You are on page 1of 13

Full Paper

Modeling the Semicontinuous Heterophase


Polymerization for Synthesizing Poly(n-butyl
methacrylate) Nanoparticles

Marı́a G. Pérez-Garcı́a, Lourdes A. Pérez-Carrillo, Eduardo Mendizábal,


Jorge E. Puig, Francisco López-Serrano*

A simple four-state model containing five-parameters, in conjunction with experimental


conversion, particle number density and average molar masses, is presented to reproduce
the semicontinuous heterophase polymerization of butyl methacrylate. The model ade-
quately describes the experimental evolutions of conversion, particle size, average number
of particles, average number of radicals per particle,
and molar mass. A coagulation mechanism was
detected and its rate constant was evaluated. The
entry coefficient and the critical radical concentration
in the aqueous phase were approximately constant
and the exit rate coefficient decreased with monomer
feed rate. Homogeneous nucleation is the dominant
mechanism of production of particles, micellar
nucleation occurs at the slowest feeding rate
examined.

1. Introduction routine fashion by batch microemulsion polymerization,


although the large amounts of surfactant and the small
Polymer nanoparticles are attracting increasing attention amounts of polymer produced have prevented its industrial
in the scientific and technological fields because of scaling.[8–10] Modifications of this process that consist of
their applications in coatings, paints, adhesives, and the semicontinuous or continuous monomer addition,
catalysis.[1–4] When properly functionalized, polymer allow increasing the amount of polymer and decreasing
nanoparticles can be used to immobilize proteins, enzymes, the surfactant/polymer ratio.[10–17]
and antibodies, as well as in the design of sensing devices A novel variation of the semicontinuous microemulsion
and in cell labeling.[2,5–7] Polymer particles in the nano- polymerization has been reported in recent years, that has
meter range can be produced with fast reaction rates in a the advantages of emulsion and microemulsion polymer-
ization processes, mainly because it allows the synthesis
of nanoparticles with similar or smaller sizes than those
F. López-Serrano
obtained by microemulsion polymerization, and yields
Departamento de Ingenierı́a Quı́mica, Facultad de Quı́mica,
large polymer contents using small amounts of surfac-
Universidad Nacional Autónoma de México, D.F. 04510, México
tant.[18–26] Here, either monomer is fed continuously drop-
E-mail: lopezserrano@unam.mx
M. G. Pérez-Garcı́a, L. A. Pérez-Carrillo, E. Mendizábal, J. E. Puig wise over a period of time (differential microemulsion
Departamento de Ingenierı́a Quı́mica, Universidad de polymerization)[18,19,22,23,25] or at a controlling rate [semi-
Guadalajara, Boul. M. Garcı́a Barragán 1451, Guadalajara, Jal. continuous heterophase polymerization (SHP)][20,21,24,26]
44430, México over a monomer-free aqueous solution containing only

ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104 1
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
M. G. Pérez-Garcı́a, L. A. Pérez-Carrillo, E. Mendizábal, J. E. Puig, F. López-Serrano

www.mre-journal.de

surfactant and initiator to achieve monomer-starved require a posteriori analysis by integrating the ODE-set to
conditions; under these conditions no emulsified or obtain the parameters’ estimates. Michalik et al.[53] for-
microemulsified monomer droplets are present in the warded another procedure with an incremental parameter
process. Furthermore, molar masses produced by this type estimation approach, where splines are used to fit the data
of processes are smaller than those obtained by micro- and differentiation of the splines provides the derivatives’
emulsion polymerization, and they can be controlled by the estimate. In their approach, a subsystem of the problem is
monomer feeding rate.[20,24,26] considered and a corresponding subset of the parameters
However, few attempts to model these semicontinuous is estimated using an iterative strategy.
processes have been reported. He et al. forwarded a model to The third methodology, the so-called differential method
reproduce the differential microemulsion polymerization where the information contained in the measurement
of methyl methacrylate (MMA)[27] and styrene (St)[28] derivate is used, has been applied mainly to reaction
assuming only homogeneous nucleation for the case of kinetics. This approach is explained in detail in text-
MMA, and both micellar and homogeneous nucleation for books,[54,55] and it has been extensively used in thermal
St. Their predictions were limited to conversion and particle analysis,[56] in the estimation of parameters,[57,58] and in
size, in which predictions overestimated particle size. model assessment in polymerization systems.[59]
Aiming to scale the production of polymer nanoparticles, Here, a simple hybrid procedure, which in principle takes
which has not been implemented to date by microemulsion advantages of the virtues of some of the approaches
differential or SHP, a mathematical model is needed for described above, is used. Inferred data is obtained from
the understanding of at least the fundamental events, time-derivative estimates to obtain an additional measure-
among the numerous phenomena, occurring in this ment (the average number of radicals per particle). In
complicated process. Due to the system complexity, even addition, instead of substituting the left hand side of
a simple model would require many parameters, usually the derivative in the ODE-set using splines[51] as performed
not reported in the literature, and also very often the earlier,[50] empirical ad hoc fits for two of the states
experimental data is not sufficient either. Therefore, (conversion and particle number in our case) are inserted
parameter estimation with the available information is a into the right hand side of the now uncoupled equations, and
crucial task. regression is performed to obtain the pursued parameters.
For ordinary differential equation (ODE) sets, typical in In this work, a model is presented that describes the
emulsion and microemulsion polymerization modeling, evolution of global conversion, size and number density of
the deterministic parameter estimation can be classified particles, average molar masses, and average particle
widely but not exclusively, into three different methodol- radical concentration for the SHP. Here, an uncoupling of
ogies: one that requires the integration of the ODE set and the ODE-set is performed and additional information
the other two that do not require such integration. These contained in the derivative of the smoothed experimental
approaches are described next. conversion data is incorporated. The model predictions are
The ODE-set integration methodology can be divided into compared with experimental data of n-butyl methacrylate
those in which optimization is decoupled from the ODE (BMA) at three monomer addition rates and the role of
system integration[29–33] and others in which the ODE-set is the different nucleation and termination mechanisms are
transformed into an algebraic system where the optimiza- examined.
tion is then performed.[34–38] The drawback of these
methodologies lies in the difficulty that they may only
reach a local optimal solution and may fail to achieve the
global optima. To overcome these limitations, two main
2. Experimental Section
approaches have been followed: one based on deterministic n-Butyl methacrylate (BMA, 99% pure from Sigma–Aldrich) was
global optimization techniques,[39–46] and the other based passed through a DHR-4 column (Scientific Polymer Products) to
on a hybrid generalized algorithm approach.[47–49] In the remove the inhibitor (hydroquinone methyl ester). The surfactant
second methodology, Varah[50] proposed an approach that sodium dodecylsulfate (SDS, 98.5% pure from Aldrich) and
does not require numerical integration of the ODE-set. In potassium persulfate (KPS, 99% pure from Alfa Aesar) were used
this approach, the measured data is fitted empirically using as received. De-ionized and bi-distilled water was used. Sodium
bicarbonate, 95% pure from Fermont, was employed as a buffer to
splines,[51] the functions of which can be time-differen-
maintain the pH at 6.5–7.0.
tiated obtaining the derivatives’ estimates. The substitu-
Polymerizations were carried out at 50 8C at a stirring speed of
tion of these estimates into the equation set reduces the 300 rpm in a Hel computer controlled polymerization reactor. The
dynamic optimization problem into a much simpler procedure for the reaction consisted of dissolving 3 g of SDS, 0.6 g of
algebraic optimization one. Varziri et al.[52] proposed an KPS, and the necessary amount of the buffer to obtain a pH of 7.0
iterative principal differential analysis approach based in 236 g of water. This solution was loaded in the reactor, bubbled
upon fitting splines to the data. These methods usually with nitrogen to remove the oxygen present, and heated to the

