NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1972, Issue 11
Displaying 1-50 of 50 articles from this issue
  • Hironobu KUNIEDA, Kozo SHINODA
    1972 Volume 1972 Issue 11 Pages 2001-2006
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It was confirmed from the studies of phase diagrams that the Schulman's so-called microemulsion is not an emulsion, but a solubilized solution. It is desirable to find conditions to produce so-called microemulsions in which the solution is stable and solubilization is large that oil and water mix to each other over wide composition range. Following conditions were found in order to increase the mutual solubility of oil and water with far less amount of solubilizer.
    (1) Optimum Hydrophile-Lipophile Balance (HLB) or Phase Inversion Temperature (PIT) of a surfactant.
    (2) Optimum mixing ratio of surfactants (solubilizer), i. e., optimum HLB (or PIT) of the mixture.
    (3) Optimum temperature for a given nonionic solubilizer.
    (4) The closer the HLBs of two surfactants, the larger the solubilization.
    (5) If the HLBs of two surfactants are close, mixtures of ionic and nonionic surfactants is better.
    (6) The larger the molecular size of solubilizer, the more efficient the solubilization.
    Download PDF (410K)
  • Hiroshi UTSUGI, Shigeoki NISHIMURA, Mitsuo SHIMAZAKI
    1972 Volume 1972 Issue 11 Pages 2007-2012
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Surface properties of the organophilic silica gels were investigated by the adsorptions of argon at 77°K, water vapour at 288°K and n-heptane at 273°K and were also investigated by the heat of immersion of water and that of n-heptane on the silica gels which have been surface treated with methanol, ethanol, butanol, pentanol, hexanol, octanol, decanol, dodecanol and tetradecanol, respectively. It was noticed that (1) the shape of isotherms of water vapour adsorption showed the type II classified by BET on both native silica gel and silica gels treated by methanol or ethanol, whereas, it showed the type I on the silica gels treated by alcohol with longer carbon chain. The adsorbed amounts of water on the surface-treated silica gels were reduced remarkably in comparison with that on the native silica gel at the same relative pressure. (2) The argon adsorption yielded the reasonable surface area of the silica gels with either hydrophilic or organophilic surface. (3) The sum of the surface areas obtained from water vapour adsorption and from n-heptane adsorption on the organophilic silica gel showed a reasonable agreement with the surface area obtained from argon adsorption. (4) The surface of the treated silica gel was composed of both the organophilic surface due to surface group and the hydrophilic surface due to silanols which were either unreacted with alcohol or uncovered by surface groups. The appreciable amounts of surface silanol. were observed to be covered by adjacent surface group and this trend was significant with increase of the length of carbon chain in a surface group. (5) On the organophilic silica gels, the heat of immersion of n-heptane increased with an increase of length of carbon chain in a surface group. The heat of immersion of water showed a linear relation with the surface areas obtained from water vapour adsorption. t Studies on the Surface Treatment of the Ultrafine Powders. XV.
    Download PDF (416K)
  • Chen-Shen CHAO, Shinya MUTO, Kenjiro MEGURO
    1972 Volume 1972 Issue 11 Pages 2013-2016
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The absorption spectra of iodine dissolved in aqueous solutions of heptaethyleneglycol dodecyl ether (C12E7) exhibit two extra bands at about 295 nm and 386 nm, when the concentration of C12E7 exceeds than its critical micelle concentration (CMC). These bands are attributable to the charge transfer interaction between iodine (an electron acceptor)and micellar C12E7 (electron donor), and can be utilized to determine the CMC of C12E7.
    The absorbance at 295 nm (and at 386 nm, too) plotted against C12E7 concentration showed an abrupt increase at 7 X 10-5 mol/l which agreed with the cmc determined by other methods. A similar behavior was observed also in the case of sodium dodecyl sulfate (SDS). Thus, the con- centration at which the absorption shows an abrupt increase may be called the apparent CMC of C12E7.
    Addition of a small amount of SDS (2 x 10-5 mol/l) effected a slight increase in the apparent CMC, together with a slight decrease in the absorbance. Upon further addition of SDS (1 X 10-4 mol/l) the apparent CMC increased a little, but the absorbance decreased significantly.
    When the concentration of SDS added (1 X 10-2 mol/l) exceeded the CMC, the absorbance increased monotonouslly with the concentration of C12E7, showing no apparent CMC. The SDS added probably formed mixed micelles with C12E7, and consequently, the charge-transfer complex between iodine and C12E7 was mostly destroyed.
    Download PDF (277K)
  • Fujio TAKAHASHI, Masayoshi OKAMOTO, Hirotoshi MAEDA
    1972 Volume 1972 Issue 11 Pages 2017-2022
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The solvent effects on the formation of charge transfer complexes of imidazole with p-nitrobenzaldehyde (NBA) and 1, 3, 5-trinitrobenzene (TNB) have been studied by the spectrophotometric method. The molar ratio of imidazole to NBA and to TNB are found to be 1: 1 and 3: 1 in any solvent respectively. The equilibrium constants for the imidazole-NBA complex in acetonitrile, acetone, N, N-dimethylformamide and chloroform are 1.2, 1.4, 2.5 and 1.4 l/mol (25°C), respectively. The absorption spectrum of the imidazole-TNB system shows two absorption maxima at 420∼430 nm and 520∼530 nm. The former might be assigned to 7r-r transition in imidazole. The latter might be due to n-x transition in which the lone pair of nitrogen of imidazole takes part. It could be thought from these results that the hydrogen bonding of imidazole with each solvent might influence the formation of charge transfer complexes.
    Amine-TNB-acetone systems form Meisenheimer type complexes. The values of logarithms of the rate constants of Meisenheimer type complex formation are proportional to pKb of amines. It is thought that the same process occurs in each amine-TNB-acetone system and imidazole behaves as a base catalyst as the other amines.
    Download PDF (335K)
  • Yoshihisa WATANABE, Makoto TAKEDA
    1972 Volume 1972 Issue 11 Pages 2023-2028
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In the presence of ruthenium trichloride, acrylonitrile is converted to a mixture of cis- and trans-1, 4-dicyanobutenes, adiponitrile and propionitrile under an atmosphere of hydrogen. The effects of the reaction temperature, RuCl3 concentration, HCl concentration, and partial pressure of hydrogen on the rate of the reaction have been investigated and the results are shown in Figs.3, 4, 5 and 8.
    The reaction rate was second order in the initial concentration of acrylonitrile as shown in Fig.5. However, the plots of hydrogen absorption vs. reaction time (Figs.2 and 3) indicate that the reaction proceeded independently of the, actylonitrile concentration.
    The cause of these apparently conflicting results was found to be that the rate is determined by the totol concentration of -C=N (Table 3). The active species as catalyst is suggested to be a ruthenium complex coordinated with two molecules of nitrites. A reaction mechanism based on the kinetic studies is proposed.
