Meldrum's acid

Last updated
Meldrum's acid
Meldrum's acid.png
Names
Preferred IUPAC name
2,2-Dimethyl-1,3-dioxane-4,6-dione
Other names
Isopropylidene malonate
Identifiers
3D model (JSmol)
ChemSpider
ECHA InfoCard 100.016.358 OOjs UI icon edit-ltr-progressive.svg
PubChem CID
UNII
  • InChI=1S/C6H8O4/c1-6(2)9-4(7)3-5(8)10-6/h3H2,1-2H3 X mark.svgN
    Key: GXHFUVWIGNLZSC-UHFFFAOYSA-N X mark.svgN
  • InChI=1/C6H8O4/c1-6(2)9-4(7)3-5(8)10-6/h3H2,1-2H3
    Key: GXHFUVWIGNLZSC-UHFFFAOYAM
  • InChI=1/C6H8O4/c1-6(2)9-4(7)3-5(8)10-6/h3H2,1-2H3
    Key: GXHFUVWIGNLZSC-UHFFFAOYAM
  • O=C1OC(OC(=O)C1)(C)C
Properties
C6H8O4
Molar mass 144.126 g·mol−1
Melting point 94 to 95 °C (201 to 203 °F; 367 to 368 K) (decomposes) [1]
Acidity (pKa)4.97
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

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(CH3)2)−O−(C=O)−(CH2)−(C=O)−].

Contents

It is a crystalline colorless solid, sparingly soluble in water. It decomposes on heating with release of carbon dioxide and acetone. [2] [3]

Properties

Acidity

The compound can easily lose a hydrogen ion from the methylene (CH2) in the ring (carbon 5); which creates a double bond between it and one of the adjacent carbons (number 4 or 6), and a negative charge in the corresponding oxygen. The resulting anion [C6H7O4] is stabilized by resonance between the two alternatives, so that the double bond is delocalized and each oxygen in the carbonyls has a formal charge of −1/2.

Meldrum-anion-resonance.png

The ionization constant pKa is 4.97; which makes it behave as a monobasic acid even though it contains no carboxylic acid groups. [2] In this and other properties, the compound resembles dimedone and barbituric acid. However, while dimedone exists in solution predominantly as the mono-enol tautomer, Meldrum's acid is almost entirely as the diketone form. [2]

The unusually high acidity of this compound was long considered anomalous—it is 8 orders of magnitude more acidic than the closely related compound dimethyl malonate. In 2004, Ohwada and coworkers determined that the energy-minimizing conformation structure of the compound places the alpha proton's σ*CH orbital in the proper geometry to align with the π*CO, so that the ground state poses unusually strong destabilization of the C-H bond. [4]

Preparation

Original synthesis

The compound was first made by Meldrum by a condensation reaction of acetone with malonic acid in acetic anhydride and sulfuric acid. [3]

Meldrum-acid-original-synthesis.png

Alternative syntheses

As an alternative to its original preparation, Meldrum's acid can be synthesized from malonic acid, isopropenyl acetate (an enol derivative of acetone), and catalytic sulfuric acid.

A third route is the reaction of carbon suboxide C3O2 with acetone in the presence of oxalic acid. [2]

Uses

Like malonic acid and its ester derivatives, and other 1,3-dicarbonyl compounds, Meldrum's acid can and serve as a reactant for a variety of nucleophilic reactions.

Alkylation and acylation

The acidity of carbon 5 (between the two carbonyl groups) allows simple derivatization of Meldrum's acid at this position, through reactions such as alkylation and acylation. For example, deprotonation and reaction with a simple alkyl halide (R−Cl) attaches the alkyl group (R−) at that position:

Meldrum-anion-alkylation.png

The analogous reaction with an acyl chloride (R−(C=O)−Cl) attaches the acyl (R−(C=O)−) instead:

Meldrum-anion-acylation.png

These two reactions allow Meldrum's acid to serve as a starting scaffold for the synthesis of many different structures with various functional groups. The alkylated products can be further manipulated to produce various amide and ester compounds. Heating the acyl product in the presence of an alcohol leads to ester exchange and decarboxylation in a process similar to the malonic ester synthesis. The reactive nature of the cyclic-diester allows good reactivity even for alcohols as hindered as t-butanol, [5] and this reactivity of Meldrum's acid and it's derivatives has been used to develop a range of reactions. [6] [7] [8] [9] Ketoesters formed from the reaction of alcohols with Meldrum's acid derivatives are useful in the Knorr pyrrole synthesis.

