3.1 Free-Radical Chain-Growth Polymerization Process
Polymerizations by free-radical mechanism are
typical free-radical reactions. That is to say, there is an
initiation, when the
radicals are formed, a propagation, when the products are
developed, and a termination, when the free-radical
chain reactions end. In the polymerizations, the propagations are
usually chain reactions. A series of very rapid repetitive steps
follow each single act of initiation, leading to the addition of
thousands of monomers.
This process of polymerization of vinyl monomers
takes place at the expense of the double bonds, −C=C– → –C–C–.
Table 3.1
illustrates the steps in this process.
Table
3.1
Illustration of a free radical
polymerization
1. Initiation
|
|
2. Propagation
|
|
3. Termination
|
|
a. By combination
|
|
b. By disproportionation
|
|
c. By transfer
|
Formation of initiating radicals is the
rate-determining step in the initiation reaction. The formation of
initiating radicals can result, as shown in Table 3.1, from cleavages of
compounds, such as peroxides, or from other sources. Actually, many
reactions lead to formations of free radicals. The initiating
radicals, however, must be energetic enough to react with the vinyl
compounds. A linear correlation exists between the affinities of
some radicals for vinyl monomers and the energy (calculated)
required to localize a π electron at the β-carbon of the monomer
[1]. By comparison to other steps
in the polymerization process, initiation is a slow step and
requires high energy of activation.
3.1.1 Kinetic Relationships in Free-Radical Polymerizations
A kinetic scheme for typical free-radical
polymerizations is pictured as follows [2]; the rate of propagation R p is equal to the rate of
polymerization R
pol, as all the monomer molecules (except one) are
consumed during this step.
Initiation
|
|
|
|
|
|
Propagation
|
|
|
Termination
|
|
|
Transfer
|
|
|
|
In the above shown kinetic scheme, M stands for
the monomer concentration, I is the concentration of the initiator,
and [R•] and [P•] mean the concentration of primary and polymer
radicals, respectively. S stands for the chain transferring agent.
R I denotes the
decomposition rate of the initiator and R P the rate of
polymerization. The rate constants, for the initiator decomposition
is k d, for the
initiation reaction is k
I, for the propagation k P, and for the termination
is k T. The
above is based on an assumption that k P and k I are independent of the
sizes of the radicals. This is supported by experimental evidence
that shows that radical reactivity is not affected by the size,
when the chain length exceeds dimer or trimer dimensions
[3]. The reactions involved in a
typical free-radical polymerization process, as stated above, are
illustrated in Table 3.1
The equation for the rate of propagation, shown
above in the kinetic scheme, contains the term [M•]. It designates
radical concentration. This quantity is hard to determine
quantitatively because its concentration is usually very low. A
steady state assumption
is, therefore, made to simplify the calculations. It is assumed
that while the radical concentration increases at the very start of
the reaction, it reaches a constant value almost instantly. This
value is maintained from then on, and the rate of change of
free-radical concentration is assumed to quickly become and remain
zero during the polymerization. At steady state, the rates of
initiation and termination are equal, or R i = R t = 2k i[M•] [4]. This assumption makes it possible to solve
for [M•] and can then be expressed as:
The rate of propagation is
The rate of propagation is approximately equal to
the total rate of polymerization. The total rate can be designated
as R pol.
Because all but one molecule are converted during the step of
propagation, we can write:
This rate of propagation applies if the kinetic
chain length is large and if the transfer to monomer is not very
efficient. The rate of monomer disappearance can be expressed as
Because many more molecules of the monomer are
involved in the propagation than in the initiation step, a very
close approximation is
The average lifetime, τ, of a growing radical under steady
state conditions can then be written as follows [3, 5],
Not all primary radicals that form attack the
monomer. Some are lost to side reactions. An initiator efficiency
factor, f, is, therefore,
needed. It is a fraction of all the radicals that form and can be
expressed as:
The rate of initiator decomposition and the rate
equation can be expressed as
According to the kinetic scheme, chain transfer
does not affect the rate of polymerization but alters the molecular
weight of the product. Also, it is important to define the average
number of monomer units that are consumed per each initiation. This
is the kinetic chain
length, and it is equal
to the rate of polymerization per rate of initiation:
At steady state conditions, ν is also equal to
k p/k t. The kinetic chain
length can also be expressed as:
By substituting the expression for [M•] the
equation becomes:
The number
average degree
of polymerization, DP, is equal to 2 ν, if the termination takes place by
coupling. It is equal to ν,
if it takes place by disproportionation. (Terminations by coupling
or disproportionation are discussed in the section on the
termination reactions) Above kinetic relationships apply in many
cases. They fail, however, to apply in all cases [2]. To account for it, several mechanisms were
advanced. They involve modifications of the initiation,
termination, or propagation steps. These are beyond the discussions
in this book.
At steady state conditions,
3.2 Reactions Leading to Formation of Initiating Free Radicals
Initiating free radical can come from many
sources. Thermal decompositions of compounds with azo and peroxy
groups are common sources of such radicals. The radicals can also
come from “redox” reactions or through various light induced
decompositions of various compounds. Ionizing radiation can also be
used to form initiating radicals.
3.2.1 Thermal Decomposition of Azo Compound and Peroxides
The azo compound and peroxides contain weak
valence bonds in their structures. Heating causes weak bonds in
these compounds to cleave and to dissociate into free radicals as
follows:
For many azo compounds such dissociations occur
at convenient elevated temperatures. One commonly used azo compound
is α,α′-azobisisobutyronitrile. An original synthesis of this
compound was reported to be as follows [3, 4]:
The final products of decomposition of this
compound are two cyanopropyl radicals and a molecule of nitrogen:
As stated earlier, not all free radicals that
form, however, initiate polymerizations. Some are lost to side
reactions. Thus, for instance, some free radicals that form can
recombine inside or outside the solvent cage, where the
decompositions take place, to yield either
tetramethylsuccinonitrile or a ketenimine [4, 5]:
Examples of other, fairly efficient, azo
initiators include the following:
1.
N-Nitrosoacylanilides,
2.
Bromobenzenediazohydroxide,
3.
Triphenylazobenzene,
The triphenylmethyl radical shown above is
resonance stabilized and unable to initiate polymerizations. The
phenyl radical, on the other hand is a hot radical. It initiates
polymerizations readily. Decomposition rates of some azonitrile
initiators are listed in Table 3.2. There are also many
peroxides available for initiating free-radical polymerizations.
These can be organic and inorganic compounds. There are, however,
many more organic peroxides available commercially than are the
inorganic ones. The organic ones include dialkyl and diaryl
peroxides, alkyl and aryl hydroperoxides, diacyl peroxides, peroxy
esters, and peracids. Hydrogen peroxide is the simplest inorganic
peroxide.
Table
3.2
Decomposition rates of some azonitrile
initiatorsa
Compound
|
Solvent
|
T(°C)
|
K
db(s−1)
|
---|---|---|---|
2,2′-azobisisobutyronitirle
|
Benzene
|
78.0
|
8 × 10−5
|
2,2′-azobis-2-ethylpropionitrile
|
Nitrobenzene
|
100.0
|
1.1 × 10−3
|
2,2′-azobis-2-cyclopropylpropionitrile
|
Toluene
|
50.0
|
8.2 × 10−5
|
1,1′-azobiscyclohexanenitrile
|
Toluene
|
80.0
|
6.5 × 10−6
|
2,2′-azobis-2-cyclohexylpropionitrile
|
Toluene
|
80.0
|
8.3 × 10−6
|
1,1′-azobiscyclooctanenitrile
|
Toluene
|
45.0
|
1.5 × 10−4
|
Syntheses, structures, and chemistry of various
peroxides were described thoroughly in the literature
[5]. Here will only be mentioned
some properties of peroxides and their performance as they pertain
to initiations of polymerizations. Decompositions of peroxides,
such as the azo compounds, are also temperature dependent
[6]. This means that the rates
increase with temperature. The rates are also influenced by the
surrounding medium, such as the solvents that imprison or “cage”
the produced pairs of free radicals. Before undergoing a net
translational diffusion out of the cage, one or both of the
radicals may or may not expel a small molecule. For instance,
benzoyl peroxide can and often does decompose into a phenyl radical
and carbon dioxide, as follows:
The resultant phenyl radicals can combine and
yield new and completely inactive species:
The above-described recombination reactions of
free radicals are some of the causes of inefficiency among
initiators. The average time for recombination of free radicals
inside a solvent cage and also the time for their diffusion out of
the cage is about 10−10 s [7]. In addition, the efficiency of the initiator
is affected by the monomer and by the solvent. It was shown that
the viscosity of the medium is inversely proportional to the
initiator efficiency because the more viscous the solution, the
greater the cage-effect [8,
9].
Numerous lists are available in the literature
that give the decomposition temperatures or the half-lives at
certain elevated temperatures of many initiators [6]. Decompositions of peroxides may proceed via
concerted mechanisms [10,
11] and the rates are structure
dependent. This can be illustrated on benzoyl peroxide. The benzoyl
groups, the two halves of this molecule, are dipoles. They are
attached, yet they repel each other. Rupture of the peroxide link
releases the electrostatic repulsion between the two dipoles.
Presence of electron donating groups in the para position increases the repulsion,
lowers the decomposition temperature, and increases the
decomposition rate. The opposite can be expected from electron
attracting groups in the same position [6]. The effect of substituents on the rate of
spontaneous cleavage of dibenzoyl peroxide was expressed
[11] in terms of the Hammett
equation, log (K/K O) = ρ σ. This is shown in
Table 3.3.
Substituent
|
K
i × 103
|
log K 1/K
|
σ
i + σ
0
|
---|---|---|---|
p,p′-dimethoxy
|
7.06
|
0.447
|
0.536
|
p-methoxyy
|
4.54
|
0.255
|
0.268
|
p,p′-dimethyl
|
3.68
|
0.164
|
0.340
|
p,p′-di-t-butyl
|
3.65
|
0.161
|
0.394
|
Parent compound
|
2.52
|
0.000
|
0.000
|
p,p′-dichloro
|
2.17
|
−0.065
|
+0.454
|
m,m′-dichloro
|
1.58
|
−0.203
|
+0.746
|
m,m′-dibromo
|
1.54
|
−0.215
|
+0.782
|
p,p′-dicyano
|
1.22
|
−0.314
|
+1.300
|
In addition, peroxides can cleave in two ways,
heterolytically and
homolytically.
Heterolytic cleavage of peroxides results in formation of ions,
but homolytic cleavage results in formation of radicals:
In the gaseous phase, the cleavage is usually
homolytic because it requires the least amount of energy
[12]. In solution, however, the
dissociation may be either one of the two, depending upon the
nature of the R groups. Heterolytic cleavage may be favored, in
some cases, if the two groups, R and R′, differ in electron
attraction.
The same is true if the reaction solvent has a
high dielectric constant. Solvation of the ions that would form due
to heterolytic cleavage is also a promoting influence for such a
cleavage:
where, S represents the solvent.
In sum total, the types and the amounts of side
reactions that can take place are a function of the structures of
the peroxides, the stability of the formed radicals, the solvent,
and the monomer that is being polymerized. The stability of the
radicals that form can also affect the amount of radicals being
captured by the monomers. Also, it was reported that while
generally the character of free radicals is neutral, some of them
are electrophilic (such as chloro) and others are nucleophilic
(such as t-butyl). This
tendency, however, is relatively slight when compared with positive
and negative ions [15].
There is much information in the literature on
the rates and manner of decomposition of many peroxides in various
media. Beyond that, diagnostic tests exist that can aid in
determining the decomposition rates of a particular peroxide in a
particular media [13].
Table 3.4
is presented to show how different solvents affect the rate of
decomposition of benzoyl peroxide into radicals.
Table
3.4
Decomposition of benzoyl peroxide in
various solvents at 79.8°C [15]
Solvent
|
Approximate% decomposition in
4 h
|
---|---|
Anisole
|
43.0
|
Benzene
|
50.0
|
Carbon tetrachloride
|
40.0
|
Chlorobenzene
|
49.0
|
Chloroform
|
44.0
|
Cyclohexane
|
84.0
|
Cyclohexene
|
40.0
|
Ethyl acetate
|
85.0
|
Ethylbenzene
|
46.0
|
Methyl benzoate
|
41.0
|
Methylene chloride
|
62.0
|
Nitrobenzene
|
49.0
|
Tetrachloroethylene
|
35.0
|
Toluene
|
50.0
|
Some initiators can function as both, thermal and
photoinitiators. Such an initiator, for instance, is
2,2′-azobisisobutyronitrile. Also, Engel and coworkers
[16] reported synthesis of an
initiator that can function both as a thermal free radical
initiator and a photoinitiator (see Sect. 3.2.4). It can be
illustrated as follows:
The claimed advantage of this initiator is that
it can be used to form block copolymers.
3.2.2 Bimolecular Initiating Systems
Decompositions of peroxides into initiating
radicals are also possible through bimolecular reactions involving
electron transfer mechanisms. Such reactions are often called
redox initiations and
can be illustrated as follows:
where, A is the reducing agent and ROOR′ is the peroxide.
The above can be illustrated on a decomposition
of a persulfate (an inorganic peroxide) by the ferrous ion:
Side reactions are possible in the presence of
sufficient quantities of reducing ions:
A redox reaction can also take place between the
peroxide and an electron acceptor:
Side reactions with an excess of the ceric ion
can occur as well:
Another example is a redox reaction of
t-butyl hydroperoxide with
a cobaltous ion [17]:
The cobaltic ion that forms can act as an
electron acceptor:
The cobaltic ion that forms can act as an
electron acceptor:
Side reactions can occur here too, such as:
Nevertheless, cobaltous ions form efficient redox
initiating systems with peroxydisulfate ions [18].
Tertiary aromatic amines also participate in
bimolecular reactions with organic peroxides. One of the unpaired
electrons on the nitrogen atom transfers to the peroxide link,
inducing decomposition. No nitrogen, however, is found in the
polymer. It is, therefore, not a true redox type initiation and the
amine acts more like a promoter of the decomposition
[19]. Two mechanisms were proposed
to explain this reaction. The first one was offered by Horner et
al. [19]:
The dimethyl aniline radical-cation, shown above,
undergoes other reactions than addition to the monomer. The benzoyl
radical is the one that initiates the polymerizations.
Presence of electron-releasing substituents on
diethyl aniline increases the rate of the reaction with benzoyl
peroxide [21]. This suggests that
the lone pair of electrons on nitrogen attack the positively
charged oxygen of the peroxide link [22].
By comparison to peroxides, the azo compounds are
generally not susceptible to chemically induced decompositions. It
was shown [23], however, that it
is possible to accelerate the decomposition of
α,α′-azobisisobutyronitrile by reacting it with
bis(-)-ephedrine-copper (II) chelate. The mechanism was postulated
to involve reductive decyanation of azobisisobutyronitrile through
coordination to the chelate [23].
Initiations of polymerizations of vinyl chloride and styrene with
α,α′-azobisisobutyronitrile coupled to aluminum alkyls were
investigated [24]. Gas evolution
measurements indicated some accelerated decomposition. Also,
additions of large amounts of tin tetrachloride to either
α,α′-azobisisobutyronitrile or to dimethyl-α,α′-azobisisobutyrate
increase the decomposition rates [25]. Molar ratios of
[SnCl4]/[AIBN] = 21.65 and
[SnCl4]/[MAIB] = 19.53 increase the rates by factors of
4.5 and 17, respectively. Decomposition rates are also enhanced by
donor solvents, such as ethyl acetate or propionitrile in the
presence of tin tetrachloride [25].
A bimolecular initiating system, based in
2,2′-azobisisobutyronitrile was reported by Michl and coworkers
[26]. It consists of weakly
solvated lithium in combination with the cyanopropyl radical (from
AIBN). The combination can initiate polymerizations of olefins. The
reaction was illustrated as follows:
3.2.3 Boron and Metal Alkyl Initiators of Free-Radical Polymerizations
These initiators were originally reported a long
time ago [27–29]. Oxygen plays an important role in the
reactions [30, 31]. It reacts with the alkyl boride under mild
conditions to form peroxides [32,
33]:
Initiating radicals apparently come from
reactions of these peroxides with other molecules of boron alkyls
[34, 35]. One postulated reaction mechanism can be
illustrated as follows [35]:
Another suggested reaction path is
[36]:
Catalytic action of oxygen was observed with
various organometallic compounds [35]. One example is dialkylzinc [37] that probably forms an active peroxide
[38]. The same is also true of
dialkylcadmium and of triethylaluminum [38]. Peroxide formation is believed to be an
important step in all these initiations. Initiating radicals,
however, do not appear to be produced from mere decompositions of
these peroxides [35].
3.2.4 Photochemical Initiators
This subject is discussed in greater detail in
Chap.
10, in the section on photo-cross-linking
reactions of coatings and films. A brief explanation is also
offered here because such initiations are used, on a limited scale,
in a few conventional preparation of polymers.
Many organic compounds decompose or cleave into
radicals upon irradiation with light of an appropriate wavelength
[38, 39]. Because the reactions are strictly light
and not heat induced, it is possible to carry out the
polymerizations at low temperatures. In addition, by employing
narrow wavelength bands that only excite the photoinitiators, it is
possible to stop the reaction by merely blocking out the light.
Among the compounds that decompose readily are peroxides, azo
compounds, disulfide, ketones, and aldehydes. A photodecomposition
of a disulfide can be illustrated as follows:
Today, many commercially prepared photoinitiators
are available. Some consists of aromatic ketones that cleave by the
Norrish reaction or are photoreduced to form free radicals. There
are also numerous other two and three component photoinitiating
systems. There are also those that decompose by irradiation with
visible light and make it possible to initiate the reactions with
longer wavelength light (see Chap.
10) Some examples of various photoinitiators are
given in Chap.
10. Many others can be found in the
literature.
As an example can be sited the work by
Barner-Kowollik and coworkers studied the photoinitiation process
in methyl methacrylate polymerization, using high-resolution
electro spray-mass spectrometry [40]. The polymerization was conducted using a
pulsed laser at temperatures ≤0°C in the presence of the
photoinitiators 2,2-dimethoxy-2-phenylacetophenone, benzoin,
benzil, benzoin ethyl ether, and 2,2-azobisisobutylnitrile. They
identified the termination products, both combination and
disproportionation with high accuracy. Both the benzoyl and acetal
fragments generated as a result of
2,2-dimethoxy-2-phenylacetophenone photocleavage were found to
initiate and highly likely terminate the polymerization. Both the
benzoyl and ether fragments produced as a result of benzoin
photocleavage were found to act as initiating and probable
terminating species, indicating that the ether radical fragment
does not act exclusively as a terminating species.
3.2.5 Initiation of Polymerization with Radioactive Sources and Electron Beams
Different radioactive sources can initiate
free-radical polymerizations of vinyl monomers. They can be
emitters of gamma rays, beta rays, or alpha particles. Most useful
are strong gamma emitters, such as 60Co or
90Sr. Electron beams from electrostatic accelerators are
also efficient initiators. The products from irradiation by
radioactive sources or by electron beams are similar to but not
identical to the products of irradiation by ultraviolet light.
Irradiation by ionizing radiation causes the excited monomer
molecules to decompose into free radicals. Ionic species also form
from initial electron captures. No sensitizers or extraneous
initiating materials are required. It is commonly accepted that
free radicals and ions are the initial products and that they act
as intermediate species in these reactions. There is still
insufficient information, however, on the exact nature of all of
these species [38, 39]. The polymerizations are predominantly by a
free-radical mechanism with some monomers and by an ionic one with
others [38, 39].
3.3 Capture of Free Radicals by Monomers
Once the initiating radical is formed, there is
competition between addition to the monomer and all other possible
secondary reactions. A secondary reaction, such as a recombination
of fragments, as shown above, can be caused by the cage effect of
the solvent molecules [41]. Other
reactions can take place between a radical and a parent initiator
molecule. This can lead to the formation of different initiating
species. It can, however, also be a dead end as far as the
polymerization reaction is concerned.
After the initiating radical has diffused into
the proximity of the monomer, the capture of the free radical by
the monomer completes the step of initiation. This is a
straightforward addition reaction, subject to steric effects:
The unpaired electron of the radical is believed
to be in the pure p-orbital of a planar, sp 2, carbon atom.
Occasionally, however, radicals with sp 3 configuration appear to
form [42–44].
Using quantum chemical calculations it was
demonstrated that nucleophilic and electrophilic alkyl or aryl
radicals attack alkenes following a tetrahedral trajectory
[45, 46]:
This means that only substituents Y at the
attacked olefinic carbon exert large steric effects [47]. In addition to the steric effects, the
rates of addition of strongly nucleophilic or electrophilic
radicals are governed mainly by polar effects of the substituents
Rx, Y, and Z [48]. In
borderline cases, however, the stabilities of the adducts and
products tend to dominate [47].
Also, it was demonstrated that acyclic radicals
can react with high stereoselectivity [45]. In order for the reactions to be
stereoselective, the radicals have to adopt preferred conformations
where the two faces of the prochiral radical centers are shielded
to different extents by the stereogenic centers. Giese and
coworkers [49] demonstrated with
the help of Electron Spin Resonance studies that ester-substituted
radicals with stereogenic centers in β-positions adopt preferred
conformations that minimize allylic strain [49] (shown below). In these conformations, large
(L) and medium sized substituents (M) shield the two faces. The
attacks come preferentially from the less shielded sides of the
radicals. Stereoselectivity, because of A-strain conformation, is
not limited to ester-substituted radicals [50]. The strains and steric control in reactions
of radicals with alkenes can be illustrated as follows
[50]:
The above considerations can be illustrated on
initiation by benzoyl peroxide, a commonly used initiating
compound. The half-life of the initial benzoyloxy radicals from
decompositions of benzoyl peroxide is estimated to be
10−4 to 10−5 s. Past that time, they
decompose into phenyl radicals and carbon dioxide [49]. This is sufficient time for the benzoyloxy
radicals to be trapped by fast-reacting monomers. Slow-reacting
monomers, however, are more likely to react with the phenyl
radicals that form from the elimination reaction. In effect, there
are two competing reactions [50]:
1.
Decomposition of the free radical:
2.
