonitrile
|
Lauryl methacrylate
|
Glycidyl methacrylate
|
Acrylamide
|
||
Aminoethyl acrylate
|
The choice of the cross-linking reaction may
depend upon desired application. It may also simply depend upon
price, or a particular company that manufactures the resin, or
simply to overcome patent restrictions. Some common cross-linking
reactions will be illustrated in the remaining portion of this
section. If the functional groups are carboxylic acids in the
copolymer or terpolymer, cross-linking can be accomplished by
adding a diepoxide.

Other reactions can also cross-link resins with
pendant carboxylic acid groups. For instance, one can add a
melamine formaldehyde condensate:

A diisocyanate, a phenolic, or a
melamine–formaldehyde resin can be used as well. Resins with
pendant hydroxyl groups can also be cross-linked by these
materials. A diisocyanate is effective in forming urethane
linkages:

When the pendant groups are epoxides, like
glycidyl esters, cross-linking can be carried out with dianhydrides
or with compounds containing two or more carboxylic acid groups
[241]. Aminoplast resins
(urea–formaldehyde or melamine–formaldehyde and similar ones) are
also very effective [242].
Pendant amide groups from terpolymers containing
acrylamide can be reacted with formaldehyde to form methylol groups
for cross-linking [243]:

6.15 Acrylonitrile and Methacrylonitrile Polymers
Polymers from acrylonitrile are used in synthetic
fibers, in elastomers, and in plastic materials. The monomer can be
formed by dehydration of ethylene cyanohydrin:

Other commercial processes exist, like
condensation of acetylene with hydrogen cyanide, or ammoxidation of
propylene:

Acrylonitrile polymerizes readily by free-radical
mechanism. Oxygen acts as a strong inhibitor. When the
polymerization is carried out in bulk, the reaction is
autocatalytic [242,
243]. In solvents, like
dimethylformamide, however, the rate is proportional to the square
root of the monomer concentration [242]. The homopolymer is insoluble in the
monomer and in many solvents.
Acrylonitrile polymerizes also by anionic
mechanism. There are many reports in the literature of
polymerizations initiated by various bases. These are alkali metal
alkoxides [246], butyllithium
[247, 248], metal ketyls [249, 250],
solutions of alkali metals in ethers [251, 252],
sodium malonic esters [232], and
others. The propagation reaction is quite sensitive to termination
by proton donors. This requires use of aprotic solvents. The
products, however, are often insoluble in such solvents. In
addition, there is a tendency for the polymer to be yellow. This is
due to some propagation taking place by 1,4 and by 3,4 insertion in
addition to the 1,2 placement [253, 254]:

Another disadvantage of anionic polymerization of
acrylonitrile is formation of cyanoethylate as a side reaction. It
can be overcome, however, by running the reaction at low
temperatures. An example is polymerizations initiated by KCN at
−50°C in dimethylformamide [254],
or by butyllithium in toluene at −78°C [255]. Both polymerizations yield white, high
molecular weight products that are free from cyanoethylation.
It was suggested that the terminations in anionic
polymerizations of acrylonitrile proceed by proton transfer from
the monomer. This, however, depends upon catalyst concentrations
[256, 257]. At low concentrations, the terminations
can apparently occur by a cyclization reaction [257] instead:

Industrially, polyacrylonitrile homopolymers and
copolymers are prepared mainly by free-radical mechanism. The
reactions are often conducted at low temperatures, in aqueous
systems, either in emulsions or in suspensions, using redox
initiation. Colorless, high molecular weight materials form. Bulk
polymerizations are difficult to control on a large scale.
Over half the polymer that is prepared
industrially is for use in textiles. Most of these are copolymers
containing about 10% of a comonomer. The comonomers can be methyl
methacrylate, vinyl acetate, or 2-vinylpyridine. The purpose of
comonomers is to make the fibers more dyeable. Polymerizations in
solution offer an advantage of direct fiber spinning.
Polyacrylonitrile copolymers are also used in
barrier resins for packaging. One such resin contains at least 70%
acrylonitrile and often methyl acrylate as the comonomer. The
material has poor impact resistance and in one industrial process
the copolymer is prepared in the presence of about 10%
butadiene–acrylonitrile rubber by emulsion polymerization. The
product contains some graft copolymer and some polymer blend. In
another process the impact resistance of the copolymer is improved
by biaxial orientation. The package, however, may have a tendency
to shrink at elevated temperature, because the copolymer does not
crystallize.
It is possible to form clear transparent
polyacrylonitrile plastic shapes by a special bulk polymerization
technique [258, 259]. The reaction is initiated with
p-toluenesulfinic
acid–hydrogen peroxide. Initially, heterogeneous polymerizations
take place. They are followed by spontaneous transformations, at
high conversion, to homogeneous, transparent polyacrylonitrile
plastics [260]. A major condition
for forming transparent solid polymer is continuous supply of
monomer to fill the gaps formed by volume contraction during the
polymerization process [261].
Methacrylonitrile,
CH2=C(CH3)CN, can also be prepared by several
routes. Some commercial processes are based on acetone cyanohydrin
intermediate and others on dehydrogenation (or oxydehydrogenation)
of isobutyronitrile. It is also prepared from isobutylene by
ammoxidation:

Just like acrylonitrile, methacrylonitrile does
not polymerize thermally but polymerizes readily in the presence of
free-radical initiators. Unlike polyacrylonitrile,
polymethacrylonitrile is soluble in some ketone solvents. Bulk
polymerizations of methacrylonitrile have the disadvantage of long
reaction time. The rate, however, accelerates with temperature. The
polymer is soluble in the monomer at ambient conditions
[262].
Emulsion polymerization of methacrylonitrile is a
convenient way to form high molecular weight polymers. With proper
choices of emulsifiers, the rates may be increased by increasing
the numbers of particles in the latexes. At a constant rate of
initiation, the degree of polymerization of methacrylonitrile
increases rather than decreases as the rate of polymerization rises
[263].
Methacrylonitrile polymerizes readily in inert
solvents. The polymer, depending on the initiator and on reaction
conditions, is either amorphous or crystalline. Polymerizations
take place over a broad range of temperatures from ambient to −5°C,
when initiated by Grignard reagents, triphenyl ethylsodium, or
sodium in liquid ammonia [264].
The properties of these polymers are essentially the same as those
of the polymers formed by free-radical mechanism.
The homopolymer, prepared by polymerization in
liquid ammonia with sodium initiator at −77°C, is insoluble in
acetone, but it is soluble in dimethylformamide [265]. When it is formed with lithium in liquid
ammonia, at −75°C, the molecular weight of the product increases
with monomer concentration and decreases with initiator
concentration. If, however, potassium initiates the reaction rather
than lithium, the molecular weight is independent of the monomer
concentration [266,
267]. Polymethacrylonitrile
prepared with n-butyllithium in toluene or in dioxane
is crystalline and insoluble in solvents like acetone
[268]. When polymerized in
petroleum ether with n-butyllithium, methacrylonitrile forms
a living polymer [269]. Highly
crystalline polymethacrylonitrile can also be prepared with
beryllium and magnesium alkyls in toluene over a wide range of
temperatures.
6.16 Polyacrylamide, Poly(acrylic acid), and Poly(methacrylic acid)
Commercially, acrylamide is formed from
acrylonitrile by reaction with water. Similarly, the preferred
commercial route to methacrylamide is through methacrylonitrile.
Acrylamide polymerizes by free-radical mechanism [270]. Water is the common solvent for
acrylamide and methacrylamide polymerizations because the polymers
precipitate out from organic solvents.
Crystalline polyacrylamide forms with metal
alkyls in hydrocarbon solvents by anionic mechanism [271]. The product is insoluble in water and in
dimethylformamide.
Both acrylic and methacrylic acids can be
converted to anhydrides and acid chlorides. The acids polymerize in
aqueous systems by free-radical mechanism. Polymerizations of these
monomers in nonpolar solvents like benzene result in precipitations
of the products.
Polymerizations of anhydrides proceed by inter-
and intramolecular propagations [272]:
where R = H, CH3.

The above shown cyclopolymerizations produce
soluble polymers rather than gels.
The acid
chlorides of both acrylic and methacrylic acids polymerize
by free-radical mechanism in dry aromatic and aliphatic solvents.
Molecular weights of the products, however, are low, usually under
10,000 [273, 274]. Polyacrylic and polymethacrylic acids are
used industrially as thickeners in cosmetics, as flocculating
agents, and when copolymerized with divinyl benzene in ion-exchange
resins.
6.17 Halogen-Bearing Polymers
The volume of commercial fluorine-containing
polymers is not large when compared with other polymers like, for
instance, poly(vinyl chloride). Fluoropolymers, however, are
required in many important applications. The main monomers are
tetrafluoroethylene, trifluorochloroethylene, vinyl fluoride,
vinylidine fluoride, and hexafluoropropylene.
6.17.1 Polytetrafluoroethylene
This monomer can be prepared from chloroform
[275]:

Tetrafluoroethylene boils at −76.3°C. It is not
the only product from the above pyrolytic reaction of
difluorochloromethane. Other fluorine by-products form as well and
the monomer must be isolated. The monomer polymerizes in water at
moderate pressures by free-radical mechanism. Various initiators
appear effective [276]. Redox
initiation is preferred. The polymerization reaction is strongly
exothermic, and water helps dissipate the high heat of the
reaction. A runaway, uncontrolled polymerization can lead to
explosive decomposition of the monomer to carbon and carbon
tetrafluoride [277]:

Polytetrafluoroethylene is linear and highly
crystalline [278]. Absence of
terminal
groups shows that few, if any, polymerization terminations occur by
disproportionation but probably all take place by combination
[279]. The molecular weights of
commercially available polymers range from 39,000 to 9,000,000.
Polytetrafluoroethylene is inert to many chemical attacks and is
only swollen by fluorocarbon oils at temperatures above 300°C. The
T m of this
polymer is 327°C and the T
g is below −100°C.

