NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1986, Issue 10
Displaying 1-20 of 20 articles from this issue
  • Hisashi UEDAIRA, Hatsuho UEDAIRA
    1986 Volume 1986 Issue 10 Pages 1265-1269
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The spin-lattice relaxation times of 2H and 23Na were measured for aqueous solutions of C6H5SO3Na, p-CH3C6H4SO3Na, m-C6H4(SO3Na)2, 1, 3, 6-C10H5(SO3Na)3, CH3SO3K, C2H5SO3Na, 1, 2-(CH2)2(SO3Na)2, and 1, 4-(CH2)4(SO3Na)2 as a function of the concentration at 25°C. The coordination numbers and the rotational correlation times of water molecule s around the sulfonate ions were estimated. It is shown that the hydration of the aromatic ring is of hydrophobic nature. The value of the rotational correlation time, τc-c0=1.66, for C2H5SO3- ion is the largest and τc-c0=1.08 for 1, 3, 6-C10H5(SO3-)3 ion is the smallest. Th e value of τc-c0 for aromatic sulfonate ions markedly decreases with increasing number of sulfonato groups. It is shown that the sulfonate group has a structure-breaking effect. The value of τc-c0 for CH3SO3- ion is equal to that for 1, 2-(CH2)4(SO3-)2 ion. This result suggests that the microviscosities of water around these two anions are equal to each other.
    The 23Na relaxation rates in the C6H5SO3Na, 1, 4 -(CH2)4(SO3Na)2 and 1, 2 -(CH2)2(SO3Na)2 solutions decrease in this order, and show a behavior similar to that in a 1-1 electrolyte solution. It is suggested from the "Na relaxation rates of m-C6H4(SO3Na)2 and 1, 3, 6-C10H5(SO3Na)3 solutions that these sulfonates form ion pairs, since both anions have small values of τc-c0 and large coordination numbers.
    Download PDF (1242K)
  • Shiro KAMBE, Tomoji KAWAI, Shichio KAWAI
    1986 Volume 1986 Issue 10 Pages 1270-1274
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Raney Ni modified with (R, R)-tartaric acid (MRNi) was loaded on a CdS particle by solution-phase mixing. Irradiation to this CdS/MRNi photocatalyst in a deaerated methyl acetoacetate (MAA)-ethanol solution leads to the formation of acetaldehyde, hydrogen, and optically active methyl 3-hydroxybutyrate (MHB) with an optical yield of 67%.
    It is concluded that the dominant reaction mechanism is not the reduction of MAA by the photocatalytically produced gaseous hydrogen, but the electrochemical reduction of MAA and proton by the photogenerated electron. The hydrogen production site on MRNi seems to be different from the MHB production site, judging from the kinetic behavior of the reaction.
    Download PDF (3772K)
  • Hiroyoshi TANABE, Seitaro FUKUSHIMA
    1986 Volume 1986 Issue 10 Pages 1275-1280
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The interface properties of the silver-sputtered electrode and (Bi2O3)07.7 (Y2O3)0.23 solid electrolyte were studied by means of d. c. and a. c. measurements in the temperature range from 400°C to 700°C and in the oxygen partial pressure region between 8.6 × 10-6 atm and 1 atm using Ar + O2 gas mixtures. In order to obtain the same surface conditions of the electrolyte, the electrolyte surface was presputtered before the silver electrode was prepared by the rf magnetron sputtering technique.
    The electrode process is influenc ed by the oxygen partial pressure. The complex impedance results in a frequency dispersion between 2 Hz and 20 kHz composed of one or two depressed semicircular arcs attributed to a different polarization process. The differential capacitance obtained from the a. c. method varied with potential, temperature, and oxygen partial pressure. These results indicate the presence of the Ag-Bi(Y)-O matching area between the electrode and the electrolyte plays an important role in the electrode reaction. On the partial pressure of oxygen, the low frequency arcs in correspondence to the electrode process at, the Ag(gas)/matching area interface is revealed the dependence of PO2-0.35 by the analysis of the dependence. The migration of O2- along the matching area plays an important role in the electrode kinetics at the lower overpotential. On the other hand, the high frequency arc in correspondence to the behaviour of the matching area shows the dependence of about PO2-0.16 at the temperature below 450°C. The influence of the matching area on the electrode process increases with decreasing temperature and/or increasing overpotential at a given PO2. Probably diffusion of Vo- and in the matching area is the essential process in the electrode kinetics.
