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 31 毫秒
1.
The non-isothermal 'kinetics of the decomposition of CoOOH powder has been studied derivatographically in a temperature range of 20–450 °C in air. The reaction proceeds in two stages: up to about 280°C with an activation energy E1 = 38–50 kcal mol?1 and above that temperature with E2 = 20–25 kcal mol?1, depending on the kinetic equations which are employed. The results have been critically discussed on the basis of certain current concepts.  相似文献   

2.
Using the DTA curves the thermal decomposition of alkali persulfates for the corresponding pyrosulfates is shown to be a second order reaction with activation energies of 72.7–75.6 kcal mol?1 for sodium persulfate and 67.7–69.1 kcal mol?1 for potassium persulfate.  相似文献   

3.
The thermal decomposition of a number of compounds Cp2TiR2 (R = aryl was studied in the solid state and in various solvents. A first-order reaction was observed and activation energies of 20–30 kcal mol?1 were found depending on the nature of R. The activation energy for Cp2Ti(C6H5)2 (20–22 kcal mol?1) appeared to be independent of the reaction medium (solid state or dissolved in cyclohexane, benzene, THF, CCl4 or in the presence of tolane). Deuteration of the phenyl groups results in a higher value of the activation energy (~29 kcal mol?1), whereas deuteration of the Cp ligands does not.A reaction mechanism is proposed in which the first and rate-determining step of the decomposition is the conversion of one of the σ-bonded ligands R to a π-bonded activated state.  相似文献   

4.
In this study, we investigated the C? H bond activation of methane catalyzed by the complex [PtCl4]2?, using the hybrid quantum mechanical/effective fragment potential (EFP) approach. We analyzed the structures, energetic properties, and reaction mechanism involved in the elementary steps that compose the catalytic cycle of the Shilov reaction. Our B3LYP/SBKJC/cc‐pVDZ/EFP results show that the methane activation may proceed through two pathways: (i) electrophilic addition or (ii) direct oxidative addition of the C? H bond of the alkane. The electrophilic addition pathway proceeds in two steps with formation of a σ‐methane complex, with a Gibbs free energy barrier of 24.6 kcal mol?1, followed by the cleavage of the C? H bond, with an energy barrier of 4.3 kcal mol?1. The activation Gibbs free energy, calculated for the methane uptake step was 24.6 kcal mol?1, which is in good agreement with experimental value of 23.1 kcal mol?1 obtained for a related system. The results shows that the activation of the C? H bond promoted by the [PtCl4]2? catalyst in aqueous solution occurs through a direct oxidative addition of the C? H bond, in a single step, with an activation free energy of 25.2 kcal mol?1, as the electrophilic addition pathway leads to the formation of a σ‐methane intermediate that rapidly undergoes decomposition. The inclusion of long‐range solvent effects with polarizable continuum model does not change the activation energies computed at the B3LYP/SBKJC/cc‐pVDZ/EFP level of theory significantly, indicating that the large EFP water cluster used, obtained from Monte Carlo simulations and analysis of the center‐of‐mass radial pair distribution function, captures the most important solvent effects. © 2011 Wiley Periodicals, Inc. J Comput Chem, 2011  相似文献   

5.
The overall thermal decomposition of poly(vinylbenzyl chloride) has been carried out in vacuum and under nitrogen. The reaction order is unity over a wide range of loss of mass. The activation energy is 40 kcal mol−1 in vacuum and 30 kcal mol−1 under nitrogen. The solid residues are extensively cross-linked. The volatile compounds have been identified by GC-MS. The main products are mono- and bicyclic bichlorinated compounds.  相似文献   

6.
Thermal decomposition of formaldehyde diperoxide (1,2,4,5-tetraoxane) in aqueous solution with an initial concentration of 6.22 × 10?3 M was studied in the temperatures range from 403 to 439 K. The reaction was found to follow first-order kinetic law, and formaldehyde was the major decomposition product. The activation parameters of the initial step of the reaction (ΔH = 15.25 ± 0.5 kcal mol?1, ΔS = ?47.78 ± 0.4 cal mol?1K?1, E a = 16.09 ± 0.5 kcal mol?1) support a mechanism involving homolytic rupture of one peroxide bond in the 1,2,4,5-tetraoxane molecule with participation of the solvent and formation of a diradical intermediate.  相似文献   

