Oxime

Last updated

General structure of oximes.svg

In organic chemistry, an oxime is an organic compound belonging to the imines, with the general formula RR’C=N−OH, where R is an organic side-chain and R' may be hydrogen, forming an aldoxime, or another organic group, forming a ketoxime. O-substituted oximes form a closely related family of compounds. Amidoximes are oximes of amides (R1C(=O)NR2R3) with general structure R1C(=NOH)NR2R3.

Contents

Oximes are usually generated by the reaction of hydroxylamine with aldehydes (R−CH=O) or ketones (RR’C=O). The term oxime dates back to the 19th century, a combination of the words oxygen and imine. [1]

Structure and properties

If the two side-chains on the central carbon are different from each other—either an aldoxime, or a ketoxime with two different "R" groups—the oxime can often have two different geometric stereoisomeric forms according to the E/Z configuration. An older terminology of syn and anti was used to identify especially aldoximes according to whether the R group was closer or further from the hydroxyl. Both forms are often stable enough to be separated from each other by standard techniques.

Oximes have three characteristic bands in the infrared spectrum, whose wavelengths corresponding to the stretching vibrations of its three types of bonds: 3600 cm−1 (O−H), 1665 cm−1 (C=N) and 945 cm−1 (N−O). [2]

In aqueous solution, aliphatic oximes are 102- to 103-fold more resistant to hydrolysis than analogous hydrazones. [3]

Preparation

Oximes can be synthesized by condensation of an aldehyde or a ketone with hydroxylamine. The condensation of aldehydes with hydroxylamine gives aldoximes, and ketoximes are produced from ketones and hydroxylamine. In general, oximes exist as colorless crystals or as thick liquids and are poorly soluble in water. Therefore, oxime formation can be used for the identification of ketone or aldehyde functional groups.

Oximes can also be obtained from reaction of nitrites such as isoamyl nitrite with compounds containing an acidic hydrogen atom. Examples are the reaction of ethyl acetoacetate and sodium nitrite in acetic acid, [4] [5] the reaction of methyl ethyl ketone with ethyl nitrite in hydrochloric acid. [6] and a similar reaction with propiophenone, [7] the reaction of phenacyl chloride, [8] the reaction of malononitrile with sodium nitrite in acetic acid [9]

A conceptually related reaction is the Japp–Klingemann reaction.

Reactions

The hydrolysis of oximes proceeds easily by heating in the presence of various inorganic acids, and the oximes decompose into the corresponding ketones or aldehydes, and hydroxylamines. The reduction of oximes by sodium metal, [10] sodium amalgam, hydrogenation, or reaction with hydride reagents produces amines. [11] Typically the reduction of aldoximes gives both primary amines and secondary amines; however, reaction conditions can be altered (such as the addition of potassium hydroxide in a 1/30 molar ratio) to yield solely primary amines. [12]

In general, oximes can be changed to the corresponding amide derivatives by treatment with various acids. This reaction is called Beckmann rearrangement. [13] In this reaction, a hydroxyl group is exchanged with the group that is in the anti position of the hydroxyl group. The amide derivatives that are obtained by Beckmann rearrangement can be transformed into a carboxylic acid by means of hydrolysis (base or acid catalyzed). And an amine by Hoffman degradation of the amide in the presence of alkali hypoclorites. Beckmann rearrangement is used for the industrial synthesis of caprolactam (see applications below).

The Ponzio reaction (1906) [14] concerning the conversion of m-nitrobenzaldoxime to m-nitrophenyldinitromethane using dinitrogen tetroxide was the result of research into TNT analogues: [15]

Ponzio Synthesis V2.svg

In the Neber rearrangement certain oximes are converted to the corresponding alpha-amino ketones.

Oximes can be dehydrated using acid anhydrides to yield corresponding nitriles.

Certain amidoximes react with benzenesulfonyl chloride to make substituted ureas in the Tiemann rearrangement: [16] [17]

Tiemann rearragement.png

Uses

In their largest application, an oxime is an intermediate in the industrial production of caprolactam, a precursor to Nylon 6. About half of the world's supply of cyclohexanone, more than a million tonnes annually, is converted to the oxime. In the presence of sulfuric acid catalyst, the oxime undergoes the Beckmann rearrangement to give the cyclic amide caprolactam: [18]

Caprolactam Synth.png

Metal extractant

Structure of Nickel bis(dimethylglyoximate). NIMGLO12.png
Structure of Nickel bis(dimethylglyoximate).

