You are on page 1of 111

CHAPTER – 4

RESULTS AND DISCUSSION

In this chapter, the simulation results obtained for the ethylene and propylene polymerization

using different metallocene catalyst systems are presented in sections 4.1 and 4.2 respectively.

Sections 4.1.1 and 4.1.2 detail the results obtained for ethylene polymerization in gas phase

with silica supported and in solution phase with in-situ-supported zirconocene catalysts

respectively. Results of solution phase propylene homopolymerization with various

metallocene catalysts are presented and discussed in Sections 4.2.1 through 4.2.7. Models

developed in Chapter 3 are simulated for validation and study of effects of reaction

parameters with data reported in open literature. Experimental conditions and corresponding

references are briefed in Table 4.1 and Table 4.2. All polymerizations in referenced works

were carried out in a stirred, semi-batch autoclave reactor with a flowmeter unit.

In all ethylene and propylene polymerization studies carried out in this work, model

equations are first solved analytically. To obtain analytical solution, the reactions of chain

propagation, spontaneous deactivation and chain transfer to monomer are considered. It was

possible to solve the model equations analytically since the concentration of monomer was

held constant (by maintaining its pressure). Differential equations describing moments of

living and dead polymer chain length distribution were analytically solved yielding Equation

4.1 through Equation 4.6.

l0  P(0). exp  kd t  (4.1)

1l 
k p  ktM P(0)
exp  kd t   exp  k d  ktM M t
ktM (4.2)

98
k  k tM P(0)
2k
l2   k tM exp  k d t   exp  k d  k tM M t
p
2 p
k tM
2k p  k tM P(0)k p M (4.3)
 t. exp  k d  k tM M t
k tM

0 
k d  ktM M P(0) 1  exp  k t 
kd
d (4.4)

 exp  k d  ktM M t 


k  k k  k M P(0)   k d  ktM M 
1  p tM d tM
 

ktM  exp  k d t  1 1 (4.5)
  
k d  ktM M  k d 
 kd

 exp  k d  ktM M t exp  k d t 



2 
k p  ktM k d  ktM M P(0)2k p  ktM  
 
 k d  ktM M  kd 

2
ktM  1 1 
 k  k M   k 
 d tM d  (4.6)
 exp  k d  ktM M t 
2k p  ktM P(0)k p M 
1



k d  ktM M  k d  ktM M 
ktM
 t. exp  k d  ktr M t 

Analytical solution described above is only possible for semibatch reactor with

invariable monomer concentration throughout the polymerization. For the parameters

estimation in all the subsequent studies in this work, this approach is utilized along with

NLDE to determine the coarse values of kinetic parameters in order to judge the range of

parameters those were required in fine optimization.

99
Table 4.1 Sections discussing results of ethylene polymerization

Sec. Catalyst system Reaction conditions Solvent Reference


4.1.1 E1: Silica supported (Me2Si[Ind]2ZrCl2) / MAO P =5 bar; Catalyst: 0.2 g; Al/Zr = 386; - Roos et al. (1997)
T = 40 °C, 50 °C, 60 °C and 70 °C
4.1.2 E2: In-situ-supported (Et[Ind]2ZrCl2) / MAO P = 80 psig; Catalyst: 6 μmol; Al/Zr = 500; Hexane Chu et al. (2000)
T = 40 °C, 60 °C, 80°C, 100°C, 120 °C.
* For the sake of brevity, above two catalysts shall be denoted as E1 and E2 respectively from here onwards.

Table 4.2 Sections discussing results of propylene polymerization

Sec. Catalyst Reaction conditions Solvent Reference


4.2.1 P1: (Me2Si[Ind]2ZrCl2) / MAO P = 2 bar; [Zr] = 10 μmol/L; T = 25 °C and 75 °C; Toluene Marques et al.
Al/Zr 500 and 2000 (2002)
4.2.2 P2: (Et(Ind)2ZrCl2) / MAO - Do - - Do -
4.2.3 P3: (Me2Si(Ind)2HfCl2) / MAO P = 2 bar; [Hf] = 10 μmol/L; T = 40 °C and 80 °C; - Do -
Al/Hf 500 and 2000
4.2.4 P4: (Et(Ind)2HfCl2) / MAO - Do - - Do -
4.2.5 P5: ([2,4,6-Me3Ind]2ZrCl2) / MAO P = 0.98 atm; [Zr] = 20 μmol/L; T = 0 °C; - Do - Yasin et al.
Al/Zr = 2000 and 4000 (2004)
4.2.6 P6: ([2,4,7-Me3Ind]2ZrCl2) / MAO P = 0.98 atm; [Zr] = 20 μmol/L; T = 0 °C; - Do -
Al/Zr = 1000, 2000 and 4000
4.2.7 P7: (Me2Si[2,4,6-Me3Ind]2ZrCl2) / MAO P = 0.98 atm; [Zr] = 20 μmol/L; Al/Zr = 2000; - Do - Yasin et al.
T = 30 °C, 50 °C and 70 °C (2005)
* For the sake of brevity, above seven catalysts shall be denoted as P1, P2, P3, P4, P6 and P7 respectively from here onwards.

100
4.1 Ethylene polymerization

Ethylene polymerization model is applied to gas phase polymerization with silica supported,

bridged Me2Si[Ind]2ZrCl2 catalyst and to solution phase polymerization with in-situ-silica

supported, bridged Et[Ind]2ZrCl2 catalyst. Based on experimental conditions, suitable

assumptions are made and a truncated form of comprehensive model is applied to both the

cases. The set of reactions considered are described in corresponding sections ahead.

4.1.1 Ethylene polymerization with Me2Si[Ind]2ZrCl2 (E1)/MAO

Ethylene polymerization model (EPM), discussed in Section 3.2.2, is applied to gas phase

polymerization of ethylene with silica- supported Me2Si[Ind]2ZrCl2 catalyst and kinetic

parameters are obtained. Experimental data for model validation are taken from Roos et al.

(1997). Simulations are carried out (with ODE-15s function provided in MATLAB™ 7.0

software) using natural logarithmic differential evolution approach of optimization to estimate

the kinetic parameters.

Model description

The gas phase production of polyethylene by silica supported Me2Si[Ind]2ZrCl2 catalyst is a

multistep process that necessarily includes initiation, propagation, and termination. Following

assumptions are made while employing the comprehensive ethylene polymerization model:

Assumptions

(i) Instantaneous formation of active sites (reaction between catalyst and cocatalyst).

(ii) First-order propagation with respect to monomer and the active site is assumed, and

the reactivity of the complex species does not depend on the length of the polymeric

chain.

101
(iii) Chain transfer takes place to monomer only and follows first order kinetics. This

assumption is induced, as the reactor was flushed with ethylene several times after

scavenging and catalyst introduction. This precludes the possibility of chain transfer to

anything but ethylene.

(iv) The same type of site generates after transfer as is primitively formed by activation of

catalyst with cocatalyst. So chain transfer step does not change the number of active

sites.

(v) Monomer consumption in initiation and chain transfer reactions is negligible as

compared to the propagation reaction (long chain assumption).

(vi) First-order deactivation of active sites.

Ethylene polymerization reactions considered based on above assumptions are

described in Table 4.3.

Table 4.3 Reactions Considered in Ethylene Polymerization

Reaction Stoichiometry Description

1. Cat  Cocat 
 P(0) Instantaneous catalyst activation

2. P0  M 
kin
P1 Chain initiation

3. Pi   M  Pi  1 Chain propagation


k
p

4. Pi  
kd
Pd 0  Di  {chain} Spontaneous catalyst

deactivation
5. P0 
kd
Pd 0 {site}

6. Pi   M k
tM
D  (i)  P1 Chain transfer to monomer

102
Estimated parameters and effect of temperature

Figure 4.1 shows the polymerization rates predicted by the model proposed by Roos et al.

(Section 2.4; Page 51) and ethylene polymerization model (Section 3.2.2) developed in this

study. Roos et al. modeled the polymerization rate assuming that deactivation of the catalyst

increases with increasing polymerization rate under isothermal conditions. Since they did not

consider initiation or any chain transfer reaction, the rate profiles are obtained exponentially

decreasing from a maximum value. For the same reason, polymerization rates are under

predicted at low temperatures (50 °C and 60 °C) and over predicted at 70 °C.

On the other hand, polymerization rates predicted by EPM exhibit a good agreement

with experimental data, at all the temperatures (40 °C, 50 °C, 60 °C and 70 °C). Estimated

kinetic parameters and objective function values F(k) are shown in Table 4.4. Close range of

objective function values (from 0.29934 to 0.6735) obtained, shows good fit with

experimental observations. Rates of initiation and propagation are increasing with increase in

temperature, as inferred from the estimated rate constants for these reactions. At the beginning

of polymerization, also referred to as induction period, the monomer is mainly consumed in

initiation followed by propagation reaction. During this period, the slope of polymerization

rate vs. time curve is indicative of the rate of initiation. In Figure 4.1, both, the initial

experimental polymerization rates and corresponding EPM predictions, clearly evince that

initiation and propagation rates are increasing with increase in temperature. Frequency of

spontaneous deactivation and chain transfer to monomer are also increasing with increase in

temperature, moreover the latter termination mechanism dominates over the former.

Increasing rates of deactivation and transfer to monomer at 50 °C, 60 °C and 70 °C are

responsible for a steeper decay in polymerization rate after reaching a maximum value.

103
0
-2
40 C
1.6x10 0
50 C
Polymerization rate (moles/L/s)

0
60 C
0
70 C

-3
8.0x10

0.0
0 100 200
Time (minutes)

Figure 4.1 Ethylene polymerization rate vs. time with Me2Si[Ind]2ZrCl2 (E1)/MAO;

Symbols: Exp data; Lines: Model Prediction ‐‐‐ (Roos et al.), — (EPM; Section 3.2.2).

[Catalyst (E1) = 0.2 g; P = 5 bar; Al/Zr = 386]

104
Model proposed by Roos et al. is not capable to predict average molecular weights and PDIs.

Molecular weight distribution obtained with EPM is shown in Table 4.5. Average molecular

weights are decreasing with increase in temperature. For an increase of 10 °C, from 40 °C to

70 °C, the weight average molecular weight ( M w ) is found to decrease by 67.40 %, 70.47 %

and 81.67 % respectively. This inverse trend is normal in chain growth polymerization and is

usually ascribed to the high activation energies for chain transfer reactions as compared to the

propagation reaction (Rudin, 1999). With increase in temperature, the rate of chain

termination to monomer increases rapidly as compared to the rate of propagation, which

results in decreasing molecular weights. Polymers prepared with metallocene catalysts

possess narrow molecular weight distribution, with polydispersities ranging from 2.0 to 2.5,

which is a consequence of the single-site feature of the metallocenes (Stevens; 1999).

Polydispersity indices of polyethylene prepared with Me2Si[Ind]2ZrCl2/MAO catalyst system

are found to be 1.999 irrespective of temperature.

Simulations are carried out further, to study the effects of ethylene pressure and

catalyst amount on polymerization rate and average molecular weights.

Effect of ethylene pressure

Polymerization rate is linearly increasing with ethylene pressure at all the temperatures as

shown in Figures 4.2. Low pressures (1-3 bar), i.e., low concentrations of monomer, reasons

low rates that are steady and maintained (due to negligible transfer reaction). At higher

pressures (5-7 bar), higher polymerization rates are obtained, but at the same time, transfer to

monomer also increases with high monomer concentrations resulting in steeper decay in

polymerization rates.

105
Table 4.4 Estimated Parameters for Me2Si[Ind]2ZrCl2 (E1)/MAO

T (°C) 40 50 60 70

k in ×10-5 (M-1.s-1) 8.4474 26.975 35.832 44.951

k p ×10-3 (M-1.s-1) 4.3141 7.2207 42.568 195.420

k d ×10-3 (s-1) 42.032 45.265 54.526 94.095

k tM (M-1.s-1) 4.9690 5.5374 6.4339 7.0468

F(k) (-) 0.29934 0.56928 0.67350 0.45939

Table 4.5 Predicted Molecular Weights & PDI with Me2Si[Ind]2ZrCl2 (E1)/MAO

T (°C) M n (g/mol) M w (g/mol) PDI

40 9.628938 × 105 1.925759 × 106 1.999970

50 3.139822 × 105 6.276838 × 105 1.999106

60 9.268304 × 104 1.853654 × 105 1.999990

70 1.698616 × 104 3.397203 × 104 1.999982

106
-2
1.5x10
Polymerization rate (moles/L/s)

-2
1.2x10

-3
9.0x10

6.0x10
-3 7 bar

5 bar
-3
3.0x10
3 bar

0.0 1 bar
0 50 100 150 200
Time (minutes)

(a)

-2
2.0x10
Polymerization rate (moles/L/s)

-2
1.6x10

-2
1.2x10

-3
8.0x10

7 bar

-3
4.0x10 5 bar

3 bar

0.0
1 bar
0 50 100 150 200
Time (minutes)

(b)

107
-2
2.4x10

Polymerization rate (moles/L/s)


-2
1.8x10

-2
1.2x10

-3
6.0x10
7 bar
5 bar
3 bar
0.0
1 bar
0 50 100 150 200
Time (minutes)

(c)

Figure 4.2 Polymerization rate vs. time: Effect of pressure.

[Catalyst (E1) = 0.2 g, Al/Zr = 386, T = (a) 50 °C, (b) 60 °C (c) 70 °C]

As shown in Figures 4.3(a), at 50 °C, average molecular weights are increasing

slightly and not affected much with a change in ethylene pressure. A similar trend is obtained

at 60 °C as shown in Figure 4.3(b). However, at 70 °C [Figure 4.3(c)], an increase in M w from

2.03457 × 104 to 3.0561 × 104 is obtained for a change in pressure from 1 bar to 3 bar. With

further increase in pressure till 7 bar, a slight increase in average molecular weights are

observed. These results come along to be consistent with the assumption that the chain

transfer to monomer reaction is bimolecular giving a unvarying M w with respect to ethylene

concentration (pressure).

108
Average molecular weight (g.mol )

5
6.0x10
-1

Mn
Mw
T = 50 C 0

5
4.0x10

5
2.0x10
0 2 4 6 8
Pressure (bar)

(a)

Mn
Mw
5
2.25x10
-1

T = 60 C
0
Average Molecular weight g.mol

5
1.50x10

4
7.50x10

0.00
1 2 3 4 5 6 7
Pressure (bar)

(b)

109
Mn
Mw

Average Molecular weight (g.mol )


4
4.0x10

-1
T = 70 C
0

4
2.0x10

0.0
1 2 3 4 5 6 7
Pressure (bar)

(c)

Figure 4.3 Effect of pressure on average molecular weights.

