Вы находитесь на странице: 1из 44

Emulsion polymerization mechanisms and kinetics

C.S. Chern
*
Department of Chemical Engineering, National Taiwan University of Science and Technology, 43 Keelung Road,
Section 4, Taipei 106, Taiwan, ROC
Received 15 August 2005; received in revised form 24 November 2005; accepted 15 February 2006
Available online 5 April 2006
Abstract
Emulsion polymerization involves the propagation reaction of free radicals with monomer molecules in a very large number of
discrete polymer particles (10
16
10
18
dm
K3
) dispersed in the continuous aqueous phase. The nucleation and growth of latex
particles control the colloidal and physical properties of latex products. This review article focused on the polymerization
mechanisms and kinetics involved in such a heterogeneous polymerization system over the preceding 10-year period. First, an
overview of the general features of emulsion polymerization was given, followed by the discussion of several techniques useful for
studying the related polymerization mechanisms and kinetics. Emulsion polymerizations using different stabilizers were studied
extensively in the last few years and representative publications were reviewed. The performance properties of some specialty
polymerizable, degradable or polymeric surfactants and surface-active initiators were also evaluated in emulsion polymerization.
At present, the particle nucleation process is still not well understood and deserves more research efforts. This article continued to
discuss the origin of non-uniform latex particles from both the thermodynamic and kinetic points of view. This was followed by the
discussion of various reaction parameters that had signicant effects on the development of particle morphology. Recent studies on
the polymerization in non-uniform polymer particles were then reviewed. Finally, the polymerization mechanisms, kinetics and
colloidal stability involved in the versatile semibatch emulsion polymerization were reviewed extensively.
q 2006 Elsevier Ltd. All rights reserved.
Keywords: Emulsion polymerization; Latex particles; Polymerization mechanisms; Particle nucleation; Particle growth; Kinetics
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
2. Emulsion polymerization mechanisms and kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445
2.1. General features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445
2.2. Some representative techniques used to study particle nucleation and growth mechanisms . . . . . . . . . . . . 450
2.3. Surfactant-free polymerization systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452
2.4. Ionic surfactant stabilized polymerization systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
2.5. Nonionic surfactant and mixed anionic and nonionic surfactants stabilized polymerization systems . . . . . 459
2.6. Polymerizable or degradable surfactant and surface-active initiator stabilized polymerization systems . . . 463
2.7. Polymeric surfactant and protective colloid stabilized polymerization systems . . . . . . . . . . . . . . . . . . . . . 464
2.8. Mathematical modeling studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468
3. Polymerization in non-uniform latex particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 470
Prog. Polym. Sci. 31 (2006) 443486
www.elsevier.com/locate/ppolysci
0079-6700/$ - see front matter q 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.progpolymsci.2006.02.001
* Tel.: C886 2 27376649; fax: C886 2 27376644.
E-mail address: cschern@mail.ntust.edu.tw
3.1. Origin of non-uniform latex particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 470
3.2. Morphology development in latex particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471
4. Semibatch emulsion polymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
4.1. Polymerization mechanisms and kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
4.2. Mathematical modeling studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 477
5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 480
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 480
1. Introduction
Emulsion polymerization is a unique chemical
process widely used to produce waterborne resins
with various colloidal and physicochemical properties.
This heterogeneous free radical polymerization process
involves emulsication of the relatively hydrophobic
monomer in water by an oil-in-water emulsier,
followed by the initiation reaction with either a water-
insoluble initiator (e.g. sodium persulfate (NaPS)) or an
oil-soluble initiator (e.g. 2-2
/
-azobisisobutyronitrile
(AIBN)) [16]. Typical monomers used in emulsion
polymerization include butadiene, styrene, acryloni-
trile, acrylate ester and methacrylate ester monomers,
vinyl acetate, and vinyl chloride. An extremely large
oilwater interfacial area is generated as the particle
nuclei form and grow in size with the progress of the
polymerization. Thus, an effective stabilizer such as
ionic and non-ionic surfactants and protective colloid
(e.g. hydroxyethyl cellulose and polyvinyl alcohol),
which can be physically adsorbed or chemically
incorporated onto the particle surface, is required to
prevent the interactive latex particles from coagulation.
Under the circumstances, satisfactory colloidal stability
can be achieved via the electrostatic stabilization
mechanism [7], the steric stabilization mechanism
[8,9] or both. The environmentally friendly latex
products comprise a large population of polymer
particles (ca. 10
1
10
3
nm in diameter) dispersed in
the continuous aqueous phase. These emulsion poly-
mers nd a wide range of applications such as synthetic
rubbers, thermoplastics, coatings, adhesives, binders,
rheological modiers, plastic pigments, standards for
the calibration of instruments, immunodiagnosis tests,
polymeric supports for the purication of proteins and
drug delivery system, etc. To gain a fundamental
understanding of polymerization mechanisms and
kinetics is a must in designing quality products that
fulll customers requirements.
Emulsion polymerization is a rather complex
process because nucleation, growth and stabilization
of polymer particles are controlled by the free radical
polymerization mechanisms in combination with
various colloidal phenomena. Perhaps, the most
striking feature of emulsion polymerization is the
segregation of free radicals among the discrete
monomer-swollen polymer particles. This will greatly
reduce the probability of bimolecular termination of
free radicals and, thereby, result in a faster polymer-
ization rate and polymer with a higher molecular
weight. This advantageous characteristic of emulsion
polymerization cannot be achieved simultaneously in
bulk or solution polymerization. Although the nuclea-
tion period is quite short, generation of particle nuclei
during the early stage of the polymerization plays a
crucial role in determining the nal latex particle size
and particle size distribution and it has also a
signicant inuence on the quality of latex products.
How to effectively control the particle nucleation
process represents a very challenging task to those
who are involved in this fascinating research area.
Transport of monomer, free radicals and surfactant to
the growing particles and partition of these reagents
among the continuous aqueous phase, emulsied
monomer droplets (monomer reservoir), monomer-
swollen polymer particles (primary reaction loci) and
oilwater interface are the key factors that govern the
particle growth stage. The colloidal properties of latex
products are of great importance from both academic
and industrial points of view. Some representative
properties include the particle size and particle size
distribution, particle surface charge density (or zeta
potential), particle surface area covered by one
stabilizer molecule, conformation of the hydrophilic
polymer physically adsorbed or chemically coupled
onto the particle surface, type and concentration of
functional groups on the particle surface, particle
morphology, optical and rheological properties and
colloidal stability.
Batch emulsion polymerization is commonly used in
the laboratory to study reaction mechanisms, develop
new latex products and obtain kinetic data for process
development and reactor scale-up. Most of the
commercial latex products are manufactured by
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 444
semibatch or continuous reaction systems due to the
very exothermic nature of free radical polymerization
and rather limited heat transfer capacity in large-scale
reactors. One major difference among the above
polymerization processes is the residence time distri-
bution of the growing particles within the reactor. The
broadness of the residence time distribution in
decreasing order is continuousOsemibatchObatch.
As a consequence, the broadness of the resultant
particle size distribution in decreasing order is
continuousOsemibatchObatch. The rate of polymer-
ization generally follows the following trend: batchO
semibatchOcontinuous. Furthermore, the versatile
semibatch and continuous emulsion polymerization
processes offer the operational exibility to produce
latex products with controlled polymer composition
and particle morphology. This may have an important
inuence on the application properties of latex
products. This feature will be illustrated by the
semibatch emulsion polymerization below. For the
continuous emulsion polymerization, more efforts
should be devoted to the development and under-
standing of different reactor designs and operating
procedures, especially related to characteristics of latex
products and start-up and product changeover strat-
egies, which will not be covered in this work. Only a
small number of journal papers dealing with these two
industrial processes are available in the open literature.
More research efforts are required to further advance
the semibatch and continuous emulsion polymerization
technology. For those who are interested in the previous
studies on semibatch and continuous emulsion
polymerizations, please refer to those cited in the
review articles [1015].
Miniemulsion, microemulsion and conventional
emulsion polymerizations show quite different particle
nucleation and growth mechanisms and kinetics. The
former two research areas have received increasing
interest recently. However, miniemulsion and micro-
emulsion polymerization processes are beyond the
scope of this work and they will not be covered here.
Recently, miniemulsion polymerization has been
reviewed by El-Aasser et al. [16,17], Capek and
Chern [18], Antonietti and Landfester [19] and Asua
[20]. The review articles [2125] focused on micro-
emulsion polymerization. Representative review or
journal articles concerning emulsion polymerization
can be found in references [1015,2638]. The
objective of this work was to review the major aspects
of emulsion polymerization mechanisms and kinetics
emphasizing progress over the preceding 10-year
period.
2. Emulsion polymerization mechanisms
and kinetics
2.1. General features
A typical emulsion polymerization formulation
comprises monomer, water, surfactant and a water-
soluble initiator. The reaction system is characterized
by the emulsied monomer droplets (ca. 110 mm in
diameter, 10
12
10
14
dm
K3
) dispersed in the continuous
aqueous phase with the aid of an oil-in-water surfactant
at the very beginning of polymerization. Monomer-
swollen micelles (ca. 510 nm in diameter, 10
19

10
21
dm
K3
in number) may also exist in the reaction
system provided that the concentration of surfactant in
the aqueous phase is above its critical micelle
concentration (CMC). Only a small fraction of the
relatively hydrophobic monomer is present in the
micelles (if present) or dissolved in the aqueous
phase. Most of the monomer molecules dwell in the
giant monomer reservoirs (i.e. monomer droplets). The
polymerization is initiated by the addition of initiator.
According to the micelle nucleation model, proposed
by Harkins [3941] and Smith and Ewart [4244] and
modied by Gardon [45,46], submicron latex particles
(ca. 0.051 mm in diameter, 10
16
10
18
dm
K3
in
number) are generated via the capture of free radicals
by micelles, which exhibit an extremely large oilwater
interfacial area. In general, monomer droplets are not
effective in competing with micelles in capturing free
radicals generated in the aqueous phase due to their
relatively small surface area. However, monomer
droplets may become the predominant particle nuclea-
tion loci if the droplet size is reduced to the submicron
range. This technique is termed the miniemulsion
polymerization [1620] and innovative miniemulsion
polymerization technology continues to spurt in recent
years [19].
Waterborne free radicals rst polymerize with
monomer molecules dissolved in the continuous
aqueous phase. This would result in the increased
hydrophobicity of oligomeric radicals. When a critical
chain length is achieved, these oligomeric radicals
become so hydrophobic that they show a strong
tendency to enter the monomer-swollen micelles and
then continue to propagate by reacting with those
monomer molecules therein. As a consequence,
monomer-swollen micelles are successfully trans-
formed into particle nuclei. These embryo particles
continue to grow by acquiring the reactant species from
monomer droplets and monomer-swollen micelles. In
order to maintain adequate colloidal stability of the
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 445
growing particle nuclei, micelles that do not contribute
to particle nucleation disband to supply the increasing
demand for surfactant. In addition, the surfactant
molecules adsorbed on monomer droplets may also
desorbs out of the droplet surface, diffuse across the
continuous aqueous phase and then adsorb on the
expanding particle surface. The particle nucleation
stage (Interval I) ends immediately after the exhaustion
of micelles. About one of every 10
2
10
3
micelles can
be successfully converted into latex particles.
The SmithEwart theory predicts that the number of
latex particles nucleated per unit volume of water (N
p
)
is proportional to the surfactant concentration and
initiator concentration to the 0.6 and 0.4 powers,
respectively. This relationship shows that the most
important parameter that controls the particle nuclea-
tion process is the surfactant concentration. Although
the particle nucleation period is relatively short (up to
about 1020% monomer conversion), it controls the
particle size and particle size distribution of latex
products. The application properties of emulsion
polymers such as rheology and lm formation are
strongly dependent on the particle size and particle size
distribution. By rule of thumb, latex products with a
large particle size can be produced by using a relatively
low surfactant concentration in the particle nucleation
stage. In addition, a narrow particle size distribution
will be achieved. This is simply because the shorter the
particle nucleation period (i.e. the lower the surfactant
concentration), the narrower the resultant particle size
distribution. It should be noted that occulation of
particle nuclei arising from inadequate stabilization of
the colloidal system may also take place during
polymerization. This will make the task of controlling
the particle size and particle size distribution of latex
products more difcult. In addition, it is much easier to
prepare latex products with high total solid contents
when a small population of particles with a large
particle size is produced at the end of polymerization.
This is closely related to the relatively mean free path
length (H/r) between two interactive particles as a
function of the total solid content [47]. H and r
represent the average interparticle distance and particle
radius, respectively. The value of H/r decreases with
increasing total solid content. Thus, the higher the total
solid content, the more crowded the colloidal system
(i.e. the greater the interaction between two approach-
ing particles). Furthermore, larger particles exhibit
larger value of H for the stationary particle packing.
Thus, at constant total solid content, the colloid system
comprising larger particles should be less crowded. In
this manner, high solid latex products with satisfactory
rheological properties are achieved. This is a very
important consideration in many industrial
applications.
After the particle nucleation process is completed,
the number of latex particles (i.e. reaction loci) per unit
volume of water remains relatively constant toward the
end of polymerization. The propagation reaction of free
radicals with monomer molecules takes place primarily
in monomer-swollen particles. Monomer droplets only
serve as reservoirs to supply the growing particles with
monomer and surfactant species. The majority of
monomer is consumed in this particle growth stage
ranging from ca. 1020 to 60% monomer conversion.
The particle growth stage (Interval II) ends when
monomer droplets disappear in the polymerization
system. SmithEwart case 2 kinetics has been widely
used to calculate the rate of polymerization (R
p
)
R
p
Zk
p
[M]
p
(nN
p
=N
A
) (1)
where k
p
is the propagation rate constant, [M]
p
the
concentration of monomer in the particles, n the
average number of free radicals per particle, and N
A
the Avogadro number. This kinetic model was
developed based on the following assumptions:
(1) Nucleation and coagulation of particles do not
occur and the number of particles per unit volume
of water remains constant during polymerization.
(2) The particle size distribution is relatively mono-
disperse.
(3) Desorption of free radicals out of the particles does
not take place.
(4) Bimolecular termination of the polymeric radical
inside the particle upon the entry of an oligomeric
radical from the aqueous phase is instantaneous.
These assumptions then lead to a scenario that, at
any moment, monomer-swollen particles contain either
only one free radical (active) or zero free radical (idle).
Under the circumstances, a value of n equal to 0.5 is
achieved for the polymerization systems that follow
SmithEwart case 2 kinetics. In addition, the concen-
tration of monomer in the particles does not vary to any
extent with the progress of polymerization in the
presence of monomer droplets. As a result, a steady
polymerization rate is attained during Interval II.
Furthermore, the polymerization kinetics is strictly
controlled by the population of particles available for
consuming monomer. SmithEwart case 2 kinetics has
been successfully applied to emulsion polymerizations
of relatively water-insoluble monomers such as styrene
and butadiene.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 446
Emulsion polymerization proceeds from Interval II
to III when all the monomer droplets disappear. In
Interval III, latex particles become monomer-starved
and the concentration of monomer in the reaction loci
continues to decrease toward the end of polymerization.
Thus, the steady polymerization rate observed in
Interval II cannot be maintained any more and the
polymerization rate decreases during Interval III. On
the other hand, the polymerization rate may increase
rapidly with increasing monomer conversion. This is
attributed to the greatly reduced bimolecular termin-
ation reaction between two polymeric radicals within
the very viscous particle provided that polymerization
is carried out at a temperature below the glass transition
temperature of the monomer-starved polymer solution.
This phenomenon is termed the gel effect [48,49]. To
minimize residual monomer in latex products is
essential for the successful product development
because of the potential hazard to end-users. Schematic
representations of the micelle nucleation model and the
rate of polymerization as a function of monomer
conversion are shown in Figs. 1 and 2, respectively.
It should be noted that some mechanisms other than
micelle nucleation must be responsible for the particle
formation process when the surfactant concentration is
below the CMC. Priest [50], Roe [51] and Fitch and
Tsai [5254] proposed the homogeneous nucleation
mechanism for the formation of particle nuclei in the
continuous aqueous phase, as shown schematically in
Fig. 3. First, waterborne initiator radicals are generated
by the thermal decomposition of initiator and they can
grow in size via the propagation reaction with those
monomer molecules dissolved in the aqueous phase.
The oligomeric radicals then become water-insoluble
when a critical chain length is reached. The hydro-
phobic oligomeric radical may thus coil up and form a
particle nucleus in the aqueous phase. This is followed
by formation of stable primary particles via the limited
occulation of the relatively unstable particle nuclei
and adsorption of surfactant molecules on their particle
surfaces. The surfactant species required to stabilize
these primary particles come from those dissolved in
the aqueous phase and those adsorbed on the monomer
droplet surfaces. The above ideas were incorporated
into the following kinetic model developed by Fitch
and Tsai [5254]
dN
p
=dt Zbr
i
KR
c
KR
f
(2)
where t is the reaction time, r
i
the rate of generation of
free radicals in the aqueous phase, b a parameter that
takes into account the aggregation of oligomeric
radicals, R
c
the rate of capture of free radicals by the
particles, and R
f
the rate of occulation of the particles.
Hansen and Ugelstad [55] proposed that a primary
particle is nucleated when the chain length of an
oligomeric radical reach a critical value (n*) and then
derived the following equations for calculating the
number of primary particles per unit volume of water Fig. 1. A schematic representation of the micelle nucleation model.
Fig. 2. Typical rate of polymerization as a function of the monomer
conversion. The three distinct intervals of the polymerization process
are also indicated in the plot.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 447
originating from homogeneous nucleation (N
1
)
N
1
(t) Z1=k
1
[k
1
r
i
n
+
t(k
2
C1)
n
+
]
1=n
+
K(k
2
C1) (3)
k
1
Zk
c
=(k
p
[M]
w
) (4)
k
2
Z(k
tw
r
i
)
1=2
=(k
p
[M]
w
) (5)
where k
c
is the average rate constant for the capture of
free radicals by the particles, k
tw
the termination rate
constant in water, and [M]
w
the concentration of
monomer in water.
To gain a better understanding of the emulsion
polymerization kinetics is essential in the process
development and reactor design. Among the kinetic
parameters, the most difcult ones to predict are the
number of latex particles per unit volume of water (N
p
)
and average number of free radicals per particle (n) in
Eq. (1). It should be noted that derivation of Eq. (1) was
based on the assumption that the propagation reaction
between polymeric radicals and monomer molecules
takes place only in monomer-swollen particles. Polymer-
ization in the continuous aqueous phase is insignicant.
As discussed above, particle nucleation controls the
particle size and particle size distribution of latex
products. It is the transport of free radicals between the
continuous aqueous phase and particles that determines
the average number of free radicals per particle during
polymerization (Fig. 4). Free radicals generated in the
aqueous phase can be absorbed by the particles during
Interval II. For example, anidle particle becomes active in
the propagation reaction immediately after the capture of
one free radical fromthe aqueous phase. In addition to the
major propagation reaction, one free radical may undergo
the bimolecular termination reaction with another free
radical if there are twoor more free radicals inthe particle.
Desorption of free radicals out of the particles may also
take place during polymerization. This process begins
with the chain transfer reaction of a polymeric radical to
monomer or chain transfer agent (if present). As a result, a
rather mobile monomeric radical or chain transfer agent
radical is produced. This is followed by the molecular
diffusion of this free radical from the interior of the
Fig. 4. A schematic representation of the transport of free radicals
between the continuous aqueous phase and the particle phase.
Fig. 3. A schematic representation of the homogeneous nucleation
mechanism.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 448
particle, across the particlewater interface, and then into
the aqueous phase. The desorbed monomeric radical may
also have the chance to be absorbed again by another
particle and reinitiate the propagation reaction therein.
Furthermore, the bimolecular termination reaction
between two approaching free radicals in the aqueous
phase may also happen in some emulsion polymerization
systems.
Smith and Ewart [42] and Haward [56] derived very
similar expressions for the population balance of the
number of latex particles containing i free radicals per
unit volume of water (N
i
, iZ0,1,2,.)
dN
i
=dt Zr
a
=N
p
(N
iK1
KN
i
) Ck
0
a
p
=v
p
[(i C1)N
iC1
KiN
i
]
Ck
tp
=v
p
[(i C2)(i C1)N
iC2
Ki(iK1)N
i
]
(6)
where r
a
is the rate of absorption of free radicals by the
particles, k
0
the rate coefcient for desorption of free
radicals out of the particles, a
p
and v
p
the surface area
and volume of a single particle, respectively, and k
tp
the
bimolecular termination reaction rate constant in the
particles. At pseudo-steady state (i.e. dN
i
/dtZ0), three
limiting cases were obtained from this system of rst-
order ordinary differential equations.
Case 1: r
a
/N
p
/k
0
a
p
/v
p
(fast desorption)
n Zr
a
v
p
=(k
0
a
p
N
p
)/0:5 (7)
Case 2: k
0
a
p
/v
p
/r
a
/N
p
/k
tp
/v
p
(no desorption and
fast termination)
n ZN
1
=(N
0
CN
1
) Z0:5 (8)
Case 3: r
a
/N
p
[k
tp
/v
p
(fast absorption and slow
termination)
n Z[r
a
v
p
=(2k
tp
N
p
)]
1=2
[1 (9)
A schematic representation of SmithEwart kinetics
cases 13 is shown in Fig. 5. The pseudo-steady-state
assumption is quite reasonable because the concen-
tration of free radicals in the colloidal system is very
low and the reactivity of free radicals is extremely high.
Stockmayer [57] derived the following equations to
calculate the steady value of n when desorption of free
radicals out of the latex particles is insignicant
(i.e. k
0
Z0)
n Za=4[I
0
(a)=I
1
(a)] (10)
a Z(8a)
1=2
(11)
a Zr
a
v
p
=(k
tp
N
p
) (12)
where I
0
(a) and I
1
(a) are the Bessel functions of the rst
kind of order 0 and 1, respectively. OToole [58]
extended this approach to take into account desorption
of free radicals out of the particles and obtained the
Fig. 5. A schematic representation of the SmithEwart kinetics cases
13.
Fig. 6. A schematic representation of the log(n) versus log(a) prole.
The three limiting cases of the SmithEwart kinetic model are also
indicated in this plot.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 449
following equations:
n Za=4[I
m
(a)=I
mK1
(a)] (13)
m Zk
0
a
p
=k
tp
(14)
A schematic representation of the log(n) versus
log(a) prole is shown in Fig. 6. Ugelstad et al. [59,60]
then incorporated the events of the bimolecular
termination reaction in the aqueous phase and
reabsorption of the desorbed free radicals by the
particles into the OToole model and the mass balance
equation for free radicals in the aqueous phase can be
expressed as
a Za
/
CmnKYa
2
(15)
a
/
Zr
i
v
p
=(k
tp
N
p
) (16)
Y Z2N
A
k
tw
k
tp
=(k
2
c
v
p
N
p
) (17)
where r
i
is the rate of generation of initiator radicals in
the aqueous phase, k
tw
the bimolecular termination
reaction rate constant in the aqueous phase, and k
c
the
rate coefcient for the capture of free radicals by the
particles. The authors then solved the above simul-
taneous equations for the average number of free
radicals per particle and plotted the values of n versus
a
/
at various values of Y and m.
The basic framework of emulsion polymerization
mechanisms and kinetics was primarily built on the
aforementioned pioneering studies and many other
excellent contributions appeared thereafter.
2.2. Some representative techniques used to study
particle nucleation and growth mechanisms
As discussed above, particle nucleation and growth
processes have signicant effects on the particle size
and particle size distribution of latex products and the
colloidal stability during polymerization and upon
transportation and storage. Thus, to gain a better
understanding of particle nucleation and growth
mechanisms guarantees successful product develop-
ment in the laboratory and scale up in the plant.
Moreover, it is important to keep a tight rein on the
amount of key ingredients (e.g. anionic and non-ionic
surfactants, protective colloids, etc.) used to control the
population of particles in the plant production. To
determine which particle nucleation mechanism
(micelle nucleation or homogeneous nucleation) pre-
dominates in a particular emulsion polymerization
system is not straightforward. For example, Roe [44]
showed that even the fact that experimental data follow
the relationship that the number of particles nucleated
per unit volume of water is proportional to the
surfactant and initiator concentrations to the 0.6 and
0.4 powers, respectively, does not necessarily conrm
the SmithEwart theory. More independent experimen-
tal data are required to distinguish micelle nucleation
from homogeneous nucleation for emulsion polymer-
izations with the surfactant concentration above the
CMC.
