'Ziegler–Natta Catalysts'. In: Encyclopedia of Polymer...

20
ZIEGLER–NATTA CATALYSTS Introduction Ziegler–Natta catalysts have had enormous impact on the polymer industry in the past 50 years, with current world production of polyolefins using Ziegler–Natta catalysis amounting to more than 50 million tons per annum. The vast advances made during the past decades stem from breakthrough discoveries made by Karl Ziegler and Giulio Natta in the early 1950s. It was in 1953 that Ziegler and co- workers, at the Max Planck Institute in M ¨ ulheim, were investigating the “Aufbau” reaction in which triethylaluminum reacts with ethylene to give higher aluminum trialkyls (1). Unexpectedly, one experiment led not to the oligomerization of ethy- lene via the Aufbau reaction, but to the formation of 1-butene. It turned out that this dimerization reaction had been catalyzed by traces of nickel present as a contaminant in the reactor. Soon afterwards, a revolutionary breakthrough was achieved when combinations of transition-metal compounds and aluminum alkyls were found that could polymerize ethylene under mild conditions, yielding high density polyethylene (2,3). In 1954 Giulio Natta and co-workers at Milan Poly- technic succeeded not only in polymerizing propylene with the Ziegler catalyst combination TiCl 4 /Al(C 2 H 5 ) 3 , but also in fractionating the resulting polymer to obtain and characterize isotactic polypropylene (4–6). This demonstration of stere- oregular polymerization led to an explosive growth of new polymers and industrial applications as the full scope of Ziegler–Natta catalysis was realized (7–9); Ziegler and Natta were jointly awarded the Nobel Prize for Chemistry in 1963. Ziegler–Natta catalysts for polyethylene and polypropylene have progressed from first-generation titanium trichloride catalysts, used in the manufacturing processes of the late 1950s and the 1960s, to the high activity magnesium chlo- ride supported catalysts used today. Improvements in catalyst performance have 517 Encyclopedia of Polymer Science and Technology. Copyright John Wiley & Sons, Inc. All rights reserved.

Transcript of 'Ziegler–Natta Catalysts'. In: Encyclopedia of Polymer...

ZIEGLER–NATTA CATALYSTS

Introduction

Ziegler–Natta catalysts have had enormous impact on the polymer industry in thepast 50 years, with current world production of polyolefins using Ziegler–Nattacatalysis amounting to more than 50 million tons per annum. The vast advancesmade during the past decades stem from breakthrough discoveries made by KarlZiegler and Giulio Natta in the early 1950s. It was in 1953 that Ziegler and co-workers, at the Max Planck Institute in Mulheim, were investigating the “Aufbau”reaction in which triethylaluminum reacts with ethylene to give higher aluminumtrialkyls (1). Unexpectedly, one experiment led not to the oligomerization of ethy-lene via the Aufbau reaction, but to the formation of 1-butene. It turned out thatthis dimerization reaction had been catalyzed by traces of nickel present as acontaminant in the reactor. Soon afterwards, a revolutionary breakthrough wasachieved when combinations of transition-metal compounds and aluminum alkylswere found that could polymerize ethylene under mild conditions, yielding highdensity polyethylene (2,3). In 1954 Giulio Natta and co-workers at Milan Poly-technic succeeded not only in polymerizing propylene with the Ziegler catalystcombination TiCl4/Al(C2H5)3, but also in fractionating the resulting polymer toobtain and characterize isotactic polypropylene (4–6). This demonstration of stere-oregular polymerization led to an explosive growth of new polymers and industrialapplications as the full scope of Ziegler–Natta catalysis was realized (7–9); Zieglerand Natta were jointly awarded the Nobel Prize for Chemistry in 1963.

Ziegler–Natta catalysts for polyethylene and polypropylene have progressedfrom first-generation titanium trichloride catalysts, used in the manufacturingprocesses of the late 1950s and the 1960s, to the high activity magnesium chlo-ride supported catalysts used today. Improvements in catalyst performance have

517Encyclopedia of Polymer Science and Technology. Copyright John Wiley & Sons, Inc. All rights reserved.

518 ZIEGLER–NATTA CATALYSTS Vol. 8

facilitated the development of efficient gas-phase and bulk processes for polyethy-lene and polypropylene, and at the same time have led to ever-increasing controlover polymer composition and properties.

Early Catalysts

Titanium Trichloride. One of the first catalysts found by Ziegler to beeffective in ethylene polymerization was the product of the reaction of titaniumtetrachloride with triethylaluminum. At low Al/Ti ratios, this reaction yields ti-tanium trichloride as a solid precipitate. TiCl3 exists in four crystalline modifica-tions, the α, β, δ, and γ forms, of which the β-modification has a linear (chain-like)structure and the α, δ, and γ forms have a layered structure (10,11). The reactionproduct of TiCl4 and AlR3 is β-TiCl3, which can be converted to the γ form byheating. The latter catalyst has much higher stereoregulating ability in propy-lene polymerization, while β-TiCl3 is an effective catalyst for the production ofcis-1,4-polyisoprene. α-TiCl3 can be prepared by reduction of TiCl4 with hydrogenor with aluminum powder. The δ form can be prepared by prolonged grinding of γ -or α-TiCl3 and has a more disordered structure as a result of sliding of Cl–Ti–Cltriple layers during mechanical activation (12).

The first-generation Ziegler–Natta catalysts used in early manufacturingprocesses for polypropylene (PP) comprised TiCl3 and cocrystallized AlCl3, result-ing from reduction of TiCl4 with Al or an aluminum alkyl. The cocatalyst used inthe polymerization process was Al(C2H5)2Cl (DEAC). Catalyst activity was rela-tively low, giving polymer yields of around 1 kg PP/g cat., necessitating removal(deashing) of catalyst residues from the polymer. In many cases, extractive re-moval of atactic polymer was also required.

Other Early Developments. In addition to the breakthrough by Ziegler,two other discoveries of ethylene polymerization catalysts were made in the early1950s. A patent by Standard Oil of Indiana, filed in 1951, disclosed reduced molyb-denum oxide or cobalt molybdate on alumina (13). At the same time, Phillipsdiscovered supported chromium oxide catalysts, prepared by impregnation of asilica–alumina support with CrO3 (14–16). Both the Phillips catalyst and tita-nium chloride based Ziegler catalysts are widely used in the production of highdensity polyethylene (HDPE).

The various discoveries made independently by different industrial researchgroups in the early 1950s resulted in intensive patent litigations (8,17,18), whichin the case of PP continued up to the 1980s, when a composition of matter patenton PP was awarded in the United States to Phillips, because a fraction of crys-talline PP was found to be present in a polymer prepared using a CrO3/Al2O3/SiO2catalyst (19). However, despite the importance of the Phillips catalyst for HDPE,it was unsuitable for PP, which is produced entirely using Ziegler–Natta catalystsand (to a much smaller extent) metallocene-based catalysts.

Second-Generation Catalysts. In the 1970s, an improved TiCl3 cata-lyst for PP was developed by Solvay (20). Catalyst preparation involved reduc-tion of TiCl4 using DEAC, followed by treatment with an ether and TiCl4. Theether treatment results in removal of AlCl3 from TiCl3 · nAlCl3, while treatmentwith TiCl4 effects a phase transformation from β- to δ-TiCl3 at a relatively mild

Vol. 8 ZIEGLER–NATTA CATALYSTS 519

temperature (<100◦C) (21). Using catalysts of this type, it was possible to obtainPP yields in the range 5–20 kg/g cat. in 1–4 h of polymerization in liquid monomer(22). Commercial implementation of second-generation catalysts was, however,overshadowed by the advent of third- and later-generation magnesium chloridesupported catalysts (discussed under Ziegler–Natta Catalysts for Polypropylene).

