Baeyer–Villiger oxidation

Last updated

Contents

Baeyer-Villiger oxidation
Named after Adolf von Baeyer
Victor Villiger
Reaction type Organic redox reaction
Reaction
Ketone
+
RCO3H or mCPBA
Ester or Lactone
Identifiers
Organic Chemistry Portal baeyer-villiger-oxidation
RSC ontology ID RXNO:0000031

The Baeyer–Villiger oxidation is an organic reaction that forms an ester from a ketone or a lactone from a cyclic ketone, using peroxyacids or peroxides as the oxidant. [1] The reaction is named after Adolf von Baeyer and Victor Villiger who first reported the reaction in 1899. [1]

Baeyer-Villiger oxidation Baeyer-Villiger oxidation.svg
Baeyer-Villiger oxidation

Reaction mechanism

In the first step of the reaction mechanism, the peroxyacid protonates the oxygen of the carbonyl group. [1] This makes the carbonyl group more susceptible to be attacked by the peroxyacid. [1] Next, the peroxyacid attacks the carbon of the carbonyl group forming what is known as the Criegee intermediate. [1] Through a concerted mechanism, one of the substituents on the ketone group migrates to the oxygen of the peroxide group while a carboxylic acid leaves. [1] This migration step is thought to be the rate determining step. [2] [3] Finally, deprotonation of the oxocarbenium ion produces the ester. [1]

Reaction mechanism of the Baeyer-Villiger oxidation. Baeyer-Villiger oxidation reaction mechanism.svg
Reaction mechanism of the Baeyer-Villiger oxidation.

The products of the Baeyer–Villiger oxidation are believed to be controlled through both primary and secondary stereoelectronic effects. [4] The primary stereoelectronic effect in the Baeyer–Villiger oxidation refers to the necessity of the oxygen-oxygen bond in the peroxide group to be antiperiplanar to the group that migrates. [4] [3] This orientation facilitates optimum overlap of the 𝛔 orbital of the migrating group to the 𝛔* orbital of the peroxide group. [1] The secondary stereoelectronic effect refers to the necessity of the lone pair on the oxygen of the hydroxyl group to be antiperiplanar to the migrating group. [4] This allows for optimum overlap of the oxygen nonbonding orbital with the 𝛔* orbital of the migrating group. [5] This migration step is also (at least in silico ) assisted by two or three peroxyacid units enabling the hydroxyl proton to shuttle to its new position. [6]

Stereoelectronic effects Baeyer-Villiger oxidation stereoelectronic effects.svg
Stereoelectronic effects

The migratory ability is ranked tertiary > secondary > aryl > primary. [7] Allylic groups are more apt to migrate than primary alkyl groups but less so than secondary alkyl groups. [5] Electron-withdrawing groups on the substituent decrease the rate of migration. [8] There are two explanations for this trend in migration ability. [9] One explanation relies on the buildup of positive charge in the transition state for breakdown of the Criegee intermediate (illustrated by the carbocation resonance structure of the Criegee intermediate). [9] Keeping this structure in mind, it makes sense that the substituent that can maintain positive charge the best would be most likely to migrate. [9] The higher the degree of substitution, the more stable a carbocation generally is. [10] Therefore, the tertiary > secondary > primary trend is observed.

Resonance structures of the Criegee intermediate Criegee intermediate resonance structures.svg
Resonance structures of the Criegee intermediate

Another explanation uses stereoelectronic effects and steric arguments. [11] As mentioned, the substituent that is antiperiplanar to the peroxide group in the transition state will migrate. [4] This transition state has a gauche interaction between the peroxyacid and the non-migrating substituent. [11] If the bulkier group is placed antiperiplanar to the peroxide group, the gauche interaction between the substituent on the forming ester and the carbonyl group of the peroxyacid will be reduced. [11] Thus, it is the bulkier group that will prefer to be antiperiplanar to the peroxide group, enhancing its aptitude for migration. [11]

Steric bulk influencing migration Criegee intermediate stereoelectronics.svg
Steric bulk influencing migration

The migrating group in acyclic ketones, usually, is not 1° alkyl group. However, they may be persuaded to migrate in preference to the 2° or 3° groups by using CF3CO3H or BF3 + H2O2 as reagents. [12]

