Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/342103287

Kinetics and mechanism of ethylene and propylene polymerizations catalyzed


with ansa-zirconocene activated by borate/TIBA

Article  in  Journal of Organometallic Chemistry · June 2020


DOI: 10.1016/j.jorganchem.2020.121366

CITATIONS READS

6 117

10 authors, including:

Amjad Ali Yintian Guo


Zhejiang University Zhejiang Research Institute of Chemical Industry Co. LTD.
18 PUBLICATIONS   51 CITATIONS    17 PUBLICATIONS   176 CITATIONS   

SEE PROFILE SEE PROFILE

Muhammad Adnan Akram Akbar Khan


Zhejiang University Zhejiang University
5 PUBLICATIONS   29 CITATIONS    10 PUBLICATIONS   80 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Active centers in the initial stage of ethylene-propylene copolymerization with MgCl2-supported Ziegler-Natta catalyst View project

Synthesis and Mechanistic Studies of olefin Polymerizations Using (α-Diimine)nickel(II) Catalysts: Influence of Temperature, Pressure, and Ligand Structure on Polymer
Properties View project

All content following this page was uploaded by Amjad Ali on 19 June 2020.

The user has requested enhancement of the downloaded file.


Journal Pre-proof

Kinetics and mechanism of ethylene and propylene polymerizations catalyzed with


ansa-zirconocene activated by borate/TIBA

Amjad Ali, Xiaoyu Liu, Yintian Guo, Muhammad Adnan Akram, Haifeng Wu, Wucan
Liu, Akbar Khan, Baiyu Jiang, Zhisheng Fu, Zhiqiang Fan

PII: S0022-328X(20)30268-0
DOI: https://doi.org/10.1016/j.jorganchem.2020.121366
Reference: JOM 121366

To appear in: Journal of Organometallic Chemistry

Received Date: 28 December 2019


Revised Date: 14 April 2020
Accepted Date: 2 June 2020

Please cite this article as: A. Ali, X. Liu, Y. Guo, M.A. Akram, H. Wu, W. Liu, A. Khan, B. Jiang, Z.
Fu, Z. Fan, Kinetics and mechanism of ethylene and propylene polymerizations catalyzed with ansa-
zirconocene activated by borate/TIBA, Journal of Organometallic Chemistry (2020), doi: https://
doi.org/10.1016/j.jorganchem.2020.121366.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


Kinetics and mechanism of ethylene and propylene polymerizations
catalyzed with ansa-zirconocene activated by borate/TIBA

Amjad Ali,1 Xiaoyu Liu,1 Yintian Guo,2* Muhammad Adnan Akram,1 Haifeng Wu,2,3 Wucan Liu,2,3
Akbar Khan,3 Baiyu Jiang, 1 Zhisheng Fu,1 Zhiqiang Fan1*

1 MOE Key Laboratory of Macromolecular Synthesis and Functionalization, Department of Polymer Science and
Engineering, Zhejiang University, Hangzhou 310027, China

2 Sinochem Lantian Zhejiang Research Institute of Chemical Industry Co., Ltd., Hangzhou, 310023, China

3 State Key Laboratory of Fluorinated Greenhouse Gases Replacement and Control Treatment, Hangzhou,
310023, China
4 Department of Polymer Engineering, National Textile University, Karachi Campus, ST-2/1, Sector 30, K.I.A,
74900, Karachi, Pakistan
*Corresponding author. E-Mail address: guoyintian@sinochem.com (Y. Guo), fanzq@zju.edu.cn (Z. Fan).

ABSTRACT: Ethylene and propylene polymerizations catalyzed with two ansa-zirconocenes (rac-Me2Si(2-
Me-4-Ph-Ind)2ZrCl2 (Mt-I) and rac-Et(Ind)2ZrCl2 (Mt-II)) activated by (Ph3C)B(C6F5)4/triisobutylaluminum
were respectively conducted under the same conditions for different duration ranging from 0.5 to 30 minutes,
and quenched with 2-thiophenecarbonyl chloride for determining the active center fraction ([C*]/[Zr]).
Variations of polymerization rate, molecular weight, isotacticity and thermal properties with time were also
studied. The [C*]/[Zr] fraction gradually increased with time in the first 20 min and then reached steady stage
in all polymerization systems. [C*]/[Zr] value in the steady stage of Mt-II catalyzed ethylene polymerization
was significantly higher than that of propylene polymerization, similar to the phenomena previously
observed in ethylene and propylene polymerizations catalyzed with Mt-II/methylaluminoxane. When Mt-
I/borate was used as the catalyst, both ethylene and propylene polymerization showed higher steady stage
[C*]/[Zr] levels than those of the Mt-II/borate systems, and the influence of zirconocene structure on
[C*]/[Zr] of propylene polymerization was much stronger than that of ethylene polymerization. Change in
zirconocene structure also significantly influenced apparent chain propagation rate constant (kp) and its time-
dependent variations. By taking into consideration the time-dependent changes of [C*]/[Zr], polymer’s
molecular weight distribution and chain structure, presence of multiple active centers in the polymerization
system and late activation of the centers with lower intrinsic reactivity are proposed to be the main reasons
for steep decay of kp with time. Structure-performance relationships of the ansa-zirconocene catalysts have
also been discussed based on the results of kinetic study.

