NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1983, Issue 5
Displaying 1-27 of 27 articles from this issue
  • Eiji IKADA, Fumio SUMIN, Michio ASHIDA
    1983 Volume 1983 Issue 5 Pages 601-607
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Little is known about the origin of the asymmetric distribution in the Davidson-Cole-type dielectric relaxations of polyhydroxy compounds such as diol and triol. The dielectric properties of the diastereomers of 2, 4-pentanediol were studied to clarify isomeric effects on the asymmetric distribution of relaxation times for liquid diols and triols.2, 4-Pentanediol was separated into the two isomers, racemic (syndiotactic) and meso (isotactic) molecules. The dielectric relaxation in the pure liquid state was measured with a transformer ratio-arm bridge in the frequency range from 23 Hz to 3 MHz. The static dielectric constants of racemic 2, 4pentanediol are equal to those of the meso diol, but they are smaller than those of nonfractionated 2, 4-pentanediol. The relaxation times decrease in the order of syndiotactic, non-fractionated, and isotactic 2, 4-pentanediol. The width of the distribution of relaxation times of non-fractionated 2, 4-pentanediol is not so different from those of the racemic and meso diols as expected from the effect of mixture. It was concluded that the origin of the distribution of relaxation times in polyhydroxy compounds could not be attributed merely to the coexistence of steric isomers in a liquid.
    Download PDF (1614K)
  • Tokisada TAKEDA, Susumu OKAZAKI
    1983 Volume 1983 Issue 5 Pages 608-611
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The adsorption of various amines in aqueous solution on TiO2 was carried out at mainly 30°C with constant stirring for several hours. The adsorptive activity of TiO2 per specific surface area was higher than those of other metal oxides (SiO2, ZrO2, Fe2O3, Al2O3) (Fig.3). The amounts of amines adsorbed on TiO2 were in the order of
    1, 3-Propanediamine>3-Amino-1-propanol>Butylamine>Benzylamine.
    The Langmuir equation was applicable to the adsorption of butylamine on TiO 2 (Fig.6). Th e activation energy was about 16.6 kJ/mol. The amount of butylamine adsorbed roughly corresponded to the amount of surface acid determined with KOH aqueous solution (Fig.7). Th ere was a almost linear relationship between the amount of adsorption and that of effective OH groups on TiO2 calcinated above 400°C (Fig.8). The ratio of the amounts of butylamine adsorbed at higher than 400°C to those of effective surface OH groups was determined to be nearly 1: 1. Thus, the adsorptions were considered to take place not on Lewi s acid sites, b ut on the surface OH groups of TiO2.
    Download PDF (999K)
  • Seiichi KONDO, Tatsuo ISHIKAWA, Hirofumi YAMAUCHI, Hitoshi YAMAOKA, Ei ...
    1983 Volume 1983 Issue 5 Pages 612-617
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The characteristics of the static and dynamic adsorption of cobalt(II) ion on silica gel were studied in order to clarify the adsorption mechanism and to obtain the basic data for the separation of this harmful ion from waste water. At high cobalt(II) ion concentration, one cobalt ion exchanged adout two protons of free silanol groups on the surface. At concentrations below several ppm, it was found that 60 to 90% of cobaltOn ions was adsorbed at pH higher than 6, and that cobalt(II) ions were removed almost completely by dynamic adsorption chromatography. The resolution of these adsorbed ions into water was markedly suppressed by sintering the adsorbent after the adsorption. These results indicate the possibility of efficient and economical separation and fixation of such harmful ions as cobalt(II) ion from waste water.
    Download PDF (1424K)
  • Hiroshi MATSUI, Takeo HISANO, Toshio TERAZAWA
    1983 Volume 1983 Issue 5 Pages 618-625
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The kinetics of ethanol oxidation at a roughened Pt electrode was investigated by a potentiostatic pulse method, and the reaction mechanism was discussed on the basis of the kinetic data obtained. Both current and surface coverage during the constant-potential electrolysis of ethanol did not reach to steady states within 6 minutes from the beginning of the electrolysis. However, the current densities per unit area of the "free" electrode surface reached to the steady state after about 200 seconds. In the potential range of 0.30 to 0.45 V vs. RHE, a Tafel relation was found to hold provided that the steady current densities were adopted. The Tafel slope observed was 0.133 V per decade. The reaction order with respect to the ethanol concentration was 0.98 in the ethanol concentration range of 0.05 to 2 mol. dm-3when the surface coverage, the potential and the pH were constant. Based on these kinetic data, it was concluded that the ethanol oxidation occurred in the following sequence in the potential range of 0.30 to 0.45 V
    C2H5OHsol →r.d.s. C2H5Oad H++ e
    C2H5Oad→CH2CHOsol H+ e where "r. d. s. " denoted the rate-determining step. At the potentials above 0.5 V, the ethanol oxidation was inhibited by acetic acid, which was one of oxidation products.
