NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1978, Issue 1
Displaying 1-33 of 33 articles from this issue
  • Yoshiaki KISO, Takane KITAO, Yoshinobu OTSUKI, Shigehisa IWAI
    1978 Volume 1978 Issue 1 Pages 1-6
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Permeabilities for amides, ureas and substituted benzenesulfonic ions in reverse osmosis process have been studied by using porous cellulose acetate membrane at a pressure of 40 kg/ cm. The solute separation of amides and ureas was in the range 6.0, -60.9% and that of substituted benezenesulfonic ions was in the range 86.5-93.8%. The solute separation increased with decreasing hydrophilicity of solute. The partition coefficient of solutes in water-1- octanol was used as the index of hydrophilicity. The solute separation data for amides and ureas were also correlated with Taft's Ea* and E E. values and the data for substituted benezenesulfonic ions with Hammett's a values. These values represent the polar or steric effect of substituents. The solute separation increased with the decrease of Ea* and EE, values for amides and ureas and with the decrease of a value for substituted benezenesulfonic ions. It was found that ln (D/10) varied linearly with Ea*. However, a definite relation between the solute separation and the electron-density on oxygen atom, which was calculated by CND0/2 method, was not obtained.
    Download PDF (339K)
  • Toshiro SUZAWA, Ryozo ISHIMOTO
    1978 Volume 1978 Issue 1 Pages 7-10
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to study the surface adsorption behavior of the polyelectrolyte onto the fiber, 4potential of Nylon 6 fiber was measured in aqueous acidic solutions (pH 3.5) of sodium polyacrylate (Na-PAA). The amount of Na-PAA adsorbed on unit area of the fiber surface was calculated on the basis of the -potential values. With increasing Na-PAA concentration, the sign of -potential changed from positive to negative and the amount of the polymer adsorbed on the fiber surface (expressed in g/cm2-fiber) increased to reach the saturated value. These results suggested that the fiber interacted with the polyelectrolyte mainly in terms of an electrostatic force. The saturated value of the polyelectrolyte adsorbed, AB, decreased with the increasing degree of polymerization. The slope a of the straight line, derived from the relationship between As and M, the molecular weight, was equal to or slightly less than zero. It was, therefore, suggested that the polyelectrolyte molecule lay flat on the fiber surface with the same dimension as those in the solution.
    Download PDF (199K)
  • Yoshiko KITA, Sachio MATSUMOTO, Daizo YONEZAWA
    1978 Volume 1978 Issue 1 Pages 11-14
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The migration of water through semi-permeable membrane from the outer aqueous phase to the inner W/O-type emulsion phase was measured. The results supported the idea that water permeates through the oil layer in the W/O/W-type multiple-phase emulsion systems. This phenomenon could be recognized clearly in the presence of a small amount of solute in the inner aqueous-phase such as sodium chloride, glucose, etc. Therefore, it seems probable that osmotic pressure plays an important role in the permeation of water through the oil layer. The mechanism of the water permeation may be interpreted as follows; (1) incorporated micelles of hydrophobic and hydrophilic emulsifying agents in the oil layer carries water so as to diminish the concentration gradient of solute between the two aqueous phases, or (2) water molecules pass through thin lamellae of the adsorbed emulsifying agents which partially form in the oil layer due to the fluctuation of the thickness.
    Download PDF (578K)
  • Akihiko NIINA, Kaoru FUJIMOTO, Taiseki KUNUGI
    1978 Volume 1978 Issue 1 Pages 15-20
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The complete oxidations of carbon monoxide and ethylene over supported palladium, particularly the effect of halogenide ions on these reactions, were investigated in detail. In the oxidation of ethylene, hysteresis phenomena in the catalytic activities were observed when the reaction temperature or the partial pressure was varied. These are probably caused by the strongly adsorbed species coming from ethylene. While in the oxidation of carbon monoxide, the activity of the catalyst did not alter (Fig.1, Table 1). At high temperatures, the activity of the catalysts containing sodium halide as an additive altered with time irreversibly, accom- panied with the outflow of hydrogen halides. The addition of sodium halides exhibited drastic effects similarly on both reactions. The degree of suppression of both reactions was in the order of NaI NaBr NaCl Blank NaF (Table 2, 3, Fig.5), which accords with the sequ- ence of the spectrochemical series of halogenide ions. Sodium bromide suppressed the adsorption of carbon monoxide, so that the oxygen adsorption was less affected by carbon monoxide (Fig.6). The followings are deduced. The halogen is adsorbed as the ion on the surface of palladium particles and acts as a modifier of palladium. Halogenide ion except fluoride ion not only decreases the surface area of palladium, but it acts as an electron donor and changes the electronic properties of palladium surface, resulting in the suppression of the complete oxidation reaction. On the contrary, fluoride ion, which is a strong electron acceptor, likely promotes the complete oxidation reaction. Pressure dependencies of both reactions are very similar too, if the difference in their adsorption constants is taken into account (Fig.2, 3, 4). It is suggested that the Langmuir-Hinshelwood mechanism is dominant in the oxidation of carbon monoxide under these reaction conditions.