Historical background

In 1899, Adolf Baeyer and Victor Villiger first published a demonstration of the reaction that we now know as the Baeyer–Villiger oxidation. [13] [14] They used peroxymonosulfuric acid to make the corresponding lactones from camphor, menthone, and tetrahydrocarvone. [14] [15]

Original reactions reported by Baeyer and Villiger Original Baeyer-Villiger oxidation reactions.png
Original reactions reported by Baeyer and Villiger

There were three suggested reaction mechanisms of the Baeyer–Villiger oxidation that seemed to fit with observed reaction outcomes. [16] These three reaction mechanisms can really be split into two pathways of peroxyacid attack – on either the oxygen or the carbon of the carbonyl group. [17] Attack on oxygen could lead to two possible intermediates: Baeyer and Villiger suggested a dioxirane intermediate, while Georg Wittig and Gustav Pieper suggested a peroxide with no dioxirane formation. [17] Carbon attack was suggested by Rudolf Criegee. [17] In this pathway, the peracid attacks the carbonyl carbon, producing what is now known as the Criegee intermediate. [17]

Proposed Baeyer-Villiger oxidation intermediates Proposed Baeyer Villiger Intermediates.png
Proposed Baeyer-Villiger oxidation intermediates

In 1953, William von Eggers Doering and Edwin Dorfman elucidated the correct pathway for the reaction mechanism of the Baeyer–Villiger oxidation by using oxygen-18-labelling of benzophenone. [16] The three different mechanisms would each lead to a different distribution of labelled products. The Criegee intermediate would lead to a product only labelled on the carbonyl oxygen. [16] The product of the Wittig and Pieper intermediate is only labeled on the alkoxy group of the ester. [16] The Baeyer and Villiger intermediate leads to a 1:1 distribution of both of the above products. [16] The outcome of the labelling experiment supported the Criegee intermediate, [16] which is now the generally accepted pathway. [1]

The different possible outcomes of Dorfman and Doering's labelling experiment Dorfman and Doering's Labelling Experiment.png
The different possible outcomes of Dorfman and Doering's labelling experiment

Stereochemistry

The migration does not change the stereochemistry of the group that transfers, i.e.: it is stereoretentive. [18] [19]

Reagents

Although many different peroxyacids are used for the Baeyer–Villiger oxidation, some of the more common oxidants include meta-chloroperbenzoic acid (mCPBA) and trifluoroperacetic acid (TFPAA). [2] The general trend is that higher reactivity is correlated with lower pKa (i.e.: stronger acidity) of the corresponding carboxylic acid (or alcohol in the case of the peroxides). [5] Therefore, the reactivity trend shows TFPAA > 4-nitroperbenzoic acid > mCPBA and performic acid > peracetic acid > hydrogen peroxide > tert-butyl hydroperoxide. [5] The peroxides are much less reactive than the peroxyacids. [2] The use of hydrogen peroxide even requires a catalyst. [7] [20] In addition, using organic peroxides and hydrogen peroxide tends to generate more side-reactivity due to their promiscuity. [21]

Limitations

The use of peroxyacids and peroxides when performing the Baeyer–Villiger oxidation can cause the undesirable oxidation of other functional groups. [22] Alkenes and amines are a few of the groups that can be oxidized. [22] For instance, alkenes in the substrate, particularly when electron-rich, may be oxidized to epoxides. [22] [23] However, methods have been developed that will allow for the tolerance of these functional groups. [22] In 1962, G. B. Payne reported that the use of hydrogen peroxide in the presence of a selenium catalyst will produce the epoxide from alkenyl ketones, while use of peroxyacetic acid will form the ester. [24]

Payne reported that different reagents will give different outcomes when there are more than one functional group Reagent Dependent Oxidation.png
Payne reported that different reagents will give different outcomes when there are more than one functional group

Modifications

Catalytic Baeyer-Villiger oxidation

The use of hydrogen peroxide as an oxidant would be advantageous, making the reaction more environmentally friendly as the sole byproduct is water. [7] Benzeneseleninic acid derivatives as catalysts have been reported to give high selectivity with hydrogen peroxide as the oxidant. [25] Another class of catalysts which show high selectivity with hydrogen peroxide as the oxidant are solid Lewis acid catalysts such as stannosilicates. [26] Among stannosilicates, particularly the zeotype Sn-beta and the amorphous Sn-MCM-41 show promising activity and close to full selectivity towards the desired product. [27] [28]

