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.
This study was undertaken to compare thermal cure kinetics of urea–formaldehyde (UF) resins, in both liquid and solid forms as a function of formaldehyde/urea (F/U) mole ratio, using multi-heating rate methods of differential scanning calorimetry. The requirement of peak temperature (T p), heat of reaction (ΔH) and activation energy (E) for the cure of four F/U mole ratio UF resins (1.6, 1.4, 1.2 and 1.0) was investigated. Both types of UF resins showed a single T p, which ranged from 75 to 118 °C for liquid resins, and from 240 to 275 °C for solid resins. As the F/U mole ratio decreased, T p values increased for both liquid and solid resins. ΔH values of solid resins were much greater than those of liquid resins, indicating a greater energy requirement for the cure of solid resins. The ΔH value of liquid UF resins increased with decreasing in F/U mole ratio whereas it was opposite for solid resins, with much variation. The activation energy (E a) values calculated by Kissinger method were greater for solid UF resins than for liquid resins. The activation energy (E α ) values calculated by isoconversional method which showed that UF resins in liquid or solid state at F/U mole ratio of 1.6 followed a multi-step reaction in their cure kinetics. These results demonstrated that thermal curing behavior of solid UF resin differed greatly from that of liquid resins, because of a greater branched network structure in the former.  相似文献   

2.
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   

3.
The thermal cure kinetics of an epoxidized linseed oil with methyl nadic anhydride as curing agent and 1-methyl imidazole as catalyst was studied by differential scanning calorimetry (DSC). The curing process was evaluated by non-isothermal DSC measurements; three iso-conversional methods for kinetic analysis of the original thermo-chemical data were applied to calculate the changes in apparent activation energy in dependence of conversion during the cross-linking reaction. All three iso-conversional methods provided consistent activation energy versus time profiles for the complex curing process. The accuracy and predictive power of the kinetic methods were evaluated by isothermal DSC measurements performed at temperatures above the glass transition temperature of the completely cured mixture (T g ). It was found that the predictions obtained from the iso-conversional method by Vyazovkin yielded the best agreement with the experimental values. The corresponding activation energy (E a) regime showed an increase in E a at the beginning of the curing which was followed by a continuous decrease as the cross-linking proceeded. This decrease in E a is explained by a diffusion controlled reaction kinetics which is caused by two phenomena, gelation and vitrification. Gelation during curing of the epoxidized linseed/methyl nadic anhydride system was characterized by rheological measurements using a plate/plate rheometer and vitrification of the system was confirmed experimentally by detecting a significant decrease in complex heat capacity using alternating differential scanning calorimetry (ADSC) measurements.  相似文献   

4.
Cure kinetics study of epoxy resin composite, employed as stator bars insulation system, was evaluated through differential scanning calorimetry using the dynamic methodology, different heating rates. These experiments provided some important information about the system as activation energy (E a = 65.4 kJ mol−1), glass transition (T g) values on different curing stages including the final cured material information and, besides that, it enables the comparison of these data with new materials under development. The activation energy value allows the determination of different energy needs of the system under evaluation, specially temperature for the material cure, improving the possibility of comparison between different insulation systems in use in the high voltage insulation business. The composite conversion degree based on the cure enthalpy (ΔH cure) at different time of cure was also subject of analysis and from that it was possible to comprehend the cure pattern which allows the cure state prediction of further samples of this type of material and the more accurate evaluation of similar samples acquired directly from stator bars.  相似文献   

5.
The effect of curing process of thermosetting powder coating consists of carboxylated polyester resin cured with triglycidyl isocyanurate has been investigated using broadband dielectric relaxation spectroscopy over a wide range of frequency (10−1-106 Hz) and temperature (70-105 °C) for different constant curing times. The molecular dynamics of the glass relaxation process (α-process) was investigated as a function of curing time, frequency, and temperature. It has been found that, only one common α-relaxation process has been observed for all measured samples of different degree of curing stages, its dynamics and broadness were found to be curing time dependent. In addition, the curing time dependence of the dielectric relaxation strength, Δε, has also been examined for the α and β-relaxation processes. The Δε for the two relaxation processes decreased strongly at the beginning of curing process and then became almost constant at longer curing times. This finding implied that the numbers of reoriented dipoles decrease with curing time as a result of the formation of three-dimensional polymer network. Furthermore, the dislocation energy, εs, calculated from the Meander model was found to be increased with increasing the curing time, i.e. the formation of a three-dimensional polymer network produces many structural defects or dislocation points. In addition, the activation energy of the curing process was calculated from the analysis of the calorimetric exothermic peaks of the curing process at different heating rates.  相似文献   

