You are on page 1of 13

Kinetic and Microstructure Studies of

Poly(ethylene-co-p-methylstyrene) Copolymers
Prepared by Metallocene Catalysts with
Constrained Ligand Geometry
T. C. CHUNG, H. L. LU
Department of Materials Science and Engineering, The Pennsylvania State University,
University Park, Pennsylvania 16802

Received 12 June 1997; accepted 27 October 1997

ABSTRACT: This paper discusses the poly(ethylene-co- p-methylstyrene) copolymers

prepared by metallocene catalysts, such as Et(Ind)2ZrCl2 and [C5Me4 (SiMe2N tBu)]TiCl2 , with constrained ligand geometry. The copolymerization reaction was examined
by comonomer reactivity (reactivity ratio and comonomer conversion versus time),
copolymer microstructure (DSC and 13C-NMR analyses) and the comparisons between
p-methylstyrene and other styrene-derivatives (styrene, o-methylstyrene and m-methylstyrene). The combined experimental results clearly show that p-methylstyrene
performs distinctively better than styrene and its derivatives, due to the cationic coordination mechanism and spatially opened catalytic site in metallocene catalysts with
constrained ligand geometry. A broad composition range of random poly(ethylene-cop-methylstyrene) copolymers were prepared with narrow molecular weight and composition distributions. With the increase of p-methylstyrene concentration, poly(ethylene-co- p-methylstyrene) copolymer shows systematical decrease of melting point and
crystallinity and increase of glass transition temperature. At above 10 mol % of pmethylstyrene, the crystallinity of copolymer almost completely disappears. q 1998 John
Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 10171029, 1998

INTRODUCTION
For many decades since the commercialization of
PE and PP, the functionalization of polyolefins 1 has
been a long scientific interest and technological important research subject to improve their adhesion
to and compatibility with other materials. Unfortunately, the chemistry to prepare functional polyolefins is very limited both in the direct and postpolymerization processes, namely due to the catalyst poison 2 and the inert nature 3 of polyolefins.
Recently, several new research approaches, including the use of late transition metals 4 and protecting
groups, 5 may provide some solutions.
Correspondence to: T. C. Chung
Journal of Polymer Science: Part A: Polymer Chemistry, Vol. 36, 10171029 (1998)
q 1998 John Wiley & Sons, Inc.
CCC 0887-624X/98/061017-13

Our functionalization chemistry has been focusing on the reactive polyolefin approach, especially the polyolefin containing borane groups.6
The borane groups can be effectively incorporated
into polyolefin by both direct 7 and post 8 polymerization processes. In turn, borane groups in polymer can be easily interconverted to various functional groups, such as {OH and halides, under
mild reaction conditions. Recently, the borane
group has been transformed to a stable radical
initiator 9 which can then carry out living radical
graft-from polymerization of monomer containing
functional group. In other words, the reactive borane group serves as the intermediate for the
transformation reaction from transition metal to
living radical polymerizations. Such a process
produces graft 10 and block 11 copolymer struc1017

/ 8G6C$$109T

02-19-98 09:38:28

polca

W: Poly Chem
97109T

1018

CHUNG AND LU

tures, containing both polyolefin and functional


polymer segments. The block and graft copolymers are preferred molecular structures which offer not only high concentration of functional
groups but also preserve the desirable polyolefin
properties (such as crystallinity, melting point
and glass transition temperature). The resulting
graft and block copolymers were shown as the effective compatibilizers to improve the adhesion 12
and compatibility 13 of polyolefins in polyolefin
blends and composites.
Recently, our reactive polyolefin approach
has extended to another type reactive comonomer,
i.e. p-methylstyrene (p-MS) groups.14 The major
advantages of p-MS are due to its commercial
availability, easy incorporation into polyolefin
and versatility 15 in the functionalization chemistry under various reaction mechanisms, including
free radical, cationic and anionic processes. The
benzylic protons are known to be facile in many
chemical reactions, such as halogenation, 16 metallation 17 and oxidation, 18 to form desirable functional group at the benzylic position under mild
reaction conditions. In addition, the benzylic protons can also be interconverted to stable anionic
initiator for living anionic graft-from polymerization.19
In general, the preparation of reactive polyolefin copolymers have been greatly enhanced by
the exciting progresses in the metallocene technology.20 The well-defined metallocene catalyst
with constrained ligand geometry, having spatially opened catalytic site, allow the effective incorporation of large comonomers. Several important examples are LLDPE copolymers 21 with
well-defined composition and molecular weight
distributions, Dows long chain branching polyethylene 22 and poly(ethylene-co-styrene) copolymer with pseudo-random23 and alternating microstructures.24