Asymmetric Baeyer-Villiger oxidation

There have been attempts to use organometallic catalysts to perform enantioselective Baeyer–Villiger oxidations. [7] The first reported instance of one such oxidation of a prochiral ketone used dioxygen as the oxidant with a copper catalyst. [23] Other catalysts, including platinum and aluminum compounds, followed. [23]

Baeyer-Villiger monooxygenases

Reaction mechanism of the flavin cofactor to catalyse the Baeyer-Villiger reaction in Baeyer-Villiger monooxygenase enzymes. BVMO reaction mechanism.png
Reaction mechanism of the flavin cofactor to catalyse the Baeyer-Villiger reaction in Baeyer-Villiger monooxygenase enzymes.

In nature, enzymes called Baeyer-Villiger monooxygenases (BVMOs) perform the oxidation analogously to the chemical reaction. [29] To facilitate this chemistry, BVMOs contain a flavin adenine dinucleotide (FAD) cofactor. [30] In the catalytic cycle (see figure on the right), the cellular redox equivalent NADPH first reduces the cofactor, which allows it subsequently to react with molecular oxygen. The resulting peroxyflavin is the catalytic entity oxygenating the substrate, and theoretical studies suggest that the reaction proceeds through the same Criegee intermediate as observed in the chemical reaction. [31] After the rearrangement step forming the ester product, a hydroxyflavin remains, which spontaneously eliminates water to form oxidized flavin, thereby closing the catalytic cycle.

BVMOs are closely related to the flavin-containing monooxygenases (FMOs), [32] enzymes that also occur in the human body, functioning within the frontline metabolic detoxification system of the liver along the cytochrome P450 monooxygenases. [33] Human FMO5 was in fact shown to be able to catalyse Baeyer-Villiger reactions, indicating that the reaction may occur in the human body as well. [34]

BVMOs have been widely studied due to their potential as biocatalysts, that is, for an application in organic synthesis. [35] Considering the environmental concerns for most of the chemical catalysts, the use of enzymes is considered a greener alternative. [29] BVMOs in particular are interesting for application because they fulfil a range of criteria typically sought for in biocatalysis: besides their ability to catalyse a synthetically useful reaction, some natural homologs were found to have a very large substrate scope (i.e. their reactivity was not restricted to a single compound, as often assumed in enzyme catalysis), [36] they can be easily produced on a large scale, and because the three-dimensional structure of many BVMOs has been determined, enzyme engineering could be applied to produce variants with improved thermostability and/or reactivity. [37] [38] Another advantage of using enzymes for the reaction is their frequently observed regio- and enantioselectivity, owed to the steric control of substrate orientation during catalysis within the enzyme's active site. [29] [35]

Applications

Zoapatanol

Zoapatanol is a biologically active molecule that occurs naturally in the zeopatle plant, which has been used in Mexico to make a tea that can induce menstruation and labor. [39] In 1981, Vinayak Kane and Donald Doyle reported a synthesis of zoapatanol. [40] [41] They used the Baeyer–Villiger oxidation to make a lactone that served as a crucial building block that ultimately led to the synthesis of zoapatanol. [40] [41]

Kane and Doyle used a Baeyer-Villiger oxidation to synthesize zoapatanol Synthesis of Zoapatanol.png
Kane and Doyle used a Baeyer-Villiger oxidation to synthesize zoapatanol

Steroids

In 2013, Alina Świzdor reported the transformation of the steroid dehydroepiandrosterone to anticancer agent testololactone by use of a Baeyer–Villiger oxidation induced by fungus that produces Baeyer-Villiger monooxygenases. [42]

Swizdor reported that a Baeyer-Villiger monooxygenase could change dehydroepiandrosterone into testololactone Dehydroepiandrosterone to testololactone.png
Świzdor reported that a Baeyer-Villiger monooxygenase could change dehydroepiandrosterone into testololactone

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">Epoxide</span> Organic compounds with a carbon-carbon-oxygen ring

In organic chemistry, an epoxide is a cyclic ether, where the ether forms a three-atom ring: two atoms of carbon and one atom of oxygen. This triangular structure has substantial ring strain, making epoxides highly reactive, more so than other ethers. They are produced on a large scale for many applications. In general, low molecular weight epoxides are colourless and nonpolar, and often volatile.