Oximes are commonly used as ligands and sequestering agents for metal ions. Dimethylglyoxime (dmgH2) is a reagent for the analysis of nickel and a popular ligand in its own right. In the typical reaction, a metal reacts with two equivalents of dmgH2 concomitant with ionization of one proton. Salicylaldoxime is a chelator in hydrometallurgy. [19]

Amidoximes such as polyacrylamidoxime can be used to capture trace amounts of uranium from sea water. [20] [21] In 2017 researchers announced a configuration that absorbed up to nine times as much uranyl as previous fibers without saturating. [22]

Other applications

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">Beckmann rearrangement</span> Chemical rearrangement

The Beckmann rearrangement, named after the German chemist Ernst Otto Beckmann (1853–1923), is a rearrangement of an oxime functional group to substituted amides. The rearrangement has also been successfully performed on haloimines and nitrones. Cyclic oximes and haloimines yield lactams.

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

Hydroxylamine is an inorganic compound with the formula NH2OH. The material is a white crystalline, hygroscopic compound. Hydroxylamine is almost always provided and used as an aqueous solution. It is consumed almost exclusively to produce Nylon-6. The oxidation of NH3 to hydroxylamine is a step in biological nitrification.

<span class="mw-page-title-main">Imine</span> Organic compound or functional group containing a C=N bond

In organic chemistry, an imine is a functional group or organic compound containing a carbon–nitrogen double bond. The nitrogen atom can be attached to a hydrogen or an organic group (R). The carbon atom has two additional single bonds. Imines are common in synthetic and naturally occurring compounds and they participate in many reactions.

In organic chemistry, a nitrile is any organic compound that has a −C≡N functional group. The name of the compound is composed of a base, which includes the carbon of the −C≡N, suffixed with "nitrile", so for example CH3CH2C≡N is called "propionitrile". The prefix cyano- is used interchangeably with the term nitrile in industrial literature. Nitriles are found in many useful compounds, including methyl cyanoacrylate, used in super glue, and nitrile rubber, a nitrile-containing polymer used in latex-free laboratory and medical gloves. Nitrile rubber is also widely used as automotive and other seals since it is resistant to fuels and oils. Organic compounds containing multiple nitrile groups are known as cyanocarbons.

<span class="mw-page-title-main">Nitro compound</span> Organic compound containing an −NO₂ group

In organic chemistry, nitro compounds are organic compounds that contain one or more nitro functional groups. The nitro group is one of the most common explosophores used globally. The nitro group is also strongly electron-withdrawing. Because of this property, C−H bonds alpha (adjacent) to the nitro group can be acidic. For similar reasons, the presence of nitro groups in aromatic compounds retards electrophilic aromatic substitution but facilitates nucleophilic aromatic substitution. Nitro groups are rarely found in nature. They are almost invariably produced by nitration reactions starting with nitric acid.

<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.

A lactam is a cyclic amide, formally derived from an amino alkanoic acid through cyclization reactions. The term is a portmanteau of the words lactone + amide.

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

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).

Barrelene is a bicyclic organic compound with chemical formula C8H8 and systematic name bicyclo[2.2.2]octa-2,5,7-triene. First synthesized and described by Howard Zimmerman in 1960, the name derives from the resemblance to a barrel, with the staves being three ethylene units attached to two methine groups. It is the formal Diels–Alder adduct of benzene and acetylene. Due to its unusual molecular geometry, the compound is of considerable interest to theoretical chemists.

<span class="mw-page-title-main">Alkyl nitrite</span> Organic compounds of the form R–O–N=O

In organic chemistry, alkyl nitrites are a group of organic compounds based upon the molecular structure R−O−N=O, where R represents an alkyl group. Formally they are alkyl esters of nitrous acid. They are distinct from nitro compounds.

The reduction of nitro compounds are chemical reactions of wide interest in organic chemistry. The conversion can be effected by many reagents. The nitro group was one of the first functional groups to be reduced. Alkyl and aryl nitro compounds behave differently. Most useful is the reduction of aryl nitro compounds.

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.

<span class="mw-page-title-main">Schmidt reaction</span> Chemical reaction between an azide and a carbonyl derivative

In organic chemistry, the Schmidt reaction is an organic reaction in which an azide reacts with a carbonyl derivative, usually an aldehyde, ketone, or carboxylic acid, under acidic conditions to give an amine or amide, with expulsion of nitrogen. It is named after Karl Friedrich Schmidt (1887–1971), who first reported it in 1924 by successfully converting benzophenone and hydrazoic acid to benzanilide. The intramolecular reaction was not reported until 1991 but has become important in the synthesis of natural products. The reaction is effective with carboxylic acids to give amines (above), and with ketones to give amides (below).