    Download PDF (1363K)
  • Katsuya INOUYE, Yoichi OZAWA, Katsumi KANEKO, Sumio OZEKI
    1986 Volume 1986 Issue 10 Pages 1281-1286
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The aqueous mixed solution of FeCl3 and Fe2(SO4)3 has been hydrolyzed at various pH (1.6, 7, and 13), temperatures (30, 50, and 100°C), and mol% of FeCl3/(FeCl3+Fe2(SO4)3)(abbr. M. F. ) (0-100%). The precipitates were examined by X-ra y diffraction, electron microscope, DTA, measurements of specific surface area, and chemical analysis. From acidic solution, jarosite, β-FeO(OH), and their mixture were obtained at 100°C under the ce rtain composition of solution (Fig.1). The crystal form of jarosite in the mixture obtained from the mixed solution of 60-75% in M. F. is thinner than that from the mixed solution containing FeCl3 less than 55% in M. F. (Fig.3). The addition of small amount of Fe2(SO4)3to aqueous FeCl3 solution (95-85% in M. F. ) made β-FeO(OH) crystals needle sh ape from spindle shape (Fig.3) and reduced content of Cl- from c. a.15% to 5% in Cl-/Fe3+ m olar ratio (Table 1). The reduction of Cl--content in β-FeO(OH) crystals led to higher shift of decomposition temperature of β-FeO(OH). The β-FeO(OH) crystals of needle shape was smaller in crystallite size and larger in specific surface area than the β-FeO(OH) crystals of spindle shape. The precipitates from neutral and basic solutions were amorphous hydrous iron oxide and α-FeO(OH), respectively, regardless of the composition of the solutio n and temperature. The effect of coexistence of and SO42- on the hydrolysis of Fe3+ was discussed.
    Download PDF (3411K)
  • Shigeru MATSUDA, Susumu OKAZAK
    1986 Volume 1986 Issue 10 Pages 1287-1291
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The amounts of the surface OH groups on α- and γ-Fe2O3 were determined by exchanging the OH groups with F- or Na+ in an aqueous solution and by the reactions with Al(C2H5)3and some polar organic reagents such as 1-butanol, butylamine and butyric acid.
    The amounts determined by exchanging with F- and Na+ remarkably decreas ed or increased with an increase in pH in both cases, respectively (Figs.2 and 3). The fact showed that the F- and Na+-exchanging methods determined the amounts of the basic or acidic surface OH groups, respectively. On the other hand, the reaction with Al(C2H5)3 is considered to be suitable for the estimation of the amount of the total surface OH groups (Fig.4). The reactions with polar organic reagents are not applicable to estimate the amount of the total OH groups, nevertheless they are useful to measure the amount of the active sites being available for the surface modification using these organic reagents.
    In general, the determination of the exact amount of the surface OH groups on γ-Fe2O3 is difficult because of the high reactivity of γ-Fe2O3 with various reagents used for the determination (Table 2).
