You are on page 1of 9

Continuous Solution Copolymerization of Ethylene and

Octene-1 with Constrained Geometry Metallocene Catalyst


WEN-JUN WANG, EDWARD KOLODKA, SHIPING ZHU, ARCHIE E. HAMIELEC
Department of Chemical Engineering, McMaster University, Hamilton, Ontario, Canada L8S 4L7

Received 26 October 1998; accepted 4 March 1999

ABSTRACT: The solution copolymerization of ethylene (1) with octene-1 (2) in Isopar E
using constrained geometry catalyst system, [C5Me4(SiMe2NtBu)]TiMe2 (CGC-Ti)/
tris(pentafluorophenyl)boron (TPFPB)/modified methylaluminoxane (MMAO), has
been carried out in a high-temperature, high-pressure continuous stirred-tank reactor
(CSTR) at 140C, 500 psig and with a mean residence time of 4 min. A series of
copolymer samples with octene-1 content up to 0.337 mole fraction were synthesized
and characterized. The estimated reactivity ratios were r1 5 7.90 and r2 5 0.099. The
CGC-Ti showed a higher ability to incorporate high a-olefins than other metallocene
catalysts investigated in the literature due to its open structure. The presence of
octene-1 lowered the catalyst activity, particularly at octene-1 levels higher than 0.45
mole fraction. Octene-1 was also found to reduce the molecular weight of polymer and
broaden the molecular weight distributions. The triad distributions were measured by
13
C-NMR. A minor penultimate effect was observed. The penultimate octene-1 unit
appeared to slow down monomer insertion rates. A comparison of the propagation rate
of octene-1 with the incorporation rate of macromonomer in the homopolymerization of
ethylene suggests that the addition of macromonomer is effectively instantaneous after
it is generated with diffusion to or from the active center reaction volume playing a
minor role. 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2949 2957, 1999
Keywords: metallocene polymerization; linear low density polyethylene; ethylene/
octene-1 copolymerization; constrained geometry catalyst; continuous solution polymerization; branching; reactivity ratio

INTRODUCTION
Linear low-density polyethylene (LLDPE) is an
important family of ethylene polymers because to
its superior processing properties. Copolymerization of ethylene with small amounts of a-olefins,
i.e. butene-1, hexene-1 and octene-1, produces
LLDPE with short chain branching which lowers
both crystallinity and density. In contrast to conventional ZieglerNatta catalysts, the metallocene catalyst technology provides copolymers
with narrower composition distributions, nar-

Correspondence to: S. Zhu


Journal of Polymer Science: Part A: Polymer Chemistry, Vol. 37, 2949 2957 (1999)
1999 John Wiley & Sons, Inc.
CCC 0887-624X/99/152949-09

rower molecular weight distributions (MWD), and


higher levels of comonomer insertion.17
The incorporation of high a-olefins, and the
molecular weight of the resultant polymers,
strongly depend on the structure of the catalyst
active centers.3,6 7 For the copolymerization of
ethylene with hexene-1, the syndiospecific metallocenes were found to be most effective for the
incorporation of high a-olefins due to their open
structures.3 The recent advent of constrained geometry catalysts (CGC) have offered new routes
in synthesizing branched polyolefins. The open
structure of CGC permits the incorporation of
macromonomers generated in situ via b-hydride
elimination and/or transfer to monomer. Such
branched polymers have an excellent combination
2949

2950

WANG ET AL.

of good mechanical properties and processibilities, the best of both worlds.8 11 The CGC active
center is bonded to a monocyclopentadienyl ring
and bridged with a heteroatom unit, forming a
constrained cyclic structure. This geometry allows
the transition metal center to readily incorporate
high a-olefins and macromonomers. In our recent
report on the kinetics of ethylene homopolymerization,11 the ethylene macromonomers were found to
be formed mainly via chain transfer to ethylene.
Branched polyethylenes were produced through the
in situ copolymerization of the macromonomer with
ethylene.
Following up the investigation of the continuous solution homopolymerization of ethylene using CGC, this paper reports the copolymerization
of ethylene with octene-1 using our high-temperature high-pressure continuous stirred-tank reactor (CSTR) system. The copolymerization kinetics
will be discussed and compared with the ethylene
homopolymerization data to further elucidate the
pathway of LCB formation.