    Download PDF (451K)
  • Yuichi MURAKAMI, Yukio SAEKI, Keiichi ITO
    1978 Volume 1978 Issue 1 Pages 21-26
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The Beckmann rearrangement of cyclohexanone oxime (CHOX)to e-caprolactam (e LTM) in vapor phase has been investigated using a tubular type flow reactor . Solid acid catalysts used were Si02Al208-H (A1, 03 25%), Si02.Al203-L (Al20, 12. 5%), A1, 03 B20, , Si02 Mg0, phosphoric acid on kieselguhr, and aluminum sulfate. It was found that SiO2 - Al20, catalysts showed the highest activity and selectivity (Figs. 1 and 2) and that Al203 - B208 was similar to the former catalysts in selectivity but lower in activity than the former (Fig . 7). The activity and selectivity of phosphoric acid on kieselguhr was lowest (Fig. 8). Decline in activity with time-on-stream by the deposit of carbonaceous substances was observed in all the catalysts tested (Fig. 4) but no change in selectivity was found (Fig. 5). Effects of the reaction temperature, the mass of catalyst and the partial pressure of CHOX on the activity and selectivity were investigated with SiO, A1, 03-H. The selectivity was little affected by the reaction conditions (Fig. 2). By a pulse technique, adsorption of e LTM (Fig. 10) on and pulse reaction of CHOX (Figs. 11 and 12) over Si02Al203-H alone and with e LTM impregnated were carried out. The results showed that e LTM adsorbed strongly on the catalyst but CHOX more strongly and that the rate of reaction was the zeroth order for CHOX (Fig . 3). The mechanism of reaction is discussed. It appears that the stronger the acid strength of catalysts is, the higher the activity for the Beckmann rearrangement of CHOX and the selectivity to e LTM are. By-products like nitrile and carbonaceous substances seem to be formed from an intermediate on the catalyst surface in the course of the rearrangement .
    Download PDF (332K)
  • Keisuke MITA
    1978 Volume 1978 Issue 1 Pages 27-31
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The initial reaction rate of liquid phase hydrogenation of acetone was found to be remarkably promoted by the addition of magnesium to nickel catalysts where the catalysts prepared from nickel nitrate and magnesium nitrate by coprecipitation were reduced at relatively low temperature without prior calcination. In order to investigate the promoting effect of magnesium, kinetic studies on the hydrogenation of acetone over nickel catalysts with various amounts of magnesium were carried out at 20°C over a wide range of acetone concentration. The kinetics were found to be successfully interpreted by the rate equation based upon Langmuir-Hinshelwood mechanism, except for the catalysts with more than 0.3 in atomic ratio of Mg/Ni. The optimum atomic ratio of Mg/Ni were found to be 0.02-0.13 where the catalytic activities were about 10 times as great as the activity of a nickel catalyst without magnesium. The apparent activation energies were decreased from 7.5-40 kcal/mol to 2-7 kcal/mol and the adsorption strength of hydrogen relative to that of acetone was increased by the addition of magnesium.
    Download PDF (303K)
  • Keizi HASHIMOTO, Shozi WATANABE, Kimio TARAMA
    1978 Volume 1978 Issue 1 Pages 32-35
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The oxidation of NO with oxygen chemisorbed on reduced MoO3-Al203 catalysts was studied: by various physico-chemical techniques. The oxidation over the catalysts was rapid at room temperature. The data of ESR measurements suggested that NO reacted with 02- but did not with 0- at the temperature lower than 100°C. The number of active sites on the catalyst was determined on the basis of the saturation value of adsorbed NO and 02 and was found to be nearly equal to the content of Mo(V), [Mo5+]w in the supported molybdenum oxide soluble in water. The variation of the number of active sites with the amount of supported molybdenum oxide and the extent of its reduction agreed well with the variation of [Mo5+]w with them. It is concluded that the oxidation activity is attributable to [Mo5+]w
    Download PDF (261K)
  • Ryuichi NAKAMURA, Kazuto ICHIKAWA, Etsuro ECHIGOYA
    1978 Volume 1978 Issue 1 Pages 36-41
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The metathesis of various olefins such as propene, 2-pentene, and 1-hexene over watertreated or untreated Re207/Al208 catalysts (Re/Al 0-7/93) was investigated and the structure and nature of active sites on the catalysts were discussed. A comparison of the catalytic activity with informations as to the structure or nature of the active sites estimated by EPR, X-ray diffraction, and kinetic studies suggests that there are three types of rhenium V10 oxides in the Re207/Al203 system. At lower Re concentrations (Re/Al1/99: K-region), Re is so strongly held in the distorted lattice of an alumina support that the catalytic activity is abnormally low without hydrogen reduction of the catalyst at elevated temperatures. At higher Re concentrations (Re/Al3/97 M-region), there is a layer of rhenium V10 oxide, which comprizes Re-O-Re bonds, and the catalytic activity is elevated by a doublypromoting effect of the Re ions. This effect should be attributed to the nature of the Re-0- Re species, that is to say, (1) the Re-O-Re species has weak bonds, so it can be easily converted to an active species by the reduction and complexation with olefins, and (2) Re7+ ions around the active center increase the positive charge on it, and consequently the adsorption of olefins is promoted. In a moderate Re concentration region (1/99Re/A13/97 L-region) the active site has a relatively strong Re-O-Al bond which does not afford the doubly-promoting effect, so that the catalytic activity is not very high. The metal oxide, if the electronegativity of the metal ion is large, can be used as a substitute for the weak bonded rhenium(W) oxide at M-region. Such a metal oxide transforms the catalytic activity and nature of the active sites at K or L-region into those at L or M-region.
