研究者業績

藤田 守文

フジタ モリフミ  (Morifumi Fujita)

基本情報

所属
兵庫県立大学 大学院 物質理学研究科 准教授
学位
工学修士(大阪大学)
工学博士(大阪大学)

J-GLOBAL ID
200901082229941660
researchmap会員ID
1000195344

外部リンク

論文

 31
  • Takeshi Kuri, Yoshihiko Mizukami, Mio Shimogaki, Morifumi Fujita
    Organic Letters 22(19) 7613-7616 2020年10月2日  査読有り
    An oxetane intermediate during a direct aldol reaction was trapped with an internal aryl group to yield trans-tetraline products. The contribution of the oxetane intermediate was confirmed by 18O-isotope labeling experiments.
  • Makoto Nakatsuji, Morifumi Fujita, Yasuaki Okamoto, Takashi Sugimura
    Catalysis Science & Technology 10(19) 6573-6582 2020年  査読有り
    <p>The substitutions at the 2′- and/or 6′-positions of the quinoline ring of cinchonidine reduce both the intrinsic enantioselectivity and adsorption strength for the enantioselective hydrogenation of α-phenylcinnamic acid over Pd/C.</p>
  • Bokeun Kim, Makoto Nakatsuji, Takuya Mameda, Takeshi Kubota, Morifumi Fujita, Takashi Sugimura, Yasuaki Okamoto
    Bulletin of the Chemical Society of Japan 93(1) 163-175 2020年1月1日  査読有り
    © 2020 The Chemical Society of Japan. Enantioselective hydrogenations of α,β-unsaturated carboxylic acids over cinchona alkaloid-modified Pd metal heterogeneous catalysts have received considerable attention because of scientific importance in molecular recognition catalysis as well as feasibility of industrial applications. In the present study, comprehensive kinetic analysis of the hydrogenation was conducted to disclose the crucial kinetic parameters controlling enantiodifferentiation and reaction rate with the combinations of four kinds of modifier and three kinds of substrate. Despite simplicity of the kinetic model, the present novel kinetic formulation allows us to describe the enantioselectivity as a function of modifier concentration, to estimate intrinsic enantioselectivity at the modified sites, to estimate respective reaction rates at the modified and unmodified sites, and to establish a correlation between the magnitude of ligand acceleration and kinetic parameters. The enantioselectivity is successfully correlated to the reaction rate. The adsorption strength of the modifier on Pd is suggested to decrease in the order, cinchonidine > cinchonine > quinine > quinidine. The roles played by benzylammine and the observed decrease in the selectivity at a high modifier concentration are also discussed. The kinetic model and formulation can be applied to analyze the catalytic behaviors and performance of Pt counterparts.
  • Makoto Nakatsuji, Takeshi Kubota, Morifumi Fujita, Yasuaki Okamoto, Takashi Sugimura
    Catalysis Letters 2020年  査読有り
    © 2020, Springer Science+Business Media, LLC, part of Springer Nature. Abstract: The asymmetric hydrogenation of (E)-2,3-di(4-methoxyphenyl)propenoic acid was conducted over Pd/C chirally modified with cinchona alkaloids methyl-substituted at 2′-position of the quinoline ring. It is revealed that the adsorption strength of the modifiers and the intrinsic enantioselectivity at the modified sites are decreased by the methyl-substitution. The intrinsic enantioselectivity of the modifier is correlated to kinetic parameters. Graphic Abstract: [Figure not available: see fulltext.]
  • Azka Azkiya Choliq, Rio Nakae, Mariko Watanabe, Tomonori Misaki, Morifumi Fujita, Yasuaki Okamoto, Takashi Sugimura
    Bulletin of the Chemical Society of Japan 92(7) 1175-1180 2019年  査読有り
    © 2019 The Chemical Society of Japan To ensure high enantiopurity of the product, enantio-differentiating hydrogenation of methyl acetoacetate over a (R,R)-tartaric acid-modified Raney nickel catalyst is normally performed under elevated H2-pressure (310 MPa). In this study, higher enantioselectivity than previously reported for methyl acetoacetate was achieved (92% ee) under low H2pressure of 0.42 MPa. Effects of reaction conditions on the enantioselectivity and hydrogenation rate were investigated using a low-pressure reaction system (<0.5 MPa of H2). It was found that impurities in the solvent greatly reduce the enantioselectivity of MAA. The low-pressure reaction system enabled a satisfactory kinetic approach. The reaction rate was well described by Langmuir-Hinshelwood formalism, verifying the previous assumption that the addition of adsorbed hydrogen to the substrate interacting with surface tartrate is a rate-determining step.

