NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1985, Issue 12
Displaying 1-25 of 25 articles from this issue
  • Hiroyoshi TANABE, Seitaro FUKUSHIMA
    1985 Volume 1985 Issue 12 Pages 2219-2225
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Nickel-metal (metal: Mo, Co, Fe, V, Ag, etc) oxide film electrodes were prepared by thermal decomposition and by radio frequency magnetron sputtering techniques, and their anodic polarization characteristics were studied in alkaline solution after the preanodization.
    The characteristics were significantly affected by the chemical state of metal oxide formed at the surface of glassy carbon substrate. The amorphous Ni-Mo oxide film electrode prepared by rf magnetron sputtering technique had the highest oxygen evolution activity among electrodes investigated. The NiMoO4 film electrode produced on glassy carbon substrate showed the best anodic properties and stability among the polycrystalline oxide electrodes prepared by thermal decomposition technique.
    A reaction mechanism was proposed for the oxygen evolution on the NiMoO4 film electrode by assuming Langmuir's adsorption.
    It was shown by XPS analysis that a highly oxidized molybdenum ion in the Ni-Mo-O site, which was definitely produced in the Ni-Mo oxide film prepared by rf magnetron sputtering technique and was not found in the film prepared by thermal decomposition technique, played a very important role in the oxygen evolution kinetics.
    Download PDF (1512K)
  • Kohshin TAKAHASHI, Teruhisa KOMURA, Hiroto IMANAGA
    1985 Volume 1985 Issue 12 Pages 2226-2230
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The specific resistance of 3-ethyl-5-[(3-ethyl-2(3H)-benzothiazolylidene)ethylidene]-2thioxo-4-thiazolidinone film is 4.5×109Ωcm in the dark and 4.2×108Ωcm under illumination, respectively. The film, therefore, is intermediate between insulator and semiconductor, and slightly photoconductive. When the thick film of 500 nm on gold is illuminated in methylviologen, FeIII EDTA complex, quinone and [Fe(CN)6]3- aqueous solutions, cathodic photocurrent is obserbed in more positive potential than the reduction potential. Anodic photocurrent, on the other hand, is observed only in more positive potential than the Dxidation potential in FeII EDTA complex, hydroquinone and [Fe(CN)6]4- solutions. The thick film, therefore, is probably a p-type semiconductor with low carrier density and the Hatband potential of 0.4∼0.5 V vs. AgCl/Ag. When the thin film of 50 nm on gold is illuminated in quinhydrone solution, short circuit anodic photocurrent is observed. The anodic photocurrent can be attributed to dye-sensitized photoreaction because the thin film is not semiconductor.
    Download PDF (929K)
  • Akihiro NAKA, Hiroshi SUGIYAMA, Shuichi HONJO, Tomoo SUGIYAMA
    1985 Volume 1985 Issue 12 Pages 2231-2236
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A coal-water slurry (CWS) of high coal content is desirable to burn without dehydration. Many types of phenolic resin multi-branched high molecular weight nonionic surface active agent (PR nonionic SAA) were synthesized and CWS with these surfactants was prepared. From the measurements of their physical properties the following conclusions were obtained:
    1) Using PR nonionic SAA, CWS of high coal content (up to 70%) can be prepared.
    2) Allowable coal content of CWS becomes higher when the phenolic resin has nonyl or dodecyl group, and when it has either 12 or 19 condensed rings.
    3) th1e ability of the synthesized surfactants is largest wh en they have ethylene oxide side chains with molecular weigh of about 3000 per a chain.
    4) Both allowable coal content of CWS and stability of the CWS are higher when PR nonionic SAA is used than when conventional anionic dispersants are used.
    5) PR nonionic SAA has a good dispersibility due to the steric struct ure of the molecule. So, using them, CWS of high coal content can be prepared.
