You are on page 1of 9

ISOTHERMAL LIQUID PROPYLENE POLYMERIZATION WITH A HIGHLY ACTIVE ZIEGLERNATTA CATALYST IN A FULLY-FILLED BATCH REACTOR

M. Al-haj Ali, B. Betlem, G. Weickert, and B. Roffel Department of Science and Technology University of Twente P O Box 217, 7500 AE Enschede, The Netherlands E-mail: m.alhajali@tnw.utwente.nl

Abstract: This paper presents an experimental kinetic study of liquid propylene polymerization in a totally-filled batch reactor using a highly-active MgCl2/TiCl4/phthalate/silane/AlR3 catalyst. The effect of reactor filling on the polymerization kinetics has been investigated. In addition, the presence of characteristic pressure profiles is used to estimate the polymerization rate for a fully-filled reactor. This approach is known as a dilatometric method. The comparison of its prediction with values obtained form the use of the calorimetric method shows good agreement. The influence of hydrogen concentration (0-10 mmol/mol) on polymerization rate profiles and molecular weight distribution has also been investigated at a temperature of 70 C. It was shown that reaction rate increases rapidly with hydrogen concentration at low concentrations. At higher hydrogen amounts (> 1.4 mmol/mol), this effect disappears and a maximum rate is reached. It was also shown that the pseudo-first order deactivation constant increases with hydrogen concentration. The molecular weight increases with hydrogen concentration but the polydispersity index can be assumed effectively as a constant. These observations can be explained quantitatively by the re-activation of dormant sites which are produced after 2,1 mis-insertions. Polymerization rate, deactivation constant and average molecular weight were modeled using the dormant site mechanism.

1. INTRODUCTION In 2003, the worldwide manufacturing capacity for polypropylene exceeded 40 mega tones (Galli and Vecellio, 2004), more than 95% of which has been produced by means of supported Z-N catalysts. Polypropylene is produced using three processes: (i) slurry process, (ii) liquid process, and (iii) gas process. Since a few decades, the liquid-phase process becomes one of the most important industrial processes in polypropylene polymerization. The kinetics of propylene liquid-pool polymerization is affected by a large number of catalyst and process variables. Catalyst variations include: catalyst and cocatalyst types and concentrations, and the presence and type of the external donor. Meanwhile, process

variables include the temperature, hydrogen concentration and polymerization time. The introduction of hydrogen gas during propylene polymerization reaction with supported Z-N catalysts results in a considerable increase in the polymerization rate, in addition to decreasing the molecular weight of the polymer produced. Because of this important role of hydrogen, many studies have been carried out to understand the enhancement of the polymerization rate caused by hydrogen. However, only a very few experimental studies have been carried out to investigate the effect of introducing hydrogen gas during liquid propylene polymerization so far; most studies were carried out in either gaseous monomer or diluents such as heptane, i.e. slurry phase polymerization, (Han-Adebekun et al., 1997 ). The first paper reporting rate profiles in catalytic
1

liquid propylene polymerization has been published by Samson et al. (1999). Besides this paper, a few more authors have measured polymerization rate profiles in liquid propylene and investigated the hydrogen effect on kinetics, see (Pater et al., 2002). In these studies, the reactor was always partially-filled with liquid propylene and charged with different hydrogen amounts, that means polymerizations were carried out in the presence of a gas phase. Depending on the details of the experimental procedure, the use of a partially filled reactor could have some drawbacks, especially, when a prepolymerization step is required. A comprehensive comparison between liquid propylene polymerization in partially-filled and fully-filled reactor can be found in (Al-haj Ali et al., 2005). The effect of hydrogen on the polymer molecular weight was studied by many authors. Natta et al., (1959) studied this dependency for a TiCl3/Al(C2H5)3 catalyst. Their results showed that the number average molecular weight for both polyethylene and polypropylene could be described by the square root of the hydrogen partial pressure in the reactor, Equation (1):
Mn = 1 K 1 + K 2 PH
2

(1)

TI

injection needles
Eurother m PID controller

Cold water supply Cold water bath

To gas analysis section

Purge

BUCHI I
Oil system for heating

Return Warm water bath

2. MATERIALS AND EXPERIMENTAL SET-UP Propylene used in the experiments was polymerization grade with a purity of more than
Fig. 1 Liquid-pool polymerization set-up