In the surfactant-free emulsion polymerization of
styrene, Feeney et al. [61] used small-angle neutron
scattering (SANS) in combination with an aqueous
polyacrylamide gel containing initiator to measure the
particle size during the early stage of polymerization.
The presence of particle nuclei (i.e. precursor particles)
with an average radius of 6 nm was observed. It was
demonstrated that SANS was a very effective technique
for the investigation of particle nucleation mechanisms
involved in emulsion polymerization. However, the
colloidal system involved is much more complicated
than conventional emulsion polymerization and the
inuence of the aqueous polyacrylamide gel on the
polymerization mechanisms should be evaluated
cautiously.
Wang and Poehlein [62,63] and Thomson et al. [64]
isolated and characterized the water-soluble oligomers
generated during emulsion polymerizations. Analytical
techniques such as FT-IR spectroscopy, mass spec-
troscopy and
13
C NMR were used to characterize these
oligomers. It would be interesting to further study the
effects of various reaction variables such as the types of
monomer and initiator on the physicochemical proper-
ties of ologomeric radicals produced in the continuous
aqueous phase. These studies may provide valuable
information on the nature of oligomeric radicals
produced early in the emulsion polymerization and
promote the understanding of particle nucleation
mechanisms.
Kuhn and Tauer [65] developed a technique of
on-line monitoring the optical transmission and
conductivity of the colloidal system to investigate the
particle nucleation mechanism involved in the surfac-
tant-free emulsion polymerization of styrene. It was
concluded that the rate of initiation in the continuous
aqueous phase played an important role in the particle
nucleation stage. Particle nucleation occurred via the
cluster formation of waterborne oligomers in the
surfactant-free emulsion polymerization of styrene.
Tauer and Deckwer [66] used the MALDI-TOF-MAS
technique to study the end-groups of polymer chains
obtained from the surfactant-free emulsion
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 450
polymerization of styrene initiated by potassium
persulfate (KPS). Surprisingly, a variety of end-groups
were identied in addition to the sulfate group
originating from the persulfate initiator. It was then
concluded that polymer chains started with oligomeric
radicals generated by side reactions in the aqueous
phase played an important role in the particle
nucleation process (e.g. homogeneous nucleation).
Moreover, they pointed out that the surface activity of
oligomeric radicals absorbed by the latex particles was
denitely not a prerequisite. Kozempel et al. [67] used
the on-line multi-angle laser light scattering technique
to study the surfactant-free emulsion polymerization of
styrene. It was proposed that the polymerization
mechanisms were characterized by three intervals.
Monomer droplets (ca. 200 nm in diameter) were
produced during Interval A. This was followed by the
formation of particle nuclei (Interval B). Immediately
after particle nucleation, latex particles absorbed
monomer from monomer droplets, thereby leading to
the depletion of monomer droplets and a reduction in
the average size of the scattering objects. Beyond
Interval B, the average size of the scattering objects
increased again as a result of the predominant growth of
latex particles (Interval C). This is presumably due to
the limited particle occulation.
Chern and Lin [68] used an extremely water-
insoluble dye as the probe for determining the loci of
particle nucleation in the emulsion polymerization of
styrene. Measurements of the weight percentage of dye
incorporated into the nal latex particles provided
valuable information on particle nucleation mechan-
isms. This work illustrated that micelle nucleation and
homogeneous nucleation competed with each other
when the surfactant concentration was higher than its
CMC. In contrast, most of the particles were generated
via homogeneous nucleation in the emulsion polymer-
ization of styrene in the absence of micelles. This dye
technique was then applied to the emulsion polymer-
ization of methyl methacrylate [69]. The water
solubility of methyl methacrylate is about 80 times as
that of styrene. The polymerization taking place in the
continuous aqueous phase and, hence, homogeneous
nucleation is greatly enhanced in the emulsion
polymerization of methyl methacrylate compared to
the styrene counterpart. Indeed, the experimental data
clearly showed that homogeneous nucleation played a
key role in the early stage of the emulsion polymer-
ization of methyl methacrylate. A mixed mode of
particle nucleation (micelle nucleation and homo-
geneous nucleation) was operative in the polymer-
ization system when the surfactant concentration was
higher than the CMC. On the other hand, homogeneous
nucleation was the predominant mechanism that
governed the population of particle nuclei produced if
there were no micelles present in the polymerization
system.
Rudschuck et al. [70] adopted uorescence spec-
troscopy to on-line investigate the nucleation and
growth of particle nuclei during the surfactant-free
emulsion polymerization of styrene. The polymer-
ization was initiated by a macroinitiator, the hydrolyzed
propenemaleic acid copolymer with t-butyl perester
groups. Pyrene molecules were incorporated into the
backbone of the macroinitiator to probe the polymer-
ization mechanisms. Four distinct regions were
observed during polymerization. The rst stage was
simply related to the initial heating period. The particle
nucleation process began with thermal decomposition
of the perester groups into t-butyl-hydroxyl radicals,
aliphatic radicals at the chain and carbon dioxide. The
free radicals attached to the macroinitiator backbone
polymerized with styrene molecules dissolved in the
continuous aqueous phase. The resultant macroinitiator
with grafted oligostyrene chains exhibited some surface
activity and contributed to the stabilization of poly-
styrene particle nuclei generated in the aqueous phase.
Adsorption of the surface-active macroinitiator was
reected in the decreased uorescence intensity ratio
I
1
/I
3
as well as the increased uorescence intensity and
the back-scattered light due to the formation of particle
nuclei. The uorescence intensity ratio I
1
/I
3
is dened
as the ratio of the peak height of the rst vibronic band
to the third vibronic band of the emission spectra of
pyrene. It represents the quantitative measure of the
non-polar nature of microenvironment in which most of
the hydrophobic pyrene molecules reside. This was
followed by the relatively constant uorescence
intensity ratio I
1
/I
3
and the increased uorescence
intensity. The former was attributed to the attachment
of macroinitiator onto the particle surface. As for the
latter, it was caused by the growth of particles at the
expense of monomer droplets. Finally, the uorescence
intensity increased rapidly toward the end of polymer-
ization as a result of the gel effect.
The way that monomer-swollen particles grow
during emulsion polymerization has a signicant
inuence on their particle morphology. Emulsion
polymerization is generally initiated by a water-soluble
initiator such as NaPS. Thermal decomposition of this
kind of initiator generates two sulfate radicals in the
continuous aqueous phase. These free radicals undergo
the propagation reaction with monomer molecules
dissolved in the aqueous phase to form oligomeric
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 451
radicals. These oligomeric radicals become surface-
active and tend to penetrate the particles and continue
to grow in size therein. The anionic sulfate end-groups
of oligomeric radicals are expected to remain near the
particle surface layer (termed the anchoring effect).
Thus, non-uniform polymerization of free radicals with
monomer molecules may not take place deep inside the
particles. This mechanism played an important role in
controlling polymerization kinetics and particle mor-
phology. Amalvy et al. [71] adopted the electron
spectroscopy imaging in an analytical TEM to study the
elemental distribution in the particles during the ab
initio batch and semibatch seeded emulsion polymer-
ization of styrene. Experimental data showed that
polymer chains remained at the position where they
were produced and the sulfate groups originating from
the persulfate initiator were distributed uniformly
within the particles. No conclusive proof for the
anchoring effect was found in this work. This
interesting subject is still debatable and more discus-
sion will be given hereinafter (see Section 3).
Tauer et al. [72] demonstrated that the heat ow
versus time or monomer conversion proles obtained
from reaction calorimetry clearly reected changes in
the recipe ingredients and reaction conditions in
emulsion polymerization. Varela De La Rosa et al.
[73] used the reaction calorimetry technique to
illustrate the signicant implication of the maximum
rate of polymerization in the emulsion polymerization
of styrene. They pointed out that the point at which the
maximum rate of polymerization occurred signied the
end of the particle nucleation stage and the disap-
pearance of monomer droplets. The authors also used
this technique to study the effect of the initial ratio of
monomer to water on the emulsion polymerization of
styrene [74]. The surfactant concentration (40 mM
sodium dodecyl sulfate (SDS)) was kept at a constant
level higher than the CMC. The experimental data
showed that, at low ratios of monomer to water (nal
total solid content&10%), micelle nucleation took
place throughout the polymerization and no constant
polymerization rate period was observed. At high ratios
of monomer to water (nal solid contentS20%),
particle nuclei formed rst via the predominant micelle
nucleation mechanism. This was followed presumably
by homogeneous nucleation immediately after the
depletion of micelles. Furthermore, the length of
homogeneous nucleation increased with increasing
the ratio of monomer to water. These particle
nucleation phenomena were supported by the polymer-
ization kinetic data. Based on these experimental
results, the authors proposed that the generally
recognized Interval II could not be characterized by
the constant number of particles per unit volume of
water and constant rate of polymerization for the
polymerization with a surfactant concentration greater
than its CMC. Interval II was characterized by the
continuous particle nucleation process and increased
rate of polymerization instead. As expected, Interval II
ended when monomer droplets disappeared. Never-
theless, particle nucleation might not cease at this time.
This study illustrates that the information on the
evolution of particle nuclei throughout the reaction
was required to monitor the whole particle nucleation
process.
Sajjadi [75] investigated the diffusion-controlled
nucleation and growth of particle nuclei in the emulsion
homopolymerizations of styrene and methyl methacryl-
ate. The polymerization started with two stratied
layers of monomer and water containing surfactant and
initiator, with the water layer being stirred gently. In
this manner, the rate of transport of monomer became
diffusion-limited. As a result, the rate of growth of latex
particles was reduced signicantly and more latex
particles were nucleated.
In emulsion polymerization, monomer can be
transported to the growing latex particles by molecular
diffusion from the continuous aqueous phase or by the
shear induced collision between the monomer droplets
and particles. Kim et al. [76] developed the vapor phase
addition method to study the molecular diffusion of
monomer from the aqueous phase to the particlewater
interface. In the absence of monomer droplets, this
technique involved only the transport of monomer from
a separate reservoir to the aqueous phase, and then into
the particles. Furthermore, the resistance to monomer
transfer at the waterair interface was minimized to the
extent that the resistance at the particlewater interface
predominated in the absorption of monomer by the
particles. The overall mass transfer coefcient of vinyl
acetate for the monodisperse polystyrene particles
stabilized by SDS was determined to be in the order
of 10
K7
cm min
K1
, and it was independent of the
agitation speed in the range of 400500 rpm, which
served as supporting evidence of the diffusion-limited
mass transfer of monomer from the aqueous phase to
the particles.
2.3. Surfactant-free polymerization systems
Surfactant-free emulsion polymerization is an
important industrial process for the manufacture of
polymeric materials with excellent water resistance and
adhesion properties. In the absence of surfactant,
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 452
limited occulation of latex particles greatly reduces
the number of particles per unit volume of water (or
increases the particle size) with the progress of
polymerization. This will inevitably make the particle
nucleation and growth mechanisms more complicated.
Furthermore, intensive coagulation of the particles to
form lterable solids and scraps adhering to the reactor
wall and agitator could become a serious problem in the
plant production. Thus, the colloidal stability issue that
has been generally ignored in the past must be
addressed from both the theoretical and practical points
of view. Micelle nucleation is generally ruled out for
the polymerization system in the absence of micelles.
Tauer et al. [77] studied the surfactant-free emulsion
polymerization of styrene initiated by KPS. At the
moment of particle nucleation, the number of particle
nuclei per unit volume of water increased rapidly to
2!10
16
dm
K3
and more than one polymer chain per
particle was obtained. It was the aggregation nucleation
(or homogeneous nucleation accompanied with limited
occulation) mechanism that controlled the particle
size and particle size distribution of latex products.
A small amount of functional monomers such as
acrylic acid and methacrylic acid are commonly
incorporated into the surfactant-free emulsion polymers
to improve the colloidal stability during polymer-
ization. Wang and Pan [78] studied the surfactant-free
emulsion polymerization of styrene with the water-
soluble comonomer 4-vinylpyridine. At the very
beginning of the copolymerization of styrene and
4-vinylpyridine in the continuous aqueous phase,
oligomers rich in the monomeric units of 4-vinylpyr-
idine were generated. It was postulated that these
surface-active oligomers formed monomer-swollen
micelles that were available for the subsequent particle
nucleation. In addition, they were capable of stabilizing
the monomer droplets. This was followed by the typical
particle growth period. Ni et al. [79] studied the effects
of adding 8 wt% ethyl acetate in combination with
slower agitation speeds (100 or 200 rpm) on the
surfactant-free emulsion copolymerization of 4-vinyl-
pyridine and styrene in order to clarify the particle
nucleation mechanism. The experimental results
showed that both the nucleation and growth of latex
particles were closely related to the monomer droplets
stemming from the oilwater interface generated by
agitation and/or from the condensation of monomer
molecules dissolved in the continuous aqueous phase.
The number of particles per unit volume of water
increased with increasing the agitation speed. In
addition to molecular diffusion of monomers through
the aqueous phase, coalescence among the particles and
monomer droplets was primarily responsible for the
transport of monomers to the reaction loci. Ou et al.
[80] investigated the effect of the hydrophilic
comonomer (vinyl acetate or methyl methacrylate) on
particle nucleation in the surfactant-free emulsion
polymerization of styrene. The GPC data clearly
showed a population of oligomers with a molecular
weight of about 1000 g mol
K1
obtained from the early
stage of polymerization. This was attributed to micelle
nucleation. Isolation of the above oligomers and
characterization of the surface activity of these
oligomers in the aqueous phase are of great interest to
gaining a fundamental understanding of the related
particle nucleation mechanisms.
Yan et al. [81] investigated the surfactant-free
emulsion copolymerization of styrene, methyl metha-
crylate and acrylic acid initiated by ammonium
persulfate. As expected, both the rates of particle
nucleation and polymerization increased with increas-
ing the concentration of acrylic acid or initiator. The
persulfate initiator and temperature were the predomi-
nant parameters that governed the particle nucleation
process (homogeneous nucleation accompanied with
limited occulation). A shell growth mechanism was
proposed to describe the particle growth stage.
Mahdavian and Abdollahi [82] carried out the
surfactant-free emulsion copolymerizations of styrene
and butadiene in the presence of various levels of
acrylic acid. It was shown that the minor comonomer,
acrylic acid, had a signicant inuence on particle
nucleation. The number of latex particles per unit
volume of water and rate of polymerization increased
with increasing the concentration of acrylic acid.
However, no signicant difference in the rate of
polymerization per particle in all the polymerizations
was observed. Furthermore, the growth of particles was
less sensitive to changes in the concentration of acrylic
acid.
Zhang et al. [83] prepared cationic emulsion
copolymers of styrene, butyl acrylate and N,N-
dimethyl,N-butyl,N-methacryloloxylethyl ammonium
bromide via the surfactant-free emulsion polymer-
ization process. Azobis(isobutyramidine hydrochlo-
ride) was used as the cationic initiator. Methanol was
employed to improve the solubility of monomers in the
continuous aqueous phase. The latex particle size
decreased with increasing the concentration of the
cationic comonomer or initiator. By contrast, the
particle size rst decreased and then increased with
increasing the level of methanol. It was postulated that
particle nuclei were generated via both the micelle
nucleation and homogeneous nucleation mechanisms
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 453
based on the particle size and polymer molecular
weight data.
Shaffei et al. [84] prepared a series of sodium
bisulte adducts by the addition reaction of sodium
bisulte on the carbonyl groups of various compounds
(e.g. acetone). Surfactant-free emulsion polymerization
with KPS as the oxidizing agent at low temperature was
used to evaluate the effectiveness of the sodium
bisulte adducts as the reducing agent. At exactly the
same equivalent weight of bisulte anion, the rate of
polymerization in decreasing order was acetone sodium
bisulteOmethyl propyl ketone sodium bisulteO
octyl aldehyde sodium bisulteOacetaldehyde sodium
bisulteObenzaldehyde sodium bisulteOsalycilalde-
hyde sodium bisulteOcyclohexanone sodium bisul-
te. In addition, the resultant latex particle size in
increasing order was octyl aldehyde sodium bisulte!
salycilaldehyde sodium bisulte!benzaldehyde
sodium bisulte!cyclohexanone sodium bisulte!
methyl propyl ketone sodium bisulte!acetone
sodium bisulte!acetaldehyde sodium bisulte.
Based on the polymerization kinetic data, among the
reducing agents, acetone sodium bisulte, methyl
propyl ketone sodium bisulte and octyl aldehyde
sodium bisulte are useful in industrial applications. It
is also interesting to verify the capability of these redox
initiator systems in manipulating the particle size of
latex products.
Sahoo and Mohapatra [85] studied the catalytic
effect of the in situ developed Cu(II)EDTA complex
with ammonium persulfate on the surfactant-free
emulsion polymerization of methyl methacrylate. The
rate of polymerization at 50 8C was proportional to the
concentrations of Cu(II), EDTA, ammonium persulfate
and methyl methacrylate to the 0.35, 0.69, 0.57 and
0.75 powers, respectively. In addition, the apparent
activation energy and activation energies of the initiator
decomposition, propagation and termination, respect-
ively, were determined to be 34.5, 26.9, 29 and
16 kJ mol
K1
. It was proposed that the complex just
acted as an effective surfactant in stabilizing the
nanoparticles nucleated during polymerization. Inde-
pendent experiments are required to verify this
speculation and clarify the related stabilization
mechanism.
As aforementioned, micelle nucleation is generally
not considered as an appropriate mechanism for the
formation of latex particles in the surfactant-free
emulsion polymerization. This point of view has been
reconrmed [77,81]. However, in recent studies on
surfactant-free emulsion polymerization [78,80,83], it
has been postulated that surface-active oligomers form
in situ and then aggregate together to form monomer-
swollen micelles in the continuous aqueous phase
during the early stage of polymerization. The polymer-
ization in the presence of functional monomers is
especially prone to follow this particle nucleation
mechanism. How to reconcile the particle nucleation
mechanisms in dispute represents a great challenge to
colloid and polymer scientists.
2.4. Ionic surfactant stabilized polymerization systems
Cutting and Tabner [86] adopted the electron spin
resonance (ESR) to measure the concentration of the
occluded free radicals in latex particles during the
emulsion copolymerization of methyl methacrylate and
butyl acrylate. SDS and KPS were used as the
surfactant and initiator, respectively. The occluded
free radicals were observed only for the polymer-
izations with the mole fraction of methyl methacrylate
greater than 0.96. The concentration of the occluded
free radicals reached a maximum at a monomer
conversion of about 95% and, thereafter, it decreased
toward the end of polymerization via the bimolecular
termination reaction mechanism. It was also shown that
the rate of consumption of butyl acrylate was much
faster than that of methyl methacrylate even though
methyl methacrylate had a greater propagation rate
constant. This may be closely related to the different
partition coefcients of the two monomers between the
continuous aqueous phase and the particle phase. The
concentration of the more hydrophobic butyl acrylate in
the particles should be higher compared to the methyl
methacrylate counterpart. Furthermore, the major
propagation reaction of free radicals with monomer
molecules takes place in the growing particles. Thus,
butyl acrylate is consumed faster and the emulsion
polymer formed earlier in the reaction is rich in the
monomeric units of butyl acrylate. The resultant
polymer particles should exhibit an onion-like struc-
ture, in which the concentration of the monomeric units
of butyl acrylate continuously decreases from the core
to the surface region.
It is well known that the presence of oxygen in the
emulsion polymerization system causes an inhibition
period and retards the rate of polymerization. Being one
of the most common impurities in the industrial
emulsion polymerization system, dissolved oxygen
results in the reduction in the reactor productivity
and, therefore, increases the process cost. Furthermore,
the safety concern of operating an industrial reactor
with potentially explosive monomer vapors in an
oxygen-rich headspace requires elimination of oxygen
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 454
from the polymerization system. Therefore, it is crucial
to gain a better understanding of the effect of dissolved
oxygen on the polymerization kinetics for minimizing
the batch-to-batch variations in the latex plant
production. Cunningham et al. [87] studied the
inuence of dissolved oxygen on the emulsion
polymerization of styrene stabilized by SDS and
initiated by KPS by varying the initial levels of oxygen
dissolved in the continuous aqueous phase correspond-
ing to 0100% of saturation. It was shown that the
length of the induction period did not vary linearly with
the initial oxygen concentration. This implied that
diffusion of oxygen from the headspace to the aqueous
phase played an important role in retarding the free
radical polymerization. At higher initial oxygen levels,
emulsion polymer with lower molecular weight formed
early in the reaction. This indicated that the retardation
of polymerization occurred in the latex particle phase.
Furthermore, higher initial oxygen levels led to smaller
latex particles presumably due to the prolonged particle
nucleation period. Based on the partitioning calcu-
lations and experimental data, the authors suggested
that oxygen should be modeled as both the water-
soluble and oil-soluble inhibitor.
Capek et al. [88] studied the effect of temperature on
the kinetics of emulsion polymerization of styrene
stabilized by SDS and initiated by NaPS. The kinetic
data showed two distinct non-stationary polymerization
rate intervals with one very short constant rate interval
in between and a shoulder occurring at high monomer
conversion. Several possible polymerization mechan-
isms such as the long particle nucleation period,
formation of additional free radicals by the thermally
induced initiation reaction, suppressed degradation of
monomer droplets, bimolecular termination of charged
free radicals in the continuous aqueous phase and
relatively narrow particle size distribution throughout
the reaction that were responsible for the non-stationary
polymerization rate behavior were discussed. The
maximal rate of polymerization or number of latex
particles per unit volume of water was proportional to
the rate of initiation to the 0.27 power, indicating a
lower efciency of particle nucleation compared to the
classical emulsion polymerization. The lower acti-
vation energy of polymerization was attributed to the
small energy barrier for the absorption of free radicals
by latex particles and it was controlled by the initiation
and propagation reactions. The high ratio of absorption
of free radicals by the particles to formation of free
radicals in the aqueous phase was attributed to the
efcient entry of uncharged free radicals and formation
of additional free radicals by the thermally induced
initiation reaction. It would be interesting to carry out
the emulsion polymerizations of other hydrophobic
acrylate monomers than styrene (e.g. butyl acrylate or
2-ethylhexyl acrylate) under exactly the same exper-
imental conditions. In this manner, the contribution of
thermal polymerization can be minimized or even
eliminated. Thus, the role of the thermally induced
initiation reaction in the emulsion polymerization of
styrene can be better elucidated by comparing the
kinetic data obtained from the polymerizations of
styrene and acrylate monomers side by side.
Fang et al. [89] investigated the effects of the seed
latex particle size, initiator concentration and tempera-
ture on the transport of free radicals between the
continuous aqueous phase and particle phase in the
seeded emulsion polymerization of styrene using SDS
and KPS as the surfactant and initiator, respectively. It
was proposed that absorption of free radicals by the
particles should be a competitive process. The
competition between molecular diffusion of free
radicals from the bulk aqueous phase to particle
water interface and the chemical reactions of free
radicals in the aqueous phase determined the prob-
ability of absorption of free radicals by the particles.
This implied that the physicochemical properties and
reactions of free radicals in the aqueous phase played an
important role in the emulsion polymerization kinetics.
It is generally accepted that emulsion polymer-
ization is only applicable to hydrophobic monomers
such as styrene, butadiene, butyl acrylate and methyl
methacrylate when conventional persulfate initiators
and anionic or non-ionic surfactants are employed. The
homopolymerization of water-soluble monomers (e.g.
2-hydroxyethyl methacrylate, N-methylolacrylamide,
acrylic acid and methacrylic acid) using the persulfate
initiator in water results in aqueous solution polymer or
hydrogel if a crosslinking mechanism is incorporated
into the polymerization system. Recently, Imroz Ali
et al. [90] studied the effects of the types of initiator and
surfactant on the emulsion polymerization of 2-hydro-
xyethyl methacrylate, methyl methacrylate or butyl
methacrylate. For 2-hydroxyethyl methacrylate, stable
latex products were produced only if hydrophobic
initiators (e.g. AIBN or benzoyl peroxide) in combi-
nation with alkyl sulfate with the alkyl chain length
greater than 10 or surface-active initiators (e.g. 2,2
/
-
azobis(N-2
/
-methylpropanoyl-2-amino-alkyl-1)-sulfo-
nate) with the alkyl chain length greater than eight. It
was illustrated that formation of stable colloidal
particles required a proper choice of initiator and
surfactant to match the hydrophobicity of major
monomers used in the recipe.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 455
Lamb et al. [91] studied how to adequately use the
reaction calorimetry technique in emulsion polymer-
ization. They strongly recommended the quenching
protocol in order to obtain accurate kinetic data. It was
shown that the enthalpy of polymerization should be
determined by using the apparatus and polymerization
system of the study, not just literature values taken from
different polymerization systems by different means.