Polymerization and Particle Growth

Polymer Chain Growth. The essential characteristic of Ziegler–Nattacatalysis is the polymerization of an olefin or diene using a combination of atransition-metal compound and a base-metal alkyl cocatalyst, normally an alu-minum alkyl. The function of the cocatalyst is to alkylate the transition metal,generating a transition-metal–carbon bond. It is also essential that the activecenter contains a coordination vacancy. Chain propagation takes place via theCossee–Arlman mechanism (23), in which coordination of the olefin at the vacantcoordination site is followed by chain migratory insertion into the metal–carbonbond, as illustrated in Figure 1.

Regulation of polyolefin molecular weight is effected by the use of hydrogenas chain-transfer agent. Chain transfer can also occur via β-hydrogen transferfrom the growing chain to the transition metal or to the monomer, and to a lesserextent via alkyl exchange with the cocatalyst (Fig. 2).

In propylene polymerization using titanium chloride catalysts, chain prop-agation takes place via primary (1,2-) insertion of the monomer. For isospecificpropagation, there must be only one coordination vacancy and the active sitemust be chiral. Corradini and co-workers have demonstrated that the asymmetric

TiCl

R

ClCl

Cl

TiCl

R

ClCl

Cl CH2

CH2

Ti CH2-CH2-RCl

ClCl

Cl

TiCl

ClCl

ClCH2

R CH2

Fig. 1. Cossee–Arlman mechanism for polymerization.

Ti-CH2-CH2-Polymer + H2 Ti-H + CH3-CH2-Polymer

Ti-CH2-CH2-Polymer + CH2 Ti-CH2-CH3 +CH2 CH2 CH-Polymer

Ti-CH2-CH2-Polymer + AlR3 Ti-R + AlR2-CH2-CH2-Polymer

Fig. 2. Chain transfer in ethylene polymerization.

520 ZIEGLER–NATTA CATALYSTS Vol. 8

Cl

Ti Cl

Ti

Cl Ti

Cl

R

Cl

Cl∗

Fig. 3. Model for stereospecific polymerization of propylene. The orientation of the grow-ing chain is influenced by the chlorine atom marked with an asterisk.

environment of the active site forces the growing chain to adopt a particular ori-entation so as to minimize steric interactions with (chlorine) ligands present onthe catalyst surface (24). This in turn leads to one particular prochiral face of theincoming monomer being preferred, as illustrated in Figure 3, leading to isotacticpolymer.

An elegant demonstration of the above mechanism has been provided byZambelli and co-workers (25), who showed that the first insertion of propyleneinto a Ti CH3 bond generated by chain transfer with Al alkyl using the sys-tem TiCl3/Al(CH3)3 is not stereospecific, whereas the second insertion (ie intoTi–isobutyl) is stereospecific. The importance of the combined effects of the stericbulk of the Ti–alkyl group and the halide ligand is apparent from the very highstereospecificity observed using TiI3 (26). Particularly high stereospecificity (butlow activity) was also found by Natta when TiCl3 was used in combination withAl(C2H5)2I (27).

In contrast to the isospecific titanium-based catalysts, vanadium-based cata-lysts give predominantly syndiotactic PP. At very low polymerization temperature(−78◦C), living polymerization can be obtained using homogeneous catalysts ob-tained by reaction of a vanadium compound (eg VCl4 or a V(III) β-diketonate)with R2AlCl (28,29). With these catalysts, syndiospecific propagation occurs viasecondary (2,1-) insertion of the monomer. The overall stereo- and regioregular-ity of the polymer is poor, comprising not only syndiotactic blocks resulting fromsecondary insertions but also short, atactic blocks arising from sequences of pri-mary insertions. This polymer has not been developed commercially, but vanadiumcatalysts are used in ethylene (co)polymerization (outlined under Ziegler–NattaCatalysts for Ethylene (Co)polymerization). Cs-symmetric metallocene catalysts(30) have been developed for the production of syndiotactic polypropylene havingsignificantly higher chain regularity.

Polymer Particle Growth. A very important feature of any heterogeneouscatalyst used in slurry and gas-phase processes for polyolefin production is par-ticle morphology. Heterogeneous Ziegler–Natta catalysts are microporous solids,with particle sizes typically in the range 10–100 µm. Each particle comprisesmillions of primary crystallites with sizes of up to about 15 nm. On contactingthe catalyst components, at the start of polymerization, cocatalyst and monomerdiffuse through the catalyst particle and polymerization takes place on the sur-face of each primary crystallite within the particle. As solid, crystalline polymer isformed, the primary crystallites are pushed apart as the particle grows, analogousto the expanding universe. The particle shape is retained, and this phenomenon

Vol. 8 ZIEGLER–NATTA CATALYSTS 521

Catalyst

Prepolymer

Polymer

Fig. 4. “Replication” phenomenon during polymerization.

is therefore referred to as replication (Fig. 4). Ideally, the catalyst particle shouldhave spherical morphology and controllable porosity. It is important that the me-chanical strength of the catalyst is high enough to prevent disintegration but lowenough to allow progressive expansion as polymerization proceeds (31). Furtherimplications of particle morphology and porosity are discussed under Ziegler–Natta Catalysts for Ethylene (Co)polymerization and also under Reactor GranuleTechnology.

Ziegler–Natta Catalysts for Ethylene (Co)polymerization

Ziegler–Natta catalysts are widely used in the production of high density andlinear low density polyethylene (HDPE and LLDPE). More than half the worldproduction of HDPE is based on Ziegler–Natta catalysts, chromium catalysts alsobeing widely used. Less than 1% of HDPE production utilizes metallocene or othersingle-site catalysts. In LLDPE production, Ziegler–Natta catalysts occupy a dom-inant position, accounting for more than 90% of the total production. Single-sitecatalysts currently account for less than 10% of this market, but increased use ofsuch catalysts is expected throughout the next decade.

The most important titanium-based catalysts for HDPE and LLDPE arethose comprising a titanium component on magnesium chloride or on a magnesiumchloride containing support. Toward the end of the 1960s, catalysts obtained byreaction of TiCl4 or a derivative thereof with a magnesium compound such asMg(OH)Cl, Mg(OH)2, or MgCl2 were found to give very high activity in ethylenepolymerization, eliminating the need for deashing of the polymer (31,32). Themost effective support was found to be active magnesium chloride, prepared byco-milling of MgCl2 and titanium halides or by chlorination of organomagnesiumcompounds (32). Numerous catalyst systems and methods of preparation havebeen disclosed (33), and the characteristics of magnesium chloride as a supportfor Ziegler–Natta catalysts are discussed in depth under Ziegler–Natta Catalystsfor Polypropylene. Magnesium chloride can also be used in combination with asilica support, for example by impregnation of the porous support with a solutionof MgCl2 and TiCl4 in tetrahydrofuran (34).

522 ZIEGLER–NATTA CATALYSTS Vol. 8

An important manufacturing process for HDPE that makes use of highmileage catalysts is the cascade process, in which polymerization reactors in seriesare used to give reactor blends with improved properties for film and pipe appli-cations (35). Broad molecular weight distribution (MWD) can be obtained by theuse of different hydrogen concentrations in each reactor. In addition, the processcan be designed to give low molecular weight homopolymer in the first reactorand a high molecular weight copolymer in the second. The high molecular weightcopolymer chains function as tie molecules linking the crystalline, homopolymerdomains, thereby leading to high stress crack resistance of the polymer. This pro-cess allows an “inverse” comonomer distribution to be obtained, in the sense thatthe comonomer is in the high molecular weight fraction, counteracting the generaltendency of Ziegler–Natta catalysts to incorporate the comonomer mainly in thelow molecular weight chains. The latter feature is an important consideration inZiegler–Natta catalyst design for LLDPE. Comonomer incorporation is highestat the most open catalytic centers, whereas sterically hindered centers will tendto give polyethylene chains with little or no comonomer. The best catalysts forLLDPE are therefore those that have relatively uniform active center distribu-tion, lacking excessively hindered or unhindered active sites.