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


2 www.MaterialsViews.com
ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Modeling the Semicontinuous Heterophase Polymerization . . .

www.mre-journal.de

Table 1. Final solid content, final average particle size, final particle number density, surface coverage ratio, and molar mass for the different
monomer feeding rates.

Fm [g  min1] Solid content [%] Dz [nm] Npmax T 1015 [mL1] r M n [Da]

0.2 21.2 30 18.9 0.21 429 400


0.49 20.3 33 12.2 0.24 712 461
0.73 19 35 10.3 0.28 651 627

reaction temperature. Then the monomer was fed to the reactor describe the behavior of the experimental data considered
using a calibrated dosing pump (Hel) at the desired addition rate. here.
Different experiments were performed, keeping the formulation
constant, the procedure and the total amount of monomer fed, but
varying the monomer addition rate. Samples were withdrawn at 3.1. Problem Statement
given times and placed into vials containing cold 0.05 M hydro-
Let us regard an isothermal monomer-starved SHP model
quinone aqueous solution to inhibit the reaction. The polymer was
precipitated from the withdrawn samples by adding methanol that includes all the relevant mechanisms considered in
(Bassel, reactive grade), recovered by filtration, purified, and dried earlier studies,[62] under the following assumptions: (i) a 0–
to measure conversion, average molar masses and molar mass 1 system is considered, which is equivalent to assuming
distributions. instantaneous termination inside the particles when a
z-Average particle size (Dz) as a function of reaction time was radical enters an active particle, (ii) monomer solubility in
measured at 25 8C by quasielastic light scattering (QLS) in a Malvern water is negligible for monomer distribution and conver-
Zetasizer Nano-ZS90 apparatus. To eliminate multiple scattering sion considerations, but it is taken into account for
and particle interactions, latexes were diluted with water up to the homogeneous nucleation mechanism description,
50-times. The particle size standard operating procedure sub- (iii) particles are spherical and of the same size to avoid a
routine of this equipment was employed to estimate the particle
population balance approach, (iv) coagulation of particles
size distribution (PSD).
may exist, (v) entry kinetic constant of radicals into micelles
Average molar masses were measured in a Waters (Breeze
model) gel permeation chromatograph with a Waters 1515
or into particles could be different, (vi) nucleation can occur
Isocratic HPLC Pump at a flow rate of 1.0 mL  min1 of by homogeneous and micellar mechanisms, (vii) values
tetrahydrofuran (THF) as the mobile phase. Two standard styrogel of kp in the particles and in the water phase are similar,
columns (Polyscience) covering the exclusion range from 2 000 to (viii) volume additivity applies, and (ix) pseudo-steady
1.7  106 g  mol1 and a Waters 2414 Refractive Index Detector state (SS) of radicals in the aqueous medium applies.
were used. The calibration was made with narrow-molar mass The resulting model is given by the following set of ODEs:
polystyrene (PS) standards (Polysciences Inc.), covering the molar
mass range of 1  103 to 1.5  106 g  mol1. Molar mass distribu- dx kp Mp nNp V
tions were calculated from the calibration curve by applying the ¼ ; xð0Þ ¼ 0 (1)
dt MT Nav
following Mark–Houwink parameters for PS and poly(butyl
methacrylate) (PBMA) in THF: kPS ¼ 1.14  104 g  mol1,
dNp
aPS ¼ 0.716, kPBMA ¼ 1.48  104 dL  g1 and aPBMA ¼ 0.664,[60] and ¼ kcm m þ kp Maq Rc1 Nav þ kt Rc1 RNav
ð1þaPSÞ
the Benoit’s correction,[61] MPBMA ¼ {(kPS MPS )/kPBMA}1/(1 þ aPBMA),
dt
(2)
kc 2 Np dV
where MPS and MPBMA are the PS and PBMA molar masses,  N  ; Np ð0Þ ¼ 0
respectively. For measurements, 100 ml of polymer samples Nav p V dt
dissolved in THF at a concentration of 5 mg  mL1 were injected
into the chromatograph at 30 8C. The fact that the sample dn m
¼ kcm þ kcp ð1  2nÞ  nke
completely dissolved in THF, ruled out the presence of gel. dt Np
Table 1 lists the microlatex and polymer particles characteristics Maq Rc1 Nav 2kc Np n2 n dV
at the end of the reaction. þ kp   ; nð0Þ ¼ 0
Np Nav V dt
(3)

3. The Model !
dV Fm 1 1 dx
¼ þ MT Mw  ; Vð0Þ ¼ V0 (4)
dt rm rp rm dt
In this section, the SHP modeling is formulated and the
proposed solution approach is presented. Our model
assessment problem consists of finding out the parts of In these equations, the dependent variables (or states) are
the general purpose model that should be retained to x, Np, n, and V, which correspond to overall conversion (–),