    Download PDF (2133K)
  • Hiroyuki NAITO, Hisashi TSUCHIDA, Akira TARAHASHI, Yoshibumi NOSHI
    1983 Volume 1983 Issue 5 Pages 626-632
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A new production method of lead( II) oxide by means of the oxygen pressurized wet milling method was proposed. In the method, surfaces of spherical granules (3-5 mm in diameter)of metallic lead were oxidized by the gaseous oxygen in the presence of water at a temperature below about 60°C in the rotary mill. The formation rate of lead(II) oxide was found to be controlled by the diffusion of oxygen in the thin layer of water over the surface of metallic lead granule. Under the conditions of high contact efficiency of these three phases-oxygen, water and lead surfaces-and rapid diffusion of oxygen in the water phase, the formation speed of lead(II) oxide was remarkably increased.
    Three kinds of lead(II) oxides, i. e., massicot (orthorhombic), litharge (tetragonal) and 3 PbO·H2O (tetragonal) were produced depending on the conditions. The particle sizes of these lead( ff ) oxides were much smaller than those of the lead(II) oxides obtained by the dry method. It was also found that these lead( II) oxide were very active in the reaction with CrO3 and showed other excellent properties as are summarized in Table 2.
    Download PDF (5829K)
  • Hisashi TSUCHIDA, Hirobumi NOSHI, Hiroyuki NAITO, Hiroshi OKAYASU, Eii ...
    1983 Volume 1983 Issue 5 Pages 633-638
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Active leadan(II) oxide was formed by the oxidation of spherical granules (3---5 mm in diameter) of metallic lead with pressurized oxygen in the presence of the liquid medium at temperatures below about 60°C in the rotating mill. The effects of liquid medium, initial oxygen pressure, mill rotation speed, lead/liquid ratio, and temperature on the formation rate of the lead( II) oxide were investigated from various standpoints. It was found that water was indispensable for the reaction and the formation rate was the highest when pure water was used as the liquid medium. As the reaction mechanism, the following electrochemical reactions were proposed:
    →PbO (Hydrolysis)
    Pb→Pb2++2e(Anode)
    Pb→1/2O2+HO2+2e→2OH- (Cathode)
    The reaction rate was increased with increasing oxygen pressure, mill rotation speed and lead/liquid ratio, and with decreasing temperature. These effects on the reaction could be explained using the two film theory shown in Fig.3. The reaction rate was found to be first order with respect to the oxygen pressure, and the values of 0.6×10-96.3×10-9(kg-mol·m2/kg·h) were obtained as an overall oxygen transfer coefficient. It was also estimated that the diffusion of oxygen in gas and liquid phases was the rate-determining step in the reaction.
    Download PDF (1262K)
  • Katsumitsu NAKAMURA
    1983 Volume 1983 Issue 5 Pages 639-645
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Boron thin films of 0.1-1.5 pm thickness have been prepared by the pyrolysis of the decaborane (B10H14) at the pressure of molecular flow region (≤10-4 Torr) and the temperature ranged from 350°C to 1200°C. Sapphire and tantalum were used as the substrate at the temperatures above and below 700°C, respectively.
    The crystal structure of the films were examined by means of X-ray and electron diffraction, and the results indicate that the films are amorphous.
    The deposition rate of the boron films was proportionate to decaborane partial pressure. The deposition rate is given as a function of the substrate temperature Ts when Ts≤T*=416°
    D=3.65×10-3·p·exp((Er -Ed)(1/Ts-1/T*)/R)(cm/s)
    where P is the pressure of decaborane, R is the gas constant. The activation energies for condensation Er and decomposition Ed were found to be 1.1 kcal/mol and 40.1 kcal/mol, respectively. When Ts≤T*, the deposition rate is dependent only on the condensation at the constant P and is given as
    D=3.65 × 10-3exp(Er(1/Ts-1/T*)/R)·P(cm/s)
    It was also found that the decaborane was almost completely pyrolised at 416°C.
