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 tensile drawing behavior of poly(L -lactide) has been studied in order to obtain high strength fibers. Elongational viscosity measurements indicated that the hot drawing can take place in two temperature regions with different activation energies. Up to 180°C, the deformation proceeds in the semicrystalline state of the polymer having an activation energy of 15–28 kJ/mol, presumably by shear deformation. In the range of 180–190°C, the deformation proceeds in the liquid state of the polymer having an activation energy of 145–165 kJ/mol, leading to a semicrystalline state by strain hardening after displacement of topological defects. By using high deformation rates during drawing in a temperature gradient (tube drawing), the deformation will principally proceed in the semicrystalline region and inhomogeneous draw will take place leading to inferior fiber properties, unless the deformation rate and drawing temperature are strictly adjusted. Homogeneous drawing can be achieved by applying low deformation rates so that the deformation may take place in the liquid state of the polymer in which individual chains can be easily aligned and topological defects can be removed. Poly(L -lactide) fibers with tensile strengths of 2.3 GPa have been produced in this way.  相似文献   

2.
A novel high‐resolution thermogravimetry (TG) technique in a variable heating rate mode that maximizes resolution and minimizes the time required for TG experiments has been performed for evaluating the thermal degradation and its kinetics of Kevlar fiber in the temperature range ∼ 25–900°C. The degradation of Kevlar in nitrogen or air occurs in one step. The decomposition rate and char yield at 900°C are higher in air than in nitrogen, but the degradation temperature is higher in nitrogen than in air. The initial degradation temperature and maximal degradation rate for Kevlar are 520°C and 8.2%/min in air and 530°C and 3.5%/min in nitrogen. The different techniques for calculating the kinetic parameters are compared. The respective activation energy, order, and natural logarithm of preexponential factor of the degradation of Kevlar are achieved at average values of 133 kJ/mol (or 154 kJ/mol), 0.7 (or 1.1), and 16 min−1 (or 20 min−1) in air (or nitrogen). The technique based on the principle that the maximum weight loss rate is observed at the minimum heating rate gives thermal degradation results that were in excellent agreement with values determined by traditional TG experiments. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 71: 565–571, 1999  相似文献   

3.
A kinetic analysis based on thermal decomposition of rejected polypropylene, plastic film and plastic pellets collected from different industrial outlet has been carried out. Non-isothermal experiments using different heating rates of 5, 10, 20, 30, 40 and 50 °C min?1 have been performed from ambient to 700 °C in a thermo-balance with the objective of determining the kinetic parameters. The values of activation energy and frequency factor were found to be in the range of 107–322 kJ/mol, 85–331 kJ/mol, 140–375 kJ/mol and 3.49E+07–4.74E+22 min?1, 3.52E+06–2.88E+22min?1, 7.28E+13–1.17E+25 min?1 for rejected polypropylene, plastic film and plastic pellets, respectively, by Coats-Redfern and Ozawa methods including different models. Kissinger method, a model free analysis is also adopted to find the kinetic parameters. Activation energy and frequency factor were found to be 108 kJ/mol, 98 kJ/mol, 132 kJ/mol and 6.89E+03, 2.12E+02, 8.06E+05 min?1 for rejected polypropylene, plastic film and plastic pellets, respectively, by using the Kissinger method.  相似文献   

4.
Given that the kinetics of the thermal degreasing process of alumina ceramics based on stereolithography apparatus (SLA) has not been investigated, the mechanism of crack generation is still not fully revealed. This paper aims to elucidate the mechanism of crack generation in the degreasing process of alumina ceramics and to establish a kinetic model for alumina ceramics. Two sintering atmospheres, air, and argon, were selected for the degreasing tests at 100°C–700°C. The reaction products and mass changes of alumina ceramics were analyzed by TG-FTIR and TG-DSC (heating rates of 5, 10, and 15°C/min, respectively). Meanwhile, Boswell, Friedman, Ozawa, and DAEM model was used to describe the nonisothermal kinetics of the SLA alumina ceramic degreasing process. The results showed that setting the holding time to 400°C–425°C could promote the slow release of heat from the alumina ceramics. The thermal degreasing stage of the ceramic generated fewer cracks in the argon atmosphere than in the air atmosphere. The corresponding average activation energy values were 105.40 kJ/mol (Boswell model), 112.48 kJ/mol (Friedman model), 108.14 kJ/mol (Ozawa model), and 101.36 kJ/mol (DAEM model). The results of the study could provide an invaluable reference for the fabrication of defect-free SLA alumina ceramics.  相似文献   

