Organoaluminium chemistry

Last updated
The ball-and-stick model of diisobutylaluminium hydride, showing aluminium as pink, carbon as black, and hydrogen as white. DIBAL-3D-balls.png
The ball-and-stick model of diisobutylaluminium hydride, showing aluminium as pink, carbon as black, and hydrogen as white.

Organoaluminium chemistry is the study of compounds containing bonds between carbon and aluminium. It is one of the major themes within organometallic chemistry. [1] [2] Illustrative organoaluminium compounds are the dimer trimethylaluminium, the monomer triisobutylaluminium, and the titanium-aluminium compound called Tebbe's reagent. The behavior of organoaluminium compounds can be understood in terms of the polarity of the C−Al bond and the high Lewis acidity of the three-coordinated species. Industrially, these compounds are mainly used for the production of polyolefins.

Contents

History

The first organoaluminium compound (C2H5)3Al2I3 was discovered in 1859. [3] Organoaluminium compounds were, however, little known until the 1950s when Karl Ziegler and colleagues discovered the direct synthesis of trialkylaluminium compounds and applied these compounds to catalytic olefin polymerization. This line of research ultimately resulted in the Nobel Prize to Ziegler.

Structure and bonding

Aluminium(III) compounds

Organoaluminium compounds generally feature three- and four-coordinate Al centers, although higher coordination numbers are observed with inorganic ligands such as fluoride. In accord with the usual trends, four-coordinate Al prefers to be tetrahedral. In contrast to boron, aluminium is a larger atom and easily accommodates four carbon ligands. The triorganoaluminium compounds are thus usually dimeric with a pair of bridging alkyl ligands, e.g., Al2(C2H5)4(μ-C2H5)2. Thus, despite its common name of triethylaluminium, this compound contains two aluminium centres, and six ethyl groups. When the organoaluminium compound contain hydride or halide, these smaller ligands tend to occupy the bridging sites. Three coordination occurs when the R groups is bulky, e.g. Al(Mes)3 (Mes = 2,4,6-Me3C6H2 or mesityl) or isobutyl. [4]

Structure of trimethylaluminium, a compound that features five-coordinate carbon. Trimethylaluminium-from-xtal-3D-bs-17.png
Structure of trimethylaluminium, a compound that features five-coordinate carbon.

Ligand exchange in trialkylaluminium compounds

The trialkylaluminium dimers often participate in dynamic equilibria, resulting in the interchange of bridging and terminal ligands as well as ligand exchange between dimers. Even in noncoordinating solvents, Al-Me exchange is fast, as confirmed by proton NMR spectroscopy. For example, at −25 °C the 1H NMR spectrum of Me6Al2 comprises two signals in 1:2 ratio, as expected from the solid state structure. At 20 °C, only one signal is observed because exchange of terminal and bridging methyl groups is too fast to be resolved by NMR. [5] The high Lewis acidity of the monomeric species is related to the size of the Al(III) center and its tendency to achieve an octet configuration.

Low oxidation state organoaluminium compounds

The first organoaluminium compound with an Al-Al bond was reported in 1988 as (((Me3Si)2CH)2Al)2 (a dialane). They are typically prepared reduction of the dialkylaluminium chlorides by metallic potassium: [6]

(R2AlCl)2 + 2 K → R2Al-AlR2 + 2 KCl

Another notable group of alanes are tetraalanes containing four Al(I) centres. These compounds adopt a tetrahedrane core, as illustrated by (Cp*Al)4 and ((Me3Si3C)Al)4. The cluster [Al12(i-Bu)12]2− was obtained from related investigations on the reduction of organoaluminium compounds. This dianion adopts an icosahedral structure reminiscent of dodecaborate ([B12H12]2−). Its formal oxidation state is less than one.

Preparation

From alkyl halides and aluminium

Industrially, simple aluminium alkyls of the type Al2R6 (R = Me, Et) are prepared in a two-step process beginning with the alkylation of aluminium powder:

2 Al + 3 CH3CH2Cl → (CH3CH2)3Al2Cl3

The reaction resembles the synthesis Grignard reagents. The product, (CH3CH2)3Al2Cl3, is called ethylaluminium sesquichloride. The term sesquichloride refers to the fact that, on average, the Cl:Al ratio is 1.5. These sesquichlorides can be converted to the triorganoaluminium derivatives by reduction:

2 (CH3CH2)3Al2Cl3 + 6 Na → (CH3CH2)6Al2 + 2 Al + 6 NaCl

This method is used for production of trimethylaluminium and triethylaluminium. [7]

The overall reaction for the production of these simple alkylaluminium compounds is thus as follows:

2Al + 6RX + 6M → Al2R6 + 6MX (where M is an alkali metal and X is a halogen)

Hydroalumination

Aluminium powder reacts directly with certain terminal alkenes in the presence of hydrogen. The process entails two steps, the first producing dialkylaluminium hydrides. Such reactions are typically conducted at elevated temperatures and require activation by trialkylaluminium reagents:

6 Al + 3 H2 + 12 CH2=CHR → 2 [HAl(CH2CHR)2]3

For nonbulky R groups, the organoaluminium hydrides are typically trimeric. In a subsequent step, these hydrides are treated with more alkene to effect hydroalumiunation:

2 [HAl(CH2CHR)2]3 + 3 CH2=CHR → 3 [Al2(CH2CHR)3

Diisobutylaluminium hydride, which is dimeric, is prepared by hydride elimination from triisobutylaluminium:

2 i-Bu3Al → (i-Bu2AlH)2 + 2 (CH3)2C=CH2

Carboalumination

Organoaluminum compounds can react with alkenes and alkynes, resulting in the net addition of one organyl group and the metal fragment across the multiple bond (carboalumination). This process can proceed in a purely thermal manner or in the presence of a transition metal catalyst. For the uncatalyzed process, monoaddition is only possible when the alkene is substituted. For ethylene, carboalumination leads to a Poisson distribution of higher alkylaluminum species. The reaction is regioselective for 1-alkenes. [8] The so-called ZACA reaction first reported by Ei-ichi Negishi is an example of an asymmetric carboalumination of alkenes catalyzed by a chiral zirconocene catalyst. [9]

The methylalumination of alkynes in the presence of Cp2ZrCl2 [10] [11] is employed for the synthesis of stereodefined trisubstituted olefin fragments, a common substructure in terpene and polyketide natural products. The synthesis of (E)-4-iodo-3-methylbut-3-en-1-ol [12] shown below is a typical application of this reaction:

Carboalumination.png

For terminal alkynes, the reaction generally proceeds with good regioselectivity (>90:10 rr) and complete syn selectivity, even in the presence of propargylic or homopropargylic heteroatom substituents. Unfortunately, extension of the zirconocene-catalyzed methylalumination to alkylalumination with higher alkyls results in lower yields and poor regioselectivities.

Laboratory preparations

Although the simple members are commercially available at low cost, many methods have been developed for their synthesis in the laboratory, including metathesis or transmetalation.

AlCl3 + 3 BuLi → Bu3Al + 3 LiCl
2 Al + 3 HgPh2 → 2 AlPh3 + 3 Hg

Reactions

The high reactivity of organoaluminium compounds toward electrophiles is attributed to the charge separation between aluminium and carbon atom.

Lewis acidity

Organoaluminium compounds are hard acids and readily form adducts with bases such as pyridine, THF and tertiary amines. These adducts are tetrahedral at Al.

Electrophiles

The Al–C bond is polarized such that the carbon is highly basic. Acids react to give alkanes. For example, alcohols give alkoxides:

AlR'3 + ROH → 1/n (R'2Al−OR)n + R'H

A wide variety of acids can be employed beyond the simple mineral acids. Amines give amido derivatives. With carbon dioxide, trialkylaluminium compounds give the dialkylaluminium carboxylate, and subsequently alkyl aluminium dicarboxylates:

AlR3 + CO2 → R2AlO2CR
R2AlO2CR + CO2 → RAl(O2CR)2

The conversion is reminiscent of the carbonation of Grignard reagents. [13] [14] [15]

Similarly, the reaction between trialkylaluminum compounds and carbon dioxide has been used to synthesise alcohols, olefins, [13] or ketones. [16]

With oxygen one obtains the corresponding alkoxides, which can be hydrolysed to the alcohols:

AlR3 + 3/2 O2 → Al(OR)3

A structurally characterized organoaluminum peroxide is [{HC[C(Me)N-C6H5]2}Al(R)-O-O-CMe3] [R=CH(SiMe3)2]. [17]

The reaction between pure trialalkylaluminum compounds and water, alcohols, phenols, amines, carbon dioxide, sulfur oxides, nitrogen oxides, halogens, and halogenated hydrocarbons can be violent. [18] [19]

Applications

Organoaluminium compounds are widely used in the production of alkenes, alcohols, and polymers. Some relevant processes include the Ziegler Process for the production of alcohols from ethylene. Several technologies exist for the oligomerization of ethylene to give alpha-olefins. [20] Organoaluminium compounds are used as catalysts for alkene polymerization to polyolefins, for example the catalyst methylaluminoxane.

