Você está na página 1de 13

Chemical Engineering Science 66 (2011) 5474–5486

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Controlled peroxide-induced degradation of polypropylene in a twin-screw


extruder: Change of molecular weight distribution under conditions
controlled by micromixing
P.D. Iedema a,n, K. Remerie b, M. van der Ham b, E. Biemond b, J. Tacx b
a
Universiteit van Amsterdam, the Netherlands
b
Sabic Europe, Geleen, the Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: Controlled degradation of polypropylene (PP) by peroxide was carried out in a laboratory twin-screw
Received 24 December 2010 extruder ZSK 18 and the change in Molecular Weight Distribution (MWD) was measured using Size
Received in revised form Exclusion Chromatography–Differential Viscosimetry (SEC–DV). The MWD results were compared to
6 May 2011
MWD predictions from a kinetic model developed and validated in earlier work (Iedema et al., 2001)
Accepted 14 June 2011
assuming ideal mixing. Clear deviations were observed – the measured MWD was broader – that could
Available online 11 August 2011
only be explained by unaccounted heterogeneity in the extruder. Incorporating the relatively narrow
Keywords: Residence Time Distribution (RTD) in the twin-screw extruder did not lead to MWD broadening. In
Polymer processing contrast, the exponential RTD of a Continuously Stirred Tank Reactor (CSTR) yielded a MWD widening
Reaction engineering
that was too extreme. A new micromixing model, based on the striation thinning model by Ottino
Mathematical modeling
(1980), was constructed partly based on Monte Carlo sampling using a monomer scission probability
Mixing
Polypropylene (Tobita, 1996). This model was adapted to the geometry of the extruder entrance and the peroxide feed
Twin-screw extruder practice consisting of introducing a few per thousand peroxide-rich PP particles among pure PP
particles. This micromixing model indeed allowed obtaining very good matches between measured and
modeled MWD. Under different experimental conditions with respect to initial PP quality and amount
of peroxide added, with a constant value for the striation thinning parameter the errors between
measured and predicted MWD were around 5%.
& 2011 Elsevier Ltd. All rights reserved.

1. Introduction become even more distributed by such macroscopic inhomogene-


ities. The same holds for inhomogeneities at a more microscopic
The controlled degradation of Polypropylene (PP) using per- level, as shown by our earlier publications on the controlled
oxides is extensively used in the Polyolefin industry to optimize degradation of Polypropylene in a Static Mixer Reactor (SMR)
both its processing capabilities and the end use properties desired (Fourcade et al., 2001a,b). Both macroscopic and microscopic
by their customers. At the same time processing conditions form inhomogeneities originate from the joint effect of flow and
an interesting situation for Chemical Engineering from a selectiv- diffusion, so attempting the detailed mathematical description
ity point of view, since fast reactions of peroxide radicals are of these phenomena using Computational Fluid Dynamics (CFD) is
paired to a relative slow diffusion of these highly reactive initiator one serious manner in solving the problem (McKenna et al.,
radicals through the polymer melt. It is interesting, since possibly 1998). However, this approach is still highly challenging as it is
the reactions proceed under spatially inhomogeneous conditions. faced with a number of serious problems, like the non-Newtonian
In the area of Polymer Reaction Engineering this problem has rheology of the polymer melt and the small scale of especially
been addressed for macroscopically inhomogeneous systems micromixing phenomena.
(Wells and Ray, 2005; Schmidt et al., 2005), where flow condi- Dealing with micromixing in the framework of CFD requires sub-
tions create compartments with different conditions. Distributed grid modeling, which has been applied to polymerization of low-
micro-structural properties such as the molecular mass, them- density Polyethylene (ldPE) using probability density functions to
selves possessing a certain spread in properties, are seen to describe the chemistry (Tsai and Fox, 1996; Kolhapure et al., 2005).
Accounting for micromixing phenomena has not frequently been
applied in Polymer Reaction Engineering. An early example using
n
Corresponding author. Tel.: þ31 20 525 6484; fax: þ31 20 525 5604. compartment modeling to describe the micromixing effect, again for
E-mail addresses: piet@science.uva.nl, p.d.iedema@uva.nl (P.D. Iedema). ldPE, is given in Marini and Georgakis (1984), and a more recent one

0009-2509/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2011.06.071
P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486 5475

by Nele and Soares (2002), who suggest micromixing to be the cause effect of RTD in a CSTR and a twin-screw extruder on the scission
of enhanced branching in metallocene-catalyzed Ethylene polymer- probability distribution is shown as a reference to compare the
ization. Micromixing has scarcely been applied to polymer degrada- micromixing results with. Finally, the effect of micromixing on
tion. Note that the mixing addressed in a recent paper on PP the MWD is presented. Proper values for the striation thinning
degradation in a twin-screw extruder (Suresh et al., 2008) is parameter a are obtained from the best fits of the MWD. These
concerned with macroscopic mixing. In a series of publications on prove the broadening of the MWD due to micromixing as
the controlled degradation of PP in a Static Mixer Reactor (SMR) we compared to ideal mixing.
did systematically address the micromixing phenomenon. There we
followed – as we will do in the present publication – the approach
developed by Ottino (1980) and Fields and Ottino (1987) using the 2. Experimental
concept of ‘‘striation thinning’’. According to it the complex
3-dimensional flow and convective diffusion problem is reduced Polypropylene with molecular weights Mw ¼370, 470, 590 and
to a one-dimensional problem with a limited number of parameters, 880 kg/mol and Mn ¼70, 105, 130, 245 kg/mole, respectively, were
like the striation thinning rate, a. At one point in the whole used as starting base powder obtained from Sabic Europe. This
procedure we ‘averaged’ the 3D-flow (Newtonian) as computed by starting powder material was of high purity and contained no
CFD and thus inferred the parameter a in a computational manner. additives. The base powder was compounded adding 0.2 wt%
In addition, a was experimentally determined by Laser Induced Irganox B225 as antioxidant and Luperox 802PP40 as peroxide
Fluorescence (LIF). When finally applied to experiments with con- to control the molecular weight. The stabilizer and Luperox
trolled PP degradation in the SMR we employed a as an empirical (concentrations of 0–0.5 wt%) were mixed with the pure PP base
parameter, obtained by fitting full MWD’s on measured ones (SEC- powder at SABIC-Europe using a Henschell mixer. This mixture
DV). Micromixing turned out to have a strong impact on the MWD was dosed to the hopper of a ZSK18 co-rotating twin-screw
shape, which deviates considerably from the shape obtained under extruder with a length/diameter ratio of 40. In all experiments a
ideal mixing conditions. throughput of 2.5 kg/h was maintained at a rotational speed of
The objective of this paper is to demonstrate the effect of 200 rpm. The temperature of the barrel is measured at various
micromixing on PP degradation in a twin-screw extruder. The places along the screw axis as shown in Fig. 4 and increases to
experiments performed in the ZSK 18 are closer to the industrial 240 1C. The temperature profile shown is assumed to be kept
practice than those with the SMR in our earlier work that has not during all experiments. The temperature of the polymer melt at
been used in industry for PP degradation until today. However, extruder exit was also measured and turns out to be slightly
here again a measurable effect of micromixing is not unimagin- higher: 242 1C. These temperatures were chosen to ensure high
able, since the addition of the peroxide happens in solid form by a peroxide conversion. The residence time distribution was mea-
few pp particles possessing a peroxide coating, which may lead to sured by introducing a PP grain with black colorant into the
locally high initiator radical concentrations (see Fig. 1). The extruder and measure the time of first coloring of the exiting melt
modeling approach is comparable to the one followed in our (td in Eq. (14)) and observing the color intensity change. Typically,
earlier work, although we replaced the deterministic approach to under experimental conditions the residence time varied from 37
compute MWD’s by an extremely fast Monte Carlo sampling to 70 s (see Fig. 7).
method (Tobita, 1996).
This paper is structured as follows.
First, a brief description of the laboratory extruder setup and 3. Modeling
the materials and measuring equipment used is given. The
modeling chapter starts with the kinetic equations used culmi- 3.1. Kinetics and ideally mixed batch reactor and CSTR
nating in the scission ‘density’ that is the central parameter of the
Monte Carlo scission algorithm. This parameter possesses a The kinetic model we use has been described in our previous
distribution of values according to the RTD in a Continuously publications. The reaction mechanisms involved are listed in Table 1,
Stirred Tank Reactor (CSTR). Then the way the laminar micro- the pre-exponential factors and activation energies, if available, are
mixing model is applied to the twin-screw extruder is explained listed in Table 2. In the previous work we employed a deterministic
in terms of a 3-dimensional description of the entrance section of population balance approach to compute the MWD. Here, however,
the extruder. Subsequently, the partial differential equations we decided to utilize a Monte Carlo sampling method, for reasons of
describing micromixing and reaction and their boundary condi- computational speed. The method has been introduced by Tobita
tions are presented. The solution of these equations provides the (1996). The method requires knowledge of the scission probability
profiles of scission probabilities across the striations, which of each monomer–monomer bonding (probably at the secondary
represent the input for the MC scission simulation procedure. radical sites, Camara et al., 2006), also called the scission density,
Section 4 starts with a presentation of a quantitative measure denoted by, rs. The Monte Carlo sampling procedure will be
of agreement between measured and predicted MWD. Then the explained for both a batch reactor and a CSTR. The latter reactor
type, as it has a much broader residence time distribution (RTD)
Initially selected monomer unit
Table 1
nleft nright Reaction mechanisms.