    Download PDF (1366K)
  • Kosaku HONNA, Hiroshi ILDA
    1986 Volume 1986 Issue 10 Pages 1292-1297
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Bifunctional catalysts, metals supported on Y zeolite exchanged by rare-earth ion(RE3+), have high catalytic activity for the synthesis of adamantane from tetrahydrodicyclopentadiene (TCD) in the presence of hydrogen and hydrogen chloride. These catalysts were prepared by the impregnation of metal salts with RE-Y zeolite. Their activity was measured by the batch reactor system. The TCD conversion consists of two reactions, the rearrangement of TCD to adamantane and hydrocracking to ring opening products (C10, H18 isomers). No tarry matertials were produced in the TCD conversion over bifunctional catalysts in the presence of hydrogen. The activity of metals supported on RE-Y for hydrocracking of TCD varied in proportion to the %d character of the metals, while for the rearrangement, it was independent of the %d character of the metals (Fig.3). In the case of a binary system of platinum and rhenium, the maximum activity for both reactions was observed at Re/(Re+Pt) =0.25. In the case of a Ni-Pt-Re system, the activity for hydrocracking of TCD decreased with increasing fraction of nickel(Ni/(Ni+Re+Pt)). Furthermore, poisoning by atomic sulfur on a bifunctional catalyst, the activity for the hydrocracking of TCD was depressed. Based on these findings, control of the surface state of platinum (e. g, surface composition)is quite important for increasing the adamantane selectivity.
    Download PDF (1484K)
  • Kosaku HONNA, Kichinari KAWAMURA, Norio KOMORI
    1986 Volume 1986 Issue 10 Pages 1298-1302
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In the conversion of tetrahydrodicyclopentadiene(TCD) with a bifunctional catalyst, the rearrangement of TCD to adamantane(AdH) and hydrocracking to ring cleavage were found to occur. The dependency of selectivity for adamantane on the reaction conditions were investigated with a fixed bed reactor in the presence of (Pt+Re+Co)/RE-Y catalyst, hydrogen and hydrogen chloride. Two reactions, the rearrangement and hydrocracking, proceed simultaneously (Fig.1). Both were first-order reactions with respect to the concentration of TCD. The dependency of the rate constant of the rearrangement on hydrogen pressure was observed to be a negative order(n=-0.13) (Fig.3). As for the dependency on hydrogen chloride concentration, it was observed that the order of the rearrangement (n=0.3) was higher than that of the hydrocracking(n=0.15) (Fig.4). From the Arrhenius plots, the apparent activa. tion energy of the rearrangement (E1=119 kJ·mol-1) was noted to be higher than that of hydrocracking (E2=107 kJ·mol-1).
    Download PDF (1240K)
  • Kosaku HONNA, Hiroshi IIDA, Nobuo SHIMO, Tetsuo ISHIBASHI, Keiji OSADA
    1986 Volume 1986 Issue 10 Pages 1303-1309
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Metal(Pt-Re-Co) catalyst supported on rare-earth(RE)-exchanged Y-zeolite and chlorinated platinum on alumina(Pt-Al2O3) were prepared and rested for the rearrangement reaction. In the course of the reaction, , the metal catalyst supported on RE-Y exhibited little deactivation in the presence of hydrogen(P(H2) =21 atm) with hydrogen chloride(2 mol %) (Fig.8 B). The chlorinated Pt-Al2O3 losts its activity gradually(Fig.4). To study the cause for deactivation of the both catalysts, the physical and chemical properties of deposit on the catalysts were studied by addition of cyclohexane to TCD feed and hydro-treatment of the used catalysts were carreid out. The chlorinated Pt-Al2O3 did not maintain its activity even in the presence of cyclohexane(Fig.6), and the used catalysts did not restore the activity after hydro-treatment(Fig.7). In contrast, the metal catalyst supported on RE-Y retained its activity by the addition of cyclohexane(Fig.9), and the used catalysts were completely regenerated by hydro-treatment(Fig.12). Based on these observations the following are concluded: deactivation of chlorinated Pt-Al2O3 may be due to the formation of stable complex of Lewis acids(surface Al-Cl species) with heavier hydrocarbons, such as the complex of AlCl3 with higher alkyl chloride, and deactivation of metal catalyst supported on RE-Y ma y possibly result from pore plugging with deposited hydrocarbons which can be almost completely regenerated by hydrocracking.