    Download PDF (427K)
  • Jae Hee OH, Tatsuru NAMIKAWA, Minoru SATOU
    1978 Volume 1978 Issue 1 Pages 42-45
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The formation of barium ferrite powder with high coercivity by the thermal decomposition of the coprecipitated oxalate from aqueous solution containing Fe2+ and Ba2+ was investigated by means of X-ray diffraction, differential thermal analysis, thermogravimetry, and measurement of magnetic behaviour. These results were compared with those of the mixture of iron and barium oxalates.
    When the mole ratio of Na2C204 to FeCl24H20(8.95g) and BaCle2H2O (1.00g) was above 1.5, the coprecipitate (Fe2+/Ba2+=11) of the complex oxalates was formed available for preparing BaFe12019 powder. The coprecipitated oxalate converted into the ferromagnetic phase of BaFe12019 by the heat treatment above 500°C. The single domain BaFe12019 particles of about 0.5 were obtained at 800°C. The obtained powder were 50 emu/g in the magnetization and 4.8 kOe in the coercivity at the maximum applied field of 10 kOe.
    Download PDF (613K)
  • Naoaki KUMAGAI, Hanzo MASE
    1978 Volume 1978 Issue 1 Pages 46-51
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reactions between phosphoryl triamide and MCl2 {M = Mn(II), Co(II), Cu(II) (as trace nutrient)} in aqueous solutions were studied, and the products were examined by chemical analysis, ion-exchange chromatography, paper chromatography, X-ray powder diffraction method, IR spectroscopy, etc.
    M {PO(NH2)3}2Cl2(M = Mn(II), Co(II)) complexes that contained a small amount of diamidophosphate (6-11%) were obtained by adding acetone to the mixture of MCl2(M = Mn, Co) and PO(NH2)3 (P/M atomic ratio 1.5-2.0) at room temperature. On the other hand, PO(NH2)3 was hydrolyzed rapidly in the initial stage of its reaction with Cu(II), resulting in a precipitation of Cu(II) phosphates. Further, when a solid mixture of PO(NH2)3 and MCl2(M = Mn(II), Co(II)) with a molar ratio of 2:1 was dried in a vacuum desiccator after it had been wetted with a small amount of water and mixed thoroughly, it also yielded M{PO(NH2)3}2Cl2 complexes as major constituent.
    In the reaction of PO(NH2)3 with MCl2(M = Mn(II), Co(II), Cu(II), though it was hydrolyzed as the reaction time became longer (-72hr) and at higher temperature (-60°C), metal salts were precipitated that consisted of mainly amidophosphate, orthophosphate and diphosphate. The percentage of diphosphate in the hydrolysis products were 27%, 33% and 25% for Mn(II), Co(II) and Cu(II), respectively.
    The hydrolysis process of PO(NH2)3 in the reaction with MCl2 is discussed.
    Download PDF (687K)
  • Kohei URANO, Yuji NAKAMICHI
    1978 Volume 1978 Issue 1 Pages 53-59
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Oxidation reaction rates of sodium sulfite solution of the initial concentration ranging from 3.3 10-4 to 6.7x10-3 mol/l were studied at 25°C varying the concentration of copper from 1.7x10-8 to 2.7 x10-7 mol/l. The concentration of dissolved oxygen was controlled by blowing the purified 25-100% oxygen and nitrogen gases into the solution (Fig.1). The reaction rates were calculated from the pH changes of the solution.
    The error and the time lag of pH measurement, the light and the difference of copper salt did not affect the reaction rate. The trace impurities in the water, however, affected considerably on the reaction rate (Fig.2). The sulfite dissolved in pure water is not oxidized by oxygen. The reaction rates are expressed by the following equations when the concentration of copper is higher than 1. Ox10-7mol/l (Fias.5.7.9.10).
    d[St]/dt=1.3x103[S032-]3/2[Cu]1/2[02]0: [S032-][Cu]8x10-12(mol/l)2
    =1.7x107[S032-]2[Cu]1/2[02]0:[S032-][Cu]8x10-12(mol/l)2
    where St shows the total sulfite and [ ] shows molar concentration. The reliable rate equation could not be obtained when the concentration of copper is lower than 1.Ox 10-7mol/l though the reaction rate was increased with addition of trace copper. When [S032-] [Cu] is larger than 8x10-12(mol/l)2, the radical chain reaction mechanism which was proposed by Backstorm is seemed to be reasonable.
