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 kinetics of decomposition of [Alg · Mn VIO42?] intermediate complex have been investigated spectrophotometrically at a constant ionic strength of 0.5 mol dm?3. The decomposition reaction was found to be first-order in the intermediate concentration. The results showed that the rate of reaction was base-catalyzed. The kinetic parameters have been evaluated and found to be ΔS? = ?103.88±6.18 J mol?1 K?1, ΔH? = 51.61 ± 1.02 kJ mol?1, and ΔG? = 82.57 ± 2.86 kJ mol?1, respectively. A reaction mechanism consistent with the results is discussed. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
Thermal polymerization of acrylamide was studied by differential scanning calorimetry. Latent heat of fusion ΔHf and enthalpy of polymerization ΔHp values were found to be 36 and ?18.0 kcal mol?1, respectively. The overall activation energy E for the polymerization was calculated to be 19 k cal mol?1 up to 60% conversion. The added free-radical inhibitor (benzoquinone) was found to desensitize the thermal polymerization of acrylamide suggesting the polymerization to be a free-radical type. The existing rate equation for the heterogeneous bulk polymerization in the presence of initiators has been modified for the thermally initiated bulk polymerization of acrylamide. The experimental overall E value was found to agree well with the calculated E value when considering only the propagation and termination steps, thereby suggesting the process to be similar to postpolymerization of acrylamide.  相似文献   

3.
The kinetics of decomposition of an [Pect·MnVIO42?] intermediate complex have been investigated spectrophotometrically at various temperatures of 15–30°C and a constant ionic strength of 0.1 mol dm?3. The decomposition reaction was found to be first‐order in the intermediate concentration. The results showed that the rate of reaction was base‐catalyzed. The kinetic parameters have been evaluated and found to be ΔS = ? 190.06 ± 9.84 J mol?1 K?1, ΔH = 19.75 ± 0.57 kJ mol?1, and ΔG = 76.39 ± 3.50 kJ mol?1, respectively. A reaction mechanism consistent with the results is discussed. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 35: 67–72, 2003  相似文献   

4.
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   

5.
Tautomerization of 2-benzylidene-4-methyl-3-oxo-pentanoic acid phenylamide has been studied by NMR and GC-MS. The two tautomers were separated on an HP-5 column, which enabled the kinetic and the thermodynamic behavior of on-column interconversion to be investigated. The enol-to-imide tautomerization was found to occur primarily in the stationary phase. By treating the column as a reactor, the interconversion was investigated as a function of retention time and oven temperature. This enabled determination of the rate constant (0.0605 s?1) by monitoring the increase of the less gas stable tautomer at a constant temperature of 260 °C and determination of the activation energy of the reaction for the net tautomerization (52.0 kJ mol?1), because it was found that the reaction obeyed pseudo first-order kinetics. The enthalpy and the entropy changes (?H=1.68 kJ mol?1, ?S=3.54 J K?1 mol?1) for the enol-to-imide reaction in the stationary phase were also obtained.  相似文献   

6.
This article is a critical analysis of kinetic dataavailable on carbocationic polymerizations. A survey of published propagation rate constant (kp) data revealed several orders of magnitude differences. In this article, an explanation of this apparent discrepancy is offered with a case study involving the carbocationic polymerization of 2,4,6‐trimethylstyrene (TMS). With the polymerization mechanism originally proposed for this system, kp = 1.35 × 104 L mol?1 s?1 was extracted from experimental data with the Predici polyreaction package. The alternative mechanism yielded kp = 1.01 × 107 L mol?1 s?1, close to that predicted by Mayr's Linear Free Energy Relationship (LFER). We propose that true rate constants can only be obtained from direct competition experiments or from kinetic interpretation based on independently proven mechanisms. The second part of this review discusses critical analysis of the temperature and concentration dependence of various living IB systems. Comparison of the temperature dependence in systems initiated with 2‐ chloro‐2,4, 4‐ trimethylpentane (TMPCl)/TiCl4 from various laboratories yielded of ΔH ~?25 and ?34.5 kJ/mol for high and low TMPCl/TiCl4 ratios, respectively. Aromatic (cumyl‐type) initiators show ΔH ~ ?40 kJ/mol, whereas H2O/TiCl4 in the presence of the strong electron‐ pair donor dimethylacetamide gave ΔH = ?12 kJ/mol. The significant differences indicate different underlying mechanisms with complex elementary reactions. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 5394–5413, 2005  相似文献   