This dependency was examined later for olefin polymerization with different catalysts. Other researchers (Keii et al., 1984; Hindryckx et al., 1998) showed that it gives the best fit for their experimental data; however, others prove that it is not applicable (Meier et al., 2001). In this contribution we focus on studying liquidpool propylene polymerization in a batch reactor, which is completely filled with liquid. These experiments were performed under nearindustrial conditions using a highly active Ziegler-Natta catalyst. First we will study the effect of reactor filling on polymerization kinetics. Then, we will investigate the effect of hydrogen concentration and reaction temperature on the polymerization kinetics, in terms of the polymerization rate and the pseudo-first order deactivation constant. In addition, the effect of these conditions on the produced polymer properties, namely the average molecular weight and the polydispersity index, will be investigated. Finally, using dormant-site theory, these variables will be modeled as functions of the studied process variables.

99.5%, with propane as a main impurity. Hydrogen used had purity higher than 99.999%; it was further purified by passing over a reduced BTS copper catalyst, and subsequent passing through three different beds of molecular sieves, with pore sizes of 13, 4 and 3 , respectively. Propylene was purified in the same way; additionally it was passed over a bed of oxidized BTS catalyst to remove CO. TiCl4 supported on MgCl2 with phthalate as internal donor and an external silane donor (Dex) was used as a catalyst with triethylaluminum (TEA) as a cocatalyst and a scavenger. The reactor used in this work is a 5-liter stainless steel jacketed batch reactor with a separately heated cover plate (Figure 1). It is equipped with an overpressure valve and rupture disc, the maximum allowable working pressure was 60 bar. For intensive mixing, the reactor is equipped with a turbine stirrer mounted on a hollow stirrer shaft. The temperature control system comprises separate PID controllers for the reactor cover plate and the reactor jacket. The cooling medium temperature is kept constant within a narrow range of 0.01 K during the isoperibolic experiments. The reactor is filled with liquid propylene and heated up to the reaction temperature. When the temperature reaches the set point, hydrogen is injected. The reactor temperature and pressure are monitored as a function of time. As soon as both became stable for an interval of three minutes, the reaction is started by injecting the prepared catalyst into the reactor. The heat of polymerization is measured under quasisteady state conditions. Data are collected every three seconds. The polymerization reaction is finally terminated by rapidly flushing the unreacted propylene. After each experiment, the resulting polymer is dried under vacuum and at 50C for 4 hours.
From liquid propylene MFC

From hydrogen MFC PI Vacuum pump From catalyst injection system

plunger pump

bottle to be injected

The molecular weight distribution has been measured by Gel Permeation Chromatography (GPC) employing a Waters Alliance GPCV 2000 apparatus with TSK columns at 155 C using 1,2,4-trichlorobenzene as a solvent. Average molecular weights above one million grams per mol have been measured using calibrated intrinsic viscosities. 3. RESULTS AND DISCUSSION 3.1 Reproducibility The reproducibility of the experiments has been tested by repeating a standard experiment at 70 C and 43 bar with 150 mg hydrogen at fixed concentrations of the catalyst, cocatalyst and external donor. Figure 2 shows the time profiles of the polymerization rates for three standard experiments. Obviously, the absence of a gas phase, i.e. absence of any mass transfer between gas and liquid phase, increases the accuracy of such rate profile measurements. Note that the polymerization rate is calculated from a quasi-steady state heat balance. However, during the first 1.5 minutes, this steady state situation has not been realized; therefore, calculations made during this period are incorrect. We will not discuss this issue in this contribution, a detailed discussion can be found in (Al-haj Ali et al., 2005).
120 100 Rp, kg/gcat. h 80 60

H2. It is clear that the filling degree has no effect in this case, the deviation falls within the experimental reproducibility at these low polymerization rates. However, the filling degree has a pronounced effect, when hydrogen is used in the polymerization reaction. In Figure 4, the reaction rate is shown at two different filling degrees using 150 mg hydrogen in both cases. When the reactor is not totally filled, hydrogen will be distributed in the gas and liquid phase. The equilibrium concentration of hydrogen in each phase depends on temperature. When the reactor is totally filled, the hydrogen concentration is higher, 0.144 mol % or 0.0149 mol/l compared to 0.122 mol % or 0.0103 mol/l in the half-filled reactor. Since hydrogen has an activation effect on the polymerization reaction, this difference in hydrogen concentration is the reason for the observed difference in the polymerization rate in both cases.
20 18 16 14 12 10 8 6 4 2 0 0 10

