9.1 Reactivity of Macromolecules
In consideration of various chemical reactions of
macromolecules, the reactivity of their functional groups must be
compared to those of small molecules. The comparisons have
stimulated many investigations and led to conclusions that
functional groups exhibit equal reactivity in both large and small
molecules, if the conditions are identical. These conclusions are
supported by theoretical evidence [1, 2].
Specifically, they apply to the following situations
[1]:
1.
Reactions that take place in homogeneous fluid
media with all reactants, intermediates, and end products fully
soluble. These conditions exist from the start to the end of the
reactions.
2.
All elementary steps involve only individual
functional groups. The other reacting species are small and
mobile.
3.
The steric factors in the low molecular weight
compounds selected for comparison must be similar to those of the
large molecules.
The above can be illustrated by a few examples.
For instance, the rates of photochemical cis–trans isomerization of azobenzene
residues on the backbones of flexible polymeric chains are
analogous to those of small molecules [3]. Another example is the activation energy for
cis–trans isomerization of azo-aromatic
polymers. It is the same for low molecular weight analogs
[4]. A third example is an
experiment in comparing conformational transitions of some eximers
in large and small molecules. A sandwich complex forms between an
excited aromatic chromophore, Ar, and a similar chromophore in the
ground state when irradiated with light of an appropriate
wavelength. The conformation required by such an excimer can
correspond to a prohibitive energy requirement for the unexcited
molecule. All conformational transitions must take place during the
lifetime of the excited state of the chromophore that is of the
type [5]:

The ratio of the fluorescence intensity of an
excimer and a normal molecule is a measure of the probability that
the conformational transition takes place during the excited
lifetime. A polyamide with only a small proportion of the following
units was used for comparison:

Emission spectra of dilute solutions of the above
polyamide and its low molecular weight analog were measured over a
range of temperatures. They showed that the activation energies of
the conformational transitions required for excimer formations are
essentially the same for both materials [5].
In addition, all bimolecular
activation-controlled reactions are independent of the degree of
polymerization [6]. Simple
SN2 reactions between reactive groups attached to chain
ends of monodisperse macromolecules in a wide range of molecular
weights are independent of the DP [7, 8] in the
range of 20–2,000 [7]. This was
shown on three different reactions. In the first one, the
reactivities of chlorine-terminated low and high molecular weight
polystyrenes towards polystyryllithium are equal in benzene and
cyclohexane solvents:

In the second one, the reactivity of primary
amine-terminated polyoxyethylenes with sulfonyl chloride-ended
polyoxyethylenes in chloroform is also the same:

In the third one, chain length dependence of the
propagation rates was measured in polymerizations of methyl
methacrylate. In the range of DPs from 130 to 14,200, the
propagation constant was shown to be independent of the chain
length [8].
On the other hand, unequal reactivity was
observed:
1.
In bimolecular reactions that are
diffusion-controlled.
2.
When neighboring group participations become
significant.
3.
When the properties of the polymers in solution
are altered by gelation.
4.
When the tacticities of the polymers affect
neighboring group interactions.
5.
When heterogeneous conditions affect
accessibility to the reactive sites.
There are special situations that can occur. For
instance, electrostatic charges carried by the polymers may extend
over long distances in solutions and may manifest themselves in
reactions with charged reagents. Sometimes, chain flexibility or
folding can cause functional groups to come together and interact,
though they may be located well apart on the polymer backbone.
Polymer solutions of this type are comparable to dispersions of
individual droplets of concentrated solutions.
Some statements above may require additional
clarification. An elaboration, therefore, follows.
9.1.1 Diffusion-Controlled Reactions
Reactions that are bimolecular can be affected by
the viscosity of the medium [9].
The translational motions of flexible polymeric chains are
accompanied by concomitant segmental rearrangements. Whether this
applies to a particular reaction, however, is hard to tell. For
instance, two dynamic processes affect reactions, like termination
rates, in chain-growth polymerizations. If the termination
processes are controlled by translational motion, the rates of the
reactions might be expected to vary with the translational
diffusion coefficients of the polymers. Termination reactions,
however, are not controlled by diffusions of entire molecules, but
only by segmental diffusions within the coiled chains
[10]. The reactive ends assume
positions where they are exposed to mutual interaction and are not
affected by the viscosity of the medium.
9.1.2 Paired Group and Neighboring Group Effects
When random, irreversible, and intramolecular reactions occur on
polymeric backbones with the functional groups adjacent to each
other, they can be expected to react. There is, however, an upper
limit to conversions. This upper limit is due to statistical
probability that some functional groups are bound to become
isolated. The limit for conversion was calculated to be 86.5%
[11].
Theoretically, quantitative conversions should be
possible with reversible
reactions of paired functional groups on macromolecular backbones.
The ability, however, of isolated reactive groups to find each
other and then pair off depends either upon particularly high
driving forces, or upon the time required to accomplish complete
conversions [12]. For reactions
initiated randomly, at different sites, the probability is high
that two groups on the terminal units will eventually meet and
react. Since the reactions are reversible, at least in theory, very
high conversions are possible.
Neighboring group participation can usually be
deduced from three types of evidence:
1.
If the reactions occur more rapidly during the
rate-determining step than can be expected from other
considerations.
2.
If the stereochemistry of the reactions suggests
neighboring group involvement.
3.
If molecular rearrangements occur and the groups
remain bonded to reaction centers, but break away from the atoms to
which they were originally attached on the substrates.
There are many examples in the literature that
describe neighboring group effects in reactions of polymers. One
example is hydrolysis of poly(p-nitrophenyl methacrylate-co-acrylic
acid). The high reaction rate at a neutral pH is due to attacks by
the carboxylic moieties upon the neighboring carbonyl carbons
[13–15]. Decomposition rates of t-butyl acrylate-styrene copolymers
[16] can serve as an another
example. Experimental data show pronounced acceleration for all
samples. This is interpreted in terms of both intra and
intermolecular interactions of the esters and the carboxylic
groups. It follows a suggestion of Cherkezyan and Litmanovich
[16, 17] that the instantaneous reactivity of any
group depends on its microenvironment. That includes (for reactions
of polymer in molten condition) two nearest units on the same chain
(internal neighbors) and two units belonging to two different
chains (external neighbors).
Another example of the neighboring group effect
is the behavior of polyacrylamides in hydrolyses. There are two
distinct and successive rates [18]. After conversions of up to 40–50% are
reached, the reactions slow down. This is due to accumulations of
negative electrostatic charges on the polymeric backbones
[18]. In alkaline media, the
increasing negative charges along the chains exert electrostatic
repulsions toward the hydroxyl ions. This results in rate
decreases.
9.1.3 Effect of Molecular Size
An example is the effect of DP on the rates of
alkaline hydrolyses of poly(vinyl acetate)s. Rapid increases in the
rates can be seen [19] in large,
but not in small molecules, as the reactions progress. Solvents
that are good for the products, like acetone–water mixtures, are
used in these reactions. These are, however, poor solvents for the
staring materials with high DP. Low molecular weight molecules are
more soluble. This means that, at the start of the reaction, the
large molecules are coiled up and the reactive sites not readily
available. As the reactions progress, the chains unravel and the
sites became more accessible with accompanying increases in the
reaction rates. Because the small molecules are more soluble, the
reactive sites are accessible from the start of the reactions, and
the rates are constant.
There are many report in the literature on the
effects of chain
conformation [19–25]. One
example is radical bromination of poly(methyl styrene)
[20] with N-bromosuccinimide-benzoyl peroxide or
Br2-K2CO3-light. 13C
NMR spectroscopy shows differences in reactivities of the methyl
groups in the 3 and 4 positions on the benzene rings between
isotactic and atactic polystyrenes [20].
The differences in reactivities in poly(vinyl
alcohol)s between isotactic (meso) and syndiotactic (dl-diol) portions of the polymers and
between cis and
trans acetals
[26–28] is another example. In extending this to
model compounds, reactions of stereo isomers of pentane-2,4-diol
and heptane-1,4,6-triol with formaldehyde take place much faster
for the meso than for the dl-diol portions [26–28]. Even
more important are the steric effects imposed by restricted
rotations. For instance, quaternizations of chloromethylated
polyether sulfones exhibit decreasing rates at high degrees of
substitution. This can be attributed to restricted rotations of the
polymeric chains, because this phenomenon is not observed with more
flexible chloromethylated polystyrene under identical conditions
[23, 24].
9.1.4 Effects of Changes in Solubility
Changes in solubility can occur during the
courses of various reactions. Such changes are observed, for
instance, during the chlorination of polyethylene in aromatic and
chlorinated solvents [29]. There
is an increase in the solubility until 30% conversion is reached.
After that, solubility decreases and reaches a minimum at 50–60%
chlorine content. Following that it increases again. This, however,
is not typical of many reactions of polymers in solutions. More
common is that the starting material is soluble, but not the
product or the opposite is true. Higher conversions are, usually,
expected when the polymers are solvated and the chains are fully
extended. In such situations, the reagents have ready access to the
reactive sites [29]. If the
products are insoluble in the reaction medium and tend to
precipitate as the reaction progresses, the sites become
increasingly less accessible. This can result in low conversions
and premature terminations. If the opposite is true and the product
is more soluble than the starting material, homogeneous limited
reactions can be controlled. When the starting material is
incompatible with the product, mutual precipitation or coiling of
the chains can take place. This can also result in limited
reactions. In addition, only minor differences in the constitutions
of two polymers can cause incompatibility. For instance, among
methacrylate polymers, there are incompatibilities in benzene
solutions that result from differences only in the amount of
branching of the alkyl groups [29].
Problems with solvent incompatibility can
sometimes be overcome by using mixtures of solvents. Those that are
good for the starting materials can be combined with those that are
good for the products. With careful experimentation, it may be
possible to develop a mixture of solvents that will keep all
components in solution [30]. In
some instances, however, insolubility of the products might be an
advantage. This is the case in alcoholysis reactions of poly(vinyl
acetate), where the polymer precipitates during the reactions and
in doing so absorbs the catalyst with it. The phenomenon permits
complete alcoholysis, particularly with the higher molecular weight
species that precipitate first.
Secondary reactions, like cross-linking and
gelation, can result in precipitations from solution. The extent of
the reactions, however, is not necessarily limited, because
diffusions of low molecular weight species are still possible.
Isolation of useful products, however, often becomes very
difficult.
9.1.5 Effects of Crystallinity
Crystallinity can only affect reactivity when the
reactions are carried out on polymers in the solid state and at
heterogeneous conditions. The differences in accessibility to the
reactive sites vary with the amount of crystallinity. Cellulose,
for instance, is often reacted in the solid state and the degree of
crystallinity is expressed in terms of reactivity to various
reagents [31]. The progress of a
reaction can sometimes be monitored by a loss of crystallinity.
What is more significant, however, is that greater accessibility to
amorphous regions results in reaction products with special
properties. An example is heterogeneous and homogeneous
chlorination of polyethylene. Two different products are obtained
[32]. The material from
heterogeneous chlorination is much less randomly substituted and
remains crystalline up to a chlorine content of 55%. The products
from the homogeneous reactions, on the other hand, are amorphous
after 35% substitution.
9.1.6 Reactions That Favor Large Molecules
Hydrophobic interactions play important roles in
many polymeric reactions. They are, for instance, significant in
the hydrolyses of low molecular weight esters when catalyzed by
polymeric sulfonic acid reagents, like poly(styrene sulfonic acid).
In these reactions, the hydrogen ions are located close to the
macromolecules [19]. The
hydrolytic cations are located in the regions of the macromolecules
and not in the bulk of the solution. The rates of the reactions are
high. Low molecular weight catalysts, on the other hand, like HCl,
have all the hydrogen ions distributed evenly throughout the
reaction medium. As a result, the rates are lower. Adsorption of
the ester groups to the polymeric sites is accompanied by an
increase in the apparent rate constant, as compared to reactions
with HCl. Examples are hydrolyses of methyl and butyl acetates
[19]. Another example is formation
of eximers and exiplexes in polyesters and methacrylate polymers
that always favor large molecules over small ones [33]. Proton transfer reaction of poly(vinyl
quinoline) [34] can serve as a
third example. The emission, excitation, and absorption spectra of
this polymer in a mixture of dioxane and water can be compared to
that of 2-methylquinoline. The emissions coming from the protonated
heterocyclic rings in the polymer occur sooner than from the low
molecular weight compound [34].
9.2 Addition Reactions
Polymers with double bonds in the backbones or in
the pendant groups can undergo numerous addition reactions. Some
are discussed in this section.
9.2.1 Halogenation
trans-1,4
and 1,2-Polybutadiene can be hydro halogenated under mild
conditions with gaseous HCl [51].
The same is true of copolymers of butadiene with piperylene and
also of isotactic trans-1,4-piperylene. The addition of
HCl to the asymmetric double bond is trans for polypiperylene and occurs in
a stereoselective way, judging from the 13C NMR
[51] spectra.
Polysilanes with alkene substituents add HCl and
HBr in the presence of Lewis acids [58]. The products are the corresponding chlorine
and bromine containing polymers with little degradation of the
polysilane backbone:

Chlorinations of rubber, however,
are fairly complex, because several reactions occur simultaneously.
These appear to be: (1) additions to the double bond; (2)
substitutions; (3) cyclizations; and (4) cross-linkings. As a
result, the additions of halogens to the double bonds are only a
minor portion of the overall reaction scheme [37, 38]. In
CCl4, the following steps are known to occur:

Halogenation reactions of unsaturated polymers
follow two simultaneous paths, ionic and free radical. Ionic
mechanisms give soluble products from chlorination reactions of
polybutadiene [42]. The
free-radical mechanisms, on the other hand, cause cross-linking,
isomerization, and addition products. If the free-radical reactions
are suppressed, soluble materials form. Natural rubber can be
chlorinated in benzene, however, with addition of as much as 30% by
weight of chlorine without cyclization [39, 40]. Also,
chlorination of polyalkenamers both cis and trans yields soluble polymers. X-rays
show that the products are partly crystalline [43, 44]. The
crystalline segments obtained from 1,4-trans polyisoprene are diisotactic
poly(erythro dichlorobutamer)s, while those obtained from the
1,4-cis isomer are
diisotactic poly(threo 1,2-dichlorobutamer)s [45].
Additive type chlorination of natural rubber can
also be carried out with phenyl iododichloride or with sulfuryl
chloride [39, 40]. Traces of peroxides must be present to
initiate the reactions. This suggests a free-radical mechanism.
Some cyclization accompanies this reaction as well [40]. In CCl4, for the first 25
chlorine atoms that add per each 100 isoprene units, 23 double
bonds disappear and only a small quantity of HCl forms. Subsequent
105 chlorinations, however, cause a loss of only 53 double
bonds.
Rubber can be brominated at 30°C. If traces of
alcohol are present, the reaction appears to go on entirely by
addition [39, 40]. Without alcohol, substitutions take place
rapidly and simultaneously with the additions to the double bonds
[41]. Exomethylene groups and
intramolecular cyclic structures form in the process. Slow
additions of bromine to vinylidene double bonds result in
formations of tri bromides,
–C5H7Br3. Also, cis and trans isomers of polyisoprene
[41] brominate differently.
Substitution reactions take place in brominations with N-bromosuccinimide. They are
accompanied by cyclizations [39].
Brominations of polybutadienes with N-bromosuccinimide yield α-brominated
polybutadienes [46, 47] that may also contain butane diylidene
units. The products act as typical alkyl halides and can undergo
Grignard-Wurz reactions:

The bromination reaction is accompanied by shifts
of the double bonds that are coupled with the sites of
substitution. Several different substituents can form. The polymers
may contain pentane diylidene, hexane diylidene, and heptane
diylidene units [46,
47].
By contrast, chlorination of polybutadiene in
benzene is a straightforward addition reaction of the halogens to
the double bonds [48,
49]:

Very little HCl is liberated until all the double
bonds are consumed. When CCl4 is used in place of
benzene, some substitutions occur during the latter stages of the
reactions. If cross-linking occurs at the same time, the
substitutions may not be extensive. The cross-linking reactions are
believed to involve carbocationic intermediates.
Polybutadiene can be halogenated readily in
tetrahydrofuran with iodine chloride or bromine [49]. The products are glassy polymers. These
products dehalogenate in reactions with organolithium compounds,
like n-butyllithium,
sec-butyllithium, and
polystyryllithium in tetrahydrofuran solution. Dehalogenation of
poly(iodo-chlorinated butadiene) with n-butyllithium yields product with
different cis/trans ratios. Also, this is accompanied
by partial cross-linking. The reactions may involve [49] halogen-metal exchanges that are followed by
intra- and intermolecular elimination of lithium halide. In
brominations of polybutadienes, both couplings and eliminations
take place. Both iodo-chlorinated and brominated polybutadienes
form graft copolymers when reacted with polystyryllithium in
tetrahydrofuran [50]. Gel
formation, however, accompanies the grafting reaction.
9.2.2 Hydrogenation
Atactic 1,4-polybutadiene and syndiotactic
1,2-polybutadiene can be hydrogenated at 100°C and 50 bar
pressure of hydrogen with a soluble catalyst
{[(Ph)3-P]3RhCl}. Complete saturation of
double bonds results [52].
Butadiene acrylonitrile copolymers can also be hydrogenated
quantitatively with this rhodium catalyst under mild conditions
[53]. The kinetics are consistent
with a mechanism where the active Rh-catalyst interacts with the
unsaturation at the polymer in the rate-determining step. The
nitrile group, however, appears to also interact with the catalyst
and inhibit the rates [53].
Hydrogenation of carbon-to-nitrogen double bonds
in polymer backbones and in the pendant groups can be carried out
with lithium borohydride [86]:

It was reported [485] that syndiotactic polystyrene can be
hydrogenated over Ni/SiO2 and Pd/BaSO4
catalysts. The Ni catalyst yields complete hydrogenenation when low
molecular weight polymer is used. Hydrogenation of high molecular
weight polystyrene, however, is incomplete. On the other hand, the
Pd catalyst yields completely hydrogenating material. The
hydrogenated syndiotactic polystyrene is a crystalline material
with good heat resistance [485].
9.2.3 Addition of Carbenes
Polyisoprenes and polybutadienes can also be
modified by reactions with carbenes. Dichlorocarbene adds to
natural rubber dissolved in chloroform in a phase transfer reaction
with aqueous NaOH [54]. A phase
transfer reagent must be used with the aqueous NaOH. Solid sodium
hydroxide can be used without a phase transfer reagent. There is no
evidence of cis–trans
isomerization and the distribution of the substituents is random
[54].
Difluorocarbene, generated under mild neutral
conditions, adds to 1,4-cis- and 1,4-trans-polybutadienes to give materials
containing cyclopropane groups [55]. The addition takes place randomly, to give
atactic stereo sequence distributions [55]:

Fluorocarbene, formed from
phenyl(fluoro,dichloromethyl)mercury by thermolysis in situ, also
adds to 1,4 cis- and
trans-polybutadienes. The
carbene can add at various levels [57]. The addition is stereospecific and
preserves the alkene geometry of the parent polybutadiene. Also,
the addition is random, showing that the reactivity of the double
bonds is independent of the sequence environment [57].
Dichlorocarbene, generated in situ from an
organomercury precursor, phenyl(bromo,dichloromethyl)mercury
(Seyferth reagent), adds to polybutadiene in a similar manner
[56]. The reactions take place
under homogeneous conditions. They can be carried out on 1%
solutions of the polymer in benzene, using 10–20% mol excess of the
reagent.
9.2.4 Electrophilic Additions of Aldehydes
These are additions to double bonds, like the
Prins reaction, and they
can be carried out on natural and synthetic rubbers [59, 60]. They
take place rapidly in the presence of acid catalysts. Aqueous
formaldehyde [61], or paraform in
CCl4 [62], can be used.
The catalysts are inorganic acids or anhydrous Lewis acids, like
boron trifluoride in acetic acid solution [63]:

The reaction takes a different path in the
absence of a catalyst [62]:

The products of the Prins reaction with rubbers
are thermoplastic polymers that possess fair resistance to acids
and bases. Free hydroxyl groups in the products are available for
cross-linking with diisocyanates [64] or by other means. The Prins reaction can be
carried out directly on rubber latexes [65]. It is also possible to just mill the rubber
together with formaldehyde and then heat the resultant mixture in
the presence of anhydrous metal chlorides [64] to get similar results [66].
Higher aldehydes also react with natural rubber
[67]. The reaction works best with
purified rubber. Additions take place without a catalyst at 180°C
or in the presence of AlCl3–NaCl at 120°C. These
reactions can be carried out in the solid phase by milling the
rubber with an aldehyde, like glyoxal [68]. Heating in a pressure vessel at above 175°C
is required to complete the reaction. Infra-red spectra of the
products from reactions in solution show presence of ether,
carbonyl, and hydroxyl groups [69]. Two types of additions appear to take place
[69]:

Products from reactions of rubber with glyoxal
have a strong tendency to become spontaneously insoluble. This is
probably due to a presence of residual aldehyde groups, because a
treatment of the product with 2,4-dinitrophenyl-hydrazine
eliminates spontaneous gelation.
Chloral adds to polyisoprene similarly. The
reaction is catalyzed by Lewis acids [70]. Both AlCl3 and BF3
are efficient catalysts. Less cross-linking is encountered with
aluminum chloride. Infra-red spectra of the products shows presence
of hydroxyl groups, chlorine atoms, and vinylidene unsaturation
[70].
9.2.5 Polar Additions
A number of polar additions to unsaturated
polymers are known. These include Michael addition, hydroboration,
1,3-dipolar additions, ene reaction, the Ritter reaction,
Diels–Alder additions, and others.
9.2.5.1 Michael Addition
Among polar additions to unsaturated polymers are
reactions of amines and ammonia with unsaturated polyesters in a
form of a Michael condensation. Thus, for instance, additions to
poly(1,6-hexanediol maleate) and poly(1,6-hexanediol fumarate)
[71] show a difference in the
reactivity of the two isomers. The maleate polyester reacts with
ammonia to yield a cross-linked product at room temperature, when
stoichiometric quantities or excess ammonia in alcohol is used. At
the same reaction conditions, the fumarate isomer only adds a few
percent of ammonia [71]. In a 1:1
mixture of chloroform and ethanol, however, approximately half of
the fumarate double bonds react. Also, the maleate polyester reacts
differently with piperidine or cyclohexylamine. In butyl alcohol at
60°C, the polymer initially isomerizes and precipitates. After the
isomerization is complete and the temperature is raised to 80°C,
the polymer redissolves. An exothermic reaction follows and Michael
type adducts form [71].
9.2.5.2 Hydroboration
Polymers and copolymers of butadiene or isoprene
with styrene can react with diborane [72]. A suitable solvent for this reaction is
tetrahydrofuran. Subsequent hydrolyses result in introductions of
hydroxyl groups into the polymer backbones. The reactions with
diborane are very rapid. Some side reactions, however, also occur
[72].
9.2.5.3 ,3-Dipolar Additions
Cyclic structures form on polymer backbones
through 1,3-dipolar additions to carbon to carbon or carbon to
nitrogen double bonds [73].
Because many 1,3-dipoles are heteroatoms, such additions can lead
to formations of five-membered heterocyclic rings. An example is
addition of nitrilimine to an unsaturated polyesters
[73]:

Also, iodine isocyanate adds to polyisoprene. The
product can be converted to methyliodocarbamate or to iodourea
derivatives [74]:

Iodine isocyanate additions result in
approximately 40% yields. The products can undergo typical
reactions of the isocyanate group [74], as for instance:
as well as:
where HX is a halogen acid. The products exhibit enhanced heat
stability [74].


Dipolar cycloadditions take place when nitrones
or nitrile oxides add to butadiene rubber. Some of the products
contain isoxazolidine rings [75]:


The above modification of butadiene rubber can be
carried out to the extent of 3.1 mol.%. The product is higher in
tensile modulus values and is greater in strength than the parent
compound [75].
A final example is a 1,3 dipolar addition to
pendant azide groups [87]. The
reaction takes place with phenyl vinyl sulfoxide in
dimethylformamide. Forty-eight hours at 110°C are required for the
azide groups to become undetectable by infra-red spectroscopy. The
product precipitates out with addition of ether [87]:

9.2.5.4 The Ene Reaction
The polymers of conjugated dienes can also be
modified via the ene reactions [76], as for instance:
where, X=Y can be O=N–, –N=N–, >C=S, >C=O, or>C=C<. An
example of this is an addition of triazolidones [76]:


This results in formation of pendant urazole
groups. The exact structure of the products, however, has not been
fully established. The tensile strength of polymers improves
considerably, but it is accompanied by a dramatic loss in molecular
weight [76]. Nevertheless, ene
reagents like C-nitroso and activated azo compounds are very
efficient in adding to rubber. They add in a few minutes at
temperatures between 100 and 140°C. In the case of the azo
compound, the addition can be greater than 90%.
Substituted aryl sulfonyl azides decompose at
elevated temperatures to nitrenes and add to natural rubber:

9.2.5.5 The Ritter Reaction
This reaction can be carried out on natural
rubber and on synthetic polyisoprenes [78]:

The carbon cation, apparently, reacts with any
nucleophile present. When the reaction is carried out in
dichloroacetic acid, chlorine atoms can be detected in the product
[78].
9.2.5.6 Diels–Alder Condensations
Crotonic acid esters of cellulose undergo
addition reaction with cycloaliphatic amines, like morpholine or
piperidine and with aliphatic primary amines [79]. Unsaturated polymers can also undergo
Diels–Alder reactions. One example is a reaction of
hexachlorocyclopentadiene with polycyclopentadiene [80]:

The addition takes place in an inert atmosphere
at 140–150°C. Over 90 mol.% conversion is achieved in
6 h.
Diels–Alder condensations of fumaric and maleic
acids polyesters with various dienes [81] can serve as another example. These
reactions require 20 h at room temperature. Diels–Alder
condensations can also be carried out on polymers of 1,3,5
hexatriene, 1,3,5-heptatriene, and 2,4,6-octatriene [82]. Sulfonate-substituted maleic anhydride adds
to low functionality hydrocarbon elastomers, like EPDM, presumably
via an Alder-ene type reaction [83]:

Thiols add to diene rubbers by a free-radical
mechanism [84]. Thus,
antioxidants, like 4-(mercaptoacetamido)-diphenylamine, add –SH
groups to the double bonds of cis-polyisoprene and polybutadiene in
the presence of free-radical initiators [84].
Thiol compounds also add photochemically to
polymers containing double bonds. For instance, unsaturation can be
introduced into polyepichlorohydrin by a partial elimination
reaction. The product then reacts with mercaptans, aided by a
photosensitizer (like benzophenone) and ultraviolet light
[85]:

9.2.5.7 Epoxidation Reactions
These addition reactions of unsaturated polymers,
like liquid polybutadiene, developed into preparations of useful
commercial materials [88–94]. Patent
literature describes procedures that use hydrogen peroxide in the
presence of organic acids or their heavy metal salts. Reaction
conditions place a limitation on the molecular weights of the
polymers, because it is easier to handle lower viscosity solutions.
A modification of the procedures is to use peracetic acid in place
of hydrogen peroxide [95–97]. The
most efficient methods rely upon formations of organic peracids in
situ with cationic exchange resins acting as catalysts
[98]. This can be illustrated as
follows:

The reaction is accompanied by formations of
by-products:

Polybutadienes that are high in 1,4-structures
tend to epoxidize more readily and yield less viscous products
[100, 101]. The epoxidation reaction can also be
carried out on poly(1,4-cyclopentadiene) [99]:

Perbenzoic acid is an effective reagent in
chloroform and in methylene chloride solutions at 0–20°C. The
conversions are high, yielding brittle materials soluble in many
solvents. The products can be cast into transparent films
[99].
Monoperphthalic and p-nitroperbenzoic acids are also
efficient epoxidizing agents. They can, however, cause
cross-linking, as is the case in epoxidation of polycyclopentadiene
[99]. The products react like
typical epoxy compounds [99]:

Some other reactions of the epoxy groups are
[99]:

9.3 Rearrangement Reactions
There are different types of polymeric
rearrangements. One of them is isomerizations of polymers with
double bonds. Others can be intramolecular cyclizations in the
polymeric backbone.
9.3.1 Isomerization Reactions
The isomerization of cis-polybutadiene can be carried out
with the aid of ultraviolet light or with gamma radiation. When
light is used, the free-radical reaction requires photoinitiators
[102, 103]. The mechanism involves freely rotating
transitory free-radicals on the polymer backbone. These form from
additions of photoinitiator fragments to the double bonds. The
adducts break up again, releasing the attached initiator fragments
and reestablishing the double bonds. The new configurations are
trans because they are more
thermodynamically stable [102,
103]. This can be shown as
follows:
where, X• represents a free-radical fragment from a
photoinitiator.

