Swern oxidation

Last updated
Swern oxidation
Named after Daniel Swern
Reaction type Organic redox reaction
Identifiers
Organic Chemistry Portal swern-oxidation
RSC ontology ID RXNO:0000154
 X mark.svgN (what is this?)    (verify)

In organic chemistry, the Swern oxidation, named after Daniel Swern, is a chemical reaction whereby a primary or secondary alcohol (−OH) is oxidized to an aldehyde (−CH=O) or ketone (>C=O) using oxalyl chloride, dimethyl sulfoxide (DMSO) and an organic base, such as triethylamine. [1] [2] [3] It is one of the many oxidation reactions commonly referred to as 'activated DMSO' oxidations. The reaction is known for its mild character and wide tolerance of functional groups. [4] [5] [6] [7]

Contents

The Swern oxidation. Swern Oxidation Scheme.png
The Swern oxidation.

The by-products are dimethyl sulfide ((CH3)2S), carbon monoxide (CO), carbon dioxide (CO2) and—when triethylamine is used as base—triethylammonium chloride (Et3NHCl). Of the volatile by-products, dimethyl sulfide has a strong, pervasive odour and carbon monoxide is acutely toxic, so the reaction and the work-up needs to be performed in a fume hood. Dimethyl sulfide is a volatile liquid (B.P. 37 °C) with an unpleasant odour at even low concentrations. [8] [9] [10]

Mechanism

The first step of the Swern oxidation is the low-temperature reaction of DMSO, 1a, formally as resonance contributor 1b, with oxalyl chloride, 2. The first intermediate, 3, quickly decomposes giving off carbon dioxide and carbon monoxide and producing chloro(dimethyl)sulfonium chloride, 4.

Dimethylchlorosulfonium chloride formation. Dimethylchlorosulfonium Formation Mechanism.png
Dimethylchlorosulfonium chloride formation.

After addition of the alcohol 5, the chloro(dimethyl)sulfonium chloride 4 reacts with the alcohol to give the key alkoxysulfonium ion intermediate, 6. The addition of at least 2 equivalents of base — typically triethylamine — will deprotonate the alkoxysulfonium ion to give the sulfur ylide 7. In a five-membered ring transition state, the sulfur ylide 7 decomposes to give dimethyl sulfide and the desired carbonyl compound 8.

The mechanism of the Swern oxidation. Swern Oxidation Mechanism.png
The mechanism of the Swern oxidation.

Variations

When using oxalyl chloride as the dehydration agent, the reaction must be kept colder than 60 °C to avoid side reactions. With cyanuric chloride [11] or trifluoroacetic anhydride [12] instead of oxalyl chloride, the reaction can be warmed to 30 °C without side reactions. Other methods for the activation of DMSO to initiate the formation of the key intermediate 6 are the use of carbodiimides (Pfitzner–Moffatt oxidation), a sulfur trioxide pyridine complex (Parikh–Doering oxidation) or acetic anhydride (Albright-Goldman oxidation). The intermediate 4 can also be prepared from dimethyl sulfide and N-chlorosuccinimide (the Corey–Kim oxidation).

In some cases, the use of triethylamine as the base can lead to epimerisation at the carbon alpha to the newly formed carbonyl. Using a bulkier base, such as diisopropylethylamine, can mitigate this side reaction.

Considerations

Dimethyl sulfide, a byproduct of the Swern oxidation, is one of the most notoriously unpleasant odors known in organic chemistry. Humans can detect this compound in concentrations as low as 0.02 to 0.1 parts per million. [13] A simple remedy for this problem is to rinse used glassware with bleach or oxone solution, which will oxidize the dimethyl sulfide back to dimethyl sulfoxide or to dimethyl sulfone, both of which are odourless and nontoxic. [14]

The reaction conditions allow oxidation of acid-sensitive compounds, which might decompose under the acidic oxidation conditions such as Jones oxidation. For example, in Thompson & Heathcock's synthesis of the sesquiterpene isovelleral, [15] the final step uses the Swern protocol, avoiding rearrangement of the acid-sensitive cyclopropanemethanol moiety.

