NIPPON KAGAKU KAISHI
Online ISSN : 2185-0925
Print ISSN : 0369-4577
Volume 1972, Issue 2
Displaying 1-50 of 50 articles from this issue
  • Masatoshi WATABE
    1972 Volume 1972 Issue 2 Pages 225-229
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The shift of the stretching vibrational frequencies of bonded hydroxyl group in 37 ring substituted phenols as acids and hexametylbenzene as a base has been studied. The concent-ration of the phenols in carbon tetrachloride solution is ca.0.005 mol/l and hexamethlbenzene 0.3 mol/l. In Fig.2 are shown the spectra of the solution of p-cresol (2-a), o-cresol (2-b) and 2, 6-xylenol (2-c) in the presence of hexamethylbenzene. Two bands due to the stretching vibration of bonded hydroxyl group are observed with a solution containing o-cresol, One of the two bands corresponds to that of p-cresol while the other to that of 2, 6-xylenol. Thus the OH group in o-cresol, even when hydrogen bond is formed, may be regarded as being coplanar with benzene ring the two bands seem to indicate the formation of cis and trans complexes. Although the presence of a good correlation between unassociated yon; and the Hammett constant σ has been pointed out, ΔvOH (=free v0H-bonded vOH) in the present study has a good correlation with the Hammett const σ, implying the occurence of the change of electronic state of hydroxyl group when hydrogen bond is formed. Plots ΔvOH against the Hammett const σ gave a slope of 55.0 cm-1 for ortho, and 46.2 cm-1 for meta and para substituted phenols.
    Download PDF (327K)
  • Akira KUBOYAMA, Masatoshi ANZE
    1972 Volume 1972 Issue 2 Pages 229-232
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Phosphorescence spectrum of phenanthrenequinone has been measured in various crystalline solutions at 77°K in connection with those in glassy ones at 77°K and n→π* absorption band at room temperature. All of the phosphorescence bands observed were assigned to π*→n12 bands on the basis of their short lifetimes (5-8 msec). Sharp phosphorescence bands were observed in n-paraffins, especially in n-hexane. Therefore, it was suggested, according to the rules in Shpol'skii effect, that phenanthrenequinone molecules are arranged in the n-paraffin crystals with their symmetry axis parallel to the longitudinal axis of the planar zig-zag form of n-paraffin molecules. Strong phosphorescence band at the shortest wavelength indicates that the geometry of the α-diketone group of phenanthrenequinone in the lowest (n, π*) triplet state closely resembles that in the ground state. That the relatively sharp phosphorescence bands with a relatively small blue-shift observed in acetic acid and in chloroform suggests that phenanthrenequinone molecules are regularly arranged in these solvent crystals. Phosphorescence spectrum in dioxane showed a large blue-shift (ca.1140 cm-1) relative to those in n-paraffins and the bands were very broad, owing to strong charge-transfer interactions between phenanthrenequinone and dioxane.
    Download PDF (290K)
  • Mitsunori TAKEHARA, Tsurutaro NAKAGAWA
    1972 Volume 1972 Issue 2 Pages 233-243
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Shear creep properties of oily gelatinous preparations were studied using a double-sandwich type apparatus. The specimens were the pasty stuffs obtained by cooling hot solutions of liquid petrolatum(LP) and solid stearic acid(SA).
    The instantaneous compliance (Jg)and the equilibrium compliance (Je) first decrease with the increase of temperature, and then after passing through a minimum, increase again.
    The compliances change in proportion to the reciprocal of 2.8 th power of the SA concentration, and this behavior was interpreted with van Dongen's model.
    Temperature dependence of the creep viscosity was examined and the activation energy of flow was calculated from the data as a function of SA concentration.
    The initial anomalous decrease of J, that is the apparent "hardening" of the substance, with the increase of temperature, was interpreted as an effect of voids contained in the system. ' The voids are gradually replaced by the thermally expanded LP. The scaffolding structure of SA crystals is impregnated with viscous and incompressible LP and hardened to gel.
    The proposed idea was tested by the examination of the thermal expansion behavior of LP, SA, gels, etc.
    Download PDF (399K)
  • Masaki DAIMON, Renichi KONDO
    1972 Volume 1972 Issue 2 Pages 238-243
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The vapor adsorption apparatus was assembled and the adsorption of methanol and cyclohexane on hardened 3CaO. SiO2 paste was measured to determine the pore size distribution.
    The pore size distribution was calculated from adsorption-desorption isotherm with cylinder and parallel plate models, using electronic computer. The weight change of a sample in the stream of gas-vapor mixture was determined, where the vapor pressure was varied by changing the ratio of dry gas and a gas saturated with the vapor of a desired solvent. This method was applicable to the measurement of the adsorption isotherms of water and organic compounds.
    Download PDF (373K)
  • Masayuki NAKAGAKI, Noriaki FUNASAKI, Kyoko FUJITA
    1972 Volume 1972 Issue 2 Pages 243-248
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Triglycerides of single and combinations of various saturated fatty acids(Cl2-C18)were synthesized, and the pressure-area curve, collapse pressure, collapse velocity and equilibrium spreading pressure of these triglycerides were measured.
    The equilibrium spreading pressure was the lower, but the collapse pressure was the higher, the greater the number of total carbon atoms in the fatty acids was, and the lower the temperature was. Trilaurin was different from the other triglycerides and the collapse pressure of trilaurin was equal to its equilibrium spreading pressure. The pressure of a trilaurin monolayer whose initial area was smaller than collapse area, decreased after the induction period passed away. The square root of initial surface concentration of collapsed trilaurin was directly proportional to the induction period.
    These results on collapse phenomena of a monolayer were explained on the basis of the following model of collapsing process: (1) molecules are forced out of a monolayer at collapse pressure, (2) the molecules are pushed up from the surface in a ridge two molecules thick, (3) this ridge finally breaks off and lies on the remaining monolayer as a collapsed fragment.
    It was suggested that intermolecular attraction force of mixed acid triglyceride was weaker than that of single acid triglyceride if the number of total carbon atoms in triglycerides was equal.
    Download PDF (391K)
  • Tomoyasu HASHINO, Takeshi KAWAI
    1972 Volume 1972 Issue 2 Pages 248-253
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A fundamental consideration of the iodide process of uranium in a closed system has been presented on the basis of the non-equilibrium thermodynamics.
    In order to simplify the treatment, the model of the discontinuous system consisting of two large reservoirs connected by a small transport path was employed. The condition of the mass transport between both sub-systems, i. e., a reaction zone and a deposition zone, in which several chemical reactions take place, was related to the thermo- dynamic forces of this system.
    It was shown that the rate of chemical transport of uranium is represented by the chemical potential difference of the iodide taking part in the transport, or by the chemical affinity of of the rate-determining step of the reaction.
    The present theory is applicable to the other iodide process such as niobium.
    Download PDF (325K)
  • Nagaaki TAKAMITSU, Toshikazu HAMAMOTO, Taichiro Nishi
    1972 Volume 1972 Issue 2 Pages 254-257
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    By the reaction of 2-tridecanol with metaboric acid in various mole ratios, three species of boric acid esters were prepared, and these esters were ascertained to be tri-2-tridecoxyboroxine [1], di-2-tridecoxyhydroxyborane(2), and tri-2-tridecylborate[3]. Thermal decomposition of these esters proceeded according to the first order law, and the initial decomposition rate constants were as follows:
    for [1], 4.82- 107 exp (-22.8 - 103/RT)min-1
    for [2], 9.27- 102exp (-11.0.103/RT)min-1
    for [3], 1.55 - 103 exp (-11.2.103/RT)min-1
    The yields of main products by the thermal decomposition of these esters were as follows: 96.2% tridecene from Li, 57.2% tridecene and 40.6% tridecanol from [2], and 49.2% tridecene and 47.2 tridecanol from [3]. It was found that the thermal behavios of tri-2-tridecoxyboroxine are different from those of the other esters.