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

Acetone peroxide is an organic peroxide and a primary explosive. It is produced by the reaction of acetone and hydrogen peroxide to yield a mixture of linear monomer and cyclic dimer, trimer, and tetramer forms. The dimer is known as diacetone diperoxide (DADP). The trimer is known as triacetone triperoxide (TATP) or tri-cyclic acetone peroxide (TCAP). Acetone peroxide takes the form of a white crystalline powder with a distinctive bleach-like odor or a fruit-like smell when pure, and can explode powerfully if subjected to heat, friction, static electricity, concentrated sulfuric acid, strong UV radiation or shock. Until about 2015, explosives detectors were not set to detect non-nitrogenous explosives, as most explosives used preceding 2015 were nitrogen-based. TATP, being nitrogen-free, has been used as the explosive of choice in several terrorist bomb attacks since 2001.

<span class="mw-page-title-main">Hydrogen peroxide - urea</span> Chemical compound

Hydrogen peroxide - urea is a white crystalline solid chemical compound composed of equal amounts of hydrogen peroxide and urea. It contains solid and water-free hydrogen peroxide, which offers a higher stability and better controllability than liquid hydrogen peroxide when used as an oxidizing agent. Often called carbamide peroxide in dentistry, it is used as a source of hydrogen peroxide when dissolved in water for bleaching, disinfection and oxidation.

In organic chemistry, ozonolysis is an organic reaction where the unsaturated bonds are cleaved with ozone. Multiple carbon–carbon bond are replaced by carbonyl groups, such as aldehydes, ketones, and carboxylic acids. The reaction is predominantly applied to alkenes, but alkynes and azo compounds are also susceptible to cleavage. The outcome of the reaction depends on the type of multiple bond being oxidized and the work-up conditions.

<span class="mw-page-title-main">Organoboron chemistry</span> Study of compounds containing a boron-carbon bond

Organoboron chemistry or organoborane chemistry studies organoboron compounds, also called organoboranes. These chemical compounds combine boron and carbon; typically, they are organic derivatives of borane (BH3), as in the trialkyl boranes.

<span class="mw-page-title-main">Peroxy acid</span> Organic acid having a peroxide bond

A peroxy acid is an acid which contains an acidic –OOH group. The two main classes are those derived from conventional mineral acids, especially sulfuric acid, and the peroxy derivatives of organic carboxylic acids. They are generally strong oxidizers.

The Criegee rearrangement is a rearrangement reaction named after Rudolf Criegee.

<span class="mw-page-title-main">Dakin oxidation</span> Organic redox reaction that converts hydroxyphenyl aldehydes or ketones into benzenediols

The Dakin oxidation (or Dakin reaction) is an organic redox reaction in which an ortho- or para-hydroxylated phenyl aldehyde (2-hydroxybenzaldehyde or 4-hydroxybenzaldehyde) or ketone reacts with hydrogen peroxide (H2O2) in base to form a benzenediol and a carboxylate. Overall, the carbonyl group is oxidised, whereas the H2O2 is reduced.

<span class="mw-page-title-main">Shi epoxidation</span>

The Shi epoxidation is a chemical reaction described as the asymmetric epoxidation of alkenes with oxone and a fructose-derived catalyst (1). This reaction is thought to proceed via a dioxirane intermediate, generated from the catalyst ketone by oxone. The addition of the sulfate group by the oxone facilitates the formation of the dioxirane by acting as a good leaving group during ring closure. It is notable for its use of a non-metal catalyst and represents an early example of organocatalysis. The reaction was first discovered by Yian Shi of Colorado State University in 1996.

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.

<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 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.

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

A seleninic acid is an organoselenium compound and an oxoacid with the general formula RSeO2H, where R ≠ H. It is a member of the family of organoselenium oxoacids, which also includes selenenic acids and selenonic acids, which are RSeOH and RSeO3H, respectively. The parent member of this family of compounds is methaneseleninic acid (CH3SeO2H), also known as methylseleninic acid or "MSA".

