Polimerización en Suspensión

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Chemical Engineering Science 60 (2005) 5574 – 5589

www.elsevier.com/locate/ces

Characteristic intervals in suspension polymerisation reactors: An


experimental and modelling study
Fatemeh Jahanzada , Shahriar Sajjadib , Brian W. Brooksa,∗
a Department of Chemical Engineering, Loughborough University, Loughborough, Leicestershire, LE11 3TU, UK
b Division of Engineering, King’s College London, London, WC2R 2LS, UK

Received 30 November 2004; received in revised form 15 April 2005; accepted 15 April 2005
Available online 29 June 2005

Abstract
Suspension polymerisation of methyl methacrylate was carried out as a model to elaborate on the evolution of particle size average and
distribution in the course of polymerisation. Four characteristic intervals in the evolution of particle size were identified as: transition,
quasi-steady-state, growth, and identification stages. The effects of stabiliser and initiator concentrations, monomer hold up, reaction
temperature, and agitation speed on the characteristic intervals, as well as the kinetics of polymerisation, were examined. The transition
stage, which has been totally ignored in the literature, was found to have significant effect on the evolution of particle size. The transition
stage is shortened by increasing the rate of polymerisation in the drops (either by increasing initiator concentration or using a higher
reaction temperature). Increasing the impeller speed and stabiliser concentration will also lead to a shorter transition period. However,
the delayed adsorption of the stabiliser on the surface of drops will prolong the transition stage. It is shown that the occurrence of the
quasi-steady state depends on the polymerisation conditions. The quasi-steady state occurs only if the balance between drop break up
and coalescence can be maintained. This requires a high rate of drop break up within a period of time during polymerisation (i.e., a low
rate of polymerisation in the drops by using a low initiator concentration and reaction temperature, a high agitation speed and a high
stabiliser concentration). The mechanisms underlying the growth stage are explained in terms of the overall rates of drop break up and
coalescence in the course of polymerisation reactions. It is also shown that the onset of growth stage cannot be defined in terms of a
critical conversation or viscosity, and it depends on the polymerisation conditions including mixing. The growth stage occurs if drops are
not sufficiently stable against both break up and coalescence. The onset of the growth stage is advanced with a decrease in the rate of drop
break up (e.g., decreasing agitation speed and stabiliser concentration). The growth stage can be totally eliminated from a polymerisation
process if dispersions with a static steady state can be formed. That requires a high concentration of stabiliser, or a low concentration of
monomer, to be used. A population balance model, which included the transition stage and the delayed adsorption of the stabiliser, was
developed that is capable of predicting the evolution of drop size in the suspension polymerisation.
䉷 2005 Elsevier Ltd. All rights reserved.

Keywords: Suspension polymerisation; Characteristic intervals; Break up; Coalescence; Particle size; Population balance modelling; Methylmethacrylate

1. Introduction distribution are two of key factors to guarantee the quality


of the product. In comparison with the number of studies
Suspension polymerisation is widely used in industry to appearing in the literature on the kinetics of polymerisa-
produce high value-added particulate products such as chro- tion and/or properties of final polymer particles, less atten-
matographic separation media, ion exchange resins and sup- tion has been paid to the study of evolution of particle size
ports for enzyme immobilisation. For all these products the and particle size distribution. The main reason for this is
average particle size and the breadth of the particle size perhaps the complexity of the phenomena that determines
particle size.
∗ Corresponding author. Tel.: +44 1509 222 510; fax: +44 1509 223 923. In principle, a balance between the rate of drop break up
E-mail address: b.w.brooks@lboro.ac.uk (B.W. Brooks). and coalescence determines the size of drops in suspension
0009-2509/$ - see front matter 䉷 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2005.04.063
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5575

polymerisation. Therefore, the drop size is a strong func- concentration of stabiliser in general, and PVA in partic-
tion of several parameters such as densities and viscosities ular, has been found to increase its ability to stabilise the
of continuous and dispersed phases, interfacial tension, type dispersion due to steric and Marangoni effects. In both sim-
and concentration of suspending agent, dispersed-phase hold ple liquid–liquid dispersion and suspension polymerisation,
up, type of impeller and stirring speed, as well as the kinetics a smaller mean drop/particle size and sharper drop/particle
of polymerisation. Suspension polymerisation reactions gen- size distribution have been obtained with increasing PVA
erally produce particles with a broad or bimodal size distri- concentration (Chatzi and Kiparissides, 1994; Konno et al.,
bution. Bead particles, which are the intended product from 1982; Lazrak et al., 1998).
a suspension polymerisation, with diameters in the range of It is always desired in industry to increase the output of
10 m to 5 mm, are usually accompanied with unintended manufacturing units at minimum cost. In suspension poly-
smaller particles. Many aspects of suspension polymerisa- merisation processes this target is usually met by increasing
tions have been reviewed in the literature (Brooks, 1990; the monomer hold up. However, the incorporation of more
Dowding and Vincent, 2000; Vivaldo-Lima et al., 1997). monomer destabilises the particles and also increases the
When two immiscible phases are brought into contact by possibility of thermal runaway. Therefore, a careful exam-
agitation, drops are formed and their size depends on many ination of interaction of parameters involved in suspension
variables including the agitation speed. An increase in the polymerisation reactors is required. Increase in monomer
agitation speed enhances the rate of drop break up and thus hold up enhances the dispersion viscosity and thus damps
favours the formation of smaller drops. However, at very the turbulence of the system, and also increases the colli-
high agitation speeds the drop size may increase with ag- sion frequency of dispersed drops. All these result in larger
itation speed due to an increase in the rate of coalescence polymer particles (Kalfas et al., 1993; Konno et al., 1982;
because of the very large surface area of drops and reduced Lazrak et al., 1998).
effectiveness of the suspending agent molecules on the in- One important aspect of monomer-hold up can be seen in
terface. A U-shape dependence of the mean drop size on association with monomer solubility in the water phase. Wa-
the agitation speed has been reported by several investiga- ter solubility of monomer is one of the factors that can affect
tors (Chatzi and Kiparissides, 1994; Johnson, 1980; Zhou the kinetics of the reaction and the quality of the product.
and Kresta, 1998). Yang et al. (2001) applied a variable When the solubility of the monomer in the aqueous phase
agitation-speed method during the suspension polymerisa- is negligible (as it is for styrene, for example) the kinetics
tion of styrene to limit the drop size enlargement during the of suspension polymerisation can be described as similar
growth stage and to control the final polymer particle size. to that of the corresponding homogeneous bulk polymeri-
One of the crucial factors in the control of particle size in sation. However, in the case of monomers with a moderate
suspension polymerisation reactors is the type and the con- to high water solubility, some of the monomer resides in
centration of the stabiliser(s) used. The majority of stabilis- the water phase and does not get involved in the polymeri-
ers used in this process are either water-soluble polymeric sation reaction. As polymerisation proceeds in the drops,
materials or inorganic particles. Suspending agents adsorb monomer transfers from the aqueous phase into the drops
at the monomer/water interface and, thus, enhance the sta- to replace the monomer that has polymerised. This effect
bility of drops against coalescence. Water-soluble stabilisers causes some deviations from the kinetics of bulk polymeri-
also facilitate the drop break up by reducing the interfacial sation. Kalfas et al. (1993) carried out a series of suspen-
tension. Partially hydrolysed poly vinyl acetate or poly vinyl sion homo- and copolymerisations of monomers with differ-
alcohol (PVA) is one of the polymeric water-soluble stabilis- ent water solubility. Their results showed that mass transfer
ers used in suspension and emulsion polymerisation pro- limitations in suspension polymerisation of partially water-
cesses. A number of studies have been reported on the sus- soluble monomers (methyl methacrylate, vinyl acetate and
pension polymerisation with PVA as stabiliser (e.g., He et al., acrylonitrile) at low monomer-to-water ratios is consider-
2002; Konno et al., 1982; Mendizabal et al., 1992; Zerfa able and leads to a lower final conversion in comparison
and Brooks, 1998). It has been shown that the best PVA with higher monomer-to-water ratios. Water solubility of
grade for using as a stabilising agent in suspension poly- monomer also causes the formation of emulsion particles in
merisation is the one with a degree of hydrolysis of 80–90% the aqueous phase. These particles are the result of the ho-
and molecular weight of above 70,000 (Castellanos et al., mogenous nucleation occurring in the water phase and be-
1991; Mendizabal et al., 1992). This grade of PVA forms a cause of segregation of radicals in these particles, the poly-
thicker and stronger layer on the water/monomer interface mer formed usually has a higher molecular weight than the
and has less tendency to be desorbed. Using PVAs with a polymer formed in larger particles of the suspension poly-
very low degree of hydrolysis (< 80%) or low molecular merisation (Cunningham, 1999).
weight results in coagulation and agglomeration of poly- The variation in reaction temperature will affect the prop-
mer particles and formation of polymer lumps, while using erties of both monomer and oil phases in a suspension poly-
PVAs with a very high degree of hydrolysis (> 90%) re- merisation. A 10◦ C temperature increase results in a two-
sults in unstable dispersions leading to formation of shape- to threefold increase in the rate of polymerisation (Odian,
less polymer bulk (Mendizabal et al., 1992). Increasing the 1991). The effects of temperature increase on the properties
5576 F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589

