Blaise ketone synthesis

Last updated
Blaise ketone synthesis
Named afterEdmond Blaise
Reaction type Coupling reaction

The Blaise ketone synthesis (named after Edmond Blaise) is the chemical reaction of acid chlorides with organozinc compounds to give ketones.

Contents

The Blaise ketone synthesis Blaise Ketone Synthesis Scheme.png
The Blaise ketone synthesis

The reaction was claimed to bring excellent yields by Blaise, however, investigators failed to obtain better than moderate yields (50%). Thus, the reaction is particularly ineffective in forming ketones from acyl chlorides.

The reaction also works with organocuprates.

Reviews have been written.

Reaction Mechanism

The mechanism is sampled from the proposed mechanism for organocadmium compounds, given that the mechanisms are identical to one another the proposed mechanism for the reaction is the same as the one for organocadmium compounds .

Proposed Mechanism for the Blaise Ketone Synthesis Reaction Blaise ketone synthesis proposed mechanism.png
Proposed Mechanism for the Blaise Ketone Synthesis Reaction

After the oxygen forms a bond with the organozinc compound, R’ shifts to the carbonyl carbon, having chlorine act as a leaving group and removing the negative charge from zinc. The chlorine that left returns to form a bond with zinc, pushing the electrons back on to oxygen and thus forming the ketone.

Variations

Blaise-Maire reaction

The Blaise-Maire reaction is the Blaise ketone synthesis using β-hydroxy acid chlorides to give β-hydroxyketones, which are converted into α,β-unsaturated ketones using sulfuric acid.

Ketone Formation from Organocadmium Compounds

This ketone formation is an identical reaction to the Blaise ketone synthesis. Only instead of organozinc compounds, organocadmium compounds are used and produce higher yields.

See also

Related Research Articles

<span class="mw-page-title-main">Ketone</span> Organic compounds of the form >C=O

In organic chemistry, a ketone is a functional group with the structure R–C(=O)–R', where R and R' can be a variety of carbon-containing substituents. Ketones contain a carbonyl group –C(=O)–. The simplest ketone is acetone, with the formula CH3C(O)CH3. Many ketones are of great importance in biology and in industry. Examples include many sugars (ketoses), many steroids, and the solvent acetone.

The Friedel–Crafts reactions are a set of reactions developed by Charles Friedel and James Crafts in 1877 to attach substituents to an aromatic ring. Friedel–Crafts reactions are of two main types: alkylation reactions and acylation reactions. Both proceed by electrophilic aromatic substitution.

In organic chemistry, an acyl chloride is an organic compound with the functional group −C(=O)Cl. Their formula is usually written R−COCl, where R is a side chain. They are reactive derivatives of carboxylic acids. A specific example of an acyl chloride is acetyl chloride, CH3COCl. Acyl chlorides are the most important subset of acyl halides.

<span class="mw-page-title-main">Simmons–Smith reaction</span>

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.

The Claisen condensation is a carbon–carbon bond forming reaction that occurs between two esters or one ester and another carbonyl compound in the presence of a strong base, resulting in a β-keto ester or a β-diketone. It is named after Rainer Ludwig Claisen, who first published his work on the reaction in 1887.

Clemmensen reduction is a chemical reaction described as a reduction of ketones to alkanes using zinc amalgam and concentrated hydrochloric acid. This reaction is named after Erik Christian Clemmensen, a Danish chemist.

<span class="mw-page-title-main">Knorr pyrrole synthesis</span>

The Knorr pyrrole synthesis is a widely used chemical reaction that synthesizes substituted pyrroles (3). The method involves the reaction of an α-amino-ketone (1) and a compound containing an electron-withdrawing group α to a carbonyl group (2).

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

1,1'-Carbonyldiimidazole (CDI) is an organic compound with the molecular formula (C3H3N2)2CO. It is a white crystalline solid. It is often used for the coupling of amino acids for peptide synthesis and as a reagent in organic synthesis.

The Reformatsky reaction is an organic reaction which condenses aldehydes or ketones with α-halo esters using metallic zinc to form β-hydroxy-esters:

Hydrazides in organic chemistry are a class of organic compounds with the formula RNHNH2 where R is acyl (R'CO-), sulfonyl (R'SO2-), or phosphoryl (R'2P(O)-). Unlike hydrazine and alkylhydrazines, hydrazides are nonbasic owing to the inductive influence of the acyl, sulfonyl, or phosphoryl substituent.

<span class="mw-page-title-main">Nucleophilic conjugate addition</span>

Nucleophilic conjugate addition is a type of organic reaction. Ordinary nucleophilic additions or 1,2-nucleophilic additions deal mostly with additions to carbonyl compounds. Simple alkene compounds do not show 1,2 reactivity due to lack of polarity, unless the alkene is activated with special substituents. With α,β-unsaturated carbonyl compounds such as cyclohexenone it can be deduced from resonance structures that the β position is an electrophilic site which can react with a nucleophile. The negative charge in these structures is stored as an alkoxide anion. Such a nucleophilic addition is called a nucleophilic conjugate addition or 1,4-nucleophilic addition. The most important active alkenes are the aforementioned conjugated carbonyls and acrylonitriles.

