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 78 毫秒
1.
In most solid state reactions the reaction velocity can be described as a product of two functionsK(T) andf(1?α) whereT is the temperature and α the degree of conversion of the solid reactant. The physical interpretation of these functions is discussed, and a systematic method is described by whichf(1?α) of a reaction is identified from its kinetic data.K(T) and the reaction mechanism are then determined. This method has been successfully applied to analyse the kinetics of the thermal decomposition of silver azide.  相似文献   

2.
The exothermal process of curing of thermoset resins in adiabatic conditions cannot be monitored by differential thermal analysis techniques such as DSC. Starting from the specific reaction rate, heat capacity as function of the temperature and the heat of reaction at some reference temperature, it is possible to design any adiabatic operation. In this paper we apply the energy balance to the curing process in adiabatic conditions and solve the basic rate law for the two empirical kinetic functionsf(α) usually used:n th-order kinetics [f(α)=(1?α)n], and autocatalytic kinetics [f(α)=αm(1?α)n], where α is the degree of conversion andn andm the reaction orders, in order to obtain the heat generation curve (dH/dt) as a function of time, as well as the change of temperature with time, the explosion time, the maximum adiabatic temperature and the rate of reaction as a function of the degree of conversion.  相似文献   

3.
4.
Synthesis, Crystal Structure, Vibrational Spectra, and Normal Coordinate Analysis of K2[OsCl5(CO)] · H2O The X-ray structure determination of K2[OsCl5(CO)] · H2O (monoclinic, space group P21/c a = 13.600(2), b = 7.122(1), c = 22.186(11) Å, β = 98.66(3)°, Z = 8) revealed two crystallographic independent bat very similar complex anions [OsCl5(CO)]2? with rough C4v point symmetry. Due to the stronger trans influence of the carbonyl group the bond lengths in the Cl? Os? CO axis Os? Cl = 2.449(2), 2.430(2) Å are langer as compared with the octahedron basis Os? Cl = 2.340-2.370 Å. The water of crystallization is coordinated to potassium (K? OH2 = 2.625-2.815 Å). Using the molecular parameters the IR and Raman spectra are assigned by normal coordinate analysis. The valence force constants are fd(CO) = 15.30, fd(OsC) = 3.88, fd(OsCl) = 1.81, fd(OsCl) = 1.36, fd(OH) = 7.65, 7.82, 7.79 mdyn/Å. The strengthening of the Os? C bond by stronger back donation of the OsIII(d5) complex in comparison with the isostructural OsIV (d4) compound is discussed.  相似文献   

5.
A new method for determination of the conversion dependence of substantial initiation rate constants k i = f(C) in free-radical polymerization processes has been developed. On the basis of the known data on k i1 = f(C) dependences for initiator I1 and the kinetic analysis of a single trivial and simple experiment, this method allows one to calculate k i2 = f(C) function for any other initiator I2 under the same conditions (monomer, temperature). The reference experiment includes measurements of polymerization rates in the presence of initiator I1 in a wide conversion range from 0 to 100% and in the presence of I2, on the condition that the rates of initiation are equal w i1 = w i2, thus ensuring equal initial rates of polymerization. The above-described approach has been approved for the polymerization of styrene, methyl methacrylate, and vinyl acetate initiated with AIBN and benzoyl peroxide.  相似文献   

6.
A method for predicting an analytical equation of state for polymer melts from surface tension and liquid state density at the freezing temperature (γf,ρf), as scaling constants, is presented. B2(T) follows a promising corresponding-states principle. Calculation of α(T) and b(T), the two other temperature-dependent constants of the equation of state, are made possible by scaling. As a result, γf and ρf are sufficient for the calculation of thermophysical properties of polymer melts. We applied the procedure to predict specific volumes of polyethylene glycol (PEG), polypropylene glycol (PPG), polypropylene (PP) and polyvinylchloride (PVC) at compressed state with temperature range from 298.15 to 423.15 K and pressures up to 200 MPa. The experimental specific volumes were correlated satisfactorily with our procedure and the results are within 3%.  相似文献   

7.
Systematic study of hyperfine structures, Zeeman and Stark effects in Sm I is performed for the lowest 7G1-6 levels belonging to the configuration 4f 66s6p by atomic-beam laser spectroscopy with fluorescence detection. The hyperfine coupling constants of 7G2-6 levels are determined. From the Zeeman splittings for the 4f 66s 2 7F2-6 ? 4f 66s6p 7G2-6 transitions, g-values are determined for the 7G2.6 levels and the precision is improved by several orders of magnitude. From the Stark splittings for the 7F0-3 ? 7G1-3 transitions, tensor polarizabilities α 2(J) are determined for the upper 7G1-3 levels. Particularly for the 7G1 level (15 650.55 cm?1) which has close-lying opposite-parity level, the isotope dependence of α 2(J) is clearly observed for the first time.  相似文献   