MISC

 116
  • FUKUZUMI S, FUJITA M, NOURA S, OHKUBO K, SUENOBU T, ARAKI Y, ITO O
    The Journal of Physical Chemistry A 105(10) 1857-1868 2001年  
  • T Okuyama, M Fujita, R Gronheid, G Lodder
    TETRAHEDRON LETTERS 41(26) 5125-5129 2000年6月  
    Alkenyl(phenyl)iodonium tetrafluoroborates dissolved in chloroform, or in the solid state, decompose thermally at 60 degrees C to yield fluoroalkenes and iodobenzene as major products via an S(N)1- or S(N)2-type reaction within the ion pair of the substrates. (C) 2000 Elsevier Science Ltd. All rights reserved.
  • T Okuyama, M Fujita, R Gronheid, G Lodder
    TETRAHEDRON LETTERS 41(26) 5125-5129 2000年6月  
    Alkenyl(phenyl)iodonium tetrafluoroborates dissolved in chloroform, or in the solid state, decompose thermally at 60 degrees C to yield fluoroalkenes and iodobenzene as major products via an S(N)1- or S(N)2-type reaction within the ion pair of the substrates. (C) 2000 Elsevier Science Ltd. All rights reserved.
  • D Laine, M Fujita, SV Ley
    JOURNAL OF THE CHEMICAL SOCIETY-PERKIN TRANSACTIONS 1 1999(12) 1639-1645 1999年6月  
    C-2 symmetric chiral 2,2'-bis(triisopropylsilyloxymethyl)bi(dihydropyran)s (S,S)-1 and (R,R)-1 were prepared from the corresponding glycidols and selectively reacted with 1,2-diols to give dispiroketals. The products of these reactions could be deprotected following treatment with fluoride, oxidation and reductive cleavage with samarium(II) iodide.
  • M Fujita, D Laine, SV Ley
    JOURNAL OF THE CHEMICAL SOCIETY-PERKIN TRANSACTIONS 1 1999(12) 1647-1656 1999年6月  
    Glycolic acid can be converted to optically active 1,2,3,4-tetraols using a dispiroketal unit as a protecting group and chiral auxiliary. Aldol reactions of dispiroketal protected glycolate with aldehydes afford one diastereoisomer preferentially with two newly formed stereogenic centres. To extend the polyol chain, the carbonyl group of the aldol product is converted to a vinyl ether by the Tebbe reagent after protection of the free alcohol. A subsequent hydroboration-oxidation protocol affords the dispiroketal protected tetraol. The final deprotection of the tetraol occurs selectively without epimerisation or migration of the silyloxy protecting groups.
  • D Laine, M Fujita, SV Ley
    JOURNAL OF THE CHEMICAL SOCIETY-PERKIN TRANSACTIONS 1 1999(12) 1639-1645 1999年6月  
    C-2 symmetric chiral 2,2'-bis(triisopropylsilyloxymethyl)bi(dihydropyran)s (S,S)-1 and (R,R)-1 were prepared from the corresponding glycidols and selectively reacted with 1,2-diols to give dispiroketals. The products of these reactions could be deprotected following treatment with fluoride, oxidation and reductive cleavage with samarium(II) iodide.
  • M Fujita, D Laine, SV Ley
    JOURNAL OF THE CHEMICAL SOCIETY-PERKIN TRANSACTIONS 1 1999(12) 1647-1656 1999年6月  
    Glycolic acid can be converted to optically active 1,2,3,4-tetraols using a dispiroketal unit as a protecting group and chiral auxiliary. Aldol reactions of dispiroketal protected glycolate with aldehydes afford one diastereoisomer preferentially with two newly formed stereogenic centres. To extend the polyol chain, the carbonyl group of the aldol product is converted to a vinyl ether by the Tebbe reagent after protection of the free alcohol. A subsequent hydroboration-oxidation protocol affords the dispiroketal protected tetraol. The final deprotection of the tetraol occurs selectively without epimerisation or migration of the silyloxy protecting groups.
  • T Sugimura, S Nagano, H Kohno, M Fujita, A Tai
    CHEMISTRY LETTERS (2) 179-180 1999年2月  
    The Diels-Alder reaction of tetracyanoethylene (TCNE) with a l-trienol unit in the tropilidenes at the 1,4-position was a quick and reversible process, whereas the 3,6-addition only proceeded in polar solvent and was irreversible.