Furthermore, a signicant reduction in the enthalpy of
polymerization was observed for the reaction system
containing very small latex particles (e.g. 55 nm in
diameter). This was attributed to the solvation and/or
surface effects.
Adequate agitation not only provides the hetero-
geneous emulsion polymerization system with satisfac-
tory mixing and heat transfer but also has a signicant
inuence on the polymerization mechanisms and kinetics
and the colloidal stability. For example, Ramirez et al.
[92] illustrated the effect of agitation on the emulsion
polymerization of styrene stabilized by SDS and initiated
byKPS. The experimental results showedthat the optimal
rate of polymerization and polymer molecular weight
were obtained from the run with the optimal range of
agitation speed, which was strongly dependent on the
hydrodynamic characteristics of the reactor and polymer-
ization conditions. By contrast, the polymerization above
the optimal range of agitation speed resulted in a smaller
number of latex particles per unit volume of water and a
lower polymer molecular weight. This was attributed to
the enhanced absorption of free radicals by the particles.
These free radicals might come from the continuous
aqueous phase or from the limited occulated particles
induced by shear force.
Mayer et al. [93] demonstrated that large-scale
reactors (1.2 dm
3
used in this work) could be operated
under both the stable and isothermal conditions for the
batch emulsion polymerization of styrene stabilized by
SDS and initiated by NaPS. First, the polymerization
was carried out at a relatively low temperature (e.g.
50 8C). When the rate of polymerization leveled off,
complete monomer conversion was achieved within an
acceptable period of time simply by raising the reaction
temperature (e.g. from 50 to 75 8C) provided that the
particles were relatively small (!150 nm in diameter).
The authors did not offer an explanation for the
observed limiting monomer conversion of about 90%
at 50 8C. This phenomenon is generally attributed to the
scenario that the polymerization temperature is well
below the glass transition temperature of the monomer-
starved polystyrene particles. Under the circumstances,
even the propagation reaction of polymeric radicals
with monomer molecules inside the particles becomes
diffusion-controlled. Thus, the rate of polymerization is
greatly retarded.
Badran et al. [94] studied the sodiumdodecyl benzene
sulfonate/polyvinyl alcohol stabilized emulsion (co)po-
lymerizations of styrene and styrene and butyl acrylate.
The initiator package included KPS and different
aromatic aldehydic sodiumbisulte adducts (methylben-
zaldehyde sodium bisulte, chlorobenzaldehyde sodium
bisulte, methoxybenzaldehyde sodium bisulte, and
sodium bisulte). The number of latex particles per unit
volume of water decreased with increasing the initiator
concentration or decreasing the surfactant concentration.
Increasing the fraction of butyl acrylate in the monomer
mixture resulted in a decrease in the number of particles
per unit volume of water. The emulsion (co)polymeriza-
tion systems investigated in this work did not follow the
SmithEwart model.
Gu et al. [95] developed a single stage emulsion
polymerization process to prepare micron-sized poly-
styrene particles. The polymerization was initiated by
KPS and the cationic cetyltrimethylammonium bro-
mide and anionic SDS were used to control the limited
occulation and colloidal stability of latex particles
during polymerization. Changes in the concentration of
initiator and agitation speed did not have signicant
effects on the particle size distribution of latex products.
By contrast, the concentration of monomer had a
signicant inuence on the average particle size, and
the time at which the cationic or anionic surfactant was
added to the polymerization system played an
important role in determining the particle size
distribution. On the other hand, Omi et al. [96]
proposed to use a two-step polymerization process to
prepare composite emulsion polymers with an average
particle size up to 1 mm. The rst stage involved the
preparation of negatively charged polymethyl metha-
crylate seed particles by using SDS and ammonium
persulfate as the surfactant and initiator, respectively.
After the removal of unreacted monomer and initiator
by dialysis, the seed particles were swollen with
dimethylaminoethyl methacrylate and styrene. This
was followed by addition of the cationic 2,2
/
-azobis(2-
amidinopropane)$2HCl to initiate the second stage
emulsion polymerization. A small amount of non-
ylphenol polyethoxylate with an average of 23 units of
ethylene oxide per molecule was used in the seeded
emulsion polymerization to improve the colloidal
stability. The maximal growth of seed particles was
achieved for the polymerization at pH 10 and with a
lower non-ionic surfactant concentration. The latex
products comprised stable aggregates of several seed
particles with dimethylaminoethyl methacrylate
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 456
incorporated among the aggregated seed particles and
deposited on the aggregated seed particle surface. The
secondary nucleation of particles rich in dimethylami-
noethyl methacrylate was promoted for the polymer-
ization at pH 9. In addition, reducing the level of non-
ionic surfactant enhanced the growth of seed particles.
Ito et al. [97] rst prepared negatively charged
poly(methyl methacrylate-co-methyl acrylate) seed
particles using SDS and ammonium persulfate as the
surfactant and initiator, respectively. This was followed
by the seeded emulsion copolymerization of styrene
and the cationic dimethylaminoethyl methacrylate
stabilized by an adequate amount of nonylphenol
polyethoxylate with an average of 23 monomeric
units of ethylene oxide per molecule and initiated by
the cationic 2,2
/
-azobis(2-amidinopropane)$2HCl. It
was shown that intensive mixing caused the polymer-
ization system to lose its colloidal stability. The smaller
seed particles resulted in a well-controlled occulation
process and a larger particle size. At pH 8.08.5, a quite
high level of coagulum was produced due to the
extensive coagulation experienced in the course of
polymerization. At pH 9.510.0, the polymerization
system was not as stable as that carried out at optimal
pH 8.759.0. Recently, the authors studied the reaction
mechanisms of the styrene emulsion polymerization in
the presence of anionic poly(methacrylate-co-methyl
acrylate) seed particles accompanied with the regulated
occulation of particles induced by the addition of
dimethylaminoethyl methacrylate and 2,2
/
-azobis(2-
methylamidinopropane)$2HCl [98]. The occulation
process was induced by the electrostatic attraction force
between the anionic seed particles and cationic
oligomeric chains or precursor particles containing
the monomeric units of dimethylaminoethylmethacry-
late and end groups derived from 2,2
/
-azobis(2-
methylamidinopropane)$2HCl. Dimethylaminoethyl
methacrylate was preferentially partitioned into the
continuous aqueous phase, whereas styrene tended to
swell the seed particles. The cationic oligomeric
radicals and, subsequently, precursor particles rich in
the monomeric units of dimethylaminoethyl methacry-
late were generated by homogeneous nucleation. The
rate of polymerization remained relatively constant
during the rst 10-h period. This was attributed to the
slow rate of the bimolecular termination reaction in the
particles along with the enhanced desorption of the
mobile free radicals generated by the chain transfer
reaction. The occulated particles remained spherical
before the polymerization temperature was raised from
60 to 80 8C. The spherical particles started to occulate
and formed irregular shaped particles during
the heating stage, and the reaction mechanisms
responsible for the changed particle morphology were
offered. All the controlled occulation processes
discussed above represent an effective tool to speed
up the particle growth rate during polymerization
without detriment to the colloidal stability of latex
products. However, the feasibility of applying these
unique processes to the plant production of such latex
products requires further evaluation because the levels
of lterable solids and total scraps adhered to the
reactor wall and agitator were not reported. The
formation of high levels of lterable solids and scraps
during polymerization increases the cost and is not
acceptable.
The oxidant t-butyl hydroperoxide in combination
with the reducing agent sodium formaldehyde sulfoxy-
late have been widely used in initiating the emulsion
polymerization. Wang et al. [99] reported that sodium
formaldehyde sulfoxylate induced the aqueous polymer-
ization of methacrylate ester monomers such as methyl
methacrylate, ethyl methacrylate, butyl methacrylate
and methacrylic acid, but the resultant products were in
the form of lumped polymer. On the other hand, the
needle-shaped crystalline acrylate sulphone dimmers
were obtainedfromthe reactionof sodiumformaldehyde
sulfoxylate with acrylate ester monomers (e.g. methyl
acrylate, ethyl acrylate and butyl acrylate) in water. The
authors then demonstrated that sodium formaldehyde
sulfoxylate in combination with SDS were capable of
initiating the emulsion polymerization of styrene under
neutral condition and stable latex products with a
relatively monodisperse particle size distribution were
achieved [100,101]. The reaction system was character-
ized by a typical free radical polymerization based on the
electron paramagnetic resonance spectra. In fact, the
authors postulated that the colloidal system was
established according to the surfactant-free emulsion
polymerization mechanism. The conductivity data for
the aqueous SDS solutions in the presence of 1.44!10
K
2
M sodium formaldehyde sulfoxylate did not exhibit
any break point in the conductivity versus surfactant
concentration plot and this was taken as supporting
evidence of the absence of monomer-swollen micelles.
The rate of polymerization was independent of the
styrene concentration. It was proportional to the
concentration of SDS or sodium formaldehyde sulfox-
ylate to the 0.7 power. The nucleation and growth of
particle nuclei in emulsion polymerization initiated by
such a novel initiator package are not clear at this time
and deserve further research efforts.
Lin et al. [102] studied the emulsion polymerization
of styrene stabilized by SDS. They observed
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 457
the existence of water-in-oil-in-water (w/o/w) reverse
micelles or double emulsion droplets, as evidenced
from the measurement of CMC and the observation
experiment of w/o/w double emulsion droplets. The
existence of monomer droplets at high monomer
conversion and relatively low mass ratio of styrene to
polystyrene in latex particles was also demonstrated.
The enhanced uniformity and stability of monomer
emulsion was attributed to the prolonged pre-emulsi-
cation period (2 h) and/or accumulation of poly-
styrene in the colloidal system at 60 8C. This would
reduce the degradation of monomer droplets, induce the
formation of double emulsion droplets and promote the
apparent monomer-starved condition during polymer-
ization. As a result, the polymerization system deviated
from the conventional SmithEwart theory. This work
illustrates the unexpected effects of characteristic
properties of monomer emulsions initially present in
the colloidal system on the subsequent polymerization
mechanisms and kinetics. How to design experiments
to verify the proposed polymerization mechanisms and
further distinguish the individual effects represents a
challenge to colloid and polymer chemists.
The growth of newly born particle nuclei in
conventional SmithEwart interval II can be achieved
without any problem by molecular diffusion of
monomer from monomer droplets to particle nuclei.
By contrast, particle nuclei formed in Interval III in the
absence of monomer droplets involve diffusion of
monomer from the existing monomer-starved polymer
particles to particle nuclei. As the number of latex
particles per unit volume of water increases with the
progress of polymerization, the rate of particle growth
decreases and the particle nucleation period is
prolonged concomitantly. As a result, a very large
population of particles can be obtained. Sajjadi [103]
used the seeded emulsion polymerization of styrene to
demonstrate this concept. Furthermore, the decreased
rate of particle growth in Interval III resulted in the
positively skewed particle size distribution in terms of
volume for the secondary particles. It was also pointed
out that cessation of particle nucleation in Interval III
could not be marked by the depletion of micelles, as it
was generally marked for the end of particle nucleation
in conventional emulsion polymerization (Interval I).
The diminished particle nucleation in Interval III was
attributed to the decreased rate of generation of free
radicals in the continuous aqueous phase, reduced
capture of free radicals by micelles and occurrence of
homo- and hetero-occulation of particle nuclei.
The rod-like 1-[u-(4
/
-methoxy-4-biphenylyloxy)oc-
tyl]pyridinium bromide was used to stabilize
the emulsion polymerization of styrene [104]. At high
surfactant concentration (S48 mM), the constant
polymerization rate period (Interval II) disappeared.
Furthermore, the latex products showed bimodal particle
size distribution. These results were attributed to the
continuous particle nucleation process, which comprised
both the conventional and coagulative particle nucleation
mechanisms. Tofurther evaluate the performance proper-
ties of this rod-like surfactant against conventional
surfactants is of great importance to the understanding
of its surface activity, adsorptionbehavior at the oilwater
interface and total free energy barrier against coagulation
between two interactive particles.
Tang et al. [105] described an emulsier-minor
emulsion polymerization process that dealt with
copolymerization of methyl methacrylate, butyl acry-
late and styrene in combination with water-soluble
anionic monomer (methacrylic acid or acrylic acid) and
non-ionic monomer (N-methylol acrylamide). It was
proposed that water-soluble monomers played an
important role in stabilizing latex particles because
they could be incorporated into the particle surface
layer and formed a protective layer that imparted the
steric and/or electrostatic stabilization effects to the
particles and prevented two approaching particles from
coagulation during polymerization. The minor surfac-
tant, sodium alkylated diphenyl ether disulfonate, at a
level greater than its CMC generated particle nuclei by
micelle nucleation and, therefore, controlled the
number of particles per unit volume of water (or
particle size and particle size distribution) and the rate
of polymerization.
Short chain alcohols (e.g. 1-pentanol) with some
surface activity have been widely used as the
cosurfactant to stabilize microemulsion polymer-
ization. Recently, Chern and Yu [106] demonstrated
that 1-pentanol also had a signicant effect on the
styrene emulsion polymerization. They studied the
inuence of 1-pentanol on the styrene emulsion
polymerization mechanisms and kinetics. It was
shown that the CMC value of styrene emulsion
stabilized by SDS rst decreased rapidly and then
leveled off when the concentration of 1-pentanol
increased from 0 to 72 mM. The effect of 1-pentanol
increased to a maximum and then decreased when the
concentration of SDS increased from 2 to 18 mM.
When the SDS concentration was set at 2 mM,
homogeneous nucleation controlled the polymerization
kinetics regardless of the 1-pentanol concentration. At a
SDS concentration of 4 mM, the effect of 1-pentanol
appeared due to the transition from homogeneous
nucleation to a mixed mode of particle nucleation
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 458
(homogeneous nucleation and micelle nucleation)
occurred when the concentration of 1-pentanol
increased from 0 to 72 mM. The effect of 1-pentanol
was the strongest at a SDS concentration of 6 mM since
particle nucleation mechanisms spanned homogeneous
nucleation (low 1-pentanol concentration), a mixed
mode of particle nucleation (homogeneous nucleation
and micellar nucleation) (medium 1-pentanol concen-
tration) and micelle nucleation (high 1-pentanol
concentration). At SDS concentrations greater than
6 mM, in which micelle nucleation predominated in the
polymerization kinetics, the effect of 1-pentanol
decreased rapidly with increasing SDS concentration.
It should be noted that this work demonstrated the
important role of monomer-swollen micelles during the
particle nucleation period. At constant SDS concen-
tration (e.g. 6 mM), the number of monomer-swollen
micelles available for particle nucleation could vary
from nil to a quite large population simply by
manipulation of the level of 1-pentanol. Apparently,
homogeneous nucleation alone could not explain the
data of the number of latex particles per unit volume
and the rate of polymerization.
Xia et al. [107] studied the ultrasonically initiated
emulsion polymerization of butyl acrylate. Based on the
experimental data, a free radical polymerization mech-
anism including the source of free radicals, formation of
free radicals and locus of polymerization was proposed.
Yin and Chen [108] studied the surfactant-free emulsion
copolymerization of butyl acrylate and acrylamide
initiated by a 20 kHz ultrasonic generator. It was
proposed that particle nucleation began with the
preferential polymerization of water-soluble acrylamide
in the continuous aqueous phase. The hydrophilicity of
oligomeric radicals continued to decrease with the
progress of polymerization. These oligomeric radicals
then had the chance to copolymerize with butyl acrylate,
followed by formation of monomer-swollen aggregates
with the micelle-like structure (particle nuclei). After the
formation of particle nuclei, butyl acrylate started to
participate the propagation reaction therein. Neverthe-
less, the polymerization of acrylamide in the aqueous
phase took place continuously due to the cavitation
bubble produced by ultrasound. The aqueous polymer-
ization of acrylamide then became the primary event
toward the end of polymerization immediately after the
depletion of butyl acrylate droplets. The rate of
polymerization increased with increasing the power
output of ultrasound. The authors then developed a
general kinetic model for the prediction of particle
nucleation in the ultrasonically initiated emulsion
polymerization [109]. This mechanistic model took into
account homogeneous nucleation, micelle nucleation and
monomer droplet nucleation. The inuence of ultrasound
on the formation of monomeric and oligomeric radicals
was also taken into consideration in the model
development. This model was assessed by experimental
data obtained from the ultrasonically initiated emulsion
polymerizationof styrene. Agoodagreement betweenthe
model simulation and experimental data was achieved.
2.5. Non-ionic surfactant and mixed anionic and non-
ionic surfactants stabilized polymerization systems
Mixed anionic and non-ionic surfactants have been
widely used in industry to manufacture latex particles.
Anionic surfactants can provide repulsive force
between two similarly charged electric double layers
to the latex particles. By contrast, non-ionic surfactants
can impart two approaching particles with the steric
stabilization mechanism. In addition, non-ionic surfac-
tants can improve the chemical and freezethaw
stability of latex products. The following studies clearly
show that the polymerization mechanisms and kinetics
involved in a variety of emulsion polymerizations
stabilized by non-ionic or mixed anionic/non-ionic
surfactants are far more complicated than the conven-
tional SmithEwart theory.
Ozdeer et al. [110] studied the emulsion copolymer-
ization of styrene and butyl acrylate using octylphenol
polyethoxylate with an average of 40 monomeric units
of ethylene oxide per molecule (Triton X-405) as the
sole surfactant. Although the levels of Triton X-405
were all well above its CMC, the substantial partition-
ing of Triton X-405 into the oil phase led to the scenario
that the concentrations of Triton X-405 in the
continuous aqueous phase were well below the CMC
for polymerizations with the lowest to intermediate
levels of Triton X-405. As a consequence, latex
products with unimodal particle size distribution were
achieved for polymerizations with the lowest and
highest levels of Triton X-405. On the other hand,
latex products with bimodal particle size distribution
were obtained from polymerizations with intermediate
levels of Triton X-405. These experimental results were
attributed to homogeneous nucleation and coagulative
nucleation for polymerizations with lower levels of
Triton X-405, to homogeneous nucleation and coagu-
lative nucleation followed by micelle nucleation for
polymerizations with intermediate levels of Triton
X-405, and then to micelle nucleation for polymer-
izations with higher levels of Triton X-405.
Capek and Chudej [111] studied the emulsion
polymerization of styrene stabilized by polyethylene
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 459
oxide sorbitan monolaurate with an average of 20
monomeric units of ethylene oxide per molecule
(Tween 20) and initiated by the redox system of
ammonium persulfate and sodium thiosulte. The
constant polymerization rate period was not observed
in this work. The maximal rate of polymerization was
proportional to the initiator and surfactant concen-
trations to the K0.45 and 1.5 powers, respectively. The
nal number of latex particles per unit volume of water
was proportional to the initiator and surfactant
concentrations to the 0.32 and 1.3 powers, respectively.
In addition, the resultant polymer molecular weight was
proportional to the initiator and surfactant concen-
trations to the 0.62 and K0.97 powers, respectively.
Some possible mechanisms were used to explain the
deviation of the polymerization system from the
classical SmithEwart theory. Lin et al. [112] investi-
gated the emulsion polymerization of styrene stabilized
by nonylphenol polyethoxylate with an average of 40
monomeric units of ethylene oxide per molecule (NP-
40) and initiated by NaPS. The rate of polymerization
versus monomer conversion curves exhibited two non-
stationary polymerization rate intervals and a vague
constant rate period in between. The rate of polymer-
ization and number of latex particles per unit volume of
water were proportional to the 1.4 and 2.4 powers,
respectively, of the NP-40 concentration. The polymer-
ization system did not follow the micelle nucleation
model and some possible reaction mechanisms respon-
sible for this deviation were discussed.
Ouzineb et al. [113] carried out emulsion (co)poly-
merizations of butyl acrylate and methyl methacrylate
with different types and concentrations of surfactants
(Triton X-405 versus SDS) to study particle nucleation
and the resultant latex particle size and particle size
distribution. The presence of the more hydrophilic
methyl methacrylate in the continuous aqueous phase
was shown to have a signicant inuence on the CMC
of Triton X-405. Furthermore, the more hydrophobic
butyl acrylate predominated in the particle nucleation
process involved in emulsion copolymerizations of
butyl acrylate and methyl methacrylate, with the nal
number of latex particles per unit volume of water very
similar to that obtained from the homopolymerization
of butyl acrylate.
Chen et al. [114] examined the general validity of
SmithEwart theory in the emulsion polymerization of
styrene stabilized by SDS and NP-40. The CMCs of
mixed surfactants were determined for various compo-
sitions at 25 and 80 8C, and the data well described by
the regular solution model for mixed micelles. Mixed
micelles exhibited a quite non-ideal behavior,
especially at lower temperature. The effect of mixed
surfactants on particle nucleation was demonstrated by
a series of emulsion polymerizations of styrene. Adding
only a small amount of SDS into the polymerization
system dramatically increased the nal number of latex
particles per unit volume of water and, therefore,
reduced the resultant particle size. Furthermore, the
polymerization system stabilized by mixed surfactants
did not follow conventional SmithEwart theory when
the level of NP-40 was relatively high. Chern et al.
[115] then showed that the polymerization system
followed SmithEwart theory only when the level of
NP-40 in the mixture of SDS and NP-40 was less than
30 wt%. However, the polymerization system deviated
from SmithEwart theory signicantly when the level
of NP-40 was higher than 50 wt%. The steric
stabilization effect provided by NP-40 alone was not
strong enough to prohibit the interactive particles from
occulation. On the other hand, mixed anionic and non-
ionic surfactants could greatly improve the colloidal
stability of the polymerization system via the synergetic
effects provided by both the electrostatic and steric
stabilization mechanisms and, thus, retard the limited
particle occulation process. The mixed surfactants of
SDS/NP-40Z20/80 (w/w) was the best choice because
it resulted in the best reproducibility of experiments and
the fastest rate of polymerization. Chern et al. [116]
studied the effect of the initiator (NaPS) concentration
on the emulsion polymerization of styrene stabilized by
SDS and NP-40. The relationship that the number of
particles nucleated per unit volume of water was
proportional to the surfactant and initiator concen-
trations to the 0.6 and 0.4 powers, respectively, was
only applicable to the polymerization system in the
absence of NP-40. At an initiator concentration of
1.38!10
K3
M, the polymerizations with 0, 50 and
80 wt% NP-40 resulted in comparable particle sizes and
relatively monodisperse particle size distributions
throughout the reaction. On the other hand, polymer-
izations stabilized by NP-40 alone showed the largest
particle sizes along with the broadest particle size
distributions. This was attributed to the long particle
nucleation period and/or limited particle occulation.
The rate of polymerization increased with increasing
the initiator concentration for polymerizations stabil-
ized by SDS. On the other hand, the rate of
polymerization remained relatively constant with an
increase in the initiator concentration for polymer-
izations stabilized by NP-40 alone. As for the
polymerizations stabilized by 50 or 80 wt% NP-40,
the rate of polymerization rst increased to a maximum
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 460
and then decreased with increasing the initiator
concentration.
Unzueta and Forcada [117] developed a mechanistic
model for the emulsion copolymerization of methyl
methacrylate and butyl acrylate stabilized by anionic
and non-ionic surfactants, which was veried by the
experimental data. This model was based on the mass
and population balance of precursor particles and the
moments of particle size distribution. It was sensitive to
such parameters as the composition of mixed surfac-
tants and surfactant concentration. A competitive
particle nucleation mechanism was incorporated into
the model to simulate the evolution of particle nuclei
during polymerization.
Boutti et al. [118] investigated the effects of the
nature of initiator (i.e. charged or electrostatically
neutral free radicals) and its interaction with the
composition of surfactants and the ux of free radicals
on the emulsion copolymerization of methyl methacryl-
ate (20 wt%) and butyl acrylate (80 wt%). In the
presence of a small amount of anionic surfactant
(sodium salt of ethoxylated fatty acid with a sulfate
end-group), the nature of initiator had an insignicant
inuence on the nal number of latex particles per unit
volume of water and rate of polymerization. However,
the electrostatically neutral initiator (hydrogen per-
oxide) did not result in stable latex products for
polymerizations in the absence of surfactant or with
the nonionic surfactant (a mixture of linear ethoxylated
fatty acids with hydroxyl end-groups) as the sole
stabilizer. Furthermore, the order of the ascorbic acid
feed and hydrogen peroxide feed showed signicant
effects on the free radical ux and, consequently,
polymerization kinetics. As expected, addition of
ascorbic acid (reducing agent) to the polymerization
system already containing hydrogen peroxide (oxi-
dizing agent) was shown to be a more effective way to
generate free radicals.
