NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1979, Issue 10
Displaying 1-28 of 28 articles from this issue
  • Ken HIGASHITSUJI, Akira WATANABE
    1979 Volume 1979 Issue 10 Pages 1287-1293
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    For a study of the mechanism of electric emulsification, the excess free energy of interface, zla(C), was calculated for a spherical interface of curvature C with the electrochemical free energy of the double layer. It is concluded that, under a specified condition, a sharp protuberance or a fine stream of jet can grow at an interface, which leads to dispersion or emulsification taking place at the pointed end. A treatment of the entire dispersion system with a new variable, N, the number of dispersed PHASE's, shows that zero or negative interfacial tension is not the necessary condition for electric emulsification. This can also be applied to spontaneous emulsifications. It is also predicted that, for W/O type emulsifications, the most probable radius of water droplets to be formed is of the order of llkw, the double layer thickness in the water phase.
    Download PDF (398K)
  • Moriyasu MURATA
    1979 Volume 1979 Issue 10 Pages 1294-1298
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Relation between C potentials iron 010 oxide (Fe2O2-) or titanium oxide (TiO2) and calcium ion or magnesium ion concentration in the solutions, containing sodium dodecylpoly(oxyethylene) sulfate (R120(E0)SO3Na, where n is 1 or 3), sodium hydroxide, sodium chloride and calcium chloride or magnesium chloride was studied, in order to examine the previously predicted dependence of the Stern charge densities (a) on calcium ion concentration. These experiments were conducted under the same conditions as the previous one as follows; in the presence of definite concentration of R, 20(E0)O2-1a and 0.3mmolg sodium hydroxide at ionic strength of O.01 and 25°
    C. With the increase in the ratio of divalent metal ion concentration to total cation concentration (Cmo. /(Cm++ C. ), where Me- indicates CO+ or Mg2+), C potentials of Fe2O2- or TiO2 increased sharply up to a maximum value and then decreased, in the presence of sufficient amounts of R12O(E0)O3Na. On the other hand, C potentials decreased exponentially in the absence of R120(E0)O3Na. The cooperative effect of R120(E0), SO2-la and divalent metal ion on C potential of solid particles in miscellaneous electrolyte solutions was shown in Figs.1 to 4. The double layer charge density (id) was calculated by the equation 3 in which C potential was used instead of the Stern potential (To). Some of the dependences of d on calcium ion concentrations were shown as the solid lines in Figs.5 and 6, together with the dotted lines which were referred to a. The ad depended on (Coo++ C. ) in the same manner as a did. These facts support the previous prediction that the mutual charge cancellation of R120. (E0)SO2-1a and divalent metal ion on solid particle might cause the characteristic dependence of surface charge density on divalent metal ion concentration in the presence of definite concentration of R120(E0)O3Na.
    Download PDF (291K)
  • Keisuke MITA, Kazuhiko HOTTA, Shoji WATANABE
    1979 Volume 1979 Issue 10 Pages 1299-1303
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The hydrogenation reaction of benzil catalyzed by Co(dmgH)2-(n-C41-19)3P, Co(dmgH)2-pyridine, Co(dmgFI), -CH, NH Co(dmgH), --(CH, ), NH, Co(dpgH), -pyridine, Co(dpgH)2-CH3NH2, or, Co(dpgH)2--(CH3)2NH system, where dmgH is monoanion of dimethylglyoxime and dpgH is monoanion of diphenylglyoxime, was studied at various NaOH concentration under an atmospheric hydrogen pressure and at 0 and 20° C. It was found that the rate of hydrogenation catalyzed by Co(dmgH)2-(n-C41-19)3P system is independent of NaOH concentration, where as that catalyzed by other amine system depends on NaOH concentration. The absorption spectrum of Co(dmgH)2-(n-C41-19)3P system in the presence of molecular hydrogen and benzil shows bands at 408, 513, and 614nm in the visible region, which agree with those of Co(111)-hydride complex, while other amine system show the absorption bands due to Co01 complexes.
    These results suggested that the reaction catalyzed by Co(drngH)2-(n-C4H9)3P system proceeds with a mechanism being different from the hydrogenase model which was proposed to the 1.17- drogenation reaction catalyzed by Co(dmgH)2-amine system. The dependence of the rate of hydrogenation catalyzed by Co(dmgH)2-pyridine system upon NaOH concentration could be explained in terms of the enzymatic model, and in this case not only pyridine (at the alkali concentration of [NaOH]/[Co]=-0-2) but also dimethylglyoximate anion (at [NaOH]/[C0]2) act as basic sites.
