NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1977, Issue 7
Displaying 1-30 of 30 articles from this issue
  • Tsuneyuki NISHIZAWA, Yoshiki TANAKA, Kozo HIROTA
    1977 Volume 1977 Issue 7 Pages 929-933
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    By using a thermal diffusion column as a reactor, catalytic decomposition of hydrogen sulfide and enrichment of the produced hydrogen by one process were attempted. We expected that a reaction accompanied with continuous separation of the products, i. e., hydrogen and sulfur, could increase the hydrogen yield in comparison with the reaction carried out in an ordinary reactor having homogeneous temperature distribution.
    Three thermal diffusion columns made of pyrex glass tubes and equipped with a quartz pipe heater were used (Fig.1 and Table 1). Surface of the hot wall was coated with catalysts as chromium sulfide, cobalt sulfide, nickel sulfide or iron sulfide. Measurement was carried out in a batch system, maximum pressure reaching ca.1.5 atm. The initial activity of the catalysts at the hot wall temperature (TO) of 500°C in Column C was in the following order: chromium sulfide> cobalt sulfide > nickel sulfide > iron sulfide (Fig.3). Catalytic activity did not decrease during a 35-45 hrs operation. Separation of hydrogen from hydrogen/hydrogen sulfide mixture (Fig.2) and effects of the hot wall temperature and the wall distance (Fig.4) were also studied. In addition, the distribution of hydrogen concentration along the column was determined on Column A (Fig.5).
    The results agreed 6with the expectation; e. g., 40% hydrogen sulfide was decomposed and 96 % hydrogen was obtained from the top of the thermal diffusion column in the best operating conditions (Column A, Th = 5 00°C, catalyst = chromium sulfide).
    Download PDF (316K)
  • Tomoyuki INUI, Takashi UEDA, Masatoshi SUEHIRO
    1977 Volume 1977 Issue 7 Pages 934-940
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Reduction rates of a supported preoxidized copper catalyst by various gases were measured by means of a micro-thermogravimetric analysis and a flow method. Then, the characteristics of the formed surfaces were compared with those observed in other oxidation reactions.
    Regardless of a variety of gases and a change in temperature, an increase in reduction ratio (x) with time (t) obeyed the Elovich equation, dxIdt=k exp(x), after an induction period, and decrease in it was observed in higher x region The passage clasification and the method of analysis shown in the previous paper on hydrogen, could be applied in the present case. The order for the apparent reduction rate constant (k) for each gas was CO>1-12>CF14> C31-18.
    The Arrhenius parameters were expressed by the following equation with a compensation effect.
    log A=EI (2.3 RT.) +log k,
    K obeyed the following rate equation of the Langmuir adsorption type with varying concentration of the reducing gas (p):
    k=xpl(i+kP).
    The reduction rate was markedly retarded in the below-mentioned order by the addition of small amount of acrylaldehyde, water, or carbon dioxide. The different reduction rate due to gaseous species was markedly dependent on the ability of retardation of the reaction products. The rate of oxygen uptake and activity of catalytic oxidation of CO or C31-16 on the surface formed by each reducing gas, corresponded all to the above-mentioned order of the reduction rate.
    Download PDF (609K)
  • Masao YAMANA, Tsutomu KAWATA
    1977 Volume 1977 Issue 7 Pages 941-946
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Stable colored films were produced on SnO2 cathode from heptylviologen dibromide [5], benzylviologen dibromide [6] and viologen polymers [7], [8] by electrochemical reduction in aqueous solution. The color of transmitted light was reddish purple for [5], [6], and blue for [7], [8]. Viologenpolymers [7] and [8] formed stable blue-colored films on the SnO2 cathode also in polar solvents. The electrochromic characteristics of [6] and [8] in aqueous solution were studied. The writing time decreased with increasing applied voltage, and tended to reach a constant value: In the case of [6], the writing time varied in inverse proportion to the concentration of C 6D. The electrochromic efficiencies were 0.12 (mC/cm2)-1 and 0.048 (mCI cm2)-1 for [6] and [8], respectively. The erasure time decreased with increasing applied voltage of oppsite polarity and was shorter than the writing time at the same applied voltage.
    Download PDF (323K)
  • Kazuo MANABE, Makoto OGAWA, Yasunobu YONEMURA, Teruichiro KUBO
    1977 Volume 1977 Issue 7 Pages 947-952
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It has been reported that the thermal decomposition of ammonium paramolybdate tetrahydrate in an air proceeds as follows:
    3 (NH4)20.7 Mo08.4 H20 (APM) 2 (NH4)20.5 MoO3 (M1)
    2 (NH4) O.8 MoO3 (M2) Mo03.