    Download PDF (3195K)
  • Katsumitsu NAKAMURA
    1983 Volume 1983 Issue 5 Pages 646-652
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Electrical and optical properties of amorphous boron films deposited by pyorolysis of decaborane at the pressure of molecular flow region have been studied.
    Electrical conductivity was measured in the temperature range from 77 K to 1000 K. It is found that the electrical conductivity varies from 3 x 10-5 S·cm-1 at 77 K to 30 S·cm-1 at 1000 K. The activation energy is 1.07 eV over the intrinsic temperature range (700-4000 K). Thermoelectric power is positive over the temperature range investgated from 500 K to 1000 K and the maximum value is about 420μV·deg-1 at 700 K.
    Absorption coefficient, refractive index and extinction coefficient of the boron films were calculated from the measured values of the transmittance and reflectance. The energy gap of indirect allowed transition was estimated to be 1.28± 0.08 eV. In the infrared spectral region the transmittance decreases steeply at about 8 pm and reaches minimum at 12.5 μm, which is attributed to be presence of short range order in the lattice.
    Download PDF (1642K)
  • Kiyoshi MUSHA, Yuji OHASHI, Sunao YAMAZAKI, Shozo TODA, Shigeyuki TANA ...
    1983 Volume 1983 Issue 5 Pages 653-658
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The ultraviolet, infrared and Raman spectra of. a series of oxo- and sulfido-bridged dinuclear molybdenum ( V ) complexes, (a) Mo2O2(μ-O, S) (dtc)2, (b) Mo2O2(μ-O, S) (dtc)2(c) Mo2O2(μ-S)2 (dtc)2 and (d) Mo2O2(μ-O, S)[S2CN(n-Bu)2] were investigated and the crystal structure of (b), which was prepared by the reaction among molybdenum(VI) oxide, sodium hydrogensulfide and dibutyldithiocarbamic acid, was solved by the 3-dimensional X-ray method using MULTAN 78 program. The crystals are triclinic, space group P1 with lattice parameters of a=11.701(2), b=13.549(3), c=9.315(3)Å, 13i, α=101.72(3), β=97.37( 4 ) and γ=87.38(2)o for Z=2. The final R factor is O.13 for 5280 reflections.
    These molybdenum complexes have characteristic and intense UV absorption maxima in the 230-280 nm region (εmax≥3.2 ×104) assigned to π→π* transitions of ligands. The UV methodmax combined with high-performance liquid chromatography is useful for the identification and selective determination of. mixed complexes. The infrared absorption bands for functional groups were characterized as follows:ν(C=N)ca.1530-1, ν(Mo=O)ca. 965 cm-1, ν(Mo=S)ca.545cm-1, ν(Mo_??_Mo) 708, 460, δ(Mo_??_Mo)541, 348cm-1 ν(Mo_??_Mo) 470cm-1, δ(Mo_??_Mo)335cm-1, The Raman lines were: ν(C=N)ca. 1530cm-1, ν(Mo=O)965 cm-1, ν(Mo=S)545cm-1, ν(Mo_??_Mo) 710, 461cm-1, and ν(Mo_??_Mo) 429cm-1The method for the identification of molybdenum complexes based on these characteristic group frequencies was discussed.
    Download PDF (1515K)
  • Chozo YOSHIMURA, Takayosi HUZINO
    1983 Volume 1983 Issue 5 Pages 659-664
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    We have previously reported the elimination of interference of anions and deoxidation effects by the addition of carbon black in sample solutions in flame atomic absorption spectrophotometry, and the effects on absorbance of molybdenum and vanadium by flameless method. This report deals with the enhancing of absorbance and atomizing effect in flameless atomic absorption spectrophotometry of beryllium by the addition of the various powder reductants such as carbon black, metal hydrides (calcium, lithium and sodium tetrahydroborate) and sulfur.
    By dispersing carbon black in the aqueous solution, the absorbance increased about twice, and by using of N, N-dimethyl formamide solution dispersed carbon black and metal hydrides and absorbance increased by about seven times that of without powder reductants. By the results of the investigation of relationship between the absorbance and amount of powder reductants, it was found that the effective amount was about 1-2% as carbon black, 0.1%as hydrides and O.01% as sulfur.
    It apears that the enhancement of absorbance is due to the deoxidation effects of powder reductants and suppressed vaporization and activated decomposition of metal salts.