    Download PDF (287K)
  • Hajime EBISAWA, Eiichi MIYOSHI, Tsuneo SHIRAI, Saburo YANAGISAWA
    1979 Volume 1979 Issue 10 Pages 1304-1308
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction mechanism of organic compounds for a Cl or C2C hydrocarbon-ammonia-water vapor system with an electric discharge treatment was discussed in our previous papar. The yield of the organic compounds, especially, amino carboxylic acids and nuclein bases is reported in this papar. Improvement of the electric discharge apparatus to be used was attempted. The injection quantities of starting gases and the addition of metallic ions to water also were investigated. Results revealed that the water was important for the production of amino carboxylic acids obtained when the hydrocarbon-ammonia-water system was subjected to an electric discharge in the type of a gas circular or non-gas circular. Similarly, the water also was important for the formation of nuclein bases. The yield of amino carboxylic acids increased linearly with the increase of initia ammonia concentration, but it did not with that of initial hydrocarbon concentration. And it was found that metallic ions dissolved in water had an effect on the production of amino carboxylic acids obtained.
    Download PDF (317K)
  • Yoshihiro HIRATA, Akio KATO
    1979 Volume 1979 Issue 10 Pages 1309-1315
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The formation of magnesium-doped and undoped 8-type alumina (8-, -alumina) powders was investigated by a spray-pyrolysis technique involving the ethanolic solution of sodium, magnesium, and aluminum nitrates. Spray-pyrolysis products were r-alumina over the reaction temperature range of 800 to 1400° C and could be converted to the 8-type alumina by heat treatment above 1100° C. The conversion temperature for 1 hr heating period worked out at 1300° C for the magnesium-undoped product with a molar ratio Na20/Al2O2- = 1/5 of the atomizing solution. Similaryl 1100° C and 1200° C for the magnesium-doped products were obtained with the respective ratio of 1/5 and 1/8. Both in the magnesium-undoped products with the ratio of 1/11 and 1/8 and the magnesium-doped product with the ratio of 1/11, the heat-treated products (1300° C) contained a-alumina, MgAl2O2- or mullite-type alumina together with 8-type alumina. The formation process of 1-type alumina from r-alumina in the magnesium-doped products differed from that involved in the magnesium-undoped products. In the magnesiumundoped product 7-alumina was converted to 8-type through the mullite-type alumina as an intermediate. In the magnesium-doped product, on the other hand, 7-alumina was converted to 8-type directly without the formation of mullite-type one.
    Download PDF (542K)
  • Shizuko HIRATA
    1979 Volume 1979 Issue 10 Pages 1316-1321
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Organic carbon, iron, manganese, zinc and copper in the sediments of the Hiro Bay contaminated with a pulp mill e uent were determined. There were positive correlations between organic carbon and iron, zinc/or copper in humic substances extracted from the sediments with 0.5 N sodium hydroxide solution. First of all, humic and fulvic acids extracted from the sediments were fractionated on Sephadex gel columns. The humic acid fraction showed a molecular weight distribution from 0 to over 200000. Humic acid with mol. wt.10000-50000 covered 48.3% of the total eluted from the column, and iron, zinc and copper were eluted down with respective proportions of 45.1, 30.2 and 47.0% along with the humic acid fraction just mentioned. Fulvic acid exhibited a relatively narrow mol. wt. range not exceeding 10000. Fulvic acid fraction having mol. wt. less than 700 was as high as 70.1% of the total eluted from the column, and iron, zinc and copper were removed along with it with proportions of 78.2, 61.1 and 60.7%, respectively.
    Download PDF (344K)
  • Hiromu SATAKE, Sanae IKEDA
    1979 Volume 1979 Issue 10 Pages 1322-1326
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    An amperometric titration measuring the reduction current of iodate ions for the differential determination of H2N4.2 HCl and NH2OHCl in mixtures has been studied with a rotating platinum electrode (2000 rpm) at the potential of +O.6 V vs. SCE. The method is based on the difference in their reaction rates with potassium iodate in the presence of hydrochloric acid and sodium chloride. N2H4.2 HCl in the mixture was titrated with potassium iodate prior to NH2OHCl, while the latter was titrated after the addition of potassium bromide.
    A series of mixtures containing between 0.2, umol to 200, umol of N21-14HCl and 2pmol to 200, umol of NH2OHCl was determined by the following two methods. The recommended procedures were as follows.
    ( 1 ) Successive determination: An equimolar mixture (ca.20 pmol each) containing N2H4.2HCl and NH2OHCl was introduced into the titration cell, and hydrochloric acid (6molg, 8 m1) and water were added to make the total volume up to 30m1. After the solution was saturated with sodium chloride (18g) at 45° C, N2H4.2HCl was titrated with 0.05m1 portions of 0.01 molf/ potassium iodate at intervals of 5sec (O.05 m//5sec). Then NH2OHCl was suc- cessively titrated with O.01 mol/l potassium iodate (O.05 ml/20 sec) at 45° C after potassium bromide (4 mol//, 4.5m1) and water (25m1) were added. Mixtures containing N2-4..2 HCl and NH2OHHCl, of which molar ratios were 1: 1 (20 pmol each) to 1: 8 (5, umol of N2H4.2HCl: 40pmol of NH2OHCl), were analysed with the relative error less than O.3%. Time required for titration was about 15 minutes.