of the water phase containing a stabiliser are complex. These weighed and then quenched by adding cold acetone. The
mainly depend on how the stabiliser distributes between the dishes were kept in the vacuum oven at 50◦ C for at least
two phases at higher temperatures. It seems that for PVA 18 h. Monomer, acetone, and water were driven off by dry-
there is an overall trend of decreasing drop stability with ing to constant weight. Then the dishes were weighed again
increasing temperature, though the magnitude of reduction to calculate the monomer conversion using a simple mass
is quite different for different types of PVA (Lazrak et al., balance.
1998). It has been shown that this effect is not due to a de-
crease in viscosity of the continuous phase with increasing 2.2.2. Interfacial tension
temperature (He et al., 2002). Interfacial tensions between MMA and aqueous solutions
In this paper the characteristic intervals of a typical sus- of PVA were measured using a Du Nouy ring tensiometer.
pension polymerisation, in terms of average drop size, are Measurement was repeated at 25◦ C three times for each sam-
clarified. The effects of variations in process and formula- ple. The average value was taken as the interfacial tension.
tion variables on these intervals will be discussed. Methyl
methacrylate (MMA) has been selected as a “model”
monomer. A mathematical model was developed to elab- 2.2.3. Monomer concentration in the aqueous phase
orate on the kinetics of polymerisation as well as the A gas–liquid chromatograph (Pye Unicam 304-FID GLC)
evolution of drop size average and distribution. was used to determine the equilibrium MMA concentration
in the aqueous phase. A general-purpose capillary column,
wall coated-fused silica-CP-SIL5, was used together with a
2. Experimental flame ionisation detector (FID). The carrier gas was helium.
Methanol was used as the internal standard and the corre-
2.1. Materials and experimental set up sponding calibration curve was made using standard meth-
ods. Samples taken from the reactor at appropriate intervals
MMA (analytical grade from Aldrich) was distilled at re- were kept in capped test tubes and centrifuged for phase sep-
duced pressure to remove inhibitors. PVA (Mw = 85, 000 − aration. Samples from the aqueous phase were withdrawn
146, 000, degree of hydrolysis = 87–89%, from Aldrich) with a syringe and weighed. A known amount of a Methanol
and lauroyl peroxide (LPO) (97%, from Aldrich) were used solution was added to each sample before injecting to the
as stabiliser and initiator, respectively, without any further GLC. Three measurements were carried out for each sample
purification. Distilled water was used as the continuous to assure about the reproducibility of the measurement.
phase. The experiments were carried out using a 1-l jack-
eted glass reactor (internal diameter 10 cm) equipped with 2.2.4. Drop/particle size distribution (DSD/PSD)
four 90◦ baffles. A four-bladed flat turbine impeller with a A particle sizer (Coulter LS130), based on Fraunhofer
diameter of 5 cm was used for agitation. The temperature diffraction, was used for DSD/PSD measurements. Prior to
of the vessel content was controlled within ± 0.5◦ C of the any measurement, the sample cell of the instrument was
desired reaction temperature by pumping water with ap- thoroughly cleaned and filled with an aqueous solution of
propraite temperature through the jacket. Nitrogen purging sodium lauryl sulphate (0.2 g/l) to prevent coalescence of
was carried out for 30 min before the monomer phase was the droplets during measurement. At desired time intervals,
added to the aqueous phase. Samples were withdrawn, at a drop of the reactor content was immediately transferred
the desired time intervals, from the reaction vessel using to the sample cell. The cell content was mixed gently by a
a hypodermic syringe to measure the monomer conversion magnetic stirrer, provided in the cell, to prevent coalescence
and drop/particle size distribution. In order to avoid irrepro- of droplets during the measurement. Sampling and size mea-
ducibility, the sampling point was fixed at 2 cm away from surements were continuously carried out in the course of the
the impeller shaft and 3 cm below the suspension level. process.
Experiments (unless otherwise stated) were carried out at Drop size measurements required the dispersion to be
70◦ C, with an agitation speed of NI =500 rpm, the monomer highly diluted in water whereupon the monomer in the
volume fraction was d = 0.20 and PVA and LPO concen- monomer drops or unreacted monomer inside the particles
trations were 1.0 g/l (based on water phase) and 1.0 wt% may diffuse out of the particles. Water-solubility of MMA
(based on monomer phase), respectively. at 25◦ C is around 15 g/l. A simple mass balance shows that
dilution of samples with 102 times the volume of water
2.2. Measurements would be enough to transfer almost all MMA monomer out
of the drops, leading to their shrinkage. To avoid underesti-
2.2.1. Monomer conversion mation of drop size, the distilled water used for the dilution
Monomer conversion was measured gravimetrically. of the samples, was saturated with the monomer prior to
Small quantities of the dispersion were withdrawn from the preparation of the surfactant solution.
reactor and transferred into the small weighed aluminium The particle sizer Coulter LS130 has 84 channels (bins)
foil dishes. The samples, contained in the dishes, were which accommodate size ranges that change logarithmically.
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5577

These channels cover the size range of 0.43–822 m. The 100
output of the particle sizer is in terms of volume of drops
in each bin size, which can be easily transformed to volume
80
frequency (f ) defined as quasi identification
v(d) teady state
f (d) = d , (1) 60
v(d)

D 32 (µm)
max
dmin

where v(d) is the volume of drops with a diameter between


40
ln(d) and ln(d)+  ln(d). dmin and dmax are the diameters of
the smallest and the largest drops measured by the particle
sizer, respectively. The normalised distribution in terms of 20
diameter can be calculated using the following equation: transition growth

f (d) 0
F (d) = , (2)
d(d) 0 30 60 90 120
(a) time (min)
d(d) is the variable bin size between ln(d) and ln(d) +
 ln(d). It should be noted that the bin size is constant only
if a logarithmic scale is considered. F (d) has the dimension 1.E+07
of inverse length unit (m−1 ) and satisfies the normalisation
Rc
condition:
1.E+04
d
Rate (1/sec.cm3)
max
F (d) · d(d) = 1. (3) Rb
dmin
1.E+01
The Sauter mean diameter (D32 ) was calculated using the
following equation:
dmax 1.E-02
d f (d)
D32 = d min . (4)
dmin (f (d)/d)
max

1.E-05
0 30 60 90 120
3. Results and discussion (b) time (min)