  • M Fujita, M Ohshiba, Y Yamasaki, T Sugimura, A Tai
    CHEMISTRY LETTERS (2) 139-140 1999年2月  
    The presence of a zwitterionic intermediate during the oxygenation of electron-rich naphthalenes with singlet oxygen is clarified by the clean and efficient formation of hydroperoxide via nucleophilic addition of the intramolecular alcohol to the intermediate and by its stereochemical analysis using an optically active linker of the alcohol.
  • T Sugimura, S Nagano, H Kohno, M Fujita, A Tai
    CHEMISTRY LETTERS (2) 179-180 1999年2月  
    The Diels-Alder reaction of tetracyanoethylene (TCNE) with a l-trienol unit in the tropilidenes at the 1,4-position was a quick and reversible process, whereas the 3,6-addition only proceeded in polar solvent and was irreversible.
  • M Fujita, M Ohshiba, Y Yamasaki, T Sugimura, A Tai
    CHEMISTRY LETTERS (2) 139-140 1999年2月  
    The presence of a zwitterionic intermediate during the oxygenation of electron-rich naphthalenes with singlet oxygen is clarified by the clean and efficient formation of hydroperoxide via nucleophilic addition of the intramolecular alcohol to the intermediate and by its stereochemical analysis using an optically active linker of the alcohol.
  • M Fujita, M Ohshiba, S Shioyama, T Sugimura, A Tai
    CHEMICAL COMMUNICATIONS (20) 2243-2244 1998年10月  
    A novel stereochemical approach is employed in anodic oxidation of naphthalene derivatives to discriminate the intramolecular radical addition vs, intermolecular radical addition paths; the contribution of the latter is revealed to be important,judging from the stereodifferentiating addition of MeOH at C-4 during the anodic oxidation of (1'S,3'R)-1-(3'-hydroxy-1'-methylbutoxy)-4-methylnaphthalene.
  • M Fujita, M Ohshiba, S Shioyama, T Sugimura, A Tai
    CHEMICAL COMMUNICATIONS (20) 2243-2244 1998年10月  
    A novel stereochemical approach is employed in anodic oxidation of naphthalene derivatives to discriminate the intramolecular radical addition vs, intermolecular radical addition paths; the contribution of the latter is revealed to be important,judging from the stereodifferentiating addition of MeOH at C-4 during the anodic oxidation of (1'S,3'R)-1-(3'-hydroxy-1'-methylbutoxy)-4-methylnaphthalene.
  • A Tai, Y Higashiura, M Kakizaki, T Naito, K Tanaka, M Fujita, T Sugimura, H Hara, N Hayashi
    BIOSCIENCE BIOTECHNOLOGY AND BIOCHEMISTRY 62(3) 607-608 1998年3月  
    (1S, 2R, 6RS)-1,2,6-Trimethyldecyl propionate, a lower homolog of the sex pheromone of known sawflies, strongly attracted Diprion nipponica, a popular species in Japan.
  • M Fujita, Y Takarada, T Sugimura, A Tai
    CHEMICAL COMMUNICATIONS (17) 1631-1632 1997年9月  
    Forward and reverse hydride transfer with rigorous diastereodifferentiation is attained using intramolecular Meerwein-Ponndorf-Verley reduction and Oppenauer oxidation of a 2-acetylphenyl ether containing an optically active 3-hydroxy-1-methylbutyl group.
  • M Fujita, Y Takarada, T Sugimura, A Tai
    CHEMICAL COMMUNICATIONS (17) 1631-1632 1997年9月  
    Forward and reverse hydride transfer with rigorous diastereodifferentiation is attained using intramolecular Meerwein-Ponndorf-Verley reduction and Oppenauer oxidation of a 2-acetylphenyl ether containing an optically active 3-hydroxy-1-methylbutyl group.
