Chemical and Pharmaceutical Bulletin
Online ISSN : 1347-5223
Print ISSN : 0009-2363
ISSN-L : 0009-2363
Volume 30, Issue 9
Displaying 1-50 of 68 articles from this issue
  • YOSHIO SASAKI, TATSUYA TAKAGI, AKIHIRO IWATA, HIDEKO KAWAKI
    1982 Volume 30 Issue 9 Pages 3069-3073
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Biological response (BR) can be expressed by the equation BR=α (σso)2+bσso+cσi+dσx+e for quantitative structure-activity relationships (QSAR) analyses. The values of σi and σx originally take+or-sign depending on the electronic character of the substituent group and can be used to describe strong drug-site interactions. On the contrary, original chemical significance is lost in |σi| and |σx|, which can be applied to weak drug-site interaction, where dipole-dipole or dispersion interaction could be predominant. The coefficient c of the BR equation becomes the same in meta- and para-substituted benzene series. When σx is real, σx/mx/p gives a better result than σx/m, p, while on the other hand, when σx is absolute, |σx|/m, p gives a good result. As regards the entropy term a (σso)2+bσso, the ortho-substituted benzene series requires the+sign for a and the-for b, whereas for meta-and para-substituted series, this is reversed. When the relative weight of the entropy term decreases and the contribution of the enthalpy term becomes predominant, the quantum chemical index could often be used as an effective scale for BR.
    Download PDF (529K)
  • SABURO SHIMABAYASHI, HITOMI FUKUDA, TOSHITAKA AOYAMA, MASAYUKI NAKAGAK ...
    1982 Volume 30 Issue 9 Pages 3074-3081
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The amount of phosphate ion, xp, adsorbed by synthetic hydroxyapatite (HAP) was measured under various conditions in water. A long time was required to reach adsorption equilibrium, and the adsorption process was largely irreversible with respect to concentration. The adsorption isotherm was not of Langmuir type, but was dominated by the ionic product of adsorbate, depending on the species of counter ion. The value of xp decreased with pH due to mutual electrostatic repulsion between adsorbed phosphate ions, which are deprotonated, and due to competitive adsorption between phosphate ion and OH-ion. The adsorption process was also found to be endothermic, because the adsorption increased with temperature. The adsorbate ion and adsorbent surface may be dehydrated when they are bound together. From all these results, it was concluded that the adsorbate phosphate ion is oriented and coordinated to calcium ion on the surface of HAP, and this is the reason why the adsorption of phosphate ion is irreversible.
    Download PDF (964K)
  • TOSHIO KOBAYASHI, MINORU MAEDA, HIROSHI KOMATSU, MASAHARU KOJIMA
    1982 Volume 30 Issue 9 Pages 3082-3087
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The synthesis of 7-fluoro-B-homo-19-norcholest-5 (10)-en-3β-ol acetate (1) was examined by utilizing diethyl (2-chloro-1, 1, 2-trifluoroethyl) amine (FAR), diethylaminosulfur trifluoride (DAST), and/or hexafluoropropene-diethylamine (FPA) as fluorinating agents for cholest-5-en-3β, 19-diol 3-acetate (2), 6β-hydroxymethyl-19-norcholest-5 (10)-en-3β-ol 3-acetate (9), 7β-hydroxy-B-homo-19-norcholest-5 (10)-en-3β-ol 3-acetate (12), and 3β-acetoxy-6β-hydroxy-5β, 19-cyclocholestane (13). The treatment of 2 and 9 with these fluorinating agents gave the 7-fluoro-B-homo-5 (10)-ene (1) in poor yield together with the cyclopropane products, 3β-acetoxy-5β, 6β-methanocholest-1 (10)-ene (3) and 3β-acetoxy-5β, 6β-methanocholest-9-ene (4). When 12 was allowed to react with FAR at -78°C, the required 1 was produced in 43% yield. The most satisfactory result, however, was obtained by the reaction of 13 with FAR at -78°C, which afforded the 7-fluoro-B-homo-5 (10)-ene (1) in 64% yield.
    Download PDF (768K)
  • TSUYOSHI TSUJIYAMA, TSUTOMU OHSAKI, KAZUTAKA MATSUBARA, AKIMITSU SANO, ...
    1982 Volume 30 Issue 9 Pages 3088-3091
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    1-(5-Oxohexyl) theobromine (I) in alkaline solution was hydrolyzed by heating to give 4-(methylamino)-1-methyl-5-(5-oxohexyl) aminocarbonylimidazole (II) and N-[4-(5-carboxy-1-methylimidazolyl)]-N-methyl-N'-(5-oxohexyl) urea (III). When acidified with HCl, II and III were cyclized to 4, 4a, 5, 6, 7, 8-hexahydro-10-oxo-1, 4, 4a-trimethyl-1H-pyrido [1, 2-a]-purine (IV) and 4, 5, 7, 8, 9, 10-hexahydro-5, 12-dioxo-1, 4, 10a-trimethyl-1H-pyrido [2', 1' : 2, 3]-imidazo [4, 5-f] oxadiazocine hydrochloride (V), respectively.
    Download PDF (602K)
  • YUJI OIKAWA, HITOSHI HIRASAWA, OSAMU YONEMITSU
    1982 Volume 30 Issue 9 Pages 3092-3096
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    When an acetonitrile solution of Meldrum's acid (1), indole (5), and an aliphatic or aromatic aldehyde (8) in the presence of a small amount of proline (except in the case of acetaldehyde) was allowed to stand at 30°C, a simultaneous condensation of three different carbon components occurred readily to give a 5-(1H-indol-3-ylalkyl)-2, 2-dimethyl-1, 3-dioxane-4, 6-dione (11) in high yield. The reaction proceeded regardless of the nature of the aldehyde. An ethanolysis of 11 with loss of acetone and carbon dioxide took place smoothly in boiling ethanol-pyridine (1 : 10) containing a small amount of copper powder to give an ethyl β-alkylindolepropionate (2). These two reactions, the condensation and the ethanolysis, were combined in a one-pot procedure, which may provide an efficient and convenient synthetic method for various ethyl indolepropionates (2).
