scispace - formally typeset
Search or ask a question
Author

Kazu Kurosawa

Other affiliations: Kyushu University
Bio: Kazu Kurosawa is an academic researcher from Kumamoto University. The author has contributed to research in topics: Manganese & Manganese(III) acetate. The author has an hindex of 22, co-authored 144 publications receiving 1409 citations. Previous affiliations of Kazu Kurosawa include Kyushu University.


Papers
More filters
Journal ArticleDOI
TL;DR: In this paper, the reactions of 1,1-disubstituted ethenes, styrene, 1 octene, 1-nonene, cyclohexene and cyclooctene with tris(2,4-pentanedionato)manganese(III) acetate in acetic acid at room temperature give 4-acetyl-3-hydroxy-3 -methyl-1,2-dioxacyclohexanes in 8-92 % yields.

62 citations

Journal ArticleDOI
TL;DR: In this article, the reactions of 1,1-disubstituted ethenes with barbituric acid and its derivatives in the presence of manganese(II) acetate and air yielded 5,5-bis(2-hydroperoxyalkyl)barbiturric acids 3aa-ac and 3ba-ja in 62-99% yields.
Abstract: The reactions of 1,1-disubstituted ethenes with barbituric acid and its derivatives in the presence of manganese(II) acetate and air yielded 5,5-bis(2-hydroperoxyalkyl)barbituric acids 3aa-ac and 3ba-ja in 62-99% yields. The structure of 3ab was determined by X-ray crystallography. The reaction of 1a and 2a was also effected by manganese(III) acetate (95%) and metallic manganese (92%), but in poor yield by cerium (IV)ammonium nitrate (50%)

58 citations

Journal ArticleDOI
TL;DR: In this article, the mechanisms of manganese-III-induced 1,2-dioxane ring formation and concomitant radical side reaction are discussed, and the reactions of 1,1-diphenylethene, styrene, 1-octene, cyclohexene, and cyclooctene with tris(2,4-pentanedionato)manganese(III) ([Mn(acacac)3]) in acetic acid at room temperature give 4-acetyl-3-methyl-1,2
Abstract: The reactions of 1,1-diphenylethene, 1,1-bis(4-chlorophenyl)ethene, 1,1-bis(4-methoxyphenyl)ethene, 1,1-bis(4-methylphenyl)ethene, 1,1-bis(4-fluorophenyl)ethene, styrene, 1-octene, cyclohexene, and cyclooctene with tris(2,4-pentanedionato)manganese(III) ([Mn(acac)3]) in acetic acid at room temperature give 4-acetyl-3-methyl-1,2-dioxan-3-ol in 8–92% yields, together with 3-acetyl-4-hydroxy-3-hexene-2,5-dione. The similar reactions of 1,1-diphenylethene with 2,4-pentanedione, 3-methyl-2,4-pentanedione, 3-ethyl-2,4-pentanedione, 1-phenyl-1,3-butanedione, acetoacetanilide, and 1,3-cyclohexanedione in the presence of manganese(III) acetate also give the corresponding cyclic peroxide in good to moderate yields. The mechanisms of manganese(III)-induced 1,2-dioxane ring formation and concomitant radical side reaction are discussed.

48 citations

Journal ArticleDOI
TL;DR: In this article, the reactions of substituted ethenes with β-keto esters in the presence of a mixture of manganese(II) and manganous(III) acetates, and molecular oxygen yielded substituted 1, 2-dioxan-3-ols 3 in 14-95% yields.
Abstract: The reactions of substituted ethenes with β-keto esters in the presence of a mixture of manganese(II) and manganese(III) acetates, and molecular oxygen yielded substituted 1, 2-dioxan-3-ols 3 in 14–95% yields. Cobalt(III) acetate, potassium permanganate, lead(IV) acetate, copper(II) acetate, chromium(VI) trioxide, thallium(III) acetate, ammonium cerium(IV) nitrate and iron(III) perchlorate were also used in place of manganese(III) acetate. Effects on the product yields of substituents in the alkenes and β-keto esters have been examined and reaction mechanisms are discussed.