    Download PDF (256K)
  • Ichiro OKURA, Tominaga KEII
    1972 Volume 1972 Issue 2 Pages 257-261
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The kinetics of ethylene dimerization with Rhodium complex catalysts has been carried out to clarify the reaction mechanism.
    In case of RhCl2-catalyst, the rate of butene formation can be represented by
    where Co is the concentration of the catalyst, P ethylene pressure and k, K are constant. It was found that ethylene retarded the butene isomerization reaction rate. Under such conditions that the butene isomerization is forbidden (under the ethylene high pressure), 1-butene, trans2-butene and cis-2-butene were formed concurrently. A reaction mechanism via r-ally1 intermediate was proposed, on the basis of the experimental results. In the case of [RhCl (C2H4)2]2-catalyst, 1-butene was the only initial product and cis- and trans-2-butene were produced at the equal rates by the isomerization of 1-butene. Since the rate of butene formation was first order with respect to ethylene pressure, the dimerization appears to occur via an alkyl intermediate which was formed between ethylene and the catalyst, and the formation of the intermediate seems rate-determining.
    Download PDF (302K)
  • Makoto EGASHIRA, Sunao URABE, Tetsuro SEIYAMA
    1972 Volume 1972 Issue 2 Pages 261-266
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Reduction of bismuth phosphate (Bi/P=2) and of stannic oxide containing small amounts of Na20 with propylene were studied at 500-800°C, in connection with their catalytic properties in the oxidative dehydroaromatization of lower olefins.
    X-ray diffraction analyses (Figs.2 and 6) showed that bismuth phosphate catalyst, which was composed of 3 Bi203.P205, 2 Bi203. P205, monazite form 0-BiPO4, high temperature form T-BiPO, and T-Bi203, and stannic oxide catalyst were reduced as follows:
    Main products from propylene during reduction were carbon dioxide and benzene in both cases, where the initial compositions were the same as those in the catalytic oxidation. These facts indicate that the oxidative dehydroaromatization proceeds by the redox mechanism of the catalyst (Bi3+ ⇔ Bi or Sn4+ ⇔ Sn). With the progress of reduction, the amount of CO2 formed decreased monotonously in every case. However, that of benzene formed showed different behaviors from CO2 as shown in Figs.1, 4 and 5. The selectivity of benzene formation decreased with the reaction time at lower temperatures where the reduction rate was relatively small, while the selectivity increased at higher temperatures. These complicated behaviors are presumably related to the concentration and the mobility of oxygen ion and/or allylic intermediate on the catalyst surface.
    Download PDF (431K)
  • Tadao ISHII, Mikio ARAMATA, Ryusaburo FURUICHI
    1972 Volume 1972 Issue 2 Pages 266-273
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The thermal behavior of the industrial vanadium catalysts for sulfur dioxide oxidation has been studied at working state by means of gas flow differential thermal analysis (gas flow DTA). The catalysts used are V205, Li2SO4-V205 (Li), Na2SO4-V205 (Na), K2SO4-V205 (K), KOH-V205 (K1), KHSO4-V205 (K2) and Cs2SO4-V205 (Cs). DTA apparatus permits gas to flow through a reference material and the sample. Four different atmospheres-air, N2, SO2 and SO2 (7%)+air are used. Heating rate is 5°C/min and gas flow rate is 50∼80 ml/min. The DTA curves of the fresh catalysts in the atmosphere of S02-air mixtures are shown in Fig.9. Exothermic peak at about 400°C corresponds to the activation process, that is, the formation of complexes containing vanadium, alkali and SO, . Another peak at about 500°C corresponds to the catalytic reaction of 2 S02+02 → 2SO3. Fig.10 shows of DTA curves of catalysts activated in the same condition as in Fig.9. Exothermic peak at about 400°C disappear. The order of the catalytic activities for sulfur dioxide oxidation, estimated from the temperature of DTA peak in Fig.10, is as follows; (Cs): 487°C (K): 509°C=z(K1): 503 °C (k2): 548°C, T (Na): 546°C (Li): 613°C. parenthesis designates catalyst and temperature means DTA peak temperature. On the other hand, the melting points of the activated catalysts complexes, were estimated from DTA curves in Fig.11, as follows; (Cs): 305°C, (K): 407°C, (K1): 420°C, (K2): 370°C, (Na): 410°C and (Li): 508°C, this temperature is closely related to the thermal behavior of the catalyst as shown in Fig.10. The diatomaceous earth used as a carrier has no effect on the DTA curves. In addition, the stability of catalytic activity is tested through the observation of the change of DTA curves with time at 600°C. The result obtained for KOH-V205 catalyst is shown in Fig.12.
    Download PDF (555K)
  • Yoshihisa MIZUKAMI, Akira YAMAMOTO, Haruo MATSUDA, Sumio MATSUDA
    1972 Volume 1972 Issue 2 Pages 273-278
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The effect of development conditions of Raney Cobalt (R-Co) or Raney Iron (R-Fe) catalyst on the alkylation and hydrogenation activities was investigated in the case of the synthesis of N-ethylpyrrolidone from succinimide, hydrogen and ethanol. The synthesis of N-ethylpyrrolidone from pyrrolidone and ethanol was also studied to find out the effect of metal oxides which existed originally in R-Co or R-Fe catalyst or formed during the reaction.
    R-Co catalyst, developed so as to contain more alumina showed high alkylation activity as well as considerably high hydrogenation activity. R-Co catalysts, developed in the usual way, contained high and low hydrogenation active sites. The R-Fe catalyst which showed high alkylation activity, showed low hydrogenation activity. Much amount of ion oxide was produced during reaction and R-Fe catalyst showed very low hydrogenation activity. The alkylation reaction was remarkably contributed by the metal oxides in R-Co or R-Fe catalyst.
    Download PDF (433K)
  • Akio TADA, Kiyoshi MIZUSHIMA
    1972 Volume 1972 Issue 2 Pages 278-283
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The acidity and catalytic activity of P205-B2O3(PB), prepared by the reaction of various amounts of phosphoric acid and boric acid, have been studied. In general PB gave solid acid catalyst, but its acid amount and catalytic activity depended on the ratio (R) of P205 to B2O3 in the catalyst.
    The acid amount increased monotonously as R increased from 1.0 to 1.6, while it increased and passed through a maximum at about R=0.66, after which, it decreased as R decreased from 1.0 to 0.23. The catalytic activity for the dehydration of n-butanol, which is presumably catalyzed by Bransted acid site (B-site) as well as Lewis acid site (L-site), was closely related to the total acid amount in the whole range of R, while for the cracking of cumene, catalyzed only by B-site, the relationship was found in the narrow range of R from 1.0 to 1.6.
    The ratios of 1-butene to 2-butene and cis-2-butene to trans-2-butene in the dehydration reaction products were higher than the equilibrium values for the PB with excess B203 and alumina, but they were nearly equal to the equilibrium values for the PB with excess P205 and solid phosphoric acid.
    Download PDF (423K)
  • Keishiro MIYAKE, Hideaki OKA, Hiroshi SAKAMOTO, Yoshio HARANO, Tatsuya ...
