Oxoammonium-catalyzed oxidation

Last updated

Oxoammonium-catalyzed oxidation reactions involve the conversion of organic substrates to more highly oxidized materials through the action of an N-oxoammonium species. Nitroxides may also be used in catalytic amounts in the presence of a stoichiometric amount of a terminal oxidant. [1] Nitroxide radical species used are either 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) or derivatives thereof.

Contents

(1)

OxoamGen.png

Mechanism and stereochemistry

One-electron oxidation of the nitroxide produces a highly electrophilic oxoammonium species, which serves as the active oxidizing agent. [2] The nitroxide can be used as a catalyst in conjunction with cheaper stoichiometric oxidants such as sodium hypochlorite [3] or bis(acetoxy)iodobenzene (BAIB). [4]

Under neutral or slightly acidic conditions (in the presence of silica gel, for instance), oxidation occurs by an initial hydrogen bond between the hydroxyl group and the oxoammonium nitrogen, followed by concerted proton transfer and hydride abstraction. [5] The need for hydrogen bonding is supported by the low reactivity of β-alkoxy and β-amino alcohols, which exhibit competitive intramolecular hydrogen bonding. The mechanism of oxidation under weakly basic (pyridine) conditions is similar, except that pyridine neutralizes the hydroxyammonium species, and this intermediate "comproportionates" with oxoammonium salt to give nitroxide radicals and pyridinium salts (see equation (3) below). Because this reaction consumes base and active oxidant, two equivalents of base and oxidant are necessary under weakly basic conditions. A unified mechanism under neutral and basic conditions in presented in a recent article. [6] The authors present a comprehensive analysis of a number of oxoammonium salt mediated oxidations.

(2)

OxoamMech1.png

Under strongly basic conditions, the deprotonated substrate reacts with the N-oxyammonium species. Attack of the substrate alkoxide on either nitrogen or oxygen may occur, although the former is believed to operate on the basis of on observations of oxidations of N-alkoxy amines (which, presumably, proceed via intermediate 1). [7] Comproportionation of the reduced product (a hydroxylamine) with the oxoammonium ion competes with oxidation; thus, an excess of the oxidizing agent is often required.

(3)

OxoamMech2.png

Nitroxide-catalyzed oxidations involve N-oxoammonium intermediates as the active oxidizing agent. The mechanism of oxidation of the nitroxide radical depends on the terminal oxidant employed. Two-electron oxidants, such as NaOCl, are able to directly convert nitroxides into oxoammoniums.

(4)

OxoamMech3.png

One-electron oxidants, such as copper(II), operate via a more complex mechanism involving dioxygen as the terminal oxidant. [8] Copper(II) oxidizes four equivalents of nitroxide to oxoammonium, two equivalents of which (blue) react with alcohols to form carbonyl compounds. The other two equivalents of oxoammonium (red) undergo comproportionation to re-form nitroxy radicals (pink). Finally, dioxygen re-oxidizes four equivalents of copper(I) back to copper(II). Overall, a single molecule of dioxygen mediates the oxidation of two equivalents of alcohol, with the formation of two equivalents of water.

(5)

OxoamMech4.png

Stereoselective variants

Enantioselective oxidations are typically either kinetic resolutions of chiral alcohols or desymmetrization reactions. These oxidations may be facilitated through the use of chiral nitroxide radicals in the catalytic mode. A good example is provided by the kinetic resolution of racemic 1-phenylethanol. [9] Oxidative desymmetrization processes employing oxoammonium oxidants, on the other hand, are relatively rare. [10]

(6)

OxoamScope7.png

Scope

Oxidations using oxoammonium salts may be carried out either in the stoichiometric or catalytic mode under acidic or basic conditions. This section describes the most commonly used conditions for the stoichiometric and catalytic oxidation of alcohols to carbonyl compounds with oxoammonium salts. Although a wide variety of alcohols may be oxidized using TEMPO, competitive oxidation of more electron-rich functionality sometimes takes place. In addition, the site selectivity of oxidation of polyols may differ depending on the conditions used.

Stoichiometric oxidations

Under mildly acidic or neutral conditions, oxoammonium salts such as Bobbitt's salt oxidize allylic, benzylic, [11] propargylic, [12] or aliphatic alcohols to the corresponding aldehydes or ketones. Secondary alcohols react faster than primary ones, although selectivity is low. A convenient experimental protocol allows for recycling of the oxoammonium salt. [12]

(7)

OxoamScope1.png

Amines, benzylic ethers, and alkenes are oxidized more rapidly than unactivated alcohols; thus, selective stoichiometric oxidation of unactivated alcohols in the presence of these functional groups is not possible. [13] Alcohols with β-nitrogen or β-oxygen substituents react sluggishly under acidic conditions. [12] Allylic and benzylic alcohols can be selectively oxidized under these conditions [13]

(8)

OxoamScope2.png

Under basic conditions, two equivalents of oxidant are needed because of competitive comproportionation between reduced nitroxide and unreacted oxoammonium (see equation (3) above). Pyridine is usually employed as the base. These are the most common conditions for nitroxide oxidations in the stoichiometric mode.