    ( 2 ) Differential determination: An equimolar mixture (ca.20, umol each) containing N21-142HCl and NH2OHCl was analysed by using two aliquots. One aliquot was used to titrate N2H4.2HCl with 0.01 mol// potassium iodate (0.05 ml/5sec) at room temperature after the addi- tion of hydrochloric acid (6 mol//, 20m1), sodium chloride (18g) and water to make the total volume up to 50m1. Another aliquot was used to titrate the sum of N2H4.2HCl and NH2OH. HCl with O.01 mol/l potassium iodate (0.05 ml/20 sec) at 45° C after the addition of hydrochloric acid (6 mol//, 7m1), potassium bromide (4 mol//, 4m1), sodium chloride (18g) and water to make the total volume up to 50m1. The amount of NH2OHCl was estimated from the difference in the two titrations. Mixtures containing N2H4.2HCl and NH2OHCl, from 5: 1 (20, umol of N2H4.2HCl: 4 pmol of NH2OHCl) to 1: 8 (5, urnol of N2H4.2HCl 40, umol of NH2OH HCl), were analysed with the relative error less than 1%. A mixture of 0.2, umol of N2H4 2HCl and 38pmol of NH2OHCl was analysed with the relative error of about 2.5%. Total titration time was about 25 minutes. The effects of concomitant compounds were examined.
    Download PDF (311K)
  • Kiichi SUZUKI, Kenichi NAKAZATO, Tetushi TAKASAKI
    1979 Volume 1979 Issue 10 Pages 1327-1331
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Acrylamide, N, N'-methylenediacrylamide and acrylonitrile were suspended in petroleum hydrocarbon and copolymerized to give spherical porous gels in the presence of water-ethanol as a diluent, in which ethanol served as a diluent.
    Water soluble materials such as acetone, ethylene glycol oligomers and dextrans were allowed to pass down the columns packed with the gels to construct calibration curves. The increasing proportin of the crosslinking agent and acrylonitrile in total monomer yielded a larger exclusion limit. Especially the increasing proportion of the crosslinking agent in a high molar ratio of acrylonitrile to acrylamide yielded a larger exclusion limit most effectively. For the gel containing 17.9 mol% crosslinking agent for the total monomer in which the molar ratio of acrylonitrile to acrylamide was O.27, its excluded molecular weight, degree of swelling (wetm//dry-mi) and amount of water regain (g/dry-g) worked out at 4 x IV, 3.74, and 2.89, respectively. On the other hand, for the gel of 16.7 mol% of crosslinking agent for the total monomer in which the molar ratio of acrylonitrile to acrylamide was O.98, its excluded molecular weight, degree of swelling and amount of water regain were 3 X 107, 4. O2- and 3.01, respectively.
    Download PDF (239K)
  • Mikio WADA, Takuji KANNO
    1979 Volume 1979 Issue 10 Pages 1332-1336
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Variation in the partition coefficients of chelating extractants, 2-thenoyltrifluoroacetone(HTTA), 8-quinolinol (HOx) and cupferron (HCup), and of their americium chelates have been studied between a number of inert solvents and dilute perchloric acid solution of ionic strength O.1. Partition behavior of cupferron was investigated by means of spectrophotometry, and its perti- tion coefficients were correlated with solubility parameters of solvents. Distribution ratio of americium was determined by the r-counting rate of Am. The extracted chelates were found to be of the type AmA3, where A is the chelating anion. Partition coefficients (PM) for carbon tetrachloride and stability constants (R8) were obtained as; log PM = 1.10, log R3=12.66 for Am(TTA).; 1.75, 22. O2- for Am (0x), ; and 80, 9.10 for Am (Cup)The PM values in other solvents were estimated from the readily determinable Pmf 3values. Variations of these Pm values in different solvents were also explained in terms of solubility parameters.
    These regular trends of the PM values were similar to those of the partition coefficients of the extractants (PHA). The linear correlations, log Pm, - n log PHA + constant, were established for the solvents examined.
    Download PDF (280K)
  • Shizunobu HASHIMOTO, Isao FURUKAWA, Hiroshi GOTO
    1979 Volume 1979 Issue 10 Pages 1337-1342
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reactions of triphenylphosphine-trichloromethyl compounds with styrene oxide were investigated. Triphenylphosphine-carbon tetrachloride as a trichloromethyl compound gave styrene dichloride C 2D, a-chlorostyrene [3], [chloro (triphenylphosphoranylidene) methyl] triphenylphosphonium chloride C4 J, triphenylphosphine oxide C 5 and chloroform.
    The reaction was accelerated in a polar solvent such as acetonitrile. The use of excess triphenylphosphine to styrene oxide in this reaction produced C2CD in an almost quantitative yield to the exclusion of OD. However, the addition of phenols to this reaction system produced C 3D in a high yield.