    Download PDF (388K)
  • Koichi KUNO, Shuji HIRAYAMA
    1972 Volume 1972 Issue 11 Pages 2028-2033
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to elucidate the role of ammonia in the ammoxidation of toluene over V205-TiO2 catalyst, various nitrogen compounds were oxidized with toluene in a flow system (Table 1). Ammoxidation of toluene with ammonia gave 65∼68% yield of benzonitrile at 355°C. However, very poor yields of benzonitrile, namely 3.5∼41.8% were obtained, when primary amines such as aniline, benzylamine, n-butylamine and mono-ethanolamine were used as nitrogen sources in place of ammonia. Benzonitrile was also formed against our expectation in the oxidation reaction of toluene with acetonitrile or dimethylformamide, although its ammount was small. With one exception, relatively high yield of benzonitrile, 54.9% was obtained in the case of formamide (Table 1).
    Oxidation of the nitrogen compounds was studied by the pulse reaction method. Ammonia was easily oxidized by catalyst oxygen to nitrogen and water.
    Nitrogen compounds adsorbed on catalyst were also oxidized to nitrogen, carbon oxides, water and others. Results indicated that C-N bond in the nitrogen compounds was more stable than C-C, C-H or N-H bond under the reaction conditions. Little differences in the benzonitrile yield between primary amines and other nitrogen compounds were explained by this relatively stable C-N bond.
    Download PDF (432K)
  • Koichi KUNO, Shuji HIRAYAMA
    1972 Volume 1972 Issue 11 Pages 2034-2040
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Ammoxidation of toluene to benzonitrile was studied by pulse reaction method. Ammonia adsorbed on a 10% V205-TiO2 catalyst was quickly oxidized to nitrogen and water at 375°C. At the same time V205 in the catalyst was apparently reduced to V204.33 (Fig.1).
    Adsorbed toluene was also oxidized by the oxygen in catalyst to toluene dimer, benzaldehyde, ben-laic acid, carbon dioxide and benzonitrile precursor, which was converted to benzonitrile by ammonia, pulse. (Fig.2, 3, 4)
    Stability of the precursor was examined at 350, 375 and 400°C. Adsorption and oxidation of benzonitrile was negligible in this experiments. The difference of selectivity for benzonitrile formation between continuous flow system and pulse reaction one was discussed.
    Download PDF (500K)
  • Ichiro NAKABAYASHI, Jiro YASUMURA, Satoshi ABE
    1972 Volume 1972 Issue 11 Pages 2041-2044
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The thickness of layer of sandwich type Raney nickel alloy was measured, and the distribution of sulfur bound to the catalyst surface was observed by means of an electron microprobe X-ray analyzer.
    The alloy layer, heated at 650°C under a nitrogen atmosphere and then pressed by a dental roller, was about 20 μ thick. The Raney nickel catalyst prepared from this alloy was active in hydrogenation of acetone. When the catalyst was immersed for 15 sec. in a saturated aqueous solution of H2S at 20°C, a large amount of sulfur was found distributed uniformly on its surface (Fig.3(c)). The catalyst thus treated lost its activity. The amount of surface sulfur detected after immersing an ordinary nickel plate, which was inactive for the hydrogenation of acetone, in a H2S solution was very small.
    When the catalyst was annealed at various temperatures up to 800°C in a hydrogen stream, the catalytic activity decreased abruptly at 400°C. The more the decrease in activity by the annealing, the less the amount of sulfur found on the surface after immersion in a H2S solution.
    These observations may lead to a conclusion that, as sulfur poisons active centers on the catalyst surface, the relation between the catalytic activity and the amount of sulfur is qualitatively parallel.
    Download PDF (192K)
  • Hideo TESHIMA, Toshiaki Sum, Noriyoshi MORITA
    1972 Volume 1972 Issue 11 Pages 2045-2048
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Catalytic activities of strong-acid cation exchage resins partially deactivated by sodium ion were studied for comparison with those of hydrochloric acid in various concentrations.
    The specific activities of resins obtained from the rate constants by correcting for the resistance to diffusion and the adsorption of ester in resin were decreased appreciably with the increase in the replacement with sodium ion, i. e., with the decrease in the acid contents of resins, and the decreases were more appreciable than those observed for free acid.
    It was found that, when the specific activities of resins were converted to the values based on the acid contents per unit volume of the swollen resins, the log-log plots of the specific activities of resins versus the acid contents formed a single straight line regardless of the degrees of crosslinking. However, the slope was about 1.25 which differed from the unity as shown by the hydrochloric acid solutions.
    Download PDF (298K)
  • Hirohisa KATO, Hiroshi KOKADO, Eiichi INOUE
    1972 Volume 1972 Issue 11 Pages 2049-2053
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It has been known that the photoconductivity of poly(N-vinylcarbazole)(PVK) can be much improved by treating it photochemically with some polyhalides. With the aim to understand mechanism of the high sensitivity of thus treated PVK, photochemical products were examined, mainly those in PVK treated with CBr4 or CHI3. Spectroscopically found were molecular halogen, an unidentified substance which absorbed near-ultraviolet light and a dye which was photodecomposed. Detailed investigation revealed that (1) the dye undergoes a reversible bleaching by ammonia, (2) PVK is subject to cross-linking by the photochemical treatment, (3)substitution reactions take place seemingly at the 3-position of the carbazole ring of PVK (from IR spectrum), and (4) the photoconductivity at any wavelength exhibits a definite dependence on the amount of the produced dye. The sensitizing substance was speculated to be a cationic dye with carbazole rings as the chromophore, and also with a linkage such as - CBr- which at the same timecrosslinks PVK. The high sensitizing efficiency of the dye would result from the tightly bound state of the dye to the photoconductive polymer.
    Download PDF (382K)
  • Nagaaki TAKEMOTO
    1972 Volume 1972 Issue 11 Pages 2053-2058
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The concentration distribution in the interfacial layer at the desalting side of the ion exchange membrane, and its change with time were investigated by the Schlieren-diagonal method reported previously by the auther, which was capable to investigate directly the change in concentration in the interfacial layer without any disturbance. As a sample, aqueous 0.1 N sodium chloride solution was used, and the measurement was carried out at 20°C. The picture of the Schlieren pattern and those of a digital stop watch and an oscilloscope indicating the change in current density with time were synthesized electronically and recorded on the video magnetic tape. This could be used to investigate small changes of the Schlieren pattern occurred within about 1/60 sec.
    The following results were obtained:
    (1) A thin layer with a great concentration gradient appeared near the membrane surface ("the thin layer"). The concentration distribution in the interfacial layer was expressed by an equation with two quadratic terms as a function of the distance from membrane surface, connecting at the end of "the thin layer" (Fig.3, equation (4) and (5)).
    (2) In the course of formation of the interfacial layer, the thickness of the layer increased with a decreasing concentration at membrane surface, and "the thin layer" appeared gradually. In the course of disappearance, on the other hand, "the thin layer" disappeared quickly, and then the concentration in the interfacial layer was recovered gradually without any change in thickness (Fig.6).