Totally-filled

Rp, kg/gcat. hr

Half-filled

Totally-filled Half-filled

20

30 Time, min

40

50

60

Fig. 3 Filling effect in the polymerization rate without using hydrogen. T = 70 C and mcat = 3.78 mg.
120

40 20 0 0 5 10 15 Time, min 20

Rp, kg/gcat. hr

Exp. 1 Exp. 2 Exp. 3

100 80 60 40 20
Half-filled

Totally-filled

25

30

Half-filled Totally-filled

Fig. 2 Polymerization rate-time profiles of repeated standard experiments. Experimental conditions: T = 70 C, mcat = 3.78 mg, mTEA = 1040 mg, mDex = 50 mg, and mH2 = 150 mg.

0 0 5 10 15 Time, min. 20 25 30

3.2 Effect of reactor filling on polymerization kinetics In investigating the filling effect, the concentration of the cocatalyst (0.2 g/l) and the donor (0.01 g/l) are kept constant; meanwhile, hydrogen concentration is varied. Two sets of experiments were performed one at 70 C; the other at 80 C. Figure 3 shows the reaction rate as a function of time for the experiments that were done without

Fig. 4 Filling effect in the polymerization rate using 150 mg H2. T = 70 C and mcat = 3.78 mg.

Figure 5 shows the polymerization profiles for equal hydrogen concentration (150 mg) at 70 and 80 C. It is clear that the polymerization rate increases considerably with an increase in temperature. However, it is known that reaction rate starts to decrease, when the polymerization temperature is increased above 70 C (Pater et al., 2002). This trend was found for partiallyfilled reactors. We checked the behavior of our
3

catalyst and found that it has a similar behavior (Figure 6). Since we keep reaction conditions the same for partially and fully-filled experiments, this suggests that the main reason for the increase in polymerization rate with temperature increase is the reactor filling.
200

150 Rp , kg/gcat . hr

80 C

100

70 C
`

50

0 0 5 10 15 Time, min. 20 25 30

Fig. 5 The effect of temperature on the polymerization reaction in a fully-filled reactor. mH2 = 150 mg. For 70 C run, mcat = 3.78 mg and for 80 C run, mcat = 1.54 mg.
100 Rp , kg/gcat . hr 80 60 40 20 0
80 C 70 C

In the case of propylene liquid-pool polymerization, the polymerization rate is usually determined using the calorimetric method. Interestingly, the production of polypropylene results in a shrinkage of the reactors content because of the large difference in polymer and monomer density. This shrinkage phenomenon is a kinetic signal and the experimental technique is known as Dilatometry. It is an accurate method for polymerization reactions with a large difference in density between monomer and polymer as long as the volume shrinkage can be related to the reaction kinetics. Since the reactor is operated filled with liquid monomer, volume shrinkage will lead to a sensitive decrease in the reactor pressure, Figure 7. This can be used to develop a new strategy to measure reaction rate profiles in liquid propylene polymerization (Weickert, 2003). This method can be implemented by either feeding the reactor continuously with monomer to keep the reactor pressure constant (compensation dilatometry) or monitoring the pressure decrease during the experiment (pressure-drop dilatometry).

60 55 P, bar
0 10 20 30 Time, min 40 50 60

81 P T
B

80 79 T, C 78 77
C

50 45
A

76 75 74

Fig. 6 The effect of temperature on the polymerization reaction in a partially-filled reactor. For 70 C run, mcat = 3.78 mg, and mH2 = 150 mg. For 80 C run, mcat = 1.54 mg and mH2 = 120 mg.

40 35 0 10 20 30 40 Time, min 50 60 70 80

A possible explanation is the overheating of the polymerizing particles in case of the partially filled reactor. In the partially filled reactors, the liquid and vapor systems are in equilibrium before catalyst injection. When the catalyst is injected, the temperature inside the particles increases due to the heat of polymerization. This will lead to the formation of gas bubbles around the polymerizing particles, consequently deactivating the active sites on the catalyst particles. However, when the polymerization is carried out in a fully-filled reactor, the polymerization reaction takes place under high pressure. Then, there is a large tolerance for temperature increase before reaching the vaporization point. More detailed discussion about this topic can be found in (Al-haj Ali et al., 2005). 3.4 Estimation of polymerization rate using the dilatometric method

Fig. 7 Experimental pressure and temperature profiles in a full-filled reactor. A. hydrogen addition. B. cocatalyst injection. C. catalyst injection.

