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 overall activation energy of the thermal degradation of polyisobutylene has been measured using factor-jump thermogravimetry to be 206±1 kJ/mole over the range 365 to 405° in N2 at 800 mm Hg pressure and flowing at 4 mm/s over the sample. This is consistent with some values reported for thermal degradation in vacuum and in solution. In 5 mm Hg of N2, an apparent activation energy of 218±2 kJ/mole was found, and in vacuum the apparent activation energy is 238±13 kJ/mole. Troublesome bubbling made the vacuum values difficult to measure. Substitution of reasonable values for the activation energies of initiation,E i , termination,E t , and the activation energy,E a , for vacuum degradation in the equationE a =E i /2E d -E t /2 yields an activation energy Ed=84 kJ/mole for the unzipping reaction. This equation presupposes a degradation mechanism of random initiation, unzipping, and bimolecular termination. Substitution of reasonable values for the heat of polymerization, ΔH, in the definition ΔH=E p ?e d suggests that the activation energy of the polymerization reaction at 375° is approximately 30 kJ/mole.  相似文献   

2.
Application of iso-temperature method of multiple rate to kinetic analysis   总被引:2,自引:0,他引:2  
A new method of the multiple rate iso-temperature was used to define the most probable mechanism g(α) of a reaction; the iterative iso-conversional procedure has been employed to estimate apparent activation energy E a, the pre-exponential factor A was obtained on the basis of E a and g(α). In this new method, the thermal analysis kinetics triplet of dehydration of calcium oxalate monohydrate is determined, which apparent activation energy E a is 82.83 kJ mol-1, pre-exponential factor A is 1.142·105-1.235·105 s-1, the most probable mechanism belongs to phase boundary reaction Rn with integral form g(α)=1-(1-α)n and differential form f(α)=n(1-α)1-(1/n), where accommodation factor n=2.40-1.40.  相似文献   

3.
The quenching of polymerization with a chromium oxide catalyst by radioactive methanol 14CH3OH enables one to determine the concentration of propagation centers and then to calculate the rate constant of the propagation. The dependence of the concentration of propagation centers and the polymerization rate on reaction time, ethylene concentration, and temperature was investigated. The change of the concentration of propagation centers with the duration of polymerization was found to be responsible for the time dependence of the overall polymerization rate. The propagation reaction is of first order on ethylene concentration in the pressure range 2–25 kg/cm2. For catalysts of different composition, the temperature dependence of the overall polymerization rate and the propagation rate constant were determined, and the overall activation energy Eov and activation energy of the propagation state Ep were calculated. The difference between Eov and Ep is due to the change of the number of propagation centers with temperature. The variation of catalyst composition and preliminary reduction of the catalyst influence the shape of the temperature dependence of the propagation center concentration and change Eov.  相似文献   

4.
The kinetics of free-radical cross-linking polymerization of methyl methacrylate (MM) in the presence of poly[2-(10-undecenoyloxy)ethyl methacrylate] (PUDEM) as a macromolecular cross-linker has been isothermally examined within the temperature range from 85–100°C using the differential scanning calorimetry (DSC). The activation energy found for this reaction, E a=89.3 kJ mol–1, exceeds slightly the literature values of activation energy obtained for the mass polymerization of MM without any cross-linking agent. The activation energy has been also determined by the isoconversion method. It has been found that E a decreases with the increase in the conversion, which may indicate a change in the reaction mechanism.This work was partly supported by the Committee for Research (KBN) in the framework of project No. 7 T08E 026 20  相似文献   

5.
The free-radical bulk polymerization of 2,2-dinitro-1-butyl-acrylate (DNBA) in the presence of 2,2′-azobisisobutyronitrile (AIBN) as the initiator was investigated by DSC in the non-isothermal mode. Kissinger and Ozawa methods were applied to determine the activation energy (E a) and the reaction order of free-radical polymerization. The results showed that the temperature of exothermic polymerization peaks increased with increasing the heating rate. The reaction order of non-isothermal polymerization of DNBA in the presence of AIBN is approximately 1. The average activation energy (92.91±1.88 kJ mol −1) obtained was smaller slightly than the value of E a=96.82 kJ mol−1 found with the Barrett method.  相似文献   

