Research in the Barron Group

 

Group 13 Organometallic and Coordination Compounds
 
 
 
Group 13 Alkyls: Structure, Properties and Reactivity
 
Structure and Bonding
Control over Volatility
 
Reactivity
Oxidation
Hydrolysis
Ligand Exchange
Siloxane Cleavage
 
Lewis Acid Properties of Group 13 Metal Compounds
Activation of Small Molecules
Activation of Metals
 
Alumoxanes
Redefining their Structure
Redefining their Mode of Activation: Latent Lewis Acidity
New catalysts
 
Group 13 Compounds as Ligands

  

Group 13 Alkyls: Structure, Properties and Reactivity

The Group 13 alkyl compounds, in particular those of aluminum, are widely used in industrial catalysis, organic synthesis and the electronics industry. As would be expected from their position in the Periodic Table, they have the general formula, MR3-nXn, where R is a hydrocarbon unit and X can be a hydride, halide, alkoxide, or related group.
 
Aluminum and indium trialkyls are ordinarily oligomeric involving alkyl bridges, except where precluded by steric interactions. Trimethyl aluminum is the archetypal electron deficient compound, while the presence of halide, alkoxide, amide and similar groups results in oligomers with octet configurations. In contrast to the aluminum and indium compounds GaR3 are monomeric, although the compounds of the formula, GaR3-nXn follow the same patterns as their aluminum analogs.
 
The most common reaction of Group 13 alkyls involves the reaction between the M-R bonds and an acid protic source, HX. However, M-R bonds also undergo insertion reactions involving a variety of small molecules.
 
Top of Page

Structure and Bonding

Our studies into the basic structure and bonding in aluminum compounds involve a desire to understand the relative magnitude of steric versus electronic effects in defining the structures of simple organometallic compounds (Organometallics, 1991, 10, 597) as well as controlling the extent of oligomerization (J. Chem. Soc., Dalton Transs., 992, 3179). Using a simple series of phosphine adducts of AlMe3 we were able to demonstrate that steric factors predominated (J. Chem. Soc., Dalton Trans., 1988, 3047), however, based upon structural (Organometallics, 988, 7, 2543), spectroscopic (J. Am. Chem. Soc., 1990, 112, 3369) and ab initio (J. Am. Chem. Soc., 1991, 113, 39) studies we proposed the presence of a weak p-type interactions between aluminum and oxygen in monomeric aluminum alkoxides and aryloxides involving the donation of electron density of the oxygen lone pair to the Al-X anti bonding orbital.
  
Our evolving understanding of the structure and bonding in Group 13 organometallic compounds has led to the development of a quantitative measure of steric bulk (Organometallics, 1999, 18, 4399), and the development of a Lewis acidity scale.
 
Other results include: the synthesis of the first 6-coordinate aluminum alkyl (J. Am. Chem. Soc., 1989, 111, 398); the first observation of the trans-influence in an aluminum compound. (Organometallics, 1989, 8, 1828); Topological reorganization of gallium sulfide clusters (Organometallics, 1992, 11, 2783).
 
Top of Page

 

 

 Control over Volatility

Industrial MOCVD processes rely almost exclusively on gaseous and liquid precursors. Systems where solid precursors are employed are generally undesirable due to the difficulty of maintaining a constant flux of source vapors over a non-equilibrium percolation (solid) process (Adv. Mater. Optic. Electron., 1993, 2, 271). Since the majority of metal-organic compounds reported in the literature are solids, alternative approaches have been used to overcome these difficulties including the synthesis of new more volatile precursor compounds. Delay in developing a rational approach to volatile compounds is, in part, due to a lack of a detailed understanding in the factors that control the volatility of metal-organic compounds.
 
Temperature of volatilization and sublimation enthalpies (DHsub) for cubane compounds [(R)Ga(E)]4 (where R = tBu, EtMe2C, Et2MeC, or Et3C and E = S, Se, or Te) have been determined (Chem. Mater., 1997, 9, 796). The temperature of volatilization was found to increase in a linear fashion with respect to increasing molecular mass, perturbations were observed that can be attributed to intermolecular ligand interactions. Sublimation enthalpies (DHsub) for each cubane are more dependent on the degree of branching of the alkyl ligand than the molecular mass effects alone. Using the TGA sublimation data vapor pressures may be calculated for each of the cubane compounds over a wide range of temperatures.
 