EXPERIMENTAL
Materials
All manipulations with organometallic compounds
were conducted using a drybox technique. Catalyst,
[C5Me4(SiMe2NtBu)]TiMe2 (Me: methyl; tBu: isobutyl; C5: cyclopentadiene) (CGC-Ti), and co-catalyst,
tris(pentafluorophenyl)boron (TPFPB), were donated by Dow Chemical as 10 and 3 wt % solution in
Isopar E, respectively. Modified methylaluminoxane (MMAO-3A) was provided by Akzo-Nobel Corporation as 7.2 wt % aluminum in heptane. The
CGC-Ti and co-catalysts were used without further
purification. Polymerization-grade ethylene with
560 ppm (in mole) of hydrogen was from Matheson
Gas and further purified by passing through columns with CuO, Ascarite, and molecular sieves.
Octene-1 from Aldrich was distilled over CaH2. Isopar E (Caledon Laboratories Ltd.) as solvent was
dried over a mixture of 5A and 13X molecular sieves
from Fisher Scientific and silica gel from Caledon
Laboratories Ltd., and de-oxygenated by sparging
with ultra-high purity nitrogen (99.999%) from
Matheson Gas.
Polymerization
The continuous solution polymerization reactor
system used for the kinetic studies of ethylene/

octene-1 copolymerization is illustrated in Figure


1. Polymerizations were carried out at 500 psig
and 140C with a mean residence time (t) of 4
min. More details regarding the continuous solution polymerization reactor system set-up and
control can be found in our previous papers.1112
The CGC-Ti/MMAO/Isopar E, TPFPB/Isopar E,
and octene-1/ Isopar E solutions were prepared
and stored separately in the catalyst, co-catalyst,
and comonomer tanks. The solvent, catalyst, cocatalyst, and comonomer tanks were kept at a
gauge pressure of 15 psi under the protection of
nitrogen. By adjusting the feed rates of four inlet
streams, the mean residence time of 4 minutes,
CGC-Ti concentration in 15 mM, TPFPB in 45
mM, and MMAO in 150 mM were obtained and
remained unchanged for all runs.
The CSTR system was heated to the desired
temperature, and controlled at the set pressure.
Isopar E was then fed into the reactor. When the
flow rate of Isopar E reached 50% of the set value,
ethylene (1) and octene-1 (2) were charged and
the catalyst and co-catalyst pumps were turned
on. The initial time of polymerization was set
when all the conditions reached their pre-set values. During the polymerization process, polymer
samples were collected and mixed with anhydrous ethanol. The samples were then washed by
methanol, filtered, and dried under vacuum.
Characterization
13

C-NMR spectra at 120C were obtained with a


Bruker AC 300 pulsed NMR spectrometer and
used for the determination of copolymer compositions and triad distributions. The acquisition parameters and procedure followed ASTM D
5017-91 method.13 A polymer sample (1.2 g) with
1.5 mL trichlorobenzene and 1.3 mL deuterated
o-dichlorobenzene was transferred into a 10 mm
NMR sample tube. The carbon signals were collected with complete 1H decoupling using broadband decoupling method to remove any coupling
between proton and carbon nuclei. The detailed
chemical shift assignments of the copolymer
chain and their calculations followed ASTM D
5017-91 method.13 The composition of octene-1
was estimated from the mean value of the integrals of a-methylene and methine peaks, and that
of ethylene from the peaks of isolated methylene
groups. Figures 2(a) and 2(b) give the spectra of
three ethylene/octene-1 copolymer samples with
octene-1 molar fractions of 0.028 (run 1), 0.125
(run 4), and 0.337 (run 7), respectively. The spec-

SOLUTION COPOLYMERIZATION OF ETHYLENE AND OCTENE-1 METALLOCENE CATALYST

Figure 1.
this work.

