首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The kinetics, mechanism, and activation energy of the isothermal decomposition of CuCrO4 was studied using an isothermal TG method and an X-ray high-temperature diffraction technique in either air or a flowing atmosphere of N2. The enthalpy change ΔH of the decomposition reaction
2CuCrO4CuO+CuO+CuCr2O4+32O2
was determined by DSC analysis. The mechanism of the thermal decomposition of CuCrO4 is well represented by the standard Avrami-Erofeev kinetic equation [?ln(1 ? α)]12 = kt. According to this mechanism, the reaction rate is controlled by the formation and growth of nuclei on the surface of the reactant. The activation energy EA of the process in air is EA = (248 ± 8) kJ mole?1, in flowing atmosphere of nitrogen EA = (229 ± 8) kJ mole?1. ΔH in air is 110 kJ mole?1, in flowing nitrogen 67 kJ mole?1. The lower values of ΔH and EA in the flowing atmosphere of nitrogen are due to the fast elimination of O2 from the reaction interface. However, the decay of the crystalline portion of CuCrO4 during its thermal decomposition, studied by the X-ray diffraction, is controlled by a different reaction mechanism (first-order kinetics). The reaction mechanism is discussed in the relation to the crystal structure of the reactants.  相似文献   

2.
The tracer diffusion coefficient, D1O, of oxide ions in LaCoO3 single crystal was determined over the temperature range of 700–1000°C by a gas-solid isotopic exchange technique using 18O tracer. For the determination, two methods, the gas phase analysis and the depth profile measurement, were employed. Under an oxygen pressure of 34 Torr, the temperature dependence of D1O in LaCoO3 was expressed by
D1O(cm2·sec?1) = 3.63 × 104exp? (74 ± 5)kcal · mole?1RT
D1O at 950°C was found to be proportional to P?0.35O2. The diffusion of oxide ions occurs through a vacancy mechanism. The activation energy for the migration of oxide ion vacancies was estimated as 18 kcal · mole?1.  相似文献   

3.
The relationship between the structure of monomer and kinetics of the radical polymerization of N-ethylmethacrylamide, N-butylmethacrylamide and N-phenylmethacrylamide in methanol and in dimethylsulphoxide was investigated. The reaction order with respect to initiator is 0·5 in all cases; the order with respect to monomer is independent of the type of substituent but depends on the solvent and on the viscosity of the reaction mixture. The polymerization rate, the value of κpt12, and the initiator efficiency decrease in the series N-phenylmethacrylamide, N-ethylmethacrylamide and N-butylmethacrylamide. The overall activation energy of polymerization for the monomers lies between 16 and 20 kcal/mol.  相似文献   

4.
The tracer diffusion coefficient, D1O, of oxide ions in LaFeO3 single crystal was determined over the temperature range of 900–1100°C by the gas-solid isotopic exchange technique using 18O as a tracer. For the determination of D1O, the depth profile of 18O was measured by means of a secondary ion mass spectrometer (SIMS). The surface exchange reaction was found to be slow and the surface exchange rate constant, k, was determined together with D1O. It was found that D1O at 950°C is proportional to P?0.58O2, where PO2 is an oxygen pressure. The vacancy mechanism was determined for the diffusion of oxide ions from the PO2 dependence. The vacancy diffusion coefficient, DV, for LaFeO3 was nearly the same as that for LaCoO3 at the same temperature. The activation energy for migration of oxide ion vacancies was 74 kJ · mole?1 for both oxides.  相似文献   

5.
The solid state reaction in a sandwich type diffusion couple of NiO and β-Ga2O3 has been investigated between 1240 and 1550°C in air and inert gas atmosphere.Optical microscope and X-ray methods are used to identify the reaction product, which is a singlephased spinel of the general formula Ni1?yGa2+2y3O4. The homogeneity range of the spinel phase was investigated by X-ray methods; there is a high solubility of β-Ga2O3 in the spinel lattice.The growth of the thickness of the reaction layer follows a parabolic rate law, and therefore a diffusion process must be rate determining. The activation energy of the rate controlling step is 82 kcal/mole.Pt-marker experiments are not sufficient for determining the reaction mechanism. Investigations with an electron-probe microanalyzer, an connection with a modified marker technique, resulted in a Wagner-mechanism of counterdiffusion of cations for formation of nickel-gallium spinel; the total amount in Ga3+ ions is lost for spinel formation before there is an appreciable solubility of gallium in NiO.  相似文献   