5.
The thermal behavior of allyl PPO and its cured resin were investigated. In the allyl PPO curing process, the specific temperatures were Tgel = 173.6°C, Tcure = 225.4°C, and Ttreat = 237.7°C, and the activation energy (Ea) was 122 kJ/mol. The average number of PPO molecular units between two crosslinking points was about 20. In the degradation process of cured allyl PPO resin, the temperature at which mass loss equaled 1% was much higher than 300°C. The Ea for degradation was calculated as 125 kJ/mol, with a degradation fraction (α) in the range of 0.15–0.65, or 117 kJ/mol with an α of 0.10–0.90. The most probable mechanism function of decomposition of the cured allyl PPO resin was f(α) = 2(1 ? α)3/2 or g(α) = (1 ? α)?1/2 ? 1. The thermocompressed laminate of the allyl PPO blending with an additive resin (made from BDM and DP) exhibited the desired properties. ©2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 4111–4115, 2006  相似文献   

6.
In the present work, the kinetics of the epoxidation of soybean oil (SBO) by peroxyacetic acid (PAA) generated in situ in the presence of sulfuric acid as the catalyst was studied at various temperatures (45, 65 and 75 °C). It was found that epoxidation with almost complete conversion of unsaturated carbon and negligible oxirane cleavage can be attained by the in situ technique. The rate constant for epoxidation of SBO was found to be of the order of 10–6 mol–1s–1 and the activation energy of epoxidation is 43.11 kJ/mol. Some thermodynamic parameters: enthalpy, entropy and free activation energy of 40.63 kJ/mol, –208.80 J/mol and 102.88 kJ/mol, respectively, were obtained for the epoxidation of SBO. The kinetic and thermodynamic parameters of epoxidation obtained from this study indicate that an increase in the process temperature would increase the rate of epoxide formation. The epoxidation of corn oil and sunflower oil were also investigated under the same conditions. The results show that the reaction rate is in the order of soybean oil > corn oil > sunflower oil.  相似文献   

7.
The catalytic ammonia decomposition over iron and iron nitride, Fe4N, under the atmosphere of ammonia–hydrogen mixtures of different amounts of ammonia in the temperature range of 400–550 °C by means of thermogravimetry has been studied. A differential tubular reactor with mixing has been used. The ammonia concentration in the gas phase during all the process was analysed. The balance between the inlet and outlet ammonia quantity has been used to determine a degree of ammonia conversion and the values of decomposition reaction rate. The activation energy of ammonia decomposition reaction over Fe and Fe4N was found to be 68 and 143 kJ/mol, respectively.  相似文献   

8.
Instant controlled pressure drop (DIC) is known to enhance drying kinetics of cork granules. The overall kinetics is limited by the internal mass transfer. Hence, the drying kinetics carried out at 50, 60, and 70°C allows the effective diffusivity values to be determined as 2.63E-10 to 4.50E-10 and 4.34E-10 to 6.99E-10 m2 s?1 for untreated and DIC-treated cork granules, respectively. Because the temperature dependence is represented by an Arrhenius-type relationship, the activation energy values are determined as 25.29–48.12 kJ/mol and 32.43–43.98 kJ/mol for untreated and DIC-treated cork granules, respectively.  相似文献   