In Figure 8 the time versus polymerization rate plot is shown. From this plot, we can see that the curves resulting from dilatometric (pressuredrop mode) and the corresponding calorimetric data are almost congruent. 3.3 Influence of H2 polymerization kinetics concentration on

The simplified model for constant monomer and hydrogen concentrations (Samson et al., 1998) is:
R p = R po exp( k d t )

(2)

It can be used, to fit a given isothermal polymerization rate profile with two parameters Rpo and kd. The experimental conditions and the
4

25 Rp, Kg/g cat . hr 20 15 10 5 0 0

Dilatometry

Calorimetry

of hydrogen. This is in agreement with results presented by Pater et al. (2002) who assumed that the concentration of the blocked sites is very low at high hydrogen concentrations - our results confirm these earlier results, now with more reliable data.
20 Time, min.
140 120 100 80 60 40 20 0 0 5
1000 150 50 25 0

10

30

40

50

Fig. 8 Pressure-drop dilatometry versus calorimetry for propylene polymerization. Experimental conditions: T = 70 C, mcat = 3.78 mg, mTEA = 1040 mg, mDex = 50 mg, and mH2 = 0.0 mg.

Rp, kg/ g cat . hr

10

15 Time, min

20

25

30

Table 1 Propylene polymerization: experimental * conditions and kinetic results.

Run no. 1 2 3 4 5 6 7

H2 mg 0 25 50 150 250 500 1000

X*103

Fig. 9 Polymerization rate-time profiles at different hydrogen concentrations.

tr min 60 60 60 35 47 45 45

R po, kg/g cat . hr

0 0.24 0.49 1.44 2.47 5.15 9.94

Rpo kg/gcat. hr 16.1 62.5 81.3 121.6 145.1 154.1 139.6

kd 1/hr 0.342 0.798 0.948 1.188 1.53 1.764 1.932


Fig. 10 The effect polymerization rate, Rpo. of hydrogen on initial
180 150 120 90 60 30 0 0 200 400 600 800 1000 1200 Hydrogen added, mg

* Other polymerization conditions: mTEA = 1040 mg,


mDex = 50 mg and mcat = 3.78 mg except for experiments labeled with where mcat = 1.54 mg.

results of the kinetic experiments are summarized in Table1. Figure 9 shows the rate profiles for different hydrogen concentrations. The presence of hydrogen in the reactor significantly increases the catalyst activity. In absence of hydrogen, the catalyst shows low activity, which increases strongly with increasing hydrogen concentration at molar fractions, X, below 0.001 corresponding to 150 mg hydrogen injection in our experiments, see Table 1 and Figure 10; however, above this value, the impact of hydrogen levels off, and the initial reaction rate reaches asymptotically a maximum value of about 155 kg/gcat.hr. The plateau effect at high hydrogen concentrations has been confirmed by several authors. Guastalla and Giannini (1983) proposed the presence of adsorption phenomena of hydrogen on the surface of the solid catalyst with a saturation at high concentrations. Rishina et al. (1994) suggested that a limiting maximum number of active sites is realized in the catalytic system in the presence

Figure 11 shows that the deactivation constant increases with hydrogen concentration. The shape of these curves is similar to the shape of the Rpo curves shown in Figure 10. The values for kd are plotted, Figure 13, as a function of Rpo for all experiments in Table 1. Remarkably, the figure shows that the runs that differ in process conditions show the same correlation between Rpo and kd. This way the deactivation can be interpreted as an activitydependent probability. Additionally one can conclude that an active site can not deactivate during its dormant state. The low deactivation at zero hydrogen is obviously caused by the high probability of being dormant, because of low reactivation rates by monomer transfer and propagation reactions. This supports the idea discussed previously by Pater et al. (2002) and Roos et al. (1997) that the deactivation of the catalyst relates to the catalyst activity independent of the reason for the activity change (temperature, pressure, hydrogen concentration etc.).