  • Morifumi Fujita, Akito Ishida, Setsuo Takamuku, Shunichi Fukuzumi
    Journal of the American Chemical Society 118(36) 8566-8574 1996年9月11日  
    Addition of alkylbenzenes with 10-methylacridinium ion (AcrH+) occurs efficiently under visible light irradiation in deaerated acetonitrile containing H2O to yield 9-alkyl-10-methyl-9,10-dihydroacridine selectively. On the other hand, the photochemical reaction of AcrH+ with alkylbenzenes in the presence of perchloric acid in deaerated acetonitrile yields 10-methyl-9,10-dihydroacridine, accompanied by the oxygenation of alkylbenzenes to the corresponding benzyl alcohols. The photooxygenation of alkylbenzenes occurs also in the presence of oxygen, when AcrH+ acts as an efficient photocatalyst. The studies on the quantum yields and fluorescence quenching of AcrH+ by alkylbenzenes as well as the laser flash photolysis have revealed that the photochemical reactions of AcrH+ with alkylbenzenes in both the absence and presence of oxygen proceed via photoinduced electron transfer from alkylbenzenes to the singlet excited state of AcrH+ to produce alkylbenzene radical cations and 10-methylacridinyl radical (AcrH·). The competition between the deprotonation of alkylbenzene radical cations and the back electron transfer from AcrH· to the radical cations determines the limiting quantum yields. In the absence of oxygen, the coupling of the deprotonated radicals with AcrH· yields the adducts. The photoinduced hydride reduction of AcrH+ in the presence of perchloric acid proceeds via the protonation of acridinyl radical produced by the photoinduced electron transfer from alkylbenzenes. In the presence of oxygen, however, the deprotonated radicals are trapped efficiently by oxygen to give the corresponding peroxyl radicals which are reduced by the back electron transfer from AcrH· to regenerate AcrH+, followed by the protonation to yield the corresponding hydroperoxide. The ratios of the deprotonation reactivity from different alkyl groups of alkylbenzene radical cations were determined from both the intra- and intermolecular competitions of the deprotonation from two alkyl groups of alkylbenzene radical cations. The reactivity of the deprotonation from alkylbenzene radical cations increases generally in the order methyl &lt ethyl &lt isopropyl. The strong stereoelectronic effects on the deprotonation from isopropyl group of alkylbenzene radical cations appear in the case of the o-methyl isomer.
  • J Otera, Y Fujita, N Sakuta, M Fujita, S Fukuzumi
    JOURNAL OF ORGANIC CHEMISTRY 61(9) 2951-2962 1996年5月  
    Mechanism of Mukaiyama-Michael reaction of ketene silyl acetal has been discussed. The competition reaction employing various types of ketene silyl acetals reveals that those bearing more substituents at the beta-position react preferentially over less substituted ones. However, when ketene silyl acetals involve bulky siloxy and/or alkoxy group(s), less substituted compounds react preferentially. The Lewis acids play an important role in these reactions. Enhanced preference for the more sterically demanding Michael adducts is obtained with Bu(2)Sn(OTf)(2), SnCl4, and Et(3)-SiClO4 in the former reaction while TiCl4 gives the highest selectivity for the less sterically demanding products in the latter case. These results are interpreted in terms of alternative reaction mechanisms. The reaction of less bulky ketene silyl acetals are initiated by electron transfer from these compounds to a Lewis acid. On the other hand, bulkier ketene silyl acetals undergo a ubiquitous nucleophilic reaction. Such a mechanistic change is discussed based on a variety of experimental results as well as the semiempirical PM3 MO calculations.
  • J Otera, Y Fujita, N Sakuta, M Fujita, S Fukuzumi
    JOURNAL OF ORGANIC CHEMISTRY 61(9) 2951-2962 1996年5月  
    Mechanism of Mukaiyama-Michael reaction of ketene silyl acetal has been discussed. The competition reaction employing various types of ketene silyl acetals reveals that those bearing more substituents at the beta-position react preferentially over less substituted ones. However, when ketene silyl acetals involve bulky siloxy and/or alkoxy group(s), less substituted compounds react preferentially. The Lewis acids play an important role in these reactions. Enhanced preference for the more sterically demanding Michael adducts is obtained with Bu(2)Sn(OTf)(2), SnCl4, and Et(3)-SiClO4 in the former reaction while TiCl4 gives the highest selectivity for the less sterically demanding products in the latter case. These results are interpreted in terms of alternative reaction mechanisms. The reaction of less bulky ketene silyl acetals are initiated by electron transfer from these compounds to a Lewis acid. On the other hand, bulkier ketene silyl acetals undergo a ubiquitous nucleophilic reaction. Such a mechanistic change is discussed based on a variety of experimental results as well as the semiempirical PM3 MO calculations.