3.1. Typical behaviour Fig. 1. Characteristic intervals in the evolution of mean drop size (a) and
theoretical predictions of overall rates of break up and coalescence (b) in
MMA suspension polymerisation at the basic conditions.
Before discussing the results in detail, it is helpful to con-
sider the characteristic intervals of a typical suspension poly-
Table 1
merisation, which were identified in this research (Jahanzad, Characteristics of different stages in a suspension polymerisation
2004) as the transition, quasi-steady-state, growth and iden-
tification stages. The size of drops in suspension polymeri- Stage Feature Result
sation, in principle, is determined by a balance between the Transition Rbreak-up  Rcoalescene Decrease in drop size
overall rates of drop break up (Rb ) and coalescence (Rc ). Initial drops formed
Depending on the magnitudes of these rates, different stages Quasi-steady Rbreak-up ≈ Rcoalescene Almost constant drop
can be observed in the course of polymerisation. Fig. 1a il- state size
Growth Rbreak-up
Rcoalescene Increase in drop size
lustrates typical behaviour of time evolution of drop size in Identification Rbreak-up = Rcoalescene = 0 Final particle size
suspension polymerisation. Fig. 1b shows the correspond- achieved
ing theoretical predictions of the overall rates of drop break
up and coalescence (per cm3 of the reactor volume) during
different intervals. The expressions used to calculate Rb and in size until they reach an almost constant average size. This
Rc have been given in Section 4 (Model Development). The stage has not received any detailed attention in the literature.
characteristics of each interval (or stage) are summarised in Quasi-steady-state stage: during this stage the rates of
Table 1. The significance of each stage is as follows. drop break up and drop coalescence are almost balanced
Transition stage: during this stage the drop size decreases leading to a quasi-steady-state drop size and distribution. At
exponentially and drop size distribution narrows dramati- the quasi-steady state, coalescence removes as many drops
cally due to a higher rate of drop break up in comparison from the dispersion as are made via the break-up mechanism.
with drop coalescence. During this stage drops are reduced Therefore, this steady state is often of a dynamic nature. The
5578 F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589

existence of this stage in suspension polymerisation has been 1


questioned by some investigators (Machado et al., 2000). 4.0 wt%
However, as we will show later, the occurrence of this stage 1.0 wt%
is not always possible and depends on the polymerisation 0.8
conditions.
Growth stage: during this stage the rate of drop break
0.25 wt% LPO
up falls considerably below the rate of drop coalescence 0.6

conversion
(Rc  Rb ) due to the viscosity build up in drops, leading to
the drop enlargement and size distribution broadening. This
stage has been traditionally called a “sticky stage” or “tacky 0.4
stage” in the literature perhaps because of the impression
that drop coalescence is boosted with the drop tackiness as
a result of increasing drop viscosity. However, the present 0.2
authors believe that this may be a misleading title for this
stage. This can be better explained if we look at the mecha-
0
nism of drop coalescence. Before two drops can participate
0 50 100 150 200
in a coalescence event, they have to approach each other,
time (min)
and must remain together long enough for the intervening
liquid film to drain to its critical thickness, whereupon film Fig. 2. Conversion-time variations for MMA suspension polymerisation
with different LPO concentrations.
rupture occurs. In fact, the rate of drop coalescence is con-
trolled by liquid drainage in the film between approaching
droplets, and more significantly, by the rigidity of the two variation with time is, in fact, a reflection of the rate of poly-
corresponding drop/water interfaces. Increasing the viscos- merisation inside the drops. Fig. 2 shows the conversion-
ity of dispersed phase in mobile-interface systems leads to time variations for the runs with different LPO concentra-
partial immobility of the interface and so decreases the prob- tions. The overall rate of polymerisation increased with LPO
ability of coalescence (Liu and Li, 1999). This implies that, concentration. Generally the onset of the gel effect, when
with increasing viscosity, particles are more stable against the rate of polymerisation increases significantly, depends
coalescence than in the earlier stages. Viscosity build up in on two factors; the volume fraction of polymer in the mix-
the polymerising drops diminishes both the rates of drop ture (or conversion) and the molecular weight of the poly-
break up and coalescence. With increasing drop viscosity, mer produced (Marten and Hamielec, 1979). Fig. 2 indicates
a point is reached where drops cannot be easily broken up that for the LPO concentrations of 0.25, 1.0, and 4.0 wt% the
but they can still undergo coalescence. As a result, drops onset of the gel effect happened at the conversions of 0.35,
start growing by coalescence. The occurrence of the growth 0.40, and 0.45, respectively. By increasing the LPO concen-
stage depends on the drop viscosity, drop stability and mix- tration shorter polymer chains were produced and the gel
ing conditions as will be discussed later. effect was postponed to higher conversions. The rate of the
Identification stage: this stage starts when the liquid–liquid polymerisation reaction during the gel effect increased sig-
dispersion fully transforms to a solid–liquid suspension nificantly with LPO concentration, as was expected. Gener-
with constant particle size average and distribution. At this ally, a higher final conversion was achieved with increasing
stage, drops have a very high viscosity and behave like solid LPO concentration.
particles so they cannot coalesce with each other and keep Fig. 3a illustrates the variations of drop Sauter mean di-
their identity for the remainder of the process. The onset of ameter (D32 ) with reaction time (t). This figure shows that
this stage has been called the “Identification point”. This the largest decrease in the average size of drops during the
point is mainly determined by the glass transition tempera- transition stage was obtained for the lowest LPO concentra-
ture of the reaction mixture, which is a strong function of tion used. This is due to a lower rate of viscosity build up in
polymer phase fraction in the particles, and is independent the drops containing a lower concentration of LPO. A high
of the mixing conditions. viscosity of drops, associated with a higher rate of polymeri-
Having clarified the characteristic intervals, in the follow- sation at high LPO concentration, damps the rate of drop
ing section we evaluate the effects of different process and break up so that the rates of drop break up and drop coales-
formulation variables on each stage. cence are balanced earlier. This also results in a larger drop
size in the beginning of the quasi-steady-state stage. During
this stage, drops underwent a slight size increase with time.
3.2. Initiator concentration Obviously, the quasi-steady state lasted longer for the lower
LPO concentration.
The kinetics of polymerisation can play an important role Note that time is not a good representative of the sta-
in the evolution of particle size and size distribution in tus of the polymerising drops when they contain different
suspension polymerisation. The viscosity of drops and its concentrations of the initiator. The evolution of D32 can
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5579

120 1

1.0 wt%
0.8
90 4.0 wt%

0.6

conversion
PVA (g/l)
D 32 (µm)

60 0.1

0.25 wt% LPO 0.4 0.2

0.5
30
0.2 1.0

4.0
0 0
0 50 100 150 200 0 30 60 90 120 150
(a) time (min)
time (min)
120
Fig. 4. Conversion-time variations for MMA suspension polymerisation
0.25 wt% LPO with different PVA concentrations.

90
400
0.1 g/ l
1.0 wt% 0.2 g/ l
D 32 (µm)

60
300

4.0 wt%
D 32 (µm)

30 200 0.5 g/ l

1 g/l
0 100
0 0.2 0.4 0.6 0.8 1 4 g/ l
(b) conversion

Fig. 3. Variations of Sauter mean diameter with time (a) and conversion 10 g/ l PVA
(b) for different LPO concentrations. 0
0 30 60 90 120 150
time (min)
be better explained if the comparison is made in terms of
monomer conversion as shown in Fig. 3b. This figure reveals Fig. 5. Variations of Sauter mean diameter with time for different PVA
that growth stage is delayed with increasing LPO concentra- concentrations.
tion. This is because low molecular weight polymer chains
are produced with increasing initiator concentration. As a
result of a low viscosity, the balance between drop break 3.3. Stabiliser concentration
up and drop coalescence can be maintained up to a higher
conversion. Fig. 4 shows the conversion-time variation for runs with
At low initiator concentrations, the rate of polymerisation different PVA concentrations. All the data points fall on the
is slow and polymerising drops are subject to more collision same curve confirming that PVA concentration does not af-
before the identification point can be achieved. As a result, fect the rate of polymerisation. This also indicates that pos-
the final size of particles obtained from the polymerisations sible grafting to PVA molecules does not have a significant
increased with decreasing LPO concentration. A recent work effect on the kinetics of polymerisation (Okaya et al., 1999).
on the effect of initiator concentration in MMA suspension Fig. 5 illustrates the variations of D32 with time for dif-
polymerisation with benzoyl peroxide as initiator showed ferent PVA concentrations. The rate of drop break up is a
similar results (Lazrak et al., 1998). strong function of interfacial tension, particularly during the
5580 F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589

16 0.03
1 min

20 min
12
interfacial tension (dyn/cm)

40 min
0.02

F (d) (1/µm)
65 min

8 70 min

0.01

0
0
0 50 100 150 200
0 2 4 6 8 10
diameter (µm)
PVA concentration (g/ l)
Fig. 7. Time evolution of particle size distribution in MMA suspension
Fig. 6. Variations of interfacial tension between MMA and aqueous PVA
polymerisation using 1.0 g/l PVA.
solution with PVA concentration at 25◦ C.