    The reactions using other trichloromethyl compounds, except he xachloroethane, resulted in the formation of the products analogous to the case of carbon tetrachloride. The use of trichloromethyl compounds having more electron-withdrawing groups increased the ratio of C2C to [3]. On the other hand, the reaction using hexachloroethane proceeded at room temperature to give C2C in a quantitative yield. On the basis of these results, the reaction mechanisum was discussed in detail.
    Download PDF (358K)
  • Tohru NISHIWAKI, Masao USUI, Kinji ANDA, Mitsuhiko HIDA
    1979 Volume 1979 Issue 10 Pages 1343-1349
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Polychlorobenzenes were photolyzed in both alkaline and neutral 2-propanol in order to investigate the positional selectivity towards dechlorination. The selectivity under neutral conditions was considerably different from that in the presence of alkali (Table 2, 3 and 4), which was elucidated on the basis of the steric and the electronic effects.
    Chlorine atoms were, classified into the following three types; 1) one having two adjacent chlorine atoms on benzene ring, 2) one having one adjacent chlorine atom, and 3) one having no adjacent chlorine atom. Chlorine atom of type 1 was most reactive. The reactivities of chlorine atoms in type 2 and 3 decreased greatly in that order (Table 5).
    The positional selectivity under neutral conditions was considered to be determined predominantly by the steric effect, caused by the adjacent chlorine atom. The electronic effect on the reactivities in the presence of alkali was estimated by the electron densities on carbon and chlorine atom of corresponding radical anion, which were calculated by use of Hiickel MO method. The reactivities in the presence of alkali were successfully explained by the electronic effect (Table 6 and 7), rather than by the steric effect.
    The reactivities of polychlorobiphenyls were similarly discussed from the viewpoints of the electronic and steric effects (Table 6 and 7).
    Download PDF (436K)
  • Masatoshi NAGAI, Kiminori SAWAHIRAKI, Toshiaki KABE
    1979 Volume 1979 Issue 10 Pages 1350-1354
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The hydrodenitrogenation of acridine on a pre-reduced MoO3-Al2O2- catalyst, as an appropriate, model reaction of residual oil, was studied to obtain mechanistic information. The study was carried out in a flow microreactor using xylene solutions containing 0.1-1.0 wt% acridine at temperatures of 200-350° C, total pressures of 25-145 atm, WHSV 10, and hydrogen flow rate of 30 //hr.
    Acridine was easily hydrogenated to 9, 10-dihydroacridine even below 200° C. Dihydroacridine was also hydrogenated succesively to 1, 2, 3, 4, 4a, 9, 9a, 10-octahydroacridines and perhydroacridines, which was denitrogenated to dicyclohexylmethane (DCM) above 300° C (Fig.2). The activation energy for the formation of DCM was found to be 32. ° kcal/mol. The rate equation was expressed as
    where, r is the rate of formation of DCM, and CpA and PH2 are the partial pressures of perhydroacridines and hydrogen, respectively. k, KPA and KH2 are constants. On the basis of the kinetics, it was found that the hydrogenation of acridine to perhydroacridines is in equilibrium at higher temperatures and the rate-determining step is the reaction of perhydroacridines with hydrogen molecule (Fig.9).
    Download PDF (301K)
  • Keizo OGINO, Masahiko ABE, Hideyuki DOBASHI
    1979 Volume 1979 Issue 10 Pages 1355-1358
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Thermal behavior of anionic surfactants, sodium alkyl sulfates, has been studied by thermogravimetry (TG), differential thermal analysis (DTA), differential scanning calorimetry (DSC), infrared absorption spectrometry, X-ray powder diffraction analysis and mass spectrometry. When alkyl sulfates were heated dynamically in air, their weights decreased due to the evapora- tion of the decomposition products and oxidation products. The temperature at which the weight loss started was raised by an increase in carbon number n of the surfactant. The temperatures for the surfactants with n=8, 10, 12, 14, 16 and 18 were 125, 132, 140, 142, 165 and 168° C, respectively. The weight loss was remarkable up to 270° C, and stopped at 550° C. The residue was composed of sodium thiosulfate and sodium sulfate. The analysis by DTA indicated that a phase transition similar to the one observed in fatty acid soap (Curd-Waxy) took place, when the surfactant, whose carbon number was higher than 12, was heated. The heat of transition was about 1 kcal/mol and was ascribed to the microbrownian or rotational motion of methylene groups.
    Download PDF (235K)
  • Katsumi KAMEGAWA, Hisayoshi YOSHIDA, Seiji ARITA
    1979 Volume 1979 Issue 10 Pages 1359-1364
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It was found that arsenite ion is easily oxidized to arsenate ion by dissolved oxygen in the presence of activated carbon. The catalytic oxidation of As(III) has been investigated at 25° C by using five species of activated carbon.