1
Keywords: Zirconocene; Borate; Ethylene; Propylene; Polymerization; Kinetics

1. Introduction
In the past three decades, metallocene catalysts have made enormous impact on both fundamental
researches of catalytic olefin polymerization and polyolefin industry [1-7]. Kinetic and mechanistic studies
on metallocene catalyzed polymerizations have been highly emphasized since the 1990’s, with the aim of
obtaining scientific knowledge for optimizing the present catalysis systems and exploring new catalysts with
better performances [8-10]. Many achievements in mechanistic studies have been made through detailed
polymer chain structure characterization [11-14]. For understanding the reaction mechanism, studies on the
polymerization kinetics have been conducted by more research groups [15-27]. It is widely accepted that
mechanistic studies based on direct measurement of the active center concentration in the metallocene
catalysis system can help the construction of a comprehensive mechanistic model. There are already many
literature reports on kinetics of metallocene catalyzed olefin polymerization that made use of different
methods of counting the number of active centers [28-46]. In most of the literatures, the determined active
center fraction ([C*]/[Mt], where [Mt] is concentration of the metallocene complex) of metallocene catalysis
systems was evidently lower than 1, showing that only a small proportion of the metallocene species can
work as the active center. In our previous study, we have compared the active center fractions and chain
propagation rate constants of ethylene and propylene polymerizations and their copolymerization with rac-
Et(Ind)2ZrCl2 (Mt-II) activated by modified methylaluminoxane (MMAO) [47]. The results showed the
presence of multiple active species in olefin polymerization with Mt-II/MMAO, among them some
transition metal species are unable to catalyze propylene homopolymerization, while some species become
dormant in ethylene homopolymerization. In ethylene-propylene copolymerization the dormant species
present in ethylene polymerization can be reactivated by propylene, resulting in nearly complete activation
of all the metallocene molecules. This work clearly shows that the knowledge of active center fractions and
kinetic parameters can greatly intensify our insights into the mechanism of metallocene catalyzed
polymerizations.

One of the main benefits of using metallocene catalysts is feasibility of regulating the catalytic behavior
by modifying the structure of ligands on the metallocene. In the cases of propylene polymerization with
racemic ansa-metallocenes, great improvements in isotacticity and molecular weight of polypropylene have
been achieved by respectively introducing a methyl at substitution positions 2 and a phenyl or alkyl group at
position 4 of bridged bisindenyl ligand of ansa-zirconocene complexes. As compared with Mt-II and rac-
Me2Si(Ind)2ZrCl2, rac-Me2Si(2-Me-4-Ph-Ind)2ZrCl2 (Mt-I) was found to produce isotactic polypropylene
with much higher isotacticity and molecular weight besides a threefold enhancement in catalytic activity
2
[48]. Ethylene polymerization activity with Mt-I was also much higher than with Mt-II [48]. However, there
are still no systematical mechanistic studies on ansa-zirconocenes with different ligand structure based on
experimental determination of the active center fraction and kinetic parameters.

In this work, systematical comparative studies on the kinetics of ethylene and propylene
polymerizations with Mt-I and Mt-II are made, using borate/alkylaluminum as the activator. The method of
counting the number of active centers adopted in this work is based on selectively quenching the transition
metal-polymer bonds with acyl chloride, which has been verified by application in the studies of olefin
polymerization with heterogeneous Ziegler-Natta, metallocene and late-transition metal catalysts [44,47,49-
63]. Such a comparative study will enable us to elucidate or clarify several important mechanistic problems
in metallocene catalyzed polymerization, especially the relationship between metallocene structure and
characteristics of its catalytic active centers. In the previous work [47], we have found that the active center
fraction ([C*]/[Zr]) was much lower in propylene polymerization than that of ethylene polymerization when
Mt-II/MMAO was the catalyst/activator combination. It will be interesting to see how changes in
metallocene structure influence the activation efficiency as well as the kinetic features of olefin
polymerization. Comparison of active center fraction and kinetic constants of the same metallocene/olefin
combination activated by aluminoxane and borate cocatalysts will also be highly interesting, since no such
comparative studies have been reported before [46].

2. Results and discussion

2.1 Ethylene polymerization


Ethylene polymerization catalyzed with two ansa-zirconocenes (Mt-I and Mt-II) activated by
borate/TIBA were firstly investigated. For each zirconocene, a series of ethylene polymerization runs were
conducted in toluene, and the active centers were quenched after different durations. Time dependent change
of chain propagation rate (Rp) with time was determined from differentiation of the curve of polymer yield
versus time. Sulfur content of the purified polymer samples were measured to determine the active center
fraction ([C*]/[Zr]) according to the equation [S] = [C*], and change of [C*]/[Zr] with reaction time was
determined. According to the rate equation Rp = kp[C*][M] that is well established for most catalytic olefin
polymerizations, the time dependent change of chain propagation rate constant (kP) was also determined,
using equilibrium ethylene concentration [M] = 0.085 mol/L [64,65] (toluene as solvent, 1 atm and 50˚C) in
the calculation. The results of polymerization, including time-dependent [C*]/[Zr], Rp, kp, polymer molecular
weight and thermal properties data, are summarized in Table 1 and Table 2, and the plots of Rp changes with
time for both catalysts are shown in Fig. 1.

3
Table 1

Results of ethylene polymerization with Mt-I/boratea

Run Time Yield Rp [C*]/[Zr] kp Mw Ɖ Tm ∆Hm


(min) (g) (mmolPE/molMt·s) (%) (L/mol·s) (kg/mol) (°C) (J/g)
1.1 0.5 0.26 30.4 6.5 5430 132 156
1.2 1 0.30 30.3 9.8 3610 163.5 4.9 135 170
1.3 1.5 0.31 30.2 10.3 3410 198.2 5.3 135 173
1.4 2 0.33 30.1 13.1 2670 135 166
1.5 5 0.42 29.5 21.7 1580 136.7 5.5 135 179
1.6 10 0.82 28.6 51.6 640 136 173
1.7 15 1.03 27.8 51.9 620 244.4 8.7 136 151
1.8 20 1.31 27.1 66.0 480 133 164
1.9 30 1.97 25.6 69.4 430 133 173
a
Reaction conditions: catalyst Mt-I 1.25 µmol, borate 2.5 µmol, TIBA 1000 µmol, ethylene pressure = 0.1 MPa, temperature = 50˚C,
solvent: toluene (50 mL), TPCC 2000 µmol.

Table 2

Results of ethylene polymerization with Mt-II/boratea

Run Time Yield Rp [C*]/[Zr] kp Mw Ɖ Tm ∆Hm


(min) (g) (mmolPE/molMt·s) (%) (L/mol·s) (kg/mol) (°C) (J/g)
2.1 0.5 0.21 34.5 3.3 12200 47.3 3.1 122 147
2.2 1 0.28 34.2 5.6 7080 124 156
2.3 1.5 0.33 33.6 6.5 6010 51.0 3.6 127 168
2.4 2 0.35 33.0 9.2 4170 130 172
2.5 5 0.51 29.5 12.6 2270 65.1 3.6 134 183
2.6 10 0.74 24.5 21.8 1310 135 187
2.7 15 0.96 20.0 41.5 560 80.8 4.4 133 190
2.8 20 1.25 16.9 43.6 450 132 202
2.9 30 1.38 11.8 49.0 280 132 180
a
Reaction conditions: Mt-II 1.25 µmol, borate 2.5 µmol, TIBA 1000 µmol, ethylene pressure = 0.1 MPa, temperature = 50˚C,
solvent: toluene (50 mL), TPCC 2000 µmol.