8.
Crystal Structure, Vibrational Spectra, and Normal Coordinate Analysis of K2[IrCl5(NH3)] The X-ray structure determination of K2[IrCl5(NH3)] (orthorhombic, space group Pnma, a = 13.426(4), b = 10.015(2), c = 6.8717(7) Å, Z = 4) revealed the Cs point symmetry of the complex anion [IrCl5(NH3)]2? (Ir? Cl = 2.337–2.365, Ir? N = 2.067(10); N? H = 0.73–0.79 Å). Using the molecular parameters the IR and Raman spectra are assigned by normal coordinate analysis. The valence force constants are fd(NH) = 5.88, fd(IrN) = 2.66, fd(IrCl) = 1.68 mdyn/Å.  相似文献   

9.
The adsorption properties of GaAs-CdS solid solutions and the constituent binary systems with respect to CO and NH3 were studies by piezoquartz microweighing, temperature-programmed desorption, and IR spectroscopy. On the basis of an analysis of the measured α p = f(T), α T = f(p), and α T = f(t) dependences, the thermodynamic and kinetic characteristics of adsorption, earlier obtained acid-base and other physicochemical characteristics of adsorbents, and the electronic properties of the adsorbate molecules, the mechanism and regularities of the adsorption processes at various conditions and compositions of the system were established. A comparison of the adsorption properties of the GaAs and CdS individual binary compounds with their (GaAs)x(CdS)1?x solutions, multicomponent systems, revealed common and distinctive features. Optimal compositions of adsorbents suitable for manufacturing primary transducers in sensors for medical and environmental purposes were determined.  相似文献   

10.
Na1?x KxTi2(PO4)3 (0 ≤ x ≤ 1) solid solutions are synthesized through ion exchange under hydrothermal conditions and a sol-gel process. The unit cell parameters are calculated for (Na,K) titanium phosphates. Cation-exchange reactions in the NaTi2(PO4)3-KTi2(PO4)3-NaCl-KCl-H2O system are studied at T = 973 K and p = 200 MPa. The solid phase with compositions in the range 0 ≤ x ≤ 0.7 is enriched with sodium; in the range 0.7 ≤ x ≤ 1.0, it is enriched with potassium. The excess functions of mixing for the solid solutions are described in terms of the Margules model. Titanium phosphates Na1?x KxTi2(PO4)3 show greater nonideality than zirconium phosphates Na1?x KxZr2(PO4)3 and lower thermodynamic stability in decay into pure components at high pressures and temperatures.  相似文献   

11.
The layered uranyl chromate, K6[(UO2)4(CrO4)7]·6H2O (1), has been synthesized by reacting UO3 with K2Cr2O7 under mild hydrothermal conditions. The structure of 1 is formed from UO22+ cations that are bound by chromate anions to yield a pentagonal bipyramidal geometry around the uranium centers. These polyhedra are bridged by chromate anions to yield one-dimensional 1[UO2(CrO4)2]2− chains. The chains are similar to the 1[UO2(CrO4)2(H2O)]2− chains found in the previously reported one-dimensional uranyl chromate, K2[UO2(CrO4)2(H2O)]·3H2O, a phase that forms concomitantly with 1. These chains are condensed with the loss of the coordinated water molecules into two-dimensional 2[(UO2)4(CrO4)7]6− layers with additional one-dimensional 1[(UO2)2(CrO4)3]2− chains. These anionic layers form a new type of anionic sheet topology and are charge balanced by both intra- and interlayer K+ cations. Crystallographic data (193 K): 1, orthorhombic, space group P21212, a=10.9583(5) Å, b=22.582(1) Å, c=7.9552(4) Å, Z=2, MoKα, λ=0.71073, R(F)=1.77% for 268 parameters with 4657 reflections with I>2σ(I).  相似文献   

12.
Cross sections for projectile and targetK x-ray emission have been measured as a function of the target thickness for the symmetric systems Ni-Ni, Cu-Cu, Nb-Nb, Ag-Ag at energies between 75 and 105 MeV. The projectileK x-rays were separated from the target ones by using the Doppler shift. Kα and Kβ energy shifts and Kβ to Kα intensity ratios were also determined and used to calculateM- andL-ionization and the correspondingK-fluorescence yield for both collision partners. At non vanishing small target thicknesses, the targetK-vacancy production cross section is generally larger than that of the projectile. By analysing the target thickness dependence of the cross sections with a two component model which takes also into account the evolution of the projectileM-,L- andK-shell population inside the solid, targetK-vacancy sharing and cross sections forKK electron capture in symmetric systems could be determined. These results are in good agreement with molecular orbital model predictions.  相似文献   