    1972 Volume 1972 Issue 2 Pages 284-289
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Ammoxidation of propylene on Fe2O3-Bi205 (44.5: 44.5: 11.0, mol%) catalysts was studied in a fixed-bed reactor, under the following conditions: reaction temperature, 410∼490°C; W/F, 0.3∼1.8 sec; particle size of catalyst, 40∼400 mesh and feed compositions; &as 0.023∼0.165, PO, 0.048∼0.303 and PNR3 0.047∼0.175 atm. Main products were acrylonitrile (AN), acetonitrile(AcN), CO, and CO. Acrolene(AL) and HCN were not detected.
    The rate equations of propylene conversion and of products formation, together with their apparent activation energies (C3H6, 21.7: AN, 23.6; AcN, 11.1: CO, 36.6: CO, , 16.0 kcal/mol), were obtained.
    From the effects of the feed composition on propylene conversion and products selectivities, it was concluded that two types of adsorption centers for propylene (I, II) exist on these catalysts and that the rates of propylene conversion on the type I and II site were proportional to (PO6/PC3H6)0 and (PO2/PC3H6)0, respectively. As the results, the following reaction scheme was suggested.
    The reaction seemes to proceed by either of the two reaction paths, depending on the oxidation state of the catalyst.
    Download PDF (369K)
  • Mamoru Ai, Sadao SUZUKI
    1972 Volume 1972 Issue 2 Pages 290-295
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The oxidation of cis-2-butene, butadiene, furan and maleic anhydride in the presence of excess air was carried out over MoO3-Bi203-P205 oxide catalysts having various Bi/Mo ratios and constant P/Mo ratio (0.2 atomic ratio), in order to elucidate the addition effect of Bi203 on the activity and selectivity of MoO3-Bi203-P205 catalysts. The activity for the oxidation of butene, butadiene and furan first rose with the increasing Bi203 content, and then decreased showing a maximum at Bi/Mo=0.10. However, the activity for the decomposition of maleic anhydride sharply increased with the Bi203 content. The results were explained on the basis of the acid-base nature of the reactant and the catalyst. Highest yielde of maleic anhydride from butene, butadiene and furan were obtained at Bi/Mo=0.1. They were 30, 60 and 76 mol% respectively. Above Bi/Mo=0.1, the yields remarkably decreased, because of the acceleration of secondary oxidation of maleic anhydride.
    Download PDF (393K)
  • Shin OHNO, Yukio UKITA
    1972 Volume 1972 Issue 2 Pages 295-301
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    When 1, 3, 3-trimethylindolinobenzopyrylspiran and its derivatives ([A]) in polystyrene matrix were bombarded with electron beam of medium energy (5-20 kV), two products were formed; a blue or magenta colored [B] and a yellow colored [C]. They returned reversibly to CAD by thermal decoloring. [B] was identical with photoproduct of the irradiation of ultraviolet light and decayed by the irradiation of visible light. [C] was formed only upon the bombardment. It was more stable than [B] and insensitive to visible light.
    Crosslinking of polystyrene, the binding matrix, took place by the bombardment, slightly disturbing the coloring and decoloring reactions because of viscosity change.
    The formation of [C] was promoted when the decoloring reaction of [B] was rapid. So it was suggeted that [C] was formed mainly by the bombardment of [B], although the possibility of direct formation of [C] from CAD was not neglected. Moreover, the activated intermediate of polystyrene on the crosslinking partiapated in the formation of [C].
    Download PDF (489K)
  • Masako SASAKI, Akio NOZAKI, Kenichi HONDA, Shin-ichi KIKUCHI
    1972 Volume 1972 Issue 2 Pages 302-309
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    An intermediate species of thephotoreduction of Cr(VI)to Cr(III)was detected with rigid matrix of ammonium dichromate solution containing reducing agents at -196°C.
    The intermediate was observed by spectrophotometry in the aqueous rigid matrix and the temperature dependence of the spectrum of the intermediate is discussed. The spectrum of blue intermediate(λ max 600mμ, e= 1000), the life time of which is more than 40 hrs at -196° C, decreased slowly with the increase of the temperature from -196° C and disappeared completly at about -140° C, near the glass transition temperature of water.
    It is confirmed that the intermediate is quite different from the final phatareductian product(Cr(III))and from the alcohol radical which could be produced by oxidation of alcohol during phatareduction of Cr(VI) to Cr(III).
    Download PDF (544K)
  • Daijiro YAMASHITA, Yoshihumi YAMAMOTO
    1972 Volume 1972 Issue 2 Pages 309-313
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The effect of the presence of antimony on the negative plate(Cd(OH)2-polyethylene)of pressed-type alkaline storage battery were studied through a capacity decrease of negative plate, X-ray diffraction of active material and potential sweep method.
    The results are summerized as follows:
    (1) Capacity decrease of negative plate on standing for long period in charged condition is prevented by adding antimony.
    (2) According to X-ray diffraction pattern of active material, discharge reaction Cd→ Cd(OH)2on standing for long period in charged condition is promoted a little, but grain growth of metallic cadmium is remarkably reduced by the presence of antimony.
    (3) According to potential sweep method, the greater part of antimony disolves into electrolytic solution on discharge and does not deposit on charge, but a small part of antimony deposits as metal with cadmium on charge. Therefore, it is considered that grain growth of cadmium crystal is prevented.
    (4) Reversibility of charge-discharge reaction is promoted by the addition of antimony.
    Download PDF (334K)
  • Takashi HAYASHI, Hajime SAITO
    1972 Volume 1972 Issue 2 Pages 314-320
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The reaction of phosphonitrilic chloride trimer(PNC)with N, N-dimethylformamide(DMF)has been studied.
    When the reaction was carried out at 60° C, solid(PNC2∼3DMF)and liquid(PNC4∼6DMF)materials were obtained. On the other hand, liquid material was only obtained at a temperature between 90° C and 120° C. These products were determined as adducts of PNC and DMF by IR spectra, 31P-NMR and elemental analysis.
    It was shown that the adducts are highly polar organic substances bymeasurement of the conductance.
    Download PDF (404K)
  • Yukio ITO
    1972 Volume 1972 Issue 2 Pages 320-328
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Reaction of thermal transformations of sulfuryl diamide at temperatures between 100 and 20° C was studied.
    The products obtained by heating sulfuryl diamide at temperatures 100, 120, 140, 160, 180, and 200° C, for 10 min, 30 min, 2 hr, 5 hr, and 8 hr respectively were analyzed by quantitative paper chromatagraphy.
    The components of the products detected by the chromatography were NH2SO2(NH4NSO2)3-NH2 and NH4SO3NH2 in addition to NH2SO2NH2, NH4N(SO2NH2)2, NH2SO2(NH4NSO2)3NH2, and(NH4NSO2)3which were reported previously.
    Results of analyses of the products indicate clearly the mode of successive transformations of sulfuryl diamide at temperatures between 100 and 200° C. Moreover, the diagram(Fig.8) of composition against reaction ratio(NH4/NH2SO2NH2)shows that sulfuryl diamide remains only10% of initial amount when the reaction ratio is 0.6, whereas the formation of trimeric ring(NH4NSO2)3 occurs at a reaction ratio from O.6 to O.97. On the other hand, the diagram also shows that selfuryl diamide changes mostly into linear oligomers at the reaction ratio below 0.5. These facts lead to the conclusion that the trimeric ring forms via linear oligomers, and the steps of thermal oligomerizatian of sulfuryl diamide are as follows:
    Download PDF (524K)
  • Yukio ITO
    1972 Volume 1972 Issue 2 Pages 329-334
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A procedure for quantitative separation of aquo ammono sulfuric acids by paper chromato-graphy was described. The paper chromatography was accomplished by ascending technique with a solvent mixture, dioxane-water-conc. aqueous ammonia(68.5: 31.4: 1), reported by Lehmann and Kempe. The chromatograms of aquo ammono sulfuric acid mixtures were developed and visualized with bromophenol blue spray reagent(yellow color of the components was appeared clearly in blue ground on standing overnight).
