Acetaldoxime

Last updated
Acetaldoxime
Acetaldoxime.svg
Acetaldoxime.png
Names
Preferred IUPAC name
N-Hydroxyethanimine
Other names
Aldoxime, Acetaldehyde oxime, Ethanal oxime, Ethylidenehydroxylamine
Identifiers
3D model (JSmol)
1209252
ChEBI
ChemSpider
ECHA InfoCard 100.003.164 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 203-479-6
PubChem CID
RTECS number
  • AB2975000
UNII
  • InChI=1S/C2H5NO/c1-2-3-4/h2,4H,1H3/b3-2+
    Key: FZENGILVLUJGJX-NSCUHMNNSA-N
  • InChI=1/C2H5NO/c1-2-3-4/h2,4H,1H3/b3-2+
    Key: FZENGILVLUJGJX-NSCUHMNNBP
  • N(/O)=C\C
Properties
C2H5NO
Molar mass 59.067 g mol−1
Appearanceclear, colorless to yellow liquid
Density 0.97 g cm−3
Melting point 25 °C (77 °F; 298 K) (average of the α and β forms)
Boiling point 115.24 °C (239.43 °F; 388.39 K)
299 g L−1
Solubility in ethanol miscible
log P -0.13
Vapor pressure 13 mmHg
Acidity (pKa)11.82
Hazards
Occupational safety and health (OHS/OSH):
Main hazards
Flammable, harmful by ingestion, irritant
NFPA 704 (fire diamond)
NFPA 704.svgHealth 2: Intense or continued but not chronic exposure could cause temporary incapacitation or possible residual injury. E.g. chloroformFlammability 2: Must be moderately heated or exposed to relatively high ambient temperature before ignition can occur. Flash point between 38 and 93 °C (100 and 200 °F). E.g. diesel fuelInstability 0: Normally stable, even under fire exposure conditions, and is not reactive with water. E.g. liquid nitrogenSpecial hazards (white): no code
2
2
0
Flash point 40 °C (104 °F; 313 K)
Safety data sheet (SDS) External MSDS
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Acetaldoxime is the chemical compound with formula C2H5NO. It is one of the simplest oxime-containing compounds, and has a wide variety of uses in chemical synthesis.

Contents

Properties

Acetaldoxime will often appear as a colorless liquid, or a white solid. Its solid can form two different needle-like crystal structures. The α-form melts at approximately 44 °C - 47 °C, while the β-form melts at 12 °C. The liquid is known to have a pungent odor, and is highly flammable. The compound can act as both an acid or a base, due to its acidic proton on the hydroxyl group and the basic nitrogen atom. The compound exists as a mixture of its Z and E stereoisomers (i.e. syn and anti, or cis and trans) in its normal form. The E stereoisomer can be isolated by slow crystallization of a distilled E/Z mixture. [1] [2]

Production

Acetaldoxime can be prepared by combining pure acetaldehyde and hydroxylamine under heating in the presence of a base. [3]

Preparation of acetaldoxime from acetaldehyde and hydroxylamine Acetaldehyde Oxime Synthesis V1.svg
Preparation of acetaldoxime from acetaldehyde and hydroxylamine

The use of CaO as a base in the preparation of oximes from various types of ketones and aldehydes under mild conditions also gave quantitative yields. [4]

Reactions

Alkylation

Deprotonation of acetaldoxime with 2 equivalents of n-butyllithium at -78 °C generates the dianion which reacts with benzyl bromide or 1-iodopropane to give excellent yields of α-alkylated (Z)-oximes. [5] α,α-Dialkylation by further alkylation in similar way has been achieved. [5] It is generally known that ketone oximes can be deprotonated and alkylated regiospecifically syn to the oxime hydroxy group. [6] [7] It is essential to perform the deprotonation and alkylation at -78 °C as otherwise no α-alkylated oximes are isolated, the major byproducts being nitriles. [6]

Rearrangement to acetamide

Heating of acetaldoxime in xylene in the presence of 0.2 mol % nickel(II) acetate [1] or silica gel [8] as catalyst caused isomerization into acetamide.

