NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1979, Issue 7
Displaying 1-29 of 29 articles from this issue
  • Shigehito DEKI, Yukio NAKATA, Yukio KANAJI
    1979 Volume 1979 Issue 7 Pages 805-810
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    For a study of interactions between ions and those between ions and water in concentrated electrolyte solutions of water and an electrolyte, the chemical behaviors of cation, anion and water in Mn(NO3)2-H2O systems were investigated with infrared, laser Raman, near infrared and visible spectroscopies for the binary systems from the dilute solution to such a concentrated solution as 9.6 mol/l.
    An analysis of Raman bands at 720-750 cm-1 with a curve resolver showed that interactions between NO3- and Mn2+changed with increasing concentration and that the transition of the free coordinate into unidentate and bidentate coordinations took place at 3.0 mol/l and 6.0 mol/l respectively. Effect of the concentration of the electrolyte on NIR spectra for the combination bands of -H2O indicated that the breakdown of the structure of the bulk water increased with increasing concentration of the electrolyte and that the amount of coordinate water increased with the increase of the concentration. Results of visible spectroscopy suggested that the covalency in the interaction of Mn2+-NO3- increased in concentrations higher than 6.0 mol/l.
    These results suggest that Mn(NO3)2-H2O systems have interactions depending on co ncentration, 0-3 mol/l, 3-6 mol/l and higher concentrations than 6 mol/l.
    Download PDF (306K)
  • Takaaki HATTORI, Katsumi KANEKO, Tatsuo ISHIKAWA, Katsuya INOUYE
    1979 Volume 1979 Issue 7 Pages 811-816
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Temperature programmed desorption (TPD) spectra of NO adsorbed on α-, β-, and γFeOOH have been determined in the temperature range 80-460°C (Figs.2, 3, and 4). The chemical analyses showed that the desorbed gas is NO. Spectra showed separate peaks at 100, about 200, and about 300°C for each FeOOH and adsorption states corresponding to each peak for each FeOOH were discussed in view of surface structures of FeOOH crystals and areas of peaks (Figs.5, 6, and 7). The high NO adsorption value found for the calcinated °C-FeOOH at 500°C (cf. this Journal, 1979, 423) was explained by the desorption result.
    Download PDF (311K)
  • Hiroshi ISHIKAWA, Eiichi ISHII, Itsuki UEHARA, Masanori NAKANE
    1979 Volume 1979 Issue 7 Pages 817-824
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    As a fundamental study on the thermochemical hydrogen production from water, bromination of Fe3O4 with gaseous HBr has been investigated.
    First, experiments were carried out in a fixed b ed reactor and the amount of evolved Br2was measured with time. The experiment showed that Fe3O4 was converted to FeBr2 according to the following successive reactions:
    Fe3O4 + 2 HBr → FeBr2 + Fe2O3 + H2O (1)
    Fe2O3 + 6 HBr → 2 FeBr2 + 3H2O + Br2 (2)
    Fe3O4 + 8 HBr → 3 FeBr2 + 4H2O + Br2
    The optimum reaction temperatures were 200-250°C from thermodynamical points of view.
    Second, the kinetics of these successive reactions was studied with powdered Fe3O4 at t emperatures of 200-400°C using a thermobalance. These two reactions were found to proceed as follows: at the initial stage, the reactant-product interface of the reaction (1) moved toward Fe3O4particles prior to the interface of the sequent reaction (2); this rate decreased as the reaction proceeded and was outrun by that of the interface of the reaction (2) when the conversion rate of Fe3O4, x, became ca.0.4. After this, both reactions (1) and (2) proceeded on the same interface, until x became 0.8. The reaction (1) was expressed by the "parabolic law equation" when x was less than 0.4, and the activation energy of the reaction was 1.2 kcal/mol. Therefore, the rate of HBr gas diffusion to the interface through pores in the particles would be the rate determing. As the reaction (2) was expressed by the "contracting cube equation" when x was less than 0.7 and its activation energy was 0.7 kcal/mol, it was concluded that chemical reactions at the interface would be the rate determing.