    For the quantitative determination, the yellow bands due to NH4-salts of aquo ammono sulfuric acid and selfaryl diamide zone in the chromatograms were cut off, and the componeat were extaracted with water. The sulfur in each separatedsubstance was conveted into SO42-/by the nitrous acid decomposition method and determined by turbidimetry.
    Pure aquo ammono sulfuric acids mixtures and reaction products containing aquo ammono sulfuric acids such as thermal oligomerization product of sulfaryl diamide, reaction product between SO2-Cl2gaseous mixture or SO2Cl2 and NH3, and reaction product between NH3and SO3 were analyzed by the present chromatographic procedure.
    The results demonstrate that the method gives reproducible results, and reliable for the quantitative analysis.
    A simultaneous extraction device(Fig.1)was also described.
    Download PDF (645K)
  • Takafumi KANAZAWA, Takao UMEGAKI
    1972 Volume 1972 Issue 2 Pages 335-338
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Dissolutions of calcium phosphates in acid solutions were studied by measuring heat of solution at 25° C. Fluorapatite(FAp), β -calcium orthophosphate(TCP), β -calcium pyrophosphate(CPP) and rhenanite(Rh), synthesized by dry methods, were introduced into hydrochloric acid and citric acid solutions.
    The amount of heat evolution was TCP> Rh> FAp> CPP. Concerning TCP, heats of solution both in O.05% acid and 2% citric acid, with pH which is 2.0, were about 20kcal/mol and those both in o.1% hydrochloric acid and 10% citric acid, with pH 1.7 were about 30kcal/mol. The heat output of TCP increased exponentiallywith a concentraton of hydrochloric acid. Those fads show that heat of solution depends upon the concentration of hydrogen ions and the dissolution of TCP proceeds by way of an exchange reaction between Ca2+ and H+. The formation of Ca-citrate complexes, however, was not revealed by the calorimetric measurement. Almost constant values of heats of solutions were obtained against different amounts of TCP dissolved in 2% citric acid, whereas in case of 10% citric acid heat of solution varied with an amount of TCP introduced. Consequently, solubilities of calcium phosphate are not always determined only by the calorimetry.
    Download PDF (322K)
  • Hideki MONMA, Takafumi KANAZAWA
    1972 Volume 1972 Issue 2 Pages 339-343
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Thermal formation of hydroxylapatite(HAp) from Ca3(PO4)2(TCP) and CaO inthepresence of water vapor was studied by X-ray diffractometry, IR absorption spectrometry and DTA. Conversion of TCP in the mixtures covering the mole ratios Cad/TCP in the range 1/3 to 5/3 was examined in the following conditions: 2.3∼23mmHg, 1000∼1300° C and O∼150min. Air stream containing water vapor with various vapor pressures was supplied to a mixture of CaCO3, which was used as a source of CaCO3, and TCF at the rate of 1OOml/min.
    The products were identined to be HAp from the analyses by X-ray diffraction and IR spectrometry and also from deficlency in the condition for oxyapatite formation in this system. The conversion isotherms obtained were found to obey the Shinriki-Kub's or Ginstling-Brounshtein's rate equation for solid-solid reactions. As shown in Fig.9, apparent activation energies obtained for the initial and intermediate stages of the reaction were 25 and 45 kcal/mal, respectively. The former value agreed closely with the value which was previously determined to be 27 kcal/mol for the initial stage of fluorapatite formation by a dry proeess. HAp was observed to form even in the air-atmosphere with 2.4mmHg water vapor pressure. Diffusion component in the above reaction was presumed to be CaO2 but not TPC.
    Download PDF (322K)
  • Hiroshi YAMAMURA, Ryoichi KIRIYAMA
    1972 Volume 1972 Issue 2 Pages 343-349
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    To clarify the roles of oxygen vacancies in the, perovskite-type, compounds of ABO8, the solid solution systems (Sr, M)FeO3-δ(M=Y3+, La3+, Bi3+ and In3+) are studied from the crystallo-chemical point of view by means of X-ray diffractometry, density measurements an chemical analysis.
    The results are summarised as follows: (1)In the (Sr, M)FeO3-δ prepared by firing at elevated temperatures in air, the lattice constants increase linearly with the elevation of M3+ content as a result of decreasing Fe4+/Fe3+ ratio to the respective critical content of M3+, then the lattice constants decrease as a result of an additional contribution of decreasing oxygen vacancies as the M3+ content further increase. When the specimens are fired in fairly high temperatures, however, A-site-vacancies in addition to the oxygen vacancies often appear and a substitutional exchange between A-site and B-site ions also occurs in a small extent. From the fact that the all tolerance factors of (Sr, M)FeO3-δ at each maximum solubility of Ms3+ are nearly equal to 1.01, we may conclude that the stability limit of the present solid solutions to be controlled by the geometrical requirement of the constituent ions. (2)In the (Sr, M)FeO3-δ prepared by firing in vacuum, brownmillerite-type phase changes to cubic perovskite-type with the elevation of M3+ content owing to increasing incorporation of oxygen. A further incorporation of oxygen into the cubic phase results in the orthorhombic symmetry. The presence of oxygen vacancies in the solid salutians seems to play an important role for the stabilization of perayskite-type structure with cubic symmetry.
    Download PDF (505K)
  • Makoto NOSHIRO, Yukio JITSUGIRI
    1972 Volume 1972 Issue 2 Pages 350-352
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    A rapid analysis for the direct determination of fluorine in glass using the Corning fluoride specific electrode without a prior separation of interfering ions is given. Among many buffer solutions, a 1M sodium citrate buffer solution was recommended for the analysis of fluorine in glass samples. Strong complexes of interfering ions such as Si6+, B3+, Al3+ with fluoride could be destroyed by an addition of the buffer solution to release the bound fluoride.
    Procedure: About 0.05∼0.4g of ground glass samples were accurately weighed into a platinum crucible. The sample was mixed with a 1.0g of sodium carbonate and fused. Distilled water was added and warmed on a steam bath to dissolve. The pH of the solution was adjusted to 6.0 using HCl solution and it was made up to 100ml. To ensure that the ionic strength and pH of the solution remained constant, 10ml of sample solution were mixed with 10ml of the buffer solution. The electrode were placed into the solution and the fluoride ion activity was measured. The concentration of fluorine ion was determined from the calibration curve obtained previously and the percentage of fluorine in the glass sample was calculated.
    The results of the proposed method were compared with the results obtained by the pyrohydrolysis-Th(NO3)4 titration method; these two results were in good agreement.
    The standard deviation of determination was about 0.009% for the fluorine 1.0% level.
    The time required for analysisis 20min.
    Download PDF (202K)
  • Eizen ISHII, Masato MAMIYA, Tetsuro MURAKAMI
    1972 Volume 1972 Issue 2 Pages 353-358
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The diffuse reflection spectra in the infrared region of the several inorganic salts were taken by the use of trial diffuse reflction equipment, which was attached to an infrared spectrophotometer, and the possibilities of this instrument for the quantitative analysis have been investigated.
    The Kubelka-Munk function can be written in the form,
    f(R∞)=(1-R∞)2/2R∞=c/k
    where k' is a constant equal to S/2.303ε(S is the scattering coefficient and ε is the molar absorptivity) and C is the molar concentration. The Kubelka-Munk function is analogous to the Beer-Lambert law of absorption spectrophotometry except b effect, (b is an average path length of beam in powdered medium). ln regard to b effect, the greater the absorption of the sample, the more intensive backscattering and incident beam reflects in the vicinity of the sample surface.However, the smaller the sample absorption, the greater the scattering and the beam penetrates into the inside of the sample reflecting by multiple scattering. Consequently, the path length of light gets averaged and b effect has an characteristic that absorption intensity is determined through an averagrd spectra over whole region in comparison with the tablet method that has constant light path length.