48 citations

Journal ArticleDOI
TL;DR: In this paper, manganese(III) acetate was used to oxidize 9-substitued xanthenes to 2-(9-thioxanthenyl)-1,3-dicarbonyl compounds in 57-91% yields.
Abstract: Oxidation of xanthenes with manganese(III) acetate in the presence of active methylene compounds such as 1,3-dicarbonyl compounds, malononitrile derivatives, acetone, and nitromethane selectively gives 9-substitued xanthene derivatives in good yields. A similar oxidation of thioxanthene also yields 2-(9-thioxanthenyl)-1,3-dicarbonyl compounds in 57-91% yields. The obtained 2-(9-xanthenyl)-1,3-dicarbonyl compounds are readily converted to 2-(9-xanthenylidene)-1,3-dicarbonyl derivatives using manganese(III) complexes or 2,3-dichloro 5,6-dicyano-1,4-benzoquinone

46 citations


Cited by
More filters
Journal ArticleDOI
TL;DR: In this article, it was shown that the same alkylhydridoplatinum(IV) complex is the intermediate in the reaction of ethane with platinum(II) σ-complexes.
Abstract: ion. The oxidative addition mechanism was originally proposed22i because of the lack of a strong rate dependence on polar factors and on the acidity of the medium. Later, however, the electrophilic substitution mechanism also was proposed. Recently, the oxidative addition mechanism was confirmed by investigations into the decomposition and protonolysis of alkylplatinum complexes, which are the reverse of alkane activation. There are two routes which operate in the decomposition of the dimethylplatinum(IV) complex Cs2Pt(CH3)2Cl4. The first route leads to chloride-induced reductive elimination and produces methyl chloride and methane. The second route leads to the formation of ethane. There is strong kinetic evidence that the ethane is produced by the decomposition of an ethylhydridoplatinum(IV) complex formed from the initial dimethylplatinum(IV) complex. In D2O-DCl, the ethane which is formed contains several D atoms and has practically the same multiple exchange parameter and distribution as does an ethane which has undergone platinum(II)-catalyzed H-D exchange with D2O. Moreover, ethyl chloride is formed competitively with H-D exchange in the presence of platinum(IV). From the principle of microscopic reversibility it follows that the same ethylhydridoplatinum(IV) complex is the intermediate in the reaction of ethane with platinum(II). Important results were obtained by Labinger and Bercaw62c in the investigation of the protonolysis mechanism of several alkylplatinum(II) complexes at low temperatures. These reactions are important because they could model the microscopic reverse of C-H activation by platinum(II) complexes. Alkylhydridoplatinum(IV) complexes were observed as intermediates in certain cases, such as when the complex (tmeda)Pt(CH2Ph)Cl or (tmeda)PtMe2 (tmeda ) N,N,N′,N′-tetramethylenediamine) was treated with HCl in CD2Cl2 or CD3OD, respectively. In some cases H-D exchange took place between the methyl groups on platinum and the, CD3OD prior to methane loss. On the basis of the kinetic results, a common mechanism was proposed to operate in all the reactions: (1) protonation of Pt(II) to generate an alkylhydridoplatinum(IV) intermediate, (2) dissociation of solvent or chloride to generate a cationic, fivecoordinate platinum(IV) species, (3) reductive C-H bond formation, producing a platinum(II) alkane σ-complex, and (4) loss of the alkane either through an associative or dissociative substitution pathway. These results implicate the presence of both alkane σ-complexes and alkylhydridoplatinum(IV) complexes as intermediates in the Pt(II)-induced C-H activation reactions. Thus, the first step in the alkane activation reaction is formation of a σ-complex with the alkane, which then undergoes oxidative addition to produce an alkylhydrido complex. Reversible interconversion of these intermediates, together with reversible deprotonation of the alkylhydridoplatinum(IV) complexes, leads to multiple H-D exchange

2,505 citations

Journal ArticleDOI

1,157 citations

Journal ArticleDOI
TL;DR: This work has reported several recently reported Cu-catalyzed C-H oxidation reactions that feature substrates that are electron-deficient or appear unlikely to undergo single-electron transfer to copper(II), and evidence has been obtained for the involvement of organocopper(III) intermediates in the reaction mechanism.
Abstract: The selective oxidation of C-H bonds and the use of O(2) as a stoichiometric oxidant represent two prominent challenges in organic chemistry. Copper(II) is a versatile oxidant, capable of promoting a wide range of oxidative coupling reactions initiated by single-electron transfer (SET) from electron-rich organic molecules. Many of these reactions can be rendered catalytic in Cu by employing molecular oxygen as a stoichiometric oxidant to regenerate the active copper(II) catalyst. Meanwhile, numerous other recently reported Cu-catalyzed C-H oxidation reactions feature substrates that are electron-deficient or appear unlikely to undergo single-electron transfer to copper(II). In some of these cases, evidence has been obtained for the involvement of organocopper(III) intermediates in the reaction mechanism. Organometallic C-H oxidation reactions of this type represent important new opportunities for the field of Cu-catalyzed aerobic oxidations.

