Alkenylaluminium compounds

Last updated

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. [1]

Contents

Introduction

Aluminium, like its congener boron, is less electronegative than carbon (Al, 1.61; C, 2.55); thus, aluminium-bound carbons in organoalanes possess partial negative charge and are nucleophilic as a result. Generally, however, organoalanes are not nucleophilic enough to transfer an organic group on their own (the exception being when carbonyl and enone acceptors are used, due to the high oxophilicity of aluminium [2] ). In most cases, nucleophilic activation of organoalanes is necessary for group transfer to take place. Like organoboranes, organoalanes possess an empty p orbital on the aluminium center that can receive electron density from an added nucleophile. The resulting negatively charged aluminate is much more nucleophilic than the neutral alane. [3]

This concept has been applied to methods for the synthesis of organic compounds from alkenyl- and alkynylalanes. The most notable applications are methods for the stereospecific synthesis of olefins. Alkenylalanes, which are easily synthesized with complete stereocontrol through alkyne hydroalumination, transfer the alkenyl unit to a variety of electrophiles. Alkynylalanes are less commonly used because alkali metal acetylides can be used for many of the same transformations as alkynylalanes. However, alkynylalanes are useful for the coupling of tertiary halides and alkynes (a reaction difficult to effect with alkali metal alkynes) [4] and for conjugate addition [5] and epoxide opening reactions. [6]

(1)

AlaneGen.png

Mechanism and stereochemistry

Hydroalumination of alkynes

Hydroalumination of alkynes may be either stereospecifically cis or trans depending on the conditions employed. When a dialkylalane such as di(isobutyl)aluminium hydride (DIBAL-H) is used, the hydrogen and aluminium delivered from the reagent end up cis in the resulting alkenylalane. [7] This stereospecificity can be explained by invoking a concerted addition of the H–Al bond across the triple bond. In the transition state, partial positive charge builds up on the carbon forming a bond to hydrogen; thus, the carbon better able to stabilize a positive charge becomes attached to hydrogen in the product alkenylalane. [8] Hydroaluminations of terminal alkynes typically produce terminal alkenylalanes as a result. Selectivity in hydroaluminations of internal alkynes is typically low, unless an electronic bias exists in the substrate (such as a phenyl ring in conjugation with the alkyne). [9]

(2)

AlaneMech1.png

Stereospecific trans hydroalumination is possible through the use of lithium aluminium hydride. The mechanism of this transformation involves the addition of hydride to the carbon less able to stabilize the developing negative charge (viz., in the β position to an electron-withdrawing group). [10] Coordination of aluminium to the resulting trans vinyl carbanion leads to the observed trans configuration of the product. [11]

(3)

AlaneMech2.png

Reactions of alkenyl- and alkynylalanes and aluminates

Neutral alanes are not nucleophilic enough to deliver organic groups to electrophilic substrates. However, upon activation by a nucleophile, the resulting aluminates are highly nucleophilic and add to electrophiles with retention of configuration at the migration carbon. Thus, stereospecific hydroalumination followed by nucleophilic attack provides a method for the stereospecific synthesis of olefins from alkynes. [3]

(4)

AlaneMech3.png

Scope and limitations

Because unsaturated alanes are oxygen- and moisture-sensitive, they are most often prepared for immediate use without isolation. However, the method of preparation determines the configuration of the intermediate unsaturated alane, which is directly related to the configuration of the product (transfer of the alkenyl group occurs with retention of configuration). Thus, an understanding of available hydroalumination methods is important for the study of reactions of unsaturated alanes. This second describes the most common methods of hydroalumination, and subsequent chemical reactions that the resulting alkenylalanes may undergo.

