+ All Categories
Home > Documents > Design of stereoselective Ziegler–Natta propene polymerization … · Design of stereoselective...

Design of stereoselective Ziegler–Natta propene polymerization … · Design of stereoselective...

Date post: 12-Feb-2021
Category:
Upload: others
View: 6 times
Download: 0 times
Share this document with a friend
6
Design of stereoselective Ziegler–Natta propene polymerization catalysts Vincenzo Busico*, Roberta Cipullo, Roberta Pellecchia, Sara Ronca, Giuseppina Roviello, and Giovanni Talarico Dipartimento di Chimica, Universita ` di Napoli Federico II, Via Cintia, 80126 Naples, Italy Edited by Tobin J. Marks, Northwestern University, Evanston, IL, and approved July 25, 2006 (received for review April 14, 2006) After five decades of largely serendipitous (albeit formidable) progress, catalyst design in Ziegler–Natta olefin polymerization, i.e., the rational implementation of new active species to target predetermined polyolefin architectures, has ultimately become a realistic ambition, thanks to a much deeper fundamental under- standing and major advances in the tools of computational chem- istry. In this article, we discuss, as a case history, a unique class of stereorigid C2 -symmetric bis(phenoxy-amine)Zr(IV) catalysts with controlled kinetic behavior. A large variety of polypropylene mi- crostructures have been obtained with these catalysts by modu- lating the steric demand of one key substituent, without altering the nature and symmetry of the ancillary ligand framework, under the guidance of computer modeling. This unusual achievement is relevant per se and for the perspective implications in catalyst discovery. enantioselectivity olefin polymerization polypropylene Z iegler–Natta olefin polymerization is probably the most effective and atom-economical large-volume industrial chemical process (1–3). The catalytic cycle (Fig. 1 for ethene) is disarmingly simple and entails merely the insertion of a mono- mer molecule into a M-C bond (where M is a coordinatively unsaturated transition metal center that can be isolated or reside at a suitable surface). What makes Ziegler–Natta catalysts ‘‘unique and marvelous’’ (4) is that several thousand cycles may close in 1 s under mild conditions before chain transfer (usually, a -H elimination event; Fig. 1), an individual M can produce several thousand chains before it ultimately deactivates, and, last but not least, the process is amenable to thorough chemocontrol, regiocontrol, and stereocontrol (1–3). The enantioselectivity in the insertion of propene, in partic- ular, is truly amazing. Although neither isotactic nor syndiotactic polypropylene are chiral molecules (Fig. 2) (5), each individual monomer insertion implies a chiral recognition, the catalytic species selecting one of the two enantiofaces of propene in the favored regiochemistry with an enantiomeric excess that, in several reported cases (6, 7), is 99.8% (an unrivaled perfor- mance for a substrate with no functional groups other than an olefinic double bond). More than in these impressive figures, though, the real fascina- tion and challenge of modern Ziegler–Natta chemistry are in the possibility of implementing catalysts for specific homopolymer or copolymer architectures. In particular, the ability to enchain pro- pene, alone or in combination with other monomers (e.g., ethene), with more or less precisely (pre)determined chemoselectivity, regioselectivity, and enantioselectivity, and the dramatic effect of chain microstructure on the physical properties of the materials (ranging from ultra-rigid thermosets to elastomers via all conceiv- able grades of thermoplastic elastomers), is a key to understanding why polypropylene production continues to grow exponentially, faster than for any other large volume resins (e.g., other polyolefins, polyesters, polyamides, polycarbonates, polyvinylchloride, etc.), and to some extent at their expense (2). Like most scientific breakthroughs, the discovery of Ziegler– Natta catalysts has been largely serendipitous since the very beginning. Without the accidental contamination of an auto- clave with colloidal nickel in Karl Ziegler’s laboratory in 1953 (4, 8), the fortuitous chlorination of MgO to MgCl 2 by TiCl 4 (late 1960s) (1), or the partial hydrolysis of AlMe 3 in a broken NMR tube to yield methylalumoxane (mid-1970s) (9), each triggering formidable chains of deductions and consequent actions, catalyst evolution and diversification would have occurred at a much slower pace. The sound argument that the chance of a lucky find is roughly proportional to the number of attempts has now been set at the foundation of catalyst discovery program by high- throughput screening technologies (10), which have already led to industrially relevant achievements (11, 12). In this scenario, catalyst design (meaning the rational imple- mentation of active organometallic species for tailored applica- tions) has been, euphemistically, marginal until now. On the other hand, the hundreds of metallocene (7, 13) and nonmetal- locene (14) catalyst structures that have been disclosed and tested in the last two decades, and the thorough theoretical studies on realistic models thereof (6, 7), highlighted a number of basic principles relating catalyst structure and performance. These can now be used successfully, if not (yet) for attempts of de novo design, certainly for catalyst fine-tuning. In this article, we illustrate, with a case history based on the current research of our own laboratory, how wide the margins of useful improve- ment can be. Before that, though, it may be useful to recall a few important notions of Ziegler–Natta stereochemistry. Stereoselectivity of Propene Polymerization Catalysts: Preliminary Remarks Most Ziegler–Natta propene polymerization catalysts are highly regioselective in favor of 1,2 (primary) insertion [M-R Author contributions: V.B. designed research; R.C., R.P., S.R., G.R., and G.T. performed research; S.R. contributed new reagentsanalytic tools; V.B., R.C., and G.T. analyzed data; and V.B. wrote the paper. The authors declare no conflict of interest. This article is a PNAS direct submission. Abbreviation: QM, quantum mechanics. Data deposition: The atomic coordinates have been deposited in the Cambridge Structural Database (CSD), Cambridge Crystallographic Data Centre, Cambridge CB2 1EZ, United Kingdom (CSD reference nos. 603920 and 603921). *To whom correspondence should be addressed. E-mail: [email protected]. © 2006 by The National Academy of Sciences of the USA Fig. 1. The catalytic cycle of Ziegler–Natta olefin polymerization. www.pnas.orgcgidoi10.1073pnas.0602856103 PNAS October 17, 2006 vol. 103 no. 42 15321–15326 CHEMISTRY SPECIAL FEATURE Downloaded by guest on July 3, 2021
Transcript
  • Design of stereoselective Ziegler–Natta propenepolymerization catalystsVincenzo Busico*, Roberta Cipullo, Roberta Pellecchia, Sara Ronca, Giuseppina Roviello, and Giovanni Talarico