    Download PDF (470K)
  • Shosaku ISHIHARA, Nobuyuki OHKUMA, Keizo UEMATSU, Nobuyasu MIZUTANI, M ...
    1978 Volume 1978 Issue 1 Pages 60-63
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Interaction of water vapor with zinc. oxide was studied mostly by Differential Thermal Gas Analysis (DTGA) installed with a thermal conductivity detector, a humidity - sensor and an oxygen sensor (Fig.1).
    On heating ZnO samples, their weight were found to decrease around 270°C(1.3 mglg ZnO) and around 500°C(0.12 mg/g ZnO) where endothermic reactions occurred (Fig.2). DTGA showed that these weight losses were due to the evolution of water vapor (Fig.3 and 4). On cooling the sample from 600°C, the weight gain of the sample occurred only, around 480°C; the interaction reaction was exothermic. The amount of water vapor adsorbed in the sample in this temperature range was equal to that evolved from the sample at 500°C on heating. However, apparently the water vapor evolved from the sample at 270°C on heating was not re-adsorbed on cooling. Powder X-ray diffraction analysis on quenched samples showed that their structure was not affected by the adsorption of water vapor. The interaction of water vapor with ZnO at 500°C was reversible. The change in enthalpy for this interaction was estimated to be 38 kcal/mol from the vant Hoff's plot of the interaction temperature against the vapor pressure of water (Fig.5).
    The effect of preheating was studied. Both the specific surface area of the sample and the amount of water vapor evolved from it at 500°C decreased with the increase of preheating temperature above 480°C, however, they were not affected by preheating below 480°C (Fig.6).
    Download PDF (285K)
  • Seizi TAKEUCHI, Kazunori FUZITA, Fumito NAKAZIMA, Yoshiziro ARIKAWA
    1978 Volume 1978 Issue 1 Pages 64-67
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A Rapid separation of amino acids in protein hydrolyzates by ion-exchange chromatography was investigated.
    Effects of resin particle diameters, flow rates of mobile phases and solvents in eluent on the separation were studied using a relation between separation of two peaks and numbers of effective theoretical plates.
    A number of effective theoretical plates of 1750 for serine was required to attain 70% of separation between threonine and serine at a separation factor of 1.10. In the present study the height equivalent to a theoretical plate for glycine of O.025 mm was obtained when resin particles of a diameter of 5-6 m were used. It was found that the optimum ethanol concentration in the eluent was 12 vol% for the separation between threonine and serine and between glycine and alanine. It was possible to separate 19 amino acids in protein hydrolyzates within 40 minutes using a single column of 2.6 mm i. d. and 150 mm long packed with a strong cationic ion-exchange resin of 5-6 pm diameter under the opti mum conditions.
    Download PDF (223K)
  • Masao MARUYAMA, Michiko KAKEMOTO, Hitoshi SHIBUYA
    1978 Volume 1978 Issue 1 Pages 68-74
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The gas chromatographic behavior and direct microdetermination of five substituted phenyl methylcarbamates were investigated by gas chromatography-mass spectrometry.
    In direct gas chromatographic measurements, the conversion of carbamates into the corresponding phenols depended on the characteristics of stationary liquid rather than injection and/ or column temperatures. When PEG 20 M used as the stationary liquid, carbamates were completely converted into the corresponding phenols, but this was not the case for FFAP and Apiezon L, where carbamates were partially converted and mostly eluted from GC column as carbamates. This phenomenon was confirmed by mass spectra as well as gas chromatographic and thin layer chromatographic retention data.
    When PEG 20 M was used as the stationary liquid, carbamates were determined as the corresponding phenols by multiple ion detector (MIS). Carbamates were selectively determined by MIS setting to their characteristic fragments. For example, m-tolyl methylcarbamate (TMC) and o-isopropylphenyl methylcarbamate (i-PPMC), which have the same gas chroma- tographic retention time, were selectively determined by setting mle to 108 for TMC and 121 for i-PPMC. A linear relationship held between the weight of carbamate and its response and a minimum detectable amount was found to be as low as 1x10-10 - 3x10-11g as carbamates. This method was applied to analysis of pesticide residues in various plants. Determination of small amount of substituted phenyl methylcarbamates can be done easily, selectively and sensitively, without complicated pretreatments such as delivatization.
    Download PDF (401K)
  • Katsuyoshi SHIBATA, Hiromitsu SADAKA, Masaki MATSUI, Yoshimi TAKASE
    1978 Volume 1978 Issue 1 Pages 75-78
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Photolysis of N-acetylhydrazobenzene [1] gave azobenzene mainly through the cleavage of the CO-N bond in preference to the N-N bond. On the contrary, the major product from N, N'-diacetylhydrazobenzene [2] was acetanilide which was resulted from the cleavage of the N-N bond. Rearrangement products of both photo-Fries type and semidine type were obtained as minor products. The semidine was only comprised of ortho compound. Benzidine rearrangement scarcely occured during the photolysis.
