Dicarbonyl

Last updated
General structure of 1,2-, 1,3-, and 1,4-dicarbonyls Dicarbonyl structure.png
General structure of 1,2-, 1,3-, and 1,4-dicarbonyls

In organic chemistry, a dicarbonyl is a molecule containing two carbonyl (C=O) groups. Although this term could refer to any organic compound containing two carbonyl groups, it is used more specifically to describe molecules in which both carbonyls are in close enough proximity that their reactivity is changed, such as 1,2-, 1,3-, and 1,4-dicarbonyls. Their properties often differ from those of monocarbonyls, and so they are usually considered functional groups of their own. These compounds can have symmetrical or unsymmetrical substituents on each carbonyl, and may also be functionally symmetrical (dialdehydes, diketones, diesters, etc.) or unsymmetrical (keto-esters, keto-acids, etc.).

Contents

1,2-Dicarbonyls

1,2-Dialdehyde

The only 1,2-dialdehyde is glyoxal, (CHO)2. Like many alkyldialdehydes, glyoxal is encountered almost exclusively as its hydrate and oligomers thereof. These derivatives often behave equivalently to the aldehydes since hydration is reversible. Glyoxal condenses readily with amines. Via such reactions, it is a precursor to many heterocycles, e.g. imidazoles.

Diacetyl, the simplest diketone Diacetyl.svg
Diacetyl, the simplest diketone

1,2-Diketones

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

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. Condensation with aromatic amines gives diketimine ((RC=NAr)2).

In the cases of 1,2-cyclohexanedione and 1,2-cyclopentanedione, the enol is about 1-3 kcal/mol more stable than the diketo form. [3]

ortho-Quinone, C4H4(CO)2, is the parent of a large family of 1,2-diketones.

1,2-Ketoaldehydes

Methylglyoxal, a well-known 2-oxoaldehyde Methylglyoxal.png
Methylglyoxal, a well-known 2-oxoaldehyde

A well-known compound of this class is methylglyoxal, CH3C(O)CHO, also known as pyruvaldehyde. These compounds are also known as 2-oxoaldehydes [4] or α-ketoaldehydes.

1,2-Diesters and diacids

Oxalic acid and its esters define this family of compounds. The diacid is produced industrially by oxidation of 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.

α-Keto- and formylcarboxylic acids

Important keto-acids: pyruvic acid (top), acetoacetic acid, and levulinic acid (bottom). Ketocarboxylic Acids General Formulae V.1.svg
Important keto-acids: pyruvic acid (top), acetoacetic acid, and levulinic acid (bottom).

α-Keto-acids and -esters are well known. Pyruvic acid (CH3C(O)CO2H) is the parent α-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), a β-dicarbonyl. 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

1,3-Diketones are also called β-diketones. An important member is acetylacetone, CH3C(O)CH2C(O)CH3. Dimedone is a cyclic 1,3-diketone. 1,3-Indandione is the cyclic 1,3-diketone fused to a benzene ring. 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.

1,3-Diketones often tautomerize to an enol and ketol. They usually exist predominantly in the enol form [ citation needed ]. The percent enol in acetylacetone, trifluoroacetylacetone, and hexafluoroacetylacetone are 85, 97, and 100%, respectively (neat, 33 °C). [5] Cyclic 1,3-diketones, such as 1,3-cyclohexanedione and dimedone, similarly exist significantly in the enol form.

AcacH.svg

Like other diketones, 1,3-diketones are versatile precursors to heterocycles. 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.

β-Keto-esters

β-Keto-esters arise readily by the condensation of a pair of esters. A well known example is ethyl acetoacetate (although it is prepared by ethanolysis of ketene).

1,4-Dicarbonyls

1,4-Dialdehydes

Succinaldehyde (CH2CHO)2 is the simplest and parent 1,4-dialdehyde. The aromatic derivative is phthalaldehyde.

1,4-Diketones

1,4-Benzoquinone is a cyclic 1,4-doubly unsaturated diketone. P-Benzochinon.svg
1,4-Benzoquinone is a cyclic 1,4-doubly unsaturated diketone.

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:

Paal-Knorr Pyrrole Synthesis.svg

This reactivity is the basis of the neurotoxicity of γ-diketones. [6] 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.

para-quinone, C4H4(CO)2, is the parent of a large family of 1,4-diketones.