Suresh et al. [119] studied the effects of different
multifunctional monomers including ethylene glycol
diacrylate and m-diisopropenylbenzene and the type
and concentration of anionic or non-ionic surfactant on
the emulsion polymerization of butyl acrylate. Experi-
mental results showed that the nal number of latex
particles per unit volume of water decreased with
increasing the ethylene glycol diacrylate concentration.
On the other hand, the opposite trend was observed for
polymerizations in the presence of m-diisopropenyl-
benzene. These results were attributed to the different
water solubility of multifunctional monomers (water
solubility: ethylene glycol diacrylateOm-diisoprope-
nylbenzene), which led to different nucleation rates of
precursor particles. In addition, the conventional
SmithEwart theory was only applicable to the
m-diisopropenylbenzene containing polymerizations;
the increased number of particles per unit volume of
water (i.e. enhanced particle nucleation rate) was
caused by the decreased rate of particle growth due to
the reduced swelling of the crosslinked polymer
particles with monomer and concomitant prolonged
particle nucleation period [120].
Water-soluble persulfate initiators have been widely
used in emulsion polymerization. Although the oil-
soluble initiators (e.g. AIBN) have been seldom used in
emulsion polymerization, they representative alternative
when non-ionic latex products are desirable. The
feasibility of using an oil-soluble initiator to polymerize
monomer emulsions has been conrmed, but the origin of
free radicals participating in the chain polymerization is
still debatable. This is because an oil-soluble initiator
molecule is thermally decomposed into two initiator
radicals inside the monomer-swollen micelle or particle
nucleus. These two neighboring free radicals may
undergo the bimolecular termination reaction before the
propagation reaction or desorption of free radicals out of
the particle can take place during the early stage of
polymerization. Some researchgroups suggestedthat free
radicals that initiatedthe nucleationandgrowthof particle
nuclei originated from the small fraction of initiator
dissolved in the continuous aqueous phase [121,122]. On
the other hand, other groups proposed that one of the two
newly born free radicals desorbed out of the micelle or
particle nucleus before the bimolecular termination
occurred. As a result, the remaining free radical under-
went the propagation reaction with monomer molecules
inside the micelle or particle nucleus [123125].
Recently, Luo and Schork [126] carried out the AIBN-
initiated conventional emulsion polymerization and
miniemulsion polymerization of butyl acrylate in the
presence or absence of a water-soluble radical scavenger.
Triton X-405 was used as the nonionic surfactant. It was
concluded that no particle nuclei were generated in the
conventional emulsion polymerization containing a
water-soluble radical scavenger and, thus, free radicals
originatingfromthe small fractionof initiator dissolvedin
the aqueous phase predominated inthe micelle nucleation
process. Free radicals generated in the particle phase did
contribute to the propagation reaction and the extent of
contribution increased with increasing the particle size.
For the particle diameter of up to approximately 100 nm,
emulsion polymerization was initiated by free radicals
originating from the aqueous phase.
Jain et al. [127] studied the emulsion copolymer-
ization of methyl methacrylate and ethyl acrylate
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 461
stabilized by SDS and initiated by KPS or mixed
initiators of KPS and AIBN. The rate of polymer-
ization was proportional to the KPS concentration
to the 0.76G0.03 and 0.72G0.04 powers for
polymerizations initiated by KPS alone and those
initiated by KPS along with a constant AIBN
concentration, respectively. Furthermore, the rate of
polymerization was proportional to the AIBN
concentration to the 0.40 power for polymerizations
initiated by AIBN along with a constant KPS
concentration. It was postulated that desorption of
free radicals out of the latex particles was rather
facile and the cage effect was important at high
monomer conversion. As expected, the polydisper-
sity index of the resultant polymer molecular
weight was governed by the type and concentration
of initiators. Polymerization kinetics data obtained
from this work strongly suggest that the initiation
efciency of KPS is much higher than that of
AIBN.
Polymerization mechanisms and kinetics involved in
the polymerization system initiated by mixed water-
soluble and oil-soluble initiators are far more compli-
cated than those initiated by either the water-soluble or
the oil-soluble initiator alone. Taking the mixed KPS
and AIBN initiators as an example, the potential loci for
generation of initiator radicals include the continuous
aqueous phase, monomer-swollen micelles and mono-
mer droplets. As would be expected, the initiator
efciency factor could vary signicantly in the
aforementioned loci. The bimolecular termination
reaction between two adjacent initiator radicals
originating from AIBN may take place rapidly inside
the tiny monomer-swollen micelles. However, the
oligomeric radical with a sulfate end-group may
penetrate the monomer-swollen micelle already con-
taining zero or two free radicals, and the oligomeric
radical or surviving free radical can convert the
monomer-swollen micelle into a particle nucleus due
to the extremely large total oilwater interfacial area
associated with micelles. On the other hand, two
initiator radicals generated by the thermal decompo-
sition of AIBN within the gigantic monomer droplet
may escape from the cage effect and then initiate the
propagation reaction therein. Nevertheless, the prob-
ability for the monomer droplet to capture a waterborne
free radical is rather slim because of the insignicant
total monomer droplet surface area. Other processes
that may govern the transport or reaction of free radicals
(e.g. desorption of free radicals out of the micelles or
particle nuclei) may also come into play in emulsion
polymerization. Further research is required to clarify
this issue.
Chern et al. [128] studied the effects of mixed
SDS and NP-40 surfactants on the emulsion
polymerization of styrene. Experimental data
showed that the rate of polymerization was faster
for the polymerization stabilized by mixed surfac-
tants with the weight percent NP-40 in the
surfactant mixture W(NP-40)Z50 or 90% in
comparison with the polymerization stabilized only
by SDS or NP-40 (W(NP-40)Z0 or 100%).
Furthermore, the effectiveness of NP-40 in the
stabilization of latex particles was greatly reduced
as the reaction temperature was increased from 60
to 80 8C. However, incorporation of a small amount
of SDS into the polymerization system made it
insensitive to changes in temperature. It was
postulated that the limited bridging occulation
played an important role in the formation of
particles during the very early stage of polymer-
ization at 60 8C with W(NP-40)Z100 or 90%. On
the other hand, particle nuclei were relatively stable
for the polymerization with W(NP-40)Z0%. The
polymerization with W(NP-40)Z50% exhibited an
intermediate particle nucleation behavior.
A common impression about the performance of
non-ionic surfactants is the extreme difculty in
synthesizing stable latex particles via the steric
stabilization mechanism alone. However, only low
particle surface charge density originating from ionic
surfactants or initiators can improve the non-ionic
surfactant stabilized emulsion polymerization signi-
cantly. Limited particle occulation plays an important
role in the particle nucleation and growth processes
and, hence, latex products with larger particle sizes are
usually obtained. This fact signies the important
synergetic effects provided by both the electrostatic
and steric stabilization mechanisms in maintaining
satisfactory colloidal stability during emulsion
polymerization. It is noteworthy that we still know
very little about the fundamental aspects of this
research area at this point of time. Furthermore,
creative ideas regarding how to prepare purely non-
ionic latex products are in great demand.
It is noteworthy that the hydrophilic part (e.g.
polyethylene oxide) of some non-ionic surfactants has a
lower critical solution temperature greater than 70 8C.
Thus, the performance properties of the polyethylene
oxide-based surfactants are quite sensitive to changes in
temperature. The CMC of such a family of surfactants
is expected to decrease with increasing temperature.
Therefore, the reaction temperature should have
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 462
a signicant inuence on the mechanisms and kinetics
of the non-ionic surfactant stabilized emulsion
polymerization. This interesting subject certainly
deserves more research efforts in the future.
2.6. Polymerizable or degradable surfactant and
surface-active initiator stabilized polymerization
systems
Conventional surfactants such as the anionic SDS
and non-ionic NP-40 are relatively small and mobile
and these surface-active molecules tend to migrate
toward the surface layer of a polymeric lm. This
phenomenon can have a negative inuence on the
application properties such as adhesion and water
resistance of the waterborne pressure-sensitive
adhesives. One approach to alleviate the surfactant
migration problem is to adopt a polymerizable
surfactant, which has a carboncarbon double bond
and can be chemically incorporated into latex particles
during polymerization.
Chern et al. [129] carried out the emulsion
copolymerization of methyl methacrylate and butyl
acrylate stabilized by a non-ionic polymerizable
surfactant NE-40 (Asahi Denka Kogyo) and initiated
by NaPS. The molecular structure of NE-40 is shown
below:
CH
2
OCH
2
CH ZCH
2
}
CH(OCH
2
CH
2
)
40
OH
}
CH
2
OC
6
H
4
C
9
H
19
The resultant latex particle size decreased with
increasing the NE-40 concentration and increased with
increasing the NaPS concentration. The dependence of
particle size on the initiator concentration did not follow
the conventional SmithEwart theory. This was attributed
to the bridging occulation of particle nuclei during the
particle nucleation period. It was postulated that the
differences in the particle nucleation and growth stages
and colloidal stability observed in the NE-40 and NP-40
stabilized polymerization systems were due to the
different distribution patterns of surfactant molecules
within the particles. Surprisingly, the particle size
decreased with increasing the electrolyte concentration
or agitation speed. The total lterable solids and scraps,
presumably caused by the bridging occulation process,
increased rapidly with the increased sodium chloride
concentration or agitation speed. As expected, latex
products stabilized by NE-40 showed excellent stability
toward added sodium salt.
Denise and Sherrington [130] synthesized and
characterized various cationic polymerizable mono-
and divalent quaternary ammonium surfactants and
evaluated the feasibility of using these surfactants to
stabilize the emulsion polymerization of styrene or
methyl methacrylate initiated by AIBN. Montoya-Goni
et al. [131] prepared cationic surfactants with reactive
maleate or succinate groups and used them to stabilize
the emulsion polymerization of styrene initiated by azo-
N-N
/
-dimethylisobutylamidinium hydrochloride. The
maleate containing surfactant was chemically incor-
porated onto the latex particle surface layer readily, and
a very low level of surfactant remained in the
continuous phase at the end of polymerization.
Furthermore, the colloidal stability of latex particles
was satisfactory during polymerization. By contrast,
cationic surfactant with succinate group tended to
retard the polymerization. Although a signicant
fraction of the succinate containing surfactant was
located on the particle surface, the concentration of the
succinate containing surfactant was higher than that of
the maleate containing surfactant in the aqueous phase.
Klimenkovs et al. [132] synthesized a series of
polymerizable surfactants obtained from the reaction of
maleic isoimides containing a hydrophobic alkyl or
benzyl group. It should be noted that maleic isoimides
could be regarded as the activated form of maleic acid
hemiamides useful for the conversion of the last species
into maleic diamides (polymerizable surfactants). The
CMCs of these polymerizable surfactants were within
the range of many conventional surfactants. The
polymerizable surfactant with a longer alkyl chain
length resulted in satisfactory colloidal stability of the
emulsion (co)polymerization of styrene in a batch
reactor or styrene and butyl acrylate in a seeded
semibatch reactor. KPS was used as the initiator. In
batch emulsion polymerization, all the polymerizable
surfactant molecules were either grafted or strongly
adsorbed onto the particle surface. However, this was
not the case for the copolymerization of styrene and
butyl acrylate in the presence of polystyrene seed
particles carried out in a semibatch reactor. This was
attributed to the particle surface layer rich in
monomeric units of butyl acrylate. Surprisingly, all
the latex products showed poor stability to the major
tests. For those who are interested in polymerizable
surfactants (or surface-active monomers) in more
detail, please refer to Ref. [32].
In an attempt to minimize the inuence of surfactant
on the application properties of latex products, Mezger
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 463
et al. [133] synthesized and characterized three types of
light-degradable surfactants comprising a hydrophobic
moiety linked to an ionic moiety via the light-splittable
diazophenyl bond. Alkylbenzeneazosulfonate surfac-
tants were used to stabilize the emulsion polymer-
ization of methyl methacrylate initiated by KPS. It was
shown that latex products lost their colloidal stability
when irradiated and, therefore, the feasibility of this
creative idea was contmed. Coagulation of latex
particles occurred through the conversion of azosulfo-
nate to the corresponding diazonium salt and mutual
charge neutralization of the cationic and anionic
species.
The surface-active initiator (or inisurf) serves as a
supplier of free radicals to initiate emulsion polymer-
ization in the absence of conventional surfactant. This
unique initiator then switches its role to stabilize
particle nuclei generated during polymerization. This
technique can minimize the total amount of water-
sensitive species in the recipe, thereby leading to better
performance properties of latex products. Aslamazova
and Tauer [134,135] studied the emulsion polymer-
ization of styrene initiated and stabilized by surface-
active initiators. The initiators investigated included
polyethylene oxide-sulfonate-azo-compounds and 2,2
/
-
azobis(N-2-methylpropanoyl-2-amino-alkyl-1)sul-
fonate compounds. It was shown that the colloidal
stability of latex particles during polymerization
depended primarily on the polyethylene oxide and
alkyl chain lengths of both types of initiators. The
surface activity of these initiators had a stronger effect
on the heterogeneous polymerization system compared
to the rate of decomposition of initiators. In general, the
higher the surface activity of initiators, the higher the
particle surface charge density. As a result, the better
colloidal stability of the polymerization system was
achieved. Furthermore, the theoretical height of the
interparticle interaction potential energy barrier against
occulation correlated quite well with the observed
colloidal stability during polymerization.
2.7. Polymeric surfactant and protective colloid
stabilized polymerization systems
Just like low molecular weight surfactants (e.g. SDS
and NP-40), polymeric surfactants form monomer-
swollen micelles (or aggregates) in the continuous
aqueous phase that are available for the subsequent
particle nucleation in emulsion polymerization. The
hydrophilichydrophobic nature of polymeric surfac-
tants has a signicant inuence on the properties of
micelles. In contrast to small surfactants, polymeric
surfactant molecules exhibit rather restricted mobility
in the heterogeneous polymerization system and this
characteristic feature may have an impact on the
polymerization mechanisms and kinetics. In general,
latex products stabilized by polymeric surfactants show
better water resistance than those stabilized by
conventional surfactants. Among various polymeric
stabilizers used in industry, amphiphilic block or graft
copolymers are the most effective in stabilizing
emulsion polymerization. Nevertheless, some water-
soluble polymers such as polyvinyl alcohol and
hydroxyethyl cellulose are commonly used as the
protective colloid in stabilizing the vinyl acetate-
based emulsion polymerizations. The grafting reaction
of the hydrophobic polyvinyl acetate chains onto the
water-soluble polymer chain results in an amphiphilic
graft copolymer chain in situ, which can be used to
stabilize particle nuclei generated in the continuous
aqueous phase. However, the contribution of adsorption
of water-soluble polymer molecules on the particle
surface to the colloidal stability during polymerization
cannot be ruled out. These vinyl acetate-based latex
products nd applications in laminating adhesives and
interior architecture coatings.
Riess [136] used polystyrenepolyethylene oxide
di- and triblock copolymers to stabilize the emulsion
(co)polymerizations of styrene and styrene and butyl
acrylate. Experimental results showed that the
efciency of these polymeric surfactants decreased
with increasing the molecular weight or polystyrene
content of block copolymers. Acrylic latex products
with particle sizes smaller than 50 nm were achieved
by using polystyrenepolyethylene oxide and poly-
methyl methacrylatepolyethylene oxide diblock
copolymers as the stabilizing species. Crosslinking
of these emulsion polymers then led to the formation
of microgels that were dispersible in organic
solvents. The anionic polymethyl methacrylate
polyacrylic acid diblock copolymers were also used
to prepare the negatively charged hairy latex
particles at high pH.
Kislenko [137] studied the graft polymerization of
methyl acrylate onto water-soluble polymers, methyl
cellulose and hydroxyethyl cellulose. The graft copo-
lymer formed during the rst stage of polymerization
and then acted as a stabilizer for latex particles, which
were the primary reaction loci. Nephelometric exper-
iments and adsorption titration of the resultant polymer
dispersions showed that the saturated adsorption of
particles changed with the progress of polymerization
and its nal value was quite large for the graft
copolymers derived from the cellulose derivatives.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 464
Hwu and Lee [138] studied emulsion polymerization
stabilized by the alkali-soluble resin, poly(styrene-co-
acrylic acid). It was shown that alkali-soluble resin
(25 wt% based on monomer) reduced the average
number of free radicals per particle and, thereby, led to
the decreased rate of polymerization of methyl
methacrylate compared to the counterpart with 3 wt%
SDS as the surfactant. The rate of emulsion polymer-
ization of butyl acrylate stabilized by the alkali-soluble
resin was even slower and the latex product was
unstable. On the other hand, the alkali-soluble resin
(25 wt%) did not retard the rate of polymerization of
styrene compared to the counterpart with 3 wt% SDS as
the surfactant. GPC measurements provided supporting
evidence of the grafting reaction that occurred in the
presence of the alkali-soluble resin. Kato et al. [139]
synthesized and characterized a series of surface-active
polyelectrolytes, poly(methyl methacrylate-co-
methacrylic acid). The effects of the copolymer
molecular weight and composition on the emulsion
polymerization of styrene initiated by KPS were
investigated. The rate of polymerization and nal
number of latex particles per unit volume of water
were only slightly dependent on the copolymer
molecular weight, and they decreased monotonously
with increasing the methacrylic acid content in the
copolymer. The maximal rate of polymerization and
nal number of particles per unit volume of water
occurred when the copolymer molecular weight was in
the range of 500010,000 g mol
K1
. This type of
emulsion polymers has been widely used in the
marketplace of printing inks and varnishes.
Rimmer and Tattersall [140] studied the emulsion
polymerization of the extremely hydrophobic dodecyl
methacrylate or octadecyl methacrylate in the presence
of b cyclodextrin. Dowfax 2A1 and KPS were used as
the surfactant and initiator, respectively. Although the
levels of coagulum were quite high, stable latex
products were produced. Emulsion polymerization is
generally not applicable to the extremely hydrophobic
monomers such as dodecyl methacrylate or octadecyl
methacrylate. On the other hand, miniemulsion
polymerization is an appropriate process for polymer-
izing monomer emulsion comprising dodecyl metha-
crylate or octadecyl methacrylate. It is interesting to
study the role of b cyclodextrin in the nucleation and
growth of particle nuclei and transport of the extremely
hydrophobic monomer molecules from monomer
droplets to the growing latex particles.
Chang and Lee [141] prepared tetracarboxylic acid
terminated sulfopolyester and used these polymeric
surfactants to stabilize the emulsion polymerization of
butyl methacrylate. KPS was used as the initiator. It
was shown that latex particles were stabilized by those
polymeric surfactants with shorter hydrophobic chain
lengths via the depletion stabilization mechanism. The
resultant latices exhibited excellent mechanical stab-
ility though high levels of coagulum were produced
during polymerization. By contrast, particles were
adequately stabilized by those polymeric surfactants
with longer hydrophobic chain lengths via the
electrostatic stabilization mechanism. In this case,
satisfactory colloidal stability was experienced in the
course of polymerization, but latex products were
sensitive to the mechanical agitation. It was postulated
that primary particles were produced via the occula-
tion of precursor particles, polymeric micelles and
existing particles. Later, Chang et al. [142] synthesized
a novel polyester surfactant (5-sulfoisophthalic acid
dimethyl ester sodium salt-modied tetracarboxylic
acid-terminated polyester) and used this polymeric
surfactant to stabilize the emulsion polymerization of
styrene. A continuous particle nucleation mechanism in
the absence of SmithEwart interval II was proposed to
explain the observed polymerization mechanisms and
kinetics. This was attributed to the very high levels of
polymeric surfactant (1030 wt%) and, hence, a vast
number of polymeric micelles (ca. 10 nm in diameter)
per unit volume of water available for particle
nucleation. As a result, a broad particle size distribution
of the latex product was obtained.
Chern and Lee [143] synthesized and characterized
an amphiphilic graft copolymer comprising monomeric
units of the polyethylene oxide containing macromo-
nomer, stearyl methacrylate and 2-hydroxyethyl metha-
crylate. They evaluated the effectiveness of this graft
copolymer in stabilizing the emulsion polymerization
of styrene initiated by NaPS. The total amount of
coagulum was quite high even for the polymerization
with the highest level of the graft copolymer (30 times
the CMC). This was attributed to the retarded molecular
diffusion of the graft copolymer from the monomer
droplet surface to the expanding latex particle surface.
Thus, the particles with a relatively low surface
polyethylene oxide density tended to coagulate with
one another. The same authors then studied the effect of
the structure of graft copolymers on the emulsion
polymerization of styrene [144]. It was shown that the
rate of polymerization increased with increasing the
graft copolymer concentration, initiator concentration,
or temperature. At a constant graft copolymer concen-
tration, the rate of polymerization increased, and the
amount of coagulum decreased with an increase in the
hydrophilicity of graft copolymers. The polymerization
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 465
system did not follow SmithEwart case 2 kinetics. The
desorption of free radicals out of the particles played an
important role in the polymerization kinetics. The
overall activation energy and the activation energy for
the free radical desorption process were determined to
be 85.4 and 34.3 kJ mol
K1
, respectively.
Burguiere et al. [145] prepared and characterized a
series of copolymers comprising the hydrophobic
polystyrene and hydrophilic polyacrylic acid blocks,
and the effectiveness of these copolymers in stabilizing
the emulsion polymerization of styrene evaluated. It
was shown that the block copolymers with high acid
content performed just like small surfactant species. It
is therefore very difcult to predict the number of latex
particles nucleated per unit volume of water. For those
copolymers with low acid content, on the other hand, all
the monomer-swollen polymeric micelles were directly
converted into latex particles during the subsequent
emulsion polymerization. It is noteworthy that this
approach provides an effective control over the particle
nucleation process and it is somewhat similar to the
ideal miniemulsion polymerization that converts all the
stable miniemulsion droplets into latex particles. In this
case, the more hydrophobic copolymers may act as both
the polymeric surfactant and costabilizer in the
stabilization of monomer-swollen polymeric micelles
(or aggregates).
Beal and Chevalier [146] synthesized a series of
amphiphilic copolymers comprising the hydrophobic
polybutyl methacrylate block and the hydrophilic
polysodium methacrylate block and evaluated the
feasibility of using these block copolymers to stabilize
the emulsion polymerization of butyl methacrylate. The
number of latex particles per unit volume of water was
proportional to the graft copolymer concentration to the
a power. The values of a increased from 0.44 to 0.73
when the hydrophilic content of the copolymer
decreased. This was correlated to the competition
between homogeneous nucleation and micelle nuclea-
tion. It was proposed that micelle nucleation was not
applicable to the current polymerization system
because the exchange of block copolymer molecules
between polymeric micelles and latex particles was too
slow. The nal number of latex particles was smaller
than that of the initial polymeric micelles per unit
volume of water was attributed to the limited
occulation of particle nuclei during the particle
nucleation period and/or the facile transfer of monomer
from polymeric micelles to the growing particles
caused by the Ostwald ripening effect. In contrast
with the small surfactant such as SDS, block
copolymers strongly adsorbed on the particle surface
and, consequently, acted as an effective stabilizer
in emulsion polymerization, even at low concentration
(!10
K1
mM) and surface coverage (!10% of the oil
water interfacial area).
Water-soluble polyethylene oxide resembles the
hydrophilic component of many families of commer-
cially available non-ionic surfactants such as the
aforementioned Triton X-405 and NP-40 and it can
be chemically incorporated onto the latex particle
surface by grafting the hydrophobic polymeric radical
onto the polyethylene oxide chain via the hydrogen
abstraction mechanism [147,148]. Thus, the surface-
active graft copolymer formed in situ is just like a non-
ionic polymeric surfactant, and it can help nucleate and
stabilize primary particles in emulsion polymerization.
Chern et al. [149] studied the emulsion copolymeriza-
tion of methyl methacrylate and butyl acrylate in the
presence of polyethylene oxide with different molecu-
lar weights (M
n
Z1500, 4600 and 8000 g mol
K1
). The
initiator used in this work was NaPS. Experimental data
showed that the colloidal stability of acrylic latices
stabilized by polyethylene oxide was primarily con-
trolled by the bridging occulation process. The nal
latex particle size increased with increasing the
concentration of initiator, polyethylene oxide or sodium
chloride. The amount of total scrap produced during
polymerization increased rapidly with increasing the
concentration of sodium chloride due to the ionic
strength effect. The resultant particle size decreased
rapidly with an increase in the agitation speed, but the
amount of total scrap was generally larger for the
polymerization operated at a higher agitation speed.