13.
Steric effects on proton transfer from, and to, hydroxylic oxygen have been studied in a series of seventeen α-methyl and a-benzyl cyclohexanols in anhydrous DMSO, under both acid and base catalysed conditions, using dynamic MNR techniques. The protonation rate constants (k1 ? 106 M-1 s-1 at 25°C) obey a Taft-Ingold relationship, containing only a steric contribution Es = EsOH + Esα, where: EsOH = 0 or 0.15 for an axial or equatorial hydroxyl respectively and Esα = ?0.070 (or ?0.115) for substituting an α-hydrogen by a methyl (or benzyl) group. An equatorial hydroxylic function is therefore 40% more reactive than its axial homologue. These kinetic data are fairly consistent with structural information resulting from IR spectroscopy (vco and vOH vibrations) and from NMR (hydroxylic chemical shifts and coupling constants).  相似文献   

14.
New potassium-conducting solid electrolytes in the mixed ferrite-aluminate K1.80(Fe1 ? x Al x )1.90V0.10O4 and K1.85(Fe1 ? x Al x )1.925P0.075O4 are synthesized and studied. In the first system, the conductivity only slightly depends on the composition; therefore, the composition of solid electrolyte can be optimized with respect to the conductivity-stability characteristics. In the second system, extremes and considerable negative deviations from the additivity are observed in the conductivity isotherms. Possible reasons for this phenomenon are discussed.  相似文献   

15.
A new technique for measuring rate constants of moderately fast electrode reactions is proposed. The input is a decreasing current ramp of the form i(t) = Δi(1 ? t/τ). The overpotential transient-response expected for the activation/diffusion model is confirmed through experiments on the system K4Fe(CN)6+K3Fe(CN)6/Pt. The rate constant and the transfer coefficient are obtaine presence of several interesting features (a maximum, a zero crossing etc.) in the η?t profile makes this technique almost unique.  相似文献   

16.
In many decomposition reactions, the reaction velocity can be described as a product of two functions: a temperature dependent part K(T) and the kinetic function f(1 – α), where T designates the temperature and α the fraction of reactant that has decomposed. The physical interpretation of these functions is discussed for both solid and homogeneous systems. A method is described by which f(1 – α) and K(T) can be determined from kinetic data. The mechanism of decomposition can subsequently be identified which should be consistent with the derived kinetic parameters. The method has been applied to analyze the kinetics of the thermal decomposition of nitromethane. © 1994 John Wiley & Sons, Inc.  相似文献   

17.
A set of constrained force constants has been derived from experimental vibrational frequency data for eighty three octahedral molecules. Superimposing the condition that the larger value for fdfdd′ be used when ffdα″ is a maximum on the six equations relating vibrational frequencies to force constants generates a seventh. This provides a uniform set of results for all 83 molecules. The values of the force constants have a simple rationale in terms of chemical bonding theory. Some preliminary calculations for SF6 show that these force constants are suitable for use in generating reliable molecular dynamical trajectory data.  相似文献   

18.
19.
The first and second protonation constants of linear polyphosphates at 25°C and ionic strength 0.1 have been evaluated: log K1 = 8.91, log K12 = 6.13 for P2O74?(P2); log K1 = 8.88, log K12 = 5.86 for P3O105?(P3); log K1 = 8.40, log K12 = 6.58 for P4O136?(P4); log K1 = 8.15, log K12 = 7.03 for P5O167?(P5); log K1 = 8.12, log K12 = 7.16 for P6O198?(P6); log K1 = 8.07, log K12 = 7.18 for P7O229?(P7). The variations of these values with the polyphosphate chain length have been discussed. A simple identification for polyphosphate species utilizing 13P NMR signal ratio of middle to end P has been proposed. By applying a micro ion-exchange technique, a rapid concentration of each polyphosphate and the subsequent preparation of the magnesium salt have been accomplished.  相似文献   

20.
Precise thermodynamic ionization constants K for 3-nitrophenol, 3,4-dichlorophenol, and 4-cyanophenol have been obtained in 1,4-dioxane-water mixtures (0–70% volume fraction in dioxane) at 25°C using a potentiometric method. The same information for another twelve cationic, neutral, and anionic phenols were taken from the literature. Three different methods were used to study the effects of the solvents on the ionization constants: one involves a single polarity parameter, E T(30); the next involves the Kamlet–Taft multiparametric method; and the last involves the preferential solvation model. The pK values follow the preferential solvation model, but the parameters obtained are highly correlated. Using the data for the phenol molecule as reference, a linear correlation between ΔpK and E T(30) has been used to develop a new method of obtaining pK values for any binary solvent composition, with only the pK in water known. The pK(s) values correlate with the molecular parameters for the dipolarity/polarizability of the solvent π* and its hydrogen-bond donor ability α. The preferential solvation parameter, f 12/1, correlates with the parameter for the hydrogen-bond donor ability of the solvent. All the phenols follow Hammett's equation and the reaction constants have been calculated for the different water–dioxane mixtures.  相似文献   

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