2.5 2 kd , 1/hr 1.5

7.5 7 PDI
0 200 400 600 800 1000 1200 Hydrogen added, mg

1 0.5 0

6.5 6 5.5 0 0.001 0.002 0.003 X 0.004 0.005 0.006

Fig. 11 The influence of H2 on the deactivation constant

Fig. 13 Effect of hydrogen concentration on the PDI of polypropylene.


2.5 2 kd , 1/hr 1.5 1 0.5 0
0 50 100 150 200

3.6 Modeling The modeling of reaction kinetics and molecular weight is based on the dormant site mechanism (Weickert, 2002) using an averaged-site approach based on the following mechanism: Chain propagation:
C j + M k C j+1
p

Rpo , kg/gcat . hr

Fig. 12 Deactivation constant versus polymerization rate, all runs, see Table 1.

initial

(3)

3.5 Molecular weight distribution The effect of hydrogen concentration on molecular weight and polydispersity index (PDI) is summarized in Table 2. Similar to other Tibased polymerization catalysts (Keii et al., 1984; Meier et al., 2001), the molecular weight decreases with increasing hydrogen concentration. This is attributed to the fact that hydrogen reacts with the growing chains, while it is in the dormant state, and terminates its propagation. Figure 13 shows the PDI of polypropylene at different hydrogen concentrations and reaction temperatures. It is clear that increasing levels of hydrogen result in a somewhat larger polydispersity index, PDI is about 6.3-7.3, because of the activation of site types that are dormant in the absence of hydrogen.
Table 2 Propylene polymerization: molecular weight distribution results.

Chain transfer to hydrogen:


C j + H 2 k C o + D j
h

(4)

Chain transfer to monomer:

C j + M k C1 + D j
m

(5)

Dormant site formation:


k C j + M S j+1
dm

(6)

Dormant sites reactivation by H2:


k S j + H 2 C o + D j
dm

(7)

Reactivation of dormant sites by monomer:


k S j + M C j+1
rem

(8)

Deactivation:
C j k D j
d

Run no.

M+ w
kg/mol

PDI

(9)

1 2 3 4 5 6 7
+

3878 1119 833 361 294 244 132

6.35 6.58 6.77 7.32 -

Note that the catalyst is activated before injecting it to the reactor; thus catalyst activation reaction is not considered in this mechanism. Using the long chain hypothesis, the polymerization rate can be described as:
R p = k p C Cm

(10)

Values in italics are calculated from intrinsic viscosity.

here kp is the propagation reaction constant. Because of active sites heterogeneity, this is an average value. C is the active catalyst site concentration. Cm is monomer concentration near the active sites, but a simplified model using the density of liquid propylene directly can be used, see for example (Banat et al., 2005). By developing mass balances over the concentrations of dormant and active sites and considering the quasi-steady state assumption, it can be proved that the polymerization rate can be defined as:
k PC m C max ( k 1 X + 1) 1 + k1 X + k 2 where Rp =
k1 = k reh ; k rem k2 = kS k rem

Table 3 summarizes the results of the non-linear estimation of kp, k1 and k2.
Table 3 Propylene polymerization kinetic constants at 70 C

kp (m /gcat. hr) 0.334

k1 15000

k2 8.02

(11)

(12)

Similarly, it can be proved that the following relationship holds:


C j = p C j1

Figure 14 shows a comparison between the fitted model (Equation 11) and the experimental data. The figure shows that the model describes the data well. Figure 12, as discussed earlier, shows an almost linear relationship between kd and Rpo. This suggests that the relationship between these variables can be described with a twoterm function. The first term describes the linear behavior, meanwhile, the second term compensates for the deviation from the linearity. The following empirical relation gives the best fit, see Figure 15:

(13)

k d = 8.38 103 R po + 0.187 (1 + 288 X)

(17)

with Cj is the concentration of active sites of chain length j. p is chain propagation probability that is defined as:
kS k p X k reh k rem + 1 p= k k k 1+ S + h X + m kp kp kp 1+

Finally, Equation (16) fits well with our experimental data, Figure 18. With the optimum parameter values, this equation can be written as:
15 103 X q = 2.08 105 + 0.0448 X + 1.28 104 1 + 15 103 X

(14)

(18)
200 R po, kg/gcat . hr 150 100 50 0 0 0.0025 0.005 X, [-] 0.0075 0.01
Rpo, exp Rpo, model

The termination (or chain transfer) probability (q) is defined by:


q = 1 p = km kh k k X + X + S 1 kp kp k p 1 + k1 X

(15)

with k1 defined by Equation (12). By considering the quasi-single-site approach, the relationship between the weight average molecular weight (Mw) and the average termination probability (q) is defined as:
q= 2 MM MW

Fig. 14 Comparison between the experimental and modeled Rpo.