These results suggested that the fraction of the particle
surface covered by polyethylene oxide and the ratio of
the thickness of the adsorbed polyethylene oxide layer
to that of the electric double layer of the particles
should play an important role in determining the
particle size and colloidal stability.
Another approach to prepare such hairy latex
particles is to polymerize monomer emulsion in the
presence of the polyethylene oxide macromonomer.
The carboncarbon double bond at the end of
macromonomer increases the probability of chemically
incorporating the polyethylene oxide chains onto the
particle surface. Liu et al. [150] prepared polystyrene
nanoparticles stabilized by the polyethylene oxide
macromonomer with an average of 9 or 23 monomeric
units of ethylene oxide. It was observed that the latex
particle size was greatly reduced immediately after the
incorporation of polyethylene oxide chains onto the
particle surface layer due to the steric stabilization
mechanism provided by polyethylene oxide. In
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 466
the presence of the cationic comonomer, 2-methacry-
loyloxyethyl trimethylammonium chloride, the resul-
tant particle size was further reduced down to 60 nm.
Increasing the level of comonomer led to the decreased
particle size. This was attributed to the enhanced
colloidal stability caused by the electrostatic stabiliz-
ation effect of the monomeric units of 2-methacryloy-
loxyethyl trimethylammonium chloride on the particle
surface. On the other hand, increasing the level of
macromonomer resulted in slightly larger particles, but
the agglomeration of primary particles was greatly
retarded. It was postulated that the particle formation
process involved the nucleation, agglomeration and
capture of primary particles generated in the continuous
aqueous phase by the existing particles. For those
who are interested in this research area, please refer to
Ref. [34].
Okaya et al. [151] studied the effects of different
initiators on the emulsion polymerization of methyl
methacrylate in the presence of polyvinyl alcohol
(degree of polymerizationZ580 and degree of hydro-
lysisZ88%). The resultant polymer particles (ca.
90 nm in diameter and 2!10
13
cm
K3
in number)
were fractionated into three parts (i.e. acetone-soluble,
water-soluble and insoluble in both solvents). The
amount of insoluble part was the largest for the
polymerization initiated by ammonium persulfate,
whereas it was the smallest for the polymerization
initiated by AIBN. The high efciency of grafting
polymethyl methacrylate radicals onto polyvinyl alco-
hol chains was observed during the early stage of
polymerization, as evidenced by the absence of
acetone-soluble part. These experimental data together
with other supplemental experiments indicated the
important role of the grafting reaction in the emulsion
polymerization system investigated. Budhlall et al.
[152] studied the role of the grafting reaction in the
nucleation of latex particles during the emulsion
polymerization of vinyl acetate stabilized by the
partially hydrolyzed polyvinyl alcohol with different
degrees of blockiness (degree of polymerization
w1750) and initiated by KPS. The polymerization
system followed the SmithEwart case 1 kinetics (n!
0.5) and did not exhibit a constant polymerization
rate period (SmithEwart interval II). In contrast to
conventional emulsion polymerization, the rate of
polymerization was not linearly proportional to the
number of particles per particle. The number of
particles per unit volume of water observed at the
maximal rate of polymerization in decreasing order was
N
p
(medium blockiness)ON
p
(high blockiness)O
N
p
(low blockiness), whereas the nal number of
particles per unit volume of water was independent of
the degree of blockiness of polyvinyl alcohol. The
resultant particle size distribution was bimodal. It was
postulated that the particle nucleation period was quite
long and the intensive limited occulation of particles
took place during polymerization. The amount of the
grafted polyvinyl acetate increased with the progress of
polymerization and the grafting reaction was nearly
complete by the time when all the monomer droplets
had disappeared (at ca. 25% monomer conversion). The
nal values of the grafted polyvinyl acetate (ca. 40%)
did not vary with the degree of blockiness of polyvinyl
alcohol. It was proposed that the grafting reaction
occurred primarily in the continuous aqueous phase.
The grafted polyvinyl acetate chains continued to grow
in the aqueous phase until they became water-insoluble
and precipitated out of the aqueous phase to form
particle nuclei.
Carra et al. [153] studied the effect of different
polyvinyl alcohols on the emulsion polymerization of
vinyl acetate. They adopted the selective solublization
procedure to obtain three fractions of polyvinyl alcohol
in latex products, free polyvinyl alcohol in the aqueous
phase and physically adsorbed and chemically grafted
polyvinyl alcohols on the particle surface. It was
concluded that the blockiness of polyvinyl alcohol did
not have a signicant inuence on the polymerization
kinetics. The fraction of the grafted polyvinyl alcohol
remained unchanged with the different polyvinyl
alcohols, whereas the fraction of the physically
adsorbed polyvinyl alcohol increased with increasing
the blockiness of polyvinyl alcohol. Furthermore, the
fraction of the grafted polyvinyl alcohol was lower than
that of the physically adsorbed polyvinyl alcohol.
Cheong et al. [154] synthesized and characterized
two water-soluble polyurethane resins comprising
isophorone diisocyanate, poly(1,2-propylene glycol)
(PPG-750 and PPG-2000) and 2,2-bis(hydroxymethyl)-
propionic acid (PUR-750: M
n
Z5600, M
w
Z11,500,
acid numberZ49.1, CMCZ4.2!10
K7
M and PUR-
2000: M
n
Z7100, M
w
Z14,900, acid numberZ30.9,
CMCZ1.1!10
K7
M), where M
n
and M
w
represent the
number-average and weight-average molecular
weights, respectively, and CMC is the critical micelle
concentration. These polymers formed micelles (or
aggregates) in water and the feasibility of using these
amphiphilic polymers in stabilizing the emulsion
polymerization of styrene was evaluated. The rate of
polymerization and nal number of latex particles per
unit volume of water were proportional to the
polyurethane concentration to the 0.250.30 and
0.600.70 powers, respectively. The polystyrene
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 467
polyurethane hybrid particles had very small particle
sizes (ca. 4060 nm in diameter) and broad particle size
distributions. Not like conventional surfactants, con-
tinuous particle nucleation was observed up to a
monomer conversion of 7080%. The characteristic
behaviors of particle nucleation and growth processes
were attributed primarily to the superior solubilization
capability and high internal viscosity of polyurethane
aggregates. Zhang and Zhang [155] prepared poly-
methyl methacrylatepolyurethane hybrid particles via
the emulsion polymerization of methyl methacrylate in
the presence of polyurethane seed particles. The
efciency of grafting polymethyl methacrylate radicals
onto the polyurethane seed particles was higher for the
polymerization initiated by g-ray than that of the
polymerization initiated by KPS.
2.8. Mathematical modeling studies
In the conventional SmithEwart interval II, the
growth of latex particles is primarily controlled by the
concentration of monomer in polymer particles and
average number of free radicals per particle (Eq. (1)).
Morton et al. [156] proposed the following equation for
calculating the equilibrium concentration of monomer
in the particle
2F
m
s=(RTr
p
)
ZK[ln(1KF
p
) C(1K1=X
n
)F
p
CcF
2
p
] (18)
where F
m
and F
p
are the volume fractions of monomer
and polymer, respectively, in the particle. s is the oil
water interfacial tension, R the gas constant, T the
absolute temperature, r
p
the particle radius, and c the
FloryHuggins interaction parameter. Decreasing the
interfacial tension (s), increasing the particle radius (r
p
)
or increasing the temperature (T) results in an increase
in the equilibrium concentration of monomer in the
particle (F
m
). During the SmithEwart interval II,
particle nuclei continue to grow in size at the expense of
monomer droplets. The expanding particle surface area
leads to a reduction in the particle surface coverage by
surfactant and, as a consequence, causes the particle
water interfacial tension to increase during polymer-
ization. The concentration of monomer in the growing
particles may thus remain relatively constant if the
effect of the increased particle size is counterbalanced
by the effect of the increased interfacial tension. The
average number of free radicals per particle is governed
by such kinetic events as the absorption of free radicals
by the particle, bimolecular termination of free radicals
in the particle, desorption of free radicals out of the
particle, bimolecular termination reaction in the
continuous aqueous phase and reabsorption of the
desorbed free radicals by the particle.
Lopez de Arbina et al. [157] carried out the seeded
styrene emulsion polymerizations initiated by a
persulfate initiator to study the growth of latex
particles. The kinetic data were used to evaluate two
mathematical models of different levels of complexity.
The rst model assumed that all the free radicals in the
aqueous phase were exactly the same, whereas the
second model distinguished between the free radicals
generated via the thermal decomposition of initiator
and the desorbed monomeric radicals. No advantage
was obtained from the more complex model based on
computer simulation results. Furthermore, the depen-
dence of the rate coefcient for absorption of free
radicals by the particle on the particle size was
consistent with both the diffusional and colloidal
mechanisms. By contrast, the dependence of the rate
coefcient for desorption of free radicals out of the
particle suggested that the anchoring of free radicals
originating from the persulfate initiator on the particle
surface might have a signicant inuence on the
desorption mechanism.
Mayer et al. [158] pointed out that the polymer-
ization taking place in the newly born particle nuclei
could not adequately described by the classical Smith
Ewart recursion relation. Once nucleated, the particle
nucleus containing only one free radical grew for a long
period of time before the desorption of that radical out
of the particle nucleus or absorption of another free
radical by the particle nucleus followed by the
bimolecular termination reaction occurred. It was
shown that desorption of one free radical out of the
freshly nucleated particle was responsible for the
cessation of the particle growth event. A simple
procedure was then developed to take into account
the contribution of the newly born particle nuclei to the
total rate of polymerization in the micelle nucleation
stage.
Coen et al. [159] developed a comprehensive model
for predicting the number of latex particles produced
per unit volume of water, particle size and particle size
distribution and secondary particle nucleation in
emulsion polymerization. This model took into account
polymer reactions in the continuous aqueous phase,
particle nucleation mechanisms, transport of free
radicals between the aqueous phase and particle
phase, and coagulation of particles. The validity of
this model was veried by experimental data obtained
from the emulsion polymerization of styrene.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 468
The computer simulation also suggested why secondary
particle nucleation took place in the emulsion polymer-
ization stabilized by polymeric surfactant. Ferguson
et al. [160] modied a mechanistic model for predicting
the particle nucleation process in surfactant-free
emulsion polymerization. The objective of this paper
was to identify the appropriate polymerization con-
ditions for the secondary particle nucleation to occur.
The model predicted that nucleation of a secondary
crop of particle nuclei was suppressed by decreasing the
seed particle size, increasing the seed particle concen-
tration and by the monomer-starved-feed condition.
Among these strategies, seed particle size was the most
effective one. The results obtained from this work are
especially useful for designing the polymer particle
morphology. For example, to eliminate secondary
particle nucleation completely is a must in order to
prepare composite particles with a perfect core/shell
structure. Lin et al. [161] observed signicant
secondary particle nucleation during the emulsion
polymerization containing polydimethylsiloxane
seed particles in an attempt to prepare composite
particles with a morphological structure of polydi-
methylsiloxane/poly(methyl methacrylate-co-butyl
acrylate) (core/shell). By contrast, latex products with
bimodal particle size distributions as a consequence of
secondary particle nucleation often offer advantageous
rheological properties, especially for those with very
high total solid contents due to the efcient particle
packing.
Mathematical modeling of emulsion polymerization
is crucial in gaining a fundamental understanding of the
very complicated heterogeneous reaction system. It is
also a valuable tool for the product development,
process assessment and reactor design. Important
features should include (1) nucleation of latex particles,
(2) transport of free radicals between the continuous
aqueous phase and the particle phase (e.g. absorption,
desorption and reabsorption of free radicals), (3)
transport of monomer from monomer droplets to latex
particle, (4) occulation of latex particles, and (5)
dependence of the bimolecular termination and propa-
gation reactions on the monomer conversion.
Sathyagal and McCormick [162] used a population
balance model to study the effect of the time dependent
particle nucleation on the evolving particle size
distribution. This model conrmed that, under the
action of particle aggregation alone, extended and
divided particle nucleation periods in some cases could
result in a very narrow particle size distribution. It also
provided strategies to manipulate the time dependent
particle nucleation to tune the average particle size and
variance of particle size distribution independently.
Herrera-Ordonez and Olayo [163] developed a
mechanistic model for simulating the emulsion
polymerization of styrene. It was shown that homo-
geneous nucleation had an insignicant inuence on the
polymerization kinetics compared to micelle nuclea-
tion. Furthermore, the rate of particle nucleation
reached a maximum with the progress of polymer-
ization due to the contribution of desorbed free radicals.
The effects of the type and chain length of free radicals
on various kinetic events (e.g. particle nucleation and
absorption, desorption and termination of free radicals)
were also investigated. The authors then used the
experimental data of the number of latex particles per
unit volume of water, particle size distribution and
reaction rate of the polymerization of methyl metha-
crylate with a surfactant concentration greater than its
CMC available in the literature to further assess the
proposed model [164]. Instead of the secondary particle
nucleation mechanism, the bimodal particle size
distribution data were attributed to different growth
rates of those particles with different sizes (i.e.
competitive particle growth). The strong gel effect
(i.e. very large number of free radicals per particle)
generally experienced in the conventional emulsion
polymerization of methyl methacrylate was greatly
retarded for the experiment with a low level of
monomer initially present in the reaction system due
to the signicant desorption of free radicals out of small
particles. It is noteworthy that the results obtained from
these computer simulation studies denitely contribute
to the fundamental understanding of the early stage of
the semibatch emulsion polymerization of methyl
methacrylate since only a small fraction of the
monomer feed has been added into the reactor
thereabout.
Kammona et al. [165] made a comprehensive
experimental and theoretical investigation of the non-
ionic surfactant stabilized emulsion copolymerization
of styrene and 2-ethylhexyl acrylate. A mechanistic
model was developed to adequately predict the time
evolution of monomer conversion and latex particle
size distribution. Both the micelle nucleation and
homogeneous nucleation mechanisms were taken into
consideration. Ginsburger et al. [166] developed a
mechanistic model for the emulsion copolymerization
of styrene and butyl acrylate carried out in a batch or
semibatch reactor. This kinetic model was capable of
predicting the polymerization kinetics, particle size and
particle size distribution, polymer molecular weight
and molecular weight distribution and polymer glass
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 469
transition temperature. The validity of this model was
conrmed by the experimental data.
Coen et al. [167] developed a kinetic model that took
into account both the pseudo-bulk (two or more free
radicals per particle) and zeroone (zero or one free
radical per particle) kinetics during the particle
nucleation stage for predicting the number of latex
particles per unit volume of water and rate of
polymerization. This model is especially applicable to
the emulsion polymerization of acrylate ester mono-
mers (e.g. butyl acrylate) that have a very large
propagation rate constant. Under the circumstances,
the polymeric radical can undergo extensive propa-
gation reaction with monomer before it is eventually
terminated upon encountering another free radical since
the concentration of monomer is much higher than that
of free radical in the particle. Thus, two or more free
radicals can coexist in a rather small particle for a
signicant period of time. The experimental data
obtained from the emulsion polymerization of butyl
acrylate were in reasonable agreement with model
predictions.
Dong [168] proposed that the rate of propagation of
oligomeric radicals with monomer molecules at the
particlewater interface was expressed as
Kd[M
+
z
]
I
=dt Zk
pI
A[M]
p
[M
+
z
]
aq
(19)
where [M
+
z
]
I
and [M
+
z
]
aq
were the concentrations of
oligomeric radicals at the particlewater interface and
in the continuous aqueous phase, respectively. [M]
p
is
the concentration of monomer in latex particles, k
pI
the
propagation rate constant for oligomeric radicals at the
particlewater interface that might be estimated by the
transient-state theory, and A the total particlewater
interfacial area. In addition, this approach might be
useful for modeling the absorption of free radicals by
the particles under the monomer-starved condition,
where the propagation reaction was the rate-limiting
step.
3. Polymerization in non-uniform latex particles
3.1. Origin of non-uniform latex particles
Multi-phase polymer particles manufactured by
emulsion polymerization nd many important com-
mercial applications such as elastomers, coatings,
adhesives and impact resistance thermoplastics. Latex
products showing non-uniform particle morphologies
are produced when two or more monomers react with
one another such that separate phases form during
polymerization. The incompatibility of different poly-
mers or the sequence and location of formation of
polymer chains can result in separate polymer phases.
Multi-stage emulsion polymerization has been widely
used to design and prepare latex products with various
particle morphologies. As soon as an appreciable
amount of the post-formed polymer is generated, a
two-phase structure will normally be observed within
the latex particle (Fig. 7). This is attributed to the
relatively incompatible nature of most polymer pairs.
The resultant emulsion polymers with a variety of
particle morphologies generally exhibit different mech-
anical and physicochemical properties. Thus, design
and control of the particle morphology are crucial to
fulll the end-use requirements. It should be noted that
very often contradictory results exist in the early
literature because the particle morphology is a very
complicated function of several physicochemical
parameters (e.g. the incompatibility of polymer pairs,
distribution of free radicals and monomers in the
polymerizing particles, degree of grafting at the
interfacial layer of polymer pairs, polymer molecular
weight, glass transition temperatures of polymer pairs,
etc.) and polymerization conditions (e.g. the recipes,
sequence and location of formation of polymer chains,
temperature, etc.).
Fig. 7. Some possible latex particle morphologies obtained from the
seeded emulsion polymerization of monomer 2 in the presence of seed
particles of polymer 1.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 470
If the seed latex is swollen with the second stage
monomer, which is different from the monomeric units
of seed polymer, phase separation will occur during
polymerization. This will then lead to various morpho-
logical structures for the resultant latex particles
[169171]. By rule of thumb, the greater the
incompatibility of polymer pairs, the greater the extent
of phase separation in the polymerizing particles.
Furthermore, the polarity of polymer is a measure of
the compatibility of polymer pairs. The larger the
difference in the polarity of polymer pairs, the greater
the incompatibility of polymer pairs. The polarity of
some representative polymers in decreasing order is
polyvinyl acetateOpolymethyl methacrylateOpoly-
styreneOpolyvinyl chloride. It should be noted that
even the complete compatibility between two polymers
does not guarantee the formation of a uniform particle
structure because the particle morphology may be
dependent on other factors such as the distribution of
free radicals and monomers in the polymerization
particles, methods of monomer addition, etc.
If the seed particle can hardly be swollen by the second
stage monomer, the subsequent seeded emulsion
polymerization will be localized near the particle surface
layer. As a result, the post-formedpolymer tends to forma
shell around the seed particle (core). An example for this
kind of particle morphology is the seeded emulsion
polymerization of methyl methacrylate with polyvinyli-
dene chloride seed particles. On the other hand, polymer
reactions can take place withinthe seedparticles provided
that the second stage monomer can swell the seed
particles. Under the circumstances, a variety of particle
morphologies (e.g. the perfect core/shell, inverted core/-
shell, dumb-bell shaped and occluded structures) can be
obtained, depending on various physicochemical par-
ameters and polymerization conditions.
The seed particle tends to precipitate inside the nal
latex particle and sometimes even escape from the
particle when the volume ratio of the second stage
monomer to the seed polymer is very high. As another
extreme, very often the post-formed polymer cannot
encapsulate the seed particle completely and the
formation of a perfect core/shell particle eventually
becomes impossible if the volume ratio of the second
stage monomer to the seed polymer is very small.
Polymerization temperature can affect the mobility of
both monomer molecules and polymer chains and,
hence, the rate of separation of polymer pairs, which
will consequently contribute to the determination of the
particle morphology.
Sundberg et al. [172] and Chen et al. [173] used
thermodynamic considerations to minimize the total
Gibbs free energy changes associated with the latex
particle morphologies. This approach shows the role of
the interfacial tension in controlling the particle
morphology and some success in predicting the particle
morphology has been achieved. It should be noted that
such a thermodynamic analysis only predicts the
ultimate particle morphology (when the aging time
approaches innity) and, unfortunately, this is generally
not the case because other physicochemical parameters
and polymerization conditions may also come into play
in determining the particle morphology.
It is noteworthy that continuous stirred tank reactors
with seed latex particles in the feed stream could be an
interesting polymerization system for the particle
morphology studies. The broad residence time distri-
bution associated with such a reactor conguration
results in a broad particle size distribution. By changing
the seed particle size distribution (monodisperse or
polydisperse) and operation conditions (mean residence
time) simultaneously, one can essentially obtain a
number of particle morphologies.
3.2. Morphology development in latex particles
Durant and Sundberg [174] developed an algorithm
to predict the thermodynamic equilibrium morphology
of latex particles as a function of monomer conversion.
It was pointed out that the predicted particle mor-
phology should match that observed experimentally for
the interfacial free energy surface with steep contours
adjacent to the minimum point. On the other hand, there
were a number of possible particle morphologies,
which possessed comparable free energies, in the
polymerization system when the minimal point was
located within a rather at region on the interfacial free
energy surface. Several case studies dealing with
composite particles comprising polystyrene and poly-
methyl methacrylate and two very different surfactants
were carried out. Concurrently, Gonzalez-Ortiz and
Asua [175] developed an interesting mechanistic model
for the migration of clusters during the particle
morphology development in emulsion polymerization.
It was postulated that the driving force for the migration
of clusters was the balance between the van der Waals
force and viscous force. Several equilibrium particle
morphologies (e.g. core/shell, inverted core/shell and
occluded structure) predicted by the dynamic model
were illustrated. Later, the authors considered compo-
site particles to be a biphasic system comprising
clusters of polymer 1 dispersed in a matrix of polymer
2 [176]. Both the polymerization and migration of
clusters were incorporated into the framework of
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 471
model. The polymerization of monomer 1 took place in
both the matrix of polymer 2 and clusters of polymer 1.
Furthermore, polymer 1 produced in the matrix of
polymer 2 diffused readily into clusters of polymer 1.
Migration of clusters was directed toward the equili-
brium particle morphology in order to minimize the
interfacial free energy of the polymerization system.
The driving forces for the migration of clusters
included van der Waals forces between the clusters
and the continuous aqueous phase and those between
the clusters themselves. In addition, the effect of the
viscosity of the matrix of polymer 2 on the migration of
clusters was also taken into consideration. They then
developed another model that took into account the
phase separation to induce the nucleation of clusters,
polymer reactions, diffusion of polymer chains, and
migration of clusters [177]. The model was used to
simulate the emulsion polymerization of methyl
methacrylate in the presence of polystyrene seed
particles. A good agreement between the experimental
results and model predictions was achieved. Further-
more, both the initial volume of clusters and the rate
coefcient for the nucleation of clusters had an
insignicant inuence on the equilibrium particle
morphology.
Stubbs et al. [178] interpreted the effects of various
experimental variables on the nal latex particle
morphology through an assessment of probabilities
for diffusion and reaction of both polymeric radicals
and monomer molecules within the particle. It was
concluded that a uniform monomer concentration
prole across the particle was achieved without any
problem, even during the slow monomer feed or for the
glassy seed particle. By contrast, polymeric radicals
were most likely restricted near the particle surface
layer when the ux of free radicals was high enough and
the rate of monomer feed was slow enough for the
glassy seed particle. Nevertheless, this might not be the
case for the soft seed particle. Thus, the polymerization
in the presence of glassy seed particles was character-
ized by the starved feed process, whereas the starved
feed process was not experienced in the polymer-
ization using soft seed particles. Under the
circumstances, the probability for the formation of
non-equilibrium particle morphologies was higher in
the former case.
Aerdts et al. [179] rst synthesized polybutadiene
seed particles (90 nm in diameter). This was followed
by the semibatch emulsion copolymerization of styrene
and methyl methacrylate in the presence of the seed
particles. Finally, glycidyl methacrylate was grafted
onto composite latex particles. Sodium dihexyl
sulfosuccinate and KPS were used as the surfactant
and initiator, respectively. The functionalized compo-
site particles, served as an impact modier, were
blended with polyamide-6 and these particles were
dispersed well in the matrix of polyamide-6.
Karlsson et al. [180] prepared and characterized a
series of latex products with heterogeneous particle
structures via the seeded emulsion polymerization. The
weight percentage of the hydrophobic seed particles
(poly(styrene-co-butyl acrylate) with glass transition
temperatures ranging from 20 to 100 8C) was kept
constant at 40% throughout this work. The second stage
polymer comprised hydrophilic monomeric units of
methyl methacrylate, butyl acrylate and methacrylic
acid with the glass transition temperature equal to
20 8C. Calculated rates of diffusion of the propagating
species during polymerization were correlated to the
observed morphological evolution and the fraction of
interphase in the composite polymer particles.
