Chemical and Pharmaceutical Bulletin
Online ISSN : 1347-5223
Print ISSN : 0009-2363
ISSN-L : 0009-2363
Volume 28, Issue 2
Displaying 1-50 of 58 articles from this issue
  • NAOHIRO SHIRAI, HISAO NAKATA, TOYO KAIYA, JINSAKU SAKAKIBARA
    1980 Volume 28 Issue 2 Pages 365-371
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The 13C-NMR signals of grayanotoxin-I, a tetracyclic diterpene with an A-nor-B-homo-ent-kaurane skeleton, were assigned by means of single-frequency off-resonance decoupling, selective proton decoupling and by comparison with spectra of derivatives.
    Download PDF (590K)
  • ATSUSHI WATANABE, YUMIKO YAMAOKA, KOJI KURODA
    1980 Volume 28 Issue 2 Pages 372-378
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    When measurement of the refractive index of crystals is used for the identification of a drug or polymorphic form, it is more convenient to determine the "key refractive index" than the principal refractive index. Two characteristic key refractive indices can be measured for crystals in plate, bladed, or lamellar form, but for acicular or long prismatic crystals only one key refractive index can usually be measured along the direction of elongation when the crystals show parallel extinction. Key refractive indices were measured for several drugs listed in J.P.IX. The relation between the key refractive indices and crystal symmetry, habit, and extinction is discussed. The immersion method was improved to obtain more reliable results by using a commercial kit of immersion oils with refractive indices between 1.47 and 1.73 at 0.005 intervals.
    Download PDF (653K)
  • TOSHIO KASAMA, KIICHIRO OSHIRO, MICHIKO UCHIDA, TAKASHI OKUBO, NORIKO ...
    1980 Volume 28 Issue 2 Pages 379-386
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Since purity and source species might influence the absorption of a long-acting monocomponent insulin (MC-insulin) preparation, studies were carried out with a preparation of mixed bovine and porcine insulin obtained by conventional recrystallization, a specially prepared preparation of mixed MC grade bovine and porcine insulin, a specially prepared preparation composed of recrystallized porcine insulin alone, and MC-insulin. These were injected subcutaneously into rabbits, and their absorption were examined using hypoglycemia as an index. The preparation composed only of porcine insulin gave a higher maximum hypoglycemic rate than the mixed bovine and porcine insulin preparation, but the blood glucose recovered to the original level earlier, indicating its more rapid absorption. This phenomenon was more strinking at larger test doses. However, there was no difference in absorption between preparations of different purity. Furthermore, the mixed MC-grade bovine and porcine insulin preparation and MC-insulin were subcutaneously administered to an in situ animal experimental model to determine the blood insulin course. MC-insulin was again found to be absorbed more rapidly. It is suggested that this occurs because porcine insulin crystals dissolve more rapidly than bovine insulin crystals. Such a difference in the solubility of the insulin crystals may be caused by configurational differences between porcine and bovine insulin, arising from the differences in the amino acid residues at positions 8 and 10 of the A-chain.
    Download PDF (1019K)
  • TAKASHI ABE, JUNKO MAYUZUMI, NORIKO KIKUCHI, SEIICHI ARAI
    1980 Volume 28 Issue 2 Pages 387-392
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The seasonal variations of skin temperature, skin pH, evaporative water loss (EWL) and skin surface lipid values were determined in the forearm skin of 24 healthy Japanese adults at four different periods over one year. The results provide new information on the changing systemic patterns of the four skin parameters. It was found that EWL, total lipid and squalene values in July were approximately 2-3 times those in January. On the other hand, the skin pH was significantly lower in July than in January, April or October. However, the extents of change of skin pH and cholesterol values were less than those of EWL, squalene and total lipid values on the basis of statistical analysis. Furthermore, significant correlations were observed between skin temperature and EWL, total lipid and squalene values. On the other hand, significant inverse correlations were found between skin pH and skin temperature and EWL values. On the basis of these results, the dependence of sebaceous lipids on skin temperature and the contribution of eccrine sweating to the skin pH are discussed.
    Download PDF (649K)
  • TOMOAKI FUKUDA, YOSHINOBU FUKUMORI, SHINJI WADA, YOSHINOBU HANYU
    1980 Volume 28 Issue 2 Pages 393-400
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The internal friction of powders was examined during tablet preparation for lactose and phenacetin as base materials and for potassium chloride, potassium bromide and sodium chloride as model samples. The coefficient of internal friction decreased with the porosity, even if the voids were filled by fragmentation of particles or by plastic flow. When the porosity had reached zero, the state of the whole bed of powder was found to be plastic, where the coefficient of internal friction was zero, for magnesium stearate, phenacetin, potassium bromide and potassium chloride. The plastic behavior of magnesium stearate and phenacetin was considered to be related to the "capping" of tablets containing them. For lactose granules, the porosity did not reach zero even at a compacting pressure of 4000 kg/cm2, the maximum used in this work, because of its crystal hardness. The coefficient of internal friction was large and decreased only gradually with the porosity. Mixtures of two components exhibited intermediate internal friction, corresponding to the composition. However, when hydroxypropyl cellulose (HPC) was added to phenacetin powder as a binder, HPC increased the yield stress from 2000 to 3100 kg/cm2 even at a content of only 2% w/w on a dry basis. In contrast, lubricants (magnesium stearate and talc) added to granules had little or no effect on the internal friction, although this small amount was sufficient to reduce the wall friction.
    Download PDF (814K)
  • KAZUHIRO KOIKE, CAROLYN BEVELLE, SUNILK. TALAPATRA, GEOFFREYA. CORDELL ...
    1980 Volume 28 Issue 2 Pages 401-405
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The cytotoxic activity of extracts of Asclepias albicans (Asclepiadaceae) was traced to uzarigenin (2), desglucouzarin (3) and a new cardenolide glycoside, uzarigenin 3-O-β-D-glucopyranosyl (1→4)-β-D-glucopyranoside (4). The latter compound also displayed significant antitumor activity against P-388 lymphocytic leukemia.
    Download PDF (696K)
  • SHINICHI HAYASHI, AKIYOSHI YOSHIDA, HITOSHI TANAKA, YOKO MITANI, KYOKO ...