Related Research Articles

<span class="mw-page-title-main">Organometallic chemistry</span> Study of organic compounds containing metal(s)

Organometallic chemistry is the study of organometallic compounds, chemical compounds containing at least one chemical bond between a carbon atom of an organic molecule and a metal, including alkali, alkaline earth, and transition metals, and sometimes broadened to include metalloids like boron, silicon, and selenium, as well. Aside from bonds to organyl fragments or molecules, bonds to 'inorganic' carbon, like carbon monoxide, cyanide, or carbide, are generally considered to be organometallic as well. Some related compounds such as transition metal hydrides and metal phosphine complexes are often included in discussions of organometallic compounds, though strictly speaking, they are not necessarily organometallic. The related but distinct term "metalorganic compound" refers to metal-containing compounds lacking direct metal-carbon bonds but which contain organic ligands. Metal β-diketonates, alkoxides, dialkylamides, and metal phosphine complexes are representative members of this class. The field of organometallic chemistry combines aspects of traditional inorganic and organic chemistry.

<span class="mw-page-title-main">Trimethylaluminium</span> Chemical compound

Trimethylaluminium is one of the simplest examples of an organoaluminium compound. Despite its name it has the formula Al2(CH3)6 (abbreviated as Al2Me6 or TMA), as it exists as a dimer. This colorless liquid is pyrophoric. It is an industrially important compound, closely related to triethylaluminium.

Organopalladium chemistry is a branch of organometallic chemistry that deals with organic palladium compounds and their reactions. Palladium is often used as a catalyst in the reduction of alkenes and alkynes with hydrogen. This process involves the formation of a palladium-carbon covalent bond. Palladium is also prominent in carbon-carbon coupling reactions, as demonstrated in tandem reactions.

<span class="mw-page-title-main">Organoboron chemistry</span> Study of compounds containing a boron-carbon bond

Organoboron chemistry or organoborane chemistry studies organoboron compounds, also called organoboranes. These chemical compounds combine boron and carbon; typically, they are organic derivatives of borane (BH3), as in the trialkyl boranes.

<span class="mw-page-title-main">Diisobutylaluminium hydride</span> Chemical compound

Diisobutylaluminium hydride (DIBALH, DIBAL, DIBAL-H or DIBAH) is a reducing agent with the formula (i-Bu2AlH)2, where i-Bu represents isobutyl (-CH2CH(CH3)2). This organoaluminium compound is a reagent in organic synthesis.

<span class="mw-page-title-main">Schwartz's reagent</span> Chemical compound

Schwartz's reagent is the common name for the organozirconium compound with the formula (C5H5)2ZrHCl, sometimes called zirconocene hydrochloride or zirconocene chloride hydride, and is named after Jeffrey Schwartz, a chemistry professor at Princeton University. This metallocene is used in organic synthesis for various transformations of alkenes and alkynes.

<span class="mw-page-title-main">Organozinc chemistry</span>

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

A carbometallation is any reaction where a carbon-metal bond reacts with a carbon-carbon π-bond to produce a new carbon-carbon σ-bond and a carbon-metal σ-bond. The resulting carbon-metal bond can undergo further carbometallation reactions or it can be reacted with a variety of electrophiles including halogenating reagents, carbonyls, oxygen, and inorganic salts to produce different organometallic reagents. Carbometallations can be performed on alkynes and alkenes to form products with high geometric purity or enantioselectivity, respectively. Some metals prefer to give the anti-addition product with high selectivity and some yield the syn-addition product. The outcome of syn and anti- addition products is determined by the mechanism of the carbometallation.

In organic chemistry, the Kumada coupling is a type of cross coupling reaction, useful for generating carbon–carbon bonds by the reaction of a Grignard reagent and an organic halide. The procedure uses transition metal catalysts, typically nickel or palladium, to couple a combination of two alkyl, aryl or vinyl groups. The groups of Robert Corriu and Makoto Kumada reported the reaction independently in 1972.

<span class="mw-page-title-main">Organozirconium and organohafnium chemistry</span>

Organozirconium chemistry is the science of exploring the properties, structure, and reactivity of organozirconium compounds, which are organometallic compounds containing chemical bonds between carbon and zirconium. Organozirconium compounds have been widely studied, in part because they are useful catalysts in Ziegler-Natta polymerization.