2951

Schematic of the high-pressure high-temperature CSTR system used in

tral assignments of unsaturated chain ends are


shown in Figure 2(b).
Molecular weight and molecular weight distribution (MWD) of the copolymers were determined
by a high-temperature SEC (Waters-Millipore
150C) at 140C. The SEC was equipped with
three linear mixed Shodex AT806MS columns
and a differential refractive index detector. 1,2,4tricholorobenzene (TCB) was used as solvent at a
flow rate of 1.0 mL/min. The column was calibrated using known monodisperse polystyrene
(PS) standards with the MarkHouwink constants K 5 2.32 3 1024 and a 5 0.653 for PS, and
K 5 3.95 3 1024 and a 5 0.726 for PE.

RESULTS AND DISCUSSION


Activities and Molecular Weights
Ethylene and octene-1 copolymers with different
compositions were synthesized at various monomer feed ratios with the same total inlet concentration. The experimental conditions and results
are summarized in Table I. Copolymers having a
wide range of compositions were prepared. Octene-1 content reached as high as F2 5 0.337 in

mole fraction (run 7). This shows that the CGC-Ti


active centers can readily incorporate high a-olefins into the copolymer chains. A high content of
octene-1 eliminates crystallinity and makes the
ethylene copolymer more rubbery. The molecular
weights of the copolymers decreased with increasing octene-1 content. The molecular weight distributions were narrow with polydispersities of
about 2.
Compared with ethylene homopolymerization
under similar experimental conditions (run 0),11
the activity of CGC-Ti for the copolymerization
decreased significantly with the increase in mole
fraction of octene-1 in the reaction feed stream
(f20). Similar results were also reported for the
homogeneous copolymerization of ethylene with
hexene-1 using Cp2ZrCl2 and Et[Ind]2ZrCl2 as
catalysts.4 Figure 3 shows the effects of f20 on
ethylene and octene-1 conversions. Ethylene gave
higher conversions than octene-1. The CGC-Ti
catalyst is more active for ethylene propagation
than for octene-1. With the increase of f20, the
conversions decreased. However, it should be
noted that the conversion of ethylene changed
slightly for f20 , 0.45 and experienced a dramatic
decrease when f20 . 0.45. The high level of oc-

2952

WANG ET AL.

tene-1 caused an appreciable reduction in the ethylene consumption rate.


With the increase of f20, the weight-average
molecular weight (Mw) decreased and MWD
broadened as shown in Table I. The Mw of polyethylene for run 0 was 1.04 3 10.5 Introducing
small amount of octene-1 in the feed (5 mol % of
octene-1) reduced the Mw to 4.44 3 10.4 Figure
2(b) shows that both terminal vinyl groups (1V,
1V*) and vinylene groups (1VE, 1VE*) existed in
the copolymer samples. The formation of the terminal vinyl and end vinylene groups was via
chain transfer and b-hydride elimination of propagating centers with ethylene and octene-1 terminal units, respectively. The increase of chain end
unsaturation with increasing octene-1 incorporation indicates significant chain transfer reactions
at high f20 values. The existence of more vinylene
groups than vinyl groups and the increase of both
1V* and 1VE* with f20 further clarifies the effect
of octene-1 on the polymerization rate. Due to the
slow propagation rate, the active centers with
octene-1 terminal and/or penult units are more
likely to undergo a transfer reaction to form the
terminal unsaturated groups. This contributes to
the reduction in Mw.
Reactivity Ratios
The high-temperature and high-pressure solution
CSTR system provides an effective approach to
evaluate the reactivity ratios for a copolymerization system. When the CSTR system reaches
steady-state operation, the monomer and polymer
compositions remain constant, it therefore avoids
the composition drifting problem in batch techniques. Moreover, the utilization of a homogeneous copolymerization system at a high temperature minimizes the effect of diffusion limitations
on the reactivity ratios.
A first-order Markov model, described by the
following equations, was applied to the copolymerization of ethylene and octene-1:
k 11

AOEE
AOE 1 E O

(1)

k 12

Figure 2. 13C-NMR spectra of copolymer samples


(runs 1, 4, and 7) measured at 75.4 MHz and 120C
using deuterium o-dichlorobenzene and 1,2,4-trichlorobenzene as solvents. Chemical shifts: (a) from 10 to 50
ppm and (b) from 100 to 160 ppm.

