NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1975, Issue 4
Displaying 1-33 of 33 articles from this issue
  • Chikakiyo Nagata, Shigeyuki Tanaka
    1975 Volume 1975 Issue 4 Pages 579-582
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The effect of substituent R in N, N-dimethylamides on the rotational barrier about the carbon-nitrogen bond is studied by carbon-13 nmr spectroscopy. The free energies of activation (Fc ≠) at coalescence temperatures (Tc) for the hindered rotation of these amides are obtained by using the Eyring's absolute rate equation. The Fc ≠ values in 20 mol% CCl4 solutions are > 18.8, 18.7, 14.9 and 14.7kcal/mol for N, N-dimethylformamide (DMF), N, N-dimethylacetamide (DMA), N, N-dimethylnicotinamide (DMN) and N, N-dimethylbenzamide (DMB), reSpectively (Table 2).
    Although the values of Fc ≠ for DMN and DMB are smaller than that for DMA, the π -bond orders calculated by CNDO/2 method are 0.4812, 0.4224, 0.4328 and 0.4323 for DMF, DMA, DMN and DMB, respectively (Table 4). This result agrees with the fact that Fc ≠ for DMF is larger than those for DMN and DMB, but it does not explain the larger value of Fc ≠ for DMA than those for DMN and DMB.
    The differences in total energies (E) between the stable conformation and transition state in which N-methyl group is rotated 90° about carbon-nitrogen bond are calculated by the CNDO/2 method. They are obtained as 4.77 and 1.60 kcal/mol for DMA and DMB, respectively (Table 5), in agreement with the experimental results.
    Download PDF (1017K)
  • Akira Yanagase, Nobuyuki Sugita, Kiyoshi Kudo, Yoshimasa Takezaki
    1975 Volume 1975 Issue 4 Pages 583-590
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Synthesis of methyl glycolate from methylene sulfate and carbon-monoxide has been studied kinetically under pressure in 1, 2-dichloroethane (EDC) solution. The rate of methyl glycolate formation is of the first order with respect to carbon-monoxide pressure and nearly of the first order to the concentration of methylene sulfate. Under the reaction conditions studied, dimeric methylene sulfate decomposes to its monomer, formaldehyde and SO3 in equilibrium.
    Based on these results, the following mechanism has been proposed: methylene sulfate monomer reacts with carbon monoxide to form a complex which is converted into methyl glycolate by the action of methanol outside of the reaction system.

    [CH2OSO3] + CO(d) → k [CH2(CO)OSO3] [CH2(CO)OSO3] + 2CH3OH → CH2(OH)COOCH3 + CH3OSO3H
    The derived rate equation is:

    H: Henry constant of CO, Pco: CO pressure (kg, cm-2), L0: Initial charge of MS (mol)/solvent (mol), G: GAM (mol) Initial charge of MS (mol),
    which is reasonably in agreement with the experimental results. The equilibrium constants, rate constants, Henry constant in EDC and the overall activation energy have been determined to be K1=1.22 × 10-2, K2=5.7 × 10-3, k=6.05 min-1, H CO ED=6.8 × 10-4 kg-1, cm2 and Ea=28.4kcal, mol-1, respectively.
    In the detailed experiment by changing dielectric constant of medium, the observed rate constant has been found to fit the Kirkwood modified relation, which holds for the reaction through dipole-dipole interaction.
    Download PDF (1756K)
  • Keizi Hashimoto, Shozi Watanabe, Kimio Tarama
    1975 Volume 1975 Issue 4 Pages 591-595
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Three kinds, [W], [M], [S], of molybdenum oxide chemically interacted with alumina were investigated in order to obtain the correlation between the chemical states and the activities. The oxide [W] is soluble in water. The oxide [M] is insoluble in water but so1uble in aqueous ammonia. The oxide [S] is insoluble in a concentrated ammonium hydroxide. The amount of oxide [W] was little in the catalyst which molybdite-content was less than 6 wt% and increased sharply with the further increase of molybdite-content. The oxide [M] showed a constant value in the amount beyond 8wt% of molybdite-content and existed in fine dispersion on the surface of alumina. The oxide [S] increased with molybdite-content.
    For ethylene polymerization the activity of the catalyst which molybdite-content was less than several weight percents was not observed. Beyond several weight percents of molybditecontent, the activity increased in proportion to the amount of the oxide [W]. However, the catalyst which contains Only oxide [W] did not have any polymerization activity. These facts indicates that the oxide [W] contributs to the polymerization in the coexistence of the other states.