transition stage when the viscosity of drops is still very low. allow for a sharp reduction in the rate of drop break up with
As a result, a shorter transition stage and a lower steady-state increasing viscosity. Note that the poor coverage of drops
D32 were achieved with increasing PVA concentration. The with PVA, as well as delayed adsorbtion (discussed later in
difference, however, became less important with increasing Section 4.1), enhance the drop instability. These cause the
PVA concentration. This is confirmed by a diminishing dif- drops to start growing at a low conversion. The higher the
ference in the interfacial tension, with PVA concentration, PVA concentration, the less was the particle enlargement
as shown in Fig. 6. Values of interfacial tension in the reac- during the growth stage. Drops with higher stabiliser cover-
tor (at 70◦ C) will be slightly different from those shown in age are less susceptible to coalescence after a collision. As
Fig. 6. a result, a sharper rise in drop sizes is observed during the
The occurrence of a steady state in a suspension polymeri- growth stage at a lower PVA concentration. The results show
sation is not always certain, and its existence depends on that no critical conversion or critical viscosity can be defined
the polymerisation kinetics, mixing condition and the type to mark the onset of a growth stage. The critical conversion
and concentration of the stabiliser used. With the PVA con- or viscosity defined in the literature (Villalobos et al., 1993)
centration of 0.1 g/l, the steady state was never reached and for the onset of this stage, thus, may be only applicable for
drops started growing, soon after the minimum drop size was fixed polymerisation conditions and cannot be generalised.
reached, until the mass coagulation of particles occurred af- It is also clear from the results that the gel effect cannot be
ter 20 min. For the PVA concentrations of 0.2 and 0.5 g/l, the considered as the sole cause of the growth stage.
quasi-steady state lasted until about 30 and 50 min, respec- The onset of the identification stage was not affected by
tively, followed by the growth stage. For the PVA concen- PVA concentration implying that it is only determined by the
trations of 1.0 g/l and more, the polymerisations produced glass temperature of the reaction mixture. However, some
almost the same steady-state drop size until about 60 min. limited flocculation was observed at this point at lower PVA
Note that if drops are reduced to a size that cannot be fur- concentrations.
ther broken by agitation and if they are also sufficiently pro- Fig. 7 shows the time evolution of particle size distribu-
tected against drop coalescence, then drops do not undergo tion for the run with PVA concentration of 1.0 g/l. A small
any transformation during mixing and a static steady state is peak appeared in each distribution at the lower end of the
established. This implies that with increasing the stabiliser size range, with volume fraction of about 0.05, but those
concentration, the nature of the steady-state stage gradually peaks have not been shown in Fig. 7 for more clarity. Their
shifts from a dynamic one towards a static one (Jahanzad et significance has been discussed elsewhere (Jahanzad et al.,
al., 2004). Fig. 5 shows that the increase in drop size in the 2004). Fig. 7 shows that the drop size distribution initially
growth stage is very small for the run with the PVA concen- narrowed and that drop sizes changed to lower values with
tration of 10.0 g/l. time (transition stage). At later times, drop sizes showed a
The onset of the growth stage was found to be influenced quasi-steady state and then shifted to a higher values dur-
by the concentration of PVA. At a low concentration of PVA, ing the growth stage. The PSD remained unchanged after
the interfacial tension is sufficiently large (see Fig. 6) to 70 min when the identification point was reached.
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5581

200 1.0
220 rpm φd

0.025 0.05
0.8
150 300 rpm 0.10 0.20

0.30 0.40
0.6

conversion
D 32 (µm)

100
500 rpm
0.4
initial final
stage stage
50 inter-
0.2
mediate
stage

0.0
0
0 50 100 150 200 0 30 60 90 120
time (min) time (min)

Fig. 8. Variations of Sauter mean diameter with time in MMA suspension Fig. 9. Conversion-time variations for MMA suspension polymerisation
polymerisation at different agitation speeds. with different monomer volume fractions.

3.4. Agitation speed contact time. The net result, however, is an increase in the
rate of drop coalescence with agitation speed.
Three different agitation speeds were used. The results
indicate that agitation speed does not have a significant ef- 3.5. Monomer hold up
fect on the rate of polymerisation, as was expected (not
shown). However, the size of drops was significantly af- Suspension polymerisations of MMA were carried out
fected by the agitation speed. Fig. 8 shows the variations of with the agitation speed of 300 rpm using different monomer
D32 with time. Generally, the approach to the steady state volume fractions (d ). Fig. 9 shows the conversion-time
was shortened with increasing agitation speed as a result of variation for this series. The rate of polymerisation can
a more extensive drop break up. Similarly, the duration of be better described if the following conversion regions are
the steady-state stage was prolonged at high impeller speed. considered:
For NI =220 rpm, the steady-state stage did not exist and the Initial stage: the initial rate of polymerisation shows an
particle growth (growth stage) started just after a minimum increase with increasing monomer volume fraction. MMA
drop size was reached. The onset of the growth stage oc- has a water solubility of 1.6 wt% at 70◦ C. The monomer in
curred approximately at the conversions of about 0.20, 0.40 the dispersed phase saturates the water continuous phase be-
and 0.60, for NI = 220, 300, and 500 rpm, respectively. It fore much polymerisation can take place. This reduces the
can be inferred from the data depicted in Fig. 8 that the con- amount of monomer available in the monomer drops. Know-
version, as well as the viscosity, at which the particle growth ing that the locus of polymerisation in suspension polymeri-
started, are not constant and could vary with the impeller sation is, in fact, the monomer drops and assuming that the
speed. polymerisation of MMA in the water phase is negligible, the
Note that the rate of drop break up is a strong function of overall rate of polymerisation is expected to decrease with
the agitation speed. At a low agitation speed, an infinitesimal decreasing d . For d = 0.025, 60% of the initial monomer
increase in the drop viscosity will lead to an imminent reduc- is dissolved in the water phase and is absent from the locus
tion in the rate of drop break up so that the balance between of polymerisation. Obviously, the rate of polymerisation was
drop break up and coalescence could not be maintained fur- mostly affected for this case. Even with d =0.05, polymeri-
ther. This will result in a gradual enlargement of drops with sation started with 30% of the initial monomer charge in the
time. The balance could be better maintained at higher im- water phase, and so the rate of polymerisation was conse-
peller speed, due to a more extensive break up, which allows quently affected. With d values higher than 0.05, the initial
a quasi-steady state to form. It should be noted that the rate rate of polymerisation became less affected by d . It is worth
of drop coalescence increases with agitation speed to main- noting that the concentration of initiator in the monomer
tain the balance as more drops are formed. The drop–drop drops maybe greater than that indicated by the overall for-
collision frequency increases with the agitation speed, but mulation because of partial solubility of the monomer in the
the efficiency of coalescence decreases due to decline in the water phase. This may slightly compensate the fall in the
5582 F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589

0.03 300
φd
φd
0.025 0.05
0.20
0.10 0.20
0.30
0.02 200 0.30 0.40

D 32 (µm)
w MMA,aq

0.01 100

0.00 0
0 30 60 90 120 0 20 40 60 80 100
time (min) time (min)

Fig. 10. Variations of MMA concentration in the aqueous phase with time Fig. 11. Variations of Sauter mean diameter with time in suspension
in suspension polymerisations with two different MMA volume fractions. polymerisations with different MMA volume fractions.