Reaction Equation

ns,left (from exp (-ρs)) > nleft ns,right (from exp (-ρs)) < nright Initiation kd
I!2I
no scission scission Recombination of initiator radicals kc
I þ I !IH
Fig. 1. Scission procedure. A linear chain is sampled from the weight distribution Hydrogen abstraction followed by scission kht n
Pn þ I !Pm

þ Pn-m þ IH
pw
n ; on this chain an arbitrary monomer unit is selected; lengths left and right are Transfer to polymer followed by scission ktr Um
compared to sampled values from the scission probability function psn ¼ exp(  rsn). Pn þ Pm !Pn þ Pml

þ Pl
Shortest pair of lengths decide whether scission event(s) took place and are added Disproportionation termination ktd
Pn þ Pm

!Pn þ Pm
to yield final length.
5476 P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486

Table 2
Kinetic coefficients.

Reaction Coefficient Pre-exponential factor Activation energy/R (K) Units

Decomposition Luperox 802PP40 kd 7.7  1015 18,600 s1


Recombination reaction initiator radicals kc 2  1017 6000 m3k mole  1 s  1
Hydrogen abstraction ki 2.7  105 – m3k mole  1 s  1
Transfer to polymer ktr 1.5  104 5998 m3k mole  1 s  1
Termination by disproportionation ktd 5  104 10 m3k mole  1 s  1

than a twin-screw extruder, might therefore seem a less relevant the latter equality using Eq. (2c). Finally, the scission probability
reference system, but we will nevertheless take it into consideration, now follows as the sum of two contributions, one from the
for it enables to clearly demonstrate the effect of RTD on the hydrogen abstraction and one from the transfer to polymer
molecular weight distribution. Before discussing the MC scission reaction, both supposed to be instantaneously followed by scis-
procedure, we will describe how the scission density is derived from sion (see Table 1). For the batch reactor it is given by the integral
the kinetic equations governing both ideally mixed reactor types. between 0 and the residence time in the extruder, t
The balance for the initiator concentration, with initiator feed Z t
concentration I2f and average residence time for a CSTR, tCSTR , rs ¼ ðkht I þ ktr l0 Þdt ð4aÞ
0
reads
As regards the CSTR we may formulate the average scission
dI2 1
¼ 2kd I2 þ ðI I Þ ð1aÞ probability in terms of the average residence time,tCSTR , as
dt tCSTR 2f 2
rs ¼ ðkht I þktr l0 ÞtCSTR ð4bÞ
Here, t represents the residence time of a fluid element in the
extruder. In a batch reactor without inflow and outflow the initiator Note that the manner in which the scission probability is used
concentration is seen to decay at a first-order rate according to in the sampling procedure CSTR requires a little more explana-
tion, to be given in a following section.
I2 ðtÞ ¼ I2 ð0Þexpfkd ðtÞtg ð1bÞ
I2(0) is the initial initiator concentration. The dissociation coefficient 3.2. Monte Carlo sampling to describe random scission
kd coefficient becomes a strong function of residence time, if the
temperature profile imposed is accounted for, due to the fairly large The procedure starts by sampling a primary polymer from the
activation energy for initiator dissociation (Table 1). In a CSTR at chain length weighted length distribution of PP, pw n . This is
steady-state Eq. (1a) yields for the initiator concentration equivalent to sampling one monomer unit out of all monomer
I2f units present in polymer chains. Then, an arbitrary monomer unit
I2 ¼ ð1cÞ
1 þ kd tCSTR on this chain is selected and the segments to the right and left of
this unit, having lengths nr and nl, are checked for an eventual
The change of initiator radicals concentration follows as the
scission point (Fig. 1). In order to detect scission events the
dynamic balance between production through initiator dissocia-
lengths at RHS and LHS are to be compared to lengths from a
tion and consumption by the hydrogen abstraction reaction:
scission length distribution that is based upon the scission prob-
dI 1  ability, rs. The probability that the unit adjacent to the initially
¼ 2kd I2 kht I m1 kc ðI Þ2  I ð2aÞ
dt tCSTR selected one has not undergone scission and hence is connected
For the batch reactor this yields equals (1  rs); for the subsequent one it is (1  rs) (1 rs); and
for the ns-th one to be connected the scission length distribution

dI function is
¼ 2kd I2 kht I m1 kc ðI Þ2 ð2bÞ
dt
ps ðns Þ ¼ ð1rs Þns rs  expðrs ns Þ ð5Þ
Here, all quantities are functions of residence time, except for the
total number of monomer units in polymer chains, m1, which Eq. (3) expresses the probability that given the scission
remains unchanged. Also, if the batch reactor operates non- probability rs, the length of a scission fragment at RHS or LHS is
isothermally, kinetic coefficients are a function of residence time. exactly ns. Note that this is a chain length weighted distribution,
In the CSTR case we gets since it is based on sampling of monomer units. Furthermore, the
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi distribution of fragments according to Eq. (3) would be found in

kht m1 þ1=tCSTR þ ðkht m1 þ 1=tCSTR Þ2 þ 4kc kd I2 practice, if scission would occur to linear chains of infinite length.
I ¼ ð2cÞ
2kc In the present case of PP with a given MWD undergoing scission,
length values sampled from the exponential distribution of Eq. (5)
The macroradical concentration results from production by
are compared to the RHS and LHS lengths as obtained in the above
hydrogen abstraction and consumption by disproportionation
described manner. If the scission length is shorter than the chain
termination
part at a certain side, scission was a real event, and the length of
dl0 1 that chain part assumes the shorter value. If it is longer, no
¼ kht I m1 ktd ðl0 Þ2  l , ð3aÞ
dt tCSTR 0 scission took place, and the length of the fragment at that side
which in the batch case gives retains its original value. Thus, a population of chain lengths
sampled from the original length distribution is processed yield-
dl0
¼ kht I m1 ktd ðl0 Þ2 ð3bÞ ing a population of chains having undergone scission.
dt Now, the procedure for a CSTR is slightly more complicated
and for the CSTR than that for the batch reactor, since we have to account for the
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi residence time distribution of individual chains in the former
kht I m1 reactor type. Chains with a longer residence time are longer
l0 ¼ ð3cÞ
ktd exposed to the conditions causing scission. The scission probability
P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486 5477