x CHCH 1 y CHCH

Metallocene
Catalyst

CH
(CHCH)x(CHCH)y

CH

Three metallocene catalysts, including a simple


Cp2ZrCl2 catalyst and two bridged ones (i.e.
Et(Ind)2ZrCl2 and [C5Me4 (SiMe2N tBu)]TiCl2 ) illustrated in Scheme 1, representing three generations of metallocene technology, were used in this
copolymerization study. Catalyst (I), without
bridge between Cp ligands, has the zirconium active site sandwiched between two Cp rings, with
Cp-Zr-Cp angle of 1807. In catalyst (II), ethylene
bridge induces the constrained indenyl ligand geometry, with Cp-Zr-Cp angle 25 of 125.87. On the
other hand, silicon bridge in catalyst (III) pulls
back both Cp and amido ligands from normal positions to form a highly constrained ligand geometry, with Cp-Ti-N angle 26 of 107.67. The spatial
opening of the catalytic site shall be (III) (II)
(I). Based on the structure/activity relationships of the metallocene catalysts, it is logical to
predict that the incorporation of p-MS will follow
the same trend of spatial opening. The bridged
catalysts (II) and (III) with the spatially opened
active sites will be more favorable than the nonbridged catalyst (I), and the one-membered silicon bridged catalyst (III) will be preferable over
the two-membered ethylene bridge in (II). Each
copolymerization reaction was systematically
evaluated by studying the conversion of p-MS versus reaction time, comonomer reactivity ratios at
low monomer conversion and molecular structure
of the resulting poly(ethylene-co- p-methylsty-

RESULTS AND DISCUSSION


In this paper, we will discuss the copolymerization reaction of ethylene and p-methylstyrene as
shown in Equation 1, especially, focusing on the
kinetic aspects of copolymerization reaction and
the resulting molecular structure of poly(ethylene-co-p-methylstyrene) copolymers.

/ 8G6C$$109T

02-19-98 09:38:28

(1)

Scheme 1.

polca

W: Poly Chem
97109T

POLY(ETHYLENE-CO-p-METHYLSTYRENE)

1019

Figure 1. GPC curves of (a) polyethylene and three poly(ethylene-co- p-methylstyrene) copolymers, containing (b) 1.08, (c) 9.82 and (d) 18.98 mol % of p-MS comonomers, prepared by [C5Me4 (SiMe2N tBu)]TiCl2 catalyst, and the molecular weights and
molecular weight distributions (insert).

rene) copolymers. In addition, p-MS was compared with styrene and other methylstyrene derivatives (i.e. o-methylstyrene and m-methylstyrene).
Synthesis of Poly(ethylene-co-p-methylstyrene)
In a typical copolymerization, the reaction was
started by the addition of the metallocene catalyst
mixture to a solution of the two monomers in solvent under an inert gas atmosphere. The appearance of the reacting polymer solution was very
dependent on the quantity of p-MS used. In the
high p-MS cases, a homogeneous solution was observed through the whole copolymerization reaction. In the low p-MS cases, a slurry solution with
white precipitates was observed right in the beginning of the reaction. The precipitation is obviously due to the crystallinity of copolymer which
has long ethylene sequences. The copolymer was
isolated by filtering and washed completely with
MeOH and dried under vacuum at 507C for 8 h.
The copolymer was firstly analyzed by GPC and
1
H-NMR measurements. Figure 1 shows the typical GPC curves of the homo- and copolymers prepared by [C5Me4 (SiMe2N tBu)]TiCl2 catalyst. The

/ 8G6C$$109T

02-19-98 09:38:28

uniform molecular weight distribution in all samples, with MV w /MV n 23, implies the single-site
polymerization mechanism. In fact, the GPC
curves show a slight reduction of molecular
weight distribution in the copolymers, from MV w /
MV n 2.86 in PE to 1.68 in poly(ethylene-co- pmethylstyrene) containing 18.98 mol % of p-MS.
The similar narrow molecular distribution results
were also observed in the copolymers prepared by
Et(Ind)2ZrCl2 catalyst. The better diffusibility of
monomers in the copolymer structures (due to
lower crystallinity) may help to provide the ideal
polymerization condition. It is very interesting to
note that the average molecular weight of copolymers maintain very high throughout the entire
composition range, which may be attributed to the
relatively high reactivity of p-MS.
Figure 2 shows the 1H-NMR spectra of two
poly(ethylene-co-p-methylstyrene) copolymers,
containing 2.11 and 18.98 mol % of p-MS, respectively. In addition to the major chemical shift at
1.35 ppm, corresponding to CH2 , there are three
minor chemical shifts around 2.35, 2.5 and 7.0
7.3 ppm, corresponding to CH3 , CH and aromatic
protons in p-MS units. The integrated intensity
ratio between the chemical shift at 1.35 ppm and

polca

W: Poly Chem
97109T

1020

CHUNG AND LU

Figure 2. 1H-NMR spectra of poly(ethylene-co- p-methylstyrene) copolymers, containing (a) 2.11 and (b) 18.96 mol % of p-MS comonomers.

the chemical shifts between 7.0 and 7.3 ppm and


the number of protons both chemical shifts represent determines the concentration of p-MS.
Comparison Among Metallocene Catalysts
Table I summarizes the copolymerization results,
including all three catalyst systems. The copolymerization efficiency follows the sequence of [C5Me4(SiMe2N tBu)]TiCl2 Et(Ind)2ZrCl2 Cp2ZrCl2 ,
which is directly relative to the spatial opening
at the active site. Between the comparable runs
P-363/P-358 and p-365/p-360, Cp2ZrCl2 (I) and
Et(Ind)2ZrCl2 (II) catalysts produced the copolymers with 2.2/5.16 and 1.84/4.76 mol % of p-MS