rate of polymerisation due to presence of monomer in the similarly. For both monomer hold ups, MMA concentration
water phase. However, the rate of polymerisation is always in water gradually decreased from about 15 g/l, the equilib-
a decreasing function of d . rium solubility of MMA in water, by diffusion of monomer
Intermediate stage: this stage, which is manifested by the molecules from the water phase into the polymerising drops.
gel effect, appears at different conversions depending on The monomer transport occurs to maintain the thermody-
the d values used. For the higher values of d , the gel namic equilibrium between phases as a result of conversion
effect started at a conversion of 0.40, which corresponds to of monomer to polymer in the polymerising drops. During
the onset of gel effect for the MMA polymerisation in bulk the gel effect, when the rate of monomer conversion in-
at the same conditions. With decreasing d , the onset of creased significantly (after about 55 min corresponding to
gel effect is shifted toward a lower conversion. For d = the conversion of 0.40), a sudden drop in the MMA con-
0.025, for example, the gel effect started at the conversion centration in the water phase is observed. This is a result
of 0.20 because a substantial fraction of the monomer was of rapid monomer molecules emigration from the water
dissolved in the aqueous phase. It should be noted that the phase to the polymer/monomer drops in order to maintain
onset of the gel effect depends on the monomer conversion the equilibrium. As the conversion reached higher values,
inside the drops and not on the overall monomer conversion the MMA concentration in the water phase dropped to a
in the reactor. This means that for the MMA suspension value as low as 0.2 wt%. Monomer transport at this stage
polymerisations, it is expected that the onset of the gel effect was quite slow due to solidification of polymer particles.
appears at the particle conversion of 0.40, independent of Fig. 11 compares the variation in D32 with time for dif-
the values for the corresponding overall conversions. ferent values of d . Generally, the number of drop encoun-
Final stage: generally a lower final conversion was ters increases with monomer hold up. This, together with a
achieved for the lower d . This is in agreement with the lower surface coverage by the stabiliser, at a constant sta-
results reported by Kalfas et al. (1993). This effect is more biliser concentration, will boost the rate of drop coalescence
significant as d decreases to very low values. The limited and result in larger drops during the transition stage and the
conversion is a result of monomer partitioning between quasi-steady-state stage. It also gives a steeper rise in the
monomer drops and water phase. Though at high conver- drop size during the growth stage. The final size of particles
sions the monomer dissolved in the water phase diffuses increased with d , similar to what was shown by other in-
into the polymer particles and reacts there, but the diffusion vestigators (Kalfas et al., 1993; Konno et al., 1982; Lazrak
process is extremely slow. et al., 1998).
The monomer concentration in the water phase shows According to Fig. 11, the drop size showed only a small
simultaneous variation with that in the polymer particles. increase with time with lower values of d , but showed an
Fig. 10 shows the time variations in the MMA weight appreciable increase during the growth stage for the runs
fraction in the aqueous phase (wMMA,aq ) during the poly- with higher values of d . For the lowest values of d used,
merisations with d = 0.20 and 0.30. Both systems behaved 0.025 and 0.05, the conversion at which drop enlargement
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5583

1 180
80 C 80 C
70 C
150
0.8
70 C
60 C
120
60 C 50 C
0.6
conversion

D 32 (µm)
90 50 C

0.4
60

0.2 30

0
0
0 50 100 150 200 250 300 350
0 50 100 150 200 250 300 350
time (min)
time (min)
Fig. 13. Variations of Sauter mean diameter with time at different tem-
Fig. 12. Conversion-time variations for MMA suspension polymerisation
peratures.
at different temperatures.

occurred was around the value of 0.20, which corresponds been reported (Lazrak et al., 1998). According to this figure
to the onset of gel effect for these runs as mentioned earlier. the reduction in D32 during the transition stage appears to be
Except for the two lowest d used, the conversion range of more appreciable at the lower temperatures. This is because
the quasi-steady-state stage seems to narrow with increasing of lower rate of polymerisation, which allows drop break up
d . This implies that with an enhancement of the rate of to be operative for longer times. For the higher temperatures,
drop coalescence with d , the balance between drop break such as 80◦ C, the rise in drop viscosity depressed the rate
up and coalescence is perturbed and drops enter the growth of drop break up and caused the transition stage to become
phase earlier. The results suggest that the onset of the growth shorter by early balancing of the rates of drop break up and
stage is advanced with increasing d . For the highest d drop coalescence. As a result of a depressed rate of break
used, 0.40, particle coalescence occurred extensively at x = up, a higher quasi-steady-state drop size was achieved with
0.45 leading to the formation of some polymer lumps. The a temperature increase.
measurement of conversion for d = 0.40 was hampered by Fig. 13 shows that D32 increases slightly during the so
gross coagulation of particles. called quasi-steady-state stage for the lower temperatures
(50◦ C and 60◦ C) after a few percent conversion. However,
3.6. Reaction temperature the quasi-steady state continued for a longer period of time.
At these temperatures, the rate of viscosity build up in drops
Experiments were carried out at different temperatures with conversion is so high (because of high molecular weight
with the agitation speed of 300 rpm. The conversion-time polymer) that the rate of drop break up is extensively im-
variations for the runs are depicted in Fig. 12. It is evident paired as reaction proceeds. As a result, the rate of drop break
from this figure that the rate of polymerisation increases up falls behind that of coalescence. This leads to a constant,
with temperature. The onset of the gel effect was recorded at though small, rise in the size of drops. At higher reaction
210, 100, 55, and 32 min (corresponding to the conversions temperatures the quasi-steady state (constant drop size) was
of 0.30, 0.35, 0.40, and 0.45, respectively) for the polymeri- more observable in terms of conversion, though it was very
sations carried out at 50, 60, 70, and 80◦ C, respectively. The short in terms of time. The molecular weight of polymer,
onset of the gel effect occurred at lower conversions with de- and the polymer viscosity, is significantly reduced at high
creasing temperature because of the high-molecular weight temperatures. Note that this is a double effect because be-
polymer formed. The duration of the gel effect, however, sides formation of low molecular weight polymers with low
was shortened with increasing temperature. This implies that viscosity, polymers are less viscous at higher temperatures.
the viscosity increase during the gel effect will occur in a But the higher rate of monomer conversion at higher tem-
short period of time at higher temperatures (about 5 min for peratures disturbs the balance of break up and coalescence
80◦ C, for example). and shortens the quasi-steady-state stage.
Fig. 13 shows the variations of D32 with time for the runs For all runs, significant rise in the drop size was observed
at different temperatures. Generally the final particle size during the growth stage, in association with the gel effect,
increased with the reaction temperature. Similar results have where a massive increase in the drop viscosity occurred
5584 F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589

within a short period of time. The growth was less signif- distribution, which is considered a normal distribution. n(u)
icant for the lower temperatures and became steeper with is the number of drops formed by breaking a drop with
a temperature increase. If the stabiliser performance does volume u, which is assumed to be 2 (binary break up). The
not vary with temperature, then a greater drop growth is ex- calculated n(dv , t), is then transformed to volume probable
pected at a lower temperature because of a longer polymeri- density function, F (dv , t), using Eq. (2).
sation time that allows more coalescence. This is in contrast The rates of drop break up and coalescence were cal-
to the results obtained. Therefore, it seems that the growth of culated using the correlations introduced by Alvarez et al.
the drops at higher temperatures could be due to loss of sta- (1994) for suspension polymerisation of styrene (Eqs. (6)
bility of the polymerising drops. This implies that PVA be- and (7)). They included the effects of viscoelastic proper-
comes less effective in protecting polymerising drops against ties of polymerising drops on the rates of drop break up and
coalescence with increasing temperature, as has been sug- coalescence:
gested before (Lazrak et al., 1998). The reasons for the loss Rate of drop break up:
of particle stability at higher temperatures are not clear. In-
creased rate of chain transfer and grafting to PVA (Okaya b(v) = b (v) exp[−b (v)]. (6.1)
et al., 1999) and dissolution of PVA in the monomer phase
kb 1/3
at higher temperatures could be possible reasons. b (v) = 2/3
, (6.2)
(1 + d )dv