2 .5

(16)
kd , 1/h r

2 1 .5 1 0 .5

3.7 Parameter estimation Equation (11) has three parameters, namely kp, k1 and k2. It should be emphasized that the true number of active sites and the true (in-situ) sorption is not taken into account and will influence the kp value.

E xp m o d el

0 0 50 100 R p o , k g /g c a t . h r 150 200

Fig. 15 Comparison of model and experimental: kd.

1.2E-3 1.0E-3 8.0E-4 6.0E-4 4.0E-4 2.0E-4 0.0E+0 0 0.005 0.01 X 0.015 0.02 q
Exp model_Dormant site

Cmax CS Cm CH2 Cj E Dj fH2 kd kh km kp

concentration Concentration of active sites Concentration of dormant sites Concentration of monomer Hydrogen concentration Growing polymer chain containing j monomers Activation energy Dead (terminated) polymer chain containing j monomers Fraction of potentially active sites Deactivation rate constant Chain transfer to hydrogen rate constant Chain transfer to monomer rate constant Propagation constant Propagation constant Dormant site reactivation by hydrogen rate constant Dormant site reactivation by monomer rate constant Number average molecular weight Viscosity average molecular weight Weight average molecular weight Mass of catalyst injected Mass of injected external donor Mass of injected TEA Mass of injected H2 Propagation probability Termination probability Polymerization rate Polymerization rate Polymerization temperature Time Reaction time Hydrogen molar ratio Yield Monomer density

mol/l mol/l mol/l mol/l J/mol 1/hr l/mol. s l/mol. s l/mol. s m3/gcat. hr l/mol. s l/mol. s g/mol g/mol g/mol mg mg mg mg mol/l. s kg/gcat hr C hr min gPP 3 kg/m

Fig. 16 Comparison of model and experimental q.

4. CONCLUSIONS In this work, liquid propylene polymerization has been studied in a fully filled batch reactor under quasi-isothermal conditions. The polymerization reaction is performed using a prepolymerized Ziegler-Natta catalyst with TEA as cocatalyst at a temperature of 70 C and a wide range of hydrogen concentrations (0-100 mmol/l). Polymerization rate profiles and molecular weight distributions have been measured and modeled. The following results are obtained from the experiments: 1. For MgCl2 supported catalysts, it is widely accepted that the propylene polymerization rate decreases with reaction temperature if it exceeds 70 C. In contrary, when the reaction is carried out in a filled reactor, the polymerization rate increases with increasing reaction temperature even if it reaches 80 C. 2. The polymerization rate increases with increasing hydrogen reaching a plateau at high hydrogen concentrations (> 150 mg). 3. The pseudo-first-order deactivation constant is linearly dependent on the initial polymerization rate. 4. The polymer molecular weight decreases with hydrogen concentration, but increases with polymerization temperature. Meanwhile, the polydispersity index is effectively constant with an average value of 6.5. 5. Both polymerization rate and the molecular weight can be modeled using the dormant sites theory.

R ,o p
kreh krem Mn Mv Mw mcat mDex mH2 mTEA p q Rp

k p
T t tr X Y m

7. REFERENCES

5. ACKNOWLEDGMENT We gratefully acknowledge the Dutch Polymer Institute fir supporting this work under the project number DPI # 114. We further wish to thank Basell for the analytical support.

6. NOMENCLATURE
C Activated catalyst mol/l

Al-haj Ali, M., B. Betlem, B. Roffel and G. Weickert (2005). "Liquid Propylene Polymerization with a Ziegler-Natta Catalyst: The Influence of Hydrogen and Temperature." Submitted to AIChE. Al-haj Ali, M., B. Betlem, B. Roffel and G. Weickert (2005). "Propylene Polymerization in A Full-Filled Batch Reactor." In Preparation. Banat, Y., U. P. Veera and G. Weickert (2005). "The Role of Equilibrium Sorption in Gas and Liquid Phase Propylene Polymerization." To be submitted to Chemical Engineering Science. Chadwick, J. C., G. Morini, G. Balbontin, V. Busico, G. Talarico and O. Sudmeijer (2000). "Advances in Propene Polymerization using
8

