NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1978, Issue 4
Displaying 1-30 of 30 articles from this issue
  • Koichi HIRANO
    1978 Volume 1978 Issue 4 Pages 481-488
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A general expression of the solvent-induced UV frequency shift, which can be applied to organic ions, is derived within the reaction field approximation as an extension of the McRae theory. It is concluded that since the charge-induced reaction field at the solute ions is zero, the electrostatic terms in Eq. (13) vanish.
    Consequently, the frequency shift is expressed by such a dispersive term as the one Eq. (48) shows. The parameters Lkt and ekl in Eq. (45), (47) was replaced by the relating quantities just like I and nD, and a semiempirical formula of Eq. (55) was obtained. Application of the formula to Eosin-Y2- anion dissolved in aprotic solvents, shows that Eq. (55) is in good agreement with the experimental results.
    Download PDF (487K)
  • Chizuko SHIBATA, Tsugio TAKEUCHI, Mitsuru YAMAZAKI, Jun NIWA
    1978 Volume 1978 Issue 4 Pages 489-495
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Electronic spectra of 2-furoyltrifluoroacetone (FTA), benzoyltrifluoroacetone (BFA), and 2-thenoyltrifluoroacetone (TTA) were measured in several solvents. Of the three β-diketones studied, FTA was the most sensitive to the solvent effect examined. The spectra of FTA in heptane showed two maxima at 320nm and 345nm, both bands being resolved into two sub-sidary bands (see Fig. 1). The spectra in acetonitrile, on the other hand, two broad bands at 320 nm and 350 nm with no subsidary bands (see Fig. 1). Moreover, the comparison of the relative intensities of two maxima in acetonitrile with those in heptane showed that they were inverted (see Fig. 4). These facts suggest the presence of the equilibrium between two or more tautomers. For the interpretation of the spectra at the low concentrations in inert solvents, equilibria between two cyclic enol structures ([A] and [B]) were postulated for these β-diketones, and spectra of those were reasonably assigned by means of an ASMO SCF-CI calculation (Table 1).
    Download PDF (327K)
  • Koichi TANIHARA, Tetsuro SEIYAMA
    1978 Volume 1978 Issue 4 Pages 496-504
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Settling behavior of three kinds of flocculated bentonite (mainly alkali ion form having swelling property) suspensions, A, B, and C, has been studied. Suspension A was prepared by dispersing bentonite in pure water and subsequent addition of inorganic electrolyte (final concentration : k), suspension B was prepared by dispersing bentonite directly in inorganic electrolyte solution (final concentration : k), and suspension C was prepared by dispersing bentonite in inorganic electrolyte solution (final concentration : α) and subsequent addition of the same inorganic electrolyte (final concentration : b, α+b=k). The results obtained are as follows :
    (1) Differences in flocculation values between A and B (Table 1) for inorganic electrolytes were scarcely observed. (2) The settling rate of B was extremely higher than that of A, but the difference between them was small when bentonite, changed into calcium form, was used.
    (3) The higher the valency and the smaller the hydration radius of cation, the higher was the settling rate of B (Fig. 3).
    (4) The effect of addition of large amounts of anions on the settling rate of B was examined by using the inorganic electrolytes of Nan+Xn- type which dissociate to produce the same amounts of sodium ions in the solution. As a result, the settling rate was found to be in the order : CO32-<OH-<SO42-<NO3-≅/≈Cl-(Fig. 4) .
    (5) The settling curve, C, appeared between those of A and B (Fig. 6). Therefore, the suppression effect of swelling of montmorillonite with dilute inorganic electrolyte solution could be evaluated by the curve, C.
    Download PDF (635K)
  • Akira MIYAMOTO, Yuichi MURAKAMI
    1978 Volume 1978 Issue 4 Pages 505-511
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Computer simulation method has been applied to the analysis of the kinetics of NO-NH3 reaction on metal oxide catalysts. Rate constants of elementary steps in the NO-NH3 reaction have been determined by the least squares method for V2O4 and Fe2O3 catalysts to elucidate the differences in their catalytic properties. The results showed that the adsorption of NH3 is faster on V2O4 than on Fe2O3, whereas the adsorption of NO and the reaction of adsorbed NO with adsorbed hydrogen are faster on Fe2O3 than on V2O4. The effects of surface oxygen on the rates of NO-NH3 and NO-H2 reactions have been studied on the vanadium oxide catalyst. It has been found that the rate of NO-NH3 reaction is markedly enhanced by surface oxygen (V=O species) whereas the rate of NO-H2 reaction is significantly decreased. The promotive effect of the surface oxygen in the former reaction has been shown to arise from the follo-wing two actions of surface oxygen ; the removal of adsorbed hydrogen dissociated from NH3 and the acceleration of the adsorption of NO. On Fe2O3, MnO2, Cr2O3, ZnO and NiO catalysts . (group A catalysts), the activity for NO-NH3 reaction has been found to be parallel to that for NO-H2 reaction. The correlation has been explained in terms of the difference in the rates of the reactions of adsorbed NO with adsorbed hydrogen derived from NH3 or H2 ; the larger the rate constants of the steps, the faster the rates of both NO-H2 and NO-NH3 reac-tions. Further, guides to the improvement of V2O4 and Fe2O3 catalysts for NO-NH3 reaction were given on the basis of the rate constants obtained.