    Download PDF (1336K)
  • Seiki YAMAGUCHI, Tsutako KONISHI, Tsutomu TSUKAMOTO, Mitsugi SENDA
    1983 Volume 1983 Issue 5 Pages 665-670
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In cathodic scan pulse polarography (cspp) with the initial potential (E') of 0 V vs. SCE, a 10-4molidm3 a-lipoic acid (1, 2-dithiolane-3-pentanoic acid) solution buffered at pH 4.5 gave a cathodic wave with a maximum at -0, 495 V which corresponds to the DC polarographic wave of a-lipoic acid. And two small, ill-defined and peak-shaped waves were produced at more negative potentials between -1.1 to -1.2 V. In anodic scan pulse polarography (aspp)with E1 of -1.5 V, α-lipoic acid gave a cathodic wave at the potential near -1.1 V and a S-shaped cathodic-anodic wave at E1/2=-0.452 V. In cspp (with E1 =0 V), dihydrolipoic acid (6, 8-dimercaptooctanoic acid) at concentrations lower than 10-4mol/dm3 gave an anodic-cathodic wave at E1/2=0.475 V (cspp- I ) and a small, round-maximum wave with the peak potential (Ep) of -1.27 V (cspp-II). With increasing dihydrolipoic acid concentration, the cspp- I wave became a peak-shaped wave with the peak height approaching a saturation value, and at concentrations higher than 3 × 10-4mol/dm3, an additional peak-shaped wave was produced at Ep=-0.394 V (cspp-III). In aspp (E1=-1.5 V) dihydrolipoic acid gave a small wave at near -1.1 V and a S-shaped anodic wave at E1/2=0.453 V. At concentrations lower than 10-4mol/dm3, the behavior of a-lipoic acid essentially agreed with that of dihydrolipoic acid.
    Download PDF (1655K)
  • Toshihiko HOSHI, Hisashi ITO, Jun OKUBO, Shoji MORI
    1983 Volume 1983 Issue 5 Pages 671-677
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The electronic absorption spectra of N, N'-diphenylcarbodiimide (DPC) and N-methyl-N'phenylcarbodiimide (MPC) were measured in cyclohexane and ethanol, and MINDO/2 calculations were performed for these compounds. It has been clarified that the structure of DPC is linear with the two benzene rings orthogonal to each other. The polarized absorption spectra of DPC and MPC were measured by means of a stretched polymer film technique, and the polarization of each electronic transition was determined. From the above experimental results together with PPP calculations extended to the 3-dimensional system, the nature of each electronic transition was clarified. The 284 nm and 202 nm electronic bands of DPC are polarized along the short molecular axis. The former band is interpreted as an intramolecular charge transfer transition from the N=C=N group to the phenyl groups, and the latter as the back charge transfer one.
    Download PDF (1635K)
  • Kazunobu HASHIMOTO, Muneyoshi YAMADA
    1983 Volume 1983 Issue 5 Pages 678-684
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The parallel paths (1) and (2) observed for the initial reaction between thiols and hydrogen atom have been investigated theoretically by the MNDO MO method.
    RSH+H→R+H2S(1)
    →RS+H2 (2)
    Calculations were performed on the simplified reaction system H2S + H. Pyramidal complex model ( I ) was assumed. Linear complex model (II) was also taken into account for reaction ( 2 ) (Figs.1 and 2). H-S-H-H(I)H-S-H-H (II)
    Three characteristic features of the potential energy surface emerge in I as the attacking hydrogen atom approaches the sulfur atom. Firstly, the reaction system crosses a low energy barrier (4.6 kJ/mol), then, goes down to a deep potential well, and again goes over a high energy barrier (68.6 kilmol). The first barrier corresponds to the initial approach of the attacking hydrogen atom to the sulfur atom (Figs.3 and 4). The deep potential well corresponds to trivalent intermediate H3S, which has Csymmetry and is 48.5 kJ/mol more stable than ELS+ H (Fig.5). The stability of H3S is estimated to be comparable to that of H4S (Fig.6). The potential energy change due to the abstraction of hydrogen atom from H3S is the same as that induced by the approach of hydrogen atom to H2S (Fig.7). The second barrier corresponds to the molecular hydrogen abstraction from H3S (Figs.8 and 9). The potential energy surface via II is found to be more simple than that of I. The hydrogen atom of hydrogen sulfide is abstracted by the attacking hydrogen atom with the barrier of 19.7 kJ/mol (Figs.10, 11 and 12). The linear complex was found to be the most stable conformation (Figs.13 and 14). Reactions ( 1 ) and ( 2 ) are concluded to take place via I and II, respectively (Fig.15).