Preparation of alkenylalanes

Stereospecific cis-hydroalumination is possible through the use of dialkylalanes. The most common reagent used for this purpose is di(isobutyl)aluminium hydride (DIBAL-H). Analogous to hydroboration reactions with R2BH, hydroalumination with R2AlH leads to the attachment of aluminium at the carbon less able to stabilize developing positive charge (anti-Markovnikov selectivity). [12] Metalation of terminal alkynes is a significant side reaction that occurs under these conditions. If metalation is desired, tertiary amine complexes of DIBAL-H are useful. [13]

(5)

AlaneScope1.png

The use of silyl acetylenes avoids the problem of competitive metalation of terminal alkenes. The stereoselectivity of hydroalumination can be altered through a change in solvent: tertiary amine solvents provide the cis alkenylalane and hydrocarbon solvents provide the trans isomer. [14]

(6)

AlaneScope2.png

Lithium aluminium hydride hydroaluminates alkynes to afford trans alkenylalanes. In equation (7) hydride adds to the terminal carbon, which places the developing negative charge next to the stabilizing phenyl substituent. [10]

(7)

AlaneScope3.png

Reactions of alkenylalanes

Alkenyl- and alkynylaluminates are most commonly generated through the addition of n-butyllithium to the alkenylalane. The alkenyl and alkynyl groups, which are better able to stabilize negative charge, are transferred in preference to the alkyl group. When these intermediates react with alkyl halides, functionalized olefins are produced. [15]

(8)

AlaneScope4.png

Treatment of alkenylaluminates with halogen electrophiles such as N-bromosuccinimide (NBS) and iodine leads to the formation of halogenated olefins. [16] These products are useful for cross-coupling reactions.

(9)

AlaneScope5.png

Zirconium-catalyzed carboalumination of alkynes by trimethylalane is a convenient method for accessing substituted alkenylalanes stereoselectively. [17] Upon exposure to aldehydes and ketones, alkenylalanes form secondary or tertiary allylic alcohols. Formaldehyde is a useful reagent in this context for the introduction of a hydroxymethyl unit. [18]

(10)

AlaneScope6.png

Alkynylalanes are primarily used in place of the corresponding alkali metal acetylides when the latter reagents are ineffective. The coupling of an acetylide and tertiary alkyl halide is an example of a reaction that cannot be accomplished with alkali metal acetylides, which displace halides in an SN2 fashion. The corresponding alkynylalanes are able to couple to tertiary halides via an SN1-like mechanism. [4]

(11)

AlaneScope7.png

Alkynyl- and alkenylalanes add in a conjugate fashion to enones in the s-cis conformation without nucleophilic activation. Enones locked in an s-trans conformation, such as cyclohexenone, are unreactive. The coordination of oxygen to aluminium is believed to be necessary for this reaction. [19]

(12)

AlaneScope8.png

When alkynes and dialkylalanes are combined in a 2:1 ratio, 1,3-dienes result. The aluminium-carbon bond of the initially formed alkenylalane adds across a second molecule of alkyne, forming a conjugated dienylalane. Protonolysis provides the metal-free diene product. [20]

(13)

AlaneScope9.png

Alkenyl- and alkynylalanes undergo transmetalation to a variety of metals, including boron, [21] zirconium, [21] and mercury. [22]

(14)

AlaneScope10.png

Experimental conditions and procedure

Typical conditions

Organoaluminium compounds are extremely pyrophoric and are handled only under inert atmospheres. Dialkylaluminium hydrides and lithium aluminium hydride are both available commercially; the former is available either neat or in solutions that may be standardized using known procedures. [23] Solvents and other reagents should be scrupulously dried. Workup of these reactions should employ extreme pH's (10% hydrochloric acid or 6 N sodium hydroxide), as moderate pH encourages the formation of gelatinous aluminium hydroxide, which renders product separation difficult.

