Sie sind auf Seite 1von 11

Hydrogen Response in Liquid Propylene

Polymerization: Towards a Generalized Model


M. Al-haj Ali, B. Betlem, B. Roffel, and G. Weickert
University of Twente, Enschede, The Netherlands

DOI 10.1002/aic.10783
Published online February 6, 2006 in Wiley InterScience (www.interscience.wiley.com).

Liquid propylene batch experiments in the absence of a gas phase have been carried out
using a highly-active MgCl2/TiCl4/phthalate/silane/AlR3 catalyst at varying temperatures
(60-80C) and molar hydrogen-monomer ratios of 0-10 mmol/mol. With increasing
hydrogen concentration the
Y polymerization rate increases rapidly, reaching a constant value at concentrations
above 1.4 mmol/mol;
Y pseudo-rst-order catalyst deactivation constant increases;
Y molecular weight decreases;
Y polydispersity decreases slightly;
but average molecular weight and polydispersity increase with increasing temperature.
Polymerization rate, deactivation constant, and average molecular weight can be modeled
based on a consistent dormant site mechanism assuming an (averaged) quasi-single-site
model. 2006 American Institute of Chemical Engineers AIChE J, 52: 1866 1876, 2006
Keywords: liquid phase propylene polymerization, polymerization kinetics, dormant-site
theory, molecular weight distribution, catalyst decay, hydrogen response

Introduction partially-lled, and polymerizations were carried out in the


Only a very few experimental studies have been carried out presence of a gas phase, which can introduce mass transfer
so far to investigate the effect of hydrogen during liquid pro- problems, see ref. 19.
pylene polymerization; most studies were performed in either a
gaseous monomer or slurry,1-15 but gas phase and slurry phase Molecular weight
results cannot easily be compared and heat and sorption effects In 1959, Natta et al.20 studied the effect of hydrogen on the
and mass transfer limitations may play a role.16 molecular weight of the produced polymer for a TiCl3/
Al(C2H5)3 catalyst. Their results showed that the number av-
Polymerization rate erage molecular weight of both polyethylene and polypro-
Some researchers measured polymerization yields in liquid pylene could be described by the square root of the hydrogen
propylene,17,18 but much more kinetic information can be ex- partial pressure in the reactor, Eq. 1:
tracted from polymerization rate proles. The rst article re-
porting rate proles in catalytic liquid propylene polymeriza- 1
Mn (1)
tion was published by Samson et al.,4 and a number of other K 1 K 2 P H2
authors investigated the hydrogen effect on kinetics (see Pater
et al.2 and Shimizu et al.3). The reactor used was always The authors concluded that hydrogen is adsorbed on the
catalyst surface where it dissociates (Eq. 2) before it reacts
Correspondence concerning this article should be addressed to G. Weickert at with the growing polymer.
g.weickert@tnw.utwente.nl.

2006 American Institute of Chemical Engineers H2 ^ H H (2)

1866 May 2006 Vol. 52, No. 5 AIChE Journal


Table 1. Most Important Publications on the Inuence of Hydrogen in Catalytic Olen Polymerization
Author System Studied Findings
Keii et al.22 TiCl4/MgCl2/C6H5COOC2H5/Al(C2H5)3 Agreed with Eq. 1
Gas phase propylene
Soga & Siono23 TiCl4/MgCl2/Al(C2H5)3/ethyl benzoate 40C; slurry Atactic PP is formed by atomic hydrogen transfer;
propylene isotactic PP results from molecular H2 transfer
Guastalla & Gianinni14 TiCl4/MgCl2/Al(C2H5)3 slurry propylene and Mw of different polymer fractions is reduced
ethylene following Eq. 1
Ross21 TiCl3 AA/DEAC Eq. 1 ts best
Gas phase propylene
Hindryckx et al.49 MgCl2-based TiCl3 Eq. 1
Slurry, ethylene
Busico et al.26 Wide range of hydrogen partial pressure studied Eq. 1
Meier et al.42 Metallocene Inverse Mw depends linearly on hydrogen
Gas phase, propylene (linearization possible in a narrow H2 range)
0 CH2/Cpropylene 0.0016
Kissin et al.24 TiCl4/dibutylphthalate/MgCl2 Mw-H2: linear at low hydrogen, but constant at
0 CH2/Cpropylene 0.08 high hydrogen; polymer H2 complex formed