6.
Flory’s gelation theory, the non-equilibrium thermodynamic fluctuation theory and the Avrami equation have been used to predict the cure behavior of epoxy resin/organo-montmorillonite (Org-MMT)/imidazole intercalated nanocomposites at various temperatures and Org-MMT loadings. The theoretical prediction is in good agreement with the experimental results obtained by a dynamic torsional vibration method. The results show that the addition of Org-MMT reduces the gelation time tg and increases the rate of the curing reaction, the value of the kinetic constant k. The half-time t1/2 of cure after the gel point decreases with increasing of cure temperature, and the value of n is around 3 at lower temperature (<90 °C) and decreases to ∼2 as the temperature increases. The addition of Org-MMT has no obvious effect on the apparent activation energy of the cure reaction. There is no special curing process required for the formation of an epoxy resin/Org-MMT/imidazole intercalated nanocomposite.  相似文献   

7.
Thermogravimetric measurements of weight loss accompanying the pyrolysis of four pitches have been made over a range of linear heating rates. For three of the samples, the data at each heating rate could be described by an integral and a differential method of analysis, assuming a simple order function for f(α), with the result that the apparent activation energy increased with heating rate. The data for all four samples could also be satisfactorily described by the Ozawa or Friedman multiple heating rate methods, and these resulted in apparent activation energies (Ea) which increased with the value of β at which they were determined. It is suggested that this tendency for the apparent activation energy to increase, as the temperature is raised, is due to a change in the relative importance of the different reactions which lead to weight loss in this system. The apparent kinetic parameters all fall on a common compensation plot which is used to explain the relative magnitude of Ea values from Ozawa and Doyle methods of analysis. The higher values of Ea from Friedman than from Ozawa analyses are also explained.  相似文献   

8.
A two-pack polyurethane coating was analyzed using thermoanalytical techniques. The curing reaction, monitored using pressure differential calorimetry (PDSC), rheometry and dynamic mechanical analysis (DMA) shows the temperature dependency of activation energy and hence rate of curing. In-situ ATR-FTIR shows the formation of urethane linkage over time. The decomposition behavior carried out under non-isothermal mode using thermogravimetric analysis (TGA) shows bimodal behavior. The activation energies of the initial step (10% decomposition) from both iso and non-isothermal experiments are in very good agreement with each other. The use of dynamic mechanical analysis (DMA) shows the difference in glass transition behavior (Tg) and elastic modulus (E') due to the different state of cure. Also the coating exhibits a very broad loss modulus peak (E') indicating higher energy dissipation with deformation.This revised version was published online in November 2005 with corrections to the Cover Date.  相似文献   

9.
The Flory's gelation theory, non-equilibrium thermodynamic fluctuation theory and Avrami equation have been used to predict the gel time t g and the cure behavior of epoxy resin/organo-montmorillonite/diethylenetriamine intercalated nanocomposites at various temperatures and organo-montmorillonite loadings. The theoretical prediction is in good agreement with the experimental results obtained by dynamic torsional vibration method, and the results show that the addition of organo-montmorillonite reduces the gelation time t gand increases the rate of curing reaction, the value of k, and half-time of cure after gelation point t1/2 decreases with the increasing of cure temperature, and the value of n is ~2 at the lower temperatures (<60°C) and decreases to ~1.5 as the temperature increases, and the addition of organo-montmorillonite decreases the apparent activation energy of the cure reaction before gelation point, but has no apparent effect on the apparent activation energy of the cure reaction after gelation point. There is no special curing process required for the formation of epoxy resin/organo-montmorillonite/diethylenetriamine intercalated nanocomposite. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

10.
The phosphorescence spectra and lifetimes of 2,4-, 2,5-, and 3,4-dimethylbenzaldehydes dispersed in durene single crystals have been measured as a function of temperature between 10 and 200 K. For all the guests involved, the vibrational structures of the spectra are found to be temperature dependent. This is interpreted in terms of two emissions that proceed from a triplet state having predominantly a ππ* character at low temperatures and from a thermally populated triplet state having essentially a nπ* character at higher temperatures. The energy gaps ΔET between 3ππ* and 3nπ* states evaluated spectroscopically are found to be 100, 70, and 340 cm?1, respectively for 2,4-, 2,5- and 3,4-dimethylbenzaldehydes.Activation energies ΔE* determined from the Arrhenius plots of the phosphorescence decay rate constants are in good agreement with the ΔET for the first two guests. In contrast, the ΔE* are higher than the ΔET for 3,4-dimethylbenzaldehydes as well as for 2,4,5-trimethylbenzaldehyde (where ΔET ≈ 400 cm?1) because of the rapid increase of radiationless transitions in the temperature range where thermal population of the upper 3nπ* state is efficient. In the low and high temperature ranges, the phosphorescence decays for all these guests are exponential. In the intermediate range, these decays are non-exponential. The origin of these non-exponential decays is discussed.  相似文献   