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. Unsaturated members include maleic and fumaric acids and their esters.

1,5-Dicarbonyls

1,5-Dialdehydes

Glutaraldehyde (CH2)3(CHO)2 is the simplest and parent 1,5-dialdehyde. It hydrates readily. The aromatic analogue is isophthalaldehyde. [7]

1,5-Diketones

These diketones have three methylene groups separating the carbonyl groups.

1,5-Diesters and diacids

Glutaric acid (CH2)3(CO2H)2 is the parent 1,5-diacid.

Hydration and cyclization

2,5-Dihydroxytetrahydrofuran, the hydrated form of succinaldehyde. SuccinaldehydeHydrate.png
2,5-Dihydroxy­tetrahydro­furan, the hydrated form of succinaldehyde.

Small aldehydes tend to hydrate. Hydration is prevalent for dialdehydes. Glyoxal forms a series of cyclic hydrates. Succinaldehyde hydrates readily to give 2,5-dihydroxy­tetrahydro­furan. The aromatic phthalaldehyde also forms hydrated.

Orthophthalaldehyde and hydrated forms 001.png

Similar hydration and cyclization equilibria apply to maleic dialdehyde, [8] [9] glutaraldehyde, and adipaldehyde.

Safety

A number of dicarbonyl compounds are bioactive. Diacetyl is known to cause the lung disease bronchiolitis obliterans in those individuals exposed to it in an occupational setting. [10] Dialdehydes, e.g. glutaraldehyde and malonaldehyde, are fixatives or sterilizers.

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 an organic compound 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 (CH3)2CO. Many ketones are of great importance in biology and in industry. Examples include many sugars (ketoses), many steroids, and the solvent acetone.

<span class="mw-page-title-main">Aldehyde</span> Organic compound containing the functional group R−CH=O

In organic chemistry, an aldehyde is an organic compound containing a functional group with the structure R−CH=O. The functional group itself can be referred to as an aldehyde but can also be classified as a formyl group. Aldehydes are a common motif in many chemicals important in technology and biology.

A diol is a chemical compound containing two hydroxyl groups. An aliphatic diol is also called a glycol. This pairing of functional groups is pervasive, and many subcategories have been identified.

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

Diethyl malonate, also known as DEM, is the diethyl ester of malonic acid. It occurs naturally in grapes and strawberries as a colourless liquid with an apple-like odour, and is used in perfumes. It is also used to synthesize other compounds such as barbiturates, artificial flavourings, vitamin B1, and vitamin B6.

<span class="mw-page-title-main">Enol</span> Organic compound with a C=C–OH group

In organic chemistry, alkenols are a type of reactive structure or intermediate in organic chemistry that is represented as an alkene (olefin) with a hydroxyl group attached to one end of the alkene double bond. The terms enol and alkenol are portmanteaus deriving from "-ene"/"alkene" and the "-ol" suffix indicating the hydroxyl group of alcohols, dropping the terminal "-e" of the first term. Generation of enols often involves deprotonation at the α position to the carbonyl group—i.e., removal of the hydrogen atom there as a proton H+. When this proton is not returned at the end of the stepwise process, the result is an anion termed an enolate. The enolate structures shown are schematic; a more modern representation considers the molecular orbitals that are formed and occupied by electrons in the enolate. Similarly, generation of the enol often is accompanied by "trapping" or masking of the hydroxy group as an ether, such as a silyl enol ether.

A sigmatropic reaction in organic chemistry is a pericyclic reaction wherein the net result is one σ-bond is changed to another σ-bond in an uncatalyzed intramolecular reaction. The name sigmatropic is the result of a compounding of the long-established sigma designation from single carbon–carbon bonds and the Greek word tropos, meaning turn. In this type of rearrangement reaction, a substituent moves from one part of a π-bonded system to another part in an intramolecular reaction with simultaneous rearrangement of the π system. True sigmatropic reactions are usually uncatalyzed, although Lewis acid catalysis is possible. Sigmatropic reactions often have transition-metal catalysts that form intermediates in analogous reactions. The most well-known of the sigmatropic rearrangements are the [3,3] Cope rearrangement, Claisen rearrangement, Carroll rearrangement, and the Fischer indole synthesis.