IsovelleralPreparationViaSwernOxidation.png

See also

Related Research Articles

<span class="mw-page-title-main">Dimethyl sulfoxide</span> Organosulfur chemical compound used as a solvent

Dimethyl sulfoxide (DMSO) is an organosulfur compound with the formula (CH3)2SO. This colorless liquid is the sulfoxide most widely used commercially. It is an important polar aprotic solvent that dissolves both polar and nonpolar compounds and is miscible in a wide range of organic solvents as well as water. It has a relatively high boiling point. DMSO has the unusual property that many individuals perceive a garlic-like taste in the mouth after DMSO makes contact with their skin.

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

Oxalyl chloride is an organic chemical compound with the formula Cl−C(=O)−C(=O)−Cl. This colorless, sharp-smelling liquid, the diacyl chloride of oxalic acid, is a useful reagent in organic synthesis.

The Kornblum oxidation, named after Nathan Kornblum, is an organic oxidation reaction that converts alkyl halides and tosylates into carbonyl compounds.

Organosulfur chemistry is the study of the properties and synthesis of organosulfur compounds, which are organic compounds that contain sulfur. They are often associated with foul odors, but many of the sweetest compounds known are organosulfur derivatives, e.g., saccharin. Nature is abound with organosulfur compounds—sulfur is vital for life. Of the 20 common amino acids, two are organosulfur compounds, and the antibiotics penicillin and sulfa drugs both contain sulfur. While sulfur-containing antibiotics save many lives, sulfur mustard is a deadly chemical warfare agent. Fossil fuels, coal, petroleum, and natural gas, which are derived from ancient organisms, necessarily contain organosulfur compounds, the removal of which is a major focus of oil refineries.

<span class="mw-page-title-main">Carbodiimide</span> Class of organic compounds with general structure RN=C=NR

In organic chemistry, a carbodiimide is a functional group with the formula RN=C=NR. On Earth they are exclusively synthetic, but in interstellar space the parent compound HN=C=NH has been detected by its maser emissions.

<span class="mw-page-title-main">Sulfoxide</span> Organic compound containing a sulfinyl group (>SO)

In organic chemistry, a sulfoxide, also called a sulphoxide, is an organosulfur compound containing a sulfinyl functional group attached to two carbon atoms. It is a polar functional group. Sulfoxides are oxidized derivatives of sulfides. Examples of important sulfoxides are alliin, a precursor to the compound that gives freshly crushed garlic its aroma, and dimethyl sulfoxide (DMSO), a common solvent.

<i>N</i>,<i>N</i>-Diisopropylethylamine Chemical compound

N,N-Diisopropylethylamine, or Hünig's base, is an organic compound that is a tertiary amine. It is named after the German chemist Siegfried Hünig. It is used in organic chemistry as a non-nucleophilic base. It is commonly abbreviated as DIPEA,DIEA, or i-Pr2NEt.

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">Johnson–Corey–Chaykovsky reaction</span> Chemical reaction in organic chemistry

The Johnson–Corey–Chaykovsky reaction is a chemical reaction used in organic chemistry for the synthesis of epoxides, aziridines, and cyclopropanes. It was discovered in 1961 by A. William Johnson and developed significantly by E. J. Corey and Michael Chaykovsky. The reaction involves addition of a sulfur ylide to a ketone, aldehyde, imine, or enone to produce the corresponding 3-membered ring. The reaction is diastereoselective favoring trans substitution in the product regardless of the initial stereochemistry. The synthesis of epoxides via this method serves as an important retrosynthetic alternative to the traditional epoxidation reactions of olefins.