    This paper deals with the effect of mechanical grinding on the processes of thermal decomposition of APM, M1, and M2. The methods of TG, DTA, X-ray diffraction, mass spectroscopy and chemical analysis were used in this study.
    Two kinds of new intermediates (named A and B) were consecutively formed by mechanical grinding of APM and M1, and a new intermediate (named C) was produced by grinding of M2. Both two intermediates A and B decompose into MoO, through the intermediate compound M, in the course of thermal decomposition, but amorphous sample, obtained by grinding of M, more than a week, decomposes directly into MoO8 without passing M, in the course of thermal decomposition. The intermediate C decomposes through a new intermediate (named D) into MoO8.
    Three kinds of intermediates, A, B and C become amorphous upon grinding for a long time, but a crystal structure of intermediate D scarcely changes upon grinding for a week.
    It was supposed that three new compounds, A, B and C would have the same composition as m (NH, ), O. n MoO3, and that an intermediate D would have the same crystal form as hexagonal MoOs, and would occlude both water and ammonia in interstitial positions of their crystal lattices.
    Download PDF (325K)
  • Shoji SHIMIZU, Masao TANIGUCHI
    1977 Volume 1977 Issue 7 Pages 953-958
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The equilibrium water vapor pressures for the dehydration of CaC2C04420, ZnC2C04 H20, BaCl2 H20, BaCl24-120, FeCl2 H20, FeCl2 H2O, CuCl2 H20, CaSO4/2 H20 and CuSO4' H20 were measured in the temperature ranges from 306 K to 490 K by transpiration method. The transpiration technique was applied for both the directions of dehydration and hydration in order to determine the equilibrium pressure. The temperature dependence of the vapor pressure was given by the general equation: log P (atm)=A B x 103/T (K) (Table 2).
    The linear equations for standard free energies of each dehydration reaction were estimated by using the enthalpy and entropy data obtained from the vapor pressure equations (Table 2). The enthalpy of dehydration ranges from 13.0 to 17.9 kcal/mol, except the anomalous value (4H°=29.6 kcal/mol) for CuSO420. The average enthalpy of dehydration giving unhydrate products was 16.3 kcal/mol, and the average enthalpy of dehydration giving lower hydrate products was 14.1 kcal/mol (Table 3). The observed enthalpy value of dehydration was compared with the thermochemically calculated value. The thermochemical estimation based on the hypothetical dehydration process gave a value of approximately 5 kcal/mol, which corresponds to the bond energy between the unhydrate or almost unhydrate salt and crystallization water. Two kinds of average entropy 33 e. u. and 41 e. u. were obtained for the dehydration reactions.
    Download PDF (383K)
  • Tsuyoshi NAKAJIMA, Hidemasa NONAKA, Nobuatsu WATANABE
    1977 Volume 1977 Issue 7 Pages 959-962
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Aluminium was dispersed in molten salts as small particles by agitating the mixture of molten aluminium and NaCl-KCl or BaCli-K. Cl containing a small amount of alumina at 700-900° C. Dispersed aluminium particles were not so stable as those of magnesium in molten salts, somewhat coalescing with each other. The specific surface areas of dispersed aluminium particles gave a curve with a maximum point against the amount of alumina added into molten salts or the experimental temperature. Addition of sodium fluoride into molten salts resulted in the decrease of the specific surface areas of dispersed aluminium. From these facts, dispersed aluminium particles seemed to be stabilized with the solid film of aluminium suboxide which was formed by the reaction of metallic aluminium with alumina. ESCA spectrum of a solid sample supported this view.
    Download PDF (561K)
  • Kazuhiko SUGASAKA, Ayako FUJII
    1977 Volume 1977 Issue 7 Pages 963-968
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Crystals of different crystal habits were separated from the copper(I) iodide and sodium iodide solution and the thermal changes of the composition of copper(I) iodide and sodium iodide complexes were studied by chemical analysis, thermal analysis and X-ray diffractometry.
    Granular and columnar crystals were determined to be copper(I) iodide and sodium iodide dihydrate by X-ray diffraction analysis, respectively. Needle crystal ( A ) which was separated from the solution at 25° C was assumed to be Na2CuI3.6 H20. ( A ) was stable in its appearance in the air, but the X-ray diffraction pattern of ( A ) changed. Needle crystal ( B ) which was recrystallized at 10° C from mother liquor after the separation of crystal ( A ) was assumed to be NaCui2.4 H2O. (B) was hygroscopic and decomposed to precipitate copper(I) iodide with moisture in the air.