  • M Fujita, S Fukuzumi, GE Matsubayashi, J Otera
    BULLETIN OF THE CHEMICAL SOCIETY OF JAPAN 69(4) 1107-1116 1996年4月  
    beta,beta-Dimethyl-substituted ketene silyl acetal (1a) reduces p-chloranil and other activated quinones with electron-withdrawing substituents to produce the carbon-oxygen adduct, the hydrolysis of which yields the corresponding hydroquinone ether. The structure of the hydroquinone ether has been determined by the X-ray crystal analysis. The reactions are significantly slowed down in benzene where the charge-transfer spectra of electron donor-acceptor complexes formed between la and the activated quinones are observed. The comparison of the observed rate constant with that predicted for the electron transfer process from la to p-chloranil indicates that the addition of 1a to p-chloranil proceeds via the electron transfer from 1a to p-chloranil. Although no reaction takes place between 1a and p-benzoquinone, the electron affinity of which is significantly smaller than that of p-chloranil, the reduction of p-benzoquinone by 1a occurs efficiently in the presence of magnesium ion. The kinetic expression of the Mg2+ catalysis changes from the first-order to second-order in [Mg2+] under the conditions that Mg2+ forms the 2 : 1 complexes with the corresponding radical anions. The catalytic effects of Mg2+ are approximately the same as those observed for the electron transfer reduction of these oxidants, demonstrating the important contribution of the Mg2+-catalyzed electron transfer process in the addition of 1a to p-benzoquinone. On the other hand, the reaction of a nonsubstituted ketene silyl acetal (1d) with p-fluoranil yields the carbon-carbon adduct rather than the carbon-oxygen adduct. The much larger rate constants of 1d than those of 1a despite the higher oxidation potential of 1d suggest that the 1,2-addition to p-fluoranil occurs via the nucleophilic attack of 1d, which is much less sterically hindered than 1a, to the positively charged carbonyl carbon of p-fluoranil rather than an alternative electron transfer pathway.
  • M Fujita, S Fukuzumi, GE Matsubayashi, J Otera
    BULLETIN OF THE CHEMICAL SOCIETY OF JAPAN 69(4) 1107-1116 1996年4月  
    beta,beta-Dimethyl-substituted ketene silyl acetal (1a) reduces p-chloranil and other activated quinones with electron-withdrawing substituents to produce the carbon-oxygen adduct, the hydrolysis of which yields the corresponding hydroquinone ether. The structure of the hydroquinone ether has been determined by the X-ray crystal analysis. The reactions are significantly slowed down in benzene where the charge-transfer spectra of electron donor-acceptor complexes formed between la and the activated quinones are observed. The comparison of the observed rate constant with that predicted for the electron transfer process from la to p-chloranil indicates that the addition of 1a to p-chloranil proceeds via the electron transfer from 1a to p-chloranil. Although no reaction takes place between 1a and p-benzoquinone, the electron affinity of which is significantly smaller than that of p-chloranil, the reduction of p-benzoquinone by 1a occurs efficiently in the presence of magnesium ion. The kinetic expression of the Mg2+ catalysis changes from the first-order to second-order in [Mg2+] under the conditions that Mg2+ forms the 2 : 1 complexes with the corresponding radical anions. The catalytic effects of Mg2+ are approximately the same as those observed for the electron transfer reduction of these oxidants, demonstrating the important contribution of the Mg2+-catalyzed electron transfer process in the addition of 1a to p-benzoquinone. On the other hand, the reaction of a nonsubstituted ketene silyl acetal (1d) with p-fluoranil yields the carbon-carbon adduct rather than the carbon-oxygen adduct. The much larger rate constants of 1d than those of 1a despite the higher oxidation potential of 1d suggest that the 1,2-addition to p-fluoranil occurs via the nucleophilic attack of 1d, which is much less sterically hindered than 1a, to the positively charged carbonyl carbon of p-fluoranil rather than an alternative electron transfer pathway.
  • M Fujita, A Shindo, A Ishida, T Majima, S Takamuku, S Fukuzumi
    BULLETIN OF THE CHEMICAL SOCIETY OF JAPAN 69(3) 743-749 1996年3月  
    Photooxygenation of 1,1-diarylethylene occurs efficiently using 10-methylacridinium ion as a photocatalyst to yield the 1,2-dioxane and/or the diaryl ketone depending on the substituents on the aryl groups. The reaction mechanism is revealed based on the dependence of the quantum yields on the concentrations of the alkene and oxygen, the fluorescence quenching of 10-methylacridinium ion by the alkene, and the direct detection of reactive intermediates by applying laser flash spectroscopy as well as pulse radiolysis. The photooxygenation proceeds via photoinduced electron transfer from the alkene to the singlet excited state of 10-methylacridinium ion. The alkene radical cation formed by the photoinduced electron transfer reacts with alkene to give the 1,4-dimer radical cation, which then reacts with oxygen to produce the oxygenated 1,6-radical cation. The subsequent one-electron reduction of the 1,6-radical cation results in formation of the 1,6-biradical which cyclizes to yield 1,2-dioxane derivative or fragmentates to yield diaryl ketone. When the 1,6-biradical is reduced by the alkene itself, the alkene radical cation is regenerated to repeat the radical chain process.