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


3
www.MaterialsViews.com ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
M. G. Pérez-Garcı́a, L. A. Pérez-Carrillo, E. Mendizábal, J. E. Puig, F. López-Serrano

www.mre-journal.de

polymer particle number density (L1), average-number of The values of all the parameters used are listed in Table 2 as
radicals per particle (–), and reactor volume (L), respectively. well as the references from where the parameters were
The independent variable is the reaction time, t (min), kp, taken.[63–69]
and kt, are the propagation and termination rate constants It is worth mentioning that for the model solution
(both in L  mol1  s1), respectively. The rate parameters for and to minimize numerical problems, the equation for
this heterogeneous process are the radicals’ entry to the time evolution of Np was made dimensionless by
micelles (kcm), to particles (kcp), and radicals exit from dividing Np by the maximum number of particles,
particles (ke), which are all considered to be of first order Npmax, for each experiment (i.e., Np ¼ Np =Np max )
(s1), and the rate of coagulation between particles kc (Table 1); m was also made dimensionless by defining
(L  mol1  s1), which is assumed to be of second order. Also, m ¼ m=Np max .
the total radical concentration in the water phase, R, and the The modeling assessment problem consists of estimating
critical-minus-one radical concentration, Rc–1, in the water the following fiveparameter grouping set: kcm, kcp, ke,
phase (mol  L1) are considered. Other variables appearing kcNpmax/Nav, and MaqRc–1Nav/Npmax (min1) using the
are the monomer concentration inside the particles Mp experimental evolution measurements of conversion y(t),
(mol  L1) and the micelle concentration m (L1). The the dimensionless particle number density z(t), and the
process parameters are the total amount of monomer MT average-number of radicals per particle ne(t) and a four-
(mol) and the rate of monomer addition Fm (g  min1). Other state dynamic model [Equation (1–4)]. It is understood that
parameters are the monomer molar mass Mw (g  mol1), the direct application of the standard regression-based
the Avogadro’s number Nav (mol1), the monomer con- integral approach is by no means a straightforward task
centration in the water phase Maq (mol  L1), and the because of the large ratio of adjustable parameters-to-
density r (g  mL1), with subindex m for monomer and p experimental measurements.[[41] and references therein]
for polymer. The micellar concentration m, can be obtained
as follows:[62]
3.2. Verifying the 0–1 Hypothesis and Inferring an
! Additional Experimental Measurement
S pD2p Np
m¼   CMC Nav =nagg (5) To validate the 0–1 assumption and to obtain an additional
Vw Mwe Nav as
experimental measurement, i.e., the average number of
radicals inside the particles (n), the following procedure was
In Equation (5), S is the amount of emulsifier (g), Vw the carried out: (i) the conversion experimental data [x(t)] were
water volume (L), Mwe the emulsifier molar mass (g  mol1), fitted with an empirical function y(t), (ii) the experimental
as the emulsifier coverage area (Å ´ 2), CMC the critical data for the dimensionless number of particles ½Np ðtÞ was
1
micellar concentration (mol  L ), Dp (cm) the particle fitted with an empirical function z(t); these two steps allow
diameter, and nagg the emulsifier aggregation number. uncoupling the model equation set [Equation (1–4)], and

Table 2. Values of parameters used in the model.

Parameter Value Ref.

Mw [g  mol1] 142.2
Nav [mol1] 6.03  1023
Mwe [g  mol1] 288
CMC [mol  L1] 0.008 [63]
2
as [Å ] 45 [64]
Nagg [–] 64 [64]
rm [g  mL1] 0.91545–9.64  104 (Temp 8C) [65]
1
rp [g  mL ] 1.19–8.07  104 (Temp 8C) [65]
kp [L  mol1  s1] 3.78  106exp(–2 754/(Temp þ273.15)) [66]
1 1
ktr [L  mol s ] 2.82  103exp(–3 717/(Temp þ273.15)) [67]
1 1 7
kt [L  mol s ] 9.43  10 exp(–830/(Temp þ273.15)) [68]
kd [min1] 1.7  104 [69]
1
Mwe [g  mol ] 288

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


4 www.MaterialsViews.com
ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Modeling the Semicontinuous Heterophase Polymerization . . .

www.mre-journal.de

(iii) the monomer mass balance (neglecting the monomer 3.3. Model Construction and Parameters’ Assessment
solubility in the water phase) can be written as:
To build the model, consider the particle evolution
mm ¼ mm0 þ Fm t  xMT Mw (6) [Equation (2)] and refer to the two homogeneous nucleation
terms, one by propagation and the other by radical
termination in the aqueous phase, followed by precipitation
Equation (6) is the definition of monomer conversion in
to produce a particle; if one considers that the radical
terms of mass. Here, mm and mm0 are the mass of monomer
concentration in the water phase lies between 109
present in the reactor at time t and its initial value (taken to
and 1010 mol  L1, that the monomer concentration in
be zero in our case), respectively; Fm the monomer feed rate
the water phase is Maq ¼ 2.5  103 M,[73] and that the
(g  min1), t the reaction time (min), x the conversion, and
propagation rate constant is about three orders of
MT (mol) is the total monomer fed. The total particle volume,
magnitude smaller than the termination one (see Table 2),
vp (mL) can be obtained assuming additivity, as follows:
then kpMaqRc–1  ktRRc–1 indicates that it is acceptable to
vp ¼ mm =rm þ mp =rp (7) neglect the homogeneous nucleation term due to radical
termination in the water phase in Equation (2).
where mp is the mass of polymer at time t. Therefore, the If a pseudo-SS is assumed for the aqueous radical
monomer concentration inside the particles Mp, under concentration, the particle evolution [Equation (2)] can be
starved conditions, can be obtained as: fitted to the experimental data for the parameters kcm,
MaqRc–1/Npmax, and kcNpmax/Nav, without solving the
Mp ¼ ðmm =Mw Þ=vp (8) equation for the average radical concentration
[Equation (3)] because, as stated in the previous section,
Following the differential step of the ID model assess- the conversion can be obtained from the function y(t)
ment approach,[70] the smoothed filtered[71] conversion mentioned above. Therefore, one can fit only Np using
 Equation (1 and 4) and y(t). This approach, which is an
data trend, y(t), and its derivative, y , are obtained in analytic
form, and Equation (1) is solved for the estimated average uncoupling technique for an equation set, yields a much
radicals per particle [ne(t)] along the reaction as: more robust and unique solution compared to solving the
entire equation set [Equation 1–4)] with five parameters.
 The resulting parameters as well as their statistical report
y MT Nav
ne ðtÞ ¼ (9) are presented in Table 3.
kp Mp zV
Similar to the approach used in the equation for Np , the
parameters appearing in the balance for the average
This estimated value is used as an additional (inferred)
number of radicals n [Equation (3)] can be obtained from
experimental measurement along the reaction. fitting the functions y(t), z(t), and ne(t) obtained before. The
The instantaneous average molar mass was estimated by
resulting parameters, along with their statistical report,
the following approximation:[72]
appear in Table 4. Therefore, Equation (2 and 3) can be
solved in an uncoupled fashion.
Rp
Mni ðtÞ ¼ Ri
Mw (10)
nNp þ Rtr
4. Results and Discussion
Here, Rp, Ri, and Rtr are the propagation, initiation, and
transfer rates, respectively. The initiation rate can be The proposed model was applied to the heterogeneous
written as Ri ¼ (2fkdmI)/V, where f is the efficiency of semicontinuous polymerization of BMA at 50 8C. Table 1
radicals generation, kd the initiator decomposition rate shows the final polymer content, final average particle size
constant (min1), and mI is the added amount of initiator (Dpmax), final particle number density (Npmax), surface
(mol). After simple algebraic manipulations, Equation (10) coverage ratio (r), and molar mass obtained for the different
becomes: monomer-feeding rates (Fm). The parameters used are
reported in Table 2. Particle number density, as well as
MT dx
dt the surface coverage ratio (r), which is the total surface
Mni ðtÞ ¼ ktr kp Mp nNp V
Mw
2fkd mI þ MT kp MT Nav
coverage by the available surfactant divided by the total
(11) surface area of the particles, were calculated from the
MT dx
dt
¼ Mw following formulae:[21]
2fkd mI þ Cm MT dx
dt