    The activated carbon shows high catalytic activity for the oxidation in acidic or alkaline solution, especially in alkaline solution (Figs.1, 4). The catalytic activity depends on the raw materials from which an activated carbon is obtained. The order of the activity in alkaline solution is as follows: activated carbon from coal>activated carbon from coconut shell activa- ted carbon from wood (Fig.4). This order is reversed in acidic solution (Fig.3). The activated carbon which adsorbs more HCl or NaOH shows the higher catalytic activity in alkaline or acidic solution, respectively (Fig.5). Therefore, it seemed that the catalytic activity depends on the amount of basic and acidic surface oxides on activated carbon. The oxidation rate of As(l11) to As(V) V was found to be proportional to the concentration of As(111) and the amount of activated carbon added (Figs.7, 9). From these results, oxidation of AWE on activated carbon was considered to proceed as follows: As(11) is oxidized to As(V) V by such active surface oxides on activated carbon as quinone or hydroperoxide type. Subsequently, the resulted such inactivated surface oxides as hydroquinone type are regenerated by oxidation with dissolved oxygen in the solution. As an overall reaction, therefore, the dissolved oxygen oxidizes As (la) to As(V) V and the active surface oxide acts as a catalyst of this reaction.
    Download PDF (397K)
  • Katsumi KAMEGAWA, Hisayoshi YOSHIDA, Seiji ARITA
    1979 Volume 1979 Issue 10 Pages 1365-1370
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Adsorption characteristic of arsenite and arsenate ions on activated carbon was studied at ambient temperature (25° C) by using five kinds of activated carbon. As(V) was determined by atomic adsorption spectrometry of molybdenum after the extraction of molybdoarsenic acid into MIBK. Total As was determined by the same method after oxydation of As(111) to As (V) with KMnO2-. All activated carbons used were deashed with HF solution before use to prevent the interference from silicon in the determination of arsenic.
    The species of As(V) adsorbed on activated carbon were considered to be 112AsO2-- and HAs O2-2-, since As(V) was adsorbed in the pH range from 4 to 9 on activated carbon prepared from coal, coconut shell and wood by steam activation method (Figs.1, 2, 3). The adsorption ability of activated carbon for As(V) was in the order: activated carbon from coalactivated carbon from coconut shellactivated carbon from wood (Fig.4). Influence of coexistent cations on adsorption of As(V) was small (Fig.6). The coexistence of anions interfered greatly with adsorption of As(V) because anions was adsorbed competitively on adsorption sites for H2AsO2-- and HAsO2-2-. The interference form anion was large in the following order: ClO2--SO42-N08-Cl- (Fig.5). It should be noted that As(V) was scarcely adsorbed on activated carbon prepared from wood by the ZnCl, activation method (Fig.3).
    As(111) was not adsorbed on all the activated carbons used, since As(111) dissolved in the undissociated form, 1-13AsO2-, below pH 8. However, As(M) was easily oxidized to As(V) with dissolved oxygen on activated carbon so that As formed was adsorbed on activated carbon. Activated carbon, therefore, having a higher activity to oxidize AWE showed a higher adsorption ability for As. The results of this study support our idea that heavy metal ions are adsorbed on activated carbon in the form of complex anion.
    Download PDF (399K)
  • Shuji KAWAI, Yoshihiro ICHIHARA, Hiroshi NARUKAWA, Haruo NAKAYASU
    1979 Volume 1979 Issue 10 Pages 1371-1375
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The melt viscosity measurements of poly (ethylene terephthalate) were carried out within the ranges of the molecular weight 9.21 x 1O2- < M< 1.48 x 105 and the temperature from 265 to 285° C. In order to obtain the molecular weight of polymer running through the viscometer the melton polymer was quenched immediately after each viscosity measurement and the intrinsic viscosities of the quenched samples were measured.
    Thus obtained data were examined employing the current emperical equations. In the logarithmic plot of the viscosity against weight average molecular weight, data were approximated by two straight lines with slopes of 3.4 in the high molecular weight range and of 1.7, in the low molecular weight range. The critical chain length calculated from the molecular weight at the intersection of the two straight lines was 211, which is simmilar to the value reported for polyesters.
    The parameters in the Berry-Fox emperical equations were discussed. In the logarithmic plot of viscosity against X, which is a function of chain length, mean square radius of gyra- tion, molecular weight, volumetric fraction and specific volume of the polymer, data were approximated by straight lines with the slope of 3.4 in the range of high values of X and of 1.7 in the range of low X. The latter value of the slope became to unity when data were converted to those at iso-free volume state.
    Polymers prepared under more oxidative conditions tend to give slightly lower viscosities. Although a definite evidence is lucking, it is considered that the difference in viscosity is attributed to the small degree of branching generated during heating of the polymers.