    (3) The agreement of calculated values of the' electric resistance at the interfacial layer, between from the concentration distribution and from the potential-current relationship, were satisfactory. The greater the ratio of electric resistance of "the thin layer" to that of whole interfacial layer the greater was current densities (Table 1).
    Download PDF (366K)
  • Shizuo SUGITA
    1972 Volume 1972 Issue 11 Pages 2059-2064
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    For the purpose of studying the thermal decomposition of the carbonate in the brine obtained by the ion exchange membrane method, laboratory tests were made by the use of a rotary evaporator at temperatures 50°, 62.5°, 75°, 87.5° and 100°C under the following conditions: (1) heating in the non-boiling state, (2) heating in the boiling state under reduced pressure, and (3) concentrating to the concentration-factor 1.8 in the boiling state. Furthermore the brines and the scales of salt-manufacturing factories were investigated. The results are as follows:
    1) In the laboratory tests, the carbonates are decomposed rapidly above 75°C.
    2) From the analytical date of the suspended substance obtained under the condition (3), it is found that calcium carbonate is formed below about 85°C, while magnesium hydroxide is formed above this temperature.
    3) The saturation indexes of calcium carbonate and magnesium hydroxide increase linearly in relations to the temperature elevation.
    4) From the Murotani's formula"), the following relation is found,
    K=(c/co-A)
    where, v is the deposition rate of alkaline scales (as CaCO30. hr), K is a rate constant, c/co is the ratio of super-saturation, and A is the limit of ratio of super-saturation in the metastable condition.
    In the above formula if the saturation index is substituted for the ratio of super-saturation, under the condition (2) and (3), v has the linear relation to the saturation index of calcium carbonate, the values of K are 0.032 and 0.076 respectively, and the value of A for the saturation index of calcium carbonate is 1.2 in both cases.
    5) The analytical results of the scales deposited from the brine of the salt manufacturing factories show that the phosphate scale was formed by the decomposition of the condensed sodium phosphate, which is added to the brine at the electrodialytic process.
    Download PDF (390K)
  • Shizuo SUGITA
    1972 Volume 1972 Issue 11 Pages 2065-2070
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In the concentration of brine by the use of the ion exchange membrane the relationship between the formation of alkaline scale and the decomposition of the condensed sodium phosphate, a scale inhibiter, was- tested. This phenomenon was studied by the laboratory and practical tests, described in the previous paper.
    In this case the phosphate ion, the decomposition product of the condensed sodium phosphate, was analyzed quantitatively by the solvent chloroform+buthanol extraction method. The results are as follows:
    1) The effects of sodium hexameta-phosphate on the inhibiter of alkaline scales in the brine with the concentration-factor 1.8, is sufficiently large at 50°C, but is extremely small at 75° and 100°C.
    2) Under the boiling and concentrating conditions the close relation between the decomposition of sodium hexameta-phosphate and that of carbonates is found; the former increases with an increase of the latter.
    3) At 100° and 50°C, the rates of decomposition of sodium hexameta-phosphate in the decarbonated brine decrease to 1/50 and 1/20 of those of the non treated brine, respectively.
    4) In concentrating the brine to the concentration-factor 1.8 at boiling temperature 100°C, 97.8, 0.98 and 1.25% of the added sodium hexameta-phosphate are transferred, to the deposited scale, the common salt, and the mother liquid, respectively.
    5) Sodium penta-calcium sulfate (Na2S0, 5 CaSO4 H20) is found in the scales deposited on the inner place of the pans which are practically operated at the concentration-factor of about 15 and at 66° and 88°C.
    Download PDF (424K)
  • Chuichi GOTO, Isao JOKO, Kyoichi TOKUNAGA
    1972 Volume 1972 Issue 11 Pages 2070-2075
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Dynamic drying of xylene, benzene, ethyl ether, ethanol and 2-propanol were studied at room temperature using a fixed bed packed with zeolite-A (ML 4 P) synthesized from halloysite. For comparison, Linde Molecular, Sieve 4 A (LMS 4 A) and 3 A (LMS 3 A) were also tested in a similar manner. The breakthrough curve was drawn from the result of each run of the ex- periments and used to calculate the height of adsorption zone (Z.) and the amount of equilibrium water adsorption.
    The height per transfer unit (HoL) was also calculated from Za, and the number of transfer units.
    The height of adsorption zone (Z.) and the height per transfer unit (HoL) of ML 4 P were shorter than those of LMS 4 A when it was used as the dryer of xylene and 2-propanol. These results were supposedly attributed to the difference of pore size between ML 4 P and LMS 4 A, the pore size of ML 4 P being larger than that of LMS 4 A. The height of adsorption zone (Za) and the height per transfer - unit (HoL) for the drying of ethyl ether with ML 4 P were longer than those of xylene and benzene, while the amount of equilibrium water adsorption for the drying of ethyl ether was less than that of xylene and benzene.
    The amount of equilibrium water adsorption for the drying of ethanol with ML 4 P was less than that of ethyl ether. ML 4 P-17 R (1.7 m/m Pellet) showed the most high amount of water adsorption among the adsorbents tested. The water content of xylene, benzene, ethyl ether, ethanol, 2-propanol were reduced to less than 5 wt. ppm by the use of ML 4 P or LMS adsorbents.
    Download PDF (387K)
  • Hirohi ITOH, Katsuzo SHIRAISHI
    1972 Volume 1972 Issue 11 Pages 2076-2081
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The photochemistry of μ-hydroxo-bis {pentaamminechromium(III)} ion, [(NH3)6Cr-OH-Cr(NH35)5+], in aqueous acid solution (0.1 N HCl04, μ=0.2) has been studied using light which showed the absorption corresponding to the ligand field bands (18.5 X103 cm-1, 27.3x 10 cm-1). When [(NH8)5Cr-OH-Cr(NH3)5]5+ in aqueous acid solution was irradiated, absorption maxima of the complex (20.0x 10 cm-1, 27.0x 10 cm-1) shifted to longer wavelength and ammonium ion as well as pH of the solution increased with exposure time. So primary process of photochemical reaction was expressed as follows:
    [(NH8)5Cr-OH-(NH3)5]5+ + H20 →[(NH3)5Cr-OH-Cr(NH3)4(H20)]5+ + NH3
    Quantum yields for the primary process were obtained by the determination of NH4+ in the solution. The value was 0.30 and was independent of pH, reaction temperature, ionic strength of the solution and irradiating wavelength.
    Mechanism of the thermal reaction of the complex ion in aqueous acid solution has been studied in relation to photochemical reaction.