Some of the most important ndings are summarized in Experimental


Table 1. All these studies were carried out either in a gas- Chemicals
phase21,22 or slurry.23-26 While most of them support the Natta
model (Eq. 1), some differ. Propylene used in the experiments was polymerization
grade and obtained from Indugas. The purity was more than
99.5%, with propane as a main impurity. Hydrogen used had a
Catalyst deactivation purity higher than 99.999%; it was further puried by passing
One of the most widely supported hypotheses for catalyst it over a reduced BTS copper catalyst, obtained from BASF,
activity decay is the over-reduction of the active centers, Ti3, and subsequently passing it through three different beds of
into inactive form, Ti2, by the cocatalyst.12,13,15,27-29 Keii et molecular sieves, with pore sizes of 13, 4, and 3 , respec-
al.15 examined the effect of cocatalyst addition, Triethylalumi- tively. Propylene was puried in the same way; additionally, it
num (TEA), whereas Busico et al.28 showed a rapid second- was passed over a bed of oxidized BTS catalyst to remove CO.
order decay of the polymerization rate explained by clustered TiCl4 supported on MgCl2 with phthalate as internal donor and
low-active Ti2 species. However, Albizzati et al.30 studied an external silane donor was used as a catalyst, with TEA as a
[6-areneTiAl2Cl8] and found that catalysts containing Ti2 cocatalyst and scavenger.
are active in propylene polymerization and the catalyst decay
cannot be explained by the reduction of active trivalent Ti. Reactor system
Shimizu et al.3 showed that higher TEA concentrations can A 5-liter stainless steel jacketed batch reactor (Buchi BEP
lead to lower decay rates. 280) with a separately heated cover plate was used; it is
Many authors proposed different chemically-based mecha- described elsewhere.2-4 For intensive mixing, the reactor was
nisms. Chien and Nozaki31 explained the polymerization rate equipped with a turbine stirrer operated at 2000 rpm. The
decay by the reaction of adjacent pairs of Ti3 with monomer. pneumatic injection system allows the introduction of liquids
Based on their observations, Guyot et al.32 suggested that and slurries into the reactor, even at high reactor pressures. The
catalyst deactivation results from the formation of inactive cooling medium temperature is kept constant within 0.01 K
-allyl species with saturation of the polymer chain end. during isoperibolic experiments.
Different authors have already claimed that a monomer mass
transfer limitation is not the cause of the rate decay12,15,28,29;
however, others still support this hypothesis.5 Also, Han-Ad- Experimental procedure
ebekun et al.6 showed that the effectiveness factor for highly The reactor was ushed with nitrogen gas ve times at 90C
crystalline propylene is smaller than 1.0. and was purged with propylene gas at the beginning of the
In this article: experiment, then lled with liquid propylene and heated up to
(1) Liquid propylene polymerization with a highly active the reaction temperature. When the temperature reached the
catalyst in an over-pressured fully lled batch reactor is inves- setpoint, hydrogen was injected. The reactor temperature and
tigated. At these conditions, all the hydrogen is dissolved in pressure were monitored as a function of time. As soon as both
liquid propylene and gas-liquid mass transfer is excluded. became stable for an interval of 3 minutes, the reaction was
(2) The inuence of polymerization temperature and hy- started by injecting the prepared catalyst into the reactor.
drogen concentration on polymerization kinetics and molecular The experiments were executed under isoperibolic condi-
weight is studied in a wide range of hydrogen concentrations. tions. Thus, just after the catalyst injection, the temperature
(3) Models are presented and compared to quantify the control system becomes active in trying to keep the jacket
effect of temperature and hydrogen concentration on the initial temperature constant; the reaction temperature increases
polymerization rate, the catalyst deactivation, and the polymer slightly, reaching a quasi-steady-state after about 1.5 minutes
molecular weight. in the case of using a fully pre-activated catalyst. The heat of

AIChE Journal May 2006 Vol. 52, No. 5 Published on behalf of the AIChE DOI 10.1002/aic 1867
mate the true initial polymerization rate. We estimated two
different initial polymerization rates: Rpo, by extrapolating the
reaction rate curve to time zero; and Rpo,ATR, which can be
calculated from the initial adiabatic temperature rise assuming
that the heat produced mainly heats up propylene during the
rst seconds, see Table 2. For almost all runs, Rpo is higher
than Rpo,ATR, but because of the uncertainties of all differential
(slope) methods combined with the assumption of adiabatic
conditions and taking into account the data acquisition fre-
quency of only 0.3 s1, we interpret these ndings in terms
of a fully pre-activated catalyst, i.e., the extrapolated value Rpo
can be used as the true initial polymerization rate.
Figure 1. Polymerization rate-time proles of repeated
It is relatively easy to check this extrapolation hypothesis
standard experiments.
by varying the reaction time. If the yield, Y, of such experi-
ments can be estimated from the area underneath the extrapo-
polymerization is measured under quasi-steady state condi- lated (isoperibolic) temperature difference, T (time curve),
tions. Data were collected every three seconds. The polymer- but for different reaction times, then this hypothesis is true.
ization reaction was nally terminated by rapidly ushing the First, this area is estimated for two T proles:
unreacted propylene. After each experiment, the resulting poly- (1) A1 from T recorded directly starting at zero rate,
mer was dried under vacuum at 50C for 4 hours. The molec- Figure 3a; and
ular weight distribution was measured by Gel Permeation (2) A2 from T by extrapolating the curve to time zero,
Chromatography (GPC) employing a Waters Alliance GPCV Figure 3b.
2000 apparatus with TSK columns at 155C using 1,2,4-tri- Clearly, it holds always that A2 A1. The results are shown
chlorobenzene as a solvent. The GPC is calibrated with the in Table 2. For example, for experiment 7 we estimate Y/A1
narrow molecular weight distribution polystyrene standard as a 2.2 and Y/A2 2.06, respectively. These results are not
reference. Average molecular weights above 1 million grams signicantly different, because of the relatively small contri-
per mol have been measured using calibrated intrinsic viscos- bution of the initial period. However, the situation changesas
ities. expectedfor a short-time experiment, carried out under sim-
ilar conditions, but stopped after 7.5 minutes, experiment 14;
Results and Discussion see Table 2 and Figure 4, with Y/A1 2.95 and Y/A2 1.95.
Reproducibility The latter value comes close to the value of the long-lasting
The reproducibility of the experiments has been tested by experiment 7, whereas the model based on A1 fails, with 30%
repeating a standard experiment at 70C and 43 bar with 150 deviation. This has two major consequences:
mg hydrogen at xed concentrations of the catalyst, cocatalyst, (1) The polymerization rate prole should be corrected by
and external donor. Figure 1 shows the reaction rate proles for extrapolating the measured prole to time zero. The yield
three standard experiments. The increase in polymerization rate measured is represented by the area underneath this corrected
during the rst 1.5 minutes is not real, but caused by the rate-time curve.
(incorrect) application of the quasi-steady-state assumption for (2) Y/A2 should be calculated for a given reactor system.
the heat balance during the initial phase. We will ignore this By means of this ratio, one can estimate yield and polymer-
effect for the moment, but will come back to this point later. ization rate on-line directly from the temperature difference
Obviously, the absence of a gas phase, i.e., absence of any during an isoperibolic experiment.
mass transfer between gas and liquid phase, increases the The kinetic behavior of different catalyst systems can be
accuracy of such rate prole measurements. very different, but the method used above can be applied to
clarify this point.
Isoperibolic initial polymerization rate
Reaction calorimeters can be classied according to their
way of controlling the reaction temperature into the following
categories: (i) isothermal, (ii) adiabatic, and (iii) isoperibolic.
Using an isoperibolic calorimeter, the jacket temperature is
kept constant, whereas the reaction temperature slightly in-
creases during the exothermic reaction. However, the polymer-
ization temperature can be kept within a small range producing
quasi-isothermal data. The isoperibolic calorimeter has many
advantages compared to the other calorimetric methods as
discussed in 33,34
The region where the rate of polymerization reaches the
quasi-steady-state near the maximum, II in Figure 2, is a
signicant kinetic ngerprint in calorimetric observations of Figure 2. Kinetic curve obtained during propylene poly-
catalytic polymerizations,35,36 but the question is how to esti- merization.