    Download PDF (443K)
  • Nobuo TAKAMIYA, Takeaki IWATSUKI, Toshie NAGAI, Sukenaga MURAI
    1979 Volume 1979 Issue 7 Pages 825-829
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Hydrodesulfurization of thiols over various poly (metal phthalocyanine) catalysts (Poly-M-Pc)was studied by a continuous flow system at an atmospheric pressure. The reaction products were hydrogen sulfide and hydrocarbon (1:1). Other reaction did not occur. In the hydrodesulfurization of ethanethiol, the percentage of yield of hydrogen sulfide depended on the central metals. The catalytic activity was found to decrease in the order, Pt, Pd, Cr<Ni, Co<Fe, Cu. The mixed type catalyst of Poly-Co-Ni-Pc was more active in the yield of hydrogen sulfide than the single type catalyst of Poly-Co-Pc or Poly-Ni-Pc. So insertion of two kinds of metal ions to the enlarged conjugated system of the catalyst gave special effect on the yield of hydrogen sulfide. When compared the catalytic activity of Poly-Pd-Pc with that of Pd(silica), the former showed the constant activity, but the latter showed the degradation of activity. In the kinetic study over Poly-Pd-Pc catalyst, the order of the reaction was found to be zero with respect to hydrogen, and 0.5 with respect to ethanethiol. The observed activation energy was 9.5 kcal/mol in the temperature range of 350-400°C. In the hydrodesulfurization of C4H10S isomers over Poly-Pd-Pc catalyst, (CH3)3CSH was most acti ve but (C2H5)2S almost inactive.
    Download PDF (233K)
  • Ikuo ABE, Katsumi HAYASHI, Mutsuo KITAGAWA, Toshihiro URAHATA
    1979 Volume 1979 Issue 7 Pages 830-835
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The characteristic adsorption curves on activated carbon with different pore size distributions have been determined in gas and liquid phases. The adsorption isotherms of benzene in gas phase were measured and the parameters in the Dubinin-Astakhov equation were calculated. The linear relationship exists between nG and EG. The adsorption volume (WG) was calculated from the limiting vapor adsorption at P/P0→1. The linear relationships exist between WOG/WG and nG or EG. In liquid phase, the adsorption isotherms of anisole from aque ous solution were measured and correlated by the following equation:
    where ε s1 =R T In Cs/C, Cs is solubility, and C an equilibrium concentration. The linear relationship exist between nL and EL, and between WOG/WG and nL or EL. The values of n and E in liquid phase are smaller than those in gas phase. The magnitude of WOL is similar to that of WOG. This suggests that in liquid phase, pores, having such large diameters as occur capilary condensation, are also utilized for an adsorption.
    Download PDF (355K)
  • Akihiko NI-INA, Kaoru FUJIMOTO, Taiseki KUNUGI
    1979 Volume 1979 Issue 7 Pages 836-841
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Behaviors of carbon monoxide and oxygen adsorbed on supported palladium catalysts were studied using a pulse reaction technique. The effects of sodium halides added to the catalyst were also studied. Adsorbed carbon monoxide reacted rapidly with oxygen in the gas phase in the temperature range from 50°C to 300°C. The amount of oxygen adsorbed was almost constant between 100 and 200°C and it corresponded to the monolayer coverage of palladium surface with atomic oxygen. A remarkable increase in adsorption of oxygen on palladium was observed above 250°C. This increase is probably attributed to the formation of palladium oxide. Although the amount of carbon monoxide adsorbed was not affected by the addition of any other sodium halide than sodium iodide in the catalyst, the reactivity of carbon monoxide decreased in the following order: Pd (216) →Pd-NaCl(100) →Pd-NaBr (31) →Pd-NaI (2). The amount of oxygen adsorbed on those catalysts at 200°C decreased in the following order: Pd (271) →Pd-NaCl(100) →Pd-NaBr (27) → Pd-NaI (4). Relative values in the product of the value of the reactivity of carbon monoxide adsorbed and the amount of oxygen adsorbed on each catalyst agreed well with the relative reactivities of the catalyst. It suggests that the role of halide ion as an inhibitor on CO oxidation is based on the poisoning of catalytic sites which are capable of adsorbing dissociated oxygen and active carbon monoxide.
    Download PDF (370K)
  • Yasuyoshi KATO, Akira IGARASHI, Yoshisada OGINO
    1979 Volume 1979 Issue 7 Pages 842-847
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The mechanism of activation of water in the steam dealkylation of in-xylene was investigated by the use of a pulse reactor. In the activation of H2O over Rh/Al2O8 catalyst the formation of active surface OH groups occurs. It was demonstrated that the mechanism of activation of H2O over Rh/MWO 4 (M=M g, Ca and Sr) differs from that over Rh/Al2O3 catalyst. Namely, the activation over Rh/MgWO4 brings about the cleavage of H2O molecule into gaseous H2 and surface oxygen ato m. The high reactivity of this surface oxygen on Rh/MgWO4 catalyst seemed to be responsible for the high steam-dealkylating activity of this catalyst. Thus, the mechanism of activation of H2O in the steam dealkylation of alkylbenzene over Rh/MWO4 catalyst is well expressed by the following two equations:
    1/dx MWO4+ R → 1/dx MO + 1/dx W O3-dx RO (a)
    1/dx MO + 1/dx WO3-dx + H2O → 1/dx MWO4 + H2 (b)
    R: Intermediate species for demethylattion on rhodium.