The Pfitzner–Moffatt oxidation, sometimes referred to as simply the Moffatt oxidation, is a chemical reaction for the oxidation of primary and secondary alcohols to aldehydes and ketones, respectively. The oxidant is a combination of dimethyl sulfoxide (DMSO) and dicyclohexylcarbodiimide (DCC). The reaction was first reported by J. Moffatt and his student K. Pfitzner in 1963.

The Corey–Kim oxidation is an oxidation reaction used to synthesise aldehydes and ketones from primary and secondary alcohols. It is named for American chemist and Nobel Laureate Elias James Corey and Korean-American chemist Choung Un Kim.

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

Sodium methylsulfinylmethylide is the sodium salt of the conjugate base of dimethyl sulfoxide. This unusual salt has some uses in organic chemistry as a base and nucleophile.

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

Methanesulfonic anhydride (Ms2O) is the acid anhydride of methanesulfonic acid. Like methanesulfonyl chloride (MsCl), it may be used to generate mesylates (methanesulfonyl esters).

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

Trifluoroacetic anhydride (TFAA) is the acid anhydride of trifluoroacetic acid. It is the perfluorinated derivative of acetic anhydride.

The Parikh–Doering oxidation is an oxidation reaction that transforms primary and secondary alcohols into aldehydes and ketones, respectively. The procedure uses dimethyl sulfoxide (DMSO) as the oxidant and the solvent, activated by the sulfur trioxide pyridine complex (SO3•C5H5N) in the presence of triethylamine or diisopropylethylamine as base. Dichloromethane is frequently used as a cosolvent for the reaction.

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.

<span class="mw-page-title-main">Sulfinyl halide</span> Class of chemical compounds

Sulfinyl halide have the general formula R−S(O)−X, where X is a halogen. They are intermediate in oxidation level between sulfenyl halides, R−S−X, and sulfonyl halides, R−SO2−X. The best known examples are sulfinyl chlorides, thermolabile, moisture-sensitive compounds, which are useful intermediates for preparation of other sufinyl derivatives such as sulfinamides, sulfinates, sulfoxides, and thiosulfinates. Unlike the sulfur atom in sulfonyl halides and sulfenyl halides, the sulfur atom in sulfinyl halides is chiral, as shown for methanesulfinyl chloride.

<i>N</i>-<i>tert</i>-Butylbenzenesulfinimidoyl chloride Chemical compound

N-tert-Butylbenzenesulfinimidoyl chloride is a useful oxidant for organic synthesis reactions. It is a good electrophile, and the sulfimide S=N bond can be attacked by nucleophiles, such as alkoxides, enolates, and amide ions. The nitrogen atom in the resulting intermediate is basic, and can abstract an α-hydrogen to create a new double bond.

<span class="mw-page-title-main">Albright–Goldman oxidation</span>

The Albright–Goldman oxidation is a name reaction of organic chemistry, first described by the American chemists J. Donald Albright and Leon Goldman in 1965. The reaction is particularly suitable for the synthesis of aldehydes from primary alcohols. Analogously, secondary alcohols can be oxidized to form ketones. Dimethyl sulfoxide/acetic anhydride serves as oxidizing agent.

Sulfonium-based oxidations of alcohols to aldehydes summarizes a group of organic reactions that transform a primary alcohol to the corresponding aldehyde (and a secondary alcohol to the corresponding ketone). Selective oxidation of alcohols to aldehydes requires circumventing over-oxidation to the carboxylic acid. One popular approach are methods that proceed through intermediate alkoxysulfonium species (RO−SMe+
2
X-
, e.g. compound 6) as detailed here. Since most of these methods employ dimethylsulfoxide (DMSO) as oxidant and generate dimethylsulfide, these are often colloquially summarized as DMSO-oxidations. Conceptually, generating an aldehyde and dimethylsulfide from an alcohol and DMSO requires a dehydrating agent for removal of H2O, ideally an electrophile simultaneously activating DMSO. In contrast, methods generating the sulfonium intermediate from dimethylsulfide do not require a dehydrating agent. Closely related are oxidations mediated by dimethyl selenoxide and by dimethyl selenide.