Synthesis of ketenes

At temperatures greater than 200 °C [10] Meldrum's acid undergoes a pericyclic reaction that releases acetone and carbon dioxide and produces a highly reactive ketene compound: [11]

Meldrum-derivative-pyrolysis.png

These ketenes can be isolated using flash vacuum pyrolysis (FVP). Ketenes are highly electrophilic and can undergo addition reaction with a range of other chemicals, particularly ketene cycloadditions, or dimerisation to diketene. With this approach it is possible to form new C–C bonds, rings, amides, esters, and acids:

Ketene-amine-addition.png            Ketene-enol-addition.png            Ketene-water-addition.png

Alternately, the pyrolysis can be performed in solution, to obtain the same results without isolating the ketene, in a one-pot reaction. The ability to form such diverse products makes Meldrum's acid a very useful reagent for synthetic chemists. [12] [13] [14]


History

Incorrect structure. Meldrum-acid-wrong-structure.png
Incorrect structure.

The compound is named after Andrew Norman Meldrum who reported its synthesis in 1908. [3] He misidentified its structure as a β-lactone of β-hydroxyisopropylmalonic acid; the correct structure, the bislactone of 1,3-dioxane was reported in 1948. [15]

Related Research Articles

<span class="mw-page-title-main">Ester</span> Compound derived from an acid

In chemistry, an ester is a compound derived from an acid in which the hydrogen atom (H) of at least one acidic hydroxyl group of that acid is replaced by an organyl group. Analogues derived from oxygen replaced by other chalcogens belong to the ester category as well. According to some authors, organyl derivatives of acidic hydrogen of other acids are esters as well, but not according to the IUPAC.

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

In organic chemistry, a ketene is an organic compound of the form RR'C=C=O, where R and R' are two arbitrary monovalent chemical groups. The name may also refer to the specific compound ethenone H2C=C=O, the simplest ketene.

<span class="mw-page-title-main">Acyl group</span> Chemical group (R–C=O)

In chemistry, an acyl group is a moiety derived by the removal of one or more hydroxyl groups from an oxoacid, including inorganic acids. It contains a double-bonded oxygen atom and an organyl group or hydrogen in the case of formyl group. In organic chemistry, the acyl group is usually derived from a carboxylic acid, in which case it has the formula R−C(=O)−, where R represents an organyl group or hydrogen. Although the term is almost always applied to organic compounds, acyl groups can in principle be derived from other types of acids such as sulfonic acids and phosphonic acids. In the most common arrangement, acyl groups are attached to a larger molecular fragment, in which case the carbon and oxygen atoms are linked by a double bond.

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.

<span class="mw-page-title-main">Dicarbonyl</span> Molecule containing two adjacent C=O groups

In organic chemistry, a dicarbonyl is a molecule containing two carbonyl 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 or unsymmetrical.

<span class="mw-page-title-main">Malonic acid</span> Carboxylic acid with chemical formula CH2(COOH)2

Malonic acid (IUPAC systematic name: propanedioic acid) is a dicarboxylic acid with structure CH2(COOH)2. The ionized form of malonic acid, as well as its esters and salts, are known as malonates. For example, diethyl malonate is malonic acid's diethyl ester. The name originates from the Greek word μᾶλον (malon) meaning 'apple'.

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">Acyl halide</span> Oxoacid compound with an –OH group replaced by a halogen

In organic chemistry, an acyl halide is a chemical compound derived from an oxoacid by replacing a hydroxyl group with a halide group.