AOE 1 O O
AOOE

(2)

k 21

AOO 1 E O
AOEO

(3)

2953

SOLUTION COPOLYMERIZATION OF ETHYLENE AND OCTENE-1 METALLOCENE CATALYST

Table I. Results of Continuous Solution Copolymerizations of Ethylene (1) and Octene-1 (2) at 140C

Run
(#)

f 20 a
(in mole)

[M 1 ] 0 b
(M)

[M 2 ] 0
(M)

[M 1 ]
(M)

[M 2 ]
(M)

f2
(in mole)

F 2c
(in mole)

MW
(310 24 )

MWD

Activity
(kg Polymer/g
Catalyst)

011
1
2
3
4
5
6
7
A1
A2

0
0.05
0.10
0.20
0.30
0.45
0.60
0.80

1.43
1.36
1.28
1.14
0.998
0.784
0.570
0.285

0
0.071
0.143
0.285
0.428
0.642
0.842
1.14

0.153
0.180
0.181
0.215
0.268
0.200
0.264
0.253
0.274
0.442

0
0.037
0.091
0.194
0.324
0.479
0.736
1.12
0.284
0.659

0
0.170
0.335
0.476
0.547
0.705
0.736
0.816
0.509
0.599

0
0.028
0.045
0.089
0.125
0.218
0.258
0.337
0.124
0.167

10.4
4.44
3.92
2.59
1.96
1.16
1.04
0.791

2.04
1.95
2.03
2.12
2.20
2.21
2.16
2.24

7.27
7.49
7.48
7.36
6.55
7.04
4.19
0.554

Activityd
(3 1024 kg
Polymer/[M]
z mol Ti)
1.56
1.13
0.901
0.589
0.362
0.339
0.137
0.013

f 20 , f 2 : molar fractions of octene-1 in feed and outlet;


[M1]0, [M2]0, [M 1 ], [M 2 ]: ethylene and octene-1 inlet and outlet concentration, respectively;
F 2 : molar fraction of octene-1 in the copolymer determined by 13C-NMR.
d
Activity: weight of copolymer/(total monomer concentrations 3 molcatalyst).
b
c

k 22

AOO 1 O O
AOOO

(4)

where AOE and AOO are the active centers with


the last inserted monomer unit being ethylene
and octene-1, respectively. k11, k12, k21, and k22
are their propagation rate constants. When the
CSTR system reached its steady-state operation,
the copolymerization rates (Rp) for each monomer
can be expressed as

R p1 5

@M 1# 0 2 @M 1#
5 ~k 11@A 1# 1 k 21@A 2#!@M 1#
t

(5)

R p2 5

@M 2# 0 2 @M 2#
5 ~k 22@A 2# 1 k 12@A 1#!@M 2#
t

(6)

where [M1]0, [M2]0, [M1] and [M2] are the inlet


and outlet concentrations of ethylene and octene-1, respectively. [A1] and [A2] are the concentrations of the active centers AOE and AOO,
respectively.
Under the long chain assumption, the cross
propagation rates [eqs. (2) and (3)] are equal,
k 12@A 1#@M 2# 5 k 21@A 2#@M 1#

(7)

Substituting eq. (7) into eqs. (5) and (6) yields


F2 5

@P 2#
@P 1# 1 @P 2#
5

Figure 3. Monomer conversions (h: ethylene, : octene-1, and F: total) versus the inlet octene-1 molar
fraction ( f20) during the copolymerization of ethylene
and octene-1 at 140C with the constrained geometry
catalyst.

r 2 f 22 1 f 2~1 2 f 2!
r f 1 2f 2~1 2 f 2! 1 r 1~1 2 f 2! 2
2
2 2

(8)

where [P1] and [P2] are the concentrations of monomeric units chemically bound to the copolymer
chains ([Pi] 5 [Mi]0 2 [Mi]), F2 is the molar fraction of octene-1 in copolymer and f2 (5[M2]/([M1]
1 [M2])) is the outlet molar fraction of octene-1.
The reactivity ratios are defined as r1 5 k11/k12
and r2 5 k22/k21.

2954

WANG ET AL.