    Download PDF (472K)
  • Akio NISHIJIMA, Toshio SATO, Yoshimichi KIYOZUMI, Minoru KURITA, Hiroy ...
    1978 Volume 1978 Issue 4 Pages 512-516
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The preparation of a-Fe2O3-Al2O3 catalysts for the reduction of NO with NH3 was studied with spherical γ-Al2O3 particles and iron(III) chloride, iron(III) nitrate or iron(III) sulfate solutions. The structure and the surface properties of the catalysts were related to their catalytic activity by means of X-ray powder diffraction, scanning electron microscope and ion microanalyzer.
    A highly dispersed iron catalyst was made in an impregnation of the Al2O3 particles with a iron(III) chloride solution, while iron(III) nitrate and iron(III) sulfate solutions did not give highly dispersed iron catalysts. The FeCl3-Al2O3 catalyst had a high activity for NO reduction because of iron ions highly dispersed (Fig. 3) and probably because of effects of ions (Fig. 2) upon the reaction rate of NO with NH3. The FeCl3-Al2O3 catalyst was, however, very volatile at the reaction temperature for practical use. In order to stabilize the iron ions on the carrier, we calcined the catalyst at high temperatures under various conditions. The activity of the calcined catalyst changed depending on its calcination conditions ; that is, the dispersion states of iron ions were modified by the calcination. A highly dispersed iron catalyst showed a high activity, and a less dispersed one a low activity (Table 2). In the case of a poorly dispersed iron catalyst, iron ions presented in a surface layer of the carrier forming large crystals (Fig. 5).
    An attempt to find an optimum condition to calcinate the catalyst was made and some effective techniques were developed to get a highly dispersed iron catalyst. FeCl3-Al2O3 catalyst should be dried completely before calcination. Then, it will be calcined in an oxygen rich atmosphere in order to prevent the crystallization of α-Fe2O3 particles in the catalyst.
    Download PDF (310K)
  • Tomoyuki INUI, Masaki FUNABIKI, Masatoshi SUEHIRO, Tadashi SEZUME, Tos ...
    1978 Volume 1978 Issue 4 Pages 517-524
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Spherically granulated silica with bimodal macro- and micro-pore stucture was employed as the catalyst support. The content of Ni supported as the substrate of the catalyst was several percent. La2O3 and Ru were incorporated with Ni in atomic ratios of 0. 2 and 0. 1 (for one Ni), respectively. CO and CO2 methanation over these catalysts was investigated by means of a continuous flow method at atmospheric pressure.
    In CO methanation over the Ni catalyst, the selectivity for methane formation decreased due to CO3 formation in higher temperature range where CO conversion was higher. On the other hand, when the Ni-La2O3 catalyst was used, the catalyst was activated at the tempera-ture several ten degrees lower than that for the Ni catalyst. Moreover, the selectivity for methane attained nearly 100% without CO2 formation over the catalyst at higher temperatures (Fig. 1). In CO2 methanation, methane was formed almost selectively, regardless of a variety of catalysts and reaction conditions, and all the catalysts gave an apparent activation energy of 19. 6 kcal/mol (Fig. 3 and Table 4).
    Over the Ni catalyst, the methanation rate of CO was greater than that of CO2 ; however, the reverse order in the rate was obtained with the catalyst containing La2O3, and with the Ru catalyst (Figs. 2 and 6). The activity of the Ni-La2O3-Ru catalyst far exceeded the sum of the activity of each single component catalyst (Fig. 3 and Table 4).
    On the basis of the amount of adsorption of reactants on the catalyst (Table 3) and partial pressure dependences of methanation rates (Figs. 4, 5, and Table 2), the composite effects of these polygenous catalysts were estimated as follows. The rate of CO2 methanation increases as a result of increase of H2 and CO2 uptakes owing to the combination of La2O3 with Ni, and in addition, as a result of the hydrogen-spillover effect consisting of Ni-La2O3 part as the hydrogen acceptor, and of Ru part as its porthole.
    Download PDF (541K)
  • Daijiro YAMASHITA, Yoshifumi YAMAMOTO, Kenji MASUSE, Hiroshi YOSHIDA
    1978 Volume 1978 Issue 4 Pages 525-529
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Effects of hydrogen absorption-desorption on the electrolytic activation of iron electrodes in 4. 7 N KOH solution were studied in order to obtain fundamental data for the preparation of the negative plate of a sintered-type alkaline storage battery.