11.
Heats of mixing and excess volumes at infinite dilution have been obtained at 25°C. for polydimethylsiloxane or its lower oligomers in various solvents by using a twin conduction microcalorimeter and from the pycnometric specific volumes. From those values, excess energies ΔEMv at constant volume have been determined. The prediction on intramolecular conformation contributions to the heat of solution as proposed by Bianchi has been evaluated by the values of ΔEMv. The heat of solution in the polymer–solvent systems was interpreted by the expression for ΔEMv derived from the Van Laar-Hildebrand work on simple liquid mixtures with the solubility parameters of polymers obtained from indirect measurements. The values of conformational intramolecular energy change calculated from dilute solution properties were difficult to rationalize with our results. Our present results suggest that systems in a nonideal state can not be distinguished for certain from those in the ideal state. This conclusion based on apparent values does not deny the possible effect of the conformational energy change.  相似文献   

12.
The title terpolymer (8-HQ5-SAMF-II) is synthesized by the condensation of 8-hydroxyquinoline 5-sulfonic acid (8-HQ5-SA) and melamine (M) with formaldehyde (F) in the presence of acid catalyst and using 2:1:3 M proportions of the reacting monomers. The synthesized terpolymer resin is then characterized by different physicochemical techniques viz. number average molecular mass determination, intrinsic viscosity determination, and spectral studies like UV–Visible, IR, 1H NMR, and 13C NMR spectra. The morphology of synthesized terpolymer was studied by scanning electron microscopy (SEM). The thermogravimetry of the terpolymer resin prepared in this study has been carried out by non-isothermal thermogravimetry technique in which sample is subjected to condition of continuous increase in temperature at linear rate. Thermal study of the resin was carried out to determine their mode of decomposition and relative thermal stabilities. Thermal decomposition curves were studied carefully with minute details. The Freeman-Carroll and Sharp-Wentworth methods have been used in the present investigation to calculate thermal activation energy and different kinetic parameter of the terpolymer resins. Thermal activation energy E a calculated with the two above-mentioned methods are in close agreement. The advantage of Freeman-Carroll method is to calculate both the order of reaction n and energy of activation in one single stage by keeping heating rate constant. By using data of thermogravimetry, various thermodynamic parameters like frequency factor Z, entropy change ΔS, free energy change ΔF, and apparent entropy S* have been determined using Freeman-Carroll method.  相似文献   

13.
Nanocomposites from nanoscale silica particles(NS),diglycidylether of bisphenol-A based epoxy(DGEBA),and 3,5-diamino-N-(4-(quinolin-8-yloxy) phenyl) benzamide(DQPB) as curing agent were obtained from direct blending of these materials.The effect of nanosilica(NS) particles as catalyst on the cure reaction of DGEBA/DQPB system was studied by using non-isothermal DSC technique.The activation energy(E_a) was obtained by using Kissinger and Ozawa equations. The E_a value of curing of DGEBA/DQPB/10%NS system showed a decrease of about 10 kJ/mol indicating the catalytic effect of NS particles on the cure reaction.The E_a values of thermal degradation of the cured samples of both systems were 148 kJ/mol and 160 kJ/mol,respectively.The addition of 10%of NS to the curing mixture did not have much effect on the initial decomposition temperature(T_i) but increased the char residues from 20%to 28%at 650℃.  相似文献   

14.
Isothermal crystallization of an As2Se3 undercooled melt was studied by differential scanning calorimetry and described using the classical theory of nucleation and crystal growth. The maximum rate of nucleation and crystal growth was observed to occur at approximately 235 and 350 °C, respectively. The activation energies of nucleation and crystal growth were determined to be ΔE D = 311 kJ mol?1 and ΔE* = 104 kJ mol?1, respectively. The temperature dependencies of both the activation free energy of nucleation, ΔG*, and the critical diameter, r*, were also calculated.  相似文献   

15.
A stirring bar type agitation system has been designed and characterized for the accelerating rate calorimeter (ARC). The device allows stirring of the contents of a standard ARC sample container at stirring rates of up to 500 rev. min?1, depending on sample viscosity. Experiments on a well-characterized thermal decomposition reaction, such as that of di-t-butyl peroxide, indicate that the device does not degrade the measurement of the energy of reaction, ΔEv, the Arrhenius activation energy, Ea, or the pre-exponential factor, A.The utility of this stirring apparatus is demonstrated by examining the runaway data of a suspension polymerization. The results indicate that a polymerization “kill” agent can be successfully used for that particular reaction.  相似文献   