    Download PDF (658K)
  • KUNIHIKO MOHRI, YUJI OIKAWA, KENICHI HIRAO, OSAMU YONEMITSU
    1982 Volume 30 Issue 9 Pages 3097-3105
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Phenylhydroxylamine (13) oxalate was quite easily acylacetylated by heating with an equimolar amount of acyl Meldrum's acid (4) in acetonitrile to give an N-acylacetylphenylhydroxylamine (14) in high yield. When 14 was treated with another equimolar amount of the same 4 in refluxing toluene, a series of reactions, O-acylacetylation, 1-aza-1'-oxa [3, 3] sigmatropic rearrangement, decarboxylation, dehydrative cyclization, and deacylation, occurred consecutively to give a 2-substituted indole (16) in fair yield, though sometimes accompanied by the formation of a 5-substituted 4-isoxazolin-3-one (17). N-Benzoyl, N-acetyl, and N-benzyloxycarbonyl derivatives of phenylhydroxylamine (18) were treated with phenylacetyl Meldrum's acid (4i) in refluxing benzene containing copper powder to give readily rearranged ortho. alkylation products (19), which were converted to the corresponding N-acyl-2-benzylindoles (20) by treatment with hydrochloric acid in boiling ethanol or with anhydrous p-toluenesulfonic acid in benzene at room temperature.
    Download PDF (923K)
  • ISAO MINAMI, YOSHIO KOZAI, HIROAKI NOMURA, TAZUKO TASHIRO
    1982 Volume 30 Issue 9 Pages 3106-3120
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Antitumor agents of a new structural type are reported. A number of cyanines with mono-, di- and tricyclic nuclei, merocyanines and oxonols have been i. p. screened for antitumor activity against P388 leukemia and B16 melanoma. Among these compounds, monomethin-, trimethin- and pentamethincyanines having naphthothiazole, naphthoxazole and benzindole nuclei resulted in a significant prolongation of the survival time of tumor-bearing mice. Replacement of the conjugated chain system between the terminal two nuclei with a saturated aliphatic chain produced a marked decrease in the antitumor activity. Structure-activity relationships are discussed.
    Download PDF (1532K)
  • HIDEKI MIKI
    1982 Volume 30 Issue 9 Pages 3121-3124
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    2, 4 (1H, 3H)-Quinazolinedione readily undergoes a von Braun-type reaction with N-alkyl cyclic amines in phosphoryl chloride through the formation of 4-oxo-3, 4-dihydroquinazolinyl dichlorophosphate and N-alkyl-N-(4-oxo-3, 4-dihydro-2-quinazolinyl) cyclic ammonium chloride in sequence ; the latter decomposes to give a new tertiary amine, 2-[N-alkyl-N-(ω-chloroalkyl) amino]-4-chloroquinazolines. A possible reaction mechanism is discussed.
    Download PDF (511K)
  • MASAYUKI TANNO, SHOKO SUEYOSHI, SHOZO KAMIYA
    1982 Volume 30 Issue 9 Pages 3125-3132
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    1-Aryl-3-cyanotriazene potassium salts (IIa-j) were synthesized by treating the corresponding arylazides (Ia-j) with potassium cyanide. When IIa-j were treated with dimethyl sulfate in the presence of dicyclohexyl-18-crown ether-6 (18-crown-6), methylation occurred at N1 and N3 of the triazeno group. The resulting two isomers, 1-aryl-3-cyano-1-methyltriazenes (IIIa-j) and 1-aryl-3-cyano-3-methyltriazenes (IVa-f) were all separable by means of chromatography on silica gel. However, several of the 3-isomers (IVg-j) were not isolated. On the other hand, the direct alkylation of arylcyanotriazene potassium salts (IIa, c, e) with dimethyl sulfate in methanol gave only the 1-methyl derivatives, 1-aryl-3-cyano-1-methyltriazenes (IIIa, c, e). The alkylated positions in these products were determined by chemical and spectral studies. A clear correlation between the Hammett constants and the 13C-nuclear magnetic resonance (13C-NMR) chemical shifts of the nitrile carbons of (1-aryl-3-cyano) methyltriazenes (IIIa-j, IVa-f) was obtained.
    Download PDF (871K)
  • SACHIYO TAHARA, MARI SHIGETSUNA, HIROTAKA OTOMASU
    1982 Volume 30 Issue 9 Pages 3133-3138
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Nitrosation of 2-methylisoquinoline-1, 3 (2H, 4H)-dione (I) with NaNO2 in AcOH afforded the 4-nitroso compound (II) in almost quantitative yield. Catalytic reduction of II produced 4-imino-2-methylisoquinoline-1, 3 (2H, 4H)-dione (III) and a small amount of 4, 4'-bi-2-methylisoquinoline-1, 3 (2H, 4H)-dione (IV). Acidic hydrolysis of III gave 2-methylisoquinoline-1, 3, 4 (2H)-trione (V). The hydrogenated mixture of II was allowed to react with methyl- and phenylisocyanate under an inert atmosphere to give 4-ureido compounds (VIIa, b) in yields of 79% (both). Dehydrogenation of VIIa, b with Pd-C afforded the 4-carbamoylimino compounds (VIIIa, b) in 60 and 87% yields, respectively. When their methanol solutions were refluxed with base catalyst, the compounds (VIIIa, b) were transformed into spiro [imidazolidin-4, 1'-isoindoline]-2, 3', 5-triones (XIIa, b) in yields of 80 and 86%, respectively.
    Download PDF (709K)
  • MASAMI SHIMAZAKI, JUNZO HASEGAWA, KAZUNORI KAN, KENJI NOMURA, YASUSHI ...
    1982 Volume 30 Issue 9 Pages 3139-3146
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A synthesis of N-(3-mercapto-2-D-methylpropanoyl)-L-proline (1, Captopril) is described in which the key intermediate, optically active 3-chloro-2-D-methylpropanoyl chloride (3b), was prepared by treating microbiologically derived optically active 3-hydroxy-2-D-methylpropanoic acid (2) with thionyl chloride. The compound 3b was coupled with L-proline to afford the chloride (4) which was directly converted into Captopril by reaction with hydrosulfide or trithiocarbonate ion in hot water with retention of the stereochemistry. The preparation of another useful intermediate, 3-acetylthio-2-D-methylpropanoic acid (9a), is also described.