The physical properties of
polytetrafluoroethylene depend upon crystallinity and on the
molecular weight of the polymer. Two crystalline forms are known.
In both cases the chains assume helical arrangements to fit into
the crystalloids. One such arrangement has 15 CF2 groups
per turn and the other has 13.
Polytetrafluoroethylene does not flow even above
its melting point. This is attributed to restricted rotation around
the C–C bonds and to high molecular weights. The stiffness of the
solid polymer is also attributed to restricted rotation. The
polymer exhibits high thermal stability and retains its physical
properties over a wide range of temperatures. The loss of strength
occurs at about the crystalline melting point. It is possible to
use the material for long periods at 300°C without any significant
loss of its strength.
6.17.2 Polychlorotrifluoroethylene
The monomer can be prepared by dechlorination of
trichlorotrifluoroethane with zinc dust and ethanol.

It is a toxic gas that boils at −26.8°C.
Polymerization of chlorotrifluoroethylene is usually carried out
commercially by free-radical suspension polymerization. Reaction
temperatures are kept between 0 and 40°C to obtain a high molecular
weight product. A redox initiation based on reactions of
persulfate, bisulfite, and ferrous ions is often used. Commercial
polymers range in molecular weights from 50,000 to 500,000.
Polychlorotrifluoroethylene exhibits greater
strength, hardness, and creep resistance than does
polytetrafluoroethylene. Due to the presence of chlorine atoms in
the chains, however, packing cannot be as tight as in
polytetrafluoroethylene, and it melts at a lower temperature. The
melting point is 214°C. The degree of crystallinity varies from 30
to 85%, depending upon the thermal history of the polymer.
Polytrifluorochloroethylene is soluble in certain chloro fluoro
compounds above 100°C. It flows above its melting point. The
chemical resistance of this material is good, but inferior to
polytetrafluoroethylene.
6.17.3 Poly(vinylidine fluoride)
The monomer can be prepared by
dehydrochlorination of 1,1,1-chlorodifluoroethane:
or by dechlorination of 1,2-dichloro-1,1-difluoroethane
[280]:


Vinylidine fluoride boils at −84°C. The monomer
is polymerized in aqueous systems under pressure. Details of the
process, however, are kept as trade secrets. Two different
molecular weight materials are available commercially, 300,000 and
6,000,000. Poly(vinylidine fluoride) is crystalline and melts at
171°C. The material exhibits fair resistance to solvents and
chemicals, but is inferior to polytetrafluoroethylene and to
polytrifluorochloroethylene.
6.17.4 Poly(vinyl fluoride)
Vinyl fluoride monomer can be prepared by
addition of HF to acetylene. The monomer is a gas at room
temperature and boils at −72.2°C. Commercially, vinyl fluoride is
polymerized in aqueous medium using either redox initiation or one
from thermal decomposition of peroxides. Pressures of up to
1,000 atm may be used. Radicals generated at temperatures
between 50 and 100°C yield very high molecular weight
polymers.
Poly(vinyl fluoride) is moderately crystalline.
The crystal melting point, T m, is approximately 200°C.
The high molecular weight polymers dissolve in dimethylformamide
and in tetramethyl urea at temperatures above 100°C. The polymer is
very resistant to hydrolytic attack. It does, however, loose HF at
elevated temperatures.
6.17.5 Copolymers of Fluoroolefins
Mary different copolymers of fluoroolefins are
possible and were reported in the literature. Commercial use of
fluoroolefin copolymers, however, is restricted mainly to
elastomers. Such materials offer superior solvent resistance and
good thermal stability.
The elastomers that are most important
industrially are vinylidine fluoride–chlorotrifluoroethylene
[260] and vinylidine
fluoride–hexafluoropropylene copolymers [282]. These copolymers are amorphous due to
irregularities in their structures and can range in properties from
resinous to elastomeric, depending upon composition [283]. Those that contain 50–70 mole percent of
vinylidine fluoride are elastomers. The T g ranges from 0 to −15°C,
also depending upon vinylidine fluoride content [284]. They may be cross-linked with various
peroxides, polyamines [284], or
ionizing radiation. The cross-linking reactions by peroxides take
place through hydrogen abstraction by primary radicals:

Copolymers of vinylidine fluoride with
hexafluoropropylene are prepared in aqueous dispersions using
persulfate initiators. Hexafluoropropylene does not homopolymerize
but it does copolymerize. This means that its content in the
copolymer cannot exceed 50%. Preferred compositions appear to
contain about 80% of vinylidine fluoride. The cross-linking
reactions with diamines are not completely understood. It is
believed that the reaction takes place in two steps [285, 286]. In
the first one, a dehydrofluorination occurs:

The above elimination is catalyzed by basic
materials. These may be in the form of MgO, which is often included
in the reaction medium. In the second step the amine groups add
across the double bonds:

Free diamines, used for cross-linking, are too
reactive and can cause premature gelation. It is common practice,
therefore, to add these diamine compounds in the form of
carbamates, like ethylenediamine carbamate or hexamethylene diamine
carbamate. The above fluoro elastomers exhibit good resistance to
chemicals and maintain useful properties from −50 to +300°C.
Copolymers of tetrafluoroethylene with
hexafluoropropylene are truly thermoplastic polyperfluoroolefins
that can be fabricated by common techniques. Such copolymers soften
at about 285°C and have a continuous use temperature of −260 to
+205°C. Their properties are similar to, though somewhat inferior
to, polytetrafluoroethylene.
6.17.6 Miscellaneous Fluorine Containing Chain-Growth Polymers
One of the miscellaneous fluoroolefin polymers is
a copolymer of trifluoronitrosomethane and tetrafluoroethylene
[287], an elastomer:

It can be formed by suspension polymerization.
One procedure is to carry out the reaction in an aqueous solution
of lithium bromide at −25°C with magnesium carbonate as the
suspending agent. No initiator is added and the reaction takes
about 20 h. Because the reaction in inhibited by hydroquinone
and accelerated by ultra-violet light, it is believed to take place
by a free-radical mechanism. Whether it is chain-growth
polymerization, however, is not certain. A 1:1 copolymer is always
formed regardless of the composition of the monomer feed, and the
copolymerization takes place only at low temperatures. At elevated
temperatures, however, cyclic oxazetidines form instead:

Two polyfluoroacrylates are manufactured on a
small commercial scale for some special uses in jet engines. These
are poly(1,1-dihyroperfluorobutyl acrylate):
and poly(3-perfluoromethoxy-1,1-dihydroperfluoropropyl acrylate):


The polymers are prepared by emulsion
polymerization with persulfate initiators.
Although many other fluorine containing polymers
were described in the literature, it is not possible to describe
all of them here. They are not utilized commercially on a large
scale. A few, however, will be mentioned as examples. One of them
is polyfluoroprene [288]:

The polymer is formed by free-radical mechanism,
in an emulsion polymerization using redox initiation. All three
possible placements of the monomer occur [267].
Polyfluorostyrenes are described in many
publications. A β-fluorostyrene can be formed by cationic mechanism
[289]. The material softens at
240–260°C. An α,β,β-trifluorostyrene can be polymerized by
free-radical mechanism to yield an amorphous polymer that softens
at 240°C [290]. Ring-substituted
styrenes apparently polymerize similarly to styrene. Isotactic
poly(o-fluorostyrene) melts
at 265°C. It forms by polymerization with Ziegler–Natta catalysts
[291]. The meta analog, however, polymerized under
the same conditions yields an amorphous material [291].
6.17.7 Poly(vinyl chloride)
Poly(vinyl chloride) is used in industry on a
very large scale in many applications, such as rigid plastics,
plastisols, and surface coatings. The monomer, vinyl chloride, can
be prepared from acetylene:

The reaction is exothermic and requires cooling
to maintain the temperature between 100 and 108°C.
The monomer can also be prepared from ethylene:

The reaction of dehydrochlorination is carried
out at elevated pressure of about 3 atm.
Free-radical polymerization of vinyl chloride was
studied extensively. For reactions that are carried out in bulk the
following observations were made [292]:
1.
The polymer is insoluble in the monomer and
precipitates out during the polymerization.
2.
The polymerization rate accelerates from the
start of the reaction. Vinyl chloride is a relatively unreactive
monomer. The main sites of initiation occur in the continuous
monomer phase.
3.
The molecular weight of the product does not
depend upon conversion nor does it depend upon the concentration of
the initiator.
4.
The molecular weight of the polymer increases as
the temperature of the polymerization decreases. The maximum for
this relationship, however, is at 30°C.
There is autoacceleration in bulk polymerization
rate of vinyl chloride [293]. It
was suggested by Schindler and Breitenbach [294] that the acceleration is due to trapped
radicals that are present in the precipitated polymer swollen by
monomer molecules. This influences the rate of the termination that
decreases progressively with the extent of the reaction, while the
propagation rate remains constant. The autocatalytic effect in
vinyl chloride bulk polymerizations, however, depends on the type
of initiator used [295]. Thus,
when 2, 2′-azobisisobutyronitrile initiates the polymerization, the
autocatalytic effect can be observed up to 80% of conversion. Yet,
when benzoyl peroxide initiates the reaction, it only occurs up to
20–30% of conversion.
When vinyl chloride is polymerized in solution,
there is no autoacceleration. Also, a major feature of vinyl
chloride free-radical polymerization is chain transferring to
monomer [296]. This is supported
by experimental evidence [297,
298]. In addition, the growing
radical chains can terminate by chain transferring to “dead”
polymer molecules. The propagations then proceed from the polymer
backbone [297]. Such new growth
radicals, however, are probably short lived as they are destroyed
by transfer to monomer [299].
The 13C NMR spectroscopy of poly(vinyl
chloride), which was reduced with tributyltin hydride, showed that
the original polymer contained a number of short four-carbon
branches [300]. This, however,
may not be typical of all poly(vinyl chloride) polymers formed by
free-radical polymerization. It conflicts with other evidence from
13C NMR spectroscopy that chloromethyl groups are the
principal short chain branches in poly(vinyl chloride)
[301, 302]. The pendant chloromethyl groups were
found to occur with a frequency of 2–3/1,000 carbons. The formation
of these branches, as seen by Bovey and coworkers, depends upon
head to head additions of monomers during the polymer formation.
Such additions are followed by 1,2 chlorine shifts with subsequent
propagations [301, 302]. Evidence from still other studies also
shows that some head to head placement occurs in the growth
reaction [303]. It was suggested
that this may be not only an essential step in formation of
branches but also one leading to formation of unsaturation at the
chain ends [303, 304]:

Poly(vinyl chloride) prepared with boron alkyl
catalysts at low temperatures possesses higher amounts of
syndiotactic placement and is essentially free from branches
[305–307].
Many attempts were made to polymerize vinyl
chloride by ionic mechanisms using different organometallic
compounds, some in combinations with metal salts [308–312].
Attempts were also made to polymerize vinyl chloride with
Ziegler–Natta catalysts complexed with Lewis bases. To date,
however, it has not been established unequivocally that vinyl
chloride does polymerize by ionic mechanism. Use of the above
catalysts did yield polymers with higher crystallinity. These
reactions, however, were carried out at low temperatures where
greater amount of syndiotactic placement occurs by the free-radical
mechanism [313]. Vinyl chloride
was also polymerized by
AlCl(C2H5)O2H5 + VO(C3H7O2)
without Lewis bases [312]. Here
too, however, the evidence indicates a free-radical
mechanism.
On the other hand, butyllithium–aluminum
alkyl-initiated polymerizations of vinyl chloride are unaffected by
free-radical inhibitors [313].
Also the molecular weights of the resultant polymers are unaffected
by additions of CCl4 that acts as a chain transferring
agent in free-radical polymerizations. This suggests an ionic
mechanism of chain growth. Furthermore, the reactivity ratios in
copolymerization reactions by this catalytic system differ from
those in typical free-radical polymerizations [313]. An anionic mechanism was also postulated
for polymerization of vinyl chloride with t-butylmagnesium in tetrahydrofuran
[314].
Commercially, by far the biggest amount of
poly(vinyl chloride) homopolymer is produced by suspension
polymerization and to a lesser extent by emulsion and bulk
polymerization. Very little polymer is formed by solution
polymerization.
One process for bulk polymerization of vinyl
chloride was developed in France where the initiator and monomer
are heated at 60°C for approximately 12 h inside a rotating
drum containing stainless steel balls. Typical initiators for this
reaction are benzoyl peroxide or azobisisobutyronitrile. The speed
of rotation of the drum controls the particle size of the final
product. The process is also carried out in a two-reactor
arrangement. In the first one approximately 10% of the monomer is
converted. The material is then transferred to the second reactor
where the polymerization is continued until it reaches 75–80%
conversion. Special ribbon blenders are present in the second
reactor. Control of the operation in the second reactor is quite
critical [315].
Industrial suspension polymerizations of vinyl
chloride are often carried out in large batch reactors or stirred
jacketed autoclaves. Continuous reactors, however, have been
introduced in several manufacturing facilities [315]. Typical recipes call for 100 parts of
vinyl chloride for 180 parts of water, a suspending agent, like
maleic acid–vinyl acetate copolymer, a chain transferring agent,
and a monomer soluble initiator. The reaction may be carried out at
100 lb/in.2 pressure and 50°C for approximately
15 h. As the monomer is consumed the pressure drops. The
reaction is stopped at an internal pressure of about
10 lb/in.2 and remaining monomer (about 10%) is
drawn off and recycled. The product is discharged.
Emulsion polymerizations of vinyl chloride are
usually conducted with redox initiation. Such reactions are rapid
and can be carried out at 20°C in 1–2 h with a high degree of
conversion. Commercial poly(vinyl chloride)s range in molecular
weights from 40,000 to 80,000. The polymers are mostly amorphous
with small amounts of crystallinity, about 5%. The crystalline
areas are syndiotactic [317,
318].
Poly(vinyl chloride) is soluble at room
temperature in oxygen-containing solvents, such as ketones, esters,
ethers, and others. It is also soluble in chlorinated solvents. The
polymer, however, is not soluble in aliphatic and aromatic
hydrocarbons. It is unaffected by acid and alkali solutions but has
poor heat and light stability. Poly(vinyl chloride) degrades at
temperatures of 70°C or higher or when exposed to sun light, unless
it is stabilized. Heating changes the material from colorless to
yellow, orange, brown, and finally black. Many compounds tend to
stabilize poly(vinyl chloride). The more important ones include
lead compounds, like dibasic lead phthalate and lead carbonate.
Also effective are metal salts, like barium, calcium, and zinc
octoates, stearates, and laurates. Organotin compounds, like
dibutyl tin maleate or laurate, also belong to that list.
Epoxidized drying oils are effective heat stabilizers, particularly
in coatings based on poly(vinyl chloride). Some coating materials
may also include aminoplast resins, like
benzoguanamine–formaldehyde condensate.
The process of degradation is complex. It
involves loss of hydrochloric acid. The reactions are free radical
in nature, though some ionic reactions appear to take place as
well. The process of dehydrochlorination results in formations of
long sequences of conjugated double bonds. It is commonly believed
that formation of conjugated polyenes, which are chromophores, is
responsible for the darkening of poly(vinyl chloride). In addition,
the polymer degrades faster in open air than it does in an inert
atmosphere. This shows that oxidation contributes to the
degradation process. All effective stabilizers are hydrochloric
acid scavengers. This feature alone, however, can probably not
account for the stabilization process. There must be some
interaction between the stabilizers and the polymers. Such
interaction might vary, depending upon a particular
stabilizer.
6.17.7.1 Copolymers of Vinyl Chloride
A very common copolymer of vinyl chloride is
vinyl acetate. Copolymerization with vinyl acetate improves
stability and molding characteristics. The copolymers are also used
as fibers and as coatings. Copolymers intended for use in moldings
are usually prepared by suspension polymerization. Those intended
for coating purposes are prepared by solution, emulsion, and
suspension polymerizations. The copolymers used in molding
typically contain about 10% of poly(vinyl acetate). Copolymers that
are prepared for coating purposes can contain from 10 to 17% of
poly(vinyl acetate). For coatings, a third comonomer may be
included in some resins. This third component may, for instance, be
maleic anhydride, in small quantities, like 1%, to improve adhesion
to surfaces.
Copolymers of vinyl chloride with vinylidine
chloride are similar in properties to copolymers with vinyl
acetate. They contain from 5 to 12% of poly(vinylidine chloride)
and are intended for use in stabilized calendaring.
Copolymers containing 60% vinyl chloride and 40%
acrylonitrile are used in fibers. The fibers are spun from acetone
solution. They are nonflammable and have good chemical
resistance.
6.17.8 Poly(vinylidine chloride)
Vinylidine chloride homopolymers form readily by
free-radical polymerization, but lack sufficient thermal stability
for commercial use. Copolymers, however, with small amounts of
comonomers find many applications.
The monomer, vinylidine chloride, can be prepared
by dehydrochlorination of 1,1,2-trichloroethylene:

It is a colorless liquid that boils at 32°C.
Also, it is rather hard to handle as it polymerizes on standing.
This takes place upon exposure to air, water, or light. Storage
under an inert atmosphere does not completely prevent polymer
formation.
Poly(vinylidine chloride) can be formed in bulk,
solution, suspension, and emulsion polymerization processes. The
products are highly crystalline with regular structures and a
melting point of 220°C. The structure can be illustrated as
follows:

This regularity in structure is probably due to
little chain transferring to the polymer backbone during
polymerization. Such regularity of structure allows close packing
of the chains and, as a result, there are no effective solvents for
the polymer at room temperature.
Copolymerization of vinylidine chloride with
vinyl chloride reduces the regularity of the structure. It
increases flexibility and allows processing the polymer at
reasonable temperatures. Due to extensive crystallization, however,
that is still present in 85:15 copolymers of vinylidine chloride
with vinyl chloride, they melt at 170°C. The copolymerization
reactions proceed at slower rates than do homopolymerizations of
either one of the monomers alone. Higher initiator levels and
temperatures are, therefore, used. The molecular weights of the
products range from 20,000 to 50,000. These materials are good
barriers for gases and moisture. This makes them very useful in
films for food packaging. Such films are formed by extrusion and
biaxial orientation. The main application, however, is in
filaments. These are prepared by extrusion and drawing. The tensile
strength of the unoriented material is
10,000 lb/in.2 and the oriented one
30,000 lb/in.2.
Vinylidine chloride is also copolymerized with
acrylonitrile. This copolymer is used mainly as a barrier coating
for paper, polyethylene, and cellophane. It has the advantage of
being heat sealable.
6.18 Poly(vinyl acetate)
Vinyl acetate monomer can be prepared by reacting
acetylene with acetic acid:

The reaction can be carried out in a liquid or in
a vapor phase. A liquid phase reaction requires 75–80°C
temperatures and a mercuric sulfate catalyst. The acetylene gas is
bubbled through glacial acetic acid and acetic anhydride. Vapor
phase reactions are carried out at 210–250°C. Typical catalysts are
cadmium acetate or zinc acetate. There are other routes to vinyl
acetate as well, based on ethylene.
Commercially, poly(vinyl acetate) is formed in
bulk, solution, emulsion, and suspension polymerizations by
free-radical mechanism. In such polymerizations, chain transferring
to the polymer may be as high as 30%. The transfer can be to a
polymer backbone through abstraction of a tertiary hydrogen:

It can also take place to the methyl proton of
the acetate group:

The polymer has a head to tail structure and is
highly branched. It is quite brittle and exhibits cold flow. This
makes it useless as a structural plastic. It is, however, quite
useful as a coating material and as an adhesive for wood. The
polymer is soluble in a wide range of solvents and swells and
softens upon prolonged immersion in water. At higher temperatures
or at extended exposures to temperatures above 70°C, the material
loses acetic acid.
A number of copolymers are known where vinyl
acetate is the major component. In coatings, vinyl acetate is often
used in copolymers with alkyl acrylates (line 2-ethylhexyl
acrylate) or with esters of maleic or fumaric acids. Such
copolymers typically contain 50–20% by weight of the comonomer and
are usually formed by emulsion polymerization in batch processes.
They are used extensively as vehicles for emulsion paints.
Shaver and coworkers [319] investigated the mechanism of
bis(imino)pyridine ligand framework for transition metal
systems-mediated polymerization of vinyl acetate. Initiation using
azobisisobutyronitrile at 120°C results in excellent control over
poly(vinyl acetate) molecular weights and polymer dispersities. The
reaction yields vanadium-terminated polymer chains which can be
readily converted to both proton-terminated poly(vinyl acetate) or
poly(vinyl alcohol). Irreversible halogen transfer from the parent
complex to a radical derived from azobisisobutyronitrile generates
the active species.
6.19 Poly(vinyl alcohol) and Poly(vinyl acetal)s
Vinyl alcohol monomer does not exist because its
keto tautomer is much more stable. Poly(vinyl alcohol) can be
prepared from either poly(vinyl esters) or from poly(vinyl ethers).
Commercially, however, it is prepared exclusively from poly(vinyl
acetate). The preferred procedure is through a transesterification
reaction using methyl or ethyl alcohols. Alkaline catalysts yield
rapid alcoholyses. A typical reaction employs about 1% of sodium
methoxide and can be carried to completion in 1 h at 60°C. The
product is contaminated with sodium acetate that must be removed.
The reaction of transesterification can be illustrated as follows:

The branches of poly(vinyl acetate) that form
during polymerization as a result of chain transferring to the
acetate groups cleave during transesterification. As a result,
poly(vinyl alcohol) is lower in molecular weight than its parent
material.
Poly(vinyl alcohol) is very high in head to tail
structures, based on NMR data. It shows the presence of only a
small amount of adjacent hydroxyl groups. The polymer prepared from
amorphous poly(vinyl acetate) is crystalline, because the
relatively small size of the hydroxyl groups permits the chains to
line-up into crystalline domains. Synthesis of isotactic poly(vinyl
alcohol) was reported from isotactic poly(vinyl ethers), like
poly(benzyl vinyl ether), poly(t-butyl vinyl ether),
poly(trimethylsilyl vinyl ether), and some divinyl compounds.
Poly(vinyl alcohol) is water soluble. The
hydroxyl groups attached to the polymer backbone, however, exert a
significant effect on the solubility. When the ester groups of
poly(vinyl acetate) are cleaved to a hydroxyl content of 87–89%,
the polymer is soluble in cold water. Further cleavage of the ester
groups results in a reduction of the solubility and the products
require heating of the water to 85°C to dissolve. This is due to
strong hydrogen bonding that also causes unplasticized poly(vinyl
alcohol) to decompose below its flow temperature. On the other
hand, due to hydrogen bonding the polymer is very tough.
Poly(vinyl acetals) are prepared by reacting
poly(vinyl alcohol) with aldehydes. Reactions of poly(vinyl
alcohol) with ketones yield ketals. These are not used
commercially.
Not all hydroxyl groups participate in formations
of acetals and some become isolated. A typical poly(vinyl acetal)
contains acetal groups, residual hydroxyl groups, and residual
acetate groups from incomplete transesterification of the parent
polymer.
Poly(vinyl acetal)s can be formed directly from
poly(vinyl acetate) and this is actually done commercially in
preparations of poly(vinyl formal). A typical reaction is carried
out in the presence of acetic acid, formalin, and sulfuric acid
catalyst at 70°C:

Poly(vinyl butyral), on the other hand, is
prepared from poly(vinyl alcohol) and butyraldehyde. Sulfuric acid
is used as the catalyst. Commercially only poly(vinyl formal) and
poly(vinyl butyral) are utilized on a large scale in coating
materials.
6.20 Review Questions
6.20.1 Section 6.1
1.
What are the two types of polyethylene that are
currently manufactured commercially?
2.
Describe the chemical structure of low-density
polyethylene produced by free-radical mechanism and show by
chemical equations how all the groups that are present form. How
can low-density polyethylene be prepared by ionic mechanism?
3.
Describe conditions and procedure for commercial
preparation of polyethylene by free-radical mechanism, the role of
oxygen, and the problems associated with oxygen.
4.
Describe a tubular reactor for preparation of
polyethylene.
5.
What are the industrial conditions for
preparations of high-density polyethylene. Describe the continuous
solution process, the slurry process, and the gas-phase
process.
6.
Show with chemical reactions how polymethylene
forms from diazomethane.
6.20.2 Section 6.2
1.
Discuss high activity catalysts for the
manufacturing of isotactic polypropylene, heterogeneous and
homogeneous.
2.
What are the current techniques for polypropylene
manufacture?
3.
How can syndiotactic polypropylene be prepared
and what are its properties?
6.20.3 Section 6.3
1.
Describe the two industrial processes for
manufacturing polybutylene.
6.20.4 Section 6.4
1.
Draw the chemical structure of isotactic
poly(butene-1). How is it prepared and used?
2.
What is TPX, how is it prepared, and what are its
properties?
6.20.5 Section 6.5
1.
Discuss copolymers of ethylene with propylene.
How are they prepared? What catalysts are used in the preparations?
How are ethylene–propylene rubbers cross-linked?
2.
What are the copolymers of ethylene with higher
α-olefins and why are they prepared and how?
3.
Discuss the copolymers of ethylene with vinyl
acetate? How are they prepared and used?
4.
What are ionomers? Describe each type. How are
they used?
5.
Describe the catalysts used in preparations of
aliphatic ketones by copolymerization of ethylene with carbon
monoxide.
6.20.6 Section 6.6
1.
Discuss polybutadiene homopolymers. How are they
prepared? What are their uses?
2.
What are popcorn polymers? What causes their
formation?
3.
Discuss liquid polybutadienes. How are they
prepared and used?
4.
How are high molecular weight polybutadienes
prepared and used?
5.
Discuss polyisoprenes. What is natural rubber?
Where does it come from? What are synthetic polyisoprenes? How are
they prepared?
6.20.7 Section 6.7
1.
What is methyl rubber?
6.20.8 Section 6.8
1.
What is chloroprene rubber? How is it made and
used?
6.20.9 Section 6.9
1.
What are poly(carboxybutadiene)s?
6.20.10 Section 6.10
1.
Discuss cyclopolymerization of conjugated
dienes.
6.20.11 Section 6.11
1.
What is SBR rubber? Explain and describe
preparation and properties.
2.
What are block copolymer elastomers? How are they
prepared and what gives them their unique properties?
3.
What is GR-N rubber? Explain and describe
preparation and properties.
6.20.12 Section 6.12
1.
How are atactic and syndiotactic polystyrenes
prepared commercially? Describe and explain.
2.
What polymers of substituted styrenes are
available commercially? How are they prepared?
6.20.13 Section 6.13
1.
What is high-impact polystyrene and how is it
prepared?
2.
Discuss ABS resins. How are they prepared?
6.20.14 Section 6.14
1.
Discuss the chemistry of free-radical
polymerization of acrylic and methacrylic esters.
2.
What are acrylic elastomers and how are they
vulcanized?
3.
How is poly(methyl methacrylate) prepared
commercially, such as Plexiglas in the form of sheets and rods? Is
poly(methyl methacrylate) prepared in any other way, how? For what
applications?
4.
Describe the thermosetting acrylic resins used in
industrial coatings. How are they prepared? How are they
cross-linked?
6.20.15 Section 6.15
1.
Discuss industrial polymers and copolymers of
acrylonitrile and methacrylonitrile. How are they prepared and
used?
6.20.16 Section 6.16
1.
Describe preparation and uses of polyacrylamide,
poly(acrylic acid), and polymethacrylic acid.
6.20.17 Section 6.17
1.
How is polytetrafluoroethylene prepared, and what
are its properties and uses?
2.
Discuss the chemistry of
polychlorotrifluoroethylene, poly(vinylidine fluoride), and
poly(vinyl fluoride).
3.
What common copolymers of fluoroolefins are used
commercially?
4.
Discuss the chemistry of poly(vinyl chloride) and
poly(vinylidine chloride).
5.
Discuss the important commercial copolymers of
vinyl chloride. What are their main uses?
6.
Discuss the chemistry of poly(vinylidine
chloride).
6.20.18 Section 6.18
1.
Discuss preparation, properties, and uses of
poly(vinyl acetate).
6.20.19 Section 6.19
1.
How is poly(vinyl alcohol) prepared, used, and
converted to poly(vinyl acetal)s?
References
1.
E.W. Fawcett, O.R. Gibson,
M.W. Perrin, J.G. Paton, and E.G, Williams, Brit. Pat. # 571,590
(1937)
2.
3.
4.
R.O. Gibson, “The Discovery
of Polythene”, Roy
. Inst . Chem ., London Lectures, 1964, 1, 1
5.
E.W. Fawcett, et al., Brit.
Pat, # 471,590 (1937)
6.
R.A. Hines, W.M.D, Bryant,
A.W. Larchar, and D.C. Pease, Ind . Eng . Chem ., 1957, 49(7), 1071
7.
A.L.J. Raum, Chapt, 4 in
Vinyl and Allied Polymers, (P.D. Ritchle, ed.),
CRC Press, Cleveland, Ohio, 1968
8.
Monsanto Co., U.S. Pat. #
2,856,395 and 2,852,501
9.
G. Downs, Chapt. 5 in
Vinyl and Allied Polymers, (P.D. Ritchle, ed.),
CRC Press, Cleveland, Ohio, 1968
10.
11.
A.H. Willbourn, J. Polymer Sci .,1959, 34, 569
12.
D.E. Dorman, E.P. Otacka,
and P.A. Bovey, Macromolecules,1972, 5, 574
13.
W.G. Oakes and R.B.
Richards, J. Chem
. Soc ., 1949, 2929
14.
K.W. Doak, “Low Density
Polyethylene (High Pressure)”, Encyclopedia of Polymer Science and Technology, Vol. 6, 2nd Ed.. (H.F.
Mark, N.M. Bikales, C.G. Overberger, and G. Menges, eds.), Wiley-
Interscience, New York, 1986
15.
K. Weissermel, H. Cherdron,
J. Berthold, B. Diedrich, K.D. Keil, K. Rust, H. Strametz, and T.
Toth, J. Polymer
Sci ., Symposium # 51, 187 (1975)
16.
D.M. Rasmussen, Chem . Eng ., 1972, 104 (Sept. 18)
17.
W. Keim, R. Appel, A.
Storeck, C. Kruger, and R. Goddard, Anew . Chem ., Int . Ed . Engl., 1981, 20 (1), 116
18.
19.
L.K. Johnson, C.M. Killian,
and M. Brookhart, J. Am. Chem.
Soc., 1995
, 117, 6441
20.
M. Brookhart et al., U.S.
Patent # 5 866 663 (Feb. 2, 1999); from Chemtech, 1999 ,
29 (2), 39
21.
B.L. Small, M. Brookhart,
A.M.A. Bennett, J. Am. Chem.
Soc., 120, 4049,
(1998); A.M.A. Bennett, Chemtech, 1999, 29 (7), 24
22.
G. J. P.; Baugh, Simon P.
D.; Hoarau, O., Gibeon, Vernon C.; Wass, D. F.; White, A. J. P.;
Williams, D. J.. Inorganica
Chimica Acta, 2003,
345, 279
23.
T. Senda, H. Hanaoka, T.
Hino,Y. Oda, H. Tsurugi and K. Mashima, Macromolecules, 2009, 42 (21), 8006
24.
J. H. Woo, Y.-S. Ha, Y.-J.
Shin, and S. C. Hong, Am.Chem.
Soc. Polymer Preprints, 2009, 50 (1),196
25.
I.D. Burdett, Chemtech, 1992, 22, 616
26.
27.
H. Pechmann, Ber., 1898, 31, 2643
28.
E. Bamberger and F.
Tschirner, Ber.,
1900, 33, 955
29.
G.D. Buckley, L.H. Cross,
and N.H. Roy, J. Chem
. Soc., 1950, 2714
30.
A. Renfrew and P. Morgan,
Polythene,
Wiley-Interscience, New York, N.Y., 1957; H. Meerwein, Angew . Chem., 1948, A60, 78
31.
C.E.H. Bawn, A. Ledwith and
P. Matthies, J. Polymer
Sci., 1959, 34, 93
32.
C.E.H. Bawn and A. Ledwith,
Chem . and