1,129 citations

Journal ArticleDOI
TL;DR: The thiol−olefin cooxidation process was applied to the total synthesis of antimalarial agent yingzhaosu A and was extended to include the more challenging 1,5-dienes, from which six-membered ring endoperoxides can be obtained.
Abstract: s a hydrogen atom from the thiol to give hydroperoxide 96 and a thiyl radical, which propagates the chain. Hydroperoxide 96 is reduced in the presence of triphenyl phosphine to give the corresponding alcohol 91. The preference for the formation of cis-3,5-disubstituted 1,2-dioxolanes is in agreement with the Beckwith−Houk transition state model for 5-exo-trig cyclizations. Similarly, the addition of thiophenol onto 5methylhepta-1,3,6-triene 97 under an atmosphere of oxygen led to 1,2-dioxolane 98, isolated in 49% as a single diastereoisomer after treatment with triphenyl phosphine, together with minor amounts of linear alcohols 99 and 100 (Scheme 49, eq b). This reaction is remarkable for a number of reasons. First, the addition of the thiyl radical occurs exclusively at the terminal position of the conjugated diene system and not at the terminal alkene, thus highlighting the higher reactivity of conjugated dienes as compared to isolated alkenes. Second, intermolecular trapping of the resulting allyl radical is reversible and regioselective under these reaction conditions. Because of the reversibility of the reaction between the allyl radical and molecular oxygen, both ratios 1,4/1,2-addition and 1,2dioxolane/linear alcohols strongly depend upon the initial concentration in thiol. Accordingly, the 1,2-dioxolanes were obtained in good yields only in highly diluted solutions. Finally, the 5-exo-trig cyclization occurs in a completely stereoselective manner, with only one of the two diastereomeric peroxyl radical intermediates (101) undergoing cyclization, while the other one (102) either leads to linear alcohol 99 or fragments back the allyl radical (Scheme 49, eq b). The reversible reaction of allyl radicals with molecular oxygen was also demonstrated for carotenoid-derived carbon-centered radical generated by Scheme 47 Scheme 48. Application to the Preparation of Functionalized 1,2,4-Trioxanes Chemical Reviews Review dx.doi.org/10.1021/cr400441m | Chem. Rev. XXXX, XXX, XXX−XXX X addition of a thiyl radical to the conjugated polyene carotene. This process has been extended to include the more challenging 1,5-dienes, from which six-membered ring endoperoxides can be obtained. Bachi and co-workers applied the thiol−olefin cooxidation process to the total synthesis of antimalarial agent yingzhaosu A (Scheme 50) and its C14epimer, as well as the preparation of a series of active analogues, from readily available limonene 103. The overall process is extremely challenging in this case due to the particular structure of the diene, with the 6-exo-cyclization process being in competition with intermolecular hydrogen atom abstraction from the thiol, and also potentially with intramolecular hydrogen abstraction from the activated allylic position by the reactive oxygen-centered radical. As previously observed, addition of the thiyl radical takes place at the less hindered position, and due to the lack of stereocontrol during the trapping of the resulting carbon-centered radical, peroxyl radical 105 is formed as a 1:1 mixture of diastereoisomers (Scheme 50). The latter undergoes 6-exo-trig cyclization to give carbon-centered radical 106. Unlike the initial trapping with molecular oxygen, the 2,3-dioxabicyclo[3.3.1]nonane system of 106 allows a highly diastereoselective reaction for the second trapping with molecular oxygen from the less hindered face to give 107. Alcohol 104 is then obtained following hydrogen abstraction from the thiol by peroxyl radical 107 and reduction of the resulting hydroperoxide with triphenylphosphine. The yields of endoperoxides remain relatively low (ca. 20−30%, calculated on the diene); however, considering the accessibility and the cost of the reactants (thiophenol, limonene, and oxygen), this approach represents a very attractive access to these structurally complex endoperoxides, some of which exhibit very promising activity for the treatment of malaria. 3.3.2. Intramolecular Trapping of the Carbon-Centered Radical. 3.3.2.a. Fragmentation Reaction: RingOpening of Vinyl Cyclopropanes. The carbon-centered radicals generated by addition of a thiyl radical onto the C C bond of vinylcypropanes have been shown to undergo cyclopropane ring-opening. The resulting radical species can then be trapped by hydrogen abstraction from the thiol. This fragmentation is a very fast process with rate constants in the range 10−10 s−1 (310 K) for most of the cyclopropylcarbinyl radicals, which allows for the fragmentation process to compete favorably with intermolecular reactions, as well as with most intramolecular processes. Alternatively, the carboncentered radical resulting from the β-fragmentation of the cyclopropylmethyl radical can engage further in carbon−carbon bond-forming processes. The allylsulfide moiety allows for the addition of radicals with concomitant release