    Download PDF (1043K)
  • Atsushi AOSHIMA, Tatsuo YAMAGUCHI
    1985 Volume 1985 Issue 12 Pages 2237-2245
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The hydrolytic behavior of various heteropolyacids has been studied both on their aqueous solutions and on neutralization. An aqueous solution of 12-molybdophosphoric acids (PMo12)(10 wt% or more) formed precipitates of MoO3⋅H2O, MoO3, or their mixture on heating at 40∼90°C (Figs.1, 2). Such a hydrolytic behavior on heating is affected chiefly by the kind of polyatoms and secondarily by the kind of heteroatoms (Table 1, Fig.3). Heteropolytungstic acids are more stable than heteropolymolybdic acids. The decreasing order of the stability with regard to the kind of heteroatoms among heteropolymolybdic acids is P>Si>Ge. Dawson type structure reveals higher stability in aqueous solutions than Keggin type ones of the same constituent elements. On the other hand, the degree of hydrolysis on neutralization in aqueous solutions is affected more markedly by heteroatoms than by polyatoms (Table 2). Raman spectra (Figs.6, 7) indicate the hydrolytic decomposition of PMo12 on heating is attributed to the formation of 18-molybdodiphosphoric acid (P2Mo18) wit h the precipitation of molybdenum oxides as expressed in Eq. (2) (1'), while 12-tungstophosphoric acid (PW12) is clarified not to form 18-tungstodiphosphoric acid (P2W18)in concentrated aqueous solutions at elevated temperatures. The study of 170 exchange rates using 17O-NMR (Table 3) suggests that a small portion of PW12 dissociates to 11-tungstophosphoric acid (PW11) which can not be converted to P2W18, while PMo12 dissociates to 9-molybdophosphori c acid (PMo9) which dimerizes to P12Mo18. The degree of hydrolysis on neutralization is correlated to the O exchange rate between terminal oxygens of heteropolyanions and dissociated species (Fig.5), and the important role of heteroatoms on the hydrolytic stability on neutralization is discussed in view of the exchange rate. The thermostability of solid heteropolyacids indicates no direct relation with the hydrolytic stabilities in aqueous solutions.
    Download PDF (2050K)
  • Shigeru NAGAHARA, Tsutomu UCHIMOTO, Wataru YANO, Yoshiharu DOI
    1985 Volume 1985 Issue 12 Pages 2246-2254
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Hydroalumination of isoprene with LiAlH4 or monoalkylaluminum hydrides such as LiAlH3(C5H11) and LiAlH3(C8H17) has been investigated in the presence of a catalytic amou n t of [Ti(cp)2Cl2] (cp=η-cyclopentadienyl). NMR analysis of the deuterolysis products with D2O indicated that the main products in the initial stage of isoprene hydroalumination with LiAlH4 ([Isoprene]/[LiAlH4].1.0) were 2-methyl-3-alumino-l-butene [5], 1-alumino-2methy1-2-butene [9], 1, 4-dialumino-2-methylbutane [14] and 4-alumino-3-methy1-1-butene [21] (Tables 1, 2, 4 and 5). On the other hand, hydroaluminati on of isoprene with monoalkylaluminum hydrides selectively gave [5] and 2-methyl-4-alumino-1-butene [6], resulting in selective formation of 2-methyl-1-butene [1] as the hydrolysis product (Table 2 and Fig.5). A kinetic study was carried out to elucidate the mechanism of the isoprene hydroalumination (Fig.4 and Table 6).
    Download PDF (1294K)
  • Hiromu SATAKE, Sanae IKEDA
    1985 Volume 1985 Issue 12 Pages 2255-2259
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Chloride ion in natural water was replaced with solid silver iodate to iodate ion (Cl-+AglO3→AgCl+IO3-), which could be titrated with potassium iodide (IO3-+5I-+6H+→+3I2+3 H2O) in the presence of sulfuric acid using a rotating platinum electrode (2000 rpm)at +O.7 V vs. SCE. Error of less than ±0.5μg was obtained with 7∼36μg of chloride ion in 15 cm3 of sample. Analytical results of chloride ion (1∼202μg cm-3) in river, rain, and well waters by the proposed method showed good agreement with those obtained by the spectrophotometrical JIS method and ion chromatographical method.
