Miniemulsion Polymerization
Miniemulsion Polymerization
net/publication/225756343
Miniemulsion Polymerization
CITATIONS READS
381 3,753
6 authors, including:
Alessandro Butte
ETH Zurich
86 PUBLICATIONS 1,901 CITATIONS
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by Yingwu Luo on 20 May 2014.
Editorial Board:
A. Abe · A.-C. Albertsson · R. Duncan · K. Dušek · W. H. de Jeu
J. F. Joanny · H.-H. Kausch · S. Kobayashi · K.-S. Lee · L. Leibler
T. E. Long · I. Manners · M. Möller · O. Nuyken · E. M. Terentjev
B. Voit · G. Wegner
Advances in Polymer Science
Recently Published and Forthcoming Volumes
With contributions by
A. Butté · P. Y. Chow · J. M. DeSimone · K. Fontenot · L. M. Gan
K. Ito · S. Kawaguchi · K. A. Kennedy · Y. Luo · M. Nomura
G. W. Roberts · J. P. Russum · F. J. Schork · W. Smulders · K. Suzuki
H. Tobita
2 3
The series presents critical reviews of the present and future trends in polymer and biopolymer science
including chemistry, physical chemistry, physics and material science. It is addressed to all scientists at
universities and in industry who wish to keep abreast of advances in the topics covered.
As a rule, contributions are specially commissioned. The editors and publishers will, however, always be
pleased to receive suggestions and supplementary information. Papers are accepted for “Advances in Polymer
Science” in English.
In references Advances in Polymer Science is abbreviated Adv Polym Sci and is cited as a journal.
ISSN 0065-3195
ISBN 3-540-22923-X Springer Berlin Heidelberg New York
DOI 10.1007/b14102
This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, re-printing, re-use of illustrations, recitation, broadcasting,
reproduction on microfilms or in any other ways, and storage in data banks. Duplication of this publication
or parts thereof is only permitted under the provisions of the German Copyright Law of September 9, 1965,
in its current version, and permission for use must always be obtained from Springer-Verlag. Violations are
liable to prosecution under the German Copyright Law.
Editorial Board
Prof. Akihiro Abe Dutch Polymer Institute
Department of Industrial Chemistry Eindhoven University of Technology
Tokyo Institute of Polytechnics PO Box 513
1583 Iiyama, Atsugi-shi 243-02, Japan 5600 MB Eindhoven, The Netherlands
[email protected]
Prof. Jean-François Joanny
Prof. A.-C. Albertsson Physicochimie Curie
Department of Polymer Technology Institut Curie section recherche
The Royal Institute of Technology 26 rue d’Ulm
S-10044 Stockholm, Sweden F-75248 Paris cedex 05, France
[email protected] [email protected]
For all customers who have a standing order to Advances in Polymer Science,
we offer the electronic version via SpringerLink free of charge. Please contact
your librarian who can receive a password for free access to the full articles by
registering at:
springerlink.com
If you do not have a subscription, you can still view the tables of contents of the
volumes and the abstract of each article by going to the SpringerLink Home-
page, clicking on “Browse by Online Libraries”, then “Chemical Sciences”, and
finally choose Advances in Polymer Science.
Preface
Miniemulsion Polymerization
F. J. Schork · Y. Luo · W. Smulders · J. P. Russum · A. Butté · K. Fontenot . . 129
Dispersion Polymerization
S. Kawaguchi · K. Ito . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
Emulsion Polymerization:
Kinetic and Mechanistic Aspects
Mamoru Nomura ( ) · Hidetaka Tobita · Kiyoshi Suzuki
Department of Materials Science and Engineering, Fukui University, Fukui, Japan
[email protected], [email protected], [email protected]
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
2 M. Nomura et al.
Abstract The current understanding of the kinetics and mechanisms of batch and continu-
ous emulsion polymerizations is summarized from the viewpoints of particle formation and
growth and polymer structure development. There are numerous factors that affect these
processes; among them, studies on the radical transfer and monomer partitioning between
phases, which are key factors for particle formation and growth, are reviewed and discussed.
Attention is also focused on the effects of initiator type, additives and impurities in the recipe
ingredients, and agitation, each of which sometimes exert crucial influences on the processes
of particle formation and growth. In relation to polymer structure development, important
aspects of the molecular weight distribution and branched/crosslinked polymer formation
are highlighted.
Abbreviations
AA acrylic acid
AAm acrylamide
AN acrylonitrile
APS ammonium persulfate
AIBN 2,2¢-azobis-isobutyronitrile
BA butyl acrylate
Bu butadiene
CCTVFR continuous Couette-Taylor vortex flow reactor
CLTR continuous loop-tubular reactor
CMC critical micellar concentration
CSTR continuous stirred tank reactor
CTR continuous tubular reactor
DVB divinylbenzene
E ethylene
KPS potassium persulfate
MA methyl acrylate
MAA methacrylic acid
MC Monte Carlo
MMA methyl methacrylate
MWD molecular weight distribution
NaLS sodium laurylsulfate
n-BA n-butyl methacrylate
PSD particle size distribution
PSPC(R) pulsed sieve plate column (reactor)
PT(R) pulsed tubular (reactor)
S-E Smith and Ewart
SEC size exclusion chromatography
St styrene
VAc vinyl acetate
VCl vinyl chloride
Am total surface area of micelles per unit volume of water
Ap total surface area of polymer particles per unit volume of water
as surface area occupied by a unit amount of emulsifier
dm diameter of a micelle
dp diameter of a polymer particle
Dw diffusion coefficient for radicals in the aqueous phase
Emulsion Polymerization: Kinetic and Mechanistic Aspects 3
1
Introduction
2
Emulsion Polymerization Kinetics
2.1
Generally Accepted Kinetics Scheme
(a)
(b)
Fig. 1 (a) Three major mechanisms of particle formation, and (b) various chemical and
physical events that occur during the process of particle growth in an emulsion polymer-
ization
6 M. Nomura et al.
2.2
Summary of the Smith-Ewart Theory
The Smith and Ewart theory (the S-E theory) describes the basic concept of
emulsion polymerization. Its main points are briefly reviewed here. Smith and
Ewart showed that the rate of emulsion polymerization, which proceeds
exclusively in the polymer particles, is given by
R = k [M ] nN– (1)
p p p T
• • •
– = ∑ nN
n ∑ Nn = ∑ nNn NT (2)
n
n=1 0 n=1
where Çw is the rate of radical production per unit volume of water, ktw is the
rate coefficient for bimolecular radical termination in the aqueous phase, and
[R*w] is the radical concentration in the aqueous phase.
On the other hand, they derived an expression that predicts the number
of polymer particles produced, NT, assuming that (i) a monomer-swollen
emulsifier micelle is transformed into a polymer particle by capturing a free
radical from the aqueous phase, (ii) the volumetric growth rate per particle m
is constant, at least during particle formation, and (iii) free radical activity does
not transfer out of a growing particle
Emulsion Polymerization: Kinetic and Mechanistic Aspects 7
3
Kinetics and Mechanisms of Emulsion Polymerization
3.1
Radical Entry
One of the most important parameters in the S-E theory is the rate coefficient
for radical entry. When a water-soluble initiator such as potassium persulfate
(KPS) is used in emulsion polymerization, the initiating free radicals are gen-
erated entirely in the aqueous phase. Since the polymerization proceeds ex-
clusively inside the polymer particles, the free radical activity must be trans-
ferred from the aqueous phase into the interiors of the polymer particles, which
are the major loci of polymerization. Radical entry is defined as the transfer of
free radical activity from the aqueous phase into the interiors of the polymer
particles, whatever the mechanism is. It is believed that the radical entry event
consists of several chemical and physical steps. In order for an initiator-derived
radical to enter a particle, it must first become hydrophobic by the addition
of several monomer units in the aqueous phase. The hydrophobic oligomer
radical produced in this way arrives at the surface of a polymer particle by mol-
ecular diffusion. It can then diffuse (enter) into the polymer particle, or its
radical activity can be transferred into the polymer particle via a propagation
reaction at its penetrated active site with monomer in the particle surface layer,
while it stays adsorbed on the particle surface.A number of entry models have
been proposed: (1) the surfactant displacement model; (2) the collisional
model; (3) the diffusion-controlled model; (4) the colloidal entry model, and;
(5) the propagation-controlled model. The dependence of each entry model on
particle diameter is shown in Table 1 [12].
However, some of these models have been refuted, and two major entry
models are currently widely accepted. One is the diffusion-controlled model,
which assumes that the diffusion of radicals from the bulk phase to the surface
8 M. Nomura et al.
3.1.1
Diffusion-Controlled Entry
Smith and Ewart [4] first proposed that the transfer of free radical activity into
the interior of a polymer particle takes place by the direct entry of a free rad-
ical into a polymer particle. They pointed out that the rate of radical entry into
a polymer particle is given by the rate of diffusion of free radicals from an in-
finite medium of concentration [R*w] into a particle of diameter dp with zero
radical concentration,
Çe /NT = 2p Dw dp[R*w] = kep [R*w] (7)
where Dw is the diffusion coefficient for the radicals in the water phase and kep
the mass transfer coefficient for radical entry into a particle. However, for
simplicity, they actually used a rate coefficient that is proportional to the square
of the diameter (the surface area). Since then, most researchers have treated the
problem of particle formation by assuming that the rate of radical entry into
a micelle and a polymer particle is proportional to the surface area (the colli-
sional entry model) [8, 13].
On the other hand, Nomura and Harada [14] proposed a kinetic model for
the emulsion polymerization of styrene (St), where they used Eq. 7 to predict
the rate of radical entry into both polymer particles and monomer-swollen
micelles. In their kinetic model, the ratio of the mass-transfer coefficient for
radical entry into a polymer particle kep to that into a micelle kem, kep/kem, was
the only one unknown parameter (Eq. 37). They determined the value of kep/kem
to be about 103 by comparing the model’s predictions with experimental re-
sults. However, the observed value of kep/kem was at least two orders of magni-
tude greater than that predicted by Eq. 7, because kep/kem=dp/dm (dm is the
Emulsion Polymerization: Kinetic and Mechanistic Aspects 9
1 Dw –1
(9)
3 = 7 (X coth X – 1) + W¢
F l Dp
where X=(dp/2){(kp[M]p+nktp/vp)/Dp}1/2 and l is the equilibrium partition co-
efficient between particles and water for radicals, W¢ is the potential energy
Fig. 2 Capture efficiency F as a function of particle size rp for different values of the parti-
tion coefficient, l, and the number of radicals in a polymer particle for St polymerization;
(a) n=0 and (b) n=1
10 M. Nomura et al.
barrier analogous to Fuchs’ stability factor, Dp is the diffusion coefficient for the
radicals inside a particle, kp is the propagation rate constant, [M]p is the
monomer concentration in particles, n is the number of radicals in a particle,
NA is Avogadro’s number and vp is the particle volume. It should be noted that
radicals are captured inside the particles only if they react therein; otherwise
they will eventually diffuse out and back to the water phase. Figure 2 shows
an example of F versus particle radius rp, calculated from Eq. 9. The important
conclusion of Eq. 9 is, as is clear from Fig. 2, that the value of F for a particle con-
taining radicals is higher than that for a particle containing no radicals.
According to Fig. 2a, it approximately holds that Fµdp2, and hence this model
gives kep/kem=(dp /dm)(Fp/Fm)=(dp /dm)3 @103, the value of which is in good
agreement with the result obtained by Nomura et al. in the emulsion polymer-
ization of St [14].
A much simpler model for the radical capture (absorption) efficiency F can
be derived by introducing the concept of radical desorption from a polymer
particle, developed in Section 3.2.1. The probability F for a radical to be cap-
tured inside a particle containing n radicals by any chemical reaction (propa-
gation or termination) is given by
kp [M]p + ktp(n/vp)
F = 99993 (10)
Ko + kp [M]p + ktp(n/vp)
where Ko is the overall radical desorption rate constant for a particle, defined
by Eq. 19 and shown later in Section 3.2.1. For simplicity, no distinction is made
here between radicals with and without initiator fragments at their ends. In the
case where Kopkp[M]p, ktp(n/vp), substitution of Eq. 19 into Eq. 10 leads to
kp [M]p md yDw
F = kp [M]p /Ko = 99 1 + 82 dp2 µ dp2 (11)
12Dw mdDp
The result of Eq. 11 agrees with Fµdp2 obtained above by Ugelstad and Hansen
[15]. Therefore, both Eq. 9 developed by Hansen and Ugelstad and Eq. 11
developed here can explain the value of kep/kem@103 found experimentally
by Nomura et al. [14], although no direct experimental confirmation of the
validity of these radical capture models have been reported yet.
Unzueta et al. [18] derived a kinetic model for the emulsion copolymeriza-
tion of methyl methacrylate (MMA) and butyl acrylate (BA) employing both
the micellar and homogeneous nucleation mechanisms and introducing the
radical absorption efficiency factor for micelles, Fm, and that for particles, Fp.
They compared experimental results with model predictions, where they em-
ployed the values of Fp=10–4 and Fm=10–5, respectively, as adjustable param-
eters. However, they did not explain the reason why the value of Fm is an order
of magnitude smaller than the value of FP. Sayer et al. [19] proposed a kinetic
model for continuous vinyl acetate (VAc) emulsion polymerization in a pulsed
Emulsion Polymerization: Kinetic and Mechanistic Aspects 11
sieve plate column reactor, where they assumed that both micellar and homo-
geneous nucleation takes place, and introduced the radical absorption efficiency
factor Fm for micelles and Fp for polymer particles, respectively. They could
explain the experimental results by employing Fm=1.0¥10–5 and Fp=3.3¥10–3 in
the model predictions, indicating that kep/kem=330. This value agrees fairly well
with the value of 100 found for the St system [14], but is 30 times less than
kep/kem=104 found for the VAc system [20].Araújo et al. [21] developed a detail-
ed dynamic mathematical model that describes the evolution of particle size
distributions (PSDs) during the emulsion copolymerization of VAc and
Veova10 in a continuous loop-tubular reactor and compared results from it
with their experimental data. They could describe the process of micellar par-
ticle formation by introducing radical absorption efficiency factors for micelles
of Fm=1.5¥10–4 and for particles of Fp=1.5¥10–3, respectively, although they also
did not provide a reason why the value of Fm is 1/10 of the value of Fp. This gives
kep/kem=(dp/dm)(Fp/Fm)@102 if one assumes that dp/dm@10. On the other hand,
Herrera-Ordóñez et al. [22] also developed a mathematical model for St emul-
sion polymerization employing Eq. 8 as the radical capture rate coefficient,
where the expression for the capture of monomeric radicals is that used by
Hansen and Ugelstad [15, 17], while a more detailed modification was made for
the entry of initiator-derived radicals.
3.1.2
Propagation-Controlled Entry
Maxwell et al. [11] proposed a radical entry model for the initiator-derived
radicals on the basis of the following scheme and assumptions. The major
assumptions made in this model are as follows: An aqueous-phase free radical
will irreversibly enter a polymer particle only when it adds a critical number
z of monomer units. The entrance rate is so rapid that the z-mer radicals can
survive the termination reaction with any other free radicals in the aqueous
phase, and so the generation of z-mer radicals from (z–1)-mer radicals by the
propagation reaction is the rate-controlling step for radical entry. Therefore,
based on the generation rate of z-mer radicals from (z–1)-mer radicals by prop-
agation reaction in the aqueous phase, they considered that the radical entry
rate per polymer particle, Ç (Ç=Çe/NT) is given by
Ç = kpw[IM*z–1][M]w NT (12)
where kpw is the propagation rate constant in the aqueous phase and [M]w
is the monomer concentration in the aqueous phase. By substituting the steady-
state concentration of (z–1)-mer radicals [IM*z–1] into Eq. 12, the approximate
expressions for Ç and the initiator efficiency, fentry are derived, respectively, as
2kd[I] d9kd[I] k41
t,w
1–z 2kd[I]
Ç = 92 926 + 1 = 721 fentry (13)
Np kp,w[M]w Np
12 M. Nomura et al.
d9kd[I] k41
t,w
1–z
fentry = 926 + 1 (14)
kp,w[M]w
There has been discussion on the value of z. Maxwell et al. [11] proposed a
semi-empirical thermodynamic model to predict the value of z for persulfate-
derived oligomeric radicals, which is given by
z @ 1 + int(–23 kJmol–1/{RT ln[Msat]w}) (15)
where the integer function (int) rounds down the quantity in parentheses to the
nearest integer value and [Msat]w is the saturation solubility of the monomer in
mol dm–3. On the other hand, Sundberg et al. [23] proposed a thermodynamic
method for estimating the critical chain length z of entry radicals with a
hydrophilic end group (such as SO4–) using a simple two-layer lattice model. The
values of z calculated by both Sundberg et al. and by Maxwell et al. (Eq. 15) are
listed in Table 2.
Several research articles have been published that deal with the methodology
for determining the radical entry rate Ç, the initiator efficiency fentry and the
actual values of z. Hawkett et al. [24] developed a method for determining the
value of Ç along with the desorption rate coefficient kf, termed the slope-and-
intercept method. This method is experimentally simple, but has several draw-
backs [25]. For example, it is only applicable to the so-called zero-one system
– ≤0.5) with negligible radical termination in the aqueous phase. It is usually
(n
very difficult to judge whether or not the radical termination in the aqueous
phase is negligible. Moreover, it gives a large error if an induction period caused
by any trace of impurity exists. Marestin et al. [26] proposed an experimental
method for directly determining the entry rate of a critical size MMA oligomer
into the polymer particle using the seeded emulsion polymerization of MMA
Monomer Z value
2-EHA – 1
Styrene (St) 2–3 2
Butyl methacrylate (BMA) 3 2
Butyl acrylate (BA) 2–3 2
Butadiene (Bu) 3 2
Ethyl acrylate (EA) – 4
Methyl methacrylate (MMA) 4–5 4
Vinyl acetate (VAc) – 5
Methyl acrylate (MA) – 8
Acrylonitrile (AN) – >10 (estimate: 12)
Emulsion Polymerization: Kinetic and Mechanistic Aspects 13
initiated by KPS. The initial seed latex used was synthesized so as to have
radical traps (TEMPO) covalently bound onto the particle surface. When an
aqueous phase-propagating radical entered a seed particle, the nitroxide
moiety led to the formation of a stable alkoxyamine. Therefore, the kinetics of
radical entry into the seed particles was followed by monitoring the decay of
the ESR signal from the nitroxide in the samples withdrawn from the reactor.
They obtained fentry=0.36 for KPS at 70 °C and fentry=0.33 for V-50 at 70 °C,
respectively. Maxwell et al. [11] obtained the value of z@2 by comparing the
model predictions with the experimental results in the emulsion polymeriza-
tion of St. Schoonbrood et al. [27] reported z@18 for a 80:20/St:MA emulsion
copolymerization system. However, there is an example where it is difficult to
explain the kinetic behavior of the emulsion copolymerization of St and AAm
without assuming z=(one St monomer unit), as shown later in Section 3.2.3.
3.1.3
Miscellaneous Kinetic Problems in Radical Entry
firmed that for conventional emulsifiers such as sodium lauryl sulfate (NaLS,
electrostatic stabilizer), the extent of the surface coverage on a particle and the
ionic strength have no effect on radical entry, within experimental error mar-
gins [10, 30]. Adams et al. [30] found that the radical entry rate is independent
not only of the surface coverage and the ionic strength, but also of differences
in the kind of charges between the particle surface and the entering free radi-
cal. Furthermore, they refuted, on the basis of experimental data, the proposals
that the rate of radical entry is controlled by surfactant displacement [7] and
by collision of free radicals with the polymer particles [31]. Penbos et al. [10] car-
ried out the seeded emulsion polymerization of St using a PSt seed latex with
negative surface charge in combination with three different initiators: persul-
fate anions (negatively charged radicals), hydrogen peroxide/iron (II) (neutral
radicals) and 2,2¢-azobis(2-amidinopropane) (V-50, cationic free radicals).
They found that the rate of radical entry showed no significant dependence on
the kind of charges of these initiating radicals. El-Aasser et al. [25] examined the
effect of an adsorbed layer of the nonionic emulsifier (steric stabilizer) TritonX-
405 (octylphenoxypolyethoxy ethanol with an average number of 40 ethylene
oxide units) on the rate of radical entry by changing the ratio of anionic (NaLS)
and nonionic emulsifiers in the seeded emulsion polymerization of St initiated
by KPS. They concluded that any variation in surface coverage by the nonionic
emulsifier did not significantly affect the value of Ç within experimental error,
suggesting that the adsorbed layer of the nonionic emulsifier on the particle
surface would not act as a barrier to radical entry. Contrary to the observations
of El-Aasser et al. [25], Kuster et al. [32] observed a large decrease in the des-
orption rate coefficient for monomeric radicals in the emulsion polymerization
of St conducted using poly(ethylene oxide) nonylphenol surfactant with an
average number of 30 ethylene oxide units as the steric stabilizer. They con-
cluded from this observation that a “hairy” layer near the particle surface would
also act as a barrier for the entry of such uncharged radicals because re-entry
of desorbed radicals is the reverse process of desorption. Wang et al. [33] con-
ducted the seeded emulsion polymerization of St using the reactive surfactant
sodium dodecyl allyl sulfosuccinate (TREM LF-40) and its polymeric counter-
part, poy(TREM), and only observed an increase in the PSt molecular weight
when poly(TREM) was used. This was considered to be a consequence of the de-
crease in the rate of bimolecular radical termination resulting from the decrease
in the rate of radical entry into the polymer particles due to the diffusion bar-
rier of the hairy layer near the particle surface to the diffusing radicals.
On the other hand, several reports have been published that point out that
when a polymeric surfactant acting as an electrosteric stabilizer is used, the rate
of radical entry into a polymer particle should decrease due to a diffusion
barrier of the hairy layer built up by the polymeric surfactant adsorbed on the
surface of the polymer particles [34–36]. Coen et al. [34] found that in the
seeded emulsion polymerization of St using a PSt seed latex stabilized elec-
trosterically by a copolymer of acrylic acid (AA) and St, the electrosteric stabi-
lizer greatly reduced the radical entry rate Ç compared to the same seed latex
Emulsion Polymerization: Kinetic and Mechanistic Aspects 15
with a conventional electrostatic stabilizer. The explanation was that the “hairy”
layer on the surface of the polymer particle acted as a diffusion barrier to
the entering radicals. Kim et al. [35] also observed the decrease in the value of
Ç in the seeded emulsifier-free emulsion polymerization of MMA conducted
using KPS as the initiator and a PSt seed latex electrosterically coated by a
copolymer of St and styrenesulfonate (NaSS), which constitutes a thick hairly
layer on the particle surface.Vorwerg et al. [36] observed a decrease in the value
of Ç, although it was dependent on the pH, in the seeded emulsion polymer-
ization of St using seed latexes electrosterically stabilized with poly(acrylic
acid). Leeman et al. [37], on the other hand, synthesized amphiphilic water-
soluble polyelectrolytes, a poly(methyl methacrylate-block-sulfonated glycidyl
methacrylate) block copolymer (PMMA-b-SPGMA; polyanionic block copoly-
mer) and a poly(methyl methacrylate-block-quaternized N,N¢-dimethylamino-
ethyl methacrylate) block copolymer (PMMA-b-QPDMAEMA; polycationic
block copolymer). They carried out the emulsion polymerization of MMA at
60 °C in the presence of these two copolymers as surfactants and with the
following four types of initiators: KPS producing anionic radicals, H2O2 and
AIBN both producing neutral radicals and AAP (2,2¢-azobis(2-amidinopropane))
which generates cationic radicals in acid medium and neutral radicals in basic
medium. There was no major difference in the rate of polymerization Rp when
the initiators producing neutral radicals were combined with either of these
two oppositely charged copolymer surfactants. But a large decrease in Rp was
observed when the charge of the entering radical was different to the charge of
the block polyelectrolyte surfactant. Considering the experimental results
of El-Aasser et al. [25], the former finding is acceptable, but the latter one is
contrary to our expectations and the conclusions given in the articles [34–36].
Therefore, clarifying whether or not the adsorbed layer of polymeric surfac-
tants on the particle surface act as a diffusion barrier to the entering radicals
is still an important problem that needs to be solved in the future.
In the emulsion copolymerization of the water-soluble monomer acryl-
amide (AAm) and the sparingly water-soluble monomer St using KPS as ini-
tiator, Kawaguchi et al. [38] observed odd kinetic behavior. Regardless of the
fraction of AAm in the monomer feed, the polymerization of both monomers
started from the beginning of the reaction, but soon the AAm polymerization
slowed down and finally stopped, while the polymerization of St continued very
smoothly to the end. Nomura et al. [39] examined the reason for this abnormal
kinetic behavior and ascribed the reason to unusual radical entry behavior.
They studied the seeded emulsion copolymerization of St and AAm at 50 °C
using PSt particles as the seed and KPS as the initiator, and found that the change
in the number of PSt seed particles Ns caused a drastic change in the kinetic
behavior of this emulsion copolymerization system.When the number of seed
particles was less than a certain critical value Nc (~2.5¥1012 particles/cm3-
water), both St and AAm started polymerizing as soon as the initiator was added.
However, when the number of seed particles was higher than Nc, an apparent
induction period suddenly appeared for AAm polymerization; in other words,
16 M. Nomura et al.
although more than 99% of the initially charged AAm existed in the aqueous
phase, the AAm polymerization did not start until the St conversion exceeded
around 75%. Therefore, the apparent induction period was the time necessary
for the St conversion to reach 75%. On the other hand, the St polymerization
started and continued very smoothly from the beginning to the end of the
reaction, independently of the AAm polymerization.We have not yet succeeded
in explaining this interesting phenomenon quantitatively, but the reason may
be explained qualitatively as follows [40]. Using values for the monomer reac-
tivity ratio and the concentrations of St and AAm in the aqueous phase, it is
clear that the addition rate of St radicals to AAm is about 100 times faster than
to St and the addition rate of AAm radicals to AAm is about 5 times faster than
to St in the aqueous phase. Therefore, once an AAm radical is formed in the
aqueous phase, this AAm radical preferentially adds AAm monomer in the
aqueous phase without entering the polymer particles due to its hydrophilicity.
When the number of initially charged PSt seed particles is higher than Nc, no
AAm polymerization occurs in the aqueous phase although all free radicals are
produced in the aqueous phase and almost all AAm monomer units exist in the
aqueous phase. This implies that KPS radicals preferentially add St monomer
units as soon as they are generated in the aqueous phase and enter the seed
particles before the resultant St radicals add one AAm monomer unit. This is
because the mean residence time of the St radicals in the aqueous phase before
they enter the seed particles is shorter than the average time necessary for these
radicals to add one AAm monomer unit. On the other hand, when the number
of seed particles is less than Nc, the mean residence time of most of the St rad-
icals would become longer than the average time necessary for a St radical to
add one AAm monomer unit, and so these St radicals are mostly transformed
into AAm radicals, although some of these radicals can enter the seed particles
and continue St polymerization inside the particles. Therefore, the St radicals
produced continue AAm polymerization in the aqueous phase. The mechanism
of radical entry proposed in this copolymerization system contradicts the con-
clusions of Maxwell et al. [11], who claimed that KPS radicals need to add at
least 2–3 St monomer units before entering the monomer-swollen polymer
particles.
3.2
Radical Desorption
3.2.1
Desorption in Homopolymer Systems
The desorption (exit) of free radicals from polymer particles into the aqueous
phase is an important kinetic process in emulsion polymerization. Smith and
Ewart [4] included the desorption rate terms into the balance equation for Nn
particles, defining the rate of radical desorption from the polymer particles
containing n free radicals in Eq. 3 as kf nNn. However, they did not give any
Emulsion Polymerization: Kinetic and Mechanistic Aspects 17
detailed discussion on radical desorption. Ugelstad et al. [41] pointed out that
the rate coefficient for radical desorption (the desorption rate coefficient) could
be a function of particle size, the rate of chain transfer to monomer, the rate
of polymerization and the diffusion coefficients involved in the transport
processes leading to desorption of radicals, and suggested that, in the emulsion
polymerization of VCl, the desorption rate coefficient, kf might be expressed in
the form
kf = kvp–2/3 = k¢dp–2 (16)
On the other hand, Nomura and Harada [14, 42–45] pointed out that radical
desorption from the polymer particles and micelles plays a decisive role in par-
ticle formation and growth, and further that there are many examples of kinetic
deviations from the S-E theory that are attributable to radical desorption. First,
they theoretically derived the desorption rate coefficient from both stochastic
[42] and deterministic approaches, [14, 42, 43] based on a scheme consisting of
the following three consecutive steps: (1) chain-transfer of a polymeric radical
to a monomer molecule or a species like CTA in a polymer particle, followed
by (2) diffusional transportation of the resulting low molecular weight radical
to the particle-water interface, and (3) successive diffusion into the bulk water
phase through a stagnant film adjacent to the surface of the particle. In mod-
eling the rate coefficient for radical desorption, the following assumptions were
made:
A1. Polymer particles contain at most one radical (zero-one system).
A2. An oligomeric radical with no more than s monomer units can desorb
from and re-enter into a polymer particle with the same rate, irrespective
of its chain length.
A3. Instantaneous termination takes place when another radical enters a
particle already containing a radical.
A4. No distinction is made between radicals with or without an initiator
fragment on its end.
A5. Water-phase reactions such as propagation, termination, and chain-trans-
fer to a monomer are negligible for the desorbed radicals. This means that
all of the desorbed radicals would re-enter particles and the loss of these
radicals occurs only through the event given by the assumption A3).
A6. The physical and chemical properties of chain transfer agent (CTA) rad-
icals are approximately equal to those of monomer radicals.
Based on these assumptions, they derived the desorption rate coefficient kf as
Çw (1 – – Çw (1 – n– )ki
n) kmf kTf [T]p
kf = KoI 9948 + K + +
KoI n– + ki[M]p n– Np kp kp [M]p KoI n– + ki[M]p kp n– Np
o 6 661 99401
S j
kp [M]p
· ∑ 9941– (17)
j=1 Kon + kp[M]p
18 M. Nomura et al.
When it is assumed that initiator-derived radicals do not exit and only mono-
meric and CTA radicals produced by chain transfer to a monomer and/or a
CTA can desorb (s=1), Eq. 17 can be simplified as
kp [M]p kmf kTf [T]p
kf = Ko 986
– 6 + 661 (18)
Kon + kp[M]p kp kp [M]p
where kmf and kTf are the chain transfer rate constants to monomer and to CTA,
respectively, and Ko is the overall desorption rate constant per particle for
monomeric (or CTA) radicals, which is approximately given by [42–44]
2pDw dp yDw –1 12Dw d
Ko = 95 1 + 91 = 93 (19)
mdvp mdDp mddp2
= 9
d
–1
2pDw dp yDw 12Dp
Ko = 95 1 + 91 2
(19a)
mdvp mdDp p
kmf 12Dwd kmf 12DwCmd –2
kf = Ko 6 = 63111
2 6 = 6311131 d p (20)
kp mddp kp md
where Cm is the monomer chain transfer constant. The validity of Eq. 20 has
been demonstrated by Nomura et al. [20, 42, 44, 47] and Adams et al. [48].
Equation 18 was derived under the assumption that the physical and chemical
properties of a CTA radical are approximately equal to those of a monomeric
radical. However, if it is necessary to take into account the differences in the
physical and chemical properties between monomeric and CTA radicals, Eq. 18
can be modified approximately as
Emulsion Polymerization: Kinetic and Mechanistic Aspects 19
kmf [M]p kTf [T]p
kf = Kom 9862– + KoT 666152
– (18a)
Kom n + kp[M]p KoT n + kiT[M]p
where Kom and KoT are the overall desorption rate constants per particle for
monomeric and CTA radicals, respectively.
Ugelstad et al. [9] later derived almost the same desorption rate coefficient
as Eq. 19 given by
= 61212
m d k¢
–1
12Dw kmf Dw 12Dwd kmf
kf = 64 2 61 1 + 6121 2
(21)
mddp k¢p mdDp
6
d p p
where k¢p is the rate constant for the reaction between a monomeric radical
formed by chain transfer and a monomer, the value of which may be different
from the value of kp, the propagation rate constant.
Asua et al. [49–51] modified Eq. 18 in the absence of a CTA to include more
general cases, taking into account the fate of the desorbed radicals (both chem-
ical reactions in the water phase and re-entry) as
Ko
kf = kmf [M]p 6121941 (22)
Kob + kp[M]p
where b stands for the fraction of desorbed radicals that cannot re-enter
because of the aqueous phase termination or propagation, and is given by
kp[M]w + ktw[T*]w
b = 999492 (23)
kp[M]w + ktw[T*]w + kaNT
where [T*]w is the total radical concentration in the water phase and ka is the
mass-transfer coefficient for radical entry, and [M]w is the monomer concen-
tration in the water phase.
On the other hand, Casey and Morrison et al. [52, 96] derived the desorption
rate coefficient for several limiting cases in combination with their radical
entry model, which assumes that the aqueous phase propagation is the rate-
controlling step for entry of initiator-derived free radicals. Kim et al. [53] also
discussed the desorption and re-entry processes after Asua et al. [49] and
Maxwell et al. [11] and proposed some modifications. Fang et al. [54] discussed
the behavior of free-radical transfer between the aqueous and particle phases
(entry and desorption) in the seeded emulsion polymerization of St using KPS
as initiator.
3.2.2
Desorption in Copolymer Systems
As we discuss later in Section 3.3.3, Nomura et al. [45, 47] first derived the rate
coefficient for radical desorption in an emulsion copolymerization system by
20 M. Nomura et al.
In the case where all the desorbed A-monomeric radicals reenter the polymer
particles, the desorption rate coefficient for A-monomeric radicals kfA is given
by
CmAArA[MA]p + CmBA[MB]p
kfA = KoA 99998641 (26)
rB {(KoA –
nt/kpAA) + [MB]p} + [MA]p
where KoA is the overall desorption rate constant per particle for A-monomeric
radicals given by Eq. 19 in Section 3.2.1 and CmBA is the chain transfer constant
for a B-radical to A-monomer.
López et al. [55] investigated the kinetics of the seeded emulsion copoly-
merization of St and BA in experiments where the diameter and number of seed
particles, and the concentration of initiator were widely varied. The experi-
mental data were fitted with a mathematical model in which they used the
desorption rate coefficient developed by Forcada et al. [56] for a copolymer-
ization system. The desorption rate coefficient for the A-monomeric radical
that they used was a modification of Eq. 22 and Eq. 23, and is given by
K0A
kfA fA = (kmf,AA fA + kmf,BA fB)[MA]p 9999862 (27)
bAK0A + k¢pAA[MA]p + k¢pAB[MB]p
where kmf,AB denotes the chain transfer constant of the A-radical to the B-mono-
mer and bA is the fraction of the desorbed A-monomeric radicals that cannot
reenter the polymer particles because of the aqueous phase termination or
propagation. Barudio et al. [57], on the other hand, developed a simulation model
for emulsion copolymerization based on the pseudo-homopolymerization
approach, where they used the average rate coefficient for radical desorption
– – –
given by kde = (12Dw z/mddp2)(kmf /kp). Saldivar et al. have presented a survey of
Emulsion Polymerization: Kinetic and Mechanistic Aspects 21
3.2.3
Miscellaneous Kinetics Problems in Radical Desorption
The rate coefficient for radical desorption was derived by assuming that the
adsorbed layer of conventional or polymeric surfactant on the surface of the
polymer particle does not act as an interfacial diffusion barrier to the desorb-
ing neutral monomeric radicals. However, Kusters et al. [32] studied the kinet-
ics of particle growth in emulsion polymerization systems with a surface-active
initiator (an “inisurf ”). The inisurf employed was the diester of 4,4¢-azobis(4-
cyanopentanoic acid), the initiator moiety, with poly(ethylene oxide) nonyl-
phenol, the surfactant moiety. They observed a large decrease (one order of
magnitude) in the desorption rate coefficient for monomeric radicals in the
emulsion polymerization of St. They ascribed the reason for the decrease in the
rate of radical desorption to a “hairy” layer of the polymeric surfactant, which
would play the role of a diffusion barrier. Coen et al. [34] also reported that
in the seeded emulsion polymerization of St using a PSt seed latex stabilized
electrosterically by a copolymer of AA and St, the electrosteric stabilizer greatly
reduced the rate of radical desorption compared to the same seed latex with an
electrostatic stabilizer. They interpreted the reason for the decrease in the rate
of radical desorption by assuming that the aqueous-phase diffusion of mono-
meric radicals is slower in the hairy layer. Recently, Vorwerg et al. [36] carried
out a kinetic study of the seeded emulsion polymerization of St using PSt seed
lattices electrosterically stabilized with poly(acrylic acid) (pAA). They found
that seed lattices with a high-coverage of pAA (above 50 mC cm–2) exhibited a
significant reduction in radical desorption (by a factor of ~3 compared to the
ionically stabilized seed) at low pH.
22 M. Nomura et al.
3.3
Particle Formation and Growth
3.3.1
Particle Formation
As we mentioned in Section 2.1 (Fig. 1a), there are three major models for par-
ticle formation in emulsion polymerization. According to these models, poly-
mer particles are formed:
1. When a free radical in the aqueous phase enters a monomer-swollen emul-
sifier micelle and polymerization proceeds therein (micellar nucleation).
2. When the chain length of a free radical growing in the aqueous phase
exceeds its solubility limit and precipitates to form a particle nucleus
(homogeneous nucleation).
3. When a free radical growing in the aqueous phase enters a monomer droplet
and polymerization proceeds therein (droplet nucleation).
However, when the resultant polymer particles become unstable and coagulate,
then whatever the mechanism of particle formation is, the final number of
polymer particles produced is determined by a limited coagulation between
existing polymer particles (coagulative nucleation).
Smith and Ewart [4] derived an expression that can predict the number of
polymer particles produced, by assuming that:
1. A monomer-swollen emulsifier micelle is transformed into a polymer par-
ticle by capturing a free radical from the aqueous phase [4, 5].
2. The volumetric growth rate per particle m is constant, at least during parti-
cle formation (m=dnp/dt=constant).
3. Free radical activity does not transfer out of a growing particle (kf@0).
4. The amount of emulsifier that dissolves in the water phase without forming
micelles and adsorbs on the surface of emulsified monomer droplets may be
neglected.
Based on these assumptions, two limiting cases were discussed.
Case A: The rate of radical entry into micelles that results in the formation
of new particles is approximately equal to the rate of radical generation in the
water phase (Çw), as long as emulsifier micelles are present; in other words,
dNT
7 = Çw (28)
dt
Particle formation stops at the time tc, when the emulsifier micelles have just
disappeared because all of the emulsifier molecules comprising the emulsifier
micelles have been transferred to the surfaces of growing polymer particles
for adsorption. The volume vp,c at time tc of a particle formed at time t is
vp,c=m(tc–t), and so the surface area ap,c of this particle at time tc is given by
Emulsion Polymerization: Kinetic and Mechanistic Aspects 23
Am Çw
dNT/dt = Çw 77 = 08 (30)
Am + Ap 1 + Ap /Am
Then, it follows that
NT = 0.37(Çw /m)0.4 (asS0)0.6 (31)
On the other hand, Nomura et al. [14] proposed a different approach for pre-
dicting the number of polymer particles produced, where the new concept of
“radical capture efficiency” of a micelle relative to a polymer particle was pro-
posed. The assumptions employed were almost the same as those of Smith and
Ewart, except that the volumetric growth rate m of a polymer particle was not
considered to be constant. It was also assumed that all of the radicals formed
in the aqueous phase enter either micelles or polymer particles with negligible
termination in the aqueous phase. In this approach, the following elementary
reactions and their respective rates were defined.
(1) Particle formation by radical entry into a micelle
R* + ms Æ N*, kemms [R*w] (32)
(2) Formation of a dead particle by radical entry into an active particle con-
taining a radical
R* + N* Æ N0, kep N*[R*w] (33)
(3) Formation of an active particle by radical entry into a dead particle con-
taining no radical
R* + N0 Æ N*, kep N0[R*w] (34)
24 M. Nomura et al.
d[R*w]
91 = Çw – kemms [R*w] – kep NT [R*w] (35)
dt
dNT
71 = kemms [R*w] (36)
dt
dNT Çw Çw
71 = kemms [R*w] = 7788 = 77241 (37)
dt 1 + (kep NT /kemms ) 1 + (eNT /Sm)
where kepNT/kemms denotes the ratio of the rate of radical entry into polymer
particles to that into micelles and is rewritten as eNT/Sm, where e=(kep/kem)Mm
and e is the one unknown parameter, which affects the number of polymer par-
ticles produced. Here, Sm is the total number of emulsifier molecules forming
micelles, and Mm is the aggregation number of emulsifier molecules per micelle,
defined by Mm=Sm/ms. By solving a set of simultaneous differential equations
describing NT, N*, the monomer conversion XM, and using the balance equation
for the number of emulsifier micelles ms , the number of polymer particles pro-
duced NT can be predicted with respect to the initial emulsifier (S0) and initia-
tor concentrations (I0) (or Çw=2kd f [I0]) as shown by NTµÇ0.3 0.7
w S 0 . In the case
of VAc emulsion polymerization [20], the authors took into account radical
desorption from the polymer particles, yielding the following expression in
place of Eq. 37.
dNT Çw + kde –
nNT Çw + kde –
nNT
= k m
em s [R*
w ] = = (38)
dt 1 + (kepNT/kemms ) 1 + (e NT/Sm)
52 7788 772411
to introduce a value of e that is far greater than that predicted by using the
diffusion theory given by Eq. 7 (kep/kem=dp/dm). A value of e=1.28¥105 was
necessary for St emulsion polymerization [14], and a value of e=1.2¥107 for
VAc emulsion polymerization [20], while the value of e predicted by diffusion
theory (Eq. 7) was ~1000 in both systems because Mm≈100 and dp/dm≈10 hold
(roughly) in Interval I of particle formation, as already discussed in Sect. 3.1.1.
The authors, therefore, proposed the concept of the “radical capture efficiency”
of a micelle relative to a polymer particle to correct for this disagreement. The
same phenomenon has been encountered by several researchers [18, 19, 21]
(Sect. 3.1.1). However, the reason for the disagreement between the predicted
and experimental values of e has not been found yet.
As we discussed in Sect. 3.1.1, Hansen et al. [15] made significant improve-
ments to the concept of the radical capture efficiency proposed by Nomura
et al. [14]. Taking this concept into consideration, they examined the effect of
radical desorption on micellar particle formation in emulsion polymerization
[65]. Assuming that radical entry is proportional to the xth power of the micelle
radius and the polymer particle radius, they proposed the following general
expression for the rate of particle formation:
Çw + kf –
x
dNT – dcemsd m nNT
52 = (Çw + kf nNT) 5886 = (39)
dt x x
dcemsd m + NTd p
59941
NT dp
1 + 53 5
dcems dm
where dce is the radical capture efficiency of a micelle relative to a polymer par-
ticle, and is related to e by dce=(Mm/e)(dp/dm)x. The condition x=1 corresponds
to the diffusional entry model, while the condition x=2 corresponds to the col-
lisional entry model. Using x=1, they calculated the effect of radical desorption
in the emulsion polymerizations of St, MMA, VAC, and VCl on the number
of polymer particles produced, and demonstrated that the following general
rule for initiator and emulsifier exponents, which was first found by Nomura
et al. [20, 43], could also be applied to the emulsion polymerizations of these
monomers.
NT µ I01–zS 0z (40)
where 0.6<z<1.0. The value of z increases from 0.6 (a common value for St) to
1.0 (a common value for VAc) with increasing radical desorption.
Particle formation below the critical micellar concentration (CMC) in
emulsion polymerization is now accepted to take place according to the ho-
mogeneous nucleation mechanism. Among several quantitative treatments of
homogeneous particle formation in emulsion polymerization, the best-known
model was that proposed by Fitch and co-workers [66]. Their model is based
on the assumption that when the chain length of a free radical growing in the
aqueous phase reaches its solubility limit (critical chain length), it precipitates
to form a primary particle, and that particle formation will be hindered if these
growing oligomers are absorbed in polymer particles formed earlier. Hansen
26 M. Nomura et al.
et al. [67] made significant improvements on the Fitch model [the HUFT
(Hansen-Ugelstad-Fitch-Tsai) model]. According to Hansen et al, the rate of
particle formation is given by
dNT
52 = kpw Mw (RIjIcr + RMjMcr) (41)
dt
where kpw and Mw are the propagation rate constant and the monomer con-
centration in the aqueous phase, respectively, and RIjcr and RMjcr are the aque-
ous phase concentrations of oligomer radicals with critical chain length derived
from initiator and monomer radicals, respectively. This equation means that
oligomers stemming from initiator and monomer radicals precipitate as par-
ticles when they propagate beyond their respective critical degree of polymer-
ization, jIcr and jMcr. The authors derived the steady-state expressions for RIj and
RMj and obtained a general equation for homogeneous particle formation by
nserting them into Eq. 41. Furthermore, in order to simplify it, they neglected
particle formation from oligomers stemming from the desorbed monomeric
radicals, along with several other assumptions, and obtained
dNT – –j
(42)
52 = Çw (1 + ktwRw /kpw Mw + kc NT /kpw Mw) Icr
dt
–
where kc is the average rate coefficient for radical entry into polymer particles,
and Rw is the total radical concentration in the aqueous phase. Assuming that,
as an approximation, radical absorption by particles may be neglected in the
calculation of Rw, one gets
droplets are very small, they become an important source of particle formation
because the monomer droplets can compete for aqueous phase free radicals with
emulsifier micelles. This mode of heterogeneous polymerization is now called
“miniemulsion polymerization” and is reviewed in another chapter.
It is accepted that particle formation below the CMC in emulsion polymer-
ization takes place by homogeneous nucleation. However, there have been claims
that homogeneous nucleation is the main particle formation mechanism, even
above the CMC. Lichti et al. [70] investigated the mechanism of particle forma-
tion in the emulsion polymerization of St using sodium dodecyl sulfate (SDS)
as an emulsifier, and proposed the concept of “coagulative nucleation”. They
measured the full PSDs by electron microscopy at consecutive times soon after
the cessation of particle formation, and found that the PSDs obtained during
particle formation (Interval I) were positively skewed, confirming the role of co-
agulation, even above the CMC. Based on this phenomenon, they concluded that
the particle formation process does not occur by either simple micellar entry
or homogeneous nucleation mechanisms. Therefore, they suggested a mecha-
nism for particle formation where the homogeneous nucleation of oligomers
in the aqueous phase creates small primary polymer particles, and these
primary particles coagulate to produce polymer particles. On the basis of this
experimental finding, Feeny et al. [71] proposed a detailed theory for coagula-
tive nucleation and the PSDs in emulsion polymerization. The theory com-
bined and extended Müller-Smoluchowski coagulation kinetics with the DLVO
theory. Expressions were provided for the time evolutions of the nucleation
rate, particle number, and PSD. They showed that with physically reasonable
values for the parameters of the coagulation kinetics, agreement was obtained
with experimental data for the St emulsion polymerization system. Richards et
al. [72] developed a mathematical model for emulsion copolymerization. The
model combined the theory of coagulative nucleation of homogeneously nu-
cleated precursors with detailed species material and energy balances to calcu-
late the time evolution of the concentration, size and colloidal characteristics of
polymer particles, the monomer conversion, the copolymer composition, and
the molecular weight in an emulsion copolymerization system.
Although it is now accepted that particle formation below the CMC in emul-
sion polymerization takes place according to the homogeneous nucleation
mechanism, there has been debate as to whether homogeneous nucleation is
still operative even above the CMC, especially when relatively water-soluble
monomers are polymerized in emulsion in the presence of emulsifier micelles.
To date, most investigators believe that in the emulsion polymerization of
partially water-soluble monomers such as MMA and methyl acrylate (MA),
polymer particles are generated not by a micellar mechanism, but by homo-
geneous nucleation even in the presence of emulsifier micelles. This is because
the emulsion polymerization involving these monomers does not follow the
Smith-Ewart theory, and moreover, because an inflection point cannot be seen
around the CMC of the emulsifier on the particle number versus initial emul-
sifier concentration curve, where an abrupt and sharp decrease in the number
28 M. Nomura et al.
mechanism. The rationale behind the method was that if particle formation
took place by homogeneous nucleation, the resultant polymer particles would
contain negligible amounts of dye because the transport of the dye species
from the monomer droplet phase to the resultant polymer particles could be
neglected due to the insolubility of the dye in the aqueous phase. If, on the
other hand, particle formation took place by micellar nucleation, the resultant
polymer particles would contain an amount of dye corresponding to that sol-
ubilized in the micelles. They carried out the semibatch emulsion polymeriza-
tion of St and of MMA in the presence of the dye. In the semibatch emulsion
polymerization of MMA, for example, the experimental results showed that
when the emulsifier (SDS) concentration is above its CMC, mixed mode parti-
cle nucleation (micellar and homogeneous nucleation) was the predominant
mechanism. However, a question raised for this study is that if the transport of
the dye species from the monomer droplets to the resultant polymer particles
can be neglected, how is the dye transported from the monomer droplet phase,
where the dye is dissolved, to the monomer-swollen micelle phase where the
dye is solubilized.
Semibatch seeded emulsion polymerizations are quite common in industrial
operations. One of the most important problems in semibatch seeded emulsion
polymerization is how to control secondary particle formation. It is well known
that the amount of emulsifier must be carefully fed during starved-fed semi-
batch seeded emulsion polymerization. Too little emulsifier leads to emulsion
instability and hence coagulation, while too much emulsifier leads to secondary
particle formation by the micellar mechanism. Wang et al. [89] developed a
method for controlling the emulsifier level in starved-fed emulsion polymer-
ization. Morrison et al. [90] studied the conditions for secondary particle
formation in emulsion polymerization systems where the amount of added
emulsifier was below the CMC. They advanced their discussion based on the
HUFT model (Eq. 41), incorporating their reaction-controlled entry model, and
deduced a simple means for determining conditions for the onset and extent
of secondary particle formation. Coen et al. [91] further extended the work
of Morrison et al. [90] and proposed an extensive model for the PSD, particle
number, particle size and amount of secondary particle nucleation in emulsion
polymerization. Prescott et al. [92] proposed a simplified model for particle
formation, which is particularly useful for exploring the conditions required
for the growth of large particles, while avoiding secondary particle formation.
Butucea et al. [93] studied the seeded emulsion polymerization of VCl to es-
tablish the conditions needed to avoid the formation of new polymer particles
(secondary nucleation), and proposed new parameters: (1) MSA, the minimum
surface area of seed particles necessary to capture all initiator-derived (ionic)
radicals generated in the aqueous phase at a given initiator concentration; (2)
MCCI, the maximum critical concentration of initiator per unit surface of seed
particles under which the formation of new polymer particles is avoided; (3)
PVR1, the polymer volume per active growing radical necessary for the radical
to be within the particle for one second.
Emulsion Polymerization: Kinetic and Mechanistic Aspects 31
It has been reported that both the surface charge density and the degree of
surface coverage by emulsifier on the seed particles affect the behavior of sec-
ondary particle nucleation in seeded emulsion polymerization because these
factors control the rate of radical entry into seed particles. Vorwerg et al. [36]
carried out a kinetic study of seeded emulsion polymerization using PSt seed
particles electrosterically stabilized with poly(acrylic acid), and studied the
effect of the degree of surface coverage by poly(acrylic acid) on both radical
entry (Ç) and desorption (kf), through which secondary particle nucleation is
influenced under the condition of a fixed number of seed particles. The be-
havior of Ç and kf for the low-coverage particles was the same as that of the PSt
seed stabilized by initiator fragments and adsorbed emulsifier. The high-cov-
erage particles, on the other hand, exhibited strongly reduced kf values (by a
factor of three) at low pH, but Ç was only slightly lower than for the ionically
stabilized seed, while at high and neutral pH, secondary particle nucleation and
a decreased polymerization rate was observed with increasing pH (despite
an increase in particle number), indicating a reduced Ç value. Therefore, an
extensive electrosteric stabilizer reduced the rate coefficient for radical entry
(and for radical desorption), inducing secondary particle nucleation. Cheong
et al. [35, 94] investigated the effects of surface charge density on the kinetics
of secondary particle formation. They carried out three emulsifier-free seeded
emulsion polymerizations of MMA using monodispersed seed particles with
different surface charge densities, prepared from the St and NaSS comonomers
using the two-stage shot-growth process. In the case of the highest surface
charge density (72.7 mC/cm2), secondary particle nucleation was observed.
They ascribed the reason for this to the reduced rate of radical entry into the
seed particles resulting from electrical repulsion between seed particles and
entering oligmeric radicals [35]. They [94] proposed a mathematical model
to explain the effects of seed particle surface charge density on secondary
particle nucleation by introducing a modified Smolchowski equation and the
DLVO theory.