4. Model development b (v) = ab (dv ), (6.3)


6 1
To date, a number of mathematical models have been re- (dv ) = + , (6.4)
Re(dv )[1 + Re(dv )Ve (dv )] W e(dv )
ported for suspension polymerisation systems. These models
 
consider the effects of viscoelastic properties of polymeris- 1− 1
Ve (dv ) =  exp − − ,
ing drops on the rates of drop break up and coalescence 2Y0 Re(dv ) 12
(Alvarez et al., 1991,1994; Chen et al., 1999; Machado et al.,
2d
2000) as well as the nonhomogeneity of the turbulent flow Y0 = , (6.5)
(Maggioris et al., 1998, 2000; Vivaldo-Lima et al., 1998). d Ed dv2
A mathematical model was developed in this research us- 
= 1 − 48Y0 ,
ing the population balance modelling. The tank Reynolds
Y0
numbers for all conditions used were always above 15,000 =
(1 − )
which implies that a fully developed turbulent flow was dom-  

inant in the system. A homogeneous turbulent flow during (1 − )2 1−


× 1+ − exp − . (6.6)
the suspension polymerisation was assumed. A brief intro- 1+ 2Y0 Re(dv )
duction to the model is presented here. For details, readers
are referred to Jahanzad (2004). In the developed model the Rate of drop coalescence:
evaluation of drop volume distribution has been carried out
by numerical solution of the population balance equation c(v, u) = c (v, u) exp[−c (v, u)], (7.1)
(PBE) for dispersed drops, which is an integro-differential
1/3 2/3 2/3
equation (Eq. (5)), to predict the evolution of drop size dur- c (dv , du ) = kc (dv2 + du2 ) (dv + du ), (7.2)
ing the polymerisation. 1 + d
 dv max c (v, u) = ac (dvu ) − bc /W e(df ),
dn(dv , t)
= (/2)(6)2/3 v 2/3 b(u)(u, v)n(u) −1
dvu (v, u) = (dv−1 + du−1 ), df = (dv3 + du3 )/dvu
2
. (7.3)
dt dv
× n(du , t) d(du ) − b(v)n(dv , t) + v 2/3 where  and  are the frequency and the efficiency of
 dv/2
the break up (with subscript b) or coalescence (with sub-
× c(v − u, u)n(dv−u , t)n(du , t) d(du )
dv min script c).  is the average energy dissipation rate of the sys-
 dvmax−v tem. Re(d) and W e(d) are the Reynolds and Weber num-
− n(dv , t) c(v, u)n(du , t) d(du ), (5) bers of a drop with diameter d, respectively. Ed ,
d and
dv min
d are Young’s elasticity modulus, viscosity and density of
where n(dv , t) is the number of drops with diameter dv the dispersed phase, respectively. Ed is considered as a lin-
(and volume v) at time t, dv min and dv max are the minimum ear function of polymer–monomer conversion, Ed = xE p ,
and maximum drop diameters, respectively, which can exist where Ep is Young’s elasticity modulus of pure polymer
in the dispersion. b and c are the rates of drop break up (=3.2 × 1010 g/cm s2 , Ferry, 1980). kb , ab , kc , ac and bc are
and coalescence, respectively. (u, v) is the probability of adjustable parameters.
forming a drop with volume v from break up of a drop The method of moments was used for prediction of the
with volume u or, in other words, the daughter drop volume kinetics of MMA bulk polymerisation (Chiu et al., 1983)
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5585

and gel and glass effects were predicted by the correlations Table 2
proposed by Baillagou and Soong (1985) for bulk polymeri- Fitted values for the adjustable parameters of the model
sation of MMA. Monomer and stabiliser partitioning were Adjustable parameter Fitted value
modelled using the modified Flory–Huggins theory for poly-
kb 1.5
mer solutions (Kalfas and Ray, 1993) and the Langmuir-
ab 0.03
type interfacial adsorption (Adamson, 1976), respectively. kc (1/cm3 ) 3 × 10−6
Variations of drop viscosity with monomer conversion was ac 1 × 10−5
estimated using the correlation presented by Baillagou and bc 3.0
Soong (1985).
The set of differential equations for the moments of poly-
mer molecular weight have been solved by a Runge–Kutta 100
4th order method. The PBE (Eq. (5)) was solved numerically
1
for n(dv , t). The integro-differential PBE was transformed
into a system of differential equations by discretising the 0.8
80

θ/θeq
range of drop diameters according to the method of classes 0.6
(Chatzi and Kiparissides, 1992). The size range was divided 0.4
into a number of equally spaced intervals. Up to 100 points 60 0.2

D 32 (µm)
were used for the numerical integration. A Runge–Kutta 4th 0
order method was also used for solving the differential equa- 0 0.2 0.4 0.6 0.8 1
40 t/teq
tions over the time. The integral terms in the PBE were ap-
proximated using a composite Simpson’s rule. For an even
number of intervals Simpson’s 1/3 rule was applied, while 20
for an odd number of intervals Simpson’s 3/8 rule was em-
ployed. The initial condition, for solving the population bal-
ance equation, was assumed to be a normal distribution with 0
a D32 equal to the first experimental point which was gener- 0 30 60 90 120
ally available within 30 seconds after the start of the mixing. time (min)
All the physical constants and parameters used in the ki- Fig. 14. Model prediction of drop size evolution in a simple MMA–water
netic model (method of moments), for MMA free radical dispersion; solid and dashed lines are model predictions with and without
polymerisation, were extracted from the literature, mostly considering Eq. (8), respectively, at the basic conditions. The small graph
after Baillagou and Soong (1985). The adjustable parame- illustrates Eq. (8).
ters (kb , ab , kc , ac and bc ), used for calculating the rates of
drop break up and coalescence, were obtained for the best
fit to the experimental D32 -time data. The parameter kb in adjustable values were estimated. The fitted values for the
the break up frequency model (Eq (6.2)) determines the rate adjustable parameters are shown in Table 2.
of eddy-drop collisions leading to the drop break up. With
increasing kb the frequency of break up increases. The pa- 4.1. Delayed stabiliser adsorption
rameter ab in Eq. (6.3) indicates the efficiency of drop break
up. The higher the values of ab , the lower the required en- The model was initially checked for MMA–water disper-
ergy for drop deformation leading to a higher probability for sions in the absence of initiator. Experimental data for non-
drop break up. The parameter kc in Eq. (7.2) determines the reactive MMA-water dispersions have been reported else-
frequency of drop–drop collision. The collision frequency where (Jahanzad et al., 2005). The model predicted a very
increases with kc . The parameters ac and bc in Eq. (7.3) quick transition stage in comparison with the experimental
indicate the importance of coalescence efficiency. A higher data as shown in Fig. 14 (dashed line). It is known that the
ac and lower bc result in a higher coalescence efficiency steady-state drop size is independent of the magnitudes of
due to lower required energy for drop deformation and film the rate of drop break up and coalescence and only depends
drainage processes. on their relative value (Rb /Rc ) (Sovova, 1981). However,
In order to estimate the values of the adjustable param- the approach to steady state depends on the magnitudes of
eters for suspension polymerisation of MMA, a two-stage Rb and Rc . Mathematically, it is possible to slow down the
approach was used. We first applied the model to the D32 - approach to the steady state by decreasing the ratio of the
time experimental data of the non-reacting low-hold up rate of drop break up over that of drop coalescence without
MMA/water dispersion (d = 0.05) while ignoring coa- affecting the steady-state value of drop size. However, this
lescence. In this way, the estimated values for break-up approach is not justified because while it produces a good
adjustable parameters to give the best fit to the experi- fit for the transition stage, it cannot give a satisfactory fit
mental data were obtained. Then using these values for a to D32 -time data during the growth stage because of under-
higher dispersed phase hold up (d = 0.20), the coalescence estimation of the rate of drop break up. There are reports
5586 F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589

in the literature that the adsorption of PVA molecules on 1


the surface of drops in liquid–liquid dispersions is a slow 4.0 wt%
process (Lazrak et al., 1998; Mendizabel et al., 1992; Zerfa
and Brooks, 1998). Zerfa and Brooks (1998) and Mendiz- 0.8
1.0 wt%
abel et al. (1992) claimed that while the adsorption of PVA
might not be slow, the reconfiguration of PVA molecules
0.25 wt% LPO
on the surface of drops, required for a better protection of 0.6

conversion
drops, will take time. A simple empirical correlation (Eq.
(8)), based on the data of Lazrak et al. (1998), was devel-
oped to predict the evolution of instantaneous surface cov- 0.4
erage ( ) with time. The equilibrium time (teq ) within which
the equilibrium value ( eq ) is reached was taken as 30 min
(Lazrak et al., 1998). The Langmuir-type interfacial adsorp- 0.2
tion (Eq. (9)) was considered.
= eq (1 + 0.14Ln(t/teq )) for t  teq ,
(8) 0
= eq for t > teq ,
0 50 100 150 200
ka csa time (min)
eq = , (9)
1 + ka csa Fig. 15. Experimental (symbol) and theoretical (line) conversion-time for
where ka and csa are the partitioning coefficient and aqueous suspension polymerisation of MMA with different LPO concentrations.
phase concentration of the suspending agent, respectively.
In Fig. 14 the solid line shows the model prediction with the integration of this term gives the total rate of coalescence:
incorporation of delayed PVA adsorption. The small graph in  dy max   dv/2
Fig. 14 illustrates the variations of / eq with dimensionless Rc (t) = v 2/3 c(v − u, u)n(dv−u , t)
time (t/teq ) according to Eq. (8). dy min dv