    The analytical procedure is as follows. Place exactly 15cm3 of a sample containing chloride ion (less than 2.4 mg⋅dm-3), 30cm3 of methanol and 0.5 cm3 of 0.05 mol⋅dm-3 sulfuric acid in the volumetric flask (50 cm3) and make up the solution to the mark with pure water. After standing for about 5 min at 0°C, add 100 mg of silver iodate and stir for 5 min. Transfer 17∼18 cm3 of the solution into the test tube and immediately centrifuge for about 30 s. Pack cotton fiber loosely on the surface of supernatant solution. Centrifuge once again until the cotton fiber sink down to the bottom of the test tube (about 1 min). Place exactly 10 cm3 of supernatant solution in the titration cell (100 cm3). Add 25cm3 of 5 mol⋅dm-3 sulfuric acid and make up to 50 cm3 with pure water. Titrate the resultant solution with standard potassium iodide solution amperometrically. Determine the blank value (y) from titration value (x) using a calibration curve (1) that is prepared with the solution of known chloride ion concentration in the same way as the sample. Estimate the concentration of chloride ion from the difference between titration and blank values.
    y=0.210+(0.220±0.02) (1/x-1.31)
    Download PDF (1089K)
  • Hiroshi MATSUSHITA, Norihisa ISHIKAWA
    1985 Volume 1985 Issue 12 Pages 2260-2264
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A linear analysis is presented for the simultaneous determination of the mixed dissociation constants and concentrations of weak acids from pH-titration data. The proposed method is based on, for example, the following linear equations in which dissociation constants, concentrations, and activity coefficients of hydrogen ion are unknown constants in acidic regions.
    where KM=mixed dissociation constant of the monobasic acid, x1=aH⋅υ, x0=aH⋅(V+υ)⋅[1+(aH/KM)], υ=volume of the titrant, V=initial volume of acid, aH=activity of the hydrogen ion, cW=concentration of acid, cB=concentration of titrant, and γH=activity coefficient of the hydrogen ion.
    The above equation is solv ed by a combination of a least squares method and an iteration technique, i. e., a successive linear regression analysis. The proposed method has such advantage that the mixed dissociation constants can be determined without the values of activity coefficients of the hydrogen and hydroxide ions, and is applied to the determination of monobasic acid concentration with an uncalibrated glass electrode.
    Download PDF (1202K)
  • Chozo YOSHIMURA, Kiyoshige MIYAMOTO
    1985 Volume 1985 Issue 12 Pages 2265-2270
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The titanometry of perchlorate and chlorate were investigated by the direct and indirect potentiometric titrations using Pt-Ag bimetallic electrodes in N, N-dimethylformamide(DMF). Titanium(III) chloride was slowly oxidized with perchlorate in DMF solution.
    The direct titration of titanium(III) chloride with perchlorate was carried out at a titration rate of 0.1cm3⋅min-1 at 80°C. In the case of the indirect titration, an excess amount of titanium(III) chloride was added to the perchlorate solution and elapsed for 40 min at about 80°C, and then, unreacted titanium(III) chloride was titrated with DMF solution of potassium dichromate potentiometrically.
    It was found that the reaction molar ratio of titanium(III) chloride to perchlorate was 8: 1 by the direct or indirect titration. The reaction molar ratio of titanium(III) chloride to chlorate was 6: 1.
    Mixed DMF soluti o n of perchlorate and chlorate were determined by the proposed method. The total amount of perchlorate and chlorate was determined by the indirect titration and the amount of chlorate was obtained by the direct titration at room temperature.
    Visual determination of the end point for the titration was possible with Methylene Blue or Methyl Red as an indicator. Water less than 10% and hydrochloric acid less than 3% gave no influence on the titration. The determinable lower limit of this method was 7×10-4 mol⋅dm-3 for perchlorate ion and 4×10-4 mol⋅dm-3 for chlorate ion, respect ively.
    Download PDF (1226K)
  • Masakichi IWASE, Takao YOTSUYANAGI, Masaichi NAGAYAMA
    1985 Volume 1985 Issue 12 Pages 2271-2276
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    An 8-quinolinolate extraction method was applied to the determination of iron(III) species in aqueous solution containing polynuclear hydroxoiron(III) complexes and iron(III) hydroxide. The extractable species which can be determined photometrically were found to be the mononuclear and dinuclear ones. Thus, the total concentration of iron(III) existed in the form of the extractable species, [Fe]ext, is given by
    [Fe]ext, = [Fe3+] + [Fe(OH)2+]+2[Fe2(OH)24+] [FeLn]
    where L is sulfate ion or chromate ion (n=1 and 2). Effect of aging on the solubility of Fe(III) species was studied by the extraction method and the solubility product of iron(III) hydroxide Kso aged for 12 mon and 36 mon were determined as
    Kso=[Fe3+][OH-]3=10-39.13±0.11(20°C, I=0) and 10-39.74±0.12(20°C, I= 0), respectively.