The Stieglitz rearrangement is a rearrangement reaction in organic chemistry which is named after the American chemist Julius Stieglitz (1867–1937) and was first investigated by him and Paul Nicholas Leech in 1913. It describes the 1,2-rearrangement of trityl amine derivatives to triaryl imines. It is comparable to a Beckmann rearrangement which also involves a substitution at a nitrogen atom through a carbon to nitrogen shift. As an example, triaryl hydroxylamines can undergo a Stieglitz rearrangement by dehydration and the shift of a phenyl group after activation with phosphorus pentachloride to yield the respective triaryl imine, a Schiff base.

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

Hydroxylammonium sulfate (NH3OH)2SO4, is the sulfuric acid salt of hydroxylamine. It is primarily used as an easily handled form of hydroxylamine, which is explosive when pure.

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

Cyclohexanone oxime is an organic compound containing the functional group oxime. This colorless solid is an important intermediate in the production of nylon 6, a widely used polymer.

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.:

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

Acetone oxime (acetoxime) is the organic compound with the formula (CH3)2CNOH. It is the simplest example of a ketoxime. It is a white crystalline solid that is soluble in water, ethanol, ether, chloroform, and ligroin. It is used as a reagent in organic synthesis.

Hydroxylamine-<i>O</i>-sulfonic acid Chemical compound

Hydroxylamine-O-sulfonic acid (HOSA) or aminosulfuric acid is the inorganic compound with molecular formula H3NO4S that is formed by the sulfonation of hydroxylamine with oleum. It is a white, water-soluble and hygroscopic, solid, commonly represented by the condensed structural formula H2NOSO3H, though it actually exists as a zwitterion and thus is more accurately represented as +H3NOSO3. It is used as a reagent for the introduction of amine groups (–NH2), for the conversion of aldehydes into nitriles and alicyclic ketones into lactams (cyclic amides), and for the synthesis of variety of nitrogen-containing heterocycles.