  • M Fujita, A Ishida, T Majima, S Takamuku
    JOURNAL OF PHYSICAL CHEMISTRY 100(13) 5382-5387 1996年3月  
    The selective eletron-transfer quenching of the radical anions of dicyanoanthracene, phenazine, and anthraquinones in the excited state by a quencher such as fumaronitrile or dicyanobenzene is investigated in N,N-dimethylformamide solution at room temperature using the pulse radiolysis-laser flash photolysis combined method. The radical anions generated by pulse radiolysis do not change upon irradiation with a laser flash at 532 nm. The radical anions in the excited state decay into the ground state within the laser flash (5 ns). Lifetimes of approximately 4 ns are estimated for three radical anions in the excited state assuming a diffusion-controlled rate constant for the electron-transfer quenching. The shorter lifetimes of 1.0-1.4 ns for methyl and chloro substituents on anthraquinone are discussed in terms of internal conversion from the excited to the ground state of the radical anions accelerated by rotation of the substituents. The energy gap between the excited and ground states of the radical anions is a significant factor for the rate of the internal conversion. The quencher radical anion-neutral molecule pair is suggested as an intermediate in the electron-transfer quenching of the radical anions during the excited state by the quencher and is discussed with respect to separation and back electron transfer in the pair.
  • M Fujita, A Shindo, A Ishida, T Majima, S Takamuku, S Fukuzumi
    BULLETIN OF THE CHEMICAL SOCIETY OF JAPAN 69(3) 743-749 1996年3月  
    Photooxygenation of 1,1-diarylethylene occurs efficiently using 10-methylacridinium ion as a photocatalyst to yield the 1,2-dioxane and/or the diaryl ketone depending on the substituents on the aryl groups. The reaction mechanism is revealed based on the dependence of the quantum yields on the concentrations of the alkene and oxygen, the fluorescence quenching of 10-methylacridinium ion by the alkene, and the direct detection of reactive intermediates by applying laser flash spectroscopy as well as pulse radiolysis. The photooxygenation proceeds via photoinduced electron transfer from the alkene to the singlet excited state of 10-methylacridinium ion. The alkene radical cation formed by the photoinduced electron transfer reacts with alkene to give the 1,4-dimer radical cation, which then reacts with oxygen to produce the oxygenated 1,6-radical cation. The subsequent one-electron reduction of the 1,6-radical cation results in formation of the 1,6-biradical which cyclizes to yield 1,2-dioxane derivative or fragmentates to yield diaryl ketone. When the 1,6-biradical is reduced by the alkene itself, the alkene radical cation is regenerated to repeat the radical chain process.
  • M Fujita, A Ishida, T Majima, S Takamuku
    JOURNAL OF PHYSICAL CHEMISTRY 100(13) 5382-5387 1996年3月  
    The selective eletron-transfer quenching of the radical anions of dicyanoanthracene, phenazine, and anthraquinones in the excited state by a quencher such as fumaronitrile or dicyanobenzene is investigated in N,N-dimethylformamide solution at room temperature using the pulse radiolysis-laser flash photolysis combined method. The radical anions generated by pulse radiolysis do not change upon irradiation with a laser flash at 532 nm. The radical anions in the excited state decay into the ground state within the laser flash (5 ns). Lifetimes of approximately 4 ns are estimated for three radical anions in the excited state assuming a diffusion-controlled rate constant for the electron-transfer quenching. The shorter lifetimes of 1.0-1.4 ns for methyl and chloro substituents on anthraquinone are discussed in terms of internal conversion from the excited to the ground state of the radical anions accelerated by rotation of the substituents. The energy gap between the excited and ground states of the radical anions is a significant factor for the rate of the internal conversion. The quencher radical anion-neutral molecule pair is suggested as an intermediate in the electron-transfer quenching of the radical anions during the excited state by the quencher and is discussed with respect to separation and back electron transfer in the pair.