    Download PDF (1779K)
  • Kosaku HONNA, Hiroshi ICHIKAWA, Michio SUGIMOTO
    1986 Volume 1986 Issue 10 Pages 1310-1316
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Dimethyladamantane(DMA) is useful as a synthetic intermediate with two functional groups. Development of heterogeneous catalysts was attempted for DMA synthesis from cyclododecatriene(CDT) by the following reactions: the transannular ring-closure reac tion of CDT(CDT rearrangement) to obtain a mixture of acenaphthene(acenaphthenes: octahydroacenaphthene etc. ), the hydrogenation of acenaphthenes to perhydroacenaphthene(PHA) and PHA rearrangement to DMA. Although the CDT rearrangement proceeds on conventional acid catalysts such as supported polyphosphoric acid, the polymerization of CDT occurs concurrently and deactivation is inevitable. Catalyst screening was thus carried out to find the catalyst which can be easily regenerated. The activities of Y-zeolites, Ca-Na-Y(Fig.5) and La-Na-Y(Fig.6) exceeded those of binary oxides (Fig.1-4) and various sulfates (Table 3), but were nearly equal to that of polyphosphoric acid(Table 2). Y-zeolite catalyst was easily regenerated by oxidation in an air stream. For the PHA rearrangement to adamantanes, a Pt-Re/rare-earth-exchanged Y-zeolite was found to be highly active(Yield 77%). These findings, may provide the way of DMA synthesis from CDT, a raw material derived petrochemical origin, in an industrial scale.
    Download PDF (1791K)
  • Junko MOTONAKA, Hiroshi KONISHI, Sanae IKEDA, Nobuyuki TANAKA
    1986 Volume 1986 Issue 10 Pages 1317-1320
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A rapid method was proposed for the determination of hemoglobin by potentiometric argentimetric titration of iodide ions, liberated in proportion to the amount of hemoglobin at constant temperature when hemoglobin was treated with an iodine-methanol solution, using an iodide-selective electrode.
    The reaction ratios of standard amino acids to iodine, effect of amounts of iodine, effect of temperature, effect of added sulfuric acid, effect of reaction time, effect of concomitant compounds and precision data of hemoglobin were discussed.
    Recommended procedure is as follows. A sample containing 0.076-19 mg hemoglobin was taken into a 50 ml-titration cell and treated with 10 ml of 50 mmol·dm-3 iodine-methanol solution. In addition, 3 ml of 1 mol·dm-3 sulfuric acid was introduced into sample solution. The volume was adjusted with distilled water to 50 ml which was then titrated potentiometrically with a 1-10 mmol·dm-3 standard silver nitrate solution. The temperature was maintained at 35°C through these procedures.
    By addition of sulfuric aci d, potential jump of titration curve was enhanced: Hemoglobin (0.076-19 mg) was determined with a coefficient of variation of less than 3% at 35°C. The calibration curve was shown by the equation of Y=0.276 X-0.011, Y being the added silver nitrate weight in mg and X hemoglobin weight in mg. The coefficient of correlation was 0.999.
    Download PDF (982K)
  • Hiroki YAMANAKA, Takayuki ARAKI, Masaki KUWABARA, Koushi FUKUNISHI, Mo ...
    1986 Volume 1986 Issue 10 Pages 1321-1328
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Preparation of RfCF=CHI [2 a-c] (Rf; a: HCF2, b: H(CF2)3, c: H(CF2)5) by the reactions of RfCF2CH2l [1a-c] with the various bases and the reactions of [2] with nucleophi les, metals, and organornetallic compounds were investigated.
    [2a] was synthesized in a good yield (80%) by the reaction of [1a] with potassium hydroxide in a mixed solvent of water and DMSO (1: 4), and [2b] and [2c] were also ob tained in 83. --89% yields by the reactions of [1b] and [1c] with sodium methoxide in a mixed solvent of ethanol and DMSO (1: 1) (see Table 1). The configuration of [2] thus obt ained was only Z-form.