    Download PDF (210K)
  • Shizunobu HASHIMOTO, Masanao YASUDA, Masami MOURI
    1978 Volume 1978 Issue 1 Pages 79-81
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The photochemical reaction of benzophenone in water was investigated by using an immersion-type 130-W high pressure mercury lamp in a stream of nitrogen.
    Benzopinacol, o-hydroxybenzophenone, and p-hydroxybenzophenone were isolated as the reaction products.
    Investigation on the effects of wavelength and pH suggested that benzopinacol was produced by an electron transfer from OH- anion to n- excited benzophenone, and o- and p-hydroxybenzophenones were produced by the nucleophilic attack of OH- anion to - excited benzophenone.
    Download PDF (174K)
  • Shizuko ISOKAWA, Shinzo TSURUOKA
    1978 Volume 1978 Issue 1 Pages 82-85
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of 1, 4-dichloro-5, 8-dinitroanthraquinone [1] with aniline in the presence of potassium acetate and copper ( II ) acetate was investigated at the temperature ranging from 100 to 155°C. The reaction products were isolated by column chromatography. Based on their physical properties, the structures were confirmed. It was observed that 1-anilino-5, 8-dichlo- ro-4-nitroanthraquinone [3] was a sole product below 100°C. But at 115-120°C, [3] accom-panied 1, 8-dianilino-4-chloro-5-nitro- [4] and 1, 5-dianilino-4-chloro-8-nitro-anthraquinone [5] as minor products.4, 5-Dianilino-1-nitro-[6] 1 1, 5-dianilino-4-nitro-[7], 1, 4, 5-trianilino-8- nitro-C [8], 1, 4, 5, 8-tetraanilino-[9], 1, 4, 5-trianilino-[10] and 1, 4-dianilino-anthraquinone [11] were obtained at 150-155°C.
    It was revealed that the reaction path of [1] with aniline was as follows: Below 100°C, the displacement of one of the nitro groups of [1] by aniline occurred and at about 115°C, each chloro substituent of [3] thus obtained began to react with aniline with almost the same reactivity, [4] and [5] being obtained. The dechlorination of [3], [4] and[5] occurred giving [11], [6] and [7], respectively, at 150-155°C. While, the displacement of chloro substituents of [4] and 5 by aniline gave [8], which was converted to [9], and of nitro groups of [6] and [7] gave [10].
    At this temperature [8] was the main product at early stage of reaction and its nitro group was not so reactive that [8] partly remained after 24 hr.
    Download PDF (223K)
  • Seisaku INADA, Shigeto HIRABAYASHI, Kazuhiro TAGUCHI, Mitsuo OKAZAKI
    1978 Volume 1978 Issue 1 Pages 86-92
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The thermal reaction of N-allyl-N-mesyl- and N-allyl-N-tosyl-substituted anilines in N, N-dibutylaniline containing a small amount of triphenylphosphine gave the Claisen-type rearranged products, o-allylsulfonanilides, in highly pure states and excellent yields, except for N-allyl-N-mesyl-p-dimethylaminoaniline. The rearrangement of the latter was accompanied by the formation of a considerable amount of the deallylated product. Anilides, whose methyl groups exist at the ortho-position, gave no para-rearranged products in contrast to expectation. As to the effect of substituento at para-position, it was found that electron donative groups promote the rearrangement and electron attractive groups retard it.
    The effect of sulfonyl group on amino nitrogen atom in the rearrangement was discussed in comparison with the thermal rearrangement of N-allylanilines.
    Download PDF (363K)
  • Takaari YUMOTO, Kozo ISEDA
    1978 Volume 1978 Issue 1 Pages 93-96
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The T-ray induced reactions between tetrahydrofuran and chlorinated methanes were studied. Main products were 4-chloro-l-butanol [1] and 5-chlorooctahydro-2, 2'-bifuran [2]. No expected substitution product between tetrahydrofuran and chlorinated methane was detected. The amounts of [1] and [2] approximately equaled, and the rates of formation were nearly proportional to a dose rate. The yields of [1], [2], and HCl changed in the following order: CCl4CHCl8CH2Cl2. The reactions induced by organic peroxides were also studied.
    Download PDF (355K)
  • Yozo OSHIMA, Fumikazu CHIDA, Hiroshi OHNUMA
    1978 Volume 1978 Issue 1 Pages 99-101
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It has been found that adductibilities of monomethylalkanes were different from those of normal alkanes because of a side methyl group of the guest molecules. In this study, precise lattice constants of 2-methyldodecane, 2-methylpentadecane, 2-methylheptadecane and 2-methyleicosane-urea adduct crystals were determined by powder X-ray diffraction method with the intention of comparing with those of normal alkane-urea adduct.
    The indexing of powder X-ray diffraction patterns of 2-methylalkane adducts showed that they belonged to hexagonal system as well as hexadecane adduct. The differences between the value of lattice constant co of hexadecane adducts and one of 2-methylalkane adducts were not significant but the values of ao of 2-methylalkane adduct were all larger than that of hexadecane adduct. Lattice constants as of 2-methylalkane adducts became smaller and tended to approach a constant value with the increasing carbon numbers of 2-methylalkane.
