Are Aluminoxanes Nanotubular? Structural Evidence from a Quantum Chemical Study

August 27, 2017 | Autor: John Severn | Categoria: Polymerization, Ab initio calculations, Nanostructures, CHEMICAL SCIENCES, Structure Elucidation
Share Embed


Descrição do Produto

Angewandte

Chemie

structures over other proposed alternatives was demonstrated by quantum chemical calculations on methyl-substituted[6, 7] and unsubstituted[8] aluminoxanes. Herein, we report a HFand B3LYP-level quantum chemical study of the structures of alkylaluminoxanes. The structural determination of MAO is further complicated by association of trimethylaluminum[1] into the {AlO(Me)} core. This led us to an unexpected observation while studying the subsequent reactions between the unsubstituted [AlO(H)]n cages and AlH3. The products of the reactions for cages with n = 6 and 8, which are the unsubstituted counterparts of two cages characterized crystallographically by Barron et al., are illustrated in Figure 1. The Polymerization Catalysts DOI: 10.1002/anie.200600197

Are Aluminoxanes Nanotubular? Structural Evidence from a Quantum Chemical Study Mikko Linnolahti,* John R. Severn, and Tapani A. Pakkanen Single-site a-olefin polymerization catalysts hold promise for a rational tailoring of the polymer microstructure and hence its properties. One of the critical components in these systems is the co-catalyst, which can profoundly influence the activity, stereoselectivity, and molecular-weight capability of the catalytic system.[1] The archetypal co-catalyst is a solution of methylaluminoxanes (MAO), the structural characterization of which has remained challenging and elusive. As a consequence, the understanding and control of the polymerization process, together with optimization of the co-catalyst, has been handicapped by the inability to determine the structure of the active component. Due to the lack of precise crystallographic and spectroscopic characterization, several structural models have been proposed for MAO. The first proposals included chains and rings with three-coordinate, highly Lewis acidic Al centers.[2] The preference for four-coordinate Al and three-coordinate O atoms was demonstrated in 1983 by Atwood et al. by a crystal structure analysis of [Al7O6Me16] .[3] Further progress towards interpretation of polyhedral cages as the most relevant structural alternative was made in the mid-90s by Sinn[4] and Barron et al.[5] Following the synthesis of [AlO(tBu)]n cages (n = 6–9 and 12),[5] the preference for cage

[*] Dr. M. Linnolahti, Prof. T. A. Pakkanen Department of Chemistry University of Joensuu P.O. Box 111, 80101 Joensuu (Finland) Fax: (+ 358) 13-251-3390 E-mail: [email protected] Dr. J. R. Severn Borealis Polymers Oy P.O. Box 330, 06101 Porvoo (Finland) Supporting Information for this article is available on the WWW under http://www.angewandte.org or from the author. Angew. Chem. Int. Ed. 2006, 45, 3331 –3334

Figure 1. The formation of aluminoxane nanotubes by association of four {AlH3} groups into [AlO(H)]6 and [AlO(H)]8 cages.

AlH3 units attach to the [AlO(H)]n core by breaking the Al O bond between {Al2O2} squares; the mechanism of this “latent Lewis acidicity”[9] has been described by Zurek and Ziegler.[10] For both cages, the reaction is exothermic for addition of up to four AlH3 groups,[11] after which all {Al2O2} squares are opened to form six-membered rings exclusively. The resulting molecular structures are striking, with sections of armchair (2,2) nanotubes capped with two AlH3 groups at each end. These species do not undergo further reaction with AlH3. Changing the hydrogen atoms to methyl groups leads a similar, although somewhat more pronounced, result: in the case of the smaller cage the reaction energy increases from 102.7 kJ mol 1 to 157.2 kJ mol 1. Next, we investigated the cage dimer of [AlO(H)]6, whose methylated counterpart has been proposed as a possible component of MAO.[12] It turns out that this cage dimer is unstable and undergoes a structural deformation. Depending on the orientations of the two cages, the dimer has three isomers; the energetically most favored one, by a margin of over 40 kJ mol 1, is shown in Figure 2. This S4-symmetric isomer, which is favored over two non-interacting cages by 340 kJ mol 1, takes the form of a capped (2,2) nanotube. As the dimer is capped with three adjacent {Al2O2} squares, oligomerization of the cages probably proceeds beyond the dimer. Alternatively, incorporation of trialkylaluminum would terminate the tube growth.