6.
The kinetic parameters of the complex reaction between phenol and formaldehyde in the presence of sodium hydroxide (NaOH) have been obtained by differential scanning calorimetry (DSC). The two dominant reactions appear to be addition of formaldehyde to phenol with formation of o-hydroxymethyl-phenol and subsequent condensation of the latter. For both reactions, the activation energy (Ea), reaction order and rate constants at different temperatures have been determined. Ea for addition changes from 23·7 to 19·3 kcal mole?1 and for condensation from 22·9 to 19·1 kcal mole?1 when the amount of NaOH is increased from 0·25 to 1·00 per cent. The reaction order for addition is 2 and for condensation 1. Thus DSC appears useful for studying the kinetics of more complex polymerization reactions.  相似文献   

7.
Experimental kinetic data on reactions of the chlorine atom with halogenated derivatives of methane and ethane (37 reactions) have been analyzed by the intersecting-parabolas method. The following five factors have an effect on the activation energy of these reactions: the enthalpy of reaction, triplet repulsion, the electronegativities of the reaction center atoms, the dipole–dipole and multidipole interactions between the reaction center and polar groups, and the effect of π electrons in the vicinity of the reaction center. The increments characterizing the contribution from each factor to the activation energy of the reaction have been calculated. The contribution from the polar interaction, ΔE μ, to the activation energy depends on the dipole moment of the polar group and obeys the following empirical equation: ln(ΔE μ/Σμ) = ?0.74 + 0.87(ΔE μ/Σμ) ? 0.084(ΔE μ/Σμ)2.  相似文献   

8.
The inhibition of ethylene polymerization with radioactive carbon monoxide (14CO) was used to obtain data on the number of active sites (CP) and propagation rate constant (kP) at ethylene polymerization in the temperature range of 35–70 °C over supported catalysts LFeCl2/Al2O3, LFeCl2/SiO2, and LFeCl2/MgCl2 (L: 2,6‐(2,6‐(Me)2C6H3N = CMe)2C5H3N) with activator Al(i‐Bu)3. The values of effective activation energy (Eeff), activation energy of propagation reaction (EP), and temperature coefficients of variation of the number of active sites (ECp = Eeff ? EP) were determined. The activation energies of propagation reaction for catalysts LFeCl2/Al2O3, LFeCl2/SiO2, and LFeCl2/MgCl2 were found to be quite similar (5.2–5.7 kcal/mol). The number of active sites diminished considerably as the polymerization temperature decreased, the ECp value being 5.2–6.2 kcal/mol for these catalysts at polymerization in the presence of hydrogen. The reactions of reversible transformations of active centers to the surface hydride species at polymerization in the presence and absence of hydrogen are proposed as the derivation of ECp. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 6621–6629, 2008  相似文献   

9.
In this work, the interaction of three Li+-doped polycyclic hydrocarbons (Li+-DPH) with H2 and H2O was calculated to investigate the effect of curvature of substrate on the interaction energy (Eint). For this purpose, the Eint and its decomposed energy components (electrostatic (Eelec), exchange (Eexch), induction (Eind), and dispersion energy (Edisp)) were calculated using DF-SAPT (DFT) methodology for the selected systems (Li+-(3,3) carbon nanotube (Li+-CNT33), Li+-(6,6) carbon nanotube (Li+-CNT66), and Li+-graphene). According to the results, Eint does not change significantly with curvature for the interaction between both H2 and H2O gases and the selected Li+-DPH. Since the variation of the Eint with the curvature of Li+-DPH is not significant, the selection of a planar Li+-DPH is a trustworthy model to develop a general force field for describing the interaction between a Li+-DPH and adsorbed gases. The results reveal that, in the case of the H2, the components Eelect, Eexch, Eind, and Edisp have shown a decreasing trend with Li+-DPH’s curvature decrement. However, for the H2O, Eelect, Eexch, and Eind decrease from the Li+-CNT33 to the Li+-CNT66 while they increase from the Li+-CNT66 to the Li+-graphene. In this case, the Edisp increases with a decrease of the curvature of Li+-DPH. Finally, it can be seen that although the variation of the Eint with the curvature of Li+-DPH is not significant, the variation trend of the interaction energy components and the amount of variation depend on the gas molecule and in some cases are not negligible.  相似文献   

10.
The template polymerization of N-vinylpyrrolidone (NVP) along syndiotactic poly(methacrylic acid) (s1-PMAA) templates has been studied by differential scanning calorimetry (DSC) using the scanning as well as the isothermal technique. The resulting Arrhenius plot covers a temperature range between 65 and 120°C and two parts can be distinguished. Below 80°C the overall activation energy, Ea, and entropy ΔS, are 76 kJ · mol?1 and ?79 J · mol?1 · K?1 respectively, in excellent agreement with previous dilatometric results. These values differ slightly from those of the blank polymerization leading to rate enhancement by a factor of only two. The small difference in activation parameters is explained by the occurrence of desolvation of st-PMAA chains during propagation of the polyvinylpyrrolidone (PVP) radicals along the template. Above 80°C, the decreasing tendency to form complexes between PVP and st-PMAA results in a decreasing template effect and a gradual change of apparent Ea and ΔS values towards those of the blank polymerization. Similar results were obtained with atactic and isotactic PMAA templates, but smaller rate enhancements were observed due to weaker complex formation.  相似文献   