Two types of radicals can add to the monomer
(where x represents any typical substituent of vinyl monomers, such
as halogens, or esters, or aromatic groups, or nitriles,
etc.):
The ratio of the rates of the two reactions,
K″/K′, (or K′″/K′) depends upon the reactivity of the
monomers. It is shown in Table 3.5 [51, 52]. The
benzoyloxy radical is used in this table as an illustration. A
similar comparison is possible for a redox initiating system. An
initiating sulfate radical ion from a persulfate initiator can
react with another reducing ion or add to the monomer:
Table
3.5
Relative reactivities of the benzoyloxy
radical at 60°C
Monomer
|
Structure
|
K′/K″ (mol/L)
|
---|---|---|
Acrylonitrile
|
0.12
|
|
Methyl methacrylate
|
0.30
|
|
Vinyl acetate
|
0.91
|
|
Styrene
|
2.50
|
|
2,5-dimethylstyrene
|
5.0
|
In Table 3.6 are shown the relative reaction rates of
SO4• with some monomers at 25°C
[53–57]. As explained, the rate of addition of a
radical to a double bond is affected by steric hindrance from bulky
substituents. Polar effect, such as dipole interactions also
influence the rate of addition.
Forbes and Yashiro studied the addition of the
initiating radials to methyl methacrylate in liquid supercritical
carbon dioxide [58]. They
demonstrated that the rate of addition of the initiating radicals
to the monomers, k
add values, can be measured in liquid
CO2.
Phenyl or methyl groups located on the carbon
atom that is under a direct attack by a free radical can be
expected to interfere sterically with the approach. For instance,
due to steric hindrance, trans-β-methylstilbene is more reactive
toward a radical attack than is its cis isomer [58]. Yet, the trans isomer is more stable of the two.
While 1,1-disubstituted olefins homopolymerize readily, the
1,2-disubstituted olefins are hard to homopolymerize
[59]. Some exceptions are vinyl
carbonate [61] and maleimide
derivatives [62]. Also,
perfluoroethylene and chlorotrifluoroethylene polymerize readily.
Table 3.6
shows the relative reaction rates of SO4• with some monomers at 25°C
[50].
Homopolymerizations of diethyl fumarate by
free-radical mechanism were reported [63]. The M n was found to be 15,000.
The same is true of homopolymerizations of several other dialkyl
fumarates and also dialkyl maleates [64–66]. The
polymerization rates and the sizes of the polymers that form
decrease with increases in the lengths of linear alkyl ester
groups. There is, however, an opposite correlation if the ester
groups are branched. Also, the maleate esters appear to isomerize
to fumarates prior to polymerization [66].
3.4 Propagation
The transition state in a propagation reaction
can be illustrated as follows:
In the above transition state, the macroradical
electron is localized on the terminal carbon. Also, the two π
electrons of the double bond are localized at each olefinic carbon.
Interaction takes place between p-orbital of the terminal atom in
the active polymer chain with associated carbon of the monomer.
This results in formation of σ-bonds [67].
The rate of the propagation reaction depends upon
the reactivity of the monomer and the growing radical chain. Steric
factors, polar effects, and resonance are also important factors in
the reaction.
Another factor that can affect the rate of
propagation is interaction between propagating radicals. Siegmann
and Beuermann [60] studied the
rate of propagation of 1H,1H,2H,2H-tridecafluorooctyl methacrylate and
compared it to the rate of propagation of methyl methacrylate. They
observed that k
p for 1H,1H,2H,2H-tridecafluorooctyl methacrylate
polymerization is 1.9 times that of k p for methyl methacrylate.
They concluded that this higher rate is due to less interactions
occurring between the propagating macroradicals.
Bowman and coworkers studied the impact of
intermolecular and intramolecular interactions on the
polymerization kinetics of monoacrylates [69]. They carried out polymerization studies in
the presence of extensive amounts of solvent. This was an attempt
to elucidate the effects of intermolecular interactions, such as
bulk medium polarity, π–π stacking, and hydrogen bonding and
characterize the contribution of intramolecular conformational
effects to monomer reactivity. Solution polymerization kinetics of
various monomers were measured in the presence of 95 wt%
1,4-dioxane. The results were compared to bulk polymerization
kinetics. The studies revealed that aliphatic acrylates such as
hexyl acrylate exhibit approximately two to threefold reduction in
reactivity upon dilution. Monomers characterized by only
hydrogen-bonding features such as hydroxyethyl acrylate exhibit an
8- to 12-fold reduction upon dilution. Monomers possessing only
aromatic ring stacking interactions such as phenyl acrylate exhibit
approximately a five to tenfold reduction upon dilution under
similar conditions. Even at a concentration of 5 wt% monomer in
1,4-dioxane, there were approximately two to fivefold differences
in reactivity observed between various acrylates. Bowman and
coworkers attributed these reactivity differences between various
acrylates, upon extensive dilution, solely to intramolecular
interactions [69].
3.4.1 Steric, Polar, and Resonance Effects in the Propagation Reaction
The steric effects depend upon the sizes of the
substituents. The resonance stabilization of the substituents has
been shown to be in the following order [70]:
The reactivities of the propagating
polymer-radicals, however, exert greater influence on the rates of
propagation than do the reactivities of the monomers. Resonance
stabilization of the polymer-radicals is a predominant factor. This
fairly common view comes from observations that a methyl radical
reacts at a temperature such as 60°C approximately 25 times faster
with styrene than it does with vinyl acetate [72]. In homopolymerizations of the two monomers,
however, the rates of propagation fall in an opposite order. Also,
poly(vinyl acetate)-radicals react 46 times faster with
n-butyl mercaptan in
hydrogen abstraction reactions than do the polystyrene-radicals
[71]. The conclusion is that the
polystyrene radicals are much more resonance stabilized than are
the poly(vinyl acetate)-radicals. Several structures of the
polystyrene-radicals are possible due to the conjugation of the
unpaired electrons on the terminal carbons with the adjacent
unsaturated groups. These are resonance hybrids that can be
illustrated as follows:
There is not such opportunity, however, for
resonance stabilization of the poly(vinyl acetate) radicals because
oxygen can accommodate only eight electrons. The effect of steric
hindrance on the affinity of a methyl radical is illustrated in
Table 3.7
[56, 57].
In vinyl monomers, both olefinic carbons are
potentially subject to free-radical attack. Each would give rise to
a different terminal unit:
The newly formed radicals can again potentially
react with the next monomer in two ways. This means that four
propagation reactions can occur:
Contrary to the above shown four propagation
modes, a “head to tail” placement shown in (3.1), strongly
predominates. This is true of most free radical vinyl
polymerizations. It is consistent with the localized energy at the
α-carbon of the monomer. Also, calculations of resonance
stabilization tend to predict head to tail additions
[68].
The free-radical propagation reactions that
correspond to conversions of double bonds into single bonds are
strongly exothermic. In addition, the rates increase with the
temperature. It is often assumed that the viscosity of the medium,
or change in viscosity during the polymerization reaction does not
affect the propagation rate or the polymer growth reaction. This is
because it involves diffusion of small monomer molecules to the
reactive sites. Small molecules, however, can also be impeded in
their process of diffusion. This can impede the growth rate
[50].
During chain growth, the radical has a great deal
of freedom with little steric control over the manner of monomer
placement. Decrease in the reaction temperature, however, lowers
mobility of the species and increases steric control over
placement. This is accompanied by an increase in stereoregularity
of the product [70, 71]. The preferred placement is trans – trans, because of lower energy
required for such placement. As a result, a certain amount of
syndiotactic arrangement is observed in polymerizations at lower
temperatures [72]. Trans–trans configurations (with respect to
the carbon atoms in the chains) yield zigzag backbones. This was
predicted from observations of steric effects on small molecules
[74, 75]. It was confirmed experimentally for many
polymers, such as, for instance, in the formation of
poly(1,2-polybutadiene) [74] and
poly(vinyl chloride) [72]. Also,
in the free-radical polymerizations of methyl methacrylate,
syndiotactic placement becomes increasingly dominant at lower
temperatures. Conversely, the randomness increases at higher
temperatures [74]. The same is
true in the free-radical polymerization of halogenated vinyl
acetate [75].
One proposed mechanism for the above is as
follows. The least amounts of steric compression within
macromolecules occur during the growth reactions if the ultimate
and the penultimate units are trans to each other. Also, if the lone
electrons face the oncoming monomers during the transition states
[80, 81], as shown below, syndiotactic placement
should be favored:
While the above model explains the formation of
syndiotactic poly(methyl methacrylate), possible interactions
between the free radicals on the chain ends and the monomers are
not considered. Such interactions, however, are a dominant factor
in syndiotactic placement, if the terminal carbons are sp 2 planar in structures
[75].
3.4.2 Effect of Reaction Medium
There were some early reports that reaction media
influences the polymerizations of vinyl chloride in aliphatic
aldehydes at 50°C [80,
81]. This was not confirmed in
subsequent studies [82–84].
Subsequently, the rate of polymerization was shown to be influenced
by the pH of the reaction medium in polymerizations of monomers
such as methacrylic acid (MAA) [85, 88]. Also,
the rate of polymerization and solution viscosities increase in
polymerizations of acrylamide and acrylic acid with an increase in
water concentration [83]. It is
not quite clear whether this is due to increases in the speeds of
propagations or due to decreases in the termination rates. In the
free-radical polymerization of vinyl benzoate, the rate of
propagation varies in different solvents in the following order
[88, 89]:
Similarly, the rate of photopolymerization of
vinyl acetate is affected by solvents [88]. In most cases, however, the rate of
polymerization is proportional to the square root of the initiator
concentration and to the concentration of the monomer
[5].
De Sterck and coworkers [90] studied solvent effect on tacticity of
methyl methacrylate in free-radical polymerization. They observed
that solvents CH3OH and (CF3)3COH,
which are H-bonded with the carbonyl oxygens and are located on the
same side of the backbone of the growing polymer radical hinder the
formation of isotactic poly(methyl methacrylate) to some extent.
Methanol is less effective in reducing the isotacticity because of
its small size and also because of the relatively loose hydrogen
bonds with the carbonyl oxygens.
There is a controversial suggestion that the
solvent affects the propagation step in some reactions by forming
“hot” radicals [91]. These
radicals are supposed to possess higher amounts of energy. At the
moment of their formation, they obtain surplus energy from the heat
of the reaction and from the activation energy of the propagation
reaction. This is claimed to provide the extra energy needed to
activate the next chain propagation step. The surplus energy may
affect the polymerization kinetics if the average lifetime of the
hot radicals is sufficient for them to react with the monomer
molecules. This surplus energy is lost by the hot radicals in
collisions with monomer and solvent molecules. There is a
difference in the rate constants of propagation for hot and
ordinary radicals so two different reaction schemes were written
[91]:
Propagation by ordinary radicals:
Propagation by hot radicals:
Energy transfer processes:
where, R* is the symbol for hot radicals.
The rate expression for the polymerization is
then written as follows [92]:
where, K
x = k
2(2k
1 f/k 4)1/2,
γ = k 1*/k 2* and γ′ = k 3*/k 2*, S = solvent,
[M] = monomer
In comparing free-radical polymerizations of
ethyl acrylates in benzene and in dimethyl formamide at 50°C
[218] the rates were found to be
proportional to the square roots of the initiator concentration.
They were not proportional, however, to the concentrations of the
monomer. This was interpreted in terms of hot radicals
[93].
Similar results, however, were interpreted by
others differently. For instance, butyl acrylate and butyl
propionate polymerizations in benzene also fail to meet ideal
kinetic models. The results, however, were explained in terms of
termination of primary radicals by chain transferring (see
Sect. 3.5 for explanation of chain
transferring).
3.4.3 Ceiling Temperature
For most free-radical polymerization reactions,
there are some elevated temperatures at which the chain-growth
process becomes reversible and depropagation takes place:
where, k d·p is
the rate constant for depropagation or depolymerization. The
equilibrium for the polymerization–depolymerization reaction is
temperature dependent. The reaction isotherm can be written:
In the above equation ΔF 0 is the free energy of
polymerization of both, monomer and polymer, in appropriate
standard states [88]. The standard
state for the polymer is usually solid (amorphous or partly
crystalline). It can also be a one molar solution. The monomer is a
pure liquid or a one molar solution. The relationships of monomer
concentration to heat content, entropy, and free energy are shown
by the following expression. This applies over a wide range of
temperatures [5].
In the above equation, T c is the ceiling temperature for the
equilibrium monomer
concentration. It is a
function of the temperature of the reaction. Because the heat
content is a negative quantity, the concentration of the monomer
(in equilibrium with polymer) increases with increasing
temperatures. There are a series of ceiling temperatures that
correspond to different equilibrium monomer concentrations. For any
given concentration of a monomer in solution, there is also some
upper temperature at which polymerization will not proceed. This,
however, is a thermodynamic approach. When there are no active
centers present in the polymer structure, the material will appear
stable even above the ceiling temperature in a state of metastable
equilibrium.
The magnitude of the heat of polymerization of
vinyl monomers is related to two effects: (1) Steric strains that
form in single bonds from interactions of the substituents. These
substituents, located on the alternate carbon atoms on the
polymeric backbones, interfere with the monomers entering the
chains. (2) Differences are in resonance stabilization of monomer
double bonds by the conjugated substituents [70].
Most 1,2 disubstituted monomers, as stated
earlier, are difficult to polymerize. It is attributed to steric
interactions between one of the two substituents on the vinyl
monomer and the β-substituent on the ultimate unit of the polymeric
chain [94]. A strain is also
imposed on the bond that is being formed in the transition
state.
The propagation reaction usually requires only an
activation energy of about 5 kcal/mol. As a result, the rate
does not vary rapidly with the temperature. On the other hand, the
transfer reaction requires higher activation energies than does the
chain-growth reaction. This means that the average molecular weight
will be more affected by the transfer reaction at higher
temperature. When allowances are made for chain transferring, the
molecular weight passes through a maximum as the temperature is
raised. At temperatures below the maximum, the product molecular
weight is lower because the kinetic chain length decreases with the
temperature. Above the maximum, however, the product molecular
weight is also lower with increases in the temperature. This is due
to increase in the transfer reactions. The above assumes that the
rate of initiation is independent of the temperature. The
relationship of the kinetic chain length to the temperature can be
expressed as follows [5]:
where, E P,
E T, and
E I are energies
of propagation, termination, and initiation, respectively. A large
E I means that
if the temperature of polymerization is raised, the kinetic chain
length decreases. This is affected further by a greater frequency
of chain transferring at higher temperatures. In addition, there is
a possibility that disproportionation may become more
significant.
3.4.4 Autoacceleration
When the concentrations of monomers are high in
solution or bulk polymerizations, typical auto-accelerations of the
rates can be observed. This is known as the gel effect or as the Trammsdorff effect, or also, as the
Norrish – Smith effect [66]. The effect has been explained as being
caused by a decrease in the rate of termination due to increased
viscosity of the medium. Termination is a reaction that requires
two large polymer-radicals to come together and this can be impeded
by viscosity. At the same time, in propagation the small molecules
of the monomer can still diffuse for some time to the radical sites
and feed the chain growth.
One should not mistake an acceleration of the
polymerization reaction due to a rise in the temperature under
nonisothermal conditions for a true gel effect from a rise in
viscosity. The gel effect can occur when the temperature of the
reaction is kept constant.
A critical analysis of the gel effect suggests
that the situation is complicated. In some polymerizations, three
different stages appear to be present when R P/[M][I]1/2 is
plotted against conversion or against time [95]. The plot indicates that during the first
stage there is either a constant or a declining rate and during the
second stage there is autoacceleration. During the third stage,
there is again a constant or a declining rate [95].
Numerous publication made a substantial case for
associating and/or attributing the gel effect to entanglement of
polymerizing chain radicals, resulting in a marked reduction in the
termination rate parameter, k T. This was often done by
using the assumption that in the neighborhood of the gel effect
k T is
controlled by polymer self-diffusion, which in turn exhibits
entangled polymer dynamics. O’Neil et al. [96], however, argued against that opinion. They
carried out a series of experiments involving bulk polymerizations
of methyl methacrylate and styrene and feel that their data
contradicts this widely held belief that the gel effect onset is
related to the formation of chain entanglements. The experimental
conditions used were such that they tended to delay or eliminate
the formation of chain entanglement. These conditions were high
initiator and/or chain transferring agent concentrations and
additions of low molecular weight polymers prior to the reactions.
The results indicated that the gel effect occurs readily in the
absence of entanglement and that delaying the onset of
entanglements does not necessarily delay the onset of the gel
effect. Also, critical examination of the molecular weights
produced in these experiments indicated values that were too low
for entanglement formation in solution (polymer plus monomer) and
sometimes even in bulk polymer, not only at the onset but also
throughout the gel effect [96].
O’Neil et al. [96], found that
even under conditions where entanglements are likely to exist, the
gel effect onset does not correlate with polymer molecular weight
of the chains produced in a manner consistent with entanglement
arguments. Whether the kinetics during the gel effect may be
affected by entanglements was left uncertain.
3.4.5 Polymerization of Monomers with Multiple Double Bonds
Polymerizations of monomers with multiple double
bonds yield products that vary according to the locations of these
bonds with respect to each other. Monomers with conjugated double
bonds, such as 1,3-butadiene and its derivatives, polymerize in two
different ways. One way is through one of the double bonds only.
Another way is through both double bonds simultaneously. Such 1,4
propagation is attributable to the effect of conjugation and
hybridization of the C2–C3 bond that involves
sp 2 hybrid
orbitals [97]. All three modes of
propagation are possible in one polymerization reaction so that the
product can, in effect, be a copolymer. The 1,4, 1,2, and 3,4,
placement in propagations can be illustration as follows:
The polymerizations and copolymerizations of
various conjugated dienes are discussed in Chap.
6.
3.4.5.1 Ring Forming Polymerization
Propagation reactions of unconjugated dienes can
proceed by an intra-intermolecular process. This usually results in
ring formation or in cyclopolymerization. It can be
illustrated as follows:
where, X can designate either a carbon or a heteroatom. An example
of such a polymerization is a free-radical polymerization of
quaternary diethyldiallylamine [98]:
Another example is a polymerization of
2,6-disubstituted 1,6-heptadiene [103]:
where, R=COOC2H5, COOCH3, or
COOH.
The intra-intermolecular propagations can result
in ring structures of various sizes. For instance, three-membered
rings can form from transanular polymerizations of
bicycloheptadiene [100,
101]:
Four-membered rings form in free-radical
polymerization of perfluoro-1,4-pentadiene [103]. The size of the ring that forms depends
mainly on the number of atoms between the double bonds:
Formation of many five-membered rings is also
known. One example is a polymerization of 2,3-dicarboxymethyl-1,6
hexadiene [98]:
where, R=COOCH3.
The polymer that forms, shown above, is
cross-linked, but spectroscopic analysis shows that 90% of the
monomer placement is through ring formation [104]. Formations of six-membered rings are also
well documented. Two examples were shown above in the
polymerization of a quaternary diethyldiallylamine and in the
polymerization of 2,6 disubstitued, 1,6-heptadiene. Many other
1,6-heptadienes yield linear polymers containing six-membered rings
[105].
This tendency to propagate intra-intermolecularly
by the unconjugated dienes is greater than can be expected from
purely statistical predictions [106]. Butler suggested that this results from
interactions between the olefinic bonds [107–109].
Ultraviolet absorption spectra of several unconjugated diolefins
does show bathochromic shifts in the absorption maxima relative to
the values calculated from Woodward’s rule [102, 104].
This supports Butler’s explanation [107, 109].
3.5 The Termination Reaction
The termination process in free-radical
polymerization is caused, as was shown early in this chapter, by
one of three types of reactions: (1) a second order radical-radical
reaction, (2) a second order radical-molecule reaction, and (3) a
first order loss of radical activity.
The first reaction can be either one of
combination or of disproportionation. In a combination reaction,
two unpaired spin electrons, each on the terminal end of a
different polymer-radical, unite to form a covalent bond and a
large polymer molecule. In disproportionation, on the other hand,
two polymer-radicals react and one abstracts an atom from other
one. This results in formation of two inactive polymer molecules.
The two differ from each other in that one has a terminal saturated
structure and the other one has a terminal double bond. Usually,
the atom that is transferred is hydrogen.
It was suggested [111] that a basic rule of thumb can be applied
to determine which termination reaction predominates in a typical
homopolymerization. Thus, polymerizations of 1,1-disubstituted
olefins are likely to terminate by disproportionation because of
steric effects. Polymerizations of other vinyl monomer, however,
favor terminations by combination unless they contain particularly
labile atoms for transferring. Higher activation energies are
usually required for termination reactions by disproportionation.
This means that terminations by combination should predominate at
lower temperatures.
For a polymer radical that simply grows by adding
monomeric units and still possesses an active center after the
growth, the number of monomeric units (r) added to a radical center during the
time interval t, according to Tobita [112], conforms to the following Poisson
distribution:
where θ is the expected
number of monomeric units added to a radical center, given by
where k p is the
propagation rate coefficient and [M] is the monomer concentration.
If the number of the added monomeric units, r, is large enough, one can approximate
that r ≅ θ. For bimolecular termination
reactions that are independent of chain length, the required time
for bimolecular termination between a particular radical pair is
also given by the following most probable distribution
[112]:
where ξ = k t/(k p[M]νN A), and k t is the bimolecular
termination rate coefficient. The imaginary time for chain stoppage
by bimolecular termination must be considered for all radical pairs
that exist in the reaction medium [112].
The third type of a termination reaction is chain
transferring. Premature termination through transferring results in
a lower molecular weight polymer than can be expected from other
termination reactions. The product of chain transferring is an
inert polymer molecule and, often, a new free radical capable of
new initiation. If, however, the new radical is not capable of
starting the growth of a new chain, then this is degenerative chain transferring. It is also referred
to as a first
- order termination reaction. The molecules that accept
the new radical sites (participate in chain transferring) can be
any of those present in the reaction medium. This includes
solvents, monomer molecules, inactive polymeric chains, and
initiators.
The ease with which chain transferring takes
place depends upon the bond strength between the labile atoms that
are abstracted and the rest of the molecule to which they are
attached. For instance, chain transferring in methyl methacrylate
polymerization to the solvent occurs in the following order
[115]:
The rate of a chain transferring reaction is,
where, k tr is
the chain transferring constant in a reaction:
Examples of molecules that have particularly
labile atoms and contribute readily to chain-transferring are
mercaptans and halogen compounds, such as chloroform, carbon
tetrachloride, etc.
A polymer that was prematurely terminated in its
growth by chain transferring may be a telomer. In most cases of telomer
formation, the newly formed radical and the monomer radical are
active enough to initiate new chain growth. Thus, the life of the
kinetic chain is maintained.
An illustration of a telomerization reaction can be
free-radical polymerization of ethylene in the presence of
chloroacetyl chloride:
Chain transferring is affected by temperature but
not by changes in the viscosity of the reaction medium
[115]. When a transfer takes
place to a monomer, it is independent of the polymerization rate
[116, 117]. When, however, transfer takes place to
the initiator, the rate increases rapidly [118].