    ( A ) and (B) were found to change by heating and or drying, respectively, as follows:
    Download PDF (317K)
  • Yosiro YASUMOTO, Yukio ITO
    1977 Volume 1977 Issue 7 Pages 969-972
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The thermal decompositions of sodium imidobis (sulfate) (TSIS) and potassium imidobis (sulfate) (TPIS) over the temperature range 20-500° C have been investigated by using DTA and TG together with X-ray powder diffraction method and paper chromatography.
    When monohydrates of TSIS and TPIS were heated, they dehydrated at 74-404° C and 54-97° C, respectively and the anhydrous products were stable up to 290 and 300° C, again respectively. The same anhydrous products, however, decompose at 300-450° C and 340-450° C, respectively, giving nitrogen, gaseous sulfur, and solid sulfate:
    2 (MSO3)2NM N2T + Si + 3 M2SO4, where M=Na or K
    The reaction is exothermic, and TSIS melts during the course of the decomposition; whereas TPIS melts, at 313° C, just below the decomposition temperature.
    Download PDF (236K)
  • Yasuo HIKICHI, Ken-iti HUKUO
    1977 Volume 1977 Issue 7 Pages 973-977
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Solid state reactions between the synthetic monazite (CePOless than 1 Az in length) and some oxides were investigated by X-ray diffractometry. The oxides used were CaO, MgO, Al203, Cr2O3 and Si02. The mixtures of the synthetic monazite and oxide with the mole ratio(oxide/ monazite) 0.7-17 were pressed into pellets of about 10 mm/ in diameter under the pressure of 1000 kg/cm' for a few minutes. These pellets were fired in the temperature range from 600° C to 1900° C for 10 to 60 minutes in air. Monazite began to react with CaO or MgO around 700-800° C, and CeO2 and calcium phosphate (Ca3(P009)) or magnesium phosphate (Mg3(PO4)2) were formed. Monazite began to react with Al203 at 1820° C, and Ce02 was formed. Monazite did not react with Cr203 or Si02. A method of decomposition of natural monazite by firing with CaO seemed to be one of available methods for the extraction of thorium and rare earths from it. The conditions of decomposition of natural monazite with CaO were determined as follows: particle size of monazite less than 62 a, mixing weight ratio of CaO to monazite 0.6, the firing temperature 1300° C and the firing time 60 minutes. After the natural monazite was decomposed, CeO2, which contained thorium and rare earths as solid solutions, and calcium phosphate were formed.
    Download PDF (286K)
  • Kousaburo OHASHI, Toshinori SUZUKI, Kaoru KUBO, Katsumi YAMAMOTO
    1977 Volume 1977 Issue 7 Pages 978-982
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The kinetic studies on the complex formation between aquachromium (III) ion and ethylenediaminetetraacetic acid (EDTA) catalyzed by sodium nitrite or sodium sulfite were carried out at 25° C, 1=0.90 mol/l, and in the pH range from 3.50 to 4.80. The reaction product was ascertained spectrophotometrically to be Cr (III) -EDTA. When the concentration of EDTA or sodium nitrite (sodium sulfite) was in large excess over aquachromium(III) ion, the reaction to form Cr (III)-EDTA was of a first-order with respect to the aquachromium OE concentration and the rate increased with increasing concentration of EDTA or sodium nitrite (sodium sulfite); but in this case, the reaction was of no longer a first-order with respect to the concentrations of these substances. In the case of sodium nitrite catalyst, the rate of formation remarkably decreased with increasing pH. The rapid coordination of nitrite or sulfite ion to aquachromium(III) ion might accelerate the substitution of coordinated water for EDTA.
    Download PDF (312K)
  • Yukio SASAKI, Masao TAKIZAWA, Kisaburo UMEMOTO, Niro MATSUURA
    1977 Volume 1977 Issue 7 Pages 983-989
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The formation of halogenide and thiocyanate complexes of Cu (II) in propylene carbonate, N, N-dimethylformamide (DMF), and dimethyl sulfoxide (DMSO) was investigated and the overall formation constants (β3) were determined.