  • FUKUZUMI S, OKAMOTO T, FUJITA M, OTERA J
    Chemical Communication (3) 393-394 1996年2月  
  • FUJITA M, ISHIDA A, TAKAMUKU S, FUKUZUMI S
    Journal of the American Chemical Society 118(36) 8566-8574 1996年  
  • M FUJITA, A ISHIDA, T MAJIMA, S FUKUZUMI, S TAKAMUKU
    CHEMISTRY LETTERS (2) 111-112 1995年2月  
    Photoinduced hydride reduction of 10-methylacridinium ion (AcrH(+)) by alkylbenzene occurs to yield 10-methyl-9,10-dihydroacridine in the presence of perchloric acid, while photoaddition occurs to yield 9-alkyl-10-methyl-9,10-dihydroacridine in the absence of perchloric acid. The hydride reduction of AcrH(+) in the presence of perchloric acid proceeds via protonation of acridinyl radical produced by photoinduced electron transfer from alkylbenzene.
  • Morifumi Fujita, Akito Ishida, Tetsuro Majima, Shunichi Fukuzumi, Setsuo Takamuku
    Chemistry Letters (2) 111-112 1995年  
  • S FUKUZUMI, M FUJITA, J MARUTA, M CHANON
    JOURNAL OF THE CHEMICAL SOCIETY-PERKIN TRANSACTIONS 2 1994(7) 1597-1602 1994年7月  
    Various NAD+ analogues have been reduced regioselectively by the tetramethylammonium salt of 2-nitropropane anion in acetonitrile at 298 K to yield the corresponding 4-alkylated NADH analogues. The one-electron oxidation potential of the tetramethylammonium salt of 2-nitropropane anion has been determined as 0.10 V (vs. SCE) by using second harmonic ac voltammetry as well as by analysing the cyclic voltammograms at various sweep rates. The rate constants for the reduction of NAD+ analogues by 2-nitropropane anion (&gt;1 x 10(6) dm3 mol-1 s-1) are much larger than those estimated for outer-sphere electron transfer from 2-nitropropane anion to NAD+ analogues based on the one-electron oxidation potential of 2-nitropropane anion and the one-electron reduction potentials of NAD+ analogues. The origin of the regioselectivity is discussed in terms of the HSAB (hard and soft acids and bases) principle.
  • S FUKUZUMI, M FUJITA, J MARUTA, M CHANON
    JOURNAL OF THE CHEMICAL SOCIETY-PERKIN TRANSACTIONS 2 1994(7) 1597-1602 1994年7月  
    Various NAD+ analogues have been reduced regioselectively by the tetramethylammonium salt of 2-nitropropane anion in acetonitrile at 298 K to yield the corresponding 4-alkylated NADH analogues. The one-electron oxidation potential of the tetramethylammonium salt of 2-nitropropane anion has been determined as 0.10 V (vs. SCE) by using second harmonic ac voltammetry as well as by analysing the cyclic voltammograms at various sweep rates. The rate constants for the reduction of NAD+ analogues by 2-nitropropane anion (&gt;1 x 10(6) dm3 mol-1 s-1) are much larger than those estimated for outer-sphere electron transfer from 2-nitropropane anion to NAD+ analogues based on the one-electron oxidation potential of 2-nitropropane anion and the one-electron reduction potentials of NAD+ analogues. The origin of the regioselectivity is discussed in terms of the HSAB (hard and soft acids and bases) principle.
  • Morifumi Fujita, Shunichi Fukuzumi
    Journal of Molecular Catalysis 90(3) L225-L229 1994年  
    The efficient oxygenation of alkylbenzenes is initiated by photoinduced electron transfer from alkylbenzenes to the singlet excited state of 10-methylacridinium ion under visible light irradiation, followed by the deprotonation of alkylbenzene radical cations and the facile addition of oxygen to the deprotonated radicals, leading to the final oxygenated products. © 1994.
  • M FUJITA, S FUKUZUMI, J OTERA
    JOURNAL OF MOLECULAR CATALYSIS 85(2) 143-148 1993年11月  
    An excellent linear relationship is obtained between the relative reactivities of the carbonyl compounds in the triethylsilyl perchlorate- and perchloric acid-catalyzed reduction by triethylsilane demonstrating that identical catalytic mechanisms are operative in both the Bronsted acid (perchloric acid) and the Lewis acid (triethylsilyl perchlorate) systems in solution.