4
Fig. 1. Change of chain propagation rate with time in ethylene polymerization.

From Table 1 and Table 2 we found that Mt-I showed higher average activity in 30 min than Mt-II.
However, as shown in Fig. 1, the initial polymerization rate of Mt-II was even slightly higher than that of
Mt-I, but the later showed more stable activity than Mt-II when the reaction duration was prolonged to 30
min. It is worth noting that both catalysts showed high initial polymerization rate, which can be taken as
evidence for fast initiation of the active centers.

Weight average molecular weight of the obtain polyethylene (PE) were lower in the initial stage, and
slightly increased with polymerization time (tp) in both catalytic systems. Molecular weight distribution
(MWD) of PE produced by Mt-I was rather broad, and became even broader with proceeding of the
reaction. Though MWD of PE produced by Mt-II was narrower than that by Mt-I, its polydispersity index
(Ɖ) was still evidently larger than 2, the theoretical polydispersity index of a true single site catalyst,
implying presence of multiple active sites in the catalytic system. As compared with molecular weight of PE
produced by Mt-II/MMAO under similar conditions [47], the Mt-II/borate catalyst produced PE with
evidently lower molecular weight but similar MWD. Chain transfer of the active centers with TIBA in the
Mt-II/borate system could be responsible for the lower molecular weight.

As shown in Fig. 2 and Fig. 3, the active center fraction of both catalytic systems underwent a continuous
increase with time in the first 20 min of polymerization, and then approached to a steady level. Both systems
had a very low initial [C*]/[Zr] level (3~6% at 0.5 min), but the Mt-I/borate system reached a higher
[C*]/[Zr] level than that of Mt-II/borate after 30 min. In contrast, the Mt-II/MMAO and Mt-II/MAO
catalyzed ethylene polymerization showed much higher initial [C*]/[Zr] level (25~33% at 2.5 min) and
relatively faster increase of [C*]/[Zr] in the polymerization process (steady [C*]/[Zr] level was reached after
5 min in Mt-II/MMAO) [44,47]. The [C*]/[Zr] level after 30 min of polymerization (73%) was also higher in
5
the Mt-II/MMAO system. of It means that metallocene activation by the borate/alkylaluminum and the
aluminoxane cocatalysts have different characteristics.

Fig. 2. Variation of [C*]/[Zr] and kp with time in ethylene polymerization with Mt-I/borate.

Fig. 3. Variation of [C*]/[Zr] and kp with time in ethylene polymerization with Mt-II/borate.

The decrease of propagation rate constant with time (see Fig. 2 and Fig. 3) may be partly attributed to
decrease of local monomer concentration due to diffusion barrier, as the rate of ethylene diffusion from the
bulk of solution to the location of active centers will rapidly decrease when the initial batch of polymer
chains are formed and aggregate around the active centers. However, since the active center concentration
was increased for more than 10 times in the first 20 min, the possible reactivity difference between the active
6
centers activated at different reaction stages may also cause significant change of the kp value. For example,
if the active centers activated in the period of 5-10 min are 50% less active than those activated in the period
of 0-5 min, and the former account for 60% of total active centers, the average kp value at tp = 10 min will be
only 70% of that at tp = 5 min (assuming no diffusion barrier in the system). Therefore, the steeper kp
decrease of Mt-II/borate than that of Mt-I/borate (ratio of kp at 30 min to kp at 5 min was 0.12 for Mt-
II/borate, in contrast to 0.27 for Mt-I/borate) should be mainly attributed to the different intrinsic catalytic
properties of the two metallocenes.

2.2 Propylene polymerization


Propylene polymerizations with the same Mt-I/borate and Mt-II/borate catalysts were conducted under
the same conditions as the ethylene polymerization, and the results are summarized in Table 3 and Table 4.
In contrast to the similar activities between ethylene polymerization with Mt-I/borate and Mt-II/borate, in
propylene polymerization, the activity of Mt-I/borate was more than one magnitude higher than that of Mt-
II/borate in the initial stage (tp = 0~5 min), and the polymer yield at 30 min of the former system was about
5 times of the latter. Similar difference of propylene polymerization activity between Mt-I and Mt-II has
been reported by W. Spaleck et al [48]. Surprisingly enough, the activity of propylene polymerization with
Mt-I was even higher than that of ethylene polymerization. According to W. Spaleck et al., activity of
ethylene polymerization with Mt-I/MAO was evidently higher than that of propylene at comparable
conditions [48]. The borate cocatalyst used in this work could be responsible for the phenomenon. The rate
curve of Mt-I/borate system (Fig. 4a) showed steep deactivation with proceeding the polymerization, but the
Rp of Mt-II/borate showed only slight change with time (Fig. 4b). The stable rate curve of Mt-II/borate is
similar to that of propylene polymerization with Mt-II/MMAO [47], but the fast-decaying rate of Mt-
I/borate catalyzed propylene polymerization has not been reported before. M. Bochmann et al. has reported
moderately decaying 1-hexene polymerization with rac-Me2Si(2-Me-Benz[e]Ind)2ZrCl2 activated by MAO
[27].

The molecular weight of polypropylene (PP) produced with the two metallocenes were also strongly
different, with PP from Mt-I/borate system showing Mw about 4 times higher than that of Mt-II. Both
catalysts showed rather small change of Mw with polymerization time, but PP from the Mt-I/borate system
has broader MWD than that from Mt-II/borate.

Micro isotacticity of the PP products (isotactic pentad fraction [mmmm]) was determined by 13C NMR
analysis, and thermal properties of PP (Tm and ∆Hm) were determined by DSC analysis. As seen in Tables 3
and 4, isotacticity of PP from Mt-I/borate was evidently higher than that from Mt-II/borate, similar to the
trend reported in literature [48]. Owing to the higher isotacticity of PP samples from Mt-I, their melting
temperature and melting enthalpy were also evidently higher than those of Mt-II. However, [mmmm] of Mt-

7
I system was not so high (~80%) in the initial stage, and high [mmmm] (> 92%) was reached after 30 min of
polymerization. In contrast, [mmmm] of PP from Mt-II was rather stable. The change of chain structure with
time in the Mt-I/borate system could be related to its unstable activity in the reaction process. Further
discussions on this phenomenon will be made in the later part of the article.