    It was concluded that the reflective absorbance was approximately linear to the logarithm of concentration of the sample in the working range. This relationship was not affected by specular reflection and the most of all vibration bands were obtained by b effect. A definite reflectivity was obtained for a same lot of matrix, regardless the condition of sample surfaces.
    Download PDF (434K)
  • Teruzo ASAHARA, Manabu SENO, Yasuyuki SHIMOZATO, Yutaka NISHI
    1972 Volume 1972 Issue 2 Pages 359-361
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    3, 4-Epoxy-3-methyl-1-butene was obtained from the reaction of isoprene with perbenzoic acid (PBA).
    The rate constants of this epoxidation were measured in various solvents at -18°C, -14°C -10°C and -6°C, and the several kinetic parameters were calculated (Table 1). The solvent order with decreasing rate constants was as follows; chloroform>1, 2-dichloroethane>chloro-benzene>toluene>carbon tetrachloride. This is the same order as that of epoxidation of butadiene.
    A close correlation was found between the rates of epoxidation and the polarographic half-wave potentials of PBA in some mixed solvent system (Table 2). Further, the rate of hydrogen-deuterium interchange of PBA in various solvents was measured by NMR, and gives greater values in oxygen-contaning solvents (Figure 2). A good correlation was observed between the rate of epoxidation and the basicity of solvents; generally, the increas-ing basicity of solvents results in the decreasing rate of epoxidation reaction.
    Download PDF (223K)
  • Nobuaki FUJIWARA, Mitsuyoshi NAKAO, Futoru YOSHIMURA
    1972 Volume 1972 Issue 2 Pages 362-368
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    9, 10-Dihydroanthracene[DHA] gave anthraqttinone and at least six intermediates, anthracene, 9-anthrol (or anthrone), anthrahydroquinone, 9-acetoxyanthracene, 10-acetoxy-9-anthrol (or 10-acetoxyanthrone) and anthraquinyl diacetate, by oxidation with Na2Cr2o72H2O [Cr (VT)] in AcOH containing various amounts of AcONa (Table 1) at 40∼60°.
    The reaction rate was proportional both to the concentrations of Cr(VI) and DHA (equation 1), and was decreased with increasing amount of AcONa in AcOH (Table 6).δH was also decreased with increasing amount of AcONa.
    A linear relationship was found between δMΔH, δMΔS, and the isokinetic temperature was calculated as 299°K.
    The oxidation process wad discussed.
    Download PDF (422K)
  • Takaari YUMOTO, Tatsuo MATSUDA
    1972 Volume 1972 Issue 2 Pages 369-375
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Methyl, ethyl, propyl and butyl alcohols were found to add to tetrachloroethylene by the irradiation with X-ray.
    Saturated and unsaturated 1: 1-adducts, whose general formulas were CHCl2-CCl2-C(R1)(R2)-OH and CC2=CCl-(R1)(R2)-OH, were obtained.
    Effects of molar ratios of reactants, dose rate, total dose and temperature on this reaction were studied. The optimum conditions were found to be: (ROH): (CCl2=CCl2)=4:1, dose rate about 4 x 105 r/hr, and total dose 4.3∼7.0 Mr.
    Almost all alcohols were found to be consumed at rates nearly proportional to dose rate, but in the case of ethyl alcohol, the rate was found to be nearly proportional to the square root of dose rate.
    Yields of adduct decreased in the order: ethyl>n-propyl>isopropyl>sec-butyl>methyl>n-butyl>isobutyl.
    As the total dose increases, the yield becomes higher but the ratio of the yields of unsaturated to saturated adducts becomes large.
    Download PDF (476K)
  • Shizunobu HASHIMOTO, Wataro KOIKE, Masaki OYAMA
    1972 Volume 1972 Issue 2 Pages 375-379
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Aryl 2-thenoates were obtained in high yields by the reaction of substituted benzenes with 2-thenoyl peroxide in the presence of cupric chloride at 80°C. The ratios of isomers and the relative rates for aromatic substrates indicated that the 2-thenoyloxylation was brought about by 2-thenoyloxy radicals possessing electrophilic properties. A plot of the partial rate factors (f) of meta and para positions vs.σ+ substituent constants satisfied a liner relation (p= -1.96). The reaction mechanisms were also discussed.
    Download PDF (354K)
  • Masahiko SAITO, Masataka DEHARA, Osamu MANABE
    1972 Volume 1972 Issue 2 Pages 380-382
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It has been found that reaction of aniline with sulfamic acid gave ammonium orthanilate through the. reaction intermediates ammonium N-phenylsulfamate.
    When a mixture of 1 mole of sulfamic acid and 4 moles of aniline was stirred at 160°C for 12 hr, orthanilic acid (0.42 mol) sulfanilic acid (0.35 mol) and a small amount of aniline disulfonic acid was obtained. The separation of orthanilic acid and sulfanilic acid was per- formed successfully by means of the difference of solubilities in aqueous pottassium hydroxide.
    The total yields of orthanilic acid and sulfanilic acid were 35.5% and 24.4% respectively based on sulfamic acid. This method can be conveniently used for the preparation of orthanilic arid.
    Download PDF (224K)
  • Kenzo HOSOKAWA, Kan INUKAI
    1972 Volume 1972 Issue 2 Pages 383-386
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    1-Chloro-2-trifluoromethylnaphthalene [1] was prepared by chlorination of 1-hydroxy-2- naphthoic acid with phosphorous pentachloride followed by fluorination with antimony trifluoride.
    Reactions of [1] with various nucleophiles (sodium methoxide, sodium phenoxide, sodium thiophenoxide, cuprous cyanide, butylamine, benzylamine, potassium fluoride and cesium fluoride) were tried.
    In the case of the reactions such as methoxide, thiophenoxide, cyanide and fluoride, aprotic polar solvents are effective for the preparation of 1-methoxy-, 1-phenyl-thio, 1-cyano- and 1-fluoro-2-trifluoromethylnaphthalene.
    When [1] was treated with amines in DMF, 2-trifluoromethylnaphthalene [2] and 1-dimethylamino-2-trifluoromethylnaphthalene were obtained. The compound [2] was also obtained by the treatment of [1] with a mixture of cuprous oxide and propionic acid in pyridine.
    Download PDF (313K)
  • Takashi KEUMI, Setsuo YAMADA, Hideaki NODA, Yoshibumi OSHIMA
    1972 Volume 1972 Issue 2 Pages 387-391
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The Friedel-Crafts benzoylation of substituted diphenylene oxides (X-D) has been investigated to elucidate the site of benzoylation in X-D.
    Reaction of benzoyl chloride (BzCl) with X-D in the presence of aluminum chloride in nitrobenzene (molar ratio, X-D: BzCl: AlCl3, 1: 1.5: 1.5) at 50∼60°C for 10 hours yielded the following compounds. (Substituents X in the reactants: products, mp °C, yield %) 2-NO2: 8-Bz-2-NO2-D, 206∼207°, 83.8, 2-NHCOCH3: 8-Bz-2-NHCOCH3-D, 212∼213°, 94.6, 2-Br: 8-Bz-2-Br-D, 159∼160°, 84.5, 2-Cl: 8-Bz-2-Cl-D, 152∼153°, 84.3, 3-NO2: 8-Bz-3-NO2-D, 208.5∼209°, 69.2, 3-NHCOCH3: 8-Bz-3-NHCOCH8-D, 188.5∼189.5°, 67.0 and 7-Bz-3-NHCOCH3- D, 203∼205°, 0.6, 3-Br: 8-Bz-3-Br-D, 147∼148.5°, 56.5 and 7-Bz-3-Br-D, 173∼174°, 31.1, 3-Cl: 8-Bz-3-Cl-D, 138.5∼139.5°, 60.5 and 7-Bz-3-Cl-D, 160∼151°, 2.9.