    1980 Volume 28 Issue 2 Pages 406-412
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A high-performance liquid chromatographic method for the separation of 8-glucosyl-rhein, sennosides A, B, C, and a new compound, sennoside G, was established. The extraction of sennosides from senna was also investigated. Sennosides, extracted from senna or pharmaceutical preparations, were directly injected and separated using a Permaphase ODS column or a Lichrosorb Rp-18 column, and the sennosides contents in Sennae folium pulv., Sennae folium, and pharmaceutical preparations containing sennoside Ca were determined.
    Download PDF (645K)
  • REIJI HIRAMATSU, TAKAHARU MIZUTANI, AKIRA MIZUTANI
    1980 Volume 28 Issue 2 Pages 413-416
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The effects of bovine thymus hypocalcemic protein (TP1) on cAMP and cGMP levels in murine thymus cells were studied. TP1 elevated cGMP levels 12.5-fold at a concentration of 0.15 mg/ml after incubation for 5 min but the amount of cAMP in thymocytes was lowered. These results suggest that TP1 stimulates the differentiation of early T-cells to mature T-cell subsets, such as helper cells, but does not stimulate that of stem cells to early T-cells.
    Download PDF (519K)
  • SHIGERU AONUMA, YASUHIRO KOHAMA, YUTAKA KOMIYAMA, SHIGEKO FUJIMOTO, MA ...
    1980 Volume 28 Issue 2 Pages 417-423
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The complete amino acid sequence of a biologically and immunologically active fragment, Fr. AA-1, of salivary gland hormone (parotin) has been established through the uses of combined dansyl-Edman procedure, manual Edman degradation and standard enzymatic and chemical techniques. The binding site of the carbchydrate moiety was determined from the amino acid composition of the carbohydrate moiety obtained from digests of Fr. AA-1 with pronase. Fr. AA-1 was shown to be a glycopeptide consisting of 58 amino acid residues. Its sequence is : H2N-Leu^^1-Tyr-Ile-Leu-Tyr-Phe-Phe-Gln-Ser-Asp^^10-Asn-Glu-Asp-Lys-Glu-Lys-Val-Val-Arg-Gln^^20-Glu-Glu-Gly-Glu-Glu-Arg-Ile-Thr-Ala-Leu^^30-Leu-Met-Asn (carbohydrate moiety)-Gly-Ser-Ala-Leu-Lys-Gln-Glu^^40-Glu-Trp-Trp-Glu-Gly-Lys-Glu-Asp-Thr-Asp^^50-Asp-Thr-Ala-Ile-Val-Leu-Leu-Lys^^58-COOH.
    Download PDF (741K)
  • KAZUKO SHICHIRI, KAZUHISA FUNAKOSHI, SEITARO SAEKI, MASATOMO HAMANA
    1980 Volume 28 Issue 2 Pages 424-430
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The reaction of quinoline 1-oxide (1) with tosyl chloride (1 eq) and a large excess of triethylamine in a mixture of chloroform and water at room temperature gave 2-tosyloxy-quinoline (2 ; trace), di (2-quinolyl) ether (3 ; 46%), N-(2-quinolyl)-2-quinolone (4 ; 35%) and carbostyril (5 ; 19%). Products 3 and 4 presumably originate from nucleophilic attack of carbostyril anions (5'a and 5'b) on the tosyl chloride adduct of 1 (11). "Di (2-quinolyl) ether" reported by Murakami and Matsumura to be obtained by heating 1 with tosyl chloride was shown in fact to be an equimolar complex of 4 and 5, and the nature of the reaction was clarified in detail.
    Download PDF (705K)
  • TADAO MAEDA, HIROSHI TAKENAKA, YOSHIYA YAMAHIRA, TAKESHI NOGUCHI
    1980 Volume 28 Issue 2 Pages 431-436
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The α and β forms of chloramphenicol palmitate (CPP) crystals were successfully obtained in similar particle sizes by utilizing wet micronization. Comparisons of the absorbability of CPP suspensions of the above crystals using conventionally fasted rabbits were unsuccessful because of the low plasma level after dosing the β form (less than 1 μg/ml). Plasma levels were then compared by dosing CPP suspensions of the α and β forms to stomach-emptying-controlled (SE-controlled) rabbits in a crossover manner. Compartive absorption studies were also performed in the SE-controlled rabbits with a capsule form of CP and suspensions of CPP in the α and β forms. It was demonstrated that absorption occurred in the order CP>CPP (α form)>CPP (β form). The correlation between in vitro extent of hydrolysis and in vivo absorbability in terms of AUC is discussed. These results suggest that the SE-controlled rabbit is a very useful animal model for preliminary bioavailability studies on oral dosage forms for human clinical use.
    Download PDF (606K)
  • JIRO KAWAMURA, TAKAO HAYAKAWA, TSUYOSHI TANIMOTO
    1980 Volume 28 Issue 2 Pages 437-446
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The reactivities of 20β-hydroxysteroid dehydrogenase [EC 1.1.1.53] from Streptomyces hydrogenans towards 64 kinds of steroids having the pregnan-20-one skeleton were investigated. The presence and nature of a substituent around the reacting 20-oxo group of the steroid played a decisive role in the interaction with the catalytic site of the enzyme. In general, steroids having bulky group (s) at C-21 and/or C-17 did not act as substrates, though steroids having a bulky group at C-3 or C-11 were utilized by the enzyme. The existence of a substituent at C-16 caused 20-oxo steroids to lose their reactivity almost completely, but the existence of a substituent at C-3, C-6, C-9, C-11, C-17 or C-21 did not. The configurational relationship between the plane of the steroid ring and the C-17 side chain might also be important in steroid recognition by the enzyme. Among the nonsubstrate steroids, 16β-methylpregn-4-ene-3, 20-dione was a competitive inhibitor with respect to both steroid substrate and coenzyme. It was inferred that the steroid ring of the inhibitor molecule competed with the substrate binding process, while the 16β-methyl substituent competed with the coenzyme action. When 16α-methylpregn-4-ene-3, 20-dione, pregn-4-ene-3, 20-dione and 17, 21-dihydroxypregn-4-ene-3, 11, 20-trione were used as substrates, the apparent Km values for NADH were 13.97, 3.91 and 4.07μM, respectively. The presence of a substituent at C-16 unfavorably affected the apparent Km for NADH, but substituents at C-21, C-17 or C-11 did not. These results suggest that the coenzyme may be located in the vicinity of C-16 of the steroid molecule in the ternary complex, steroid-coenzyme-enzyme.