    The results were summarized as follows : (1) The rest potential was significantly less noble in the electrolytic reduction at -1.30 V and showed a nearly constant potential of -0.990 V when electrolytic reduction was carried out for more than 10 min.
    (2) The current-potential curves and potential-time curves at constant current showed that a new peak probably due to the oxydation of absorbed atomic hydrogen appeared at -0. 85 V on anodic scanning from the hydrogen evolution potential. Iron electrodes activated by the adsorption-desorption of hydrogen exhibited an increased discharge capacity. The electrode potential at the initial stage of discharge decreased with drop in the charging potential. This caused higher discharge capacities.
    ( 3 ) When the iron electrodes were activated by electrolytic reduction, a swelled or disintegrated surface were observed on electronmicrophotograph, but no distinct change in X-ray diffraction pattern was found.
    Download PDF (436K)
  • Hiroshi ISHIKAWA, Eiichi ISHII, Itsuki UEHARA, Masanori NAKANE, Yoshiz ...
    1978 Volume 1978 Issue 4 Pages 530-534
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    As a fundamental study on thermochemical production of hydrogen from water, the hydro-lysis of iron (II) bromide
    3 FeBr2+4 H2O →Fe3O4 + 6 HBr+ H2 (1)
    FeBr2 + H2O → FeO + 2 HBr (2)
    has been investigated. First, solid FeBr2 was hydrolyzed at ca. 490°C according to the equation (1), and the maximum amount of hydrogen in the highest yield was obtained at 620°C. The hydrolysis reaction was carried out by using a fixed bed reactor (25 mmΦ) charged with 120 g of FeBr2 at 620°C. The maximum conversion rate of water was 10% at 30 g/hr of feed rate of H2O, and the rate increased gradually with decreasing feed rate of H2O, i. e., 11% at 20 g/hr and 13% at 10 g/hr. These higher values than that predicted by thermodynamic calculation, i. e., 9. 2%, suggested that sublimed FeBr2 was also hydrolyzed.
    Second, the hydrolysis of molten FeBr2 was carried out by bubbling vaporized water through the molten salt at 700°C. At the initial stage of the reaction, FeBr2 was hydrolyzed according to the equation (2) resulting FeO, and the conversion rates of water were higher than those in solid-gas reaction mentioned above. But continuous run of the reaction was difficult because resulting iron oxide closed the water feed pipe.
    Third, the hydrolysis of vaporized FeBr2 was also carried out in a flow system at 800∼900°C. In this case, FeBr2 was hydrolyzed in terms of the both equation, (1) and (2). The fractions of FeBr2 hydrolyzed according to these competitive reactions depended on a H2O to FeBr2 ratio.
    Download PDF (282K)
  • Naoaki KUMAGAI, Mitsuse UEMURA, Hanzo MASE
    1978 Volume 1978 Issue 4 Pages 535-540
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The hydrolysis of phosphoryl triamide was studied mainly in an acetic acid-acetate buffer solution and an aqueous sodium hydroxide solution by means of ion-exchanged chromatography. General acid-catalysis was effected in the hydrolysis of PO(NH2)3 by acetic acid in acetate buffer, and its catalysis constant was 2. 29 × 10-3 /.mor-1.sec-1 (25°C), while the basic catalysis by acetate ion was not detected. In the buffers of 0. 2, 0. 4 mol.l-1(pH 5∼6), the reaction rate was the first-order function of the concentration, and the first order rate constants for the hydrolysis increased with increasing concentration of acetic acid (Fig. 5). In the buffers of 0. 1 mol.l-1 (pH 4∼6), the rate was proportional to the concentration of hydrogen ions, and the rate constants increased with increasing concentration of the ions (Fig. 6). In the former case, the acid-catalysis worked as a general type one, and the latter as a special type, in spite of the fact that the initial transfer of a protone was supplied from the undissociated acid. The activation energy and entropy for the hydrolysis of PO(NH2)3 in the buffer were 12. 3 kcal . mol-1, -42.0 e. u. (25°C), respectively. In the aqueous solutions of pH 7 to 10 region, induction periods of 6∼24 hr were observed for the hydrolysis of PO (NH2)3 (Fig. 7) . The initial rates of hydrolysis of PO(NH2)3 in the aqueous solutions of sodium hydroxide accorded with a first-order equation, and the rates increased in proportion to the increase of the concentration of sodium hydroxide (Fig. 9).
    The mechanism of hydrolysis of PO(NH2)3 in acetate buffer was discussed through the protolytic mechanism.