    Download PDF (1141K)
  • Masakichi IWASE, Takao YOTSUYANAGI, Masaichi NAGAYAMA
    1985 Volume 1985 Issue 12 Pages 2277-2282
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to elucidate the aging reaction mechanism of hydrolyzed iron(III) solution, behaviour of extractable iron (III) species by 8-quinolinolate extraction method were examined. The rate law for aging reaction (t>7 d) in perchlorate and in nitrate solutions was and that in sulfate and in chromate solutions was where L=SO42-, m=0.26∼0.40 and n= -0.1, and L=CrO42-, m=0∼0.20 and n=0, respectively.
    It was found that the polynuclear aqua hydroxoiron(III) complexes (polymer) formed at the earlier stage of hydrolysis reaction of iron(III) solution diminish on aging.
    The rate law for depolymerization reaction of the polymer was at pH 1.3-2.1, where [Fe]p is the total (atomic) concentration of iron (III) exist as polymers. The order of the magnitude of kp is well correlated with that of stability constant of 1: 1 complex of Fe(III)-L where L=SO42-, Cl, NO3- and C14-.
    The model of hydrolysis of reaction of iron(III) was propos ed.
    Download PDF (1181K)
  • Muneyoshi YAMADA, Tohru KAMO, Jie TANG, Yozo OSHIMA, Akira AMANO
    1985 Volume 1985 Issue 12 Pages 2283-2290
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Low pressure pyrolysis of deuterium-labelled 1-butanethiol (1, 1-d2, 2, 2-d2, 4, 4, 4-d3), 2-butanethiol and 2-methyl-2-propanethiol was carried out at 10-2 Pa and below 1110K (Table 1), in order to investigate the following reaction mechanism proposed for pyrolysis of 1-butanethiol in our previous paper.
    n-C4H9SH→ CH3(CH2)2CH2⋅+⋅SH
    →C2H5CH=CH2+H2S
    CH3(CH2)2CH2⋅→C2H4+C2H5
    C2H5⋅→C2H4+H
    Molecular weights of ethylene and 1-butene from deuterium-labelled 1-butanethiol were measured. The formers were 28, 29, 30 for 1, 1-d2, 28, 30 for 2, 2-d2 and 28, 29, 30for 4, 4, 4-d3 respectively (Table 2). This distribution can be accounted for by 1, 4hydrogen shift of butyl radical in this reaction. The latters were 58 for 1, 1-d2, 57for 2, 2-d2 and 59 for 4, 4, 4-d3 respectively. This distribution suggests that H2S elimination proceeds through withdrawal of mercapto group and β-hydrogen.
    2-Butanethiol decomposed above 900 K (Fig.1) to yield propylene, butenes, methane and hydrogen sulfide (Figs.2 and 3). Compared with other experiments, which had been carried out at higher pressure than the present experiment, the distinctive feature of low pressure pyrolysis of 2-butanethiol consists in the formation of propylene. Estimating the kinetic feature of s-butyl radical based on the reported values, s-butyl radical, once produced, decomposes selectively to yield propylene and methyl radical under these conditions. Therefore, it is suggested that propylene and butenes observed reflect competition of two pathways, i. e. C-S bond cleavage and H2S elimination respectively. The RRK/2 calculation supported this suggestion (Fig.4). It is also noticed that the composition of butenes in this reaction is similar to that of butenes produced by pyrolysis of 2-chlorobutane. This similarity suggests that butenes are produced by the withdrawal of mercapto group and β-hydrogen of 2-butanethiol.
    2-Methyl-2-propaneth iol decomposed above 850 K to yield 2-methylpropene and hydrogen sulfide stoichiometrically. Since 2-methylpropene can be formed through not only t-butyl radical but also H2 Selimination, no differentiation can be made experimentally. The RRKM calculation (Table 3) and the amount of hydrogen sulfide observed, however, seem to suggest that the main reaction is H2S elimination.
    It is concluded that butanethiols decom pose through two pathways. The ratio of two reactions seems to depend on molecular structure, because the proportion of I-12S elimination increases in the order of 1-butanethiol<2-butanethiol<2-methyl-2-propanethiol, as is the same order of H2O elimination in pyrolysis of alcohols and hydrogen halide elimination in pyrolysis of alkyl halides.