9.
This work was focused on the removal of phosphate ions using polypyrrole‐coated sawdust as a novel cost‐effective sorbent. The phosphate uptake followed the Langmuir sorption isotherm, and the sorption capacity at 20, 35, and 50°C was found to be 17.33, 23.41, and 30.39 mg/g, respectively; this indicated favorable sorption at higher temperatures. The kinetic uptake data were modeled with the Lagergren equation, first‐order and second‐order kinetic models, and the simple Elovich model. The results indicated that the Lagergren model best described the uptake data. The intraparticle diffusion coefficient, as determined for 250–211‐ and 630–600‐μm sorbent particles at 20°C, was found to be 287.3 × 10?2 and 228.3 × 10?2 mg g?1 min?1, respectively. The intraparticle diffusion was also confirmed with the Bangham equation. The sorption mean free energy, calculated with the Dubinin–Radushkevich equation, was found to be 10.98 kJ/mol, thus confirming an ion‐exchange regulated sorption process. The positive value of the enthalpy change (i.e., 4.23 kJ/mol) confirmed the endothermic nature of the sorption process. The negative values of the change in the standard free energy were indicative of the spontaneous nature of the sorption process. Finally, the activation energy of the sorption process for 250–212‐μm particles, determined with the Arrhenius equation, was found to be 41.68 J/mol. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

10.
The free radicals produced by γ irradiation in the polymer network are formed with ionizing radiation. The decay rates of radicals in the temperature range were used to estimate the diffusion coefficient of oxygen into polymeric spheres by an electron spin resonance (ESR) technique. The ESR results showed that the activation energy of the diffusion of oxygen into poly(methyl methacrylate) (PMMA) in the temperature range 10–40°C is 29.6 kJ/mol. There are two activation energies of the diffusion of oxygen into poly(vinyl acetate) (PVAc) in the temperature range 25–50°C. The activation energies below and above 35°C, which is the glass transition temperature of PVAc, were found to be 16.8 and 82.5 kJ/mol, respectively. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 71: 1203–1207, 1999  相似文献   

11.
Polytriazole polyethylene oxide-tetrahydrofuran (PTPET) elastomer was prepared based on the click-chemistry reaction between the  CCH and  N3 groups. The PTPET exhibits improved thermal stability with an activation energy of 204 kJ/mol compared with hydroxyl-terminated polyethylene oxide-tetrahydrofuran (PET) elastomer. TG-FTIR and pyrolysis-gas chromatography–mass spectrometry (Py-GC–MS) were used to analyze the pyrolysis mechanism and products of the PTPET elastomer. The pyrolytic mechanism of PTPET mainly involves breakage of the polyether chains. The main pyrolytic products were identified. The PTPET underwent for 180 days at 50°C and 60°C. The tensile properties, dynamic mechanical analysis (DMA) and TGA of the aged PTPET were investigated. The tensile stress and strain of the aged PTPET increased until 120 days (at 60°C) or 150 days (at 50°C), which could be attributed to the postcuring of the elastomer. The scission of the PTPET chains due to aging could dominate after aging for 120 days (at 60°C) or for 150 days (at 50°C), resulting in the reduction of mechanical properties.  相似文献   

12.
Reactions of vinyl trimethoxysilane grafting onto low-density polyethylene (LDPE) were investigated using Fourier transform infrared spectroscopy, differential scanning calorimetry, and thermal gravimetric analysis. The silane grafting reactions were induced by a fixed amount of dicumyl peroxide at 0.2 part of reagent per hundred parts with respect to LDPE. Fourier transform infrared data demonstrated that the extent of the silane grafting reaction was increased as the amount of silane used, the reaction time, or the reaction temperature was increased. The apparent activation energy of the silane grafting reaction was 9.7 kJ mol−1. Differential scanning calorimetry was used to follow the silane grafting reactions in situ at a heating rate of 20°C per minute. The silane grafting reaction was exothermic starting at about 150°C and ending at about 230°C, indicating a completion of the reaction in 4 min. The grafting reaction heat has linear relations to the amount of silane used. The grafting reaction heat of about 1 J/g of sample was generated during reaction per part of reagent per hundred parts of silane used. The reaction heat of silane grafting onto LDPE per mol of silane used was 14.5 kJ mol−1 silane, and the reaction heat of peroxide that reacted with LDPE was −12 kJ mol−1 peroxide. © 1998 John Wiley & Sons, Inc. J Appl Polym Sci 69: 255–261, 1998  相似文献   