    Download PDF (278K)
  • Yasuko ANDO, Toshio IWAMOTO, Yuko ATSUMI, Noriko SAKAI, Jiro KOMIYAMA, ...
    1979 Volume 1979 Issue 10 Pages 1376-1383
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The binding isotherms of p-aminoazobenzene or p-[Dis(2-hydroxyethyl) amino]azobenzene to sodium poly (styrenesulfonate) having various degrees of sulfonation have been determined in acetate buffer at pH 5.65 and 25° C using the dynamic dialysis technique. The apparent binding constant, calculated on the basis of free residue of the polymer and free and bound dye, was dependent on the dye concentration in all cases examined. The positive cooperativity in the dye binding was suggested by sigmoidal increases (Fig.5 and Fig.8) of the binding density, the molar ratio of bound dye to polymer residue, with rise of free dye concentration and by the concave upward shape of the Scatchard plots (Fig.7 and Fig.9). Application of the McGhee and von Hippel's theory to the present results gave the following values: the number of consecutive residues covered by a bound dye, n=75-90; the intrinsic binding constant, K =25-55 in liter/residue mol of the polymer; and the cooperativity parameter, w=120-300 (Table 2). A speculative model of the interaction was presented in terms of the three parameters.
    Download PDF (467K)
  • Kiyoshi JUNA, Takeo SUGANO, Fujio TANINO
    1979 Volume 1979 Issue 10 Pages 1384-1389
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to elucidate the mechanism of dissolution of polyacrylonitrile (PAN) in concentrated aqueous solution of Mg(SCN)2 or Ca(SCN)2, the interaction among alkaline-earth metal salts and the following three aliphatic dinitriles, succinonitrile, glutaronitrile, or adiponitrile, was investigated by IR and Raman spectroscopic measurements as well as by potentiometric measu- rements with a divalent cation-selective and a thiocyanate ion-selective electrode.
    The results are as follows: (a) In the presence of Mg(NO3)2 or Ca (NO3)2, the IR spectra of glutaronitrile and adiponitrile show a new strong band (v'), assignable to the coordinated nitrile groups in the range 2270-2290 crn-', in addition to a normal peak (v) for the free cyano groups near 2250 cm-' (Fig.1 and Table 2). (b) In aqueous solution of Mg(ClO2-)2 or Ca (ClO2-)2, the Raman spectra of glutaronitrile also show the characteristic band v' in the range 2270'-'2290 cm' (Fig.2). Further, when succinonitrile was added to an aqueous solution of MgCl2 or CaCl2, the concentration of free Mg2or Ca decreases (Fig.3). (c) Similarly, upon addition of succinonitrile to an aqueous solution of Mg(SCN)2 or Ca(SCN)2, the concen- tration of free SCN- decreases (Fig.4).
    From these results, it was inferred that Mg2+ or CO+ can form complexes with these aliphatic dinitriles through the medium of their cyano groups, and that SCN- also can bind to the nitriles. The dissolution of PAN in aqueous solution of Mg(SCN)2 or Ca(SCN)2, therefore, is explained in terms of the complex formation of Mg2 or Ca2 with PAN and also of the binding of SCN- to PAN.
    Download PDF (416K)
  • Yoshio SEOKA, Masato SATOMURA, Akira UMEHARA
    1979 Volume 1979 Issue 10 Pages 1390-1394
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The polymerizations of 4-vinylquinoline (VQ) and 9-vinylacridine (VA) with excess amounts of alkyl halides were found to be initiated under mild conditions with appreciable conversions. Partially quaternized polymers were obtained in the polymerization of VQ with butyl bromide but quaternization did not occur during the polymerization of VA. With AIBN as a radical initiator, the polymerization rate of VQ was rather high but that of VA was extremely small. As in the case of the polymerization of 4-vinylpyridine (VP), the polymerization of VA with alkyl halides is assumed to proceed via an ionic mechanism rather than a free radical mechanism. Quaternization abilities of these polymers obtained with radical initiators decrease in the order of poly-VP, poly-VQ, and poly-VA. The polymerization of VP or VQ with unsaturated alkyl halides indicates the possibility for one-step synthesis of water-soluble photo-sensitive polymers.
    Download PDF (293K)
  • Mitsuaki TANAKA, Taizo UNO
    1979 Volume 1979 Issue 10 Pages 1395-1399
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    An investigation was made of the relation of the charging rate constant (Ks) to the concentration, the apparent molecular distance and the system viscosity for a conducting substance (2-ethylhexyl m-aminobenzoate) which was uniformly dispersed into dodecylbenzene (liquid phase) or ethylcellulose (solid phase).
    The relation of K= AC between IC and the concentration (C) of a conductive substance is obtained, where n is 7.6 for solid phase and 3.6 for liquid phase. IC increases remarkably when the molecular distance is about 9.6A in the solid phase and about 12.0A in the liquid phase. KS in both phases are about 3.3 min-1. IC decreases having a maximum peak in the range of 200 cp with the increases in the viscosity of the system. The IC of soda-lime silicate glass, which is known as a material free of electrification troubles, is 5.8 min-'.