    Download PDF (326K)
  • Yoshiharu HISAMATSU, Masaakira IGUCHI
    1978 Volume 1978 Issue 4 Pages 541-548
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    On the absorption reaction of hydrogen by pentacyanocobaltate(II) ion, [Co(CN)5]3-, in aqueous solutions, the effect of various neutral salts has been studied by kinetic and spectroscopic methods. The order of the absorption reaction is 1 for a CN/Co molar ratio of 4 and 5, and 2 for a CN/Co molar ratio of 6 and 8 in the absence of neutral salts, respectively. In the former case, the absorption reaction involves multiple equilibria for the complexation with the formation of precipitates. In the latter case, the reaction is depended on the equilibrium for the formation of dimer as the reacting species with hydrogen. The orders of the absorption reaction are not influenced by added cations. The rate of the absorption reaction increases with the increase in the effective charge on added metal cations and with the increase of the ionic radii of the homologous cations. The apparent rate constants are depended on the concentrations of added salts. It is suggested that added cations promote the dimerization of [Co(CN)5]3-ions by the reduction of the electrostatic force of repulsion between the ions. Lanthanum ions and cesium ions affect differently on the complexation and hydrogen absorption.
    Download PDF (425K)
  • Akio YONEDA, Takatugu Azumi
    1978 Volume 1978 Issue 4 Pages 549-554
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Photodecomposition of butyl nitrite in the air was carried out by irradiating light from 100 W high pressure mercury lamp with the maximum intensity at 365 nm. The decomposition didn't proceed in dark.
    However the concentration of butyl nitrite decreased smoothly with time under irradiation, and 87% of butyl nitrite, of which initial concentration was 6430 ppm, was decomposed after 5 hours irradiation. The reaction followed a pseudo first-order rate law, and the effect of temperature was not observed between 30 and 50°C. The rate of decomposition increased with the decrease of the initial butyl nitrite concentration between 34 and 6430 ppm. No difference in the reaction was observed when oxygen and air was used as dilutant. However the reaction was suppressed by the presence of nitrogen or nitrogen monoxide and deviated from first-order rate law. The reaction products were butyraldehyde, 1-butanol, butyl nitrate, butyl butyrate and acidic products. On the basis of the experimental data and known atmospheric reaction, a mechanistic model was proposed.
    Download PDF (321K)
  • Shigeto KITAHARA, Haruto MURAISHI
    1978 Volume 1978 Issue 4 Pages 555-560
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The kinetics of adsorption of soluble silica on magnesium hydroxide surface was studied. Magnesium hydroxide powder (0.5g) was suspended in a silica solution of 200 cm3 at 30∼80°C. The rate of adsorption was determined by measuring the decrease of silica concentration in the solution as a function of tin. A parabolic law was observed as C=k.t1/2 where C is the amount of adsorbed silica and t is the reaction time. The value of the rate constant k for a commercial magnesium hydroxide in a solution of 114 ppm of the initial silica concentration was 0.15 mmol.g-1.hr-1/2 at 30°C. The value of k depended on the kinds of magnesium hydroxide, the initial silica concentration, and the reaction temperature. The apparent activation energy of about 5 kcal/mol was caluculated from the temperature dependence of the rate constant, and it was independent of the kinds of magnesium hydroxide. The rate of adsorp- tion was increased by the increase in pH of the solution. The presence of Al3+ ions in the solution strongly promoted the rate of adsorption. The adsorption was a chemisorption, and the IR spectrum of silica adsorbed on magnesium hydroxide suggested the formation of the-Mg-O-Si-bond.
    Download PDF (341K)
  • Chozo YOSHIMURA, Yoshinori NODA
    1978 Volume 1978 Issue 4 Pages 561-564
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It has previously reported that interferences of phosphate or fluoride ions is eliminated by dispersing carbon black in the sample solution on the flame atomic absorption spectrophoto- metry of metals. This report deals with the investigation of the deoxidation effect which is considered as a reason of the elimination. Powder of metal oxides (Cu2O, CuO, PbO2, ZnO, ZnO2) were used as samples of direct atomization, dispersed in the sample solutions with the carbon black, atomized and combustioned with a fuel gas.
    The obtained absorbance was compared between the presence of carbon black and activated carbon or 1-butanol and the difference of the deoxidation effect was estimated. By the results, activated carbon and 1-butanol were less effective than carbon black and the difference of the effect were about 40∼50%. But lower absorbance was obtained for the flame atomic absorp- lion spectrophotometry of metal oxide powder by direct atomization with dispersing carbon black than that of general atomic absorptiometry for respective ions.