Table 3

Results of propylene polymerization with Mt-I/boratea

Run Time Yield Rp [C*]/[Zr] kp Mw Ɖ Tm ∆Hm [mmmm] b


(min) (g) (mmolPP/molMt·s) (%) (L/mol·s) (kg/mol) (°C) (J/g) (mol%)
3.1 0.5 0.90 115.2 20.3 1510 65.5 3.1 151 87 78.7
3.2 1 0.92 110.6 20.5 1430 144 67
3.3 1.5 0.98 106.2 23.8 1190 45.0 3.6 144 50 83.3
3.4 2 1.28 102.0 24.1 1130 143 53
3.5 5 1.86 80.0 26.0 820 57.4 3.6 146 70 84.7
3.6 10 2.89 54.0 29.3 490 144 65
3.7 15 3.73 36.5 47.6 200 52.3 4.4 143 87 87.0
3.8 20 4.44 24.5 51.6 130 145 62 87.0
3.9 30 4.67 11.6 52.5 60 64.3 4.2 144 73 92.6
a
Reaction conditions: catalyst Mt-I =1.25 µmol, Borate = 2.5 µmol, TIBA = 1000 µmol, propylene pressure = 0.1 MPa, solvent:
toluene = 50 mL, TPCC = 2000 µmol. b Fraction of isotactic pentads determined by 13C NMR analysis.

Table 4

Results of propylene polymerization with Mt-II/boratea

Run Time Yield Rp [C*]/[Zr] kp Mw Ɖ Tm ∆Hm [mmmm] b


(min) (g) (mmolPP/molMt·s) (%) (L/mol·s) (kg/mol) (°C) (J/g) (mol%)

4.1 0.5 0.16 8.3 6.0 360 13.9 2.0 103 24 77.5
4.2 1 0.20 8.2 7.6 290 113 41
4.3 1.5 0.23 8.1 8.4 260 12.7 2.0 116 47 74.6
4.4 2 0.25 8.0 8.7 250 121 50
4.5 5 0.28 7.8 8.9 230 13.2 2.0 122 53 71.4
4.6 10 0.48 7.6 14.8 140 12.7 2.0 118 56
4.7 15 0.53 7.6 17.2 120 12.0 2.1 119 52 71.9
4.8 20 0.61 7.5 18.8 110 118 51
4.9 30 0.92 7.4 19.3 100 119 54 72.5
a
Reaction conditions: catalyst Mt-II =1.25 µmol, Borate = 2.5 µmol, TIBA = 1000 µmol, propylene pressure = 0.1 MPa, solvent:
toluene = 50 mL. TPCC = 2000 µmol. b Fraction of isotactic pentads determined by 13C NMR analysis.

8
Fig. 4. Change of chain propagation rate with time in propylene polymerization (a. Mt-I; b. Mt-II).

The active center fraction of Mt-I catalyzed propylene polymerization was around 20% in the initial stage,
and it continuously increased with time to over 50% at tp = 30 min (see Fig. 5). In comparison with ethylene
polymerization by Mt-I, we can see that the propylene polymerization system had a much higher initial
[C*]/[Zr] value than ethylene polymerization (20% vs. 6.5%), but a smaller [C*]/[Zr] fraction at the steady
stage (52% in PP vs. 69% in PE system at tp = 30 min). On the other hand, the Mt-II catalyzed propylene
polymerization showed evidently lower [C*]/[Zr] values in the initial stage ([C*]/[Zr] = 6.0% at tp = 0.5 min)
and the steady stage ([C*]/[Zr] = 19.3% at tp = 30 min) than the Mt-I system. Ethylene polymerization with
the same Mt-II catalyst showed lower initial [C*]/[Zr] value ([C*]/[Zr] = 3.3% at tp = 0.5 min) than
propylene polymerization, but reached a much higher [C*]/[Zr] in the steady stage ([C*]/[Zr] = 49% at tp = 30
min). These intriguing effects of monomer type and metallocene structure on activation of active centers may
imply complicated mechanism of the polymerization systems.

9
Fig. 5. Variation of [C*]/[Zr] and kp with time in propylene polymerization with Mt-I/borate.

Fig. 6. Variation of [C*]/[Zr] and kp with time in propylene polymerization with Mt-II/borate.

Decrease of kp with time was also observed in propylene polymerization systems. As seen in Fig. 5 and
Fig. 6, the extent of kp decrease in 30 min was significantly larger in Mt-I than that in Mt-II. This is contrary
to the situation of ethylene polymerization with the two metallocenes (see Fig. 2 and Fig. 3), where the Mt-
II showed larger extent of kp decrease than Mt-I.

It is interesting to compare the kinetic behaviors of Mt-II/borate catalyzed propylene polymerization


with those of the Mt-II/MMAO system reported in our previous work [47]. The Mt-II/MMAO catalyzed
propylene polymerization showed slightly higher [C*]/[Zr] fraction and similar kp value in the steady stage
([C*]/[Zr] = 27.7% and kp = 83 L·mol-1·s-1at tp = 30 min) as the Mt-II/borate system. The isotacticity of PP
produced with Mt-II/borate was also very close to that with Mt-II/MMAO. Considering that the

10
polymerization conditions adopted in this work are the same as that used in the previous work, it can be said
that changing the type of cocatalyst can exert limited influences on efficiency of metallocene activation and
kinetic behaviors of propylene polymerization catalyzed by Mt-II.

2.3 Discussions on the polymerization mechanism


In our previous studies on ethylene polymerization with Mt-II/MAO and Mt-II/MMAO catalysts, the
evident rises of PE molecular weight and polydispersity index with polymerization time were explained by
later activation of active centers with shorter cation-anion distance (active centers composed of tight ion
pairs) than those with longer cation-anion distance (loose ion pairs) [44,47]. The loose ion pairs were found
to show higher reactivity for chain propagation and produce PE with lower molecular weight, since
interferences of the counter anion on monomer coordination to the cationic active center are weaker, and
conformational transformation of the propagation chain for β-hydrogen transfer becomes easier in the
relatively open ion pair (see Scheme 1). The ion pairs with shorter cation-anion distance will be less active
for ethylene coordination owing to strong interferences of the counter anion on monomer coordination. The
short cation-anion distance in the tight ion pairs will also cause their retarded activation, since approaching
and coordination of the first monomer will be hindered by the counter anion more seriously than in the loose
ion pairs.