    Download PDF (1412K)
  • Hiroshi Sakaguchi, Etsuo Niki, Yoshio Kamiya
    1975 Volume 1975 Issue 4 Pages 596-601
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The conversion rate and yields of products in the decomposition of t-butyl hydroperoxide over boron trioxide under vacuum have been studied over the temperature range of 200 to 310 C, in order to clarify the mechanism of boric acid catalyzed autoxidation of tertiary aliphatic hydrocarbons. A novel procedure has been devised where t-butyl hydroperoxide was transfered from trap to trap through boron trioxide under vacuum (Fig.1). Under these conditions, primary products were transfered immediately from the catalyst surface to a trap, so that the secondary reactions of primary products could be suppressed and the boron trioxide was kept at its initial state throughout the reaction. Boron trioxide was quite effective in the decomposition of t-butyl hydroperoxide, and the major products were isobutene, water, acetone, and methanol (Fig.2). Isobutene was supposed to be produced by the thermal decomposition of t-butyl metaborate formed directly from the interaction of t-butyl hydroperoxide and boron trioxide. Therefore, little t-butanol was formed under the reaction conditions employed. Boron trioxide was also effective as a Lewis acid to give equimolar amount of acetone and methanol. The thermal decomposition of t-butyl hydroperoxide was completely suppressed by the boron trioxide below 250 C (Table 1, Fig.2). At higher temperature than 280 C, however, both the rate of esterification and acid-catalyzed decomposition decreased rapidly and that of thermal decomposition increased. A comprehensive mechanism has been proposed.
    Download PDF (1796K)
  • Yoshiharu Matsuda, Toshio Fujisawa, Shuzo Fujikawa, Yoshio Takasu, Yos ...
    1975 Volume 1975 Issue 4 Pages 602-606
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Cyanide ion was oxidized in alkaline solution by bubbling a mixture of ozone and oxygen. At 25 C, the apparent initial rate of decomposition of cyanide changed according to the equation:
    v=k[CN-]0.5 [O3]
    An activation energy of 10.6kcal/mol was obtained for the oxidation of cyanide at pH 13.5from measurements of the reaction rates in the temperature range 10 - 30 C. The effect of Cu2+, Zn2+ and Sn2+ on the rate of decomposition of cyanide were studied and Cu2+ was revealed to accelerate the decomposition greatly.
    Download PDF (1411K)
  • Susumu Okazaki, Kazuo Watanabe, Harumi Arakawa
    1975 Volume 1975 Issue 4 Pages 607-610
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The catalytic effect of TiO2 for the dehydrofluorination of polyhydrofluoroethylene i. e. polyvinylidenefluoride (PVDF) and polyvinylfiuoride (PVF), was investigated by heating TiO2with these polymers in the stream of N2.
    The effect of calcinating temperature of TiO2 on the catalytic activity was examined and the order of calcinating temperature for the catalytic activity was confirmed as follows,
    500° C > 600° C > 400° C > 300° C > 800° C
    (for PVDF dehydrofiuorination)
    and
    500° C > 600° C > 400° C ≈ 700° C > 300° C
    (for PVF dehydrefiuorination).
    These orders coincided almost with the acid amounts (H0 ≤ 4.8) of TiO2. In addition to previous results with respect to the promoting effects of various metal fluorides, the effectiveness of the solid acidity for the decomposition of polyhydrofluoroethylene was thus revealed.
    The induction period of dehydrofluorination reaction was reduced for both polymers as the acid amount of TiO2 increased. Consequently, the acid site seemed to work as a special point for the promotion of the first cleavage of C-F of these polyhydrofluoroethylene.
    Download PDF (1119K)
  • Setsuko Wada, Takuji Ito
    1975 Volume 1975 Issue 4 Pages 611-616
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Ion association constants (kSn 2+, SO4 2-) of tin(II) sulfate which were expressed in terms of concentration, e, g., moles per liter, were determined by measuring the emfs of the cell, Sn-Hg | Sn(ClO4)2, HClO4, NaClO4, H2SO4 || NaClO4 || NaCl || SCE, at various ionic strength(I=0.037 - 1.0) in aqueous solutions at different temperatures between 15.0 C and 45.0 ° C
    kSn 2+, SO4 2- decreases with the increase of ionic strength. The value of log kSn 2+, SO4 2-, in which kSn 2+, SO4 2- was the thermodynamic equilibrium constant for the ion association of tin(II) sulfate at 25.0 C, was calculated from kSn 2+, SO4 2- obtained at each ionic strength. The average value of log kSn 2+, SO4 2- corrected for the ion pair tin(II)bisulfate(Sn 2+, HSO4 -) is 3.24 ± 0.07.
    The standard Gibbs energy, the standard enthalpy and the standard entropy for the reaction Sn 2+ + SO4 2- Sn2+, SO4 2- were calculated from the temperature-dependence of the ion associtation constants. The values obtained are as follows, 4G°=-4.42 kcal/mol, H°=4.05kcal/mol, S°=28.4cal/deg, mol.