Vanadium catalysts have also been developed for polyethylene and ethylene-based copolymers, particularly ethylene–propylene–diene rubbers (EPDM). Ho-mogeneous (soluble) vanadium catalysts produce relatively narrow MWDpolyethylene, whereas supported vandium catalysts give broad MWD (36). Poly-merization activity is strongly enhanced by the use of a halogenated hydrocarbonas promoter in combination with a vanadium catalyst and aluminum alkyl cocat-alyst (36,37).

Ethylene polymerization, in contrast to the polymerization of propylene andother alpha-olefins, is often affected by diffusion limitations, which occur if themonomer reactivity in polymerization is high relative to diffusivity through thecatalyst particle. This can result in the formation of an “onion” particle structureas polymerization first takes place at the external surface of the particle, particlegrowth occurring step by step as the monomer reaches the inner parts of thecatalyst particle. This mechanism of particle growth is associated with a kineticprofile 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.

Ziegler–Natta Catalysts for Polypropylene

Worldwide manufacture of PP, currently around 30 million tons per annum, isdominated by high activity MgCl2-supported Ziegler–Natta catalysts. The first-and second-generation TiCl3 catalysts have all but disappeared, and the recentlydeveloped metallocene catalysts still account for less than 1% of all PP produced,although they are likely to grow in importance. The development and implemen-tation of MgCl2-supported catalysts in bulk (liquid monomer) and gas-phase pro-cesses has led to the advent of simple, low-cost (nondeashing, nonextracting) man-ufacturing processes for PP (18).

The basis for the development of the high activity supported catalysts lay inthe discovery, in the late 1960s, of “activated” MgCl2 able to support TiCl4 and give

Vol. 8 ZIEGLER–NATTA CATALYSTS 523

high catalyst activity, and the subsequent discovery, in the mid-1970s, of electrondonors (Lewis bases) capable of increasing the stereospecificity of the catalyst sothat (highly) isotactic PP could be obtained (32,38,39). A further feature that hascontributed greatly to the commercial success of MgCl2-supported catalysts is thedevelopment of spherical catalysts with controlled particle size and porosity (40),which not only replicate their morphology during polymerization as the polymerparticle grows, but which have now opened the way to a broad range of homo-and copolymers and multiphase polymer alloys via what has been termed ReactorGranule Technology (41).

Catalyst Structure and Composition. In the early stages of MgCl2-supported catalyst development, activated magnesium chloride was prepared byball milling in the presence of ethyl benzoate, leading to the formation of verysmall (≤3 nm thick) primary crystallites within each particle (21). Nowadays,however, the activated support is prepared by chemical means such as complexformation of MgCl2 and an alcohol or by reaction of a magnesium alkyl or alkoxidewith a chlorinating agent or TiCl4. Many of these approaches are also effective forthe preparation of catalysts having controlled particle size and morphology. Forexample, the cooling of emulsions of molten MgCl2 · nC2H5OH in paraffin oil givesalmost perfectly spherical supports, which are then converted into the catalysts(18). A typical catalyst preparation involves reaction of the MgCl2 · nC2H5OHsupport with excess TiCl4 in the presence of an “internal” electron donor. Temper-atures of at least 80◦C and at least two TiCl4 treatment steps are normally used,in order to obtain high performance catalysts in which the titanium is mainlypresent as TiCl4 rather than the TiCl3OC2H5 generated in the initial reactionwith the support. Catalysts obtained via chemical routes generally have a BETsurface area of around 300 m2/g and pore volumes in the range 0.3–0.4 cm3/g (18).

High activity Ziegler–Natta catalysts comprising MgCl2, TiCl4, and an in-ternal donor are typically used in combination with an aluminum alkyl cocat-alyst such as Al(C2H5)3 and an “external” electron donor added in polymeriza-tion. The first catalyst systems containing ethyl benzoate as internal donor wereused in combination with a second aromatic ester such as methyl p-toluate asexternal donor (39). These were followed by catalysts containing a diester (egdiisobutyl phthalate) as internal donor, used in combination with an alkoxysi-lane external donor of type RR′Si(OCH3)2 or RSi(OCH3)3 (42). The combinationMgCl2/TiCl4/phthalate ester–AlR3–alkoxysilane is currently the most widely usedcatalyst system in PP manufacture. The most effective alkoxysilane donors forhigh isospecificity are methoxysilanes containing relatively bulky alkyl groupsbranched at the position alpha to the silicon atom (43–46). Typical examples in-clude cyclohexyl(methyl)dimethoxysilane and dicyclopentyldimethoxysilane (47).Of these, the latter gives particularly high stereospecificity (48) and broader MWD(49). High PP stereoregularity and broad MWD have also been obtained by theuse of dimethoxysilanes containing polycyclic amino groups (50,51).

The function of the internal donor in MgCl2-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 TiCl4 in the final catalyst.Activated magnesium chloride has a disordered structure comprising very smalllamellae. Giannini (32) has indicated that, on preferential lateral cleavage sur-faces, the magnesium atoms are coordinated with four or five chlorine atoms, as

524 ZIEGLER–NATTA CATALYSTS Vol. 8

Cl

Mg

(110)cut

(100)cut

Fig. 5. Model of a MgCl2 layer showing the (100) and (110) lateral cuts. Based on Ref. 31.

opposed to six chlorine atoms in the bulk of the crystal. These lateral cuts corre-spond to (110) and (100) faces of MgCl2, as shown in Figure 5.

It has been proposed that bridged, dinuclear Ti2Cl8 species can coordinate tothe (100) face of MgCl2 and give rise to the formation of chiral, isospecific activespecies (Fig. 6), it being pointed out that Ti2Cl6 species formed by reduction oncontact with Al(C2H5)3 would resemble analogous species in TiCl3 catalysts (52,53). Accordingly, it has been suggested (18) that a possible function of the internaldonor is preferential coordination on the more acidic (110) face of MgCl2, suchthat this face is prevailingly occupied by donor and the (100) face is prevailinglyoccupied by Ti2Cl8 dimers.

Analytical studies (54) have indicated that a monoester internal donor suchas ethyl benzoate is coordinated to MgCl2 and not to TiCl4. In the search fordonors giving catalysts having improved performance, it was considered (55) thatbidentate donors should be able to form strong chelating complexes with tetra-coordinate Mg atoms on the (110) face of MgCl2, or binuclear complexes withtwo pentacoordinate Mg atoms on the (100) face. This led to the developmentof the MgCl2/TiCl4/phthalate ester catalysts, used as indicated above in combi-nation with an alkoxysilane as external donor. The requirement for an externaldonor when using catalysts containing a benzoate or phthalate ester is due to

Vol. 8 ZIEGLER–NATTA CATALYSTS 525

Fig. 6. Model showing dimeric and monomeric Ti species on a (100) lateral cut of MgCl2.Based on Ref. 31.

the fact that, when the catalyst is brought into contact with the cocatalyst, alarge proportion of the internal donor is lost as a result of alkylation and/or com-plexation reactions. In the absence of an external donor, this leads to poor stere-ospecificity because of increased mobility of the titanium species on the catalystsurface. When the external donor is present, contact of the catalyst componentsleads to replacement of the internal donor by the external donor, as has beenshown (56,57) with MgCl2/TiCl4/ethyl benzoate–Al(C2H5)3–methyl p-toluate andwith MgCl2/TiCl4/dibutyl phthalate–Al(C2H5)3–C6H5Si(OC2H5)3. The most activeand stereospecific systems were those which allowed the highest incorporation ofexternal donor (58), the effectiveness of a catalyst system depending more on thecombination of donors rather than on the individual internal or external donor. Forexample, the use of a monoester rather than an alkoxysilane as external donorwith a phthalate-containing system is ineffective (59), as in this case very lit-tle of the external donor is adsorbed (58). Further studies (60,61) showed that aphthalate-containing catalyst adsorbed alkoxysilanes to a greater extent than acatalyst without internal donor.