    There is a difference among the complex species formed by Cl-, Br- and I-, SCN- ions in propylene carbonate. As for I- and SCN- ions, the complexes of CuI, and Cu (SCN), are first formed and reduced rapidly to form precipitates of CuI and CuSCN, respectively. The precipitates dissolve completely to produce complexes of CuI - and Cu(SCN) 2- with increasing concentration of I- and SCN- ions, respectively. The formation of complexes of the type CuI42- and Cu(SCN)42- was also observed polarographically. As for I- and SCN- ions, both Cu(I) and Cu (II) complexes are formed in propylene carbonate. The fact that the complexes of the type CuI42- and Cu(SCN)42- are formed in propylene carbonate is very different from that in water. On the other hand, the formation of precipitates of Cu(I) complexes with Cl- and Br- ions in propylene carbonate was not observed. The complexes of the type CuCl42- and CuBr41- in the presence of excess Cl- and Br- ions are formed. A remarkable difference was also observed in the potentiometric titration curves between I- or SCN- and Cl- or Br- ions, respectively. In DMF and DMSO, the complexes of the type CuX42- and Cu(SCN)42- in the presence of excess X- and SCN- ions are formed, respectively.
    The overall formation constants log 80 for three solvents are show below:
    For propylene carbonate, For DMF, For DMSO, Cl- (42.5) >SCN- (36.7) >Br- (35.6) >1- (34.5) Cl- (32.5) >Br- (27.2) >SCIsT- (25.2) >1- (21.9) Cl- (23.1) >Br- (17.9) >SCINI- (17.6) >115. (15.4)
    On the basis of these facts, it was recongnized that the order of the overall formation constants between Br- and SCN- ions is reversed by a kind of a solvent. A similar tendency was also observed with regard to log β3.
    Download PDF (400K)
  • Masao MARUYAMA, Kazuo MURAKAMI
    1977 Volume 1977 Issue 7 Pages 990-997
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The electrochemical reduction of pyridazine and cinnoline in N, N-dimethylformamide (DMF), DMF-water containing 0-100 vol% water and Britton-Robinson buffer solution at a mercury electrode was investigated by voltammetry, controlled potential coulometry and spectroscopic methods.
    A series of experimental data indicates the following mechanisms for the electrode reaction of pyridazine and cinnoline in DMF-water and buffer solution.
    Dihydropyridazine and dihydrocinnoline formed by the controlled potential electrolysis of pyridazine and cinnoline (0.4 mmol/dms) in DMF containing 2% or 4% H20 at a mercury pool cathode were confirmed by gas chromatography-mass spectrometry. In the case of concen-trated solution (20 mmolidms) of cinnoline in DMF-H20 (above 80%) and buffer solution, however, precipitates were deposited by the controlled potential electrolysis with total charge passage of 0.8-1.2 electron per molecule. Mass spectra revealed that these precipitates were made up of polymerized cinnoline.
    Download PDF (496K)
  • Toyokichi KITAGAWA, Kazuyuki FUJIKAWA
    1977 Volume 1977 Issue 7 Pages 998-1002
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    2-Carboxy-1-pyrrolidinecarbodithioic acid (CPCD Hcpcd) reacts with Ni( II ) and Co (II) to form soluble complexes in aqueous solution. Both complexes show almost identical absorption spectra giving their maximum absorption at 324 nm. The molar absorption coefficients of Niand Co-cpcd complexes are 35200 and 25600 (mol//) -1cm-i, respectively. The compositions of these complexes are determined to be Ni: cpcd = 1: 2 and Co: cpcd= 1: 3 by the continuous variation method at 324 nm.
    The formation reactions of the complexes were studied spectrophotometrically by using a stopped-flow technique under the following conditions: excess CPCD, 0.01 moladm-g acetate buffer and A= 0.1 (NaCl04) at 25° C. The reaction rate for Co (II) is smaller than that for Ni (II), because Co (II) is oxidized to Co (III) by dissolved oxygen. The ratio of the observed first order rate constants, koNsiggs, is about 70 in air saturated solution at pH 5.49.
    Ni( III ) and Co (II) can be determined simultaneously by means of the following logarithmic extrapolation method. When the binary mixtures undergo pseudo-first-order reactions, the difference of the absorbances at time t=oo and t can be described as
    As=eNILINi(lE )10 exp(II) -Fec0LIC0(II) exP(-14: sit)
    where [Ni(II) ]o and [Co (II)] are the initial concentrations. Since the reaction rate of Ni (II) is far greater than that of Co (II), a plot of log (As, As) vs. t yields two straight lines. The extrapolated intercept at t=0 gives the value of log (e00L3[Co (II)]a) and the value of [Ni (II)], is obtained by subtracting the absorbance at the intercept from the Asa, . The calibration curve for each element was obtained with good linearity in the concentration range from 2x 10-6 to 8 x 10-6 molidm-3.