    From these results it was suggested that urea channels of 2-methylalkane adducts were expanded to a axis from that of hexadecane adduct owing to a side methyl group of guest 2-methylalkane molecule. The influence of a side methyl group on expansion of urea channel decreased with increasing of the chain, length of guest molecules. The each expansion was due to the stretching of one kind of N-H...0 bond of urea channel of normal alkane adducts. The latter estimation was supported by comparing the stretching energy of N-H...0 bond with the difference between heat of adduct formation of normal alkane and one of 2-methylalkane.
    Download PDF (216K)
  • Toshihiko ISHIKAWA, Akio SUGIHARA, Susumu NAGAI, Nobutami KASAI
    1978 Volume 1978 Issue 1 Pages 102-107
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to elucidate the appreciable splitting of the equatorial reflection (d=4. 1 A) in the X-ray photograph of the Nylon 12 sheets specimens drawn at temperatures 140-460°C with draw ratios of 2. 4-3. 0, crystallite orientation in Nylon 12 sheets after the uniaxial drawing, drawing by rolls, hot-pressing and roll-pressing has been investigated by means of X-ray diffraction. The split reflections are assigned to 001(+201) and 200, respectively, by Weissenberg photograph (Fig. 7). Specimens drawn uniaxially under the aforementioned conditions showed a sign of biaxial orientation (Fig. 1(b)), the (200) planes of the crystallites being arranged parallel to the plane of the sheet (selective uniplanar orientation), whereas those drawn uniaxially at temperatures lower than 130°C with draw ratios 2. 4-4. 0 did not show any sign of biaxial orientation (Fig. 1(a)). The splitting appears to be caused by the selective uniplanar orientation as well as by the shortening of d-spacing and sharpness of the 200 reflection (Fig. 5, Table 2). In addition, with the specimen drawn out to about 4 times its original length at high temperatures (140-460°C), the splitting became obscure due to the decrease in intensity of the 200 reflection and increase in the width of the two reflections (Fig. 9, 10, Table 1). Furthermore, the modes of orientation of crystallites after various processings are discussed (Fig. 1(a), (b), (c), 3, 4, 6).
    Download PDF (886K)
  • Masakatsu YONESE, Hideya TSUGE, Hiroshi KISHIMOTO
    1978 Volume 1978 Issue 1 Pages 108-112
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The molal osmotic coefficients (q) of chondroitinsulfate A and C (ChS-A and -C) were measured by vapor pressure osmometry. The measurements were made at 25, 37 and 60°C in the concentration range from about 0. 005 to O. 2 monomol per kg of water. ChS-A salts used in this work Li-, Na-, K-, NH4-, (CH3)4N-, Mg-, Ca-, Ba- and AlChS-A. ChS-C salts were Li-, Na-, K-, Mg-, Ba- and AlChS-C.
    Experimental / values were found to be about O. 45 at the valency of counterion zg =1, about 0. 2 at zg = 2 and about O. 15 at zg = 3, showing that as the valency of counterion increased the molal osmotic coefficient decreased. As there were nearly no difference between values of ChS-A and of ChS-C which are structural isomers with each other, it was concluded that the interaction between macroion and counterion did not depend on the position of the sulfate group.
    Experimental values well agreed with the theoretical ones calculated by Manning's limiting law when the distance between the neighbouring charges of macroion, b, was assumed to be 6. 3A. It was concluded that the interaction of macroion with counterion at ChS-A and -C salts could be explaind by an electrostatic effect.
    Download PDF (284K)
  • Masanori HIROSE, Yoshio IMAMURA
    1978 Volume 1978 Issue 1 Pages 113-116
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The system consisting of an organic solvent-H20-KMnO4-a quaternary ammonium salt was examined for the condensation polymerization of 2, 4, 6-tribromophenol (TBP) promoted by a phase-transfer catalyst. The yield of the polymer in the reaction varied depending on the combination of an organic solvent and a quaternary ammonium salt, and reached a constant value after addition of a very small amount of (n-C4H9)4NBr. The effect of stirring rate in the reaction system and of the choice of the quaternary ammonium salt indicated that a phase-transfer mechanism was predominant in the condensation polymerization of TBP in the system of C6H6-H20-KMn04- (n-C4H9)6NBr. However, this mechanism seemed to be less important, when other organic solvents such as CS2 or CCl2=CCl2 were used together with quaternary ammonium salts.