    Download PDF (940K)
  • SHOJI TAKUMA, YASUMASA HAMADA, TAKAYUKI SHIOIRI
    1982 Volume 30 Issue 9 Pages 3147-3153
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The extent of racemization and the coupling efficiency during the coupling of benzyloxycarbonyl-L-phenylalanyl-L-valine with L-proline tert-butyl ester were conveniently and rapidly determined by the use of high performance liquid chromatography with high accuracy. An extensive examination of various known coupling methods revealed that the coupling methods using diphenyl phosphorazidate (DPPA) and, especially, diethyl phosphorocyanidate (DEPC) afforded the best results. Reported results using DPPA and DEPC as peptide coupling reagents are also summarized (Table I).
    Download PDF (865K)
  • TOSHIAKI MIURA, HIDETOSHI TAKAGI, KENICHI ADACHI, MICHIYA KIMURA
    1982 Volume 30 Issue 9 Pages 3154-3159
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The behavior of testosterone (I) and 17β-hydroxyandrosta-4, 6-dien-3-one (VII) in sulfuric acid was investigated by absorption and 1H- and 13C-nuclear magnetic resonance (NMR) spectroscopic studies. I and VII were demonstrated to give initially the diprotonated species, II and VIII, respectively and then the dication (x-300, III), which is the intermediate in the color and fluorescence reaction of I with strong acids. The conjugate base of x-300, 17-methyl-18-norandrosta-4, 8 (14), 13 (17)-trien-3-one (IV), was isolated from the reaction mixture of VII with 97% sulfuric acid. Thus, the chemical structure of x-300 was fully elucidated and the reaction mechanism previously proposed was confirmed.
    Download PDF (628K)
  • HIROSHI TAKAHASHI, YUJI SUZUKI, HIDEKAZU INAGAKI
    1982 Volume 30 Issue 9 Pages 3160-3166
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Optically pure 1-aryl-N-(2'-hydroxy-1'-isopropylethyl)-2-phenylethylamines (3a-c) were synthesized by the reactions of N-(2-hydroxy-1-isopropylethyl) arylmethylideneamines (2a-c) with benzylmagnesium chloride. The diastereomers (5a-c) were prepared from N-(2-hydroxy-1-isopropylethyl) phenylethylideneamine (4) and aryllithium compounds. The optical purity of 3a was elucidated and the absolute configuration was determined by synthesis via an alternative route. The characteristic methyl signals in the nuclear magnetic resonance (NMR) spectra suggested the configurations of the chiral amines (3a-c and 5a-c).
    Download PDF (847K)
  • YOSHIYASU TERAO, NOBUYUKI IMAI, KAZUO ACHIWA, MINORU SEKIYA
    1982 Volume 30 Issue 9 Pages 3167-3171
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Efficient synthesis of pyrrolizidines and indolizidines has been achieved by reacting tetracyclic hexahydro-1, 3, 5-triazines with olefinic and acetylenic compounds in the presence of trimethylsilylmethyl trifluoromethanesulfonate and cesium fluoride. This reaction was applied to the synthesis of several pyrrolizidine alkaloids, (±)-trachelanthamidine, (±)-supinidine, and (±)-isoretronecanol.
    Download PDF (515K)
  • TOSHIKI MUTO, JUN TANAKA, TOSHIAKI MIURA, MICHIYA KIMURA
    1982 Volume 30 Issue 9 Pages 3172-3177
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Oxy-functionalization of cholesteryl acetate (1a) occurred, giving 3β-acetoxy-5, 6-epoxycholestane (2a), 3β-acetoxycholest-5-en-7-one (3a), 3β-acetoxycholest-5-en-7-ol (4a), and an unidentified product (5), when 1a was oxidized by a system consisting of Fe (acac)3 and the hydroperoxide of an unsaturated long-chain fatty acid (LH) such as oleic, linolic, or linolenic acid (Table I). The epoxidation of stilbene by the same system was found to be non-stereospecific. These results and the fact that the reaction in this system was inhibited by a radical scavenger (BHT) were fairly compatible with those obtained with the Fe (acac)3-tBuOOH system, which is assumed to generate oxy and peroxy radicals. Autoxidation of 1a and cholesterol (1b) in the presence of Fe (acac)3 and LH proceeded after a time-lag of several hours and was also inhibited by BHT. The marked stereoselectivity of β-epoxidation (β/α+β=0.72) and the extent (about 30-40%) of allylic oxidation in the autoxidation were in fair agreement with those found for 1a in the Fe(acac)3-LOOH system (Table II). Autoxidation of stilbene in the presence of Fe (acac)3 and LH also led to non-stereospecific epoxidation. Thus, the autoxidation of cholesterols (1a and 1b) in the Fe (acac)3-LH system was assumed to be a radical reaction in which LOO·and LO·are the attacking species.
    Download PDF (749K)
  • TOMIHIKO OHSAWA, TAKASHI TAKAGAKI, FUMIAKI IKEHARA, YASUO TAKAHASHI, T ...
    1982 Volume 30 Issue 9 Pages 3178-3186
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A system of potassium metal-crown ether-diglyme has been proved to be highly effective for reductive removal of the sulfonyl group from O-sulfonates or sulfonamides. The reaction proceeds even with mesylamides of dialkyl amines, which resist reductive cleavage by naphthalene radical anion. It has also been found that a simple combination of potassium metal or sodium-potassium alloy with isopropanol in diglyme or toluene is effective for the cleavage of dialkylamine sulfonamides which are relatively susceptible to reduction. In particular, the use of sodium-potassium alloy gave satisfactory results.
    Download PDF (1010K)
  • FUMIO YONEDA, MIYUKI MOTOKURA, MUTSUKO KAMISHIMOTO, TOMOHISA NAGAMATSU ...
    1982 Volume 30 Issue 9 Pages 3187-3196
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Treatment of 1, 3, 6-trimethyl-5-nitrouracil with aryl aldehydes in the presence of piperidine results in condensation to the 5-nitro-6-styryluracil derivatives, followed by intramolecular cyclization including piperidine-catalyzed oxidation-reduction to give rise to the corresponding 3-hydroxy-4, 6-dimethylpyrrolo [3, 2-d] pyrimidine-5, 7 (4H, 6H)-dione (9-hydroxy-9-deazatheophylline) derivatives in a single step. Some properties of these compounds and an X-ray crystal structure determination of 8-(p-chlorophenyl)-9-methoxy-7-methyl-9-deazatheophylline are described.
