You are on page 1of 162

4.

09
Olefin Polymerizations with Group IV Metal
Catalysts
L Resconi, Basell Polyolefins, Ferrara, Italy
J C Chadwick, Eindhoven University of Technology, Eindhoven, The Netherlands
L Cavallo, University of Salerno, Salerno, Italy
2007 Elsevier Ltd. All rights reserved.
4.09.1 Introduction 1006
4.09.2 Pre-catalysts by Chemical Type and Reaction Principles 1007
4.09.2.1 MC as Propagating Species/Activation 1008
4.09.2.2 Monomer Coordination and Insertion Reactions 1010
4.09.2.3 Concepts of Stereo-, Regio-, and Enantioselectivity 1015
4.09.2.3.1 Regio- and stereochemistry of monomer insertion 1015
4.09.2.3.2 Definition of stereoregular polymers 1016
4.09.2.3.3 Elements of chirality 1016
4.09.2.3.4 Mechanism of stereocontrol 1018
4.09.2.3.5 Symmetry rules for stereocontrol 1020
4.09.2.4 Mechanism of Regiocontrol and Stereochemistry of Regioirregular Insertions 1023
4.09.2.5 Chain-release and Isomerization Reactions 1023
4.09.2.6 Kinetics 1028
4.09.3 ZieglerNatta Polymerizations with Heterogeneous Catalysts 1031
4.09.3.1 Catalyst Structure and Characterization 1031
4.09.3.2 Polymer Particle Growth 1033
4.09.3.3 Mechanistic Studies of ZieglerNatta Catalysts 1034
4.09.3.3.1 Oxidation state 1034
4.09.3.3.2 Number of active centers 1035
4.09.3.3.3 Internal/external donor effects and the nature of the active species 1035
4.09.3.3.4 Effects of hydrogen 1037
4.09.3.3.5 Effects of temperature 1038
4.09.3.4 Polyolefins Accessible from ZieglerNatta Catalysts 1038
4.09.3.5 Polymerization of Acyclic Internal Olefins 1040
4.09.3.6 Major Industrial Processes 1040
4.09.4 Polymerizations with Metallocene Catalysts 1041
4.09.4.1 Ethylene Polymers 1041
4.09.4.1.1 Polyethylene 1041
4.09.4.1.2 Ethylene/c-olefin co-polymers 1043
4.09.4.1.3 Ethylene/propylene co-polymers and ethylene/propylene/diene terpolymers 1045
4.09.4.1.4 Ethylene co-polymerization with c,c
/
-disubstituted and internal olefins 1047
4.09.4.1.5 Ethylene co-polymers with cycloolefins 1047
4.09.4.1.6 Ethylene/styrene co-polymers 1049
4.09.4.2 Propylene Polymers 1051
4.09.4.2.1 Amorphous polypropylene 1052
4.09.4.2.2 Isotactic polypropylene 1056
4.09.4.2.3 Low isotacticity: from flexible to elastomeric isotactic polypropylene 1064
4.09.4.2.4 Syndiotactic crystalline and elastomeric polypropylene 1070
4.09.4.2.5 Semicrystalline propylene/ethylene co-polymers 1073
4.09.4.2.6 Propylene/butene co-polymers 1075
1005
4.09.4.2.7 Propylene/higher c-olefin co-polymers 1076
4.09.4.2.8 Propylene co-polymerization with macromonomers 1077
4.09.4.3 Polybutene 1078
4.09.4.4 Poly(c-olefins) from Monomers Higher than Butene 1080
4.09.4.5 Polystyrene 1081
4.09.4.6 Cyclopolymers 1084
4.09.4.7 Polymers of Cyclic Olefins 1084
4.09.4.8 Polymerization of Conjugated Dienes 1084
4.09.5 Polymerization of Ethylene, Propylene, and Higher c-Olefins with other
Single-Center Catalysts 1086
4.09.5.1 Complexes with Coordination Number 4 1086
4.09.5.1.1 Ligands with coordinating OO atoms 1086
4.09.5.1.2 Ligands with coordinating NN atoms 1087
4.09.5.1.3 Other ligands 1090
4.09.5.2 Complexes with Coordination Number 5 1091
4.09.5.2.1 Ligands with coordinating OO atoms 1091
4.09.5.2.2 Ligands with coordinating NO atoms 1091
4.09.5.2.3 Ligands with coordinating NN atoms 1092
4.09.5.2.4 Other ligands 1095
4.09.5.3 Complexes with Coordination Number 6 1095
4.09.5.3.1 Ligands with coordinating OO atoms 1095
4.09.5.3.2 Ligands with coordinating NO atoms: phenoxyimine-catalysts for polyethylene 1096
4.09.5.3.3 Ligands with coordinating NO atoms: phenoxyimine catalysts for syndiotactic polypropylene 1115
4.09.5.3.4 Ligands with coordinating NO atoms: phenoxyimine catalysts for isotactic polypropylene 1126
4.09.5.3.5 Other ligands with coordinating NO atoms 1127
4.09.5.3.6 Complexes with NN chelate ligands 1138
4.09.5.3.7 Other ligands 1142
4.09.5.3.8 Olefin co-polymerizations with post-metallocene catalysts 1143
4.09.5.3.9 Polystyrene and olefinstyrene co-polymerization with post-metallocene catalysts 1145
References 1146
4.09.1 Introduction
This chapter covers the polymerization of alkenes with homogeneous and heterogeneous catalysts based on group 4
metals, including the underlying reaction principles and the relationship between catalyst structure and polymer
properties. Applications of related complexes in CC bond-forming reactions in organic synthesis are covered in
Chapter 00125. The use of transition metal catalysts in polymer synthesis is more widely discussed in chapter 11.06.
Catalytic olefin polymerization by means of groups 4 and 5 (ZieglerNatta) or group 6 (Phillips) metal catalysts is
one of the major chemical industries in the world. Polyethylene (PE) (both high density (HDPE) and linear low
density (LLDPE) and polypropylene ((PP); including propylene-rich co-polymers and heterophasic co-polymers) are
the two major thermoplastic polymers, with world productions of about 40 and 36 million tons/year, respectively (2003
figures). Titanium-based, heterogeneous ZieglerNatta catalysts dominate PP production and also play a leading role
in the manufacture of HDPE and LLDPE. Chromium-based Phillips catalysts are also widely used in HDPE
production, while metallocene and related single-site catalysts are making significant inroads in LLDPE produc-
tion. The total market for industrial polyolefin catalysts is estimated to exceed 6000 tons/year.
In the last 20 years or so, thanks to the development of the metallocene and single-site organometallic catalysts,
catalytic olefin polymerization has further evolved into one of the most actively studied branches of catalysis. (The
term single-site catalyst is widely used; however, in order to avoid confusion with coordination sites, and to
underline the chemical uniformity of the active species in metallocene catalysts, we prefer the term single-center
catalyst.) While characterized as mature about 10 years ago, and despite its cyclic nature, the polyolefin business is
recognized today as a healthy and growing business, thanks to continuing technology innovations, and significant
1006 Olefin Polymerizations with Group IV Metal Catalysts
expansions in the Asian market. The huge commercial success of polyolefin materials has, in turn, fueled research
activities in academia and industrial R&D institutions. In addition to the continuing expansion of established
technologies, such as the Spheripol and Unipol processes, several new processes have been developed, and new
plants built, in order to fulfill the ever-growing market request for new polyolefin-based materials. Most recent
examples are those of Basells new Spherizone gas-phase process for PP, Basells new two-reactor polybutene plant,
and Dows and Exxons solution processes for the production of propylene-based plastomers and elastomers. Without
diminishing the importance of process and material design, polymer science, and obviously market economics, the
success of these new technologies is to a great extent due to catalyst development.
Despite the heterogeneous and multi-component nature of the industrial MgCl
2
- or silica-supported ZieglerNatta
catalysts, which hampers the understanding of the elementary steps and kinetics of monomer insertion, chain growth,
and termination mechanisms, significant progress has been made, especially in the elucidation of fundamental
aspects of stereoregulation and molecular mass control. New and more efficient catalyst modifiers (donors) that
enable the tuning of chain stereoregularity, molecular mass distribution, and co-monomer incorporation in isotactic
polypropylene (iPP) have been found.
On the other hand, in order to simplify the nature of the active species and better unravel the many elementary
steps simultaneously operating during catalytic polymerization, group 4 bis(cyclopentadienyl) complexes
1
were
studied by Natta and Breslow as early as 1957 as soluble and structurally well-defined models for TiCl
3
-based
heterogeneous ZieglerNatta catalysts.
2,3
However, for many years, these complexes remained just models due to
their uncompetitively low catalytic activities. At the end of the 1970s, the pioneering work of Brintzinger on the
synthesis of chiral metallocenes,
413
combined with Sinn and Kaminskys seminal discovery of methylalumoxane
(MAO) as a superior activator for metallocene catalysts,
14
suddenly turned zirconocenes from model catalysts into
highly effective ethylene polymerization systems, endowed also with an unprecedented co-monomer incorporation
ability. These discoveries, and Ewens subsequent groundbreaking work on ligand effects in stereoselective poly-
merization, marked the birth of a new era in catalytic olefin polymerization: that of well-defined, purposely designed,
single-center organometallic catalysts.
Organometallic chemists have played a key role in designing new ligands, organometallic complexes, and catalyst
systems, understanding their activation chemistry, and determining the mechanisms of olefin interaction with
transition metals and the stereochemical implications of chain growth. In addition to a much clearer understanding
of the chemistry involved in polymerization catalysis, detailed mechanistic investigations have also generated a
wealth of new polyolefin materials, new applications, and ultimately markets, that were inaccessible with the
heterogeneous ZieglerNatta catalysts.
Many extensive reviews and books have been recently dedicated to the field of catalytic olefin polymerization,
both for ZieglerNatta catalysts
1517
and for metallocene and other single-center catalysts.
1820
Nevertheless, the
pace of development is so quick that a new, comprehensive review appears timely. In the following, we describe the
evolution of ZieglerNatta catalysts, the revolution of single-center catalysts, and their application most at
laboratory level only to the synthesis of novel or improved polyolefins in the last 10 years.
4.09.2 Pre-catalysts by Chemical Type and Reaction Principles
The most common geometries adopted by group 4 catalysts are depicted in Scheme 1. In all practical cases, the active
center is a cationic, strongly electrophilic metal complex capable of activating the CTC double bond of the inserting
monomer. This positive charge of the complex cation is counterbalanced by a weakly (or non-) coordinating
Coordination 4
Tetrahedral
Coordination 5
Trigonal bipyramid
P
P
M
P = growing polymer chain;
M M
L
1
L
1
+ + +
L
1
L
2
L
2
L
2
L
4
L
3
L
3
L = generic ligand
P
X

X

= counterion;
X

Coordination 6
Octahedral
Scheme 1
Olefin Polymerizations with Group IV Metal Catalysts 1007
counteranion. The active center must have two coordination sites in mutually cis-positions in order to enable the
transfer of the growing polymeryl chain to the coordinated monomer. In the absence of the monomer, one of these
cis-coordination positions is usually saturated by the counterion. The ligand(s) must confer the required steric and
electronic properties, which control the microstructure and the molecular mass of the produced polymers. The
generic ligands L can be based on anionic aromatic groups such as the cyclopentadienyl (Cp, j
5
-C
5
H
5
) ring and its
derivatives, as well as on anionic or neutral o,:-donors usually based on heteroatoms, such as O, N, S, and P. Overall,
the set of coordinating L ligands is usually dianionic. Finally, the metal atom most often is a d
0
-metal in the oxidation
state IV.
A brief listing of the most typical pre-catalysts used in catalytic olefin polymerizations is represented in Figure 1.
These are the systems that will be discussed in much more detail in the following sections. Examples of catalysts
based on 1 are dialkoxide- and diamide-based tetrahedral systems. Introduction of an extra neutral donor ligand as in
2 results in pentacoordinate catalysts. Structures with a piano-stool geometry such as 3 are usually denominated half-
sandwich complexes. Pre-catalysts such as 4 include the remarkably interesting class of ansa-monocyclopentadienyl
amido complexes (also known as constrained-geometry catalysts or CGC), while the well-known bis-cyclopentadienyl
metallocenes correspond to pre-catalysts of generic formula 5. Systems 68 present an octahedral coordination
geometry at the metal atom, and include systems with two unbridged chelating ligands as in 6, and the most well-
known complexes of this kind are the bis(phenoxyimine) complexes of titanium. Pre-catalysts with tetradentate
ligands as in 7 include the bridged bis(phenoxyamine)-based catalysts, while pre-catalysts such as 8 are character-
ized by a tridentate ligand with an extra donor arm. Systems such as 35, which contain at least one Cp ligand, are
discussed in Section 4.09.4, while systems 1, 2, 6, 7, and 8, which can be broadly defined as non-metallocene
catalysts, are discussed in Section 4.09.5
4.09.2.1 MC as Propagating Species/Activation
The propagating active site in olefin polymerizations mediated by group 4 catalysts is the MC(polymer) bond of a
metalalkyl complex.
2133
Although a few neutral group 4 catalysts, such as complexes 9 (M=Zr, Hf
34
) and 10,
3539
have been synthesized, almost all effective group 4 complexes are inactive in polymerization if not activated by a
suitable co-catalyst.
9 10
M

5
-C
2
B
9
H

11
Me
M
+
R
B(C
6
F
5
)
3
L
MX
2
L
1
L
1
MX
2
L
1
L
2
2
X
M
X
X
R
5
3
MX
2
L
R
5
4
MX
2
R
5
R
5
5
L
2
L
1
MX
2
2
6
M
X
L
1
L
2
L
2
L
1
X
7
L
1
M L
1
L
2
L
3
X
X
8
Figure 1 Structure of the most typical pre-catalysts used in catalytic olefin polymerizations.
1008 Olefin Polymerizations with Group IV Metal Catalysts
Activation and formation of the cationic species are accomplished through a suitable activating species, the co-
catalyst, and thus the importance of the co-catalyst in olefin polymerizations with group 4 systems is fundamental.
The activator becomes an anion after the activation process, forming a cationanion pair, which is now accepted to be
the real catalytically active polymerization species. With different activators, dramatic differences in activity are
possible for a given pre-catalyst structure.
40,41,86,86a
Furthermore, the counteranion was demonstrated to influence
the stability and activity of the catalyst, as well as the molecular masses and even stereoregularity of the polymers
produced. It was the discovery of MAO by Sinn and Kaminsky
14
that started the metallocene revolution, although the
complexity of the catalytic system did not allow conclusions about the structure of the active species. After the
cationic nature of the active catalytic species was established,
2232
several other activators were designed, most of
them based on non-coordinating borates and aluminates. Excellent reviews on the subject have appeared.
40,41
Selected examples of activators are shown in Figure 2.
In order to produce an active catalyst upon reaction with the activator, the pre-catalyst has to be alkylated
either during its synthesis or in situ by an aluminum alkyl compound. Al-alkyls and Al-alkyl chlorides are
important components of heterogeneous ZieglerNatta systems. However, their inability to efficiently activate
group 4 metallocenes has for a long time limited developments in this field, until the arrival of MAO, which is
now the most widely used activator. The structure of MAO is still rather undefined. In solution, MAO exists as an
equilibrium of species with different aggregation numbers and structures.
4244
Proposed structures include linear
chains, cyclic rings, three-dimensional clusters, and cage structures.
40,4556
MAO as co-catalyst has some draw-
backs: low solubility in aliphatic solvents, poor long-term stability in solution, the high content of MAO residues
(alumina) in the final product, and the relatively high cost, not least in view of the rather large amount needed for
effective activation (the typical Al/M molar ratio needed for homogeneous systems is 10
3
: 110
4
: 1, although in
supported systems, ratios around 100 : 1 are sufficient). This is especially true for systems of not very high activity.
F B
F F
3
11
F
F F
F F
F F
F F
B
F F F F
2
C
6
F
5
F
F
12
F
B
F F
3
F F
F
F
13
(C
6
F
5
)
2
B B(C
6
F
5
)
2
F F
14
F
F
F
F
B(C
6
F
5
)
2
B(C
6
F
5
)
2
15
(C
6
F
5
)
2
(C
6
F
5
)
2
F
F
F
F
F
F
F
F
16
F BH
F F
2
2
17
F
F F
F F
Ph
3
C
+
B
4

18
F
F F
F F F F F F
R
3
NH
+
B
4

19
F
F F
Ph
3
C
+
Al
4

20
Si
i
Pr
3
F F
Ph
3
C
+
B
4

21
F
F
F
Ph
3
C
+
B
F
F
F
F
F
2

22
F
F F
O
F F

Ph
3
C
+
Al
4
23
(C
6
F
5
)
3
B C N B(C
6
F
5
)
3

Ph
3
C
+
24
B(C
6
F
5
)
3
(C
6
F
5
)
3
B
R
3
NH
+

25
N
B(C
6

+
F
5
)
3
26
Figure 2 Selected examples of activators.
Olefin Polymerizations with Group IV Metal Catalysts 1009
Finally, the danger inherent in the use of extremely pyrophoric AlMe
3
has to be taken into account. Surrogates of
MAO include ethylalumoxane and isobutylalumoxane synthesized from AlEt
3
and AlBu
i
3
, although they do not
perform as well as MAO.
5763
To solve the above problems, modified MAOs have been investigated. The patent
literature reports the use of MAO/AlBu
i
3
mixtures,
64
or the hydrolysis products of AlBu
i
3
and other branched
Al-alkyls.
61,65,66
The presence of residual AlMe
3
is another problem associated with MAO. Several authors
showed that increasing the AlMe
3
/MAO ratio or replacing AlMe
3
with AlEt
3
or AlBu
i
3
results in a decrease of
both activity and molecular masses.
44,6772
MAOs which contain much less residual AlMe
3
have been developed
and are claimed to exhibit better performances than conventional MAO.
73,74
Several other approaches have been
proposed to reduce the amount of AlMe
3
in MAO.
7577
A different strategy toward stoichiometric co-catalysts has been the use of perfluoroaryl boranes such as 1115 and
17. Ewen and Marks independently introduced the already known strongly Lewis-acidic borane B(C
6
F
5
)
3
11 as
activator for olefin polymerizations with group 4 metallocenes.
7881
Reaction of B(C
6
F
5
)
3
with group 4 dimethyl
metallocenes (Figure 3) is rapid and quantitative at room temperature in non-coordinating solvents. Crystal structures
of the products show that the methyl group of the [MeB(C
6
F
5
)
3
]

moiety remains coordinated to the cationic


metallocene.
78,79
Other perfluoroaryl borane activators were developed, such as the bifunctional borane
[HB(C
6
F
5
)
2
]
2
17,
82
and the sterically encumbered perfluorobiphenyl and perfluoronaphthyl boranes.
8385
Trityl
and ammonium borates such as 18, 19, 21, and 22 and aluminate salts such as 20 and 23 are other classes of widely
used activators.
40,83,86,86a90
Different approaches include the cyano-bridged 24, the weakly coordinating 25, and the
pyrrole-based 26.
86,86a,91,92
Although the [B(C
6
F
5
)
4
]

-based activators are highly effective in olefin polymerization,


9398
they have some drawbacks. They are poorly soluble in many hydrocarbon solvents and can have limited thermal
stability, which results in short catalytic lifetimes.
89
On the other hand, whereas MAO and related co-catalysts are
used in large stoichiometric excess, for borane, borate, and similar co-catalysts, a 1 : 1 molar ratio of activator and
dialkyl pre-catalyst is sufficient. In some cases, [Ph
3
C][B(C
6
F
5
)
4
] used in excess over the metallocene can
significantly increase the productivity of some propylene polymerization catalysts, in particular, those with high
activity systems such as constrained-geometry titanium complexes.
99,100
Since the catalyst activator has been shown to exert a remarkable influence on the performance of olefin
polymerization catalysts,
41,99,101106
the search for new co-catalysts is an active field of research. This, however, is
beyond the scope of this review.
86,86a,107113
The products of activation with the three main classes of activators described above are shown in Figure 3. We only
add that in order to have more reproducible results, and to reduce the amount of catalyst needed for optimum activity,
adding small amounts of AlR
3
(such as AlBu
i
3
and AlEt
3
) to the reaction system is a common practice to scavenge
impurities and, with metallocene dihalide precursors, to alkylate the metal.
114,115
It is worth noting that small
aluminum alkyls such as AlMe
3
and AlEt
3
form heterobinuclear complexes with metallocene alkyl cations, of the
type [L
2
M(j-Me)
2
AlMe
2
]

, such that high concentrations of these aluminum alkyls reduce the catalyst activity.
31,33
However, there is no evidence that bulky aluminum alkyls such as AlBu
i
3
form similar adducts with group 4
metallocene catalysts.
Upon activation, the metalalkyl cation and the counteranion form an ion pair. In the low-polarity solvents used in
olefin polymerizations, the interaction between the cation and the anion is rather strong. Methyl borates derived from
11 (activation reaction (b) in Figure 3) represent an example of a tight ion pair, with a bridging Me
group.
40,116
Conversely, ion pairs with a tetrakis(perfluoroaryl) borate counterion (Figure 3, reactions(c) and (d))
represent examples of less tightly bound ion pairs and the anion in an outer-sphere position.
40,101,116
The exact
mechanism and energetics of ion pair formation (pre-catalyst activation) have been widely investigated by several
groups.
40,117120,122,126,127
The structure and dynamics of ion pairs is conveniently investigated by spectroscopic
NMR techniques.
116
Finally, the possible aggregation of ion pairs to form species such as ion quadruples, hextuples,
and higher-order aggregates has also been investigated.
101,105,121125
The main conclusion seems to be that, at the
concentrations typically used in olefin polymerizations, catalyst ion pairs are unlikely to be present as higher
aggregates.
105,123,124
These aspects have been summarized in pertinent reviews.
41,116
4.09.2.2 Monomer Coordination and Insertion Reactions
The fundamental reaction in catalytic olefin polymerizations is monomer insertion into an MC bond, schematically
described in Scheme 2. The general mechanistic features are well covered in two reviews.
126,127
The mechanism generally accepted for the chain-growth reaction of Scheme 2 is reported in Figure 4. Cossee
originally proposed this mechanism, now known as the ArlmanCossee mechanism.
128,129
It substantially occurs in
1010 Olefin Polymerizations with Group IV Metal Catalysts
two steps: (i) olefin coordination to the metal, (ii) alkyl migration of the o-coordinated growing chain to the
:-coordinated olefin. Green, Rooney, and Brookhart slightly modified this mechanism with the introduction of
a stabilizing c-agostic interaction,
130
which would facilitate the insertion reaction.
131133
The role of c-agostic
interactions in olefin insertion has been rationalized by Grubbs and Coates.
134
M Me
2
Si
MAO-X

MMe
2
Me
2
Si
MMe
2
Me
2
Si
+
Me
MX
2
Me
2
Si
M Me
2
Si
-MeB(C
6
F
5
)
3

+
Me
M Me
2
Si
[B(C
6
F
5
)
4
]

+
Me
+ MAO
+ B(C
6
F
5
)
3
+ [Ph
3
C][B(C
6
F
5
)
4
]
Ph
3
CMe
MMe
2
Me
2
Si M Me
2
Si
+
Me
+
N
B(C
6
F
5
)
3

N
B(C
6

F
5
)
3
+
CH
4
MMe
2
Me
2
Si M Me
2
Si
[B(C
6
F
5
)
4
]

+
Me
+ [R
3
NH][B(C
6
F
5
)
4
]
NR
3
,CH
4
(a)
(b)
(c)
(g)
(d)
M Me
2
Si M Me
2
Si
+
+ B(C
6
F
5
)
3
(C
6
F
5
)
3
B
MMe
2
Me
2
Si M Me
2
Si
-MeAl(C
6
F
5
)
3

+
-MeAl(C
6
F
5
)
3

+ 2Al(C
6
F
5
)
3
(e)
(f)
Figure 3 Selected examples of metallocene activation processes.
Olefin Polymerizations with Group IV Metal Catalysts 1011
The first step of the insertion reaction requires that the active metal center has an available coordination site for the
incoming monomer. For many years, it was commonly accepted that olefin coordination to the cationic metal was an
easy process, with a low activation energy possibly connected to the displacement of a weakly coordinated solvent
molecule or of a weakly agostic interaction between the metal and a CH bond of the growing polymer chain. In
recent years, this view has changed. Certainly, with coordinating anions like [MeB(C
6
F
5
)
3
]

, olefin coordination
requires anion displacement, and it has even been suggested that olefin coordination could represent the rate-limiting
step.
135,136
The second step of the chain-growth reaction, the insertion step, occurs via chain migration to the closest carbon of
the olefin double bond, which undergoes cis-opening with formation of the new metalcarbon and carboncarbon
bonds.
137
Consequently, at the end of the reaction, the new Mchain o-bond is on the site previously occupied by the
coordinated monomer molecule (chain-migratory mechanism). At the end of the reaction, the coordination position
previously occupied by the growing chain is then occupied by the counteranion. This mechanism is schematically
represented in Figure 4(b). It is important to note that the inclusion of the anionic counterion does not pertain to
heterogeneous ZieglerNatta catalytic systems since no anionic co-catalysts are used in this case.
The overall activation energy of the reaction is the result of different contributions, from counterion displacement
to the breaking and forming of the MC bonds. Of course, the strength of the ion pair interaction contributes
sensitively to the overall activation barrier, and it explains why catalysts with tightly bound counteranions such as
[MeB(C
6
F
5
)
3
]

show lower activities relative to catalysts with weakly bound counteranions such as [B(C
6
F
5
)
4
]

. In
some cases, it has been suggested that the nature of the monomer influences the position of the transition state for
monomer insertion, with anion displacement being important in propene polymerization, while with 1-hexene alkyl
transfer to the coordinated monomer was found to be rate determining, independent of the anion.
138
Further details
on this topic can be found in a critical review.
41
After insertion, the growing chain can swing back to the coordination position occupied before insertion. This
isomerization mechanism, represented in Scheme 3, is usually referred to as site isomerization or backskip of the
M P P
n
+ Monomer M
n +1
Scheme 2
M
H
P
n
H
M
H
P
n
H
M
H
P
n
H
M
H
P
n
H
M
P
n
+C
2
H
4
M
H
P
n
H
M
H
P
n
H
M
H
P
n
H
M
H
P
n
H
M
P
n
+C
2
H
4
A

+
+ + + +
A

(a)
(b)
+ + + + +

Figure 4 (a) Modified Cossee mechanism for olefin polymerizations with group 4 transition metals; (b) modified mechanism in
the presence of an anionic counterion.
M
P
n
M
P
n
chain
backskip
Scheme 3
1012 Olefin Polymerizations with Group IV Metal Catalysts
growing chain.
139
The backskip of the growing chain can have an effect on the sequence of enantioselective
steps which determine the microstructure of the resulting polymer in the case of prochiral olefins.
104,139
While the
chain-migratory mechanism is commonly accepted, there are cases in which regular (or predominant) chain migration
at each insertion step is not operative. In this case, the growing chain returns to the original coordination position at
the end of each insertion reaction, and olefin coordination occurs predominantly at one coordination site. This last
mechanism was shown to occur in some particular cases, and its occurrence is highly dependent on the nature of the
counteranion.
104
We refer to it as chain-retention mechanism.
Detailed quantum mechanics calculations have indicated that agostic interactions occur between the growing chain
and the metal atom. The most typical are shown in Figure 5. Calculations on gas-phase metal alkyl cations indicate
that the u-agostic interaction is the most stable, with the -agostic interaction roughly 25 kcal mol
1
higher in
energy, and the less stable c-agostic interaction about 10 kcal mol
1
higher in energy.
140,141
Quantum mechanics calculations indicated that olefin coordination to the naked cationic catalyst is a barrierless
and exothermic process that leads to the olefin coordination intermediate of Figure 6. The coordination intermediate
eventually evolves to the four-center Cossee-like transition state of Figure 6, and then collapses into the products that
resemble the agostically bound alkyl species of Figure 5.
140146
Interestingly, these quantum mechanics calculations
confirmed that the transition state is assisted by c-agostic interactions, as proposed by Green, Rooney and
Brookhart.
130
Quite a small energy barrier (15 kcal mol
1
) has been calculated for the insertion step in the case of
the naked cationic catalyst.
141,144146
While the naked cation could be a model of a catalyst with a completely non-coordinating counteranion, the energy
profile in the presence of a tightly coordinating counterion such as [MeB(C
6
F
5
)
3
]

is remarkably different. The first


issue is how the olefin enters the metal coordination sphere. The three different olefin approaches shown in
Scheme 4 have been investigated with quantum mechanics approaches.
-Agostic -Agostic
-Agostic
2.36
2.17
2.31
Zr Zr Zr
Figure 5 Agostic interactions between an isobutyl group (simulating a growing chain) and the Zr atomin [Me
2
Si(1-Ind)
2
Zr-Bu
i
]

;
distances in

A.
Olefin
coordination
Insertion
transition state
Olefin
Olefin
2.26
2.36
2.38
Zr
2.10
1.42 1.35
3.01
2.76
2.73
Zr
2.29
Growing
chain
Growing
chain
Figure 6 Olefin coordination intermediate and transition state for insertion of propylene into the ZrBu
i
bond of [Me
2
Si(1-Ind)
2
-
Zr-Bu
i
]

; the Bu
i
group simulates the growing chain; distances in

A.
Olefin Polymerizations with Group IV Metal Catalysts 1013
For the [H
2
Si(Cp)(NBu
t
)TiCH
3
][MeB(C
6
F
5
)
3
] system, olefin coordination/insertion along path A is slightly
favored over paths B and C because it requires less cationanion separation. In any case, olefin coordination in the
presence of the counterion requires that a sizeable energy barrier must be overcome.
147149
Modeling ethylene
insertion on the [Me
2
Si(Cp)(NBu
t
)TiCH
3
][MeB(C
6
F
5
)
3
] system confirmed that ethylene approach pathways A and B
are of very similar energy, but they also indicated that for insertion into longer Tialkyl bonds, such as insertion into
the TiPr
n
bond, path B is favored.
136
More interestingly, they also suggested that the rate-limiting step could be
olefin coordination and not olefin insertion.
135,136
Ethylene insertion into the [Cp
2
ZrC
2
H
5
]

cation with both the


[MeB(C
6
F
5
)
3
]

and [B(C
6
F
5
)
4
]

counterions has been modeled; for these systems too, the approach along path B was
found to be favored.
117
Many experimental mechanistic studies have been devoted to clarify the role of the counterion in monomer
insertion (and thus on catalyst activity).
41,55,99,101106,117,119,135,136,147,148,150156
Based on the results of studies on
the competitive coordination to the metal atom of the counterion versus an added Lewis base, it has been proposed
that the tight ion pair is unable to insert the monomer, and that displacement of the counterion has to occur. After
dissociation, one (or possibly more) olefin molecules may insert into the Mchain bond before the counterion
recoordinates to the metal, and chain growth is stopped until the counterion dissociates.
157
This mechanism closely
resembles that proposed by Fink based on early studies on the kinetics of ethylene oligomerization, which led to a
mechanistic scheme where the polymer chain-growth process could be interrupted at any stage by the reversible
formation of a resting state, the so-called intermittent mechanism.
158
On the other hand, studies on the polymerization of 1-hexene polymerization catalyzed by rac-C
2
H
4
(Ind)
2
ZrMe-
(j-Me)B(C
6
F
5
)
3
showed that monomer insertion, anion displacement, and anion recoordination are part of a
concerted process, for which the term continuous mechanism was suggested.
154
NMR studies on metallocene
ion pairs bearing a longer-chain alkyl ligand as polymeryl model, [rac-Me
2
Si(Ind)
2
ZrCH
2
SiMe
3

], indicated that
a mechanism of the continuous type is operative for the tightly bound counteranion [MeB(C
6
F
5
)
3
]

. By contrast, if
the cation is paired with the very weakly coordinating [B(C
6
F
5
)
4
]

, the counteranion does not enter the inner


coordination sphere of the metallocenium cation, and as a result the inserting monomer does not have to compete
with the counteranion for coordination to the metal. The catalysts differ therefore structurally, with [MeB(C
6
F
5
)
3
]

forming an inner-sphere ion pair (ISIP), while [B(C


6
F
5
)
4
]

gives an outer-sphere ion pair (OSIP). Nevertheless,


although the degree of anion coordination and the catalyst structures are strongly anion dependent, both insertion
mechanisms are similar in the sense that both involve an exchange of the alkyl ligand and anion positions after each
insertion step, following the principle outlined in Figure 4(b).
101
This mechanistic model is in agreement with Finks
original concept of an intermittent process.
158
On the basis of combined X-ray and NMR studies, it has been
suggested that the different binding capability of the two counterions results in different resting states, involving an
c-agostic methyl interaction with tightly bound [MeB(C
6
F
5
)
3
]

, and a -agostically bonded alkyl chain in OSIPs with


non-bonded counterions such as [B(C
6
F
5
)
4
]

(see Scheme 5).


41,101
P
n
M
Cp Cp
X

+
A
B
C
Scheme 4
M
Cp Cp
C
+
Tightly bound ion pair
M
Cp Cp
X

+
P
n
Me
Weakly bound ion pair
(a) (b)
P
n
Me
H
H
B(C
6
F
5
)
3
H
Scheme 5
1014 Olefin Polymerizations with Group IV Metal Catalysts
4.09.2.3 Concepts of Stereo-, Regio-, and Enantioselectivity
While ethylene insertion can occur in a single mode, insertion of c-olefins can occur in the four geometrically
different modes represented in Scheme 6. Thus, polymerization of prochiral monomers requires the definition of a
few terms.
4.09.2.3.1 Regio- and stereochemistry of monomer insertion
The regiochemistry of insertion (the catalyst regioselectivity) defines whether olefin insertion is primary or secondary
(also called 1,2 or 2,1 insertions, respectively). Any catalyst will insert some olefin molecules with the wrong
regiochemistry. Regioirregular insertions (regioerrors or regiomistakes) mean occasional secondary (primary) insertion
if propagation is prevailingly primary (secondary). Monomer insertion is mostly primary for metallocene catalysts (the
amount of regiomistakes being usually <1%), regioirregular for some dialkoxotitanium complexes, and mostly
secondary for some non-metallocene catalysts (the amount of regiomistakes usually <12%).
137,159162
While the
preference for primary insertion in metallocenes is mainly steric in nature,
163,164
for most non-metallocene catalysts it
is the result of a subtle interplay of steric and electronic effects.
165
The ability to insert preferentially one and the same olefin enantioface (or enantioface selectivity) defines the
stereochemistry of each insertion (the catalyst stereoselectivity). Every insertion of a monosubstituted prochiral
c-olefin creates a new stereogenic center, and the stereoregularity or tacticity of the polymer is determined by the
stereochemical relationships between successive tertiary carbon atoms in the polymer chain. The term stereo-
specific polymerization refers to the process leading to a tactic (stereoregular) polymer produced with a stereo-
selective catalyst. The terms stereoselective (or enantioselective) and regioselective are also used to refer to the
single insertion event. Given their widespread use, the terms aspecific, isospecific, and syndiospecific, referring to the
type of enantioselectivity of a catalyst,
162
are also used.
In summary, the stereoregularity or tacticity of a polymer chain is determined by the stereochemical relationship(s)
between neighboring tertiary carbon atoms in the chain. The stereoselectivity of a catalyst, that is, the reactivity and
selectivity of its active sites, is determined by the metal, the geometry of the ligands, and the structure of the metal-
bound polymer chain end. Thus, different structures of the last inserted monomer unit, or the chirality of its
stereogenic carbon (defined by the primary or secondary, left or right insertion) increase the number of active sites
on a given metal center. The sites can, and often do, differ in reactivity, regioselectivity, and enantioface selectivity:
as a result, the active metal center itself changes in chemical structure during a single chain growth, but statistically
keeps the same overall reactivity from one polymer chain to the next. From these considerations, it appears that these
catalysts are better defined as single-center catalysts, a term we prefer over the commonly used description as
single-site catalysts.
In the framework of the chain-migratory mechanism, olefin insertion occurs on the two different coordination sites
available. This implies that the two different geometric situations, corresponding to an exchange between the
coordination positions of the growing chain and of the monomer, are related by the overall symmetry of the metal
Scheme 6
Olefin Polymerizations with Group IV Metal Catalysts 1015
:-ligands framework. The most typical situations are shown in Figure 7. If the pre-catalyst has an overall
C
1
-symmetry, the two coordination positions are sterically and electronically different. The activity and catalytic
properties of the two possible sites can be completely different. If the pre-catalyst has an overall C
2
-symmetry, the
two coordination positions are homotopic and thus the activity and catalytic properties of the two sites are identical. If
the pre-catalyst has an overall C
S
-symmetry, two cases are possible. If the local mirror plane contains the coordination
positions available to the growing chain and the monomer, the two coordination positions are not symmetry related,
and, by symmetry, they are both non enantioselective. More interesting is the case in which the local mirror plane
relates the coordination positions available to the growing chain and to the monomer. In fact, in this case the two
coordination positions are enantiotopic, and thus the possible asymmetric induction (in the framework of the chain-
migration mechanism) is opposite at each insertion step.
4.09.2.3.2 Definition of stereoregular polymers
The structure of perfectly regular isotactic and syndiotactic PP(iPP and sPP) is shown in Figure 8. Isotactic PP is
characterized by a sequence of tertiary C atoms with the same local spatial arrangement (a single configuration) in the
polymer chain. Syndiotactic PP is characterized by a sequence of tertiary C atoms with alternate configuration in the
polymer chain. Atactic PP(aPP) has no regularity in the sequence of the configuration of the tertiary C atoms. While
the iPP is of major industrial interest, and sPP- and aPP have found some applications, hemiisotactic PP, in which one
in every two tertiary C atoms is isotactic while the other is atactic, also shown in Figure 8, is an academic curiosity. If
the relative configuration of two successive tertiary C atoms (a diad) is considered, an isotactic polymer can be
considered as composed by a sequence of m (meso) diads, while a syndiotactic polymer can be considered to be
composed by a sequence of r (racemic) diads. Finally, in the case of diolefins, stereoselective polymerization can lead
to diisotactic and disyndiotactic polymers, whose structure is also reported in Figure 8.
166
4.09.2.3.3 Elements of chirality
Stereoselective c-olefin polymerization is the result of a sequence of asymmetric reactions (c-olefin coordination/
insertion). The main elements of chirality are as follows.
(i) Monomer coordination. Coordination of the two enantiofaces of a prochiral c-olefin gives rise to chiral si and re c-
olefin coordinations.
167
Isotactic polymers are generated by multiple insertions of c-olefin molecules with the
same enantioface (either re or si), while syndiotactic polymers are generated by a regular alternation of insertions
of re- and si-coordinated monomers.
(ii) Chirality of the active site. Different cases are present here. (a) The chirality can arise from coordination of
prochiral ligands. In this case, the notation (R) or (S), in parenthesis, according to the CahnIngoldPrelog rules as
extended by Schlo gl can be used.
168,169
As an example, the (R,R) chirality of coordination of the H
2
C(1-Ind)
2
ligand, labeled according to the absolute configurations of the bridgehead carbon atoms marked by arrows, is
shown in Figure 9(a). In the case of complexes with two bidentate ligands, the relative orientations of the two
bidentate ligands can be chiral and generate chirality at the metal. This chirality can be labeled with the notation
or , defined for octahedral coordination compounds (Figure 9(b)).
170
(c) An intrinsic chirality at the central
metal atom, which for tetrahedral or pseudo-tetrahedral situations can be labeled with the notation R or S of
C
1
-symmetric
pre-catalyst
M M M M
C
2
-symmetric
pre-catalyst
C
s
-symmetric
pre-catalysts
Figure 7 Schematic representation of the most common symmetries of group 4 olefin polymerization catalysts. Gray
rectangles define the space occupied by the organic ligand. Hollow squares represent the coordination positions available to the
growing chain and to the monomer. Dashed lines represent the local mirror plane of the two C
S
-symmetric catalysts.
1016 Olefin Polymerizations with Group IV Metal Catalysts
A A A A A A A A A A
B B B B B B B B B
A A A A A A A A A A B B B B B B B B B
A B
A B A B
A B
A B
A B
A
A B A
B A B
(a)
(b)
(c)
(d)
(e)
(f)
(g)
m m m m m m m m m
r r r r r r r r r
m r r r r r m r m
Figure 8 Segments of isotactic (a), syndiotactic (b), atactic (c), and hemiisotactic polypropylene (d) chains. Segments of
erythro-diisotactic (e), threo-diisotactic (f), and disyndiotactic (g) poly-diolefin chains. The modified Fischer projection is shown.
For parts, (a)(c) a zigzag representation is also reported.
Monomer
Monomer
Growing
chain Growing
chain
Mt Mt
Mt
(R )
(a) (b) (c)
Growing
chain
Monomer
R (R)
Figure 9 Schematic representation of the chirality at the active site in the case (a) of a C
2
-symmetric pseudo-tetrahedral
metallocene, (b) of a C
2
-symmetric octahedral model for heterogeneous catalysts, and (c) of a syndiospecific C
S
-symmetric
pseudo-tetrahedral metallocene.
Olefin Polymerizations with Group IV Metal Catalysts 1017
CahnIngoldPrelog rules as extended by Stanley and Baird.
168,171
For instance, the diastereoisomer with
intrinsic R configuration at the central metal atom is shown in Figure 9(c), for the case of a metallocene with a
H
2
C(Cp)(9-Flu) ligand. It is important to note that chirality of type (c) requires that the two coordination
positions available for the growing chain and the monomer are occupied by different ligands. This implies that
compounds such as the dichloride pre-catalysts are not chiral.
One or more of these kinds of chirality of the site can be present in the active site. However, for the case of
catalytic complexes in which the two ligands are tightly connected through chemical bonds and which are called
hereafter as stereorigid, only the chirality of kind (c) can change during the polymerization reaction.
(iii) Chirality of the growing chain. The last tertiary C atom of the growing chain is chiral, and its configuration is
determined by the chirality of monomer coordination in the last insertion step. The R/S CahnIngoldPrelog
nomenclature can be used. However, it is common to label the two configurations as si- or re-ending growing
chains, according to the configuration of the monomer during coordination/insertion.
168,172
4.09.2.3.4 Mechanism of stereocontrol
It is well accepted that two mechanisms of stereocontrol (the chiral induction responsible for selecting the monomer
enantioface) are operative in stereoselective c-olefin polymerizations. In the simpler cases, the discrimination
between the two faces of the prochiral monomer may be dictated either by the configuration of the asymmetric
tertiary C atom of the last inserted monomer unit or by the chirality of the catalytic site. These two different
mechanisms of stereocontrol are named chain-end stereocontrol and enantiomorphic-site or site stereocontrol. In the
case of chain-end stereocontrol, the selection between the two enantiofaces of the incoming monomer is operated by
the chiral environment provided by the last inserted tertiary C atom of the growing chain, whereas in the case of site
stereocontrol this selection is operated by the chirality of the catalytic site. The origin of stereocontrol in olefin
polymerization has been reviewed extensively.
162,172178
The distribution of steric defects along the polymer chain may be indicative of which kind of stereocontrol is
operative. The type and amount of stereomistakes (enantioface insertion errors) is measured by solution
13
C NMR
spectroscopy, a sensitive technique that is able to see the steric environment of a given propylene unit up to
undecads (five propylene units on each side of the observed monomeric unit). Routine analysis is usually performed
at the pentad level (two propylene units on each side of the observed monomeric unit).
162,179
The microstructures
which result from stereomistakes are shown in Scheme 7.
Any catalyst will make some enantioface insertion mistakes (stereoerrors or stereomistakes). In the case of
stereomistakes in propylene polymerization, the chemical shift of the methyl groups is highly sensitive to the relative
stereochemistry of neighboring monomer units. The degree of tacticity can be given as the pentad, triad, or diad
content (% mmmm, % mm, % m for isospecific polymerization, and % rrrr, % rr, % r for syndiospecific polymerization,
respectively). In the case of low stereoregularity, the diad excess, % m-r (% r-m) better represents the degree of iso-
(syndio-) tacticity.
180
Isolated insertion errors unambiguously identify the polymerization mechanism.
Stereomistakes are easily and quantitatively detected by NMR spectroscopy, and useful relationships for the four
stereospecific polymerization mechanisms discussed above are reported in Table 1. Additionally, the relationships
2[rr]/[mr] =1 and 2[mm]/[mr] =1 identify iso- and syndiospecific site control, respectively, whereas the relationship
4[mm][rr]/[mr]
2
=1 identifies chain-end control. The average stereoregular block lengths are 2[m]/[r] 1 and 2[r]/
[m] 1 for isotactic and syndiotactic polymers, respectively.
It is worth noting that in the case of syndiospecific propagation, the microstructure of the polymer is also affected
by other secondary reactions such as the backskip of the growing chain, or backside attack of the olefin. For example,
a backskip of the growing chain followed by correct enantioface selection and regular chain migration would originate
a microstructure identical to that generated from a stereomistake in the case of chain-end stereocontrol.
104,179,181,182
The origin of stereocontrol in c-olefin polymerizations has been clarified in detail, and it was shown to be
essentially driven by steric effects.
172,174176,178
The commonly accepted mechanism was developed by the
Corradini school, and it is called the mechanism of the chiral orientation of the growing chain. It was proposed
at the beginning of the 1980s to explain the stereospecificity of heterogeneous catalysts.
183188
In the case of primary
propagation, the chiral environment provided by the chirality of the complex (in the case of site stereocontrol) or by
the chirality of the tertiary C atom of the last inserted monomeric unit (in the case of chain-end stereocontrol)
imposes a chiral orientation to the growing chain in order to minimize the steric interaction between the ligand
skeleton and the growing chain. It is this chiral orientation of the growing chain that selects between the two
enantiofaces of the incoming monomer molecule. The preferred enantioface is the one that minimizes steric
1018 Olefin Polymerizations with Group IV Metal Catalysts
interaction with the chirally oriented growing chain. Usually, this is the enantioface that places the methyl group of
the propylene trans to (i.e., away from) the chirally oriented growing chain (Figure 10).
172,174,175,183,185,187,189192
In the case of secondary propagation, the mechanism of the chiral orientation of the growing chain is still operative
for stereoflexible compounds. For example, Cavallo and Guerra suggested that for the bis(phenoxyimine)Ti
catalysts, the secondary growing chain assumes a chiral orientation in order to establish an c-agostic interaction
with the metal.
172,193
This chiral orientation imposes a configuration to the stereoflexible complex, and it is the
chirality of the complex (imposed by the growing chain) that selects between the two enantiofaces of the secondary
inserting propylene. This mechanism was originally proposed for the V-based systems.
194
Finally, in the case of
secondary propagation and stereorigid complexes, stereoselectivity is simply determined by steric interactions
between the ligand and the monomer molecule. The growing chain plays no role here.
The mechanism of the chiral orientation of the growing chain has strong experimental support. The
13
C NMR
analysis of the PP end groups performed by Zambelli and co-workers showed that propylene insertion is essentially
non-enantioselective in the first polymerization step (when the alkyl group bonded to the metal is a methyl group),
whereas it is enantioselective in successive insertion steps when an isobutyl group is bonded to the metal. This holds
good for both heterogeneous
195
and homogeneous
196
ZieglerNatta catalysts. The same mechanism predicts that re
insertion of the monomer is favored in case of (R,R) chirality of coordination of the C
2
H
4
(1-Ind)
2
ligand. This is in
agreement with optical activity measurements by Pino
197,198
on saturated propylene oligomers obtained with this
M
P
n
Isolated stereomistake
Stereomistake propagated
Isolated stereomistake
Stereomistake propagated
isospecific
syndiospecific
isospecific
syndiospecific
site control
chain-end control
primary
insertion
Scheme 7
Table 1 Relationship between the microstructures which result from the stereomistakes
shown in Scheme 7
Mechanism Misinsertions (triads) Misinsertions (pentads)
Isospecific chiral-site control [mr] =2[rr] [mmmr] =[mmrr] =2[mrrm]
Syndiospecific chiral-site control [mr] =2[mm] [rrrm] =[mmrr] =2[rmmr]
Isospecific chain-end control mr only [mmmr] =[mmrm]
Syndiospecific chain-end control mr only [rrrm] =[rrmr]
Olefin Polymerizations with Group IV Metal Catalysts 1019
kind of catalyst, proving that re insertion of the monomer is indeed favored in case of (R,R) chirality of coordination of
the C
2
H
4
(1-Ind)
2
ligand. Moreover, deuteration and deuterio-oligomerization studies of c-olefins (propylene,
1-pentene, 4-methyl-pentene) using catalysts based on (R,R) C
2
H
4
(H
4
-1-Ind)
2
zirconium derivatives showed that the
R enantioface of the olefin is predominantly involved in dimerizations and oligomerizations whereas the S enantioface is
favored in the deuterations.
197
These results confirm that the growing chain plays a primary role in enantioface
discrimination.
199
Results relative to deuteration and deuterodimerization experiments on isotopically chiral 1-pentene,
as well as on propylene insertion with betaine derivatives of classical metallocenes, also agree with a mechanism
involving a chiral orientation of the growing chain.
200,201
In conclusion, in site-controlled stereoselective polymerizations, it is accepted and proved that the site chirality is
unable to select directly between the two enantiofaces of the inserting monomer. Instead, it is accepted and proved
that the site chirality can force a chiral orientation of the growing chain, which in turn is able to select between the
two enantiofaces of the inserting monomer. Thus, the growing chain acts as a messenger to transfer the chiral
information from the catalytic site to the monomer.
172
4.09.2.3.5 Symmetry rules for stereocontrol
The key to understanding the variety of different tacticities obtained in c-olefin polymerizations is the chain-
migratory mechanism with the site-switching mechanism of Figure 4, and the fact that insertion of the olefin can
actually occur on two different active sites. It is the symmetry relationship between the two situations corresponding
to an exchange of the positions of the growing chain and of the monomer that determines the tacticity of the resulting
poly-c-olefin. Considering the space around the metal center divided in four quadrants, the steric bulkiness of the
ligands shapes a chiral pocket that can result in asymmetric insertion. The relationship between metallocene
symmetry and polymer stereochemistry has been fully understood. In the quadrants representation, a gray quadrant
means that space in this quadrant is occupied by the ligand, and thus is scarcely accessible to either the monomer or
the growing chain. In this representation, it does not matter if the geometry of coordination around the metal atom is
tetrahedral or octahedral, since space occupation is relevant. The most typical symmetries and the (possibly
expected) microstructures of the resulting poly-c-olefins are shown in Scheme 8. These rules apply to stereorigid
catalysts operating under site-control mechanism, and for primary monomer insertion.
Representative examples of aspecific C
2v
-symmetric pre-catalysts are Me
2
Si(Cp)
2
MCl
2
and Me
2
Si(9-Flu)
2
MCl
2
27. The ligands of these two catalysts have very different space occupation, and in the quadrants representation can
be considered to correspond to systems with all white quadrants or all gray quadrants, respectively. Another example
of an aspecific catalyst is based on the C
s
-symmetric meso-C
2
H
4
(1-Ind)
2
MCl
2
complex 28.
202
In the quadrant
representation, this catalyst can be considered a combination of the two C
2v
-symmetric catalysts just described, and
thus it is characterized by two white quadrants on the side of the Cp rings and by two gray quadrants on the side of the
six-membered rings of 28.
Growing
chain
Growing
chain
Indenyl
Zr
Zr
Fluorenyl
Cyclopentadienyl
Transition state for isotactic propagation
re-Propene
re-Propene
Indenyl
(a)
Transition state for syndiotactic propagation
(b)
Figure 10 Transition states for primary insertion of propylene (a) with the isospecific Me
2
Si(1-Ind)
2
Zr system and (b) with the
syndiospecific Me
2
C(Cp)(9-Flu)Zr systems.
1020 Olefin Polymerizations with Group IV Metal Catalysts
ZrCl
2
Si
27
ZrCl
2
28
Examples of isospecific C
2
-symmetric catalysts are the pseudo-tetrahedral rac-C
2
H
4
(1-Ind)
2
MCl
2
complex
29
6,202,203
and the rac-H
2
C(3-Bu
t
-1-Ind)
2
MCl
2
complex 30.
204
In the quadrants representation of Scheme 8,
sterically encumbered gray quadrants correspond to the six-membered rings in 29, and to the Bu
t
groups in 30.
Another relevant system is the rac-C
2
H
4
(3-Me-1-Ind)
2
MCl
2
complex 31, which is substantially aspecific although
it is C
2
-symmetric.
205
The effect of the increase of the bulkiness in different quadrants is best understood
comparing the three C
2
-symmetric systems of Scheme 9. While the parent bis(indenyl)-based catalyst is iso-
specific, the presence of a methyl group in position 3 of the Cp ring substantially counterbalances the bulkiness
of the six-membered rings. In the quadrant representation, all quadrants are gray, and the catalyst is substantially
aspecific because it is unable to impose a chiral orientation on the growing chain. Further increase of the steric
bulkiness by introduction of Bu
t
groups in position 3 of the Cp rings again corresponds to an isospecific catalyst,
but the quadrants occupied by the Bu
t
groups are substantially forbidden to the growing chain (black quadrants).
Theoretical calculations have in fact indicated that the Bu
t
group is able to induce a stronger steric pressure on
the growing chain relative to the flat indenyl group, and that for the same configuration of the complexes,
catalysts based on 29 and 30 impose opposite chiral orientations to the growing chain, which results in the
insertion of opposite propylene enantiofaces.
206,207
C
2v
-symmetric, achiral
C
2
-symmetric, chiral
C
1
-symmetric, chiral
C
s
-symmetric, achiral
C
s
-symmetric, prochiral
Diastereotopic sites
Enantiotopic sites
X M X X M X
X M X X M X X M X
X M X X M X
Homotopic sites
Homotopic sites
Isotactic polymer
From hemiiso- to isotactic polymer
From hemiiso- to isotactic
Atactic polymer
Syndiotactic polymer
Scheme 8
Olefin Polymerizations with Group IV Metal Catalysts 1021
ZrCl
2
29
ZrCl
2
30
ZrCl
2
31
Returning to Scheme 8, an example of syndiospecific C
s
-symmetric catalysts is the Me
2
C(Cp)(9-Flu)ZrCl
2
system 32,
208
characterized by enantiotopic coordination sites that favor opposite chiral orientations of the
growing chain, which results in the insertion of opposite propylene enantiofaces. It is the (almost) regular chain
migration between the two coordination sites that rationalizes the syndiospecificity of this kind of catalysts.
Finally, C
1
-symmetric catalysts can range from hemiisospecific, as for the Me
2
C(3-MeCp)(9-Flu)ZrCl
2
system
33,
209211
to isospecific as for the Me
2
C(3-Bu
t
-Cp)(9-Flu)ZrCl
2
structure, 34.
212
The effect of the bulkiness of
the substituents on the parent Me
2
C(Cp)(9-Flu)ZrCl
2
complex is summarized in Scheme 10. Introduction of a
single methyl substituent on position 3 of the Cp ring counterbalances the steric effect of the Flu group. In the
quadrants representation, there are three gray quadrants, and when the growing chain is located in the coordina-
tion position between the Flu and the Me group, insertion is non-stereoselective because the catalyst is unable to
impose a chiral orientation to the growing chain. In the framework of the chain-migratory mechanism, this C
1
-
symmetric catalyst is hemiisospecific, because there is an (almost) regular alternance between non-selective and
stereoselective propylene insertions. Instead, a single Bu
t
substituent on position 3 of the Cp ring sterically
dominates the Flu ligand, and the quadrant occupied by the Bu
t
group is forbidden to the growing chain. Thus, at
each insertion step, the catalyst imposes the same chiral orientation to the growing chain, and the same propylene
enantioface is selected. This results in an isotactic polymer. Detailed theoretical calculations have rationalized
these effects at molecular level.
175,213,214
ZrCl
2
32
ZrCl
2
33
ZrCl
2
34
X X M X X
Zr Zr
Isotactic polymer Isotactic polymer Atactic polymer
Zr
ZrCl
2
ZrCl
2 ZrCl
2
M
C
2
-symmetric, chiral homotopic sites
X X M
Scheme 9
1022 Olefin Polymerizations with Group IV Metal Catalysts
These rules, now referred to as Ewens symmetry rules, are the result of milestone papers by Brintzinger,
6
Ewen,
181,202,205,208,215,216
Kaminsky,
203
and co-workers. Although they have been developed for primary propylene
insertion with pseudo-tetrahedral metallocenes, they can be extended to octahedral catalysts as well. On the other
hand, in the case of secondary monomer insertion with unbridged octahedral systems, the tacticity of the polymer
produced is the result of a delicate balance between the chiralities of the chain end and of the fluxional active species.
This particular case has been discussed in a review.
172
4.09.2.4 Mechanism of Regiocontrol and Stereochemistry of Regioirregular Insertions
The origin of regiochemistry in propylene polymerization by group 4 catalysts has been investigated in great detail.
While the preference for primary insertion in metallocenes is mainly steric in nature,
163,164
for most post-metallocene
catalysts it is a subtle interplay of steric and electronic effects.
165
In the case of the prototype-based catalysts 29 and 32,
the most favored transition states leading to secondary propylene insertion are shown in Figures 11(a) and 11(b),
respectively. These transition states are disfavored with respect to the corresponding transition states for primary
propylene insertion due to some steric interaction between the methyl group of the inserting propylene molecule and
the nearby Cp ring of the metallocene ligand. These steric interactions are at the origin of the preference for primary
propagation with metallocene-based catalysts. Transition states of Figures 11(c) and 11(d) correspond to secondary
insertion of the other propylene enantioface, and are much higher in energy due to strong steric interactions between
the methyl group of the inserting propylene molecule and the closer and bulkier (relative to the Cp ring) six-membered
ring of the metallocene ligand. This implies that insertion of the secondary propylene is enantioselective, and that for
C
2
-symmetric metallocenes opposite enantiofaces are favored for primary and secondary insertion (compare Figures
10(a) and 11(a)), whereas in the case of C
S
-symmetric metallocenes the same propylene enantioface is favored in
primary and secondary insertion (compare Figures 10(b) and 11(b)). This fact has been used to develop a kinetic model
that links the regioselectivity of a given metallocene to its stereoselectivity, and that rationalizes the higher regioselec-
tivity usually exhibited by syndiospecific C
S
-symmetric metallocenes relative to C
2
-symmetric metallocenes.
164
4.09.2.5 Chain-release and Isomerization Reactions
The average degree of polymerization P

n
of a polyolefin (its molecular mass), produced under steady-state conditions,
with a non-living process, is determined by the ratio between propagation rates and chain release rates (Equation (1)).
Syndiotactic polymer Hemiisotactic polymer Isotactic polymer
Zr Zr
X X M
X X M
C
1
-symmetric C
1
-symmetric C
s
-symmetric
X X M
ZrCl
2
ZrCl
2
ZrCl
2
Zr
Scheme 10
Olefin Polymerizations with Group IV Metal Catalysts 1023
P

n
=
R
p
R
r
(1)
The structure and the mechanisms of formation of end groups have been reviewed.
217
The most important chain-
release reactions are u-H transfer after primary or secondary insertion (either to the metal or to a coordinated
monomer molecule), u-Me transfer, and chain transfer to the aluminum co-catalyst or scavenger, when present.
For a given set of polymerization conditions, the rates of these chain-release reactions are obviously inherent to the
steric and electronic nature of the active species. In other words, the molecular mass, as is the case with stereo-
regularity, strongly depends on the structure of the catalyst. Since propagation and chain-release reaction rates often
have different dependencies on temperature and monomer concentration, the molecular mass of polyolefins is also
strongly affected by the polymerization conditions. This important point is further discussed in Section 4.09.4.2.2.
In cases in which the catalyst produces too high molecular mass polymers for a given process or application,
molecular masses can be controlled more effectively by hydrogenolysis (chain transfer to hydrogen),
218221
or, in
specific cases, by chain transfer to ethylene after a primary insertion.
222,223
It is worth noting that the activity of most
propylene polymerization catalysts is increased by both hydrogen and ethylene. Thus, designing catalysts able to
produce polyolefins with much higher molecular masses than needed, then requiring hydrogen (or ethylene) for
molecular mass control, provides the additional advantage of increasing the catalyst activity.
224
The chain-transfer and -release reactions occurring with Ti-based heterogeneous ZieglerNatta catalysts are discussed
in Section 4.09.3. In the following, the most important chain-release reactions occurring at metallocene and other single-
center group IV catalysts are summarized. Chain transfer to ethylene is also addressed in Sections 4.09.4.1 and 4.09.4.2.
For ethylene polymerization, the picture is fairly simple, including bimolecular u-hydride transfer to a coordinated
ethylene monomer
225227
and transfer to the aluminum co-catalyst.
42,228
Formation of internal unsaturations has
been reported,
227,229,230
often connected to the formation of hydrogen (see below). With some catalysts, isomeriza-
tion and formation of ethyl and longer branches have also been observed, as a consequence of this chain-release
reaction. This aspect is described in Section 4.09.4.1.
Growing
chain
C
2
-symmetric
isospecific catalyst
C
s
-symmetric
syndiospecific catalyst
Growing
chain
Growing
chain
Growing
chain
Steric
interaction
Severe steric
interaction
Severe steric
interaction
Steric
interaction
si-Propylene
si-Propylene
re-Propylene
re-Propylene
(a)
(c) (d)
(b)
Figure 11 Favored transition states for the secondary insertion of propylene with (a) the isospecific Me
2
Si(1-Ind)
2
Zr systemand
with (b) the syndiospecific Me
2
C(Cp)(9-Flu)Zr system. High-energy transition states for the secondary insertion of propylene with
(c) the isospecific Me
2
Si(1-Ind)
2
Zr system and (d) the syndiospecific Me
2
C(Cp)(9-Flu)Zr system.
1024 Olefin Polymerizations with Group IV Metal Catalysts
The most ubiquitous chain-release reactions occurring in c-olefin polymerizations are the unimolecular and
bimolecular u-hydride transfers after primary insertion.
217,231236
These are shown in Scheme 11.
Unimolecular u-H transfer to the metal in propylene polymerization is key to understanding growing-chain-end
isomerization
237
and formation of internal vinylidenes.
217
Detailed studies have unambiguously shown that in the case of zirconocenes,
237246
and other single-center
catalysts,
247,248
isotacticity decreases at lower propylene concentrations due to unimolecular primary-growing-
chain-end epimerization, which scrambles the chirality of the last chirotopic methine of the growing chain.
The now-accepted mechanism of epimerization, first proposed by Busico and Cipullo,
238
and recently confirmed
by Yoder and Bercaw by means of an elegant double-labeling study,
237
involves the reaction product of u-H transfer
to the metal, a metal cationpolymeryl olefin complex, and is shown in Scheme 12. Allylic activation on the same
metal cationolefin complex has been proposed to generate internal double bonds in PE
229
and vinylidene unsatura-
tions in PP (Scheme 13).
249
An important phenomenon occurs in ethylenepropylene co-polymerization, especially in liquid propylene poly-
merization: addition of small amounts (<30 mol%) of ethylene causes a strong decrease in PP molecular masses. This
L
n
M
H
H
CH
3
H
P
+
L
n
M
+
H
P
H

L
n
M
CH
3
H CH
3
H
P
+
L
n
M
+
P
Unimolecular
-H transfer
Bimolecular
-H transfer
P
Insertion
(initiation)
Insertion
(chain growth)
Scheme 11 Primary u-H transfer reactions.
B
A
+
(R,R)Zr
(R,R)Zr
H
P
1
3
+
H
P
1
3
Primary
insertion
Primary
insertion
In-plane olefin
rotation
Me
P
H
Secondary
insertion
-H from C(3)
Secondary
insertion
-H from C(1)
P
Me
In-plane olefin
rotation
H
-H
P
Me
H
-H
Me
P
H
3
1
1 3
P
Me
3
1
Scheme 12 Epimerization via unimolecular u-hydride transferdouble bond in-plane rotationreinsertion.
Olefin Polymerizations with Group IV Metal Catalysts 1025
has been observed with isospecific and aspecific metallocenes,
222,223,250253
and has limited the development of
industrial metallocene catalysts for isotactic polypropylene (iPP).
In order to avoid this limitation, new (and more complex) bis-indenyl ligands have been developed, and are
described in Section 4.09.4.2.5.
When the catalyst is not fully regioselective, chain release by a u-H transfer after a secondary insertion with
formation of internal double bonds is often observed. This has been reported for ethylene/c-olefin co-poly-
mers,
229,254
PP,
255
and other polyolefins,
232
as well as for 1-hexene polymerization with dialkoxide catalysts.
256
The reaction is shown in Scheme 14 for the case of propylene, where kinetic studies have shown it to be a bimolecular
process, following the rate law
s
R
u-H
=
s
k
u-H
[
s
Zr][m].
217,257
[
s
Zr] refers to the concentration of active Zr centers
bearing a growing chain having a secondary propylene unit linked to the metal.
Differently, in the case of secondary chain growth with bis(phenoxy-imine)titanium catalysts, Coates and co-
workers reported that chain release occurs exclusively by u-H hydride transfer from the terminal methyl. This
generates an allylic end group as shown in Scheme 15,
160
and it has been utilized to produce functionalized
syndiotactic propylene oligomers.
258
Chain release by u-CH
3
transfer to the metal is unimolecular (Scheme 16)
32,207,259
and obviously limited to the
presence of propylene or other 2-methyl-substituted c-olefins, such as isobutene,
260
and 2-methyl-1,5-hexadiene.
261
No u-alkyl transfer has been reported for higher alkyls, except for u-trimethylsilyl transfer
262
and cases in which a
strained ring is formed.
263265
u-Methyl transfer is an important and sometimes prevalent cause of molecular mass
depression in the case of propylene polymerization with sterically hindered metallocenes
262,266,267
or high polymer-
ization temperatures combined with low propylene concentrations.
268
Eisen reported this mechanism to be the
P P
L
2
Zr
+
P
P
L
2
Zr
+
L
2
Zr
+
P
P
L
2
Zr
+
H
Allylic
activation
P
Allylic
activation
C
3
H
8
H
2
n
L
2
Zr
+
A
B
Scheme 13 Formation of internal vinylidene unsaturations. A: via bimolecular u-H transfer; B: via unimolecular u-H transfer.
L
n
M
CH
3
H
H
H
P
+
-CH
3
transfer
L
n
M
+
+ P CH
3
Scheme 16
L
n
Ti
CH
3
P
n +
L
n
Ti
P
n +
H +
-H
transfer
Scheme 15 u-H transfer to the metal from terminal methyl of a secondary chain.
H

L
n
Zr
CH
3
H H
H
3
C
P
+ -H
transfer
L
n
Zr
+
+ P
Scheme 14 u-H transfer to a coordinated propylene monomer after a secondary propylene insertion.
1026 Olefin Polymerizations with Group IV Metal Catalysts
exclusive chain-release reaction with bis(benzamidinate)-based catalysts.
247,248
u-Methyl transfer has been used for
the production of allyl-terminated propylene oligomers, which are themselves polymerizable monomers and have
been used for the production of long-chain-branched (LCB) polyolefins.
269,270
Figure 12 shows the most common terminal unsaturations of iPP, generated by chain-release reactions after
primary insertion. The internal vinylidene, likely produced by allylic activation, is also shown.
217,246
Figure 13 shows the terminal unsaturations of iPP, generated by chain-release reactions after a secondary insertion.
246,257
In addition to chain transfer to the coordinated monomer, there are three other relevant chain-transfer reactions:
(i) Chain transfer to the aluminum alkyl species (usually the co-catalyst or the AlR
3
species used as scavenger of
impurities) or other organometallic species, usually zinc alkyls.
271,272
Chain transfer to aluminum usually occurs
4.6 4.8 5.0 5.2 5.4 5.6 5.8
ppm
Figure 13 Olefinic region of a
1
H NMR spectrum (120

C, solvent C
2
D
2
Cl
4
at 5.95ppm) showing terminal and internal
unsaturations in iPP (catalyst rac-C
2
H
4
(4,7-Me
2
Ind)
2
ZrCl
2
/MAO).
4.6 4.8 5.0 5.2 5.4 5.6 5.8
ppm
Figure 12 Olefinic region of a
1
H NMR spectrum (120

C, solvent C
2
D
2
Cl
4
at 5.95ppm) showing terminal and internal
unsaturations in iPP (catalyst rac-Me
2
C(3-Bu
t
Ind)
2
ZrCl
2
/MAO).
Olefin Polymerizations with Group IV Metal Catalysts 1027
with the residual AlMe
3
present in MAO, and has been used to prepare hydroxo-terminated PP.
273,274
Using
borate activators in combination with higher Al alkyls such as AlEt
3
or AlBu
i
3
effectively reduces chain transfer
to Al (Scheme 17).
Transfer to Al was reported to be operative with several non-metallocene catalysts. It is the only chain-release
mechanism operative with the diamido complexes MCl
2
{ArN(CH
2
)
n
NAr} catalysts, as well as with the mono-
and tris(benzamidinate) catalysts, since no olefinic resonances were observed in the
1
H or
13
C NMR spectra of
these polymers.
275,276
This chain-release reaction is also dominant with bis(phenoxyimine)zirconium cat-
alysts,
277
as well as with tris(pyrazolyl)borate-based catalysts.
278,279
(ii) Chain transfer to hydrogen is the most widely used means of molecular mass control, in both classic Ziegler
Natta catalysis
224,280,281
and metallocenes.
218221,282
Although molecular masses can be reduced by either
reducing monomer concentration (in propylene and higher olefin polymerizations) or increasing polymerization
temperature, these two experimental parameters can be varied only within a narrow range in a given polymer-
ization process. Hence, a good hydrogen response is a must for an industrial polymerization catalyst.
(iii) Chain transfer to organosilanes, introduced by Marks for lanthanides, corresponds to the silanolytic Mchain
reaction shown in Scheme 18. This chain-transfer process efficiently produces silyl end-capped PEs as well as
ethylene/1-hexene and ethylene/styrene co-polymers.
283,284
4.09.2.6 Kinetics
The kinetics of olefin polymerization are the subject of several studies,
104,153156,162,182,221,226,240,241,246,252,255,266,285312
and of an excellent book by Keii.
17
The most relevant studies will be discussed below. However, we first note
that the precise description of the kinetics of catalytic olefin polymerization under industrially relevant poly-
merization conditions has proved to be very difficult. For a given catalytic system, one has to identify all possible
insertion, chain-release, and chain-isomerization reactions, and their dependence on the polymerization para-
meters (most importantly, temperature and monomer concentration). Once the kinetic laws for each elementary
step have been determined, they have to be combined in one model in order to be able to predict the catalyst
performance. This has been attempted for both ethylene
226,285
and propylene polymerizations. The case of
propylene polymerization with a chiral, isospecific zirconocene is shown in Figure 14.
162
Both polymerization temperature and monomer concentration usually have a strong influence on catalyst activity
and polyolefin molecular mass. With respect to monomer concentration, the rate of monomer insertion must obey the
simple first-order law
R
p
= k
p
[C[[m[ (2)
L
n
M
+
Al CH
3
R
2
Al P L
n
M
+
CH
3
H
+
HO
O
2
+
H
2
O
2
/NaOH
n
n
n
Scheme 17 Chain transfer to aluminum (the case of methylated aluminum species, such as Al(CH
3
)
3
, is shown).
L
n
M
P
n
+ R
3
SiH L
n
M
H
+ R
3
SiP
n
R = Bu
n
, Bn, Ph
Scheme 18
1028 Olefin Polymerizations with Group IV Metal Catalysts
Here, [C] is the concentration of active catalyst and [m] is the concentration of the monomer. In practice, polymer
production rates have instead been found to deviate substantially from first-order behavior, and follow the non-linear
dependence on monomer concentration
R
p
= k
p
9[m[
n
(3)
for both ethylene
286
and propylene,
241,246,255,288
with 1 <n <2. This behavior has been rationalized by the coex-
istence of two different active species differing in monomer insertion rates.
289
In this model, the two species are in
equilibrium with each other and the position of the equilibrium is determined by the concentration of the monomer,
with each active species obeying the rate law of Equation (2). This results in the law expressed by Equation (4).
R
p
=
k
1
[C[[m[ k
2
[C[[m[
2
k
3
k
4
[m[
(4)
Chain-release
reactions
Chain-isomerization Insertion
reactions
reactions
Primary growing chain
+
P
epimerization
R
e
= k
e
[C]
P
-H
R
-H
= k
-H
[C][M]
(R,R)Zr
(R,R)Zr
(R,R)Zr
(R,R )Zr
(R,R)Zr
+
P
1,2 re
2,1 si
1,2 si
+
P
+
P
+
P
P
-CH
3
or R
-H
= k
-H
[C]
Secondary growing chain
threo (t ) Regioinversion
P
3,1-Insertion
P
P
erythro (e) Regioinversion
P
1
2
3
1
2
3
1
2
(R,R)Zr
+
P
1,2 re
1,2 si
2
1
3
3
3,1-isomerization
1,2 re
s
R
is
=
s
k
is
[
s
C]
P
-H
s
R
-H
=
s
k
-H
[
s
C][M]
R
-Me
= k
-Me
[C]
Figure 14 The most relevant elementary steps observed at the (R,R)-enantiomer of a chiral, C
2
-symmetric, isospecific
zirconium center with a primary growing chain end (top) and a secondary growing chain end (bottom). The (S,S)-enantiomer
produces the opposite stereochemistry of each single event, but overall the same polymer chains and the same insertion
mistakes. In practice, in the case of C
2
-symmetric metallocenes, the racemic mixture (R,RS,S) is always used. P=growing
polypropylene chain; [C] =concentration of active primary centers; [
s
C] =concentration of active secondary centers.
Olefin Polymerizations with Group IV Metal Catalysts 1029
Propagation rates of first order in monomer concentration have been reported for ethylene
226,287
and for propylene
in the case of aspecific metallocenes
266
as well as for propylene polymerization with isospecific metallocenes
activated with MAO, B(C
6
F
5
)
3
, and [Ph
3
C][B(C
6
F
5
)
4
].
290,297
Moreover, first-order kinetics were also observed for
1-hexene polymerization with the [rac-C
2
H
4
(1-Ind)
2
ZrMe][MeB(C
6
F
5
)
3
].
156
Higher orders affect the laws determining the degree of polymerization as a function of propylene concentration.
The dependence of molecular mass on propylene concentration is given by
P

n
=
a[m[ b[m[
2
c d[m[
(5)
If the propagation rate is first order in propylene concentration, then Equation (5) reduces to Equation (6), which
can be linearized (Equation (7)).
266
P

n
=
k
p
[m[
k
t
0
k
t
1
[m[
(6)
1
P

n
=
k
t
1
k
p

k
t
0
k
p
[m[
(7)
Models able to rationalize and predict catalyst performances for propylene polymerization with C
1
- and
C
s
-symmetric metallocenes such as Me
2
C(Cp)(1-Ind)ZrMe
2
and Me
2
C(Cp)(9-Flu)ZrCl
2
have been proposed.
291,292
A kinetic model has been proposed based on microstructural analysis, including both chain-epimerization and site-
epimerization reactions in both C
2
- and C
s
-symmetric metallocenes, and rationalizing the observed pseudo-second-
order kinetics of propylene polymerization promoted by C
2
-symmetric metallocene catalysts.
182,293
This point has
been extended to co-polymers.
298
A thorough study of propylene polymerization with the Me
2
C(Cp)(9-Flu)ZrCl
2
system in the presence of a large series of different counterions that rationalized the correlation between the nature of
ion pair and the microstructure of the resulting PPs has been performed.
104
As regards a comparison between initiation (i.e., insertion of the monomer into the MCH
3
bond) and propagation
(i.e., insertion of 1-hexene into the Mchain bond), in the case of 1-hexene polymerization with the [rac-C
2
H
4
(1-
Ind)
2
ZrMe][MeB(C
6
F
5
)
3
] catalyst, initiation is about 70 times slower than propagation.
156
An even more pronounced
effect was found in the polymerization of propylene with the rac-Me
2
Si(1-Ind)
2
zirconocene; in the presence of MAO,
initiation is about 800 times slower than propagation, and this value increases to 6000 when polymerizing in the
presence of the [B(C
6
F
5
)
4
]

counterion.
290
The amount of metal centers actually active during polymerization has been investigated using quenched-flow
kinetic techniques. The concentration of active sites in MAO-activated systems is significantly lower (about 10%)
than the analytical concentration of the metal, and a large fraction of metal species are in a dormant state.
290,297
In
the case of propylene polymerization, it was suggested that the resting state after secondary propylene insertion was a
major contributor to the dormant-state concentration, since insertion of a new propylene molecule into the M
(secondary alkyl) is greatly slowed by steric effects.
240,290,294,295,299301
However, any other situation that slows chain
propagation may contribute to the dormant state. These conclusions have been extended also to [Ph
3
C][B(C
6
F
5
)
4
]-
activated systems, and the higher activity exhibited by borate-activated systems relative to the MAO-activated
systems has been ascribed to the weaker coordination ability of borate relative to MAO, rather than to the differences
in the active site concentrations.
290
Finally, the mole fraction of active species is about 80% when the
[Ph
3
C][B(C
6
F
5
)
4
]-activated bis(phenoxyamine)Zr catalyst is considered.
296
Using deuterium-labeling experiments, about 100% of the metal was shown to be active in 1-hexene polymeriza-
tions with the [rac-C
2
H
4
(1-Ind)
2
ZrMe][MeB(C
6
F
5
)
3
] catalyst,
922
and the reactivity of M(secondary alkyl) bonds at
80

C was comparable to that of primary alkyl metallocenes.