min
The partitioning of PVA molecules and its effect on in-
terfacial tension ( ) and drop surface coverage ( ) can be × n(du , t) d(du ) d(dy ). (13)
calculated by the following equations along with Eq. (9):
cs = csa + a c∞ , (10) The above expressions were used to produce the predictions
shown in Fig. 1b.
= 0 − k , (11)
4.3. Model predictions
where a is the interfacial area per water volume unit, cs is the
total concentration of the stabiliser, c∞ is the concentration
Here a few examples of model predictions in compari-
of the stabiliser (mass/area) at = 1, 0 is the interfacial
son with experimental data are presented. Figs. 15 and 16
tension in the absence of the stabiliser and k is constant. We
compare the experimental and theoretical conversion-time
used the same values for ka and k as obtained by Lazrak
for MMA suspension polymerisation with different LPO
et al. (1998), using their experimental data for interfacial
concentrations and monomer volume fractions, respectively.
tension of the same system at 70◦ C (10.56 l/g and 10.40 for
These figures show a good agreement between the model
ka and k , respectively). The value of c∞ (=1×10−6 g/cm2 )
predictions and the experimental kinetics data. Fig. 16 shows
was estimated from the data reported by Van Den Boomgaard
the capability of the model to predict the kinetics of suspen-
et al. (1978).
sion polymerisation for monomers with partial water solu-
bility. The small graph in Fig. 16 compares the experimental
4.2. Estimation of Rb and Rc and model results for weight fraction of MMA in aqueous
phase during the suspension polymerisation with monomer
The overall rate of drop break up at time t, Rb (t), can be volume fraction of 0.2.
calculated by integrating the rate of break up for drops with Figs. 17 and 18 illustrate the experimental data and model
diameter dv at time t (the second term in the right hand side predictions of drop size evolution in suspension polymerisa-
of Eq. (5)) over the range of drop size of dv min to dv max : tion of MMA with different PVA and LPO concentrations,
 dv max respectively. As it is seen, the model can predict all four char-
Rb (t) = [b(v)n(dv , t)] d(dv ). (12) acteristic intervals and the effects of stabiliser and initiator
dv min
concentrations on these intervals. In Fig. 19 the experimen-
The third term of the right hand side of Eq. (5) gives the tal and model results for the evolution of PSD are compared
number of coalescence events at time t which form drops for one of the experiments. Again the agreement between
with diameter dv from coalescing smaller drops. Therefore the theoretical and experimental results is acceptable.
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5587

1.0 150
φd
0.025
0.8 0.20 0.25 wt% LPO
model
1.0 wt%
100
0.6
4.0 wt%
conversion

D 32 (µm)
0.4
0.03
WMMA,aq

50
0.2
0
0 120
time (min)
0.0
0 30 60 90 120
time (min) 0
0 50 100 150 200
Fig. 16. Experimental (symbol) and theoretical (line) conversion-time for time (min)
suspension polymerisations with different MMA hold ups. The small
graph shows the monomer weight fraction in water phase during the Fig. 18. Experimental (symbol) and theoretical (line) D32 for suspension
polymerisation. polymerisations with different LPO concentrations.

200 0.05

0.5 g /l PVA 40 min


0.04
150

0.03 0.5 min


F(d) (1/µm)
D 32 (µm)

100 final PSD


1 g/l
0.02

50 10 g/l
0.01

0 0.00
0 50 100 150 200 0 50 100 150 200
time (min) diameter (µm)

Fig. 17. Experimental (symbol) and theoretical (line) D32 for suspension Fig. 19. Experimental (symbol) and theoretical (line) evolution of particle
polymerisations with different PVA concentrations. size distribution for MMA suspension polymerisation with 1.0 g/l PVA.
The curves for 0.5 and 40 min and Final PSD represent approximately the
transition stage, quasi-steady-state, and identification stage, respectively.

5. Conclusions
showed that:
Suspension polymerisations can be characterised by up
to four intervals. They are: transition stage during which • The transition stage is shortened by increasing the rate
drops are reduced in size, quasi-steady-state stage with a of polymerisation in the drops. This could be done by
rather constant drop size, growth stage with a sharp increase increasing the initiator concentration or using a higher
in the size of drops, and identification stage where the full reaction temperature. Increasing the impeller speed and
transformation of drops to particles occurs. The existence, PVA concentration will also lead to a shorter transition
duration, and the boundaries of these intervals are subject to period. However, the delayed adsorption of the stabiliser
massive change with polymerisation conditions. The results on the surface of drops will prolong the transition stage.
5588 F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589

• The quasi-steady state occurs only if the balance between Ed Young’s elasticity modulus of
drop break up and coalescence can be maintained. This re- monomer–polymer mixture
quires that a high rate of drop break up should be operative Ep Young’s elasticity modulus of pure polymer
in the course of a polymerisation. This can be assured via f volume frequency, Eq. (1)
a low rate of polymerisation in the drops, i.e., low initiator F normalised distribution, cm−1 , Eq. (2)
concentration and reaction temperature, which facilitates ka partitioning coefficient, cm3 g−1 , Eq. (9)
drop break up. Alternatively, a high rate of drop break up kb , kc adjustable parameters of break up and
can be maintained by using a high agitation speed and collision frequencies, respectively
high stabiliser concentration. Low impeller speed and k constant in Eq. (11), dyn cm−1
low PVA concentration, for example, may remove the n average number of daughter drops formed by
quasi-steady-state stage completely and drops start grow- breaking of a mother drop
ing right after the transition stage. At a high stabiliser NI agitation speed, rpm
concentration, all stages overlapped so that the drops Rb , Rc overall rates of drop break up and
formed after the transition stage were maintained in the coalescence, respectively, s−1 cm−3
reaction medium and were gradually converted to poly- Re Reynolds number
mer beads with the least interaction with other particles t time, min
(static steady state). v drop/particle volume, cm3
• A growth stage occurs if drops are not sufficiently sta- Ve elasticity parameter, defined by Eq. (6.5)
ble against both break up and coalescence. The onset of We Weber number
the growth stage is advanced with a decrease in the rate wMMA,aq MMA weight fraction in the aqueous phase
of drop break up, i.e., decreasing stirring speed and PVA x conversion of monomer to polymer
concentration. The onset of the growth stage was delayed Y0 elasticity parameter, defined by Eq. (6.5)
with decreasing initiator concentration and temperature.
The growth stage can be totally eliminated from a poly-
merisation process if dispersions with a static steady state Greek letters
can be formed. That requires a high concentration of sta-
biliser, or a low concentration of monomer to be used. elasticity parameter, defined by Eq. (6.6)
• The onset of the identification stage depends only on  daughter drop volume distribution, Eq. (5)
the variables which alter the rate of polymerisation such  average energy dissipation rate, cm2 s−3
as temperature and initiator concentration. Mixing condi- instantaneous surface coverage of
tions do not play an important part here. drops, Eq. (8)
eq equilibrium surface coverage of drops,
A population balance modelling approach was used to model Eq. (9)
the suspension polymerisation of MMA. Predicted kinetics  elasticity parameter, defined by Eq. (6.6)
features and drop size behaviour were compared with ex- b , c efficiencies of drop break up and
perimental data. coalescence, respectively, Eqs. (6.3)
and(7.3)

d viscosity of dispersed phase, g cm−1 s−1
d density of dispersed phase, g cm−3
Notation
0 interfacial tension in the absence of the
stabiliser, dyn cm−1
ab , ac adjustable parameters of break up (Eq. (6.3)) d volume fraction of dispersed phase
and coalescence (Eq. (7.3)) efficiencies, b , c drop break up and collision frequency,
respectively respectively, s−1 , Eqs. (6.2) and (7.2)
b rate of drop break up, s−1 , Eq. (5)  efficiency parameter, defined by Eq. (6.4)
bc adjustable parameter of coalescence
efficiency, Eq. (7.3)
c rate of drop coalescence, s−1 , Eq. (5) References
csa stabiliser concentration in the aqueous
phase, g cm−3 Adamson, A.W., 1976. Physical Chemistry of Surfaces. Wiley, New York.
c∞ stabiliser concentration (mass/area) at Alvarez, J., Alvarez, J.J., Martinez, R.E., 1991. Conformation of the
= 1, g cm−2 particle size distribution in suspension polymerisation. The role of
kinetics, polymer viscosity and suspension agent. Journal of Applied
d drop/particle diameter, cm
Polymer Science 49, 209–221 (Applied Polymer Symposium).
dv max maximum drop size, cm Alvarez, J., Alvarez, J.J., Hernandez, M., 1994. A population balance
dv min minimum drop size, cm approach for the description of particle size distribution in suspension
D32 Sauter mean diameter, cm, Eq. (4) polymerisation reactors. Chemical Engineering Science 49 (1), 99–113.
F. Jahanzad et al. / Chemical Engineering Science 60 (2005) 5574 – 5589 5589