    Download PDF (348K)
  • Kunishige NAITO, Shigeru TAKAGI, Hideto EBATA, Shinsuke TAKEI
    1979 Volume 1979 Issue 7 Pages 848-854
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The kinetic study of the oxidation of iron (II) ion in sulfuric acid solution by oxygen has been carried out using activated carbon as a catalyst. The oxidation rate was found to increase with an increase of the initial concentration of iron (II) ion and of an amount of the activated carbon catalyst, but to decrease with the increase of the initial concentration of iron (III) ion or sulfuric acid. It was newly observed that the oxidation rate decreased with a further increase of the initial concentration of iron(II) ion when the activated carbon with high catalytic activity was used. No linear relations were obtained between the oxidation rate and the amount of the catalyst. It was considered that such a result was caused by insufficient input of oxygen to the reaction system due to its slow dissolving rate in the solution. The main course of the oxidation process is estimated by the following surface reactions,
    It was concluded that the activated carbon could act not only as an electrode for cell reactions between iron (II) ion and oxygen and/or between iron (II) ion and hydrogen peroxide, but also a catalyst for decomposition of hydrogen peroxide. Experimental data indicate that the oxidation rate follows the rate equation,
    derived from the Langmuir-Hinshellwood mechanism postulating the forementioned surface reactions as a rate-determing step.
    Download PDF (406K)
  • Yasuaki Osumi, Hiroshi SUZUKI, Akihiko KATO, Masanori NAKANE, Yoshizo ...
    1979 Volume 1979 Issue 7 Pages 855-860
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Absorption-desorption characteristics of hydrogen have been investigated for titanium-cobalt alloys TiCo, Ti1-xAxCo and TiCo1-xAx. TiCo can easily be activated and reacts readily with hydrogen to form a hydride TiCoH, .4 (tetragonal structure, hydrogen content 1.3 wt%) under 30 atm hydrogen pressure and at 60-80°C. The standard enthalpy change in the hydride formation is given -13.8 kcal/mol H2, as determined from the dissociation isotherms (Fig.3)for the TiCo-H system and the dissociation pressure at 150°C yields 2.0 atm. A value of 10.6 kcal/mol H2 tames out as the apparent activation energy of hydrogen desorption. Th e hydrogen absorption-desorption cycles are repeated 30 times, but the variation in hydrogen absorption-desorption ability was hardly noticed.
    Both Ti0.05A0.05Co and TiCo0.95A0.05 (A: Cr, Cu, Fe, La, Mn, Ni and V) have been found to have the same cubic structure as TiCo. They react readily with hydrogen to form hydrides (hydrogen content 0.9-1.4 w t f) under 30 atm hydrogen pressure and below 200°C. TiCo0.95Mn0.05, and TiCo0.95Fe0.05 react readily with hydrogen to form hydrides of TiCo0.95Mn0.05H1.4and TiCo0.95Fe0.05H1.2(hydrogen content 1.3, 1.2 wt %). Dissociation pr essures of these hydrides at 150°C yield 3.1 and 3.4 atm, respectively, and the rate of hydrogen absorpt iondesorption becomes increased markedly in comparison with that of TiCo. The hydrides of TiCo0.95Mn0.05 and TiCo0.95Fe0.05h ave properties suitable as a hydrogen storage material.
    Download PDF (355K)
  • Masanori TAKENOUCHI, Tadaaki MORI, Takeaki IIDA, Takashi MITAMURA
    1979 Volume 1979 Issue 7 Pages 861-868
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Formation mechanism of titanium nitride (TiN) by the reaction of TiCl3 and N2 in the presence of Al has been investigated by means of thermogravimetry, X-ray diffractometry and EPMA-SEM.
    In a TiCl3-Al-N2 system, the reduction of TiCl3 by Al began at ca.350°C in preference to the disproportionation of TiCl3, and Ti was formed above ca.500°C by way of the formation of TiCl2. At this temperature Ti reacted immediately with N2 and with residual Al by competitive reaction to form TiN and A1Ti, respectively. TiN was formed by the reaction of AlTi and TiCl2 under N2 at 500°C, too. On the other hand, in a TiCl3-Al system AlTi was formed at 500°C and reacted with TiCl2 above 700°C to form Ti2Al. This shows that a portion of TiCl2 is reduced by Al in AlTi to Ti which has structural imperfections and is active against N2.