Cyclohexanone monooxygenase (EC 1.14.13.22, cyclohexanone 1,2-monooxygenase, cyclohexanone oxygenase, cyclohexanone:NADPH:oxygen oxidoreductase (6-hydroxylating, 1,2-lactonizing)) is an enzyme with systematic name cyclohexanone,NADPH:oxygen oxidoreductase (lactone-forming). This enzyme catalyses the following chemical reaction

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

The White–Chen catalyst is an Iron-based coordination complex named after Professor M. Christina White and her graduate student Mark S. Chen. The catalyst is used along with hydrogen peroxide and acetic acid additive to oxidize aliphatic sp3 C-H bonds in organic synthesis. The catalyst is the first to allow for preparative and predictable aliphatic C–H oxidations over a broad range of organic substrates. Oxidations with the catalyst have proven to be remarkably predictable based on sterics, electronics, and stereoelectronics allowing for aliphatic C–H bonds to be thought of as a functional group in the streamlining of organic synthesis.

<span class="mw-page-title-main">Stereoelectronic effect</span> Affect on molecular properties due to spatial arrangement of electron orbitals

In chemistry, primarily organic and computational chemistry, a stereoelectronic effect is an effect on molecular geometry, reactivity, or physical properties due to spatial relationships in the molecules' electronic structure, in particular the interaction between atomic and/or molecular orbitals. Phrased differently, stereoelectronic effects can also be defined as the geometric constraints placed on the ground and/or transition states of molecules that arise from considerations of orbital overlap. Thus, a stereoelectronic effect explains a particular molecular property or reactivity by invoking stabilizing or destabilizing interactions that depend on the relative orientations of electrons in space.

<span class="mw-page-title-main">Alkenyl peroxides</span> Organic compounds of the form R2C=C(R)OOR

In organic chemistry, alkenyl peroxides are organic compounds bearing an alkene residue directly at the peroxide group, resulting in the general formula R2C=C(R)OOR. They have very weak O-O bonds and are thus generally unstable compounds.

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

Trifluoroperacetic acid is an organofluorine compound, the peroxy acid analog of trifluoroacetic acid, with the condensed structural formula CF
3
COOOH
. It is a strong oxidizing agent for organic oxidation reactions, such as in Baeyer–Villiger oxidations of ketones. It is the most reactive of the organic peroxy acids, allowing it to successfully oxidise relatively unreactive alkenes to epoxides where other peroxy acids are ineffective. It can also oxidise the chalcogens in some functional groups, such as by transforming selenoethers to selones. It is a potentially explosive material and is not commercially available, but it can be quickly prepared as needed. Its use as a laboratory reagent was pioneered and developed by William D. Emmons.

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

MoOPH, also known as oxodiperoxymolybdenum(pyridine)-(hexamethylphosphoric triamide), is a reagent used in organic synthesis. It contains a molybdenum(VI) center with multiple oxygen ligands, coordinated with pyridine and HMPA ligands. It is an electrophilic source of oxygen that reacts with enolates and related structures, and thus can be used for alpha-hydroxylation of carbonyl-containing compounds. Other reagents used for alpha-hydroxylation via enol or enolate structures include Davis oxaziridine, oxygen, and various peroxyacids. This reagent was first utilized by Edwin Vedejs as an efficient alpha-hydroxylating agent in 1974 and an effective preparative procedure was later published in 1978.