Ferguson et al. [181] explored how to prepare latex
particles with a core/shell (polystyrene/polyvinyl
acetate) structure. To minimize the secondary particle
nucleation, polyvinyl acetate was chosen as the seed
particles and the seeded emulsion polymerization of the
second stage monomer styrene was carried out. The
objective of this work was to prepare ideal core/shell
particles via the inversion of the second stage
polystyrene formed during the seeded emulsion
polymerization and the rst-stage polyvinyl acetate
seed particles, as would be expected by thermodynamic
predictions. It was shown that evolution of the target
particle morphology was promoted by fast diffusion of
polyvinyl acetate chains toward the particle surface
layer during the seeded emulsion polymerization of
styrene. A number of methods (e.g. reducing the
molecular weight and degree of branching of polyvinyl
acetate, minimizing the grafting reaction of polystyrene
radicals onto polyvinyl acetate chains and increasing
the hydrophilicity of polyvinyl acetate) to achieve this
goal were proposed. Zhao et al. [182] investigated the
morphological evolution of the emulsion polymer-
ization of butyl acrylate in the presence of polyvinyl
acetate seed particles stabilized by mixed anionic and
non-ionic surfactants and initiated by KPS at 70 8C. The
polymerization system was operated under the mono-
mer-starved condition. The inverted core/shell (poly-
butyl acrylate/polyvinyl acetate) structure was the
thermodynamically preferred morphology. Neverthe-
less, the multi-particle morphology was observed,
which was attributed to the restricted polymer chain
mobility closely related to the high viscosity inside the
particles. It was postulated that polybutyl acrylate rst
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 472
formed around polyvinyl acetate seed particles and then
migrated toward the core of seed particles.
Pan et al. [183] rst prepared crosslinked poly(-
butylacrylate-co-2-ethylhexyl acrylate)/poly(methyl
methacrylate-co-styrene) (core/shell) particles by the
multi-stage emulsion polymerization. The core/shell
seed particles were then used in the emulsion
polymerization of vinyl chloride. The efciency of
grafting vinyl chloride radicals onto the seed particles
increased with increasing the amount of seed particles.
The particle morphology was transformed from the
perfect core/shell structure into irregular sandwich-like
structure when the weight percentage of the monomeric
units of styrene in the shell was greater than 70%.
Furthermore, the composite particles prepared by the
emulsion polymerization of vinyl chloride in the
presence of these seed particles exhibited a three-
layered core/shell structure. The compatibility between
polyvinyl chloride and composite seed particles was
good and seed particles were dispersed uniformly in the
polyvinyl chloride matrix.
4. Semibatch emulsion polymerization
4.1. Polymerization mechanisms and kinetics
Semibatch emulsion polymerization is an import-
ant process for the manufacture of latex products
such as coatings, adhesives and synthetic elastomers.
In addition to its operational exibility for products
with controlled polymer composition and particle
morphology, the semibatch emulsion polymerization
process can easily remove the enormous heat
generated during the reaction. The most striking
difference between the semibatch and batch emulsion
polymerization processes is that reaction ingredients
such as monomer, surfactant, initiator or water can
be added to the semibatch reaction system through-
out the polymerization (Fig. 8). Thus, the residence
time distribution of particle nuclei is broader for
semibatch emulsion polymerization. All these
features make the semibatch emulsion polymer-
ization mechanisms and kinetics more complicated
in comparison with the batch counterpart. The
studies dealing with the effects of reaction variables
on the polymerization mechanisms and kinetics and
colloidal stability are very useful for the product and
process development. The large number of journal
papers dealing with semibatch emulsion polymer-
ization over the preceding 10-year period cited in
this work more or less reects the industrial
importance of this research area.
Gugliotta et al. [184] developed a new approach to
estimate the monomer conversion and copolymer
composition in semibatch emulsion polymerization
based on reaction calorimetric measurements. The
validity of this technique was conrmed by the
semibatch emulsion copolymerizations of both the
styrenebutyl acrylate and vinyl acetatebutyl acrylate
pairs.
Unzueta and Forcada [185] carried out the semi-
batch emulsion copolymerization of methyl methacry-
late and butyl acrylate stabilized by mixed anionic and
non-ionic surfactants (SDS and dodecyl polyethoxy-
late) under the monomer-starved condition. The
polymerization stabilized by the non-ionic surfactant
alone resulted in a slower rate of polymerization and a
larger latex particle size. At low anionic surfactant
concentration, the nal number of particles per unit
volume of water increased with increasing the total
surfactant concentration for the polymerization stabil-
ized by mixed anionic and non-ionic surfactants. On the
other hand, at high anionic surfactant concentration and
a ratio of anionic surfactant to non-ionic surfactant
between 1:1 and 1:3, a smaller population of particles
was produced. During the early stage of polymer-
ization, the particle size distribution was positively
skew for the polymerization containing no surfactant or
low surfactant concentration, narrower and Gaussian-
Semibatch emulsion polymerization
Initial reactor charge
water, surfactant, monomer, initiator
Primary particles
(Particle growth)
(Secondary nucleation)
(Flocculation)
Monomer addition
Monomer, surfactant,
initiator, water
Filterable solids
Reactor scrap
Stable
latex particles
(Nucleation period)
Fig. 8. Flow chart for a typical semibatch emulsion polymerization
process.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 473
like for the polymerization containing intermediate
surfactant concentration, and broader for the polymer-
ization containing high surfactant concentration. Fur-
thermore, latex products with larger particle sizes and
narrower particle size distributions were obtained from
polymerizations stabilized by mixed anionic and non-
ionic surfactants compared to the polymerization
stabilized by anionic surfactant alone.
Chern and Hsu [186] studied the semibatch emulsion
copolymerization of methyl methacrylate and butyl
acrylate stabilized by the mixed surfactants of SDS and
NP-40 and initiated by NaPS. The concentration of
SDS in the initial reactor charge controlled the nal
latex particle size. NP-40 only acted as an auxiliary
stabilizer. Both the initiator and monomer concen-
trations in the initial reactor charge and the ratio of
methyl methacrylate to butyl acrylate showed very little
effects on the nal particle size. In addition, the
resultant particle size increased with increasing the
electrolyte concentration in the initial reactor charge or
agitation speed. Chern et al. [187] focused on the effects
of various reaction variables on the colloidal stability of
acrylic latex particles in the course of polymerization.
SDS was used as the sole surfactant. The amount of
coagulum produced by intensive particle coagulation
was greatly reduced when the level of SDS in the
monomer emulsion feed increased. The increased level
of SDS in the initial reactor charge led to an increase in
the particle volume change due to the limited particle
occulation later in the polymerization process. The
larger the ratio of methyl methacrylate to butyl acrylate
in the copolymer, the greater the amount of coagulum
produced. Both the amount of coagulum and particle
volume change increased with increasing the electro-
lyte concentration. The agitation speed (500800 rpm)
showed an insignicant inuence on the particle
coagulation process. Furthermore, latex particles lost
their colloidal stability rapidly above 40% total solid
content due to the crowding effect.
Chern and Lin [188] illustrated that latex particles
were capable of maintaining appreciable colloidal
stability via the limited particle occulation during
the semibatch surfactant-free emulsion polymerization
of butyl acrylate. However, intensive agitation (e.g.
800 rpm in a 1-dm
3
reactor) resulted in a signicant
amount of coagulum in addition to the limited particle
occulation. At a common reaction time, the particle
size increased with increasing the monomer feed rate.
At any time during polymerization, the particle size rst
increased to a maximum and then decreased with
increasing the NaPS concentration. The polymerization
with the lowest level of initiator exhibited the worst
colloidal stability because of the insufcient supply of
sulfate end-groups to stabilize the growing particles.
Nevertheless, a signicant amount of coagulum was
also observed for the polymerization with the highest
level of initiator due to the ionic strength effect. The
optimal colloidal stability occurred at a point close to
an initiator level of 0.19%. The effects of sodium
bicarbonate (buffer) on the particle size and colloidal
stability were signicant, again, owing to the ionic
strength effect.
To overcome the poor water resistance generally
experienced with waterborne polymers, the level of
surfactant used in emulsion polymerization can be
reduced. However, the colloidal stability of latex
particles can be greatly reduced and a signicant
amount of coagulum can form during the monomer
addition period. Furthermore, the particles can grow in
size by relatively mild particle agglomeration in order
to maintain appreciable colloidal stability, in addition
to polymerizing the imbibed monomer in the particles.
The improved latex stability is attributed to the
decreased particle surface area and, hence, the
increased particle surface charge density associated
with such a limited particle occulation process
[35,36,38]. The limited particle occulation process
makes the task of particle size control more difcult.
This is a critical issue since control of the particle size is
the key to guaranteeing the quality of latex products. To
get around the dilemma between satisfactory product
properties and smooth plant production, a small amount
of functional monomers such as acrylic acid can be
incorporated into the emulsion polymer to greatly
enhance the colloidal stability of latex products. The
ionized carboxyl group that is chemically incorporated
into the emulsion polymer can increase the particle
surface charge density and, therefore, increase the
electric repulsion force among the interactive particles.
Chern and Lin [189] studied the effects of functional
monomers (acrylic acid, methacrylic acid and hydro-
xyethyl methacrylate) on the semibatch emulsion
polymerization of butyl acrylate stabilized by the
mixed surfactants of SDS and NP-40 and initiated by
NaPS. Experimental results showed that the concen-
tration of SDS in the initial reactor charge was the most
important parameter in controlling the latex particle
size during polymerization. The nal particle size
decreased with increasing the concentration of SDS,
NP-40 or functional monomer. Among the functional
monomers investigated, acrylic acid was the most
efcient one in nucleating and then stabilizing the
particles. The particle size rst decreased to a minimum
and then increased with increasing the concentration of
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 474
sodium bicarbonate (a neutralizing agent). The optimal
sodium bicarbonate concentration for achieving the
smallest particle size occurred at a point close to 0.15
0.29%. Furthermore, the aqueous phase polymerization
played an important role during the particle nucleation
period. The colloidal stability of the particles during the
semibatch emulsion polymerization of butyl acrylate in
the presence of 010% acrylic acid was also investi-
gated [190]. Incorporation of only 5% acrylic acid into
the emulsion polymer greatly improved the latex
stability. The amount of coagulum was greatly reduced
by increasing the level of SDS in the monomer
emulsion feed. On the other hand, increasing the
concentration of SDS in the initial reactor charge
resulted in an increase in the percentage of the particle
volume change. Both the coagulation and secondary
particle nucleation processes caused a signicant
deviation from the Novak model [191].
In an attempt to improve the colloidal stability and,
therefore, gain a better control of the latex particle size,
a small quantity of acrylic acid or methacrylic acid was
incorporated into the emulsion polymer in the
semibatch surfactant-free emulsion polymerization of
butyl acrylate [192]. Indeed, the limited particle
occulation phenomenon was depressed signicantly
for the polymerization in the presence of carboxylic
monomers, thereby leading to the decreased particle
size. The nal latex particle size rst increased to a
maximum around 2% acrylic and then decreased with
increasing the concentration of acrylic acid. This was
attributed to the polyelectrolyte effect in the particle
nucleation period. In addition, the particle size was
independent of the type of carboxylic monomers. As
expected, a signicant fraction of the monomeric units
of acrylic acid was present near the particle surface
layer. On the other hand, the more hydrophobic
monomeric units of methacrylic acid were distributed
more uniformly in the particles.
Alternatively, conventional surfactants used to
nucleate and stabilize latex particles can be replaced
by a reactive surfactant (or polymerizable surfactant).
In addition to the basic surfactant properties (e.g.
lowering of surface tension and formation of micelles
in water), the reactive surfactant molecules can be
chemically incorporated onto the particle surface
during polymerization. Thus, the water sensitivity of
latex products arising from the immobilized surfactant
molecules near the particle surface layer can be
minimized. Chern and Chen [193] carried out the
semibatch emulsion polymerization of butyl acrylate
stabilized by sodium dodecyl allyl sulfosuccinate and
initiated by NaPS. It was shown that the reactive
sodium dodecyl allyl sulfosuccinate played a similar
role in the particle nucleation and growth stages to the
conventional SDS. The nal number of latex particles
per unit volume of water was proportional to the
concentration of sodium dodecyl allyl sulfosuccinate in
the initial reactor charge (the most important parameter
with regard to particle nucleation) to the 0.720.80
power. The saturated particle surface area occupied by
one molecule of sodium dodecyl allyl sulfosuccinate
was determined to be 0.36 nm
2
for the polybutyl
acrylate particles prepared by the surfactant-free
emulsion polymerization. Furthermore, the value of
the saturated particle surface area increased with
increasing the particle surface polarity for the acrylic
latices stabilized by the reactive surfactant. It was also
demonstrated that the ionic strength had a signicant
impact on the colloidal stability during polymerization
[194]. Both the amount of coagulum and percentage of
the particle volume change increased with increasing
the electrolyte concentration. In addition, the particle
volume change increased signicantly with an increase
in the concentration of sodium dodecyl allyl sulfosuc-
cinate in the initial reactor charge, and the amount of
coagulum increased rapidly with increasing the
agitation speed from 400 to 800 rpm. Finally, the
kinetic studies indicated that an induction period or
even a complete inhibition of the polymerization was
observed for the experiments with relatively high
sodium dodecyl allyl sulfosuccinate concentration or
low NaPS concentration [195]. This was attributed to
the intensive chain transfer of free radicals to the
reactive surfactant. The rate of polymerization was
primarily controlled by the monomer feed rate. The rate
of polymerization increased with increasing the
initiator concentration, whereas it was relatively
insensitive to changes in the reactive surfactant
concentration.
Xu and Chen [196] prepared two polymerizable
surfactants, sodium 4-(u-acryloyloxyalkyl)oxy ben-
zene sulfonate with the alkyl chain length equal to 8
or 10, and used them to stabilize the semibatch
emulsion copolymerization of butyl methacrylate.
A redox initiator system of ammonium persulfate and
tetramethylethylenediamine was used to start the
polymerization at room temperature. The latex particle
size increased continuously and the number of particles
per unit volume of water remained relatively constant
with the progress of polymerization. This was attributed
to the predominant micelle nucleation mechanism.
X-ray photoelectron spectroscopy data showed that
polymerizable surfactant molecules were preferably
located near the particle surface layer.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 475
The seeded technique (i.e. the number of latex
particles per unit volume of water in the reaction system
is known immediately before the start of monomer
feed) has been widely used to study the particle growth
mechanisms in the semibatch emulsion polymerization
in order to avoid the complicated effects by particle
nucleation. Furthermore, from the industrial point of
view, the semibatch emulsion polymerization in the
presence of a constant seed particle concentration
results in better quality control of latex products.
Chern et al. [197] studied the particle growth
mechanisms involved in the semibatch surfactant-free
emulsion polymerization of butyl acrylate in the
presence of polybutyl acrylate seed particles. Experi-
mental data showed that the limited particle occula-
tion, often observed in the polymerization containing
no seed particles, did not take place in this work. This
was attributed to the extensive formation of coagulum
during polymerization. Agitation speed was the most
important parameter in determining the level of total
scrap produced, followed by the initiator concentration,
monomer feed rate and then buffer concentration.
Furthermore, the level of coagulum for the polymer-
izations using carboxylic seed particles was much
lower than that for the polymerizations using ordinary
polybutyl acrylate seed particles. Again, limited
particle occulation was not observed during polymer-
ization. In this case, nucleation of a second crop of
primary particles during the monomer addition period
was conrmed. Chern et al. [198] prepared and
characterized a series of highly carboxylated seed
latices comprising methacrylic acid or acrylic acid,
dodecyl methacrylate and methyl methacrylate. Dode-
cyl mercaptan was used to regulate the polymer
molecular weight. These neutralized latices were then
used as the seed particles for the subsequent semibatch
surfactant-free emulsion polymerization of butyl acryl-
ate. The seed latices containing 50 wt% methacrylic
acid or acrylic acid at pH 89 were very effective for
this purpose due to the increased particle surface charge
density.
Sajjadi and Brooks [199] showed that the concen-
tration of monomer in the initial reactor charge had a
signicant inuence on the nal number of latex
particles per unit volume of water in the semibatch
emulsion polymerization of butyl acrylate. The
polymerization system operated under the monomer-
starved condition promoted the rate of particle
nucleation due to the retarded particle growth. As a
result, more particles were nucleated for the polymer-
ization with a concentration of monomer in the initial
reactor charge below its critical value. The largest
population of particles was obtained from the polymer-
ization without any monomer in the initial reactor
charge. In addition, the steady state rate of polymer-
ization was only slightly dependent on the number of
particles per unit volume of water. They also
demonstrated that bimodal particle size distribution
could be obtained from the semibatch emulsion
polymerization of butyl acrylate using the monomer
emulsion feed with a constant rate and no surfactant in
the initial reactor charge [200]. It was shown that
variations in the monomer emulsion feed rate,
concentrations of monomer, surfactant and initiator
and their distribution between the initial reactor charge
and the feed stream had signicant effects on the
resultant particle size distribution. Furthermore, the rate
of secondary particle nucleation was inversely pro-
portional to the primary particle nucleation. The
particle size distribution of latex products became
broader when the monomer feed rate decreased. The
aqueous phase polymerization played an important role
in the secondary particle nucleation process. The
authors then studied the semibatch emulsion polymer-
ization of butyl acrylate with the neat monomer feed
[201]. SDS and KPS were used as the surfactant and
initiator, respectively. The parameters investigated
included the monomer feed rate, concentrations of
surfactant and initiator, temperature and pre-period
time. At steady state, the polymerization system
followed the Wessling model [202]. It was shown that
the time required to reach the near steady state
increased with increasing the monomer feed rate for
the polymerization with the particles not swollen with
monomer, whereas it decreased with increasing the
monomer feed rate for the polymerization with the
monomer-swollen particles. Nevertheless, the time
required to reach the real steady state was always
minimal at the lowest monomer feed rate regardless of
the seed particle compositions.
Latex products with high solid contents and
acceptable rheological properties are desirable in
many industrial applications such as adhesives, coat-
ings and caulks and sealants. Such unique colloidal
systems are generally manufactured by the semibatch
emulsion polymerization processes and characterized
by the extremely efcient packing of the polymer
particles with broad particle size distributions. Thus,
how to manipulate the complex reaction variables to
precisely control the particle size distribution is crucial
for the product development program.
Chern et al. [203] demonstrated that injecting SDS
into the reaction medium induced a second crop of tiny
primary particles in the semibatch seeded emulsion
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 476
polymerization of acrylic monomers. As a result, latex
products with bimodal particle size distributions were
obtained. The concentration of seed particles was the
most important parameter that controlled the particle
size distribution, followed by the time when the
surfactant was injected into the reactor. Retarded
secondary particle nucleation during the monomer
addition period was achieved when the concentration
of seed particles or the time when the surfactant was
injected into the reactor increased. The total seed
particle surface area was shown to be a very useful
parameter for manipulating the resultant particle size
and particle size distribution. Schneider et al. [204]
proposed two methods for preparing high solid content
(over 65%) latex products with bi- and trimodal particle
size distributions in the semibatch emulsion copoly-
merization of methyl methacrylate, butyl acrylate and a
small amount of acrylic acid. The key parameter in
determining the colloidal stability of the latex particles,
formation of coagulum and rheology was the particle
size distribution that could be modied effectively by
the secondary particle nucleation. Nucleation of the
third crop of particles for the latex product with a
trimodal particle size distribution was best accom-
plished with mixed anionic and non-ionic surfactants
since anionic surfactant alone generally resulted in
detrimental changes in the particle size distribution as a
consequence of excessive particle occulation. Boutti
et al. [205] prepared high solid content (over 70%) latex
products without resort to the use of intermediate seed
particles in the semibatch emulsion copolymerization
of acrylic monomers. An electrostatically neutral
initiator hydrogen peroxide and relatively low levels
of mixed anionic and non-ionic surfactants were used to
start the polymerization in the initial stage in order to
avoid stabilizing small particle nuclei. This was
followed by the use of a persulfate initiator in the
second half of the polymerization process. In this
manner, a desirable population of small particle nuclei
was generated and adequately stabilized.
Sebenik and Krajnc [206] studied the effect of the
soft segment (polyester polyol) chain length on the
reaction kinetics of the semibatch emulsion copoly-
merization of methyl methacrylate, butyl acrylate and
a small amount of acrylic acid in the presence of
polyurethane seed particles. The waterborne poly-
urethane seed particles were prepared by the conden-
sation reactions of isophorone diisocyanate (M
n
Z1000
or 2000), polyneopentyladipate and dimethylol pro-
pionic acid. The rigidity of polyurethane was
controlled by varying the soft segment chain length.
The weight ratio of acrylic monomers to polyurethane
was kept constant at 1:1. It was shown that the
polyurethane seed particles containing the higher
molecular weight polyol were swollen with acrylic
monomers to a greater extent. In the pseudo-steady
state, the number of latex particles per unit volume of
water remained relatively constant and it was
comparable to the number of polyurethane seed
particles present in the initial reactor charge. The
polymerization system showed intermediate behavior
between SmithEwart cases 1 and 2.
4.2. Mathematical modeling studies
Chern [207] developed a mechanistic model based
on the diffusion-controlled reaction mechanisms to
predict the kinetics of the semibatch emulsion
polymerization of styrene. Reasonable agreement
between the model predictions and experimental data
available in the literature was achieved. Computer
simulation results showed that the polymerization
system approached SmithEwart case 2 kinetics (nZ
0.5) when the concentration of monomer in the latex
particles was close to the saturation value. By contrast,
the polymerization system under the monomer-starved
condition was characterized by the diffusion-controlled
reaction mechanisms (nO0.5). The author also
developed a model to predict the effect of desorption
of free radicals out of the particles on the kinetics of the
semibatch emulsion polymerization of methyl acrylate
[208]. The validity of the kinetic model was conrmed
by the experimental data for a wide range of
monomer feed rates. The desorption rate constant
for methyl acrylate at 50 8C was determined to be 4!
10
K12
cm
2
s
K1
.
Wang et al. [209] used the Langmuir site adsorption
model in combination with surface tension measure-
ments to control the level of surfactant (sodium
dodecylbenzene sulfonate) fed to the semibatch seeded
emulsion copolymerization of methyl methacrylate and
butyl acrylate with the goal to avoid the secondary
particle nucleation or coagulation. The Langmuir site
adsorption model was developed to relate the surface
tension to surfactant concentration in the polymer-
ization system from rst principles. It was used to
predict the partitioning of the added surfactant and
mobile in situ surfactant between the continuous
aqueous phase and the particlewater interface in the
presence of anchored sulfate end-groups originating
from the persulfate initiator. When the surface tension
was maintained at 4557 dyn cm
K1
, monodisperse
latex products with average particle diameters of
0.53 mm were obtained.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 477
Chern and Kuo [210] developed a kinetic model to
simulate the shear-induced particle coagulation process
during the semibatch seeded emulsion polymerization
of acrylic monomers. DLVO theory was used to
calculate the total potential energy barrier against the
coagulation of latex particles. The coagulation rate
constant was postulated to be proportional to an
exponential function of the relatively mean free path
length between two particles. Agreement between the
model predictions and experimental data was good. The
model was then used to study the effects of important
reaction variables such as the total solid content, seed
particle size and concentration, initiator concentration
and surfactant feed prole on the colloidal stability of
the particles nucleated in a semibatch reactor.
Based on the coagulative particle nucleation
mechanism, a two-step model was developed for the
semibatch surfactant-free emulsion polymerization of
butyl acrylate in the absence or presence of a small
amount of acrylic acid [211]. During Stage 1, precursor
particles formed by phase separation of oligomeric
radicals in the continuous aqueous phase, followed by
the limited occulation of unstable precursor particles
to form stable primary particles. During the early part
of Stage 2, the rate of limited occulation of precursor
particles to form primary particles was slower than the
rate of coagulation of primary particles. This would
result in the decreased particle concentration with time.
Later on, the particle concentration started to level off
and nally reached a steady value. The model predicted
the experimental data of the particle concentration and
particle size with the progress of polymerization
reasonably well.