    [2a] reacted with RONa (R =CH3, C2H5, CF3CH2) at room temperature in DMSO to give 1 H, ωH-1-iodo-2-alkoxyperfluoro-1-alkenes [5a] in 70-73% yields and not to give 1 H, ωH-1-alkoxyperfluoro-1-alkenes. However, the reactions of [2] with RSNa gave only [7], which was brought about by exchanging both of fluorine and iodine bonded to the olefin ic carbons of [2] with two RS groups.
    [2] were treated with two e quivalents of Grignard reagent, followed by adding benzaldehyde to afford the corresponding 2-alkyn-1-ol [9] and similarly, by adding carbon dio xide to give the corresponding carboxylic acid [10] in good yields, respectively.
    The ultrasonic-assisted carboxylation of [2] and [5a] with carbon dioxide in the presence of zinc gave the corresponding polyfluorinated α, β-unsaturated carboxylic acids, [11] and [12a] D, in considerable yields, respectively. This reaction proceeded stereospecifically with retention of the configuration of the starting materials.
    Download PDF (2204K)
  • Kiyoshi KITAHARA, Tadashi MURAKAMI, Takashi MITAMURA, Hisao NISHI
    1986 Volume 1986 Issue 10 Pages 1329-1332
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Starting from diethyl 2, 6-dioxo-1, 4-cyclohexanedicarboxylate [1] three routes to diethyl triphenodithiazine-6, 13-dicarboxylates [4] were examined, and it was found that the following route was the best. The reaction of 2-aminobenzenethiolates [2] with diethyl 3, 6dibromo-1, 4-benzoquinone-2, 5-dicarboxylate [5] which was prepared from [1] gave directly four derivatives of [4] in 35-, 55% yields.[4 a-d] were green needles showing the absorption maxima in the visible region of 560-586 nm. The introduction of ester groups at 6 and 13 positions to triphenodithiazine resulted in a lower melting point and higher solubility in organic solvents as compared with triphenodithiazine.
    Reduction of diethyl triphenodith iazine-6, 13-dicarboxylate [4a] with phenylhydrazine afforded diethyl 7, 14-dihydrotriphenodithiazine-6, 13-dicarboxylate[8].
    Furthermore, diethyl triphenodiselenazine-6, 13- dicarboxylate [11] was synthesized by the reaction of zinc 2-aminobenzeneselenolate [9] with [5] and its reaction mechanism was discussed.
    Download PDF (877K)
  • Tadamasa TERAI
    1986 Volume 1986 Issue 10 Pages 1333-1337
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Grayanotoxins are toxic diterpenoids found in Leucothoe grayana MAX., and are unstable in mineral acids. In this paper, the rate of acetalization and dehydration of G-II [1] and GM-III [2] with some metal sulfates were studied by TLC method. Generally, the rate of dehydration of C-10-OH was faster than that of acetalization of C-5-OH and C-6-OH. The rate of dehydration of C-16-OH was slower than that of acetalization. The optimum conditions for conversion of G-III into its acetal was as follows: metal sulfate, NiSO4; reaction temp., 65.0°C; reaction time, 30 min. The optimum conditions for conversion of G-II into G-II acetal was as follows; metal sulfate, Fe2(SO4)3 or NiSO4; reaction temp., 65.0°C; reaction time, 70-100 min.
    Download PDF (1041K)
  • Yoshio ARAI, Hiroki MATSUDA
    1986 Volume 1986 Issue 10 Pages 1338-1341
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Chromium(III) nitrate enneahydrate was found to initiate polymerization of methyl methacrylate (MMA) in an aqueous solution containing a nonionic surfactant under a nitrogen. atmosphere. The effective surfactants for the polymerization were confined to those as follows: Tween series (except Tween 85) and Brij 35.2, 2-Dipheny1-1-picrylhydrazyl inhibited the polymerization. The copolymer composition curve for the copolymerization of MMA with styrene agreeded with that of an ordinary radical mechanism. It was considered from the above result that the polymerization of MMA proceeded through a radical mechanism.