Ind . (London), 1957, 1180
33.
G.D. Buckley and N.H. Ray,
J. Chem . Soc ., 1952, 3701
34.
A.G. Nasini, G. Saini, and
T. Trossarelli, J. Polymer
Sci., 1960, 48, 435; Pure and Appl . Chem ., 1962, 4, 255
35.
M.G, Krakonyak and S.S.
Skorokhodov, Vysokomol
. Soyed ., 1969, A11(4), 797; H. Pechler, B. Firnhaber,
D.Kroussis, and A, Dawallu, Makromol . Chem ., 1964, 70, 12
36.
G.A. Mortimer and L.C.
Arnold, J. Polymer
Sci ., 1964, A2, 4247
37.
G. Natta, P. Pino, G
Mazzanti, and P. Longi, Gazz . Chim . Ital .,1957, 87, 570 (from a private
translation)
38.
G. Natta, I. Pasquon, A.
Zambelli, and G. Gatti, J.
Polymer Sci
.,1961, 51, 387
39.
Avison Corp. U.S. Patents #
3,134,642 (May 26, 1964);
2,980,664 (April 18, 1961);
3,055,878 (Sept,25, 1962); 3,216,987 (Nov. 9, 1596); 3,303,179
(Feb. 7, 1967); 3,328,375 (June 27, 1967); 3,362,916 (Jan 9 1964);
3,313,791 (April 11, 1967);
3,255,167 (June 7, 1966); German Patent # 1,236,789 (March 16,
1967); 1,234,218 (April 6, 1967)
40.
Eastman Kodak Co., U.S.
Patents # 3,232,919 (Feb. 1, 1966); 3,058,696 (Oct. 10, 1962);
3,081,287 (March 12, 1963) 2,962,487 (Nov. 29, 1960); 3,143,537
(July 21, 1964); 3,004,015 (Oct. 10, 1965); 3,149,097 (Sept. 15,
1964); 3,201,379 (Aug. 17, 1965); 3,213,073 (Oct. 19, 1965);
3,230,208 (Jan. 18, 1966); 3,149,097 (Sept. 15, 1964); 3,186,977
(June 1, 1965); 3,189,590 (June 15, 1965); 3,072,629 (Jan 8, 1963);
3,178,401 (April 13, 1965) 3,284,427 (Nov. 8, 1966); British
Patents # 921,039 (March 13, 1963); 930,633 (July 3, 1963);
1,007,030 (Oct. 13, 1965); 1,00,348 (Aug. 4, 1965); 1,000,720 (Aug
11, 1965) French Patents # 1,315,782 (Jan 25, 1963)
41.
Montecantini-Edison Sp.A.,
U.S. Patents # 3,112,300 (Nov. 26, 1963); 3,259, 613 (July 5,
1966); 3,139,418 (June 30, 1964); 3,141,872 (July 21, 1964);
3,277,069 (Oct 4, 1969); 3,252,954 (May 25, 1966); Brit. Patents #
1,014,944 (Jan 10, 1964); Canad. Patents # 649,164 (Sept. 25,
1962); German Patents # 1,238,667 (April 13, 1967); 1,214 000 (June
6, 1961); Italian Patents # (from Chem. Abstr.) 646,950 (Sept. 20,
1965)
42.
Phillips Petroleum Co. U.S.
Patents # 3,280,092 (Oct. 18, 1966); 3,119,798 (Jan. 28, 1964);
3,147,241 (Sept. 1, 1964); 3,182,049 (May 4, 1965): 3,210,332 (Oct.
5, 1965); 3,317,502 (May 2, 1967): Brit. Patents # 1,017,988 (Jan
26, 1966): 1,034,155 (June 29,1966); 1,119,033 (March 7, 19680);
Belgian Patent # 695,060 (June 9, 1967) German Patent # 1,266,504
(April 18, 1968)
43.
Shell Oil Co. U.S. Patents #
3,147,238 (Sept. 1, 1964); 3,240,773 (March 15, 1966); 3,282,906
(Nov.1,1966): 3,398,130 (Aug. 20, 1968); 3,311,603 (March 28,
1967): 3,394,118 (July 23, 1968); 3,264,277 (Aug, 2, 1966)
Brit.Patent # 1,006,919 (Oct, 6, 1965)
44.
U. Giannini, A Casata, P.
Lorgi, and R. Mazzocchi, Germ. Pat. Appl. 2,346,577 (1974)
45.
H.R. Sailors and J.P. Hogam,
J. Macromol . Sci .- Chem., 1981 ,
A15, 1377
46.
P. Pino and R. Mulhaupt,
Angew . Chem ., Int . Ed ., Engl ., 1980, 19, 857
47.
K. Weissermel, H. Cherdron,
J. Berthold, B. Diedrich, K.D. Keil, K. Rust, H. Strimetz, and T.
Toth, J. Polymer
Sci., Sympos . 1975, #51, 187
48.
S. Sivaram, Ind . Eng . Chem . Prod . Res . Dev., 1977, 16(2), 121
49.
E.E. Vermel, V.A. Zakharov,
Z.K. Bukatova, G.P. Shkurina, S.G. Echevskaya, E.M. Moroz, and S.V.
Sudakova, Vysokomol
. Soyed ., 1980, 22(1), 22
50.
T.W. Campbell and A.C.
Haven, Jr., J. Appl.
Polymer Sci .,1959, 1, 73; P. Carradini, V. Busico, and G.
Guerra, “Monoalkene Polymerization: Stereochemistry,” Chap. 3 in
Comprehensive Polymer
Science, Vol.4, (G.C.
Eastmond, A. Ledwith, S. Russo, P. Sigwalt, eds.), Pergamon Press,
Oxford, 1989.
51.
U. Giannini, Makromol. Chem ., Suppl.,1981, 5, 216
52.
K.H. Reichert and H.U.
Moritz, J. Appl.
Polymer Sci., 1981, 36, 151
53.
J. Lewis, P.E. Okieimen, and
G.S. Park, J. Macromol.
Sci .- Chem., 1982, A17, 1021
54.
55.
D. O’Sullivan, Chem. Eng. News, p. 29 (July 4, 1983)
56.
57.
58.
B. Rieger. X. Mu, D.T.
Mallin, M.D. Rausch, and J.C.W. Chien, Macromolecules, 1990, 23, 3559
59.
H.N. Cheng and J.A. Ewen,
Makromol. Chem., 1989, 190, 823
60.
A. Grassi, A. Zambelli, L.
Resconi, E. Albizzati, and R. Mazzocchi, Macromolecules, 1988, 21, 617CrossRef
61.
G.G. Arzonmanidis and N.M.
Karayannis, Chemtech,
1993, 23, 43
62.
63.
B., Vincenzo; C., V. Van
Axel; A., P. C., Roberta; S. Annalaura; T. Giovanni; and M.
Vacatello, J. Am. Chem
Soc., 2003,
125(18), 5451—5460
64.
J.C.W. Chien, M.D. Iwamoto,
M.D. Rausch, W. Wedler, and H.H. Winter, Macromolecules, 1997, 30, 3447CrossRef
65.
J. Boor Jr. and E.A.
Youngman, J. Polymer
Sci.,1966, A-1,4, 1861
66.
67.
A. Zambelli, G. Natta, and
I. Pasquon, J. Polymer
Sci., 1963, C,4,411
68.
V.W. Buls and T.L. Higgins,
J. Polymer Sci., Polymer Chem. Ed., 1973, 11, 925; Y. Doi, S. Ueki, and T. Keii,
Macromolecules,
1979, 12(5), 814
69.
I.G. Farbenindustri, A.-G.,
German Pat. # 634,2 78 (1936)
70.
J.Brandrup and E.H.
Immergut, Polymer Handbook,
Wiley-Interscience, New York, 1989
71.
T. Asanuma, Y. Nishimori, M.
Ito, N. Uchikawa, and T. Shiomura, Polym. Bull.,1991, 25, 567
72.
R.M. Thomas, W.J. Sparks,
P.K. Frolich, M. Otto and M. Mueller-Cuuradi, J. Am. Chem. Soc.,1940, 62, 276
73.
R.B. Cundall, The Chemistry of Cationic Polymerization, (P.H. Plesch, ed.),
Pergamon Press, Macmillan Co., New York, 1963
74.
J.P. Kennedy in Polymer Chemistry of Synthetic Elastomers; Part 1, J.P. Kennedy and
E.G.M. Tornquist, ed. Wiley-Interscience, New York, 1968
75.
A. Gandini and H. Cheradame,
“Cationic Polymerization” in Encyclopedia of Polymer Science and Engineering, Vol. 2, H.F. Mark,
N.M. Bikales, C.G. Overberger, and G. Menges, eds., Wiley
Interscience, New-York, 1985
76.
N.G. Gaylord and H.F. Mark,
“Linear and Stereoregular Addition Polymers”, Interscience, N.Y.,
1959
77.
78.
79.
G. Natta, G. Mazzanti, A.
Valvassori, and G. Pajaro, Chim. Ind. (Milan), 1957, 39, 733
80.
K.R. Dunham, J.
Vanderberghe, J.W.H. Faber, and L.E. Contois, J. Polymer Sci., 1963, A-1,1, 751
81.
G. Natta, P. Carradini, and
I.W. Bassi, Gazz.
Chim. Ital.,1959, 89, 784
82.
H.W. Coover Jr., and F.B.
Joyner, J. Polymer
Sci., 1965, A-1,3, 2407
83.
84.
W. Passmann, Ind. and Eng. Chem., 1970, 62(5), 48
85.
H. Yasuda, E. Ihara, M.
Morimoto, M. Nodono, S. Yoshioka, and M. Furo, Macromol. Symp. 1995 ,
95, 203CrossRef
86.
W. Marconi, S. Cesea, and G.
Della Foruna, Chim.
Ind. (Milan). 1964 ,
46, 1131
87.
S. Cesea, S. Arrighetti, and
W. Marconi, Chim.
Ind. (Milan) 1968, 50, 171
88.
S. Cesea, G. Bertolini, G.
Santi, and P.V. Duranti, J.
Polymer Sci.,
1971, A-1,9, 1575
89.
E.K. Gladding, B.S. Fischer,
and J.W. Collette, Ind.
Eng. Chem., Prod. Res. Devel. 1962,, 1, 65
90.
E.W. Duck and W. Cooper,
23-rd IUPAC Congress. Macromol. Prepr., 1971, 2, 722
91.
E.K. Easterbrook, T.J.
Brett, F.C. Loveless, and D.N. Mathews, 23-rd IUPAC. Congress, Macromol. Prepr., 1971 ,
2, 712
92.
93.
A. Akimoto, J. Polymer Sci., Polymer Chem. Ed., 1972, 10, 3113
94.
C.A. Lukach and H.M.
Spurlin, in Copolymerization, (G.E. Ham, ed.),
Interscience, N.Y., 1964
95.
G. Natta, G. Mazzanti, A.
Valvassori, G. Sartori, and A. Barbagallo, J. Polymer Sci., 1961 ,
51, 411, 429
96.
G. Bier and G. Lehmann, in
Copolymerization, G.E. Ham,
ed., Interscience, N.Y., 1964.
97.
J,G, Hefner and B.W.S.
Kolthammer, U.S. Patent 5,554,777 (Sept 10, 1996)
98.
Chemical and Engineering
News, October 5, 1998,
p.18
99.
P.G. Hustad, J.D. Weinhord,
G.R. Marchnd, E.I. Garcia-Maitin, and P.L. Robers, Macromolecules DOI:10/1021/ma9002819
from Chem.Eng.News, May 4,