/ 8G6C$$109T

02-19-98 09:38:28

concentrations, respectively, and 26.2/51.1 and


7.72/32.8% overall p-MS conversions, respectively. On the other hand, the [C5Me4 (SiMe2N tBu)]TiCl2 catalyst offers the best of the three catalysts
in the preparation of poly(ethylene-co- p-methylstyrene) copolymers. In run P-377, about 90% of
p-MS was incorporated into copolymer in 1 h. In
run P-383, the reaction produced the copolymer containing 40 mol % of p-MS, which is close to the ideal
50 mol % (as will be discussed in the reactivity ratio
and copolymer microstructure studies, the consecutive insertion of p-MS in all three catalyst systems
is almost impossible). High concentration of p-MS
in copolymer and high p-MS conversion were
achieved by catalyst (III) which must provide excel-

polca

W: Poly Chem
97109T

/ 8G6C$$109T

02-19-98 09:38:28

polca

II/17
II/17
II/17
II/17
II/17

p-357
p-392
p-360
p-362
p-375
45
45
45
45
45
45
10

45
45
45
45
45

45
45
45
45
45

45
45

Ethylene
(psi)c

0
0.447
0.912
0.447
0.912
1.82
1.82

0.085
0.456
0.678
1.36
2.03

0.085
0.508
0.678
1.36
2.03

0.678
0.678

p-MS
(mol/L)

Toluene
Hexane
Hexane
Toluene
Toluene
Toluene
Hexane

Toluene
Toluene
Toluene
Toluene
Toluene

Hexane
Hexane
Hexane
Hexane
Hexane

Hexane
Toluene

Solvent

30
30
30
30
30
30
30

50
50
50
50
50

50
50
50
50
50

50
50

Temperature
(7C)

4.4
12.0
15.5
13.0
17.4
24.2
15.9

5.70
8.63
15.1
19.5
19.4

6.5
19.8
21.9
18.9
20.8

24.2
8.44

Yield
(g)

440.0
1200.0
1550.0
1300.0
1740.0
2420.0
1590.0

335.3
507.6
888.2
1147.1
1141.2

382.4
1164.7
1288.2
1111.8
1223.5

1423.5
496.5

Catalyst
Efficiency
(kg P/mol Mrh)

13.5
22.6
10.9
21.6
32.8
40.0

1.30
3.85
4.76
6.36
8.49

1.83
3.94
5.16
7.20
8.94

2.20
1.84

p-ms in
Copolymer
(mol %)

90.3
81.3
83.8
86.9
75.8
54.6

29.9
23.2
32.8
27.0
22.8

47.2
48.7
51.1
29.0
25.4

26.2
7.72

Conversion
of p-MS
(%)

I: Cp2ZrCl2/MAO; II: Et(Ind)2ZrCl2/MAO; III: [(C5Me4)SiMe2N(t-Bu)]TiCl2/MAO.


Polymerization conditions: Al/Zr, Al/Ti 1500; solvent: 100 mL; reaction time: 1 h.
c
45 psi ethylene 0.309 mol/L in toluene, 0.424 mol/L in hexane at 507C; 0.398 mol/L in toluene, 0.523 mol/L in hexane at 307C. 10 psi ethylene 0.116 mol/L in
hexane at 307C.

III/10
III/10
III/10
III/10
III/10
III/10
III/10

II/17
II/17
II/17
II/17
II/17

p-356
p-372
p-358
p-361
p-371

p-270
p-377
p-378
p-267
p-379
p-380
p-383

I/17
I/17

Catalysta
(mmol)

Summary of The Copolymerization Reactionsb Between Ethylene and p-Methylstyrene

p-363
p-365

Run
Number

Table I.

POLY(ETHYLENE-CO-p-METHYLSTYRENE)

W: Poly Chem

97109T

1021

1022

CHUNG AND LU

lent spatial freedom for p-MS to access to the propagating chain end.
In general, the catalyst activity systematically
increases with the increase of p-MS content,
which was also observed in the 1,4-hexadiene copolymerization reactions 8 and could be a physical
phenomenon relative to the improvement of
monomer diffusion in the lower crystalline copolymer structures. In [C5Me4 (SiMe2N tBu)]TiCl2
case, the catalyst activity attains a value of more
than 2.4 1 10 6 g of copolymer/mol of Ti 1 h in
run P-380, which is about 6 times the value for
the homopolymerization of ethylene in run P-270
under similar reaction conditions. It is very interesting to note that a very small solvent (hexane
and toluene) effect to the catalyst (III) activity
was observed in the comparative runs (P-377/P267) and (P-378/P-379), despite the significant
difference in the beginning of reaction conditions
(heterogeneous in hexane and homogeneous in
toluene). However, the solvent effect is very significant in both catalyst (I) and (II) systems. All
comparative reaction pairs (P-363/P-365, P-356/
P-357, P-358/P-360, P-361/P-362 and P-371/P375), carried out under similar reaction conditions, consistently show higher p-MS incorporation in hexane solution. The explanation of solvent effect is not clear.

min by using [C5Me4 (SiMe2N tBu)]TiCl2 catalyst.