    Download PDF (1599K)
  • Tsuyoshi Nakajima, Koichiro Nakanishi, Nobuatsu Watanabe
    1975 Volume 1975 Issue 4 Pages 617-621
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The dispersion of metallic lithium in various molten salts was studied at constant partial pressures of nitogen or oxygen gas. The increase of dispersed lithium amounts in the presence of nitrogen gas was more rapid than that in the presence of oxygen gas. A large amounts of lithium dispersed in LiCl and LiCl-KCl eutectic mixture where the solubilities of lithium oxide or nitride were almost equal to those in liquid lithium. On the other hand, the amounts of dispersed lithium were small in LiF-LiCl, LiF-KF and NaCl-CaCl2 systems, where the solubilities of oxide or nitride were much different from those in lithium.
    Download PDF (1081K)
  • Ken-iti Hukuo, Yasuo Hikichi
    1975 Volume 1975 Issue 4 Pages 622-626
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The precipitates from the mixed solution of cerium salt and orthophosphoric acid or phosphate were examined by X-ray diffraction analysis, chemical analysis and electron microscopic observation.
    Crystalline hexagonal cerium phosphate was detected for the precipitates obtained from the solution of the pH between 1 and 3 (Fig.1 and 3). The minute crystals showed prismatic habit (Fig.4). Amorphous precipitate was obtained when the pH of solution was above 3. At the pH below 1, hexagonal cerium phosphate coprecipitated with monoclinic form at 90 ° C, but precipitation was not proceeded at room temperature (Fig.2). The chemical formula of hexagonal and amorphous cerium phosphates was given as CePO4, 0.5H2O.
    Download PDF (1422K)
  • Takehiko Takahashi, Hideaki Itoh, Shoichiro Noguchi
    1975 Volume 1975 Issue 4 Pages 627-631
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Zirconium nitride (ZrN) was obtained on graphite or silicon substrate by chemical vapor deposition technique using a gas mixture of ZrCl4, N2, H2 and Ar. The correlation between growth condition and crystal morphology was investigated in detail in the temperature range of 950° C to 1300° C. The maximum weight of ZrN deposited on graphite was obtained at 1200° C. The optimum gas composition for the deposition of ZrN was found > 1.0 and > 40mol % in terms of N2/ZrCl4 mole ratio and H2 concentration, respectively (H2 flow rate: 4ml/sec), where the deposition rate was about 2 mg/hr, cm2. Preferred orientation to the direction of (100) was confirmed for the films obtained at higher temperature than 1250° C, especially deposited on the (100) silicon substrate. In addition to the growth of homogeneously coated layer, ZrN whiskers up to the length of 500μ m were grown at higher temperatures. The growth direction of the whisker was confirmed to be < 111> or < 110>
    Download PDF (1895K)
  • Yshinori Kidani, Reiko Saito, Hisashi Koike
    1975 Volume 1975 Issue 4 Pages 632-635
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Copper (II) complexes of six aminocyclohexanecarboxylic acid isomers have been prepared and investigated by means of magnetic susceptibilities, visible diffuse-reflectance spectra and infrared spectra. These complexes have been classified into two groups: Group I, Cu (cis-2 ACC')2, 2H2O, and Cu(trans-2 ACC')2, 2H2O, Group II, Cu(cis-3 ACC')2, Cu(trans-3 ACC')2, Cu (cis-4 ACC')2, and Cu (trans-4 ACC')2, where ACC' stands for aminocyclohexanecarboxylate anion. Octahedral complexes of group I are blue, and magnetically slight subnormal. Square planar complexes of group II, synthesized in DMF, are violet and magnetically subnormal. The properties of group II complexes are similar to copper (II) complexes of 4-aininobutyric acid and 5-aminovaleric acid (ABA, and AVA).
    Infrared spectral analysis suggests that the conformations of the ligands in regard to the cyclohexane ring are not changed by complexation. In view of the above properties, group II complexes might have a structure in which the amino and the carboxyl groups were simultaneously coordinated to the different copper atoms.
    Download PDF (1172K)
  • Yoshitaka Masuda, Kazuo Koya, Seizo Misumi
    1975 Volume 1975 Issue 4 Pages 636-641
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The μ -Thio-bis(1, 2-diphenylethanedithion)-copper(I), [Cu2S6C4(C6H5)4], has been prepared. From elemental analysis, molecular weight determination by the use of "Vapor pressure osmometer", result of IR absorption spectra and analysis of TGA, it was concluded that [Cu2S6C4(C6H5)4] is a sulfur bridged compound with structure(II), complexed by ligand having dithioketonic character.
    It was found that its complex is diamagnetic from the measurement of its magnetic moment.
    When the copper(I) complex reacts with iodine, two molecules of copper iodide and sulfur compound, not containing copper, are formed.
    Plot of C=C stretching frequency of the copper (I) complex (1445 cm-1) vs. C=C character (%) calculated by Schrauzer suggests that this complex has about 75% double bond character. And the peak at 328 cm-1 in the far-infrared region is probably due to the Cu-S stretching frequency.