With gamma radiation, there is no need for any
additives [104, 105]. Here, the mechanism of isomerization is
believed to involve direct excitation of the π-electrons of the
double bonds to anti-bonding orbitals where free and geometric
interconversions are possible. In benzene solutions, energy
transfers take place from excited benzene molecules to the polymer
double bonds.
When 1,4-polyisoprene films are irradiated with
light [122], cis–trans isomerizations occur. In the
process, the quantities of 1,4-unsaturations decrease. Also, vinyl
and vinylidene double bonds and cyclopropyl groups form. This can
be illustrated as follows:
and




9.3.2 Cyclizations and Intramolecular Rearrangements
Cyclization reactions of natural rubber and other
polymers from conjugated dienes have been known for a long time.
The reactions occur in the presence of Lewis and strong protonic
acids. They result in loss of elastomeric properties and some
unsaturation. Carbon cations form in the intermediate step and
subsequent formation of polycyclic structures [108, 109]:


In a similar manner, polymers with pendant
unsaturation undergo cyclization reactions in benzene in the
presence of BF3 or POCl3, yielding ladder
structures. The exact nature of the initiation process is not
clear. Water may be needed for the initiation step [110, 111]:

The reactions result in formations of
six-membered monocyclic and fused polycyclic units. These reactions
of carbon cations should also lead to molecular rearrangements,
like 1,2 shifts of protons, resulting in formations of
five-membered rings and spiro cyclopentane repeat units
[111].
Cyclization reactions of polyisoprene can be
catalyzed by TiCl4 [112] and by sulfuric acid [113–115,
117]. The products appear similar
in the infra-red spectra [110]
with only a few minor differences. Also, there are only a small
number of fused rings in the product.
Polyacrylonitrile converts to a red solid when
heated above 200°C [116]. Only a
small amount of volatile material is given off. Further heating of
the red residue to about 350°C or higher converts it to a black
brittle material. This black material has a ladder structure:

Heating of polyacrylonitrile in the presence of
oxygen yields some quinone-type structures:

Further heating of the polymer, at very high
temperatures, in excess of 2,000°C, results in formation of
graphite-like structures [116].
All, or almost all nitrogen is lost, probably as HCN and
N2.
Migration of double bonds is a well-established
phenomenon in polymers from conjugated dienes. It occurs, for
instance, during vulcanization of rubber (discussed in a later
section). It also occurs upon simply heating some polymers
[117–119]. Thus, the double bonds shift in natural
rubber when it is heated to temperatures of 150°C or above
[120]:

Hydro chlorination of rubber in solution causes a
different kind of double bond shirt [121]:

Many other polymeric structures can rearrange
under proper conditions. For instance, poly(4,4′-diphenylpropane
isophthalate) rearranges upon irradiation with UV light to a
structure containing o-hydroxybenzophenones [126]:

The mechanism is believed to be that of a
photo-induced Fries
rearrangement. It may be similar to one catalyzed by Lewis
acids. Another example is a spontaneous rearrangement in the solid
state of poly(α-phenylethylisonitrile) [128]:

Based on studies of molecular models, it was
concluded [128] that the
substituents force progressive, one-handed twisting of the helix
shape of the molecules. This occurs to such an extent that the
vicinal imino double bonds are out of conjunction with each other.
As a result, each benzylic hydrogen atom becomes localized over the
electrons of the imino groups. It proceeds in a constant screw
direction around the axis of the helix. Such steric confinement
causes tautomeric rearrangements, as shown above, which results in
some relaxation of the compression. An additional driving force is
a gain in conjugation between the imino double bonds and the
aromatic rings of each substituent.
Intramolecular rearrangements of polymers with
ketone groups were subjects of several studies. The Schmidt and Beckmann rearrangements were carried
out on copolymers of ethylene and carbon monoxide [129]:

The starting materials were low and medium
molecular weight copolymers. Infra-red spectra of the products from
the Schmidt reaction show a high degree of conversion
[129]. The yields of the oximes
and subsequent Beckmann rearrangements are also high.
A Schmidt reaction on poly(p-acetyl styrene) yields a product
containing acetyl amino styrene groups in high yields and the
products are surprisingly pure [130]. The solvent is acetic acid and the
converted material precipitates out as a sticky mass:

Another rearrangement reaction is isomerization
of unsaturated polyesters upon heating. Thus, for instance, maleate
polyesters rearrange to the fumarate analogs at elevated
temperatures [131]:

It was reported that polymers with cis-stilbene moieties in the backbone
can undergo photocylization that is almost quantitative
[473]:

9.4 Substitution Reactions
Many substitution reactions are carried out on
polymers in order to replace atoms in the backbones or in the
pendant groups with other atoms or groups of atoms. These reactions
do not differ much from those of the small molecules.
9.4.1 Substitution Reactions of Saturated Polymeric Hydrocarbons
It is often desirable to replace hydrogens with
halogens. Fluorination of
polyethylene can be carried out in the dark by simply
exposing the polymer, either in sheet or in powder form, to
fluorine gas. The reaction is exothermic and it is best to dilute
the gas with nitrogen 9:1 to allow a gradual introduction of the
fluorine and avoid destruction of the polymer [132, 133]. In
this manner, however, only the surface layers are fluorinated and
the substitutions occur only a few molecular layers deep.
In surface fluorination with vacuum ultraviolet
glow discharge plasma, the photon component of the plasma enhances
the reactivity of the polymer surfaces toward fluorine gas
[134]:


The surface free-radicals can also cause
elimination of hydrogen radicals and formation of double bonds.
Whether as free-radicals or through unsaturation, the units are now
more reactive toward fluorine.
A film that is 3 mil thick can be completely
fluorinated on a 100-mesh phosphor bronze gauze, if the reaction is
allowed to proceed for several days [135]. Fluorination can also be carried out with
mercuric or cupric fluorides in hydrofluoric acid. The reaction
must be carried out at 110°C for 50 h. As much as 20% of
fluorine can be introduced [136].
In direct fluorination of powdered high-density
polyethylene with the gas, diluted with helium or nitrogen, the
accompanying exotherm causes partial fusion. In addition, there is
some destruction of the crystalline regions [137]. On the other hand, fluorination of single
crystals of polyethylene can result in fluorine atoms being placed
on the carbon skeleton without disruption of the crystal structure.
The extent of cross-linking, however, is hard to assess
[138]. The reaction has all the
characteristics of free-radical mechanism [139]:

Chlorinations of
polyethylene can be carried out in the dark or in the
presence of light. The two reactions, however, are different,
though both take place by free-radical mechanism. When carried out
in the dark at 100°C or higher, no catalyst is needed, probably
because there are residual peroxides from oxidation of the starting
material. Oxygen must be excluded because it inhibits the reaction
and degrades the product [140].
The reaction is catalyzed by traces of TiCl4
[141]. Such trace quantities can
even be residual titanium halide from a Ziegler-Natta catalyst left
over in the polymer from the polymerization reaction. When it is
carried out at 50°C in chlorobenzene, –CHCl– groups form
[142]:

This slows the chlorination of adjacent
groups.
Trace amounts of oxygen catalyze chlorinations in
the presence of visible light [140]. The same reaction in ultraviolet light is
accompanied by cross-linking. The photochemical process can be
illustrated as follows:
etc.

Chlorination of polyethylene can result in
varying amounts of hydrogen atoms being replaced by chlorine. It is
possible to form a product that contains 70% by weight of chlorine.
The amount of chlorination affects the properties of the product.
At low levels of substitution, the material still resembles the
parent compound. When, however, the level of chlorine reaches
30–40%, the material becomes an elastomer. At levels exceeding 40%,
the polymer stiffens again and becomes hard.
Commercial chlorinations of polyethylene are
usually conducted on high-density (D > 0.96) linear polymers. The
molecular weights of the starting materials vary. High molecular
weight polymers form tough elastomers. Low molecular weight
materials, however, allow easier processing of the products. The
reactions are carried out in carbon tetrachloride, methylene
dichloride, or chloroform at reflux temperatures of the solvents
and at pressures above atmospheric to overcome poor solubility. The
solubility improves with the degree of chlorination. Industrially,
two different procedures are used. In the first one, the reactions
are conducted at 95–130°C. When the chlorinations reach a level of
about 15%, the polymers become soluble and the temperatures are
lowered considerably [143]. In
the second one, the reactions are conducted on polymers suspended
in the solvent. When the chlorine content reaches 40% and the
polymers become soluble, chlorinations are continued in solution.
By continuing the reaction, a chlorine content of 60% can be
reached. The products from the two processes differ. The first one
yields a homogeneous product with the chlorine atoms distributed
uniformly throughout the molecules. Chlorination in suspension, on
the other hand, yields heterogeneous materials with only segments
of the polymeric molecules chlorinated. Some commercial
chlorinations are conducted in water suspensions. These reactions
are carried out at 65°C until approximately 40% levels of chlorine
are achieved. The temperatures are then raised to 75°C to drive the
conversions further. In such procedures, agglomerations of the
particles can be a problem. To overcome that, water is usually
saturated with HCl or CaCl2 [144]. Problems with agglomeration are also
encountered during suspension chlorinations in solvents, like
CCl4. Infra-red spectra of chlorinated polyethylenes
show presence of various forms of substitutions. There are
–CHCl–CHCl– as well as –CCl2– groups present in the
materials [145, 146]. Surface photo chlorination of polyolefin
films [146] considerably improves
the barrier properties of the films to permeations of gases.
Chlorinations of
polypropylene usually result in severe degradations of the
polymer. When TiCl4 is the chlorination catalyst,
presumably, less degradation occurs [140]. Studies of bromination of polypropylene (atactic)
show that when the reaction is carried out in the dark, in
CCl4 at 60°C, the substitution reactions proceed at the
rate of 0.5%/h [147].
Chlorosulfonation or polyethylene is a
commercial process. The reaction resembles chlorination in the step
of hydrogen abstraction by chlorine radicals. It is catalyzed by
pyridine [148–150] and can be pictured as follows
[147]:

The amount of SO2 vs. chlorine in the
reaction mixture affects the resultant ratios of chlorosulfonation
vs. chlorination of the polymer. These ratios and the amounts of
conversion vary with the temperature [149].
Commercially produced chlorosulfonated
polyethylene contains approximately 26–29% chlorine and 1.2–1.7%
sulfur [151]. The material is an
elastomer that remains flexible below −50°C. It is commonly
cross-linked (vulcanized) through the sulfonyl chloride groups.
Heating it, either as a solid or in solution, to 150°C results in
loss of SO2 and HCl. If the material is heated for
2 h at 175°C, all SO2Cl groups are removed. In
gamma radiation-induced chlorosulfonations and sulfoxidations of
polyethylene powders [153] at
room temperature, the ratios of SO2Cl groups to Cl
groups decrease with increases in radiation.
9.4.2 Substitution Reactions of Halogen-Bearing Polymers
Procedures for commercial chlorinations of poly(vinyl chloride)
vary. Low temperature chlorinations are done on aqueous dispersions
of the polymers that are reacted with chlorine gas in the presence
of swelling agents, like chloroform. These are light catalyzed
reactions, usually carried out at about 50°C. They result in
substitutions of methylene hydrogens [158, 159].

Some breakdown of the polymers accompanies the
reactions in suspension [154].
Irregularity in the structures and the release of HCl significantly
contribute to this. It appears that the degradations are initiated
by HCl that is released as a result of the chlorination. In the
process, double bonds form. Immediately upon their formation, they
become saturated with chlorines.
A study was conducted on solution chlorination of
poly(vinyl chloride) in the presence of free radicals
[155] generated from
azobisisobutyronitrile. The reaction appears to proceed in two
stages. The first one takes place until 60% of all the
CH2 groups have reacted. After that, in the second stage
the original –CHCl groups are attacked. Some unreacted
–CH2– groups, however, remain [155]. Photo chlorination can achieve 90%
conversion [156, 157]. Typical commercially chlorinated
poly(vinyl chloride), however, has a chlorine content of 66–67%.
The material contains methylene groups and is in effect a copolymer
of vinyl chloride and 1,2-dichloroethylene.
Chlorination raises the softening temperature of
poly(vinyl chloride). The products exhibit poorer heat stability
than the parent material and are higher in melt viscosity.
High-temperature chlorinations in chlorinated solvents at 100°C
result in extensive substitutions of the methylenic hydrogens as
well. The reactions, however, are accompanied by extensive chain
scissions. The products are soluble in solvents like acetone and
methylene chloride and have low softening points, low impact
strength, and poor color stability.
Chlorination of
poly(vinylidene chloride) can be carried out with sulfuryl
chloride using azobisisobutyronitrile [155]. The reaction appears to proceed in three
steps:
1.
Formation of SO2Cl• free radicals by abstraction.
2.
Formation of free radicals on the polymeric
backbones:

3.
Transfer reactions of the active sites:

Chlorination of
poly(vinyl fluoride) yields a product with 40–50% chlorine
content [158]. Fluorination of poly(vinylidene
fluoride) was reported [160].
When mixtures of fluorine and nitrogen gases are used, the
reactions are limited by the amount of diffusion of fluorine into
the polymer network. X-ray photoelectron spectroscopy shows
presence of –CF2–, –CHF–, and –CH2– groups in
the product [161].
Many attempts at other modifications of
poly(vinyl chloride) were reported in the literature. Often the
reactions are based on expectations that the polymers will react
like typical alkyl halides. Unfortunately, in place of nucleophilic
substitutions, the polymers often undergo rapid and sequential
eliminations of HCl along the chains. Nevertheless, many
substitution reactions are still possible and can be successfully
carried out. One example is a replacement of 43% of the chlorine
atoms with azide groups [162]:

The reaction can be carried out at 60°C for
10 h and the polymer does not cross-link. It proceeds faster
in DMF than in other solvents. Once substituted, the polymer
becomes photosensitive and cross-links when irradiated with
ultraviolet light. The presence of azide groups allows further
modifications. One such modification is a reaction with an
isocyanate [162]:

By a similar reaction, phosphinimine groups can
be formed on the polymer backbone [162]:

Note: The
illustration of the reactions with LiAlH4, as shown
above, implies that all chloride atoms remain intact on the polymer
backbones. It appears likely, however, that some of them might get
removed and double bonds might form instead in the backbone.
Poly(vinyl chloride) reacts with various thiols
in ethylene diamine to produce monosulfide derivatives
[165]:

When sulfur is added to the reaction mixture,
disulfides form instead [165]:

The thiolate ions can serve as strong
nucleophiles and also as weak bases [166]. When poly(vinyl chloride) is suspended in
water in the presence of swelling agents and phase transfer
catalysts, it reacts with mercaptans, like 3-[N-(2-pyridyl)carbamoyl]-propylthiol:

Acetoxylation of poly(vinyl chloride) can be
carried out under homogeneous conditions [167]. Crown ethers, like 18-crown-6, solubilize
potassium acetate in mixtures of benzene, tetrahydrofuran, and
methyl alcohol to generate unsolvated, strongly nucleophilic
“naked” acetate anions. These react readily with the polymer under
mild conditions [167].
Substitutions of the chlorine atoms on the polymeric backbones by
anionic species take place by SN2 mechanism. The
reactions can also proceed by SN1 mechanism. That,
however, requires formations of cationic centers on the backbones
in the rate-determining step and substitutions are in competition
with elimination reactions. It is conceivable that anionic species
may (depending upon basicity) also facilitate elimination reactions
without undergoing substitutions [167].
Reactions of poly(vinyl chloride) with aromatic
amines, amino alcohols, or aliphatic amines in DMF solution result
in both substitutions and in eliminations [168]. Reactions with aniline yield the
following structure [168]:

Carbanionic reaction sequences can be used to
introduce various photosensitive groups [169]. The process consists of creating
carbanionic centers on the backbone and then reacting them with
various halogenated derivatives:





A redox cyclopentadienyl iron moiety can also be
introduced into the poly(vinyl chloride) backbone by a similar
technique [170]. Many other
attempts were reported at replacing the halogens of poly(vinyl
chloride), poly(vinyl bromide), and poly(vinyl iodide) with an
alkali metal or with a hydrogen. For instance, in an effort to form
poly(vinyl lithium), the polymers were reacted with organolithium
compounds and with metallic lithium. The reactions with
alkyllithium, however, resulted in substitutions by the alkyl
groups, similarly to the reactions shown previously [171]:

Reactions with metallic lithium lead to
formations of polyenes [171]. On
the other hand, when poly(vinyl chloride) is reacted with metal
hydrides, like lithium aluminum hydride in a mixture of
tetrahydrofuran and decalin at 100°C, macroalkanes form
[172]:

Replacement of the chlorine with N,N-dialkyl dithiocarbamate was reported
to occur at 50–60°C in DMF solvent [173]:

The reaction is catalyzed with ethylene diamine
[174].
9.4.3 Substitution Reactions of Polymers with Aromatic Rings
There are some interesting reports in the
literature on reactions carried out on the backbones of
polystyrenes. There are also many reports in the literature on
aromatic substitution reactions of polystyrene. Only a few,
however, are in industrial practice.
9.4.3.1 Reactions of Polystyrene
Photo
chlorination of polystyrene involves replacement of
hydrogens at the α and β positions [175, 176]. It
was believed in the past that the chlorine atoms react
preferentially at the α-position until the chlorine content of the
product reaches 20% by weight. After that, it was thought that the
chlorines are introduced into the other position. Later, however,
this was contradicted [177]. In
fact, when polystyrene is photo-chlorinated in carbon tetrachloride
at low temperatures, like 13°C, it is substituted equally at both
positions. At higher temperatures, like 78°C, substitutions at the
β-position actually predominate [177].
Chlorinations of poly(p-methyl styrene) are somewhat more
selective for the pendant methyl groups and result in di- and
tri-substitutions at the p-methyl position. Only small amounts
of chlorine are introduced into the polymer backbones
[178]. Substitutions at the
backbones, however, are possible with the use of
SO2Cl2 as the chlorinating agent. In this
case, half of the chlorines still replace the methyl hydrogens, but
the other half replace hydrogens on the backbone.
Free radical additions of maleic anhydride to
polystyrene backbones can be carried out with the help of either
peroxides or ultraviolet light [164]. Approximately 2% of the anhydride can be
introduced. If, however, the additions are carried out on
α-brominated polystyrene, the anhydride content of the polymer can
be raised to 15% [164]:

Note: the
extra hydrogen shown above on the maleic anhydride moiety of the
product presumably comes from chain transferring.
Bromination of polystyrene with
N-bromosuccinimide and
benzoyl peroxide in CCl4 at room temperature can achieve
a 61% conversion in 4 h. Considerable degradation, however,
accompanies this reaction [180].
9.4.3.2 Chloromethylation Reactions
Chloromethylation reactions of the aromatic
rings of polystyrene and styrene copolymers are being
carried out extensively. Chlorodimethyl ether (a carcinogen) is a
good solvent for these polymers. It is, therefore, commonly
employed as the reagent [181,
182]. Laboratory preparations can
be carried out in mixtures of carbon disulfide and ether, using
zinc chloride as the catalyst. A 9-h reaction at room temperature
yields 10% substitution [183].
The chloromethylation process [184] occurs in two steps. Benzyl methyl ether
forms as an intermediate. Cross-linking reaction between the
aromatic nuclei and the formed CH2Cl group occurs as
side reaction. There are strong indications that the
chloromethylation takes place only at one position on the ring
[185]. The same is true of
bromomethylation [185].
Stannic chloride is a very effective catalyst for
this Friedel-Craft reaction [186]. Iodomethylation can also be carried out
in the same manner with similar results [179]. When the reactions are carried out on
cross-linked styrene copolymers with chlorodimethyl ether and
stannic chloride catalyst, they are accompanied by strong
morphological changes [187]. If
these reactions are carried out with low levels of
chloromethylating agents or catalysts, they occur more or less
homogeneously. Larger levels of either of them, however, result not
only in greater levels of chloromethylation, but also in higher
degrees of secondary cross-linkings and in uneven distributions of
the chloromethyl groups [188].
Another technique of chloromethylating
polystyrene is to react it with methylal and thionyl chloride in
the presence of zinc chloride [189]:

Chloromethyl-substituted polystyrenes can also be
prepared from poly(p-methyl
styrene)s by treating them with aqueous sodium hypochlorite in the
presence of a phase transfer catalysts, like benzyltriethylammonium
chloride [190]. The conversions
of methyl to chloromethyl groups can be as high as 20% without any
detectable morphological changes [187]. If these reactions are carried out with
low levels of chloromethylating agents or catalysts, they occur
more or less homogeneously. Larger levels of either of them,
however, result not only in greater levels of chloromethylation,
but also in higher degrees of secondary cross-linkings and in
uneven distributions of the chloromethyl groups [188].
Another technique of chloromethylating
polystyrene is to react it with methylal and thionyl chloride in
the presence of zinc chloride [189].
Techniques for chloromethylating polyarylether
sulfones, polyphenylene oxide, phenolic resins, and model compounds
were described recently [191].
When the subsequent products are converted to quaternary amines,
there is a decrease in the quaternization rate with increase in
degree of substitution. This may be due to steric effects imposed
by restricted rotation of the polymeric chains [191]. This phenomenon was not observed in
quaternization of poly(chloromethyl styrene). The chloromethylation
reaction of a polysulfone with chloromethyl ether, catalyzed by
stannic chloride, can be illustrated as follows:

Vinyl benzyl chloride monomer is available
commercially. It is possible, therefore, to simply prepare the
chloromethylated polystyrene or copolymers from the monomer without
the chloromethylation reactions.
9.4.3.3 Reactions of Halomethylated Polymers
Many known reactions of the halomethyl groups on
polymers are possible. One can, for instance, convert
poly(chloromethyl styrene) to poly(hydroxymethyl styrene)
[183]. Also, iodomethylated
polystyrene can be treated with triethyl phosphite in order to
carry out an Arbuzov
reaction [192]:

Chloromethylated polystyrene can also be
converted to a phosphonium salt for use in the Wittig reaction [193]:

The product of a reaction of chloromethylated
polystyrene and triphenylphosphine can also convert to nucleophiles
[194]. In addition, use of a
phase transfer catalyst converts soluble chloromethylated
polystyrenes to phosphine oxides. Reactions with dioctylphosphine
can serve as an example [195].
Sometimes phase transfer reactions are easier to carry out than
conventional ones. This is the case with a Witting reaction. Both
linear and cross-linked chloromethylated polystyrenes react
smoothly with triphenylphosphine to give derivatives that react
with various aldehydes [196,
197]. Phase transfer catalysts
can also be used in carrying out nucleophilic substitutions with
the aid of sulfides, like tetrahydrothiophine [198]:

When chloromethylated cross-linked polystyrene is
reacted with potassium superoxide, the yield depends upon the type
of solvent used. In dimethylsulfoxide, in the presence of
18-crown-6 ether, the conversion to hydroxymethyl groups is 45%. In
benzene, however, it is only 25%. High conversions are obtained by
catalyzing the reaction with tetrabutylammonium iodide in a mixture
of solvents. This results in 85% conversions to hydroxymethyl
groups, while the rest become iodide groups [199].
Quaternary salts are more effective than crown
ethers in reactions with salts of oxygen-anions, such as
carboxylate and phenolate [200].
On the other hand, lipophilic crown ethers, like
dicyclohexyl-18-crown-6, exhibit higher catalytic activity than the
quaternary salts in reactions with salts of the sulfur anions.
Also, the catalytic activity of the phase transfer catalysts toward
nitrogen anions is intermediate between that toward oxygen and that
toward sulfur anions. Solid–liquid two-phase systems generally give
higher degrees of conversion than do liquid–liquid systems. When,
however, lipophilic phase transfer catalysts are used with
lipophilic reagents, high degrees of substitutions are achieved in
liquid–liquid two-phase systems [200].
Conversion of chloromethylated styrene to anionic
exchange resins is done commercially by amination reactions to form
quaternary ammonium groups [201,
202]. This reaction can be
illustrated as follows [203]:

The kinetics of amination of chloromethylated
polystyrene with monohydroxy dialkyl tertiary amines shows that the
reactions proceed in two steps, at two different rates. The rate
changes take place at conversions of 45–50% [205]. These rates are favorably influenced by
increases in the dielectric constants of the solvents
[204]. Two different rate
constants also exist in reactions with
3-alkylaminopropionitrile.
9.4.3.4 Friedel-Craft Alkylation Reactions
Friedel-Craft acylations of polystyrene can be
carried out in CS2 or in CCl4 at reflux
temperatures of the solvents. The yields are high [209]:

Chloromethylated copolymers of styrene with
divinyl benzene undergo Friedel-Craft type reactions in
condensations with 1-phenyl-3-methylpyrazolone or with
1-phenyl-2,3-dimethylpyrazolone-5 in the presence of either
ZnCl2, BF3, or SnCl4
[208]:

The above condensation takes place in
dichloroethane, with stannic chloride catalyst at 50°C
[210]. The maximum reaction rate
varies with both, the initial concentration of
1,4-dimethyl-2,5-dichloromethylbenzene, shown above, and the
initial concentration of SnCl4. Cross-linked polystyrene
particles, or beads also form by Friedel-Craft suspension
cross-linking of polystyrene with 1,4-dichloromethyl-2,5
dimethyl-benzene [211]. The
polymer is dissolved in nitrobenzene and a two-phase reaction
occurs in 70% by weight of an aqueous suspension of
ZnCl2. Poly(vinyl alcohol) can be used as the suspending
agent.
9.4.3.5 Sulfonation Reactions
Sulfonation reactions of polystyrene and its
copolymers with divinyl benzene are carried out commercially to
prepare ion exchange resins. Partial sulfonations of polystyrenes
are achieved in the presence of ethers. When more than 50% of the
aromatic rings are sulfonated, the polymers become water-soluble.
At lesser amounts of sulfonation, 25–50%, the polymers are
solvent-soluble [212,
213].
When polystyrene is sulfonated in chlorinated
hydrocarbons with a complex of dioxane-SO3, the polymer
precipitates from solution at low concentrations [214, 215].
Complexes of ketones with SO3 can also be used to
sulfonate polystyrene in halogenated solvents [216]. The ratio of sulfonation is more
favorable for poly(vinyl toluene) than it is for polystyrene at the
same conditions [217]. Also,
sulfur dioxide swells polystyrene. The polymer can be sulfonated in
this medium with sulfur trioxide or with chlorosulfonic acid
[218]. Polystyrene, sulfonated in
CS2 with aluminum chloride catalyst, is water-insoluble
in a free acid form [219].
9.4.3.6 Nitration, Reduction, and Diazotization
Nitration of polystyrene was originally carried
out a long time ago [220]. A
nitrating mixture of nitric and sulfuric acids dissolves the
polymer and a nitro derivative forms at 50°C within 3 h
[221]:

The reaction is accompanied by a loss of
molecular weight. Nitration of isotactic polystyrene yields a more
crystalline product (about 1.6 NO2/ring) than the parent
compound [222]. Here too,
however, a loss in molecular weight accompanies the reaction
[223]. Polystyrene can be
nitrated under mild conditions using acetyl nitrate. The product
contains approximately 0.6 nitro groups per each benzene ring
[224].
The nitro groups of polynitrostyrene are reduced
by phenyl hydrazine that acts as a hydrogen donor [225]:

Polyaminostyrene can undergo typical reactions of
aromatic amines, such as diazotization [226]. The diazonium salt decomposes with
ferrous ions to yield polymeric free-radicals:

9.4.3.7 Metalation Reactions
Functional polystyrene derivatives are starting
materials for further reactions in many multistep syntheses. An
example is a metalation of polystyrenes for use as intermediates
[227]:

Some other reactions of lithiated polystyrene are
[227]:

When polystyrenelithium is aminated by a reagent
prepared from methoxyamine and methyllithium, two reaction
mechanisms are possible. One may proceed via nitrene intermediates
and the other one via electrophilic nitrenium ions [228]. Many other reactions of
polystyrenelithium can be found in the literature [229–232].
Sodium metalated polystyrene reacts in a similar manner
[228]:

9.4.4 Reactions of Acrylic, Methacrylic, and Related Polymers
The functional groups of polymers from acrylic
and methacrylic esters can undergo all the typical reactions of
such groups. There are, therefore, numerous reports in the
literature on such reactions.
9.4.4.1 Reduction of the Ester Groups
Perhaps the most reported reactions of these
polymers are reductions of
the functional groups. Among them is the reaction with lithium
aluminum hydride to reduce the ester groups. The success, however,
depends upon the reaction medium. Poly(methyl methacrylate) can be
reduced to poly(methallyl alcohol) in ether solvents
[236]:

The results, however, are inconclusive, because
combustion analyses fail to match the theoretical composition for
poly(methallyl alcohol). It is impossible to tell to what extent
the reduction takes place [236].
Inconclusive results are also obtained in similar reductions of
poly(methyl acrylate) in mixtures of tetrahydrofuran and benzene.
The product of such reduction is acetylated with acetic anhydride
in pyridine [237] as follows:


Hydrolysis in water of the product of
acetylation, followed by treatment with hot m-cresol, and subsequent extraction
with hydrochloric acid to remove the suspended inorganic matter
[237], yields a material that is
still only soluble in pyridine and m-cresol.
Somewhat similar results are obtained in
reductions of high molecular weight poly(methyl acrylate) with
lithium aluminum hydride [238] in
tetrahydrofuran. The reaction yields a product that is only soluble
in mixtures of hydrochloric acid with either methyl alcohol,
dioxane, or tetrahydrofuran. The problem is apparently due to some
residual aluminum that is hard to remove [239]. If, however, the reduction is carried out
in a N-methylmorpholine
solution, followed by addition of potassium tartrate, a pure
product can be isolated [240].
N-methylmorpholine is a
good solvent for reductions of various macromolecules with metal
hydrides [236]. In addition, the
solvent permits use of strong NaOH solutions to hydrolyze the
addition complexes that form. Other polymers that can be reduced in
it are those bearing nitrile, amide, imide, lactam, and oxime
pendant groups. Reduction of polymethacrylonitrile, however, yields
a product with only 70% of primary amine groups [241].
Complete reductions of pendant carbonyl groups
with LiAlH4 in solvents other than N-methylmorpholine, however, were
reported. Thus, a copolymer of methyl vinyl ketone with styrene was
fully reduced in tetrahydrofuran [242].
Reductions with metal hydrides are often
preliminary steps for additional reactions. For instance, a product
of LiAlH4 reduction of syndiotactic poly(methyl
methacrylate) can be reacted with succinic anhydride and then
converted to an amide [243]:

Substituted succinic anhydride can be used as
well. When poly(dimethyl itaconate) is reduced with
LiAlH4 in THF, the product contains some ash, but 93% of
the functional groups are reduced [239]:

9.4.4.2 Nucleophilic and Electrophilic Substitutions
Many other conversions of functional groups of
acrylic and methacrylic resins were reported. One of them is a
conversion of methyl acrylate to a hydrazide by a direct reaction
with hydrazine [244]:

The above reaction requires a 10:1 ratio of
hydrazine to the ester groups. In the laboratory, it can be carried
out on a steam bath over a period of 2–3 h. Approximately
60–80% of the ester groups convert [244]. The hydrazides can form various
hydrazones through reactions with aldehydes and ketones:

The hydrazide can also be converted to an azide
[244]:

A Hoffman
reaction of the azide yields a cross-linked polymer
[244]. Syndiotactic poly(methyl
methacrylate) converts to a hydrazide in a similar manner
[245].
Nucleophilic
substitution reactions can be carried out on poly(methyl
methacrylate) with heterocyclic organolithium reagents
[246]. The reactions are
conducted in homogeneous solutions in tetrahydrofuran or in benzene
combined with hexamethyl-phosphoramide. Copolymers will form with
tautomeric keto-β-heterocyclic structures. Following heterocyclic
reagents are useful [246]:
2-picolinyllithium, [(4,4-dimethyl-2-oxazole-2-yl)methyl]lithium,
quinaldinyllithium, and [2-thiazole-2-yl-methyl]lithium.
In attempts to carry out Arndt-Eister reactions on
poly(methacryloyl chloride), the polymer was reacted with
diazomethane in various molar ratios and at different temperatures
[247]. Initially, acid chloride
groups do react with diazomethane as expected. The products,
however, undergo subsequent reactions with neighboring acid
chloride groups and form cyclic structures [247]:

When Curtius and Lossen rearrangement reactions are
attempted on poly(acryloyl chloride) [248], the products are fairly regular
polyampholytes:

Somewhat similar results are obtained with a
Hoffman reaction on
polyacrylamide [279]. A
Schmidt reaction on
poly(acrylic acid) also yields mixed results [250]. When it is run in acetic acid, the
intermolecular reactions appear to predominate over the
intramolecular ones. Also, the products formed in acetic acid have
higher nitrogen content that those formed in dioxane
[250]. The NMR spectra show
presence of some acid anhydride groups. This has an additional
effect of lowering the yield.
Diels-Alder
reactions can be carried out on poly(furfuryl methacrylate)
with dienophiles like maleic anhydride or a maleimide
[252]. Dilute solutions (10%) of
the polymers in benzene can be used, requiring up to 30% molar
excess of the dienophiles:

The additions take place at room temperature and
the reactions take from 7 to 30 days [252] to complete.
Polyacrylonitrile reacts with hydroxylamine and
the product can be metalated by elements from Group IV (Sn, Ge, and
Si). This is a convenient route to formation of polymers with such
pendant organometallic groups [253]:

The yields are high when the reactions are
carried out on dilute solutions of polyacrylonitrile in
dimethylformamide at 75°C. The solutions must contain 1.5 mol
each of hydroxylamine hydrochloride and sodium carbonate per mole
of acrylonitrile groups [253].
A Ritter
reaction can be carried out on atactic polyacrylonitrile with
N-hydroxymethylamides of
acetic, benzoic, and benzene-sulfonic acids [254]. When the same reactions are carried out
with N-hydroxymethylimides
of succinic or phthalic acids in tetramethylene sulfone, there is a
stronger tendency toward cross-linking.
Copolymers of methacroyl chloride will undergo an
Arbuzov rearrangement in
reactions with triethyl phosphite in dimethylformamide, dioxane or
benzene at 75°C [255]. The
conversions are high, ranging between 96 and 98%.
The aldehyde groups of polyacrolein can be
reduced by the Meerwein-Ponndorf
reaction. There is a limit, however, to the amount of
alcoholate that can be used and to the concentrations of free
aldehyde groups in the starting material [256]. Also, ester condensations take place
(Tischenko reaction) at the
same time as the reductions occur [256].

9.4.5 Substitution Reactions of Poly(vinyl alcohol)
There are many practical uses for products from
reactions of poly(vinyl alcohol). Among them are commercial
preparations of poly(vinyl acetal)s formed through condensations
with aldehydes. Two materials that are currently being marketed are
poly(vinyl formal) and poly(vinyl butyral). The first one is formed
from partially hydrolyzed poly(vinyl acetate) that is dissolved in
aqueous acetic acid and excess formaldehyde. The mixture is heated,
sulfuric acid is added, and the reaction is allowed to proceed at
70–90°C for 6 h. Sulfuric acid is then neutralized and the
formal precipitates out.
Two different industrial processes are used for
preparations of the butyral. In both of them, acetate free
poly(vinyl alcohol) is used. In the first one, 10% solutions of the
starting material are treated with butyraldehyde and sulfuric acid.
The mixtures are heated to 90°C for 1½ h and the products
precipitate. They are neutralized, washed, and dried. In the second
one, poly(vinyl alcohol) is suspended in ethanol/ethyl acetate and
butyraldehyde together with a strong mineral acid is added. The
solutions are then neutralized. The butyrals separate out. They are
neutralized and the resin are washed and dried.
If poly(vinyl alcohol) films are reacted with
formaldehyde in water containing salt and an acid catalyst
(heterogeneous formalization), cross-linking occurs. The number of
the cross-links increases with decreasing acid concentration and
fixed amounts of formaldehyde and salt [257].
Direct reactions of poly(vinyl alcohol) with
aldehydes in the Kornblum
reaction result in formations of acetals that also contain
residual hydroxyl group and often acetate groups. The acetate
groups can be there from incomplete hydrolysis of the parent
poly(vinyl acetate) that was used to form the poly(vinyl alcohol).
Reactions of poly(vinyl alcohol) with ketones yield similar ketals.
At present, no ketals are offered commercially.
Alkyl etherification of poly(vinyl alcohol)
occurs when the polymer is combined with n-alkyl halides in dimethylsulfoxide
combined with pyridine [258,
259]. It was suggested that the
alkyl halides convert to aldehydes and acids and then act as
intermediates in the dimethylsulfoxide-pyridine solution
[258, 259]:

Many modifications of poly(vinyl alcohol) were
carried out to form photosensitive materials. Thus, unsaturation
was introduced into the pendant groups for photocross-linking. One
example is a condensation with pyridinium and quinolinium salts
[260]:

The material cyclodimerizes on exposure to light
[261] (see Chap.
10 for additional discussion of this
subject).
Schotten-Baumann
esterifications of poly(vinyl alcohol) are used extensively
in preparations of various derivatives. The reactions appear to
proceed well when acid chlorides are employed in two-phase systems
[262]. The polymers are dissolved
in water and the solutions are blended in 1:1:1 equal volume with
NaOH solutions and cyclohexanone. They are then mixed thoroughly
with solutions of the acid chlorides in mixtures of cyclohexanone
and toluene. The reaction mixture is stirred vigorously for about
90 min at −5 to 5°C to obtain the desired product.
The Ritter
reaction on poly(vinyl alcohol) yields soluble products.
Only some of the hydroxyl groups, however, are converted to amide
structures [264]:

Also, it is possible that neighboring group
interactions may lead to cyclizations and formations of
1,3-oxazines [264].
9.4.6 Miscellaneous Exchange Reactions
Many miscellaneous exchange reactions are
reported in the literature. A few are presented here. One such
reaction is reduction of pendant carbonyl groups of poly(vinyl
methyl ketone) with metal hydrides [242, 265]:

Another one is introduction of unsaturation into
pendant groups by a Wittig reaction on pendant carbonyls
[266]:

The same reaction can also be carried out on a
copolymer of ethylene and carbon monoxide [266].
Polycaprolactam can be treated with either
SO2Cl2, POCl3, or PCl5
at 70°C to introduce ionic chlorine groups [267]. The main product is
poly(α,α-dichloro-caprolactam):

Also, when Nylon 6,6 is reacted with
trifluoroacetic anhydride, trifluoroacetyl nylon forms
[268]:

Syndiotactic poly(2-methallyl hydrogen phthalate)
can be prepared and amidated according to the following scheme
[269]:
poly(pyridine ether sulfone)s in nitrobenzene [70]. The reaction can be illustrated as follows:


A 20% mol excess of the alkylating reagent is
required and the reaction must be conducted for 6 h at
80°C.
Dichloroketene, generated by the
ultrasound-promoted dechlorination reaction of thrichloroacetyl
chloride with zinc, adds to the carbon–carbon double bonds of
poly(methyl-1-phenyl-1-silane-cis-pent-3-ene) [271]. This can be illustrated as follows:

9.5 Cross-linking Reactions of Polymers
These reactions are quite numerous and have been
utilized for a long time. They also include all thermosetting
processes of polymers. Many are discussed in previous
chapters.
9.5.1 Vulcanization of Elastomers
Cross-linking of natural rubber was discovered by
Goodyear back in 1839. Sulfur, which was the original cross-linking
agent, is still utilized today in many processes. Early studies
demonstrated that the cross-links are mainly polysulfides:

The reactions take place at all temperatures, but
industrially they are carried out from 50 to 75°C and above. At
lower temperatures, however, the process may take days to complete.
At temperatures of 135–155°C, approximately 8% of sulfur (by weight
of rubber) reacts [272]. Also,
sulfur dissolves in unvulcanized rubber even at room temperature.
The overall mechanism of the reaction is still being studied. Most
evidence points to an ionic mechanism and a sulfonium ion
intermediate [272]. It was shown
[273] that a straightforward
reaction of sulfur with rubber is insufficient. Somehow, between 40
and 100 atoms of sulfur must be combined in order to obtain one
cross-link. Out of 40–100 atoms, only 6–10 are actually engaged in
the formation of the cross-links. The rest of the atoms form cyclic
sulfide units that become spread along the main chain
[273].
To improve the efficiency of the vulcanization
reaction, various accelerators were developed. Among them are zinc
oxide combined with fatty acids, and/or amines. Zinc oxide forms
zinc mercaptides like (XS)2ZnL2 where X is an
electron withdrawing substituent and L is a ligand from a carboxyl
or an amine group. The function of the ligand is to render the
complex soluble. The mercaptide complexes are believed to react
with additional sulfur to form zinc perthiomercaptides.
The accelerators that are most commonly used are
derivatives of 2-mercaptobenzothiazole. They are very effective
when used in combinations of metal oxides with fatty acids
(referred to as activators). The favorite activators
are zinc oxide combined with stearic acid. The combinations permit
rapid vulcanizations that take minutes compared to hours when
sulfur is used alone. In the process of vulcanization,
2,2′-dithiobisbenzthiazole forms initially and then reacts with
sulfur to form polysulfides [273]:

The products from reactions with sulfur in turn
react with natural or synthetic rubber at any allylic hydrogen.
This is a concerted reaction that results in formation of sulfur
containing adducts of the polymers:

Once the cross-links are formed, further
transformations take place. Some of them consist of reactions that
shorten the polysulfide links:

Also, some cross-links are lost through
elimination reactions:

In addition, some cyclic sulfur compounds form in
the process [273]:

9.5.2 Cross-linking of Polymers with the Aid of Peroxides
The reaction can be utilized with many polymers
such as polyolefins, polymers of dienes, and others. The reactions
with natural rubber can be illustrated as
follows [274]:


The extent of the cross-linking, as shown above,
is not clear. It is known, however, that cleavage reactions that
are followed by free-radical recombinations can take place
[274]:

Polymeric chains bearing free radicals combine
with each other to give branched structures. Additions of chains
with freeradical to double bonds result in formations of
cross-links [274].
9.5.3 Miscellaneous Cross-linking Reactions of Polymers
Many other miscellaneous cross-linkings of
polymeric materials are reported in the literature. For instance,
poly(acryloyl chloride) can be cross-linked with diamines
[275]:

In a similar manner, polymers with pendant
chlorosulfonate groups cross-link when reacted with diamines or
with glycols [275].
9.6 Graft Copolymers
This is an important part of polymer syntheses
that is used in many industrial processes. In 1967, Battaerd and
Tregear [282] published a book on
the subject that contains 1,000 references to journal publications
and 1,200 references to patents. In addition, there are several
monographs and many review papers [283]. The synthetic methods developed to date
range from using free radical attacks on polymeric backbones to
highly refined ionic reactions. There are examples where these
ionic reactions attach the side-chains at well-designated
locations.
9.6.1 Free-Radical Grafting by Chain-Transferring Process
This technique is probably one of the simplest
ways to form graft copolymers. It consists of carrying out
free-radical polymerizations of monomers in the presence of
polymers preformed from different monomers. A prerequisite for this
synthesis is that the active sites must form on the polymeric
backbones during the course of the reactions. Ideally, this occurs
if the steps of initiations consist only of attacks by the
initiating radicals on the backbones:

Propagations then precede from the backbone
sites:

Terminations can take place in many ways. Of
course, termination by combination will lead to cross-linked
insoluble polymers and that is undesirable. An ideal termination
takes place by chain transferring to another site on a polymer
backbone to initiate another chain growth.
The above, however, is an ideal picture. In
reality, the efficiency of grafting by this technique depends upon
the following:
1.
Competitions between the different materials
present in the reaction mixture, such as monomer, solvent, and
polymer backbone for the radical species. This includes competition
between chains growth and chain transferring to any other species
present.
2.
Competition between the terminating processes,
such as disproportionation and chain transferring.
The conditions can vary considerably and it is
possible to carry out the reactions in bulk, solution, or in
emulsion. When the reaction take place in emulsion, the success
depends greatly on experimental techniques. The rate of diffusion
is a factor and anything that affects this rate must be considered.
Because grafting efficiency depends upon chain transferring to the
backbone, knowledge of the chain-transferring constants can help
predict the outcome of the reactions. Sometimes, the information on
the chain-transferring constants is not available from the
literature. It may, however, be possible to obtain the information
from reactions of low molecular weight compounds with similar
structures [284–286]. One assumes equal reactivity toward
attacking radicals. The validity of such an assumption was
demonstrated on oligomers [281–283].
The reactivity of the initiating radicals toward
the backbones can vary and this can also vary the efficiency of
grafting. Benzoyl peroxide-initiated polymerizations of methyl
methacrylate monomer, for instance, in the presence of polystyrene
[284] yield appreciable
quantities of graft copolymers. Very little graft copolymers,
however, form when di-t-butyl peroxide initiates the same
reactions. Azobisisobutyronitrile also fails to yield appreciable
quantities of graft copolymers. This is due to very inefficient
chain transferring to the polymer backbones by t-butoxy and isobutyronitrile
radicals.
A better approach is work by Chung and coworkers
[288] who grafted maleic
anhydride to polypropylene with the use of Borane/O2
initiator. This initiator is claimed to form in situ a monooxidized
adduct (R–O* *O–BR2). These adducts then carry out
hydrogen abstractions from the polypropylene chains at ambient
temperature. This results in formation of stable tertiary polymer
radicals that react with maleic anhydride to form graft
copolymers:
Not all chain transferring to the backbones
results in formations of graft copolymers. An example is
polymerization of vinyl acetate in the presence of poly(methyl
methacrylate). No graft copolymers form and this is independent of
the reactivity of the initiators [285]. In fact, grafts of poly(vinyl acetate) to
poly(methyl methacrylate) and to polystyrene cannot be prepared by
this technique [286].
Grafting efficiency may increase with temperature
[287]. This could be due to
higher activation energy of the transfer reaction than that of the
propagation reaction [288].
Meaningful effects of temperature, however, are not always
observed. In grafting polystyrene to poly(butyl acrylate) in
emulsion, for instance, there is no noticeable difference between
60 and 90°C by this technique [289].
The presence of sites with high transfer
constants on the polymeric backbone enhances the efficiency of
grafting. Such sites can be introduced deliberately. These can, for
instance, be mercaptan groups [290] that can be formed by reacting
H2S with a polymer containing epoxy groups:

Free-radical polymerizations of acrylic and
methacrylic esters in the presence of the above backbones result in
high yield of graft copolymers [291, 292].
Another example is formation of mercaptan groups
on cellulose in order to form graft copolymers [293]:

Pendant nitro groups are also effective in chain
transfer grafting reactions. Thus, graft copolymers of polystyrene
with cellulose acetate p-nitrobenzoate [294] and with poly(vinyl p-nitrobenzoate) [295] form readily. Nitro groups appear to be
more effective in formations of graft copolymers by radical
mechanism than are double bonds located as pendant groups
[294].
9.6.2 Free-Radical Grafting Reactions to Polymers with Double Bonds
Carbon to carbon double bonds, either in the
backbone or in the pendant groups, are potential sites for
free-radical attacks. In addition to the double bonds, the hydrogen
atoms on the neighboring carbons are allylic and potential sites
for chain transferring. Because rubbers, natural and synthetic,
possess such unsaturations, they are used extensively as backbones
for various grafting reactions. Whether the reactivity of the
initiating radical is important in determining grafting efficiency
is not completely established. Graft copolymers of poly(methyl
methacrylate) on gutta-percha, however, form in good yields when
the initiator is benzoyl peroxide [296, 297].
Yet, when azobisisobutyronitrile is used, only a mixture of
homopolymers forms. Work with 14C-labeled initiators
shows that the primary radicals react both by addition to the
double bonds and by transfer to the methylene group [298]. Grafting reactions to polybutadiene,
however, proceed via chain transferring from the growing chain
radical to the backbone [299].
Nevertheless, strong evidence also shows that the initiator
radicals can interact directly with polymeric backbones
[299, 300].
When graft copolymers of polystyrene to natural
rubber form, the chain length of the attached branches equals to
the chain lengths of the unattached polystyrene homopolymer that
forms simultaneously [301]. This
led to the following conclusions [303]:
1.
As the concentration of rubber increases, the
length of the grafted branches diminishes, while their number
remains the same.
2.
When the concentration of the initiator
increases, the length of the branches diminishes, but the number of
branches increases.
3.
When the concentration of monomers increases, the
length of the branches also increases, but their number
diminishes.
4.
When the polymerization time increases, the
length of the branches remains the same, but their number
increases.
9.6.3 Preparation of Graft Copolymers with the Aid of Macromonomers
The chain-transferring methods for preparing
graft copolymers suffer from the disadvantages of low efficiency
and contamination by homopolymers. The efficiency in forming graft
copolymers, however, increases with the use of macromonomers. A macromonomer is a
macromolecular monomer, an oligomer, or a polymer with a
polymerizable end group. When macromonomers are copolymerized with
other monomers, comb-shaped polymers form [302, 303].
The copolymerizations can be free radical or ionic in mechanism.
Some examples of macromonomers are presented in
Table 9.1. A preparation of graft copolymers with the
aid of macromonomers can be illustrated as follows [304]:
Table
9.1
Some macromonomers reported in the
literature
Macromonomer
|
References
|
---|---|
![]() |
[320]
|
![]() |
[321]
|
![]() |
[322]
|
![]() |
[323]
|
![]() |
[324]
|
![]() |
[325]
|
![]() |
[326]
|
![]() |
[327]
|
![]() |
[328]
|
![]() |
[329]
|
![]() |
[330]
|
![]() |
[331]
|

Many variations on the above technique are
possible.
9.6.4 Initiations of Polymerizations from the Backbone of Polymers
High degrees of grafting by free-radical
mechanism can be attained when polymerizations are initiated from
the backbones of the polymer. One way this can be done is to form
peroxides on the backbone structures. Decompositions of such
peroxides can yield initiating radicals. Half of them will be
attached to the backbones. An example is preparation of graft
copolymers of polystyrene [305,
306]:

Thermal cleavage of the above peroxides leads to
both macromolecules with free-radicals sites. Hydroxyl radicals
also form and initiate formations of homopolymers. Decompositions
of the peroxides by redox mechanisms, however, increase the yields
of graft copolymers, but do not stop all formations of hydroxy
radicals [303]:

Some homopolymers still form [307].
Air oxidation of polypropylene can result in
formation of hydroperoxide units at the sites of the tertiary
hydrogens [383]. The polymer can
also be oxidized when dissolved in cumene that contains some cumene
hydroperoxide at 70–80°C. A product containing 0.8% oxygen by
weight as a hydroperoxide [308]
can be formed and can subsequently be reacted to form graft
copolymers. Various monomers [309–311] can
be used, such as vinyl acetate or 2-vinyl pyrrolidone.
Many other hydroperoxidations of polymers were
reported in the literature. The materials are used in formations of
graft copolymers. One example is hydroperoxidation of poly(ethylene
terephthalate) and poly(ε-caproamide). The products yield graft
copolymers with various acrylic and methacrylic esters and acrylic
and methacrylic acids [312–314].
Ozone treatment of polymers can also cause
hydroperoxidation of labile hydrogens. It can, however, also cause
extensive degradation of the backbone polymers, because attacks by
ozone on double bonds in the backbones convert them to unstable
ozonides. Starch can be ozonized to form graft copolymers
[315, 316]. The same is true of cellulose
[317], poly(vinyl chloride)
[318], and polyethylene
[319]. Hydroperoxides form
without noticeable degradation. This allows subsequent preparations
of graft copolymers.
In a similar manner, it is possible to start with
copolymers of acryloyl or methacryloyl chloride and react them with
hydroperoxides [332]. This can be
illustrated as follows:

The decomposition of the pendant peroxide in the
presence of vinyl monomers yields mixtures of graft copolymers and
homopolymers.
Preparation and subsequent decomposition of
polymers with diazonium salts can also be used to form
graft-copolymers. An example is nitrated polystyrene, reduced to
the amine derivative and then diazotized [333]. The decomposition of the diazonium salt
results in formation of radicals:

The radical sites are capable of initiating
polymerizations of monomers. A similar approach can be taken with
cellulose [334]. Mercerized
cotton and sodium salt of carboxymethyl cellulose will react with
p-aminophenacyl chloride:

The material can be converted to diazonium salts
and then decomposed with ferrous ions in the presence of some vinyl
monomers to form graft copolymers. Acrylonitrile forms graft
copolymer readily without formation of any homopolymers. Styrene
and vinyl acetate, however, do not. A modification of this
technique is to conduct the diazotization reaction in the presence
of emulsifiers [335]. The amounts
of graft copolymers that form with acrylic and methacrylic monomers
and N-vinylpyrrolidone
depend upon the nature and pH of the emulsifiers, the reaction
time, and the temperature.
Ceric ions form graft copolymers with various
macromolecules by a redox mechanism. The reactions can be
illustrated as follows:

The almost exclusive formation of free radicals
on the polymeric backbones results in formations of many products
that are close to being free from homopolymers [346, 350].
The reactions are widely used in formations of graft copolymers of
poly(vinyl alcohol) and particularly of cellulose and starch. The
grafting reaction fails, however, when attempted on polysaccharides
that lack free hydroxyl groups on the second and third carbons.
This led to speculation [351]
that the bond between these carbons cleaves. In the process, free
radicals, presumably, form on the second carbons and aldehyde
structures on the third carbons of the glucose units. This point of
view, however, is not generally accepted. Instead, it was proposed
that more likely positions for attacks by the ceric ions are at the
C1 carbons of the glucoses at the end of the
polysaccharide chains [352]. This
is supported by observation that oxidation of cellulose is an
important prerequisite for the formation of graft copolymers
[353].
Graft copolymerizations by redox mechanism can
also take place with the aid of other ions. This includes grafting
on cellulose backbones with ferrous ions and hydrogen peroxide
[354]. Redox grafting reactions
can also take place on nylon and on polyester. For instance, graft
copolymers of methyl methacrylate on nylon 6 can be prepared with
manganic, cobaltic, and ferric ions [355]. Another example is grafting poly(glycidyl
methacrylate) on poly(ethylene terephthalate) fibers with the aid
ferrous ion–hydrogen peroxide. The reaction depends on the
concentration of the monomer, hydrogen peroxide, time, and
temperature [356].
9.6.5 Photochemical Syntheses of Graft Copolymers
Photo labile groups on polymers can serve as
sites for photoinitiated graft copolymerizations. For instance,
when polymers and copolymers of vinyl ketone decompose in
ultraviolet light in the presence of acrylonitrile, methyl
methacrylate or vinyl acetate graft copolymers form [357]:

The free radicals that are unattached to the
backbone polymers, like the methyl radical shown above, also
initiate polymerizations and considerable amounts of homopolymers
form as well.
In some instances, graft copolymers form as a
result of chain transferring that takes place after
photodecomposition of the photo labile materials. An example is
formation of graft copolymers of polyacrylamide on natural rubber,
poly(vinyl pyrrolidone), or dextrin with the aid of benzophenone
and ultraviolet light [358]. The
free radicals from photodecomposition of benzophenone react with
the polymers by chain transferring. The growth of acrylamide is
subsequently initiated from the polymer backbones. Photo tendering
dyes can be used in this manner with cellulose [359]. Thus, anthraquinone dyes can be adsorbed
to cellulose. Upon irradiation, proton abstractions take place,
creating initiating radicals on the backbone polysaccharide:

It is believed that the above dye mono radicals
disproportionate to hydroquinones and quinones. Transfer reactions
to solvent lead to formations of homopolymers. This gives high
yields of graft copolymers of methyl methacrylate with cellulose.
The same is true of acrylonitrile [360]. On the other hand, only small quantities
of graft copolymers form with styrene or vinyl acetate monomers
[360].
It is also possible to form graft copolymers on
the surface of fibers by coating them with photoinitiators, like
benzophenone together with a monomer and then irradiating them with
ultraviolet light [414]. Similar
to the action of the anthraquinone dyes shown above, benzophenone
in the excited triplet state mainly abstracts hydrogens and forms
radicals on the surface [415].
9.6.6 Graft Copolymer Formation with the Aid of High-Energy Radiation
High-energy radiation sources include gamma rays
from radioactive elements, electron beams from accelerators, and
gamma rays from nuclear reactors. The energy radiated by these
sources is sufficiently high to rupturing covalent bonds. This
results in formations of free radicals. Several different methods
are used to form graft copolymers. One of them is irradiation in open air. The free
radicals that form scavenge oxygen and form peroxides and
hydroperoxides on the polymeric chains. These are subsequently
decomposed in the presence of monomers to form graft copolymers.
When they are decomposed thermally [361], the hydroperoxides yield much greater
quantities of homopolymers then do the peroxides. When, however,
the decompositions are done at room temperature by redox
mechanisms, formations of homopolymers are reduced [361, 362].
Another method is irradiation
in vacuum. This results in formations of trapped radicals on
the polymer backbones. After irradiation, the polymers are heated
in the absence of oxygen and in the presence of vinyl monomers. The
best results are obtained when irradiations are done at low
temperatures, below T
g of the polymers. High degree of crystallinity is also
beneficial, because mobility of the chains results in loss of
trapped radicals. When the monomers are added, however, heat must
be applied, but this can result in loss of some of the radicals.
The third method is mutual
irradiation in an inert atmosphere of polymers and monomers
together. The polymers are either swelled or dissolved by the
monomers. The relative sensitivities of the two species, the
monomer and the polymer to radiation, can be important factors in
this third procedure. Efficiency of grafting depends upon
formations of free radicals on the polymer backbones. If only a
small number of free radicals form, the irradiations produce mainly
homopolymers. Also, if the polymers tend to degrade from the
irradiation, block copolymers form instead. Presence of solvents
and chain-transferring agents tends to lower the amount of the
grafting [363].
It was also shown [409] that energetic heavy ions can also produce
graft copolymers. The results appear similar to those obtained by
electron beams. Also, many papers reported use of plasma for
surface modification of films. The process can result in formation
of graft copolymers when it is accompanied by an introduction of a
monomer or monomers. One such example is a use of argon plasma to
graft polyacrylamide to polyaniline films [410]. The near ultraviolet light plasma induces
the reaction. Other monomers that can be grafted by this reaction
are 4-styrenesulfonic acid and acrylic acid [410].
9.6.7 Preparation of Graft Copolymers with Ionic Chain-Growth and Step-Growth Polymerization Reactions
Both anionic and cationic mechanisms can be used
to form graft copolymers. A typical example of graft copolymer
formation by anionic mechanism is grafting polyacrylonitrile to
polystyrene [364]:

Another example of ionic graft copolymerization
in a reaction carried out on pendant olefinic groups using
Ziegler-Natta catalysts in a coordinated anionic type
polymerization [365]. The
procedure consists of two steps. In the first one, diethylaluminum
hydride is added across the double bonds. The product is
subsequently treated with a transition metal halide. This yields an
active catalyst for polymerizations of α-olefins. By this method,
polyethylene and polypropylene can be grafted to butadiene styrene
copolymers. Propylene monomer polymerization results in formations
of isotactic polymeric branches:

Another example is formation of graft copolymers
of formaldehyde with starch, dextrin, and poly(vinyl alcohol)
[366, 367]. This procedure is also carried out in two
steps. Potassium naphthalene is first reacted with the backbone
polymer in dimethylsulfoxide. The formaldehyde is then introduced
in gaseous form to the alkoxide solution.
A similar reaction can be used to form graft
copolymers of poly(ethylene oxide) on cellulose acetate
[391]. Poly(ethylene oxide) can
also be grafted to starch. For instance, a preformed polymer
[392] terminated by chloroformate
end groups can be used with potassium starch alkoxide:

The products are water-soluble. The efficiency of
the coupling process, however, decreases with an increase in the DP
of poly(ethylene oxide).
Lithiated polystyrene reacts readily with
halogen-bearing polymers like polychlorotrifluoroethylene
[411]. This can be utilized in
formation of graft copolymers. The reactions can be conducted in
solutions as well as in preparations of surface grafts on films
[411].
An example of a cationic grafting reaction is
formation of graft copolymers of polyisobutylene on polystyrene
backbones [393]. Polystyrene is
chloromethylated and then reacted with aluminum bromide in carbon
disulfide solution. This is followed by introduction of
isobutylene:


The above, however, yields only 5–18% of a graft
copolymer, even at −60°C. It is possible that considerable amounts
of cross-linking occur at the reaction conditions and may, perhaps,
be the reason for the low yield [393].
Another example is grafting to cellulose.
BF3 can be adsorbed to the surface of the polymer. It
then reacts with hydroxy groups and yields reactive sites for
cationic polymerization of α-methyl styrene and isobutylene
[402]. These reactions are
carried out at −80°C.
Cationic graft copolymerizations of trioxane can
be carried out with the help of reactive C–O–C links in a number of
polymers, like poly(vinyl acetate), poly(ethylene terephthalate),
and poly(vinyl butyral) [403].
Many graft copolymers can also be formed by ring opening
polymerizations [404]. The
reactions with active hydrogens on the pendant groups, like
hydroxyl, carboxyl, amine, amide, thiol, and others, can initiate
some ring opening polymerizations. An example is preparation of
graft copolymers of ethylene oxide with styrene [405]. Copolymers of styrene with 1–2% of
hydrolyzed vinyl acetate (vinyl alcohol after hydrolysis) can
initiate polymerizations of ethylene oxide and graft copolymers
form.
Recently, solutions of polysilanes were treated
with controlled amounts of triflic acid
(CF3SO2OH) in CH2Cl2
and afterwards with tetrahydrofuran. This yielded a graft copolymer
of poly(tetramethylene oxide) on polysilane backbones
[412].
An interesting series of papers were published by
Kennedy and coworkers on use of alkylaluminum compounds as
initiators of graft copolymerizations [366, 367].
Allylic chlorines form very active carbon cations in the presence
of this initiator. This is also true of macromolecular carbon
cation sources [402]. As a
result, very high grafting efficiency is achieved in many different
polymerizations using macromolecular cationogens and alkylaluminum
compounds. In some instances, formation of graft copolymers is
greater than 90%. The grafting reaction can be illustrated as
follows [367].

The temperatures of the reactions and the nature
of the aluminum compounds are the most important synthetic
variables [367].
On the other hand, many other graft
copolymerizations by cationic mechanism suffer from low grafting
efficiencies. They are also often accompanied by large formations
of homopolymers. Use, however, of living cationic processes appears
to overcome this drawback. An illustration of this can be another
preparation of a graft copolymer of polyisobutylene on a
polystyrene backbone [413]:

9.6.8 Miscellaneous Graft Copolymerizations
In a rather interesting reaction, ethylene oxide
can be graft-copolymerized with nylon 6,6 [406]. Formation of the graft copolymer greatly
enhances flexibility of the material, while the high melting point
of the nylon is still maintained. Thus, nylon 6,6 that contains as
much as 50% by weight of grafted poly(ethylene oxide) still melts
at 221°C and has an apparent T g below −40°C. It also
maintains flexibility and other useful properties over a wide range
of temperatures [406]:

An entirely different procedure can be used to
form graft copolymers by a step-growth polymerization
[347]. Formaldehyde is condensed
with either phenol, p-cresol, or p-nonyl phenol and the resin is
attached to either nylon 6, nylon 6,6, nylon 6,10, or nylon 11
backbones. Initially, the formaldehyde is prereacted with an excess
of phenol in the presence of the nylon, but without any catalyst,
at temperatures high enough to cause condensation. This is followed
by addition of toluenesulfonic acid at lower temperatures. At that
point, when free formaldehyde is no longer present in the reaction
mixture to cause gelation, the novolac molecules attach themselves
to the nylon backbones. The excess phenol is washed away, leaving
pure graft copolymers [347].
Yagci and coworkers reported a special
preparation of perfectly alternating poly(p-phenylene) amphilitic graft
copolymers by combination of controlled free-radical polymerization
and Suzuki coupling process [345].
9.7 Block Copolymers
These polymers consist of two or more strands of
different polymers attached to each other. There does not appear to
be any general stipulation to the minimum size of each block. There
does appear to be, however, a general agreement that each sequence
should be larger than just a few units. In describing a block
copolymer, it is helpful if the following structural parameters are
available to characterize the material:
1.
Copolymer sequence distribution as well as the
length and the number of blocks.
2.
The chemical nature of the blocks.
3.
The average molecular weight and the molecular
weight distribution of the blocks and of the copolymer.
Block copolymers, particularly of the A—B—A type,
can exhibit properties that are quite different from those of
random copolymers and even from mixtures of homopolymers. The
physical behavior of block copolymers is related to their solid
state morphology. Phase separation occurs often in such copolymers.
This can result in dispersed phases consisting of one block
dispersed in a continuous matrix from a second block. Such
dispersed phases can be hard domains, either crystalline or glassy,
while the matrices are soft and rubber-like.
An interesting example of block copolymers is
work by de Ruijter et al. [348],
who prepared a series of block copolymers that contain rigid liquid
crystal forming blocks of poly(p-phenylene terephthalamide) and
flexible blocks of hexamethylene adipamide. The polymers have been
prepared in a one-pot procedure by a low-temperature
polycondensation reaction in N-methyl-2-pyrrolidone.
9.7.1 Block Copolyesters
Two polyester homopolymers can react and form
block copolymers in a molten state at temperatures high enough for
ester interchange [414]. As the
reaction mixtures are stirred and heated, the interchanges
initially involve large segments. With time, however, smaller and
smaller segments form as the transesterifications continue. To
prevent eventual formation of random copolymers, the reactions
should be limited in time.
Ester interchange can be retarded, particularly
when esterification catalysts like zinc or calcium acetate are
present by addition of phosphorous acid or triphenyl phosphite
[415]. This improves the chances
of forming block copolymers. The procedure can be applied to
preparation of block copolymers of poly(ethylene terephthalate)
with poly(ethylene maleate), poly(ethylene citraconate), and
poly(ethylene itaconate) [416].
With ester interchange catalysts, like titanium alkoxides or their
complexes, melt randomization may be inhibited by adding arsenic
pentoxide that deactivates them [417].
Block copolyesters also form in reactions between
hydroxy and acid chloride-terminated prepolymers [419]. This can take place in the melt or in
solution in such solvents as chlorobenzene or o-dichlorobenzene [418]. For relatively inactive acid chlorides,
like terephthaloyl chloride, high reaction temperatures are
required. Phosgene also reacts with hydroxy-terminated polyesters
to form block copolymers. The reactions must be carried out in
inert solvents. Block copolyethers also form readily by ester
interchange reactions with low molecular weight diesters
[348]:


Acetates of tin, lead, manganese, antimony, and
zinc as well as esters of orthotitanates catalyze the reactions
[421]. Optimum temperatures for
these reactions are between 230 and 260°C at 0.03–1 mm Hg
pressure [421]. Block copolymers
can also form by ring opening polymerizations of lactones, when
carboxyl-terminated macromolecular initiators are used
[422]:

9.7.2 Block Copolyamides
Simple melt blending reactions can also be
applied to preparations of block copolyamides, similarly to the
process for polyesters. With time, total equilibrium conditions
also are gradually achieved in the melt [423]. Interfacial polycondensation is also
useful in preparation of block copolymers. When mixed diacid
chlorides and/or mixed diamines are reacted, the more active diacid
chlorides and/or diamines react preferentially and blocks form. In
addition, it is possible to carry out the growth of one of the
segments first, to a fairly large size and follow it by addition of
the other comonomers [424].
9.7.3 Polyurethane-Polyamide Block Copolymers
These block copolymers can be formed in many
ways. One technique is to prereact a diamine with a diacid
chloride. The polyamide that forms is then treated with
bischloroformate to attach to polyurethane blocks [425]. The process can be reversed and the
polyurethane can be formed first and then attached to polyamide
blocks [425].
9.7.4 Polyamide-Polyester Block Copolymers
Block copolymers consisting of polyamide and
polyester blocks can form through melt blending [426]. The reactions probably involve aminolyses
of the terminal ester groups of the polyesters by the terminal
amine groups of the polyamides. Ester interchange catalysts
accelerate the reaction [427].
Block polyester-polyamides also form through
initiation of ring opening polymerizations of lactones by the
terminal amine groups of the polyamides [428]:

If the polyamide has terminal amine groups at
both ends, then triblock copolymers form.
9.7.5 Polyurethane Ionomers
These materials were reviewed as a special class
of block copolymers [433]. They
are linear polyaddition products of diisocyanates containing
nonrandom distributions of ionic centers. The preparations are
similar to those of polyurethane elastomers that are described in
Chap.
6. One example is a material prepared from high
molecular weight polyester that is free from ionic centers and that
is terminated by isocyanate groups at each end. The prepolymer is
coupled with N-methylamino-2,2′-diethanol to form a
segmented polymer:

Similar products form from isocyanate-terminated
polyethers. This material can be cross-linked with difunctional
quaternizing agents, such as 1,4-bis(chloromethyl) benzene
[434]:

The products are cationic ionomers. Anionic
ionomers form very similarly through coupling of chains with
bifunctional anionic “chain lengtheners” [435]:

Because of interactions between the chains, the
polyurethane ionomers are similar in properties to cross-linked
elastomers. In solution, they are strongly associated.
9.7.6 Block Copolymers of Poly(α-Olefin)s
These block copolymers form readily when
appropriate Ziegler-Natta catalysts are used [436]. This is discussed in Chaps.
4 and 6. In addition, there is a special
technique for preparations of such block copolymers. At the outset
of the reaction, only one monomer is used in the feed. A typical
catalyst might be
α-TiCl3/(C2H5)3Al.
After the first monomer has been bubbled in for a short period,
perhaps 5 min, the addition is stopped and the unreacted
monomer removed by evacuation or by flushing. The second monomer is
then introduced and may also be bubbled in for the same period. The
addition of the second monomer may then be stopped, the second
monomer evacuated, and the whole thing might be repeated. If equal
length of each block is desired, the addition times of each monomer
may be varied to adjust for different rates of polymerization
[436].
Kuhlman and Klosin [436] reported forming multiblock copolymers of
polyethylene and tuning block composition by catalyst selection.
Chain shuttling, the fast exchange of growing polymer chains among
catalyst centers mediated by a chain shuttling agent (CSA), has
enabled the production of ethylene-based olefin block copolymers.
Diethyl zinc was used as a chain shuffling agent. Thus,
copolymerization of ethylene and α-olefin by certain catalyst pairs
in the pretense of a CSA produces blocks that are alternately
highly or lightly branched. Chain shutling polymerizations are bet
conducted in continuous reactors.
A catalyst system derived from titanocene complex
and methyl-aluminoxane (see Chap.
5) has been used to polymerize propylene and,
depending on polymerization conditions, produce a block copolymer
of crystalline and amorphous, elastomeric polypropylene
[440]:

In the above process, the formation of
crystalline domains involves consecutive insertions from one of the
lateral coordination sites of the catalyst so as to give rise to
isotactic sequences, whereas consecutive insertions at the other
site (2) give rise to atactic amorphous sequences. Interconversion
between these two states must occur within the lifetime of a given
polymer chain in order to generate a physically cross-linked
network and is believed to occur via occasional isomerizations of
the polymer chains (i.e., interconversion at the metal center).
Preparation of an oscillating catalyst that yields an elastomeric
polypropylene was also reported by others [441].
9.7.7 Simultaneous Use of Free Radical and Ionic Chain-Growth Polymerizations
This technique allows formation of many different
types of block copolymers [437].
Lithium metal can be used to initiate polymerizations in solvents
of varying polarity. Monomers, like styrene, α-methylstyrene,
methyl methacrylate, butyl methacrylate, 2-vinylpyridine, 4-vinyl
pyridine, acrylonitrile, and methyl acrylate, can be used. The
mechanism of initiation depends upon formation of ion-radicals
through reactions of lithium with the double bonds:

Propagation reactions proceed from both active
sites, the radical and the carbanion. When two different monomers
are present, free-radical propagation favors formation of
copolymers, while propagation at the other end favors formation of
homopolymers. There is a tendency, therefore, to form AB—B type
block copolymers.
Several publications appeared recently that
describe use of controlled/“living” radical polymerizations to form
block copolymers. Thus, Jerome et al. [435] described formation of block copolymers by
using an initiator capable of initiating simultaneously dual living
polymerizations:

In a similar manner, Yoshida and Osagawa
[436] synthesized
poly(ε-caprolactone) with 2,2,6,6-tetramethylpiperdine-1-oxyl
(TEMPO) at one end by anionic polymerization of caprolactone using
an aluminum tri(4-oxy-TEMPO) initiator. The TEMPO-supported
polycaprolactone behaved as a polymeric counter radical for a
controlled/“living” radical polymerization of styrene to form block
copolymers [436].
Also, Kotani et al. [437] reported using controlled/“living” atom
transfer radical polymerization (ATRP) to form block copolymers of
ethyl and n-butyl
methacrylates. A ternary initiating system that consists of carbon
tetrachloride, tris(triphenyl-phosphine)ruthenium dichloride
[RuCl2(PPh3)3], and aluminum
compounds produced ABA triblock copolymers [437].
Huang and coworkers [437] reported preparation of a series of
well-defined amphiphilic block copolymers containing conjugated
poly(fluorene) (PF) block and coil like poly(2-(dimethylamino)ethyl
methacrylate) (PDMAEMA). The block copolymers were synthesized
through ATRP. The reactions were initiated by a 2-bromoisobutyrate
end-capped macroinitiator using
CuCl/1,1,4,7,10,10-hexamethyltriethylenetetramine as the
catalyst.
Matron and Grubbs formed block copolymers by
combining ring opening metathesis polymerization with ATRP
[437]. Use was made of fast
initiating ruthenium metathesis catalyst to form three different
monotelechelic poly(oxa)norbornenes. The ends were functionized and
ATRP polymerizations of styrene and tert-butyl acrylate followed.
Coca et al. [438] showed a general method of transforming
living ring opening metathesis polymerization into
controlled/“living” atom transfer polymerizations to form block
copolymers. Ring opening polymerizations of norbornene or
dicyclopentadiene were followed by Witting-like reactions with
p-(bromomethyl)benzaldehyde
to form efficient (ATP) macroninitiators for formation of block
copolymers with styrene [478]:

Other cationic ring opening polymerizations can
also be transformed to ATRP to yield block copolymers
[439]. Thus, formation of block
copolymers was initiated by poly(tetramethylene glycol) containing
one bromopropionyl end group. These were used to form block
copolymers by ATP polymerization of styrene, methyl methacrylate,
and methyl acrylate.

9.7.8 Preparation of Block Copolymers by Homogeneous Ionic Copolymerization
Formation of block copolymers by this method
depends upon the ability to form “living” chain ends. Among the
anionic systems, the following polymerizations fit this
requirement:
1.
Polymerizations of nonpolar monomers with alkali
metal-aromatic electron transfer initiators in ethers
[398].
3.
Acrylonitrile polymerizations in dimethyl
formamide initiated by sodium triethylthioisopropoxyaluminate at
−78°C [340].
5.
Polymerization of alkyl isocyanates initiated by
organoalkali species in hydrocarbons at −78°C [342].
Among the cationic “living” polymerizations that
can be used for block copolymer formation are:
1.
2.
3.
Polymerization of p-methyl styrene, N-vinylcarbazole, and indene with
appropriate catalysts.
The preparations by anionic mechanism of A—B—A
type block copolymers of styrene and butadiene can be carried out
with the styrene being polymerized first. Use of alkyl lithium
initiators in hydrocarbon solvents is usually a good choice, if one
seeks to form the greatest amount of cis-1,4 microstructure [346]. This is discussed in Chap.
4. It is more difficult, however, to form block
copolymers from methyl methacrylate and styrene, because “living”
methyl methacrylate polymers fail to initiate polymerizations of
styrene [347]. The poly(methyl
methacrylate) anions may not be sufficiently basic to initiate
styrene polymerizations [345].
A “living” cationic polymerization of
tetrahydrofuran, using BH3 as the initiator in the
presence of epichlorohydrin and
3,3-bis(chloromethyl)oxacyclo-butane [348], results in formation of block copolymers.
Two types form. One is an A—B type. It consists of
polytetrahydrofuran blocks attached to blocks of
poly(3,3-bis(chloromethyl)oxacyclo-butane). The other one is an
A—AB—B type [348].
The preparation of well-defined sequential
copolymers by anionic mechanism has been explored and utilized
commercially for some time now. Initially, the cationic methods
received less attention until it was demonstrated by Kennedy
[424] that a large variety of
block copolymers can be formed. The key to Kennedy’s early work is
tight control over the polymerization reaction. The initiation and
propagation events must be fundamentally similar, although not
identical [424]:
-
Ion generation:
-
Cationization and propagation:In this scheme, chain transfer to monomer must be absent and the termination is well defined.
-
Termination:
This allows formation of macromolecules with
terminal halogens. They can be used to initiate new and different
polymerizations.
Three methods were developed to overcome transfer
to monomer [424]. These are: (1)
use of inifers; (2) use of
proton traps; and (3) establishing conditions under which the rate
of termination is much faster than the rate of transfer to monomer.
The first one, the inifer method, is particularly useful in
formation of block copolymers. It allows preparation of
head and end (α and ω) functionalized telechelic
polymers. Bifunctional initiators and transfer agents (inifers) are used. The following
illustrates the concept [424]:
-
Ion generation:
-
Cationization and propagation:
In the above scheme, the inifer, XRX, is usually
an organic dihalide. If chain transferring to the inifer is faster
than chain transferring to the monomer, the polymer end groups
become exclusively terminated with halogens.
It is also possible to carry out “living”
cationic polymerization of isobutylene, initiated by a difunctional
initiator [435]. This results in
a formation of bifunctional “living,” segments of polyisobutylene
that are soft and rubbery. Upon completions of the polymerization,
another monomer, one that yields stiff segments and has a high
T g, like
indene, is introduced into the living charge. Polymerization of the
second monomer is initiated from both ends of the formed
polyisobutylene. When the reaction is complete, the polymerization
is quenched. Preparations of a variety of such triblock and star
block polymers have been described [435].
A technique was developed, by introducing
cationic to anionic transformation [438]. A “living” carbocationic polymerization
of isobutylene is carried out first. After it is complete, the ends
of the chains are quantitatively transformed to
polymerization-active anions. The additional blocks are then built
by an anionic polymerization. A triblock polymer of poly(methyl
methacrylate)-polyisobutylene-poly(methyl methacrylate) can thus be
formed. The transformation takes several steps. In the first one, a
compound like toluene is Friedel-Craft alkylated by
α,ω-di-tert-chloropolyisobutylene. The
ditolylpolyisobutylene, which forms, is lithiated in step two to
form α,ω-di-benzyllithium polyisobutylene. It is then reacted with
1,1-diphenylethylene to give the corresponding dianion. After
cooling to −78°C and dilution, methyl methacrylate monomer is
introduced for the second polymerization [438] in step 3.
Formation of block copolymers from polymers with
functional end groups has been used in many ways. In anionic
polymerization, various technique were developed for terminating
chain growth with reactive end groups. These end groups allow
subsequent formations of many different block copolymers. One such
active terminal group can be toluene diisocyanate [439]. The isocyanate group located ortho to the
methyl group is considerably less reactive toward the lithium
species due to steric hindrance [438]. The unreacted isocyanate group can be
used for attachment of various polymers that are terminated by
hydroxy, carboxy, or amine groups. Other functional compounds that
can be used in such reactions are alkyl or aryl halides, succinic
anhydride, n-bromophthalimide [448], and chlorosilanes [449].
Because block copolymers can often offer
properties that are unattainable with simple blends or random
copolymers [364], many efforts
were made to combine dissimilar materials, like hydrophilic with
hydrophobic, or hard with soft segments, as was shown earlier. One
paper [432] describes formation
of block copolymers containing helical polyisocyanide and an
elastomeric polybutadiene. Compound
[(η3-C3H5)-Ni(OC(O)CF3)]2
was used to carry out “living” polymerization of butadiene and then
followed by polymerization of tert-butyl isocyanide to a helical
polymer.
9.7.9 Special Reactions for Preparation of Block Copolymers
A special case is the use of the Witting reaction. Poly(p-phenylene pentadienylene)
[415] is prepared by this
reaction first. This is utilized in a preparation of a block
copolymer [456] according to the
following scheme:

A preparation of block copolymers by “living”
ring-opening olefin metathesis polymerization was reported
[417]. Initially, norbornene or
exo-dicyclopentadiene are
polymerized by bis(η5-cyclopentadienyl)titanacyl butane.
The resulting living polymers are then reacted with
terephthaldehyde to form polymers with terminal aldehyde groups.
The aldehyde groups in turn initiate polymerizations of
t-butyldimethylsilyl vinyl
ether by aldol-group transfer polymerizations [418]. Following is an illustration of the
process:

Subsequently, the terminal aldehyde group is
reduced with NaBH4 and the silyl groups cleaved off by
treatment with tetrabutylammonium fluoride to produce a
hydrophobic-hydrophilic A—B diblock copolymers.
Living metathesis type polymerization was also
employed to form block copolymers from norbornene and its
derivative with bimetallic ruthenium catalysts,
(PR3)2Cl2Ru(=CH-p-C6H4CH=)RuCl2(PR3)2
[434]. This can be illustrated as
follows [434]:

Pitet and Hillmyer [436] formed triblock copolymers from
1,5-cycloocatadiene and dl-Lactide by combining ring opening
metathesis polymerization with cyclic ester ring opening
polymerization.