References

  1. 1 2 3 4 5 6 7 8 9 Kürti, László; Czakó, Barbara (2005). Strategic Applications of Named Reactions in Organic Synthesis . Burlington; San Diego; London: Elsevier Academic Press. p.  28. ISBN   978-0-12-369483-6.
  2. 1 2 3 Krow, Grant R. (1993). "The Baeyer-Villiger Oxidation of Ketones and Aldehydes". Organic Reactions. 43 (3): 251–798. doi:10.1002/0471264180.or043.03. ISBN   0471264180.
  3. 1 2 Carey, Francis A.; Sundberg, Richard J. (2007). Advanced Organic Chemistry: Part B: Reactions and Synthesis (5th ed.). New York: Springer. p. 1135. ISBN   978-0387683546.
  4. 1 2 3 4 Crudden, Cathleen M.; Chen, Austin C.; Calhoun, Larry A. (2000). "A Demonstration of the Primary Stereoelectronic Effect in the Baeyer-Villiger Oxidation of α-Fluorocyclohexanones". Angew. Chem. Int. Ed. 39 (16): 2851–2855. doi:10.1002/1521-3773(20000818)39:16<2851::aid-anie2851>3.0.co;2-y. PMID   11027987.
  5. 1 2 3 4 Myers, Andrew G. "Chemistry 115 Handouts: Oxidation" (PDF). Harvard University.
  6. Yamabe, Shinichi (2007). "The Role of Hydrogen Bonds in Baeyer−Villiger Reactions". The Journal of Organic Chemistry. 72 (8): 3031–3041. doi:10.1021/jo0626562. PMID   17367197.
  7. 1 2 3 4 ten Brink, G.-J.; Arends, W. C. E.; Sheldon, R. A. (2004). "The Baeyer-Villiger Reaction: New Developments toward Greener Procedures". Chem. Rev. 104 (9): 4105–4123. doi:10.1021/cr030011l. PMID   15352787.
  8. Li, Jie Jack; Corey, E. J., eds. (2007). Name Reactions of Functional Group Transformations. Hoboken, NJ: Wiley-Interscience.
  9. 1 2 3 Hawthorne, M. Frederick; Emmons, William D.; McCallum, K. S. (1958). "A Re-examination of the Peroxyacid Cleavage of Ketones. I. Relative Migratory Aptitudes". J. Am. Chem. Soc. 80 (23): 6393–6398. doi:10.1021/ja01556a057.
  10. Jones, Jr., Maitland; Fleming, Steven A. (2010). Organic Chemistry (4th ed.). Canada: W. W. Norton & Company. p.  293. ISBN   978-0-393-93149-5.
  11. 1 2 3 4 Evans, D. A. "Stereoelectronic Effects-2". Chemistry 206 (Fall 2006-2007).
  12. Sanyal, S.N. (2003). Reactions, Rearrangements and Reagents (4 ed.). p. 90. ISBN   978-81-7709-605-7.
  13. Baeyer, Adolf; Villiger, Victor (1899). "Einwirkung des Caro'schen Reagens auf Ketone". Ber. Dtsch. Chem. Ges. 32 (3): 3625–3633. doi:10.1002/cber.189903203151.
  14. 1 2 Hassall, C. H. (1957). "The Baeyer-Villiger Oxidation of Aldehydes and Ketones". Organic Reactions. 9 (3): 73–106. doi:10.1002/0471264180.or009.03. ISBN   0471264180.
  15. Renz, Michael; Meunier, Bernard (1999). "100 Years of Baeyer-Villiger Oxidations". Eur. J. Org. Chem. 1999 (4): 737–750. doi:10.1002/(SICI)1099-0690(199904)1999:4<737::AID-EJOC737>3.0.CO;2-B.
  16. 1 2 3 4 5 6 Doering, W. von E.; Dorfman, Edwin (1953). "Mechanism of the Peracid Ketone-Ester Conversion. Analysis of Organic Compounds for Oxygen-18". J. Am. Chem. Soc. 75 (22): 5595–5598. doi:10.1021/ja01118a035.
  17. 1 2 3 4 Doering, W. von E.; Speers, Louise (1950). "The Peracetic Acid Cleavage of Unsymmetrical Ketones". Journal of the American Chemical Society. 72 (12): 5515–5518. doi:10.1021/ja01168a041.
  18. Turner, Richard B. (1950). "Stereochemistry of the Peracid Oxidation of Ketones". J. Am. Chem. Soc. 72 (2): 878–882. doi:10.1021/ja01158a061.