(9)

OxoamScope3.png

Tertiary allylic alcohols can also be stoichiometrically oxidized by oxoammonium salts to enones in a variation of the Babler-Dauben reaction. [14]

Catalytic oxidations

Catalytic oxoammonium oxidation may be facilitated using sodium hypochlorite as the terminal oxidant. The pH must be maintained below 10 using a buffer for the reaction to proceed. The active oxidizing agent of nitroxide is hypobromite anion; hence, potassium bromide is used as an additive. [3] No epimerization of α-stereogenic centers in carbonyl-containing products takes place.

(10)

OxoamScope4-corrected.png

The use of chlorites as terminal oxidants in conjunction with both hypochlorites and TEMPO gives carboxylic acids without chlorination side products. [15] The reaction is usually carried out in two steps in the same pot: partial oxidation is effected with TEMPO and hypochlorite, then chlorite is added to complete the oxidation. Only primary alcohol oxidation is observed. In conjunction with Sharpless dihydroxylation, this method can be used to generate enantiopure α-hydroxy acids. [16]

(11)

OxoamScope8.png

A significant limitation of both of the above methods is incompatibility with free amine or alkene functionality, both of which undergo competitive oxidation. The use of bis(acetoxy)iodobenzene (BAIB) as the terminal oxidant avoids this problem. BAIB is unable to oxidize the nitroxide radical directly, and initial formation of oxoammonium is believed to be due to acid-catalyzed disproportionation. BAIB may then oxidize the resulting hydroxylamine to an oxoammonium salt. Although the reaction is conducted under acidic conditions (acetic acid is a byproduct, and is often added to facilitate disproportionation), selectivity for primary alcohol oxidation is substantial. [4] Base-sensitive functional groups, such as epoxides, are tolerated under these conditions. [17]

(12)

OxoamScope5.png

Other two-electron terminal oxidants used with TEMPO include mCPBA (secondary oxidation is favored, although side reactions may occur), [18] N-chlorosuccinimide, [19] and Oxone. [20]

Copper(II), both as the free chloride salt and as a complex with bidentate ligands, oxidizes TEMPO to its oxoammonium salt. In these reactions, air serves as the terminal oxidant. [21] It is unclear whether air oxidizes copper(I) to copper(II), or whether alcohol oxidation is partially mediated by copper and air oxidizes the resulting hydroxylamine back to the oxoammonium salt. The former occurs during the Wacker process, but the latter explains why copper complexes and a few other metal complexes are able to oxidize alcohols in conjunction with TEMPO.

(13)

OxoamScope6.png

Activated manganese dioxide, which oxidizes allylic and benzylic alcohols, is cheaper than TEMPO and operationally simple to use. [22] Chromium-based reagents such as pyridinium chlorochromate can also be used to convert alcohols to carbonyl compounds; although the stoichiometric generation of chromium wastes is a disadvantage. [23] Oxidations employing dimethyl sulfoxide, such as the Swern and Moffatt reactions, involve no heavy metals and oxidize a wide variety of substrates. [24] Oxoammonium oxidations are preferred to DMSO methods for reactions of diols and acetylenic alcohols. Dess-Martin periodinane is a highly selective, mild oxidant of alcohols, whose primary disadvantages are difficulties with preparation and safety. [25]

Related Research Articles

Ceric ammonium nitrate Chemical compound

Ceric ammonium nitrate (CAN) is the inorganic compound with the formula (NH4)2Ce(NO3)6. This orange-red, water-soluble cerium salt is a specialised oxidizing agent in organic synthesis and a standard oxidant in quantitative analysis.

Pyridinium chlorochromate Chemical compound

Pyridinium chlorochromate (PCC) is a yellow-orange salt with the formula [C5H5NH]+[CrO3Cl]. It is a reagent in organic synthesis used primarily for oxidation of alcohols to form carbonyls. A variety of related compounds are known with similar reactivity. PCC offers the advantage of the selective oxidation of alcohols to aldehydes or ketones, whereas many other reagents are less selective.

<i>N</i>-Bromosuccinimide Molecule

N-Bromosuccinimide or NBS is a chemical reagent used in radical substitution, electrophilic addition, and electrophilic substitution reactions in organic chemistry. NBS can be a convenient source of Br, the bromine radical.