Sajjadi [95] examined the conditions for secondary particle formation and
coagulation in the seeded semibatch emulsion polymerization of BA under
monomer-starved conditions. They arrived at very interesting conclusions: (1)
particle coagulation occurred if the particle surface coverage (qcr1) dropped
below qcr1=0.25±0.05; (2) secondary particle formation occurred above a
critical surface coverage of qcr2=0.55±0.05, indicating that the presence of mi-
celles in the reaction vessel is not the only prerequisite for micellar nucleation
to occur; (3) the number of polymer particles remained approximately con-
stant if the critical surface coverage was within (qcr1=0.25)<q<(qcr2=0.55), and;
(4) this surface coverage band is equivalent to the surface tension band of
42.50±5.0 dyne/cm that is required to avoid particle formation and coagulation
in the course of polymerization. Sajjadi [96] also carried out an experimental
investigation on particle formation under monomer-starved conditions in the
semibatch emulsion polymerization of St. They observed that the number of
polymer particles formed increased with a decreasing monomer feed rate, and
32 M. Nomura et al.
that a much larger number of particles (1–2 orders of magnitude greater) than
that generally expected from a conventional batch emulsion polymerization
was obtained. It is clear from Eq. 29 that any variation in the formulation or
process variables that results in a reduction in the volumetric growth rate per
particle m will enhance particle formation as long as polymer particles are
generated from emulsifier micelles. A depressed rate of particle growth can
be achieved either by the starvation of polymer particles or by a reduction in
the capability of polymer particles to swell monomer, with the exception of
impurities (see Sect. 3.5) and radical desorption (see Sect. 3.2). In this case,
the former was the reason for an increase in particle formation. The latter
could be achieved, for example, if a small amount of crosslinking agent was
used in the formulation. It is well known that crosslinking will decrease the
extent of particle swelling by a monomer and thereby reduce the rate of parti-
cle growth [97]. Sajjadi [98] also investigated particle nucleation in the seeded
emulsion polymerization of St in the presence of Aerosol-MA emulsifier mi-
celles and in the absence of monomer droplets (Interval III). A larger number
of polymer particles were found to form in Interval III than in the corre-
sponding seeded batch operation in the presence of monomer droplets. The in-
crease in the number of polymer particles could be attributed to the reduced
rate of growth of new particles, which retarded the depletion of the emulsifier
micelles.
There are an enormous variety of commercial emulsifiers that are employed
in emulsion polymerization. Emulsifiers are generally categorized into four
major classes: anionic, cationic, nonionic and zwitterionic (amphoteric). The an-
ionic and nonionic emulsifiers are the most widely used. In addition, mixtures
of emulsifiers are also often used. Since the effects of the molecular structure
and chemical and physical properties of an emulsifier on particle formation are
still far from being well understood, numerous experimental investigations on
particle formation have been carried out to date with various nonionic emulsi-
fiers [99–102], mixed emulsifiers (ionic and nonionic emulsifiers) [18, 103–106]
and reactive surfactants [33, 107–110]. Recently, polymeric surfactants have
become widely used and studied in emulsion polymerizations [111–116]. A
general review of polymeric surfactants was published in 1992 by Piirma [117].
Recently, emulsion polymerization stabilized by nonionic and mixed (ionic and
nonionic) emulsifiers was reviewed by Capek [118].
Özdeğer et al. studied the role of the nonionic emulsifier Triton X-405 (octyl-
phenoxy polyethoxy ethanol) in the emulsion homopolymerization of St [99]
and n-butyl acrylate (n-BA) [100], and in the emulsion copolymerization of St
and n-BA [101]. In the emulsion homopolymerization of St, they noted two
separate nucleation periods, resulting in bimodal PSDs.Although the total con-
centration of the emulsifier was maintained at a level above its CMC based on
the water phase in the recipe, the portion of the emulsifier initially present in
the aqueous phase was below the CMC due to partitioning between the oil and
aqueous phases. Due to the nature of this emulsifier, the first of the two nucle-
ation periods was attributed to homogeneous nucleation, while the second was
Emulsion Polymerization: Kinetic and Mechanistic Aspects 33
attributed to micellar nucleation. In the case of n-BA, contrary to the case of St,
the emulsifier was found to partition primarily into the aqueous phase, leading
to nucleation in the presence of micelles and unimodal PSDs. Particle formation
was accompanied by limited aggregation in the early stages of the reaction with
particles being formed past 50% monomer conversion in some cases. In the
emulsion copolymerization of St and n-BA, unimodal PSDs were observed at
the lowest (4.2 mM) and the highest (12.5 mM, 16.2 mM) levels of the emulsi-
fier, while bimodal PSDs were produced at intermediate levels. These results
were also attributed to emulsifier partitioning between the oil and aqueous
phases. Lin et al. [102] also studied the kinetics of the emulsion polymerization
of St in the presence of the nonionic emulsifier NP40 (nonylphenol poly-
ethoxylate with an average of 40 oxyethelene units per molecule). The initially
charged emulsifier concentration was well above its CMC. The number of poly-
mer particles produced was proportional to the 2.4th power of the total emul-
sifier concentration. This deviation of particle formation kinetics from the
S-E theory (the 0.6th power) was attributed to the low water solubility of the
emulsifier (higher solubility in the monomer droplets), the increased agglom-
eration of polymer particles for the system with the lower amount of emulsi-
fier, and the increased contribution of miniemulsion polymerization kinetics
to the system with a higher emulsifier level.
Anionic emulsifiers provide electrostatic stability to the resultant polymer
particles. The efficiency of anionic emulsifiers is dependent on various para-
meters, such as ionic strength and pH, but this can be a major drawback in
terms of the stability of the resultant polymer particles. On the other hand,
nonionic polymeric emulsifiers provide the steric stabilization provided by the
thermodynamically-favored steric repulsion between particles. It is therefore
practical to use mixtures of anionic and nonionic emulsifiers in emulsion poly-
merization to take advantage of these two different stabilization mechanisms.
Chern et al. [104, 105] studied the CMC of the mixed emulsifier SDS/NP40 for
various compositions at 25 °C and at 80 °C by performing surface tension mea-
surements. They examined the effect of the mixed emulsifier SDS/NP40 on par-
ticle formation in the emulsion polymerization of St at 80 °C. They found that
adding only a small amount of SDS to NP40 dramatically increased the number
of polymer particles produced, and that the emulsion polymerization of St with
the mixed emulsifier SDS/NP40 did not follow the conventional S-E theory. The
number of polymer particles produced, NT, was described by the expression
NTµS0a, where S0 is the concentration of mixed emulsifier, and the values of a
were 0.60, 0.76, 1.3 and 1.1 for experiments with molar concentrations of NP40
of 0%, 40%, 70% and 100%, respectively. Colombié et al. [106] investigated the
role of mixed anionic-nonionic emulsifier systems in particle formation in the
emulsion polymerization of St. They carried out the emulsion polymerization
of St using a mixture of SDS and Triton X-405 at 70 °C. They found that adding
just 1 mM SDS to 6.4 mM Triton X-405 produced a dramatic increase in the rate
of polymerization and the number of polymer particles produced. The increase
in the number of polymer particles was 17 times that without SDS. When in-
34 M. Nomura et al.
creased amounts of SDS (3 and 5 mM) were used in combination with 6.4 mM
Triton X-405, the number of polymer particles produced increased. No sec-
ondary particle nucleation was noted upon the disappearance of the droplets,
and the resulting latexes were stable. They attributed this behavior to the change
in the partitioning of the Triton X-405 between the oil and aqueous phases by
changing the amount of SDS added. When no SDS was used, 95% of the Triton
X-405 was associated with the oil phase as opposed to 78% when 5 mM SDS was
used. Therefore, increasing the amounts of Triton X-405 in the water phase
by increasing the amount of SDS added allowed the formation of mixed emul-
sifier micelles, resulting in an increase in the number of polymer particles
produced and the rate of polymerization. Unzueta et al. [103] carried out the
semicontinuous emulsion copolymerization of MMA and n-BA using mixed
emulsifier systems (anionic sodium lauryl sulfate and nonionic polyethylene
oxide lauryl ether (Brij35)) and found that narrower PSDs with larger average
particle sizes were obtained with mixed emulsifier systems than those obtained
with single anionic systems. Furthermore, they developed a mathematical
model for the emulsion copolymerization of MMA and n-BA with mixed
anionic/nonionic emulsifier systems, where the CMC and micelle composition
of mixed emulsifiers was predicted using the thermodynamics of nonideal
mixtures [18].
Reactive surfactants have also been used in emulsion polymerization [33,
107–110, 114]. This is because the disadvantages of the surfactants that are
typically used in emulsion polymerization, such as instability of the latex and
surfactant migration during film formation, can be overcome in theory by
using a reactive surfactant. Studies of the use of reactive surfactants in het-
erophase polymerizations (up to 1997) have already been extensively reviewed
[107]. Amalvy et al. [108] investigated the particle formation process in the
emulsion polymerizations of St, MMA and VAc stabilized by sodium dodecyl
sulfopropyl maleate, a polymerizable surfactant (surfmer), focusing their at-
tention on whether the reactivity of the surfmer with the main monomer(s)
and the polymerization locus play critical roles in the particle formation
in emulsion polymerization systems. The results obtained suggested that the
presence of the surfmer did not affect the particle formation mechanism. They
concluded from the shape of the logNT vs logS0 (surfmer concentration) plot
that the polymer particles were formed by micellar nucleation in the case of St
and by homogeneous nucleation in the case of MMA and VAc. Wang et al. [33,
109, 110, 114] studied the emulsion polymerization of St using the reactive sur-
factant sodium dodecyl allyl sulfosuccinate (TREM LF-40) and its polymeric
counterpart (poly(TREM)) as anionic polymeric emulsifiers in terms of the
polymerization kinetics. The use of TREM LF-40 gave NTµS 00.5–0.6 and RpµN T0.7
at constant initiator concentration. The reasons for the unusual kinetics com-
pared to those with SDS (RpµN T1.0) were ascribed to chain transfer to TREM
LF-40, copolymerization of St with TREM LF-40, and the influence of the ho-
mopolymer TREM LF-40 (poly(TREM)) and/or the copolymer (poly(TREM-
co-St)) on the entry and the exit rates of the free radicals. In contrast, by
Emulsion Polymerization: Kinetic and Mechanistic Aspects 35
varying the initiator concentration, the kinetics were found to have the same
dependencies as the conventional emulsifier (RpµN T1.0µI 00.4). In the case of poly-
(TREM), the dependencies of Rp and NT on S0 and I0 varied depending on
experimental conditions (RpµN T1.0µS 00.2–0.4, and RpµN T1.0µI 0.6–0.8
0 ). It was inferred
that homogeneous nucleation was dominant when using poly(TREM), even at
concentrations exceeding its CMC. This was different from the monomeric
TREM LF-40 emulsifier.
Recently, polymeric surfactants have received considerable attention in
industry. They provide the steric repulsion between interacting particles,
which gives the latex excellent stability against high electrolyte concentration,
freeze-thaw cycling and high shear rates. Cochin et al. [111] carried out a
comparative study of the emulsion polymerization of St using conventional,
polymerizable and polymeric emulsifiers. Ayoub et al. [112] investigated the
emulsion polymerization of St with amphiphatic copolymers [of VAc and
methoxy polyoxyethlene (PVAc-b-MPOE)(35:65, 27:73, 19:81 wt/wt) prepared
with a macroradical initiator in the presence of benzoyl peroxide] as the emul-
sifier. The experimental results for the number of polymer particles produced
(NT) versus emulsifier concentration (S0) were as follows: NTµS01.82 (65%),
NTµS 02.1 (73%) NTµS01.66 (81%). They [113] also studied the emulsion polymer-
izations of VAc and St using the polymeric emulsifier prepared from poly-
oxyethylene methylether (POE, 66%) and St (34%). They did not measure the
number of polymer particles produced, but the rate of polymerization was
found to be proportional to the 0.9th and 0.76th powers of the initiator (KPS)
concentrations, and to the 0.77th and 0.66th powers of the emulsifier concen-
trations for VAc and St monomers, respectively. Kato et al. [115] investigated the
emulsion polymerization of St using poly(methyl methacrylate (MMA)-co-
methacrylic acid (MAA)) with different copolymer compositions as polymeric
emulsifiers. They examined the effect of the copolymer compositions, molec-
ular weights and MAA contents of the polymeric emulsifiers on the number of
polymer particles produced. They found that the number of polymer particles
produced showed a slight dependence on the copolymer molecular weight,
having a maximum when the molecular weight was in the range 5,000–10,000,
and decreasing monotonously with the content of MMA in the copolymer.
Cheong et al. [116] studied the kinetics of particle nucleation and growth in the
emulsion polymerization of St using water-soluble polyurethane resins (PUR)
as the emulsifier. They found that the number of polymer particles produced
became constant in the early stage of polymerization when the concentration
of the initially charged PUR was lower. However, the monomer conversion
where the particle number became constant increased with increasing the
initial PUR concentration. The constant particle number observed (NT) was
correlated as NTµ[PUR]00.6–0.7[KPS]00.4 (where [PUR]0 and [KPS]0 are the PUR
and KPS concentrations, respectively). These dependencies are almost the same
as those predicted by the S-E theory.
36 M. Nomura et al.
3.3.2
Particle Growth in Homopolymer Systems
As is clear from Eq. 1, the rate of particle growth (Rp/NT) is proportional to the
monomer concentration, [M]p and the average number of radicals per particle,
–
n, respectively. Thus, –n is one of the basic parameters that characterize the
kinetic behavior of particle growth in an emulsion polymerization system.
Early researchers devoted their efforts to deriving a quantitative description of
–
n by solving Eq. 3 for –n defined by Eq. 2 [4, 119, 120].
Smith and Ewart [4] did not obtain a general solution to Eq. 3, but rather
solved it for three limiting cases at steady-state conditions, that is, dNn /dt=0.
where ktw is the termination rate constant in the water phase, Çw is the rate of
radical generation per unit volume of water, [Rw*] is the concentration of radi-
cals in the water phase, and md is the partition coefficient of radicals between
the water and the polymer particle phases. However,
b. When termination in the polymer particles is dominant,
–
n = (Ç /2k N )1/2 O 0.5 (45)
w f T
Moreover, when both radical termination in the water phase and radical de-
sorption from the particles are negligible, Eq. 47 is reduced to
–
n = (Ç v /2k N )1/2 p 0.5 (48)
w p tp T
In this case, the kinetic behavior is quite similar to that of suspension poly-
merization, except that the polymer particles are supplied with free radicals
from the external water phase. When the polymerization proceeds according
to Eq. 48, the system is sometimes referred to as obeying “pseudo-bulk” kinet-
ics.
A general solution to Eq. 3 was provided by Stockmayer [121] with minor
corrections by O’Toole [119]. On the other hand, Ugelstad et al. [120] proposed
the most useful and widely applicable expression for – n given by
Im (a) 2a
n– = (a/4) 921 = (1/2) 99966
Im –1(a) 2a
m + 9977 (49)
2a
m + 1 + 9171
m+2+…
where Im(a) is the modified Bessel function of the first kind, m=kfvp/ktp , and
a=a2/8=Çevp/ktpNT. On the other hand, the radical balance in the water phase
(Eq. 4) leads to the following relationship using the non-dimensional parame-
ters, a, aw, m and Y.
a = a + mn – – Ya2 (50)
w
2 1/2 1/2
– 1 aw aw aw 1 aw 1
n=3 aw + 5 + 2 aw + 5 – aw + 5 + 3 + 5 – 3 (51)
2 m m m 4 2 2
The values predicted by Eq. 51 agree well with those predicted by Eq. 49 within
less than 4%. This type of plot is called a “Ugelstad plot” and has been applied
as a criterion to determine whether a system under consideration obeys either
zero-one kinetics (n – ≤0.5) or pseudo-bulk kinetics (n
– >0.5).
Nomura et al. [42, 43, 64] showed that when the value of the term ktp/vp is
very large (the rate of bimolecular termination in the polymer particles is very
rapid), –n is expressed by
– C 2 + lll
n = (– C + dllll 2C)/2 (52)
38 M. Nomura et al.
sented by Nomura et al. [123]. Although this approach is rather laborious and
time-consuming, they could successfully determine the propagation rate con-
stant kp and the value of m, from which the desorption rate coefficient kf could
then be deduced, in seeded and unseeded emulsion polymerization of VDC at
50 °C. On the other hand, Hawkett et al. [24] developed a method for deter-
mining Ç and kf termed the slope-and-intercept method. This method is sim-
ple and straightforward, but has several drawbacks (as stated in Sect. 3.1.2), so
care must be taken when this method is used. Asua et al. [12, 124] proposed a
new approach for the estimation of kinetic parameters such as the entry and
desorption rate coefficients, the termination rate constant in the aqueous
phase, the rate coefficient for initiator decomposition and the propagation rate
constant in emulsion homopolymerization systems under zero-one conditions.
40 M. Nomura et al.
They claim that accurate values for the parameters are obtainable with this
approach provided that a sufficient number of experiments with a minimum
range of variation are available.
Recently, several modeling papers have been published which are useful for
the design and operation of emulsion homopolymerization processes [22, 80, 81,
125–127]. Mendoza et al. [125] developed a mathematical model that could pre-
dict the monomer conversion, particle diameter, number of polymer particles
produced, and the number-average and weight-average molecular weights in the
unseeded emulsion polymerization of St using n-dodecyl mercaptan as CTA.
This model was validated by fitting the experimental data to the model’s pre-
dictions. Kiparissides et al. [126] proposed a comprehensive mathematical
model to quantify the effect of the oxygen concentration on the polymerization
rate and PSD in the unseeded emulsion polymerization of VCl. Particle forma-
tion was assumed to proceed by both the homogeneous and micellar nucleation
mechanisms.Asua et al. [127] developed a mathematical model for seeded emul-
sion polymerization stabilized with polymerizable surfactants (surfmers). The
model included the most distinctive features of surfmer polymerization, in-
cluding partitioning of the unreacted surfmer between the surface of the poly-
mer particles and the aqueous phase, and the surfmer burying itself inside the
polymer particles. The model also included the possibility of having radical
concentration profiles in the polymer particles. Herrera-Ordóñez et al. [22, 80,
81] proposed a detailed mathematical model of the kinetics of St emulsion
polymerization, which was a modification and adaptation of previous works
reported in the literature. By comparing model predictions with experimental
results, they arrived at the conclusion that initiator-derived radicals with only
one monomeric unit also make a significant contribution to the rate of radical
capture by polymer particles, which contradicts the conclusion obtained by the
Sydney Group [11, 91]. They applied the model to the emulsion polymerization
of MMA above the CMC of the emulsifier to discuss the mechanism of particle
formation and growth in this system, and concluded that particle formation by
micellar nucleation is at all times at least ten orders of magnitude greater than
the homogeneous one, although MMA is moderately water-soluble [81].
Although the emulsion polymerization of VAc is already one of the most
studied systems, research articles on this topic are still being published [128–
132]. Gilmore et al. [128, 129] presented a mathematical model for particle
formation and growth in the isothermal semibatch emulsion polymerization
of VAc stabilized with poly(vinyl alcohol) (PVA). The model accommodated
grafting onto the PVA backbone during particle formation, and polymeric sta-
bilization.When the emulsion recipe, process conditions and kinetic parameters
are supplied, the model can predict the various species concentrations along
with the monomer conversion and particle size and number profiles. In Part II
of a series of papers [129], model predictions were compared with semibatch
and batch experimental results. Budhiall et al. [130] investigated the role of
grafting in particle formation and growth during the emulsion polymerization
of VAc with partially hydrolyzed PVA as the emulsifier and KPS as the initiator.
Emulsion Polymerization: Kinetic and Mechanistic Aspects 41
They found that: (1) the number of polymer particles produced was dependent
on the PVA blockiness; (2) the PVA with the higher degree of blockiness led
to the formation and stabilization of more polymer particles; (3) particle for-
mation continued to high conversions for the more random PVA (Poval 217),
whereas it appeared to stop at intermediate conversions for the blockier PVA
(PVA217EE); (4) all systems exhibited limited aggregation of the polymer par-
ticles during the polymerization process, and; (5) the greatest amount of graft-
ing of the PVA stabilizer onto the polymer particles occurred early in the
reactions (XM<25%), presumably contributing primarily to the stabilization of
the particles. Shaffie et al. [131] studied the kinetics of the emulsion polymer-
ization of VAc initiated by redox initiation systems of different persulfate cations
such as KPS, sodium persulfate (NaPS), and ammonium persulfate (APS); each
of them was coupled with a developed acetone sodium bisulfate adduct as the
reducing agent. In emulsion polymerization, the exhaustion of the separate
monomer droplet phase (the onset of Interval III) is usually followed by a
decrease in the rate of polymerization due to the decrease in monomer concen-
tration in the polymer particles. This is not the case for the emulsion polymer-
ization of VAc, where the rate of polymerization remains constant throughout
most of Interval III (ca. 20~90% conversion). In the case of VAc, Interval III starts
from around 20% conversion [20, 64]. In order to explain the reason for the
independence of the polymerization rate from the monomer concentration,
Nomura et al. [20] developed a model that takes into account the particles con-
taining at most two free radicals, where no instantaneous termination inside
the particles is assumed. Based on this model, they ascribed the reason to
an increase in the value of – n due to the gel effect, which compensates for the
decrease in monomer concentration with conversion. The decrease in the
termination rate constant inside the polymer particles due to an increase in
viscosity with conversion (the gel effect) prolongs the coexisting time of two
free radicals in the same particle, thereby increasing the value of –n. Chern et al.
[133] also developed a model that includes particles containing at most two free
radicals. However, they attributed the reason to a decrease in the rate coefficient
for radical desorption due to an increase in viscosity with conversion, which
results in an increase in the value of –
n. On the other hand, Bruyn et al. [132] pro-
posed a kinetic model that considers a zero-one system with instantaneous rad-
ical termination inside the particles. The model assumes that radical loss is by
transfer to a monomeric species which is very slow to propagate and whose
radical activity is lost by desorption and termination, either in the aqueous
phase or when it enters a particle containing a radical. Since the transfer step
is rate determining, the rate of this process is proportional to monomer con-
centration, which then cancels the dependence on the monomer concentration
in the overall polymerization rate expression. This model also predicts that the
radical loss rate coefficient should be either ktrCp (Limit 2a [1]) or 2ktrCp
(Limit 2b [1]), where ktr is the rate coefficient for transfer to monomer and Cp
is the monomer concentration in the particles. They [134] also studied the
kinetics and mechanisms of the emulsion polymerization of vinyl neo-decanete
42 M. Nomura et al.
3.3.3
Particle Growth in Copolymer Systems
Ballard et al. [136] presented an extended S-E theory that provides a description
of the emulsion copolymerization system during Interval II and III and sug-
gested the possibility of using an “average” rate coefficient to treat the copoly-
merization system. On the other hand, Nomura et al. [45, 47, 137] first developed
an approach to generalize the S-E theory for emulsion homopolymerization to
emulsion copolymerization by introducing “average (or mean) rate coefficients”
for propagation, termination and radical desorption. This methodology was
termed the “pseudo-homopolymerization approach” [138] or the “pseudo-kinetic
rate constant method” [139] and is now widely applied, not only to emulsion
copolymerization systems, but also to other homogeneous free radical copoly-
merization systems. Nomura et al. [47, 122a, 140] demonstrated that the equa-
tions derived so far for emulsion homopolymerization can also be applied with-
out any modification to a binary emulsion copolymerization system with
monomers A and B by substituting the following “average rate coefficients” for
the corresponding rate constants for emulsion homopolymerization.
The polymerization rate for the A-monomer is expressed as
–
r = k [M ] –
pA pA nN
A p t T (53)
where [MA]p is the concentration of A-monomer in the polymer particles and
– – =–
nt is the average number of total radicals per particle (n nA + –
nB). The over-
t
all rate of copolymerization is defined by
–– –
R = k [M] –
p p n N = r + r = (k [M ] + k [M ] )n– N
p t T pA pB pA A p pB B p t T (54)
Emulsion Polymerization: Kinetic and Mechanistic Aspects 43
–
where kp is the overall propagation rate coefficient defined by Eq. 54 and is a
function of the propagation rate constant, monomer reactivity ratio and mole
fraction of each monomer in the polymer particles. In the case of a binary
emulsion copolymerization system, for example, the average rate coefficients
are defined as follows:
–
1. The average rate coefficient for the propagation of the A-monomer, kpA, is
given by
–
kpA = kpAA fA + (kpBB/rB)fB (55)
Here
–
nA kpBBrA[MA]p
fA = 5
– = 999953 (25)
nt kpAArB[MB]p + kpBBrA[MA]p
where kpBA denotes the rate constant for the propagation of a B-radical to an A-
monomer, fA is the fraction of A-radicals in the particle phase (fA+fB=1), and
rB is the B-monomer reactivity ratio.
–
2. The average rate coefficient for radical desorption, kf , is defined using the
equations
– /n
kf = kfA(n – – –
A t) + kfB(nB/nt) = kfA fA + kfB fB (24)
and
CmAArA[MA]p + CmBA[MA]p
kfA = KoA 99998641 (26)
rB {(KoA –
nt/kpAA) + [MB]p} + [MA]
where kfA is the desorption rate coefficient for the A-monomeric radicals, CmBA
is the chain transfer constant of a B-radical to an A-monomer, and KoA is the
desorption rate constant for A-monomeric radicals, given by Eq. 19 in
Sect. 3.2.1.
–
3. The average rate coefficient for radical termination in the particle phase, ktp,
is defined by
–
ktp = ktpAA f A2 + 2ktpAB fA f B + ktpBB f B2 (56)
where ktpAB is the bimolecular radical termination rate constant between A- and
B-radicals. Other average coefficients, such as the average chain transfer coef-
ficient, can be defined, if necessary, using the same principle. This approach can
be easily extended to multimonomer emulsion polymerization systems [138,
141]. Based on an exact mathematical treatment, Giannetti [142] concluded that
the pseudo-homopolymerization approach represents a suitable approximation
for most copolymerization systems of practical interest, except for very special
cases.
44 M. Nomura et al.
tinous reactors and the corrsponding model predictions. Vega et al. [61, 147]
and Dubé et al. [144] both developed mathematical models for the emulsion
copolymerization of AN and Bu initiated by a redox initiator system, with the
aim of simulating an industrial process and of improving the final polymer
quality. Due to the high solubility of AN in water, the following effects were
included: (a) the homopolymerization of AN in the aqueous phase; (b) the
desorption of AN radicals from the polymer particles, and; (c) homogeneous
particle nucleation. Saldívar et al. [58–60, 143] carried out extensive investiga-
tions on emulsion copolymerization. In the first paper, they presented a de-
tailed mathematical modeling of emulsion copolymerization reactors along
with comprehensive reviews of earlier models [58]. Then, they validated their
model with experimental results obtained in the emulsion copolymerization of
St and MMA, and demonstrated the generality of the model by applying it to
three illustrative problems [143]. Furthermore, they performed a systematic
experimental study of ten binary and three ternary emulsion copolymerization
systems involving St, MMA, BA, Bu,VAc,AA and E [59]. The predictions for the
evolution of conversion and average particle diameter in batch emulsion
copolymerizations from the model were compared with experimental data for
four emulsion copolymerizations of St with the comonomers MMA, BA, Bu and
AA.After data fitting for unknown or uncertain parameters, the model was ca-
pable of quantitatively explaining the experimental observations for conversion
evolution, but could only qualitatively explain the particle size evolution [59].
In industrial emulsion polymerization processes, a small amount of water-
soluble carboxylic monomer (such as AA) is often added to improve the col-
loidal stability and surface properties of the resulting latex particles. Therefore,
numerous studies have been carried out to date to clarify the influence of the AA
monomer on the kinetic behavior of the emulsion copolymerization of St and
AA [148–152] and of emulsion terpolymerizations including AA [79, 153, 154].
Shoaf et al. [148] presented a kinetic model that describes the reaction behavior
of emulsion copolymerization systems where significant polymerization occurs
in both the particle and aqueous phases. The model was applied to two seeded
carboxylated emulsion copolymerization systems, AA-St and methacrylic acid
(MAA)-St. They observed that the reaction behavior is greatly affected by the
type of acid monomer used, the partitioning of the monomer between the
various phases, and the locus of polymerization, and furthermore that the
mechanism for the AA-St system is more complicated than that for the MAA-
St system. They suggested that the primary reaction locus in the AA-St system
shifts from the particles to the aqueous phase after the hydrophobic monomer,
St, has been consumed. Yang et al. [149, 150] studied the effects of the initial
comonomer (AA) concentration on the monomer conversion and particle
number (NT) in the emulsifier-free emulsion copolymerization of St and AA.
They proposed an end-chain extension model to explain the experimental
results where the monomer conversion to the power of 2/3 is proportional to
the reaction time and log(NT)=36.00+9.44[AA]w. Slawinski et al. [151] investi-
gated the influence of AA on the particle growth process, with pH as the main
46 M. Nomura et al.
on the time evolution of the conversion, terpolymer composition, and the to-
tal number of polymer particles were investigated. The experimental results
were analyzed by means of a mathematical model that incorporated the main
features of the system. Ge et al. [162] studied the inverse emulsion copolymer-
ization of (2-methacryloyloxyethyl) trimethyl ammonium chloride and AAm
initiated with KPS. Aqueous monomer solutions were emulsified in kerosene
with a blend of two emulsifers (Span80 and OP10). Particle formation was
supposed to take place by monomer droplet nucleation. The observed rate of
polymerization is represented by RpµI 00.52S 00.38M 01.50, where M0 is the monomer
concentration.
3.3.4
Monomer Concentration in Polymer Particles
1 – XM
[M]p = [M]pc 94 (58)
1 – XMc
DFip 1 2
2Vmgj p1/3
7 = ln(1 – jp) + jp 1 – 31
– + cj p + 97 = 0 (59)
RT Pn r0RT
where DFip is the partial molar free energy of the monomer in the polymer
–
particles, jp is the volume fraction of polymer in the polymer particles, Pn is
48 M. Nomura et al.
Maxwell et al. [166] discussed the effects of several factors on the saturation and
partial swelling of polymer particles by monomers using Eq. 59 and the Vanzo
equation [168] that deals with the partial swelling of polymer particles in In-
terval III. By comparing theory and experiments for the MA and poly(MA-co-
St) system, the authors showed that the monomer partitioning was insensitive
to temperature, particle radius, copolymer composition, polymer molecular
weight, polymer cross-linking, the value of the Flory-Huggins interaction pa-
rameter, and the particle-water interfacial tension, and that the conformational
entropy of mixing of monomer and polymer was the significant term in de-
termining the degree of partial particle swelling by the monomer. Contrary to
Maxwell et al. [166],Antonietti et al. [167] observed a pronounced dependence
of the swelling ratio on particle size where absolute values of swelling were
much lower than those described by the classical Morton equation. In order
to explain this phenomenon, the authors presented a modified description that
considered size-relevant effects (such as the Kelvin pressure and depletion)
using an additional osmotic pressure term, which increases with the inverse of
the particle size. They also studied the effect of different types of covalently
bound surface stabilizing groups on the degree of swelling, and found that elec-
trically-stabilized particles resulted in higher swelling ratios and significant
lower values for the interfacial energy as compared to sterically stabilized
particles.
In an emulsion copolymerization, monomer partitioning between the mono-
mer droplet, polymer particle and aqueous phases plays a key role in deter-
mining the rate of copolymerization and the copolymer composition. Two
approaches (empirical and thermodynamic) have been proposed to predict the
monomer concentrations in the polymer particles in an emulsion copolymer-
ization system. In the emulsion copolymerization of St and MMA, Nomura et al.
[45, 122, 140] first proposed an empirical approach for predicting the saturated
concentration of each monomer in the polymer particles as a function of the
monomer composition in the monomer droplets, as shown by
1
[Mi]p = 9182 (61)
ai + bi /Wi,m
Emulsion Polymerization: Kinetic and Mechanistic Aspects 49
where the subscript i denotes monomer i, ai and bi are the numerical constants
particular to monomer i, and Wi,m is the weight fraction of monomer i in the
monomer droplets. Figure 5 is an example that shows good agreement between
the predictions from Eq. 61 and experimental results.
The authors demonstrated experimentally that the saturation monomer
concentration in the polymer particles was insensitive to particle radius and
copolymer composition, and also that the weight fraction of the monomer i in
the polymer particles (Wi,m) was approximately equal to that in the monomer
droplets (Wi,p); in other words,
Wi,m = Wi,p (62)
Thermodynamic methods were developed based on the extended equation by
Ugelstad et al. [169], and have been further dealt with by various researchers
[170–181]. Maxwell et al. [170] worked on the partitioning of two monomers
between the polymer particle, monomer droplet and aqueous phases in an
50 M. Nomura et al.
where jp is the volume fraction of polymer in the latex particles, jpi, jdi, jpj
and jdj respectively represent the volume fractions of monomers i and j in the
polymer particles and monomer droplet phases, cij, cip and cjp are the Flory-
Huggins interaction parameters between each of the respective monomers i
and j and the polymer, mij is the ratio of the molar volumes of monomers i and
j (so mij=Vmi/Vmj, where Vmi and Vmj are the molar volumes of monomers i and
j, respectively), [Mi]w is the concentration of monomer i in the aqueous phase,
and [Mi]w,sat is its saturation concentration value if there are no other mono-
mers present. The derivations of Eq. 63a and Eq. 63b involve the reasonable
assumption that mip and mjp, the ratios of the respective molar volumes of
monomers i and j and the molar volume of polymer are negligible compared
to all other terms. Furthermore, they made the following three assumptions to
simplify Eqs. 63a and 63b.
1. For many pairs of monomers, the differences between the molar volumes of
the monomers is slight. If this is the case, the ratio of the molar volumes of
monomer i and j is well approximated by unity, so mij=mji=1.
2. The contribution to the partial molar free energy arising from the residual
(enthalpic and non-conformational entropic) partial molar free energy of
mixing of the two monomers is small relative to all other terms in the
monomer droplet phase.
3. The interaction parameters for each monomer with the same polymer are
equal (cip=cjp).
They finally obtained the following simple expressions for saturation swelling.
jpi jdi
= (64)
jpj 42
42 jdj
Emulsion Polymerization: Kinetic and Mechanistic Aspects 51
where fpi, fdi, fpj and fdj represent the volume fractions of monomers i and j in
the polymer particle and monomer droplet phases, respectively. This equation
relates the ratios of the volume fractions (or concentrations) of monomers i
and j in the monomer droplet and particle phases. Eq. 65 is basically the same
as Eq. 62 derived empirically by Nomura et al. [122(a)]. The validity of Eq. 65
was experimentally demonstrated in the St-MA, St-BA and MA-BA systems us-
ing seed polymer particles with different copolymer compositions and diam-
eters. Based on Eq. 65 and experimental results, they finally proposed the fol-
lowing simple empirical expressions that could predict the concentration of
monomers i and j in the polymer particles.
Ci = fdi [(Ci,m – Cj,m)fd,i + Cj,m] (66a)
Cj = fdj[(Cj,m – Ci,m)fd,j + Ci,m] (66b)
where Ci and Cj are the concentrations of monomers i and j in the polymer
particles, and Ci,m and Cj,m the maximum saturations of monomers i and j in
the polymer particles (the homo-monomer swelling concentrations in the par-
ticles). By comparing the predictions from Eqs. 64 and 65 with experimental
data from the St-MA, St-BA, MA-BA systems, they demonstrated that Eqs. 64
and 65 could provide adequate predictions for the monomer concentrations
in the polymer particles. In these discussions, the ratio of the molar volumes
of monomers i and j (mij) was assumed to be unity. On the other hand, Schoon-
brood et al. [171] examined the validity of the assumption mij=1, made by
Maxwell et al. [170], and demonstrated that this assumption could be used with
systems where mij deviated from 1, at least up to a value of 2.
Noël et al. [172] experimentally determined both the saturation and partial
swelling of MA-VAc copolymer latex particles by MA-VAc monomer mixtures.
Monomer partitioning at saturation swelling could be predicted using the
simplified relationships developed by Maxwell et al. [170]. On the basis of the
work by Maxwell et al. [166, 170], Noël et al. developed an extended thermo-
dynamic model for monomer partitioning at the partial swelling of latex par-
ticles by two monomers with limited water solubility in order to predict the
monomer concentrations and fractions within the different phases, and con-
firmed the validity of this model by showing that the model’s predictions were
in good agreement with the observed monomer partitioning.
On the other hand, Schoonbrood et al. [173] investigated multimonomer
partitioning in latex particles and derived simple equations describing mono-
mer partitioning among the latex particle, monomer droplet and aqueous
phases during Intervals II and III in emulsion copolymerization with any num-
ber of low to moderately water-soluble monomers, by extending the approaches
developed by Maxwell et al. [170] and Noël et al. [172]. They showed that it is
mainly the conformational entropy from mixing the monomer and polymer
52 M. Nomura et al.
that governs the partitioning behavior, and that other contributions to the free
energies of the monomers in the polymer particles are marginal. They con-
firmed that all of the assumptions made in this study were valid using experi-
mental results for St, MMA and MA, and confirmed that the simple equations
proposed describe the monomer partitioning with these three monomers in
Intervals II and III very well. In this approach, the only parameters needed to
calculate the monomer concentrations in all of the phases were the saturation
concentrations of each monomer in the polymer particles, and the saturation
concentrations of each monomer in the aqueous phase.
By combining thermodynamically-based monomer partitioning relation-
ships for saturation [170] and partial swelling [172] with mass balance equa-
tions, Noël et al. [174] proposed a model for saturation and a model for partial
swelling that could predict the mole fraction of a specific monomer i in the
polymer particles. They showed that the batch emulsion copolymerization
behavior predicted by the models presented in this article agreed adequately
with experimental results for MA-VAc and MA-Inden (Ind) systems. Karlsson
et al. [176] studied the monomer swelling kinetics at 80 °C in Interval III of
the seeded emulsion polymerization of isoprene with carboxylated PSt latex
particles as the seeds. The authors measured the variation of the isoprene sorp-
tion rate into the seed polymer particles with the volume fraction of polymer
in the latex particles, and discussed the sorption process of isoprene into the
seed polymer particles in Interval III in detail from a thermodynamic point of
view.
These thermodynamic equations provide the most complete description
of the swelling of polymer particles by monomers, but include a rather large
number of parameters whose accurate estimation requires extensive work.
Considering this, Gugliotta et al. [175] presented a criterion for choosing which
monomer partitioning models developed so far in the mathematical modeling
of emulsion copolymerization should be applied to a given system. In the math-
ematical simulations, the seeded emulsion copolymerization of four monomer
systems with a wide variety of reactivity ratios and water-solubilities were con-
sidered: BA-St, VAc-MA, VAc-BA and St-MAA. They investigated the effect of
the complexity of the monomer partitioning equations, the type of process, the
solid contents, and the amount of seed on the time evolution of the conversion
and copolymer composition, and tabulated a summary of recommended
monomer partitioning models.
In the industrial production of structured AN-Bu-St (ABS) latex particles,
the grafting copolymerization of AN and St on crosslinked polybutadiene (PB)
seed latex is carried out in emulsion polymerization. Therefore, information on
the effect of PB crosslinking density on the swelling of PB latex particles by a
St-AN monomer mixture is very important for the production of ABS copoly-
mers with desired properties. Mathew et al. [177] studied the effect of several
thermodynamic parameters, such as the crosslinking density, particle size and
monomer mixture composition on the swelling behavior of PB latex particles
by pure St and AN, and St-AN mixtures of various compositions. They reported
Emulsion Polymerization: Kinetic and Mechanistic Aspects 53
that, in the case of mixtures, the higher the AN concentration in the mixture,
the lower the maximum swelling by St, and the opposite effect was observed for
AN swelling. The parameters describing the interaction between the two
monomers were found to be functions of the composition in the initial mixture.
Liu and Nomura et al. [178–180] carried out a series of investigations on the
swelling behaviors of St-AN (SAN) and ABS latex particles by St-AN monomer
mixtures. In the first article, Nomura et al. [178] examined the effects of copoly-
mer composition and its compositional inhomogeneity in SAN latex particles
on their swelling behavior, and found that both the copolymer composition and
the compositional inhomogeneity in SAN latex particles had little or no influ-
ence on the swellability of SAN latex particles with a St-AN monomer mixture,
as long as the weight fraction of AN monomer units in SAN latex particles was
less than a certain value (between 0.6 and 0.8). Based on the experimental data,
they proposed semiempirical equations that could predict the saturation con-
centration of each monomer in the SAN latex particles as a function of the
comonomer composition in the monomer droplets and the overall copolymer
composition in the SAN latex particles. In the follow-on article, Liu et al. [179]
investigated the possibility of a thermodynamic correlation between both
the partial and saturation swelling of SAN latex particles by St-AN monomer
mixtures. First, they determined the three unknown Flory-Huggins interaction
parameters between each monomer and homopolymer, cA,PA, cA,PS, and cS,PS (S:
styrene, PA: polyacrylonitrile, PS: polystyrene) by fitting the thermodynamic
swelling equations to the experimentally observed monomer concentrations in
SAN latex particles. Then, they showed that the AN concentrations predicted
by using these interaction parameters agreed fairly well with those observed.
However, agreement between the predicted and observed St concentrations
was somewhat worse than that for the AN concentrations. On the basis of the
preceding studies, Liu et al. [180] further studied the saturation swelling of ABS
latex particles by a St-AN monomer mixture. In order to describe the observed
saturation swelling behavior, they proposed a two-phase swelling model
based on the assumptions that in ABS latex particles, St-AN (SAN) copolymer
domains were randomly dispersed in a continuous PB matrix, and further that
thermodynamic equilibrium was attained among the SAN copolymer domain,
PB matrix and monomer droplet phases. By using the proposed model, the
effects of various factors on the saturation concentrations of each monomer
in the ABS latex particles were experimentally and theoretically discussed.
The factors examined were the polymer crosslinking density M 3C, the interfa-
cial tension between the PB latex particle and aqueous phases g, the ratio of the
molar volumes of the St and AN monomers mab, the weight fraction of AN units
in the SAN copolymer domains HA, and the weight fraction of PB in the ABS
latex particles HPB. It was found that the saturation concentration of each
monomer in the SAN latex particles was insensitive to M 3C, g, mAB and HA, but
that HPB had a large influence on the saturation concentration of the AN
monomer but almost no influence on the saturation concentration of the St
monomer. They finally concluded that the two-phase swelling model developed
54 M. Nomura et al.
3.3.5
Reaction Calorimetry
integral of the heat of reaction curve, and (2) nearly continuous information
is obtained. Using this information, a more detailed examination of the poly-
merization kinetics can be made, allowing the observation of important fea-
tures which cannot be seen using any other technique (such as gravimetry and
gas chromatography). Therefore, reaction calorimetry provides a powerful tool
for investigating heterophase polymerization [82]. Recently, many papers have
been published on reaction calorimetry studies into the kinetics of emulsion
polymerization. The kinetics of the emulsion polymerization of St was rein-
vestigated in detail [82–84, 184]. The effect of surface charge density on the
kinetics of the seeded emulsion polymerization of St was studied [185]. The de-
pendence of the reaction rate profiles on the water solubility of the monomers,
on the presence of CTA, on the types and concentrations of the stabilizer and
the initiator, and on the polymerization temperature was investigated [187].
The influence of oxygen on the kinetics of chemically initiated seeded emulsion
homopolymerization of St and the seeded emulsion copolymerization of St and
BA was investigated [188]. Reaction calorimetry has been used to estimate the
parameters in emulsion copolymerization systems [63], to control monomer
conversion and copolymer composition in semi-batch unseeded emulsion
copolymerization on-line [186, 189–191], and to monitor the copolymer com-
position, the average molecular weight, and the average degree of branching in
the semi-batch unseeded emulsion copolymerization of AN and Bu [147].
Most reaction calorimeters work according to heat-flow calorimetry prin-
ciples. The heat of reaction Qr evolved from a reaction mixture running at Tr
under isothermal conditions is transferred to the fluid in the cooling jacket
according to the equation
Qr = Rp (– DHp) = UAconst (Tr – Tj) (67)
where Rp is the reaction rate, DHp is the heat of reaction per unit amount of
reactant, U is the overall heat transfer coefficient, Aconst is the constant heat
transfer area, and Tj is the fluid temperature in the cooling jacket. The flow rate
of the cooling fluid is so high that the cooling fluid temperature at any position
in the jacket is considered to be, to a good approximation, equal to Tj. The tem-
perature of the cooling fluid Tj is adjusted to keep the reaction temperature
constant at Tr, and the heat of reaction Qr is found by calculating the value of
UAconst(Tr–Tj). Here, the heat transfer area Aconst and the value of U (determined
by calibration before and after the reaction) are treated as constant during
the reaction. However, the value of U is likely to change during the reaction
wherever the viscosity of the reaction mixture varies and/or the deposition of
scale on the heat transfer surface occurs during the reaction.
To avoid the drawbacks mentioned above, a novel calorimeter was developed
[192], as shown in Fig. 6, which can accurately measure the heat of reaction
independently of the variation of U during a reaction.
The working principle is as follows. A cooling fluid at temperature Ti at the
inlet is fed into the cooling jacket at a constant mass flow rate Fin, and the wet-
ted heat transfer area Avar in the jacket (which can be varied in this calorime-
56 M. Nomura et al.
ter) is controlled to keep the temperature of the reaction mixture at the desired
reaction temperature Tr by adjusting the fluid level (c) in the jacket, which is
achieved by regulating the flow rate of the cooling fluid at the outlet with a com-
puter-controlled throttle valve positioned there. The fluid temperature in the
cooling jacket is considered to be Tj throughout the jacket because of satisfac-
tory mixing due to the kinetic energy of the cooling fluid entering through a
nozzle. Therefore, the temperature of the cooling fluid flowing through the out-
let of the jacket is also Tj. To offset the heat loss from the reaction calorimeter
Qloss, a constant heat flux Qh(>Qloss) is passed into the reaction mixture through
an immersed electrical heater (f) in order to maintain the reaction mixture at
a constant temperature Tr even in the absence of the reaction. A reference run
is carried out with no reaction before measuring the heat of reaction. Then, the
steady-state heat balance for the reaction mixture in the calorimeter at tem-
perature Tr is given by
where Qm0 is the heat flux transferred to the cooling fluid across the reactor wall
and Tj is the cooling fluid temperature at the outlet of the jacket. Since the
values of Fin, Ti and Tj are measurable, the heat flux Qm0 is obtained by calcu-
lating the value of FinCp(Ti–Tj), where Cp is the specific heat of the cooling fluid.
Therefore, this calculated value of Qm0 gives a base-line reading and is constant
Emulsion Polymerization: Kinetic and Mechanistic Aspects 57
as long as Qloss remains constant. On the other hand, when the reaction is
allowed to take place at the constant temperature Tr, one gets the steady-state
heat balances given by
Qh + Qr = Qm + Qloss (70)
Qm = FinCpi (Ti – Tj) = UAvar (Tr – Tj) (71)
where Qm is the heat flux corresponding to the new reading when the reaction
is taking place. As long as the value of Qloss is kept constant by maintaining
the temperature around the region of the calorimeter constant, one gets the
following expression from Eqs. 68 to 71.
Qr = Qm – Qm0 (72)
Therefore, one can derive the heat of reaction Qr independently of the value of
U from the difference between the new and base-line readings.
3.4
Effect of Initiator Type
There are two types of chemical initiator that can be used to initiate emulsion
polymerization. They are water-soluble initiators (such as KPS, hydrogen per-
oxide-iron (II) redox system) and oil-soluble initiators (like azobis-isobuty-
ronitrile (AIBN), benzoil peroxide (BPO), benzoil peroxide-N,N-dialkylaniline
redox system). Water-soluble initiators are more commonly used in emulsion
polymerization than oil-soluble initiators. However, oil-soluble initiators are
sometimes used when the fragments derived from ionic water-soluble initia-
tors are not desirable either in the latex serum or on the surface of the polymer
particles. Water-soluble initiators produce almost all of the free radicals in
the water phase, because the amount of initiator partitioned into the organic
phases is usually negligible. Contrary to water-soluble initiators, oil-soluble
initiators distribute among the four phases: monomer-swollen micelles, mono-
mer-swollen polymer particles, monomer droplets (if any), and the water
phase. In the case of oil-soluble initiators, only a small fraction of radicals are
produced in the water phase because the amount of initiator partitioned into
the water is usally very small. Therefore, it is useful to find out whether the
different principal initiator loci of polymerization systems with water-soluble
and oil-soluble initiators brings about any differences in the kinetics and mech-
anisms of polymerization between both initiator systems.
Several researchers have carried out experimental and/or theoretical inves-
tigations on emulsion polymerizations initiated with oil-soluble initiators and
reported that the kinetics of the emulsion polymerizations is basically similar
to that initiated with water-soluble initiators [193–202]. Breitenbach et al. [193]
carried out the emulsion polymerization of St initiated by BPO at 50 and 60 °C.
The authors interpreted the experimental results by assuming a relatively
rapid exchange of low molecular weight radicals between the micelle-polymer
58 M. Nomura et al.
particle and water phases. Van der Hoff [194] conducted the emulsion poly-
merization of St initiated by cumen hydroperoxide (CHP) and suggested three
possible mechanisms. One possibility is the entry of single radicals generated
from the fraction of the initiator dissolved in the water phase. A second possi-
bility is that single radicals are formed by desorption of one of a pair radicals
(that form within the particles or by a side reaction) into the water phase. An-
other possible mechanism is that pairs of radicals are produced in the emulsi-
fier layer, and only the organic radical (C9H11O·) enters the particle, while the
inorganic initiator fragment (OH·) remains in the water phase where it can
undergo further reaction. The author stated that there was no direct evidence
that any of these three mechanisms in fact came into play. Dunn et al. [195]
carried out the emulsion polymerization of St at 60 °C using octadecyl sulfate
as the emulsifier and AIBN as the initiator. They found that the number of
polymer particles produced varied approximately with the 0.4th power of the
initially charged initiator concentration. This behavior is quite similar to that
usually found in the emulsion polymerization of St initiated with KPS. They
ascribed this kinetic similarity to desorption of either of the primary radicals
that formed as a pair into the water phase, leaving a single radical inside the
polymer particle for initiation. It was also found that only 4% of the whole ini-
tially charged initiator was effective in the emulsion polymerization in contrast
with an efficiency of ~50% found in bulk or solution polymerizations. Barton
et al. [196] investigated the effect of an oil-soluble initiator (AIBN) on the
kinetics and mechanism of the emulsion polymerization of BMA at 60 °C in
the presence of the anionic emulsifier disodium dodecylphenoxybenzene
disulfonate. They compared the results obtained with the course of the emul-
sion polymerization of BMA initiated by KPS, and proposed that the radicals
produced by decomposition in the aqueous phase determine the kinetics of the
polymerization.
On the other hand, Il’menev et al. [197] carried out the emulsion polymer-
ization of St at 50 °C using oil-soluble initiators such as AIBN, BPO and lauryl
peroxide (LPO). The water-solubilities of AIBN, BPO and LPO at 20 °C are 3.6,
0.1 and 0.01 mol/dm3-water, respectively. The rate of polymerization conducted
at 50 °C with 0.025 mol/dm3-St of each initiator was found to be (fastest to
slowest): AIBN, BPO and LPO; in other words, in the order of decreasing
water-solubility. They estimated the average times for a primary radical to ter-
minate, propagate, and desorb into the aqueous phase, respectively, when a pair
of radicals are generated in a micelle and a polymer particle. They concluded
that the contribution to the polymerization (particle formation and growth)
from the free radicals that are produced in pairs in the micelles and polymer
particles is almost negligible, because they are very likely to cause rapid gem-
inate termination, and that the free-radicals generated in the monomer droplets
also play only a small part in the polymerization, because their desorption into
the water phase can be ignored. This view was strongly supported later by the
theoretical and experimental work of Nomura et al. [198–202]. Nomura et al.
[198] proposed a theoretical approach by which the effects of various factors
Emulsion Polymerization: Kinetic and Mechanistic Aspects 59
1. The latex (polymer) particles are generated from the emulsifier micelles and
the number of latex particles produced is proportional to the 0.70th power
of the initial concentration of the emulsifier forming micelles and to the
0.30th power of the concentration of initially charged AIBN. This behavior
is very similar to that observed when the water-soluble initiator KPS is used.
2. The polymerization takes place both in the monomer droplets and in the
latex particles produced. The polymerization inside the monomer droplets
proceeds according to the kinetics of suspension polymerization until the
60 M. Nomura et al.
Therefore, they showed both theoretically and experimentally that the kinetic
behavior of the emulsion polymerization of St initiated by AIBN is basically sim-
ilar to that initiated by KPS, and concluded that this similarity is mainly due to
the radicals produced from the water-soluble fraction of the initiator, because
the radicals produced pair-wise inside the small volume of a monomer-swollen
latex particle or a monomer-swollen micelle are very likely to recombine.
Several researchers have also experimentally and theoretically investigated
the reasons for this kinetic similarity [203–208]. Asua et al. [203] proposed a
mathematical model that can predict the average number of radicals per par-
ticle –n in seeded emulsion polymerization initiated by oil-soluble initiators.
Their model includes the parameter fw that denotes the fraction of the initia-
tor dissolved in the aqueous phase, and the following various kinetic events: (i)
generation of radicals inside the particles, (ii) desorption of primary initiator
radicals from the polymer particles before reacting with a monomer molecule,
(iii) termination of radicals by bimolecular reaction in the particles, (iv) de-
sorption of single-unit monomeric radicals produced by chain transfer to
monomer molecules, (v) absorption of radicals from the aqueous phase into the
particles, (vi) termination of radicals in the aqueous phase, and (vii) generation
of radicals in the aqueous phase by decomposition of the initiator dissolved in
that phase. They calculated the average number of radicals per particle in a typ-
ical example of the seeded emulsion polymerization of St, using their model that
distinguishes between desorption of primary initiator radicals and single-unit
monomeric radicals. The effect of increasing the water-soluble fraction fw of the
initiator from 0 to 0.1 was calculated for various particle diameters in the range
Emulsion Polymerization: Kinetic and Mechanistic Aspects 61
23–231 nm. For a fixed particle size, the value of – n was found to be essentially
independent of the fraction fw of the initiator present in the aqueous phase,
even for fw=0. Moreover, the plot of – n versus the seed particle diameter was
quite similar to that found for the emulsion polymerization initiated by KPS.
The authors therefore concluded that the kinetic similarity mainly originated
from desorption of the initiator radicals produced in the particles rather than
from decomposition of the initiator present in the aqueous phase. Mørk et al.