    The spectral properties of this copper(I) complex was discussed and compared them with those of the related bis(benzene-1, 2-dithiolato)copper (II) complexes with planar structure.
    The obtained AC polarogram and voltammograms show the presence of [Cu2IS6C4(C6H5)4]+, [Cu2IS6C4(C6H5)4]-, and [Cu2IS6C4(C6H5)4]2-. The half-wave potentials of reduction and oxidation waves obtained at rotating platinum electrode are shown at -0.823 and +0.383 V vs. SCE, respectively. It was found from both electronic spectra and polarography that this copper (I) complex is the charge transfer complex.
    The electrode reaction of anodic wave can be expressed as follows:
    [Cu2IS6C4(C6H5)4] → [Cu2IS6C4(C6H5)4]+ + e-
    Download PDF (1490K)
  • Kyoichi Sekiguchi, Tsuneo Takishima
    1975 Volume 1975 Issue 4 Pages 642-646
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The ultraviolet spectra of river water were studied to characterize them reference to nitrate ion concentration.
    1) Ultraviolet spectra of hundreds of river water samples were classified into four types: A, B, C and D. The classification was made on the basis of wave-length of absorption maximum and the analysis of inflection pqint at 200± a few nm.
    2) The relationship between the absorbance at 210nm and nitrate ion concentration in river waters was experimentally studied for 99 river water samples. It was concluded that for samples with E210/E220 1.75 and E210 2.06, E210 gave an accurate measure for nitrate concentration.
    3) Coefficients of correlation the nitrate ion concentrations and E210 valuse were calculated for samples with four different ranges of E210/E220 ratios, namely 1.43, 1.75, 1.95 and 2.15. The result revealed that the nitrate ion concentration of river water samples with E210/E220 1.75 can be successfully calculated from E210 values.
    4) For samples fulfilling the criterion that E210/E220 1.75, E210 was calculated as E210=0.505x + 0.096, where x=concentration of NO3-N(ppm). r=0.980 and P < 0.001 were obtained. Thus, direct ultraviolet spectrometric determination of nitrate ion in river water samples was possible when the samples have E210/E220 1.75.
    Download PDF (1526K)
  • Yuji Yanagida, Masaaki Matsubara, Masakatsu Nomura, Shoichi Kikkawa
    1975 Volume 1975 Issue 4 Pages 647-651
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Bromine addition to bicyclo[2.2.2]oct-5-ene-2-endo-3-exo-dicarbonitrile [1] was investigated. Two adducts, [2a] and [2b] were isolated and the addition was concluded to proceed via trans addition mechanism on the basis of NMR coupling constants analysis. The measurement of dipole moments also revealed that the observed values of [2a] and [2b] agreed with those calculated by assuming trans configuration.
    In the previous investigation on the bromine addition to trans-5, 6-dicyano-2-norbornene [3]in acetic acid, exo-cis adduct (24% of total adducts) and only one kind of trans adduct (76%)were obtained.
    In the bromination of [1], on the contrary, exo-cis adduct was not obtained, but two trans adducts were obtained. These results might be attributed to the difference of internal strain between [1] and [3].
    Download PDF (1227K)
  • Yuji Yanagida, Toshio Yokota, Masakatsu Nomura, Shoichi Kikkawa
    1975 Volume 1975 Issue 4 Pages 652-656
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    1, 3-Dipolar cycloaddition reaction of diazomethane to various 5, 6-disubstituted 2-norbornenes were investigated. All the adducts obtained were 1-pyrazoline derivatives. In the reaction of trans-5, 6-dicyano(or dicarbomethoxy)-2-norbornene, two isomers were isolated and the structures were determined by NMR analysis.
    Kinetic study in diethyl ether showed that the rate increased in the order, exo-cis > trans > endo-cis. And electron withdrawing substituents increased the rates. The field effect was suggested to explain these results.
    Download PDF (1249K)
  • Yuji Yanagida, Hideo Shigesato, Masakatsu Nomura, Shoichi Kikkawa
    1975 Volume 1975 Issue 4 Pages 657-660
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Epoxidation of various 5, 6-disubstituted 2-norbornenes with perbenzoic acid was investigated. NMR spectral analysis showed that all the products are 2, 3-exo-cis epoxides. The reaction rates of epoxidation for the three 5, 6-disubstituted isomers increased in the following order, 5, 6-exo-cis < 5, 6-tra < 5, 6-endo-cis isomer. As the steric hindrance of 5, 6-endo-substituents will not exert any influence upon such an 2, 3-exo-cis addition reaction, the result of this ikinetic study suggests that the substituent effect is predominantly polar effect. Thus, it was concluded that the polar effect of 5, 6-exo-substituent is greater than that of 5, 6-ende-one in this reaction.