Magnesium Chloride-Supported Catalysts." ACS SYMPOSIUM SERIES 749: Olefin Polymerization Emerging Frontiers: 50-65. Galli, P. and G. Vecellio (2004). "Polyolefins: The Most Promising Large-Volume Materials for the 21st Century." Journal of Polymer Science: Part A: Polymer Chemistry 42: 396-415. Guastalla, G. and U. Gianinni (1983). "The Influence of Hydrogen on the Polymerization of Propylene and Ethylene with an MgCl2 Supported Catalyst." Macromolecular Chemie-Rapid Communications 4: 519-527. Han-Adebekun, G. C., M. Hamba and W. H. Ray (1997). "Kinetic Study of Gas Phase Olefin Polymerization with a TiCl4/MgCl2 Catalyst I. Effect of Polymerization Conditions." Journal of Polymer Science: Part A: Polymer Chemistry 35: 2063-2074. Hindryckx, F., P. Dubois, R. Jrme and M. G. Marti (1998). "Ethylene Polymerization by a High Activity MgCl2 Supported Ti Catalyst in the Presence of Hydrogen and/or 1-Octene." Polymer 39(3): 621-629. Keii, T., Y. Doi, E. Suzuki, M. Tamura, M. Murata and K. Soga (1984). "Propene Polymerization with a Magnesium ChlorideSupported Ziegler- Catalyst, 2: Molecular Weight Distribution." Macromolecular Chemistry and Physics 185: 1537-1557. Kissin, Y. V., R. Ohnishi and T. Konakazawa (2004). "Propylene Polymerization with TitaniumBased Ziegler-Natta Catalysts: Effects of Temperature and Modifiers on Molecular Weight, Molecular Weight Distribution and Stereospecifity." Macromolecular Chemistry and Physics 205(3): 284-301. Meier, G. B., G. Weickert and W. P. M. Van Swaaij (2001). "Gas-Phase Polymerization of Propylene: Reaction Kinetics and Molecular Weight Distribution." Journal of Polymer Science: Part A: Polymer Chemistry 39(4): 500-513. Natta, G., G. Mazzanti, P. Longi and F. Bernardini (1959). "Influenza Dell'idrogeno Sulla Polimerizzazione Anionica Coordinata del Propilene e Dell' Etilene." La Chimica E L'Industria 41: 519-526. Pater, J. T. M., G. Weickert and V. S. W.P.M. (2002). "Polymerization of Liquid Propylene with A Fourth Generation Ziegler-Natta Catalyst: Influence of Temperature, Hydrogen, Monomer Concentration, and Prepolymerization Method on Polymerization Kinetics." Chemical Engineering Science 57(16): 3461-3477. Rishina, L. A., E. I. Vizen, L. N. Sosnovskaja and F. S. Dyachkovsky (1994). "Study of the Effect of Hydrogen in Propylene Polymerization with MgCl2-Supported Ziegler-Natta Catalyst-Part 1. Kinetics of

Polymerization." European Polymer Journal 30(11): 1309-1313. Roos, P., G. B. Meier, J. J. C. Samson and G. Weickert (1997). "Gas Phase polymerization of Ethylene with a SilicaSupported metallocene Catalyst: Influence of Temperature on Deactivation." Macromolecular Chemie-Rapid Communications 18: 319-324. Samson, J. J. C., P. J. Bosman, G. Weickert and K. R. Westerterp (1999). "Liquid-Phase Polymerization of Propylene with a Highly Active Ziegler-Natta Catalyst. Influence of Hydrogen, Cocatalyst, and Electron Donor on Reaction Kinetics." Journal of Polymer Science: Part A: Polymer Chemistry 37: 219-232. Samson, J. J. C., G. Weickert, A. Heerze and K. R. Westerterp (1998). "Liquid-Phase Polymerization of Propylene with a Highly Active Catalyst." AIChE Journal 44: 14241437. Weickert, G. (2002). Modeling of Polymerization Rate and Molecular Weight: H2 Response. EC Project Report GRD2-2000-30189 "Polyprop", Milan, Italy. Weickert, G. (2003). High Precision Polymerization Rate profiles. 3rd International Workshop on Heterogeneous Ziegler-Natta Catalysts, Japan.

You might also like