6.
The solid-state reaction of the second kind in a sandwich type diffusion couple of Co1?zGa2z3O and β-Ga2O3 has been investigated between 1249 and 1550°C in air. The quantity z, which corresponds to the saturation concentration of β-Ga2O3 in CoO, was determined as a function of temperature by X-ray methods and the optical microscope; the homogeneity range of the spinel phase Co1?yGa2+2y3O4 was investigated also. The growth of the thickness of the reaction layer follows a parabolic rate law; the activation energy is 71.6 kcal/mole. A comparison of reaction rate constants of the first and second kind in connection with experimental results, achieved with a modified marker technique, leads to confirmation of the Wagner mechanism for the formation of CoGa2O4 spinel as supposed before by Laqua. Reaction rate constants of the second kind, calculated from interdiffusion profiles in CoO-β-Ga2O3 diffusion couples, are in good agreement with experimental values. Presented data are used for estimating interdiffusion coefficients for the CoO-β-Ga2O3 system according to theoretical aspects developed by Pelton, Schmalzried, and Greskovich.  相似文献   

7.
An apparatus for measuring the time dependence of light scattering intensity (LSI) in the μsec range was developed. Using this equipment, the degradation of polyphenylvinylketone (PPVK) in benzene solution and that of polymethylmethacrylate (PMMA) in acetone solution were investigated at room temperature. PPVK was irradiated with 25 nsec flashes of 3471 Å light and PMMA with 2 μsec pulses of 15 MeV electrons. The following results were obtained:PPVK: the quantum yield for main chain scission φ(S) is 0·4–·6. Quencher experiments yielded Stern-Volmer constants which agree with literature data. Half lives (τ12) of the change of LSI amount to about 20 μsec and are not influenced by quenchers. The dependence on molecular weight M? is τ12M0·22 ± 0·02. It is assumed that τ12 corresponds to disentanglement diffusion. Oxygen has no influence on φ(S) and τ1d2.PMMA: the G-value is about 2 for oxygen-free solutions and 0·7 in the presence of O2. The time dependence of LSI is significantly influenced by oxygen and presumably determined by the life time of activated centres in the polymer chains. From the decay curves, evidence was obtained for two different intermediates: one decaying rapidly with τ12 < 100 μsec, the other decaying relatively slowly with τ12 = 4·0±0·2 msec (rate constant: 170/sec. The slowly decaying species is scavenged completely by oxygen and is presumably a macroradical whereas the fast decay might indicate a scission process via excited states.  相似文献   

8.
A problem of trap diffusion, that is diffusion of point defects in crystals participating in a solid-phase chemical reaction with motionless impurity ions, is solved. Time dependences of the reaction-front displacement, Xf, and its steepness, (?C?X)f are determined analytically for N0 ? C0 and numerically for all relations of N0 and C0xf2=2N0C0Dt; (acax)f=0.3C032(gD)12>where C0 and N0 are the initial concentration of impurity and the eqilibrium defect concentration, respectively, D is a diffusion coefficient, and g is a chemical reaction constant. Dependence of Xf vs C0 and t is confirmed for oxygen annealing of corundum crystals doped with titanium which, reacting with the point defects, changes its valency. The data are obtained for dependence of displacement Xf upon partial oxygen pressure and thermotreatment temperature as well as upon the sign of the constant electric field applied to the sample. From these data we conclude that the reaction of titanium impurity, changing from the three-valent to the tetravalent state at the activation energy of 80 ± 8.5 kcal/mole is due to anisotropic diffusion of charged aluminum vacancy and holes in the valence band. The diffusion coefficient for that process at 1500°C is estimated to be larger than 10?5 cm2/sec. Using the trap-diffusion features, the concentration of optical centers of the 0.315-μm absorption band in ruby is also estimated.  相似文献   