[208], however, pointed out that the almost identical values of – n found for a
completely water-insoluble initiator (fw=0) appeared to contradict calculations
performed by Mørk et al. [208] and Nomura et al. [198], and that calculations by
Asua et al. [203] could not be taken as evidence that a desorption mechanism is
the reason for the similarity. Alduncin et al. [204] studied the seeded emulsion
polymerization and the miniemulsion polymerization of St using an oil-soluble
initiator (AIBN) in an attempt to elucidate the main locus of radical formation
in emulsion polymerization initiated by an oil-soluble initiator. The monomer/
water weight ratio (M/W) was varied while keeping the monomer/initiator
ratio constant. They found that the average number of radicals per particle (n –)
increased as the M/W ratio increased. This was taken as evidence that the over-
all rate of radical entry into a particle increased when the M/W ratio increased.
They claimed that this phenomenon could only be explained by assuming that
the radicals responsible for emulsion polymerization initiated by oil-soluble
initiators are mainly those produced from the initiator partitioned into the
polymer particles, followed by desorption into the water phase. Mørk et al. [208]
concluded, on the basis of their calculations, that the argument provided by
Alduncin et al. [204] was not strong enough to resolve the issue of the similar-
ities between the kinetic behaviors of emulsion polymerization with oil-solu-
ble and water-soluble initiators.
Mørk et al. [206–208] recently published a series of theoretical works. In the
first article of this series [206], the authors aimed to develop expressions that
would allow easy and rapid calculation of the average number of radicals per
particle in emulsion polymerizations with a constant number of reaction loci
containing an oil-soluble initiator. Taking into account pairwise formation of
radicals in the particles, desorption and reabsorption, water phase termination,
solubility of the initiator in the water phase, and the possible formation of a
single radical species, they derived the recurrence relation that determined the
stationary state distribution of radicals in a particle. The calculation was based
on a probabilistic analysis leading to a third-order recurrence relation solved
using confluent, hypergeometric Kummer functions. The calculated results
confirmed the previous finding of Nomura et al. [198] that the kinetics of emul-
sion polymerizations carried out with oil-soluble initiators are quite similar to
those with water-soluble initiators, provided that the oil-soluble initiator is not
completely insoluble in the water phase. The main intention of the second
article [207] was to develop equations that make it relatively easy to assess
the effects of the most common experimental variables on the stationary state
average number of radicals per particle in a bidispersed seeded emulsion poly-
62 M. Nomura et al.
but at a considerably reduced emulsifier level. They extended this study, mainly
to clarify the effects of varying the input ultrasound intensity [213], and found
that: (1) a marked increase in the rate of polymerization was seen as the input
power was increased; (2) despite the increase in the rate of polymerization, the
increasing intensity did not affect the resultant polymer particle sizes, which
were in all cases 40–50 nm, and; (3) increases in both the concentration of NaLS
and the reaction temperature resulted in an increased rate of polymerization
at a fixed input intensity, but the particle sizes were invariant.
Cooper et al. [214] carried out the emulsion homopolymerization of BA and
of VAc, and also the emulsion copolymerization of BA and VAc at 30 °C (±5 °C)
using 20-kHz ultrasound as the initiator with SDS and Aerosol AT as the emul-
sifiers, respectively. The homopolymerization rate of VAc (10 wt%) was much
lower than that of BA (10 wt%), and interestingly, lower rates of BA emulsion
homopolymerization were observed at higher temperatures. The reason for
such a large difference in the rate of polymerization between the BA and VAc
systems was explained by the greater evaporation of the more volatile VAc
monomer into the cavity, suppressing cavitation and thereby reducing the rate
of radical production. The average particle sizes produced in the BA system and
the copolymerization system with 50:50 wt% BA and VAc were very small;
around 15–20 nm, respectively. But the average particle sizes produced in the
emulsion homopolymerization of VAc were much larger, showing a size of
around 300 nm. The reason for producing smaller polymer particles in both the
BA homopolymer and BA-VAc copolymer systems than in the VAc homopoly-
mer system even at low emulsifier concentrations was attributed to a high rate
of particle formation due to a large number of very small monomer-emulsion
droplets that were to be transformed into polymer particles.
Grieser et al. [215] investigated the kinetics and mechanisms of the emulsion
polymerization of MMA and of BA at 30 °C (±5 °C) using ultrasonic irradiation
(20-kHz horn sonifier) and a cationic emulsifier, dodecyltrimetylammonium
chloride (DTAC). They observed the formation of stable dispersions with par-
ticle diameters in the range of 40–150 nm and with polymer molecular weights
greater than 106 g mol–1. In the case of MMA, the average particle size was found
to be constant throughout the reaction time (sonication time) and independent
of the initial DTAC concentration. The final particle size decreased as the initial
DTAC concentration was decreased, but the rate of polymerization was ap-
proximately the same over the concentration range of DTAC examined. In the
case of BA, the kinetic behavior was basically the same as that of MMA, except
that the average particle size was constant (~30 nm) up to 50 min of sonication,
after which a dramatic increase in size (100–140 nm) was observed when the
initial TDAC concentration was comparatively low. Based on their experimen-
tal data, the authors proposed the kinetics and mechanisms of the ultrasonic
(sonochemical) initiation in this polymerization process. When ultrasound
is applied in a liquid medium, the cavitation event that occurs as ultrasound
travels through the liquid medium, producing microbubbles in the solution.
When the microbubbles rapidly collapse, this leads to high local temperatures
Emulsion Polymerization: Kinetic and Mechanistic Aspects 65
of the order of 4000–5000 K within the bubble and at least 1250 K in the liquid
immediately surrounding the interfacial region. In an aqueous medium, such
high temperatures lead to the homolysis of water, creating hydroxyl (·OH) and
hydrogen (·H) radicals. The authors assumed that primary organic radicals
produced from MMA and BA were unlikely to play a major role at 20 kHz even
though MMA and BA are volatile and could enter cavitation bubbles to produce
a variety of primary organic radicals by thermal decomposition. The hydroxyl
and hydrogen radicals generated in the aqueous phase add several monomer
units and then enter the miniemulsion droplets produced by ultrasonication,
initiating polymerization. Therefore, the results obtained strongly support a
polymerization process involving a miniemulsion polymerization system,
where continuous formation of polymer particles takes place throughout the
polymerization.
Chou and Stoffer [216, 217] carried out the ultrasonically-initiated free rad-
ical emulsion polymerization of MMA at ambient temperature using NaLS as
the emulsifier, and published two articles on this topic. In the first article [216],
the authors studied: (1) the nature and source of the free radicals for the initia-
tion process; (2) the effects of different types of cavitations, and; (3) the depen-
dence of the polymerization rate, the number of polymer particles generated,
and the polymer molecular weight on the acoustic intensity, argon gas flow
rate, surfactant concentration, and the initial monomer concentration. They
found that, in the absence of argon gas flow, no polymerization took place, and
that, contrary to Grieser et al. [215], the source of the free radicals for the ini-
tiation process came from the degradation of the NaLS, presumably in the
aqueous phase. The molecular weight of the poly(MMA) obtained varied from
(2.5–3.5)¥106 g mol–1, and the monomer conversion was up to 70%. The rate of
polymerization was found to be proportional to the acoustic intensity to the
power of 0.98, to the argon gas flow to the power of 0.086, and to the emulsifier
concentration to the power of 0.08 in the emulsifier concentration range of
0.035–0.139 M. The number of polymer particles was found to be proportional
to the acoustic intensity to the power of 1.23, to the argon gas flow to the power
of 0.16, and to the emulsifier concentration to the power of 0.3 in the emulsifier
concentration range of 0.035–0.139 M. In the second article [217], the radical
generation process was studied. Based on this experimental study, the authors
tried to explain the kinetic data obtained in the previous work. In this study, rad-
ical trapping experiments were used to investigate the effects of acoustic inten-
sity, argon gas flow rate, and NaLS concentration on the extent of free radical
generation in aqueous NaSL solutions. Aqueous solutions of NaLS were ul-
trasonically irradiated in the presence of a radical scavenger. The NaLS mol-
ecules then decomposed by ultrasound to form free radicals in the aqueous
phase. It was found that the extent of free radical generation increased as: (1)
the 0.6th power of the acoustic intensity, (2) the 0.44th power of the argon gas
flow rate, (3) the 0.35th power of the emulsifier concentration in the emulsifier
concentration range of 0.035–0.139 M. These experimental results were found
to explain the effects of acoustic intensity, argon gas flow rate, surfactant con-
66 M. Nomura et al.
3.5
Effect of Additives and Impurities
It has been recognized that the presence of oxygen during emulsion poly-
merization can have detrimental effects on the course of a reaction, causing
inhibition periods and retarding the reaction rate. Relatively few publications
have addressed the issue of the effects of oxygen in emulsion polymerization
[126, 188, 220–223]. With the intention of clarifying the effect of stirring on
emulsion polymerization, Nomura et al. [220] carried out St emulsion poly-
merization under three nitrogen atmospheres with different purities (contain-
ing a trace of oxygen) and at different stirring speeds. They observed that the
faster the stirring speed, the longer the retardation period. They attributed the
result to the diffusion limited transfer of oxygen from the headspace into the
water phase through the liquid surface, which was controlled by stirring. Fur-
thermore, they found that the polymerization rate following a long retardation
period was often greater than that after a shorter retardation period, indicating
that the final number of polymer particles produced with a long retardation
period was higher than that with a shorter retardation period. The reason for
this is discussed later. The same trend was also observed by other researchers
[188, 222]. Cunningham et al. [222] examined the effects of oxygen on the in-
duction period, conversion kinetics, molecular weight and particle size during
the emulsion polymerization of St, by varying the initial dissolved oxygen con-
centration in the aqueous phase. They found that the length of the induction
period did not vary linearly with the initial oxygen level, suggesting diffusion
from the reactor headspace to the aqueous phase could have a significant impact
on rates of particle formation and growth. Furthermore, the higher the initial
dissolved oxygen level, the longer the induction period and the smaller the
average diameter of polymer particles in the final latex product, which indicates
that the longer the induction period, the greater the number of polymer parti-
cles produced. Their experimental results suggested that, during the induction
and retardation period, the oxygen molecules in the reactor headspace were
continuously transferred into the aqueous phase, and some of them are con-
sumed by the radicals in the aqueous phase, but most of them diffuse further
into both the monomer-swollen micelles and polymer particles. Therefore, the
oxygen molecules that have diffused into the monomer-swollen micelles and
polymer particles inhibit the growth of radicals within them, thereby reducing
the volumetric growth rate per particle, m and resulting in an increase in the
number of polymer particles produced according to Eq. 29.
Arbina et al. [188] investigated the influence of oxygen on the kinetics of the
chemically-initiated seeded emulsion homopolymerization of St and the seeded
emulsion copolymerization of St and BA using reaction calorimetry. They dis-
cussed whether oxygen behaved kinetically as an ideal inhibitor. In the experi-
ments, they observed that oxygen not only caused an inhibition period, but also
behaved like a retarder by reducing the polymerization rate. Their explanation
for this seemingly contradictory behavior was the existence of mass-transfer
limitations from the reactor headspace to the latex, resulting in a gradual and
continuous flow of oxygen into the aqueous phase. Their own experiments
showed that this induction period decreased with increasing initiator concen-
68 M. Nomura et al.
tration. When the headspace to aqueous phase ratio was decreased, the induc-
tion period was reduced and the polymerization rate increased. The length of
the retardation period caused by oxygen in the seeded emulsion polymerization
is known to depend on the kind of monomer. In seeded emulsion polymeriza-
tions, the inhibition period may be followed by a retardation period during
which the polymerization rate increases to a steady state value. The retardation
period observed in the seeded emulsion polymerization of VAc is unusually long
compared to that of St or MMA. Bruyn et al. [223] tried to quantitatively explain
the reason for this unusually long retardation in terms of the initiator effi-
ciency, fentry, proposed by Maxwell et al. [11]. They argue that this unusually
long retardation is due to the high radical entry efficiency of the aqueous phase
oligomeric radicals, which allows latex particles to compete with dissolved oxy-
gen for these initiating radicals. In the case of VAc, this is due to the high value
of the product of the propagation rate constant and the water solubility.As oxy-
gen is consumed, the competition increasingly favors the entry of initiating rad-
icals into polymer particles and the rate of polymerization gradually increases.
In another case, Kiparissides et al. [126] considered the different effects of the
presence of oxygen on the kinetics and PSD in the emulsifier-free emulsion
polymerization of VCl. Oxygen is capable of reacting with primary initiator rad-
icals and the resulting oligomeric radicals in the aqueous phase to produce vinyl
polyperoxides. Taking this into account, they developed a mathematical model
into which the combined role of oxygen as an inhibitor and a radical generator,
through the formation and subsequent decomposition of vinyl polyperoxides,
was incorporated.
CTAs are used not only to reduce the molecular weight of the polymer pro-
duced, but also to limit the extent of the branching and crosslinking of the poly-
mer produced in diene-polymerization. It is well known that an ideal CTA is able
to reduce the molecular weight of the polymer produced in a homogeneous bulk
or solution free radical polymerization, without affecting the overall rate of
polymerization. For a long time it has been known that even an ideal CTA such
as mercaptan could affect not only the molecular weight of the polymer pro-
duced in emulsion polymerization, but also the rate of the polymerization [224],
but details about the effects of CTAs (including mercaptans) on the kinetics of
heterophase radical polymerizations, like emulsion polymerizations, have only
been revealed recently [225–229].
Whang et al. [225] observed that the rate of polymerization decreased in the
seeded emulsion polymerization of St in the presence of carbon tetrabromide,
which functions as an ideal CTA in the homogeneous bulk or solution free
radical polymerization of St. They pointed out that this appeared to result from
the enhanced rate of radical desorption of the free radicals from the polymer
particles. Also, a substance which behaves as an ideal CTA in homogeneous
polymerizations might apparently function as a retarder in heterogeneous free
radical polymerization. Lichti et al. [227] advanced the discussion by Whang
et al. [225] and argued that the increase in the rate of radical desorption which
brought about the decrease in the rate of polymerization paralleled the increase
Emulsion Polymerization: Kinetic and Mechanistic Aspects 69
in the chain transfer constant for the additives: CBr4>CCl4>St. The efficiency
of desorption of free radicals formed by chain transfer from the latex particles
followed the inverse order CBr4<CCl4<St, and this reflected the reactivities of
the low-molecular weight free radicals (formed by atom abstraction) with the
monomer.
At almost the same time, Nomura et al. [226] carried out an extensive ex-
perimental study of the effect of typical CTAs such as CCl4, CBr4 b, and four pri-
mary mercaptans (C2, n-C4, n-C7 and n-C12) on particle formation and growth
processes in the unseeded emulsion polymerization of St. They found that these
CTAs, which had almost no effect on the rate of bulk polymerization of St, de-
creased the rate of polymerization per particle m and so increased the number
of polymer particles produced (see Eq. 29). From a theoretical point of view,
they suggested that these effects could be enhanced by increasing the rate of
radical desorption from the polymer particles by adding a CTA with a higher
value of the chain transfer constant and/or with higher water-solubility.
Nomura et al. [20, 43] pointed out that the number of polymer particles pro-
duced NT could be expressed as NTµS 0zI 01–z in an emulsion polymerization sys-
tem following micellar particle formation. They also showed that, by increasing
the rate of radical desorption from the polymer particles in the interval of par-
ticle formation with the help of CTA, the emulsifier dependence exponent, z,
would increase from about 0.6 to 1.0, thereby decreasing the initiator depen-
dence exponent from about 0.4 to 0. This was also confirmed experimentally.
Therefore, they demonstrated that the effect of CTA on particle formation and
growth in the emulsion polymerization of St could be explained in terms of
desorptions of chain-transferred radicals from the polymer particles.
Maxwell et al. [228] discussed the effect of CTA, such as mercaptan, on ini-
tiator efficiency and extended their quantitative model for initiator efficiency
to take into account the effect of adding CTA. They assumed the following
model. The effect of the CTA on the entry rate occurs by facilitating the pro-
duction of aqueous-phase free radical species (CTA radicals) via transfer be-
tween species such as ·MnSO4– (where M is a monomer entity and n<z) and the
CTA in the aqueous phase. The CTA radicals will be formed at a reasonable rate
provided that the CTA is not too water-insoluble (as in C12H25SH) and the
resultant CTA radical is able to enter the latex particles rapidly because of this
relative insolubility in water. If the monomer-derived ·MnSO4– tends to suffer
aqueous-phase termination rather than entry, the overall rate of entry (and
hence initiator efficiency) will increase. They claimed that this mechanism
could explain the accelerating (promoting) effect of intermediate molecular
weight CTAs (C10–C12) on the emulsion polymerization of monomers such
as Bu, where the z value is large and so initiator efficiency is very low in the
absence of CTA, because most of the ·MnSO4– undergoes termination rather
than entry into the latex particles. However,Weerts et al. [229] studied the “pro-
moting effect” in the emulsion polymerization of Bu using SDS, potassium
stearate and potassium oleate as the emulsifiers and sodium or potassium per-
oxodisulfate, 4,4¢-azobis-(4-cyanopentanoic acid and AIBN as the initiators, and
70 M. Nomura et al.
Fig. 7 A typical example of the effect of water-soluble (APS) and oil-soluble (DBP) initiators
on the progress of the emulsion polymerization of St in the presence of a water-soluble
radical inhibitor (Fremy’s salt, FS); for [APS]=5¥10–4 mol/dm3, empty circles indicate [FS]=0,
filled circles indicate [FS]=10–4 mol/dm3; for [DBP]=5¥10–4 mol/dm3, empty squares indicate
[FS]=0, filled squares indicate [FS]=10–4 mol/dm3
72 M. Nomura et al.
inhibition period was observed, but after the end of the inhibition period the
conversion versus time curve was almost the same as that encountered in the
absence of Fremy’s salt. On the other hand, when the emulsion polymerization
of St was initiated by the oil soluble initiator DBP, the monomer conversion
versus time curve observed in the presence of Fremy’s salt was identical to that
seen in the absence of Fremy’s salt. In the case of MMA, the results were basi-
cally the same as those with St. Also, the rate of emulsion polymerization ini-
tiated by DBP was almost the same as the rate of bulk polymerization initiated
by DBP. This indicated that in the emulsion polymerization initiated by DBP,
the polymerization did not proceed in the monomer-swollen micelles and the
resultant polymer particles according to emulsion polymerization kinetics, but
in the monomer droplets according to bulk kinetics. This implies that neither
member of a radical-pair generated in a monomer-swollen micelle initiates
polymerization, either because geminate termination took place before either
of the pair of radicals desorbs from the micelle into the aqueous phase, or
because both of the radicals desorb as soon as they are generated and are scav-
enged in the aqueous phase. This finding is closely related to the claim [200]
that in the emulsion polymerization of St initiated by oil-soluble initiators, the
free radicals generated from the fraction of the initiator dissolved in the aque-
ous phase mainly participate in particle formation from the monomer-swollen
micelles. Barton et al. claimed that, although many studies had been performed
to clarify the effect of a variety of water-soluble and oil-soluble inhibitors on
emulsion polymerization, no unequivocal results had been obtained for their
effects on its kinetics, and specifically for their effect on particle formation, and
that this problem is still open for discussion.
Ignoring their side effects, chain transfer agents (CTAs) were originally
used in emulsion polymerization as additives to regulate the molecular weight
distribution of the resultant polymers and to limit the extent of branching and
crosslinking of the polymer produced in diene-polymerization. Recently, sev-
eral investigations based on this point of view have been published [57, 125,
233–238]. Barudio et al. [57] studied the effect of CTAs (tert-butanethiol and
n-dodecanethiol) on the microstructures of copolymers (the molecular weight
distribution (MWD) and glass transition temperature (Tg)) and the diameters
of polymer particles produced in the batch and semibatch emulsion copoly-
merizations of St and BA. The experimental results were interpreted in terms
of enhanced radical desorption and diffusion limitations of CTA between the
monomer droplet and particle phases. They proposed a kinetic model that was
able to successfully compute the kinetic constants, the number of radicals per
particle, the GPC/SEC diagram and the DSC thermogram related to the MWD
and Tg, respectively. Salazar et al. [234] developed a mathematical model that
included the effect of CTA (tert-nonyl mercaptan) on particle formation and
the average molecular weights in the batch and monomer-starved emulsion
polymerizations of St. Asua and co-workers [125, 233, 235–237] published
several reports on the effects of CTAs on the kinetics and the microstructures
of the resultant polymers in seeded and unseeded emulsion homo- and co-
Emulsion Polymerization: Kinetic and Mechanistic Aspects 73
3.6
Effects of Other Important Factors
Fig. 8 A typical example of the effect of stirring on the progress of the emulsion polymer-
ization of St in the presence of a typical inhibitor oxygen (40 °C, initiator: H2O2)
Emulsion Polymerization: Kinetic and Mechanistic Aspects 75
ture from nitrogen atmosphere in which the polymerization was carried out.
Evans et al. [244] carried out the emulsion polymerization of VDC at 36±1 °C
using NaLS as the emulsifier and APS as the initiator, and found that: (1) the
first stage polymerization rate decreased with increasing stirring speed; (2) the
second stage polymerization rate increased with increasing stirring speed, and;
(3) the third stage polymerization rate was independent of stirring speed. To
explain their results, they suggested two factors through which stirring affected
the polymerization rate. The first factor was the reduced levels of effective
emulsifier available for the formation of polymer particles, caused by the
adsorption of emulsifier molecules onto monomer droplets finely dispersed by
the stirring (in the first stage). The other factor was the effect of monomer-
transport from the monomer droplets to the polymer particles where the poly-
merization proceeded (in the second stage) upon the rate. Omi et al. [245] came
to the contrary conclusion that when the monomer used was St, stirring did not
influence emulsion polymerization as long as initial emulsification conditions
were not changed. They considered that stirring affected the polymerization
only through the former of the two factors suggested by Evans et al. [244]. Later,
Nomura et al. [75] carried out a kinetic study of the batch emulsion polymer-
ization of VDC at 50 °C with KPS as the initiator and NaLS as the emulsifier,
aiming to derive a quantitative explanation for the effect of stirring observed
by Evans et al. [244]. Unfortunately, however, the polymerization proceeded
very smoothly and the abnormal kinetic behavior caused by a change in the
stirring speed was not observed. Therefore, they concluded that the effect of
stirring observed by Evans et al. [244] in the VDC emulsion polymerization
must be a very special case. Therefore, the reasons for such abnormal kinetic
behavior remain a mystery.
These results illustrate that investigations into the effects of stirring on
emulsion polymerization, have produced inconsistent results and conclusions.
Further research was therefore needed to elucidate the effect of stirring in more
detail. Nomura et al. [220] carried out an extensive investigation into the effect
of stirring on the emulsion polymerization of St initiated by KPS at 50 °C with
NaLS as the emulsifier, with the intention of explaining the effects of stirring
quantitatively by showing how agitation affects emulsion polymerization, what
steps of the polymerization are affected by stirring, and whether a suitable
range of agitation exists in emulsion polymerization. The reactor used was
a cylindrical glass vessel with a dished bottom, fitted with four baffle plates
located at 90° intervals, and a four-bladed paddle type impeller. They concluded
that the effect of stirring on emulsion polymerization appears through the
following four factors:
On the other hand, Weert et al. [246] investigated the effects of stirring on the
kinetics of the emulsion polymerization of Bu at 60 °C using NaLS as the emul-
sifier. They carried out the polymerizations in a 2.3-liter reactor fitted with four
baffle plates located at 90° intervals and a twelve-flat-bladed turbine impeller.
In all of the experiments, the system was pre-emulsified by stirring for a few
minutes at 400 rpm, before adjusting the stirring speed n to the desired level.
The number of polymer particles produced, NT, was found to remain constant
within experimental error beyond a sufficiently high value of n, while a discon-
tinuous increase in NT became apparent towards lower values of n. The change
in NT was significant, especially at low n. They ascribed the reason for this
change to the fact that the level of emulsifier available for particle formation and
stabilization was influenced by the degree of agitation, as already pointed out by
Nomura et al. [220]. They discussed the effect of stirring in connection with the
flow conditions in the reactor, which are closely related to the stirring speed, and
Emulsion Polymerization: Kinetic and Mechanistic Aspects 77
finally arrived at the conclusions that the stirring speed influenced this poly-
merization system by reducing the effective emulsifier concentration available
for particle formation and stabilization at higher n, and by limiting the diffu-
sion of monomer to the polymer particles at low n. Arai et al. [247] studied the
effect of agitation on the kinetics of the soapless emulsion polymerization
of MMA in water at 65 °C. As expected, the stirring influenced the monomer
conversion versus time history, the molecular weight of the polymer produced,
and the number of polymer particles produced versus time. They found that
agitation was an important factor that affected the rate of monomer-transport
from the monomer droplets to the water phase. The authors proposed a quan-
titative kinetic model that could predict the effect of monomer transport on
the rate of polymerization. Kostov et al. [158] also studied the effects of poly-
merization conditions, including stirring, on the emulsion copolymerization of
tetrafluoroethlene (TFE) and propylene (P) with ammonium perfluorooc-
tanoate, initiated by a redox initiator system containing tert-butylperbenzoate.
They found that both the rate of copolymerization and the molecular weight
of the copolymer produced increased as the stirring speed increased up to
450 rpm, but then became independent of the speed above 450 rpm. The ex-
planation for this was that the stirring affected the rate of mass-transfer of TFE
and P from the gaseous to aqueous phases, which was the rate-controlling step
for speeds less than 450 rpm. Kim et al. [248] also reported the importance of
agitation in the semi-batch emulsion polymerization of TFE carried out using
a chemical initiator (APS) and a fluorinated surfactant (FC-143). The rate of
polymerization was found to increase linearly with the stirring speed. Based on
the experimental findings, they concluded that the diffusion or dissolution
of TFE into the aqueous phase was the rate-determining step through which
agitation affected the polymerization. Özdeģer et al. [249] investigated the ef-
fect of stirring speed and impeller type (axial and radial flow impellers) on the
kinetics of the emulsion copolymerization of St and n-BA using Triton X-405
(octylphenoxy polyethoxy etanol) as the emulsifier.At low solids content (30%),
the impeller type and speed did not have any significant effect on both the fi-
nal number of polymer particles and the overall rate of polymerization. The
PSDs were unimodal in all cases. For high solids content (50%), the rate of poly-
merization carried out with the axial flow impeller was slower, indicating that
fewer polymer particles were produced. Bimodal PSDs were obtained for both
cases. These differences were attributed to the partitioning of the emulsifier.
The axial flow impeller created more shear than the radial flow impeller. This
resulted in more monomer droplets being formed, leading to more of the emul-
sifier being associated with them. This also resulted in fewer emulsifier micelles
being available for particle formation, thereby leading to the lower number of
polymer particles produced.
In industrial emulsion polymerizations, CTAs like mercaptan are often used
to regulate the molecular weight of the polymer produced. In some cases, other
ingredients that directly participate in the polymerization reaction are used
to modify the properties of the polymer latex produced. In these cases, these
78 M. Nomura et al.
reacting species must be transported from one phase, for example, the monomer
droplets, via the aqueous phase, to the monomer-swollen polymer particles
where the reaction takes place. Therefore, when designing a latex product to
have particular properties, it is important to quantitatively elucidate the diffu-
sional behavior of these reacting species when they move between the phases
in an emulsion polymerization system. However, only a few researchers [125,
234, 239, 250–255] have presented quantitative discussions on the mass-trans-
fer problem involved in emulsion polymerization. Brooks [250] discussed the
monomer diffusion rate in an emulsion polymerization system. The author
calculated the maximum diffusion rate from monomer droplets to a polymer
particle via the water phase by using a simple diffusion equation and showed
that this rate was usually far greater than the rate of polymerization per parti-
cle. He also suggested that an adsorbed emulsifier layer on the surface of the
polymer particles would not impede monomer transfer to the particles. Finally,
he concluded that in most systems the diffusional processes that occurred in
the water phase would not affect the course of the polymerization. However,
Nomura et al. [220] pointed out a possibility that the monomer-transport step
from the monomer droplets to the water phase could control the polymeriza-
tion rate when the intensity of agitation was comparatively low.
Nomura et al. [239] discussed the mass-transfer problem in more detail for
the seeded emulsion polymerization of St initiated by KPS at 50 °C using NaLS
as the emulsifier and CTAs as the diffusing species. They carried out the seeded
emulsion polymerization of St using five normal aliphatic mercaptans with dif-
ferent molecular weights (n-C7, n-C8, n-C9, n-C10, and n-C12) under the condi-
tions of 400 rpm (stirring speed), NT=1.4¥1014 particles/cm3-water (the num-
ber of seed polymer particles), dpo=48 nm (the average diameter of seed
particles), and Tmo=8.44¥10–6 mol/cm3-water (the concentration of initially
charged mercaptan per unit volume of water), and measured the consumption
rate of each mercaptan. They found that the consumption rate decreased dras-
tically with the molecular weight of mercaptan, although the consumption rate
of n-C8 was not so different from that of n-C7. They analyzed these experimen-
tal data using a proposed diffusion model derived on the basis of the so-called
two-films theory, and concluded that the concentration of the CTA in the poly-
mer particles during the polymerization dropped to a value much lower than
the one that would be attained if thermodynamic equilibrium for the CTA were
reached between the monomer droplets and the polymer particles. This was
due mainly to the CTA molecules’ resistance to transfer across the diffusion
film at the interface between the monomer droplet and water phases. The au-
thors suggested that the proposed model can also be used to predict the rate of
mass-transfer of any sparingly water-soluble reacting species from monomer
droplets to the polymer particles where this species participate directly in the
polymerization (this may include the monomer itself).
It is well known that emulsion polymerizations of highly water-insoluble
monomers such as octadecyl methacrylate (OM), dodecyl methacrylate (DM),
and stearyl acrylate (SA) are generally not feasible using traditional surfactant
Emulsion Polymerization: Kinetic and Mechanistic Aspects 79
systems. This is because monomer transport from the monomer droplets to the
water phase is diffusion limited [239]. However, almost simultaneously, Rimmer
et al. [256, 257], Leyer et al. [258] and Lau et al. [259] reported that these highly
water-insoluble monomers could be emulsion polymerized in the presence of
b-cyclodextrin (b-CD). These studies uncovered a very interesting phenome-
non. Rimmer et al. [256, 257] successfully conducted the emulsion polymeriza-
tion of OM and DM at 70 °C using KPS as the initiator and Dowfax 2A1 as the
emulsifier. They claimed that the reason for this successful polymerization was
that the use of b-CD appeared to aid monomer transport from the monomer
droplets to the polymer particles across the aqueous phase by increasing the
apparent water-solubility of these monomers, because the CDs apparently
solubilize the hydrophobic compounds in aqueous media. On the other hand,
Leyrer et al. [258] reported that they also succeeded in emulsion-polymerizing
SA in the presence of methyl-b-CD, and claimed that in this case the CD served
as a phase transfer agent, because the ability of the CD to form a water-soluble
complex with hydrophobic molecules made it easier for the SA molecules to
leave the monomer droplets and to be released from the complex after arriv-
ing at the surfaces of the growing polymer particles. The authors reported that
only 5 wt% of CD was necessary to polymerize almost 100% of SA.
Recently, Soares and Hamielec [252] presented a review article on the study
of transport phenomena in emulsion polymerization and introduced a case
study on how to increase the amount of ethylene (E) content in the copolymer
produced by the emulsion copolymerization of E and VAc under the conditions
of a mass-transfer-controlled polymerization rate. Zubitur et al. [255] studied
the effect of agitation on the batch and semicontinuous emulsion polymeriza-
tions of St in a reactor equipped with a four-paddle type stirrer and dodecyl
mercaptan as the CTA. Here, the CTA mass-transfer from the monomer droplets
to the aqueous phase was the rate-controlling step. They showed that the mole-
cular weights of polymers decreased as the stirring speed was increased because
of the improvement in the CTA mass transfer from the monomer droplets to the
aqueous phase due to the improved emulsification of the monomer droplets. For
semicontinuous polymerization, the instantaneous conversion increased as the
stirring speed was increased for speeds less than 150 rpm because the system
was monomer diffusion controlled, whereas at stirring speeds higher than
150 rpm, the agitation was strong enough for the polymerization rate to be
kinetically controlled. They [251] further studied the effect of agitation on the
monomer and CTA transport step from the monomer droplets to the polymer
particles in the semicontinuous emulsion polymerization of St and BA with
KPS as the initiator and NaLS as the emulsifier, respectively. Polymerizations
were carried out in a 2 dm3 glass reactor fitted with a stainless-steel anchor-
type stirrer. It was found that when neat monomer addition was used, a mild
degree of agitation (0.1 kW/m3) was needed to overcome monomer mass trans-
fer limitations. However, a moderate degree of agitation (0.3 kW/m3) was not
enough to avoid mass transfer limitations when dodecyl mercaptan (CTA) was
present. Preemulsification of the feed was used to minimize the mass transfer
80 M. Nomura et al.
limitations of both the monomer and the CTA, even for a gentle degree of
agitation (0.01 kW/m3). The molecular weights of the polymers produced de-
pended on the presence of the CTA. In the presence of the CTA, the molecular
weights decreased with the stirring speed, whereas they increased in the absence
of the CTA.
Salazar et al. [234] investigated how the molecular weight could be controlled
in a starved emulsion polymerization of St using tert-dodecyl mercaptan and
tert-nonyl mercaptan (more water-soluble) as the CTAs. The authors showed
that in a starved polymerization with tert-dodecyl mercaptan, a mass-transfer
resistance to the mercaptan was required to fit the observed PSt molecular
weights to the model predictions, but this extra mass-transfer resistance could
be neglected in the case of the more water-soluble tert-nonyl mercaptan. Cun-
ningham et al. [253, 254] investigated the seeded emulsion polymerization of St
in order to study the effects of n-dodecly mercaptan on the polymer molecu-
lar weight distribution. In the emulsion polymerization of St with n-dodecly
mercaptan as the CTA, the transport of the CTA from the monomer droplets
to the polymer particles is diffusion limited, meaning that it was difficult to
calculate molecular weights, except perhaps by using empirical approaches.
They developed a methodology that used the mass transfer model developed
by Nomura et al. [239], which allowed the CTA concentration within the poly-
mer particles to be determined, regardless of whether or not the CTA was at
its equilibrium value, and validated the essential correctness of the approach
by comparing experimental molecular weight distributions with the model’s
predictions. The authors further suggested that the methodology used might
be amenable to online applications. Mendoza et al. [125] carried out a study of
the kinetics of St emulsion polymerization using n-dodecly mercaptan as the
CTA. In this study, it was found that the CTA had no effect on the polymeriza-
tion rate, but had a substantial effect on the molecular weight distribution
(MWD). The efficiency of the CTA in reducing the MWD was lowered by the
mass transfer limitations. The process variables affecting the CTA mass-trans-
fer were also investigated. For example, the average molecular weight of the
polymers produced was found to decrease with increasing stirring speed. The
authors developed a mathematical model to predict monomer conversion, the
number of polymer particles, and the number-average molecular weights, and
then validated the proposed model by fitting it to the experimental data.
During an emulsion polymerization, one often encounters the formation
of coagulum, which is sometimes fatal for products such as paints. Moreover,
it may prevent the scale-up of commercially-acceptable latex. Therefore, the
formation of coagulum during an emulsion polymerization is an important
industrial problem and may be closely related to agitation. Although some
researchers [260–262] postulated that coagulum may form during emulsion
polymerization due to tangential or shear stresses, which originate in the re-
action mixture due to agitation, the literature has very little information on any
quantitative experimental data in this field. Vanderhoff [260] discussed this
problem and proposed two mechanisms for the formation of coagulum in
Emulsion Polymerization: Kinetic and Mechanistic Aspects 81
the emulsion polymerization process: (i) a failure of the stability of the latex,
giving rise to flocculation and growth of the aggregates to macroscopic size
(lumps), and (ii) a different mechanism of polymerization, for example poly-
merization in large monomer droplets or a separate monomer layer in the
vapor space above the latex and on the reactor surfaces. Lowry et al. [261] stud-
ied the phenomenon of shear-induced coagulation in emulsion polymerization
carried out in a stirred tank reactor. Here, the authors correlated the coagulum
formation for different emulsion polymerizations to various agitation parame-
ters. For a low Reynolds number, it was shown that the stirring speed is impor-
tant, whereas, for a high Reynolds number, power consumption is the important
parameter. Matejicek et al. [262] studied the influence of agitation on the
creation of coagulum during the semicontinuous emulsion terpolymerization
of St-BA-AA carried out in 25 dm3 and 5 m3 reactors, respectively, and gave the
relationship between the amount of coagulum formed and the intensity of agita-
tion. The authors found that the amount of coagulum formed (Y%) was cor-
related to the specific impeller power input ei introduced by agitation, showing
that the dependence of the coagulum content in the dispersion on impeller
speed passes through a minimum.
4
Kinetic Aspects in Polymer Structure Development
4.1
Molecular Weight Distribution (MWD)
4.1.1
Monte Carlo (MC) Simulation Method
Fig. 9 Schematic drawing illustrating the simulation method based on the competition tech-
nique
4.1.2
Instantaneous Molecular Weight Distribution
In this part, the instantaneous MWD formed over a very small time interval
is considered. This is equivalent to considering the distribution of polymer
84 M. Nomura et al.
chains formed in a certain fixed environment. Note that, because the polymer/
monomer ratio is kept approximately constant during Interval II, the instanta-
neous MWD may be a reasonable approximation for the linear polymers formed
during Interval II. This is not the case, however, for nonlinear polymer forma-
tion, as discussed later.
Assuming that the polymer particles have a uniform size, a single statistical
polymer particle that is representative of the whole population of particles can
be considered. This polymer particle is a kind of imaginary micro-reactor that
does not modify either the volume or the polymer/monomer ratio, even after
polymer chains are produced, so the reaction environment is kept constant
except that the number and chain length distribution of the macroradicals are
changed stochastically. By producing a large number of polymer molecules
consecutively with this imaginary polymer particle, the instantaneous MWD
can be determined.
After the nucleation period, three types of kinetic processes determine the
kinetics of emulsion polymerization: radical entry, radical desorption, and
polymer chain formation in the polymer particles. The kinetics of emulsion
polymerization are fully described by the following five dimensionless para-
meters:
1. Radical entry
Çe 1
e = 961 = 941 (73)
kp[M]pNT kp[M]p –
te
where Çe is the rate of radical entry into polymer particles, including the
reentry of desorped radicals, NT is the number of polymer particles, and –
te is
the average time interval between radical entry.
Assuming a random entry of radicals to all polymer particles, the imagi-
nary chain length Pe shown in Fig. 9 follows the most probable distribution,
and can be determined by using a random number between 0 and 1, y, as fol-
lows [273]:
Pe = (1/e) ln(1/y) (74)
2. Chain transfer
[CTA]p
Cf = Cm + CfCTA 93 (75)
[M]p
where Cm is the monomer transfer constant, and CfCTA is the constant of transfer
to chain transfer agents.
In the same way as in Eq. 74, the imaginary chain lengths Pf1 and Pf2 shown
in Fig. 9 can be determined by using a random number between 0 and 1, y, as
follows:
Pf = (1/Cf) ln(1/y) (76)
Emulsion Polymerization: Kinetic and Mechanistic Aspects 85
3. Bimolecular termination
2ktp
x = 93251 (77)
kp[M]p vpNA
where ktp is the bimolecular termination rate constant, in which the termina-
tion rate is represented by Rt=2ktp[R*]2, not Rt=ktp[R*]2 used in [263–273]. NA
is Avogadro’s number, and vp is the volume of a swollen polymer particle. Note
that the bimolecular termination rate depends on the particle size in emulsion
polymerization, and is larger for smaller polymer particles.
Assuming that the bimolecular terminations are independent of chain-
length, the imaginary chain length P12 shown in Fig. 9 can be determined by:
P12 = (1/x) ln(1/y) (78)
Note that Eq. 78 must be considered for all possible radical pairs.
4. Radical desorption
K0
d = 93 (79)
kp[M]p
where K0 is the desorption rate coefficient for an oligomeric radical.Any chain-
length-dependent radical desorption can be accounted for using the MC
method, in principle. However, it is often reasonable to assume that only
monomeric radicals can exit. For such cases, the average desorption rate coef-
ficient for all radicals in polymer particles, kf, which appears in the Smith-Ewart
equation [4] can be approximated by [43, 44, 122] kf @K0Cf. In this article, the
simulated results for such simple cases are shown.
If a chain transfer reaction is the actual event, as shown in Fig. 9, a mono-
meric or a CTA radical is formed. Neglecting the difference between mono-
meric and CTA radicals, the probability of radical exit when the chain transfer
reaction occurs is given by:
d
Pdes = 99961 (80)
1 + Cf + (n – 1) x + d
where n is the number of radicals in the polymer particle.
5. Type of bimolecular termination
ktp,c
jc = 99 (81)
ktp,c + ktp,d
where ktp,c and ktp,d are the bimolecular termination rate constants by combi-
nation and by disproportionation, respectively.
If bimolecular termination is the actual event, the probability that the
termination is by combination is equal to jc.
86 M. Nomura et al.
Fig. 10 Calculated number- and weight-average chain lengths as a function of the average
number of radicals in a polymer particle, –
n, with e =2¥10–4 and Cf =d =0
Emulsion Polymerization: Kinetic and Mechanistic Aspects 87
Fig. 11 Calculated chain length distribution on a number and weight basis, for –
n=0.500
88 M. Nomura et al.
Fig. 12 Instantaneous chain length distribution on a number and weight basis, where
dead chains are formed by disproportionation termination. The value of –
n is increased by
decreasing the bimolecular termination rate
Emulsion Polymerization: Kinetic and Mechanistic Aspects 89
4.1.3
Effect of Chain-Length-Dependent Bimolecular Termination
Fig. 13 Effect of particle size on the instantaneous chain length distribution, where bi-
molecular terminations are chain-length dependent and are by disproportionation
Emulsion Polymerization: Kinetic and Mechanistic Aspects 91
As the particle size increases, the bimolecular termination rate decreases. For
cases with chain-length independent bimolecular termination, the oligomeric
peak in the number fraction distribution moves toward larger chain length, as
shown in Fig. 12. On the other hand, if the bimolecular terminations are highly
diffusion controlled, the oligomeric peak location does not move, but the peak
height becomes smaller due to an increased amount of dead polymer formation
from chain transfer to the monomer, which is accounted for in this simulation.
Note that we should not expect these oligomeric peaks to be detected via usual
SEC analysis, represented on the basis of the weight fraction distribution, as
shown in the lower panel of Fig. 13. The termination mode may not be distin-
guished from the SEC data.
The chain-length-dependence of bimolecular termination reactions needs
to be taken into account in order to be able to accurately estimate the MWD
formed, except when very small polymer particles are formed and/or chain
transfer reactions dominate over dead polymer chain formation.
4.1.4
Accumulated Molecular Weight Distribution
Fig. 14 Accumulated weight fraction distribution development with and without terminal
double bond polymerization
average time interval between entries for radicals generated in the water phase
(excluding reentry of desorped radicals) is 50 s. The value of np is the total num-
ber of monomeric units bound into polymer chains in a polymer particle, and
therefore, np increases as the reaction proceeds. In the simulation, the poly-
mer/monomer ratio in the polymer particle is kept constant. Both with and
without TDBP, the MWD profiles do not change significantly during polymer-
ization (as long as monomer droplets exist); however, much larger polymer
molecules can be formed by the TDBP and this can change the MWD profile
significantly. In emulsion polymerization, the enhanced accidental branching
caused by a higher polymer concentration cannot be neglected, even for reac-
tion systems in which the branching reactions are not significant in the bulk
polymerization.
4.1.5
Determination of Monomer Transfer Constants from MWD
Fig. 15 Monte Carlo simulation results for emulsion polymerization that involves polymer
transfer reactions, under the conditions Cm=5¥10–5, e =5¥10–7 and Cfp=5¥10–5, without
radical desorption
4.2
Branched and Crosslinked Polymer Formation
4.2.1
Long-Chain Branched Polymers
4.2.1.1
Compartmentalization of Radicals
4.2.1.2
Higher Polymer Concentration Effects
Fig. 16 Development of the average branching density a and the branching density distri-
bution b for emulsion and bulk polymerizations. The conversion at which monomer droplets
disappear in emulsion polymerization is xc=0.5
the polymer concentration stays the same during the lifetimes of the monomer
droplets, the ratio of the branching and propagation reaction rates is kept
constant, resulting in a constant average branching density until the deple-
tion of the monomer droplets, as shown in Fig. 16a. Emulsion polymerizations
enhance the frequency of branching and crosslinking reactions due to the
higher polymer concentration associated with the existence of monomer
droplets.
Figure 16b shows the development of the branching density distribution,
where Çb(q,y) shows the expected branching density of the primary polymer
molecule born at conversion x=q, when the conversion at the present time
is x=y. The primary chains formed in the early stages of polymerization are
subjected to branching reactions for a longer period of time, and therefore, the
expected branching density is higher than those chains formed in the later
Emulsion Polymerization: Kinetic and Mechanistic Aspects 97
4.2.1.3
Limited Space Effects
Fig. 17 Hypothetical example that illustrates the differences between the polymer transfer
reactions that occur in bulk polymerization a and in emulsion polymerization b
98 M. Nomura et al.
If all of these species exist in the same reaction locus as in Fig. 17a, it would
be highly probable that all of the radicals would attack the largest chain. In
other words, the chain transfer rate of the polymer chain with chain length P,
vfp,Pp, is proportional to its chain length:
vfp,PP = kfp[R*] r[PP] (86)
where kfp is the rate constant for chain transfer to polymer, [R·] is the total
radical concentration, and [PP] is the concentration of polymer molecules with
chain length P in the whole reaction mixture.
On the other hand, suppose that each of these polymer molecules is isolated
into different particles, and that each particle contains one radical, as shown
in Fig. 17b. If the radical causes the polymer transfer reaction, the partner
must be the polymer molecule that happens to exist in the same particle (so it
cannot partner a larger polymer molecule that exists in a different polymer
particle).As a consequence, the expected size of the polymer molecule attacked
by a radical is smaller for emulsion systems than for the homogeneous model
shown in Fig. 17a.
Equation 86 is commonly used for homogeneous reaction systems, but it is
not exact in emulsion polymerization. The value of [PP] is different for each
polymer particle, and the value obtained for [PP] when all of the particles are
combined cannot be used either. Strictly, one needs to determine a discrete
distribution function of polymer molecules in each polymer particle.
Figure 18 shows the simulated MWD profiles that clearly demonstrate the
effects of limited space [273].
Because the total number of monomeric units bound into polymer molecules
is np=4¥105 for the present calculation, the high molecular tail can never exceed
4¥105. Figure 18 shows that the model that does not account for the effects of
Fig. 18 Comparison of the calculated weight fraction distribution with Cf=Cfp=5¥10–4 and
xc=0.5. For the emulsion polymerization model, the total number of polymerized monomeric
units in a polymer particle np=4¥105, which is equal to the size of a dried polymer particle
Emulsion Polymerization: Kinetic and Mechanistic Aspects 99
4.2.1.4
Formation of the Bimodal Molecular Weight Distribution
xc [CTA]p 1
t = Cm + Cfp 93 + CfCTA + 75 + 753 (88)
(1 – xc) [M]p kp[M]p–
te
Cfp xc /(1 – xc )
Pb = 9999999998 (89)
Cm + Cfp xc /(1 – xc ) + CfCTA[CTA]p /[M]p + 1/(kp[M]p–
te)
A model analysis was conducted, which assumed that (1) both t and Pb do not
change during polymerization, (2) all polymer particles are formed instanta-
neously, and (3) the number of primary chains in a polymer particle, npc, is
the same for all particles. Important conclusions were that (i) bimodal MWDs
(represented in terms of W(logP)) are formed if Pb is larger than 0.5, and (ii) for
Pb>0.5, the weight-average molecular weight increases without limits over the
whole course of polymerization. Note that the second conclusion does not
indicate that gelation occurs. Figure 21 shows the calculated development of
the weight-average chain length [311].
For Pb<0.5, the weight-average chain length reaches a constant value that
– – –
is given by Pw = Pwp /(1 – 2Pb), where Pwp is the weight-average chain length of
the primary chains. On the other hand, the weight-average molecular weight
increases without limit for Pb>0.5, but it increases very gradually and it takes
an infinitely long time to reach an infinitely large polymer molecule. It was
found [313] that the formed MWD is a power-law distribution [312] that pos-
sesses fractal characteristics is formed.
For the emulsion-polymerized polyethylene shown in Fig. 20, which clearly
shows a bimodal MWD, Pb=0.813>0.5. For another example investigated in
[311], Pb=0.711>0.5, and the MWD is bimodal.
On the other hand, for the emulsion polymerization of vinyl acetate, Friis
et al. reported that (i) the weight-average molecular weight does not increase
significantly until the monomer droplets are depleted, and that (ii) the MWDs
are unimodal [314, 315]. They considered the TDBP in addition to the polymer
transfer reactions; however, the contribution of the TDBP is minor and a qual-
itative discussion on the MWD shape could be made without needing to involve
the TDBP. The parameters they used are Cm=2.32¥10–4, Cfp=3.98¥10–4 and
xc=0.2, which gives Pb=0.3.With Pb=0.3, the theoretical analysis in [311] showed
102
Fig. 21 Calculated development of the weight-average chain length during the model emulsion polymerization
M. Nomura et al.
Emulsion Polymerization: Kinetic and Mechanistic Aspects 103
that (i) the weight-average chain length reaches a steady state value rather
quickly, and that (ii) the MWDs formed are unimodal, both of which agree with
experimental observations.
Although the kinetic behavior during Interval III was not considered in
[311], Pb=0.5 is an important consideration when we look at the possibility of
forming bimodal MWDs in emulsion polymerization that involves chain trans-
fer to polymer.
4.2.2
Crosslinked Polymers
4.2.2.1
Crosslinked Structure
Assuming that classical chemical kinetics are valid and that the crosslinking
reaction rate is proportional to the concentrations of polymer radicals and
pendant double bonds, it was shown theoretically that the crosslinked polymer
formation in emulsion polymerization differs significantly from that in cor-
responding bulk systems [270, 316]. To simplify the discussion, it is assumed
here that the comonomer composition in the polymer particles is the same as
the overall composition in the reactor, and that the weight fraction of polymer
in the polymer particle is constant as long as the monomer droplets exist. These
conditions may be considered a reasonable approximation to many systems, as
shown both theoretically [316] and experimentally [271, 317]. First, consider
Flory’s simplifying assumptions for vinyl/divinyl copolymerization [318]; that:
(1) the reactivities of all types of double bonds are equal, (2) all double bonds
Fig. 22 Crosslinking density distribution development during bulk and emulsion poly-
merization under Flory’s simplifying assumptions, where the initial mole fraction of divinyl
monomer is 0.01, and xc=0.4 for emulsion polymerization
104 M. Nomura et al.
react independently, and (3) there are no cyclization reactions in finite mole-
cules. Under these simplifying assumptions, a completely homogeneous net-
work is formed in homogeneous batch polymerization [319].
Figure 22 shows the calculated results for the expected crosslinking density
of primary chains formed at conversion q when the conversion at the present
time is y, under Flory’s simplifying assumptions for bulk and emulsion copoly-
merization.
The expected crosslinking density is the same for all primary chains at any
stage of polymerization in a bulk system, indicating that a statistically homo-
geneous network is formed. On the other hand, the crosslinking density dis-
tribution is completely different in emulsion polymerization. In particular,
the crosslinking densities of the primary chains formed in the earlier stages
of polymerization are very high, and the variance of the crosslinking density
distribution is significant in emulsion polymerization. The bending at q=0.4 oc-
curs because the conversion at which monomer droplets disappear is assumed
to be 0.4 in these calculations. Homogeneous networks cannot be formed, even
under Flory’s simplifying assumptions in emulsion crosslinking copolymer-
ization.
Figure 23 shows the average crosslinking density development under Flory’s
simplifying assumptions.
The average crosslinking density is higher for emulsion polymerization due
to the higher polymer concentration effects inside the polymer particles, which
are the loci of polymerization. The fact that the average crosslinking density
is high even from a very early stage of polymerization has been confirmed
experimentally, and this agrees with theoretical calculations adequately [271,
272, 317]. The fact that the average crosslinking densities in some emulsion
polymerizations may not change much during emulsion polymerization, as
shown in Fig. 23, is sometimes misunderstood as producing a homogeneous
polymer network [320]. However, this argument may not be true, as shown in
the crosslinking density distribution profile shown in Fig. 22.
4.2.2.2
Unique Molecular Weight Distributions
Fig. 24 Experimental results for the weight-average molecular weight development during
the emulsion crosslinking copolymerization of styrene (St)/divinylbenzene (DVB) and
methyl methacrylate (MMA)/ethylene glycol dimethacrylate (EGDMA). Data from [323]
106 M. Nomura et al.
using SEC even when each polymer particle forms a single crosslinked polymer
molecule.
In a typical recipe for microgel formation, a sufficient amount of crosslinker
is used. Because the crosslinking density tends to be high from a very early
stage of polymerization, the MC simulation results showed [270, 271] that each
polymer particle tends to consist of a single large crosslinked polymer mole-
cule even from the early stage of polymerization. In such cases, it was shown
[270, 271] that (1) the weight-average molecular weight grows linearly with con-
version, and (2) the MWD formed is narrow and the distribution shifts to larger
molecular weights, preserving the narrow profile, as polymerization proceeds.
Experimental data [323–325] in which the MWDs are measured by SEC with
on-line multiangular laser light scattering (MALLS) show the above trend [326].
Figure 24 shows some examples of the weight-average molecular weight de-
velopment, which show linear increases with respect to conversion.
The y-intercept might approximately correspond to the weight-average
molecular weight of primary chains. It was further confirmed [326] that the
MWD of polymer molecules and that of whole polymer particles are the same,
showing that each polymer particle essentially consists of one crosslinked
polymer molecule. Figure 25 shows such an example, and the MWD formed
agrees reasonably well with the particle size distribution represented in terms
of molecular weights.