    These results might be rationalized in terms of F effect, which was tentatively estimated by the Kirkwood-Westheimer model.
    Download PDF (1074K)
  • Kiyonori Shinoda, Shuzo Anzai
    1975 Volume 1975 Issue 4 Pages 661-664
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Co-pyrolysis of 1, 1, 2-trichloroethane (TCE) and methanol on various activated alumina catalysts was carried out at 2700C. As the base amount of catalysts increased, the conversion of TCE and the ratio of 1, 1-dichloroethylene to 1, 2-dichloroethylene were found to increase.
    The results of pyrolysis of TCE with several different alcohols and ethers on activated alumina at 280° C showed the decreasing reactivity in the following order: Et20 > n-PrOH > EtOH > MeOH > n-Bu20 > n-BuOH > t-BuOH > s-BuOH > i-BuOH > i-Pr20 > i-PrOH.
    The higher homologs of the normal alcohol series increased the conversion of TCE. The nucleophi1icity of the alcohol adsorbed on the basic site of activated alumina might be appreciably enhanced by the inductive effect of alkyl group. However, the reactivity of higher alcohols was decreased owing to the steric hindrance of the bulky alkyl group.
    On the other hand, it was found that branched alcohols were readily dehydrated to form olefins and the conversion of TCE was lower than that of normal alcohols. The co-pyrolysis of TCE and ethers was analogous to that of alcohols.
    Download PDF (852K)
  • Hiroshi Kimoto, Hiroshige Muramatsu, Kan Inukai
    1975 Volume 1975 Issue 4 Pages 665-671
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Seven 1, 2, 2-trifluoro-1-trifluoromethylindanes [3] were synthesized by the one-step procedure of cycloaddition of alkylbenzenes [1] (toluene, ethylbenzene, cumene, p-methoxytoluene, p-xylene, p-chlorotoluene) to hexafluoropropene in the presence of di-t-butyl peroxide. The dehydrofuorination and photo-bromination of [3] gave the corresponding 1, 2-difluoro-1-trifluoromethylindenes [7] and 3-bromo-1, 2, 2-trifiuoro-1-trifluoromethylindane [9] together with 3, 3-dibromo-1, 2, 2-trifluoro-1-trifluoromethylindane [10], respectively.
    Download PDF (1774K)
  • Yoshihiro Muramoto, Hideyuki Asakura
    1975 Volume 1975 Issue 4 Pages 672-677
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The rates of desulfonation of cresolsulfonic acids were determined and the separation method of pure m-cresol from the mixture of m- and P-cresol was improved by means of the sulfonation-desulfonation process.
    For the rate determination, the desulfonations were carried out by heating 2-methylphenol-6-[1] and -4-sulfonic acids [2], 3-methylphenol-6-[3] and -4-sulfonic acids [4], and 4-methylphenol-2-sulfonic acid [5] with 47% hydrobromic acid, and the reactions were found to obey pseudo-first-order kinetics.
    The desulfonation rate constants of [1] - [5] at 100° C were 10.1, O.635, 15.1, 26.7, 5.99 × 10-3 min-1 respectively.
    m- And p-cresols were sulfonated with dilute sulfuric acid at low temperature, and m-isomer was sulfonated 3 to 5 times as fast as P-isomer was.
    The improved method for separation of almost pure m-cresol from the m-/p-mixture was as follows: the mixture was partially sulfonated with 90% sulfuric acid at 20° C for 2 hr, then the temperature was gradually raised to 100° C and kept at that temperature for 3 hr. The quantity of sulfuric acid was controlled to give 65% spent acid by sulfonating all of the m-cresol and one-fifth of the p-cresol in the mixture. The unsulfonated cresol was extraced with benzene and the sulfonated m-cresol was des with super-heated steam at 120 - 135° C. The yield of m-cresol by this process was 80 - 90% and its purity was 97 - 99%.
    Download PDF (1374K)
  • Yoichi Koiwa, Masayuki Teranishi, Keizo Tatsukawa, Yasuo Fujimoto
    1975 Volume 1975 Issue 4 Pages 678-682
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The mechanism of γ -methyl L-glutamate N-carboxy anhydride(MLG, NCA) formation by use of phosgene gas was studied in a solvent system of monochlorobenzene-dioxan(15: 1 in volume ratio) at 90° C.
    At the first stage of the reaction, the reaction temperature was observed to rise significantly. This exothermic phenomenon was also observed in just the same pattern when hydrogen chloride gas was added at the rate twice as that of phosgene. Therefore, this phenomenon could be attributed to neutralization reaction reaction between MLG and hydrogen chloride. By the time when temperature lowered, one third of MLG was converted to NCA and two thirds of MLG was converted into its hydrochloride, that is, hydrogen chloride formed when one third of MLG was converted to NCA reacted quantitatively with two thirds of MLG. This first stage of reaction was very fast, and the rate of the reaction should be determined by the rate of addition of phosgene.