    The absorption maxima of the complex ion shifted to shorter wavelength as the thermal reaction proceeded. Thermal reaction was considered to be a rupture of Cr-0 bond as follows:
    [(NH8)5Cr-OH-(NH3)5]5+ +H20 →2 [Cr (NH3)5(H2O)3+]
    The reaction was of first order and its rate was expressed as follows:
    v = kobs {(NH8)5Cr-OH-(NH8)5(H2O)5+} and kobs was dependent on reaction temperature (10.3X 10-3 min-1. at 45°C, 2.3x103 min-1 at 35°C and 0.4 X 10-3 min-1 at 25°C in 0.4 N HCl04) and on pH of the solution.
    Download PDF (366K)
  • Kensaku HARAGUCHI, Saburo ITO
    1972 Volume 1972 Issue 11 Pages 2082-2086
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Solvent extraction technique was utilized to study the ligand substitution reaction kinetics of bis (2-methyl-8-quinolinolato)copper (ll) (CuL2) with ethylenediaminetetraacetate ion (EDTA). The reaction is described as follows:
    The rate of back extraction of the CuL2 complex from chloroform layer to the aqueous layer by EDTA was determined spectrophotometrically measuring absorption band of the CuL2 com- plex in the organic phase at 400 nm. The rate of back extraction increases quite rapidly by the elevation of shaking speed, and reaches constant value at and above the shaking speed of 340 strokes per minute. Within this "plateau" region, the observed rate of back extraction was considered to be equal to the rate of chemical reaction in the aqueous phase. The pH and EDTA concentration dependences of the pseudo first order rate constant, kips, were analyzed over a pH range from 6.2 to 7.6 at 25°C and ionic strength of 0.1.
    The initial rate of the reaction can be expressed by the following equation:
    The estimated rate constants, kkh2y, kH3Y, kY are 1.1X104, 4.6X10-2, 3.8X10-21.mol-1, sec-1, respectively.
    Download PDF (316K)
  • Tamio KAMIDATE, Takao YOTSUYANAGI, Kazuo AOMURA
    1972 Volume 1972 Issue 11 Pages 2087-2091
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It was found that the extraction efficiency of Cu(I) ion for dialkyl sulfides increased extremely by an addition of MgBr2 into CuBr2 solution, and, dialkyl sulfides from C, to Cl0 were quanti- tatively extracted into the aqueous phase. In order to elucidate the extraction characteristics of bromo-copper(1I) complex ions for dialkyl sulfides, dimethylsulfide (DMS) was extracted with the solution of Cu2Br-. It was found that CuBr+ was mainly concerned for the extraction. CuBr+ reacted with DMS and formed white solid having a composition of Cu(I): Br: DMS =1: 1: 1, in which copper(I) was reduced to Cu(I). However, in the concentrated solution of bromide ion or CuBr2, only water soluble extract exists in the aqueous phase. These results can be explained that bromo-copper(I) complex ion, [Cu(I)BrMS](-n) +(n = 2, 3) forms in the aqueous solution when bromide ion concentration is high, and that polynuclear complex ion, [Cu(II)Brn+iCu(I)DMS](2-n)+(n=1, 2) forms when the concentration of CuBr2 is high.
    Download PDF (362K)
  • Akira UEJIMA, Hideaki MUNAKATA
    1972 Volume 1972 Issue 11 Pages 2092-2095
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Esters of carboxylic acid can be used as a alkylating reagent forter tiary amines, as well as alkyl halides. This reaction is remarkably accelerated in the presence of methyl alcohol or phenol. In this experiment, the reaction of dimethyl benzylamine with alkyl benzoates has been studied kinetically.
    Download PDF (264K)
  • Koichi MURAI, Chikai KIMURA
    1972 Volume 1972 Issue 11 Pages 2096-2100
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Competitive and consecutive quaternization (1) of N, N, N', N'-tetramethyl-a, w-alkanediamines with butyl and ethyl bromides was studied in propylene carbonate at different temperatures. The rate constants were calculated by the Frost-Schwemer method and the correlation between the reactivity and the chain length of diamines was discussed. The ratios of rate constants, 11x=k1lk2, were found to be larger than unity for all cases and varied with the chain length of diamines in the following order: n=2>3>4>6∼10. Such tendency in 1/k values was found to be due largely to the fact that the variation in k2 values with increase of n value is far larger than in k1.
    Similarly to the tendency of the rate data, the dependence of activation parameters, E and ΔS, on n values was larger for the 2nd step than for the 1st step of the reaction. Both parameters for the 2nd step were much smaller when n is 2 or 3 than the other n values. The isokinetic relationship between E and 4, 94 was observed for the 1st step, but this relation did not hold for the 2 nd step.
    From the kinetic results it was concluded that the specific features of the quaternization of diamines appeared in the 2nd step were caused by the steric, polar and electrostatic field effects of the 1st quaternized amino group and the statistical factor with the concentration of amino group.
    Download PDF (327K)
  • Teruto TANDA, Shaya Fujii
    1972 Volume 1972 Issue 11 Pages 2100-2103
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of benzoic anhydride in toluene in the presence of cupric oxide to give tolyl benzoates, cresols, phenyl benzoate and phenol has been studied. Tolyl benzoates are produced by the reaction of benzoic anhydride with toluene, while phenyl benzoate is the decomposition product of cupric benzoate. Cresols and phenol are formed by hydrolysis of the corresponding esters.
    Total yield (I) of tolyl benzoates and cresols increased with increased temperatures and the amounts of cupric oxide. The highest value of (I) is 53.1 mol% (260°C, 5 hr, cupric oxide/ benzoic anhydride= 6: molar ratio). Total yield (II) of phenyl benzoate and phenol is fairly small.
    The ratio of (I) to (II) is found to be higher for larger amounts of cupric oxide, and reaches to 75.5 (260°C, 0.5 hr, cupric oxide/benzoic anhydride= 6: molar ratio) showing that the reaction with toluene proceeds, with relatively high selectivity.
    Download PDF (239K)
  • Akihiro YAMAGUCHI, Mitsuo OKAZAKI
    1972 Volume 1972 Issue 11 Pages 2103-2107
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    α-Bromo-4-nitrcstilbenes and diphenylacetylenes were prepared by utilizing the modified Wittig reaction.
    Diethyl a-bromo-p-nitrobenzylphosphonate [2] was obtained quantitatively by bromination of diethyl p-nitrobenzylphosphonate [1].
    Treatment of [2] with aromatic aldehydes in the presence of an equivalent amount of sodium alkoxide in alcohol at room temperature gave the corresponding a-bromo-4-nitrostilbenes in good yield.
    The use of two equivalent amounts of base gave diphenylacetylenes. Similarly the reaction of α-iodo-p-nitrobenzylphosphonate with aromatic aldehydes was investigated.
    Download PDF (343K)
  • Yukio KIKUCHI, Tsuneo IKAWA
    1972 Volume 1972 Issue 11 Pages 2107-2112
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to determine the relative reactivity of hydrocarbons and ketones toward their peroxy radicals, oxidation reaction of hydrocarbons in the presence of ketones was investigated.