In every monomer feed ratio, p-MS consistently
shows more than 30% higher incorporation than
the corresponding styrene.
Table II summarizes the experimental results,
including a series of comparative copolymerization reactions at various reaction temperatures.
In each side-by-side comparisons (runs p-451/442
and p-463/p-443 at 207C, runs p-418/p-412 and
p-416/p-411 at 407C and runs p-448/p-433, p-444/
p-434 and p-446/p-436 at 607C, respectively), pMS all shows better incorporation than styrene.
The significantly higher p-MS incorporation must
due to the electronic donation of p-methyl group,
which is favorable in the cationic polymerization
mechanism.27 Sterically, the methyl group at
para-substitution doesnt effect the monomer insertion.
The p-MS incorporation was also compared
with its derivatives, o-MS and m-MS. Table III
compares the copolymerization results which
were carried out in toluene at 307C for 60 min
by using [C5Me4 (SiMe2N tBu)]TiCl2 catalyst. All
comonomers show no retardation to catalyst activity. However, p-MS consistently shows better incorporation than o-MS and m-MS. Both isomers
(o-MS and m-MS) were not receiving the full benefits of the electronic and steric effects existed
in p-MS.

Comparison Among Styrene Derivatives


In the evaluation of p-MS comonomer, it is very
interesting to compare p-MS with styrene and
methylstyrene isomers. The consumption of comonomer during the reaction is an useful way to
understand the dynamic of copolymerization reaction. Figure 3(a) shows a comparative plots of
p-MS and styrene incorporation versus reaction
time in the batch copolymerization reaction of ethylene (29 psi) and comonomer (0.356 mol/L), using [C5Me4 (SiMe2N tBu)]TiCl2 catalyst at 607C in
toluene. The conversion of comonomer was calculated from the yield of copolymer and the concentration of comonomer in copolymer (determined
by 1H-NMR). The significant better p-MS incorporation starts in the beginning of copolymerization reaction. After 30 min, the difference became
relatively constant due to the depletion of p-MS.
About 1 h, more than 80% of p-MS and only less
than 60% of styrene were incorporated in the ethylene copolymers. Figure 3(b) compares the incorporated comonomer concentration (mol %) in copolymer versus comonomer ratio. Each copolymerization reaction was carried out at 407C for 15

/ 8G6C$$109T

02-19-98 09:38:28

Reactivity Ratios
The best way to investigate a copolymerization is
to measure the reactivity ratio of the comonomers.
To obtain meaningful results, a series of experiments were carried out by varying monomer feed
ratio and comparing the resulting copolymer composition at low monomer conversion ( 10%). The
reactivity ratios between ethylene ( r1 k11 /k12 )
and p-MS (r2 k22 /k21 ) are estimated by the
KelenTudos method. The calculation 28 is based
on the equation 2.

h r1j 0 r2 / a(1 0 j )
h G/ a / F and j F/ a / F
where x [ethylene]/[comonomer] in feed and y
d[ethylene]/d[comonomer] mol ratio in polymer, G x(y 0 1)/y, F x 2 /y, a (Fm 1 FM ) 1/2 ,
and Fm and FM are the lowest and highest values
of F. Figure 4 shows a typical plot h versus j and
the least squares best fit line. In general, a good
fit of the experimental results in the straight line

polca

W: Poly Chem
97109T

POLY(ETHYLENE-CO-p-METHYLSTYRENE)

1023

Figure 3. Plots of (a) comonomer conversion versus reaction time and (b) comonomer
concentration in copolymer versus comonomer ratio, during the copolymerization reactions of ethylene with p-methylstyrene and with styrene using [C5Me4 (SiMe2N tBu)]TiCl2 catalyst.

was observed in most cases. The extrapolation to


j 0 gives 0r2 / a and r1 . Table IV summarizes
the calculated values of r1 and r2 .
In [C5Me4 (SiMe2N tBu)]TiCl2 cases, high r1 (r1
19.6 and 21.4 for 20 and 607C, respectively)

and low r2 (r2 0.04 and 0.08 for 20 and 607C,


respectively) indicate the strong tendency of ethylene consecutive insertion and very low possibility in continuous p-MS insertion. The values of r1
1 r2 are close to unity in both temperatures, which

Table II. The Comparison of Copolymerization Reactions Between Ethylene (M1) and p-Methylstyrene or
Styrene (M2) Using [(C5Me4)SiMe2N(t-Bu)]TiCl2/MAO Catalysta

Run
Number

Reaction
Temperature
(7C)

Reaction
Time
(min.)

Ethylene [M1]
(mol/L)

[M2]b Comonomer
(mol/L)

Yield
(g)

M2 Concentration
(mol %)

p-451
p-463
p-442
p-443

20
20
20
20

60
60
60
60

0.294
0.294
0.294
0.294

p-MS/0.890
p-MS/1.186
St./0.894
St./1.192

0.726
0.667
0.425
0.454

13.1
16.6
8.2
10.6

p-418
p-416
p-412
p-411

40
40
40
40

15
15
15
15

0.22
0.22
0.22
0.22

p-MS/0.22
p-MS/0.88
St./0.22
St./0.88

2.557
3.130
1.463
2.438

5.40
18.96
2.39
12.38

p-448
p-444
p-446
p-433
p-434
p-436

60
60
60
60
60
60

15
15
15
15
15
15

0.178
0.178
0.178
0.178
0.178
0.178

p-MS/0.093
p-MS/0.178
p-MS/0.534
St./0.096
St./0.183
St./0.548

3.22
3.36
3.10
2.44
3.15
3.75

2.34
4.77
12.24
0.95
2.06
8.56

a
b

[Ti] 2.5 1 1005 mol/L, Al/Ti 3,000, Solvent: 100 mL toluene.


p-MS p-methylstyrene, St. styrene.