Recently, research on MgCl2-supported catalysts has led to systems not re-quiring the use of an external donor. This required the identification of bidentateinternal donors that not only had the right oxygen–oxygen distance for effectivecoordination with MgCl2 but that, unlike phthalate esters, were not removed fromthe support on contact with Al(C2H5)3 and that were unreactive with TiCl4 dur-ing catalyst preparation. It was found (55,62–64) that certain 2,2-disubstituted1,3-dimethoxypropanes met all these criteria. The best performance was obtainedwhen bulky substituents in the 2-position resulted in the diether having a mostprobable conformation (65) with an oxygen–oxygen distance in the range 2.8–3.2A. The successive “generations” of high activity MgCl2-supported catalyst systemsfor PP are summarized below:

(1) MgCl2/TiCl4/ethyl benzoate–AlR3–aromatic ester(2) MgCl2/TiCl4/phthalate ester–AlR3–alkoxysilane(3) MgCl2/TiCl4/diether–AlR3

Catalyst performance has improved considerably with each generation. ThePP yield obtained under typical polymerization conditions (liquid monomer, inthe presence of hydrogen, 70◦C, 1–2 h) has increased from 15–30 kg/g cat. for

526 ZIEGLER–NATTA CATALYSTS Vol. 8

the third-generation ethyl benzoate containing catalysts to 30–80 kg/g cat. forthe fourth-generation phthalate-based catalysts. With the recently developedfifth-generation catalysts containing a diether as internal donor, yields of 80–160 kg/g cat. can be achieved. These different catalysts also display differentkinetic profiles in propylene polymerization. The catalysts containing a dietheras internal donor exhibit very stable activities during polymerization. A lowrate of catalyst decay during polymerization is also obtained with the cata-lyst system MgCl2/TiCl4/phthalate ester–AlR3–alkoxysilane, whereas the systemMgCl2/TiCl4/ethyl benzoate–AlR3–aromatic ester has a very high initial polymer-ization activity but also a high decay rate, limiting the final polymer yield. Therapid decay in activity can at least partially be ascribed to the use of an ester asexternal as well as internal donor, the ester being able to react with titanium–hydrogen bonds formed in chain transfer with hydrogen, generating Ti O bondsinactive for chain propagation (66).

Most recently, a further family of MgCl2-supported catalysts has been devel-oped (67,68), in which the internal donor is a succinate rather than a phthalateester. As is the case with the phthalate-based catalysts, an alkoxysilane is usedas external donor. The essential difference between these catalysts is that thesuccinate-based systems produce PP having much broader MWD (discussed un-der Catalyst/Polymer Relationship).

Mechanistic Aspects. It is well established that effective external donorsnot only increase the isotactic index of the polymer (the proportion of polymer in-soluble in boiling heptane or in xylene at 25◦C), but can also increase in absoluteterms the amount of isotactic polymer formed. This has been demonstrated byKashiwa (69) for the catalyst system MgCl2/TiCl4–Al(C2H5)3. An increase in themolecular weight and stereoregularity of the isotactic fraction was also noted.Similar trends are apparent with catalyst systems of type MgCl2/TiCl4/phthalateester–AlR3–alkoxysilane (70). Kakugo (71) has used elution fractionation todemonstrate that the external donor not only decreases “atactics” formation butalso increases the degree of steric control at isospecific sites. Soga has reportedthat an almost aspecific MgCl2/TiCl3 catalyst, with a very low content of (proba-bly isolated, monomeric) Ti species, could be rendered isospecific by the additionof ethyl benzoate as external donor (72) or by using Cp2Ti(CH3)2 as cocatalyst(73). It was suggested (74) that in both cases aspecific sites having two coordina-tion vacancies could be converted to isospecific sites by blocking one of the twovacancies.

A powerful technique to study the effects of electron donors on site selectivityin Ziegler–Natta catalysts is the determination of the stereoregularity of the firstinsertion step in propylene polymerization. Sacchi and co-workers (60,75) haveinvestigated the effect of Lewis bases on the first-step stereoregularity resultingfrom propylene insertion into a Ti C2H5 bond formed via chain transfer with13C-enriched Al(C2H5)3. First-step stereoregularity is particularly sensitive to thesteric environment of the active center, because the stereospecificity of the firstmonomer insertion is always lower than that of the following propagation steps.For example, with a MgCl2/TiCl4/diisobutyl phthalate catalyst it was found (60)that the mole fraction of erythro (isotactic) placement in the isotactic polymerfraction was 0.67 with no external donor, 0.82 with CH3Si(OC2H5)3, and 0.92with C6H5(OC2H5)3. It could be concluded that the alkoxysilane external donor

Vol. 8 ZIEGLER–NATTA CATALYSTS 527

C C C C C C

C C

C C C C C C

C

C

(erythro) (threo)

Fig. 7. Erythro (isotactic) and threo (syndiotactic) placement resulting from insertion intoa Ti–ethyl unit.

was present in the environment of at least part of the isospecific centers (Fig. 7).Subsequent studies (76,77) indicated that similar considerations apply to dietherdonors.

Recently, significant advances have been made in understanding the funda-mental factors determining the performance of state-of-the-art MgCl2-supportedcatalysts. Studies by Busico and co-workers (78) have shown that the chain irreg-ularities in isotactic PP prepared using heterogeneous catalysts are not randomlydistributed along the chain but are clustered. The chain can therefore contain, inaddition to highly isotactic blocks, sequences that can be attributed to weakly iso-tactic (isotactoid) and to syndiotactic (syndiotactoid) blocks. This implies that theactive site can isomerize very rapidly (during the growth time of a single polymerchain, ie in less than a second) between three different propagating species. Thesame sequences are present, but in different amounts, in both the soluble and theinsoluble fractions. The polymer can therefore be considered to have a stereoblockstructure in which highly isotactic sequences alternate with defective isotactoidand with syndiotactoid sequences. A mechanistic model has been formulated inwhich the relative contributions of these sequences can be related to site transfor-mations involving the presence or absence of steric hindrance in the vicinity of theactive species. 13C NMR studies have indicated (79) the presence of C1-symmetricactive species in MgCl2-supported catalysts, with a mechanism of isotactic prop-agation which is analogous to that for certain C1-symmetric metallocenes, in thesense that propylene insertion at a highly enantioselective site tends to be followedby chain “back-skip” rather than a less regio- and stereoselective insertion whenthe chain is in the coordination position previously occupied by the monomer. Theprobability of chain “back-skip” will increase with decreasing monomer concen-tration, and it has indeed been confirmed that increased polymer isotacticity isobtained at low monomer concentration. It is proposed (78) that a (temporary)loss of steric hindrance from one side of an active species with local C2-symmetry,giving a C1-symmetric species, may result in a transition from highly isospecificto moderately isospecific propagation. Loss of steric hindrance on both sides canlead to syndiospecific propagation in which chain-end control becomes operative.The model is illustrated in Figure 8.

If it is considered that the steric hindrance in the vicinity of the active speciescan result from the presence of a donor molecule, and that the coordination ofsuch a donor is reversible, the above model provides us with an explanation forthe fact that strongly coordinating, stereorigid donors typically give stereoregularpolymers in which the highly isotactic sequences predominate. Several types ofactive species in which the presence of a donor in the vicinity of the active Ti atomis necessary for high isospecificity have been proposed (80), although the exactstructure of the active species is still by no means resolved. Isospecific active

528 ZIEGLER–NATTA CATALYSTS Vol. 8

S2

S1

L2

L1

(b) (c)(a)

Fig. 8. Model of possible active species for (a) highly isotactic, (b) isotactoid, and (c)syndiotactic propagation (78).

species not requiring the presence of a donor for high stereospecificity have alsobeen proposed (81).

