共查询到20条相似文献,搜索用时 15 毫秒
1.
The formation of the spinel CuAl2O4 from the oxides CuO and Al2O3 has been studied at 1000 and 950°C in air by measuring the fraction reaction completed as a function of time. In the experiments the molar oxide ratios were CuO/Al2O3 = 0.5, 1.0 and 2.0 and the grain size for CuO was 1–3 μ throughout while for Al2O3 fractions of 40–60 μ, 71–100 μ and 100–125 μ, were used. The rate of reaction could be explained quite well assuming a three-dimensional diffusion mechanism. 相似文献
2.
KNO2 III below ?13°C is monoclinic, space group P21 or P21/m, with a0, b0, c0 = 4.677, 9.650, 6.395 Å, β = 93.8° at ?35°C. There is a further phase transformation between ?35°C and ?100°C to a new phase KNO2 VII, which is also monoclinic, space group P21 or P21/m: with a0, b0, c0 = 8.397, 4.773, 7.644 Å, β = 112° at ?100°C. Both these phases appear to be ordered. 相似文献
3.
The partial enthalpies of liquid lead(II) oxide and of tetragonal germanium dioxide in lead germanate melts were measured calorimetrically at 900°C in the range 0 to 65 mole-% GeO2. The corresponding integral enthalpies of mixing were calculated. A relatively sharp dependence of the partial enthalpies on composition in the range from N = 0.3 to 0.45 probably is due to the formation of Ge2O. 相似文献
4.
Raj T. Natarajan Theodore Davidson 《Journal of Polymer Science.Polymer Physics》1972,10(11):2209-2222
The rate at which PTFE transforms from the triclinic to the hexagonal phase has been measured under isothermal conditions. Samples in a series of decreasing density and crystallinity show increased isothermal rates of transformation. Observed kinetic data are interpreted on the basis of a modified Avrami-Johnson-Mehl treatment. A model for the transformation in which planes of helix-hand reversal propagate through the lattice is shown to fit the experimental results. The transformation rate is observed to be proportional to the total (001) surface area in the polytetrafluoroethylene specimens, suggesting that nucleation of the transformation takes place at grain boundaries. 相似文献
5.
6.
A kinetic and mechanistic study of the autoxidation of liquid pentaerythrityl tetraheptanoate (PETH) at 180–220°C has been carried out utilizing a stirred-flow reactor. The results are consistent with the occurrence of a chain reaction scheme similar to that proposed for n-hexadecane autoxidation, namely, the formation of monohydroperoxides by the intermolecular abstraction reaction (3), the formation of α,γ- and α,δ-dihydroperoxides and α,γ- and α,δ-hydroperoxyketones by intramolecular peroxy radical abstraction reactions (4) and (4*), the bimolecular termination of peroxy radicals, reaction (6), and the rapid conversion of α,γ-hydroperoxyketones to the corresponding cleavage acids and methyl ketones, reaction (7). Comparisons of various rate parameters for the n-hexadecane and PETH systems reveal that the values of k7 and (k3/H atom)/(2 k6)1/2 are within experimental uncertainties identical for the two systems at 180°C. The proposed reaction scheme includes the concurrent formation of hydroxy radicals and hydroperoxyketone species. The results of kinetic analysis and the experimentally observed isomer distributions of primary and secondary monohydroperoxide products at high and low oxygen pressures suggest that ≈60% of the hydrogen abstractions from PETH at high oxygen pressures occur by hydroxy radicals. 相似文献
7.