302
These relative monomer insertion rates appear
strongly ligand specific. However, when these comparisons are made, it must be borne in mind that different authors
use very different catalysts, as well as different definitions of the term active center: Landis defined the active
species as the product of the first monomer insertion, whereas Bochmann
290
and Busico
297
used the term to describe
the proportion of total [Zr] actively engaged in chain growth based on kinetic measurements. Similarly important is
the fact that very different activities and polymerization mechanisms are operative with the weakly coordinating
1030 Olefin Polymerizations with Group IV Metal Catalysts
MAO and [B(C
6
F
5
)
4
]

counterions on one hand,


290,297
and the tighly bound [MeB(C
6
F
5
)
3
]

counterion on the
other.
101,156,302
4.09.3 ZieglerNatta Polymerizations with Heterogeneous Catalysts
4.09.3.1 Catalyst Structure and Characterization
As indicated above, ZieglerNatta catalysts occupy a dominant position in polyolefins manufacture, and have
evolved from first- and second-generation titanium trichloride catalysts, developed up until the 1970s
313315
and
described in COMC (1982),
316
to the high-activity MgCl
2
-supported catalysts used in modern industrial processes.
317
PP catalysts, often termed third-, fourth-, and fifth-generation, according to the catalyst performance and catalyst
composition, comprise magnesium chloride as support material, a titanium component (usually TiCl
4
), and an
electron donor (Lewis base). The basis for the development of the high-activity supported catalysts lay in the
discovery, in the late 1960s, of activated MgCl
2
able to support TiCl
4
and give high catalyst activity, and the
subsequent discovery, in the mid-1970s, of electron donors (Lewis bases) capable of increasing the stereospecificity
of the catalyst so that (highly) isotactic PP could be obtained.
318321
A further feature which contributed greatly to the
commercial success of MgCl
2
-supported catalysts was the development of spherical catalysts with controlled particle
size and porosity.
317
High-activity ZieglerNatta catalysts comprising MgCl
2
, TiCl
4
, and an internal electron donor are typically
used in combination with an aluminum alkyl co-catalyst such as AlEt
3
and an external electron donor added in
polymerization. The third-generation catalyst systems contained ethyl benzoate as internal donor and a second
aromatic ester as external donor, whereas the now widely used fourth-generation catalysts contain a diester (e.g.,
diisobutyl phthalate) as internal donor and are used in combination with an alkoxysilane external donor of type
RR
1
Si(OMe)
2
or RSi(OMe)
3
. The most effective alkoxysilane donors for high stereospecificity are methoxysilanes
containing relatively bulky groups c to the silicon atom.
322325
In the early stages of MgCl
2
-supported catalyst
development, activated magnesium chloride was prepared by ball milling in the presence of ethyl benzoate,
leading to the formation of very small (_3 nm thick) primary crystallites within each particle.
315
High resolution
transmission electron microscopy has shown that the MgCl
2
crystal structure is severely distorted by ball milling,
changing the structure from crystalline to amorphous.
326
Nowadays, the activated support is prepared by chemical
means, such as via complex formation of MgCl
2
and an alcohol, or by reaction of a magnesium alkyl or alkoxide
with a chlorinating agent or TiCl
4
. A chemical rather than a physical route to catalyst preparation can also lead to
a more uniform titanium distribution.
327
Many of these approaches are also effective for the preparation of
catalysts having controlled particle size and morphology. For example, the cooling of emulsions of molten
MgCl
2
?nEtOH in paraffin oil gives almost perfectly spherical supports, which are then converted into the
catalysts.
317
Characterization of a number of adducts of magnesium chloride and ethanol has been described by
Bart and Roovers
328
, while Sozzani et al.
329
have recently reported the use of advanced solid-state NMR
techniques to determine the various components present in MgCl
2
?nEtOH adducts. It was found that n values
of 1.5 and 2.8 represented stable and well-defined complexes.
A typical catalyst preparation involves reaction of the MgCl
2
?nEtOH support with excess TiCl
4
in the presence of
an internal electron donor. Temperatures of at least 80

C and at least two TiCl


4
treatment steps are normally used,
in order to obtain high-performance catalysts in which the titanium is mainly present as TiCl
4
rather than the
TiCl
3
OEt generated in the initial reaction with the support.
The function of the internal donor in MgCl
2
-supported catalysts is twofold. One function is to stabilize small
primary crystallites of magnesium chloride; the other is to control the amount and distribution of TiCl
4
in the final
catalyst, affecting stereoselectivity. Activated magnesium chloride has a disordered structure comprising very small
lamellae. X-ray diffraction studies have revealed rotational disorder in the stacking of the ClMgCl triple
layers.
330,331
Small MgCl
2
crystallite size and large rotational disorder appear to give high catalyst activity.
332
Giannini
319
has indicated that, on preferential lateral cleavage surfaces, the magnesium atoms are coordinated with
four or five chlorine atoms, as opposed to six chlorine atoms in the bulk of the crystal. These lateral cuts correspond to
the (110) and (100) faces of MgCl
2
, as shown in Figure 15. It has been proposed that bridged, binuclear Ti
2
Cl
8
species
can coordinate to the (100) face of MgCl
2
and give rise to the formation of chiral, isospecific active species (Figure 16),
it being pointed out that Ti
2
Cl
6
species formed by reduction on contact with AlEt
3
would resemble analogous species
in TiCl
3
catalysts.
333,334
An extended X-ray absorption fine structure (EXAFS) investigation of a MgCl
2
/TiCl
4
catalyst has indicated the presence of dimeric TiCl
4
complexes on the (100) face of MgCl
2
.
335
It has been
Olefin Polymerizations with Group IV Metal Catalysts 1031
suggested
317
that a possible function of the internal donor is preferential coordination on the more acidic (110) face of
MgCl
2
, such that this face is prevailingly occupied by donor and the (100) face is prevailingly occupied by Ti
2
Cl
8
dimers. However, as outlined in Section 4.09.3.3.3, this is by no means the only likely mode of coordination of Ti
species to the support.
Analytical studies have indicated that a monoester internal donor such as ethyl benzoate is coordinated to MgCl
2
and not to TiCl
4
.
336
In the search for donors giving catalysts with improved performance, it was considered
337
that
bidentate donors should be able to form strong chelating complexes with tetracoordinate Mg atoms on the (110) face
of MgCl
2
, or binuclear complexes with two pentacoordinate Mg atoms on the (100) face. This led to the development
of the MgCl
2
/TiCl
4
/phthalate ester catalysts, used, as indicated above, in combination with an alkoxysilane as
external donor. The requirement for an external donor when using catalysts containing a benzoate or phthalate
ester is due to the fact that, when the catalyst is brought into contact with the co-catalyst, a large proportion of the
internal donor is lost as a result of alkylation and/or complexation reactions, which in the absence of an external donor
would lead to poor stereospecificity. When the external donor is present, contact of the catalyst components leads to
replacement of the internal donor by the external donor. The most active and stereospecific systems were found to be
those which allowed the highest incorporation of external donor,
338
the effectiveness of a catalyst system depending
more on the combination of donors rather than on the individual internal or external donor.
Recently, research on MgCl
2
-supported catalysts has led to systems not requiring the use of an external donor. This
required the identification of bidentate internal donors which not only had the right oxygenoxygen distance for
effective coordination with MgCl
2
but which, unlike phthalate esters, were not removed from the support on contact
with AlEt
3
, and which were unreactive with TiCl
4
during catalyst preparation. It was found
337,339341
that certain
100
110
CI
Mg
cut
cut
Figure 15 Model of a MgCl
2
layer showing the (100) and (110) lateral cuts.
Figure 16 Model showing dimeric and monomeric Ti species on a (100) lateral cut of MgCl
2
.
1032 Olefin Polymerizations with Group IV Metal Catalysts
2,2-disubstituted-1,3-dimethoxypropanes met all these criteria. The best performance was obtained when bulky
substituents in the 2-position resulted in the diether having a most probable conformation
342
with an oxygenoxygen
distance in the range 2.83.2

A. The successive generations of high-activity MgCl
2
-supported catalyst systems for
PP are summarized below:
(i) MgCl
2
/TiCl
4
/ethyl benzoateAlR
3
aromatic ester,
(ii) MgCl
2
/TiCl
4
/phthalate esterAlR
3
alkoxysilane, and
(iii) MgCl
2
/TiCl
4
/dietherAlR
3
.
Catalyst performance has improved considerably with each generation. The PP yield obtained under typical
polymerization conditions (liquid monomer, in the presence of hydrogen, 70

C, 12 h) has increased from


1530 kg g
1
catalyst for the third-generation ethyl benzoate-containing catalysts to 3080 kg g
1
catalyst for the
fourth-generation phthalate-based catalysts. With the recently developed fifth-generation catalysts, containing a
diether as internal donor, yields of 80160 kg g
1
catalyst can be achieved. These different catalysts also display
different kinetic profiles in propylene polymerization. The catalysts containing a diether as internal donor exhibit
very stable activities during polymerization. A low rate of catalyst decay during polymerization is also obtained with
the catalyst system MgCl
2
/TiCl
4
/phthalate esterAlR
3
alkoxysilane, whereas the system MgCl
2
/TiCl
4
/ethyl benzo-
ateAlR
3
aromatic ester has a very high initial polymerization activity but also a high decay rate, limiting the final
polymer yield. The rapid decay in activity can at least partially be ascribed to the use of an ester as external as well as
internal donor, the ester being able to react with titaniumhydrogen bonds formed in chain transfer with hydrogen,
generating TiO bonds inactive for chain propagation.
343
Most recently, a further family of MgCl
2
-supported catalysts has been developed,
344,345
in which the internal donor
is a succinate rather than a phthalate ester. As is the case with the phthalate-based catalysts, an alkoxysilane is used as
external donor. The essential difference between these catalysts is that the succinate-based systems produce PP
having much broader molecular mass distribution, discussed in Section 4.09.3.4.
4.09.3.2 Polymer Particle Growth
An important feature of olefin polymerization using heterogeneous catalysts concerns the characteristics of particle
growth during the course of polymerization, taking into account such aspects as mass and heat transfer. Ineffective
monomer mass transfer can limit the catalyst productivity, while ineffective removal of heat of polymerization from
the growing particle in the early stages of polymerization can cause the formation of hot spots, which may in turn lead
to catalyst decay. In the absence of pre-polymerization, exotherms of up to 20

C have been measured for individual


catalyst particles in the early stages of polymerization.
346
Typically, heterogeneous ZieglerNatta catalysts have particle sizes in the range 10100 mm. Each particle
comprises millions of primary crystallites with sizes of up to about 15 nm. On contacting the catalyst components,
at the start of polymerization, co-catalyst and monomer diffuse through the catalyst particle, and polymerization takes
place on the surface of each primary crystallite within the particle. As solid, crystalline polymer is formed, the primary
crystallites are pushed outward and apart as the particle grows, analogous to the expanding universe. The particle
shape is retained, and this phenomenon is therefore referred to as replication (Figure 17). Ideally, the catalyst particle
should have spherical morphology and controllable porosity. It is important that the mechanical strength of the
catalyst is high enough to prevent disintegration, but low enough to allow progressive expansion as polymerization
proceeds.
347
Polymer
Pre-polymer
Catalyst
Figure 17 Replication phenomenon in polymerization.
Olefin Polymerizations with Group IV Metal Catalysts 1033
Under appropriate polymerization conditions, polymer particles can be obtained with an internal morphology that
can range from a compact to a porous structure.
348
In what is termed reactor granule technology, porous polymer
particles can be produced which can then function as a microreactor for the polymerization of other monomers within
the solid matrix. A PP skin encloses the second polymer phase, thereby preventing coalescence of particles in which
the second phase is an amorphous, low-melting material. Reactor granule technology is able to provide products
ranging from superstiff, high-fluidity PP homopolymers to stiff/impact or clear/impact heterophasic co-polymers and
supersoft alloys, produced using the Catalloy process.
345,347,348
The feasibility of producing heterophasic alloys
containing up to 70% of elastomeric co-polymers arises from the use of a controlled porosity catalyst and the ability to
control the porosity of the growing polymer particle during the early stages of polymerization. Pre-polymerization is
applied to give the particles sufficient heat capacity to withstand injection into a gas-phase polymerization step.
Various models describing particle growth during olefin polymerization have been developed and recently
reviewed by McKenna and Soares.
349
One of the most popular models is the multigrain model, put forward by
Ray and co-workers,
350
in which the monomer diffuses into the catalyst macroparticle and polymerizes on the surface
of the microparticles within, causing progressive expansion of the macroparticle as polymerization proceeds. An
investigation by Kakugo et al.
351
of nascent polymer morphology obtained using a TiCl
3
catalyst showed that the
polymer particle comprised numerous globules, each of which contained some tens of much smaller primary particles.
Recently, a model for particle growth with MgCl
2
-supported catalysts has been proposed by Cecchin,
345,352
who has
also provided evidence for polymer subglobule formation within the growing particle. Again, these subglobules
originate from clusters of primary crystallites, but the crystallites themselves are pushed to the surface of each
subglobule as the polymer forms. This model explains the fact that, in the preparation of heterophasic co-polymers
via propylene homopolymerization followed by ethylenepropylene co-polymerization, the rubbery EP co-polymer
is formed at the surface of the homopolymer subglobules, gradually filling up the pores within the particle. Evidence
for drifting of catalyst microparticles to the surface of polymer (sub)globules has been provided by scanning electron
microscopy studies of pre-polymerized catalyst particles.
353
Extensive fragmentation and uniform particle growth are key features in the replication process and are dependent
on a high surface area, a homogeneous distribution of catalytically active centers throughout the particle, and free
access of the monomer to the innermost zones of the particle. In the case of ethylene polymerization, the latter is not
always the case, and it is frequently observed that polymer growth starts at and near the particle surface, leading to the
formation of a shell of PE around the catalyst particle. This imposes a diffusion limitation, preventing free access of
the monomer to active sites within the particle. Polymerization then takes place layer by layer, as the monomer
gradually diffuses through the outer layers to the core, resulting in an onion-type internal morphology.
354
This
mechanism of particle growth is associated with a kinetic profile in which an initial induction period is followed by an
acceleration period, after which, in the absence of chemical deactivation, a stationary rate is obtained. Each
polymerizing particle can be considered as a microreactor with its own mass and heat balance.
355
Ethylene poly-
merization activity can be greatly increased by first carrying out pre-polymerization with propylene, which results in a
significant lowering in the monomer diffusion barrier in the subsequent ethylene polymerization.
356
In a further
refinement of particle growth models, effects of not only diffusion but also monomer convection have been taken into
account, the convection being driven by a pressure gradient created by monomer consumption within the particle.
357
Serious mass-transfer limitations are less common in propylene polymerization. Studies with MgCl
2
-supported
ZieglerNatta catalysts have revealed uniform polymer particle growth in the early stages of polymerization, the
catalyst support undergoing even and progressive fragmentation.
358
4.09.3.3 Mechanistic Studies of ZieglerNatta Catalysts
4.09.3.3.1 Oxidation state
It is generally accepted
315
that in the TiCl
3
/AlEt
2
Cl system, the active species are Ti(III). In the case of MgCl
2
-
supported catalysts, in which the titanium is generally tetravalent prior to contact with the co-catalyst, the dominant
active species in polymerization also appears to be Ti(III), although active Ti(IV) and Ti(II) species may also be
present. Most reports indicate that both Ti(III) and Ti(II) species are active for ethylene polymerization, whereas
Ti(II) has little or no activity in propylene polymerization.
359
The immobilization of a titanium chloride on a
magnesium chloride support lowers the susceptibility to (over)reduction on contact with an Al alkyl,
360
but
Kashiwa
361
found that treatment of an MgCl
2
/TiCl
4
/ethyl benzoate catalyst with excess AlEt
3
for 2 h at 60

C
resulted in reduction to 80% Ti(II) and 20% Ti(III). The catalyst was moderately active for ethylene polymerization
but not for propylene, although co-polymerization was possible. After reoxidation, the catalyst was active for
1034 Olefin Polymerizations with Group IV Metal Catalysts
propylene polymerization and showed increased activity in ethylene polymerization. With a similar catalyst but at
milder reaction conditions and in the presence of an external donor (methyl p-toluate), Chien
362
observed that 85% of
the initial Ti
4
was reduced to Ti
3
and 15% to Ti
2
, while in subsequent studies Ti
4
(35%), Ti
3
(25%), and Ti
2
(40%) could be detected.
363
Brief (2 min) treatment of a similar catalyst with AlBu
i
3
, with or without ethyl benzoate as
external donor, resulted in almost exclusively Ti
3
, about 30% of which could be detected by ESR and the remainder
only after contact with pyridine.
364
A combination of ESR and titration techniques was used by Chien to estimate the
Ti oxidation states in a catalyst containing a phthalate as internal donor.
365
Contact of the catalyst with AlEt
3
/
phenyltriethoxysilane resulted in oxidation states Ti
4
: Ti
3
: Ti
2
in the proportions 28.1 : 38.5 : 33.4, whereas in the
absence of phenyltriethoxysilane the proportions were 7.0 : 73.7 : 19.3.
X-ray photoelectron spectroscopy (XPS) has also been used to investigate the nature of the species formed on
contacting MgCl
2
-supported catalysts with Al-alkyls. Terano and co-workers
366
observed that the binding energy of
MgCl
2
-supported TiCl
4
was unaffected by the presence of an internal donor, indicating no direct interaction of the
donor with the titanium, but contact with AlR
3
resulted in a shift to lower binding energy, indicative of reduction of
Ti
4
.
4.09.3.3.2 Number of active centers
Various methods, generally based on kinetic and radiotagging methods, have been developed for the determination of
the number of active centers in ZieglerNatta catalysts.
367
Labeling methods determine the proportion of total metal
that carries a polymeryl chain, whereas kinetic methods can be used to measure the proportion of Mpolymeryl chains
actively engaged in chain growth. Hence, very different values must be expected, arising from different definitions.
For MgCl
2
-supported catalysts, active center (C
*
) values of around 4% (of total Ti present) have been obtained from
stopped-flow experiments.
368
C
*
values obtained by other techniques, notably
14
CO quenching of propylene
polymerization, have ranged from 1% or less
369
to more than 20%.
370
Clearly, there are large differences in C
*
values
obtained by different groups, but it is consistently found that the major proportion of the Ti present in ZieglerNatta
catalysts is inactive. The propagation rate constant (k
p
) values for isospecific active sites are around an order of
magnitude greater than those for weakly specific sites.
369,370
The value of k
p
increases significantly in the presence of
hydrogen,
371
in accordance with the reactivation of dormant (2,1-inserted) centers by chain transfer (vide infra).
4.09.3.3.3 Internal/external donor effects and the nature of the active species
The characteristics of ZieglerNatta catalysts for polypropylene are, to a large extent, dependent on the types of
donors present, both in the solid catalyst and added as external donor in polymerization. It is well established that
effective external donors not only increase the isotactic index of PP (the proportion of polymer insoluble in boiling
heptane or in xylene at 25

C), but can also increase in absolute terms the amount of isotactic polymer formed. This
has been demonstrated by Kashiwa
372
for the catalyst system MgCl
2
/TiCl
4
AlEt
3
. An increase in the molecular mass
and stereoregularity of the isotactic fraction was also noted. Similar trends are apparent with catalyst systems of type
MgCl
2
/TiCl
4
/phthalate esterAlR
3
alkoxysilane.
373
Kakugo
374
has used elution fractionation to demonstrate that the
external donor not only decreases atactics formation but also increases the degree of steric control at isospecific
sites. Soga has reported that an almost aspecific MgCl
2
/TiCl
3
catalyst, with a very low content of (probably isolated,
monomeric) Ti species, could be rendered isospecific by the addition of ethyl benzoate as external donor
375
or by
using Cp
2
TiMe
2
as co-catalyst.
376
It was suggested
377
that in both cases, aspecific sites having two coordination
vacancies could be converted to isospecific sites by blocking one of the two vacancies.
A powerful technique to study the effects of electron donors on site selectivity in ZieglerNatta catalysts is the
determination of the stereoregularity of the first insertion step in propylene polymerization. Sacchi et al.
378,379
have
investigated the effect of Lewis bases on the first-step stereoregularity resulting from propylene insertion into a
TiEt bond formed via chain transfer with
13
C-enriched AlEt
3
. First-step stereoregularity is particularly sensitive to
the steric environment of the active center, due to the fact that the stereospecificity of the first monomer insertion is
always lower than that of the following propagation steps. For example, with an MgCl
2
/TiCl
4
/diisobutyl phthalate
catalyst, it was found
379
that the mole fraction of erythro- (isotactic) placement in the isotactic polymer fraction was
0.67 with no external donor, 0.82 with MeSi(OEt)
3
, and 0.92 with PhSi(OEt)
3
. It could be concluded that the
alkoxysilane external donor was present in the environment of at least part of the isospecific centers. Subsequent
studies
380,381
indicated that similar considerations apply to diether donors.
Recently, significant advances have been made in understanding the fundamental factors determining the
performance of state-of-the-art MgCl
2
-supported catalysts. Studies by Busico et al.
382
have shown that the chain
irregularities in iPP prepared using heterogeneous catalysts are not randomly distributed along the chain but are
Olefin Polymerizations with Group IV Metal Catalysts 1035
clustered. The chain can therefore contain, in addition to highly isotactic blocks, sequences which can be attributed
to weakly isotactic (isotactoid) and to syndiotactic blocks. This implies that the active site can isomerize very rapidly
(during the growth time of a single polymer chain, i.e., in less than a second) between three different propagating
species. The same sequences are present, but in different amounts, in both the soluble and the insoluble fractions.
The polymer can therefore be considered to have a stereoblock structure in which highly isotactic sequences
alternate with defective isotactic (isotactoid) and with syndiotactoid sequences. A mechanistic model has been
formulated in which the relative contributions of these sequences can be related to site transformations involving
the presence or absence of steric hindrance in the vicinity of the active species. The
13
C NMR studies have
indicated
383
the presence of C
1
-symmetric active species in MgCl
2
-supported catalysts, with a mechanism of isotactic
propagation which is analogous to that for certain C
1
-symmetric metallocenes, in the sense that propylene insertion at
a highly enantioselective site tends to be followed by chain backskip rather than a less regio- and stereoselective
insertion when the chain is in the coordination position previously occupied by the monomer. The probability of
chain backskip will increase with decreasing monomer concentration, and it has indeed been confirmed that
increased polymer isotacticity is obtained at low monomer concentration. It is proposed
382
that a (temporary) loss of
steric hindrance from one side of an active species with local C
2
-symmetry, giving a C
1
-symmetric species, may result
in a transition from highly isospecific to moderately isospecific propagation. Loss of steric hindrance on both sides can
lead to syndiospecific propagation in which chain-end control becomes operative. The model is illustrated in
Figure 18.
If it is considered that the steric hindrance in the vicinity of the active species can result from the presence of a
donor molecule, and that the coordination of such a donor is reversible, the above model provides us with an
explanation for the fact that strongly coordinating, stereorigid donors typically give stereoregular polymers in which
the highly isotactic sequences predominate.
384
It has been suggested
385
that the high stereospecificity obtained using
silanes having one or more bulky hydrocarbyl groups is due to the silane stabilizing fluctuating isospecific sites, the
bulky hydrocarbyl groups protecting the silane from removal from the catalyst surface via complexation with
aluminum alkyl.
Several types of active species in which the presence of a donor in the vicinity of the active Ti atom is necessary for
high isospecificity have been proposed,
386
although the exact structure of the active species is still by no means
resolved. In particular, it is still uncertain whether the active Ti is located on the (100) face of MgCl
2
, as originally
proposed,
333,334
or on the (110) face. There is strong evidence that the dominant coordination mode of diether donors
to MgCl
2
is via bidentate coordination on the (110) face.
387,388
It has also been shown
389
that the use of a diether as
external donor in combination with a MgCl
2
/TiCl
4
/phthalate ester catalyst gives active species which are very similar
to those present when the diether is used as internal donor. This could imply that the dominant coordination mode of
phthalate esters is also bidentate coordination on the (110) face of MgCl
2
. Bearing in mind the evidence for the
presence of a donor in the environment of at least part of the isospecific active centers, it thus seems likely that active
Ti is also located on the (110) face of MgCl
2
, or on an edge position adjacent to the (110) face.
386
In contrast to the
diether-based catalysts, molecular modeling studies have indicated the succinate ester complexes with both the (110)
and the (100) faces of MgCl
2
, via three or four different modes of interaction.
390
This is likely to be a contributory
factor to the formation of a broad range of active species in the succinate-based system.
Isospecific active species not requiring the presence of a donor for high stereospecificity have been proposed by
Boero,
391
while Terano
392
has suggested various active species in which coordination of AlEt
3
or AlEt
2
Cl in the
vicinity of the titanium atom can lead to high stereospecificity. Pre-contact of a MgCl
2
/TiCl
4
catalyst with AlEt
3
prior
Ti
Mg
CI
Ti
Mg
CI
Ti
Mg
CI
L1
L2
L2
Figure 18 Model of possible active species for highly isotactic, isotactoid, and syndiotactic propagation. Reproduced with
permission from Macromolecules 2003, 26, 26162622. 2003, American Chemical Society.
1036 Olefin Polymerizations with Group IV Metal Catalysts
to polymerizations carried out under stopped-flow conditions led to the formation of a small proportion of highly
isotactic material in the resulting polymer. Subsequent studies with a catalyst containing ethyl benzoate as internal
donor showed
393
that the highly isotactic polymer fraction increased when the catalyst was briefly pre-contacted with
AlEt
3
, but extended pre-contact led to a loss of isospecificity due to extraction of the internal donor. Solid-state
27
Al NMR spectroscopy has been used to study the surface aluminum compounds formed during the treatment of a
MgCl
2
/TiCl
4
catalyst with aluminum alkyls; it was proposed that the proportion of Ti in active state was dependent
on adsorption/desorption of Al species. Strong adsorption of AlEt
2
Cl, blocking active centers, was put forward to
explain the low catalytic activity obtained using AlEt
2
Cl as co-catalyst.
394
It has also been suggested
384
that the
catalyst system MgCl
2
/TiCl
4
/ethyl benzoateAlEt
3
aromatic ester contains species for which high selectivity is not
dependent on the presence of an electron donor in the immediate vicinity of the active site, taking into account that it
is unlikely that monoester donors can provide the stable coordination shown by highly stereoregulating diether and
silane donors.
4.09.3.3.4 Effects of hydrogen
In PP production, hydrogen is used as a chain-transfer agent for polymer molecular mass (melt flow rate) control.
Recent work by Terano and co-workers
395
has shown that, under stopped-flow conditions, hydrogen is only effective
as chain-transfer agent when catalyst and co-catalyst have been pre-contacted. These and subsequent
396,397
results
indicated that effective chain transfer with hydrogen requires the presence of species able to promote the dissociation
of H
2
to atomic hydrogen.
The effect of hydrogen (concentration) on polymer molecular mass is dependent on the catalyst system. An
advantage of catalysts containing a diether donor, in addition to very high activity, is high sensitivity to hydrogen,
so that relatively little hydrogen is required for molecular mass control. This effect can be ascribed to chain transfer
after the occasional secondary (2,1) rather than the usual primary (1,2) insertion, a 2,1-insertion slowing down a
subsequent monomer insertion and therefore increasing the probability of chain transfer.
398
Reactivation of dor-
mant (2,1-inserted) species via chain transfer with hydrogen also explains the frequently observed activating effect
of hydrogen in propylene polymerization, giving yields which may be around 3 times those observed in the complete
absence of hydrogen.
399
These conclusions have been based on the
13
C NMR determination of the relative
proportions of i-Bu- and n-Bu-terminated chains, resulting from chain transfer with hydrogen after primary and
secondary insertion, respectively Equations (8) and (9).
Ti-CH
2
CH(CH
3
)[CH
2
CH(CH
3
)]
n
Pr + H
2
Ti-H + i-Bu-CH(CH
3
)[CH
2
CH(CH
3
)]
n 1
Pr (8)
Ti-CH(CH
3
)CH
2
[CH
2
CH(CH
3
)]
n
Pr + H
2
Ti-H + n-Bu-CH(CH
3
)[CH
2
CH(CH
3
)]
n 1
Pr (9)
The probability of the 2,1-insertion is much higher for the first insertion of propylene into a TiH bond. When
followed by a series of 1,2-insertions, this results in the formation of a 2,3-dimethylbutyl terminal unit, as opposed to
n-propyl when the first insertion is primary. Significant proportions of 2,3-dimethylbutyl end groups have been found
in polymers prepared using both metallocene
219,400
and ZieglerNatta catalysts.
401
The relatively low propagation
rate of 2,1-inserted species also applies to the Ti-isopropyl reaction product of 2,1-insertion into TiH, making a
further transfer reaction with hydrogen (generating propane) more probable than chain growth to form a 2,3-
dimethylbutyl end group.
401
It has been suggested
402
that the hydrogen activation effect is related to the hydro-
genolysis of the Tiisopropyl unit, but this seems unlikely, as in this case hydrogen is involved in both the formation
and removal of such dormant species.
The introduction of hydrogen in ethylene polymerization with a ZieglerNatta catalyst causes a significant
decrease in the polymerization rate, in contrast to what is observed in propylene polymerization. Kissin
403407
has
proposed that this can be explained by the formation of dormant TiCH
2
CH
3
species formed by insertion of
ethylene into TiH after chain transfer with hydrogen. The low propagation activity of the TiEt species was
ascribed to u-agostic stabilization. However, this proposal was not supported by Garoff,
408
who found that increasing
the amount of hydrogen present in polymerization tended to slow down the activation of catalysts in which the
titanium was present as Ti(IV).
A strong deactivating effect of hydrogen has been noted in ethylene polymerization with vanadium-based catalysts
such as MgCl
2
/VCl
4
AlBu
i
3
.
409
This deactivation was reversible, activity being restored when hydrogen was
removed. It was proposed that the deactivating effect of hydrogen arose from the reaction of VH species with the
Olefin Polymerizations with Group IV Metal Catalysts 1037
co-catalyst to give VR and AlR
2
H, the latter component then blocking the active sites via the formation of VHAl-
bridged complexes. Hydrogen had no effect on the number of active centers but decreased the propagation rate constant,
in line with the mechanism proposed.
410
A comparison of vanadium- and titanium-based catalysts for ethylene poly-
merization has shown that hydrogen is a more effective chain-transfer agent with the vanadium catalysts.
411
A half-order
dependency of chain transfer on the hydrogen concentration was found, in line with dissociative adsorption of hydrogen.
For propylene polymerization, it has been demonstrated that the chain transfer is dependent not only on regio-
but also on stereo-selectivity.
412
This is in keeping with the tendency that, with catalyst systems of the type
MgCl
2
/TiCl
4
/phthalate esterAlR
3
alkoxysilane, the silanes which give the most stereoregular isotactic polymer also
give relatively low hydrogen response.
4.09.3.3.5 Effects of temperature
PP production using ZieglerNatta catalysts is generally carried out at a polymerization temperature of around 70

C.
The effect of temperature on catalyst stereospecificity is dependent on the nature of the catalyst. With TiCl
3
catalysts,
PP stereoregularity decreases as the polymerization temperature is increased,
315
but with MgCl
2
-supported catalysts the
opposite trend has been found, an increase in polymerization temperature leading not only to higher proportions of
isotactic polymer but also to increased stereoregularity of the isotactic fraction.
413,414
This effect was ascribed to a greater
increase in productivity, with increasing temperature, for highly isospecific as opposed to weakly isospecific centers.
An increase in polymerization temperature from 70 to 100

C has been found to lead to significant decreases in both


catalyst productivity and polymer molecular mass, chain transfer to Al being particularly rapid at high temperature using
AlEt
3
as co-catalyst in combination with a catalyst containing dioctyl phthalate as internal donor and with an alkoxysilane
as external donor.
415
Decreased polymer melting point was also obtained, ascribed to the formation and co-polymeriza-
tion of a small amount of ethylene, produced by the decomposition of AlEt
3
; this could be avoided by the use of AlBu
i
3
as
co-catalyst.
416
It has recently been reported that the use of AlBu
i
3
together with a MgCl
2
/TiCl
4
/diether catalyst gives
similar polymerization activities at 70 and 100

C, whereas with AlEt


3
decreased activity was observed at 100

C.
417
4.09.3.4 Polyolefins Accessible from ZieglerNatta Catalysts
ZieglerNatta catalysts are multi-site systems, which contain a range of different active centers having different
relative rates of chain propagation and chain transfer. Each individual site produces a SchulzFlory distribution with
M

w
,M

n
= 2 and with M

z
,M

n
= 1.5, and the overall polymer molecular mass distribution represents a combination of
these individual distributions and is therefore broader than that which is produced by metallocene and related single-
center catalysts. Kissin, via deconvolution of GPC traces, estimated that the overall distribution could be described by
considering the presence of five different active centers.
418
In ethylene/1-hexene co-polymerization, the centers
giving the highest molecular masses were those giving negligible hexene incorporation. That the relatively broad
molecular mass distribution of ZieglerNatta polyolefins is due to active center distribution and not to monomer
diffusion or other effects has been demonstrated by the use of stopped-flow polymerization,
419
which showed that
molecular mass distribution (MWD) was unaffected by the polymerization time.
By varying the nature of the electron donors (esters, silanes, diethers) present in ZieglerNatta catalyst systems for
PP, it is possible to control the polymer tacticity, molecular mass, and molecular mass distribution to produce a range
of polymers having the processing and end-use properties required for very different applications. The donors present
in the catalyst system play an active role in the formation or modification of isospecific sites, and the polymer
molecular mass distribution depends on the relative contribution and hydrogen response (i.e., sensitivity to chain
transfer with hydrogen) of each type of active site. The incorporation of an external donor into the catalyst system
generally leads to an increase in molecular mass, the magnitude of the molecular mass (MW) increase depending on
the nature of the donor. The characteristics of different catalyst systems with regard to PP molecular mass distribu-
tion are as follows:
345
Internal donor External donor
M

w
,M

n
Diether 55.5
Phthalate Alkoxysilane 6.58
Succinate Alkoxysilane 1015
1038 Olefin Polymerizations with Group IV Metal Catalysts
It can be seen that the diether-based catalysts give relatively narrow molecular mass distribution. A narrow
molecular mass distribution, and relatively low molecular mass, are advantageous in fiber-spinning applications. In
contrast, extrusion of pipes and thick sheets requires high melt strength, and therefore relatively high molecular mass
and broad molecular mass distribution. A broad molecular mass distribution, along with high isotactic stereoregularity,
is also beneficial for high crystallinity and therefore high rigidity. The new succinate-based catalysts enable very
broad molecular mass distribution PP homopolymers to be produced in a single reactor and are also of interest for the
production of heterophasic co-polymers having an improved balance of stiffness and impact strength, taking into
account that the incorporation of a rubbery (ethylene/propylene, EP) co-polymer phase into a PP homopolymer
matrix increases impact strength but at the same time leads to decreased stiffness.
The relatively narrow PP molecular mass distributions obtained using diether-based catalysts can be attributed to
the fact that in these systems even the most highly stereospecific active sites are not totally regiospecific. A
proportion of approximately one secondary insertion for every 2000 primary insertions at highly isospecific sites has
been noted for the system MgCl
2
/TiCl
4
/dietherAlR
3
.
398
The probability of chain transfer with hydrogen after a
secondary insertion is such that this is sufficient to prevent the formation of very high molecular mass chains, taking
into account that the highest molecular mass fraction of the polymer is formed on the active species having the
highest isospecificity. The broader molecular mass distributions obtained with catalysts containing ester internal
donors are likely to be due to the presence of (some) isospecific active sites having very high regiospecificity and
therefore lower hydrogen sensitivity.
420
Such results illustrate the profound effect of catalyst regio- and stereo-
specificity distribution on both molecular mass control and polymer molecular mass distribution and properties.
In addition to their dominant position in PP manufacture, ZieglerNatta catalysts are widely used in the production
of HDPE and LLDPE. More than half the world production of HDPE is based on ZieglerNatta catalysts, chromium
catalysts also being widely used. In LLDPE production, ZieglerNatta catalysts account for more than 90% of the
total production, although single-center catalysts are making strong inroads into this market.
In HDPE production, high-mileage ZieglerNatta catalysts are used in the cascade process, in which polymeriza-
tion reactors in series are used to give reactor blends with improved properties for film and pipe applications.
421
Broad
molecular mass distribution can be obtained by the use of different hydrogen concentrations in each reactor. In
addition, the process can be designed to give a low molecular mass homopolymer in the first reactor and a high
molecular mass co-polymer in the second. The high molecular mass co-polymer chains function as tie molecules
linking the crystalline, homopolymer domains, thereby leading to high stress crack resistance of the polymer. This
process allows an inverse co-monomer distribution to be obtained, in the sense that the co-monomer is in the high
molecular mass fraction, counteracting the general tendency of ZieglerNatta catalysts to incorporate the co-mono-
mer mainly in the low molecular mass chains. The latter feature is an important consideration in ZieglerNatta
catalyst design for LLDPE. Co-monomer incorporation is highest at the most open catalytic centers, whereas
sterically hindered centers will tend to give PE chains with little or no co-monomer. The best catalysts for
LLDPE are therefore those that have relatively uniform active center distribution, lacking excessively hindered or
unhindered active sites.
In addition to titanium-based ZieglerNatta catalysts, vanadium-based systems have also been developed for PE
and ethylene-based co-polymers, particularly ethylenepropylenediene rubbers (EPDM). Homogeneous (soluble)
vanadium catalysts produce relatively narrow molecular mass distribution PE, whereas supported V catalysts give
broad molecular mass distribution.
422
Polymerization activity is strongly enhanced by the use of a halogenated
hydrocarbon as promoter in combination with a vanadium catalyst and aluminum alkyl co-catalyst.
422,423
In addition to their widespread use in the production of PE and PP, ZieglerNatta catalysts play an important role
in the production of poly-1-butene, for which both TiCl
3
-based and MgCl
2
-supported catalysts have been developed.
TiCl
3
catalysts have been used with dialkylaluminum halide co-catalysts, AlEt
2
I giving higher isotacticity than
AlEt
2
Cl.
315
Very high isotacticity has been obtained using TiCl
3
in combination with Cp
2
TiMe
2
.
424
Much higher
polymerization activity, as well as high isotacticity and broad molecular mass distribution, is obtained using MgCl
2
-
supported catalysts, for example the catalyst system MgCl
2
/TiCl
4
/diisobutyl phthalateAlEt
3
alkoxysilane.
425
ZieglerNatta catalysts have also been developed for the polymerization of 4-methyl-1-pentene
426
and higher
c-olefins. Polymerization activity decreases with increasing steric bulk of the monomer. For example, with the
catalyst system MgCl
2
/TiCl
4
/ethyl benzoateAlEt
3
ethyl benzoate, the relative activities in propylene, 1-butene, and
4-methyl-1-pentene polymerization were 100 : 80 : 15.
427
For catalyst systems of type MgCl
2
/TiCl
4
/phthalate
esterAlR
3
alkoxysilane, the type of silane required is dependent on the steric bulk of the monomer. An active
center having high stereospecificity in propylene polymerization may be too sterically hindered for effective poly-
merization of a bulkier monomer, propylene/1-butene co-polymerization studies having shown
428
that the incorpora-
tion of 1-butene into the polymer chain decreases with increasing site stereospecificity. This phenomenon is also
Olefin Polymerizations with Group IV Metal Catalysts 1039
illustrated by the fact that non-bulky alkoxysilanes such as Me
3
SiOMe are effective donors in 4-methyl-1-pentene
polymerization,
429
whereas such donors are relatively ineffective in propylene polymerization.
4.09.3.5 Polymerization of Acyclic Internal Olefins
Acyclic internal olefins such as cis- and trans-2-butene can be polymerized using Brookhart-type Ni(II) and Pd(II)
complexes
430
but in the absence of isomerization do not homopolymerize using ZieglerNatta catalysts, although
examples of co-polymerizations of ethylene with an internal olefin are known.
313
The application of ZieglerNatta catalysts to monomer isomerization polymerization, in which an olefin such as 2-
butene is first isomerized to 1-butene and then polymerized, has been described in an extensive series of papers by
Endo and Otsu.
431
The non-participation of 2-butene in ZieglerNatta polymerization (taking precautions to exclude
the possibility of carbocationic polymerization)
432
also allowed the selective polymerization of 1-butene from
mixtures of butene isomers.
433
It was proposed that isomerization and polymerization took place on Ti(II) and
Ti(III) species, respectively.
434
The relative rates of isomerization and polymerization were dependent on the catalyst
system, polymerization being more rapid with TiCl
3
/AlEt
3
than with Ti(OBu)
4
/AlEt
3
.
With the system TiCl
3
/AlEt
3
, the rate of isomerization of cis-2-butene to 1-butene can be increased by the
incorporation of a late transition metal component such as NiCl
2
.
435
Isomerization polymerization using 2-butene
has also been shown to be useful for the synthesis of co-polymers of 1-butene with c-carbon-branched 1-alkenes
having low polymerization activity.
436
The difference in reactivity between the branched monomer and 1-butene is
compensated for by the low concentration of 1-butene formed by isomerization from 2-butene.
4.09.3.6 Major Industrial Processes
The major processes for polyolefins production using ZieglerNatta catalysts involve polymerization in the gas phase
or in slurry, including bulk liquid monomer in the case of propylene. LLDPE is also produced via a solution process
operating at temperatures in the range 130250

C.
Gas-phase processes operating in the range 70115

C and pressures of 2030 bar are widely used for the production
of LLDPE and HDPE. Fluidized bed reactors, such as those used in the Unipol process,
437
typically comprise a
vertical cylindrical reactor containing a fluidized bed of catalyst/polymer particles above a perforated plate (gas
distribution plate), the velocity of the circulating gas being sufficient to fluidize the bed but low enough to prevent
particle entrainment from the reactor. Cooling is provided by the fluidizing gas stream, which is recycled to the bed
by means of a compressor or blower, via one or more heat exchangers. Additional cooling can be provided by
operating in the condensed mode, which involves the injection via the recycle stream of a condensed monomer
and/or a volatile liquid such as isopentane, to give evaporative cooling.
Slurry processes in hydrocarbon diluent are used in the production of HDPE, including bimodal polymers
produced in the cascade process in which different hydrogen concentrations are applied in two or more reactors in
series. Liquid loop reactors are generally used with a light hydrocarbon diluent such as isobutane, whereas heavier
hydrocarbon diluents are typically used in continuous stirred tank reactors.
The development of the Borstar PE process, by Borealis, is a relatively recent development in multi-reactor
processes. The foundation of this process is the utilization of supercritical propane as diluent in the slurry loop
reactor.
438
Operating the slurry loop in a supercritical condition provides several advantages over the tradition diluent
(isobutane). The solubility of PE drops markedly at the supercritical point of propane, allowing the process to operate
at higher temperatures and reducing the risk of reactor fouling. Supercritical process conditions have also been
developed for propylene polymerization.
439
In PP manufacture, modern bulk (liquid monomer) and gas-phase processes have largely replaced the earlier slurry
processes in which polymerization was carried out in hydrocarbon diluent. The most widely adopted process for PP is
Basells Spheripol process.
317
Homopolymer production involves a pre-polymerization step at relatively low tem-
perature, followed by polymerization in a loop reactor using liquid propylene; random co-polymers are produced by
introducing small quantities of ethylene into the feed. The pre-polymerization step gives a pre-polymer particle with
the capacity to withstand the reaction peak, which occurs on entering the main loop reactor. The addition of one or
two gas-phase reactors for EP co-polymerization makes it possible to produce heterophasic co-polymers containing up
to 40% of E/P rubber within the homopolymer matrix.
The development of the Spheripol process was based on the use of MgCl
2
-supported ZieglerNatta catalysts
having spherical particle morphology. Further catalyst and process development, including the manufacture and use
1040 Olefin Polymerizations with Group IV Metal Catalysts
of catalysts having different degrees of porosity, led to the Catalloy process. This is a highly sophisticated modular
technology, based on three mutually independent gas-phase reactors in series. Random PP co-polymers containing up
to 15% co-monomer can be obtained, as well as heterophasic alloys containing high proportions of multi-monomer co-
polymers. The feasibility of producing reactor-grade polymer blends containing up to 65% rubber phase arises from
the use of a controlled porosity catalyst and the ability to control the porosity of the growing polymer particle during
the early stages of polymerization. Again, pre-polymerization is applied to give particles with sufficient heat capacity
to withstand injection into a gas-phase polymerization step.
The most recent major development in polyolefin process technology has been the introduction of the Spherizone
process, based on a gas-phase multi-zone circulating reactor.
440
In this process, the growing polymer granule is
continuously circulated between two polymerization zones: upward, by fast fluidization, in the riser leg, and
downward, by means of gravity in a packed bed, in the downer leg. The monomer composition and the hydrogen
concentration in the two legs can be different, and the multiple short passes of the growing particle between the two
zones leads to intimate and effective mixing of very different polymer compositions.
4.09.4 Polymerizations with Metallocene Catalysts
4.09.4.1 Ethylene Polymers
4.09.4.1.1 Polyethylene
PE homopolymer (HDPE) differs from the two other major polyethylene materials, the highly branched low-density
PE (LDPE, produced under supercritical conditions in high pressure reactors by means of radical initiators, and out of
the scope of this review) and the LLDPE, (the family of co-polymers of ethylene with 1-butene, 1-hexene, or
1-octene, produced by both heterogeneous ZieglerNatta and single-center catalysts, which is discussed in Section
4.09.4.1.2), by being predominantly linear. Contrary to what one might expect, however, HDPE is a rather sophis-
ticated polymer.
355
Most importantly, its molecular mass distribution (width and shape of the distribution) deter-
mines the rheological behavior and mechanical properties of PE. Secondly, HDPE is seldom perfectly linear, and the
number and length of the few side-branches also affect PE properties and applications. A review on metallocene-
catalyzed ethylene polymerization is available.
441
Metallocene catalysts have been recognized early on to be capable
of extremely high ethylene polymerization activities, to have excellent co-monomer incorporation ability, and to be
prone to generating both long- and short-chain branches. The ansa-cyclopentadienylamido titanium complexes
developed by Dow (dubbed CGCs)
442,443
have dominated the development of metallocene PE and are the most
efficient catalysts at generating these LCBs,
444446
although LCBs have been detected in HDPE from bis(cyclo-
pentadienyl) complexes as well.
446448
Methyl branches in PE made with Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
at high
temperature have also been detected.
445
Formation of ethyl branches has been observed in PE from sterically encumbered complexes such as meso-
C
2
H
4
(Ind)
2
ZrCl
2
,
449,450
meso-Me
2
C(Ind)
2
ZrCl
2
,
451
and rac-CH
2
(1-R-2-Ind)
2
ZrCl
2
complexes.
452
These branches
are formed via chain transfer to ethylene followed by reinsertion of the coordinated polymerene chain, as shown
in Scheme 19.
449
Very likely, the bulkiness of the ligands slows release of the double-bond-terminated PE chain,
increasing the chance of its reinsertion in the formed MC
2
H
5
bond.
A similar mechanism can be invoked to explain formation of LCBs, as shown in Scheme 20. This mechanism seems
kinetically more probable than the one involving formation of a TiH intermediate (Scheme 21) or the o-CH metathesis
mechanism (Scheme 22
146
). In support of the mechanism of Scheme 19 and Scheme 20, we note that both ethyl branches
L
n
M
C
2n
CH
2
=CH
2
L
n
M
C
2n
L
n
M
-H
C
2n
Scheme 19
Olefin Polymerizations with Group IV Metal Catalysts 1041
and butyl branches have been detected in PE produced with a linked homobimetallic zirconium ansa-Cpamido
complex,
103
and both ethyl branches and LCBs are formed in PE from siloxy-substituted bis(indenyl) zirconocenes.
447
Concerning the performance of ansa-Cpamido complexes, many papers discuss the influence of the :-ligand type
and substitution,
453457
bridge type,
458464
o-ligand type and substituents,
463,465,466
and the different metals.
455
An interesting approach has been the development of linked homo- and heterobimetallic cyclopentadienylamido
complexes, that show enhanced LCB formation
467
and produce higher molecular mass PE,
103
as well as increased
co-monomer incorporation (see following section).
Several studies aimed at evaluating the influence of substituents on the cyclopentadienyl ligands,
61,447,468484
of
the nature of the co-catalyst,
61,485
and of the metal
486
of Cp9
2
MX
2
complexes have appeared.
Broad or bimodal molecular mass distributions have been obtained using mixtures of metallocenes,
452,487491
linked homo- and heterobimetallic bis(cyclopentadienyl) complexes for broad molecular mass distribution,
492
or
amino-functionalized bis-Cp complexes.
476
L
n
M
C
2n
CH
2
=CH
2
L
n
M
C
2n
n CH
2
=CH
2
L
n
M
C
2n
C
2m
L
n
M
C
2m
C
2n
Scheme 20 LCB formation by u-H transfer to the ethylene monomer.
L
n
M
C
2n
L
n
M
C
2n
n CH
2
=CH
2
L
n
M
C
2n
C
2m
L
n
M
C
2n
H
C
2m
Scheme 21 LCB formation at a site with two growing chains.
L
n
M
C
2n
L
n
M
H
2
C
H
C
2n
(CH
2
)
m
L
n
M
C
2n
(CH
2
)
m
CH
3
nCH
2
=CH
2
C
2n
(CH
2
)
m
CH
3
L
n
M
Scheme 22 LCB formation by o-bond metathesis.
146
1042 Olefin Polymerizations with Group IV Metal Catalysts
In addition to ansa-Cpamido and bis(cyclopentadienyl) complexes, several monocyclopentadienyl complexes
have also been evaluated for ethylene polymerization.
493502
Concerning the evaluation of catalyst activity, some papers have cast a shadow on the reliability of published
activity data by pointing out the dramatic influence of diffusion limitation in ethylene polymerization.
86,86a,470
4.09.4.1.2 Ethylene/c-olefin co-polymers
Ethylene has been co-polymerized with virtually any conceivable c-olefin, from propylene to vinyl-terminated PE and PP
macromonomers. Ethylene/propylene (E/P) copolymerization to produce saturated rubbers and ethylene/propylene/diene
(EPD) terpolymerization to produce unsaturated, vulcanizable rubbers will be discussed in Section 4.09.4.1.3. 1-Butene,
1-hexene, and 1-octene are the most commonly used co-monomers for the production of LLDPE. Ethylene/octene
co-polymers, developed by Dow and marketed under the Engage tradename, have been shown to have improved thermal
properties compared to ethylene/butene and ethylene/hexene co-polymers.
503
In ethylene/c-olefin (E/O) co-polymeriza-
tions, the critical parameters are co-monomer reactivity and co-monomer distribution. The former is most conveni-
ently described by the relative reactivity parameter, R, defined as the ratio between polymer composition and reactor
medium composition.
Co-monomer distribution is defined as the product of reactivity ratios, r
E
r
O
, and describes the statistics of
intramolecular co-monomer distribution in the polymer chain:
(i) The product r
E
r
O
=1 indicates perfect randomness (Bernoullian co-monomer distribution, with diads (deter-
mined by
13
C NMR analysis) obeying the relationship 4[EE][OO]/[EO]
2
=1).
(ii) The product r
E
r
O
<1 indicates tendency to alternation (comonomer homosequences are below the Bernoullian
value).
(iii) The product r
E
r
O
1 indicates blockiness (comonomer homosequences are above the Bernoullian value).
The product r
E
r
O
is independent of the concentration of the monomers and temperature, but the single r
E
and r
O
values might change with temperature.
504
Several r
E
, r
O
, and r
E
r
O
values have been listed in the literature.
229,254,505507
There are different methods for estimating r
E
and r
O
. Two methods based on NMR analysis use equations
based on the concentrations of diads
508
or triads:
509,510
r
E
=
2[EE[
[EO[

[O[
[E[
and r
O
=
2[OO[
[EO[

[E[
O
r
E
=
2[EEE[ [EEO[
2[EOE[ [OOE[

E
and r
O
=
2[OOO[ [OOE[
2[EOE[ [OOE[

O
.
Here,
E
and
O
are the mole fractions of the two monomers in the polymerization medium.
In addition to the above, the intermolecular co-monomer distribution, that is the extent of the co-monomers
incorporation in different molecular mass fractions of the polymer, is an important aspect. When feasible, this is
evaluated by fractionating the polymer and measuring the co-monomer content in each fraction, by solubility
(TREF),
511
crystallizability (CRYSTAF),
512,513
or by GPC-IR.
514
Whereas ZN catalysts tend to produce LLDPE
co-polymers of broad co-monomer distribution (with more co-monomer incorporated in the lower molecular mass
fractions), single-center catalysts always have a nearly perfect intermolecular co-monomer distribution, and this
aspect is taken as implicit in most co-polymerization studies with the latter systems.
Co-monomer type, amount, and distribution in turn determine the key properties of E/O co-polymers such as
density, melting point (T
m
), and glass transition temperature (T
g
). A few examples of the latter two properties as a
function of co-monomer type and content are shown in Figure 19 (T
m
) and Figure 20 (T
g
). It appears that differences
in melting points are fairly independent of the length of the olefin, while 1-octene is superior in lowering the glass
transition temperatures at low incorporation. Larger differences in T
g
can be appreciated beyond 50 mol% of
incorporated c-olefin, where melting temperatures also become dependent on the polymer tacticity (see Section
4.09.4.2.5). The scatter of the data can be due to several factors, including different DSC instruments, scan rates, and
calibrations, different co-monomer distributions produced by the different metallocene catalysts, differences
between IR and NMR measurements of co-monomer contents, and last but not least, sample inhomogeneity due
to possible varying monomer concentrations during the polymerization experiments.
Olefin Polymerizations with Group IV Metal Catalysts 1043
Co-monomer incorporation is highly dependent on the size of the co-monomer and the steric constraints of the
active catalyst center, but also on temperature, activator, and solvent. Almost invariably, addition of c-olefin
co-monomers decreases molecular masses
254,516
and increases catalyst activity. The latter effect, for which a clear
explanation is still lacking, is observed up to a given co-monomer content which depends on the type of catalyst.
E/O co-polymerization reports with bis(cyclopentadienyl) complexes include ethylene/1-butene,
229,233,505,515,520,521
ethylene/1-pentene,
254
ethylene/4-methyl-1-pentene,
516,522,523
ethylene/1-hexene,
229,230,254,446,447,478,481,484,509,512515,
517519,522,524530
ethylene/1-octene,
254,504,505,515,522,531535
ethylene/1-decene,
254,520
ethylene/1-dodecene,
254,536
ethy-
lene/1-tetradecene,
254,532
ethylene/1-hexadecene,
511,512,518,525
ethylene/1-octadecene.
532,536
The most relevant E/O co-polymerization reports with ansa-Cpamido titanium complexes include ethylene/1-
butene,
537,538
ethylene/1-hexene,
539,540
and ethylene/1-octene.
91,442,454,456,457,464,506,541
Ethylene/4-methyl-1-pen-
tene co-polymerizations have been reported as well.
457,542
The thermal and physical properties of the octene
co-polymers produced with ansa-Cpamido titanium complexes have been studied.
543546
The importance of the
activator has been pointed out in many instances.
91,108
Marks has shown that linked homo- and heterobimetallic ansa-Cpamido zirconium complexes produce increased
co-monomer incorporation and higher molecular masses compared to their mononuclear analogs.
103
The MAO-activated, unbridged Cp
*
TiCl
2
(O-2,6-Pr
i
2
C
6
H
3
) has been reported to produce ethylene/1-butene with
activities and molecular masses similar to the ansa-Cpamido catalyst Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
/MAO under similar
conditions (70

C in toluene).
537
A series of (Cp9)(O-2,6-Pr
i
2
C
6
H
3
)TiCl
2
bearing Cp with different substitution
patterns have been used for the co-polymerization of ethylene with 1-hexene, 1-octene, and 1-decene.
547,548
0
20
40
60
80
100
120
140
0 5 10 15 20 25
Co-monomer (mol%)
T
m

(

C
)
Figure 19 Melting point (T
m
) for ethylene/c-olefin co-polymers in the
range 75mol%<ethylene <100mol%.
254,454,484,504,507,512,515519
^: propylene;
M : 1-butene;
&
: 1-hexene; : 1-octene;
&
: 1-decene; v: 1-dodecene.
80
60
40
20
0
Co-monomer (mol%)
T
g

(

C
)
0 10 30 40 50 60 70 80 90 100 20
Figure 20 Glass transition temperature (T
g
) for ethylene/c-olefin co-polymers in the range
0 mol%<ethylene <100mol%.
454,504,507,515
. : propylene; N : 1-butene; ^: 1-hexene; : 1-octene.
1044 Olefin Polymerizations with Group IV Metal Catalysts
A broad range of ansa-(Cp9)(aryloxo)TiCl
2
has been reported for ethylene/1-hexene co-polymerization.
549
Cp
*
(nitroxide)TiMe
2
has been shown to incorporate 1-hexene efficiently, and to produce higher molecular mass
co-polymer compared to Me
2
Si(C
5
Me
4
)(NBu
t
)TiMe
2
, although it is less active.
539
More exotic co-monomers include vinylcyclohexane,
550
4-vinylcyclohexene,
551
divinylbenzene,
552
N-(4-vinylphenyl)carbazole,
553
2-vinylnaphthalene,
554
5-hexen-1-ol,
555
10-undecen-1-ol,
556,557
and dialkylalumi-
num-protected allyl alcohol, R
2
Al(OCH
2
CHTCH
2
).
558
Non-conjugated dienes have also been co-polymerized with ethylene. Examples are 1,4-pentadiene,
559
1,5-hexadiene,
560563
1,7-octadiene,
562564
and 7-methyl-1,6-octadiene.
562
In particular, ethylene co-polymerization with 1,5-hexadiene, 1,7-octadiene, and 7-methyl-1,6-octadiene using
classical metallocenes such as Cp
2
ZrCl
2
and C
2
H
4
(Ind)
2
HfCl
2
has been reported.
562,563,565
1,5-Hexadiene is mainly
incorporated to give a cyclopentane ring, while the two octadienes are incorporated mostly as 1,2-units, thus
generating branches carrying a terminal unsaturation. The vinyl-terminated branches obtained from 1,7-octadiene
can cross-link different polymer chains. Similar results were obtained in the co-polymerization of ethylene with 1,4-
hexadiene, that presents an internal CTC double bond. As expected, catalyst activity and co-monomer incorporation,
as well as molecular masses, were low.
566
Examples of terpolymerizations have also been reported: ethylene/styrene/1,5-hexadiene,
561
ethylene/propylene/
hexene and ethylene/propylene/1-hexadecene,
512
ethylene/butene/1-decene,
520
ethylene/butene/octadecene.
521
In order to introduce LCBs of different structures, vinyl-terminated olefin oligomers have been used in co-
polymerization with ethylene (as well as propylene, see Section 4.09.4.2.8). Such macromonomers can be prepared
by ethylene oligomerization,
531,567
E/P co-oligomerization,
568
or propylene oligomerization.
569
In the last case, a
catalyst with a very high selectivity for u-Me transfer is necessary, in order to produce oligomers with the required
c-olefinic terminal.
Several articles deal with the so-called tandem co-polymerization, where two catalysts are mixed, one oligomer-
izing ethylene to c-olefins, the second co-polymerizing these c-olefins with ethylene.
490,567,570573
The require-
ments for efficient tandem co-polymerization are quite strict: the two catalysts must have compatible activation
chemistry and temperature response; the oligomerization catalyst must have a very high selectivity for c-olefins to
avoid formation and buildup of non-polymerizable oligomers, and no or extremely small co-polymerization capability;
the polymerization catalyst on the other hand must have an extremely high capability of inserting the higher
oligomers.
4.09.4.1.3 Ethylene/propylene co-polymers and ethylene/propylene/diene terpolymers
Ethylene/propylene co-polymers (usually called EPRs for ethylenepropylene rubbers, or EPMs for ethylene
propylene monomers) are amorphous polyolefins when the propylene content is in the range 3070%. Despite the
typical unreactivity of saturated polyolefins, ethylene-rich EP co-polymers can be made highly elastic by radical
cross-linking, but in order to make the rubber vulcanizable, a diene (5-ethylidene-2-norbornene, 1,4-hexadiene, or
dicyclopentadiene) is added, which leaves one unreacted double bond available for subsequent cross-linking. These
latter materials are called EPDMs (for ethylenepropylenediene monomers).
Although a large share of the EP co-polymers and EPD terpolymers is still manufactured with vanadium-based
catalysts at low temperature, metallocene catalysts have added a whole new dimension to EP co-polymerization, and
to the range of material properties that can be achieved, as has been the case for the ethylene-based low-density co-
polymers discussed in the previous section. The subject of EP co-polymerization with group 4 bis(cyclopentadienyl)
complexes has been reviewed in detail up to 1998.
59
The two most important properties of amorphous EP co-polymers are the molecular mass and the content of co-
monomer: both strongly depend on reaction conditions (e.g., temperature and concentration of the monomers) and
catalyst structure. The third parameter, that mostly depends on the catalyst structure, is the type of distribution of the
co-monomers: metallocenes allow the preparation of alternating (r
E
r
P
<1),
59,574576
random (r
E
r
P
~1),
577,578
and blocky (r
E
r
P
1)
59,577,579
co-polymers.
In addition to variables such as molecular mass and co-monomer distribution, metallocenes can produce EP
co-polymers varying in the regiochemistry of propylene insertion
578
and tacticity of propylene sequences, from
isotactic
579
to atactic.
250,578,580
Alternating atactic
59,574,576,581
and isotactic
575,576,581,582
co-polymers have been produced with either C
2v
-sym-
metric or C
1
-symmetric ansa-bis(cyclopentadienyl) complexes. Quite revealing from a mechanistic standpoint is the
observation that the hemiisoselective complexes Me
2
C(CpR)(9-Flu)ZrMe
2
and Me
2
Si(CpR)(9-Flu)ZrCl
2
(R=Me,
Pr
i
) produce isotactic PEPEP sequences and mostly atactic EPPE sequences, and the syndioselective metallocenes
Olefin Polymerizations with Group IV Metal Catalysts 1045
Me
2
C(Cp)(9-Flu)ZrMe
2
and Me
2
Si(Cp)(9-Flu)ZrCl
2
again produce isotactic PEPEP sequences (in addition to the
expected syndiotactic PP and EPPE sequences).
575
This observation is indicative of an alternating insertion
mechanism where propylene inserts only or prevailingly at the most hindered stereoselective site of a C
1
-symmetric
metallocene, while ethylene inserts at the other. The mechanism (main reaction path only) proposed by Soga et al.
582
and Leclerc and Waymouth
575,581
is shown in Scheme 23. Still, the driving force for alternation, especially in the case
of a C
s
-symmetric ligand frame,
575
is not fully understood. Of more practical interest is the discovery that the Si-
bridged complexes produce higher molecular mass co-polymers compared to their C-bridged analogs, although data at
polymerization temperatures higher than 0

C are not available.