    In the TiCl3 -Al-N2 system, it was confirmed that TiN is formed above ca.500°C by nitridation of Ti formed by Al reduction of TiC13 and reaction of AlTi with TiCl2. Above 700°CTiN was obtained as a single phase. The crystallite size and the amount of TiN increased with increasing temperature. The product obtained by this reaction at 1000°C for 2 hr had the composition of TiN0.90and the lattice constant of a0=4.234A.
    Download PDF (488K)
  • Tsugio SATO, Seiyu MATANI, Taijiro OKABE
    1979 Volume 1979 Issue 7 Pages 869-878
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Department of Applied Chemistry, Faculty of Engineering, Tohoku University; Aramaki, Sendai-shi 9 80 Japan Analytical methods for the determination of nitrogen-sulfur compounds produced by the reaction of aqueous sodium sulfite solution with nitrogen dioxide were developed. The oxidation of sodium sulfite with nitrogen dioxide proceeded by bubbling nitrogen dioxide gas (630 to 2000ppm) at a flow rate of 2.5 /min through the aqueous sulfite solutions (0.240-0.380 mol/l), which ranged 25 to 70°C in temperature and pH 5.92 to 10.60. More than 99% of nitrogen dioxide was absorbed under all the conditions tested, producing various nitrogen-sulfur compounds such as NO, -, HON (SO3)22-, etc. The reaction products differed remarkably according to the reaction conditions applied. Main products in the pH range 6.60-10.6 at 25°C consisted of SO42-, S2O62-, NO2- and HON(SO3)22-, and those produced at 70°C in the pH range of 6.60-7.20 consisted of SO42-, S2O62- and HN(SO3)22-. Any appreciable amount of nitrat e was not produced in all the cases examined. These results indicated that the reaction between nitrogen dioxide and sulfite solution at the first stage should not be such that
    2 NO2 +H2O=NO2-+NO3-+2H+but proceeded in the following way:
    2 NO2+HSO3-+H2O=2NO2-+SO42-+3H+
    2NO2+2H SO3-=2NO2-+S2O62-+2H+Subsequent reaction of NO2- and HS08- led to the formation of HON (SO3)22-. The reactions above were competitive, so that the amount of S082- produced increased with increasing pH values and temperature. In acid media HON (SO3)22-, N(SO3)33- and HN(SO3)22- became oxidized by NO2- to form sulfate. Also, NO2- partially hydrolyzed to yield NO, NO2 and NO3-. Based on these observations an possible oxidation pathway of sulfite in solution by nitrogen dioxide was suggested.
    Download PDF (594K)
  • Soei SATO, Seiichiro HIGUCHI, Shigeyuki TANAKA
    1979 Volume 1979 Issue 7 Pages 879-884
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    For the direct analysis of alkylxanthogenates (RO-CSS-K+), a new method with laser Raman spectrometry was developed. The Raman spectra of eleven C1-C5 alkylxanthogenates with normal and branched chains in both solid states and solution were measured in the 1200-200cm-1 region, and their identification by use of each of the characteristic Raman l ines was attempted. The Raman spectra were measured with 488 nm line of Ar+ laser with 200 mW power. The Raman line due to vs(-CS2-) stretching vibrations and skeletal deformations of alkyl groups were mainly observed in the 700-200 cm-1 region. These Raman spectra exhibited characteristic patterns for the alkylxanthogenates. The Raman lines observed for solution samples were shown to be useful for the identification of isomers. A characteristic line table for the identification of each of the alkylxanthogenates was made up from the spectra. In addition, the characteristic lines was used to determine of alkylxanthogenates. The relation between the peak intensities and concentrations of the alkylxanthogenates was linear, when thiocyanate ion was used as an internal standard. It was shown that the simultaneous determination of some of the alkylxanthogenates is possible. This Raman spectrometric method was applied to the analysis of the practical samples used as the collectors in flotation, and satisfactory results were obtained. It was proved that this spectrometric method is useful for the qualitative and quantitative analysis of alkylxanthogenates.
    Download PDF (364K)
  • Kunio NAKANO, Hiroshi TADANO, Shigeyuki SUGIMOTO
    1979 Volume 1979 Issue 7 Pages 885-890
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It has already been reported that measuring the wave height of a two-step integral (h) gave better results for quantitative measurement of ESR signal than measuring the peak-to-peak height of the ordinary signal of derivative type (This journal, 1972, 2453). However, since signals which are recorded under different conditions cannot be compared with each other directly, this method is not simply applicable. In order to solve this problem, a method to normalize the measured values has now been examined.