of a thiyl radical, and very elegant processes using only substoichiometric amounts of a source of thiyl radicals have been developed for the rearrangement of vinylcyclopropanes (see section 5.2.1.e). In particular, under nonreducing conditions and in the presence of an external olefin, efficient annulation reactions have been achieved, giving access to polycyclic compounds. The carboncentered radicals generated by the thiol-mediated ring-opening Scheme 49 Scheme 50 Chemical Reviews Review dx.doi.org/10.1021/cr400441m | Chem. Rev. XXXX, XXX, XXX−XXX Y of vinylcyclopropanes could also be trapped to form a new carbon−heteroatom bond. Here again, annulations taking advantage of the allylsulfide moiety have been developed (see section 5.2.1.e). Landais, Renaud, and co-workers used vinyl cyclopentenes such as 108, easily prepared by monocyclopropanation of silylcyclopentadienes, as radical acceptors for photogenerated thiyl radicals. The reversible addition of the thiyl radical onto the CC bond of 108 leads eventually to cyclopropylcarbinyl radical 110, which undergoes fragmentation to give carboncentered radical 111, stabilized by the neighboring ester group. Hydrogen atom abstraction from the thiol then furnishes cyclopentene 109 and regenerates a thiyl radical that propagates the chain (Scheme 51, eq a). The addition of the thiyl radical at the β-carbon center takes place in a highly stereoselective manner, opposite to the bulky silyl group. The fate of the stabilized carbon-centered radical resulting from the fragmentation process depends upon the reaction conditions. For instance, Naito and co-workers reported the use of vinylcylopropyl oxime ethers such as 112 in domino reactions promoted by a thiol or a disulfide in the presence of triethylborane. The ring-opening of the cyclopropyl moiety is initiated by addition of a thiyl radical onto the terminal position of vinylcyclopropyl oxime ether 112. The stabilized carboncentered radical resulting from the fragmentation process reacts with triethylborane to form a boryl enamine 115 (Scheme 51, eq b). Depending on the reaction conditions, the latter can engage further in a radical oxygenation process, leading eventually to α-hydroxy oxime ether 113 after reduction of peroxyl radical 116 by the thiol (Scheme 51, eq b). Alternatively, 113 can react with aldehydes in an ionic aldol process to give β-hydroxy oxime ethers in a stereoselective manner, as illustrated by the preparation of 117 from 112 (Scheme 51, eq c). In the aforementioned reactions, the allylsulfide moieties generated upon addition of a thiyl radical onto the vinylcyclopropane unit remain intact at the end of the reaction. However, radical reactions taking advantage of the fragmentation of allyl sulfides upon addition of radical species are also well documented. Some examples of intermolecular additions, as well as cyclization and annulation processes, will be described in section 5.2.1.e. 3.3.2.b. Rearrangement and Cyclization of Nonconjugated Dienes. In the addition of thiyl radicals onto nonconjugated dienes, the CC bonds can either react independently or lead to rearrangements through intramolecular trapping of the carbon-centered radical generated in the initial addition step. In many cyclic dienes, addition occurs selectively at the more strained double bond, and products resulting from rearrangements are often observed. For example, the addition of thiophenol to 5-methylene-norbornene led to the exo addition products 118 and 119, together with tricyclic adduct 120. The latter results from the rearrangement of homoallyl radical intermediate 121 into cyclopropylcarbinyl radical 122 (Scheme 52). Similar rearrangements have been observed in norbornadiene derivatives where substitution at C-7 can influence facial selectivity, while substitution of the methylene bridge in 7,7-dimethylnorbornene proved to have no effect in directing the addition of thiophenol. The formation of cyclopropylcarbinyl radical intermediates in norbornadiene derivatives can also lead to other skeletal rearrangements, as illustrated by the addition of thiophenol to hexachloronorbornadiene 123, which results in the formation of 125, beside the expected 1:1 addition product 124 (Scheme 53, eq a). Following addition of the thiyl radical, presumably from the less hindered endo-face, and subsequent 3-exo-trig cyclization onto the neighboring CC bond, cyclopropylcarbinyl radical 126 undergoes fragmentation to give the more stable α-chlorosubScheme 51 Scheme 52 Chemical Reviews Review dx.doi.org/10.1021/cr400441m | Chem. Rev. XXXX, XXX, XXX−XXX Z stituted carbon-centered radical 127. The latter then abstracts a hydrogen atom from the thiol to give 125 (relative configuration not established) and a thiyl radical, which goes on to propagate the chain. Similar rearrangement was observed i n t h e add i t i on o f t BuSH on to 1 , 2 , 3 , 4 , 7 , 7 hexamethylbicyclo[2.2.1]heptadiene. Likewise, Hodgson and co-workers have observed complete skeletal rearrangements in the addition of thiophenol to 7-azabicyclo[2.2.1]heptadienes such as 128 (Scheme 53, eq b). Transa

665 citations