    Download PDF (289K)
  • Jun-ichi IWAMURA
    1977 Volume 1977 Issue 7 Pages 1003-1008
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Certain saturated secondary alcohols were found to be easily oxidized with 2, 3-dichloro-5, 6- dicyano-1, 4-benzoquinone (DDQ) to the corresponding ketones.
    Steric hindrance around the hydroxyl groups of the alcohols appears to enhance the rates of the oxidation significantly. Thus d1-8-norborneol [1 a] having a less hindered or unhindered hydroxyl group resisted the oxidation, whereas d1-a-norborneol [1 ID] in which the hydroxyl group is hindered was oxidized to the corresponding ketone in a satisfactory yield.
    The selative rates of the oxidation of, [1 a], [1 b], 1-isoborneol [3 a], d-borneol [3]- 13- f enchyl alcohol [2 a], d-a-fenchyl alcohol [2 b], 1-isoverbanol [4 ], 1-isopinocampheol [ 5 ], 1- neothujyl alcohol [6], cyclopentanol [7], cyclohexanol [8], monoalkylcyclohexanols ([ 9 ]-[17]), and dialkylcyclohexanols ([18]-[21]) were measured and the sesults discussed in terms of the steric hindrance around the hydroxyl groups.
    Download PDF (331K)
  • Jun-ichi IWAMURA, Nobuyuki IWAMOTO, Nenokichi HIRAO
    1977 Volume 1977 Issue 7 Pages 1009-1012
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of substituted benzaldehyde ethylene acetals with 2, 3-dichloro-5, 6-dicyano-1, 4- benzoquinone (DDQ) has been investigated. When the acetals were refluxed with DDQ in 1: 1 molar ratio in benzene for 3 hrs, the corresponding aldehydes, 2, 3-dichloro-5, 6-dicyanohydroquinone bis[2-(substituted benzoyloxy)ethyl] ethers (addition products) and 2, 3-dichloro-5, 6- dicyanohydroquinone (DDQH2) were isolated by column or gas chromatography (GC) from the reaction mixture. Identity of the known compounds was established by comparison of their rnp, tR in GC, Rf in thin-layer chromatography, MS and IR spectra with those of the authentic samples. The structure of the unknown compounds was determined on the basis of the spectral properties and chemical treatment. The reactivity of the acetals appears to depend on the inductive and steric effect of the substituents: p-anisaldehyde ethylene acetal reacted easily to give the addition product in a good yield.
    The formation of the corresponding aldehyde is reduced, when the methyl group is introduced at the 4- or 4, 5-position of the 1, 3-dioxolane moiety.
    Download PDF (192K)
  • Hideaki HAMADA, Takehiko MATSUZAKI, Katsuhiko WAKABAYASHI
    1977 Volume 1977 Issue 7 Pages 1013-1017
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The vapor phase ammonolysis of e-caprolactone to obtain e-caprolactam over copper-solid acid catalysts was investigated to elucidate the reaction mechanism. The reaction was carried out in a flow type under atmospheric pressure. The products obtained were e-caprolactam, 6-hydroxyhexanenitrile, 6-hydroxyhexanamide, adiponitrile and polymers; their yields were dependent on the catalysts and reaction conditions. With solid acid catalysts 6-hydroxyhexanenitrile was formed and copper scarcely showed a catalytic activity toward e-caprolactone. When the binary copper-solid acid catalysts were reduced before use and employed in the presence of hydrogen, e-caprolactam was mainly obtained; but otherwise, the main product was 6-hydroxyhexanenitrile. In the case of copper-titania catalyst, e-caprolactam was obtained in the maximum yield at 260-270°C. The catalytic activity decreased rapidly with time on stream, probably due to the covering of the catalyst surface with polymer. The addition of water to the feed increased the yield of e-caprolactam. The obtained results suggested the following reaction pathway: e-Caprolactone reacted with ammonia over solid acid to convert it into 6-hydroxyhexanamide, which was then dehydrated to form e-caprolactam over reduced copper. In the absence of copper, the intermediate was dehydrated to form 6-hydroxyhexanenitrile over solid acid.
    Download PDF (324K)
  • Jun-ichi IWAMURA, Minoru KAMEDA, Koichiro KOMAI, Nenokichi HIRAO
    1977 Volume 1977 Issue 7 Pages 1018-1020
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    An essential oil was obtained from tubers of Mizugayatsuri (Cyperus serotinus Rottb.) by steam distillation in O.02% yield, whose property being dr 1.067, 40 1.4949, and [ex]=14-0° (c= 0.9, CHCl3).