Baillagou, P.E., Soong, D.S., 1985. Free-radical polymerisation of methyl- Liu, S., Li, D., 1999. Drop coalescence in turbulent dispersions. Chemical
methacrylate in tubular reactors. Polymer Engineering and Science 25 Engineering Science 54, 5667–5675.
(4), 212–231. Machado, R.A.F., Pinto, J.C., Araujo, P.H.H., Bolzan, A., 2000.
Brooks, B.W., 1990. Basic aspects and recent developments in suspension Mathematical modeling of polystyrene particle size distribution
polymerisation. Makromolecular Chemistry, Makromolecular produced by suspension polymerisation. Brazilian Journal of Chemical
Symposium 35,36, 121–140. Engineering 17, 395–405.
Castellanos, J.R., Mendizabal, E., Puig, J.E., 1991. A quick method for Maggioris, D., Goulas, A., Alexopoulos, A.H., Chatzi, E.G., Kiparissides,
choosing a protecting colloid for suspension polymerisation. Journal of C., 1998. Use of CFD in prediction of particle size distribution in
Applied Polymer Science 49, 91–101 (Applied Polymer Symposium). suspension polymer reactors. Computers and Chemical Engineering 22
Chatzi, E.G., Kiparissides, C., 1992. Dynamic simulation of bimodal (Suppl. S), S315–S322.
drop size distributions in low-coalescence batch dispersion systems. Maggioris, D., Goulas, A., Alexopoulos, A.H., Chatzi, E.G., Kiparissides,
Chemical Engineering Science 47 (2), 445–456. C., 2000. Prediction of particle size distribution in suspension
Chatzi, E.G., Kiparissides, C., 1994. Drop size distributions in high holdup polymerisation reactors: effect of turbulance nonhomogeneity.
fraction dispersion systems: effect of the degree of hydrolysis of PVA Chemical Engineering Science 55, 4611–4627.
stabiliser. Chemical Engineering Science 49 (24B), 5039–5052. Marten, F.L., Hamielec, A.E., 1979. High-conversion diffusion-controlled
Chen, Z., Pauer, W., Moritz, H.U., Pruss, J., Warnecke, H., 1999. Modeling polymerisation of methyl methacrylate. American Chemical Society,
of the suspension polymerisation process using a particle population Symposium Series 104, 43.
balance. Chemical Engineering and Technology 22, 609–616. Mendizabal, E., Castellanos-Ortega, J.R., Puig, J.E., 1992. A method for
Chiu, W.Y., Carratt, G.M., Soong, D.S., 1983. A computer model for the selecting a polyvinyl alcohol as stabiliser in suspension polymerisation.
gel effect in free-radical polymerisation. Macromolecules 16, 348–357. Colloids and Surfaces 63, 209–217.
Cunningham, M.F., 1999. Microsuspension polymerisation of methyl Odian, G., 1991. Principles of Polymerisation. 3rd ed. Wiley, New York.
methacrylate. Polymer Reaction Engineering 7 (2), 231–257. Okaya, T., Suzuki, A., Kikuchi, K., 1999. Importance of grafting in the
Dowding, P.J., Vincent, B., 2000. Suspension polymerisation to form emulsion polymerisation of MMA using PVA as a protective colloid.
polymer beads. Colloids and Surfaces A: Physicochemical and Effect of initiators. Colloids and Surfaces A: Physicochemical and
Engineering Aspects 161, 259–269. Engineering Aspects 153, 123–125.
Ferry, J.D., 1980. Viscoelastic Properties of Polymers. 3rd ed. Wiley, Sovova, H., 1981. Breakage and coalescence of drops in a batch stirred
New York. vessel—I. Comparison of continuous and discrete models. Chemical
He, Y., Howes, T., Litster, J.D., Ko, G.H., 2002. Experimental study Engineering Science 36 (1), 163–171.
of drop-interface coalescence in the presence of polymer stabilisers. Van Den Boomgaard, T., King, T.A., Tadros, T.F., Tang, H., Vincent, B.,
Colloids and Surfaces A: Physicochemical and Engineering Aspects 1978. The influence of temperature on the adsorption and adsorbed
207, 89–104. layer thickness of various molecular weight fractions of poly(vinyl
Jahanzad, F., 2004. Evolution of particle size distribution in suspension alcohol) on polystyrene latex particles. Journal of Colloid and Interface
polymerisation reactions. Ph.D. Thesis, Loughborough University, Science 66 (1), 68–76.
Loughborough, UK. Villalobos, M.A., Hamielec, A.E., Wood, P.E., 1993. Bulk and suspension
Jahanzad, F., Sajjadi, S., Brooks, B.W., 2004. On the evolution of polymerisation of styrene in the presence of n-pentane. An evaluation of
particle size average and size distribution in suspension polymerisation monofunctional and bifunctional initiation. Journal of Applied Polymer
processes. Macromolecular Symposia 206, 255–262. Science 50, 327–343.
Jahanzad, F., Sajjadi, S., Brooks, B.W., 2005. Comparative study of Vivaldo-Lima, E., Wood, P.E., Hamielec, A.E., Penlidis, A., 1997. An
particle size in suspension polymerization and corresponding monomer- update review on suspension polymerisation. Industrial and Engineering
water dispersion. Industrial and Engineering Chemistry Research 44 Chemistry Research 36, 939–965.
(11), 4112–4119. Vivaldo-Lima, E., Wood, P.E., Hamielec, A.E., Penlidis, A., 1998.
Johnson, G.R., 1980. Effects of agitation during VCM suspension Calculation of the particle size distribution in suspension
polymerisation. Journal of Vinyl Technology 2, 138–143. polymerisation using a compartment-mixing model. The Canadian
Kalfas, G., Ray, W.H., 1993. Modeling and experimental studies Journal of Chemical Engineering 76, 495–505.
of aqueous suspension polymerisation processes. I. Modeling and Yang, B., Takahashi, K., Takeishi, M., 2001. Unsteady stirring method used
simulation. Industrial and Engineering Chemistry Research 32, in suspension polymerisation of styrene. Journal of Applied Polymer
1822–1830. Science 82, 1873–1881.
Kalfas, G., Yuan, H., Ray, W.H., 1993. Modeling and experimental studies Zerfa, M., Brooks, B.W., 1998. Experimental investigation of PVA
of aqueous suspension polymerisation processes. II. Experiments in adsorption at the vinyl chloride water interface in monomer suspension.
batch reactors. Industrial and Engineering Chemistry Research 32, Colloids and Surfaces A: Physicochemical and Engineering Aspects
1831–1838. 132 (2–3), 267–273.
Konno, M., Arai, K., Saito, S., 1982. The effect of stabiliser on coalescence Zhou, G., Kresta, S.M., 1998. Correlation of mean drop size and minimum
of dispersed drops in suspension polymerisation of styrene. Journal of drop size with the turbulence energy dissipation and the flow in an
Chemical Engineering of Japan 15 (2), 131–135. agitated tank. Chemical Engineering Science 53 (11), 2063–2079.
Lazrak, N., Le Bolay, N., Ricard, A., 1998. Droplet stabilization in high
holdup fraction suspension polymerisation reactors. European Polymer
Journal 34, 1637–1647.

You might also like