  19. Gallagher, T. F.; Kritchevsky, Theodore H. (1950). "Perbenzoic Acid Oxidation of 20-Ketosteroids and the Stereochemistry of C-17". J. Am. Chem. Soc. 72 (2): 882–885. doi:10.1021/ja01158a062.
  20. Cavarzan, Alessandra; Scarso, Alessandro; Sgarbossa, Paolo; Michelin, Rino A.; Strukul, Giorgio (2010). "Green Catalytic Baeyer–Villiger Oxidation with Hydrogen Peroxide in Water Mediated by Pt(II) Catalysts". ChemCatChem. 2 (10): 1296–1302. doi:10.1002/cctc.201000088. S2CID   98508888.
  21. Schweitzer-Chaput, Bertrand; Kurtén, Theo; Klussmann, Martin (2015). "Acid-Mediated Formation of Radicals or Baeyer-Villiger Oxidation from Criegee Adducts". Angewandte Chemie International Edition. 54 (40): 11848–11851. doi:10.1002/anie.201505648. PMID   26267787.
  22. 1 2 3 4 Grant R. Krow (1991). Trost, Barry M.; Fleming, Ian (eds.). Comprehensive Organic Synthesis – Selectivity, Strategy and Efficiency in Modern Organic Chemistry, Volumes 1 – 9. Elsevier. pp. 671–688. ISBN   978-0-08-035930-4.
  23. 1 2 3 Seymour, Craig. "Page 1 The Asymmetric Baeyer-Villiger Oxidation" (PDF). scs.illinois.edu.
  24. Payne, G. B. (1962). "A Simplified Procedure for Epoxidation by Benzonitrile-Hydrogen Peroxide. Selective Oxidation of 2-Allylcyclohexanone". Tetrahedron. 18 (6): 763–765. doi:10.1016/S0040-4020(01)92726-7.
  25. ten Brink, Gerd-Jan; Vis, Jan-Martijn; Arends, Isabel W. C. E.; Sheldon, Roger A. (2001). "Selenium-Catalyzed Oxidations with Aqueous Hydrogen Peroxide. 2. Baeyer−Villiger Reactions in Homogeneous Solution". J. Org. Chem. 66 (7): 2429–2433. doi:10.1021/jo0057710. PMID   11281784.
  26. Ferrini, Paola; Dijkmans, Jan; Clercq, Rik De; Vyver, Stijn Van de; Dusselier, Michiel; Jacobs, Pierre A.; Sels, Bert F. (2017). "Lewis acid catalysis on single site Sn centers incorporated into silica hosts". Coordination Chemistry Reviews. 343: 220–255. doi:10.1016/j.ccr.2017.05.010.
  27. Corma, A; Navarro, MT; Nemeth, L; Renz, M (7 November 2001). "Sn-MCM-41—a heterogeneous selective catalyst for the Baeyer-Villiger oxidation with hydrogen peroxide". Chemical Communications (21): 2190–1. doi:10.1039/B105927K. ISSN   1364-548X. PMID   12240094.
  28. Renz, M; Blasco, T; Corma, A; Fornés, V; Jensen, R; Nemeth, L (18 October 2002). "Selective and shape-selective Baeyer-Villiger oxidations of aromatic aldehydes and cyclic ketones with Sn-beta zeolites and H2O2". Chemistry: A European Journal. 8 (20): 4708–17. doi:10.1002/1521-3765(20021018)8:20<4708::AID-CHEM4708>3.0.CO;2-U. ISSN   1521-3765. PMID   12561111.
  29. 1 2 3 Leisch, Hannes; Morley, Krista; Lau, Peter C. K. (13 July 2011). "Baeyer−Villiger Monooxygenases: More Than Just Green Chemistry". Chemical Reviews. 111 (7): 4165–4222. doi:10.1021/cr1003437. ISSN   0009-2665. PMID   21542563.
  30. Sheng, Dawei; Ballou, David P.; Massey, Vincent (1 September 2001). "Mechanistic Studies of Cyclohexanone Monooxygenase: Chemical Properties of Intermediates Involved in Catalysis". Biochemistry. 40 (37): 11156–11167. doi:10.1021/bi011153h. ISSN   0006-2960. PMID   11551214.
  31. Polyak, Iakov; Reetz, Manfred T.; Thiel, Walter (8 February 2012). "Quantum Mechanical/Molecular Mechanical Study on the Mechanism of the Enzymatic Baeyer–Villiger Reaction". Journal of the American Chemical Society. 134 (5): 2732–2741. doi:10.1021/ja2103839. ISSN   0002-7863. PMID   22239272.