    The integrator used was an analogue-type dev ice with two operational amplifiers, constructed by the authors. The experimental results were as follows. With regard to the conditions of integration, "h" could be normalized by multiplying it by CiR, and C, R2, the time constants of the first step and the second step of the integrator, and by dividing it by R4/(R8+R4) and R6/(R5+R6), the resistor ratio of each of the output attenuators. Moreover, "h" could a lso be normalized by dividing it by the square of the time of the integral period. Although it is generally- thought that the signal intensity can be norm alized by dividing it by the square root of microwave power (P0.5), the results, of the present experiments revealed difficulty in normalizing by dividing by Pk or log P. Therefore the microwave power should be kept constant during the measurements. Concerning other conditions of the ESR spectrometer, "h" could be normalized by dividing it by gain and modulation amplitude, but thepeak-to-peak height of a sharp signal could not be normalized when the modulations amplitude was large.
    Although the peak-to-peak height of a sharp signal was strongly affected by response time change even under constant microwave power, "h" was affected very little by crystal current and response time change when the microwave power was constant. Thus, the wave heights of a two-step integral measured even under different conditions can be compared with each other provided that microwave power is kept constant.
    Download PDF (332K)
  • Toru NOZAKI, Masatomi SAKAMOTO, Junji NAKAGI, Masao MIYAKE
    1979 Volume 1979 Issue 7 Pages 891-895
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A Polarographic Study of the 2-Aminobutanoate Complexes of Copper(II), Lead(II), Cadmium(II) and Manganese(II)Toru NOZAKI, Masatomi SAKAMOTO, Junji NAKAGI and Masao MIYAKE Department of Industrial Chemistry, Faculty of Engineering, Ehime University; Bunkyo-cho, Matsuyama-shi 790 Japan Polarographic behaviors of copper (II), lead (II), cadmium (II) and manganese (II) in 2aminobutanoate solutions were studied at ionic strength of 0.4 adjusted with sodium perchlorate. The reversible two-electron reduction waves were obtained in the pH ranges from 1.2 to 4.2and 6.1 to 11.9 for the copper system, from 1.2 to 6.8 and 11.3 to 13.1 for the lead system, from 1.3 to 11.0 for the cadmium system and from 5.7 to 9.7 for the manganese system. Metal complexes, MX+, MX2, MX3-, M(OH)X, M(OH)X2-, M(OH)2X-, M(OH)2X22- for copper, the species of the same types, M(OH)3X2- and M(OH)4x3- for lead, MX+, MX2, MX, M (OH) X, M(OH)x2- for cadmium, and MX+ for manganese were assumed and their stability constants were calculated by Schaap and DeFord-Humes' method. The logarithms of the overall stability constants at 25°C were 8.2, 12.0, 14.3, 11.2, 14.7, 14.9, 16.8 for the forementioned copper complexes, 4.6, 8.8, 12.5, 11.9, 15.1, 15.5, 17.4, 17.6, 18.5 for the lead complexes, 3.8, 6.3, 9.0, 8.2, 10.5 for the cadmium complexes, 3.1 for the manganese complex. The square-wave polarograms suitable for the determination of Ni and Co(II), Zn(II), Mn(II), Cu (II) and Cd(II), Pb(II) at pH 9.3, 12.3, 7.7, 8.9, 13.1, respectively, were obtained in the 2-aminobutanoate solutions of 4.0×103- mol/l. The lower limits for the determination of the metals were 0.6, 0.4, 0.4, 0.2, 0.02, 0.05 and 0.05 ppm, respectively.
    Download PDF (269K)
  • Kazuo ITOH, Yoshiaki FURUYA, Kumiko TSUJI
    1979 Volume 1979 Issue 7 Pages 896-900
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The 4- (phenylazo) -1-naphthol [1] and its derivatives were easily etherified in the presence of sulfuric acid to form 1-methoxy-4- (phenylazo) naphthalene and its derivatives. The irreversible reactions were traced in terms of visible spectrophotometry. The rate equation was of first-order with respect to J. The Hammett plots of derivatives with electron releasing group and with electron-attracting group gave two straight lines with slopes + 2.58and +0.40, respectively. The isokinetic plots gave two straight lines similarly. The B-values were 430K and 396K for electron-releasing and attracting group, respectively.
    A probable mechanism, that the rate-determining being the a dhesion of methanol on the protonated tautomer (I) in the case of the electron-releasing group and that the rate-determining being the deprotonation from methanol adhesive substance toward (I) and dehydration in the case of the electron-attracting group, was proposed and discussed.