    Download PDF (210K)
  • Masayo MUROZUMI, Seiji NAKAMURA, Tatsushi IGARASHI, Hiroyuki TSUBOTA
    1978 Volume 1978 Issue 4 Pages 565-570
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Concentrations of copper, cadmium, and lead in sea water can be determined by the isotope dilution-surface ionization mass spectrometry using 65Cu, 116Cd, and 206Pb as a triplicate spike. The fact, that Cu+, Cd+, and Pb+ ionic currents are emitted from the ion source of the mass spectrometer separately at different temperatures, makes possible to measure the intensity of their mass spectra, successively. After being extracted from the sea water sample into chloroform as dithizonates, three elements were treated with both nitric and perchloric acids to synthesize respective nitrates. An aliquot of aqueous solution of these nitrates was loaded onto the centre of a rhenium ionization filament and silica gel and phosphoric acid were added as stabilizers. The present method could detect the amounts of 10-11, 10-13∼10-12, and 10-13∼, 10-12g of copper, cadmium, and lead, respectively. The effective detection limit was 10-9, 10-10, and 10-10g for copper, cadmium, and lead respectively, because the sample suffered contaminations through laboratory works which amounted up to the level of 1.5. ± 0.1, 0.10 ± 0.07, and 0.3 ± 0.02 ng, again respectively. The accuracy for determinations of isotope ratios was 0.3∼2.8, 0.2∼1.9, and 0.2∼1.8% in the coefficient of variation for copper, cadmium, and lead, respectively. Replicate analyses could be performed within the error, 0.6∼8.5, 0.3∼7.5, and 0.0∼2.5% for copper, cadmium, and lead, respectively. In comparison with the method using a single spike, the present method could save analytical time, labor, expense, and the amount of important sample.
    Download PDF (397K)
  • Seisaku INADA, Yoshitoshi SODEYAMA, Mitsuo OKAZAKI
    1978 Volume 1978 Issue 4 Pages 571-577
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The amido-Claisen rearrangement of N-allyl-N-mesyl derivatives of aniline, 1- and 2- naphthylamines, and 2-, 3- and 9-aminophenanthrenes was kinetically investigated. A good linear relationship between In K and the bond localization energy (BLE) was observed in the rearrangement. In the corresponding oxy-Claisen rearrangement, the kinetical data described in literature were similarly correlated with BLE. These results may demonstrate that the reactivities of these aromatic Claisen rearrangements depend on the difference of the resonance energy between the reactant and the transition state.
    The rearrangement of the N-naphthylamides and N-phenanthrylamides, as well as the corresponding oxy-Claisen rearrangement, gave experimentally no position isomers. The pro- babilities of the isomer formation were calculated from the correlated expressions.
    Download PDF (376K)
  • Masayoshi OKANO, Tatsuru KINOSHITA, Yoshihiro MURAMOTO, Takaaki ARATAN ...
    1978 Volume 1978 Issue 4 Pages 578-581
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Chlorination of 1-hexene, cyclohexene, anisole, phenetole, cyclohexane and toluene by using manganese(III) chloride, which was prepared by mixing manganese(III) acetate and calcium chloride in acetic acid solution, was carried out to develop a convenient method for chlorination. Furthermore, bromination was also carried out by using calcium bromide instead of calcium chloride. The yield of halogenation were very good except for a few examples such as chlorination of phenetole, cyclohexane and toluene. This halogenation proceeded in the absence of the catalyst such as light and iron powder.
    Download PDF (262K)
  • Noboru MATSUMURA, Yasuhide MATSUYAMA, Yoshio OTSUJI, Eiji IMOTO
    1978 Volume 1978 Issue 4 Pages 582-586
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of N-(methoxycarbonyl)saccharin [1a] with various nucleophiles proceeded through two reaction paths, A and B: path A involved the attack of nucleophiles to the ring carbonyl carbon to yield sulfonamide derivatives [3a∼e] and the nucleophilic substitution took place in path B at the carbonyl carbon of the methoxycarbonyl group to form carbonic ester derivatives [4 d∼h] along with saccharin C 61. The selectivity for the reaction paths depended on the nature of nucleophiles: ( 1 ) alkoxide and ethanethiolate ions reacted exclu- sively via path A, ( 2 ) amines reacted both via paths A and B, and ( 3 ) phenoxide, benzenethiolate and carboxylate ions reacted via path B. The use of [1 a] and N-(isopropoxycarbonyl)saccharin [1 b] for the synthesis of amide and peptides was examined. The mechanisms for these reactions were discussed in terms of the specific interactions such as a CT interaction between [1 a] and the nucleophiles in solution.
    Download PDF (319K)
  • Yoshishige MURATA
    1978 Volume 1978 Issue 4 Pages 587-591
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Thermal decomposition of maleic anhydride in liquid phase has been studied at 200°C in a closed system under the atomosphere of nitrogen. Carbon dioxide, carbon monoxide, acety- lene, and ethylene were generated in the course of the reaction and the generation rates of these gases were about 5 × 10-2, 5 × 10-3, 8 × 10-5 and 3 × 10-5 mol%/hr, respectively. The kind of these gaseous products was the same as that obtained by the thermal decomposition in gas phase, but the product distribution was very much different. The low polymer was obtained after removing the unreacted maleic anhydride from the reaction mixture and was confirmed to undergo the thermal decomposition to give almost the same amount of CO2 observed in thermal decomposition of maleic anhydride (Fig. 1).