    Oxygen absorption rates observed are plotted against the concentration (Fig.1∼4). Relative reactivities of peroxy radicals were calculated from Equation ( 1 ) by means of co-oxidation rate method (Table 4). The reactivities of peroxy radicals from ketones seem to depend on their structure.
    The shape of the curve of oxygen absorption rate was determined by three parameters r A, r, 3 and in Equation ( 1 ), that is, the value of in Equation (7). The relative reactivities were also determined by the methods of Mayo-Lewis and of FinemanRoss (Table 6), and the reliability of their values was discussed.
    Download PDF (347K)
  • Sei-ichiro IMAMURA, Toshiaki BANBA, Masaaki TERAMOTO, Hiroshi TERANISH ...
    1972 Volume 1972 Issue 11 Pages 2113-2118
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The experiment on the decomposition of 1, 2, 3, 4-tetrahydro-1-naphthyl hydroperoxide was carried out in pyridine solution using copper salts as catalyst.
    The cuprous salt was more active than the cupric one. The addition of water, one of the products in the reaction, increased the catalytic activity of the cupric salt but, on the contrary, deactivated the cuprous catalyst.
    From these facts and the consideration based on the ESR measurement, it was concluded that the cuprous salt decomposed the hydroperoxide through radical mechanism and the cupric salt through ionic one, namely, ionic dehydration of the hydroperoxide.
    The products consisted of fair amount of a-tetralone and water in pyridine solution bu':. considerable amount of a-tetralol was obtained in acetonitrile solution. This showed that the products distribution varied with change of basicity of the solvents. In connection with this, the experiment on the photo-decomposition of 1, 2, 3, 4-tetrahydro-1-naphthyl hydroperoxide was carried out and the behavior of the alkoxy radical in basic solvents was discussed.
    Download PDF (377K)
  • Haruo SHIBATANI, Akira AMANO
    1972 Volume 1972 Issue 11 Pages 2119-2126
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Thermal decomposition of 2-pentene was studied in a flow apparatus under an atmospheric pressure, at temperatures ranging from 480 to 730°C, and with 'residence times from 0.05 to 35 sec and nitrogen/2-pentene mole ratios 6 and 15. The overall reaction obeys the first order rate equation with the rate constant being log k (sec-1)=12.3-53000/4.575 T. The main reaction products were approximately equimolar amounts of methane and butadiene, consisting 50∼60 mol% of 2-pentene decomposed. Other reaction products were hydrogen, ethylene, ethane, propene, 1-butene, 2-butene, 1, 3-pentadiene, 3-methyl-1-butene and 4-methyl-2-pentene.
    The reaction can be accounted for by a free-radical chain process with complication due to competition between decomposition and isomerization of alkenyl as well as alkyl radicals. Ex- perimentally observed product distributions as well as reaction rates are found to be in good agreement with those calculated on the basis of estimated kinetic parameters of the elementary reactions in the proposed chain scheme.
    Download PDF (579K)
  • Kazuo KITAMURA, Shigenobu KOBAYASHI
    1972 Volume 1972 Issue 11 Pages 2127-2131
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The dye penetration coefficients, the starting temperatures of contraction and other factors in the carrier dyeing of poly(ethylene terephthalate)fibre, were determined as a function of carrier concentration in the dyeing bath.
    The dependence of dye penetration coefficient on the concentration of carrier and the temperature was found to be shown formally by the Williams-Landel-Ferry equation,
    Here p and p, are the dye penetration coefficient at absolute temperature, T and that at temperature, Ts, respectively. The subscripts, (c) and (0) indicate the concentration of carrier added and that of zero, respectively. In the above equation, T8, (0) and Ts, (e) are the temperatures which are higher by 25°C than the starting temperature of contraction in each dyeing bath.
    Download PDF (279K)
  • Satoshi KISHIMOTO, Osamu MANABE, Hachiro HIYAMA, Nenokichi HIRAO
    1972 Volume 1972 Issue 11 Pages 2132-2139
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The coupling reaction of p-substituted benzenediazonium salts with 1-naphthol has been studied kinetically.
    The products were 2-(I), and 4-azo dyes (II) and a small amount of 2, 4-diazo dye (Table 4). The I / II ratio depended on pH, buffer composition of the reaction medium, and the electronegativity of the p-substituent.
    The formation of II was found to subject to a general base catalysis, while the formation of I was little affected by base (Fig.3). The catalytic effect decreased as the p-substituent becomes more electronegative in the order of OCH3<CH3<H<Cl<NO2 (Table 5); it was hardly observed in the case of p-nitrobenzenediazonium ion. It was assumed that in the coupling on the 2-position of 1-naphthol, the formation of the a-complex may be the rate determining step. On the other hand, deprotonation from the a-complex by bases is the rate determining step to form II. As the p-substituent in the diazo-component becomes more electronegative, the rate of deprotonation becomes relatively greater than that of the formation of the a-complex, thus in the case of p-nitrobenzenediazonium salt, the formation of the a-complex is the rate determining step.
    The kinetic data agree well the Yukawa-Tsuno relationship (Fig.4): r =0.58, p=3.7 (for the reaction at 2-position), p=3.2 (for the reaction at 4-position).
    Download PDF (465K)
  • Shin-ichi FUJITA, Yasuo KIMURA, Toyohide IGUCHI, Rikisaku SUEMITSU, Ya ...
    1972 Volume 1972 Issue 11 Pages 2140-2143
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of l-linalool [1] with chloranil [2] in benzene was investigated. The reaction products were R-myrcene [3], α-terpinene [19], dipentene [4], d-lirnonene [5], r-terpinene [20], cis-ocimene [17], p-cymene [16], terpinolene [21], trans-ocimene [18], 1-methyl-4- isopropenylbenzene[6], terpinen-4-ol[8], myrcenol[9], cis-ocim6nol[10], trans-ocimenoI [11], d-α-terpinel[12], nerol[13], and geraniol[14].
    The reaction of dl-linaloo1[1'], geraniol[14], nerol[13], orα-terpineol[12] with chloranil [2] was also tested, and the mechanism of the formation of the above mentioned products was discussed.
    Download PDF (255K)
  • Takahiko ISOBE, Tadao KAMIKAWA, Takashi KUBOTA
    1972 Volume 1972 Issue 11 Pages 2143-2148
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Constituents of the leaves of Isodon species have been examined. From I. longitubus KUDO, nodosin, lasiodin, isodocarpin and oridonin were isolated. From I. shikokianus HARA var. intermedius MURATA, a new bitter component, isodomedin, was isolated. From I. Kameba OKUYAMA, kamebanin and mebadonin, from I. inflexus KUDO, inflexin, from I. umbrosus HARA, umbrosin and from I. effusus HARA, effusin were isolated as new components. Parts of structures have been assigned to these new compounds from spectroscopic evidence.