/ 8G6C$$109T

02-19-98 09:38:28

polca

W: Poly Chem
97109T

1024

CHUNG AND LU

Table III. The Comparison of Copolymerization Reactionsa Between Ethylene [M1] and Methylstyrene
Derivatives [M2] Using [(C5Me4)SiMe2N(t-Bu)]TiCl2/MAO Catalyst
Run
Number

M2b
(mmol)

Yield
(g)

M2 Concentration
(mol %)

M2 Conversion
(%)

Tm
(7C)

xc
(%)

p-277
p-267
p-275
p-274

none
p-MS (46.6)
o-MS (46.6)
m-MS (46.6)

4.27
13.0
12.9
5.43

0
11.0
4.52
2.36

84.3
38.7
9.53

133.7
76.0
98.3
119.1

51.5
5.40
12.1
23.4

a
b

Catalyst: 10 umol, reaction temp.: 307C, reaction time: 60 min. ethylene (M1) pressure: 45 psi.
p-MS: p-methylstyrene; o-MS: o-methylstyrene; m-MS: m-methylstyrene.

suggest the near ideal random copolymerization


reactions and very small probability to find two
adjacent p-MS units in the polymer chain. In
other words, the p-MS units shall be homogeneously distributed in the polymer chain. In
Et(Ind)2ZrCl2 cases, the copolymerization reactions exhibit even higher r1 (r1 60), very strong
favorable for ethylene incorporation, and almost
no possibility of p-MS consecutive insertion (r2
0). The less opened active site in Et(Ind)2ZrCl2

catalyst may sterically prohibit p-MS consecutive


insertion.
Molecular Structure of Copolymers
The other way to study the copolymerization reaction is to examine the copolymer molecular structure. Both DSC (differential scanning calorimetry) and 13C-NMR measurements were carried out
to evaluate the macroscopic and microscopic

Figure 4. KelenTudos plots for copolymerization of ethylene/ p-methylstyrene at


207C by using [C5Me4 (SiMe2N tBu)]TiCl2 catalyst.

/ 8G6C$$109T

02-19-98 09:38:28

polca

W: Poly Chem
97109T

POLY(ETHYLENE-CO-p-METHYLSTYRENE)

1025

Table IV. A Summary of Reactivity Ratios in The Copolymerization Reactions


Between Ethylene [r1] and p-Methylstyrene [r2]

Catalyst

Reaction
Temperature
(7C)

r1

r2

r 1 1 r2

[(C5Me4)SiMe2N(t-Bu)]TiCl2
[(C5Me4)SiMe2N(t-Bu)]TiCl2
Et(Ind)2ZrCl2
Et(Ind)2ZrCl2
Et(Ind)2ZrCl2

20
60
20
40
60

19.6
21.4
89.0
72.0
64.1

0.039
0.082
0
0
0

0.77
1.76

structures of the copolymers. Figure 5 shows the


comparison of DSC curves between PE homopolymer and poly(ethylene-co-p-methylstyrene) copolymers prepared by [C5Me4 (SiMe2N tBu)]TiCl2
catalyst at 407C in toluene. Even a small amount
( 1 mol %) of p-MS comonomer incorporation
has significant effect to the crystallization of polyethylene. Overall, the melting point ( Tm ) and
crystallinity ( xc ) of copolymer are strongly relative to the density of comonomer, the higher the
density the lower the Tm and xc . Only a single
peak is observed throughout the whole composition range and the melting peak completely disappears at 10 mol % of p-MS concentration. The
similar general trend was also observed in the
DSC curves of poly(ethylene-co- p-methylstyrene) copolymers prepared by Et(Ind)2ZrCl2 catalyst. The systematic decrease of Tm and uniform
reduction of crystalline curve imply the homogeneous reduction of PE consecutive sequences. It
is interesting to note that the Tm peak (ranging
80457C shown in Fig. 5e) of the copolymer con-

Figure 5. The comparison of DSC curves between (a)


polyethylene homopolymer and poly(ethylene-co- pmethylstyrene) copolymers with (b) 1.08, (c) 2.11, (d)
5.40, (e) 9.82 and (f ) 18.98 mol % of p-MS comonomer.

/ 8G6C$$109T

02-19-98 09:38:28

taining 9.82 mol % of p-MS covers the similar Tm


range (45557C) of paraffin wax. This poly(ethylene-co-p-methylstyrene) copolymer is consisted
of in average 2022 consecutive methylene units,
which is in the molecular weight range of solid
paraffin wax.
The removal of crystallinity provides the opportunity to obtain elastic properties in many thermoplastic polymers. Figure 6 shows the glass
transition temperature (Tg ) of poly(ethylene-cop-methylstyrene) copolymers with 18.98, 32.8
and 40 mol % p-MS, respectively. The lowest observed Tg (Fig. 6, a) is 05.77C in the copolymer
containing 18.98 mol % p-MS. With the increase
of p-MS, the Tg of copolymer systematically increases as shown in Figures 6(b) and (c). The
single Tg with sharp thermal transition is indicative of homogeneous copolymer structure. It is
clear that poly(ethylene-co-p-methylstyrene) copolymer is difficult to be a good elastomer due to
the high Tg of poly(p-methylstyrene). To achieve
the elastic polymer, the system requires a third
monomer which can provide low Tg . The detail