1868 DOI 10.1002/aic Published on behalf of the AIChE May 2006 Vol. 52, No. 5 AIChE Journal
Table 2. Conditions and Results of Liquid-Pool Propylene Polymerization

Run T H2 X 10 3 Catalyst t reaction Y Mw R po kg/ R po,ATR kd Y/ Y/
No. C mg [] mg min gram kg/mol PDI gcat hr kg/gcat hr 1/hr A1 A2
1 60 0 0 3.78 60 35.9 3284 8.5 11 0.288 1.91 1.93
2 60 150 1.33 3.78 60 132.9 266 6.5 54.7 62.7 0.900 2.21 2.12
3 60 1000 9.81 1.54 50 54 106 75.8 53.1 1.446 2.26 2.14
4 70 0 0 3.78 60 115 3878 16.1 9.6 0.342 1.92 1.86
5 70 25 0.24 3.78 60 163 1119 6.35 62.5 37.9 0.798 2.2 2.05
6 70 50 0.49 3.78 60 214 833 6.58 81.3 74.9 0.948 2.29 2.21
7 70 150 1.44 3.78 35 169.5 361 6.77 121.6 56.5 1.188 2.2 2.06
8 70 250 2.47 3.78 47 247.7 294 145.1 90.5 1.53 2.32 2.17
9 70 500 5.15 1.54 45 95.5 244 7.32 154.1 144.7 1.764 2.43 2.29
10 70 1000 9.94 1.54 45 90.3 132 139.6 89.9 1.932 2.05 1.92
11 75 150 1.56 3.78 35 232 208.5 120 2.26 2.01 1.95
12 80 0 0 3.78 60 81.5 2818 27.9 30.6 0.378 2.17 2.08
13 80 150 1.9 1.54 52 92 393 205.4 169.2 2.034 2.4 2.24
14 80 1000 12.4 1.54 32 92.3 143 5.7 252.4 130.4 3.21 2.48 2.3
15 70 150 1.50 3.78 7.5 47.8 NA NA NA NA 2.95 1.95

Other polymerization conditions: 1040 mg TEA, 50 mg Donor.


Values in italics are calculated from intrinsic viscosity data.

Primary parameter estimation Polymerization rate


The simplied model for constant monomer and hydrogen Figure 5 shows the rate-time proles for different tempera-
concentrations2,3,37 is: tures and two hydrogen concentrations, namely, 0.0 mg (Runs
1, 4, and 12) and 150 mg (Runs 2, 7, and 13). It seems that the
R p R po expkd t (3) time between catalyst injection and reaching the maximum is
quite uncertain at low polymerization rates (zero hydrogen),
and can be used to t a given isothermal polymerization rate whereas at high polymerization rates this time interval slightly
prole with two parameters, Rpo and kd. The results of all 15 decreases with increasing temperature. Obviously, this is of
kinetic experiments are summarized in Table 2. minor importance for the kinetic interpretation: Y/A2 has a
value of approximately 2.0 for all experiments. Another effect
seems to be more important: the higher the polymerization rate,
the faster the catalyst decay, as will be seen in the sequel.
Figure 6 presents the Arrhenius plot of the initial polymer-
ization rates Rpo at different hydrogen concentrations. As ex-
pected, the polymerization rate decreases with decreasing tem-
perature for different hydrogen concentrations, but two issues
should be emphasized:
The activation energy is virtually independent of the hy-
drogen concentration. The calculated activation energies are
58.6, 63.6, and 56.7 kJ/mol for 0.0, 150, and 1000 mg H2,
respectively, in the temperature range of 60-80C. These values
are consistent with the literature data.2,8,38 The relative hydro-
gen response effect in terms of polymerization rate is the same
for different temperatures.

Figure 3. Area underneath the isoperibolic T curve of


run 7: (a) A1; (b) A2.
A1 left, A2 right hand side. [Color gure can be viewed in the
online issue, which is available at www.interscience. Figure 4. Temperature proles for the long-term poly-
wiley.com.] merization experiment and the short-term one.