575
The statistics of co-polymerization are rather complicated: most of the co-polymerizations do not follow simple
Bernoullian statistics, but are better described by terminal (first-order Markovian) or penultimate (second-order
Markovian) statistics.
59,574
The influence of co-monomer concentration and type of solvent for the catalyst rac-C
2
H
4
(H
4
Ind)
2
ZrCl
2
/tetra-
isooctylalumoxane has been studied.
583
The influence of the transition metal on a given ligand frame has been investigated for both isospecific
584
and
syndiospecific
585
metallocenes, with the finding that hafnium tends to give more blocky sequences compared to
zirconium, while titanium tends to be more alternating compared to zirconium in the syndioselective systems
585
.
The co-polymerization with highly isoselective zirconocenes has also been studied.
253
Co-polymerization results
with state-of-the-art highly isoselective C
2
-symmetric zirconocenes providing high molecular mass co-polymers under
industrially relevant conditions can be found in the patent literature.
586588
The ansa-Cpamido complexes of titanium have met with the largest success in industry thanks to their high-
temperature stability, which allows the use of high-temperature solution polymerization processes.
442
In addition to
producing almost perfect random co-polymers,
578
they give a relatively high content of regioinverted propylene units
and allow the formation of LCBs.
589
Since the early report by Kaminsky and Miri on the terpolymerization of ethylene, propylene, and 5-ethylidene-2-
norbornene with Cp
2
ZrMe
2
in 1985,
590
other terpolymerization studies have appeared.
286
Addition of a diene reduces
Zr
R
P
Zr
R
Zr
R
P
CH
2
CH
2
Zr
R
P
CH
2
CHCH
3
P
R = H, Me, Pr
i
Scheme 23
1046 Olefin Polymerizations with Group IV Metal Catalysts
the activity of the catalyst and the molecular mass of the polymer. The best performing catalysts are again the ansa-
Cpamido complexes of titanium, thanks to their high diene incorporation ability.
591,592
4.09.4.1.4 Ethylene co-polymerization with c,c
/
-disubstituted and internal olefins
Several non-polymerizable olefins have been successfully co-polymerized with ethylene, the most successful
results being achieved with the ansa-Cpamido catalysts. Relevant cases are those of isobutene,
260
2-methyl-1-
pentene,
593
and 2-butene. Typical C
2
- and C
s
-symmetric metallocenes like 29 and 32 have been reported to
selectively co-polymerize ethylene with cis- and trans-2-butene, respectively. Working at low ethylene concentration,
up to 25% and 14%mol of butene could be incorporated into the co-polymers obtained with 29 and 32, respectively.
Independent of the symmetry of the catalyst, the inserted 2-butene units undergo chain-isomerization reactions that
lead to isolated methyl groups in the case of trans-2-butene co-polymerization, and to mainly isolated ethyl groups
and a minor amount of isolated methyl groups in the case of cis-2-butene insertion, as shown in Scheme 24.
594,595
4.09.4.1.5 Ethylene co-polymers with cycloolefins
Ethylene has been co-polymerized witha range of cycloolefins, including cyclobutene,
596
cyclopentene,
596
cyclohexene,
597
norbornene (NB),
596
5-phenyl-2-norbornene,
598
5-vinyl-2-norbornene,
599
5-ethylidene-2-norbornene,
600
dimethano-octa-
hydro-naphthalene,
601
phenyldimethano-octahydro-naphthalene,
598
and norbornadiene.
602
Several non-conjugated cyclo-
diolefins have been co-polymerized as well.
603
Ph
Norbornene 5-Phenyl-2-norbornene 5-Vinyl-2-norbornene 5-Ethylidene-2-norbornene
Dimethano-octahydronaphthalene
Several examples of ethylene/cyclopentene co-polymers prepared with metallocene catalysts have been
reported.
596,603,604
As is the case of many other co-polymers, for ethylene/cyclopentene co-polymers, the glass transition
temperature increases linearly with the cyclopentene content,
605
in the range 2550 mol% of cyclopentane units.
Waymouth reported the synthesis of alternating ethylene/cyclopentene co-polymers with up to 50% cyclopentane
units. The achiral Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
produces atactic poly(cis-1,2-cyclopentane-alt-ethylene) co-polymers,
while the isotactic poly(cis-1,2-cyclopentane-alt-ethylene) co-polymers are obtained with the chiral
Me
2
Si(Ind)(NBu
t
)TiCl
2
pre-catalyst. Remarkably, there seems to be no influence of stereoregularity on melting
points (see Figure 21).
606
Isolated methyl
Isolated ethyl
Zr
Pn
Zr Pn
Zr
Pn
Zr Pn
cis-2-Butene
trans-2-Butene
Isolated methyl
Scheme 24
Olefin Polymerizations with Group IV Metal Catalysts 1047
Within the family of cycloolefin co-polymers, the most important from a material properties standpoint, are the
ethylene/norbornene co-polymers. These co-polymers, dubbed COC for cycloolefin co-polymers, are produced by
Ticona and Mitsui under the tradenames Topas

and Apel

, respectively. An overview of properties and applications


(for example, blisters for pills) can be found on Ticonas Topas homepage.
607
Detailed ethylene/norbornene co-
polymerization studies with different C
1
-symmetric and ansa-Cpamido catalysts, with listing of co-polymerization
parameters, have been published.
608611
NB is inserted exclusively in the cis-2,3-exo-mode (Scheme 25), and most of
the metallocene catalysts tend to produce alternating co-polymers,
609,612
due to the low reactivity of the MNB
intermediate toward further NB insertion. This mode of NB insertion prevents u-H transfer, and thus ethylene/
norbornene co-polymers have increasing molecular masses at increasing NB content.
611
A tendency to alternation means a limitation to the amount of inserted NB co-monomer, at or slightly above
50 mol%. Some of the C
1
-symmetric zirconocenes, and notably those with a higher NB homopolymerization activity,
however, were found to be able to incorporate more NB (up to 70 mol% under the conditions investigated), thus
leading to NB dyads and triads. The least sterically encumbered zirconocenes, MeCH(Cp)
2
ZrCl
2
, showed the highest
NB incorporation.
609
Analogous ansa-bis(cyclopentadienyl)zirconium complexes based on the 2,5-dimethylcyclo-
pentadienyl ligand have been reported to have a good NB incorporation ability with good activities.
613616
Reviews on some catalytic aspects of ethylene/norbornene co-polymerization have appeared.
596,601
The influence
of substituents on ansa-Cpamido complexes
617
and the microstructure of the co-polymers,
618,619
have been the
subject of several studies. Contrary to most ansa-Cpamido complexes of titanium, which were found to produce
alternating ethylene/norbornene co-polymers,
609,617,620
the Me
2
Si(9-Flu)(NBu
t
)TiMe
2
/Ph
3
CB(C
6
F
5
)
4
/AlOct
3
cata-
lyst has been reported to produce random ethylene/norbornene containing up to 82 mol% of NB and a correspond-
ingly high T
g
of 237

C.
621
The correlation between composition and glass transition temperature in ethylene/norbornene co-polymers is
shown in Figure 22.
In the ethylene/5-vinyl-2-norbornene co-polymerization, only the endocyclic double bond undergoes insertion (up
to 14 mol% with the catalyst and under the conditions investigated), leaving the exocyclic vinyl bond accessible for
further reactions, leading to functionalized PEs (Scheme 26).
599
A similar approach uses the co-polymerization of ethylene with 5-ethylidene-2-norbornene, followed by
hydroboration/oxidation of the unreacted vinyl group. The hydroxylic functions in the co-polymer are then converted
into OAlEt
2
groups and used as catalysts for -caprolactone polymerization, thus leading to poly(ethylene-co-ENB)-
graft-polycaprolactone co-polymers.
600
0
40
80
120
160
200
Cyclopentene (mol%)
T
m

(

C
)
0 10 20 40 50 30
Figure 21 Melting points of alternating ethylene/cyclopentene co-polymers, showing apparent independence of melting point
from tacticity.
&
: atactic poly(cis-cyclopentane-alt-ethylene) co-polymers, Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
/MAO; isotactic poly(cis-
cyclopentane-alt-ethylene) co-polymers, Me
2
Si(Ind)(NBu
t
)TiCl
2
/MAO.
606
The data have been fitted with a third-grade poly-
nomial, with the only purpose of providing a guide to the eye.
P
L
n
M
+
P
L
n
M
+
Scheme 25
1048 Olefin Polymerizations with Group IV Metal Catalysts
4.09.4.1.6 Ethylene/styrene co-polymers
The successful co-polymerization of ethylene with styrene is another polymeric material conquest of homogeneous
catalysis. At low styrene incorporation, the co-polymer is substantially a functionalized crystalline polyethylene.
Increasing the styrene content results in a decrease of crystallinity while elastomeric properties arise,
623
up to about
80 mol% styrene; for higher styrene contents, the co-polymer becomes a glassy amorphous material. A comprehensive
review on the properties of ethylene/styrene co-polymers is available.
624
The main drawback of the elastomeric co-polymers is their rather high T
g
, which is always higher than
LLDPE co-polymers and increases with the styrene content, reaching values above 0

C beyond 60 wt% of
styrene.
624
In the styrene range of 2050 mol%, the co-polymer can be either amorphous elastomeric or
crystalline. This kind of crystallinity occurs when the ES dyad is stereoregular.
625
Ethylene/styrene co-poly-
mers can act as compatibilizer for polyethylenepolystyrene blends. A review on ethylene/styrene co-polymer-
ization has appeared.
626
Even at high styrene incorporation, the co-polymers are formed by ethylene blocks and isolated styrene
units.
627
Half-sandwich titanium complexes such as 3539 have also been reported to be active in the ethyl-
ene/styrene co-polymerization. The performance of the MAO-activated complex 35 is highly dependent on the
Al/Ti ratio. At a ratio of 100, a co-polymer composed of polyethylene blocks with essentially isolated styrene units
could be fractionated from the homopolymers. By contrast, at Al/Ti ratios of 1000, a co-polymerization at the
same feed ratio resulted in the production of only homopolymers, or co-polymers composed of long PE and sPS
blocks at most.
628
Subsequent
13
C NMR analysis of the co-polymers obtained at 20

C indicated that up to
36 mol% of styrene was incorporated.
629
However, under very similar conditions, only formation of the homo-
polymers was reported.
630,631
This may be reasonable since catalytic systems 35/MAO and 36/MAO give rise to
several active species with different catalytic properties. Thus, remarkably different results can be obtained with
small differences in the experimental procedure.
0
50
100
150
200
250
300
350
400
0 10 20 30 40 50 60 70 80 90 100
NB (mol%)
T
g

(

C
)
Figure 22 Correlation between norbornene content and glass transition temperature in ethylene/norbornene
co-polymers ( ).
621,622
P
Cp
2
Zr
+
toluene
OH
O
m-Cl-C
6
H
4
COOOH
toluene
i, 9-BBN
ii, NaOH/H
2
O
2
Scheme 26
Olefin Polymerizations with Group IV Metal Catalysts 1049
Ti
Cl
Cl
Cl
35
Ti
Cl
Cl
Cl
36
Ti
OBn
BnO
BnO
37
Ti
Cl
Cl
Cl
38
Ti
O
Cl
Cl
Pr
i
Pr
i
39
The B(C
6
F
5
)
3
-activated complex 36, which is highly active in the syndiospecific styrene polymerization, yields
ethylene-styrene (E-S) copolymers at polymerization temperatures 25

C. Increasing amounts of ES units were


obtained at increasing styrene concentrations in the feed. However, besides a larger amount of ES units, the
production of sPS becomes favored. Nevertheless, the THF-soluble fraction of the materials obtained comprised
ES co-polymers with a highly alternating and, interestingly, atactic microstructure.
626,632
The B(C
6
F
5
)
3
-activated complexes 35 and 36 yield PES with 4-aryl-1-butyl branches as shown in Scheme 27. The
4-phenyl-1-butyl branches were shown to originate from the formation of ethylene/styrene co-oligomers such as
6-phenyl-1-hexene which is subsequently incorporated in the PE.
633,634
Other authors have investigated titanium half-sandwich complexes by varying polymerization conditions,
635,636
or
by varying the groups bound to the metal, or the Cp ring substitution, as in complexes, 3739,
637640
or by using
theoretical approaches.
641
The above systems always yield a mixture of co-polymer and homopolymers, and the co-polymer has to be
extracted with solvents such as THF or MEK. Thus, the discovery that ansa-Cpamido Ti complexes are able to
co-polymerize ethylene and styrene was a clear step forward. In fact, these catalysts do not homopolymerize
styrene.
642
Structural characterization of the co-polymers obtained with the Me
2
Si(Cp)(NBu
t
)TiCl
2
/MAO catalyst
showed that only 35 mol% of styrene was incorporated in the co-polymer, even at a styrene feed of 91 mol%. NMR
analysis indicated the absence of SS sequences.
643
A series of MAO-activated ansa-Cpamido Ti complexes has
been investigated, and all catalysts produced random co-polymers without any regioregular or stereoregular micro-
structure. The highest activity corresponds to Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
, while the higher styrene incorporation was
obtained with complexes Me
2
Si(3-Me
3
Si-Indenyl)(NBu
t
)TiCl
2
and Me
2
Si(C
5
Me
4
)(N-Bn)TiCl
2
.
644
Using a fluor-
enyl- and Zr-based catalyst, Me
2
Si(fluorenyl)(NBu
t
)ZrCl
2
/MAO, only 1 mol% of styrene was incorporated.
644
A very
similar catalyst, Me
2
Si(9-Flu)(NBu
t
)TiMe
2
/[Ph
3
C][B(C
6
F
5
)
4
], yields an almost perfectly alternating and isotactic ES
co-polymer with a T
m
of 118

C.
645
Other studies on ethylene/styrene co-polymerization with ansa-Cpamido
complexes have been reported.
646652
Ethylene/styrene co-polymerizations using the MAO-activated C
2
-symmetric rac-C
2
H
4
(1-Ind)
2
ZrCl
2
complex 29
or the C
s
-symmetric Me
2
C(Cp)(9-Flu)ZrCl
2
complex 32 lead to random co-polymers with a styrene content of up to
45 mol%.
653
NMR characterizations of these co-polymers indicated that insertion of styrene is not completely
regioregular.
654
Almost perfectly stereoregular and alternating ethylene/styrene co-polymers using the same
C
2
- and C
S
-symmetric metallocenes 29 and 32 were achieved by lowering the co-polymerization temperature to
25

C. The regular microstructure allowed crystallization of the co-polymers.


625
Interestingly, both the co-polymers
obtained with 29 and 32 resulted in an isotactic arrangement of the styrene units.
655
Co-polymers composed of
isotactic PS and PE blocks were also synthesized utilizing the bulky H
2
C(3-Bu
t
-indenyl)ZrCl
2
complex 30. The
blocky nature of the co-polymers obtained with 30 is due to the high tendency of this complex to induce primary
insertion of 1-olefins, including styrene. This is an interesting result because secondary insertion of styrene is usually
favored.
656
The synthesis at 50

C of random ethylenestyrene co-polymers using complex 40 has been reported. The


resulting co-polymers showed a high content of isotactic ES units and of occasional regioregularly arranged
Scheme 27
1050 Olefin Polymerizations with Group IV Metal Catalysts
head-to-tail styrenestyrene units.
657
The consequences of the C
1
-symmetry of metallocenes such as 41 in the
ethylene/styrene co-polymerization was investigated. A Ph group on the Cp ring has a beneficial effect on the activity
of the catalyst, while styrene insertion is similar to that obtained with the unsubstituted analogs. Conversely, an alkyl
substituent in the same position depresses styrene insertion and overall activity of the catalyst.
658
The mechanism of
ethylene and styrene co-polymerization has also been investigated with combined experimental/theoretical
approaches.
648,659,660
Isotactic polystyrene (iPS) polymers with ethylene units from roughly 0% to 50% can be
produced using the 30/MAO catalyst. Interestingly, at low ethylene contents, the isolated ethylene units inhibit
crystallization of the iPS segments.
661
40
ZrCl
2 Me
2
C Me
2
C ZrCl
2
41
Other complexes that have been used in ethylenestyrene co-polymerizations are 42, which is inactive in styrene
homopolymerization but has been claimed to produce ethylene/styrene co-polymers with styrene content in the
range 3587 mol%.
662
Living ethylene/styrene co-polymerization can be achieved using the MAO-activated complex
43. Although a relatively low amount of styrene was incorporated (about 10 mol%), NMR analysis indicated that trace
amounts of pseudo-random tail-to-tail SS or SES sequences were observed.
663
Marks and co-workers showed
that bimetallic catalysts based on 44 can effectively yield ethylene/styrene co-polymers and, in contrast to the
monometallic CGCs, styrene incorporation can be higher than 50%.
664
Zr
P
Ph
Ph
42
Ti
N
Cl
Cl
Bu
t
Bu
t
43
Ti
N
Me
Me
SiMe
2
Bu
t
Ti
N
Me
Me
Me
2
Si
Bu
t
44
Finally, it is worth noting that chloro- or methyl-substitued styrenes can be co-polymerized with ethylene using
either Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
or rac-C
2
H
4
(1-Ind)
2
ZrCl
2
. Results indicated that p-methyl-styrene is incorporated
far more effectively than styrene.
653,665,666
The Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
complex allows the terpolymerization of
ethylene, propylene, and p-methyl-styrene as well as of ethylene 1-octene and p-methyl-styrene.
667
Besides the
ethylene and styrene terpolymerization with propylene and 1-octene, the terpolymerization of ethylene and styrene
with norbornene or 1,5-hexadiene has been reported.
561,668
4.09.4.2 Propylene Polymers
Propylene is the simplest prochiral olefin, and the different chain microstructures that can be generated by its four
possible insertion modes (all made possible by single-center catalysts, albeit to largely different extents) have been
described in detail in Section 4.09.2.3. Two rather different strategies can be followed in order to tailor the physical
properties of PP: (i) co-polymerization with ethylene or other c-olefins and (ii) varying the enantioface selectivity of
the active sites of the single-center catalyst by ligand variation. The co-polymerization approach is described in
Sections 4.09.4.2.5 (ethylene as co-monomer) and 4.09.4.2.6 and 4.09.4.2.7 (c-olefin co-monomers). In Sections
4.09.4.2.14.09.4.2.4, we describe the different PP homopolymers accessible from group 4 metallocene catalysts.
The degree of PP chain stereoregularity can be modified from totally absent to almost perfect (in either direction,
isotactic or syndiotactic) by altering the monomer enantioface selectivity of each active site of the catalyst: in the case
of metallocenes, this means varying the substitution pattern of the two cyclopentadienyl ligands. Metallocenes are by
far the most versatile systems for the production of PPs of different chain stereoregularity and molecular masses.
Controlling the type and number of monomer insertion mistakes leads to controlling the crystallinity and the
melting point of PP, which, together with its molecular mass, define its physical and mechanical properties.
Olefin Polymerizations with Group IV Metal Catalysts 1051
Quoting from a previous review: in the case of highly crystalline iPP, metallocenes are unlikely to replace the
newest heterogeneous Ti/MgCl
2
catalysts in any foreseeable future [for the production of commodity polypropylene].
So, why use metallocenes to produce polypropylene? In one sentence, because polypropylene properties can be
tailored! For example, iPP can be made from fully amorphous to highly crystalline and anything in between.
162
This
is shown pictorially in Figure 23. In addition, the very high co-monomer incorporation ability (especially for higher
c-olefins) and very good inter- and intramacromolecular co-monomer distribution achievable with metallocenes
represents an additional, very powerful tool for tailoring PP properties.
One additional and important requirement for a catalyst in order to make it useful in practice is its molecular mass
capability, which must be as high as possible and coupled with a good hydrogen response. This simultaneous control
over both the stereoselectivity and the molecular mass capability of a PP catalyst in the range of industrial
polymerization temperatures, which in addition must have high activity (to give a rough indication, 100 kg PP
(g metallocene)
1
) and a good co-monomer incorporation ability, and last but not least, an efficient and cost-effective
metallocene synthetic procedure, are the considerable challenges to be faced when developing a new catalyst.
In the following sections, we review the recent developments in metallocene catalyst structures, that made it
possible to produce PPs ranging from flexible and elastomeric with varying degrees of crystallinity to highly crystal-
line, highly stereoregular PP on the one hand, and a very broad family of propylene-based co-polymers on the other,
all these polymers having molecular masses in the range of industrial applications.
4.09.4.2.1 Amorphous polypropylene
As shown in Figure 23, PP is no longer able to crystallize when the stereoregularity of the chains is reduced below a
threshold value (belowabout 70% mdiad, or 40% mmmmpentad content for iPP, or belowabout 60% rrrr pentad content
for sPP), and it becomes amorphous (amPP). When statistical randomness in the sequence of chirotopic methynes in the
polymer chain is reached, the polymer is called atactic (aPP). In this case the pentad distribution is perfectly random
Bernoullian: mmmm: mmmr : rmmr : mmrr : (rmrr mrmm) : mrmr : rrrr : rrrm: mrrm~1 : 2 : 1 : 2 : 4 : 2 : 1 : 2 : 1.
0 10 20 30 40 50 60 70 80 90 100
% m
155
165 C
T
m
(C)
iPP
sPP
aPP
m = r
60
Amorphous PP
C
2v
C
s
C
2
C
1
Ph
2
C ZrCl
2
Me2Si ZrCl
2
HfMe
2
Me2Si
TiMe2
N
N
Me2Si
ZrCl
2
Figure 23 Schematic representation of melting point dependence on PP microstructure (% m dyads), catalyst symmetries
with ranges of accessible stereoregularities, and selected examples of representative metallocene structures.
1052 Olefin Polymerizations with Group IV Metal Catalysts
While aPP is obviously amorphous, an aPP is not necessarily atactic. The methyl pentad region of a fully
regioregular, nearly perfect aPP is shown in Figure 24 with pentad assignments.
The most important consequences of the absence of crystallinity are softness, tackiness (the property of a material
to adhere to itself), a complete solubility in most low-polarity organic solvents, including ethers and aliphatic
hydrocarbons, higher transparency, and lower density with respect to crystalline PP. Other physical properties depend
also on the molecular mass of aPP.
670
Despite the insolubility of aPP in liquid propylene, its tackiness makes it
impossible to produce it in bulk or gas-phase processes, with a solution process at medium temperature likely being
the only viable manufacturing process.
Amorphous PP was a byproduct of iPP production in the early slurry processes, and was isolated from the aliphatic
solvents in the solvent recycle section.
This material, often incorrectly referred to as atactic, is actually neither atactic nor fully amorphous, but is a
mixture of chains of different stereoregularity and molecular masses, as proved by polymer fractionation.
315,374,671,672
The methyl pentad region of the
13
C NMR spectrum of such a ZN PP clearly shows both isotactic and syndiotactic
stereoblocks. Although the molecular masses of this material are low, amPPs have been used as bitumen additives
and in hot-melt adhesives. With the improvement of the stereospecificity of the ZN catalysts for iPP, amPP was no
longer available as a byproduct. Fully or largely amorphous PP from ZN has since been produced on purpose, also by
adding co-monomers to further reduce crystallinity.
673680
Atactic or nearly atactic PP can be made with four types of metallocene catalysts: (i) the achiral, unbridged
metallocenes lacking stereorigidity (e.g., Cp
2
ZrCl
2
, (MeCp)
2
ZrCl
2
, Ind
2
ZrCl
2
, (2-MeInd)
2
ZrCl
2
, and the
like)
180,250,681686
and the bridged, stereorigid C
2v
-symmetric metallocenes such as Me
2
SiFlu
2
ZrCl
2
;
250,266,687,688
(ii) the meso-isomers of ansa-metallocenes (e.g., meso-C
2
H
4
(1-Ind)
2
ZrCl
2
, meso-C
2
H
4
(H
4
Ind)
2
ZrCl
2
, meso-Me
2
Si(2-
Me-4-Ar-Ind)
2
ZrCl
2
);
689,690
(iii) the ansa-C
2
-symmetric metallocenes having the bridge between the 2,29-position of
the two indenyl moieties, such as rac-C
2
H
4
(1-R-2-indenyl)MX
2
477
and rac-H
2
C(1-R-2-indenyl)MX
2
(R=Me,
CH
2
Ph, Bu
t
, Me
3
Si; M=Zr, Hf; X=Cl, Me)
669
; (iv) some monocyclopentadienyl complexes, such as CpTiX
3
(X=Me, Cl, OR)
691,692
and Cp
*
TiX
3
(X=Me, Cl, OAr, NAr, j
2
-ONR
2
).
497,547,693695
A selection of structures together with the most relevant polymerization results is reported in Table 2 (mono-
cyclopentadienyl titanium complexes) and Table 3 (zirconocenes).
CpTiR
3
complexes with pendant phenyl moieties attached to the Cp show, with respect to Cp
*
TiMe
3
, reduced
propylene polymerization activities and molecular masses upon activation with B(C
6
F
5
)
3
in toluene at T
p
20 to
60

C, in accordance with the formation of 16-electron resting states upon j


6
-arene coordination.
696
On the other
hand, (Me
2
NCH
2
CH
2
Cp)TiCl
3
shows much enhanced activity with respect to the practically inactive CpTiCl
3
/MAO
system.
286,697,698
Quite interestingly, polymerization of propylene with (Me
2
NCH
2
CH
2
Cp)TiCl
3
/MAO at 25

C
produces fairly high molecular mass aPP and shows a rate of polymerization with a monomer concentration
dependence of 1.8, a behavior similar to that of some stereoselective ansa-metallocenes.
289
19.5 20.0 20.5 21.0 21.5 22.0 ppm
m
m
m
m
m
m
m
r

r
m
m
r

m
m
r
r

m
m
r
m

+

r
m
r
r
m
r
m
r

r
r
r
r

r
r
r
m

m
r
r
m

Figure 24 Methyl pentad region of the
13
C NMR spectrum (100 MHz, 120

C, C
2
D
2
Cl
4
) of aPP with pentad assignments.
Catalyst: rac-CH
2
(1-Me
3
C-2-Ind)
2
HfMe
2
/MAO.
669
Olefin Polymerizations with Group IV Metal Catalysts 1053
Table 2 Propylene polymerization with monocyclopentadienyl Ti complexes
Pre-catalyst Activator Conditions
Activity (kg
PP
mmol
Ti
1
h
1
) M

w
Tacticity References
Ti
Me
Me
Me
B(C
6
F
5
)
3
Toluene,
propylene 1.9
mol l
1
, 80 to
60

C, 30 min
1.5 7.4 10
5
8.5%
mm~2%
2,1
691
Ti
PhCH
2
O
OCH
2
Ph
PhCH
2
O
MAO
Toluene,
1.30 10
5
Pa
propylene,
60

C, 1 h
0.9 2.3 10
4
10% mm 692
B(C
6
F
5
)
3
Liquid propylene,
45

C, 240 s
5.2 4 10
6
amorphous 693
Ti
Me
Me
Me
B(C
6
F
5
)
3
Toluene, 1 atm
propylene,
25

C, 5 min
0.5 2.6 10
5
48.5%
m12%
2,1
694
Ph
3
C

B(C
6
F
5
)
4

90% Propylene in
toluene, 20

C,
20 min
0.7 1.1 10
6
Amorphous 695
Ti
CH
2
CHCH
2
O
OCH
2
CH CH
2
CH
2
CHCH
2
O
MAO
Toluene, 1.3 atm
propylene,
40

C, 1 h
0.1 27.3 10
4
11% mm 702
Ti
X
X
X
Me
2
N
MAO
Toluene, 20 psi
propylene,
20

C, 30 min
1.5 4.5 10
5
Amorphous 698
N
N
N
Ti
CH
2
Ph
CH
2
Ph
B(C
6
F
5
)
3
Toluene, 3 bar
propylene,
50

C, 30 min
14.4 8.1 10
4
23% mm 496
Ti
Cl
Cl
N
SiMe
3
MAO
Toluene, 7 atm
propylene, 0

C,
30 min
0.051 91.3 10
4
Atactic 497
(Continued)
1054 Olefin Polymerizations with Group IV Metal Catalysts
Cp
*
TiMe
3
/B(C
6
F
5
)
3
shows a quasi-living behavior at T
p
_20

C, producing aPP with molecular mass above 10


6
and very narrow molecular mass distribution.
693
The same system, investigated at higher temperature, was shown to
produce aPP with up to 15% regioerrors;
694
these regioerrors do not affect the rate of further monomer insertion as
evidenced by the lack of hydrogen activation, which however efficiently lowers PP molecular mass.
699
A similar living
behavior at low temperatures has been reported for Cp9
2
MMe
2
/B(C
6
F
5
)
3
/AlR
3
(Cp9 =C
5
H
5
, C
5
Me
5
; M=Zr, Hf;
R=
i
Bu,
n
Oct).
700,701
Interestingly, the CpTi(OR)
3
, CpTiX
3
, and Cp
*
Ti(OR)
3
complexes produce regio-block
aPP.
692,702,703
The very high molecular mass of the aPP obtained with (2-MeInd)
2
ZrCl
2
/MAO, well above 3 10
6
at T
p
_0

C,
250
is remarkable. Amorphous PP, deviating from perfect atacticity in either direction, but still lacking any long-range
stereochemical order, is obtained by several of the above four systems when different co-catalysts or low polymeriza-
tion temperatures are applied,
682,691,704
and by three other classes of catalysts: (v) the ansa-Cpamido Ti metallo-
cenes, such as Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
and its analogs, which produce very high molecular mass amPPs with a bias
toward syndiotacticity;
30,705708
(vi) some C
2
-symmetric metallocenes, such as rac-Me
2
C(3-Pr
i
Ind)
2
ZrCl
2
;
222,709
and
(vii) several C
1
-symmetric metallocenes.
576
These two latter cases will be discussed in Section 4.09.4.2.3.
In addition to these, it is worth mentioning that also thio-bis(aryloxy)titanium compounds give aPP of high
molecular masses, but with low productivities.
159
Regioselectivities vary widely among different catalysts, with the
bis(cyclopentadienyl) complexes being the most regioselective and the thio-bis(aryloxy) compounds producing
highly regioirregular aPP.
Appending asymmetric aryl groups to Cp in CpMCl
3
(M=Ti, Zr) has been shown to induce the formation of
PP-containing aPP/iPP stereoblocks, the length of which strongly depends on the polymerization temperature.
710,711
The silyl-bridged Cpamido titanium catalysts, first developed for ethylene homo- and co-polymerization (see
Section 4.09.4.1), have also been extensively studied in propylene polymerization. In general, these complexes
produce remarkably high molecular mass amPP even at elevated temperatures. Microstructures are slightly biased
toward either syndiotactic or isotactic, depending on the Cp substitution pattern. Significant deviations from
atacticity have been reported for some complexes, and will be discussed in the next section. Variation of the
substituent on the amido nitrogen also has an influence on tacticity and molecular mass.
705,706,712
The influ-
ence of bridge type and length has been studied by Hessen
461
and Marks.
460
Selected structures and propylene
polymerization results are shown in Table 4.
The C
2v
-symmetric R
2
Si(9-Flu)
2
ZrCl
2
and the C
1
or C
s
-symmetric ansa-Cpamido titanium complexes so far seem
to be the most efficient catalysts for the production of high molecular mass amPP, due to their ability to maintain high
activity and produce high molecular masses at relatively high polymerization temperatures (T
p
50

C), and high


molecular masses also at relatively low propylene concentration in hydrocarbon solution polymerization. The activity
recently reported for Cp
*
TiMe
2
(ONEt
2
)/[Ph
3
C][B(C
6
F
5
)
4
]
695
is clearly outstanding.
Table 2 (Continued)
Pre-catalyst Activator Conditions
Activity (kg
PP
mmol
Ti
1
h
1
) M

w
Tacticity References
Ti
Cl
Cl
N MAO
Toluene, 7 atm
propylene, 0

C,
30 min
2.6 n.r. Atactic 498
Ti
O
Me
Me
N
Ph
3
C

B(C
6
F
5
)
4

90% propylene in
toluene, 20

C,
20 min
153 2.5 10
6
Atactic,
~3% 2,1
695
Olefin Polymerizations with Group IV Metal Catalysts 1055
4.09.4.2.2 Isotactic polypropylene
Isotactic PP is the second major thermoplastic in terms of volumes, and the vast majority of it is produced by means of
heterogeneous Ti catalysts. As described in Section 4.09.3, the heterogeneous, donor-modified MgCl
2
/TiCl
4
/AlR
3
catalysts produce very high molecular mass, highly isotactic PP with generally broad molecular mass distributions and
containing small (210%) amounts of stereoirregular, amorphous PP, both telltales of the multi-site nature of such
catalysts. Careful fractionation studies have demonstrated that, even in the more stereoregular, insoluble fraction of
ZN iPP, the stereodefects are not randomly distributed in the chains but are rather cumulated in stereoblocks.
713
Therefore, ZN iPP has a relatively high crystallinity, a T
m
of 163167

C, and relatively high elastic modulus (1500


2000 MPa).
Table 3 Propylene polymerization with bis(cyclopentadienyl) Zr complexes
Pre-catalyst Activator Conditions
Activity
(kg
PP
mmol
Zr
1
h
1
) M

w
Tacticity References
ZrCl
2 MAO
Liquid propylene,
50

C, 1 h
5.7 oligomers ~25% mm 180
Me Me
3
Si
Me
SiMe
3
ZrCl
2 MAO
Liquid propylene,
50

C, 1 h
63 3.9 10
4
17.6% mm 669
ZrCl
2
MAO
Liquid propylene,
0

C, 1 h
1.8 2.6 10
5
56% m 250
MAO
Liquid propylene,
50

C, 1 h
18 ~1 10
4
30% mm 180
ZrCl
2
MAO
Liquid propylene,
0

C, 1 h
8 3.2 10
6
36% mm 250
MAO
Liquid propylene,
50

C, 1 h
10
2.2 10
4
35% mm 250
ZrCl
2
MAO
Liquid propylene,
50

C, 1 h
17.7 1.4 10
5
20% mm 250
ZrCl
2
Bu
2
Si
MAO
Liquid propylene,
50

C, 1 h
32.7 4.6 10
5
15% mm 250
1056 Olefin Polymerizations with Group IV Metal Catalysts
Table 4 Propylene polymerization with ansa-Cpamido and related complexes
Pre-catalyst Activator Conditions
Activity
(kg
PP
mmol
Ti
1
h
1
) M

w
Tacticity References
TiCl
2
Me
2
Si
N
MAO
Liquid propylene,
60

C, 1 h
10.9 650 10
3
10.5% mm 708
TiMe
2
Me
2
Si
N
MAO
Liquid propylene,
60

C, 1 h
4.1 135 10
3
24.8% mm 708
TiMe
2
Me
2
Si
N
MAO
Liquid propylene,
60

C, 1 h
6.2 550 10
3
15.6% mm 708
TiMe
2
Me
2
Si
N
MAO
Liquid propylene,
60

C, 1 h
30.8 194 10
3
13.8% mm 708
Ti(CH
2
Ph)
2
N
CH
3
B(C
6
F
5
)
3
Toluene, 4.7 10
5
Pa
propylene, 30

C,
30 min
0.72 1,050 10
3
14% mm 461
TiCl
2
N
CH
3
MAO
Toluene, 4.7 10
5
Pa
propylene, 30

C,
30 min
0.29 720 10
3
17% mm 461
(Continued)
Olefin Polymerizations with Group IV Metal Catalysts 1057
In the case of metallocene catalysts, due to their single-center nature, the stereochemical or regiochemical errors
are randomly distributed in all PP chains. This randomness leads to PP of lower melting points (_160

C) and
stiffness, but at the same time better transparency, compared with ZN iPP.
The earlier, simple chiral ansa-bisindenyl zirconocenes and hafnocenes were found to be less stereoselective than
ZN catalysts, producing PP of low melting point and very low molecular mass. The long development from the early
Brintzinger and Ewen ethylene-bridged chiral ansa-bis(indenyl) structures
202,203,714
to metallocenes showing perfor-
mances close to those of ZN catalysts has been described in detail up to the end of 1999.
162
Since then, three major directions of industrial research have become evident: the introduction of pendant
715,716
or
condensed heterocycles onto one or both Cp ligands,
63,717720
the development of pseudo-racemic C
1
-symmetric
structures aimed at removing the molecular mass drop induced by ethylene,
586,721,722
and the evolution of improved
supportation techniques. This latter aspect has been extensively reviewed
722724
and will not be further discussed
here. Several new C
2
-symmetric structures have been described in recent years, mostly in the patent literature, and
for the latter case polymerization results refer to the silica-supported catalysts.
A selection of these C
2
-symmetric structures, together with polymerization results, is shown in Table 5. Although
structural complexity has grown considerably, no major improvement in propylene homopolymerization has been
obtained so far.
Stereoerrors in PP chains produced with catalyst systems based, for example, on Me
2
Si(2-Me-4-ArInd)
2
ZrCl
2
are
close to the detection limit by NMR. The low melting point (T
m
150160

C) of PP is due to the presence of


regioirregularities (of the 2,1-erythro-type). Two
13
C NMR spectra of the methyl region of iPP having only regio-
defects (2,1-erythro) and only stereodefects are shown in Figure 25.
It is worth noting here that in the early systems (such as rac-C
2
H
4
(Ind)
2
ZrCl
2
, rac-C
2
H
4
(H
4
Ind)
2
ZrCl
2
, rac-
C
2
H
4
(4,7-Me
2
Ind)
2
ZrCl
2
, and their Me
2
Si-bridged analogs), 2,1-insertions have been indicated as the cause of
lower molecular masses and activities, due to the formation of a less reactive catalyst state, and to fast u-H transfer
after a 2,1-unit (see Section 4.09.2.4). On the other hand, the zirconocenes of the class Me
2
Si(2-Me-4-ArInd)
2
ZrCl
2
,
despite an even lower regioselectivity, show the highest molecular masses and activities. In fact, the 2-Me-4-Ph
substitution pattern increases stereoselectivity partially at the expense of regioselectivity,
164
but without increasing
chain transfer rates: here, a 2,1-unit does not seem to adversely affect the rate of monomer insertion.
From the examples in Tables 57, it is clearly apparent that, of the different evolutionary directions of
isospecific metallocene structures, the one based on the 2-methyl-4-aryl-indenyl type,
734
has by far been the most
successful.
The influence of polymerization conditions on the performance of isoselective, C
2
-symmetric metallocenes has
been described in detail already.
162
Two further studies on catalysts rac-Me
2
Si(2-Me-4-PhInd)
2
ZrCl
2
and rac-
Me
2
Si(2-Me-4-NaphthInd)
2
ZrCl
2
have appeared recently.
305,735
Busico has investigated the kinetics of propylene
polymerization with rac-Me
2
Si(2-Me-4-PhInd)
2
ZrCl
2
/MAO.
297
Table 4 (Continued)
Pre-catalyst Activator Conditions
Activity
(kg
PP
mmol
Ti
1
h
1
) M

w
Tacticity References
Ti Cl
2
N
CH
3
MAO
Toluene, 2 bar
propylene, 50

C,
30 min
2.4 110 10
3
22%
mm<0.5%
2,1
462
Ti(CH
2
Ph)
2
O
Ph
3
C

B(C
6
F
5
)
4

Toluene, 1 atm
propylene, 25

C,
5 min
3.8 23.6 10
3
22% mm 460
1058 Olefin Polymerizations with Group IV Metal Catalysts
Table 5 Selected C
2
-symmetric isospecific zirconocenes with a 2-alkyl-4-arylindenyl ligand framework
Zirconocene Activator
T
p
(

C)
M

w
(10
3
)
T
m
(

C)
NMR
a
Notes References
Me
2
Si ZrCl
2
45
MAO 70 228 144
96.3% mm, 93.8%
mmmm 0.4%
2,1
Unsupported, liquid propylene 725
MAO 40 257 145.4
92.5 % mmmm 1%
2,1
Silica-supported, hexane,
[propylene]=1.83 mol
1
726
Me
2
Si ZrCl
2
46
MAO 70 1,184 156
99.1% mmmm
0.5% 2,1
Unsupported, liquid propylene 720
MAO 60 600 149.2
Silica-supported, liquid
propylene
727
Me
2
Si ZrCl
2
47
MAO 50 380 156.2
98.6 % mmmm
0.3% 2,1
Unsupported, toluene, 1 atm
propylene
728
Me
2
Si ZrCl
2
48
MAO 50 400 159.8
99.2% mmmm
ca. 0.8% 2,1
Unsupported, toluene, 1 atm
propylene
728
Me
2
Si ZrCl
2
49
MAO 70 900 151
Silica-supported, liquid
propylene
729,730
(Continued)
Olefin Polymerizations with Group IV Metal Catalysts 1059
We recall here the two major effects:
(i) Propylene concentration has a non-linear influence on both iPP molecular mass and isotacticity. The cause of this
behavior is the onset of competitive growing-chain-end epimerization. The mechanism of this reaction has been
discussed in Section 4.09.2.4. Selected examples are shown in Figure 26 (isotacticity) and Figure 27 (molecular
mass).
The rate of epimerization at a given monomer concentration [m] depends on the polymerization temperature and
on the nature of the ansa-:-ligand. The dependence of isotacticity on [m] can be described by Equation
(10).
162,241
The equilibrium constant K
eq
=[M?m]/([M][m]) is according to the model described in Ref: 162.
b
obs
1b
obs
=
0.5 bK
eq
[m[
0.5 (1b)K
eq
[m[
(10)
The observed (apparent) enantioselectivity parameter b
obs
is usually lower than the true probability of a correct
enantioface insertion (averaged over the two sites of the catalyst) in the absence of epimerization, defined by the
Table 5 (Continued)
Zirconocene Activator
T
p
(

C)
M

w
(10
3
)
T
m
(

C)
NMR
a
Notes References
Me
2
Si ZrCl
2
50
MAO 70 550 152.7
Silica-supported, liquid
propylene
727
Si
ZrCl
2
51
MAO 60 711 156.2
Silica-supported, liquid
propylene
587
Si
ZrCl
2
52
MAO 60 381 160.8
Silica-supported, liquid
propylene (low activity)
587
a
When allowed by data description, %mmmm refers to primary insertions only.
1060 Olefin Polymerizations with Group IV Metal Catalysts
Bernoullian probability parameter b: the value of b depends on the active site structure and T
p
, but is independent
of [m]. In liquid propylene, for most metallocenes b
obs
b, at least at the polymerization temperature of 50

C.
Concerning the dependence of molecular mass on monomer concentration, three examples are shown in
Figure 27. In these examples, the main chain-release mechanism is different for each catalyst: Brintzingers
catalyst, rac-C
2
H
4
(Ind)
2
ZrCl
2
/MAO, allows u-H transfer both after a 2,1-insertion; and after a 1,2-insertion;
the fully regioselective catalyst, rac-Me
2
C(3-Bu
t
Ind)ZrCl
2
/MAO, shows mainly u-Me transfer;
246
while rac-
Me
2
Si(2-Me-4-PhInd)ZrCl
2
, despite being the least regioselective of the three, gives mainly u-H transfer after
a 1,2-insertion (even in liquid monomer), with u-Me transfer becoming predominant at the lowest monomer
concentrations.
268
Two facts are worth pointing out: (a) testing metallocene catalysts under reduced pressure might lead to gross
underestimation of their molecular mass capability and isoselectivity, and to missing possibly relevant differences
between different catalysts; and (b) regioselectivity also affects the crystallinity of iPP.
736738
For example, rac-
Me
2
Si(2-Me-4-PhInd)ZrCl
2
/MAO is more isospecific than rac-Me
2
C(3-Bu
t
Ind)ZrCl
2
/MAO at any propylene
concentration, but less regioselective, the net result being that iPP from rac-Me
2
Si(2-Me-4-PhInd)ZrCl
2
and
Me
2
C(3-Bu
t
Ind)ZrCl
2
have very similar melting points.
(ii) Increasing the polymerization temperature causes a drop, sometimes dramatic, in isoselectivity and molecular
mass.
267
This effect has sometimes been overestimated due to concomitant decrease in monomer concentration,
when the experiments were performed at constant pressure, rather than at constant monomer concentration.
739
Data detailing the influence of polymerization temperature for polymerizations in liquid propylene on iPP
isotacticity and molecular mass are available for a few C
2
-symmetric bisindenyl zirconocenes.
204,267
On the other hand, the above phenomena do not apply to C
1
-symmetric (nor to syndiospecific C
s
-symmetric)
metallocenes: for these, decreasing monomer concentration either increases the isotacticity and melting point of iPP
or has no relevant effect. This is due to the mechanism of site epimerization (also referred to as chain backskip,
Scheme 28), in which the chain, at the lower monomer concentrations, has a higher chance to migrate to the less
hindered site, which is usually also the more stereoselective. For the same reason, increasing the polymerization
temperature either increases the melting point of an isotactic poly(c-olefin), or has no relevant effect.
725
17 18 19 20 21 22 ppm
m
m
r
r

m
m
m
m

m
m
m
r

m
r
r
m
1 2
5
4 3
m m m r r
1 2
3
4 5
m
Figure 25
13
C NMR spectra of the methyl region of iPP having almost only 2,1-erythro-regiodefects (top) and only
stereodefects (bottom). The relative intensity of the mmmr, mmrr, and mrrm pentads is 2: 2 : 1, the fingerprint of site-controlled
isotactic chain growth.
Olefin Polymerizations with Group IV Metal Catalysts 1061
Table 6 Selected C
2
-symmetric isospecific zirconocenes containing heterocycles
Zirconocene Activator
T
p
(

C)
M

w
( 10
3
) T
m
(

C) NMR
a
Notes References
S
S
Me
2
Si ZrCl
2
53
MAO 70 445 156 0.41% mrrm 0.3%
2,1
Unsupported, liquid
propylene
720
S
S
Me
2
Si ZrCl
2
54
MAO 70 604 160 0.35% mrrm 0.2%
2,1
Unsupported, liquid
propylene
720
Ph
3
CB(C
6
F
5
)
4
/
TIBA
70 1,165 162 Unsupported, liquid
propylene
720
S
S
Me
2
Si ZrCl
2
55
MAO 70 795 160 0.26% mrrm,
0.2% 2,1
Unsupported, liquid
propylene
720
N
ZrCl
2
N
Me
2
Si
56
MAO 70 69.3 154.4 97% mmmm
0.17% 2,1
Unsupported, liquid
propylene, 1% H
2
in gas phase
63
N
ZrCl
2
N
Me
2
Si
57
MAO 70 55.5 156.3 97.7% mmmm
0.34% 2,1
Unsupported, liquid
propylene, 1% H
2
in gas phase
63
Si
ZrCl
2
O
O
58
MAO 30 733, 000 159.1 97.3% mmmm Toluene 716,731
a
When allowed by data description, %mmmm refers to primary insertions only.
1062 Olefin Polymerizations with Group IV Metal Catalysts
C
1
-symmetric structures having one of the two Cp ligands endowed with bilateral symmetry (such as fluorenyl)
have several synthetic advantages, the main being the absence of a meso-isomer (Figure 28). We recall here that the
latter is instead formed as the undesired byproduct in the synthesis of the C
2
-symmetric chiral ansa-metallocenes.
In general, C
1
-symmetric structures based on fluorene give lower molecular masses than the best C
2
-symmetric
ones, and are also less active, with very few exceptions. Two types of C
1
-symmetric structures have been developed,
one based on fluorene (or related ligands with bilateral symmetry) and substituted cyclopentadienes, and the other
based on the same bilaterally symmetric ligands linked to substituted indenyls. Most of the latter have been
developed to generate elastomeric PPs of low isotacticity and are discussed in the next section. Zirconocene
complexes containing substituted indenyls usually give higher molecular mass iPP compared to the ones having
substituted cyclopentadienyls. The zirconocene complexes based on dithienocyclopentadienyl also produce higher
Table 7 Selected C
2
-symmetric isospecific zirconocenes based on azulenyl and related rings
Zirconocene Activator
T
p
(

C)
(M

w
(10
3
) T
m
(

C) NMR Notes References


Me
2
Si ZrCl
2
59
MAO 70 350 151 96.8% mmmm
Unsupported, liquid
propylene
732
Me
2
Si ZrCl
2
60
MAO 70 370 156 96.0% mmmm
Unsupported, liquid
propylene
732
Me
2
Si HfCl
2
61
MAO 70 2,500 160
Unsupported, liquid
propylene
Iwama
JOMC
2005
Oct
n
Oct
n
Me
2
Si HfCl
2
62
MAO 70 90 Amorphous
83.5% mmmm
12.5% 3,1
Unsupported, liquid
propylene
733
Olefin Polymerizations with Group IV Metal Catalysts 1063
molecular masses compared to the ones based on fluorenyl. Selected C
1
-symmetric structures and polymerization
results related to the high isoselectivity structures are listed in Table 8 (cyclopentadienyls) and Table 9 (indenyls).
4.09.4.2.3 Low isotacticity: from flexible to elastomeric isotactic polypropylene
Metallocenes purposely designed for the synthesis of stereodisordered, low melting, soft and highly flexible PPs have
been reviewed up to the end of 1999.
162
In subsequent years, the line of research aimed at the preparation of low-
crystallinity plastomeric or elastomeric iPP has aroused increasing interest, both from the perspective of catalyst
50
60
70
80
90
100
0 4 8 10 12
[Propylene] (mol l
1
)
m
m
m
m

(
%
)
6 2
Figure 26 Influence of propylene concentration on iPP isotacticity (%mmmm). ^: rac-C
2
H
4
(Ind)
2
ZrCl
2
/MAO at 50

C in
toluene; *: rac-C
2
H
4
(4,7-Me
2
Ind)
2
ZrCl
2
/MAO at 50

C in toluene;
&
: rac-Me
2
C(3-Bu
t
Ind)ZrCl
2
/MAO at 50

C in pentane;
N : rac-Me
2
Si(2-Me-4-PhInd)ZrCl
2
/MAO at 60

C in hexane.
0
2,000
4,000
6,000
80,00
10,000
0 4 8 10 12
[Propylene] (mol l
1
)
A
v
e
r
a
g
e

d
e
g
r
e
e

o
f

p
o
l
y
m
e
r
i
z
a
t
i
o
n
2 6
Figure 27 Influence of propylene concentration on iPP molecular mass. : rac-C
2
H
4
(Ind)
2
ZrCl
2
/MAO at 50

C in toluene;
: rac-Me
2
C(3-Bu
t
Ind)ZrCl
2
/MAO at 50

C in pentane; : rac-Me
2
Si(2-Me-4-PhInd)ZrCl
2
/MAO at 60

C in hexane.
722
Mt
P
R
Mt
R
P
site
isomerization
Scheme 28
1064 Olefin Polymerizations with Group IV Metal Catalysts
design as well as from the standpoint of material properties. In addition to manufacturing difficulties of these
polymers in slurry or gas-phase processes, the most important limitation to the practical use of these materials is
the relatively high glass transition temperature of PP (T
g
~0

C), which prevents the use at sub-ambient tempera-


tures, where PP becomes brittle. We recall that the T
g
of a polyolefin, being a property of the amorphous phase, is not
(or very little)
222
affected by tacticity.
The most successful classes of metallocene catalysts studied for low-tacticity iPP are: (i) the fluxional bis(2-
arylindenyl) metallocenes first conceived and demonstrated by Waymouth and Coates
748
and recently
reviewed;
749,750
(ii) a few examples of C
2
-symmetric, 3-alkyl-substituted ansa-bis(indenyl) zirconocenes;
222,709,751
and (iii) several types of C
1
-symmetric catalysts.
Of all the catalyst types, the sterically fluxional bis(2-arylindenyl) complexes
749,752768
produce elastomeric PP
with the best combination of properties, that is, relatively high melting points and very low crystallinity, due to their
stereoblock nature,
769772
but unfortunately their activity and molecular mass capability are too low at industrial
polymerization temperatures (6080

C). The mechanism of stereoblock formation originally proposed by the


inventors
748
has recently been questioned and an active role of the counterion has been proposed to better
account for the heptad distribution in the
13
C NMR spectra of the elastomeric, stereoblock PP.
773,773a
These
polymers are PP reactor blends, since they can be fractionated with solvents into low- and high-tacticity
components.
162,767,769,770,774,775
The C
2
-symmetric zirconocenes of type (ii) have two key structural features: a single carbon bridge, and a
3-isopropyl substituent on indene. For any alkyl substituent on C(3) of indene, all complexes are fully regioselective
(through
13
C NMR at 100 MHz), and the size of the 3-R substituent dramatically affects PP microstructure and
molecular mass, as shown in Table 10.
A comparison of the physical properties of these polymers with that of fully amorphous PP has been reported.
722
These C
2
-symmetric zirconocenes are made from relatively inexpensive ligands, but have some limitations: their
synthesis produces an isomer ratio rac/meso <1 and their catalytic activities are not very high. In addition, the
elastomeric PP cannot be produced with controlled morphology in liquid propylene and, as is the case of the bis(2-
arylindenyl) systems mentioned above, do not give sufficiently high molecular masses in solution at industrial
polymerization temperatures.
Of the three classes designed for elastomeric or plastomeric PP, the most successful seems to be the class of C
1
-
symmetric structures. These are based on the bilaterally symmetric fluorenyl ligands, first developed by Ewen,
209
and have received a great deal of attention due mainly to three facts: (i) they are far simpler to synthesize than the
chiral isospecific C
2
-symmetric metallocenes; (ii) they can cover a very broad range of stereoselectivity by structural
modification of one ligand only; and (iii) due to the presence of two different active sites, they offer a more potent
mechanistic tool and intellectual challenge.
The propylene polymerization performances of the cyclopentadienylfluorenyl
212,740,741,777
and cyclopenta-
dienyldithienocyclopentadienyl
742,778,779
systems have been recently reviewed.
742
The major drawbacks limiting
the use of these catalysts are an often low activity and low molecular masses, with only a couple of exceptions.
A selection of results related to the low stereoselective catalysts of this type is shown in Table 11.
Ph
2
C
Me
2
C
Me
2
C
Me
2
C
ZrCl
2
ZrCl
2
ZrCl
2
ZrCl
2
Syndiotactic
181
Isotactichemiisotactic
740
Isotactic
741
Amorphous Amorphous T
m
132144 C T
m
130150 C
Hemiisotactic
209
Figure 28 Influence of the Cp-C(3) substituent in C
1
-symmetric fluorenyl-ansa-cyclopentadienyl complexes.
Olefin Polymerizations with Group IV Metal Catalysts 1065
Table 8 Propylene polymerization results with MAO-activated (3-RCp) C
1
-symmetric structures
a
Pre-catalyst T
p
(

C) mmmm (%) T
m
(

C)
M

v
References
Me
2
C ZrCl
2
60 77.5 127 62,000 (M

w
) 741
ZrCl Me
2
C
2
60 80.1 129 47,800 742
Me
2
C ZrCl
2
60 86.3 144 402,000 (M

w
) 741
Me
2
C ZrCl
2
70 154 321,000 (M

w
) 743
Me
2
C
S S
ZrCl
2
60 82.9 139 50,100 742
Me
2
Si ZrCl
2
60 148 68,000 744
Ph
2
Si ZrCl
2
70 144 47,000 (M

w
) 745
Me
2
Si
S S
ZrCl
2
60 92.1 148 11,100 742
S S
Me
2
Si ZrCl
2
SiMe
3
70 93.8 150 39,900 742
a
All data from liquid propylene polymerizations.
1066 Olefin Polymerizations with Group IV Metal Catalysts
Table 9 Propylene polymerization results with MAO-activated indenyl C
1
-symmetric structures
Pre-catalyst T
p
(

C) mm (%) T
m
(

C)

M
w
References
ZrCl
2
70 91 151 27,000 746
ZrCl
2
70 92 142 50,000 747
S S
Me
2
Si ZrCl
2
70 93.8 148 107,400 (M

v
) 725
S S
Me
2
Si ZrCl
2
70 96.2 156 139,900 (M

v
) 725
a
All data from liquid propylene polymerizations.
Table 10 Influence of the bridge and indenyl C(3) substituents on PP structure. Liquid propylene polymerization
results at 50

C
222,776
Me
2
C
ZrCl
2
Me
2
C ZrCl
2
Me
2
C ZrCl
2
Me
2
C ZrCl
2
mmmm (%) 94.8 15.6 n.a.
a
80.7
T
m
(

C) 124.6 Amorphous None 127


M

w
111,400 164,000 ~8,000 ~12,000
H
2
C ZrCl
2
H
2
C ZrCl
2
ZrCl
2
H
2
C ZrCl
2
mmmm (%) 97 25.5 20.0 71.4
T
m
(

C) 162 Elastomeric Amorphous 110


M

w
236,800 100,600 ~21,000 3,100
a
n.a. =not available.
Olefin Polymerizations with Group IV Metal Catalysts 1067
Complexes based on the indenylfluorenyl,
576,725,785,786
heteroindenylfluorenyl,
787
and indenyldithienocyclo-
pentadienyl systems
725,738,788
show improved catalytic performance, producing PPs with often higher molecular
masses. Some of these catalysts have very high activities in liquid propylene. Selected structures with polymerization
results are shown in Table 12.
Table 11 C
1
-symmetric cyclopentadienylfluorenyl and cyclopentadienyldithienocyclopentadienyl complexes
Me
2
C ZrCl
2
Me
2
C ZrCl
2
Me
2
Si ZrCl
2
Me
2
C ZrCl
2
SiMe
3
[Propylene] 2 bar 2 bar Liquid Liquid
T
p
(

C) 70 70 60 60
mm (%) 52.7 63.2 77.3
mmmm (%) 31.4 44.0 64.4 67.0
T
m
(

C) 109
M

v
62,000 (M

W
) 64,000 (M

n
)
References 212 212 780 781,782
Me
2
C ZrCl
2 ZrCl
2
Me
2
C ZrCl
2
Ph
2
C ZrCl
2
[Propylene] Liquid 1.29 mol l
1
Liquid Liquid
T
p
(

C) 20 30 20 20
mm (%) 50.1 70.8 50.6 43.5
mmmm (%) 31.8 57.5 31.4 25.3
T
m
(

C) 149 n.r. 88 125


M

W
535,000 (M

W
) 270,000 (M

v
) 81,900 (M

W
) 435,000 (M

W
)
References 740 783 740 740
Ph
2
C HfCl
2
S S
Me
2
C ZrCl
2
Me
2
C ZrCl
2
Ph
2
C ZrCl
2
diastereoisomer
[Propylene] Liquid Liquid Liquid Liquid
T
p
(

C) 20 40 70 20
mm (%) 45.0 44.6 60.2 38.6
mmmm (%) 24.0 23.2 18.5
T
m
(

C) 135 147
M

W
806,000 128,000 29,000 390,000
References 740 778 784 740
S S
Me
2
C ZrCl
2
S S
Me
2
Si ZrCl
2
S S
Me
2
C ZrCl
2
S S
Me
2
C ZrCl
2
[Propylene] Liquid Liquid Liquid Liquid
T
p
(

C) 60 60 60 60
mm (%) 58.7 84.0 73.2 66.5
mmmm (%) 42.3 73.2 59.9 49.24
M

n
45,500 140,000 80,000 113,000
T
m
(

C) 80 123 103 80
References 742 742 742 742
1068 Olefin Polymerizations with Group IV Metal Catalysts
The average isotactic sequence length determines the average crystal lamella thickness, which in turn is directly
related to the heat of fusion and melting temperature of iPP. The correlation between microstructure and DSC
melting point of a series of fully regioregular, as-polymerized (neither moulded nor extruded) iPP samples is shown in
Figure 29. Several detailed studies on the influence of the amount and type of stereoerrors on highly defective,
Bernoullian (random) PP have been reported.
738,786,789,793
Table 12 C
1
-symmetric indenylfluorenyl, heteroindenylfluorenyl, and indenyldithienocyclopentadienyl complexes
ZrCl
2 ZrCl
2
HfMe
2
Me
2
Si ZrCl
2
[Propylene] Liquid Liquid Liquid
a
Liquid
T
p
(

C) 70 50 50 70
mm (%) 57 71
mmmm (%) 27 34
T
m
(

C) 114 117 87

M
W
40,000 200,000 700,000 50,000
References 746 789 789 747
Me
2
Si ZrCl
2
Me
2
Si ZrCl
2
ZrCl
2
O
O
HfMe
2
S
[Propylene] Liquid 5.1 M 5.1 M Liquid
a
T
p
(

C) 70 30 30 30
mm (%) 77.4 62
mmmm (%) 65.2 44 35.8 77.7
T
m
(

C) 110

M
W
152,000 (M

v
) 158,000 115,000 254,000
References 725 790 791 792
HfCl
2
HfCl
2
Me
2
Si
ZrCl
2
N Me
2
Si ZrCl
2
S
[Propylene] Liquid Liquid Liquid Liquid
T
p
(

C) 40 40 60 70
mm (%) 81.7
mmmm (%) 58 72 71.4 56
T
m
(

C) 117 125

M
W
300,000 300,000 236,700 149,000
References 793 793 725,787 787
S S
Me
2
Si ZrCl
2
S S
Me
2
Si ZrCl
2
S S
Me
2
Si ZrCl
2
S S
Me
2
Si ZrCl
2
[Propylene] Liquid Liquid Liquid Liquid
T
p
(

C) 70 70 70 70
mm (%) 79.8 67.0 82.2 90.2
mmmm (%) 68.6 51.1 72.1 84.2
T
m
(

C) 116 84 119 138


M

v
231,500 123,400 209,000 169,600
References 725 725 725 725
a
Activated with CPh
3
[B(C
6
F
5
)
4
].
Olefin Polymerizations with Group IV Metal Catalysts 1069
4.09.4.2.4 Syndiotactic crystalline and elastomeric polypropylene
The report by Ewen and co-workers that a metallocene of C
s
-symmetry was able to produce a highly syndiotactic,
fully regioregular PP by site-controlled primary polyinsertion provided the first clear-cut evidence that metallocene
catalysts operate by a dual-site chain-migratory mechanism.
181,208
Figure 30 shows the
13
C NMR methyl region of
sPP produced with 65/MAO at 50

C in liquid propylene. This spectrum clearly shows the fingerprint of stereoerrors


due to enantioface misinsertion (rrrm=rrmm=2rmmr) and of chain backskip (site isomerization) leading to the
formation of an isolated m dyad (rrmr =rrrm). Since this spectrum is at the level of heptad resolution, the rrrrmr and
rrrrmm heptads are separated. The sequence of insertions leading to the different stereosequences is shown in
Scheme 29. Chain-end epimerization (see Section 4.09.5.2.1) does not seem to occur with C
s
-symmetric metallo-
cenes.
182,794
In addition, syndioselective catalysts are also highly regioselective.
164,795
In contrast to the isospecific C
2
-symmetric metallocenes, which required 20 years of structural improvement
in order to achieve acceptable molecular mass capability and high isotacticity, in the case of C
s
-symmetric com-
plexes, the evolution from the original Me
2
C(Cp)(9-Flu)ZrCl
2
(65, Table 13) resulted very early in the complex
50
70
90
110
130
150
170
30 40 50 60 70 80 90 100
Isotacticity, m diad or mmmm pentad (%)
T
m

(

C
)
mmmm
m
Figure 29 Melting point, T
m
(second melt, heating rate 10

min
1
), of regioregular iPPs made with C
1
- and C
2
-symmetric
metallocene catalysts, as a function of isotacticity ( : mmmm %; : m %,).
725,742
19.5 20.0 20.5 21.0 21.5 22.0
ppm
r
m
m
r
m
m
r
r
r
r
m
r
r
r
r
r
m
m
r
r
r
r
r
r
r
r
m
r

Figure 30 Methyl region of
13
C NMR spectra of sPP obtained from 65/MAO at 50

C in liquid propylene (see Table 13).


1070 Olefin Polymerizations with Group IV Metal Catalysts
Ph
2
C(Cp)(9-Flu)ZrCl
2
(66, Table 13),
745,796
which on MAO activation produces sPP with very high molecular
masses and requires H
2
for molecular mass control. Nevertheless, the syndioselectivity of 66 not being very high,
the quest for a more syndioselective catalyst (hence producing sPP with a higher melting point and hopefully a
faster sPP crystallization rate) has continued in recent years, with some improvements with respect to the parent
cyclopentadienylfluorenyl-based structures (selected data are reported in Table 13).
All aspects of syndioselective propagation with C
s
-symmetric catalysts, the influence of ligand,
794,797807
metal,
806,808,809
and counteranion
100,104
variations, as well as the influence of the polymerization conditions,
805,810
have been studied in detail and reviewed.
162,181,209,288,811
Syndiotactic PP is a thermoplastic with a slightly lower melting point (150155

C), lower crystallization rate, and


higher flexibility (elastic modulus ca. 500) than iPP. It also has lower density, lower heat sealing temperature, but also
better impact properties and better transparency.
812,813
One of the difficult aspects of sPP as a material, but a great
playground for polymer chemists and physicists, is its polymorphic behavior: sPP has complex thermal properties and
can crystallize in four different forms, some of which can interconvert.
814821
A different, possibly more productive, research direction has been the development of less syndioselective
catalysts, based on modifications of the silyl-bridged cyclopentadienylamido titanium complex.
Based on the symmetry requirements for syndioselectivity, the C
s
-symmetry of most of ansa-Cpamido titanium
complexes should render them syndiospecific by site control. However, Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
shows only minor
syndioselectivity (about 50% rr triads), very likely because the bulky tetramethylcyclopentadienyl moiety cannot
easily accommodate the methyl group of the coordinated propylene molecule in its syndioselective placement
(methyl opposite to the preferred conformation of the growing chain, which is away from the C
5
Me
4
ligand).
Replacing tetramethylcyclopentadienyl by fluorenyl does increase syndiotacticity, at least for polymerizations in
heptane at 0

C.
822
However, several titanium complexes with more expanded Cp rings have shown improved
syndioselectivity. Examples include structures with t-butylamide linked to 2,7-di-tert-butylfluorene,
782
3,6-di-tert-
butylfluorene,
781,823
indenoindole,
824
or indenopyrrole.
824
Syndioselectivity of these complexes never reaches that of
the cyclopentadienylfluorenyl complexes described above, but nevertheless it is high enough to produce PP which
is partially crystallizable and highly elastic due to the formation of small crystalline domains in the predominantly
amorphous phase. The heterocyclic Ti complexes produce PP of very high molecular masses, even at polymerization
temperatures as high as 80

C. Syndiotactic pentad contents range from 48% to 57% rrrr, with very low amounts of
P
Stereoregular
insertion
P
Site isomerization
(chain backskip)
C
S
-M P C
S
-M
C
S
-M
C
S
-M
P
Stereoregular
insertions
Stereoregular
insertions
Stereoregular
insertions
Stereoregular
insertions
Stereoirregular
insertions
.... r r r r m r r r r r .... .... r r r r m m r r r r....
Scheme 29 Insertion sequences leading to the two major types of enchainment defects in propylene polymerization at a
C
s
-symmetric active center: site isomerization (left) and enantioface misinsertion (right).
Olefin Polymerizations with Group IV Metal Catalysts 1071
Table 13 Selected results from syndioselective bis-cyclopentadienyl complexes
Pre-catalyst Activator
Polym.
conditions Syndiotacticity
Average
molecular mass
(M

v
or M

w
)
T
m
(

C) References
Me
2
C
ZrCl
2
65
MAO
0

C, liquid
propylene
88.1 rrrr 140,000 145 162
50

C, liquid
propylene
82 rrrr 10
5
137 162
C
ZrCl
2
66
MAO
20

C, liquid
propylene
98.7 rr 8.4 10
5
147 794
MAO
40

C, liquid
propylene
92.7 rr 3.8 10
5
137 794
MAO 60

C, liquid
propylene
91.0 rr 2.7 10
5
132 794
C
ZrCl
2
67
MAO 60

C 88.5 rrrr 5.1 10


5
143 741
Me
2
C
ZrCl
2
68
MAO
0

C, liquid
propylene
95.5 rr 5.3 10
5
154 806
ZrCl
2
Me
2
Si Me
2
Si
69
MAO
20

C, liquid
propylene
96.9 rr 1.2 10
6
151 794
50

C, liquid
propylene
95.5 rr 3.3 10
5
140 794
70

C, liquid
propylene
89.5 rr 1.6 10
5
119 794
S
S
Me
2
C
ZrCl
2
70
MAO
50

C, liquid
propylene
84.0 rr 10
5
110 717,742
1072 Olefin Polymerizations with Group IV Metal Catalysts
regioerrors. These PPs have been studied in detail in terms of their physical properties by De Rosa and co-
workers.
825827
The high molecular masses also contribute to improve elasticity. For example, catalyst 73/MAO
produces sPP with a tensile modulus <30 MPa and elongation at break 600%.
Some examples are shown in Table 14, together with a selection of polymerization results. It is worth noting the
higher syndioselectivity of the Zr complex 74 compared to that of 73, and the even higher syndioselectivity of the Zr
complex, which bears the expanded octamethyloctahydrodibenzofluorenyl ligand (99% rrrr, T
m
=165

C at 15

C
in liquid propylene).
942
The very low activities and sPP molecular masses of 74/MAO however render this complex of
little interest.
Figure 31 shows the correlation between melting points and stereoregularity in the syndiotactic domain. As is the
case of isoselective single-center catalysts, the degree of stereocontrol decreases by increasing the polymerization
temperature for the syndioselective ones. This behavior is shown in Figure 32.
4.09.4.2.5 Semicrystalline propylene/ethylene co-polymers
On reading Section 4.09.4.2, it must have appeared evident that the very beauty of metallocene catalysts tunability
of selectivity and molecular mass by ligand design is also their major weakness, at least from a practical (industrial)
perspective. In fact, each single-center catalyst produces its own polymer, and the polymer properties can be
modified only to a limited extent by varying the polymerization conditions, especially when the molecular mass of
the polymer and the activity of the catalyst have to be kept within the range of practical applications. Developing a
catalyst for an industrial process is a very expensive and time-consuming exercise. In addition to being highly active,
toxicologically safe, inexpensive, and fit for the polymerization processes at which it is aimed, a successful catalyst
must also be versatile. Although the dream of one catalyst fit for all polymers and all processes will hardly ever
come true, an ideal catalyst must have high performance in both liquid and gas phase, a high molecular mass
capability, a good hydrogen response, and must react efficiently with more than one monomer. In other words, it
must be able to produce as many different polyolefin materials as possible.
These are the main reasons that highlight the true value of single-center catalysts and metallocenes in particular,
with respect to Ti-based ZieglerNatta and Cr-based Phillips systems: a very good co-monomer distribution ability
(from random to alternating), coupled with the generally much higher reactivity toward the higher and less reactive
olefinic co-monomers.
It is obviously much easier to tune polyolefin properties by adding co-monomers, rather than modifying the
polymerization conditions or the catalyst itself. In addition, co-monomers enable the tuning of such properties, like
the glass transition temperature, that cannot be changed by changes in tacticity, in the case of poly(c-olefins). This
aspect has been extensively demonstrated in the case of both ethylene co-polymers (see Section 4.09.4.1) and
propylene co-polymers. Below, we describe a few recent examples of the use of ethylene as co-monomer, and the
following sections will cover the use of higher olefins.
Ethylene is used to modify the crystallinity of iPP: small amounts (110%) of ethylene reduce T
m
, heat of fusion
and T
g
of iPP, thus increasing transparency and improving ductility, flexibility, and impact properties of iPP. This
effect is stronger in metallocene-made propylene/ethylene co-polymers compared to ZN ones, due to a much better
co-monomer distribution, and the absence of the highly modified, low molecular mass amorphous fraction typically
formed with ZN catalysts. The co-polymer becomes soluble in hydrocarbons between 5 and 15 mol% ethylene,
depending on stereoregularity, and is fully amorphous above 20 mol%. Several metallocene catalytic systems have
been investigated for the production of plastomeric propylene/ethylene co-polymers, with the most relevant results
being limited to the patent literature. The three key physical properties of these plastomeric co-polymers are melting
point and heat of fusion (crystallinity), solubility in hydrocarbons (usually determined as the amount of co-polymer
soluble in xylene at room temperature), and the glass transition temperature. The most apparent, and easier to
measure, effect of ethylene on the structure of iPP is on the melting point. The decrease of melting point and heat of
fusion induced by ethylene incorporation on PP from a typical highly stereoselective C
2
-symmetric metallocene is
shown in Figure 33.
The main drawback of propylene/ethylene co-polymerization with metallocenes is the frequent strong decrease of
molecular mass at increasing ethylene incorporation. The available mechanistic explanation is a fast chain transfer to
ethylene after a propylene insertion (Scheme 30), and has been reported for metallocenes of different symmetries: C
2 v
-
symmetric,
250
C
2
-symmetric,
222,251253
and fluxional bis(2-arylindenyl).
223
This phenomenon has strongly limited the
development of isoselective metallocenes for iPP until the introduction of the 2-isopropyl-indenyl ligands designed to
solve the problem.
586,721,722,828,829
An alternative solution to low molecular masses is the use of hafnocenes, but
activities are lower.
584
Olefin Polymerizations with Group IV Metal Catalysts 1073
Table 14 Selected results from syndioselective cyclopentadienyl amide complexes
Pre-catalyst Activator
Polym.
conditions Syndiotacticity
Average
molecular mass
(M

v
or M

w
)
T
m
(

C) References
Me
2
Si
TiCl
2
N
71
MAO
60

C, liquid
propylene
50.2 rr 6.5 10
5
Amorphous 824
Me
2
Si
TiCl
2
N
72
MAO
60

C, liquid
propylene
63.9 rrrr 5 10
5
Amorphous 782
Me
2
Si
TiCl
2
N
73
MAO
60

C, liquid
propylene
75.8 rrrr
86.2 rr
3.5 10
5
105 781
Me
2
Si
ZrCl
2
N
74
MAO
60

C, liquid
propylene
85.7 rrrr
93.7 rr
6 10
3
145 781
Me
2
Si
Ti Me
2
N
N
75
MAO
80

C, liquid
propylene
52 rrrr
71.2 rr
10
6
62 824
N
Et
Me
2
Si
TiMe
2
N
76
MAO
70

C, liquid
propylene
54.1 rrrr
72.8 rr
1.3 10
6
824
1074 Olefin Polymerizations with Group IV Metal Catalysts
4.09.4.2.6 Propylene/butene co-polymers
Isotactic C
3
/C
4
co-polymers with C
4
<15% are made with ZieglerNatta catalysts in order to improve transparency,
increase flexibility, and lower the melting point of iPP.
830832
C
3
/C
4
co-polymers with higher content of butene are
commercialized by Mitsui under the trade name TAFMER XR or XM.
833
. According to data from the Elastomers
Division of Mitsui Chemicals, such co-polymers have excellent compatibility with various polyolefins and can easily
be blended with them. In general, TAFMER modifiers improve the flexibility, impact resistance, and heat sealability
60
80
100
120
140
160
50 60 70 80 90 100
rr (%)
T
m

(

C
)
Figure 31 Melting point of sPPs made with C
s
-symmetric catalysts, versus the rr triad content. Note that samples with
rr <75%are amorphous in the second DSC heating scan, and require several days of annealing at room temperature to develop
measurable crystallinity.
40
50
60
70
80
90
100
20 30 40 50 60 70 80 90
Tp (C)
r
r

%
Figure 32 Temperature dependence of the stereoselectivity of selected C
s
-symmetric metallocenes.
&
: Me
2
Si(C
5
Me
4
)(NBu
t
)-
TiMe
2
/MAO,
824
N : Me
2
Si(2,5-dimethyl-5,6-dihydroindeno-indol-6-yl)(NBu
t
)TiMe
2
/MAO,
824
^: Me
2
Si(3,6-Bu
t
2
Flu)(NBu
t
)TiMe
2
/
MAO 63,
781
v: (1,2-SiMe
2
)
2
(C
5
H
2
)(C
5
H-3,5-Pr
i
2
)ZrCl
2
/MAO 64.
794
Olefin Polymerizations with Group IV Metal Catalysts 1075
of PP and PE, and are used in packaging applications such as films, extrusion laminates, and sheets. Isotactic
propylene/butene co-polymers show higher melting points and higher elastic moduli at the same molar co-monomer
content compared to isotactic propylene/ethylene co-polymers, allowing for higher use temperatures. On the other
hand, they have higher glass transition temperatures, limiting applications at sub-ambient temperatures. Both T
g
and
modulus can be further lowered by addition of ethylene as termonomer.
Obviously, as is the case of ethylene/propylene co-polymers, the melting point of the co-polymers is a function of
both catalyst stereoselectivity and butene content. A major difference between propylene/ethylene and propylene/
butene co-polymers is that while the former become amorphous at ethylene content above about 25 mol%, the latter
can develop crystallinity in the whole range of composition. This is true for both the isotactic
315,834
and
syndiotactic
835,836
co-polymers due to an easier inclusion of the co-monomer in the crystalline lamellae.
Syndiotactic co-polymers have been prepared with both Me
2
C(Cp)(Flu)ZrCl
2
and Ph
2
C(Cp)(Flu)ZrCl
2
, and thus
differ mainly by their molecular mass.
Isotactic propylene/butene co-polymers have been produced with a much broader array of metallocene catalysts,
with greatly differing isotacticities and molecular masses.
719,832,834,837840
The melting point/composition relation-
ship for these isotactic co-polymers is shown in Figure 34.
4.09.4.2.7 Propylene/higher c-olefin co-polymers
Propylene has been co-polymerized with a broad set of higher olefins to isotactic co-polymers, including 1-pen-
tene,
842845
4-methyl-1-pentene,
842,846,847
1-hexene,
834,840,842,848852
1-heptene,
842,847
1-octene,
840,842,846,853857
1-decene,
856
1-tetradecene,
856
1-octadecene,
856,858
and norbornene,
859862
as well as a series of non-conjugated
dienes like 1,5-hexadiene,
863
7-methyl-1,6-octadiene,
851,864,865
6-phenyl-1,5-hexadiene and isocitronellene,
864
and
1,9-decadiene.
851
0
20
40
60
80
100
120
140
160
0 2 4 6 8 10 12 14 16 18 20
C
2
(mol%)
T
m

(

C
)
,

H

(
J

g

1
)
H
f
T
m
Figure 33 Isotactic propylene/ethylene co-polymers: influence of ethylene content on crystallinity (DSC data, heating rate
20

Cmin
1
).
-H transfer
to propylene
P
x
E
y
Cp
2
M
+
P
x
E
y
Cp
2
M
+
P
x
E
y
Cp
2
M
+
P
x
E
y
Cp
2
M
+
P
x
E
y
Cp
2
M
+
P
x
E
y
P
x
E
y
Cp
2
M
+
P
x
E
y
Cp
2
M
+
P
x
E
y
-H transfer
to propylene
-H transfer
to ethylene
P
x
E
y
-H transfer
to ethylene
P
x
E
y
Reduced by 2-methyl-
indenyl substitution
Increased by 2-methyl-
indenyl substitution
Not observed
Not observed
Scheme 30 Dashed bond: CHH or CHCH
3
.
1076 Olefin Polymerizations with Group IV Metal Catalysts
Isotactic propylene/1-hexene co-polymers have received the most attention: 1-hexene has been found to be more
efficient than butene in lowering the melting point of the co-polymer
840,847
(see Figure 35), and their structure and
physical properties have been investigated.
852
Syndiotactic propylene/1-octene co-polymers have been prepared.
866
In propylene/norbornene co-polymerization, cata-
lyst activity strongly decreases with the increase of norbornene in the feed. Very high T
g
co-polymers have been produced at
highnorbornene incorporation. Surprisingly, T
g
differences withthe ethylene/norbornene co-polymers are small (Figure 36).
The co-polymerization between propylene and styrene is not feasible, unless minor amounts of ethylene are added
to the system.
867,868
Several attempts at producing oxygen-functionalized polypropylenes by co-polymerization of
propylene with olefinic alcohols or esters have been reported.
869,870
4.09.4.2.8 Propylene co-polymerization with macromonomers
LCB PP shows enhanced melt strength. Polypropylene long chain branches (PP-LCBs) are generated by electron
beam irradiation of iPP in the solid state.
317
LCBs can also be formed by propylene co-polymerization with vinylene-
terminated macromonomers, taking advantage of the high co-monomer reactivity of the ansa-Cpamido Ti com-
plexes and some chiral metallocenes. These macromonomers can be generated by different catalysts and with
different monomers and microstructures. In the case of propylene, some sterically bulky metallocenes give rise to
u-methyl elimination as the main chain-release reaction (see Section 4.09.2.4): under the proper conditions, this
chain-release reaction allows the production of allyl-terminated iPP
268
and aPP macromonomers,
871
which in turn can
80
100
120
140
160
0 5 10 15 20 25 30
Butene (mol%)
T
m

(

C
)
Figure 34 Melting points of isotactic propylene/butene co-polymers as a function of butene molar content. v: C
3
/C
4
fromrac-C
2
H
4
(Ind)
2
HfCl
2
/MAO,
834,841
&
: C
3
/C
4
from rac-Me
2
Si(2-Me-Benz[e]ind)
2
ZrCl
2
,
837,838
, N : C
3
/C
4
from rac-Me
2
Si(2-methyl-4-aryl indenyl)
and thiopentalenyl zirconocenes,
719,839,840
*: rac-Me
2
Si(2-Me-4-NaphthInd)
2
ZrCl
2
/MAO,
832
^: C
3
/C
4
from ZN catalysis.
830,831,839
0
20
40
60
80
100
120
140
160
0 5 10 15 20 25 30
Co-monomer (mol%)
T
m