[Catalyst (E1) = 0.2 g, Al/Zr = 386, T = (a) 50 °C, (b) 60 °C (c) 70 °C]

Effect of catalyst Amount

A steady increase in polymerization rate with increase in catalyst amount at constant

temperature and pressure is observed as shown in Figures 4.4. At higher temperatures, the rate

reaches to a maximum within no time and then decreases steeply, suggesting that both, the

initiation and the termination rates increase staggeringly with catalyst amount. With variation

in initial catalyst amount, average molecular weights and PDIs are not affected and the values

are upheld to those given in Table 4.5. In literature, for metallocene catalysts average

molecular weights are reported to be decreasing with increase in catalyst concentration

[Estrada and Hamielec (1994); Rieger and Janiak (1994); Kaminsky (1996); Zohuri et al.

(2005); Cheng and Tang (2010)]. This phenomenon is commonly explicated by the fact that

there is an increase in rate of chain transfer reactions which are dependent on the active site

concentration i.e. chain transfer to monomer, chain transfer to cocatalyst and β-hydride
110
elimination. The model applied to the present case is based on the assumption that the chain

transfer takes place to monomer only. Increase in catalyst amount seems to affect both, the

propagation rate and chain transfer rate equivalently at a given temperature and ethylene

pressure, due to which an invariant MWD is obtained.

-2
2.0x10
Polymerization rate (moles/L/s)

-2
1.6x10

-2
1.2x10 Zr
0.5 g

8.0x10
-3
0.4 g

0.3 g
-3
4.0x10 0.2 g

0.1 g
0.0
0 50 100 150 200
Time (minutes)

(a)

-2
2.5x10

-2
Polymerization rate (moles/L/s)

2.0x10

-2
1.5x10
Zr
-2
0.5 g
1.0x10
0.4 g
0.3 g
-3
5.0x10
0.2 g
0.1 g
0.0
0 50 100 150 200
Time (minutes)

(b)

111
-2
3.6x10

-2
3.0x10

Polymerization rate (moles/L/s)


-2
2.4x10

-2
1.8x10
Zr
-2
1.2x10 0.5 g
0.4 g
-3
0.3 g
6.0x10
0.2 g
0.1 g
0.0
0 50 100 150 200
Time (minutes)

(c)
Figure 4.4 Polymerization rate vs. time: Effect of catalyst amount.

[P = 5 bar, Al/Zr = 386, T = (a) 50 °C, (b) 60 °C (c) 70 °C]

4.1.2 Ethylene polymerization with in-situ-supported Et[Ind]2ZrCl2 (E2)/MAO

Ethylene polymerization model is applied to solution phase polymerization of ethylene with

in-situ-silica supported Et[Ind]2ZrCl2 catalyst and kinetic parameters are obtained. Data for

the model validation are taken from Chu et al. (2000).

Model description

Some rudimentary assumptions made while employing the model are described hereunder:

Assumptions

(i) Constant and uniform temperature and ethylene concentration inside the reactor.

(ii) Instantaneous vapor-liquid equilibrium. In solution polymerization, the concentration

of monomer in the liquid phase is dependent upon the solubility of the gas in the

solvent. Atiqullah et al. (1998), compared various equations of state for predicting
112
ethylene solubility in toluene under similar reaction conditions and observed that

Peng-Robinson (PR) and Soave–Redlich–Kwong (SRK) equations of state are equally

good for the purpose. Here, Peng-Robinson equation of state (PR-EOS) is used to

determine the concentration of ethylene in solvent at different temperatures and

monomer pressures.

(iii) Negligible polymer volume as compared to the total volume of the reactor, i.e.

constant gas-phase volume in the reactor.

(iv) Instantaneous activation of catalyst sites.

(v) First-order propagation with respect to monomer and the active site, and the reactivity

of the activated complex species does not devolve on the length of the polymer chain.

(vi) First-order deactivation of active sites. Spontaneous catalyst deactivation in a

propagating chain produces dead chain without terminal double bond.

(vii) Chain transfer to monomer, catalyst and cocatalyst produce dead chains with terminal

double bond.

(viii) Chain transfer to a dead polymer with terminal double bond produces long chain

branching.

Ethylene polymerization reactions considered based on above assumptions are

described in Table 4.6.

Estimated parameters and effect of temperature

Model is simulated in order to estimate the kinetic parameters at different temperatures using

ethylene flow rate as a measure of polymerization rate. A very large population size (120

times the dimension) is used to make certain of receiving optimized estimates of parameters.

Table 4.7 summarizes the parameters estimates at 40 °C, 60 °C, 80 °C, 100 °C and 120° C at

fixed catalyst amount 6μmols, Al/Zr = 500 and ethylene pressure 80 psig. Figure 4.5 shows

113
the model predictions of polymerization rate at different temperatures, which are very close to

the experimental values.

Table 4.6 Reactions Considered in Ethylene Polymerization

Reaction Stoichiometry Description

1. Cat  Cocat 
 P(0) Instantaneous catalyst activation

2. P0 M 
k in
P1 Chain initiation

3. Pi   M  Pi  1 Chain propagation


pk

4. Pi  
kd
Pd 0  Di  {chain} Spontaneous catalyst

deactivation
5. P0 
kd
Pd 0 {site}

6. Pi   M k
tM
D  (i)  P1 Chain transfer to monomer

7. Pi  Cocat k


tCo
 D (i)  P0 Chain transfer to cocatalyst

8. Pi   D  (i)  P0 β-hydride elimination


k

9. Pi   D ( j ) k
lcb
Pi  j  Long-chain branching

As defined in Equation 3.99 and reported in Table 4.7, the values of objective function

indicate that the model is capable of predicting kinetic behaviour for all the temperatures

expeditiously. Function values obtained are ranging closely, with a least value of 1.1186 at

80 °C representing the best fit to experimental observations as compared to highest value of

1.6063 at 120 °C. Significantly lower values of propagation rate constants at 40 °C and 60 °C

are obtained relating to very low polymerization rate and catalyst activity with respect to those

at higher temperatures. At 80 °C, high propagation rate as compared to those at lower

temperatures trace higher polymerization rate with highest activity of the catalyst.

Polymerization rates are increasing with increase in temperature from 40 °C to 120 °C.

114
Declining rate profiles at 100 °C and 120 °C can be explained by serious catalyst deactivation,

transfers to monomer and β-hydride elimination at higher temperatures as pointed out by the

values of these parameters obtained.

Figure 4.6 shows the predicted decrease in concentration of active catalyst sites with

time at different temperatures. At 40 °C, 60 °C and 80 °C, all the active catalyst sites are

occupied within initial 15 minutes, whereas at higher temperatures (100 °C and 120 °C)

certain fraction of those could not attach a monomer to initiate the chain.

Experimental kinetic data at different operating conditions, like different catalyst

amount, ethylene pressure and cocatalyst to catalyst mole ratio, are utilized to validate the

model at fixed temperature of 60 °C. Parameters estimated at 60 °C, 80 psig and 6 μmol

catalyst amount with Al/Zr = 500 are used to verify the model responses at various conditions.

Table 4.7 Estimated Parameters for Et[Ind]2ZrCl2 (E2)/MAO

T k in ×103 kp k d ×105 k tM ×103 k tCo ×106 k  ×106 k lcb ×105 F(k)


(°C) (-)
(M-1.s-1) (M-1.s-1) (s-1) (M-1.s-1) (M-1.s-1) (s-1) (M-1.s-1)

40 7.9660 17.1863 1.5377 3.3232 78.1863 1.3426 1.6298 1.3832

60 12.2164 44.9804 2.4541 19.5830 43.1982 7.9277 5.6526 1.1674

80 14.8433 188.9502 4.2504 35.4385 4.1940 8.6007 14.1722 1.1186

100 21.4832 529.4216 8.5169 56.6386 1.6294 1107.979 16.3020 1.3586

120 63.7560 938.188 74.0553 86.9489 1.4610 1307.70 30.3491 1.6063

115
0
-3
40 C
3.0x10 0
60 C

Polymerization rate (moles/L/s)


0
T 80 C
0
100 C
0
120 C
-3
2.0x10

-3
1.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.5 Effect of temperature on ethylene polymerization rate with in-situ-supported


Et[Ind]2ZrCl2 (E2)/MAO; solid lines are model predictions.
[Catalyst (E2) = 6 μmol, Al/Zr = 500 and P = 80 psig]

0
40 C
0
60 C
Active catalyst sites (mol)

0
80 C
4 0
100 C
0
120 C

0 10 20 30 40 50 60
Time (minutes)

Figure 4.6 Active catalyst sites vs. reaction time.


[Catalyst (E2) = 6 μmol, Al/Zr = 500 and P = 80 psig]

116
Effect of ethylene pressure

The model also predicted the proportional changes in polymerization rate at different ethylene

pressures as shown in Figure 4.7. The trend vindicates the first order dependence of rate on

ethylene concentration as no declining regions are seen. Figure 4.8 shows that, at 40 psig

pressure active catalyst sites stayed available for all the polymerization time suggesting the

non-initiation of some active site. At 80 psig pressure, active site were occupied within 10

minutes and for higher pressures within 20 minutes.

Effect of catalyst amount

Figure 4.9 shows a good match between experimental observations and model predictions at

3, 12 and 18 μmol of initial catalyst amount taken. The model adequately captures the features

of polymerization with in-situ-supported metallocene catalyst by following the sustained

polymerization rate with time and increase in the same with the increase in catalyst amount.

Figure 4.10 portrays that all the active catalyst sites were occupied within 10 minutes,

irrespective of initial amount of catalyst used.

Effect of cocatalyst to catalyst mole ratio

Effect of cocatalyst to catalyst mole ratio on the model predictions of polymerization rate are

shown in Figure 4.11. Use of high Al/Zr ratio brings in higher polymerization rate and for the

entire range of ratios, the maximum rate was reached within 10 minutes. Figure 4.12 depicts

that for lower ratios (250 and 500), active catalyst sites disappeared within first 10 minutes,

whereas for higher ratios (above 500) these decreased with time but remained available for

entire polymerization time. This is possible for high amount of cocatalyst delays the

spontaneous catalyst deactivation and also facilitates the regeneration of deactivated catalyst.

117
-3
3.0x10
40 psig
80 psig
120 psig

Polymerizatiom rate (moles/L/s)


-3
2.4x10
160 psig

-3
1.8x10

-3
1.2x10

-4
6.0x10

0.0
0 10 20 30 40 50 60

Time (minutes)

Figure 4.7 Effect of pressure on ethylene polymerization rate; solid lines are model
predictions.
[Catalyst (E2) = 6 μmol, Al/Zr = 500 and 60 °C]

7.5

6.0
40 psig
Activared catalyst sites (mol)

80 psig
120 psig
4.5 160 psig

3.0

1.5

0.0

0 10 20 30 40 50 60
Time (minutes)

Figure 4.8 Active catalyst sites vs. reaction time.


[Catalyst (E2) = 6 μmol, Al/Zr = 500 and 60 °C]

118
-3
2.0x10
Initial catalyst amount
18 mol
-3

Polymerization rate (moles/L/s)


1.6x10

-3
1.2x10
12 mol

-4
8.0x10

6 mol
-4
4.0x10

3 mol
0.0

0 10 20 30 40 50 60
Time (minutes)

Figure 4.9 Effect of catalyst amount on ethylene polymerization rate; solid lines are
model predictions.
[Al/Zr = 500, 60 °C, and 80 psig]

20

15
Active catalysts sites (mol)

Initial catalyst amount


3 mol
6 mol
10 12 mol
18 mol

0 10 20 30 40 50 60
Time (minutes)

Figure 4.10 Active catalyst sites vs. reaction time.


[Al/Zr = 500, 60 °C and 80 psig]

119
-3
3.0x10

Al / Zr
2.5x10
-3 250

Polymerization rate (moles/L/s)


500
1000
2.0x10
-3 2000
4000

-3
1.5x10

-3
1.0x10

-4
5.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.11 Effect of Al/Zr mole ratio on ethylene polymerization rate; solid lines are
model predictions.
[Catalyst (E2) = 6 μmol, T = 60 °C and P = 80 psig]

6.0
Active catalyst sites (mol)

4.5

3.0
Al / Zr
250
500
1.5 1000
2000
4000

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.12 Active catalyst sites vs. reaction time.


[Catalyst (E2) = 6 μmol, T = 60 °C and P = 80 psig]

120
Polyethylene properties

Table 4.8 presents model cyphered properties of polyethylene at different sets of operating

conditions. Following experimental observations, average molecular weights of polyethylene

obtained from the model are found to be decreasing with increase in temperature, whereas the

change in catalyst amount, cocatalyst to catalyst mole ratio or ethylene pressure brought

insignificant effect.

The calculated number average molecular weight values of polyethylene are ranging

from 35127.01 to 38166.63 for varation in catalyst amount, from 35127.01 to 39089.35, for

varation in cocatlyst to catalyst mole ratio and from 35127.01 to 37860.78 for varation in

ethylene pressure, which advises that change in catalyst amount, cocatalyst to catalyst mole

ratio or ethylene pressure have insignificant effect on molecular weight and its distribution of

the polymer produced by in-situ supported Et[Ind]2ZrCl2 catalyst.

For homogeneous and otherwise supported metallocene catalysts, average molecular

weights have been reported to be decreasing with increase in catalyst concentration or

cocatalyst to catalyst mole ratio by many researchers [Estrada and Hamielec (1994); Rieger

and Janiak (1994); Kaminsky (1996); Zohuri et al. (2005); Cheng and Tang (2010)]. This

phenomenon is usually explained by the fact that there is an increase in rate of chain transfer

reactions which are dependent on the active site concentration. It is worth noticing that, the

rate of propagation is also dependent on the active site concentration and when the active site

concentration is maintained during polymerization (as for in-situ supported catalyst),

propagation rate increases proportionally, resulting effectively no change in average

molecular weight.