AIChE Journal May 2006 Vol. 52, No. 5 Published on behalf of the AIChE DOI 10.1002/aic 1869
Figure 7. Polymerization rate-time proles for different
Figure 5. Polymerization rate-time proles at different H2 concentrations and temperatures.
temperatures. (a) 70C (Runs 4-10), (b) 60C (Runs 1-3), (c) 80C (Runs
13-15). [Color gure can be viewed in the online issue, which
(a) 0.0 mg H2 (Runs 1, 4, and 12), (b) 150 mg H2 (Runs 2, 7,
11, and 13). [Color gure can be viewed in the online issue, is available at www.interscience.wiley.com.]
which is available at www.interscience.wiley.com.]
value of about 75, 155, and 252 kg/gcat.hr for 60, 70, and 80C,
In the fully lled reactor, the initial polymerization rate respectively. Compared to the initial polymerization rates at
does not atten out or decrease with increasing temperature at zero hydrogen, this provides acceleration factors of 8.8, 9.7,
relatively high temperatures. Note that this is not always the and 9 for 60, 70, and 80C, respectivelya constant, temper-
case for experiments in a partially lled reactor in the presence ature independent value of about 9. This can be interpreted as
of a monomer gas phase operated under similar conditions (see follows:
2,38). For this catalyst, about 1/9 (11%) of the potentially active
Figure 7 shows the rate proles for different hydrogen con- sites are really active if one assumes that all potentially active
centrations at 60, 70, and 80C. The presence of hydrogen in sites are active at the hydrogen plateau. This is in agreement
the reactor signicantly increases the catalyst activity for all with the kinetic data discussed by Rishina and coworkers.9 This
three temperatures studied. part is temperature independent, at least within the temperature
In the absence of hydrogen, the catalyst shows low activity, and concentration ranges measured in this research. The acti-
which increases strongly with increasing hydrogen concentra- vation energy of the initial polymerization rate is hydrogen
tion at molar fractions, X, below 0.001, corresponding to 150 independent; see Figure 8. Therefore, it can be assumed that the
mg hydrogen injection in our experiments (see Table 2 and same ratio of potentially active sites is always blocked at zero
Figure 8); however, above this value, the impact of hydrogen hydrogen. Adding a given amount of hydrogen leads to similar
levels off, and the initial reaction rate reaches a maximum acceleration effects at different temperatures, for example, at

Figure 6. Arrhenius plot for the initial reaction rate at Figure 8. The effect of hydrogen on the initial reaction
different H2 concentrations. rate, Rpo.
[Color gure can be viewed in the online issue, which is [Color gure can be viewed in the online issue, which is
available at www.interscience.wiley.com.] available at www.interscience.wiley.com.]

1870 DOI 10.1002/aic Published on behalf of the AIChE May 2006 Vol. 52, No. 5 AIChE Journal
Figure 9. Hydrogen function for different temperatures.
Figure 11. Deactivation constant versus initial polymer-
ization rate, all runs, see Table 2.
150 mg hydrogen, the initial polymerization rates are 54.7, [Color gure can be viewed in the online issue, which is
121.6, and 205.4 kg/gcat hr for 60, 70, and 80C, respectively. available at www.interscience.wiley.com.]
If one divides these values by the corresponding values at zero
hydrogen, then one would get 6.4; 7.5, and 7.3 for 60, 70, and
80C, respectivelyagain, a nearly constant (temperature in- tion at zero hydrogen is caused by the high probability of being
dependent) acceleration factor of about 7 under these condi- dormant, because of low reactivation rates by monomer trans-
tions. Obviously, the curves shown in Figure 8 are shifted by fer and propagation reactions. This supports the idea discussed
adding hydrogen. Normalizing these curves leads to the mas- previously by Pater et al.2 and Roos et al.39 that the deactivation
ter curve shown in Figure 9, which provides the evidence of of the catalyst relates to the catalyst activity independent of the
the temperature independence of this shift effect that can be reason for the activity change (temperature, pressure, hydrogen
considered as a kinetic ngerprint for the given catalyst. concentration, etc.). These experimental ndings need to be
The plateau effect at high hydrogen concentrations has been interpreted.
conrmed by several authors.2,4,9,14 Guastalla and Giannini14 Figure 12 shows the Arrhenius plot of the lumped rst-order
proposed the presence of adsorption phenomena of hydrogen deactivation rate constant, kd. The plot is linear for all hydro-
on the surface of the solid catalyst, with a saturation at high gen concentrations measured; again, there is no change of the
concentrations. Rishina et al.9 suggested that a limiting maxi- slope at high temperatures. The results clearly indicate an
mum number of active centers is realized in the catalytic increase in the activation energy with the increasing level of
system in the presence of hydrogen. This is in agreement with hydrogen at low hydrogen concentrations, but a constant value
results presented by Samson et al.4 and Pater et al.,2 who near 40 kJ/mol if more than 150 mg H2 is used. Interestingly,
assumed that the concentration of the blocked sites is very low at zero hydrogen, the activation energy is in the range of a
at high hydrogen concentrations. Our results conrm these diffusion (limited) process.40,41 At zero hydrogen, the reactiva-
earlier results, now with more reliable data. tion of dormant sites is realized via monomer transfer or
The deactivation constant increases with hydrogen concen- propagation reactions. These reactions are very slow and show
tration, as can be seen in Figure 10. The shape of the curves is a low activation energy of 13.5 kJ/mol. The slope difference
quite similar to the shape of the Rpo curves shown in Figure 8. compared to high hydrogen concentrations is an indication that
Figure 11 shows a linear relationship between Rpo and kd for a second deactivation step dominates the temperature depen-
all experiments: the more monomer converted by an active site, dence of kd shown in Figure 12. According to Guyot et al.,32
the higher the probability of deactivation, but independent of the coordinated propylene monomer is either inserted into the
the reason for the higher polymerization rate. This way the growing polymer or decomposes into an inactive -allyl spe-
deactivation can be interpreted as activity-dependent proba- cies and H atom that is added to the growing chain. The
bility. Additionally, one can conclude that an active site formation of -allyl species, according to the authors, accounts
cannot deactivate during its dormant state. The low deactiva-

Figure 10. The inuence of hydrogen on the deactivation Figure 12. Arrhenius plot for deactivation constant at
constant. different H2 concentrations.
[Color gure can be viewed in the online issue, which is [Color gure can be viewed in the online issue, which is
available at www.interscience.wiley.com.] available at www.interscience.wiley.com.]