Sublimation enthalpies (DHsub) for M(b-diketonate)n complexes were also shown to be dependent on the number and type of intermolecular interactions, rather to be more substantial than molecular mass effects. The relationship between the DHsub of the substituted b-diketonate derivatives as compared to the values for parent M(acac)n may be used to predict either quantity for a range of M(b-diketonate)n complexes where the values for M(acac)n are known (Adv. Mater. Optics Electron., 2000, 10, 223).
 
 

Top of Page

 

Reactivity

The types of reactions that Group 13 compounds undergo (oxidation, hydrolysis/solvolysis, ligand exchange, etc.) are well known, however, a detailed understanding of the mechanisms, intermediates and product structures is often missing. We have attempted to gain a fundamental understanding of the reactivity of Group 13 organometallic compounds and apply our understanding across a wide range of applications: from surface stabilization of nanoparticles to the chemical control over the structure of a solid material.

  

Oxidation

The pyrophoric nature of Group 13 trialkyls (MR3) means that oxidation reactions are often uncontrolled if not catastrophic! However, if the steric bulk of the alkyl group (R) is sufficient it is possible to isolate the intermediates. For example, reaction of In(tBu)3 with dioxygen results in the isolation of [(tBu)2In(OOtBu)]2 (J. Am. Chem. Soc., 1989, 111, 8966). The gallium analog is made in a similar manner and their use as mild oxidation agents has been explored (Organometallics, 1993, 12, 4908).
 
In a related series of reactions, the interaction of Ga(tBu)3 with elemental sulfur, selenium, and tellurium was explored (Organometallics, 1992, 11, 1055). In the case of sulfur the alkyldisulfide derivative, [(tBu)2Ga(SStBu)]2, is formed as an intermediate to the [(tBu)Ga(S)]4 cubane. For selenium and tellurides the cubanes are formed via [(tBu)2Ga(EtBu)]2 (E = Se, Te).

Top of Page

 Hydrolysis 

The hydrolysis of aluminum alkyls is ubiquitous in the handling of these compounds. Partial hydrolysis is the most common cause of impurities, however, the purposeful hydrolysis of AlR3, to alkylalumoxanes is a vital part of the new generation of highly active polymerization catalysts.
 
The hydrolysis of Al(tBu)3 yields the trimeric hydroxide compound, which may be converted to a mixture of alumoxanes, [(tBu)AlO]n (n = 6, 7, 8, 9, 12) upon mild thermolysis. Based on spectroscopic evidence, and confirmed by the X-ray crystallographic structural determinations of [(tBu)AlO]6, [(tBu)AlO]8 and [(tBu)AlO]9, we have shown that these compounds have three-dimensional cage structures (J. Am. Chem. Soc., 1993, 115, 4971 and Organometallics, 1994, 13, 2957). In addition, we have demonstrated that partial hydrolysis of Al(tBu)3 allows for the isolation of the tetra-alumoxane, [(tBu)2Al{OAl(tBu)2}]2, whose structure contains the three-coordinate aluminum center that has been proposed to be active in olefin polymerization.
  
The oxidation and hydrolysis of (Me2InPPh2)2 to yield the first 13/16 cubane compound (Polyhedron, 1988, 7, 2091). Related to hydrolysis is the reaction of Ga(tBu)3 with H2S which yields the first gallium hydrosulphido complex [(tBu)2Ga(SH)]2, thermolysis of which results in the formation of the first Ga-S analog of an alumoxane, [(tBu)GaS]4 (J. Chem. Soc., Chem. Commun., 1991, 1315).

 

 

Ligand Exchange 

While ligand exchange of alkyls, alkoxides, and halides is common for Group 13 compounds, we have discovered a chalcogenide exchange unique to gallium-telluride cubanes.
 