as expressed by Eq. (4a) is also valid for a CSTR, but as the (macro- diffusion equation. At Sx scales, the evolution of the concentration
)radical concentrations I and l0 now have assumed constant of a species, s, while following Sx in the flow (Lagrangian
values, so the scission density becomes a linear function of description), is given by the convection diffusion equation with
residence time reaction source term
Z t
rs ðtÞ ¼ ðkht I þktr l0 Þdt ¼ ðkht I þ ktr l0 Þt ð4cÞ @Cs @Cs @ 2 Cs
0 aðX,tÞx ¼ Ds 2 þ Rs ð6Þ
@t @x @x
Thus, instead of the constant value for the scission probability
in a batch reactor, we have a distribution of scission probabilities where x is the direction perpendicular to the striations, Cs is the
according to the distribution of residence times, t, which in the concentration of the species, Rs is the rate of production of the
case of a CSTR has an exponential shape species from chemical reaction, Ds is the diffusion coefficient, and
a is the rate of striation evolution (thinning in the case of mixing)
Pðrs Þ ¼ ðkht I þ ktr l0 Þt expðt=tCSTR Þ ð4dÞ in Sx. The lamella-mixing model employs non-adjustable para-
This implies that the MC scission procedure for the CSTR meters that have a real, physically clear meaning to describe the
requires an extra sampling step: that from the residence time basic interplay between diffusion, reaction and hydrodynamics at
distribution, yielding a value for t and subsequently one for rs. small scales. The chemical source term is obtained from the
Note that sampling of t happens independently from sampling of scission model.
the length, n. Furthermore, the value obtained for rs holds for the As regards the application of the micromixing model to the
length comparisons at both RHS and LHS. The MC scission situation of molten Polypropylene in a twin-screw extruder we
procedure for batch reactor and CSTR proceeds very fast; a employ the same principle as in our previous publications con-
population of 1 million chains takes around 1 s of CPU-time. cerning the Static Mixer Reactor. Most importantly, this implies
The method is compared with the deterministic procedure based that diffusion is supposed to be limited to the low-molecular
on the Galerkin-hp method of PREDICIs (Wulkow, 1996) and species, initiator and initiator radicals, the Cs in Eq. (6). Polymer
identical results were obtained. liquid layers are assumed to be deforming and thinning, but
diffusion of polymer chains in and between striations is thought
to be negligible. The previous section has shown that the scission
3.3. Application of laminar micromixing model to twin-screw probability rs is the key variable determining scission, so the
extruder most important question to be answered in this section is how
micromixing affects this variable. Eq. (4a) expresses the scission
The melting of solid particles under the conditions at the probability in terms of (macro-)radical concentrations I and l0,
extruder entrance is an extremely complex problem to describe hence mathematically speaking one desires the solution of the
exactly. The micromixing model as we will discuss in this section micromixing equation (6) for these two species. The remainder of
is a simplification of the real process. This model must be this section, therefore, is devoted to defining the parameters and
regarded as an engineering approach to quantitatively describe boundary conditions involved in Eq. (6) and subsequently solving
the competition between a fast dissociation reaction and relatively the differential equation.
slow diffusion of low-molecular compounds. We do not claim that it According to the lamellar micromixing model, mixing is
is a correct 3D description of the rheologically complex flow described as the one-dimensional convection and diffusion in a
around a melting and deforming object. striation layer of decreasing thickness. Hence, we have to define
Laminar micromixing approaches such as developed by Ottino the starting situation of a striation as related to the 3D-situation
(1980) and Baldyga et al. (1998) employ a Lagrangian framework at the extruder entrance; it is depicted in Fig. 3. In the experi-
for the description of simultaneous deformation and diffusion in mental twin-screw extruder peroxide is introduced as peroxide-
fluid elements. In the lamella-mixing model, the flow is seen as a rich polymer grains with average diameter dp and peroxide
population of microflow elements Sx, in which the flow field is weight fraction wp(0). The initial values of the parameters
homogeneous. In each of these microflow elements the mixture describing the striation, its thickness, s0, and the fraction of it
has a lamellar structure, which is described by so-called striations having this high concentration of diffusing species, d, are related
with a certain thickness (Fig. 2). This approach enables to describe
the complex folding and deformation process in 3 dimensions
(x1, x2, x3 in Fig. 2) in terms of a 1-dimensional (x) convective

x3
x1
x2
SX

Fig. 2. Microflow element with an ideal lamellar structure of the mixture and Fig. 3. Starting point of a striation as related to the 3D-situation at the extruder
definition of striation thickness, s. entrance.
5478 P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486

to the real geometrical situation as shown in Fig. 3. First, one added in the micromixing model is implemented by changing the
should realize that the fraction d equals the ratio between the initial striation thickness s0, reflecting that the average distance
volume flows of peroxide-rich polymer grains and bulk polymer between peroxide layers is the manipulated quantity.
grains. Thus, the overall weight fraction averaged over all grains For reasons of computational efficiency in solving the differential
follows as: equations given by Eq. (6) we have performed two variable
transformations. The first one was already proposed by Ottino
wp ð0Þ ¼ dwp ð0Þ ð7Þ
(1980). It consists of transforming the space variable x into a
Now, Fig. 3 shows that we may consider the micromixing dimensionless variable x at a scale between 0 and 1 based on the
process around each peroxide-rich particle, while traveling actual value, s, of the striation thickness. In addition, the time
through the extruder, to take place inside a sphere around it, variable t is transformed into a new variable, t, called ‘warped
having a diameter equal to the average separation distance time’. The second transformation is meant to increase the resolution
between the particles. Note that the concentration development inside the thinning striation in the most interesting, peroxide-rich
in the expanding spheres is the result of diffusion and convection side by implementing a finer numerical solution grid for dimension-
in deforming and folding fluid elements as described by the less space, x. The two transformations of Eq. (6) result into a set of
lamellar micromixing model. With this picture in mind we decided coupled ordinary differential equations (ODE). As the procedure is
to take half the inter-particle separation as the initial striation practically identical to that of our previous publication (Fourcade
thickness, s0. Realizing that the ratio between the volumes of the et al., 2001b), we summarize it in the Appendix.
particles and the surrounding spheres again is expressed by d, we
find the value for the inter-particle separation and hence s0 to be 3.4. Monte Carlo scission procedure in the micromixing model
related to the power  1/3 of this fraction

s0 ¼ d
ð1=3Þ
dp =2 ð8Þ Solving the micromixing equation (6) through its numerical
implementation as described in the Appendix leads to solutions
For instance, if the particle diameter dp equals 0.75 mm, with a for the profiles in time and space of the radical concentrations
volume (flow) fraction d ¼2.5  10  3, the average inter-particle I and l0. This is the outcome of the concept that PP chains at
separation becomes 5.5 mm and s0 ¼ 2.8 mm. It should be noted initially different distances from the peroxide-rich regions experi-
that the extent of degradation is controlled by changing the ence a different history of being exposed to these scission
average weight fraction, wp ð0Þ, which in practice simply is carried enforcing radicals. Numerically, this heterogeneity is expressed
out by adjusting the number of peroxide-rich particles in the feed by striation nodes, i, each with different time profiles of I and l0
stream. This implies that while varying the average weight and consequently, scission probability rs. In line with Eq. (4d) this
fraction, wp ð0Þ, the initial particle weight fraction, wp ð0Þ, remains leads to different chain length distributions after scission as
unchanged. However, changing the number of particles does alter computed using the MC scission procedure based on the scission
the volume fraction, d, and hence the initial striation thickness. density as the key variable. Accounting for the proportions of the
Increasing the average weight fraction thus corresponds to a different striation regions (Eq. (B.11) in Appendix), this leads to an
decrease of inter-particle distance, which under these circum- overall chain length distribution, which is the one to be compared
stances implies a better mixing situation. More general, the with the experimentally measured distributions.
adopted concept as described above is logically in line with the
obvious fact that more and finer grains corresponds to better
mixing. 4. Results
Next, the solution of Eq. (6) under proper boundary conditions
is discussed. These apply to the diffusing species, i.e. the peroxide In the previous section we have presented models describing
species, I, and the peroxide radical species, I . At the low- scission of PP for different mixing situations. In this section we
concentration part of the striation, x ¼s, a no-flux boundary want to compare the results of these models in terms of
condition is imposed Molecular Weight Distribution to those obtained from experi-
ments with the ZSK-18. This requires defining a quantitative
@I @I
¼ 0; ¼0 ð9Þ manner to rate how well a computed MWD matches to a
@x @x
measured one. Furthermore, one of the questions to be answered
The boundary conditions at the high-concentration side of the is which mixing regime is applicable to the twin-screw extruder
striation, x ¼0, are obtained from mass balances over the volume conditions. We have to describe the strategy to determine this
fraction d of peroxide-rich polymer, in which a diffusion flux term and consequently identify the proper model. After this, the
to the remaining polymer appears. The boundary conditions for remainder of the section is devoted to comparing the results of
I and I thus become this model to the experimental MWD.
@I2 @I2
dsðtÞ ¼D þ dsðtÞkd I2 at x ¼ 0 ð10Þ
@t @x 4.1. Quantifying a match between MWD’s