AIChE Journal May 2006 Vol. 52, No. 5 Published on behalf of the AIChE DOI 10.1002/aic 1871
for the high deactivation observed at high temperature. In terms kd
of our results, this means: if -allyl species are responsible for Deactivation Cj Dj (10)
the second deactivation step mentioned above, then both the
higher deactivation and higher activation energy in the pres-
ence of hydrogen can be explained. Using the long chain hypothesis, the polymerization rate can
be described as:
Molecular weight distribution
The effects of hydrogen concentration and temperature on Rp kp C Cm (11)
molecular weight are summarized in Table 2. Similarly to other
Ti-based polymerization catalysts,8,21,22,24,42 the molecular A sorption model should be used to describe the monomer
weight decreases with increasing hydrogen concentration. Con- concentration near the active sites, but a simplied model
trary to what is usually stated in different articles,8,43,44 the using the concentration of liquid propylene directly can be
molecular weight of the produced polymer increases with tem- used.34,48 The active catalyst site concentration, C, is the dif-
perature. It is evident that the activation energy of the propa- ference between the maximum concentration of active sites,
gation reaction is higher than that of the transfer reaction under Cmax, and the concentration of dormant sites, CS:
our conditions. Chadwick et al.45 found that a 2,1-inserted
center remains in the dormant state for a signicant period of
C C max CS (12)
time even in the presence of hydrogen, and that the temperature
increase can reduce the barrier to chain propagation after this
misinsertion. This implies, according to the authors, that the The concentration of dormant sites can be calculated assum-
molecular weight of the isotactic fraction may increase with ing the quasi-steady-state:
increasing polymerization temperature, which is in agreement
with what we found. The polydispersity trend indicates a slight R S 0 k S C C m k reh CS CH2 krem CS Cm (13)
narrowing of the MWD with increasing polymerization tem-
perature, as can be seen in Table 2. This also indicates chang- Rearrangement of these equations results in:
ing contributions of different active sites (different activation
energies) with changing temperature; see 43,45,46.
kS C
CS (14)
k reh X krem
Modeling
The modeling of reaction kinetics and molecular weight is where X is the hydrogen-monomer molar ratio:
based on the dormant site mechanism47 using an averaged-
site approach based on the following mechanism:
C H2
X (15)
Cm
kp

Chain propagation Cj M Cj1 (4)


Combining Eqs. 12 and 14 leads to:

C max 1 k1 X
kh
C (16)
Chain transfer to hydrogen Cj H2 Co Dj (5) 1 k2 k1 X

with
km

Chain transfer to monomer Cj M O


C1 Dj (6) k reh kS
k1 ; k2 (17)
krem krem

ks This way the polymerization rate can be described as a


Dormant site formation Cj M Sj1 (7) function of three parameters, k1, k2, and kp:

k P C m C max k1 X 1
RP (18)
kreh 1 k1 X k2
Dormant sites reactivation by H2 Sj H2 O
Co Dj (8)
or

krem R p k p C max Cm fH2 (19)


Reactivation of dormant sites by monomer Sj M O
Cj1
with the hydrogen dependent function f H2
(9) dened as:

1872 DOI 10.1002/aic Published on behalf of the AIChE May 2006 Vol. 52, No. 5 AIChE Journal
1 k1 X
f H2 (20)
1 k2 k1 X

f H2 represents the fraction of sites that are active under certain


conditions, showing a minimum at X 0 and a maximum,
fH2,max 1, at high values of X where all sites are activated.
The quasi-single-site approach results in an average ter-
mination probability:

2 MM
q (21)
MW Figure 13. Calculation of propagation constant for pro-
pylene polymerization.
For dormant sites of chain length j, CS,j, it holds that
iments without hydrogen, Runs 1, 4, and 12, Eq. 18 can be
R S, j 0 k S C j1 C m k reh CS, j CH2 krem CS, j Cm rewritten as:
(22)
kp
Rp,o (28)
which gives: 1 k2 m

k S C j1 With known values for m (monomer density, kgm3), one


C S, j (23) can easily estimate the ratio kp/(1 k2), and a good Arrhenius
k reh X krem
t (Figure 13) resulted with:


The balance of the active sites of length j, Cj, can be written
as: Ep
k p k po 1 k 2 exp (29)
RT
R C, j 0 kp Cj Cm kp Cj1 Cm kS Cj Cm
with
krem CS, j Cm kh Cj CH2 km Cj Cm (24)
k po 6.41 10 8 m3 /gcat hr (30)
Substituting CS,j in Eq. 23 into Eq. 24 gives:
E p 67.22 10 3 J/mol (31)
C j p C j1 (25)
Nonlinear estimation of k1 and k2 gives the following results:
with the chain propagation probability, p:
k 1 32.8 T2 2.26 104 T 3.86 106
k S/k p
1 k 2 8.02
X k reh/krem 1 (32)
p (26)
kS k h km
1 X Figure 14 shows a comparison between the tted model and
kp kp kp
the experimental data at 70 and 60C.
Figure 11, as discussed earlier, shows an almost linear rela-
The termination (or chain transfer) probability q is dened
by:

q1p
km kh ks
X
kp kp
k1 X
kp 1 k1 X (27)

with k1 dened by Eq. 17.


One can model the weight average molecular weight Mw by
combining Eqs. 21 and 27.

Parameter estimation
Equation 18 has three parameters to be determined, namely,
kp, k1, and k2. It should be emphasized that kp is an average,
because of the active sites heterogeneity; additionally, the true Figure 14. Comparison of model and experiment.
number of active sites and the true (in-situ) sorption is not [Color gure can be viewed in the online issue, which is
taken into account and will inuence the kp value. For exper- available at www.interscience.wiley.com.]