2009
100.
Kenichi Kakinuki, Michiya
Fujiki, and Kotohiro Nomura, Macromolecules, 2009, 10, 1921/ma900576v
101.
102.
103.
R.F. Jordan, Am. Chem. Soc. Polymer Preprints,
2010, 51 (2), 372
104.
J. Furukawa, Angew. Makromol. Chem., 1972 ,
23, 189
105.
J. Furukawa, S. Tsuruki,
and J.Kiji J. Polymer Sci.,
Polymer Chem. Ed. ,1973 ,
11, 1819
106.
107.
M.J. Wisotsky, A.B. Kober,
and I.A. Zlochower, J.
Appl. Polymer Sci.,
1971, 15, 1737
108.
T.E. Ferrington and A.V.
Tobolsky, J. Polymer
Sci., 1958, 31, 25
109.
110.
111.
112.
F.P. Gintz, Chapt. 11
Vinyl and Allied Polymers, P.D. Ritchie, ed., Vol
I, CRC Press, Cleveland, Ohio, 1968
113.
V.D. Mochel, J.. Polymer Sci., 1972, A-1,10, 1009
114.
L. Annunziata, S.
PragIlola, D.Pappalardo, C. Tedesco, and C. Pellecchia,
Macromolecules,
2011, 44(7), 1934
115.
W.M. Saltman et al,
J. Am. Chem. Soc., 1958 ,
80, 5615
116.
W.M. Saltman, J. Polymer Sci., 1963, A-1,1, 373
117.
118.
A.V. Tobolsky, J. Polymer Sci., 1959, 40, 73
119.
F.W. Stavely et al,
Ind. and Eng. Chem., 1956, 48(4), 778
120.
G.A. Thomas and W.H.
Carmody, J. Am.
Chem. Soc .,1933, 55, 3854
121.
J.C. D’Ianni, Kautschuk Gummi Kunststoffe, 1966, 19(3), 138; J.D. D’ianni, F.J. Naples,
and J.E. Field, Ind.
and Eng. Chem.,1950, 41, 95; S.. Kostjuk, S. Ouardad, F.
Peruch, A. Deffieux, C. Absalon, J. E. Puskas, and F. Ganachaud,
Macromolecules,
2011, 44, 1372
122.
J.E. Field and M. Feller,
U.S. Patent # 2,728,758 (Dec. 27, 1955)
123.
H.E. Adams, R.S. Stearns,
W.A. Smith, and J.L Binder, Ind. Eng. Chem., 1958, 50, 1507
124.
S. Minekawa et al, Japanese
Patent # 45-36519 (Nov. 20, 1970)
125.
T.F. Yen, J. Polymer Sci., 1959, 33, 535
126.
O. Solomon and C. Amrus,
Rev. Chem. (Romania), 1958, 9, 150
127.
M. Gippin, U.S. Patent
3,442,878 (May 6, 1969)
128.
A.L. Klebanski,
J. Polymer Sci.,
1958, 30, 363; Rubber Chem.
Technol., 1959, 32, 588
129.
J.T. Maynard and W.E.
Mochel, J. Polymer
Sci., 1954 ,
13, 235, 251
130.
R.R. Garrett, C.D.
Hargreaves, II, and D.N. Robinson, J. Macromol. Sci.-Chem.1970 A4(8), 1679
131.
132.
C.E.N. Bawn, D.G.T. Cooper,
and A.M. North, Polymer,
1966, 7(3), 113
133.
W. Marconi, A. Mazzei, S.
Cucinella, and M. Cesari, J.
Polymer Sci.,
1964, A-1,2, 426
134.
G. Natta, M. Farina, M.
Donati, and M. Peraldo, Chim. Ind. (Milan), 1960, 42, 1363
135.
136.
G. Natta, P, Corradini, and
P. Canis, J. Polymer
Sci., 1965, A-1,3, 11; G. Natta, L. Porri, P.
Corradini, G. Zanini, and F. Ciampelli, J. Polymer Sci., 1961 ,
51, 463; G. Natta, L. Porri, G. Stoppa, G. Allegra, and F,
Ciampelli, J. Polymer
Sci., 1963, B,1, 67.
137.
G. Natta, L. Porri, A.
Carbonaro, F. Ciampelli, and G. Allegro, Makromol. Chem., 1962, 51, 229CrossRef
138.
139.
G. Natta and L. Porri,
A.C.S. Polymer Preprints, 5(2), 1163 (1964); K. Bujadoux, J.
Josefonvicz and J. Neel, Eur. Polymer J., 1970, 6, 1233
140.
A.L. Halasa, D.F. Lohr, and
J.E. Hall, J. Polymer
Sci., Polymer Chem. Ed., 1981, 19, 1357
141.
N.G. Gaylord, I. Kossler,
M. Stolka, and J. Vodehnal, J.
Polymer Sci.
,1964 , A-1,2, 3969
142.
N.G. Gaylord, I, Kossler,
and M. Stolka, J. Macromol.
Sci.,- Chem., 1968, A2(2), 421
143.
K. Hasegawa and R. Asami,
J. Polymer Sci., Polymer Chem. Ed. 1978, 16,1449; K. Hasegawa, R. Asami, and T.
Higashimura, Macromolecules
, 1977, 10, 585,592
144.
I.I. Yermakova, Ye.N.
Kropacheva, A.1. Kol’tsov, and B.A. Dolgoplosk, Vysokomol. Soyed., 1969, A11(7), 1639
145.
N.G. Gaylord, M. Stolka, V.
Stepan, and I. Kossler, J.
Polymer Sci.,1968, C(23), 317
146.
E.R. Moore, (ed.), “Styrene
Polymers” in Encyclopedia
of Polymer Science
and Engineering, Vol. 16,
2nd ed., (H.F Mark, N.M. Bikales, C.G. Overberger, and G. Menges,
eds.), Wiley-Interscience, New York, 1989
147.
T. Takamatsu, A. Ohnishi,
T. Nishikida and J. Furukawa, Rubber Age,1973, 23
148.
E. Simon, Ann., 1839, 31, 265
149.
H. Stobbe and G. Posnjack,
Ann.,1909, 259
150.
G.V. Schulz, A. Physik. Chem. (Leipzig), 1939, B44, 277
151.
152.
153.
154.
F.R. Mayo, J. Am. Chem. Soc., 1968 ,
90, 1289
155.
156.
F.A. Bovey and I.M.
Kolthoff, Chem.
Rev., 1948 ,
42, 491
157.
O. Roehm, German Pat, #
656,469 (1938)
158.
N. Platzer, Ind. and Eng. Chem., 1970 ,
62(1), 6
159.
I.G.Farbenindustri, A.
-G.,German Patent #634,278 (1936)
160.
Dow Chem, Co. U.S. Patent #
2,727,884 (1955)
161.
J.P. Kennedy, J. Macromol. Sci .- Chem., 1969, A3, 861
162.
J.P. Kennedy, J. Macromol. Sci .- Chem., 1969, A3, 885
163.
F.S. D’yachkovskii, G.A.
Kazaryan, and N.S. Yenikolopyan, Vysokomol. Soyed., 1969, A11, 822
164.
A.A. Morton and E.
Grovenstein Jr., J. Am.
Chem. Soc., 1952, 74, 5435
165.
J.L.R. Wllliams, T.M.
Laasko, and W.J. Dulmage, J.
Org. Chem.,
1958, 23, 638, 1206
166.
D. Baum, W. Betz, and W.
Kern, Makromol.
Chem.,1960, 42, 89
167.
168.
K.F. O’Drlscoll and A.V.
Tobolsky, J. Polymer
Sci., 1959,35, 259
169.
R.C.P. Cubbon and D.
Margerison, Proc.
Roy. Soc. (London), 1962, A268, 260
170.
A.A. Morton and L.D.
Taylor, J. Polymer
Sci., 1959 ,
38, 7
171.
172.
G. Natta, J. Polymer Sci., 1955, 16, 143
173.
C.G. Overberger,
J. Polymer Sci., 1959, 35. 381
174.
A.C. Shelyakov,
Dokl. Akad. Nauk USSR, 1958, 122, 1076
175.
176.
V.A. Kargin, V.A. Kabanov
and II. Mardenko, Polym.
Sci.., USSR, 1960, 1, 41
177.
178.
179.
A. Grassi and A. Zambelli,
Am. Chem. Soc. Polymer
Preprints, 1996,
37 (2), 523
180.
181.
182.
183.
D.J. Worsford and S.
Bywater, Can. J. Chem., 1960, 38, 1881
184.
D.P. Wyman and T. Altares
Jr., Makromol. Chem., 1964 72. 68; T. Altares Jr., D.P. Wyman, and
V.R. Allen, J. Polymer Sci., 1965, A3, 4131
185.
F. Wenger and S.P.S. Yen,
Am. Chem. Soc. Polymer Preprints, 1962, 3, 163,
186.
187.
T. Altares Jr. and E.L.
Clark, Ind. Eng. Chem., Prod. Res. Dev .
1970, 9 (2),
168
188.
189.
190.
Y. Sakurada, J. Polymer Sci., 1963, A-1,1, 2407
191.
F.S. Dalton and R.H.
Tomlinson, J. Chem.
Soc., 1953, 151
192.
D.O. Jordan and A.R.
Mathieson, J. Chem.
Soc., 1952, 2354
193.
D.J. Worsford and S.
Bywater, J. Am.
Chem. Soc., 1957, 9, 491
194.
195.
G. Natta, SPE J., 1959, 15, 373
196.
J.A. Price, M.R. Lytton and
B.G. Ranby, J. Polymer
Sci. 1961 ,
51, 541
197.
R.F. Boyer In Vol.13 of
Encyclopedia of
Polymer Science
and Technology, (H.F. Mark,
N.G. Gaylord, and N.M.Bikales, eds.), Wiley-Interscience, New York,
N.Y., 1970
198.
199.
R.H. Boundy and R.F. Boyer
“Styrene, Its Polymers, Copolymers, and Derivatives”, Reinhold, New York,
1952
200.
L.J. Young, Chapt. 8 in
Copolymerization (G.E. Ham,
ed.), Interscience, New York, N.Y. 1964
201.
L.A. Wood, J. Polymer Sci., 1958 28, 319
202.
J.G. Fox and S. Gratch,
Ann, N.Y. Acad. Sci., 1953, 57, 367
203.
R.H. Wiley and G.M. Brauer,
J. Polymer Sci.,
1948, 3, 455, 647,; ibid, 1949, 4, 351
204.
J.G. Bevington, H.W.
Melville, and R.P. Taylor, J.
Polymer Sci.
1954 , 12, 449
205.
G.E. Molan, J. Polymer Sci., 1965, A-1,3, 126
206.
G.E. Molan, J. Polymer Sci., 1965 A-1,3, 4235
207.
U.D. Standt and J. Klein,
Makrom, Makromol.
Chem., Rapid Comm., 1981, 2, 41
208.
G.E, Schildknecht,
Vinyl and Related Polymers, Wiley, New York,
1952
209.
G.E. Barnes, R.M. Olofson,
and G.O, Jones, J. Am.
Chem. Soc., 1950, 72, 210
210.
H.W. Melville, Proc. Roy. Soc. (London), 1938, A
167, 99
211.
G.D. Dixon, J. Polymer Sci, Polymer Chem. Ed., 1974, 12, 1717
212.
213.
D.L. Glusker, E. Stiles,
and Yoncoskie, J. Polymer
Sci . 1961, 49, 197
214.
215.
W.E. Goode, F.H. Owens,
R.P. Fellman, W.H. Snuder, and J.E. Moore, J. Polymer Chem., 1960, 46, 317
216.
217.
F.J. Welch, U.S. Patent #
3,048,572 (Aug, 7, 1962)
218.
219.
S. Smith, J. Polymer Sci., 1959, 38, 259