    Download PDF (900K)
  • KOICHIRO YAMADA, MIKIO TAKEDA, NOBUO ITOH, HISAO OHTSUKA, AKIRA TSUNAS ...
    1982 Volume 30 Issue 9 Pages 3197-3201
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A convenient synthetic route to (±)-5, 7-dihydroxy-1-(3, 4, 5-trimethoxybenzyl)-1, 2, 3, 4-tetrahydroisoquinoline (racemic TA-073), an orally active bronchodilator, through the cyclization of the amine 4 or its N-acyl derivative (10b, c) is presented. Catalytic reduction of 4 over 10% Pd-C in EtOH-H2O in the presence of oxalic acid (6 eq) afforded racemic TA-073 in 65% yield. The acid-catalyzed reaction of 10b, c was also investigated. Cyclization of 10b, c with 1 N HCl (1 eq) in THF gave the corresponding N-acyl enamine 12b, c in quantitative yield. Further treatment of 12b with conc. HCl in THF afforded the pavine derivative 14 in 80% yield. Cyclization of 10b with neat HCOOH, however, resulted in the formation of the isopavine derivative 11b in 80% yield. This result parallels the reported observation by McDonald et al.
    Download PDF (661K)
  • YUKINOBU IKEYA, NORIYUKI OOKAWA, HEIHACHIRO TAGUCHI, ITIRO YOSIOKA
    1982 Volume 30 Issue 9 Pages 3202-3206
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Three new dibenzocyclooctadiene lignans, angeloylgomisin O (1), and angeloyl-(2) and benzoylisogomisin O (3) were isolated from the fruits of Schizandra chinensis BAILL. (Schizandraceae). Their absolute structures were elucidated by means of chemical and spectral studies.
    Download PDF (664K)
  • YUKINOBU IKEYA, HEIHACHIRO TAGUCHI, ITIRO YOSIOKA
    1982 Volume 30 Issue 9 Pages 3207-3211
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A new dibenzocyclooctadiene lignan, gomisin R (4), was isolated from the fruits of Schizandra chinensis BAILL. (Schizandraceae) together with wuweizisu C (1)2) and schisantherin D (12).3) The absolute structure of 1, the relative structure of which has been elucidated by Liu et al., was defined by chemical correlation with dimethylgomisin J (2). The absolute structure of 4 was determined by spectral studies and chemical correlation with 1.
    Download PDF (692K)
  • HIDEO YAMAGUCHI, MASAO ARIMOTO, MARIKO TANOGUCHI, TOSHIMASA ISHIDA, MA ...
    1982 Volume 30 Issue 9 Pages 3212-3218
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Two kinds of lignans were isolated from the seeds of Hernandia ovigera L. (Hernandiaceae) besides the previously reported desoxypodophyllotoxin (DPT), desoxypicropodophyllin, bursehernin and podorhizol, and the structures of the lignans were clarified. One, C23H24O8, named hernandin, was identified as 5-methoxy-desoxypodophyllotoxin (I) by both chemical and X-ray crystallographic methods. The other, C22H18O7, was obtained by purification of a previously isolated impure substance, mp 270-275°C. This compound was identified as 1, 2, 3, 4-dehydrodesoxypodophyllotoxin (II). This is the first report of the natural occurrence of II in plants.
    Download PDF (699K)
  • MASAMICHI FUKUOKA
    1982 Volume 30 Issue 9 Pages 3219-3224
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    5-O-Caffeoylshikimic acid (dactylifric acid) was isolated from bracken fern as a major constituent of its acutely toxic fraction, which causes depression of leucocytes and thrombocytes in calves. 5-O-Caffeoylshikimic acid exhibited an antithiamine effect in vitro, but had no hematuric effect in guinea pigs.
    Download PDF (710K)
  • MASAICHIRO MASUI, YORIE KAIHO, TAKAHIRO UESHIMA, SHIGEKO OZAKI
    1982 Volume 30 Issue 9 Pages 3225-3230
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The effects of a wide range of bases on the oxidation potentials of hydroxamic acids and 4-fluorophenol were examined by cyclic voltammetry mainly in acetonitrile and methylene chloride at a glassy-carbon electrode. Linear relationships between ΔEp and the pKa of the added bases were obtained for pyridines and for aliphatic amines, where ΔEp=Epl-Epe ; Epl is the potential for the original first wave and Epe is that for the extra wave developed in the presence of the base. The results can be largely explained in terms of the formation of hydrogen-bonded complexes between the acids and the bases.
    Download PDF (637K)
  • KAZUHIRO WATANABE, ITSUO YOSHIZAWA
    1982 Volume 30 Issue 9 Pages 3231-3238
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    For the direct assay of the enzymatic O-methylation products of 2-hydroxyestradiol 17β-sulfate (2) and 17β-glucuronide (3), the corresponding guaiacol estrogens have been prepared as authentic specimens and their high-performance liquid chromatography (HPLC) was investigated. The materials synthesized were : potassium 3-hydroxy-2-methoxyestra-1, 3, 5 (10)-trien-17β-yl sulfate (7), potassium 2-hydroxy-3-methoxyestra-1, 3, 5 (10)-trien-17β-yl sulfate (13), potassium [3-hydroxy-2-methoxyestra-1, 3, 5 (10)-trien-17β-yl-β-D-glucopyranosid] uronate (9), and potassium [2-hydroxy-3-methoxyestra-1, 3, 5 (10)-trien-17β-yl-β-D-glucopyranosid]-uronate (15). These sulfates and glucuronides were separated quantitatively by reversed-phase HPLC. The separation was performed with a mixture of acetate buffer (50 mM, pH 5.0) and methanol (50 : 50) as the mobile phase on a column of ODS SIL. The eluates were monitored in terms of the absorbance at 280 nm. Calibration curves between the amounts of conjugated guaiacols and the peak heights on chromatograms were all linear. The results obtained by proposed HPLC method for the quantification of O-methylated products obtained by the incubation of 2 and/or 3 with purified rat liver catechol O-methyltransferase in the presence of (3H3C)-S-adenosyl-L-methionine were in good agreement with the results obtained by a different procedure, the reverse isotope dilution method.