    Download PDF (1700K)
  • Miyoko KASHIYAE, Suehiko YOSHITOMI
    1983 Volume 1983 Issue 5 Pages 685-689
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The oxidation of 1-methylnaphthalene by cobalt(III) acetate has been carried out in acetic acid and in several mixtures of carboxylic acids in the absence of oxygen. The rate of oxidation in acetic acid increased fairly by adding a small amount of a strong acid, and the reduction rate of cobalt(III) acetate was related to pKa, value of the acid added (Table 3). In the other acids with similar pKa value to acetic acid, the rate of oxidation was smilar to that in acetic acid, and the kinds of products were the same as those in acetic acid. The molar ratios of 1-naphthylmethyl carboxylates formed as the main products were comparable to those of carboxylic acids in the solvent used (Table 5). These facts suggest that the acetate is produced by the reaction with acetate anion in the solvent and not with acetate ligands of cobalt (III) acetate. In acetonitrile-acetic acid mixtures, the reduction rate of cobalt (III) acetate decreased with increasing acetonitrile (Table 1), that is due to the substitution of the ligand of cobalt(III) ion into nitrile which has strong coordination ability (Fig.1).
    Download PDF (973K)
  • Nobuyuki SUTOH, Munehiko KATO, Touru IKEGAMI, Akitsugu OKUWAKI, Akira ...
    1983 Volume 1983 Issue 5 Pages 690-696
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Oxidation of Yallourn coal, consisting 66.8% carbon on dryash-free basis, was carried out using a conventional 300 cm3-autoclave under reaction conditions varied in the following ranges reaction temperature, 225-300°C (Fig.1); concentration of NaOH, 15-, 35 mol/kg-H2O(Fig.2); O2S pressure, 10-50 kg/cm2 (Figs.3 and 4); revolution of magnetically driven stirrer, 400-1200 rpm (Figs.5 and 6); and initial coal/NaOH weight ratio, 0.4 (Fig.7). The reaction afforded CO2S, oxalic acid, benzenepolycarboxylic acids (BPCA), and water-insoluble acids. Maximal oxalic acid yield of 88.5 wt%, corresponding to 42.3% on the basis of carbon recovery, was attained when the oxidation was undertaken at 250°C for 2 h under the following set of reaction conditions: concentration of NaOH, 25 mol/kg-H2O; O2S pressure, 50 kg/cm2, revolution of stirrer, 1200 rpm; and coal/NaOH ratio, 0.2 (Fig.3). It was revealed, throughout the course of the oxidation, that the fraction of coal converted to oxalic acid was nearly equal to that converted to CO2S (Fig.10). It was also =revealed that relative abundance of BPCA's decreased in the order: tri->tetra->di- pentacarboxylic acids (Table 4). The oxidation took place only slowly at 225°C. Upon raising the temperature above 250°C, however, the oxidation to produce oxalic acid and CO2S became significantly faster thereby emerging globular suspensions composed of coal, water-insoluble acids, Na2C2O4 and Na2CO3. To explain this oxidation at high temperature a reaction model, in which the diffusion of dissolved O2S in the globular body plays rate-determining role, is proposed. The model is compatible with the effects of O2S pressure as well as mechanical agitation (Fig.8). Both products yield and composition of BPCA were compared with those previously reported on nitrate oxidation of Yallourn coal (Fig.11, Table 5).
    Download PDF (1622K)
  • Yasuo Mixi, Yoshikazu SUGIMOTO
    1983 Volume 1983 Issue 5 Pages 697-703
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Fused-silica capillary chromatography combined with mass spectrometry has been used for the qualitative and quantitative analysis of anthracene oil. The analytical system consists of Hewlett-Packard 5880 gas chromatograph with a SP-2100 or OV-1 capillary column (50 m X 0.2 mm) and a Hewlett-Packard 5992 mass spectrometer. Standard mixtures were obtained by alkylation or hydrogenation of pure polynuclear aromatics. The gas chromatogram by an FID-detector is given in Fig.1 and identification and quantitation of the peaks are listed in Table 2. More than 140 components were separated and identified. Nitrogen-containing compounds were analysed by using an NP-detector. The results are shown in Fig.2 and Table 3. The acidic compounds in anthracene oil were also analysed to give the results shown in Fig.3 and Table 4.