    After this stage, reaction proceeded between MLG, HCl and phosgene. This reaction was very slow compared with the first stage, and apparently of first order with respect to MLG, HCl.
    Download PDF (1324K)
  • Tatsuo Sekiguchi, Motoo Tanaka, Makoto Kobayashi
    1975 Volume 1975 Issue 4 Pages 683-686
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The process for producing Cu-Pc-sulfonic acid from Cu-Pc-disulfate by Baking process was studied. The results obtained are as follows:
    1)A mixture of Cu-Pc-monosulfonic acid and Cu-Pc-disulfonic acid was obtained when Cu-Pc-disulfate was baked in the powdered form at 190 - 300° C(A method).
    2)Besides Cu-Pc-monosulfonic acid and Cu-Pc-disulfonic acid, a small amount of Cu-Pc-trisulforlic acid was also obtained when Cu-Pc was dispersed in two equivalent amounts of dilute sulfuric acid and baked directly with agitation after water was vaporized(B method).
    3) In A method, a content of Cu-Pc-monosulfonic acid in the product increased as a ratio of combined H2SO4 per Cu-Pc in Cu-Pc-sulfate became less than 2mol/mol.
    4) The components of products were separated by their adsorption chromatography withsilica gel.
    5) It was made clear that the sulfonic group of Cu-Pc-monosulfonic acid obtained by Baking Process is introduced into the 4-position of the benzene ring of Cu-Pc.
    Download PDF (1123K)
  • Saburo Nakanishi, Keiji Ito
    1975 Volume 1975 Issue 4 Pages 687-691
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Photo-oxidation of anthracene on activated alumina was studied. The oxidizing activity of alumina produced by calcination of aluminum hydroxide varied with its calcinating temperature in such a way that two maximal activities were obtained: the maximal activities were obtained when alumina was calcinated at 300-350° C and at 620° C. The oxidizing activity of alumina was also proportional to the acidity of its Lewis acid site. The reactivity of anthracene on alumina was little affected by the wavelength of the light illuminated. Identification of the reaction products and measurements of the diffuse reflection spectrum of anthracene during the reaction revealed that (1) a primary product of the reaction was 9, 10-epidioxYanthracene which was appreciably stable on the surface of alumina, (2) this product was converted into cis-9, 10-dihydroxyanthracene and anthraquinone, and (3) the former was oxidized to anthraquinone and the latter was oxidized to mono-hydroxyanthraquinone or alizarin.
    From the above results, it was suggested that the reaction proceeds through a chargetransfer complex between anthracene and alumina.
    Download PDF (1470K)
  • Yukuo Katayama, Takuji Hosoi, Gen Takeya
    1975 Volume 1975 Issue 4 Pages 692-696
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In the last paper, a new approach was proposed to analyze structural factors of pitch materials such as number of naphthene rings and ring compactness factor by using model compounds.
    This new analytical method was, in this paper, applied to structural analysis of real pitches as coal tar pitches and various petroleum heavy oil ends. The analytical result was compared with those from presently known methods (the Brown-Ladner method and the Krevelen method)(Table 3). The new method was confirmed to be quite satisfactory and furthermore certified to make up the defect of the Brown-Ladner and the Krevelen methods.
    Download PDF (1168K)
  • Takao Kishimoto, Yoshiharu Matsubara
    1975 Volume 1975 Issue 4 Pages 697-700
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of alicyclic saturated alcohols (cyclopentanol [1], cyclohexanol [2] and cyclooctanol [3]) and aliphatic saturated alcohols (n-hexanol [4], 2-ethylhexanol [5], n-octanol [6] and 4-methyl-2-pentanol [7]) with paraformaldehyde (PFA) in the presence of acetic anhydride and sodium acetate was investigated. Three new products ([1 a]-[7 a], 1 b-[7 b] and [1 c]-[7 c]) were isolated by distillation or gas chromatography.
    Their structures were assigned on the basis of their physical constants, IR, NMR and mass spectra as alkyl acetoacetate [1 a]-[7 a], alkyl dioxymethyleneacetoacetate [1 b]-[7 b] and alkyl trioxymethyleneacetoacetat, e [1 c]-[7 c].
    Download PDF (1010K)
  • Takao Kishimoto, Yoshiharu Matsubara
    1975 Volume 1975 Issue 4 Pages 701-704
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of acyclic and monocyclic terpene alcohols (nerol [1], geraniol [2], citronellol [3], dihydrocitronellol [4], perillyl alcohol [5] and homoperillyl alcohol [6]) with paraformaldehyde (PFA) was carried out in the presence of acetic anhydridei and sodium acetate.