(

C
)
C
3
C
6
C
3
C
4
Figure 35 Molecular mass depression in propylene/1-hexene copolymers compared to propylene/butene copolymers.
M : data from Ref: 852; ^: data from Ref: 840.
Olefin Polymerizations with Group IV Metal Catalysts 1077
be used as co-monomers in propylene co-polymerization. Thus, PPs with isotactic backbone and either atactic
871
or
isotactic PP-LCB
270
have been prepared, as well as PPs with atactic backbone and crystalline PP-LCB.
872
Other approaches include the use of difunctional olefins such as 1,7-octadiene,
873,874
1,9-decadiene,
851
or para-
(3-butenyl)styrene.
875
While the former method also generates chain cross-linking (thus unprocessable polymer gels),
the latter leads only to LCB formation through hydrogenolysis after a secondary styrene insertion. Tandem Zr/Fe
catalysis has been used as well.
876
The preparation of iPP with PS branches has been achieved by co-polymerization
of propylene with allyl-terminated PS macromonomers.
877
4.09.4.3 Polybutene
The major producer of isotactic poly(1-butene) (iPB) is Basell, which recently started a 45 kiloton-per-year plant in
the Netherlands.
878
A smaller producer is Mitsui. The properties and applications of iPB, and the early process for its
production have been reviewed
315,879882
.
About 70% of PB production is used for the manufacturing of pipes for cold and hot water supply, and under-floor
heating pipes. The remaining 30% of the production volume includes butene/ethylene random co-polymers and finds
application in highly specialized and fragmented specialties markets, such as components in easy-peel films, process
aids, and many others.
878
TiCl
3
catalysts for 1-butene polymerization have relatively low activity.
883
The old Shell process made use of a
first-generation ZieglerNatta catalyst, and required de-ashing of the final polymer. The new Basell process makes
use of a much improved high yield MgCl
2
-supported Ti-based ZN catalyst and a two-reactor setup.
425,884
Improved
distribution of co-monomer in 1-butene co-polymers with ethylene or propylene can also be achieved using such
catalysts.
885
Metallocenes are far more versatile in controlling polymer stereochemistry compared to ZieglerNatta catalysts, as
extensively demonstrated in the case of PP. Also in 1-butene polymerization, all kinds of chain microstructures can be
obtained with different metallocenes. The
13
C NMR pentad analysis of polybutene is somewhat less immediate than
that of PP, and has been reported for both ZN
886,887
and metallocenes.
180
The
13
C NMR spectrum of atactic
polybutene, with pentad assignments of the C(3) methylene region, is shown in Figure 37.
An early report by Kaminsky and Brintzinger showed that the isospecific metallocene rac-C
2
H
4
(H
4
Ind)
2
ZrCl
2
/
MAO catalyst is able to produce iPB.
203
Kaminsky also reported the influence of polymerization temperature on T
m
and molecular mass, in the polymerization of 1-butene with rac-C
2
H
4
(H
4
Ind)
2
ZrCl
2
/MAO catalyst in the range 15
to 60

C. Melting points decrease with increasing T


p
(11978

C), as do molecular masses (M

v
=45 000 to 5000).
888
The rac-C
2
H
4
(Ind)
2
HfCl
2
/MAO catalyst produces iPB of lower isotacticity (T
m
(II) =90

C), but higher molecular mass


(M

n
= 123000) compared to the Zr analog (T
m
(II) =101

C, M

n
= 12000) at 20

C in 1 : 1 toluene : 1-butene.
Activities were low.
889
An important observation was made by Kioka, who reported the presence of 4,1-units in iPB from rac-
C
2
H
4
(Ind)
2
ZrCl
2
/MAO, and a very strong hydrogen activation effect.
300
This fact was later confirmed by studies of
0
50
100
150
200
250
300
350
400
0 10 20 30 40 50 60 70 80 90 100
NB (mol%)
T
g

(

C
)
Figure 36 Correlation between norbornene content and glass transition temperature in ethylene/norbornene ( )
621
and
propylene/norbornene (^) co-polymers.
861,862
The best linear fit for the propylene/norbornene data extrapolating at T
g
=330

C
for polynorbornene is also shown.
1078 Olefin Polymerizations with Group IV Metal Catalysts
butene polymerizaton in the presence of hydrogen with both rac-C
2
H
4
(Ind)
2
ZrCl
2
/MAO and rac-Me
2
Si(Ind)
2
ZrCl
2
/
MAO catalysts.
301,890
Low molecular mass iPB was produced with rac-Me
2
Si(4,5,6,7-H
4
Ind)ZrCl
2
/MAO at 100

C in order to determine
the structure of end groups and regioirregular units.
232
rac-Me
2
Si(2-Me-Benz[e]ind)
2
ZrCl
2
/MAO produces iPB of low
molecular mass (47 10
3
g mol
1
) and melting point (98

C).
515
The C
1
-symmetric zirconocene Me
2
C(3-MeCp)(9-
Flu)ZrCl
2
, which is hemiisospecific in propylene polymerization, becomes more isoselective in the case of 1-butene,
producing moderately isotactic PB with 46% mmmm.
891
Syndiotactic PB has been produced with Me
2
Si(Cp)(9-Flu)ZrCl
2
/MAO catalyst,
892,893
but only low molecular mass
PB was obtained. sPB crystallizes very slowly and has a low melting point (4050

C). Pure atactic, high molecular


mass PB has been produced with Me
2
Si(9-Flu)
2
ZrCl
2
/MAO.
894
The rac- and meso-Me
2
Si(2,3,5-Me
3
Cp)
2
ZrCl
2
produce isotactic/atactic PB mixtures in low yields and with rather
low molecular masses.
895
The meso-isomer is less active than the racemic isomer. Interestingly, [Ph
3
C][B(C
6
F
5
)
4
] as
activator coupled with TIBA as scavenger seems to produce a more active catalyst than MAO (at Al/Zr =1000). A
similar investigation was performed by the same authors on rac-Me
2
Si(2-MeInd)
2
ZrCl
2
(rac : meso =17 : 83) at 40

C in
toluene (Al/Zr =1000).
896
Also in this case, the meso-isomer was less active than the racemic isomer, the latter also
producing higher molecular mass PB. The best co-catalyst, TIBA/[Ph
3
C][B(C
6
F
5
)
4
] gives iPB of M

n
250 000 and
M

w
,M

n
= 2.9, with an activity of 10 kg
PB
mmol
Zr
1
h
1
(or 22 kg
PB
g
1
h
1
metallocene).
More recently, Kaminsky studied the behavior of rac/meso-mixtures of Me
2
Si(Ind)
2
ZrCl
2
, Me
2
Si(2-Me-4,5-
Benz[e]Ind)
2
ZrCl
2
, and Me
2
Si(2-Me-4-PhInd)
2
ZrCl
2
in the polymerization of 1-butene.
897
The most relevant
observation is that, opposite to what is observed for propylene, the meso-isomer of the indenyl-substituted complexes
are more active than the racemic ones, a fact that might be linked to the regioselectivity of the two isomers, the meso-
isomer being likely more regioselective than the racemic one. Idemitsu Petrochemical
898
produced PB of low
isotacticity, being low melting or amorphous, with doubly bridged metallocenes. Atactic, highly regio-irregular PB
of high molecular mass has been prepared with a half-sandwich complex, Cp
*
Ti(OCH
2
CHTCHPh)
3
/MAO, at
relatively low (050

C) polymerization temperatures.
899
More recently, high molecular mass, partially stereoregular
m
m
m
m

m
m
m
r

r
m
m
r

+

m
m
r
r

m
m
r
m

+

r
m
r
r

r
r
r
r

m
r
m
r
m
r
r
r

m
r
r
m

1
1
2
3
4
2
40 35 30 25 20 15 ppm
3
4

Figure 37
13
C NMR spectrum (100 MHz) of atactic PB. A trace amount of iPB has been added to show the mmmmmm
heptad, marked with H .
Olefin Polymerizations with Group IV Metal Catalysts 1079
PB has been obtained with a similar catalyst system, Cp
*
Ti(OCH
2
Ph)
3
/MAO.
900
MAO-activated thiobis-2,29-(4-
methyl-6-t-butyl-phenoxy)titanium dichloride produces low molecular mass, atactic PB when the co-catalyst is MAO-
modified with isobutyl groups. When MAO pretreated with water is used instead, a very high molecular weight PB is
obtained, with molecular weights above 3 million at T
p
of 40

C.
74
In addition, this water-modified MAO endows the
catalyst with a minor enantioselectivity: the polybutene produced at 25

C has mmmm pentad content of 25%. A


slightly higher productivity and stereoselectivity (30% mmmm at T
p
=40

C) has been reported by the same authors


with a MAO co-catalyst modified with water and pentafluorophenol.
901
The NMR data are not sufficient to
determine the mechanism of enantioselectivity, and whether the microstructure is Bernoullian or stereoblock. No
physical properties have been reported for these polybutenes, but given the pentad content and the molecular
weight, this polymer should be amorphous with some elastomeric properties.
The above elastomeric polybutene might be similar in properties, if not in constitution, to the one produced with
Al
2
O
3
-supported tetrabenzyl and tetraneophyl metals (Ti, Zr, Hf), which is a blend of atactic PB and isotacticatactic
stereoblock PB,
902
that can be fractionated by diethyl ether extraction.
A few reports have highlighted the increase in enantioface selectivity on going from propylene to 1-butene, in
both chain-end-controlled
180,808
and site-controlled polymerizations.
301,891,903
Basell researchers have recently
found that several C
2
- and C
1
-symmetric metallocenes are able to produce high molecular weight PB under
industrial conditions.
903909
A few examples of 1-butene co-polymers with higher c-olefins have been reported. Addition of 1-hexene stabilizes
form II of polybutene and is included in the crystals.
315
C
4
/C
6
random co-polymers are soft or semirigid resins with
characteristics that markedly differ from those of, for example, C
4
/C
2
random co-polymers.
908
In general terms, C
4
/C
6
co-polymers have been reported to have higher crystallinity compared to C
4
/C
2
, that is,
better thermal resistance and lower stickiness, higher flexibility and transparency, excellent workability, and good
impact shock resistance. Co-polymers with 1-octene and 4-methyl-1-pentene have also been reported in the patent
literature.
910
4.09.4.4 Poly(c-olefins) from Monomers Higher than Butene
As far as laboratory-scale model studies are concerned, polymers from linear c-olefins higher than 1-butene combine
the advantages of the liquid monomer (no high-pressure reactors needed) and polymers which are acceptably well
characterizable by
13
C NMR, having a sufficiently resolved c-methylene pentad region.
911
In addition, 1-pentene
and 1-hexene are also low boiling (easily removed after polymerization), facilitating polymer purification. However,
they are amorphous and sticky, thus making their handling quite difficult and messy. Only a few 1-pentene
polymerization studies have appeared,
897,912916
the higher-boiling 1-hexene being largely preferred as
monomer.
891,913,917919
In many instances, 1-hexene has been employed for kinetic
138,154156,920922
and activation studies,
923927
and to
prove the livingness of a catalyst system.
700,928931
Higher c-olefins like 1-octene,
457,846,913915,932
1-decene,
913915
1-tetradecene,
915
and 1-octadecene
914,915
have also been polymerized with different metallocenes. An interesting
aspect of higher olefin polymerization has been the production of ultrahigh molecular weight polymers of 1-hexene
and 1-octene under high pressure.
933935
Syndiotactic poly(1-pentene), poly(1-hexene), and poly(1-octene) have been described.
936
Concerning the branched olefins, the most studied, and the only one with relevant industrial interest, is 4-methyl-
1-pentene. Isotactic poly(4-methyl-1-pentene) has one of the highest melting points (230240

C) among polyolefins.
Its production, properties, and uses have been reviewed.
426,937
As a material, such a high melting point makes it
somewhat difficult to process, and hence it is produced as a co-polymer with minor amounts of higher olefins. These
polymers are manufactured and commercialized by Mitsui under the tradename TPX: TPX has excellent transpar-
ency, heat resistance, and release properties, and is used in a broad range of application areas. These include
industrial materials such as release film, release paper, sheath and mandrels used for the manufacture of high-
pressure rubber hoses, LED molds, food-packaging materials such as heat-resistant-cooking wrap film and bags for
retaining the freshness of vegetables and fruits, as well as conventional applications such as laboratory ware, medical
instruments, and microwave ovenware.
938
The polymerization of 4-methyl-1-pentene with metallocene catalysts has received only occasional attention.
Concerning
13
C NMR spectroscopy, the carbon most sensitive to stereosequences seems to be the backbone
methine (C2), but neither pentads nor triads have been assigned for this polymer. Also the side-group methylene
(C3), which is usually the most sensitive to stereosequences in linear poly(c-olefins),
911
shows extensive overlapping
1080 Olefin Polymerizations with Group IV Metal Catalysts
of the CH
2
pentad signals. Thus, stereoregularity of poly(4-methyl-1-pentene) can only be given at most at the diad
level. The
13
C NMR spectra of the three limit structures of poly(4-methyl-1-pentene), isotactic, syndiotactic, and
atactic, are shown in Figure 38. The carbons are assigned according to the literature.
893,939
Isotactic poly(4-methyl-1-pentene) from C
2
-symmetric zirconocenes melts at lower temperature, _230

C,
846,916
compared with the polymer from ZN catalysts, and has much lower molecular weights. C
1
-symmetric isospecific
zirconocenes can produce poly(4-methyl-1-pentene) of higher isotacticity and correspondingly higher melting points,
but still with low molecular weights at the higher polymerization temperatures.
916
ansa-Indenylamido complexes of titanium produce poly(4-methyl-1-pentene) of low stereoregularity and
molecular weights, which depend on the indenyl substitution pattern. The related fluorenyl complexes produces
higher molecular weights, but with a much lower activity.
542
Highly syndiotactic poly(4-methyl-1-pentene) melts at 210215

C and has been prepared with Ewens catalyst


Me
2
C(Cp)(9-Flu)ZrCl
2
/MAO
892,893,940,941
and the ansa-Cpamido complex Me
2
Si(octamethyloctahydrodibenzo-
fluorenyl)(NBu
t
)ZrCl
2
(OEt
2
).
457,942
Atactic poly(4-methyl-1-pentene) has been prepared with C
2
H
4
(9-Flu)
2
ZrCl
2
/
MAO and is a glassy polymer with a T
g
of about 40

C.
Other branched monomers that have been homopolymerized with metallocenes include 3-methyl-1-butene,
943,944
3-methyl-1-pentene,
943,945,946
4-methyl-1-hexene,
947
allylbenzene,
948,949
vinylcyclohexane,
929
and 1-vinylcyclohexene.
950
4.09.4.5 Polystyrene
Syndiotactic polystyrene (sPS) represents an important achievement in olefin polymerization catalysis. Syndiotactic
PS is an industrially relevant thermoplastic material produced by Dow Chemical and Idemitsu Kosan Co. under the
tradenames Questraand Xarec, respectively. Industrial interest on sPS originates from the remarkable properties
exhibited by this highly crystalline polymer. The high melting temperature, 270

C, the relatively fast crystallization


rate (at least much faster than that of iPS), the high heat resistance, the low dielectric constant, the high elastic
modulus, and an excellent resistance to chemicals explain the industrial interest for this material. Syndiotactic PS was
considered as an innovative new resin option for the automotive, electrical, and electronic markets, appliances such as
3
1
1
2
2
3
4
4
6
5
5, 6
48 46 44 42 40 38 36 34 32 30 28 26 24 ppm
Figure 38
13
C NMR of poly(4-methyl-1-pentene). Isotactic, from rac-H
2
C(3-Bu
t
Ind)
2
ZrCl
2
/MAO (bottom); syndiotactic, from
Me
2
C(Cp)(9-Flu)ZrCl
2
/MAO (middle); atactic, from C
2
H
4
(9-Flu)
2
ZrCl
2
/MAO (top).
Olefin Polymerizations with Group IV Metal Catalysts 1081
microwave oven inner parts, hot water filter and pump cases, films, fibers, packaging materials, housewares, and even
medical applications. Due to low material cost and excellent properties, sPS was considered to be able to compete
with most crystalline materials ranging from polybutylene terephthalate to high-temperature nylons. However, sPS
did not experience the broad market penetration that was anticipated, and in December 2004 Dow Chemical
announced it was going to discontinue sPS production in 2005. Conversely, Idemitsu Kosan Co. announced that it
will resume a sPS plant in Japan in 2007 with a yearly production capacity of 5000 tons.
Syndiotactic PS was discovered at Idemitsu Kosan Co. in 1986.
951953
The first literature report disclosed very few
details about the catalyst used. Later reports indicated that MAO-activated mono-Cp Ti complexes were used. One
year later, Zambelli and co-workers reported that MAO-activated Ti(CH
2
Ph)
4
and Zr(CH
2
Ph)
4
complexes also
promote syndiospecific styrene polymerization.
954,955
These two groups contributed considerably to the develop-
ment and mechanistic understanding of this kind of catalysis.
956964
The syndiospecific polymerization of styrene has
been reviewed.
958,965968
Almost any Ti and many Zr compounds are able to yield sPS, although activity and
syndiotacticity of the resulting polymer is dependent on the metal and on the ligands. Zr-based catalysts are usually
less active and less syndiospecific than the corresponding Ti-based catalysts. Simple halide, benzyl, and alkoxides
such as TiCl
4
, TiBr
4
, Ti(OCH
3
)
4
, Ti(CH
2
Ph)
4
, activated with MAO, yield sPS. Mono-Cp-based systems are however
the most widely investigated systems. The performances of different half-sandwich complexes are compared in
Table 15. It is clear that F atoms bound to the Ti atom in the pre-catalyst remarkably improve activity, probably a
leaving group effect, although a clear rationalization has been provided yet. Permethylated Cp-based catalysts are
notably less active than those with unsubstituted Cp rings. Conjugated aromatic ligands, however, improve activity
and yield reasonably high molecular masses.
969971
In almost all cases, rather low M

w
,M

n
values, typical of single-site
catalysis, are observed.
Molecular masses obtained with alkoxides such as Ti(OBu
n
)
4
in combination with MAO and using toluene as
solvent can reach values up to M

w
= 600000. Considerably lower M

w
values are obtained at 87

C, although the
catalyst system still exhibits good catalytic activity.
963
Both MAO and fluoroborate counterions have been shown to
be suitable co-catalysts for syndiotactic styrene polymerization. Activity up to 10 kg
sPS
mmol
Ti
1
h
1
can be reached
both with MAO and with perfluoroborate co-catalysts such as Ph
3
C[B(C
6
F
5
)
4
] or B(C
6
F
5
)
3
. The sPS yields are
substantially independent of the Al/Ti ratio, provided that ratios higher than 50 : 1 are used.
963
In all cases, the
syndiotacticity of the resulting polystyrenes is remarkably high.
957
Styrene polymerization with bridged bis-metallo-
cenes has also been investigated. In this case, however, very low activity was observed.
956
Besides mono-Cp systems,
958,959,973989
other half-sandwich complexes with different aromatic ligands were
investigated.
969971,982,983,985,990993
Other studies were concerned with polymerization conditions, including
catalyst heterogenization.
990,9941006
Other catalysts such as mono-
276
and bis-benzamidinate
1007
complexes,
bis(phenolato) Ti complexes,
159,1008
and salen-based Ti complexes,
1009
are also able to polymerize styrene to a
highly syndiotactic polymer. Finally, it is noteworthy that bimetallic catalysts based on complex 44 can yield atactic
Table 15 Polymerization of styrene in the presence of different catalytic systems in toluene
T
p
Productivity T
m
Catalyst
(

C) g
sPS
mmol
Ti
1
h
1
M

w
M

w
,M

n
(

C) References
CpTiCl
3
50 1,100 140,000 1.9 258 972
CpTiF
3
50 3,000 100,000 2.0 265 972
CpTi(OBu
n
)
3
45 1,600 40,000 258 973
Ti(OBu
n
)
4
50 568,000 4.1 963
(C
5
Me
5
)TiCl
3
50 15 169,000 3.6 275 972
(C
5
Me
5
)TiF
3
50 690 660,000 2.0 275 972
Ti
Cl Cl
Cl
Ph
50 40,000 277,000 268 971
Ti
Me Cl
Cl
50 38,000 270 970
1082 Olefin Polymerizations with Group IV Metal Catalysts
polystyrene (aPS) with good activity, whereas the corresponding monometallic catalyst shows extremely low activ-
ity.
664
Similar results have been obtained by Do and co-workers utilizing the bimetallic complex 77.
1010
Ti
Cl
Cl
Cl
Ti
Cl
Cl
Cl
77
Comparison of different catalytic systems under different polymerization temperatures is reported in Table 16. It is
clear that CpTiCl
3
is able to give quite high amounts of sPS in the polymerization temperature range from 17 to
90

C. Remarkably, the stereoregularity and the T


m
of the sPS are quite high also at 90

C. Although less active, the


permethylated Cp
*
TiCl
3
/MAO catalyst performs even better in terms of stereoregularity, since the resulting sPS is
substantially perfectly syndiotactic whatever polymerization temperature was used. The Ti(CH
2
Ph)
4
/MAO catalyst
also exhibits remarkably good performances.
Several mechanistic studies, mostly from Zambellis school, have been devoted to understanding the syndio-
specific polymerization of styrene with group-4-based catalysts.
954,960,961,963,964,10111019
Most of these mechanistic
studies have been reviewed.
968
It has been demonstrated that the syndiospecificity is chain-end controlled, since
stereomistakes occur as isolated m diads.
954,960,1011,1012
Regiochemistry of styrene propagation is secondary, and the
same regiochemistry is observed in the initiation as well as in chain-end groups.
960,961
Solvent effects play a role in
syndiospecific styrene polymerization. In the case of aromatic solvents, electron-donating groups (as in mesitylene)
reduce activity, while electron-withdrawing groups (as in 1,2,4-trichlorobenzene) increase activity. Solvents of high
polarity such as CH
2
Cl
2
increase activity, although they have detrimental effects on the syndiotacticity of the
resulting polymers.
1020
As for the oxidation state of the active Ti atoms, ESR experiments, and subsequent combined
ESR and NMR monitoring of the reaction with
13
C-enriched reactants, demonstrated that the most active Ti atoms
are in the oxidation state III.
10211023
However, this topic was rather debated in the literature.
968,1017,1021,1024,1025
In
fact, a carbocationic mechanism was initially suggested, while subsequent investigations clearly demonstrated the
occurrence of an insertion mechanism.
1020,1026
Soon after syndiospecific styrene polymerization, attention was directed to the homopolymerization of substituted
styrenes as well as to their co-polymerization with styrene.
956,957,964,10271029
Mono-Cp-based Ti systems are capable
of homopolymerizing methyl-substituted styrenes and p-chlorostyrene, as well as co-polymerizing them with styrene.
The general trend that emerged is that electron-withdrawing Cl substituents decrease the reactivity relative to
styrene, whereas electron-releasing Me groups increase it. In both cases, syndiotactic co-polymers were obtained.
Table 16 Polymerization of styrene in the presence of different catalytic systems in
toluene (all data from Ref:1011)
T
p
T
m
Catalyst (

C) % Atactic
a
P
r
b
% rrr (

C)
CpTiCl
3
/MAO 17 Traces 1 100 257
0 Traces 0.96 91 258
15 14 0.96 92 255
50 25 0.94 88 243
70 33 0.93 81 243
90 47 0.92 79 242
Cp
*
TiCl
3
/MAO
c
50 Traces 1 265
70 Traces 1 264
90 Traces 1 264
Ti(CH
2
Ph)
4
/MAO 50 Traces 0.97 256
90 10 0.98 260
a
Acetone soluble fraction.
b
Statistical parameter of the syndiotactic propagation.
c
Cp* =C
5
Me
5
.
Olefin Polymerizations with Group IV Metal Catalysts 1083
4.09.4.6 Cyclopolymers
The cyclopolymerization of non-conjugated terminal diolefins, such as 1,5-hexadiene, has been devised by
Waymouth as an additional tool to probe the enantioselectivity of active metallocene sites, in addition to being a
source of novel polyolefinic structures. The regioregular cyclopolymers have four structures of maximum order
(shown in Scheme 31 for poly(methylenecyclopentane)). The trans-isotactic structure, having no mirror planes of
symmetry, is chiral by virtue of its main-chain stereochemistry.
1030
Cyclopolymerization with zirconocene complexes
has been reviewed.
1031
This reaction has also a general mechanistic importance; for example, it rules out the
occurrence of two simultaneously coordinated olefins at the same site, proposed by some authors to explain higher
reaction orders in monomer,
313,1032
since such an arrangement would lead only to cis-rings in the case of 1,5-
hexadiene.
1033
1,5-Hexadiene has also been cyclopolymerized with the CGC Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
/MAO, with good cyclo-
polymerization efficiency.
561
A monocyclopentadienyl complex, Cp
*
TiCl
2
(O-2,6-Pr
i
2
C
6
H
3
), is also active in the
polymerization of 1,5-hexadiene upon activation with MAO, but cyclization efficiency was rather low, producing a
polymer with 2533 mol% buten-1-yl side groups, depending on diene concentration.
1034
This observation contrasts
an earlier report of a high cyclization efficiency shown by [Cp
*
TiMe
2
][MeB(C
6
F
5
)
3
].
1035
The cyclo-co-polymerization of 1,5-hexadiene and ethylene catalyzed by C
1
-symmetric ansa-zirconocenes gives
access to alternating co-polymers,
1036
and lends further supporting evidence for the proposed mechanism of alterna-
tion in ethylene/propylene co-polymerization.
575
The living cyclopolymerization of 1,5-hexadiene has been achieved with monocyclopentadienyl monobenzamidi-
nato complexes Cp
*
{MeC(NR)
2
}ZrMe
2
activated by [PhMe
2
NH][B(C
6
F
5
)
4
] at 10

C.
1037
These catalysts also show
a very high (98100%) selectivity for cyclization. The living nature of these polymerizations allowed the preparation
of iso-poly(1-hexene)/poly(methylenecyclopentane) block co-polymers. Selectivity for cyclization over insertion
decreases as the monomer length increases. 1,7-Octadiene has been shown to cyclize only partially, and the
selectivity is shown to be lowest for the syndioselective zirconocene Ph
2
C(Cp)(9-Flu)ZrCl
2
.
1038
4.09.4.7 Polymers of Cyclic Olefins
The isomerizationpolymerization of cyclopentene to poly(1,3-cyclopentane) has been reviewed.
596
Following
the original work of Kaminsky, reviewed up to 1997,
596
the addition polymerization of NB has been further
studied.
10391042
Polynorbornene from vinyl addition is an amorphous polymer of outstandingly high glass transition
temperature, probably exceeding its decomposition temperature. For a perspective view, see Refs: 10431043a.
Metallocene catalysts are far less active in the vinyl polymerization of NB than late transition metal catalysts.
1040
The
co-polymerization of norbornadiene with ethylene
602
and 1-hexene
1044
has been reported.
4.09.4.8 Polymerization of Conjugated Dienes
Polybutadiene rubber is a technologically important elastomer. Conjugated dienes such as butadiene and isoprene,
which are typically polymerized by anionic lanthanides and late transition metal catalysts, are generally strong
inhibitors for group 4 cationic metal catalysts, and even in co-polymerization with c-olefins strongly depress catalyst
activity. The scope of metallocene catalysts in the polymerization of butadiene, isoprene, and other conjugated
dienes, and the mechanistic implications, have been reviewed in detail.
1045
The homopolymerization of butadiene,
with a series of monocyclopentadienyl titanium trichlorides and trifluorides,
1046
to high molecular weight rubbers
(M

w
<10
6
), with a prevalence of 1,4-cis-butadiene insertion and T
g
in the range 88 to 97

C, has been achieved.


cis-Diisotactic cis-Disyndiotactic
trans-Diisotactic trans-Disyndiotactic
Scheme 31
1084 Olefin Polymerizations with Group IV Metal Catalysts
The same catalysts homopolymerize isoprene. Preparation of syndiotactic styrene/butadiene block co-polymers by
means of CpTiX
3
/MAO catalysts (X=Cl, F) and Cp
*
TiMe
3
/MAO has been reported.
1047
Co-polymerization of conjugated dienes with ethylene by means of metallocene catalysts has been investigated by
several authors. The co-polymerization of ethylene with 1,3-butadiene, 4-vinyl-cyclohexene, 1,4- and 1,5-hexadiene
with rac-Me
2
Si(Ind)
2
ZrCl
2
has been reported.
1048,1049
Subsequently, a detailed
13
C NMR analysis of ethylene/
butadiene co-polymers obtained with the above-mentioned metallocenes was described, which confirmed that
butadiene is incorporated in the 1,4-trans-configuration or is cyclopolymerized to a trans-1,2-methylene-cyclopentane
unit, whereas 1,4-cis- and 1,2-units were not detected. Up to 5%mol butadiene units were incorporated. Cp
2
ZrMe
2
incorporated higher amounts of butadiene relative to rac-C
2
H
4
(H
4
Ind)
2
ZrCl
2
, but molecular masses were lower. In
any case, activity was reduced.
1050
The cyclopentane units are formed as shown in Scheme 32.
Ethylene/butadiene co-polymers with a butadiene content up to 20% have been prepared with
Me
2
Si(C
5
Me
4
)(NBu
t
)TiCl
2
.
1051
The co-polymer with 7 mol% of butadiene has a T
g
of 28

C. The butadiene
units are incorporated partly as 1,2-units, and mainly in the 1,4-trans-configuration. This was rather surprising,
since the same catalyst promotes the homopolymerization of 1,3-butadiene to mainly 1,4-cis- and 1,2-units. The
authors suggest that the relatively high amount of vinyl groups could allow vulcanization. iPP-ran-1,3-butadiene co-
polymers have been obtained using the rac-Me
2
Si(Ind)
2
ZrCl
2
metallocene. By performing a metathesis reaction of
the 1,4-butadiene units in the co-polymer with ethylene, short iPP segments with both chain ends having vinyl bonds
have been obtained.
1052
The synthesis of functionalized ethylene and propylene containing 1,2-methylene-cyclopropane and 1,2-methylene-
cyclopentane units has been achieved, as shown in Scheme 33 for PE.
1053
These structures were obtained from ethylene/1,3-butadiene co-polymerizations at room temperature using
sterically hindered metallocenes such as rac-H
2
C(3-Bu
t
-Ind)
2
ZrCl
2
. Working with ethylene and 1,3-butadiene con-
centrations of 0.24 and 0.62 mol 1
1
, respectively, about 4 mol% of butadiene units are inserted in the resulting PE.
About 67% of the butadiene is present as cyclopropane units, while the remaining 33% is present as cyclopentane
units. In both cases, and in agreement with previous findings,
590,1050
a trans-stereochemistry of the rings, as shown in
Scheme 33, was observed. By contrast, with the less hindered H
2
C(Ind)
2
ZrCl
2
metallocene, no cyclopropane units
were observed in the resulting polymer. The inserted butadiene units (about 21 mol%) were present as cyclopentane
(22%) and 1,4-butadiene (78%) units.
10531055
Furthermore, on increasing the polymerization temperature to 73

C
and at low ethylene concentrations (0.05 mol 1)
1
, the inserted butadiene units are present as 1,1- (76%) and 1,3-
(24%) butadiene units, as shown in Scheme 34.
10541056
L
2
M
+
P
L
2
M
+
P
L
2
M
+
P
L
2
M
+
P
Scheme 32
(a) (b)
Scheme 33
1
2
3
4
(a) (b)
1
2
3
4
Scheme 34
Olefin Polymerizations with Group IV Metal Catalysts 1085
Co-polymerization of ethylene with different conjugated dienes, such as 1,4-pentadiene, yields methylene-1,2-
cyclobutane units in the co-polymers.
559
In the case of propylene/1,3-butadiene co-polymerization with rac-
H
2
C(3-Bu
t
-Ind)
2
ZrCl
2
, a low amount of butadiene units is present as cycles in the resulting co-polymer, since the
butadiene inserts prevailingly as an c-olefin, thus resulting in 1,2-inserted butadiene units. At high butadiene
concentration, sequences of cis-1,4-butadiene units were observed. Interestingly, the co-polymerization of butadiene
forces some regioirregular secondary insertion of propylene.
1057
Co-polymerization of ethylene with cyclic dienes such as 1,3-cyclopentadiene, dicyclopentadiene, and 4-vinyl-1-
cyclohexene using rac-C
2
H
4
(Ind)
2
ZrCl
2
showed that dicyclopentadiene was the most reactive co-monomer. 1,3-
Cyclopentadiene rapidly dimerizes to dicyclopentadiene, and thus ethylene/1,3-cyclopentadiene co-polymerization
actually resulted in ethylene/1,3-cyclopentadiene terpolymers with dicyclopentadiene. Co-polymers with more than
9 mol% of the co-monomer did not show a melt transition.
1058
Ethylene/butadiene cyclopolymerization was used to probe the dual-site nature of C
1
-symmetric ansa-metallo-
cenes.
1059
4.09.5 Polymerization of Ethylene, Propylene, and Higher c-Olefins with other
Single-Center Catalysts
The broadly used term non-metallocene (or post-metallocene) catalysts now comprises a large variety of
different ligands and of metals with different coordination numbers. A simple classification is thus not straightfor-
ward, even if this classification is restricted to catalysts containing group 4 metals. The one we adopted here is based
on the classification used by Do and co-workers in a recent review on the same topic.
1060
Other reviews on this
subject were published a few years ago.
1061,1062
To keep track of the different geometrical and chemical environ-
ments in the pre-catalysts, the complexes are divided into sections according to the metal coordination number, and
then into subsections according to the atoms that coordinate to the metal. However, for the sake of consistency,
pre-catalysts that represent small variations from a largely explored class are discussed within the main class they were
derived from, even if they should be discussed in a different section.
Before discussing catalyst activities, a caveat is in order. Experimentally determined activities strongly depend on
the specific conditions (temperature, polymerization time, solvent, stirring, reactor, and even addition of the different
components to the reactor) and thus a safe comparison of different catalysts is not easy. Following the approach used
by Gibson and co-workers, we report the activities in g
polymer
(mmol metal)
1
h
1
bar
1
, whenever possible, and we
order the catalyst activities according to the scale reported in Table 17.
1061,1062
Finally, it must also be considered that
under similar conditions, metallocenes can easily show activity around 10
4
g
polymer
(mmol metal)
1
h
1
bar
1
.
4.09.5.1 Complexes with Coordination Number 4
In the case of group 4 metals, coordination 4 at the metal atom corresponds to a tetrahedral or pseudo-tetrahedral
geometry of coordination. This geometry is analogous to that presented by metallocenes and CGCs.
4.09.5.1.1 Ligands with coordinating OO atoms
One of the most thorough studies was performed in 1995, when the catalytic behavior of a series of MAO-activated
dialkoxide complexes was tested in the polymerization of ethylene and 1-hexene.
256
At 20

C, MAO-activated 7883
and analogs exhibited the moderate average activity of about 5090 g
PE
(mmol M)
1
h
1
bar
1
. Under the same
conditions, the catalyst based on 84 showed high activity, about 3 10
2
g
PE
(mmol M)
1
h
1
bar
1
, whereas the
catalyst based on 82 exhibited the low activity of 6 g
PE
(mmol M)
1
h
1
bar
1
. Unfortunately, the few molecular
Table 17 Qualitative performance assignment for catalyst activities
Performance Activity (g
polymer
(mmol metal)
1
h
1
bar
1
)
Very low <1
Low 110
Moderate 1010
2
High 10
2
10
3
Very high 10
3
1086 Olefin Polymerizations with Group IV Metal Catalysts
masses reported are rather low, while rather large molecular mass distributions were observed (M

n
<50 000 M

w
/
M

n
>7.0).
256
Catalysts based on 78 and 79 yield highly regioregular and isotactic poly-1-hexene with rather high
molecular mass and narrow molecular mass distribution (M

w
= 675000, M

w
,M

n
= 2.2). Finally, catalysts based on
82 only yield oligomers.
256
Theoretical DFT studies of model systems related to 82 and 83 predicted insertion
barriers around 1015 kcal mol
1
for M=Ti and Zr, respectively. The presence of the CH
2
bridge was shown to have
a minor role. Interestingly, the resting state is stabilized by a -agostic interaction between the growing chain and the
metal.
1063,1064
This is rather different from metallocenes, where the u-agostic interaction is usually 45 kcal mol
1
more stable than the -agostic interaction.
141
O
ZrCl
2
O
SiMe
3
SiMe
3
78
O
ZrCl
2
O
SiMePh
2
SiMePh
2
O
ZrCl
2
O
SiPh
3
SiPh
3
80
O
TiBn
2
O
SiMe
3
SiMe
3
81
MeO
MeO
O
TiBn
2
O
Bu
t
Bu
t
82
79
O
TiBn
2
O
Bu
t
Bu
t
83
O
ZrBn
2
O
SiMePh
2
SiMePh
2
84
Another catalyst with coordination 4 at the metal atom and coordinating O atoms is based on complex 85. This
complex shows high activity in ethylene polymerization (1 10
2
3 10
2
g
PE
(mmol M)
1
h
1
). The tetrameric analog
exhibits moderate activity.
1065
N
O
O
O
Ti
Ph
Ph
Ph
Cl
85
4.09.5.1.2 Ligands with coordinating NN atoms
Pre-catalysts with coordination 4 at the metal and a coordinating NN ligand include complexes with chelating and
dianionic diamide ligands. Activation of these pre-catalysts leads to a class of systems with remarkable catalytic
properties and, not surprisingly, these systems have been quite deeply investigated.
275,10661084
The MAO- and
B(C
6
F
5
)
3
-activated diamide complexes 86 and 87, known as McConville catalysts, were introduced for 1-hexene
polymerization in 1996.
1066,1069
N
TiMe
2
N
86
N
TiMe
2
N
Pr
i
Pr
i
Pr
i
Pr
i
87
Olefin Polymerizations with Group IV Metal Catalysts 1087
At 68

C and with a MAO/Ti ratio of 250, these complexes yield poly-1-hexene with the remarkable activity of
about 3 10
5
g
PH
(mmol M)
1
h
1
. The polymers produced had moderate molecular masses (M

w
=30 00050 000)
with M

w
,M

n
close to 2, indicative of single-site behavior.
1066
Activation with [Ph
3
C][B(C
6
F
5
)
4
] salts also leads to
quite active catalysis (activity up to 10
5
g
PH
(mmol M)
1
h
1
), whereas activation with B(C
6
F
5
)
3
is much less
effective, although activity remains high (<2 10
3
g
PH
(mmol M)
1
h
1
). However, it is worth noting that polymer-
ization times were shorter than 1 min,
1067
and that extending the polymerization time to about 30 min resulted in only
moderate activity.
275
Solvent effects were shown to be relevant. Activity of 87/MAO in the presence of toluene
decreases to 8 10
2
g
PH
(mmol M)
1
h
1
at 22

C. It was hypothesized that toluene can compete with the monomer


for coordination to the metal atom.
1066
On the other hand, small amounts of CH
2
Cl
2
added to the aliphatic solvent
increased activity by a factor of 10. Probably, the higher polarity of the media helped to separate the borate counterion
from the active species.
275
In the case of MAO activation,
1
H and
13
C NMR spectra of oligomers indicated that the
only chain-termination mechanism is chain transfer to the Al co-catalysts, since no olefinic resonances were
observed.
275
This could rationalize the fact that B(C
6
F
5
)
3
and [Ph
3
C][B(C
6
F
5
)
4
] activation leads to living polymer-
ization of c-olefins such as 1-hexene, 1-octene, and 1-decene.
275,1067
In 10 min polymerization time, poly-1-hexene
with an M

n
of 164 000 and an M

w
,M

n
ratio of 1.07 was obtained.
275
The B(C
6
F
5
)
3
-activated 86 and 87 present
different stabilities. While the activated 86/B(C
6
F
5
)
3
species is stable for days, 87/B(C
6
F
5
)
3
deactivates as shown in
Scheme 35.
1068
In all cases, atactic polymers were obtained. NMR analysis of the resulting poly-1-hexenes indicated
that the regiochemistry of monomer insertion into the Tigrowing-chain bond is primary.
1067
McConville complexes 86 and 87 have also been used for propylene polymerization, and it has been shown
that the nature of the resulting PPs is highly dependent on the polymerization conditions.
10701072
While complex
86 only yields aPP, whatever AlR
3
species is used,
1072
catalysts based on 87/B(C
6
F
5
)
3
/AlR
3
yield aPP only with
AlMe
3
and AlEt
3
. Instead, with AlBu
i
3
(or R bulkier than Bu
i
), the polymers produced could be fractionated into
atactic and isotactic fractions.
1072
The weight percentage of the isotactic fraction increases with propylene
concentration from roughly 0% at 0.41 mol l
1
up to 45% at 2.76 mol l
1
.
1072
This suggested that two propylene
molecules could coordinate to the metal atom since double monomer coordination would increase steric bulkiness
and could induce the stereospecific behavior.
1072
Further support for this hypothesis was provided by propylene
polymerization in the presence of cyclohexene. NMR analysis indicated that cyclohexene was not inserted in the
polymers, but even at low propylene pressure (1 atm) the iPP fraction could increase from roughly 0%, with no
cyclohexene, up to 77% with a cyclohexene concentration of 8.8 mol l
1
. These findings added further support for
the isospecific active species depicted in Scheme 36.
1070
To date, the origin of stereoselectivity with these
catalysts is still unclear.
N
Zr
N
Pr
i
Pr
i
Pr
i
Pr
i
Me
-MeB(C
6
F
5
)
3
N
Zr
N
Pr
i
Pr
i
Pr
i
Pr
i
Me
Me
N
Zr
N
Pr
i
Pr
i
Pr
i
Pr
i
C
6
F
5
+B(C
6
F
5
)
3
B(C
6
F
5
)
2

+
CH
4
Scheme 35
N
Ti
N
Pr
i
Pr
i
Pr
i
Pr
i
P
N
Ti
N + +
Pr
i
Pr
i
Pr
i
Pr
i
P
Scheme 36
1088 Olefin Polymerizations with Group IV Metal Catalysts
NMR analysis of the iPP produced indicated that a rather good level of stereoregularity (% mmmm up to 80%) is
reached, and that enantiomorphic site control is operative. Insertion is essentially primary and regiomistakes were not
detected.
1071
Activity is in the range of 4 101 10
3
g
PP
(mmol M)
1
h
1
when the polymerization temperature
is varied between 0 and 150

C. In the same polymerization temperature range, the M

n
of the polymers is between
194 000 and 51 000, which indicates that even at 150

C reasonably high molecular masses are obtained. In all


cases, M

w
,M

n
was slightly higher than 2. The weight percentage of the iPP fraction increases from 8% at 0

C,
reaches a maximum of 79% at 40

C, and decreases to 36% at 150

C.
1072
Since the only chain-termination
mechanism operative with 86 and 87 is transfer to MAO, it was of interest to test AlMe
3
-free MAOs for
the polymerization of propylene. Indeed, at 0

C activation of 87 with dried MMAO leads to living propylene


polymerization.
1073,1074
Several modifications of the aromatic rings on the N atoms as well as of the C
3
spacer of the classical McConville
complex have been explored.
1075,1076
While the 89/[Ph
3
C][B(C
6
F
5
)
4
]/AlBu
i
3
catalyst yields PE with high activity,
3 10
2
g
PE
(mmol M)
1
h
1
bar
1
, the catalyst 88/MMAO gave only moderate activity (15 g
PE
(mmol M)
1
h
1
bar
1
at
room temperature).
1075
Similarly, poor activity was reported for 90/MAO, as well as for the Zr and Hf analogs.
1076
N
ZrCl
2
N
Si
Si
88
N
ZrMe
2
N
Pr
i
Si
Pr
i
Pr
i
Pr
i
Si
Pr
i
Pr
i
89
N
TiCl
2
N
Ph Ph
Bu
t
Ph
Ph
Bu
t
90
Catalysts with a rigid naphthalene bridge, such as those based on complexes 9294, have been also
explored.
1075,1077,1078
MMAO-activated 92 and 93 exhibited moderate ethylene polymerization activity
(1040 g
PE
(mmol M)
1
h
1
bar
1
),
1075,1077
while 94/MMAO in heptane showed a reasonably high activity
(~2 10
2
g
PE
(mmol M)
1
h
1
bar
1
at 60

C). The PE produced had moderate molecular masses (M

w
=121 000
214 000) with rather broad and even bimodal distributions.
1078
Also, in this case, using toluene as solvent depresses
activity, but it almost doubles the molecular masses.
1078
N
ZrCl
2
N
Si
Si
91
N
TiCl
2
N
Si
Si
92
N
ZrCl
2
N
Pr
i
Si
Pr
i
Pr
i
Pr
i
Si
Pr
i
Pr
i
93
N
TiCl
2
N
Bu
t Si
Bu
t
Si
94
Other variations on the NN bridge were reported.
10791085
Reducing the bridge to a C
2
spacer, 91, resulted in low
activity toward ethylene polymerization.
1085
The 95/MMAO catalyst exhibited moderate to low activity in ethylene,
propylene, and 1-hexene polymerization (1030 g
polymer
(mmol M)
1
h
1
).
1079
Catalysts based on 96 (13 g
PE
(mmol
M)
1
h
1
bar
1
and 70 gPP (mmol M)
1
h
1
bar
1
) and 97 also exhibited moderate activity in ethylene and propylene
polymerization.
10801082
Instead, the 98/MAO catalyst exhibited high activity in ethylene, propylene, and
1-hexene polymerization (5 10
3
, 8 10
2
and 2 10
2
g
Polymer
(mmol M)
1
h
1
, respectively). Unfortunately, mole-
cular masses were rather low (M

w
= 161000. 42 000. and 26 000, respectively).
1083
Activation of the same complexes
with B(C
6
F
5
)
3
resulted in a remarkable decrease of activity for ethylene, while propylene and 1-hexene
were polymerized with similar activity. However, molecular masses were notably higher
(M

w
= 165000. 305000 and 145 000, respectively).
1083
Olefin Polymerizations with Group IV Metal Catalysts 1089
N
TiCl
2
N
Si
Si
95
N
ZrBn
2
N
Ph
Ph
96
N
ZrMe
2
N
Pr
i
Pr
i
Pr
i
Pr
i
B
B
Me
2
N
Me
2
N
97
N
ZrMe
2
N
Pr
i Si
Pr
i
Pr
i
Pr
i
Si
Pr
i
Pr
i
98
The aminopyrrolate complex 99 is another case of a complex with coordination 4 and two coordinating N atoms.
In the temperature range 050

C, the 99/MAO catalyst oligomerizes ethylene (M

w
<8 000, M

w
/M

n
<2.8) with
high activity (<8 10
2
g
PE
(mmol M)
1
h
1
bar
1
), whereas the monoligated imino-pyrrolate complex 100 shows
moderate polymerization activity (19 g
PE
(mmol M)
1
h
1
bar
1
).
1086
99 100
N
N r( ) CH
2
Ph
2
PhCH
2
Z
N
N r(CH
2
Ph)
3
Z
Another class of catalysts with two coordinating N atoms is represented by the bis-phosphinimide Ti catalysts.
10871091
Complex 101 activated with either B(C
6
F
5
)
3
and [Ph
3
C][B(C
6
F
5
)
4
] is an highly active catalyst for ethylene
polymerizations. At 25

C and 1 bar ethylene, activity is about 1 10


3
g
PE
(mmol M)
1
h
1
bar
1
. At the much
more drastic conditions of about 100 bar ethylene and 160

C, 101 still shows the interesting activity of about


6 10
2
g
PE
(mmol M)
1
h
1
bar
1
. Considering the high polymerization temperature, the molecular masses
obtained (M

w
= 80000 and M

w
/M

n
= 1.9) are reasonably high, while their narrow distribution indicates that
single-site catalysis is achieved. Similar Zr-based catalysts showed rather lower activity.
1088,1089
In contrast, catalysts
based on bidentate phosphinimide complexes such as 102 resulted in disappointingly lower activity.
1091
The reaction
of 101 with AlMe
3
was investigated in detail, and was shown to proceed via competitive CH activation and
metathesis reactions.
1090
N
Ti
Me
Me
Bu
t
P
Bu
t
Bu
t
N
Bu
t
P
Bu
t
Bu
t
101
N
Ti
Me
Me
P
Bu
t
Bu
t
N
P
Bu
t
Bu
t
102
4.09.5.1.3 Other ligands
Other complexes with coordination 4 at the metal atom that have been tested in olefin polymerizations are the
monodicarbollide systems of general formula (j
5
-C
2
B
9
H
11
)(C
5
Me
5
)M(CH
3
) (M=Zr, Hf) or (j
5
-C
2
B
9
H
11
)-
M(NEt
2
)(NHEt
2
), (M=Ti, Zr). These catalyst systems were proved to be active in ethylene polymerization
with a moderate activity comparable to that of CpTiCl
3
.
34,1092
Other systems investigated are ZrCl
4
esters or
1090 Olefin Polymerizations with Group IV Metal Catalysts
ethers that were proved to oligomerize and polymerize ethylene and propylene,
1093,1094
and TiCl
2
(2-OC
6
H
4
OCH
3
)
2
which was shown to polymerize propylene to an elastomeric polymer.
1095
4.09.5.2 Complexes with Coordination Number 5
4.09.5.2.1 Ligands with coordinating OO atoms
Dialkoxides with an extra donor atom have been also tested as ethylene, propylene, and 1-hexene polymerization
catalysts.
159,10961098
The first reports showed that an additional sulfur donor results in quite high ethylene and
propylene polymerization activity (about 10
3
g
PE
(mmol M)
1
h
1
bar
1
), and very high molecular masses.
159,1096
Catalysts based on 103, known as the Kakugos catalysts, yield a regioregular but atactic poly-1-hexene of rather
low molecular mass (M

w
= 10
4
).
256
Theoretical DFT studies showed that the S atom donates electron density to the
metal, which reduces the activation barrier relative to the complexes with a CH
2
bridge such as the simple dialkoxide
83.
1063,1064
A modified version of the Kakugo catalyst, based on complex 104, resulted in rather poor activity
(1 g
PE
(mmol M)
1
h
1
bar
1
).
1099
The MAO-activated complexes 105107 showed low activity in either ethylene
or c-olefin polymerization.
1097,1098
Nevertheless, while the meso-complex 106 yields atactic poly-1-hexene of low
molecular mass (M

n
= 10
3
), the racemic analog 107 yields isotactic poly-1-hexene of reasonably high molecular mass
(M

n
up to 300 000).
1098
The catalyst 108/MAO was shown to yield polyethylene with very high activity
(10
3
g
PE
(mmol M)
1
h
1
bar
1
) in the polymerization temperature range 065

C, whereas its Ti analog exhibited


disappointingly low activity (4 gPE (mmol M)
1
h
1
bar
1
).
1100
S
O
TiCl
2
O
Bu
t
Bu
t
103
S
O
ZrCl
2
THF
O
Bu
t
Bu
t
Bu
t
Bu
t
104
N
O
O
TiCl
2
Ph
Ph
Ph Ph
105
N
O
O
TiCl
2
Ph
Ph
Ph
106
N
O
O
TiCl
2
Ph
Ph
Ph
107
N
O
O
Zr
Bu
t
Bu
t
Cl
THF
Cl
108
4.09.5.2.2 Ligands with coordinating NO atoms
Phenoxyamide complexes with an extra neutral donor, such as the monometallic complexes with trichelate ligands
of Figure 39, can be effective ethylene polymerization catalysts.
1101
The best performances were obtained with the
catalyst based on the complex with a diphenylphosphine group. Besides the highest activity in the series, this
complex is thermally robust and shows optimum activity at 70

C.
In the polymerization temperature range 2570

C, the catalyst based on the dimeric 109 activated with MAO


exhibits moderate to high activity in ethylene polymerization (activity =1010
3
g
PE
(mmol M)
1
h
1
bar
1
). The PE
produced has reasonably high molecular masses (M

w
= 680 000 at 25

C and 120 000 at 70

C). This system was not


active in propylene polymerization.
1102
Olefin Polymerizations with Group IV Metal Catalysts 1091
O
Ti
O
Ti
N
N
NMe
2
NMe
2
Me
2
N
Me
2
N
Bu
t
Bu
t
109
4.09.5.2.3 Ligands with coordinating NN atoms
Complexes with dianionic diamide ligands show remarkable catalytic properties also in the presence of an extra
neutral donor, and thus tridentate diamide systems have been extensively investigated.
311,312,11031114
The geometry
of coordination at the metal atom is trigonal bipyramidal, with the two amide N atoms in the equatorial plane.
Initiation consists in the reaction of a tridentate dialkyl complex with B(C
6
F
5
)
3
or a borate salt such as
[Ph
3
C][B(C
6
F
5
)
4
] or [PhNMe
2
H][B(C
6
F
5
)
4
]. These catalysts, known as Schrocks catalysts, usually require low
polymerization temperatures (around 0

C) to avoid catalyst decomposition. Performances strongly depend on the


nature of the diamide ligand, on the nature of the metal, and on the activation procedure. At 0

C/1 h, the catalyst 110/


[PhNMe
2
H][B(C
6
F
5
)
4
] polymerizes 1-hexene in a living fashion.
1112,1114
The resulting poly(1-hexene)s have
M

n
~ 2500035 000 and M

w
,M

n
~ 1.021.05. Raising the polymerization temperature to 22

C yields poly(1-
hexene) with an activity of 2 10
2
gPH(mmol M)
1
h
1
, and M

n
= 45000, but M

w
,M

n
rises to 1.2.
1114
The
regiochemistry of 1-hexene insertion is primary.
1115
The Hf analog of 110 also polymerizes 1-hexene in a
living fashion, while the Ti analog does not exhibit catalytic activity.
1112
The kinetics of these systems were
investigated in detail.
311
At 22

C, 110/B(C
6
F
5
)
3
yields a PE with the high activity of 1 10
2
gPE (mmol
M)
1
h
1
bar
1
, and changing the initiator to a borate salt, 110/[PhNMe
2
H][B(C
6
F
5
)
4
], increases the activity to
8 10
2
gPE (mmol M)
1
h
1
bar
1
.
1114
O
N
N
Zr
Me
Me
Bu
t
Bu
t
110
Complexes 111114 represent other examples of Zr catalysts that contain a diamido ligand with an extra O-donor
group. After activation with [Ph
3
C][B(C
6
F
5
)
4
], these systems polymerize 1-hexene at 0

C.
1111,1113
Interestingly,
structures 111113 exhibit substantially similar activity, although the molecular mass decreases sensibly with the
bulkiness of the substituents ortho to the amido N atoms 111: M

w
= 18 000, M

w
,M

n
= 1.6; 112: M

w
= 11 000,
1 bar ethylene, MAO/Ti = 2,000, heptane, 25 C, 30 min
O
N
TiBn
2
Bu
t
Bu
t
9 10
0.44
306,500
3.7
1 10
2
0.48
617,500
3.5
4 10
3
3.53
594,700
3.6
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
w
M
w
/M
n
2 10
4
1.95
1,803,000
3.8
L
N SPh N OPh N N PPh
2
N
N L
Figure 39 Ethylene polymerization with mono-phenoxyamide-based catalysts.
1092 Olefin Polymerizations with Group IV Metal Catalysts
M

w
,M

n
= 1.2; 113: M

w
= 400, M

w
,M

n
= 1.7). The values of M

w
,M

n
are rather narrow, but living polymerization
was not reported.
1113
Similar behavior was observed for the Hf analogs. Introduction of the THF ring in 114 results in
an increase of the molecular mass of the polymers and in narrower molecular mass distribution M

n
~ 3500045 000
and M

w
,M

n
~1.11.5). It was suggested that the THF ring introduces rigidity and steric hindrance, which prevents
u-H elimination reaction.
1111
Me
N
Me
Zr
N
O
111
Me
N
Me
Zr
N
Et
Et
Et
Et
O
112
Me
N
Me
Zr
N
Pr
i
Pr
i
Pr
i
Pr
i
O
113
Me
N
Me
Zr
N
O
114
Complexes 115121 represent examples of diamido ligands with an extra N-donor group.
1103,11051108,1110
Both
115 and 116 exhibit reasonable activity, but molecular masses obtained with 115 (M

w
~ 20000, M

w
,M

n
~ 1.3) are
approximately the half of those obtained with 116 (M

w
~ 40 000, M

w
,M

n
~ 1.4). These catalysts also exhibited
good stability at room temperature, while increasing the temperature to about 65

C resulted in rapid irreversible


catalyst decomposition.
1110
One of the deactivation paths is a CH activation reaction that involves an Me group of
the ligand, as shown in Scheme 37.
312,1107
A clever and simple solution to catalyst deactivation was the synthesis of
117, which at 0

C and in 1 h yields living 1-hexene polymerization (M

w
,M

n
~ 1.02) and M

w
up to 80 000.
1103,1107
The Hf analog of 117, instead, is prone to significant u-H elimination which implies low molecular masses.
1103
Me
N
NH
Me
Zr
N
115
Me
N
N
Me
Zr
N
116
Me
N
N
Me
Zr
N
Cl
Cl
Cl
Cl
117
N
N
N
Zr
Me
Me
Me
3
Si
Me
3
Si
SiMe
3
118
Me
N
N
Me
Zr
N
Cl
Cl
Cl
Cl
120
Me
N
N
Me
Zr
N
119 121
Bu
i
N
N
Bu
i
Zr
N
The B(C
6
F
5
)
3
-activated complex 118 is a moderately active ethylene polymerization catalyst, although it suffers
from degradation due to a CH activation of an SiMe
3
group adjacent to the metal center.
1116
This degradation

Me
N
N
Zr
N
+
N
N
Zr
N
+
CH
4
Scheme 37
Olefin Polymerizations with Group IV Metal Catalysts 1093
mechanism is extremely similar to that observed in related tridentate diamide complexes (see Scheme 37).
1103
Another variation in tridentate complexes with a diamido ligand is represented by 119 and 120.
1106,1108,1109
Differently from the case of 116 and 117, the complex 120 with halogen atoms ortho to the N amido atoms shows
lower activity and yields polymers with lower molecular mass than the catalyst based on 119.
1106
Comparison of the
pyridyldiamido complex 119 with 121, in which the two Me groups were replaced by the two i-Bu groups, allowed
investigation of the role of the alkyl group bound to the metal. The poorer catalytic behavior of 119 was ascribed to
the formation of largely inactive cationic dimeric species through Me bridges.
1105,1108,1109
Finally, in the attempt to
achieve stereospecific polymerization, asymmetric complexes such as 122 were synthesized. Unfortunately, these
complexes also only yield atactic poly (1-hexene).
1104
Me
N
Me
Zr
N
Pr
i
Pr
i
O
122
u-Diketiminato zirconium complexes such as 123 and 124 exhibited moderate activity in the polymerization of
ethylene (about 50 gPE (mmol M)
1
h
1
bar
1
), but molecular masses were disappointingly low (M

n
~10 000).
1117,1118
Introduction of a CF
3
group as in 125 resulted in very high activity (2 10
3
gPE (mmol
M)
1
h
1
bar
1
) but still low molecular masses (M

n
= 80 000).
1117
N
ZrCl
3
N
Pr
i
Pr
i
Pr
i
Pr
i
123
N
ZrCl
3
N
124
N
ZrCl
3
N
CF
3
CF
3
125
A different class of ligands with coordination 5 and two N atoms bound to the metal is represented by the Hf
pyridyl amine complexes 126128 discovered in 2003. These systems are characterized by an HfC o-bond between
the aromatic ligand and the metal. The original patent concerns the preparation of iPP co-polymers with ethylene or
other unsaturated co-monomers.
1119
Nevertheless, a catalyst solution composed of complex 127/a borate salt/MAO in
the ratio 1 : 1 : 200 (Al) is able to homopolymerize propylene to an isotactic polymer at 110

C for at least 10 min. The


catalyst exhibited very high activity, around 5000 gPP (mmol M)
1
h
1
bar
1
. The iPP produced had a T
m
=144.4

C
and high molecular mass, M

w
= 293000. The M

w
,M

n
= 2.9 indicated that under these rather drastic conditions the
catalyst shows single-site behavior. The production of a highly isotactic PP with a C
1
-symmetric catalyst is particularly
remarkable since it is obtained at a high polymerization temperature.
1119
Me
Hf
N
Me
N
Pr
i
Pr
i
H
126
Me
Hf
N
Me
N
Pr
i
Pr
i
H
127
Me
Hf
N
Me
N
Pr
i
Pr
i
H
128
1094 Olefin Polymerizations with Group IV Metal Catalysts
4.09.5.2.4 Other ligands
Complex 129 activated with [Ph
3
C][B(C
6
F
5
)
4
] is a catalyst for aspecific propylene polymerization. The system
showed low activity (7 g
PP
(mmol M)
1
h
1
bar
1
) and yielded polymers of rather low molecular masses
(M

n
= 21000).
1120
Zr
N
Me
Bn Bn
129
4.09.5.3 Complexes with Coordination Number 6
4.09.5.3.1 Ligands with coordinating OO atoms
The phenoxyether complexes of Figure 40 are active ethylene polymerization catalysts.
1121
MAO- and
[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al-activated complexes show rather similar behavior. This is different from phenoxyimine-
based complexes, where Bu
i
3
Al is accepted to react with the imine functionality. In all cases, the molecular masses of
the PEs produced are very high (M

v
2 10
6
). Activity and M

v
strongly depend on the size of the substituent R.
Activity is poor with the small R=H substituent, while it is high with the bulky R=adamantyl. Bulky substituents
probably prevent formation of a tightly coupled ion pair. Surprisingly, higher molecular masses are obtained with the
smaller R=H substituent. This was rather unexpected, since steric congestion at the active center usually reduces
chain-release reactions.
Similar complexes were investigated in less detail by other authors, although rather good performances were
observed. Complex 130 is a highly active ethylene polymerization catalyst that yields very broad molecular mass
distribution (M

w
,M

n
= 16). The Zr analog of 130 behaves similarly.
1122
Other complexes tested in ethylene
polymerization are 131 and 132. While the AlEt
3
/AlEt
2
Cl-activated 131 shows high polymerization activity (about
2 10
2
g
PE
(mmol M)
1
h
1
bar
1
), the similarly activated 132 performs even better (about 3 10
3
g
PE
(mmol
M)
1
h
1
bar
1
).
1123
O
O iCl
2
2
R
H n-Bu Adamantyl
4.10
0.11
7,240,000
310
2
0.11
5,420,000
310
3
1.17
2,040,000
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
v
25 C, 1 bar ethylene, MAO/Ti = 250, [Ph
3
C][B(C
6
F
5
)
4
]/ Ti = 1.2, Bu
i
3
Al /Ti = 50,
toluene, polymerization time: R = H, 30 min; R = n-Bu and adamantyl, 5 min
6.10
0.14
5,440,000
110
3
0.61
3,120,000
310
3
1.45
2,600,000
Yield (g)
M
v
[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al as co-catalyst
MAO as co-catalyst
R
T
Figure 40 Ethylene polymerization with phenoxyether-based catalysts.
Olefin Polymerizations with Group IV Metal Catalysts 1095
130
O
O iCl
2
Bu
t
2
T
131
O
O iCl
2
2
Me
T
132
O
O i(OEt)
2
2
O
T
The acetylacetonate 133/MAO catalyst, instead, yields PP with low activity (310 g
PP
(mmol M)
1
h
1
bar
1
) and
moderate molecular mass (M

n
up to 125 000). The resulting PP is slightly isotactic (mmmm 20%) and exhibits
elastomeric properties.
1124
Finally, AlEt
3
/AlEt
2
Cl-activated dimeric Ti complexes such as 134 yield PE with very
high activity (about 10
5
g
PE
(mmol M)
1
h
1
bar
1
), and with high molecular masses.
1125
O
TiCl
2
O
2
133
Me
O
S
Bu
t
Ti
O
Bu
t
Me
O
S
Bu
t
Ti
O
Bu
t
O
O
Et
Et
Bu
t
Bu
t
Bu
t
Bu
t
134
4.09.5.3.2 Ligands with coordinating NO atoms: phenoxyimine-catalysts for polyethylene
Catalysts based on salicylaldiminato ligands (commonly referred to as phenoxyimine ligands) were introduced for
ethylene polymerization in 1998.
1126
This remarkable class of highly active catalysts can lead to an unique variety of
linear PEs.
1121,11271144
Several reviews have been dedicated to these systems.
1140,11451149
The performances of
some of these bis(phenoxyimine) systems are reported in Table 18. The most typical and most widely investigated
systems in this class are complexes 135138.
135
O
N rCl
2
2
Z
136
O
N iCl
2
2
T
137
O
N iCl
2
2
F
F
F
F
F
T
138
O
N iCl
2
2
F
F
F
F
F
T
The MAO-activated prototype system (135) exhibits an activity higher than the Cp
2
ZrCl
2
systems under the same
conditions.
11281130
Several modifications of the basic prototype system 135 have been tested in ethylene polymer-
ization. This class of catalysts represents by far the most widely investigated post-metallocene systems, and these
modifications are centered on positions R
1
, R
2
, and R
3
of 139.
O
N rCl
2
R
1
R
2
2
R
3
139
Z
1096 Olefin Polymerizations with Group IV Metal Catalysts
Table 18 Ethylene polymerization with bis(phenoxyimine)-based catalysts
Entry Pre-catalyst Co-catalyst
Co-catalyst
M
T
P
a
(

C)
P
E
b
(bar)
(Solvent)
t
P
c
(min)
Productivity
(g
PE
(mmol
M)
1
h
1
bar
1
)
Yield
(g) M

w
/M

n
M

v
References
1 MAO 250 25 1 (Tol) 10 3 10
3
2.8 516,000 1127
2 MAO 250 50 1 (Tol) 10 4 10
3
3.3 546,000 1127
3 MAO 250 75 1 (Tol) 10 4 10
3
3.1 440,000 1127
4 MAO 250 25 1 (Tol) 5 3 10
3
1.4 510,000 1127
5 MAO 6,250 25 1 (Tol) 5 4 10
3
0.1 326,000 1133
6
O
N
TiCl
2
Bu
t
2
MAO 1,250 75 9 (nC7) 15 5 10
3
12.0 664,000 1127
7 [Ph
3
C][B(C
6
F
5
)
4
]/
Bu
i
3
Al
2/30 25 1 (Tol) 5 2 10
2
0.1 4,810,000 1135
8
[Ph
3
C][B(C
6
F
5
)
4
]/
Bu
i
3
Al
2/30 50 1 (Tol) 5 4 10
2
0.2 5,860,000 1135
9 [Ph
3
C][B(C
6
F
5
)
4
]/
Bu
i
3
Al
2/30 75 1 (Tol) 5 7 10
2
0.3 3,920,000 1135
10 MAO 2,500 50 9 (Tol) 30 4 10
3
11.2 2.4 464,000 1143
11 MgCl
2
/Bu
i
m
Al(OR)
3m
a
800/4,800 50 9 (Tol) 30 4 10
3
9.1 2.7 509,000 1143
12 DMAO 250 25 1 (Tol) 30 3 10
3
8.1 2.6
M

w
1,281,000
1138
13
O
N
TiCl
2
Ph
2
MAO 250 25 1 (Tol) 5 5 10
3
1.9 604,000 1127
14
O
N
TiCl
2
SiPh
2
Me
2
MAO 250 25 1 (Tol) 10 3 10
3
2.1 375,000 1127
(Continued)
Table 18 (Continued)
Entry Pre-catalyst Co-catalyst
Co-catalyst
M
T
P
a
(

C)
P
E
b
(bar)
(Solvent)
t
P
c
(min)
Productivity
(g
PE
(mmol
M)
1
h
1
bar
1
)
Yield
(g) M

w
/M

n
M

v
References
15
O
N
TiCl
2
2
DMAO 250 25 1 (Tol) 30 4 10
1
0.1 3.1 M

w
66,000 1138
16
O
N
TiCl
2
Me
2
DMAO 250 25 1 (Tol) 30 3 10
2
0.6 1.5 M

w
402,000 1138
17
O
N
TiCl
2
SiMe
3
2
DMAO 250 25 1 (Tol) 30 3 10
3
6.7 2.5
M

w
1,105,000
1138
18 MAO 6,250 25 1 (Tol) 5 2 10
5
3.3 2.2 8,000 1129
19 MAO 15,625 25 1 (Tol) 5 5 10
5
0.9 2.1 10,000 1128
20 MAO 62,500 25 1 (Tol) 5 6 10
5
0.9 2.1 9,000 1129
21 MAO 125,000 25 1 (Tol) 5 5 10
5
0.9 2.0 7,000 1129
22 MAO 312,500 25 1 (Tol) 5 5 10
5
0.8 2.0 7,000 1129
23 [Ph
3
C][B(C
6
F
5
)
4
]/
Bu
i
3
Al
2/25 50 1 (Tol) 5 1 10
4
2.3 5,050,000 1128
24
O
N
ZrCl
2
Bu
t
2
[Ph
3
C][B(C
6
F
5
)
4
]/
Bu
i
3
Al
2/40 25 1 (Tol) 5 4 10
3
0.8 3,830,000 1129
25
[Ph
3
C][B(C
6
F
5
)
4
]/
Bu
i
3
Al
2/40 50 1 (Tol) 5 1 10
4
2.3 5,050,000 1129
26 MAO 78,125 50 9 (Tol)h 30 1 10
5
10.3 74,000 1143
27 MgCl
2
/
Bu
i
m
Al(OR)
3m
d
12,500/
50,000
50 9 (Tol)h 30 2 10
4
2.0 91,000 1143
28 MAO 25,000 50 9 (nC7) 15 1 10
5
14.9 47,000 1130
29 MAO 25,000 75 9 (nC7) 15 2 10
4
2.6 100,000 1130
30 MAO 62,500 75 1 (Tol) 5 1 10
5
2.4
M

w
8,100
1144
31 MAO 12,500 25 1 (Tol) 5 2 10
5
1.9
M

w
7,300
1139
32
O
N
ZrCl
2
n-C
6
H
13
Bu
t
2
MAO 6,250 50 9 (nC7) 15 3 10
4
17.3 11,000 1130
33
O
N
ZrCl
2
Bu
t
2
MAO 12,500 50 9 (nC7) 15 8 10
4
18.8 15,000 1130
34 MAO 6,250 25 1 (Tol) 5 1 10
5
1.9 1.8 M

w
13,800 1152
35
O
N
ZrCl
2
Bu
t
2
MeO
MAO 12,500 50 9 (nC7) 15 3 10
4
8.2 363,000 1130
36 MAO 62,500 75 1 (Tol) 5 8 10
4
11.1 M

w
28,600 1144
(Continued)
Table 18 (Continued)
Entry Pre-catalyst Co-catalyst
Co-catalyst
M
T
P
a
(

C)
P
E
b
(bar)
(Solvent)
t
P
c
(min)
Productivity
(g
PE
(mmol
M)
1
h
1
bar
1
)
Yield
(g) M

w
/M

n
M

v
References
37
O
N
ZrCl
2
n-C
6
H
13
Bu
t
2
MeO
MAO 25,000 50 9 (nC7) 15 5 10
4
5.7 99,000 1130
38
O
N
ZrCl
2
Bu
t
2
MeO
MAO 25,000 50 9 (nC7) 15 9 10
4
10.2 168,000 1130
39 MAO 25,000 75 9 (nC7) 15 1 10
5
18.3 54,000 1130
40
O
N
ZrCl
2
2
MeO
MAO 6,250 75 9 (nC7) 15 3 10
5
15.3 95,000 1130
41 O
N
ZrCl
2
2
Ph
MeO
n-C
6
H
13
MAO 12,500 75 9 (nC7) 15 7 10
5
17.7 39,000 1130
42
O
N
ZrCl
2
2
Ph
MeO
MAO 250,000 75 9 (nC7) 15 7 10
5
9.1 104,000 1130
43
Me
O
N
ZrCl
2
Bu
t
2
MAO 62,500 25 1 (Tol) 5 3 10
5
0.5 2.0 7,000 1129
44 MAO 62,500 75 1 (Tol) 5 7 10
4
4.1 M

w
19,100 1144
45
O
N
ZrCl
2
Me
2
MAO 250 25 1 (Tol) 5 4 10
2
0.2 2.3 3,000 1129
46
O
N
ZrCl
2
Pr
i
2
MAO 250 25 1 (Tol) 5 9 10
2
0.4 2.5 6,000 1129
(Continued)
Table 18 (Continued)
Entry Pre-catalyst Co-catalyst
Co-catalyst
M
T
P
a
(

C)
P
E
b
(bar)
(Solvent)
t
P
c
(min)
Productivity
(g
PE
(mmol
M)
1
h
1
bar
1
)
Yield
(g) M

w
/M

n
M

v
References
47
O
N
ZrCl
2
2
Me
MAO 62,500 25 1 (Tol) 5 7 10
5
1.2 2.7 12,000 1129
48
O
N
ZrCl
2
2
Ph
Me
MAO 125,000 25 1 (Tol) 5 2 10
6
1.7 7.2 18,000 1129
49
O
N
ZrCl
2
2
Ph
Me
MAO 250,000 25 1 (Tol) 5 4 10
6
1.8 1.9 15,000 1129
50
O
N
ZrCl
2
Bu
t
2
MAO 2,500 25 1 (Tol) 5 4 10
4
1.7 2.1 320,000 1129
51
O
N
ZrCl
2
Bu
t
2
Bu
t
MAO 250 25 1 (Tol) 30 1 10
2
0.2 2,740,000 1129
52
O
N
ZrCl
2
Bu
t
2
Bu
t
Bu
t
MAO 12,500 25 1 (Tol) 5 2 10
5
2.0 1.8 26,000 1129
53
O
N
ZrCl
2
Bu
t
2
Bu
t
MAO 12,500 25 1 (Tol) 5 3 10
5
2.3 2.0 7,000 1129
54
O
N
ZrCl
2
Bu
t
2
MAO 6,250 25 1 (Tol) 5 4 10
4
0.6 1.8 M

w
4,700 1152
(Continued)
Table 18 (Continued)
Entry Pre-catalyst Co-catalyst
Co-catalyst
M
T
P
a
(