121
Table 4.8 Polyethylene Properties with Et[Ind]2ZrCl2 (E2)/MAO

S. No. T Zr Al/Zr P Mn Mw Mw PDI PDI f (=) lcb/1000C


(°C) (μmoles) (-) (psig) (Exp) (Model) (Model) (Model)
(Model) (Exp) (Model)
1. 60 3 500 80 38166.63 74038.10 76256.93 2.50 1.998 0.870 4.633 × 10-9
2. 60 6 500 80 35127.01 74234.10 70211.57 2.59 1.998 0.986 7.238 × 10-10
3. 60 12 500 80 35127.08 67110.50 70211.94 2.22 1.998 0.986 1.447 × 10-9
4. 60 18 500 80 35135.82 70916.20 70230.15 2.26 1.998 0.986 2.168 × 10-9
5. 60 6 250 80 38143.88 78666.70 76211.49 2.60 1.998 0.986 5.428 × 10-10
6. 60 6 500 80 35127.01 74234.10 70211.57 2.59 1.998 0.986 7.238 × 10-10
7. 60 6 1000 80 39089.35 80000.00 78022.34 2.13 1.996 0.987 2.501 × 10-14
8. 60 6 2000 80 38969.59 72000.00 77783.32 2.40 1.996 0.864 2.244 × 10-15
9. 60 6 4000 80 36296.29 69333.30 72483.69 2.20 1.997 0.781 9.308 × 10-16
10. 40 6 500 80 56975.08 - 112696.71 - 1.978 0.909 1.439 × 10-7
11. 60 6 500 80 35140.93 74234.10 70211.57 2.59 1.998 0.986 7.238 × 10-10
12. 80 6 500 80 29273.46 - 53102.05 - 1.814 0.972 1.265 × 10-11
13. 100 6 500 80 19669.60 - 39299.87 - 1.998 0.985 3.034 × 10-12
14. 120 6 500 80 13965.66 - 27931.31 - 2.000 0.992 5.615 × 10-12
15. 60 6 500 40 36419.99 59165.70 69598.61 2.44 1.911 0.887 8.061 × 10-12
16. 60 6 500 80 35127.01 74234.10 70211.57 2.59 1.998 0.986 7.238 × 10-10
17. 60 6 500 120 37412.19 75418.70 72579.66 2.61 1.940 0.845 3.105 × 10-8
18. 60 6 500 160 37860.78 73472.40 75607.98 2.62 1.997 0.965 5.798 × 10-9

122
As monomer is also acting as a chain transfer agent, the insight of the effect of ethylene pressure

on molecular weight may be gained by probing Equation 4.7 for number average degree of

polymerization. Estimated parameters (Table 4.7) indicate that chain transfer to monomer is

dominating over other transfer reactions, for which the contribution of other transfer reactions

may be neglected in the denominator of Equation 4.7. For such a situation, degree of

polymerization and so the molecular weight show up to be independent of monomer

concentration.

k p M 
DPn 
ktM M   ktCo Cocat   k   klcb 0
(4.7)

The effect of temperature is widely acknowledged with the reason that the activation

energy for chain transfer is greater than that for propagation. Consequently, early termination of

the chains results in lower average moelcular weight. Polydispersity in all the cases is obtained

very close to 2.

As inferred from calculated fraction of dead chains with double bond at the end, major

modes of chain termination, nearly for all sets of conditions, are believed to be chain transfer to

monomer, chain transfer to cocatalyst and β-hydride elimination. Relative magnitude of different

transfer reactions is evident from estimated parameters at various temperatures. Long chain

branching frequency is detected to be negligibly low, except only for low (40 °C) temperature

and high (120 and 160 psig) ethylene pressures, suggesting that the product is comprising of

linear chains and posseses high density.

123
4.2 Propylene polymerization

Model description

Propylene polymerization model with some elementary assumptions is applied to the solution

phase production of polypropylene catalyzed with various zirconium and hafnium based catalyst

systems. In all the studies hereinafter, simulations are carried out using 'natural logarithmic

differential evolution' approach of optimization.

Assumptions

(i) Instantaneous formation of active sites.

(ii) First-order propagation with respect to monomer and the active site, and the reactivity of

the activated complex species does not devolve on the length of the polymer chain.

(iii) First-order deactivation of active sites.

(iv) Chain transfer following a propagation by primary (1, 2) insertion takes place by β-

hydride transfer to the monomer, producing a vinylidene-terminated dead chain and

liberating an active center.

(v) Chain transfer to catalyst (Zr) takes place by β-hydride elimination, producing a dead

chain and a hydride activated complex which reinitiates and follows the features of a

primitive propagating chain.

(vi) A secondary (2, 1) insertion brings about a mis-inserted chain, which can undergo

propagation and terminate by β-hydride transfer to the monomer to produce a dead chain

with butenyl-end group.

(vii) Monomer consumption in the initiation, secondary (2, 1) insertion and all chain transfer

reactions is negligible as compared to the propagation reaction via primary (1, 2)

insertion.

124
(viii) Chain transfer to cocatalyst (MAO) occurs, producing a dead chain and a methyl

activated complex which reinitiates a new chain.

Polymerization reactions considered based on the above assumptions are described in

Table 4.9.

Table 4.9 Reactions Considered in Propylene Polymerization

Reaction Stoichiometry Description

1. Cat  Cocat 
 P(0) Instantaneous catalyst activation

2. P0 M 
k in
P1 Chain initiation

3. Pi   M  Pi  1 Chain propagation


k
p

4. Pi  
kd
Pd 0  Di  {chain}
Spontaneous catalyst deactivation
5. P0 
kd
Pd 0 {site}

6. Pi  M k
tM
D(i)  P1 Chain transfer to monomer

7. Pi  
 D(i)  PH* 0 β-Hydride elimination
k
 ,H

8. PH* 0  M  Reinitiation after β-H elimination


r k
P(1)

9. Pi  M 
ks
R(i  1) Secondary (2, 1) insertion

10. Ri  M  P(i  1) Propagation after (2, 1) insertion


k
sp

11. Ri  M k


sM
 D(i)  P(1) Chain transfer after (2, 1) insertion

Pi  Cocat 


 D(i)  PMe 0
12. t , Alk * Chain transfer to cocatalyst

13. *
PMe 0  M k
rAl
P1 Reinitiation after transfer to
cocatalyst

125
4.2.1 Propylene polymerization with Me2Si[Ind]2ZrCl2 (P1)/MAO

Propylene polymerization model is applied to solution phase polymerization of propylene with

Me2Si[Ind]2ZrCl2/MAO catalyst system and kinetic parameters are obtained. Data for the model

validation are taken from Marques et al. (2002).

Propylene concentration in toluene is calculated from Equation 4.8 for a specific temperature

(in °C) and pressure. [Bravakis et al. (1998) referred and used in Marques et al. (2002)].

   
0.179  3.37  10 3 T  1.83  10 5 T 2 at 15 psi

X C3H 6    
 0.354  6.75  10 3 T  3.66  10 5 T 2 at 30 psi (4.8)
   
0.536  1.01  10 2 T  5.49  10 5 T 2
 at 45 psi

Estimated parameters and effect of temperature

Kinetic parameters for Me2Si[Ind]2ZrCl2 (P1)/MAO catalyst system are estimated by simulating

the model with experimental data at 25 °C and 75 °C with Al/Zr molar ratio of 500.

Experimental data for Al/Zr ratio of 2000 are used to validate the model at both temperatures.

Figure 4.13 and Figure 4.14 present experimental and model predicted propylene polymerization

rates at 25 °C and 75 °C respectively. Experimental observations reveal that both, the

temperature and MAO to catalyst molar ratio have a significant effect on polymerization rate.

Model predictions are in very close agreement with experimental observations at both the

temperatures and Al/Zr molar ratios of 500 & 2000.

Figure 4.13 and Figure 4.14 show that with an increase in Al/Zr molar ratio,

polymerization rate increases at both the temperatures considered. Larger Al/Zr ratio gives rise to

higher concentration of activated complex at the beginning of the reaction which increases the

rate of initiation and consequently the rate of propagation. At very high Al/Zr ratio, the
126
decreasing polymerization rate after reaching a maximum may be explained by large rate of

chain transfer to cocatalyst.

Figure 4.15 explains that the temperature has a profound effect on polymerization rate.

Maximum rate is seen to increase four folds at 75 °C when compared with that at 25 °C, at fixed

catalyst concentration, Al/Zr ratio and pressure. Increase in propagation rate constant k p with

temperature is also in coherence with this observed fact. As understood by k d and ktM values

obtained, spontaneous deactivation and chain transfer to monomer are activated hugely with

increase in temperature. Ascribing to which, the polymerization rate is decreasing steeply at 75

°C after reaching a maximum.

Kinetic parameters and objective function [F(k)] values which indicate the extent of fit of

experimental data with model prediction are given in Table 4.10.

Active catalyst site concentration drops to near zero within 5 minutes and 10 minutes at

25 °C and 75 °C respectively as shown in Figure 4.16. Solubility of propylene in solvent

decreases with increase in temperature at fixed pressure (Equation 4.8), thereby reducing the

concentration of monomer in the reaction mixture. Further, the fact, k in .[M ] = 2.2388×10-3 s-1 at

25 °C vs. k in .[M ] = 0.2753×10-3 s-1 at 75 °C explains that chain initiation rate is slower at 75

°C. Active sites are deactivated spontaneously as well, frequency of which is obtained higher (cf.

k d , Table 4.10) at 75 °C.

127
2.5
Zr:Al
1:500
2.0 1:2000

Polymerization rate (mol/L/s)


1.5

1.0

0.5

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.13 Effect of Al/Zr mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P1) = 10 μM, T = 25 °C and P = 30 psi]

Zr:MAO
6
1:500
Polymerization rate (moles / L / s)

1:2000

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.14 Effect of Al/Zr mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P1) = 10 μM, T = 75 °C and P = 30 psi]
128
6.0

T
25 °C

Polymerization rate (mol/L/s)


4.5
75 °C

3.0

1.5

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.15 Effect of temperature on propylene polymerization rate; solid lines are model
predictions.
[Catalyst (P1) = 10 μM, Al/Zr = 500 and P = 30 psi]

The overall rate of disappearance of active sites is higher at 25 °C due to which these are

exhausted little early. Fractional disappearance of active sites by spontaneous deactivation may

be determined by kd / kd  kin .[M ] . This fraction is found less at lower temperatures and as per

calculations 10.28% active sites at 25 °C whereas, 53.1% active sites at 75 °C are figured to be

deactivated spontaneously.

Maximum concentration of hydride activated complex P 0


*
H is noted to be

3.5409×10-9 μM (in 7 minutes) at 25 °C and 4.0866×10-7 μM (in 14 minutes) at 75 °C as shown

in Figure 4.17 and Figure 4.18 respectively.

129
Table 4.10 Estimated Parameters for Me2Si[Ind]2ZrCl2 Table 4.11 Predicted Properties with Me2Si[Ind]2ZrCl2
(P1)/MAO (P1)/MAO

T (°C) 25 75 T (°C) 25 75
k in ×103 (M-1.s-1) 5.1845 10.7586 Exp Model Exp Model

M n ×10-4 - 13.549 - 0.9925


k p × 10-4 (M-1.s-1) 8.566 39.4588
(g/mol)
k d × 104 -1
(s ) 2.5665 3.1176
M w ×10-4 12.3 27.257 1.5 1.9984

k tM (M-1.s-1) 2.5699 10.2576 (g/mol)


PDI 1.8 2.011 2.8 2.013
6 -1 1.55277 11.7565
k  , H ×10 (s ) - 74.1 - 86.3
f v (%)
k r ×10-2 (M-1.s-1) 1.5294 152.700
f b (%) - 5.2 - 3.4
k s ×107 (M-1.s-1) 1.3249 11.6351 - 20.7 - 10.3
f i (%)
k sp ×104 (M-1.s-1) 1.4583 118.967

k sM (M-1.s-1) 0.047 1.0394

k t , Al (M-1.s-1) 426.688 141.243

k rAl (M-1.s-1) 8.574 13.0285

F(k) (-) 0.272 0.03713

130
10
T
25 °C

Active catalyst sites (moles/L)


75 °C

0 10 20 30 40 50 60
Time (minutes)

Figure 4.16 Active catalyst site concentration vs. time.


[Catalyst (P1) = 10 μM, Al/Zr = 500 and P = 30 psi]

With P1/MAO catalyst system, β-H elimination is observed to occur negligibly ( k  , H of

the order of 10-6) as compared to all other modes of chain termination. Increase in temperature

enhances the frequency of β-H elimination. Due to higher concentration of monomer at 25 °C

than that at 75 °C, reinitiation rate after β-H elimination is higher, which is reflected by the

falling concentration of hydride activated complex after reaching a maximum. At 75 °C, due to


lower reinitiation rate, PH* 0 is almost steady beyond maximum. 
Concentration of methyl activated complex P 0 is found to reach a maximum of
*
Me

0.8142 μM (1.8 min) followed by a steep decrease at 25 °C as compared to a maximum of

0.2329 μM (3.58 min) at 75 °C followed by slow decrease as shown in Figure 4.19.

131
-9
4.0x10

Hydride activated complex (moles/L)


-9
3.0x10

-9
2.0x10

-9
1.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.17 Hydride actived complex concentration vs. time.


[Catalyst (P1) = 10 μM, Al/Zr = 500, T = 25 °C and P = 30 psi]
Hydride activated complex (moles/L)

-7
4.0x10

-7
2.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.18 Hydride activated complex concentration vs. time.


[Catalyst (P1) = 10 μM, Al/Zr = 500, T = 75 °C and P = 30 psi]

132
Chain transfer to cocatalyst is decreasing with increase in temperature. This trend is also

observed with the subsequent cases studied in this work. MAO is an oligomer and consists

mainly of units of the basic structure [Al4O3Me6], which contains four aluminium, three oxygen

atoms and six methyl groups. As the aluminium atoms in this structure are co-ordinatively

unsaturated, the basic units join together and give rise to different structures as described in

Figure 2.10. The observed trend of decrease in chain transfer to MAO with increase in

temperature suggests the existence of different molecular structures of MAO at different

temperatures. It seems that MAO changes its structure from simple (linear or cyclic) at low

temperatures to a congested one (ladder or cage) at high temperatures and thereby offering a

steric hindrance for chain transfer. At 75 °C, the rectivation rate, after transfer to cocatalyst is

slow due to low concentrations of monomer and methyl activated complex, for which the

*
concentration of PMe 0 is seen decreasing slowly opposite to that at 25 °C.

0.9

T
Methyl activated complex (moles/L)

25 °C
75 °C
0.6

0.3

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.19 Methyl activated complex concentration vs. time.


[Catalyst (P1) = 10 μM, Al/Zr = 500 and P = 30 psi]

133
Average molecular weights, PDI and percentage of butenyl- & isobutyl-terminated chains

predicted by the model are given in Table 4.11. Degree of polymerization depends upon relative

rate of propagation against various termination rates as given by Equation 4.9.

k p M 
DPn 
ktM M   kt , Al Cocat   k  ,H  kd  k s M 
(4.9)

Agreement between experimental observations and predicted results for MWD is shown

in Table 4.9. Low molecular weights obtained at higher (75 °C) temperature, may be attributed to

the relatively high frequency of termination via transfer to monomer, spontaneous deactivation

and β-H transfer. The model predicts Schulz-Flory distribution with a polydispersity index very

around 2 for all reaction conditions.

For P1/MAO, chain termination is understood to take place majorly via spontaneous

catalyst deactivation, transfer to monomer and β-Hydride transfer because f v values in Table

4.11 point the dominant presence of chains with vinylidene end group. Low f b values suggest

that the fraction of chains with butenyl end group which are produced by termination after

secondary insertion is miserable, which is expected for highly isotactic polypropylene. However,

with increase in temperature, f b is observed to decrease. This is due to relatively higher increase

in k p than the increase in k s values with increase in temperature. Further, it has been noted that

increased regioirregular (2,l) insertions slow down chain propagation and inhibit chain transfer to

the monomer. This effect is consistent with previous results reported in the literature for the

studied catalyst system [Busico et al. (1998)].