A chain transferring reaction to a monomer can be
illustrated as follows:
A transfer reaction can also occur from the
terminal group of the polymer-radical to a location on the
polymeric backbone. This is known as backbiting:
The new free-radical site on the polymer backbone
starts chain growth that results in formation of a branch. The same
reaction can take place between a polymer-radical and a location on
another polymer chain. In either case, fresh chain growth results
in formation of a branch.
Whether chain transferring can take place to an
initiator depends upon the initiator’s chemical structure. It was
believed in the past that chain transferring to
α,α′-azobisisobutyronitrile does not occur. Later it was shown that
chain transferring to this initiator does occur as well, at least
in the polymerizations of methyl methacrylate [118, 119].
The amount of chain transferring that takes place
to monomers is usually low because the reaction requires breaking
strong carbon-hydrogen bonds. Monomers, however, such as vinyl
chloride and vinyl acetate have fairly large chain transferring
constants. In the case of vinyl acetate, this is attributed to the
presence of an acetoxy methyl group. This explanation, however,
cannot be used for vinyl chloride.
The chain transferring constants, are usually
defined as:
The values can be found in handbooks and other
places in the literature. Presence of chain transferring agents in
a polymerization reaction requires redefining the degree of
polymerization to include the chain termination terms. The number
average degree of polymerization has to be written as follows:
It can also be expressed in terms of the chain
transferring constants as follows:
This can also be written in still another form:
When a polymerization reaction is conducted in a
concentrated solution, or in complete absence of a solvent, the
viscosity of the medium increases with time, (unless the polymer
precipitates out). This impedes all steps in the polymerization
process, particularly the diffusions of large polymer-radicals
[54]. The decreased mobility of
the polymer-radicals affects the termination process. It appears
that this is common to many, though not all, free-radical
polymerizations. All molecular processes in the termination
reactions are not fully understood, particularly at high
conversions [119] This is a
complex process that consists of three definable steps. These can
be pictured as follows. First, two polymer radicals migrate
together by means of translational diffusion. Second, the radical
sites reorient toward each other by segmental diffusion. Third, the
radicals overcome the small chemical activation barriers and react.
The termination reaction is, therefore, diffusion controlled. At
low concentrations, this will be segmental diffusion while at
medium or high concentrations it will be translational
diffusion.
Present theories of terminations suggest that at
intermediate conversions, terminations are dominated by
interactions between short chains formed by transfer and entangled
long chains [121]. When
terminations are diffusion controlled, most termination events
involve two highly entangled chains whose ends move by the
“reaction-diffusion” process [119]. In this process, terminations occur
because of the propagation-induced diffusion of the chain ends of
growing macroradicals. This means that the rates of terminations
depend upon the chain lengths [113].
Diffusion theories have been proposed that relate
the rate constant of termination to the initial viscosity of the
polymerization medium. The rate-determining step of termination,
the segmental diffusion of the chain ends, is inversely
proportional to the microviscosity of the solution [123]. Yokota and Itoh [124] modified the rate equation to include the
viscosity of the medium. According to that equation, the overall
polymerization rate constant should be proportional to the square
root of the initial viscosity of the system.
The number average termination rate constants in
a methyl methacrylate polymerization were measured with an in-line
ESR spectrometer. This was done by observing the radical decay
rates [120]. The results are in
disagreement with the concept of termination by
propagation-diffusion that is expected to be dominant at high
conversion rates. Instead, the termination rate constants decrease
dramatically in the posteffect period at high conversions.
Actually, a fraction of the radicals were found trapped during the
polymerization. Thus, there are two types of radicals in the
reaction mixture, trapped and free radicals. In the propagations
and termination reactions, the two types of radical populations
have very different reactivities [120].
Shipp and coworkers [120] described a method for analyzing the chain
length dependence of termination rate coefficients of the reacting
radicals in low conversion free radical polymerizations. Their
method involves comparing experimental molecular weight
distributions of polymers formed in pulsed laser photolysis
experiments with those predicted by kinetic simulation. The method
is enabled by direct measurements of the concentration of radicals
generated per laser pulse. Knowledge of the radical concentrations
should mean that the only unknowns in the simulations are the
termination rate coefficients. They concluded that the analysis
demonstrates the need for chain length dependent termination rate
constants in describing polymerization kinetics.
Free-radical photopolymerizations (see
Chap.
10) of multifunctional acrylic monomers result in
cross-linked polymeric networks. The kinetic picture of such
polymerizations varies from ordinary linear polymerization because
the diffusion of free radicals and functional groups becomes
severely restricted. This causes growing polymer chains to rapidly
cyclize and cross-link into clusters (microgels). The clusters
become linked up into networks. Many free radicals become trapped,
but terminations take place by combinations and by chain
transferring. The cumulative chain length in such polymerizations
can be calculated from the following equation [125]:
where, χ is the conversion
of functional groups and n
m0 is the initial number of functional groups and
n rg is the
total number of radicals generated.
3.6 Copolymerization
If more than one monomer species is present in
the reaction medium, a copolymer or an interpolymer can result from
the polymerization reaction. Whether the reaction products will
consist of copolymers or just a mixture of homopolymers of both,
however, depends largely upon the reactivity of the monomers. A
useful and a simplifying assumption in kinetic analyses of
free-radical copolymerizations is that the reactivity of polymer
radicals is governed entirely by the terminal monomer units
[52]. For instance, a growing
polymer radical that contains a methyl methacrylate terminal unit,
is considered, in terms of reactivity, as a poly(methyl
methacrylate) radical. This assumption although not always adequate
[52] can be used to predict
satisfactorily the behavior of many mixtures of monomers. Based on
this assumption, the copolymerization of a pair of monomers
involves four distinct growth reactions and two types of polymer
radicals.
3.6.1 Reactivity Ratios
In a reaction of two monomers, designated as
M1 and M2, four distinct reactions can be
written as follows:
The ratios of k 11/k 12 and k 22/k 21 are called monomer reactivity ratios. They can be written as
follows:
The relationship can be express in terms of the
ratio of the monomers, [M1]/[M2] that end up
in the formed polymer, R
p:
where, R m is
equal to [M1]/[M2]. Table 3.8 illustrates a few
typical reactivity ratios taken from the literature. Many more can
be found [128].
These reactivity ratios represent the relative
rates of reactions of polymer radicals with their own monomers vs.
that with the comonomers. When r 1 > 1, the radical
~M1• is reacting with monomer M 1 faster than it is with
the comonomer M
2. On the other hand, when r 1 < 1, the opposite is
true. Based on the r
values, the composition of the copolymers can be calculated
from a copolymerization
equation [52] shown below:
In an ideal copolymerization reaction,
If r
1 and r
2 values are equal to or approach zero, each polymer
radical reacts preferentially with the other monomer. This results
in an alternating copolymer, regardless of the composition of the
monomer mixture. That is, however, a limiting case. In the majority
of instances, r
1 × r
2 is greater than zero and less than one. When the
polymer radicals react preferentially with their own monomer, and
r
1 × r
2 > 1, then mainly a mixture of homopolymers forms
and only some copolymerization takes place.
Reactivity of vinyl monomers is very often
determined experimentally by studying copolymerizations. Values of
many free-radical reactivity ratios have been tabulated for many
different monomer pairs [126].
Also, the qualitative correlations between copolymerization data
and molecular orbital calculations can be found in the literature
[136].
Some general conclusion about monomer reactivity
toward attacking radicals was drawn by Mayo and Walling
[130]:
1.
The alpha substituents on a monomer have the
effect of increasing reactivity in the following order.
2.
The effect of a second alpha substituent is
roughly additive.
Giese and coworkers [131–136]
developed a special technique for studying the effects of the
substituents upon the relative reactivity of vinyl monomers toward
free radicals. Briefly it is as follows. Free radicals are produced
by reducing organomercury halides with sodium borohydride. The
radicals undergo competitive additions to pairs of various
substituted olefins. The adducts are in turn trapped by hydrogen
transfers from the formed organomercury hydrides. Relative
quantities of each product are then determined
This method was applied in copolymerization of
acrylonitrile and methyl acrylate [129]. It showed that the ratios of the rate
constants for each of the two monomers are independent of their
concentrations.
Copolymerization reactions are affected by
solvents. One example that can be cited is an effect of addition of
water or glacial acetic acid to a copolymerization mixture of
methyl methacrylate with acrylamide in dimethyl sulfoxide or in
chloroform. This causes changes in reactivity ratios
[131]. Changes in r values that result from changes in
solvents in copolymerizations of styrene with methyl methacrylate
are another example [133,
134]. The same is true for
styrene acrylonitrile copolymerization [132]. There are also some indications that the
temperature may have some effect on the reactivity ratios
[135], at least in some
cases.
3.6.2 Q and e Scheme
Though molecular orbital calculations allow
accurate predictions of reactivity ratios [133], many chemists also rely upon the
Price–Alfrey Q
– e equations [140]. These are based on: (1) the polarity of
the double bonds of the monomers or measures of the propagating
chain ends, (2) mesomerism of the substituents with the double
bonds or with the chain ends, and (3) the steric hindrance of the
substituents. This relationship is expressed in the following
equation [149]:
it can also be written as follows:
where K 12
represents the rate constant for the reaction of the propagating
radical ~M1• with monomer M2, P1
represents the general reactivity of the polymer radical with the
terminal unit of monomer M1, Q1 and
Q2 are the reactivities of the monomer M1 and
M2, and e
1 and e
2 are measures the polar characters of the
monomers.
It is possible to calculate the Q and e values from r 1 and r 2, or, conversely,
r values can be obtained
from the Q and e values. The relationship is as
follows [136]:
The Q and
e scheme is based on a
semiempirical approach. Nevertheless, some attempts were made to
develop theoretical interpretations. Thus, Schwann and Price
[141] developed the following
relationship:
In the above equation, q represents the resonance of
stabilization (kcal/mol), ε
is the electrical charge of the transition state, and γ is the distance between the centers
of the charge of the radical and the monomer, D stands for the effective dielectric
constant of the reaction field. The values of q and e are derived by calculation. In
addition, more rigorous molecular orbital calculations
[138] show a relationship between
Q and the localized energy
of a monomer, and between e
and the electron affinity. Also, a scale of Q and e values was deduced from essentially
molecular orbital considerations [143]. In addition, a Huckel treatment of the
transition state for the monomer–radical reaction in a free-radical
copolymerization was developed [142]. The resulting reactivity ratios compared
well with those derived from the Q and e scheme. This scheme is regarded by
some as a version of the molecular orbital approach [145]. Nevertheless, the scheme should only be
considered as an empirical one. The precision of calculating
Q and e values can be poor because steric
factors are not taken into account. It is good, however, for
qualitative or semiquantitative results.
A revised reactivity scheme was proposed by
Jenkins, that he called U,V scheme [148]. It is claimed to be more accurate and
also capable of application to both copolymerizations and to
transfer reactions. The scheme retains much of the format of the
Q and e scheme. In this one, the intrinsic
radical reactivity is quantified by reference to the rate of
reaction of the radical with styrene monomer. The original approach
for this scheme was based on copolymerizations of styrene with
acrylonitrile and with other acrylic monomers [149]. In developing a more general approach,
however, Jenkins concluded that while in principle, a general
procedure involves fewer assumptions, in practice much of the
utility is lost. He proposed that it is convenient in practice to
employ k is:
from styrene copolymerizations.
In writing the original scheme for styrene
copolymerization with acrylic monomers in benzene, Jenkins
introduced the term “S” to
denote styrene σ
p represents the Hammett sigma constant for a
substituent in the para
position. In his revised scheme, he still uses k is and writes it as
follows:
The term k 11 is substituted from
both sides of the equation to yield:
The above equation represents a postulate on the
same bases as does the Q
and e scheme. This scheme
contains an assumption that the intrinsic reactivity of a radical
is measured by the value of k 1s and its polarity
σ 1 (or
π 1). Thus,
It follows that r s2 = −ν 2 and σ 1 = π S = u S = ν S = 0. The final equation
becomes:
k
is represents the intrinsic reactivity of the styrene
polymer radical (assuming that it is monomer 1)
ν
2 represents the intrinsic reactivity of monomer 2
σ
1 represents the polarity of the polymer radical derived
from monomer 1, and
u
2 represent polarity of monomer 2
Whether this scheme is applicable to general
copolymerizations is not clear at this point.
3.6.3 Solvent Effect on Copolymerization
It was reported by Barb in 1953 that solvents can
affect the rates of copolymerization and the composition of the
copolymer in copolymerizations of styrene with maleic anhydride
[145]. Later, Klumperman also
observed similar solvent effects [145]. This was reviewed by Coote and coworkers
[145]. A number of complexation
models were proposed to describe copolymerizations of styrene and
maleic anhydride and styrene with acrylonitrile. There were
explanations offered for deviation from the terminal model that
assumes that radical reactivity only depends on the terminal unit
of the growing chain. Thus, Harwood proposed the “bootstrap model”
based upon the study of styrene copolymerized with MAA, acrylic
acid, and acrylamide [146]. It
was hypothesized that solvent does not modify the inherent
reactivity of the growing radical, but affects the monomer
partitioning such that the concentrations of the two monomers at
the reactive site (and thus their ratio) differ from that in
bulk.
Hutchinson and coworkers investigated
[147] effects of solvent on
free-radical copolymer composition and propagation kinetics in
copolymerizations of styrene with three methacrylates,
2-hydroxyethyl methacrylate, glycidyl methacrylate, and
n-butyl methacrylate Three
different solvents, n-butanol, toluene, and N,N-dimethylformamide were used. They
found that all three solvents effected the composition of
styrene-2-hydroxyethyl methacrylate copolymer. Liang and Hutchinson
[147] also observed variations in
monomer reactivity ratios with solvent polarity. Butanol was the
only one that affected styrene butyl methacrylate copolymer
composition. None of the solvents appeared to effect the
composition of the styrene–glycidyl methacrylate copolymer.
3.7 Terpolymerization
A quantitative treatment of terpolymerization,
where three different monomers are interpolymerized, becomes
complex. Nine growth reactions take place [155]:
Reaction
|
Rate
|
---|---|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
The rates of disappearance of the three monomers
are given by
By assuming steady state conditions for the three
radicals, M1•, M2•, and M3• it is
possible to write:
An equation for terpolymer composition was
developed from the rate expressions by expressing the steady state
with the relationships [94]:
It is claimed that this terpolymerization composition
equation is often in
good agreement with experimental results. Other, more complicated
equations also exist, but apparently they yield results that are
similar to those obtained from the above shown expression
[94, 149].
One example of other equations is an early
theoretical relationship for terpolymerization that was written by
Alfrey and Goldfinger [150].
where a, b, and c are the quantities of each monomer
found in the resultant terpolymer and A, B, and C are the quantities of the monomer in
the feed system. Needless to say, a copolymerization of four or
even more monomers becomes progressively more difficult to treat
rigorously.
3.8 Allylic Polymerization
Compounds possessing allylic structures
polymerize by free-radical mechanism only to low molecularweight
oligomers. In some cases the products consist mostly of dimers and
trimers. The DP for poly(allyl acetate), for instance, is only
about 14. This is due to the fact that allylic monomer radicals are
resonance-stabilized to such an extent that no extensive chain
propagations occur. Instead, there is a large amount of chain
transferring. Such chain transferring essentially terminates the
reactions [151]. The resonance
stabilization can be illustrated on an allyl alcohol radical:
The hydrogen transfer takes place from the
allylic hydrogen, as shown on allyl acetate:
Hydrogen transfer can also take place to the
acetate moiety:
The above described chain transferring is called
degradative chain
transferring. Other
monomers, such as methyl methacrylate and methacrylonitrile, also
contain allylic carbon–hydrogen bonds. They fail to undergo
extensive degradative chain transferring, however, and do form
high-molecular-weight polymers. This is believed to be due to lower
reactivity of the propagating radicals that form from these
monomers [5].
Yamasaki et al. [152], reported that they successfully performed
the radical polymerization of allylbiguanide hydrochloride in a
concentrated, acid solution using either hydrochloric acid or
phosphoric acid in the presence of a radical initiator at 50°C. The
polymer was precipitated from the reaction solution through the
addition of an excess amount of acetone. The molecular weight
average of the product was 10,340–113,200, with a low
polydispersity 1.04–1.68.
In spite of degradative chain transferring,
polyallyl compounds can be readily polymerized by a free-radical
mechanism into three-dimensional lattices. High DP is not necessary
to achieve growth in three dimensions. An example of such polyallyl
compounds is triallyl phosphate:
Many other polyallyl derivatives are offered
commercially for use in cross-linked films and are described in the
trade literature.
3.9 Inhibition and Retardation
Free-radical polymerizations are subject to
inhibition and retardation from side reactions with various
molecules [54]. Such
polymerization suppressors are classified according to the effect
that they exert upon the reaction. Inhibitors are compounds that react
very rapidly with every initiating free radical as it forms. This
prevents any polymerization reaction from taking place until the
inhibitor is completely consumed in the process. The reactions of
inhibitors with initiating radicals result in formations of new
free radicals. The newly formed free radicals, however, are too
stable to initiate chain growths. As a result, well-defined
induction periods exist. After the inhibitors are used up,
polymerizations proceed at normal rates.
Retarders are compounds that also
react with initiating radicals. They do not react, however, as
energetically as do the inhibitors, so some initiating radicals
escape and start chain growth. This affects the general rate of the
reaction and slows it down. There is no induction period and
retarders are active throughout the course of the
polymerization.
The efficiency of an inhibitor depends upon three
factors: (1) the chain transfer constant of an inhibitor with
respect to a particular monomer, (2) the reactivity of the
inhibitor radical that forms, (3) the reactivity of the particular
monomer.
Phenols and arylamines are the most common chain
transfer inhibitors. The reaction of phenols, though not fully
elucidated, is believed to be as follows [153]
Quinones are effective inhibitors for many
polymerization reactions. The reaction occurs either at an oxygen
or at a ring carbon [153–156]:
The reaction, however, is not always strict
inhibition. Thus, for instance, hydroquinone acts as an efficient
inhibitor for the methyl methacrylate radical but only as a
retarder for the styrene radical [155]. Hydroquinone is often employed as an
inhibitor; it requires, however, oxygen for activity
[156, 157]:
Oxygen, however, can also act as a comonomer in a
styrene polymerization:
It causes marked retardation, however, in the
polymerizations of methyl methacrylate [158]. The same is true of many other
free-radical polymerizations.
The ability of phenols to inhibit free-radical
polymerizations appears to increase with the number of hydroxyl
groups on the molecules [157].
The locations of these hydroxyl groups on the benzene rings in
relationships to each other is important. For instance, catechol is
a more efficient inhibitor than is resorcinol [158].
Aromatic nitro compounds can act as strong
retarders. Their effect is proportional to the quantity of the
nitro groups per molecule [160,
161].
Figure 3.1 illustrates the effect of inhibitors and
retarders on free-radical polymerization [162]. The equation that relates rate data to
inhibited polymerizations is
where Z is the inhibitor or the retarder in chain-growth
termination:
Fig.
3.1
Illustration of the effects of inhibitors
and retarders. (A) Normal polymerization rate, (B) effect of a
retarder, (C) effect of an ideal inhibitor, and (D) effect of a non
ideal inhibitor. The time between A and C is the induction period
caused by an ideal inhibitor
To simplify the kinetics it is assumed that Z•
and ~M nZ• do
not initiate new chain growth and do not regenerate Z upon
termination.
3.10 Thermal Polymerization
A few monomers, such as styrene and methyl
methacrylate, will, after careful purification and presumably free
from all impurities, polymerize at elevated temperatures. It is
supposed that some ring-substituted styrenes act similarly. The
rates of such thermal self-initiated polymerizations are slower
than those carried out with the aid of initiators. Styrene, for
instance, polymerizes only at a rate of 0.1% per hour at 60°C and
only 14% at 127°C. The rate of thermal polymerization of methyl
methacrylate is only about 1% of the rate for styrene
[163, 164]. Several mechanisms of initiation were
proposed earlier. The subject was reviewed critically
[165]. More recently, the
initiation mechanism for styrene polymerization has been shown by
ultraviolet spectroscopy to consist of an initial formation of a
Diels–Alder dimer. The dimer is believed to subsequently transfer a
hydrogen to a styrene molecule and as in doing so form a free
radical [166]:
Thermal polymerization of methyl methacrylate, on
the other hand, appears to proceed through an initial dimerization
into a diradical [167]. This is
followed by a hydrogen abstraction from any available source in the
reaction mixture.
3.11 Donor–Acceptor Complexes in Copolymerization
Polar interactions of electron donor monomers
with electron acceptor monomers lead to strong tendencies toward
formations of alternating copolymers. Also, some alternating
copolymerizations might even result from compounds that by
themselves are not capable of conventional polymerization. An
example is copolymerization of dioxene and maleic anhydride. Two
reaction mechanisms were proposed. One suggests that the
interactions of donor monomers with acceptor radicals or acceptor
monomers with donor radicals lead to decreased energies of
activation for cross-propagations [168]. The transition state is stabilized by a
partial electron transfer between the donor and acceptor species
[169]. The second mechanism
suggests that the interactions result in formations of
charge-transfer complexes [170].
An electron is completely transferred from the donor monomer to the
acceptor monomer. After the transfer, the complex converts to a
diradical that subsequently polymerizes by intermolecular coupling.
For instance, while many believe that the Diels–Alder reaction
takes place by a concerted mechanism, the intermediate was
postulated by some to be a charge-transfer complex. An electron is
transferred from the donor to the acceptor and a charge-transfer
complex forms [171]. This can be
illustrated on a Diels–Alder reaction between butadiene and maleic
anhydride:
If the reaction mixture is irradiated with high
energy radiation, such as gamma rays, instead of being heated, an
alternating copolymer forms. The complex converts to a diradical
[1, 171] that homopolymerizes:
Alternating copolymerization of styrene with
maleic anhydride is also explained by donor acceptor interactions
[171]. A charge-transfer complex
is seen as the new monomer, a diradical, which polymerizes through
coupling [171–174].
Charge-transfer complexes are also claimed to be
the intermediates in free-radical alternating copolymerization of
dioxene or vinyl ethers with maleic anhydride [176–179]:
where, R• is a polymerization-initiating radical. Here, a third
monomer can be included to interpolymerize with the complex that
acts as a unit. The product is a terpolymer [176, 177]. A
diradical intermediate was also postulated in sulfur dioxide
copolymerizations and terpolymerizations with bicycloheptene and
other third monomers [173]. These
third monomers enter the copolymer chain as block segments, while
the donor–acceptor pairs enter the chains in a one-to-one molar
ratio. This one-to-one molar ratio of the pairs is maintained,
regardless of the overall nature of the monomer mixtures.