On the basis of the MC simulation results [270], it is expected that if the
amount of divinyl monomer is reduced to the level of, say, several crosslinkers
per primary chain, a bimodal MWD (as shown in Fig. 26) may result.
The arrows indicate the molecular weight of a dried polymer particle, which
gives an upper limit to the molecular weight of the polymers attainable. Qual-
itatively speaking, the high molecular weight peaks are formed because the
crosslinked polymer molecules want to grow further due to the higher chain
Fig. 25 MWD of polymer molecules and dried polymer particles in the emulsion copoly-
merization of St and DVB (20 mol%) [326]
Emulsion Polymerization: Kinetic and Mechanistic Aspects 107
Fig. 26 MC simulation results that show bimodal MWDs in the emulsion copolymerization
of vinyl/divinyl monomers [270]
connectivity, but they cannot because of the limitation of a small particle size.
Experimentally, such unique bimodal distributions were obtained in the
styrene/divinyl benzene system [327] by setting the number of divinylbenzene
molecules per chain to ~5. An example of a bimodal MWD, together with the
particle size distribution [327], is shown in Fig. 27.
It was found that the locations of these peaks can be controlled indepen-
dently. The location of a high molecular weight peak is mainly controlled by the
particle size, and the location of a low molecular weight peak is controlled by
the chain lengths of the primary polymer molecules.
The MC simulation method is particularly suitable for investigating emul-
sion polymerization that involves various simultaneous kinetic events with a
very small locus of polymerization. The MC simulation method will become a
standard mathematical tool for the analysis of complex reaction kinetics, both
for linear and nonlinear emulsion (co)polymerization.
5
Continuous Emulsion Polymerization
where ap is the surface area of a polymer particle and E(t) is the residence time
distribution function. Combining Eqs. 91 and 92 gives
Çwq
NT = 99984 (93)
1 + (4.36 Çw m 2/3q 5/3asS0)
Thus, the rate of polymerization Rp can be predicted using Eqs. 1, 51 and 93.
Omi et al. [335] and Nomura et al. [163] pointed out that Eq. 93 suggests the
existence of a maximum number of polymer particles NT,max at the optimum
residence time qmax, which is given by
qmax = 0.53(asS0/Çwm 2/3)3/5 (94)
NT,max = 0.21(Çw /m )0.4(aSS0)0.6 (95)
110 M. Nomura et al.
By comparing Eq. 31 with Eq. 95, they also pointed out that NT,max is only 58%
of that produced in a batch operation with the same recipe and temperature,
and also that
qmax = 0.83tc (96)
where tc is the time at which particle formation stops due to complete deple-
tion of micelles in the water phase, in a batch operation with the same recipe
as that used in a continuous operation.
Gershberg and Longfield [330] carried out the continuous emulsion poly-
merization of St at 70 and 100 °C in a train of three CSTRs, and compared the
experimental results with theoretical predictions. They presented the following
conclusions and suggestions;
1. The theoretical equation NTµRpµÇw0S 01.0q–2/3, derived from Eqs. 1 and 93,
correctly predicts the effect of the operating variables upon the rate of poly-
merization in the stirred-tank reactors at any stage at 70 °C for large q
values.
2. The number of polymer particles produced at 70 °C (and hence, the rate of
polymerization) was higher than that at 100 °C, as Eq. 93 predicted.
3. The theory and the experimental data from this study demonstrates that in
a train of CSTRs, essentially all of the particles form in the first reactor.
Therefore, it is possible to maximize the monomer conversion in the latex
leaving the first reactor by keeping the temperature and the residence time
at the first reactor as low as possible in order to produce the maximum num-
ber of polymer particles and so increase the rate of polymerization in the
succeeding stages. This is the so-called “pre-reactor” concept.
4. NT,max certainly occurs at low q, as the theory predicts.
5. Sustained oscillation can take place in the monomer conversion.
6. The residence time in the first stage should be long enough to overcome the
retarding effects of traces of oxygen and/or impurities in the feed-stream
upon particle formation and growth.
Poehlein and Degraff [336] extended the derivation of Gershberg and Long-
field [330] to the calculation of both molecular weight and particle size distri-
bution in the continuous emulsion polymerization of St in a CSTR. On the
other hand, Nomura et al. [163] carried out the continuous emulsion poly-
merization of St in a cascade of two CSTRs and developed a novel model for the
system by incorporating their batch model [14], which introduced the concept
that the radical capture efficiency of a micelle relative to a polymer particle
was much lower than that predicted by the diffusion entry model (Ç µd1.0).
The assumptions employed were almost the same as those of Smith and
Ewart (Sect. 3.3), except that the model did not assume a constant value of m.
The elementary reactions and their rate expressions employed in the first stage
are as follows:
Emulsion Polymerization: Kinetic and Mechanistic Aspects 111
Fig. 28 Effect of the mean residence time in the first reactor q on the number of polymer
particles produced (S0 (NaLS)=12.5 g/dm3-water, I0 (KPS)=1.25 g/dm3-water, M0 (St)=
500 g/dm3-water; 50 °C. Experimental data: empty circles, first reactor; filled circles, second
reactor
model, but for MA only the particle number varied in the manner predicted by
the Gershberg and Longfield model. This showed that a relatively water-solu-
ble monomer does not correlate with the model based on S-E theory. The
existence of multiple steady-states for the isothermal operation of continuous
emulsion polymerization in a CSTR was first demonstrated by Gerrens et al.
[338, 339]. It is now well understood from a kinetic point of view that this
phenomenon takes place as a consequence of the so-called “gel-effect” (the
Trommsdorff-Norrish effect).
Oscillations in the number of polymer particles, the monomer conversion,
and the molecular weight of the polymers produced, which are mainly observed
in a CSTR, have attracted considerable interest. Therefore, many experimental
and theoretical studies dealing with these oscillations have been published
[328]. Recently, Nomura et al. [340] conducted an extensive experimental study
on the oscillatory behavior of the continuous emulsion polymerization of VAc
in a single CSTR. Several researchers have proposed mathematical models
that quantitatively describe complete kinetics, including oscillatory behavior
[341–343]. Tauer and Müller [344] proposed a simple mathematical model for
the continuous emulsion polymerization of VCl to explain the sustained oscil-
lations observed. Their numerical analysis showed that the oscillations depend
on the rates of particle growth and coalescence. However, it still seems to be
difficult to quantitatively describe the kinetic behavior (including oscillations)
of the continuous emulsion polymerization of monomers, especially those with
relatively high solubility in water. This is mainly because the kinetics and mech-
Emulsion Polymerization: Kinetic and Mechanistic Aspects 113
Fig. 29 Taylor vortices between two concentric cylinders (Couette-Taylor vortices): inner
cylinder rotating, outer cylinder at rest
116 M. Nomura et al.
wbRi b
Ta = 9 4 (100)
v Ri
where Ri is the inner cylinder radius, b is the radial clearance between two con-
centric cylinders, n is the kinematic viscosity, and w is the angular velocity of
the inner cylinder.When the Taylor number exceeds a certain value between 46
and 60, called the critical Taylor number, Tac, a transition occurs from pure
Couette flow to a flow regime in which toroidal vortices are regularly spaced
along the cylinder axis, as shown in Fig. 29, which is the so-called “Couette-Tay-
lor vortex flow”. They adopted the tank-in-series model that has only one
parameter, N, the number of tanks in series, to characterize the deviation of the
flow in the CCTVFR from plug flow. The relationship between the number
of tanks N and the Taylor number Ta was determined by using a stimulus-
response method. Figure 30 shows an interesting example of the monomer con-
version versus reaction time, where the steady-state monomer conversion was
increased from 27 to 35% simply by decreasing the rotational speed of the
inner cylinder from 145 to 30 rpm.
This can be explained by the fact that the flow in the CCTVFR became closer
to plug flow as the Taylor number was dropped closer to Tac . Therefore, the
steady-state particle number and the steady-state monomer conversion could
be arbitrarily varied by simply varying the rotational speed of the inner cylin-
der. Moreover, no oscillations were observed, and the rotational speed of the
inner cylinder could be kept low, so that the possibility of shear-induced co-
agulation could be decreased. Therefore, a CCTVFR with these characteristics
is considered to be highly suitable as a pre-reactor for a continuous emulsion
polymerization process. In the case of the continuous emulsion polymerization
of VAc carried out with the same CCTVFR, however, the situation was quite
different [365]. Oscillations in monomer conversion were observed, and almost
no appreciable increase in steady-state monomer conversion occurred even
when the rotational speed of the inner cylinder was decreased to a value close
to Tac . Why the kinetic behavior with VAc is so different to that with St cannot
be explained at present.
Ohmura et al. [361, 362] carried out the continuous emulsion polymeriza-
tion of St and of VAc in a single CCTVFR from the standpoint of controlling
the particle size distribution of the latex produced. They also observed no
oscillations in the monomer conversion for the St system, but did observe
them in the system with VAc. They concluded that a self-sustained oscillation
would be useful for controlling the size of latex particles and for raising the
monomer conversion. On the other hand, Schmidt et al. [363] studied the con-
tinuous emulsion polymerization of n-BA in a single CCTVFR. The special
flow pattern and residence time distributions of the reactor were investigated
at different flow conditions. Both the hydrodynamics and the monomer con-
version of the CCTVFR were modeled using computational fluid dynamics
simulations. The model predictions were in good agreement with experi-
mental results.
In order to increase the performance and the productivity of the whole
continuous reactor system, it has been suggested that a pre-reactor operating
as a seed generator (a seeding reactor) should be placed upstream of a reactor
train. Even when it is optimized to have the mean residence time as qmax, the
steady-state number of polymer particles in a continuous stirred-tank pre-re-
actor, reaches only 58% of the number produced in an ideal plug flow reactor
or in a batch reactor using the same recipe and temperature conditions [163,
335]. Therefore, it is generally desirable to place a pre-reactor with plug flow
behavior (such as tubular, PSPC, PPC and CTVF reactors) upstream of the main
reactors, because this can produce a similar number of polymer particles to
that produced in a batch reactor if properly operated. Instead of incorporating
a pre-reactor, Nomura et al. [331, 332] have devised an operational technique
called “split-feed operation”, by which the number of polymer particles pro-
duced could be increased to much higher than that in a batch reactor under
118 M. Nomura et al.
the same recipe conditions. Figure 31 indicates the general concept of the split-
feed operation.
In a batch emulsion polymerization of St, they found that the number of
polymer particles produced increased when the amount of initially charged
monomer was decreased below a critical value, Mc, as shown in Fig. 32, where
the solid line shows the values predicted by
NT = (a3sÇ2p /36p)S03 M 0–2I 00 (101)
where Çp is the polymer density, and S0, M0 and I0 are the emulsifier, monomer
and initiator concentrations, respectively.
When a monomer split-feed operation based on the experimental result
shown in Fig. 32 was applied, for example, to a continuous tubular pre-reactor
with some backmixing, the number of polymer particles increased by about
30% at MF1=0.02 g/cm3-water, compared to the number produced in a batch re-
actor, as shown in Fig. 33.
The monomer split-feed operation was also shown to work in a continuous
stirred-tank pre-reactor. When certain conditions are fulfilled, the split-feeds
Fig. 33 Effect of monomer concentration MF1 fed to the first tubular seeding reactor with
back mixing on the number of polymer particles produced (S0 (NaLS)=6.25 g/dm3-water,
I0 (KPS)=1.25 g/dm3-water, MF1 (St)=variable; 50 °C. Experimental data: empty circles, par-
ticle number observed at t=40 min in a batch reactor; filled circles, steady-state particle num-
ber observed in the first tubular seeding reactor operated with mean residence time
t =40 min)
of emulsifier, initiator and water could effectively increase the number of poly-
mer particles produced. The split-feed operation has additional advantages in
that the volume of a pre-reactor can be made smaller, and oscillations can be
eliminated when it is applied to a continuous stirred-tank pre-reactor. Penlides
et al. [333] have also discussed the advantages of the split-feed operation.
Despite the industrial importance, very little work on the kinetics of con-
tinuous emulsion copolymerization has been reported. Poehlein et al. [366, 367]
carried out the continuous emulsion copolymerization of St-AM and St-AN us-
ing a continuous reactor system comprised of a tubular reactor followed by one
or two stirred-tank reactors, and developed a steady-state model by employing
the kinetic model proposed by Nomura et al. [14] for batch emulsion copoly-
merization. Their model could predict the latex particle size distribution,
average number of radicals per particle, rate of copolymerization, and copoly-
mer composition. They also investigated the continuous emulsion copolymer-
ization of a moderately water-soluble monomer, ethyl acrylate, with a com-
pletely water-soluble monomer, methacrylic acid. Several continuous processes
involving a tubular reactor and/or a CSTR were designed and utilized so as to
produce a latex product with properties similar to the batch product [368, 369].
Nomura et al. [370] carried out the continuous emulsion copolymerization of
a sparingly water-soluble monomer, St, with a moderately water-soluble
monomer, MMA, in a single CSTR, in order to experimentally elucidate how the
kinetic behavior of the continuous emulsion homopolymerization of a spar-
ingly water-soluble monomer changes when it is copolymerized with a mod-
erately water-soluble monomer.
120 M. Nomura et al.
6
Concluding Remarks
The kinetics and mechanisms of particle growth and polymer structure de-
velopment are comparatively well understood compared to those of particle
nucleation. Therefore, the rate of polymerization and the properties of the poly-
mer produced can be (roughly) estimated as long as the number of polymer
particles produced is known (for example, in seeded emulsion polymerization).
However, the prediction of the number of polymer particles produced is still
far from being an established technique. Therefore, further efforts are needed
to qualitatively and quantitatively clarify the effects of numerous factors that
affect the process of particle formation in order to gain a more quantitative
understanding of emulsion polymerization.
References
73. Sütterlin N (1980) In: Fitch RM (ed) Polymer colloids II. Plenum, New York, p 583
74. Nomura M, Fujita K (1994) Polym React Eng 2:317
75. Nomura M, Kodani T, Ojima J, Kihara Y, Fujita K (1998) J Polym Sci Pol Chem 36:1919
76. Morrison BR, Maxwell IA, Gilbert RG, Napper DH (1992) In: Daniels ES, Sudol ED,
El-Aasser MS (eds) Polymer Latexes. ACS Symp Ser. 492. Am Chem Soc, Washington
DC, p 28
77. Sajjadi S, Brooks BW (1999) J Polym Sci Pol Chem 37:3957
78. Nomura M, Sakai H, Kihara Y, Fujita K (2002) J Polym Sci 40:1275
79. Yuan X-Y, Dimonie VL, Sudol ED, Roberts JE, El-Aasser MS (2002) Macromolecules
35:8356
80. Herrera-Ordóñez J, Olayo R (2000) J Polym Sci Pol Chem 38:2219
81. Herrera-Ordóñez J, Olayo R (2001) J Polym Sci Pol Chem 39:2547
82. Varela de la Rosa L, Sudol ED, EL-Aasser MS, Klein A (1996) J Polym Sci Pol Chem
34:461
83. Varela de la Rosa L, Sudol ED, EI-Aasser MS, Klein A (1999) J Polym Sci Pol Chem
37:4066
84. Varela de la Rosa L, Sudol ED, EI-Aasser MS, Klein A (1999) J Polym Sci Pol Chem
37:4073
85. Nomura M, Satpathy US, Kouno Y, Fujita K (1988) J Polym Sci Pol Lett 26:385
86. Nomura M, Takahashi K, Fujita K (1990) Makromol Chem–M Symp 35/36:13
87. Chern C-S, Lin C-H (1998) Polymer 40:139
88. Chern C-S, Lin C-H (2000) Polymer 41:4473
89. Wang Z, Paine AJ, Rudin A (1995) J Polym Sci Pol Chem 33:1597
90. Morrison BR, Gilbert RG (1995) Macromol Symp 92:13
91. Coen EM, Gilbert RG, Morrison BR, Leube H, Peach S (1998) Polymer 39:7099
92. Prescott SW, Fellows CM, Gilbert RG (2002) Macromol Theor Simul 11:163
93. Butucea V, Sarbu A, Georgescu C (1998) Angew Makromol Chem 255:37
94. Cheong I-W, Kim J-H (1998) Macromol Theor Simul 7:49
95. Sajjadi S (2000) J Polym Sci Pol Chem 38:3612
96. Sajjadi S (2001) J Polym Sci Pol Chem 39:3940
97. Nomura M, Fujita K (1993) Polym Int 30:483
98. Sajjadi S (2002) J Polym Sci Pol Chem 40:1652
99. Ozdeger E, Sudol ED, El-Aasser MS, Klein A (1997) J Polym Sci Pol Chem 35:3813
100. Ozdeger E, Sudol ED, El-Aasser MS, Klein A (1997) J Polym Sci Pol Chem 35:3827
101. Ozdeger E, Sudol ED, El-Aasser MS, Klein A (1997) J Polym Sci Pol Chem 35:3837
102. Lin S-Y, Capek I, Hsu T-J, Chern C-S (1999) J Polym Sci Pol Chem 37:4422
103. Unzueta E, Forcada J (1995) Polymer 36:4301
104. Chen L-J, Lin S-Y, Chern C-S, Wu S-C (1997) Colloid Surface A 122:161
105. Lin S-Y, Capek I, Hsu T-J, Chern C-S (2000) Polym J 32:932
106. Colombie D, Sudol ED, El-Aasser MS (2000) Macromolecules 33:7283
107. Asua JM, Schoonbrood HAS (1998) Acta Polym 49:671
108. Amalvy JI, Unzue MJ, Schoonbrood HAS, Asua JM (1998) Macromolecules 31:5631
109. Wang X, Sudol ED, El-Aasser MS (2001) J Polym Sci Pol Chem 39:3093
110. Wang X, Sudol ED, El-Aasser MS (2001) Macromolecules 34:7715
111. Cochin D, Laschewsky A, Nallet F (1997) Macromolecules 30:2278
112. Ayoub MMH, Nasr HE, Rozik NN (1998) J Macromol Sci Pure 35:1415
113. Ayoub MMH (1998) J Elastom Plast 30:207
114. Wang X, Sudol ED, El-Aasser MS (2001) Langmuir 17:6865
115. Kato S, Nomura M (1999) Colloid Surface A 153:127
116. Cheong I-W, Nomura M, Kim J-H (2001) Macromol Chem Phys 202:2454
Emulsion Polymerization: Kinetic and Mechanistic Aspects 123
117. Piirma I (1993) Polymer surfactants (Surfactant Science Series No 42). Marcel Dekker,
New York
118. Capek I (2002) Adv Colloid Interfac 99:77
119. O’Toole JT (1965) J Appl Polym Sci 9:1291
120. Ugelstad J, Mørk PC, Aasen JO (1967) J Polym Sci A1 5:2281
121. Stockmayer WH (1957) J Polym Sci 24:314
122. (a) Nomura M, Fujita K (1985) Makromol Chem Suppl 10/11:25, (b) Nomura M (1987)
Kobunshi Kagaku 36:680
123. Sakai H, Kihara Y, Fujita K, Kodani T, Nomura M (2001) J Polym Sci Pol Chem 39:1005
124. Asua JM, Adams ME, Sudol ED (1990) J Appl Polym Sci 39:1183
125. Mendoza J, de la Cal JC, Asua JM (2000) J Polym Sci Pol Chem 38:4490
126. Kiparissides C, Achilias CDS, Frantzikinakis CE (2002) Ind Eng Chem Res 41:3097
127. de la Cal JC, Asua JM (2001) J Polym Sci Pol Chem l39:585
128. Gilmore CM, Poehlein GW, Schork FJ (1993) J Appl Polym Sci 48:1449
129. Gilmore CM, Poehlein GW, Schork FJ (1993) J Appl Polym Sci 48:1461
130. Budhlall BM, Sudol ED, Dimonie VL, Klein A, El-Aasser MS (2001) J Polym Sci Pol Chem
39:3633
131. Shaffie KA, Moustafa AB, Mohamed ES, Badran AS (1997) J Polym Sci Pol Chem 35:3141
132. Bruyn HD, Gilbert RG, Ballard MJ (1996) Macromolecules 29:8666
133. Chern C-S, Poehlein GW (1987) J Appl Polym Sci 33:2117
134. Bruyn HD, Miller CM, Bassett DR, Gilbert RG (2002) Macromolecules 35:8371
135. Matsumoto A, Kodama K, Aota H, Capek I (1999) Eur Polym J 35:1509
136. Ballard MJ, Napper DH, Gilbert RG (1981) J Polym Sci Pol Chem 19:939
137. Nomura M, Kubo M, Fujita K (1981) Mem Fac Eng Fukui Univ 29:167
138. Storti G, Carra S, Morbidelli M, Vita G (1989) J Appl Polym Sci 37:2443
139. Tobita H, Hamielec AE (1989) Macromolecules 22:3098
140. Nomura M, Horie I, Kubo M, Fujita K (1989) J Appl Polym Sci 37:1029
141. Chen S-A, Wu K-W (1988) J Polym Sci Pol Chem 26:1487
142. Giannetti E (1989) Macromolecules 22:2094
143. Saldivar E, Ray WH (1997) Ind Eng Chem Res 36:1322
144. Dube MA, Penlidis A, Mutha RK, Cluett WR (1996) Ind Eng Chem Res 35:4434
145. Martinet F, Guillot J (1999) J Appl Polym Sci 72:1627
146. Vicente M, Leiza JR, Asua JM (2001) AIChE J 47:1594
147. Vega JR, Gugliotta LM, Meira GR (2002) Polym React Eng 10:59
148. Shoaf GL, Poehlein GW (1991) J Appl Polym Sci 42:1213
149. Yang B-Z, Chen L-W, Chiu W-Y (1997) Polym J 29:744
150. Yang B-Z, Chen L-W, Chiu W-Y (1997) Polym J 29:737
151. Slawinski M, Schellekens MAJ, Meuldijk J, Herk AMV, German AL (2000) J Appl Polym
Sci 76:1186
152. Wang PH, Pan C-Y (2001) Colloid Polym Sci 279:98
153. Santos AMD, Mckenna TF, Guillot J (1997) J Appl Polym Sci 65:2343
154. Yan C, Cheng S, Feng L (1999) J Polym Sci Pol Chem 37:2649
155. Henton DE, Powell C, Reim RE (1997) J Appl Polym Sci 64:591
156. Xu Z, Yi C, Cheng S, Zhang J (1997) J Appl Polym Sci 66:1
157. Fang S-J, Fujimoto K, Kondo S, Shiraki K, Kawaguchi H (2000) Colloid Polym Sci
278:864
158. Kostov GK, Petrov PC (1994) J Polym Sci Pol Chem 32:2229
159. Petrov PC, Kostov GK (1994) J Polym Sci Pol Chem 32:2235
160. Noel LFJ, Altveer JLV, Timmermans MDF, German AL (1996) J Polym Sci Pol Chem
34:1763
124 M. Nomura et al.
245. Omi S, Shiraishi Y, Sato H, Kubota H (1969) J Chem Eng Jpn 2:64
246. Weerts PA, van der Loos JLM, German AL (1991) Makromol Chem 192:1993
247. Arai K, Arai M, Iwasaki S, Saito S (1981) J Polym Sci Pol Chem 19:1203
248. Kim CU, Lee JM, Ihm SK (1999) J Appl Polym Sci 73:777
249. Ozdeger E, Sudol ED, El-Aasser MS, Klein A (1998) J Appl Polym Sci 69:2277
250. Brooks BW (1971) Brit Polym J 3:269
251. Zubitur M, Asua JM (2001) J Appl Polym Sci 80:841
252. Soares JBP, Hamielec AE (1997) In: Asua JM (ed) Polymeric dispersions (NATO ASI Ser
E:335). Kluwer Academic, London, p 289
253. Cunningham MF, Ma JW (2000) J Appl Polym Sci 78:217
254. Ma JW, Cunningham MF (2000) Macromol Symp 150:85
255. Zubitur M, Mendoza J, de la Cal JC, Asua JM (2000) Macromol Symp 150:13
256. Rimmer S, Tattersall P (1999) Polymer 40:5729
257. Rimmer S (2000) Macromol Symp 150:149
258. Leyrer RJ, Machtle W (2000) Macromol Chem Phys 201:1235
259. Lau W (1994) US Patent 5 521 266, assigned to Rohm and Haas Co, Philadelphia, PA, USA
260. Vanderhoff JW (1981) ACS Sym Ser 165:199
261. Lowry V, El-Aasser MS, Vanderhoff JW, Klein A (1984) J Appl Polym Sci 29:3925
262. Matejicek A, Ditl P, Pivonkova A, Kaska J, Formanek L (1988) J Appl Polym Sci
35:583
263. Tobita H, Takada Y, Nomura M (1994) Macromolecules 27:3804
264. Tobita H, Takada Y, Nomura M (1995) J Polym Sci Pol Phys 33:441
265. Tobita H (1995) Macromolecules 28:5128
266. Tobita H (1994) Polymer 35:3023
267. Tobita H (1994) Polymer 35:3032
268. Tobita H (1997) J Polym Sci Pol Phys 35:1515
269. Tobita H, Nomura M (1999) Colloid Surface A 153:119
270. Tobita H, Yamamoto K (1994) Macromolecules 27:3389
271. Tobita H, Uemura Y (1996) J Polym Sci Pol Phys 34:1403
272. Tobita H, Yoshihara Y (1996) J Polym Sci Pol Phys 34:1415
273. Tobita H (1995) Acta Polym 46:185
274. Min KW, Ray HW (1974) J Macromol Sci R M C 11:177
275. Lin CC, Chiu WY (1979) J Appl Polym Sci 23:2049
276. Lichti G, Gilbert RG, Napper DH (1980) J Polym Sci Pol Chem 18:1297
277. Lichti G, Gilbert RG, Napper DH (1982) In: Piirma I (ed) Emulsion polymerization.
Academic, New York, p 93
278. Katz S, Shinnar R, Saidel GM (1969) Adv Chem Ser 91:145
279. Giannetti E, Storti G, Morbidelli M (1988) J Polym Sci Pol Chem 26:1985
280. Storti G, Polotti G, Cociani M, Morbidelli M (1992) J Polym Sci Pol Chem 30:731
281. Ghielmi A, Storti G, Morbidelli M (1998) Macromolecules 31:7172
282. Benson SW, North AM (1962) Am Chem Soc 84:935
283. Ito K (1974) J Polym Sci Pol Chem 12:1991
284. Mahabadi HK (1985) Macromolecules 18:1319
285. Olaj OF, Zifferer G, Gleizner G (1986) Makromol Chem 187:977
286. Russell GT, Gilbert RG, Napper DH (1992) Macromolecules 25:2459
287. O’Shaughnessy B, Yu J (1994) Macromolecules 27:5067
288. Buback M, Egorov M, Kaminsky V (1999) Macromol Theor Simul 8:520
289. Russell GT (1994) Macromol Theor Simul 3:439
290. Clay PA, Gilbert RG (1995) Macromolecules 28:552
291. Wulkow M (1996) Macromol Theor Simul 5:393
Emulsion Polymerization: Kinetic and Mechanistic Aspects 127
332. Nomura M (1986) In: Reichert K-H, Geisler W (eds) Polymer eeaction engineering.
Hüthig & Wepf, Basel, p 41
333. Penlidis A, MacGregor JF, Hamielec AE (1989) Chem Eng Sci 44:273
334. Poehlein GW, Dougherty DJ (1977) Rubber Chem Technol 50:601
335. Omi S, Ueda T, Kubota H (1969) J Chem Eng Jpn 2:193
336. Degraff AW, Poehlein GW (1971) J Polym Sci A2 9:1955
337. Gerrens H, Kuchner K (1970) Brit Polym J 2:18
338. Gerrens H, Kuchner K, Ley G (1971) Chem Ing Tech 43:693
339. Gregor L, Gerrens H (1974) Macromol Chem 175:563
340. Nomura M, Sasaki S, Xue W, Fujita K (2002) J Appl Polym Sci 86:2748
341. Kiparissides C, MacGregor JF, Hamielec AE (1979) J Appl Polym Sci 23:401
342. Rawlings JB, Ray WH (1988) Polym Eng Sci 28:237
343. Rawlings JB, Ray WH (1988) Polym Eng Sci 28:257
344. Tauer K, Muller I (1995) DECHEMA Monogr 131:95
345. Greens RK, Gonzalez RA, Poehlein GW (1976) ACS Sym Ser 24:341
346. Feldon M, McCann RF, Laundrie RW (1953) India Rubber World 128:51
347. Geddes K (1983) Chem Ind 21:223
348. Geddes KR (1989) Brit Polym J 21:443
349. Rollin AL, Patterson I, Huneault R, Bataille P (1977) Can J Chem Eng 55:565
350. Iabbadene A, Bataille P (1994) J Appl Polym Sci 51:503
351. Lee D-Y, Kuo J-F, Wang J-H, Chen C-Y (1990) Polym Eng Sci 30:187
352. Lee D-Y, Wang J-H, Kuo J-F (1992) Polym Eng Sci 32:198
353. Abad C, de la Cal JC, Asua JM (1995) Polymer 36:4293
354. Abad C, de la Cal JC, Asua JM (1994) Chem Eng Sci 49:5025
355. Abad C, de la Cal JC, Asua JM (1995) J Appl Polym Sci 56:419
356. Abad C, de la Cal JC, Asua JM (1995) DECHEMA Monogr 131:87
357. Paquet DA Jr, Ray WH (1994) AIChE J 40:73
358. Mayer MJJ, Meuldijk J, Thoenes D (1996) Chem Eng Sci 51:3441
359. Scholetens CA, Meuleijk J, Drinkenburg AAH (2001) Chem Eng Sci 56:955
360. Imamura T, Saito K, Ishikura S, Nomura M (1993) Polym Int 30:203
361. Kataoka K, Ohmura N, Kouzu M, Okubo Y (1995) Chem Eng Sci 50:1409
362. Ohmura N, Kataoka K, Watanabe S, Okubo M (1998) Chem Eng Sci 53:2129
363. Schmidt W, Kossak S, Langenbuch J, Moritz H-U, Herrmann C, Kremeskötter J (1998)
DECHEMA Monogr 134:509
364. Wei X, Takahashi H, Sato S, Nomura M (2001) J Appl Polym Sci 80:1931
365. Xue W, Yoshikawa K, Oshima A, Nomura M (2002) J Appl Polym Sci 86:2755
366. Mead RN, Poehlein GW (1988) Ind Eng Chem Res 27:2283
367. Mead RN, Poehlein GW (1989) Ind Eng Chem Res 28:51
368. Shoaf GL, Poehlein GW (1989) Polym Plast Tech Eng 28:289
369. Poehlein GW (1995) Macromol Symp 92:179
370. Fang S-J, Xue W, Nomura M (2003) Polym React Eng 11:815
Miniemulsion Polymerization
F. Joseph Schork ( )1 · Yingwu Luo 2 · Wilfred Smulders1 · James P. Russum1 ·
Alessandro Butté 3 · Kevin Fontenot 4
1
School of Chemical and Biomolecular Engineering, Georgia Insitute of Technology,
Atlanta Georgia30332–0100 USA
[email protected], [email protected],
[email protected]
2 The State Key Laboratory of Polymer Reaction Engineering, College of Material
and Chemical Engineering, Zhejiang University, Hangzhou, 310027 P. R. China
[email protected]
3 Prof. M. Morbidelli Group, ETH Zürich, Institute for Chemical- and Bioengineering,
ETH Hönggerberg/HCI F 135, 8093 Zürich, Switzerland
[email protected]
4
Eastman Chemical Company, P.O. Box 1972, Kingsport TN 37662, USA
[email protected]
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
1.1 Dispersed-Phase Polymerization . . . . . . . . . . . . . . . . . . . . . . . 133
1.2 The Concept of Miniemulsion Polymerization . . . . . . . . . . . . . . . . 135
1.3 Publication History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
4.1 Robust Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
4.1.1 Effect of Initiation and Inhibition . . . . . . . . . . . . . . . . . . . . . . . 178
4.1.1.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
4.1.1.2 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
4.1.2 Particle Size Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
4.1.2.1 Hexadecane as Costabilizer . . . . . . . . . . . . . . . . . . . . . . . . . . 184
4.1.2.2 Dodecyl Mercaptan as Costabilizer . . . . . . . . . . . . . . . . . . . . . . 186
4.1.2.3 Polymethyl Methacrylate as Costabilizer . . . . . . . . . . . . . . . . . . . 187
4.1.2.4 Influence of the Amount of the Surfactant . . . . . . . . . . . . . . . . . . 187
4.1.2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
4.1.3 Shear Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
4.1.3.1 Relative Shear Stability of Miniemulsion and Macroemulsion Latexes . . . 190
4.1.3.2 Effect of Large Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
4.1.3.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
4.2 Monomer Transport Effects . . . . . . . . . . . . . . . . . . . . . . . . . . 194
4.2.1 Polymerization of Highly Water-Insoluble Monomers . . . . . . . . . . . . 194
4.2.2 Copolymer Composition Distribution . . . . . . . . . . . . . . . . . . . . 195
4.2.2.1 Batch Copolymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
4.2.2.2 Semibatch Copolymerization . . . . . . . . . . . . . . . . . . . . . . . . . 200
4.2.2.3 Copolymerization in a Continuous Stirred Tank Reactor . . . . . . . . . . 203
4.2.3 Interfacial Polymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
4.3 Multiphase Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
4.3.1 Hybrid Miniemulsion Polymerization . . . . . . . . . . . . . . . . . . . . . 208
4.3.1.1 Alkyds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
4.3.1.2 Polyester . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
4.3.1.3 Polyurethane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
4.3.1.4 Other Hybrids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
4.3.1.5 Artificial Miniemulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
4.3.2 Nanoencapsulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
4.4 Controlled Free Radical Polymerization . . . . . . . . . . . . . . . . . . . . 216
4.4.1 Nitroxide-Mediated Polymerization . . . . . . . . . . . . . . . . . . . . . . 216
4.4.1.1 Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
4.4.1.2 Effects Of Segregation And Heterogeneity . . . . . . . . . . . . . . . . . . 219
4.4.1.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
4.4.2 Atom Transfer Radical Polymerization . . . . . . . . . . . . . . . . . . . . 223
4.4.2.1 Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.4.2.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
4.4.3 Reversible Addition Fragmentation Polymerization . . . . . . . . . . . . . 228
4.4.3.1 Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
4.4.3.2 Effects of RAFT and Transfer Agents on Emulsion Polymerization Kinetics 230
4.4.3.3 Application of RAFT in Miniemulsion . . . . . . . . . . . . . . . . . . . . 234
4.4.4 Colloidal Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
4.5 Other Applications and Future Directions . . . . . . . . . . . . . . . . . . 242
4.5.1 Anionic/Cationic Polymerization in Miniemulsions . . . . . . . . . . . . . 242
4.5.2 Polycondensation in Miniemulsions . . . . . . . . . . . . . . . . . . . . . . 243
Miniemulsion Polymerization 131
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
nm nanometer
NMA n-methylol acrylamide
NMCRP nitroxide-mediated controlled radical polymerization
NMR nuclear magnetic resonance spectroscopy
PFR plug flow reactor
pMS paramethyl styrene
PS polystyrene
PSD particle size distribution
PST polystyrene
PVA polyvinyl alcohol
PVAc polyvinyl acetate
QSSA quasi-steady state approximation
r radius
ri reactivity ratio
R gas constant
RAFT reversible addition-fragmentation chain transfer
Rp rate of polymerization
RX organic halide
SE Smith Ewart
SPS sodium persulfate
T temperature
Tg glass transition temperature
–
Vi partial molar volume of i
VAc vinyl acetate
VD vinyl n-decanoate
VEH vinyl 2-ethylhexanoate
VEOVA vinyl neo-decanoate (vinyl versatate)
VH vinyl hexanoate
VS vinyl stearate
W Fuchs stability ratio
X halogen
cij Flory Huggins interaction parameter
–
DGi partial molar free energy or phase i
g interfacial tension
ji volume fraction of component i
m viscosity
mi chemical potential of component i
mm micron
%wt percent by weight
1
Introduction
Over the past 25 years, miniemulsion polymerization has grown from being
the subject of a single paper to being the focus of a great deal of academic
and industrial research. During that time, some products have been commer-
cialized based on this technology, and in the next few years a number of new
Miniemulsion Polymerization 133
1.1
Dispersed-Phase Polymerization
Free radical polymerization may be carried out in various media. Bulk poly-
merization is the simplest, but while the reactants (monomers) are most often
liquid, the product (polymer) is solid. This leads to problems when removing
the polymer from the reactor. In addition, since most free radical polymeriza-
tions are highly exothermic, the high viscosity of the monomer/polymer mix
inhibits the removal of the heat of reaction. Solution polymerization will
reduce, to some extent, the viscosity of the polymerizing mass, but it brings
with it the environmental and health issues of organic solvents. In addition, the
solvent reduces the monomer concentration, and hence the rate of polymer-
ization. Finally, recovery and recycling of the solvent can add substantially
to the cost of the process. Nevertheless, solution polymerization of vinylic
monomers is used in a number of commercial processes.
An alternative to solution polymerization is the whole realm of dispersed-
phase polymerization. In this class of processes, the liquid monomer is dispers-
ed in a second, continuous phase, usually water. As the monomer polymerizes,
the viscosity of the dispersion remains low, aiding the removal of the heat of
polymerization. If the dispersed phase is water, the high thermal conductivity
provides a very effective heat transfer medium. The high specific heat and large
latent heat of vaporization provide a large safety margin in the event of a run-
away polymerization. In addition, water is plentiful, nontoxic, environmentally
friendly, and inexpensive.
If an oil-soluble monomer is dispersed in a continuous aqueous phase with-
out the use of surfactants, suspension polymerization results. The viscosity of
the resulting suspension will remain essentially constant over the course of the
polymerization. Oil-soluble free radical initiators are used to effect polymer-
ization. The monomer is dispersed into beads by the action of an agitator. Since
little or no surfactant is used, no emulsification takes place, and, if the agita-
tion is stopped, the monomer will form a separate bulk phase, usually above
the aqueous phase. The monomer is polymerized by the initiator within
the droplets, forming polymer beads of approximately the same size as the
monomer droplets (0.1–10 mm diameter). The product can be readily sepa-
rated from the aqueous phase (via filtration or decantation) in the form of
macroscopic particles or beads, which can be easily packaged and/or trans-
ported. Heat transfer is facilitated by the presence of the continuous aqueous
phase. Blocking agents such as clays or talcs are used to prevent particle ag-
134 F. J. Schork et al.
1.2
The Concept of Miniemulsion Polymerization
Interval I Emulsified
Monomer
swollen monomer
micelle droplet
~ 100 Å ~ 10 μ
Emulsifier
Monomer
swollen
polymar
particle
~ 500 Å
Continuous aqueous phase
Interval II Emulsified
monomer
droplet
~ 10 μ
Monomer
swollen
polymar
particle
~ 500 Å
Interval III
Monomer
swollen
polymar
particle
~ 1000 Å
b
Fig. 1 Macroemulsion polymerization (a) versus miniemulsion polymerization (b)
1.3
Publication History
2
Miniemulsion Polymerization
2.1
The Mechanism of Macroemulsion Polymerization
Rp = kp[M]pNpn–/NA
2.1.1
Interval I – Particle Nucleation
Nucleation mechanisms are generally divided into three types: micellar, homo-
geneous, and droplet. Statistically, all three types can occur simultaneously in
every reaction. However it is the preponderance of one mechanism above the
others in a given system that causes authors to consider only one in their
studies. Numerous extensions have been made to the SE micellar nucleation
theory in an effort to furnish explanations for experimental results observed
Miniemulsion Polymerization 139
for monomers with slight water solubilities. Detailed reviews of these exten-
sions are readily available [10, 11, 14, 15]. Monomer droplet nucleation is nor-
mally neglected, except when considering mini- and microemulsions, due to its
insignificant contribution to the particle number and the size distribution.
2.1.1.1
Micellar Nucleation
2.1.1.2
Homogeneous Nucleation
Although the SE micellar nucleation theory explains data for certain systems,
it fails for others. This has led some authors to propose a different mechanism
for nucleation. In the homogeneous nucleation theory, aqueous phase radicals
polymerize to form oligomers. These continue to grow until they reach a criti-
cal chain length, the size of a primary particle, and then precipitate. Through-
out the growth process, the oligomers may also flocculate or coagulate. This
theory is typically employed for relatively water-soluble monomers. Slight va-
riations of this theory have also been postulated.
Prior to 1952, little evidence for homogeneous nucleation existed [16, 17]. In
1952, Priest [18] studied the polymerization of vinyl acetate and presented a
qualitative theory for homogeneous nucleation. He concluded from experi-
mental work that aqueous phase nucleation is important in systems with
monomers that have a relatively high water solubility. Primary particle forma-
tion occurs throughout the course of the reaction. During later periods of
the reaction, large monomer-swollen polymer particles act as sinks for these
primary particles, encouraging coagulation.
In 1968, Roe [19] developed the SE limiting case equations for particle num-
ber from the homogeneous nucleation theory. He showed that the SE equation
140 F. J. Schork et al.
for particle nucleation is not unique to micellar nucleation, but results from
the SE assumptions. By assuming that (i) the nucleation stops upon depletion
of micelles, (ii) the volumetric growth rate is constant, and (iii) the radical
absorption is strictly a function of radical generation, he showed that the SE
dependency on radical flux and surfactant concentration could be generated
from homogeneous nucleation theory.
Fitch and Tsai [20, 21] developed a quantitative theory for homogeneous
nucleation. By using the collision theory for radical capture, Fitch [22] has shown
that the rate of radical capture is a function of radical production, particle
number, particle size, and diffusion distance. Primary particles may coagulate
with each other because of their small size and lower surface charge. As par-
ticles coagulate, the surface to volume ratio decreases, which causes an increase
in surface potential. When particles become sufficiently large, coagulation
ceases due to insufficient kinetic energy to overcome the biparticle surface
repulsion. Fitch and Tsai have provided experimental support for this theory by
polymerizing MMA with different initiators.
Ugelstad and Hansen [11] proposed that free radicals in the aqueous phase
propagate with dissolved monomer. Primary particles form by precipitation
when a critical chain length is reached. During growth from a monomer radical
to a primary particle, each oligomer can (i) terminate with other radicals, (ii)
precipitate if its length exceeds the critical chain length, or (iii) be captured by
particle.
A fundamental extension to the homogeneous nucleation theory was pro-
posed by Lichti et al. [23] and Feeney et al. [24]. Their theory is based on the posi-
tive skewness of the particle size distribution (PSD) as a function of volume
during Interval II. This implies that the rate of nucleation during Interval I
increases with time until it eventually drops off at the cessation of nucleation.
Lichti and Feeney claim that micellar nucleation or one step homogeneous
nucleation incorrectly predict either decreasing or constant nucleation rates.
This theory has been given the name coagulative nucleation. According
to Lichti and Feeney’s mechanism, precipitated “precursor particles” undergo
coagulation to form “true” or “mature” latex particles. A precursor particle is
unstable and said to be formed in either a micelle or the aqueous phase. Due
to their small size and hydrophilic nature, the precursors have low swelling
capacity and high radical desorption rates. Consequently, the propagation rate
is low and these precursors tend to coagulate with other precursors or mature
latex particles. These conclusions then rule out micellar nucleation as a possible
mechanism for precursor formation.
More recently, Maxwell et al. [25] suggested that the values to be used for
the critical chain length are much smaller than originally thought. They also
suggested that oligomeric radical capture is independent of particle size and
limited by the rate of propagation of the radical in the aqueous phase. However,
this theory does not consider variations in other parameters with particle size.
Miniemulsion Polymerization 141
2.1.1.3
Droplet Nucleation
Table 1 Variation of number of micelles with droplet size for MMA droplets and SLS
surfactant in a basic emulsion recipe
Base case parameters: temperature=50 °C; monomer/water ratio=0.4 g/g; surfactant CMC=
0.004 mol/L(aq); surfactant concentration=0.020 mol/L; surfactant surface area=0.57 nm2/
molecule; molecules per micelle=62; monomer watersaturation=0.137 mol/L(aq)
142 F. J. Schork et al.
2.1.1.4
Competition for Radicals
2.1.2
Interval II – Particle Growth
2.1.3
Interval III – Gel and Glass Effects
Interval III begins when all monomer droplets have vanished and/or the aque-
ous phase becomes unsaturated. Since each droplet in a macroemulsion actual-
ly absorbs radicals, they cannot disappear but rather shrink to a point where
they have no excess monomer. The monomer in the aqueous phase decreases
corresponding to the decrease in the particles. The conversion at which Inter-
val III begins varies for different monomers and systems, but is typically around
40 to 50%. However, it may not be as distinguishable in miniemulsions due to
early initiation of the gel effect.
As the monomer within the particles is consumed by polymerization, the
viscosity rises within the particles and the diffusion rate of the polymeric radi-
cals decreases. This causes a reduction in the rate of termination, which cor-
responds to a dramatic increase in the radical concentration. A higher radical
number within the particle results in an “auto acceleration” in the rate of poly-
merization. Common practice is to model this auto acceleration or gel effect by
decreasing the termination rate constant byseveral orders of magnitude as a
function of percent monomer in the particle.A free volume approach has been
used by Sundberg et al. [34]. Gilbert and coworkers [35] suggest a completely
empirical approach from precise experimental data. Empirical correlations
used in modeling the gel effect in bulk or solution polymerization have also
been modified for use in emulsion processes [36–38].
The problem with applying correlations derived from other systems to
emulsion polymerization is twofold. First, normal macroemulsion particles
are said to be created with 30 to 40% monomer in them and so the unbiased
(zero conversion) termination rate is unknown. Secondly, the diffusional
limitations in particles might be quite different from those observed in bulk or
suspension polymerizations. It is for these reasons that an empirical approach
is suggested.
144 F. J. Schork et al.
2.2
The Mechanism of Miniemulsion Polymerization
smaller droplets is larger than the gain in interfacial area of the larger ones. The
reduction in interfacial energy is the driving force for degradation of the small
droplets.
If Ostwald ripening is allowed to continue unchecked, creaming of the
monomer will occur as the droplet sizes become large enough for Stokes law
creaming to occur. This will occur in a matter of seconds to minutes. If the
system is initiated, bulk polymerization of the monomer layer will occur. If the
emulsion is stirred, an emulsion of large monomer droplets (of the order of
those of a macroemulsion) will result, and if the stirred emulsion is initiated,
macroemulsion polymerization will take place. A costabilizer functions to
prevent Ostwald ripening by retarding monomer diffusion from the smaller
droplets to the larger. Costabilizers should be highly insoluble in the aqueous
phase (so that they will not diffuse out of the droplets), and highly soluble in
the monomer droplets. Under these conditions, diffusion of the monomer out
of the smaller droplets results in an increase in the concentration of the co-
stabilizer in those particles (since, by definition costabilizers are too insoluble
in the aqueous phase to leave the droplet). The increase in free energy associ-
ated with the concentration of the costabilizer balances the decease from
reduced interfacial area caused by Ostwald ripening, and, at some point, ripen-
ing stops. Since all costabilizers are somewhat water-soluble, Ostwald ripening
will proceed, but on a timescale of months, which is unimportant since the
timescale of polymerization is minutes to hours. This phenomenon is shown
in Fig. 3 [42]. Here a macroemulsion and a miniemulsion of methyl methacry-
late (same recipe, but with no costabilizer, and no high shear for the macro-
emulsion) are shown after three hours without mixing. The macroemulsion is
completely creamed, since Stokes law creaming has taken place on the large
monomer droplets. The miniemulsion has not creamed, since Brownian motion
is sufficient to prevent creaming of the submicron monomer droplets. Similar
miniemulsions have remained stable for six months or more.
Ugelstad et al. [45] have applied this equation to various monomers and sur-
factants. It is clear from this equation that the free energy increases as the
phase diameter decreases. The smaller the monomer droplet, the less stable it
is. Therefore, a driving force exists for the monomer to diffuse from a small
droplet to a larger one. Over time, non-monodisperse systems of droplets of
pure monomer will decrease in number as the smaller droplets swell the larger
ones and then disappear. Jannson [46] has shown that this occurs in unagitated
systems, and that the timescale for diffusional instability can be on the order
of seconds.
Prior to 1962, droplets below 1 mm were considered too unstable to partici-
pate in the nucleation process. In 1962, Higuchi and Misra [47] proposed that
the addition of a water insoluble compound to the monomer will enhance the
stability of small droplets by prohibiting diffusion. In 1973, Ugelstad et al. [48]
showed how submicron styrene droplets could be made stable enough to par-
ticipate in the nucleation processes by adding small amounts of cetyl alcohol.
Later, Ugelstad [48] used Eq. 2 to explain these experimental observations.
It can be shown [49] for two phases in equilibrium that the partial molar free
energies must be equal. In an emulsion (or miniemulsion) there are three
phases: monomer droplets, the aqueous phase and polymer particles. Since
monomer is soluble in all of these phases, the equilibrium condition requires
that the three phases have equal partial molar free energies. In the presence of
monomer droplets, emulsion polymer particles contain 30–80% monomer in
them. Therefore, they are said to be “swollen” with monomer. Ugelstad et al. [48]
and Azad and Fitch [50] have shown that addition of a third water-insoluble
component to a swollen polymer particle can increase the monomer to polymer
ratio. They have shown that an optimum chain size for the additive exists since
the solubility of the additive increases as the chain size decreases. They found
Miniemulsion Polymerization 147
that the optimum hydrocarbon stabilizer is hexadecane. Others have found that
if a fatty alcohol is used as the stabilizer, the minimum chain length required
is 12 carbon atoms [51].
Ugelstad et al. [52] have shown how this theory may be used to devise a
method to prepare large monodisperse particles of predetermined size. By
using the appropriate amount of cosurfactant, polymer particles can be swollen
with monomer to the desired size. Polymerization in conditions that prevent
additional nucleation results in large monodisperse polymer particles of size
1–100 µm. This method has been criticized by other groups as being in error
due to measurement selectivity.
If Ostwald ripening is retarded by using a costabilizer, predominant droplet
nucleation can be achieved. This is the basis of miniemulsion polymerization.
One of the first comprehesive studies of miniemulsion polymerization was
done on styrene by Choi et al. [53].
2.3
Mathematical Modeling of Miniemulsion Polymerization
3
Properties of Miniemulsion Polymerization
3.1
Shear Devices
3.2
Choice of Surfactant
3.3
Choice of Costabilizer
Early work in miniemulsion polymerization [5, 48–50] used either cetyl alcohol
(CA) or hexadecane (HD) to retard Ostwald ripening in submicron monomer
droplets. Both CA and HD, referred to here as costabilizers, have the requisite
properties for a costabilizer: high monomer solubility, low water solubility and
low molecular weight. The need for these properties can be seen from Eq. 2.
High monomer solubility will give a large Flory Huggins interaction parame-
ter between the costabilizer and the monomer (cij). Low water solubility will
ensure a distribution coefficient for the costabilizer that very strongly favors the
monomer drops, giving a higher volume fraction of costabilizer in the droplet.
Low molecular weight will give a high ratio of costabilizer molecules to mo-
nomer molecules (mij) in the droplet.All of these factors will enhance swelling,
or retard monomer loss via Ostwald ripening.
With cetyl alcohol, there is the complication that the polarity of the mole-
cule may cause it to reside at the surface of the droplet, imparting additional
colloidal stability. Here, the surfactant and costabilizer form an ordered struc-
ture at the monomer-water interface, which acts as a barrier to coalescence and
mass transfer. Support for this theory lies in the method of preparation of the
emulsion as well as experimental interfacial tension measurements [79]. It is
well known that preparation of a stable emulsion with fatty alcohol costabiliz-
ers requires pre-emulsification of the surfactants within the aqueous phase
prior to monomer addition. By mixing the fatty alcohol costabilizer in the water
prior to monomer addition, it is believed that an ordered structure forms from
the two surfactants. Upon addition of the monomer (oil) phase, the monomer
diffuses through the aqueous phase to swell these ordered structures. For long
chain alkanes that are strictly oil-soluble, homogenization of the oil phase is
required to produce a stable emulsion.Although both costabilizers produce re-
152 F. J. Schork et al.
latively stable emulsions, Azad et al. [80] have shown that the alkanes will pro-
duce emulsions of higher stability. A 1:3 molecular ratio of surfactant to co-
stabilizer has been shown to provide optimal stability in emulsion systems
where the costabilizer is a fatty alcohol. Shah [81] postulated that this ratio is
due to an optimum alignment of surfactant and costabilizer molecules at the
interface in microemulsions. Hallworth and Carless [82] have proposed that
the stability of an emulsion containing long chain alkanes of fatty alcohols
comes from a film at the interface which makes collisions at the interface more
elastic.
Various researchers [83–88] have concluded from experimental data that
liquid crystals of surfactants exist at the interface. These observations have
been suggested through birefringence, interfacial tension, and viscosity mea-
surements. Lack et al. [87] studied the formation of liquid crystals with fatty
alcohol costabilizers as a function of concentration, chain length, and ratio of
emulsifiers using birefringence measurements. Tertiary phase diagrams were
presented which show two regions of normal micelle formation. There are also
three regions of homogeneous anisotropic mesophases. These three meso-
phases, are (i) hexagonal rodlike aggregates of mixed micelles, (ii) lamellar
double layers with overlapping tails, and (iii) lamellar double layers dispersed
in the aqueous phase. The concentration of surfactants used in miniemulsions
is found to fall within region (iii). These “liquid crystals” were shown to form
more easily at higher emulsifier concentrations, with shorter chain alcohols,
and with sonication.
Although evidence exists for liquid crystal formation with fatty alcohol
costabilizers, it does not for systems with long chain alkanes. Delgado et al. [89]
have presented evidence that the role of hexadecane costabilizer in miniemul-
sion polymerizations is one of diffusional control. Rodriguez [90] and Delgado
[91] have reported that no optional ratio of hexadecane to SLS exists in the
preparation of miniemulsions. This provides evidence for the lack of crystal
formation. Ugelstad et al. [45] have presented evidence that alkanes are more
likely to follow the diffusion mechanism.