<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 Dakin–West reaction is a chemical reaction that transforms an amino-acid into a keto-amide using an acid anhydride and a base, typically pyridine. It is named for Henry Drysdale Dakin (1880–1952) and Randolph West (1890–1949). In 2016 Schreiner and coworkers reported the first asymmetric variant of this reaction employing short oligopeptides as catalysts.

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

Nucleophilic acyl substitution describes a class of substitution reactions involving nucleophiles and acyl compounds. In this type of reaction, a nucleophile – such as an alcohol, amine, or enolate – displaces the leaving group of an acyl derivative – such as an acid halide, anhydride, or ester. The resulting product is a carbonyl-containing compound in which the nucleophile has taken the place of the leaving group present in the original acyl derivative. Because acyl derivatives react with a wide variety of nucleophiles, and because the product can depend on the particular type of acyl derivative and nucleophile involved, nucleophilic acyl substitution reactions can be used to synthesize a variety of different products.

In organic chemistry, the Arndt–Eistert reaction is the conversion of a carboxylic acid to its homologue. Named for the German chemists Fritz Arndt (1885–1969) and Bernd Eistert (1902–1978), the method entails treating an acid chlorides with diazomethane. It is a popular method of producing β-amino acids from α-amino acids.

<span class="mw-page-title-main">Peterson olefination</span> Chemical reaction

The Peterson olefination is the chemical reaction of α-silyl carbanions with ketones to form a β-hydroxysilane (2) which eliminates to form alkenes (3).

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

Diphenylketene is a chemical substance of the ketene family. Diphenylketene, like most stable disubstituted ketenes, is a red-orange oil at room temperature and pressure. Due to the successive double bonds in the ketene structure R1R2C=C=O, diphenyl ketene is a heterocumulene. The most important reaction of diphenyl ketene is the [2+2] cycloaddition at C-C, C-N, C-O, and C-S multiple bonds.

<span class="mw-page-title-main">Diketene</span> Organic compound with formula (CH2CO)2

Diketene is an organic compound with the molecular formula C4H4O2, and which is sometimes written as (CH2CO)2. It is formed by dimerization of ketene, H2C=C=O. Diketene is a member of the oxetane family. It is used as a reagent in organic chemistry. It is a colorless liquid.

<span class="mw-page-title-main">Ethenone</span> Organic compound with the formula H2C=C=O

In organic chemistry, ethenone is the formal name for ketene, an organic compound with formula C2H2O or H2C=C=O. It is the simplest member of the ketene class. It is an important reagent for acetylations.

Nitrile anions is jargon from the organic product resulting from the deprotonation of alkylnitriles. The proton(s) α to the nitrile group are sufficiently acidic that they undergo deprotonation by strong bases, usually lithium-derived. The products are not anions but covalent organolithium complexes. Regardless, these organolithium compounds are reactive toward various electrophiles.

The Buchner–Curtius–Schlotterbeck reaction is the reaction of aldehydes or ketones with aliphatic diazoalkanes to form homologated ketones. It was first described by Eduard Buchner and Theodor Curtius in 1885 and later by Fritz Schlotterbeck in 1907. Two German chemists also preceded Schlotterbeck in discovery of the reaction, Hans von Pechmann in 1895 and Viktor Meyer in 1905. The reaction has since been extended to the synthesis of β-keto esters from the condensation between aldehydes and diazo esters. The general reaction scheme is as follows:

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

Ethyl cyanoacetate is an organic compound that contains a carboxylate ester and a nitrile. It is a colourless liquid with a pleasant odor. This material is useful as a starting material for synthesis due to its variety of functional groups and chemical reactivity.

