Dicarbonyl

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Smokefoot (talk | contribs) at 12:06, 16 October 2019 (→‎1,4-Diketones: layout). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Diacetyl, the simplest diketone

A dicarbonyl is a molecule containing two carbonyl (C=O) groups. Their properties often differ from those of monocarbonyls. The simplest are symmetrical dialdehydes, diketones, and diesters. Related to the latter are dicarboxylic acids and diamides. Unsymmetrical dicarbonyls are also prevalent, e.g. keto-aldehydes, keto acids, etc.

1,2-Dicarbonyls

1,2-Dialdehydes

A single 1,2-dialdehyde exists, glyoxal, (CHO)2. Like most alkyldialdehydes, it is rarely encountered as such but instead is encountered almost exclusively as its hydrate and oligomers thereof. These derivatives often behave like the aldehyde. Glyoxal condenses readily with amines. Via such reactions, it is a precursor to many heterocycles, e.g. imidazoles.

1,2-Diketones

The principal diketone is diacetyl, also known as 2,3-butanedione, CH3C(O)C(O)CH3. These compounds are often generated by oxidation (dehydrogenation) of the diols:[1]

RCH(OH)CH(OH)R → RC(O)C(O)R + 2 H2

2,3-Butanedione, 2,3-pentanedione, and 2,3-hexanedione are found in small amounts in various foods. They are used as aroma components in alcohol-free beverages and in baked goods.[1] Benzil, (PhCO)2, is the corresponding diphenyl derivative.

A distinctive feature of 1,2-diketones is the long C-C bond linking the carbonyl groups. This bond distance is about 1.54 Å, compared to 1.45 Å for the corresponding bond in 1,3-butadiene. The effect is attributed to repulsion between the partial positive charges of the carbonyl carbon atoms.[2]

1,2-Diketones condense with many bifunctional nucleophiles, such as urea and thiourea to give heterocycles. Upon condensation with aromatic amines, they convert to the diketimine.

1,2-Diesters and diacids

Oxalic acid and its esters define this family of compounds. The diacid is produced industrially by oxidation waste sugars. It occurs naturally (as the conjugate base), notably in members of the plant species Oxalis. Condensation of the diesters with diamines gives cyclic diamides.

1,2-Keto- and formylcarboxylic acids

Pyruvic acid (CH3C(O)CO2H) is the parent 2-ketoacid. Its conjugate base, pyruvate (CH3C(O)CO2-), is a component of the citric acid cycle and product of glucose metabolism (glycolysis). The corresponding aldehyde-acid is glyoxalic acid (HC(O)CO2H).

1,3-Dicarbonyls

1,3-Dialdehydes

The parent 1,3-dialdehyde is malondialdehyde (CH2(CHO)2). Like most dialdehydes, it is rarely encountered as such. Instead it is handled almost exclusively as its hydrate, methyl acetal, and oligomers thereof. These derivatives often behave like the parent. Many 2-substituted derivatives are known. They are often prepared by alkylation of the enolate of malondialdehyde.

1,3-Diketones

An important member is acetylacetone, CH3C(O)CH2C(O)CH3. Dimedone is a cyclic 1,3-diketone. Acetylacetone is prepared industrially by the thermal rearrangement of isopropenylacetate.[1] Another cyclic 1,3-diketone is 2,2,4,4-tetramethylcyclobutanedione, which is a precursor to a useful diol.

CH2(CH3)COC(O)Me → MeC(O)CH2C(O)Me

1,3-Diketones that can tautomerize to an enol that is conjugated to the other carbonyl usually exist predominantly in the enol form, and especially when the product can be further stabilized by a six-membered ring containing a hydrogen bond. For example, the percent enol in acetylacetone, trifluoroacetyacetone, and hexafluoroacetylacetone are 85, 97, and 100%, respectively (neat, 33 °C).[3]

Scheme 1. Tautomerism of 2,4-pentanedione

Like other diketones, 1,3-diketones are versatile precursors to heterocycles. Hydrazine, for example, condenses to give pyrazoles.

The conjugate base derived from 1,3-ketones can serve as ligand s to form metal acetylacetonate coordination complexes. In the DeMayo reaction 1,3-diketones react with alkenes in a photochemical pericyclic reaction to form (substituted) 1,5-diketones.

Classically, 1,3-diketones are prepared by the Claisen condensation of a ketone with an ester.

1,3-Diesters and diacids

Malonic acid and its esters are the parent members of this class of dicarbonyls. Also common are the 2-substituted derivatives with the formula RCH(CO2R)2, which arise by C-alkylation of the conjugate base (the enolate) NaCH(CO2R)2.

1,4-Dicarbonyls

1,4-Dialdehydes

Succinaldehyde (CH2CHO)2 is the simplest and parent 1,4-dialdehyde. It hydrates readily to give 2,5-dihydroxytetrahydrofuran.

1,4-Diketones

Diketones with two methylene groups separating the carbonyl groups, also called γ-diketones, typically coexist with their enol tautomers. The preeminent member is acetonylacetone. 1,4-Diketones are useful precursors to heterocycles via the Paal–Knorr synthesis, which gives pyrroles:

Dehydration of 1,4-diketones gives furans.

This reactivity is the basis of the neurotoxicity of γ-diketones.[4] 1,4-Diketones are also precursor to furans and thiophenes. The condensation of 1,4-diketones (and related substrates) with hydrazines afford dihydropyridazines, which can be converted to pyridazines.

1,4-Diesters and diacids

Succinic acid and its esters are the parent members of this family of 1,4-dicarbonyls. Succinic acid is notable as a component in the citric acid cycle. It forms a cyclic acid anhydride, succinic anhydride.

See also

References

  1. ^ a b c Hardo Siegel; Manfred Eggersdorfer (2007). "Ketones". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. p. 16. doi:10.1002/14356007.a15_077. ISBN 978-3527306732.
  2. ^ Eriks, K.; Hayden, T. D.; Yang, S. Hsi; Chan, I. Y. (1983). "Crystal and molecular structure of biacetyl (2,3-butanedione), (H3CCO)2, at -12 and -100 °C". J. Am. Chem. Soc. 105 (12): 3940–3942. doi:10.1021/ja00350a032.
  3. ^ Jane L. Burdett; Max T. Rogers (1964). "Keto-Enol Tautomerism in β-Dicarbonyls Studied by Nuclear Magnetic Resonance Spectroscopy. I. Proton Chemical Shifts and Equilibrium Constants of Pure Compounds". J. Am. Chem. Soc. 86: 2105–2109. doi:10.1021/ja01065a003.
  4. ^ Stephen R Clough; Leyna Mulholland (2005), "Hexane", Encyclopedia of Toxicology, vol. 2 (2nd ed.), Elsevier, pp. 522–525