Dess–Martin periodinane Chemical reagent

Dess–Martin periodinane (DMP) is a chemical reagent used in the Dess–Martin oxidation, oxidizing primary alcohols to aldehydes and secondary alcohols to ketones. This periodinane has several advantages over chromium- and DMSO-based oxidants that include milder conditions, shorter reaction times, higher yields, simplified workups, high chemoselectivity, tolerance of sensitive functional groups, and a long shelf life. However, use on an industrial scale is made difficult by its cost and its potentially explosive nature. It is named after the American chemists Daniel Benjamin Dess and James Cullen Martin who developed the reagent in 1983. It is based on IBX, but due to the acetate groups attached to the central iodine atom, DMP is much more reactive than IBX and is much more soluble in organic solvents.

<i>N</i>-Oxoammonium salt

N-Oxoammonium salts are a class of organic compounds with the formula [R1R2+N=O]X. The cation [R1R2+N=O] is of interest for the dehydrogenation of alcohols. Oxoammonium salts are diamagnetic, whereas the nitroxide has a doublet ground state. A prominent nitroxide is prepared by oxidation of (2,2,6,6-tetramethylpiperidin-1-yl)oxyl, commonly referred to as [TEMPO]+. A less expensive analogue is Bobbitt's salt.

Dihydroxylation is the process by which an alkene is converted into a vicinal diol. Although there are many routes to accomplish this oxidation, the most common and direct processes use a high-oxidation-state transition metal. The metal is often used as a catalyst, with some other stoichiometric oxidant present. In addition, other transition metals and non-transition metal methods have been developed and used to catalyze the reaction.

Oppenauer oxidation, named after Rupert Viktor Oppenauer, is a gentle method for selectively oxidizing secondary alcohols to ketones.

Cornforth reagent Chemical compound

The Cornforth reagent or pyridinium dichromate(PDC) is a pyridinium salt of dichromate with the chemical formula [C5H5NH]2[Cr2O7]. This compound is named after the Australian-British chemist Sir John Warcup Cornforth (b. 1917) who introduced it in 1962. The Cornforth reagent is a strong oxidizing agent which can convert primary and secondary alcohols to aldehydes and ketones respectively. In its chemical structure and functions it is closely related to other compounds made from hexavalent chromium oxide, such as pyridinium chlorochromate and Collins reagent. Because of their toxicity, these reagents are rarely used nowadays.

Rhodium(II) acetate Chemical compound

Rhodium(II) acetate is the coordination compound with the formula Rh2(AcO)4, where AcO is the acetate ion (CH
3
CO
2
). This dark green powder is slightly soluble in polar solvents, including water. It is used as a catalyst for cyclopropanation of alkenes. It is a widely studied example of a transition metal carboxylate complex.

Oxidation of primary alcohols to carboxylic acids

The oxidation of primary alcohols to carboxylic acids is an important oxidation reaction in organic chemistry.

Alcohol oxidation is a class of organic reactions in which the alcohol functional group is converted into another functional group in which carbon carries a higher oxidation state.

Oxidation with chromium(VI) complexes involves the conversion of alcohols to carbonyl compounds or more highly oxidized products through the action of molecular chromium(VI) oxides and salts. The principal reagents are Collins reagent, PDC, and PCC. These reagents represent improvements over inorganic chromium(VI) reagents such as Jones reagent.

Oxidation with dioxiranes refers to the introduction of oxygen into organic molecules through the action of a dioxirane. Dioxiranes are well known for their oxidation of alkenes to epoxides; however, they are also able to oxidize other unsaturated functionality, heteroatoms, and alkane C-H bonds.

Nucleophilic epoxidation is the formation of epoxides from electron-deficient double bonds through the action of nucleophilic oxidants. Nucleophilic epoxidation methods represent a viable alternative to electrophilic methods, many of which do not epoxidize electron-poor double bonds efficiently.

Manganese-mediated coupling reactions are radical coupling reactions between enolizable carbonyl compounds and unsaturated compounds initiated by a manganese(III) salt, typically manganese(III) acetate. Copper(II) acetate is sometimes used as a co-oxidant to assist in the oxidation of intermediate radicals to carbocations.

TEMPO Chemical compound

(2,2,6,6-Tetramethylpiperidin-1-yl)oxyl or (2,2,6,6-tetramethylpiperidin-1-yl)oxidanyl, commonly known as TEMPO, is a chemical compound with the formula (CH2)3(CMe2)2NO. This heterocyclic compound is a red-orange, sublimable solid. As a stable aminoxyl radical, it has applications in chemistry and biochemistry. TEMPO is used as a radical marker, as a structural probe for biological systems in conjunction with electron spin resonance spectroscopy, as a reagent in organic synthesis, and as a mediator in controlled radical polymerization.