    The structures of the products were elucidated to be terpenyl acetoacetate [1 a]-[6 a], terpenyl dioxymethyleneacetoactate [1 b]-[6 b]. and terpenyl trioxymethyleneacesoacetate [1c]-[6c] on the basis of their physiscal data.
    The yields of these adducts were 32 - 50% on the paraformaldehyqe used(Table 1).
    Download PDF (945K)
  • Yoshiyasu Tanaka, Tadashi Nakamura
    1975 Volume 1975 Issue 4 Pages 705-710
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The pyrolysis of pentoses was carried out at 190-300° C by using DTA and TG.
    DTA curve consisted of three peaks: the first, endothermic melting point, the second, endothermic decomposition and the third, exothermic peak beginning at about 300° C. The beginning of decomposition of pentoses occurred at 194° C. Differences between the decomposition temperatures and the melting points of pentoses were 100° C for ribose, 60° C for lyxose, 50° C for xylose, and 40° C for arabinose.
    The weight loss curve obtained by TG showed two types: the one, for D-arabinose and D-ribose and the other, for D-xylose and D-lyxose. TG curves showed two steps of weight loss at 240° C.
    Ten ml aqueous solution was prepared from pyrolysis product of 30.0 mg pentose for each 10° C rise in temperature between 190° C and 300° C. The aqueous solution of pyrolysis (250° C)products of D-xylose showed the lowest value of transmittance at 520 nm. The solutions of pyrolysis (270° C) products of D-lyxose, D-ribose, and D-arabinose showed the lowest value at 52O nm. This showed the decomposition of pentoses was completed at these temperatures. The solutions, treated with Schales reagent showed the transmittance at 420 nm. which relates with the reducing power of them. The aqueous solutions of pyrolysis (220° C) products of pentoses showed the highest transmittance. The transmittance of these solutions markedly decreased at 220-230° C (from 90% to 30% for D-xylose). The solutions in which pentose could be detected contained finally 1.48% D-xylose, O.94% D-ribose, 1.40% D-arabinose and O.71% D-lyxose in the solution of pyrolysis (230° C), (240° C), (240° C) and (240° C) products, respectively.
    Download PDF (1454K)
  • Yasuji Fujita, Shin-ichi Fujita, Takeyoshi Hasegawa
    1975 Volume 1975 Issue 4 Pages 711-713
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    (±)-Linalool was administered to various living plants, such as Fatsia japonica, Iris Pseudacorus and Quercus phillyraeoides. And the formation of α -terpineol, limonene, geraniol, nerol, trans- and cis-linalooloxides was recognized.
    These results have a great significance for the biogenesis of mono-terpenoids in plants, as linalool can be considered as a precursor.
    It is of special interest to note that we have obtained the optically pure specimen of (-)α -terpineol with the optical rotation of[α ]22D -110.3° in the case of Fatsia japonica.
    Download PDF (837K)
  • Hiroaki Omichi, Tetsuo Miyakoshi, Shojiro Saito
    1975 Volume 1975 Issue 4 Pages 714-719
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The Prins reaction of methylallyl chloride(MC) with formaldehyde(HCHO) was reinvestigated and by-products were shown to be α -isobutylene chlorohydrin [1], cis-3-chloro-4-hydroxy-4methyltetrahydropyran[3], trans-3-chloro-4-hydroxy-4-methyltetrahydropyran[4] and 4chloromethyl-4-hydroxytetrahydropyran [5] (Table 1, Fig.2). Treatment of [4] and [5]with NaOH gave corresponding epoxides [16] and [17] respectively (Table 2).
    4-Chloromethyl-4-methyl-1, 3-dioxane (1, 3-dioxane) [2] was prepared as a main product in 66% yield when a molar ratio of HCHO/MC is 2 in the presence of 50% H2SO4(Table 4, Fig.5, 6).
    The reaction in acetic acid was also carried out and by-products were shown to be those above described and their acetates, as well as dehydration products of 4-chloro-3-methyl-1, 3butandiol monoacetate (monoacetate) [15] (Table 3, Fig.4).
    Either 1, 3-dioxane [2] or monoacetate [15] was obtained as a main product according to a molar ratio of HCHO/MC used(Table 5). No 4-chloro-3-rnethy1-1, 3-butandiol diacetate (diacetate), which has been reported previously, was obtained.
    Download PDF (1324K)
  • Masakuni Yoshihara, Tsunehito Eda, Yoshio Matsubara, Tochihisa Maeshim ...
    1975 Volume 1975 Issue 4 Pages 720-723
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Copolymerization of p-substituted derivatives [1], [2], and [3] (M2) of 1-phenyl-2-methyl-1, 2, 3, 3, 6-tetrahydro-3, 6-pyridazindiones with styrene (St) and methyl methacrylate (MMA) have been carried out in dimethylformamide (DMF) by thermal, photo, or radical initiation. Homopolymer of pyridazindione was not obtained despite the change in reaction condition.