The propagation and termination steps in the
above reactions are claimed [175,
181–183] to be related. As stated, an interaction
and coupling between two diradicals is a propagation step. When
such interactions result in disproportionations, however, they are
termination steps. This means the charge-transfer mechanisms are
different from conventional free-radical polymerizations. They
involve interactions not only between growing polymer-radicals and
monomers but also between polymer-radicals and complexes. In
addition, the polymer radicals react with each other
[175, 181–184].
Li and Turner [185], reported copolymerization of maleic
anhydride with trans-stilbene:
where, R1, R2, and R3 are either
methyl groups or hydrogens
The stability of charge-transfer complexes
depends upon internal resonance stabilization. This degree of
stabilization determines how easily the diradicals open up
[183]. Consequently, the
stability also determines how the copolymerization occurs. It can
occur spontaneously, or under the influence of light or heat, or
because of an attack by an initiating free radical.
There are many examples of spontaneous reactions.
When, for instance, isobutylene is added to methyl α-cyanoacrylate,
a spontaneous copolymerization in a 1:1 ratio takes place at room
temperature. This was explained by the following scheme
[175, 181–183]:
The same happens when sulfur dioxide is added to
bicycloheptene at −40°C [184].
Another example is a room temperature 1:1
copolymerization of vinylidene cyanide with styrene [185]:
Still another example is a reaction of 1,3
dioxalene with maleic anhydride [183].
Examples of stable complexes are reactions of
sulfur dioxide with styrene [175], or vinyl ethers with maleic anhydride
[184], also α-olefins with maleic
anhydride [179–181]. Also, a reaction of trans-stilbene with maleic anhydride
[182], In these reactions
charge-transfer complexes form. They are stable and their existence
can be detected by spectroscopic means. Additional energy, such as
heat or a free-radical attack, converts them to diradicals and
polymerizes them into alternating copolymers [175, 185–191].
Examples of intermediates between the two
extremes in stability are reaction products of sulfur dioxide with
conjugated dienes [192]. In this
case, the reaction results in formation of mixtures of alternating
copolymers and cyclic adducts:
The yield of the polymer increases at the expense
of the cyclic structure when heat or radiation is applied. A
free-radical attack has the same effect [192].
If donor–acceptor interactions and subsequent
polymerizations occur upon irradiation with ultraviolet light, the
reactions can be very selective. An example is a triphenylphosphine
interaction with acrylic monomers [193]:
This reaction does not occur, however, between
triphenylphosphine and styrene or vinyl acetate [193].
The nature and the amount of solvent can
influence the yield and the composition of the copolymers in these
copolymerizations. Thus, copolymerization of phenanthrene with
maleic anhydride in benzene yields a 1:2 adduct. In dioxane,
however, a 1:1 adduct is obtained. In dimethyl formamide, no
copolymer forms at all [193].
Another example is a terpolymerization of acrylonitrile with
2-chloroethyl vinyl ether and maleic anhydride or with p-dioxene-maleic anhydride. The amount
of acrylonitrile in the terpolymer increases with an increase in
the π-electron density of the solvent in the following order
[194]:
xylene ≫ toluene ≫ benzene > chlorobenzene ≫
chloroform.
The ratio of maleic anhydride to the vinyl ether
in the product remains, however, equimolar.
Whether the concept of charge-transfer complexes
in copolymerizations is fully accepted is not certain. Much of the
accumulated evidence, to date, such as UV and NMR spectroscopy,
does support it in many systems [195]. Further support comes from the strong
tendencies to form alternating copolymers over a wide range of feed
compositions, and also from high reaction rates at equimolar feed
compositions [171]. On the other
hand, as shown above, it was claimed in the past that
copolymerization of styrene with maleic anhydride involves
charge-transfer complexes [171,
181–183]. This, however, is now contradicted in a
publication of a study of radical copolymerization of maleic acid
with styrene. The reaction was carried out in a dioxane solution at
70°C. The authors reported that UV spectroscopy fails to show
presence of a charge transfer and formation of a complex between
the two monomers in the copolymerization system [196].
3.12 Polymerization of Complexes with Lewis Acids
Some polar vinyl monomers such as methyl
methacrylate or acrylonitrile interact and complex with Lewis
acids. They subsequently polymerize at a faster rate and to a
higher molecular weight than can be expected otherwise. The
effective Lewis acids are ZnCl2, AlCl3,
Al(C2H5)2Cl,
AlCl2(C2H5), SnCl4, and
some others [197–199]. Complexes can form on an equimolar basis
and undergo homopolymerizations after application of heat. A
free-radical attack or irradiation with ultraviolet light or with
gamma rays is also effective [198].
It was proposed [198–201] that
the greater reactivity is due to delocalization of the electrons of
the complexed molecules:
Molecular orbital calculations support this
[200, 201]. Polymerizations can also be carried out
with less than stoichiometric ratios of the Lewis acids to the
monomers. As an illustration, coupling of an acrylic or a
methacrylic monomer with a Lewis acid can be shown as follows
[203, 204]
where, MeXn represents a Lewis acid.
Vinyl pyridine and vinyl imidazole also form
complexes with zinc chloride. Here, conjugation with the metal salt
results in a super delocalizability and leads to spontaneous
thermal polymerization [203,
204].
The increased reactivity toward polymerization by
some monomers, even in the presence of less than equimolar amounts
of Lewis acids, was suggested to be due to one of two
possibilities. It may be due to greater reactivity of the portion
of the monomer that complexes toward the growing chain end. It may
also be due to formation of new complexes between uncomplexed
monomer and complexed ones [171]:
The picture is more complicated when the Lewis
acids are used in combinations with donor–acceptor monomers. The
donor–acceptor complexes are believed to form first and then react
with Lewis acids according to the following scheme [203, 204]:
A structure of methyl methacrylate and styrene
complexed with SnCl4 was shown to be as follows
[204]:
In formations of ternary complexes, the acceptor
vinyl compound must have a double bond conjugated to a cyano or to
a carbonyl group. Such acceptors are acrylonitrile,
methacrylonitrile, acrylic and methacrylic esters and acids, methyl
vinyl ketone, acrylamide, etc. Donor monomers are styrene, α-methyl
styrene, butadiene, 2–3-dimethyl butadiene, isoprene, chloroprene,
etc.
One proposed mechanism [171, 204] is
that such charge-transfer polymerizations are in effect
homopolymerizations of the charge-transfer complexes [DA_…MeXn]. In other words, the
metal halide is complexed with the electron acceptor monomer and
acts as an acceptor component.
The above opinion, however, is not universal.
Others hold that the increased susceptibility to ultraviolet
radiation or to initiating radicals [205] is due to increased reactivity of the
propagating radicals of complexed monomers toward incoming
uncomplexed ones.
Arguments against the ternary complex mechanism
are as follows: (1) the physical evidence that proves the existence
of the ternary molecular complexes is weak; (2) the ternary
molecular complexes can have no bearing on the copolymerizations
because the equilibrium concentration of the complexed monomers is
low, compared to the uncomplexed ones [206].
A third opinion is that a complex of an acceptor
monomer with a Lewis acid copolymerizes alternately with the donor
monomer and with an uncomplexed acceptor monomer [207, 208].
This presumably takes place according to the conventional
chain-growth polymerization scheme of radical copolymerization. The
alternate placement of monomers is due to highly enhanced values of
cross-propagation constants. It results from complexing acceptor
monomers with Lewis acids. Such a mechanism fails to explain
satisfactorily the completely alternating incorporation of monomers
and the inefficiency of chain-transfer reagents. It also fails to
explain the spontaneous initiation of alternating
copolymerization.
Kabanov suggested [209] that during the primary free-radical
formation of the Lewis acid–monomer complex, both the uncomplexed
and the complexed monomers may participate in chain propagation.
This would result in appearance of complexed propagating radicals
besides the usual ones. In the complexed ones, the last unit
carrying the valence is a ligand of coordination complex:
It excludes, however, all electron transfer
reactions that may take place due to ultraviolet light
irradiation.
3.13 Steric Control in Free-Radical Polymerization
In free-radical polymerization reactions, the
propagating radical chain has a great amount of freedom. Atactic
polymers, therefore, are usually formed. Some control that the
reaction conditions exercise over the propagating species increases
at lower temperatures due to lower mobilities. This leads to
increased syndiotactic placement, as was discussed in the section
on propagation. Special techniques, however, such as the use of
canal complexes can be employed to form stereoregular polymers by
free-radical mechanism. Urea and thiourea were used originally for
such purposes [210,
211]. Monomers such as butadiene
or others form complexes within the voids, or canals of the crystal
lattices of these compounds. Brief exposure to high energy
radiation initiates chain growth. In the canals, the monomer
molecules are held in fixed positions, so chain growth is
restricted in one direction only. Steric control is exercised
because in these fixed positions the monomer molecules tend to
align uniformly. It was suggested that in the canal complexes the
monomers are not just lined up end to end, but packed in an
overlapping arrangement. For molecules such as isobutylene or
vinylidene chloride, it may be possible for the monomers to lie
directly on top of each other, resembling a stack of coins. Such
stacking greatly facilitates reactions between guest molecules
[210, 211].
Polymerizations in thiourea canal complexes
yields high melting crystalline trans-1,4 polybutadiene,
2,3-dimethylbutadiene, 2,3-dichlorobutadiene, and
1,3-cyclohexadiene. Cyclohexadiene monoxide, vinyl chloride, and
acrylonitrile also form stereoregular polymers. On the other hand,
polymerizations of isobutylene and of vinylidene chloride fail to
yield stereospecific polymers.
Sodium montmorillonite can also be used to
polymerize polar monomers between the lamellae. Here too, the
organization of monomer molecules within the monolayers influences
the structure of the resultant polymers [212, 213].
Poly(methyl methacrylate) formed in sodium montmorillonite is
composed of short, predominantly isotactic stereosequences
[211]. The percentage of
isotactic component increases with an increase in the ion
exchanging population on the surface of the mineral and is
independent of the temperature between 20 and 160°C. In this way,
it is possible to vary the population of isotactic triads at will
up to 50% composition [205].
Perhydrotriphenylene also forms channel-like
inclusions with conjugated dienes. Polymerization of these dienes
yields some steric control [216,
218].
Uemura and coworkers [217] carried out radical polymerizations of
vinyl monomers (styrene, methyl methacrylate, and vinyl acetate)
within various nanochannels of porous coordination polymers. They
studied the relationships between the channel size and
polymerization behaviors, such as monomer reactivity, molecular
weight, and stereostructures. They reported that in these
polymerization systems, the polymer-growing radicals were
remarkably stabilized by efficient suppression of the termination
reactions within the channels, resulting in relatively narrow
molecular weight distributions. A significant nanochannel effect on
the polymer stereoregularity was also seen, leading to a clear
increase of isotactic placement in the resulting polymers.
There were attempts at controlling steric
placement by a technique called template polymerization. An example
is methyl methacrylate polymerization in the presence of isotactic
poly(methyl methacrylate) [208,
209]. Thus template
polymerization is a process of polymerizing a monomer in the
presence of a polymer, usually from a different monomer. The
presence of template polymers, however, only results in
accelerating the rates of polymerizations [219].
3.14 Controlled/“Living” Free-Radical Polymerization
Living
polymerizations are chain-growth reactions where the
propagating centers on the growing chains do not terminate and do
not undergo chain transfer. Such polymerizations are noted for
preparations of polymers with controlled molecular weights, desired
end groups and low polydispersities. In addition, the preparations
of polymers with predetermined molecular weights and narrow
molecular weight distributions require fast initiations and fast
exchanges between sites of variable activities and variable
lifetimes. Such chain-growth reactions, ionic in nature, are
discussed in Chap.
4. In typical homogeneous free radical
polymerizations, however, bimolecular terminations between two
growing radicals cannot be avoided and, therefore, typical living
free radical polymerization cannot be fully realized. Also, in
conventional free radical polymerizations, the initiations are
slow, while high-molecular-weight polymers form shortly after the
start of the reactions. As the reactions progress, polydispersities
increase, while the molecular weights actually decrease. It is
possible, however, to adjust conditions of some radical
polymerizations in such a way that polymers with controlled
molecular weights and relatively low polydispersities form
[220, 221]. These are not true living polymerization
as such because termination reactions do occur. They possess,
however, some characteristics that are similar to living
polymerizations and are referred to by many as controlled /“ living ” polymerizations. Such reactions yield
polymers with controlled molecular weights, exhibit increase in
molecular weight with conversion, yield narrow molecular weight
distributions, and can be used to form copolymers. Some examples of
such polymerizations include nitroxyl radical-mediated
polymerizations of styrene [222–225],
atom transfer polymerizations controlled by ruthenium-(II)/aluminum
[226, 227] or by copper/bipyridine complexes
[228], Co(II)-mediated
polymerizations of methacrylates and acrylates [229], and polymerization of styrene using a
degenerative transfer method [230], as well as others. Some features are
unusual for radical processes and the radical nature of some of
these reactions might be questioned, as for instance,
polymerizations catalyzed by transition metals. Evidence has been
presented, however, that strongly indicates radical nature in at
least in atom transfer polymerizations [231]. The evidence, however, is not
unambiguous.
Some initial attempts at producing “living”
polymerizations made use of iniferters . This term appears to come from the word
inifer, a bifunctional
compound that brings about both initiation and chain transfer.
“Living” cationic polymerizations make use of inifers to form block
copolymers. This is discussed in Chap.
9. The term iniferter was proposed by Otsu and
Yoshida in 1982 [232]. Iniferters
used in controlled/living free-radical polymerizations are
sulfur-centered free radicals that can be generated from
sulfur-containing molecules such as dithiocarbamates. The radicals
react reversibly with growing polymeric chain ends, thereby
controlling the concentration of the radical species. Many of these
sulfur centered radicals, however, can also initiate new polymer
chains. This can lead to uncontrolled growth. To overcome these
difficulties, other approaches were also developed [233].
Deactivation of growing radicals with stable
radicals can be carried out with the aid of various nitroxyl
radicals, protected phenoxy radicals, dithiocarbamate, trityl, and
benzhydryl derivatives. Growing radicals can also be deactivated
with nonradicals in the presence of organometallic compounds that
form stabilized hyper-coordinated radicals. The polymerizations
with the aid of reversible degradative chain transferring are
unique in that they requires very rapid and “clean” chain transfers
without side reactions. The enhanced control of polymerization
process relies on reduction in the ratio of the rate of termination
to that of propagation, due to low instantaneous concentration of
growing radicals. This means that initiation and propagation
reactions must proceed at similar rates due to application of the
initiators resembling polymer end groups in their dormant state.
Also, in these polymerization reactions, there must be a low
proportion of chains marked by uncontrolled termination and/or
transfer due to relatively low molecular weights.
Homogenous controlled/“living” free radical
polymerizations are based, therefore, on the reversible
deactivations of growing radicals. Early, Matyjaszewski divided
such polymerizations into three classes [240, 241].
These were:
1.
Deactivations of growing radicals with stable
radicals by reversible formations of dormant covalent species,
followed by homolytic cleavages:
2.
Reversible deactivations of growing radicals with
“nonradical” species by formation of dormant persistent radicals:
3.
Reversible degenerative transfers based on
thermodynamically neutral exchange reactions between growing
radicals and transfer agents:
3.14.1 Cobalt Mediated Polymerizations
Catalytic chain transfer polymerizations can
utilize metals such as low spin cobalt(II) compounds as chain
transferring agents. The mechanism is believed to involve repeated
disturbing of each propagating step by abstraction of hydrogen
atoms from the propagating polymers. This yields chains with
unsaturated terminal units and hydrogen transfer agent adducts,
Co(III)–H. Subsequent transfers of hydrogens to the growing chains
result in reinitiating the processes [241]. The β-hydrogen abstractions from the
growing radicals and the formations of metal hydrides
[229] can be illustrated as
follows:
where Mt = Co
Fig.
3.2
Cobalt mediated controlled/“living”
polymerization
This catalytic cycle for a cobalt mediated
polymerization, using bis-dimethyl-glyoximate cobalt boron fluoride
catalyst was illustrated Haddleton et al. [233, 234]. A
similar illustration of the process is shown in
Fig. 3.2.
The affinities of metals for hydrogen abstractions, and/or their
abilities to lose electrons depend on their oxidation states and
the nature of their ligands. Such reactions can, therefore, be
suppressed. This can be done by choosing high oxidation state
metals, ligands that protect the metal from the abstraction of the
P–H atoms, and by controlling the position of the
oxidation–reduction equilibrium.
The metal cobalt(II) is usually chelated. This
can be cobalt porphyrin, cobalt phthalocyanin, or cobalt oxime, as
well as others. The polymer molecules that form have, as stated
earlier, terminal double bonds and can be illustrated as follows,
The cobalt hydride in turn reacts with a new
monomer molecules to regenerate the Co(II) [235].
One publication describes a chain transferring
agents that can be used in controlled polymerization of
methacrylate monomers where reductive elimination of cobalt hydride
from the neighboring methyl group deflects further chain growth
[236]. The agent was illustrated
as follows:
A described example is a reaction conducted at
60°C in deoxygenated benzene, using neopentylcobalt with
tetramesityl-porphyrin ligand and methyl acrylate monomer
[236]. A slow polymerization
yields 66% conversion in 38 h. The product is a narrow
molecular weight distribution polymer of M n = 144,000. The
polymerization is even slower with less hindered phenyl
substituents on the porphyrin ligand. Both homopolymers and block
copolymers can be formed.
Catalytic chain transfer by a cobalt(ll)
porphyrin in radical polymerization of MAA in water was studied by
Wayland and coworkers [237].
Cobalt tetrasulfonatophenylporphyrin was found by them to be
exceptionally effective in the catalytic chain transfer for the
radical polymerization of MAA in water. A remarkable feature of
this process is that the increase in the degree of polymerization,
with conversion requires that more monomer be consumed in chain
growth of the existing macromonomers than in initiation and
propagation of new chains through chain transfer to monomer.
Reinitiation of oligomer olefins and chain growth are significant
inherent reactivity features of the cobalt++-porphyrin
catalyzed chain transfer process.
3.14.2 Atom Transfer Radical Polymerizations
Atom transfer radical polymerizations (ATRP) were
reported simultaneously by two groups: (1) Matyjaszewski et al.
[218] and (2) Sawamoto and
coworkers [226]. Matyjaszewski et
al. utilized a Cu/bipyridine complex as a halogen transfer agent that functions between
dormant and active polymer chains. Formation of polymers with
predetermined molecular weight of up to M n ≈ 10 [5] and polydispersity as narrow as 1.05 was
reported [238, 239]. This type of polymerization appears to
offer the possibility of preparing a broad range of polymeric
materials [240–242]. The reactions proceed under conditions
that could make the process commercially attractive. Thus, for
instance, by using nonionic surfactants, such as poly(oxyethylene
oleyl ethers) it is possible to prepare polymers from butyl
methacrylate, methyl methacrylate, styrene, and butyl acrylate in
aqueous emulsions. In addition, by using multidentate ligand such
as tris[(2-dimethyl-amino)ethyl]amine the atom transfer
polymerizations can be made to proceed rapidly at room temperature
[242, 243]. The atom transfer polymerization reaction
can be illustrated as follows:
Polymerizations of styrene using 2,2′-dipyridyl
as the ligand indicated that they proceed first order with respect
to the concentration of initiator, and 0.4 and 0.6 orders with
respect to the concentration of Cu(l) halide and ligand
[218, 229]. The copper bipyridyl complexes mentioned
above were pictured by Haddleton et al. [233, 234] as
follows:
Recently, Matyjaszewski has summarized the
mechanism of these polymerizations [245].
Matyjaszewski and coworkers [246–250]
reported that small amounts of air present in the reaction mixture
can be consumed by addition of sufficient amounts of an appropriate
reducing agent, such as tin(II)2-ethylhexanoate or ascorbic acid.
In this process, the cuprous ions are initially oxidized by oxygen
to the cupric ions, but then in turn reduced by the reducing agent.
The cuprous ions activate the reaction. There is an induction
period until all the oxygen is consumed. This is referred to as
(ARGENT) ATRP. Also, they have subsequently
reported that polymerizations of 2-(dimethylamino)ethyl
methacrylate does not require any addition of a reducing agent as
the tertiary amine group presumably serves as an internal one
[251].
In addition, Percec and coworkers [253] reported that polymerizations in polar
solvents in conjunction with copper and appropriate ligands allow
ultrafast syntheses of high-molecular-weight polymers at ambient
temperature. The process is referred to as Single Electron Transfer - Living Radical Polymerization (SET - LRP ). The mechanism proposed is based on
disproportionation of cuprous ions to cupric ions and metallic
copper. This is catalyzed by the polar solvents and the appropriate
ligands. The proposed mechanism can be illustrated as follows:
The work by Percec and coworkers included an
investigation of various solvents and ligands for the catalyst
activity and their ability to disproportionate the cuprous ion.
They demonstrated that addition of 10 mol% of phenol as ligand
leads to spontaneous disproportionation to metallic copper and
cupric ions [253]. An alternative
to the proposed Percec’s mechanism was proposed by Matyjaszewski
[254]. According to this
mechanism, metallic copper acts as a reducing agent for the cupric
ions and yields active cuprous ions that catalyze the
polymerization. This mechanism is similar to one proposed for the
reactions that utilize ascorbic acid or tin based reagents to
reduce cupric ions to cuprous ones [255, 256].
The mechanism can be illustrated as follows:
Haddleton and coworkers [256] investigated use of toluene as a solvent
with phenol as an additive for use in living/controlled
polymerizations. They demonstrated a direct relationship between
the reaction time and the amount of phenol added. The optimum
amount found by them is 20 equivalents of phenol with respect to
the initiator. Their products were narrow molecular weight polymers
with MWD ranging between 1.05 and 1.25.
Removal of copper from ATP products can sometimes
be a problem [257]. Honigfort and
coworkers reported that they found that when the ligands were
supported on Janda Jel (see Chap.
10) resins, easy removal of the catalyst complex
was possible from the reaction mixture. The Janda Jel ligands were
used in ATRP of methyl methacrylate, styrene, and
2-(dimethylamino)ethyl methacrylate. The methyl methacrylate and
2-(dimethylamino)ethyl methacrylate polymerizations proceeded
quickly to high conversion (>90%) and were well controlled. The
styrene polymerization, however, was found by them to be sluggish
and proceeded only to 63% conversion. After polymerizations were
complete, the catalyst ligand complex was easily removed by
filtration. Zhu and coworkers [257] claim to have a simple and effective
method for purification of an ATP product using catalyst
precipitation and microfiltration. The method relies on the
precipitation of the Cu + Br ligand catalyst complex by the
additions of Cu++Br2. The precipitate thus formed is
effectively retained by a 0.14-μm PTFE filter, resulting in up to
99.9% of the catalyst being removed from the polymer. The resulting
clear polymer filtrate contains little residual copper, down to
10 ppm.