    The components were separated by means of column and gas chromatography, and as a main component, 8-elemene, caryophyllene, ce-humulene, 8-selinene, methyl trans, trans-farnesate, trans, trans-farnesyl acetate, or a-cyperone, and as a minor one, ct-copaene, 8-santalene, guaiene, 8-cadinene, calamene, or cyperotundone was identified in terms of IR, MS and NMR spectra.
    For comparison, the essential oil of Cyperus rotundus L. was analyzed, but methyl farnesate or farnesyl acetate could not be found.
    Download PDF (176K)
  • Hiroaki OMICHI, Hiroshi MACHIDA, Tetsuo MIYAKOSHI, Shojiro SAITO
    1977 Volume 1977 Issue 7 Pages 1021-1025
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The synthesis of dl-mevalonolactone from 4-chloro-3-methyl-1, 3-butanediol [1] or 4-chloro3-hydroxy-3-methylbutyl acetate [2] is described.
    By the reaction of [1] or [2] with KCN, the corresponding nitriles were obtained in 77 or 66% yield (Table 1, 3), respectively and upon alkaline hydrolysis of the above products, dlmevalonolactone [14] was obtained in 85 or 76% yield, respectively (Table 2, 4).
    Some by-products obtained by these reactions were also identified.
    Download PDF (314K)
  • Akio KARUBE
    1977 Volume 1977 Issue 7 Pages 1026-1028
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Hydrogenation of r-tetrahydrosantonin 2 and sodium santoninate (8) was studied. Upon hydrogenation of ( 2 ), a mixture was obtained which was separated into the known isomer [5] and a new isomer, (11 S)-3 6 ce-dihydroxy-4 aH-eudesman-12-oic acid 6, 12- lactone [4].
    Upon hydrogenation of [8] in the presence of excess sodium hydroxide, a new carboxylic acid [13] and the known hydroxy acid [9] were obtained. The yield of the former acid increased with increasing concentration of alkali. The structure of [13] was elucidated.
    Download PDF (196K)
  • Katsuhiro MIZOGUCHI, Tamotsu KAMIYA, Satoshi MATSUURA, Eishun TSUCHIDA ...
    1977 Volume 1977 Issue 7 Pages 1029-1034
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The electric properties of polymer composites with highly conductive 7, 7, 8, 8-tetracyanoquinodimethane (TCNQ) salts such as quinolinium-TCNQ complex salt (HQ, Qj), acridiniumTCNQ complex salt (HQ.CQ22) and N-methylacridinium-TCNQ complex salt (MA, Q2) have been studied. Polyacrylonitrile (PAN), Poly (N-vinylcarbazole) (PNVCz) and Poly (4-vinylpyridine) (P 4 VP) were chosen as matrix polymers.
    The specific resistance (p) of the PAN-HQ, Q2 system was O.37 S2cm at the HQ, Q2' content of 20 wt% in the film. When the content of HQ, Q2' was increased up to 40 wt%, a separation of the needle crystals was observed and the p value was increased. When HA, Q, ' was dispersed into PAN or PNVCz, the separation was also observed and the samples showed low conductivity. Uniform films with fine crystals were obtained in the PAN-MA, Q2, PNVCzMA, Q2r and P 4, VP-MA. CCW systems, and the values of p were O.37, 1.05 and 3.40cm, respectively. The properties of these composites were influenced by the combination of the TCNQ salts and the polymers. Uniformity of the composites was necessary to get high conductivity.
    Download PDF (439K)
  • Naoki NEGISHI, Masayuki TAKAHASHI, Akira IWASAWA, Kentarou MATSUYAMA, ...
    1977 Volume 1977 Issue 7 Pages 1035-1040
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    In order to elucidate the influence of hydrophobic interaction on photomechanical effect in water, copolymers, consisting of 2-hydroxyethyl methacrylate (HEMA) and N-vinylpyrrolidone (VPy), were prepared. As to the amphiphilic HEMA-VPy copolymer, which is water soluble non-polyelectrolyte, the hydrophobic interaction markedly increased with increasing HEMA composition. The study on the influence of hydrophobic interaction on the photoviscosity dehavior was made in the aqueous copolymer-dye solution.
    The copolymer-dye complex was formed upon addition of anionic azo dye to the aqueous coplymer solution, and the viscosity behavior of the complex was quite similar to that of polyelectrolytes. The copolymer chain expanded because of the electrostatic repulsion between the negative charges of the dye, and the extent of the copolymer chain remarkably increased with increasing HEMA composition in the copolymer. This showed that the hydrophobic interaction contributed to the complex formation.