Proto et al. [434] living isoselective coordination
polymerization of styrene to form isoselective block copolymers.
This was accomplished by sequential monomer addition.
9.7.10 Miscellaneous Block Copolymers
Polyamide-polyether block copolymers can be
formed by a variety of techniques. One of them consists of initial
preparation of amine-terminated polyethers. This can be done by
reacting hydroxy-terminated polyethers with acrylonitrile and then
reducing the nitrile groups to amines [429]:

The products of the reduction condense with
carboxylic acid-terminated polyamides to form block
copolymers.
Another one is to form the polyethers with
terminal chloride groups [430].
Hydroxy-group-terminated polyethers, for instance, can be converted
to halogen-terminated polyethers. The products will react with
ammonia and the amine-terminated polymer will react with carboxylic
acid-terminated polyamides [430].
A British patent describes preparations of block
copolymers in two steps. In the first one, two different salts of
hexamethylene diamine are formed; one with carboxylic
acid-terminated polyoxyethylene and the other one with adipic acid
(nylon 6,6 salt). In the second step, the two salts are reacted in
the melt. Caprolactam can be used in place of the second salt
[431]. Also, a Japanese paper
describes formations of block copolymers by reacting
polyoxyethylene in melt condensation reactions with caprolactam in
the presence of dicarboxylic acids [432].
Polyamide-6 (nylon 6) can form block copolymers
with rubber [419] and with
poly(dimethylsiloxane) [420]. In
the latter case, the polysiloxane forms first by “living”
polymerization and is terminated by an acylated caprolactam. The
caprolactam portion of the molecule is then polymerized with the
aid of lithium caprolactamate:

This diblock copolymer can be melt annealed at
ca. 250°C. It exhibits superior mechanical properties to nylon 6
homopolymer [420].
9.7.11 Mechanochemical Techniques for Formation of Block Copolymers
These techniques rely upon high shear to cause
bond scissions. Ruptured bonds result in formations of free radical
and ionic species [413]. When
this application of shear is done in the presence of monomers,
block copolymers can form. This approach is exploited fairly
extensively. Such cleavages of macromolecules can take place during
cold mastication, milling, and extrusion of the polymers in the
visco-elastic state. Both homolytic and heterolytic scissions are
possible. The first one yields free radical and the second one
ionic species. Heterolytic scissions require more energy, but
should not be written off as completely unlikely [413]. Early work was done with natural rubber
[413]. It swells when exposed to
many monomers and forms a visco-elastic mass. When this swollen
mass is subjected to shear and mechanical scission, the resultant
radicals initiate polymerizations. The mastication reaction was
shown to be accompanied by formation of homopolymers
[413]. Later the technique was
applied to many different polymers with many different monomers
[414].
9.8 Processes in Polymer Degradation
There are many causes of degradation of polymers.
The chief ones among them are heat, oxidation, light, ionic
radiation, hydrolysis, and mechanical shear. The effect can be
discoloration, loss of molecular weight, cross-linking, or
cyclization. The loss of molecular weight can be so severe that the
polymer is degraded to a mixture of monomers and oligomers. The
other effects can be blackening or charring and loss of useful
properties. On the other hand, mechanical shearing is often applied
to some polymers like rubbers to deliberately reduce molecular
weight for commercial processing. In the environment, synthetic
polymers generally degrade due to man-made environmental pollutants
in the atmosphere, like carbon monoxide, sulfur dioxide, nitrogen
oxide, and oxidizing smog rich in ozone.
Molecular weight loss occurs through the breaking
of primary valence bonds. Such chain scissions may occur at random
points along the polymer backbone or they may take place at the
terminal ends of the polymer where monomer units are released
successively. This last effect can be compared to unzipping. The response of any
particular polymeric material to specific causes of degradation
depends upon the chemical structure.
This chapter presents the degradation processes
in some typical commercial polymers. The materials selected were
those that received major attention in the literature.
9.8.1 Thermal Degradation of Common Chain-Growth Polymers
Thermal degradation of polymers is conveniently
studied by pyrolytic methods. The polymer literature contains many
reports on such studies conducted at various temperatures in inert
atmospheres, in air, or in vacuum. The volatile products are
usually monitored with accompanying measurements of the weight loss
per unit time. The reaction rates are thus measured by:
1.
Loss of molecular weight as a function of
temperature and the extent of degradation
2.
The quantity and the composition of the volatile
and nonvolatile products of degradation
3.
The activation energy of the degradation
process
A general scheme for thermal degradations of
chain-growth polymers by free-radical reactions can be written as
follows:
-
Random type chain scission:
-
Chain-end type scission:
In a thermal chain depolymerization, the
degradation can also be random. Rupture occurs at various points
along the chain. The products are various size fragments of the
polymer, usually larger than individual monomers. Both reactions
shown above can take place simultaneously in the same polymer chain
or only one of them might take place exclusively. Also, chain
depropagation may not necessarily be initiated from the terminal
ends of the macromolecules, though such depolymerizations are more
common. They may instead start from some points of imperfection in
the chain structures. These imperfections might consist of
incorporated initiator fragments, or peroxides, or ether links.
They might have formed as a result of oxygen molecules being
present during chain growths. Weak points in polymer backbones can
also be at locations of some tertiary hydrogens. Each individual
polymer will depolymerize at its own specific rate and the
degradation products will be peculiar to the particular chemical
structure of the polymer [446–452]. For
instance, poly(methyl methacrylate) can be converted almost
quantitatively back to the monomers. The depolymerization occurs
from the terminal end in an unzipping fashion with the overall
molecular weight decreasing slowly in proportion to the amount
volatilized. Polyethylene, on the other hand, undergoes scission
into longer olefin fragments. Very little monomer is released. At
the same time, the molecular weight tends to decrease rapidly and
only a small amount of volatilization takes place. The two polymers
are good examples of extreme behavior of chain-growth polymers.
Most of the chain-growth polymers, however, fall between the two.
The depolymerization reaction of chain-growth polymers generally
occurs by a free-radical mechanism and the reactions are similar.
If depropagation is the major portion of the degradation process,
then the molecular weight reduction is proportional to the quantity
of the monomer that forms. If, on the other hand, chain
transferring is the major portion of the degradation process, then
there is rapid loss of molecular weight with little accompanying
monomer accumulation and the reaction products are large segments
of the chains. The rate of depolymerization exhibits a
maximum.
When depropagation takes place at an elevated
temperature, at a rate that is equal to the propagation in a
free-radical polymerization, then the temperature of the reaction
is a ceiling temperature (see Chap.
3). Termination can take place by
disproportionation. Secondary reactions, however, may occur in the
degradation process depending upon the chemical structure of the
polymer. Such side reactions can, for instance, be successive
eliminations of hydrochloric acid, as in poly(vinyl chrolide), or
acetic acid as in poly(vinyl acetate).
9.8.2 Thermal Degradation of Polyolefins and of Polymers from Conjugated Dienes
One early study of thermal degradation of polyethylene was
carried out on low molecular weight polymers [453]. Later the work was repeated with
high-density polyethylene [454].
The volatile products were identified by gas chromatography. The
biggest portion of the volatiles was found to be propylene. The
remaining materials were methane, ethane, propane, and a number of
unsaturated and saturated higher hydrocarbons. The quantity of
ethylene increased slightly with an increase in the temperature of
pyrolysis. Formation of propylene can occur in two ways. The first
one is through intramolecular radical transfer to the second
carbon, followed by a decomposition reaction. This, however, is not
believed to contribute significantly to the quantity of propylene
obtained, based on theoretical consideration [454]. The second one can occur as a result of
scissions of C–C bonds located in the positions β to the terminal
double bonds. Such double bonds probably form in large quantities
during decomposition and this process is thought to be the main
source of propylene [454]:
It was observed that isotactic polypropylene decomposes
thermally by a mechanism that varies at different temperatures and
conditions [455]. Thus, at 340°C
the major volatile product is propane, while at 380°C it is
n-pentane, and at 420°C it
is propylene. The propane is believed to originate from some weak
spot in the polymeric chain. Formation of n-pentane involves a radical
abstraction and a six-membered ring formation in a backbiting
process. Propylene may come from a free-radical depolymerization
process or a cyclic six-membered ring formation involving a
terminal double bond [455].
The thermal degradation of diene polymers was the
subject of several studies [456,
457]. The scheme for polyisoprene
and polybutadiene degradation was postulated in part by Golub and
Garguila [458, 460]. It is based on infra-red spectra and NMR
studies of the products:

In addition, spectroscopic evidence shows that
cis–trans isomerizations as
well as cyclizations occur in the process of thermal degradation of
1,4-polyisoprene. It is interesting that the cis–trans isomerizations were observed
at temperatures as low as 200°C [460].
9.8.3 Thermal Degradation of Polystyrene and Polystyrene-Like Polymers
Thermal degradation of atactic polystyrene
results in formation of volatile products that contain as much as
42% of the monomer and small quantities of toluene, ethylbenzene,
and methylstyrene. The rest of the volatile material is made up of
dimers, trimers, and tetramers. No large fractions were isolated.
This suggests that the main mechanism of decomposition is
depropagation. The rate, however, exhibits a maximum and there is a
rapid decline in molecular weight. This indicates scission of the
chains. Some evidence was presented that rapid decrease in
molecular weight is a result of scission of weak points in the
polymer and is independent of the free-radical depropagation
reaction [461–465]. Some of the weak spots are believed to be
occasional “head-to-head” placement of monomers [19]. It was suggested that some weak spots may
also be structures that form as a result of monomer additions to
the aromatic ring during the chain growth (see Chap.
6) [466]. The
following scheme of chain scission was, therefore, proposed
[466]:


The identity of the weak spots, however, has not
been established with certainty [467]. The evidence does indicate that the “weak
points” are head-to-head bonds, branch points, or unsaturated
structure [467].
The rate of conversion of polystyrene into
volatiles was measured in numerous studies.
In contrast to polystyrene, poly(α-methyl
styrene) yields in vacuum pyrolyses at temperatures between 200 and
500°C 95–100% monomer. By comparison, polystyrene only yields about
40.6% [457]. The difference can
be attributed to the fact that hydrogen transfer is completely
blocked from the sites of chain scission by the methyl groups in
the α-positions [457]. As a
result, the terminal free-radicals unzip into monomers and
dimers.
Polystyrenes that are substituted on the benzene
ring, like poly(vinyl toluene), behave similarly to polystyrene
when pyrolyzed [457]. Also,
poly(m-methylstyrene) at
350°C yields 44.4% monomer as compared to polystyrene that yields
40.6% monomer at these conditions. The rate, however, for
polystyrene at this temperature is 0.24 mol.%/min, while for
poly(m-methylstyrene) it is
0.9 [457].
9.8.4 Thermal Degradation of Methacrylic and Acrylic Polymers
The thermal degradation of polymers of acrylic and methacrylic alkyl esters is a process
of depolymerization to monomers at temperatures up to 250°C,
provided that the alkyl group is small, less than butyl
[468]. Poly(t-butyl methacrylate) yields
quantitatively isobutene instead. It was shown that thermal
depolymerization to monomers is probably common to all
poly(methacrylate ester)s. As the size of the alkyl group
increases, however, particularly within secondary or tertiary
structures, there is increased tendency for the alkyl group to also
decompose. This decomposition interferes with the depolymerization
process
Thermal stability of poly(methyl methacrylate)
appears to vary with the molecular weight. For instance, a sample
of the polymer of molecular weight of 150,000 when heated in vacuum
for 30 min at 318°C yields 74.3% volatiles. By comparison, a
sample of this polymer of molecular weight of 5,100,000 when heated
for 30 min at 319°C yielded only 35.2% volatiles
[456].
The thermal stability of copolymers of
long-chained diol dimethacrylates was investigated [583]. These copolymers included 1,4-butane-,
1,5-pentane-, 1,6-hexane-, 1,8-octane-, 1,10-decane-, or
1,12-dodecanediol dimethacrylates, respectively, as well as
2,2-bis[4-(2-hydroxy-3-methacryloyl-oxypropoxy)-phenyl]propane and
triethylene glycol dimethacrylate. The polymers were found to be
thermally stable up to ≈250°C, as shown by the initial
decomposition temperature and their degradation profiles.
A quantitative investigation of the thermal
degradation of poly(ethyl acrylate), poly(n-propyl acrylate), poly(isopropyl
acrylate), poly(n-butyl
acrylate), and poly(2-ethylhexyl acrylate) demonstrated that the
principle volatile products are carbon dioxide, olefin, and alcohol
corresponding to the alkyl group [469, 470].
The following mechanism of degradation was proposed [469, 470]:

More recently, another study was carried out on
the thermal decomposition of homopolymers of ethyl methacrylate,
n-butyl methacrylate, and
2-hydroxyethyl methacrylate as well as their copolymers
[471]. The copolymers of
hydroxyethyl methacrylate with ethyl methacrylate and butyl
methacrylate were found to degrade by unzipping to yield the
monomers similarly to poly(methyl methacrylate). In addition, there
is competition between unzipping and cross-linking in binary
copolymers of hydroxyethyl methacrylate with ethyl methacrylate and
in n-butyl
methacrylate.
Thermal degradation of nitrile polymers, particularly
acrylonitrile, was studied in detail [472–478]. It
was shown that there can be two paths of degradation, depending
upon the temperature (see also Chap.
8). One reaction takes place at a low
temperature, between 100 and 200°C, and the other one occurs above
240°C. At the low temperature, the polymers develop dark color.
There is, however, very little evolution of volatiles. At the high
temperature, on the other hand, there are evolutions of volatiles
and thermally stable residues develop.
The low temperature darkening process of
polyacrylonitrile was shown to be intramolecular cyclization and
polymerization of the cyanide groups [475, 476].
The overall reaction can be illustrated as follows:

The above shown ladder structure is actually a
vary idealized picture. Nevertheless, the formation of fused
tetrahydropyridine rings was demonstrated by IR data
[475, 476]. At the same time, there are many
irregularities in the above shown structure. Also, it was
demonstrated that the longer are the isotactic sequences in the
polymeric structures, the longer are the sequences of ladder
structures that form [479].
High temperature degradation of polyacrylonitrile
leads to formation of oligomers. The general form of the
oligomerized material can be shown as follows [480]:
where, x = 0–2 and
n = 0–5.

The above structure occurs as a result of an
initiation and termination process of cyclization at frequent
intervals along the chain. The growth of the ladder structure,
however, terminates due to hydrogen transfer. This process is
associated with the atactic sequences in the polymer chain:

Polymethacrylonitirle develops color
upon heating as a result of linking up of adjacent carbons and
nitrogen atoms in intramolecular cyclization reactions similar to
acrylonitrile.
This reaction is initiated primarily by
impurities that are often present in the polymer both at the end of
the chains and at various locations at the backbones. Due to this
ring formation, the amount of monomer that can be obtained from the
polymer at 200°C is approximately 50%. If, however, the polymer is
prepared from highly purified monomer, the yield of monomer upon
thermal degradation at 300°C is 100%. Also, a yellow color does not
develop from such a polymer at temperatures of 120–220°C
[457].
9.8.5 Thermal Degradation of Chlorocarbon and Fluorocarbon Polymers
The thermal instability of poly(vinyl chloride) is a cause for
concern commercially and has, therefore, stimulated extensive
investigations. The overall process of degradation is complex and
still not completely resolved. Some of the questions that remain
are:
1.
Is the overall process of degradation ionic or
free radical in nature?
2.
What is the exact mechanism of initiation?
It is easilly observable that the degradation is
accompanied by evolution of HCl and blackening of the polymer.
Also, at elevated temperatures, poly(vinyl chloride) not only gives
off hydrogen chloride, but this dehydrochlorination is also
accompanied by rapid depolymerization [483]. The rate of decomposition decreases with
increasing molecular weight and is highest in oxygen and lowest in
helium atmosphere. It was claimed that HCl exerts no catalytic
effect upon the rate of decomposition [483]. This contrasts with evidence that was
presented earlier that HCl does accelerate thermal degradation
[484]. The earlier evidence was
obtained from autocatalytic decmposition of films. Based on that
eveidence, it was hypothesized that the chloride ion acts as a base
(supporting an ionic mechanism of decomposition) [484]:

In a more recent study, HCl was also found to be
essential for the initiation of the unzipping process, but may or
may not be essential to the depolymerization reaction itself, once
the process has been started [485]. The above shown ionic mechanism was
disputed in a study where ESR signals were recorded during thermal
decomposition of the polymer at elevated temperatures
[486]. This strongly supports a
free-radical mechanism. A free-radical mechanism was proposed
earlier by Bamford and Feuton [487]. This mechanism illustrates the formation
of hydrochloric acid and is based on rupture of carbon–chlorine
bonds:
-
Initiation:
-
Propagation:
Subsequent interaction of macromolecular radicals
leads to cross-linking. Many additional investigations demonstrated
that thermal decompositions of vinyl polymers with pendant
electronegative groups that we can designate as X, result, after
elimination of HX and formation of macromolecular residues, with
polyene structures [488]. In
addition to that, at higher temperatures the polyene sequences that
form rearrange into a large numbers of aromatic hydrocarbons
[489–492]. Formation of polyenes is common to
poly(vinyl chloride), poly(vinyl bromide), poly(vinyl alcohol), and
poly(vinyl acetate) [493]. The
polyene structures partly decompose at formation due to bond stress
and molecular reorganization processes. The stresses result from
formation of conjugated structures along the polymeric chains. When
the chains fracture, the remaining portions of polyene sequences
form aromatic compounds [47]. It
was postulated that the process initially involves reactions of
enone groups, which are present in poly(vinyl chloride) as
anomalies, with chlorine units of a neighboring polymeric chains
[494, 495]:

The first step is formation of a dihydropyran
ring. This is followed by a retro-Diels-Alder splitting and leads
to regeneration of the α,β-unsaturated ketone and to formation of a
double bond in the poly(vinyl chloride) molecule. The double bond
initiates a subsequent elimination of hydrogen chloride from that
molecule [495].
An intramolecular initiation process that
explains constant rate of dehydrochlorination was also proposed
[496]:

It was pointed out that poly(vinyl chloride) is,
in a sense, its own worst enemy, in that all the structural defects
that are known to contribute appreciably to its thermal
instability. They are formed in the polymer molecules during the
process of polymerization by routes involving hydrogen abstraction
from the polymer backbone [584].
Gupper et al. used micro Raman spectroscopy to
study thermal degradation of poly(vinyl chloride) containing
various additives [586]. They
observed a linear increase in conjugated sequences in the process
of dehydrochlorination.
Carty and coworkers investigated thermal
decomposition of chlorinated poly(vinyl chloride) [496]. The thermal decomposition of pure
chlorinated poly(vinyl chloride) (without stabilizer or lubricant)
was studied by dynamic thermogravimetric analysis at heating rates
from 5 to 100°C/min in atmospheres of nitrogen, air, and oxygen. In
each case, a two-step decomposition was observed, similar to that
for poly(vinyl chloride) where dehydrochlorination is followed by
pyrolysis/oxidation of the carbonaceous residue. The rate of
dehydrochlorination was dependent on atmosphere, occurring slightly
slower in nitrogen than in air, and slightly more quickly in oxygen
than in air. The decomposition of the residual char was clearly
dependent on the conditions that it formed in. Under dynamic
conditions, chars formed at high heating rates appeared more
resistant to oxidative degradation than those formed more slowly.
However, when chars were formed by heating at different rates and
then held at 500°C, the char formed at the slowest heating rate was
the slowest to be oxidized. The uptake of oxygen by the char
appears to be rate limiting. At low heating rates, char oxide is
similar in both air and oxygen. As the heating rate is raised, the
rate of mass loss of char in air becomes progressively closer to
that in nitrogen until at 100°C/mm they are almost identical.
The rates of thermal decompositions of
poly(vinylidine chloride)s
were shown to depend upon the method by which the polymers were
prepared [497]. Those that were
formed from very pure monomers by mass polymerization are most
stable. Polymers prepared by emulsion polymerization, on the other
hand, degrade fastest. The mechanism of degradation of
poly(vinylidine chloride) was proposed to be as follows
[498–500]:
1.
Hydrochloric acid is eliminated in a chain
reaction.
2.
Conjugated sequences condense to form
cross-linked structures.
Some support for the above mechanism came from
stepwise heating studies of poly(vinylidine chloride)
[501].
Although polytetrafluoroethylene has the
reputation for being quite stable thermally, it does degrade at
elevated temperatures. The polymer upon pyrolysis yields almost
100% of monomer. The mechanism is believed to be free-radical
unzipping of the chains until the entire chain is consumed. This
can be illustrated as follows [457]. Initially, the chain ruptures:
this is followed by formation of monomers:


Polychlorotrifluoroethylene is less
stable thermally than polytetrafluoro-ethylene. It yields as much
as 86.0% volatiles in 30 h at 331.8°C [11]. These volatiles contain large amounts of
monomer. A mechanism that resembles the postulated degradation
mechanism of polytetrafluoroethylene was proposed [457].
Unlike polytetrafluoroethylene, polyvinyl fluoride, poly(vinylidine
fluoride), and polytrifluoroethylene yield primarily
on heating HF [457]. Among these
three, poly(vinylidine fluoride) yields larger amounts of HF than
do the other two polymers with an accompanying formation of double
bonds.
9.8.6 Thermal Degradation of Poly(Vinyl Acetate)
Thermal decomposition of poly(vinyl acetate)
results in a loss of acetic acid. The reaction is typical of
thermal cleavages of esters. It is facilitated by formation of
pseudo six-membered rings as a result of interactions between the
β-hydrogens of the alcohol residues and the carboxylic groups:

When double bonds form, adjacent methylene groups
become activated. The loss of acetic acid is the main product at
temperatures up to 200–250°C. Beyond these temperatures, aromatic
pyrolytic compounds form.
Studies of thermal degradation of copolymers of
vinyl chloride with vinyl acetate showed that the copolymers are
thermally less stable than the homopolymers [502, 503].
The ratio of hydrochloric acid to that of acetic acid that
volatilize remains constant during the degradation. This indicates
that neither is evolved preferentially, once the reaction begins
[502, 503]. It is interesting to note that
degradation studies of a copolymer of vinyl chloride and styrene
also demonstrated that the copolymer is less stable than each of
the homopolymers [504].
9.9 Thermal Degradation of Common Step-Growth Polymers
The thermal decomposition of step-growth polymers
cannot take place by a chain reaction like that of chain-growth
polymers. As a result, these materials degrade in a random fashion,
rupturing at the weakest bonds first.
9.9.1 Thermal Degradation of Polyoxides
Polyoxymethylene depolymerizes into
formaldehyde at 220°C. This was found to be a first-order reaction
with the rate varying from 0.42 to 5.8%/min, depending upon
conditions of polymer preparation and the molecular weight of the
polymer [457].
Poly(ethylene
oxide) decomposes upon heating at lower temperatures than
does polyethylene. Among the volatile products are found
formaldehyde, ethanol, ethylene oxide, carbon dioxide, and water.
Poly(propylene oxide) is
also less heat stable than polypropylene. Isotactic poly(propylene
oxide) is somewhat more stable than the atactic one.
9.9.2 Thermal Degradation of Polyesters
Poly(ethylene
terephthalate) decomposes upon heating through a series of
different reactions. These run either concurrently or
consecutively. The result is a complex mixture of volatile and
nonvolatile products. It was found that when poly(ethylene
terephthalate) is maintained in molten condition under an inert
atmosphere at 282–323°C, it slowly converts to a mixture of gaseous
low molecular weight fragments [581]. The major products from pyrolysis of
poly(ethylene terephthalate) are carbon dioxide, acetaldehyde and
terephthalic acid. In addition, there can be detected trace amounts
of anhydrides, benzoic acid, p-acetylbenzoic acid, acetophenone,
vinyl benzoate, water, methane, ethylene, acetylene, and some
ketones [505]. The following
mechanism of degradation was postulated [505]:

The vinyl end groups that form from cleavage of
the ester groups decompose further in a number of ways:


The thermal degradation of poly(butylene terephthalate) was
examined with the aid of a laser microprobe and mass spectrometry
[506]. A complex multistage
decomposition mechanism was observed that involves two reaction
paths. The initial degradation takes place by an ionic mechanism.
This results in an evolution of tetrahydrofuran. This is followed
by concerted ester pyrolyses reactions that involve intermediate
cyclic transition states and result in formation of 1,3-butadiene.
Simultaneous decarboxylations occur in both decomposition paths.
The latter stages of decomposition are characterized by evolutions
of carbon monoxide and various aromatic compounds, like toluene,
benzoic acid, and terephthalic acid. The first step can be shown as
follows [506]:

The subsequent decomposition, shown below, can
actually take place at lower temperatures:

There are indications that there is moisture
among the decomposition products. This may imply that acid
hydrolysis plays a part in tetrahydrofuran formation:

In oxidations of hydrocarbons, oxygen is believed
to act as a diradical in the ground state. This would explain
radical combination reactions:
and the subsequent hydrogen abstraction reaction:


The rate of formation of the peroxy radical is
much higher than is the rate of hydrogen abstraction
[521]. The overall rate of
oxidation of polymeric materials by atmospheric oxygen is strongly
affected by light, heat, oxygen concentration, moisture, and the
presence of traces of impurities. The impurities, however, can act
as either catalysts or as inhibitors of oxidation.
The degradation of poly-α-esters was studied on
poly(isopropylidine carboxylate) [518] over a range of temperatures, from 200 to
800°C. Among the decomposition products were found tetramethyl
glacolide, acetone carbon monoxide, and to a lesser extent
methacrylic acid. The primary decomposition product appears to be
tetramethylene glacolide that becomes an intermediate upon further
pyrolysis:
also:


9.9.3 Thermal Degradation of Polyamides
The thermal degradation of polyamides starts with
free-radical cleavage of nitrogen–carbon bonds [520]. Degradation of nylon 6 can be illustrated as follows:

When nylon 6 is heated for 100 h at 305°C,
half of the nitrogen escapes from the polymer and a small amount of
carbon dioxide forms [457]. This
was postulated by Kamerbeek et al. as being the result of reactions
of two terminal amine groups [579]:

In decomposition of nylon 6,6, there is the additional
tendency for ring closure by adipic acid [509]:

The cyclization reaction can also occur as
follows [509]
or, perhaps, as a bimolecular reaction [509]:


At higher temperatures, the carbamoyl ketone
might convert to the Schiff base by a monomolecular process
[509]:

The route to the Schiff base may also be through
elimination of an isocyanate [509]:

Rather than thermal degradation of nylon 6, it is
possible to depolymerize this polyamide. Kamimura and Yamamoto
reported that it is possible to depolymerize this nylon back to
caprolactam [510]:

The best yield of caprolactam was 86% obtained
with N-methyl, N-propyl piperidinium
bis(trifluoromethyl-sufonyl)imide as the solvent and N,N-dimethylpyridine as the
catalyst.
In the past, it was believed that at high
temperatures nylons degrade at a faster rate at elevated humidity.
This assumption, however, was shown to be erroneous by Bernstein et
al. [457].
9.9.4 Thermal Degradation of Epoxy Resins
The thermal degradations of Bisphenol A-based epoxy resins
cross-linked with diaminodiphenylmethane or with phthalic anhydride
were studied with the aid of pyrolysis and radiochemical gas
chromatography technique [511].
The products of degradation depend upon the temperature. At 400°C,
the amine cross-linked resins yield hydrogen, methane, and water.
The resins cross-linked with phthalic anhydride yield hydrogen,
methane, carbon dioxide, and water. The number of degradation
products increases with temperature. It was concluded that the most
important nonscission reactions in these resins are competing with
dehydration and dehydrogenation reactions [511]:

A nonscission reaction that yields methane is:

The scission reactions can take place at various
weak spots. The breakdown of bisphenol A segments produces phenol
[511]:

The main differences that were observed between
amine-cured and anhydride-cured resins are [511]:
1.
The amine cross-linked resin generated more water
and hydrogen, because they contained more
–CH2–CHOH–CH2– groups.
2.
Scissions of anhydride-reacted resins tend to
regenerate the anhydride and release CO and CO2 in large
quantities
3.
The aliphatic segments of the amine cross-linked
resins yield more acetaldehyde than acetone. The reverse is true of
the anhydride-cured resins. This is thought to be due to
preferential rupture of carbon–nitrogen bonds [511]:

The anhydride-cured compositions may tend to
break up symmetrically instead:

9.9.5 Thermal Degradation of Polyimides, Polyoxidiazoles, and Polyquinoxalines
The technique mentioned in the previous section
of pyrolysis and radiochemical gas chromatography was also applied
in a study to thermal degradation of aromatic polyimides [512]. Aromatic polyimides are more stable
thermally than previously discussed polymers and require higher
temperature for decomposition. Based on the experimental results,
the following mechanism was proposed [512]. Initially, hydrogen radicals form. They
come from various places of the polymer backbone during the
pyrolytic decomposition. Also, decomposition of amic acid may yield
water (see Chap.
6).