Synthesis of heterocycles

Chlorination of acetaldoxime with N-chlorosuccinimide [9] or chlorine gas [7] [10] in chloroform affords acetohydroxamic acid chloride, which suffers dehydrochlorination with triethylamine to give acetonitrile N-oxide. The latter 1,3-dipole undergoes 1,3-dipolar cycloaddition to alkenes giving 2-isoxazolines in a one-pot procedure. [9] This reaction is also suitable for the construction of more complex molecules such as the conversion of a 6-ethylideneolivanic acid derivative into the corresponding spiroisoxazoline. [10]

Uses

Aldoximes such as acetylaldoxime are using during chemical synthesis processes as intermediates for chemical reactions. It is especially notable for its commercial application as an intermediate for the production of pesticides. [11]

Related Research Articles

<span class="mw-page-title-main">Hydrazone</span> Organic compounds - Hydrazones

Hydrazones are a class of organic compounds with the structure R1R2C=N−NH2. They are related to ketones and aldehydes by the replacement of the oxygen =O with the =N−NH2 functional group. They are formed usually by the action of hydrazine on ketones or aldehydes.

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">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.

<span class="mw-page-title-main">Enolate</span> Organic anion formed by deprotonating a carbonyl (>C=O) compound

In organic chemistry, enolates are organic anions derived from the deprotonation of carbonyl compounds. Rarely isolated, they are widely used as reagents in the synthesis of organic compounds.

<span class="mw-page-title-main">Wacker process</span> Chemical reaction

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride and copper(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

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

The Wolff rearrangement is a reaction in organic chemistry in which an α-diazocarbonyl compound is converted into a ketene by loss of dinitrogen with accompanying 1,2-rearrangement. The Wolff rearrangement yields a ketene as an intermediate product, which can undergo nucleophilic attack with weakly acidic nucleophiles such as water, alcohols, and amines, to generate carboxylic acid derivatives or undergo [2+2] cycloaddition reactions to form four-membered rings. The mechanism of the Wolff rearrangement has been the subject of debate since its first use. No single mechanism sufficiently describes the reaction, and there are often competing concerted and carbene-mediated pathways; for simplicity, only the textbook, concerted mechanism is shown below. The reaction was discovered by Ludwig Wolff in 1902. The Wolff rearrangement has great synthetic utility due to the accessibility of α-diazocarbonyl compounds, variety of reactions from the ketene intermediate, and stereochemical retention of the migrating group. However, the Wolff rearrangement has limitations due to the highly reactive nature of α-diazocarbonyl compounds, which can undergo a variety of competing reactions.

<span class="mw-page-title-main">Stork enamine alkylation</span> Reaction sequence in organic chemistry

The Stork enamine alkylation involves the addition of an enamine to a Michael acceptor or another electrophilic alkylation reagent to give an alkylated iminium product, which is hydrolyzed by dilute aqueous acid to give the alkylated ketone or aldehyde. Since enamines are generally produced from ketones or aldehydes, this overall process constitutes a selective monoalkylation of a ketone or aldehyde, a process that may be difficult to achieve directly.

The Rubottom oxidation is a useful, high-yielding chemical reaction between silyl enol ethers and peroxyacids to give the corresponding α-hydroxy carbonyl product. The mechanism of the reaction was proposed in its original disclosure by A.G. Brook with further evidence later supplied by George M. Rubottom. After a Prilezhaev-type oxidation of the silyl enol ether with the peroxyacid to form the siloxy oxirane intermediate, acid-catalyzed ring-opening yields an oxocarbenium ion. This intermediate then participates in a 1,4-silyl migration to give an α-siloxy carbonyl derivative that can be readily converted to the α-hydroxy carbonyl compound in the presence of acid, base, or a fluoride source.

The total synthesis of quinine, a naturally-occurring antimalarial drug, was developed over a 150-year period. The development of synthetic quinine is considered a milestone in organic chemistry although it has never been produced industrially as a substitute for natural occurring quinine. The subject has also been attended with some controversy: Gilbert Stork published the first stereoselective total synthesis of quinine in 2001, meanwhile shedding doubt on the earlier claim by Robert Burns Woodward and William Doering in 1944, claiming that the final steps required to convert their last synthetic intermediate, quinotoxine, into quinine would not have worked had Woodward and Doering attempted to perform the experiment. A 2001 editorial published in Chemical & Engineering News sided with Stork, but the controversy was eventually laid to rest once and for all when Williams and coworkers successfully repeated Woodward's proposed conversion of quinotoxine to quinine in 2007.