    Dipartimento di Chimica, Università di Napoli Federico II, Via Cintia, 80126 Naples, Italy

    Edited by Tobin J. Marks, Northwestern University, Evanston, IL, and approved July 25, 2006 (received for review April 14, 2006)

    After five decades of largely serendipitous (albeit formidable)progress, catalyst design in Ziegler–Natta olefin polymerization,i.e., the rational implementation of new active species to targetpredetermined polyolefin architectures, has ultimately become arealistic ambition, thanks to a much deeper fundamental under-standing and major advances in the tools of computational chem-istry. In this article, we discuss, as a case history, a unique class ofstereorigid C2-symmetric bis(phenoxy-amine)Zr(IV) catalysts withcontrolled kinetic behavior. A large variety of polypropylene mi-crostructures have been obtained with these catalysts by modu-lating the steric demand of one key substituent, without alteringthe nature and symmetry of the ancillary ligand framework, underthe guidance of computer modeling. This unusual achievement isrelevant per se and for the perspective implications in catalystdiscovery.

    enantioselectivity � olefin polymerization � polypropylene

    Z iegler–Natta olefin polymerization is probably the mosteffective and atom-economical large-volume industrialchemical process (1–3). The catalytic cycle (Fig. 1 for ethene) isdisarmingly simple and entails merely the insertion of a mono-mer molecule into a M-C � bond (where M is a coordinativelyunsaturated transition metal center that can be isolated or resideat a suitable surface). What makes Ziegler–Natta catalysts‘‘unique and marvelous’’ (4) is that several thousand cycles mayclose in �1 s under mild conditions before chain transfer(usually, a �-H elimination event; Fig. 1), an individual M canproduce several thousand chains before it ultimately deactivates,and, last but not least, the process is amenable to thoroughchemocontrol, regiocontrol, and stereocontrol (1–3).

    The enantioselectivity in the insertion of propene, in partic-ular, is truly amazing. Although neither isotactic nor syndiotacticpolypropylene are chiral molecules (Fig. 2) (5), each individualmonomer insertion implies a chiral recognition, the catalyticspecies selecting one of the two enantiofaces of propene in thefavored regiochemistry with an enantiomeric excess that, inseveral reported cases (6, 7), is �99.8% (an unrivaled perfor-mance for a substrate with no functional groups other than anolefinic double bond).

    More than in these impressive figures, though, the real fascina-tion and challenge of modern Ziegler–Natta chemistry are in thepossibility of implementing catalysts for specific homopolymer orcopolymer architectures. In particular, the ability to enchain pro-pene, alone or in combination with other monomers (e.g., ethene),with more or less precisely (pre)determined chemoselectivity,regioselectivity, and enantioselectivity, and the dramatic effect ofchain microstructure on the physical properties of the materials(ranging from ultra-rigid thermosets to elastomers via all conceiv-able grades of thermoplastic elastomers), is a key to understandingwhy polypropylene production continues to grow exponentially,faster than for any other large volume resins (e.g., other polyolefins,polyesters, polyamides, polycarbonates, polyvinylchloride, etc.), andto some extent at their expense (2).

    Like most scientific breakthroughs, the discovery of Ziegler–Natta catalysts has been largely serendipitous since the verybeginning. Without the accidental contamination of an auto-

    clave with colloidal nickel in Karl Ziegler’s laboratory in 1953 (4,8), the fortuitous chlorination of MgO to MgCl2 by TiCl4 (late1960s) (1), or the partial hydrolysis of AlMe3 in a broken NMRtube to yield methylalumoxane (mid-1970s) (9), each triggeringformidable chains of deductions and consequent actions, catalystevolution and diversification would have occurred at a muchslower pace. The sound argument that the chance of a lucky findis roughly proportional to the number of attempts has now beenset at the foundation of catalyst discovery program by high-throughput screening technologies (10), which have already ledto industrially relevant achievements (11, 12).