    Download PDF (952K)
  • MATAJIRO KOYAMA, ATSUSHI TAKADA, TAKEO UEDA
    1982 Volume 30 Issue 9 Pages 3239-3243
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The color reaction of benzaldehyde with 1-naphthol in concentrated sulfuric acid was investigated to clarify the reaction mechanism. Two reaction products, bis (4-hydroxy-1-naphthyl) phenylmethane and (4-hydroxy-1-naphthyl) (1-hydroxy-2-naphthyl) phenylmethane were isolated from the colored reaction mixture. These compounds were found to be the key intermediates to the colored species, since their colored solutions in concentrated sulfuric acid showed the same absorption maxima as the colored solution obtained from the reaction of benzaldehyde with 1-naphthol in the acid. The compounds were considered to be transformed into the cation radicals, which were responsible for the colorations.
    Download PDF (533K)
  • KAZUHIRO INAGAKI, ICHITOMO MIWA, TAMOTSU YASHIRO, JUN OKUDA
    1982 Volume 30 Issue 9 Pages 3244-3254
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The development of potent aldose reductase inhibitors as therapeutic agents for diabetic complications is highly desirable. The inhibitory action of 54 hydantoin derivatives consisting of 25 hydantoins, 21 2-thiohydantoins and 8 2-alkylthiohydantoins was therefore tested on rat and bovine lens aldose reductases in vitro. 1-(Phenylsulfonyl)-hydantoin (18) and its derivatives, 1-[(substituted phenyl) sulfonyl] hydantoins, were found to be potent inhibitors of the enzymes. 1-[(p-Bromophenyl) sulfonyl] hydantoin (49) was the most potent among them. It inhibited purified rat and bovine lens aldose reductases by 50% at 7×10-7M and 3.7×10-7M, respectively. Inhibition of rat and bovine lens aldose reductases by this compound (49) was due to its non-ionized form, but not the ionized form, and was of a non-competitive type with respect to DL-glyceraldehyde as a substrate.
    Download PDF (1163K)
  • TAKUO KOSUGE, KUNIRO TSUJI, KOICHI HIRAI
    1982 Volume 30 Issue 9 Pages 3255-3259
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Isolation of neosurugatoxin, which is the causative agent of intoxication resulting from the ingestion of the toxic Japanese ivory shell, was accomplished by six steps of chromatography, to give 4 mg of the principle from 20 kg of the shellfish.
    Download PDF (574K)
  • SUSUMU TSUSHIMA, YOSHIO YOSHIOKA, SEIICHI TANIDA, HIROAKI NOMURA, SHOS ...
    1982 Volume 30 Issue 9 Pages 3260-3270
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Twenty-seven alkyl analogs of lysophospholipid were synthesized and their structureantimicrobial activity relationships were examined. These analogs differed in the structures of the long-chain alkyl moiety at position 1 and the β-N-substituted aminoethylphosphoryl moiety at position 3, and in the presence or absence of the 2-methoxy group of the glycerol moiety. Many of the alkyl lysophospholipids were found to possess antimicrobial activities much more potent than those of naturally occuring lysolecithin and lecithin against Tetrahymena pyriformis W and a variety of fungi, including human pathogens. The maximal activity was observed with 2-methyl-1-tetradecylglycero-3-phosphocholines. 1-Alkyl-2-methylglycero-3-phosphocholines with longer as well as shorter alkyl chains tended to have lower antimicrobial activity. Alkyl lysophospholipids with pyridinioethyl instead of the choline group showed potent antifungal activity comparable to alkyl glycerophosphocholines with the corresponding alkyl group but lower antiprotozoal activity. The tetradecyl congeners in these two classes of compounds showed potent inhibitory activity against Trichophyton species, comparable to that of clotrimazole. In contrast, alkyl lysophospholipids with an ethanolamine moiety in the polar head group showed decreased activity. Changing the molecular backbone from glycerol to 1, 3-propanediol had little effect upon the activity, and the resulting 1-alkyl-2-deoxyglycero-3-phosphocholines displayed antimicrobial properties similar to those of 1-alkyl-2-methylglycero-3-phosphocholines.
    Download PDF (1243K)
  • TAKASHI ABIKO, HIROSHI SEKINO
    1982 Volume 30 Issue 9 Pages 3271-3277
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The octadecapeptide, H-Arg-Lys-Asp-Val-Tyr-Val-Glu-Leu-Tyr-Leu-Gln-Ser-Leu-Thr-Ala-Leu-Lys-Arg-OH, corresponding to positions 32 to 49 of the revised amino acid sequence of bovine thymopoietin II was synthesized by the azide condensation of three fragments, (32-36), (37-41) and (42-49), followed by deprotection with hydrogen fluoride in the presence of anisole-thioanisole-o-cresol. The in vitro addition of the synthetic thymopoietin II (32-49) significantly restored the low E-rosette forming capacity of cells from an aged patient with chronic renal failure to normal levels. The in vitro effect of [Gln38, Thr43, Val47]-thymopoietin II (32-49) on the low E-rosette forming capacity of cells from the aged patient with chronic renal failure was also compared with that of the synthetic thymopoietin II (32-49). The [Gln38, Thr43, Val47]-thymopoietin II (32-49) was approximately equipotent with the synthetic thymopoietin II (32-49) at a concentration of 100 μg/ml.
    Download PDF (983K)
  • ISAO KIJIMA, KUNIO EZAWA, SATOSHI TOYOSHIMA, KIMIO FURUHATA, HARUO OGU ...
    1982 Volume 30 Issue 9 Pages 3278-3283
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    We investigated the effects of α- and β-methylglycosides of N-acetylneuraminic acid and its disaccharide derivatives on the proliferation and immunological functions of murine lymphocytes. The Con A-induced increase of deoxyribonucleic acid (DNA) synthesis was enhanced by some of these neuraminic acid derivatives when they were added to a culture of murine spleen lymphocytes, and the disaccharide nucleosides, 5-fluoro-2', 3'-isopropylidene-5'-O-(4-N-acetyl-2, 4-dideoxy-3, 6, 7, 8-tetra-O-acetyl-1-methoxycarbonyl-D-glycero-α-D-galactooctapyranosyl) uridine (compound 9) and 2', 3'-di-O-acetyl-4-N-acetyl-2, 4-dideoxy-3, 6, 7, 8-tetra-O-acetyl-1-methoxycarbonyl-D-glycero-α-D-galactooctapyranosyl) inosine (compound 10), were especially effective. The above disaccharide nucleosides and their starting materials, 5-fluoro-2', 3'-isopropylidene uridine (FIU) and 2', 3'-di-O-acetylinosine (DAI), suppressed in vitro primary antibody response toward sheep red blood cells (SRBC). In vivo antibody response toward SRBC was also suppressed when compound 9 or 10 was injected intraperitoneally into mice with SRBC. Moreover, lymphocytes incubated with compound 9 or 10 showed suppressor activity on primary anti-SRBC antibody response in vitro. On the other hand, FIU and DAI did not induce the suppressor activity of murine lymphocytes in vivo or in vitro. The induction of suppressor cells by compounds 9 and 10 was abolished by pretreatment of the lymphocytes with anti Thy-1 antiserum plus complement. These results suggest that these disaccharide nucleosides can induce suppressor T cells.