In organic chemistry, the Ei mechanism, also known as a thermal syn elimination or a pericyclic syn elimination, is a special type of elimination reaction in which two vicinal (adjacent) substituents on an alkane framework leave simultaneously via a cyclic transition state to form an alkene in a syn elimination. This type of elimination is unique because it is thermally activated and does not require additional reagents, unlike regular eliminations, which require an acid or base, or would in many cases involve charged intermediates. This reaction mechanism is often found in pyrolysis.

Selenoxide elimination is a method for the chemical synthesis of alkenes from selenoxides. It is most commonly used to synthesize α,β-unsaturated carbonyl compounds from the corresponding saturated analogues. It is mechanistically related to the Cope reaction.

<span class="mw-page-title-main">Organoindium chemistry</span> Chemistry of compounds with a carbon-indium bond

Organoindium chemistry is the chemistry of compounds containing In-C bonds. The main application of organoindium chemistry is in the preparation of semiconducting components for microelectronic applications. The area is also of some interest in organic synthesis. Most organoindium compounds feature the In(III) oxidation state, akin to its lighter congeners Ga(III) and B(III).

Alcohol oxidation is a collection of oxidation reactions in organic chemistry that convert alcohols to aldehydes, ketones, carboxylic acids, and esters where the carbon carries a higher oxidation state. The reaction mainly applies to primary and secondary alcohols. Secondary alcohols form ketones, while primary alcohols form aldehydes or carboxylic acids.

The α-ketol rearrangement is the acid-, base-, or heat-induced 1,2-migration of an alkyl or aryl group in an α-hydroxy ketone or aldehyde to give an isomeric product.

In organic chemistry, α-halo ketones can be reduced with loss of the halogen atom to form enolates. The α-halo ketones are readily prepared from ketones by various ketone halogenation reactions, and the products are reactive intermediates that can be used for a variety of other chemical reactions.

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

The White catalyst is a transition metal coordination complex named after the chemist by whom it was first synthesized, M. Christina White, a professor at the University of Illinois. The catalyst has been used in a variety of allylic C-H functionalization reactions of α-olefins. In addition, it has been shown to catalyze oxidative Heck reactions.

<span class="mw-page-title-main">Enders SAMP/RAMP hydrazone-alkylation reaction</span>

The Enders SAMP/RAMP hydrazone alkylation reaction is an asymmetric carbon-carbon bond formation reaction facilitated by pyrrolidine chiral auxiliaries. It was pioneered by E. J. Corey and Dieter Enders in 1976, and was further developed by Enders and his group. This method is usually a three-step sequence. The first step is to form the hydrazone between (S)-1-amino-2-methoxymethylpyrrolidine (SAMP) or (R)-1-amino-2-methoxymethylpyrrolidine (RAMP) and a ketone or aldehyde. Afterwards, the hydrazone is deprotonated by lithium diisopropylamide (LDA) to form an azaenolate, which reacts with alkyl halides or other suitable electrophiles to give alkylated hydrazone species with the simultaneous generation of a new chiral center. Finally, the alkylated ketone or aldehyde can be regenerated by ozonolysis or hydrolysis.

<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.

References

  1. 1 2 Field, L.; Hughmark, P. B.; Shumaker, S. H.; Marshall, W. S. J. Am. Chem. Soc.1961, 83, 1983.
  2. "Oximes - Encyclopedia".
  3. "Acetaldoxime".
  4. Sharghi, H., & Sarvari, M. H.. A mild and versatile method for the preparation of pximes by use of calcium oxide. J. Chem. Research, 2000, pp. 24—25.
  5. 1 2 Gawley, R. E.; Nagy, T. TL. 1984, 25, 263.
  6. 1 2 Kofron, W. G.; Yeh, M. K. J. Org. Chem.1976, 41, 439.
  7. 1 2 Mukerji, S. K.; Sharma, K. K.; Torssell, K. B. G. T. 1983, 39, 2231.
  8. Chattopadhyaya, J. B.; Rama Rao, A. V. T. 1974, 30, 2899.
  9. 1 2 Larsen, K. E.; Torssell, K. B. G. T. 1984, 40, 2985.
  10. 1 2 Corbett, D. F. J. Chem. Soc.1986, 421.
  11. "Production of acetaldehyde oxime".