In organometallic chemistry, a migratory insertion is a type of reaction wherein two ligands on a metal complex combine. It is a subset of reactions that very closely resembles the insertion reactions, and both are differentiated by the mechanism that leads to the resulting stereochemistry of the products. However, often the two are used interchangeably because the mechanism is sometimes unknown. Therefore, migratory insertion reactions or insertion reactions, for short, are defined not by the mechanism but by the overall regiochemistry wherein one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

Zirconocene dichloride is an organozirconium compound composed of a zirconium central atom, with two cyclopentadienyl and two chloro ligands. It is a colourless diamagnetic solid that is somewhat stable in air.

<span class="mw-page-title-main">Organocobalt chemistry</span> Chemistry of compounds with a carbon to cobalt bond

Organocobalt chemistry is the chemistry of organometallic compounds containing a carbon to cobalt chemical bond. Organocobalt compounds are involved in several organic reactions and the important biomolecule vitamin B12 has a cobalt-carbon bond. Many organocobalt compounds exhibit useful catalytic properties, the preeminent example being dicobalt octacarbonyl.

<span class="mw-page-title-main">Diethylaluminium chloride</span> Chemical compound

Diethylaluminium chloride, abbreviated DEAC, is an organoaluminium compound. Although often given the chemical formula (C2H5)2AlCl, it exists as a dimer, [(C2H5)2AlCl]2 It is a precursor to Ziegler-Natta catalysts employed for the production of polyolefins. The compound is also a Lewis acid, useful in organic synthesis. The compound is a colorless waxy solid, but is usually handled as a solution in hydrocarbon solvents. It is highly reactive, even pyrophoric.

Reactions of alkenyl- and alkynylaluminium compounds involve the transfer of a nucleophilic alkenyl or alkynyl group attached to aluminium to an electrophilic atom. Stereospecific hydroalumination, carboalumination, and terminal alkyne metalation are useful methods for generation of the necessary alkenyl- and alkynylalanes.

<span class="mw-page-title-main">Ethylaluminium sesquichloride</span> Chemical compound

Ethylaluminium sesquichloride, also called EASC, is an industrially important organoaluminium compound used primarily as a precursor to triethylaluminium and as a catalyst component in Ziegler–Natta type systems for olefin and diene polymerizations. Other applications include use in alkylation reactions and as a catalyst component in linear oligomerization and cyclization of unsaturated hydrocarbons. EASC is a colourless liquid, spontaneously combustible in air and reacts violently when in contact with water and many other compounds.

<span class="mw-page-title-main">Triisobutylaluminium</span> Chemical compound

Triisobutylaluminium (TiBA) is an organoaluminium compound with the formula Al(CH2CH(CH3)2)3. This colorless pyrophoric liquid is mainly used to make linear primary alcohols and α-olefins.

The zirconium-catalyzed asymmetric carbo-alumination reaction was developed by Nobel laureate Ei-ichi Negishi. It facilitates the chiral functionalization of alkenes using organoaluminium compounds under the influence of chiral bis-indenylzirconium catalysts. In a first step the alkene inserts into an Al-C bond of the reagent, forming a new chiral organoaluminium compound in which the aluminium atom occupies the lesser hindered position. This intermediate is usually oxidized by oxygen to form the corresponding chiral alcohol. The reaction can also be applied to dienes, where the least sterically hindered double bond is attacked selectively.

<span class="mw-page-title-main">Cyclopentadienyliron dicarbonyl dimer</span> Chemical compound

Cyclopentadienyliron dicarbonyl dimer is an organometallic compound with the formula [(η5-C5H5)Fe(CO)2]2, often abbreviated to Cp2Fe2(CO)4, [CpFe(CO)2]2 or even Fp2, with the colloquial name "fip dimer". It is a dark reddish-purple crystalline solid, which is readily soluble in moderately polar organic solvents such as chloroform and pyridine, but less soluble in carbon tetrachloride and carbon disulfide. Cp2Fe2(CO)4 is insoluble in but stable toward water. Cp2Fe2(CO)4 is reasonably stable to storage under air and serves as a convenient starting material for accessing other Fp (CpFe(CO)2) derivatives (described below).

In organic chemistry, the Ziegler process is a method for producing fatty alcohols from ethylene using an organoaluminium compound. The reaction produces linear primary alcohols with an even numbered carbon chain. The process uses an aluminum compound to oligomerize ethylene and allow the resulting alkyl group to be oxygenated. The usually targeted products are fatty alcohols, which are otherwise derived from natural fats and oils. Fatty alcohols are used in food and chemical processing. They are useful due to their amphipathic nature. The synthesis route is named after Karl Ziegler, who described the process in 1955.