3.3.1
Polymeric Costabilizers
The use of polymer as a costabilizer was first reported by Reimers et al. in 1995
[92]. Conventional thinking has been that effective costabilizers must be highly
water-insoluble, highly monomer-soluble, and of low molecular weight, as
required by Eq. 2. Polymer made from the same monomer from which the
miniemulsion is to be made will be highly water-insoluble, and most polymers
are quite soluble in their own monomers. The requirement that the costabilizer
must be of low molecular weight is based on reported swelling experiments and
theoretical swelling calculations [93]. Data from Schork and Reimers [94]
demonstrate that it is possible to create miniemulsion latexes with a poor co-
stabilizer (polymer). The inclusion of a small amount (~4%wt) of a monomer-
Miniemulsion Polymerization 153
Polymeric materials are not costabilizers in the sense that costabilizers cause
superswelling. Rather, they slow the onset of Ostwald ripening and preserve the
number of monomer droplets, if not their size. However, this review will take
a functional, rather than thermodynamic definition of a costabilizer, and in-
clude a discussion of the use of polymers as agents to enhance droplet nuclea-
tion under the heading of costabilizers.
3.3.2
Monomeric Costabilizers
3.3.3
Other Costabilizers
The use of a chain transfer agent (CTA) as a costabilizer opens up new possibi-
lities for molecular weight control. Macroemulsion polymerizations which util-
ize higher molecular weight mercaptan chain transfer agents exhibit retarded
transport of the CTA from the monomer droplet into the growing polymer par-
ticles. This results in slower delivery of the mercaptan to the reaction sites over
the course of the polymerization. (In some commercial recipes this retarded
transport is used to “meter” the highly reactive CTA to the reaction site.) If the
mercaptan were at the site of polymerization, as in a miniemulsion, new degrees
of freedom in selecting chain transfer agents would exist. That is, the relative re-
activities of chain transfer versus propagation can be used to select the CTA, with-
out relying on retarded mass transfer. This may increase the efficiency of chain
transfer (since CTA will not be “trapped”in the shrinking monomer droplets near
the end of Interval II), or at least allow the chemist additional degrees of freedom
in tailoring the molecular structure by manipulating the reaction conditions.
Miniemulsion Polymerization 155
3.3.4
Enhanced Nucleation
Miller et al. [109–111] report that the addition of a small amount (as small
as 0.05%wt) of polystyrene (PS) to the styrene phase of a miniemulsion poly-
merization of styrene causes an increase in both the rate of polymerization and
the number of final polymer particles. This is not just a polymeric costabilizer
effect, since these emulsions were also stabilized with what are known to be
effective levels of HD or CA, although the effect was more pronounced with CA.
With the addition of 1%wt styrene, the number of final particles was nearly the
same as the original number of droplets, indicating 100% droplet nucleation.
This was not the case for equivalent polymerizations without the PS. For poly-
merizations without the PS, the final particle number varied with the initial
initiator concentration to the power of 0.31.With 1%wt PS, the particle number
was independent of the initiator concentration, which is very clear evidence of
100% droplet nucleation. Miller hypothesized that miniemulsions prepared
from polystyrene in styrene solutions resemble the polymer particles formed
in normal (no polymer) miniemulsion polymerizations at early conversions.
This being the case, these polymer-containing droplets would be able to effec-
tively compete with growing polymer particles for free radicals, whereas their
counterparts that contain no polymer are not, and as a result a greater fraction
of the initial droplets become polymer particles. Based on this mechanism,
Miller speculated that the presence of the polymer increases the capture effi-
ciency of the droplets by modifying either their interior (by increasing the
interior viscosity, thereby increasing the probability of a radical propagating
rather than exiting) or the droplet/water interface (by disrupting the surfactant/
CA interfacial barrier to radical entry). Experimental results were reported
which support the latter explanation. (It should be noted that the effect was
most pronounced with CA, and that other investigators have reported near
100% droplet nucleation with HD and without added polymer.)
In a follow-on set of papers, Blythe et al. [112–115] studied the effect on the
polymerization kinetics of changing the properties of the polymer used to en-
hance nucleation. Miniemulsions were formed from CA with PS as the polymer
additive, and CA as the costabilizer. Varying the molecular weight of the PS
from 39,000 and 206,000 in systems containing 1% PS did not change the ki-
netics. Also, changing the end group of the polymer chain from a hydrophobic
end group to a hydrophilic end group had no effect on the kinetics in 1%
polymer systems. However, predissolving 1% PS in a miniemulsion always
results in a significant enhancement in the kinetics compared to similar sys-
tems that do not contain predissolved polymer. The authors conclude that this
enhancement in the kinetics is not due to either a change in the interior vis-
cosity of the droplets or a disruption of the condensed phase formed by cetyl
alcohol and sodium lauryl sulfate (since these effects would be altered by the
variations in the PS above). Instead, it was suggested that the enhancement can
primarily be attributed to a preservation of the droplet number due to the pre-
Miniemulsion Polymerization 157
3.4
Choice of Initiator
3.5
Robust Nucleation
3.6
Monomer Transport Effects
3.7
Droplet Stability
sion) seem to indicate a correlation between process and final particle size of
the polymer particle dispersion, this is usually not the case. On the contrary,
these techniques should rather be distinguished by either their nucleation
process or the stability of the initial oil (the monomer) dispersion. These two
aspects of the emulsion polymerization processes, nucleation and stability, are
strictly related and depend upon each other.
To clarify this point, let us first discuss the case of a conventional emulsion
recipe, or macroemulsion. In this case, the monomer is initially dispersed in
water under agitation in the presence of surfactants [121], resulting in a rather
coarse oil dispersion. Droplet size typically depends upon rate of stirring, and
the resulting droplet size is generally in the 1–10 mm range (the reason for the
name “macroemulsion”). This dispersion is unstable and, if the stirring is
stopped, the monomer phase separates very quickly. Because of the large sizes
of the droplets, the total surface area of the dispersion is small (the surface-to-
volume ratio for spherical droplets is proportional to the inverse of the droplet
diameter) and most of the surfactant is not used to stabilize the droplets; it
aggregates in water to form micelles. These micelles, and not the droplets, are
the main loci for polymerization, while the droplets just act as monomer reser-
voirs. In other words, the final particles in the macroemulsion do not correspond
to the initial droplets. The most important consequence is that, in order to have
a successful process, the monomer, as well as all other possible comonomers
and coreactants, must be water-soluble enough to diffuse from droplets to
particles.
For historical reasons [122], a stable dispersion of sub-micron droplets is
called a miniemulsion. Miniemulsions do not form spontaneously and require
high shear devices to form. The resulting dispersion is usually quite fine, with
the droplet size ranging from 50 to 500 nm [3], a large droplet surface area, with
all of the surfactant used to stabilize the droplets and with no more micelles
present in the system.As a result, the droplets become the predominant loci for
nucleation and polymerization. That is, in an ideal miniemulsion, there is a 1:1
correspondence between initial monomer droplets and final polymer particles.
(This 1:1 correspondence is not always attained, and remains a point of contro-
versy [1].) This was experimentally demonstrated for the first time by Ugelstad
and coworkers in 1973 [5]. The consequence of this virtual copying process
from droplets to particles is twofold: (i) final particle size is given by the initial
droplet dispersion, and both surfactant coverage and surface tension do not sig-
nificantly change during the process; (ii) any kind of hydrophobic component
can be conveniently included in the recipe, since we can be sure that it is going
to participate to the polymerization process.
A third type of emulsion process is the so-called microemulsion [123]. In
microemulsions, the polymerization starts in droplets as well. However, these
are thermodynamically stable and, in contrast to miniemulsions, they form
spontaneously by gentle stirring. They consist of large amounts of surfactants
or mixtures of them, and they possess an interfacial tension close to zero at the
water/oil interface, with droplet sizes usually ranging between 5 and 50 nm. In
Miniemulsion Polymerization 161
3.7.1
Stability of Monomer Dispersions
There are two ways by which a dispersion of monomer droplets can degrade:
(i) by droplet coalescence, and (ii) by diffusion degradation (often referred to
as Ostwald ripening).While the first mechanism of degradation can be avoided
by adding enough surfactant to the system, when two monomer droplets of dif-
ferent sizes, and stabilized by a surfactant, are put in water, they will start ex-
changing monomer without even making direct contact – through monomer
diffusion across the water (continuous) phase.
This process of molecular diffusion is governed by the difference in chem-
ical potentials of the monomer in the two droplets. Morton’s equation has been
successfully used to describe the swelling of polymer particles with monomer
[32]. According to this equation, the chemical potential of the monomer in a
droplet of radius rp(µ (d)
m ) in the presence of polymer is given by:
–
µ(d)
m 1 2
2gVm
= ln (jm) + 1 – 7 p j + c j
m,p p + (3)
RT mpm rpRT
6 9
where jm and jp represent the volume fraction of monomer and polymer, re-
spectively, in the particle (so jm+jp=1); mpm is the ratio of equivalent number
of molecular segments between monomer and polymer; cm,p is the interaction
parameter between monomer and polymer; g is the interfacial tension at the
–
water/oil interface; Vm is the molar volume of the monomer; R is the universal
gas constant; T the temperature. In the case of pure monomer droplets, the
partial molar free energy of mixing is zero, which is represented in Eq. 3 by the
first two terms, accounting for the entropy of mixing, and by the third term,
accounting for the enthalpy of mixing. Ugelstad and Hansen showed that this
expression could be satisfactorily extended to the case where any other species,
not necessarily a polymer, is involved [11].Accordingly, we will use Eq. 3 to con-
veniently describe the monomer chemical potential in a miniemulsion droplet.
According to Eq. 3, in the case of a macroemulsion of pure monomer
droplets, the monomer chemical potential is given by the last term with (jm=1),
162 F. J. Schork et al.
Fig. 6 Monomer chemical potential for a pure monomer droplet as a function of droplet
–
radius (g=25 Mn/M; Vm=1.1·10–4 m3 mol; T=298.15 K). Point A represents an unstable equi-
librium
accounting for the contribution of the interfacial energy. Therefore, given two
droplets with the same interfacial tension (we assume that surface tension is a
function of the degree of coverage only, and that this is the same for the two
droplets), their chemical potential is a function of the droplet radius only and
it is only the same if the two droplets have the same size. Therefore, when two
droplets of different radii rd,1 and rd,2 are put together, the difference in their
chemical potential is given by:
Dµ(d)
m,1–2 2gṼm 1 1 1 1
02 = 9 5 – 5 = y 5 – 5 (4)
RT RT rd,1 rd,2 rd,1 rd,2
Fig. 7 Schematic representation of the role of the hydrophobe in halting the ripening
process in a miniemulsion
Dµ(d)
m,1–2 jm,1 1 1
= ln 7 + (1– mm,h)(jm,2 – jm,1) + y 5 – 5 (5)
RT
02 jm,2 rd,1 rd,2
where the indexes 1 and 2 refer to the first and the second droplet, respectively,
subscript h identifies the costabilizer, and y is defined as in Eq. 4. Knowing that
lnjm≈–(1–jm) for small volume fractions of the costabilizer, it is possible to
simplify the previous equation as follows:
Dµ(d)
m,1–2 1 1
= mm,h (jm,1 – jm,2) + y 5 – 5 (6)
RT
02 rd,1 rd,2
Since the two droplets have the same initial composition (jm,1=jm,2), it follows
from this equation that the difference in chemical potential is given by the dif-
ference in size only, and (as in a macroemulsion) monomer will flow from the
smaller to the bigger droplet. The main consequence of this process is that
the costabilizer is concentrated in the smaller droplet, while the big droplet
becomes more and more dilute. This creates a gradient in composition (see
Fig. 7b), which is represented by the first term on the right hand side of Eq. 6.
164 F. J. Schork et al.
∂ µ(9)
m y
6 6 = – 32 (9)
∂rd RT rd
Let us check the conditions under which the derivative of the chemical
potential is positive in a miniemulsion. Under the hypothesis of small concen-
trations of costabilizer in the droplet (jmÆ1), and of ideal behavior of the
mixture (cm,h=0), the condition of local stability leads to the following expres-
sion:
∂ µ(9)
m 9mm,hVh y
6 6 = 04 –4 (10)
∂rd RT 4p r 4d rd2
9mm,hVh
rd < r–d = 04 (11)
4py
The previous expression returns the value of the droplet radius corresponding
to the maximum of the monomer chemical potential curve (r–d), and says that,
for a given amount of costabilizer in the particle (Vh), the droplet radius must
be smaller than r–d in order to be locally stable. The same equation can be
expressed in terms of the critical costabilizer volume fraction, j–h, as follows:
3/2
y
jh > j–h = 9 (12)
3mrh
where rh=(3Vh/4p)1/3. Both these two critical quantities, r–d and j–h, are functions
of costabilizer concentration, costabilizer type and surface tension, as well as
the corresponding value of the monomer chemical potential, which is given by
the following simple expression:
µ(d)
m y 3/2
1 1/2
=2 4 (13)
RT rd =r–d 3 mm,h rh
6 01
but the amount of costabilizer in the droplets is varied while the droplet radius
is kept constant and the costabilizer volume fraction varies. Again, as the co-
stabilizer concentration decreases, the system moves from a region of complete
stability to a region of instability. By observing Fig. 9b, it is also possible to
distinguish between local and global equilibrium stability. The curve corre-
sponding to jh=0.04 clearly satisfies the requirements of local equilibrium sta-
bility. In fact, if the system is slightly perturbed, the equilibrium point is always
168 F. J. Schork et al.
Fig. 9c,d (C) effect of surface tension (g=5, 15, 25 and 35 mN/m); and (D) effect of hy-
drophobe to monomer segment ratio (1/mm,h=1, 1.29, 2, 5 and 10). Other parameters (oth-
erwise differently indicated): mm,h=1; cm,h=0; g=25 mN/m; Ṽm=1.1 · 104 m3 mol; T=298.15 K;
jh=0.04; rd=100 nm
Miniemulsion Polymerization 169
3.7.2
Experimental Validation
8 kbT
b=8 (14)
3µW
where m represents the water viscosity and W is Fuch’s stability ratio (see [129]
for a further explanation of how this value has been computed). In Fig. 10b, the
quantity 1/bNd is reported, which expresses the characteristic time for coales-
cence. We can see that even small surface potentials are enough to give great
colloidal stability. These values correspond to rather low surfactant coverage
(around 5%) [124, 129], which is well below the typical coverage measured
experimentally for miniemulsions [127]. It is therefore apparent that colloidal
stability is a very strong function of droplet size, and it is quite difficult to
explain the formation of bimodalities or monomer separation by colloidal
stability only.
On the other hand, previous analyses of the monomer chemical potential in
miniemulsions may justify some of the previous results. In Fig. 8, it is shown
that miniemulsion formulations are often very close to instability, with positive
monomer chemical potentials (which corresponds to conditions of monomer
oversaturation, or superswelling). Experimental proof of this was reported by
Landfester et al. [127], who said that the Laplace pressure is much larger than
the osmotic pressure at equilibrium. We must also consider that, in reality, we
are never dealing with perfectly monodispersed distributions of droplets, as the
previous analysis supposed. Therefore, it is realistic to suppose that in the pre-
sence of a droplet size distribution, a fraction of the droplets lie in the unstable
region and can lead to the formation of either pronounced bimodalities,
monomer phase separation, or droplet flocculation [124]. Moreover, from the
analysis of Fig. 9a to d, we observe that critical stability is almost reached when
small droplets are formed. Therefore, it is not surprising to observe that smaller
droplets require larger surfactant coverage to be stable, since lower surface
tension helps the droplets to move away from the instability region. The same
influence of surface tension on the chemical potential of the monomer could
also explain why, on the addition of surfactant, the sizes of the droplets do not
change significantly after emulsification. In fact, by adding surfactant, one
decreases the interfacial tension and stabilizes the miniemulsion ripening.
172 F. J. Schork et al.
Fig. 10 Coagulation rate constant (b) and characteristic time for coagulation (1/bNd) as a
function of droplet size for various droplet surface potential values (z=15, 25 and 40 Mv),
as computed by DLVO theory. In order to compute Nd, a system with 20% oil fraction was
supposed
Miniemulsion Polymerization 173
3.8
Semibatch and Plug Flow Reactors
dent of the method of addition, and was equal to the seed particles plus the
number of miniemulsion droplets added. Therefore, with predissolved polymer,
the droplet number (but not the droplet size distribution as inferred from the
final PSD) was preserved. This result reinforces the idea of polymer as an agent
for preserving droplet number.
Macro- and miniemulsion polymerization in a PFR/CSTR train was modeled
by Samer and Schork [64]. Since particle nucleation and growth are coupled for
macroemulsion polymerization in a CSTR, the number of particles formed in
a CSTR only is a fraction of the number of particles generated in a batch re-
actor. For this reason, their results showed that a PFR upstream of a CSTR has
a dramatic effect on the number of particles and the rate of polymerization in
the CSTR. In fact, the CSTR was found to produce only 20% of the number of
particles generated in a PFR/CSTR train with the same total residence time as
the CSTR alone. By contrast, since miniemulsions are dominated by droplet
nucleation, the use of a PFR “prereactor” had a negligible effect on the rate of
polymerization in the CSTR. The number of particles generated in the CSTR
was 100% of the number of particles generated in a PFR/CSTR train with the
same total residence time as the CSTR alone.
Durant [133] has used a PFR to achieve a miniemulsion with a solids con-
tent in the industrially relevant range. Ouzineb and McKenna [134] have used
a PFR to obtain miniemulsion latexes with high solids contents.
3.9
Continuous Stirred Tank Reactors
Observation (i) above can be understood in terms of droplet nucleation and the
lack of competition between nucleation and growth.A mechanistic understand-
ing of observation (ii) above was provided by Samer and Schork [64]. Nomura
and Harada [136] quantified the differences in particle nucleation behavior for
macroemulsion polymerization between a CSTR and a batch reactor. They
started with the rate of particle formation in a CSTR and included an expression
for the rate of particle nucleation based on Smith Ewart theory. In macroemul-
sion, a surfactant balance is used to constrain the micelle concentration, given
the surfactant concentration and surface area of existing particles. Therefore,
they found a relation between the number of polymer particles and the resi-
dence time (reactor volume divided by volumetric flowrate). They compared this
relation to a similar equation for particle formation in a batch reactor, and con-
cluded that a CSTR will produce no more than 57% of the number of particles
produced in a batch reactor. This is due mainly to the fact that particle forma-
tion and growth occur simultaneously in a CSTR, as suggested earlier.
An approach similar to that taken by Nomura and Harada was used by
Samer to quantify the effects of droplet nucleation on emulsion polymerization
kinetics in a CSTR. In their simplified analysis, it was assumed that radical cap-
ture by particles and droplets is proportional to the ratio of particle and droplet
diameters. This assumption is reliable at low to moderate residence times, when
polymer particles still closely resemble monomer droplets with respect to com-
position and surface characteristics. For predominant droplet nucleation, the
maximum particle generation is limited by the concentration of monomer
droplets in the feed. In Fig. 11 the steady state particle generation is given as
a function of the residence time and temperature. Nucleation efficiency is
defined as the number of particles divided by the number of droplets in the
Fig. 11 Model predictions for the number of particles in CSTR miniemulsion polymeriza-
tion expressed as the number of particles divided by the number of droplets in the feed
(from [64])
176 F. J. Schork et al.
feed and is shown to increase with increasing temperature. It can be seen in this
figure that the nucleation efficiency approaches unity for residence times
greater than 30 minutes. Therefore, whereas the particle number is limited by
surfactant in macroemulsion emulsion polymerization in a CSTR, in mini-
emulsion polymerization, the particle number is limited by monomer droplet
concentration, but to a much smaller extent. Therefore, since nucleation effi-
ciencies in miniemulsion polymerization approach 100%, rather than the 57%
predicted for macroemulsion polymerization, the steady-state conversion in
a miniemulsion (proportional to particle number) found by Barnette [36] is
approximately twice that in a macroemulsion.
Samer and Schork [69, 137] confirmed Barnette’s findings on the lack of
oscillations in the CSTR miniemulsion polymerization of MMA.Aizpurua and
Barandiaran [138] confirmed the lack of oscillations in CSTR miniemulsion
polymerization over a wide range of surfactant and initiator concentrations
for VAc. They also observed near-identical MWD for macroemulsion and mini-
emulsion polymerizations under the same conditions. This is understandable,
since MWD should be determined by particle size (and number of radicals per
particle) rather than by particle nucleation mechanism. Aizpurua et al. [139]
successfully used polymeric costabilizers in the CSTR miniemulsion polymer-
ization of VAc at high solids levels. Neither sonication alone, nor the presence
of costabilizer alone, was able to eliminate the oscillations found in macro-
emulsion polymerization. However, sonication and costabilizer together were
capable of eliminating oscillations, indicating droplet nucleation. The results
cited above are particularly significant, since, as fresh miniemulsion droplets
are introduced into a CSTR, they must compete to retain their monomer with
existing particles that may contain more than 50% polymer (50% conversion).
Therefore, for instance, miniemulsion droplets stabilized with 5% polymer
would need to compete with existing particles containing 50% polymer. The
fact that the particle number approaches the droplet number in the feed sug-
gests that the droplet number (if not droplet size distribution) is conserved.
Samer also demonstrates the existence of multiple steady states in isothermal
miniemulsion polymerization in a CSTR. This is not surprising, since multipli-
city is a function of gel or Trommsdorf effect, and not of nucleation mechanism.
Miniemulsion copolymerization in a CSTR involves some very interesting
features. However, in the interest of clarity, these systems will be discussed
along with results for batch copolymerization.
4
Applications
4.1
Robust Nucleation
4.1.1
Effect of Initiation and Inhibition
4.1.1.1
Results
unknown. If we plot the log of the nitrite concentration against the log of the
number of particles, we find a linear relationship between them, as depicted by
Fig. 12. The slope of this line is 0.153±0.009. This value is close to the value of
0.2 reported for the initiator dependence, perhaps implying that the function
of the water-phase retarder is simply to reduce the effective radical flux to the
particles. The polymer-stabilized miniemulsions are far less sensitive to the
presence of the retarder than are the macroemulsions. The retarder has little ef-
fect on the particle number. Particle numbers remained fairly constant up to a
concentration of 5 mMaq. Up to this amount, the dependence of the retarder
concentration on the number of particles was calculated to be 0.020±0.007
(Fig. 13). This is significantly less than the value found for macroemulsions.
(It is presumed that the highest level of retarder prevents a large fraction of the
droplets from ever being nucleated.) Monomer conversions exhibit prolonged
nucleation periods, but the rates are not significantly affected.Again the nitrite
is acting as a retarder, since no induction period is observed.
Macroemulsion polymerizations carried out in the presence of an oil-phase
inhibitor (DPPH) resulted in an increase in the number of particles. Presum-
ably initiator radicals that enter droplets are terminated by the inhibitor,
resulting in dead particles. These particles do not grow, and hence do not con-
sume surfactant to stabilize their increasing surface area, until they absorb
another radical. The surfactant not adsorbed by dead particles is available to
Miniemulsion Polymerization 181
stabilize new particles, thereby increasing the total number of particles. Since
the nucleation period is lengthened, the polydispersity increases. Figure 14
shows that the dependence of the inhibitor concentration on the number of
particles is 0.176±0.010. Conversion time curves indicate that an induction
period results from the presence of the inhibitor. Since polymer-stabilized
miniemulsion polymerization occurs via droplet nucleation, it should be less
sensitive to oil-phase inhibition. Initiator radicals will enter the droplet one
after the other until all of the inhibitor is used up, and the monomer polymer-
izes. This does not affect the number of droplets or particles.As seen in Fig. 15,
the number of particles is proportional to the DPPH concentration raised to the
power of 0.0031±0.0001. Therefore, the number of particles is essentially
independent of the presence of inhibitor.
4.1.1.2
Summary
Shifting the site of nucleation to the droplets greatly enhances the robustness
of the nucleation process to recipe variations, inhibition levels, and changes in
operating procedure (initiation rate and/or agitation rate).As a result of droplet
nucleation, polymer-stabilized miniemulsion polymerizations are far less sen-
Miniemulsion Polymerization 183
X a (macroemulsion) a (miniemulsion)
4.1.2
Particle Size Distribution
There has been a belief that, due to the fact that the original miniemulsion
droplets are formed by a shear process, the droplet size distribution will be
broad, and so the resulting PSD will have a large polydispersity (as measured
by the polydispersity index, defined as the mass average over the number av-
erage particle radius). In a recent note [142] Landfester et al. discuss particle
size polydispersity in miniemulsions and attempt to dispel the idea that
miniemulsions necessarily have broader PSD than the equivalent macroemul-
sions. Rather, they argue that the PSD of a miniemulsion can be either broader
or narrower than its macroemulsion counterpart, and that, in most cases, the
miniemulsion will have a polydispersity equal to, or only very slightly greater
than, the equivalent macroemulsion.
184 F. J. Schork et al.
4.1.2.1
Hexadecane as Costabilizer
Macroemulsion – 1.04
0.5 min 135 1.01
1 min 112 1.03
2 min 96 1.00
5 min 87 1.03
10 min 84 1.02
20 min 83 1.01
Macroemulsion 98 1.04
0.33 109 1.03
0.66 108 1.01
1.66 108 1.01
3.33 102 1.04
5 100 1.03
6.66 99 1.05
8.33 95 1.01
persity. It was also found that the droplet size is initially a function of the
amount of mechanical agitation. The droplets are rapidly reduced in size
throughout sonication in order to approach a pseudo-steady state [143]. Once
this state is reached, the size of the droplet does not change. Higher sonication
time causes a slight reduction in polydispersity.
After halting sonication, a rather rapid equilibration process must occur.
Since the droplet number after sonication is fixed, this process does not in-
fluence the average size, but the droplet size distribution usually undergoes
very rapid change. It was found that steady-state miniemulsification results in
a system “with critical stability”; in other words the droplet size is the product
of a rate equation of fission by ultrasound and fusion by collisions, and the
droplets are as small as possible for the timescales involved. The equality of
droplet pressures makes such systems insensitive against net mass exchange by
diffusion processes (after the very fast equilibrium process at the beginning),
but the net positive character of the pressure makes them sensitive to all changes
in the droplet size. Steady-state homogenized miniemulsions, which are criti-
cally stabilized, undergo droplet growth on a timescale of hundreds of hours,
presumably by collisions or by costabilizer exchange. As can be seen from
Table 7, during this growth, the polydispersity does not change significantly.
186 F. J. Schork et al.
Table 7 Influence of time delay between the ultrasonication and the polymerization
(from [143])
0 82 1.01
1 87 1.05
6 108 1.03
48 152 1.03
96 164 1.04
4.1.2.2
Dodecyl Mercaptan as Costabilizer
Mouron et al. [105] and Wang et al. [106] have used dodecyl mercaptan (DDM)
as the costabilizer in styrene and MMA miniemulsion polymerizations, re-
spectively. Some of the results are shown in Table 8 and Table 9. For styrene
(Table 8), the macroemulsion is compared with miniemulsions containing
varying levels of DDM (costabilizer). In this case, the macroemulsion has a
broader particle size distribution than all but one of the miniemulsions. For
MMA (Table 9), miniemulsions and the equivalent macroemulsions have been
compared at varying initiator concentrations. In this case, the macroemulsions
Macroemulsion 1.02
1 1.01
2 1.01
3 1.02
4 1.04
Table 9 Dodecyl mercaptan as costabilizer with methyl methacrylate monomer (from [105])
all have narrower particle size distributions, although the difference is hardly
significant.
4.1.2.3
Polymethyl Methacrylate as Costabilizer
Reimers and Schork [144] have used polymethyl methacrylate as the co-
stabilizer for methyl methacrylate miniemulsion polymerization. A portion of
the results are shown in Table 10. In this case, the miniemulsion has a narrower
particle size distribution than the equivalent macroemulsion.
4.1.2.4
Influence of the Amount of the Surfactant
Polydispersity index
Macroemulsion 1.02
Miniemulsion 1.01
4.1.2.5
Summary
Based on the data above, it would appear that it is possible, via miniemulsion
polymerization, to make a polymer latex with a particle size distribution that
approaches that made by macroemulsion polymerization. In some cases, the
miniemulsion product may be even narrower than the macroemulsion. There
are two significant mechanisms leading to this narrowness. First, the monomer
droplet size distribution is to some extent determined by the thermodynamics
of swelling, and not solely by the droplet size distribution induced by the soni-
cator or homogenizer. For this to be true, the process should include a ripening
time between sonication and polymerization. During this ripening time, the
droplets will come to swelling equilibrium. Studies show that the ripening time
is of the order of seconds to minutes, and is naturally included in the prepara-
tion of batch polymerizations. Second, the narrowness of the particle size dis-
tributions depends on the ability to nucleate nearly all of the droplets over a
short period of time. If droplet nucleation takes place over a longer period of
time, some particles will have polymerized for a longer time, and some droplets
will lose monomer by mass transfer to growing particles before the droplets
begin to polymerize. Using hexadecane or polymer as a costabilizer will facili-
tate one hundred percent droplet nucleation, while the use of cetyl alcohol does
not. Miller et al. [109] have shown that a small amount of polymer dissolved in
the monomer droplets enhances droplet nucleation.Also, the initiator flux must
be high enough to nucleate all of the droplets within a short time interval.
In summary, the miniemulsion route to polymer latexes should not be dis-
missed solely due to a requirement for narrow particle size distribution, par-
Miniemulsion Polymerization 189
4.1.3
Shear Stability
The shear stabilities of mini- and macroemulsion latexes were compared and
quantitatively evaluated with respect to their particle size distributions by
Rodrigues and Schork [146]. Although miniemulsion latexes exhibit many of
the properties of macroemulsion latexes, there may be subtle differences in
particle size distribution and surface characteristics due to differences in their
polymerization mechanisms. To study the effects of these differences on the
shear stabilities of the miniemulsions, a quantitative approach was developed
where changes in the average diameter and total number of particles have been
related to the particle size distribution before and after shearing.
Two pairs of MMA mini- and macroemulsion latexes were polymerized for
this study. HD was used as the costabilizer for the miniemulsions, and the
polymerizations were carried out at 60 °C. Efforts were made to make the main-
and macroemulsion in each pair as similar as possible. The two pairs were:
Pair I
Macroemulsion (Sample A)
0.02 mol. SLS/Laq; 0.0115 mol. KPS/Laq; 2 g DDM;
PSD range: 141–188 nm (diameter)
Miniemulsion (Sample E)
0.02 mol. SLS/Laq; 0.0115 mol. KPS/Laq; 2 g DDM;
PSD range: 96–123 nm (diameter)
Pair II
Macroemulsion (Sample H)
0.01 mol. SLS/Laq; 0.0115 mol. KPS/Laq; 4 g DDM
PSD range: 167–241 nm (diameter)
Miniemulsion (Sample D)
0.01 mol. SLS/Laq; 0.0115 mol. KPS/Laq; 4 g DDM
PSD range: 145–209 nm (diameter)
The samples were sheared using a rotational viscometer with a coaxial cylinder
system, based on the Searle-type, where the inner cylinder (connected to a
sensor system) rotates while the outer cylinder remains stationary. The outer
cylinder surrounding the inner one was jacketed, allowing good temperature
control, and the annular gap was of constant width. The sensor system used was
the NV type, with a rotor with a recommended viscosity range of 2¥103 mPa,
a maximum recommended shear stress of 178 Pa, and a maximum recom-
mended shear strain rate of 2700 s–1; this rotor could work with volumes from
10–50 ml. Flow was laminar.
190 F. J. Schork et al.
The optimum shear rate for each pair was arrived at by trial and error, such
that the shearing produced aggregation, but not massive coagulation. That is,
the shear rate was not increased beyond the point at which the particle dia-
meters stopped increasing (and maybe even started decreasing again). All
the tests were conducted at 25 °C; each miniemulsion latex within a pair was
sheared for the same time interval as its macroemulsion latex counterpart. Only
changes in particle sizes were followed during these series of experiments, with
the particle size distribution (analyzed by dynamic light scattering, DLS) re-
corded before and after shearing for each of the shear experiments.
The ratio of total number of particles initially present to the total number
of particles after shearing (Ni/Nf) was computed and used both to compare the
two types of latexes, and to determine the extent of aggregation and the nature
of the aggregates formed.
4.1.3.1
Relative Shear Stability of Miniemulsion and Macroemulsion Latexes
The particle size range and average particle diameter before and after shear-
ing, and the shear rate and time of shear used for each of the sample pairs is
shown in Table 13. For both pairs, the shifts in the particle size range and in the
average diameter are substantially greater for the macroemulsion latex than the
corresponding miniemulsion latex. In all cases, the particle size distribution
broadened after shearing. The percentage change in average diameter is greater
for the macroemulsions as well.
The ratio of the initial total number of particles to the final total number
of particles after shearing (N0/Nf) is shown in Table 14. It is clear that the
macroemulsion latexes showed greater shear instability. The ratio No/Nf gives
an indication of the extent of aggregation in each of the latexes; a No/Nf ratio
of two would indicate that average aggregation up to doublet formation has
taken place, and so on. It can be seen from Table 14 that aggregation has taken
place in all the latexes essentially up to doublet formation, with slightly higher
Table 13 Particle size ranges and average particle diameters before and after shearing, and
the shear rates and times of shear for samples A, E, H, and D (from [146])
Sample Shear Time of Initial PSD Average PSD range Average % Change
rate shear range diameter after diameter in average
(s–1) (s) (diameter) before shearing after diameter
(nm) shearing (nm) shearing (nm)
(nm) (nm)
Table 14 Ratio of the initial total number of particles to the final number of particles after
shearing for samples A, E, H, and D (from [146])
4.1.3.2
Effect of Large Particles
Table 15 Particle size ranges and average particle diameters before and after shearing,
and the shear rates and times of shear for samples A, E, H, and D with large particles added
(from [146])
Sample Shear Time of Initial PSD Average PSD range Average % Change
rate shear range diameter after diameter in average
(s–1) (s) (diameter) before shearing after diameter
(nm) shearing (nm) shearing (nm)
(nm) (nm)
that in all four latexes used here, the size differential between these latexes and
the two latexes used as large particles (Samples C and G) is not very large.
Therefore, the particles that qualify as large particles by definition lie at the
upper end of the particle size distribution of samples C and G; in other words,
only a small fraction of the externally-added larger latex actually functions as
large particles in terms of influencing shear stability.
The results shown in Table 16 support the observations drawn from Table 15.
For all of the latexes, there is an increase in No/Nf value as the percentage
of externally added large particles increases, up to approximately 5–10%wt
of large particles. Any further addition of large particles decreases the No/Nf
value.
Miniemulsion Polymerization 193
Table 16 Ratio of the initial total number of particles to the final number of particles after
shearing for samples A, E, H, and D with large particles added (from [146])
4.1.3.3
Summary
4.2
Monomer Transport Effects
4.2.1
Polymerization of Highly Water-Insoluble Monomers
4.2.2
Copolymer Composition Distribution
kp11 kp22
r1 = 6 r2 = 6 (16)
kp12 kp21
r1 f1
6 +1
dM1 F1 f2
7 = 4 = 02 (17)
dM2 F2 r2 f2
6 +1
f1
1 1
1+ 7 1+ 7
K2y K1y
r¢1 = r1 03 r¢2 = r2 03 (18)
1 1
1+ 7 1+ 7
K1y K2y
The Mayo Lewis equation, using reactivity ratios computed from Eq. 18, will
give very different results from the homogenous Mayo Lewis equation for mini-
or macroemulsion polymerization when one of the comonomers is sub-
stantially water-soluble. Guillot [151] observed this behavior experimentally for
the common comonomer pairs of styrene/acrylonitrile and butyl acrylate/vinyl
acetate. Both acrylonitrile and vinyl acetate are relatively water-soluble (8.5 and
2.5%wt, respectively) whereas styrene and butyl acrylate are relatively water-
insoluble (0.1 and 0.14%wt, respectively). However, in spite of the fact that
styrene and butyl acrylate are relatively water-insoluble, monomer transport
across the aqueous phase is normally fast enough to maintain equilibrium
swelling in the growing polymer particle, and so we can use the monomer parti-
tion coefficient.
Schuller’s equation is appropriate when one of the monomers has significant
water solubility and the other does not. In the case where one monomer is water
insoluble, and the other is extremely water insoluble, Schuller’s equation does
not hold. For the comonomer pair of MMA/VS (vinyl stearate), K1K2, so
Schuller’s equation predicts that, for macroemulsion polymerization, the poly-
mer formed early in a batch reaction will be richer in vinyl acetate (VA) than
in the homogenous case. The opposite was found by Reimers [102]. It has long
been accepted that monomer transport is not a rate-limiting step in the con-
ventional emulsion polymerization of relatively water-insoluble monomers,
such as MMA, styrene, and butyl acrylate (BA). However, the water solubility
of VS is as much as three orders of magnitude smaller than these typical emul-
sion polymerization monomers. In this case, VS cannot readily cross the
aqueous phase to saturate the growing polymer particles. (The major transport
resistance is actually from the monomer droplets into the aqueous phase.)
Therefore, Schuller’s partition coefficient model cannot be used here, since
it assumes that the particles are saturated with both monomers. The homo-
geneous polymerization model is not useful either since the monomer con-
centration in the particles is not identical to the bulk monomer concentration.
Therefore, a pseudo-partition coefficient k2 was proposed by Samer [137] for
extremely water-insoluble comonomers in order to interpret the experimental
data. k2 is not a partition coefficient and does not have any physical signi-
ficance.As defined by Samer, it is simply an adjustable parameter that replaces
K2 in Eq. 18. If K1 is set to infinity and k2 is set to unity, Eq. 17 and Eq. 18 can be
used to correlate the copolymer composition with the total monomer conver-
sion data in the MMA/VS (and other) systems.
Extremely water insoluble comonomers are only selectively used in emulsion
polymerization because of concerns about monomer transport limitations.
Miniemulsion Polymerization 197
4.2.2.1
Batch Copolymerization
Reimers [102] carried out batch copolymerizations in both macro- and equiv-
alent miniemulsions. MMA was used as the main monomer. The MMA was
copolymerized in macroemulsion- or miniemulsion with p-methylstyrene
(pMS), vinyl hexanoate (VH), vinyl 2-ethylhexanoate (VEH), vinyl n-decanoate
(VD) or vinyl stearate (VS). The comonomers were copolymerized at 10%wt
comonomer, 90%wt MMA. SLS was used as the surfactant and KPS as the
initiator. The comonomers (all highly water insoluble) were used as the co-
stabilizer. Miniemulsions were sonicated, while equivalent macroemulsions were
only subjected to vigorous mixing. Polymerizations were carried out at 60 °C.
Copolymer composition was obtained by integrating characteristic peaks in
the 1H NMR spectra. The integrated peaks correspond to signals produced by
the methyl ester protons (three) of MMA, 3.56 ppm, the aromatic protons (four)
of pMS, 6.7 ppm, and the a-proton of the vinyl esters at 4.8 to 5.6 ppm. The
relative ratios of the peaks allowed the mole fraction of each comonomer in the
copolymer to be assessed. Samples were taken throughout the reaction at
different times to monitor the incorporation of comonomer as a function of the
extent of total monomer conversion. These results were then compared with the
results for average comonomer fraction given by an integrated copolymer
(Mayo Lewis) equation [149], using r1 and r2 from bulk polymerization. The
mole fractions used were based on the total moles of monomer in the droplets
and were corrected for the water solubility.
Both the mini- and macroemulsion copolymerizations of pMS/MMA tend
to follow bulk polymerization kinetics, as described by the integrated copoly-
mer equation. MMA is only slightly more soluble in the aqueous phase, and
the reactivity ratios would tend to produce an alternating copolymer. The mini-
emulsion polymerization showed a slight tendency to form copolymer that
is richer in the more water-insoluble monomer. The macroemulsion formed
a copolymer that is slightly richer in the methyl methacrylate than the co-
198 F. J. Schork et al.
4.2.2.2
Semibatch Copolymerization
emulsifier SLS was used as the initiator. The semibatch miniemulsion poly-
merization involved two stages, a miniemulsion batch stage and semibatch
stage. The miniemulsion batch stage was performed as above. When the
monomer conversion was estimated to be about 80%, the feeding stage was
initiated by pumping monomer emulsion (miniemulsion or macroemulsion,
which was continuously formed while feeding) and initiator solution at set
flow rates simultaneously into the reactor. The composition of the monomer
feed was identical to the composition of the monomer in the batch stage.
Copolymer composition was determined by NMR, MWD by GPC and PSD by
dynamic light scattering.
Miniemulsion and macroemulsion copolymerizations of VAc and
comonomers with extremely different physical and kinetic properties (BA,
NMA and DOM), were investigated in batch and semibatch systems. The results
for polymer particle size and number, monomer conversion, composition and
molecular weights of copolymers indicated that there was an obvious diver-
gence between macroemulsion and miniemulsion copolymerization. In all
cases, the particle size was smaller and the particle number was higher in
macroemulsion copolymerization than in miniemulsion copolymerization. For
the systems VAc/BA and VAc/DOM, the particle number increased with in-
creasing conversion throughout the reaction for both batch macroemulsion
and miniemulsion runs. This was taken to indicate that the nucleation of new
particles takes place via homogeneous nucleation throughout these reactions.
For the batch runs, the rate of polymerization of the macroemulsion polymer-
ization runs was faster than that of the miniemulsion.
An investigation of the copolymer composition demonstrated the impor-
tant effect of monomer transport on the copolymerization. The droplets in
the macroemulsion act as monomer reservoirs. In this system, the effect of
monomer transport will be predominant when an extremely water-insoluble
comonomer, such as DOM, is used. In contrast with the macroemulsion system,
the miniemulsion system tends to follow the integrated Mayo Lewis equation
more closely, indicating less influence from mass transfer.
Likewise, for the semibatch operation, the influence of monomer was seen
in the differences between macro- and miniemulsion feeds. For extremely wa-
ter-insoluble monomers, the miniemulsion-feed mode lessens the departure of
the copolymer composition from the feed composition during semi-starved
semibatch polymerization. However, this is accomplished by simultaneously
broadening the PSD. Results from the GPC analysis indicated that the polymers
with lower molecular weight and broader distribution were formed in the
semibatch process, in contrast to the batch run.
Wu [159] also investigated the miniemulsion and macroemulsion copoly-
merization of VAc and vinyl versitate (VEOVA). At room temperature, the
water solubility of VAc is 2.58%wt, and vinyl versitate (also known as neo-
decanoate, one of the isomers in VEOVA-10) is 7.5¥10–4%wt. The extreme dif-
ference in water solubility between the two comonomers may impact on co-
polymer composition and the properties of the final polymer, due to the mass
202 F. J. Schork et al.
4.2.2.3
Copolymerization in a Continuous Stirred Tank Reactor
4.2.3
Interfacial Polymerization
In order to gain evidence for interfacial initiation, the redox initiator system
was compared with a water-soluble initiator (VA-044) in terms of the emulsion
polymerization behavior of butyl acrylate (BA)/[2-(methacryloyoxy)ethyl]tri-
methyl ammonium chloride (MAETAC). It was found that for the water-solu-
ble initiator system, only homopoly(MAETAC) was formed and BA did not
polymerize at all. In the case of VA-044, it was suggested that it may be difficult
for polymeric free radicals in the aqueous phase to penetrate the viscous
surfactant layer to initiate the polymerization of the BA monomer. On the other
hand, it has also been found that BA could be rapidly polymerized under the
same conditions if VA-044 is replaced with CHP/TEPA, indicating that radicals
are formed in the interface, where they do not need to penetrate through vis-
cous surfactant layer.
The polymerization kinetics of BA/MAETAC macroemulsion and mini-
emulsion copolymerization was investigated with the interfacial redox initiator
system. It was found that adding MAETAC had a complex effect on the poly-
merization kinetics of BA, as shown in Figs. 17 and 18 [170].
In comparison with the homopolymerization of BA, adding 5%wt MAETAC
greatly increases the BA polymerization rate for both macroemulsion and
miniemulsion polymerization. However, when the MAETAC level is increased
further, the polymerization rate begins to decrease. At higher MAETAC levels,
the effect of MAETAC is different for macroemulsion and miniemulsion
polymerizations. For macroemulsion polymerization at high MAETAC level,
there seems to be an induction period before polymerization begins.At higher
MAETAC levels, the length of the induction period increases. For miniemulsion
polymerization, there is no induction period; instead, the BA polymerization
rate decreases to zero at less than full conversion. BA homopolymerization with
CHP/TEPA levels off at ~30% conversion. At higher MAETAC levels, the BA
conversions levels off at substantially higher conversion. The leveling-off has
been ascribed to the depletion of TEPA. Comparing Fig. 17 with Fig. 18, it is
clear that the influence of MAETAC levels on macroemulsion polymerization
is much greater than on miniemulsion polymerization, especially in the early
stage of polymerization. With increasing MAETAC levels, the polymerization
rate of BA, especially in the early stage of polymerization, rapidly drops off
in the macroemulsion polymerization. It is suggested that adding MAETAC
would interfere with the nucleation of macroemulsion polymerization. At low
MAETAC levels, introducing MAETAC leads to homogeneous nucleation, so
the polymerization rate increases. However, at high MAETAC level, adding
MAETAC did not result in homogeneous nucleation but suppressed micellar
nucleation. It has been suggested that polymer formed in the aqueous phase
is too hydrophilic to lead to homogeneous nucleation at high MAETAC. The
suppression of micelle nucleation can be illustrated by Fig. 19. The hydrophobic
molecules such as BA and CHP will be present in the micelles at the beginning
of the polymerization. When TEPA was added to the reactor, many of the free
radicals will be formed by the following mechanism in the surface of a micelle,
where CHP and TEPA meet each other [173].
Because of their different hydrophilicities, the two free radicals formed at the
same time can separate from each other quickly which can eliminate the cage
effect. In a micelle, the local BA concentration may be quite high. Once a micelle
is initiated, a number of BA molecules may be added quickly. As a result, some
short BA blocks would be incorporated into a poly(MAETAC) chain to form
something like multi-block copoly(MAETAC-BA), as shown in Fig. 19 [170].
Surfactant should stabilize the BA blocks so that the block copolymer remains
in the aqueous phase.
The BA blocks in the copolymer, which is surrounded by surfactant, swell
with BA monomer, which polymerizes there until particles form. Therefore,
micellar nucleation is diminished or even eliminated. As a result, micellar
nucleation is retarded, and Np decreases. It is clear that the nucleation is very
Miniemulsion Polymerization 207
4.3
Multiphase Particles
4.3.1
Hybrid Miniemulsion Polymerization
4.3.1.1
Alkyds
Water-based coatings have become more widely used over the past few decades
because they are environmentally friendly, offer easy clean-up, and their pro-
perties and application performance characteristics have improved. Solvent-
based systems such as alkyd resins and polyurethanes have remained impor-
tant for some applications because of superior properties, such as gloss and
hardness. This is due to the curing mechanism of oil-based coatings in which
the oils react with atmospheric oxygen to form very hard crosslinked materials.
This mechanism is generally lacking in water-based coatings, which tend to
be soft and pliable, due to the fact that the coatings are made soft to allow film
formation, and since there is no curing chemistry available, remain soft on
drying. Several researchers have focused on the use of the hybrid miniemulsion
polymerization of acrylic monomers in the presence of alkyd and polyurethane
resins to develop alternative coatings which have the advantages of water-based
systems (like low VOC) with the drying (air cure) properties of solvent-based
systems.Alkyd/acrylate coatings are targeted as replacements for solvent-based
architectural coatings, and oil-modified polyurethane (OMPU)/acrylate coat-
ings may provide a low VOC alternative to solvent-based clear coats. Since
U.S. architectural coating sales in 1995 amounted to 625 million gallons [174],
a conservative estimate of the VOC reduction if all of these coatings were water-
based is approximately 500 million pounds of solvent that would not be re-
leased into the air.
Nabuurs and German [175] developed an alkyd-acrylic hybrid system via
emulsion polymerization. They were able to produce a stable product using
MMA as the acrylic.When the alkyds were functionalized by sulfonation, there
was no evidence of heterogeneity in the particles.When unfunctionalized alkyd
was used, the MMA appeared as microdomains within the particles. Grafting
of acrylic to alkyd was low. The presence of the alkyd led to low rates of poly-
merization and limited conversion that were both attributed to retardation
through radical delocalization following radical transfer to the unsaturated
groups in the fatty acids of the alkyd. Although this work was not reported to
be miniemulsion polymerization, the emulsions were subjected to high shear
prior to polymerization. No costabilizer was added, but the alkyd presumably
functioned as such. Since the alkyd-acrylic droplet size was probably quite
small, droplet nucleation was probably the dominant nucleation mechanism.
This would explain the good colloidal stability of the resulting system, and the
fact that both alkyd and acrylic domains were found in the same particles.
Wang et al. [98] carried out macroemulsion and miniemulsion polymeriza-
tion of acrylic monomers in the presence of alkyd resins. Miniemulsion and
macroemulsion polymers were produced using a commercial medium soya-
linseed alkyd and a mix of acrylic monomers consisting of 50% BA, 49% MMA,
and 1% acrylic acid (AA). PMMA polymer with a weight average molecular
weight of 100,000 was used as the costabilizer. Alkyd levels were 5, 30, 60 or
210 F. J. Schork et al.
100% based on total acrylic monomer. SLS was used as the surfactant, and
sodium persulfate (SPS) was used as the initiator. The miniemulsions were pre-
pared by dispersing the desired amount of monomer-PMMA-alkyd solution in
the aqueous SLS solution by mixing with stirring at room temperature. The re-
sulting emulsion was sheared further by sonication. Polymerization was carried
out at 60–80 °C. The reaction was followed by gravimetric conversion analysis.
Macroemulsion polymerizations were carried out in the same manner except
that no sonication process was used and the PMMA costabilizer was not em-
ployed. Droplet and particles sizes were determined by dynamic light scatter-
ing, and double bond content of the alkyd by NMR. Grafting was determined
by extraction, and degree of crosslinking by exhaustive extraction.
The monomer miniemulsions with PMMA as costabilizer were prepared
with different amounts of alkyd resin. The PMMA costabilizer was effective in
the preparation of stable miniemulsions, especially in conjunction with the
alkyd. The size of monomer droplets was below 300 nm. After five days, the
unpolymerized macroemulsions with alkyd separated into three phases,
monomer on the top, clear water in the middle, and alkyd resin on the bottom.
The miniemulsion without alkyd showed two phases, monomer and water.
All miniemulsions with alkyd resin appear to remain uniform. Very stable
miniemulsions were obtained when the alkyd content was higher than 30%.
The shelf life of macroemulsions was only 2–8 minutes.
The polymerization rate in the presence of alkyd was slower than that with-
out alkyd. Doubling the initiator and emulsifier concentration increased the
reaction rate, but not to the level achieved with the miniemulsion polymeriza-
tion without alkyd. This retardation (as reported also by Nabuurs) increased
with increasing alkyd level. The latexes obtained from the miniemulsion poly-
merization of the alkyd-acrylate mixtures were uniform emulsions, and no
coagulation occurred during polymerization. Macroemulsion polymerization
with alkyd resulted in colloidal instability, probably due the inability of the
alkyd to reach the locus of polymerization.
NMR analysis indicated that approximately 30% of the alkyd double bonds
had been consumed in the polymerization process. This is important in that it
indicates that a substantial fraction of the double bonds remain for oxidative
crosslinking while a coating made from this material is dried. Selective ex-
traction indicated that approximately 60% of the acrylate was grafted to alkyd.
Exhaustive extraction indicated less than 5% crosslinked material. The poly-
merized latex formed good films with acceptable hardness.
On the whole, the miniemulsion polymerization process proved to be effec-
tive for incorporating an alkyd resin into acrylic coated copolymers. The reac-
tion produced stable, small particle size latexes that contain graft copolymer of
the acrylic and alkyd components. Attempts at macroemulsion hybrid poly-
merization were unsuccessful.
Van Hamersveld et al. [176, 177] carried out hybrid miniemulsion poly-
merization of MMA, using HD as the costabilizer. In an attempt to encourage
grafting, oxidized triglycerides (such as sunflower oil) were used as initiators.
Miniemulsion Polymerization 211
4.3.1.2
Polyester
4.3.1.3
Polyurethane
4.3.1.4
Other Hybrids
El-Aasser and coworkers [185, 186] have carried out miniemulsion polymeriza-
tion of styrene monomer in the presence of Kraton D1102 thermoplastic
elastomer, to form hybrid composite latexes approximately 100–150 nm in size.
A costabilizer other than the rubber was used. The miniemulsification was
carried out via homogenization, and resulted in a very broad droplet size dis-
tribution. This resulted in the formation of inhomogeneous hybrid composite
particles due to monomer diffusion during polymerization. When an oil-
soluble initiator was used, an induction period was observed, resulting from the
presence of radical scavengers such as antioxidants and UV stabilizers within
the Kraton-styrene droplets. This also lead to inhomogeneous particles, but was
eliminated with a water-soluble initiator. TEM showed domains of polystyrene
within the rubber particles. Some evidence of homogeneous (or micellar)
nucleation was found.