134
f i value (Table 4.11) shows, 20.7% terminated chains bear isobutyl end group at 25 °C, which

represent a significant chain transfer to cocatalyst. Further, increase in temperature results in a

decreased rate of chain transfer to cocatalyst which is indicated by the decrease in f i (10.3% at

75 °C). This chain transfer dominates at lower temperature, due to depressed, competing β-H

elimination rates [Amin (2007)].

Effect of Pressure

Figure 4.20 shows that an increase in monomer pressure results in a steady increase in

polymerization rate up to a maximum ranging in between 0.685 moles/L/s (15 psi) to 2.056

moles/L/s (45 psi) at 25 °C and fixed Zr = 10 μM, Al/Zr = 500. Increase in pressure seems to

increase the monomer concentration at reaction site and thereby increasing the propagation rate

proportionally as expected and obtained from simulation results. At 75 °C, higher polymerization

rates are observed {2.489 moles/L/s (15 psi) to 8.487 moles/L/s (45 psi)}, which do not continue

maintained due to high rates of termination at this temperature (Figure 4.21).

A marginal increase of 2.22 % (15-30 psi) and 0.87 % (30-45 psi) in weight average

 
molecular weight M w is observed with increase in pressure at 25 °C (Figure 4.22). Similarly,

at 75 °C, a maximum increment of 10.31 % in M w is seen (Figure 4.23). Therefore, an increase

in monomer pressure is believed to increase the polymerization rate appreciably with an incresed

intake of monomer but brings negligible effect on polymer molecular weights.

135
2.5

45 psi

Polymerization rate (moles/L/s)


2.0

1.5
30 psi

1.0

15 psi

0.5

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.20 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P1) = 10 μM, Al/Zr = 500, and T = 25 °C]

8
Polymerization rate (moles/L/s)

4 45 psi

30 psi
2
15 psi

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.21 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P1) = 10 μM, Al/Zr = 500 and T = 75 °C]

136
5
3x10

Molecular weight (g/mol)


Mn
5
2x10 Mw

5
1x10

15 30 45
Pressure (psi)

Figure 4.22 Effect of pressure on average molecular weights.


[Catalyst (P1) = 10 μM, Al/Zr = 500 and T = 25 °C]

4
2.5x10

4
2.0x10
Molecular weight (g/mol)

Mn
4
1.5x10 Mw

4
1.0x10

3
5.0x10

15 30 45
Pressure (psi)

Figure 4.23 Effect of pressure on average molecular weights.


[Catalyst (P1) = 10 μM, Al/Zr = 500 and T = 75 °C]

137
Effect of catalyst concentration

Polymerization rates at various catalyst concentrations (10, 20, 40 and 80 μM) at 25 °C and 75

°C are shown in Figure 4.24 and Figure 4.25 respectively. On doubling the catalyst

concentration, maximum polymerization rate is almost doubled, showing a linear proportional

dependence. Higher polymerization rates with decreasing trend after reaching a maximum are

obtained at 75 °C, which can be explained with the fact that along with propagation, termination

rates also increase profusely at high temperatures. Average molecular weights are found to be

insensitive to catalyst concentration as shown in Figure 4.26 and Figure 4.27. Catalysts

concentration may affect the degree of polymerization via β-H transfer only. Significantly low

values of k  , H (Table 4.10), as compared to other transfer rate constants, warrant that change in

catalyst concentration has no effect on molecular weights.

12
[Zr] = 80M

10
Polymerization rate (moles/L/s)

6 [Zr] = 40M

4
[Zr] = 20M

2
[Zr] = 10M

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.24 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 500, T = 25 °C and P = 30 psi]

138
50

Polymerization rate (moles/L/s)


40

30

[Zr] = 80M
20

[Zr] = 40M
10
[Zr] = 20M

[Zr] = 10M
0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.25 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 500, T = 75 °C and P = 30 psi]

5
3x10
Molecular weight (g/mol)

Mn
5
2x10 Mw

5
1x10

0 20 40 60 80
[Zr] (M)

Figure 4.26 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 500, T = 25 °C and P = 30 psi]

139
4
2.5x10

4
2.0x10

Molecular weight (g/mol) Mn


4
1.5x10 Mw

4
1.0x10

3
5.0x10
0 20 40 60 80
[Zr] (M)

Figure 4.27 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 500, T = 75 °C and P = 30 psi]

4.2.2 Propylene polymerization with Et(Ind)2ZrCl2 (P2)/MAO

Propylene polymerization model is applied to solution phase polymerization of propylene with

Et(Ind)2ZrCl2/MAO catalyst system and kinetic parameters are obtained. Simulations are carried

out numerically using 'natural logarithmic differential evolution' approach of optimization. Data

for the model validation are taken from Marques et al. (2002).

Estimated parameters and effect of temperature

Model predictions of reaction rate captures the typical behavior of experimental rate profiles for

propylene polymerization with Et(Ind)2ZrCl2/MAO catalyst system as shown in Figures. 4.28

and 4.29. Estimated kinetic parameters and objective function values are given in Table 4.12. At

75 °C, parameters are evaluated at an Al/Zr molar ratio of 500 and the experimental data at a

140
ratio of 2000 are used to verify the model prediction for polymerization rate. Excellent

predictions are obtained as shown in Figures. 4.28 and 4.29.

Figure 4.30 presents the concentration profiles of active catalyst site at 25 °C and 75 °C.

Active sites are found to decrease exponentially from 10 μM to almost nil in 27 minutes at 25 °C

and within 15 minutes at 75 °C at fixed pressure 30 psi and Al/Zr = 2000. k in .[M ] = 3.61×10-4

at 25 °C vs. k in .[M ] = 9.39×10-4 at 75 °C suggest that chain initiation rate is faster at 75 °C and

therefore active catalyst sites take lesser time to exhaust than that at 25 °C. Spontaneous

deactivation of catalyst site is also increased staggeringly at 75 °C with aids to lower

concentration of active sites observed. As low as 0.35% active catalyst sites at 25 °C against

23.9% at 75 °C are disappearing by spontaneous deactivation.

141
2

Polymerization rate (mol/L/s)

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.28 Polymerization rate vs. time; solid lines are model predictions.
[Catalyst (P2) = 10 μM, Al/Zr = 2000, T = 25 °C and P = 30 psi]

Zr:MAO
3 1:500
1:2000
Polymerization rate (moles/L/s)

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.29 Polymerization rate vs. time; solid lines are model predictions.
[Catalyst (P2) = 10 μM, T = 75 °C and P = 30 psi]

142
Table 4.12 Estimated Parameters for Et(Ind)2ZrCl2 (P2)/MAO

T (°C) 25 75
k in ×10 3
(M .s ) 1.7348
-1 -1 1.7686

k p ×10-5 (M-1.s-1) 1.1175 2.0135

k d ×106 (s-1) 1.2638 294.8404

k tM ×104 (M-1.s-1) 9.4135 536.7625

k  , H ×106 (s-1) 6.9652 9.6698


k r ×10-2 (M-1.s-1) 2.8886 3.4166

k s ×105 (M-1.s-1) 9.9996 11.6170

k sp (M-1.s-1) 4.5639 68.0812

k sM ×104 (M-1.s-1) 4.2017 46.3481

k t , Al ×10-4 (M-1.s-1) 6.4558 1.5515

k rAl (M-1.s-1) 12.9809 93.7468


F(k) (-) 0.2995 0.3615

Table 4.13 Predicted Polypropylene Properties with Et(Ind)2ZrCl2 (P2)/MAO

T (°C) 25 75
Al/Zr 2000 500 2000
Exp Model Exp Model Exp Model
M n ×10-4 (g/mol) - 2.5785 - 0.49795 - 0.3316

M w ×10-4 (g/mol) 5.4 5.3249 1.2 1.0353 0.8 0.7434


PDI (-) 2.2 2.065 2.1 2.079 1.9 2.242
f v (%) - 51.2 - 71.2 - 64.9
f b (%) - 4.6 - 4.1 - 3.8
f i (%) - 44.2 - 24.7 - 31.3

143
12

T
25 °C
75 °C

Active catalyst sites (moles/L)


8

0 10 20 30 40 50 60
Time (minutes)

Figure 4.30 Active catalyst site concentration vs. time.


[Catalyst (P2) = 10 μM, Al/Zr = 2000 and P = 30 psi]

Like P1/MAO catalyst system, with Et(Ind)2ZrCl2 (P2)/MAO also, β-H elimination is not

observed significant at both the temperatures considered. Maximum concentration of hydride

activated complex P 0 *


H is got to be at 2.2354×10-8 μM (in 14 minutes) at 25 °C and

1.7542×10-7 μM (in 27 minutes) at 75 °C as shown in Figure 4.31 and Figure 4.32 respectively.

k  , H values (Table 4.10) suggest that the rate of β-H elimination is low at 25 °C, for which the

concentration of PH* 0 is low at 25 °C as compared to that at 75 °C. Higher reinitiation rate after


β-H elimination at 25 °C is obtained due to which PH* 0 decreases after reaching the maxima. 

144
-8
2.25x10

Hydride activated complex (moles/L)


-8
1.50x10

-9
7.50x10

0.00
0 10 20 30 40 50 60
Time (minutes)

Figure 4.31 Hydride actived complex concentration vs. time.


[Catalyst (P2) = 10 μM, Al/Zr = 2000, T = 25 °C and P = 30 psi]

-7
1.8x10
Hydride activated complex (moles/L)

-7
1.2x10

-8
6.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.32 Hydride actived complex concentration vs. time.


[Catalyst (P2) = 10 μM, Al/Zr = 2000, T = 75 °C and P = 30 psi]

145
Chain transfer to cocatalyst is found to be very high at Al/Zr = 2000 at both the temperatures. At

25 °C, concentration of methyl activated complex is found to reach a maximum of 5.484 μM

(3.37 min) as compared to 3.932 μM (4 min) at 75 °C as shown in Figure 4.33. Methyl activated

complex concentration after reaching the maximum is decreasing steeply to a negligible value at

6.2 min (25 °C) and 7.63 min (75 °C). The rectivation rate constant, after transfer to cocatalyst

is high but monomer concentration is less at 75 °C. Consequently, the reactivation rates are

comparable and high at both the temperatures ( krAl M   2.7013 at 25 °C, krAl M   4.9779 at

*
75 °C), which clarifies the quick decrease in concentration of PMe 0 with time.

T
Methyl activated complex (moles/L)

25 °C
75 °C

0 10 20 30 40 50 60
Time (minutes)

Figure 4.33 Methyl actived complex concentration vs. time.


[Catalyst (P2) = 10 μM, Al/Zr = 2000 and P = 30 psi]

146
Average molecular weights, PDI and percentage of vinylidene-, butenyl- & isobutyl-terminated

chains predicted by the model for Et(Ind)2ZrCl2/MAO catalyst system are presented in Table

4.13. Model predictions are in very close correspondence with the experimental values. Effect of

temperature on MWD is obtained similar to that discussed for P1/MAO system in Section

4.2.1.1, however P2 yields low molecular weight product at matched reaction conditions. PDIs

at all reaction conditions are obtained near 2.0.

For P2/MAO, f v values in Table 4.13 indicate that chain termination takes place mainly

via spontaneous catalyst deactivation, transfer to monomer and β-Hydride and increasing with an

increase in temperature. With an increase in Al/Zr ratio, f v decreases due to enhanced chain

transfer to cocatalyst. Low f b values obtained advise that the fraction of chains with butenyl

end group is very less as compared to others. f b is noted to decrease with increase in

temperature as well as Al/Zr mole ratio. 44.2 % chains at 25 °C and 31.3 % chains at 75 °C are

found to terminate via chain transfer to cocatalyst as indicated by f i values (Table 4.13, Al/Zr =

2000). This trend of decreasing rate of chain transfer to cocatalyst with increase in temperature is

consistent with that obtained with P1/MAO system. The rate of chain transfer to cocatalyst is

proportional to the catalyst concentration. f i value is found to increase from 23.6 % to 43.2 %

with increase in Al/Zr mole ratio from 500 to 2000 respectively at 75 °C.

Effect of Pressure

With increase in monomer pressure an increase in polymerization rate up to a stabilized

maximum of 0.8808 moles/L/s (15 psi) and 2.678 moles/L/s (45 psi) at 25 °C and fixed Zr = 10

μM, Al/Zr = 2000 is observed as shown in Figure 4.34. At 75 °C, little higher polymerization

147
rates {Rpmax: 1.083 moles/L/s (15 psi) and 3.96 moles/L/s (45 psi)} with decreasing trend are

observed (Figure 4.35). The changes in average molecular weights with variation in pressures at

25 °C (Figure 4.36) and at 75 °C (Figure 4.37) are trifling. The discussion on the trends of

resultant rate profiles and average molecular weights at varied pressures given in Section 4.2.1.1

adjudges equivalently appropriate for the results obtained for Et(Ind)2ZrCl2/MAO catalyst

system.

3
45 psi
Polymerization rate (moles/L/s)

2
30 psi

1 15 psi

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.34 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P2) = 10 μM, Al/Zr = 2000 and T = 25 °C]

148
4

Polymerization rate (moles/L/s)


3

45 psi
2

30 psi

1 15 psi

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.35 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P2) = 10 μM, Al/Zr = 500 and T = 75 °C]

4
6.0x10
Molecular weight (g/mol)

4
4.5x10
Mn
Mw
4
3.0x10

4
1.5x10

15 30 45
Pressure (psi)

Figure 4.36 Effect of pressure on average molecular weights.


[Catalyst (P2) = 10 μM, Al/Zr = 2000 and T = 25 °C]

149
4
1.2x10

Molecular weight (g/mol)


3
9.0x10

Mn
Mw

3
6.0x10

3
3.0x10

15 30 45
Pressure (psi)

Figure 4.37 Effect of pressure on average molecular weights.


[Catalyst (P2) = 10 μM, Al/Zr = 500 and T = 75 °C]

Effect of catalyst concentration

Polymerization rates at different catalyst (Et(Ind)2ZrCl2) concentrations at 25 °C and 75 °C are

shown in Figure 4.38 and Figure 4.39 respectively. Like Me2Si[Ind]2ZrCl2 (P1), with this catalyst

system also, polymerization rate is found to be linearly dependent on the catalyst concentration.

However P2 offers a lower polymerization rates than P1 under similar reaction conditions. To

compare, at Zr = 10 μM and 75 °C, a maximum rate of 2.45 moles/L/s with P2 is obtained

against 8.475 mole/L/s with P1. Additionally, average molecular weights are observed to be

unaffected by catalyst concentration as shown in Figure 4.40 and Figure 4.41, which can be

attributed to very low values of k  , H (Table 4.11), as compared to other transfer rate constants.