    Download PDF (278K)
  • Mitsuaki TANAKA, Taizo UNO
    1979 Volume 1979 Issue 10 Pages 1400-1403
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Studies were made of the effects of the chemical structure, the basicity and the hydrogen bonding of amino compounds on the charging rate constants,
    The conductivity did not generally depend on the basicity of amino compounds and was especially high for 2-hydroxyethylamino compounds. The results indicate that the intermole- cular hydrogen bonding due to -OH. -N- of 2-hydroxyethylamino groups contributes to the reduction of the intermolecular distance between conducting substances.
    Download PDF (206K)
  • Akio YONEDA, Takatugu AZUMI
    1979 Volume 1979 Issue 10 Pages 1404-1409
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Photochemical reaction of butyl nitrite (BN) with 2-chloropropane (CP) in an air was carried out under irradiating light from 100 W high pressure mercury lamp. The reaction didn't proceed, in a dark place. However, BN decomposed under irradiation and the reaction followed a pseudo-first-order rate law in a similar way as the photolysis of BN alone. CP, which did not decompose in a simple gas photochemically, reacted in the presence of BN. The reaction products were butylaldehyde, 1-butanol, butyl nitrate, and acetone. Acetone was presumed to be formed by photochemical decomposition of CP. In order to obtain the information about this acetone formation, the photochemical reactions among CP and NO, NO2, or O were carried out. And its mechanism was proposed on the basis of the experimental data and known atmospheric reactions. As the results, it was considered that oxygen atom plays a major role. In the photochemical reaction of BN-CP system in nitrogen, the formation of 2-chloro-2-nitrosopropane was assumed to be an intermediate.
    Download PDF (324K)
  • Toshiro MATSUMURA, Akio TANIMURA, Eiji HIGUCHI, Noboru YAMATE, Mamoru ...
    1979 Volume 1979 Issue 10 Pages 1410-1414
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    We have established an analytical method for N-nitrosodimethylamine in air by the Tenax GC adsorption method. Air was sampled in a glass collection tube (8mm in diameter and 16 cm in length) packed with Tenax GC (O.8g) at a flow rate of 0.51/mmn for 40-60min at room temperature. The collection tube was set in a desorption device. The N-nitrosodimethylamine adsorped on the Tenax GC was desorbed at higher temperature in an argon flow (200° C, 40 m//min) and trapped in another glass tube (4mm in diameter and 18cm in length, GC tube) packed with Tenax GC (O.3g) at an argon flow rate of 40m1/min for 5min. Then the GC tube with the amine was set in the carrier gas line of a gas chromatograph, and the GC tube was heated for 1 min from room temperature to 200° C. The N-nitrosodimethylamine was determined with a gas chromatography-thermal energy analyzer. The present method was applied to determine N-nitrosodimethylamine in indoor and urban air. The N-nitrosodimethyl- amine levels found varied from 9 to 560 ng/m3.
    Download PDF (276K)
  • Shigeru TANAKA, Tadashi KUSUBE, Hiroshi IWASHITA, Yoshikazu HASHIMOTO
    1979 Volume 1979 Issue 10 Pages 1415-1419
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A turbidimetric technique using 2-aminoperimidine was investigated for the determination of sulfates in the atmosphere, and applied to the field measurement of sulfates. This method is simpler in procedure and is more sensitive and accurate than the conventional barium sulfate method. A suitable range of measurement of the method was from 0.5 to 5 ppm of sulfate concentrations in the solution, and the lower limit of detection was O.2 ppm (Fig.5). No stabilizing agents were required, since the turbidity due to the sulfate-reagent compound was very stable. The coefficient of variation in the measurements (n=5) was 4.0% for sulfate solution of 1 ppm (Table 1). The recommended procedures were as follows: the ambient aerosols were collected on a Millipore filter (AAWP O2-700), and the sulfates in aerosols were extracted in distilled water, and filtered off. The filtrate was made up to 10 ml and 2 ml of 2-aminoperimidine (0.5 w/v%) was added and the produced turbidity was measured at 600 nm (Fig.1). The diurnal variation of sulfates concentration in ambient aerosols were observed at Hiyoshi, Yokohama-shi, in May, 1978. The coefficeint of variation of the analytical results was less than 10% (Table 3). This method made it possible to measure diurnal variation of sulfate concentrations in the atmosphere. The peak in the concentration of sulfates appeared in the afternoon, and its diurnal variation were similar to that of sulfur dioxide (Fig.6).