    Download PDF (1050K)
  • AKIRA TAKAMIZAWA, YOSHIHIRO MATSUSHITA, HIROSHI HARADA
    1980 Volume 28 Issue 2 Pages 447-452
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Dialkyl acylphosphonate acted as an acylating agent for the 2-methyl group in 1, 3, 4-thiadiazolium salts, as well as 2-methylthiazolium and thiazolinium salts, to give 2-acylidenethiadiazoline derivatives. The reaction of diethyl benzoylphosphonate (1) with 5-amino-2, 3-dimethyl-1, 3, 4-thiadiazolium iodide (2c) gave the N-monobenzoate inner salt (6), C-monobenzoate (7), C, N-dibenzoate (8) and 5-benzoylamido-2-styryl-thiadiazolium hydroxide inner salt (9) ; the extent of N-acylation was greater than that of C-acylation. The structures of the inner salt products (6 and 9) were confirmed by an alternative synthesis. A proposed mechanism for the acylphosphonate reaction with 2-methyl-1, 3, 4-thiadiazolium salts is presented, and is consistent with the 13C-nuclear magnetic resonance chemical shifts of the 2-methyl groups of the 1, 3-azolium salts.
    Download PDF (677K)
  • YUKIO FUJITA, KENJI SEMPUKU, KOJI KITAGUCHI, TAMIKI MORI, HIROMU MURAI ...
    1980 Volume 28 Issue 2 Pages 453-458
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Various amides of tetrahydroabietic, Δ8-dihydroabietic, abietic and dehydroabietic acids were prepared and tested for hypocholesterolemic activity in cholesterol-fed rats. The introduction of an aromatic ring into the amine moiety of secondary amides markedly enhanced the activity of the parent acids. The secondary amides having an aliphatic ring were slightly less active than those having an aromatic ring, but those having an alkyl or allyl group were completely inactive. Tetrahydroabietic and dihydroabietic acids appear to be preferable (in terms of activity) to abietic or dehydroabietic acid as the acid moiety of these amide derivatives. N-Phenyltetrahydro (18)- or N-phenyl-Δ8-dihydroabietamide (19) and N-benzyltetrahydro (20)- or N-benzyl-Δ8-dihydroabietamide (21) are considered to be promising parent compounds for potent hypocholesterolemic drugs. In the case of benzyl derivatives of Δ8-dihydroabietamide, N-(4-methoxybenzyl)-Δ8-dihydroabietamide (49) and N-(α-benzylbenzyl)-Δ8-dihydroabietamide (58) were the most active, being more than 10 times as potent as the corresponding parent compounds.
    Download PDF (711K)
  • ATSUO NAGAMATSU, JINICHI INOKUCHI, SHINJI SOEDA
    1980 Volume 28 Issue 2 Pages 459-464
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Angiotensin I-converting enzyme of hog kidney exists in two different forms, peak A and peak B, as shown by hydroxyapatite column chromatography. These enzymes were found to possess identical molecular weights by filtration through Sephadex G-200. The sialic acid and neutral sugar contents of peak A were roughly twice those of peak B. The peak A preparation gave a single protein component on SDS-gel electrophoresis, while peak B showed three components. After neuraminidase treatment, however, both enzyme preparations showed a single protein band with an estimated molecular weight of 90000 as determined by SDS-gel electrophoresis. It is suggested that the different mobilities of the two forms on electrophoresis depend upon the relative amounts of sialic acid in their molecules, and that these enzymes consist of two polypeptide chains.
    Download PDF (804K)
  • KAZUYUKI WACHI, ATSUSUKE TERADA
    1980 Volume 28 Issue 2 Pages 465-472
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A new synthetic method for primary 2-aminopyridine derivatives is described. Treatment of the imidoyl chlorides of 1, 3-benzoxazines (1a-i) with pyridine N-oxides resulted in the introduction of an oxazine moiety into the α-position of the pyridine ring through rearrangement of the initially formed reaction adduct. Acid hydrolysis of the rearrangement products afforded 2-aminopyridine derivatives in excellent yields. When methoxypyridine N-oxides were used, products of a different type (10 and 14) were obtained.
    Download PDF (747K)
  • YASUJI YAMADA, SUKEJI SUZUKI, KAZUO IGUCHI, HIROYUKI KIKUCHI, YASUMASA ...
    1980 Volume 28 Issue 2 Pages 473-478
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Three new polyhydroxylated sterols (II), (IV) and (VI), along with the known sterol (I), were isolated from the Japanese soft coral Lobophytum pauciflorum (Ehrenberg). The structures of II, IV and VI were established to be 24ζ-methylcholestane-3β, 5α, 6β, 25-tetrol, 24ζ-methylcholestane-1β, 3β, 5α, 6β-tetrol and 24-methylenecholestane-1β, 3β, 5α, 6β-tetrol, respectively.
    Download PDF (634K)
  • EISUKE KAJI, SHONOSUKE ZEN
    1980 Volume 28 Issue 2 Pages 479-486
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    An improved method for the synthesis of 4-substituted-3, 5-bis (methoxycarbonyl)-isoxazoline N-oxides (3) was developed by one-step cyclization of aldehydes with methyl nitroacetate in N, N-dimethylacetamide. The reaction mechanism was found to involve intramolecular nitrite ion displacement by the nitronate anion (9) ; this was confirmed by the formation of 3, 5-bis (methoxycarbonyl)-4-phenylisoxazoline N-oxide-4-d (11), starting from benzaldehyde-1-d and methyl nitroacetate. Conversion of isoxazoline N-oxides (3) to the corresponding 3, 5-bis (butylcarbamoyl) isoxazoles (7) is also discussed.