The reactivity ratios of r1 5 7.90 and r2 5 0.099


were estimated by the GaussNewton algorithm.
Figures 4(a) and 4(b) present the fitting of experimental data and the 95% confidence region14 for
the reactivity ratios, respectively. Compared to
the reactivity ratios of other metallocene systems
for the copolymerization of ethylene with octene-1
or hexene-1, summarized in Table II, the higher
value of r2 confirms the ability of CGC-Ti to incorporate high a-olefin monomers.
The inverse of the reactivity ratio r1 (k12/k11)
gives a measure of the copolymerization tendency of
ethylene. For octene-1, the ratio is 0.127. This value
is higher but not appreciably different from that for
ethylene with ethylene macromonomer found in our
previous work (50.081).11 The macromonomers in
the ethylene homopolymerization were generated
in situ via transfer to ethylene and/or b-hydride
elimination of ethylene chains. Such macromonomers had molecular weights at the same levels as
polyethylene backbones, i.e., about 105. Considering
the tremendous difference in the molecular weights
between octene-1 and ethylene macromonomer, and
the similarity in their k12/k11 values, we are convinced that in the ethylene homopolymerization, an
ethylene macromonomer must be incorporated into
a polymer chain before it diffuses away from the
vicinity of an active center. Once a macromonomer
is generated either via transfer to ethylene or b-hydride elimination, it has two fates: be incorporated
into a growing polymer chain to form a branch or
diffuse out from the vicinity of the active center. We
believe that once a macromonomer with high molecular weight migrates away from an active center,
it is too difficult for the chain to go back to participate in further propagation. This is why LCB metallocene polyethylenes can be made by gas-phase

Figure 4. (a) Fitting of F2 versus f2 experimental


data (E) by eq. (8) with the estimated reactivity ratios
r1 5 7.90 and r2 5 0.099. (b) 95% confidence region
(solid line) for the reactivity ratios (1) of r1 5 7.90 and
r2 5 0.099 regressed using the experimental data.

Table II. A Summary of Reactivity Ratios in the Copolymerization Between


Ethylene (r 1 ) and Octene-1 or Hexene-1 (r 2 )
Catalyst

Temperature (C)

Comonomer

r1

r2

r1r2

[C5Me4(SiMe2NtBu)]TiMe2/TPFPB/MMAO
meso-Me2Si(2-Me-1-Ind)2ZrCl2/MAO5
rac-Me2Si(2-Me-1-Ind)2ZrCl2/MAO5
rac-Me2Si(Ind)2ZrCl2/MAO7
rac-Me2Si(2-MeInd)2ZrCl2/MAO7
rac-Me2(Si(Benz[e]Ind)2ZrCl2/MAO7
rac-Me2Si(2-MeBenz[e]Ind)2ZrCl2/MAO7
Cp2ZrCl2/MAO3
Et(IndH4)2ZrCl2/MAO3
i-Pr(Cp)(Flu)ZrCl2/MAO3

140
60
60
40
40
40
40
40
40
40

1-Octene
1-Octene
1-Octene
1-Octene
1-Octene
1-Octene
1-Octene
1-Hexene
1-Hexene
1-Hexene

7.90

18.9
19.5
10.7
10.1
21.9
12.9
5.6

0.099

0.014
0.013
0.076
0.118
;0
0.027
0.051

0.78
0.008
0.14
0.27
0.25
0.81
1.20

0.35
0.29

2955

SOLUTION COPOLYMERIZATION OF ETHYLENE AND OCTENE-1 METALLOCENE CATALYST

Table III. The Triad Distributions for Ethylene/Octene-1 Copolymers Synthesized in Continuous Solution
Copolymerizations at 140C

Run (#)

f1
(in molar)

P(EOE)
(%)

P(EOO 1 OOE)
(%)

P(OOO)
(%)

P(OEO)
(%)

P(OEE 1 EEO)
(%)

P(EEE)
(%)