Kawahara et al. [187] prepared acrylic/epoxy composite latexes via hybrid
miniemulsion polymerization. Landfester et al. [188] have incorporated PMMA
Miniemulsion Polymerization 215
4.3.1.5
Artificial Miniemulsions
4.3.2
Nanoencapsulation
4.4
Controlled Free Radical Polymerization
4.4.1
Nitroxide-Mediated Polymerization
4.4.1.1
Mechanism
Since the nitroxide and the carbon-centered radical diffuse away from each
other, termination by combination or disproportionation of two carbon-cen-
tered radicals cannot be excluded. This will lead to the formation of “dead”
polymer chains and an excess of free nitroxide. The build-up of free nitroxide
is referred to as the Persistent Radical Effect [207] and slows down the poly-
merization, since it will favor trapping (radical-radical coupling) over propa-
gation. Besides termination, other side reactions play an important role in
nitroxide-mediated CRP. One of the important side reactions is the decompo-
sition of dormant chains [208], yielding polymer chains with an unsaturated
end-group and a hydroxyamine, TH (Scheme 3, reaction 6).Another side reac-
tion is thermal self-initiation [209], which is observed in styrene polymeriza-
tions at high temperatures. Here two styrene monomers can form a dimer,
which, after reaction with another styrene monomer, results in the formation
of two radicals (Scheme 3, reaction 7). This additional radical flux can compen-
sate for the loss of radicals due to irreversible termination and allows the poly-
4.4.1.2
Effects Of Segregation And Heterogeneity
4.4.1.3
Results
A review article by Qiu et al. [212] and references herein [217–226] covers
NMCRP in miniemulsions up to 2001. Cunningham wrote a related review
in 2002, also covering controlled radical polymerization in dispersed phase sys-
tems [227]. Here, the main results reported in the Qiu review will be sum-
marized, and new developments in the field since then will be reviewed.
For several reasons, miniemulsion polymerization is the preferred technique
for NMCRP in aqueous dispersed systems. Since NMCRP in general requires
high reaction temperatures (above 100 °C) for the thermal polymerization of
220 F. J. Schork et al.
styrene, the presence of a monomer phase is not desired. This would lead to
colloidal instability. Furthermore, the presence of a large monomer reservoir
makes the nitroxide partition out of the loci of polymerization, which will
reduce control of the polymerization. Another reason for using miniemulsion
polymerization is the fact that pre-formed alkoxyamines or nitroxide-termi-
nated oligomers are often used. Generally, these are too water-insoluble to be
transported through the aqueous phase, which excludes the use of conventional
or seeded emulsion polymerization. Finally, the poorly understood and
complex particle nucleation step is avoided in miniemulsion polymerization,
which allows the use of oil-soluble initiators and makes results easier to
understand.
The first results for NMCRP in miniemulsion were reported by Propdan
et al. [218, 221] and Macleod et al. [219]. Both groups performed miniemulsion
polymerizations of styrene at high temperatures (above the boiling point of
water) using high pressure reactors. Propdan et al. used oil-soluble benzoyl
peroxide (BPO) free-radical initiator and TEMPO, 1 (see Scheme 4), in a Dow-
fax 8390 surfactant system at 125 °C, while Macleod et al. used water-soluble
potassium persulfate (KPS) initiator and sodium dodecylbenzene sulfonate
(SDBS) surfactant at 135 °C. The Propdan system resulted in stable latexes and
90% conversion was reached in 12 hours. The polydispersity was 1.15–1.60 and
molecular weights were up to 40 kg/mol. The Macleod system, if a proper
nitroxide/initiator ratio was chosen, also gave stable latexes with good control,
with polydispersities in the range of 1.1–1.2. Conversion reached 87% in six
hours, although a later publication [225] showed that this relatively fast poly-
merization also resulted in a large proportion of dead chains.
Later on, these same research groups started using pre-formed TEMPO-
terminated polystyrene as a one component initiator system instead of a bi-
component nitroxide/initiator system. Pan et al. [224] prepared TEMPO-termin-
ated polystyrene in bulk and isolated this to use it as the initiator in their
miniemulsions. This led to slower polymerization rates, molecular weights
lower than predicted and relatively broad molecular weight distributions.
Keoshkerian et al. [225], on the other hand, reported very high conversions in
six hours (99.6%) and narrow polydispersities (1.15) by preparing TEMPO-
terminated polystyrene in bulk up to a conversion of about 5% and applying
Miniemulsion Polymerization 221
4.4.2
Atom Transfer Radical Polymerization
4.4.2.1
Mechanism
The initiators used are typically alkyl halides with similar structures to that
of the monomer being employed. Of the halogens, chlorine and bromine have
been shown to produce the best overall results with respect to molecular weight
control [256]. However, iodine works well in certain cases, for example, rhe-
nium-mediated styrene polymerizations [246] and copper-mediated acrylate
polymerizations [257]. The C–F bond with fluorine is too strong to cleave homo-
lytically, ruling out its use as an effective initiator. In general, any alkyl halide
with activating substituents on the a-carbon (aryl, carbonyl or allyl groups) can
act as an ATRP initiator. In addition, polyhalogenated compounds (like CCl4
or CHCI3) and compounds with a weak R–X bond (like N–X, S–X) can also
be used.
There are several requirements that are generally recognized as essential to
an effective ATRP catalyst [256, 258, 259, 260]. The metal center should be able
to assume at least two oxidation states, separated by one electron, like Cu(I) and
Cu(II). It should also be attractive to halogens, it should possess an expandable
226 F. J. Schork et al.
coordination sphere such that when oxidized it can contain the halogen, and it
should have a low affinity for alkyl radicals and the hydrogen atoms on alkyl
groups.
Conventional free radical polymerizations in miniemulsions benefit kineti-
cally from the effects of radical segregation. In solution, any radical could ter-
minate with another theoretically. However, when the radicals are segregated
into isolated reaction loci (such as miniemulsion droplets or particles) termi-
nation is no longer possible. Because the total concentration of radicals is
distributed throughout the particles, the probability that any two radicals will
terminate bi-molecularly is greatly reduced. The ideal situation is one in which
a lone radical in a particle can terminate only with a radical that enters the
particle. This is known as the “zero-one” limit [121], and in this case the rate of
bimolecular termination is controlled by radical entry alone. In the absence of
other effects, this lowering of the incidence of bi-molecular termination events
tends to increase the overall rate of polymerization while simultaneously nar-
rowing the molecular weight distribution. However, because of the mechanism
involved, the same benefit is not seen with ATRP in miniemulsion. With ATRP,
the dominating rate-controlling factor is the equilibrium between the dormant
and active species. Since the equilibrium heavily favors the dormant species, the
lifetime of the active species is extremely short. As such, the concentrations of
these active radicals are always minute [261]. Therefore, the probability that a
water phase radical will enter a particle containing another radical and termi-
nate is exceedingly low. Any kinetic benefit that might otherwise be gained
from segregation is overshadowed by this low radical concentration. It should
also be noted that, because of the ATRP mechanism, increases in reaction rate,
whatever their origin, will come at the expense of the deactivator species. The
resulting decrease in the concentration of deactivator tends to produce broader
polydispersities.
4.4.2.2
Results
all rate was much slower than with AIBN. Additionally, the progression of Mn,
though linear, was much higher than found theoretically. In this case, the ini-
tiator decomposed faster but the radicals formed tended to terminate in the
aqueous phase, contributing both to the slower overall rate and the higher than
predicted actual molecular weights. The authors performed direct ATRP with
an oil-soluble initiator, ethyl 2-bromoisobutyrate (EBiB). At a lower tempera-
ture than the reverse process, 70 °C vs 90 °C, the polymerization rate of the
direct process was significantly faster. The polydispersity remained relatively
flat throughout the experiment, with a final value of approximately 1.3. The
molecular weight evolution, though roughly linear, was higher than predicted.
Additionally, a semilog plot of the conversion data revealed some curvature,
indicating a larger than expected number of chain termination events. It was
postulated that the cause lay in the partitioning of the Cu(II) species into the
aqueous phase. Recent studies with ATRP in aqueous dispersions lend credence
to this argument [263]. They observed that the polymerization rate was insen-
sitive to the size and number of particles and controlled entirely by the atom
transfer equilibrium. The researchers also studied the effect of removing the
costabilizer, hexadecane, from the recipe in order to determine if the hydro-
phobic dNbpy ligand alone would act as a sufficient droplet stabilizer. However,
it was noted that the droplet size increased dramatically in the absence of hexa-
decane, indicating that the osmotic pressure would be insufficient to prevent
Ostwald ripening.
Li and Matyjaszewski [264], building on the earlier work mentioned here
[262], conducted reverse ATRP in miniemulsions using n-butyl methacrylate
(BMA) with a more active catalyst system and a faster initiator. The solids
content was roughly double that of the previous effort, jumping from approx-
imately 13% to over 20%. More importantly, the surfactant (Brij 98) concen-
tration was reduced from 13.5%wt to 2.3%wt based on monomer, decreasing
the likelihood of micellar nucleation. Because of their high activities in bulk
and solution ATRP, complexes of hexasubstituted tris(2-aminoethyl)amine
(TREN) with Cu/Br2 were utilized as the metal activator/deactivator. The water-
soluble initiator used was 2,2¢-azobis[2-(2-imidazolin-2-yl)propane] dihydro-
chloride (VA-044). The VA-044 was chosen for its fast decomposition rate,
which facilitates a well-controlled ATRP and contributes to colloidal stability.
The authors looked at the effect of the ligand and noted that there was no
induction period with the TREN ligand as compared to dNbpy. While linear
progressions of Mn with conversion were seen, the authors found that the mole-
cular weight was not quantitative in initiator concentration. Instead Mn tended
to be much higher than that calculated based on the initiator concentration and
more closely followed a trajectory calculated from the initial concentration of
deactivator, Cu(II).Assuming 100% initiator efficiency, the authors explain the
deviation in terms of an excess radical concentration over the concentration of
Cu(II), leading to the generation and subsequent termination of some of the
oligomeric chains. Since a 1:1 molar ratio of [Cu(II)]0/[I]0 was used in most
of the experiments, in theory there would be two initiator radicals for every
228 F. J. Schork et al.
4.4.3
Reversible Addition Fragmentation Polymerization
4.4.3.1
Mechanism
The third (and also the most recently developed) controlled free radical tech-
nique discussed in this review is RAFT. In 1998 Rizzardo et al. published a novel
“controlled” free-radical polymerization technique, which they designated the
RAFT process [265–267] because the mechanism involves Reversible Addition-
Fragmentation chain Transfer. This technique allowed the production of poly-
mer with a narrow molecular weight distribution. In fact, this concept was not
entirely new, and stemmed from the same researchers’ previously published
work to produce block copolymers using methacrylate macromonomers as
reversible addition-fragmentation chain transfer agents in 1995 [268]. However,
these macromonomers were not very effective RAFT agents. The breakthrough
came with the discovery of a more reactive double bond species, S=C(Z)SR.
During styrene polymerization, the propagating radicals were very reactive
to the dithioesters and to a much lesser extent to the xanthates [269]. A brief
description of the RAFT process is given below, and a schematic representation
is given in Scheme 7.
A conventional free-radical initiator is added (contrary to some other con-
trolled free-radical polymerization techniques) that generates radicals, which
can add either to the monomer or the S=C moiety of the RAFT agent (step 1). In
most cases the addition of small carbon-centered radicals to the RAFT agent is
rapid and is not rate determining. Therefore, step (1) involves polymeric radical
addition to 1 to form an intermediate radical species 2 that will fragment back
to the original polymeric radical species or fragment to a dormant species 3
Miniemulsion Polymerization 229
and a small radical, R•. R• can then further propagate to form a polymeric rad-
ical (step 2 in Scheme 7), rather than adding to 3. The dormant polymeric RAFT
agent acts in a similar way to a RAFT agent, so growing polymeric radicals can
also add to the dithiocarbonyl double bond of the polymeric RAFT agent,
thereby forming an intermediate radical 4. This intermediate has an equal
probability of fragmenting back into its starting species or into a dormant poly-
meric RAFT agent and a polymeric radical, in which the dithiocarbonate
moiety has been exchanged between the active and dormant polymer chains
of the starting species. This equal probability of fragmenting to either side of
the equilibrium is a result of the symmetry of 4. There might be a difference in
the chain length of both sides, but this will not have an effect, unless one of the
two sides is extremely short. This mechanism of addition of radicals to the
dithiocarbonyl double bond and fragmentation of the intermediate was shown
by Moad et al. [270], who observed the intermediate radical directly by ESR.
Overall, polymer chains with a dithiocarbonate end-group are formed. If
addition to the dithiocarbonyl double bond is fast compared to propagation,
230 F. J. Schork et al.
4.4.3.2
Effects of RAFT and Transfer Agents on Emulsion Polymerization Kinetics
As far back as 1948, Smith and Ewart [13] included the effects of radical desorp-
tion in emulsion polymerization kinetics, and in 1965 Romatowski et al. [282–
284] showed that radicals resulting from chain transfer to monomer indeed
escape from the particles.
Nomura et al. [285] and Lichti et al. [286] studied the effects of transfer
agents on the kinetics of ab initio and seeded emulsion polymerization of
styrene, respectively. Nomura et al. found that the polymerization rate per
particle decreased with increasing amounts of carbon tetrachloride, carbon
tetrabromide and primary mercaptans, and that the effects were stronger when
Miniemulsion Polymerization 231
the transfer constant or the water-solubility was higher. Lichti et al. observed
the same in seeded experiments with carbon tetrachloride and carbon tetra-
bromide as chain transfer agents. They were the first to actually measure the
exit rate coefficient, using g-radiolysis relaxation data. They found an increas-
ing exit rate coefficient with increasing amounts of transfer agents and a higher
exit rate coefficient when the transfer constant was higher. They also found that
the entry rate coefficient increased with increasing amounts of carbon tetra-
bromide. Due to the counterbalancing effects of an increased exit rate and an
increased entry rate the polymerization rate passed through a minimum.
Maxwell et al. [287] extended their own model for entry by taking the effect
of transfer agent into account and used this model to explain the increase
observed in emulsion polymerizations with monomers with a high critical
chain length z and thiols of intermediate chain length. They also used this
model to show that longer chain thiols are too water-insoluble to have an effect
and that short chain thiols might suffer aqueous phase termination and
increase the exit rate, and so they can reduce the polymerization rate instead
of increasing it. No effect was expected for styrene from Maxwell’s model,
which was confirmed by the work of Asua et al. [288]. They found that n-
dodecyl mercaptan had no effect on the polymerization rate.
The work of Monteiro et al. [289] showed that when RAFT agents are applied
in emulsion, the rate of polymerization is significantly retarded. This effect
is stronger when a RAFT agent with a more water-soluble leaving group is used.
Exit from the particles after fragmentation was proposed to be the main
reason for the observed retardation. Because of the high reactivity of the RAFT
agents used, it is expected that all of the RAFT agent is consumed after a few
percent conversion, and so it should no longer should have an effect. However,
it was observed that the rate of polymerization decreased with increasing RAFT
in interval II. Monteiro et al. claimed that this was due to transport limitation
of RAFT from the monomer droplets to the particles, meaning that there is a
constant flux of RAFT agent to the particles, even if all of the RAFT agent has
been consumed in the particles. Therefore, not all of the chains start to grow
simultaneously, resulting in broad polydispersities. However, retardation was
also observed in Interval III, and this could not be ascribed to exit and trans-
port limitations.When this work was published, intermediate radical termina-
tion [275, 276] had not yet been put forward by Monteiro et al. as a source of
retardation. However, since the system might not be under zero-one conditions
in Interval III due to the increased particle size, intermediate radical termina-
tion might explain these results.
Another observation Monteiro et al. made was that a red layer was observed
during Interval II, consisting of low molecular weight dormant chains, swollen
with monomer.At the crossover to Interval III, this red layer coalesced, forming
red coagulant. The same red layer was also observed by De Brouwer et al. [290]
in miniemulsions stabilized with ionic surfactants. When polymer was used
as the so-called cosurfactant, this polymer was not present in the red layer,
indicating that this layer was not due to droplet coalescence.Also, the use of an
232 F. J. Schork et al.
oil-soluble initiator did not reduce the formation of the red layer. Using a
higher radical flux (to enhance droplet nucleation) did not have an effect.
Indeed, the formation of the red layer was correlated to the polymerization rate,
which indicates that the product formed during the polymerization plays a
crucial role in the destabilization. However, when nonionic surfactants were
used, destabilization did not occur, and controlled miniemulsion polymeriza-
tions could be performed without destabilization. Later, Luo et al. suggested
that the instabilities are the result of the large number of oligomers formed in
the early stages of RAFT miniemulsions, causing a superswelling state [128],
which could be prevented by increasing the amount of cosurfactant.
Moad et al. [291] showed that the type of RAFT agent is important. Using a
very reactive RAFT agent (with a transfer constant of about 6000), similar to
that used in the work of De Brouwer and Monteiro, resulted in a broad poly-
dispersity in ab initio styrene polymerizations with ionic surfactant, which was
ascribed to the fact that the RAFT agent was not uniformly dispersed in the
polymerization medium. The use of less reactive RAFT agents (with transfer
constants of 10–30) did not result in destabilization and the final polymer had
a polydispersity close to 1.4.
Prescott et al. [292] used acetone to transport a water insoluble RAFT agent
to the seed particles. The polymerization was initiated in Interval III after
removing the acetone. No destabilization was observed, which according to our
previous discussion might indicate that the transfer constant of the RAFT agent
used was not extremely high. However, it was high enough to result in a linear
increase in molecular weight, and polydispersities between 1.2 and 1.4. Al-
though the RAFT agent is consumed at the beginning of the reaction (the mole-
cular weight follows the theoretical linear increase), a reduction in rate is ob-
served throughout the reaction. In these experiments, a small seed was used
and the amount of monomer was such that the particle size does not increase
much, which means that the system is likely to be under zero-one conditions
throughout the polymerization and so intermediate radical termination can-
not explain the retardation observed.
Monteiro et al. [293] also studied the effect of xanthates (RAFT agents with
low transfer constants) with styrene, in ab initio styrene polymerizations.
Again rate retardation was observed throughout the polymerization. This is not
surprising, since the low transfer constants of these RAFT agents mean that
they are present during the whole polymerization, which results in an increased
exit rate throughout the reaction.
This was later confirmed by Smulders et al. [294], who experimentally deter-
mined the exit rate in similar systems using g-relaxation experiments. The exit
rate was found to increase linearly with the RAFT concentration, although the
decrease in rate could not be ascribed to the increase in exit rate alone.
Summarizing, we know that RAFT can be applied in emulsion, although the
mechanism for this is not yet fully understood. Highly reactive RAFT agents
can lead to destabilization, although the use of nonionic surfactants seems to
prevent this destabilization.Rate retardation is observed in all cases. This can
Miniemulsion Polymerization 233
be partly ascribed to the increased exit rate, although the retardation is still
observed even when all of the RAFT agent has been consumed. In that case,
intermediate radical termination might explain the reduction in rate. However,
even when the system is under zero-one conditions and all RAFT is consumed,
retardation still occurs. This cannot be ascribed to intermediate radical termi-
nation or to exit. In fact, an explanation for this might be quite simple, as shown
by Smulders [295]. Retardation with RAFT in zero-one systems in which all
of the RAFT has been consumed cannot be ascribed to increased exit rate any-
more, since the leaving groups of the dormant polymer chains cannot exit.
Intermediate radical termination is also not a dominant mechanism, since each
particle contains only one radical. However, the fact that each particle contains
only one radical explains why retardation is observed in these systems. This
one radical is either present as a “normal” radical, R•, capable of propagating
and so consuming monomer, or as a “intermediate” radical, I•.While the radical
is in the intermediate state it does not consume monomer, which in turn leads
to retardation. Since the system is under zero-one conditions, the system does
not reach steady state at the microscopic level (inside a particle), because a
particle contains either no radical, one “normal” radical, or one “intermediate”
radical. The lifetimes of R• and I• are given by:
1 1
t R• = 0005 t I • = 91 (19)
kadd [dormant chains] 2k–add
tI• =15.2 s. This means that zero-one polymerization should not proceed,
because the radicals are present as intermediate radicals for more than 99.99%
of time. Since we have experimental evidence that RAFT systems do proceed
under zero-one conditions [292], these rate parameters seem highly unlikely,
although it should be noted that Prescott et al. used a RAFT agent with a less
stable intermediate. Zero-one experiments with dithiobenzoate RAFT agents
might be the key to closing the “six-orders-of-magnitude-gap” for the frag-
mentation rate constant.
4.4.3.3
Application of RAFT in Miniemulsion
4.4.4
Colloidal Stability
emulsion polymerization has much bigger particle size and much broader
particle size distribution than that obtained using regular miniemulsion poly-
merization. One can also see that colloidal instability is very sensitive to the
polymerization recipe. Monomer, surfactant, dormant agent, and levels of
surfactant and costabilizer all have large influences on the stability. High levels
of surfactant and costabilizer, and use of a nonionic surfactant and an oligo-
meric control agent all proved to aid stability, although larger particle sizes and
broader particle size distributions were still seen. On the other hand, the litera-
ture indicates that instability is a general problem in controlled free radical
miniemulsion polymerization, regardless of the living control mechanism,
monomer, surfactant system (except for polymeric surfactant), or initiator sys-
tem. It seems reasonable to assume that the instability is caused by the “living”
nature of the systems. In a controlled free radical polymerization system, the
kinetics of polymer chain formation is totally different from that for classical
free radical polymerization. In classical free radical polymerization, a polymer
chain is fully polymerized in about 1 s. At the very beginning of polymeriza-
tion, a few polymer chains of high molecular weight are formed. During poly-
merization, monomer is consumed to form more and more large polymer
chains. However, in controlled free radical polymerization, a large number of
oligomers are formed at the beginning of polymerization. During polymer-
ization, the oligomers gradually grow into large polymers. The feature common
to all of the controlled free radical systems is the presence of large concentra-
tions of oligomers early in the polymerization. In miniemulsion polymeriza-
tion, polymerization occurs in the particles (around 100 nm in size). Ugelstad
et al. [40] showed that oligomers are very efficient swelling agents, and
hence the existence of oligomers may dramatically modify the state of the
miniemulsion. A theoretical model has been developed by Luo et al. [128] to
simulate the swelling of oligomers formed in the controlled miniemulsion poly-
merization.
The chemical potential of monomer droplets is determined by [128]
–
1 2 c +
2V1
µd = RT (ln jd1 + 1 – 5 jd2 + j d2 12 RT (20)
m2 g rd
6
1
m2 1
µd = RT ln jd1 + 1 – 5 jd2 + 1 – 5 jp3 + j 2p2 c12 + j p3
m3
2
c13
–
c13 2V1
+ jp2 jp3 c12 + c13 – 5 + 6 RT (21)
m2 g rd
Miniemulsion Polymerization 239
Fig. 21 Variation in the droplet chemical potential (top) and particle chemical potential
(bottom) during particle swelling (from [170])
The monomer chemical potential in the particles is lower than that in the
droplets, so that the monomer in the droplets will diffuse across the aqueous
phase and into the particles, leading to changes in the monomer chemical po-
tentials of the the particles and droplets. The change in the monomer chemical
potential is illustrated in Fig. 21, where Y is defined as the swelling capacity: the
ratio of the weight of a swollen particle to its weight before it is swollen.
As shown in Fig. 21, the particles swell with monomer diffusion (increase
in Y). During swelling, the monomer chemical potential in the particles first
rapidly increases and then decreases gradually down to zero with more and
more monomer swelling. On the other hand, the droplets shrink and the co-
stabilizer is concentrated since it cannot (by definition) diffuse out with the
monomer. The monomer chemical potential decreases monotonically. During
the process of monomer diffusion, if the monomer chemical potential in the
droplets is equal to that of the particles, equilibrium is established and monomer
diffusion ceases. In Fig. 21, the formation of high MW polymer is shown for
contrast. Three intersections of the droplet and particle chemical potential curves
can be seen. Two of these occur at low Y and the third at a much higher Y. As
is often the case with three equilibrium points, the middle point is unstable.
When the system arrives at the first intersection during swelling (lowest Y),
monomer transfer stops and the system reaches an equilibrium state that is
called the normal swelling state. In this case, the other two equilibrium points
will never be reached. However, in the case of controlled polymerization, the
formation of oligomers rather than high molecular weight polymer leads to a
lower mixing free energy so that the monomer in the particles has a lower
chemical potential. If the effect is large enough so that the chemical potential
of the droplets remains higher than that of the particles at the peak of the
particle chemical potential curve, the system will move to the right-most equi-
librium point. This will be denoted as the super-swelling state. In this case, a
large amount of monomer will transfer from the droplets to the particles.
240 F. J. Schork et al.
Fig. 22 Plot of monomer conversion versus the number of input free radicals in a droplet
(dotted: D=50 nm; solid: D=100 nm) (from [170])
Fig. 23 The monomer conversion in particles with different initial droplet sizes at various
average free radical fluxes (average of ten runs, in s–1). Filled squares: 0.02; filled triangles:
0.05; filled circles: 0.1 (from [107])
sizes where segregation is not effective. The same argument is suitable for ATRP
in miniemulsion.Additionally, it is well-known that it takes some time for SFRP
and ATRP to set up the propagation/dormant equilibrium. It is likely that the
time required to build up the equilibrium is droplet-size dependent. In such
cases, the scenarios above might also occur. In fact, it has been reported that the
reverse ATRP in miniemulsion has a higher colloidal stability than the direct
ATRP in miniemulsion [264].
The above argument suggests that the initial droplet size distribution may
play an important role in super-swelling or colloidal instability. This indicates
that one should monitor the emulsification procedure for controlled free
radical emulsion polymerization closely.
4.5
Other Applications and Future Directions
Recently, many new reactions have been carried out in miniemulsions. Most of
these are polymerizations, but a number of nonpolymerization reactions have
been proposed. This section will survey these applications, and close with some
speculation on the future of miniemulsions.
4.5.1
Anionic/Cationic Polymerization in Miniemulsions
Maitre et al. [314] carried out anionic polymerization of phenyl glycidyl ether
(PGE) in miniemulsion using didodecyldimethylamonium hydroxide as an
inisurf (combination initiator and surfactant). Long chain alcohols were used
as the costabilizer and stable miniemulsions were created by sonication.
Monomer conversion was low, as was the degree of polymerization, which only
Miniemulsion Polymerization 243
4.5.2
Polycondensation in Miniemulsions
4.5.3
Other Polymerizations in Miniemulsions
4.5.4
Other Miniemulsion Applications
4.5.5
Future Directions
It would seem that miniemulsions have finally moved from being a laboratory
curiosity to being a viable commercial process and a useful synthetic technique
for producing interesting materials with nano-scale structure. It would appear
that polymer-polymer hybrids and polymer-inorganic hybrids achieved via
miniemulsion polymerization will result in new classes of water-borne mate-
rials. Other, traditionally solvent-based, polymerization chemistries may soon
be carried out routinely via the miniemulsion route due to improvements in
polymerization catalysts. The use of miniemulsions in ROMP has been cited
above. Metallocene polymerization of ethylenic monomers has been carried
out in macroemulsion. A short review and discussion of this work is given in
[338]. If these water-tolerant polymerization chemistries are successful, it
cannot be long before they are ported into miniemulsions. Controlled radical
polymerization, particularly using RAFT chemistry, is a natural application for
miniemulsion technology. Perhaps most importantly, miniemulsion techniques
will be used in a variety of nonpolymerization technologies to produce nano-
scale, highly structured materials.
References
1. Asua JM (2002) Prog Polym Sci 27:1283
2. Guyot A (2001) Curr Trends Polym Sci 6:47
3. Antonietti M, Landfester K (2002) Prog Polym Sci 27:689
4. Trommsdorf E, Schlidknecht E (1956) Suspension polymerization in polymer
processes. Interscience, New York
5. Ugelstad J, El-Aasser MS, Vanderhoff JW (1973) J Polym Sci Pol Lett 11:503
6. Gardon JL (1970) Brit Polym J 2:1
7. Poehlein GW (1982) In: Piirma I (ed) Emulsion polymerization. Academic, New York,
p 357
8. Poehlein GW (1985) ACS Sym Ser 285:131–150
Miniemulsion Polymerization 247
51. El-Aasser MS, Lack CD, Choi YT, Min TI, Vanderhoff JW, Fowkes FM (1984) Colloids
Surface 12:79
52. Ugelstad J, Kaggerud KH, Fitch RM (1980) In: Fitch RM (ed) Polymer colloids II.
Plenum, New York, p 83
53. Choi YT, El-Aasser MS, Sudol ED, Vanderhoff JW (1985) J Polym Sci Pol Chem 23:
2973
54. Chamberlain BJ, Napper DH, Gilbert RG (1982) J Chem Soc Farad T 1 78:591
55. Chen CM, Gothjelpsen L, Schork FJ (1986) Polym Proc Eng 4:1
56. Delgado J, El-Aasser MS, Silebi CA, Vanderhoff JW (1986) Polym Mat Sci Eng 54:444
57. Rodriguez VS, Delgado J, Silebi CA, El-Aasser MS (1988) Polym Mat Sci Eng 58:761
58. Delgado J, El-Aasser MS, Vanderhoff JW (1986) J Polym Sci Pol Chem 24:861
59. Asua JM, Rodriguez VS, Silebi CA, El-Aasser MS (1990) Makromol Chem M Symp
35–36:59
60. Rodriguez VS, Delgado J, Silebi CA, El-Aasser MS (1989) Ind Eng Chem Res 28:65
61. Rodriguez VS, Asua JM, El-Aasser MS, Silebi CA (1991) J Polym Sci Pol Physics 29:483
62. Fontenot K, Schork FJ (1992) Polym Reaction Engineering 1:75
63. Fontenot K, Schork FJ (1993) Polym React Eng 1:289
64. Samer CJ, Schork FJ (1997) Polym React Eng 5:85
65. Ma JW, Cunningham MF, McAuley KB, Keoshkerian B, Georges MK (2003) Macromol
Theor Simul 12:72
66. Ma JW, Cunningham MF, McAuley KB, Keoshkerian B, Georges M (2003) Chem Eng Sci
58:1177
67. Ma JW, Smith JA, McAuley KB, Cunningham MF, Keoshkerian B, Georges MK (2003)
Chem Eng Sci 58:1163
68. Ma JW, Cunningham MF, McAuley KB, Keoshkerian B, Georges MK (2002) Macromol
Theor Simul 11:953
69. Samer CJ, Schork FJ (1999) Ind Eng Res 38:1801
70. Landfester K, Bechthold N, Tiarks F, Antonietti M (1999) Macromolecules 32:2679
71. Bechthold N, Tiarks F, Willert M, Landfester K, Antonietti M (2000) Macromol Symp
151:549
72. Bradley M, Grieser F (2002) J Colloid Interf Sci 251:1
73. Wang S, Schork FJ (1994) J Appl Polym Sci 54:2157
74. Chern CS, Chen TJ (1997) Colloid Polym Sci 275:1060
75. Chern CS, Liou YC (1999) Polymer 40:3763
76. Wu XQ, Schork FJ (2001) J Appl Polym Sci 81:1691
77. Graillat C, Guyot A (2003) Macromolecules 36:6371
78. Boisson F, Uzulina I, Guyot A (2001) Macromol Rapid Comm 22:1135
79. Lack CD, El-Aasser MS, Vanderhoff JW, Fowkes FM (1985) ACS Sym Ser 272:345
80. Azad ARM, Ugelstad J, Hansen FK (1976) ACS Sym Ser 24:1
81. Hawkett BS, Napper DH, Gilbert RG (1980) J Chem Soc Farad T I 76:1323
82. Hallworth GW, Carless JE (1974) Theory and practice of emulsion technology. Acade-
mic, New York, p 305
83. Choi YT, El-Asser MS, Sudol ED, Vanderhoff JW (1985) J Appl Polym Sci 23:2973
84. Choi YT (1986) PhD Thesis, Lehigh University, Bethlehem, PA
85. Friberg SE, Neogi P (1986) Disp Sci Tech J 7:50
86. Kislalioglu S, Friberg S (1976) Theory and practice of emulsion technology. Academic,
New York, p 257
87. Lack CD, El-Aasser MS, Silebi CA, Vanderhoff JW, Fowkes FM (1987) Langmuir 3:1155
88. Shah DO, Schechter RS (1977) Improved oil recovery by surfactant and polymer flood-
ing. Academic, New York
Miniemulsion Polymerization 249
89. El-Aasser MS, Lack CD, Choi YT, Min TI, Vanderhoff JW, Fowkes FM (1984) Colloids
Surface 12:79
90. Rodriguez VS (1988) PhD Thesis, Lehigh University, Bethlehem, PA
91. Delgado J, El-Aasser MS, Vanderhoff JW (1986) J Polym Sci Pol Chem 24:861
92. Reimers JL, Skelland AHP, Schork FJ (1995) Polym React Eng 3:235
93. Ugelstad J (1980) Adv Colloid Interfac 13:101
94. Reimers JL, Schork FJ (1996) J Appl Polym Sci 60:251
95. Reimers JL, Schork FJ (1996) J Appl Polym Sci 59:1833
96. Aizpurua I, Amalvy JI, Barandiaran MJ (2000) Colloids Surface A 166:59
97. Lelu S, Novat C, Graillat C, Guyot A, Bourgeat-Lami E (2003) Polym Int 52:542
98. Wang ST, Schork FJ, Poehlein GW, Gooch JW (1996) J Appl Polym Sci 60 2069
99. Wu XQ, Schork FJ, Gooch JW (1999) J Polym Sci Pol Chem 37:4159
100. Tsavalas JG, Gooch JW, Schork FJ (2000) J Appl Polym Sci 75:916
101. Dong H, Gooch JW, Schork FJ (2000) J Appl Polym Sci 76:105
102. Reimers JL, Schork FJ (1996) Polym React Eng 4:135
103. Chern C-S, Sheu J-C (2000) J Polym Sci Pol Chem 38:3188
104. Samer CJ, Schork FJ (1999) Ind Eng Res 38:1792
105. Mouron D, Reimers J, Schork FJ (1996) J Polym Sci Pol Chem 34:1073
106. Wang S, Poehlein GW, Schork FJ (1997) J Polym Sci Pol Chem 35:595
107. Reimers JL, Schork FJ (1997) Ind Eng Resh 36:1085
108. Asua J, Alduncin J, Forcada J (1994) Macromolecules 27:2256
109. Miller CM, Blythe PJ, Sudol ED, Silebi CA, El-Aasser MS (1994) J Polym Sci Pol Chem
32:2365
110. Miller CM, Sudol ED, Silebi CA, El-Aasser MS (1995) Macromolecules 28:2754
111. Miller CM, Sudol ED, Silebi CA, El-Aasser MS (1995) Macromolecules 28:2765
112. Miller CM, Sudol ED, Silebi CA, El-Aasser MS (1995) Macromolecules 28:2772
113. Blythe PJ, Morrison BR, Mathauer KA, Sudol ED, El-Aasser MS (1999) Macromolecules
32:6944
114. Blythe PJ, Klein A, Sudol ED, El-Aasser MS (1999) Macromolecules 32:6952
115. Blythe PJ, Klein A, Sudol ED, El-Aasser MS (1999) Macromolecules 32:4225
116. Blythe PJ, Morrison BR, Mathauer KA, Sudol ED, El-Aasser MS (2000) Langmuir
16:898
117. Ghazaly HM, Daniels ES, Dimonie VL, Klein A, El-Aasser MS (2001) J Appl Polym Sci
81:1721
118. Luo Y, Schork FJ (2002) J Polym Sci Pol Chem 40:3200
119. Alduncin JA, Asua JM (1994) Polymer 35:3758
120. Blythe PJ, Klein A, Phillips JA, Sudol ED, El-Aasser MS (1999) J Polym Sci Pol Chem
37:4449
121. Gilbert RG (1995) Emulsion polymerization: A mechanistic approach.Academic, London
122. Kabalnov AS, Pertzov AV, Shchukin ED (1987) J Colloid Interf Sci 118:590
123. Candau F, Pabon M, Anquetil J-Y (1999) Colloid Surface A 153:47
124. Sood A, Awasthi KJ (2003) Appl Polym Sci 88:3058
125. Erdem B, Sully Y, Sudol ED, Dimonie VL, El-Aasser MS (2000) Langmuir 16:4890
126. Landfester K, Bechthold N, Föster S,Antonietti M (1999) Macromol Rapid Comm 20:81
127. Landfester K, Bechthold N, Tiarks F, Antonietti M (2000) Macromolecules 32:5222
128. Luo Y, Tsavalas JG, Schork FJ (2001) Macromolecules 34:5501
129. Melis S, Kemmere M, Meuldijk J, Storti G, Morbidelli M (2000) Chem Eng Sci 55:3101
130. Leiza JR, Sudol ED, El-Aasser MS (1997) J Appl Polym Sci 64:1797
131. Tang PL, Sudol ED, Adams M, El-Aasser MS, Asua JM (1991) J Appl Polym Sci 42:
2019
250 F. J. Schork et al.
172. Gilbert RG, Anstey JF, Subramaniam N, Monteiro MJ (1999) ACS Polym Prepr Div
Polym Chem 40:102
173. Blackley DC (1975) Emulsion polymerization. Wiley, New York, p 244
174. Dong H, Gooch JW, Poehlein GW, Wang ST, Wu X, Schork FJ (2001) ACS Sym Ser
766:8–17
175. Nabuurs T, Baijards RA, German AL (1996) Prog Org Coat 27:163
176. Van Hamersveld EMS, van Es GS, Cuperus FP (1999) Colloids Surface A 153:285
177. Van Hamersveld EMS,Van Es GS, German AL, Cuperus FP,Weissenborn P, Hellgren AC
(1999) Prog Org Coat 35:235
178. Tsavalas JG, Luo Y, Hudda L, Schork FJ (2003) Polym React Eng 11:277
179. Tsavalas JG, Luo Y, Schork FJ (2003) J Appl Polym Sci 87:1825
180. Tsavalas JG, Schork FJ, Landfester K (2003) J Coating Technol (in press)
181. Shoaf GL, Stockl RR (2003) Polym React Eng 11:319
182. Li, Daniels ES, Dimonie VL, Sudol ED, El-Aasser MS (2001) Polym Mat Sci Eng 85:258
183. Wang C, Chu F, Graillat C, Guyot A (2003) Polym React Eng 11:541
184. Barrere M, Landfester K (2003) Macromolecules 36:5119
185. El-Aasser MS, Li M, Jeong P, Daniels ES, Dimonie VL, Sudol ED (2001) DECHEMA
Monographien (7th Int Workshop on Polymer Reaction Engineering), Hamburg,
Germany, 8–10 October 2001, 137:1
186. Jeong P, Dimonie VL, Daniels ES, El-Aasser MS (2002) ACS Sym Ser 801:357
187. Kawahara H, Goto T, Okamoto Y, Kage H, Ogura H, Matsuno Y (2002) Kagaku Kogaku
Ronbun 28:175
188. Landfester K, Dimonie VL, El-Aasser MS (1998) DECHEMA Monographien 134 (6th
Int Workshop on Polymer Reaction Engineering 1998), Berlin, 5–7 October 1998,
134:469
189. Roberts JE, Marcu I, Dimonie V, Daniels E, El-Aasser MS (2003) ACS Polym Prepr Div
Polym Chem 44:277
190. Marcu I, Daniels ES, Dimonie VL, Hagiopol C, Roberts JE, El-Aasser MS (2003) Macro-
molecules 36:328
191. El-Aasser MS, Vanderhoff JW, Poehlein GW (1977) Coating Plastic Prepr 37:92
192. El-Aasser MS, Hoffman JD, Manson JA,Vanderhoff JW (1980) Org Coating Plast Chem
43:136
193. Vanderhoff JW, El-Aasser MS, Hoffman JD (1978) US Patent 4070323
194. Noh MH, Jang LW, Lee DC (1999) J Appl Polym Sci 74:179
195. Noh MH, Lee DC (1999) J Appl Polym Sci 74:2811
196. Garcés JM, Moll DJ, Bicerano J, Fibinger R, McLeod DG (2000) Adv Mater 12:1835
197. Van Herk AM, German AL (1999) Microsph Microcaps Lipos 1:457
198. Erdem B, Sudol ED, Dimonie VL, El-Aasser MS (2000) J Polym Sci Pol Chem 38:4419
199. Erdem B, Sudol ED, Dimonie VL, El-Aasser MS (2000) J Polym Sci Pol Chem 38:4431
200. Erdem B, Sudol ED, Dimonie VL, El-Aasser MS (2000) J Polym Sci Pol Chem 38:4441
201. Tiarks F, Landfester K, Antonietti M (2001) Langmuir 17:5775
202. Antonietti M, Landfester K (2002) Chem Ing Tech 74:543
203. Landfester K, Montenegro R, Scherf U, Guntner R, Asawapirom U, Patil S, Neher D,
Kietzke T (2002) Adv Mater 14:651
204. Ramirez LP, Landfester K (2003) Macromol Chem Phys 204:22
205. Solomon DH, Rizzardo E, Cacioli P (1985) Eur Pat Appl EP 135280
206. Georges MK, Moffat KA, Veregin RPN, Kazmaier PM, Hamer GK (1993) Polym Mater
Sci Eng 69:305
207. Fischer H (1997) Macromolecules 30:5666
208. Gridnev AA (1997) Macromolecules 30:7651
252 F. J. Schork et al.
209. Moad G, Solomon DH (1995) The chemistry of free radical polymerization, 1st edn.
Elsevier, Amsterdam
210. Fukuda T, Terauchi T, Goto A, Ohno K, Tsujii Y, Miyamoto T, Kobatake S, Yamada B
(1996) Macromolecules 29:6393
211. Fukuda T (2002) Handbook of radical polymerization. Wiley, New York, Ch 9
212. Qiu J, Charleux B, Matyjaszewski K (2001) Prog Polym Sci 26:2083
213. Butte A, Storti G, Morbidelli M (1998) In: Reicher KH, Moritz HU (eds) 6th Int Work-
shop on Polymer Reaction Engineering. DECHEMA Monographien, Berlin, 5–7 October
1998, 134:497
214. Butte A, Storti G, Morbidelli M (2000) Macromolecules 33:3485
215. Charleux B (2000) Macromolecules 33:5358
216. Pan G, Sudol ED, Dimonie VL, El-Aasser MS (2002) Macromolecules 35:6915
217. Lansalot M, Farcet C, Charleux B, Vairon J-P, Pirri R, Tordo P (2000) ACS Sym Ser
768:138
218. Prodpan T, Dimonie VL, Sudol ED, El-Aasser M (1999) Polym Mater Sci Eng 80:534
219. MacLeod PJ, Keoshkerian B, Odell P, Georges MK (1999) Polym Mater Sci Eng 80:539
220. Lansalot M, Charleux B, Vairon J-P, Pirri R, Tordo P (1999) ACS Polym Prepr 40:317
221. Prodpan T, Dimonie VL, Sudol ED, El-Aasser M (2000) Macromol Symp 155:1
222. MacLeod PJ, Barber R, Odell P, Keoshkerian B, Georges MK (2000) Macromol Symp
155:31
223. Farcet C, Lansalot M, Charleux B, Pirri R, Vairon JP (2000) Macromolecules 33:8559
224. Pan G, Sudol ED, Dimonie VL, El-Aasser MS (2001) Macromolecules 34:481
225. Keoshkerian B, MacLeod PJ, Georges MK (2001) Macromolecules 34:3594
226. Farcet C, Charleux B, Pirri R (2001) Macromolecules 34:3823
227. Cunningham MF (2002) Prog Polym Sci 27:1039
228. Tortosa K, Smith JA, Cunningham MF (2001) Macromol Rapid Comm 22:957
229. Farcet C, Belleney J, Charleux B, Pirri R (2002) Macromolecules 35:4912
230. Farcet C, Charleux B, Pirri R (2002) Macromol Symp 182:249
231. Keoshkerian B, Szkurham AR, Georges MK (2001) Macromolecules 34:6531
232. Cunningham MF, Tortosa K, Lin M, Keoshkerian B, Georges MK (2002) J Polym Sci Pol
Chem 40:2828
233. Cunningham MF, Xie M, McAuley KB, Keoshkerian B, Georges MK (2002) Macromole-
cules 35:59
234. Ma JW, Cunningham MF, McAuley KB, Keoshkerian B, Georges MK (2002) Macromol
Theor Simul 11:953
235. Ma JW, Cunningham MF, McAuley KB, Keoshkerian B, Georges MK (2003) Macromol
Theor Simul 12:72
236. Matyjaszewski K, Wang J (1995) J Am Chem Soc 117:5614
237. Matyjaszewski K, Wang J (1995) Macromolecules 28:7901
238. Matyjaszewski K, Xia J, Patten T, Abernathy T (1996) Science 272:866
239. Sawamoto M, Kamigaito M, Higashimura T, Kato M (1995) Macromolecules 28:1721
240. Sawamoto M, Kamigaito M, Ando T, Kato M (1996) Macromolecules 29:1070
241. Sawamoto M, Kamigaito M, Kotani Y, Kato M (1996) Macromolecules 29:6979
242. Uegaki H, Kotani Y, Kamigato M, Sawamoto M (1998) Macromolecules 31:6756
243. Moineau G, Minet M, Dubois P, Teyssié P, Sennenger T, Jérôme R (1999) Macromolecules
32:27
244. Lecomte P, Drapier I, Dubois P, Teyssié P, Jérôme R (1997) Macromolecules 30:7631
245. Brandts JAM, van de Geijn P, van Faassen EE, Boersma J, van Koten G (1999) J Organo-
met Chem 584:246
246. Kotani Y, Kamigaito M, Sawamoto M (1999) Macromolecules 32:2420
Miniemulsion Polymerization 253
286. Lichti G, Sangster DF, Whang BCY, Napper DH, Gilbert RG (1982) J Chem Soc Farad T
1 78:2129
287. Maxwell IA, Morrison BR, Napper DH, Gilbert RG (1992) Makromol Chem 193:303
288. Mendoza J, De La Cal JC, Asua JM (2000) J Polym Sci Pol Chem 38:4490
289. Monteiro MJ, Hodgson M, De Brouwer HJ (2000) Polym Sci Pol Chem 38:3864
290. de Brouwer H, Tsavalas JG, Schork FJ, Monteiro MJ (2000) Macromolecules 33:9239
291. Moad G, Chiefari J, Chong YK, Krstina J, Mayadunne RTA, Postma A, Rizzardo E, Thang
SH (2000) Polym Int 49:993
292. Prescott SW, Ballard MJ, Rizzardo E, Gilbert RG (2002) Macromolecules 35:5417
293. Monteiro MJ, de Barbeyrac J (2001) Macromolecules 34:4416
294. Smulders WW, Gilbert RG, Monteiro MJ (2003) Macromolecules 36:4309
295. Smulders WW (2002) PhD Thesis, Technische Universiteit Eindhoven, The Nether-
lands
296. Tsavalas JG, Schork FJ, de Brouwer H, Monteiro MJ (2001) Macromolecules 34:3938
297. Vosloo JJ, De Wet-Roos D, Tonge MP, Sanderson RD (2002) Macromolecules 35:4894
298. Lansalot M, Davis TP, Heuts JPA (2002) Macromolecules 35:7582
299. Uzulina I, Kanagasbapathy S, Claverie J (2000) Macromol Symp 150:33
300. Monteiro MJ, Hodgson M, de Brouwer H (2000) J Polym Sci Pol Chem 38:3864
301. Bon SAF, Bosveld M, Klumperman B, German AL (1997) Macromolecules 30:324
302. Marestin C, Noel C, Claverie J (1998) Macromolecules 31:4041
303. Lansalot M, Farcet C, Charleux B, Vairon JP, Pirri R, Tordo O (2000) ACS Sym Ser
768:138
304. Jousset S, Qiu J, Matyjaszewski K, Granel C (2001) Macromolecules 34:6641
305. De Brouwer H, Tsavalas JG, Schork FJ, Monteiro MJ (2000) Macromolecules 33:9239
306. Prodpan T, Dimonie VL, Sudol ED, El-Aasser M (1999) Polym Mater Sci Eng 80:534
307. MacLeod PJ, Keoshkerian B, Odell P, Georges MK (1999) Polym Mater Sci Eng 80:539
308. Pan GF, Sudol ED, Dimonie VL, El-Aasser MS (2002) Macromolecules 35:6915
309. Matyjaszewski K, Qui J, Shipp DA, Gaynor SG (2000) Macromol Symp 155:15
310. Tonge MP, McLeary JB, Vosloo JJ, Sanderson RD (2003) Macromol Symp 193:289
311. Vosloo JJ, Roos DW, Tonge MP, Sanderson RD (2002) Macromolecules 35:4894
312. Ausa JM (2002) Prog Polym Sci 27:1283
313. Landfester K, Bechthold N, Tiarks F, Antonietti M (1999) Macromolecules 32:5222
314. Maitre C, Ganachaud F, Ferreira O, Lutz JF, Paintoux Y, Hemery P (2000) Macromole-
cules 33:7730
315. Barrere M, Maitre C, Dourges MA, Hemery P (2001) Macromolecules 34:7276
316. Barrere M, Ganachaud F, Bendejacq D, Dourges MA, Maitre C, Hemery P (2001) Polymer
42:7239
317. Limouzin C, Caviggia A, Ganachaud F, Hemery P (2003) Macromolecules 36:667
318. Cauvin S, Sadoun A, Santos R, Belleney J, Ganachaud F, Hemery P (2002) Macromole-
cules 35:7919
319. Barrere M, Landfester K (2003) Polymer 44:2833
320. Landfester K, Rothe R, Antonietti M (2002) Macromolecules 35:1658
321. Willert M, Landfester K (2002) Macromol Chem Phys 203:825
322. Marie E, Rothe R, Antonietti M, Landfester K (2003) Macromolecules 36:3967
323. Marie E, Landfester K, Antonietti M (2002) Biomacromolecules 3:475
324. Taden A, Antonietti M, Landfester K (2003) Macromol Rapid Comm 24:512
325. Claverie JP, Viala S, Maurel V, Novat C (2001) Macromolecules 34:382
326. Landfester K (2001) Adv Mater 13:765
327. Montenegro R, Antonietti M, Mastai Y, Landfester K (2003) J Phys Chem B 107:5088
328. Taden A, Landfester K (2003) Macromolecules 36:4037
Miniemulsion Polymerization 255
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
258 P. Y. Chow · L. M. Gan
Abstract This review describes how the unique nanostructures of water-in-oil (W/O), oil-
in-water (O/W) and bicontinuous microemulsions have been used for the syntheses of some
organic and inorganic nanomaterials. Polymer nanoparticles of diameter approximately
10–50 nm can easily be obtained, not only from the polymerization of monomers in all three
types of microemulsions, but also from a Winsor I-like system. A Winsor I-like system with
a semi-continuous process can be used to produce microlatexes with high weight ratios of
polymer to surfactant (up to 25). On the other hand, to form inorganic nanoparticles, it is
best to carry out the appropriate chemical reactions in W/O- and bicontinuous micro-
emulsions.
Recent developments in the cross-polymerization of the organic components used in
bicontinuous microemulsions ensure the successful formation of transparent nanostruc-
tured materials. Current research into using polymerizable bicontinuous microemulsions
as a one-pot process for producing functional membranes and inorganic/polymer nano-
composites is highlighted with examples.
Abbreviations
AN Acrylonitrile
AA Acrylic acid
AM Acrylamide
APTAC 2-acrylamido-2-propane trimethylammonium chloride
AUTMAB (Acryloyloxy)-undecyltrimethylammonium bromide
AUDMAA (Acryloyloxy)-undecyldimethylammonioacetate
AIBN Azobisisobutyronitrile
AOT 1,4-Bis(2-ethylhexyl)sulfosuccinate sodium salt
APS Ammonium persulfate
BL g-butyrolactone
BUA Butylacrylate
C1-PEO-C11-MA-40 w-methoxypoly(ethylene oxide)40-undecyl-a-methacrylate
CTAB Cetyltrimethylammonium bromide
DMC Dimethyl carbonate
DTAB Dodecyltrimethylammonium bromide
EGDMA ethyleneglycol dimethacrylate
EC Ethylene carbonate
HC Hydrocarbon
HA Hydroxyapetite
HEMA 2-hydroxylethylmethacrylate
MADQUAT 2-Methacryloyloxyethyltrimethyl ammonium chloride
MMA Methyl methacrylate
NaA Sodium acrylate
NaAMPS Sodium-2-acrylamido-2–2-methylpropane sulfonate
NaSS Sodium styrenesulfonate
NIPAM N-isopropylacrylamide
NP Poly(oxyethylated alkylphenyl ether)
OTAC Octyldecyltrimethylammonium chloride
PC Propylene carbonate
PMMA Poly(methyl methacrylate)
PNIPAM Poly(N-isopropylacrylamide)
Microemulsion Polymerizations and Reactions 259
PS Poly(styrene)
SDS sodium dodecyl sulfate
SEAAU Sodium 11-(N-ethylacrylamido)-undecanoate
St Styrene
TEM Transmission electron microscope
THFM Tetrahydrofurfuryl methacrylate
TTAB Tetradecyltrimethylammonium bromide
VA Vinyl acetate
VBSLi 4-Vinylbenzene sulfonate, lithium salt
1
Introduction
Microemulsions are transparent liquid systems consisting of at least ternary mix-
tures of oil, water and surfactant. Sometimes a cosurfactant is needed for the
formation of a thermodynamically-stable microemulsion. A transparent micro-
emulsion is, in fact, heterogeneous (nanostructured) on a molecular scale. Micro-
emulsion domains fluctuate in size and shape and undergo spontaneous coales-
cence and break-up [1]. They can exhibit water continuous and bicontinuous
structures, with typical equilibrium domain sizes ranging from about 10 to 100 nm
[2]. The transparent microemulsions can be in the form of nano-globules of
oil-swollen micelles dispersed in the water continuous phase as oil-in-water (O/W)
microemulsions, or water-swollen micellar globules dispersed in oil as water-in-
oil (W/O) microemulsions. In between the regions of O/W and W/O microemul-
sions, there may also exist a composition region of bicontinuous (sponge-like
structure) microemulsions [3], whose oil and water domains are randomly dis-
persed in two phases, as can be seen from the sketched phase diagram of Fig. 1.