    Taking these facts into consideration, the reaction scheme for the thermal decomposition of maleic anhydride in liquid phase and the structure (Fig. 3) of the low polymer obtained were proposed.
    Download PDF (310K)
  • Toyohiko NAKAJIMA, Toshiko KENJO, Chikato TANOBE, Taro MATSUMOTO
    1978 Volume 1978 Issue 4 Pages 592-596
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A structural investigation has been made of an asphalt-like substance of the Jomon era excavated from the Korekawa ruins in Aomori by means of proton magnetic resonance spectroscopy, measurement of mean molecular weight, elementary analysis and infrared spectroscopy.
    NMR and IR spectra of the substance are very similar to those of heavy fractions of petroleum. The specimen has a relatively high hydrogen content (9.3%) and a relatively low oxygen content (4.7%). These data shows that it has the characteristics of heavy fractions of petroleum rather than those of coal.
    The fundamental structure of the substance consists of the condensed aromatic ring sheets and its condensation form is of an intermediate one of peni- and kata-type. More than 30% of the peripheral protons on the aromatic rings are substituted with paraffinic chains having about 7 carbon atoms. The substance is thus low in aromaticity.
    The substance is presumed to have been in contact with oxygen and moisture in air for a long period of time in the course of which the light fractions volatilized and the remaining heavy fractions were subjected to concentration as well as a series of reactions such as oxida- tion, dehydrogenation, condensation and polymerization. Consequently, the precursor of the substance studied should have been lower in oxygen content, in condensation degree of aro- matic rings and in aromaticity than the substance itself. These speculations further suggest that the substance studied was not derived from coal source but from petroleum source.
    Download PDF (288K)
  • Kenji NEGORO, Masanori YAMADA
    1978 Volume 1978 Issue 4 Pages 597-601
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    [2-(p-Alkylphenylsulfonylamino) ethyl]trimethylammonium chlorides (alkyl: C4H9, C8H17, C10H21, C12H25, C14H20, C16H33) were prepared and identified by elementaly and instrumental analysis. Studies on the surface activities and antimicrobial properties of the samples were carried out according to the ordinary methods. It was proved that the samples, R 10 and R 12, were excellent as dispersing, emulsifying, solubilizing, wetting and antimicrobial agents.
    Download PDF (270K)
  • Kenji NEGORO, Kunio NISHIHARA
    1978 Volume 1978 Issue 4 Pages 602-608
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Sodium 2'-hydroxy-4'-alkoxybenzophenonesulfonates (alkyl: n-C8H17, R8;n-C12H25, R12;n-C16H33, R16) were prepared and identified by elemental and instrumental analysis. Studies on the physico-chemical properties were carried out on aqueous solutions of these samples. From these results, it was found that R 12 and R 16 were fairly active in surface tension depression, as well as in emulsification of liquid paraffin and solubilization of Orange OT into water. Studies on the photofading of Congo Red in the presence of the above samples in aqueous solution were also carried out. Photodegradation of Congo Red occurred as the results of irradiation by a Hg lamp to aqueous Congo Red solution, and the changes of concentrations of Congo Red with elapsed time were measured by spectrophotometric (UV) method.
    Decreasement of comparative specially fixed absorption of Congo Red in the aq. solution of salts of 1-naphthol-4-sulfonic acid, naphthionic acid and above 4 samples addition were compared at the same conditions. Above all, naphthionic acid was the most effective, as it seems that this result was caused by the similar chemical structure between substrate and the additive. The ability of 4 samples as retarder against photofading of Congo Red was as follows: R 4>R 8>R 16>R 12.
    Download PDF (368K)
  • Hideya TSUGE, Masakatsu YONESE, Hiroshi KISHIMOTO
    1978 Volume 1978 Issue 4 Pages 609-613
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Densities of, aqucous solutions of chondroitinsulfate A (ChS-A) were measured on a precision densitometer at 298.15, 310.15 and 323.15 K in the concentration ranging from ca.0.005 to 0.11 monomol/dm3. The counterions used in this work were Li, Na, K, NH4, (CH3)4 N, Ca, Ba and Al.
    The apparent molar volumes, Φ, calculated from the above results were extrapolated to zero concentration to give the partial molar volumes at infinite dilution, The values were found to be nonadditive, which could he explained on the basis of counterion site binding. From the analysis of the fraction of counterion site binding, βSD, was found to he ca.0.30 for alkali metal salts, ca.0.43 for alkali earth metal salts and 0.53 for Al salt. The βSD value was compared with the fraction of ion binding, (1-Φ), derived from the molal osmotic coefficient, Φ. The former was found to he nearly equal to half the latter. From the discussion of βSD for alkali metal salts, it was tentatively proposed that the counterion site binding mainly occurred at the carboxyl groups.