where Cm is the chain transfer coefficient (ktr/kp) and 6xMT Mw


Np ¼ (12)
Mn can be obtained applying the mean value theorem. prp D3p V

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


5
www.MaterialsViews.com ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
M. G. Pérez-Garcı́a, L. A. Pérez-Carrillo, E. Mendizábal, J. E. Puig, F. López-Serrano

www.mre-journal.de

Table 3. Estimated values for kcm, Maq Rc–1 Nav/Npmax, and kc Npmax/Nav, their standard deviation and overall R2, fitting Np using Equation (2),
(4), y(t), and z(t) for the three monomer addition rates.

Fm kcm Maq Rc–1 Nav/Npmax kc Npmax/Nav Overall Overall Npmax


[g  min1] [min1] [min1] [min1] [mL1]

Std dev Std dev Std dev R2 Std dev

0.2 4.266  105 1.237  107 4.104  103 0.999 6.336  103 1.89  1016
8.727  105 1.356  108 1.105  103
7
0.49 0 2.625  10 6.563  103 0.995 3.855  102 1.22  1016
8 3
1.074  10 1.412  10
7
0.73 0 3.019  10 1.975  103 0.999 1.221  102 1.03  1016
8 3
1.122  10 1.951  10

Table 4. Estimated values for kcp, and ke, their standard deviation and overall R2, fitting n using Equation (3), (4), y(t), z(t), and ne(t) for the
three monomer addition rates.

Fm [g  min1] kcp [min1] ke [min1] Overall R2 Overall


Std dev Std dev Std dev

0.2 3.856  101 13.521 0.684 1.950  102


4.965  101 16.61
0.49 5.526  101 4.914 0.769 5.639  102
8.141  102 5.916  101
1
0.73 1.181  10 3.996 0.946 1.823  102
9.217  103 6.097  102

 pD2p Np

S
  CMC as Nav particles (ne), obtained with Equation (9), are shown in
Vw Mwe Nav as
r¼ (13) Figure 2–4 as discrete data (circles) for graphical purposes
pD2p Np
only, although they are continuous. These results reveal
that the 0–1 hypothesis is valid because in all cases, ne was
Figure 1 depicts the monomer concentration inside the
particles calculated with Equation (8). This figure indicates,
as expected, that the monomer concentration inside the
particles becomes smaller as Fm decreases. The monomer
concentration decreases rapidly up to 40% conversion and
remains approximately constant thereafter. The initial
value corresponds to ca. 5 mol  L1, which is of the order of
magnitude of reported values for the maximum swelling in
conventional emulsion polymerization of BMA at 50 8C,
that range from 3.2 to 3.7 mol  L1.[73,74]
Figure 2–4 depict the model predictions and experi-
mental data (or inferred in the case of n) for the three-
monomer feeding rates, respectively, of the time evolutions
of conversion (x), average number of radicals inside the
particles (n), dimensionless average particle size (Dp ), and
dimensionless number density of particles (Np ), obtained
from the solution of the complete Equation set (1–8) with
the parameters depicted in Table 2–4. The role and meaning
Figure 1. Calculated monomer concentration inside the particles
of these parameters will be discussed below. [Equation (7)] neglecting the monomer water solubility.
To confirm first that the 0–1 hypothesis is valid, Fm ¼ 0.73 g  min1 (continuous line), Fm ¼ 0.49 g  min1 (dashed
the inferred average-number of radicals (circles) inside line), and Fm ¼ 0.20 g  min1 (dotted line).

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


6 www.MaterialsViews.com
ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Modeling the Semicontinuous Heterophase Polymerization . . .

www.mre-journal.de

Figure 2. Model results [Equation set (1–7)] applying values of Figure 4. Model results [Equation set (1–7)] applying values of
parameters from Table 3 and 4 (Fm ¼ 0.20 g  min1). Symbols are parameters from Table 3 and 4 (Fm ¼ 0.73 g  min1). Symbols are
experimental (or inferred in the case of ne) data. Lines refer to experimental (or inferred in the case of ne) data. Lines refer to
model predictions. model predictions.

always lower than 0.5, which is the maximum possible indicates that the estimation of the micellar nucleation
value in a 0–1 system. The evolution of the average radicals mechanism (kcm) has significant uncertainties (ca. 200%), as
per particle (ne), obtained by the ID method [Equation (9)], stated by its standard deviation; however, this nucleation
describes a slightly bell-shaped curve, especially for the 0.49 mechanism is only important at the slowest monomer feed
and 0.73 g  min1 addition rates (see Figure 2 and 3). rate because at faster feed rates, there is practically no
Table 3 lists the results for the parameter triplet: (kcm, Maq micellar nucleation (kcm ffi 0). The homogeneous nucleation
Rc–1/Npmax, and kcNpmax/Nav) and its statistical report term, involving the critical-minus-one radical concentra-
obtained from the fitted dimensionless number of particles tion (Rc–1), is approximately constant and much larger than
Np , the empirical functions, y(t) and z(t), and the monomer the micellar one, showing a slight increase with monomer
mass balance. This table shows that the overall fitting feed rate, with less than 10% estimation uncertainties.
was very good (0.995 < R2 < 0.999) with standard devia- The coagulation rate parameter appears to be somewhat
tions that were between 0.63 and 3.9%. This table also invariant with feed rate, with uncertainties lying between
20 and 100%. For this parameter, the uncertainty is highest
for the slower monomer feed rate.
The results of solving Equation (3) for the parameter pair
kcp–ke, using ne as an experimental (inferred) measure-
ment, along with the auxiliary functions y(t) for conversion,
z(t) for the particle number density and the parameters
listed in Table 3, are shown in Table 4, which also includes
their statistical report. The overall fit is not as good
as the one obtained in Table 3 (0.684 < R2 < 0.946), being
the largest for the fastest monomer addition rate. The
entrance of radicals to polymer particles parameter (kcp) can
be assumed to be approximately constant, considering that
the uncertainties (standard deviations) lie between 8 and
130%, being the largest again for the fastest feed rate. The
exit rate parameter and its standard deviation decrease
with increasing feed rates.
It is worth mentioning that uncoupling the equations
helped to solve the model system. When the obtained
Figure 3. Model results [Equation set (1–7)] applying values of
parameters from Table 3 and 4 (Fm ¼ 0.49). Symbols are exper-
parameters, listed in Table 3 and 4, were set as the initial
imental (or inferred in the case of ne) data. Lines refer to model guess in an overall regression, the solution led to unrealistic
predictions. values of kcp and ke. The reason is that for a 0–1 system, the