Another study of thermal degradation of aromatic
polyimides led to similar conclusions [513]. At lower temperatures, the degradation of
polyimides proceeds by a hydrolytic mechanism that is greatly
influenced by presence of uncyclized units. At higher temperatures,
the degradation is mainly by decarbonylation of the imide ring and
breaking down of the polymer chain through “weak” bonds.
The thermal degradation of polyoxidiazoles was
shown to proceed mainly through the heterocyclic rings that are
apparently the weak spots [514,
515]:

The process by which aromatic polyquinoxalines
decompose is random. At 500–600°C, it involves the opening of the
heterocyclic ring with loss of fragment [516]:

At still higher temperatures, between 640 and
690°C dehydrogenation of nitrogen residues takes place.
9.9.6 Thermal Degradation of Aromatic Polysulfones
In aromatic
polysulfones, the weakest link was found to be the
carbon–sulfur bond [517]. These
polymers are processed at about 300°C. At that temperature, the
thermodegradative processes as well as thermo-oxidative ones can
occur. The rupture of the C–S bonds produces molecules of
SO2 and fragments of the polymer molecules with terminal
phenyl radicals [517]:

The phenyl radicals then abstract hydrogens from
the methyl groups that are present when the polysulfone is prepared
from bisphenol A:

9.9.7 Thermal Degradation of Polyethers
The aromatic
polyethers are subject to rupture at the ether link at
around 300°C [518]. Products from
such ruptures are somewhat similar to those obtainable from
decomposition of the polysulfones, shown above.
The thermal degradation of polyoxymethylene was found to be
initiated at the chain ends. There appear to be three possible
sites for initiation of homolytic bond cleavages [519]:
9.9.8 Thermal Degradation of Cellulosic Materials
The degradation of cellulose triacetate in vacuum
was analyzed with the aid of chromatography, mass spectrometry,
infra-red, and NMR spectroscopy [517]. The mechanism of degradation was proposed
by Scotney to consist primarily of deacetylation in the polymer
chain and scission of the chain at the 1,4 glycosidic linkage
between the pyranose rings [517].
The type of products that are formed depends upon the degradation
temperature. At temperatures above 250°C but below 350°C, unstable
intermediates form. Continued heating results in loss of carbon
monoxide, carbon dioxide, and acetic acid, and eventually end in
formation of tar. Heating above 350°C causes condensation of
radical intermediates, cross-linked aromatic, and hydroaromatic
ring systems. This can be illustrated as follows:


9.9.9 Hydrolytic Degradation of Polymers at Elevated Temperatures
Hydrolytic degradation is only significant in
polymers with chain links that can react with water, such as
polyesters. A recent study reports on depolymerization of
poly(ethylene terephthalate) during processing, if the material is
not dried thoroughly, prior to melting [530]. The hydrolytic depolymerization of
poly(ethylene terephthalate) was carried out in a stirred batch
reactor at 235, 250, and 265°C above the polymer melting point and
under autogenous pressure. The solid products were mainly composed
of terephthalic acid. The liquid products were mainly composed of
ethylene glycol and a small amount of its dimer. Moreover, an
autocatalytic mechanism was detected. That indicates that some of
the hydrolytic depolymerization of PET is catalyzed by the
carboxylic acid groups produced during the reaction. The dependence
of the rate constant on the reaction temperature was correlated by
the Arrhenius equation [530].
9.9.10 Oxidative Degradation of Polymers
This section deals not only with oxidative
reactions of polymers at room temperature, but also with
thermo-oxidative degradation at elevated temperatures. Although
this type of degradation resembles photo-oxidative degradation, for
the purpose of maintaining clarity, the latter is discussed
separately in the next section.
The simplest oxidative reactions occur in
hydrocarbon polymers. This free-radical process generally follows
the same path as the oxidation of low molecular weight compounds.
The difference, however, is in the propagation of the reaction. In
oxidation of low-molecular weight hydrocarbons, each step of chain
propagation results in transfer of active center from on molecule
to another. In polymers, however, the probability of such a
transfer is low. Instead, the oxidation propagates along the
polymer backbone [520].
9.9.11 Oxidation of Chain-Growth Polymers
Polymers that lack double bonds, like
polyethylene, can be considered high molecular weight paraffin.
They are slow to oxidize in the absence of UV light, much like the
low molecular weight hydrocarbons. On the other hand, polymeric
materials with double bonds oxidize rapidly. Nevertheless, polymers
like polyethylene may oxidize rapidly as well when contaminated
with metallic ions because such ions catalyze the decomposition of
peroxides.
The chemical structure of the polyolefins
determines their susceptibility to oxidative degradation. Linear
polyethylene, in the absence of additives, is more resistant to
oxidation that polypropylene that oxidizes rather readily due to
the presence of labile tertiary hydrogens. It was demonstrated, for
instance, that the molecular weigh of polypropylene sheets in a
138°C oven can drop from 250,000 to approximately 10,000 in
3 h [522]. The process of
oxidation was shown to take place according to the following scheme
[522]:


Further oxidative degradation of fragments leads
to formation of a carboxylic acid, an ester, and a γ-lactone
[522]. It was also found that the
main oxidation products of polyethylene are an acid and a ketone.
On the other hand, polypropylene yields upon oxidation
approximately equal quantities of an acid, a ketone, an aldehyde,
an ester, and a γ-lactone [522].
In order for a polymer molecule to be attacked by
oxygen, it must come in contact with it. This means that oxygen
must be able to permeate into the material. Otherwise, all the
oxidation will occur only at the surface. It was shown that
oxidation occurs more readily in amorphous regions of the polymers
where permeation of oxygen is not hindered by the chains being
packed tightly together in the crystallites. That is only true, of
course, at temperatures below T m of the polymer.
Among the chain-growth polymer, oxidation of
polystyrene was
investigated thoroughly. It was found that the rate of oxygen
absorption and the number of chain scissions remain constant up to
a high degree of reaction. There is no evidence of cross-linking
under these circumstances [523].
During the degradation process, carbonyl groups accumulate in the
polymer [524]. Among the
degradation products were identified benzaldehyde [525] and a number of ketones [526].
The primary oxidation and chain scission process
in polystyrene at room temperature is as follows [527]:
the reaction can also proceed in this manner:


Continuation of the process reduces the polymer
to small fragments that include benzaldehyde and methyl phenyl
ketone [527].
When solid polystyrene is subjected to an
ozone attack, carbonyl,
peroxide, and carboxylic acid groups form on the surface of the
polymer [528]. The reaction rate
is proportional to the concentration of the ozone and the surface
of the sample. As a result of the ozone action, intramolecular
cross-linking takes place. The reaction mechanism of the ozone
attack on polystyrene can be shown as follows [528]:

In poly(vinyl
toluene), the initial oxidation steps consist of formations
of radicals at the tertiary carbon atoms as they do in polystyrene.
The radicals subsequently form peroxides that decompose into
ketones and aldehydes.
9.9.12 Oxidation of Step-Growth Polymers
Oxidation of step-growth polymers follows the
paths that are similar to oxidation reactions of organic molecules.
Thus, oxidation of poly(ethylene terephthalate) was shown to
proceed as follows [529]:


During thermal oxidation of polyamides, the N-vicinal methylene group is the
preferred side of attack [531].
The reactivity depends, apparently, on the conformation of the
amide group because an interaction takes place between the
π-electronic system of the carbonamide group and its C–H bonds. In
polycaprolactam, there exists a statistical distribution of both
gauche and trans-conformations. The N-vicinal methylene groups in the
trans-conformations to the
amide groups are much more reactive than gauche [531].
When the oxidation of a polyamide occurs through
an attack by NO2 or ozone, then the amide groups
themselves are subject to the attack [532]:

This appears to take place with hydrogen-bonded
amide groups as well. The reaction, however, is inhibited by
benzaldehyde or by benzoic acid. When the degradation does take
place, it occurs at random [532].
Thermo-oxidative degradation of polyoxyglycols is reported to occur
typically by a free-radical mechanism with the formation of
hydroperoxides. The decomposition of the peroxides leads to
formation of acids and carbonyl compounds [533]. A study of thermal oxidation of oligomers
(molecular weight approximately 1,900) of polyoxypropylene glycols
showed that at 70–120°C the end hydroxyl groups are not responsible
for degradation and that chain ruptures occur through decomposition
of the hydroperoxides [534]: Two
types of cleavage are possible. One can take place through the
carbon to carbon single bonds as follows:

The other type of cleavage can occur through the
oxygen–carbon bonds:

Both long and short molecular chains undergo
degradation in an identical manner. The secondary radicals that
form as a result of the decomposition of the alkoxy radical
accelerate the oxidation process. It transforms into a chain
process with an accompanying formation of various compounds. The
products are alcohols, aldehydes, and ketones. The aldehydes are
easily oxidized into peracids that degrade further into
radicals.
Thermal–oxidative degradation of polysulfones was studied at 280°C
[535]. The oxidation is a radical
chain reaction with the initiation consisting of hydrogen
abstraction from a methyl group:

The process then continues with formation of
alkyl peroxide radicals, isomerization, breakdown, and formation of
oxygen-containing compounds like aldehydes, and ketones.
In the oxidative pyrolysis of poly-p-xylene, the oxygen attacks
methylene groups first to form hydroperoxides [536]. This is followed by chain cleavage:

9.9.13 Photo-Degradation of Macromolecules
The quantum energies associated with sunlight in
the violet and near ultraviolet portions of the spectrum are of the
magnitude sufficient to rupture chemical bonds present in most
polymers [537]. The mechanism of
photo-degradation, however, is more complex than would be
visualized by the simple bond rupture, because photo-absorptions
are complicated by various factors, including crystallinity that
causes scattering of light. In addition, it was demonstrated that
little light is absorbed above the wavelength of 2,800 Å. Yet,
this wavelength represents the lower limit of sunlight reaching the
earth. On the other hand, presence of impurities that can act as
photosensitizes can markedly accelerate the degradation
process.
Polymers that are in current commercial use on a
large scale fall roughly into three categories, depending upon
their ability to withstand photo-degradation [538]:
1.
Polymers that are resistant to sunlight attacks
outdoors, like polyethylene and poly(methyl methacrylate).
2.
Moderately stable polymers, like poly(ethylene
terephthalate) and polycarbonate.
3.
Polymers that are unstable in sunlight and
require ultraviolet light stabilizer, like poly(vinyl chloride),
polypropylene, nylons, rubbers, and cellulose.
In addition to structural instability of some
polymer molecules in the ultraviolet light, degradation may also be
accelerated by chromophores that can form from oxidation during
preparation or processing. These compounds can act as excited
donors and transfer the energy to the polymers that may act as
acceptors. An intramolecular energy transfer may, actually, occur
within the same polymer molecule. This can take place between an
excited chromophore that is present in one segment of the chain and
an acceptor at another segment. Also, chromophores, like carbonyl
groups, can undergo Norrish Type I or a Norrish Type II
reaction.
Several studies were carried out on photolysis of polyethylene
[539, 540]. The primary process of photo-degradation,
however, is still being elucidated. Direct irradiation of pure
low-density polyethylene with ultraviolet light from a mercury lamp
results in formation of free-radicals that were identified by ESR.
They are alkyl radicals of the type:
~CH2–CH2• and
~CH2–CH•–CH2~ [539, 540].
The true absorption bands of polyethylene, however, are located at
wavelengths shorter than 200 nm. It is difficult, therefore,
to accept that the radiation from the lamp caused chain scissions,
because the output of the lamp is much longer in wavelength. It is
suspected, therefore, that the degradation is a result of
photo-oxidation. This is discussed in the next section.
A similar photo-degradation process is believed
to take place in polypropylene [541]. The formation of free-radicals is
ascribed to presence of oxidized molecules that form during
processing. The oxidation products are carbonyl compounds and
hydroperoxides [542]. The
photolysis of the carbonyl derivatives is as follows
[541]:

And the photolysis of chains carrying
hydroperoxide group can be shown in this manner [541]:

Photo-degradation of 1,2-polybutadiene was
studied on a film [543]. Among
the degradation products were found hydrogen, compounds with methyl
groups, vinyl groups, and cross-linked material. The following
mechanism of photo-initiation of the degradation reaction was
proposed [543]:

The mechanism of photo-degradation of
poly(p-methylstyrene) was
studied by several investigators [544–547]. A
gas evolution was observed during the irradiation with ultraviolet
light. This gas contains hydrogen as its major portion and methane,
ethane as the minor portions. There are also traces of styrene,
p-methylstyrene, and
toluene [101]. The gas evolution
is accompanied by cross-linking. The start of the process is
pictured as follows [547]:

The products react to further yield hydrogen,
cross-linked material, monomer, and cyclohexadiene.
Photolysis of poly(p-isopropyl styrene) with
ultraviolet light of 254 nm in vacuum at
10−6 mbar and room temperature yields hydrogen as
the main product and a small quantity of methane, ethane, and a
trace of propane [548].
Poly(methyl
methacrylate) depolymerizes at elevated temperature under
the influence of ultraviolet light of 259.7 nm [549]. Irradiation of polyacrylonitrile,
however, leads to chain scission at the acrylonitrile units. The
difference between thermal and ultraviolet light degradation of
polyacrylonitrile is principally in the different sites of
initiation and the fact that the reaction occurs at 160°C in the
presence of light and at 280°C in the darkness [550].
When a copolymer of methyl methacrylate and
n-butyl acrylate is
irradiated in vacuum at elevated temperature, like 165°C with light
of 253.7 nm, the gaseous products are minor [551]. These are hydrogen, carbon monoxide, and
methane. The liquid products are predominantly methyl methacrylate,
n-butyl acrylate,
n-butyl alcohol, and
n-butyraldehyde. The
degradation is explained by Grassie and coworkers [551] in terms of a radical process that is
initiated by scission of pendant acrylate units. The propagation is
actually a combination of depropagation of the chains and intra-
and intermolecular chain transfer process. The relative importance
of each phase of depropagation reaction depends upon the
composition of the copolymer [551].
Poly(vinyl
acetate) was shown to undergo simultaneous cross-linking and
chain scission when irradiated in vacuum with ultraviolet light
[552]. The following mechanism
was proposed as the route to chain cleavage [552]:

Also, intramolecular hydrogen abstraction is
assumed to occur through a seven-membered ring transition state.
This can provide a route to chain scission:

In subsequent studies of photo-degradation of
poly(vinyl acetate) and poly(vinyl propionate) in vacuum, it was
found that acids, aldehydes, and carbon dioxide were the principal
products [553]. Two mechanisms,
one involving hydrogen abstractions, were postulated. One of them
takes place by abstraction by acyl radicals formed through Norrish
Type I cleavages and the other by intramolecular hydrogen
abstraction by excited carbonyl groups followed by fragmentation
[553]:


A second, perhaps more viable mechanism, was
pictured for the production of aldehydes involving hydrogen
abstraction as follows [553]:

The main chain scissions are envisioned as
involving fragmentation of the polymer alkoxy radicals that form
[553]:

When cast poly(vinyl chloride) films are
irradiated by ultraviolet light at various intensities and
temperatures under a nitrogen atmosphere, dehydrochlorination
reaction takes place [554]. This
reaction occurs in two parts. During the first hour, the reaction
is dependent on the intensity of light and the temperature, but
after that it becomes independent of these two parameters.
When poly(vinyl chloride) films are irradiated in
the presence of benzophenone, the initiation is a result of
hydrogen abstraction from the polymer by the excited triplet of the
aromatic ketone [555]: This is
followed by degradation that takes place by a chain mechanism:

In the process, the chlorine radicals abstract
hydrogens and form HCl. The quantum yield of hydrogen chloride from
this reaction is ΦHCl = 0.11 [556]. This indicates that only 1 in every 100
photons that are absorbed causes the dehydrochlorination and is
followed by formation of polyenes. It led to the conclusion that a
photochemical process must take place between the polyene sequences
and HCl [556, 557].
When poly(phenyl
vinyl ketone) is irradiated by ultraviolet light of
365 nm, it degrades and exhibits a loss of molecular weight
[558]. No benzaldehyde was found
among the products. That excludes a Norrish Type I reaction and
indicates that the polymer cleaves by a Norrish Type II mechanism
[558].
Day and Wiles have pictured the primary steps of
photochemical decomposition of poly (ethylene terephthalate) as
follows [559]:

The mechanism of formation of carboxylic acid
groups at the terminal ends of the degradation products was
explained by a Norrish Type II reaction [559]:

In addition to the above, photo-Fries
rearrangements occur in various aromatic polyester films upon
ultraviolet light irradiation [560, 561].
Also, photo-Fries rearrangements are believed to occur in
polycarbonate resins as well [562]. This can be illustrated as follows:

The same is true of aromatic polyamides where
photo-Fries rearrangements are done according to the following
mechanism [563]:

Photolytic decomposition of polycarbonate films was shown to
produce products that are also consistent with the photo-Fries
reaction. These are salicylic acid and bisphenol type species
[580]. This is in agreement with
earlier studies that showed that a variety of processes, including
rearrangements (photo-Fries) and photo-oxidation, can occur when
bisphenol A-based polycarbonate is photolyzed [581, 582].
9.9.14 Photo-Oxidative Degradations of Polymers
Singlet oxygen, the lowest excited state of
molecular oxygen, 1Δg O2, is known
to cause rapid degradation of polymers. Singlet oxygen forms by
several processes, including one of two mechanisms: (1) through
photosensitization by some impurity; (2) through energy transfer
from an excited triplet state of a chromophore.
Polyolefins that are unprotected from weathering
react very readily with oxygen in the presence of sunlight even at
room temperature. This leads to loss of molecular weight and often
embrittlement. The mechanism of photo-oxidation of polyethylene was
demonstrated to involve carbonyl groups that developed in the
polymer via oxidation during processing [564, 565].
These carbonyl groups act as chromophores and their triplet states
(n → π*) are quenched by molecular oxygen in
the ground triplet state. Electronically excited singlet oxygen
molecules form as a result [564].

The Norrish Type II reaction then takes place
with a cleavage of carbon to carbon bonds, formation of olefins
and, subsequently, formation of hydroperoxides:

The hydroperoxides apparently have no photo
inductive effect on the overall oxidation of polyethylene
[566]. Photolysis of
hydroperoxides, however, initiates new vinylidine oxidation.
It was shown that the kinetics of photo-oxidation
of polyethylene is characterized by the superposition of two
phenomena. The first corresponds to an exponential increase in the
concentration of the carbonyl groups with time and is observed when
the kinetics are controlled by the diffusion of oxygen. The second
one is not controlled by diffusion, but corresponds to a linear
increase of the carbonyl concentration with time and takes place in
degraded samples. This is explained in terms of chain rupture in
the amorphous regions of the polymer, allowing free access of
oxygen [567].
Commercial polypropylene can photo-degrade
rapidly. This is due to a presence of chromophore groups that form
during polymerization. In addition, chromophores may also form
through oxidation by singlet oxygen during extrusion in air. At
temperatures of 250–300°C, light of 300 nm will initiate
oxidation by atmospheric oxygen [568]. The hydroperoxide groups that form are
not stable and decompose to alkoxy radicals to form ketones on the
backbones and at the ends of the chains. These ketones in turn
undergo Norrish Type I and Type II reactions when irradiated with
ultraviolet light [568]:

Photo-oxidation of poly(vinyl chloride) appears to be
greatly enhanced by imperfections in the polymer structure that
form during polymerization and processing. Formation of peroxide
and carbonyl groups is very difficult to prevent. The initiation of
photo-oxidation probably results from the free-radical that forms
as these groups decompose upon irradiation [568]. In addition, double bonds that form from
loss of HCl will sensitize formation of free radicals. These
radicals are further oxidized and also promote cross-linking that
results in gels.
Photo-oxidation of cis-polybutadiene was also shown to
involve singlet oxygen [570].
Attacks by singlet oxygen on double bonds with formation of allylic
hydroperoxides and shifts of these double bonds according to the
ene reaction were confirmed
by several other studies [571].
On the other hand, a study of model compounds for photo-oxidation
of polyisoprene failed to show formation of endo peroxides by
1,4-cycloaddition [572].
Aromatic polyurethanes exhibit a strong tendency
to yellow in sunlight. This may be accounted for by two mechanisms
based on formation of different photoproducts. The first one was
found to be a photo-Fries reaction caused by a short wave length
light [572]. The second one,
caused by longer wave length light, is formation of quinonoid
structures [573]:

Photo-oxidation of Nylon 6 was also investigated
[574]
The subject of photo-oxidation of silicones is seldom
raised because these polymers are known for their stability to
oxidation. Nevertheless, an investigation of photo-oxidative
degradation of these materials showed that these materials can be
affected by combined attacks of oxygen and ultraviolet irradiation
[585]. The groups that are
affected are the functional groups that are more fragile than the
backbones. Thus, the silicon hydride group was found to be readily
photo-oxidized, perhaps due to its low bond energy. The reaction
can be illustrated as follows [585]:

The main process that occurs in silicones that
contain vinyl groups is photo-scission. Also, the vinyl groups are
photo-oxidized into silanols, carboxylic acids, and esters. Part of
the silanols are condensed into high molecular weight silicones
[585].
9.9.15 Degradation of Polymeric Materials by Ionizing Radiation
Ionizing radiation is not absorbed selectively
like the ultraviolet. Instead, the absorption of gamma rays is a
function of electron density in the path of the radiation. After a
random ray strikes an atom, however, and energy transfer occurs,
free-radicals may form and subsequent reactions, like oxidation and
chain reactions, may proceed in the same manner as after UV light
irradiation. In addition to creating free-radicals, ionizing
radiation also gives rise to ions, electrons, and excited molecules
in the medium. This is followed by reactions, like, chain scission,
formation of unsaturation, and cross-linking that is accompanied by
formation of volatile products.
Due to the high energy of ionizing radiation, if
the irradiation dose is sufficiently high, all polymers will
degrade to form low molecular weight fragments. The difference is
in the rates of degradation and in the amounts of
cross-linking.
Among vinyl polymers, those that tend to unzip
and yield large quantities of monomers in thermal degradation are
also likely to loose molecular weight rapidly upon irradiation. On
the other hand, those vinyl polymers that yield low quantities of
monomers during pyrolysis tend to cross-link instead. Also, it was
observed that polymeric materials that exhibit high heats of
polymerization during their formation tend to cross-link, while
those that do not tend to degrade [575].
It was reported back in 1952 that polyethylene could be cross-linked by
irradiation in a controlled manner [577]. The reaction is accompanied by evolution
of considerable quantities of volatile gases that consist mainly of
hydrogen. The portion of hydrogen that is released, however, is
lower and the amount of volatile hydrocarbons is greater from
polyethylene that is branched [578]. There appears to be an inverse
relationship between the amount of branching and the quantity of
hydrogen gas that is released [578].
Polypropylene was reported to
cross-link under irradiation [579]. At the same time, there is a large
evolution of gas. If the irradiation is carried out in air, there
is marked degradation even at fairly low doses.
When polyisobutylene is irradiated, the molecular
weight decreases rapidly until a viscous, low molecular weight
liquid remains. At the same time, a mixture of hydrogen and methane
(95% of total gas) is released [582].
Generally, aromatic structures are more stable to
ionizing radiation than are the aliphatic ones. Accordingly,
polystyrene is more resistant than polyethylene [575]. In addition, cross-linking predominates
over chain scission.
The radiolysis of poly(methyl methacrylate) results in a
rapid loss of molecular weight. This degradation increases with the
intensity of the radiation [579].
The volatile products that form are hydrogen, carbon dioxide,
carbon monoxide, methane, propane, and methyl methacrylate monomer.
This varies with the temperature and the type of ionizing radiation
that the polymer is exposed to.
Polytetrafluoroethylene is extremely
sensitive to radiation and exhibits marked damage to its mechanical
properties. It was shown that fluoride ions evolve from the polymer
not only during irradiation, but also for long periods afterward
[134]. The molecular weight of
the polymer drops steadily with each dose. Many double bonds form
in the polymer [580].
9.10 Review Questions
9.10.1 Section 9.1
1.
What are the necessary conditions for a fair
comparison of the reactivity of functional groups on macromolecules
with those on small molecules.
2.
When is unequal reactivity observed between large
and small molecules?
3.
Discuss the effect of diffusion-controlled
reactions on reactivity of macromolecules.
4.
Discuss how paired group and neighboring group
effects influence random irreversible reactions.
5.
Discuss reactions that favor large
molecules.
9.10.2 Section 9.2
1.
Explain, showing chemical equations, how
hydrochlorination of natural rubber is often accompanied by
cyclization.
2.
Discuss chlorination of natural rubber. How can
natural rubber and polybutadiene be brominated?
3.
Discuss with the aid of chemical equations
hydrogenation of 1,4-polybutadiene.
4.
How can polyisoprene and polybutadiene be
modified by additions of carbenes? Explain and discuss, showing the
structures of the starting materials and the products.
5.
Illustrate the Prins reactions of rubber with
formaldehyde and with glyoxal. How can this reaction be carried out
in the solid phase?
6.
Discuss polar additions to unsaturated polymers.
Give examples Include the ene reaction.
7.
How can the Ritter reactions be carried out on
isoprenes?
8.
How can the Diels-Alder reactions are carried out
with unsaturated polymers. Use chemical equations to
illustrate.
9.
Explain how polybutadiene and isoprene can be
epoxidized, giving reagents and showing all by-products.
9.10.3 Section 9.3
1.
Show how, and explain why, cis-polybutadiene rearranges to
trans as a result of
irradiation by gamma rays or ultraviolet light. Also, show what
happens to polyisoprene when it is irradiated in the ultraviolet
light.
2.
Discuss with the aid of chemical equations the
cyclization reactions of rubber.
3.
How are polyacrylonitrile fibers converted to
graphite-like fibers. Explain, showing all the steps in thermal
cyclization of polyacrylonitrile.
4.
Discuss migration of double bonds in
polymers.
5.
How does poly(4,4′-diphenylpropane isophthalate)
rearrange upon irradiation with ultraviolet light?
6.
Give examples of the Schmidt and Beckmann
rearrangements.
9.10.4 Section 9.4
1.
Discuss with the help of chemical equations the
fluorination reactions of polyethylene.
2.
Discuss chlorination of polyethylene and
polypropylene.
3.
How is chlorosulfonation of polyethylene carried
out industrially. Explain and write the chemical equations for the
reactions. How is the product used?
4.
Discuss chlorination of poly(vinyl
chloride).
5.
Discuss chlorination of poly(vinylidine chloride)
and poly(vinyl fluoride).
6.
Describe the reactions of sodium azide with
poly(vinyl chloride) and the subsequent reactions of the azide
group. Use chemical equations.
7.
Describe reactions of poly(vinyl chloride) with
sulfur compounds.
8.
What are the products of reactions of poly(vinyl
chloride) with aniline? Show and explain.
9.
Explain how carbanionic centers can be formed on
the backbones of poly(vinyl chloride) molecules. Show what
subsequent reactions can take place.
10.
What happens when poly(vinyl chloride) is treated
with organolithium compounds. Explain with the help of chemical
equations.
11.
Discuss photo chlorination of polystyrene.
12.
Discuss chloromethylation reactions of
polystyrene and its copolymers, showing all chemical
reactions.
13.
Discuss various reactions of chloromethylated
polystyrene with the help of chemical equations.
14.
Show examples of the Friedel-Craft reactions of
polystyrene and chloromethylated polystyrene.
15.
Discuss sulfonation of polystyrene.
16.
Discuss nitration of polystyrene, reduction of
nitropolystyrene, and the subsequent diazotization reaction.
17.
Discuss matalation of polystyrene and subsequent
reactions.
18.
Discuss reduction reactions of the polymers of
methacrylic and acrylic esters with metal hydrides. What solvents
give optimum conditions?
19.
Discuss nucleophilic substitution reactions of
poly(methyl methacrylate); the Arndt-Eister reactions of
poly(methacryloyl chloride) and the Curtius and Lossen
rearrangements of poly(acryloyl chloride)
20.
Show a Diels-Alder reaction of poly(furfuryl
methacrylate) with maleic anhydride.
21.
What are the industrial processes for preparation
of poly(vinyl butyral)? Describe the process, showing all chemical
reactions.
22.
How can the Schotten-Baumann esterifications of
poly(vinyl alcohol) are carried out? Explain.
9.10.5 Section 9.5
1.
Discuss the vulcanization reactions of rubbers.
Show all the chemical reactions.
2.
Discuss ionizing radiation cross-linking of
polymers.
3.
Show the reaction of photocross-linking of
poly(vinyl cinnamate) and explain the mechanism.
4.
Show the reactions of photocross-linking of
polymers bearing azide groups.
5.
Draw examples of polymers with pendant benzoin
and furoin groups.
6.
Explain the mechanism of photocross-linking by
2π + 2π type dimerization. Give examples.
7.
Discuss typical ultraviolet light-curable coating
compositions by free radical and by cationic mechanisms.
8.
Why are photoinitiators preferred to
photosensitizers in UV-curable materials?
9.
Discuss with the help of chemical equations the
mechanism of photoreduction of aromatic ketone photoinitiators
hydrogen donors and by electron donors.
10.
Discuss photoinitiators for photocross-linking by
cationic mechanisms, showing all reactions and products.
11.
Discuss cross-linking reactions by electron
beams.
9.10.6 Section 9.6
1.
What is meant by microporous and macroreticular
supports.
2.
Discuss some applications of polymeric reagents.
Show examples.
3.
How are enzymes immobilized? Discuss, showing
chemical reactions.
4.
Discuss nonenzymatic catalysts.
9.10.7 Section 9.7
1.
What is the mechanism of free-radical graft
copolymer formation by chain-transferring technique. Explain by
examples and discuss advantages and disadvantages.
2.
Discuss free-radical grafting reactions to
polymers with double bonds. Give examples and show reactions.
3.
How can macromonomers be used to form graft
copolymers? Give several examples.
4.
How can polymerizations be initiated from the
backbones of the polymers? Explain and give examples.
5.
Discuss photochemical preparation of graft
copolymers.
6.
How is high-energy radiation used to form graft
copolymers?
7.
Discuss formations of graft copolymers by ionic
chain-growth and step-growth polymerizations.
9.10.8 Section 9.8
1.
Discuss advantages of block copolymers over
polymer blends.
2.
How can block copolyesters be formed? Explain
with chemical equations.
3.
How can block copolyamides be formed? Explain as
in Question 2.
4.
How can polyurethane-polyamide and
polyamide-polyester block copolymers be formed? Explain and show
chemical reactions.
5.
What are polyurethane ionomers? How are they
prepared?
6.
Describe the technique for preparation of block
copolymers of poly(α-olefins).
7.
How do block copolymers form in simultaneous free
radical and ionic chain-growth polymerizations? Explain and give
examples.
8.
Discuss formations of block copolymers by
homogeneous ionic copolymerization.
9.
How can block copolymers form by mechanochemical
techniques?
9.10.9 Section 9.9
1.
What is meant by polymer degradation and what are
the various causes?
9.10.10 Section 9.10
1.
What is the common technique for studying thermal
degradation of polymers?
2.
Explain the two thermal degradation processes in
chain-growth polymers?
3.
Describe the thermal degradation of
polyolefins.
4.
Describe the thermal degradation of
polybutadienes.
5.
What is the thermal degradation process in
polystyrene and polystyrene-like polymers?
6.
How do the methacrylic and acrylic esters degrade
thermally?
7.
What happens when polyacrylonitrile is heated?
Describe
8.
Discuss the thermal degradation of poly(vinyl
chloride)
9.
Discuss the thermal degradation of fluorocarbon
polymers
10.
Discuss the thermal degradation of poly(vinyl
acetate)
9.10.11 Section 9.11
1.
Can the step-growth polymers degrade in the same
manner as the chain-growth polymers?
2.
How do the polyoxides degrade thermally?
3.
Discuss the thermal degradation of
polyesters.
4.
Discuss the thermal degradation of
polyamides.
5.
Discuss the thermal degradation of the epoxy
resins
6.
How do the polyimides and the polyquinoxalines
degrade thermally?
7.
Discuss the thermal degradation of
polysulfones
8.
How do cellulosic materials degrade
thermally?
9.10.12 Section 9.12
1.
How does hydrolytic degradation of polymers take
place?
9.10.13 Section 9.13
1.
Describe oxidative degradation in hydrocarbon
polymers
2.
Describe oxidative degradation of
polystyrene
3.
Describe oxidative degradation of poly(ethylene
terephthalate)
4.
How do the polyamides degrade oxidatively?
9.10.14 Section 9.14
1.
What is meant by photo-degradation of
polymers?
2.
Describe photolysis of polyethylene.
3.
Describe photo-degradation of polybutadiene and
poly(p-methylstyrene)
4.
How does poly(methyl methacrylate) degrade at
elevated temperature under the influence of ultraviolet
light?
5.
Describe photo-degradation of poly(vinyl acetate)
and poly(vinyl chloride)
6.
What is the mechanism of photo-degradation of
aromatic polyesters?
9.10.15 Section 9.15
1.
What causes photo-oxidative degradation and how
does it take place in polyolefins?
2.
How does commercial polypropylene degrade
photo-oxidatively?
9.10.16 Section 9.16
1.
If you had to select polymeric materials for use
near a nuclear reactor, what types of materials would you select?
Explain.
References
1.
P.J. Flory, Principles of Polymer Chemistry,
Cornell University Press, Ithaca, New York, 1953, Chapt.II.
2.
T. Alfrey Jr., Chemical Reactions of Polymers, (E.M.
Fetters,ed)., Wiley-Interscience, New York, 1964, Capt. IA.
3.
4.
5.
H. Morawetz, J . Macromol. Sci.-Chem., A13, 311 (1979)
6.
M. Kamachi, Makromol. Chem., Suppl ., 14, 17 (1985)
7.
A. Okamoto, Y. Shimanuki, I,
Mita, Eur. Polym. J.,
18, 545 (1982); ibid, 19, 341 (1983)
8.
K.F. O’Driscoll and H.K.
Mahabadi, J. Polymer Sci., Chem.
Ed., 11, 869
(1976)
9.
10.
M. Kamachi, Adv. Polym. Sci., 38, 56 (1981)
11.
P.J. Flory, J. Am. Chem. Soc., 61, 1518 (1939); T.H.K. Barron and E.A.
Boucher, Trans. Faraday
Soc., 65 (12), 3301
(1969)
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
N. Plate, Vysokomol. Soyed., A10 (12), 2650 (1968)
23.
W.H. Daly, J. Macromol. Sci.-Chem., A22, 713 (1985)
24.
A.R. Mathieson and J.V.
McLaren, J. Polymer Sci.,
A,3, 2555 (1965)
25.
J.C. Layte and M. Mandel,
J. Polymer Sci.,
A,2, 1879 (1964)
26.
F. Fleming and G.R. Marshal,
Tetrahedron Lett., 2403
(1970)
27.
28.
E.Bondi, M. Fidkin, and A.
Patchornik, Israel J.
Chem., 6, 22
(1968)
29.
O. Fuchs, Angew. Chem., Intern. Ed., 7 (5), 394 (1968)
30.
M. Tsuda, Makromol. Chem., 72, 174 183 (1964)
31.
D.M. Jones, Adv. Carbohydrate Chem., 19, 229 (1964)
32.
E.M. Fetters, Crystalline Olefin Polymers, Part II,
(R.A. Raff and K.W. Doaks, eds.), Interscience, New York,
1964
33.
S. Tazuke, Makromol. Chem., Suppl., 14, 145 (1985)
34.
35.
A.D. Roberts, (ed.)
Natural Rubber Science and
Technology, Oxford University Press, New York, 1988
36.
H. Staudinger and H.
Staidinger, J. Prakt.
Chem., 162, 148
(1943); Rubber Chem.
Technology, 17, 15
(1944)
37.
C.S. Ramakrishnan, D.
Raghunath, and J.B. Ponde, Trans.
Inst. Rubber Ind., 29,
190 (1953); ibid
., 30, 129 (1954)
38.
39.
C.F. Broomfield,
J. Chem. Soc., 1944, 114; Rubber Chem. Technol., 17, 759 (1944)
40.
C.S. Ramakrishnan, D.
Raghunath, and J.B. Ponde, Trans.
Instit. Rubber Ind., 30, 129 (1954); Rubber Chem. Technol., 28, 598 (1955)
41.
I.A.Tutorskii, L.V.
Sokolova, and B.A. Dogadkin, Vysokomol. soyed., A13 (4), 952 (1971)
42.
43.
G. Dall’Asta, P. Meneghini,
and U. Gennaro, Makromol.
Chem, 154, 279
(1972)
44.
G. Dall’Asta, P. Meneghini,
I.W. Bassi, and U. Gennaro, Makromol. Chem., 165, 83 (1973)
45.
46.
F. Stelzer, K. Hummel, C.
Graimann, J. Hobisch, M.G. Martl, Makromol. Chem., 188, 1795 (1987)CrossRef
47.
K. Hummel, J. Mol. Cat., 28, 381 (1987); K. Hummel, Pure Appl. Chem., 54, 351 (1982)
48.
Y.G. Garbachev, K.A.
Garbatova, O.N. Belyatskaya, and V.E. Gul, Vysokomol. Soed., 7, 1645 (1965)
49.
50.
E. Ceausescu, R. Bordeianu,
E. Buzdugan, I. Cerchez, P. Ghioca and R. Stancu, J. Macromol. Sci. - Chem., A22, 803 (1985)
51.
M. Bruzzone and A.
Carbonaro, J. Polymer Sci., Chem.
Ed ., 23, 139 (1985)
52.
53.
54.
H. Bradbury and M.C. Semale
Perera, Brit. Polymer J.,
18, 127 (1986); A. Konietzny
and U. Bietham, Angew
. Makromol. Chem., 74, 61 (1978)
55.
56.
57.
58.
59.
H.J. Prins, Chem. Weekblad, 17, 932 (1917); ibid., 16, 1510 (1919)
60.
F. Kirchhoff, Chem. Ztg., 47, 513 (1923)
61.
J. McGavack, U.S. Patent #
1,640,363 (Aug. 30, 1927)
62.
D.F. Twiss and F.A. Jones,
Brit. Patent # 348 303 (March 24, 1930)
63.
D.F. Twiss and F.A. Jones,
Brit. Patent # 523 734 (July 22, 1940)
64.
G.H. Latham, U.S. Patent #
2,417,424 (March 18, 1947)
65.
S. Hirano and R. Oda,
J. Soc. Chem. Ind. (Japan),
47, 833 (1944); from
Chem. Abstr., 42, 7982 (1948)
66.
D.F. Twiss and F.A. Jones,
U.S. Patent # 1,915,808 (June 27, 1933)
67.
C. Harris, Ber., 37, 2708 (1904)
68.
C. Pinazzi and R. Pautrat,
Compt. Rend., 254, 1997 (1962); from Chem. Abstr., 56, 15647 (1962)
69.
C. Pinazzi and R. Pautrat,
Rev. Gen. Coutchuc,
39, 799 (1962)
70.
C. Pinazzi, R. Pautrat, and
R. Cheritat, Compt. rend
., 256, 2390, 2607 (1963)
71.
S.K. Lee. Am. Chem. Soc. Polymer Preprints,
8, 700 (1967)
72.
73.
G. Caraculacu and I.
Zugravescu, J. Polymer Sci.,
Polymer Letters, 6,
451 (1968)
74.
C.G. Gubelein, J. Macromol. Sci. - Chem.,
A5, 433 (1971)
75.
76.
K.W. Leong and G.B. Butler,
J. Macromol. Sci. - Chem
., A14,287 (1980); C,S,L. Baker and D.
Bernard, Am. Chem
. Soc., Polymer Preprints, 26 (2), 29 (1985)
77.
D.A. Bansleben and F.
Jachimowicz, Am. Chem. Soc.,
Polymer Preprints, 26
(1), 106 (1985)
78.
79.
J.W. Mench and B. Fulkerson,
Ind. and End. Chem.,
7, 2 (1968)
80.
J.P. Kennedy and R.A. Smith
J. Polymer Sci., Chem. Ed.,
18, 1523 (1980)
81.
B. Ivan, J.P. Kennedy and
V.S.C. Chang, J.
Polymer Sci., Chem. Ed
.. 18, 3177 (1980)
82.
J.P. Kennedy, V.S.C. Chang,
and W.P. Francik, J. Polymer Sci.,
Chem. Ed., 20, 2809
(1982)
83.
84.
E. Ceausescu, S. Bittman, V.
Fieroiu, E. Badea, E. Gruber, A. Ciupitoiu, and V. Apostol,
J. Macroml. Sci.-Chem.,
A22 (5-7), 525
(1985)CrossRef
85.
86.
87.
H.L. Cohen, J. Polymer Sci., Chem. Ed.,
22, 2293 (1984)
88.
G.J. Carlson, J.R. Skinner,
C.W. Smith, and C.H. Wilcoxen Jr.,U.S. Patent # 2,870,171 (May 6,
1958)
89.
C.W. Smith and G.P. Payne,
U.S. Patent # 2,786,854 (March 26, 1957)
90.
C.M. Gable, U.S. Patent #
2,870,171 (Jan 20, 1959)
91.
W.C. Smith, Rec. Trav. Chim., 29, 686 (1930); from Chem. Abstr., 24, 4261 (1930)
92.
F.P. Greenspan and R.E.
Light Jr., U.S. Patent # 2,829,130 (April 1958)
93.
S. Tocker, U.S. Patent #
3,043,818 (July 1962)
94.
R.W. Rees, U.S. Patent #
3,050,507 (August 1962)
95.
H. Lee and K. Neville,
Handbook of Epoxy Resins,
Mc Grow-Hill, New York, 1967
96.
W.G. Potter, Epoxy Resins, Springer-Verlag, New
York, 1970
97.
C.A. May, (ed.) Epoxy Resins: Chemistry and Technology,
2nd. ed., Dekker, New York, 1988
98.
A.F. Chadwick, D.O. Barlow,
A.A. D’Adieco, and J.W. Wallace, J. Am. Oil Chemists Soc ., 35, 355 (1958)CrossRef
99.
K. Meyersen and J.Y.C. Wang,
J. Polymer Sci ., A,5, 725 (1967)
100.
H.S. Makowski, M. Lynn, and
D.H. Rotenberg, J. Macromol.
Sci.-Chem., A4, 1563
(1980)
101.
102.
103.
104.
105.
106.
M.A. Golub and C.L.
Stephens, J. Polymer Sci.,
C-6 (16), 765 (1967)
107.
108.
109.
M.A. Golub and J. Hiller,
Tetrahedron Letters, 2137
(1963)
110.
R.J. Angelo, M.L. Wallach,
and R.M. Ikeda, Am. Chem. Soc.
Polymer Preprints, 8
(1), 221 (1967)
111.
M.A. Golub and J. Heller,
J. Polymer Sci., Polymer
Letters, 4, 469
(1966)
112.
113.
N.G. Gaylord, I. Kossler,
M. Stolka, and J. Vodehnal, J.
Polymer Sci., A-1,2,
3969 (1964)
114.
M. Stolka, J. Vodehnal, and
I. Kossler, J. Polymer
Sci., A-1,2, 3987
(1964)
115.
I. Kossler, J. Vodehnal,
and M. Stolka, J Polymer
Sci., A-1,3, 2081
(1965)
116.
P. Ehrburger and J.B.
Donnet, High Technology
Rubbers, Part A, (M. Lewin and J. Preston, eds.), Dekker,
New York, 1985; G. Henrici-Olive’ and S. Olive’, Adv. Polym. Sci., 51, 1 (1983); E. Fitzer, Angew. Chem., Intern. Ed ., 19, 375 (1980)
117.
118.
119.
120.
121.
G. Salmon and Chr. van der
Schee, J. Polymer Sci.,
14, 287 (1954)
122.
M.A. Golub and C.L.
Stephens, J. Polymer Sci.,
A-1,6, 763 (1968)
123.
M.A. Golub, J. Polymer Sci., Polymer Letters,
12, 615 (1974); ibid., 12, 295 (1974)
124.
M.A. Golub and J.L.
Rosenberg, J. Polymer Sci., Chem.
Ed., 18, 2548
(1980)
125.
M.A. Golub, J. Polymer Sci., Chem. Ed.,
19, 1073 (1981)
126.
R.W. Lenz, K. Ohata, and J.
Funt, J. Polymer Sci., Chem.
Ed., 11, 2273
(1973)
127.
S.B. Maerov, J. Polymer Sci., A-1,3, 487 (1963)
128.
F. Millich and R. Sinclair,
Chem. and Eng. News, p.30,
(June 26, 1967)
129.
130.
131.
J.A. Banchette and J.D.
Cotman Jr., J. Org. Chem.,
23, 117 (1958)
132.
133.
H. Morawetz and G.
Gordimer, J. Am. Chem. Soc
., 92, 7532 (1970)
134.
135.
A.J. Rudge, Brit. Patent #
710,523 (June 16, 1954)
136.
I. Vogt and W. Krings,
German Patent # 1,086,891 (Aug. 11, 1960)
137.
M. Okada and K. Makuuchi,
Kagyo Kagaku Zashi,
73, 1211 (1970) (from a
private English translation)
138.
139.
140.
J.R. Myles and P.J. Garner,
U.S. Patent # 2,422,919 (June 24, 1947)
141.
W.N. Baxter, U.S. Patent #
2,849,431 (Aug. 26, 1958)
142.
L.B. Krentsel, A.D.
Litmanovich, I.V. Patsukhova, and V.A. Agasandyan, Vysokomol. soyed ., A13, 2489 (1971)
143.
M. Goren, Polim. Vehomerim Plast., #1, 8
(1971)
144.
R.S. Taylor, U.S. Patent #
2,592,763 (April 20, 1952)
145.
H.W. Thompson and P.
Torkington, Trns. Faraday
Soc., 41, 254
(1945)
146.
147.
148.
G.D. Jones, Chemical Reactions of Polymers, (E.M.
Fetters, ed.),Wiley- Interscience, New York, 1964
149.
Brit. Patent # 811,848
(April 15, 1959); from Chem.
Abstr., 53, 23100f
(1951)
150.
H. Noeske and O. Roeleau,
U.S. Patent # 2, 889,259 (June 2, 1959)
151.
A. McAlevy, U.S. Patent #
2,405,971 (August 20, 1946)
152.
R.E. Brooks, E.D. Straom
and A. McAlevy, Rubber
World, 127, 791
(1953)
153.
G. Natta, G. Mazzanti, and
M. Buzzoni, Ital. Patent # 563,508 (April 24, 1956)
154.
155.
P. Bertican, J.J. Bejat,
and G. Vallet, J. Chim. Phys,
Physicochim. Biol., 67, 164 (1979)
156.
157.
158.
C. Decker, J. Polymer Sci., Polymer Letters,
25, 5 (1987)
159.
160.
M. Kolinsky, D.
Doskocilova, B. Schneider, and J. Stokr, J. Polymer Sci., A-1,9, 791 (1971)
161.
Backskai, L.P. Lindeman,
and J.W. Adams. J. Polymer
Sci., A-1,9, 991
(1971)
162.
G.G. Scherer, P. Pfluger,
H. Braun, J. Klein, H. Weddecke, Makromol. Chem., Rapid Commun.,
5, 611 (1984)
163.
M. Takeishi and M. Okawara,
J. Polymer Sci., Polymer Letters, 7, 201 (1969)
164.
M. Takeishi and M. Okawara,
J. Polymer Sci., Polymer
Letters, 8, 829
(1970)
165.
M. Okawara, T. Endo, and Y.
Kurusu, Progress of Polymer Sci.,
Japan, (K. Imahori, ed.) Vol. 4, p. 105, 1972
166.
K. Mori and Y. Nakamura,
J. Polymer Sci., Polymer
Letters, 9, 547
(1971)
167.
G. Levin, Makromol Chem., Rapid Commun.,
5, 513 (1984)
168.
M.K. Naqvi and P. Josph,
Polymer Communication,
27, 8 (1986)
169.
Z. Wolkober and I. Varga,
J. Polymer Sci ., C (16), 3059 (1967)
170.
H. Calvayrac and J. Gole,
Bull. Soc. Chim. France,
1986, 1076
171.
E. Roman, G. Valenzuela, L.
Gargallo, D. Radic, J. Polymer
Sci., 21 2057
(1983)
172.
173.
174.
M. Okawara, Am. Chem.Soc. Coatings and Plastics Chemistry
Preprints, 40, 39
(1979)
175.
Y. Nakamura, K. Mori, and
M. Kaneda, Nippon Kagaku
Kaishi, 1620 (1976); from ref. # 164.
176.
177.
R.A. Haldon and J.N. Hay,
J. Polymer Sci.,
A-1,5, 2295 (1967)
178.
179.
R. Tarascon, M. Hartney,
M.J. Bowden, Am. Chem. Soc.
Polymer Preprints, 25
(1), 289 (1984)
180.
T. Saegusa and R. Oda,
Bull. Inst. Chem. Research Kyoto
Univ., 33, 126 (1955);
from Chem. Abstr.,
50, 1357 (1956)
181.
F. Helfferich, Ion Exchange, McGraw-Hill, N.Y., 1962;
G.A. Olah and W.S. Tolgyesi, Friedel-Craft and Related Reactions,
(G.A. Olah, ed.), Wiley-Interscience, New York, 1964; W.E. Wright,
E.G. Topikar, and S.A. Svejda, Macromolecules, 24, 5879 (1991)
182.
183.
184.
R. Hauptman, F. Wolf, and
D. Warnecke, Plaste Kaut.,
18, 330 (1971)
185.
E.E. Ergojin, S.P. Rafikov,
and B.A. Muhidtina, Izv. Akad.
Nauk Kaz. S.S.R., 19
(16), 49 (1969)
186.
G.Z. Esipov, L.A.
Derevyanki, V.P. Markidin, R.G. Rakhuba, and K.L. Poplavskii,
Vysokomol. soyed.,
B12, 274 (1970)
187.
188.
S.R. Rafikov, G.N.
Chelnokova, and G.M. Dzhilkibayeva, Vysokomol. soyed., B10, 329 (1968)
189.
L. Galeazzi, German Pat. #
2,455,946 (June 1975)
190.
W.H. Daly, J. Macromol. Sci., Chem., A22, 713 (1985)
191.
192.
S.R. Rafikov, G.M.
Dzhilkibayeva, G.N. Chelkova, and G.B. Shaltuper, Vysokomol. soyed., A12 (7) 1608 (1970)
193.
194.
M. Hassanein, A. Akelah,
Am. Chem. Soc. Polymer
Preprints, 26 (1), 88
(1985)
195.
T.D. N’guyen, J.C. Gautier,
and S. Boileau, Am. Chem. Soc.
Polymer Preprints, 23,
143 (1982)
196.
P. Hodge, B.J. Hunt, E.
Khoshdel, and J. Waterhouse, Am.
Chem. Soc. Polymer Preprints, 23, 147 (1982)
197.
P. Hodge, B.J. Hunt, J.
Waterhouse, and A. Wightman, Polymer Communications, 24, 70 (1983)
198.
M. Takeishi, N. Umeta,
Makromol. Chem., Rapid
Commun., 3, 875
(1982)
199.
200.
201.
202.
203.
F. Marinola and G. Naumann,
Angew. Makromolekulare
Chem., 4/5,185
(1968)
204.
S. Dragan, I. Petrariu and
M. Dima, J. Polymer Sci., Chem.
Ed., 10, 3077
(1986)
205.
S. Dragan, V. Barboiu, D.
Csergo, I. Petrariu, M. Dima, Makromol. Chem., 187, (1986)