AIChE Journal May 2006 Vol. 52, No. 5 Published on behalf of the AIChE DOI 10.1002/aic 1873
Figure 16. Test of different molecular weight models.
Figure 15. Comparison of model and experiment: kd at [Color gure can be viewed in the online issue, which is
available at www.interscience.wiley.com.]
60, 70, and 80C.
[Color gure can be viewed in the online issue, which is
available at www.interscience.wiley.com.]
soning for this deviation is still unclear. More research is
required to interpret these ndings.
tionship between kd and Rpo. This suggests that the relationship According to the literature search above, one of the follow-
between these variables can be described by a two-term func- ing models can be used to describe the average molecular
tion. The rst term describes the linear behavior; meanwhile, weight:
the second term compensates for the deviation from the linear- (1) Square-root model, that is, Natta model;
ity. The following empirical relationship gives the best t, as (2) Linear model, that is, Mayo model;
shown in Figure 15: (3) The dormant site model.34,48
The dormant site model gives the best t; see Table 3 and


k d k d,1 R po k d,2 exp
Ek d
RT
1 kd,3 X (33)
Figure 16, with:

q 2.077 10 5 0.0448 X
where 15000 X
1.28 10 4 (38)
3
1 15000 X
k d,1 8.38 10 gcat/kg (34)
Conclusions
k d,2 1.7 10 4 1/hr (35)
Quasi-isothermal liquid propylene polymerization has been
studied in a fully lled (over-pressured) batch reactor using a
k d,3 288.2 (36)
prepolymerized Ziegler-Natta catalyst with TEA as cocatalyst
at different temperatures (60-80C) and hydrogen concentra-
E kd 20 103 J/mol (37) tions (0-100 mmol/L). Polymerization rate proles and molec-
ular weight distributions have been measured and modeled
In Figure 15 the experimental data are compared to predic- with the following results:
tions using Eq. 33. As can be seen, the t between model and (1) The adiabatic temperature rise in isoperibolic operation
experimental data is good at the three polymerization temper- allows the estimation of initial polymerization rates after about
atures. Interestingly, this gure shows, also, that at low hydro- 10 seconds. The catalyst used shows no signicant activation
gen concentration, 150 mg, the dependence of kd on Rpo can after injection.
be described reasonably well using a linear relationship. How- (2) Isoperibolic mode allows on-line estimations of the
ever, increasing hydrogen concentration introduces deviations polymerization rate.
from linearity; where such a deviation becomes noticeable (3) The polymerization rate increases with increasing hy-
depends on the polymerization temperature. The physical rea- drogen, reaching a plateau at high hydrogen concentrations that

Table 3. Parameters of Different Average Molecular Weight Models


Model A B C D R2 SSE*
5
Linear model q A B X 3.8 10 0.0545 0.94 4.13 108
Natta model q A B X/C m 2.0 105 0.616 0.96 2.73 108
(mol/l)0.5
This article q A B X 2.077 105 0.0448 1.28 104 15000 0.99 8.02 109
C [D X/(1 D X)]