    Download PDF (267K)
  • Hirofusa SHIRAI, Yutaka Nio, Akio KUROSE, Sadao HAYASHI, Nobumasa HOJO
    1978 Volume 1978 Issue 1 Pages 117-123
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The overall stability constant of the Fe (iii)-poly (vinyl alcohol) (PVA) complex is about 1010 times as large as that of the Cu (ii)-PVA complex. The complex formation of PVA with the Fe (iii) ion proceeds smoothly in an acid solution, while the corresponding reaction with the Cu(ii) ion take place in the neutral or alkaline region. Therefore, a new Cu (ii) and Fe (iii)-mixed complex is expected to be formed by increasing the pH of the solution of the Fe (iii) -PVA complex in the presence of the Cu (ii) ion. Electronic spectra, pH-titrations and viscosity have been measured on the PVA aqueous solutions containing the Cu (ii) and Fe (iii) ions. Under a limited metal concentration ( TFe3+/THLO.014, TCu2+/THL0.08), the mixed metal complexes are obtained as expected. The viscosity measurements have revealed that the Cu(ii) ion combines with the residual OH groups in the Fe (iii)-PVA complexes and that intra-molecular bonding is formed. The Fe (iii)-PVA complexes containing a larger amount of Fe (iii) ion (Fe2+/ THL over 0.014) have a compact structure, and therefore additional complex formation with Cu (ii) ion is not expected; the Cu (ii) complexes are in unstable complex forms. To clarify these points, formation constants and thermodynamic parameters for the complex formations of the Cu (ii) ion with the residual OH groups in Fe(iii)-PVA complexes have been determined by applying the modified Bjerrum's and Ringbom's methods. The second succesive stability constants (log k2), and values for the complex formations decrease with increasing TFe3+/THL, in the Fe (iii) -PVA complexes. These mixed complexes which have the Cu (ii) ion in the incompletely four-coordinated planar structure are found to display considerable activity for the decomposition reaction of hydrogen peroxide.
    Download PDF (440K)
  • Shinichi TOYOSAWA, Yoshihiko TAKEUCHI, Mitsuru FUJIMURA, Hidenari INOU ...
    1978 Volume 1978 Issue 1 Pages 124-127
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The size distribution of rubber particles has been measured in a highway atmosphere in the neighborhood of a tunnel in Tokyo. Since zinc oxide (ZnO) exists in tire tread compounds as a vulcanisation activator, the distribution of zinc in particles was also measured and compared with the size distribution of rubber particles. The samples were collected by using 8 stage Andersen impactors during 8 days. vulcanized rubber components collected were extracted with o-dichlorobenzene in a stream of oxygen at 180°C. After evaporation of o-dichlorobenzene, the rubber components obtained were analyzed by pyrolysis-gas chromatography. Tlie size distribution, curves of rubber particles show two maxima at 5 pm and ca.0.5 Aim. Zinc extracted with acids from the samples was determined by atomic absorption spectrophoto- metry. The distribution curve of zinc in particles also shows amaximum at 5 pm, however this maximum was not observed in the atmosphere far apart from highways. The comparison of zinc amount determined with that calculated in the rubber particles showed that the maximum of the distribution of zinc in particles at 5 gm can be ascribed to the amount of zinc 'contained, in the tire treads.
    Download PDF (242K)
  • Juziro NISHIJO
    1978 Volume 1978 Issue 1 Pages 128-130
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    As a part of a series of studies on the complex formation of organic acids with amino acids, the heats of reaction (H) for formation of solid complexes of glycine with oxalic acid, maleic acid, malonic acid, glutaric acid and aminomalonic acid were determined. The values for H, were found 1-5 kcal/mol from the measurements of the heats of solution of the solid complexes and their components. The correlation of heats of reaction with the PKa values of the organic acids was discussed.
    Download PDF (195K)
  • Koichi WADA, Takao YOTSUYANAGI, Kazuo AOMURA
    1978 Volume 1978 Issue 1 Pages 131-133
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Nickel complex, Ni1, 22-, was extracted quantitatively with Qion into chloroform as an ion pair of Q2[NiL2] over the pH range from 6.2 to 9.6. The composition of the ion pair was confirmed by the molar ratio method and elemental analysis of the isolated crystals. The ion pair had two absorption maxima at 548 nm (= 4.66 X 104) and 589 nm (e = 4.72 X 104), whereas the reagent ion pair, QHL, gave only a very small absorbance at 589 nm (e = 31). Beer's law was confirmed over the nickel concentration range from 0 to 13.6 mmol. m-8, and the sensitivity for A= 0.001 was 0.0012 pg iNi cm-2. The extraction constants, Kex= [Q, 13], , [Q+]-n[Bn-]-(where jo indicates the species in chloroform), for Q2[NiL2] and QHL were estimated to be 101770 and 10977003, respectively (at 25°C, 1=0.1).
    Download PDF (166K)
  • Takaari YUMOTO
    1978 Volume 1978 Issue 1 Pages 134-137
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Photochemical addition reaction of alcohols (methyl, ethyl, propyl, isopropyl) and ethers (tetrahydrofuran, tetrahydropyran, 1, 4-dioxane, dibutyl ether, diisopropyl ether) with tetrachloroethylene in the presence of di-t-butyl peroxide at a room temperature was investigated.The products were saturated 1 : 1-adducts [ 1 ] and unsaturated compounds [ 2] containing CHCl2-CCl2- and CCl2=CCl- group, respectively.