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


7
www.MaterialsViews.com ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
M. G. Pérez-Garcı́a, L. A. Pérez-Carrillo, E. Mendizábal, J. E. Puig, F. López-Serrano

www.mre-journal.de

Table 5. Summary of obtained parameter values.

Fm [g  min1] kcm [s1] Rc–1 [mol  L1] kc [m3  s1] kcp [s1] ke [s1]

0.2 7.11  107 1.53  109 3.66  1027 6.43  103 2.25  101
9 27 3
0.49
0 2.21  10 8.61  10 9.21  10 8.19  102
0.73
0 2.06  109 3.201027 1.97  103 6.66  102

SS value for the average radical number per particle is magnitude with previous works for the microemulsion
nss ffi kcp/(2kcp þ ke), and when ke >> kcp, which occurs in polymerization of styrene;[23,24] moreover, its value is much
our case (see Table 5), the solution is nss ffi kcp/ke, leading smaller than the rate of entry to particles (kcp).
to an infinite number of values for the kcp–ke pair that The value of the second-order coagulation rate constant
produce the same value of nss. This was the reason for obtained here is about nine orders of magnitude smaller
the equation set uncoupling technique presented above, than the reported values for rapid coagulation;[80,81]
where the true experimental behavior (conversion and nevertheless, its value is important and should be taken
particle number evolutions) was anchored by ad hoc into account for modeling purposes. The particle coales-
empirical equations. cence has been considered in semi-batch microemulsion
Figure 2–4 also reveal that the model accurately predicts polymerization and the values obtained here agree with the
the experimental overall conversion and the number ones reported for butyl acrylate.[82]
density of particles, even though the experimental data Figure 5 depicts the surface coverage ratio r, calculated
for the number density of particles exhibit a rather large with Equation (13), along the reaction. This figure indicates
dispersion, although evenly distributed. However, the that r diminishes from values larger than one, indicating
regression-based model (dotted line) adequately depicts that at early stages of the reaction there is enough
the overall trajectory of n in Figure 2–4. Due to the scatter, surfactant for particle stabilization, to values ca. 0.23 at
the statistical report (Table 4) is not as good as the one for later reaction stages due to the increasing number of
the parameters for the Np -evolution (Table 3). Thus, the particles. Another important finding of the present model is
entry and/or exit coefficients probably present a particle that r is around 1 at a conversion of ca. 0.25, and then this
size,[72–74] or vitreous effects state dependency. ratio decreases to values smaller than 1 (Figure 5),
Figure 2–4 also indicate that the model slightly under- indicating that the surfactant molecules do not cover the
predicts the experimental average particle size for the three whole particles’ surface, and that coagulation may occur.
monomer addition rates. Table 5 reports a summary of the The estimation of this ratio along the reaction can be useful
individual parameters obtained by the model shown for obtaining smaller particle size microlatex by finding
in Table 3 and 4, as lumped parameters. Results in this ways to diminish coagulation.
Table indicate that the assumption of disregarding the
homogeneous nucleation term due to water radical
termination and precipitation (ktRc–1RNav) was correct
[see Equation (2)], as confirmed by the low value of the
concentration of Rc–1, obtained by the model (see Table 5);
hence, one would expect that the total radical concentration
(R) would be of the same order as Rc–1.
Regarding the capture of radicals by micelles and
particles, some authors have assigned different values to
these two parameters.[25,75–78] The values for kcp coincide in
order of magnitude with the ones reported for seeded
emulsion polymerization of styrene,[27,61,79]) and for butyl
methacrylate.[70,71] However, it is not a simple task to
compare the coefficients of entry to micelles with literature
values because in some cases, the value is not reported
explicitly and some works calculate a second order
coefficient (i.e., kcm ¼ Rk0cm ). It is interesting to note in
Table 5 that the rate constant of radical capture by micelles Figure 5. Emulsifier coverage area ratio, calculated with
obtained by this methodology is non-zero only for the Equation (13); Fm ¼ 0.73 g  min1 (continuous line), Fm ¼ 0.49
lowest feed rate and that its value agrees in order of g  min1 (long dash), and Fm ¼ 0.20 g  min1 (short dash).

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


8 www.MaterialsViews.com
ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Modeling the Semicontinuous Heterophase Polymerization . . .

www.mre-journal.de

Figure 6 depicts that lower molar masses were obtained run and will increase with monomer addition rates as
as the monomer feed rate diminished, which are consistent observed.
with experimental data (Table 1). Equation (11) predicts Figure 7 shows the importance of the lumped terms
that molar masses are constant throughout the reaction appearing in Equation (2) (rate evolution of Np ) for the three
at a given Fm and that they decrease as the addition rate monomer-feed rates studied. This figure indicates that the
(or dx/dt) is slower. Unfortunately, the experimental data
of molar masses do not add any additional robustness
to solve Equation set (1–4) because, as evident from
Equation (11), none of the parameters appearing in our
Equation set is repeated in the equation for Mni. The molar
masses were obtained with the parameters listed in Table 2
and 5, fitting only the initiator decomposition efficiency ( f).
The results for the initiator efficiency factor ( f) were from
0.44, 0.87, and 1.12 as the monomer addition rates were
increased. Obviously, the efficiency cannot exceed the value
of 1.0. One possible explanation for a value higher than 1.0 is
that the decomposition initiator constant is higher than the
one used here (Table 2), but it can also be due to the fact that
absolute molar masses of the PBMA were obtained with the
Benoit correction and the Mark–Houwink parameters for PS
and PBMA (see Section 2). In addition, it has been reported
that the presence of SDS increases the KPS decomposition
rate.[83] However, the calculated decrease of f with
decreasing addition rate, predicted by the model, is realistic
because the initiator efficiency diminishes with increasing
reaction time, which becomes longer as the addition rate
diminishes. This is because the probability for a radical to be
terminated in the aqueous phase, before it enters a particle
or initiates a new particle, increases due to the lower
monomer concentration as monomer feed rate diminishes.
As evident in Equation (11), both the numerator and the
denominator (if the half-life time of the initiator is long
enough) are somewhat constant. Therefore, the average
molecular weight will be approximately constant for each

Figure 6. Number average molecular weight (Mn ) as calculated


integrating Equation (11), for each monomer addition rate: Figure 7. Magnitude comparison of the terms (min1) appearing
Fm ¼ 0.2 g  min1 (squares; continuous line), Fm ¼ 0.49 g  min1 in Equation (2), for each monomer addition rate: (a) Fm ¼ 0.2
(triangles; long dash), and Fm ¼ 0.73 g  min1 (circles; short dash). g  min1, (b) Fm ¼ 0.49 g  min1, and (c) Fm ¼ 0.73 g  min1.