*SSE: Sum of squares due to error.50

1874 DOI 10.1002/aic Published on behalf of the AIChE May 2006 Vol. 52, No. 5 AIChE Journal
is about 9 times higher than polymerization rates at zero tion method on polymerization kinetics. Chem Eng Sci. 2002;57:3461-
hydrogen. This effect does not depend on temperature. 3477.
3. Shimizu F, Pater JTM, Van Swaaij WPM, Weickert G. Kinetic study
(4) The pseudo-rst-order deactivation constant is linearly of a highly active MgCl2-supported Ziegler-Natta catalyst in liquid
dependent on initial polymerization rates. pool propylene polymerization. II. The inuence of alkyl aluminum
(5) The rate of polymerization and the deactivation constant and alkoxysilane on catalyst activation and deactivation. J Applied
increase with temperature up to 80C. The average activation Polym Sci. 2002;83:2669-2679.
energy of the polymerization rate is 58 kJ/mol, while the 4. Samson JJC, Bosman PJ, Weickert G, Westerterp KR. Liquid-phase
polymerization of propylene with a highly active Ziegler-Natta cata-
average activation energy of the deactivation process is con- lyst. Inuence of hydrogen, cocatalyst, and electron donor on reaction
siderably lower and is dependent on the hydrogen concentra- kinetics. J Polym Sci A: Polym Chem. 1999;37:219-232.
tion. 5. Lim S, Choung S. Studies on the catalytic death in propylene poly-
(6) The polymer molecular weight decreases with hydrogen merization. Appl Catalysis A: General. 1997;153:103-118.
6. Han-Adebekun GC, Hamba M, Ray WH. Kinetic study of gas phase
concentration, but increases with polymerization temperature.
olen polymerization with a TiCl4/MgCl2 catalyst I. effect of poly-
(7) The polydispersity index increases slightly with the merization conditions. J Polym Sci A: Polym Chem. 1997;35:2063-
hydrogen concentration, but decreases with temperature. 2074.
(8) Both the polymerization rate and the molecular weight 7. Forte MC, Coutinho FMB. The inuence of catalyst system and
can be modeled using the dormant sites theory. polymerization conditions on polypropylene properties. European
Polym J. 1996;32:605-611.
8. Soares JB, Hamielec A. kinetics of propylene polymerization with a
non-supported heterogeneous Ziegler-Natta catalyst: effect of hydro-
Acknowledgments gen on rate of polymerization, stereoregularity, and molecular weight
This work has been funded by Dutch Polymer Institute (DPI), project distribution. Polymer. 1996;37:4607-4614.
DPI #114. We greatly acknowledge R. Emonds and K. van Bree for the 9. Rishina LA, Vizen EI, Sosnovskaja LN, Dyachkovsky FS. Study of the
technical assistance. The authors further wish to thank Basell for the effect of hydrogen in propylene polymerization with MgCl2-supported
analytical support. Ziegler-Natta catalyst: Part 1. Kinetics of polymerization. European
Polym J. 1994;30:1309-1313.
10. Chien J, Weber S, Hu Y. Magnesium chloride supported catalysts for
Notation olen polymerization. XIX. Titanium oxidation states, catalyst deac-
tivation, and active site structure. J Polym Sci A: Polym Chem. 1989;
C activated catalyst concentration, mol/L 27:1499-1514.
Cmax concentration of active sites, mol/L 11. Chien J, Kuo C. Magnesium chloride supported high-mileage catalysts
CS concentration of dormant sites, mol/L for olen polymerization. X. Effect of hydrogen and catalytic site
Cm concentration of monomer, mol/l deactivation. J Polym Sci A: Polym Chem. 1986;24:2707-2727.
CH2 hydrogen concentration, mol/l 12. Chien J, Kuo C. Magnesium chloride supported high mileage catalyst
Cj growing polymer chain containing j monomers for olen polymerization. VII. Determination of concentration of tita-
E activation energy, J/mol nium polymer and aluminum-polymer bonds. J Polym Sci: Polym
Dj dead (terminated) polymer chain containing j monomers Chem Ed. 1985;23:731-760.
fH2 fraction of potentially active sites 13. Chien J, Kuo C. Magnesium chloride supported high mileage catalyst
k1 kreh/krem for olen polymerization. VIII. Decay and transformation of active
k2 ks/krem sites. J Polym Sci Polym Chem. 1985;23:761-786.
kd deactivation rate constant, 1/h 14. Guastalla G. Gianinni U. The inuence of hydrogen on the polymer-
kh chain transfer to hydrogen rate constant, l/mol. s ization of propylene and ethylene with an MgCl2 supported catalyst.
km chain transfer to monomer rate constant, l/mol. s Macromol Chemie-Rapid Commun. 1983;4:519-527.
kp propagation constant, l/mol. s 15. Keii T, Suzuki E, Tamura M, Murata M, Doi Y. Propene polymeriza-
kp propagation constant, m3/gcat. h tion with a magnesium chloride supported Ziegler catalyst: 1. Principal
kreh dormant site reactivation by hydrogen rate constant, l/mol. s
kinetics. Macromol Chem Phys. 1982;183:2285-2304.
krem dormant site reactivation by monomer rate constant, l/mol. s
16. Yuan HG, Taylor TW, Choi KY, Ray WH. Polymerization of olens
MM monomer molecular weight, g/mol
through heterogeneous catalysis. I. Low pressure propylene polymer-
Mn number average molecular weight, g/mol
ization in slurry with Ziegler-Natta catalyst. J Appl Polym Sci. 1982;
Mv viscosity average molecular weight, g/mol
27:1691-1706.
Mw weight average molecular weight, g/mol
17. Naga N, Mizunuma K. Stereochemical control in propylene polymer-
p propagation probability
ization with non-bridged metallocene dichloride/MAO. Polymer.
q termination probability
1998;39:2703-2708.
PDI polydispersity index
18. Spaleck W, Kuber F, Bachmann B, Fritze C, Winter A. New bridged
Rp polymerization rate, mol/l. s
zirconocenes for olen polymerization: binuclear and hybrid struc-
Rp polymerization rate, kg/gcat h
tures. J Molecular Catalysis A: Chem. 1998;128:279-287.
Rp,ATR polymerization rate based on adiabatic raise in temperature,
19. Al-haj Ali M, Betlem B, Roffel B, Weickert G. Propylene polymer-
kg/gcat. h
ization in a full-lled batch reactor. In preparation.
T polymerization temperature, C
20. Natta G, Mazzanti G, Longi P, Bernardini F. Inuenza dellidrogeno
t time, h
sulla polimerizzazione anionica coordinata del propilene e dell
X hydrogen molar ratio
etilene. La Chimica E LIndustria. 1959;41:519-526.
Y yield, gPP
21. Ross JF. Hydrogen as a chain terminator in continuous gas phase
m monomer density, kg/m3
polymerization of propylene. J Polym Sci A: Polym Chem. 1984;22:
2255-2263.
22. Keii T, Doi Y, Suzuki E, Tamura M, Murata M, Soga K. Propene
Literature Cited polymerization with a magnesium chloride-supported Ziegler catalyst,
1. Pater JTM, Weickert G, van Swaaij WPM. Propene bulk polymeriza- 2: Molecular weight distribution. Macromol Chem Phys. 1984;185:
tion kinetics: role of prepolymerization and hydrogen. AIChE J. 2003; 1537-1557.
49:180-193. 23. Soga K, Siono T. Effect of hydrogen on the molecular weight of
2. Pater JTM, Weickert G, van Swaaij WPM. Polymerization of liquid polypropylene with Ziegler-Natta catalysts. Polym Bull. 1982;8: 261-
propylene with a fourth generation Ziegler-Natta catalyst: inuence of 268.
temperature, hydrogen, monomer concentration, and prepolymeriza- 24. Kissin YV, Rishina LA, Vizen EI. Hydrogen effect in propylene