    Download PDF (258K)
  • Masanao OYA, Tomoko TAKAHASHI, Yasuo SUNAGA
    1978 Volume 1978 Issue 4 Pages 614-618
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Polymer blends consisting of a synthetic polypeptide, i. e., poly(L-leucine), poly(L-phenylalanine), poly(L-valine), and poly(γ-benzyl L-glutamate), and a thermoplastic, i. e., polycarbonate, polysulfone, polychloroprene, ethylcellulose and poly (vinyl chloride) have been obtained by polymerization of N-carboxy-eα-amino acid anhydrides in the solution of the thermoplastic. Infrared spectra (Fig.1∼3) of cast film of the polymer blends show the peaks of both the synthetic polypeptide and the thermoplastic. The picture of the blended films observed under microscope and electron micrograph showed that the synthetic polypeptides formed from NCA's were finely dispersed in the thermoplastic.
    Download PDF (535K)
  • Shuzo TOKUNAGA, Kitoshi UEMATSU
    1978 Volume 1978 Issue 4 Pages 619-625
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Calcium silicon proved to be very effective in removing mercury from aqueous solutions of 0.1 to 10000 ppm Hg.
    The mercury removal rate increased with decreasing particle size of calcium silicon, and calcium silicon particles of 250 to 590, μm in diameter yielded the best results. The removal of mercury was conducted satisfactorily in acid solutions, but it is recommended to adjust the pH of the solutions at about 4 to prevent the consumption of calcium silicon by acid.
    Hg2+ ion in the solutions was reduced chiefly by the calcium moiety to Hg2+ ion and Hg which precipitated from the solutions. A part of the Hg volatilized into the air. The volatilization fraction of Hg decreased with increasing mercury concentration and amount of calcium silicon.
    The removal of mercury was also affected by other factors than reduction with calcium silicon. The maximum mercury removal capacity of calcium silicon was as large as 40 m-atom Hg/g calcium silicon, which was much larger than theoretical values calculated from the amount of dissolution of calcium and silicon. Residual mercury concentrations were found to be lower than the solubilities of the reaction products.
    When a calcium silicon packed column was used, the problem of clogging of column due to disintegration of calcium silicon was noticed.
    Download PDF (429K)
  • Hiroshi HASUI, Hisaya MIKI, Michio HASHIMOTO, Naruaki NAKAHARA, Osamu ...
    1978 Volume 1978 Issue 4 Pages 626-633
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Electrolytic technique was examined for the reproduction of Fe (II)-edta complex, formed from Fe(III)-edta complex, the oxidation product of Fe(II)-edta complex, whose aqueous solution was used as an absorbing solution for SO2+NOx simultaneously in a boiler flue gas. First, the voltammetric study showed that the electrolytic reduction took place in the following order: Fe (III)-edta complex, hydrogensulfite ion and Fe(II) (NO)-edta complex, and that the reduction current of Fe(III)-edta was pH-dependent.
    Second? the electrolytic reduction in terns of a battery type electrolyzer, whose cathode and anode chambers were separated by a cation-exchange membrane and, whose anolyte was a dilute sulfuric acid solution brought about the selective reduction of Fe(III)-edta complex at a potential higher than -0.35 V vs. SCE. The stainless steel electrode modified by carbon cloth was found to be a good cathode for the reduction of Fe(III)-edta complex of low concentration, giving rise to the high current density and the high current efficiency. Furthermore, the following relation was found between cathodic current density, Dc (mA/ cm2), and linear velocity of catholyte, U (cm/sec)
    Dc cc U0.53∼0.57
    when using the filter-press type electrolyzer, whose electrode and membrane being separated by the distance, 0.3∼5.0 mm. No potential distribution was detected on the cathode surface. Thus, this type of electrolyzer was found to be effective for the purpose of the present investigation.
    Finally, the continuous operation of the apparatus, in the presence of SO2+NOx for simultaneous removal of SO2+NOx by absorption, for electrolysis of the solution, and for release of absorbed SO2+NOx revealed that Fe(III)-edta complex was selectively reduced on the carbon cloth modified electrode at a potential of -0.30 V vs. SCE and that the absorbed SO2+NOx was quantitatively released as a highly concentrated gas.