    In this scenario, catalyst design (meaning the rational imple-mentation of active organometallic species for tailored applica-tions) has been, euphemistically, marginal until now. On theother hand, the hundreds of metallocene (7, 13) and nonmetal-locene (14) catalyst structures that have been disclosed andtested in the last two decades, and the thorough theoreticalstudies on realistic models thereof (6, 7), highlighted a numberof basic principles relating catalyst structure and performance.These can now be used successfully, if not (yet) for attempts ofde novo design, certainly for catalyst fine-tuning. In this article,we illustrate, with a case history based on the current researchof our own laboratory, how wide the margins of useful improve-ment can be. Before that, though, it may be useful to recall a fewimportant notions of Ziegler–Natta stereochemistry.

    Stereoselectivity of Propene Polymerization Catalysts:Preliminary RemarksMost Ziegler–Natta propene polymerization catalysts are highlyregioselective in favor of 1,2 (primary) insertion [M-R �

    Author contributions: V.B. designed research; R.C., R.P., S.R., G.R., and G.T. performedresearch; S.R. contributed new reagents�analytic tools; V.B., R.C., and G.T. analyzed data;and V.B. wrote the paper.

    The authors declare no conflict of interest.

    This article is a PNAS direct submission.

    Abbreviation: QM, quantum mechanics.

    Data deposition: The atomic coordinates have been deposited in the Cambridge StructuralDatabase (CSD), Cambridge Crystallographic Data Centre, Cambridge CB2 1EZ, UnitedKingdom (CSD reference nos. 603920 and 603921).

    *To whom correspondence should be addressed. E-mail: [email protected].

    © 2006 by The National Academy of Sciences of the USA

    Fig. 1. The catalytic cycle of Ziegler–Natta olefin polymerization.

    www.pnas.org�cgi�doi�10.1073�pnas.0602856103 PNAS � October 17, 2006 � vol. 103 � no. 42 � 15321–15326

    CHEM

    ISTR

    YSP

    ECIA

    LFE

    ATU

    RE

    Dow

    nloa

    ded

    by g

    uest

    on

    July

    3, 2

    021

  • CH2�CH(Me)�M-CH2-CH(Me)(R)], because of electronicand�or steric effects (6). For the insertion to also be enantio-selective, at least one element of chirality needs to combine withre or si propene coordination (Fig. 3) and give rise to appreciablefree energy differences between the competing diastereomerictransition states. This element can be the configuration of thegrowing polypropylene chain (and in particular of the asymmet-ric C in the last-inserted monomeric unit) and�or the possiblechirotopicity of the active site(s) (6, 7).

    The two cases are commonly referred to as ‘‘chain-end con-trol’’ and ‘‘site control,’’ respectively (6, 7). In the latter, theenantiodiscrimination is traceable to direct nonbonded contactsbetween the incoming monomer and the ancillary ligand(s), or,much more frequently, it is mediated by the growing chain,sterically constrained in a chiral conformation favoring propeneinsertion with the enantioface that directs the methyl substituentanti to the first chain C-C bond (Fig. 4; P � polymer chain) (6,7, 15).

    Although, in principle, all said elements can have nonnegli-gible effects, in general one is largely predominant, and theconfiguration of all known stereoregular polypropylenes ob-tained so far with single-center catalytic species, determined bymeans of high-resolution 13C NMR in terms of normalizedstereosequence distributions up to the heptad�nonad level (6),has been well reproduced in terms of chain propagation modelsbased on the limiting approximation of pure chain-end or sitecontrol, the latter in a more or less elaborated version dependingon the chirotopicity of the active sites (6). Without going intodetails, let us define the following approximated stochasticmatrix of chain propagation states:

    The matrix formulation assumes complete regioselectivityand (up to) first-order Markov configurational statistics. Therows are addressed to the configuration of the last-insertedmonomeric unit, and the columns are addressed to that of themonomeric unit to be generated; therefore, � and � are theprobabilities that a monomeric unit of R or S configuration,respectively, is followed by a new one of R configuration. Pure

    site control requires � � �, whereas � � (1 � �) correspondsto pure chain-end control (6). Matrix multiplication codes forthe statistical analysis of the 13C NMR stereosequence distri-bution, ending up with the best-fit values of � and �, have beendescribed (6).

    Understandably, symmetry is a key issue for catalysts withchirotopic sites. The evolution of metallocenes teaches thatC2-symmetric active species with homotopic sites tend to beisotactic-selective, whereas Cs-symmetric ones with enantiotopicsites are often syndiotactic-selective (6, 7, 13). On the otherhand, recent results in the fast-growing area of nonmetallocenecatalysts (14), many of which are easier to make and amenableto structural amplification with a parallel synthetic approach (10,11), demonstrate that the relationship between symmetry andstereoselectivity can be far less obvious (16).