2
Polymerizations in Globular and Bicontinuous Microemulsions
for Producing Microlatexes
Over the past two decades, free radical polymerization studies have mainly
been carried out in globular microemulsions (both O/W microemulsions and
W/O microemulsions, also known as inverse microemulsions). Each milliliter
of the globular microemulsion usually contains 1015–1017 W/O- or O/W nano-
sized (5–10 nm in diameter) droplets (globules). The enormous number of
nano-globules are potential loci for fast polymerization, producing microlatex
particles much less than 50 nm in diameter. Though small polymer particles,
molecular weights exceeding one million can easily be obtained from these
polymerization systems. Most of the microemulsion polymerization studies [5]
have dealt with hydrophobic monomers, such as styrene (St) or methyl meth-
acrylate (MMA), within oil cores of O/W microemulsions [9] and with the poly-
merization of water-soluble monomers, such as acrylamide (AM), within
aqueous cores of inverse microemulsions [10, 11]. For both O/W and inverse
microemulsion systems, the amount of monomer was usually restricted to less
than 10 wt% with respect to the total weight of microemulsion. Moreover,
Microemulsion Polymerizations and Reactions 261
higher amounts of surfactant (5–15 wt%) were normally needed for the stabil-
ity of the polymerization. For those microemulsions requiring a cosurfactant,
the compatibility between the cosurfactant and the polymers formed becomes
an issue. For instance, if styrene is polymerized within an O/W microemulsion
that contains an alcohol cosurfactant, this alcohol will not dissolve polystyrene.
However, this is not the case for the polymerization of acrylamide (AM) in
alcohol-free inverse microemulsion.
2.1
Inverse Microemulsion Polymerization
2.2
Bicontinuous Microemulsion Polymerization
2.3
Polymerization in Oil-in-Water Microemulsions
droplets. Therefore, the newly formed latex particles continue to grow in the
microlatex through the constant supply of monomer from nucleated droplets
until the chain in the particle is terminated by chain transfer to monomer [42,
49, 50]. However, Kaler’s group recently consistently obtained very high mole-
cular weights of ~15¥106 Da for styrene O/W microemulsion polymerizations
[51]. Such molecular weights are considerably higher than those typically
obtained via free-radical polymerizations. In the free-radical polymerization
of styrene, the chain transfer to monomer is limited and hence can only produce
molecular weights of up to ~2¥10–6 Da, as suggested by the most recent mea-
surements of chain-transfer constants by Kukulj et al [52]. Therefore, the
diffusion-limited exit of monomer radicals to the aqueous phase coupled with
chain transfer to polymer are probable reasons for the enhanced molecular
weight of polystyrene [51].
In contrast to emulsion polymerization, the reaction kinetics of micro-
emulsion polymerization is characterized by two polymerization rate intervals;
the interval of constant rate characteristic of emulsion polymerization is miss-
ing [42, 49, 53], as shown in Fig. 2. Polymer particles are generated continuously
during the reaction by both micellar and homogeneous mechanisms. As the
solubility of the monomer in the continuous domain increases, homogeneous
Fig. 2 Experimental and model rate versus conversion profiles for the polymerization of
hexylmethacrylate in a microemulsion stabilized by the surfactant DTAB. The two curves are
for initiator concentrations of 0.045 wt% (top) and 0.015 wt% (bottom) relative to the
amount of monomer in the microemulsion. The solid lines are predictions from the Mor-
gan model [56]
Microemulsion Polymerizations and Reactions 265
Fig. 3 Changes in PMMA particle size during long term storage at 60 °C for microlatexes
stabilized by different surfactants: (filled triangles) TTAB; (filled squares) TTAC; (filled
circles) CTAB; (empty triangles) OTAC
2.4
Microemulsion Polymerization for Microlatexes
with High Polymer-to-Surfactant Weight Ratios
The new system consists of a pure styrene upper phase placed on top of a
ternary O/W microemulsion. Strictly speaking, this new system is not identical
to the Winsor I system. Hence the new polymerization system has been referred
as a “Winsor I-like” system, which can be prepared by simply topping up a
ternary O/W microemulsion with hydrophobic monomer without disturbing
the microemulsion phase. Styrene polymerization took place only in the initial
ternary microemulsion containing 0.5–1.0 wt% styrene, 1.0 wt% DTAB and a
redox initiator of ammonium persulfate/tetramethylethylenediamine (APS/
TMEDA). The upper styrene phase acted only as a monomer reservoir to conti-
nuously supply monomer to the polymerization loci in the lower micro-
emulsion phase through diffusion without stirring. PS particles initially formed
in the microemulsion phase became the seed for the further growth of PS
particles to a more uniform size, up to about 80 nm in diameter. Therefore,
microlatexes of about 15 wt% PS could be obtained using only 1 wt% DTAB.
It should also be noted that the Winsor I-like system with a small amount of
monomer in the upper phase becomes a milky emulsion on stirring. But this
milky emulsion will transform into a transparent or bluish microlatex after
polymerization.
The polymerization rate of the Winsor I-like system was very slow due to the
low monomer diffusion rate through the limited interface to the polymeriza-
tion loci for the unstirred system. In order to increase the interfacial areas for
the infusion of monomer to the polymerization loci, hollow fiber monomer
feeds were later used to polymerize not only styrene, but also MMA and butyl-
acrylate [66]. Such a set-up using continuous monomer feeding via a hollow-
fiber is shown in Fig. 4. About 100 polypropylene (PP) hollow fibers (pore size
70 nm) were bundled together. One of the openings of the hollow-fiber bundle
was used for monomer feeding and the other end was sealed with epoxy resin.
The initial microemulsion usually consisted of 0.5 wt% styrene or BA, or 2 wt%
MMA, 1 wt% CTAB or 1.5 wt% SDS together with 0.2 wt% 1-pentanol, 4 mM
equimolar redox initiator (APS/TMEDA), and water to make it up to 100 wt%.
After polymerization of the microemulsion for about 15 min at room temper-
ature with about 97–98% conversion, the additional monomer was then contin-
uously introduced to the polymerization system via the infusion of monomer
from hollow fibers connected to monomer reservoir. The rate of monomer
infusion into the microemulsion could be regulated by the external nitrogen
pressure applied to the monomer reservoir. The slow bubbling by nitrogen gas
into the polymerizing microemulsion was to ensure homogeneous mixing.
When the infusion rate of monomer was optimal, latex particles would grow to
more uniform sizes at the expense of forming secondary particles via homo-
geneous nucleation. This method was able to produce almost uniform micro-
latexes of various sizes (15–65 nm) at high polymer/surfactant weight ratios up
to about 15 with high molecular weights (106 g/mol) within 2–3 h.
Though the continuous addition of monomer via hollow fibers into a
Winsor I-like system is rather novel, it is not as convenient as the drop-wise
addition of monomer under a semi-continuous process. Ming et al [67, 68]
268 P. Y. Chow · L. M. Gan
Fig. 4 Schematic presentation of the set-up (not to scale) for the polymerization of mo-
nomer in a microemulsion via hollow-fiber feeding of additional monomer
slightly modified the Winsor I-like method by directly adding monomer very
slowly to pre-polymerized ternary microemulsions for producing microlatexes
of PS, PMMA, poly(butyl methacrylate), or poly(methacrylate) (PMA). With
1 wt% DTAB, 24 wt% PMMA microlatexes of 33–46 nm in size were obtained.
They showed that this semi-continuous process was very effective for produc-
ing small microlatexes (15 nm) containing up to 30 wt% PMA at a high PMS to
SDS weight ratio of 25:1. The semi-continuous process works at the monomer-
starved condition to ensure that no empty micelles exist during polymerization.
The polymerization of styrene in Winsor I-like systems by semi-continuous
feeding of monomer stabilized by either DTAB, TTAB or CTAB has been sys-
tematically investigated by Gan and coworkers [69a]. Rather monodisperse
polystyrene microlatexes of less than 50 nm with molecular weights of over one
million were obtained at a polymer/surfactant weight ratio of 14:1. The Win-
sor I-like (micro)emulsion polymerization of styrene stabilized by non-ionic
surfactant and initiated by oil-soluble initiators has also been reported very re-
cently [69b]. The sizes of the large monomer-swollen particles decreased with
conversion and they merged with growing particles at about 40–50% conversion.
High PMMA content (30–40%) microlatexes [70] stabilized with low con-
centrations of anionic SDS were also prepared by microemulsion polymeriza-
Microemulsion Polymerizations and Reactions 269
3
Bicontinuous-Microemulsion Polymerization for Nanostructured
Solid-Materials
surfactants, such as sodium dodecyl sulfate (SDS), and their polymerized micro-
emulsions were usually found to be opaque and phase-separated. By incorpo-
rating polymerizable acrylic acid as co-surfactant in a microemulsion, stable
transparent polymeric materials can be obtained but only at low solid contents
(<15 wt%) [85]. However, the polymerized microemulsions did not reveal any
microstructures when viewed by scanning electron microscope (SEM). But
when SDS was replaced by a polymerizable surfactant, transparent nanoporous
polymeric materials were obtained.
3.1
Nanostructured Polymers Produced by Bicontinuous-Microemulsion
Polymerizations
3.1.1
Ion-Conductive Membranes
Fig. 8 SEM micrograph of the polymerized microemulsion solid, that contains the poly-
merizable non-ionic surfactant C1-PEO-C11-MA-40 [96], after ethanol extraction
lyte membranes were rather high, in the range of (2–3)¥10–3 S/cm. This rela-
tively high conductivity is attributed to the inclusion of the polymerizable ionic
salt VBSLi and the existence of the interconnected aqueous channels in the
polymerized bicontinuous network, which facilitate ionic conduction for the
membrane electrolyte [94]. However, some unusual ionic conduction pheno-
mena were also observed for these ion-containing membranes – large ions
exhibited higher mobilities than smaller ones [95]. A possible explanation
comes from the larger hydration shells of the lighter cations. This is further
supported by the sharp drop in conductivity when the system is cooled below
the freezing point of water.
The morphology of the microemulsion-polymerized solid after ethanol
extraction, as revealed by SEM, is shown in Fig. 8. The extracted samples show
globular microstructures and voids (pores). These pores might be derived from
the interconnected water-filled voids generated from numerous coalescences of
growing particles during polymerization.
The water in the membranes [96] could easily be replaced by polar organic
solvents, such as BL, PC-EC, or EC-DMC, by immersion. The conductivities of
the membranes decreased sharply (by about two orders of magnitude) with the
use of organic solvents.After soaking these membranes in electrolyte solutions
of 1 M LiSO3CF3/PC-EC, 1 M LiBF4/BL, or 1 M LiClO4/EC-DMC, their conduc-
tivities were restored to 10–3 S/cm [96]. This demonstrates that the water con-
tent in these microporous membranes can be freely exchanged for organic
solvents or electrolyte solutions, indicating that they may be further developed
as composite polymeric electrolytes for lithium/lithium ion rechargeable
batteries.
274 P. Y. Chow · L. M. Gan
3.1.2
Proton-Exchange Membranes
Fig. 9 Membrane performance of a single fuel cell (area: 5 cm2, Pt loading on anode/cathode:
0.4 mg/cm2) using commercial membranes 1 and 2, and microemulsion-synthesized mem-
branes HC-1 and HC-2
3.2
Polymer Nanocomposites Produced by Bicontinuous-Microemulsion
Polymerizations
3.2.1
Ruthenium (II) Complexes in Polymerized Bicontinuous-Microemulsions
3.2.2
Aligned Nanocomposites of Ferrite-Polymer from Bicontinuous-Microemulsion
Polymerization
b
Fig. 11 TEM micrographs of polymerized ferrite-microemulsion nanocomposites contain-
ing 2 wt% NiFe2O4 . a No magnetic field during polymerization; b under the influence of
H=50 Oe field during polymerization
3.3
Synthesis of Nanocomposites Via In-Situ Microemulsion Polymerization
Some non-oxide nanoparticles such as PbS and CdS can be used to prepare
polymer-inorganic nanocomposites by a double-microemulsion process [103].
In this case, two precursor microemulsions must be prepared separately first
and then mixed together for polymerization. Using CdS-polymer nanocom-
278 P. Y. Chow · L. M. Gan
4
Microemulsion Reactions for Processing Inorganic Nanomaterials
The existence of microdomains or droplets with large interfacial area per unit
volume in inverse and bicontinuous microemulsions opens up possibilities for
controlling the inorganic reaction rates, the pathways, the stereochemistry, and
the morphology of the products. By employing microemulsions, monodisperse
particles of a few nanometers in diameter can be prepared for the study of
quantum size effects. Many new and unusual physical and chemical properties
also arise as particles attain nano-sized dimensions [105–107]. There is increas-
ing recognition that aqueous synthesis offers growth control capabilities that
can be conveniently exploited to prepare these desirable fine particles [107,
108]. The dependence of the size and polydispersity of the particles obtained
on the size and concentration of the microemulsion droplets, the concentration
Microemulsion Polymerizations and Reactions 279
of the dispersed reactants, the exchange rate between the droplets, and the
reactions and interactions between surfactants and metal ions or particles has
already been partly addressed [109–114].
Compared to conventional solid-state reaction methods, solution-based syn-
thesis results in higher levels of chemical homogeneity. In solution systems,
mixing of the starting materials is achieved at the molecular level, and this is
especially important when multi-component oxides are being prepared. As a
solution-based materials synthesis technique, the microemulsion-mediated
method offers a unique ability to affect particle synthesis and particle stabil-
ization in one step. A wide variety of nanosize metal oxides – binary oxides
from alumina to zirconia, as well as complex metal oxides – can be prepared by
exploiting the ability of microemulsions to solubilize, compartmentalize, and
concentrate reactants and products. Most of the focus in the past has been on
the successful synthesis of specific metal oxide compounds as nanoparticles.
Water-in-oil microemulsions (inverse microemulsions) have been the most
widely used for the preparation of spherical nanoscale particles. Reagents dis-
solved in the water domains of an inverse microemulsion can react via inter-
micellar communication through dynamic collision processes. Nucleation and
growth occur as the process proceeds.
Many potential applications of microemulsion-derived materials appear
in the literature. The list includes catalysts, nanoporous membranes, nanocom-
posites and precursor powders for functional ceramics [115–139]. However,
before microemulsion-derived materials become more than laboratory curios-
ities, two key issues must be addressed: product recovery and yield. Materials
generated in microemulsion media can be used in two ways: as prepared in the
fluid phase, or as recovered solids. A common practice found in the literature
involves phase destabilization of microemulsions via the addition of polar
solvents such as ethanol and acetone to precipitate the products. This approach
works well as a laboratory technique that permits particles to be recovered
for later characterization. However, it is unlikely to be an economically viable
approach for practical processing, as we eventually need to separate the added
solvents in order to reformulate the original microemulsion. The susceptibility
of microemulsion to destabilization by electrolytes severely limits the highest
concentrations that can be used for precipitation reactions. This has an adverse
impact on prospects for large-scale microemulsion processing. The feasibility
of microemulsion-derived synthesis depends on the availability of surfactant/
oil/water formulations that give stable microemulsions during reaction, but not
during precipitation reactions. Therefore, investigations that emphasize the
effects of reactants and products on the stability domains of microemulsions
will play an important role. Hopefully, such investigations will lead to guide-
lines for formulating microemulsion compositions that are compatible with
large-scale material synthesis.
280 P. Y. Chow · L. M. Gan
4.1
Synthesis of Inorganic Nanoparticles in Inverse Microemulsion
Table 2 (continued)
4.2
Materials Systems
4.2.1
Doped and Un-Doped CdS/ZnS Nanoparticles
containing lead nitrate (0.1 M) in the aqueous phase, was also needed [142].
This is to encourage the formation of PbS shells around the CdS particle cores,
as depicted Fig. 15. Pb2+ ions displaced some Cd2+ in the Cd–S bonds to form
Pb–S bonds on the surfaces of the CdS particles. The thicknesses of the PbS
shells in the PbS-coated nanocomposite could be controlled by changing the
Pb2+ ion concentration. Absorption measurements of the mixed systems as a
function of the concentration of lead nitrate solution indicated a continuous
trend toward the PbS spectrum with increasing Pb2+ concentration. The par-
ticle sizes of the CdS, PbS and CdS/PbS composites are shown in Fig. 16. The
resulting NLO properties of the CdS and PbS and CdS/PbS nanocomposite
particles were determined by a femtosecond Z-scan technique. A large refrac-
tive nonlinearity in these nanocomposite particles was observed due to the
optical Stark Effect and strong interfacial and inter-particle interactions. These
nanoparticles with large refractive nonlinearities may find application in opti-
cal devices, such as those for optical limiting and switching.
Mn-doped ZnS powders were also prepared by inverse microemulsion under
hydrothermal treatment at 120 °C [143]. This produced ultrafine and agglom-
erate-free particles 5–20 nm in diameter. In contrast, the particles prepared via
conventional aqueous solutions often coagulated, forming large aggregated
clusters. Furthermore, compared with Mn-doped ZnS materials synthesized
through conventional aqueous reactions, the nanoparticles prepared in mi-
croemulsion showed significantly enhanced photoluminescence. In particular,
the photoluminescence of particles prepared in microemulsion under hydro-
thermal treatment was found to be 60 times higher than that for material
Microemulsion Polymerizations and Reactions 285
obtained through the direct aqueous reaction at room temperature (as shown
in Fig. 17). This dramatic increase in photoluminescence yield is attributed to
the surface passivation of nanoparticles by the adsorption of surfactants, the
formation of sphelerite with cubic zinc blende structure, and Mn migration
into the interior lattice of the ZnS host.
4.2.2
Magnetic Ferrites
Nano-sized magnetic ferrite particles are the subject of intensive research be-
cause their physical properties are quite different from those of the bulk mate-
rial. The magnetic characteristics of particles used for recording media crucial-
ly depend on their sizes and shapes. So, the material used for high-quality
recording media should be ultrafine, chemically homogeneous, and stable, with
a narrow particle size distribution a predetermined shape. These requirements
demand a reliable and reproducible preparation technique.
The formation of microhomogeneous nanoparticles of nickel and barium
ferrite was carried out in a three-component microemulsion consisting of NP-
5/NP-9 as the surfactant, cyclohexane as the oil, and an aqueous solution of
Microemulsion Polymerizations and Reactions 287
nickel nitrate and ferric nitrate [144].An immediate precipitation was affected
by adding ammonia to the microemulsion, and the resulting amorphous pre-
cipitate was transformed into nickel ferrites by calcining at 600 °C. Discrete
nickel ferrite particles with polyhedron shapes and an average size range
from 10–20 nm were obtained. The sol-particles were identified as NiFe2O4 by
Mossbauer spectroscopy, electron diffraction and X-ray spectroscopy. Magnetic
measurement revealed a saturation magnetization value of 43.0 emu/g for the
sample calcined at 500 °C.
Barium ferrite particles were obtained by mixing a ferric ion-containing
microemulsion and a barium-containing microemulsion [145]. Precipitation of
the hydroxide precursor was affected by adding ammonia solution dropwise
into the microemulsions under stirring. The resulting barium ferrite particles
formed at a calcination temperature of 950 °C. The round particles synthesized
were well dispersed, although a limited degree of particle agglomeration was
also observed. The average size of the calcined particles was in the range of
100–200 nm, which was about ten times bigger than the average size of the
precursor particles. A high saturation magnetization of 69.74 emu/g and an
intrinsic coercivity of 5639 Oe were obtained for the powder calcined at 950 °C
(Fig. 18). These results are comparable to some of the best ever reported for fine
barium ferrite powders prepared via chemistry-based processing routes. The
much improved saturation magnetization and coercivity can be explained by
the high phase purity and well-defined crystallinity of BaFe12O19 developed in
the microemulsion-derived precursor when calcined at a temperature that is
high enough.
288 P. Y. Chow · L. M. Gan
4.2.3
Silica and Silica-Supported Ru-Cu Oxides
4.2.4
Perovskites
The complex oxides that belong to the group known as perovskites have the
general formula ABO3, where the ionic charges on the metals can assume the
form A+B5+O3,A2+B4+O3 and A3+B3+O3. The first microemulsion-mediated pro-
cess to synthesize a perovskite used a double microemulsion protocol based on
a nonionic surfactant Genapol OX/decane/water system [154]. A microemul-
sion containing BaCl2 and TiCl4 in the dispersed phase was mixed with an
oxalic acid-containing microemulsion. However, X-ray diffraction could not
confirm the presence of the intermediate barium titanyl oxalate and the calcin-
ed solid did not yield barium titanate.
The double inverse microemulsion method was also used to synthesize per-
ovskite-type mixed metal oxides [155]. One microemulsion solution contained
nitrate salts of either Ba(NO3)2/Pb(NO3)2, La(NO3)3/Cu(NO3)2 or La(NO3)3/
Ni(NO3)2, and the other microemulsion contained ammonium oxalate or
oxalic acid as the precipitant. These metal oxalate particles of about 20 nm were
readily calcined into single phase perovskite-type BaPbO3, La2CuO4 and LaNiO3.
The calcinations required for the microemulsion-derived mixed oxalates were
100–250 °C below the temperatures used for the metal oxalates prepared by a
conventional aqueous solution precipitation method.
4.2.5
Zirconia, Lead Zirconate and Lead Zirconate Titanate
powders were much finer and has less particle-particle agglomeration. The
synthesis of PbZrO3 powders could also be performed via a polyaniline-medi-
ated microemulsion process [160].A small amount (~4 wt%) of polyaniline was
retained in the microemulsion-derived oxalate precursor by an in situ poly-
merization of lead oxalate and zirconate oxalate. The in situ polymerization of
aniline on the surfaces of the oxalate particles resulted in the formation of a
well-dispersed precursor powder. Upon calcination at 800 °C, ultrafine lead zir-
conate powder was obtained.
Low temperature synthesis of lead zirconate titanate (PZT) can also be obtain-
ed via a microemulsion process [161]. The microemulsion, containing cations
of lead zirconium and titanium in the aqueous phase, was coprecipitated as
hydroxide precursors by the addition of ammonium solution. Crystalline
tetragonal PZT powders were then obtained by calcining the precursors at a
temperature as low as 450 °C in air without forming any intermediate phases.
4.2.6
Hydroxyapetite (HA)
the high production yield using smaller amounts of oil and surfactant. In this
study, HA was prepared by reacting CaCl2 and (NH4)2HPO4 in three reaction
systems: a conventional aqueous solution, a micellar solution containing
3.2 wt% of the nonionic surfactant KB6ZA and 1.0 M CaCl2 solution, and O/W
emulsions containing KB6ZA with varying amounts of petroleum ether and
1.0 M CaCl2 solution. For the emulsion system, the oil phase residing in the
hydrophobic cores of the oil-swollen micelles lead to a large expansion in the
micellar dimensions, in the form of emulsion droplets. This undoubtedly in-
creases the reaction sites between CaCl2 and (NH4)2HPO4 at the interfaces, as
shown schematically in Fig. 19. Moreover, the emulsion droplets are more stable
than the surfactant micelles and the average life-span of an emulsion droplet
is much longer than that of a micelle. All of these factors make an emulsion
system much more suitable than a micellar system for forming HA particles of
high crystallinity. The enhanced crystallinity of HA derived from the emulsion
composition derives from the complexation of Ca2+ by oxyethylene groups of
the nonionic surfactant. The resulting HA precursor powders underwent little
growth in crystallite and particle sizes upon calcination at 650 °C for 6 h.
4.2.7
PtRu/C Catalysts
PtRu alloys are currently the most active anode catalysts for the oxidation of
methanol or CO-contaminated H2, (like H2 derived from reformed methanol)
in low temperature solid polymer electrolyte fuel cells such as direct methanol
fuel cells (DMFC) or indirect methanol fuel cells (IMFC). High surface area
catalysts are generally prepared by co-impregnation,coprecipitation, absorbing
alloy colloids, or surface organometallic chemistry techniques. For both alloy
and oxide promoted catalytic systems it is important that Pt and the second
metal (or metal oxide) are in intimate contact. This close association of the
platinum and the co-catalyst can be difficult to achieve using conventional
catalyst preparation techniques because the active components may be depo-
sited at different sites on the support surface.As the preparation details control
the final composition, surface structure and morphology of the catalysts, it is
not surprising to learn that catalytic activity is strongly dependent on prep-
aration conditions.
Nano-sized PtRu catalysts supported on carbon have been synthesized
from inverse microemulsions and emulsions using H2PtCl6 (0.025 M)/RuCl3
(0.025 M)/NaOH (0.025 M) as the aqueous phase, cyclohexane as the oil phase,
and NP-5 or NP-9) as the surfactant, in the presence of carbon black suspended
in a mixture of cyclohexane and NP-5+NP-9 [164]. The titration of 10% HCHO
aqueous solution into the inverse microemulsions and emulsions resulted
in the formation of PtRu/C catalysts with average particle sizes of about 5 nm
and 20 nm respectively. The RuPt particles were identified by X-ray diffraction,
X-ray photoelectron, and BET techniques. All of the catalysts prepared show
characteristic diffraction peaks pertaining to the Pt fcc structure. XPS analysis
292 P. Y. Chow · L. M. Gan
Fig. 19 Micelle and emulsion droplets in CaCl2 aqueous solutions: The calcium-rich shell
in the adsorbed state, at the: a micelle-water interface; b emulsion droplet surface. The surfac-
tant molecules are represented by the zig-zag lines with Ca2+ heads
Microemulsion Polymerizations and Reactions 293
also revealed that the catalysts contained mostly Pt(0) and Ru(0), with a little
Pt(II), Pt(IV) and Ru(IV). The double microemulsion-derived PtRu/C catalysts
had higher electrocatalytic activities for methanol oxidation than that of the
emulsion-derived PtRu/C electrocatalyst. The peak current density for metha-
nol oxidation at PtRu/C obtained from the transparent inverse microemulsion
was about four times higher that of the emulsion-derived PtRu/C.
4.2.8
Polymer-Coated Inorganic Nanoparticles
5
Conclusions
ration, or proton exchange membranes for fuel cell applications. The surface
characteristics of the nanostructured membranes can be modified further us-
ing suitable functional monomers. New polymerizable microemulsion systems
will be devised for specific copolymers and inorganic/polymer nanocompos-
ites. Some tailor-made nanostructure polymeric materials may therefore be ob-
tained by the polymerizable bicontinuous-microemulsion approach.
References
1. Zana R, Lang J (1987) In: Frieberg SE, Bothorel (eds) Microemulsions: structure and
dynamics. CRC, Boca Raton, FL, Ch 6
2. Biasia J, Clin B, Laolanne P (1987) In: Friberg SE, Bothorel P (eds) Microemulsions:
structure and dynamics. CRC, Boca Raton, FL, Ch 1
3. Frieberg SE, Bothorol P (1986) In: Friberg SE, Bothorel P (eds) Microemulsions: struc-
ture and dynamics. CRC, Boca Raton, FL
4. Winsor PA (1948) T Faraday Soc 44:376
5. Candau F (1999) In: Kumar P, Mittal KL (eds) Handbook of microemulsion science and
technology. Marcel Dekker, New York, Ch 22, p 679
6. Gan LM, Chew CH (2001) In: Nalwa HS (ed) Advanced functional molecules and
polymers. Gordon and Breach, New York, Ch 2, p 35
7. (a) Lopez-Quintela MA, Rivas J (1993) J Colloid Interf Sci 158:446; (b) Aikawa K, Kaneko
K, Tamura T, Fujitsu M, Ohbu K (1999) Colloids Surface A 150:95; (c) Zarur AJ,Ying JY
(2000) Nature 403:65; (d) Rymes J, Ehret G, Hilaire L, Boutonnet M, Jiratova K (2002)
Catal Today 75:297
8. Klier J, Tucker CJ, Kalantar TH, Green DP (2000) Adv Mater 18:1751
9. Gan LM, Chew CH (1996) In: Salamone (ed) Polymeric materials encyclopedia. CRC,
Boca Raton, FL, M4321
10. Candau F (1992) In: Paleos CM (ed) Polymerization in organized media. Gordon and
Breach, New York, p 215
11. Candau F (1998) In: Kumar P, Mittal KL (eds) Microemulsions: fundamental and
applied aspects. Marcel Dekker, New York
12. Herrera JR, Peralta RD, Lopez RG, Cesteros LC, Puig JE (2003) Polymer 44:1795
13. Candau F, Leong YS, Pouyet G, Candau SJ (1984) J Colloid Interf Sci 101:167
14. Candau F, Yong YS, Fitch RM (1985) J Polym Sci Pol Chem 23:193
15. Candau F (1987) In: Mark H, Bikales NM, Overberger CG, Menges G (eds) Encyclopedia
of polymer science and engineering. Wiley, New York, 9:718
16. Vaskova V, Juranicova V, Barton J (1990) Makromol Chem 191:717
17. Vaskova V, Hlouskova Z, Barton J, Juranicova V (1990) Makromol Chem 193:267
18. Lezovic M, Ogino K, Sato H, Capek I, Barton J (1998) Polymer Int 46:269
19. Barton J, Kawamoto S, Fujimoto K, Kawaguchi H, Capek I (2000) Polymer Int 49:358
20. Barton J, Capek I (2000) Macromolecules 33:5353
21. Candau F (1989) In: El-Nokaly M (ed) ACS Sym Ser 384, Ch 4
22. Canadu F, Anquetil JY (1998) In: Shah DO (ed) Micelles, microemulsions and mono-
layers. Marcel Dekker, New York, p 193
23. Candau F, Zekhnini Z, Durand JP (1987) Prog Colloid Polym Sci 73:33
24. Corpart JM, Candau F (1993) Colloid Polym Sci 271:1055
25. Corpart JM, Selb J, Candau F (1993) Polymer 34:3873
26. Candau F, Buchert P (1990) Colloids Surface 48:107
27. Candau F, Pabon M, Anqueti (1999) Colloids Surface A 153:47
Microemulsion Polymerizations and Reactions 295
28. Sosa N, Peralta RD, Lopez RG, Ramos LF, Katine I, Certeros C, Mendizabal E, Puig JE
(2001) Polymer 42:6923
29. Braun O, Selb J, Candau F (2001) Polymer 42:8499
30. Candau F, Braun O, Essler F, Stahler K, Selb J (2002) Macromol Symp 179:13
31. Stoffer JO, Bone T (1980) J Disper Sci Technol 1:37
32. Atik SS, Thomas JK (1981) J Am Chem Soc 103:4279
33. Tang HI, Johnson PL, Gulari E (1984) Polymer 25:1357
34. Kuo PL, Turro NJ, Tseng CM, El-Aasser MS, Vanderhoft JW (1987) Macromolecules
20:1216
35. Feng L, Ng KY (1990) Macromolecules 23:1048
36. Gan LM, Chew CH, Lye I (1992) Makromol Chem 193:1249
37. Gan LM, Chew CH, Frieberg SE (1983) J Macromol Sci Chem A 19:739
38. Ferrick MR, Murtagh J, Thomas JK (1989) Macromolecules 22:1515
39. Antonietti M, Bremser W, Muschenborn D, Rosenaur C, Schupp B, Schmidt M (1991)
Macromolecules 24:6636
40. Rodriguez-Guadarrama LA, Mendizabal E, Puig JE, Kaler EW (1993) J Appl Polym Sci
48:775
41. Gan LM, Chew CH, Ng SC, Loh SE (1993) Langmuir 9:2799
42. Gan LM, Chew CH, Lee KC, Ng SC (1993) Polymer 34:3064
43. Larpent C, Tados RF (1991) Colloid Polym Sci 269:1171
44. Full AP, Puig JE, Gron LU, Kaler EW, Minter JR, Meurey TH, Texter J (1992) Macro-
molecules 25:5157
45. Capek I, Potisk P (1995) Eur Polym J 31:1269
46. Capek I, Fouassier JP (1997) Eur Polym J 33:173
47. Capek I, Juranicova V (1996) J Polym Sci Pol Chem 34:575
48. Capek I, Juranicova V, Barton J, Asua JM, Ito K (1997) Polymer Int 43:1
49. Guo JS, Sudol ED, Vanderhoff JW, El-Aasser MS (1992) J Polym Sci Pol Chem 30:691
50. Puig JE, Perez-Luna VH, Macias ER, Rodriguez BE, Kaler EW (1993) Colloid Polym Sci
271:114
51. Co CC, Cotts P, Burauer S, deVries R, Kaler EW (2001) Macromolecules 34:3245
52. Kukulj D, Davis TP, Gilbert RG (1998) Macromolecules 31:994
53. Hentze HP, Kaler EW (2003) Curr Opin Coll Interf Sci 8(2):164–178
54. Mendizabal E, Flores J, Puig JE, Katime I, Lopez-Serrano F, Alvarez J (2000) Macromol
Chem Phys 201:1259
55. Co CC, Kaler EW (1998) Macromolecules 31:3203
56. Morgan JD, Kaler EW (1998) Macromolecules 31:3197
57. Sanghvi PG, Pokhriyal NK, Devis (2002) J Appl Polym Sci 84:1832
58. Gan LM, Chew CH, Lee KC, Ng SC (1994) Polymer 35:2659
59. Loh SE, Gan LM, Chew CH, Ng SC (1995) J Macromol Sci Pure A32:1681
60. Gan LM, Lee KC, Chew CH, Tok ES, Ng SC (1995) J Polym Sci Pol Chem 33:1161
61. Sutterline N, Kurth HJ, Markett G (1976) Angew Makromol Chem 177:1549
62. Aguiar A, Gonzales-Villegas S, Rabelero M, Mendizabal E, Puig JE (1999) Macromole-
cules 32:6767
63. Jang J, Ha H (2002) Langmuir 18:5613
64. Jang J, Lee K (2002) Chem Comm 1098
65. Gan LM, Lian N, Chew CH, Li GZ (1994) Langmuir 10:2197
66. Xu XJ, Siow KS, Wong MK, Gan LM (2001) Langmuir 17:4519
67. Ming W, Jones FN, Fu SK (1998) Polym Bull 40:749
68. Ming W, Jones FN, Fu SK (1998) Macromol Chem Phys 199:1075
69. (a) Xu XJ, Chew CH, Siow KS, Wong MK, Gan LM (1999) Langmuir 15:8067; (b) Chudej
J, Capek I (2002) Polymer 43:1681
296 P. Y. Chow · L. M. Gan
Dispersion Polymerization
Seigou Kawaguchi ( ) 1 · Koichi Ito2
1
Department of Polymer Science and Engineering, Faculty of Engineering,
Yamagata University, 4–3–16 Jonan, 992–8510 Yonezawa, Japan
[email protected]
2
Department of Materials Science, Toyohashi University of Technology,
1–1 Tempaku-cho, 441–8580 Toyohashi, Japan
[email protected]
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
1
Introduction
Micron-size monodisperse polymeric microspheres are used in a wide variety
of applications, such as toners, instrument calibration standards, column pack-
ing materials for chromatography, spacers for liquid crystal displays, and bio-
medical and biochemical analysis [1–3]. Because of the commercial and scien-
tific interest in these particles, research into their preparation has been active
302 S. Kawaguchi · K. Ito
for the past two decades. Micron-size monodisperse particles were usually dif-
ficult to obtain because this size is in-between the diameter range of particles
produced by conventional emulsion polymerization (0.06–0.7 mm) in a batch
process and suspension polymerization (50–1000 mm). Vanderhoff et al. [4, 5]
used the successive seeding method to obtain micron-size monodisperse
polymer particles. The particles were also prepared by Ugelstad et al. [6, 7] by
means of the two-stage swelling method. Omi et al. [8] and Kamiyama [9] used
modified suspension polymerizations, and Okubo et al. [10] developed the
dynamic monomer swelling method (DSM).
Dispersion polymerization is an attractive and promising alternative to
other polymerization methods that affords micron-size monodisperse particles
in a single batch process. Dispersion polymerization may be defined as a type
of precipitation polymerization in which one carries out the polymerization of
a monomer in the presence of a suitable polymeric stabilizer soluble in the
reaction medium. The solvent selected as the reaction medium is a good sol-
vent for both the monomer and the steric stabilizer polymers, but a non-solvent
for the polymer being formed. Dispersion polymerization, therefore, involves
a homogeneous solution of monomer(s) with initiator and dispersant, in which
sterically stabilized polymer particles are formed by the precipitation of the
resulting polymers.As a continuous medium, the properties of the solvent also
change with increasing monomer conversion. Under favorable circumstances,
the polymerization can yield, in a batch step, polymer particles of 0.1–15 mm
in diameter, often of excellent monodispersity. This dispersant polymer can be
formed as a reactive, polymerizable macromonomer. It can be a block copoly-
mer in which one block has an affinity for the surface of the precipitated poly-
mer, or it can be a soluble polymer (a “stabilizer precursor”) to which grafting
is thought to occur during the polymerization reaction. In all instances, this sol-
uble dispersant polymer – a hairy layer – plays a crucial role in the dispersion
polymerization process. By adsorbing or becoming incorporated onto the
surface of the newly-formed precipitated polymers, it acts as a steric stabilizer,
directing the particle size and colloidal stability of the system. This feature of
dispersion polymerization is widely appreciated and well understood (Fig. 1).
Dispersion polymerization in organic hydrocarbon media was first devel-
oped by Osmond and coworkers at ICI [11]. They polymerized acrylic and
vinylic monomers in hydrocarbons with oil soluble polymer stabilizers to pro-
duce nonaqueous dispersions (NAD) of polymer particles. Later, Almog et al.
[12] extended the concept to dispersion polymerization in polar solvents
as a method of forming monodisperse polymeric microspheres. Ober et al.
[13–16], Tseng et al. [17], Okubo et al. [18, 19], and Paine et al. [20–24], among
other authors, studied this technique in order to control particle size and
achieve a narrow particle size distribution. A great deal of research has been
devoted to dispersion polymerization during past two decades, as reviewed by
Croucher and Winnik [25], Guyot and Tauer [26], Cawse [27], Pichot et al. [28],
Asua and Schoonbrood [29], and Ito et al. [30–32]. The present article is
intended to discuss state-of-the-art design of microspheres obtained by dis-
Dispersion Polymerization 303
2
Microsphere Syntheses by Linear and Block Polymer Dispersants
2.1
Functional Microspheres
Functional particles
PS, PMMA derivatives HEMA 2-Butanol/toluene [47]
Cellulose acetate HEMA Alcohol/toluene [48]
butyrate
PS-b-PB 4VP DMF/toluene [49]
PVP GMA MeOH/water or DMF [50]
PVP GMA/DVB Ethanol [51]
PVP ST/GMA Ethanol/water [52]
Cellulose acetate GMA DMF/methanol [53]
PVME AAm t-BuOH/water [54]
None AAm t-BuOH/water [55]
PVP CMS Ethanol/DMSO [56]
PAA CMS Methoxyethanol/MeOH [57]
PS-b-P(PP-alt-E) VPy/EDMA Cyclohexane [58]
PVP ST/Boc-AMST i-Propanol [59]
PVP ST/NVC Ethanol [60]
PVA-co-PVAc ST Methanol [61]
PDMAEMA-b-PBMA ST Alcohols [62]
Hybrid particles
Poly(amic acid) ST/4VBAC Ethanol/water [63]
PVP ST/polyimide prepolymer i-Propanol [64]
PVP GMA/iron oxide Alcohol/water [65]
Cellulose acetate HEMA/EDMA/iron oxide Alcohol/toluene [66]
butyrate
PVP HEMA/GMA/iron oxide Alcohol/toluene [67]
PVP ST/SiO2 Ethanol/water [68]
None 4VP/HPMA/SiO2 Water [69]
Crosslinked particles
PVP DVB Acetonitrile or Ethanol [70]
None DVB Acetonitrile [71]
None DVB/CMS Acetonitrile [72]
Chitosan NIPAM Acetic Acid [73]
PS-b-P(PP-alt-E) Oxazoline methacrylate Heptane [74]
PVP-Aerosol-OT ST/urethane acrylate Ethanol [75]
PVP ST/2,2¢-oxy-bisethanol Ethanol/heptane [76]
diacrylate
PVP MMA/EDMA Ethanol/water [77]
Dynamic swelling method (DSM)
None ST/DVB Ethanol/water [10, 78–83]
Dispersion Polymerization 305
2.2
Living Dispersion Polymerization
Dispersion polymerization has also been applied to the ring opening poly-
merization of e-caprolactone and lactide in heptane-dioxane (4/1 v/v) with
poly(dodecyl methacrylate)-g-poly(e-caprolactone) as stabilizer [97]. Diethyl-
aluminium ethoxide and tin(II) 2-ethylhexanoate were used as initiators in
these two systems, respectively, to obtain functional microspheres with a
narrow particle size distribution and a narrow molecular weight distribution
[98]. Table 2 provides an overview of microspheres obtained by living disper-
sion polymerization.
2.3
Microspheres from Non-Vinyl Monomers
Oxidative polymerization
PVP Aniline Alcohol/water [100, 101]
PVME Aniline Alcohol/water [102, 103]
PVA Aniline Water/alcohol [104–106]
Poly(styrenesulfonic acid) Aniline Water [107]
Methycellulose Aniline Water/alcohol [108]
HPC Aniline Water [109]
PVA 3,5-Xylidine Water [110, 111]
Ethylhydroxycellulose Pyrrole Ethanol/water [112, 113]
PVME Pyrrole Ethanol/water [114]
Enzymatic polymerization
PVME, PVA, PEG Phenol Dioxane/phosphate [115, 116]
buffer
PVME p-Phenylphenol Dioxane/phosphate [115, 116]
buffer
Polyurethane
Poly(lauryl methacrylate) diol TDI/EG Paraffin oil [117]
Polyester
Poly(2-ethylhexyl acrylate-co- 4-Acetoxybenzoic – [118]
styrene-co-acrylic acid) acid + 2,6-aceto-
xynaphthoic acid
308 S. Kawaguchi · K. Ito
3
Microsphere Syntheses by Reactive Dispersants, Macromonomers,
Inimers, and Transurfs
monomers was first carried out in aliphatic hydrocarbon media with the hy-
drophobic macromonomers 1 and 2 [11]. These were copolymerized with
MMA or other polar monomers to produce comb-graft copolymers, which
have limited solubility in pure aliphatic hydrocarbons but adequate solubility
in hydrocarbon-monomer mixtures. It is particularly effective in stabilizing
PMMA NAD particles. Polyethylene (PE) macromonomers 3 have been used
310 S. Kawaguchi · K. Ito
This technique has been extended to polar media, especially alcohols and
their mixtures with water as a continuous phase.Kobayashi and Uyama et al.
[121– 124] reported that poly(2-oxazoline) macromonomers such as 5 and 6 are
very effective for dispersion copolymerization with styrene, MMA, and N-vinyl-
formamide in methanol, ethanol, and mixtures of these alcohols with water.
They reported that the particle size decreased with increasing initial macro-
monomer concentration and that graft-copolymerized poly(2-oxazoline) chains
are concentrated on the particle surface to act as steric stabilizers.
Dispersion copolymerizations that use the poly(ethylene oxide) (PEO)
macromonomers 7–13 in alcoholic media have been intensively studied by
Dispersion Polymerization 311
The polyelectrolyte macromonomers 17, 20, 22, and 23 [142, 143, 147, 148]
were prepared and applied to dispersion copolymerizations to produce poly-
meric particles covered with polyelectrolyte chains. Evidently, the dependence
of the conformational properties of polyelectrolyte brush chains attached to
the latex surface on the pH, the degree of neutralization, and the salt concen-
tration have been the subject of growing experimental and theoretical effort.
314 S. Kawaguchi · K. Ito
Some novel water soluble macromonomers, 24, have been synthesized by the
oxyanionic polymerization [149] of 2-(dimethylamino)ethyl, 2-(diisopropyl-
amino)ethyl, and 2-(N-morpholino)ethyl methacrylate, and conducted to dis-
persion copolymerization of styrene in alcohol media [150]. Sufobetaine-based
macromonomer was prepared by the polymer reaction of 24 (R=CH3) with
propane sultone, and was found to be useful in the dispersion polymerization
of styrene even at high electrolyte levels (up to 1 M NaCl). Ito et al. [151] syn-
thesized new PEO macromonomers with a cationic charge at the w-end, 25, and
examined the influence of the charge on the particles’ size in dispersion copoly-
merization with styrene in alcohol media.
4
Particle Size Control in Dispersion (Co-)Polymerization
4.1
Theoretical Model
Fig. 3 Schematic model for the particle nucleation and growth of sterically-stabilized
particles in dispersion polymerization
In 1990, Paine [24] first developed a multibin kinetic model for the aggregation
of precipitated radicals or unstabilized particles in dispersion polymerization,
based on diffusion-controlled particle aggregation, which was described by
Smoluchowski [160] and Frenklach [161]. The stabilized particles whose sur-
faces are completely covered with grafted polyvinylpyrrolidone (PVP) chains
do not aggregate. This was the first model that quantitatively simulated the role
of the stabilizer molecules at the particle formation stage.Assuming ideal core-
shell structures in which the main monomer forms the core, and dispersant the
shell, one can readily establish a basic relationship between microsphere size
and weights of main core polymer and stabilizer polymer attached on the
particle surface, as given by the following equations:
4
W = Woq = 3 p R3 ÇN (1)
3
and
WDoqD NAS
nS = 08 = 4p R2 (2)
NMD
2 1 1 1
1 3Wo 32 MD 22 0.386k2 62 kt 12
22
R= q 32 (4)
ÇNA
7 CSWDoScrit
06 4pkt
02 2kd f [I]o
04
1 3Wo 32 MDr1 22 0.386k2 62 kt 12
22
R= q 32 (5)
ÇNA
7 WDoScrit
03 4pkp
02 2kd f [I]o
04
1 1 1 1
– 21 3Wo 32 MDScrit 22 4pkp 62 2kd f [I]o 12
22 q
S= q 3 (6)
Ç
7NA rW
02
1 Do 0.386k2
02 04
kt qD
4
4.2
Comparison of Experiment with Theory
Fig. 4 Double logarithmic plots of average particle radius (R) as a function of weight (WDo)
of PEO macromonomer. a Styrene 7 (m=4 and n=45), Wo=100 g/L, [I]0=0.0122 mol/L, q=1
at 60 °C. The straight line is a theoretical curve calculated from Eq. 5 with the parameters in
the text. b MMA 7 (m=1 and n=45), Wo=100 g/L, [I]0=0.012 mol/L, q=1 at 60 °C in methanol:
water=8:2 (empty triangles), =7:3 (empty circles), =6:4 (empty squares), and =5:5 (filled
squares)
Fig. 4b. The significant change in the value of the exponent with the polarity of
the continuous phase cannot be simply explained by the current model, so
further refinements [163–165] are needed.
The criteria for designing PEO macromonomers to be used as efficient dis-
persants in dispersion polymerization have been thoroughly studied by Ito
et al. [128, 131, 135, 151], who examined the effects of the length of spacer (m),
the degree of polymerization of PEO chain (n) of 7, and charge group at the
w-end of 25 on particle size.
Dispersion Polymerization 321
5
Chain Conformation of Grafted Polymer Chains at the Particle Surface
Fig. 6 Plots of mobile fraction of surface anchored PEO chains against the estimated mean
separation D between PEO anchor points on the surface of the particles. The D values were
calculated from the particle size and number, assuming that all PEO chains were located at
the surface
occurs when, say, D≈<S2>1/2, as expected by theory [162, 169, 170]. This subject
should lead to a better understanding of the true nature of the steric stabiliza-
tion that exists in many dispersion systems.