Reaction of [(tBu)GaTe]4 and elemental sulfur or selenium, results in the stoichiometric formation of the appropriate cubane, [(tBu)GaE]4 (E = S, Se), and metallic tellurium (Organometallics, 1998, 17, 5310). Each of the intermediate cubane compounds, [(tBu)4Ga4ExTe4-x] (x = 0 - 4; E = S, Se) has been characterized. The rate of the chalcogenide exchange is not only dependent on the chalcogen but also the allotropic form of the chalcogen. The chalcogen exchange reaction is first order with respect to the cubane, and the DH and DS have been determined. The exchange reaction is heterogeneous in nature and involves the activation of the cubane via surface absorption followed by cage opening.
 

Top of Page

Siloxane cleavage 

We have shown that the products from the cleavage of poly(diorganosiloxanes) with AlMe3 are dimeric aluminum siloxides resulting from methyl transfer from aluminum to silicon (Organometallics, 1990, 9, 2137). An extension of this concept involves the reaction of R2AlH with cyclic siloxanes leading to rupture of the silicon-oxygen framework and the formation of aluminum polysiloxides (Organometallics, 1999, 18, 5395). Some examples of the unusual compounds formed by this route include: R2Al(OSiMe2H)(OSiMe2OSiMe2H)AlR2 (R = tBu, iBu), Me2Al(OSiMe2H)AlMe2(OSiMe2O)Me2Al(OSiMe2H)AlMe2, (tBu)2Al(OSiMe2H)(OSiMe2OSiMe2OSiMe2H)Al(tBu)2, R2Al(OSiMe2H)AlR2(OSiMe2OSiMe2O)R2Al(OSiMe2H)AlR2 (R = tBu, iBu). The reaction pathway for the cleavage of polysiloxanes has been explored.
 

Top of Page

Lewis Acid Properties of Group 13 Metal Compounds

Activation of Small Molecules
Our study of the activation of small molecules by Group 13 Lewis acids was initiated by the study of the p-face selectivity of coordinated ketones to nucleophilic addition. In addition to the importance of aluminum-oxygen p-bonding we noted that the reduction potential of the ketone is reduced by up to 1 V upon coordination (J. Am. Chem. Soc., 1990, 112, 3446). This result led to a general study of the interaction of organic carbonyls with sterically crowded aryloxide compounds of aluminum (Organometallics, 1990, 9, 3086).  A unique organic transformation that takes place on sterically demanding aluminum alkyl compounds is the direct conversion of an aldehyde to a ketone (Tet. Lett., 1990, 31, 323). 
 
As part of our study on the hydrolysis mechanism we have investigated alcohol and secondary amine complexes of tri-tert-butyl aluminum. The observation of enhanced stability through intra-molecular hydrogen bonding (J. Chem. Soc., Dalton Trans., 1997, 3129) offers an insight into the activation of an alcohol (or water) upon coordination to a Lewis acid. The formation of the Lewis acid-base adduct activates the a-proton by increasing its acidity, as measured by a decrease in its pKa by about 7. However, the aluminum alkyl is deactivated upon coordination of a Lewis base. This has led to our proposal of an intermolecular mechanism for the hydrolysis (solvolysis) of aluminum alkyls.

 

 

Activation of Metals 
 
In contrast to the well known reactivity of Group 13 halides, the Lewis acidic nature of Group 12 halides, in particular those of mercury, has been much less studied. However, the chemistry of mercury(II) salts with aromatic hydrocarbons is well developed and Hg...arene complexes are well established as important intermediates, although simple complexes have only been characterized spectroscopically. Based on the possibility that Group 13 halide Lewis acids could "activate" other weaker Lewis acids we have investigated the effect of AlCl3 and GaCl3 on the stability of Hg...arene complexes. 
 
The reaction of HgCl2 with two molar equivalents of MCl3 (M = Al, Ga) in a substituted aromatic solvent (C6H6-xMex) yields a colored solution, from which crystalline material may be obtained in moderate to high yield of [Hg(arene)2(MCl4)2] for C6H5Me, C6H5Et, o-C6H4Me2, and C6H3-1,2,3-Me3 (Angew. Chem. Int. Ed., 2000, 39, 4117) In contrast, reaction of HgCl2 with two molar equivalents of AlCl3 in benzene, m-C6H4Me2, p-C6H4Me2 yields liquid clathrates.
 