Matyjaszewski and coworkers developed a process
[255] for an electrochemically
mediated ATRP. They use applied voltage to drive the production of
Cu + ions that catalyze the polymer formation. Because the rate of
the reaction is controlled by a redox equilibrium between cuprous
and cupric ions, electrochemistry permits the regulation of the
concentration of each species.
A similar ATP process is one that uses iron(II)
bis(triphenylphosphine)-dichloride[FeCl2(PPh3)2].
It induces “living” polymerization of monomers such as methyl
methacrylate in conjunction with organic halides as initiators in
the presence and in the absence of Al(OiPr)3
in toluene at 80°C. The molecular weight distributions of the
products are 1.1–1.3 [269]. The
following mechanism is visualized [269]:
The ATP process developed by Sawamoto and
coworkers [226], uses an
initiating system consisting of carbon tetrachloride,
dichlorotri(triphenyl-phosphine)-ruthenium (II) and methylaluminum
bis(2,6-di-tert-butylphenoxide) to polymerize
methyl methacrylate [226]. The
polymerization involves reversible and homolytic cleavages of
carbon-halogen terminal groups assisted by transition metal
complexes [226].
The ruthenium(II) complexes interact with
CCl4 and are oxidized in the process to become Ru(III)
and radicals CCl3• that add to molecules of methyl
methacrylate. The polymerization proceeds via repetitive additions
of methyl methacrylate molecules to the radical species that are
repeatedly generated from the covalent species with carbon-halogen
terminal groups [226]. Suwamoto
also reported [226] that addition
of a halogen donor, Ph3C–Cl aids the shift of the
equilibrium balance to dormant species. The reaction of
polymerization can be illustrated as follows:
Klumperman and coworkers [259] observed that while it is lately quite
common to treat living radical copolymerization as being completely
analogous to its radical counterpart, small deviations in the
copolymerization behavior do occur. They interpret the deviations
on the basis of the reactions being specific to controlled/living
radical polymerization, such as activation—deactivation equilibrium
in ATRP. They observed that reactivity ratios obtained from atom
transfer radical copolymerization data, interpreted according to
the conventional terminal model deviate from the true reactivity
ratios of the propagating radicals.
Velazquez and coworkers [260], developed a kinetic model incorporating
effects of diffusion-controlled reactions on atom-transfer radical
polymerization. The reactions considered to be diffusion-controlled
are monomer propagation, bimolecular radical termination, chain
transfer between propagating radicals and catalyst, and transfer to
small molecules. Model predictions indicate that a
diffusion-controlled propagation reduces the “living” behavior of
the system, but a diffusion-controlled termination enhances its
livingness. Also, diffusion-controlled transfer between chains and
catalyst is the same in the forward and in the reverse directions.
The “livingness” of the system is enhanced, but if one of them is
kept unchanged the other is increased, and the “livingness” of the
system is reduced. When diffusion-controlled termination is
important, their simulations show that the overall effect of
diffusion-controlled phenomena in ATRP is to enhance the livingness
of the system.
Preparation of gradient copolymer of styrene and
n-butyl acrylate was
reported by the use of ATRP [261]. Gradient copolymers are copolymers
with sequence distributions varying in a well-defined order as
functions of chain lengths. It is suggested that gradient
copolymers have the potential of outperforming block and
alternating copolymers in some instances [261].
3.14.3 Nitroxide-Mediated Radical Polymerizations
A nitroxide mediated polymerization of styrene
was first reported in 1985 [262].
This reaction, however, was studied extensively only since 1993.
The monomer conversion rates vs. temperatures are much slower than
they are in conventional styrene polymerization. Also, the
polydispersities of the products are not as narrow as obtained in
anionic polymerization but, generally, the polydispersities
produced by this process are proportional to the molecular weights
of the polymers produced. In fact, a linear relationship between
polydispersity and the molecular weight of the polystyrene product
was demonstrated [263].
After the initial nonstationary period, typical
alkene polymerizations in the presence of alkoxyamines proceed
according to the first order kinetics with the molecular weights
increasing with conversion. The dispersity of the products and the
contribution of the nonstationary periods depend upon the
temperature, the particular initiating system and on the nature of
the monomers. Styrene polymerizations can be carried out in the
presence of stable nitroxyl radicals, such as the
2,2,6,6-tetramethylpiperydinyl-1-oxy radical, commonly referred to
as TEMPO
[264] or ditertiary butyl
nitroxide, referred to as DTB
N .
Such radicals are incapable of initiating
polymerizations by additions to the double bonds, but react
selectively with growing radicals to form reversibly covalent
species [265]. In addition, the
reactions of growing radicals with dormant species occur via
degenerative transfer:
The position of the equilibrium constant in
reactions with TEMPO depends on the nature of the radical, the
solvent and the temperature. These polymerizations can be initiated
by either bimolecular initiators or by unimolecular ones. The
bimolecular initiators utilize common free radical sources such as
benzoyl peroxide or azobisisobutyronitrile to start the reaction.
The carbon-centered initiating radicals that form in turn react
with TEMPO. This can be illustrated as follows:
Various descriptions of different unimolecular
initiators can be found in the literature. A presence of α-methyl
groups on the alkoxyamines appears to be essential [266]. These compounds yield, upon dissociation,
both stable radicals and initiating ones and can be shown as
follows [267].
The optimal amount of the radical initiator
depends on the efficiency of the initiation. Ideally the
concentration of the radicals generated from the initiator should
be slightly higher than the concentration of the scavenger.
At higher temperatures, such as 120°C, the
polymerizations of styrene tend to exhibit ideal behavior. Also, at
higher temperatures narrower molecular weight distributions are
obtained, indicating sufficiently high exchange rates.
A low-temperature method for the preparation of
unimolecular initiators was reported [268]. In this method, oxidation is used to
generate carbon radicals in the presence of nitroxide traps such as
TEMPO.
A variation in controlled/“living” polymerization
of vinyl acetate by the use of a bidentate ligand, 2,2′-bipyridyl
and TEMPO composition in 2:1.2 ratio that was reported by Mardare
and Matyjaszewski [267]. The
following mechanism was proposed.
(1) Pentacoordinated complexes (I) are formed at
a molar ratio of 1.1
(2) The irreversible attacks by TEMPO on the
pentacoordinated complexes, (Al(iBu)3):BPy (I), lead to
relatively stable and delocalized radicals(II). TEMPO also reacts
with some short-lived radicals present at stage (1) to form
alkoxyamines and pentacoordinated complexes of type IV. The
radicals II could be in equilibrium with tiny amounts of very
reactive radicals R• capable of initiations and subsequent
propagations.
Aldabbagh and coworkers [269] reported that carrying out the
nitroxide-mediated polymerization in supercritical carbon dioxide
allows improved control of the reaction.
Nesvadba and coworkers [270] used nitrones in controlled radical
polymerization of vinyl monomers. This was the beginning of the in
situ NMP concept. The
alkoxyamines were prepared by reaction of free radicals obtained
from decomposition of azo-initiators, such as
azobisisobutyronitrile or l,1′-azobis(cyclohexanecarbonitrile) with
selected nitrones:
The alkoxyamines were utilized in radical
polymerization of acrylates and styrene in bulk or in solution
between 100 and 145°C. Low molecular weight polymers,
3,000–14,000 g/mol formed rapidly with polydispersity,
M w/M n between 1.2 and 3.4.
High styrene conversion was observed together with a low
polydispersity.
Subsequently, nitroxides and parent alkoxyamines
were formed directly in the polymerization medium (in situ
NMP) by reaction of the
nitrone with the free radical initiators [270]. Two types of reactions were carried out.
One was a reaction before monomer addition and the other one after
the addition. In either case, a prereaction was systematically
carried out at temperatures ranging from 60 to 80°C. This was
followed by polymerizations at 130°C. The in situ-formed nitroxides
and alkoxyamines controlled the radical polymerizations of
n-butyl acrylate yielding,
however, low molecular weight polymers, of M n < 10,000 and
M w/M n equal to 1.65–2.0.
A patent was issued to Wertmer and coworkers
[271] for controlled radical
(co)polymerization of vinyl monomers mediated by nitrones
substituted by longer alkyl groups that contained as much as 18
carbon atoms. The nitrone was simply heated in the presence of
peroxide and a monomer, such as styrene at 130°C for 24 h.
High-molecular-weight polystyrene, M n = 98,000–146,000 was
formed. The ratio of M
w/M
n, however, was not disclosed
Recently, Grubbs and coworkers [272] have synthesized an active alkoxyamine by
reaction of 2-methyl-2-nitrosopropane with 1-bromoethylbenzene,
catalyzed by ligated CuBr in the presence of metallic copper. A
purified alkoxyamine was used to initiate the radical
polymerization of styrene and isoprene. Well-defined low
polydispersity polymers formed with M w/M n = 1.14 for polystyrene
and 1.28 for polyisoprene. Subsequently, Grubbs and coworkers
[273] used this alkoxyamine and
successfully controlled the radical polymerization of n-butyl acylate at 125°C. Lower ratio
of M
w/M n
was observed when the alkoxyamine was preheated at temperatures up
to 125 for 30 mm prior to adding the monomer. This prereaction
was needed for an excess of free nitroxide to be formed in situ and
for polymerization to be controlled.
3.14.4 Reversible Addition-Fragmentation Chain Transfer Polymerization
Another type of “living”/controlled radical
polymerization involves reversible addition - fragmentation chain transfer. It was named, therefore,
RAFT polymerization.
Great versatility and effectiveness was shown for the process
[274]. The process is said to be
compatible with a very wide range of monomers including functional
monomers containing such functional groups as acids, acid salts,
and hydroxyl or tertiary amine groups. The conditions of
polymerization are those used in conventional free-radical
polymerizations. They can be carried out in bulk, solution,
emulsion and suspension (see Sect. 3.16). The usual azo or
peroxide initiators are employed [274]. The reaction was originally illustrated
as follows [274]:
The RAFT process depends upon rapid
addition—fragmentation equilibrium reaction between propagating
(Pn•) as well as intermediate radicals, and chain
activity and dormancy, as shown below in the reaction scheme. The
concentrations of each of the species within the equilibrium is
dependent on the relative rate coefficients for addition of a
propagating radical to the RAFT agent (K add) and fragmentation of
the formed intermediate radical (K frag). This equilibrium
applies correctly only for polymeric chains that are present in
significant concentrations after an initialization period. During
the initiation period there are mainly shorter chains present. The
important part of this equilibrium is the relatively stable radical
intermediates.
It was reported that RAFT-mediated polymerization
reactions typically contain anomalies, such as an “inhibition”
period and rate retardation. The rate retardations or reductions in
the polymerization rates apparently occur in the presence of RAFT
agents, and are not observed when RAFT agents are absent. Examples
are dithiobenzoate-mediated polymerization reactions
[275].
Tonge and coworkers [276] investigated the reactions of short-chain
species during the initial period of cumyl dithiobenzoate mediated
polymerization of styrene at 84°C. Using electron spin resonance
and hydrogen and carbon NMR spectroscopies they were able to
demonstrate that the reactions are very specific during the initial
stages. There is a strong preference to add single monomer species.
This is followed by fragmentation and release of shorter radicals
prior to formation of longer chains.
The effectiveness RAFT agents were investigated
by Moad and coworkers [277].
These RAFT agents, such as thiocarbonylthio compounds, depend in
effectiveness on the nature of the group, Z and R (shown below)
that modify the reactivity of the thiocarbonyl group toward free
radical addition. R is the free radical leaving group
[277]:
These RAFT agents are based on the structure
[S=C(Ph)S–R] [277] They found
that the effectiveness of these agents also depends upon the nature
of the monomer and on the polymerization conditions. For the
polymerization of styrene, methyl methacrylate, butyl acrylate, and
methyl acrylate at 60°C, the effectiveness of R decreases in the
following order [277]:
In addition, among the above compounds, only when
R=C(CH3)2CN or
C(CH3)2Ph did these thiocarbonylthio
compounds yield polymers with narrower polydispersities in batch
polymerizations. Also, only these compounds allowed molecular
weight control that may be expected from a living polymerization.
The reaction mechanism was proposed by Moad and coworkers as
follows [277]:
Moad and coworkers concluded [277] that a major factor that determines the
transfer coefficient of dithiobenzoate derivatives is the way the
intermediate 3 (see the above in equation) partitions between
starting materials and products. This in turn is determined by the
relative ability of the leaving group R• and by the propagating
radical. Steric factors and radical stability of R• are also
important. They conclude that more stable, more electrophilic, and
bulkier radicals are better leaving groups. The partitioning of R•
between the monomers (to reinitiate) or by adding to polymeric RAFT
can also have a significant effect on the rate of consumption of
RAFT agent [277].
Vana and coworkers [278] studied conversion vs. time and molecular
weight distributions in conjunction with the kinetic scheme for the
RAPT process. In particular, conditions leading to inhibition and
rate retardation were examined to act as a guide to optimizing the
reaction. They demonstrated that there is an inhibition period of
considerable length. It is induced by either slow fragmentation of
the intermediate RAFT radicals appearing in the preequilibrium or
is due to slow reinitiation of the leaving group radicals from the
initial RAFT agent. The absolute values of the rate coefficients
governing the core equilibrium of the RAFT process (at a fixed
value of the equilibrium constant) are found to be crucial in
controlling the polydispersity of the resulting M w/M n values. Higher
interchange frequency effects narrower distributions. They also
demonstrated that the size of the rate coefficient controlling the
addition reaction of propagating radicals to polymer-RAFT agent,
K β, is mainly
responsible for optimizing the control of the polymerization. The
fragmentation rate coefficient, K−β, of the macro RAFT
intermediate radical, on the other hand, may be varied over orders
of magnitude without affecting the amount of control exerted over
the polymerization. Based on the basic RAFT mechanism, shown above,
its value mainly governs the extent of rate retardation in RAFT
polymerizations [278].
Calitz, Tonge, and Sanderson reported the results
of a study of RAFT polymerization by means of electron spin
resonance spectroscopy [276].
They observed intermediate radical signals that were not consistent
with current RAFT theory [276].
Sawamoto and coworkers reported obtaining
simultaneous control of molecular weight and steric structure in
RAFT polymerization of N-isopropylacrylamide by addition of
rare earth metal, Y(O-tetrafluoromethanesulfonate)3,
Lewis acid. The M
w/M n
ratio of the products ranged between 1.4–1.9 and the isotactic
content was 80–84% [277].
Goto et al. [279] developed a process that they describe as
reversible living chain transfer
radical polymerization [278], where they us Ge, Sn, P, and N compounds
iodides in the iodide mediated polymerizations.ref In
this process, a compound such as GeI4 is a chain
transferring agent and the polymer-iodide is catalytically
activated via a RFT process. They proposed that the new reversible
activation process be referred to as RTCP [279]. The process can be illustrated by them as
follows [279]:
3.14.4.1 Combinations of Click Chemistry and ATP as Well as ATP and RAFT Polymerizations
In the last few years, “click reactions,” as
termed by Sharpless et al. [280]
received attention due to their high specificity, quantitative
yields, and good fidelity in the presence of most functional
groups. The “click chemistry” reaction includes a copper-catalyzed
Huisgen dipolar cycloaddition reaction between an azide and an
alkyne leading to 1,2,3-triazole. Recent publications on this
“click reaction” indicate that it is a useful method for
preparation of functional polymers [281].
Matyjaszewski and Gao synthesized functional
polymers by combining ATRP and the “click reactions.” They also
prepared telechelic polymers, star polymers and brush polymers
[282]. Formation of telechelic
polymers was illustrated as follows:
Also, a synthesis of an iniferter that consists
of a trithiocarbonate moiety and two bromine chains ends was
reported [282]. This iniferter
was used to conduct either independently or concurrently both ATRP
and reversible addition-fragmentation chain transfer
polymerizations. The iniferter was illustrated as follows:
RAFT polymerizations with this iniferter of
N-butyl acrylate and
styrene yielded polymers with M w/M n equal to 1.15 and 1.16
respectively. Polymerization of methyl methacrylate, however,
yielded a polymer with a broad M w/M n ratio. On the other
hand, polymerization in the presence of CuBr/TMPA by ATRP
exclusively through the bromine chain ends yielded a polymer with
narrow M
w/M n
ratio [282].
3.14.5 Special Types of Controlled/“Living” Polymerizations
It was reported that it is possible to employ
persistent phosphorus-based radicals in controlled/living
free-radical polymerization [283,
284]. Also, in cases of low
stability of the hyper coordinated radicals, the ligand exchanges
become facile and some organoaluminum, organoboron, and other
compounds have been used successfully as transfer agents in
polymerization of styrene, acrylics, and vinyl acetate
[283, 284].
Chung and coworkers [286] described a “living” radical initiator
that is based on oxidation adducts of alkyl-9-borabicyclononanes,
such as hexyl-9-borabicyclononane. The “living” radical
polymerizations take place at room temperature. The initiators form
in situ by reactions with oxygen:
The alkoxy radicals are very reactive and
initiate radical polymerizations readily On the other hand, the
borinate radicals are stabilized by the empty p-orbitals of boron
through back-donating electron density and are too stable to
initiate polymerizations. During the polymerization, the borinate
radicals may form weak and reversible bonds with the growing
chains.
Boroxyl mediated living radical polymerizations
were subsequently described by Chung [287] in a review article. The chemistry is
centered on the living radical initiator, i.e.,
alkylperoxydialkylborane (C—O—O—BR) species, similar to the one
shown above, that are formed by mono oxidation of an asymmetric
trialkylborane with oxygen. In the presence of polar monomers
(including acrylates and methacrylates), the C—O—O—BR., undergoes a
spontaneous hemolytic cleavage at ambient temperature to form an
active alkoxy radical and a stable boroxyl radical The alkoxyl
radical is active in initiating the polymerization of vinyl
monomers, and the stable boroxyl radical forms a reversible bond
with the propagating radical site to prevent undesirable
termination reactions. The living polymerizations were
characterized by predictable polymer molecular weight, narrow
molecular weight distributions, and by formation of telechelic
polymers and block copolymers through sequential monomer addition.
Furthermore, this living radical initiator system benefits from two
unique features of trialkylborane. These are (a) easy incorporation
into polymer chains (chain ends or side chains) and (b) in situ
auto-transformation to living radical initiators.
Lacroix and coworkers reported a reverse iodine
transfer polymerization (RITP), where elemental iodine is used as a
control agent in living radical polymerization [288]. Styrene, butyl acrylate, methyl acrylate,
and butyl α-fluoroacrylate were homopolymerized, using a radical
catalyst and I2 as a chain transfer agent. Methyl
acrylate was also copolymerized with vinylidene chloride using this
process.
3.14.6 Kinetics of Controlled/Living Free-Radical Polymerizations
Several papers were published to describe the
kinetics of controlled free radical polymerization. Goto and Fukuda
[289] postulate two activation
processes for nitroxy/styrene polymerization systems:
The equilibrium constant, K = K d/K c. Stationary-state
concentrations of P• and X• are
The stationary concentration of P• and X• are
determined by different mechanisms. [P•] is determined by the
balance of the initiation rate R I and the termination rate
k t [P•]
[2]. This is the same as in
conventional free radical polymerization systems. [X•] is
determined, however, from the equilibrium equation shown for
process I. It depends, therefore upon the equilibrium constant
K and on the concentration
of the adduct [P–X] and [P•] [289]. The rate of polymerization during the
stationary state is
The polymer-nitroxyl adduct P–X reversibly
dissociates thermally, in process I into the polymer radical P• and
the nitroxyl radical X•. The rate constants of dissociation and
combination are k
d and k
c, respectively. The, so-called, “degenerative transfer”
takes place in process II. The second-order rate constant for
active species in either direction is k ex. Here all the rate
constants are assumed to be independent of chain length. Since the
frequency of cleavage of the P–X bond is proportional to [P–X] in
process I and to [P•1[P′–X]] in process II, the overall frequency,
f a per unit
time and per unit volume, of the bond-cleaving or activation
reactions, may be expressed by [277]:
with
where k
a is the overall activation rate constant, viewed as a
first-order reaction. Goto and Fukuda concluded that it may be more
convenient to represent the above equation in the form
[289]:
and show the general expression of the time-averaged k a for a batch system:
Matyjaszewski et al. wrote the kinetic equation
for atom transfer polymreization [290, 291]. It
is based on the ATP reaction mechanism that was described above. By
assuming fast initiations, insignificant termination reactions and
steady concentrations of the propagating radicals, the following
relationship was derived [290,
291]:
3.15 Thermodynamics of the Free-Radical Polymerization Reaction
3.15.1 Effects of Monomer Structure on the Thermodynamics of the Polymerization
There is a close relationship between monomer
structure and changes in free energy, in enthalpy and in entropy.
Thus, for instance, knowledge of changes in enthalpy will allow
appropriate thermal control of the reaction and yield proper rate
of propagation and molecular weight distribution. The quantities of
ΔF, ΔH, and ΔS relate only to the rate of
propagation because initiation and termination are single steps,
while propagation consists of multiple steps.
Free radical polymerization is generally
exothermic because it involves conversion of π bonds to σ bonds.
Thus, the change in enthalpy ΔH is negative. Also, because there is a
decrease in randomness in conversion of monomers to polymer, the
change in entropy ΔS is
also negative. The overall change in free energy of the free
radical polymerization process is,
The free energy is generally negative for the
free-radical polymerization process. Variations in monomer
structures have a significant effect on the values of ΔH for the following reasons. These are
differences in resonance stabilizations due to differences in
conjugation and hyperconjugation. Also, due to steric strains that
arise from bond angle deformation and bond stretching, as well as
variations in secondary bond forces, such as hydrogen bonding and
dipole interactions.
3.15.2 Thermodynamics of the Constrains of the Free-Radical Polymerization Reaction
Free-radical polymerization reactions are
equilibrium reactions. The equilibrium between the monomer and the
growing polymer is subject to thermodynamic conditions. At
equilibrium, therefore, the change in free energy is zero:
The change in free energy for the reaction can,
therefore, be written;
In the above equation, ΔF 0, ΔH 0, and ΔS 0 represent statistical
variations in the changes in free energy, enthalpy, and entropy,
representing the transition that the monomer undergoes by being
placed into the polymeric chain.