Example procedure

(15)

AlaneEx.png

To 2.76 g (25.0 mmol) of 1-octyne was added 25.0 mL of a 1.07 M solution of diisobutylaluminium hydride (26.8 mmol) in n-hexane while the temperature was maintained at 25–30° by means of a water bath. The solution was stirred at room temperature for 30 minutes and then was heated at 50° for 4 hours. The resultant alkenylalane was cooled to –30°, diluted with 15 mL of dry ether, and treated with 5.35 g (30.1 mmol) of N-bromosuccinimide while keeping the temperature below –15°. The reaction mixture was gradually warmed to room temperature and stirred for 1 hour before being poured slowly into a mixture of 6 N hydrochloric acid (50 mL), n-pentane (10 mL), and ice (10 g). The layers were separated, and the aqueous phase was extracted with pentane. The combined organic extract was washed successively with 1 N sodium hydroxide, 10% sodium sulfite, and saturated aqueous sodium chloride and then was treated with a few crystals of BHT to inhibit isomerization of the alkenyl bromide. After drying over magnesium sulfate, distillation afforded 3.72 g (78%) of (E)-1-bromo-1-octene, bp 67° (5 mm), nD26 1.4617. This compound, which contained 4% of 1-bromo-1-octyne, was stored over a few crystals of BHT. 1H NMR (CDCl3): δ 0.92 (m, 3 H), 1.1-1.7 (m, 8 H), 1.90-2.34 (m, 2 H), 5.85-6.30 (m, 2 H).

Related Research Articles

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

The Simmons–Smith reaction is an organic cheletropic reaction involving an organozinc carbenoid that reacts with an alkene to form a cyclopropane. It is named after Howard Ensign Simmons, Jr. and Ronald D. Smith. It uses a methylene free radical intermediate that is delivered to both carbons of the alkene simultaneously, therefore the configuration of the double bond is preserved in the product and the reaction is stereospecific.

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.

The Corey–House synthesis is an organic reaction that involves the reaction of a lithium diorganylcuprate with an organic halide or pseudohalide to form a new alkane, as well as an ill-defined organocopper species and lithium (pseudo)halide as byproducts.

In organic chemistry, hydroboration refers to the addition of a hydrogen-boron bond to certain double and triple bonds involving carbon. This chemical reaction is useful in the organic synthesis of organic compounds.

<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.

The Negishi coupling is a widely employed transition metal catalyzed cross-coupling reaction. The reaction couples organic halides or triflates with organozinc compounds, forming carbon-carbon bonds (C-C) in the process. A palladium (0) species is generally utilized as the metal catalyst, though nickel is sometimes used. A variety of nickel catalysts in either Ni0 or NiII oxidation state can be employed in Negishi cross couplings such as Ni(PPh3)4, Ni(acac)2, Ni(COD)2 etc.

<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.

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

Aluminium hydride (also known as alane and alumane) is an inorganic compound with the formula AlH3. Alane and its derivatives are common reducing (hydride addition) reagents in organic synthesis that are used in solution at both laboratory and industrial scales. In solution—typically in etherial solvents such tetrahydrofuran or diethyl ether—aluminium hydride forms complexes with Lewis bases, and reacts selectively with particular organic functional groups (e.g., with carboxylic acids and esters over organic halides and nitro groups), and although it is not a reagent of choice, it can react with carbon-carbon multiple bonds (i.e., through hydroalumination). Given its density, and with hydrogen content on the order of 10% by weight, some forms of alane are, as of 2016, active candidates for storing hydrogen and so for power generation in fuel cell applications, including electric vehicles. As of 2006 it was noted that further research was required to identify an efficient, economical way to reverse the process, regenerating alane from spent aluminium product.

<span class="mw-page-title-main">Organocopper chemistry</span> Compound with carbon to copper bonds

Organocopper chemistry is the study of the physical properties, reactions, and synthesis of organocopper compounds, which are organometallic compounds containing a carbon to copper chemical bond. They are reagents in organic chemistry.

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

Organoaluminium chemistry is the study of compounds containing bonds between carbon and aluminium. It is one of the major themes within organometallic chemistry. 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.

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">Nozaki–Hiyama–Kishi reaction</span> Coupling reaction used in organic synthesis

The Nozaki–Hiyama–Kishi reaction is a nickel/chromium coupling reaction forming an alcohol from the reaction of an aldehyde with an allyl or vinyl halide. In their original 1977 publication, Tamejiro Hiyama and Hitoshi Nozaki reported on a chromium(II) salt solution prepared by reduction of chromic chloride by lithium aluminium hydride to which was added benzaldehyde and allyl chloride:

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.