    Download PDF (827K)
  • SOJI SUGAI, KAZUO TOMITA
    1980 Volume 28 Issue 2 Pages 487-492
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Heating of 2, 5-dimethyl-4-isoxazolin-3-thione (IIa) in acidic media gave 2, 5-dimethyl-4-isothiazolin-3-thione (IIIa) and bis (5-methyl-3-isoxazolyl) disulfide (IVa) in poor yields. On the other hand, 2-methyl-5-phenyl-4-isoxazolin-3-thione (IIe) afforded only bis (5-phenyl-3-isoxazolyl) disulfide (IVb). The reactions of 4-isoxazolin-3-thiones (IIa-d) with hydrogen sulfide or thioacetic acid in 48% hydrobromic acid gave 4-isothiazolin-3-thiones (IIIa-d) in moderate yields, while 5-phenyl-4-isoxazolin-3-thiones (IIe, f) afforded 3-imino-5-phenyl-1, 2-dithiols (VIe, f) in addition to 5-phenyl-4-isothiazolin-3-thiones (IIIe, f). A reaction mechanism is proposed. The thiol (VIe) was treated with base to give IIIe. An improved synthesis of the disulfides (IVa, b) was developed by the reaction of 2-methyl-4-isoxazolin-3-thiones (IIa, b) or 2-benzyl-4-isoxazolin-3-thiones (IIg, h) with bromine.
    Download PDF (667K)
  • KAZUKO SHICHIRI, KAZUHISA FUNAKOSHI, SEITARO SAEKI, MASATOMO HAMANA
    1980 Volume 28 Issue 2 Pages 493-499
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Reactions of N-oxides of lepidine (1a) and 4-methyl- (1b), 4-chloro- (1c) and 6-methoxyquinoline (1d) with tosyl chloride (1 eq) and triethylamine (ca. 10 eq) in a mixture of chloroform and water at room temperature gave the corresponding di (2-quinolyl) ethers (3a-d) and N-(2-quinolyl)-2-quinolones (4a-d). The efficiency of this type of reaction depends upon the nature and position of the substituents. Whereas the reaction of 1c with carbostyril under the same conditions gave small amounts of 4-chloro-2-tosyloxyquinoline (2c) and N-(4-chloro-2-quinolyl)-4-chloro-2-quinolone (4c), that of 1d afforded 6-methoxy-2-quinolyl 2'-quinolyl ether (15) and N-(6-methoxy-2-quinolyl)-2-quinolone (16) in 47 and 29% yields, respectively.
    Download PDF (763K)
  • NAOKO MORISAKI, JUN FURUKAWA, SHIGEO NOZOE, AKIKO ITAI, YOICHI IITAKA
    1980 Volume 28 Issue 2 Pages 500-507
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Cyclopropylarbinyl alcohols with marasmane and isomarasmane carbon skeletons (compounds 21 and 26) were synthesized from the azo compound 14 by a route involving an intramolecular carbene insertion reaction. The major product 21 was found to have an isomarasmane type stereostructure by X-ray crystallographic analysis of compound 23.
    Download PDF (915K)
  • TAMIKO SAKURAI, SEISHI TSUCHIYA, HIDEO MATSUMARU
    1980 Volume 28 Issue 2 Pages 508-513
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    To investigate the protein binding characteristics of thiamylal (TA), thiopental (TP) and 6-n-propyl-2-thiouracil (PTU), the effects of pH, ionic strength and temperature on the binding of these three drugs were compared by equilibrium dialysis. The bindings between each of the three drugs and bovine serum albumin (BSA) were similarly dependent on pH. However, the binding constants of TA and TP at pH 7.4 were scarcely influenced by changes of the ionic strength made by the addition of KCl or K2SO4, while that of PTU decreased with increase of the ionic strength. On the basis of these results and the considerable changes in the pKa of bound PTU (previous paper), it is possible that the binding of PTU involves a larger contribution from the interaction with the charged region on BSA than those of TA and TP. A thermodynamic study showed major contributions of ΔS° in the cases of TA and TP and of ΔH° in the case of PTU to ΔG°, suggesting that the binding sites of TA and TP might be different from those of PTU. The pH dependency of binding was also studied by fluorometric titration employing 8-anilinonaphthalene-1-sulfonate (ANS) as an indicator. The displacement ratios of TA and TP increased with increase of pH, in accord with the results of equilibrium dialysis, but that of PTU was only slightly dependent on pH.
    Download PDF (653K)
  • MASAYOSHI ARUGA, SHOJI AWAZU, MANABU HANANO
    1980 Volume 28 Issue 2 Pages 514-520
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Kinetics studies were performed on the non-enzymatic decomposition of glutathione (GSH) in anaerobic aqueous solution. The apparent first-order rate constants of decomposition were determined in the pH range of 1-12, ionic strength 0.5, at 60° and the pH profile was constructed from the buffer-free rate constants and pH values. The decomposition of GSH took place rapidly in acidic solution, but the rate constant showed a minimum in the pH range of 5-6, a maximum at pH 8.5 and a plateau in the pH range of 10-12. The effects of ionic strength and dielectric constant were almost negligible. The apparent activation energies of decomposition of GSH were all in the range of 19-21 kcal/mol. The rate constants of decomposition were related to the mole fractions of ionic species of GSH, and could be expressed as a function of hydrogen ion activity at arbitrary pH.
    Download PDF (645K)
  • MASAYOSHI ARUGA, SHOJI AWAZU, MANABU HANANO
    1980 Volume 28 Issue 2 Pages 521-528
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Kinetic studies on the peptide bond (γ-glutamyl bond) cleavage and desulfurization of glutathione (GSH) in anaerobic aqueous solution were performed in the range of pH 1-12 at 60°. The desulfurization of GSH in neutral solution was very slow, but the rate increased markedly with increase of pH. In the range of pH 10-12, the pH of the reaction solution had little effect on the desulfurization rate. The cleavage rate of γ-glutamyl bond showed a minimum in the range of pH 5-6, a maximum at pH 8.5 and a plateau in the range of pH 10-12. The effects of ionic strength and dielectric constant on both γ-glutamyl bond cleavage and desulfurization were almost negligible. The apparent activation energy for cleavage of the γ-glutamyl bond was 19-21 kcal/mol and that for desulfurization was about 21 kcal/mol. The apparent first-order rate constants of both reactions were related to the mole fractions of ionic species of GSH and could be expressed as a function of hydrogen ion activity at arbitrary pH. In the range of pH 6.5-10, the pH profile of the logarithmic cleavage rate constant of the γ-glutamyl bond was bell-shaped ; it is likely that cleavage of the γ-glutamyl bond is accelerated, at least in the range of pH 7.5-10, by intramolecular catalysis involving the NH2 and SH groups of GSH.