    Download PDF (641K)
  • YOSHIHITO WATANABE, TETSUO ADACHI, YOSHIMASA ITO, KAZUYUKI HIRANO, MAM ...
    1982 Volume 30 Issue 9 Pages 3284-3287
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Human α-fetoprotein was purified from human cord serum by employing an immunoadsorbent column, consisting of Sepharose 4B coupled with the antibody to human α-fetoprotein. By using 0.2 M Na2CO3 solution (pH 11.6) as an eluent, a large amount of α-fetoprotein was obtained from the immunoadsorbent column in high purity. After gel filtration on Sephadex G-150, the purified α-fetoprotein was demonstrated to be homogeneous by polyacrylamide gel electrophoresis, SDS-gel electrophoresis and two-dimensional immunoelectrophoresis.
    Download PDF (786K)
  • MASAMI MORITA, SADAO HIROTA
    1982 Volume 30 Issue 9 Pages 3288-3296
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Crystallinity in a drug has been recognized as an important factor affecting stability, bioavailability and scaling-up of processing operations. In the field of pharmaceutical technology, crystallinity has not yet been studied by Ruland's method, one of the X-ray diffraction methods. The main reason why it is not used may be that extensive and time-consuming calculations are required. We have developed a modified Ruland's method for determining the degree of crystallinity and the disorder parameter from a powder diffraction pattern of a single sample, suitable for a routine preformulation test. The degrees of crystallinity and the disorder parameters in several samples were measured by Ruland's method and the effects of integral upper limits, integral lower limits, and normalization constants on the results were examined. The integral upper limits seemed to affect the results calculated from the data of a single sample very much, and so did the separation of the crystalline scattering intensity from the total scattering intensity. We completed a computer program for the automated separation of the noncrystalline scattering intensity and the determination of integral upper limits with a single sample. The results can be obtained from a single sample with satisfactory accuracy for the practical purpose of pharmaceutical technology.
    Download PDF (830K)
  • YOICHI SAWAYANAGI, NAOKI NAMBU, TSUNEJI NAGAI
    1982 Volume 30 Issue 9 Pages 3297-3301
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The permeation of several drugs such as promethazine hydrochloride, chlorpromazine hydrochloride, diethazine hydrochloride, triflupromazine hydrochloride, flufenamic acid, ketoprofen, indomethacin, and tolbutamide through a chitosan membrane was investigated as part of a series of studies on pharmaceutical applications of chitin and chitosan. The membrane constant of the chitosan membrane was found to be 0.0443, which is close to the reported value for cellulose membrane. The permeability decreased with increase in the molecular volume of the drugs. The activation energy of the permeation was found to be close to that of diffusion of usual organic drugs in water. Greater permeabilities were observed for acidic drugs than basic drugs. The observed effect of pH on the permeation of drugs through the chitosan membranes was considered to be attributable to the cationic state of the chitosan membrane. These results suggested that the permeation of drugs through the chitosan membrane is controlled mainly by diffusion through pores, especially in the case of drugs of smaller molecular volume, and depends upon the cationic state of the chitosan membrane.
    Download PDF (627K)
  • HIDEYA YAGINUMA, YUTAKA KOHNO, SOICHI ITOH, KAZUYOSHI KUBO, MASARU YAM ...
    1982 Volume 30 Issue 9 Pages 3302-3309
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The effect of the concurrent administration of amino acid (s) with diclofenac sodium on mucosal damage was investigated in rats. An administration of the base alone (Witepsol H-15) induced an only slight desquamation of epithelial cells at the luminal border, without any other changes in the mucosa. However severe histological damage was observed after administration of diclofenac sodium in Witepsol H-15. On the other hand, the concurrent administration of diclofenac sodium with L-form amino acids decreased the damage to the rectal mucosa. The L-forms of methionine and phenylalanine were the most effective, but in the protective effect of their D-forms on the rectal mucosa was much less. Rectal absorption of diclofenac sodium was investigated when it was administered with amino acids in rabbits, dogs and rats. The absorption was not influenced by the concurrent administration of amino acid (s).
    Download PDF (1376K)
  • HIROSHI FUJIWARA, SUSUMU KAWASHIMA, YUTAKA YAMADA, KUNIHIKO YABU
    1982 Volume 30 Issue 9 Pages 3310-3318
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The degradation of ampicillin was found to be inhibited due to the formation of a 1 : 1 molar adduct between ampicillin and furfural in alkaline solution. Further, specific base catalysis for the ampicillin-furfural adduct was smaller, about one hundredth, than that for ampicillin-benzaldehyde adduct. Since the adduct formation was not observed at pH<6.00 and the formation constant of the adduct increased with increase of pH, the α-amino group of ampicillin was concluded to participate in the adduct formation. Thus, to estimate the structure of the adduct, various physical properties of samples prepared by freeze-drying of alkaline solution containing ampicillin and an aldehyde (furfural or benzaldehyde) were studied and compared. Nuclear magnetic resonance (NMR) spectra (1H and 13C), infrared (IR) spectra and the chemical properties of the adducts suggested that they have a Schiff base structure formed from the aldehyde moiety and the free amino group of ampicillin.