    Download PDF (471K)
  • Masayuki NAKAGAKI, Saburo SHIMABAYASHI
    1972 Volume 1972 Issue 11 Pages 2149-2153
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In the process to reach the membrane-equilibrium in the aqueous solution of poly(vinylpyrrolidone)(PVP) and sodium dodecyl sulfate(SDS), it was found that osmosis of water occurred through the membrane as well as diffusion of SDS. By the osmosis through the membrane, water penetrated into the inner solution which contained PVP and the volume of the inner solution increased at the membrane equilibrium regardless of the initial condition of the ex. - periments. It was concluded that the penetration of water into the PVP-containing side oc- curred by the additional osmotic pressure which was due to the sodium ion dissociated from the SDS-PVP complex formed. This effect on the osmosis was small for aqueous PVP solutions without addition, or for the PVP solutions containing CaCl2 or dodecyl amine hydrochloride (DAH) which did not interact with PVP. On the other hand, the aqueous solutions of polyelectrolyte, such as sodium carboxymethyl cellulose(CMC-Na), showed this effect more markedly than the above mentioned systems, even without any added salt. It was, therefore, concluded that counter-ions dissociated from the polyelectrolytes or from the complex formed contributed to this osmotic phenomenon.
    Tryptophan(Tay.) formed complex with PVP, but amount of bound Try. was smaller than that of SDS, and the free concentration in equilibrium was larger than that of SDS, so the intensity of osmosis was small. This phenomenon was similar to the fact that when NaCI was added to aqueous solution of CMC-Na, intensity of osmosis decreased. It was, therefore, concluded that the osmotic pressure due to counter-ions from the polymer decreased and, accordingly, the osmotic phenomenon became less pronounced, when the concentration of added salt in polyelectrolyte solution or that of the free solute in complex-forming solution increased.
    Download PDF (390K)
  • Sakae KON-YA, Eiichi HIROTA, Masaaki YOKOYAMA
    1972 Volume 1972 Issue 11 Pages 2154-2157
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The polycondensation reaction of bis[p-(chloroformyl)phenyl]phosphine oxide [1] with four kinds of heterocyclic diamines [2] were carried out in dimethylacetamide. When [1] and [2] were reacted in a 1: 1 molar ratio at 0°C for more than 4 hr in the presence of tertiary amine, white or pale-yellow solids were obtained in ca.65% yield, which had inherent viscosities of 0.06-0.18 dlIg and softening points of the temperature range of 260∼318°C. As an acid acceptor, N, N'-diethylaniline having a pKa value of 6.56 was found to be satisfactory for the polycondensation reaction. But the inherent viscosities of the products did not exceed more than 0.18. These polyamides were generally heat stable, flame-resistant, and readily soluble in polar solvents compared with the corresponding polyamides from terephthaloyl dichloride.
    Download PDF (267K)
  • Eishun TSUCHIDA, Kei SANADA, Kazuhiko MORIBE
    1972 Volume 1972 Issue 11 Pages 2158-2161
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The occurrence of water absorption of polycationic polymers of an integral type, which have cationic sites in the polymer backbone, has been examined and discussed in terms of relative humidity and hygroscopic equilibrium, and also compared it with some cationic polymers of a pendant type.
    The water sorption capacities of cationic polymers remarkably increased at higher relative humidities than those of about 50%. At lower relative humidities, only localized water absorption occurred, while at higher relative humidities, dissolution occurred.
    It was concluded that water absorption became larger in the pendant type than in the integral type and even larger due to the existence of non-ionic hydrophilic groups in macromolecules.
    Download PDF (257K)
  • Masaaki MIWA, Kei SANADA, Eishun TSUCHIDA
    1972 Volume 1972 Issue 11 Pages 2161-2165
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Formation of polyionic polymer complexes (PIC) by the reaction of sodium poly(styrenesulfonate) (NaSS) with some cationic polymers of an integral type, which have cationic sites in the polymer backbone, have been studied in comparison with polycationic species of a pendant type. The reaction in aqueous solution was rapid and stoichiometric, and led to the formation of the precipitate of PIC. The infrared spectra of the precipitates coincided with that of the equimolar mixture of polycations and polyanions. The solubility of each component was completely different, so an ionic cross-linked structure was suggested. By the use of cationic polymers of an integral type as polycationic components, PIC once formed could be dissolved again when excess NaSS was added. This phenomenon seemed to be due to electrostatic interaction between polycationic sites which were not protected by hydrophobic moiety and other polyanionic sites of NaSS to loose salt-like bond.
    The behavior in resolution of PIC was different according to the structure of each polycation. Phase diagram of PIC containing cationic polymer of an integral type for a ternary solvent (water-organic solvent-simple salt electrolyte) had fairly smaller dissolution region than that of PIC of the pendant-pendant type.
    Download PDF (349K)
  • Kuniharu KOJIMA, Takeshi HABU, Susumu IWABUCHI, Masako YOSHIKUNI
    1972 Volume 1972 Issue 11 Pages 2165-2171
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The polymerization of methyl methacrylate (MMA) by the catalysts systems of tributylborane (TBB) with pyridine and its derivatives, such as hydroxypyridine, cyanopyridine, nicotinic acid, nicotinamide, nicotinic ester, picoline, lutidine and quinoline was studied. The relative initiating activities of the pyridine and its derivatives were found to be in the following order: nicotinic acid> nicotinic ester> nicotinamide> pyridine>3-hydroxypyridine>3-cyanopyridine> quinoline > 3-picoline > acridine > 2, 6-lutidine In the polymerization of MMA initiated by TBB/methyl nicotinate (MN) system, the rate of polymerization was found to be proportional to the concentration of MMA and to the square root of TBB/MN concentrations, respectively, and the apparent activation energy was 4.1 kcal/ mol. And the reaction of TBB with pyridine was also studied by NMR and ESR methods. From these results, it was found that the polymerization proceeds via a radical mechanism involving a co-ordination between TBB and pyridine or its derivatives.
    Download PDF (464K)
  • Hisao ISHIKAWA, Katsumi OKUBO, Tae OKI
    1972 Volume 1972 Issue 11 Pages 2171-2177
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A purified mucin solution isolated from natto bean was composed of fructan and poly-DLglutamic acid in the ratio of 22.1% and 77.6%. A aqueous solution of the natto mucin had higher viscoelasticity and spinnability than those of the solutions of the other polyelectrolytes. Spinnability of the mucin solution was remarkably decreased on heating over 60°C or by vigor- ous stirring, and this property was closely related with surface tension of the solution.
    Spinnability, surface tension, and viscosity of natto mucin and poly-L-glutamic acid solutions had maxima in the range of neutral pH, while the helix content of poly-L-glutamic acid or poly-DL-glutamic acid in the mucin decreased remarkably with increasing pH from 4.7 to 8.5.
    The above results indicate that the spinnability of natto mucin in an aqueous solution arises from the formation of the network structure of randomly coiled poly-DL-glutamic acid by the intermolecular hydrogen bonding in the presence of fructan.