1
2
3
4
5
6
7
A1
A2

0.170
0.335
0.476
0.547
0.705
0.736
0.816
0.509
0.599

2.66
3.73
7.04
9.08
12.6
14.4
11.4
9.76
11.1

0.193
0.692
2.30
3.84
10.5
13.2
17.7
4.03
6.55

0.002
0.013
0.099
0.267
1.05
1.56
3.47
0.203
0.413

0.050
0.309
0.707
1.36
3.18
4.23
7.93
1.03
2.02

3.61
8.63
11.9
15.8
21.3
23.4
26.7
14.1
19.0

93.5
86.6
78.0
69.6
51.4
43.3
32.8
70.8
60.9

and slurry processes. The ethylene macromonomers, once generated, are incorporated into polymer chains to form branches before they diffuse out
of the active center reaction volume.

k 221

AOEOO
AOOO 1 E O

(15)

k 222

AOOO 1 O O
AOOOO

Penultimate Effect
In a copolymerization process, if the reactivity of
an active center is determined by both terminal
and penultimate units, the propagation reactions
should be written as

(16)

and the four reactivity ratios are


r 1p 5 k 111/k 112

(17)

r91p 5 k 211/k 212

(18)

r 2p 5 k 222/k 221

(19)

r92p 5 k 122/k 121

(20)

k 111

AOEE 1 E O
AOEEE

(9)

k 112

AOEE 1 O O
AOOEE

(10)

k 211

AOEO 1 E O
AOEEO

(11)

k 212

AOOEO
AOEO 1 O O

(12)

k 121

AOOE 1 E O
AOEOE

(13)

k 122

AOOE 1 O O
AOOOE

(14)

where A-ji is the active center with the j terminal


and i penultimate units, kijk is the rate constant of
monomer k to insert the active center A-ji. rjp and
r9jp are the reactivity ratios for the active centers
of j terminal unit with j and i penultimate unit,
respectively.
Note that the pair of EOO and OOE triads, as
well as EEO and OEE, are indistinguishable from
each other. We can only estimate their sums.
With the long chain assumption, the amounts of
EOO and EEO triads are the same as those of
OOE and OEE, respectively. The relationships
between the reactivity ratios and the triad distributions, based on the mass balances for the
steady-state CSTR system, are thus established
as described by eqs. (21)(24).

2956

WANG ET AL.

Figure 5. Dependence of the triad distribution ratios on the monomer ratio of ethylene and octene-1 for a series of copolymerizations carried out with different monomer
ratios at 140C. (a) 2P(EEE)/P(OEE 1 EEO) versus (12f2)/f2 for the estimate of r 1p , (b)
P(OEE 1 EEO)/P(OEO) versus 2(1-f2)/f2 for the estimate of r91p (c) 2P(OOO)/P(OOE
1 EOO) versus f2/(12f2) for the estimate of r2 p, (d) P(OOE1EOO)/P(EOE) versus
2 f2/(12f2) for the estimate of r92p.

r 1p 5

2f 2P(EEE)
~1 2 f 2!P~OEE 1 EEO!

(21)

r91p 5

f 2P~OEE 1 EEO!
2~1 2 f 2!P~OEO)

(22)

r 2p 5

2~1 2 f 2!P~OOO!
f 2P~OOE 1 EOO!

(23)

r92p 5

~1 2 f 2!P~OOE 1 EOO!
2f 2P~EOE!

(24)

triad distributions for the copolymer samples in


this work.
The plots of the triad contributions versus
monomer fractions are shown in Figure 5(a d).
The slopes are the reactivity ratios. The reactivity
ratios and their 95% confidence ranges were calculated as:
r 1p 5 10.6 6 0.41

where P(ijk) are the triad ijk distributions from


the 13C-NMR spectra.15 Table III provides the

r91p 5 7.39 6 0.21


r 2p 5 0.088 6 0.009

SOLUTION COPOLYMERIZATION OF ETHYLENE AND OCTENE-1 METALLOCENE CATALYST

r92p 5 0.175 6 0.012


The fact that r91p , r1p and r2p , r92p indicates that
the activities of active centers with octene-1 penultimate unit are lower than those with ethylene
penultimate unit. However, it also shows that the
penultimate effects are minor because the values
of r1p and r91p, as well as r2p and r92p, are rather
close to each other. The copolymerization seems
to follow first-order Markov statistics. Even after
the insertion of a high a-olefin, the active center is
still highly active for the incorporation of other
monomers. This explains why an ethylene monomer can be inserted into the active center in succession after the incorporation of a macromonomer to form a trifunctional LCB.