    Download PDF (872K)
  • JUNJI HIRATE, JUN WATANABE, KIKUO IWAMOTO, SHOJI OZEKI
    1982 Volume 30 Issue 9 Pages 3319-3327
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    ^<14>C-Thiourea was given by intravenous or oral administration at low dose (100 μCi/0.16 mg/kg) or high dose (100 μCi/160 mg/kg) to rats, then plasma levels were determined and whole-body autoradiography was carried out. The gel filtration on Sephadex G-10 of plasma obtained following intravenous administration of 14C-thiourea showed that the radioactive label was covalently bound to macromolecular fractions such as protein or peptide. Plasma levels of radioactivity due to unchanged thiourea and the tissue distribution patterns of radioactivity following intravenous administration of 14C-thiourea were considerably different after a low dose and a high dose. The elimination of thiourea from plasma in the case of high dose (t1/2β ; 7.0 h) was greatly delayed as compared with that after the low dose (t1/2β ; 0.69 h). The whole-body autoradiograms indicated that, at low dose, very high radioactivity was presented in the lung, thymus and Harder's gland at 60 min following intravenous administration, whereas when the high dose was administered the radioactivity was distributed almost homogeneously throughout the body. The remarkable dose dependency in thiourea distribution might be brought about by the saturation of irreversible covalent binding sites of plasma and tissue macromolecules in the case of high dose administration.
    Download PDF (1479K)
  • MAYUMI NAKAJIMA, ATSUKO FUKUHARA, SHIRO MORIMOTO
    1982 Volume 30 Issue 9 Pages 3328-3332
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The present study was carried out to assess the correlations among the results of four different assays of urinary kallikrein in the rat. The peptidase, esterase, vasodilator and kinin-forming activities were employed for the determination of kallikrein activity. In the urine from normal rats, the assay of vasodilator activity correlated well with those of peptidase activity (γ=0.86) and esterase activity (γ=0.89). Similarly, a good correlation (γ=0.99) was obtained between the peptidase and kinin-forming activities. However, in the case of rats under furosemide treatment, the esterase activity was poorly correlated with the peptidase and vasodilator activities, although a good correlation (γ=0.93) was obtained between the latter two activities. These results indicate that the peptidase and vasodilator activities should be employed for the estimation of urinary kallikrein excretion in the urine from rats under furosemide treatment.
    Download PDF (614K)
  • MIKIO ARISAWA, JUNKO OHSHIMA, HIROMI B. MARUYAMA
    1982 Volume 30 Issue 9 Pages 3333-3339
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Nine 2-aminothiazolyl methoxyiminoacetamidocephems (ATOICs) as well as several 4-furyl methoxyiminoacetamidocephems (FOICs) were compared for stability to and inhibition of various β-lactamases. Although ATOICs are generally less stable to cefuroxymases (CXases) from Proteus vulgaris or Bacteroides fragilis, the relative susceptibility among them was greatly affected by the substitution at the 3-position : the compound unsubstituted at C-3 was most stable, while thiomethyleno-2-methyl-6-hydroxytriazine-5-one substitution gave the most labile compound. A similar tendency was also seen with FOICs, which, however, were in general more susceptible to those CXases than were ATOICs. The substitution at C-3 had lesser effects on the stability to some cephalosporinases (CSases), and also had little effect on the inhibitory activity of ATOICs and FOICs on various CSases, though the compound unsubstituted at C-3 (ceftizoxime) exhibited the least inhibition among ATOICs tested. The Ki values of typical ATOICs except ceftizoxime were 10-8-10-9M for enzymes from Citrobacter freundii, and Enterobacter sp. and 10-6-10-7M for those from Escherichia coli, Serratia marcescens and Pseudomonas aeruginosa.
    Download PDF (680K)
  • HARUHISA KIZU, TSUYOSHI TOMIMORI
    1982 Volume 30 Issue 9 Pages 3340-3346
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Three triterpenoid prosapogenins tentatively named CP0 (I), CP2a (VI) and CP3a (VIII) have been isolated from the alkaline hydrolysate of the crude saponin obtained from the root of Clematis chinensis OSBECK. On the basis of chemical and physicochemical evidence, they were characterized as follows : I, hederagenin 23-O-α-L-arabinopyranoside ; VI, hederagenin 23-O-β-D-glucopyranoside ; VIII, olean-12-ene-28-oic acid-3β, 24-diol (4-epihederagenin) 3-O-α-L-rhamnopyranosyl-(1→2)-α-L-arabinopyranoside. I and VI are the first examples of 23-O-glycosides of oleanane-type triterpenes to be isolated from nature. Six less polar genuine saponins were isolated and they appeared to be identical with prosapogenins CP4, CP6, CP7, CP8, CP9 and CP10, 1, 4) on the basis of the results of high performance thin-layer chromatography (HPTLC).
    Download PDF (907K)
  • KAZUO ITO, TOSHIYUKI IIDA, KAZUHIKO ICHINO, MASA TSUNEZUKA, MASAO HATT ...
    1982 Volume 30 Issue 9 Pages 3347-3353
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Two new biphenyl ether lignans, named obovatol (1) and obovatal (2) were isolated from the leaves of Magnolia obovata THUNB. (Magnoliaceae) and their chemical structures were determined to be 4', 5-diallyl-2, 3-dihydroxybiphenyl ether and 3, 4-dihydroxy-5-(pallylphenoxy) cinnamic aldehyde by means of chemical and spectral studies. The former substance showed antibacterial activity against a cariogenic bacterium, Streptococcus mutans, but was less active than magnolol (5) and honokiol (6).
    Download PDF (808K)
  • KAZUYUKI KIZUKI, MASAFUMI KAMADA, MASAHIKO IKEKITA, HIROSHI MORIYA
    1982 Volume 30 Issue 9 Pages 3354-3361
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The presence of two main components of porcine pancreatic prokallikrein (prokallikreins A and B) was observed on anion-exchange chromatography. One of them, prokallikrein A, was purified by water extraction, followed by a combination of ammonium sulfate fractionation, ion-exchange chromatographies, preparative disc gel electrophoresis and gel filtration in the presence of various protease inhibitors. Prokallikrein A reacted immunologically with the antibody against kallikrein A obtained from autolyzed porcine pancreas. It migrated slightly more slowly than kallikreins A and B on immunoelectrophoresis. The molecular weight of prokallikrein A was estimated to be 3.6-3.9×104 by high performance liquid chromatography. On the other hand, those of kallikreins A and B were estimated to be 3.3-3.4×104 by the same method. Thus, it was speculated that at least large fragment would not be liberated from the prokallikrein molecule during the activation by protease (s).
    Download PDF (907K)
  • KAZUMOTO CHIBA, MIEKO SAKAMOTO, MEGUMI ITO, NAOMI YAGI, HITOSHI SEKIKA ...