    Download PDF (724K)
  • SHOZO KAMIYA, MASAYUKI TANNO
    1980 Volume 28 Issue 2 Pages 529-534
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    4-(3-Cyano-1-triazeno) pyridazine 1-oxides (IIIa-d), 3, 6-dimethyl-4-(3-cyano-1-triazeno) pyridazine 2-oxide and 6-(3-cyano-1-triazeno) tetrazolo [1, 5-b] pyridazine were synthesized by treating the corresponding azides with potassium cyanide, followed by acidification with hydrochloric acid. 4-(3-Cyano-1-triazeno) pyridazine 1-oxide potassium salts (IIa, IIb) gave pyridazine 1-oxides (Va, Vb) on treatment with a mixture of hydrochloric acid and ethanol, and on treatment with hydrochloric acid alone, they gave a mixture of 4-chloropyridazine 1-oxides (VIa, VIb) and 4-aminopyridazine 1-oxides (VIIa, VIIb). The reaction of the cyanotriazene potassium salt (IIa) and hydroxylamine hydrochloride gave 4-[3-(N2-hydroxyamidino)-1-triazeno] pyridazine 1-oxide (X), and a similar compound (XIIa) was also prepared. An azourethan, 3, 6-dimethyl-4-[3-(ethoxycarbonyl)-1-triazeno] pyridazine 2-oxide (XIIb) and an azourea, 6-(3-carbamoyl-1-triazeno) tetrazolo [1, 5-b] pyridazine (XV) were synthesized from the corresponding cyanotriazene potassium salts. Arylazopyridazine 1-oxides (VIIIa-g) and 2-oxides (IXa, b) were synthesized by treating the cyanotriazene potassium salts (IIa-d, XI) with a mixture of 2-naphthol or diphenylamine and ethanol containing hydrochloric acid.
    Download PDF (699K)
  • YOSHIRO OZEKI, YUKIHISA KURONO, TOSHIHISA YOTSUYANAGI, KEN IKEDA
    1980 Volume 28 Issue 2 Pages 535-540
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Using the reaction rate of p-nitrophenyl acetate (NPA) with human serum albumin (HSA) as an index, the inhibition modes and binding sites of various anionic drugs were investigated in pH 7.4 phosphate buffer at 25°. Eleven anionic drugs including two specific fluorescence probes were chosen on the basis of Sudlow's classification of the drug binding sites, Site I and Site II (Sudlow et al., Mol. Pharmacol., 12, 1052 (1976)). The types of inhibition caused by these drugs suggested that at least three binding sites were present on HSA. The primarily reactive site (R site) of HSA corresponded to Sudlow's Site II, which corresponds to tyrosine-411 in Brown's HSA sequence. Site I was not involved in the esterase activity of HSA, which required tryptophan-214 (U site). In addition to the R site, HSA had a secondarily reactive site (T site) for NPA, which could also bind the drugs. The binding affinities of the drugs to the R site were estimated and some structural features of the R site were deduced.
    Download PDF (750K)
  • RIICHI TAWA, SHINICHI SHIMIZU, SHINGO HIROSE
    1980 Volume 28 Issue 2 Pages 541-545
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Some tertiary amines were determined by measurement of the absorbance of their π-complex (I) with tetracyanoethylene (TCNE) as a function of time. The initial absorbance values of I formed (corresponding to the added concentrations of tertiary amines) were calculated by a numerical method with a pseudo first-order rate equation for the decomposition of I using the rate constant and the absorbance values as a function of time. Calibration plots gave good straight lines for each tertiary amine. The average errors of determination were from -1.23% to +3.03% and the relative standard deviations were less than 1.82% (n=3). The recovery of each amine was from 98.0% to 103.0%. For mixtures of primary, secondary and tertiary amines, the interference of primary and secondary amines could be eliminated by addition of acetic anhydride before the reaction with TCNE. The average errors of determination were from -2.96% to +5.79% and the relative standard deviations were less than 1.15% (n=3). The recovery of each tertiary amine was from 97.0% to 105.8%. The presented method is accurate, rapid and highly sensitive for many tertiary amines.
    Download PDF (544K)
  • SHIGERU OHMIYA, HIROTAKA OTOMASU, JOJU HAGINIWA, ISAMU MURAKOSHI
    1980 Volume 28 Issue 2 Pages 546-551
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Nine known lupin alkaloids, (+)-matrine, (+)-matrine N-oxide, (+)-sophoranol, (+)-sophoranol N-oxide, (-)-sophoridine (1), (-)-cytisine, (-)-N-methylcytisine, (-)-N-formylcytisine, (-)-baptifoline, and (-)-sophoridine N-oxide (2a), were isolated from the aerial parts of Euchresta japonica BENTH., in addition to two alkaloids, (+)-5, 17-dehydromatrine N-oxide and (-)-12-cytisineacetic acid, reported previously. 2a was one of the two diastereoisomeric sophoridine N-oxides with different configurations of N-oxide nitrogen. Seasonal variations of the alkaloid contents in the aerial and ground parts of E. japonica were also examined. The N-oxidation of sophoridine with m-chloroperbenzoic acid gave two N-oxides in a ratio of ca. 1 : 5. The minor N-oxide was identical with 2a. The major N-oxide (2b) was characterized as the other diastereomer of sophoridine N-oxide by MS spectroscopy and deoxygenation to sophoridine. The stereochemical assignments of the two isomers were carried out by analysis of their 1H- and 13C-NMR spectra.
    Download PDF (736K)
  • SOJI SUGAI, KAZUO TOMITA
    1980 Volume 28 Issue 2 Pages 552-557
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    3-Mercaptoisoxazoles (XI) were synthesized from 3-allylsulfinylisoxazoles (II). The reaction of II with triphenylphosphine in the presence of acetic anhydride gave the corresponding 3-acetylthioisoxazoles (VII). The thioacetates (VII) were converted into 3-mercaptoisoxazoles (XI) by reaction with silver nitrate and subsequent treatment of the resulting silver salts (X) with hydrogen sulfide. The thiols (XI) were oxidized in air to give the corresponding disulfides (XIV). In alkaline solution, 3-mercapto-5-phenylisoxazole (XIb) was decomposed into benzoylacetonitrile (IX) and sulfur.