@I 
@I n o
dsðtÞ ¼D þ dsðtÞ 2kd I2 kht I m1 kc ðI Þ2 at x ¼ 0 ð11Þ The quantitative evaluation of the match between MWD’s is
@t @x
based on the representation of the semi-logarithmic MWD usual in
The left-hand side of these equations represents the concen- Gel Permeation Chromatography, where a normalized concentra-
tration changes over time in the peroxide-rich layer. The first tion PMM2 is plotted versus log10 M, if M is the weight of a polymer
term right-hand side is the flux of the peroxide (or peroxide molecule with n monomer units of molecular weight Mmonomer.
radical species) into the polymer at x¼0 and the second term This MWD has exactly the same shape as the distribution of
refers to the production and consumption of peroxide and per- monomer units per chain, Pnn2, versus log10 n. The normalization
oxide radicals in that layer. The term ds(t) refers to the absolute enforces the surface under the MWD curve to equal 1, so that the
thickness of the peroxide-rich layer given by the product of its normalization factor follows as the summation:
volume fraction d and the striation thickness s, decreasing over nX n o
¼1
10 10 1 nX ¼1
time. Obviously, at the starting time, t¼ 0, the absolute thickness pn n2 log ðn þ1Þlog ðnÞ  pn n ð12Þ
equals ds0. Note again that variations in the amount of peroxide n¼1
lnð10Þ n ¼ 1
P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486 5479

The error e between a calculated distribution pcn and an experi- deviation from the experimental MWD: the latter ones are
mentally measured distribution pen consequently is expressed as broader than those calculated assuming ideal mixing. In terms
  of the error e defined above in Eq. (13) the deviation amounts to
 pcn pen 
e ¼ n2 lnð10ÞPn ¼ 1 c
 Pn ¼ 1 e

 ð13Þ around 17%. Hence, we conclude that our data – in contrast to e.g.
n¼1 p n n n¼1 p n n
Berzin’s data – are not well represented by an ideal mixing model.
This error lies between 0 and 1.

4.4. Thermal inhomogeneities in lateral direction


4.2. Strategy for finding the relevant mixing regime
In the previous section the effect of the thermal inhomogene-
The task for the model is to properly describe the observed ities in axial direction has been discussed. In this paragraph we
change in MWD of degraded PP. It is not obvious on beforehand will pay attention to such inhomogeneities in lateral direction,
that mixing will play a role at all. For instance, the experiments which is perpendicular to the screw axes. The importance of this
performed by Berzin and Vergnes (2000), also in a twin-screw aspect has been put forward by Berzin and Vergnes (2006) and
extruder turned out to produce MWDs that could perfectly be Suresh et al. (2008). The latter authors have developed a detailed
described by relatively simple random scission kinetics and ideal flow model of a twin-screw extruder and computed lateral
mixing in a batch reactor. Note also, that the kinetic rate temperature profiles, indicating that maximum temperature dif-
coefficients we use, were found using a kinetic model developed ferences in lateral direction may amount up to 20 K. Similar
by ourselves (Iedema et al., 2001) and Berzin’s MWD’s; this model information is available in Berzin and Vergnes (2006).
and these coefficients formed the basis for our earlier work on PP We will address the effect of lateral temperature inhomogene-
degradation in an SMR (Fourcade, 2001). The first question now ities using a simple compartment approach. According to this
is: can the MWDs measured by us equally well be explained by concept the extruder is a system of ideally mixed reactors in
ideal mixing as Berzin’s data? parallel, where polymer is exposed to the same initial peroxide
concentration and subjected to the same shape of temperature
4.3. The ideal mixing reference in relation to thermal profile as depicted in Fig. 4, but at various absolute levels. No
inhomogeneities in axial direction interchange or interaction exists between the compartments.
Obviously, here our objective is not to give a realistic description
At this point it is necessary to specify the ‘‘ideal mixing’’ of transport phenomena in an extruder, such as developed by
reference. It is decided that this reference is to be the ideal plug Suresh et al. (2008). It does, however, allow us to check whether
flow reactor without any axial or transversal dispersion of the existence of any temperature difference in principle could
temperature or concentrations. In that case all polymer material lead to MWD broadening.
has the same residence time in the extruder and there is no One series of computations have been performed with the
residence time distribution. This implies that the batch reactor measured temperature profile (Fig. 4) as the lowest in one
model described above applies to this situation, with the scission compartment and with DT higher levels in 10 further compart-
probability computed from Eq. (4a). As the temperature profile ments, such that the highest level is 30 K higher than the
the measured temperature in the 6 ZSK compartments has been measured temperature everywhere along the extruder. Thus, in
adopted, an example is shown in Fig. 4. Hence, the reference the coldest compartment the temperature increases from 200 1C
system is not an isothermal one. Now, one might wonder whether at time 15 s (Fig. 4) to 240 1C at the end and the hottest from
the shape of the axial temperature profile already would influence 230 1C to 270 1C. This must be considered as fairly extreme in
the shape of the MWD in the reference system. We have checked view of the data concerning lateral temperature differences in
this and found that the MWD resulting from any temperature Suresh et al. (2008) and Berzin and Vergnes (2006), around 20 K
profile could exactly be reproduced by an MWD at a certain at maximum, but here this approach serves well our goal to check
effective ‘‘average’’ processing temperature in an isothermal for MWD broadening. The computations were based on the
system. This implies that the heterogeneity in axial direction experimental shift of PP with an MW of 930 kg/mole to a product
does not alter the MWD shape and cannot explain any broadening with an MW of 125 kg/mole. The result is shown in Fig. 5 as the
of the MWD in PP degradation. MWD indicated with the average temperature T melt ¼ Tmeas þ15. It
The MWD has been calculated using the batch model to serve turns out to be close to the MWD for ideal mixing and to possess
as the reference for all the various experimental conditions
applied, as shown in Figs. 12–15. We observe a systematical Tmelt = Tmeas+15, per = 98%
1.2
Tmelt = Tmeas−5, per = 78%
250 1 Ideal mixing, per = 93%
Tmelt = Tmeas−15; ηper = 58%
Concentration

200 0.8
Measured
Temperature °C

0.6 SEC-DV Virgin PP


150 Mw = 125 kg/mole Mw = 930 kg/mole
0.4
100
0.2

50 0
104 105 106
Molar Mass
0
0 5 10 15 20 25 30 35 40 45 Fig. 5. Effect of thermal heterogeneities in lateral direction. No MWD broadening
Residence time in extruder and stronger MW shift due to higher conversion (98%) for average 15 K higher
melt temperature compared to ideal mixing case (93%). Broadening occurs at
Fig. 4. Temperature profile in extruder. lower temperatures and conversions and weaker shift.
5480 P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486

the same shape, so no broadening is observed in this case. The


1.2
small shift to lower MW is caused by the higher peroxide Batch reactor
conversion, Zper, being 98% instead of 93% in the ideal mixing
1 SEC-DV
case. More important, however, is to notice that it is the Mw = 125,000
Virgin PP

Concentration
difference in conversion due to the temperature differences 0.8
between the compartments that leads to different scission prob-
abilities. In this case the conversion level in all compartments is 0.6
high and consequently the differences in scission intensity small,
CSTR
which does not lead to significant broadening. The measured 0.4
MWD of the PP product after the considerable strong peroxide
shift in this experiment suggests a high peroxide conversion 0.2
indeed. Our conclusion here is that eventual higher temperature
0
regions can have only a minor effect on the conversion that is 103 104 105 106 107
already at a high level in view of the high overall temperature Molar Mass
level as measured in the extruder. Hence, such thermal hetero-
geneities may be very important for the product quality in itself, Fig. 6. Effect of residence time distribution on MWD; comparison of batch reactor
and CSTR.
but they will not lead to significant heterogeneities in shift
intensity and consequently neither lead to MWD broadening.
From the foregoing one may expect that stronger differences in
conversion in principle still may lead to significant broadening.
This is only possible, if the overall temperature and conversion
are lower, which in the experimental example taken would imply
that parts of the polymer melt in the extruder on average are at
lower than the measured temperature. Since this is not very
probable the issue of conversion differences under these condi-
tions is of mainly academic interest. However, our simple com-
partment model does allow us to check the effect of such lower
conversion. Fig. 5 shows the results of these exercises: one for, on
average, 5 K and one for 15 K lower than the measured tempera-
ture. Now, for these conditions we do observe considerable
broadening, but indeed, as expected, at the expense of much
lower conversions: 78% and 53%, respectively. It is also clear that
the lower peroxide conversion leads to considerably smaller shift
than the experimentally observed one. Our conclusion from this
exercise is that indeed thermal heterogeneities may lead to MWD Fig. 7. Residence time distribution in a twin-screw extruder yielding distribution
broadening, but only if the overall conversion is low. Hence, we of scission probabilities.
observe that the temperature differences cannot explain the
observed MWD broadening in our experiments.