    Download PDF (1470K)
  • Yasuo MIKI, Yoshikazu SUGIMOTO
    1983 Volume 1983 Issue 5 Pages 704-712
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The qualitative and quantitative analyses of hydrogenated anthracene oil have been carried out on the analytical system consisting of a Hewlett-Packard 5880 gas chromatograph with a SP-2100 fused-silica capillary column (50 x 0.2 mm) and a 5992 mass spectrometer. The chromatographic operating conditions are summarized in Table 1. Thirty five polynuclear aromatics, the main components of anthracene oil, were hydrogenated and used as the standard samples for the gas chromatographic analsis of hydrogenated anthracene oil. The gas chromatogram of a mixture obtained by hydrogenation of pyrene and the identified molecular structures are shown in Fig.1. Perhydro-compounds and partially hydrogenated aromatics were obtained and separated. A similar chromatogram was obtained from each hydrogenated mixture of standard aromatics. The gas chromatogram of hydrogenated anthracene oil is shown in Fig.4. The chromatogram was analysed in reference to those of the standard mixtures and of anthracene oil and by comparison with the GC-MS analysis. The results are listed in Table 2. Out of 335 compounds, 284 were separated chromatographically.
    Download PDF (1604K)
  • Katsutoshi NAGAI, Akira HARADA, Masahiro OYAMADA
    1983 Volume 1983 Issue 5 Pages 713-718
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The synthesis of poly(keto sulfoxide)s by the reaction of polymers containing ester groups, poly(methyl methacrylate) (PMMA) and MMA-styrene copolymer (MSC), with sodium methylsulfinylmethanide (CH3SOCH2Na) in dimethyl sulfoxide and the subsequent acidification was investigated. The structure of the resulting poly(keto sulfoxide)s was determined by infrared and 'H-nuclear magnetic resonance spectroscopies and elemental analyses.
    The reaction with PMMA was carried out at a molar ratio of CH, SOCH, Na/COOCH3 group =2.3 and 4.1 at 25 and 40°C. The resulting poly(keto sulfoxide)s contained β-keto sulfoxide sidceh a(-CO-CO-CHSOCH3) [3] as well as β, β'-diketo sulfoxide side chains(-CO-CO-CHSOCH3) [4]. The latter groups resulted from an attack of the a-keto-a-sulfinylmethanide anion to neighboring ester groups. Poly (keto sulfoxide) with a composition of [3]/[4] =2.0, containing no ester groups, was obtained by the reaction at 40°C for 18 h at a molar ratio of CH3SOCH2Na/COOCH3 group =4.1. The results obtained in the present study showed that the ratio of [4]/[3] was increased to a greater extent by a rise in the reaction temperature rather than an increase in molar ratio of CH3SOCH2Na/COOCH3 group.
    The reaction with MSC having no sequential MMA units (MMA content, 28.4 mol%) gave a product which ester groups of the copolymer were completely converted to only 8-keto sulfoxide side chains [3].
    Download PDF (1366K)
  • Akira KAI, Tomoki KOSEKI, Yasuji KOBAYASHI
    1983 Volume 1983 Issue 5 Pages 719-723
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Relation between change in the morphology and the crystallization of cellulose gels produced by Acetobacter xyilnum has been studied from the aspects of the resistance of the never-airdried cellulose gels to alkali. The yield of cellulose was determined as the function of time of incubation and the observation of fibrils was carried out by an electron microscope. The never-air-dried cellulose gels from the glucose medium incubation for 5 hours at 28°C consisted of about 30 wt% of microscopic fibrils and about 70 wt% of fibrils without microscopic fibrils. The fraction of microscopic fibrils increased to 80 wt% when incubation was continued for 12 hours. The never-air-dried cellulose gels from the complex medium incubation in exponential bacteria growth stage consisted of about 40 wt% of fibrils without microscopic fibrils and about 60 wt% of microscopic fibrils and these values were independent of the incubation period.
    X-ray diffraction and IR results showed that, when the concentration of NaOH was higher than 11 wt%, the resistance of the never-air-dried cellulose gels from glucose and complex medium to alkali was equivalent to that of purified and dried sample. However, when the concentration of NaOH was lower than 11 wt%, the absorbance of the 1430 cm-1 band in IR spectrum of the alkali-treated sample made from the never-air-dried cellulose gels of the short incubation time was lower than that of the purified and dried sample. X-ray photographs of all the never-air-dried samples treated with 11 wt% NaOH showed the diffraction of Cellulose I but not that of Cellulose II.
    It was concluded that the microscopic fibril formed from the initial fibril later had Cellulose I crystal already. It is assumed that the fibrils before transforming into microscopic fibrils have Cellulose I crystals already or the structure which is easily to be converted into Cellulose I crystals.