Scheme 1. Active centers of metallocene catalyst with different cation-anion distances (Pn denotes the
propagating polymer chain, and X denotes (R•MAO) or B(C6F5)4).

Since similar rises of PE molecular weight and polydispersity index with tp have been observed in
ethylene polymerization with Mt-I/borate and Mt-II/borate catalysts, we tend to explain the phenomena with
the same mechanistic model of Scheme 1. As shown in the MWD curves of PE produced at different tp (see
Fig. 7), the high molecular weight component of the MWD curve was continuously enhanced with
proceeding of the polymerization in both catalytic systems. This time-dependent changes of MWD in tp = 0.5
~ 15 min can be reasonably attributed to the marked increase of active center fraction in the same period (see
Fig. 2 and Fig. 3), assuming that the later activated active centers have shorter cation-anion distance and
produce PE with higher molecular weight.

11
Fig. 7. MWD curves of polyethylene synthesized with (a) Mt-I/borate and (b) Mt-II/borate at different time.
(The dashed lines have been added to indicate the peak position)

The different kinetic behaviors of Mt-I/borate and Mt-II/borate systems can be tentatively explained by
this model. As shown in Table 1 and Table 2, kp value of the Mt-II catalyzed ethylene polymerization was
higher than that of the Mt-I system in most of the reaction period, meanwhile it produced PE with lower
molecular weight. It means that active centers in Mt-II/borate catalyst has longer cation-anion distances than
the Mt-I/borate system. Presence of methyl and phenyl substituents in Mt-I could be responsible for its
tighter ion pairs.
The higher [C*]/[Zr] values of Mt-I/borate than that of Mt-II/borate in both the initial and steady stages
of ethylene polymerization can be tentatively attributed to lower fraction of Zr−CH2CH3 type dormant site in
the former. In our previous work, evidences supporting presence of such dormant sites have been presented
[47]. As shown in Scheme 2, the Zr−CH2CH3 species formed via ethylene insertion in Zr−H can turn into the
dormant state by β-agostic interaction between the methyl hydrogens and the cationic Zr center. The strength
of the β -agostic interaction will be weakened when electron deficiency of the Zr center is lower. Considering
that Mt-I has electron donating substituents (methyl and phenyl) on the ligand, its Zr center should have
lower degree of electron deficiency, resulting in lower fraction of dormant sites.

12
Scheme 2. Model of dormant site formation in ethylene polymerization (the anionic counter ion has been
omitted for clarity).

In the case of propylene polymerization, similar type of dormant sites can be formed via 2,1-insertion of
a propylene monomer in Zr−H that has been formed by β-hydrogen transfer (see Scheme 3) [66,67]. The β-
agostic interaction between the methyl hydrogens and the cationic Zr center of Zr−CH(CH3)2 species makes
them inaccessible to the incoming monomer. Although the more favorable 1,2- insertion of a propylene
monomer in Zr−H leads to Zr−CH2CH2CH3 species that are active for propylene polymerization, quick
accumulation of the Zr−CH(CH3)R (R is an alkyl or a polyolefin chain) type dormant sites can still make a
large population of transition metal species that are nearly inactive to propylene coordination [68]. This
factor should be responsible for the presence of inactive zirconocene species (accounting for 50~80% of the
total metallocene load) in both Mt-I and Mt-II catalyzed propylene polymerization (see Fig. 5 and Fig. 6).
The inactive zirconocene species can thus be attributed to Zr−CH(CH3)2 type dormant sites. Among the two
zirconocenes, Mt-I/borate system had much lower proportion of dormant sites than Mt-II/borate, which can
also be explained by lower electron deficiency of the Zr center in Mt-I.

13
Scheme 3. Model of dormant site formation in propylene polymerization (the anionic counter ion has been
omitted for clarity).

The very steep decay of apparent propagation rate constant (kp) with time in Mt-I/borate catalyzed
propylene polymerization (kp value declined to about 1/15 of the initial level in 30 min) is somewhat strange,
since kp value of the Mt-II/borate system was only moderately lowered in the 30 min reaction process (see
Fig. 5 and Fig. 6). This sharp contrast between the two zirconocenes could be explained by the presence of
multiple active centers in Mt-I/borate, as its PP products have broader MWD than that of Mt-II/borate (see
Table 3 and Table 4). The polydispersity index of PP from Mt-II/borate was nearly 2, meaning that there is
only one type of active center. If the monomer concentration gradient in the nascent polymer/catalyst
particles keeps constant in the polymerization process (namely, the influence of diffusion barrier on local
monomer concentration is negligible), the kp value should not vary with time in a single-site system. As seen
in Fig. 6, the experimentally determined kp value decreased for about 70% in the 30 min reaction process.
This kp decay can only be attributed to gradual development of monomer concentration gradient in the
polymer/catalyst particles. In contrast, the polydispersity index of PP from Mt-I/borate was larger than 3,
indicating coexistence of multiple active centers. In this case the experimentally determined kp value should
be a statistical average based on the propagation rate constant and number of different active centers. When
the active centers with lower kp value are activated in the later stage of polymerization, the average kp value
will become lower in the later stage. This effect will make additional kp decay on the basis of kp decay caused
by diffusion barrier. When we consider the factor that the extent of diffusion barrier enhances with