CONCLUSION
Ethylene and octene-1 copolymers with narrow
MWDs of about 2 and high octene-1 contents (octene-1 mole fraction up to F1 5 0.337) were produced with the CGC-Ti catalyst system using our
high-temperature
high-pressure
continuous
stirred-tank reactor. The CGC-Ti catalyst appeared to be a single site type for the copolymerization and showed high incorporation rates for
high a-olefin. The activity of CGC-Ti decreased
with increase in octene-1 concentration, particularly at high octene-1 concentration level. With
the increase of octene-1 feed ratio, the molecular
weight decreased and MWD also broadened. The
ethylene and octene-1 copolymerization with
CGC-Ti followed first-order Markov statistics.
The estimated reactivity ratios were r1 5 7.90 and
r2 5 0.099. The penultimate effect was also investigated using 13C-NMR measurements of triad
distributions. The reactivity ratios were r1p
5 10.6, r91p 5 7.39 for ethylene and r2p 5 0.088, r92p
5 0.175 for octene-1 terminal units, respectively.
These values demonstrated the existence of a minor penultimate effect in the ethylene/octene-1
copolymerization under the present experimental
conditions. It was also found that the value of
k12/k11 5 0.127 for the ethylene/ octene-1 system
is comparable to that of k12/k11 5 0.081 for the
ethylene/macromonomer system found in our pre-

2957

vious work.11 This indicates that the long chain


branching in ethylene homopolymerization was
via an instantaneous incorporation of ethylene
macromonomer right after it was generated and
before it diffused out from the active center reaction volume.
The authors thank the National Science and Engineering Research Council of Canada (NSERC) for support
for this research. Thanks to Prof. Brian G. Sayer for
help in NMR measurements.

REFERENCES AND NOTES


1. Michelotti, M.; Altomare, A.; Ciardelli, F.; Ferrarini, P. Polymer 1996, 37, 5011.
2. Chung, T. C.; Lu, H. L. J Polym Sci Part A Polym
Chem 1998, 36, 1017.
3. Uozumi, T.; Soga, K. Makromol Chem 1992, 193,
823.
4. Chien, J. C. W.; Nozaki, T. J Polym Sci Part A
Polym Chem 1993, 31, 227.
5. Uozumi, T.; Miyazawa, K.; Sano, T.; Soga, K. Macromol Rapid Commun 1997, 18, 883.
6. Schneider, M. J.; Mulhaupt, R. Macromol Chem
Phys 1997, 198, 1121.
7. Schneider, M. J.; Suhm, J.; Mulhaupt, R.; Prosenc,
M.-H.; Brintzinger, H.-H. Macromolecules 1997,
30, 3164.
8. Lai, S. Y.; Wilson, J. R.; Knight, G. W.; Stevens,
J. C.; Chum, P. W. S. (to Dow Chemical Co.). Elastic
substantially linear olefin polymers, U.S. Patent
5,272,236, 1993.
9. Wang, W.-J.; Yan, D.; Charpentier, P. A.; Zhu, S.;
Hamielec, A. E.; Sayer, B. G. Macromol Chem Phys
1998, 199, 2409.
10. Yan, D.; Wang, W.-J.; Zhu, S. Polymer 1999, 40,
1737.
11. Wang, W.-J.; Yan, D.; Zhu, S.; Hamielec, A. E.
Macromolecules 1998, 31, 8677.
12. Charpentier, P. A.; Zhu, S.; Hamielec, A. E.; Brook;
M. A. Ind Eng Chem Res 1997, 36, 5074.
13. ASTM D 5017-96, Determination of Linear Low
Density Polyethylene (LLDPE) Composition by
Carbon-13 Nuclear Magnetic Resonance, In 1998
Annual Book of ASTM Standards, Volume 08.03,
1998; p. 286.
14. Behnken, D. W. J Polym Sci Part A 1964, 2, 645.
15. Randall, J. C. Rev Macromol Chem Phys 1989,
C29, 201.

You might also like