11.
The formation of 2,3,4,5-tetraphenylnickelole-bis(triphenylphosphine) (IIIa) and 2,3,4,5-tetraphenylnickelole-bis(1,2-diphenylphosphino)ethane (IIIb), either from (E,E)-1,2,3,4-tetraphenyl-1,3-butadien-1,4-ylidenedilithium (I) and the corresponding nickel(II) chloride-phosphine complexes (II) or from the reduction of η4-tetraphenylcyclobutadienenickel(II) bromide dimer (XII) in the presence of phosphines, proceeds in good yields. Nickelole IIIa displays physical and chemical properties consistent with its structure and is a catalyst for the trimerization of diphenylacetylene. Nickelole IIIb is a highly associated structure but in its chemical response to alkynes, HOAc, O2, Br2, NaAlEt2H2 and heat displays the properties of a nickelole, rather than a cyclobutadienenickel(0) complex. Attempts to generate IIIb photochemically from η4-1,5-cyclooctadiene(η4-tetraphenylcyclopentadienone)nickel and diphos failed, but it was shown that structural types, such as η4-tetraphenyl-cyclopentadienone(diphos)nickel (a model for the structure suggested by Hoberg and Richter for IIIb), are unstable.Oligomerizations of diphenylacetylene by bis(1,5-cyclooctadiene)nickel were retarded by conducting the reaction in THF or in the presence of diphos. This retardation permitted the interception of products (cis-stilbene and (E,E)-1,2,3,4-tetraphenyl-1,3-butadiene) diagnostic for the intermediacy of nickelirenes and nickeloles. Deuterium labeling verified the presence of carbon-nickel bonds. These trapping experiments, together with findings on the thermal behavior of nickeloles, are combined into a comprehensive view of the cyclotrimerization, cyclotetramerization and linear polymerization of alkynes by nickel(0).  相似文献   

12.
The aqueous emulsifier-free emulsion polymerization of methyl methacrylate (MMA) was studied under the catalytic effect of in situ developed bivalent transition metal-EDTA complex with ammonium persulfate (APS, (NH4)2S2O8) as initiator. Out of these, Cu(II)-EDTA system was selected for detailed kinetic and spectrometric study of polymerization. The apparent activation energy Ea, 34.5 kJ/mol, activation energy of initiator decomposition Ed, 26.9 kJ/mol, energy of propagation Ep, 29 kJ/mol and energy of termination Et, 16 kJ/mol were reported. The emulsion polymer (PMMA) latex was characterized through the determination of the size and morphology by scanning electron microscopy, the average molecular weight by GPC and viscosity methods and the sound velocity by ultrasonic interferometer. From the kinetic results, the rate of polymerization, Rp at 50 °C was expressed by
  相似文献   

13.
By introducing equivalence classes of critical points of potential energy hypersurfaces a unique topological space, Reaction Topology (3N E, T C ) is defined over an Euclidean nuclear configuration space 3N E for a system of N nuclei and k electrons. Relations between the topological concepts of molecular structure and reaction mechanism are analyzed. Topological equivalences between Euclidean and Riemannian representations of nuclear configuration spaces are exploited for the analysis of quantum mechanical reaction networks of all possible chemical reactions over the given potential energy hypersurface.  相似文献   

14.
《Fluid Phase Equilibria》1988,39(2):193-209
Published values of the thermodynamic excess functions (gE, hE, cEP) are used to fit the temperature-dependent parameters of an empirical gE model. A modified Redlich-Kister approach with six concentration-dependent coefficients is chosen as a model equation for the excess Gibbs energy. The temperature dependence of the coefficients is defined in such a way that the result is a quadratic dependence of the molar excess heat capacity cEP as a function of the temperature. The fitting is carried out over the temperature range 25–150°C using the maximum likelihood method combined with the Gauss-Newton procedure (Anderson et al.). The hE and cEP data, in particular, are clearly reproduced much better than when a fit based only on VLE data is used.  相似文献   