    It was considered that the charge transfer complex might affect the reactivity of St system. Pyridazindione was found to have poor copolymerizability with MMA. A good linear relationship was observed between log 1/r1 and Hammett's σ values for any copolymerization system.
    Low reactivity of pyridazindiene was discussed in terms of the effects due to ring-strain and conjugation.
    Download PDF (901K)
  • Yuji Minoura, Masaru Sakai, Yasou Hirai
    1975 Volume 1975 Issue 4 Pages 724-727
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It was previously found that the accelerating effect on the polymerization of methyl methacrylate (MMA) in the presence of various cyclic ethers [i. e., 3, 3-bis(chloromethyl)oxetane (BCMO), epichlorohydrin (ECH) and etc. ] was due to the decrease in the termination rate constant (kt) of propagating chains resulted from the increase in viscosity of the polymerization system and assumedly to the increase in the propagation rate constant (kp) of MMA in the BCMO system.
    In the present investigation in order to see the change in propagation rate of MMA in BCMO, the copolymerization of styrene[M1]-MMA[M2] was investigated in it and the kp and kt were mesured by the rotating sector method. Values of r1 and r2 derived from Fineman and Ross method were O.52 and O.50 in benzene, and O.30 and O.61 in BCMO, respectively. The values of kp and kt obtained by the above sector method in BCMO and in benzene are as follows:
    kp(l/mol, sec) 2kt(l/mol, sec)
    186 8.6 × 106 in BCMO
    144 12.1 × 106 in Benzene
    Download PDF (1096K)
  • Keizo Makuuchi, Masahahu Asano, Toshihiko Abe
    1975 Volume 1975 Issue 4 Pages 728-732
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Poly(vinylidene fluoride) (PVdF), dissolved in. N, N'-dimethylformamide, N, N'-dimethylacetamide, 1-methyl-2-pyrrolidinone, and dimethyl sulfoxide, has been irradiated with Co-60 γ -rays.
    Although PVdF is crosslinked in the solid state, this crosslinking was not observed in the solution. The number-average molecular weight of PVdF decreased with increasing dose. The G value for main-chain scission G (S), was independent of the kind of solvent (Table 1), polymer concentration (Table 2), and the initial molecular weight distribution, it was 0.3 O.05. This value was smaller than the G (S) for solid PVdF. The molecular weight distribution varied from initial bread distribution to final random one as the dose increased (Figs.1 and 2). The results were interpreted in terms of energy and charge transfers from PVdF to solvent. The PVdF itradiated in the solvents except dimethyl sulfoxide acquires color. Its UV spectra are different from those of PVdF irradiated in the solid state (Fig.3). This coloration is caused by dehydrofluorination catalyzed by amines formed by radiolysis of these solvents.
    Download PDF (1479K)
  • Tadashi Yurugi, Hiro Iwata, Michio Okuma
    1975 Volume 1975 Issue 4 Pages 733-737
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The early stage of the pyrolysis of cellulose up to 400° C under nitrogen has been studied in terms of differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), infrared spectra and X-ray diffraction. Thermal degradation behaviors of highly crystalline cellulose (acid-hydrolysis residue) having lattice types I and II (C. C. I, C. C. II) and amorphous (non-crystalline) cellulose prepared by saponification of cellulose triacetate in non-aqueous medium were compared with one another.
    It has been found that both highly crystalline and amorphous cellulose show no structural change such as melting or recrystallization before thermal degradation.
    Furthermore, it has been observed that amorphous cellulose decomposes more readily at lower temperature (140-300° C), than crystalline cellulose which shows no appreciable change below 300° C and decomposes abruptly at 300-350° C. The decomposition temperature of C. C. II has been found to be slightly higher than that of C. C. I which is attributed to the difference of stability between the lattice structures of two types.
    Download PDF (1136K)
  • Fujio Takahashi, Tatsuo Ogasa, Shigeru Tanaka
    1975 Volume 1975 Issue 4 Pages 738-740
    Published: April 10, 1975
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    For the purpose to utilize keratin as a functional polymer, the reduced keratin was prepared by the reaction of human hair with sodium sulfide and thioglycolate. The molecular states of the reduced keratin in solution have been studied by viscometry as well as ORD spectrophotometry and its precipitates by the enzyme-probe method.
    The molecular weight of the reduced keratin was estimated to be about 5000, and the amount of SH group was 0.53 m-equivalent per g of protein. The reduced keratin was insoluble in aqueous solution whose pH lies within the range of 2 - 5.2. It was plausible that the reduced keratin formed a temporary network by the molecular interaction and contained 8 - 16% α -helices. It seemed that upon heating at 60 - 75° C the volume of molecules diminished and the molecular interaction increased. This molecular interaction might have influence on the enzyme molecule surrounded with the precipitates and interfere with the denaturation of enzyme.
    Download PDF (734K)
feedback
Top