    Download PDF (341K)
  • Shizunobu HASHIMOTO, Isao FURUKAWA, Shouzo FUJIMOTO
    1972 Volume 1972 Issue 2 Pages 391-395
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The rates of deoxygenation by triethyl phosphite have been measured for α, N-diphenylnitrone and its mono-substituted derivatives. The kinetic expression of the reaction is of first order with both triethyl phosphite and nitrones respectively. The rate was enhanced by electronwithdrawing substituents, and retarded by electron-donating substituents on either phenyl group of the nitrones. A good Hammett correlation with σp substituent constants was obtained.
    (ρx= +1.18, r = 0.996; ρy= +1.47, r = 0.987)
    Introduction of methyl group at ortho-position of N-phenyl group on the nitrone slowed down the reaction rate because of the steric hindrance effect. All these data seem to suggest that the deoxygenation reaction proceeds through nucleophilic attack by triethyl phosphite on the nitrogen atom of nitrone.
    Download PDF (295K)
  • Akira ARASE, Yuzuru MASUDA, Masaaki ITOH, Mitsuomi ITOH, Akira SUZUKI
    1972 Volume 1972 Issue 2 Pages 395-401
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Thermal treatment of trialkylboranes was studied in the presence of polar compounds, including DMSO, acetophenone, DMF, benzaldehyde and furfural. Remarkable elimination, of an alkyl group as olefin was observed at temperatures higher than 120°C, and it was shown that in DMSO, acetophenone and DMF one of the three alkyl groups of a trialkyl- borane was eliminated, while in benzaldehyde and furfural two of the alkyl groups were eliminated. In all cases a secondary alkyl groups was eliminated preferentially from mixed primary-secondary alkylboranes. Further, neither isomerization in alkyl groups which remained on boron atom nor exchange of alkyl groups between trialkylborane molecules was observed. Only internal olefins were formed from secondary alkyl groups. These selective and stepwise decompositions were found to be useful for analysis of a composition of the isomers in trialkylboranes.
    Download PDF (436K)
  • Shoichi KIKKAWA, Takatoshi HAYASHI, Kazuhide FUJITA
    1972 Volume 1972 Issue 2 Pages 402-407
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Phenothiazine used as high-temperature antioxidant is regarded as a donor from the viewpoint of molecular complex. An interesting synergism for the inhibition of autoxidation was revealed in many combinations of the donor and suitable acceptors.
    The synergism was also recognized evidently in several other cases in which the donor is phenothiazine, one of its homologous, diphenylamine, 2, 6-di-tert-butyl-p-cresol, or tetrabase, and the acceptor is, for example, chloranil, 1, 3, 5-trinitrobenzene or 2, 4, 7-trinitrofluorenone.
    The synergism showed continuous synergistic index change according to the mole fraction change of donor to acceptor and did not take a stoichiometric maximum value.
    These results suggest that this pronounced synergistic effect may depend upon the formation of π-complex having high inhibition effects.
    Download PDF (382K)
  • Hajimu NISHIOKA, Shozo WADA
    1972 Volume 1972 Issue 2 Pages 408-412
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The liquid phase hydrogenation of 2-ethylhexanal was carried out with the nickel-kieselguhr catalyst under the high pressure of hydrogen to identify the by-products, and to investigate the effects of the reaction conditions and some additives on the side-reaction.
    It was confirmed that 2-ethylhexanal (di-2-ethylhexyl) acetal was produced during 2-ethylhexanal hydrogenation and the yield of the acetal increased during the hydrogenation up to a maximum value at a point of 60∼70% conversion of aldehyde, and then decreased.
    The yield of bis-(2-ethylhexyl)ether increased with increase in the reaction temperature, that of the acetal decreasing. This reason was discussed in the paper. This side-reaction was promoted by acids contained in 2-ethylhexanal, but was supressed by water.
    Download PDF (310K)
  • Nobuyoshi SASAKI, Tetsuo MORI, Yukio YAMAGUCHI, Tetsuo YOKOYAMA, Takeh ...
    1972 Volume 1972 Issue 2 Pages 413-422
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    It was found that urethane prepolymers, that is, diisocyanate adducts of polyether glycol, in DMF solution were rapidly crosslinked by the catalytic action of sodium cyanide. By evapolating to dryness, three-dimensional polymers were obtained. Anionic catalysts such as nbutyl lithium, sodium benzophenone ketyl and sodium benzoate were also found to be effective for the crosslinking of prepolymers. The crosslinking reactions of prepolymers with sodium cyanide in DMA solution were studid by determining the gel point. Gelation time decreased with increasing catalyst concentration and prepolymer concentration. The apparent activation energy of gelation was ca.18 kcal for polypropylene glycol-tolylene diisocyanate at 0∼25°C and ca.35 kcal for polypropylene glycol-hexametylene diisocyanate at 55∼80°C. The crosslinking reaction of prepolymers, prepared from polypropylene glycol and tolylene diisocyanate, in DMA solution with sodium cyanide catalyst was followed at 25°C by the titration of unreacted isocyanate groups, and was found to obey apparent third-order. The infrared spectrum of the resulting polymer shows an absorption at 1410 cm-1 corresponding to isocyanurate ring. From these results, it was considered that the crosslinking is effected by trimerization of isocyanate groups. The resulting network polymers were rubbery at room temperature and gel fractions of these polymers were mostly near 1. Densities, swelling properties, glass transition temperature, Young's modulus, tensile strengths and elongations were also measured. Polyurethanes prepared from hexametylene diiscyanate have good mechanical properties (tensile strength 130∼320 kg/cm2 and elongation 450∼1000%). DTA and TGA measurement showed these polymers were stable up to 280∼320°C.
    Download PDF (654K)
  • Katsutoshi OHKUBO, Hiroyuki SAKAMOTO, Keisuke MORITA
    1972 Volume 1972 Issue 2 Pages 423-426
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The nucleophile-initiated polymerization of α, α-disubstituted β-lactones (R, R'-BPL) was studied with the extended Hückel molecular orbital theory.
    First, the results of the calculations well explained the experimental findings concerning the reaction course between acid or base catalyst and R, R'-BPL. Secondly, relatively small overlap population of the β-C-O bond in the lactone ring, 0.43∼0.47, suggests weakness of this bond. There exists a parallelism between the overlap population of the above bond in R, R'- BPL and the initial rate constant (k1) of the reaction of R, R'-BPL initiated by quaternary ammonium carboxylates. Finally, we discussed the reactivity of R, R'-BPL toward strong base catalysts.
    Download PDF (256K)
  • Ryota FUJIO, Minoru KOJIMA, Shiro ANZAI, Akira ONISHI
    1972 Volume 1972 Issue 2 Pages 426-433
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Copolymerizations of butadiene and styrene were investigated with n-buthllithium (BuLi)polar organic compound mixture and BuLi-alkaline metal complexes which were used as catalyst at 40°C or 50°C in n-hexane. The weight ratio of butadiene to styrene was 75: 25 and that of the monomer to solvent was 1: 4 in all cases.