    Download PDF (1838K)
  • Satoshi KISHIMOTO, Yasutaka KATSUMA, Tsuneo OKUSHI, Tsuneaki HIRASHIMA
    1985 Volume 1985 Issue 12 Pages 2291-2295
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The effect of p-cyclodextrin on coupling reaction of 1-naphthol in aqueous solution and the property of the inclusion compound have been studied.
    The equilibrium constant for the formation of the inclusion compound was very small (K<3.6×102), and the solubility of the inclusion compound in water was considerably small. The inclusion compound deposited as colloidal solid was comprised of one or two molecules of 1-naphthol with one molecule of β-cyclodextrin, and the pKa value of included 1-naphthol was relatively large (∼12), as compared with the pKa value of free 1naphthol (∼10).
    From th e coupling reaction of 1-naphthol as inclusion compound (88∼96% of the total 1-naphthol) with p-methyl- or p-nitrobenzenediazonium salt, the active species of 1-naphthol was concluded to the included colloidal 1-naphtholate ion (pKa≅12) other than free 1naphtholate ion (pKa≅10). The significant differences in the regioselectivity of both naphtholate ions to 2- and 4-position in the coupling of 1-naphthol were not recognized, that is, the ratios of 2-azo dye were increased only slightly (1∼3%).
    Download PDF (877K)
  • Tadahiro YAMAMOTO, Yukiyasu NAKASHIO, Hideyuki ONISHI, Masayoshi HIROT ...
    1985 Volume 1985 Issue 12 Pages 2296-2301
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Decompositions of di-t-butylperoxide (DTBP) were carried out in nine monosubstituted benzenes and seven normal alkanes at 125°C. And the first order rate constants (kd) were measured in the range of the concentrations from 10-3 to 10-2 mol/l. The values of kd in various solvents were also collected from literatures.
    The kd varied in these solvents: kd in ace tonitrile was 3.8 times greater than that in tetradecane. The kd increased with an increase in the polarity of the solvent and the plot of log kd vs. the polarity parameter Et for each solvent gave a good straight line. The activation entropies (ΔS_??_) of the decompositions in five solvents decreased with an increase in the polarity of the solvent and the complementarity was found between zie ΔS_??_ and the activation energy of the decomposition. Moreover, the straight line with a positive slope was obtained between kd in normal alkanes and reciprocal of the viscosity as well as by the use of Pryor's equation on the viscositydependence. However, the kd in monosubstituted benzenes was independent of the viscosity.
    Thus, it was found that the decomposition rates of DTBP in solvents apparently depend upon the polarity of solvents, and also the decomposition rates in nonpolar solvents decrease on account of the nonpolarity and the viscosity dependence. The kd in corresponding solvents for the decompositions of DTBP, bis(1-methyl-1-phenylethyl)peroxide and azobisisobutyronitrile were plotted and a linear relationship was obtained to each other. These results suggest the similarity in the solvent effect caused by each solvent.
    Download PDF (1224K)
  • Shinzo KOHJIYA, Kazuyuki MAEDA, Yoshito IKUSHIMA, Yoshio ISHIHARA, Shi ...
    1985 Volume 1985 Issue 12 Pages 2302-2309
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Polymerizations of anionic or cationic monomers, i. e., 2-acrylamido-2-methyl-l-propanesulfonic acid (AMPS), sodium salt of AMPS (NaAMPS), and 2-(trimethylammonio) ethyl methacrylate (QDM) and copolymerizations of the three monomers with acrylonitrile (AN)were carried out to synthesize polycations or polyanions (polyelectrolytes). Monomer reactivity ratios for AMPS (M1) and AN (M2) were found: r1=0.175, r2=1.89. The viscosity of all the polymers showed the polyelectrolyte behavior, and satisfied the Fuoss-Strauss equation. Dissociations of sulfonic acid in AMPS and ammonium group in QDM were found perfect in water or methanol, but imperfect (not 100% dissociation) when these ionic groups were incorporated into polymers. The polyion-complex formation between poly(AMPS) and poly(QDM) was followed by the turbidimetric titration, the result of which suggested that the anions and the cations reacted stoichiometrically. When the complexation was carried out in the presence of catalase, the immobilization of the enzyme was achieved. The immobilized catalase showed higher thermal stability. Furthermore, the stability and activity of the immobilized catalase was improved by incorporating a coexistent (albumin, histone or lecithin).