Sajjadi [212] developed two mechanistic models for
the particle nucleation involved in the semibatch
emulsion polymerization of styrene under the mono-
mer-starved condition. The rst model involving some
simplications resulted in analytical solutions. Smith
Ewart theory was extended to take into account the
particle nucleation under the monomer-starved con-
dition. The number of latex particles per unit volume of
water was proportional to the surfactant concentration,
the rate of initiator decomposition and the rate of
monomer addition, respectively, to the 1.0, 2/3 and
K2/3 powers. The second model considered the
aqueous phase polymerization kinetics and its effect
on the efciency of capture of free radicals by the
monomer-swollen micelles. It successfully predicted
some features of the particle nucleation process.
Immanuel et al. [213] developed a population
balance model for the non-ionic surfactant-stabilized
emulsion copolymerization of vinyl acetate and butyl
acrylate. The initiator package included t-butyl hydro-
gen peroxide and sodium formaldehyde sulfoxylate.
The model took into account the effects of the
nucleation, growth and occulation of latex particles
on the evolution of particle size distribution. Sajjadi and
Yianneskis [214] studied the particle nucleation and
growth mechanisms involved in the semibatch emul-
sion polymerization reactors operated under the
monomer-starved condition. Among the parameters
investigated, the monomer feed rate together with the
level of monomer in the initial reactor charge were the
primary factors that controlled the particle nucleation
process. Reducing the rate of particle growth prolonged
the particle nucleation period and, therefore, slowed
down the rate of micelle depletion. As a result, a larger
population of particles was produced during
polymerization.
5. Summary
Emulsion polymerization involves nucleation and
growth of particle nuclei, followed by consumption of
residual monomer in a heterogeneous reaction system.
The propagation reaction of free radicals with monomer
molecules takes place primarily in the latex particles
and the emulsied monomer droplets only serve as a
reservoir to supply the growing particles with mono-
mer. Adequate colloidal stability of the polymerization
system is generally achieved by using surfactant or
protective colloid. There is no doubt that the relatively
short particle nucleation stage predominates in the
evolution of particle size and particle size distribution
and, therefore, has a signicant effect on the colloidal
and physical properties of latex products. Furthermore,
from the practical point of view, to gain a better
understanding of the particle nucleation mechanisms is
crucial for the product development and scale up and
quality control. This review article was aimed at the
emulsion polymerization mechanisms and kinetics over
the preceding 10-year period.
First, an overview of the general features of
emulsion polymerization was given. The pioneering
studies such as the well-known SmithEwart theory,
homogeneous nucleation mechanism, transport of free
radicals and monomer molecules among different
phases (the continuous aqueous phase, monomer-
swollen micelles, monomer droplets and monomer-
swollen polymer particles) and a variety of chemical
reactions involving these species and the particle
growth mechanisms were discussed briey.
Several techniques useful for studying the polymer-
ization mechanisms and kinetics were then presented.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 478
Some methods involved isolation and characterization
(e.g. FT-IR and
13
C NMR) of water-soluble oligomers
and precursor particles (or particle nuclei) to investigate
the particle nucleation process [6264]. On-line
monitoring techniques such as the optical transmission
in combination with conductivity [65], multi-angle
laser light scattering [67], uorescence spectroscopy
[70] and reaction calorimetry [7274] and other
delicately designed experiments were developed to
study the particle nucleation mechanisms and polymer-
ization kinetics. The end-groups of polymer chains and
their distribution in the particles were investigated by
using MALDI-TOF-MAS [66] and electron spec-
troscopy imaging in an analytical TEM [71], respect-
ively. Experimental results thus obtained were used to
interpret the role of water-soluble initiator radicals in
the particle nucleation and growth processes. Water-
insoluble probe molecules such as dye [68,69] and
pyrene [70] were employed to determine the particle
nucleation loci in emulsion polymerization.
Emulsion polymerizations stabilized by different
species (e.g. sulfate group derived from the persulfate
initiator, surface-active initiators, functional mono-
mers, anionic and non-ionic surfactants, polymerizable
surfactants, degradable surfactants, polymeric surfac-
tants and protective colloids) were studied extensively
in the last few years. Although the polymerization
systems investigated became more complicated than
ever, the basic principles behind a variety of physico-
chemical phenomena still remained unchanged.
Depending on the recipes and reaction conditions, one
or more than one of the particle nucleation mechanisms
(micelle nucleation, homogeneous nucleation and
monomer droplet nucleation) are operative in emulsion
polymerization. It is noteworthy that in some cases the
limited particle occulation process occurring in the
course of polymerization cannot be ignored, especially
for those polymerizations in the absence of surfactant
or with low levels of surfactant. As a result, the
evolution of the latex particle size and particle size
distribution becomes more complicated. At present, the
particle nucleation process is still not well understood.
How to distinguish between the particle nucleation
mechanisms and quantitatively determine the fraction
of latex particles originating from each nucleation
mechanism places a great challenge before the colloid
and polymer chemists.
The number of studies dealing with the growth of
latex particles during polymerization is relatively small
in comparison with particle nucleation. This is probably
due to the fact that the polymerization kinetics (Eq. (1))
is not the only factor that controls the particle growth
process. Beyond the particle nucleation stage, primary
particles may grow in size via the propagation reaction
of free radicals with the imbibed monomer molecules,
limited occulation among the interactive particles and,
perhaps, coalescence between monomer droplets and
polymer particles. Semibatch emulsion polymerization
in the presence of seed particles that eliminates the
particle nucleation stage and with the regulated
concentration of monomer in the reaction system
certainly represents an effective tool to investigate the
particle growth mechanisms. The studies dealing with
the diffusion-controlled nucleation and growth of
particle nuclei [75] and the molecular diffusion of
monomer from the aqueous phase to the particlewater
interface [76] represent two creative examples for
exploring the polymerization mechanisms and kinetics.
Furthermore, the number of publications dealing with
the polymerizable, degradable or polymeric surfactants
and surface-active initiators is quite few and further
research is required to advance our knowledge of
effectively using these specialty surfactants and
initiators to nucleate and stabilize latex particles.
The development of new applications represents
valuable opportunities for the related industries.
Polymerization in non-uniform latex particles is a
potential candidate for offering various application
properties. This article rst discussed the origin of non-
uniform latex particles from both the thermodynamic
and kinetic points of view. This was followed by the
discussion of various reaction parameters that had
signicant effects on the development of particle
morphology. Recent studies on the polymerization in
non-uniform polymer particles were then reviewed. It
was shown that the interfacial tension between polymer
pairs and the particlewater interfacial tension played
an important role in the development of particle
morphology. Furthermore, the predicted particle mor-
phology during polymerization via a thermodynamic
analysis should match that observed experimentally for
the interfacial free energy surface with steep contours
adjacent to the minimal point [174]. On the other hand,
there were a number of possible particle morphologies,
which possessed comparable interfacial free energies,
in the polymerization system when the minimal point
was located within a rather at region on the interfacial
free energy surface. It should be noted that such a
thermodynamic analysis only gives the ultimate
particle morphology (when the aging time approaches
innity). Nevertheless, this is generally not the case
because other physicochemical parameters and
polymerization conditions may also come into play in
determining the particle morphology. This critical issue
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 479
was conrmed experimentally. In addition to the
advancement of theoretical treatment, a number of
papers investigated how to manipulate the reaction
variables to effectively control the development of
particle morphology with the progress of polymer-
ization and many successful case studies were
achieved.
Semibatch emulsion polymerization is an important
process for the manufacture of a variety of latex
products because of its operational exibility for
products with controlled polymer compositions and
particle morphologies and easy removal of the
enormous heat generated during the reaction. This
versatile process is especially useful for studying the
particle growth mechanisms provided that the seed
particles are present initially in the polymerization
system to eliminate the particle nucleation stage. The
polymerization mechanisms, kinetics and colloidal
stability involved in the semibatch emulsion polymer-
ization were reviewed extensively. The major reaction
variables studied included the compositions of mono-
mer and surfactant, concentrations of surfactant,
initiator and monomer in the initial reactor charge,
the feed rates of monomer, surfactant, initiator and
water (if monomer emulsion feed used), and agitation
speed. The continuous or secondary particle nucleation
and limited particle nucleation were shown to have
important effects on the resultant particle size and
particle size distribution. Latex products with high solid
contents (above 70%) and bimodal particle size
distributions and some innovative hybrid emulsion
polymers prepared by semibatch emulsion polymer-
ization were demonstrated.
In order to be in compliance with the health and
environment regulations, minimizing the levels of
residual monomer in latex products within a reasonable
batch cycle time is required for the successful product
development. It is very important to understand the
partitioning of residual monomer among the continuous
aqueous phase, particlewater interface and particle
phase before an effective strategy can be established to
alleviate the residual monomer problem. How to select
the adequate initiator package and reaction temperature
prole during the later stage of polymerization is
probably the key of a successful monomer scavenging
technique. This subject of industrial importance
deserves more fundamental research efforts.
Acknowledgements
The nancial support from National Science Council
of Taiwan, Republic of China is greatly appreciated.
References
[1] Bovey FA, Kolthoff IM, Medalia AI, Meehan EJ. Emulsion
polymerization. New York: Interscience Publishers; 1965.
[2] Blakely DC. Emulsion polymerization. Theory and practice.
London: Applied Science; 1975.
[3] Eliseeva VI, Ivanchev SS, Kuchanov SI, Lebedev AV.
Emulsion polymerization and its applications in industry.
New York: Consultants Bureau; 1981.
[4] Barton J, Capek I. Radical polymerization in disperse systems.
New York: Ellis Horwood; 1994.
[5] Gilbert RG. Emulsion polymerization: a mechanistic approach.
London: Academic Press; 1995.
[6] Fitch RM. Polymer colloids: a comprehensive introduction.
London: Academic Press; 1997.
[7] Verwey EJW, Overbeek JThG. Theory of the stability of
lyophobic colloids. New York: Elsevier; 1943.
[8] Sato T, Ruch R. Stabilization of colloidal dispersions by
polymer adsorption. New York, NY: 1980.
[9] Napper DH. Polymeric stabilization of colloidal dispersions.
London: Academic Press; 1983.
[10] Poehlein GW, Dougherty DJ. Continuous emulsion polymer-
ization. Rubber Chem Technol 1977;50:60138.
[11] Li B, Brooks BW. Semi-batch processes for emulsion
polymerization. Polym Int 1992;29:416.
[12] Snuparek J. Principles and limits of polymer latex tailoring.
Prog Org Coat 1996;29:22533.
[13] Gao J, Penlidis A. Mathematical modeling and computer
simulator/database for emulsion polymerizations. Prog Polym
Sci 2002;27:403535.
[14] Chern CS. Polymerization of monomer emulsions. In:
Hubbard A, editor. Encyclopedia of surface and colloid
science. New York: Marcel Dekker; 2002. p. 422041.
[15] Nomura M, Tobita H, Suzuki K. Emulsion polymerization:
kinetic and mechanistic aspects. Adv Polym Sci 2005;175:
1128.
[16] El-Aasser MS, Miller CM. Preparation of latexes
using miniemulsions. In: Asua JM, editor. Polymeric disper-
sions. Principles and applications. Dordrecht: Kluwer; 1997. p.
10926.
[17] Sudol ED, El-Aasser MS. Miniemulsion polymerization. In:
Lovell PA, El-Aasser MS, editors. Emulsion polymerization
and emulsion polymers. Chichester: Wiley; 1997. p. 699722.
[18] Capek I, Chern CS. Radical polymerization in direct
miniemulsion systems. Adv Polym Sci 2001;155:10165.
[19] Antonietti M, Landfester K. Polyreactions in miniemulsions.
Prog Polym Sci 2002;27:689757.
[20] Asua JM. Miniemulsion polymerization. Prog Polym Sci 2002;
27:1283346.
[21] Candau F. Polymerization in microemulsions. In: Paleos CM,
editor. Polymerization in organized media. Philadelphia:
Gordon & Breach; 1992. p. 21582.
[22] Candau F. Polymerization in microemulsions. In: Kumar P,
Mittal KL, editors. Handbook of microemulsion science and
technology. New York: Marcel Dekker; 1999. p. 679712.
[23] Capek I. Radical polymerization of polar unsaturated mono-
mers in direct microemulsion systems. Adv Colloid Interface
Sci 1999;80:85149.
[24] Capek I. Microemulsion polymerization of styrene in the
presence of anionic emulsier. Adv Colloid Interface Sci 1999;
82:25373.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 480
[25] Chern CS. Microemulsion polymerization. In: Encyclopedia of
polymer science and technology; 2003.
[26] Alexander AE, Napper DH. Emulsion polymerization. In:
Jensen DP, editor. Progress in polymer science. Oxford:
Pergamon Press; 1971. p. 14597.
[27] Barrett KE. Dispersion polymerization in organic media. New
York: Wiley; 1975.
[28] Ugelstad J, Hansen FK. Kinetics and mechanism of emulsion
polymerization. Rubber Chem Technol 1976;49:536609.
[29] Vanderhoff JW. Mechanism of emulsion polymerization.
J Polym Sci Polym Symp 1985;72:16198.
[30] Wang Q, Fu S, Yu T. Emulsion polymerization. Prog Polym
Sci 1994;19:70353.
[31] Aslamazova TR. Emulsier-free latexes and polymers on their
base. Prog Org Coat 1995;25:10967.
[32] Nagai K. Radical polymerization and potential applications of
surface-active monomers. Trends Polym Sci 1996;4:1227.
[33] Guyot A. Recent advances and challenges in the synthesis of
polymer colloids. Colloid Surf A: Physicochem Eng Aspects
1999;153:1121.
[34] Capek I. Radical polymerization of polyoxyethylene macro-
monomers in disperse systems. Adv Polym Sci 1999;145:155.
[35] Kawaguchi H. Functional polymer microspheres. Prog Polym
Sci 2000;25:1171210.
[36] Capek I. On the role of oil-soluble initiators in the radical
polymerization of micellar systems. Adv Colloid Interface Sci
2001;91:295334.
[37] Capek I. Sterically and electrosterically stabilized emulsion
polymerization. Kinetics and preparation. Adv Colloid Inter-
face Sci 2002;99:77162.
[38] Cunningham MF. Living/controlled radical polymerizations
in dispersed phase systems. Prog Polym Sci 2002;27:
103967.
[39] Harkins WD. A general theory of the reaction loci in emulsion
polymerization. J Chem Phys 1945;13:3812.
[40] Harkins WD. A general theory of the reaction loci in emulsion
polymerization. II. J Chem Phys 1946;14:478.
[41] Harkins WD. A general theory of the mechanism of emulsion
polymerization. J Am Chem Soc 1947;69:142844.
[42] Smith WV, Ewart RH. Kinetics of emulsion polymerization.
J Chem Phys 1948;16:5929.
[43] Smith WV. The kinetics of styrene emulsion polymerization.
J Am Chem Soc 1948;70:3695702.
[44] Smith WV. Chain initiation in styrene emulsion polymer-
ization. J Am Chem Soc 1949;71:407782.
[45] Gardon JL. Emulsion polymerization. I. Recalculation and
extension of the SmithEwart theory. J Polym Sci A 1 1968;6:
62341.
[46] Gardon JL. Emulsion polymerization. II. Review of exper-
imental data in the context of the revised SmithEwart theory.
J Polym Sci A 1 1968;6:64364.
[47] Chern CS, Hsu H, Lin FY. Stability of acrylic lattices in a
semibatch reactor. J Appl Polym Sci 1996;60:130111.
[48] Friis N, Hamielec AE. Gel-effect in emulsion polymer-
ization of vinyl monomers. In: Piirma I, Gardon JL,
editors. ACS symposium series no. 24, Washington, DC;
1976. p. 8291.
[49] Russell GT, Gilbert RG, Napper DH. Chain-length-dependent
termination rate processes in free-radical polymerization. 1.
Theory. Macromolecules 1992;25:245969.
[50] Priest WJ. Particle growth in the aqueous polymerization of
vinyl acetate. J Phys Chem 1952;56:107783.
[51] Roe CP. Surface chemistry aspects of emulsion polymer-
ization. Ind Eng Chem 1968;60:2033.
[52] Fitch RM, Tsai CH. Particle formation in polymer colloids.
III. Prediction of the number of particles by a homo-
geneous nucleation theory. In: Fitch RM, editor. Polymer
colloids. New York: Plenum Press; 1971. p. 73102.
[53] Fitch RM, Tsai CH. Homogeneous nucleation of polymer
colloids. IV. The role of soluble oligomeric radicals. In:
Fitch RM, editor. Polymer colloids. New York: Plenum Press;
1971. p. 10316.
[54] Fitch RM. The homogeneous nucleation of polymer colloids.
Br Polym J 1973;5:46783.
[55] Hansen FK, Ugelstad J. Particle nucleation in emulsion
polymerization. I. A theory for homogeneous nucleation.
J Polym Sci Polym Chem 1978;16:195379.
[56] Haward RN. Polymerization in a system of discrete particles.
J Polym Sci 1949;4:27387.
[57] Stockmayer WH. Note on the kinetics of emulsion polymer-
ization. J Polym Sci 1957;24:3137.
[58] OToole JT. Kinetics of emulsion polymerization. J Appl
Polym Sci 1965;9:12917.
[59] Ugelstad J, Mork PC, Aasen JO. Kinetics of emulsion
polymerization. J Polym Sci A 1 1967;5:22818.
[60] Ugelstad J, Mork PC. A kinetic study of the mechanism of
emulsion polymerization of vinyl chloride. Br Polym J 1970;2:
319.
[61] Feeney PJ, Geissler E, Gilbert RG, Napper DH. SANS study of
particle nucleation in emulsion polymerization. J Colloid
Interface Sci 1988;121:50813.
[62] Wang ST, Poehlein GW. Characterization of water-soluble
oligomer in acrylic-styrene emulsion copolymerization. J Appl
Polym Sci 1993;50:217383.
[63] Wang ST, Poehlein GW. Studies of water-soluble oligomers
formed in emulsion copolymerization. J Appl Polym Sci 1994;
51:593604.
[64] Thomson B, Wang Z, Paine A, Lajoie G, Rudin A. A mass
spectrometric investigation of the water-soluble oligomers
remaining after the emulsion polymerization of
methyl methacrylate. J Polym Sci Polym Chem Ed 1995;33:
2297304.
[65] Kuhn I, Tauer K. Nucleation in emulsion polymerization: a
new experimental study. 1. Surfactant-free emulsion polymer-
ization of styrene. Macromolecules 1995;28:81228.
[66] Tauer K, Deckwer R. Polymer end groups in persulfate-
initiated styrene emulsion polymerization. Acta Polym 1998;
49:4116.
[67] Kozempel S, Tauer K, Rother G. Aqueous heterophase
polymerization of styrenea study by means of multi-angle
laser light scattering. Polymer 2005;46:116979.
[68] Chern CS, Lin CH. Using a water-insoluble dye to probe the
particle nucleation loci in styrene emulsion polymerization.
Polymer 1998;40:13947.
[69] Chern CS, Lin CH. Particle nucleation loci in emulsion
polymerization of methyl methacrylate. Polymer 2000;41:
447381.
[70] Rudschuck S, Adams J, Fuhrmann J. Online observation of
emulsion polymerization by uorescence technique. Nucl
Instrum Meth Phys Res B 1999;151:3415.
[71] Amalvy JI, Asua JM, Leite CAP, Galembeck F. Elemental
mapping by ESI-TEM, during styrene emulsion polymer-
ization. Polymer 2001;42:247989.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 481
[72] Tauer K, Muller H, Schellenberg C, Rosengarten L. Evaluation
of heterophase polymerizations by means of reaction calori-
metry. Colloid Surf A: Physicochem Eng Aspects 1999;153:
14351.
[73] Varela De La Rosa L, Sudol ED, Dimonie VL, Klein A, El-
Aasser MS. Emulsion polymerization of styrene using
reaction calorimeter. II. Importance of maximum in rate of
polymerization. J Polym Sci A: Polym Chem 1999;37:
406672.
[74] Varela De La Rosa L, Sudol ED, El-Aasser MS, Klein A.
Emulsion polymerization of styrene using reaction calorimeter.
III. Effect of initial monomer/water ratio. J Polym Sci A:
Polym Chem 1999;37:407389.
[75] Sajjadi S. Diffusion-controlled particle growth and its effects
on nucleation in stirred emulsion polymerization reactors.
Macromol Rapid Commun 2004;25:8827.
[76] Kim JI, Vanderhoff JW, Klein A. Vapor phase addition method
to investigate the monomer transport at the latex particle
interface. Polym Test 1995;14:289306.
[77] Tauer K, Deckwer R, Kuhn I, Schellenberg C. A
comprehensive experimental study of surfactant-free emul-
sion polymerization of styrene. Colloid Polym Sci 1999;
277:60726.
[78] Wang YM, Pan CY. Study of the mechanism of the emulsier-
free emulsion polymerization of the styrene/4-vinylpyridine
system. Colloid Polym Sci 1999;277:65865.
[79] Ni H, Ma G, Nagai M, Omi S. Effects of ethyl acetate on the
soap-free emulsion polymerization of 4-vinylpyridine and
styrene. II. Aspects of the mechanism. J Appl Polym Sci 2001;
82:2692708.
[80] Ou JL, Yang JK, Chen H. Styrene/potassium persulfate/water
systems: effects of hydrophilic comonomers and solvent
additives on the nucleation mechanism and the particle size.
Eur Polym J 2001;37:78999.
[81] Yan C, Cheng S, Feng L. Kinetics and mechanism of
emulsier-free emulsion copolymerization: styrenemethyl
methacrylateacrylic acid system. J Polym Sci A: Polym
Chem 1999;37:264956.
[82] Mahdavian AR, Abdollahi M. Investigation into the effect of
carboxylic acid monomer on particle nucleation and growth in
emulsier-free emulsion copolymerization of styrene, buta-
diene and acrylic acid. Polymer 2004;45:32339.
[83] Zhang J, Zou Q, Li X, Cheng S. Soap-free cationic emulsion
copolymerization of styrene and butyl acrylate with co-
monomer in the presence of alcohols. J Appl Polym Sci
2003;89:27917.
[84] Shaffei KA, Ayoub MMH, Ismail MN, Badran AS. Kinetics
and polymerization characteristics for some polyvinyl acetate
emulsions prepared by different redox pair initiation systems.
Eur Polym J 1998;34:5536.
[85] Sahoo PK, Mohapatra R. Synthesis and kinetic
studies of PMMA nanoparticles by non-conventionally
initiated emulsion polymerization. Eur Polym J 2003;39:
183946.
[86] Cutting GR, Tabner BJ. Reaction temperature proles and
radical concentration measurements on batch emulsion
copolymerization of methyl methacrylate and butyl acrylate.
Eur Polym J 1995;31:12159.
[87] Cunningham MF, Geramita K, Ma JW. Measuring the effects
of dissolved oxygen in styrene emulsion polymerization.
Polymer 2000;41:538592.
[88] Capek I, Lin SY, Hsu TJ, Chern CS. Effect of temperature on
styrene emulsion polymerization in the presence of sodium
dodecyl sulfate. II. J Polym Sci A: Polym Chem 2000;38:
147786.
[89] Fang SJ, Wang K, Pan ZR. Behavior of free radical
transfer between aqueous phase and latex particles in
emulsion polymerization. Polymer 2003;44:138590.
[90] Imroz Ali AM, Tauer K, Sedlak M. Comparing emulsion
polymerization of methacrylate-monomers with different
hydrophilicity. Polymer 2005;46:101723.
[91] Lamb DJ, Fellows CM, Morrison BR, Gilbert RG. A critical
evaluation of reaction calorimetry for the study of emulsion
polymerization systems: thermodynamic and kinetic aspects.
Polymer 2005;46:28594.
[92] Ramirez JC, Herrera-Ordonez J, Maldonado-Textle H. Kin-
etics of the styrene emulsion polymerization above cmc. II.
Agitation effect on molecular weight. Polym Bull 2005;53:
3337.
[93] Mayer MJJ, Meuldijk J, Thoenes D. High conversion emulsion
polymerization in large scale. ChemEng Sci 1995;50:332930.
[94] Badran AS, Ayoub MMH, El-Ghaffar MAA, Naser HE, El-
Hakim AAA. Preparation and characterization of polystyrene
and styrenebutyl acrylate copolymer latices. Eur Polym J
1997;33:53741.
[95] Gu S, Mogi T, Konno M. Preparation of monodisperse,
micronsized polystyrene particles with single-stage polymer-
ization in aqueous media. J Colloid Interface Sci 1998;207:
1138.
[96] Omi S, Fujiwara K, Nagai M, Ma GH, Nakano AK. Study of
particle growth by seed emulsion polymerization with counter-
charged monomer and initiator system. Colloid Surf A:
Physicochem Eng Aspects 1999;153:16572.