    Download PDF (1105K)
  • Akihiro NAKA, Yoshihisa NISHIDA, Osamu MURAKAMI, Hiroshi SUGIYAMA
    1986 Volume 1986 Issue 10 Pages 1342-1347
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The highly-loaded coal-water mixture (CWM) suitable for practical use is reguired to have no settling of coal particles as well as to be highly loaded and low viscous. Thus, in this study, anionic and nonionic surface active agents possessing a dispersing power were added to CWM samples and their stabilizing ability was examined, and the following results were obtained.
    1) Sodium salt of formalin-naphthalenesulfonate condensate (NSF-Na) had a great stabilizing effect at a molecular weight of 7500 or more to keep a Saxon Vale coal CWM stable for 14 d.
    2) S odium salt of polystyrenesulfonate (PSTS-Na) had a great stabilizing effect at a molecular weight of 50000 to 120000 to keep a Saxon Vale coal CWM stable for 21 d.
    3) When the pH value of CWM samples was set from 9 to 10 by the combined uses of NSF-Na and NaOH, a Saxon Vale coal CWM remained stable for 28 d.
    4) When the agent produced by the addition polymerization of prop ylene oxide (PO) and ethylene oxide (EO) to polyethylenimine (PEI) was used, its stabilizing effect was higher as its molecular weight per the side chain increased. It kept a Saxon Vale coal CWM stable for 28 d at its best. The ethylene oxide content in its side chain was preferable to be 80% to 90%.
    5) The stabilizing effect of NSF-Na tended to vary depending on the type of coal, while that of the adduct of PEI with PO-EO was invariable with any type of coal.
    Download PDF (1530K)
  • Takeo SAITO
    1986 Volume 1986 Issue 10 Pages 1348-1351
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    This paper describes Cd, Pb, Cu, Ni, Zn, Cr and Mn concentrations in the two well-known pumices (Futatsudake Pumice: FP and Hassaki Pumice: HP) originating from Haruna volcano. Those pumices being the famous Quatenary, Tephras in the Northern Kanto district were collected systematically zone by zone at the outcrops of Mizusawa as FP samples (25)and of Yokomuro as HP samples (10).
    The samples were decomposed b y acid digestion and analyzed by atomic absorption spectrometry with diethyldithiocarbamate-isobutyl methyl ketone extraction for Cd, Pb, Cu, Ni and Zn, and directly with no extraction for Cr and Mn.
    The average concentrations of these elements in FP and HP samples were 0.56+0.18 ppm for Cd, 24.2±6.3 ppm for Pb, 16.3±7.3 ppm for Cu, 37.7±20.6 ppm for Ni, 81.0±12.6 ppm for Zn, 67.0±17.9 ppm for Cr and 978 ±269 ppm for Mn.
    FP was compared with HP in average concentrations and the former showed considerably lower values than HP for Cu, Cr and Mn. The relative values on the average concentrations of FP and HP proved significantly small for Cu, and considerably large for Cd and Zn in comparison with the 35 tephras in the Northern Kanto district. Haruna and Akagi volcanos in contrast to Asama one had smaller relative values for Cu and Ni.
    Download PDF (962K)
  • Takashi SHIRAKASHI, Kazuhisa NAKAYAMA, Kazuo KAKII, Mitsuo KURIYAMA
    1986 Volume 1986 Issue 10 Pages 1352-1356
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The removal of mercury from laboratory wastewater is very difficult, because it frequently contains various substances which stabilize Hg2+ as Hg(II) complexes. In this paper, the removal of mercury from wastewater containing some ligands has been *studied by a Fepowder method. The apparatus used is shown in Fig.1. The effluent from a Fe-column was aerated and both the volatilized mercury (Hg0) and the residual mercury after aeration were measured. The maximum recovery of mercury was obtained at pH 2, but a large amount of Hg0 was detected in the solution passed through the column in the range of pH 1 to 6(Fig.4). In the absence of ligands, 100 mg/l of Hg2+ in the influent was reduced to 33 μgl by the Fe-column treatment and the subsequent aeration. When the effluent from the column was aerated at pH 8, the residual mercury concentration in the filtrate was reduced to 4.4μgl (Fig.7). Even in the presence of any ligand used in our laboratory (Table 1), the recovery of mercury was over 93%, but in the case of some ligands, the residual mercury concentration in the solution after aeration at pH 2 was over the effluent standard value (Table 2). However, the residual mercury concentration in the filtrate was reduced to less than 2 μgl by aeration at pH 8 except for the case of KI or peptone (Table 2). These results led us to the conclusions that the Fe-column can efficiently recover the mercury contained in laboratory wastewater and that a large portion of the mercury remained in the effluent from the column can be removed by the aeration and precipitation at pH 8.