    The polar organic compounds used were classified as follows: i) ether compounds, ii) acetallike compounds, iii) tertiary amine compouns, iv) thioether compounds and v) hetero ring compounds. When the styrene content of copolymers obtained by initiating with BuLi and polar organic compounds of group i), ii) or iii) and by terminating at 20∼60% conversion, was correlated with 1, 2-addition content of butadiene units, a fairly good linear relationship was observed among 24 compounds. While, an addition of any compound similar to thioether or hetero ring compounds to BuLi exhibited little or no effect on both the polymerization rate and the course of copolymerization.
    It was found that the complexes obtained by the reaction of BuLi with alkaline metals(except Na) in aromatic solvent gave random copolymers of butadiene and styrene. Copolymerizations were carried out by using the following combined catalysts with various mole ratio to BuLi: (a) BuLi-K complex, (b) BuLi-tetrahydrofuran (or diethylether) and (c) BuLi-o-dimethoxybenzene. A good linear relationship was also observed between 1, 2-addition content and styrene content of copolymers obtained, whose slopes were in the order: (a)>>(b) (cf. Fig.2).
    The styrene sequence distribution for a random copolymer obtained by (a) was discussed on the basis of NMR and far-infrared spectroscopic measurements. The molecular weight distribution of this copolymer was measured by gel-permeation chromatography.
    Download PDF (528K)
  • Ryota FUJIO, Minoru KOJIMA, Shiro ANZAI, Akira ONISHI
    1972 Volume 1972 Issue 2 Pages 434-439
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The effects of some methyl substituted pyridine compounds (including prydine, quinoline and β-collidine) on n-butyllithium (BuLi)-initiated copolymerization of butadiene and styrene were examined at 40°C or 50°C in n-hexane (solvent). The weight ratio of butadiene to styrene was 75: 25 and that of the monomer to solvent was 1: 4 in all cases.
    Addition of a small amount of pyridine (or quinoline) retarded the rate of polymerization but did not change the styrene inclusion profile. When more than equimolar pyridine (or quinoline) was added to BuLi, polymerization did not occur. Addition of 2, 6-lutidine or 3, 4- lutidine was effective in increasing both the rate and the styrene content. Addition of 3, 4- lutidine with same 1, 2-addition content gave far more styrene content than that of 2, 6-lutidine when conversion was less than 60%. A detailed comparison of rates of polymerization with copolymer structures obtained was carried out when 2, 4, 6-collidine, β-collidine and triethylamine (TEA) were used as additives to BuLi. The rate of polymerization and the 1, 2-addition content of copolymer decreased in the following order: 2, 4, 6-collidine>8-collidine>TEA, while the order of effectiveness for inclusion of styrene in the copolymer chain was 8-collidine > 2, 4, 6-collidine > TEA.
    From the oxidative degradation it was found that addition of 2, 4, 6-collidine, β-collidine or 3, 4-lutidine was effective to produce a random butadiene styrene copolymer. Grass transition temperatures (Tg) of these copolymers were also examined.
    Butadiene-styrene copolymerizations were carried out bs7 using fifteen methyl substituted pyridines as additives to BuLi under the same conditions. It was found the polymerization rate showed as a whole negative correlation with pKb of each compound and the plots of the 1, 2-addition content vs. the styrene content of copolymer (10∼50% conversion) split into three lines because of a steric effect of the additives (cf. Fig.6).
    Download PDF (361K)
  • Ryota FUJIO, Minoru KOJIMA, Shiro ANZAI, Akira ONISHI
    1972 Volume 1972 Issue 2 Pages 440-446
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Homopolymerization of butadiene and isoprene and copolymerization of butadiene with styrene were investigated in hydrocarbon solvents by using three organoalkaline-earth metal compounds, i.e., calciumzinc-tetraethyl (Ca⋅ZE), strontiumzinc-tetraethyl (Sr⋅ZE) and barium-zinc-tetrabutyl (Ba⋅ZB) as catalysts. The weight ratio of monomer to solvent was 1 : 4 in all cases.
    In carrying out the experiments in toluene, polybutadienes were formed containing 75∼65% of trans 1, 4-addition, 15∼25% of cis 1, 4-addition and 10% of 1, 2-addition. In isoprene polymers there were 80-70% of 1, 4-addition and 20∼30% of 3, 4-addition but without 1, 2-addition.
    The copolymerization of butadiene with styrene were examined by using Ca⋅ZE. Sr⋅ZE, Ba⋅ZB, dibutylmagnesium, dibutylzinc n-butyllithium and phenylsodium as catalysts (cf. Table 3).
    When mixtures of 55 wt.% of butadiene and 45 wt.% of styrene were copolymerized with Ca⋅ZE, Sr⋅ZE and Ba⋅ZB, the initial copolymers contained 9 wt.%, 14 wt.% and 32 wt.% styrene, respectively. These results were discussed in relation to the electronegativity of each ion. Among cyclohexane, n-hexane and toluene, there was found little difference in the course of copolymerization of butadiene and styrene with Ca⋅ZE, Sr⋅ZE, or Ba⋅ZE, but the polymerization rate was the fastest in toluene.
    The influence of reaction conditions, such as catalyst concentration, polymerization tempe-rature and addition of tetrahydrofuran, on the butadiene-styrene copolymerization by using Ba⋅ZB as a catalyst, was examined in detail. Oxidative degradation and glass transition temperature showed that random butadiene-styrene copolymers were produced by Ba⋅ZB.
    Download PDF (437K)
  • Ryota FUJIO, Minoru KOJIMA, Shiro ANZAI, Akira ONISHI
    1972 Volume 1972 Issue 2 Pages 447-453
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Copolymerizations of butadiene and styrene were investigated by using n-butyllithium (BuLi) and alkaline-earth metal compound as a catalyst. The weight ratio of butadiene to styrene was 75: 25 and that of the monomer to solvent was 1: 4 in all cases.
    Calcium compound retarded the rate of BuLi-initiated copolymerization but did scarcely change the microstructure of the copolymer. The addition of strontium t-buthoxide (B⋅Sr) to BuLi somewhat increased styrene inclusion and the trans-1, 4-addition content. The addition of barium compound, especially barium t-buthoxide (B⋅Ba), was the most efficient in accelerating the styrene incorporation and in increasing the trans-1, 4-addition content, whereby there was a negligible change in the 1, 2-addition content. For example, the microstructure of the copolymer obtained with BuLi-B. Ba mixture (BuLi: B⋅Ba=5: 1∼2: 1) contained about 65% trans-1, 4-addition, about 25% cis-1, 4-addition and about 10% 1, 2-addition. The polymerization rate was always markedly increased by elevating polymerization temperature. As the results shown above, the features of butadiene-styrene copolymerization by using BuLi-B⋅Ba mixture [1] as catalyst agreed well with those by bariumzinctetrabutyl as a catalyst. Accordingly, the presence of common catalyst species such as dibutyl-barium was suggestible.
    Time vs. conversion, conversion vs. styrene content and conversion vs. inherent viscosity were plotted for copolymerizations in n-hexane, benzene and toluene with [1]. Oxidative degradation and glass transition temperature showed that random copolymers, were produced in these cases.
    The styrene sequence distribution for a random copolymer obtained with [1] was discussed on the basis of NMR and far-infrared spectroscopic measurements, where it seemed that the copolymer with [1] contained more microblocks of styrene than a random copolymer obtained with a BuLi-tetrahydrofuran mixture [2]. Molecular weight distribution by gel permeation chromatography of the copolymer with [1] was far more broad than that of the copolymer with [2].