    Results and DiscussionIn the last few years, we have undertaken a systematic study(17–19) of Zr(IV) catalysts bearing a stereorigid tetradentatebis(phenoxy-amine) ancillary ligand (Fig. 5), originally intro-duced by Kol and coworkers (20). We had been attracted by thisligand framework because it closely mimics the coordinationenvironment of the surface Ti atoms in the classical heteroge-neous Ziegler–Natta systems for the industrial production ofisotactic polypropylene (1, 4, 6). In the following, we shall seehow a simple modulation in the steric demand of one keysubstituent (R1 in Fig. 5) without altering the C2 symmetry of thecomplex makes it possible to play with the mechanisms of stereoregulation almost at will and to achieve polypropylenes withdramatically different microstructures. We shall also discuss theconcurrent effect of this fine-tuning on the propensity to chaintransfer, which opened the door to the first highly isotacticpropene block copolymerizations.

    The activation of the neutral (ONNO)ZrBn2 precursors of Fig.5 to [(ONNO)ZrR]� active cations can be carried out withmethylalumoxane, or proper boranes or borate salts similarly towhat is done with metallocenes (21). Solution NMR studies onthe resulting ion couples indicate that the tetradentate coordi-

    Fig. 2. Sawhorse representations of an isotactic (Upper) and syndiotactic(Lower) poly(1-alkene) chain in the (hypothetical) all-trans conformation.Consecutive monomeric units with equal or opposite relative configurationsat the stereogenic tertiary C atoms are denoted as meso (m) and racemo (r)diads, respectively, according to a nomenclature originally introduced byBovey (5).

    Fig. 3. The chirality of coordination of propene to a metal center.

    Fig. 4. Schematic representation of the ‘‘growing chain orientation mech-anism of stereocontrol,’’ originally proposed by Corradini et al. (15).

    R S

    R α 1-α

    S β 1-β

    15322 � www.pnas.org�cgi�doi�10.1073�pnas.0602856103 Busico et al.

    Dow

    nloa

    ded

    by g

    uest

    on

    July

    3, 2

    021

  • nation and the C2-symmetric trans(O,O), cis(N,N) configurationis retained.

    An extensive quantum mechanics (QM) computationalscreening of the active species was aimed at predicting theirregioselectivity, enantioselectivity, and molecular mass capabil-ity in propene polymerization (for technical details see Materialsand Methods and Supporting Text, which is published as support-ing information on the PNAS web site). The results led us toconclude that in general propene insertion with 1,2 regiochem-istry is dominant, for reasons that have been described (22). Theyalso revealed that substituent R2, far from the active pocket, hasvirtually no effect on the catalytic behavior, whereas the stericbulk of R1, more to the catalyst front, is the key to control at thesame time the enantioselectivity of 1,2 insertion and the pro-pensity of the growing chain undergo �-H elimination to themonomer, which is by far the main chain transfer pathway in theabsence of Al-trialkyls (17–19, 23). It is important to note thatthe above is at odds with what is known for C2-symmetricansa-metallocene cations, whose enantioselectivity and chaintransfer properties are independently affected by different sub-stituents (7). According to our computations (Table 1), site-controlled enantiodiscrimination in 1,2 propene insertion ap-pears for R1 � tert-butyl, but for a high catalyst isotacticselectivity and molecular mass capability an even higher stericdemand is needed to enforce the growing chain orientationmechanism (Figs. 4 and 6 Left) and destabilize the six-centertransition state of �-H elimination to propene (Fig. 6 Right) (24,25). On the other hand, the calculations indicate that when R1is too bulky (e.g., R1 � trityl) direct interference occurs betweenthe same R1 and the methyl group of a monomer moleculeinserting with the correct regiochemistry and the enantiofacefavored by the chain, which is detrimental to the enantioselec-tivity, and also, less intuitively, to the regioselectivity.

    When we started our work, two (ONNO)ZrBn2 complexes,with R1 (� R2) � methyl (1) and tert-butyl (3), had been reportedand upon activation with B(C6F5)3 found to mediate the poly-merization of 1-hexene, the latter in isotactic and living fashion(20). These systems appeared ideal entries for a validation of our

    modeling; therefore, we prepared and tested them in propenepolymerization. In agreement with the QM prediction, the chiralpocket of 1 turned out to be too open to recognize the twoenantiofaces of propene, and a syndiotactically enrichedpolypropylene was obtained because of chain-end control. Forpolymerizations carried out at 25°C in toluene, 13C NMRstereosequence analysis on the polymers ended up indeed with� � (1 � �) � 0.387, independently of propene concentration(Table 2, entry 1). With R1 � tert-butyl, on the other hand, thechiral pocket is tighter, and we were pleased to find out that QMmodeling had been right to predict that site control would bedominant. In fact, depending on its � or � configuration, at thetwo homotopic active sites the active species of 3 has a fairlystrong preference for one or the other monomer enantiofaces,which wipes out the influence of chain end. Quantitatively, bystatistical analysis of the 13C NMR stereosequence distributionfor polypropylene samples prepared with 3 at 25°C in toluene wemeasured � � � � 0.955; the polymer is therefore isotactic,albeit with 4.5 mol% of randomly distributed steroirregular units(. . . mmmrrmmm . . .), which results into a fairly modest meltingtemperature (Table 2, entry 3).