 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

3331

Communications

Figure 2. The formation of a capped (2,2) armchair aluminoxane nanotube by dimerization of face-bridged [AlO(H)]6 cages.

Encouraged by these observations, we studied the tubular section of the unsubstituted aluminoxanes by means of periodic ab initio calculations for the open-ended infinite tubes. In analogy with carbon nanotubes,[13] zigzag, armchair, and chiral nanotubes can be derived. The relative stabilities of the tubes are listed in Table 1 using the preferred [AlO(H)]n cage (T-symmetric Al28O28H28[8]) as a reference. The favored structures for each family are illustrated in Figure 3. Table 1: Diameters [nm] and stabilities [kJ mol 1] relative to the preferred cage (Al28O28H28[8]) for [AlO(H)]n nanotubes of infinite length. [AlO(H)]n nanotube

Diameter

(2,2) (3,3) (4,4) (3,0) (4,0) (5,0) (6,0) (2,1) (3,1) (3,2) (4,1) (4,2)

0.69 0.87 1.03 0.56 0.75 0.81 0.95 0.55 0.71 0.74 0.79 0.88

DE (nHF) 1.4 2.3 3.3 3.8 10.6 10.6 7.2 15.7 5.0 4.4 8.8 3.7

DE (nB3LYP) 3.3 4.1 1.0 1.6 11.5 11.4 8.2 13.3 6.5 6.1 9.9 5.3

The aluminoxane nanotubes are favored over the cage whatever the method applied. The tubes reach their energy minimum at diameters somewhat below 1 nm: armchair in (3,3), zigzag in (4,0), and chiral in (4,1). The preference for zigzag tubes is due to the repulsion between the hydrogen substituents. The zigzag arrangement allows the largest separation between the neighboring substituents (3.76 E in the case of the preferred (4,0), compared to 2.98 E for armchair (3,3) and 3.27 E for chiral (4,1)). The preference for relatively thin tubes, on the other hand, is due to the optimal curvature for sp3-hybridized Al. While the Al28O28H28 cage is a representative of optimal curvature, the cages are handicapped by the {Al2O2} squares necessary for cage closure. In this regard, nanotubular shapes appear more reasonable owing to their larger relative proportion of favorable sixmembered rings. To verify that the preference for the tubular shape is not characteristic for unsubstituted aluminoxanes alone, we repeated the cage dimerization study (Figure 2) for methylsubstituted cages. The substitution pattern does not affect our conclusions, the dimerization energy in the case of methyl substituents being 330 kJ mol 1 (compared to 340 kJ mol 1

3332

www.angewandte.org

Figure 3. Optimized structures of the favored [AlO(H)]n nanotubes of infinite length for each family: armchair (3,3) (top), zigzag (4,0) (middle), and chiral (4,1) (bottom).

for the unsubstituted case). We then optimized the methyland tert-butyl-substituted zigzag (4,0) and (3,0) nanotubes, respectively, and compared their stabilities with those of the correspondingly substituted cages. The calculations of Ziegler et al. suggest that Al12O12Me12 is the favored [AlO(Me)]n cage,[6] therefore we selected this as the reference structure. No data for the relative stabilities of tert-butyl-substituted cages are available, therefore a cage synthesized by Barron et al., namely [AlO(tBu)]6, was selected as a reference.[5] The optimized structures of the substituted cages and nanotubes, together with their relative stabilities, are given in Figure 4. The methylaluminoxane nanotube, in line with the unsubstituted ones, is clearly favored over the cage, the difference in relative energy being almost 20 kJ mol 1 per {AlO(Me)} unit. The opposite is observed for tert-butyl substituents, however, as the marked overcrowding due to the vicinity of the bulky substituents results in significant destabilization—the [AlO(tBu)]6 cage is favored by more than 70 kJ mol 1 per {AlO(tBu)} unit. One should note that the energy difference is due to destabilization of the tube rather than stabilization of the cage owing to the presence of tert-butyl substituents. Preliminary studies on the (4,0) nanotube suggest that the tubes