9.
In order to elucidate the defect structure of the perovskite-type oxide solid solution La1?xSrxFeO3?δ (x = 0.0, 0.1, 0.25, 0.4, and 0.6), the nonstoichiometry, δ, was measured as a function of oxygen partial pressure, PO2, at temperatures up to 1200°C by means of the thermogravimetric method. Below 200°C and in an atmosphere of PO2 ≥ 0.13 atm, δ in La1?xSrxFeO3?δ was found to be close to 0. With decreasing log PO2, δ increased and asymptotically reached x2. The log(PO2atm) value corresponding to δ = x2 was about ?10 at 1000°C. With further decrease in log PO2, δ slightly increased. For LaFeO3?δ, the observed δ values were as small as <0.015. It was found that the relation between δ and log PO2 is interpreted on the basis of the defect equilibrium among Sr′La (or V?La for the case of LaFeO3?δ), V··O, Fe′Fe, and Fe·Fe. Calculations were made for the equilibrium constants Kox of the reaction
12O2(g) + V··o + 2FexFe = Oxo + 2Fe·Fe
and Ki for the reaction
2FexFe = FeFe + Fe·Fe·
Using these constants, the defect concentrations were calculated as functions of PO2, temperature, and composition x. The present results are discussed with respect to previously reported results of conductivity measurements.  相似文献   

10.
The reaction between nitric oxide and vibrationally excited ozone was studied in a fast flow reactor by monitoring the visible emission from electronically excited NO21. The antisymmetric mode (ν3) of O3 was excited with a Q-switched 9.6 μm CO2 laser, and a laser-induced signal was detected, with a rise rate constant of (4.0 ± 0.5) × 1011 cm3/mole sec and a decay rate constant of (1.1 ± 0.1) × 1011 cm3/mole sec for an NO-rich mixture. The latter was unaffected by addition of large amounts of He or Ar, indicating that the signal was not a thermal effect. Most of the measurements were made at 350°K; however, the He and Ar dilution results suggest that the enhanced reaction rate is not very sensitive to temperature. In order to explain the observed rise times, it was necessary to postulate an intermediate step prior to the chemical reaction. A model which is consistent with our data has energy transferred from ν3 to ν2 (the bending mode) at a rate of (2.9 ± 0.5) × 1011 cm3/mole sec for NO and a rate of (1.1 ± 0.2) × 1011 cm3/mole sec for He. According to this model, the rate constant for the reaction of NO with O3 (ν2= 1) producing vibrationally excited ground state NO22,
NO + O32 (010) 3 NO22 + O2
is (1.5 ± 0.2) × 1011 cm3/mole sec, and the relative rate for the reaction of O3 (ν2 = 1) and O32 = 0) with NO was estimated to be k3(1)k3(0) ≈ 22.  相似文献   

11.
12.
13.
14.
The scattering function P0 for classical elastic light scattering is specified for several molecular weight distribution functions (Schulz-, Square root-, Maxwell-, Normal-, Poisson-, Logarithmic-normal-distribution function). Precision measurements on anionic polymerized polystyrene from Pressure Chemical Co. (Mw = 2 · 106 and Mw = 6,7 · 106) and radical initiated polystyrene were performed in transdecalin with the light scattering photometer Fica 50. Comparison of experimental results with theoretical curves indicate that the anionic polystyrenes exhibit a broader distribution than given by Pressure Chemical Co. The distribution of the radical polystyrene conforms with distributions found by other methods.  相似文献   

15.
A new hydrate of tungsten trioxide, WO3 · 13H2O has been obtained by hydrothermal treatment at 120°C of an aqueous suspension of either tungstic acid gel or crystallized dihydrate. This hydrate has been characterized by different methods. A crystallographic study was carried out from X-ray powder diffraction. The hydrate crystallizes in the orthorhombic system: a = 7.359(3) Å, b = 12.513(6) Å, c = 7.704(5) Å, Z = 12. The existence of structural relationships between the hydrate, WO3 · 13H2O, and the product of dehydration, hexagonal WO3, has permitted us to propose a structural model in agreement with the experimental data. WO3 · 13H2O must be regarded as an interesting compound because its dehydration leads to a new anhydrous tungsten trioxide, hexagonal WO3.  相似文献   

16.
The Gibbs energy of formation of V2O3-saturated spinel CoV2O4 has been measured in the temperature range 900–1700 K using a solid state galvanic cell, which can be represented as Pt, Co + CoV2O4 + V2O3(CaO)ZrO2Co + CoO, Pt. The standard free energy of formation of cobalt vanadite from component oxides can be represented as CoO (rs) + V2O3 (cor) → CoV2O4 (sp), ΔG° = ?30,125 ? 5.06T (± 150) J mole?1. Cation mixing on crystallographically nonequivalent sites of the spinel is responsible for the decrease in free energy with increasing temperature. A correlation between “second law” entropies of formation of cubic 2–3 spinels from component oxides with rock salt and corundum structures and cation distribution is presented. Based on the information obtained in this study and trends in the stability of aluminate and chromite spinels, it can be deduced that copper vanadite is unstable.  相似文献   