    Download PDF (496K)
  • Tetsuya IMAMURA, Fumikatsu TOKIWA
    1972 Volume 1972 Issue 11 Pages 2177-2184
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Fe203-deposition onto polyester, nylon and cotton fabrics in Na5P3O16 or NaCl solution of an ionic strength of 1 x 10-3 was discussed in terms of their potential energies. Potential energies due to the interaction of the dissimilar electrical double layers (Fig.3) and the interaction of the van der Waals forces (Fig.4) between Fe203 and fabrics, were calculated by an assumption of interaction between a sphere and an infinite flat plate, on the basis of the theory of heterocoagulation. Zeta potentials (C) of Fe203 and fabrics (Table 1) were measured by a method of streaming potential. In Na5P3O16 solution, the potential energy between Fe203 and fabric of polyester or nylon was large repulsive (Table 3) due to each large negative C potential, and Fe203 was not deposited onto fabrics (Fig.7). Hewever, in NaCl solution, the potential energy was small (repulsive or attractive) due to each small negative C potential and Fe203 was deposited onto fabrics considerably (Fig.7). The relation between the deposition of Fe203 onto cotton fabric and their potential energies showed a similar tendency, but the deposition seemed to be influenced by not only their potential energies but also the complicated fine. structure of cotton fabric.
    Download PDF (588K)
  • Toshiro IIJIMA, Du Jin CHUNG
    1972 Volume 1972 Issue 11 Pages 2184-2188
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    To elucidate the dependence of diffusion coefficient of disperse dye in poly(ethyleneterephthalate) (PET) on concentration, p-nitroaniline (PNA) was used as a model penetrant of the dye over the wide range of concentration at 70, 75, 80, 85, 90 and 95°C. Diffusion coefficients were measured by cylindrical film roll method. The dependence of diffusion coefficient on penetrant concentration was clearly observed(Fig.2), and empirically represented by the equation (1), for the lower concentration (up to about 8 X 10-2 mol/kg-1) and (2) for the higher concentration,
    D=D0Cn(1)
    D=D0'exp(C)(2)
    where C is concentration of penetrant and n and 8 are constants. PNA penetrated into an easily diffusible amorphous region at the lower concentration. At the higher concentration of PNA, the penetrant diffused into more closely packed amorphous part without changing the total amounts of amorphous region available for dyeing.
    It was suggested that the increase of diffusion coefficient was caused by the minute qualitative change of amorphous region of PET, which could not be detected by the IR spectrum (Table 2), small angle X-ray scattering and dichroism of the dyed film. With increasing PNA concentration the decrease of relaxation modulus of the PET film was observed (Fig.6), and this tendency corresponded to the increase of diffusion coefficient mentioned above (Fig.2).
    Download PDF (329K)
  • Tsutomu INA, Toshihide KUROSAWA, Takeshi KOMATSU, Tetsuhisa YAMAMOTO, ...
    1972 Volume 1972 Issue 11 Pages 2188-2193
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The fractionation of a large amount of the copolymer, ethylacrylate and 2-chloroethylacrylate (95/5) (Table 1), was carried out by the use of large scale column (Fig.1). Ethanol was used as a elution solvent. A concentrated ethanol solution of the polymer was introduced in the column and deposited on the glass beads by lowering the temperature of the column (Table 2). The ethanol solution containing 350 g of polymer occupied the whole space of the column. The time required to complete the deposition of polymer on glass beads was shorter than that re- ported before. The polymer concentration in eluted solution was maintained below a certain value through the fractional process by controlling the temperature elevation of the column due to the increase in concentration (Fig.3, Fig.4).350 g of the polymer was fractionated efficiently using a column 10 cm in diameter and 150 cm in length (Table 4). The fractional equation of Mark-Sakurada-Howink showing the relation between' intrinsic viscosity [η] (dl/g) and number average molecular weight Mn was used as follows:
    [η] = 2.1 x 10-4 Mn65 (acetone, 30°C)
    Download PDF (429K)
  • Takeo YANO, Toru SUETAKA, Tadashi UMEHARA
    1972 Volume 1972 Issue 11 Pages 2194-2199
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The kinetics of the oxidation of cuprous chloride with oxygen was studied in the solution of hydrochloric acid whose concentration range was 0.2-4.8 molg. Since the reaction was fairly fast, it was difficult to study the reaction by the ordinary kinetic method. Accordingly, the reactions li'vere observed in a stirred vessel with a flat gas-liquid interface in the diffused state. The observed chemical absorption rates of oxygen were analyzed by using the equations for chemical absorption obeyed (m+ n) th-order irreversible kinetics, and it was found that the reaction was of the first-order with respect to cuprous chloride and oxygen. The over-all reaction rate constants were determined from the experimental data by using the equations for chemical absorption obeyed second-order kinetics. It was found that cupric chloride in the solution retarded the reaction rates. The retardation due to cupric chloride was studied and found that the concentration of cuprous chloride was decreased due to the formation of a complex with cupric chloride, which brought about the depression of the rate.
    Download PDF (430K)
  • Isao YASUMOTO, Minoru IWAKURA, Keizo NAGAI
    1972 Volume 1972 Issue 11 Pages 2200-2201
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A device for temperature control of a liquid bath was presented, using a thermoregulator, an FET, a pulse generator, an SCR, and a heater.
    In the device, the FET was used instead of a vacuum-tube, and a triggering signal was applied to turn the SCR on. In order to generate pulse, any of the following three methods was used; the astable multivibrator, the blocking-oscillator circuit and the relaxation-oscillator circuit.
    The detail of the circuits was presented and the operation of them was described.
    Download PDF (115K)
  • Fumio NOZAKI, Ichiro INAMI
    1972 Volume 1972 Issue 11 Pages 2202-2204
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The catalytic selectivity of the three uranium oxides, UO3, Us08 and UO2, for dehydrogenation and dehydration of ethanol was investigated using the pulse-reaction method.
    As is shown in Fig.1, the catalytic activities varied initially as the injection of reactant was repeated, and then reached nearly stationary values. During the course of reaction, the uranium oxides except UO2 were reduced to oxides of uranium of lower oxidation states. An interesting relationship was found between the oxidation state of uranium and the catalytic selectivity, as is summarized in Table 1.
    Download PDF (236K)
  • Yoshi KADOTA, Yutaka FUKUDA, Kozo SONE
    1972 Volume 1972 Issue 11 Pages 2204-2206
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The formation of the mixed chelates [Cu AB]n in aqueous solutions containing [Cu A2]2+ (A= bip or phen) and [Cu B] (B=1, 3-propanediamine (tn), β-alaninate ion (β-ala), malonate ion (mal), or pyrocatechol-3, 5-disulfonate ion (tiron)) was studied spectrophotometrically. When B was tn, the mixed chelates were found to be unstable, while in all other cases the formation of the mixed chelates were nearly quantitative, and their Vmax values, shown in Table 1, clearly indicated the same trend as those of the other [Cu AB]n chelates described previously. Solid chelates shown in Table 2 were also obtained and characterized.