AIChE Journal May 2006 Vol. 52, No. 5 Published on behalf of the AIChE DOI 10.1002/aic 1875
polymerization reactions with titanium-based Ziegler-Natta catalysts. 38. Shimizu F. Liquid Pool Polymerization of Propylene and Ethylene
II. Mechanism of the chain-transfer reaction. J Polym Sci A: Polym Using a Highly Active Ziegler-Natta Catalyst, PhD Diss. Dept of
Chem. 2002;40:1899-1911. Science and Technology. University of Twente, Enschede, The Neth-
25. Mori H, Endo M, Tashino K, Terano M. Study of activity enhancement erlands, 2002..
by hydrogen in propylene polymerization using stopped-ow and 39. Roos P, Meier GB, Samson JJC, Weickert G. Gas phase polymeriza-
conventional methods. J Molecular Catalysis A-Chem. 1999;145:153- tion of ethylene with a silica-supported metallocene catalyst: inuence
158. of temperature on deactivation. Macromol Chemie-Rapid Commun.
26. Busico V, Cipullo R, Corradini P. Ziegler-Natta oligomerization of 1997;18:319-324.
1-alkenes: a catalysts ngerprint, 1. Macromol Chem Phys. 1993; 40. Snijder ED, te Riele MJM, Versteeg GF, van Swaaij WPM. Diffusion
194:1079-1093. coefcients of several aqueous alkanolamine solutions. J Chem Eng
27. Kashiwa N, Yoshitake J. The inuence of the valance state of titanium Data. 1993;38:475-480.
in MgCl2-supported titanium catalysts on olen polymerization. Mak- 41. Scheible EG. Liquid diffusivities. Industrial Eng Chem Research.
romolekulare Chemie. 1984;185:1133-1138. 1954;46:2007-2010.
28. Busico V, Corradini P, Ferraro A, Proto A. Polymerization of propene 42. Meier GB, Weickert G, Van Swaaij WPM. Gas-phase polymerization
in the presence of MgCl2-supported Ziegler-Natta catalysts, 3: catalyst of propylene: reaction kinetics and molecular weight distribution. J
deactivation. Makromolekulare Chemie. 1986;187:1125-1130. Polym Sci A: Polym Chem. 2001;39:500-513.
29. Albizzati E, Gianinni U, Collina G, Noristi L, Resconi L. Catalysts and 43. Kissin YV, Ohnishi R, Konakazawa T. Propylene polymerization with
polymerizations. In: Polypropylene Handbook. Moore E, ed. Munich: titanium-based Ziegler-Natta catalysts: effects of temperature and
Carl Hanser Verlag; 1996:74-89. modiers on molecular weight, molecular weight distribution and
30. Albizzati E, Giannini U, Balbontin G, Camurati I, Chadwick JC, stereospecicity. Macromol Chem Phys. 2004;205:284-301.
Dallocco T, Dubitsky Y, Galimberti M, Morini G, Maldotti A. Pro- 44. Yang CB, Hsu CC. Propene polymerization with MgCl2-supported
pylene polymerization with catalysts containing divalent titanium. J TiCl4/dioctylphthalate catalyst. III. Effects of polymerization condi-
Polym Sci A: Polym Chem. 1997;35:2645-2652. tions on molecular weights and molecular weight distribution. J Appl
31. Chien J, Nozaki T. High activity magnesium chloride supported cat- Polym Sci. 1995;58:1245-1254.
alysts for olen polymerization. XXIX. Molecular basis of hydrogen 45. Chadwick JC, Morini G, Balbontin G, Busico V, Talarico G, Sudmei-
activation of magnesium chloride supported Ziegler-Natta catalysts. J jer O. Advances in propene polymerization using magnesium chloride-
Polym Sci A: Polym Chem. 1991;29:505-514. supported catalysts. ACS Symposium Series 749: Olen Polymeriza-
32. Guyot A, Spitz R, Dassaud J, Gomez C. Mechanism of deactivation tion Emerging Frontiers. 2000;50-65.
and reactivation of Ziegler-Natta catalysts for propene polymerization. 46. Chien J. Olen polymerizations and polyolen molecular weight dis-
J Molecular Catalysis. 1993;82:29-36. tribution. J Polym Sci A: General Papers. 1963;1:1839-1856.
33. Moritz HU. Polymerization calorimetrya powerful tool for reactor 47. Weickert G. Modeling of Polymerization Rate and Molecular Weight:
control. In: Polymer Reaction Engineering. Berlin: VCH Weinheim; H2 Response. Internal report. Enschede, The Netherlands: IPP group,
1989:248-266. Twente University; 2002.
34. Weickert G. High precision polymerization rate proles. Presented at 48. Banat Y, Veera UP, Weickert G. The role of equilibrium sorption in
3rd International Workshop on Heterogeneous Ziegler-Natta Catalysts, gas and liquid phase propylene polymerization. To be submitted to
Kanazawa, Japan, 2003. Chem Eng Sci.
35. Tait PJT, Watkins ND. Monoalkene polymerization: mechanisms. In: 49. Hindryckx F, Dubois P, Jerome R, Marti MG. Ethylene polymerization
Comprehensive Polymer Science. Oxford: Pergamon Press; 1989:533- by a high activity MgCl2 supported Ti catalyst in the presence of
573. hydrogen and/or 1-octene. Polymer. 1998;39:621-629.
36. Keii T. Kinetics of Ziegler-Natta Polymerization. Tokyo: Kodansha 50. Constantinides A, Mostou N. Numerical Methods for Chemical En-
LTD; 1972. gineers with MATLAB Applications. Upper Saddle River, NJ: Prentice
37. Samson JJC, Weickert G, Heerze A, Westerterp KR. Liquid-phase Hall PTR; 1999.
polymerization of propylene with a highly active catalyst. AIChE J.
1998;44:1424-1437. Manuscript received July 6, 2005, and revision received Dec. 13, 2005.

1876 DOI 10.1002/aic Published on behalf of the AIChE May 2006 Vol. 52, No. 5 AIChE Journal

Das könnte Ihnen auch gefallen