The equilibrium constant can then be written as:
where, PM*n+1 and PM*n are concentrations of
species. Assuming that they are practically equal, one can write:
it would then mean that:
The ceiling temperature can then be written as
3.16 Polymer Preparation Techniques
Four general techniques are used for preparation
of polymers by free-radical mechanism: polymerization in
bulk, in solution, in suspension and in emulsion. The bulk or mass polymerization is probably the
simplest of the four methods. Only the monomer and the initiator
are present in the reaction mixture. It makes the reaction simple
to carry out, though the exotherm of the reaction might be hard to
control, particularly if it is done on a large scale. Also there is
a chance that local hot spots might develop. Once bulk
polymerization of vinyl monomers is initiated, there can be two
types of results, depending upon the solubility of the polymer. If
it is soluble in the monomer, the reaction may go to completions
with the polymer remaining soluble throughout all stages of
conversion. As the polymerization progresses, the viscosity of the
reaction mixture increases markedly. The propagation proceeds in a
medium of associated polymeric chains dissolved in or swollen by
the monomer until all the monomer is consumed.
If the polymer is insoluble, it precipitates out
without any noticeable increase in solution viscosity. Examples of
this type of a reaction can be polymerizations of acrylonitrile or
vinylidene chloride. The activation energy is still similar to most
of the polymerizations of soluble polymers and the initial rates
are proportional to the square root of initiator concentration.
Also, the molecular weights of the polymerization products are
inversely proportional to the polymerization temperatures and to
initiator concentrations. Furthermore, the molecular weights of the
resultant polymers far exceed the solubility limits of the polymers
in the monomers. The limit of acrylonitrile solubility in the
monomer is at a molecular weight of 10,000. Yet, polymers with
molecular weights as high as 1,000,000 are obtained by this
process. This means that the polymerizations must proceed in the
precipitated polymer particles, swollen and surrounded by monomer
molecules.
The kinetic picture of free-radical
polymerization applies best to bulk polymerizations at low points
of conversion. As the conversion progresses, however, the reaction
becomes complicated by chain transferring to the polymer and by gel
effect. The amount of chain transferring varies, of course, with
the reactivity of the polymer radical.
Bulk polymerization is employed when some special
properties are required, such as high molecular weight or maximum
clarity, or convenience in handling. Industrially, bulk
polymerization in special equipment can have economic advantages,
as with bulk polymerization of styrene. This is discussed in
Chap.
6.
Solution polymerization differs
from bulk polymerization because a solvent is present in the
reaction mixture. The monomer may be fully or only partially
soluble in the solvent. This, the polymer may be (1) completely
soluble in the solvent, (2) only partially soluble in the solvent,
and (3) insoluble in the solvent.
When the monomer and the polymer are both soluble
in the solvent, initiation and propagation occur in a homogeneous
environment of the solvent. The rate of the polymerization is
lower, however, than in bulk. In addition, the higher the dilution
of the reactants the lower is the rate and the lower is the
molecular weight of the product. This is due to chain transferring
to the solvent. In addition, any solvent that can react to form
telomers will also combine with the growing chains.
If the monomer is soluble in the solvent, but the
polymer is only partially soluble or insoluble, the initiation
still takes place in a homogeneous medium. As the chains grow,
there is some increase in viscosity that is followed by
precipitation. The polymer precipitates in a swollen state and
remains swollen by the diffused and adsorbed monomer. Further
propagation takes place in these swollen particles.
Because propagation continues in the precipitated
swollen polymer, the precipitation does not exert a strong effect
on the molecular weight of the product. This was demonstrated on
polymerization of styrene in benzene (where the polymer is soluble)
and in ethyl alcohol (where the polymer is insoluble). The average
molecular weight obtained in benzene at 100°C was 53,000 while in
ethyl alcohol at the same temperature it was 51,000 [280]. When the monomers are only partially
soluble and the polymers are insoluble in the solvents the products
might still be close in molecular weights to those obtained with
soluble monomers and polymers. Polymerization of acrylonitrile in
water can serve as an example. The monomer is only soluble to the
extent of 5–7% and the polymer is effectively insoluble. When
aqueous saturated solutions of acrylonitrile are polymerized with
water-soluble initiators, the systems behave initially as typical
solution polymerizations. The polymers, however, precipitate out
rather quickly as they form. Yet, molecular weights over 50,000 are
readily obtainable under these conditions.
There are different techniques for carrying out
solution polymerization reactions. Some can be as simple as
combining the monomer and the initiator in a solvent and then
applying agitation, heat and an inert atmosphere [292]. Others may consist of feeding into a
stirred and heated solvent the monomer or the initiator, or both
continuously, or at given intervals. It can be done throughout the
course of the reaction or through part of it [293]. Such a set up can be applied to
laboratory preparations or to large-scale commercial preparations.
It allows a somewhat better control of the exotherm during the
reaction.
In both techniques the initiator concentration
changes only a few percent during the early stages of the reaction,
if the reaction temperature is not too high. The polymerization
may, therefore, approach a steady state character during these
early stages. After the initial stages, however, and at higher
temperatures, the square root dependence of rates upon the
initiator concentration no longer holds. This is a result of the
initiator being depleted rapidly. The second technique, where the
initiator, or the monomer and the initiator are added continuously
was investigated at various temperatures and rates of addition
[294–299]. If the initiator and monomer are
replenished at such a rates that their ratios remains constant,
steady state conditions might be extended beyond the early stages
of the reactions. How long they can be maintained, however, is
uncertain.
Suspension polymerization
[298] can be considered as a form
of mass polymerization. It is carried out in small droplets of
liquid monomer dispersed in water or some other media and caused to
polymerize to solid spherical particles. The process generally
involves dispersing the monomer in a non solvent liquid into small
droplets. The agitated stabilized medium usually consists of
nonsolvent (often water) containing small amounts of some
suspending or dispersing agent. The initiator is dissolved in the
monomer if it is a liquid or it is included in the reaction medium,
if the monomer is a gas.
To form a dispersion, the monomer must be quite
insoluble in the suspension system. To decrease the solubility and
to sometimes also increase the particle size of the resultant
polymer bead, partially polymerized monomers or prepolymers may be
used. Optimum results are obtained with initiators that are soluble
in the monomer. Often, no differences in rates are observed between
polymerization in bulk and suspension. Kinetic studies of styrene
suspension polymerization have shown that all the reaction steps,
initiation, propagation, and termination, occur inside the
particles [299].
The main difficulty in suspension polymerization
is in the forming and in the maintaining uniform suspensions. This
is because the monomer droplets are slowly converted from thin
immiscible liquids to sticky viscous materials that subsequently
become rigid granules. The tendency is for the sticky particles to
attach to each other and to form one big mass. The suspending
agent’s sole function is to prevent coalescing of the sticky
particles. Such agents are used in small quantities (0.01–0.5% by
weight of the monomer). There are many different suspending agents,
both organic and inorganic. The organic ones include
methylcellulose, ethyl cellulose, poly(acrylic acid),
poly(methacrylic acid), salts of these acids, poly(vinyl alcohol),
gelatins, starches, gums, alginates, and some proteins, such as
casein or zein. Among the inorganic suspending agents can be listed
talc, magnesium carbonate, calcium carbonate, calcium phosphate,
titanium and aluminum oxides, silicates, clays, such as bentonite,
and others. The diameter of the resultant beads varies from 0.1 to
5 mm and often depends upon the rate of agitation. It is
usually inversely proportional to the particle size. Suspension
polymerization is used in many commercial preparations of
polymers.
Zhang, Fu, and Jiang, reported a study of factors
influencing the size of polystyrene microspheres in dispersion
polymerization [300]. The found
that that the size of polystyrene microspheres decreased with an
increasing amount of stabilizer and also increased with increasing
the amount of monomer and initiator. The amount of stabilizer and
monomer concentration were the major factors influencing the size
distribution of polystyrene microspheres. The size of the
microspheres decreased with an increase of the solvency of reaction
media. The size distribution, however, hardly changed. The size of
polystyrene microspheres increased with an increase in the reaction
temperature. but the size distribution hardly changed.
Emulsion polymerization is used
widely in commercial processes [300, 301].
The success of this technique is due in part to the fact that this
method yields high-molecular-weight polymers. In addition, the
polymerization rates are usually high. Water is the continuous
phase and it allows efficient removal of the heat of
polymerization. Also, the product from the reaction, the latex, is
relatively low in viscosity, in spite of the high molecular weight
of the polymer. A disadvantage of the process is that water-soluble
emulsifiers are used. These are hard to remove completely from the
polymers and may leave some degree of water sensitivity.
The reaction is commonly carried out in water
containing the monomer, an emulsifier or a surface-active agent,
and a water-soluble initiator. Initiation may be accomplished
through thermal decomposition of the initiator or through a redox
reaction. The polymer forms as a colloidal dispersion of fine
particles and polymer recovery requires breaking up the
emulsion.
The full mechanism of emulsion polymerization is
still not completely worked out. It is still not clear why a
simultaneous increase in the polymerization rate and in the
molecular weight of the product is often observed. Also, in
emulsion polymerization, at the outset of the reaction the monomer
is in a form of finely dispersed droplets. These droplets are about
1 μ in diameter. Yet, during the process of a typical
polymerization, they are converted into polymer particles that are
submicroscopic, e.g., 1,000 Å in diameter.
At the start of the reaction the emulsifier
exists simultaneously in three loci: (a) as a solute in water; (b)
as micelles; (c) and as a stabilizing emulsifier at the interface
between the monomer droplets and the water. The bulk of the
emulsifier, however, is in the micelles. The monomer is also
present in three loci: (a) in the monomer droplets that are
emulsified and perhaps 1–10 μ in diameter; (b) it is
solubilized in the micelles, perhaps 50–100 Å in diameter; (c)
and it is present as individual molecules dissolved in the water.
The bulk of the monomer is in the droplets. There are on the
average 1018/mL of monomer-swollen micelles in the
reaction mixture at the outset of the reaction [302]. At the start of the reaction there are
also on the average 1012/mL monomer droplets that act as
reservoirs. The monomer is supplied from the droplets to
radical-containing micelles when the reaction progresses by a
process of diffusion through the aqueous phase
(Fig. 3.3).
Fig.
3.3
Early stages of emulsion polymerization
(from ref. [306])
The first hypothesis of the mechanism of emulsion
polymerization was formulated by Harkins [305]. According to this hypothesis, the
water-soluble initiator decomposes in the aqueous phase. This
results in formation of primary radicals. The primary radicals in
turn react with the monomer molecules dissolved in the water
(though their number may be quite small). Additional monomer
molecules may add to the growing radicals in the water until the
growing and propagating chains of free radicals acquire
surface-active properties. At that stage, the growing radicals
consist of inorganic and organic portions:
These growing radical-ions tend to diffuse into
the monomer-water interfaces. The probability that the diffusion
takes place into monomer-swollen micelles rather than into monomer
droplets is backed by the considerations of the relative surface
areas of the two. There are on the average 1018 micelles
in each milliliter of water. These are approximately 75 Ā in
diameter and each swollen micelle contains on the average 30
molecules of the monomer. At the same time, the diameters of the
monomer droplets are approximately 1 μ. and it is estimated
that there are only approximately 1012 such droplets per
milliliter of water. Thus, the micelles offer 60 times more
surfaces for penetration than do the droplets. The initiating
radicals are almost always generated in the water phase. After
formation in the water phase, a number of free radicals may be lost
due to recombination. Termination is also possible after reaction
of free radicals with some of the monomers dissolved in the
water.
Several theories tried to explain the entry
process. Thus, a “diffusion control” model [307, 308]
supposes that diffusion of aqueous-phase radicals into the particle
surface is the rate-controlling step for entry. Another theory
suggests that displacement of surfactant from the particle surface
is the rate-determining step [309]. A third one assumes that the entry can be
thought of as a colloidal interaction between a latex particle and
primary phase oligomeric aqueous-phase radical. These are the
radicals formed through reactions of initiating radicals and
monomer molecules dissolved in water [310]. The most accepted entry model appears to
be the “control by aqueous-phase growth” model of Maxwells et al.
[311]. This theory postulates
that free radicals generated in the aqueous phase propagate until
they reach a critical degree of polymerization (let us call it
z), at which point they
become surface-active and their only fate is irreversible entry
into a latex particle; the rate of entry of z-mers into a particle is assumed to be
so fast as not to be rate-determining. An efficiency of less than
100% arises if there is significant aqueous-phase termination of
the propagating radicals.
The entry model of Maxwells et al. was derived
from and/or supported by data on the influence of particle surface
characteristics (charge, size) on the entry rate coefficient
[312]. It was assumed that the
aqueous radicals became surface active when the degree of
polymerization reached 2–3. This was based on thermodynamic
considerations of the entering species.
Further data on the Maxwell et al. entry model
was obtained by Gilbert and coworkers [313] who studied the effects of initiator and
particle surface charges. They obtained kinetic data for radical
entry in the emulsion polymerization of styrene and concluded that
their data further supports the Maxwell et al. entry model and
refutes the alternative models mentioned above.
Once the radicals penetrate the micelles,
polymerization continues by adding monomers that are inside. The
equilibrium is disturbed and the propagation process proceeds at a
high rate due to the concentration and crowding of the stabilized
monomers. This rapidly transforms the monomer-swollen micelles into
polymer particles. The changes result in disruptions of the
micelles by growths from within. The amount of emulsifier present
in such changing micelles is insufficient to stabilize the polymer
particles. In trying to restore the equilibrium, some of the
micelles, where there is no polymer growth, disintegrate and supply
the growing polymer particles with emulsifier. In the process many
micelles disappear per each polymer particle that forms. The final
latex usually ends up containing about 1015 polymer
particles per milliliter of water. By the time conversions reach
10–20% there are no more micelles present in the reaction mixtures.
All the emulsifier is now adsorbed on the surface of the polymer
particles. This means that no new polymer particles are formed. All
further reactions are sustained by diffusion of monomer molecules
from the monomer droplets into the growing polymer particles. The
amount of monomer diffusing into the particles is always in excess
of the amount that is consumed by the polymerization reaction due
to osmotic forces [297]. This
extra monomer supplied is sufficient for equilibrium swelling of
the particles [298]. As a result,
the rate of polymerization becomes zero order with respect to
time.
When conversion reaches about 70%, all the
remaining monomer is absorbed in the polymer particles and there
are no more monomer droplets left. At this point the reaction rate
becomes first order with respect to time.
The qualitative approach of Harkins was put on a
quantitative basis by Smith and Ewart [314–316].
Because 1013 radicals are produced per second and can
enter between 1014 and 1015 particles, Smith
felt that a free radical can enter a particle once every
10–100 s. It can cause the polymerization to occur for
10–100 s before another free radical would enter and terminate
chain growth [317]. A period of
inactivity would follow that would last 10–100 s and then the
process would repeat itself. Such a “stop and go” mechanism implies
that a particle contains a free radical approximately half of the
time. It can also be said that the average number of radicals per
particle is 0.5. This is predicted on conditions that (a) the rate
of chain transfer out of the particle is negligible and (b) the
rate of termination is very rapid compared with the rate of radical
entry into the particle.
The kinetic relationships derived by Smith and
Ewart for the system are as follows:
where, k P is
the constant for propagation, [M] is the concentration of monomer,
N is the number of
particles containing n
radicals (~0.5) and the expression for the number of particles
formed:
where, μ is the volume
increase of the particles, A S is the area occupied by
one emulsifier molecule. S
is the amount of emulsifier present. K is a constant = 0.37 (based on the
assumption that the micelles and polymer particles compete for free
radicals in proportion to their respective total surface areas).
K can also be equal to 0.53
(based on the assumption that the primary radicals enter only
micelles, as long as there remain micelles in the reaction
mixture). ρ is the rate of
entry into the particles. The kinetic chain length can be written
as:
The Smith-Ewart mechanism does not take into
account any polymerization in the aqueous phase. This may be true
for monomers that are quite insoluble in water, such as styrene,
but appears unlikely for more hydrophilic ones such as methyl
methacrylate or vinyl acetate. In addition, it was calculated by
Flory that there is insufficient time for a typical cation-radical
(such as a sulfate ion radical) to add to a dissolved molecule of
monomer such as styrene before it becomes captured by a micelle
[317]. This was argued against,
however, on the ground that Flory’s calculations fail to consider
the potential energy barrier at the micelle surfaces from the
electrical double layer. This barrier would reduce the rate of
diffusion of the radical-ions into the micelles [316].
Considerably different mechanisms were proposed
by several groups [317,
318]. They are based on a concept
that most polymerizations must take place at the surface of the
particles or in their outer “shell” and not within the particles.
It is claimed that the interiors of the particles are too viscous
for free radicals to diffuse inside at a sufficiently fast rate.
Two different mechanisms were proposed to explain why
polymerization takes place preferentially in the shell layer. One
of them suggests that the monomer is distributed nonuniformly in
the polymer particles. The outer shell is rich in monomer
molecules, while the inside is rich in polymer molecules
[319]. The other explanation is
that the radical ions that form from the water-soluble initiator
are too hydrophilic to be able to penetrate the polymer particles
[320].
Surfactant - free emulsion polymerization are carried out in
the absence of a surfactant [321]. The technique requires the use of
initiators that yields initiating species with surface-active
properties and imparts them to the polymer particles. Examples of
such initiators are persulfates. The lattices that form are
stabilized by chemically bound sulfate groups that are derived from
persulfate ions. Because the surface-active groups are chemically
bound, the lattices are easier to purify and free the product from
unreacted monomer and initiator. Generally, the particle number per
milliliter from a surfactant-free emulsion polymerization is
smaller than the particle number from typical emulsion
polymerization.
In an inverse emulsion polymerization an aqueous solution
of a hydrophilic monomer is emulsified in an organic solvent and
the polymerization is initiated with a solvent soluble initiator.
This type of emulsion polymerizations is referred to as
water in oil polymerization. Inverse
emulsion polymerization is used in various commercial
polymerizations and copolymerization of water-soluble monomers.
Often nonionic emulsifiers are utilized. The product emulsions are
often less stable than the oil in water emulsions.
A special approach to emulsion polymerization is
called miniemulsion
polymerization [322].
These reactions contain both micelles and monomer droplets, but the
monomer droplets are smaller than in macrosystems. Usually, a
water-soluble surfactant is used for emulsification. An example of
such a surfactant can be sodium dodecyl sulfate. In addition, a
highly water-insoluble costabilizer is added, such as hexadecanol.
Thus, miniemulsions are dispersions of critically stabilized oil
droplets with a size between 50 and 500 nm prepared by
shearing a system containing oil, water, a surfactant and a
hydrophobic material. Polymerizations in such miniemulsions, when
carefully prepared, result in latex particles which have about the
same size as the initial droplets. An appropriate formulation of a
miniemulsion suppresses coalescence of droplets. The polymerization
of miniemulsions extends the possibilities of the widely applied
emulsion polymerization and provides advantages with respect to
copolymerization reactions of monomers with different polarity,
incorporation of hydrophobic materials or with respect to the
stability of the formed latexes. Although labeled “emulsion,” it
appears that some may involve a combination of emulsion and
suspension polymerizations. It was reported [323] that by using a difunctional alkoxyamine
as an initiator for the homopolymerization of butyl acrylate in
miniemulsion, to increase the achievable molar mass and to use the
polymer as a difunctional macroinitiator for the synthesis of
triblock copolymers in aqueous dispersed systems. Well-defined
polymers with one alkoxyamine functionality at each end were
obtained, providing that monomer conversion was kept below 70%.
Beyond this conversion, extensive broadening of the molar mass
distribution was evidenced, as the consequence of termination and
transfer to polymer.
Tsavalas et al. [324] reported that a phenomenon seemingly
unique to hybrid miniemulsion polymerization was observed by them,
where monomer conversion would either plateau at a limiting value
or quickly switch to a dramatically lesser rate. They attributed
this phenomenon to a combination of three factors. The first one is
the degree to which the monomer and resinous component are
compatible. The second is the resultant particle morphology after
approximately 80% monomer conversion, which roughly corresponds to
the portion of reaction where this morphology is established. The
third factor is the degree of interaction between the growing
polymer and the resin (a grafting reaction). Of these three, the
first two factors were found by them to be much more significant in
contributing to the limiting conversion.
RAFT
emulsion polymerization is a new development
that has attracted considerable attention. It be carried out in a
regular emulsion polymerization [325] and in a reverse emulsion polymerization
[326].
Also, recently, several reports in the literature
have described miniemulsion RAFT polymerizations. In some
instances, use is made of water-soluble RAFT agents to control
polymer molecular weight [327].
Also, Hawkett and coworkers reported using surface active RAFT
agents to emulsify the dispersed phase, stabilize the particles and
also control the molecular weight. This yielded polymer latexes
that were free from surfactant and costablilizer [328]. One of these special RAFT agents was
illustrated as follows:
The other two RAFT agents used by them had
similar structures. A surface active iniferter was also reported by
Choe and coworkers [330,
331]:
This RAFT agent allowed polymerization of methyl
methacrylate initiated by ultraviolet light irradiation in the
absence of added surfactant or initiator.
Rieger and coworkers [332] reported a surfactant free RAFT emulsion
polymerization of butyl acrylate and styrene using
poly(N,N-dimethylacrylamide) trithiocarbonate
macromolecular transfer agent. They observed that the
polymerizations were fast and controlled with molar masses that
matched well the theoretical values and low polydispersity indexes.
Monomer conversions close to 100% were reached and the
polymerizations behaved as controlled systems, even at 40% solids
contents. The products were poly(N,N-dimethyl
acrylamide)-b-poly(n-butyl
acrylate) and poly(N,N-dimethylacrylamide)-b-polystyrene
amphiphilic diblock copolymers formed in situ.
3.17 Review Questions
3.17.1 Section 3.1
1.
What are the three steps in free-radical
polymerization? Illustrate each step in free-radical chain
polymerization.
2.
What is the rate-determining step in free-radical
polymerization?
3.
Write the kinetic expressions for initiation,
propagation, termination, and transfer.
4.
What is the steady state assumption? How is it
expressed? Why is it necessary?
5.
What is the expression for the rate of
propagation? Rate of monomer disappearance? The average lifetime of
a growing radical under steady state conditions?
6.
Why is an initiator efficiency factor needed for
the rate equation? What is a kinetic chain length?
7.
In polymerization of styrene in benzene at 60°C
using 0.1 mol benzoyl peroxide initiator and 1 mol of
monomer, k d is
1.38 × 10─5, assume steady state and calculate the
free-radical concentration during the reaction. If k p is 176/mol what is the
rate of propagation? What is the lifetime of a growing radical if
k t is
7.2 × 10─5? What is the rate of propagation if the
initiator efficiency is 72%?
3.17.2 Section 3.2
1.
What sources of initiating free radicals do you
know? Illustrate the decomposition reaction of
α,α′-azobisisobutyronitrile. Illustrate how free radicals can
recombine inside or outside the solvent cage and be lost to the
initiation process.
2.
Illustrate one or more inorganic and also one or
more organic peroxides.
3.
How does solvent “cage” affect the initiating
free radicals? Explain and illustrate.
4.
Explain homolytic and heterolytic cleavage of
peroxides.