    The number average polymerization degree (Pn) of thepolypropylene made with 1 and 3 is rather low (Table 2) andindependent of propene concentration, as is the case when �-Helimination to propene is the dominant chain transfer pathway(Pn � Rp�Rt � kins[C3H6]�(kH,M[C3H6]), where Rp and Rt are theoverall rates of chain propagation and transfer, and kins and kH,Mare the kinetic constants of 1,2 monomer insertion and of �-Helimination to monomer, respectively. These results are nicelyconsistent with the QM indications.

    Table 1. Main results of the QM modeling of propenepolymerization at active cations derived from the (ONNO)ZrBn2precursors of Fig. 5, with an iso-butyl group simulating thegrowing chain

    R1 R2 �Eregio# * �Eenantio

    # † �Et-p# ‡

    H H 3.5 �0 0Methyl Methyl 3.6 �0 0.2tert-butyl tert-butyl 3.6 1.7 1.81-Adamantyl Methyl 4.1 4.1 3.5�,�-dimethylbenzyl (cumyl) Methyl 3.5 3.7 2.2Triphenylmethyl (trityl) Methyl 1.6 �0 3.8

    Difference in internal energy (in kcal�mol�1) between the lowest transitionstates for: *2,1 and 1,2 propene insertion; †1,2 propene insertion with the twopossible enantiofaces; ‡chain transfer via �-H elimination to propene andchain propagation via 1,2 propene insertion. The latter is scaled to the case ofR1 � R2 � H set to �Et-p

    # � 0.

    Fig. 5. The general structure of the (ONNO)ZrBn2 precatalysts studied in thepresent work (Bn � benzyl).

    Fig. 6. QM models of the transition states of 1,2 propene insertion (Left) and �-H elimination to propene (Right) at an active [(ONNO)ZrP]� cation, in whichthe growing polymer chain (P) is simulated with an iso-butyl group (in bright green). Substituents R1 of Fig. 5 are generically represented as yellow spheres. Themain nonbonded interactions are indicated by dotted lines.

    Busico et al. PNAS � October 17, 2006 � vol. 103 � no. 42 � 15323

    CHEM

    ISTR

    YSP

    ECIA

    LFE

    ATU

    RE

    Dow

    nloa

    ded

    by g

    uest

    on

    July

    3, 2

    021

  • Moving from these facts, we undertook a program of catalystfine-tuning aimed primarily at the achievement of highly isotac-tic, high-molecular-mass polypropylene. We also wanted tosynthesize one or more catalysts for which chain-end and sitecontrol are both nonnegligible and verify the QM-predictedexistence of a threshold in the steric bulk of R1 above which thesubstituent effect on the regioselectivity and enantioselectivitywould be detrimental.

    Not surprisingly, we found out that the range in whichchain-end and site control are concurrent corresponds to R1 ofa size intermediate between methyl and tert-butyl, which is thecase, e.g., of R1 � cyclohexyl. The x-ray molecular structure ofthe dibenzyl precatalyst 2 (R2 � methyl) is shown in Fig. 7 Left.The coordination around Zr is distorted-octahedral with tran-s(O,O), cis(N,N) configuration [C-Zr-C � 111.6(1), O-Zr-O �166.3(1), N2-Zr-O2 � 74.8(1)], similarly to what was observedbefore for 3 (20). Apart from the distortions traceable to theconstraints imposed by the chelation of the tetradentate(ONNO) ligand, the local symmetry at the metal deviates fromC2 for the arrangement of the two benzyl moieties, one of which

    is bound in �1 fashion, whereas the other approaches �2 coor-dination [Zr-C1b-C13b � 97.1(2)°; Zr-C13b � 0.2880(4) nm;C1a-Zr-C1b-C13b � 2.9(3)°]. The high-resolution 1H NMRspectrum (Supporting Text) clearly indicates that the trans(O,O),cis(N,N) configuration is maintained in solution, where themolecular motions result into an average C2 symmetry.

    The chiral active pocket of 2 is tight enough to moderatelyfavor propene insertion with one enantioface, but this preferenceconflicts with the preference of the growing chain for analternation. As a result, the two controls can be either ‘‘in phase’’or ‘‘out of phase,’’ depending on the configuration of thelast-inserted monomeric unit, which is shown unambiguously bypolymer stereosequence analysis ending up with � different from� and � different from (1-�) (entry 2 of Table 2 and Table 3,which is published as supporting information on the PNAS website); to the best of our knowledge, this case was still unprece-dented in Ziegler–Natta propene polymerization (6). The cor-responding polypropylene configuration is really intriguing, withthe copresence of relatively long isotactic and syndiotacticstrands ([mmmm] � 4.6%, [rrrrrr] � 7.0%).