This has been predicted that catlyst P2 produces lower molecular weight polypropylene when

compared to P1 for alike reaction conditions (Table 4.9 and Table 4.11). The effect of catalyst

150
concentration upon rate and molecular weights are qualitatively similar to that discussed for

catalyst P1 in Section 4.2.1.1.

16
[Zr] = 80M
Polymerization rate (moles/L/s)

12

8
[Zr] = 40M

4 [Zr] = 20M

[Zr] = 10M

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.38 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 2000, T = 25 °C and P = 30 psi]

20
Polymerization rate (moles/L/s)

15

[Zr] = 80M

10

[Zr] = 40M
5
[Zr] = 20M

[Zr] = 10M
0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.39 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 500, T = 75 °C and P = 30 psi]
151
4
6.0x10

4
5.0x10
Molecular weight (g/mol)
Mn
4
4.0x10 Mw

4
3.0x10

4
2.0x10

0 20 40 60 80
[Zr] (M)

Figure 4.40 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 2000, T = 25 °C and P = 30 psi]

4
1.2x10
Molecular weight (g/mol)

3
9.0x10
Mn
Mw

3
6.0x10

0 20 40 60 80
[Zr] (M)

Figure 4.41 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 500, T = 75 °C and P = 30 psi]

152
4.2.3 Propylene polymerization with Me2Si(Ind)2HfCl2 (P3)/MAO

Simulation results obtained with propylene polymerization model applied to solution phase

polymerization of propylene with Me2Si(Ind)2HfCl2/MAO catalyst system are presented and

discussed in this section. Data for the model validation were taken from Marques et al. (2002).

Estimated parameters and effect of temperature

Experimental rate data at Al/Hf = 2000 are regressed with the model to determine kinetic

parameters. Further, the model is simulated to predict the polymerization rate at Al/Hf = 500.

Estimated kinetic parameters and objective function [F(k)] values are provided in Table 4.14.

Figures. 4.42 and 4.43 depict a good agreement between the experimental data and model

predictions of polymerization rate with Me2Si(Ind)2HfCl2 (P3)/MAO catalyst system at 40 °C

[F(k) = 1.5175] and 80 °C [F(k) = 0.3161] respectively. Experimental observations reveal that at

40 °C, polymerization rate is weakly affected by a change in Al/Hf ratio as shown in Figure 4.42.

Whereas at 80 °C, polymerization rate increases with increase in cocatalyst to catalyst mole ratio

(Figure 4.43).

Polymerization rate is observed to increase with increase in temperature. At fixed catalyst

concentration, Al/Hf ratio (2000) and monomer pressure, maximum rate is found to be 0.4068

moles/L/s (29 minutes) at 40 °C against 0.9378 moles/L/s (7 minutes) at 80 °C. A rapid gain in

propagation rate at 80 °C is evident from lower frequency of spontaneous catalyst deactivation

(Table 4.12) and a higher initiation rate at this temperature, which provides high concentration of

initiated chains during initial period. Hafnium based metallocene catalysts are known for

producing high molecular weight polymers in comparison to their zirconium analogues, but at

153
the expense of substantially reduced catalytic activity [Ewen et al. (1987); Nakayama and Shiono

(2005)].

Figure 4.44 lays out the concentration profiles of active catalyst site at 40 °C and 80 °C.

Active sites are found to decrease slowly from 10 μM to 1.37 μM in entire polymerization time

at 40 °C and from 10 μM to 0.0124 μM within 19.6 minutes at 80 °C at fixed pressure 30 psi

and Al/Zr = 2000. Chain initiation rate is faster at 80 °C ( k in .[M ] = 1.0898×10-4 s-1 vs. 4.34×10-

5
s-1 at 40 °C, Table 4.14) and the frequency of spontaneous deactivation of catalyst sites is also

found increasing with temperature (cf. k d , Table 4.14). Therefore active catalyst sites are

consumed in very less time at 80 °C than that at 40 °C.

Table 4.14 Estimated Parameters for Me2Si(Ind)2HfCl2 (P3)/MAO

T (°C) 40 80
k in ×103 (M-1.s-1) 0.3045 2.2591

k p ×10-4 (M-1.s-1) 4.0398 5.5376

k d ×104 (s-1) 1.8879 5.3001

k tM ×102 (M-1.s-1) 0.2771 8.7269

k  , H ×106 (s-1) 9.2795 188.81

k r ×10-3 (M-1.s-1) 1.4199 22.8680

k s ×108 (M-1.s-1) 0.2002 1.4798

k sp ×105 (M-1.s-1) 1.1091 684.3200

k sM ×108 (M-1.s-1) 1.2084 102.4800

k t , Al (M-1.s-1) 183.5400 99.5750

k rAl ×10-3 (M-1.s-1) 1.0293 2.4608


F(k) (-) 1.5175 0.3161

154
Table 4.15 Predicted Properties with Me2Si(Ind)2HfCl2 (P3)/MAO

T (°C) 40 80
Al/Zr 500 2000 500 2000
Exp Model Exp Model Exp Model Exp Model
M n ×10-4 (g/mol) - 10.7115 - 12.1355 - 2.4346 - 1.6478

M w ×10-4 (g/mol) 34.6 30.0459 27.1 24.3074 6.2 4.8698 3.9 3.2919
PDI 2.9 2.8050 2.8 2.003 1.9 2.0002 1.9 1.9977
f v (%) - 86.6 - 87.2 - 90.2 - 90.7
f b (%) - 4.2 - 3.1 - 3.5 - 2.4
f i (%) - 9.2 - 9.7 - 6.3 - 6.9

Hf:MAO
1:500
1:2000
0.4
Polymerization rate (moles/L/s)

0.2

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.42 Effect of Al/Hf mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P3) = 10 μM, T = 40 °C and P = 30 psi]

155
1.0

Hf:MAO
1:500
0.8

Polymerization rate (moles/L/s)


1:2000

0.6

0.4

0.2

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.43 Effect of Al/Hf mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P3) = 10 μM, T = 80 °C and P = 30 psi]

T
10 40 °C
Active catalyst sites (moles/L)

80 °C

0 10 20 30 40 50 60
Time (minutes)

Figure 4.44 Active catalyst site concentration vs. time.


[Catalyst (P3) = 10 μM, Al/Hf = 2000 and P = 30 psi]

156
Spontaneous deactivation rate of active catalyst sites is high with comparison to the rate of

initiation at both temperatures. 81.3% active sites at 40 °C and 82.9% at 80 °C are calculated to

disappear due to spontaneous deactivation.

With Me2Si(Ind)2HfCl2 (P3)/MAO system, frequency of β-H elimination is found to increase

from 9.2795×10-6 s-1 at 40 °C to 1.8881×10-4 s-1 at 80 °C (Table 4.14).

At 40 °C, concentration of hydride activated complex reaches a maximum of 1.906×10-7

μM (in 28 minutes) and decreases with a slower rate, on the other hand, at 80 °C it reaches a

maximum of 2.327×10-6 μM (in 8.85 minutes) and decreases appreciably as shown in Figure

4.45. Reinitiation rate after β-H elimination is high ( k r >103) at both the temperatures, so


considerable decrease in PH* 0 is seen at higher concentration. 

-6
3.0x10

T
Hydride activated complex (moles/L)

-6
2.5x10
40 °C
80 °C
-6
2.0x10

-6
1.5x10

-6
1.0x10

-7
5.0x10

0.0

0 10 20 30 40 50 60
Time (minutes)

Figure 4.45 Hydride actived complex concentration vs. time.


[Catalyst (P3) = 10 μM, Al/Hf = 2000 and P = 30 psi]

157
Chain transfer to cocatalyst is seen significant at Al/Hf = 2000 at both the temperatures. At 40

°C, concentration of methyl activated complex is observed to reach a maximum of 6.96×10-4 μM

(26.8 min) as compared to 1.836×10-5 μM (26.4 min) at 80 °C as shown in Figure 4.46 and

Figure 4.47 respectively. Reactivation rates are high and comparable at both the temperatures

hence methyl activated complex concentration after reaching the maximum is decreasing

appreciably.

Molecular weight distribution and percentage of vinylidene-, butenyl- & isobutyl-

terminated chains predicted by the model for Me2Si(Ind)2HfCl2 (P3)/MAO system are given in

Table 4.15.

-4
8.0x10
Methyl activated complex (moles/L)

-4
6.0x10

-4
4.0x10

-4
2.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.46 Methyl actived complex concentration vs. time.


[Catalyst (P3) = 10 μM, Al/Hf = 2000, T = 40 °C and P = 30 psi]

158
-5
2.0x10

Methyl activated complex (moles/L)


-5
1.6x10

-5
1.2x10

-6
8.0x10

-6
4.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.47 Methyl actived complex concentration vs. time.


[Catalyst (P3) = 10 μM, Al/Hf = 2000, T = 80 °C and P = 30 psi]

Model predictions are in good agreement with the experimental values. Effect of

temperature on MWD is obtained qualitatively similar to that observed for P1/MAO and

P2/MAO systems (Sections 4.2.1.1 and 4.2.2.1). It is worth noting that Me2Si(Ind)2HfCl2

(P3)/MAO system yields a way higher molecular weight polypropylene when compared to its Zr

analogue i.e. P1/MAO. Model predicted value of PDI at 40 °C with Al/Hf molar ratio is 2.8 (exp.

value = 2.9) and for all other reaction conditions is about 2.0.

Like other catalyst systems discussed so far, with P3/MAO also, the major pathway of

chain termination is via spontaneous catalyst deactivation, transfer to monomer and β-Hydride

elimination. Among these β-Hydride elimination is found increasing dominantly whereas

spontaneous deactivation and transfer to monomer are seen to decrease a bit, with an increase in

159
temperature. f v values are increasing with increase in temperature at a constant Al/Hf molar ratio

as shown in Table 4.15, but there is negligible increase with an increase in Al/Hf ratio at a given

temperature. Fraction of chains with butenyl end group is very less as compared to others as

inferred by low f b values obtained, which are also noticed to decrease with increase in

temperature as well as Al/Hf mole ratio.

9.2 % chains at 40 °C and 6.3 % chains at 80 °C ( f i values at Al/Hf = 500) are found to

terminate via chain transfer to cocatalyst which increases with increase in Al/Hf molar ratio. The

trend of decreasing rate of chain transfer to cocatalyst with increase in temperature is similar to

that obtained with its Zr analogue (P1/MAO system) but the amount of this termination is quite

less with P3/MAO system.

Effect of Pressure

Polymerization rates up to a maximum of 0.131 moles/L/s (15 psi), 0.38 moles/L/s (30 psi) and

0.693 moles/L/s (45 psi) at 40 °C and fixed Zr = 10 μM, Al/Hf = 2000 are obtained as shown in

Figure 4.48. At 80 °C, maximum rates of 0.382 moles/L/s at 15 psi, 0.855 moles/L/s at 30 psi

and 1.350 moles/L/s at 45 psi are obtained (Figure 4.49). A steady increase in polymerization

rate with an increase in monomer pressure is seen at both the temperatures. Higher initiation rate

at 80 °C, provide larger concentration of live chains which in turn alleviate to reach high

polymerization rate in less time.

Effect of monomer pressure on molecular weight is found potential at lower temperature.

A significant increase of 71.6 % (15-30 psi) and 30.72 % (30-45 psi) in weight average

 
molecular weight M w is observed with increase in pressure at 40 °C (Figure 4.50). At 80 °C,

160
relatively lower increase of 18.8 % (15-30 psi) and 6.73 % (30-45 psi) in M w is observed with

increase in pressure (Figure 4.51).

0.75

Polymerization rate (moles/L/s)

0.50 45 psi

30 psi

0.25

15 psi

0.00
0 10 20 30 40 50 60
Time (minutes)

Figure 4.48 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P3) = 10 μM, Al/Hf = 2000 and T = 40 °C]

1.50

1.25
Polymerization rate (moles/L/s)

1.00

45 psi
0.75

30 psi
0.50

15 psi
0.25

0.00
0 10 20 30 40 50 60
Time (minutes)

Figure 4.49 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P3) = 10 μM, Al/Hf = 2000 and T = 80 °C]
161
5
3.5x10

5
2.8x10

Molecular weight (g/mol)


Mn
5
2.1x10 Mw

5
1.4x10

4
7.0x10

15 30 45
Pressure (psi)

Figure 4.50 Effect of pressure on average molecular weights.


[Catalyst (P3) = 10 μM, Al/Hf = 2000 and T = 40 °C]

4
4.0x10

4
Molecular weight (g/mol)

3.0x10
Mn
Mw

4
2.0x10

4
1.0x10

15 30 45
Pressure (psi)

Figure 4.51 Effect of pressure on average molecular weights.


[Catalyst (P3) = 10 μM, Al/Hf = 2000 and T = 80 °C]

162
Effect of catalyst concentration

Polymerization rates at different catalyst concentrations (10, 20, 30, 40 and 50 μM) at 40 °C and

80 °C are shown in Figure 4.52 and Figure 4.53 respectively. With increase in catalyst

concentration, more number of active catalyst sites are rendered, which increase the rate of

initiation even if monomer concentration is invariant. Propagation rate is also dependent of

concentration of live chains. On this account, a higher polymerization rate is expected with

increase in catalyst concentration, which is correctly predicted by the model at both the

temperatures. Due to higher initiation rate at 80 °C, maximum polymerization rate is achieved

earlier than that at 40 °C at corresponding catalyst concentrations. Since high catalyst

concentration promotes β-H elimination, polymerization rate diminishes at a faster rate as seen in

Figure 4.52 and Figure 4.53 respectively. This effect is particularly pronounced at 80 °C due to

very high k  , H value. Further, the model predicts a proportional dependence of propagation rate

on catalyst concentration.

With P3/MAO catalyst system, average molecular weights are noticed to be decreasing

with increase in catalyst concentration as shown in Figure 4.54 and Figure 4.55. An inverse

relationship between polymer molecular weight and catalyst (Z-N/metallocene) concentration

have been obtained and reported in literature [Breslow and Newburg (1959), Chien (1959),

Brintzinger (1995)]. With P3/MAO, this effect is more influencing at 40 °C, where higher

molecular weights are obtained. A decrease of 35.1% (Hf = 10-20 μM), 21.5% (Hf = 20-30

μM), 14% (Hf = 30-40 μM) and 9.1% (Hf = 40-50 μM) in M w at 40 °C (Figure 4.54), whereas

at 80 °C, 2.1 - 2.3% (Figure 4.55) decrease is obtained.

163
2.0

[Zr]

Polymerization rate (moles/L/s)


1.5 50M

40M

1.0
30M

20M
0.5

10M

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.52 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Hf = 2000, T = 40 °C and P = 30 psi]

4.0

3.5
Polymerization rate (moles/L/s)

3.0

2.5
[Zr]

2.0 50M
40M
1.5
30M
1.0 20M

0.5 10M

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.53 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Hf = 2000, T = 80 °C and P = 30 psi]

164
5
2.8x10

Mn
5
2.1x10 Mw

Molecular weight (g/mol)


5
1.4x10

4
7.0x10

0.0
0 10 20 30 40 50
[Hf] (M)

Figure 4.54 Effect of catalyst concentration on average molecular weights.