206.
R.E. Kesting, Synthetic Polymeric Membranes,
McGraw-Hill, New York, 1971
207.
208.
G.V. Kharitonov and L.P.
Belikova, Vysokomol.
soyed., A11 (11), 2424
(1969)
209.
W.O. Kenyon and G.P. Waugh,
J. Polymer Sci.,
32, 83 (1958); C,C, Unruh,
J. Appl. Polymer Sci.,
2, 358 (1959)
210.
211.
D.G. Barar, K.P. Staller,
and N.A. Peppas, J. Polymer Sci.,
Chem. Ed., 21, 1013
(1983)
212.
R. Singer, A. Demagistri,
and C. Miller, Makromol.
Chem., 18/19, 139
(1956)
213.
R. Singer and A.
Demagistri, J. Chim. Phys.,
47, 704 (1950)
214.
M. Baer, U.S. Patent #
2,533,211 (Dec. 12, 1950)
215.
R. Singer, U.S. Patent #
2,604, 456 (July 22, 1952)
216.
B. Blaser, M. Rugenstein
and T. Tischbirek, U.S. Patent # 2,764,576 (Sept. 25,1956)
217.
218.
J. Eichhorn, U.S. Patent #
2,877,213 (Mar. 10, 1959)
219.
C.F.H. Aleen and L.M.
Minsk, U.S. Patent #2,735,841 (Feb. 21, 1956)
220.
J. Blyth and A.W. Hofman,
Ann. Chem., 53, 316 (1984)
221.
H. Zenftman, J. Chem. Soc., 1950, 820
222.
223.
224.
F. Cramer, H. Helbig, H.
Hettler, K.H. Scheit and H. Seliger, Angew. Chem., 78, 64 (1966); ibid., Intern. Edit., 5, 601 (1966)
225.
B. Philipp, G. Reinisch,
and G. Rafler, Plaste und
Kautschuk, 3, 190
(1972)
226.
W.Kern, R.C. Schulz, and D.
Braun, Chemiker Ztg./Chem.
Apparatur, 84 (12),
385 (1960)
227.
228.
229.
230.
231.
232.
D. Braun, Chimia, 14, 24 (1960)
233.
234.
235.
K. Chander, R.C. Anand,
I.K. Varma, J. Macromol. Sci. -
Chem., A20, 697
(1983)
236.
J. Petit and B. Houel,
Comp. rend., 246, 1427 (1958)
237.
B. Houel, Comp. rend., 246, 2488 (1958)
238.
R.C. Schulz, P. Elzer, and
W. Kern, Chimia,
13, 237 (1959)
239.
240.
241.
242.
243.
244.
245.
246.
247.
248.
M. Vrancken and G. Smets,
J. Polymer Sci., 14, 521
(1954)CrossRef
249.
250.
A. Ravve, J. Polymer Sci., A-1,6, 2889 (1968)
251.
252.
H. Boudevska, Izv. Bulgar. Acad. of Sci.,
3, 303 (1970)
253.
254.
255.
N.S. Bondareva and E.N.
Rastovskii, Zh. Prikl. Khim.
(Leningrad), 43 (1),
215 (1970)
256.
I.V. Andreyeva, MM. Kotin,
L.Ya. Modreskay, Ye.I. Pokrovskii, and G.V. Lyubimova, Vysokomol. soyed., A14 (7), 1565 (1972)
257.
258.
K. Imai, T. Shiomi, T.
Tsuchida, C. Watanabe, and S. Nishioka, J. Polymer Sci., A-1,21, 305 (1983)
259.
K. Imai, T. Shiomi, Y.
Tezuka, M. Miya, J. Polymer Sci.,
Chem. Ed., 22, 841
(1984)
260.
K. Ichimura, J. Polymer Sci., Chem. Ed.,
A-1,20, 1411 (1982)
261.
K. Ichimura and S.
Watanabe, J. Polymer Sci., Chem.
Ed., A-1,20, 1419
(1982)
262.
263.
264.
265.
266.
L.X. Mallavarapu and A.
Ravve, J. Polymer Sci.,
A,3, 593 (1965)
267.
G. Reinisch and Dietrich,
Eur. Polymer J.,
6, 1269 (1970)
268.
269.
J. Klein and E. Klesper,
Makromol. Chem., Rapid
Commun., 5, 701
(1984)
270.
271.
272.
H.L. Fisher, Chemistry of Natural and Synthetic
Rubber, Reinhold, New York, 1957
273.
P.W. Allen, D.Barnard, and
B. Saville, Chem. Brit.,
6, 382 (1970)
274.
275.
N.G. Gaylord and F.S. Ang,
Chemical Reactions of
Polymers, (E.M. Fetters, ed.), Wiley-Interscience, New York,
1964
276.
277.
278.
S. Colonna, R. Fornasier,
and U. Pfeiffer, J. Chem. Soc.,
Perkin Trans., 1, 8
(1978)
279.
E. Chiellini and R. Solaro,
J. Chem. Soc., Chem.
Commun., 1977,
231
280.
A. Mastagli, A. Froch, and
G. Durr, C.R. Acad. Sci.,
Paris, 235, 1402
(1952)
281.
E.C. Blossey, D.C. Neckers,
C. Douglas, A.L. Thayer, and A.P. Schaap, J. Am. Chem. Soc., 95, 582 (1973)CrossRef
282.
H.A.J. Battaerd and G.W.
Tregear, Graft Copolymers,
Wiley-Interscience, New York, 1967
283.
P. Dreyfus and R.P. Quirk,
“Graft Copolymers”, in Encyclopedia of Polymer Science and
Engineering, Vol 7, 2nd. ed., (H.F. Mark, N.M. Bikales, C.G.
Overberger, and G. Menges, eds.), Wiley-Interscience, New York,
1986
284.
S. Basu, J.N. Sen, and S.R.
Palit, Proc. Roy. Soc.
(London), A202, 485
(1950);
285.
286.
D. Lim and O. Wichterle,
J. Polymer Sci.,
5, 606 (1961)
287.
288.
S. Okamura and K. Katagiri,
Makromol. Chem.,
28, 177 (1958); Z.M. Wang, H.
Hong. and T.C. Chung, Macromolecules, 2005, 38, 8966
289.
G.V. Schulz, G. Henrici,
and S. Olive’, J. Polymer
Sci., 17, 45
(1955)
290.
291.
292.
293.
294.
295.
M.S. Gluckman, M.J. Kampf,
J.L. O’Brien, T.G. Fox, and R.K. Graham, J. Polymer Sci., 37, 411 (1959)CrossRef
296.
T.G. Fox, M.S. Guckman,
F,B, Gornick, R.K. Graham, and G. Gratch, J. Polymer Sci., 37, 397 (1959)
297.
298.
299.
S. Nakamura, H. Sato, and
K. Matsuzaki, J. Polymer Sci.,
Polymer Letters, 11,
221 (1973)
300.
301.
302.
R.J. Ceresa, (ed.)
Block and Graft Copolymers,
Wiley Interscience, New York, 1973
303.
304.
305.
R.E. Wetton, J.D. Moore,
and B.E. Fox, Makromol.
Chem., 132, 135
(1970)
306.
L.V. Valentine and C.B.
Chapman, Ric. Sci. Suppl.,
25, 278 (1955)
307.
308.
G.I. Simionescu and S.
Dumitriu, J. Polymer Sci.,
C (37), 187 (1972)
309.
310.
R.M. Livshits, V.P.
Alachev, M.V. Prokofeva, and Z.A. Rogovin, Vysokomol. soyed., 6 (4), 655 (1964)
311.
Y. Iwakura, T. kurosaki,
and Y. Imai, J. Polymer
Sci., A-1, 3, 1185
(1965); Y.Iwakura, Y. Imai, and K. Yagi, J. Polymer Sci., A-1, 6, 801 (1968); ibid., A-1, 6, 1625 (1968); Y. Imai, E.
Masuhara, and Y. Iwakura, J.
Polymer Sci., Polymer Letters, 8, 75 (1970)
312.
Y. Ogiwara, Y. Ogiwara, and
H. Kubota, J. Polymer Sci.,
A-1,6, 1489 (1968)
313.
A.A. Gulina, R.M. Livshits,
and Z.A. Rogovin, Vysokomol.
soyed., 7 (9) 1693
(1965)
314.
315.
A. Hebeish, S. Shalaby, A.
Waly, and A. Bayazeed, J. Polymer
Sci., 28, 303
(1983)
316.
J.E. Guillet and R.G.W.
Norrish, Nature, 173, 625 (1954); Proc. Roy. Soc. (London), A 233, 172 (1956)
317.
318.
N, Geacintov, V. Stannett,
and E.W. Abrhamson, Makromol.
Chem., 36, 52 (1959);
A. Chapiro, J. Polymer
Sci., 29, 321 (1958);
ibid., 34, 439 (1959)
319.
320.
321.
Y. Hachihama, S. Takamura,
Technol. Report Osaka
Univ., 11 (485) 431
(1961); from Chem. Abstr.,
57, 6179 (1962)
322.
323.
A. Noshay and J.E. McGrath,
Block Copolymers, Academic
Press, New York, 1977
324.
325.
326.
327.
R. Milkovich, Am. Chem. Soc. Polymer Preprints,
21, 40 (March 1980)
328.
R. Milkovich and M.T.
Chiang, U.S. Patent # 3,786,116 (1974)
329.
M. W. Thompson and F.A.
Waite, Brit. Patent # 1,096,912 (1967)
330.
B.W. Jackson, U.S. Patent #
3,689,593 (1972)
331.
332.
J.P. Kennedy, Preprints, 5th Intern. Symposium on Cationic
and Other Ionic Polymerizations, 6, (1980)
333.
334.
335.
P. Rempp, Polymer Prepr. Jpn., 29, 1397 (1980)
336.
Y. Tanizaki, K. Minagawa,
S. Takase, and K. Watanabe, Abstracts of the A.C.S.-C.S.J. Chemical
Congress,INDE, 136 (1979)
337.
338.
Y. Yamashita, J. Appl. Polymer Sci., Applied Polymer
Symposia, 36, 193
(1981)
339.
340.
D.J. Metz and R.B.
Mesrobian, . J. Polymer
Sci., J16, 345
(1955)
341.
342.
343.
D. Makulova and D. Berek,
Sb. Prac. Chem. Fak. Sloven
Vysokej. Skily Tech., 141, (1961); from Chem Abstr., 57, 8729 (1962)
344.
J.J. Wu, Z.A. Rogovin, and
A.A. Konkin, Khim Volokna,
6, 11 (1962); from
Chem. Abstr., 58, 11509 (1963)
345.
346.
347.
A. Ravve and C.W. Fitko,
J. Polymer Sci.,
A-1,4, 2533 (1966)
348.
C. de Ruijter, W. F. Jager,
J. Groenewold, and S. J. Picken, Macromolecules, 2006, 39, 3824
349.
Schollenberger,
Polymer Technology,
Interscience, New York, 1969
350.
K.C. Gupta and K.
Khandekar, Am. Chem. Soc. Polymer
Preprints, 2005,
46 (1), 654
351.
352.
353.
354.
M. Tsuda, J. Polymer Sci., 7, 259 (1969);
355.
H.G. Heine, H.J.
Resenkranz, and H. Rudolph, Angew.
Chem., Int. Edit.Engl., 11, 974 (1974);
356.
C.J. Groeneboom, H.J.
Hageman, T. Overeem, and A.J.M. Weber, Makromol. Chem., 183, 281 (1982)
357.
J. Hutchison, M.C. Lambert,
and A. Ledwith, Polymer,
14, 250 (1973)CrossRef
358.
359.
W.A. Pryor, Free Radicals, McGraw-Hill, New York,
1966
360.
M.J. Gibian, Tetrahedron Letters, 5331 (1967)
361.
G. Rausing and S. Sunner,
Tappi, 45 (1), 203A (1962)
362.
363.
364.
365.
S.G. Cohen, H.S. Haas, and
H. Slotnick, J. Polymer
Sci., 11, 913 (1953);
H.S. Haas, S.G. Cohen, A.C. Oglesby, and E.R. Karlin, J. Polymer Sci., 15, 427 (1955); S.T. Rafikov, G.N.
Chelnokova, I.V. Zhuraleva, and P.N. Gribkova, J. Polymer Sci., 53, 75 (1961)
366.
J.P. Kennedy, ed.,
J. Appl. Polymer Sci., Applied
Polymer Symposia, 30,
(1977); J.P. Kennedy, U.S. Patent # 3,349,065 (Oct. 24, 1967); J.P.
Kennedy in Polymer Chemistry of
Synthetic Elastomers, ( J.P. Kennedy and E. Tornquist,
ed.)Wiley-Interscience, New York, 1968; J.P. Kennedy and J.K.
Gillham, Adv. Polymer Sci.,
10, 1 (1972)
367.
J.P. Kennedy and M. Nakao,
J. Appl. Polymer Sci., Appl.
Polymer Symposia, 30,
73 (1977)
368.
369.
E.T. Kang, K.G. Neoh, K.L.
Tan, Y. Uyama, N. Morikawa, and Y. Ikada, Macromolecules, 25, 1959 (1992)CrossRef
370.
B.U. Kolb, P.A. Patton, and
T.J. McCarthy, Am. Chem. Soc.,
Polymer Preprints, 28
(2), 248 (1987)
371.
K Matyjaszewski, YGnanou,
and L. Leibler, eds. ‘Macromolecular Engineering, Wiley, New
York, 2007
372.
J. Hirkach, K. Ruehl, and
K. Matyjaszewski, Am. Chem. Soc.,
Polymer Preprints, 29
(2), 112 (1988)
373.
Y. Jiang and J.M.J.
Frechet, Am. Chem. Soc., Polymer
Preprints, 30 (1), 127
(1989)
374.
J.F. Kenney, Polymer Eng. and Sci ., 8 (3), 216 (1968); W.H.Charch and J.C.
Shivers, Textile Research
J., 29, 536
(1959)
375.
Imperial Chemical
Industries Ltd. Brit. Patent # 802,921 (Nov. 23, 1955)
376.
A. Noshay and J.E. McGrath,
Block Copolymers: Overview and
Critical Survey, Academic Press, New York, 1977
377.
D.H. Solomon, Chapter 1 in
Step Growth Poymerization,
(D.H. Solomon, ed.), Dekker, New York, 1972
378.
J.A. Brydson, Plastic Materials, 4th ed., Butterworth
Scientific, London 1982
379.
General Electric Co., Brit.
Patent # 984,522 (June 1, 1960)
380.
Schollenberger,
Polymer Technology. (P.F.
Bruins, ed.) Interscience, New York, 1969
381.
R. Ukielski and H.
Wojcikiewicz, Intern. Polymer
Science and Technology, 3 (11), T65 (1976); G.K. Hoexchele and
W.K. Witsiepe, Angew. Makromol.
Chem., 29-30, 267
(1973)
382.
Y. Yamashita and T. Hane,
J. Polymer Sci., Polym. Chem.
Ed., 11, 425 (1973);
Y. Yamashita, Y. Murase, and K. Ito, J. Polymer Sci., Polym. Chem. Ed.,
11, 435 (1973)
383.
V.V. Korshak and T.M.
Frunze, Synthetic Hetero-Chain
Polyamides, S. Manson, Jerusalem. 1987
384.
P.W. Morgan, J. Polymer Sci., C4, 1075 (1964); D.J. Lyman and S.L.
Yung, J. Polymer Sci.,
40, 407 (1959)
385.
386.
V.V. Korshak, S.V.
Vinogradova, M.M. Teplyakov, R.D. Fedorova, and G.Sh. Papara,
Vysokomol Soyed.,
8 (12), 2155 (1966);
V.V.Korshak, S.V.Vinogradova, M.M.Teplyakov, and A.D.Maksomov,
Izv. Vysok. Uchebn. Zaved Khim. i
Khim. Technol.,
10 (6), 688 (1967)
387.
V.V. Korshak, S.V.
Vinogradova, and M.M. Teplyakov, Vysokomol. Soyed., 7 (8), 1406 (1965)
388.
Y. Goto and S. Miwa,
Chem. High Polymers, Japan,
25, 595 (1968)
389.
K.L. Mittal, Polyamides: Syntheses, Characterizations and
Applications, Vols. 1 and 2, Plenum, New York, 1984
390.
Imperial Chemical
Industries, Brit. Patent # 807,666 (Jan 4, 1957)
391.
Imperial Chemical
Industries, Brit. Patent # 1,148,068 (Aug. 9, 1965)
392.
M. Tsuruta, F. Kabayashi,
and K. Matsuyata, Kogyo Kagaku
Zasshi, 62, 1084, 1087
(1959); from Chem. Abstr.,
57, 13977 (1962)
393.
D. Dieterich, W. Keberle,
and H. Witt, Angew. Chem., Intern.
Ed., 9 (1), 40
(1970)
394.
D. Dieterich. O. Bayer, J.
Peter, and E. Muller, Ger. Patent # 1,156,977 (June 5, 1962)
395.
D. Dieterich and O. Bayer,
Brit. Patent # 1,078,202; W. Keberle and D. Dieterich, Brit. Patent
# 1,078,688
396.
C.A. Lukach and H.M.
Spurlin, Copolymerization,
(G.E. Ham, ed.), Interscience, New York, 1964; H.F. Mark and S.M.
Atlas,Petrol. Refiner,
39, 149 (1960); G. Natta,
J. Polymer Sci.,
34, 21, 531 (1959)
397.
A.V. Tobolsky and K.F.
O’Driscoll, DAS 1 114 323, June 24, 1959
398.
M. Szwarc and J. Smid,
Progress in Reaction
Kinetics, Vol. II, p. 219, Pegamon Press, Oxford, 1964; M.
Szwarc, Nature,
178, 1168 (1956)
399.
400.
R. Chiang, J.H. Rhodes, and
R.A. Evans, J. Polymer
Sci., A-1,4, 3089
(1966)
401.
N.L. Madison, Amer. Chem. Spc. Polymer Preprints,
7, 1099 (1966)
402.
403.
404.
405.
406.
S. Bywater, Macromolecular Chemistry, Buterworths,
Longond, 1962; M. Morton, Anionic
Polymerization: Principles and Practice, Academic Press, New
York, 1983
407.
408.
409.
L.J. Fetters, J. Polymer Sci., C (26), 1 (1969)
410.
J.P. Kennedy, Makromol. Chem., Suppl., 3, 1 (1979)
411.
R.J. Ceresa, Block and Graft Copolymers,
Butterworths, Washington D.C., 1962
412.
H. Staudinger, Ber., 57, 1203 (1924); Kautchuk, 5, 128 (1929)
413.
414.
415.
B. Gordon III and L.F.
Hancock, Am.Chem.Soc. Polymer
Preprints, 26 (1), 98
(1985)
416.
P.J. Hans and B. Gordon
III, Am.Chem.Soc. Polymer
Preprints, 28 (2), 310
(1987)
417.
W. Risse and R.H. Grubbs,
Am.Chem.Soc. Polymer
Preprints, 30, 193
(1989)
418.
419.
D. Patit, R. Jerome, and
Ph. Teyssie, J. Polymer Sci.,
Polymer Chem. Ed., 17,
2903 (1979); W.T. Allen and D.E. Eaves, Angew. Makromol. Chem., 58/59, 321 (1977); B.H. Wondraczek and
J.P. Kennedy, J. Polymer Sci.,
Polymer Chem. Ed., 20,
173 (1982)
420.
C.A. Veith and R.E. Cohen,
Am.Chem.Soc. Polymer
Preprints, 31 (1), 42
(1990)
421.
J.P. Kennedy, B. Keszler,
Y. Tsunogae, and S. Midha, Am.
Chem. Soc. Polymer Preprints, 32 (1), 310 (1991)
422.
423.
T. Higashimura, S. Aoshima,
and M. Sawamoto, Makromol. Chem.
Macromol Symp .,
13/14, 457 (1988)
424.
J.P Kennedy and J.L. Price,
Polym. Mater. Sci.,
1991, 64; J.P. Kennedy, J.L.
Price, and K. Koshimura, Macromolecules, 24, 6567 (1991)
425.
426.
M.G. Kanatzidis,
Chem. and Eng. News, p. 38
(Dec. 3, 1990); L.Y. Chiang, P.M. Chaikin, D.O. Cowan,(eds.),
Advanced, Organic Solid State
Materials, Mat. Res. Soc.
Symp. Proc .,
173, (1990); J.R. Reynolds,
Chemtech, 18, 440 (1988)
427.
X.-F. Sun, S.B. Clough, S.
Subrmanyam, A. Blumstein, and S.K. Tipathy, Am. Chem. Soc. Polymer Preprints,
33 (1), 576 (1992)
428.
429.
R.D. McCullough and R.D.
Lowe, Am. Chem. Soc. Polymer
Preprints, 33 (1), 195
(1992)
430.
W. Rutsch, G. Berner, R.
Kirchmayr, R. Husler, G. Rist, and N. Buhler, Organic Coatings-Science and
Technology, (G.D Parfitt and A.V. Patsis, eds.), Dekker, New
York, Vol. 8, 1986
431.
K. Dietliker, M. Rembold,
G. Rist, W. Rutsch, and F. Sitek, Rad-Cure Europe 87, Proceedings
of the 3rd Conference Association of Finishing Processes SME;
Dearborn, MI, 1987
432.
G. Rist, A. Borer, K.
Dietliker, V. Desobry, J.P. Fouassier, and D. Ruhlmenn,
Macromolecules,
25, 4182 (1992)CrossRef
433.
B.J. Sun and G.C. Bazan,
. Am. Chem. Soc. Polymer
Preprints, 36 (1), 253
(1995)
434.
M. Weck, P. Schwab, and
R.H. Grubbs, Macromolecules, 29, 1789 (1996); A. Proto, A. Avagliano,
D. Saviello, R. Ricciardi and C. Capacchione, Micrornolecules, 2010, 43 5919
435.
R. Jerome, J.L. Hedrick,
and C.J. Hawker, Angew. Chem, Int
Ed. Engl. 37, 1274
(1998); C.J. Hawker, J.L. Hedrick, E.E. Malmstrom, M. Tollsas, D.
Mecerreyes, G. Moineau, Ph. Dubois and R. Jerome, Macromolecules, 31, 213 (1998)
436.
E. Yoshida and Y. Osagawa,
Macromolecules,
31, 1446 (1998); R.L.Kuhlman
and J. Klosin, Macromolecules, 2010, 45, 7903; L.M. Pitet and M.A. Hillmyer,
Macromolecules,
2009, 42, 3674
437.
Y. Kotani, M. Kato, M.
Kamigaito, and M. Sawamoto, Macromolecules, 29, 6979 (1996); S. Lu, Q.-L. Fan, S.-J.
Chua, and W. Huang, Macromolecules, 2003, 36, 304; J.B. Matson and R.H. Grubbs, ,
Macromolecules,.2008,
ASAP article10.1021/ma800980p
438.
S. Coca, H.-j. Paik, and K.
Matyjaszewski, Macromolecules, 30, 6513 (1997)
439.
440.
441.
442.
J.Vohlidal, J. Sedlacek,
and M. Zigon, Kovine, Zlitine,
Tehnol. 31(6), 581
(1997); from Chem. Abstr.
129, 637 (1998)
443.
444.
W. Liu, J. Kumar, S.
Trypathy, K.J. Senecal, L. Samuelson, J. Am. Chem. Soc., 121, 71 (1999)CrossRef
445.
446.
L.A. Wall, S.L. Madorsky,
D.W. Brown, S. Straus, and R. Simka, J. Am. Chem. Soc., 76, 3430 (1954)
447.
L.A. Wall and S. Straus,
J. Polymer Sci.,
1960, 44, 313
448.
S. Madorsky, J. Polymer Sci., 1952, 9, 133
449.
S. Madorsky, J. Polymer Sci., 1953, 11, 491
450.
S. Madorsky, S. Straus, D.
Thompson, and L. Williamson, J.
Polymer Sci., 1949,
4, 639
451.
S. Madorsky and S. Straus,
J. Polymer Sci.,
1959, 36, 183
452.
S. Madorsky and S. Straus,
J. Res. Natn. Bur. St.,
1948, 40, 417
453.
W.G. Oaks and R.B.
Richards, J..Chem. Soc.
(London), 1949,
2929
454.
Y. Tsuchiya and K. Sumi,
J. Polymer Sci.,
1968, A-1,6, 415
455.
Chem. and Eng. News, p. 41,
Sept. 28, 1964
456.
S. Madorsky, Thermal Degradation of Organic
Polymers, Interscience, New York, 1964
457.
E.M. Bavilacqua, in
Thermal Stability of
Polymers, Vol. 1., R.T. Conley (ed.), Dekker, New York,
1970; R. Bernstein, D.R. Derzon, and K.T. Gillen, Am. Chem. Soc. Polymer Preprints,
2007, 48(1), 611
458.
459.
460.
G.G. Cameron and N.
Grassie, Polymer,
19961, 2, 367
461.
H.H.G. Jellinek,
J. Polymer Sci.,
1948, 3, 850
462.
H.H.G. Jellinek,
J. Polymer Sci.,
1949, 4, 1
463.
H.H.G. Jellinek,
J. Polymer Sci.,
1949, 4, 13
464.
465.
466.
467.
468.
N. Grassie and J.G.
Speakman, J. Polymer Sci.,
1971, A-1,9, 419
469.
N. Grassie, J.G. Spaakman,
and T.J. Davis, J. Polymer
Sci., 1971,
A-1,9, 931
470.
S. Choudhary and K.
Lederer, Eur. Polymer J.,
1982, 18, 1021
471.
W.J. Burlant and J.L.
Parsons, J. Polymer Sci.,
1956, 22, 156
472.
J. Schulz, H. Bayzer, W.
Stubchen, and J. Schurz, Makromol.
Chem. 1957,
23, 152; ibid., 1959, 29, 156
473.
S.L. Ma472orsky amd S.
Strais. J. Res. Natl. Bur.
Std.. 1958,
61, 77
474.
N. Grassie and J.N. Hay,
J. Polymer Sci.,
1962, 56, 189
475.
476.
J.N. Hay, J. Polymer Sci., 1968. A-1,6, 2127
477.
478.
N.A. Kubasova, Din Suan
Din, M.A. Heiderich, and M.V. Shishkina, Vysokomol. soyed. , 1979, A13 (1), 162
479.
N. Grassie and R. McGuchan,
Eur. Polymer. J.,
1970, 6, 1277; ibid., 1971, 7, 1091; ibid., 1971, 7,1357
480.
H. Gilbert, S.J. Averill,
F.E. Miller, R.F. Schmidt, F.D. Stewart, and H.L. Trumbull,
J. Am. Chem. Soc.,
1954, 76, 1074CrossRef
481.
482.
P. Bataille and B.T. Van,
J. Polymer Sci.,
1972, A-1,10, 1097
483.
484.
J.D. Danforth, J. Spiegel,
and J. Bloom, J. Macromol. Sci. –
Chem., 1982,
A17 (7), 1107
485.
S.A. Liebman, D.H.
ahlstrom, E.J. Quinn, A.G. Geigley, and J.T. Melluskey,
J. Polymer Sci.,
1971, A-1,9, 1921.
486.
I. Mita in Aspects of degradation and Stabilization of
Polymers, H.H. Jellinek (ed.), Elsevier, Amsterdam,
1978
487.
488.
E.P.Chang and R.Salovey,
J. Polymer Sci., Chem. Ed.,
1974, 12, 2927
489.
D.O. Hummel, H.J. Dussel,
and K. Rubenacker, Makromol.
Chem., 1971,
145, 267
490.
R.M. Lum487. M.M. O’Mara,
J. Polymer Sci., Chem. Ed.,
1970, A-1,8, 1887, J. Polymer Sci., Chem. Ed.,
1977, 15, 489
491.
A. Ballistreri, S. Foti, G.
montaudo, and E. Scamporrino, J.
Polymer Sci., Chem. Ed., 1980, 18, 1147
492.
J. Svetly, R. Lucas, J.
Michalcova, and M. Kolinsky, Makromol. Chem.,Rapid Commun.,
1980, 1, 247
493.
J. Svetly, R. Lucas, S.
Pokorny, and M. Kolinsky, Makromol. Chem.,Rapid Commun.,
1981, 2, 149
494.
L. Valko, P. Simon, V.Kello
Makromol. Chem.,
1982, 183, 3057
495.
496.
R.F. Boyer, J. Phys. Coll. Chem., 1947, 51, 80; E. Elakesh, T.R. Hull, D.
Price, P.Carty, J. of Vinyl and
Additive, 2003,
9(3), 116
497.
498.
C,B. Havens, N.B.S. Cicrular # 525, 1953, 107
499.
F.H. Winslow, W.O. Baker,
and Y.A. Yager, Proceeding of
First and Second Carbon Conferences, Univ. of Buffalo Press,
1956
500.
501.
502.
R.A. Haldeon, J. Polymer Sci., 1968, A-1,6, 951
503.
E.P. Goodings, Soc. Chem. ind. (London), Monograph,
1961, 13, 211
504.
R.M. Lum, J. Polymer Sci., 1979, 17, 203
505.
506.
G.J. Sutton, B.J. Tighe,
and m. Roberts, J. Polymer
Sci.,Chem. Ed.,
1973, 11, 1079
507.
A. Davis and J.H. Golden,
J. Chem. Soc., B, (London),
1968, 45
508.
509.
510.
511.
Ye,P. Krasnov, V.P.
Aksenova, S.N. Kharkov, and S.A. Baranova, Vysokomol. soyed., 1970, A12 (4), 873
512.
V.V. Rode. Ye.M.
Bondarenko, V.V. Korshak, Ye.S. Krongauz, and A.L. Rusanov,
Dokl. ANSSSR, 1966, 171, 355
513.
V.V. Rode. Ye.M.
Bondarenko, Vysokomol.
soyed., 1967,
A9,(12), 218; W. Wrasidlo,
J. Polymer. Sci.,
1970, A-1,8, 1107
514.
A. Scotney, Eur. Polymer J. 1972, 8, 175; ibid., 1972, 8, 185
515.
I.I. Levantovskaya, G.V.
Dralyuk, O.A. Mochalova, I.A. Yurkova, M.s. Akutin, and B.V.
Kovarkaya, Vysokomol.
soyed, 1971,
A13 (1), 8.
516.
V.V. Kopylov and A.N.
Pravednilkov, Visokomol.
Soyed., 1969,
A11, 849
517.
518.
Yu. Sclyapnikov, T.A.
Bagaenskay, S.G. Kiryushkin, and T.V. Monakhova, Eur. Polymer J. 1979, 15, 737CrossRef
519.
.W.L. Hawkins and F.H.
Winslow in Chemical Reactions of
Polymers, E.M. Fetters (ed.), Interscience, NY.1964
520.
J.H. Adams, J. Polymer Sci., 1970, A-1,8, 1077
521.
V.M. Yurev, A.A.
Pravednikov, and S.S. Medvedev, Dokl.Akad. Nauk S.S.S.R., 1959, 124, 335
522.
523.
B.G. Achhammer, M.J.
Reiney, and F.W. Reinhart, J.
Polymer Sci., 1952,
8, 555
524.
A. Votinova, P. Kobenko,
and F. Marei, Zh. fiz.
Khim., 1952,
16, 106
525.
C. Crouzet and J. Marchal,
J. Appl. Polymer Sci., Appl
Symposia, 1979,35, 151
526.
S.D. Rasumovskii, O.N.
Karpukhin, A.A. Kefeli, T.V. Pokholok, and G.Ye. Zaikov,
Vysokomol. soyed.,
1971, A13, 782
527.
528.
K.,C.-Yu, W.,B.-Zu; C.
W.-Hsun, lnd. Eng. Chem.
Res. 1998, 37(4),
1228
529.
530.
H.H.G. Jellinek and A.K.
Chandhuri, J. Polymer Sci.,
1972, A-1,10, 1773 88. Synthetic Lubricants and Liquids,
R.S. Gunderson and A.V. Khart, (eds.), Chemistry publishing House,
New York, 1965
531.
P.A. Okunev and O.G.
Tarakanov, Vysokomol.
soyed. 1968,
A10, (1), 173
532.
I.I.. Levantovskaya, G.V.
Dralyuk, O.A. Mochalova, I.A. Yurkova, M.S. Akutin, and
B.V.Kovarkaya, Vysokomol.
soyed., 1971,
A13 (1), 8
533.
H.H.G. Jellinek and S.H.
Ronel, J. Polymer Sci.,
1971, A-1,9, 2605
534.
Fitton,.N. Howard, and G.R.
Williamson, Brit. Polymer
J.,1970, 2, 217
535.
B. Ranby and J.F. Rabek,
J. Appl. Polymer Sci., Appl.
Symposia, 1979,
35, 243
536.
Y. Hama, K. Hosono, Y.
Furui, and K. Shinohara, J.
Polymer Sci, 1969,
B7, 839
537.
T. Takeshita, K. Tsuji, and
T. Keiki, J. Polymer Sci.,
1972, A-1,9, 1411
538.
K. Tsuji and T. Seiki,
J. Polymer Sci., Letters,
1972, 10, 139
539.
D.J. Carlson and D.M.
Wiles, Macromolecules,
1969, 2, 587; 597
540.
T. Kagiya and K. Takemoto,
J. Macromol. Sci.,
1976, 5, 795
541.
T. Takashita, K. Tsuji, and
T. Keiki, J. Polymer Sci.,
1972, A-1,10, 2315
542.
I.M. Rozman, Izv. Akad. Nauk S.S.S.R., 1958, 22, 60
543.
G. Oldham, A.R. Ware, and
S.A. Blackwell, Nucl.
Energy, 1967,
20, 158
544.
N.A. Weir, J. Appl. Polymer Sci., 1973, 17, 401
545.
N.A. Weir and T.H. Milkie,
eur. Polymer J.,
1980, 141
546.
N. Grassie and B.J.D.
Torrance, J. Polymer Sci.,
1968, A-1,6, 3303; 3315
547.
548.
N. Grassie and R.H.
Jenkins, Eur. Polymer
J.,1973, 9, 697
549.
550.
551.
552.
E.D. Owens and R.J. Bailey,
J. Polymer Sci.,
1972, A-1,10, 113
553.
C. Decker and M. balander,
Eur. Polymer J.,
1982, 18, 1085
554.
E.D. Owens and J.I.
Williams, J. Polymer Sci.,Chem.
Ed., 1974,
12, 1933
555.
C. David, W. Demarteau, and
G. Gueskens, Polymer,
1967, 497
556.
M. Day and D.M. Wiles,
J. Polymer Sci.,Letters,
1971, 9, 665
557.
S.B. Maerov, J. Polymer Sci., 1965, A-1,3, 487
558.
S.M. Cohen, R.H. Young, and
A.H. Markhart, J. Polymer
Sci., 1971,
A-1,9, 3263
559.
J.F. Rabek, G. Ramme, G.
Cauback, b. Ranby, and V.T. Kagiya, Eur. Polymer J., 1979, 15, 339
560.
D.J. Carlsson, R.D.
Parnell, and D.M. Wiles, J.
Polymer Sci.,Letters, 1973, 11, 149
561.
562.
M.L. Kaplan and P.O.
Keller, J. Polymer
Sci.,Letters, 1971,
9, 565
563.
J.M. Ginhac, J.L Gardette,
R. Arnaud, and J. Lemaire, Makromol. Chem., 1981, 182, 1017
564.
A. Huvet, J. Phillipe, and
j. Verdu, Eur. Polymer J.,
1978, 14, 709
565.
B. Ranby and J.F. Rabek,
J. Appl. Polymer Sci., Appl.
Symposia, 1979,
35, 243
566.
J.F. Rabek, G. Cauback, and
B. Ranby, J. Appl. Polymer Sci.,
Appl. Symposia, 1979,
35, 299
567.
M.L. Kaplan and P.G.
Keller, J. Polymer Sci.,
1970, A-1,8, 3163
568.
J.F. Rabek, G. Ramme, G.
Cauback, B. Ranby, and V.T. Kagiya, Eur. Polymer J., 1979, 15, 339
569.
J. Chaineaux and C.
Tanielian, J. Appl. Polymer Sci.,
Appl. Symposia, 1979,
35, 337
570.
571.
C.S. Schollenberger and
F.d. Stewart, Advances in Urethane
Science and Technology, K.C. Frisch and S.L. Reegen (eds.),
Technology Publishing Co., inc. Westport, Conn., 1976
572.
L. Reich and S.S. Stivala,
Elements of polymer
Degradation, McGraw-Hill Book Co., New York, 1971
573.
A. Chapiro, Irradiation of Polymers, R.F. Gould
(ed.), Advances in Chemistry
Series, # 66, Am. Chem. Soc., Washington, D.C., 1967
574.
575.
F. Harlen, W. Simpson, F.B.
Waddington, J.D. Waldon, and A.C. Baskett, J. Polymer Sci., 1955, 18, 589
576.
A.A. Miller, E.J. Lawton,
and J.S. Balwit, J. Polymer
Sci.,1954,
14, 503
577.
A Chapiro, Radiation Chemistry of Poymeric
Systems, Interscience-Wiley, New York. 1962
578.
E.P. Goodings, Thermal Degradation of Polymers,
Soc. Chem. Ind., Monogr.
1961, No. 13, 211
579.
G. Kamerbeek, H. Kroes, and
W. Grolle Thermal Degradation of
Polymers, Soc. Chem. Ind.,
Monogr. 1961,
No. 13, 357
580.
K. Pielichowski; D. Bogdal, J,
Pielichowski, A. Boron, Thermochim. Acta 1997, 307(2), 155
581.
W.H. Starnes, Jr., H.
Chung, B.J. Wojciechowski, D.E. Skillcorn, and G.M. Benedikt,
Am. Chem. Soc. Polymer
Preprints, 1993,(2),
34, 114
582.
J. Lacoste and Y. Israeli,
Am. Chem. Soc. Polymer
Preprints, 1993,(2),
34, 127
583.
C.R. Hoyle, I.B. Rufus, and
H. Shah, Am. Chem. Soc. Polymer
Preprints, 1993,(2),
34, 131
584.
A. Rivaton, D. Sallet, J.
Lemaire, J. Polymer Degrad.
Stab. 1986,
14,1
585.
586.
A. Gupper, P. Wilhem, and
M. Schiller, Polymers and Polymer
Composites, 2003,
11(2), 123