The nitrone-olefin [3+2] cycloaddition reaction is the combination of a nitrone with an alkene or alkyne to generate an isoxazoline or isoxazolidine via a [3+2] cycloaddition process. This reaction is a 1,3-dipolar cycloaddition, in which the nitrone acts as the 1,3-dipole, and the alkene or alkyne as the dipolarophile.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur–carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

<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).

<span class="mw-page-title-main">Vinyl iodide functional group</span>

In organic chemistry, a vinyl iodide functional group is an alkene with one or more iodide substituents. Vinyl iodides are versatile molecules that serve as important building blocks and precursors in organic synthesis. They are commonly used in carbon-carbon forming reactions in transition-metal catalyzed cross-coupling reactions, such as Stille reaction, Heck reaction, Sonogashira coupling, and Suzuki coupling. Synthesis of well-defined geometry or complexity vinyl iodide is important in stereoselective synthesis of natural products and drugs.

References

  1. Zweifel, G.; Miller, J. A. Org. React. 1984, 32, 375. doi : 10.1002/0471264180.or032.02
  2. Collins, P. W.; Dajani, E. Z.; Bruhn, M. S.; Brown, C. H.; Palmer, J. H.; Pappo, R. Tetrahedron Lett.1975, 4217.
  3. 1 2 Negishi, E.-i. J. Organometal. Chem. Lib.1976, 1, 93.
  4. 1 2 Negishi, E.-i.; Baba, S. J. Am. Chem. Soc.1975, 97, 7385.
  5. Hooz, J.; Layton, R. B. J. Am. Chem. Soc.1971, 93, 7320.
  6. Danishefsky, S.; Kitahara, T.; Tsai, M.; Dynak, J. J. Org. Chem.1976, 41, 1669.
  7. Wilke, G.; Müller, H. Chem. Ber.1956, 89, 444.
  8. Bundens, J. W.; Francl, M. M. Organometallics1993, 12, 1608.
  9. Eisch, J. J.; Gopal, H.; Rhee, S.-G. J. Org. Chem.1975, 40, 2064.
  10. 1 2 Slaugh, L. H. Tetrahedron1966, 22, 1741.
  11. House, H. O. Modern Synthetic Reactions, 2nd ed., Benjamin/Cummings, Menlo Park, CA, 1972.
  12. Eisch, J. J.; Kaska, W. C. J. Am. Chem. Soc.1963, 85, 2165.
  13. Binger, P. Angew. Chem. Int. Ed. Engl.1963, 2, 686.
  14. Eisch, J. J.; Foxton, M W. J. Org. Chem.1971, 36, 3520.
  15. Zweifel, G.; Lynd, R. A. Synthesis1976, 816.
  16. Palei, B. A.; Gavrilenko, V. V.; Zakharkin, L. I. Izv. Akad. Nauk SSSR, Ser. Khim.1969, 2760 [C.A., 72, 79143s (1970)].
  17. Yoshida, T.; Negishi, E.-i. J. Am. Chem. Soc.1981, 103, 4985.
  18. Negishi, E.-i.; King, A. O.; Klima, W.; Patterson, W.; Silveira, Jr., A. J. Org. Chem.1980, 45, 2526.
  19. Hooz, J.; Layton, R. B. Can. J. Chem.1973, 51, 2098.
  20. Zweifel, G.; Snow, J. T.; Whitney, C. C. J. Am. Chem. Soc.1968, 90, 7139.
  21. 1 2 Negishi, E.-i.; Boardman, L. D. Tetrahedron Lett.1982, 3327.
  22. Negishi, E.-i.; Jadhav, K. P.; Daotien, N. Tetrahedron Lett.1982, 2085.
  23. Brown, H. C. Organic Syntheses via Boranes, Wiley, New York, 1975.