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


9
www.MaterialsViews.com ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
M. G. Pérez-Garcı́a, L. A. Pérez-Carrillo, E. Mendizábal, J. E. Puig, F. López-Serrano

www.mre-journal.de

(dash-dot line), but just in a short time interval at earlier


polymerization times. This finding is consistent with recent
results of Wang et al.[25] where, for the polymerization
of MMA by semicontinuous addition with ammonium
persulfate as initiator, they reported that the main particle
formation mechanism was homogeneous nucleation.
Regarding the model by He et al., they only predicted
conversion and particle size and, a quantitative evaluation
for the homogeneous and micellar nucleation contributions
was not reported.[27,28] The coagulation term (dotted line) is
important in magnitude, indicating that coagulation occurs
and that it has to be taken into account. The dashed line
depicts that the term (Np =V)(dV/dt) decreases with reaction
time revealing that the particle concentration decreases
along the reaction due to monomer addition and particles
coagulation.
Figure 8 depicts the relative magnitude of the
lumped terms appearing in Equation (3) (average radicals
per particle rate of change). Here, it is shown that the
term that takes into account homogeneous nucleation
(kpMaqRc–1Nav Npmax/Np ) is very important at the beginning
of the reaction, diminishes asymptotically (dashed line)
along the reaction, but it does not vanish completely. The
particle radical capture [kcp(1–2n)] (solid upper line in
Figure 8) and the radical exit from particles (ke n) (solid
lower line) are the most important terms throughout the
whole reaction in all cases. The remaining terms appearing
in Equation (3), such as micellar nucleation [(kcm m)/Np ],
particle coagulation (2kcNpn2/Nav) and evolution of volume
change [(n/V)(dV/dt)], are very small compared to the
others and practically do not affect the average radicals per
particle rate.

5. Conclusion

A simple four-state model reproduced experimental data


for the SHP of BMA at several monomer-fed rates. The
trajectory of the average radicals per particle (ne) was
inferred, allowing an additional experimental measure-
ment with the aid of the integro-differential method,
and the assumption that the reaction behaves as a 0–1
system was validated. With auxiliary functions for the time
evolutions of conversion and particle density, an approach
to decouple the equation-set was applied, obtaining
Figure 8. Magnitude comparison of the terms (min1) appearing five parameters, three (kcm, Rc–1, and kc) from the particle
in Equation (3), for each monomer addition rate: (a) Fm ¼ 0.2 size evolution and two more (kcp and ke) from the average
g  min1, (b) Fm ¼ 0.49 g  min1, and (c) Fm ¼ 0.73 g  min1.
radical number evolution. The estimation of the critical-
minus-one radical concentration (Rc–1) made it possible to
definitely disregard the termination in the water phase,
most important term, in terms of its magnitude, is the proposed initially as an assumption. The rate parameters
homogeneous nucleation one, given by kpMaqRc–1Nav were of the same order of magnitude compared to the ones
(upper line). Only at the slowest addition rate (Figure 7a), reported in the literature. The entry rate coefficient to
a small contribution due to micellar nucleation appears micelles, when different from zero, was several orders of

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


10 www.MaterialsViews.com
ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Modeling the Semicontinuous Heterophase Polymerization . . .

www.mre-journal.de

magnitude smaller than that to particles. It was found Np (Npa/Npmax) dimensionless total particle concen-
that particle formation occurred mainly by homogeneous tration (–)
nucleation, which agrees with recent reports for differential N1 concentration of particles with one radical (L1)
microemulsion polymerization.[25] It was found by the Nav Avogadro’s number (mol1)
model predictions that particle coagulation occurred, which Npmax final value for Np in each run (L1)
could assist in further experiments to design conditions to R total radical concentration in water phase
obtain particles with tailor-made-sizes. The results pre- (mol  L1)
sented in this work provide one step toward the under- Rc–1 radical critical–1 concentration in water phase
standing of a complex polymerization process in which (mol  L1)
many phenomena are occurring. The model for SHP was rp particle radius (cm)
capable of describing rather accurately the trends of S surfactant in system (g)
conversion, particle number density and average particle t time (s)
size, average radicals per particle, and number average V reactor volume (L)
molar mass along the reaction. V0 initial reactor volume (L)
Vw water volume (L)
vp total particle’s volume (L)
6. Nomenclature y(t) empirical function describing conversion time
evolution
as surfactant coverage area (Å2) x overall conversion
aPS Mark–Houwink exponent z(t) empirical function describing particle density
aPBMA Mark–Houwink exponent time evolution
f initiator efficiency
Cm (ktr/kp) chain transfer coefficient (–)
Fm rate of monomer addition (g  min1)
kd initiator decomposition rate constant (s1) 7. Abbreviations
kp propagation rate constant (L  mol1  s1)
kt termination rate constant (L  mol1  s1) BMA n-butyl methacrylate
kcm capture by micelles rate constant (s1) HPLC high pressure liquid chromatograph
kcp capture by particles rate constant (s1) ID integro-differential
ke exit of radicals from particles rate constant (s1) KPS potassium persulfate
kc coagulation of particles rate constant PBMA poly(n-butyl methacrylate)
(L  mol1  s1) PS polystyrene
kPS Mark–Houwink constant QLS quasielastic light scattering
kPBMA Mark–Houwink constant SDS sodium dodecylsulfate
m micelle concentration (L1) SHP semicontinuous heterophase polymerization
mI initiator charge (mol) SS steady state
mm total monomer mass (g) THF tetrahydrofuran
mm0 initial monomer charged to the reactor (g)
M monomer concentration (mol  L1)
Mni instantaneous number polymer molar mass
(g  mol1) 8. Greek Letters
Maq monomer concentration in water phase
(mol  L1) rm monomer density (g  mL1)
Mp monomer concentration inside particles rp polymer density (g  mL1)
(mol  L1)
MT total moles of monomer (mol)
Mw monomer molecular mass (g  mol1)
Mwe surfactant molecular mass (g  mol1)
MPS molar mass Acknowledgements: Funds for this work, provided by UNAM
MPBMA molar mass (PAPIIT IN114212), CONACYT (Grant CB-2007-82437) and the
n average radical number per particle (–) interchange program between U de G and UNAM, are gratefully
acknowledged.
ne inferred average radical number per particle (–)
nagg aggregation micellar number (–) Received: February 24, 2013; Published online: DOI: 10.1002/
Np total particle concentration (L1) mren.201300104