    Download PDF (636K)
  • KUNIO HIROI, SHUKO SATO, KAZUHIDE MATSUO
    1980 Volume 28 Issue 2 Pages 558-566
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The oxidative cleavage of 2-alkyl-2-(1-hydroxyalkyl)-1, 3-dithianes (1-5) and 2-(1-hydroxyethyl)-2-methyl-1, 3-dithiolane (11) with Pb (OAc)4 gave 3-alkyl-1, 4-dithiepan-2-ones (7a-e) and 1, 4-dithian-2-one (12), respectively, in fairly good yields. Analogously, the treatment of 2-alkylidene-1, 3-dithianes (13a-e) with Pb (OAc)4 resulted in a ring expansion to give 7a-e, whereas the reaction of 2-benzylidene-1, 3-dithiane (13f) with Pb (OAc)4 did not produce the expected ring expansion product, instead giving 2-(α-acetoxybenzylidene)-1, 3-dithiane (8).
    Download PDF (962K)
  • YOSHIHISA OKAMOTO, KANAME TAKAGI, TAKEO UEDA
    1980 Volume 28 Issue 2 Pages 567-570
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The reaction of 4-amino-1H-1, 5-benzodiazepine-3-carbonitrile hydrochloride (1) with hydroxylamine gave a ring-opened compound, 3-amino-3-(o-aminoanilino)-2-cyano-2-propenaloxime (5). Compound 5 was converted to 5-(o-aminoanilino)-4-cyanoisoxazole (6) on treatment with hydrochloric acid. Compound 1 was also reacted with methoxyamine to give 2-(1'-cyano-2'-methoxyaminovinyl) benzimidazole (8). On the other hand, hydrolysis of 1 at pH 1 gave 2-cyanomethylbenzimidazole hydrochloride (12).
    Download PDF (511K)
  • TAKAO SAKAMOTO, TAKEJI SAKASAI, HIROSHI YAMANAKA
    1980 Volume 28 Issue 2 Pages 571-577
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Pyrimidine derivatives in which both the 2- and 4-positions are free exhibited site selectivity in their reactions with radicals generated in redox systems. Namely, 6-phenyl-(I), 6-methylpyrimidine (XV), and 5, 6, 7, 8-tetrahydroquinazoline (XVII) reacted with radicals such as RCO, R2NCO, EtOCO, and CH2OH to give predominantly the 4-substituted products. Except in the reaction of I with the N, N-dimethylcarbamoyl radical, the formation of the corresponding 2-substituted isomers was not observed.
    Download PDF (865K)
  • MAKOTO OKAMURA, MANABU HANANO, SHOJI AWAZU
    1980 Volume 28 Issue 2 Pages 578-584
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Two morphologically different crystals, column-shaped and ramified, were sorted under a microscope from a recrystallized batch of 5-nitroacetylsalicylic acid. Their degradation rates in the solid state were quantitatively studied and the progress of decomposition was observed under a polarizing microscope. Plots of the ratio decomposed against the reaction time were sigmoidal under all reaction conditions, and Avrami's kinetic equation was applicable to this reaction below 40% degradation. The reactivity of the ramified crystals was greater than that of the column-shaped crystals under all experimental conditions. Although the parameter B of Avrami's equation, which is assumed theoretically to be proportional to the product of the number of latent reaction nuclei and the voluminal growth rate varied by a factor of 104 at maximum under various experimental conditions, the ratio of B between ramified and column-shaped crystals remained fairly constant under all the conditions tested, and was about five. The results of observation of the decomposing crystals under a polarizing microscope agreed well with Avrami's theory of the formation and growth of reaction nuclei.
    Download PDF (1387K)
  • KIKUO TEJIMA, SHOJI OZEKI
    1980 Volume 28 Issue 2 Pages 585-593
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The interrelationships between fatty acid and probe binding to bovine serum albumin (BSA) were investigated by a fluorescence method, using 1-anilinonaphthalene-8-sulfonate (ANS) and N-phenyl-1-naphthylamine (NPN) as fluorescent probes. The fatty acidinduced quenching of probe fluorescence is probably due to the displacement of probes from BSA binding sites. It appears that NPN and ANS are bound to BSA at the same sites. Various molar ratios of capric acid, lauric acid and palmitic acid were added to defatted BSA in the presence of ANS, and the association constants of the fatty acids were calculated. Measurements of ANS fluorescence (λmax 465nm) and the fluorescence (λmax 340nm) of the ANS-BSA complex as well as the circular dichroism spectra of BSA in the presence of fatty acids in pH 7.4 phosphate buffer indicated that fatty acids induce different conformational states of the albumin molecule. One of the strongest binding sites of long-chain fatty acids may be the site which is labeled by TNBS (2, 4, 6-trinitrobenzenesulfonic acid), located in the carboxyl-terminal half of the molecule, which does not contain tryptophan residues. It is suggested that the mechanism of inhibition of the probe-BSA binding by capric acid may involve competition for the same binding sites. In contrast, the decreases in probe binding with BSA in the presence of lauric acid. palmitic acid or stearic acid are due not only to a displacement of the probes by fatty acids from the probebinding sites but also to a simultaneous conformational change of the probe-binding sites caused by the interaction of the probes and fatty acids.
    Download PDF (963K)
  • HIROFUMI KUSHIMA, YUKIHIRO SHOYAMA, ITSUO NISHIOKA
    1980 Volume 28 Issue 2 Pages 594-598
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A quantitative analysis procedure for tetrahydrocannabinolic acid, cannabichromenic acid, cannabidiolic acid and cannabigerolic acid monomethyl ether was established by gas chromatography and by a combined gas chromatography-preparative thin-layer chromatography method using cholestane as an internal standard. The variations of cannabinoid contents with leaf-age, season and sex were investigated in three kinds of "physiological varieties" of Cannabis sativa L., the Mixican, the Minamioshihara No. 1 and the CBDA strain.
    Download PDF (532K)
  • HIROSHI MIWA, MAGOBEI YAMAMOTO, TSUTOMU MOMOSE
    1980 Volume 28 Issue 2 Pages 599-605
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Aqueous or aqueous ethanolic samples of aliphatic and aromatic acids were coupled with 2-nitrophenylhydrazine hydrochloride to give acid hydrazides, which showed intense violet colors in an alkaline medium. Dicyclohexylcarbodiimide was used as a coupling agent, and a small amount of pyridine markedly enhanced the reaction. The brown blank solution was decolorized to a faint yellow by heating the alkaline solution at 60°. The reaction conditions were investigated in detail with acetic, stearic, benzoic and citric acids, and sensitive methods for the detection and determination of carboxylic acids were established. The limits of detection and the absorption maxima of the alkaline colors obtained with many carboxylic acids were determined. Most amino acids, together with hydroxybenzoic acids and aminobenzoic acids, were detected with reduced sensitivity by this method. Salicylic acid was shown by high-performance liquid chromatography and gas chromatography-mass fragmentography to form mainly its ehyl ester in the coupling reaction, whereas benzoic acid, which could be sensitively detected, formed only a small amount of the ester. p-Aminobenzoic acid also formed a large amount of its ethyl ester.