References

  1. "Meldrum's Acid". The Merck Index . 14th. Vol. edition. Merck Research Laboratories. 2006. p. 1005. ISBN   978-0-911910-00-1.
  2. 1 2 3 4 McNab, Hamish (1978). "Meldrum's acid". Chemical Society Reviews. 7 (3): 345–358. doi:10.1039/CS9780700345.
  3. 1 2 3 Norman Meldrum, Andrew (1908). "A β-lactonic acid from acetone and malonic acid". Journal of the Chemical Society, Transactions. 93: 598–601. doi:10.1039/CT9089300598.
  4. Nakamura, Satoshi; Hirao, Hajime; Ohwada, Tomohiko (2004). "Rationale for the Acidity of Meldrum's Acid. Consistent Relation of C−H Acidities to the Properties of Localized Reactive Orbital". J. Org. Chem. 69 (13): 4309–4316. doi:10.1021/jo049456f. PMID   15202884.
  5. Oikawa, Yuji; Sugano, Kiyoshi; Yonemitsu, Osamu (1978). "Meldrum's acid in organic synthesis. 2. A general and versatile synthesis of β-keto esters". J. Org. Chem. 43 (10): 2087–2088. doi:10.1021/jo00404a066.
  6. Ghosh, Santanu; Purkait, Anisha; Jana, Chandan K. (2020). "Environmentally benign decarboxylative N -, O -, and S -acetylations and acylations". Green Chemistry. 22 (24): 8721–8727. doi:10.1039/D0GC03731A. S2CID   229195097.
  7. Li, Jiang-Sheng; Da, Yu-Dong; Chen, Guo-Qin; Yang, Qian; Li, Zhi-Wei; Yang, Fan; Huang, Peng-Mian (13 February 2017). "Solvent-, and Catalyst-free Acylation of Anilines with Meldrum's Acids: A Neat Access to Anilides". ChemistrySelect. 2 (5): 1770–1773. doi:10.1002/slct.201601965.
  8. Davis, Garrett J.; Sofka, Holly A.; Jewett, John C. (3 April 2020). "Highly Stable Meldrum's Acid Derivatives for Irreversible Aqueous Covalent Modification of Amines". Organic Letters. 22 (7): 2626–2629. doi:10.1021/acs.orglett.0c00597. PMC   7679203 . PMID   32191483.
  9. Heard, David M.; Lennox, Alastair J. J. (7 May 2021). "Dichloromeldrum's Acid (DiCMA): A Practical and Green Amine Dichloroacetylation Reagent". Organic Letters. 23 (9): 3368–3372. doi:10.1021/acs.orglett.1c00850. hdl: 1983/89034a97-d592-4fcb-925d-459b9efd9a40 . PMID   33844547. S2CID   233223351.
  10. Gaber, Abd El-Aal M.; McNab, Hamish (2001). "Synthetic Applications of the Pyrolysis of Meldrum's Acid Derivatives". Synthesis. 2001 (14): 2059–2074. doi:10.1055/s-2001-18057.
  11. Dumas, Aaron M.; Fillion, Eric (2009). "Meldrum's Acids and 5-Alkylidene Meldrum's Acids in Catalytic Carbon–Carbon Bond-Forming Processes". Acc. Chem. Res. 43 (3): 440–454. doi:10.1021/ar900229z. PMID   20000793.
  12. Oikawa, Yuji; Hirasawa, Hitoshi; Yonemitsu, Osamu (1978). "Meldrum's acid in organic synthesis. 1. A convenient one-pot synthesis of ethyl indolepropionates". Tetrahedron Letters. 19 (20): 1759–1762. doi:10.1016/0040-4039(78)80037-9.
  13. Lipson, Victoria V.; Gorobets, Nikolay Yu. (2009). "One hundred years of Meldrum's acid: Advances in the synthesis of pyridine and pyrimidine derivatives". Mol. Divers. 13 (4): 399–419. doi:10.1007/s11030-009-9136-x. PMID   19381852. S2CID   20523615.
  14. Bonifácio, Vasco D. B. (2004). "Meldrum's Acid". Synlett. 2004 (9): 1649–1650. doi: 10.1055/s-2004-829539 .
  15. Davidson, David; Bernhard, Sidney A. (1948). "The Structure of Meldrum's Supposed β-Lactonic Acid". Journal of the American Chemical Society. 70 (10): 3426–3428. doi:10.1021/ja01190a060. PMID   18891879.

Further reading