    Download PDF (3050K)
  • Makoto SHIRAISHI, Shunichiro HORIO, Kentaro TOYOSHIMA
    1983 Volume 1983 Issue 5 Pages 724-730
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Relationships between defoaming rate of aqueous poly(vinyl alcohol) solution and preparation methods of poly(vinyl alcohol) were studied.
    Poly(vinyl acetate) prepared by emulsion or suspension polymerization is more superior in the defoaming properties than that prepared by solution polymerization in methanol. Poly(vinyl alcohol)s which have the broader composition distribution of residual acetoxyl groups and/or block copolymers of vinyl acetate and vinyl alcohol had low foaming stability. Defoamation rate of aqueous poly(vinyl alcohol) solutions increased greatly by adding a small amount of less hydrolyzed poly(vinyl alcohol). It is concluded that, to increase the defoaming ability of aqueous poly(vinyl alcohol) solutions, it is necessary to lower the crystllinity of poly (vinyl alcohol) s.
    Download PDF (1631K)
  • Michiaki MATSUMOTO, Kazuo KONDO, Fumiyuki NAKASHIO
    1983 Volume 1983 Issue 5 Pages 731-737
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to investigate the behavior of liquid surfactant membranes in a stirred tank, the membranes containing chelating reagents were used to concentrate copper(11). At the same time, the water permeation through the membranes, the breakdown of W/O emulsion drop and the entrainment of external aqueous solution into the emulsion were studied by using Ni (II) and Mn(II) as the tracers, which were not extracted by the chelating reagents. The following results were obtained:
    (1) The efficiency of separation and the concentration of copper(II) ion were influenced by the water permeation through the membranes, the breakdown of the membranes and the entrainment of the external aqueous solution. However, it was found that the entrainment in the stirred tank was negligibly small.
    ( 2 ) The behavior of water permeation through the membranes was influenced by the species of chelating reagents used as the carrier. In the cases of benzoylacetone and (E)-2hydroxy-5-nonylbenzophenone oxime, water permeated through the membranes from the external aqueous phase into the internal phase, and the efficiency of the concentration of cupric ion was lowered. On the other hand, in the case of N-8-quinolyl-p-dodecylbenzenesulfonamide(henceforth DBSAQ), the direction of water permeation through the membranes changed during the experiment, that is, in the initial step, water permeated from the external to the internal phase, after that, water permeated from the internal into the external aqueous phase. This change was explained by the difference of permeation rate of hydrochloric acid through the membranes. Different from benzoylacetone and (E)-2-hydroxy-5-nonylbenzophenone oxime, DBSAQ reacts with hydrochloric acid to form the product which is soluble in the organic phase. As the results, DBSAQ plays the role of a carrier for hydrochloric acid.
    Download PDF (1674K)
  • Akira OKU, Hiroyuki UEDA, Hiroaki TAMATANI, Hiroshi TAKAI
    1983 Volume 1983 Issue 5 Pages 738-742
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Chlorine atoms in octachloronaphthalene (OCN) were almost completely removed as chloride ion when OCN was treated with Na dihydronaphthylide (Na+C10H8-[I]) in THF. The minimum amount of[1] necessary to remove all the chlorine atoms was 1.2 mol/Cl (dechlorination efficiency: 99.1-99.8%). Other polychlorinated naphthalenes (technical grade) were also dechlorinated by the same procedure. When OCN was treated with a deficient amount of either C I J or K dihydronaphthylide (0.5 mol/Cl), the dechlorination efficiency exceeded 50% and reached 68-76% which could not be explained stoichiometrically. Comparing the reaction by unsolvated hydroxide ion with that by hydrated one, this was ascribed to the nucleophilic displacement reaction due to unsolvated (activated) hydroxide ion which was formed during the preparation ofCIin THF. Analysis of the products from the nucleophilic displacement of OCN is also described.
    Download PDF (1319K)
  • Masatoshi YAMADA, Akio FURUTA, Hiroo MATSUOKA
    1983 Volume 1983 Issue 5 Pages 743-745
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The fine structure of γ-alumina prepared by different procedures was observed by electron microscope. The specimens were prepared by ultra thin sectioning method to preserve their fine structure.