[Al/Hf = 2000, T = 40 °C and P = 30 psi]

4
4.0x10

4
Molecular weight (g/mol)

3.0x10

Mn
Mw
4
2.0x10

4
1.0x10

0 10 20 30 40 50
[Hf] (M)

Figure 4.55 Effect of catalyst concentration on average molecular weights.


[Al/Hf = 2000, T = 80 °C and P = 30 psi]

165
4.2.4 Propylene polymerization with Et(Ind)2HfCl2 (P4)/MAO

Propylene polymerization model is applied to solution phase polymerization of propylene with

Et(Ind)2HfCl2 catalyst and kinetic parameters are obtained. Data for the model validation were

taken from Marques et al. (2002).

Estimated parameters and effect of temperature

Kinetic parameters for Et(Ind)2HfCl2 (P4)/MAO catalyst system are estimated by simulating the

model with experimental data at 40 °C and 80 °C with Al/Hf molar ratio of 500. The model

predictions are affirmed with experimental data for Al/Hf ratio of 2000 at both temperatures.

Kinetic parameters and objective function [F(k)] values obtained for P4/MAO system are given

in Table 4.16. Figure 4.56 and Figure 4.57 present experimental and model predicted propylene

polymerization rates at 40 °C and 80 °C respectively. Model predictions are in conformation

with experimental observations at both the temperatures and Al/Hf molar ratios of 500 & 2000.

Figure 4.56 and Figure 4.57 show that with an increase in Al/Hf molar ratio,

polymerization rate increases at both the temperatures. The trend observed is coherent with the

previously discussed studies. At high Al/Hf ratio, large rate of chain transfer to cocatalyst

induces the decrease in polymerization rate after reaching a maximum. In general, increase in

temperature is found to upshoot in higher polymerization rate.

With P4/MAO system, Rp,max is found to increase from 0.5216 moles/L/s at 40 °C to

1.092 moles/L/s at 80 °C at Al/Hf molar ratio of 2000, following the general trend. But with this

catalyst system at Al/Hf ratio of 500, Rp,max is found to decrease meagerly from 0.1952 moles/L/s

at 40 °C to 0.1148 moles/L/s at 80 °C.

166
Hf:MAO
0.6 1:500
1:2000

Polymerization rate (moles/L/s) 0.4

0.2

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.56 Effect of Al/Hf mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P4) = 10 μM, T = 40 °C and P = 30 psi]

Hf:MAO
1:500
1.0
1:2000
Polymerization rate (moles/L/s)

0.5

0.0

0 10 20 30 40 50 60
Time (minutes)

Figure 4.57 Effect of Al/Hf mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P4) = 10 μM, T = 80 °C and P = 30 psi]

167
This surprising drift indicates that Al/Hf ratio of 500 is too low to efficiently activate this catalyst

at 80 °C. Owing to this fact a lower value of propagation rate constant ( k p ) is obtained at 80 °C

than that at 40 °C. The model is able to capture the experimental rate profile adequately well at

Al/Hf mole ratio of 2000 with the kinetic parameters determined at a ratio of 500.

Figure 4.58 shows that active catalyst site concentration decreases to negligible value

within 2 minutes at 80 °C, on the contrary, at 40 °C, active catalyst site concentration decreases

slowly. This fact is also evident from very high k in value at 80 °C (Table 4.16). Despite a

decrease in spontaneous deactivation frequency of active catalyst with temperature, extremely

high initiation rate at 80 °C ( kin .[M ] = 1.018×10-3 vs. 4.08×10-5 at 40 °C ) copiously contribute

for overall rapid decrease in active catalyst site concentration. For the same reasons, fractional

disappearance of active sites by spontaneous deactivation is found to be 93.8% at 40 °C and 16.4

% at 80 °C.

Maximum concentration of hydride activated complex P 0


*
H is obtained to be

7.0652×10-3 μM (in 33.5 minutes) at 40 °C and 1.9034 μM (in 17 minutes) at 80 °C as shown in

Figure 4.59 and Figure 4.60 respectively. With P4/MAO catalyst system, β-H elimination is

observed to occur at a higher frequency when compared to other catalyst systems discussed so

far and following the previous trends, increases with increase in temperature.

Reinitiation rate after β-H elimination is higher at 40 °C, which is evident from the

falling concentration of hydride activated complex after reaching a maximum. At 80 °C, due to

 
lower reinitiation rate, PH* 0 is almost stabilized after maximum.

168
Table 4.16 Estimated Parameters for Et(Ind)2HfCl2 (P4)/MAO

T (°C) 40 80

k in ×104 (M-1.s-1) 2.8629 211.0300

k p ×10-3 (M-1.s-1) 3.4717 28.5890

k d ×104 (s-1) 2.0049 6.1684

k tM ×104 (M-1.s-1) 3.4470 5.3048

k  , H ×105 (s-1) 3.0296 127.7600

k r ×103 (M-1.s-1) 1.6685 7.7346

k s ×109 (M-1.s-1) 1.2524 37.0180

k sp ×105 (M-1.s-1) 5.6388 17.9600

k sM ×103 (M-1.s-1) 5.6042 10.6520

k t , Al (M-1.s-1) 177.79 34.7210

k rAl ×10-3 (M-1.s-1) 2.4626 3.3866

F(k) (-) 1.3866 0.5143

Table 4.17 Predicted Properties with Et(Ind)2HfCl2 (P4)/MAO

T (°C) 40 80
Al/Zr 500 2000 500 2000
Exp Model Exp Model Exp Model Exp Model
M n ×10-8 (g/mol) - 8.2808 - 8.0613 - 1.3440 - 0.5421

M w ×10-8 (g/mol) 22.1 20.8312 17.7 16.1219 2.9 2.6891 2.0 1.0805
PDI 1.83 2.5156 1.85 1.9999 2.78 2.0008 2.18 1.9931
f v (%) - 81.20 - 80.50 - 86.10 - 83.90
f b (%) - 6.30 - 5.40 - 4.60 - 3.90
f i (%) - 12.50 - 14.10 - 9.30 - 12.2

169
12

T
40 °C

Active catalyst sites (moles/L)


80 °C
9

0 10 20 30 40 50 60
Time (minutes)

Figure 4.58 Active catalyst site concentration vs. time.


[Catalyst (P4) = 10 μM, Al/Hf = 500 and P = 30 psi]

-3
8.0x10
Hydride activated complex (moles/L)

-3
6.0x10

-3
4.0x10

-3
2.0x10

0.0

0 10 20 30 40 50 60
Time (minutes)

Figure 4.59 Hydride actived complex concentration vs. time.


[Catalyst (P4) = 10 μM, Al/Hf = 500, T = 40 °C and P = 30 psi]

170
2.0

Hydride activated complex (moles/L)


1.5

1.0

0.5

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.60 Hydride actived complex concentration vs. time.


[Catalyst (P4) = 10 μM, Al/Hf = 500, T = 80 °C and P = 30 psi]

With P4/MAO system, chain transfer to cocatalyst is obtained very high at 40 °C than

*
that at 80 °C. Concentration of methyl activated complex PMe 0 is found to reach a maximum  
of 0.4011 μM (17.3 min) followed by a steep decrease at 40 °C as compared to a maximum of

1.8862×10-5 μM (1.4 min) at 80 °C followed by slow decrease as shown in Figure 4.61 and

Figure 4.62 respectively. Due to low concentrations of monomer and methyl activated complex

at 80 °C, the rectivation rate, after transfer to cocatalyst is slow and therefore the concentration

*
of PMe 0 is decreasing slowly. At 40 °C, high concentration of methyl activated complex
*
enhances the reactivation rate for which a rapid decrease in PMe 0 is observed after reaching a  
maximum.

171
Methyl activated complex (moles/L)
0.4

0.2

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.61 Methyl actived complex concentration vs. time.


[Catalyst (P4) = 10 μM, Al/Hf = 500, T = 40 °C and P = 30 psi]

-5
2.0x10
Methyl activated complex (moles/L)

-5
1.5x10

-5
1.0x10

-6
5.0x10

0.0

0 10 20 30 40 50 60
Time (minutes)

Figure 4.62 Methyl actived complex concentration vs. time.


[Catalyst (P4) = 10 μM, Al/Hf = 500, T = 80 °C and P = 30 psi]

172
For P4/MAO system, model predicted molecular weight distribution and percentage of vinyl-,

butenyl- and isobutyl-terminated chains are given in Table 4.17. A good match between

experimental and predicted results for MWD is obtained.

Polymerization rates obtained with this catalyst system are comparable with P3/MAO but

very less when compared with P1/MAO and P2/MAO. P4/MAO yielded very high molecular

weights (~ 108 g/mol) of polypropylene as compared to its zirconium analogue (P2) as well as

other catalysts (P1 and P3) discussed heretofore.

Molecular weight is found to decrease with increase in temperature (from 40 °C to 80 °C)

and also with increase in Al/Hf molar ratio (from 500 to 2000). At 80 °C, chain termination via

β-H elimination is highly dominating followed by the transfer to monomer, than that at 40 °C,

which causes a significant drop in molecular weight at higher temperature. Polydispersity index

predicted by the model for Al/Hf ratio of 500 at 40 °C is 2.5156, whereas for all other reaction

conditions it is around 2.0.

At all the temperatures and Al/Hf ratios considered, the highest percentage of terminated

chains hold vinylidene end group ( f v > 80%), followed by isobutyl end group (9.3 < f i < 14.1)

and butenyl end group (3.9 < f b < 6.3) as shown in Table 4.17. At a fixed Al/Hf molar ratio,

with increase in temperature, f v is increasing whilst f b and f i are found decreasing. Increase in

Al/Hf ratio increases the rate of chain transfer to cocatalyst and thereby f i increases at a given

temperature.

Though these trends for P4/MAO are consistent with those catalyst systems discussed

earlier but all the termination rates are notably less, due to which high molecular weight product

is raised.

173
Effect of Pressure

Figure 4.63 and Figure 4.64 show the effect of monomer pressure on polymerization rate at 40

°C and 80 °C respectively with P4/MAO system at fixed Zr = 10 μM, Al/Zr = 500. For a change

in monomer pressure of 30 psi to 45 psi, at 40 °C, Rp,max increases from 0.1720 mol/L/s to 0.3186

mol/L/s and at 80 °C, Rp,max increases from 0.07838 mol/L/s to 0.1176 mol/L/s. This indicates

that an increase in monomer pressure results in linear increase in polymerization rate. As

observed experimentally and discussed earlier, this catalyst system affords lower polymerization

rates at 80 °C with an Al/Hf molar ratio of 500.

With increase in pressure, molecular weights are found to increase. 2.92 fold increase in

M w for a change in pressure from 15 to 30 psi and 1.86 fold increase for a change from 30 to 45

psi is found at 40 °C (Figure 4.65). Similarly, at 80 °C, 1.94 fold increase in M w for a change in

pressure from 15 to 30 psi and 2.82 fold increase for a change from 30 to 45 psi is observed

(Figure 4.66).

Effect of catalyst concentration

Polymerization rates at different catalyst concentrations (10, 20, 30, 40 and 50 μM) at 40 °C and

80 °C with P4/MAO system are shown in Figure 4.67 and Figure 4.68 respectively.

Polymerization rate is linearly increasing with increase in catalyst concentration at both the

temperatures and in tune with the discussion for P3/MAO system.

With P4/MAO catalyst system also, an inverse relationship between polymer molecular

weight and catalyst concentration is obtained as shown in Figure 4.69 and Figure 4.70.

174
0.35

0.30

Polymerization rate (moles/L/s)


0.25

45 psi
0.20

0.15 30 psi

0.10

15 psi
0.05

0.00
0 10 20 30 40 50 60
Time (minutes)

Figure 4.63 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P4) = 10 μM, Al/Hf = 500 and T = 40 °C]

0.15

45 psi
Polymerization rate (moles/L/s)

0.10

30 psi

0.05

15 psi

0.00
0 10 20 30 40 50 60
Time (minutes)

Figure 4.64 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P4) = 10 μM, Al/Hf = 500 and T = 80 °C]

175
9
4.5x10

Molecular weight (g/mol)


9
3.0x10
Mn
Mw

9
1.5x10

0.0
15 30 45
Pressure (psi)

Figure 4.65 Effect of pressure on average molecular weights.


[Catalyst (P4) = 10 μM, Al/Hf = 500 and T = 40 °C]

8
4.5x10
Molecular weight (g/mol)

8
3.0x10 Mn
Mw

8
1.5x10

0.0
15 30 45
Pressure (psi)

Figure 4.66 Effect of pressure on average molecular weights.


[Catalyst (P4) = 10 μM, Al/Hf = 500 and T = 80 °C]

176
At 40 °C, M w is decreasing from 2.0831 × 109 to 1.757 × 109 (15.6%) for a change in catalyst

concentration from 10 μM to 20 μM. On further increase in catalyst concentration (30, 40 and 50

μM), M w is not observed to decrease much. This outcome again emphasizes that at this

temperature, Al/Hf ratio of 500 is not sufficient to activate catalyst efficaciously and therfore

increase in catalyst concentration does not really increase the number of active chains. On the

contrary, at 80 °C M w is found to decrease with each increment in catalyst concentration. As

shown in Figure 4.70, for every 10 μM increase in catalyst concentration, 44% (10-20 μM),

28.1% (20-30 μM), 19.95% (30-40 μM) and 17.94% (40-50 μM) decrease in M w is obtained.

1.00

[Zr]
Polymerization rate (moles/L/s)

0.75
50M

40M
0.50
30M

20M
0.25

10M

0.00
0 10 20 30 40 50 60
Time (minutes)

Figure 4.67 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Hf = 500, T = 40 °C and P = 30 psi]

177
[Zr]
0.4
50M

Polymerization rate (moles/L/s)


0.3 40M

30M
0.2

20M

0.1
10M

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.68 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Hf = 500, T = 80 °C and P = 30 psi]

9
2.5x10

9
2.0x10
Molecular weight (g/mol)

9
1.5x10

Mnbar
Mwbar
9
1.0x10

8
5.0x10

0.0
0 10 20 30 40 50
[Hf] (M)

Figure 4.69 Effect of catalyst concentration on average molecular weights.


[Al/Hf = 500, T = 40 °C and P = 30 psi]

178
8
3.0x10

Mnbar
Mwbar

Molecular weight (g/mol) 8


2.0x10

8
1.0x10

0.0
0 10 20 30 40 50
[Hf] (M)

Figure 4.70 Effect of catalyst concentration on average molecular weights.