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

Acetylacetone is an organic compound with the chemical formula CH3−C(=O)−CH2−C(=O)−CH3. It is classified as a 1,3-diketone. It exists in equilibrium with a tautomer CH3−C(=O)−CH=C(−OH)−CH3. The mixture is a colorless liquid. These tautomers interconvert so rapidly under most conditions that they are treated as a single compound in most applications. Acetylacetone is a building block for the synthesis of many coordination complexes as well as heterocyclic compounds.

<span class="mw-page-title-main">Michael addition reaction</span> Reaction in organic chemistry

In organic chemistry, the Michael reaction or Michael 1,4 addition is a reaction between a Michael donor and a Michael acceptor to produce a Michael adduct by creating a carbon-carbon bond at the acceptor's β-carbon. It belongs to the larger class of conjugate additions and is widely used for the mild formation of carbon-carbon bonds.

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. The reaction produces 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. The reaction has often been displaced by diketene-based chemistry, which affords acetoacetic esters.

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

Meldrum's acid or 2,2-dimethyl-1,3-dioxane-4,6-dione is an organic compound with formula C6H8O4. Its molecule has a heterocyclic core with four carbon and two oxygen atoms; the formula can also be written as [−O−(C 2)−O−(C=O)−(CH2)−(C=O)−].

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

The organic compound ethyl acetoacetate (EAA) is the ethyl ester of acetoacetic acid. It is a colorless liquid. It is widely used as a chemical intermediate in the production of a wide variety of compounds. It is used as a flavoring for food.

The benzilic acid rearrangement is formally the 1,2-rearrangement of 1,2-diketones to form α-hydroxy–carboxylic acids using a base. This reaction receives its name from the reaction of benzil with potassium hydroxide to form benzilic acid. First performed by Justus von Liebig in 1838, it is the first reported example of a rearrangement reaction. It has become a classic reaction in organic synthesis and has been reviewed many times before. It can be viewed as an intramolecular redox reaction, as one carbon center is oxidized while the other is reduced.

The Stetter reaction is a reaction used in organic chemistry to form carbon-carbon bonds through a 1,4-addition reaction utilizing a nucleophilic catalyst. While the related 1,2-addition reaction, the benzoin condensation, was known since the 1830s, the Stetter reaction was not reported until 1973 by Dr. Hermann Stetter. The reaction provides synthetically useful 1,4-dicarbonyl compounds and related derivatives from aldehydes and Michael acceptors. Unlike 1,3-dicarbonyls, which are easily accessed through the Claisen condensation, or 1,5-dicarbonyls, which are commonly made using a Michael reaction, 1,4-dicarbonyls are challenging substrates to synthesize, yet are valuable starting materials for several organic transformations, including the Paal–Knorr synthesis of furans and pyrroles. Traditionally utilized catalysts for the Stetter reaction are thiazolium salts and cyanide anion, but more recent work toward the asymmetric Stetter reaction has found triazolium salts to be effective. The Stetter reaction is an example of umpolung chemistry, as the inherent polarity of the aldehyde is reversed by the addition of the catalyst to the aldehyde, rendering the carbon center nucleophilic rather than electrophilic.

In organic chemistry, the Paal–Knorr synthesis is a reaction used to synthesize substituted furans, pyrroles, or thiophenes from 1,4-diketones. It is a synthetically valuable method for obtaining substituted furans and pyrroles, which are common structural components of many natural products. It was initially reported independently by German chemists Carl Paal and Ludwig Knorr in 1884 as a method for the preparation of furans, and has been adapted for pyrroles and thiophenes. Although the Paal–Knorr synthesis has seen widespread use, the mechanism wasn't fully understood until it was elucidated by V. Amarnath et al. in the 1990s.

<span class="mw-page-title-main">Gould–Jacobs reaction</span> Gould-Jacobs reaction explained

The Gould–Jacobs reaction is an organic synthesis for the preparation of quinolines and 4‐hydroxyquinoline derivatives. The Gould–Jacobs reaction is a series of reactions. The series of reactions begins with the condensation/substitution of an aniline with alkoxy methylenemalonic ester or acyl malonic ester, producing anilidomethylenemalonic ester. Then through a 6 electron cyclization process, 4-hydroxy-3-carboalkoxyquinoline is formed, which exist mostly in the 4-oxo form. Saponification results in the formation of an acid. This step is followed by decarboxylation to give 4-hydroxyquinoline. The Gould–Jacobs reaction is effective for anilines with electron‐donating groups at the meta‐position.