5.
Explain and give chemical equations for
redox initiations with
Fe++, Co++, and Ce++++ ions in the reaction mixture.
6.
How are peroxides such as benzoyl peroxide
decomposed by aromatic tertiary amines. Show the two postulated
mechanisms for the reaction of benzoyl peroxide with dimethyl
aniline.
3.17.3 Section 3.3
1.
Describe the reaction of the initiating free
radical with the monomer. Show this reaction with equations, using
a phenyl initiating radical and styrene monomer as an
example.
2.
Do the same as question 1, but with a redox
mechanism, showing a sulfate ion-radical adding to vinyl
acetate.
3.17.4 Section 3.4
1.
Illustrate the transition state in the
propagation reaction.
2.
Explain the steric. polar and resonance effects
in the propagation reaction.
3.
Explain why there is a tendency for a
trans–trans placement in the propagation
reactions when carried out at low temperatures.
4.
How does the reaction medium affect the
propagation reaction?
5.
What is ceiling temperature and what is the
kinetic expression for this phenomenon?
6.
Explain what is meant by autoacceleration and how
does it manifest itself.
7.
What is cyclopolymerization? Explain and give
several examples.
3.17.5 Section 3.5
1.
What are the three termination processes in
free-radical polymerization?
2.
What is meant by degenerative chain transferring?
Illustrate back-biting. The telomerization reaction.
3.
What is meant by chain transferring
constants?
4.
Write the equation for the degree of
polymerization including all the chain transferring constants. In a
benzoyl peroxide initiated polymerization of 2 moles of styrene in
benzene at 85°C (K
d = 8.94 × 10−5 L/mol-s at 85°C). How
much benzoyl peroxide will be required in the polymerization
solution to attain an average molecular weight of 250,000? Assume
that termination occurs only by recombination and no
chain-transferring takes place.
5.
In the above polymerization, the transferring
constant to monomer, C
M × 10─4 = 3.74, the transferring constant to
solvent, C
S = 5.6 × 10─4 and transferring constant to
the initiator, C
I = 0.75. Assuming that f = 0.72, and k t is
7.2 × 10─5, k
p is 176/mol, (k
d is shown above) calculate R p and the average
DP.
3.17.6 Section 3.6
1.
Explain what is meant by reactivity ratios and
how they are derived.
2.
Write the copolymerization equation. In a
copolymerization of 1 mol of styrene with 1 mol of
butadiene, r
1 = 0.78 and r
2 = 1.39, what is the expected composition of the
copolymer at the early stages of the polymerization?
3.
How do substituents on the monomer molecules
affect reactivity of the monomers toward attacking radicals?
4.
Explain the Q and e scheme and write the Price–Alfrey
equation.
5.
How can r
1 and r
2 be derived from the Q and e values. Show the relationship.
6.
From chemical structures alone predict the
products from free-radical copolymerizations of pairs of (1)
styrene and methyl methacrylate, (2) styrene and vinyl acetate, (3)
methyl methacrylate and vinyl chloride. Consult
Table 3.8
for reactivity ratios.
3.17.7 Section 3.7
1.
How many reactivity ratios are there in a
terpolymerization?
2.
Write the equation for the terpolymerization
reaction.
3.17.8 Section 3.8
1.
What is allylic polymerization? If allyl alcohol
does not polymerize to a high-molecular-weight polymer by
free-radical polymerization, why does triallyl cyanurate form a
high-molecular-weight network structure by the same
mechanism?
3.17.9 Section 3.9
1.
What is inhibition and retardation?
Explain.
2.
Give an example of a good inhibitor and a good
retarder and show by chemical equations the reaction with free
radicals.
3.
Show the reaction of quinone with free
radicals.
4.
Write the equation that relates rate data to
inhibited polymerizations.
5.
Calculate R p for the polymerization
of 1 mol of styrene containing 0.01 mol of hydroquinone
inhibitor. with k
p = 176/mol and k
t = 7.2 × 10─5.
3.17.10 Section 3.10
1.
What is thermal polymerization? Show by chemical
equations the postulated mechanism of formation of initiating
radicals in styrene thermal polymerization.
3.17.11 Section 3.11
1.
Show the proposed charge-transfer mechanism for
copolymerization of styrene with maleic anhydride and dioxene with
maleic anhydride.
2.
What determines the stability of charge-transfer
complexes? Explain and give examples.
3.17.12 Section 3.12
1.
How do some polar monomers complex with Lewis
acids? How does that affect polymerization of these monomers?
Copolymerization?
3.17.13 Section 3.13
1.
How can canal complexes be used for steric
control in free-radical polymerization? Give examples.
3.17.14 Section 3.14
1.
How do controlled/“living” polymerizations differ
from typical living polymerizations?
2.
List the different types of controlled/“living”
free-radical polymerizations that you can think of. What are the
three classes of homogeneous controlled/“living” polymerizations as
described by Matyjaszewski? Illustrate.
3.
Describe cobalt mediated polymerizations.
Illustrate.
4.
Describe and illustrate atom transfer
polymerizations controlled by copper/bipyridine complex and by
carbon tetrachloride, dichloro(triphenyl-phosphine)-ruthenium(II),
and methylaluminum bis(2,6-di-tert-butyl-phenoxide). Explain what
(ARGENT)ATRP and (SET-LRP) mean. Illustrate the proposed
Percec mechanism and the Matyjaszewski mechanisms.
5.
Describe nitroxyl radical mediated
polymerizations. Illustrate TEMPO controlled polymerization of
styrene. What is meant by in situ NMP?
6.
Explain and illustrate a reversible
addition-fragmentation chain transfer polymerization (RAFT).
7.
Write the Matyjaszewski proposed kinetic equation
of ATP polymerization.
3.17.15 Section 3.15
1.
What is meant by bulk or mass polymerization?
Explain and discuss.
2.
What are some of the techniques for carrying out
solution polymerizations?
3.
Give a qualitative picture of emulsion
polymerization as described by Harkins.
4.
How did Smith and Ewart put the Harkins picture
of emulsion polymerization on a quantitative basis? What is the
equation for the rate of emulsion polymerization?
5.
Describe a surfactant-free emulsion
polymerization, and inverse emulsion polymerization and a
miniemulsion polymerization.
3.17.16 Section 3.16
1.
Discuss the effect of monomer structure on the
thermodynamics of the free-radical polymerization process.
3.17.17 Recommended Reading
-
G. Moad and D.H. Solomon, The Chemistry of Radical Polymerization, Pergamon Press, Oxford, 1995.
References
1.
T. Fueno, T. Tsuruta, and J.
Furukawa, J. Polymer
Sci ., 1959, 40, 487
2.
G.E. Scott and E. Serrogles,
J. Macromol . Sci .
- Rev . of
Macromol .
Chem ., 1973, C9(1), 49; C.H. Bamford, “Radical
Polymerization” in Encyclopedia
of Polymer Science
and Engineering, Vol.13,
H.F.Mark, N.M. Bikales, C.G. Overberger, and G. Menges. (eds.),
Wiley-Interscience, New York, 1988
3.
S. Inoue, Chapter 5 in
Structure and Mechanism in Vinyl Polymerization, T. Tsuruta and K.
F. O’Driscoll, (eds.) Dekker, New York, 1969
4.
J. Thiele and K. Hauser,
Ann ., 1896, 290,
5.
D. Swern. ed, Organic Peroxides, Wiley-Interscience,
New York, 1970; S. Patai,
(ed)., The Chemistry
of Peroxides,
Wiley-Interscience, New York, 1983
6.
I.A. Saad and F. R. Eirich,
Am . Chem . Soc .,Polymer Preprints, 1960, 1, 276,
7.
8.
F.H. Solomon and G. Moad,
Makromol . Chem .,Macromol . Symp .,1987, 10/11, 109
9.
P.D. Bartlett and R. R.
Hiatt, J. Am . Chem . Soc .,1958, 80, 1398,
10.
11.
A.V.Tobolsy and
R.B.Mesrobian, “Organic Peeroxides”,Wiley, New York, 1954
12.
M.G. Evans, Trans. Faraday Soc.,1946, 42, 101
13.
C.G. Swain, W.T. Stockmayer,
and J.T. Clarke, J. Am
. Chem . Soc .,
1950, 72, 5726
14.
P.B. Bartlett and K. Nozaki,
J. Am. CHem. Soc., 1946, 68, 1686; ibid., 1947, 69, 2299
15.
T. Koenig, Decomposition of Peroxides and Azoalkanes, Ch.3 in Free-Radicals, Vol.I, J.K. Kochi, (ed.),
Wiley, New York, 1973
16.
17.
A.T. Blomquist and A.J.
Buselli, J. Am . Chem . Soc ., 1951, 73, 3883; C.S. Sheppard, “Peroxy
Compounds”, in Encyclopedia
of Polymer Science
and Engineering, Vol.11,
H.F. Mark, N.M. Bikales, C.G. Overberger, and G. Menges, eds.,
Wiley-Interscience, New York, 1988
18.
M.G. Evans, Trans . Faraday Soc .,1946, 42, 101
19.
L. Horner and E. Schlenk,
Angew. Chem, 1949, 61, 711
20.
M. Imoro and S. Choe,
J. Polymer Sci., 1955,15, 482
21.
S.S. Harihan and M.
Maruthamuthu, Makromol
. Chem ., 1979, 180, 2031
22.
23.
V. Horanska, J. Barton, and
Z. Manasek, J. Polymer
Sci ., 1972, A-1,10, 2701
24.
25.
B. Yamada, H. Kamei, T.
Otsu, J. Polymer
. Sci.,Chem. Ed., 1980, 18, 1917
26.
L. Michl, S. Korbe,K.
Vyakaranam, and J. B. Barbour, J.
Am. Chem. Soc.,
2006, 128, 5610 and 5680
27.
G.S. Kolesnikov and L.S.
Fedorova, Izv . Akad . Nauk
SSSR,Otd
. Khim . Nauk, 1957, 236
28.
J. Furukawa and T. Tsuruta,
and S.Inoue, J. Polymer
Sci .,1957, 26, 234
29.
N. Ashikari and N.
Nishimura, J. Polymer
Sci ., 1958, 28, 250
30.
31.
J. Furukawa and T. Tsuruta,
J. Polymer Sci ., 1958, 28, 227
32.
C.E.H. Bawn, D. Margerison,
and N.M. Richardson, Proc
. Chem . Soc . (London), 1959, 397
33.
34.
35.
S. Inoue, Chapter 5 in
Structure and Mechanism of Vinyl Polymerization, T. Tsuruta and
K.F. O’Driscoll, eds., Dekker, New York, 1969
36.
A.L. Barney, J.M. Bruce Jr.,
J. N. Coker, H.W. Jackobson, and W. H. Sharkey, J. Polymer Sci., 1966, A-1,4, 2617
37.
Y. Nakoyama, T. Tsuruta, and
J. Furukawa, Makromol
. Chem .,1960, 40, 79
38.
A. Charlseby,
Chap.1.,“Irradiation of Polymers”, N. Platzer, (ed.), Am. Chem. Soc., Advancement in Chem. Series #66, Washington, 1967; J.G. Calvin and J.N. Pitts, Jr.,
Photochemistry, Wiley, New
York, 1967; N. J. Turo,
Molecular Photochemistry,
Benjamin, New York, 1967
39.
J.E. Wilson, Radiation Chemistry of Monomers, Polymers, and Plastics, Dekker, New York,
1974
40.
Z. Szablan, T. M. Lovestead,
T. P. Davis, M. H. Stenzel, and C. BarnerKowollik, Macromolecules, 2007, 40, 26CrossRef
41.
J. Frank and E. Rabinowitch,
Trans . Faraday Soc .,1934,30, 120
42.
W.A. Pryor Free Radicals, McGraw-Hill, New York,
N.Y., 1966
43.
J.C. Martin, “Solvation and
Association”, Chapt.20 in Free-Radicals, Vol.II, J.K. Kochi, (ed.),
Wiley New York, 1973
44.
45.
B. Giese, Am. Chem. Soc. Polymer Preprints,1997, 38 (1), 639
46.
H. Zipse, 1. He, K. N. Houk,
B. Giese, 113 J. Am. Chem. Sec. 1991, 4324
47.
B. Giesce, J. He, W. Mehl,
Chem. Ber., 1988, 121, 2063
48.
B. Giese. Angew. Chem. Int. Ed. Engl. 1983, 22, 753,(1983)
49.
B. Giese, W. Damm, R. Batra,
Chemtracts 1994, 7, 355
50.
B. Giese, M. Bulliard, J.
Dickhaut, R. Halbach, C. Hassler, U. Hoffmann, B. Hinzen, M.Senn,
Synlen 1995, 116
51.
J.C. Bevington and J. Toole,
J. Polymer Sci .,1958, 28, 413
52.
J.C. Bevington, “Radical
Polymerization”, Academic Press, London, 1961
53.
54.
J.M. Tedder, “Reactivity of
Free-Radicals”, Chapt.2, in “Reactivity, Mechanism, and Strucure in Polymer Chemistry”, A.D. Jenkins and A.
Ledwith, (eds.), Wiley-Interscience, New York, 1974
55.
R.J. Orr and H. L. Williams,
J. Am . Chem . Soc .,1955, 77, 3715
56.
57.
F. Leavitt, M. Levy, M.
Szwarc, and V. Stannett, J.
Am . Chem . Soc .,1955, 77, 5493
58.
59.
60.
61.
M. Hagiwara, H. Okamoto, T.
Kagiya, and T. Kagiya, J.
Polymer Sci
.,1970, A-1,8, 3295
62.
G. Smets and K. Hayashi,
J. Polymer Sci ., 1958, 27, 626,
63.
N.D. Field, J.R. Schaefgen,
J. Polymer Sci .,1962, 58, 533,
64.
M. Yamada, I. Takase,
Kobunshi Kagaku,
22, 626, (1965); from
Chem . Abstr ., 1966, 64, 19803g
65.
66.
B.M. Cubertson, “Maleic and
Fumaric Polymers” in “Encyclopedia
of Polymer Science
and Technology”, Vol.9,
H.F. Mark, N.M. Bikalis, C.G. Overberger, and G.Menges, (eds.),
Wiley- Interscience, New York, 1987
67.
T. Otsu, O. Ito, N. Toyoda,
and S. Mori, Macromol
. Chem .,Rapid
Commun .,1981, 2, 725
68.
T. Otsu, O. Ito and N.
Toyoda, Makromol
. Chem .,Rapid
Commun ., 1981, 2, 729
69.
70.
C. Walling and E. H. Huyser,
Vol. 13, Organic Reactions,
A.C. Cope, (ed.), Wiley, New York, 1963
71.
C.H. Bamford, W.G. Barb,
A.D. Jenkins, and P.F. Onyon, Kinetics of Vinyl Polymerization by Radical Mechanism, Butterworth, London,
1958
72.
G.S. Hammond, C. Wu, O.D.
Trapp, J. Warkentin, and R.T. Keys, Am Chem . Soc . Polymer Preprints, 1960, 1 (1), 168
73.
K. Hayashi, T. Yonezawa, C.
Nagata, S. Okamura, and F. Fukii, J. Polymer Sci .,1956, 20, 537
74.
K.F. O’Driscoll and T.
Yonezawa, Reviews of
Macromolecular
Chemistry,1966,
1 (1), 1
75.
M.L. Huggins, J. Am . Chem . Soc .,1961, 66, 1991
76.
C.G. Fox, B.S. Garret, W.E.
Goode, S. Gratch, J.F. Kinkaid, A. Spell, and J.D. Stroupe,
J. Am . Chem . Soc .,1958, 80, 1968
77.
J.W.L. Fordham, J. Polymer Sci ., 1959, 39, 321
78.
D.J. Cram, J. Chem . Ed .,1960, 37, 317
79.
A.M. North and D.
Postlethwaite, Chap. 4, Structure
and Mechanism in
Vinyl Polymerization, T.
Tsuruta and K.F. O’Driscoll, (eds), Dekker, New York, 1969
80.
C.H.E. Bawn, W.H. Janes, and
A.M. North, J. Polymer
Sci ., 1963, C,4 427
81.
P.H. Burleigh, J. Am . Chem . Soc .,1960, 82, 749
82.
I. Rosen, P.H. Burleigh, and
J.F. Gillespie, J. Polymer
Sci ., 1961, 54, 31
83.
G.M. Burnett and F.L. Ross,
J. Polymer Sci ., 1967, A-1,5,
84.
85.
T. Uryu, H. Shiroki, M.
Okada, K. Hosonuma and K. Matsuzaki, J. Polymer Sci ., 1971, A-2,9, 2335
86.
G. Blauer,J. Polymer Sci.,1953, 11, 189
87.
F.C. De Schryver, G. Smets,
and Van Thielen, J. Polymer
Sci., Polymer Letters,1968, 6, 54
88.
M. Kamachi, J. Satoh, S.
Nozakura, J. Polymer
Sci., Polymer Chem . Ed ., 1978, 16 (8), 1789
89.
T. Yamamoto, T. Yamamoto, A.
Moto, and M. Hiroto, Nippon
Kagaku Kaishi,
1979 (3), 408; from
Chem . Abstr ., (June 1979) 90 (12), Macromol. Section, 5,
90.
B. De Sterk, K.Hamelsoet, V.
Van Spaybroeck and M. Waroquier, Macromolecules, 2010, 43, 5602CrossRef
91.
F. Tudos, Acta Chim . Budapest, 1965,43, 397; 44, 403
92.
93.
W, Wunderlich, Makromol . Chem ., 1976, 177, 973
94.
G. Odian, Principles of Polymerization, 3-rd edit. J. Wiley,
New York, 1991
95.
J.M. Dionisio, H.K.
Mahabadi, J.H. O’Driscoll, E. Abuin, and E.A. Lissi, J. Polymer Sci., Chem . Ed ., 1979, 17, 1891; J.M. Dionisio and J.H.
O’Driscoll, J. Polymer
Sci., Chem . Ed ., 1980,18, 241; R. Sack, G.V. Schulz, and G.
Meyerhoff, Macromolecules,1988, 21, 3345
96.
97.
N.F. Phelan and M. Orchin,
J. Chem . Educ .,1968,45, 633
98.
99.
100.
P.J. Graham, E.L. Buhle,
and N.J. Pappas, J. Org
. Chem .,1961, 26, 4658
101.
R.H. Wiley, W.H. Rivera,
T.H. Crawford, and N.F. Bray J.
Polymer Sci
., 1962, 61, 538; Polymer Sci ., 1964, A-1,2, 5025
102.
N.L. Zutty, J. Polymer Sci ., 1963, A-1,1, 2231
103.
J.E. Fearn, D.W. Brown, and
L.A. Wall, J. Polymer
Sci.,1966,A-1,4, 131
104.
105.
R.J. Cotter and M. Matzner,
Ring-Forming Polymerizations, Academic
Press, New York, 1969
106.
G.B. Butler and M.A.
Raymond, J. Polymer
Sci ., 1965, A-1,3, 3413
107.
G.B. Butler, J. Polymer Sci ., 1960, 48, 279
108.
G.B. Butler, Am . Chem . Soc.,
Polymer Preprints,
1967, 8 (1), 35
109.
G.B. Butler and M.A.
Raynolds, J. Org
. Chem ., 1965, 30, 2410: G.B. Butler, T.W. Brooks,
J. Org . Chem ., 1963, 28, 2699
110.
S. Erkoc and A. Ersin Acar,
Macromoleculles,
2008, ASAP
article,10.1021/ma801492a, web release Nov.
6
111.
G. Oster and N.L. Yang,
Chem . Reviews, 1968, 68 (2), 125
112.
113.
S.R. Chatterjee, S.N.
Kanna, and S.R. Palit, J.
Indian Chem
. Soc ., 1964, 41, 622
114.
115.
116.
117.
118.
G. H. Schulz, G.
Henrici-Olive, Z.
Electrochem.,Ber.
Bunsenges Physik
. Chem ., 1956, 60, 296
119.
120.
D.A. Shipp, D.H. Solomon,
T.A. Smith, and G. Moad, Macromolecules, 2003, 36, 2032
121.
122.
123.
G. Odian, “Principles of
Polymerization”, Wiley, New York, IV ed., 2003
124.
K. Yukota amd M. Itoh,
J.Polymer Sci.,Polymer Letters, 1968, 6, 825,
125.
A. Ravve, “Light Associated
Reactions of Synthetic Polymers”, Springer, New York, 2006
126.
J.B. Brandrup and E.H.
Immergut, eds., Polymer
Handbook, 3 rd edit., J. Wiley, New York, 1989
127.
J. Furukawa, T. Tsuruta,T.
Fueno, R. Sakata, and K. Ito, Macromol. Chem.,1959, 30, 109
128.
L.J. Young, J. Polymer Sci ., 1961, 54, 411; ibid ., 1962, 62, 515
129.
K.F. O’Driscoll and T.
Yonezawa, Reviews of
Macromol . Chem ., 1966, 1 (1), 1
130.
131.
132.
133.
134.
B. Giese and K. Heuck,
Tetr . Lett ., 1980, 21, 1829
135.
136.
B. Giese and W. Zwick,
Angew . Chem .,Int . Ed ., 1978, 17, 6,130.;B. Giese and J. Meixner,
Chem . Ber ., 1977, 110, 2588
137.
M. Jacob, G. Smets, and F.
de Schryver. J. Polymer
Sci ., 1972, A-1,10, 669
138.
139.
140.
T. Alfrey, Jr. and C.C.
Price, J. Polymer
Sci ., 1947, 2, 101
141.
T.C. Schwann and C. C.
Price, J. Polymer
Sci ., 1959, 40, 457
142.
G.S. Levinson, J. Polymer Sci ., 1962, 60, 43
143.
K. Hayashi, T. Yonezawa, S.
Okamura, and K. Fukui, J.
Polymer Sci
., 1963, A-1,1, 1405
144.
G.G. Cameron and D. A.
Russell, J. Macromol
. Sci.,- Chem ., 1971, A5 (7), 1229
145.
A.D. Jenkins, J. Macromol. Sci.,- Pure and Applied Chem., 2000, A37, 1547. a.W.G. Barb, J. Polymer Sci.,1953, 11,117; b. Klumperman and I.R. Kraeger,
Macromolecules,1994, 27,1559; c. Coote, M. L.; Davis, T. P.;
Klumperman, B.; Monteiro, M. J.
Rev. Mocromol. Chem.
Phys.1998, 38, 567-593
146.
Harwood, H. J. Mokromol. Chem.,Mokromol. Symp. 1987, 10111, 331-354
147.
K. Liang, M. Doasit, D.
Moscatelti; and R.A. Hutchinson, Macromolecules, 2009, 42 (20), 7736; K.Liang and R.A.