  • M FUJITA, S FUKUZUMI
    JOURNAL OF THE CHEMICAL SOCIETY-CHEMICAL COMMUNICATIONS (19) 1528-1529 1993年10月  
    The isopropyl group has no significant stereoelectronic effect on inter- and intra-molecular competition in deprotonation from alkylbenzene radical cations in the photoaddition of alkylbenzene derivatives with 10-methylacridinium ion via photoinduced electron transfer from alkylbenzene derivatives to the singlet excited state of 10-methylacridinium ion, followed by the deprotonation of alkylbenzene radical cations.
  • S FUKUZUMI, M FUJITA, J OTERA
    JOURNAL OF ORGANIC CHEMISTRY 58(20) 5405-5410 1993年9月  
    Photoaddition of beta,beta-dimethyl-substituted ketene silyl acetals to 10-methylacridone occurs efficiently under irradiation of the visible light in benzene as well as acetonitrile to yield the siloxy adduct. The carbon-oxygen bond of the adduct is readily cleaved by an acid to yield the corresponding 9-alkylated 10-methylacridinium ion. The comparison of the observed rate constants determined from the dependence of the quantum yields on the concentrations of ketene silyl acetals as well as the fluorescence quenching by ketene silyl acetals with those predicted for the electron transfer processes indicates that the photoaddition proceeds via the photoinduced electron transfer from beta,beta-dimethyl-substituted ketene silyl acetals to the singlet and triplet excited states. No photoaddition of nonsubstituted ketene silyl acetal to 10-methylacridone occurs, since the electron donor ability is too weak to transfer an electron to the excited states of 10-methylacridone as expected from the higher oxidation potential compared with the beta,beta-dimethyl-substituted analogues.
  • S FUKUZUMI, M FUJITA, G MATSUBAYASHI, J OTERA
    CHEMISTRY LETTERS (8) 1451-1454 1993年8月  
    A beta,beta-dimethyl-substituted ketene silyl acetal reduces p-chloranil to produce the carbon-oxygen adduct, the hydrolysis of which yields the corresponding hydro-quinone ether (3). The structure of 3 has been determined by the X-ray crystal analysis. On the other hand, the reaction of a nonsubstituted ketene silyl acetal with p-fluoranil yields the carbon-carbon adduct selectively. The mechanistic difference between the C-O and C-C bond formation is elucidated based on the kinetic study.
  • S FUKUZUMI, M FUJITA, G MATSUBAYASHI, J OTERA
    CHEMISTRY LETTERS (8) 1451-1454 1993年8月  
    A beta,beta-dimethyl-substituted ketene silyl acetal reduces p-chloranil to produce the carbon-oxygen adduct, the hydrolysis of which yields the corresponding hydro-quinone ether (3). The structure of 3 has been determined by the X-ray crystal analysis. On the other hand, the reaction of a nonsubstituted ketene silyl acetal with p-fluoranil yields the carbon-carbon adduct selectively. The mechanistic difference between the C-O and C-C bond formation is elucidated based on the kinetic study.
  • S FUKUZUMI, M FUJITA, S NOURA, J OTERA
    CHEMISTRY LETTERS (6) 1025-1028 1993年6月  
    Various NAD+ analogues are reduced regioselectively by ketene silyl acetals to afford the corresponding 2-, 6-, or 4-alkylated NADH analogues. The regioselectivity of the 1,2- (or 1,6-) vs. 1,4- reduction is finely controlled by the beta-methyl-substitution of ketene silyl acetals, the position of methyl-substitution of NAD+ analogues, and the addition of Et4NCl or Bu4NF.
  • FUJITA Morifumi, FUKUZUMI Shunichi
    Chemistry Letters 1993(11) 1911-1914 1993年  
  • Shunichi Fukuzumi, Morifumi Fujita, Junzo Otera
    Journal of the Chemical Society, Chemical Communications (20) 1536-1537 1993年  
    Thermal reduction of 10-methylacridinium ion by allylic stannanes occurs via a polar mechanism to yield the dihydroacridines allylated at the γ-position exclusively, while the photoallylation of 10-methylacridinium ion proceeds via the photoinduced electron transfer from allylic silanes and stannanes to the singlet excited state of 10-methylacridinium ion to yield mainly the α-adducts.
  • Journal of the Chemical Society, Perkin Transaction 2 1993(10) 1915 1993年  
  • FUJITA Morifumi, FUKUZUMI Shunichi
    Chemistry letters 1993(11) 1911-1914 1993年  
  • Shunichi Fukuzumi, Morifumi Fujita, Souta Noura, Junzo Otera
    Chemistry Letters (6) 1025-1028 1993年  

講演・口頭発表等

 1

共同研究・競争的資金等の研究課題

 17