In PP production, hydrogen is used as a chain-transfer agent for polymermolecular weight (melt-flow rate) control. The effect of hydrogen (concentration)on polymer molecular weight is dependent on the catalyst system. An advantageof catalysts containing a diether donor, in addition to very high activity, is highsensitivity to hydrogen, so that relatively little hydrogen is required for molecularweight control. This effect can be ascribed to chain transfer after the occasionalsecondary (2,1-) rather than the usual primary (1,2-) insertion, a 2,1-insertionslowing down a subsequent monomer insertion and therefore increasing the prob-ability of chain transfer (82). Reactivation of “dormant” (2,1-inserted) species viachain transfer with hydrogen also explains the frequently observed activating ef-fect of hydrogen in propylene polymerization, giving yields which may be aroundthree times those observed in the complete absence of hydrogen. These conclusionshave been based on the 13C NMR determination of the relative proportions of i-C4H9 and n-C4H9 terminated chains, resulting from chain transfer with hydrogenafter primary and secondary insertion respectively:

Ti CH2 CH(CH3) [CH2 CH(CH3)]nPr + H2 → Ti H + i−C4H9 CH(CH3)[CH2(CH3)]n− 1Pr

Ti CH(CH3) CH2 [CH2 CH(CH3)]nPr + H2 → Ti H + n−C4H9 CH(CH3)[CH2(CH3)]n− 1Pr

Other studies have demonstrated that chain transfer is dependent not onlyon regio- but also on stereoselectivity (48). This is in keeping with the ten-dency that, with catalyst systems of type MgCl2/TiCl4/phthalate ester–AlR3–alkoxysilane, the silanes that give the most stereoregular isotactic polymer alsogive relatively low hydrogen response.

Catalyst/Polymer Relationship. By varying the catalyst composition,and in particular the nature of the electron donors (esters, silanes, diethers)present in the catalyst system, it is possible to control the PP tacticity, molec-ular weight, and MWD to produce a range of polymers having the processing andend-use properties required for very different applications. Ziegler–Natta cat-alysts typically give broader MWDs than are obtained with homogeneous

Vol. 8 ZIEGLER–NATTA CATALYSTS 529

(metallocene) catalysts. This is because Ziegler–Natta catalysts contain a range ofdifferent active centers, each center giving different relative rates of chain propa-gation and chain transfer. Each individual site produces a Schulz–Flory distribu-tion with Mw/Mn = 2 and Mz/Mw = 1.5, and the overall polymer MWD representsa combination of these individual distributions. The dominant effect of active cen-ter distribution has been demonstrated by the use of stopped-flow polymerization(83), where the polymer MWD was shown to be unaffected by the polymeriza-tion time. Stopped-flow polymerization has also been used to determine activesite concentration (C∗) and propagation rate constants, kp. For MgCl2-supportedcatalysts, C∗ values of around 4% (of total Ti present) have been obtained fromstopped-flow experiments (84). C∗ values obtained by other techniques, notably14CO quenching of propylene polymerization, have ranged from 1% or less (85) tomore than 20% (86). Clearly, there are large differences in C∗ values obtained bydifferent groups, but it is consistently found that the major proportion of the Tipresent in Ziegler–Natta catalysts is inactive. The kp values for isospecific activesites are around an order of magnitude greater than those for weakly specific sites(85,86). The value of kp increases significantly in the presence of hydrogen (87),in accordance with the reactivation of “dormant” (2,1-inserted) centers by chaintransfer.

Recent work by Terano and co-workers (88) has shown that, under stopped-flow conditions, hydrogen is only effective as chain-transfer agent when catalystand cocatalyst have been precontacted. These and subsequent (89,90) results indi-cated that effective chain transfer with hydrogen requires the presence of speciesable to promote the dissociation of H2 to atomic hydrogen.

The donors present in the catalyst system play an active role in the formationor modification of isospecific sites, and the polymer MWD depends on the relativecontribution and hydrogen response (ie sensitivity to chain transfer with hydro-gen) of each type of active site. The incorporation of an external donor into thecatalyst system generally leads to an increase in molecular weight, the magnitudeof the MW increase depending on the nature of the donor. The characteristics ofdifferent catalyst systems with regard to PP MWD are as follows (68):

Internal donor External donor Mw/Mn

Diether — 5–5.5Phthalate Alkoxysilane 6.5–8Succinate Alkoxysilane 10–15

It can be seen that the diether-based catalysts give relatively narrow MWD. Anarrow MWD, and relatively low molecular weight, is advantageous in fiber spin-ning applications. In contrast, extrusion of pipes and thick sheets requires highmelt strength, and therefore relatively high molecular weight and broad MWD. Abroad MWD, along with high isotactic stereoregularity, is also beneficial for highcrystallinity and therefore high rigidity. The new succinate-based catalysts enablevery broad MWD PP homopolymers to be produced in a single reactor and are alsoof interest for the production of heterophasic copolymers having an improved bal-ance of stiffness and impact strength, taking into account that the incorporation

530 ZIEGLER–NATTA CATALYSTS Vol. 8

of a rubbery (ethylene/propylene) copolymer phase into a PP homopolymer matrixincreases impact strength but leads at the same time to decreased stiffness.

The relatively narrow PP molecular weight distributions obtained usingdiether-based catalysts can be attributed to the fact that in these systems eventhe most highly stereospecific active sites are not totally regiospecific. A propor-tion of approximately one secondary insertion for every 2000 primary insertionsat highly isospecific sites has been noted for the system MgCl2/TiCl4/diether–AlR3(82). The probability of chain transfer with hydrogen after a secondary insertion issuch that this is sufficient to prevent the formation of very high molecular weightchains, taking into account that the highest molecular weight fraction of the poly-mer is formed on the active species having the highest isospecificity. The broaderMWDs obtained with catalysts containing ester internal donors are likely to be dueto the presence of (some) isospecific active sites having very high regiospecificityand therefore lower hydrogen sensitivity. Such results illustrate the profound ef-fect of catalyst regio- and stereospecificity distribution on both molecular weightcontrol and polymer MWD and properties.

Reactor Granule Technology. As indicated in the section Polymer Par-ticle Growth, particle morphology and porosity are very important features of aZiegler–Natta catalyst used in modern bulk (liquid monomer or gas-phase) poly-merization processes. Under appropriate polymerization conditions, polymer par-ticles can be obtained that have an internal morphology that can range from acompact to a porous structure (91). In what is termed Reactor Granule Technol-ogy (RGT), porous polymer particles can be produced, which can then functionas a microreactor for the polymerization of other monomers within the solid ma-trix. A polypropylene skin encloses the second polymer phase, thereby preventingcoalescence of particles in which the second phase is an amorphous, low-meltingmaterial (92). RGT has been defined as “controlled, reproducible polymerization ofolefinic monomers on an active MgCl2-supported catalyst, to give a growing, spher-ical granule that provides a porous reaction bed within which other monomers canbe introduced and polymerized to form a polyolefin alloy” (93).

Today, RGT is able to provide products ranging from superstiff, high fluid-ity PP homopolymers to stiff/impact or clear/impact heterophasic copolymers andsupersoft alloys, produced using the Catalloy process (31,68). The feasibility ofproducing heterophasic alloys containing up to 70% of multimonomer copolymersarises from the use of a controlled porosity catalyst and the ability to control theporosity of the growing polymer particle during the early stages of polymeriza-tion. Prepolymerization is applied to give the particles sufficient heat capacity towithstand injection into a gas-phase polymerization step.