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


11
www.MaterialsViews.com ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
M. G. Pérez-Garcı́a, L. A. Pérez-Carrillo, E. Mendizábal, J. E. Puig, F. López-Serrano

www.mre-journal.de

Keywords: differential; heterophase polymerization; modeling; [28] G. He, Q. Pan, G. L. Rempel, J. Appl. Polym. Sci. 2007, 105,
nanoparticles; semicontinuous 2129.
[29] M. Hwang, J. H. Seinfeld, AIChE J. 1972, 18, 90.
[30] N. Kalogerakis, R. Luus, AIChE J. 1983, 29, 858.
[31] G. Maria, Can. J. Chem. Eng. 1989, 67, 825.
[32] I. W. Kim, M. J. Liebman, T. F. Edgar, Comput. Chem. Eng. 1991,
15, 663.
[33] P. Bilardello, X. Joulia, J. M. LeLann, H. Delmas, B. Koehret,
[1] C. F. Lee, J. Appl. Polym. Sci. 2003, 88, 312. Comput. Chem. Eng. 1993, 17, 517.
[2] C. Larpent, ‘‘Colloidal Biomolecules, Biomaterials and Biome- [34] N. Baden, J. Villadsen, Chem. Eng. J. 1982, 23, 1.
dical Applications’’, Surfactant Series Vol. 115, A. Elaissari, Ed., [35] M. J. Liebman, T. F. Edgar, L. S. Lasdon, Comput. Chem. Eng.
M. Dekker, New York 2003, Ch. 7, p. 145. 1992, 16, 963.
[3] M. Rabelero, S. López-Cisneros, M. Puca, E. Mendizábal, [36] I. B. Tjoa, L. T. Biegler, Ind. Eng. Chem. Res. 1991, 30, 376.
J. Esquena, C. Soláns, R. G. López, J. E. Puig, Polymer 2005, [37] B. Van Den Bosch, L. Hellinckx, AIChE J. 1974, 20, 250.
46, 6182. [38] J. Villadsen, M. L. Michelsen, ‘‘Solution of Differential
[4] L. A. Pérez-Carrillo, M. Puca, M. Rabelero, F. López-Serrano, R. G. Equation Models by Polynomial Approximation’’, Prentice-
López, J. E. Puig, E. Mendizábal, Polymer 2007, 48, 1212. Hall, Inc., Englewood Cliffs, NJ 1978.
[5] J. K. Oh, D. I. Lee, J. M. Park, Prog. Colloid Polym. Sci. 2009, 34, [39] C. S. Adjiman, I. P. Androulakis, C. A. Floudas, A. A. Neumaier,
126. Comput. Chem. Eng. 1998, 22, 1137.
[6] C. Cannizo, S. Amigoni-Gerbier, M. Frigoli, C. Larpent, J. Appl. [40] R. P. Byrne, I. D. L. Bogle, Ind. Eng. Chem. Res. 2000, 39, 4296.
Polym. Sci. 2008, 46, 3375. [41] W. R. Esposito, C. A. Floudas, Ind. Eng. Chem. Res. 2000, 39,
[7] H. Sun, K. Almdal, T. L. Andersen, Chem. Commun. 2011, 47, 1291.
5268. [42] S. Lee, I. E. Grossmann, Comput. Chem. Eng. 2003, 27, 1557.
[8] C. C. Co, R. de Vries, E. W. Kaler, ‘‘Reactions and Synthesis in [43] V. Dua, K. P. Papalexandri, E. N. Pistikopoulos, J. Global Optim.
Surfactant Systems’’, J. Texter, Ed., Surfactant Sci. Series, Vol. 2004, 30, 59.
100, Marcel Dekker, New York 2001, Ch, 22. [44] L. Liberti, C. C. Pantelides, J. Global Optim. 2006, 36, 161.
[9] C.-S. Chern, ‘‘Principles of Emulsion Polymerization’’, Wiley, [45] A. B. Singer, J. W. Taylor, P. I. Barton, W. H. Green, J. Phys.
Hoboken N.J 2008, Ch. 6. Chem. A 2006, 110, 971.
[10] J. E. Puig, E. Mendizábal, F. López-Serrano, R. G. López, ‘‘Surfactant- [46] Y. Lin, M. A. Stadther, AIChE J. 2007, 53, 866.
Assisted Polymerization Processes’’, P. Somasundaram, Ed., [47] T. Park, G. F. Froment, Comput. Chem. Eng. 1998, 22, S103.
Encyclopedia of Surface and Colloid Science, 2nd edition Taylor [48] S. Katare, A. Bhan, J. M. Caruthers, W. N. Delgass,
and Francis, New York DOI: 10.1081/E-ESCS-120047408. V. Venkatasubramanian, Comput. Chem. Eng. 2004, 28,
[11] C. P. Yong, L. M. Gan, Adv. Polym. Sci. 2005, 175, 257. 2569.
[12] X. J. Xu, L. M. Gan, Curr. Opin. Colloid Interface Sci. 2005, 10, [49] M. Rodriguez-Fernandez, P. Mendes, J. R. Banga, BioSystems
239. 2006, 83, 248.
[13] M. Rabelero, M. Zacarı́as, E. Mendizábal, J. E. Puig, Polym. Bull. [50] J. M. Varah, SIAM J. Sci. Stat. Comput. 1982, 3, 28.
1997, 8, 695. [51] S. D. Conte, C. de Boor, ‘‘Elementary Numerical Analysis. An
[14] X. J. Xu, C. H. Chew, K. S. Siow, M. K. Wong, L. M. Gan, Algorithmic Approach’’, 3th edition, McGraw-Hill, New York
Langmuir 1999, 15, 8067. 1980.
[15] W. H. Ming, F. N. Jones, K. S. Fu, Macromol. Chem. Phys. 1998, [52] M. S. Varziri, A. A. Poyton, K. B. McAuley, P. J. McLellan, J. O.
199, 1075. Ramsay, Comp. Chem. Eng. 2008, 32, 3011.
[16] W. H. Ming, F. N. Jones, K. S. Fu, Polym. Bull. 1998, 40, 749. [53] C. Michalik, B. Chachuat, W. Marquardt, Ind. Eng. Chem. Res.
[17] N. Sosa, R. D. Peralta, R. G. López, L. F. Ramos, I. Katime, 2009, 48, 5489.
C. Cesteros, E. Mendizábal, J. E. Puig, Polymer 2001, 42, 6923. [54] O. Levenspiel, ‘‘Chemical Reaction Engineering’’, 2nd edition
[18] G. He, Q. Pan, G. L. Rempel, Macromol. Rapid Commun. 2003, Wiley, New York 1972.
24, 585. [55] H. S. Fogler, ‘‘Elements of Chemical Reaction Engineering’’, 4th
[19] G. He, Q. Pan, Macromol. Rapid Commun. 2004, 25, 1545. edition, Prentice Hall, New Jersey, USA 1999.
[20] R. Ledezma, M. E. Treviño, L. E. Elizalde, L. A. Pérez-Carrillo, [56] A. K. Galwey, Thermochim. Acta 2004, 413, 139.
E. Mendizábal, J. E. Puig, R. G. López, J. Polym. Sci., Part A: [57] F. López-Serrano, J. E. Puig, J. Álvarez, Ind. Eng. Chem. Res.
Polym. Chem. 2007, 45, 1463. 2004, 43, 7361.
[21] S. Sajjadi, Langmuir 2007, 23, 1018. [58] F. López-Serrano, J. E. Puig, J. Álvarez, AIChE J. 2004, 50, 2246.
[22] C. Norakankorn, Q. Pan, G. L. Rempel, S. Kiatkamjornwong, [59] F. López-Serrano, J. E. Puig, J. Álvarez, Ind. Eng. Chem. Res.,
Macromol. Rapid Commun. 2007, 28, 1029. 2007, 46, 2455.
[23] C. Norakankorn, Q. Pan, G. L. Rempel, S. Kiatkamjornwong, [60] D. A. Hutchinson, D. A. Paquet, J. H. McMinn, S. Beuermann,
J. Appl. Polym. Sci. 2009, 113, 375. R. E. Fuller, DECHEMA Monographs 1995, 131, 467.
[24] J. Aguilar, M. Rabelero, S. M. Nuño-Donlucas, E. Mendizábal, [61] P. Hiemenz, T. Lodge, ‘‘Polymer Chemistry’’, 2nd edition CRC
A. Martı́nez-Richa, R. G. López, J. E. Puig, M. R. Arellano, J. Appl. Press, Boca Raton 2007, Ch. 8, p. 369.
Polym. Sci. 2011, 119, 1827. [62] R. G. Gilbert, ‘‘Emulsion Polymerization: A Mechanistic
[25] H. Wang, Q. Pan, G. L. Rempel, Eur. Polym. J. 2011, 47, 973. Approach’’, Academic Press, London 1995.
[26] M. G. Pérez-Garcı́a, M. Rabelero, S. M. Nuño-Donlucas, [63] T. Staicu, M. Micutz, G. Cristescu, M. Leca, Rev. Roumaine
E. Mendizábal, A. Martı́nez-Richa, R. G. López, M. R. Arellano, Chim. 2008, 53, 481.
J. E. Puig, J. Macromol. Sci., A: Pure Appl. Chem. 2012, 49, 539. [64] A. F. Santos, E. L. Lima, J. C. Pinto, C. Graillat, T. McKenna,
[27] G. He, Q. Pan, G. L. Rempel, Ind. Eng. Chem. Res. 2007, 46, J. Appl. Polym. Sci. 2003, 90, 1213.
1682. [65] D. Li, N. Li, R. A. Hutchinson, Macromolecules 2006, 39, 4366.