7.
The kinetic of D,L-lactide polymerization in presence of biocompatible zirconium acetylacetonate initiator was studied by differential scanning calorimetry in isothermal mode at various temperatures and initiator concentrations. The enthalpy of D,L-lactide polymerization measured directly in DSC cell was found to be ΔH=−17.8±1.4 kJ mol−1. Kinetic curves of D,L-lactide polymerization and propagation rate constants were determined for polymerization with zirconium acetylacetonate at concentrations of 250–1000 ppm and temperature of 160–220 °C. Using model or reversible polymerization the following kinetic and thermodynamic parameters were calculated: activation energy Ea=44.51±5.35 kJ mol−1, preexponential constant lnA=15.47±1.38, entropy of polymerization ΔS=−25.14 J mol−1 K−1. The effect of reaction conditions on the molecular weight of poly(D,L-lactide) was shown.  相似文献   

8.
Tannase has been extensively applied to synthesize gallic acid esters. Bioimprinting technique can evidently enhance transesterification-catalyzing performance of tannase. In order to promote the practical utilization of the modified tannase, a few enzymatic characteristics of the enzyme and its kinetic and thermodynamics properties in synthesis of propyl gallate by transesterification in anhydrous medium have been studied. The investigations of pH and temperature found that the imprinted tannase holds an optimum activity at pH?5.0 and 40?°C. On the other hand, the bioimprinting technique has a profound enhancing effect on the adapted tannase in substrate affinity and thermostability. The kinetic and thermodynamic analyses showed that the modified tannase has a longer half-time of 1,710?h at 40?°C; the kinetic constants, the activation energy of reversible thermal inactivation, and the activation energy of irreversible thermal inactivation, respectively, are 0.054?mM, 17.35?kJ?mol?1, and 85.54?kJ?mol?1 with tannic acid as a substrate at 40?°C; the free energy of Gibbs (??G) and enthalpy (??H) were found to be 97.1 and 82.9?kJ?mol-1 separately under the same conditions.  相似文献   

9.
The kinetics of reduction of hexachloroplatinate(IV) by dithionite have been examined spectrophotometrically in sodium acetate?Cacetic acid buffer medium in the temperature range 288?C303?K. The reaction is first order in both platinum(IV) species and dithionite. H+ ion has an inhibiting effect on the rate in the pH range 3.68?C4.80. The pseudo-first order rate constant increased upon increasing both ionic strength and dielectric constant. The suggested mechanism involves an initial transition state between two like charged ions, which then decomposes to give SO3 2? through the intermediate formation of free radicals. The presence of free radicals was confirmed by performing the reaction in the presence of acrylamide. PtCl6 2? is finally reduced to PtCl4 2?, as confirmed by thermogravimetric analysis and IR spectrophotometry. The values of ?H?? and ?S?? associated with the rate-determining step have been calculated as 33?±?4?kJ?mol?1 and ?141?±?7?JK?mol?1, respectively. The values of ?H° and ?S° for the dissociation of HS2O4 ? are 16?±?4?kJ?mol?1 and ?14?±?7?JK?mol?1, respectively.  相似文献   

10.
Adsorption of molecular hydrogen on single-walled carbon nanotube (SWCNT), sulfur-intercalated SWCNT (S-SWCNT), and boron-doped SWCNT (BSWCNT), have been studied by means of density functional theory (DFT). Two methods KMLYP and local density approximation (LDA) were used to calculate the binding energies. The most stable configuration of H2 on the surface of pristine SWCNT was found to be on the top of a hexagonal at a distance of 3.54 Å in good agreement with the value of 3.44 Å reported by Han and Lee (Carbon, 2004, 42, 2169). KMLYP binding energies for the most stable configurations in cases of pristine SWCNT, S-SWCNT, and BSWCNT were found to be ?2.2 kJ mol?1, ?3.5 kJ mol?1, and ?3.5 kJ mol?1, respectively, while LDA binding energies were found to be ?8.8 kJ mol?1, ?9.7 kJ mol?1, and ?4.1 kJ mol?1, respectively. Increasing the polarizability of hydrogen molecule due to the presence of sulfur in sulfur intercalated SWCNT caused changes in the character of its bonding to sulfur atom and affected the binding energy. In H2-BSWCNT system, stronger charge transfer caused stronger interaction between H2 and BSWCNT to result a higher binding energy relative to the binding energy for H2-SWCNT.  相似文献   