    Table 2. Experimental results of propene polymerization with catalysts derived from complexes 1-7 (at 25°C in toluene; forexperimental details, see Materials and Methods and Supporting Text)

    Precatalyst�entry R1 R2 mm, % rr, % � � 2,1 units, % Tm,°C Pn � 10�2 Rp

    1 Methyl Methyl 12 30 0.387 1-� 0.2 Amorphous 1.0 6.42 Cyclohexyl Methyl 18 41 0.517 0.844 0.3 Amorphous 1.5 3.43 tert-butyl tert-butyl 87 4.3 0.955 � 0.3 120 1.4 2.54 1-Adamantyl Methyl 98.8 0.4 0.996 � 0.4 152 18 4.85 Cumyl Methyl 97.6 0.8 0.980 � 0.8 140 50 736 Diphenylmethyl tert-butyl 86 4.0 0.93 � 1.0 114 9.1 7.47 Trityl tert-butyl 24 37 �0.5* �0.5* 10 Amorphous 20 1.8

    mm and rr � 13C NMR fractions of mm and rr triads in the polymer. � and � � stochastic probabilities of chain propagation, as obtained by statistical analysisof 13C NMR stereosequence distribution (for the model, see text). 2,1 units � 13C NMR fraction of regioirregular (2,1) monomeric units in the polymer. Tm � DSCmelting temperature of the polymer. Pn � 1H NMR number average degree of polymerization. Rp � rate of polymerization, in kg(polymer)�mol(Zr)�1 [C3H6]�1�h�1.*Difficult to quantify more precisely by 13C NMR, caused by the extensive overlapping between resonances arising from regio and stereo defects.

    Fig. 7. X-ray molecular structures of the dibenzyl precatalysts of Fig. 5 with R1 � cyclohexyl, R2 � methyl (2; Left) and R1 � trityl, R2 � tert-butyl (7; Right). Thermalellipsoids are shown at 25% probability level. For clarity, all hydrogen atoms were omitted. More details are provided in Supporting Text.

    15324 � www.pnas.org�cgi�doi�10.1073�pnas.0602856103 Busico et al.

    Dow

    nloa

    ded

    by g

    uest

    on

    July

    3, 2

    021

  • At entries 4–7 of Table 2, in turn, we illustrate the effect onthe extent of site control of increasing the steric demand of R1beyond that of a tert-butyl. The catalyst obtained from thedibenzyl precursor 4, with R1 � 1-adamantyl (and R2 � methyl),has an exceedingly high enantioselectivity (at 25°C, we measured� � � � 0.996, corresponding to an enantiomeric excess �99.2%) and yields a highly isotactic polypropylene with a meltingtemperature and enthalpy of 152°C and 130 J�g�1, respectively;these values are among the highest ever reported for propenepolymerization in homogeneous phase (6, 7, 13, 14). On theother hand, in accordance with the computational analysis, assoon as R1 becomes bigger the regioselectivity and enantiose-lectivity start fading out. An extreme case is that of complex 7,with R1 � trityl (and R2 � tert-butyl). The steric crowding at theZr center is evident from the x-ray molecular structure (Fig. 7Right). Although the configuration is the same as 2 and 3, thehuge trityl groups intrude into the volume occupied by the twobenzyl moieties, which are both forced to bind in a ‘‘slim’’ �1fashion. The NMR characterization in solution (see Fig. 9, whichis published as supporting information on the PNAS web site),while confirming the trans(O,O), cis(N,N) configuration, pointedout a diffuse substituent interlocking, as shown in particular bythe broad resonances of most trityl protons in the solution 1HNMR spectrum at room temperature. At the 13C NMR charac-terization, the polypropylene prepared with this catalyst turnedout to be practically atactic, and also poorly regioregular, with�10 mol% of randomly distributed head-to-head�tail-to-tailenchainments (Table 2, entry 7, and Fig. 10, which is publishedas supporting information on the PNAS web site).

    Fig. 8 shows the calculated transition states of the four lowestenergy propene insertion paths identified for a model activecation of 7 (R2 � Me), with the growing chain simulated with aniso-butyl group (for details, see Computational Section in Sup-porting Text). In Fig. 8 A and B, the first C-C bond of the iso-butylpoints away from the nearest-in-space trityl substituent; a pro-pene molecule that inserts with 1,2 regiochemistry cannot avoidbumping with the methyl substituent either into the trityl (Fig.8A) or, with the other enantioface, into the first chain C-C bond

    (Fig. 8B). The alternative 1,2 insertion shown in Fig. 8C is morefavorable as far as the nonbonded contacts of propene areconcerned, but it is the chain now to interfere with the trityl. Allthree transition states are practically at the same energy, whichgives reason for the observed poor enantioselectivity in thedominant insertion regiochemistry; we may define this situationas ‘‘internally conflicting site control,’’ in the sense that the directand indirect (chain-mediated) nonbonded steering of the(ONNO) ligand on the inserting monomer are divergent. Underthese circumstances, it is not at all surprising that 2,1 insertioncan be competitive; the transition state in Fig. 8D, in particular,was calculated to be higher in energy than the other three by only�1.6 kcal�mol�1 only, in very nice agreement with the poorregioselectivity observed experimentally.

    A very important issue that still needs to be discussed is theeffect of R1 on the ease of �-H elimination to propene. Onceagain in agreement with the QM indication (Table 1 and Fig. 6Right), high-molecular-mass polypropylenes were obtained withcatalysts derived from the encumbered complexes 4–7 (Table 2).As already observed for 1 and 3, Pn turned out to be independentof propene concentration, which confirms that �-H eliminationto propene is indeed the main chain transfer pathway for thesecatalysts, too.