References

  1. Omura, K.; Swern, D. (1978). "Oxidation of alcohols by "activated" dimethyl sulfoxide. A preparative, steric and mechanistic study". Tetrahedron . 34 (11): 1651–1660. doi:10.1016/0040-4020(78)80197-5.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  2. Mancuso, A. J.; Brownfain, D. S.; Swern, D. (1979). "Structure of the dimethyl sulfoxide-oxalyl chloride reaction product. Oxidation of heteroaromatic and diverse alcohols to carbonyl compounds". J. Org. Chem. 44 (23): 4148–4150. doi:10.1021/jo01337a028.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  3. Mancuso, A. J.; Huang, S.-L.; Swern, D. (1978). "Oxidation of long-chain and related alcohols to carbonyls by dimethyl sulfoxide "activated" by oxalyl chloride". J. Org. Chem. 43 (12): 2480–2482. doi:10.1021/jo00406a041.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  4. Dondoni, A.; Perrone, D. (2004). "Synthesis of 1,1-Dimethyl Ethyl-(S)-4-formyl-2,2-dimethyl-3-oxazolidinecarboxylate by Oxidation of the Alcohol". Organic Syntheses .{{cite journal}}: CS1 maint: multiple names: authors list (link); Collective Volume, vol. 10, p. 320
  5. Bishop, R. (1998). "9-Thiabicyclo[3.3.1]nonane-2,6-dione". Organic Syntheses .; Collective Volume, vol. 9, p. 692
  6. Leopold, E. J. (1990). "Selective hydroboration of a 1,3,7-triene: Homogeraniol". Organic Syntheses .; Collective Volume, vol. 7, p. 258
  7. Tojo, G.; Fernández, M. (2006). Oxidation of alcohols to aldehydes and ketones: A guide to current common practice. Springer. ISBN   0-387-23607-4.
  8. Mancuso, A. J.; Swern, D. (1981). "Activated dimethyl sulfoxide: Useful reagents for synthesis". Synthesis (Review). 1981 (3): 165–185. doi:10.1055/s-1981-29377.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  9. Tidwell, T. T. (1990). "Oxidation of alcohols to carbonyl compounds via alkoxysulfonium ylides: The Moffatt, Swern, and related oxidations". Org. React. (Review). 39: 297–572. doi:10.1002/0471264180.or039.03. ISBN   0471264180.
  10. Tidwell, T. T. (1990). "Oxidation of alcohols by activated dimethyl sulfoxide and related reactions: An update". Synthesis (Review). 1990 (10): 857–870. doi:10.1055/s-1990-27036.
  11. De Luca Lidia (2001). "A Mild and Efficient Alternative to the Classical Swern Oxidation". The Journal of Organic Chemistry. 66 (23): 7907–7909. doi:10.1021/jo015935s. PMID   11701058.
  12. Omura, Kanji; Sharma, Ashok K.; Swern, Daniel (1976). "Dimethyl Sulfoxide-Trifluoroacetic Anhydride. New Reagent for Oxidation of Alcohols to Carbonyls". J. Org. Chem. 41 (6): 957–962. doi:10.1021/jo00868a012.
  13. Morton, T. H. (2000). "Archiving Odors". In Bhushan, N.; Rosenfeld, S. (eds.). Of Molecules and Mind. Oxford: Oxford University Press. pp. 205–216.
  14. Atkins, William J. Jr.; Burkhardt, Elizabeth R.; Matos, Karl (2006). "Safe Handling of Boranes at Scale". Org. Process Res. Dev. 10 (6): 1292–1295. doi:10.1021/op068011l.
  15. Thompson, S. K.; Heathcock, C. H. (1992). "Total synthesis of some marasmane and lactarane sesquiterpenes". J. Org. Chem. 57 (22): 5979–5989. doi:10.1021/jo00048a036.