14
polymer/catalyst mass ratio [54,61], the markedly larger extent of kp decay in Mt-I than Mt-II becomes more
understandable, since polymer/catalyst mass ratio of the former system was much larger than the latter.
In propylene polymerization with Mt-I/borate, the isotacticity of produced PP continuously increased
with time, meanwhile in the Mt-II/borate system the isotacticity became lower in the later stage. As
proposed in our previous work [47], the phenomenon of declining isotacticity with time may be attributed to
time-dependent decline of monomer concentration for the rising diffusion barrier. The negative correlation of
[mmmm] with propylene concentration has been reported in literatures [69,70]. However, unlike the single-
site behaviors of the Mt-II/borate system, the Mt-I/borate system clearly showed multi-site behaviors in
propylene polymerization. In this case the stereoselectivity of the later activated active centers may impose
stronger influence on the [mmmm] of PP produced in the later stage. When the later activated active centers
have markedly higher stereoselectivity than those activated in the early stage, it will be possible to see
[mmmm] rising with time. This is likely the case of Mt-I/borate system, in which the later activated active
centers produce PP with much higher isotacticity than that produced by the early activated centers. The
shorter anion-cation distance of the later activated active centers could be responsible for their higher
stereoselectivity, as strong interactions from the counterion could prevent “incorrect” propylene coordination
on the Zr center that leads to stereoerror in the propagation chain. This explanation is in accord with the
previous discussions on polymerization kinetics, as the kinetic results suggest that the later activated active
centers in Mt-I/borate have lower kp value than the early activated centers, and lower kp value is
characteristic of the active centers with tighter ion pairs.

3. Conclusions
In ethylene and propylene polymerizations catalyzed with Mt-II/borate, the active center fraction
([C ]/[Zr]) gradually increases with time in the first 20 min and then reaches a steady stage. The [C*]/[Zr]
*

value in the steady stage of ethylene polymerization was significantly higher than that of propylene
polymerization, similar to the phenomena previously observed in ethylene and propylene polymerizations
catalyzed with Mt-II/MMAO. When ansa-zirconocene Mt-I/borate was used as the catalyst, both the
ethylene and propylene polymerizations showed higher steady stage [C*]/[Zr] levels than those of the Mt-
II/borate systems, and the influence of zirconocene structure on [C*]/[Zr] value of propylene polymerization
was much stronger than that of ethylene polymerization. Change in zirconocene structure also significantly
influenced apparent chain propagation rate constant (kp) and its time-dependent variations in the
polymerizations of both ethylene and propylene. In ethylene polymerization, the decrease of kp with time in
Mt-I/borate system was relatively moderate as compared with the Mt-II/borate system. In contrast, the
decrease of kp with time in propylene polymerization with Mt-I/borate was much steeper than the Mt-
II/borate system. By taking into consideration the changes of [C*]/[Zr], polymer’s MWD and chain structure

15
with polymerization time (tp), the relationships between metallocene structure and kp~tp profile of the
polymerization process can be reasonably explained. Besides the effect of ascending monomer diffusion
barrier with time, presence of multiple active centers in the polymerization system and late activation of the
centers with lower intrinsic reactivity are considered the main reasons for steep decay of kp with time.

4. Experimental
All manipulations of air- and/or moisture-sensitive compounds were performed under a nitrogen
atmosphere using standard Schlenk techniques or glove box.

4.1 Materials
The metallocene catalyst rac-Me2Si(2-Me-4-Ph-Ind)2ZrCl2 (Mt-I) was kindly donated by Shanghai
Research Institute of Chemical Industry, Shanghai, China. rac-Et(Ind)2ZrCl2 (Mt-II) was purchased from
Sigma-Aldrich. Borate (Ph3C)B(C6F5)4 was synthesized according to literatures [71-73].
Triisobutylaluminum (TIBA) was purchased from Albemarle Co., and 2-thiophenecarbonyl chloride (TPCC,
98%) was purchased from J&K Scientific, China. Ethylene (polymerization grade, 99.9% purity) and
propylene (polymerization grade 99.9% purity) were purchased from Zhejiang mixing Gas Co. (Hangzhou,
China) and further purified by passing through columns containing deoxygenize agent and 4Å molecular
sieves in a gas purification system (Dalian Samat Chemicals Co., Ltd., China). n-Heptane (Shanghai Titan
Scientific Co., Ltd., China) was purified by passing through columns of molecular sieve and deoxygenate
agent, refluxed over sodium-benzophenone and distilled under nitrogen prior to use. Toluene (HPLC grade,
Jiangsu Yonghua Fine Chemical Co., Ltd., China) were refluxed over sodium-benzophenone and distilled
under nitrogen before use. Solution of cocatalyst TIBA (2M in n-heptane) was made by injecting required
quantity of neat TIBA into a sealed bottle containing n-heptane. Solution of the quenching agent TPCC was
distilled and diluted with n-heptane to 2 M before use.

4.2 Polymerization
All the polymerization runs were performed in a 100-mL Schlenk flask with a magnetic stirrer. The
flask was first filled with about 50 mL toluene at 50ºC and then saturated with gaseous ethylene or propylene
of 0.1 MPa. TIBA solution was first added to the reactor, followed by toluene solutions of the borate and the
metallocene. The time of adding metallocene was taken as the start point of polymerization. Ethylene or
propylene with pressure of 0.1 MPa was continuously fed to the reactor during the polymerization. After
designed time, the polymerization was quenched by adding TPCC solution (TPCC/Al = 2) for 5 min at the
same temperature. Then excess amount of ethanol (containing about 2% HCl) was injected to decomposed
the catalyst species and the unreacted quencher, and the polymer was precipitated. The polymer samples
were thoroughly purified by a three-step process according to our previous work [49,50] and dried in
vacuum.
16
4.3 Characterization
Sulfur content of the quench-labeled polymer samples was measured by a YHTS-2000 ultraviolet
fluorescence sulfur analyzer (produced by Jiangyan Yinhe Instrument Co., Jiangyan, China) with a lower
detection limit of 0.05 ppm. Three parallel measurements for sulfur content were made for each polymer
sample, and their average value was taken as the sulfur content. The S content of a blank polyethylene
sample synthesized at the same conditions without the quenching step was found to be nearly 0, in contrast
to 5-25 ppm S content of the quench-labeled samples.
Molecular weight and molecular weight distribution (MWD) of polymer samples were measured by
GPC with a PL 220 GPC instrument (Polymer Laboratories) equipped with three PL-gel 10 µm MIXED-B
columns using 1,2,4-trichlorobenzene as the eluent at a flow of 1.0 mL/min at 150ºC. Universal calibration
against narrow polystyrene standards was adopted.