    An interesting feature of the whole catalyst family is anunusually low [for Ziegler–Natta catalysis (26–28)] turnoverfrequency, as if the polymerization occurred ‘‘at slow motion.’’In particular, for the most sterically congested catalysts 4, 6, and7 we measured average chain growth times of several hours at25°C (under the conditions described in Materials and Methods).This controlled kinetic behavior (29) is highly advantageous forapplication in block copolymerization (29, 30); we have dis-cussed specifically this aspect in ref. 19, where we documentedthe use of catalyst systems based on complex 4 for the firstsyntheses of well defined and fully characterized samples of(highly isotactic polypropylene)-block-(polyolefin).

    ConclusionsIn our opinion, the results presented in the previous section arerelevant for (at least) two reasons. First, they demonstrate how

    Fig. 8. QM transition states for the lowest energy 1,2 and 2,1 propene insertions into a growing polypropylene chain (simulated with an iso-butyl group) ata model active cation of 7 with � configuration. (A) 1,2 insertion�re�anti. (B) 1,2 insertion�si�syn. (C) 1,2 insertion�si�anti. (D) 2,1 insertion�si. The mainnonbonded interactions relevant for the enantioselectivity are explicitly indicated by dotted lines.

    Busico et al. PNAS � October 17, 2006 � vol. 103 � no. 42 � 15325

    CHEM

    ISTR

    YSP

    ECIA

    LFE

    ATU

    RE

    Dow

    nloa

    ded

    by g

    uest

    on

    July

    3, 2

    021

  • far one can get in the molecular control of Ziegler–Nattapropene polymerizations with applications of computer-aidedcatalyst design. But we do not want to hide that some importantlimitations still need to be overcome; in particular, the ab initioprediction of catalytic activity is somehow at a pioneering stage,mainly because most active species in single-center Ziegler–Natta olefin polymerizations are cationic (21), and realisticcalculations of counterion and solvent effects have becomefeasible only recently (31–34). However, computational chem-istry is progressing so fast that it is plausible to anticipate thatmost residual problems will be solved in a few years.

    Second, the results are direct proof that, at least in favorablecases, a proper fine-tuning of a single basic catalyst structurewithout altering the nature and the symmetry of the ancillaryligand framework can provide access to practically the full rangeof polypropylene microstructures. This finding is at odds with thelesson of metallocene catalysts (6, 7, 13) and provides a strongargument in support of catalyst diversification programs basedon parallel synthesis and structural amplification (10, 11).

    Altogether, we believe that the above points to high-throughput computation as the next-generation tool for Ziegler–Natta catalyst discovery (ref. 35 and references therein).

    Materials and MethodsAll reactants and solvents used for precatalyst synthesis werepurchased from Aldrich (St. Louis, MO) and used as received.Air-sensitive compounds were manipulated under argon, usingSchlenk techniques and�or a Braun (Melsungen, Germany)LabMaster 130 glove box. Complexes 1 and 3–5 were preparedas described (19, 20). Complexes 2, 6, and 7 were synthesizedwith similar procedures, as described in Supporting Text.

    The molecular structures of complexes 2 and 7 were deter-mined by x-ray diffraction on single crystals grown from asaturated solution of the complex in benzene (for 2) or toluene(for 7) at 293 K. Data collection was performed in flowing N2at 173 K on a Bruker-Nonius (Delft, The Netherlands) kap-paCCD diffractometer (MoK� radiation, CCD rotation images,thick slices, � scans � � scans to fill the asymmetric unit). Fulldetails can be found in Supporting Text. Crystallographic data inCIF format are available from the Cambridge CrystallographicData Centre.

    Propene polymerization runs were carried out at 25°C intoluene at two different concentrations, [C3H6] � 1.3 M and 5.3M, mainly to check for the effect of this parameter on polymeraverage molecular mass. The experiments at lower propeneconcentration were carried out in a 250-ml magnetically stirred,two-necked, jacketed Pyrex reactor; those at higher concentra-tion were carried out in a 2-liter magnetically stirred, stainless-steel reactor (Brignole AU-2). In all cases, the precatalyst wasdissolved in toluene and activated with methylalumoxane[Crompton (Middlebury, CT) 10% (wt�wt) solution in toluene]added with 2,6-di-tert-butylphenol (Aldrich) to trap ‘‘free’’ tri-methylaluminum (23). The general procedures can be found inSupporting Text.

    Quantitative 1H and 13C NMR spectra of all polypropylenesamples were recorded at 120°C, on 35 mg�ml solutions intetrachloroethane-1,2-d2, with a Bruker DRX 400 Avance spec-trometer operating at 100.6 MHz with a 5-mm BBO probe.Conditions for 1H NMR were: 90° pulse; acquisition time, 4.0 s;relaxation delay, 2.0 s; 32 transients. Conditions for 13C NMRwere: 80° pulse; acquisition time, 1.6 s; relaxation delay, 3.0 s;10,000 transients. Broad-band proton decoupling was achievedwith a modified WALTZ16 sequence (BI�WALTZ16�32 byBruker). Peak integration by full spectral simulation and best-fitcalculations of stereosequence distributions were carried out byusing the SHAPE2004 and CONFSTAT (version 3.1 for Win-dows) software packages, respectively. For more detail onpolymer microstructures and statistical analysis thereof, seeSupporting Text.