 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Angew. Chem. Int. Ed. 2006, 45, 3331 –3334

Angewandte

Chemie

Figure 4. Structures and relative stabilities of selected methyl- and tertbutyl-substituted aluminoxane cages and zigzag nanotubes.

become increasingly overcrowded as a function of the tube diameter. The structural differences between MAO and tert-butylsubstituted aluminoxanes can now be discussed in a new light. The [AlO(tBu)]n cages could actually be considered to be very short tubes for which the tube growth is prematurely interrupted after two or three repetitive cycles as a consequence of repulsion between the bulky substituents. In the case of MAO, no interruption of the tube growth occurs as the repulsion between the methyl groups is insignificant. One should note that the current interpretation of the structure of MAO being in a dynamic equilibrium of variably sized cages[7] is basically based on BarronHs tert-butyl-substituted polyhedral cages.[5] Taking into account the significantly different structural preferences of methyl- and tert-butyl-substituted aluminoxanes, the structure of MAO might involve a dynamic equilibrium mostly between nanotubes of variable lengths and thicknesses, perhaps containing fractions of proposed cages and other structural fragments. In order to understand the function of nanotubular MAO as a co-catalyst in polymerization catalysis, further studies are required to determine its average molar mass, tube growth mechanism, capping of the tube ends, and association of trimethylaluminum into the AlO(Me) core. As far as the molar mass is concerned, various experimental values have been reported.[1] Recently, the size of the [Me MAO] anion was determined from a pulsed field-gradient NMR study by Babushkin and Brintzinger.[14] The observed mean effective hydrodynamic radius of 12.2–12.5 E, assuming a spherical structure, corresponds to about 150–200 Al atoms in each MAO molecule. Possibly due to its size, catalytic activity declines strongly at MAO concentrations below an Al/Zr ratio of 200–300:1.[15] The sizes of the preferred MAO cages reported previously[7] are about an order of magnitude lower, hence they are unlikely to account for the structure of MAO. Instead, the presence of hundreds of Al atoms in nanotubular shapes could be a plausible explanation. In the case of the zigzag (4,0) tube, 200 Al atoms would correspond to 50 Angew. Chem. Int. Ed. 2006, 45, 3331 –3334

repetitive {Al4O4Me4} rings. As the tubular section of MAO is inert towards reactions with trimethylaluminum, it is likely that trimethylaluminum becomes associated into the tube ends (see Figure 1). Building on this theory, one could reason that the functional sites of MAO as a co-catalyst are exclusively at the tube ends, hence a MAO molecule containing hundreds of Al atoms arranged into a nanotubular shape would possess, at most, two active sites, namely both ends of the tube. The relatively few active sites of large MAO molecules given the tasks of the co-catalyst, that is, activation of the catalyst precursor and scavenging of impurities, would explain why thousandfold Al/Zr ratios are generally required in catalytic polymerizations. In summary, we have provided new evidence concerning the structures of aluminoxanes, among which methylaluminoxane is particularly important due to its application as a co-catalyst in single-site homogeneous polymerization catalysis. The data presented suggest that MAO is nanotubular in shape. The aluminoxanes are capable of adopting armchair, zigzag, and chiral analogues of the well-known carbon nanotubes and prefer diameters of about 1 nm. Trimethylaluminum is likely to become associated at the tube ends, which act as functional sites in MAO.