    Download PDF (427K)
  • Yuriko NITTA, Yasuaki OKAMOTO, Toshinobu IMANAKA, Shiichiro TERANISHI
    1978 Volume 1978 Issue 4 Pages 634-635
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The hydrogenation of p- and m-substituted acetophenones to the corresponding alcohols was studied in ethanol at 30°C over four kinds of nickel catalysts. For all the catalysts, the Hammett plots of reaction rates showed volcano type curves with the maximum rate at the point corresponding to acetophenone. The decreasing order of specific rates of hydrogenation per unit surface area of the catalysts was Ni-B>D-Ni>R-Ni>Ni-P for each acetophenone, which was consistent with that of electron density of nickel metal. These features can be explained in terms of the adsorption strength of reactants on the catalysts.
    Download PDF (127K)
  • Jae Hee OH, Tatsuru NAMIKAWA, Minoru SATOU
    1978 Volume 1978 Issue 4 Pages 636-638
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The particle growth and its crystallographic orientation in the formation of BaFe, 20, 9 by the thermal decomposition of coprecipitated oxalate of iron and barium with the molar ratio of 11 in terms of Fe2+/Ba2+ were investigated by means of electron microscopy and electron diffractometry.
    When the coprecipitate was heated at 450° C in air, the decomposition product was observed to be well oriented crystalline aggregate keeping the external shape of the coprecipitate. The electron diffraction pattern of the aggregate showed a net pattern corresponding to a-Fe203 and that of the aggregate heat-treated at 550° C showed the net patterns of a-Fe203 and BaFe12019. Therefore the latter aggregate consisted of highly oriented et-Fe203 and BaFe120, 9 phases. The crystallographic relationships between the coprecipitate and its decomposed products were determined as follows;
    [O10]Coprectpitatell[00011a-F0203/0001]BaFet2019
    [100]coprecipito, te//[1010]a-Fe203//[2110]BaFe12019
    [001]Coprecipitate/02101a-Fez03//[0110]
    When the temperature raised to 600° C the aggregate changd to many isolated particles, and converted into the single phase of BaFe12019 at 800° C.
    Download PDF (463K)
  • Toru NOZAKI, Masatomi SAKAMOTO
    1978 Volume 1978 Issue 4 Pages 639-642
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Kinetics of the substitution reactions of lead (II) complexes of ethylenediaminetetraacetate (edta), N-(2-hydroxyethy)ethylenediamine-N, N', N'-triacetate (edtaoh), 1, 2-cyclohexanediaminetetraacetate (cydta), and diethylenetriamine-N, N, N', N", N"-pentaacetate (dtpa) with cobalt (II) ions has been studied in the pH range 2.6∼4.9, at an ionic strength of 0.1, in the presence of excess lead (II) ions ( (0.04∼2.0) × 10-4 mol.dm-8), and at 20°C. The reverse reactions have been kinetically studied in the presence of excess cobalt (II) ions ((0.1∼2.0) × 10-4 mol. dm-8) in a similar way to the above study.
    The initial rate equations for the forward and reverse reactions have been obtained from the measurement of change in absorbance in the region 240∼250 nm. The rate constants for the formation of CoHX(n-s)- where Xn- represents the ligand groups, KCoHX, were estimated; the values of KCoHX were 1.2 × 107, 1.4 × 107 and 4.3 × 107 mol-1 dms2. sec-1 for edta, edtaoh and cydta, respectively.
    Download PDF (218K)
  • Shozo FUJII, Kan INUKAI
    1978 Volume 1978 Issue 4 Pages 643-645
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    This paper presents a preparative method of tetrachloroterephthalonitrile [3] as a precursor of tetrafluoroterephthalonitrile. The method consists of halogenation of 1, 2, 4, 5-tetrachlorobenzene [1] followed by replacement of the introduced halogen atoms (Br or I) by cyano groups.
    In fuming sulfuric acid [1] showed a high reactivity toward iodine and gave 1, 4-diiodotetrachlorobenzene [2b] in 92% yield while reaction of [1] with bromine afforded 1, 4-dibromotetrachlorobenzene [2a] only in 60% yield.
    In aprotic polar solvent such as DMF, [2a] was converted into [3] with copper(I) cyanide in less than 58% yield, the reactivity difference between bromine and chlorine within the molecule being relatively small. With [2b] the iodine atoms were selectively replaced by cyano groups, and [3] was isolated in 88% yield.
    Download PDF (159K)
  • Takeshi KATAOKA, Tadaaki NISHIKI, Masami OKAMOTO, Koretsune UEYAMA
    1978 Volume 1978 Issue 4 Pages 646-647
    Published: April 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The rates of extraction of uranyl(VI) nitrate with trioctylamine in benzene were investigated. Theoretical equations for the extraction rates were derived under the assumption that UO22+, UO2NO32 and UO2(NO3)2 present as uranyl(VI) complexes in the aqueous nitrate phase and that an electric field affects the rates. The extraction rates were calculated numerically.
    It was found that the extraction rates were little affected by the electric field, and could be sufficiently predicted by the theoretical equations neglecting the presence of UO22+ and UO2(NO3)2.
    Download PDF (115K)
feedback
Top