C)
P
E
b
(bar)
(Solvent)
t
P
c
(min)
Productivity
(g
PE
(mmol
M)
1
h
1
bar
1
)
Yield
(g) M

w
/M

n
M

v
References
55
O
N
ZrCl
2
Bu
t
2
MAO 6,250 25 1 (Tol) 5 3 10
4
0.5 1.5 M

w
2,100 1152
56
O
N
ZrCl
2
Bu
t
2
MAO 6,250 25 1 (Tol) 5 1 10
5
1.9 1.8 M

w
3,800 1152
57
O
N
ZrCl
2
2
Ph Ph
MAO 62,500 40 1 (Tol) 5 6 10
5
24.8 133,200 1142
58 MAO 62,500 40 1 (Tol) 8 4 10
5
43 231,800 1142
59
O
N
TiCl
2
Bu
t
2
F MAO 6,250 25 1 (Tol) 5 4 10
3
0.1 419,000 1133
60 MAO 3,125 50 1 (Tol) 5 5 10
3
0.2 2.2 M

n
128,000 1134
61
O
N
TiCl
2
Bu
t
2
F F
MAO 6,250 25 1 (Tol) 5 3 10
4
0.6 623,000 1133
62 MAO 2,500 50 1 (Tol) 1 3 10
4
0.3 1.8 M

n
129,000 1134
63
MAO 6,250 25 1 (Tol) 5 4 10
4
0.7 378,000 1133
64
O
N
TiCl
2
Bu
t
2
F F
F
MAO 2,500 50 1 (Tol) 1 4 10
4
0.4 2.0 M

n
98,000 1134
65
O
N
TiCl
2
Bu
t
2
CF
3
MAO 6,250 25 1 (Tol) 5 3 10
3
0.1 542,000 1133
(Continued)
Table 18 (Continued)
Entry Pre-catalyst Co-catalyst
Co-catalyst
M
T
P
a
(

C)
P
E
b
(bar)
(Solvent)
t
P
c
(min)
Productivity
(g
PE
(mmol
M)
1
h
1
bar
1
)
Yield
(g) M

w
/M

n
M

v
References
66
O
N
TiCl
2
Bu
t
2
CF
3
F
3
C
MAO 6,250 25 1 (Tol) 5 4 10
4
0.7 1,365,000 1133
67 MAO 2,500 50 1 (Tol) 1 4 10
4
0.3 1.13 M

n
424,000 1134
68
O
N
TiCl
2
Bu
t
2
F F
F
F
F
MAO 1,250 75 1 (Tol) 1 3 10
4
0.5 1.15 M

n
272,000 1140
69 MAO 625 90 1 (Tol) 1 1 10
4
0.5 1.3
M

n
167,000
1140
70
O
N
TiCl
2
Bu
t
2
F
F
F
MAO 3,125 50 1 (Tol) 5 2 10
3
0.1 1.3 M

n
145,000 1134
71
O
N
TiCl
2
Bu
t
2
F
F
MAO 1,250 50 1 (Tol) 5 8 10
2
0.1 1.05 M

n
64,000 1134
72
O
N
TiCl
2
Bu
t
2
F
MAO 250 50 1 (Tol) 5 1 10
2
0.1 1.06 M

n
13,000 1134
73
O
N
ZrCl
2
Bu
t
2
F F
F
F
F
MAO 12,500 25 1 (Tol) 5 3 10
5
2.8 1.9 M

w
157,200 1139
(Continued)
Table 18 (Continued)
Entry Pre-catalyst Co-catalyst
Co-catalyst
M
T
P
a
(

C)
P
E
b
(bar)
(Solvent)
t
P
c
(min)
Productivity
(g
PE
(mmol
M)
1
h
1
bar
1
)
Yield
(g) M

w
/M

n
M

v
References
74
O
N
HfCl
2
Bu
t
2
F F
F
F
F
MAO 2,500 25 1 (Tol) 5 3 10
4
1.2 2.7 M

w
409,600 1139
75
O
N
HfCl
2
Bu
t
2
MAO 2,500 25 1 (Tol) 5 2 10
4
0.9 M

w
16,800 1139
a
T
P
=polymerization temperature.
b
P
E
=ethylene pressure.
c
t
P
=polymerization time.
d
R=2-ethyl-hexyl.
Stereochemistry of the complex. Bis(phenoxyimine) and related systems are characterized by an octahedral geometry
at the metal atom. The five geometrical isomers of Figure 41 are possible for octahedral complexes with two
unbridged bidentate ligands. Only isomers AC are potentially active in polymerization because they present
the X groups in mutually cis-positions. The tendency of bis(phenoxyimine)-based catalysts to assume a geometry
with a cis-N, trans-O, and cis-X disposition of the coordinating atoms (isomer A in Figure 41) is proved by the
X-ray structures of several pre-catalysts,
11271129,1132,1134,1135,1145
as well as by quantum mechanics
calculations.
1127,1129,1135,1145
Nevertheless, an equilibrium between isomers AC was invoked to explain the multi-
modal molecular mass distribution observed in some cases.
1142,1144
Activity: ligand-substitution effects. Steric and electronic properties of the phenoxyimine ligands can be tuned to
regulate the catalytic behavior of the corresponding catalyst. Regarding steric effects, bulky R
2
groups ortho to the
phenoxy O atom usually enhance activity (see Figure 42).
1129
It was suggested that bulky R
2
groups enhance catalytic
activity due to a more effective separation between the cationic catalyst and the anionic counterion.
1129
A weakly bound
ion pair results in an easier coordination of ethylene
1085
to a more electrophilic (and thus reactive) metal atom.
1150
Conversely, the size of the R
1
group has less influence on the activity of the catalyst (see Figure 43). Nevertheless,
in the case of R
1
=aromatic ring, bulky groups in position 2 of the phenyl ring depress activity (see
Figure 43).
1129,1151
M
O
O
X
N
N
M
N
N X X X
X
O
O
M
N
O O O
N
X
X
M
N
O
X
N
O
M
O
N
N
X
X
A
cis-N
trans-O
cis-X
B
trans-N
cis-O
cis-X
C
cis-N
cis-O
cis-X
D
cis-N
cis-O
trans-X
E
trans-N
trans-O
trans-X
Figure 41 Octahedral-based geometric isomers for the bis(phenoxyimine) pre-catalysts.
Ethylene polymerization activities: effect of the R
2
substituent
O
N rCl
2
Ph
R
2
2
4 10
2
9 10
2
2 10
5
Activity
(g
PE
(mmol M)
1
h
1
bar
1
)
Me Pr
i
Bu
t
O
N rCl
2
Ph
R
2
Bu
t
R
2
R
2
2
3 10
5
7 10
5
2 10
6
Z Z
Figure 42 Ethylene polymerization activity (g
PE
(mmol M
1
) h
1
bar
1
) for a series of bis(phenoxyimine) Zr catalysts with
different R
2
substituents. Left to right, the systems shown correspond to entries 45, 46, 18, 43, 47, and 48 in Table 18.
Ethylene polymerization activities: effect of the R
1
substituent
O
N rCl
2
R
1
Bu
t
n-C
6
H
13
R
1
2
4 10
4
8 10
4
1 10
5
1 10
2
4 10
4
2 10
5
Activity
(g
PE
(mmol M)
1
h
1
bar
1
)
Ethylene 1 bar, 25 C Ethylene 9 bar, 50 C
Z
Figure 43 Ethylene polymerization activity (g
PE
(mmol M
1
) h
1
bar
1
) for a series of bis(phenoxyimine) Zr catalysts with
different R
1
substituents. Left to right, the systems shown correspond to entries 32, 33, 28, 51, 50, and 18 in Table 18.
Olefin Polymerizations with Group IV Metal Catalysts 1109
Electronic effects have also been shown to influence catalytic properties. It has been suggested that electron-
donating groups on the phenoxyimine ligands can reduce decomposition of the active species and thus improve
overall productivity. Electron-donating groups improve the stability of the active species because they can strengthen
the metalligand bonds. This effect is particularly important at high temperature due to the low thermal stability of
the active species.
1130
Catalysts with bis(phenoxyimine) ligands bearing the OMe group in position R
3
usually have
similar activity relative to their unsubstituted analogs (see Figure 44(a) and 44(b)). Instead, electron-withdrawing
groups as F or CF
3
on the phenyl group in the R
2
position have a profound influence on the catalyst activity.
1133,1134
This substitution is particularly effective when the F or CF
3
groups are in the meta-positions (see Figure 44(c)).
Moreover, it was hypothesized that electron-withdrawing groups generate a more electrophilic metal center, which
results in increased reactivity.
1133
Molecular mass: ligand-substitution effects. Steric and electronic effects can also be used to control the molecular mass
of the PEs produced. Bis(phenoxyimine)-based catalysts can yield polymers ranging from very low to extremely high
molecular mass (M

w
<10
4
or M

w
>10
6
).
1152
The MAO-activated prototype 135 system yields PEs with very low M

v
(about 700010 000; see entries 1922 in Table 18).
1128,1129
As is typical for c-olefin polymerizations, increasing the
bulkiness around the metal atom is the key to achieving high molecular masses.
11531155
Bulkiness of the substituents
ortho to the phenoxy O atom can be partially used to control molecular mass (see Figure 45(a)). However, the key to
high molecular mass polymers is the bulkiness of the R
1
group. Small R
1
groups yield very low molecular mass (see
Figure 45(b)). These low molecular mass (M

w
<5000) PEs contain a large amount of vinyl terminations (about 90%
of the chain ends),
1152
and are potentially useful because they can be transformed by chain-end functionalization to
terminally functionalized polymers.
11561158
Increasing the size of substituents on the phenyl group of the prototype
139 system yields PEs with high M

v
(200 000). Substitution on the ortho-position is the most effective (see
Figure 45(c)). Unfortunately, as already discussed, this kind of substitution also reduces activity. It was suggested
that these bulky alkyl substituents diminish chain-transfer reactions.
1159
This proposal is supported by theoretical
calculations.
1155,1159
Most spectacular, however, is the effect of perfluorophenyl rings in position R
1
. Molecular
O
N ZrCl
2
n-C
6
H
13
Bu
t
O
N ZrCl
2
n-C
6
H
13
Bu
t
O
N ZrCl
2
Bu
t
O
N iCl
2
Bu
t
MeO
O
N ZrCl
2
Bu
t
MeO
O
N TiCl
2
Bu
t
F
O
N TiCl
2
Bu
t
CF
3
O
N iCl
2
Bu
t
O
N TiCl
2
Bu
t
F
O
N TiCl
2
Bu
t
F F
F
F F
3
C CF
3
F F F
2 2 2
Ethylene polymerization activities: electronic effect
2
(a) (b)
2
2 2 2 2 2
versus
(c)
5 10
4
8 10
4
9 10
4
4 10
4
3 10
4
3 10
4
4 10
3
2 10
3
3 10
3
4 10
4
versus versus
Activities in (g
PE
(mmol M)
1
h
1
bar
1
) Activities in (g
PE
(mmol M)
1
h
1
bar
1
)
Activities in (g
PE
(mmol M)
1
h
1
bar
1
)
T
F
T
Figure 44 Ethylene polymerization activity (g
PE
(mmol M
1
) h
1
bar
1
) for a series of bis(phenoxyimine) Zr catalysts with
substituents having different electronic properties. Left to right, top to bottom, the systems shown correspond to entries 32, 37,
33, 38, 1, 70, 65, 61, 67, and 66 in Table 18.
1110 Olefin Polymerizations with Group IV Metal Catalysts
masses 10
5
are easily achieved with Zr- and Hf-based catalysts
1139
(see entries 73 and 74 in Table 18). In the case of
Ti-based catalysts, chain-transfer reactions are so much depressed that living polymerization is achieved.
1160
Metal effects. The Ti- and Hf-based catalysts are remarkably less active than the corresponding Zr-based catalysts
(see Figure 46). For Ti-based catalysts, the lower activity is compensated by a remarkable increase in the molecular
mass. Instead, the Hf-based catalysts yield molecular masses comparable to those obtained with the corresponding
Zr-based catalyst.
1149
Co-catalyst effects. The effect of different cocatalysts on various polymerization parameters is very remarkable. The
activity of MAO-activated systems shows only a small dependence on the Al/M ratio, Figure 47. Similarly, molecular
mass is also substantially independent of the Al/M ratio, which suggests that transfer to MAO is not the dominant
chain transfer reaction.
1129
Group 4 bis(phenoxyimine) catalysts display surprisingly different catalytic behavior when activated with
[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al.
1129,1135,11611163
While the MAO-activated prototype system 135 yields PES with an M

v
O
N ZrCl
2
C
6
H
6
O
N rCl
2
Bu
t
O
N rCl
2
Bu
t
Bu
t
O
N rCl
2
Bu
t
O
N rCl
2
Bu
t
Bu
t
Bu
t
Me
O
N rCl
2
Bu
t
Bu
t
2
Ethylene polymerization molecular masses: ligand substitution effect
2 2 2 2 2
(c)
(a) (b)
M
v
7,000 M
v
12,000 M
v
18,000 M
w
2,100 M
v
3,000
O
N ZrCl
2
C
6
H
6
O
N ZrCl
2
C
6
H
6
O
N ZrCl
2
C
6
H
6
O
N ZrCl
2
Bu
t
2 2 2 2
M
v
7,000 M
v
26,000 M
v
320,000 M
v
> 2,740,000 M
v
8,000
Z Z Z Z Z
Figure 45 PE molecular masses obtained with a series of bis(phenoxyimine) Zr catalysts. Left to right, top to bottom, the
systems shown correspond to 45, 43, 47, 48, 55, 18, 53, 52, 50, and 51 in Table 18.
O
N
MCl
2
Bu
t
Ethylene polymerization: metal effect
2
M Activity M
v
Ti 3 10
3
510,000
Zr 5 10
5
10,000
Hf 3 10
4
17,000
Cocatalysts MAO, ethylene 1 bar, 25 C
Activities in gPE (mmol M)
1
h
1
bar
1
Figure 46 Activities (g
PE
(mmol M)
1
h
1
bar
1
) and molecular masses obtained in ethylene polymerization with a series of
bis(phenoxyimine) catalysts (see Ref: 1149).
Olefin Polymerizations with Group IV Metal Catalysts 1111
of about 700010 000, activation of the same system with [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al yields PES with extremely high
molecular mass (M

v
~4 10
6
5 10
6
; see entries 2325 in Table 18). The activity decreases considerably, but
reasonable polymerization activity is preserved. The same behavior is observed with the corresponding Ti system
136 (entries 79 in Table 18).
The observed difference in the molecular masses and catalytic activity between MAO and [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al
co-catalysts can hardly be traced to different interaction between the cationic active species and the counterion.
Consequently, it was proposed that activation with MAO and [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al leads to chemically different
active species. Bu
i
3
Al can both alkylate and reduce the imine group. The
1
H NMR experiments indicated that Bu
i
3
Al
reduces the phenoxyimine ligand to a phenoxyamine ligand, which suggests the formation of a bis(phenoxyamine)
active species. Unfortunately, the resulting phenoxyamine complex was not isolated. Nevertheless, addition of
[Ph
3
C][B(C
6
F
5
)
4
] and ethylene to the bis(phenoxyimine)MCl
2
/Bu
i
3
Al mixture yields PE. Following these
observations, it was suggested that for bis(phenoxyimine)-based complexes, activation with MAO and
[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al proceeds as shown in Scheme 38. Theoretical calculations on a model of a bis(phenoxy
amine) Zr active species suggested that the amine substituents near to the active metal atom introduce additional steric
congestion that disfavors space-demanding chain-release reactions.
1162
MgCl
2
/R
n
Al(OR
1
)
3n
-activated bis(phenoxyimine) catalysts are highly active in ethylene polymerization (see
entry 11 in Table 18). Performances of the MgCl
2
-activated catalysts are rather similar to those of the corresponding
MAO-activated catalysts. It was suggested that the R
n
Al(OR
1
)
3n
species act as both alkylating agents and scaven-
gers, while MgCl
2
acts as a Lewis acid to generate the cationic active species from the alkylated catalyst.
Interestingly, PEs produced with the MgCl
2
/R
n
Al(OR
1
)
3n
-activated bis(phenoxyimine) Zr catalysts show morphol-
ogies with well-defined particles, whereas MAO-activated catalysts yield poorly defined morphologies. On this basis,
O
N
ZrCl
2
Bu
t
Ethylene polymerization: MAO effect
2
Al/Zr Yield Activity M
v
6,250 3.3 2 10
5
8,000
15,625 0.9 5 10
5
10,000
62,500 0.9 5 10
5
9,000
125,000 0.9 5 10
5
7,000
312,500 0.8 5 10
5
7,000
+ MAO
Figure 47 Ethylene polymerization yield (g) and activity (g
PE
(mmol cat)
1
h
1
) and M

v
for a series of bis(phenoxyimine)
catalysts. Top to bottom, the systems shown correspond to entries 1822 in Table 18.
O
N
MCl
2
R
1
R
O
N
M
R
1
R
B(C
6
F
5
)
4

R
O
N
M
R
1
R
MAO

+
+
R
i -Bu
2
Al
MAO
[Ph
3
C][(C
6
F
5
)
4
]/Bu
i
3
Al
Bis(phenoxy-imine) active species
Bis(phenoxy-amine) active species
2
2
2
Scheme 38
1112 Olefin Polymerizations with Group IV Metal Catalysts
it has been hypothesized that the complex exists on the surface of solid MgCl
2
/Al-alkoxide crystallites.
1143
However,
it is unclear how supportation would proceed. Finally, bis(phenoxyimine) catalysts can be activated also by
heteropoly- Mo compounds/R
3
Al, and clays/R
3
Al. Activating (135) with (Ph
3
C)
m
H
n
[PMo
12
O
40
]/Et
3
Al or mica/Et
3
Al
results in rather active species.
1137
Temperature effects. Bis(phenoxyimine)-based catalysts possess considerable activity in the range 075

C. The
MAO-activated prototype system (135) shows a remarkable activity, 3 10
5
g
PE
(mmol M)
1
h
1
bar
1
, already at
0

C. The activity maximum is at ca. 40

C, ~6 10
5
g
PE
(mmol M)
1
h
1
bar
1
.
1129
Similar good activities in the
range 2575

C are shown also by other complexes (see Figure 48), independent of the co-catalyst used. Interestingly,
the low thermal stability can be alleviated by including electron-donating groups in position 4 (see Figure 48).
1130
Finally, although the lifetime of these catalysts tends to be limited, ethylene polymerizations over periods of 5, 15,
and 30 min indicated that MAO-activated 135 has a catalytic lifetime of at least 30 min. Solvents seem to have a minor
effect on the polymerization performance.
Molecular mass distributions. The molecular mass distributions from bis(phenoxyimine) catalysts depend strongly
on both the metal and the ligand that compose the actual catalyst. The large majority of Zr-based catalysts show
M

w
,M

n
ratios around 2, typical of single-site catalysts. However, in some cases, rather large M

w
,M

n
values (from 4 to
40) have been observed (see entries 48, 57, and 58 in Table 18).
1142,1144
GPC curves of the PEs with high M

w
,M

n
ratios indicated multimodal (bi- and trimodal) molecular mass distribution. It was suggested that the high M

w
,M

n
values arise from an equilibrium between the three active isomers AC that bis(phenoxyimine) catalysts can assume
(see Figure 41). The multimodality stems from the different catalytic properties of different isomers. Equilibrium
between different isomers of the bis(phenoxyimine) catalysts was confirmed by NMR experiments
1142,1144
and
supported by DFT calculations.
1144
Deconvolution of the bimodal GPC peaks led to two symmetrical peaks. This
indicated that the bimodal PE consists of two unimodal fractions, and both fractions are consistent with single-site
polymerization behavior.
1144
Examination of different complexes indicated that a methyl or methoxy substituent
para to the phenoxyoxygen results in bimodal behavior. Cumyl substituents ortho and para to the phenoxyoxygen
result in trimodal behavior and perfluorophenyl substituents invariably lead to unimodal behavior.
1142,1144
Living polymerizations. Of greater interest, however, is the fact that fluorophenyl-substituted bis(phenoxyimine) Ti
complexes easily show extremely narrow M

w
,M

n
ratios, around 1, indicative of living polymerization
behavior.
1132,1134,1140,1160
At 50

C, the perfluorophenyl-substituted complex 137 is more active than the correspond-


ing non-fluorinated complex 136 (4 10
4
vs. 3 10
3
g
PE
(mmol M)
1
h
1
bar
1
; entries 67 and 4 in Table 18), and
yields a PE with an extremely narrow molecular mass distribution (M

w
,M

n
1.13).
1134
Analogous perfluorophenyl Zr
and Hf complexes result in M

w
,M

n
close to 2, and thus do not present living behavior (see entries 74 and 75 in
Table 18). Remarkably low M

w
,M

n
values are also obtained at higher temperature (T
P
75

C, M

w
,M

n
= 1.15; T
P
90

C, M

w
,M

n
= 1.30). Plots of M

n
and M

w
,M

n
versus polymerization time indicate that M

n
increases linearly with
time, and that M

w
,M

n
remains close to 1 for at least 15 min.
1134
O
N
TiCl
2
Bu
t
O
N
TiCl
2
Bu
t
O
N
ZrCl
2
Bu
t
MeO
Ethylene polymerization: temperature effect
2
T
P Activity
M
v
25 516,000
50 546,000
75 440,000
(a)
2
(b) (c)
+ MAO + [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al
2
+ MAO
3 10
3
4 10
3
4 10
3
T
P Activity
M
v
50 168,000
75 54,000
9 10
4
2 10
5
T
P Activity
M
v
25 4,810,000
50 5,860,000
75 3,920,000
2 10
2
4 10
2
7 10
2
Figure 48 Ethylene polymerization activity (g
PE
(mmol M)
1
h
1
bar
1
) and M

v
versus polymerization temperature, T
P
(

C), for a
series of bis(phenoxyimine)-based catalysts. Top to bottom and left to right, the systems shown correspond to entries 13,
79, and 38 and 39 in Table 18.
Olefin Polymerizations with Group IV Metal Catalysts 1113
Further evidence of the living polymerization nature was obtained by the fact that the GPC peaks of the PE
produced shift to higher molecular mass on increasing the polymerization time. The monomodal shape is retained,
and no shoulders or low molecular mass tails are detected.
1134
The stability of the living polymer chain was
investigated utilizing the MAO-activated complex 137 at 25

C.
1134
First, the activated complex is treated with
ethylene-saturated toluene for 65 min. The values of M

n
versus time clearly indicate that after 3 min all the ethylene
is consumed. After 65 min under an N
2
atmosphere, ethylene gas was fed to the system for 2 additional min. The
M

w
,M

n
value resulting after the additional 2 min ethylene feed is 1.14, which indicates that no termination reaction
occurred for at least 60 min in the absence of ethylene. This remarkable result opens the route to the controlled
synthesis of ethylene-based block co-polymers.
To investigate the origin of the living behavior, several differently fluoro-substituted Ti-based catalysts were
synthesized.
1134,1140
The main results are presented in Figure 49. Most importantly, ligands without F atoms in the
ortho-position do not exhibit living polymerization behavior (i.e., M

w
,M

n
close to 2). By contrast, even one single F
atom in an ortho-position confers living behavior to the corresponding catalyst, although very low activity is observed.
Nevertheless, activity can easily be improved by increasing the number of F atoms on the aryl ring. The more active
catalysts give marginally broader molecular mass distributions. Fujita and co-workers also remarked that a single
methyl group in the ortho-position does not lead to living behavior, since M

w
,M

n
= 2.14 after 30 min poly-
merization.
1138
Fujita and co-workers proposed that hydrogen-bonding interactions between the ortho-F atoms and H atoms on the
u-C atom of the growing chain might be responsible for the living behavior.
1140
This electrostatic interaction
between the negatively charged F atom and the positively charged u-H atom of the growing chain was thought to
stabilize the active species and depress the most likely u-H transfer termination reaction.
1140
The NMR and X-ray
characterizations of related Zr model compounds did indeed indicate that this kind of interaction is established both
in solution as well as in the solid state.
1164
On the other hand, based on combined quantum mechanics/molecular
mechanics calculations, Talarico and co-workers suggested that the main role of the ortho-F substituents was to
increase steric bulkiness around the metal atom, which disfavors the chain transfer to monomer, in analogy with
similar calculations on metallocenes as well as on Ni-based Brookharts type catalysts.
943,1153,1154,1165
Their calcula-
tions indicated the presence of an ortho-F u-H interaction in the transition state, but the strength of this interaction
was estimated to be only about 1 kcal mol
1
. Consistent with this, these calculations also indicated that a methyl
group in the ortho-position should depress chain termination.
1155
At the same time, Fujita and co-workers showed that a Cl atom in the ortho-position also yields a narrow
polydispersity (M

w
,M

n
= 1.23).
1134
They attributed this effect to the lone electron pairs on the Cl atom.
However, it must be noted that Cl is by far less electronegative than F, and that Cl has roughly the same steric
bulkiness as an Me group.
This topic has also been investigated by Coates and co-workers who reported that even the unfluorinated catalyst
of Figure 50 polymerizes ethylene in a living fashion in the temperature range 050

C, although the activity and the


molecular masses were low compared to those obtained with the perfluorinated analog.
1166
These contrasting results
indicate that the ortho-F effect cannot be considered as the only rationalization of the living behavior exhibited by
bis(phenoxyimine) Ti catalysts, and that steric effects play an important role. In conclusion, the real strength of this
attractive F-interaction, as well as the real origin of the living behavior, remain an open topic.
F
F
F
F
F
F
F
F F F F
F
F
F
F
F F
R
1
M
w
/M
n
M
n
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Polymerization time (min)
1.13
424,000
4 10
4
1
1.25
145,000
2 10
3
5
1.05
64,000
8 10
2
5
1.06
13,000
1 10
2
5
1.99
98,000
4 10
4
1
1.78
129,000
3 10
4
1
2.18
128,000
5 10
3
5
O
N iCl
2
R
1
Bu
t
2
T
Figure 49 Ethylene polymerization results with fluorinated bis(phenoxyimine) Ti-based catalysts activated with MAO at 50

C.
Left to right, the systems shown correspond to entries 67, 70, 71, 72, 64, 62 and 60, respectively, in Table 18.
1114 Olefin Polymerizations with Group IV Metal Catalysts
Remarkably, mixed-ligand titanium complexes with one phenoxyimino and one iminopyrrolato ligand, such as
pre-catalysts 140 and 141, show very high ethylene polymerization activities, about 5 10
3
and
9 10
4
g
PE
(mmol M)
1
h
1
bar
1
(MAO, 20

C). Thus, the heteroligated complexes, give activities substantially


higher than those of the corresponding homoligated complexes, with either two phenoxyimine or two imino
pyrrolate ligands under comparable conditions.
1167
A similar beneficial effect on activity, due to heteroligation, has
also been reported for propylene polymerization.
1168
Ti
N
O
Cl
N
Cl
N H
5
C
6
H
5
C
6
Bu
t
140
Ti
N
O
Cl
N
Cl
N F
5
C
6
H
5
C
6
Bu
t
141
4.09.5.3.3 Ligands with coordinating NO atoms: phenoxyimine catalysts for syndiotactic
polypropylene
Bis(phenoxyimine) catalysts have also been used extensively in the syndiotactic polymerization of
propylene.
160,1131,1138,1140,1147,1162,1163,11691177
The perfluoro Ti-based complexes activated with MAO produce
highly syndiotactic PP ([r] 90%) in a living manner and with rather good activities (see entries 10, 11, and 1723 in
Table 19).
1131,1173,1174
This result was rather surprising considering that C
2
-symmetric pre-catalysts should produce
iPPs. Aside from the rrrr pentad peak, the
13
C NMR analysis of the sPP revealed rrrm and rmrr peaks in a 1 : 1 ratio
and no detectable rmmr peak.
1174
The presence of isolated m stereomistakes indicated that these catalysts yield sPP
through a chain-end stereocontrol.
1173,1174
Moreover, the narrow M

w
,M

n
ratio (about 1.1) indicated that living
polymerization was achieved.
1173
Subsequent experiments indicated that the regiochemistry of chain propagation
is secondary.
160,161,1169,1172
Activity and syndiospecificity: ligand effect. The three R
1
, R
2
, and R
3
substituents can be used to control several
properties of the PPs produced. As in the case of ethylene polymerization, the R
1
substituent is the key to conferring
living behavior, and also has a remarkable effect on the molecular mass as well as on the stereoregularity of the PPs
(see Figure 51). The perfluorophenyl-based catalyst is remarkably more syndiospecific than the non-fluorinated
analog.
1177
The ortho- and para-F atoms seem to be important for high % rrrr values. This suggests that the
syndiospecificity is determined by both steric and electronic effects. The presence of ortho-F substituents is again
the key for living behavior. As for ethylene polymerization, the perfluorophenyl system is more active and yields
higher M

n
than the non-fluorinated analog.
1138,1172,1177
The bulkiness of the R
2
substituent controls several parameters of the PPs produced (see Figure 52). Generally, small R
2
groups result in rather active catalysts that produce low molecular mass PPs with low stereoregularity. The bulky R
2
substituent controls stereoregularity. In the case of the prototype 137 system, an almost linear relationship was established
betweenthe percentage of rr diads of the PPs and the volume occupiedby the R
2
substituents (see Figure 52).
1172
The non-
fluorinated analogs behave similarly (see entries 1215 in Table 19).
1138
The T
m
of the resulting PPs reflects of course the
high level of syndiotacticity, and the complex with R
2
=SiMe
3
at 0

C yields an sPP with a T


m
of 156

C (rr =94%).
1172
Finally, the nature of the R
3
substituent has almost no effect on the polymerization behavior (see entries 18 and 24 in
O
N iCl
2
Ph
Bu
t
Bu
t
2
0
210
3
0.73
3
78,400
1.16
20
410
3
0.42
1
48,500
1.07
50
310
3
0.37
1
44,500
1.10
Temperature (C)
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
Polymerization time (min)
M
n
M
w
/M
n
MAO/Ti = 150, toluene, ethylene 0.69 bar
T
Figure 50 Ethylene polymerization results with non-fluorinated bis(phenoxyimine) Ti-based catalysts activated with MAO.
Olefin Polymerizations with Group IV Metal Catalysts 1115
Table 19 Syndiospecific propylene polymerization catalyzed by bis(phenoxyimine) catalysts
Entry Pre-catalyst T
P
a
(

C)
P
P
b
(bar)
(Solvent) t
P
c
(h)
Productivity
(g
PP
(mol M
1
)
h
1
bar
1
) Yield (g) M

w
/M

n
M

n
T
m
d
(

C) rr
e
(%) rrrr
f
(%) References
1
O
N
TiCl
2
Bu
t
2
Bu
t
0 2.8 (Tol) 24 600 4.20 2.14 9,910 78 1177
2
O
N
TiCl
2
Bu
t
2
F
Bu
t
0 2.8 (Tol) 24 60 0.38 1.07 3,220 52 1177
3
O
N
TiCl
2
Bu
t
2
F
Bu
t
0 2.8 (Tol) 24 1,100 7.20 1.75 18,600 78 1177
4
O
N
TiCl
2
Bu
t
2
CF
3
Bu
t
0 2.8 (Tol) 24 1,000 6.40 2.19 19,280 81 1177
5
O
N
TiCl
2
Bu
t
2
F
F
Bu
t
0 2.8 (Tol) 24 80 0.56 1.06 16,410 83 1177
6
O
N
TiCl
2
Bu
t
2
F F
F
Bu
t
0 2.8 (Tol) 24 2200 14.30 2.03 9,150 83 1177
7
O
N
TiCl
2
Bu
t
2
F F
Bu
t
0 2.8 (Tol) 24 3,600 23.40 1.91 13,580 81 1177
8
O
N
TiCl
2
Bu
t
2
CF
3
CF
3
Bu
t
0 2.8 (Tol) 24 1,900 12.10 2.17 14,090 51 1177
(Continued)
Table 19 (Continued)
Entry Pre-catalyst T
P
a
(

C)
P
P
b
(bar)
(Solvent) t
P
c
(h)
Productivity
(g
PP
(mol M
1
)
h
1
bar
1
) Yield (g) M

w
/M

n
M

n
T
m
d
(

C) rr
e
(%) rrrr
f
(%) References
9
O
N
TiCl
2
Bu
t
2
F
F
F
Bu
t
0 2.8 (Tol) 24 400 2.48 1.08 43,420 95 1177
10
O
N
TiCl
2
Bu
t
2
F F
F
F
F
Bu
t
0 2.8 (Tol) 5.2 3,800 5.34 1.11 95,900 96 1177
11 0 2.8 (Tol) 24 2,500 16.20 1.26 216,610 96 1177
12
O
N
TiCl
2
2
1 3.7 (Tol) 6 100 0.24 2.93,1.51
M

w
735,000
M

w
15,000
nd 1138
13
O
N
TiCl
2
Me
2
1 3.7 (Tol) 6 1,800 4.02 1.47 M

w
101,000 nd 1138
14
O
N
TiCl
2
Bu
t
2
1 3.7 (Tol) 6 250 0.57 1.38 M

w
6,000 97 63 1138
15
O
N
TiCl
2
SiMe
3
2
1 3.7 (Tol) 6 400 0.96 1.73 M

w
14,000 140 84 1138
16
O
N
TiCl
2
Bu
t
2
1 3.7 (Tol) 6 250 0.56 1.39 M

w
7,000 101 66 1138
17 0 1 (Tol) 5 2,800 0.144 1.05 23,600 136 1172
18 25 1 (Tol) 5 3,600 0.183 1.11 28,500 137 87 1172
19
O
N
TiCl
2
Bu
t
2
F F
F
F
F
50 1 (Tol) 5 2,600 0.148 1.37 16,400 130 1172
(Continued)
Table 19 (Continued)
Entry Pre-catalyst T
P
a
(

C)
P
P
b
(bar)
(Solvent) t
P
c
(h)
Productivity
(g
PP
(mol M
1
)
h
1
bar
1
) Yield (g) M

w
/M

n
M

n
T
m
d
(

C) rr
e
(%) rrrr
f
(%) References
20 25 6 (Tol) 1 2,600 0.158 1.07 30,900 135 1172
21 25 6 (Tol) 2 2,600 0.312 1.10 52,800 1172
22 25 6 (Tol) 3 2,500 0.460 1.10 73,800 135 1172
23 25 6 (Tol) 5 2,400 0.713 1.14 108,000 135 1172
24
O
N
TiCl
2
Bu
t
2
F F
F
F
F
25 1 (Tol) 5 2,300 0.115 1.11 16,500 140 1172
25
O
N
TiCl
2
SiMe
3
2
F F
F
F
F
25 1 (Tol) 5 5,800 0.293 1.08 47,000 152 93 1172
26
O
N
TiCl
2
SiEt
3
2
F F
F
F
F
25 1 (Tol) 5 3,500 0.174 1.16 24,400 151 1172
27
O
N
TiCl
2
2
F F
F
F
F
25 1 (Tol) 5 30,600 1.534 1.51 189,000 n.d. 43 1172
28
O
N
TiCl
2
2
F F
F
F
F
25 1 (Tol) 5 69,000 3.440 1.22 260,000 n.d. 50 1172
29
O
N
TiCl
2
Pr
i
2
F F
F
F
F
25 1 (Tol) 5 31,000 1.555 1.16 153,700 n.d. 75 1172
a
T
P
=polymerization temperature.
b
P
P
=propylene pressure.
c
t
P
=polymerization time.
d
T
m
=melting temperature of the PPs produced.
e
% rr =amount of rr triads in the PPs produced.
f
% rrrr =amount of rrrr pentads in the PPs produced.
Table 19), probably because the R
3
substituent is too distant fromthe metal. As in the case of ethylene, effective propylene
living polymerization, or more generally low M

w
,M

n
, requires the presence of F atoms on the R
1
aromatic ring (see
Figure 51).
1177
The perfluorinated catalyst (137) exhibits living behavior for a period of at least 5 h at 25

C.
1172
Moreover,
longer polymerization times give higher molecular masses. Unimodal molecular mass distributions were observed in all
cases.
1172
Bulky R
2
groups also reduce the amount of regiomistakes, measured as head-to-head units.
Interestingly, heteroligated living/non-living bis(phenoxyimine) Ti-based catalysts give higher activity and
higher molecular masses relative to the symmetric and homoligated catalysts. For example, MAO-activated pre-catalysts
142 and 143 exhibit propylene polymerization activities around 10
4
g
PP
(mmol M)
1
h
1
bar
1
, whereas the activities of
the corresponding homoligated living and non-living pre-catalysts are one order of magnitude smaller.
1168
Similar
beneficial effect on activity, due to heteroligation, has also been reported for ethylene polymerization.
1167
O
Ti
N
Cl
N
O
Cl
Bu
t
Bu
t
Bu
t
Bu
t
F
F
F
F
F F
3
C
F
F
142
O
Ti
N
Cl
N
O
Cl
Bu
t
Bu
t
Bu
t
Bu
t
F
F F
F
F
143
O
N iCl
2
R
1
Bu
t
F
F
F
F
F
F
F
F F F
F
F
F F F
R
1
2
96
1.26
216,610
2
95
1.08
43,420
4 10
1
83
2.03
9,150
2
83
1.06
16,410
7 10
2
81
1.91
13,580
3
78
2.14
9,910
6 10
1
% rrrr
M
w
/M
n
M
n
Activity (g
PP
(mmol M)
1
h
1
bar
1
)
T
Figure 51 Propylene polymerization results with fluoro-substituted bis(phenoxyimine) Ti-based catalysts activated with MAO
at 0

C. Effect of the R
1
substituent.
R
2
SiMe
3
Bu
t
Pr
i
Me H
152
1.08
47,000
6
93
4
3
3
137
1.11
28,500
4
87
10
3
4
n.d.
1.16
153,700
3 10
75
22
3
8
n.d.
1.22
260,000
7 10
50
42
8
9
n.d.
1.51
189,000
3 10
43
46
11
10
T
m
(C)
M
w
/M
n
M
n
Activity g
PP
(mmol M)
1
h
1
bar
1
[rr] (%)
[mr] (%)
[mm] (%)
Head-to-head units (%)
O
N iCl
2
C
6
F
5
R
2
2
25 C, 1 bar propylene, MAO/Ti = 250, toluene, 5 h
T
Figure 52 Propylene polymerization results with fluoro-substituted bis(phenoxyimine) Ti-based catalysts. Effect of the R
2
substituent.
1122 Olefin Polymerizations with Group IV Metal Catalysts
Activity and syndiospecificity: metal effect. Zirconium and hafnium catalysts are more active than their titanium
congeners (see Figure 53).
1171
However, Zr and Hf only oligomerize propylene. The effect is particularly dramatic
for the non-fluorinated catalysts. While consideration of tacticity in the case of the Zr/MAO systems is meaningless
due to the extremely low degree of polymerization, the perfluoro Hf/MAO catalyst yields substantially atactic or
slightly syndiotactic PPs (% mm/mr/rr =9/47/44 at 25

C and 10/48/42 at 50

C). These data also indicate that tacticity


is substantially independent of the polymerization temperature.
1171
Activity and syndiospecificity: the effect of the polymerization conditions. The propylene pressure as well as the solvent
utilized have been suggested to influence the syndiotacticity of the resulting PP.
1178
The rr of PPs produced with the
MAO-activated prototype catalyst 136 decreases from 78% to 70% on increasing the propylene pressure from 2 to
6 bar. Similar behavior is shown by the fluorinated catalyst (137) (rr from 95% to 83% on increasing propylene
pressure from 1 to 6 bar). Replacing the commonly used toluene with CH
2
Cl
2
results in the synthesis of substantially
stereoirregular PPs (rr as low as 30%).
1178
Loss of syndiospecificity with increasing dielectric constant of the solvent
was already reported for propylene polymerization with the classical Me
2
C(Cp)(9-Flu)ZrCl
2
/MAO zirconocene, and
was attributed to isomerization of solvent-separated free zirconocene species via migration of the growing chain
before the next insertion.
1179
However, it is difficult to imagine a common mechanism for these two catalysts. In fact,
in the case of the bis(phenoxyimine) systems, the regiochemistry of monomer insertion is secondary and the
stereoselectivity is chain-end controlled, whereas in the case of the zirconocene the regiochemistry of monomer
insertion is primary and stereoselectivity is site-controlled.
Regiochemistry of propylene insertion. The regiochemistry of propylene polymerization with MAO-activated bis-
(phenoxyimine) Ti catalysts is prevailingly secondary.
160,161,1169,1172
NMR analysis of chain-initiation and -termi-
nation groups indicated that the MAO-activated prototype (137) leads exclusively to n-propyl, isobutyl, and isopentyl
end groups (see Scheme 39). Considering that the initiation occurs on a TiMe bond, it was concluded that primary
propylene insertion into the TiMe bond is strongly favored. Moreover, the ratio of the isopentyl versus isobutyl
NMR peaks also showed that the first 1,2-propylene insertion is followed by 70% of 2,1- and 30% of 1,2-propylene
insertions. This indicates that propylene insertion on a Tiprimary alkyl bond is highly regioirregular. Similar analysis
on the chain-end termination groups confirmed that propylene polymerization with catalysts based on (137) is
prevailingly secondary, and that the PP has a regioblock structure.
1169,1172
Further NMR analysis of chain-end groups of PPs produced with similar catalysts provided additional evidence for
the prevailingly secondary propylene propagation with this class of catalyst. In fact, it was shown that the main chain-
release reaction is u-H elimination, and that propylene insertion into the TiH bond in the initiation step is almost
exclusively primary.
160,161
Moreover, NMR analysis of a co-polymer of propylene with a small amount (<2 mol%) of
1-
13
C-ethylene, obtained with the perfluorinated catalyst (137), showed that the large majority of ethylene units in
the co-polymer was present as two methylene units (see Scheme 40). This clearly indicated that ethylene units bridge
blocks of propylene units with opposite regiochemistry, which is consistent with and further supports the whole
mechanistic scenario.
161
However, the regiochemistry of propylene insertion is dependent on both the metal and the bulkiness of the
substituents ortho to the O atoms. In fact, complexes with an ortho-methyl group, which lead to poorly syndiotactic
PPs, enchain the monomer via opposite regiochemistry
277
(see Figure 54). The
13
C NMR analysis indicated the
presence of isobutyl (initiation) and n-propyl (termination) end groups for the Ti-based catalysts, consistent with the
prevailingly secondary propagation already discussed for the titanium phenoxyimine systems, whereas the Zr
analogs showed the almost exclusive presence of isobutyl end groups, indicative of primary propylene insertion in
M
O
N Cl
2
R
1
Bu
t
2
R
1
Ti
Zr
Zr
Hf
Hf
C
6
F
5
C
6
H
5
C
6
F
5
C
6
H
5
C
6
F
5
0.18
1.72
1.07
0.89
5.71
4
7 10
2
4 10
2
6 10
4 10
4
Activity
Activity in g
PP
(mmol M)
1
h
1
bar
1
, 25 C, 1 bar propylene,
MAO/M = 250, toluene, Ti = 5 h, Zr = 0.5 h, Hf = 1.5 h
Yield M
w
/M
n
M
n
1.11
1.55
2.34
1.62
2.73
28,500
150
1,340
370
9,990
M
Figure 53 Performance of various phenoxyimine catalysts in propylene polymerization.
Olefin Polymerizations with Group IV Metal Catalysts 1123
both the initiation and the termination steps. This suggested that primary insertion is also the prevailing propagation
mode. Further support for this proposal was achieved by co-polymerization of propylene with small amounts (~0.4%)
of 1-
13
C-ethylene. The remarkably larger proportion of three methylene versus two methylene sequences (79% vs.
21%) was a clear indication that for the Zr-based catalysts primary propagation is favored (see Scheme 40).
277
The
different regioselectivity exibited by the Ti and Zr catalysts has been rationalized by DFT calculations.
165
Polymerization of c-.-diolefins has also been investigated to further explore the unusual regiochemistry exhibited
by bis(phenoxyimine) Ti catalysts. Cyclopolymerization of 1,6-heptadiene (see Figure 55) produced a polymer with
no observable unsaturations. This indicated quantitative cyclization of the monomer. The NMR analysis of the
polymer indicated the presence of ethylene-1,2-cyclopentane units and of methylene-1,3-cyclohexane units in almost
O
N Ti
F
5
C
6
Bu
t
2
Me
O
N Ti
F
5
C
6
2
O
N Ti
F
5
C
6
2
O
N Ti
F
5
C
6
2
O
N Ti
F
5
C
6
2
O
N Ti
F
5
C
6
2
O
N Ti
F
5
C
6
2
Isobutyl
(observed)
Isopentyl
(observed)
Ethyl
(not observed)
sec-Butyl
(not observed)
1,2-insertion
1,2-insertion
1,2-insertion
2,1-insertion
2,1-insertion
2,1-insertion
Scheme 39
C
C
C C
C
C C C C C C C
C C
C C C C C C C C C C
C C
>>
C C
(a) (b)
Scheme 40
18 C, 1 bar propylene, MAO/M = 150, toluene, 120 min
M
4.95
0.67
8,300
2.3
Yes
Yes
1.40
0.71
268,900
2.9
Yes
No
Zr Ti
O
N Cl
2
C
6
F
5
Me
2
Yield (g)
P
r
M
w
M
w
/M
n
Isobutyl end groups
n-Propyl end groups
M
Figure 54 Propylene polymerization with methyl substituted phenoxyimine Ti and Zr catalysts.
1124 Olefin Polymerizations with Group IV Metal Catalysts
equal amounts. The presence of both cyclopentane and cyclohexane units clearly demonstrated that secondary
insertion is easy with these catalysts since cyclopentane rings can only form following a secondary insertion of the first
double bond of 1,6-heptadiene.
160
Mechanism of stereocontrol. The chain-end stereocontrol exhibited by MAO-activated bis(phenoxyimine) Ti
catalysts was rather surprising, considering the C
2
-symmetry of the pre-catalyst (see Scheme 41). In fact, the PPs
produced by a catalyst such as that of Scheme 41 should lead to an iPP, and stereoselectivity should be site controlled.
The evidence that chain-end stereocontrol was instead operative led to the suggestion that the catalyst site could
invert the configuration at each insertion step.
1174
This was confirmed by subsequent quantum mechanics/molecular
mechanics calculations. The key of the mechanism is that the chiral chain end imposes one of the two configurations
on the chiral active site. The imposed configuration of the active site in turn selects between the two enantiofaces of
the incoming monomer through direct steric interaction between the R
1
substituent and the methyl group of the
secondary inserting propylene molecule.
193
Of course, this mechanism requires that site isomerization is faster than
propylene insertion. A direct consequence of this mechanism is that active species that cannot isomerize rapidly
should lead to isotactic polymers.
193
Although for these systems inversion between the two configurations is not experimentally proved, fluxionality in
solution has been reported for some Zr bis(phenoxyimine) complexes.
1142
Moreover, existence of these fluxional
equilibria in related neutral complexes was demonstrated by NMR experiments, which gave barriers for the
interconversion close to 15 kcal mol
1
.
1180,1181
Site isomerization was proposed to occur through the dissociative
mechanism of (Scheme 42). Quantum mechanics calculations have predicted a TiN bond-dissociation energy
slightly lower than 15 kcal mol
1
in complexes related to those of Figure 44.
172
Finally, it has been shown that some fluorinated catalysts, though not particularly effective in high molecular mass
polymerization, are instead excellent catalysts for the synthesis of allyl-terminated sPP oligomers (M

n
~30004000).
These new short sPP polymers with functionalized chain end could be used to make LCB polymers as well as other
new polymeric materials with sPP segments.
258
O
N Ti
F
5
C
6
2
1,2-insertion
P
2,1-insertion
P
O
N Ti
F
5
C
6
2
O
N Ti
F
5
C
6
2
P
1,6-cyclization
1,5-cyclization
Ethylene-1,2-cyclopentane
unit
Methylene-1,3-cyclohexane
unit
Figure 55 Cyclopolymerization of 1,6-heptadiene with bis(phenoxy-imine) Ti based catalysts.
Ti
Cl Cl
O
O
R
R
2
R
3
R
3
R
1
R
1
2
N
N
Ti
Cl Cl
O
O
R
R
2
2
R
3
R
3
R
1
R
1
N
N
site
isomerization
complex complex
Scheme 41
Olefin Polymerizations with Group IV Metal Catalysts 1125
4.09.5.3.4 Ligands with coordinating NO atoms: phenoxyimine catalysts for isotactic
polypropylene
While MAO activation of phenoxyimine-based complexes results in the syndiospecific polymerization of propylene
(see the previous Section), activation of the same complexes with [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al shows completely different
behavior.
1162,1163
The Ti-based catalyst leads to substantially atactic PP with extremely high molecular mass and rather
broad M

w
,M

n
, while the Zr- and Hf-based analogs afford moderately isotactic PPs, with moderately high molecular
masses and polydispersities typical of single-site behavior (see Figure 56).
1162
In agreement with this, the PP produced
with the Ti-based catalyst shows no melt transition T
m
, whereas PPs produced with Zr and Hf catalysts do show a T
m
,
indicative of stereoregular polymers. The Hf catalyst is remarkably more stereospecific than the Zr congener, although
in both cases only moderately isotactic PPs are formed. The NMR analysis of the Zr- and Hf-produced PP at pentad
level (Zr: mmrr 9.1%, mrrm 6.1%, mmmr 6.5%; Hf: mmrr 11.8%, mrrm 5.7%, mmmr 11.3%) revealed the presence of
isolated rr triads, indicative of a site-controlled mechanism. Increasing propylene pressure to 4 bar resulted in a
remarkable increase in molecular masses (Zr: M

w
= 698 000, M

w
,M

n
= 2.4; Hf: M

w
= 1 460 000, M

w
,M

n
= 2.3).
1162
The remarkable effect of [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al activation on the behavior of phenoxyimine-based catalysts
was ascribed to a reaction between Bu
i
3
Al and the imino functionality of the complexes. As already discussed in
Section 4.09.5.3.2, it is commonly accepted that Bu
i
3
Al reduces the imino group to an amino group, thus converting a
phenoxyimine catalyst into a phenoxyamine structure.
1162
Using well-established concepts, the isotacticity of the Zr- and Hf-produced polymers was improved by increasing
the bulkiness of the alkyl group ortho to the phenoxy O atom (see Figure 57). The resulting PPs had a multimodal
MWD. Possible explanations were that the Bu
i
3
Al does not completely reduce the imino functionality and/or that the
active species correspond to different geometrical isomers. In all a cases, the polymers had remarkably high T
m
,
indicative of the high stereoregularity. The PP fraction insoluble in boiling hexane had relatively narrow M

w
,M

n
(<5), and NMR analysis confirmed the high isotacticity (R
1
=Cy: Zr 96.9% mmmm, Hf 96.8% mmmm). NMR analysis
of the chain-end groups suggested that polymerization occurs mainly by primary monomer insertion.
1163
O
Mt
O
N
X
N
X
O
Mt
O
N
X
N
X
O
Mt
O
X
N
X
N
-Complex
6-Coordinated
Achiral Complex
5-Coordinated
- Complex
6-Coordinated
Scheme 42
M
Ti Zr Hf
5 10
0.08
8,286,000
4.15
No peak
22.9
45.7
27.7
9 10
0.16
209,000
2.42
103.5
45.8
28.6
25.6
8 10
0.13
412,000
2.15
123.8
69.0
19.1
10.9
Activity (g
PP
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
w
M
w
/M
n
T
m
(C)
mm (%)
mr (%)
rr (%)
O
N Cl
2
C
6
H
5
Bu
t
2
25 C, 1 bar propylene, [Ph
3
C][B(C
6
F
5
)
4
]/M = 2, Bu
i
3
Al/M = 30, toluene, 20 min
M
Figure 56 Propylene polymerization with [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al-activated phenoxyimine catalysts.
1126 Olefin Polymerizations with Group IV Metal Catalysts
The synthesis of iPP was also achieved using the MAO-activated phenoxyketimine complexes of Figure 58.
These complexes combine isotacticity and living behavior in propylene polymerization.
1182
All the catalysts except
for R
2
=R
3
=Bu
t
form PP with narrow molecular mass distributions indicative of living behavior (M

w
/M

n
<1.17).
Polymerizations with R
2
=R
3
=Me at 0

C were run for different reaction times (up to 10 h). Analysis of the resulting
PPs indicated a linear relationship between M

n
and yield, which supported the living polymerization behavior. As
with the aldimine catalysts, small R
2
groups result in higher activity. The
13
C NMR analysis showed the presence of
stereomistakes (mmmr, mmrr, and mrrm pentads in a 2 : 2 : 1 ratio), consistent with an enantiomorphic site-control
enchainment mechanism. Isotactic stereocontrol is only moderate, and the highest isotacticity is obtained with the
relatively small R
2
=Me group. To explain the low isotacticity observed with the biggest and smallest R
2
=Bu
t
and H
groups, it has been suggested that the very bulky R
2
groups result in unfavorable steric interaction between R
2
and
the incoming monomer, while R
2
groups that are too small give ineffective enantioselectivity. Regioirregular
insertions are not negligible (<5%). The absence of olefinic signals in the NMR spectra suggested that u-H and
u-alkyl chain-release reactions do not occur. Cyclopolymerization of c,.-dienes suggested that the regiochemistry
of monomer insertion is secondary. To explain the isotacticity obtained, it has been proposed that the presence of
the Ph group on the imine C atom prevents catalyst racemization.
1182
4.09.5.3.5 Other ligands with coordinating NO atoms
Of course, several variations of the original bis(phenoxyimine) ligand have been proposed. For example, the
coordinating N atom can be enclosed into a pyridine ring, and it has been reported that the bis(phenoxypyridine)
Ti complexes of Figure 59 polymerize ethylene. Although they are active catalysts, their activities are remarkably
R
1
Zr Hf
200,000
4.72
163.3
40 60 25 650
530,000
14.6
164.8
607,000
18.4
154.9
Activity (g
PP
(mmol M)
1
h
1
bar
1
)
M
w
M
w
/M
n
T
m
(C)
25 C, 1 bar propylene, [Ph
3
C][B(C
6
F
5
)
4
]/M = 2, Bu
i
3
Al/M = 30, toluene, 20 min
1,453,000
2.99
160.1
O
N MCl
2
R
1
2
Cy Ph
Zr Hf
Figure 57 PP polymerization with [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al-activated phenoxyimine catalysts. Role of the R
1
substituent.
0.2
0.16
bimodal
broad
n.d.
7
26
Activity (g
PP
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
n
M
w
/M
n
T
m
(C)
mmmm (%)
rrrr (%)
O
N TiCl
2
C
6
F
5
R
2
2
0 C, 2 bar propylene, toluene, MAO/Ti = 150, 3 h
Ph
R
3
Bu
t
Bu
t
Pr
i
Pr
i
2.2
0.40
7,290
1.17
n.d.
45
<1
Et
Me
6.1
1.11
27,940
1.11
69.5
53
<1
Me
Me
6.8
1.22
35,440
1.12
n.d.
8
13
H
Me
R
2
=
R
3
=
0.7
0.13
2,710
1.12
n.d.
46
<1
Figure 58 Propylene polymerization with MAO-activated phenoxyketimine catalysts. Role of the R
2
substituent.
Olefin Polymerizations with Group IV Metal Catalysts 1127
lower than those exhibited by the similarly substituted bis(phenoxyimine)Ti catalysts. Bulky groups ortho to the
phenoxy O atom do not improve activity. Based on the quantum mechanics calculations, it has been suggested that
bis(phenoxypyridine) Ti-based catalysts adopt a trans-N, cis-O, cis-X geometry. This geometry would locate bulky
groups ortho to the phenoxy O atoms away from the active metal center, and thus would reduce their impact on
activity.
1183
Poor activity is shown by other complexes with a coordinating N-pyridine atom. In fact, complexes 144 and 147 are
not active in ethylene polymerization,
1181,1184,1185
whereas complexes 145 and 146 show low to moderate activity
(about 1030 g
PE
(mmol M)
1
h
1
bar
1
) and yield PE with very low M

n
(<7000).
1184
However, these complexes
also contain a more strained five-membered chelating ligand relative to the six-membered chelating ligands of
bis(phenoxyimine) catalysts.
N
O
ZrBn
2
2
144
N
O
ZrBn
2
2
CF
3
CF
3
145
N
O
ZrBn
2
2
CF
3
146
N
O
N
O
Cl Cl
X
X
X
X
X = H, Br
147
Another example of a system with a five-membered chelating ligand is the iminephenoxy complex of Figure 60.
MAO activation leads to moderately high active catalysts while the molecular masses of the PEs are very high.
Activation with [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al results in slightly better activity and produces high molecular mass PEs.
Thus, the polymerization behavior of the iminephenoxy complex is totally different from that of the classical
phenoxyimine complexes, which exhibit higher activity with MAO than with [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al. In strict
analogy with the phenoxyimine complexes, NMR studies suggested that the Bu
i
3
Al reacts with the imine function-
ality and reduces the iminephenoxy complex to an aminephenoxy complex.
1136
The MAO-activated phenoxyketimine complexes of Figure 61 are highly active ethylene polymerization cata-
lysts, although activities are lower than those of analogous phenoxyaldimine complexes (R=H).
1166
In all cases,
rather low molecular masses (M

n
<50 000) are obtained. With the exception of the complex with R=CF
3
, rather
narrow molecular mass distributions (M

w
/M

n
<1.1) are found, indicative of living behavior. Polymerizations with
R=Ph at 0

C were run for different reaction times. Analysis of the resulting PEs showed a linear relationship
between M

n
and yield, which supported the living polymerization behavior. These experiments led to the conclusion
that the presence of ortho-F atoms in the complex is not a mandatory requirement for living behavior.
1166
R
H
Bu
t
25
4 10
4,167,000
25
9
3,679,000
T
P
(C)
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
M
v
1 bar ethylene, MAO/Ti = 250,[Ph
3
C][B(C
6
F
5
)
4
] /Ti = 1.2, Bu
i
3
Al / Ti = 50, toluene, 30 min
[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al
as co-catalyst
MAO
as co-catalyst
O
N
Ti(NMe
2
)
2
R
2
R
H u
t
25
2 10
2
3,231,000
25
6
2,000,000
75
580,000
75
1 10
2
1,130,000
T
P
(C)
M
v
75
200,000
75
190,000
B
1 10
2
4 10
2
2 10
2
Figure 59 Ethylene polymerization with phenoxypyridine-based catalysts.
1128 Olefin Polymerizations with Group IV Metal Catalysts
The MAO-activated u-enaminoketonato complexes of Figure 62 are moderately to highly active ethylene poly-
merization catalysts.
1186
However, molecular masses are not very high (M

n
<100 000). PEs produced with the first
two systems, that is, R
1
=CF
3
and R
2
=Ph or 2-theonyl, show rather narrow molecular mass distribution (M

w
/
M

n
<1.35), indicative of quasi-living behavior. The post-polymerization experiments indicated that the catalyst
lifetime is longer than 60 min even in the absence of ethylene. On the other hand, the third system, that is, R
1
=CF
3
and R
2
=2-furyl, produces PEs with rather broad polydispersities, although the only difference with the 2-theonyl-
based catalyst is the replacement of an S atom with an O atom.
O
N iCl
2
2
Ph
Bu
t
Bu
t
50
0.25
2,320,000
75
0.27
2,130,000
T
P
(C)
Activity (g
PE
(mmol Mt)
1
h
1
bar
1
)
Yield (g)
M
v
1 bar ethylene, MAO/Ti = 250, [Ph
3
C][B(C
6
F
5
)
4
]/Ti = 1.2, Bu
3
Al/Ti = 50, toluene
i
MAO
as co-catalyst
[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al
as co-catalyst
50
1 10
2
1 10
2
1 10
2
0.58
1,170,000
75
6 10
3
2.41
560,000
T
Figure 60 Ethylene polymerization with iminephenoxy-based catalysts.
R
Me Et
0.7 bar ethylene, MAO/Ti = 150, toluene, 20 C,
polymerization time: R = H,1 min; all other systems, 3 min
O
N
TiCl
2
Bu
t
2
Ph
Bu
t
Ph CF
3
R
410
3
0.42
48,500
1.07
610
2
0.20
24,000
1.06
510
2
0.15
21,300
1.04
110
3
0.39
44,500
1.08
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
n
M
w
/M
n
510
2
0.18
28,000
1.41
H
Figure 61 Ethylene polymerization with phenoxyketimine-based catalysts.
2
O
TiCl
2
N
R
1
Ph
R
2
R
1
=
Ph
CF
3
1 10
3
0.11
61,000
1.27
7 10
2
0.17
50,000
1.34
3 10
2
0.08
8,000
9.76
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
n
M
w
/M
n
25 C, 1 bar ethylene, MAO/ Ti = 2,000, toluene, 5 min
CF
3
CF
3
Me
Ph
Me
S
CF
3
O
R
2
=
1 10
3
0.27
61,000
2.56
1 10
2
0.05
Figure 62 Ethylene polymerization with u-enaminoketonato-based catalysts.
Olefin Polymerizations with Group IV Metal Catalysts 1129
The tridentate NOP complexes of Figure 63 showed good performance as ethylene polymerization catalysts.
Whereas the NOP complex with two chelating ligands showed no polymerization activity, the two monochelated
complexes exhibited high activity even at 100

C. Reasonably high molecular masses were obtained.


1187
Some authors tried to fix the stereoflexibility of bis(phenoxyimine) catalysts through the introduction of a bridge
between the two imine N atoms. The two phenoxyimine groups have been linked with a biaryl bridge. However,
the MAO-activated 148 exhibited no polymerization activity.
O
Zr
O
N
N
CH
2
Ph
CH
2
Ph
Bu
t
Me
Bu
t
Me
148
The main reason for this is that this complex decomposed through an intramolecular reduction of the imine
functionality by migration of one of the benzyl groups to the C atom of the imino moiety, as shown in Scheme 43.
1188
O
N
TiCl
2
PPh
2
O
N
TiCl
3
PPh
2
2
O
NH
TiCl
3
PPh
2
50
1 10
3
1.23
236,000
2.1
50
4 10
3
0.69
225,000
3.3
50
Trace
T
P
(C)
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
w
M
w
/M
n
Bu
t
Bu
t
Bu
t
Bu
t
Bu
t
Bu
t
100
5 10
2
0.52
73,000
2.4
100
2 10
3
0.36
77,000
1.5
Figure 63 Ethylene polymerization with tridentate NOP based catalysts.
O
Zr
O
N
N
CH
2
Ph
CH
2
Ph
Bu
t
Me
Bu
t
Me
O
Zr
O
N
N
CH
2
Ph
Bu
t
Me
Bu
t
Me
benzyl
migration
Scheme 43
1130 Olefin Polymerizations with Group IV Metal Catalysts
A solution to this problem was the introduction of a methyl group ortho to the imine group, as in complex 149, that
sterically protects the complex from benzyl migration. The MAO-activated 149 is remarkably stable, and overall
activity is moderate (about 50 g
PE
(mmol M)
1
h
1
bar
1
).
1188
O
Zr
O
N
N
CH
2
Ph
CH
2
Ph
Bu
t
Bu
t
Me
Me
149
Complexes 150 and 151, with cyclohexyl bridge between the imine groups, have also been tested. The catalyst
150/MAO polymerized ethylene with moderate activity (up to 30 g
PE
(mmol M)
1
h
1
bar
1
), while 151/MAO
exhibited low activity (10 g
PE
(mmol M)
1
h
1
bar
1
). In both cases, reasonably high molecular masses were obtained.
Perfluoroborate activation was found to be quite ineffective. These catalysts were substantially inactive for propylene
polymerization.
1189
O
Zr
O
N
N
CH
2
Ph
CH
2
Ph
Bu
t
Bu
t
Bu
t
Bu
t
O
Zr
O
N
N
Cl
Cl
Bu
t
Bu
t
Bu
t
Bu
t
150 151
While all the 6-coordinated NO complexes described so far have ligands with a negatively charged O and neutral N
atoms, a fewcomplexes with the Natomnegatively charged and with neutral Oatoms, such as 152, have been synthesized.
However, MAO activation of 152 polymerizes ethylene with rather low activity (3 g
PE
(mmol M)
1
h
1
bar
1
).
1190
Zr
N N
O
O
C
6
F
5
F
5
C
6
Cl
Cl
152
A remarkable class of catalysts with coordinating NO atoms is that based on the ligands of Scheme 44, introduced
for 1-hexene polymerization in 1999.
11911204
These catalysts differ from the bis(phenoxyimine) catalyst and analogs
N
X
R R
OH HO
N
R
OH
N
HO
R
Me Me
X = Me, NMe
2
, OMe, SMe, furan, THF
Scheme 44
Olefin Polymerizations with Group IV Metal Catalysts 1131
because the coordinating N atom is an amino N instead of an imino N atom. This family of catalysts can be divided
into two categories depending on the topology of the ligand (see Scheme 44). In the ligands on the left, the N atom
connects the two phenolate rings, and further coordination can occur through X donor atoms on the side-arm.
Conversely, in the tetradentate ligand on the right, the branched mode of connectivity (i.e., the side-arm) is
replaced with a consequential connectivity mode through the diamine bridge that connects the two phenolate rings.
In both cases, the final pre-catalysts are octahedral complexes with a trans-O, cis-N(X), cis-R geometry of
coordination at the metal atom, as shown in Scheme 45. Of course, in the case of X=Me, the n-Pr side-arm is not
coordinated to the metal, and the complexes present a 5-coordinated metal atom. The cis-disposition of the R groups
is of fundamental relevance for the ability of these complexes to act as polymerization catalysts.
Activation of these complexes with equimolar amounts of B(C
6
F
5
)
3
at room temperature resulted in active catalysts
for 1-hexene polymerization. The performance of some of the complexes with the extra side-arm is reported in
Table 20. First of all, it is important to note that complexes with no donor atom in the extra side-arm give low activity,
and only yield oligomers (entries 1 and 3 in Table 20).
1194,1195
Conversely, a donor group in the side-arm, for
example, an NMe
2
donor group, results in catalysts that present rather different behavior. For example, the Ti
complex of entry 2 in Table 20 exhibits a rather low polydispersity, indicative of quasi-living polymerization
behavior,
1195
while the Zr complex of entry 4 is remarkably active and shows no living character.
1199
Comparison
of the Zr and Hf series with donor groups NMe
2
, OMe, and SMe (Table 20; entries 4, 8, 12 for Zr, and 6, 10, 13 for Hf)
revealed no clear trend in either activity or molecular mass.
1199
In all the cases, reasonably high molecular mass
poly(1-hexene) was produced. The M

w
,M

n
values obtained were typical of single-site polymerization catalysts, and
no signs of living polymerization could be observed. Diluting the monomer in heptane results in more controlled
polymerization reactions, and in many instances the polymers produced exhibit remarkably high M

w
.
1199
Enclosing
the donor O atom in a THF or furan ring has no major impact on the polymerization behavior of the Zr- and Hf-based
catalysts (Table 20; entries 9, 15 for Zr, and 10, 16 for Hf). On the other hand, in the case of Ti replacement of the
THF ring by furan improves the activity to about 10
5
g
PH
(mmol M)
1
h
1
(entries 14 and 17).
The following general conclusions emerge. Catalyst activity mostly depends on the nature of the metal atom, with
the following approximate order: Zr _ Hf Ti. For the Zr- and Hf-based catalysts, very high activities in the range
10
4
10
5
g
PH
(mmol M)
1
h
1
and M

w.
of about 100 000 can be easily achieved. The Ti-based catalysts, on the
other hand, possess moderate activity, about 1030 g
PH
(mmol M)
1
h
1
. However, and very remarkably, while the
Zr and Hf catalysts yield M

w
,M

n
values around 2, the Ti systems can exhibit living character for prolonged periods of
time.
For the systems with an N-donor atom in the side-arm, the length of the side-arm or the nature of the N atom have
major effects on activity.
1196
In comparison, steric effects of substituents ortho to the phenolate O atoms, although
very close to the active site, have a much weaker influence. The most active catalysts contain a C
2
spacer between the
two coordinating N atoms, which corresponds to five-membered chelate rings (compare entries 4 and 20 as well as 22
and 23 in Table 20). Changing the hybridization of the extra donor atom from sp
3
to sp
2
or reducing the size of the
substituents on the aromatic rings had minor effects on reactivity (entries 4 and 22 vs. 4 and 18 in Table 20).
Increasing the bulkiness of the substituents on the N-donor reduced the reactivity dramatically and led to oligomers
(compare entries 4 and 21 in Table 20).
Another remarkable feature of these catalysts is their ability to promote living or quasi-living polymerization of
olefins. This behavior is exhibited by almost all the B(C
6
F
5
)
3
-activated Ti catalysts. At room temperature, the catalyst
with an NMe
2
group yields polymers with M

w
= 16500 and M

w
,M

n
= 1.18 after 30 min of polymerization.
1195
Insertion of the OMe group as donor results in much better living behavior, since the corresponding catalyst yields
polymers with M

w
= 445000 and M

w
,M

n
= 1.12 after 31 h of polymerization. The effect of the polymerization
temperature on the living behavior has been tested. One-hour polymerizations were carried out from room
O
M
R
N
R
N
O
M
R
X
R
N
trans-O
cis-N,X
cis-R
trans-O
cis-N
cis-R
O
O
Scheme 45
1132 Olefin Polymerizations with Group IV Metal Catalysts
Table 20 Selected results from bis(phenolate) complexes with an extra donor arm
Entry Ligand M Solvent
Productivity
(g
PH
(mmol
M)
1
h
1
) M

w
M

w
,M

n
References
1
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
Ti 1-Hexene 50 1,500 2.0 1195
2
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
NMe
2
Ti 1-Hexene 30 14,000 1.18 1195
3
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
Zr 1-Hexene 20 Oligomers 1194
4
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
NMe
2
Zr 1-Hexene 20,000 35,000 3.5 1199
5
N
OHHO
Bu
t
Bu
t
Bu
t
Bu
t
NMe
2
Zr n-C7 800 170,000 2.0 1199
6
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
NMe
2
Hf 1-Hexene 2,000 13,000 3.2 1199
7
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
NMe
2
Hf n-C7 200 20,000 1.6 1199
(Continued)
Olefin Polymerizations with Group IV Metal Catalysts 1133
Table 20 (Continued)
Entry Ligand M Solvent
Productivity
(g
PH
(mmol
M)
1
h
1
) M

w
M

w
,M

n
References
8
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
OMe
Zr 1-Hexene 50,000 80,000 3.0 1199
9
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
OMe
Zr n-C7 1,000 160,000 1.4 1199
10
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
OMe
Hf 1-Hexene 20,000 101,000 2.2 1199
11
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
OMe
Hf n-C7 700 140,000 1.8 1199
12
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
SMe
Zr 1-Hexene 9,000 195,000 2.0 1199
13
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
SMe
Hf 1-Hexene 30,000 66,000 2.5 1199
14
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
O
Ti 1-Hexene 10 316,000 1.05 1201
(Continued)
1134 Olefin Polymerizations with Group IV Metal Catalysts
Table 20 (Continued)
Entry Ligand M Solvent
Productivity
(g
PH
(mmol
M)
1
h
1
) M

w
M

w
,M

n
References
15
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
O
Zr PhCl 2,000 118,000 2.1 1201
16
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
O
Hf PhCl 1,000 60,000 1.5 1201
17
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
O
Ti 1-Hexene 200 550,000 1.37 1205
18
N
OH HO
Me Me
Me Me
NMe
2
Zr 1-Hexene High 100,000 1.9 1196
19
N
OH HO
Me Me
Me Me
NEt
2
Zr 1-Hexene 600 6,700 3.3 1196
20
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
NMe
2
Zr 1-Hexene Negligible Oligomers 1196
21
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
NEt
2
Zr 1-Hexene 60 1,100 1.6 1196
(Continued)
Olefin Polymerizations with Group IV Metal Catalysts 1135
temperature up to 65

C for the catalysts with an OMe donor group. The resulting polymers had M

w
in the range
15 00020 000, while the M

w
,M

n
increased from 1.07 to 1.30. Thus, even at rather high polymerization temperatures,
these catalysts yield narrow polydispersities.
1197
Exceptionally long-lived polymerization was achieved by introduc-
tion of a THF as donor group. Polymers with M

w
= 200000 and M

w
,M

n
= 1.06 were obtained after 2 days of
polymerization. Extending the polymerization time to 6 days still results in extremely narrow polydispersity
(M

w
,M

n
= 1.09), with polymer growth continuing almost linearly up to M

w
= 816000. Considering the extremely
long-living behavior of these catalysts, eventual chain-release reactions were attributed to technical sources such as
water or oxygen traces.
1201
Even small variations in the donor group can lead to substantially different catalytic
behavior. In fact, replacing the THF group with a furan group corresponds to polymers having M

w
>200 000 after 2 h
of polymerization, with very narrow molecular mass distribution M

w
,M

n
= 1.051.09).
1200
More detailed structure/activity relationships were obtained with the pre-catalysts 153155, as well as
156159.
1202
In the case of X=OMe, the B(C
6
F
5
)
3
-activated pre-catalysts 153155 exhibit rather similar activities,
in the range 1050 g
PH
(mmol M)
1
h
1
. The large t-Bu substituents make 153 more long-lived, with M

w
,M

n
= 1.12
after more than 24 h. By contrast, complexes 154 and 155, with smaller Me and Cl substituents, lose living character
after less than 1 h. The NMR analysis of the resulting poly(1-hexene)s indicated that the polymers were substantially
atactic and that a rather large fraction of regiomistakes was incorporated (4% for 153, 16% for 154, and 20% for 155).
Bz
Ti
N
X
O
Bz
O
X = OMe
Bu
t
Bu
t
Bu
t
Bu
t
153
Bz
Ti
N
X
O
Bz
O
X = OMe
154
Bz
Ti
N
X
O
Bz
O
X = OMe
Cl
Cl
Cl
Cl
155
The behavior of the complexes with X=NMe
2
is remarkably different. First of all, complexes 157 and 158, with
relatively small substituents ortho to the phenolate O atom, yield very high activities (1 kg
PH
(mmol M)
1
h
1
),
noticeably higher than that of the classical complex 156 with t-Bu groups (about 30 g
PH
(mmol M)
1
h
1
). Even
higher activity is exhibited by complex 159 (69 kg
PH
(mmol M)
1
h
1
), although the Cl groups are essentially similar
to Me groups from a steric point of view. It has been suggested that the very high activity of 159 is due to the
electron-withdrawing properties of the Cl atoms. The NMR analysis indicated that atactic polymers were also
produced in these cases. On the other hand, the amount of regiomistakes is noticeably smaller relative to the related
Table 20 (Continued)
Entry Ligand M Solvent
Productivity
(g
PH
(mmol
M)
1
h
1
) M

w
M

w
,M

n
References
22
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
N
Zr 1-Hexene 6,000 15,000 4.5 1196
23
N
OH HO
Bu
t
Bu
t
Bu
t
Bu
t
N
Zr 1-Hexene 60 102,000 1.7 1196
1136 Olefin Polymerizations with Group IV Metal Catalysts
complexes with an X=OMe group (1% for 156 and about 45% for 157159. The higher regioselectivity of the
complexes with X=NMe
2
can be reasonably ascribed to the additional Me group on the X donor atom that increases
the steric pressure on the transition state leading to a regiomistake. Interestingly, 157159 do not show characteristics
of living behavior. In fact, polymerization in chlorobenzene resulted in M

w
,M

n
in the range 1.52.0. However, while
M

w
in the range of 25 00050 000 was obtained with 157 and 158, exceptionally high molecular masses (M

w
up to
4 million in 1 h) were obtained with 159.
Bz
Ti
N
X
O
Bz
O
X = NMe
2
Bu
t
Bu
t
Bu
t
Bu
t
156
Bz
Ti
N
X
O Bz
O
X = NMe
2
157
Bz
Ti
N
X
O
Bz
O
X = NMe
2
158
Bz
Ti
N
X
O Bz
O
X = NMe
2
Cl
Cl
Cl
Cl
159
With regard to complexes with a tetradentate ligand, the C
2
-symmetric complexes 160163 were first
synthesized.
1193,1204
After activation with 1.1 equiv. of B(C
6
F
5
)
3
, 160 yields living and stereospecific 1-hexene
polymerization (activity =18 g
PH
(mmol M)
1
h
1
; reaction time =30 min; M

w
= 12000; M

w
,M

n
= 1.15). The
NMR analysis showed that the poly-1-hexene produced was highly isotactic.
1193
This was a major achievement
since the complexes with the extra donor on the side-arm only yield atactic polymers. Similar activation of 161
resulted in a slightly more active catalyst (activity =35 g
PH
(mmol M)
1
h
1
) but polymerization was not living
(M

w
= 23000; M

w
,M

n
= 1.57) and the resulting polymer was atactic.
1193
These results clearly indicated that the
living and isospecific behavior of these systems is strongly influenced by the groups ortho to the phenolate O atoms.
As expected, further substitution of the aromatic rings with Cl atoms, as in 162 and 163, resulted in much higher
activites (5000 and 200 g
PH
(mmol M)
1
h
1
, respectively).
1204
However, while the Zr-based complex 162 only yields
oligomers (M

w
= 9100), the Ti complex 163 yields very high molecular masses within 40 min, with some living
polymerization character (M

w
>550 000; M

w
,M

n
= 1.2). Extending the polymerization time to 19 h yielded exceed-
ingly high molecular masses (M

w
>1.9 million), but the living character was lost.
1204
Bn
Zr
N
N
O
Bn
O
Bu
t
Bu
t
Bu
t
Bu
t
160
Bn
Zr
N
N
O
Bn
O
Me
Me
Me
Me
161
Bn
Zr
N
N
O
Bn
O
Cl
Cl
Cl
Cl
162
Bn
Ti
N
N
O
Bn
O
Cl
Cl
Cl
Cl
163
The behavior in propylene polymerization of a series of catalysts based on the bis(phenoxyamino) skeleton has
been investigated. The pre-catalyst 160, activated with [Ph
3
C][B(C
6
F
5
)
4
] is able to polymerize propylene to an
isotactic polymer with a moderate activity (about 1 g
PP
(mmol M)
1
h
1
mol(propylene)
1
) in the temperature range of
050