220.
W.G. Gall and N.G. McCrum,
J. Polymer Sci., 1961, 50, 489
221.
M.R. Miller and G.E.
Rauhut, J. Am. Chem. Soc., 1958, 80, 4115
222.
M.R. Miller and G.E.
Rauhut, J. Polymer
Sci., 1959, 38, 63
223.
W.E. Goode, F.H. Owens, and
W.L. Myers, J. Polymer
Sci., 1960, 47, 75
224.
225.
W.E. Goode, R.P. Fellman,
and F.H. Owens in Vol.I of Macromolecular Syntheses (C.G.
Overberger, ed.), Wiley, New York, N.Y., 1963
226.
C.F. Ryan and J.J. Gormley,
in Vol.I of Macromolecular
Syntheses, (C.G. Overberger, ed.), Wiley, New York, N.Y.,
1963
227.
A. Yamamoto, T. Shimizu,
and I. Ikeda, Polymer J.,
1979, 1, 171 (Information was obtained from
A. Yamamoto and T. Yamamoto, Macromol. Rev .), 1978, 13, 161
228.
H. Abe, K. Imai, and M.
Matsumoto, J. Polymer
Sci., 1965, B3, 1053
229.
A. Yamamoto and T.
Yamamoto, Macromol.
Chem., 1978, 13, 161
230.
N. Yamazaki and T. Shu,
Kogyo Kagaku Zashi, 74, 2382 (1971) (from A. Yamamoto and T.
Yamamoto), Macromol.
Chem., 1978, 13, 161
231.
S.S. Dixit, A.B. Deshpande,
and S.L. Kapur, J. Polymer
Sci. 1970 ,
A-1,8, 1289
232.
A. Ravve and J.T. Khamis,
U.S. Patents # 3,306,883 (Feb 28, 1967); 3,323,946 (June 6,
1967)
233.
R.H. Yokum, Functional Monomers, Their Preparations, Polymerization and Application, Vol.I and II, Dekker, New
York, N.Y., 1973
234.
A. Ravve and J.T. Khamis,
J. Macromol. Sci., - Chem .
1967, A1(8),
1423
235.
A. Ravve and K.H. Brown,
J. Macromol. Sci., - Chem., 1979 A13(2), 285
236.
H. Yasuda, H. Yamamoto, K.
Yokota, S. Miyake, and A. Nakamura, J. Am. Chem. Soc. 1992 ,
114, 4908CrossRef
237.
H. Yasuda, H. Yamamoto, M.
Yamashita, K. Yokota, A. Nakamura, S. Miyake, Y. Kai, N. Kanehisa,
Macromolecules 1993, 26, 7134CrossRef
238.
239.
B.M. Novak and S. Boffa,
Am. Chem. Soc. Polymer
Preprints, 1997, 38
(1), 442
240.
A. Leblanc, E. Grau, J.-P.
Broyer, C. Boisson, R. Spitz, and V. Montei, Macromolecules, 2011 ,
44, 3293CrossRef
241.
A. Roig, J.E. Figueruelo,
and E. Liano, J. Polymer Sci., 1968, C(16), 4141
242.
D.P. Kelly, G,J.H. Melrose,
and D.H. Solomon, J. Appl.
Polymer Sci., 1963, 7, 1991
243.
C.H. Bamford and A.D.
Jenkins, Proc. Royal Soc., (London), Ser A, 1953, A216, 515
244.
W.M. Thomas and J.J.
Pellon, J. Polymer
Sci., 13, 329 (1954)
245.
A. Zilkha, B.A. Feit, and
M. Frankel, J. Chem.
Soc., 1959, 928
246.
M. Frankel, A. Ottolenghi,
M. Albeck, and A. Zilkha, J.
Chem. Soc.,
1959, 3858
247.
A. Ottolenghi and A.
Zilkha, J. Polymer
Sci., 1963, A-1,1, 647
248.
249.
A. Zilkha, P. Neta, and M.
Frankel, J. Chem.
Soc., 1960, 3357
250.
F.S. Dainton, E.M. Wiles,
and A.N. Wright, J. Polymer
Sci., 1960, 45, 111
251.
J.L. Down, J. Lewis, B.
Moore, and G. Wilkinson, Proc. Chem. Soc., 1957, 209
252.
R.B. Cundall, D.D. Eley,
and R. Worrall, J. Polymer
Sci., 58, 869 (1962)
253.
N.S. Wooding and W.C.S.
Higginson, J. Chem.
Soc., 1952, 774
254.
C.S.H. Chen, N. Colthup, W.
Deichert, and R.L. Webb, J.
Polymer Sci
. 1960,, 45, 247
255.
M.L. Miller, J. Polymer Sci., 1962, 56, 203
256.
W.E. Hanford and R.M.
Joyce, J. Am. Chem. Soc., 1946, 68, 2082
257.
R. Kiyama, J. Osugi, and S.
Kusuhara, Rev. Phys. Chem., Japan, 1957, 27, 22
258.
N. Shavit, A. Oplaka, and
M. Levy, J. Polymer
Sci., 1966, A-1,4, 2041
259.
N. Shavit, M. Konigsbuch,
and A. Oplaka, U.S. Patent #3,345,350
260.
N. Shavit and M.
Konigsbuch, J. Polymer
Sci., 1967, C(16), 43
261.
S. Amdur and N. Shavit,
J. Polymer Sci. 1967 ,
A-1,5, 1297
262.
263.
L.E. Ball and J.L. Greene
in Vol. 15 Encyclopedia of
Polymer Science
and Technology, (H.F. Mark,
N.G. Gaylord, and N.M. Bikales, eds.), Wiley-Interscience, New
York, N.Y. 1971
264.
265.
O.H. Bullitt, Jr., U.S.
Patent # 2,608,555 (1952)
266.
C.G. Overberger, E.M.
Pearce, and N. Mayers, J.
Polymer Sci.,
1948, 34, 109
267.
C.G. Overberger, H. Yuki,
and N. Urakawa, J. Polymer
Sci.,1960, 45, 127
268.
W.K. Wilkinson, U.S. Patent
# 3,087,919 (1963)
269.
Ben-Ami Feit, E. Heller,
and A. Zilkha, J. Polymer
Sci. 1966 ,
A-1,4, 1151
270.
W.M. Thomas and D.W. Wang,
Acrylamide Polymers in
Encyclopedia of
Polymer Science
and Engineering Vol.1, H.F.
Mark, N.M. Bikalis, C.G. Overberger, and G. Menges, eds., Wiley-
Interscience, New York, 1985
271.
K. Butler, P.R. Thomas, and
G.J. Tyler, J. Polymer
Sci. 1960, 48, 357
272.
J.F. Jones, J. Polymer Sci., 1958, 33, 15
273.
S. Boyer and A. Rondeau,
Bull. Soc. Chim. France, [5] 25
274.
L.A.R. Hall, W.J. Belanger,
W. Kirk, Jr., and Y.A. Sundatrom, J. Appl. Polymer Sci., 1959, 2, 246
275.
S.V. Gangal,
“Tetrafluoroethylene Polymers” in Encyclopedia of Polymer Science and Engineering, Vol. 16, (H.F Mark,
N.M. Bikalis, C.G. Overberger, and G. Menges, eds.),
Wiley-Interscience, New York, 1989
276.
R. Kiyama, J. Osugi, and S.
Kusuhara, Rev. Phys. Chem.,
Japan, 1957, 27,22
277.
278.
279.
E.T. McBee, H.M. Hill, and
G.B. Bachman, Ind.
Eng. Chem., 1949, 41, 70
280.
L.E. Robb, F.J. Honn, and
D.R. Wolf, Rubber Age,
1957, 82, 286
281.
J.S. Rugg and A.C.
Stevenson, Rubber Age,
1957, 82, 102
282.
283.
C.B.Griffins and J.C.
Montermoso, Rubber Age,
1959, 77, 559
284.
J.F. Smith, Rubber World, 1960, 142(3), 102
285.
J.F.Smith and G.T. Perkins,
J. Appl. Polymer Sci., 1961, 5(16), 460
286.
287.
288.
J. Ulbricht and R.
Sourisseau, Faserforsch.
Textiltech., 1955, 12, 547
289.
290.
D.I. Livingston, P.M.
Kamath, and R.S. Corley, J.
Polymer Sci.,
1956, 20, 485
291.
D. Sianeai and P.
Carradini, J. Polymer
Sci., 1959, 43, 531
292.
G. Talanini and E. Peggion,
Chapt. 5, Vinyl
Polymerization, (G. Ham, ed.), Dekker, New York, N.Y,
1967; J. Lewis, P.E.
Okieimen, and G.S. Park, J.
Macromol. Sci.-Chem., (1982, A17, 1021)
293.
W.I. Bengough and R.G.W.
Norrish, Proc. Roy. Soc. (London), 1950, A
200, 301
294.
A. Schindler and J.W.
Breitenbach, Ric.
Sci. Suppl., 1955, 25, 34
295.
E.J. Alerman and W.M.
Wagner, J. Polymer
Sci., 1951, 9, 541
296.
297.
298.
C.H. Bamford, W.G. Barb,
A.D. Jenkins, and P.F. Onyon, The
Kinetics of Vinyl
Polymerization by
Free-Radical Mechanism, Butterworth, London,
1954
299.
T. Hjertberg and E. Sorvik,
J. Polymer Sci., Polymer Letters, 1981, 19, 363
300.
A. Caraculacu, E.C.
Burniana, and G. Robila, J.
Polymer Sci,
Polymer Chem., Ed., 1978, 16, 2741
301.
G. Vancso, 0. Egyed, S.
Pekker, and A. Janossy, Polymer, 1982, 23, 14
302.
K.H. Reichert and H.U.
Moritz, J. Appi. Polymer Sci., 1981, 36, 151
303.
304.
J.W.L. Fordham, P.H.
Burleigh, and C.L. Sturm, J.
Polymer Sci.,
1959, 41, 73
305.
U. Giannini and S. Cesca,
Chim. Ind. (Milan), 1962, 44, 371
306.
V.G. Gason-Zade, V.V.
Mazurek, and V.P. Sklizkova, Vysokomol. Soedin., 1968, A10(3), 479
307.
A. Gyot and Pham-Qung-Tho,
J. Chim. Phys., 1966, 63, 742
308.
A. Gyot, D.L. Trung, and R.
Ribould, C.R. Acad.
Sci., (Paris), 266,
1968 , C 1139
309.
V. Jisova, M. Kolinsky, and
D. Lim, J. Polymer
Sci., 1970, A-1,8, 1525
310.
N. Yamazaki, K. Sasaki, and
S. Kambara, J. Polymer
Sci., Polymer
Letters, 1964,
2, 487
311.
A. Guyot and Pham Aung Tho,
J. Polymer Sci., 1964, C,4, 299
312.
L.F. Albright, Chem. Eng., 85 (July 3. 1967)
313.
314.
K.K. Kikukawa, S. Nozakura,
and S. Murahashi, Kobunshi
Kagaku 1964
, 25, 19 (from
Chem. Abstr .), 1968 69, 1968x
315.
K. Noro, Chapt. 6,
Polyvinyl Alcohol, C.A.
Finch, ed., Wiley, New York, N,Y., 1973
316.
J.A. Brydson Plastic Materials, Van Norstrand,
London, 1966
317.
318.
G. Natta and P. GarradIni,
J. Polymer Sd., 1956,
20,262
319.
L. E. N Allan, E. D. Cross
T. W. Francjs-Pranger, M. E. Hanhan, M. R. Jones, J. K, Pearson, M.
R. Perry, T. Storr, and M. P. Shaver, Macromolecules, 2011, 44 (11), 4072