    Download PDF (1637K)
  • Katsutoshi NAGAI, Yoshinaka YAMAGAI
    1985 Volume 1985 Issue 12 Pages 2310-2314
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Synthesis of a conjugated polymer from an alternating copolymer of 1, 3-cyclohexadiene (1, 3-CHD) with a-chloroacrylonitrile (CAN) was investigated. The conjugated polymer was obtained by addition of bromine to the copolymer, followed by pyrolysis: nearly complete bromination was accomplished when the molar ratio of bromine to the cyclohexene ring in the copolymer was chosen at 13 and at 25°C for 48 h in 1, 2-dichloroethane. The brominated copolymer underwent an aromatization of 1, 3-CHD unit and a simultaneous dehydrochlorination of CAN unit in the copolymer by pyrolysis at 250°C for 1 to 2 h. The conjugated polymer was an insoluble, black solid being stable up to ca.350°C.
    Download PDF (995K)
  • Katsumi KANEKO, Sumio OZEKI, Katsuya INOUYE
    1985 Volume 1985 Issue 12 Pages 2315-2322
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Activated carbon fibers (ACF) from different origins were treated with iron(M) salt solutions under the same conditions as α-FeO(OH) (α-treatment) and β-FeO(OH) (β-treatment)were prepared. The ACF pre-oxidized with 6 mol⋅dm-3 HNO3 was also treated with an iron(III) sulfate solution at pH =13.0 and 30°C (α-(ox)-treatment). The NO adsorption characteristics of these iron-treated ACF's were examined by an adsorption isotherm (at 30°C), by a pulse method with 1 ml NO (at 30°C), and by a flow adsorption method with a 300 ppm NO-N2 mixed gas (at 100°C). The treatment of the ACF in iron(III) salt solutions improved its adsorption activity for NO and the degree of improvement depended on the origin of ACF and treating methods. The amount of saturated adsorption (Langmuir b) of the“α-(ox)”treated ACF was several times larger than that of the original ACF. The Langmuir b values for “α-(ox)”treated PAN (polyacrylonitrile) and phenol-resin based ACF reached 150 mg/g. The amount of pulse adsorption of the a-treated ACF were larger than the original ACF by a factor of 7. Both β- and α-(ox)-treatments increased the amount of flow adsorption. The breakthrough time was also increased by the α-(ox)treatment. It is assumed that the increase in the adsorption activity for NO is caused by deposition of fine FeO(OH)-like materials having active surface oxygens which can combine with NO molecules.
    Download PDF (1816K)
  • Tadao SEKINE, Michiko YONEMURA, Hisashi UEDA
    1985 Volume 1985 Issue 12 Pages 2323-2325
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A method for generating V4++ ion on V2O5/silica gel has been studied. Thermal hydrogen reduction method does not yield concentrated V4++ ion in this system. UV reduction of V2O5/silica gel in aqueous systems had been tried. Two conditions were found for production of V4++ on V2O5/silica gel. They are: ( 1 ) An alcohol must be added to the irradiation mixture to make it work as an electron donor. (2) An alkali carboxylate must be added to the mixture to stabilize V4++ ion formed on the silica gel surface.
    Download PDF (470K)
  • Hiroki OHURA, Yoichi KISAKI, Sumio YAMASAKI
    1985 Volume 1985 Issue 12 Pages 2326-2330
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of [Fe(CN)6]3- with BrO- in alkaline solution was studied by means of chemical and UV spectral analysis.
    In the concentration rang e of NaOH from 0.05 to 0.5 mol⋅dm-3 and at the temperature of 75°C, cyanide ligands in [Fe(CN)6]3- were oxidized to OCN- and Fe(III) in [Fe(CN)6]3- was converted to iron(III) hydroxide.
    The following reaction w as proposed as a main equation
    [Fe(CN)6]3-+6BrO-+3 H2O→Fe(OH)3+6OCN-+6Br-+3H+ (1)
    But, the increase of the NaOH concentration above 1 mol⋅dm-3 reduced [Fe(CN)6]3- to [Fe(CN)6]4- by the reaction with hypobromite ion. The reaction did not obey Eq. (1).
    Download PDF (830K)
  • Takehiko MATSUZAKI, Kiyoshi OHSUGA, Yoshihiro SUGI, Yasuo TAKAMI, Juic ...