C. The
13
C NMR analysis showed 80% mmmm pentads, with randomly distributed rr diads, typical of site
control. The mmmr/mmrr/mrrm pentad ratio of 2 : 2 : 1 further confirmed that the stereocontrol was determined by the
catalytic site. The iPP was remarkably regioregular, and analysis of the chain-end groups indicated that propagation is
primary. It has been suggested that 160 is a good model of the isotactic species active in classical ZieglerNatta
polymerization.
236
Once again, steric hindrance by the alkyl group ortho to the phenoxy O atom was found to be the
key to the isotactic stereocontrol. In fact, 161, with Me groups in place of t-Bu groups, yields a slightly syndiotactic
Olefin Polymerizations with Group IV Metal Catalysts 1137
PP, and stereoselectivity is chain-end controlled.
236
The effect of the alkyl group ortho to the phenoxy O atom has
been investigated using the complexes 164166. Complex 164 activated with [Ph
3
C][B(C
6
F
5
)
4
] exhibited remark-
able iso- and regiospecificity (98.5% mmmm, 0.4% 2,1-regiomistakes), and the resulting iPP has a T
m
of 151

C.
Activity of 164 is moderate (4 g
PP
(mmol M)
1
h
1
mol(propylene)
1
) but it is significantly higher for catalyst 165
(70 g
PP
(mmol M)
1
h
1
mol(propylene)
1
). However, the latter system results in lower stereocontrol (mmmm 89%).
By contrast, catalyst 166 resulted in substantially atactic PP. Based on the quantum mechanics/molecular mechanics
calculations, the poor performance of 166 has been ascribed to the flat nature of the anthracenyl group, which results
in ineffective space occupation around the metal. Another important achievement is the increase in M

n
and in the
average chain-growth time obtained with the bulky adamantyl groups in 164 (M

n
about 110 000 and chain-growth
time about 9 h) relative to 160 (M

n
about 4000, chain growth time <0.5 h).
1206
Bn
Zr
N
N
O
Bn
O
Me
Me
164
Bn
Zr
N
N
O
Bn
O
Me
Me
165
Bn
Zr
N
N
O
Bn
O
Me
Me
166
According to the quantum mechanics/molecular mechanics calculations, the opposite regiochemistry exhibited in
propylene polymerization by bis(phenoxyimine) Ti and bis(phenoxyamine) Zr catalysts (secondary and primary,
respectively) is the result of both steric and, mainly, electronic effects. As shown in Scheme 46, an antibonding interaction
occurs between the molecular orbital of the incipient MC bond and the lone electron pair of the N atom trans to it. This
trans-effect is stronger for the bis(phenoxyamine) systems because the molecular orbital corresponding to the lone electron
pair of theamine Natomis higher inenergy thanthe corresponding orbital of the imine Natom. Inthe case of the secondary
propylene insertion, the donating effect of the methyl group on the electron density of the incipient MC bond is larger,
which results in turn in a stronger antibonding interaction that disfavors the secondary propylene insertion.
165
4.09.5.3.6 Complexes with NN chelate ligands
Several catalysts with coordinating NN chelate ligands have been proposed. For example, the iminopyrrolate systems
of Figure 64 have been investigated by several authors.
1086,12071213
The bis(iminopyrrolate) Ti complexes of Figure 64
polymerize ethylene with high activity and yield high molecular masses. Crystal structures of the complexes reveal that
the dichloride complexes adopt a cis-Cl, cis-N(imine), and trans-N(pyrrolide) geometry. MAO and [Ph
3
C][B(C
6
F
5
)
4
]/
Bu
i
3
Al activation of these complexes results in activities comparable to those shown by similarly activated bis(phenoxy
imine) Ti catalysts. For these systems, too, [Ph
3
C][B(C
6
F
5
)
4
] activation yields significantly higher molecular masses.
1210
MAO-activated iminopyrrolate Zr complexes such as 167 also polymerize ethylene, although the activity is
moderate (20 g
PE
(mmol M)
1
h
1
bar
1
). Average molecular mass is M

w
= 94 000 and the very broad polydispersity
(M

w
,M

n
= 28.4) suggests multiple active species. Interestingly, the X-ray structure of this complex is C
1
-symmetric
and reveals one benzylic group trans to a pyrrolate N atom, while the other benzylic group is trans to an imine N
atom.
1086,1214
The related MAO-activated mono-iminopyrrolate pre-catalyst 168 also exhibits moderate ethylene
polymerization activity.
1214
Propylene Propylene
Growing
Chain
Imine N atom Amine N atom
N N
O
O
O
O
C C
C
C
Mt Mt
C
C
Growing
Chain
N N
Scheme 46
1138 Olefin Polymerizations with Group IV Metal Catalysts
Zr
N
N
CH
2
Ph
N
CH
2
Ph
N
Ar
Ar
Ar = 2,6-Pr
i
2
-C
6
H
3
167
Zr
N
N
Cl
Cl
Cl
Cl Ar
Ar = 2,6-Pr
i
2
-C
6
H
3
N
Ar
Li
OEt
2
OEt
2
168
The Zr- and Hf-based iminopyrrolide catalysts of Figure 65 have also been investigated. While the complex with
R
1
=4-Bu
t
-C
6
H
4
is not active, the other two systems with R
1
=Bu
t
and adamantyl are active ethylene polymerization
catalysts. For the Hf compounds, MAO and B(C
6
F
5
)
3
/Bu
i
3
Al activation results in similar activities and rather broad
MWDs. In the case of the Zr catalysts, MAO activation gives slightly better activities, and the PEs had polydisper-
sities typical of single-site catalysts. It has been suggested that the Hf catalysts produced multimodal PE by gradual
generation of the active species and/or isomerization of the active species under the polymerization conditions.
1212
More surprisingly, in the case of M=Hf and R
1
=Bu
t
and using [Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al as co-catalyst, solvent
effects seem to be remarkable. Switching from toluene to n-heptane yields high molecular mass PE (M

w
= 896000)
with a very broad M

w
,M

n
of 21.
1211
A quantum mechanics/molecular mechanics study has indicated that in the
presence of a tigtly bound counterion such as [MeB(C
6
F
5
)
3
]

, the rate-limiting step of chain growth is probably


ethylene coordination. The barrier for chain growth in the Ti-based catalysts is predicted to be around 13 kcal mol
1
.
A slightly lower barrier, about 9 kcal mol
1
, has been predicted for the Zr analog.
151
The bis(indolideimine) Ti complexes of Figure 66 polymerize ethylene with moderate to high activities. The
best performance is shown by the perfluoro-substituted complex. These complexes exhibit living polymerization
behavior. Differently from bis(phenoxyimine) Ti-based catalysts, living behavior is also shown by the system with
no F atoms on the R
1
substituent. It was hypothesized that the bulkiness and rigidity of the indolideimine ligand is
responsible for suppressing chain-termination reactions.
1160,1215
Catalysts that contain mono(benzamidinate) or bis(benzamidinate) ligands have been investigated.
247,248,276,1216
The bis(benzamidinate) complexes assume an octahedral geometry of coordination at the metal atom, as shown in
Scheme 47. These complexes are C
2
-symmetric, and coordination of the benzamidinate ligands leaves two mutually
cis-coordination positions. Thus, these complexes possess the requirements for isospecific propylene polymerization.
The only drawback of these structures is the lack of stereorigidity, usually provided by a tether that bridges the two
chelating ligands. In principle, there might be fast isomerization between and configurations, as shown in
R
1
C
2
H
5
n-C
6
H
13
6 10
3
0.50
75,000
4 10
2
0.03
412,000
8 10
2
0.07
441,000
1 10
4
1.18
2,601,000
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
v
25 C,1 bar ethylene, MAO/Ti= 1,250, [Ph
3
C][B(C
6
F
5
)
4
]/Ti = 6, Bui
3
Al/Ti = 250, toluene, 5 min
N
N iCl
2
R
1
2
2 10
3
0.16
4,739,000
2 10
3
0.14
4,654,000
1 10
3
0.13
4,901,000
2 10
3
0.17
4,029,000
Yield (g)
M
v
[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al as cocatalyst
MAO as co-catalyst
T
Figure 64 Ethylene polymerization with bis(iminopyrrolate) Ti-based catalysts.
Olefin Polymerizations with Group IV Metal Catalysts 1139
B(C
6
F
5
)
3
/Bu
i
3
Al
Trace
MAO
Trace
Co-catalyst
Co-catalyst
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
w
M
w
/M
n
N
N (CH
2
Ph)
2
R
1
2
R
1
Bu
t
B(C
6
F
5
)
3
/Bu
i
3
Al
2 10
3
0.93
88,000
9.19
MAO
1 10
3
0.53
87,000
6.05
B(C
6
F
5
)
3
/Bu
i
3
Al
2 10
3
0.87
96,000
12.15
MAO
6 10
2
0.26
36,000
6.66
M = Hf
B(C
6
F
5
)
3
/Bu
i
3
Al
Trace
MAO
Trace Yield (g)
M
w
M
w
/M
n
B(C
6
F
5
)
3
/Bu
i
3
Al
7 10
2
0.31
24,000
2.25
MAO
8 10
3
3.37
29,000
2.25
B(C
6
F
5
)
3
/Bu
i
3
Al
2 10
2
0.73
63,000
2.05
MAO
9 10
3
3.67
87,000
2.08
M = Zr
M
5.0 mol Complex , 25 C, 1 bar ethylene, MAO/Hf = 250, B(C
6
F
5
)
3
/Hf = 2, Bu
i
3
Al/Hf = 50, toluene, 5 min
Figure 65 Ethylene polymerization performances of Hf- and Zr-based pyrrolideimine-based catalysts.
5 10
0.04
13,800
1.14
6 10
0.05
12,400
1.13
3 10
2
0.24
46,500
1.11
1 10
3
0.95
323,000
1.93
Activity (g
PE
(mmol M)
1
h
1
bar
1
)
Yield (g)
M
w
M
w
/M
n
25 C, 1 bar ethylene, MAO/Ti = 250, toluene, 10 min
N
N iCl
2
R
1
2
MAO as co-catalyst
F F
F
F
F
F
F
F
F
F
R
1
T
Figure 66 Ethylene polymerization with indolideimine-based catalysts.
Zr
N
N
Cl
Cl
N
N
R
1
R
1
R
1
R
1
R
2
R
2
Zr
N
N
Cl
Cl
N
N
R
1
R
1
R
1
R
1
R
2
R
2
Inversion of
configuration
-Site -Site
Scheme 47
1140 Olefin Polymerizations with Group IV Metal Catalysts
Scheme 47, which would nullify any stereoinduction which might be operative at each insertion step. These systems
exhibit reasonable catalytic activity only under propylene pressure (5 bar).
At 20

C and at 5 bar propylene feed, the MAO-activated mono(benzamidinate) 169 yields poorly stereoregular iPP
(mmmm~2530%) with the moderate activity of about 20 g
PP
(mmol M)
1
h
1
bar
1
, and M

v
= 100000. Raising the
polymerization temperature to 60

C, and increasing the propylene pressure to 7 bar, gives very low activity
(1 g
PP
(mmol M)
1
h
1
bar
1
) that only yields oligomers (M

v
= 28000).
276
N
N
SiMe
3
TiCl
3
(THF)
169
By contrast, at 25

C and at propylene pressures 5 bar the MAO-activated benzamidinato complex 170 in CH


2
Cl
2
yields substantially isotactic polymers (mmmm~98%, M

v
in the range 20,00080,000). Activity is moderate, about 1050
g
PP
(mmol M)
1
h
1
bar
1
. The Ti analogue 171, on the other hand, yields a substantially elastomeric PP (mmmm
~20%) in the same conditions.
248
Removing the Me substituent in the para position of the phenyl ring has beneficial
effects on both activity and molecular masses (up to M

w
= 2 million with 172, and up to 300,000 with 173). The
resulting PP are poorly stereoregular (mmmm in the range 10%30%), although they possess elastomeric properties.
The structures of these elastomeric polymers were suggested to be consistent with longer isotactic domains disposed
between domains with a large number of stereodefects.
248
In an attempt to explore the effects of C
1
-symmetry in this
class, Eisen and coworkers synthesized complex 174. MAO activation of 174 yields highly stereoregular iPP (mmmm
in the range 93%99%) with very low activity (1 g
PE
(mmol M)
1
h
1
bar
1
) and M

v
= 1000030000
.
276
N
N
SiMe
3
SiMe
3
ZrMe
2
2
170
N
N
SiMe
3
SiMe
3
TiMe
2
2
171
N
N
SiMe
3
SiMe
3
ZrMe
2
2
172
N
N
SiMe
3
SiMe
3
TiMe
2
2
173
N
N
SiMe
3
TiCl
2
2
174
All the iPPs produced with this class of complexes had M

w
,M

n
around 2, which is consistent with the single-site
behavior. The NMR analysis of the iPP revealed equal amounts of mmmr and mmrr pentads, consistent with a
mechanism of enantioselectivity controlled by the site chirality.
276
To explain the increased isotacticity at increasing
propylene pressure, it has been proposed that monomer insertion competes with chain epimerization. At increased
propylene pressure, chain-epimerization reactions would be depressed by faster monomer insertion.
247,248,276
Other 6-coordinated pre-catalysts with chelating NN atoms have been tested in olefin polymerization. However,
complexes such as 175 and 176 yield PE with low to very low activity.
1217,1218
N
N
Zr
N
N
Bn
Cl
175 176
N
N
N
Cl Cl
X = H, Me
N
Zr
X
X
Olefin Polymerizations with Group IV Metal Catalysts 1141
Ti and Zr complexes with the tris(pyrazolyl)borate skeleton such as 177 and 178 exhibited reasonably high activity
(<10
3
g
PE
(mmol M)
1
h
1
bar
1
).
1219
Bulky mesityl groups on the tris(pyrazolyl)borate skeleton, as in the Zr
complexes 179 and 180, lead to very highly active catalysts for ethylene polymerization.
278,1220
The MAO-activated
179 and 180 yield PE with activity around 10
5
g
PE
(mmol M)
1
h
1
bar
1
. This activity is comparable to that of
bis(phenoxyimine)-based catalysts. Furthermore, 179 and 180 exhibit this activity at 75

C and for at least 20 min.


The resulting PE had remarkably high molecular mass (M

v
up 8.5 10
6
), while polydispersities around 2 indicated
single-site behavior.
278
The Ti analog of 180 possesses rather low activity.
279
These catalysts have also been
supported on silica.
1221
177
N N
N ZrCl
3
N
HB
N
N
N N
N ZrCl
3
N
N HB
N
178
N
N N
HB
N
N ZrCl
3
N
179
N
N N
ZrCl
3
N
N HB
N
180
A few other complexes with NN coordinating atoms have been tested as olefin polymerization catalysts. The
aminopyridinato complex 181 is only able to oligomerize propylene and 1-butene (M

n
<3500) with low activity.
1222
The u-diketiminato Zr complex 182 exhibits moderate ethylene polymerization activity (~50 g
PE
(mmol
M)
1
h
1
bar
1
), but molecular masses were disappointingly low (M

n
up to 30 000).
1117
Finally, at 100

C, the
MAO-activated Ti complex 183, with a tridentate triazacyclononane ligand, yields PE with very high activity
(10
4
g
PE
(mmol M)
1
h
1
bar
1
) but rather low molecular masses and broad polydispersity (M

n
= 39 500;
M

w
,M

n
= 7.0).
1223
N
Ti
N
Me
3
Si
NEt
2
NHEt
2
Cl
Cl
181
N
ZrCl
2
N
2
Ph
Ph
182
Ti
N
Cl N
Cl
N N
Me Me
Me
Bu
t
183
4.09.5.3.7 Other ligands
Some six-coordinate complexes have been reported that fall outside the previous sections. The catalyst 184/MAO
yields polypropylene with low activity (5 g
PP
(mmol M)
1
h
1
bar
1
) and moderate molecular mass (M

n
up to
105 000). The resulting PP is slightly isotactic (mmmm 35%) and exhibits elastomeric properties.
1224
The catalyst
185/[HNMe
3
][B(C
6
F
5
)
4
]/AlBu
i
3
yields PE with very high activity (5 10
4
g
PE
(mmol M)
1
h
1
bar
1
), rather high
molecular masses (M

n
about 10
6
), and low polydispersity. It also yields poly(1-hexene) with moderate activity
(10 g
PH
(mmol M)
1
h
1
) but with high molecular masses (M

v
~500 000).
1225
The MAO-activated Ti complex 186,
with a crown-type macrocyclic phosphine ligand, yields PE with moderate activity (30 g
PE
(mmol M)
1
h
1
bar
1
).
1226
TiCl
2
N
P
P
Ph
Ph
Ph
Ph
2
184
Zr(O-C
6
F
5
)
2
N
C
N
2
Me
Ph
185
Ti
P P
P
Cl Cl
Cl
186
1142 Olefin Polymerizations with Group IV Metal Catalysts
4.09.5.3.8 Olefin co-polymerizations with post-metallocene catalysts
4.09.5.3.8.(i) Ethylene/propylene co-polymers
Post-metallocene catalysts have been extensively used for the co-polymerization of olefins.
605,1139,1175,1186,1187,12271231
For example, the 187/[Ph
3
C][B(C
6
F
5
)
4
]/Bu
i
3
Al catalyst is capable of co-polymerizing ethylene and propylene at
70

Cunder 9 bar of total ethylene/propylene pressure with moderate activity, 70 g


co-polymer
(mmol M)
1
h
1
bar
1
.
The result is an amorphous ethylene/propylene co-polymer having a propylene content of 23.7mol%. The GPC analysis
indicated that the co-polymer is of extremely high molecular mass and it is produced by a single active species
(M

w
= 10.2 10
6
; M

w
,M

n
= 2.52). The
13
C NMR spectroscopy revealed that 187 yields a substantially alternating
co-polymer, indicating that the catalyst favors ethylene rather than propylene (EEE: 46.3mol%; EEP: 28.2 mol%; PEP:
5.2 mol%; EPE: 18.5mol%; PPE: 1.8 mol%; PPP: 0).
1228
187
O
N rCl
2
SiEt
3
2
Me
OMe
Z
The MAO-activated complexes 188191 have been also tested in ethylene/propylene co-polymerization.
1139
The
perfluorinated complexes 188 and 189 resulted in somewhat higher activities and molecular masses than the
corresponding non-fluorinated complexes 190 and 191 (188: M

w
= 64000; 189 M

w
= 193000; vs. 190:
M

w
= 4000; 191: M

w
= 9000). Moreover, the fluorinated complexes 188 and 189 incorporate propylene better
than the non-fluorinated complexes 190 and 191 ((188: P=12.9 mol% 189: P=9.5 mol% vs. 190: P=6.0 mol%, 191:
P=7.2 mol%). This finding has been correlated to the higher electrophilicity of the metal in the fluorinated
complexes.
1139
188
O
N rCl
2
C
6
F
5
Bu
t
2
Z
190
O
N rCl
2
C
6
H
5
Bu
t
2
Z
189
O
N fCl
2
C
6
F
5
Bu
t
2
H
191
O
N fCl
2
C
6
H
5
Bu
t
2
H
The MAO-activated 138 and 137, effective in the living homopolymerization of ethylene and propylene, also
promote their block co-polymerization.
1175,1227
This approach broadens remarkably the utility of living catalysts
because it allows the preparation of block co-polymers with high glass or melting transition blocks from common
commercial monomers such as ethylene and propylene. These materials could have applications as compatibilizers
and elastomers.
1232,1233
Using complex 138, propylene has been first homopolymerized to sPP for 2 h in toluene at
0

C (M

n
= 38400; M

w
,M

n
= 1.11). Then, an ethylene overpressure was applied, and in 1 additional h an
sPP-block-poly(E-co-P) diblock co-polymer was obtained (M

n
= 145 100, M

w
,M

n
= 1.12). The microstructure of
this diblock co-polymer is shown in Scheme 48. This co-polymer has a T
m
of 131

C while the ethylenepropylene


block (E=33 mol%) has a T
g
of 45

C.
1175
A detailed morphological and thermodynamic characterization of these
co-polymers has been reported.
1234
Using the MAO-activated complex 137, an ethylene/propylene monodisperse co-polymer with a propylene content
in the range 5015 mol% has been synthesized.
1227
Details are shown in Figure 67.
Olefin Polymerizations with Group IV Metal Catalysts 1143
Catalysts based on complexes 150 and 151 have also been tested for ethylenepropylene co-polymerizations but
poor activity was reported. The NMR analysis of the co-polymers indicated random insertion of the monomers, with
no stereoregularity in the propylene homosequences.
1189
Catalysts based on the Hf pyridyl amine complexes 126128 have been used for the preparation of ethylene/
propylene co-polymers as well as of ethylene/propylene/1-octene terpolymers. These co-polymers are character-
ized by having at least 60 wt% propylene units, M

w
around 300 000, and M

w
,M

n
in the range 2.02.4. The NMR
analysis of these co-polymers showed that the propylene sequences are remarkably isotactic (mm90%) and
showed the presence of regioirregularly inserted propylene units(<0.5%mol). The most interesting property of
catalysts based on 126128 is their high thermal stability.
1119
Using modifications of isospecific bis(phenoxy
amine)-based catalysts, such as complex 164, the controlled synthesis of iPP-block-poly(E-co-P) diblock co-poly-
mers has been achieved. This is a remarkable result since iPP and PE are both polymeric materials of extreme
industrial relevance.
1206
4.09.5.3.8.(ii) Ethylene/higher-c-olefin co-polymers
Using perfluorinated bis(phenoxyimine) Ti catalysts, and techniques developed for ethylenepropylene co-poly-
merizations,
1139,1227,1228
the random and block co-polymerization of ethene with higher c-olefins as 1-hexene,
1-octene, and 1-decene has been reported.
1231
The diamide-based catalyst 94/MMAO was tested in ethylene1-
butene copolymerization at 70

C. Activity and 1-butene incorporation were rather good (3001000 g


co-polymer
(mmol
M)
1
h
1
bar
1
; 1-butene 1240 mol%), but rather low molecular masses and fairly broad polydispersities were
obtained (M

n
=700033 000; M

w
,M

n
=4.06.4). The
13
C NMR analysis of the random poly(ethylene-co-1-butene)
showed regioselective insertion of 1-butene.
1078
The tridentate NOP complexes of Figure 63 promote the random co-polymerization of ethylene1-hexene with
high activity. 1-Hexene incorporation was in the range 540 mol%.
1187
Catalysts based on complexes 150 and 151
have been tested for ethylene co-polymerization with 1-hexene. As for ethylenepropylene, the co-polymerization
activity was poor.
1189
Ethylene1-hexene co-polymers have been obtained with catalyst 179/MAO.
278
This
catalyst exhibits very high activity (about 2 10
5
g
copolymer
(mmol M)
1
h
1
) and good co-monomer incorporation
(1-hexene 1826 mol%). The resulting co-polymers have high molecular mass (M

w
up 1.5 million) with
M

w
,M

n
= 2. The NMR analysis of the co-polymer indicated a tendency toward alternating ethylene and 1-
hexene insertions, and no sign of head-to-head regiomistakes. The resulting co-polymers are amorphos, with T
g
slightly below 0

C.
278
Ethylene1-octene co-polymers have also been prepared with bis(phosphinimide) com-
plexes such as 101.
1089,1235
m n-x n x
Scheme 48
5/95
20
0.95
19,600
1.07
47.9
E/P feed
Time (min)
Yield (g)
M
n
M
w
/M
n
P (mol%)
25 C, 1 bar E/P, toluene, MAO/Ti = 100
30/70
10
2.69
42,500
1.08
28.9
80/20
5.3
5.25
83,600
1.13
14.7
10/90
30
3.55
52,300
1.11
38.2
50/50
5
2.85
42,700
1.08
21.1
Figure 67 Ethylene/propylene co-polymerization results using the MAO-activated complex 137.
1144 Olefin Polymerizations with Group IV Metal Catalysts
4.09.5.3.8.(iii) Co-polymers of ethylene with internal alkenes
The co-polymerization of ethylene and 2-butene with the McConville catalyst 87 has been reported. NMR analysis
of the co-polymer indicated that only the trans-isomer of 2-butene was incorporated.
1236
Ethylene and cylopentene
have been co-polymerized with the MAO-activated complex 138 working at very low ethylene pressure.
605
The
almost regularly alternating poly(ethylene-alt-cis-1,2-cyclopentene) co-polymer exhibited a T
g
of 10

C. This co-
polymer is a crystalline plastomer with high structural disorder.
1237
Furthermore, multi-block co-polymers were
synthesized by changing the ethylene pressure during polymerization. This way, the various blocks are characterized
by a different amount of cyclopentene units.
605
The co-polymerization of ethylene and NB has been extensively investigated.
1186,1187,1230
Using the u-enamino-
ketonato-based quasi-living catalysts of Figure 62, almost regularly alternating E/N co-polymers can be synthesized
(N content in the range of 3549%). Activities are substantially similar to those observed in the ethylene homo-
polymerization.
1186
The GPC analysis gave M

n
values in the range 150 000570 000, while some character of quasi-
living behavior was retained up to reaction times of 20 min (M

w
,M

n
=1.171.35). The NMR analysis of the resulting
co-polymers indicated the presence of similar amounts of iso- and syndiotactic NENEN sequences. The T
g
of the co-
polymers was in the range 121135

C. The quasi-living behavior of these catalysts was also used for the preparation
of PE-block-poly(E-co-N) diblock co-polymers. After 5 min of ethylene homopolymerization, NB was added to the
reaction feed for an additional 5 min. GPC profiles indicated the synthesis of a diblock co-polymer with
M

w
,M

n
= 1.38. The diblock co-polymer possesses an NB content of 11 mol% and a T
m
of 137

C.
1186
Random
ethylene/norbornene co-polymerization with the tridentate NOP-based complexes of Figure 63 gave NB incorpora-
tions in the range of 435 mol%.
1187
The bis(pyrrolideimine) Ti complexes of Figure 64 co-polymerize ethylene and NB with high activities.
1230
Changing the N/E ratio from 0.3/1 to 3/1 resulted in higher N inclusion (26 vs. 45 mol%) and lower M

n
(652 000 vs.
285 000) and M

w
,M

n
(1.28 vs. 1.11). NMR analysis revealed an almost regularly alternating structure even at 45 mol%
N incorporation. Isolated N units (E-N-E
n _2
) (3.7 mol%) and NN sequences (0.9 mol%) were however detected.
Almost equal amounts of alternating iso- and syndiotactic sequences were also found. NMR analysis of low molecular
mass co-polymers showed the presence of CH
3
-N initiation sequences, while no peaks assignable to the CH
3
-E
initiation sequences were observed. Regarding the termination steps, N-E sequences were found, while no peaks
assignable to E-E or E-N sequences were observed. The quasi-living nature of the pyrrolideimine catalysts was
employed to synthesize PE-block-poly(E-co-N) as well as poly(E-co-N)
1
-block-poly(E-co-N)
2
diblock co-polymers. In
the latter, each co-polymer block contains a different amount of N units. These co-polymers had M

n
in the range
424 000783 000, with polydispersities in the range of 1.231.66. The resulting diblock co-polymer had two glass
transitions. For example, a co-polymer with a content of N=19 and 30 mol% in the two blocks showed a T
g
of 17

C
for the first block and of 123

C for the second block.


Finally, ethylene/norbornene co-polymerization with the dicarbollide Ti and Zr complexes has been also reported.
1238
4.09.5.3.8.(iv) Propylene/higher-c-olefin co-polymers
Complex 87 was used for propylene1-hexene co-polymerization. It has been suggested that blocks of iPP and poly-
1-hexene were joined to form a poly(iPP-co-iPH) block co-polymers.
1070
Propylene and 1,5-hexadiene have been co-
polymerized with the 138/MAO catalyst.
1229
The NMR spectroscopy indicated that the propylene homosequences
are remarkably syndiotactic. As in the case of hexadiene homopolymerization with the same catalyst (see Figure 55),
the hexadiene is incorporated as methylene-1,3-cyclopentane or 3-vinyl-tetramethylene units. Lower propylene
concentration favored the formation of methylene-1,3-cyclopentane units, while higher propylene concentration
favored 3-vinyl-tetramethylene units. Furthermore, an sPP-block-poly(P-co-HD) diblock co-polymer was synthesized
by first polymerizing propylene to sPP for 2 h and then adding 1,5-hexadiene for 6 additional h.
1229
4.09.5.3.9 Polystyrene and olefinstyrene co-polymerization with post-metallocene catalysts
Isotactic PS was discovered in 1955, and until a few years ago it was synthesized almost exclusively with heterogeneous
ZieglerNatta catalysts.
1239
Recently, it has been shown that homogeneous post-metallocene catalysts based on the
MAO-activated bis(thiophenolate) complexes 192 and 193 can be used for the isospecific polymerization of styr-
ene.
12401243
The Ti-based catalyst exhibited very high activity (more than 1 10
3
g
PS
(mmol M)
1
h
1
) and the
molecular mass of the resulting isotactic PS is remarkably high, while the molecular mass distribution indicates the
single-site behavior (M

n
= 2 654 000; M

w
,M

n
= 2.0). The Zr and Hf analogs are notably less active and yield rather
low molecular masses. Introduction of a longer spacer between the two S atoms such as in 194 had detrimental effects.
Olefin Polymerizations with Group IV Metal Catalysts 1145
O
Ti
O
S
S
Cl
Cl
Bu
t
Bu
t
Bu
t
Bu
t
192
O
Zr
O
S
S
CH
2
Ph
CH
2
Ph
Bu
t
Bu
t
Bu
t
Bu
t
193
O
S
S
Cl
Cl
Bu
t
Bu
t
Bu
t
Bu
t
Ti
O
194
The catalyst system 192/MAO has been used for the synthesis of multi-block propylenestyrene co-polymers. The
resulting co-polymers have a propylene content in the range 133 mol%, and the molecular masses are in the range
M

w
= 210060000. Lower M

w
values were observed for higher propylene content. Interestingly, NMR analysis
indicated that the regiochemistry of insertion of the two monomers is opposite. In particular, the iPP blocks are
formed by primary insertion of propylene, whereas the iPS blocks are formed by secondary insertion of styrene.
1244
Effective ethylenestyrene co-polymerization with the post-metallocene catalysts has been achieved with the
pentacoordinated complexes 195 and 196.
159,1245
Pseudo-random ethylenestyrene co-polymers (i.e., co-polymers
with no regioregularly arranged styrenestyrene units)
642
were synthesized. The synthesis of an alternating crystal-
line co-polymer with isotactic styrene units and a T
m
of 116