7.
Hydrazinium monoperchlorate (HP-1) has been shown to decompose thermally in the solid state according to the chemical equation:
The activation energy for the evolution of HCl as determined mass spectrometrically is 8.05 kcal mol?1 in the temperature range of 80 to 120°C. The rate of decomposition is seen to be altered by doping HP-1 with small concentrations of SO2?4, Ca2+ and Al3+.  相似文献   

8.
The rate of decomposition of 2-pentoxy radical to acetaldehyde and n-propyl radical has been studied in the presence of NO in competition with nitrite formation at and above 200 kPa pressure over the temperature range of 363-413 K. The rate coefficient for the decomposition is given as log(kla/s?1) = (14.2 ± 0.4) - (13.8 ± 0.8) kcal mol?1/RT ln 10. Isomerization of 2-pentoxy radical by 1,5-hydrogen shift has been investigated in the range 279–385 K in competition with the decomposition in a static system, with methyl radicals present in high concentration to ensure trapping of the isomerized free radicals. The rate coefficient for isomerization is given as log(k3/s?1) = (11.1 ± 0.7) - (9.5 ± 1.1) kcal mol?1/RT ln 10. The implications of the results for atmospheric chemistry are discussed.  相似文献   

9.
The thermal decomposition of pure ammonium perchlorate (AP)was investigated in various gaseous atmospheres at pressures up to 51 atm, using a technique of thermal analysis. It is concluded that the first and second stage decompositions of AP in an atmosphere of oxygen or nitrogen are appreciably accelerated as the pressure is increased. Platinum has a catalytic effect in the high-temperature decomposition and suppresses the sublimation of AP at high temperature ranges in helium atmosphere even at 1 atm. The reaction heat for the high-temperature decomposition of AP in the platinum cell was calculated from the peak temperatures of DTA curves at various pressures to be 77.9 kcal mol?1. The activation energies of the sublimation in helium at 1 atm and of the high-temperature decomposition in the platinum cell at various pressures of helium have also been obtained, giving similar values of 23–25 kcal mol?1.  相似文献   

10.
The 1,3-dipolar cycloreversion pathway of the pentazole anion (N5?) to the azide anion (N3?) plus dinitrogen (N2) has been investigated using ab initio methods. At the MP4SDQ/6–31 + G* level of theory plus zero-point energy contributions, the pentazole anion is predicted to lie at 31 kcal mol?1 above the N3? + N2 system but the energy barrier for decomposition is 22 kcal mol?1. This indicates that the pentazole anion could be isolated in an inert matrix at low temperature. Comparison between extended Hückel calculations on the (N5)M(CO)3 and (C5H5)M(CO)3 complexes (with M = Fe2+, Mn+ and Cr) suggests that the N5? complexes would be formed if the fragments could be brought together. Predicted vibrational frequencies of the N5? anion are also reported.  相似文献   

11.
The adsorption and the mechanism of the oxidative dehydrogenation (ODH) of propane over VO2‐exchanged MCM‐22 are investigated by DFT calculations using the M06‐L functional, which takes into account dispersion contributions to the energy. The adsorption energies of propane are in good agreement with those from computationally much more demanding MP2 calculations and with experimental results. In contrast, B3LYP binding energies are too small. The reaction begins with the movement of a methylene hydrogen atom to the oxygen atom of the VO2 group, which leads to an isopropyl radical bound to a HO? V? O intermediate. This step is rate determining with the apparent activation energy of 30.9 kcal mol?1, a value within the range of experimental results for ODH over other silica supports. In the propene formation step, the hydroxyl group is the more reactive group requiring an apparent activation energy of 27.7 kcal mol?1 compared to that of the oxy group of 40.8 kcal mol?1. To take the effect of the extended framework into account, single‐point calculations on 120T structures at the same level of theory are performed. The apparent activation energy is reduced to 28.5 kcal mol?1 by a stabilizing effect caused by the framework. Reoxidation of the catalyst is found to be important for the product release at the end of the reaction.  相似文献   