    R-H + CCl2=CCl2→RCCl2-CCl2H+RCCl=CCl2+HCl
    These results were compared with those obtained by using di-t--butyl peroxide with or without T-ray irradiation, particularly in terms of relative yields of [ 1 ] and [ 2 ]. The yields of products changed in the following order with regard to alcohols : propylethylmethylisopropyl alcohol and with regard to cyclic ethers : tetrahydrofurantetrahydropyran1, 4-dioxane.
    Download PDF (201K)
  • Sotaro MIYANO, Hisazumi WATANABE, Hidehiko USHIYAMA, Yasuhisa YAMADA, ...
    1978 Volume 1978 Issue 1 Pages 138-140
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Chlorination of alcohols with copper (II) chloride and triphenylphosphine is described (Table 1). Aliphatic alcohols afforded the corresponding chlorides in fairly good to high yields, but cyclohexanol and menthol gave poor results because of the olefin-forming side reaction.Phenol was not chlorinated under the reaction conditions. (-) - (R) - 2-Octanol gave (+) - (S)-2-chlorooctane with 81% net inversion. The reaction was explained in terms of the nucleophilic attack of chloride ion on the alkoxyphosphonium intermediate [ 2 ] (Eq. (3)) . Bromides were also obtained by the use of copper (II) bromide in the place of the chloride (Table 2) .
    Download PDF (158K)
  • Shizuko ISOKAWA, Masaru HAGIWARA, Shinzo TSURUOKA
    1978 Volume 1978 Issue 1 Pages 141-143
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The oxidation of 1-amino-4-nitroanthraquinone [4] 1with ammonium peroxomonosulfate followed by chromium trioxide gave 1, 4-dinitroanthraquinone [6] (mp 328.5-333.5°C (decomp. ), yellow needle-like crystal) in 13.6% overall yield (Scheme 1). Its chlorinated derivatives were obtained by the chlorination of [6], and their structures were confirmed by IR and NMR spectra and elemental analysis. These products seem to be useful as intermediates of dyes. At 55°C, 1, 2, 3, 4-tetrachloro-5, 8-dinitroanthraquinone [14] was obtained in 68.2% yield. While, 1, 4-dichloro-5, 8-dinitro-[12], 1, 2, 4-trichloro-5, 8-dinitroanthraquinone [13], and [14] were obtained at 40°C (Table 1). Similarly, [12] was synthesized from 1-amino-5, 8-dichlor0- 4-nitroanthraquinone [10] prepared from 5-acetamido-1, 4-dichloroanthraquinone [8] by nitration followed by hydrolysis.
    Download PDF (188K)
  • Seiko NAN'YA, Eturo MAEKAWA
    1978 Volume 1978 Issue 1 Pages 144-146
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The benzene solutions of 1, 2-bis(2-substituted 3-phenyl-1-isoindoly1) ethylene [1 a-d] show a yellow fluorescence that disappears on standing for several days. From these solutions 1- (3-phenyl-1-isobenzofurany1)-2-(2-substituted 3-phenyl-1-isoindolyl) ethylene N-oxide [ 4 ] and N-substituted 2-benzoyl-benzamide [ 5 ] were isolated.1, 1'-Bis(phenylimino)-3, 3'-diphenyl2, 2'-biindene [ 2] and 3, 3'-diphenyl-1'-phenylimino-2, 2'-biinden-l-one [ 3 ] were the main reaction products when the N-substituent was the phenyl group. When the solutions were acidified, 3, 3'-diphenyl-2, 2'-hiindene-1, 1'-dione [6] was obtained in addition to [2], [ 3] or [ 4 ].
    Download PDF (125K)
  • Tsuneo MATSUDA, Noriaki ISHUIN, Shigeru SUZUKI, [in Japanese], Takashi ...
    1978 Volume 1978 Issue 1 Pages 147-149
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Oxidation of propane was carried out over various catalysts, including those used for the oxidation of lower olefines. The conversion of propane considerably enhanced at about 500°C. It was found that the addition of MoO3 suppressed the complete oxidation of propane and the oxides of the 8 th group metals such as Ni and Co also interfered with the cracking of propane.
    Download PDF (162K)
  • Kazuo SUGIYAMA, Tetsuo TABUCHI, Toshihisa MAESHIMA
    1978 Volume 1978 Issue 1 Pages 150-152
    Published: January 10, 1978
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The rates of thermal decomposition of 1, 4-dipheny1-1, 4-dimethyl-2-tetrazene and its five derivatives [1 a-f] were determined in acetonitrile by means of UV spectroscopy. The rate measurements were carried out in the temperature range 55-150°C. A Hammett plot against al- gives a line with a negative p. The tIH* value for the decomposition decreaies with electron-releasing substituents. The effect of [1 a-e] on the radical polymerization of acryl- onitrile was studied in N, N-dimethylformamide at 60°C. The rate of decomposition of [1 a-e] induced by poly (acrylonitrile) radicals increases with electron-releasing substituents. A linear relationship is obtained between chain-transfer constants (CT) of C. I a-0 and the avalues (Fig.2), when T was regarded to be 2.60. On the basis of the above results, the effect of substituents on the basic nature of anilino radicals is discussed.
    Download PDF (153K)
feedback
Top