Li‐Peng Hou Xue‐Qiang Zhang Bo‐Quan Li Qiang Zhang 《Angewandte Chemie (International ed. in English)》2020,59(35):15109-15113
Stable operation at elevated temperature is necessary for lithium metal anode. However, Li metal anode generally has poor performance and safety concerns at high temperature (>55 °C) owing to the thermal instability of the electrolyte and solid electrolyte interphase in a routine liquid electrolyte. Herein a Li metal anode working at an elevated temperature (90 °C) is demonstrated in a thermotolerant electrolyte. In a Li|LiFePO4 battery working at 90 °C, the anode undergoes 100 cycles compared with 10 cycles in a practical carbonate electrolyte. During the formation of the solid electrolyte interphase, independent and incomplete decomposition of Li salts and solvents aggravate. Some unstable intermediates emerge at 90 °C, degenerating the uniformity of Li deposition. This work not only demonstrates a working Li metal anode at 90 °C, but also provides fundamental understanding of solid electrolyte interphase and Li deposition at elevated temperature for rechargeable batteries. 相似文献
8.
Experimental results are reported for the solid-liquid phase equilibria of the ternary system KCl? MgCl2? H2O at 130, 140, and 150°C. An “analytic” method, described previously, was used. 相似文献
9.
D. M. De Leeuw C. A. H. A. Mutsaers G. P. J. Geelen C. Langereis 《Journal of solid state chemistry》1989,80(2)
The subsolidus phase diagram of the system La2O3---SrO---CuO at 950°C under 1 bar of pure oxygen has been investigated and a new ternary compound, La1+xSr2−xCu2O5.5+δ with 0.05 ≤ x ≤ 0.15, was isolated. This compound crystallizes in an orthorhombic unit cell with lattice constants related to the lattice constant of the perovskite cubic unit cell, ap, by a = 3.80 Å ap, B = 11.48 Å 3ap, and c = 20.23 Å 5ap. The structure is isotypical to that of LnSr2Cu2O5.5+δ with Ln = Sm, Eu, or Gd. Reported data on the crystal chemistry of the equilibrium compounds in the system La2O3---SrO---CuO have been summarized and compared with the present data. The structure of all compounds is built up of a La---O rock-salt layer separated by a number of LaCuO3 perovskite layers. The general formula is (La---O)(LaCuO3)n where La can be replaced either partly or completely by Sr. Compounds are found for n = 1, 2, and ∞. The structures of the compounds show different types of oxygen vacancy ordering. 相似文献
10.
Noboru Kimizuka Francisco Brown Maria J. R. Flores Masaki Nakamura Yuichi Michiue Takahiko Mohri 《Journal of solid state chemistry》2000,150(2):276
The phase relations in the system In2O3–TiO2–MgO at 1100 and 1350°C are determined by a classical quenching method. In this system, there are four pseudobinary compounds, In2TiO5, MgTi2O5 (pseudobrookite type), MgTiO3 (ilmenite type), and Mg2TiO4 (spinel type) at 1100°C. At 1350°C, in addition to these compounds there exist a spinel-type solid solution Mg2−xIn2xTi1−xO4 (0≤x≤1) and a compound In6Ti6MgO22 with lattice constants a=5.9236(7) Å, b=3.3862(4) Å, c=6.3609(7) Å, β=108.15(1)°, and q=0.369, which is isostructural with the monoclinic In3Ti2FeO10 in the system In2O3–TiO2–MgO. The relation between the lattice constants of the spinel phase and the composition nearly satisfies Vegard's law. In6Ti6MgO22 extends a solid solution range to In20Ti17Mg3O67 with lattice constants of a=5.9230(5) Å, b=3.3823(3) Å, c=6.3698(6) Å, β=108.10(5)°, and q=0.360. The distributions of constituent cations in the solid solutions are discussed in terms of their ionic radius and site preference effect. 相似文献
11.
12.