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


12 www.MaterialsViews.com
ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!
Modeling the Semicontinuous Heterophase Polymerization . . .

www.mre-journal.de

[66] S. Beuermann, M. Buback, T. P. Davis, R. G. Gilbert, R. A. [76] I. A. Penboss, D. H. Napper, R. G. Gilbert, J. Chem. Soc. Faraday
Hutchinson, A. Kajiwara, B. Klumperman, G. T. Russell, Trans. 1984, 1(79), 1257.
Macromol. Chem. Phys. 2000, 201, 1355. [77] J. S. Guo, E. D. Sudol, J. W. Vanderhoff, M. S. El-Aasser, J. Polym.
[67] D. F. Sangster, J. Feldthusen, J. Strauch, Macromol. Chem. Phys. Sci., Part A: Polym. Chem. 1992, 30, 703.
2008, 209, 1612. [78] J. Herrera-Ordóñez, R. Olayo, J. Polym. Sci., Part A: Polym.
[68] S. Beuermann, M. Buback, Prog. Polym. Sci. 2002, 27, 191. Chem. 2000, 38, 2201.
[69] M. M. Breuer, A. D. Jenkins, Trans. Faraday Soc. 1963, 59, 1310. [79] V. M. Ovando-Medina, E. Mendizábal, R. D. Peralta, Polym.
[70] F. López-Serrano, J. E. López-Aguilar, E. Mendizábal, J. E. Puig, Bull. 2005, 54, 129.
J. Álvarez, Ind. Eng. Chem. Res. 2008, 47, 5924. [80] B. S. Hawkett, D. H. Napper, R. G. Gilbert, J. Chem. Soc. Faraday
[71] F. López-Serrano, J. E. López-Aguilar, E. Mendizábal, J. E. Puig, Trans. 1980, 1(76), 1323.
J. Álvarez, Macromol. Symp. 2008, 271, 94. [81] D. J. Lamb, C. M. Fellows, R. G. Gilbert, Polymer 2005, 46,
[72] G. Odian, Principles of Polymerization, Wiley, New York 2004. 7874.
[73] L. F. Halnan, D. H. Napper, R. G. Gilbert, J. Chem. Soc. Faraday [82] I. A. Maxwell, D. H. Napper, R. G. Gilbert, J. Chem. Soc. Fraday
Trans. 1984, 1(80), 2851. Trans. 1987, 83, 1449.
[74] J. L. Gardon, J. Polym. Sci., Part A-1 1968, 6, 623. [83] H. Holthoff, S. U. Egelhaaf, M. Borkovec, P. Schurtenberger,
[75] J. Ugelstad, F. K. Handsen, Rubb. Chem. Technol. 1976, 49, 536. H. Sticher, Langmuir 1996, 12, 5541.

Macromol. React. Eng. 2013, DOI: 10.1002/mren.201300104


13
www.MaterialsViews.com ß 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
R

Early View Publication; these are NOT the final page numbers, use DOI for citation !!

You might also like