    Download PDF (448K)
  • Minoru HASHIMOTO, Toshio KUNUGI, Nobusuke OTAGIRI, Koji AMEMIYA
    1972 Volume 1972 Issue 2 Pages 454-458
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    On the un-treated, pre-treated and after-treated specimens of 6-nylon fibers, the orientation factors of Crystallites (fc) and amorphous chains (fa), and the orientation functions of (200) and (002) planes have been measured. The changes of orientation with drawing and various treatments were investigated by comparing the results with other factors of fine texture. The following results were obtained: -
    (1) f and fa of the drawn specimens were increased by after-treatment under tensionless state. This phenomenon found in 6-nylon, which differs from other linear polymers seems to be due to the presence of hydrogen bonds.
    (2) In the region of low draw ratio, the factors (fc) of pre-treated specimens were smaller than those of un-treated specimens, but the other factors were larger. The main factors affecting it are the growth of crystallites and the elimination of strains, resulted from pretreatment before drawing. In the region of high draw ratio, it became to be difficult gradually to increase fe and fa, owing to the formation of the secondary network structure with the destruction of crystallites.
    (3) The selective orientation of two crystal-planes was slight in the un-treated or aftertreated specimens, but was remarkable in the pre-treated specimens. Such planar orientations are brought about by cleavage and gliding of hydrogen bonded planes in the crystals, and those degrees are influenced by the direction of folding of molecular chains in lamellar crystals.
    Download PDF (382K)
  • Yoshitaka OGIWARA, Hitoshi KUBOTA, Atushi SAKAMOTO, Toshisada KURUSU, ...
    1972 Volume 1972 Issue 2 Pages 459-464
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Fine fibers (80 mesh over and 300 mesh-screening residue, S4) were isolated from highly beaten pulp (NBKP, 92° SR) and their adsorbing characteristics were compared with those of beaten and unscreened sample (S2) and long fibers (80 mesh-screening residue, S3). The metallic ion-adsorbing characteristics of fine fibers were clearly distinguished among those of samples. Namely, the beating degree became lower by the addition of ferric ion to each sample, and above all, the rate of decrease of the beating degree was remarkable for S2 and S4. When ferric ion was compared with aluminium ion, it was observed that the rate of decrease of the beating degree was larger for ferric ion, and aluminium ion affected the decrease of the beating degree in the lower concentration range than ferric ion. When ferric ion was used as a metallic ion, the amount of adsorbed ferric ion was influenced by the liquor ratio, concentration and time. Especially, it was found that the largest difference of the amount of adsorption among samples was attained by selecting the liquor ratio properly. Among the samples, S2 showed the largest amount of adsorption. It was also observed that the activity of S4 in the adsorbing reaction was extremely high as compared to that of S3. As for the interactions of fine fibers with fillers and size, S4 was also observed to have remarkable characteristics for fixing. The contents of a few functional groups of the samples were determined, neither of which was found to be related with their adsorbing characteristics closely.
    Download PDF (384K)
  • Jiro OSUGI, Hiroshi NAKATANI
    1972 Volume 1972 Issue 2 Pages 465-467
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The rate constants of complexation reactions of copper(II) with asparagine and glutamine were determined by the temperature jump method. The reactions were of the type,
    At 25°C and ionic strength 0.1, the rate constants are: asparagine, k1 =4.0 x 109 M-1 sec-1, k-1=50 sec-1, k2=2.5 x 108 x M-1 sec-1, k-2=71 sec-1; glutamine, k1 =3.3 x 109M-1 sec-1, k-1=59 sec-1, k2=1.4 x 108M-1 sec-1, k2= 44 sec-1.
    The rate constant of formation reaction of CuL+ is smaller in glutamine than in asparagine, but the effect of carbon chains of these amino acid homologues is small in comparison with that of glycine homologues.
    Download PDF (181K)
  • Tadaaki TANI
    1972 Volume 1972 Issue 2 Pages 467-469
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    The degrees of the increases of the surface speeds due to the presence of phenosafranine and to sulfur sensitization have been measured in the AgBrI emulsions prepared by various procedures, and compared with the ratios of internal/surface speeds of the corresponding undyed and unsensitized (control) emulsions. It has been found that there are strong correlation between them both in the case of the addition of phenosafranine to the emulsion and in the case of sulfur sensitization. It has been suggested that the sensitivity centers in sulfur sensitiza- tion as well as phenosafranine molecules increase the surface speed of the emulsions by acting as electron traps, and that the degree of the increase of surface speed increases with increasing the ratio of internal/surface speed of the control emulsion.
    Download PDF (194K)
  • Shun ARAKI, Shigetaka SUZUKI, Masaru KITANO
    1972 Volume 1972 Issue 2 Pages 469-471
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Some phenones and benzoates were irradiated with ultra violet rays in the vapor phase (2∼6 mmHg) at cell temperature 100°C to identify alkyl and phenyl groups adjacent to the carbonyl and carboxyl groups.
    Super high pressure mercury arc was used as a light source. Benzaldehyde and phenones were irradiated with ultra violet rays transmitting through Toshiba UVD-25 filter (2500∼4000 Å), and photolysis of benzoates were carried out with the whole radiation of the arc.
    Small amounts of H2S were useful to inhibit secondary reactions of alkyl radicals except hydrogen abstraction reaction.
    The photolysis products produced through the reactions (1)∼(9) were analyzed by gas chromatograph connected to the reaction system. And they showed a good correlation with the original alkyl and phenyl groups adjacent to the carbonyl and carboxyl groups.
    Download PDF (171K)
  • Kozo KUROKAWA, Teruo KONDO
    1972 Volume 1972 Issue 2 Pages 472-474
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Gel permeation chromatography (GPC) was used to separates the components of Gach Saran residue and. its hydrodesulfurized residue. GPC separation was made with a cross-linked polystyrene gel, using benzene as a solvent. The fractions obtained were separated further by silicagel adsorption chromatography into four generic components. The method involves solvent deasphaltening for recovery of asphaltenes, followed by elution chromatography to yield saturates, naphthene-aromatics, aromatics and polar-aromatics. Gel permeation chromato- graphy results offer challenging opportunities to gain an insight into the chemical mechanisms resulting in hydrodesulfurization.
    Download PDF (183K)
  • Tsutomu INA, Toshihide KUROSAWA, Takeshi KOMATSU, Teruyoshi NAKAMURA, ...
    1972 Volume 1972 Issue 2 Pages 475-477
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    Ethyl acrylate (EA) and 2-chloroethyl acrylate (CEA) have been polymerized at 50°C in benzene using azobisisobutyronitrile as an initiator. The rate of polymerization (Rp: mol/l. sec) for both monomers has been expressed as follows.
    Rp= k[AIBN] 1/2[M]3/2
    k=7.54 x 10-4 (EA), k=1.49 x 10-3 (CEA),
    where [AIBN] and [M] are the concentrations of initiator and monomer, respectively. The kinetic constant k1/kp2 for EA has been determined at 50, 60 and 70°C. The activation energys (Ep-Et/2) for EA and CEA were caluculated as 2.8 and 2.6 kcal/mol, respectively.
    Download PDF (174K)
  • Sohei SUGA, Tadashi NAKAJIMA, Takashi ISHIHARA
    1972 Volume 1972 Issue 2 Pages 477-479
    Published: February 10, 1972
    Released on J-STAGE: May 30, 2011
    JOURNAL FREE ACCESS
    he UV-light or Free-radical Induced Polymerization of Propylene Sulfide (PS) in the Presence of maleic anhydride (MAH) has been studied. It was found that polythioether was the only product. No adduct of MAH to polythioether was formed. The polymer (F=12∼17) did not contain mercapto groups. The polymerization reaction seems to proceed cationically by the formation of a charge transfer complex of PS and MAH, as in the case of propylene oxide.
    Download PDF (180K)
feedback
Top