Several models have been put forward to explain the mechanism of particlegrowth during polymerization. One of the most popular models is the “multigrainmodel,” put forward by Ray and co-workers (94), in which the monomer diffusesinto the catalyst macroparticle and polymerizes on the surface of the microparti-cles within, causing the macroparticle to expand progressively as polymerizationproceeds. An investigation by Kakugo and co-workers (95) of nascent polymermorphology obtained using a TiCl3 catalyst showed that the polymer particlecomprised numerous globules, each of which contained some tens of much smallerprimary particles. Recently, a model for particle growth with MgCl2-supported cat-alysts has been proposed by Cecchin (68,96), who has also provided evidence for

Vol. 8 ZIEGLER–NATTA CATALYSTS 531

Polymer grain Polymersubglobule

Fig. 9. Particle growth model for propylene polymerization with a MgCl2-supported cat-alyst (96).

polymer “subglobule” formation within the growing particle. Again, these subglob-ules originate from clusters of primary crystallites, but the crystallites themselvesare pushed to the surface of each subglobule as the polymer forms. This model,illustrated in Figure 9, explains the fact that, in the preparation of heteropha-sic copolymers via propylene homopolymerization followed by ethylene/propylenecopolymerization, the rubbery ethylene/propylene copolymer is formed at the sur-face of the homopolymer subglobules, gradually filling up the pores within theparticle. Clearly, the higher the porosity of the homopolymer matrix, the greaterthe proportion of (rubbery) copolymer that can be incorporated into the particlewithout running into problems of stickiness if the rubber phase blooms to thesurface. 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 (97).

Polymerization of Other Monomers Using Ziegler–Natta Catalysts

In addition to their widespread use in the production of polyethylene andpolypropylene, Ziegler–Natta catalysts play an important role in the productionof poly-1-butene and are also widely used in the manufacture of synthetic rubberssuch as cis-1,4-polybutadiene and cis-1,4-polyisoprene, the synthetic equivalent ofnatural rubber. Ziegler–Natta catalysts for the manufacture of butadiene rubber,based on titanium, cobalt, nickel, or neodymium systems, are described elsewhere(see BUTADIENE POLYMERS). Isoprene rubber is produced using β-TiCl3 (98), typ-ically prepared by reaction of approximately equimolar quantities of TiCl4 andAl-i-(C4H9)3 in the presence of a small quantity of an ether. Increased catalystactivity can be obtained by incorporation of a sterically hindered phenoxyalu-minum cocatalyst component (99). The latter component also gives increased ac-tivity in propene polymerization using TiCl3 (100); in both cases the improvement

532 ZIEGLER–NATTA CATALYSTS Vol. 8

in catalyst performance can be attributed to selective complexation of the catalystpoison RAlCl2. In addition to aluminum alkyls, poly(N-alkylaluminoxanes) havebeen found to be effective cocatalysts in isoprene polymerization (101). These com-ponents have cage structures (eg [HAlN-i-C3H7]6) in which both Al and N atomsare tetracoordinated (102).

Cis-1,4-polymerization of conjugated dienes requires the presence of two co-ordination vacancies on the transition-metal atom, allowing bidentate coordina-tion of the diene. β-TiCl3 has a fiber-like structure in which the titanium atoms inthe lattice are octahedrally coordinated to six chlorine atoms. The terminal tita-nium atoms are, however, incompletely coordinated and are linked to four or fivechlorine atoms. Alkylation of the tetracoordinated titanium atoms will generatethe double-vacancy species active in isoprene polymerization. Stereospecificity indiene polymerization can change dramatically if one of the coordination vacanciesis blocked by a Lewis base. An interesting illustration of this (103) is that theaddition of an external donor in isoprene polymerization with TiCl4–Al(C2H5)3or MgCl2/TiCl4–Al(C2H5)3 changes the catalyst stereospecificity to give mainlytrans-1,4- rather than cis-1,4-polyisoprene. At the same time, a notable increasein isospecificity in propylene polymerization is observed.

TiCl3-based and MgCl2-supported catalysts have been developed for the pro-duction of poly-1-butene. TiCl3 catalysts are used with dialkylaluminum halidecocatalysts, Al(C2H5)2I giving higher isotacticity than Al(C2H5)2Cl (104). Veryhigh isotacticity has been obtained using TiCl3 in combination with Cp2Ti(CH3)2(105). Much higher polymerization activity, as well as high isotacticity and broadMWD, is obtained using MgCl2-supported catalysts, for example the catalyst sys-tem MgCl2/TiCl4/diisobutyl phthalate–Al(C2H5)3–alkoxysilane (106).

Ziegler–Natta catalysts have also been developed for the polymerization of4-methyl-1-pentene (107) and higher alpha-olefins. Polymerization activity de-creases with increasing steric bulk of the monomer. For example, with the cat-alyst system MgCl2/TiCl4/ethyl benzoate–Al(C2H5)3–ethyl benzoate the relativeactivities in propylene, 1-butene, and 4-methyl-1-pentene polymerization were100:80:15 (108). For catalyst systems of type MgCl2/TiCl4/phthalate ester–AlR3–alkoxysilane, the type of silane required is dependent on the steric bulk of themonomer. An active center having high stereospecificity in propylene polymer-ization may be too sterically hindered for effective polymerization of a bulkiermonomer, propylene/1-butene copolymerization studies having shown (109) thatthe incorporation of 1-butene into the polymer chain decreases with increasing sitestereospecificity. This phenomenon is also illustrated by the fact that nonbulkyalkoxysilanes such as (CH3)3SiOCH3 are effective donors in 4-methyl-1-pentenepolymerization (110), whereas such donors are relatively ineffective in propylenepolymerization.

Concluding Remarks

The dominant position of Ziegler–Natta catalysts in the manufacture of poly-olefins, in particular PP, is likely to continue for a considerable length of time,despite the many developments taking place in the field of metallocene and othersingle-site catalysis. Indeed, the range of polymer types and grades is so varied

Vol. 8 ZIEGLER–NATTA CATALYSTS 533

that there is ample scope for further development and implementation of bothZiegler–Natta and single-site catalysts.

It will be clear that the composition and characteristics of a Ziegler–Nattacatalyst must be tailored such that the required polymer molecular structure andproperties are obtained. Different catalysts are required for different polymerapplications, and the recent development and implementation of MgCl2-supportedcatalysts containing diether and succinate donors, for the production of narrowand broad MWD PP respectively, illustrates the ongoing activity in Ziegler–Nattacatalyst research. The ability to control catalyst particle size, morphology, andporosity has allowed the development of advanced and versatile polymerizationprocess technologies, so that the characteristics of the catalyst can be tuned toboth process and product requirements.

Ziegler–Natta catalysts are complex systems and are still by no means fullyunderstood, but significant advances in basic understanding have recently beenmade. This will continue, with both experimental and molecular modeling studiesproviding additional mechanistic insight, which in turn can be applied in thefurther development and implementation of these catalysts.

BIBLIOGRAPHY

1. G. Wilke, in G. Fink, R. Mulhaupt, and H. H. Brintzinger, eds., Ziegler Catalysts. Re-cent Scientific Innovations and Technological Improvements, Springer-Verlag, Berlin,1995, p. 1.

2. Ger. Pat. 973626 (Filed Nov. 17, 1953), K. Ziegler, H. Breil, E. Holzkamp, and H.Martin; Chem. Abstr. 54, 14794b (1960).

3. K. Ziegler, E. Holzkamp, H. Breil, and H. Martin, Angew. Chem. 67, 541 (1955).4. Ital. Pat. 535712 (Filed June 8, 1954), G. Natta (to Montecatini).5. G. Natta, J. Polym. Sci. 16, 143 (1955).6. G. Natta, P. Pino, P. Corradini, F. Danusso, E. Mantica, E. Mazzanti, and G. Moraglio,

J. Am. Chem. Soc. 77, 1708 (1955).7. P. Pino and R. Mulhaupt, Angew. Chem., Int. Ed. Engl. 19, 857 (1980).8. P. Pino and G. Moretti, Polymer 28, 683 (1987).9. P. Corradini, Macromol. Symp. 89, 1 (1995).