[97] Ito F, Ma G, Nagai M, Omi S. Study of particle growth by
seeded emulsion polymerization accompanied by electrostatic
coagulation. Colloid Surf A: Physicochem Eng Aspects 2002;
201:13142.
[98] Ito F, Ma G, Nagai M, Omi S. Study on preparation of irregular
shaped particle in seeded emulsion polymerization
accompanied with regulated electrostatic coagulation by
counter-ion. Colloid Surf A: Physicochem Eng Aspects 2003;
216:10922.
[99] Wang CC, Kuo JF, Chen CY. Dimerizations of acrylate
monomers with sodium hydroxymethanesulnate and charac-
terization of the products. Macromol Chem Phys 1994;195:
1493502.
[100] Wang CC, Kuo JF, Chen CY. Emulsion polymerization of
butyl acrylate induced by sodium lauryl sulfate/hydroxy-
methane sulnate. Die Angew Macromol Chem Phys 1995;
231:1524.
[101] Wang CC, Kuo JF, Chen CY. Polymerization of styrene
initiated by a novel initiator sodium formaldehyde sulfoxylate
and sodium lauryl sulfate. Eur Polym J 2000;36:96574.
[102] Lin SY, Chern CS, Hsu TJ, Hsu CT, Capek I. Emulsion
polymerization of styrene: double emulsion effect. Polymer
2001;42:148191.
[103] Sajjadi A. Particle formation in Interval III of the emulsion
polymerization of styrene with Aerosol-MA as an emulsier.
J Polym Sci A: Polym Chem 2002;40:165263.
[104] Zaragoza-Contreras EA, Navarro-Rodriguez D. On the role of
an unconventional rigid rodlike cationic surfactant on the
styrene emulsion polymerization. Kinetics, particle size and
particle size distribution. Polymer 2003;44:55416.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 482
[105] Tang L, Yang J, Zhang S, Yang J, Wu Y. Emulsier-minor
emulsion copolymerization of BA-MMA-St-MAA (or AA)-N-
MA. J Appl Polym Sci 2004;92:29239.
[106] Chern CS, Yu TC. Effect of 1-pentanol on the styrene emulsion
polymerization. Polymer 2005;46:1899904.
[107] Xia H, Wang Q, Liao Y, Xu X, Baxter SM, Slone RV, et al.
Polymerization rate and mechanism of ultrasonically initiated
emulsion polymerization of n-butyl acrylate. Ultrason Sono-
chem 2002;9:1518.
[108] Yin N, Chen K. Ultrasonically initiated emulsier-free
emulsion copolymerization of n-butyl acrylate and acryl-
amide. Part I: polymerization mechanism. Polymer 2004;45:
358794.
[109] Yin N, Chen K. Particle formation mechanism and kinetic
model of ultrasonically initiated emulsion polymerization. Eur
Polym J 2005;41:135772.
[110] Ozdeer E, Sudol ED, El-Aasser MS, Klein A. Role of the
nonionic surfactant Triton X-405 in emulsion polymerization.
III. Copolymerization of styrene and n-butyl acrylate. J Polym
Sci A: Polym Chem 1997;35:383746.
[111] Capek I, Chudej J. On the ne emulsion polymerization of
styrene with non-ionic emulsier. Polym Bull 1999;43:
41724.
[112] Lin SY, Capek I, Hsu TJ, Chern CS. On the emulsion
polymerization of styrene in the presence of a nonionic
emulsier. J Polym Sci A: Polym Chem 1999;37:442231.
[113] Ouzineb K, Fortuny Heredia M, Graillat C, McKenna TF.
Stabilization and kinetics in the emulsion copolymerization of
butyl acrylate and methyl methacrylate. J Polym Sci A: Polym
Chem 2001;39:283246.
[114] Chen LJ, Lin SY, Chern CS, Wu SC. Critical micelle
concentration of mixed surfactant SDS/NP(EO)
40
and its role
in emulsion polymerization. Colloid Surf A: Physicochem Eng
Aspects 1997;122:1618.
[115] Chern CS, Lin SY, Chen LJ, Wu SC. Emulsion polymerization
of styrene stabilized by mixed anionic and nonionic surfac-
tants. Polymer 1997;38:197784.
[116] Chern CS, Lin SY, Chang SC, Lin JY, Lin YF. Effect of
initiator on styrene emulsion polymerization stabilized by
mixed SDS/NP-40 surfactants. Polymer 1998;39:22819.
[117] Unzueta E, Forcada J. Modeling the effect of mixed emulsier
systems in emulsion copolymerization. J Appl Polym Sci 1997;
66:44558.
[118] Boutti S, Zafra RD, Graillat C, McKenna TF. Interaction of
surfactant and initiator types in emulsion polymerizations: a
comparison of ammonium persulfate and hydrogen peroxide.
Macromol Chem Phys 2005;206:135572.
[119] Suresh KI, Othegraven J, Raju KVSN, Bartsch E. Mechanistic
studies on particle nucleation in the batch emulsion polymer-
ization of n-butyl acrylate containing multifunctional mono-
mers. Colloid Polym Sci 2004;283:4957.
[120] Nomura M, Fujita K. Particle nucleation in emulsion
copolymerization containing multifunctional monomers.
Polym Int 1992;30:4839.
[121] Barton J, Karpatyova A. Emulsion polymerization of butyl
methacrylate initiated by 2,2
/
-azobisisobutyronitrile. Makro-
mol Chem 1987;188:693702.
[122] Nomura M, Fujita K. Kinetics and mechanism of emulsion
polymerization initiated by oil-soluble initiators. 1. The
average number of radicals per particle. Makromol Chem
Rapid Commun 1989;10:5817.
[123] Asua JM, Rodrigues VS, Sudol ED, El-Aasser MS. The free
radical distribution in emulsion polymerization using oil-
soluble initiators. J Polym Sci A: Polym Chem 1989;27:
356987.
[124] Alduncin JA, Forcada J, Barandiaran MJ, Asua JM. On the
main locus of radical formation in emulsion polymerization
initiated by oil-soluble initiators. J Polym Sci A: Polym Chem
1991;29:126570.
[125] Blythe PJ, Klein A, Phililips JA, Sudol ED, El-Aasser MS.
Miniemulsion polymerization of styrene using the oil-soluble
initiator AMBH. J Polym Sci A: Polym Chem 1999;37:
444957.
[126] Luo Y, Schork FJ. Emulsion and miniemulsion polymer-
izations with an oil-soluble initiator in the presence and
absence of an aqueous-phase radical scavenger. J Polym Sci A:
Polym Chem 2002;40:320011.
[127] Jain M, Vora RA, Satpathy US. Kinetics of emulsion
copolymerization of methylmethacrylate and ethylacrylate:
effect of type and concentration of initiator in unseeded
polymerization system. Eur Polym J 2003;39:206976.
[128] Chern CS, Lin SY, Chang SC. Reaction kinetics of styrene
emulsion polymerization stabilized by mixed surfactants of
SDS and NP-40. J Chin Inst Chem Eng 1999;30:23342.
[129] Chern C, Shi Y, Wu J. Acrylic lattices stabilized by
polymerizable non-ionic surfactant. Polym Int 1996;40:
12939.
[130] Denise J, Sherrington DC. Novel polymerizable mono- and
divalent quaternary ammonium cationic surfactants: 2. Surface
active properties and use in emulsion polymerization. Polymer
1997;38:142738.
[131] Montoya-Goni A, Sherrington DC, Schoonbrood HAS,
Asua JM. Reactive surfactants in heterophase polymerization.
XXIV. Emulsion polymerization of styrene with maleate- and
succinate-containing cationic surfactants. Polymer 1999;40:
135966.
[132] Klimenkovs I, Zhukovska I, Uzulina I, Zicmanis A, Guyot A.
Maleic diamide polymerizable surfactants. Applications in
emulsion polymerization. C R Chim 2003;6:1295304.
[133] Mezger T, Nuyken O, Meindl K, Wokaun A. Light
decomposable emulsiers: application of alkyl-substituted
aromatic azosulfonates in emulsion polymerization. Prog Org
Coat 1996;29:14757.
[134] Aslamazova T, Tauer K. On the colloidal stability of
polystyrene particles prepared with surface active initiators.
Adv Colloid Interface Sci 2003;104:27383.
[135] Aslamazova T, Tauer K. On the colloidal stability of
polystyrene particles prepared with different kinds of surface
active initiators. Colloid Surf A: Physicochem Eng Aspects
2004;239:310.
[136] Riess G. Block copolymers as polymeric surfactants in latex
and microlatex technology. Colloid Surf A: Physicochem Eng
Aspects 1999;153:99110.
[137] Kislenko VN. Emulsion graft polymerization: mechanism of
formation of dispersions. Colloid Surf A: Physicochem Eng
Aspects 1999;152:199203.
[138] Hwu HD, Lee YD. Studies of alkali soluble resin as a surfactant
in emulsion polymerization. Polymer 2000;41:5695705.
[139] Kato S, Sato K, Maeda D, Nomura M. A kinetic investigation
of styrene emulsion polymerization with surface active
polyelectrolytes as the emulsier. II. Effects of molecular
weight and composition. Colloid Surf A: Physicochem Eng
Aspects 1999;153:12731.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 483
[140] Rimmer S, Tattersall P. The inclusion of b cyclodextrin
provides a supramolecular solution to the problem of
polymerization of dodecyl and octadecyl methacrylates in
aqueous emulsion. Polymer 1999;40:572931.
[141] Chang YH, Lee YD. Structural effects of a novel polymeric
emulsierSSIPM modied tetracarboxylic acid terminated
polyester on the emulsion polymerization of butyl methacry-
late. Polymer 1999;40:592937.
[142] Chang YH, Lee YD, Karlsson OJ, Sundberg DC. Particle
nucleation mechanism for the emulsion polymerization of
styrene with a novel polyester emulsier. J Appl Polym Sci
2001;82:106170.
[143] Chern CS, Lee C. Emulsion polymerization of styrene
stabilized with an amphiphilic PEG-containing graft copoly-
mer. Macromol Chem Phys 2001;202:27509.
[144] Chern CS, Lee C. Effect of the structure of amphiphilic
poly(ethylene glycol)-containing graft copolymers on styrene
emulsion polymerization. J Polym Sci A: Polym Chem 2002;
40:160824.
[145] Burguiere C, Chassenieux C, Charleux B. Characterization of
aqueous micellar solutions of amphiphilic block copolymers of
poly(acrylic acid) and polystyrene prepared via ATRP. Toward
the control of the number of particles in emulsion polymer-
ization. Polymer 2003;44:50918.
[146] Beal L, Chevalier Y. Adsorption of block copolymers at latex
surface and their utilization in emulsion polymerization.
Polymer 2005;46:1395405.
[147] Bromley CWA. Preparation of sterically stabilised aqueous
latices using polyethylene oxide. Colloid Surf 1986;17:111.
[148] Brown R, Stutzel B, Sauer T. Steric stabilization by grafting
and copolymerization of water-soluble oligomers and poly-
mers. Macromol Chem Phys 1995;196:204764.
[149] Chern CS, Liou YC, Tsai WY. Emulsion polymerization of
acrylic monomers stabilized by poly(ethylene oxide).
J Macromol Sci Pure Appl Chem 1996;A33:106375.
[150] Liu Z, Xiao H, Wiseman N, Zheng A. Poly(ethylene oxide)
macromonomer-grafted polymer nanoparticles synthesised by
emulsier-free emulsion polymerization. Colloid Polym Sci
2003;281:81522.
[151] Okaya T, Suzuki A, Kikuchi K. Importance of grafting in the
emulsion polymerization of MMA using PVA as a protective
colloid. Effect of initiators. Colloid Surf A: Physicochem Eng
Aspects 1999;153:1235.
[152] Budhlall BM, Sudol ED, Dimonie VL, Klein A, El-Aasser MS.
Role of grafting in the emulsion polymerization of vinyl
acetate with poly(vinyl alcohol) as an emulsier. I. Effect of the
degree of blockiness on the kinetics and mechanism of
grafting. J Polym Sci A: Polym Chem 2001;39:363354.
[153] Carra S, Sliepcevich A, Canevarolo A, Carra S. Grafting and
adsorption of poly(vinyl) alcohol in vinyl acetate emulsion
polymerization. Polymer 2005;46:137984.
[154] Cheong IW, Nomura M, Kim JH. Water-soluble polyurethane
resins as emulsiers in emulsion polymerization of styrene:
nucleation and particle growth. Macromol Chem Phys 2001;
202:245460.
[155] Zhang G, Zhang Z. The
60
Co-g ray-initiated seeded emulsion
polymerization of methyl methacrylate in the presence of
waterborne polyurethane seeds. Radiat Phys Chem 2004;71:
2714.
[156] Morton M, Kaizerman S, Altier MW. Swelling of latex
particles. J Colloid Sci 1954;9:30012.
[157] Lopez de Arbina L, Barandiaran MJ, Gugliotta LM, Asua JM.
Emulsion polymerization: particle growth kinetics. Polymer
1996;37:590716.
[158] Mayer MJJ, van den Boomen FHAM, Paquet DA, Meuldijk J,
Thoenes D. The polymerization rate of freshly nucleated
particles during emulsion polymerization with micellar
nucleation. J Polym Sci A: Polym Chem 1996;34:174751.
[159] Coen EM, Gilbert RG, Morrison BR, Leube H, Peach S.
Modelling particle size distributions and secondary particle
formation in emulsion polymerization. Polymer 1998;39:
7099112.
[160] Ferguson CJ, Russell GT, Gilbert RG. Modelling
secondary particle formation in emulsion polymerization:
application to making coreshell morphologies. Polymer
2002;43:455770.
[161] Lin M, Chu F, Guyot A, Putaux JL, Bourgeat -Lami E.
Silicone-polyacrylate composite latex particles. Particles
formation and lm properties. Polymer 2005;46:13317.
[162] Sathyagal AN, McCormick AV. Effect of nucleation prole on
particle-size distribution. AIChE J 1998;44:231223.
[163] Herrera-Ordonez J, Olayo R. On the kinetics of styrene
emulsion polymerization above CMC. I. A mathematical
model. J Polym Sci A: Polym Chem 2000;38:220118.
[164] Herrera-Ordonez J, Olayo R. Methyl methacrylate emulsion
polymerization at low monomer concentration: kinetic model-
ing of nucleation, particle size distribution, and rate of
polymerization. J Polym Sci A: Polym Chem 2001;39:
254756.
[165] Kammona O, Pladis P, Frantzikinakis CE, Kiparissides C. A
comprehensive experimental and theoretical investigation of
the styrene/2-ethylhexyl acrylate emulsion copolymerization.
Macromol Chem Phys 2003;204:98399.
[166] Ginsburger E, Pla F, Fonteix C, Hoppe S, Massebeuf S,
Hobbes P, et al. Modelling and simulation of batch and semi-
batch emulsion copolymerization of styrene and butyl acrylate.
Chem Eng Sci 2003;58:4493514.
[167] Coen EM, Peach S, Morrison BR, Gilbert RG. First-principles
calculation of particle formation in emulsion polymerization:
pseudo-bulk systems. Polymer 2004;45:3595608.
[168] Dong Y. Radical entry in emulsion polymerization: propa-
gation at latex particles/water interfaces. J Colloid Interface Sci
2005;288:3905.
[169] Okubo M, Katsuta Y, Matsumoto T. Rupture of anomalous
composite particles prepared by seeded emulsion polymer-
ization in aging period. J Polym Sci Polym Chem Ed 1980;18:
4816.
[170] Okubo M, Yamada A, Matsumoto T. Estimation of mor-
phology of composite polymer emulsion particles by the soap
titration method. J Polym Sci Polym Chem Ed 1980;18:
321928.
[171] Lee DI. Morphology of two-stage latex particles, polystyrene
and styrene-butadiene copolymer pair systems. In: Bassett DR,
Hamielec AE, editors. Emulsion polymers and emulsion
polymerization. Washington, DC: ACS Symposium Series,
No. 165; 1981. p. 40514.
[172] Sundberg DC, Casassa AP, Pantazopoulos J, Muscato MR,
Kronberg B, Berg J. Morphology development of polymeric
microparticles in aqueous dispersions. I. Thermodynamic
considerations. J Appl Polym Sci 1990;41:142542.
[173] Chen YC, Dimonie V, El-Aasser MS. Effect of interfacial
phenomena on the development of particle morphology in a
polymer latex system. Macromolecules 1991;24:377987.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 484
[174] Durant Y, Sundberg DC. An advanced computer algorithm for
determining morphology development in latex particles. J Appl
Polym Sci 1995;58:160718.
[175] Gonzalez-Ortiz LJ, Asua JM. Development of particle
morphology in emulsion polymerization. 1. Cluster dynamics.
Macromolecules 1995;28:313545.
[176] Gonzalez-Ortiz LJ, Asua JM. Development of particle
morphology in emulsion polymerization. 2. Cluster dynamics
in reacting systems. Macromolecules 1996;29:3839.
[177] Gonzalez-Ortiz LJ, Asua JM. Development of particle
morphology in emulsion polymerization. 3. Cluster nucleation
and dynamics in polymerizing systems. Macromolecules 1996;
29:45207.
[178] Stubbs J, Karlsson O, Jonsson JE, Sundberg E, Durant Y,
Sundberg DC. Non-equilibrium particle morphology develop-
ment in seeded emulsion polymerization. 1: penetration of
monomer and radicals as a function of monomer feed rate
during second stage polymerization. Colloid Surf A: Physi-
cochem Eng Aspects 1999;153:25570.
[179] Aerdts AM, Groeninckx G, Zirkzee HF, van Aert HAM,
Geurts JM. Preparation of of epoxy-functionalized methyl
methacrylatebutadienestyrene coreshell particles and inves-
tigation of their dispersion in polyamide-6. Polymer 1997;38:
424752.
[180] Karlsson OJ, Hassander H, Colombini D. The effect of rst-
stage polymer T
g
on the morphology and thermomechanical
properties of structured polymer latex particles. C R Chim
2003;6:123344.
[181] Ferguson CJ, Russell GT, Gilbert RG. Synthesis of latices with
hydrophobic cores and poly(vinyl acetate) shells. 2. Use of
poly(vinyl acetate) seeds. Polymer 2003;44:260719.
[182] Zhao K, Sun P, Liu D, Dai G. The formation mechanism of
poly(vinyl acetate)/poly(butyl acrylate) core/shell latex in two-
stage seeded semi-continuous starved emulsion polymerization
process. Eur Polym J 2004;40:8996.
[183] Pan M, Zhang L, Wan L, Guo R. Preparation and characterization
of composite resin by vinyl chloride grafted onto poly(BA-
EHA)/poly(MMA-St). Polymer 2003;44:71219.
[184] Gugliotta LM, Arotcarena M, Leiza JR, Asua JM. Estimation
of conversion and copolymer composition in semicontinuous
emulsion polymerization using calorimetric data. Polymer
1995;36:201923.
[185] Unzueta E, Forcada J. Semicontinuous emulsion copolymer-
ization of methyl methacrylate and n-butyl acrylate: 2. Effect
of mixed emulsiers in unseeded polymerization. Polymer
1995;36:43018.
[186] Chern CS, Hsu H. Semibatch emulsion copolymerization of
methyl methacrylate and butyl acrylate. J Appl Polym Sci
1995;55:57181.
[187] Chern CS, Hsu H, Lin FY. Stability of acrylic latices in a
semibatch reactor. J Appl Polym Sci 1996;60:130111.
[188] Chern CS, Lin CH. Semibatch surfactant-free emulsion
polymerization of butyl acrylate. Polym J 1995;27:1094103.
[189] Chern CS, Lin FY. Semibatch emulsion polymerization of
butyl acrylate: effect of functional monomers. J Macromol Sci
Pure Appl Chem 1996;A33:107796.
[190] Chern CS, Lin FY. Stability of carboxylated poly(butyl
acrylate) latices during semibatch emulsion polymerization.
J Appl Polym Sci 1996;61:9891001.
[191] Novak RW. Mechanism of acrylic emulsion polymerizations.
Adv Org Coat Sci Technol Ser 1988;10:547.
[192] Chern CS, Lin CH. Semibatch surfactant-free emulsion
polymerization of butyl acrylate in the presence of carboxylic
monomers. Polym J 1996;28:34351.
[193] Chern CS, Chen YC. Semibatch emulsion polymerization of
butyl acrylate stabilized by a polymerizable surfactant. PolymJ
1996;28:62732.
[194] Chern CS, Chen YC. Stability of the polymerizable surfactant
stabilized latex particles during semibatch emulsion polymer-
ization. Colloid Polym Sci 1997;275:12430.
[195] Chern CS, Chen YC. Kinetics of semibatch emulsion
polymerization of butyl acrylate stabilized by a reactive
surfactant. J Macromol Sci Pure Appl Chem 1998;A35:
96583.
[196] Xu XJ, Chen F. Semi-continuous emulsion copolymerization
of butyl methacrylate with polymerizable anionic surfactants.
Polymer 2004;45:480110.
[197] Chern CS, Lin CH, Chen TJ. Latex stability in semibatch
surfactant-free seeded emulsion polymerization of butyl
acrylate. Polym J 1997;29:24954.
[198] Chern CS, Huang CF, Wang CL. Effect of highly carboxylated
seed latices on semibatch surfactant-free emulsion polymer-
ization of butyl acrylate. J Dispersion Sci Technol 1998;19:
118.
[199] Sajjadi S, Brooks BW. Semibatch emulsion polymerization of
butyl acrylate. I. Effect of monomer distribution. J Appl Polym
Sci 1999;74:3094110.
[200] Sajjadi S, Brooks BW. Unseeded semibatch emulsion
polymerization of butyl acrylate: bimodal particle size
distribution. J Polym Sci A: Polym Chem 2000;38:52845.
[201] Sajjadi S, Brooks BW. Semibatch emulsion polymerization
reactors: polybutyl acrylate case study. Chem Eng Sci 2000;55:
475781.
[202] Wessling RA. Kinetics of continuous addition emulsion
polymerization. J Appl Polym Sci 1968;12:30919.
[203] Chern C, Chen TJ, Wu SY, Chu HB, Huang CF. Semibatch
seeded emulsion polymerization of acrylic monomers: bimodal
particle size distribution. J Macromol Sci Pure Appl Chem
1997;A34:122136.
[204] Schneider M, Graillat C, Guyot A, McKenna TF. High solids
content emulsions. III. Synthesis of concentrated latices by
classic emulsion polymerization. J Appl Polym Sci 2002;84:
191634.
[205] Boutti S, Graillat C, McKenna TF. New routes to high solid
content latexes: a process for in situ particle nucleation and
growth. Macromol Symp 2004;206:38398.
[206] Sebenik S, Krajnc M. Seeded semibatch emulsion copolymer-
ization of methyl methacrylate and butyl acrylate using
polyurethane: effect of soft segment length on kinetics. Colloid
Surf A Physicochem Eng Aspects 2004;233:5162.
[207] Chern CS. Diffusion-controlled semibatch emulsion polymer-
ization of styrene. J Appl Polym Sci 1995;56:22130.
[208] Chern CS. Desorption of free radicals in semibatch emulsion
polymerization of methyl acrylate. J Appl Polym Sci 1995;56:
2318.
[209] Wang Z, Paine AJ, Rudin A. Control of surfactant level in
starved fed emulsion polymerization: 4. Mathematical model
and experimental test. J Colloid Interface Sci 1996;177:
60212.
[210] Chern CS, Kuo YN. Shear-induced coagulation kinetics of
semibatch seeded emulsion polymerization. Chem Eng Sci
1996;51:107987.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 485
[211] Chern CS, Lin CH. Particle nucleation in semi-batch
surfactant-free emulsion polymerization of acrylic monomers.
Polym Int 1997;42:40921.
[212] Sajjadi S. Particle formation under monomer-starved con-
ditions in the semibatch emulsion polymerization of styrene.
Part II. Mathematical modelling. Polymer 2003;44:22337.
[213] Immanuel CD, Doyle III FJ, Cordeiro CF, Sundaram SS.
Population balance PSD model for emulsion polymerization
with steric stabilizers. AIChE J 2003;49:1392404.
[214] Sajjadi S, Yianneskis M. Analysis of particle formation under
monomer-starved conditions in emulsion polymerization
reactors. Macromol Symp 2004;206:20114.
C.S. Chern / Prog. Polym. Sci. 31 (2006) 443486 486

Вам также может понравиться