    When the complex solution was exposed to and then screened from the light, alternately, the reversible conformational change in the complex was observed, being accompanied by isomerization of the dye. Such photoinduced conformational changes increased with increasing HEMA composition in the copolymer. These results suggested that photomechanical effect in water increased with increasing hydrophobic interaction of the polymer.
    Download PDF (345K)
  • Takashi OHTSUBO, Yasushi HAMANAKA, Yukihiro HOKAMURA
    1977 Volume 1977 Issue 7 Pages 1041-1045
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The depolymerization of polyhexanamide catalyzed by sodium hydroxide under steam blowing was studied.
    The depolymerization reaction of polyhexanamide under steam blowing catalyzed by sodium hydroxide was of the zero-order. The rate constant (k) for the formation of 6-hexanelactam consisted of the catalytic term k_0000 due to the acidic effect of carboxyl end groups of polyhexanamide and the non-catalytic term (k000, 0) due to the neutralization of carboxy 1 end groups with sodium hydroxide; k=k_000II +kONs% The rate constants for depolymerization at 310°C were calculated from the experimental results as follows:
    k-0oo11=0.36 x 10-2 [mmolfgrd-mol. hr]/[-NH2 meq/kg]
    k-000tin.=0.013 x 10-2 [mmol/grd-mol hr]/[-NIII meg/kg]
    The appatent activation energy of the depolymerization catalyzed by 2.00 wt % sodium hydroxide was 27.5 kcal/mol.
    Download PDF (320K)
  • Hachiro NAKANISHI, Masaki HASEGAWA, Hiroshi KIRIHARA, Tadashi YURUGI
    1977 Volume 1977 Issue 7 Pages 1046-1050
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Four-center type photopolymerization of various diolefinic compounds was investigated with particular reference to morphological change, under a polarizing microscope. Various changes have been observed: monomer crystals converted into aggregates of polymer crystallites keeping outer shape of the original crystal, certain monomers into fine fibrils, and some monomers into amorphous pseudomorphs. By closer examination of the present results, the polymerization behaviors, and the crystallographic data, reported previously, it was revealed that crack formation at the initial stage of the reaction which gave polymer crystals, highly depended on the cleavage-capability of a monomer crystal itself, and the shape change at the latter stage did on the magnitude of molecular movements during the reaction. The formation of pseudomorphs has been the characteristics of the reactions which occurred in an amorphous state, despite the reaction was topochemical or not.
    Download PDF (625K)
  • Yasuo KIKUCHI, Hiroshi FUKUDA
    1977 Volume 1977 Issue 7 Pages 1051-1054
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A water-insoluble precipitate was obtained by mixing sodium dextran sulfate (DS) solution and 2- (diethylamino)ethyldextran hydrochloride (EA) solution in various ranges of pH. The composition ratios of structural unit EA/DS in each polyelectrolyte complex (PEC), from A-1 to D-2, became 1.38-4.24. Thus, it was found that the mixing ratio EA/DS and pH values played an important role in determining the composition ratio in the PEC produced.
    The IR spectroscopic studies, the blood clotting test, the elemental analysis, the colorimetric reaction with toluidine blue, and the solubilities of PEC, revealed that the structures of PEC of A, B and C series (shown in Table 1), formed at higher pH, were different from those of D series, formed at lower pH, despite all of these PEC consisted of the same constituents. It seems that the intermolecular hydrogen bonds between OH groups would break at lower pH and form bonds between SO3H and OH groups.
    Download PDF (283K)
  • Takashi SAWAGUCHI, Takeshi KUROKI, Tatsuo ISONO, Nobuhiko IKEBAYASHI, ...
    1977 Volume 1977 Issue 7 Pages 1056-1062
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Behavior of products having high molecular weight formed upon thermal degradation of polystyrene have been investigated.
    As a result of the qualitative analysis of the above products, three kinds of dimer ([2] cis1, 3-dipheny1-2-butene (isomer), [3] 2, 4-diphenyl-1-butene, and [ 5 ] trans-1, 3-dipheny1-2- butene (isomer)) and three kinds of trimer ([6] isomer (M+=312), [ 7 ] 2, 4, 6-tripheny1-1- hexene, and [ 8 ] isomer (M+=312)) were identified.
    It was considered that the initial products, having high molecular weight, formed during degradation of polystyrene (see Table 1), resulted from the intramolecular transfer reaction (back-biting) of secondary radical, which formed upon random scission of main chain of polystyrene.