10. G. Natta, P. Corradini, and G. Allegra, J. Polym. Sci. 51, 399 (1961).11. J. Boor, Ziegler–Natta Catalysts and Polymerizations, Academic Press, Inc., New York,

1979.12. G. Guidetti, R. Zannetti, D. Ajo, A. Marigo, and M. Vidali, Eur. Polym. J. 16, 1007

(1980).13. U. S. Pat. 2692257 (Filed Apr. 28, 1951), A. Zletz (to Standard Oil of Indiana).14. Belg. Pat. 530617 (Filed Jan. 27, 1953), J. P. Hogan and R. L. Banks (to Phillips

Petroleum Co.).15. H. R. Sailors and J. P. Hogan, J. Macromol. Sci., Chem. A 15, 1377 (1981).16. M. P. McDaniel, in D. D. Eley, H. Pines, and P. B. Weisz, eds., Advances in Catalysis,

Vol. 33, Academic Press, Inc., New York, 1985, p. 47.17. H. Martin, in G. Fink, R. Mulhaupt, and H. H. Brintzinger, eds., Ziegler Catalysts. Re-

cent Scientific Innovations and Technological Improvements, Springer-Verlag, Berlin,1995, p. 15.

18. E. P. Moore, Polypropylene Handbook. Polymerization, Characterization, Properties,Processing, Applications, Hanser Publishers, Munich, 1996.

534 ZIEGLER–NATTA CATALYSTS Vol. 8

19. U.S. Pat. 4376851 (Filed Jan. 11, 1956), J. P. Hogan and R. L. Banks (to PhillipsPetroleum Co.).

20. U.S. Pat. 4210738 (Filed Mar. 21, 1972), J. P. Hermans and P. Henrioulle (to Solvay).21. B. L. Goodall and S. van der Vens, Polypropylene and Other Polyolefins. Polymerization

and Characterization, Elsevier, Amsterdam, 1990, Chapt. “1”.22. A. Bernard and P. Fiasse, in T. Keii and K. Soga, eds., Catalytic Olefin Polymerization,

Elsevier, Amsterdam, 1990, p. 405.23. E. J. Arlman and P. Cossee, J. Catal. 3, 99 (1964).24. P. Corradini, V. Busico, and G. Guerra, Comprehensive Polymer Science, Vol. 4, Perg-

amon Press, Oxford, U.K., 1988, p. 29.25. A. Zambelli, M. C. Sacchi, P. Locatelli, and G. Zannoni, Macromolecules 15, 211

(1982).26. I. Tritto, M. C. Sacchi, and P. Locatelli, Makromol. Chem. 187, 2145 (1986).27. G. Natta, I. Pasquon, A. Zambelli, and G. Gatti, J. Polym. Sci. 51, 387 (1961).28. Y. Doi, S. Suzuki, and K. Soga, Makromol. Chem., Rapid Commun. 6, 639 (1985).29. A. Zambelli, I. Sessa, F. Grisi, R. Fusco, and P. Accomazzi, Macromol. Rapid Commun.

22, 297 (2001).30. J. A. Ewen, R. L. Jones, A. Razavi, and J. D. Ferrara, J. Am. Chem. Soc. 110, 6255

(1988).31. T. Simonazzi and U. Giannini, Gazz. Chim. Ital. 124, 533 (1994).32. U. Giannini, Makromol. Chem., Suppl. 5, 216 (1981).33. K.-Y. Choi and W. H. Ray, J. Macromol. Sci., Rev. Macromol. Chem. Phys. C 25, 1

(1985).34. Eur. Pat. 4647 (Filed Mar. 30, 1979), G. L. Goerke, B. E. Wagner, and F. J. Karol (to

Union Carbide Corp.).35. L. L. Bohm, D. Bilda, W. Breuers, H. F. Enderle, and R. Lecht, in G. Fink, R. Mulhaupt,

and H. H. Brintzinger, eds., Ziegler Catalysts. Recent Scientific Innovations and Tech-nological Improvements, Springer-Verlag, Berlin, 1995, p. 387.

36. F. J. Karol, K. J. Cann, and B. E. Wagner, in W. Kaminsky and H. Sinn, eds., TransitionMetals and Organometallics as Catalysts for Olefin Polymerization, Springer-Verlag,Berlin, 1988, p. 149.

37. H. L. Hsieh, M. P. McDaniel, J. L. Martin, P. D. Smith, and D. R. Fahey, in R. B.Seymour and T. Cheng, eds., Advances in Polyolefins, Plenum Press, New York, 1985,p. 153.

38. P. Galli, L. Luciani, and G. Cecchin, Angew. Makromol. Chem. 94, 63 (1981).39. P. C. Barbe, G. Cecchin, and L. Noristi, Adv. Polym. Sci. 81, 1 (1987).40. P. Galli, P. C. Barbe, and L. Noristi, Angew. Makromol. Chem. 120, 73 (1984).41. P. Galli, Macromol. Symp. 78, 269 (1994); Macromol. Symp. 112, 1 (1996).42. Eur. Pat 45977 (Filed Aug. 13, 1981), S. Parodi, R. Nocci, U. Giannini, P. C. Barbe, and

U. Scata (to Montedison).43. J. V. Seppala, M. Harkonen and L. Luciani, Makromol. Chem. 190, 2535 (1989).44. M. Harkonen, J. V. Seppala, and T. Vaananen, in T. Keii and K. Soga, eds., Catalytic

Olefin Polymerization, Elsevier, Amsterdam, 1990, p. 87.45. A. Proto, L. Oliva, C. Pellecchia, A. J. Sivak, and L. A. Cullo, Macromolecules 23, 2904

(1990).46. T. Okano, K. Chida, H. Furuhashi, A. Nakano, and S. Ukei, in T. Keii and K. Soga,

eds., Catalytic Olefin Polymerization, Elsevier, Amsterdam, 1990, p. 177.47. Eur. Pat. 350170 (Filed June 13, 1989), N. Ishimaru, M. Kioka, and A. Toyota (to

Mitsui).48. J. C. Chadwick, G. M. M. van Kessel, and O. Sudmeijer, Macromol. Chem. Phys. 196,

1431 (1995).49. J. C. Chadwick, Macromol. Symp. 173, 21 (2001).

Vol. 8 ZIEGLER–NATTA CATALYSTS 535

50. Eur. Pat. 841348 (Filed Nov. 3, 1997), S. Ikai, H. Ikeuchi, H. Satoh, T. Inoue, and H.Sano (to Grand Polymer).

51. S. Yao and Y. Tanaka, Macromol. Theory Simul. 10, 850 (2001).52. P. Corradini, V. Busico, and G. Guerra, in W. Kaminsky and H. Sinn, eds., Transition

Metals and Organometallics as Catalysts for Olefin Polymerization, Springer-Verlag,Berlin, 1988, p. 337.

53. V. Busico, P. Corradini, L. De Martino, A. Proto, V. Savino, and E. Albizzati, Makromol.Chem. 186, 1279 (1985).

54. M. Terano, T. Kataoka, and T. Keii, Makromol. Chem. 188, 1477 (1987).55. E. Albizzati, U. Giannini, G. Morini, C. A. Smith, and R. C. Zeigler, in G. Fink, R.

Mulhaupt, and H. H. Brintzinger, eds., Ziegler Catalysts. Recent Scientific Innovationsand Technological Improvements, Springer-Verlag, Berlin, 1995, p. 413.