12.
《Analytical letters》2012,45(11):1519-1525
The thermal decomposition behavior and kinetics of pyridoxine in nitrogen-only and air atmospheres were studied using thermogravimetry analysis (TGA). Kinetic interpretation of thermal analysis data for pyridoxine decomposition was carried out using Ozawa and ASTM E698 isoconversional methods. The activation energy of the decomposition process varied with the degree of decomposition and was different in the nitrogen and air atmospheres. At a 5% decomposition level, the activation energy and the pre-exponential factor were found to be 28.3 kcal mol?1 and 1.2 × 1014 min?1, respectively, in the nitrogen-only atmosphere. Thermal stability was determined by calculating the time for 5% of the pyridoxine vitamer to decompose at 25°C. The calculated shelf life for the pyridoxine vitamer obtained via TGA was surprisingly smaller in nitrogen (0.9 years) than in air (1.5 years). This is speculated to be the result of a more complex decomposition mechanism in air, involving thermo-oxidative decomposition in the presence of oxygen.  相似文献   

13.
The thermal degradation of poly(3-pentylidene carboxylate) has been studied kinetically over the temperature range 200–300°C using thermogravimetry, gas evolution analysis, and rheogoniometry together with isolation and analysis of the reaction products. The observed behavior is completely different from that previously reported for poly(isopropylidene carboxylate) and poly(methylene carboxylate). Whereas in the latter cases the decomposition occurs by a first-order intramolecular ester interchange process characterized by an activation energy in the region of 27 kcal mole?1, poly(3-pentylidene carboxylate) decomposition occurs by random chain scission superimposed on a first-order hydrogen abstraction process. The activation energy associated with this decomposition reaction is in the region of 47 kcal mole?1, and the major degradation products are cis- and trans-2-ethyl crotonic acid.  相似文献   

14.
The dissolution of UO2 in carbonate-bicarbonate solutions containing sodium hypochlorite as an oxidant has been investigated. The effect of temperature, sodium hypochlorite concentration and stirring speed was examined. In the temperature range of 303 to 318 K, the leaching reaction displayed linear kinetics. Apparent activation energy obtained from the differential approach was found to be 57 kJ mol?1. This relatively high activation energy value indicates a chemically controlled behavior of UO2 dissolution. The order of reaction with respect to sodium hypochlorite concentration was found to be unity.  相似文献   

15.
Long B  Long ZW  Wang YB  Tan XF  Han YH  Long CY  Qin SJ  Zhang WJ 《Chemphyschem》2012,13(1):323-329
The formic acid catalyzed gas‐phase reaction between H2O and SO3 and its reverse reaction are respectively investigated by means of quantum chemical calculations at the CCSD(T)//B3LYP/cc‐pv(T+d)z and CCSD(T)//MP2/aug‐cc‐pv(T+d)z levels of theory. Remarkably, the activation energy relative to the reactants for the reaction of H2O with SO3 is lowered through formic acid catalysis from 15.97 kcal mol?1 to ?15.12 and ?14.83 kcal mol?1 for the formed H2O ??? SO3 complex plus HCOOH and the formed H2O ??? HCOOH complex plus SO3, respectively, at the CCSD(T)//MP2/aug‐cc‐pv(T+d)z level. For the reverse reaction, the energy barrier for decomposition of sulfuric acid is reduced to ?3.07 kcal mol?1 from 35.82 kcal mol?1 with the aid of formic acid. The results show that formic acid plays a strong catalytic role in facilitating the formation and decomposition of sulfuric acid. The rate constant of the SO3+H2O reaction with formic acid is 105 times greater than that of the corresponding reaction with water dimer. The calculated rate constant for the HCOOH+H2SO4 reaction is about 10?13 cm3 molecule?1 s?1 in the temperature range 200–280 K. The results of the present investigation show that formic acid plays a crucial role in the cycle between SO3 and H2SO4 in atmospheric chemistry.  相似文献   