1.2
4.5. The effect of residence time distribution on MWD 44 s
38.5 s
SEC-DV

1
At this point we will investigate the effect of residence time
Concentration

0.8
distribution. From an academic point of view one might be 58 s
interested in the results for the case of most extreme RTD: the
0.6
CSTR. The model equations applying to the CSTR have been
derived in the previous section, yielding the exponential distribu- 0.4
tion of scission probabilities as expressed by Eq. (4d). The MWD
for PP scission in a CSTR is shown in Fig. 6. One observes the 0.2
resulting MWD to be much broader than the one from the batch
reactor, but also much broader than all the experimental MWD 0
104 105 106
(the sample with average molecular weight Mw ¼125,000). The Molar Mass
error is even larger than that in the batch case: around 40%.
Hence, we conclude that RTD does have a strong impact on the Fig. 8. MWD computed for fractions of low, average and high residence time in a
twin-screw extruder. Overall MWD coincides with that for average t without
shape of the MWD indeed. However, the RTD of a CSTR is
visible broadening. Comparison with measured MWD.
evidently far from being representative for a twin-screw extruder.
The next issue is: what is the influence of a more realistic RTD
on the MWD of PP degradation in a twin-screw extruder. From
in which this range of t results, according to Eq. (4a). The MWD is
Poulesquen and Vergnes (2003) we adopt a general expression for
calculated for fractions of low (38.5 s) and high (58 s) t within
such an RTD
this range, the result is shown in Fig. 8. The overall MWD was
 
ðttd Þ2 ðttd Þ computed from all residence time fractions. This MWD turned out
EðtÞ ¼  3 exp ð14Þ to practically coincide with the one for the average t of 44 s,
2 ðttd Þ=3 ðttd Þ=3
without any visible broadening. Thus, we observe that where the
with t the average residence time and td the minimum value, in RTD of a CSTR leads to important broadening, the RTD of a twin-
the laboratory ZSK18 being 44 and 38.5 s, respectively. Fig. 7 screw extruder turns out to be too narrow for significant
shows this RTD together with the variation in scission probability, broadening.
P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486 5481

4.6. Effect of micromixing on scission probability kmole/m3 Concentration peroxide radicals


10-8
The micromixing model as described above has been applied
to compute the effect of micromixing on the MWD of PP 10-10
degradation in a twin-screw extruder, using a proper parameter

Concentration
set. The experimental conditions in the ZSK18 fix the initial
10-12 Stri
striation thickness as half the average distance between perox- atio
n no
ide-rich PP particles in the entrance section, s0, which depends on des
1-3
0
the amount of peroxide added and the volume fraction peroxide- 10-14
rich PP through Eq. (8). The striation thinning parameter, a, is not
known on beforehand, but will act as the fitting parameter in 10-16
order to match the calculated MWD to the measured ones. The
number of nodes in the striation is taken such that an increase of
10-18
this number does no longer change the overall MWD resulting. In 0 5 10 15 20 25 30 35 40 45
our previous publication this was seen to be at 11 nodes, in the Residence time (s)
present case 30 nodes were required. Note that as a consequence
kmole/m3 Concentration macroradicals
of using a Monte Carlo scission procedure the number of stria- 10-5
tions does not influence the required CPU time, as long as the
total size of the population of chains sampled – in the order of
millions – remains unchanged. Sampling using a larger number of Stri
atio
different scission probabilities simply implies fewer chains n no
10-10

Concentration
des
1-3
sampled per striation node, still yielding sufficiently smooth 0
distributions.
Before showing the resulting MWD under micromixing condi-
tions, first the consequences of micromixing for the key variable 10-15
scission probability will be presented. In Figs. 9 and 10 the
calculated concentration profiles over the extruder in all the
striations as defined above are shown. Discontinuities are due
to the discontinuous temperature profile imposed, see Fig. 4. 10-20
0 5 10 15 20 25 30 35 40 45
Employing these concentration profiles in the integration of
Residence time (s)
Eq. (4c) the variation of scission probability across the striations
Fig. 10. Concentrations versus residence in twin-screw extruder in all striations of
peroxide radicals (upper plot) and macroradicals (lower plot). Discontinuity caused
kmole/m3 Peroxide concentration by discontinuous temperature profile applied (see Fig. 4). Striation thinning rate:
100 a ¼0.06 s  1.

2500
Distribution of ρs
Concentration

10-2 Str from RTD in twin-


iati
on
nod 2000 screw extruder
Normalized frequency

es 1
-30

1500
Distribution of ρs due
10-4
to micromixing
1000

Exponential
10-6 -2 500 distribution of ρs
10 10-1 100 101 from RTD in CSTR
Residence time (s)
0
kmole/m3 Peroxide concentration 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
100 Scission probability ρs x 10-3

Fig. 11. Frequency distribution of scission probability for RTD of ideally mixed (no
micromixing) in twin-screw extruder and CSTR as compared to scission frequency
distribution due to micromixing (striation thinning rate: a ¼ 0.06 s  1).
Concentration

10-2
Stri
atio
n no
des
1-3 is determined. It is now interesting to compare the spread in
0
scission intensity as a result from RTD in a CSTR (exponential) and
10-4
in a twin-screw extruder (Fig. 7) under ideal mixing conditions to
the scission probability variation due to micromixing. We observe
that the scission intensity is extremely narrowly distributed if the
10-6 RTD in the twin-screw extruder is accounted for, as compared to
0 5 10 15 20 25 30 35 40 the exponential RTD that leads to a much broader (equally
Residence time (s) exponential) distribution in scission in the CSTR. Quite remark-
Fig. 9. Peroxide concentration versus residence in twin-screw extruder in all
able is the spread in scission density due to micromixing, which
striations, double-logarithmic (upper plot) and semi-logarithmic (lower plot). appears to be in between the two ideally mixed cases. It is still
Striation thinning rate: a ¼ 0.06 s  1. much wider than the spread caused by RTD in the extruder.
5482 P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486

Ideal mixing
SEC-DV α > 0.2 s-1, ideal mixing 1.2 SEC-DV
1.2 Mw = 125 kg/mole Micromixing model Mw = 225 kg/mole
0.08
ε = 0.056
1 0.06 1
εideal = 0.16
Concentration

Concentration
0.05 Virgin PP
0.8 0.8
Mw = 125 kg/mole
0.04
ε = 0.025 Virgin PP
0.6 0.03 0.6 εideal = 0.17 Mw = 930 kg/mole

0.4 -1 0.4
.01s
α=0
0.2 0.2
0
103 104 105 106 107 0
103 104 105 106 107
Molar Mass
Molar Mass
Fig. 12. Influence of striation parameter a on shape of MWD. For a 4 0.2 s  1 no
further changes are observed, MWD shape is identical to that of ideal mixing in a
batch reactor (Fig. 6). 1.2 SEC-DV
Ideal mixing
Micromixing model
Mw = 510 kg/mole
1 ε = 0.056
Although this RTD in the extruder is present and adds to the εideal = 0.15