13
C NMR spectra of the polypropylene samples were recorded on a Varian Mercury plus 300
spectrometer operating at 75 MHz in pulse Fourier transform mode. o-Dichlorobenzene-d4 was used as
solvent, and concentration of the polymer solution was 10 wt%. The spectra were recorded at 120ºC using
hexamethyldisiloxane as internal chemical shift reference. Cr(acac)3 was added to reduce the relaxation time
of carbon atoms, and the pulse delay time was set at 3 s. The pulse angle was 90º. The acquisition time was
0.8 s. The spectral width was 8000 Hz. Inverse gated decoupling was performed for integration. Typically,
4000 scans were collected.

Differential scanning calorimetry (DSC) analysis was made with a TA Q200 instrument calibrated with
indium and water. 4 to 6 mg of each sample were weighed and sealed into aluminum pan. The sample was
first heated to 150ºC and held for 5 min for eliminating thermal history, and then cooled to 20ºC at a rate of
10ºC/min. Finally, the sample was heated again to 180ºC at a rate of 10 ºC/min to record the melting curve.

Acknowledgments

This work is financially supported by the National Natural Science Foundation of China (Grant No.
51773178) and the Department of Science and Technology, Zhejiang Provincial Government. The authors
are grateful to Dr. Yucai Cao of SRICI, China for donating Mt-I.

References

[1] W. Kaminsky. Macromol. Chem. Phys. 197 (1996) 3907–3945.

[2] S. Bensason, J. Minick, A. Moet, S. Chum, A. Hiltner, E. Baer. J. Polym. Sci., Part B: Polym. Phys. 34 (1996)
1301–1315.

17
[3] W. Kaminsky. J. Chem. Soc., Dalton Trans. (1998) 1413–1418.

[4] G. W. Coates. Chem. Rev. 100 (2000) 1223–1252.

[5] L. Resconi, L. Cavallo, A. Fait, F. Piemontesi. Chem. Rev. 100 (2000) 1253–1346.

[6] W. Kaminsky. J. Polym. Sci., Part A: Polym. Chem. 42 (2004) 3911–3921.

[7] Y. Choi, J. B. P. Soares. Can. J. Chem. Eng. 90 (2012) 646–671.

[8] W. Kaminsky, Macromol. Chem. Phys. 209 (2008) 459–466.

[9] M. Bochmann, J. Organomet. Chem. 689 (2004) 3982–3998.

[10] A.A. Altaf, A. Badshah, N. Khan, S. Marwat, S. Ali, J. Coord. Chem. 64 (2011) 1815–1836.

[11] I. Tritto, L. Boggioni, D.R. Ferro, Coordin. Chem. Rev. 250 (2006) 212–241.

[12] V. Busico, R. Cipullo, F. Cutillo, M. Vacatello, V. Van Axel Castelli, Macromolecules 36 (2003) 4258–4261.

[13] V. Busico, R. Cipullo, F. Cutillo, M. Vacatello, Macromolecules 35 (2002) 349–354.

[14] M. Galimberti, M. Destro, O. Fusco, F. Piemontesi, I. Camurati, Macromolecules 32 (1999) 258–263.

[15] W. Kaminsky, A. Bark, R. Steiger, J. Mol. Catal. 74 (1992) 109–119.

[16] D. Fischer, R. Mulhaupt, J. Organomet. Chem. 417 (1991) C7–C11.

[17] N. Naga, K. Mizunuma, Macromol. Chem. Phys. 199 (1998) 113–118.

[18] Q. Wang, J. Weng, L. Xu, Z. Fan, L. Feng, Polymer 40 (1999) 1863–1870.

[19] S. Lin, C.D. Tagge, R.M. Waymouth, M. Nele, S. Collins, J.C. Pinto, J. Am. Chem. Soc. 122 (2000) 11275–11285.

[20] J. Liu, J.A. Støvneng, E. Rytter, J. Polym. Sci., Part A: Polym. Chem. 39 (2001) 3566–3577.

[21] Y.S. Ko, S.I. Woo, Eur. Polym. J. 39 (2003) 1609–1614.

[22] P.J. Chirik, J.E. Bercaw, Organometallics 24 (2005) 5407–5423.

[23] F. di Lena, E. Quintanilla, P. Chen, Chem. Commun. (2005) 5757–5759.

[24] Y.V. Kissin, A.S. Goldman, Macromol. Chem. Phys. 210 (2009) 1942–1956.

[25] M. Bochmann, Organometallics 29 (2010) 4711–4740.

[26] S. Mehdiabadi, J.B.P. Soares, Macromolecules 45 (2012) 1777–1791.

[27] F. Ghiotto, C. Pateraki, J.R. Severn, N. Friederichs, M. Bochmann, Dalton. Trans. 42 (2013) 9040–9048.

[28] J.C.W. Chien, B.P. Wang, J. Polym. Sci., Part A: Polym. Chem. 27 (1989) 1539–1557.

[29] T.K. Han, Y.S. Ko, J.W. Park, S.I. Woo, Macromolecules 29 (1996) 7305–7309.

[30] V. Busico, R. Cipullo, V. Esposito, Macromol. Rapid. Comm. 20(1999) 116–121.

18
[31] Z.X. Liu, E. Somsook, C.R. Landis, J. Am. Chem. Soc, 123 (2001) 2915–2916.

[32] Z.X. Liu, E. Somsook, C.B. White, K.A. Rosaaen, C.R. Landis, J. Am. Chem. Soc. 123 (2001) 11193–11207.

[33] F.Q. Song, R.D. Cannon, M. Bochmann, J. Am. Chem. Soc. 125 (2003) 7641–7653.

[34] F. Song, M.D. Hannant, R.D. Cannon, M. Bochmann, Macromol. Symp. 213 (2004) 173–186.

[35] F. Song, R.D. Cannon, S.J. Lancaster, M. Bochmann, J. Mol. Catal. A: Chem. 218 (2004) 21–28.

[36] F.Q. Song, R.D. Cannon, M. Bochmann, Chem. Commun. (2004) 542–543.

[37] M. Bochmann, R.D. Cannon, F. Song, Kinet. Catal. 47 (2006) 160–169.

[38] J.A.M. Awudza, P.J.T. Tait, J. Polym. Sci., Part A: Polym. Chem. 46 (2008) 267–277.