    Download PDF (193K)
  • Yasuhiro TAKAGI, Masatada SATAKE
    1972 Volume 1972 Issue 11 Pages 2207-2208
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to save time required for a low-temperature ashing (usually 18∼20 hrs) for the determination of cadmium in unpolished rice, pre-treatment of rice samples in an electric furnace at different temperatures was conducted and the efficiency of ashing was examined. The opti- mum conditions of ashing was found as follows: 5 g of unpolished rice was weighed and ashed for hrs in a chamber of a low-temperature asher under 150 ml/min oxygen flow and consuming 200 W of RF power, after treating 3 hrs in an electric furnace (260∼280°C).
    The ashed sample was dissolved in an acid, and trace amounts of cadmium were determined by assuming that the loss did not occur by ashing. The present method was applied to unpolished rice samples and the results are shown.
    Download PDF (127K)
  • Koichi MURAI, Chikai KIMURA
    1972 Volume 1972 Issue 11 Pages 2209-2211
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The electrostatic and non-electrostatic contributions of solvent to Eyring-Laidler equation ( 1 ) were discussed in the solvent effect of the quaternization of pyridine with dodecyl bromide. The rate constant, log k, showed better correlation with solubility parameter, 8, than with D-1/2 D+1. The linear relations were observed respectively among the aliphatic alcohols and the aprotic solvents except tetrahydrofuran, dioxane and benzene. Similar relations were observed for the plots of ET vs. log k and E vs.4S. These results indicate the importance of the mode of solvent-solute interactions as well as the electrostatic and non-electrostatic factors in the solvent effect.
    Download PDF (182K)
  • Shunzo YAMAMOTO, Yorinobu TAKATA, Norio NISHIMURA, Shigeo HASEGAWA
    1972 Volume 1972 Issue 11 Pages 2212-2214
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In hydroxylic solvents, 4-diethylaminostilbenes are much stronger base than the corresponding dimethylaminostilbenes. In acetonitrile, however, the difference, ApKa, between the basicities of these two series of compouns is very small. Furthermore, in the excited state the ilpKa value is considerably smaller even in hydroxylic solvents. The anomalously high base strength of 4-diethylaminostilbene in hydroxylic solvents can be explained in terms of the steric inhibition to hydrogen bonding between the free base and the solvent.
    Download PDF (192K)
  • Nobuo ISHIKAWA, Akira SEKIYA
    1972 Volume 1972 Issue 11 Pages 2214-2216
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In KF- or KHF2-DMF system hexafluoropropene was oligomerized at room temperature and under atmospheric pressure to give a mixture of dimers and trimers. Total yield of these oligomers was almost quantitative. The molar ratios of dimers to trimers, and the com- positions of isomers [1]-[5] were determined by gas chromatography (Table 1). In KF- or KHF2-MeCN system, dimer [1] was formed almost exclusively.
    By the action of KF-DMF at high temperature, a part of dimer [1] was isomerized into thermodynamically more stable dimer [2], whereas trimer [4] was isomerized even at room temperature into trimer [5].
    Download PDF (192K)
  • Tadao TAKAHASHI
    1972 Volume 1972 Issue 11 Pages 2216-2218
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Hydrogen treatment of C4 hydrocarbon fraction obtained from naphtha cracking the residue of butadiene extraction, has been studied over supported Pd-catalysts in a flow reactor under atmospheric pressure. The effects of catalyst composition and reaction conditions have been examined. Satisfactory results have been obtained with Pd-alkali-alumina catalysts to remove butadiene and to isomerize 1-butene to 2-butene. The optimum temperature for the isomerization was 100°C. The isomerization was greatly dependent upon a few mole percent of hydrogen, and no reaction took place without hydrogen under the experimental conditions.1- Butene appears to be isomerized via the half hydrogenated state of double bond.
    Download PDF (162K)
  • Kazuo KITAMURA, Saburo YOSHIDA, Shigeo TSUBAKI
    1972 Volume 1972 Issue 11 Pages 2219-2221
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The diffusion rates of Dispersol Fast Scarlet B from the dyed polyethylene terephthalate (PET) film to an undyed one in the atmosphere with relative humidity (RH) of 0, 50, 75 and 100% were determined at various temperatures.
    A cylindrical film roll was used and the dyed film is firstly coiled round on it repeatedly and secondly the undyed one.
    In the cases of 0 and 100% RH, the temperature dependence of diffusion coefficient obeyed the equation which has been observed in the diffusion of disperse dyes in PET.
    Download PDF (194K)
  • Yasuo KIKUCHI, Masahiko NAKAISHI
    1972 Volume 1972 Issue 11 Pages 2221-2222
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Electroinitiated polymerization of indene was studied in various solvents and it was found that the polymerization proceed only in an acetic anhydride solution containing lithium perchlorate. From the results of the IR and UV spectroscopic studies, it was presumed that the structure of polymers formed in the acetic anhydride solution were similar to that produced by the usual cationic polymerization, and that a small number of acyl and/or acetyl groups existed in the polymers. Moreover, it was speculated the polymerization initiated by electro- lysis was proceed by a cationic mechanism.
    Download PDF (141K)
  • Hirofusa SHIRAI, Nobumasa HOJO
    1972 Volume 1972 Issue 11 Pages 2223-2225
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Infrared spectra of poly (vinyl alcohol) (PVA) films complexed with copper On ion in the varying Tc, i241 THL ratio at pH 10.5 were measured in the range of 4000 to 200 cm-1 (Fig.1). The carbonyl-oxygen absorption bands are slightly shifted towords lower frequencies at 1440, 1326, 1093 and 849 cm-1 by the complex formation. The magnitude of this shift increased with the increasing of Tco+/ THL ratio (Fig.2). The value of 4430/D2945, D1326/D2945, D1093/D2945 also increased with the increasing of Tc4 Tin ratio (Fig.3). By the copper (11) complex formation, the new bands appeared at 878 cm-1 and 605 cm-1, seemed to be Cu-0 stretching frequency or chelate ring deformation between copper (II) ion and PVA unit (Fig.4).
    Download PDF (229K)
  • Toru NISHIWAKI, Atsuyuki NINOMIYA, Shizuo YAMANAKA, Kinzi ANDA
    1972 Volume 1972 Issue 11 Pages 2225-2226
    Published: November 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Polychlorinated biphenyl (PCB) can be successfully dechlorinated by the UV-irradiation in alkaline 2-propanol solution.400 ml of 2-propanol solution containing 1.84 x 10-2 mol of PCB and 2.40 X 10-2 mol of sodium hydroxide was irradiated with 100 W-high pressure mercury lamp under the flow of bubbling nitrogen at about 30°C. Percentages of dechlorination of PCB were about 80%, 95% and 100% after 5 min, 10 min and 15 min, respectively. Biphenyl and sodium chloride were identified in completely dechlorinated solution.
    Download PDF (143K)
feedback
Top