    Download PDF (690K)
  • MASUO MORISAKI, MAYUMI SHIBATA, CARMENZA DUQUE, NOBUTAKA IMAMURA, NOBU ...
    1980 Volume 28 Issue 2 Pages 606-611
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The Li2CuCl4-catalyzed Grignard reaction of the p-toluenesulfonates of 3β-tetrahydropyranyloxybisnorchol-5-en-22-ol (9) and 3β-tetrahydropyranyloxychol-5-en-24-ol (3) afforded several cholesterol analogs with a modified side chain (10). The reaction of 9 with 3-methyl-2-butenylmagnesium chloride gave a mixture of desmosterol (11) and its allylic isomer (12) in a ratio of 2 : 1. 25-Methylcholesterol (7) was prepared from the cholanal derivative (5) by reaction with t-butylmagnesium bromide followed by deoxygenation of the resulting 24-carbinol 6.
    Download PDF (651K)
  • TERUAKI OHKURA, HARUHISA UEDA, NAOKI NAMBU, TSUNEJI NAGAI
    1980 Volume 28 Issue 2 Pages 612-618
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    As an approach to an understanding of the combined effects of drugs, the competitive adsorption of aminopyrine (AM) and barbituric acid derivatives (BA) by carbon black (CB) from aqueous solution was investigated in relation to the formation of molecular compounds between AM and BA. The formation of molecular compounds of AM with various BA compounds was confirmed by differential scanning calorimetry (DSC), powder X-ray diffractometry and infrared absorption spectrometry. All adsorption isotherms of AM and BA were in good accordance with Langmuir's equation. From the results of perturbation experiments (20°to 40°transition), it was considered that the mechnaisms of adsorption of AM and BA by CB were "chemisorption" and "physical adsorption, " respectively. In competitive adsorption, the decrease of BA adsorbed was larger for compounds that formed molecular compounds with AM than for those that did not. It was considered that adsorption was influenced by an interaction between AM and BA molecules at the surface of CB ; the greatest effect in competitive adsorption was seen with AM and barbital, which also formed a molecular compound most readily.
    Download PDF (679K)
  • ISAMU MIYATA, HIROSHI KISHIMOTO
    1980 Volume 28 Issue 2 Pages 619-626
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The viscosities of cholesteryl myristate (ChM), cholesteryl palmitate (ChP) and cholesteryl oleate (ChO) were measured with a cone-plate viscometer at various shear rates with continuous heating. After a rapid decrease near the smectic-cholesteric transition temperature, the viscosities decreased gradually in the cholesteric and isotropic phases, with a peak at the cholesteric-isotropic transition. The viscosity peak temperature agrees reasonably well with the values obtained by differential scanning calorimetry (DSC) and polarized light microscopy. From the shear rate vs. shearing stress relations at various temperatures, it was concluded that the flow in the isotropic phase was Newtonian, while the flow in the cholesteric phase was non-Newtonian. With respect to the cholestericisotropic transition temperature, the effect of alkanols on the viscosity of ChM was the same as that on the DSC curve. However, the viscosity vs. temperature curve showed a stronger transition peak than the DSC curve, in spite of the addition of alkanols. The change of ChM viscosity under stress was also followed at various temperatures ; the "static viscosity" was obtained by extrapolation of the viscosity to zero time and the "kinetic viscosity" from the steady-state value under stress. Comparison between the viscosity measured with continuous heating and the kinetic viscosity confirmed the validity of the former method, especially in the isotropic and cholesteric-isotropic transition states.
    Download PDF (800K)
  • KATSUMI TAKATA, OSAMU FUJISHITA, HIROMASA KAMEYA, TOSHINORI TSURUTA, M ...
    1980 Volume 28 Issue 2 Pages 627-631
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A microbial barrier faucet with dual pipes, having a multiple capillary nozzle attached to the internal pipe and with clean air blown between the internal and external pipes, was manufactured and its performance was tested. This faucet prevented airborne microbial contamination when clean air was blown at a velocity of 4.5 m/sec. Water splashes occurring at initial water flow through the faucet were easily subject to contamination by direct contact with the contaminated external pipe end, and should be removed before subsequent use of the faucet. The principle of this faucet should be applicable for supplying not only distilled water but also sterile pharmaceutical solutions. Further application to sterile bottling in the beverage industry may also be possible.
    Download PDF (610K)
  • TADAHIRO TAKEDA, SEIICHI TAKABE, YUKIO OGIHARA
    1980 Volume 28 Issue 2 Pages 632-634
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Methyl 2, 3-(15), 2, 4-(11) and 3, 4-di-O-methyl-α-D-fucopyranosides (8) were prepared from methyl α-D-fucopyranoside via methyl 3, 4-O-isopropylidene-α-D-fucopyranoside. These substances were required in connection with methylation analysis of saponins containing D-fucose residues. Compounds 8 and 11 had not previously been synthesized at the outset of the present work.
    Download PDF (460K)
  • TAKAHARU MIZUTANI, POFENG KUO
    1980 Volume 28 Issue 2 Pages 635-638
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The effects of a bovine parotid hypocalcemic substance (pure parotin) on several components in rabbit serum were investigated after injection into the aural vein. The concentration of inorganic phosphate was significantly increased at 4-6 hr after the injection of the hypocalcemic protein, and that of serum calcium was decreased. The hypocalcemic protein also significantly reduced the concentration of hydroxyproline at 4 hr after the injection. These results suggest that the protein might act on bone tissues, directly or via mediators. The concentrations of sodium, potassium and magnesium ions, and of alkaline phosphatase in the serum did not change significantly after the injection.