Figure 6. DSC curves of (a) poly(ethylene-co- pmethylstyrene) copolymers with (a) 18.98, (b) 32.8 and
(c) 40 mol % of p-MS comonomer.

polca

W: Poly Chem
97109T

1026

CHUNG AND LU

Figure 7. (top) The comparison of two DSC curves


of poly(ethylene-co- p-methylstyrene) copolymers (containing 1 mol % of p-MS comonomer) prepared by (a)
[C5Me4 (SiMe2N tBu)]TiCl2 and (b) Et(Ind)2ZrCl2 catalysts. (bottom) The comparison of DSC curves between
two polyethylene copolymers containing 3 mol % (a) pmethylstyrene and (b) styrene comonomers, respectively, prepared by [C5Me4 (SiMe2N tBu)]TiCl2 catalyst.

results of elastic poly(ethylene-ter-propylene-terp-MS) and poly(ethylene-ter-1-octene-ter- p-MS)


terpolymers will be discussed in the near future.
It is very interesting to compare the DSC
curves of the copolymers, having similar comonomer concentration but prepared under different reaction conditions (i.e. catalyst, comonomer and temperature). In general, the temperature effect (ranging between 20 and 607C) is very
small, but the catalyst and comonomer effects are
significant. Figure 7 (top) compares the DSC
curves of two poly(ethylene-co- p-methylstyrene)
copolymers (containing 1 mol % of p-MS) prepared
by
[C5Me4 (SiMe2N tBu)]TiCl2
and
Et(Ind)2ZrCl2 , respectively, at 407C in toluene for
15 min. All results, including lower Tm (118.88
vs. 123.067C), smaller xc (81.99 vs. 98.58 J/g) and
sharper melting peak in Figure 7a (top), suggest

/ 8G6C$$109T

02-19-98 09:38:28

the more uniform distribution of p-MS units in


the copolymer prepared by [C5Me4 (SiMe2N tBu)]TiCl2 . This result is consistent with the random
copolymerization discussed in the reactivity ratio
studies. Figure 7 (bottom) compares the DSC
curves of poly(ethylene-co-p-methylstyrene) and
poly(ethylene-co-styrene) copolymers containing
3 mol % of comonomer (i.e. p-MS and styrene),
which
are
prepared
by
the
same
[C5Me4 (SiMe2N tBu)]TiCl2 catalyst at 507C in
hexane. The significant differences of Tm (111.9
vs. 117.27C) imply that p-MS comonomers are
more effectively to prevent the crystallization of
polyethylene sequences, possibly more uniformly
incorporated in copolymer, than styrene comonomers.
The molecular scale sequence distribution can be
quantitatively determined by 13C-NMR measurements. Figure 8 shows the 13C-NMR spectrum (with
the expanded aliphatic region) of poly(ethylene-cop-methylstyrene) copolymer containing about 10
mol % p-MS. At this concentration level, the PE
copolymer losses all its crystallinity, as shown in
DSC result. It is logical to expect that the methyl
group substitution at the para-position will have
very little effect on the chemical shifts of methylene
and methine carbons in the polymer backbone. In
other words, the aliphatic chemical shifts of poly(ethylene-co-p-methylstyrene) can be direct compared with that 23,24,29 of poly(ethylene-co-styrene).
In general, fewer chemical shifts shown in the poly(ethylene-co-p-methylstyrene) sample imply more
homogeneous copolymer microstructure, and every
chemical shift can be easily assigned to polyethylene and the isolated p-MS unit. In addition to the
two chemical shifts (21.01 and 29.80 ppm), corresponding to the methyl carbon from p-methylstyrene and methylene carbons from ethylene, respectively, there are three well-resolved peaks (27.74,
37.04 and 45.77 ppm) corresponding to methylene
and methine carbons from p-methylstyrene units
which are separated by multiple ethylene units
along the polymer chain. On the other hand, the
spectrum of poly(ethylene-co-styrene) shows much
more complicated methylene and methine carbon
species. Many tail-to-tail styrene sequences 23,24,29
clearly exist in the polymer chain. The reason for
the better separation of p-methylstyrene units, with
no detectable tail-to-tail sequences, may be due to
the better regioselectivity of 2,1-insertion 24 of p-MS,
with electron donating p-methyl group favorable for
cationic catalytic site. After the insertion of an
ethylene unit (no consecutive p-MS insertion allowed), the electronic donation of p-methyl group

polca

W: Poly Chem
97109T

POLY(ETHYLENE-CO-p-METHYLSTYRENE)

1027

Figure 8. 13C-NMR spectra of poly(ethylene-co-p-methylstyrene) with 10.9 mol % pMS and the expanded aliphatic region (insert).

may increase the interaction of the aromatic group


with the cationic catalytic site to further reduce
the space opening around it, as illustrated in
Scheme 2. The selective 2,1-insertion assures no
tail-to-tail p-MS sequence in the copolymer, and the
steric hindrance in the catalytic site may further
promote the consecutive ethylene insertion.