Methods All structures, including the periodic ones, were fully optimized. Aluminoxane cages were constrained to the symmetries in question, namely T for [AlO(H)]28, T for [AlO(Me)]12, and C3v for [AlO(tBu)]6, and were characterized as true minima by frequency calculations. Periodic calculations of unsubstituted aluminoxanes were performed at the HF and B3LYP levels of theory; in all other examples the HF method was applied. Periodic calculations generally require optimized basis sets. The optimized 8-5-11G* and 8-411G* basis sets were used for aluminum and oxygen,[16] respectively, together with the standard 6-31G** basis set for hydrogen, as reported in a previous study.[8] For carbon, a 6-21G* basis set with a modified outer sp exponent was adopted.[17] Identical basis sets were applied for the clusters and the periodic tubes. The calculations were performed with Gaussian 03 software.[18] Received: January 17, 2006 Published online: April 7, 2006

.

Keywords: ab initio calculations · aluminoxanes · nanostructures · polymerization · structure elucidation

[1] a) E. Y.-X. Chen, T. J. Marks, Chem. Rev. 2000, 100, 1391 – 1394; b) J.-N. PNdeutour, K. Radhakrishnan, H. Cramail, A. Deffieux, Macromol. Rapid. Commun. 2001, 22, 1095 – 1123; c) M. Bochmann, J. Organomet. Chem. 2004, 689, 3982 – 3998. [2] S. Pasynkiewicz, Polyhedron 1990, 9, 429 – 453. [3] J. L. Atwood, D. C. Hrncir, R. D. Priester, R. D. Rogers, Organometallics 1983, 2, 985 – 989. [4] H. Sinn, Macromol. Symp. 1995, 97, 27 – 52. [5] a) M. R. Mason, J. M. Smith, S. G. Bott, A. R. Barron, J. Am. Chem. Soc. 1993, 115, 4971 – 4984; b) C. J. Harlan, M. R. Mason, A. R. Barron, Organometallics 1994, 13, 2957 – 2969. [6] E. Zurek, T. K. Woo, T. K. Firman, T. Ziegler, Inorg. Chem. 2001, 40, 361 – 370. [7] E. Zurek, T. Ziegler, Prog. Polym. Sci. 2004, 29, 107 – 148.

 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.angewandte.org

3333

Communications [8] M. Linnolahti, T. N. P. Luhtanen, T. A. Pakkanen, Chem. Eur. J. 2004, 10, 5977 – 5987. [9] C. J. Harlan, S. G. Bott, A. R. Barron, J. Am. Chem. Soc. 1995, 117, 6465 – 6474. [10] E. Zurek, T. Ziegler, Inorg. Chem. 2001, 40, 3279 – 3292. [11] The energy of AlH3 was taken for its lowest energy form, which is crystalline aluminum hydride. See reference [8] for details. The energy of AlMe3 was taken as being that of its dimer, Al2Me6. [12] P. L. Bryant, C. R. Harwell, A. A. Mrse, E. F. Emery, Z. Gan, T. Cladwell, A. P. Reyes, P. Kuhns, D. W. Hoyt, L. S. Simeral, R. W. Hall, L. G. Butler, J. Am. Chem. Soc. 2001, 123, 12 009 – 12 017. [13] S. Iijima, Nature 1991, 354, 56 – 58. [14] D. E. Babushkin, H.-H. Brintzinger, J. Am. Chem. Soc. 2002, 124, 12 869 – 12 873. [15] H.-H. Brintzinger, D. Fischer, R. MPlhaupt, B. Rieger, R. M. Waymouth, Angew. Chem. 1995, 107, 1255 – 1283; Angew. Chem. Int. Ed. Engl. 1995, 34, 1143 – 1170. [16] M. Catti, G. Valerio, R. Dovesi, M. CausR, Phys. Rev. A 1994, 49, 179 – 187. [17] M. Catti, A. Pavese, R. Dovesi, V. R. Saunders, Phys. Rev. B 1993, 47, 9189 – 9198. [18] Gaussian 03, Revision C.02, M. J. Frisch, et al.; see Supporting Information.

3334

www.angewandte.org

 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Angew. Chem. Int. Ed. 2006, 45, 3331 –3334

Lihat lebih banyak...

Comentários

Copyright © 2017 DADOSPDF Inc.