  32. van Berkel, W. J. H.; Kamerbeek, N. M.; Fraaije, M. W. (5 August 2006). "Flavoprotein monooxygenases, a diverse class of oxidative biocatalysts". Journal of Biotechnology. 124 (4): 670–689. doi:10.1016/j.jbiotec.2006.03.044. hdl: 11370/99a1ac5c-d4a4-4612-90a3-4fe1d4d03a11 . ISSN   0168-1656. PMID   16712999.
  33. Iyanagi, Takashi (1 January 2007). "Molecular Mechanism of Phase I and Phase II Drug‐Metabolizing Enzymes: Implications for Detoxification". International Review of Cytology. Academic Press. 260: 35–112. doi:10.1016/S0074-7696(06)60002-8. ISBN   9780123741141. PMID   17482904.
  34. Fiorentini, Filippo; Geier, Martina; Binda, Claudia; Winkler, Margit; Faber, Kurt; Hall, Mélanie; Mattevi, Andrea (15 April 2016). "Biocatalytic Characterization of Human FMO5: Unearthing Baeyer–Villiger Reactions in Humans". ACS Chemical Biology. 11 (4): 1039–1048. doi:10.1021/acschembio.5b01016. ISSN   1554-8929. PMID   26771671.
  35. 1 2 Fürst, Maximilian J. L. J.; Gran-Scheuch, Alejandro; Aalbers, Friso S.; Fraaije, Marco W. (6 December 2019). "Baeyer–Villiger Monooxygenases: Tunable Oxidative Biocatalysts". ACS Catalysis. 9 (12): 11207–11241. doi: 10.1021/acscatal.9b03396 .
  36. Fürst, Maximilian J. L. J.; Romero, Elvira; Gómez Castellanos, J. Rúben; Fraaije, Marco W.; Mattevi, Andrea (7 December 2018). "Side-Chain Pruning Has Limited Impact on Substrate Preference in a Promiscuous Enzyme". ACS Catalysis. 8 (12): 11648–11656. doi: 10.1021/acscatal.8b03793 . PMC   6345240 . PMID   30687578.
  37. Fürst, Maximilian J. L. J.; Boonstra, Marjon; Bandstra, Selle; Fraaije, Marco W. (2019). "Stabilization of cyclohexanone monooxygenase by computational and experimental library design". Biotechnology and Bioengineering. 116 (9): 2167–2177. doi: 10.1002/bit.27022 . ISSN   1097-0290. PMC   6836875 . PMID   31124128.
  38. Li, Guangyue; Garcia-Borràs, Marc; Fürst, Maximilian J. L. J.; Ilie, Adriana; Fraaije, Marco W.; Houk, K. N.; Reetz, Manfred T. (22 August 2018). "Overriding Traditional Electronic Effects in Biocatalytic Baeyer–Villiger Reactions by Directed Evolution". Journal of the American Chemical Society. 140 (33): 10464–10472. doi:10.1021/jacs.8b04742. ISSN   0002-7863. PMC   6314816 . PMID   30044629.
  39. Levine, Seymour D.; Adams, Richard E.; Chen, Robert; Cotter, Mary Lou; Hirsch, Allen F.; Kane, Vinayak V.; Kanojia, Ramesh M.; Shaw, Charles; Wachter, Michael P.; Chin, Eva; Huettemann, Richard; Ostrowski, Paul (1979). "Zoapatanol and Montanol, Novel Oxepane Diterpenoids, from the Mexican Plant Zoapatle (Montanoa tomentosa)". J. Am. Chem. Soc. 101 (12): 3405–3407. doi:10.1021/ja00506a057.
  40. 1 2 Kane, Vinayak V.; Doyle, Donald L. (1981). "Total Synthesis of (±) Zoapatanol: A Stereospecific Synthesis of a Key Intermediate". Tetrahedron Lett. 22 (32): 3027–3030. doi:10.1016/S0040-4039(01)81818-9.
  41. 1 2 Kane, Vinayak V.; Doyle, Donald L. (1981). "Total Synthesis of (±) Zoapatanol". Tetrahedron Lett. 22 (32): 3031–3034. doi:10.1016/S0040-4039(01)81819-0.
  42. Świzdor, Alina (2013). "Baeyer-Villiger Oxidation of Some C19 Steroids by Penicillium lanosocoeruleum". Molecules. 18 (11): 13812–13822. doi: 10.3390/molecules181113812 . PMC   6270215 . PMID   24213656.