    1985 Volume 1985 Issue 12 Pages 2331-2334
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Catalytic effects of kaolin on the synthesis of methyl ethers of hydroquinone, catechol, resorcinol, 1- and 2-naphthols with methanol were studied in vapor phase at 210∼330°C. Mono- and dimethyl ethers of hydroquinone were produced in high yield at 230∼250°C, whereas dimethyl ether was decreased at high temperature with increasing by-products: C-methylated derivatives and others. Calcination of kaolin at high temperature diminished the conversion of hydroquinone to ethers, which corresponded to the decrease of hydroxyl group of kaolin. The reaction seems to proceed over protonic acid sites by nucleophilic substitution mechanism.
    For other p henols and naphthols, the O-methylation required higher temperature. Catechol gave only monomethyl ether with the selectivity of 80% at 290°C. However, resorcinol and naphthols gave predominantly C-methylated derivatives. The calcination of kaolin promoted the catalyst activity for catechol and resorcinol, or increased the selectivity of ethers for naphthols.
    Download PDF (754K)
  • Kazuo SUGIYAMA, Tadao NAKAYA
    1985 Volume 1985 Issue 12 Pages 2335-2337
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    New functional polyionenes [6] containing diazacrown ether stractural units along the main chain were prepared from 4, 7, 13, 16, 21, 24-hexaoxa-1, 10-diazabicyclo[8.8.8]hexacosan and α-(1, 3, 2-dioxaphospholan-2-yl)-ω-(1, 3, 2-dioxaphospholan-2-yloxy)poly(oxyethylene) P, P'-dioxides [3] with different degrees of polymerization. Polyionenes [5] were also prepared from the reaction of [3] with 1, 4-diazabicyclo[2.2.2]octane. The uptake of K+ from potassium picrate by [6a] was studied in benzene solution using UV spectroscopy.
    Download PDF (548K)
  • Hiroyuki YAMAMOTO, Ayako NISHIDA
    1985 Volume 1985 Issue 12 Pages 2338-2341
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The photoisomerization of azo dyes in solvents has been investigated using a mercury lamp (400 W) with the interference filters. Azobenzene in N, N-dimethylformamide showed a typical photoisomerization from trans to cis upon irradiation of the ultraviolet light, and from cis to trans upon irradiation of the visible light. In several systems, however, the photoisomerization was from trans to cis upon irradiation of the ultraviolet or visible light, and from cis to trans when kept in the dark.2-(4-Phenylazophenyl)hydrazine-1-sulfonic acid caused a photoconversion in water. The electron-attracting and electron-releasing substituent groups in azobenzene exerted an influence on the photoisomerization, together with the permanent dipole moment of solvent.
    Download PDF (682K)
  • Akio KATO, Akihiko TAKAYAMA, Yukiko MORIMITSU
    1985 Volume 1985 Issue 12 Pages 2342-2344
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Preparation of spherical silver particles was examined by using spray-pyrolysis technique from AgNO3 solution. The solvent was a mixture of H2O and C2H5OH (25-100 vol%). Although Ag particles were produced by pyrolysis above 650°C, the particles became solid and spheikir when the pyrolysis was done at temperatures higher than melting point of silver. The particle sizes were determined mainly by the degree of the splitting of atomized droplets during the heating process.
    Download PDF (1918K)
  • Hideki TATSUMOTO, Takeo HATTORI, Masahiko SUMITA
    1985 Volume 1985 Issue 12 Pages 2345-2347
    Published: December 10, 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    An electric vacuum furnace connected with special cold traps was used to heat samples under vacuum. Firstly water was collected at about 200°C. Subsequently liquid mercury compounds were trapped in the range of 200 to 300°C.
    Mercury concentrations in the liquid mercury compounds were found to be ∼196 mg/lwith an atomic absorption spectrophotometer. According to the dissolution tests carried out with the same kinds of dry batteries, it is found that mercury dissolved out ranges from 0.0016 (for SUM-3) to O.0035 (for SUM-1) mg/l for manganese dry batteries. Alkali batteries give O.0027 (for SUM-3) to O.0047 (for SUM-1) mg/l. The values are lower than 0.005 me of regulation for landfill.
    Download PDF (499K)
  • 1985 Volume 1985 Issue 12 Pages 2348a
    Published: 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Download PDF (68K)
  • 1985 Volume 1985 Issue 12 Pages 2348b
    Published: 1985
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Download PDF (68K)
feedback
Top