EXPERIMENTAL
Instrumentation and Materials
All room and high temperature 1H-NMR and 13CNMR were recorded on a Bruker WP-200 or a
Bruker AM-300 spectrometer with the DISNMR
software. The molecular weight was determined

using a Waters 150C which was operated at


1407C. The columns used were mStyragel HT of
. A flow rate of 0.7 mL/
10 6 , 10 5 , 10 4 and 10 3 A
min was used and the mobile phase was trichlorobenzene. Narrow molecular weight polyethylene
samples were used as standards. Differential
scanning calorimetry (DSC) was measured (with
a heating rate of 207C/min) on a Perkin Elmer
DSC-7, TAC-7 instrument controller.
All O2 and moisture sensitive manipulations
were performed inside an argon filled Vacuum Atmosphere dry box equipped with a MO-40-1 drytrain. Cp2ZrCl2 (Aldrich) and MAO (Ethyl) were
used as received. Et(Ind)2ZrCl2 30 and [C5Me4 (SiMe2N tBu)]TiCl2 23 were prepared as described in
the literature. High purity grade ethylene (MG Industries) was used as received. All styrene derivatives (Wiley Organics) was dried over CaH2 before
distillation. CP grade hexane and toluene were deoxygenated by argon purge before refluxing for 48
h and then distilled over potassium and CaH2 , respectively.

Copolymerization of Ethylene and Methylstyrene


Scheme 2.

/ 8G6C$$109T

02-19-98 09:38:28

In a typical ethylene copolymerization condition,


the comonomer (i.e. p-MS, o-MS, m-MS or sty-

polca

W: Poly Chem
97109T

1028

CHUNG AND LU

rene) was mixed with solvent (toluene or hexane)


and methylaluminoxane (MAO) (30 wt % in toluene) needed in a Parr 450 mL stainless autoclave
equipped with a mechanical stirrer. The sealed
reactor was then saturated with 3.1 1 10 5 Pa (45
psi) ethylene gas at 507C before adding catalyst
solution (i.e. Cp2ZrCl2 , Et(Ind)2ZrCl2 or [C5Me4(SiMe2N tBu)]TiCl2 in toluene) to initiate the polymerization. Additional ethylene was fed continuously into the reactor by maintaining a constant
pressure about 3.1 1 10 5 Pa during the whole
course of the polymerization. After 60 min, the
reaction was terminated by adding 100 mL of dilute HCl solution in MeOH. The polymer was isolated by filtering and washed completely with
MeOH and dried under vacuum at 507C for 8 h.

CONCLUSION
A new class of random poly(ethylene-co- p-methylstyrene) copolymers, with narrow molecular
weight and composition distributions, have been
developed by metallocene catalysts with constrain
ligand geometry. The experimental results clearly
show that p-methylstyrene performs distinctively
better than styrene, o-methylstyrene and mmethylstyrene, in the ethylene copolymerization
reaction. The combination of spatially opened catalytic site and cationic coordination mechanism
in [C5Me4 (SiMe2N tBu)]TiCl2 catalyst provides a
very favorable reaction condition for p-methylstyrene to be randomly incorporated in the poly(ethylene-co-p-methylstyrene) structure. In turn, the
copolymers are very useful intermediates in the
preparation of functional polyolefins which will be
discussed in the near future.
Authors would like to thank the Polymer Program of
the National Science Foundation and Dow Chem. Co.
for financial support.

REFERENCES AND NOTES


1. (a) M. D. Baijal, Plastics Polymer Science and
Technology, John Wiley & Sons, New York, 1982;
(b) E. C. Carraher Jr. and J. A. Moore, Modification of Polymers, Plenum, Oxford, 1982; (c) M. D.
Purgett, Ph.D. Thesis, University of Massachusetts, 1984; (d) K. J. Clark and W. G. City, U.S.
Patent 3,492,277, 1970.
2. J. Boor Jr., Ziegler-Natta Catalysts and Polymerizations, Academic Press, New York, 1979.