Hutchinsone, Macromolecules, 2010, 43, 6311
148.
A.D. Jenkins, J. Polymer. Sci., Chem. Ed.,1996, 34, 3495
149.
G.E. Ham, General Aspects
of Free-Radical Polymerization, in Vinyl Polymerization, G.E.Ham,
ed., Dekker, New York, 1971
150.
T. Alfrey, Jr., and G.
Goldfinger, J. Polymer
Sci., 1944, 12, 322
151.
152.
S. Yamasaki, S.Tasaki, H.
Kudoh, and M. Matsunaga, Mutsuo, Journal of Polymer Science,Part A: Polymer Chemistry 2004, 42(7), 1707.
153.
154.
B.L. Funt and F.D.
Williams, J. Polymer
Sci ., 1960, 46, 139
155.
C. Bevington, N.A. Ghanem
and H.W. Melville, J. Chem
. Soc ., 1955, 2822
156.
J.W. Breitenbach, A.
Springer, and K. Hoseichy, Ber ., 1938, 71, 1438
157.
R.G. Caldwell and J.L.
Ihrig, J. Polymer
Sci ., 1960, 46, 407; J. Am . Chem . Soc ., 1962, 84, 2878
158.
159.
160.
S.C. Foord, J. Chem . Soc ., 1950, 68, 48
161.
162.
163.
164.
165.
166.
C.C. Price, J. Polymer Sci ., 1978, 3, 772; C.E. Walling, E.R.Briggs, K.B.
Wolfstern, and F.R. Mayo, J.
Am . Chem . Soc ., 1948, 70, 1537, 1544
167.
J. Lingman and G.
Meyerhoff, Macromolecules,
1984, 17, 941; Makromol . Chem ., 1984, 587
168.
H. Hirai, J. Polymer Sci .,Macromol . Rev ., 1976, 11, 47; I. Kunz, N.F. Chamberlain, and
F.J. Stelling, J. Polymer
Sci .,Polymer Chem . Ed ., 1978, 16, 1747
169.
G.D. Russell, R.G. Gilbert,
and N.H. Napper, Macromolecules, 1992, 25, 2459 ()
170.
N.G. Gaylord, Am . Chem . Soc . Polymer Preprints, 1969, 10 (1), 277
171.
N.G. Gaylord and A.
Takahashi, Chapt.6, Addition
and Condensation
Polymerization Processes, Advances in Chemistry Series #91, Am. Chem. Soc., Washington,
1969
172.
Y. Tsuda, J. Sakai, and Y.
Shinohara, IUPAC Int
. Symp . on
Macromol .
Chem ., Tokyo-Kyoto, 1966, Preprints, III, 44
173.
J.M.G. Cowie, “Alternating
Copolymerization”, in Comprehensive Polymer Science. Vol.4, G. Eastman, A. Ledwith,
S. Russo, and P. Sigwaldt, (eds.), Pergamon Press, Oxford,
1989
174.
J. Furukawa, “Alternating
Copolymers” in “Encyclopedia
of Polymer Science
and Engineering”, Vol.4,
2nd.ed., H.F. Mark,N.M. Bikales, C.G. Overberger, and G. Menges,
eds., Wiley-Interscience, New York, 1986
175.
H.G. Heine, H.J.
Rosenkranz, and H. Rudolph, Angew . Chem . ,Intern . Ed ., 1972, 11 (11), 974
176.
S. Iwatsuki, K. Nishio and
Y. Yamashito, Kagyo Kagaku
Zashi, 1967, 7 0, 384; (From Ref. 107)
177.
S. Iwatsuki and Y.
Yamashito, J. Polymer
Sci ., 1967, A-1,5, 1753
178.
179.
180.
N.L. Zutty, C.W. Wilson,
G.H. Potter, D.C. Priest, and C.J. Whitworth, J. Polymer Sci ., 1965, A-
1,3, 2781
181.
J.M.G. Cowie, “Alternating
Copolymerization”, in Comprehensive Polymer Science. Vol.4, G. Eastman, A. Ledwith,
S. Russo, and P. Sigwaldt, (eds.), Pergamon Press, Oxford,
1989
182.
J. Furukawa, “Alternating
Copolymers” in “Encyclopedia
of Polymer Science
and Engineering”, Vol.4,
2nd.ed., H.F. Mark, N.M. Bikales, C.G. Overberger, and G. Menges,
eds., Wiley-Interscience, New York, 1986
183.
I. Kuntz, N.F. Chamberlain,
and F.J. Steling. J.
Polymer Sci;,Pol.Chem.
Ed, 1978, 16, 1747
184.
H. Gilbert, F.F. Miller,
S.J. Averill, E.J. Carlson, V.L. Folt, H.J. Heller, F.J. Stewart,
R.T. Schmidt and H.L. Trumbull, J.
Am . Chem . Soc.,1956, 7 0, 1669CrossRef
185.
Y.Li and S.R. Turner,
Am. Chem. Soc. Polymer Preprints, 2009,50(2),4830
186.
187.
M. Imoto, T.Otsu, and M.
Nakabayashi, Makromol.
Chemie, 1963,
65, 194
188.
189.
R.M. Thomas and W.J.
Sparks, U.S. Patent # 2,373,067 (April 3, 1945)
190.
191.
E.M. Fetters and F.O.
Davis, Chapt. 15, Polyethers, Part III, Polyalkylene Sulfides and Other Polythioethers, N.G. Gaylord, (ed.),
Interscience, New York, 1962
192.
193.
Y. Nakayama, K. Hayashi,
and S. Okamura, J. Macromol
Sci .,-
Chem ., 1968, A2, 701
194.
S. Iwatsuki and Y.
Yamashita, J. Polymer
Sci ., 1967, A-1,5, 1753
195.
D.R. Bassett and A.E.
Hamielec, (eds.), Emulsion
Polymers and
Emulsion Polymerization, Am. Chem. Soc. Symp.
#165, Washingotn, 1981
196.
197.
C.H. Bamford, S. Brumby,
and R.P. Wayne, Nature,
1966, 209, 292
198.
199.
200.
M. Imoto, T. Otsu, and M.
Nakabayashi, Makromol
. Chem ., 1963, 65, 194 ()
201.
N.G. Gaylord and B.
Patnaik, J. Polymer
Sci., Polymer Letters, 1971, 8, 411
202.
N.G. Gaylord, J. Macromol . Sci .,Chem ., 1972, A6 (2), 259
203.
S. Tazuke and S. Okamura,
J. Polymer Sci ., 1967, B5, 95
204.
205.
V.P. Zubov, M.B. Lachinov,
V.B. Golubov, V.P. Kulikova, V.A. Kabanov, L.S. Polak, and V.A.
Kargin, IUAPC Intern
. Symp . Macromol . Chem ., Tokyo-Kyoto, Japan, 1966, 2, 56
206.
T. Ikegami and H. Hirai,
J. Polymer Sci ., 1970, A-1,8, 195
207.
208.
S. Yabumoto, K. Ishii, and
K. Arita, J. Polymer
Sci ., 1969, A-1,7, 1577
209.
V.A. Kabanov, J. Polymer Sci., Polymer Symposia, 1980, 67, 17
210.
211.
212.
A. Blumstein, R. Blumstein,
and T. H. Vanderspurt, J.
Colloid . Sci ., 1969, 31, 237
213.
A. Blumstein, S.L.
Malhotra, and A.C. Wattson, J.
Polymer Sci
., 1970, A-1,8, 1599
214.
A. Blumstein, K.K. Parikh,
and S.L. Malhotra, 23rd
Intern . Conf . of
Pure Appl
. Chem .,Macromol . Preprints, 1971, 1, 345
215.
M. Farina, U. Pedretti,
M.T. Gramegna, and G. Audisio, Macromolecules, 1970, 3 (5), 475 ()
216.
A. Colombo and G. Allegra,
Macromolecules,
1971, 4 (5), 579
217.
T. Uemura, Y.Ono,K.
Kitagawa, and s. Kitagawa, Macromolecules, 2007, ASAP Article 10.1021
218.
R. Buter, Y. Y. Tan and G.
Challa, J. Polymer
Sci ., 1972, A-1, 10, 1031 (); ibid ., 1973, A-1,11, 1013, 2975
219.
J. Gons, E. J. Vorenkamp
and G. Challa, J. Polymer
Sci., Polymer Chem . Ed ., 1975, 13, 1699, and 1977, 15, 3031
220.
221.
222.
S.G. Gaynor, D. Greszta, T.
Shigemoto, and K. Matyjaszewski, Am. Chem. Soc. Polymer Preprints,1994, 35,(2), 585
223.
224.
I. Li, B.A. Howell, R.
Koster, and D.B. Priddy, Am.
Chem. Soc. Polymer
Preprints,1994, 35 (2), 517
225.
226.
T. Ando, M. Kato,
M.Kamigaito, and M. Sawamoto, Macromolecules, 1996, 29,1070; M. Ouchi, S. Tokuoka, and M.
Sawamoto, Macromolecules,
2008, Web. Article
10.1021/ma70245d; M. Sawamoto,
Am. Chem. Soc. Polymer Preprints, 2010, 51(1), 5; Y. Fukuzaki, Y. Tomita, T.
Terashima, M. Ouchi and M. Sawamoto, Maromolecules, 2010, 43, 5989
227.
228.
J.Xia and K. Matyjaszewski,
Am. Chem. Soc. Polymer Preprints, 1996, 37 (2), 513
229.
230.
D. Greszta and K.
Matyjaszewski, J. Polymer
Sci., Chem. Ed.,1997, 35,1857
231.
J. Chiefari, Y.K. Chong, F.
Ercole, J. Krstina, J. Jeffery, T.P.T. Lee, R.T.A. Mayadunne, G.F.
Meijs, C.L. Moad, G. Moad, E. Rizzardo, and S.H. Thang,
Macromolecules,1998, 31, 5559CrossRef
232.
T. Otsu and M. Yoshida,
Makromol.
Chem.,Rapid Comun.
1982, 3, 133
233.
D.M. Haddleton, C.B.
Jasieczek, M.J. Hannon, and A.J. Shooter, Macromolecules, 1997, 30, 2190
234.
K.G. Suddaby, D.R. Maloney,
and D.M. Haddleton, Macromolecules, 30, 702 (1997); E.J. Kelly, D.M.
Haddleton, and E. Khoshdel, Am.
Chem. Soc. Polymer
Preprints,1998, 39 (2), 453
235.
A. Goto and T. Fukuda,
Macromolecules,
1997, 30, 4272, 5183
236.
T.C. Chung, H.L. Lu, and W.
Janvikul, Am. Chem.
Soc. Polymer Preprints, 1995, 36 (1), 241
237.
Li, Yunying; Lu, Zhi;
Wayland, Bradford B., Am.
Chem. Soc.,
Polymer Preprints
2003, 44(2), 780
238.
Qiu,J.Charleux, B,
Matyjaszewski, K. Prog.
Polymer Sci.,2001,26,2083
239.
240.
K. Matyjaszewski, B.E.
Woodworth, X. Zhang, S.G. Gaynor, and Z. Metzner, Macromolecules,31, 595 (1998)
241.
242.
243.
D. Mardare and K.
Matyjaszewski Macromolecules, 27, 645 (1994)
244.
R.P.N. Veregin, P.G. Odell,
L.M. Michalak. And M.K. Georges, Macromolecules, 29, 2746 (1996)
245.
Braunecker. W. A.;
Matyjaszewski, K. Prog
Polym Sd
2007 , 32, (1), 93-146
246.
Min, K.; Gao, H.;
Matyjaszewski, K. I Am.
Chem. Soc.
2005, 127, 3825—3830.:
247.
Jakubowski, W.;
Matyjaszewski, K. Macromolecules 2005, 38, 4139-4146.
248.
249.
K. Matyjaszewski and T.P.
Davis, ‘Handbook of Radical
Polymerization’, Wiley, New York, 2002 CrossRef
250.
251.
Matyjaszewski, K.; Dong,
H.; Jakubowski, W.;Pietrasik, J.; Kusumo, A. Langmuir 2007, 23, 4528—4531CrossRef
252.
253.
Percec, V; Guliashvili, T;
Ladislaw, J S.; Wistrand, A; Stjemdahl, A.;Sienkowska, M J.;
Monteiro, M J; Sahoo, S J.Am
Chem Soc
2006,128, (43), 14156-14165
254.
Jakubowski, W; Mm, K.;
Matyjaszewski, K. Macromolecules 2006, 39,39-45
255.
Matyjaszewski, K.
Macromolecules 2007, 40, (6), 1789-1791; K. Matyjaszewski,
et al, Science,
2011, DOI:10,1126
256.
P.M. Wright, G. Mantovani,
and D.M. Haddleton, Polymer
Preprints ,
2007,48, 132
257.
Honigfort, Mical E.;
Bnttain, William J. OH 44325 USA). Macromolecules 2003, 36(9), 3111; b: S. Faucher, P. Okrutny
and S. Zhu, Macromolecules,
2006, 39, 3
258.
259.
Klumperman, Bert; Chambard,
Gregory; Brinkhuis, ACS
Symposium Series,
2003, 854
260.
261.
J. Chiefari, Y.K. Chong, F.
Ercole, J. Krstina, J. Jeffery, T.P.T. Lee, R.T.A. Mayadunne, G.F.
Meijs, C.L. Moad, G. Moad, E. Rizzardo, and S.H. Thang,
Macromolecules,31, 5559 (1998)CrossRef
262.
263.
I. Li, B.A. Howell, R.
Koster, and D.B. Priddy, Am.
Chem. Soc. Polymer
Preprints,35 (2), 517 (1994)
264.
265.
P.G. Griffiths, Rizzardo,
E., Solomon, D.H., Tetrahedron
Lett. , 23, 1309
(1982)
266.
C.J. Hawker, G.G. Barclay,
R.B. Grubbs, and J.M.J. Frechet, Am. Chem. Soc. Polymer Preprints,37 (1) 513 (1996)
267.
D. Mardare and K.
Matyjaszewski Macromolecules, 27, 645 (1994)
268.
M.Fryd et al, U.S. Patent
#5 708 102 (Jan 13, 1998)
269.
270.
P.Nesvadba, A. Kramer, A.
Steinmann, and W. Stauffer, U.S. Patent 6,262,206 (2001)
271.
Wertmer, H.; Simon, D.;
Pfaender, R. PCT mt. Appl. WO 2005/021630 Al, 2005
272.
Grubbs, R. B.; Wegrzyn, J.
K.; Xia, Q. Chem. Commun.
2005, 80
273.
Xia, Q.; Grubbs, R. B.
J Polym. Sci., Part A: Polym. Chem. 2006, 44, 5128
274.
S.V. Arehart, D. Grezta,
and K. Matyjaszewski, Am.
Chem. Soc. Polymer
Preprints, 38 (1), 705 (19777)
275.
Kwak, Y.; Goto, A.; Tsujii,
Y.; Murata, Y.; Komatsu, K.; Fukuda, T. Macromolecules 2002, 35, 3026—3029CrossRef
276.
F. M. Calitz, J. B.
McLeary, J. M. McKenzie, M. P. Tonge, B. IUumperman, and R. D.
Sanderson, Macromolecules,
2003, 36, 9687; F. M. Calitz, M. P. Tonge,
and R. D. Sanderson, Macromolecular Symposia, 2003, 193
277.
(Bill) Y.K. Chong, J.
Kratina, T.P.T, Lee, G, Moad. A. Postma, E. Rizzardo, and S.H,
Thang, Macromolecules,
2003, 35, 2256
278.
Vana, Pbilipp; Davis,
Thomas P.; Barner—Kowollik, Christopher Macromolecular Theory and Simulations 2002, 11(8), 823—835
279.
Goto, A.; Zushi, H.; Kwak,
Y.; Fukuda, T. ACS Symp.
Ser. 2006, 944, 595-603; Goto, A.; Zushi, H.;
Hirai, N.; Walcada, T.; Kwak, Y.; Fukuda, T. Macromol. Symp. 2007,248, 126-131. Goto, A.; Zushi, H.;
Hirai, N.; Wakada, T.; Tsujii, Y.; Fukuda, T. J. Am. Chem. Soc. 2007, 129, 13347-13354.
280.
KoIb, H. C.; Finn, M. G.;
Sharpless, K. B. Angew.
Chem.,Int. Ed.
2001,40, 2004
281.
Lutz, J.-F. Angew. Chem.
tnt. Ed. 2007,46, 1018.;
Binder, W. H.; Sachsenhofer, R. Macromol. Rapid Commun.
2007, 28, 15CrossRef
282.
H. Gao and K.
Matyjaszewski, Am.Chem.
Soc. Polymer Preprints,2007, 48,130
283.
284.
E.B. Milovskaya,
Russ. Chem. Rev. ,42 (5), 1979, (1973)
285.
D.H. Solomon, E. Riz zardo,
P. Cacioli, Eur. Pat.
Appl,1985, EP
135280
286.
T.C. Chung, H.L. Lu, and W.
Janvikul, Am. Chem.
Soc. Polymer Preprints, 36 (1), 241 (1995)
287.
T.C.Chung, Polymer News 2003,28(8), 238—244
288.
Lacroix-D., Patrick S.,
Romain; O.B.; Boutevin, B. (A.C.S.
PolymerPreprints . 2003, 44(2),683
289.
290.
291.
K.A.Davis, H-j. Paik, and
K. Matyjaszewski, Macromolecules,32, 1767 (1999)
292.
W. R. Sorenson and T. W.
Campbell, Preparative
Methods of Polymer
Chemistry, 2nd Edit.,
Wiley-Interscience, New York 1968
293.
294.
E. H. Riddle, Monomeric Acrylic Esters, Reinhold, New York, 1961
295.
J. Coupek, M. Kolinsky, and
D. Lim, J. Polymer
Sci ., C4, 126 (1964)
296.
297.
A.Ravve, J.T. Khamis, and
L.X. Mallavarapu, J.
Polymer Sci
., A-1,3, 1775 (1965)
298.
M. Munzer and E.
Trammsdorff, “Polymerization in Suspension”, Chapt. 5 in
Polymerization Processes,
C.E. Schieldknecht, (with I. Skeist) (ed.), Wiley-Interscience, New
York, 1977
299.
300.
K.Zhang, Q.Fu, and L.
Jiang, Cailiao Yanjiu
Xuebao 2003, 17(1), 107—112; from Chem. Abstr. Synthetic Polymers, 2004, 140, 1000
301.
Farcet, C.;B; Pirri, R.;
Guerret, O., Am. Chem.
Soc. Polymer Preprints, 2002,43(2), 98—99.
302.
303.
D.C. Blakely, Emulsion Polymerization, Theory and Practice, Applied Science, London
1965
304.
D.R. Bassett and A.E.
Hamielec, (eds.), Emulsion
Polymers and
Emulsion Polymerization, Am. Chem. Soc. Symp.
#165, Washingotn, 1981
305.
306.
A. Ravve Organic Chemistry of Macromolecules, Dekker, New York,
1967
307.
Vanderhoff, J. W. In Vinyl
Polymerization; Ham, G., Ed.; Marcel Dekker: New York, 1969; Vol.
7, Part 2.Fitch, R. M.; Tsai, C. H. In Polymer Colloids; Fitch, R.
M., Ed.; Plenum: New York, 1971; p 73.
308.
Hansen, F. K.; Ugelstad, J.
In Emulsion Polymerization; Piirma, I., Ed.; Academic: New York,
1982
309.
Yeliseyeva, V. I. In
Emulsion Polymerization; Piirma, I., Ed.;Academic: New York,
1982
310.
Ottewill, R. O. In Emulsion
Polymerization; Piirma, I., Ed.; Academic: New York, 1982; p
1.
311.
Maxwell, I. A.; Morrison,
B. R.; Napper, D. H.; Gilbert, R. G. Macromolecules 1991, 24, 1629; J.P.A. Heuts, R.G. Gilbert,
and I.A. Maxwells, Macromolecules, 30, 726 (1997)
312.
313.
Kim,Y.,Berkel,G., Russel T,
and Gilbert RG., Macromolecules, 2003, 36, 3921
314.
R.D. Athley Jr.,
Emulsion Polymer
Technology’, Dekker, New York, 1991
315.
316.
B.M.E. Van der Hoff, Am.
Chem. Soc., Advances in Chemistry Series, #34, Washington,
1962
317.
P.J. Flory, Principles of Polymer Chemistry, Cornell Univ. Press.
N.Y. 1953
318.
319.
H. J. van den Hul and J. W.
Vanderhoff, Symposium Preprints, University of Manchester, Sept.
1969
320.
S.S. Medvedev, Intern. Symposium on Macromolecular Chemistry, Pergamon
Press, New York, 1959; S.S.
Medvedev, Ric . Sci .,Suppl ., 1955, 25, 897
321.
M.R. Grancio and D.J.
Williams, J. Polymer
Sci ., 1970, A-1,8, 2617; C.S. Chern and G.W.
Poehlein, J. Polymer
Sci., Polymer Chem . Ed ., 1987, 25, 617
322.
M. Chainey, J. Hearn, and
M.C. Wilkinson, J.
Polym. Sci.,Chem. Ed., 1987, 25, 505
323.
J. Delgado, M.S. El-Aaser,
C.A. Silebi, and J.W. Vanderhoff, J. Polym. Sci., Chem. Ed., 1989, 27, 193
324.
B. Liu, Am. Chem. Soc. Polymer Preprints, 2008, 49 (1), 521
325.
D.E. Brouwer, H.; Tsavalas,
J. G.; Schork, F. J.; Monteiro, M. J. Macromolecules 2000, 33, 9239— 9246; Tsavalas, J. G.;
Schork, F. J.; de Brouwer, H.; Monteiro, M. J. Macromolecules 2001, 34, 3938—3946; Butte, A.; Storti, G.;
Morbidelli, M. Macromolecules 2001, 34, 5885—5896; Lansalot
326.
M.; Davis, T. P.; Heuts, J.
P. A. Macromolecules
2002, 35, 7582—7591
327.
Y. Luo, X. Wang, B-G.Li,
and S. Zhu, Macromolecules,
2011, 44, 221
328.
329.
Binh T. T. Pham, Duc
Nguyen, Christopher J. Ferguson, Brian S. Hawkett, Algirdas K.
Serelis, and Christopher H. Such, Macromolecules, 2003, 36, 8907
330.
Sang Eun Shim, Yoda Shin,
Jong Won Jun, Kangseok Lee, Hyejun Jung, and Soonja Choe,
Macromolecules,
2003, 36, 7994
331.
B. Ray, Y. Isobe, K.
Matsumoto,S. Hibaue, Y. Okamoto, M. Kaigaito, and M. Sawamoto,
Mamolecules, 2004, 37, 1702CrossRef
332.