James M. Bobbitt American chemist and professor

James McCue Bobbitt was an American chemist and academic who taught chemistry at the University of Connecticut from 1956 to 1991 and developed the Bobbitt reaction.

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.

The Stahl oxidation is a copper-catalyzed aerobic oxidation of primary and secondary alcohols to aldehydes and ketones. Known for its high selectivity and mild reaction conditions, the Stahl oxidation offers several advantages over classical alcohol oxidations.

Babler oxidation

The Babler oxidation, also known as the Babler-Dauben oxidation, is an organic reaction for the oxidative transposition of tertiary allylic alcohols to enones using pyridinium chlorochromate (PCC):

References

  1. Bobbitt, J. M.; Bruckner, C.; Merbouh, N. Org. React. 2009, 74, 103. doi : 10.1002/0471264180.or074.02
  2. Merbouh, N.; Bobbitt, J. M.; Brückner, C. J. Org. Chem.2004, 69, 5116.
  3. 1 2 Sheldon, R. A.; Arends, I. W. C. E.; ten Brink, G. J.; Dijksman, A. Acc. Chem. Res.2002, 35, 774. doi : 10.1021/ar010075n
  4. 1 2 De Mico, A.; Margarita, R.; Parlanti, L.; Vescovi, A.; Piancatelli, G. J. Org. Chem.1997, 62, 6974.
  5. Bailey, W. F.; Bobbitt, J. M.; Wiberg, K. B. J. Org. Chem.2007, 72, 4504.
  6. Hamlin, T. A.; Kelly, C. B.; Ovian, J. M.; Wiles, R. J.; Tilley, L. J.; Leadbeater, N. E. J. Org. Chem.2015, 80, 8150.
  7. Semmelhack, M. F.; Schmid, C. R.; Cortés, D. A. Tetrahedron Lett.1986, 27, 1119.
  8. Semmelhack, M. F.; Schmid, C. R.; Cortés, D. A.; Chou, C. S. J. Am. Chem. Soc.1984, 106, 3374.
  9. Rychnovsky, S. D.; McLernon, T. L.; Rajapakse, H. J. Org. Chem.1996, 61, 1194.
  10. Tanaka, H.; Kawakami, Y.; Goto, K.; Kuroboshi, M. Tetrahedron Lett.2001, 42, 445.
  11. Miyazawa, T.; Endo, T.; Shiihashi, S.; Okawara, M. J. Org. Chem.1985, 50, 1332.
  12. 1 2 3 Bobbitt, J. M. J. Org. Chem.1998, 63, 9367.
  13. 1 2 Bobbitt, J. M.; Merbouh, N. Org. Synth.2005, 82, 80.>
  14. Shibuya, Masatoshi; Tomizawa, Masaki; Iwabuchi, Yoshiharu (2008). "Oxidative Rearrangement of Tertiary Allylic Alcohols Employing Oxoammonium Salts". The Journal of Organic Chemistry. 73 (12): 4750–4752. doi:10.1021/jo800634r. ISSN   0022-3263. PMID   18500838.
  15. Song, Z. J.; Zhao, M.; Desmond, R.; Devine, P.; Tschaen, D. M.; Tillyer, R.; Frey, L.; Heid, R.; Xu, F.; Foster, B.; Li, J.; Reamer, R.; Volante, R.; Grabowski, E. J. J.; Dolling, U. H.; Reider, P. J.; Okada, S.; Kato, Y.; Mano, E. J. Org. Chem.1999, 64, 9658.
  16. Sharpless, K. B.; Amberg, W.; Bennani, Y. L.; Crispino, G. A.; Hartung, J.; Jeong, K. S.; Kwong, H. L.; Morikawa, K.; Wang, Z. M.; Xu, D.; Zhang, X. L. J. Org. Chem.1992, 57, 2768.
  17. De Mico, A.; Margarita, R.; Parlanti, L.; Vescovi, A.; Piancatelli, G. J. Org. Chem.1997, 62, 6974.
  18. Ganem, B. J. Org. Chem.1975, 40, 1998.
  19. Einhorn, J.; Einhorn, C.; Ratajczak, F.; Pierre, J.-L. J. Org. Chem.1996, 61, 7452.
  20. Bolm, C.; Magnus, A. S.; Hildebrand, J. P. Org. Lett.2000, 2, 1173.
  21. Sheldon, R. A.; Arends, I. W. C. E. Adv. Synth. Catal.2004, 346, 1051.
  22. Taylor, R. J. K.; Reid, M.; Foot, J.; Raw, S. A. Acc. Chem. Res.2005, 38, 851.
  23. Luzzio, F. A. Org. React.1998, 53, 1.
  24. Tidwell, T. T. Org. React.1990, 39, 297.
  25. Dess, D. B.; Martin, J. C. J. Am. Chem. Soc.1991, 113, 7277.