    Download PDF (202K)
  • Tadataka ZAIMA, Keiryo MITSUHASHI
    1979 Volume 1979 Issue 7 Pages 901-905
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of 1, 1'-bis(methoxycarbonyl)divinylamine (BDA) with hydrazines was investigated as an example of nucleophilic attack on a cross-conjugated dienamine.
    The reactions of BDA with 80.5% hydrazine hydrate, methylh ydrazine, methoxycarbonylhydrazine and phenylhydrazine gave pyruvic acid hydrazide hydrazone [3], methyl pyruvate methylhydrazone [4], methyl pyruvate methoxycarbonylhydrazone [5] and methyl pyruvate phenylhydrazone [6], respectively, in high yields. The 'H-NMR spectra of BDA in various solvents indicated the predominant existence of enamine form between enamine and imine tautomers. In addition, the reaction of N-methyl-1, 1'-bis(methoxycarbonyl)divinylamine [8] with phenylhydrazine also gave methyl pyruvate phenylhydrazone [6].
    An appropriate mechanism yielding above hydrazones, [3]-[6], is considered to involve a nucleophilic attack of hydrazines at the electron-deficient α-carbon atom of enamine form of BDA, followed by an elimination of methyl 2-aminoacrylate.
    Download PDF (268K)
  • Hiroaki TANIGUCHI
    1979 Volume 1979 Issue 7 Pages 906-910
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The influence of solvent, temperature, and the types of olefins on the equilibration reaction of a mixture of norbornylboranes were examined. In heptane or neat, the distribution rate was much smaller than that in THF-solution and the content of di-exo-norbornylborane under the norbornene: borane ratio of 2 was less favorable (in haptane, 47-50%; neat, 40-43%)than that of THF solution (55-57%). At higher temperatures, the distribution rate was considerably accelerated (the rate under THF-reflux was 102 times faster than that at room temperature); but the equilibrium distribution itself did not change significantly. The yield of dialkylborane in the redistribution reaction (R3B+BH3→)was remarkably influenced by the types of- olefins. It was possible to obtain an almost quantitative yield of R2BH using 1-
    Download PDF (268K)
  • Taketoshi KITO, Koki OTA
    1979 Volume 1979 Issue 7 Pages 911-914
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Primary alcohols (RCH2CH2OH) having more than three carbon atoms condensed in the presence of alkali metal salts of naphthols or phenol to give 2-alkyl-l-alkanols (RCH2CH2CHRCH2OH) in good yields. In these reactions, carboxylic acids were scarcely form ed.
    Download PDF (197K)
  • Tetsuo ODA, Toshihisa MAESHIMA, Kazuo SUGIYAMA
    1979 Volume 1979 Issue 7 Pages 915-920
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    1, 4-Dimethyl-1, 4-bis(methyl-substituted pheny1)-2-tetrazenes [1] can be classified into two conformational groups according to spectral data: trans-planar types [1 a, b, c, d] having the methyl groups at 3 or/and 4 positions on phenyl groups, and trans-orthogonal types [1 e, f, g]with the methyl groups at 2 or/and 6 positions on phenyl groups. The rates of thermal decomposition of C have been determined in various protic solvents by means of UV spectroscopy. Rate constants (k) which increase with the polarity of the solvents are in the order [1e]>>[1d]>[1b]> [1c]> [1a]>> [1g]>[1f] in all solvents. A linear relation is found for a plot of ΔH against the solvent polarity parameter (ET). A compensating change is found in ΔH and ΔS for each [1] in various solvents, showing that a single mechanism is operative for the thermal decomposition of [1] in various solvents. On the other hand, the reaction mode differs somewhat between trans-planar and trans-orthogonal types of [1]. The decomposition of [1] appears to proceed via conformational change of trans-planar trans-orthogonal→ cis-orthogonal. Steps a and b are rate-determining step for the decomposition of [1 a, b, c, d] and [1 e, f, g], respectively
    Download PDF (333K)
  • Takahiro NAKAMURA, Shizuo NAGAHAMA
    1979 Volume 1979 Issue 7 Pages 921-924
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Effect of the addition of oxidizing agents, such as nitrobenzene, o-chloronitrobenzene, mdinitrobenzene, acetaldehyde, SnCl4 and I, on the yield of 8-chloro-2-methylquinoline in the Doebner-Miller reaction between crotonaldehyde or paraldehyde and o-chloroaniline in the presence of 30% aqueous HCl was studied.