    Download PDF (307K)
  • Koji TSUCHIMOTO, Masaomi SUZUKI, Naoomi YAMAKI
    1979 Volume 1979 Issue 10 Pages 1420-1426
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The catalytic activities of various metal carbides such as TiC, Cr3C2C, NbC, Mo2C, TaC, WC etc. for the reduction of nitrogen oxides in automobile exhaust have been investigated. A fixed-bed flow reactor in atmospheric pressure was used.
    A Cr3C2C catalyst showed the highest activity in the reduction of NO with CO, that is, more than 90% NO conversion at 500° C, and a TiC catalyst had the next highest activity (Fig.1). One of transition metals such as Cu, Co, Fe and Ni was added to the Cr, C, and TiC. The catalytic activities of the TiC catalysts were affected by the metal added (Fig.7 and 8), but the activities of the Cr, C, catalysts were hardly affected by the added metal.
    The reduction of NO with a model automobile exhaust gas consisting of NO, CO, hydrocarbons, H2, CO, and H2O has been also studied over the metal carbide and carbide-metal alloy catalysts. The VC, CI-3C2C and TiC catalysts showed high NO conversion (Fig.9), but the amount of NH3 formed was very large for these catalysts. The structure of the TiC catalyst, however, was partially broken by the partial oxidation in the presence of H2O and the activity of the VC catalyst remarkably decreased in the addition of O2- to the model exhaust gas (Fig.10). The activities of the Cr, C, -metal alloy catalysts changed with the added metal (Fig.12). Especially, the CraC, -Ni alloy catalyst revealed a high NO conversion and a very small NH3 formation. In the presence of O.25% O2-, the NH3 formation rather decreased, and the NO conversion remained unchanged (Fig.13).
    Download PDF (423K)
  • Masaaki SUGIURA, Toshio SHINBO, Noriyuki TAKIZAWA
    1979 Volume 1979 Issue 10 Pages 1427-1429
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The uphill transport of thallium( I) and potassium ions through a 1, 2-dichloroethane liquid membrane containing chelating agents such as crown ethers and valinomycin has been studied. The membrane was set to separate two aqueous phases, one (Phase 1) containing 5x 10-4 mol/ thallium or potassium sulfate and 0.05mol// sodium sulfate and the other (Phase 2) containing 5x10-4molg thallium or potassium sulfate and 0. lmolg sodium perchlorate, nitrate or chloride. The thallium and potassium ions transported against the concentration gradient through the membrane and accumulated in Phase 1. The rates of transport of thallium and potassium ions were affected by the type of chelating agents used and counter-ions for sodium ion in Phase 2 and by their concentrations.
    Download PDF (137K)
  • Toru NOZAKI, Koichi NODA, Masatomi SAKAMOTO
    1979 Volume 1979 Issue 10 Pages 1430-1432
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Kinetics of the substitution reactions of lead On complexes of 3, 6-dioxaoctane-1, 8-diamine N, N', N'-tetraacetate with copper(II), cadmium (11) and nickel (II) ions were studied at 20 ° C in the pH range of 3.0-4.8 and at an ionic strength of O.1, in the presence of excess lead(I) ions ((0.05-1.0) x 10-3mol. dm-3). The reverse reactions were kinetically studied in the presence of excess copper(I), cadmium(II) and nickel(I) ions (1.0 x 10-9-1.0 x 10-1mo'. dm-9), respectively, in a similar way to the above study.
    The initial rate equations for the forward and reverse reactions were obtained from the measurement of change in absorbance at wavelengths of 260 and 235 nm by stoppe-flow method. The rate constants for the formation of CuHX-, CdHX- and NiHX- where X4- denotes the completely-deprotonated ligand anion, were estimated to be (9.6 0.5) x 109, (3.4 0.2) x 109 and (2.6 0.7) X 105dm3. mol-lec-, respectively.
    Download PDF (165K)
  • Kiyoshi NARUCHI, Seiji TANAKA, Hiroyuki KOBAYASHI, Kazutoshi YAMADA
    1979 Volume 1979 Issue 10 Pages 1433-1435
    Published: October 10, 1979
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Mixed calcium salt of acrylic and methacrylic acid was prepared by crystallization from an aqueous solution of the acids in an equimolar ratio and calcium hydroxide. On heating the mixed salt in a solid state at 200° C for 2 hr, a-methyleneadipate ( 3 ) and a-methylene-3-methyladipate ( 2) were formed in yields of 10.4, 9.2%, respectively. ( 2) is a dimer of methacrylic acid, and ( 3 ) a dimer resulted from acrylic and methacrylic acid. ( 3 ) was obtained for the first time, and, in a series of investigation by the authors, the structure was elucidated by elementary analysis, mass and NMR spectra. On the contrary, oligomerization of a mechani- cal mixture of salts at 200° C for 2 hr prepared from equimolar amounts of calcium acrylate and calcium methacrylate gave (2) exclusively. From observation by X-ray diffraction, a formation of new crystal structure was confirmed in the mixed salt prepared from an aqueous solution.
    Download PDF (158K)
feedback
Top