References

  1. D. F. Shriver; P. W. Atkins (2006). Inorganic Chemistry. Oxford University Press. ISBN   978-0199264636.
  2. M. Witt; H. W. Roesky (2000). "Organoaluminum chemistry at the forefront of research and development" (PDF). Curr. Sci. 78 (4): 410. Archived from the original (PDF) on 2014-10-06.
  3. Hallwachs, W.; Schafarik, A. (1859). "Ueber die Verbindungen der Erdmetalle mit organischen Radicalen". Liebigs Ann. Chem. 109 (2): 206–209. doi:10.1002/jlac.18591090214.
  4. Elschenbroich, C. (2006). Organometallics (3rd ed.). Weinheim: Wiley-VCH. ISBN   978-3-527-29390-2.
  5. Cotton, Frank Albert; Wilkinson, Geoffrey (1980). Advanced Inorganic Chemistry. p. 343. ISBN   978-0-471-02775-1.
  6. Uhl, W. (2004). Organoelement Compounds Possessing Al---Al, Ga---Ga, In---In, and Tl---Tl Single Bonds. Advances in Organometallic Chemistry. Vol. 51. pp. 53–108. doi:10.1016/S0065-3055(03)51002-4. ISBN   9780120311514.
  7. Michael J. Krause, Frank Orlandi, Alfred T. Saurage and Joseph R. Zietz "Aluminum Compounds, Organic" in Ullmann's Encyclopedia of Industrial Chemistry, 2005, Wiley-VCH, Weinheim. doi : 10.1002/14356007.a01_543
  8. Barry M. Trost; Martin F. Semmelhack; Ian Fleming (1992). Comprehensive Organic Synthesis: Additions to and substitutions at C-C[pi]-Bonds. Pergamon. ISBN   9780080405957.
  9. Negishi, Ei-ichi (2011). "Discovery of ZACA reaction : Zr-catalyzed asymmetric carboalumination of alkenes". Arkivoc. 2011 (viii): 34–53. doi: 10.3998/ark.5550190.0012.803 . hdl: 2027/spo.5550190.0012.803 .
  10. Negishi, Ei-ichi; Wang, Guangwei; Rao, Honghua; Xu, Zhaoqing (2010-05-14). "Alkyne Elementometalation−Pd-Catalyzed Cross-Coupling. Toward Synthesis of All Conceivable Types of Acyclic Alkenes in High Yields, Efficiently, Selectively, Economically, and Safely: "Green" Way". The Journal of Organic Chemistry. 75 (10): 3151–3182. doi:10.1021/jo1003218. PMC   2933819 . PMID   20465291.
  11. Negishi, Ei-ichi (2002). Organometallics In Synthesis: A Manual (Ed. M. Schlosser). Chichester, West Sussex, UK: Wiley. pp. 963–975. ISBN   978-0471984160.
  12. Rand, Cynthia L.; Horn, David E. Van; Moore, Mark W.; Negishi, Eiichi (2002-05-01). "A versatile and selective route to difunctional trisubstituted (E)-alkene synthons via zirconium-catalyzed carboalumination of alkynes". The Journal of Organic Chemistry. 46 (20): 4093–4096. doi:10.1021/jo00333a041.
  13. 1 2 Yur'ev, V.P.; Kuchin, A.V.; Tolstikov, G.A. (1974). "Reaction of aluminum trialkyls with carbon dioxide". Organic and Biological Chemistry. 23 (4): 817–819. doi:10.1007/BF00923507.
  14. Ziegler, K. (1956). "Neue Entwicklungen der metallorganischen Synthese". Angew. Chem. 68 (23): 721–729. Bibcode:1956AngCh..68..721Z. doi:10.1002/ange.19560682302.
  15. Zakharkin, L.I.; Gavrilenko, V.V.; Ivanov, L.L. (1967). Zh. Obshch. Khim. 377: 992.{{cite journal}}: Missing or empty |title= (help)
  16. David W. Marshall, US patent US3168570, assigned to Continental Oil
  17. W. Uhl; B. Jana (2008). "A persistent alkylaluminum peroxide: Surprising stability of a molecule with strong reducing and oxidizing functions in close proximity". Chem. Eur. J. 14 (10): 3067–71. doi:10.1002/chem.200701916. PMID   18283706.
  18. Cameo Chemicals SDS
  19. Handling Chemicals Safely 1980. p. 929
  20. Schmidt, Roland; Griesbaum, Karl; Behr, Arno; Biedenkapp, Dieter; Voges, Heinz-Werner; Garbe, Dorothea; Paetz, Christian; Collin, Gerd; Mayer, Dieter; Höke, Hartmut (2014). "Hydrocarbons". Ullmann's Encyclopedia of Industrial Chemistry. pp. 1–74. doi:10.1002/14356007.a13_227.pub3. ISBN   9783527306732.