    All possible transition states of propene insertion and �-Htransfer to propene monomer at model active cations werecomputed by means of full QM methods. Stationary points onthe potential energy surface were calculated with the Amster-dam Density Functional program system, release 2004.01 (doc-umentation, including user manuals and licensing, is available atwww.scm.com). More information can be found in SupportingText and Fig. 11, which is published as supporting information onthe PNAS web site.

    This research was supported by the Italian Ministry of Education,University and Research (PRIN 2004) and the Regional Government ofCampania (Legge 5, 2003). The NMR polymer characterizations andQM calculations were carried out at the Centro Interdipartimentale diMetodologie Chimico-Fisiche of the University of Naples Federico II.

    1. Moore EP, Jr, ed (1996) Polypropylene Handbook: Polymerization, Character-ization, Properties, Applications (Hanser, Munich).

    2. Moore EP, Jr (1998) The Rebirth of Polypropylene: Supported Catalysts (Hanser,Munich).

    3. Scheirs V, Kaminsky W, eds (1999) Metallocene-Based Polyolefins (Wiley,Chichester, UK), Vols 1 and 2.

    4. Boor J, Jr (1979) Ziegler–Natta Catalysts and Polymerizations (Academic, NewYork).

    5. Bovey FA (1972) High-Resolution NMR of Macromolecules (Academic, NewYork).

    6. Busico V, Cipullo R (2001) Progr Polym Sci 26:443–533.7. Resconi L, Cavallo L, Fait A, Piemontesi F (2000) Chem Rev 100:1253–1345.8. Ziegler K (1954) Brennst-Chem 35:321–325.9. Sinn H, Kaminsky W (1980) Adv Organomet Chem 18:99–149.

    10. Boussie TR, Diamond GM, Goh C, Hall KA, LaPointe AM, Leclerc M, LundC, Murphy V, Shoemaker JAW, Tracht U, et al. (2003) J Am Chem Soc125:4306–4317.

    11. Boussie TR, Diamond GM, Goh C, Hall KA, LaPointe AM, Leclerc MK,Murphy V, Shoemaker JAW, Turner H, Rosen RK, et al. (2006) Angew ChemInt Ed 45:3278–3283.

    12. Arriola DJ, Carnahan EM, Hustad PD, Kuhlman RL, Wenzel TT (2006)Science 312:714–719.

    13. Alt HG, Koppl A (2000) Chem Rev 100:1205–1222.14. Gibson VC, Spitzmesser SK (2003) Chem Rev 103:283–315.15. Corradini P, Guerra G, Cavallo L (2004) Acc Chem Res 37:231–241.16. Makio H, Kashiwa N, Fujita T (2002) Adv Synth Catal 344:477–493.

    17. Busico V, Cipullo R, Ronca S, Budzelaar PHM (2001) Macromol RapidCommun 22:1405–1410.

    18. Busico V, Cipullo R, Friederichs N, Ronca S, Togrou M (2003) Macromolecules36:3806–3808.

    19. Busico V, Cipullo R, Friederichs N, Ronca S, Talarico G, Togrou M, Wang B(2004) Macromolecules 37:8201–8203.

    20. Tshuva ET, Goldberg I, Kol M (2000) J Am Chem Soc 122:10706–10707.21. Chen, EY-X, Marks TJ (2000) Chem Rev 100:1391–1434.22. Talarico G, Busico V, Cavallo L (2003) J Am Chem Soc 125:7172–7173.23. Busico V, Cipullo R, Cutillo F, Friederichs N, Ronca S, Wang B (2003) J Am

    Chem Soc 125:12402–12403.24. Deng L, Woo TK, Cavallo L, Margl P, Ziegler T (1997) J Am Chem Soc

    119:6177–6186.25. Margl P, Deng L, Ziegler T (1999) J Am Chem Soc 121:154–162.26. Mori H, Terano M (1997) Trends Polym Sci 5:314–321.27. Busico V, Cipullo R, Esposito V (1999) Macromol Rapid Commun 20:116–121.28. Song F, Cannon RD, Bochmann M (2003) J Am Chem Soc 125:7641–7653.29. Busico V, Talarico G, Cipullo R (2005) Macromol Symp 226:1–16.30. Coates GW, Hustad PD, Reinartz S (2002) Angew Chem Int Ed 41:2236–2257.31. Chan MSW, Ziegler T (2000) Organometallics 19:5182–5189.32. Lanza G, Fragalà IL, Marks TJ (2002) Organometallics 21:5594–5612.33. Busico V, Van Axel Castelli V, Aprea P, Cipullo R, Segre AL, Talarico G,

    Vacatello M (2003) J Am Chem Soc 125:5451–5460.34. Correa A, Cavallo L (2006) J Am Chem Soc 128:10952–10959.35. Miteva MA, Lee WH, Montes MO, Villoutreix BO (2005) J Med Chem

    48:6012–6022.

    15326 � www.pnas.org�cgi�doi�10.1073�pnas.0602856103 Busico et al.

    Dow

    nloa

    ded

    by g

    uest

    on

    July

    3, 2

    021


Recommended