    It has been found that the method of preparation of r-alumina affects its fine structure significantly. JRC-ALO-1 alumina is composed of two kinds of crystallites: filamentous and columnar ones in outer and inner parts, respectively. The shape of their aggregates is the same as that of its raw material (Fig.1). JRC-ALO-4 alumina has a simple texture. Most crystallites have same size and shape (Fig.4). JRC-ALO-5 alumina is spherical granule, which consists of filamentous r-alumina covered with well crystallized r-alumina (Fig.5).
    Download PDF (3551K)
  • Shigeru OKADA, Tetsuzo ATODA
    1983 Volume 1983 Issue 5 Pages 746-748
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The single crystals of Cr5Si3, CrSi and CrSi2 were prepared by the reaction between elementary chromium and silicon in molten tin under an argon atmosphere.
    The conditions for the preparation are as follows: temperature of molten tin, 1400°C; reaction time, 10 h; mixing atomic ratios of raw materials, Si/Cr =- O.6 and Sn/Cr =6.4 for Cr5Si3, Si/Cr=1.0 and Sn/Cr =6.4 for CrSi, and Si/Cr=2.0 and Sn/Cr=6.4 for CrSi2. Under these conditions, the Cr5Si3 single crystals of polyhedral shape and of needle shape extending to‹100› direction, the CrSi single crystal of polyhedral shape and the CrSi2 single crystal of hexagonal prismatic shape extending to ‹0001› direction were obtained as a single phase, respectively. Each crystal exhibited grayish metallic appearance. Lattice constants measured are as follows: c0=9, 155± O.002Å and ao=4.636+ 0.003Å for Cr5Si3, a0 =4.626±0.002Å for CrSi, and c0 =4.426 ± 0.002Å and c0=6.375± 0.002Å for CrSi2.
    Download PDF (2791K)
  • Hiroshi MATSUSHITA, Norihisa ISHIKAWA
    1983 Volume 1983 Issue 5 Pages 749-752
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A simple potentiometric determination of multivalent ions for which no suitable ion-selective electrodes are available is proposed.
    When a sample solution containing an analyte ion(N) is added to the ML-chelate solution in which a M-selective electrode is immersed, where M and L denote a metal ion and chelateforming ligand respectively, M is released by displacement reaction. N is determined by use of the calibration curve, i. e., the electrode potential vs. the negative logarithm of analyte concentration (pcx) plots. According to the theoretical considerations on the linearity of the calibration curve, the lower limit of the linear range decreases as IC(ML) increases and the upper limit increases as the ratio of K'(NL)/Kl(ML) or the initial concentration of ML increases, where K' is a conditional stability constant. By using Cu-edta and a Cu(II)-selective solid-state electrode, the linear calibration curves were obtained in the pc z range 3.00-5.30for Fe(III), 3.30-5.26 for Bi(II), and 2.94-4.80 for Cr(III).
    Download PDF (1067K)
  • Masahiro SAITO, Tadao TAKAHASHI
    1983 Volume 1983 Issue 5 Pages 753-755
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The Disproportionation of propylene Over a tungsten oxide-silica catalyst in the presence of carbon tetrachloride, isobutylamine and their mixture was investigated using a fixed-bed flow reactor under atmospheric pressure. The activity of the catalyst was more greatly enhanced by the addition of isobutylamine to the reaction system afser the addition of CC!, than by adding CCl4 or isobutylamine alone, but the enhanced activity gradually decreased during the reaction in the presence of isobutylamine (Fig.1). The addition of a mixture of CCl4 and isobutylamine stabilized the catalyst activity, which was slightly lower than the activity obtained immediately after adding isobutylamine following the addition of CCl4 (Fig.2, 3).
    Download PDF (528K)
  • Takayuki HAYASHI, Takao TOKUMITSUT, Tsuyoshi OKAMOTO, Mitsue NISHIYAMA
    1983 Volume 1983 Issue 5 Pages 756-760
    Published: May 10, 1983
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of o-(2, 4-dimethylbenzoyl)benzoic acid [1] in polyphosphoric acid at 100-120°C gave 1-methyl-3-[2, 4-dimethy1-10-oxo-9(10H)-anthrylidenemethyl]anthraquinone [3] and 1[2, 4-dimethy1-10-oxo-9(10H)-anthrylidenemethyl]-3-methylanthraquinone 4 along with 1, 3dimethylanthraquinone [2], but gave only [2] at 60-80°C. The structures of [3] and [4]were determined based on their MS, IR, NMR and UV spectra, and their chemical behavior. Mechanisms for the formation of [3] and [4] were discussed.
    Download PDF (1038K)
feedback
Top