11.
The standard Gibbs free energy, enthalpy, and entropy of complex formation of five solid molecular complexes of iodine have been determined by comparing the e.m.f.'s of galvanic cells having either solid iodine or the iodine complex as cathode. All of the complexes were found to have a negative enthalpy of formation, which was in the range ?5 to ?14 kJ mol?1, except for one very weak complex. The relative stability of the complexes was largely determined by the standard entropy of formation which varied from +18 J K?1 mol?1, for the most stable of the complexes studied, to ?21 J K?1 mol?1.  相似文献   

12.
Pyrolytic characteristics and kinetics of pistachio shell were studied using a thermogravimetric analyzer in 50?C800?°C temperature range under nitrogen atmosphere at 2, 10, and 15?°C?min?1 heating rates. Pyrolysis process was accomplished at four distinct stages which can mainly be attributed to removal of water, decomposition of hemicellulose, decomposition of cellulose, and decomposition of lignin, respectively. The activation energies, pre-exponential factors, and reaction orders of active pyrolysis stages were calculated by Arrhenius, Coats?CRedfern, and Horowitz?CMetzger model-fitting methods, while activation energies were additionaly determined by Flynn?CWall?COzawa model-free method. Average activation energies of the second and third stages calculated from model-fitting methods were in the range of 121?C187 and 320?C353?kJ?mol?1, respectively. The FWO method yielded a compatible result (153?kJ?mol?1) for the second stage but a lower result (187?kJ?mol?1) for the third stage. The existence of kinetic compensation effect was evident.  相似文献   

13.
In order to understand the mobility of uranium it is very important to know about its sorption kinetics and the thermodynamics behind the sorption process on soil. In the present study the sorption kinetics of uranium was studied in soil and the influence parameters to the sorption process, such as initial uranium concentration, pH, contact time and temperature were investigated. Distribution coefficient of uranium on soil was measured by laboratory batch method. Experimental isotherms evaluated from the distribution coefficients were fit to Langmuir, Freundlich and Dubinin?CRadushkevich (D?CR) models. The sorption energy for uranium from the D?CR adsorption isotherm was calculated to be 7.07?kJ?mol?1.The values of ??H and ??S were calculated to be 37.33?kJ?mol?1 and 162?J?K?1?mol?1, respectively. ??G at 30?°C was estimated to be ?11.76?kJ?mol?1. From sorption kinetics of uranium the reaction rate was calculated to be 1.6?×?10?3?min?1.  相似文献   

14.
The kinetics of formation of the 1?:?1 complex of chromium(III) with 1,3-propanediamine-N,N′-diacetate-N,N′-di-3-propionate (1,3-pddadp) were followed spectrophotometrically at λ max?=?557?nm. The reaction was first-order in chromium(III). Increasing the 1,3-pddadp concentration from 2.2?×?10?2 to 0.11?mol?dm?3 accelerated the reaction rate. Increasing the hydrogen ion concentration from 1.995?×?10?5 to 6.31?×?10?4 mol?dm?3 retarded the reaction rate. The reaction rate was also retarded by increasing ionic strength and dielectric constant of the reaction medium. A mechanism was suggested to account for the results obtained which involves ion-pair formation between the various reactants. Values of 22?kJ?mol?1 and ?115?J?K?1 mol?1 were obtained for the energy and the entropy of activation, respectively, which indicate an associative mechanism. The logarithm of the formation constant of the 1?:?1 complex formed was 11.3.  相似文献   

15.
The kinetics of the thermal polymerization of N-tert-butylacrylamide were investigated in 1,4-dioxane as solvent, in the 65–80°C temperature range. It was found that the overall rate of polymerization which was determined by a gravimetric method is proportional to the 1.9 power of monomer concentration at 70°C. The rate of initiation was determined by ESR spectroscopy using DPPH as an inhibitor, and it was found that the order of initiation rate is 1.8 with respect to monomer concentration at 70°C. The overall activation energy for the thermal polymerization of N-tert-butylacrylamide was found to be 64 ± 9 kJ mol?1 in the 65–80°C temperature range. The activation energy for the rate of initiation was also determined and it was found to be 90 ± 23 kJ mol-1.  相似文献   

16.
The kinetics of the binding of the neurotoxin acrylamide to the cysteine residue of glutathione has been studied. At 37 °C and pH 7.3 the second order rate constant has been determined to be 0.72 ± 0.06 mol?1 dm3 min?1 by thermospray mass spectrometry. The critical energy at pH 11.5 measured over the temperature range 10–37°C by fast atom bombardment mass spectrometry was measured as 24.6 kJ mol?1.  相似文献   