    The addition of nitrobenzene or o -chloronitrobenzene increased the yield, whereas the other oxidizing agents did not. Nitrobenzene did not increase the yield of 8-nitro-2-methylquinoline in the reaction of o-nitroaniline and crotonaldehyde. Further quantitative studies were made on the Doebner-Miller reaction for the synthesis of 8-chloro-2-methylquinoline from crotonaldehyde or paraldehyde and o-chloroaniline in the presence of nitrobenzene or o-chloronitrobenzene.
    Download PDF (227K)
  • Nobuyoshi NAMBU, Kaneo YAMAMOTO, Sawako HAMANAKA, Masaya OGAWA
    1979 Volume 1979 Issue 7 Pages 925-930
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The complex layers obtained by the reactions of p-diethylbenzene(p-DEB) and m-diethylbenzene(m-DEB) with the AlCl3-HCl system were investigated by using 13C-NMR in order to elucidate the mechanism of transalkylation of alkylbenzenes. m-DEB was readily protonated to give its benzenium ion (Fig.2(a)), but p-DEB was rapidly transalkylated into ethylbenzene and 1, 2, 4-triethylbenzene, the latter being further converted to 1, 3, 5-triethylbenzene(s-TEB)which in turn was protonated to give its very stable benzenium ion (Table 1 and Fig.1).
    The transalkylation appeared to consist generally of successive dealkylation and alky lation and that the dealkylation occurs at the position of the ring carbon having the higher electron density, since the protonation of the ring carbon was the first step of this reaction. Therefore, the different behavior between p-DEB and m-DEB was explained by comparison of the 13C chemical shifts of their ring carbons which were in parallel with the electron density (Fig.5).
    Namely, the stability of protonated ethylbenzenes decreased in the order m-DEB s-TEB >ethylbenzene>l, 2, 4-triethylbenzene>p-DEB, whereas their reactivity gave the opposite sequence, as observed in the reaction carried out by adding successively AlCl3- and HCl to the diethylbenzene mixture (Table 3 and Fig.3). These results showed that the transalkylation of diethylbenzenes might occur according to the reaction courses given in Fig.6
    Download PDF (324K)
  • Kiyoshi NARUCHI, Seiji TANAKA, Masatoshi MIURA
    1979 Volume 1979 Issue 7 Pages 931-935
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Barium methacrylate undergoes facile oligomerization on heating in a solid state. The effect of heating temperature on the formation of dimers and trimers was investigated in the range 150240°C. At 150-460°C α-methylene-ε-methyladipate [2] was formed. At 170°C and above, however, α-methylene-γ, γ-dimethylglutarate [1] was formed together with [2], and yields of both dimers increased with a rise in temperature. At any temperature, yield of [2]was greater than that of [1]. The heating of barium methacrylate containing a small amount of water at 220°C in a liquid state resulted in remarkable decrease in the yield of C [1] and in large increase of [2].
    Change of t he crystal lattice after the thermal oligomerization was observed by X-ray diffraction. A part of the crystal lattice of the monomer salt was transferred to another one with formation of oligomers without destruction of the crystal. When total oligomer formation reached to about 50% of the monomer, all of the crystal lattice changed to the new lattice, which was maintained during further formation of oligomer. From these results, it is concluded that [2] is formed in both the crystalline and amorphous states, and [1] only under lattice control.
    Download PDF (314K)
  • Takaaki TANAKA
    1979 Volume 1979 Issue 7 Pages 936-940
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Vinyl monomer(MMA) began to polymerize by TiCl3 in water containing oxygen after a certain induction period. The polymerization did not occur in water free from oxygen. The yield of polymer increased with increasing oxygen content(Fig.1). The concentration of TiCl3decreased rapidly by the reaction of it with dissolved oxygen, and the mole ratio of reacted TiCl3 to oxygen was 4(Fig.3) in the absence of MMA. In the presence of MMA, how ever, the mole ratio was about 2(Figs.4, 5), and polymerization occurred after oxygen was almost consumed. These results suggest that the reaction of TiCl3 with oxygen proceeds by radical process, and the reaction mechanism during the induction period was considered. Some of oxygen changed to peroxides during the induction perio d(Table 1), and it was assumed that the redox reaction of these peroxides with TiCl3 initiated the polymerization.
    Download PDF (270K)
  • Yoshio SATO, Toshihiro NAKAI, Kenji TASUMI
    1979 Volume 1979 Issue 7 Pages 941-945
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Transport properties of benzenesulfonate and methyl-substituted benzenesulfonate ions in electrodialysis were studied.