16.
The thermal decomposition process and non-isothermal decomposition kinetic of glyphosate were studied by the Differential thermal analysis (DTA) and Thermogravimetric analysis (TGA). The results showed that the thermal decomposition temperature of glyphosate was above 198?°C. And the decomposition process was divided into three stages: The zero stage is the decomposition of impurities, and the mass loss in the first and second stage may be methylene and carbonyl, respectively. The mechanism function and kinetic parameters of non-isothermal decomposition of glyphosate were obtained from the analysis of DTA?CTG curves by the methods of Kissinger, Flynn?CWall?COzawa, Distributed activation energy model, Doyle and ?atava-?esták, respectively. In the first stage, the kinetic equation of glyphosate decomposition obtained showed that the decomposition reaction is a Valensi equation of which is two-dimensional diffusion, 2D. Its activation energy and pre-exponential factor were obtained to be 201.10?kJ?mol?1 and 1.15?×?1019?s?1, respectively. In the second stage, the kinetic equation of glyphosate decomposition obtained showed that the decomposition reaction is a Avrami?CErofeev equation of which is nucleation and growth, and whose reaction order (n) is 4. Its activation energy and pre-exponential factor were obtained to be 251.11?kJ?mol?1 and 1.48?×?1021?s?1, respectively. Moreover, the results of thermodynamical analysis showed that enthalpy change of ??H ??, entropy change of ??S ?? and the change of Gibbs free energy of ??G ?? were, respectively, 196.80?kJ?mol?1,107.03?J?mol?1?K?1, and 141.77?kJ?mol?1 in the first stage of the process of thermal decomposition; and 246.26?kJ?mol?1,146.43?J?mol?1?K?1, and 160.82?kJ?mol?1 in the second stage.  相似文献   

17.
Carbon-13 and proton NMR data of macrocyclic diaromatic dilactones are presented. The observed behaviour of the spectra as a function of temperature shows that the energy barrier for the re-orientation of the side chains is lower than 49 kJ mol?1 (12 kcal mol?1) and that the energy barrier for the rotation of the aromatic rings is larger than 99 kJ mol?1 (24 kcal mol?1). Hence, chiral substituted dilactones of this type will be resolvable, and the enantiomers can be easily handled at room temperature.  相似文献   

18.
The inhibiting effect of ammonia vapours on the kinetics of the thermal decomposition of ammonium perchlorate(AP) in the temperature range 215–270°C has been investigated. An initial ammonia pressure of about 200 Torr is necessary for the practically full suppression of the decomposition of the orthorhombic crystals at temperatures close to the point of AP polymorphic transformation (240°C). With the cubic crystals, 0.5 Torr is the corresponding pressure required. In the case of complete inhibition of the decomposition in the presence of ammonia, AP crystals become yellowish. The activation energy of decomposition of the orthorhombic modification is 29 ± 0.6 kcal mole?1 in the absence of ammonia, and 38 ± 1.1 kcal mole?1 under ammonia vapour pressure of 6.5 Torr. A kinetic analysis of the traditional proton model of AP decomposition has been made showing that the increase of the activation energy in the presence of ammonia may be derived from this model.  相似文献   

19.
Thermal decomposition of bis(trifluoromethyl) peroxydicarbonate has been studied. The mechanism of decomposition is a simple bond fission, homogeneous first‐order process when the reaction is carried out in the presence of inert gases such as N2 or CO. An activation energy of 28.5 kcal mol?1 was determined for the temperature range of 50–90°C. Decomposition is accelerated by nitric oxide because of a chemical attack on the peroxide forming substances different from those formed with N2 or CO. An interpretation on the influence of the substituents in different peroxides on the O? O bond is given. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 35: 15–19, 2003  相似文献   

20.
The kinetics of the dehydroxylation of talc have been measured in the temperature interval 1100–1160 K by means of isothermal weight-change determinations. The reaction follows first-order kinetics. Over the indicated temperature range the enthalpy of activation was found to be 101±4 kcal mol?1, and the entropy of activation was found to be 16±4 cal mol?1 K?1. The error estimates correspond to one standard deviation. The enthalpy necessary to break the MgOH bond was estimated from the heat of reaction for MgOH(g) → Mg(g)+OH(g). This turns out to be 97 kcal mol?1 in reasonable agreement with the measured enthalpy of activation.These activation parameters are consistent with the mechanism proposed for dehydroxylation of talc consisting of MgOH bond scission and subsequent migration of magnesium. These results contradict a previous report on the kinetics of talc dehydroxylation in which a diffusion-controlled expression was claimed to represent the rate of talc weight loss. It is suggested that the presence of adsorbed water on the talc used in the previous investigation is responsible for the discrepancy.  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