Concentration
Mw = 160 kg/mole
widening due to micromixing, we decided to carry out all 0.8 ε = 0.064
micromixing calculations at the average value of the residence εideal = 0.15
time, since the additional broadening is expected to be practically 0.6 Virgin PP
invisible, as was already noticed from Fig. 7. Mw = 930 kg/mole
0.4
4.7. Effect of micromixing on MWD
0.2
In the previous sections it was shown that the kinetic model in
0
combination with the assumption of ideal mixing did not lead to a 103 104 105 106 107
good match between the calculated and measured MWD. Now, Molar Mass
the variability in scission probability due to micromixing lying in
Fig. 13. MWD of degraded PP with initial Mw ¼ 930 kg/mole (MFI 0.15) with
between that resulting from the RTD in a twin-screw extruder
(a) 0.486 and 0.193 wt% peroxide, yielding Mw ¼ 125 and 225 kg/mole (MFI-values
and in a CSTR, indeed turns out to produce good fits, errors e in 134 resp. 15.7), and (b) 0.314 and 0.007 wt% peroxide, yielding Mw ¼ 160 and
the order of 5%. Therefore, the remainder of this paper is devoted 510 kg/mole (MFI-values 44 resp. 1.4). MWD deviation between SEC-DV and
to a discussion on the key micromixing parameter in the actual micromixing model (a ¼0.06 s  1): e ¼ 2.5–6.4% (ideal mixing e ¼ 16%).
context of PP degradation. The experimental parameters varied in
this study were the Melt Flow Index (MFI) of the starting PP was allowed to vary, but only minor variations in a were observed
material and the amount of peroxide added; in one experiment (o0.01) rather than a clear trend in this parameter. Hence, it was
the number of rounds per minute of the extruder was increased. decided to employ the model with one a-value in combination
A higher starting MFI by definition implies lower viscosity and with the initial striation thickness, s0, depending the peroxide
different flow conditions at which the phenomena take place. amount fed through Eq. (8). We noted, besides, that the absolute
Also, a larger amount of peroxide leads to a higher extent of level of a-value producing the best fit was strongly correlated to
degradation and a stronger variation of rheological conditions. the value of s0 applied.
One might expect both parameters to affect the micromixing In order to still get an idea of the quantitative sensitivity of the
phenomena as well. In the micromixing model the only freely MWD-shape to the value of a, we varied it around 0.06 down-
adjustable parameter with respect to flow and mixing to our wards to 0.01, yielding broader distributions, and upward leading
disposal is the striation thinning parameter, a. Clearly, the initial to narrower MWD, ultimately to the ones obtained for ideal
MWD of an experimental sample is used as the initial MWD in the mixing, see Fig. 12. One observes that above values of 0.2 the
model calculations, so this variation is accounted for, but still an MWD coincides with that of ideal mixing. Evidently, a has to vary
additional effect on the striation thinning rate a might be by at least one order of magnitude to arrive at visible differences
possible. Furthermore, as explained in the previous paragraph, in MWD. This is in line with our finding that changing flow
the variation in peroxide amount is not expressed in terms of conditions by modifying the extent of degradation did not
initial feed concentrations, but rather as the average distance noticeably change a. In other words, this parameter might change
between peroxide-rich particles in the feed, appearing as the a little, but not strong enough to change MWD. This turned out
initial striation thickness s0 in the micromixing model. In this also to be the case for a changed rotational speed of the extruder,
manner the micromixing model already accounts for changes in which was in a few experiments increased from the standard
peroxide feed amounts. Nevertheless, as the a-value might be a 200 rpm to 600 rpm. Again, no noticeable change in the best fit
function of the rheological conditions referred to above, in yielding a-value was obtained.
addition to the effect accounted for through s0 the a-value In Table 3 the whole range of experimental conditions in terms
possibly varies as well with peroxide amount. What we found, of initial weight average molecular weight Mw and MFI as well as
however, was that the a-values giving the best fits, that is the the amounts of peroxide added in weight percentage of the
lowest error e according to Eq. (13), did not considerably change polymer throughput in the extruder is listed. In Figs. 13–16 the
neither with initial MWD nor with peroxide amount and the MWD as computed with both the ideal mixing and the micro-
resulting extent of degradation. Using a constant a-value of 0.06 mixing model with a constant a-value of 0.06 s  1 are compared
produced errors between 2.5% to 8.5%, as goes from Figs. 13–16. to the ones measured with SEC-DV. The deviation between model
For some samples slightly better (1%) fits could be obtained, if a and experiment as calculated by Eq. (13) is shown in the figures
P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486 5483

1.2 SEC-DV
1.2 SEC-DV
Micromixing model
Micromixing model Ideal mixing Ideal mixing
1 Mw = 350 kg/mole 1 Mw = 280 kg/mole
ε = 0.084 ε = 0.025

Concentration
εideal = 0.17 εideal = 0.045
Concentration

0.8 Mw = 120 kg/mole 0.8


ε = 0.074 Mw = 140 kg/mole
εideal = 0.19 0.6 ε = 0.040
0.6 εideal = 0.095

0.4 Virgin PP 0.4 Virgin PP


Mw = 590 kg/mole Mw = 380 kg/mole

0.2 0.2

0 0
103 104 105 106 107 103 104 105 106 107
Molar Mass Molar Mass

Fig. 16. MWD of degraded PP with initial Mw ¼ 380 kg/mole (MFI 3.1) with 0.23
SEC-DV and 0.03 wt% peroxide, yielding Mw ¼ 140 and 280 kg/mole (MFI-values 119 resp.
1.2
Micromixing model 14.7). MWD deviation between SEC-DV and micromixing model (a ¼ 0.06 s  1):
Ideal mixing
e ¼ 2.5–4% (ideal mixing e ¼ 4.5–9.5%).
1 Mw = 500 kg/mole
ε = 0.049
Concentration

0.8 εideal = 0.090 Table 3


Mw = 180 kg/mole
ε = 0.049 MWD fits of ideal mixing and micromixing model (with a ¼ 0.06 s  1) to SEC-DV
0.6 εideal = 0.16 measurements of PP samples degraded in ZSK 18.

Virgin PP Initial Mw MFIa Final Mw MFIa wt% peroxide e eideal


0.4
Mw = 590 kg/mole
380,000 3.06 280,000 14.65 0.03 0.025 0.045
0.2 140,000 119.1 0.23 0.04 0.095
480,000 1.23 200,000 20.98 0.13 0.065 0.10
0 145,000 66.23 0.26 0.039 0.11
103 104 105 106 107 590,000 0.49 500,000 0.95 0.012 0.049 0.09
Molar Mass 350,000 5.63 0.07 0.084 0.17
180,000 39.6 0.23 0.049 0.16
Fig. 14. MWD of degraded PP with initial Mw ¼590 kg/mole (MFI 0.49) with 120,000 135.11 0.45 0.074 0.19
(a) 0.45 and 0.07 wt% peroxide, yielding Mw ¼120 and 350 kg/mole (MFI-values 930,000 0.15 510,000 1.4 0.045 0.056 0.15
135 resp. 5.6), and (b) 0.23 and 0.0012 wt% peroxide, yielding Mw ¼ 180 and 225,000 15.71 0.193 0.056 0.16
500 kg/mole (MFI-values 40 resp. 0.95). MWD deviation between SEC-DV and 160,000 44.07 0.314 0.064 0.15
micromixing model (a ¼0.06 s  1): e ¼ 4.9–8.5% (ideal mixing e ¼9–19%). 125,000 134.07 0.486 0.025 0.16

a
MFI: Melt Flow Index.
1.2 SEC-DV
Micromixing model Ideal mixing
Mw = 200 kg/mole
1
ε = 0.065
(around 10%), while the fits of the micromixing model are equally
εideal = 0.10 around 5%. Note again that the best fit between modeled and
Concentration

0.8 experimental MWD obtained by allowing freedom in all para-


Mw = 145 kg/mole
ε = 0.039 meters employed never was lower than 2%. We suspect that this
0.6 εideal = 0.11 deviation is close to the total experimental error in measuring
Virgin PP the MWD.
0.4
Mw = 480 kg/mole

0.2
5. Conclusion
0
103 104 105 106 107 We have performed a series of controlled degradation experi-
Molar Mass ments on a ZSK 18 twin-screw extruder with Polypropylene of
Fig. 15. MWD of degraded PP with initial Mw ¼480 kg/mole (MFI 1.2) with 0.26 varying Mw under addition of various amounts of peroxides and
and 0.13 wt% peroxide, yielding Mw ¼ 145 and 200 kg/mole (MFI-values 66 resp. have measured the initial and final MWD with SEC-DV. Models
21). MWD deviation between SEC-DV and micromixing model (a ¼0.06 s  1): assuming ideal mixing while introducing different degrees of
e ¼ 3.9–6.5% (ideal mixing e ¼10–11%). heterogeneity by accounting for different distribution function
of residence time did reveal a strong impact on MWD. In absence
and listed in Table 3. We clearly observe that assuming a of RTD the model predicted MWD turned out to be too narrow,
micromixing effect strongly improves the MWD fit as a conse- even accounting for the typical steep temperature profile existing
quence of the MWD broadening involved with the extra hetero- in the extruder. Assuming a realistic RTD for a twin-screw
geneity due to this effect. Quantitatively speaking, the extruder (Fig. 7) did not lead to noticeable broadening, but the
improvement consists of a reduction of the modeling deviation exponential RTD of a CSTR did indeed strongly widen the MWD;
from 16% for the ideal mixing to only 2.5–7.5% for the micromix- the latter produced an MWD that was too broad. However,
ing model. In addition, one may observe that the impact of neither of the several RTD options produced a good fit between
micromixing is strongest for the samples with highest initial Mw model and experiments. Clear deviations between the MWD’s
that obviously underwent the strongest degradation: those with were visible, while the calculated (Eq. (13)) deviation in surface
starting Mw of 930 and 590 kg/mole. For the samples with lower overlap amounted to 15% for the batch reactor and to 40% for the
initial Mw the ideal mixing assumption leads to smaller errors CSTR. The effect of thermal heterogeneities in lateral direction
5484 P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486