/ 8G6C$$109T

02-19-98 09:38:28

3. G. Ruggeri, M. Aglietto, A. Petragnani, and F. Ciardelli, Eur. Polymer J., 19, 863 (1983).
4. (a) J. W. Eshuis, Y. Y. Tan, A. Meetsma, J. H. Teuben, J. Renkema, and G. G. Evens, Organometallics, 11, 362 (1992); (b) L. K. Johnson, S. Mecking,
and M. Brookhart, J. Am. Chem. Soc., 118, 267
(1996).
5. M. R. Kesti, G. W. Coates, and R. M. Waymouth,
J. Am. Chem. Soc., 114, 9679 (1992).
6. (a) T. C. Chung, G. J. Jiang, and D. Rhubright,
U.S. Patents 5,286,800 (1994); (b) T. C. Chung,
G. J. Jiang, and D. Rhubright, U.S. Patents
5,401,805 (1995).
7. (a) T. C. Chung and D. Rhubright, Macromolecules, 26, 3019 (1993); (b) T. C. Chung, H. L. Lu,
and C. L. Li, Polymer International, 37, 197 (1995).
8. (a) T. C. Chung and D. Rhubright, J. Polymer Sci.,
Polymer. Chem. Ed., 31, 2759 (1993); (b) T. C.
Chung, H. L. Lu, and C. L. Li, Macromolecules, 27,
7533 (1994).
9. T. C. Chung, H. L. Lu, and W. Janvikul, J. Am.
Chem. Soc., 118, 705 (1996).
10. (a) T. C. Chung and G. J. Jiang, Macromolecules,
25, 4816 (1992); (b) T. C. Chung, D. Rhubright,
and G. J. Jiang, Macromolecules, 26, 3467 (1993).
11. T. C. Chung, H. L. Lu, and W. Janvikul, Polymer,
38, 1495 (1997).
12. (a) W. Chinsirikul, T. C. Chung, and I. Harrison, J.
Thermoplastic Composite Materials, 6, 18 (1993); (b)
S. H. Lee, C. L. Li, and T. C. Chung, Polymer, 35,
2980 (1994).
13. (a) T. C. Chung and D. Rhubright, Macromolecules, 27, 1313 (1994); (b) T. C. Chung, W. Janvikul, R. Bernard, and G. J. Jiang, Macromolecules,
27, 26 (1994); (c) T. C. Chung, W. Janvikul, R.
Bernard, R. Hu, C. L. Li, S. L. Liu, and G. J. Jiang,
Polymer, 36, 3565 (1995).
14. (a) T. C. Chung and H. L. Lu, U.S. Pat. 5,543,484
(1996); (b) T. C. Chung and H. L. Lu, J. Polym.
Sci. A, Polym. Chem. Ed., 35, 575 (1997).
15. (a) J. M. J. Frechet, Crown Ethers and Phase
Transfer Catalystsin Polymer Science, Plenum
Press, New York, 1984; (b) K. W. Powers, H. C.
Wang, T. C. Chung, A. J. Jias, and J. A. Olkusz,
U.S. Patents 5,162,445, 1992.
16. (a) S. Mohanraj and W. Ford, Macromolecules, 19,
2470 (1986); (b) R. G. Jones and Y. Matsubayashi,
Polymer, 33, 1069 (1992); (c) D. Pini, R. Settambolo, A. Raffaelli, and P. Salvadori, Macromolecules, 20, 58 (1987); (d) I. Piirma and J. R. Lenzotti, Brit. Polym. J., 21, 45 (1989).
17. (a) Y. Nagasaki and T. Tsuruta, Makromol. Chem.,
Rapid Commun., 7, 437 (1986); (b) F. Bonaccorsi,
A. Lezzi, A. Prevedello, L. Lanzini, and A. Roggero,
Polymer International, 30, 93 (1993).
18. (a) A. Onopchenko, J. G. D. Schulz, and R. J. Seekircher, J. Org. Chem., 37, 1414 (1972); (b) L. P.

polca

W: Poly Chem
97109T

POLY(ETHYLENE-CO-p-METHYLSTYRENE)

19.

20.

21.
22.
23.

24.

Ferrari and H. D. Stover, Macromolecules, 24, 6340


(1991).
(a) T. C. Chung, H. L. Lu, and R. D. Ding, Macromolecules, 30, 1272 (1997); (b) F. Bonaccorsi, A.
Lezzi, A. Prevedello, L. Lanzini, and A. Roggero,
Polym. Intern., 30, 93 (1993).
(a) W. Kaminsky, K. Kulper, and H. Brintzinger,
Angew. Chem., Int. Ed. Engl., 24, 507 (1985); (b)
J. A. Ewen, J. Am. Chem. Soc., 106, 6355 (1984);
(c) L. H. Slaugh and G. W. Schoenthal, U.S. Pat.
4,665,047 (1987); (d) H. W. Turner, U.S. Pat.
4,752,597 (1988).
J. M. Canich, U.S. Pat. 5,026,798 (1991).
S. Y. Lai, J. R. Wilson, G. W. Knight, J. C. Stevens,
and P. W. Chum, U.S. Pat. 5,272,236 (1993).
J. C. Stevens, F. J. Timmers, J. R. Wilson, G. F.
Schmidt, P. N. Nickias, R. K. Rosen, G. W. Knight,
and S. Y. Lai, Eur. Pat. Appl. 416,815 A2 (1991).
(a) P. Longo, A. Grassi, and L. Oliva, Makromol.
Chem., 191, 2387 (1990); (b) L. Oliva, L. Caporaso,

/ 8G6C$$109T

02-19-98 09:38:28

25.
26.
27.

28.
29.
30.

polca

1029

C. Pellecchia, and A. Zambelli, Macromolecules, 28,


4665 (1995); (c) C. Pellecchia, D. Pappalardo, M.
DArco, and A. Zambelli, Macromolecules, 29, 1158
(1996); (d) L. Oliva, P. Longo, L. Izzo, and M. DiSerio, Macromolecules, 30, 5616 (1997).
J. A. Ewen, R. L. Jones, A. Razavi, and J. L. Ferrara, J. Am. Chem. Soc., 110, 6255 (1988).
J. C. Stevens, Stud. Surf. Sci. Catal., 89, 277
(1994).
(a) R. F. Jordan, J. Chem. Edu., 65, 285 (1988);
(b) J. J. Eshuis, Y. Y. Tan, A. Meetsma, and J. H.
Teuben, Organometallics, 11, 362 (1992); (c) X.
Yang, C. L. Stern, and T. J. Marks, J. Am. Chem.
Soc., 116, 10015 (1994).
T. Kelen and F. Tudos, React. Kinet. Catal. Lett.,
1, 487 (1974).
F. G. Sernetz, R. Mulhaupt, and R. M. Waymouth,
Macromol. Chem. Phys., 197, 1071 (1996).
I. M. Lee, W. J. Gauthier, J. M. Ball, B. Iyengar,
and S. Collins, Organometallics, 11, 2115 (1992).

W: Poly Chem
97109T

You might also like