[Al/Hf = 500, T = 80 °C and P = 30 psi]

4.2.5 Propylene polymerization with (2,4,6-Me3Ind)2ZrCl2 (P5)/MAO

Propylene polymerization model is applied to solution phase polymerization of propylene with

an ansa-metallocene catalyst and kinetic parameters are obtained. Data for the model validation

were taken from Yasin et al. (2004).

Estimated parameters and effect of Al/Zr mole ratio

Model predicted and experimental polymerization rate profiles for propylene polymerization

with [2,4,6-Me3Ind]2ZrCl2(P5)/MAO catalyst system at Al/Zr molar ratios of 2000 and 4000 are

presented in Figure 4.71. Kinetic parameters are estimated at 0 °C with experimental rate data at

0.98 atm pressure, 20 μM catalyst and Al/Zr molar ratio of 2000. Experimental rate data at Al/Zr

mole ratio of 4000 are employed to verify the model prediction for polymerization rate. Model

179
predictions of polymerization rate are in good agreement with the experimental polymerization

rate at both the cocatalyst to catalyst mole ratios considered. Values of estimated kinetic

parameters and objective function F(k) are given in Table 4.18.

When compared with P1-P4/MAO systems, polymerization rates with P5/MAO system are

found very less, so is the k p value estimated by the model. At 0 °C, Low polymerization rates

(Rp,max = 2.593 × 10-3 moles/L/s) at Al/Zr mole ratio of 2000 and high rates (Rp,max = 4.921× 10-3

moles/L/s) at Al/Zr mole ratio of 4000 are echoing the trend obtained with other catalyst

systems. Also a sharp decrease in polymerization rate after reaching a maximum is seen at high

Al/Zr ratio implying a larger frequency of chain transfer to cocatalyst.

-3
6.0x10

Zr:MAO
Polymerization rate (moles/L/s)

1:2000
-3
1:4000
4.0x10

-3
2.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.71 Effect of Al/Zr mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P5) = 20 μM, T = 0 °C and P = 0.98 atm]

180
Table 4.18 Estimated Parameters for [2,4,6-Me3Ind]2ZrCl2 (P5)/MAO

T 0 °C
k in (M-1.s-1) 7.3213 × 10-3

k p (M-1.s-1) 71.2665

k d (s-1) 3.8988 × 10-4

k tM (M-1.s-1) 2.6484

k  , H (s-1) 8.2268 × 10-3

k r (M-1.s-1) 628.343

k s (M-1.s-1) 8.8514 × 10-4

k sp (M-1.s-1) 1.6957 × 10-4

k sM (M-1.s-1) 0.12125
3
k t , Al (M-1.s-1) 7.9402 × 10

k rAl (M-1.s-1) 32.4558

F(k) (-) 0.27301

Table 4.19 Predicted Properties with [2,4,6-Me3Ind]2ZrCl2 (P5)/MAO

T (°C) 0
Al/Zr 2000 4000
Exp Model Exp Model
4 4 4 4
M n (g/mol) 3.02 × 10 2.6715 × 10 2.71 × 10 2.3537 × 10
M w (g/mol) 7.7 × 104 6.2516 × 104 6.77 × 104 5.4211 × 104

PDI 2.55 2.34 2.50 2.30


f v (%) - 81.2671 - 76.4243

f b (%) - 4.4303 - 4.8034

f i (%) - 14.3026 - 18.7723

181
Active catalyst site concentration decreases to 0.03 μM from 20 μM in 8.6 minutes as shown in

Figure 4.72. Further, the values, k in .[M ] = 1.2792×10-2 s-1 vs. k d = 3.8988 × 10-4 s-1 suggest

that 2.96 % active sites are deactivating spontaneously and rest are utilized to initiate the chains.

With [2,4,6-Me3Ind]2ZrCl2 (P5)/MAO system, frequency of β-H elimination is found to

be 8.2268 × 10-3 (Table 4.16). Concentration of hydride activated complex reaches a maximum

of 1.4165×10-4 μM rapidly (in 4.6 minutes) and decrease thereafter as shown in Figure 4.73.

Fast reinitiation rate after β-H elimination ( k r = 628.343) causes the decrease in PH* 0 .  
Figure 4.74 shows that the concentration of methyl activated complex reaches a

maximum of 8.5575 μM within 1.37 minutes followed by a sharp decrease to a negligible value.

The rate of chain transfer to cocatalyst and reactivation after transfer both are quite high with this

catalyst system at 0 °C.

25
Active catalyst concentration (moles/L)

20

15

10

0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.72 Active catalyst site concentration vs. time.


[Catalyst (P5) = 20 μM, Al/Zr = 2000, T = 0 °C and P = 0.98 atm]
182
-4
1.5x10

Hydride actived catalyst concentration (moles/L)


-4
1.0x10

-5
5.0x10

0.0

0 10 20 30 40 50 60
Time (minutes)

Figure 4.73 Hydride actived complex concentration vs. time.


[Catalyst (P5) = 20 μM, Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

9
Methyl actived catalyst concentration

6
(moles/L)

0 10 20 30 40 50 60
Time (minutes)

Figure 4.74 Methyl actived complex concentration vs. time.


[Catalyst (P5) = 20 μM, Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

183
Table 4.19 presents the molecular weights of polypropylene and percentage of variously

terminated chains predicted by the model for [2,4,6-Me3Ind]2ZrCl2 (P5)/MAO system at 0 °C. At

Al/Zr mole ratio of 2000, M w is found to be 6.2516 × 104 which is decreased by 13.3% on

raising the ratio to 4000. Model predicted molecular weights and PDIs closely match with the

experimentally determined values. Predicted value of PDI is almost unchanged with an increase

in Al/Zr mole ratio, and is found to be 2.34 and 2.30 at an Al/Zr molar ratio of 2000 and 4000

respectively.

Chain termination is via spontaneous catalyst deactivation, transfer to monomer and β-Hydride

elimination remain in force followed by chain transfer to cocatalyst as inferred by f v and f i

values obtained (Table 4.19). With increase in Al/Zr ratio, f i is found to increase while f v

decreases, showing higher transfer to cocatalyst at higher ratio. Secondary insertions are found to

be less with this catalyst system and unresponsive to Al/Zr ratio as evident from small

percentage (4.43 - 4.8 %) of butenyl end groups obtained (Table 4.19).

Effect of Pressure

Monomer pressure is doubled and tripled from 0.98 atm (experimental) and polymerization rates

are prognosticated with the model as given in Figure 4.75. The maximum rate observed at 0.98

atm, 1.97 atm and 2.96 atm are 0.0022 moles/L/s, 0.0061 moles/L/s and 0.0133 moles/L/s

respectively at 0 °C and fixed Zr = 20 μM, Al/Zr = 2000. With increase in monomer pressure, a

higher initiation and polymerization rates are noted, which is consistent with trend observed

earlier with any catalyst system.

Figure 4.76 depicts that no appreciable increase in molecular weight is obtained with increase in

monomer pressure with this catalyst system. 0.9% and 0.4% increase in weight average

184
molecular weight is recorded for a change of 0.98 to 1.97 atm and from 1.97 to 2.96 atm in

monomer pressure. High monomer concentration not only increases the initiation and

propagation rates but also raises various chain termination rates in which monomer is involved.

Therefore eventually for this catalyst system, despite high polymerization rate no considerable

change in molecular weight is observed.

-2
1.5x10
Polymerization rate (moles/L/s)

-2
1.0x10

-3
5.0x10
2.96 atm

1.97 atm
0.98 atm
0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.75 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P5) = 20 μM, Al/Zr = 2000 and T = 0 °C]

185
4
8x10

Molecular weight (g/mol)


4
6x10
Mn
Mw

4
3x10

0
1 2 3
Pressure (atm)

Figure 4.76 Effect of pressure on average molecular weights.


[Catalyst (P5) = 20 μM, Al/Zr = 2000 and T = 0 °C]

Effect of catalyst concentration

Polymerization rates at different catalyst concentrations (10, 20, 40 and 80 μM) at 0 °C are

shown in Figure 4.77. Increase in catalyst concentration is resulting in increasead polymerization

rate which reaches to a maximum and decreases afterwards due to elevated termination rates. On

doubling the catalyst concentration, maximum polymerization rate is found to be doubled,

showing a linear proportional dependence. The trend is similar to those obtained with other

catalyst systems discussed earlier.

186
-2
1.00x10

Polymerization rate (moles/L/s)


-3
7.50x10

-3
5.00x10

[Zr]

80M
-3
2.50x10

40M
20M
10M
0.00
0 10 20 30 40 50 60
Time (minutes)

Figure 4.77 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

4
7x10
Molecular weight (g/mol)

4
Mn
4x10
Mw

4
2x10

0 20 40 60 80
[Zr] (M)

Figure 4.78 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

187
With P5/MAO catalyst system, average molecular weights are noticed to be unaltered with

increase in catalyst concentration as shown in Figure 4.78. An inverse relationship between

polymer molecular weight and catalyst concentration is obtained with hafnium based

metallocenes (P3 and P4), whereas Zr based metallocenes (P1 and P2) have shown a similar

trend as obtained for P5.

4.2.6 Propylene polymerization with [2,4,7-Me3Ind]2ZrCl2) (P6)/MAO

Propylene polymerization model is applied to solution phase polymerization of propylene with

an ansa-metallocene catalyst and kinetic parameters are obtained. Data for the model validation

were taken from Yasin et al. (2004).

Estimated parameters and effect of Al/Zr mole ratio

Figure 4.79 shows the model prediction for the experimental behaviour of [2,4,7-Me3Ind]2ZrCl2

catalyst in propylene polymerization for different Zr/MAO ratios. Experimental polymerization

rate data at Al/Zr molar ratio of 2000 are regressed to determine the kinetic parameters at 0 °C

and other sets of rate data are used to verify the simulation results. A good match between the

model predicted and experimental rate is obtained at all the Al/Zr ratios studied. Values of

estimated kinetic parameters and objective function [F(k)] are given in Table 4.20.

188
-3
3.0x10
Zr:MAO

Polymerization rate (moles/L/s)


1:1000
1:2000
1:4000
-3
2.0x10

-3
1.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.79 Effect of Al/Zr mole ratio on propylene polymerization rate; solid lines are
model predictions.
[Catalyst (P6) = 20 μM, T = 0 °C and P = 0.98 atm]

Polymerization rates with P6/MAO are observed very less when compared with P1-

P4/MAO systems, but comparable with P5/MAO system at identical reaction conditions. The

k p value estimated by the model is slightly higher (106.06 M-1.s-1 vs. 71.27 M-1.s-1) than

P5/MAO system. At Al/Zr mole ratio of 1000, polymerization rate is all less (Rp,max = 1.24 × 10-3

moles/L/s) than those observed at higher ratios. At Al/Zr ratio of 2000, polymerization rate is

found to reach a maximum of 2.9389 × 10-3 moles/L/s, but with further increase in Al/Zr ratio to

4000 polymerization rate is not observed to increase and yields a maximum rate of 2.6332 × 10-3

moles/L/s. This suggests that Al/Zr ratio as high as 4000 is sufficient to activate this catalyst and

higher ratios would bring no further increase in polymerization rate.

189
Table 4.20 Estimated Parameters for [2,4,7-Me3Ind]2ZrCl2) (P6)/MAO

T 0 °C
k in (M-1.s-1) 6.2848 × 10-3

k p (M-1.s-1) 1.0606 × 102

k d (s-1) 3.4959 × 10-4

k tM (M-1.s-1) 1.0114

k  , H (s-1) 9.8943 × 10-4

k r (M-1.s-1) 3.1658 × 102

k s (M-1.s-1) 1.0107 × 10-3

k sp (M-1.s-1) 2.0210 × 10-5

k sM (M-1.s-1) 0.2694
2
k t , Al (M-1.s-1) 5.4686 × 10

k rAl (M-1.s-1) 4.2644

F(k) (-) 0.92561

Table 4.21 Predicted Properties with [2,4,7-Me3Ind]2ZrCl2) (P6)/MAO

T 0 °C
Al/Zr 1000 2000 4000
Exp Model Exp Model Exp Model
M n (g/mol) - 4.9605 × 104 - 2.8396 × 10 4
- 2.9229 × 104

M w (g/mol) - 1.2994 × 105 - 7.8196 × 104 - 6.1089 × 104


PDI - 2.62 - 2.75 - 2.09
f v (%) - 86.4288 - 83.4627 - 77.8977
f b (%) - 6.2102 - 6.3333 - 6.4323
f i (%) - 7.361 - 10.204 - 15.670

190
Active catalyst site concentration decreases from 20 μM to 0.0522 μM in 9.3 minutes as shown

in Figure 4.80. Further, the values, k in .[M ] = 2.1687×10-3 s-1 vs. k d = 3.4959 × 10-4 s-1 suggest

that 13.88 % active sites are deactivating spontaneously while rest are utilized to initiate the

chains.

With [2,4,7-Me3Ind]2ZrCl2 (P6)/MAO system, frequency of β-H elimination is found to

be 9.8943 × 10-4 (Table 4.18). Concentration of hydride activated complex reaches a maximum

of 3.0554×10-5 μM in 13 minutes and decrease thenceforth as shown in Figure 4.81. Due to

 
rapid reinitiation rate after β-H elimination ( k r =316.58 M-1.s-1), PH* 0 decreases after

reaching the maximum.

The rate of chain transfer to cocatalyst and reactivation after transfer both are quite high

with P6/MAO catalyst system at 0 °C. k t , Al and k rAl values with this catalyst system are much

higher than those obtained with (P1-P4)/MAO but less than that obtained with P5/MAO system.

Figure 4.82 points that the concentration of methyl activated complex reaches a maximum of

11.84 μM within 3.58 minutes followed by a sharp decrease to a negligible value.

Table 4.21 gives the molecular weights of polypropylene and percentage of variously

terminated chains predicted by the model for [2,4,7-Me3Ind]2ZrCl2 (P6)/MAO system at 0 °C

and different Al/Zr ratios.

191
20

Active catalyst concentration


15

(moles/L)
10

0 10 20 30 40 50 60
Time (minutes)

Figure 4.80 Active catalyst site concentration vs. time.


[Catalyst (P6) = 20 μM, Al/Zr = 2000, T = 0 °C and P = 0.98 atm]
Hydride activated complex concentration

-5
3.0x10

-5
(moles/L)

2.0x10

-5
1.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.81 Hydride actived complex concentration vs. time.


[Catalyst (P6) = 20 μM, Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

192
12

Methyl activated complex concentration


9

(mol/L)
6

0 10 20 30 40 50 60
Time (minutes)

Figure 4.82 Methyl actived complex concentration vs. time.


[Catalyst (P6) = 20 μM, Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

Increase in Al/Zr molar ratio is resulting in decrease in molecular weights. At Al/Zr mole

ratio of 1000, M w is found to be 1.2994 × 105 which is decreased by 39.82% on raising the ratio

to 2000. On increasing Al/Zr ratio from 2000 to 4000, M w is further decreased by 21.88%.