C was also claimed. In these first studies, the importance


of the S atom was noted, since the CH
2
-bridged complex 197 showed no co-polymerization activity. However, further
studies demonstrated that the S atom is not fundamental, since complex 198 also promotes ethylenestyrene
co-polymerization.
627,1246
Moreover, it was shown that complexes 195 and 196 incorporate a rather minor amount
of styrene units (about 610 mol%) in the co-polymer relative to complex 198 (about 35 mol%), but they are roughly
12 order of magnitude more active than 198.
S
O
O
Ti(OPr
i
)
2
Bu
t
Bu
t
195
S
O
O
Ti(OPr
i
)
2
Bu
t
Bu
t
O
196
O
O
Ti(OPr
i
)
2
Bu
t
Bu
t
197
O
O
Ti(OPr
i
)
2
Bu
t
Bu
t
198
References
1. Wilkinson, J.; Birmingham, J. M. J. Am. Chem. Soc. 1954, 76, 4281.
2. Natta, G.; Pino, P.; Mazzanti, G.; Giannini, U. J. Am. Chem. Soc. 1957, 79, 2975.
3. Breslow, D. S.; Newburg, N. R. J. Am. Chem. Soc. 1957, 79, 5072.
4. Schnutenhaus, H.; Brintzinger, H.-H. Angew. Chem. 1979, 91, 837.
5. Smith, J. A.; Brintzinger, H. H. J. Organomet. Chem. 1981, 218, 159.
6. Wild, F. R. W. P.; Zsolnai, L.; Huttner, G.; Brintzinger, H. H. J. Organomet. Chem. 1982, 232, 233.
7. Schwemlein, H.; Brintzinger, H. H. J. Organomet. Chem. 1983, 254, 69.
8. Schwemlein, H.; Zsolnai, L.; Huttner, G.; Brintzinger, H. H. J. Organomet. Chem. 1983, 256, 285.
9. Wild, F. R. W. P.; Wasiucionek, M.; Huttner, G.; Brintzinger, H. H. J. Organomet. Chem. 1985, 288, 63.
10. Wochner, F.; Zsolnai, L.; Huttner, G.; Brintzinger, H. H. J. Organomet. Chem. 1985, 288, 69.
11. Wochner, F.; Brintzinger, H.-H. J. Organomet. Chem. 1986, 309, 65.
12. Ro ll, W.; Zsolnai, L.; Huttner, G.; Brintzinger, H.-H. J. Organomet. Chem. 1987, 322, 65.
13. Scha fer, A.; Karl, E.; Zsolnai, L.; Huttner, G.; Brintzinger, H.-H. J. Organomet. Chem. 1987, 328, 87.
14. Sinn, H. J.; Kaminsky, W. Adv. Organomet. Chem. 1980, 18, 99.
15. Cecchin, G.; Morini, G.; Piemontesi, F. ZieglerNatta Catalysts. In Kirk-Othmer Encyclopedia of Chemical Technology; Wiley: New York, 2003.
1146 Olefin Polymerizations with Group IV Metal Catalysts
16. Chadwick, J. C. ZieglerNatta Catalysis. In Encyclopedia of Polymer Science and Technology; WileyNew York, 2003.
17. Keii, T. Heterogeneous Kinetics: Theory of ZieglerNattaKaminsky Polymerization; Springer: Berlin, 2004; Vol. 77.
18. Benedikt, G. M., Goodall, B. L., Eds. Metallocene-catalyzed Polymers: Materials, Properties, Processing & Markets, Plastics Design Library: New
York, 1998.
19. Scheirs, J., Kaminsky, W., Eds. Metallocene-based Polyolefins. Preparation: Properties and Technology; Wiley: New York, 2000.
20. Gladysz, J. A., Ed. .Chem. Rev. 2000, 100(4), 2000; .
21. Dyachkovskii, F. S.; Shilova, F. S.; Shilov, A. K. J. Polymer Sci., Part C 1967, 16, 2333.
22. Eisch, J. J.; Piotrowski, A. M.; Brownstein, S. K.; Gabe, E. J.; Lee, F. L. J. Am. Chem. Soc. 1985, 107, 7219.
23. Bochmann, M.; Wilson, L. M. J. Chem. Soc., Chem. Commun. 1986, 1610.
24. Jordan, R. F.; Bajgur, C. S.; Willet, R.; Scott, B. J. Am. Chem. Soc. 1986, 108, 7410.
25. Jordan, R. F.; LaPointe, R. E.; Bajgur, C. S.; Echols, S. F.; Willett, R. J. Am. Chem. Soc. 1987, 109, 4111.
26. Jordan, R. F. J. Chem. Ed. 1988, 65, 285.
27. Bochmann, M.; Wilson, L. M.; Hursthouse, M. B.; Motevalli, M. Organometallics 1988, 7, 1148.
28. Bochmann, M.; Jaggar, A. J.; Nichols, J. C. Angew. Chem., Int. Ed. Engl. 1990, 29, 780.
29. Turner, H. W.; Hlatky, G. G.; Eckman, R. R. (Exxon Chemical). US Patent 5384299, 1993.
30. Turner, H. W.; Hlatky, G. G.; Canich, J. A. M. (Exxon Chemical). World Pat. Appl. WO 93/19103, 1993.
31. Bochmann, M.; Lancaster, S. J. Angew. Chem., Int. Ed. Engl. 1994, 33, 1634.
32. Guo, Z.; Swenson, D.; Jordan, R. F. Organometallics 1994, 13, 1424.
33. Bochmann, M.; Lancaster, S. J. J. Organomet. Chem. 1995, 497, 55.
34. Crowther, D. J.; Baenziger, N. C.; Jordan, R. F. J. Am. Chem. Soc. 1991, 113, 1455.
35. Bochmann, M.; Lancaster, S. J.; Robinson, O. B. J. Chem. Soc., Chem. Commun. 1995, 2081.
36. Ruwwe, J.; Erker, G.; Fro hlich, R. Angew. Chem., Int. Ed. Engl. 1996, 35, 80.
37. Sun, Y.; Spence, R. E. v. H.; Piers, W. E.; Parvez, M.; Yap, G. P. A. J. Am. Chem. Soc. 1997, 119, 5132.
38. Song, X.; Bochmann, M. J. Organomet. Chem. 1997, 545546, 597.
39. Lancaster, S. J.; Thornton-Pett, M.; Dawson, D. M.; Bochmann, M. Organometallics 1998, 17, 3829.
40. Chen, E. Y.-X.; Marks, T. J. Chem. Rev. 2000, 100, 1391.
41. Bochmann, M. J. Organomet. Chem. 2004, 689, 3982.
42. Resconi, L.; Bossi, S.; Abis, L. Macromolecules 1990, 23, 4489.
43. Pasynkiewicz, S. Polyhedron 1990, 9, 429.
44. Giannetti, E.; Nicoletti, G. M.; Mazzocchi, R. J. Polym. Sci., Polym. Chem. Ed. 1985, 23, 2117.
45. Sinn, H., Kaminsky, W., Hoker, H., Eds. Alumoxanes, Huthig & Wepf: Heidelberg, 1995.
46. Srinivasa Reddy, S.; Sivaram, S. Prog. Polym. Sci. 1995, 20, 309.
47. Sinn, H. Macromol. Symp. 1995, 97, 27.
48. Mason, M. R.; Smith, J. M.; Bott, S. G.; Barron, A. R. J. Am. Chem. Soc. 1993, 115, 4971.
49. Imhoff, D. W.; Simeral, L. S.; Sangokoya, S. A.; Peel, J. H. Organometallics 1998, 17, 1941.
50. Sugano, T.; Matsubara, K.; Fujita, T.; Takahashi, T. J. Mol. Catal. 1993, 82, 93.
51. Siedle, A. R.; Lamanna, W. M.; Newmark, R. A.; Stevens, J.; Richardson, D. E.; Ryan, M. F. Makromol. Chem., Macromol. Symp. 1993, 66,
215.
52. Siedle, A. R.; Newmark, R. A. J. Organomet. Chem. 1995, 497, 119.
53. Siedle, A. R.; Lamanna, W. M.; Newmark, R. A.; Schroepfer, J. N. J. Mol. Catal. A: Chemical 1998, 128, 257.
54. Zurek, E.; Woo, T. K.; Firman, T. K.; Ziegler, T. Inorg. Chem. 2001, 40, 361.
55. Zurek, E.; Ziegler, T. Organometallics 2002, 21, 83.
56. Zurek, E.; Ziegler, T. Prog. Polym. Sci. 2004, 29, 107.
57. Mo hring, P. C.; Coville, N. J. J. Organomet. Chem. 1994, 479, 1.
58. Galimberti, M.; Destro, M.; Fusco, O.; Piemontesi, F.; Camurati, I. Macromolecules 1999, 32, 258.
59. Galimberti, M.; Piemontesi, F.; Fusco, O. Metallocenes as Catalysts for the Copolymerization of Ethene with Propene and Dienes. In
Metallocene-based Polyolefins: Preparation, Properties and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 1, p. 309.
60. Janiak, C.; Lange, K. C. H.; Scharmann, T. G. Appl. Organomet. Chem. 2000, 14, 316.
61. Resconi, L.; Giannini, U.; Dall9Occo, T. MAO-free Metallocene Catalysts for Ethylene (Co)Polymerization. In Metallocene-based Polyolefins:
Preparation, Properties and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 1, 69.
62. Kallio, K.; Kauhanen, J. Activation of Siloxy-substituted Compounds and Homopolymerization of Ethylene by Different Soluble
Alumoxane Cocatalysts. In Organometallic Catalysts and Olefin Polymerization: Catalysts for a New Millenium; Blom, R., Follestad, A.,
Rytter, E., Tilset, M., Ystenes, M., Eds.; Springer: Berlin, 2001; p 14.
63. van Baar, J. F.; Horton, A. D.; de Kloe, K. P.; Kragtwijk, E.; Mkoyan, S. G.; Nifant9ev, I. E.; Schut, P. A.; Taidakov, I. V. Organometallics
2003, 22, 2711.
64. Tsutsui, T.; Yoshitsugu, K.; Ueda, T. (Mitsui PC). Eur. Pat. Appl. EP 452920, 1991.
65. Resconi, L.; Galimberti, M.; Piemontesi, F.; Guglielmi, F.; Albizzati, E. (Montell Technology Company). U.S. Patent 5910464, 1999.
66. Dall9Occo, T.; Galimberti, M.; Resconi, L.; Albizzati, E.; Pennini, G. (Montell Technology Company). U.S. Patent 5849653, 1998.
67. Tritto, I.; Sacchi, M. C.; Locatelli, P.; Li, S. X. Macromol. Chem. Phys. 1996, 197, 1537.
68. Tritto, I.; Mealares, C.; Sacchi, M. C.; Locatelli, P. Macromol. Chem. Phys. 1997, 198, 3963.
69. Chien, J. C. W.; Wang, B. P. J. Polym. Sci., A: Polym. Chem. 1988, 26, 3089.
70. Srinivasa Reddy, S.; Shashidhar, G.; Sivaram, S. Macromolecules 1993, 26, 1180.
71. Tritto, I.; Donetti, R.; Sacchi, M. C.; Locatelli, P.; Zannoni, G. Macromolecules 1997, 30, 1247.
72. Tomotsu, N.; Ishihara, N.; Newman, T. H.; Malanga, M. T. J. Mol. Catal. 1998, 128, 167.
73. Smith, G. M.; Palmaka, S. W.; Roger, J. S.; Malpass, D. B.; Monfiston, D. J. (Akzo-Nobel). World Pat. Appl. WO 97/23288, 1997.
74. Fujita, M.; Seki, Y.; Miyatake, T. J. Polym. Sci. A: Polym. Chem. 2004, 42, 1107.
75. Hasan, T.; Ioku, A.; Nishii, K.; Shiono, T.; Ikeda, T. Macromolecules 2001, 34, 3142.
76. Obrey, S. J.; Bott, S. J.; Barron, A. R. Organometallics 2001, 20, 5162.
77. Busico, V.; Cipullo, R.; Cutillo, F.; Friederichs, N.; Ronca, S.; Wang, B. J. Am. Chem. Soc. 2003, 125, 12402.
78. Yang, X.; Stern, C.; Marks, T. J. J. Am. Chem. Soc. 1991, 113, 3623.
Olefin Polymerizations with Group IV Metal Catalysts 1147
79. Yang, X.; Stern, C.; Marks, T. J. J. Am. Chem. Soc. 1994, 116, 10015.
80. Ewen, J. A.; Elder, M. J. (Fina Technology). Eur. Pat. Appl. EP 427697, 1991.
81. Ewen, J. A.; Elder, M. J. (Fina Technology). U.S. Patent US 5561092, 1996.
82. Piers, W. E.; Chivers, T. Chem. Soc. Rev. 1997, 26, 345.
83. Li, L.; Stern, C. L.; Marks, T. J. Organometallics 2000, 19, 3332.
84. Li, L.; Marks, T. J. Organometallics 1998, 17, 3996.
85. Chen, E. Y.-X.; Stern, C. L.; Yang, S.; Marks, T. J. J. Am. Chem. Soc. 1996, 118, 12451.
86. Lancaster, S. J.; Walker, D. A.; Thornton-Pett, M.; Bochmann, M. Chem. Commun. 1999, 1533.
86a. Zhou, J.; Lancaster, S. J.; Walker, D. A.; Beck, S.; Thornton-Pett, M.; Bochmann, M. J. Am. Chem. Soc. 2001, 123, 223.
87. Metz, M. V.; Sun, Y.; Stern, C. L.; Marks, T. J. Organometallics 2002, 21, 3691.
88. Metz, M. V.; Schwartz, D. J.; Stern, C. L.; Nickias, P. N.; Marks, T. J. Angew. Chem., Int. Ed. 2000, 39, 1312.
89. Jia, L.; Yang, X.; Ishihara, A.; Marks, T. J. Organometallics 1995, 14, 3135.
90. Jia, L.; Yang, X.; Stern, C. L.; Marks, T. J. Organometallics 1997, 16, 842.
91. LaPointe, R. E.; Roof, G. R.; Abboud, K. A.; Klosin, J. J. Am. Chem. Soc. 2000, 122, 9560.
92. Focante, F.; Camurati, I.; Nanni, D.; Leardini, R.; Resconi, L. Organometallics 2004, 23, 5135.
93. Chien, J. C. W.; Tsai, W.-M.; Rausch, M. D. J. Am. Chem. Soc. 1991, 113, 8570.
94. Ewen, J. A.; Elder, M. J. (Fina Technology). Eur. Pat. Appl. EP 0426637, 1991.
95. Turner, H. W. (Exxon Chemical Co.). Eur. Pat. Appl. EP 0277004 A1, 1988.
96. Hlatky, G. G.; Upton, D. J.; Turner, H. W. (Exxon Chemical). World Pat. Appl. WO 91/09882, 1991.
97. Yang, X.; Stern, C. L.; Marks, T. J. Organometallics 1991, 10, 840.
98. Elder, M. J.; Ewen, J. A. (Fina Technology). Eur. Pat. Appl. EP 0573403, 1993.
99. Song, F.; Cannon, R. D.; Lancaster, S. J.; Bochmann, M. J. Mol. Catal. A: Chem 2004, 218, 21.
100. Al-Humydi, A.; Garrison, J. C.; Youngs, W. J.; Collins, S. Organometallics 2005, 24, 193.
101. Song, F.; Lancaster, S. J.; Cannon, R. D.; Schormann, M.; Humphrey, M. B.; Zuccaccia, C.; Macchioni, A.; Bochmann, M. Organometallics
2005, 24, 1315.
102. Rodriguez-Delgado, A.; Hannant, M. D.; Lancaster, S. J.; Bochmann, M. Macromol. Chem. Phys. 2004, 205, 334.
103. Li, H.; Li, L.; Marks, T. J. Angew. Chem., Int. Ed. 2004, 43, 4937.
104. Chen, M.-C.; Roberts, J. A. S.; Marks, T. J. J. Am. Chem. Soc. 2004, 126, 4605.
105. Li, H.; Li, L.; Marks, T. J.; Liable-Sands, L. M.; Rheingold, A. L. J. Am. Chem. Soc. 2003, 125, 10788.
106. Chen, M.-C.; Marks, T. J. J. Am. Chem. Soc. 2001, 123, 11803.
107. Chakraborty, D.; Chen, E. Y.-X. Organometallics 2002, 21, 1438.
108. Chen, E. Y.-X.; Kruper, W. J.; Roof, G.; Wilson, D. R. J. Am. Chem. Soc. 2001, 123, 745.
109. Chen, E. Y.-X.; Abboud, K. A. Organometallics 2000, 19, 5541.
110. Lancaster, S. J.; Bochmann, M. Organometallics 2001, 20, 2093.
111. Dahlmann, M.; Erker, G.; Fro hlich, R.; Meyer, O. Organometallics 2000, 19, 2956.
112. Erker, G. Acc. Chem. Res. 2001, 34, 309.
113. Karl, J.; Erker, G. J. Mol. Catal. A: Chem. 1998, 128, 85.
114. Tsai, W.-M.; Rausch, M. D.; Chien, J. C. W. Appl. Organomet. Chem. 1993, 7, 71.
115. Matsumoto, J.; Okamoto, T.; Watanabe, M.; Ishihara, N. (Idemistsu Kosan Co.). Eur. Pat. Appl. EP 513380, 1992.
116. Macchioni, A. Chem. Rev. 2005, 105, 2039.
117. Nifant9ev, I. E.; Ustynyuk, L. Y.; Laikov, D. N. Organometallics 2001, 20, 5375.
118. Vanka, K.; Chan, M. S. W.; Pye, C. C.; Ziegler, T. Organometallics 2000, 19, 1841.
119. Vanka, K.; Ziegler, T. Organometallics 2001, 20, 905.
120. Xu, Z.; Vanka, K.; Firman, T.; Michalak, A.; Zurek, E.; Zhu, C.; Ziegler, T. Organometallics 2002, 21, 2444.
121. Beck, S.; Prosenc, M. H.; Brintzinger, H.-H. J. Mol. Catal. A: Chem. 1998, 128, 41.
122. Beck, S.; Lieber, S.; Schaper, F.; Geyer, A.; Brintzinger, H.-H. J. Am. Chem. Soc. 2001, 123, 1483.
123. Zuccaccia, C.; Stahl, N. G.; Macchioni, A.; Chen, M.-C.; Roberts, J. A.; Marks, T. J. J. Am. Chem. Soc. 2004, 126, 1448.
124. Stahl, N. G.; Zuccaccia, C.; Jensen, T. R.; Marks, T. J. J. Am. Chem. Soc. 2003, 125, 5256.
125. Babushkin, D. E.; Brintzinger, H.-H. J. Am. Chem. Soc. 2002, 124, 12869.
126. Brintzinger, H.-H.; Fischer, D.; Mu lhaupt, R.; Rieger, B.; Waymouth, R. M. Angew. Chem., Int. Ed. 1995, 34, 1143.
127. Bochmann, M. J. Chem. Soc., Dalton Trans. 1996, 255.
128. Cossee, P. J. Catal. 1964, 3, 80.
129. Cossee, P. The mechanism of Ziegler-Natta Polymerization. II. Quantum-Chemical and Crystal-Chemical aspects. In The Stereochemistry of
Macromolecules; Ketley, A. D., Ed.; Dekker: New York, 1967; Vol. 1, pp p 145176.
130. Green, M. L. H. Pure Appl. Chem. 1972, 30, 373.
131. Dawoodi, Z.; Green, M. L. H.; Mtetwa, V. S. B.; Prout, K. J. Chem. Soc., Chem. Commun. 1982, 1410.
132. Brookhart, M.; Green, M. L. H. J. Organomet. Chem. 1983, 250, 395.
133. Laverty, D. T.; Rooney, J. J. J. Chem. Soc., Farady Trans. 1 1983, 79, 869.
134. Grubbs, R. H.; Coates, G. W. Acc. Chem. Res. 1996, 29, 85.
135. Chan, M. S. W.; Ziegler, T. Organometallics 2000, 19, 5182.
136. Xu, Z.; Vanka, K.; Ziegler, T. Organometallics 2004, 23, 104.
137. Zambelli, A.; Giongo, M. G.; Natta, G. Makromol. Chem. 1968, 112, 183.
138. Song, F.; Cannon, R. D.; Bochmann, M. Chem. Commun. 2004, 542.
139. Guerra, G.; Cavallo, L.; Moscardi, G.; Vacatello, M.; Corradini, P. Macromolecules 1996, 29, 4834.
140. Yoshida, T.; Koga, N.; Morokuma, K. Organometallics 1995, 14, 746.
141. Lohrenz, J. C. W.; Woo, T. K.; Ziegler, T. J. Am. Chem. Soc. 1995, 117, 12793.
142. Kawamura-Kuribayashi, H.; Koga, N.; Morokuma, K. J. Am. Chem. Soc. 1992, 114, 8687.
143. Kawamura-Kuribayashi, H.; Koga, N.; Morokuma, K. J. Am. Chem. Soc. 1992, 114, 2359.
144. Woo, T. K.; Fan, L.; Ziegler, T. A Combined Density Functional and Molecular Mechanics Study on Olefin Polymerization by
Metallocene Catalysts. In Ziegler Catalysts; Fink, G., Mu lhaupt, R., Brintzinger, H.-H., Eds.; Springer: Berlin, 1995, p 291.
1148 Olefin Polymerizations with Group IV Metal Catalysts
145. Woo, T. K.; Margl, P. M.; Blo chl, P. E.; Ziegler, T. J. Am. Chem. Soc. 1996, 118, 13021.
146. Woo, T. K.; Margl, P.; Ziegler, T.; Blo chl, P. E. Organometallics 1997, 16, 3454.
147. Lanza, G.; Fragala` , I. L.; Marks, T. J. J. Am. Chem. Soc. 1998, 120, 8257.
148. Lanza, G.; Fragala` , I. L.; Marks, T. J. Organometallics 2001, 20, 4006.
149. Lanza, G.; Fragala` , I. L.; Marks, T. J. Organometallics 2002, 21, 5594.
150. Lanza, G.; Fragala` , I. L.; Marks, T. J. J. Am. Chem. Soc. 2000, 122, 12764.
151. Vanka, K.; Xu, Z.; Ziegler, T. Organometallics 2004, 23, 2900.
152. Vanka, K.; Xu, Z.; Ziegler, T. Can. J. Chem. 2003, 81, 1413.
153. Sillars, D. R.; Landis, C. R. J. Am. Chem. Soc. 2003, 125, 9894.
154. Landis, C. R.; Rosaaen, K. A.; Sillars, D. R. J. Am. Chem. Soc. 2003, 125, 1710.
155. Landis, C. R.; Rosaaen, K. A.; Uddin, J. J. Am. Chem. Soc. 2002, 124, 12062.
156. Liu, Z.; Somsook, E.; White, C. B.; Rosaaen, K. A.; Landis, C. R. J. Am. Chem. Soc. 2001, 123, 11193.
157. Schaper, F.; Geyer, A.; Brintzinger, H.-H. Organometallics 2002, 21, 473.
158. Fink, G.; Fenzl, W.; Mynott, R. Z. Naturforsch. Teil B 1985, 40b, 158.
159. Miyatake, T.; Mizunuma, K.; Kakugo, M. Makromol. Chem., Macromol. Symp. 1993, 66, 203.
160. Hustad, P. D.; Tian, J.; Coates, G. W. J. Am. Chem. Soc. 2002, 124, 3614.
161. Lamberti, M.; Pappalardo, D.; Zambelli, A.; Pellecchia, C. Macromolecules 2002, 35, 658.
162. Resconi, L.; Cavallo, L.; Fait, A.; Piemontesi, F. Chem. Rev. 2000, 100, 1253.
163. Guerra, G.; Cavallo, L.; Moscardi, G.; Vacatello, M.; Corradini, P. J. Am. Chem. Soc. 1994, 116, 2988.
164. Guerra, G.; Longo, P.; Cavallo, L.; Corradini, P.; Resconi, L. J. Am. Chem. Soc. 1997, 119, 4394.
165. Talarico, G.; Busico, V.; Cavallo, L. J. Am. Chem. Soc. 2003, 125, 7172.
166. IUPAC Commission of Macromolecular Nomenclature Pure Appl. Chem. 1979, 51, 1101.
167. Corradini, P.; Paiaro, G.; Panunzi, A. J. Polym. Sci., Polym. Symp. 1967, 16, 2906.
168. Cahn, R. S.; Ingold, C.; Prelog, V. Angew. Chem., Int. Ed. Engl. 1966, 5, 385.
169. Schlo gl, K. Top. Stereochem. 1966, 1, 39.
170. IUPAC Nomenclature of Inorganic Chemistry Pure Appl. Chem. 1971, 18, 77.
171. Stanley, K.; Baird, M. C. J. Am. Chem. Soc. 1975, 97, 6598.
172. Corradini, P.; Guerra, G.; Cavallo, L. Acc. Chem. Res. 2004, 37, 231.
173. Farina, M. Top. Stereochem. 1987, 17, 1.
174. Corradini, P.; Guerra, G.; Cavallo, L. Top. Stereochem. 2003, 24, 1.
175. Corradini, P.; Cavallo, L.; Guerra, G. Molecular Modeling Studies on Stereospecificity and Regiospecificity of Propene Polymerization by
Metallocenes. In Metallocene-based Polyolefins: Preparation, Properties and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester,
2000; Vol. 2, p 3.
176. Rappe , A. K.; Skiff, W. M.; Casewit, C. J. Chem. Rev. 2000, 100, 1435.
177. Coates, G. W. Chem. Rev. 2000, 100, 1223.
178. Angermund, K.; Fink, G.; Jensen, V. R.; Kleinschmidt, R. Chem. Rev. 2000, 1000, 1457.
179. Busico, V.; Cipullo, R. Prog. Polym. Sci. 2001, 26, 443.
180. Resconi, L.; Abis, L.; Franciscono, G. Macromolecules 1992, 25, 6814.
181. Ewen, J. A.; Elder, M. J.; Jones, R. L.; Curtis, S.; Cheng, H. N. Syndiospecific Propylene Polymerizations with iPr[CpFlu]ZrCl
2
. In
Catalytic Olefin Polymerization, Studies in Surface Science and Catalysis; Keii, T., Soga, K., Eds.; Elsevier: New York, 1990; p 439.
182. Busico, V.; Cipullo, R.; Cutillo, F.; Vacatello, M.; Van Axel Castelli, V. Macromolecules 2003, 36, 4258.
183. Corradini, P.; Barone, V.; Fusco, R.; Guerra, G. Eur. Polym. J. 1979, 15, 1133.
184. Corradini, P.; Guerra, G.; Fusco, R.; Barone, V. Eur. Polym. J. 1980, 16, 835.
185. Corradini, P.; Barone, V.; Fusco, R.; Guerra, G. J. Catal. 1982, 77, 32.
186. Corradini, P.; Barone, V.; Guerra, G. Macromolecules 1982, 15, 1242.
187. Corradini, P.; Barone, V.; Fusco, R.; Guerra, G. Gazz. Chim. Ital. 1983, 113, 601.
188. Corradini, P.; Guerra, G.; Barone, V. Eur. Polym. J. 1984, 20, 1177.
189. Corradini, P.; Guerra, G.; Vacatello, M.; Villani, V. Gazz. Chim. Ital. 1988, 118, 173.
190. Cavallo, L.; Corradini, P.; Guerra, G.; Vacatello, M. Polymer 1991, 32, 1329.
191. Cavallo, L.; Guerra, G.; Vacatello, M.; Corradini, P. Macromolecules 1991, 24, 1784.
192. Cavallo, L.; Guerra, G.; Corradini, P. Gazz. Chim. Ital. 1996, 126, 463.
193. Milano, G.; Cavallo, L.; Guerra, G. J. Am. Chem. Soc. 2002, 124, 13368.
194. Corradini, P.; Guerra, G.; Pucciariello, R. Macromolecules 1985, 18, 2030.
195. Zambelli, A.; Sacchi, M. C.; Locatelli, P.; Zannoni, G. Macromolecules 1982, 15, 211.
196. Longo, P.; Grassi, A.; Pellecchia, C.; Zambelli, A. Macromolecules 1987, 20, 1015.
197. Pino, P.; Galimberti, M.; Prada, P.; Consiglio, G. Makromol. Chem. 1990, 191, 1677.
198. Pino, P.; Cioni, P.; Wei, J. J. Am. Chem. Soc. 1987, 109, 6189.
199. Cavallo, L.; Guerra, G.; Corradini, P.; Vacatello, M. Chirality 1991, 3, 299.
200. Gilchrist, J. H.; Bercaw, J. E. J. Am. Chem. Soc. 1996, 118, 12021.
201. Dahlmann, M.; Erker, G.; Nissinen, M.; Fro hlich, R. J. Am. Chem. Soc. 1999, 121, 2820.
202. Ewen, J. A. J. Am. Chem. Soc. 1984, 106, 6355.
203. Kaminsky, W.; Ku lper, K.; Brintzinger, H. H.; Wild, F. R. W. P. Angew. Chem., Int. Ed. Engl. 1985, 24, 507.
204. Resconi, L.; Balboni, D.; Baruzzi, G.; Fiori, C.; Guidotti, S.; Mercandelli, P.; Sironi, A. Organometallics 2000, 19, 420.
205. Ewen, J. A.; Haspeslagh, L.; Elder, M. J.; Atwood, J. L.; Zhang, H.; Cheng, H. N. Propylene Polymerizations with Group 4 Metallocene/
alumoxane Systems. In Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.; Springer:
Berlin, 1988; p 281.
206. Toto, M.; Cavallo, L.; Corradini, P.; Moscardi, G.; Resconi, L.; Guerra, G. Macromolecules 1998, 31, 3431.
207. Moscardi, G.; Resconi, L.; Cavallo, L. Organometallics 2001, 20, 1918.
208. Ewen, J. A.; Jones, R. L.; Razavi, A.; Ferrara, J. D. J. Am. Chem. Soc. 1988, 110, 6255.
Olefin Polymerizations with Group IV Metal Catalysts 1149
209. Ewen, J. A.; Elder, M. J.; Jones, R. L.; Haspeslagh, L.; Atwood, J. L.; Bott, S. G.; Robinson, K. Makromol. Chem., Macromol. Symp. 1991, 48/
49, 253.
210. Antberg, M.; Dolle, V.; Klein, R.; Rohrmann, J.; Spaleck, W.; Winter, A. Propylene Polymerization by Stereorigid Metallocene Catalysts:
Some New Aspects of the Metallocene Structure/polypropylene Microstructure Correlation. In Catalytic Olefin Polymerization, Studies in
Surface Science and Catalysis; Keii, T., Soga, K., Eds.; Kodansha-Elsevier: Tokyo, 1990; p 501.
211. Dolle, V.; Rohrmann, J.; Winter, A.; Antberg, M.; Klein, R. (Hoechst). Eur. Pat. Appl. EP 399347, 1990.
212. Kleinschmidt, R.; Reffke, M.; Fink, G. Macromol. Rapid Commun. 1999, 20, 284.
213. Yoshida, T.; Koga, N.; Morokuma, K. Organometallics 1996, 15, 766.
214. van der Leek, Y.; Angermund, K.; Reffke, M.; Kleinschmidt, R.; Goretzki, R.; Fink, G. Chem. Eur. J. 1997, 3, 585.
215. Ewen, J. A.; Elder, M. J. Makromol. Chem., Macromol. Symp. 1993, 66, 179.
216. Cheng, H. N.; Ewen, J. A. Makromol. Chem. 1989, 190, 1931.
217. Resconi, L.; Camurati, I.; Sudmeijer, O. Top. Catal. 1999, 7, 145.
218. Blom, R.; Dahl, I. M. Macromol. Chem. Phys. 1999, 200, 442.
219. Moscardi, G.; Piemontesi, F.; Resconi, L. Organometallics 1999, 18, 5264.
220. Blom, R.; Dahl, I. M. Macromol. Chem. Phys. 2001, 202, 719.
221. Andersen, A.; Blom, R.; Dahl, I. M. Macromol. Chem. Phys. 2001, 202, 726.
222. Balboni, D.; Moscardi, G.; Baruzzi, G.; Braga, V.; Camurati, I.; Piemontesi, F.; Resconi, L.; Nifantev, I. E.; Venditto, V.; Antinucci, S.
Macromol. Chem. Phys. 2001, 202, 2010.
223. Fan, W.; Waymouth, R. M. Macromolecules 2001, 34, 8619.
224. Han-Adebekun, G. C.; Hamba, M.; Ray, W. H. J. Polym. Sci., A 1997, 35, 2063.
225. Thorshaug, K.; Rytter, E.; Ystenes, M. Macromol. Rapid Commun. 1997, 18, 715.
226. Thorshaug, K.; Stvneng, J. A.; Rytter, E.; Ystenes, M. Macromolecules 1998, 31, 7149.
227. Dang, V. A.; Yu, L.-C.; Balboni, D.; DallOcco, T.; Resconi, L.; Mercandelli, P.; Moret, M.; Sironi, A. Organometallics 1999, 18, 3781.
228. Naga, N.; Mizunuma, K. Polymer 1998, 39, 5059.
229. Karol, F. J.; Kao, S. C.; Wasserman, E. P.; Brady, R. C. New J. Chem. 1997, 21, 797.
230. Hasegawa, S.; Sone, M.; Tanabiki, M.; Sato, M.; Yano, A. J. Polym. Sci. A: Polym. Chem. 2000, 38, 4641.
231. Janiak, C.; Lange, K. C. H.; Marquardt, P. Macromol. Rapid Commun. 1995, 16, 643.
232. Rossi, A.; Odian, G.; Zhang, J. Macromolecules 1995, 28, 1739.
233. Rossi, A.; Zhang, J.; Odian, G. Macromolecules 1996, 29, 2331.
234. Janiak, C.; Lange, K. C. H.; Marquardt, P.; Kruger, R.-P.; Hanselmann, R. Macromol. Chem. Phys. 2002, 203, 129.
235. Harney, M. B.; Keaton, R. J.; Sita, L. R. J. Am. Chem. Soc. 2004, 126, 4536.
236. Busico, V.; Cipullo, R.; Ronca, S.; Budzelaar, P. H. M. Macromol. Rapid Commun. 2001, 22, 1405.
237. Yoder, J. C.; Bercaw, J. E. J. Am. Chem. Soc. 2002, 124, 2548.
238. Busico, V.; Cipullo, R. J. Am. Chem. Soc. 1994, 116, 9329.
239. Leclerc, M. K.; Brintzinger, H.-H. J. Am. Chem. Soc. 1995, 117, 1651.
240. Busico, V.; Cipullo, R. J. Organomet. Chem. 1995, 497, 113.
241. Resconi, L.; Fait, A.; Piemontesi, F.; Colonnesi, M.; Rychlicki, H.; Zeigler, R. Macromolecules 1995, 28, 6667.
242. Leclerc, M. K.; Brintzinger, H.-H. J. Am. Chem. Soc. 1996, 118, 9024.
243. Busico, V.; Caporaso, L.; Cipullo, R.; Landriani, L.; Angelini, G.; Margonelli, A.; Segre, A. L. J. Am. Chem. Soc. 1996, 118, 2105.
244. Busico, V.; Brita, D.; Caporaso, L.; Cipullo, R.; Vacatello, M. Macromolecules 1997, 30, 3971.
245. Busico, V.; Cipullo, R.; Caporaso, L.; Angelini, G.; Segre, A. L. J. Mol. Catal. A: Chem. 1998, 128, 53.
246. Camurati, I.; Fait, A.; Piemontesi, F.; Resconi, L.; Tartarini, S. Transfer and Isomerization Reactions in Propylene Polymerization with the
Isospecific, Highly Regiospecific rac-Me2C(3-t-Bu-1-Ind)2ZrCl2/MAO Catalyst. In Transition Metal Catalysis in Macromolecular Design ACS
Symposium Series; Boffa, L. S., Novak, B. M., Eds.; American Chemical Society: Washington, DC, 2000; Vol. 760, 174.
247. Volkis, V.; Shmulinson, M.; Averbuj, C.; Lisovskii, A.; Edelmann, F. T.; Eisen, M. S. Organometallics 1998, 17, 3155.
248. Volkis, V.; Nelkenbaum, E.; Lisovskii, A.; Hasson, G.; Semiat, R.; Kapon, M.; Botoshansky, M.; Eishen, Y.; Eisen, M. S. J. Am. Chem. Soc.
2003, 125, 2179.
249. Resconi, L. J. Mol. Catal. A: Chem. 1999, 146, 167.
250. Resconi, L.; Piemontesi, F.; Jones, R. L. High-molecular-weight Atactic Polypropylene from Metallocene Catalysts. Influence of Ligand
Structure and Polymerization Conditions on Molecular Weight. In Metallocene-catalyzed Polymers - Materials, Properties, Processing & Markets;
Benedikt, G. M., Goodall, B. L., Eds.; Plastics Design Library: New York, 1998.
251. Sugano, T.; Tayano, T.; Uchino, H.; Ioku, A.; Iwama, N.; Endo, J.; Osano, Y. In Proceedings of the SPO 99; Schotland Business Res. Ed.:
Houston, 1999; p 31.
252. Voegele, J.; Troll, C.; Rieger, B. Macromol. Chem. Phys. 2002, 203, 1918.
253. Tynys, A.; Saarinen, T.; Hakala, K.; Helaja, T.; Vanne, T.; Lehmus, P.; Lo fgren, B. Macromol. Chem. Phys. 2005, 206, 1043.
254. Camurati, I.; Cavicchi, B.; Dall9Occo, T.; Piemontesi, F. Macromol. Chem. Phys. 2001, 202, 701.
255. Ju ngling, S.; Mu lhaupt, R.; Stehling, U.; Brintzinger, H.-H.; Fischer, D.; Langhauser, F. J. Polym. Sci. Polym. Chem. 1995, 33, 1305.
256. van der Linden, A.; Schaverien, C. J.; Meijboom, N.; Ganter, C.; Guy Orpen, A. J. Am. Chem. Soc. 1995, 117, 3008.
257. Resconi, L.; Piemontesi, F.; Camurati, I.; Balboni, D.; Sironi, A.; Moret, M.; Rychlicki, H.; Zeigler, R. Organometallics 1996, 15, 5046.
258. Cherian, A. E.; Lobkovsky, E. B.; Coates, G. W. Macromolecules 2005, 38, 6259.
259. Eshuis, J. J. W.; Tan, Y. Y.; Meetsma, A.; Teuben, J. H.; Renkema, J.; Evens, G. G. Organometallics 1992, 11, 362.
260. Shaffer, T. D.; Canich, J. A. M.; Squire, K. R. Macromolecules 1998, 31, 5145.
261. Kesti, M.; Waymouth, R. M. J. Am. Chem. Soc. 1992, 114, 3565.
262. Resconi, L.; Piemontesi, F.; Franciscono, G.; Abis, L.; Fiorani, T. J. Am. Chem. Soc. 1992, 114, 1025.
263. Bunel, E.; Burger, B. J.; Bercaw, J. E. J. Am. Chem. Soc. 1988, 110, 976.
264. Yang, X.; Jia, L.; Marks, T. J. J. Am. Chem. Soc. 1993, 115, 3392.
265. Yang, X.; Seyam, A. M.; Fu, P.-F.; Marks, T. J. Macromolecules 1994, 27, 4625.
266. Resconi, L.; Jones, R. L.; Rheingold, A. L.; Yap, G. P. A. Organometallics 1996, 15, 998.
267. Resconi, L.; Piemontesi, F.; Camurati, I.; Sudmeijer, O.; Nifant9ev, I. E.; Ivchenko, P. V.; Kuzmina, L. G. J. Am. Chem. Soc. 1998, 120,
2308.
1150 Olefin Polymerizations with Group IV Metal Catalysts
268. Weng, W.; Markel, E. J.; Dekmezian, A. H. Macromol. Rapid Commun. 2000, 21, 103.
269. Weng, W.; Markel, E. J.; Peacock, A. J.; Dekmezian, A. H. Macromolecules 2000, 33, 8541.
270. Weng, W.; Markel, E. J.; Dekmezian, A. H. Macromol. Rapid Commun. 2001, 22, 1488.
271. Shiono, T.; Kurosawa, H.; Soga, K. Macromolecules 1994, 27, 2635.
272. Bhriain, N. N.; Brintzinger, H.-H.; Ruchatz, D.; Fink, G. Macromolecules 2005, 38, 2056.
273. Fan, G.; Dong, J.-Y. J. Mol. Catal. A: Chem. 2005, 236, 246.
274. Mogstad, A.-L.; Waymouth, R. M. Macromolecules 1992, 25, 2282.
275. Scollard, J. D.; McConville, D. H. J. Am. Chem. Soc. 1996, 118, 10008.
276. Averbuj, C.; Tish, E.; Eisen, M. S. J. Am. Chem. Soc. 1998, 120, 8640.
277. Lamberti, M.; Gliubizzi, R.; Mazzeo, M.; Tedesco, C.; Pellecchia, C. Macromolecules 2004, 37, 276.
278. Michiue, K.; Jordan, R. F. Organometallics 2004, 23, 460.
279. Michiue, K.; Jordan, R. F. Macromolecules 2003, 36, 9707.
280. Hindryckx, F.; Dubois, P.; Jerome, R.; Marti, M. G. Polymer 1998, 39, 621.
281. Kissin, Y. V.; Rishina, L. A.; Vizen, E. I. J. Polym. Sci. A: Polym. Chem. 2002, 40, 18991911.
282. Carvill, A.; Tritto, I.; Locatelli, P.; Sacchi, M. C. Macromolecules 1997, 30, 7056.
283. Fu, P.-F.; Marks, T. J. J. Am. Chem. Soc. 1995, 117, 10747.
284. Koo, K.; Marks, T. J. J. Am. Chem. Soc. 1998, 120, 4019.
285. Wester, T. S.; Johnsen, H.; Kittilsen, P.; Rytter, E. Macromol. Chem. Phys. 1998, 199, 1989.
286. Chien, J. C. W.; Yu, Z.; Marques, M. M.; Flores, J. C.; Rausch, M. D. J. Polym. Sci. A: Polym. Chem. 1998, 36, 319.
287. Chien, J. C. W.; Wang, B. P. J. Polym. Sci., A: Polym. Chem. 1990, 28, 15.
288. Fink, G.; Herfert, N.; Montag, P. The Relationship Between Kinetics and Mechanisms. In Ziegler Catalysts; Fink, G., Mu lhaupt, R.,
Brintzinger, H.-H., Eds.; Springer: Berlin, 1995; p 159.
289. Fait, A.; Resconi, L.; Guerra, G.; Corradini, P. Macromolecules 1999, 32, 2104.
290. Song, F.; Cannon, R. D.; Bochmann, M. J. Am. Chem. Soc. 2003, 125, 7641.
291. Nele, M.; Mohammed, M.; Xin, S.; Collins, S.; Dias, M. L.; Pinto, J. C. Macromolecules 2001, 34, 3830.
292. Mohammed, M.; Nele, M.; Al-Humydi, A.; Xin, S.; Stapleton, R. A.; Collins, S. J. Am. Chem. Soc. 2003, 125, 7930.
293. Busico, V.; Cipullo, R.; Cutillo, F.; Vacatello, M. Macromolecules 2002, 35, 349.
294. Busico, V.; Cipullo, R.; Corradini, P. Makromol. Chem., Rapid Commun. 1993, 14, 97.
295. Busico, V.; Cipullo, R.; Chadwick, J. C.; Modder, J. F.; Sudmeijer, O. Macromolecules 1994, 27, 7538.
296. Busico, V.; Cipullo, R.; Romanelli, V.; Ronca, S.; Togrou, M. J. Am. Chem. Soc. 2005, 127, 1608.
297. Busico, V.; Cipullo, R.; Esposito, V. Macromol. Rapid Commun. 1999, 20, 116.
298. Monaco, G. Macromolecules 2001, 34, 4406.
299. Tsutsui, T.; Kashiwa, N.; Mizuno, A. Makromol. Chem., Rapid Commun. 1990, 11, 565.
300. Kioka, M.; Mizuno, A.; Tsutsui, T.; Kashiwa, N. 1-Butene Polymerization with Ethylenebis-(1-Indenyl)Zirconium Dichloride and
Methylaluminoxane Catalyst System. In Catalysis in Polymer Synthesis, ACS Symposium Series; Vandenberg, E. J., Salamone, J. C., Eds.;
American Chemical Society: Washington, DC, 1992; Vol. 496, p 72.
301. Borriello, A.; Busico, V.; Cipullo, R.; Fusco, O.; Chadwick, J. C. Macromol. Chem. Phys. 1997, 198, 1257.
302. Landis, C. R.; Sillars, D. R.; Batterton, J. M. J. Am. Chem. Soc. 2004, 126, 8890.
303. Mynott, R.; Fink, G.; Fenzl, W. Angew. Makromol. Chem. 1987, 154, 1.
304. Mori, H.; Terano, M. Trends Polym. Sci. 1997, 5, 314.
305. Lahelin, M.; Kokko, E.; Lehmus, P.; Pitka nen, P.; Lo fgren, B.; Seppa la , J. Macromol. Chem. Phys. 2003, 204, 1323.
306. Belov, G. P.; Gyulumyan, H. R.; Khrapova, I. M.; Maryin, V. P.; Korneev, N. N. J. Mol. Catal. A: Chem 1997, 115, 155.
307. Lin, M.; Spivak, G. J.; Baird, M. C. Organometallics 2002, 21, 2350.
308. Marques, M. M.; Costa, C.; Lemos, F.; Ribeiro, F. R.; Dias, A. R. React. Kinet. Catal. Lett. 1997, 62, 9.
309. Marques, M. M.; Tait, P. J. T.; Mejlik, J.; Dias, A. R. J. Pol. Sci., Part A: Pol. Chem. 1998, 36, 573.
310. Blom, R.; Swang, O.; Heyn, R. H. Macromol. Chem. Phys. 2002, 203, 381.
311. Goodman, J. T.; Schrock, R. R. Organometallics 2001, 20, 5205.
312. Schrodi, Y.; Schrock, R. R.; Bonitatebus, P. J., Jr., Organometallics 2001, 20, 3560.
313. Boor, J., Jr. Ziegler-Natta Catalysts and Polymerizations; Academic Press: New York, 1979.
314. Kissin, Y. V. Isospecific Polymerization of Olefins; Springer: New York, 1985.
315. van der Ven, S. Polypropylene and other Polyolefins. Polymerization and Characterization; Elsevier: Amsterdam, 1990.
316. Gavens, P. D.; Bottrill, M.; Kelland, J. W.; McMeeking, J. Ziegler-Natta Catalysis. In Comprehensive Organometallic Chemistry I; Wilkinson,
G., Stone, F. G. A., Abel, E. W., Eds.; Pergamon: Oxford, 1982; Vol. 3, p 475.
317. Moore, E. P. Jr. Polypropylene Handbook: Polymerization, Characterization, Properties, Applications; Hanser: Munich, 1996.
318. Kashiwa, N. J. Polym. Sci. A: Polym. Chem. 2004, 42, 1.
319. Giannini, U. Makromol. Chem. Suppl. 1981, 5, 216.
320. Galli, P.; Luciani, L.; Cecchin, G. Angew. Makromol. Chem. 1981, 94, 63.
321. Barbe` , C.; Cecchin, G.; Noristi, L. Adv. Polym. Sci. 1987, 81, 1.
322. Seppa la , J. V.; Ha rko nen, M.; Luciani, L. Makromol. Chem. 1989, 190, 2535.
323. Ha rko nen, M.; Seppa la , J. V.; Va a na nen, T. In Catalytic Olefin Polymerization; Keii, T., Soga, K., Eds.; Elsevier: Amsterdam, 1990, p 87.
324. Proto, A.; Oliva, L.; Pellecchia, C.; Sivak, A. J.; Cullo, L. A. Macromolecules 1990, 23, 2904.
325. Okano, T.; Chida, K.; Furuhashi, H.; Nakano, A.; Ukei, S. In Catalytic Olefin Polymerization; Keii, T., Soga, K., Eds.; Elsevier: Amsterdam,
1990; p 177.
326. Mori, H.; Sawada, M.; Higuchi, T.; Hasebe, K.; Otsuka, N.; Terano, M. Macromol. Rapid Commun. 1999, 20, 245.
327. Mori, H.; Hasebe, K.; Terano, M. Macromol. Chem. Phys. 1998, 199, 2709.
328. Bart, J. C. J.; Roovers, W. J. Mat. Sci. 1995, 30, 2809.
329. Sozzani, P.; Bracco, S.; Comotti, A.; Simonutti, R.; Camurati, I. J. Am. Chem. Soc. 2003, 125, 12881.
330. Galli, P.; Barbe` , P. C.; Guidetti, G. P.; Zannetti, R.; Martorana, A.; Marigo, A.; Bergozza, M.; Fichera, A. Eur. Polym. J. 1983, 19, 19.
331. Gerbasi, R.; Marigo, A.; Martorana, A.; Zannetti, R.; Guidetti, G. P.; Baruzzi, G. Eur. Polym. J. 1984, 20, 967.
332. Marigo, A.; Marega, C.; Zannetti, R.; Morini, G.; Ferrara, G. Eur. Polym. J. 2000, 36, 1921.
Olefin Polymerizations with Group IV Metal Catalysts 1151
333. Corradini, P.; Busico, V.; Guerra, G. Possible Models for the Steric Control in the Heterogeneous High-yield and Homogeneous Ziegler
Natta Polymerizations of 1-Alkenes. In Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W., Sinn, H., Eds.;
Springer: Berlin, 1988; p 337.
334. Busico, V.; Corradini, P.; De Martino, L.; Proto, A.; Savino, V.; Albizzati, E. Makromol. Chem. 1985, 186, 1279.
335. Potapov, A. G.; Kriventsov, V. V.; Kochubey, D. I.; Bukatov, G. D.; Zakharov, V. A. Macromol. Chem. Phys. 1997, 198, 3477.
336. Terano, M.; Kataoka, T.; Keii, T. Makromol. Chem. 1987, 188, 1477.
337. Albizzati, E.; Giannini, U.; Morini, G.; Smith, C. A.; Zeigler, R. Advances in Propylene Polymerization with MgCl
2
Supported Catalysts. In
Ziegler Catalysts; Fink, G., Mu lhaupt, R., Brintzinger, H.-H., Eds.; Springer: Berlin, 1995; p 413.
338. Sacchi, M. C.; Tritto, I.; Shan, C.; Mendichi, R.; Noristi, L. Macromolecules 1991, 24, 6823.
339. Albizzati, E.; Barbe` , P. C.; Noristi, L.; Scordamaglia, R.; Barino, L.; Giannini, U.; Morini, G. (Himont). Eur. Pat. Appl. EP 361494, 1989.
340. Morini, G.; Cristofori, A. (Montell). Eur. Pat. EP 361494, 1990.
341. Albizzati, E.; Giannini, U.; Morini, G.; Galimberti, M.; Barino, L.; Scordamaglia, R. Macromol. Symp. 1995, 89, 73.
342. Barino, L.; Scordamaglia, R. Macromol. Symp. 1995, 89, 101.
343. Albizzati, E.; Galimberti, M.; Giannini, U.; Morini, G. Makromol. Chem., Macromol. Symp. 1991, 48/49, 223.
344. Morini, G.; Balbontin, G.; Gulevich, Y.; Duijghuisen, H.; Kelder, R.; Klusener, P. A.; Korndorffer, F. (Basell). World Pat. Appl. WO 00/
63261, 2000.
345. Cecchin, G.; Morini, G.; Pelliconi, A. Macromol. Symp. 2001, 173, 195.
346. Pater, J. T. M.; Weickert, G.; van Swaaij, W. P. M. AIChE J. 2003, 49, 450.
347. Simonazzi, T.; Giannini, U. Gazz. Chim. Ital. 1994, 124, 533.
348. Galli, P. Prog. Polym. Sci. 1994, 19, 959.
349. McKenna, T. F.; Soares, J. B. P. Chem. Eng. Sci. 2001, 56, 3931.
350. Hutchinson, R. A.; Chen, C. M.; Ray, W. H. J. Appl. Polym. Sci. 1992, 44, 1389.
351. Kakugo, M.; Sadatoshi, H.; Kojima, K.; Yokoyama, M. Macromolecules 1989, 22, 547.
352. Cecchin, G.; Marchetti, E.; Baruzzi, G. Macromol. Chem. Phys. 2001, 202, 1987.
353. Pater, J. T. M.; Weickert, G.; Loos, J.; van Swaaij, W. P. M. Chem. Eng. Sci. 2001, 56, 4107.
354. Galli, P.; Haylock, J. C. Prog. Polym. Sci. 1991, 16, 443.
355. Bo hm, L. Angew. Chem., Int. Ed. 2003, 42, 5010.
356. Zakharov, V. A.; Bukatov, G. D.; Barababov, A. A. Macromol. Symp. 2004, 213, 19.
357. Parasu Veera, U.; Weickert, G.; Agarwal, U. S. AIChE J. 2002, 48, 1062.
358. Noristi, L.; Marchetti, E.; Baruzzi, G. J. Polym. Sci. A: Polym. Chem. 1994, 32, 3047.
359. Fragonese, D.; Mortara, S.; Bresadola, S. J. Mol. Catal. A: Chem. 2001, 172, 89.
360. Baulin, A. A.; Novikova, Y. I.; Malkova, G. Y.; Maksimov, V. L.; Vyshinskaya, L. I.; Ivanchev, S. S. Polym. Sci. USSR 1980, 22, 205.
361. Kashiwa, N.; Yoshitake, J. Makromol. Chem. 1984, 185, 1133.
362. Chien, J. C. W.; Wu, J. C. J. Polym. Sci. A: Polym. Chem. 1982, 20, 2461.
363. Weber, S.; Chien, J. C. W.; Hu, Y. J. Polym. Sci. A: Polym. Chem. 1989, 27, 1499.
364. Sergeev, S. A.; Polubayarov, V. A.; Zakharov, V. A.; Anufrienko, U. F.; Bukatov, G. D. Makromol. Chem. 1986, 187, 243.
365. Chien, J. C. W.; Hu, Y. J. Pol. Sci., Part A: Polym. Chem. 1989, 27, 897.
366. Mori, H.; Hasebe, K.; Terano, M. J. Mol. Catal. A: Chem. 1999, 140, 165.
367. Mejzlik, J.; Lesna, M.; Kratochvila, J. Adv. Polym. Sci. 1986, 81, 83.
368. Liu, B.; Matsuoka, H.; Terano, M. Macromol. Rapid Commun. 2001, 22, 1.
369. Bukatov, G. D.; Zakharov, V. A. Macromol. Chem. Phys. 2001, 202, 2003.
370. Tait, P. J. T.; Zohuri, G. H.; Kells, A. M.; McKenzie, I. D. In Ziegler Catalysts. Recent Scientific Innovations and Technological Improvements;
Fink, G., Mu lhaupt, R., Brintzinger, H. H., Eds.; Springer: Berlin, 1995; p 343.
371. Bukatov, G. D.; Goncharov, V. S.; Zakharov, V. A. Macromol. Chem. Phys. 1995, 196, 1751.
372. Kashiwa, N.; Yoshitake, J.; Toyota, A. Polym. Bull. 1988, 19, 333.
373. Chadwick, J. C. Effects of Electron Donors in Super High Activity Catalysts for Polypropylene. In Ziegler Catalysts; Fink, G., Mu lhaupt, R.,
Brintzinger, H.-H., Eds.; Springer: Berlin, 1995; p 427.
374. Kakugo, M.; Miyatake, T.; Naito, Y.; Mizunuma, K. Macromolecules 1988, 21, 314.
375. Soga, K.; Park, J. R.; Shiono, T.; Kashiwa, N. Makromol. Chem., Rapid Commun. 1990, 11, 117.
376. Soga, K.; Park, J. R.; Uchino, H.; Uozumi, T.; Shiono, T. Macromolecules 1989, 22, 3824.
377. Soga, K. Makromol. Chem., Macromol. Symp. 1993, 66, 43.
378. Sacchi, M. C.; Shan, C.; Locatelli, P.; Tritto, I. Macromolecules 1990, 23, 383.
379. Sacchi, M. C.; Forlini, F.; Tritto, I.; Mendichi, R.; Zannoni, G.; Noristi, L. Macromolecules 1992, 25, 5914.
380. Sacchi, M. C.; Forlini, F.; Tritto, I.; Locatelli, P.; Morini, G.; Baruzzi, G.; Albizzati, E. Macromol. Symp. 1995, 89, 91.
381. Morini, G.; Albizzati, E.; Balbontin, G.; Mingozzi, I.; Sacchi, M. C.; Forlini, F.; Tritto, I. Macromolecules 1996, 29, 5770.
382. Busico, V.; Cipullo, R.; Monaco, G.; Talarico, G.; Vacatello, M.; Chadwick, J. C.; Segre, A. L.; Sudmeijer, O. Macromolecules 1999, 32, 4173.
383. Busico, V.; Cipullo, R.; Talarico, G.; Segre, A. L.; Chadwick, J. C. Macromolecules 1997, 30, 4786.
384. Chadwick, J. C.; Morini, G.; Balbontin, G.; Camurati, I.; Heere, J. J. R.; Mingozzi, I.; Testoni, F. Macromol. Chem. Phys. 2001, 202, 1995.
385. Ha rko nen, M.; Seppa la , J.; Ch uj o, R.; Kogure, Y. Polymer 1995, 36, 1499.
386. Barino, L.; Scordamaglia, R. Macromol. Theory Simul. 1998, 7, 407.
387. Scordamaglia, R.; Barino, L. Macromol. Theory Simul. 1998, 7, 399.
388. Toto, M.; Morini, G.; Guerra, G.; Corradini, P.; Cavallo, L. Macromolecules 2000, 33, 1134.
389. Sacchi, M. C.; Forlini, F.; Tritto, I.; Locatelli, P.; Morini, G.; Noristi, L.; Albizzati, E. Macromolecules 1996, 29, 3341.
390. Cecchin, G.; Morini, G.; Piemontesi, F.; Ferraro, A.; News, J.; Cavallo, L. In EUPOC 2003. European Polymer Conference on Stereospecific
Polymerization and Stereoregular Polymers: Milan, Italy, 2003.
391. Boero, M.; Parrinello, M.; Hu ffer, S.; Weiss, H. J. Am. Chem. Soc. 2000, 122, 501.
392. Liu, B.; Nitta, T.; Nakatani, H.; Terano, M. Macromol. Chem. Phys. 2002, 203, 2412.
393. Liu, B.; Nitta, T.; Nakatani, H.; Terano, M. Macromol. Chem. Phys. 2003, 204, 395.
394. Potapov, A. G.; Terskikh, V. V.; Zakharov, V. A.; Bukatov, G. D. J. Mol. Catal. A: Chem. 1999, 145, 147.
395. Mori, H.; Tashino, K.; Terano, M. Macromol. Rapid Commun. 1995, 16, 651.
1152 Olefin Polymerizations with Group IV Metal Catalysts
396. Mori, H.; Tashino, K.; Terano, M. Macromol. Chem. Phys. 1996, 197, 895.
397. Mori, H.; Iizuka, T.; Tashino, K.; Terano, M. Macromol. Chem. Phys. 1997, 198, 2499.
398. Chadwick, J. C.; Morini, G.; Albizzati, E.; Balbontin, G.; Mingozzi, I.; Cristofori, A.; Sudmeijer, O.; van Kessel, G. M. M. Macromol. Chem.
Phys. 1996, 197, 2501.
399. Guastalla, G.; Giannini, U. Makromol. Chem., Rapid Commun. 1983, 4, 519.
400. Randall, J. C.; Ruff, C. J.; Vizzini, J. C.; Speca, A. N.; Burkhardt, T. J. Initial Insertion in Metallocene Polymerizations of Polypropylene. In
Metalorganic Catalysts for Synthesis and Polymerisation; Kaminsky, W., Ed. Springer: Berlin, 1999; p 601.
401. Chadwick, J. C.; Heere, J. J. R.; Sudmeijer, O. Macromol. Chem. Phys. 2000, 201, 1846.
402. Kissin, Y. V.; Rishina, L. A. J. Polym. Sci. A: Polym. Chem. 2002, 40, 1353.
403. Kissin, Y. V.; Mink, R. I.; Nowlin, T. E.; Brandolini, A. J. Top. Catal. 1999, 7, 69.
404. Kissin, Y. V.; Mink, R. I.; Nowlin, T. E. J. Polym. Sci. A: Polym. Chem. 1999, 37, 4255.
405. Kissin, Y. V.; Brandolini, A. J. J. Polym. Sci. A: Polym. Chem. 1999, 37, 4273.
406. Kissin, Y. V.; Mink, R. I.; Nowlin, T. E.; Brandolini, A. J. J. Polym. Sci. A: Polym. Chem. 1999, 37, 4281.
407. Kissin, Y. V. Macromol. Theory Simul. 2002, 11, 67.
408. Garoff, T.; Johansson, S.; Pesonen, K.; Waldvogel, P.; Lindgren, D. Eur. Polym. J. 2002, 38, 121.
409. Mikenas, T. B.; Zakharov, V. A.; Echevskaya, L. G.; Matsko, M. A. Macromol. Chem. Phys. 2001, 202, 475.
410. Matsko, M. A.; Bukatov, G. D.; Mikenas, T. B.; Zakharov, V. A. Macromol. Chem. Phys. 2001, 202, 1435.
411. Czaja, K.; Bialek, M. J. App. Polym. Sci. 2001, 79, 356.
412. Chadwick, J. C.; van Kessel, G. M. M.; Sudmeijer, O. Macromol. Chem. Phys. 1995, 196, 1431.
413. Chadwick, J. C.; Morini, G.; Balbontin, G.; Sudmeijer, O. Macromol. Chem. Phys. 1998, 199, 1873.
414. Kissin, Y. V.; Ohnishi, R.; Konakazawa, T. Macromol. Chem. Phys. 2004, 205, 284.
415. Kojoh, S.-I.; Kioka, M.; Kashiwa, N.; Itoh, M.; Mizuno, A. Polymer 1995, 36, 5015.
416. Kojoh, S.-I.; Kioka, M.; Kashiwa, N. Eur. Polym. J. 1999, 35, 751.
417. Zhong, C.; Gao, M.; Mao, B. J. Appl. Pol. Sci. 2003, 90, 3980.
418. Kissin, Y. V. J. Polym. Sci. A: Polym. Chem. 2003, 41, 1745.
419. Keii, T.; Terano, M.; Kimura, K.; Ishii, K. Makromol. Chem., Rapid Commun. 1987, 8, 583.
420. Chadwick, J. C.; van der Burgt, F. P. T. J.; Rastogi, S.; Busico, V.; Cipullo, R.; Talarico, G.; Heere, J. J. R. Macromolecules 2004, 37,
9722.
421. Bo hm, L. L.; Bilda, D.; Breuers, W.; Enderle, H. F.; Lecht, R. In Ziegler Catalysts. Recent Scientific Innovations and Technological Improvements;
Fink, G., Mu lhaupt, R., Brintzinger, H. H., Eds.; Springer: Berlin, 1995; p 387.
422. Karol, F. J.; Cann, K. J.; Wagner, B. E. In Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W., Sinn, H.,
Eds.; Springer: Berlin, 1988; p 149.
423. Hsieh, H. L.; McDaniel, M. P.; Martin, J. L.; Smith, P. D.; Fahey, D. R. In Advances in Polyolefins; Seymour, R. B., Cheng, T., Eds.; Plenum
Press: New York, 1985; p 153.
424. Soga, K.; Yanagihara, H. Makromol. Chem. 1988, 189, 2839.
425. Cecchin, G.; Collina, G.; Covezzi, M. (Basell). U.S. Patent 6306996, 2001.
426. Lopez, L. C.; Wilkes, G. L.; Stricklen, P. M.; White, S. A. J. M. S. - Rev. Macromol. Chem. Phys. 1992, C32, 301.
427. Kashiwa, N.; Yoshitake, J. Polym. Bull. 1984, 11, 485.
428. Sacchi, M. C.; Tritto, I.; Locatelli, P.; Fan, Z.-Q.; Forlini, F. Macromol. Chem. Phys. 1994, 195, 2805.
429. Kashiwa, N.; Fukui, K. (Mitsui). U.S. Patent 4659792, 1985.
430. Leatherman, M. D.; Brookhart, M. Macromolecules 2001, 34, 2748.
431. Endo, K.; Otsu, T. J. Polym. Sci. A: Polym. Chem. 1995, 33, 79.
432. Endo, K.; Otsu, T. Makromol. Chem., Rapid Commun. 1990, 11, 663.
433. Otsu, T.; Endo, K. Makromol. Chem., Rapid Commun. 1991, 12, 147.
434. Otsu, T.; Endo, K. Makromol. Chem. 1991, 192, 261.
435. Endo, K.; Ueda, R.; Otsu, T. Makromol. Chem. 1993, 194, 2623.
436. Endo, K.; Fujii, K.; Otsu, T. Makromol. Chem., Rapid Commun. 1991, 12, 409.
437. Karol, F. J. Macromol. Symp. 1995, 89, 563.
438. Ahvenainen, A.; Sarantila, K.; Andtsjo, H. (Neste Oy). World Pat. Appl. WO 92/12181, 1992.
439. Andtsjoe, H.; Pentti, I.; Harlin, A. (Borealis A/S). World Pat. Appl. WO 97/13790, 1997.
440. Covezzi, M.; Mei, G. Chem. Eng. Sci. 2001, 56, 4059.
441. Hamielec, A. E.; Soares, J. B. P. Prog. Polym. Sci. 1996, 21, 651.
442. Stevens, J. C. Stud. Surf. Sci. Catal. 1994, 89, 277.
443. Stevens, J. C. Stud. Surf. Sci. Catal. 1996, 101, 11.
444. Wang, W.; Yan, D.; Charpentier, P. A.; Zhu, S.; Hamielec, A. E. Macromol. Chem. Phys. 1998, 199, 2409.
445. Wang, W.-J.; Yan, D.; Zhu, S.; Hamielec, A. E. Macromolecules 1998, 31, 8677.
446. Malmberg, A.; Liimatta, J.; Lehtinen, A.; Lo fgren, B. Macromolecules 1999, 32, 6687.
447. Kokko, E.; Lehmus, P.; Leino, R.; Luttikhedde, H. J.; Ekholm, P.; Na sman, J. H.; Seppa la , J. Macromolecules 2000, 33, 9200.
448. Kokko, E.; Wang, W.-J.; Seppa la , J.; Zhu, S. J. Polym. Sci. A: Polym. Chem. 2002, 40, 3292.
449. Izzo, L.; Caporaso, L.; Senatore, G.; Oliva, L. Macromolecules 1999, 32, 6913.
450. Melillo, G.; Izzo, L.; Zinna, M.; Tedesco, C.; Oliva, L. Macromolecules 2002, 35, 9256.
451. Izzo, L.; De Riccardis, F.; Alfano, C.; Caporaso, L.; Oliva, L. Macromolecules 2001, 34, 2.
452. Resconi, L. (Montell). World Pat. Appl. WO 00/29416, 2000.
453. Brown, S. J.; Gao, X.; Harrison, D. G.; Koch, L.; Spence, R. E. v. H.; Yap, G. P. A. Organometallics 1998, 17, 5445.
454. Xu, G.; Ruckenstein, E. Macromolecules 1998, 31, 4724.
455. Alt, H. G.; Fo ttinger, K.; Milius, W. J. Organomet. Chem. 1999, 572, 21.
456. Klosin, J.; Kruper, W. J., Jr.,; Nickias, P. N.; Roof, G. R.; De Waele, P. Organometallics 2001, 20, 2663.
457. Irwin, L. J.; Reibenspies, J. H.; Miller, S. A. J. Am. Chem. Soc. 2004, 126, 16716.
458. Foster, P.; Chien, J. C. W.; Rausch, M. D. J. Organomet. Chem. 1997, 545546, 35.
459. Gomes, P. T.; Green, M. L. H.; Martins, A. M.; Mountford, P. J. Organomet. Chem. 1997, 541, 121.
Olefin Polymerizations with Group IV Metal Catalysts 1153
460. Chen, Y.-X.; Fu, P.-F.; Stern, C. L.; Marks, T. J. Organometallics 1997, 16, 5958.
461. Sinnema, P.-J.; Hessen, B.; Teuben, J. H. Macromol. Rapid Commun. 2000, 21, 562.
462. van Leusen, D.; Beetstra, D. J.; Hessen, B.; Teuben, J. H. Organometallics 2000, 19, 4084.
463. Zhang, Y.; Mu, Y.; Lu, C.; Li, G.; Xu, J.; Zhang, Y.; Zhu, D.; Feng, S. Organometallics 2004, 23, 540.
464. Ryabov, A. N.; Voskoboynikov, A. Z. J. Organomet. Chem. 2005, 690, 4213.
465. Amor, F.; Butt, A.; du Plooy, K. E.; Spaniol, T. P.; Okuda, J. Organometallics 1998, 17, 5836.
466. Okuda, J.; Eberle, T.; Spaniol, T. P.; Piquet-Faure , V. J. Organomet. Chem. 1999, 591, 127.
467. Wang, J.; Li, H.; Guo, N.; Li, L.; Stern, C. L.; Marks, T. J. Organometallics 2004, 23, 5112.
468. Kaminsky, W.; Engehausen, R.; Zoumis, K.; Spaleck, W.; Rohrmann, J. Makromol. Chem. 1992, 193, 1643.
469. Tian, J.; Huang, B. Macromol. Rapid Commun. 1994, 15, 923.
470. Janiak, C.; Versteeg, U.; Lange, K. C. H.; Weimann, R.; Hahn, E. J. Organomet. Chem. 1995, 501, 219.
471. Reetz, M. T.; Bru mmer, H.; Kessler, M.; Kuhnigk, J. Chimia 1995, 49, 501.
472. Luttikhedde, H. J. G.; Leino, R. P.; Wile n, C. E.; Na sman, J. H.; Ahlgre n, M. J.; Pakkanen, T. A. Organometallics 1996, 15, 3092.
473. Leino, R.; Luttikhedde, H.; Lehtonen, A.; Wile n, C.-E.; Na sman, J. H. J. Organomet. Chem. 1997, 545546, 219.
474. Leino, R.; Luttikhedde, H.; Lehtonen, A.; Sillanpa a , R.; Penninkangas, A.; Strande n, J.; Mattinen, J.; Na sman, J. H. J. Organomet. Chem.
1998, 558, 171.
475. Thorshaug, K.; Stvneng, J. A.; Rytter, E. Macromolecules 2000, 33, 8136.
476. Mu ller, C.; Lilge, D.; Kristen, M. O.; Jutzi, P. Angew. Chem., Int. Ed. 2000, 39, 789.
477. Schaverien, C. J.; Ernst, R.; Schut, P.; Dall9Occo, T. Organometallics 2001, 20, 3436.
478. Ekholm, P.; Lehmus, P.; Kokko, E.; Haukka, M.; Seppa la , J. V.; Wile n, C.-E. J. Polym. Sci. A: Polym. Chem. 2001, 39, 127.
479. Bryliakov, K. P.; Semikolenova, N. V.; Yudaev, D. V.; Zakharov, V. A.; Brintzinger, H. H.; Ystenes, M.; Rytter, E.; Talsi, E. P. J. Organomet.
Chem. 2003, 683, 92.
480. Mo ller, A. C.; Heyn, R. H.; Blom, R.; Swang, O.; Go rbitz, C. H.; Kopf, J. J. Chem. Soc., Dalton Trans. 2004, 1578.
481. Thornberry, M. P.; Reynolds, N. T.; Deck, P. A.; Fronczek, F. R.; Rheingold, A. L.; Liable-Sands, L. M. Organometallics 2004, 23, 1333.
482. Aitola, E.; Surakka, M.; Repo, T.; Linnolahti, M.; Lappalainen, K.; Kervinen, K.; Klinga, M.; Pakkanen, T. A.; Leskela , M. J. Organomet.
Chem. 2005, 690, 773.
483. Wang, B.; Mu, B.; Deng, X.; Cui, H.; Xu, S.; Zhou, X.; Zou, F.; Li, Y.; Yang, L., Li, Y., et al. Chem. Eur. J. 2005, 11, 669.
484. Bruaseth, I.; Bahr, M.; Gerhard, D.; Rytter, E. J. Polym. Sci. A: Polym. Chem. 2005, 43, 2584.
485. Liu, J.; Stvneng, J. A.; Rytter, E. J. Polym. Sci. A: Polym. Chem. 2001, 39, 3566.
486. Quijada, R.; Dupont, J.; Silveira, D. C.; Miranda, M. S. L.; Scipioni, R. B. Macromol. Rapid Commun. 1995, 16, 357.
487. Welch, M. B.; Palackal, S. J.; Geerts, R. L.; Pettijohn, T. M. (ConocoPhillips). Eur. Pat. Appl. EP 705851, 1996.
488. DallOcco, T.; Resconi, L.; Balbontin, G.; Albizzati, E. (Montell). Eur. Pat. Appl. EP 821011, 1999.
489. Lee, D. L.; Yoon, K.-B. Macromol. Rapid Commun. 1996, 17, 639.
490. Beigzadeh, D.; Soares, J. B. P.; Duever, T. A. Macromol. Rapid Commun. 1999, 20, 541.
491. Reb, A.; Alt, H. G. J. Mol. Catal. A: Chem 2001, 174, 35.
492. Mitani, M.; Oouchi, K.; Hayakawa, M.; Yamada, T.; Mukaiyama, T. Polym. Bull. 1995, 35, 677.
493. Alt, H. G.; Fo ttinger, K.; Milius, W. J. Organomet. Chem. 1998, 564, 115.
494. Doherty, S.; Errington, R. J.; Jarvis, A. P.; Collins, S.; Clegg, W.; Elsegood, M. R. J. Organometallics 1998, 17, 3408.
495. Anti nolo, A.; Carrillo-Hermosilla, F.; Corrochano, A.; Ferna ndez-Baeza, J.; Lara-Sanchez, A.; Ribeiro, M. R.; Lanfranchi, M.; Otero, A.;
Pellinghelli, M. A., Portela, M. F., et al. Organometallics 2000, 19, 2837.
496. Kretschmer, W. P.; Dijkhuis, C.; Meetsma, A.; Hessen, B.; Teuben, J. H. Chem. Commun. 2002, 608.
497. Nomura, K.; Fujii, K. Organometallics 2002, 21, 3042.
498. Nomura, K.; Fujii, K. Macromolecules 2003, 36, 2633.
499. Hessen, B. J. Mol. Catal. A: Chem. 2004, 213, 129.
500. Sanz, M.; Cuenca, T.; Galakhov, M.; Grassi, A.; Bott, R. K. J.; Hughes, D. L.; Lancaster, S. J.; Bochmann, M. Organometallics 2004, 23, 5324.
501. Beddie, C.; Hollink, E.; Wei, P.; Gauld, J.; Stephan, D. W. Organometallics 2004, 23, 5240.
502. Martins, A. M.; Marques, M. M.; Ascenso, J. R.; Dias, A. R.; Duarte, M. T.; Fernandes, A. C.; Fernandes, S.; Ferreira, M. J.; Matos, I.,
Oliveira, M. C., et al. J. Organomet. Chem. 2005, 690, 874.
503. Cady, L. D. Plastics 1987, 25.
504. Suhm, J.; Schneider, M. J.; Mu lhaupt, R. J. Polym. Sci. A: Polym. Chem. 1997, 35, 735.
505. Suhm, J.; Schneider, M. J.; Mu lhaupt, R. J. Mol. Catal. A: Chem. 1998, 128, 215.
506. Wang, W.-J.; Kolodka, E.; Zhu, S.; Hamielec, A. E. J. Polym. Sci. A: Polym. Chem. 1999, 37, 2949.
507. Bonazza, A.; Camurati, I.; Guidotti, S.; Mascellani, N.; Resconi, L. Macromol. Chem. Phys. 2004, 205, 319.
508. Uozumi, T.; Soga, K. Makromol. Chem. 1992, 193, 823.
509. Reybuck, S. E.; Meyer, A.; Waymouth, R. M. Macromolecules 2002, 35, 637.
510. Fink, G.; Richter, W. J. Copolymerization Parameters of Metallocene-catalyzed Copolymerizations. In Polymer Handbook; Brandrup, J.,
Immergut, E. H., Grulke, E. A., Eds.; Wiley: New York, 1999; Vol. 1, p 329.
511. Koivuma ki, J.; Seppa la , J. V. Polymer 1993, 34, 1958.
512. Lehtinen, C.; Starck, P.; Lo fgren, B. J. Polym. Sci. A: Polym. Chem. 1997, 35, 307.
513. Sarzotti, D. M.; Soares, J. B. P.; Penlidis, A. J. Polym. Sci. B: Polym. Phys. 2002, 40, 2595.
514. Miyata, H.; Yamaguchi, M.; Akashi, M. Polymer 2001, 42, 5763.
515. Ma der, D.; Heinemann, J.; Walter, P.; Mu lhaupt, R. Macromolecules 2000, 33, 1254.
516. Simanke, A. G.; Galland, G. B.; Freitas, L. L.; da Jornada, J. A. H.; Quijada, R. Macromol. Chem. Phys. 2001, 202, 172.
517. Wigum, H.; Solli, K.-A.; Stvneng, J. A.; Rytter, E. J. Polym. Sci. A: Polym. Chem. 2003, 41, 1622.
518. Lehmus, P.; Ha rkki, O.; Leino, R.; Luttikhedde, H. J.; Na sman, J. H.; Seppa la , J. Macromol. Chem. Phys. 1998, 199, 1965.
519. Kim, I.; Kim, S. Y.; Lee, M. H.; Do, Y.; Won, M.-S. J. Polym. Sci. A: Polym. Chem. 1999, 37, 2763.
520. Seppa la , J. V.; Koivuma ki, J.; Liu, X. J. Polym. Sci. A: Polym. Chem. 1993, 31, 3447.
521. Koivuma ki, J.; Seppa la , J. V. Macromolecules 1994, 27, 2008.
522. Mauler, R. S.; Galland, G. B.; Scipioni, R. B.; Quijada, R. Polym. Bull. 1996, 37, 469.
523. Simanke, A. G.; Alamo, R. G.; Galland, G. B.; Mauler, R. S. Macromolecules 2001, 34, 6959.
1154 Olefin Polymerizations with Group IV Metal Catalysts
524. Koivuma ki, J.; Seppa la , J. Macromolecules 1993, 26, 5535.
525. Lehmus, P.; Kokko, E.; Ha rkki, O.; Leino, R.; Luttikhedde, H. J. G.; Na sman, J. H.; Seppa la , J. V. Macromolecules 1999, 32, 3547.
526. Wigum, H.; Tangen, L.; Stvneng, J. A.; Rytter, E. J. Polym. Sci. A: Polym. Chem. 2000, 38, 3161.
527. Britto, M. L.; Galland, G. B.; dos Santos, J. H. Z.; Forte, M. C. Polymer 2001, 6355.
528. Dankova, M.; Waymouth, R. M. Macromolecules 2003, 36, 3815.
529. Bruaseth, I.; Rytter, E. Macromolecules 2003, 36, 3026.
530. Reybuck, S. E.; Waymouth, R. M. Macromolecules 2004, 37, 2342.
531. Soga, K.; Uozumi, T.; Nakamura, S.; Toneri, T.; Teranishi, T.; Sano, T.; Arai, T. Macromol. Chem. Phys. 1996, 197, 4237.
532. Koivuma ki, J. Polym. Bull. 1996, 36, 7.
533. Schneider, M. J.; Suhm, J.; Mu lhaupt, R.; Prosenc, M. H.; Brintzinger, H.-H. Macromolecules 1997, 30, 3164.
534. Kaminsky, W.; Freidanck, F. Macromol. Symp. 2002, 183, 89.
535. Kaminsky, W.; Piel, C. J. Mol. Catal. A: Chem. 2004, 213, 15.
536. Koivumki, J.; Fink, G.; Sepp, J.a . Macromolecules 1994, 27, 6254.
537. Nomura, K.; Naga, N.; Miki, M.; Yanagi, K. Macromolecules 1998, 31, 7588.
538. Sahoo, S. K.; Zhang, T.; Reddy, D. V.; Rinaldi, P. L.; McIntosh, L. H.; Quirk, R. P. Macromolecules 2003, 36, 4017.
539. Mahanthappa, M. K.; Cole, A. P.; Waymouth, R. M. Organometallics 2004, 23, 836.
540. Chen, Y.-X.; Marks, T. J. Organometallics 1997, 16, 3649.
541. Ashe, A. J. III; Fang, X.; Kampf, J. W. Organometallics 1999, 18, 1363.
542. Xu, G.; Cheng, D. Macromolecules 2001, 34, 2040.
543. Sehanobish, K.; Patel, R. M.; Croft, B. A.; Chum, S.; Kao, C. I. J. Appl. Polym. Sci. 1994, 51, 887.
544. Minick, J.; Moet, A.; Hiltner, A.; Baer, E.; Chum, S. P. J. Appl. Polym. Sci. 1995, 58, 1371.
545. Bensason, S.; Minick, J.; Moet, A.; Chum, S.; Hiltner, A.; Baer, E. J. Polym. Sci. B: Polym. Phys. 1996, 34, 1301.
546. Bensason, S.; Stepanov, E. V.; Chum, S.; Hiltner, A.; Baer, E. Macromolecules 1997, 30, 2436.
547. Nomura, K.; Naga, N.; Miki, M.; Yanagi, K.; Imai, A. Organometallics 1998, 17, 2152.
548. Nomura, K.; Komatsu, T.; Imanishi, Y. J. Mol. Catal. A: Chem. 2000, 159, 127.
549. Hanaoka, H.; Senda, T.; Yoshikawa, E.; Kobayashi, S. (Sumitomo Chemical). Eur. Pat. Appl. EP 1475383, 2004.
550. Mani, R.; Burns, C. M. Polymer 1993, 34, 1941.
551. Kaminsky, W.; Arrowsmith, D.; Winkelbach, H. Polym. Bull. 1996, 36, 577.
552. Chung, T. C.; Dong, J. Y. Macromolecules 2002, 35, 2868.
553. Naga, N.; Toyota, A.; Ogino, K. J. Polym. Sci. A: Polym. Chem. 2005, 43, 911.
554. Naga, N.; Toyota, A. Polymer 2004, 45, 7513.
555. Hagihara, H.; Tsuchihara, K.; Takeuchi, K.; Murata, M. J. Polym. Sci. A: Polym. Chem. 2004, 42, 52.
556. Aaltonen, P.; Lo fgren, B. Macromolecules 1995, 28, 5353.
557. Aaltonen, P.; Fink, G.; Lo fgren, B.; Seppa la , J. Macromolecules 1996, 29, 5255.
558. Imuta, J.; Kashiwa, N.; Toda, Y. J. Am. Chem. Soc. 2002, 124, 1176.
559. Napoli, M.; Costabile, C.; Pragliola, S.; Longo, P. Macromolecules 2005, 38, 5493.
560. Bergemann, C.; Cropp, R.; Luft, G. J. Mol. Catal. A: Chem. 1997, 116, 317.
561. Sernetz, F. G.; Mu lhaupt, R.; Waymouth, R. M. Polym. Bull. 1997, 38, 141.
562. Pietika inen, P.; Va a na nen, T.; Seppa la , J. V. Eur. Polym. J. 1999, 35, 1047.
563. Kokko, E.; Pietikainen, P.; Koivunen, J.; Seppa la , J. J. Polym. Sci. A: Polym. Chem. 2001, 39, 3805.
564. Naga, N.; Toyota, A. Macromol. Rapid Commun. 2004, 25, 1623.
565. Pietika inen, P.; Seppa la , J. V.; Ahjopalo, L.; Pietila , L.-O. Eur. Polym. J. 2000, 36, 183.
566. Jin, H.-J.; Choi, C.-H.; Park, E.-S.; Lee, I.-M.; Yoon, J.-S. J. Appl. Polym. Sci. 2002, 84, 1048.
567. Barnhart, R. W.; Bazan, G. C. J. Am. Chem. Soc. 1998, 120, 1082.
568. Kolodka, E.; Wang, W.-J.; Zhu, S.; Hamielec, A. Macromol. Rapid Commun. 2003, 24, 311.
569. Shiono, T.; Moriki, Y.; Ikeda, T.; Soga, K. Macromol. Chem. Phys. 1997, 198, 3229.
570. Quijada, R.; Rojas, R.; Bazan, G.; Komon, Z. J. A.; Mauler, R. S.; Galland, G. B. Macromolecules 2001, 34, 2411.
571. Beigzadeh, D.; Soares, J. B. P.; Duever, T. A. Macromol. Symp. 2001, 173, 179.
572. de Wet-Roos, D.; Dixon, J. T. Macromolecules 2004, 37, 9314.
573. Zhang, Z.; Cui, N.; Lu, Y.; Ke, Y.; Hu, Y. J. Polym. Sci. A: Polym. Chem. 2005, 43, 984.
574. Arndt, M.; Kaminsky, W.; Schauwienold, A. M.; Weingarten, U. Macromol. Chem. Phys. 1998, 199, 1135.
575. Fan, W.; Leclerc, M. K.; Waymouth, R. M. J. Am. Chem. Soc. 2001, 123, 9555.
576. Heuer, B.; Kaminsky, W. Macromolecules 2005, 38, 3054.
577. Lehtinen, C.; Lo fgren, B. Eur. Polym. J. 1997, 33, 115.
578. Galimberti, M.; Mascellani, N.; Piemontesi, F.; Camurati, I. Macromol. Rapid Commun. 1999, 20, 214.
579. Galimberti, M.; Piemontesi, F.; Mascellani, N.; Camurati, I.; Fusco, O.; Destro, M. Macromolecules 1999, 32, 7968.
580. Wang, W.-J.; Zhu, S. Macromolecules 2000, 33, 1157.
581. Leclerc, M. K.; Waymouth, R. M. Angew. Chem., Int. Ed. 1998, 37, 922.
582. Jin, J.; Uozumi, T.; Sano, T.; Teranishi, T.; Soga, K.; Shiono, T. Macromol. Rapid Commun. 1998, 19, 337.
583. Galimberti, M.; Piemontesi, F.; Baruzzi, G.; Mascellani, N.; Camurati, I.; Fusco, O. Macromol. Chem Phys. 2001, 202, 2029.
584. Djupfors, R.; Starck, P.; Lo fgren, B. Eur. Polym. J. 1998, 34, 941.
585. Longo, P.; Siani, E.; Pragliola, S.; Monaco, G. J. Polym. Sci. A: Polym. Chem. 2002, 40, 3249.
586. Schottek, J.; Oberhoff, M.; Bingel, C.; Fischer, D.; Weiss, H.; Winter, A.; Fraaije, V. (Basell). World Pat. Appl. WO 01/48034, 2001.
587. Kuchta, M. C.; Stehling, U. M.; Li, R. T.; Haygood, W. T.; Burkhardt, T. J. (Exxon-Mobil). World Pat. Appl. WO 02/02575, 2002.
588. Resconi, L.; Ferrari, P.; Cecchin, G. (Basell Polyolefins). World Pat. Appl. WO 05/023889, 2005.
589. Wang, W.-J.; Zhu, S.; Park, S.-J. Macromolecules 2000, 33, 5770.
590. Kaminsky, W.; Miri, M. J. Poyml. Sci., Part A: Polym. Chem. 1985, 23, 2151.
591. Parikh, D. R.; Edmondson, M. S.; Smith, B. W.; Winter, J. M.; Castille, M. J.; Magee, J. M.; Patel, R. M.; Karajala, T. P. Structure and
Properties of Single-site Constrained Geometry Ethylene-Propylene-Diene (EPDM) Elastomers. In Metallocene-catalyzed Polymers
Materials, Properties, Processing & Markets; Benedikt, G. M., Goodall, B. L., Eds.; Plastics Design Library: New York, 1998; p 113.
Olefin Polymerizations with Group IV Metal Catalysts 1155
592. Ho, T.; Martin, J. M. Structure, Properties and Applications of Polyolefin Elastomers Produced by Constrained Geometry Catalysts. In
Metallocene-based Polyolefins: Preparation, Properties and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 2; 175.
593. Nomura, K.; Itagaki, K.; Fujiki, M. Macromolecules 2005, 38, 2053.
594. Guerra, G.; Longo, P.; Corradini, P.; Cavallo, L. J. Am. Chem. Soc. 1999, 121, 8651.
595. Longo, P.; Grisi, F.; Guerra, G.; Cavallo, L. Macromolecules 2000, 33, 4647.
596. Kaminsky, W.; Arndt-Rosenau, M. Homo- and Copolymerization of Cycloolefins by Metallocene Catalysts. In Metallocene-based Polyolefins:
Preparation, Properties and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 2, p 89.
597. Wang, W.; Fujiki, M.; Nomura, K. J. Am. Chem. Soc. 2005, 127, 4582.
598. Kaminsky, W.; Noll, A. Polymerization of Phenyl-substituted Cyclic Olefins with Metallocene/MAO Catalysts. In Ziegler Catalysts; Fink, G.,
Mu lhaupt, R., Brintzinger, H.-H., Eds.; Springer: Berlin, 1995; p 149.
599. Marathe, S.; Sivaram, S. Macromolecules 1994, 27, 1083.
600. Wang, T.-Y.; Lin, C.-H.; Jiang, G. J. Polym. Prepr. 1996, 37, 641.
601. Kaminsky, W.; Beulich, I.; Arndt-Rosenau, M. Macromol. Symp. 2001, 173, 211.
602. Radhakrishnan, K.; Sivaram, S. Macromol. Chem. Phys. 1999, 200, 858.
603. Naga, N. J. Polym. Sci. A: Polym. Chem. 2005, 43, 1285.
604. Jerschow, A.; Ernst, E.; Wolfgang, H.; Mu ller, N. Macromolecules 1995, 28, 7095.
605. Fujita, M.; Coates, G. W. Macromolecules 2002, 35, 9640.
606. Lavoie, A. R.; Ho, M. H.; Waymouth, R. M. Chem. Commun. 2003, 864.
607. Ticona Engineering polymers. www.ticona.com.
608. Ruchatz, D.; Fink, G. Macromolecules 1998, 31, 4669.
609. Ruchatz, D.; Fink, G. Macromolecules 1998, 31, 4674.
610. Ruchatz, D.; Fink, G. Macromolecules 1998, 31, 4681.
611. Ruchatz, D.; Fink, G. Macromolecules 1998, 31, 4684.
612. Cherdron, H.; Brekner, M.-J.; Osan, F. Angew. Macromol. Chem. 1994, 223, 121.
613. Lee, B. Y.; Kim, Y. H.; Won, Y. C.; Han, J. W.; Suh, W. H.; Lee, I. S.; Chung, Y. K.; Song, K. H. Organometallics 2002, 21, 1500.
614. Lee, H.; Hong, S.-D.; Park, Y.-W.; Jeong, B.-G.; Nam, D.-W.; Jung, Y.; Jung, M. W.; Song, K. H. J. Organomet. Chem. 2004, 689, 3402.
615. Lee, S.-G.; Hong, S.-D.; Park, Y.-W.; Jeong, B.-G.; Nam, D.-W.; Jung, H. Y.; Lee, H.; Song, K. H. Polymer 2004, 689, 2586.
616. Cho, E. S.; Joung, U. G.; Lee, B. Y.; Lee, H.; Park, Y.-W.; Lee, C. H.; Shin, D. M. Organometallics 2004, 23, 4693.
617. McKnight, A. L.; Waymouth, R. M. Macromolecules 1999, 32, 2816.
618. Arndt-Rosenau, M.; Beulich, I. Macromolecules 1999, 32, 7335.
619. Tritto, I.; Boggioni, L.; Ferro, D. R. Macromolecules 2004, 37, 9681.
620. De Rosa, C.; Buono, A.; Auriemma, F.; Grassi, A. Macromolecules 2004, 37, 9489.
621. Hasan, T.; Ikeda, T.; Shiono, T. Macromolecules 2004, 37, 8503.
622. MacKnight, W. J.; Rische, T.; Waddon, A. J.; Dickinson, L. C. Macromolecules 1998, 31, 1871.
623. DAniello, C.; de Candia, F.; Oliva, L.; Vittoria, V. J. Appl. Polym. Sci. 1995, 58, 1701.
624. Guest, M. J.; Cheung, Y. W.; Diehl, C. F.; Hoenig, S. M. Structure, Properties and Applications of EthyleneStyrene Interpolymers. In
Metallocene-based Polyolefins: Preparation, Properties and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 2, p 271.
625. Venditto, V.; De Tullio, G.; Izzo, L.; Oliva, L. Macromolecules 1998, 31, 4027.
626. Pellecchia, C.; Oliva, L. Rubber Chem. Technol. 1999, 72, 553.
627. Sernetz, F. G.; Mu lhaupt, R.; Fokken, S.; Okuda, J. Macromolecules 1997, 30, 1562.
628. Longo, P.; Grassi, A.; Oliva, L. Makromol. Chem. 1990, 191, 2387.
629. Oliva, L.; Mazza, S.; Longo, P. Macromol. Chem. Phys. 1996, 197, 3115.
630. Aaltonen, P.; Seppa la , J. Eur. Polym. J. 1994, 30, 683.