17.
Reaction of BaO, Nb2O5, and Nb in mole ratios of 2.4:1.6:1 in an evacuated silica capsule at 1250°C produces a mixture of at least two products, one of which has the probable composition Ba6+xNb14Si4O47 (x ? 0.23). This compound has an hexagonal unit cell of dimensions a = 9,034 ± 0.004 Å, c = 27.81 ± 0.02 Å, probable space group P63mcm, Z = 2. Its structure has been determined from 942 independent reflections collected by a counter technique and refined by least squares methods to a conventional R value of 0.062. The basic structure consists of strings of four NbO6 octahedra sharing opposite corners, each string joined to the next by edge sharing of the end octahedra, so that the c axis corresponds to the length of a strand of seven corner-linked octahedra. Chains of three such strands are formed by corner sharing between the strands. The chains in turn are joined by NbO6 octahedra and Si2O7 groups in which the SiOSi linkage is linear. Barium atoms are in sites between the chains coordinated by 13 oxygen atoms. A second site, 15 coordinated, probably has a small amount of barium as well; the fractional occupancy for barium in this site is 0.076.  相似文献   

18.
The degradation of isotactic polypropylene in the range 390–465°C was studied using factor-jump thermogravimetry. The degradations were carried out in vacuum and at pressures of 5 and 800 mm Hg of N2, flowing at 100–400 standard mL/s. At 800 mm Hg this corresponds to linear rates of 1–4 mm/s. In vacuum bubbling in the sample caused problems in measuring the rate of weight loss. The apparent activation energy was estimated as 61.5 ± 0.8 kcal/mol (257 ± 3 kJ/mol). In slowly flowing N2 at 800 mm Hg pressure the activation energy was 55.1 ± 0.2 kcal/mol (230 ± 0.8 kJ/mol) for isotactic polypropylene and 51.1 ± 0.5 kcal/mol (214 ± 2 kJ/mol) for a naturally aged sample of atactic polypropylene. For isotactic polypropylene degrading at an external N2 pressure of 5 mm Hg the apparent activation energy was 55.9 ± 0.3 kcal/mol (234 ± 1 kJ/mol). A simplified degradation mechanism was used with estimates of the activation energies of initiation and termination to give an estimate of 29.6 kcal/mol for the ß-scission of tertiary radicals on the polypropylene backbone. Initiation was considered to be backbone scission ß to allyl groups formed in the termination reaction. For initiation by random scission of the polymer backbone, as in the early stages of thermal degradation, an overall activation energy of 72 kcal/mol is proposed. The difference between vacuum and in-N2 activation energies is ascribed to the latent heat contributions of molecules which do not evaporate as soon as they are formed. At these imposed rates of weight loss the average molecular weights of the volatiles in vacuum and in 8 and 800 mm Hg N2 are in the ratios 1–1/2–1/9.  相似文献   

19.
The kinetics of the thermal polymerization of styrene have been studied over the range 60–250°. The overall energy of activation was 86 ± 2 kJ/mole, a value identical to that obtained for the thermal polymerization of styrene in diethyl adipate. As expected, the molecular weight of the polymer decreases with increase in the temperature of the polymerization, and the ratio MwMn becomes greater than 2 for polymer formed at above 140°. The plot of log (1Mn) against (absolute temperature)?1 can be represented by two straight lines yielding 24.5 and 32,0 kJ/mole for the activation energies at temperatures below 120° and above 140°, respectively. The former value is in keeping with the molecular weight being controlled by chain transfer with monomer; the latter value would be that expected if the termination process controls the molecular weight of the polymer. Mark Houwink relationships between intrinsic viscosity and Mn and Mw have been found to apply to polymer samples when the molecular weight averages were determined by osmometry and by light scattering. However, deviations were found for low molecular weight material when measured using gel permeation chromatography. The K values were considerably lower, and the α values higher than reported in the literature.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号