Weight average molecular weights obtained with P6/MAO are very close to those obtained with

P5/MAO catalyst system at identical conditions, however, polydispersity indices are little high

with P6/MAO evincing broader molecular weight distribution.

f v values in Table 4.21 indicate that chain termination is majorly via spontaneous

catalyst deactivation, transfer to monomer and β-Hydride elimination. With increase in Al/Zr

ratio, f i is found increasing while f v decreasing, exhibiting the trends similar to that discovered

with P5/MAO. Secondary insertions with P6/MAO catalyst system are higher than those with

193
P5/MAO but less in comparison with (P1-P4)/MAO systems. Percentage of butenyl end groups

are almost unchanged with variation in Al/Zr ratio.

Effect of Pressure

Like P5/MAO and other catalysts studied up to now, P6/MAO also evidences a higher initiation

rate and increased polymerization rate with increase in monomer pressure. Polymerization rates

are predicted with the model at 0.98, 1.97 and 2.96 atm monomer pressures as shown in Figure

4.83. The maximum rates observed at 0.98 atm, 1.97 atm and 2.96 atm are 0.0030 moles/L/s,

0.0081 moles/L/s and 0.0179 moles/L/s respectively at 0 °C and fixed Zr = 20 μM, Al/Zr = 2000.

Polymerization rates at studied monomer pressures are very comparable with those obtained with

P1/MAO system.

Figure 4.84 shows that a little increase in molecular weight is obtained with increase in monomer

pressure with P6/MAO catalyst system. In weight average molecular weight, a 5.18% and 2.97%

increase is seen respectively for a change of 0.98 to 1.97 atm and from 1.97 to 2.96 atm in

monomer pressure. The small increase in M w conforms to the effect of pressure discussed for

P5/MAO system.

194
-2
1.8x10

Polymerization rate (moles/L/s) 1.2x10


-2

2.96 atm
-3
6.0x10

1.97 atm

0.98 atm
0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.83 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P6) = 20 μM, Al/Zr = 2000 and T = 0 °C]

5
1x10

4
8x10
Molecular weight (g/mol)

Mn
Mw
4
5x10

4
3x10

0
1 2 3
Pressure (atm)

Figure 4.84 Effect of pressure on average molecular weights.


[Catalyst (P6) = 20 μM, Al/Zr = 2000 and T = 0 °C]

195
Effect of catalyst concentration

With P6/MAO also, polymerization rates are increasing linearly with increase in catalyst

concentration. Figure 4.85 shows polymerization rates at different catalyst concentrations (10,

20, 40 and 80 μM) at 0 °C. Since high initiation rates prevail at high catalyst concentrations,

maximum rate is achieved shortly at higher catalyst concentration.

Like P5/MAO catalyst system, average molecular weights with P6/MAO are also not changing

with increase in catalyst concentration as shown in Figure 4.86.

-2
1.5x10
Polymerization rate (moles/L/s)

-2
1.0x10

Zr
-3
5.0x10
80M

40M
20M

0.0 10 M
0 10 20 30 40 50 60
Time (minutes)

Figure 4.85 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

196
5
1x10

4
8x10

Molecular weight (g/mol)


4
6x10 Mn
Mw

4
4x10

4
2x10

0 20 40 60 80
[Zr] (M)

Figure 4.86 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 2000, T = 0 °C and P = 0.98 atm]

4.2.7 Propylene polymerization with Me2Si[2,4,6-Me3Ind]2ZrCl2) (P7)/MAO

Propylene polymerization model is applied to solution phase polymerization of propylene with

an ansa-metallocene catalyst and kinetic parameters are obtained. Simulations are carried out

using natural logarithmic differential evolution and the experimental data for the model

validation are taken from Yasin et al. (2005).

Estimated parameters and effect of temperature

Experimental and model predicted polymerization rate profiles for propylene polymerization

with Me2Si[2,4,6-Me3Ind]2ZrCl2(P7)/MAO catalyst system at 30 °C, 50 °C and 70 °C are

presented in Figure 4.87. Kinetic parameters are estimated at different temperatures with

197
experimental rate data at 0.98 atm pressure, 20 μM catalyst and Al/Zr molar ratio of 2000.

Estimated kinetic parameters and objective function F(k) values are given in Table 4.22.

A decent agreement between experimental and model predicted polymerization rates is

received at 30 °C and 70 °C. At 50 °C from 2.75 minutes to 12 minutes, very few experimental

data points are available and a large number of data is available for regression only after 12

minutes. For this reason, model is found to under predict the polymerization rate in this time

range which includes the peak corresponding to maximum polymerization rate. Due to possible

unreliableness, model outputs at 50 °C are excluded in further discussion.

Figure 4.87 shows that model predicted maximum polymerization rates at 30 °C and 70

°C are 7.103 × 10-3 (experimental: 8.38 × 10-3) and 8.747.103 × 10-3 (experimental: 8.95 × 10-3)

respectively which explicates that polymerization rate with P7/MAO catalyst system is increased

to a little extent with increase in temperature. Despite an increase in k p value with increase in

temperature, the low increase in polymerization rate may be explained with the decreased

solubility of propylene in toluene at higher temperature at fixed monomer pressure. Temperature

is rather found to affect initiation rates, which are increasing with increase in temperature and

thereby achieve maximum rate earlier. Chain termination rates via different routes are

sufficiently high at all the temperatures to induce a descending polymerization rate profile after a

maximum.

Both initiation rate ( kin .[M ] = 2.441×10-2 s-1 vs. 4.849×10-3 s-1 at 30 °C) and frequency of

spontaneous catalyst deactivation ( kd = 4.0949×10-4 s-1 vs. 6.1255×10-5 s-1 at 30 °C) are high at

70 °C. That's why active catalyst site concentration decreases to a negligible value within 3.5

minutes at 70 °C, whereas at 30 °C, it takes 19.8 minutes as shown in Figure 4.88.

198
-2
1.0x10

0
30 C
Polymerization rate (moles/L/s)

-3
8.0x10
0
50 C
0
70 C
-3
6.0x10

-3
4.0x10

-3
2.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.87 Effect of temperature on propylene polymerization rate; solid lines are model
predictions.
[Catalyst (P7) = 20 μM, Al/Zr = 2000 and P = 0.98 atm]

199
Table 4.22 Estimated Parameters for Me2Si[2,4,6-Me3Ind]2ZrCl2) (P7)/MAO

T (°C) 30 70
-3 -2
k in (M-1.s-1) 8.075708 × 10 9.985286 × 10
2
k p (M-1.s-1) 9.361976 × 10 2.211042 × 103

kd (s-1) 6.125482 × 10-5 4.094922 × 10-4

k tM (M-1.s-1) 0.2562587 12.5457

k  , H (s-1) 1.955819 × 10-4 2.307301 × 10-3

k r (M-1.s-1) 101.5226 407.3533

ks (M-1.s-1) 1.747163 × 10-3 1.255236 × 10-2


-4 -2
k sp (M-1.s-1) 1.264525 × 10 1.523233 × 10
-3
k sM (M-1.s-1) 1.33222 × 10 1.001232 × 103

F(k) (-) 0.5969 1.0339

Table 4.23 Predicted Properties with Me2Si[2,4,6-Me3Ind]2ZrCl2) (P7)/MAO

T (°C) 30 70
Exp Model Exp Model
4
M n (g/mol) - 3.9672 × 10 - 7.9313 × 103

M w (g/mol) - 7.9344 × 104 - 1.5823 × 104


PDI (-) - 2.000 - 1.995
f v (%) - 99.8648 - 99.9097
f b (%) 0.37 0.1352 0.10 0.0903

200
Fractional disappearance of active sites by spontaneous deactivation is as low as 1.24% at 30 °C

and 1.6% at 70 °C.

20 T
0
30 C
0
70 C
Active catalyst sites (moles/L)

15

10

0 10 20 30 40 50 60
Time (minutes)

Figure 4.88 Active catalyst site concentration vs. time.


[Catalyst (P7) = 20 μM, Al/Zr = 2000 and P = 0.98 atm]

Concentration of hydride activated complex PH* 0 is found to increase to a maximum  


 
followed by a decrease. Maximum PH* 0 achieved is 4.112×10-5 μM (in 6.4 minutes) at 30 °C

and 3.367 μM (in 5.85 minutes) at 70 °C as shown in Figure 4.89 (a) and (b) respectively. With

P7/MAO catalyst system, frequency of β-H elimination is substantial and increases with increase

in temperature from 30 °C to 70 °C. Reinitiation rate after β-H elimination is much lower at 70

°C than that at 30 °C due to which a high value of maximum concentration of hydride activated

complex is obtained.

201
-5
4.0x10

Hydride activated catalyst sites


(moles/L)

-5
2.0x10

0.0
0 10 20 30 40 50 60
Time (minutes)

(a)

3
Hydride activated catalyst sites

2
(moles/L)

0
0 10 20 30 40 50 60
Time (minutes)

(b)

Figure 4.89 Hydride actived complex concentration vs. time. (a) 30 °C, (b) 70 °C
[Catalyst (P7) = 20 μM, Al/Zr = 2000 and P = 0.98 atm]

202
Model predicted properties of polypropylene synthesized with Me2Si[2,4,6-Me3Ind]2ZrCl2)

(P7)/MAO are given in Table 4.23. Molecular weights are found to decrease with increase in

temperature. M w estimated at 30 °C is 7.9344 × 104 which is decreased by 80% at 70 °C. PDIs

at all reaction conditions are obtained almost 2.0 representing standard molecular weight

distribution.

Polypropylene synthesized using P7/MAO system is highly isotactic. Experimental

values of f b , which are closely predicted by the model also (Table 4.23) suggest that the fraction

of dead chains with butenyl end group is very less (< 0.5%) speculating high isotacticity. Since

transfer to cocatalyst is not considered in the model applied here, high f v values account for all

other types of chain termination i.e. via spontaneous catalyst deactivation, transfer to monomer

and β-Hydride elimination.

Effect of Pressure

Monomer pressure is varied from 0.98 atm to 1.97 atm and 2.96 atm to study the effect on

polymerization rate and polymer molecular weight at fixed Zr = 20 μM, Al/Zr = 2000. Figure

4.90 and Figure 4.91 depict the rate profiles at 30 °C and 70 °C respectively. Following the

earlier trends, polymerization rate is found increasing with increase in monomer pressure.

With increase in monomer pressure a linear increase in polymerization rate with a

maximum of 8.61142 × 10-3 moles/L/s (0.98 atm), 1.9996 × 10-2 moles/L/s (1.97 atm) and

3.3598 × 10-2 moles/L/s (2.96 atm) at 30 °C is observed as shown in Figure 4.90. At 70 °C also

a linear relationship between rate and monomer pressure is seen, with a maximum rate of 9.0595

× 10-3 moles/L/s (0.98 atm), 2.4396 × 10-2 moles/L/s (1.97 atm) and 4.0954 × 10-2 moles/L/s

(2.96 atm). Higher rates are predicted at 70 °C at corresponding monomer pressures.

203
No appreciable change in average molecular weights with variation in monomer pressures is

observed at both 30 °C (Figure 4.92) and 70 °C (Figure 4.93) temperature.

-2
3.0x10
Polymerization rate (moles/L/s)

-2
2.0x10

2.96 atm

-2
1.0x10
1.97 atm

0.98 atm

0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.90 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P7) = 20 μM, Al/Zr = 2000 and T = 30 °C]

-2
4.0x10
Polymerization rate (moles/L/s)

-2
3.0x10
2.96 atm

-2
2.0x10

1.97 atm
-2
1.0x10

0.98 atm
0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.91 Polymerization rate vs. time: Effect of pressure.


[Catalyst (P7) = 20 μM, Al/Zr = 2000 and T = 70 °C]
204
4
8.0x10

Molecular weight (g/mol) Mn


4
Mw
6.0x10

4
4.0x10

1 2 3
Pressure (atm)

Figure 4.92 Effect of pressure on average molecular weights.


[Catalyst (P7) = 20 μM, Al/Zr = 2000 and T = 30 °C]

4
2x10
Molecular weight (g/mol)

4
2x10
Mn
Mw

4
1x10

3
5x10
1 2 3
Pressure (atm)

Figure 4.93 Effect of pressure on average molecular weights.


[Catalyst (P7) = 20 μM, Al/Zr = 2000 and T = 70 °C]

205
Effect of catalyst concentration

Polymerization rates at different catalyst (P7) concentrations at 30 °C and 70 °C are shown in

Figure 4.94 and Figure 4.95 respectively. On doubling the catalyst concentration, predicted

maximum polymerization rate is found to be doubled showing a linear dependence at 30 °C as

well as at 70 °C. However at 70 °C, polymerization rate is noted to be 23% higher than that at 30

°C at corresponding catalyst concentration. All the termination reactions are highly activated at

70 °C, therefore polymerization rate is decreasing steeply after reaching a maximum at all

catalyst concentrations.

Frequency of β-H elimination is calculated to be negligibly low as compared to the chain transfer

to monomer ( k  , H vs. k tM M  in Table 4.22) at both the temperatures considered. For this

reason, average molecular weights are unchanged with increase in catalyst concentration as

shown in Figure 96 and Figure 97.


Polymerization rate (moles/L/s)

-2
3.0x10

-2
2.0x10
[Zr]
80M

-2
1.0x10
40M
20M

0.0 10M
0 10 20 30 40 50 60
Time (minutes)

Figure 4.94 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 2000, T = 30 °C and P = 0.98 atm]

206
-2
4.0x10

[Zr]

Polymerization rate (moles/L/s)


-2 80M
3.0x10

-2
2.0x10

40M

-2
1.0x10
20M
10M
0.0
0 10 20 30 40 50 60
Time (minutes)

Figure 4.95 Polymerization rate vs. time: Effect of catalyst concentration.


[Al/Zr = 2000, T = 70 °C and P = 0.98 atm]

4
8.0x10
Molecular weight (g/mol)

Mn
4
6.0x10 Mw

4
4.0x10

0 20 40 60 80
[Zr] (M)

Figure 4.96 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 2000, T = 30 °C and P = 0.98 atm]

207
4
2x10

Molecular weight (g/mol)


4
2x10

Mn
Mw

4
1x10

3
5x10
0 30 60 90
[Zr] (M)

Figure 4.97 Effect of catalyst concentration on average molecular weights.


[Al/Zr = 2000, T = 70 °C and P = 0.98 atm]

Summary of the chapter: Models developed for ethylene and propylene polymerization in

previous chapter have simulated and results were discussed in this chapter. Gas phase and

solution phase ethylene polymerization with silica-supported, bridged zirconocene catalysts were

discussed first. Solution phase propylene polymerization with various bridged/unbridged

zirconocene and hafnocene catalysts were discussed in ulterior sections. Kinetic parameters were

determined for each catalyst system with 'natural logarithmic differential evolution' approach of

optimization.

208

You might also like