631. Aaltonen, P.; Seppa la , J. Eur. Polym. J. 1995, 31, 79.
632. Pellecchia, C.; Pappalardo, D.; D9Arco, M.; Zambelli, A. Macromolecules 1996, 29, 1158.
633. Pellecchia, C.; Pappalardo, D.; Oliva, L.; Mazzeo, M.; Gruter, G.-J. Macromolecules 2000, 33, 2807.
634. Pellecchia, C.; Mazzeo, M.; Gruter, G.-J. Macromol. Rapid Commun. 1999, 20, 337.
635. Xu, G.; Lin, S. Macromolecules 1997, 30, 685.
636. Chu, P. P.; Tseng, H. S.; Chen, Y. P.; Yu, D. D. Polymer 2000, 41, 8271.
637. Wu, Q.; Ye, Z.; Gao, Q. H.; Lin, S. G. Macromol. Chem. Phys. 1998, 199, 1715.
638. Lee, D.-H.; Yoon, K.-B.; Kim, H.-J.; Woo, S.-S.; Noh, S. K. J. Appl. Polym. Sci. 1998, 67, 2187.
639. Nomura, K.; Komatsu, T.; Imanishi, Y. Macromolecules 2000, 33, 8122.
640. Nomura, K.; Okumura, H.; Komatsu, T.; Naga, N. Macromolecules 2002, 35, 5388.
641. Munoz-Escalona, A.; Cruz, V.; Mena, N.; Martinez, S.; Martinez-Salazar, J. Polymer 2002, 43, 7017.
642. Stevens, J. C.; Timmers, F. J.; Wilson, D. R.; Schmidt, G. F.; Nickias, P. N.; Rosen, R. K.; Knight, G. W.; Lai, S. Y. (Dow Chemical Co.).
Eur. Pat. Appl. EP 416815, 1991.
643. Sernetz, F. G.; Mu lhaupt, R.; Waymouth, R. M. Macromol. Chem. Phys. 1996, 197, 1071.
644. Sernetz, F. G.; Mu lhaupt, R.; Amor, F.; Eberle, T.; Okuda, J. J. Polym. Sci. A: Polym. Chem. 1997, 35, 1571.
645. Xu, G. Macromolecules 1998, 31, 2395.
646. Sukhova, T. A.; Panin, A. N.; Babkina, O. N.; Bravaya, N. M. J. Polym. Sci. A: Polym. Chem. 1999, 37, 1083.
647. Nomura, K.; Okumura, H.; Komatsu, T.; Naga, N.; Imanishi, Y. J. Mol. Catal. A: Chem. 2002, 190, 225.
648. Yang, S. H.; Jo, W. H.; Noh, S. K. J. Chem. Phys. 2003, 119, 1824.
649. Skeril, R.; Sindelar, P.; Salajka, Z.; Varga, V.; Cisarova, I.; Pinkas, J.; Horacek, M.; Mach, K. J. Mol. Catal. A: Chem. 2004, 224, 97.
650. Gentil, S.; Pirio, N.; Meunier, P.; Gallou, F.; Paquette, L. A. Eur. Polym. J. 2004, 40, 2241.
651. Martinez, S.; Exposito, M. T.; Ramos, J.; Cruz, V.; Martinez, M. C.; Lopez, M.; Munoz-Escalona, A.; Martinez-Salazar, J. J. Polym. Sci. A:
Polym. Chem. 2005, 43, 711.
652. Ramos, J.; Munoz-Escalona, A.; Martinez, S.; Martinez-Salazar, J.; Cruz, V. J. Chem. Phys. 2005, 122, 074901.
653. Inoue, N.; Shiomura, T.; Kouno, M. (Mitsui Toatsu Chemicals Inc.). Eur. Pat. Appl. EP 0108824, 1993.
654. Oliva, L.; Caporaso, L.; Pellecchia, C.; Zambelli, A. Macromolecules 1995, 28, 4665.
655. Oliva, L.; Longo, P.; Izzo, L.; Di Serio, M. Macromolecules 1997, 30, 5616.
1156 Olefin Polymerizations with Group IV Metal Catalysts
656. Caporaso, L.; Izzo, L.; Sisti, I.; Oliva, L. Macromolecules 2002, 35, 4866.
657. Arai, T.; Ohtsu, T.; Suzuki, S. Macromol. Rapid Commun. 1998, 19, 327.
658. Albers, I.; Kaminsky, W.; Weingarten, U.; Werner, R. Catal. Commun. 2002, 3, 105.
659. Martinez, S.; Cruz, V.; Munoz-Escalona, A.; Martinez-Salazar, J. Polymer 2002, 43, 295.
660. Exposito, M. T.; Martinez, S.; Ramos, J.; Cruz, V. L.; Lopez, M.; Munoz-Escalona, A.; Haider, N.; Martinez-Salazar, J. Polymer 2004, 45,
9029.
661. Capacchione, C.; DAcunzi, M.; Motta, O.; Oliva, L.; Proto, A.; Okuda, J. Macromol. Chem. Phys. 2004, 205, 370.
662. Ishiyama, T.; Miyoshi, K.; Nakazawa, H. J. Mol. Catal. A: Chem. 2004, 221, 41.
663. Zhang, H.; Nomura, K. J. Am. Chem. Soc. 2005, 127, 9364.
664. Guo, N.; Li, L.; Marks, T. J. J. Am. Chem. Soc. 2004, 126, 6542.
665. Chung, T. C.; Lu, H. L. J. Polym. Sci. A: Polym. Chem. 1997, 35, 575.
666. Chung, T. C.; Lu, H. L. J. Polym. Sci. A: Polym. Chem. 1998, 36, 1017.
667. Lu, H. L.; Hong, S.; Chung, T. C. Macromolecules 1998, 31, 2028.
668. Sernetz, F. G.; Mu lhaupt, R. J. Polym. Sci. A: Polym. Chem. 1997, 35, 2549.
669. Resconi, L. Polym. Prepr. 2002, 43, 303.
670. Resconi, L.; Silvestri, R. Polypropylene, Atactic (High Molecular Weight). In The Polymeric Materials Encyclopedia; Salamone, J. C., Ed.
CRC Press: Boca Raton, 1996; p 6609.
671. Kawamura, H.; Hayashi, T. Y.; Inoue, Y.; Ch uj o, R. Macromolecules 1989, 22, 2181.
672. Busico, V.; Corradini, P.; De Martino, L.; Graziano, F.; Iadicicco, A. Makromol. Chem. 1991, 192, 49.
673. Kakugo, M.; Miyatake, T.; Kawai, K.; Shiga, A.; Mizunuma, K. (Sumitomo Chemical). Eur. Pat. Appl. EP 241560, 1987.
674. Kakugo, M.; Miyatake, T.; Mizunuma, K.; Yagi, Y. (Sumitomo Chemical). Eur. Pat. Appl. EP 371411, 1990.
675. Smith, T.; Ames, W.; Holliday, R.; Pearson, N. (Eastman Kodak). Eur. Pat. Appl. EP 232201, 1987.
676. Smith, C. (Himont). Eur. Pat. Appl. EP 423786, 1991.
677. Pellon, B.; Allen, G. (Rexene). Eur. Pat. Appl. EP 475307, 1992.
678. Job, R. (Shell Oil). U. S. Patent 5270410, 1993.
679. Pellon, B. J. In Proceedings of the SPO 93; Schotland Business Res. Ed.; Skillman, 1993; p 401.
680. Robe, G. R. Adhesive Age 1993, 26 February.
681. Schmidt, R.; Alt, H. G. J. Organomet. Chem. 2001, 621, 304.
682. Hagihara, H.; Shiono, T.; Ikeda, T. Macromol. Chem. Phys. 1998, 199, 2439.
683. Schmidt, R.; Deppner, M.; Alt, H. G. J. Mol. Catal. A: Chem. 2001, 172, 43.
684. Yoon, J.-S.; Kwon, O.-J.; Lee, H.-K.; Lee, I.-M.; Lee, G.-Y.; Kang, M.-S.; Xue, M. Eur. Polym. J. 1998, 34, 879.
685. Naga, N.; Mizunuma, K. Polymer 1998, 39, 2703.
686. Zhang, F.; Mu, Y.; Zhao, L.; Zhang, Y.; Bu, W.; Chen, C.; Zhai, H.; Hong, H. J. Organomet. Chem. 2000, 613, 68.
687. Chen, Y.-X.; Rausch, M. D.; Chien, J. C. W. Macromolecules 1995, 28, 5399.
688. Resconi, L. Synthesis of Atactic Polypropylene Using Metallocene Catalysts. In Metallocene-based Polyolefins: Preparation, Properties and
Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 1, p 467.
689. Collins, S.; Gauthier, W. J.; Holden, D. A.; Kuntz, B. A.; Taylor, N. J.; Ward, D. G. Organometallics 1991, 10, 2061.
690. Winter, A.; Antberg, M.; Bachmann, B.; Dolle, V.; Ku ber, F.; Rohrmann, J.; Spaleck, W. (Hoechst). Eur. Pat. Appl. EP 584609, 1994.
691. Csok, Z.; Liguori, D.; Sessa, I.; Zannoni, C.; Zambelli, A. Macromol. Chem. Phys. 2004, 205, 1231.
692. Wu, Q.; Ye, Z.; Gao, Q.; Lin, S. J. Polym. Sci., A: Polym. Chem. 1998, 36, 2051.
693. Sassmannshausen, J.; Bochmann, M.; Ro sch, J.; Lilge, D. J. Organomet. Chem. 1997, 548, 23.
694. Ewart, S. W.; Sarsfield, M. J.; Jeremic, D.; Tremblay, T. L.; Williams, E. F.; Baird, M. C. Organometallics 1998, 17, 1502.
695. Dove, A. P.; Xie, X.; Waymouth, R. M. Chem. Commun. 2005, 2152.
696. Sassmannshausen, J.; Powell, A. K.; Anson, C. E.; Wocadlo, S.; Bochmann, M. J. Organomet. Chem. 1999, 592, 84.
697. Flores, J. C.; Chien, J. C. W.; Rausch, M. D. Organometallics 1994, 13, 4140.
698. Flores, J. C.; Chien, J. C. W.; Rausch, M. D. Macromolecules 1996, 29, 8030.
699. Ewart, S. W.; Parent, M. A.; Baird, M. C. J. Polym. Sci. A: Polym. Chem. 1999, 37, 4386.
700. Fukui, Y.; Murata, M.; Soga, K. Macromol. Rapid Commun. 1999, 20, 637.
701. Fukui, Y.; Murata, M. Appl. Catal. A: Gen. 2002, 237, 110.
702. Xie, M.; Wu, Q.; Lin, S. Macromol. Rapid Commun. 1999, 20, 167.
703. Zambelli, A.; Csok, Z.; Sessa, I. Macromol. Rapid Commun. 2005, 26, 519.
704. Erker, G.; Fritze, C. Angew. Chem., Int. Ed. Engl. 1992, 31, 199.
705. Canich, J. A. M. (Exxon). PCT Int. Appl. WO 96/00244, 1996.
706. McKnight, A. L.; Masood, M. A.; Waymouth, R. M.; Strauss, D. A. Organometallics 1997, 16, 2879.
707. Kamigaito, M.; Lal, T. K.; Waymouth, R. M. J. Polym. Sci. A: Polym. Chem. 2000, 38, 46494660.
708. Resconi, L.; Camurati, I.; Grandini, C.; Rinaldi, M.; Mascellani, N.; Traverso, O. J. Organomet. Chem. 2002, 664, 5.
709. De Rosa, C.; Auriemma, F.; Perretta, C. Macromolecules 2004, 37, 6843.
710. Longo, P.; Amendola, A. G.; Fortunato, E.; Boccia, A. C.; Zambelli, A. Macromol. Rapid Commun. 2001, 22, 339.
711. De Rosa, C.; Auriemma, F.; Circelli, T.; Longo, P.; Boccia, A. C. Macromolecules 2003, 36, 3465.
712. Kleinschmidt, R.; Griebenow, Y.; Fink, G. J. Mol. Catal. A: Chem. 2000, 157, 83.
713. De Rosa, C.; Auriemma, F.; Spera, C.; Talarico, G.; Tarallo, O. Macromolecules 2004, 37, 1441.
714. Ewen, J. A.; Haspeslagh, L.; Atwood, J.; Zhang, H. J. Am. Chem. Soc. 1987, 109, 6544.
715. Ushioda, T.; Fujita, H.; Saito, J. In Proceedings of the SPO 97; Schotland Business Res. Ed.: Skillman, 1997, p 101.
716. Nakano, M.; Ushioda, T.; Yamazaki, H.; Uwai, T.; Kimura, M.; Ohgi, Y.; Yamamoto, K. (Chisso). German Patent DE 10125356, 2002.
717. Ewen, J. A.; Jones, R. L.; Elder, M. J.; Rheingold, A. L.; Liable-Sands, L. M. J. Am. Chem. Soc. 1998, 120, 10786.
718. Ewen, J. A.; Jones, R. L.; Elder, M. J. Expanding the Scope of Metallocene Catalysis: Beyond Indenyl and Fluorenyl Derivatives. In
Metalorganic Catalysts for Synthesis and Polymerization; Kaminsky, W., Ed.; Springer: Berlin, 1999, p 150.
719. Ewen, J. A.; Elder, M. J.; Jones, R. L. (Basell). World Pat. Appl. WO 01/44318, 2001.
720. Ewen, J. A.; Elder, M. J.; Jones, R. L.; Rheingold, A. L.; Liable-Sands, L. M.; Sommer, R. D. J. Am. Chem. Soc. 2001, 123, 4763.
721. Elder, M. J.; Okumura, Y.; Jones, R. L.; Richter, B.; Seidel, N. Kinet. Catal. 2006, 47, 192.
Olefin Polymerizations with Group IV Metal Catalysts 1157
722. Resconi, L.; Fritze, C. Metallocene Catalysts for Propylene Polymerization. In Polypropylene Handbook, 2nd ed.; Pasquini, N., Ed. Hanser:
Munich, 2005, p 107.
723. Hlatky, G. G. Chem. Rev. 2000, 100, 1347.
724. Fink, G.; Steinmetz, B.; Zechlin, J.; Przybyla, C.; Tesche, B. Chem. Rev. 2000, 100, 1377.
725. Resconi, L.; Guidotti, S.; Camurati, I.; Frabetti, R.; Focante, F.; Nifantev, I. E.; Laishevtsev, I. P. Macromol. Chem. Phys. 2005, 206, 1405.
726. Ju ngling, S.; Koltzenburg, S.; Mu lhaupt, R. J. Polym. Sci. A: Polym. Chem. 1997, 35, 1.
727. Burkhardt, T. J.; Hart, J. R.; Haygood, W. T.; Li, R. T. (Exxon-Mobil). World Pat. Appl. WO 03/050131, 2003.
728. Kashiwa, N.; Kojoh, S.; Imuta, J.; Tsutsui, T. Characterization of PP Prepared with the Latest Metallocene and MgCl
2
-Supported TiCl
4
Catalyst Systems. In Metalorganic Catalysts for Synthesis and Polymerization; Kaminsky, W., Ed.; Springer: Berlin, 1999; p 30.
729. Bingel, C.; Goeres, M.; Fraaije, V.; Winter, A. (Targor). World Pat. Appl. WO 98/40416, 1998.
730. Resconi, L.; Ciaccia, E.; Fait, A. (Basell). World Pat. Appl. WO 04/092230, 2004.
731. Ushioda, T.; Nakano, M. In Proceedings of the MetCon: Houston, Texas, 2002.
732. Iwama, N.; Uchino, H.; Osano, Y. T.; Sugano, T. Organometallics 2004, 23, 3267.
733. Iwama, N.; Osano, Y. T. Organometallics 2005, 24, 132.
734. Spaleck, W.; Ku ber, F.; Winter, A.; Rohrmann, J.; Bachmann, B.; Antberg, M.; Dolle, V.; Paulus, E. F. Organometallics 1994, 13, 954.
735. Frediani, M.; Kaminsky, W. Macromol. Chem. Phys. 2003, 204, 1941.
736. Fischer, D.; Mu lhaupt, R. Macromol. Chem. Phys. 1994, 195, 1433.
737. Thomann, R.; Wang, C.; Kressler, J.; Mu lhaupt, R. Macromolecules 1996, 29, 8425.
738. De Rosa, C.; Auriemma, F.; Di Capua, A.; Resconi, L.; Guidotti, S.; Camurati, I.; Nifantev, I. E.; Laishevtsev, I. P. J. Am. Chem. Soc. 2004,
126, 17040.
739. Rieger, B.; Mu, X.; Mallin, D. T.; Chien, J. C. W.; Rausch, M. D. Macromolecules 1990, 23, 3559.
740. Miller, S. A.; Bercaw, J. E. Organometallics 2002, 21, 934.
741. Razavi, A.; Bellia, V.; De Brauwer, Y.; Hortmann, K.; Peters, L.; Sirole, S.; Van Belle, S.; Thewalt, U. Macromol. Chem. Phys. 2004, 205,
347.
742. Nifantev, I. E.; Laishevtsev, I.; Ivchenko, P. V.; Kashulin, I. A.; Guidotti, S.; Piemontesi, F.; Camurati, I.; Resconi, L.; Klusener, P. A. A.,
Rijsemus, J. J. H., et al. Macromol. Chem. Phys. 2004, 205, 2275.
743. Kawai, K.; Yamashita, M.; Tohi, Y.; Kawahara, N.; Michiue, K.; Kaneyoshi, H.; Mori, R. (Mitsui Chemicals). Eur. Pat. Appl. EP 1138687,
2001.
744. Ewen, J. A.; Elder, M. J. Isospecific Pseudo-helical Zirconocenium Catalysts. In Ziegler Catalysts; Fink, G., Mu lhaupt, R., Brintzinger, H.-H.,
Eds.; Springer: Berlin, 1995, p 99.
745. Spaleck, W.; Aulbach, M.; Bachmann, B.; Ku ber, F.; Winter, A. Macromol. Symp. 1995, 89, 237.
746. Thomas, E. J.; Chien, J. C. W.; Rausch, M. D. Macromolecules 2000, 33, 1546.
747. Thomas, E. J.; Rausch, M. D.; Chien, J. C. W. Organometallics 2000, 19, 4077.
748. Coates, G. W.; Waymouth, R. M. Science 1995, 267, 217.
749. Lin, S.; Waymouth, R. M. Acc. Chem. Res. 2002, 35, 765.
750. Bravaya, N. M.; Nedorezova, P. M.; Tsvetkova, V. I. Russ. Chem. Rev. 2002, 71, 49.
751. Auriemma, F.; De Rosa, C.; Boscato, T.; Corradini, P. Macromolecules 2001, 34, 4815.
752. Bruce, M. D.; Coates, G. W.; Hauptman, E.; Waymouth, R. M.; Ziller, J. W. J. Am. Chem. Soc. 1997, 119, 11174.
753. Kravchenko, R.; Masood, A.; Waymouth, R. M. Organometallics 1997, 16, 3635.
754. Petoff, J. L. M.; Bruce, M. D.; Waymouth, R. M.; Masood, A.; Lal, T. K.; Quan, R. W.; Behrend, S. J. Organometallics 1997, 16, 5909.
755. Kravchenko, R.; Masood, A.; Waymouth, R. M.; Myers, C. L. J. Am. Chem. Soc. 1998, 120, 2039.
756. Lin, S.; Hauptman, E.; Lal, T. K.; Waymouth, R. M.; Quan, R. W.; Ernst, A. B. J. Mol. Catal. A: Chem. 1998, 136, 23.
757. Petoff, J. L. M.; Agoston, T.; Lal, T. K.; Waymouth, R. M. J. Am. Chem. Soc. 1998, 120, 11316.
758. Lin, S.; Waymouth, R. M. Macromolecules 1999, 32, 8283.
759. Tagge, C. D.; Kravchenko, R. L.; Lal, T. K.; Waymouth, R. M. Organometallics 1999, 18, 380.
760. Witte, P.; Lal, T. K.; Waymouth, R. M. Organometallics 1999, 18, 4147.
761. Lin, S.; Kravchenko, R.; Waymouth, R. M. J. Mol. Catal. A: Chem. 2000, 158, 423.
762. Lin, S.; Tagge, C. D.; Waymouth, R. M.; Nele, M.; Collins, S.; Pinto, J. C. J. Am. Chem. Soc. 2000, 122, 11275.
763. Nele, M.; Collins, S.; Dias, M. L.; Pinto, J. C.; Lin, S.; Waymouth, R. M. Macromolecules 2000, 33, 7249.
764. Dreier, T.; Erker, G.; Fro hlich, R.; Wibbeling, B. Organometallics 2000, 19, 4095.
765. Wilmes, G. M.; Lin, S.; Waymouth, R. M. Macromolecules 2002, 35, 5382.
766. Wilmes, G. M.; Polse, J. L.; Waymouth, R. M. Macromolecules 2002, 35, 6766.
767. Wiyatno, W.; Chen, Z.-R.; Liu, Y.; Waymouth, R. M.; Krukonis, V.; Brennan, K. Macromolecules 2004, 37, 701.
768. Wilmes, G. M.; France, M. B.; Lynch, S. R.; Waymouth, R. M. Organometallics 2004, 23, 2405.
769. Carlson, E. D.; Krejchi, M. T.; Shah, C. D.; Terakawa, T.; Waymouth, R. M.; Fuller, G. G. Macromolecules 1998, 31, 5343.
770. Hu, Y.; Krejchi, M. T.; Shah, C. D.; Myers, C. L.; Waymouth, R. M. Macromolecules 1998, 31, 6908.
771. Scho nherr, H.; Wiyatno, W.; Pople, J.; Frank, C. W.; Fuller, G. G.; Gast, A. P.; Waymouth, R. M. Macromolecules 2002, 35, 2654.
772. De Rosa, C.; Auriemma, F.; Circelli, T.; Waymouth, R. M. Macromolecules 2002, 35, 3622.
773. Busico, V.; Cipullo, R.; Kretschmer, W. P.; Talarico, G.; Vacatello, M.; Van Axel Castelli, V. Angew. Chem., Int. Ed. 2002, 41, 505.
773a. Busico, V.; Van Axel Castelli, V.; Aprea, P.; Cipullo, R.; Segre, A. L.; Talarico, G.; Vacatello, M. J. Am. Chem. Soc. 2003, 125, 5451.
774. Carlson, E. D.; Fuller, G. G.; Waymouth, R. M. Macromolecules 1999, 32, 8100.
775. Carlson, E. D.; Fuller, G. G.; Waymouth, R. M. Macromolecules 1999, 32, 8094.
776. Resconi, L.; Piemontesi, F.; Camurati, I.; Rychlicki, H.; Colonnesi, M.; Balboni, D. Polym. Mat. Sci. Eng. 1995, 73, 516.
777. Yano, A.; Kaneko, T.; Sato, M.; Akimoto, A. Macromol. Chem. Phys. 1999, 200, 2127.
778. Ewen, J. A.; Jones, R. L.; Elder, M. J.; Camurati, I.; Pritzkow, H. Macromol. Chem. Phys. 2004, 205, 302.
779. Nifant9ev, I. E.; Guidotti, S.; Resconi, L.; Laishevtsev, I. (Montell). World Pat. Appl. WO 01/47939, 2001.
780. Spaleck, W.; Antberg, M.; Aulbach, M.; Bachmann, B.; Dolle, V.; Haftka, S.; Ku ber, F.; Rohrmann, J.; Winter, A. New Isotactic
Polypropylenes via Metallocene Catalysts. In Ziegler Catalysts; Fink, G., Mu lhaupt, R., Brintzinger, H.-H., Eds.; Springer: Berlin, 1995; p 83.
781. Razavi, A.; Thewalt, U. J. Organomet. Chem. 2001, 621, 267.
1158 Olefin Polymerizations with Group IV Metal Catalysts
782. Razavi, A.; Bellia, V.; Brauwer, Y. D.; Hortmann, K.; Lambrecht, M.; Miserque, O.; Peters, L.; Belle, S. V. Syndiotactic and Isotactic
Specific Metallocene Catalysts with Hapto-flexible Cyclopentadienyl-fluorenyl Ligand. In Metalorganic Catalysts for Synthesis and
Polymerization; Kaminsky, W., Ed.; Springer: Berlin, 1999; p 236.
783. Kaminsky, W.; Werner, R. New C
1
Symmetric Metallocenes for the Polymerization of Olefins. In Metalorganic Catalysts for Synthesis and
Polymerization; Kaminsky, W., Ed.; Springer: Berlin, 1999; p 170.
784. Halterman, R. L.; Fahey, D. R.; Marin, V. P.; Dockter, D. W.; Khan, M. A. J. Organomet. Chem. 2001, 625, 154.
785. Voegele, J.; Dietrich, U.; Hackmann, M.; Rieger, B. Design of Ethylene-bridged ansa-Zirconocene Dichlorides for a Controlled Propene
Polymerization Reaction. In Metallocene-based Polyolefins: Preparation, Properties and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley:
Chichester, 2000; Vol. 1, p 485.
786. Mu ller, G.; Rieger, B. Prog. Polym. Sci. 2002, 27, 815.
787. Jones, R. L.; Elder, M. J.; Ewen, J. A. In Ninth International Business Forum on Specialty Polyolefins; Schotland,, Ed.; Schotland: Houston,
Texas, 1999; p 141.
788. Resconi, L.; Guidotti, S.; Camurati, I.; Nifantev, I. E.; Laishevtsev, I. Polym. Mat. Sci. Eng. 2002, 87, 76.
789. Rieger, B.; Troll, C.; Preuschen, J. Macromolecules 2002, 35, 5742.
790. Kukral, J.; Lehmus, P.; Feifel, T.; Troll, C.; Rieger, B. Organometallics 2000, 19, 3767.
791. Kukral, J.; Lehmus, P.; Klinga, M.; Leskela , M.; Rieger, B. Eur. J. Inorg. Chem. 2002, 1349.
792. Deisenhofer, S.; Feifel, T.; Kukral, J.; Klinga, M.; Leskela , M.; Rieger, B. Organometallics 2003, 22, 3495.
793. Cobzaru, C.; Deisenhofer, S.; Harley, A.; Troll, C.; Hild, S.; Rieger, B. Macromol. Chem. Phys. 2005, 206, 1231.
794. Veghini, D.; Henling, L. M.; Burkhardt, T. J.; Bercaw, J. E. J. Am. Chem. Soc. 1999, 121, 564.
795. Busico, V.; Cipullo, R.; Talarico, G.; Segre, A. L.; Caporaso, L. Macromolecules 1998, 31, 8720.
796. Winter, A.; Rohrmann, J.; Antberg, M.; Dolle, V.; Spaleck, W. (Hoechst). Eur. Pat. Appl. EP 387690, 1990.
797. Alt, H. G.; Zenk, R. J. Organomet. Chem. 1996, 522, 39.
798. Alt, H. G.; Zenk, R. J. Organometal. Chem. 1996, 522, 177.
799. Alt, H. G.; Zenk, R. J. Organomet. Chem. 1996, 518, 7.
800. Alt, H. G.; Zenk, R.; Milius, W. J. Organomet. Chem. 1996, 514, 257.
801. Patsidis, K.; Alt, H. G.; Milius, W.; Palackal, S. J. Organomet. Chem. 1996, 509, 63.
802. Miyake, S.; Bercaw, J. E. J. Mol. Catal. A: Chem. 1998, 128, 29.
803. Gentil, S.; Dietz, M.; Pirio, N.; Meunier, P.; Gallucci, J. C.; Gallou, F.; Paquette, L. A. Organometallics 2002, 21, 5162.
804. Razavi, A.; Bellia, V.; De Brauwer, Y.; Hortmann, K.; Peters, L.; Sirole, S.; Van Belle, S.; Marin, V.; Lopez, M. J. Organomet. Chem. 2003,
684, 206.
805. Mu ller, F.; Hopf, A.; Kaminsky, W.; Lemstra, P. J.; Loos, J. Polymer 2004, 45, 1815.
806. Miller, S. A.; Bercaw, J. E. Organometallics 2004, 23, 1777.
807. Ashe, A. J. III; Fang, X.; Hokky, A.; Kampf, J. W. Organometallics 2004, 23, 2197.
808. Grisi, F.; Longo, P.; Zambelli, A.; Ewen, J. A. J. Mol. Catal. A: Chem. 1999, 140, 225.
809. Kaminsky, W.; Hopf, A.; Piel, C. J. Organomet. Chem. 2003, 684, 200.
810. Song, W.; Rausch, M. D.; Chien, J. C. W. J. Polym. Sci. A: Polym. Chem. 1996, 34, 2945.
811. Shiomura, T.; Uchikawa, N.; Asanuma, T.; Sugimoto, R.; Fujio, I.; Kimura, S.; Harima, S.; Akiyama, M.; Kohno, M.; Inoue, N.
Metallocene-catalyzed Syndiotactic Polypropylene: Preparation and Properties. In Metallocene-based Polyolefins: Preparation, Properties and
Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 1, p 467.
812. Shamshoum, E. S.; Sun, L.; Reddy, B. R.; Turner, D. In Metcon 94: Houston, TX, USA, 1994.
813. Shamshoum, E. S.; Sun, L.; Kim, S. In Proceedings of the SPO 95; Schotland: Houston, 1995; p 299.
814. De Rosa, C.; Corradini, P. Macromolecules 1993, 26, 5711.
815. Auriemma, F.; De Rosa, C.; Corradini, P. Macromolecules 1993, 26, 5719.
816. Lovinger, A. J.; Lotz, B.; Davis, D. D.; Schumacher, M. Macromolecules 1994, 27, 6603.
817. Rodriguez-Arnold, J.; Bu, Z.; Cheng, S. Z. D. J. M. S. - Rev. Macromol. Chem. Phys. 1995, C35, 117.
818. De Rosa, C.; Auriemma, F.; Corradini, P. Macromolecules 1996, 29, 7452.
819. De Rosa, C.; Auriemma, F.; Vinti, V. Macromolecules 1998, 31, 7430.
820. De Rosa, C.; Auriemma, F.; Vinti, V.; Galimberti, M. Macromolecules 1998, 31, 6206.
821. Auriemma, F.; Ruiz de Ballesteros, O.; De Rosa, C. Macromolecules 2001, 34, 4485.
822. Nishii, K.; Matsumae, T.; Dare, E. O.; Shiono, T.; Ikeda, T. Macromol. Chem. Phys. 2004, 205, 363.
823. Busico, V.; Cipullo, R.; Cutillo, F.; Talarico, G.; Razavi, A. Macromol. Chem. Phys. 2003, 204, 1269.
824. Grandini, C.; Camurati, I.; Guidotti, S.; Mascellani, N.; Resconi, L.; Nifant9ev, I. E.; Kashulin, I. A.; Ivchenko, P. V.; Mercandelli, P.;
Sironi, A. Organometallics 2004, 23, 344.
825. De Rosa, C.; Auriemma, F.; Ruiz de Ballesteros, O.; Resconi, L.; Fait, A.; Ciaccia, E.; Camurati, I. J. Am. Chem. Soc. 2003, 125, 10913.
826. De Rosa, C.; Auriemma, F.; Ruiz de Ballesteros, O. Macromolecules 2003, 36, 7607.
827. De Rosa, C.; Auriemma, F.; Ruiz de Ballesteros, O. Macromolecules 2004, 37, 1422.
828. Okumura, Y.; Oberhoff, M.; Schottek, J. (Basell). World Pat. Appl. WO 03/045551, 2003.
829. Okumura, Y.; Seidel, N.; Koelling, L. (Basell). World Pat. Appl. WO 04/106351, 2004.
830. Chatterjee, A. M.; Campbell, R. N. In 52nd ANTEC; SPE,, Ed.; SPE: Houston, 1994, p 1977.
831. Chatterjee, A. M.; Campbell, R. N. J. Plastic Film & Sheeting 1994, 10, p 344.
832. Hosoda, S.; Hori, H.; Yada, K.; Nakahara, S.; Tsuji, M. Polymer 2002, 43, 7451.
833. Mitsui Chemicals America, INC. www.Mitsuichemicals.com.
834. Arnold, M.; Henschke, O.; Knorr, J. Macromol. Chem. Phys. 1996, 197, 563.
835. De Rosa, C.; Auriemma, F.; Orlando, I.; Talarico, G.; Caporaso, L. Macromolecules 2001, 34, 1663.
836. Naga, N.; Mizunuma, K.; Sadatoshi, H.; Kakugo, M. Macromolecules 1997, 30, 2197.
837. Kersting, M.; Langhauser, F.; Kerth, J.; Schweier, G. (BASF). World Pat. Appl. WO 94/28039, 1994.
838. Fischer, D.; Langhauser, F.; Lilge, D.; Hingmann, R.; Schweier, G. (BASF). World Pat. Appl. WO 97/10286, 1997.
839. Ueda, T.; Mizuno, A.; Kawasaki, M.; Fukuoka, D.; Kiso, Y.; Tanizaki, T.; Hashimoto, M.; Sugi, M.; Tsutsui, T. (Mitsui Petrochemical).
Eur. Pat. Appl. EP 682042, 1999.
Olefin Polymerizations with Group IV Metal Catalysts 1159
840. Karandinos, A. G.; Lohse, D. J.; Georjon, O. J. F.; Lewtas, K.; Tancrede, J. M.; Harrington, B. A.; Nelson, K. A. (Exxon). World Pat. Appl.
WO 01/46278, 2001.
841. Tsutsui, T.; Yoshitsugu, K.; Toyota, A. (Mitsui Petrochemical). Eur. Pat. Appl. EP 495 099, 1991.
842. Wahner, U. M.; Tincul, I.; Joubert, D. J.; Sadiku, E. R.; Forlini, F.; Losio, S.; Tritto, I.; Sacchi, M. C. Macromol. Chem. Phys. 2003, 204, 1738.
843. Sacchi, M. C.; Forlini, F.; Losio, S.; Tritto, I.; Wahner, U. M.; Tincul, I.; Joubert, D. J.; Sadiku, E. R. Macromol. Chem. Phys. 2003, 204, 1643.
844. Costa, G.; Stagnaro, P.; Trefiletti, V.; Sacchi, M. C.; Forlini, F.; Alfonso, G. C.; Tincul, I.; Wahner, U. M. Macromol. Chem. Phys. 2004, 205, 383.
845. Sacchi, M. C.; Forlini, F.; Tritto, I.; Stagnaro, P. Macromol. Chem. Phys. 2004, 205, 1804.
846. Arnold, M.; Bornemann, S.; Ko ller, F.; Menke, T. J.; Kressler, J. Macromol. Chem. Phys. 1998, 199, 2647.
847. van Reenen, A. J. Macromol. Symp. 2003, 193, 57.
848. Forlini, F.; Fan, Z. Q.; Tritto, I.; Locatelli, P.; Sacchi, M. C. Macromol. Chem. Phys. 1997, 198, 2397.
849. Forlini, F.; Tritto, I.; Locatelli, P.; Sacchi, M. C.; Piemontesi, F. Macromol. Chem. Phys. 2000, 201, 401.
850. Forlini, F.; Princi, E.; Tritto, I.; Sacchi, M. C.; Piemontesi, F. Macromol. Chem. Phys. 2002, 203, 645.
851. Song, F.; Pappalardo, D.; Johnson, A. F.; Rieger, B.; Bochmann, M. J. Polym. Sci, Part A: Polym. Chem. 2002, 40, 1484.
852. Poon, B.; Rogunova, M.; Hiltner, A.; Baer, E.; Chum, S. P.; Galeski, A.; Piorkowska, E. Macromolecules 2005, 38, 1232.
853. Schneider, M. J.; Mu lhaupt, R. Macromol. Chem. Phys. 1997, 198, 1121.
854. Juengling, S.; Koltzenburg, S.; Mu lhaupt, R. J. Polym. Sci. A: Polym. Chem. 1997, 35, 1.
855. Fan, Z.-Q.; Yasin, T.; Feng, L.-X. J. Polym. Sci. A: Polym. Chem. 2000, 38, 4299.
856. van Reenen, A. J.; Brull, R.; Wahner, U. M.; Raubenheimer, H. G.; Sanderson, R. D.; Pasch, H. J. Polym. Sci. A: Polym. Chem. 2000, 38, 4110.
857. Poon, B.; Rogunova, M.; Chum, S. P.; Hiltner, A.; Baer, E. J. Polym. Sci. B: Polym. Phys. 2004, 42, 4357.
858. Palza, H.; Lopez-Majada, J. M.; Quijada, R.; Benavente, R.; Pe rez, E.; Cerrada, M. L. Macromol. Chem. Phys. 2005, 206, 1221.
859. Henschke, O.; Koller, F.; Arnold, M. Macromol. Rapid Commun. 1997, 18, 617.
860. Naga, N.; Imanishi, Y. J. Polym. Sci. A: Polym. Chem. 2003, 41, 441.
861. Boggioni, L.; Bertini, F.; Zannoni, G.; Tritto, I.; Carbone, P.; Ragazzi, M.; Ferro, D. R. Macromolecules 2003, 36, 882.
862. Hasan, T.; Ikeda, T.; Shiono, T. Macromolecules 2005, 38, 1071.
863. Kono, H.; Ichiki, T.; Mori, H.; Nakatani, H.; Terano, M. Polym. Int. 2001, 50, 568.
864. Hackmann, M.; Repo, T.; Jany, G.; Rieger, B. Macromol. Chem. Phys. 1998, 199, 1511.
865. Hackmann, M.; Rieger, B. Macromolecules 2000, 33, 1524.
866. Ju ngling, S.; Mu lhaupt, R.; Fischer, D.; Langhauser, F. Angew. Makromol. Chem. 1995, 229, 93.
867. Caporaso, L.; Izzo, L.; Oliva, L. Macromolecules 1999, 32, 7329.
868. Caporaso, L.; Izzo, L.; Zappile, S.; Oliva, L. Macromolecules 2000, 33, 7275.
869. Hakala, K.; Lo fgren, B.; Helaja, T. Eur. Polym. J. 1998, 34, 1093.
870. Hagihara, H.; Tsuchihara, K.; Takeuchi, K.; Murata, M.; Ozaki, H.; Shiono, T. J. Polym. Sci. A: Polym. Chem. 2003, 42, 52.
871. Shiono, T.; Azad, S. M.; Ikeda, T. Macromolecules 1999, 32, 5723.
872. Weng, W.; Dekmezian, A. H.; Markel, E. J.; Peters, D. L. (Exxon). U.S. Patent 6184327, 2001.
873. Walter, P.; Trinkle, S.; Lilge, D.; Friedrich, C.; Mu lhaupt, R. Macromol. Mater. Eng. 2001, 286, 309.
874. Agarwal, P. K.; Somani, R. H.; Weng, W.; Mehta, A.; Yang, L.; Ran, S.; Liu, L.; Hsiao, B. S. Macromolecules 2003, 36, 5226.
875. Langston, J.; Dong, J. Y.; Chung, T. C. Macromolecules 2005, 38, 5849.
876. Ye, Z.; Zhu, S. J. Polym. Sci. A: Polym. Chem. 2003, 41, 1152.
877. Henschke, O.; Neubauer, A.; Arnold, M. Macromolecules 1997, 30, 8097.
878. Basell www.basell.com, 2004.
879. Rubin, I. D. Poly(1-Butene) Its Preparation and Properties; Gordon and Breach: New York, 1968.
880. Foglia, A. J. Appl. Polym. Symp. 1969, 11, 1.
881. Chatterjee, A. M. Butene polymers. In Encyclopedia of Polymer Science and Engineering; Kroschwitz, J. I., Klingsberg, A., Muldoon, J.,
Salvatore, A., Eds.; Wiley: New York, 1985; Vol. 2, p 590.
882. Luciani, L.; Seppa la , J.; Lo fgren, B. Prog. Polym. Sci. 1988, 13, 37.
883. Abedi, S.; Sharifi-Sanjani, N. J. Appl. Polym. Sci. 2000, 78, 2533.
884. Vitale, G.; Morini, G.; Cecchin, G. (Basell). World Pat. Appl. WO 03/099883, 2003.
885. Morini, G.; Piemontesi, F.; Vitale, G.; Bigiavi, D.; Pelliconi, A.; Garagnani, E.; Baita, P. (Basell). World Pat. Appl. WO 04/048424, 2004.
886. Icenogle, R. D.; Klingensmith, G. B. Macromolecules 1987, 20, 2788.
887. Busico, V.; Corradini, P.; De Biasio, R. Makromol. Chem. 1992, 193, 897.
888. Kaminsky, W.; Niedoba, S.; Mo ller-Lindenhof, N.; Rabe, O. In Catalysis in Polymer Synthesis, ACS Symposium Series; Vandenberg, E. J.,
Salamone, J. C., Eds.; American Chemical Society: Washington, DC, 1992; Vol. 496, p 63.
889. Kioka, M.; Tsutsui, T.; Ueda, T.; Kashiwa, N. Stereospecific Polymerization of c-Olefin with an Ethylene Bis(1-indenyl)hafnium
Dichloride and Methyl-aluminoxane Catalyst System. In Catalytic Olefin Polymerization, Studies in Surface Science and Catalysis; Keii, T.,
Soga, K., Eds.; Elsevier: New York, 1990, p 483.
890. Busico, V.; Cipullo, R.; Borriello, A. Macromol. Rapid Commun. 1995, 16, 269.
891. Herfert, N.; Fink, G. Makromol. Chem., Macromol. Symp. 1993, 66, 157.
892. Albizzati, E.; Resconi, L.; Zambelli, A. (Himont). Eur. Pat. Appl. EP 387609, 1990.
893. Asanuma, T.; Nishimori, Y.; Ito, M.; Uchikawa, N.; Shiomura, T. Polym. Bull. 1991, 25, 567.
894. Resconi, L.; Jones, R. L. (Montell). Eur. Pat. Appl. EP 604908, 1994.
895. Naga, N.; Mizunuma, K. Macromol. Rapid Commun. 1997, 18, 581.
896. Naga, N.; Shiono, T.; Ikeda, T. Macromol. Chem. Phys. 1999, 200, 1587.
897. Vathauer, M.; Kaminsky, W. Macromolecules 2000, 33, 1955.
898. Minami, Y.; Kanamaru, M.; Kakigami, K.; Funabashi, H. t. (Idemitsu Petrochemical). Eur. Pat. Appl. EP 1260525, 2002.
899. Huang, Q.; Wu, Q.; Zhu, F.; Lin, S. J. Polym. Sci. A: Polym. Chem. 2001, 39, 4068.
900. Huang, Q.; Zhu, F.; Wu, Q.; Lin, S. Polym. Int. 2001, 50, 45.
901. Fujita, H.; Seki, Y.; Miyatake, T. Macromol. Chem. Phys. 2004, 205, 884.
902. Collette, J. W.; Tullock, C. W. (Du Pont). U.S. Patent 4298722, 1981.
903. Resconi, L.; Camurati, I.; Malizia, F. Macromole Chem. Phys. accepted for publication.
904. Resconi, L. (Basell). Macromol. Chem. Phys. accepted for publication.
1160 Olefin Polymerizations with Group IV Metal Catalysts
905. Resconi, L. (Basell). World Pat. Appl. WO 02/100909, 2002.
906. Resconi, L. (Basell). World Pat. Appl. WO 03/042258, 2003.
907. Resconi, L.; Cascio Ingurgio, A. (Basell). World Pat. Appl. WO 04/050724, 2004.
908. Resconi, L.; Pelliconi, A.; Garagnani, E. (Basell). World Pat. Appl. WO 04/050713, 2004.
909. Resconi, L. (Basell). World Pat. Appl. WO 04/099269, 2004.
910. Kohyama, M.; Muranaka, T.; Fukui, K.; Kashiwa, N. (Mitsui Petrochemical). U.S. Patent 4801672, 1989.
911. Asakura, T.; Demura, M.; Nishiyama, Y. Macromolecules 1991, 24, 2334.
912. Galimberti, M.; Balbontin, G.; Camurati, I.; Paganetto, G. Macromol. Rapid Commun. 1994, 15, 633.
913. Kim, I.; Zhou, J.-M.; Chung, H. J. Polym. Sci. A: Polym. Chem. 2000, 38, 1687.
914. Grumel, V.; Bruell, R.; Pasch, H.; Raubenheimer, H. G.; Sanderson, R.; Wahner, U. Macromol. Mater. Eng. 2001, 286, 480.
915. Bruell, R.; Kgosane, D.; Neveling, A.; Pasch, H.; Raubenheimer, H. G.; Sanderson, R.; Wahner, U. Macromol. Symp. 2001, 165, 11.
916. Mosia, M. R., Ph.D. Thesis, Eindhoven University of Technology, Eindhoven, The Netherlands, 2004.
917. Chien, J. C. W.; Gong, B. M. J. Polym. Sci. A: Polym. Chem. 1993, 31, 1747.
918. Babu, G. N.; Newmark, R. A.; Chien, J. C. W. Macromolecules 1994, 27, 3383.
919. Frauenrath, H.; Keul, H.; Ho cker, H. Macromol. Rapid Commun. 1998, 19, 391.
920. Frauenrath, H.; Keul, H.; Ho cker, H. Macromol. Chem. Phys. 2001, 202, 3543.
921. Frauenrath, H.; Keul, H.; Ho cker, H. Macromol. Chem. Phys. 2001, 202, 3551.
922. Liu, Z.; Somsook, E.; Landis, C. R. J. Am. Chem. Soc. 2001, 123, 2915.
923. Coevoet, D.; Cramail, H.; Deffieux, A. Macromol. Chem. Phys. 1996, 197, 855.
924. Coevoet, D.; Cramail, H.; Deffieux, A. Macromol. Chem. Phys. 1998, 199, 1451.
925. Coevoet, D.; Cramail, H.; Deffieux, A. Macromol. Chem. Phys. 1998, 199, 1459.
926. Coevoet, D.; Cramail, H.; Deffieux, A. Macromol. Chem. Phys. 1999, 200, 1208.
927. Pe deutour, J.-N.; Cramail, H.; Deffieux, A. J. Mol. Catal. A: Chem 2001, 176, 87.
928. Sita, L. R.; Jayaratne, K. C. J. Am. Chem. Soc. 2000, 122, 958.
929. Keaton, R. J.; Jayaratne, K. C.; Henningsen, D. A.; Koterwas, L. A.; Sita, L. R. J. Am. Chem. Soc. 2001, 123, 6197.
930. Nomura, K.; Fudo, A. J. Mol. Catal. A: Chem. 2004, 209, 9.
931. Zhang, Y.; Sita, L. R. J. Am. Chem. Soc. 2004, 126, 7776.
932. Siedle, A. R.; Newmark, R. A.; Duerr, B. F.; Leung, P. C. J. Mol. Catal. A: Chem. 2004, 214, 187.
933. Yamaguchi, Y.; Suzuki, N.; Fries, A.; Mise, T.; Koshino, H.; Ikegami, Y.; Ohmori, H.; Matsumoto, A. J. Polym. Sci. A: Polym. Chem.
1999, 37, 283.
934. Suzuki, N.; Masubuchi, Y.; Yamaguchi, Y.; Kase, T.; Miyamoto, T. K.; Horiuchi, A.; Mise, T. Macromolecules 2000, 33, 754.
935. Suzuki, N.; Yamaguchi, Y.; Fries, A.; Mise, T. Macromolecules 2000, 33, 4602.
936. Hoff, M.; Kaminsky, W. Macromol. Chem. Phys. 2004, 205, 1167.
937. Krentsel, B. A.; Kissin, Y. V.; Kleiner, V. J.; Stotskaya, L. L. Polymers and Copolymers of Higher a-Olefins; Hanser: Munich, 1997.
938. Mitsui Chemicals www.mitsui-chem.co.jp, 2005.
939. Zambelli, A.; Ammendola, P.; Proto, A. Macromolecules 1989, 22, 2126.
940. De Rosa, C.; Venditto, V.; Guerra, G.; Corradini, P. Macromolecules 1992, 25, 6938.
941. De Rosa, C.; Grassi, A.; Capitani, D. Macromolecules 1998, 31, 3163.
942. Irwin, L. J.; Miller, S. A. J. Am. Chem. Soc. 2005, 127, 9972.
943. Stehling, U.; Diebold, J.; Kirsten, R.; Ro ll, W.; Brintzinger, H.-H.; Ju ngling, S.; Mu lhaupt, R.; Langhauser, F. Organometallics 1994, 13, 964.
944. Borriello, A.; Busico, V.; Cipullo, R.; Chadwick, J. C.; Sudmeijer, O. Macromol. Rapid Commun. 1996, 17, 589.
945. Oliva, L.; Longo, P.; Zambelli, A. Macromolecules 1996, 29, 6383.
946. Sacchi, M. C.; Barsties, E.; Tritto, I.; Locatelli, P.; Brintzinger, H.-H.; Stehling, U. Macromolecules 1997, 30, 1267.
947. Chien, J. C. W.; Vizzini, J. C.; Kaminsky, W. Macromol. Chem. Rapid Commun. 1992, 13, 479.
948. Byun, D.-J.; Shin, D.-K.; Liu, J.; Kim, S. Y. Polym. Bull. 1999, 42, 265.
949. Byun, D.-J.; Shin, D.-K.; Kim, S. Y. Polym. Bull. 1999, 42, 301.
950. Longo, P.; Grassi, A.; Grisi, F.; Milione, S. Macromol. Rapid Commun. 1998, 19, 229.
951. Ishihara, N.; Seimiya, T.; Kuramoto, M.; Uoi, U. Macromolecules 1986, 19, 2464.
952. Ishihara, N.; Kuramoto, M.; Uoi, U. Eur. Pat. EP 210615, 1987.
953. Ishihara, N.; Kuramoto, M.; Uoi, U. U.S. Patent 4680353, 1987.
954. Grassi, A.; Pellecchia, C.; Longo, P.; Zambelli, A. Gazz. Chim. Ital. 1987, 117, 249.
955. Ammendola, P.; Pellecchia, C.; Longo, P.; Zambelli, A. Gazz. Chim. Ital. 1987, 117, 65.
956. Ishihara, N.; Kuramoto, M.; Uoi, M. Macromolecules 1988, 21, 3356.
957. Ishihara, N.; Kuramoto, M. Stud. Surf. Sci. Catal. 1994, 89, 339.
958. Ishihara, N. Macromol. Symp. 1995, 89, 553.
959. Tomotsu, N.; Ishihara, N. Stud. Surf. Sci. Catal. 1999, 121, 269.
960. Pellecchia, C.; Longo, P.; Grassi, A.; Ammendola, P.; Zambelli, A. Makromol. Chem., Rapid Commun. 1987, 8, 277.
961. Zambelli, A.; Longo, P.; Pellecchia, C.; Grassi, A. Macromolecules 1987, 20, 2035.
962. Ammendola, P.; Shijing, X.; Grassi, A.; Zambelli, A. Gazz. Chim. Ital. 1988, 118, 769.
963. Oliva, L.; Pellecchia, C.; Cinquina, P.; Zambelli, A. Macromolecules 1989, 22, 1642.
964. Zambelli, A.; Pellecchia, C.; Oliva, L.; Longo, P.; Grassi, A. Makromol. Chem. 1991, 192, 223.
965. Zambelli, A.; Pellecchia, C.; Proto, A. Macromol. Symp. 1995, 89, 373.
966. Po, R.; Cardi, N. Prog. Polym. Sci. 1996, 21, 47.
967. Tomotsu, N.; Ishihara, N.; Newman, T. H.; Malanga, M. T. J. Mol. Catal. A: Chem. 1998, 128, 167.
968. Pellecchia, C.; Grassi, A. Top. Catal. 1999, 7, 125.
969. Ready, T. E.; Day, R. O.; Chien, J. C. W.; Rausch, M. D. Macromolecules 1993, 26, 5822.
970. Ready, T. E.; Chien, J. C. W.; Rausch, M. D. J. Organomet. Chem. 1999, 583, 11.
971. Schneider, N.; Prosenc, M.-H.; Brintzinger, H.-H. J. Organomet. Chem. 1997, 545546, 291.
972. Kaminsky, W.; Lenk, S.; Scholz, V.; Roesky, H. W.; Herzog, A. Macromolecules 1997, 30, 7647.
973. Chien, J. C. W.; Salajka, Z. J. Pol. Sci., Part A: Pol. Chem. 1991, 29, 1253.
Olefin Polymerizations with Group IV Metal Catalysts 1161
974. Sun, X.; Xie, J.; Zhang, H.; Huang, J. Eur. Polym. J. 2004, 40, 1903.
975. Schellenberg, J. Eur. Polym. J. 2004, 40, 2259.
976. Qian, Y.; Zhang, H.; Zhou, J.; Zhao, W.; Sun, X.; Huang, J. J. Mol. Catal. A: Chem. 2004, 208, 45.
977. Nomura, K.; Fujita, K.; Fujiki, M. Catal. Commun. 2004, 5, 413.
978. Qian, X.; Huang, J.; Qian, Y.; Wang, C. Appl. Organomet. Chem. 2003, 17, 277.
979. Nomura, K.; Fudo, A. Catal. Commun. 2003, 4, 269.
980. Jamanek, D.; Woyda, A.; Skupinski, W. Appl. Organomet. Chem. 2002, 16, 575.
981. Lee, B. Y.; Han, J. W.; Seo, H.; Lee, I. S.; Chung, Y. K. J. Organomet. Chem. 2001, 627, 233.
982. Soga, K.; Kawabe, M.; Murata, M.; Kase, T.; Ozaki, H.; Fukui, Y.; Hoang, T. B.; Jin, J.; Miyazawa, A.; Hagihara, H., et al. (Japan as
Represented by Secretary of Agency of Industrial Science and Technology, Japan; Japan Chemical Innovation Institute; Soga, Hisae).
World Pat. Appl. WO 0158964, 2001.
983. Xu, G.; Cheng, D. Macromolecules 2000, 33, 2825.
984. Schellenberg, J. J. Polym. Sci. A: Polym. Chem. 2000, 38, 2428.
985. Rhodes, B.; Rausch, M. D.; Chien, J. C. W. J. Polym. Sci. A: Polym. Chem. 2000, 39, 313.
986. Liu, J.; Ma, H.; Huang, J.; Qian, Y. Eur. Polym. J. 2000, 36, 2055.
987. Xu, G. Macromolecules 1998, 31, 586.
988. Liu, J.; Ma, H.; Huang, J.; Qian, Y.; Chan, A. S. C. Eur. Polym. J. 1998, 35, 543.
989. Kaminsky, W.; Lenk, S. Macromol. Symp. 1997, 118, 45.
990. Kawabe, M.; Murata, M.; Kase, T.; Ozaki, H.; Fukui, Y.; Juang, T.; Jin, J.; Miyazawa, T.; Hagiwara, H.; Tsuchihara, K., et al. (Agency of
Industrial Sciences and Technology, Japan; Zaidan Hojin Kagaku Gijitsu Senryakusuishin Kiko). Jpn. Kokai Tokkyo Koho JP 2000319323,
2000.
991. Ready, T. E.; Chien, J. C. W.; Rausch, M. D. J. Organomet. Chem. 1996, 519, 21.
992. Tian, G.; Xu, S.; Zhang, Y.; Wang, B.; Zhou, X. J. Organomet. Chem. 1998, 558, 231.
993. Qian, Y.; Zhang, H.; Qian, X.; Chen, B.; Huang, J. Eur. Polym. J. 2002, 38, 1613.
994. Huang, B.; Cao, K.; Li, B.-G.; Zhu, S. J. Appl. Polym. Sci. 2004, 94, 1449.
995. Schellenberg, J.; Knoll, S.; Nord, G.; Leukefeld, W. Eur. Polym. J. 2003, 39, 2351.
996. Choi, K. Y.; Chung, J. S.; Woo, B. G.; Hong, M. H. J. Appl. Polym. Sci. 2003, 88, 2132.
997. Pasquet, V.; Spitz, R. Macromol. Chem. Phys. 2001, 202, 2346.
998. Kawabe, M.; Murata, M. Macromol. Chem. Phys. 2001, 202, 2440.
999. Klosin, J.; Kruper, W. J., Jr.; Nickias, P. N.; Roof, G. R.; Soto, J. (Dow Chemical). World Pat. Appl. WO 01/042315, 2001.
1000. Pasquet, V.; Spitz, R. Macromol. Chem. Phys. 1999, 200, 1453.
1001. Kaminsky, W.; Arrowsmith, D.; Strubel, C. J. Polym. Sci. A: Polym. Chem. 1999, 37, 2959.
1002. Xu, J.; Zhao, J.; Fan, Z.; Feng, L. Eur. Polym. J. 1998, 35, 127.
1003. Hu, A. T.; Lee, J. H.-J.; Chen, T.-S. J. Appl. Polym. Sci. 1998, 70, 1747.
1004. Yu, G.; Chen, H.; Zhang, X.; Jiang, Z.; Huang, B. J. Polym. Sci. A: Polym. Chem. 1996, 34, 2237.
1005. Yim, J.-H.; Chu, K.-J.; Choi, K.-W.; Ihm, S.-K. Eur. Polym. J. 1996, 32, 1381.
1006. Liu, M. T.; Baker, W. E.; Schytt, V.; Jones, T.; Baird, M. C. J. Appl. Polym. Sci. 1996, 62, 1807.
1007. Flores, J. C.; Chien, J. C. W.; Rausch, M. D. Organometallics 1995, 14, 1827.
1008. Okuda, J.; Masoud, E. Macromol. Chem. Phys. 1998, 199, 543.
1009. Kim, I.; Ha, Y. S.; Zhang, D. F.; Ha, C.-S.; Lee, U. Macromol. Rapid Commun. 2004, 25, 1319.
1010. Lee, M. H.; Kim, S. K.; Do, Y. Organometallics 2005, 24, 3618.
1011. Longo, P.; Proto, A.; Zambelli, A. Macromol. Chem. Phys. 1995, 196, 3015.
1012. Zambelli, A.; Pellecchia, C.; Oliva, L.; Han, S. J. Polym. Sci. 1988, 26, 365.
1013. Longo, P.; Grassi, A.; Proto, A.; Ammendola, P. Macromolecules 1988, 21, 24.
1014. Pellecchia, C.; Proto, A.; Zambelli, A. Macromolecules 1992, 25, 4450.
1015. Pellecchia, C.; Grassi, A.; Immirzi, A. J. Am. Chem. Soc. 1993, 115, 1160.
1016. Wang, Q.; Quyoum, R.; Gillis, D. J.; Tudoret, M.-J.; Jeremic, D.; Hunter, B. K.; Baird, M. C. Organometallics 1996, 15, 693.
1017. Williams, E. F.; Murray, M. C.; Baird, M. C. Macromolecules 2000, 33, 261.
1018. Minieri, G.; Corradini, P.; Zambelli, A.; Guerra, G.; Cavallo, L. Macromolecules 2001, 34, 2459.
1019. Minieri, G.; Corradini, P.; Guerra, G.; Zambelli, A.; Cavallo, L. Macromolecules 2001, 34, 5379.
1020. Quyoum, R.; Wang, Q.; Tudoret, M.-J.; Baird, M. C. J. Am. Chem. Soc. 1994, 116, 6435.
1021. Chien, J. C. W.; Salajka, Z.; Dong, S. Macromolecules 1992, 25, 3199.
1022. Grassi, A.; Saccheo, S.; Zambelli, A.; Laschi, F. Macromolecules 1998, 31, 5588.
1023. Grassi, A.; Zambelli, A.; Laschi, F. Organometallics 1996, 15, 480.
1024. Ready, T. E.; Gurge, R.; Chien, J. C. W.; Rausch, M. D. Organometallics 1998, 17, 5236.
1025. Mahanthappa, M. K.; Waymouth, R. M. J. Am. Chem. Soc. 2001, 123, 12093.
1026. Pellecchia, C.; Pappalardo, D.; Oliva, L.; Zambelli, A. J. Am. Chem. Soc. 1995, 117, 6593.
1027. Grassi, A.; Longo, P.; Proto, A.; Zambelli, A. Macromolecules 1989, 22, 104.
1028. Soga, K.; Nakatani, H.; Monoi, T. Macromolecules 1990, 23, 953.
1029. Zambelli, A.; Pellecchia, C. Stud. Surf. Sci. Catal. 1994, 89, 209.
1030. Coates, G. W.; Waymouth, R. M. J. Am. Chem. Soc. 1993, 115, 91.
1031. Miller, S. A.; Waymouth, R. M. Stereo- and Enantioselective Polymerization of Olefins with Homogeneous ZieglerNatta Catalysts. In
Ziegler Catalysts; Fink, G., Mu lhaupt, R., Brintzinger, H.-H., Eds.; Springer: Berlin, 1995; p 441.
1032. Ystenes, M. J. Catal. 1991, 129, 383.
1033. Cheng, H. N.; Khasat, N. P. J. Appl. Polym. Sci. 1988, 35, 825.
1034. Nomura, K.; Hatanaka, Y.; Okumura, H.; Fujiki, M.; Hasegawa, K. Macromolecules 2004, 37, 1693.
1035. Jeremic, D.; Wang, Q.; Quyoum, R.; Baird, M. C. J. Organomet. Chem. 1995, 497, 143.
1036. Choo, T. N.; Waymouth, R. M. J. Am. Chem. Soc. 2002, 124, 4188.
1037. Jayaratne, K. C.; Keaton, R. J.; Henningsen, D. A.; Sita, L. R. J. Am. Chem. Soc. 2000, 122, 10490.
1038. Naga, N.; Shiono, T.; Ikeda, T. Macromol. Chem. Phys. 1999, 200, 1466.
1162 Olefin Polymerizations with Group IV Metal Catalysts
1039. Tritto, I.; Boggioni, L.; Sacchi, M. C.; Locatelli, P. J. Mol. Catal. A: Chem. 1998, 133, 139.
1040. Arndt, M.; Gosmann, M. Polym. Bull. 1998, 41, 433.
1041. Hasan, T.; Nishii, K.; Shiono, T.; Ikeda, T. Macromolecules 2002, 35, 8933.
1042. Hasan, T.; Ikeda, T.; Shiono, T. Macromolecules 2004, 37, 7432.
1043. Mu lhaupt, R. Macromol. Chem. Phys. 2003, 204, 289.
1043a. Janiak, C.; Lassahn, P. G. J. Mol. Catal. A: Chem. 2001, 166, 193.
1044. Yanjarappa, M. J.; Sivaram, S. Macromol. Chem. Phys. 2004, 205, 2055.
1045. Porri, L.; Giarrusso, A.; Ricci, G. Metallocene Catalysts for 1,3-Diene Polymerization. In Metallocene-based Polyolefins: Preparation, Properties
and Technology; Scheirs, J., Kaminsky, W., Eds.; Wiley: Chichester, 2000; Vol. 2, p 115.
1046. Kaminsky, W.; Scholz, V. New Half-sandwich Titanocenes for the Polymerization of Butadiene. In Organometallic Catalysts and Olefin
Polymerization. Catalysts for a New Millenium; Blom, R., Follestad, A., Rytter, E., Tilset, M., Ystenes, M., Eds.; Springer: Berlin, 2001; p 346.
1047. Zambelli, A.; Caprio, M.; Grassi, A.; Bowen, D. E. Macromol. Chem. Phys. 2000, 201, 393.
1048. Welborn, H. C. (Exxon). World Pat. Appl. WO 88/04672A1, 1988.
1049. Welborn, H. C.; Austin, R. G. (Exxon). World Pat. Appl. WO 8804674, 1987.
1050. Galimberti, M.; Albizzati, E.; Abis, L.; Bacchilega, G. Makromol. Chem. 1991, 192, 2591.
1051. Kaminsky, W.; Hinrichs, B. Macromol. Symp. 2003, 195, 39.
1052. Ishihara, T.; Shiono, T. Macromolecules 2003, 36, 9675.
1053. Pragliola, S.; Milano, G.; Guerra, G.; Longo, P. J. Am. Chem. Soc. 2002, 124, 3502.
1054. Pragliola, S.; Costabile, C.; Magrino, M.; Napoli, M.; Longo, P. Macromolecules 2004, 37, 238.
1055. Longo, P.; Napoli, M.; Pragliola, S.; Costabile, C.; Milano, G.; Guerra, G. Macromolecules 2003, 36, 9067.
1056. Longo, P.; Pragliola, S.; Milano, G.; Guerra, G. J. Am. Chem. Soc. 2003, 125, 4799.
1057. Pragliola, S.; Costabile, C.; Di Bartolomeo, F.; Longo, P. Macromol. Rapid Commun. 2004, 25, 995.
1058. Simanke, A. G.; Mauler, R. S.; Galland, G. B. J. Polym. Sci. A: Polym. Chem. 2002, 40, 471.
1059. Choo, T. N.; Waymouth, R. M. J. Am. Chem. Soc. 2003, 125, 8970.
1060. Park, S. J.; Han, Y.; Kim, S. K.; Lee, J. Y.; Kim, H. K.; Do, Y. J. Organomet. Chem. 2004, 689, 4263.
1061. Britovsek, G. J. P.; Gibson, V. C.; Wass, D. F. Angew. Chem., Int. Ed. 1999, 38, 428.
1062. Gibson, V. C.; Spitzmesser, S. K. Chem. Rev. 2003, 103, 283.
1063. Froese, R. D. J.; Musaev, D. G.; Matsubara, T.; Morokuma, K. J. Am. Chem. Soc. 1997, 119, 7190.
1064. Froese, R. D. J.; Musaev, D. G.; Morokuma, K. Organometallics 1999, 18, 373.
1065. Sudhakar, P.; Amburose, C. V.; Sundararajan, G.; Nethaji, M. Organometallics 2004, 23, 4462.
1066. Scollard, J. D.; McConville, D. H.; Payne, N. C.; Vittal, J. J. Macromolecules 1996, 29, 5241.
1067. Scollard, J. D.; McConville, D. H.; Vittal, J. J.; Payne, N. C. J. Mol. Catal. A: Chem. 1998, 128, 201.
1068. Scollard, J. D.; McConville, D. H.; Rettig, S. J. Organometallics 1997, 16, 1810.
1069. Scollard, J. D.; McConville, D. H.; Vittal, J. J. Organometallics 1997, 16, 4415.
1070. Uozumi, T.; Tsubaki, S.; Jin, J.; Sano, T.; Soga, K. Macromol. Chem. Phys. 2001, 202, 3279.
1071. Tsubaki, S.; Jin, J.; Ahn, C.-H.; Sano, T.; Uozumi, T.; Soga, K. Macromol. Chem. Phys. 2001, 202, 482.
1072. Jin, J.; Tsubaki, S.; Uozomi, T.; Sano, T.; Soga, K. Macromol. Rapid Commun. 1998, 19, 597.
1073. Hagimoto, H.; Shiono, T.; Ikeda, T. Macromol. Chem. Phys. 2004, 205, 19.
1074. Hagimoto, H.; Shiono, T.; Ikeda, T. Macromol. Rapid Commun. 2002, 23, 73.
1075. Lee, C. H.; La, Y.-H.; Park, J.-W. Organometallics 2000, 19, 344.
1076. Lorber, C.; Donnadieu, B.; Choukroun, R. Organometallics 2000, 19, 1963.
1077. Lee, C. H.; La, Y.-H.; Park, S. J.; Park, J. W. Organometallics 1998, 17, 3648.
1078. Nomura, K.; Naga, N.; Takaoki, K. Macromolecules 1998, 31, 8009.
1079. Jeon, Y.; Heo, J.; Lee, W.; Chang, T.; Kim, K. Organometallics 1999, 18, 4107.
1080. Cloke, F. G.; Geldbach, T. J.; Hitchcock, P. B.; Love, J. B. J. Organomet. Chem. 1996, 506, 343.
1081. Patton, J. T.; Feng, S. G.; Abboud, K. A. Organometallics 2001, 20, 3399.
1082. Patton, J. T.; Bokota, M. M.; Abboud, K. A. Organometallics 2002, 21, 2145.
1083. Jeon, Y.-M.; Park, S. J.; Heo, J.; Kim, K. Organometallics 1998, 17, 3161.
1084. Shafir, A.; Arnold, J. Organometallics 2003, 22, 567.
1085. Mack, H.; Eisen, M. S. J. Organomet. Chem. 1996, 525, 81.
1086. Tsurugi, H.; Yamagata, T.; Tani, K.; Mashima, K. Chem. Lett. 2003, 32, 756.
1087. Gue rin, F.; Stewart, J. C.; Beddie, C.; Stephan, D. W. Organometallics 2000, 19, 2994.
1088. Stephan, D. W.; Gue rin, F.; Spence, R. E. v. H.; Koch, L.; Gao, X.; Brown, S. J.; Swabey, J. W.; Wang, Q.; Xu, W., Zoricak, P., et al.
Organometallics 1999, 18, 2046.
1089. Brown, S. J.; Gao, X.; Harrison, D. G.; McKay, I.; Koch, L.; Wang, Q.; Xu, W.; Von Haken Spence, R. E.; Stephan, D. W. (Nova Chemicals
Corporation). World Pat. Appl. WO 00/005238, 2000.
1090. Kickham, J. E.; Gue rin, F.; Stephan, D. W. J. Am. Chem. Soc. 2002, 124, 11486.
1091. Hollink, E.; Stewart, J. C.; Wei, P.; Stephan, D. W. J. Chem. Soc., Dalton Trans. 2003, 3968.
1092. Saccheo, S.; Gioia, G.; Grassi, A.; Bowen, D. E.; Jordan, R. F. J. Mol. Catal. A: Chem. 1998, 128, 111.
1093. Young, D. A. J. Mol. Catal. 1989, 53, 433.
1094. Proto, A.; Capacchione, C.; Motta, O.; De Carlo, F. Macromolecules 2003, 36, 5942.
1095. Motta, O.; Capacchione, C.; Proto, A.; Acierno, D. Polymer 2002, 43, 5847.
1096. Miyatake, T.; Mizunuma, K.; Seki, Y.; Kakugo, M. Makromol. Chem., Rapid Commun. 1989, 10, 349.
1097. Mack, H.; Eisen, M. S. J. Chem. Soc., Dalton Trans. 1998.
1098. Manivannan, R.; Sundararajan, G. Macromolecules 2002, 35, 7883.
1099. Natrajan, L. S.; Wilson, C.; Okuda, J.; Arnold, P. L. Eur. J. Inorg. Chem. 2004, 3724.
1100. Chan, M. C. W.; Tam, K.-H.; Pui, Y.-L.; Zhu, N. J. Chem. Soc., Dalton Trans. 2002, 3085.
1101. Oakes, D. C. H.; Kimberley, B. S.; Gibson, V. C.; Jones, D. J.; White, A. J. P.; Williams, D. J. Chem. Commun. 2004, 2174.
1102. Rhodes, B.; Chien, J. C. W.; Wood, J. S.; Chandrasekaran, A.; Rausch, M. D. J. Organomet. Chem. 2001, 625, 95.
1103. Schrock, R. R.; Adamchuk, J.; Ruhland, K.; Lopez, L. P. H. Organometallics 2005, 24, 857.
Olefin Polymerizations with Group IV Metal Catalysts 1163
1104. Tonzetich, Z. J.; Lu, C. C.; Schrock, R. R.; Hock, A. S.; Bonitatebus, P. J., Jr. Organometallics 2004, 23, 4362.
1105. Mehrkhodavandi, P.; Schrock, R. R.; Pryor, L. L. Organometallics 2003, 22, 4569.
1106. Schrock, R. R.; Adamchuk, J.; Ruhland, K.; Lopez, L. P. H. Organometallics 2003, 22, 5079.
1107. Schrock, R. R.; Bonitatebus, P. J., Jr.; Schrodi, Y. Organometallics 2001, 20, 1056.
1108. Mehrkhodavandi, P.; Schrock, R. R. J. Am. Chem. Soc. 2001, 123, 10746.
1109. Mehrkhodavandi, P.; Bonitatebus, P. J., Jr.; Schrock, R. R. J. Am. Chem. Soc. 2000, 122, 7841.
1110. Liang, L.-C.; Schrock, R. R.; Davis, W. M.; McConville, D. H. J. Am. Chem. Soc. 1999, 121, 5797.
1111. Flores, M. A.; Manzoni, M. R.; Baumann, R.; Davis, W. M.; Schrock, R. R. Organometallics 1999, 18, 3220.
1112. Schrock, R. R.; Baumann, R.; Reid, S. M.; Goodman, J. T.; Stumpf, R.; Davis, W. M. Organometallics 1999, 18, 3649.
1113. Aizenberg, M.; Turculet, L.; Davis, W. M.; Schattenmann, F.; Schrock, R. R. Organometallics 1998, 17, 4795.
1114. Baumann, R.; Davis, W. M.; Schrock, R. R. J. Am. Chem. Soc. 1997, 119, 3830.
1115. Baumann, R.; Schrock, R. R. J. Organomet. Chem. 1998, 557, 69.
1116. Horton, A. D.; de With, J.; van der Linden, A. J.; van de Weg, H. Organometallics 1996, 15, 2672.
1117. Vollmerhaus, R.; Rahim, M.; Tomaszewski, R.; Xin, S.; Taylor, N. J.; Collins, S. Organometallics 2000, 19, 2161.
1118. Jin, X.; Novak, B. M. Macromolecules 2000, 33, 6205.
1119. Stevens, J. C.; Vanderlende, D. D. (Dow Chemical). World Pat. Appl. WO 03/040201, 2003.
1120. Bouwkamp, M.; Van Leusen, D.; Meetsma, A.; Hessen, B. Organometallics 1998, 17, 3645.
1121. Suzuki, Y.; Inoue, Y.; Tanaka, H.; Fujita, T. Macromol. Rapid Commun. 2004, 25, 493.
1122. Matilainen, L.; Klinga, M.; Leskela, M. J. Chem. Soc., Dalton Trans. 1996, 219.
1123. Sobota, P.; Przybylak, K.; Utko, J.; Jerzykiewicz, L. B.; Pombeiro, A. J. L.; Silva, M. F. C. G. d.; Szczegot, K. Chem. Eur. J. 2001, 7, 951.
1124. Shmulinson, M.; Galan-Fereres, M.; Lisovskii, A.; Nelkenbaum, E.; Semiat, R.; Eisen, M. S. Organometallics 2000, 19, 1208.
1125. Janas, Z.; Jerzykiewicz, L. B.; Przybylak, K.; Sobota, P.; Szczegot, K. Eur. J. Inorg. Chem. 2004, 1639.
1126. Fujita, T.; Tohi, Y.; Mitani, M.; Matsui, S.; Saito, J.; Nitabaru, M.; Sugi, K.; Makio, H.; Tsutsui, T. (Mitsui Chemicals). Eur Pat. Appl. EP
874005 A1, 1998.
1127. Matsui, S.; Tohi, Y.; Mitani, M.; Saito, J.; Makio, H.; Tanaka, H.; Nitabaru, M.; Nakano, T.; Fujita, T. Chem. Lett. 1999, 10, 1065.
1128. Matsui, S.; Mitani, M.; Saito, J.; Tohi, Y.; Makio, H.; Tanaka, H.; Fujita, T. Chem. Lett. 1999, 12, 1263.
1129. Matsui, S.; Mitani, M.; Saito, J.; Tohi, Y.; Makio, H.; Matsukawa, N.; Takagi, Y.; Tsuru, K.; Nitabaru, M., Nakano, T., et al. J. Am. Chem.
Soc. 2001, 123, 6847.
1130. Matsukawa, N.; Matsui, S.; Mitani, M.; Saito, J.; Tsuru, K.; Kashiwa, N.; Fujita, T. J. Mol. Catal. A: Chem. 2001, 169, 99104.
1131. Mitani, M.; Yoshida, Y.; Mohri, J.; Tsuru, K.; Ishii, S.; Kojoh, S.-I.; Matsugi, T.; Saito, J.; Matsukawa, N.; Matsui, S.; et al. (Mitsui
Chemicals). World Pat. Appl. WO 01/55231, 2001.
1132. Saito, J.; Mitani, M.; Mohri, J.-I.; Yoshida, Y.; Matsui, S.; Ishii, S.-I.; Kojoh, S.-I.; Kashiwa, N.; Fujita, T. Angew. Chem., Int. Ed. 2001, 40,
2918.
1133. Ishii, S.-i.; Saito, J.; Mitani, M.; Mohri, J.-i.; Matsukawa, N.; Tohi, Y.; Matsui, S.; Kashiwa, N.; Fujita, T. J. Mol. Catal. A: Chem. 2002, 179,
11.
1134. Mitani, M.; Mohri, J.-I.; Yoshida, Y.; Saito, J.; Ishii, S.; Tsuru, K.; Matsui, S.; Furuyama, R.; Nakano, T., Tanaka, H., et al. J. Am. Chem. Soc.
2002, 124, 3327.
1135. Saito, J.; Mitani, M.; Matsui, S.; Tohi, Y.; Makio, H.; Nakano, T.; Tanaka, H.; Kashiwa, N.; Fujita, T. Macromol. Chem. Phys. 2002, 203, 59.
1136. Suzuki, Y.; Kashiwa, N.; Fujita, T. Chem. Lett. 2002, 3, 358.
1137. Bando, H.; Nakayama, Y.; Sonobe, Y.; Fujita, T. Macromol. Rapid Commun. 2003, 24, 732.
1138. Furuyama, R.; Saito, J.; Ishii, S.-i.; Mitani, M.; Matsui, S.; Tohi, Y.; Makio, H.; Matsukawa, N.; Tanaka, H.; Fujita, T. J. Mol. Catal. A: Chem.
2003, 200, 31.
1139. Ishii, S.-i.; Furuyama, R.; Matsukawa, N.; Saito, J.; Mitani, M.; Tanaka, H.; Fujita, T. Macromol. Rapid Commun. 2003, 24, 452.
1140. Mitani, M.; Nakano, T.; Fujita, T. Chem. Eur. J. 2003, 9, 2396.
1141. Mitani, M.; Mohri, J.-i.; Furuyama, R.; Ishii, S.; Fujita, T. Chem. Lett. 2003, 32, 238.
1142. Tohi, Y.; Makio, H.; Matsui, S.; Onda, M.; Fujita, T. Macromolecules 2003, 36, 523.
1143. Nakayama, Y.; Bando, H.; Sonobe, Y.; Fujita, T. J. Mol. Catal. A: Chem. 2004, 213, 141.
1144. Tohi, Y.; Nakano, T.; Makio, H.; Matsui, S.; Fujita, T.; Yamaguchi, T. Macromol. Chem. Phys. 2004, 205, 1179.
1145. Matsui, S.; Fujita, T. Catal. Today 2001, 66, 6373.
1146. Makio, H.; Kashiwa, N.; Fujita, T. Adv. Synth. Catal. 2002, 344, 477.
1147. Matsukawa, N.; Ishii, S.-i.; Furuyama, R.; Saito, J.; Mitani, M.; Makio, H.; Tanaka, H.; Fujita, T. e-Polymers 2003, Paper No. 21.
1148. Suzuki, Y.; Terao, H.; Fujita, T. Bull. Chem. Soc. Jpn. 2003, 76, 1493.
1149. Mitani, M.; Saito, J.; Ishii, S.-I.; Nakayama, Y.; Makio, H.; Matsukawa, N.; Matsui, S.; Mohri, J.; Furuyama, R., Terao, H., et al. Chem. Rec.
2004, 4, 137.
1150. Deck, P. A.; Beswick, C. L.; Marks, T. J. J. Am. Chem. Soc. 1998, 120, 1772.
1151. Strauch, J.; Warren, T. H.; Erker, G.; Fro hlich, R.; Saarenketo, P. Inorg. Chim. Acta 2000, 300, 810.
1152. Ishii, S.-I.; Mitani, M.; Saito, J.; Matsuura, S.; Kojoh, S.-I.; Kashiwa, N.; Fujita, T. Chem. Lett. 2002, 7, 740.
1153. Cavallo, L.; Guerra, G. Macromolecules 1996, 29, 2729.
1154. Deng, L.; Woo, T. K.; Cavallo, L.; Margl, P. M.; Ziegler, T. J. Am. Chem. Soc. 1997, 119, 6177.
1155. Talarico, G.; Busico, V.; Cavallo, L. Organometallics 2004, 23, 5989.
1156. Mu lhaupt, R.; Duschek, T.; Fischer, D.; Setz, S. Polym. Adv. Tech. 1993, 4, 439.
1157. Koo, K.; Marks, T. J. J. Am. Chem. Soc. 1999, 121, 8791.
1158. Xu, G.; Chug, T. C. J. Am. Chem. Soc. 1999, 121, 6763.
1159. Matsui, S.; Mitani, M.; Saito, J.; Matsukawa, N.; Tanaka, T.; Nakano, T.; Fujita, T. Chem. Lett. 2000, 29, 554.
1160. Matsugi, T.; Matsui, S.; Kojoh, S.-I.; Takagi, Y.; Inoue, Y.; Fujita, T.; Kashiwa, N. Chem. Lett. 2001, 6, 566.
1161. Saito, J.; Mitani, M.; Matsui, S.; Kashiwa, N.; Fujita, T. Macromol. Rapid Commun. 2000, 21, 1333.
1162. Saito, J.; Onda, M.; Matsui, S.; Mitani, M.; Furuyama, R.; Tanaka, H.; Fujita, T. Macromol. Rapid Commun. 2002, 23, 1118.
1163. Prasad, A. V.; Makio, H.; Saito, J.; Onda, M.; Fujita, T. Chem. Lett. 2004, 33, 250.
1164. Kui, S. C. F.; Zhu, N.; Chan, M. C. W. Angew. Chem., Int. Ed. 2003, 42, 1628.
1165. Ittel, S. D.; Johnson, L. K.; Brookhart, M. Chem. Rev. 2000, 100, 1169.
1164 Olefin Polymerizations with Group IV Metal Catalysts
1166. Reinartz, S.; Mason, A. F.; Lobkovsky, E. B.; Coates, G. W. Organometallics 2003, 22, 2542.
1167. Pennington, D. A.; Coles, S. J.; Hursthouse, M. B.; Bochmann, M.; Lancaster, S. J. Chem. Commun. 2005, 3150.
1168. Mason, A. F.; Coates, G. W. J. Am. Chem. Soc. 2004, 126, 10798.
1169. Saito, J.; Mitani, M.; Onda, M.; Mohri, J.-I.; Ishii, S.-I.; Yoshida, Y.; Nakano, T.; Tanaka, H.; Matsugi, T., Kojoh, S.-I., et al. Macromol. Rapid
Commun. 2001, 22, 1072.
1170. Mitani, M.; Furuyama, R.; Mohri, J.-i.; Saito, J.; Ishii, S.; Terao, H.; Kashiwa, N.; Fujita, T. J. Am. Chem. Soc. 2002, 124, 7888.
1171. Makio, H.; Tohi, Y.; Saito, J.; Onda, M.; Fujita, T. Macromol. Rapid Commun. 2003, 24, 894.
1172. Mitani, M.; Furuyama, R.; Mohri, J.-i.; Saito, J.; Ishii, S.; Terao, H.; Nakano, T.; Tanaka, H.; Fujita, T. J. Am. Chem. Soc. 2003, 125, 4293.
1173. Saito, J.; Mitani, M.; Mohri, J.-I.; Ishii, S.-I.; Yoshida, Y.; Matsugi, T.; Kojoh, S.-I.; Kashiwa, N.; Fujita, T. Chem. Lett. 2001, 6, 576.
1174. Coates, G. W.; Tian, J. Angew. Chem., Int. Ed. 2000, 39, 3626.
1175. Tian, J.; Hustad, P. D.; Coates, G. W. J. Am. Chem. Soc. 2001, 123, 5134.
1176. Coates, G. W.; Hustad, P. D.; Reinartz, S. Angew. Chem., Int. Ed. 2002, 41, 2236.
1177. Mason, A. F.; Tian, J.; Hustad, P. D.; Lobkovsky, E. B.; Coates, G. W. Isr. J. Chem. 2002, 42, 301.
1178. Lamberti, M.; Pappalardo, D.; Mazzeo, M.; Pellecchia, C. Macromol. Chem. Phys. 2004, 205, 486.
1179. Herfert, N.; Fink, G. Makromol. Chem. 1992, 193, 773.
1180. Harrod, J. F.; Taylor, K. J. Chem. Soc., Chem. Commun. 1971, 696.
1181. Bei, X.; Swenson, D. C.; Jordan, R. F. Organometallics 1997, 16, 3282.
1182. Mason, A. F.; Coates, G. W. J. Am. Chem. Soc. 2004, 126, 16326.
1183. Inoue, Y.; Nakano, T.; Tanaka, H.; Kashiwa, N.; Fujita, T. Chem. Lett. 2001, 10, 1060.
1184. Tsukahara, T.; Swenson, D. C.; Jordan, R. F. Organometallics 1997, 16, 3303.
1185. Kim, I.; Nishihara, Y.; Jordan, R. F.; Rogers, R. D.; Rheingold, A. L.; Yap, G. P. A. Organometallics 1997, 16, 3314.
1186. Li, X.-F.; Dai, K.; Ye, W.-P.; Pan, L.; Li, Y.-S. Organometallics 2004, 23, 1223.
1187. Hu, W.-Q.; Sun, X.-L.; Wang, C.; Gao, Y.; Tang, Y.; Shi, L.-P.; Xia, W.; Sun, J.; Dai, H.-L., Li, X.-Q., et al. Organometallics 2004, 23, 1684.
1188. Knight, P. D.; Clarke, A. J.; Kimberley, B. S.; Jackson, R. A.; Scott, P. Chem. Commun. 2002, 352.
1189. Cuomo, C.; Strianese, M.; Cuenca, T.; Sanz, M.; Grassi, A. Macromolecules 2004, 37, 7469.
1190. OConnor, P. E.; Morrison, D. J.; Steeves, S.; Burrage, K.; Berg, D. J. Organometallics 2001, 20, 1153.
1191. Tshuva, E. Y.; Versano, M.; Goldberg, I.; Kol, M.; Weitman, H.; Goldschmidt, Z. Inorg. Chem. Commun. 1999, 2, 371.
1192. Ziniuk, Z.; Goldberg, I.; Kol, M. Inorg. Chem. Commun. 1999, 2, 549.
1193. Tshuva, E. Y.; Goldberg, I.; Kol, M. J. Am. Chem. Soc. 2000, 122, 10706.
1194. Tshuva, E. Y.; Goldberg, I.; Kol, M.; Weitman, H.; Goldschmidt, Z. Chem. Commun. 2000, 379.
1195. Tshuva, E. Y.; Goldberg, I.; Kol, M.; Goldschmidt, Z. Inorg. Chem. Commun. 2000, 3, 611.
1196. Tshuva, E. Y.; Goldberg, I.; Kol, M.; Goldschmidt, Z. Organometallics 2001, 20, 3017.
1197. Tshuva, E. Y.; Goldberg, I.; Kol, M.; Goldschmidt, Z. Chem. Commun. 2001, 2120.
1198. Tshuva, E. Y.; Goldberg, I.; Kol, M.; Goldschmidt, Z. Inorg. Chem. 2001, 40, 4263.
1199. Tshuva, E. Y.; Groysman, S.; Goldberg, I.; Kol, M.; Goldschmidt, Z. Organometallics 2002, 21, 662.
1200. Groysman, S.; Goldberg, I.; Kol, M.; Genizi, E.; Goldschmidt, Z. Organometallics 2003, 22, 3013.
1201. Groysman, S.; Goldberg, I.; Kol, M.; Genizi, E.; Goldschmidt, Z. Inorg. Chim. Acta 2003, 345, 137.
1202. Groysman, S.; Tshuva, E. Y.; Goldberg, I.; Kol, M.; Goldschmidt, Z.; Shuster, M. Organometallics 2004, 23, 5291.
1203. Yeori, A.; Gendler, S.; Groysman, S.; Goldberg, I.; Kol, M. Inorg. Chem. Commun. 2004, 7, 280.
1204. Segal, S.; Goldberg, I.; Kol, M. Organometallics 2005, 24, 200.
1205. Jeske, G.; Lauke, H.; Mauermann, H.; Schumann, H.; Marks, T. J. J. Am. Chem. Soc. 1985, 107, 8111.
1206. Busico, V.; Cipullo, R.; Friederichs, N.; Ronca, S.; Talarico, G.; Togrou, M.; Wang, B. Macromolecules 2004, 37, 8201.
1207. Matsuo, Y.; Mashima, K.; Tani, K. Chem. Lett. 2000, 29, 1114.
1208. Yoshida, Y.; Matsui, S.; Takagi, Y.; Mitani, M.; Nitabaru, M.; Nakano, T.; Tanaka, H.; Fujita, T. Chem. Lett. 2000, 29, 1270.
1209. Yoshida, Y.; Saito, J.; Mitani, M.; Takagi, Y.; Matsui, S.; Ishii, S.-i.; Nakano, T.; Kashiwa, N.; Fujita, T. Chem. Commun. 2002, 1298.
1210. Yoshida, Y.; Matsui, S.; Takagi, Y.; Mitani, M.; Nakano, T.; Tanaka, H.; Kashiwa, N.; Fujita, T. Organometallics 2001, 20, 4793.
1211. Matsui, S.; Spaniol, T. P.; Takagi, Y.; Yoshida, Y.; Okuda, J. J. Chem. Soc., Dalton Trans. 2002, 4529.
1212. Matsui, S.; Yoshida, Y.; Takagi, Y.; Spaniol, T. P.; Okuda, J. J. Organomet. Chem. 2004, 689, 1155.
1213. Yoshida, Y.; Nakano, T.; Tanaka, H.; Fujita, T. Isr. J. Chem. 2002, 42, 353.
1214. Dawson, D. M.; Walker, D. A.; Thornton-Pett, M.; Bochmann, M. J. Chem. Soc., Dalton Trans. 2000, 459.
1215. Matsugi, T.; Matsui, S.; Kojoh, S.-I.; Takagi, Y.; Inoue, Y.; Nakano, T.; Fujita, T.; Kashiwa, N. Macromolecules 2002, 35, 4880.
1216. Richter, J.; Edelmann, F. T.; Noltemeyer, M.; Schmidt, H.-G.; Shmulinson, M.; Eisen, M. S. J. Mol. Catal. A: Chem. 1998, 130, 149.
1217. Martin, A.; Uhrhammer, R.; Gardner, T. G.; Jordan, R. F.; Rogers, R. D. Organometallics 1998, 17, 382.
1218. Carpentier, J.-F.; Martin, A.; Swenson, D. C.; Jordan, R. F. Organometallics 2003, 22, 4999.
1219. Nakazawa, H.; Ikai, S.; Imaoka, K.; Kai, Y.; Yano, T. J. Mol. Catal. A: Chem. 1998, 132, 33.
1220. Murtuza, S.; Osvaldo, L.; Casagrande, J.; Jordan, R. F. Organometallics 2002, 21, 1882.
1221. Gil, M. P.; Santos, J. H. Z. d.; Osvaldo, L.; Casagrande, J. J. Mol. Catal. A: Chem. 2004, 209, 163.
1222. Fuhrmann, H.; Brenner, S.; Arndt, P.; Kempe, R. Inorg. Chem. 1996, 35, 6742.
1223. Adams, N.; Arts, H. J.; Bolton, P. D.; Cowell, D.; Dubberley, S. R.; Friederichs, N.; Grant, C. M.; Kranenburg, M.; Sealey, A. J., Wang, B.,
et al. Chem. Comm. 2004, 434.
1224. Ku hl, O.; Koch, T.; Somoza, F. B., Jr.; Junk, P. C.; Hey-Hawkins, E.; Plat, D.; Eisen, M. S. J. Organomet. Chem. 2000, 604, 116.
1225. Benetollo, F.; Carta, G.; Cavinato, G.; Crociani, L.; Paolucci, G.; Rossetto, G.; Veronese, F.; Zanella, P. Organometallics 2003, 22, 3985.
1226. Baker, R. J.; Edwards, P. G. J. Chem. Soc., Dalton Trans. 2002, 2960.
1227. Kojoh, S.-I.; Matsugi, T.; Saito, J.; Mitani, M.; Fujita, T.; Kashiwa, N. Chem. Lett. 2001, 8, 822.
1228. Ishii, S.; Saito, J.; Matsuura, S.; Suzuki, Y.; Furuyama, R.; Mitani, M.; Nakano, T.; Kashiwa, N.; Fujita, T. Macromol. Rapid Commun. 2002,
23, 693697.
1229. Hustad, P. D.; Coates, G. W. J. Am. Chem. Soc. 2002, 124, 11578.
1230. Yoshida, Y.; Mohri, J.-i.; Ishii, S.-i.; Mitani, M.; Saito, J.; Matsui, S.; Makio, H.; Nakano, T.; Tanaka, H., Onda, M., et al. J. Am. Chem. Soc.
2004, 126, 12023.
1231. Furuyama, R.; Mitani, M.; Mohri, J.; Mori, R.; Tanaka, H.; Fujita, T. Macromolecules 2005, 38, 1546.
Olefin Polymerizations with Group IV Metal Catalysts 1165
1232. Bates, F. S. Science 1991, 251, 898.
1233. Holden, G. In Encyclopedia of Polymer Science and Engineering; Kroschwitz, J. I., Salvatore, A., Klingsberg, A., Muldoon, J., Eds.; Wiley: New
York, 1986; Vol. 5, p 416.
1234. Ruokolainen, J.; Mezzenga, R.; Fredrickson, G. H.; Kramer, E. J.; Hustad, P. D.; Coates, G. W. Macromolecules 2005, 38, 851.
1235. Brown, S. J.; Gao, X.; Harrison, D. G.; McKay, I.; Koch, L.; Wang, Q.; Xu, W.; Von Haken Spence, R. E.; Stephan, D. W. (Nova Chemicals
Corporation). U.S. Patent 2001007895, 2001.
1236. Ahn, C.-H.; Tahara, M.; Uozumi, T.; Jin, J.; Tsubaki, S.; Soga, K. Macromol. Rapid Commun. 2000, 21, 385.
1237. Auriemma, F.; Rosa, C. D.; Esposito, S.; Coates, G. W.; Fujita, M. J. Am. Chem. Soc. 2005, 127, 2850.
1238. Grassi, A.; Maffei, G.; Milione, S.; Jordan, R. F. Macromol. Chem. Phys. 2001, 202, 1239.
1239. Natta, G.; Pino, P.; Corradini, P.; Danusso, F.; Mantica, E.; Mazzanti, G.; Moraglio, G. J. Am. Chem. Soc. 1955, 77, 1708.
1240. Capacchione, C.; Proto, A.; Ebeling, H.; Mu lhaupt, R.; Mo ller, K.; Spaniol, T. P.; Okuda, J. J. Am. Chem. Soc. 2003, 125, 4964.
1241. Beckerle, K.; Capacchione, C.; Ebeling, H.; Manivannan, R.; Mu lhaupt, R.; Proto, A.; Spaniol, T. P.; Okuda, J. J. Organomet. Chem. 2004,
689, 4636.
1242. Capacchione, C.; Proto, A.; Ebeling, H.; Mulhaupt, R.; Moller, K.; Manivannan, R.; Spaniol, T. P.; Okuda, J. J. Mol. Catal. A: Chem. 2004,
213, 137.
1243. Capacchione, C.; Manivannan, R.; Barone, M.; Beckerle, K.; Centore, R.; Oliva, L.; Proto, A.; Tuzi, A.; Spaniol, T. P.; Okuda, J.
Organometallics 2005, 24, 2971.
1244. Capacchione, C.; De Carlo, F.; Zannoni, C.; Okuda, J.; Proto, A. Macromolecules 2004, 37, 8918.
1245. Kakugo, M.; Miyatake, T.; Mizunuma, K. Stud. Surf. Sci. Catal. 1990, 56, 517.
1246. Fokken, S.; Spaniol, T. P.; Okuda, J.; Sernetz, F. G.; Mu lhaupt, R. Organometallics 1997, 16, 4240.
1166 Olefin Polymerizations with Group IV Metal Catalysts

You might also like