17.
A kinetic study of the persulphate initiated polymerization of methacrylamide in natural rubber latex has shown that the overall course of polymerization resembles the analogous aqueous polymerization. However, the rate of polymerization increases with increasing rubber concentration in a complex manner, together with a concomitant decrease in induction period. The overall activation energy of polymerization decreases from 18.4 to 11.8 kcal mole?1 as the rubber concentration is increased from 0–306 gl?1, at constant monomer and initiator concentrations. Activation of polymerization probably occurs by increase in the rate of initiation perhaps by interaction of the negatively charged latex particles with the persulphate dianion. The kinetic results indicate that, in contrast to earlier studies with oil soluble monomers, the site of polymerisation is the aqueous phase or surface of the rubber particles rather than the rubber interior.  相似文献   

18.
The present study deals with the immobilization of Aspergillus nidulans SU04 cellulase onto modified activated carbon (MAC). The effect of contact time, cellulase concentration, MAC dosage, and temperature for maximum immobilization percentage and immobilization capacity is investigated. The equilibrium nature of immobilization is described by Langmuir and Freundlich isotherms. The kinetic data were tested using the pseudo first order. The activation energy of immobilization was evaluated to be 11.78?J?mol?1. Results of the thermodynamic investigation indicate the spontaneity (?G <0), slightly endothermic (?H >0), and irreversible (?S >0) nature of the sorption process. Entropy and enthalpy were found to be 41.32 J?mol?1?mg?1 and 10.99?kJ?mol?1, respectively. The Gibbs free energy was found to be ?22.79?kJ?mol?1. At 80?rpm, 323?K, 2?h, 5?mg of MAC, immobilization capacity was 4.935?mg cellulase per mg of MAC from an initial cellulase concentration of 16?mg?ml?1 with retention of 70% of native cellulase activity up to 10 cycles of batch hydrolysis experiments. The diffusion studies that were carried out revealed the reaction rate as ??mol?min?1. At optimized conditions, immobilized cellulase had a higher Michaelis?CMenten constant, K m of 1.52?mmol and a lower reaction rate, V max of 42.2???mol?min?1, compared with the free cellulase, the K m and V max values of which were 0.52?mmol and 18.9???mol?min?1, respectively, indicating the affinity of cellulase for MAC matrix.  相似文献   

19.
[RuCl2(NCCH3)2(cod)], an alternative starting material to [RuCl2(cod)] n for the preparation of ruthenium(II) complexes, has been prepared from the polymer compound and isolated in yields up to 87% using a new work-up procedure. The compound has been obtained as a yellow solid without water of crystallization. The complexes [RuCl2(NCR)2(cod)] spontaneously transform into dimers [Ru2Cl(μ-Cl)3(cod)2(NCR)] (R?=?Me, Ph). 1H NMR kinetic experiments for these transformations evidenced first-order behavior. [RuCl2(NCPh)2(cod)] dimerizes slower by a factor of ten than [RuCl2(NCCH3)2(cod)]. The following activation parameters, ΔH #?=?114?±?3?kJ?mol?1 and ΔS #?=?66?±?9?J?K?1?mol?1 for R?=?CH3CN (ΔG #?=?94?±?5?kJ?mol?1, 298.15?K) and ΔH #?=?122?±?2?kJ?mol?1 and ΔS #?=?75?±?6?J?K?1?mol?1 for R?=?Ph (ΔG #?=?100?±?4?kJ?mol?1, 298.15?K), have been calculated from the first-order rate constants in the temperature range 294–323?K. The kinetic parameters are in agreement with a two-step mechanism with dissociation of acetonitrile as the rate-determining step. The molecular structures of [Ru2Cl(μ-Cl)3(cod)2(NCR)] (R?=?Me, Ph) have been determined by X-ray diffraction.  相似文献   

20.
A combined system of potassium permanganate and pyruvic acid was found to initiate radical polymerization of vinyl monomers, especially acrylamides. From kinetic investigations of the polymerization of methacrylamide, it was found that this initiator induced a radical polymerization which proceeded with an overall activation energy of 15.7 kcal/mol. The rate is given by

Rp=K[methacrylamide] 1 [pyruvic acid]° [KMnO4]1 in aqueous and water-DMF mediums. In the presence of DMF the initial rate was found to decrease but the kinetic equation remained the same. The investigations were done at 35 ± 0.2°C in nitrogen.

Besides the clinical importance of pyruvic acid found in blood, urine, muscles, etc., it is a good initiator in conjunction with KMnO4 for vinyl polymerization. It is therefore interesting to study the polymerization of methacrylamide using the KMnO4-pyruvic acid redox couple in aqueous systems in order to find whether this system follows the same kinetic features of vinyl polymerization by a radical mechanism.  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