(perpendicular to screw axis) has also been investigated using a Ds diffusion coefficient of species s Eq. (5)
simple compartment model. This turns out to lead to MWD I2 initiator concentration
broadening only at lower overall conversion levels, which then IK initiator radical concentration
lead to weaker MW shift than experimentally observed. Hence, in kc initiator radical combination coefficient
the present case thermal inhomogeneities could not explain the kd initiator dissociation coefficient
observed broadening. kht hydrogen abstraction coefficient
Micromixing represented a further possible phenomenon ktd termination by disproportionation coefficient
introducing heterogeneity to be explored. Based on the principles ktr transfer to polymer coefficient
of an earlier developed, fully deterministic micromixing model M, Mmonomer molecular weight of polymer, resp, monomer
(Fourcade et al., 2001a,b) we constructed a partly Monte Carlo n number of monomer units in polymer chain
sampling based model. Ottino’s striation thinning approach is pn concentration of polymer with n monomer units
applied to the geometrical conditions of a twin-screw extruder, PM concentration of polymer with weight M
where peroxide is fed by a few (one in a few thousand) PP Rs reaction rate of species s Eq. (5)
particles with 40 weight % peroxide. This yields the initial s0 initial striation thickness Eq. (7)
striation thickness as half the average distance between such wp peroxide weight fraction Eq. (6)
particles in the extruder entrance section. The variation of the
scission probability over the total striation thickness (calculated
at 30 striation nodes) turned out to be considerable (see Fig. 10). Greek letters
Thus, as expected, the effect of this variation on the MWD
calculated was also significant, allowing matches between mea- a striation thinning rate Eq. (5)
sured and modeled MWD of very good optical agreement, while d volume fraction peroxide-rich part Eq. (7)
surface deviations reduced to around 5%. It turned out that all the l0 total concentration of macroradicals
measured MWD could be reproduced with the micromixing m1 total concentration of monomer units in polymer
model within this error using a constant value for the key rs scission probability
parameter, the striation thinning rate, a ¼0.06 s  1. This was so t, t, td residence time (RT), average RT, minimum RT in
despite the different conditions under which the experiments extruder
were carried out: several qualities of the starting material (Mw
varying from 380 to 930 kg/mole) and varying peroxide amounts
(between 0.012 and 0.486 wt%). Neither did one experiment at Appendix. Variable transformations
higher rotational speed (600 rpm instead of 200 rpm) alter the
a-value yielding the best fit. The transformations discussed in this Appendix are meant to
The broader experimental MWD’s as compared to those reduce the computational effort of solving the differential equa-
calculated assuming ideal mixing can only lead to the conclusion tion describing micromixing, Eq. (5) of the main text. Thus,
that more heterogeneous conditions in the extruder must exist solutions are found for the profiles of I and l0 that finally yield
than those resulting from the temperature profile and the narrow the scission probability rs under micromixing conditions.
RTD. We consider the good fit between the predicted MWD from
the micromixing model to the SEC-DV data as an (indirect) proof A. Co-ordinate transformation according to Ottino
that micromixing phenomena are the cause of the extra hetero-
geneity. If micromixing thus is observed to affect MWD and A1. Formulation of the problem
henceforth product properties, the question arises, whether this A co-ordinate transformation for the space and time coordi-
phenomenon might be utilized as an extra handle to control nates, x and t, was applied as described by Ottino (1980). New
properties. For instance, if one would desire a broad or even variables x and t were defined as follows:
bimodal product (see Fig. 11), it might be realized by even x x
x¼ ¼ eat ðA1Þ
‘‘worse’’ micromixing (lower a). Now, the present study did not st s0
reveal how to manipulate micromixing, for instance, by creating Z 
t
stronger or weaker striation thinning (a-value), so more research D 0 D e2at 1
t¼ dt ¼ ðA2Þ
in this direction is required to investigate this issue. Further 0 st0 2 s0 2 2a
research should also make clear whether micromixing plays a role where the striation thickness at time t, st, is
in other polymer degradation systems. In the present study we Z t
saw the impact of micromixing to become stronger with a greater st ¼ s0 exp aðt 0 Þdt 0 ¼ s0 eat ðA3Þ
MWD shift. In reactive modification of Polyethylene smaller shifts 0

(even to higher MW, due to branching) are observed, so the In terms of the new variables x and t, the convection diffusion
impact of micromixing might be weaker. On the other hand, the equation Eq. (6) of the main text is transformed into the following
grafting on Polypropylene (or another polymer) of other polymer form:
chains, where grafting, scission and diffusion are in competition,
@Cs @ 2 Cs
might be even more sensitive to micromixing. In general, polymer ¼ þ KRs ðA4Þ
@t @x
2
systems where selectivity is an important issue might be very
interesting to investigate with respect to micromixing. with
s0 2
K¼ ðA5Þ
Nomenclature 2as0 2 t þ D
This form is evidently much simpler than the original one, no
Latin letters longer containing a product of a time/space dependent variable
and a derivative. It should be noted that x defines a dimensionless
Cs concentration of species s Eq. (5) space scale between 0 and 1 based on the actual value of the
dp particle diameter striation thickness t is called warped time. Eq. (A4) implies that in
P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486 5485

the warped time scale t and space scale x the apparent rate of resolution at the high peroxide concentration side, we apply a
reaction is changed by a factor K. The no flux boundary conditions mesh transformation, g, which transforms variable x into a new
Eq. (9) of the main text become variable xn, to solve Eq. (A4) on equidistant grid nodes in the
@I2 @I xn-space, which corresponds to non-equidistant grid nodes in the
¼ 0 and ¼0 at x ¼ 1 ðA6Þ x-space
@x @x
g : ½0,1-½0,1
Boundary condition Eqs. (9) and (10) of the main text,
respectively, become x/xn ¼ gðxÞ ðB1Þ

@I @I2 In terms of variables x and t, Eq. (A4) becomes


n

d ¼ þK dkd I2 at x ¼ 0 ðA7Þ
@t @x @Cs @Cs @2 Cs
¼ g 00 ðxÞ n ðxÞ þ ðg0ðxÞÞ2 n2 ðxÞ þ KRs ðB2Þ
n o @t @x @x
@I @I
d ¼ þ K d 2kd I2 kht I m1 kc ðI Þ2 at x ¼ 0 ðA8Þ
@t @x where g 0 ðxÞandg 00 ðxÞstand, respectively, for the first and the
second derivative of g. The following functions were considered
for the mesh transformation g:

A2. Solution x ¼ g 1 ðxn Þ ¼ axn2 þ ð1aÞxn ðB3Þ


The solution of Eqs. (A4)–(A8) is obtained using the method of
where a is a parameter between 0 and 1. The density of nodes at
lines. The striation is discretized in N nodes (Fig. A1) and the
the vicinity of the peroxide-rich layer is related to a in such a way
diffusion term is approximated using centered finite differences.
that higher values of a correspond to higher density. All the
We define I2i and Ii as the concentration of the two (diffusing)
simulations in this paper have been performed with a value of
peroxide species I2 and I at node i, while the (non-diffusing)
i 0.7 for a. The functions taking part in Eq. (B2) are thus
macroradical concentration, l0, at node i is l0 .
The coupled set of ordinary differential equations (ODE) ð1aÞ þðð1aÞ2 þ 4axÞ1=2
gðxÞ ¼
resulting from the discretization of Eq. (A4) is 2a
@li0 i
g 0 ðxÞ ¼ ðð1aÞ2 þ4axÞ1=2
¼ KRðl0 Þ
@t g 00 ðxÞ ¼ 2aðð1aÞ2 þ 4axÞ3=2 ðB4Þ
@I2i @2 I2i
@t ¼ 2 þKRðI2i Þ i ¼ 1,N ðA9Þ
@x The boundary conditions, Eqs. (A6)–(A8), respectively, become
@Ii @2 Ii
@t ¼ þ KRðIi Þ @I @I
@x2 n
n ¼ 0 and n ¼0 at x ¼ 1 ðB5Þ
i
@x @x
Here Rðl0 Þ,RðI2i Þ
and RðIi Þ
represent the chemical reaction terms for
these species as defined above. Obviously, no diffusion term is @I @I n
d ¼ g 0 ð0Þ n þK dkd I2 at x ¼ 0 ðB6Þ
appearing in the equations for the macroradical. The diffusion term @t @x
2
@2 I2i =@x is calculated with the finite difference approximation
@I @I n o
¼ g 0 ð0Þ n þ K d 2kd I2 kht I m1 kc ðI Þ2
n
@2 I2i1 d at x ¼ 0 ðB7Þ
¼ 2 ðI2i1 2I2i þI2i þ 1 Þ ðA10Þ @t @x
2 h
@x
The resulting set of ODE’s to be solved in the (xn,t) space is
and similarly that for the initiator radicals. The derivative now
w.r.t. x terms in Eq. (A7) is approximated by the following finite
@li0 i
difference: @t ¼ KRðl0 Þ
 @I2i i @Ii i @2 I2i
@I2  3I0 4I21 þ I22 @t ¼ g 00 ðx Þ @x2n þ g 0 ðx Þ2 þ KRðI2i Þ i ¼ 1,N ðB8Þ
¼ 2 ðA11Þ @xn2
@x x ¼ 0 2h @Ii i @Ii i 2 @2 Ii
@t
00
¼ g ðx Þ @xn þ g ðx Þ 0
n2 þ KRðIi Þ
@x
where I20 is the peroxide concentration at x ¼0; a similar expression
is valid for I0 . with the finite difference approximations for the peroxide con-
centration derivatives (and similarly for initiator radicals)
B. Transformation to non-equidistant space grid @I2i
@xn
¼ 1
ðIi þ 1 I2i1 Þ
2h 2
@2 I2i
ðB9Þ
In the derivation above diffusion is dealt with using a centered n2 ¼ 1
h2
ðI2i1 2I2i þ I2i þ 1 Þ
@x
finite difference approximation in combination with an equidi-
stant mesh of the space co-ordinate x. In order to increase the The solution of Eqs. (B8) yields profiles over warped time, t, of
the initiator radical and macroradical concentrations I and l0 ,
respectively, at all striation nodes i. Since it is the scission
1 k-1 k k+1 7
probability rs we are interested in, we infer the proper expression
to compute it for all nodes from Eq. (4a) of the main text as
0 h 1
Z tb Z 1
i i
ris ¼ ðkht Ii þktr l0 Þdt ¼ KðtÞðkht Ii þ ktr l0 Þdt ðB10Þ
0 0
1 k-1 k k+1 7 Here tb is the residence time in the extruder (in real time, t) and
K(t) is the transformation function as defined by Eq. (A4). Note
1
0 hk that obviously the integrals over real time, t, and warped time, t,
should lead to identical scission probabilities.
Peroxide-rich polymer Virgin polymer Now, the scission probabilities for each striation node i by
Fig. A1. Meshes of the striation with 7 nodes. Upper part: equidistant mesh, lower applying the Monte Carlo scission procedure lead to a chain
part: mesh from 2nd transformation with higher resolution near peroxide-rich layer. length distribution after scission for each striation node, Pni . From
5486 P.D. Iedema et al. / Chemical Engineering Science 66 (2011) 5474–5486

this, finally, the overall chain length distribution is computed Kolhapure, N.H., Fox, R.O., Daiss, A., Mahling, F.O., 2005. PDF simulations of
using hi, the length of the ith element in the grid (see Fig. A1) ethylene decomposition in tubular LDPE reactors. AIChE J. 51 (2), 585–606.
X Marini, L., Georgakis, C., 1984. The effect of imperfect mixing on polymer
Pn ¼ Pni hi ðB11Þ quality in low density Polyethylene vessel reactors. Chem. Eng. Sci. 30 (6),
i ¼ 1,N 361–375.
McKenna, T.F., Cokljat, D., Wild, P., 1998. CFD modeling of heat transfer during
gas phase olefin polymerisation. Comput. Chem. Eng. Vol. 22 (Suppl),
285–292.
Nele, M., Soares, J.B.P., 2002. Long-chain branching with metallocene catalysts: is a
References
purely kinetic mechanism for terminal branching sufficient? Macromol.
Theory Simul. 11 (9), 939–943.
Baldyga, J., Rozen, A., Mostert, F., 1998. A model of laminar micromixing with Ottino, J.M., 1980. Lamellar mixing models for structured chemical reactions and
application to parallel chemical reactions. Chem. Eng. J. 69 (1998), 7–20. their relationship to statistical models; macro- and micromixing and the
Berzin, F., Vergnes, B., 2000. Modelling of peroxide initiated controlled degradation problem of averages. Chem. Eng. Sci. 35, 1377–1391.
of polypropylene in a twin screw extruder. Polym. Eng. Sci. 40, 344–356. Poulesquen, A., Vergnes, B., 2003. A study of residence time distribution in co-
Berzin, F., Vergnes, B., 2006. Modeling of reactive systems in twin-screw extru- rotating twin-screw extruders. part i: theoretical modeling. Polym. Eng. Sci. 43
sion:challenges and applications. C. R. Chim. 9, 1409–1418. (12), 1841–1848.
Camara, S., Gilbert, B.C., Meier, R.J., van Duin, M., Whitwood, A.C., 2006. EPR Schmidt, C.U., Busch, M., Lilge, D., Wulkow, M., 2005. Detailed molecular structure
studies of peroxide decomposition, radical formation and reactions relevant to modeling—a path forward to designing application properties of ldPE. Macro-
cross-linking and grafting in polyolefins. Polymer 47, 4683–4693. mol. Mater. Eng. 290 (4), 404–414.
Fields, S.D., Ottino, J.M., 1987. Effect of striation thickness distribution on the
Suresh, A., Chakraborty, S., Kargupta, K., Ganguly, S., 2008. Low-dimensional
course of an unpremixed polymerization. Chem. Eng. Sci. 42, 459–465.
models for describing mixing effects in reactive extrusion of polypropylene.
Fourcade, E., Wadley, R., Hoefsloot, H.J.C., Green, A., Iedema, P.D., 2001a. Laminar
Chem. Eng. Sci. 63 (14), 3788–3801.
striation thinning in static mixer reactors from computational fluid dynamics
Tobita, H., 1996. Random degradation of branched polymers. 1. Star polymers.
and laser induced fluorescence. Chem. Eng. Sci. 56, 6729–6741.
Fourcade, E., Hoefsloot, H.C.J., van Vliet, G., Bunge, W., Mutsers, S.M.P., Iedema, P.D., Macromolecules 29, 3000–3009.
2001b. The influence of micromixing on molecular weight distribution during Tsai, K., Fox, R.O., 1996. PDF modelling of turbulent-mixing effects on initiator
controlled Polypropylene degradation in a static mixer reactor. Chem. Eng. Sci. efficiency in a tubular LDPE reactor. AIChE J. 42 (10), 2926–2940.
56, 6589–6603. Wells, G.J., Ray, W.H., 2005. Mixing effects on performance and stability of low-
Iedema, P.D., Willems, C., Van Vliet, G., Bunge, W., Mutsers, S.M.P., Hoefsloot, H.C.J., density polyethylene reactors. AIChE J. 51 (12), 3205–3218.
2001. Using molecular weight distributions to determine the kinetics of Wulkow, M., 1996. The simulation of molecular weight distributions in polyreac-
peroxide-induced degradation of Polypropylene. Chem. Eng. Sci. 56, tion kinetics by discrete Galerkin methods. Macromol. Theory Simul. 5,
3659–3669. 393–416.

Você também pode gostar