    1982 Volume 30 Issue 9 Pages 3362-3369
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Interactions between two sulfonamides and ibuprofen or mepirizole were investigated in dogs by pharmacokinetic analysis. When sulfamethizole was coadministered with an acidic non-steroidal anti-inflammatory agent, ibuprofen, the β-elimination half-life (t1/2 (β)) for sulfamethizole was increased approximately 10 times compared to the control, whereas the coadministration of sulfamethizole with a basic agent, mepirizole, resulted in slightly prolonged or unchanged t1/2 (β) compared to the control. In the case of sulfanilamide, the time course of plasma level was not altered by coadministration of either antiinflammatory agent. An attempt was made to elucidate the mechanism by which ibuprofen alters the pharmacokinetics of sulfamethizole. In renal clearance experiments, the clearance ratio of sulfamethizole was markedly decreased after ibuprofen infusion. No significant alteration of protein binding of sulfamethizole was found when ibuprofen was added to dog plasma. These results suggest that the increased terminal half-life of sulfamethizole caused by ibuprofen is mainly a result of competitive interactions between them at the renal secretory level.
    Download PDF (744K)
  • TAKEHISA KUNIEDA, WENCHIEH TSAI, MASATO NARUSHIMA, MASAAKI HIROBE
    1982 Volume 30 Issue 9 Pages 3370-3373
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    1, 11-Dioxa [11] paracyclophane (2, 12-dioxabicyclo [11, 2, 2] heptadeca-13, 15, 16-triene) derivatives were thermally equilibrated in cholesteric mesophases. The enantiomeric and diastereomeric equilibria were biased to the extents of 0.4 and 3.6% (e.e.), respectively, independently of the macrostructural handedness of the mesophases.
    Download PDF (511K)
  • CHIAKI TANAKA, KEIKO NASU, NORIKO YAMAMOTO, MEGUMI SHIBATA
    1982 Volume 30 Issue 9 Pages 3374-3376
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Treatment of 2-amino-5-phenyl-2-oxazoline (I) with trifluoroacetic anhydride gave 1, 3-bis (2, 2, 2-trifluoroacetyl)-1-[2-(2, 2, 2-trifluoroacetyloxy)-2-phenylethyl] urea (II). Compound II was pyrolyzed at 120°C to give 2, 2, 2-trifluoroacetyl isocyanate (III) and 2, 2, 2-trifluoro-N-[2-(2, 2, 2-trifluoroacetyloxy)-2-phenylethyl] acetamide (IV). Compound II was readily hydrolyzed to give 1-(2, 2, 2-trifluoroacetyl)-3-[2-(2, 2, 2-trifluoroacetyloxy)-2-phenylethyl] urea (VI). Compound VI was pyrolyzed at 230°C to give 2, 2, 2-trifluoroacetamide (VII) and 2, 2, 2-trifluoro-N-(E)-styrylacetamide (VIII). It was considered that the formation of VIII proceeded through the intermediate 2-(2, 2, 2-trifluoroacetyloxy)-2-phenylethyl isocyanate (XI) or 1-(2, 2, 2-trifluoroacetyl)-3-(E)-styrylurea (XII), as shown in Chart 1.
    Download PDF (481K)
  • KOSAKU HIROTA, KAZUO MARUHASHI, TETSUJI ASAO, SHIGEO SENDA
    1982 Volume 30 Issue 9 Pages 3377-3379
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Thermolysis of 6-azido-1, 3-dimethyluracil (1) in formamide gave 1, 3, 6, 8-tetramethylpyrimido [5, 4-g] pteridine-2, 4, 5, 7 (1H, 3H, 6H, 8H)-tetrone (3), while the same reaction in N, N-dimethylformamide (DMF) gave 3-(5-amino-1, 3-dimethyluracil-6-yl)-4, 6-dimethyl-[1, 2, 3] triazolo [4, 5-d] pyrimidine-5, 7 (4H, 6H)-dione (4), which was converted into 3 in refluxing formamide. Compound 4 was also obtained by the treatment of 1 with 4, 6-dimethyl [1, 2, 3] triazolo [4, 5-d] pyrimidine-5, 7 (4H, 6H)-dione (5) in refluxing DMF. The mechanism of these reactions is discussed.
    Download PDF (430K)
  • SHIGEHIRO MORI, IZUMI SAKAI, TOYOHIKO AOYAMA, TAKAYUKI SHIOIRI
    1982 Volume 30 Issue 9 Pages 3380-3382
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The stable and safe compound trimethylsilyldiazomethane (TMSCHN2), which is useful for organic synthesis, was prepared in high yield by diazo-transfer reaction of trimethylsilylmethylmagnesium chloride with diphenyl phosphorazidate [DPPA, (C6H5O)2-P (O) N3].
    Download PDF (460K)
  • YOSHIO ITOH, SETSUZO TEJIMA
    1982 Volume 30 Issue 9 Pages 3383-3385
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A facile synthesis of the title compound (9) was achieved. Tritylation of 1, 6-anhydro-β-N-acetylglucosamine (4) with 3 molar equivalents of trityl chloride in pyridine at 85-90°C for 24 h afforded the corresponding 4-O-trityl ether (5) and 3-O-trityl ether (6) in 66.9 and 8.3% yields, respectively, after column chromatography. Benzylation of 5 followed by detritylation provided 9 via 3 steps from 4 in ca. 47% yield. Preparation of 2-acetamido-1, 6-anhydro-4-O-benzyl-2-deoxy-β-D-glucopyranose (10) from 6 was also carried out.
    Download PDF (513K)
  • YASUO KIKUGAWA, MASATO KASHIMURA
    1982 Volume 30 Issue 9 Pages 3386-3388
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    On catalytic transfer hydrogenation using formic acid and palladium black, 2, 3-dihydrotryptophan derivatives were obtained from tryptophan derivatives. The results showed that care is necessary in the removal of protecting groups in peptides containing tryptophans.
    Download PDF (371K)
  • YOSHIO MATSUBARA, TOSHIYUKI NAKAMURA, MASAKUNI YOSHIHARA, TOSHIHISA MA ...
    1982 Volume 30 Issue 9 Pages 3389-3391
    Published: September 25, 1982
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Several new S-alkylated thioformimidates were synthesized and their applications to the synthesis of thiolformates and thiols were investigated.
    Download PDF (349K)
feedback
Top