Relative rate constants for the reaction of OH radicals with a series of n-alkanes have been determined at 299 ± 2 K, using methyl nitrite photolysis in air as a source of OH radicals. Using a rate constant for the reaction of OH radicals with n-butane of 2.58 × 10?12 cm3 molecule?1s?1, the rate constants obtained are (X1012 cm3 molecule?1 s?1): propane 1.22 ± 0.05, n-pentane 4.13 ± 0.08, n-heptane 7.30 ± 0.17, n-octane 9.01 ± 0.19, n-nonane 10.7 ± 0.4, and n-decane 11.4 ± 0.6. The data for propane, n-pentane, and n-octane are in good agreement with literature values, while those for n-heptane, n-nonane, and n-decane are reported for the first time. These data show that the rate constant per secondary C—H bond is ∽40% higher for —CH2— groups bonded to two other —CH2— groups than for those bonded to a —CH2— group and a —CH3 group. 相似文献
13.
Phase relations at 1500°C in the ternary system ZrO2–Gd2O3–TiO2 have been determined by the powder X-ray diffraction of samples prepared by standard solid state reaction. A large area of this ternary oxide system centered on the Gd2Ti2O7–Gd2Zr2O7 join was shown to exhibit the pyrochlore and defect fluorite structures. The pyrochlore structure was observed for stoichiometries as far from the ideal M4O7 as M4O6.7 and M4O7.4, although the degree of disorder seemed much higher at these stoichiometries. On further deviation from the ideal M4O7 stoichiometry a smooth transition to fluorite average structure was observed for Zr-rich compositions. None of the other binary phases were observed to show significant extent of solid solution into the ternary region. 相似文献
14.
Ajit Kumar Chaudhuri Santi R. Palit 《Journal of polymer science. Part A, Polymer chemistry》1968,6(8):2187-2196
The mode of termination in the vinyl acetate–isobutyl methacrylate (VA–IBMA) and methyl methacrylate–methyl acrylate (MMA–MA) copolymerization systems has been investigated at 60°C. by using the dye-interaction technique for functional endgroup estimation. The results show that pairs of poly(vinyl acetate) radicals interact almost exclusively through a disproportionation mechanism. In the homopolymerization of methyl methacrylate and methyl acrylate, about 1.16 and 1.22 carboxyl-containing endgroups per polymer molecule have been estimated, which shows the predominance of disproportionation over combination in these termination reactions. In poly(isobutyl methacrylate) about 1.55 tagged initiator fragments per chain indicate that 29% of the total radicals terminate through the disproportionation mechanism. Cross termination in the (VA–IBMA) copolymerization system occurs almost entirely through combination for monomer feeds richer in isobutyl methacrylate content while for the MMA–MA system, combination is more important at intermediate monomer feed ratios. These results have been discussed in the light of different explanations for the reaction mechanism. 相似文献
15.
Studies on the equilibrium in the system MCl2? CH3OH? H2O at 25°C and 50°C (M = Sr2+, Ba2+) show that the dehydration in the water-methanol systems proceeds stepwise and all possible lower crystal hydrates may be obtained at 25°C depending on the molar ratio for the mixed solvent. The dehydration and solvation processes in the three-component system MCl2? CH3OH? H2O (M = Mg2+, Ca2+, Sr2+, Ba2+) have been considered in general and compared with those in the bromide system. 相似文献
16.
Temperature effects on γ-initiated polymerization of styrene. III. Conversion in the range 150–200°C
M. Taherzadeh J. W. Hodgins 《Journal of polymer science. Part A, Polymer chemistry》1975,13(2):425-435
Experimental data are presented for the polymerization of commercial styrene in a γ-radiation flux of 20–51 rad/sec in the temperature range of 150–200°C. At radiation intensities above 20 rad/second, conversion rate is independent of dose rate over the range. Above 165°C, radiation does not enhance the conversion rate but does produce more rapid elimination of residual monomer. Molecular weights of polymer product in this temperature range are too low to be of commercial interest. An optimal temperature range of 110–120°C is suggested for the process. 相似文献
17.