56. V. Busico, P. Corradini, L. De Martino, A. Proto, and E. Albizzati, Makromol. Chem.187, 1115 (1986).

57. L. Noristi, P. C. Barbe, and G. Baruzzi, Makromol. Chem. 192, 1115 (1991).58. M. C. Sacchi, I. Tritto, C. Shan, R. Mendichi, G. Zannoni, and L. Noristi, Macro-

molecules 24, 6823 (1991).59. R. Spitz, C. Bobichon, and A. Guyot, Makromol. Chem. 190, 707 (1989).60. M. C. Sacchi, F. Forlini, I. Tritto, R. Mendichi, G. Zannoni, and L. Noristi, Macro-

molecules 25, 5914 (1992).61. M. Kioka and N. Kashiwa, J. Mol. Catal. 82, 11 (1993).62. Eur. Pat. 361494 (Filed Sept. 29, 1989), E. Albizzati, P. C. Barbe, L. Noristi, R. Scor-

damaglia, L. Barino, U. Giannini, and G. Morini (to Himont).63. Eur. Pat. 728724 (Filed Feb. 21, 1996), G. Morini and A. Cristofori (to Montell).64. E. Albizzati, U. Giannini, G. Morini, M. Galimberti, L. Barino, and R. Scordamaglia,

Macromol. Symp. 89, 73 (1995).65. L. Barino and R. Scordamaglia, Macromol. Symp. 89, 101 (1995).66. E. Albizzati, M. Galimberti, U. Giannini, and G. Morini, Makromol. Chem., Macromol.

Suppl. 48/49, 223 (1991).67. Int. Pat. WO 00/63261 (Filed Apr. 12, 2000), G. Morini, G. Balbontin, Y. Gulevich, H.

Duijghuisen, R. Kelder, P. A. Klusener, and F. Korndorffer (to Basell).68. G. Cecchin, G. Morini, and A. Pelliconi, Macromol. Symp. 173, 195 (2001).69. N. Kashiwa, J. Yoshitake, and A. Toyota, Polym. Bull. (Berlin) 19, 333 (1988).70. J. C. Chadwick, in G. Fink, R. Mulhaupt, and H. H. Brintzinger, eds., Ziegler Cata-

lysts. Recent Scientific Innovations and Technological Improvements, Springer-Verlag,Berlin, 1995, p. 427.

71. M. Kakugo, T. Miyatake, Y. Naito, and K. Mizunuma, Macromolecules 21, 314 (1988).72. K. Soga, J. R. Park, T. Shoino, and N. Kashiwa, Macromol. Chem., Rapid Commun.

11, 117 (1990).73. K. Soga, J. R. Park, H. Uchino, T. Uozumi, and T. Shiono, Macromolecules 22, 3824

(1989).74. K. Soga, Makromol. Chem., Macromol. Symp. 66, 43 (1993).75. M. C. Sacchi, C. Shan, P. Locatelli, and I. Tritto, Macromolecules 23, 383 (1990).76. M. C. Sacchi, F. Forlini, I. Tritto, P. Locatelli, G. Morini, G. Baruzzi, and E. Albizzati,

Macromol. Symp. 89, 91 (1995).77. G. Morini, E. Albizzati, G. Balbontin, I. Mingozzi, M. C. Sacchi, F. Forlini, and I. Tritto,

Macromolecules 29, 5770 (1996).78. V. Busico, R. Cipullo, G. Monaco, G. Talarico, M. Vacatello, J. C. Chadwick, A. L. Segre,

and O. Sudmeijer, Macromolecules 32, 4173 (1999).79. V. Busico, R. Cipullo, G. Talarico, A. L. Segre, and J. C. Chadwick, Macromolecules

30, 4786 (1997).80. L. Barino and R. Scordamaglia, Makromol. Theory Simul. 7, 407 (1998).

536 ZIEGLER–NATTA CATALYSTS Vol. 8

81. M. Boero, M. Parrinello, S. Huffer, and H. Weiss, J. Am. Chem. Soc. 122, 501 (2000).82. J. C. Chadwick, G. Morini, E. Albizzati, G. Balbontin, I. Mingozzi, A. Cristofori, O.

Sudmeijer, and G. M. M. van Kessel, Macromol. Chem. Phys. 197, 2501 (1996).83. T. Keii, M. Terano, K. Kimura, and K. Ishii, Makromol. Chem., Rapid Commun. 8,

583 (1987).84. B. Liu, H. Matsuoka, and M. Terano, Macromol. Rapid Commun. 22, 1 (2001).85. G. D. Bukatov and V. A. Zakharov, Macromol. Chem. Phys. 202, 2003 (2001).86. P. J. Tait, G. H. Zohuri, A. M. Kells, and I. D. McKenzie, in G. Fink, R. Mulhaupt, and H.

H. Brintzinger, eds., Ziegler Catalysts. Recent Scientific Innovations and TechnologicalImprovements, Springer-Verlag, Berlin, 1995, p. 343.

87. G. D. Bukatov, V. S. Goncharov, and V. A. Zakharov, Macromol. Chem. Phys. 196,1751 (1995).

88. H. Mori, K. Tashino, and M. Terano, Macromol. Rapid Commun. 16, 651 (1995).89. H. Mori, K. Tashino, and M. Terano, Macromol. Chem. Phys. 197, 895 (1996).90. H. Mori, T. Iizuka, K. Tashino, and M. Terano, Macromol. Chem. Phys. 198, 2499

(1997).91. P. Galli, Prog. Polym. Sci. 19, 959 (1994).92. J. C. Chadwick, Encyclopedia of Materials: Science and Technology, Elsevier, 2001, p.

7703.93. P. Galli and J. C. Haylock, Makromol. Chem., Macromol. Symp. 63, 19 (1992).94. R. A. Hutchinson, C. M. Chen, and W. H. Ray, J. Appl. Polym. Sci. 44, 1389 (1992).95. M. Kakugo, H. Sadatoshi, M. Yokoyama, and K. Kojima, Macromolecules 22, 547

(1989).96. G. Cecchin, E. Marchetti, and G. Baruzzi, Macromol. Chem. Phys. 202, 1987 (2001).97. J. T. M. Pater, G. Weickert, J. Loos, and W. P. M. van Swaaij, Chem. Eng. Sci. 56, 1

(2001).98. E. Schoenberg, H. A. Marsh, S. J. Walters, and W. M. Saltman, Rubber Chem. Technol.

52, 526 (1979).99. Eur. Pat. 107871 (Filed Sept. 30, 1983), J. C. Chadwick and B. L. Goodall (to Shell).

100. B. L. Goodall, in W. Kaminsky and H. Sinn, eds., Transition Metals andOrganometallics as Catalysts for Olefin Polymerization, Springer-Verlag, Berlin, 1988,p. 361.

101. A. Mazzei, S. Cucinella, and W. Marconi, Makromol. Chem. 122, 168 (1969).102. A. Balducci, M. Bruzzone, S. Cucinella, and A. Mazzei, Rubber Chem. Technol. 48,

736 (1975).103. K. Soga, T. Sano, K. Yamamoto, and T. Shiono, Chem. Lett. (Japan) 425 (1982).104. S. van der Ven, Polypropylene and other Polyolefins. Polymerization and Characteri-

zation, Elsevier, Amsterdam, 1990, Chapt. 11.105. K. Soga and H. Yanagihara, Makromol. Chem. 189, 2839 (1988).106. U. S. Pat. 6306996 (Filed Mar. 4, 1999), G. Cecchin, G. Collina, and M. Covezzi (to

Basell).107. L. C. Lopez, G. L. Wilkes, P. M. Stricklen, and S. A. White, J. Macromol. Sci., Rev.

Macromol. Chem. Phys. C 32, 301 (1992).108. N. Kashiwa and J. Yoshitake, Polym. Bull. (Berlin) 11, 485 (1984).109. M. C. Sacchi, Z.-Q. Fan, F. Forlini, I. Tritto, and P. Locatelli, Macromol. Chem. Phys.

195, 2805 (1994).110. U. S. Pat. 4659792 (Filed Dec. 23, 1985), N. Kashiwa and K. Fukui (to Mitsui).

JOHN C. CHADWICK

Dutch Polymer Institute,Eindhoven University of Technology