Each toluene in [Hg(arene)2(MCl4)2], is bound in a highly asymmetric h2 manner with the shortest Hg-C (ca. 2.3 - 2.4 Å), being significantly shorter than observed for the intra-molecular coordination discussed above (ca. 3.2 Å). The stability of the Hg...arene interaction is as a consequence of the activation of the mercury by the Group 13 halide.
 
Solution NMR could not be obtained for [Hg(arene)2(MCl4)2]. For example, dissolution of the toluene complexes in C6D6 results in the rapid quantitative formation of C6D5Me and C6D5H. It is important to note that if C6H5Me and C6D6 are mixed in the presence of < 0.1 mol% of [Hg(arene)2(MCl4)2] complete scrambling of the aromatic hydrogen/deuteriums occurs; indicating that the H/D exchange reaction is catalytic.

 

Top of Page

Alumoxanes: Opening the Black Box

 

Redefining their Structure
 
Methylalumoxane (MAO), the product from the hydrolysis of AlMe3, has significant industrial application without industry having a clear picture as to what the structure of MAO is! We set out to develop a cohesive view of the structure of MAO by the study of the tert-butyl analogs.
 
Our isolation of the tert-butyl alumoxanes, [(tBu)AlO]n (n = 6, 7, 8, 9, 12), was the first step towards the structural characterization of MAO.
 
 
Redefining their Mode of Activation: Latent Lewis Acidity
 
Based upon conventional wisdom, the cage alumoxanes, [(tBu)AlO]n should be inactive as co-catalysts with Cp2ZrMe2, while [(tBu)2Al{OAl(tBu)2}]2 would be expected to be an active co-catalyst, however, the obverse is true: the electron precise cage compounds are active co-catalysts with metallocenes (J. Am. Chem. Soc., 1995, 117, 6465). Spectroscopic characterization of the catalytically active complex, [Cp2ZrMe][(tBu)6Al6O6Me], has led to our proposal that alumoxane's activity is derived from their "latent Lewis acidity".
 
 
Latent Lewis acidity is defined as the ability of a electron precise molecule, e.g., a cage alumoxane, to undergo cage opening, via heteroleptic bond cleavage, to generate a Lewis acidic site. For a given bond type (i.e., an Al-O dative bond in alumoxanes) the relative magnitude of the latent Lewis acidity is related to the relative strain present in the cage. Thus, in general four-membered Al2O2 rings are more strained than there six-membered Al3O3 homologues, and hence exhibit higher latent Lewis acidity. Based upon the angular distortions of the cage atoms from an ideal geometry a semi-qualitative value for the latent Lewis acidity may be obtained, allowing a prediction of the relative reactivity of a series of alumoxane cage structures (Organometallics, 1996, 15, 5514).
 
Top of Page  
 
 
New Catalysts
 
Once an understanding is obtained of the structure and reactivity of alumoxanes, new catalyst systems may be developed: alumoxanes as co-catalysts in palladium catalysed co-polymerization of carbon monoxide and ethylene (Organometallics, 1996, 15, 2213 and Macromolecules, 1996, 29, 1110) and the latent Lewis acid catalyzed polymerization of [R,S]-b-butyrolacetone (J. Chem. Soc., Chem. Commun., 1997, 2183). Furthermore, routes to highly active alumoxanes may be developed (Organometallics, 2001, 20, 460).
 
 
Group 13 Compounds as Ligands
 
We have recently started to explore the use of Group 13 compounds as ligands to transition and main group metals. Initial results are aimed at using gallium-(neopentane diol) compounds (J. Chem. Soc., Dalton Trans., 2000, 2151) as cryptan ligands to transition metals.
  
 
As part of our work in understanding the mechanism by which borate anion cross links guar polymers, we have demonstrated that borate moieties act as ligands to Group 1 metals (J. Chem. Soc., Dalton Trans., 2000, 3100).
 

Top of Page

 

Return to Research in the Barron Group

 

Return to Barron Group Home Page