Victor E. Meyer 《Journal of polymer science. Part A, Polymer chemistry》1967,5(6):1289-1296
From the conversion–composition data of Gruber and Elias, the reactivity ratios of styrene (M1) and methyl methacrylate (M2) were calculated to be r1 = 0.55 ± 0.02 and r2 = 0.58 ± 0.06 at 90°C. The least-squares method was then used on these and literature values at other temperatures to obtain the Arrhenius expressions: In r1 = 0.04736 – (235.45/T), and ln r2 = 0.1183 – (285.36/T). Using literature values for the homopolymerization steps, A11 = 2.2 × 107l./mole-sec., E11 = 7.8 kcal./mole, and A22 = 0.51 × 107 l./mole-sec.?1, E22 = 6.3 kcal./mole, activation energies and frequency factors were then calculated for the cross-polymerization steps: A12 = 2.1 × 107 l./mole-sec., E12 = 7.3 kcal./mole, and A21 = 0.45 × 107 l./mole-sec., E21 = 5.7 kcal./mole. 相似文献
18.
Relative rate constants for the gas-phase reactions of OH radicals with a series of alkyl nitrates have been determined at 299 ± 2 K, using methyl nitrite photolysis in air as a source of OH radicals. Using a rate constant for the reaction of OH radicals with cyclohexane of 7.57 × 10?12 cm3/molec·s, the rate constants obtained are (× 1012 cm3/molec·s): 2-propyl nitrate, 0.18 ± 0.05; 1-butyl nitrate, 1.42 ± 0.11; 2-butyl nitrate, 0.69 ± 0.10; 2-pentyl nitrate, 1.87 ± 0.12; 3-pentyl nitrate, 1.13 ± 0.20; 2-hexyl nitrate, 3.19 ± 0.16; 3-hexyl nitrate, 2.72 ± 0.22; 3-heptyl nitrate, 3.72 ± 0.43; and 3-octyl nitrate, 3.91 ± 0.80. These rate constants, which are the first reported for the alkyl nitrates, are significantly lower than those for the parent alkanes, and a formula, based on the numbers of the various types of C? H bonds in the alkyl nitrates, is derived for rate constant estimation purposes. 相似文献
19.
Michael J. Monteiro Tamaz Guliashvili Virgil Percec 《Journal of polymer science. Part A, Polymer chemistry》2007,45(10):1835-1847
A mechanistic comparison of the ATRP and SET‐LRP is presented. Subsequently, simulation of kinetic experiments demonstrated that, in the heterolytic outer‐sphere single‐electron transfer process responsible for the SET‐LRP, the activation of the initiator and of the propagating dormant species is faster than of the homolytic inner‐sphere electron‐transfer process responsible for ATRP. In addition, simulation experiments suggested that in both polymerizations the rate of deactivation is similar. In SET‐LRP, the Cu(II)X2/L deactivator is created by the disproportionation of Cu(I)X/L inactive species, while in ATRP its concentration is mediated by the bimolecular termination. The combination of higher rate of activation with the creation of deactivator via disproportionation provides, via SET‐LRP, an ultrafast synthesis of polymers with very narrow molecular weight distribution at room temperature. SET‐LRP is mediated by a catalytic amount of Cu(0), and under suitable conditions, bimolecular termination is virtually absent. Kinetic and simulation experiments have also demonstrated that the amount of water available in commercial solvents and monomers is sufficient to induce the disproportionation of Cu(I)X/L into Cu(0) and Cu(II)X2/L and, subsequently, to change the polymerization mechanism from ATRP to SET‐LRP. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 1835–1847, 2007. 相似文献
20.
The thermal reaction of hydrogen–butene-2-cis mixtures has been studied in a static system at low extent of reaction around 500°C. Hydrogen does not affect the thermal reaction itself of the olefin, but gives rise to new stoichiometries of hydrogenolysis and hydrogenation, which are specified: The reaction is described in terms of a molecular and free-radical mechanism. It is shown that the key process for the hydrogenolysis–hydrogenation reaction is (1) and that the rate constant of this process can be determined from either propylene, or methane, or butene-1 formations: with θ = 4.57 × 10?3 T kcal/mol. Other rate constants are estimated and agree with literature data. 相似文献