    Download PDF (531K)
  • KIYOSHIGE OCHI, YOSHIHIKO WATANABE, KIYOSHI OKUI, MINORU SHINDO
    1980 Volume 28 Issue 2 Pages 638-641
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Sulfation of sucrose with pyridine-sulfur trioxide was carried out in dimethylformamide and pyridine, and the degree of sulfation of the sodium salt of sucrose sulfate thus obtained was estimated from the ratio of sulfur to carbon content (S/C). The sulfates prepared with 9-15 molar equivalents of pyridine-sulfur trioxide in dimethylformamide and those prepared with 5 and 9 equivalents in pyridine were identified as sucrose octasulfate. Potassium, cesium, rubidium, and ammonium salts of sucrose octasulfate were obtained as crystals.
    Download PDF (546K)
  • TOSHIO KINOSHITA, JUNICHIRO MURAYAMA, KAZUYO MURAYAMA, AKIO TSUJI
    1980 Volume 28 Issue 2 Pages 641-644
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A fluorometric assay method for proteins based on the reaction of peptide groups is described. The N-H groups of peptides are chlorinated with sodium hypochlorite, and the resulting N-chloropeptides are allowed to react with thiamine, giving fluorescent thiochrome. The present method is highly sensitive and is applicable to 1-10μg of protein. Little variability of fluorescence intensity was observed among proteins.
    Download PDF (524K)
  • JUICHI SATO, EIJI OWADA, KEIJI ITO, TOSHIRO MURATA
    1980 Volume 28 Issue 2 Pages 645-647
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    In the previous paper, acetazolamide (AZA) was reported to shorten the hexobaribital sleeping time, while prolonging the norhexobarbital (NHB) sleeping time in mice. Since AZA was found to prolong the sleeping time induced with metharbital (MET) and mephobarbital (MB) as well as barbital (BA) and phenobarbital (PHB) in this study, N-methylation of barbiturates does not appear to be a critical factor in the action of AZA. The induction times of the five barbiturates, including NHB, were shortened by AZA pretreatment. At the moment of awakening from sleep, the brain level of BA was the same as that without the AZA pretreatment, which suggests unchanged brain sensitivity to the barbiturate. At that time, however, the plasma level was reduced in the case of AZA pretreatment. Similar results were obtained with PHB, MET, and MB. Thus, the prolongation of barbiturate sleep and shortening of the induction time are probably due to a decrease in the rate of removal of the barbiturates from the brain due to the reduction of cerebrospinal fluid flow and/or the inhibition of a system for acid transport out of the brain.
    Download PDF (458K)
  • KIMIO HIGASHIYAMA, HIROTAKA OTOMASU
    1980 Volume 28 Issue 2 Pages 648-651
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Spiro[oxindole-3, 4'-(4'H-pyran)]compounds (IIIa-j), were prepared by the Michael reaction of 3-cyanomethyleneoxindoles (Ia, b) with active methylene compounds. Similarly, the reactions of (Ib, c) with cyclic active methylene compounds, e.g. cyclohexane-1, 3-dione, 3-methyl-1-phenyl-5-pyrazolone and barbituric acid, afforded the corresponding spiro compounds (V-VII) containing the condensed pyran system.
    Download PDF (484K)
  • YOSHINOBU NAKAI, SHINICHIRO NAKAJIMA, KEIJI YAMAMOTO, KATSUHIDE TERADA ...
    1980 Volume 28 Issue 2 Pages 652-656
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    The infrared (IR) spectra of aspirin, benzoic acid, and salicylic acid, which exhibit intermolecular hydrogen bonding in their crystals, were measured in dilute CCl4 solution. Two carboxyl carbonyl stretching bands were observed in these spectra, due to monomeric and dimeric species. IR spectral changes of carbonyl stretching bands attributed to hydrogen bonding between monomeric medicinals and 1-butanol were also measured. This hydrogen-bonded band was always observed at higher frequency than the intermolecular hydrogen-bonded band (dimer), and at lower frequency than the band of the monomeric medicinal. The shifts of the carbonyl stretching bands measured in the ground mixtures were compared with those of the solution spectra. In the ground mixtures, it was confirmed that the medicinal molecules were dispersed monomolecularly and were hydrogen-bonded with hydroxyl groups of cellulose or β-cyclodextrin molecules. The new band observed at 1748 cm-1 in the β-cyclodextrin dispersed system of aspirin was considered to be due to overlapping of (i) the band of the acetoxyl group of aspirin hydrogen-bonded with hydroxyl groups of β-cyclodextrin, and (ii) the free carboxyl band of aspirin.
    Download PDF (539K)
  • MASAAKI MORI, NADAO KINOSHITA, HIDETOSHI YOSHIMURA
    1980 Volume 28 Issue 2 Pages 657-659
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Benzo (α) pyren-3-yl hydrogen sulfate was synthesized from chlorosulfonic acid and benzo (α) pyren-3-ol, and crystallized as the potassium salt. It was hydrolyzed by aryl sulfatase. The UV, IR, and fluorescence spectra of the potassium salt are presented.
    Download PDF (398K)
  • KAZUMASA HIRAUCHI, AKIKO FUJISHITA, TAMEYUKI AMANO
    1980 Volume 28 Issue 2 Pages 660-663
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    A phosphorimetric method for the determination of α-[(tert-butylamino)methyl]-3, 4-dihydroxybenzyl alcohol 3, 4-di (p-toluate) (I) was established. The method is based on the native phosphorescence of I in an alkaline extract after separation on a thin-layer chromatographic plate, and can be applied satisfactorily to the determination of I in the concentration range of 0.02-0.5μg per spot.
    Download PDF (420K)
  • RYOHEI KIMURA, TOSHIRO MURATA
    1980 Volume 28 Issue 2 Pages 664-666
    Published: February 25, 1980
    Released on J-STAGE: March 31, 2008
    JOURNAL FREE ACCESS
    Theanine (L-N-ethylglutamine, a constituent of Japanese green tea), L-glutamine or caffeine did not affect adenosine 3', 5'-monophosphate (cAMP) formation in cerebral cortex slices from rats. In addition, theanine and L-glutamine did not show any effect on cAMP formation in the presence of caffeine. Theanine did show an inhibitory effect on norepinephrine-stimulated cAMP formation in the presence or absence of caffeine. Theanine also showed an inhibitory effect on histamine-stimulated cAMP formation in the absence of caffeine, but not in the presence of caffeine. It appears that theanine does not antagonize caffeine in cAMP metabolism, but does inhibit the stimulation of cAMP formation by norepinephrine.
    Download PDF (417K)
feedback
Top