16.
The quantum-chemical model formerly designed is described, which explains the high reactivity of the d 0-transition metal-carbon σ-bond in concerted reactions including polymerization of olefins. In this model, the energy of transition from the ground singlet state to the excited singlet or triplet state corresponding to the transfer of electron density from the metal-carbon bond to vacant d-atomic orbitals correlates with its relative elongation after which a change in the valence state of a metal begins. This change is caused by the difference in the geometries of valence s-, p-, and d-atomic orbitals having close energies; as a result, at a certain bond elongation, partial uncoupling of electrons involved in bonding takes place so that one of them becomes localized on d-atomic orbital. This process facilitates formation of the reactive state of the bond of the biradical type and leads to a reduction in the energy barrier of the insertion of an olefin molecule into this bond. Lower energies of this barrier correspond to lower values of ΔE(S 0S j ) and ΔE(S 0T j ). As shown at the example of zirconocenes Cp2ZrMe2 and Me2CCp2ZrMe2, alongside with a reduction in the energy barrier of olefin insertion into the Zr-Me bond of the cationic complex being formed, the characteristic absorption band in the spectrum of the corresponding neutral derivative shifts to the long-wave region. This band is attributed to the transfer of the electron density from the Zr-Me bond to the Zr atom. Analysis is performed of the causes for bathochromic shifts of the long-wave absorption band for adduct formation in the systems including metallocene and Al-containing cocatalyst.  相似文献   

17.
The infrared photodissociation spectra of C2H4 and CH3OH dimers are measured for different internal excitation energies ΔE. The dimers are size-selected in a scattering process with He and the internal energy is varied by using (1) different collision energies, (2) different scattering angles and (3) by measuring the complete energy loss spectra with and without laser radiation by time-of-flight analysis of the scattered clusters. For (CH3OH)2 the width Γ of the spectrum increases strongly with ΔE, while the cross section at the maximum is constant, a normal behaviour for an inhomogeneously broadened system. For (C2H4)2 Γ is nearly constant after a sort of threshold and the cross section increases with increasing ΔE. These results are explained by hot band excitation and the coupling to the darkv 10-mode.  相似文献   

18.
Ab initio RHF calculations with the 3–21G basis set have been carried out on cycloadditions of CF2 and CCl2 with ethylene. Although π-complex intermediates are predicted for both reactions at this level, MP2/3-21G calculations imply that there are no complexes in reactions of CCl2 or more reactive carbenes with ethylene or substituted alkenes. Nevertheless, negative activation energies can be observed, since ΔG reaches a maximum due to the increase in —TΔS for these bimolecular reactions. The apparent “entropy control” for reactive carbenes results from the rapid decrease in ΔH. As the reactivity of the alkene increases, the transition state shifts to an earlier point on the free energy surface, where —TΔS3 is lower, but ΔH3 is higher than for less reactive alkenes. Model potentials are developed for ΔH and —TΔS which reproduce experimental behavior, without the assumption of intermediates.  相似文献   

19.
To probe the dependence of particle size on the safety of nitroamine explosives, coarse RDX and HMX were comminuted to nanometer particles by an improved superfine mill. Their thermolysis characteristics were studied by thermal analysis and described via calculating some thermodynamic and kinetic parameters such as the activation free-energy (ΔG ), activation enthalpy (ΔH ), activation entropy (ΔS ), apparent activation energy (E), critical temperature of thermal explosion (T b), and critical heating rate of thermal explosion ( $ ({\text{d}}T/{\text{d}}t)_{{T_{\text{b}} }} $ ). After comminuted, the values of T b and $ ({\text{d}}T/{\text{d}}t)_{{T_{\text{b}} }} $ were increased. However, the values of ΔH , ΔS , ΔG , and E for nanoexplosives were close to those of microexplosives, which mean decreasing particle size into nanometer did not distinctly influence the thermolysis characteristic of nitroamines. The safety of the nanoexplosives was practically assessed by testing their impact, friction, and shock sensitivities. Results indicated that nano nitroamines presented obviously higher safety than the micro-counterparts. Especially in Small Scale Gap Test, the shock sensitivity of nano-RDX and nano-HMX decreased by about 45 and 56% compared with that of micro-RDX and micro-HMX, respectively.  相似文献   

20.
The results of dynamic and quasi-isothermal thermodesorption of water from a silica gel surface at low furnace heating rates in the temperature range 20–125° are presented. From the experimental results, the water film pressure π on silica gel surface, the activation energy ΔE and the evaporation heat ΔH were calculated. An interpretation of π changes in relation to the film thickness and wetting process has been proposed. It is concluded that the characteristic film pressure values correspond to the work of spreading, and immersional, adhesional and adhesional-cohesional wetting. From the determined film pressure values, the average value of the silica gel polar component, γ s P , was calculated to be 114.67 mJ/m2.  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