[39] K.A. Novstrup, N.E. Travia, G.A. Medvedev, C. Stanciu, J.M. Switzer, K.T. Thomson, W.N. Delgass, M.M. Abu-
Omar, J.M. Caruthers, J. Am. Chem. Soc. 132 (2010) 558–566.

[40] M.D. Christianson, E.H.P. Tan, C.R. Landis, J. Am. Chem. Soc. 132 (2010) 11461–11463.

[41] B.M. Moscato, B. Zhu, C.R. Landis, J. Am. Chem. Soc. 132 (2010) 14352–14354.

[42] C.-H. Chen, W.-C. Shih, C. Hilty. J. Am. Chem. Soc. 137 (2015) 6965–6971.

[43] D.L. Nelsen, B.J. Anding, J.L. Sawicki, M.D. Christianson, D.J. Arriola, C.R. Landis. ACS Catal. 6 (2016) 7398–
7408.

[44] Y. Guo, F. He, Z. Zhang, A. Khan, Z. Fu, J. Xu, Z. Fan. J. Organomet. Chem. 808 (2016) 109–116.

[45] Cueny, E.S., H.C. Johnson, C.R. Landis, ACS Catal. 8 (2018) 11605–11614.

[46] X. Desert, J.-F. Carpentier, E. Kirillov, Coord. Chem. Rev. 386 (2019) 50–68.

[47] Y. Guo, Z. Zhang, W. Guo, A. Khan, Z.S. Fu, J.T. Xu, Z.Q. Fan. J. Polym. Sci., Part A: Polym. Chem. 55 (2017)
867–875.

[48] W. Spaleck, F. Kuber, A. Winter, J. Rohrmann, B. Bachmann, M. Antberg, V. Dolle, E.F. Paulus. Organometallics
13 (1994) 954–963.

[49] X.R. Shen, J. Hu, Z.S. Fu, J.Q. Lou, Z.Q. Fan, Catal. Commun. 30 (2013) 66–69.

[50] X.R. Shen, Z.S. Fu, J. Hu, Q. Wang, Z.Q. Fan, J. Phys. Chem. C 117 (2013) 15174–15182.

[51] B. Zhang, L.T. Zhang, Z.S. Fu, Z.Q. Fan, Catal. Commun. 69 (2015) 147–149.

[52] T. Xu, H.R. Yang, Z.S. Fu, Z.Q. Fan, J. Organomet. Chem. 798 (2015) 328–334.

[53] H.R. Yang, L.T. Zhang, Z.S. Fu, Z.Q. Fan, J. Appl. Polym. Sci. 132 (2014) 41264.

[54] A. Khan, Y.T. Guo, Z.S. Fu, Z.Q. Fan, J. Appl. Polym. Sci. 134 (2017) 45187.

[55] A. Khan, Y.T. Guo, Z. Zhang, A. Ali, Z.S. Fu, Z.Q. Fan, J. Appl. Polym. Sci. 135 (2018) 46030.

19
[56] Y.T. Guo, P.J. Yang, S.J. Zhang, B.Y. Jiang, A. Khan, L. Zhu, Z.S. Fu, Z.Q. Fan, Iran. Polym. J. 27 (2018) 153–
159.

[57] B.Y. Jiang, Y.H. Weng, S.J. Zhang, Z. Zhang, Z.S. Fu, Z.Q. Fan, J. Catal. 360 (2018) 57–65.

[58] P.J. Yang, Z.S. Fu, Z.Q. Fan, Mol. Catal. 447C (2018) 13–20.

[59] Y.H. Weng, B.Y. Jiang, Z.S. Fu, Z.Q. Fan, J. Appl. Polym. Sci. 135 (2018) 46605.

[60] B.Y. Jiang, X.Y. Liu, Y.H. Weng, Z.S. Fu, A.H. He, Z.Q. Fan, J. Catal. 369 (2019) 324–334.

[61] B.Y. Jiang, F. He, P.J. Yang, Z. Zhang, Y.H. Weng, Z.M. Cheng, Z.S. Fu, Z.Q. Fan, Catal. Commun. 121 (2019)
38–42.

[62] Z. Zhang, B.Y. Jiang, F. He, Z.S. Fu, Z.Q. Fan, Polymers 11 (2019) 358.

[63] Z. Zhang, B.Y. Jiang, B. Zhang, Z.S. Fu, Z.Q. Fan, Chinese J. Polym. Sci. 37 (2019) 1023–1030.

[64] M. Atiqullah, H. Hammawa, H. Hamid, Eur. Polym. J. 34 (1998) 1511–1520.

[65] L.S. Lee, H.J. Ou, H.I. Hsu, Fluid Phase Equ. 231 (2005) 221–230.

[66] Y.V. Kissin, J. Catal. 292 (2012) 188–200.

[67] Z.Q. Fan, L.T. Zhang, S.J. Xia, Z.S. Fu, J. Mol. Catal. A: Chem. 351 (2011) 93–99.

[68] Y. Yu, V. Busico, P. H. M. Budzelaar, A. Vittoria, R. Cipullo, Angew. Chem. Int. Ed. 55 (2016) 8590–8594.

[69] V. Busico, R. Cipullo, J. Am. Chem. Soc. 116 (1994) 9329–9330.

[70] M. J. Schneider, E. Kaji, T. Uozumi, K. Soga, Macromol. Chem. Phys. 198 (1997) 2899–2904.

[71] E. Ihara, V.G. Young Jr., R.F. Jordan, J. Am. Chem. Soc. 120 (1998) 8277–8278.

[72] M. Lehmann, A. Schulz, A. Villinger, Angew. Chem. Int. Ed. 48 (2009) 7444 –7447.

[73] M. Kuprat, M. Lehmann, A. Schulz, A. Villinger, Organometallics 29 (2010) 1421–1427.

20
Highlights

► Ethylene, propylene polymerizations with two ansa-zirconocenes were compared.

► rac-Me2Si(2-Me-4-Ph-Ind)2ZrCl2 showed higher activity than rac-Et(Ind)2ZrCl2.

► rac-Me2Si(2-Me-4-Ph-Ind)2ZrCl2 showed higher [C*]/[Zr] than EBI.

► [C*]/[Zr] ascended with time and then leveled off in each polymerization system.

► Apparent kP decayed for large extent in systems with multiple active centers.
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

View publication stats

You might also like