6
Conclusions and Future
References
1. Ugelstad J, Mørk PC, Mfutakanba HR, Soleimany E, Nordhuus I, Schmid R, Berge A,
Ellingsen T, Aune O, Nustad K (1983) In: Poehlein GW, Ottewill RH, Goodwin JW (eds)
Science and technology of polymer colloids. Nijhoff, Boston, MA
2. Ober CK, Lok KP (1986) US Pat 4617249
3. Ugelstad J, Berge A, Ellingsen T, Schmid R, Nilsen T-N, Mørk PC, Stenstad P, Hornes E,
Olsvik Ø (1992) Prog Polym Sci 17:87
4. Vanderhoff JV, El-Aasser MS, Micale FJ, Sudol ED, Tseng CM, Silwanowicz A, Kornfeld
DM, Vicente FA (1984) J Dispers Sci Technol 5:231
5. Vanderhoff JV, El-Aasser MS, Micale FJ, Sudol ED, Tseng CM, Silwanowicz A, Sheu HR,
Kornfeld DM (1986) Polym Mater Sci Eng 54:1986
6. Ugelstad J, Mørk PC, Kaggerud KH, Ellingsen T, Berge A (1980) Adv Coll Interfac Sci
13:101
7. Ugelstad J, Mørk PC, Kaggerud KH, Ellingsen T, Khan AA (1982) In: Piirma (ed) Emul-
sion polymerization. Academic, New York, Ch 11
8. Omi S, Katami K, Yamamoto A, Iso M (1994) J Appl Polym Sci 51:1
9. Kamiyama M, Koyama K, Matsuda H, Sano Y (1993) J Appl Polym Sci 50:107
10. Okubo M, Shiozaki M, Tsujihiro M, Tsukada Y (1991) Coll Polym Sci 269:222
11. Barret KEJ (ed) (1975) Dispersion polymerization in organic media. Wiley, London
12. Almog Y, Reich S, Levy M (1982) Brit Polym J 14:131
13. Lok KP, Ober CK (1985) Can J Chem 63:209
14. Ober CK, Lok KP, Hair ML (1985) J Polym Sci Pol Lett 23:103
15. Ober CK, Hair ML (1987) J Polym Sci Pol Chem 25:1395
16. Ober CK, Lok KP (1987) Macromolecules 20:268
17. Tseng CM, Lu YY, El-Aasser MS, Vanderhoff JW (1986) J Polym Sci Pol Chem 24:2995
18. Okubo M, Ikegami K, Yamamoto Y (1989) Colloid Polym Sci 267:193
19. Okubo M, Katayama Y, Yamamoto Y (1991) Colloid Polym Sci 269:217
20. Paine AJ, McNulty J (1990) J Polym Sci Pol Chem 28:2569
21. Paine AJ (1990) J Colloid Interf Sci 138:157
22. Paine AJ (1990) J Polym Sci Pol Chem 28:2485
23. Paine AJ, Luymes W, McNulty (1990) Macromolecules 23:3104
24. Paine AJ (1990) Macromolecules 23:3109
25. Croucher MD, Winnik MA (1990) In: Candau F, Ottewill RH (eds) An introduction to
polymer colloids. Kluwer Academic, Dordrecht, p 35
26. Guyot A, Tauer K (1994) Adv Polym Sci 111:43
27. Cawse JL (1997) In: Lovell PA, El-Aasser MS (eds) Emulsion polymerization and emul-
sion polymers. Wiley, Chichester, UK, p 743
28. Pichot C, Delair T, Elaissari A (1997) In: Asua JM (ed) Polymeric dispersions: principles
and applications. Kluwer Academic, Dordrecht, p 515
29. Asua JM, Schoonbrood HAS (1998) Acta Polym 49:671
30. Ito K (1998) Prog Polym Sci 23:581
31. Ito K, Kawaguchi S (1999) Adv Polym Sci 142:129
32. Ito K, Cao J, Kawaguchi S (2002) In: Arshady R, Guyot A (eds) Functional colloids (MML
Series, Vol 4). Citus, London, p 109
33. Sáenz J, Asua J M (1995) J Polym Sci Pol Chem 33:1511
34. Sáenz J, Asua J M (1995) J Polym Sci Pol Chem 34:1977
35. Bamnolker H, Margel S (1996) J Polym Sci Pol Chem 34:1857
36. Horák D, Švec F, Fréchet JMJ (1995) J Polym Sci Pol Chem 33:2961
37. Kiatkamjornwong S, Kongsupapsiri C (2000) Polym Int 49:1395
Dispersion Polymerization 325
38. Yang W, Yang D, Hu J, Wang C, Fu S (2001) J Polym Sci Pol Chem 39:555
39. Takahashi K, Nagai K (1996) Polymer 37:1257
40. Shen S, Sudol ED, El-Aasser MS (1994) J Polym Sci Pol Chem 32:1087
41. Horák D, Švec F, Fréchet JMJ (1995) J Polym Sci Pol Chem 33:2329
42. Takahashi K, Uyama H, Kobayashi S (1998) J Macromol Sci Pure A35:1473
43. Ho C-H, Chen S-A, Amiridis MD, Zee JWV (1997) J Polym Sci Pol Chem 35:2907
44. Chen Y, Yang H-W (1992) J Polym Sci Pol Chem 30:2765
45. Arhady R (1999) In: Arshady R (ed) Microspheres, microcapsules & liposomes, vol 1.
Citus, London, p 11
46. Margel S (1999) In: Arshady R (ed) Microspheres, microcapsules & liposomes, vol 2.
Citus, London, p 11
47. Takahashi K, Miyanori S, Uyama H, Kobayashi S (1996) J Polym Sci Pol Chem 34:175
48. Horák D (1999) J Polym Sci Pol Chem 37:3785
49. Takahashi K, Miyamori S, Uyama H, Kobayashi S (1997) Macromol Rapid Commun
18:471
50. Takahashi K, Uyama H, Kobayashi S (1998) Polym J 30:684
51. Park KY, Jeong WW, Suh KD (2003) J Macromol Sci A50:617
52. Yang W, Hu J, Tao Z, Li L, Wang C, Fu S (1999) Colloid Polym Sci 277:446
53. Horák A, Shapoval P (2000) J Polym Sci Pol Chem 38:3855
54. Ray B, Mandal BM (1999) J Polym Sci Pol Chem 37:493
55. Lee K, Lee S, Song B, Lee D (2000) Polymer (Korea) 24:629
56. Margel S, Nov E, Fisher I (1991) J Polym Sci Pol Chem 29:347
57. Bahar T, Tuncel A (1999) Polym Eng Sci 39:1849
58. Horák A, Kryštůfek M, Spĕváček J (2000) J Polym Sci Pol Chem 38:653
59. Covolan VL, D’Antone S, Ruggeri G, Chiellini E (2000) Macromolecules 33:6685
60. Yang W, Zhou H, Tao Z, Hu J, Wang C, Fu S (2000) J Macromol Sci Pure A37:659
61. Dawkins JV, Neep DJ, Shaw PL (1994) Polymer 35:5366
62. Baines FL, Dionisio S, Billingham NC, Armes SP (1996) Macromolecules 29:3096
63. Watanabe S, Ueno K, Kudoh K, Murata M, Masuda Y (2000) Macromol Rapid Commun
21:1323
64. Omi S, Saito M, Hashimoto S, Nagai M, Ma G-H (1998) J Appl Polym Sci 68:897
65. Horák D (2001) J Polym Sci Pol Chem 39:3707
66. Horák D, Boháček J, Šubrt M (2000) J Polym Sci Pol Chem 38:1161
67. Horák D, Semenyuk N, Lednicky F (2000) J Polym Sci Pol Chem 41:1848
68. Bourgeat-Lami E, Lang J (1998) J Colloid Interf Sci 197:293
69. Percy MJ, Michailidou V, Armes SP, Perruchot C, Watts JF, Greaves SJ (2003) Langmuir
19:2072
70. (a) Li K, Stöver HDH (1993) J Polym Sci Pol Chem 31:2473; (b) Li K, Stöver HDH (1993)
J Polym Sci Pol Chem 31:3257
71. (a) Downey JS, Frank RS, Li W-H, Stöver HDH (1999) Macromolecules 32:2838;
(b) Li W-H, Stöver HDH (2000) Macromolecules 33:4354
72. Li WH, Li LK, Stöver HDH (1999) J Polym Sci Pol Chem 37:2295
73. Lee C-F, Wen C-J, Chiu Q-Y (2003) J Polym Sci Pol Chem 41:2063
74. Hölderle M, Bar G, Mülhaupt R (1997) J Polym Sci Pol Chem 35:2539
75. Kim J-W, Suh K-D (1998) Colloid Polym Sci 276:878
76. Huang J, Zhang H, Hou J, Jiang P (2002) React Funct Polym 53:1
77. Huang J-X, Yuan X-Y, Yu X-L, Zhang H-T (2003) Polym Int 52:819
78. Okubo M, Nakagawa T (1992) Colloid Polym Sci 270:853
79. Okubo M, Minami H (1996) Macromol Symp 101:509
80. Okubo M, Minami H (1996) Colloid Polym Sci 274:433
326 S. Kawaguchi · K. Ito
156. (a) Bourgeat-Lami E, Guyot A (1997) Colloid Polym Sci 275:716; (b) Schipper ETWM,
Sindt O, Hamaide T, Lacroix-Desmazes P, Müller B, Huyot A, van den Enden MJWA,
Vidal F, van Es JJGS, German AL, Montaya Goñi AM, Sherrington DC, Schoonbrood
HAS, Asua JM, Sjöberg M (1998) Colloid Polym Sci 276:402
157. Feeney PJ, Napper DH, Gilbert RG (1987) Macromolecules 20:2922
158. Fitch RM, Tsai CH (1971) In: Fitch RM (ed) Polymer colloids. Plenum, New York
159. Gilbert RG (1995) In: Emulsion polymerization: A mechanistic approach. Academic,
New York
160. Smoluchowski MV (1917) Z Phys Chem 192:129
161. Frenklach M (1985) J Colloid Interf Sci 108:237
162. de Gennes PG (1980) Macromolecules 13:1069
163. Yasuda M,Yokoyama H, Seki H, Ogino H, Ishimi K, Ishikawa H (2001) Macromol Theor
Simul 10:54
164. Yasuda M, Seki H,Yokoyama H, Ogino H, Ishimi K, Ishikawa H (2001) Macromolecules
34:3261
165. Lacroix-Desmazes P, Guillot J (1998) J Polym Sci Pol Phys 36:325
166. Cosgrove T Griffiths PC (1992) Adv Colloid Interfac 42:175
167. Fleer GJ, Cohen Stuart MA, Scheutjens JMHM, Cosgrove T, Vincent B (1993) In: Poly-
mers at interfaces. Chapman & Hall, London
168. Kawaguchi M, Takahashi A (1992) Adv Colloid Interfac 37:219
169. de Gennes PG (1987) Adv Colloid Interfac 278:189
170. Alexander S (1977) J Phys (Paris) 38:983
171. Auroy P, Auvray L, Leger LL (1991) Physica A 172:269
172. Taunton HJ, Toprakcioglu C, Fetters LJ, Klein J (1990) Macromolecules 23:571
173. Cairns RJR, Ottewill RH, Osmond DWJ, Wagstaff I (1976) J Colloid Interf Sci 54:45
174. Jeong BJ, Lee JH, Lee HB (1996) J Colloid Interf Sci 178:757
175. Werts MPL, van der Vegte EW, Hadziioannou G (1997) Langmuir 13:4939
176. Wu C, Akashi M, Chen MQ (1997) Macromolecules 30:2187
177. Kawaguchi S, Winnik MA, Ito K (1996) Macromolecules 29:4465
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
Abstract In recent years, carbon dioxide has been investigated for its potential to replace the
aqueous and organic solvents used in polymerization processes. Carbon dioxide has several
benefits, including being environmentally benign, a tunable solvent, resistant to chain trans-
fer, and of low viscosity which facilitates high initiator efficiencies and fluid handling. Homo-
and copolymerization of fluoroolefins such as tetrafluoroethylene (TFE), vinylidene fluoride
(VF2), hexafluoropropylene (HFP), and perfluorinated vinyl ethers in carbon dioxide are
particularly interesting because of the significant waste reduction and the elimination of po-
tentially harmful processing agents. This review focuses on recent developments in poly-
merizing fluoroolefins via heterogeneous polymerization in carbon dioxide. At least one
CO2-based polymerization process has recently been commercialized, and additional re-
search is underway to understand the kinetics associated with polymerizing fluoroolefins
330 K. A. Kennedy et al.
Abbreviations
AIBN 2,2¢-azobis(isobutyronitrile)
APFO ammonium perfluorooctanoate
BPPP Bis(perfluoro-2-N-propoxypropionyl) peroxide
C8 ammonium perfluorooctanoate
CMC Critical micelle concentration
CSTR Continuous stirred tank reactor
CTFE Chlorotetrafluoroethylene
DEPDC Diethyl peroxydicarbonate
EPA Environmental Protection Agency
FEP Poly(tetrafluoroethylene-co-hexafluoropropylene)
FMG Fluoropolymers Manufacturers Group
FTIR Fourier transform infrared spectroscopy
HFP Hexafluoropropylene
LDPE Low-density polyethylene
Mn Number-average molecular weight
MAC Maleic acid
MAN Maleic anhydride
MFI Melt flow index
MWD Molecular weight distribution
NMR Nuclear magnetic resonance spectroscopy
PFA Poly(tetrafluoroethylene-co-perfluoro(propyl vinyl ether))
PMVE Perfluoro(methyl vinyl ether)
PPVE Perfluoro(propyl vinyl ether)
PTFE Poly(tetrafluoroethylene)
PVDF Poly(vinylidene fluoride)
scCO2 Supercritical carbon dioxide
SCF Supercritical fluid
TFE Tetrafluoroethylene
VF Vinyl fluoride
VF2 Vinylidene fluoride
1
Introduction
Table 1 Critical temperatures and pressures of substances useful as supercritical fluids [9]
2
Conventional Polymerization of Fluoroolefins
2.1
Tetrafluoroethylene
2.2
Vinylidene Fluoride
used to control the molecular weight [13]. The resulting latex of dispersed
particles is coagulated, washed, and spray dried to form a fine powder. The
powder consists of spherical agglomerates 0.2–0.5 mm in diameter.
The suspension polymerization is conducted using monomer-soluble per-
oxy initiators.Water-soluble polymers, such as poly(vinyl alcohol), are typically
used as suspending agents to reduce the coalescence of the polymer particles
[17].A slurry of polymer particles 30–100 mm in diameter is formed during the
polymerization. The particles are washed and dried before further processing.
The effect of polymerization conditions on the polymer morphology in
PVDF has been studied [16, 21]. Regioisomer defects in the form of reversed
head-to-tail addition are common in the polymerization of vinyl monomers
[22]. These defects affect the crystallinity and final properties of the polymer.
The fraction of defects in PVDF ranges from 3.5 to 6 mol%, increasing with
reaction temperature [22]. The a-phase crystal is the most common crystal
observed in PVDF. However, an increase in defects leads to a greater amount of
b-phase crystals in the polymer [16, 23], which gives PVDF its piezoelectric
properties. An increase in the amount of b crystals formed was observed for
different initiating systems [23].Additionally, enhancement of the b crystals in
PVDF is achieved by applying tensile forces or mechanical stretching to the
polymer [24]. The g and d crystal phases may be generated from the a-phase
using heat or an electric field, respectively. The g-phase crystal has also been
formed by crystallization under high pressure [25].
Commercial PVDF is sold in powder form or may be compression molded
into pellets. Depending on the polymerization conditions, differing properties
of PVDF may be produced. PVDF exhibiting a multimodal molecular weight
distribution has been reported due to its improved melt flow characteristics
[26] and its use in lithium batteries [24]. Dohany [26] reports the formation of
a distinct bimodal molecular weight distribution that is strongly dependent on
the addition of the initiator to the reaction. The initiator is continuously fed to
the reaction until about 50% of the total monomer feed is added to the reactor.
At that point, the addition of the initiator is stopped and the remaining
monomer is continuously fed to the reactor. The high molecular weight peak
measured by gel permeation chromatography is more than 30% of the overall
molecular weight distribution. The high molecular weight portion allows pro-
cessing of the polymer at high shear rates for improved productivity. The use
of these polymers in lithium batteries is largely the result of their relatively
high defect structure, which promotes ionic conductivity in the electrochem-
ical cell [24].
2.3
Copolymers
PVDF and PTFE have wide application in industry due to their strength, chem-
ical and wear resistance, and dielectric properties. However,VF2 and TFE may
be copolymerized with comonomers such as hexafluoropropylene (HFP) and
Heterogeneous Polymerization of Fluoroolefins in Supercritical Carbon Dioxide 335
vinyl ethers to modify the properties of the polymer. For example, copolymer-
izing VF2 with HFP improves the flexibility of the polymer without compro-
mising the electrical properties. Therefore,VF2-HFP copolymers are used more
commonly in cable insulation than VF2 homopolymer. Copolymers of TFE
with perfluoroalkyl vinyl ethers have properties very similar to the TFE homo-
polymer except that the copolymer may be processed by injection molding
because of its lower melt viscosity. Copolymers of TFE and VF2 can also be
made to form fluorinated elastomers and perfluorinated elastomers when cure
site monomers are used. Important examples of such elastomers include
terpolymers of VF2, TFE, and either HFP or PMVE along with a cure site
monomer such as bromotetrafluorobutene [19].
Copolymers are produced by aqueous copolymerization of the monomers
in a manner similar to the homopolymerizations. Nonaqueous polymerizations
of TFE with PPVE may be conducted in water or in fluorinated solvents at lower
temperatures (30–60 °C) using a soluble organic initiator such as perfluoro-
propionyl peroxide. The molecular weight is controlled by addition of chain
transfer agents such as methanol in nonaqeuous polymerizations or hydrogen
in aqeous polymerizations.
3
Supercritical Carbon Dioxide as a Polymerization Medium
3.1
Properties of Supercritical Carbon Dioxide
Supercritical fluids possess characteristics that make them interesting for use
as polymerization media.A supercritical fluid exists at temperatures and pres-
sures above its critical values. In the supercritical state, the fluid exhibits phys-
ical and transport properties intermediate between the gaseous and liquid
state. This is illustrated in Table 2. SCFs have liquid-like densities, but gas-like
diffusivities. These intermediate properties can provide advantages over liquid-
based processes. In particular, the higher diffusivities of SCFs reduce mass
transfer limitations in diffusion-controlled processes.Additionally, lower energy
is required for processing the supercritical fluid because its viscosity is lower
than that of most liquids, and because the need to vaporize large quantities of
liquid is avoided.
SCFs have a tunable density that may offer further advantages in reaction
and processing applications. This tunability is illustrated in Fig. 1 for carbon
dioxide. Near the critical point, even small changes in the temperature or
pressure of carbon dioxide dramatically affect its density. Similarly, the viscos-
ity, dielectric constant, and diffusivity are also tunable parameters, which allows
specific control of systems involving supercritical fluids.
Carbon dioxide is inexpensive and widely available. The main source of
carbon dioxide is from chemical manufacturing. For example, it is a major
byproduct in the production of ammonia, ethanol, and hydrogen. Finally, car-
bon dioxide is non-toxic, nonflammable, and easily recycled.
3.2
Advantages of Using Supercritical Carbon Dioxide
Supercritical carbon dioxide is a very good solvent for small molecules, but a
poor solvent for most high molecular weight polymers at mild conditions
(T<100 °C, P<350 bar). Amorphous fluoropolymers and silicones are the only
polymers known to be soluble in CO2 at mild conditions [6]. This difference in
solubilities is an advantage for CO2-based polymerizations, as it can be used to
reduce the energy requirements necessary to separate and purify a polymer
after synthesis. Consider, for example, a batch precipitation polymerization in
Heterogeneous Polymerization of Fluoroolefins in Supercritical Carbon Dioxide 337
4
Precipitation Polymerization in Supercritical Carbon Dioxide
4.1
Introduction
4.2
Polymerization of Tetrafluoroethylene
Several alternative polymerization media have been proposed for reducing the
amount of unstable end groups in poly(tetrafluoroethylene). These include
chlorofluorocarbons, which are detrimental to the environment, perfluoro-
carbons, hydrofluorocarbons, and perfluoroalkyl sulfide acids, which are all
expensive. Supercritical carbon dioxide has been identified as a viable alter-
native to aqueous and fluorocarbon reaction media [31]. Furthermore, mixing
Heterogeneous Polymerization of Fluoroolefins in Supercritical Carbon Dioxide 339
4.3
Polymerization of Vinylidene Fluoride
4.3.1
Continuous Polymerization
4.3.1.1
Polymer Properties
4.4
Synthesis of Copolymers
Copolymerizations of TFE [31, 47, 57–61] and VF2 [57, 58, 61, 62] in scCO2 have
been reported. Romack and coworkers [47, 60] did early work to perform
precipitation polymerizations of TFE with HFP and TFE with perfluoropropyl
vinyl ether (PPVE) at 35 °C and moderate pressures (<133 bar).A chain transfer
agent, methanol, was necessary to limit the molecular weight to a range of com-
mercial interest. Hydrogen chloride and hydrogen bromide have also been re-
ported as effective chain transfer agents for these polymerizations in CO2 [58].
Romack and coworkers concluded that the significantly higher molecular
weights produced in CO2 were due to the fact that chain transfer and b-scission
reactions occurred to a lesser extent in scCO2, especially at the lower tempera-
tures used. These undesirable reactions are commonplace in the copolymeriza-
tion of TFE/PPVE conducted at elevated temperatures in conventional solvents.
The limited b-scission reactions were attributed to the enhanced diffusion of
TFE into the polymer phase to improve the rates of radical cross-propagation
with TFE. It was demonstrated that conducting the polymerization in CO2 at
lower temperatures improved the effectiveness of the propagation reaction over
that of the chain transfer reactions to produce high molecular weight polymers.
The research group of Shoichet [57, 59] also has demonstrated that using
CO2 as a polymerization medium can improve the effectiveness of the propaga-
tion reaction over chain transfer reactions in fluoroolefin synthesis. Shoichet’s
344 K. A. Kennedy et al.
group polymerized vinyl acetate with fluoroolefins in CO2 with [59] and with-
out [57] surfactants. The copolymers formed in CO2 were linear, compared to
the highly branched copolymers formed by aqueous polymerization. They con-
cluded that because radical hydrogen abstraction from vinyl acetate was signi-
ficantly reduced in CO2, branching of the polymer was practically eliminated.
There has been research on new, novel fluoroolefin-based copolymers that
benefit from synthesis in scCO2. In 2000,Wheland and Brothers [61] described
novel copolymers of fluoroolefins with maleic anhydride (MAN) and maleic
acid (MAC) for use as coatings and as adhesives or compatibilizing agents for
fluoropolymers. Currently, maleic anhydride must be grafted onto existing
fluoropolymers to achieve the desired properties. Maleic acid does not readily
copolymerize with fluoroolefins in the presence of water. Therefore, polymer-
ization of MAN or MAC with fluoroolefins in CO2 provides a way to synthesize
the desired fluoropolymer directly. There is an interest in using 157 nm photo-
resists in microchip manufacturing to create smaller and smaller features
in circuitry. Zannoni and DeSimone [62] have copolymerized fluoroolefins
with norbornene and a norbornene analog for use as 193 nm and 157 nm pho-
toresists. Ultimately, these photoresists may be synthesized, processed and de-
veloped entirely in CO2, with dramatically reduced solvent use, revolutionizing
the manufacture of microchips.
5
Conclusions
References
1. Swallow JC (1960) The history of polyethene. In: Renfrew A, Morgan P (eds) Polyethene:
The technology and uses of ethylene polymers. Interscience, New York, p 1
2. Kirby CF, McHugh MA (1999) Chem Rev 99:565
3. Folie B, Radosz M (1995) Ind Eng Chem Res 34:1501
4. Guan Z, Combes JR, Menceloglu YZ, DeSimone JM (1993) Macromolecules 26:2663
5. McCoy M (1999) Chem Eng News 77:10
6. Canelas DA, DeSimone JM (1997) Adv Polym Sci 133:103
7. Cooper AI (2000) J Mater Chem 10:207
8. Kendall JL, Canelas DA, Young JL, DeSimone JM (1999) Chem Rev 99:543
9. Manivannan G, Sawan SP (1998) The supercritical state. In: McHardy J, Sawan SP (eds)
Supercritical fluid cleaning: fundamentals, technology and applications. Noyes, West-
wood, NJ, p 1
10. EPA (2003) Federal Register: Perfluorooctanoic acid (PFOA) and fluorinated telomers.
US EPA, Washington, DC (http://www.epa.gov/opptintr/pfoa/)
11. Lee J (2003) Chemical might pose health risk to younger women and girls. New York
Times, New York
12. Ring K-L, Kalin T, Kishi A (2002) Fluoropolymers. Chemical economics handbook. SRI
International, Menlo Park, CA
13. Feiring AE (1994) Fluoroplastics. In: Banks RE, Smart BE, Tatlow JC (eds) Organofluorine
chemistry: Principles and commercial applications. Plenum, New York, p 339
14. Gangal SV (1985) Polytetrafluoroethylene, homopolymer of tetrafluoroethylene. In:
Mark HF, Bikales NM, Overberger CG, Menges G (eds) Encyclopedia of polymer science
and engineering, vol 16. Wiley, New York, p 577
15. Tervoort T, Visjager J, Graf B, Smith P (2000) Macromolecules 33:6460
16. Lovinger AJ (1982) Poly(vinylidene fluoride). In: Bassett DC (ed) Developments in crystal-
line polymers, vol 1. Applied Science, London, p 195
17. Dohany JE (1994) Poly(vinylidene fluoride). In: Kroschwitz JI (ed) Kirk-Othmer encyclo-
pedia of chemical technology, vol 11. Wiley, New York, p 694
18. Seiler DA (1997) PVDF in the chemical process industry. In: Scheirs J (ed) Modern fluoro-
polymers: High performance polymers for diverse applications. Wiley, Chichester, UK,
p 487
19. Scheirs J (2001) Fluoropolymers: Technology, markets, and trends. Rapra Technology
Limited, Shawbury, UK
20. Atofina Chemicals (2003) Internet brochure: Kynar PVDF for lithium batteries. Atofina
Chemicals, Philadelphia, PA (http://www.atofinachemicals.com/kynarglobal/kynar-
literature.cfm)
21. Kochervinskii VV, Danilyuk TY, Madorskaya LY (1986) Polym Sci USSR 28:690
22. Dohany JE, Humphrey JS (1985) Vinylidene fluoride polymers. In: Mark HF, Kroschwitz
JI (eds) Encyclopedia of polymer science and engineering, vol 17.Wiley, New York, p 532
23. Jo SM, Lee WS, Oh HJ, Park S, Ahn BS, Park KY (1999) Polymer (Korea) 23:800
24. Wille RA, Burchill MT (1998) WO 9838687
25. Doll WW, Lando JB (1968) J Macromol Sci Phys 2:219
26. Dohany JE (1978) US 4076929
27. Savage PE, Gopalan S, Mizan TI, Martino CJ, Brock EE (1995) AIChE J 41:1723
28. DeSimone JM (2002) Science 297:799
29. DuPont (2002) Press Release: DuPont introduces fluoropolymers made with supercritical
CO2 technology. DuPont, Wilmington, DE
30. DeSimone JM (1996) US 5496901
346 Heterogeneous Polymerization of Fluoroolefins in Supercritical Carbon Dioxide
de, Abajo, J. and de la Campa, J. G.: Processable Aromatic Polyimides. Vol. 140, pp. 23–60.
Abetz, V. see Förster, S.: Vol. 166, pp. 173–210.
Adolf, D. B. see Ediger, M. D.: Vol. 116, pp. 73–110.
Aharoni, S. M. and Edwards, S. F.: Rigid Polymer Networks. Vol. 118, pp. 1–231.
Albertsson, A.-C. and Varma, I. K.: Aliphatic Polyesters: Synthesis, Properties and Appli-
cations. Vol. 157, pp. 99–138.
Albertsson, A.-C. see Edlund, U.: Vol. 157, pp. 53–98.
Albertsson, A.-C. see Söderqvist Lindblad, M.: Vol. 157, pp. 139–161.
Albertsson, A.-C. see Stridsberg, K. M.: Vol. 157, pp. 27–51.
Albertsson, A.-C. see Al-Malaika, S.: Vol. 169, pp. 177–199.
Al-Malaika, S.: Perspectives in Stabilisation of Polyolefins. Vol. 169, pp. 121–150.
Améduri, B., Boutevin, B. and Gramain, P.: Synthesis of Block Copolymers by Radical
Polymerization and Telomerization. Vol. 127, pp. 87–142.
Améduri, B. and Boutevin, B.: Synthesis and Properties of Fluorinated Telechelic Monodis-
persed Compounds. Vol. 102, pp. 133–170.
Amselem, S. see Domb, A. J.: Vol. 107, pp. 93–142.
Andrady, A. L.: Wavelenght Sensitivity in Polymer Photodegradation. Vol. 128, pp. 47–94.
Andreis, M. and Koenig, J. L.: Application of Nitrogen–15 NMR to Polymers. Vol. 124,
pp. 191–238.
Angiolini, L. see Carlini, C.: Vol. 123, pp. 127–214.
Anjum, N. see Gupta, B.: Vol. 162, pp. 37–63.
Anseth, K. S., Newman, S. M. and Bowman, C. N.: Polymeric Dental Composites: Properties
and Reaction Behavior of Multimethacrylate Dental Restorations. Vol. 122, pp. 177–218.
Antonietti, M. see Cölfen, H.: Vol. 150, pp. 67–187.
Armitage, B. A. see O’Brien, D. F.: Vol. 126, pp. 53–58.
Arndt, M. see Kaminski, W.: Vol. 127, pp. 143–187.
Arnold Jr., F. E. and Arnold, F. E.: Rigid-Rod Polymers and Molecular Composites. Vol. 117,
pp. 257–296.
Arora, M. see Kumar, M. N. V. R.: Vol. 160, pp. 45–118.
Arshady, R.: Polymer Synthesis via Activated Esters: A New Dimension of Creativity in
Macromolecular Chemistry. Vol. 111, pp. 1–42.
Auer, S. and Frenkel, D.: Numerical Simulation of Crystal Nucleation in Colloids. Vol. 173,
pp. 149–208.
Bahar, I., Erman, B. and Monnerie, L.: Effect of Molecular Structure on Local Chain
Dynamics: Analytical Approaches and Computational Methods. Vol. 116, pp. 145–206.
Ballauff, M. see Dingenouts, N.: Vol. 144, pp. 1–48.
Ballauff, M. see Holm, C.: Vol. 166, pp. 1–27.
Ballauff, M. see Rühe, J.: Vol. 165, pp. 79–150.
348 Author Index Volumes 101–175
Baltá-Calleja, F. J., González Arche, A., Ezquerra, T. A., Santa Cruz, C., Batallón, F., Frick, B.
and López Cabarcos, E.: Structure and Properties of Ferroelectric Copolymers of
Poly(vinylidene) Fluoride. Vol. 108, pp. 1–48.
Barnes, M. D. see Otaigbe, J.U.: Vol. 154, pp. 1–86.
Barshtein, G. R. and Sabsai, O. Y.: Compositions with Mineralorganic Fillers. Vol. 101,
pp. 1–28.
Baschnagel, J., Binder, K., Doruker, P., Gusev, A. A., Hahn, O., Kremer, K., Mattice, W. L.,
Müller-Plathe, F., Murat, M., Paul, W., Santos, S., Sutter, U. W. and Tries, V.: Bridging
the Gap Between Atomistic and Coarse-Grained Models of Polymers: Status and
Perspectives. Vol. 152, pp. 41–156.
Batallán, F. see Baltá-Calleja, F. J.: Vol. 108, pp. 1–48.
Batog, A. E., Pet’ko, I.P. and Penczek, P.: Aliphatic-Cycloaliphatic Epoxy Compounds and
Polymers. Vol. 144, pp. 49–114.
Barton, J. see Hunkeler, D.: Vol. 112, pp. 115–134.
Bell, C. L. and Peppas, N. A.: Biomedical Membranes from Hydrogels and Interpolymer
Complexes. Vol. 122, pp. 125–176.
Bellon-Maurel, A. see Calmon-Decriaud, A.: Vol. 135, pp. 207–226.
Bennett, D. E. see O’Brien, D. F.: Vol. 126, pp. 53–84.
Berry, G. C.: Static and Dynamic Light Scattering on Moderately Concentraded Solutions:
Isotropic Solutions of Flexible and Rodlike Chains and Nematic Solutions of Rodlike
Chains. Vol. 114, pp. 233–290.
Bershtein, V. A. and Ryzhov, V. A.: Far Infrared Spectroscopy of Polymers. Vol. 114,
pp. 43–122.
Bhargava R., Wang S.-Q. and Koenig J. L: FTIR Microspectroscopy of Polymeric Systems.
Vol. 163, pp. 137–191.
Biesalski, M.: see Rühe, J.: Vol. 165, pp. 79–150.
Bigg, D. M.: Thermal Conductivity of Heterophase Polymer Compositions. Vol. 119,
pp. 1–30.
Binder, K.: Phase Transitions in Polymer Blends and Block Copolymer Melts: Some Recent
Developments. Vol. 112, pp. 115–134.
Binder, K.: Phase Transitions of Polymer Blends and Block Copolymer Melts in Thin Films.
Vol. 138, pp. 1–90.
Binder, K. see Baschnagel, J.: Vol. 152, pp. 41–156.
Binder, K., Müller, M., Virnau, P. and González MacDowell, L.: Polymer+Solvent Systems:
Phase Diagrams, Interface Free Energies, and Nucleation. Vol. 173, pp. 1–104.
Bird, R. B. see Curtiss, C. F.: Vol. 125, pp. 1–102.
Biswas, M. and Mukherjee, A.: Synthesis and Evaluation of Metal-Containing Polymers.
Vol. 115, pp. 89–124.
Biswas, M. and Sinha Ray, S.: Recent Progress in Synthesis and Evaluation of Polymer-
Montmorillonite Nanocomposites. Vol. 155, pp. 167–221.
Bogdal, D., Penczek, P., Pielichowski, J. and Prociak, A.: Microwave Assisted Synthesis,
Crosslinking, and Processing of Polymeric Materials. Vol. 163, pp. 193–263.
Bohrisch, J., Eisenbach, C.D., Jaeger, W., Mori H., Müller A.H.E., Rehahn, M., Schaller, C.,
Traser, S. and Wittmeyer, P.: New Polyelectrolyte Architectures. Vol. 165, pp. 1–41.
Bolze, J. see Dingenouts, N.: Vol. 144, pp. 1–48.
Bosshard, C.: see Gubler, U.: Vol. 158, pp. 123–190.
Boutevin, B. and Robin, J. J.: Synthesis and Properties of Fluorinated Diols. Vol. 102.
pp. 105–132.
Boutevin, B. see Amédouri, B.: Vol. 102, pp. 133–170.
Boutevin, B. see Améduri, B.: Vol. 127, pp. 87–142.
Bowman, C. N. see Anseth, K. S.: Vol. 122, pp. 177–218.
Author Index Volumes 101–175 349
Boyd, R. H.: Prediction of Polymer Crystal Structures and Properties. Vol. 116, pp. 1–26.
Briber, R. M. see Hedrick, J. L.: Vol. 141, pp. 1–44.
Bronnikov, S. V., Vettegren, V. I. and Frenkel, S. Y.: Kinetics of Deformation and Relaxation
in Highly Oriented Polymers. Vol. 125, pp. 103–146.
Brown, H. R. see Creton, C.: Vol. 156, pp. 53–135.
Bruza, K. J. see Kirchhoff, R. A.: Vol. 117, pp. 1–66.
Budkowski, A.: Interfacial Phenomena in Thin Polymer Films: Phase Coexistence and
Segregation. Vol. 148, pp. 1–112.
Burban, J. H. see Cussler, E. L.: Vol. 110, pp. 67–80.
Burchard,W.: Solution Properties of Branched Macromolecules. Vol. 143, pp. 113–194.
Butté, A. see Schork, F. J.: Vol. 175, pp. 129–255.
Calmon-Decriaud, A., Bellon-Maurel, V., Silvestre, F.: Standard Methods for Testing the
Aerobic Biodegradation of Polymeric Materials.Vol 135, pp. 207–226.
Cameron, N. R. and Sherrington, D. C.: High Internal Phase Emulsions (HIPEs)-Structure,
Properties and Use in Polymer Preparation. Vol. 126, pp. 163–214.
de la Campa, J. G. see de Abajo, J.: Vol. 140, pp. 23–60.
Candau, F. see Hunkeler, D.: Vol. 112, pp. 115–134.
Canelas, D. A. and DeSimone, J. M.: Polymerizations in Liquid and Supercritical Carbon
Dioxide. Vol. 133, pp. 103–140.
Canva, M. and Stegeman, G. I.: Quadratic Parametric Interactions in Organic Waveguides.
Vol. 158, pp. 87–121.
Capek, I.: Kinetics of the Free-Radical Emulsion Polymerization of Vinyl Chloride.Vol. 120,
pp. 135–206.
Capek, I.: Radical Polymerization of Polyoxyethylene Macromonomers in Disperse
Systems. Vol. 145, pp. 1–56.
Capek, I.: Radical Polymerization of Polyoxyethylene Macromonomers in Disperse
Systems. Vol. 146, pp. 1–56.
Capek, I. and Chern, C.-S.: Radical Polymerization in Direct Mini-Emulsion Systems.
Vol. 155, pp. 101–166.
Cappella, B. see Munz, M.: Vol. 164, pp. 87–210.
Carlesso, G. see Prokop, A.: Vol. 160, pp. 119–174.
Carlini, C. and Angiolini, L.: Polymers as Free Radical Photoinitiators. Vol. 123, pp. 127–
214.
Carter, K. R. see Hedrick, J. L.: Vol. 141, pp. 1–44.
Casas-Vazquez, J. see Jou, D.: Vol. 120, pp. 207–266.
Chandrasekhar, V.: Polymer Solid Electrolytes: Synthesis and Structure. Vol. 135, pp.
139–206.
Chang, J. Y. see Han, M. J.: Vol. 153, pp. 1–36.
Chang, T.: Recent Advances in Liquid Chromatography Analysis of Synthetic Polymers.
Vol. 163, pp. 1–60.
Charleux, B. and Faust R.: Synthesis of Branched Polymers by Cationic Polymerization.
Vol. 142, pp. 1–70.
Chen, P. see Jaffe, M.: Vol. 117, pp. 297–328.
Chern, C.-S. see Capek, I.: Vol. 155, pp. 101–166.
Chevolot, Y. see Mathieu, H. J.: Vol. 162, pp. 1–35.
Choe, E.-W. see Jaffe, M.: Vol. 117, pp. 297–328.
Chow, P. Y. and Gan, L. M.: Microemulsion Polymerizations and Reactions. Vol. 175,
pp. 257–298.
Chow, T. S.: Glassy State Relaxation and Deformation in Polymers. Vol. 103, pp. 149–190.
Chujo, Y. see Uemura, T.: Vol. 167, pp. 81–106.
350 Author Index Volumes 101–175
Dalton, L.: Nonlinear Optical Polymeric Materials: From Chromophore Design to Commer-
cial Applications. Vol. 158, pp. 1–86.
Dautzenberg, H. see Holm, C.: Vol. 166, pp. 113–171.
Davidson, J. M. see Prokop, A.: Vol. 160, pp. 119–174.
Desai, S. M. and Singh, R. P.: Surface Modification of Polyethylene. Vol. 169, pp. 231–293.
DeSimone, J. M. see Canelas D. A.: Vol. 133, pp. 103–140.
DeSimone, J. M. see Kennedy, K. A.: Vol. 175, pp. 329–346.
DiMari, S. see Prokop, A.: Vol. 136, pp. 1–52.
Dimonie, M. V. see Hunkeler, D.: Vol. 112, pp. 115–134.
Dingenouts, N., Bolze, J., Pötschke, D. and Ballauf, M.: Analysis of Polymer Latexes by Small-
Angle X-Ray Scattering. Vol. 144, pp. 1–48.
Dodd, L. R. and Theodorou, D. N.: Atomistic Monte Carlo Simulation and Continuum Mean
Field Theory of the Structure and Equation of State Properties of Alkane and Polymer
Melts. Vol. 116, pp. 249–282.
Doelker, E.: Cellulose Derivatives. Vol. 107, pp. 199–266.
Dolden, J. G.: Calculation of a Mesogenic Index with Emphasis Upon LC-Polyimides.
Vol. 141, pp. 189–245.
Domb, A. J., Amselem, S., Shah, J. and Maniar, M.: Polyanhydrides: Synthesis and Char-
acterization. Vol. 107, pp. 93–142.
Domb, A. J. see Kumar, M. N. V. R.: Vol. 160, pp. 45118.
Doruker, P. see Baschnagel, J.: Vol. 152, pp. 41–156.
Dubois, P. see Mecerreyes, D.: Vol. 147, pp. 1–60.
Dubrovskii, S. A. see Kazanskii, K. S.: Vol. 104, pp. 97–134.
Dunkin, I. R. see Steinke, J.: Vol. 123, pp. 81–126.
Dunson, D. L. see McGrath, J. E.: Vol. 140, pp. 61–106.
Dziezok, P. see Rühe, J.: Vol. 165, pp. 79–150.
Hadjichristidis, N., Pispas, S., Pitsikalis, M., Iatrou, H. and Vlahos, C.: Asymmetric Star Poly-
mers Synthesis and Properties. Vol. 142, pp. 71–128.
Hadjichristidis, N. see Xu, Z.: Vol. 120, pp. 1–50.
Hadjichristidis, N. see Pitsikalis, M.: Vol. 135, pp. 1–138.
Hahn, O. see Baschnagel, J.: Vol. 152, pp. 41–156.
Hakkarainen, M.: Aliphatic Polyesters: Abiotic and Biotic Degradation and Degradation
Products. Vol. 157, pp. 1–26.
Hakkarainen, M. and Albertsson, A.-C.: Environmental Degradation of Polyethylene. Vol.
169, pp. 177–199.
Hall, H. K. see Penelle, J.: Vol. 102, pp. 73–104.
Hamley, I.W.: Crystallization in Block Copolymers. Vol. 148, pp. 113–138.
Hammouda, B.: SANS from Homogeneous Polymer Mixtures: A Unified Overview.Vol. 106,
pp. 87–134.
Han, M. J. and Chang, J. Y.: Polynucleotide Analogues. Vol. 153, pp. 1–36.
Harada, A.: Design and Construction of Supramolecular Architectures Consisting of
Cyclodextrins and Polymers. Vol. 133, pp. 141–192.
Haralson, M. A. see Prokop, A.: Vol. 136, pp. 1–52.
Hassan, C. M. and Peppas, N. A.: Structure and Applications of Poly(vinyl alcohol) Hydro-
gels Produced by Conventional Crosslinking or by Freezing/Thawing Methods.Vol. 153,
pp. 37–65.
Hawker, C. J.: Dentritic and Hyperbranched Macromolecules Precisely Controlled Macro-
molecular Architectures. Vol. 147, pp. 113–160.
Hawker, C. J. see Hedrick, J. L.: Vol. 141, pp. 1–44.
He, G. S. see Lin, T.-C.: Vol. 161, pp. 157–193.
Hedrick, J. L., Carter, K. R., Labadie, J. W., Miller, R. D., Volksen, W., Hawker, C. J., Yoon, D.Y.,
Russell, T. P., McGrath, J. E. and Briber, R. M.: Nanoporous Polyimides. Vol. 141, pp. 1–44.
Author Index Volumes 101–175 353
Hedrick, J. L., Labadie, J. W., Volksen, W. and Hilborn, J. G.: Nanoscopically Engineered Poly-
imides. Vol. 147, pp. 61–112.
Hedrick, J. L. see Hergenrother, P. M.: Vol. 117, pp. 67–110.
Hedrick, J. L. see Kiefer, J.: Vol. 147, pp. 161–247.
Hedrick, J. L. see McGrath, J. E.: Vol. 140, pp. 61–106.
Heine, D. R., Grest, G. S. and Curro, J. G.: Structure of Polymer Melts and Blends: Comparison
of Integral Equation theory and Computer Sumulation. Vol. 173, pp. 209–249.
Heinrich, G. and Klüppel, M.: Recent Advances in the Theory of Filler Networking in
Elastomers. Vol. 160, pp. 1–44.
Heller, J.: Poly (Ortho Esters). Vol. 107, pp. 41–92.
Helm, C. A.: see Möhwald, H.: Vol. 165, pp. 151–175.
Hemielec, A. A. see Hunkeler, D.: Vol. 112, pp. 115–134.
Hergenrother, P. M., Connell, J. W., Labadie, J. W. and Hedrick, J. L.: Poly(arylene ether)s
Containing Heterocyclic Units. Vol. 117, pp. 67–110.
Hernández-Barajas, J. see Wandrey, C.: Vol. 145, pp. 123–182.
Hervet, H. see Léger, L.: Vol. 138, pp. 185–226.
Hilborn, J. G. see Hedrick, J. L.: Vol. 147, pp. 61–112.
Hilborn, J. G. see Kiefer, J.: Vol. 147, pp. 161–247.
Hiramatsu, N. see Matsushige, M.: Vol. 125, pp. 147–186.
Hirasa, O. see Suzuki, M.: Vol. 110, pp. 241–262.
Hirotsu, S.: Coexistence of Phases and the Nature of First-Order Transition in Poly-N-iso-
propylacrylamide Gels. Vol. 110, pp. 1–26.
Höcker, H. see Klee, D.: Vol. 149, pp. 1–57.
Holm, C., Hofmann, T., Joanny, J. F., Kremer, K., Netz, R. R., Reineker, P., Seidel, C., Vilgis, T. A.
and Winkler, R. G.: Polyelectrolyte Theory. Vol. 166, pp. 67–111.
Holm, C., Rehahn, M., Oppermann, W. and Ballauff, M.: Stiff-Chain Polyelectrolytes. Vol.
166, pp. 1–27.
Hornsby, P.: Rheology, Compounding and Processing of Filled Thermoplastics. Vol. 139,
pp. 155–216.
Houbenov, N. see Rühe, J.: Vol. 165, pp. 79–150.
Huber, K. see Volk, N.: Vol. 166, pp. 29–65.
Hugenberg, N. see Rühe, J.: Vol. 165, pp. 79–150.
Hui, C.-Y. see Creton, C.: Vol. 156, pp. 53–135.
Hult, A., Johansson, M. and Malmström, E.: Hyperbranched Polymers. Vol. 143, pp. 1–34.
Hünenberger, P. H.: Thermostat Algorithms for Molecular-Dynamics Simulations. Vol. 173,
pp. 105–147.
Hunkeler, D., Candau, F., Pichot, C., Hemielec, A. E., Xie, T. Y., Barton, J., Vaskova, V., Guillot, J.,
Dimonie, M. V. and Reichert, K. H.: Heterophase Polymerization: A Physical and Kinetic
Comparision and Categorization. Vol. 112, pp. 115–134.
Hunkeler, D. see Macko, T.: Vol. 163, pp. 61–136.
Hunkeler, D. see Prokop, A.: Vol. 136, pp. 1–52; 53–74.
Hunkeler, D. see Wandrey, C.: Vol. 145, pp. 123–182.
Kaetsu, I.: Radiation Synthesis of Polymeric Materials for Biomedical and Biochemical
Applications. Vol. 105, pp. 81–98.
Kaji, K. see Kanaya, T.: Vol. 154, pp. 87–141.
Kajzar, F., Lee, K.-S. and Jen, A. K.-Y.: Polymeric Materials and their Orientation Techniques
for Second-Order Nonlinear Optics. Vol. 161, pp. 1–85.
Kakimoto, M. see Gaw, K. O.: Vol. 140, pp. 107–136.
Kaminski, W. and Arndt, M.: Metallocenes for Polymer Catalysis. Vol. 127, pp. 143–187.
Kammer, H. W., Kressler, H. and Kummerloewe, C.: Phase Behavior of Polymer Blends –
Effects of Thermodynamics and Rheology. Vol. 106, pp. 31–86.
Kanaya, T. and Kaji, K.: Dynamcis in the Glassy State and Near the Glass Transition of
Amorphous Polymers as Studied by Neutron Scattering. Vol. 154, pp. 87–141.
Kandyrin, L. B. and Kuleznev, V. N.: The Dependence of Viscosity on the Composition of
Concentrated Dispersions and the Free Volume Concept of Disperse Systems. Vol. 103,
pp. 103–148.
Author Index Volumes 101–175 355
Macko, T. and Hunkeler, D.: Liquid Chromatography under Critical and Limiting Condi-
tions: A Survey of Experimental Systems for Synthetic Polymers. Vol. 163, pp. 61–136.
Majoros, I., Nagy, A. and Kennedy, J. P.: Conventional and Living Carbocationic Polymeriza-
tions United. I. A Comprehensive Model and New Diagnostic Method to Probe the
Mechanism of Homopolymerizations. Vol. 112, pp. 1–113.
Makhija, S. see Jaffe, M.: Vol. 117, pp. 297–328.
Malmström, E. see Hult, A.: Vol. 143, pp. 1–34.
Malkin, A. Y. and Kulichkhin, S. G.: Rheokinetics of Curing. Vol. 101, pp. 217–258.
Maniar, M. see Domb, A. J.: Vol. 107, pp. 93–142.
Manias, E. see Giannelis, E. P.: Vol. 138, pp. 107–148.
Martin, H. see Engelhardt, H.: Vol. 165, pp. 211–247.
Marty, J. D. and Mauzac, M.: Molecular Imprinting: State of the Art and Perspectives.
Vol. 172, pp. 1–35.
Mashima, K., Nakayama, Y. and Nakamura, A.: Recent Trends in Polymerization of a-Olefins
Catalyzed by Organometallic Complexes of Early Transition Metals. Vol. 133, pp. 1–52.
Mathew, D. see Reghunadhan Nair, C.P.: Vol. 155, pp. 1–99.
Mathieu, H. J., Chevolot, Y, Ruiz-Taylor, L. and Leónard, D.: Engineering and Character-
ization of Polymer Surfaces for Biomedical Applications. Vol. 162, pp. 1–35.
Matsumoto, A.: Free-Radical Crosslinking Polymerization and Copolymerization of
Multivinyl Compounds. Vol. 123, pp. 41–80.
Matsumoto, A. see Otsu, T.: Vol. 136, pp. 75–138.
Matsuoka, H. and Ise, N.: Small-Angle and Ultra-Small Angle Scattering Study of the
Ordered Structure in Polyelectrolyte Solutions and Colloidal Dispersions. Vol. 114,
pp. 187–232.
Matsushige, K., Hiramatsu, N. and Okabe, H.: Ultrasonic Spectroscopy for Polymeric
Materials. Vol. 125, pp. 147–186.
Mattice, W. L. see Rehahn, M.: Vol. 131/132, pp. 1–475.
Mattice, W. L. see Baschnagel, J.: Vol. 152, pp. 41–156.
Mattozzi, A. see Gedde, U. W.: Vol. 169, pp. 29–73.
Mauzac, M. see Marty, J. D.: Vol. 172, pp. 1–35.
358 Author Index Volumes 101–175
O’Brien, D. F., Armitage, B. A., Bennett, D. E. and Lamparski, H. G.: Polymerization and
Domain Formation in Lipid Assemblies. Vol. 126, pp. 53–84.
Ogasawara, M.: Application of Pulse Radiolysis to the Study of Polymers and Polymeri-
zations. Vol. 105, pp. 37–80.
Okabe, H. see Matsushige, K.: Vol. 125, pp. 147–186.
Okada, M.: Ring-Opening Polymerization of Bicyclic and Spiro Compounds. Reactivities
and Polymerization Mechanisms. Vol. 102, pp. 1–46.
Okano, T.: Molecular Design of Temperature-Responsive Polymers as Intelligent Materials.
Vol. 110, pp. 179–198.
Okay, O. see Funke, W.: Vol. 136, pp. 137–232.
Onuki, A.: Theory of Phase Transition in Polymer Gels. Vol. 109, pp. 63–120.
Oppermann, W. see Holm, C.: Vol. 166, pp. 1–27.
Oppermann, W. see Volk, N.: Vol. 166, pp. 29–65.
Osad’ko, I. S.: Selective Spectroscopy of Chromophore Doped Polymers and Glasses.
Vol. 114, pp. 123–186.
Osakada, K., Takeuchi, D.: Coordination Polymerization of Dienes, Allenes, and Methyl-
enecycloalkanes. Vol. 171, pp. 137–194.
Otaigbe, J. U., Barnes, M. D., Fukui, K., Sumpter, B. G. and Noid, D. W.: Generation, Charac-
terization, and Modeling of Polymer Micro- and Nano-Particles. Vol. 154, pp. 1–86.
Otsu, T. and Matsumoto, A.: Controlled Synthesis of Polymers Using the Iniferter Tech-
nique: Developments in Living Radical Polymerization. Vol. 136, pp. 75–138.
Quirk, R. P., Yoo, T., Lee, Y., M., Kim, J. and Lee, B.: Applications of 1,1-Diphenylethylene
Chemistry in Anionic Synthesis of Polymers with Controlled Structures. Vol. 153,
pp. 67–162.
Ramaraj, R. and Kaneko, M.: Metal Complex in Polymer Membrane as a Model for
Photosynthetic Oxygen Evolving Center. Vol. 123, pp. 215–242.
Rangarajan, B. see Scranton, A. B.: Vol. 122, pp. 1–54.
Author Index Volumes 101–175 361
Spange, S., Meyer, T., Voigt, I., Eschner, M., Estel, K., Pleul, D. and Simon, F.: Poly(Vinyl-
formamide-co-Vinylamine)/Inorganic Oxid Hybrid Materials. Vol. 165, pp. 43–78.
Stamm, M. see Möhwald, H.: Vol. 165, pp. 151–175.
Stamm, M. see Rühe, J.: Vol. 165, pp. 79–150.
Starodybtzev, S. see Khokhlov, A.: Vol. 109, pp. 121–172.
Stegeman, G. I. see Canva, M.: Vol. 158, pp. 87–121.
Steinke, J., Sherrington, D. C. and Dunkin, I. R.: Imprinting of Synthetic Polymers Using
Molecular Templates. Vol. 123, pp. 81–126.
Stenberg, B. see Jacobson, K.: Vol. 169, pp. 151–176.
Stenzenberger, H. D.: Addition Polyimides. Vol. 117, pp. 165–220.
Stephan, T. see Rühe, J.: Vol. 165, pp. 79–150.
Stevenson,W. T. K. see Sefton, M. V.: Vol. 107, pp. 143–198.
Stridsberg, K. M., Ryner, M. and Albertsson, A.-C.: Controlled Ring-Opening Polymer-
ization: Polymers with Designed Macromoleculars Architecture. Vol. 157, pp. 27–51.
Sturm, H. see Munz, M.: Vol. 164, pp. 87–210.
Suematsu, K.: Recent Progress of Gel Theory: Ring, Excluded Volume, and Dimension.
Vol. 156, pp. 136–214.
Sugimoto, H. and Inoue, S.: Polymerization by Metalloporphyrin and Related Complexes.
Vol. 146, pp. 39–120.
Suginome, M. and Ito, Y.: Transition Metal-Mediated Polymerization of Isocyanides. Vol.
171, pp. 77–136.
Sumpter, B. G., Noid, D. W., Liang, G. L. and Wunderlich, B.: Atomistic Dynamics of
Macromolecular Crystals. Vol. 116, pp. 27–72.
Sumpter, B. G. see Otaigbe, J. U.: Vol. 154, pp. 1–86.
Sun, H.-B. and Kawata, S.: Two-Photon Photopolymerization and 3D Lithographic Micro-
fabrication. Vol. 170, pp. 169–273.
Suter, U. W. see Gusev, A. A.: Vol. 116, pp. 207–248.
Suter, U. W. see Leontidis, E.: Vol. 116, pp. 283–318.
Suter, U. W. see Rehahn, M.: Vol. 131/132, pp. 1–475.
Suter, U. W. see Baschnagel, J.: Vol. 152, p. 41–156.
Suzuki, A.: Phase Transition in Gels of Sub-Millimeter Size Induced by Interaction with
Stimuli. Vol. 110, pp. 199–240.
Suzuki, A. and Hirasa, O.: An Approach to Artifical Muscle by Polymer Gels due to Micro-
Phase Separation. Vol. 110, pp. 241–262.
Suzuki, K. see Nomura, M.: Vol. 175, pp. 1–128.
Swiatkiewicz, J. see Lin, T.-C.: Vol. 161, pp. 157–193.
Tagawa, S.: Radiation Effects on Ion Beams on Polymers. Vol. 105, pp. 99–116.
Takata, T., Kihara, N. and Furusho, Y.: Polyrotaxanes and Polycatenanes: Recent Advances
in Syntheses and Applications of Polymers Comprising of Interlocked Structures.
Vol. 171, pp. 1–75.
Takeuchi, D. see Osakada, K.: Vol. 171, pp. 137–194.
Tan, K. L. see Kang, E. T.: Vol. 106, pp. 135–190.
Tanaka, H. and Shibayama, M.: Phase Transition and Related Phenomena of Polymer Gels.
Vol. 109, pp. 1–62.
Tanaka, T. see Penelle, J.: Vol. 102, pp. 73–104.
Tauer, K. see Guyot, A.: Vol. 111, pp. 43–66.
Teramoto, A. see Sato, T.: Vol. 126, pp. 85–162.
Terent’eva, J. P. and Fridman, M. L.: Compositions Based on Aminoresins. Vol. 101,
pp. 29–64.
Theodorou, D. N. see Dodd, L. R.: Vol. 116, pp. 249–282.
364 Author Index Volumes 101–175
Thomson, R. C., Wake, M. C., Yaszemski, M. J. and Mikos, A. G.: Biodegradable Polymer
Scaffolds to Regenerate Organs. Vol. 122, pp. 245–274.
Thünemann, A. F., Müller, M., Dautzenberg, H., Joanny, J.-F. and Löwen, H.: Polyelectrolyte
complexes. Vol. 166, pp. 113–171.
Tieke, B. see v. Klitzing, R.: Vol. 165, pp. 177–210.
Tobita, H. see Nomura, M.: Vol. 175, pp. 1–128.
Tokita, M.: Friction Between Polymer Networks of Gels and Solvent. Vol. 110, pp. 27–48.
Traser, S. see Bohrisch, J.: Vol. 165, pp. 1–41.
Tries, V. see Baschnagel, J.: Vol. 152, p. 41–156.
Tsuruta, T.: Contemporary Topics in Polymeric Materials for Biomedical Applications.
Vol. 126, pp. 1–52.
Uemura, T., Naka, K. and Chujo, Y.: Functional Macromolecules with Electron-Donating
Dithiafulvene Unit. Vol. 167, pp. 81–106.
Usov, D. see Rühe, J.: Vol. 165, pp. 79–150.
Uyama, H. see Kobayashi, S.: Vol. 121, pp. 1–30.
Uyama, Y: Surface Modification of Polymers by Grafting. Vol. 137, pp. 1–40.
Yagci, Y. and Endo, T.: N-Benzyl and N-Alkoxy Pyridium Salts as Thermal and Photo-
chemical Initiators for Cationic Polymerization. Vol. 127, pp. 59–86.
Yannas, I. V.: Tissue Regeneration Templates Based on Collagen-Glycosaminoglycan Co-
polymers. Vol. 122, pp. 219–244.
Yang, J. S. see Jo, W. H.: Vol. 156, pp. 1–52.
Yamaoka, H.: Polymer Materials for Fusion Reactors. Vol. 105, pp. 117–144.
Yasuda, H. and Ihara, E.: Rare Earth Metal-Initiated Living Polymerizations of Polar and
Nonpolar Monomers. Vol. 133, pp. 53–102.
Yaszemski, M. J. see Thomson, R. C.: Vol. 122, pp. 245–274.
Yoo, T. see Quirk, R. P.: Vol. 153, pp. 67–162.
Yoon, D. Y. see Hedrick, J. L.: Vol. 141, pp. 1–44.
Yoshida, H. and Ichikawa, T.: Electron Spin Studies of Free Radicals in Irradiated Polymers.
Vol. 105, pp. 3–36.