    Download PDF (1345K)
  • Hiroshi SUZUKI, Keisuke OGURO, Akihiko KATO, Yasuaki Osumi
    1986 Volume 1986 Issue 10 Pages 1357-1362
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The hydrogen absorption and desorption characteristics of TiCo0.5Mn0.5M'0.05 and TiCo0.5Fe0.5·M'0.05 (M'= V, Zr) were investigated. These alloys reacted readily with hydrogen to form hydrides such as TiCo0.5 Mn0.5 V 0, 05H1.4, TiCo0.5Mn0.5 Zr0.05H1.6, TiCo0.5Fe0.5V0.05H1.4 and TiCo0.5·Fe0.5 Zr0.05H1.4 (hydrogen content: 1. 3, 1.4, 1.3 and 1.3 wt%, respectively) under hydrogen pressure of 4 MPa at 120°C.
    For TiCo0.5Mn0.5V0.05-H, TiCo0.5Mn0.5V0.05-H, TiCo0.5 Fe0.5 V0.05-H and TiC0.5Fe0.5Zr0.05-H systems, the dissociation pressures at 120°C were 0.2, 0.15, 0.41 and 0.28 MPa, respectively and the enthalphy changes on hydride formation determined from the dissociation isotherms were -41.8, -46.0, -41.8 and -46.0 kJ/mol H2, respectively.
    The pressure-composition isotherm of TiCo0.5 Fe0.5Zr0.05 had a part of plateau, the slope of which was very close to zero. This alloy proved to have suitable property as a heat storage material in a heat pump system.
    Download PDF (1245K)
  • Chozo YOSHIMURA, Naofumi SHINYA
    1986 Volume 1986 Issue 10 Pages 1363-1365
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    This report deals with atomizing effect of carbon black for the direct determination of aluminum silicate by graphite furnace atomic absorption spectrometry, and aluminum in andalusite was determined with this practical method.
    It has been observed that the aluminum absorbance was increased about 1.4 times by addition of 1 w/v% carbon black in aluminum silicate and the sensitivity of aluminum silicate with 1 wfv% carbon black addition was 0.54 ng/1% abs.
    The analytical results of aluminum in andalusite determined with this method was 44.2 %as aluminum oxide.
    Download PDF (503K)
  • Kazuo SUGIYAMA, Yasuharu KIHIRA
    1986 Volume 1986 Issue 10 Pages 1366-1368
    Published: October 10, 1986
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The dielectrical relaxation and electrical conductivity of undoped and iodine doped poly [2-[3-methoxy-4-(methacryloyloxy)pheny1]-3-methyl-4-thiazolidinone] (poly-/) were stu died over a temperature range between 10 and 170°C and the frequency range between 30 Hz and 300 kHz. The conductivities of undoped [poly-1] and iodine-doped sample [poly-1]/I2 were found to be 2.3 x10-14 and 3.1 x 10-13 (S·cm-1), respectively. The values of the relaxation strength Δε of [poly-1] and [poly-1]/I2 were found to be 1.15 (at 100°C) and 12.47 (at 60°C), respectively. As a conclusion, it was found that [poly-1]/I2 shows a ferroelectric property.
    Download PDF (658K)
feedback
Top