    In the presence of steam, results of pyrolyses of [3] and [7] are represented as follows:
    Dimer and trimer having terminal methylene groups isomerized (a, c) rather than degraded (b, d, e). As the back-biting and intermolecular transfer reaction of polymer radicals as well as the above isomerization were predominant at a low temperature, various kinds of isomer formed in the degraded polymer fraction (see Scheme 2).
    Under the same conditions, the ratio of isomer to dimer or trimer, was much greater in steam dilution than that in nitrogen dilution (see Fig.5). Water participated in isomerization and a mechanism of isomerization, via cyclic-transition state under which HO molecules were adsorbed to double bonds, was estimated (see Scheme 3).
    Download PDF (399K)
  • Hideyuki SUZUKI, Takao IWAI, Toshio HASEGAWA, Kunio OKUSHIMA, Tyuichi ...
    1977 Volume 1977 Issue 7 Pages 1063-1065
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Hydrous titanium (N) oxide were prepared by the homogeneous from urea and TiCl4 solution and by the conventional heterogeneous precipitation, and the change of surface area and crystal structure due to the heating were investigated by means of thermal analysis, X-ray diffractometry and surface area measurement.
    The product prepared by homogeneous process changed to anatase at 900°C and to rutile at 1000°C, while the product prepared by conventional precipitation changed to the mixture of anatase and rutile at 900°C and to rutile at 1000°C.
    The surface area of the former treated at 800°C for 5 hrs was 40 m2/g and that of the latter was 20 m2/g.
    After the heat treatment at 850°C for 240 hrs, the surface area of the former did not decrease, while that of the latter decreased remarkably. These results show that titanium (N) oxide prepared by homogeneous precipitation is thermally more stable than titanium (W) oxide prepared by heterogeneous precipitation.
    Download PDF (174K)
  • Arihiro HAMANO, Kazunari HAMADA, Kimio KUMAMOTO
    1977 Volume 1977 Issue 7 Pages 1066-1067
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Studies were made on the phase diagrams of NaCl0-NaCl-H20 system at 0, 5, 10°C. The mixture of H20, NaClo-. H20, and NaCl was stirred at least 2 hours in a thermostat; then a saturated solution was obtained. This mixture was settled for 3 hours in the thermostat, then the solution and wet crystals were analyzed respectively.
    The amount of Cl0- was determined by iodometry and the total amount of Cl was determined by Mohr's method after Cl0- was decomposed to Cl- by the addition of H2O2-. Then the amount of Cl- in the original saturated solution was calculated by the following equation: Total Cl- Cl0-.
    NaCl0-. H2O and NaCl were stable at these temperatures, and no solid solution between these compounds was found. It was found that the mole fraction of Cl0- in the eutectic composition increased as the temperature raised.
    Download PDF (89K)
  • Takaari YUMOTO
    1977 Volume 1977 Issue 7 Pages 1068-1070
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Di-t-butyl peroxide- and r-ray-induced substitution reactions of alkyl cellosolves RR/CH-0- CH, CH2OH (ethyl-, propyl-, isopropyl- and butyl cellosolve) [1] with tetrachloroethylene gave the corresponding substitution products CCl2=CClRWC-O-CH2CH2OH C2C. The yields of C2CD decreased in the order: ethyl>propyl>butyl>isopropyl.
    Download PDF (134K)
  • Takeshi MEGURO, Naochika TORIKAI, Atsushi MORI
    1977 Volume 1977 Issue 7 Pages 1071-1072
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    An adsorbent was prepared from Japanese brown coal (16-30 mesh) by treatment with an: aqueous NaOH solution. Although the above NaOH treated coal was a good cation exchange material, it had a defect of dissolving humic acid when immersed in water. In order to improve this defect, the NaOH treated coal was heated under nitrogen atmosphere at different tempera- tures between 100 and 600°C.
    No dissolution of humic acid was observed with the samples heat-treated at 100°C for 60 minutes.
    The detailed mechanism of these phenomena has not yet been elucidated.
    Download PDF (133K)
  • Moriyuki SATO, Sakae KON-YA, Masaaki YOKOYAMA
    1977 Volume 1977 Issue 7 Pages 1073-1074
    Published: July 10, 1977
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Polyglutamate having theophylline residue was prepared by the condensation of poly[y-(2- chloroethyl)-L-glutamate] with sodium salt of theophylline. The degree of substitution of polyglutamate with theophylline was up to 75%. The product was a white solid substance being soluble in such polar solvent as dimethyl sulfoxide or 1, 2-dichloroethane. The helix content of the product decreased with increasing degree of substitution with theophylline.
    Download PDF (124K)
feedback
Top