    The preferential p ermeability of benzenesulfonateions decreased with the increasing of applied current density, the concentration decreaseing of dialysing solution and the increasing in the number of methyl group contained. The permeability was related to the Stanton number (Se), a dimensionless operating parameter, defined as the ratio of ionic flux in the direct ion to an electrodialysis membrane to ionic flux in the direction of the bulk flow of electrolyte. As the Stanton number increased, the permeability decreased and showed definite values for St>10-3. The permeability coefficient calculated from the mobility ratio of the ions in ionexchange membrane in a single-solute-solution and the composition of ions in membrane in a binary-solute-solution was far less than that obtained for desalting experiments. It is suggested that chloride ion, which is more mobile and present much more in the membrane than benzenesulf onate ions, causes an acceleration of the migrating velocity of each less mobile benzenesulfonate ion.
    Download PDF (290K)
  • Satoshi TOMIZAWA, Kunio KAMIYA, Mitsutaka KAWAMURA, Shigeki TOYAMA
    1979 Volume 1979 Issue 7 Pages 946-950
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A study was carried out to clarify some properties of substances contained in a dust sample collected by an electric precipitator (E. P. dust) in an incineration plant of municipal refuses. Chemical species and chemical components of the substances and the solubility of each component at various pH values were investigated. The main constituents found w ere 46% K3Na(SO4)2, 19% NaCl and 8% KCl. α-Quartz was also found in the dust. The amount of toxic components in the dust was 300 ppm Cd, 260ppm Cr, 5100 ppm Pb, 31 ppm As, and 1.4 ppm Hg.
    The main constituents were completely dissolved in water at pH 8.4. The percentage in weight of insoluble residues was 25% of the dust. Pb, Cr, As and Hg were separated as insoluble residues together with SiO2, Al, Mg, Fe, Ti, Zn, Sn, Cu and Ni from the soluble substances in water leaching. A part of Cd was dissolved in water.
    Download PDF (264K)
  • Noriho HARUMIYA, Hajime MIYAMA
    1979 Volume 1979 Issue 7 Pages 951-952
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Quantum Yield of photonitrosation of cyclododecane in carbon tetrachloride was measured. Obtained value at 365 nm was about 0.6 irrespective of the concentration of cyclododecane. Fluctuation of values obtained at 578 nm was very high. However, by choosing suitable concentration of nitrosyl chloride and cyclododecane, high value of 0.8 or above was obtained.These results were discussed in terms of the mechanistic viewpoint.
    Download PDF (94K)
  • Yasuo HIKICHI, Ken-iti HUKUO, Jiro SHIOKAWA, Masao MIZUNO, Tetsuo NOGU ...
    1979 Volume 1979 Issue 7 Pages 953-954
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The freezing point of the synthetic RPO4(R=La, Pr, Nd, Sm, Y, or Er) was measured in terms of the cooling curve of molten specimen by using the specular reflection method with a heliostat type solar furnace. The results obtained were as follows: 2072+20°C(LaPO4), 1938±20°C(PrPO4), 1975±20°C(NdPO4), 1916±20°C(SmPO4), 1995±20°C (YPO4) and 1896±20°C(ErPO4).
    Download PDF (134K)
  • Yoshiharu MATSUBARA, Katsumi YOKOI
    1979 Volume 1979 Issue 7 Pages 955-957
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    (+)-endo-Fenchol [1] was dehydrated with KHSO4 to give δ-, cyclo-, γ-, ε-, β-, and a-fenchenes ([2]-[8]). (+)-exo-Isofenchol [9] was obtained by the hydration of these fenchenes. (-)-endo-Isofenchol (11) was synthesized selectively by reduction with LiAlH4 of isofenchone [10] which in turn was obtained by oxidation of [9]. The structu res of [9] and [11] were confirmed by the NMR measurement of pseudocontact shift due to an added shift reagent.
    The sample of [9] has elegant fragrance similar to nojigiku alcohol [A] or 6-exo-hydroxyexoisocamphane [B] whereas [11] has borneol-like odor.
    Download PDF (189K)
  • Hiroaki EGAWA, Hiroshi HARADA
    1979 Volume 1979 Issue 7 Pages 958-959
    Published: July 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Kumamoto University; Kurokami, Kumamoto-shi 860 Japan It was found that the chelating resins (RNH) containing amidoxime groups show the selective adsorption ability for uranium in sea water. The favorable chelating resins were prepared by the reaction of macroreticular acryronitrile-divinylbenzene copolymer beads with hydroxylamine in methanol. When sea water (104 times as large as the resin volume) was passed through the RNH column at the space velocity of 60 hr-1, 80% of uranium on that in sea water was recovered. The elution of uranium adsorbed on RNH was accomplished by passing 1 N sulfuric acid solution through it. The recycle of adsorption and elution was found to be satisfactory.
    Download PDF (95K)
feedback
Top