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

Succinaldehyde or succindialdehyde is an organic compound with the formula (CH2CHO)2. Typical of other dialdehydes, succinaldehyde is highly reactive and is rarely observed as the dialdehyde. Usually, it is handled as the hydrates or methanol-derived acetal. It is a precursor to tropinone. Succinaldehyde can be used as a crosslinking agent for proteins, but it is less widely used than the related dialdehyde glutaraldehyde.

Diimines are organic compounds containing two imine (RCH=NR') groups. Common derivatives are 1,2-diketones and 1,3-diimines. These compounds are used as ligands and as precursors to heterocycles. Diimines are prepared by condensation reactions where a dialdehyde or diketone is treated with amine and water is eliminated. Similar methods are used to prepare Schiff bases and oximes.

<span class="mw-page-title-main">2,3,4-Pentanetrione</span> Chemical compound

2,3,4-Pentanetrione (or IUPAC name pentane-2,3,4-trione, triketopentane or dimethyl triketone) is the simplest linear triketone, a ketone with three C=O groups. It is an organic molecule with formula CH3COCOCOCH3.

<span class="mw-page-title-main">1,1,1-Trifluoroacetylacetone</span> Chemical compound

1,1,1-Trifluoroacetylacetone is the organofluorine compound with the formula CF3C(O)CH2C(O)CH3. It is a colorless liquid. Like other 1,3-diketones, it is used as a precursor to heterocycles, e.g. pyrazoles, and metal chelates. It is prepared by condensation of esters of trifluoroacetic acid with acetone.

α,β-Unsaturated carbonyl compound Functional group of organic compounds

α,β-Unsaturated carbonyl compounds are organic compounds with the general structure (O=CR)−Cα=Cβ-R. Such compounds include enones and enals, but also carboxylic acids and the corresponding esters and amides. In these compounds, the carbonyl group is conjugated with an alkene. Unlike the case for carbonyls without a flanking alkene group, α,β-unsaturated carbonyl compounds are susceptible to attack by nucleophiles at the β-carbon. This pattern of reactivity is called vinylogous. Examples of unsaturated carbonyls are acrolein (propenal), mesityl oxide, acrylic acid, and maleic acid. Unsaturated carbonyls can be prepared in the laboratory in an aldol reaction and in the Perkin reaction.

References

  1. 1 2 3 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. Jana, Kalyanashis; Ganguly, Bishwajit (2018). "DFT Study to Explore the Importance of Ring Size and Effect of Solvents on the Keto–Enol Tautomerization Process of α- and β-Cyclodiones". ACS Omega. 3 (7): 8429–8439. doi:10.1021/acsomega.8b01008. PMC   6644555 . PMID   31458971.
  4. "CHEBI:27659 - 2-oxo aldehyde". Chemical Entities of Biological Interest (ChEBI).
  5. 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.
  6. Stephen R Clough; Leyna Mulholland (2005), "Hexane", Encyclopedia of Toxicology, vol. 2 (2nd ed.), Elsevier, pp. 522–525
  7. Ackerman, J. H.; Surrey, A. R. (1967). "Isophthalaldehyde". Organic Syntheses. 47: 76. doi:10.15227/orgsyn.047.0076.
  8. Hardy, P. M.; Nicholls, A. C.; Rydon, H. N. (1972). "The Hydration and Polymerisation of Succinaldehyde, Glutaraldehyde, and Adipaldehyde". Journal of the Chemical Society, Perkin Transactions 2 (15): 2270. doi:10.1039/P29720002270.
  9. D. M. Burness (1960). "2,5-Dihydro-2,5-Dimethoxyfuran". Organic Syntheses. 40: 29. doi:10.15227/orgsyn.040.0029.
  10. Kreiss, Kathleen (August 2017). "Recognizing occupational effects of diacetyl: What can we learn from this history?". Toxicology. 388: 48–54. doi:10.1016/j.tox.2016.06.009. PMC   5323392 . PMID   27326900.