15.
The results of experimental investigation on the study of dissolution kinetics of a Nigerian galena ore in hydrochloric acid solution were discussed. The influence of acid concentration, temperature, particle size, stirring speed and solid/liquid ratio on the extent of dissolution was examined. The elemental analysis by XRF showed that the galena ore is composed mainly of PbS with metals such as Sn, Fe and Zn occurring as minor elements and Mn, Rb, Sr and Nb as traces. The XRD analysis indicated galena as the dominant mineral phase, with the presence of associated minerals, such as α-quartz (SiO2), sphalerite (ZnS), cassiterite (SnO2), pyrite (FeS2) and manganese oxide (MnO2).Results of leaching studies showed that galena dissolution in HCl solution increases with increasing acid concentration and temperature; while it decreases with particle diameter and solid/liquid ratio at a fixed stirring rate of 450 rpm. The study showed that 94.8% of galena was dissolved by 8.06 M HCl at 80 °C within 120 min with initial solid/liquid ratio of 10 g/L. The corresponding activation energy, Ea was calculated to be 38.74 kJ/mol. Other parameters such as reaction order, Arrhenius constants, reaction and dissociation constants were calculated to be 0.28, 73.69 s?1, 1.73 ± 0.13 × 103 and 1.37 ± 0.024 × 104 mol L?1 s?1, respectively. The mechanism of dissolution of galena was established to follow the shrinking core model for the diffusion controlled mechanism with surface chemical reaction as the rate controlling step for the dissolution process. Finally, the XRD analysis of the post-leaching residue showed the presence of elemental sulphur, lead chloride and α-quartz.  相似文献   

16.
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   

17.
The isothermal compressibilities KT for cyclohexane + benzene, cyclohexane + toluene and benzene + toluene systems at 25, 35, 45 and 60°C have been used to test the Prigogine-Flory theory using Van der Waals and Lennard-Jones energy potentials. Flory's energy parameter X 12 was calculated for these systems at the four temperatures. From X 12 for the equimolar mixture, the following excess functions were calculated: (?VE/?p)T which is related to K T E , the heat of mixing H E , and the excess volume V E . The theory and any of the two potentials give (?VE/?p)T which fit the experimental data, but H E and V E , calculated using the same X 12 parameter, depart appreciably from the experimental data even though they agree in sign and have the essential features of the excess functions. The departure is apparent in both magnitude (in particular for the cyclohexane + benzene, and cyclohexane + toluene systems) and in the temperature dependence. The conclusion is that the X 12 parameter does not predict the thermodynamic properties of these systems and the Lennard-Jones potential, involving a more complicated expression, does not contribute any improvement over the Van der Waals potential.  相似文献   

18.
The temperature dependence of the rate constant for the reactions of HO2 with OH, H, Fe2+ and Cu2+ has been determined using pulse radiolysis technique. The following rate constants, k (dm3 mol−1 s−1) at 20°C and activation energies, Ea (kJ mol−1) have been found. The reaction with OH was studied in the temperature range 20–296°C (k=7.0×109, Ea=7.4) and the reaction with H in the temperature range 5–149°C (k=8.5×109, Ea=17.5). The reaction with Fe2+ was studied in the temperature range 16–118°C (k=7.9×105, Ea=36.8) and the reaction with Cu2+ in the temperature range 17–211°C (k=1.1×108, Ea=14.9).  相似文献   

19.
The solvated ion pair [(C8H17)3NCH3]+[RhCl4], formed from aqueous rhodium trichloride and Aliquat®-336 in a two-phase liquid system, was shown to hydrogenate α,β-unsaturated ketones and esters selectively at the C-C double bonds. The reduction of benzylideneacetophenone was found to follow first-order kinetics in the substrate only below 0.2 M, and to approach second-order in H2; at partial pressures of < 0.12 atm. The catalysis also proved to depend on the nature of the solvent, the phase transfer catalyst and the stirring rate. The observed activation energy Ea = 12.4 kcal mol−1 suggests that the process is both chemically and diffusion controlled.  相似文献   

20.
The kinetic law dα/dr=k(1-α)[1-(1-α)2b]1,2 proposed for the pyrolysis of polystyrene is shown to be valid for the pyrolysis of polypropylene taking into account only the percentage of isotactic polymer.As the experimental activation energy of 265 kJ mol?1 is of the same order of magnitude the as the theoretical energy calculated by the equation E = 1 2 (Eir-Er)-Er it can be concluded that the decomposition mechanism is governed by a depolymerization reaction as the principal products obtained are compounds with 3n carbon atoms.  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