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

Thiophenol is an organosulfur compound with the formula C6H5SH, sometimes abbreviated as PhSH. This foul-smelling colorless liquid is the simplest aromatic thiol. The chemical structures of thiophenol and its derivatives are analogous to phenols except the oxygen atom in the hydroxyl group (-OH) bonded to the aromatic ring is replaced by a sulfur atom. The prefix thio- implies a sulfur-containing compound and when used before a root word name for a compound which would normally contain an oxygen atom, in the case of 'thiol' that the alcohol oxygen atom is replaced by a sulfur atom.

<span class="mw-page-title-main">Favorskii rearrangement</span>

The Favorskii rearrangement is principally a rearrangement of cyclopropanones and α-halo ketones that leads to carboxylic acid derivatives. In the case of cyclic α-halo ketones, the Favorskii rearrangement constitutes a ring contraction. This rearrangement takes place in the presence of a base, sometimes hydroxide, to yield a carboxylic acid but most of the time either an alkoxide base or an amine to yield an ester or an amide, respectively. α,α'-Dihaloketones eliminate HX under the reaction conditions to give α,β-unsaturated carbonyl compounds.

The Hunsdiecker reaction is a name reaction in organic chemistry whereby silver salts of carboxylic acids react with a halogen to produce an organic halide. It is an example of both a decarboxylation and a halogenation reaction as the product has one fewer carbon atoms than the starting material and a halogen atom is introduced its place. The reaction was first demonstrated by Alexander Borodin in his 1861 reports of the preparation of methyl bromide from silver acetate. Shortly after, the approach was applied to the degradation of fatty acids in the laboratory of Adolf Lieben. However, it is named for Cläre Hunsdiecker and her husband Heinz Hunsdiecker, whose work in the 1930s developed it into a general method. Several reviews have been published, and a catalytic approach has been developed.

<span class="mw-page-title-main">Grignard reagent</span> Organometallic compounds used in organic synthesis

A Grignard reagent or Grignard compound is a chemical compound with the general formula R−Mg−X, where X is a halogen and R is an organic group, normally an alkyl or aryl. Two typical examples are methylmagnesium chloride Cl−Mg−CH3 and phenylmagnesium bromide (C6H5)−Mg−Br. They are a subclass of the organomagnesium compounds.

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 compound</span>

Organozinc compounds in organic chemistry contain carbon (C) to zinc (Zn) chemical bonds. Organozinc chemistry is the science of organozinc compounds describing their physical properties, synthesis and reactions.

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

An organocadmium compound is an organometallic compound containing a carbon to cadmium chemical bond. Organocadmium chemistry describes physical properties, synthesis, reactions and use of these compounds. Cadmium shares group 12 with zinc and mercury and their corresponding chemistries have much in common. The synthetic utility of organocadmium compounds is limited.

<span class="mw-page-title-main">Carbonyl reduction</span>

In organic chemistry, carbonyl reduction is the organic reduction of any carbonyl group by a reducing agent.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

References

    1. ^ Blaise, E. E.; Koehler, A. (1910). "Synthèse au moyen des dérivés organo-métalliques mixtes du zinc (II)". Bull. Soc. Chim. Fr. 7: 215–227.
    2. ^ Blaise, E. E. (1911). "Sur les dérivés organo-métalliques mixtes du zinc et leur emploi dans la synthèse organique". Bull. Soc. Chim. Fr. 9: 1.
    3. ^ Posner, G. H.; Whitten, C. E. (1976). "Secondary and Tertiary Alkyl Ketones from Carboxylic Acid Chlorides and Lithium Phenylthio(Alkyl)Cuprate Reagents: tert-Butyl Phenyl Ketone". Organic Syntheses . 55: 122.; Collective Volume, vol. 6, p. 248
    4. ^ Fujisawa, T.; Sato, T. (1988). "Ketones from Carboxylic Acids and Grignard Reagents: Methyl 6-Oxodecanoate". Organic Syntheses . 66: 116.; Collective Volume, vol. 8, p. 441
    5. ^ Cason, J. (1947). "The Use of Organocadmium Reagents for the Preparation of Ketones". Chem. Rev. 40 (1): 17. doi:10.1021/cr60125a002. PMID   20287882.
    6. ^ Shirley, D. A. (1954). "The Synthesis of Ketones from Acid Halides and Organometallic Compounds of Magnesium, Zinc, and Cadmium". Organic Reactions. 8: 29. doi:10.1002/0471264180.or008.02. ISBN   0471264180.
    7. ^ Blaise, E. E.; Maire, M. (1907). "Synthèse au moyen des dérivés organo-métalliques mixtes du zinc. Cétones non saturées αβ-acycliques". Compt. Rend. 145: 73-75.
    8. ^ Iqbal, M.; Baloch, I. B.; Baloch, K. (2012). "An efficient method for the preparation of benzyl γ-ketohexanoates". Chemical Papers-Slovak Academy of Sciences: 10. doi:10.2478/s11696-012-0282-8.
    9. ^ Vogel, A. I. (1989). Vogel's Textbook of Practical Organic Chemistry (PDF) (5 ed.). Prentice Hall. pp. 33–34. ISBN   0-582-46236-3.