References

  1. The name "oxime" is derived from "oximide" (i.e., oxy- + amide). According to the German organic chemist Victor Meyer (1848–1897) – who, with Alois Janny, synthesized the first oximes – an "oximide" was an organic compound containing the group (=N−OH) attached to a carbon atom. The existence of oximides was questioned at the time (ca. 1882). (See page 1164 of: Victor Meyer und Alois Janny (1882a) "Ueber stickstoffhaltige Acetonderivate" (On nitrogenous derivatives of acetone), Berichte der Deutschen chemischen Gesellschaft, 15: 1164–1167.) However, in 1882, Meyer and Janny succeeded in synthesizing methylglyoxime (CH3C(=NOH)CH(=NOH)), which they named "Acetoximsäure" (acetoximic acid) (Meyer & Janny, 1882a, p. 1166). Subsequently, they synthesized 2-propanone, oxime ((CH3)2C=NOH), which they named "Acetoxim" (acetoxime), in analogy with Acetoximsäure. From Victor Meyer and Alois Janny (1882b) "Ueber die Einwirkung von Hydroxylamin auf Aceton" (On the effect of hydroxylamine on acetone), Berichte der Deutschen chemischen Gesellschaft, 15: 1324–1326, page 1324: "Die Substanz, welche wir, wegen ihrer nahen Beziehungen zur Acetoximsäure, und da sie keine sauren Eigenschaften besitzt, vorläufig Acetoxim nennen wollen, …" (The substance, which we – on account of its close relations to acetoximic acid, and since it possesses no acid properties – will, for the present, name "acetoxime," … )
  2. Reusch, W. "Infrared Spectroscopy". Virtual Textbook of Organic Chemistry. Michigan State University. Archived from the original on 21 June 2010. Retrieved 6 July 2009.
  3. Kalia, J.; Raines, R. T. (2008). "Hydrolytic stability of hydrazones and oximes". Angew. Chem. Int. Ed. 47 (39): 7523–7526. doi:10.1002/anie.200802651. PMC   2743602 . PMID   18712739.
  4. Fischer, Hans (1943). "2,4-Dimethyl-3,5-dicarbethoxypyrrole". Organic Syntheses .; Collective Volume, vol. 2, p. 202
  5. Fischer, Hans (1955). "Kryptopyrrole". Organic Syntheses .; Collective Volume, vol. 3, p. 513
  6. Semon, W. L. & Damerell, V. R. (1943). "Dimethoxyglyoxime". Organic Syntheses .; Collective Volume, vol. 2, p. 204
  7. Hartung, Walter H. & Crossley, Frank (1943). "Isonitrosopropiophenone". Organic Syntheses .; Collective Volume, vol. 2, p. 363
  8. Levin, Nathan & Hartung, Walter H. (1955). "ω-chloroisonitrosoacetophenone". Organic Syntheses .; Collective Volume, vol. 3, p. 191
  9. Ferris, J. P.; Sanchez, R. A. & Mancuso, R. W. (1973). "p-toluenesulfonate". Organic Syntheses .; Collective Volume, vol. 5, p. 32
  10. Suter, C. M.; Moffett, Eugene W. (1934). "The Reduction of Aliphatic Cyanides and Oximes with Sodium and n-Butyl Alcohol". Journal of the American Chemical Society. 56 (2): 487. doi:10.1021/ja01317a502.
  11. George, Frederick; Saunders, Bernard (1960). Practical Organic Chemistry (4th ed.). London: Longman. p. 93 & 226. ISBN   9780582444072.
  12. Hata, Kazuo (1972). New Hydrogenating Catalysts. New York: John Wiley & Sons Inc. p. 193. ISBN   9780470358900.
  13. Clayden, Jonathan; Greeves, Nick; Warren, Stuart (2012). Organic chemistry (2nd ed.). Oxford University Press. p. 958. ISBN   978-0-19-927029-3.
  14. Ponzio, Giacomo (1906). "Einwirkung von Stickstofftetroxyd auf Benzaldoxim". J. Prakt. Chem. 73: 494–496. doi:10.1002/prac.19060730133.
  15. Fieser, Louis F. & Doering, William von E. (1946). "Aromatic-Aliphatic Nitro Compounds. III. The Ponzio Reaction; 2,4,6-Trinitrobenzyl Nitrate". J. Am. Chem. Soc. 68 (11): 2252–2253. doi:10.1021/ja01215a040.
  16. Tiemann, Ferdinand (1891). "Ueber die Einwirkung von Benzolsulfonsäurechlorid auf Amidoxime". Chemische Berichte . 24 (2): 4162–4167. doi:10.1002/cber.189102402316.
  17. Plapinger, Robert; Owens, Omer (1956). "Notes – The Reaction of Phosphorus-Containing Enzyme Inhibitors with Some Hydroxylamine Derivatives". J. Org. Chem. 21 (10): 1186. doi:10.1021/jo01116a610.
  18. Ritz, Josef; Fuchs, Hugo; Kieczka, Heinz; Moran, William C. "Caprolactam". Ullmann's Encyclopedia of Industrial Chemistry . Weinheim: Wiley-VCH. doi:10.1002/14356007.a05_031.pub2. ISBN   978-3527306732.
  19. Smith, Andrew G.; Tasker, Peter A.; White, David J. (2003). "The structures of phenolic oximes and their complexes". Coordination Chemistry Reviews. 241 (1–2): 61–85. doi:10.1016/S0010-8545(02)00310-7.
  20. Rao, Linfeng (15 March 2010). "Recent International R&D Activities in the Extraction of Uranium from Seawater". Lawrence Berkeley National Laboratory.{{cite journal}}: Cite journal requires |journal= (help)
  21. Kanno, M (1984). "Present status of study on extraction of uranium from sea water". Journal of Nuclear Science and Technology. 21 (1): 1–9. Bibcode:1984JNST...21....1K. doi: 10.1080/18811248.1984.9731004 .
  22. Dent, Steve (17 February 2017). "Endless nuclear power can be found in the seas". Engadget. Retrieved 22 February 2017.
  23. Rowe, Aaron (27 November 2007). "New Nerve Gas Antidotes". Wired .
  24. Kassa, J. (2002). "Review of oximes in the antidotal treatment of poisoning by organophosphorus nerve agents". Journal of Toxicology: Clinical Toxicology. 40 (6): 803–16. doi:10.1081/CLT-120015840. PMID   12475193. S2CID   20536869.
  25. Johannes Panten and Horst Surburg "Flavors and Fragrances, 2. Aliphatic Compounds" in Ullmann's Encyclopedia of Industrial Chemistry, 2015, Wiley-VCH, Weinheim. doi : 10.1002/14356007.t11_t01