13.
The synthesis of oleyl oleate with immobilized 1,3-specific lipase from Mucor miehei is presented in this article. Oleyl alcohol was esterified with oleic acid in the presence of a Mucor miehei lipase (lipozymeIM) to obtain oleyl oleate. The effects of various temperatures and various enzyme/substrate ratios have been investigated to determine optimal conditions for the esterification process. The highest conversion of oleic acid 86.9 % was obtained at 50°C. The optimal addition of lipase to substrates was determined to be 0.1 g per gram of reaction mixture. The esterification can be modeled successfully as a reverse second-order reaction. Thermodynamic properties of the reaction system at 50°C were also determined. Activation energy was 14.65 kJ/mol, entropy of activation –0.8 J/mol·K and free energy of activation was 98.568 kJ/mol.  相似文献   

14.
Rheological behavior of blends of poly (ethylene‐acrylic acid) (EAA) and ethylene vinyl acetate (EVA) copolymer have been carried out at various temperatures, namely, 100, 110, and 120°C, and different shear rates from 61.33 to 613.30 s?1 using a Monsanto Processability Tester. The melt viscosity of the blends shows synergism during processing. The activation energy of the blends is in the range 20.7–44.6 kJ/mol. Highest activation energy was observed for the blends containing 40–60% of EVA by weight. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 98: 1947–1954, 2005  相似文献   

15.
Polyhedral oligomeric silsesquioxanes epoxy resin (POSSER) was prepared from 3-glycidypropyl-trimethoxysilane (GTMS) and tetramethylammonium hydroxide (TMAH) by hydrolytic condensation. POSSER was characterized using Fourier-transformed infrared spectroscopy (FTIR), 1H-NMR, and liquid chromagraphy/mass spectrometry (LC/MS). The epoxy value of POSSER is 0.50 mol/100 g. The LC/MS analysis indicated that T10 is the majority and contain some amount of T8, besides, a trace T9 also exists. The curing kinetics of POSSER with 4,4′-diaminodipheny sulfone (DDS) as a curing agent was investigated by means of differential scanning calorimetry (DSC). The curing reaction order n is 0.8841 and the activation energy Ea is 61.06 kJ/mol from dynamic DSC analysis. Thermal stability and kinetics of thermal degradation were also studied by thermal gravimetric analysis (TGA). TGA results indicated that the temperature of POSSE/DDS system 5% weight loss is approximately 377.0°C, which is higher by 12.6°C than that of pure POSSER, and the primary degradation reaction (300–465°C) followed first order kinetics; the activation energy of degradation reaction is 75.81 kJ/mol.  相似文献   

16.
Curing of a novolac‐type phenolic resin was studied by DSC. The kinetic analysis was performed by means of the dynamic Ozawa method at heating rates of 5, 10, 15, and 20°C/min. This analysis was used to determine the kinetic parameters of the curing process. The activation energy was found to be 144 kJ/mol. It was found that the Ozawa exponent values decreased with increasing reaction temperature from 3.5 to 1, suggesting a change in the reaction mechanism from microgel growth to diffusion‐controlled reaction. The reaction rate constant was found to range from 123.0 to 33.6 (°C/min)n. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 90: 1678–1682, 2003  相似文献   

17.
Mullite formation from kaolinite was studied by means of high-temperature in situ powder neutron diffraction by heating from room temperature up to 1370 °C. Neutron diffractometry under this non-isothermal conditions is suitable for studying high-temperature reaction kinetics and to identify short-lived species which otherwise might escape detection. Data collected from dynamic techniques (neutron diffraction, DTA, TGA and constant-heating rate sintering) were consistent with data gathered in static mode (conventional X-ray diffraction and TEM). The full process occurs in successive stages: (a) kaolinite dehydroxylation yielding metakaolinite in the ∼400–650 °C temperature range, (b) nucleation of mullite in the temperature range ∼980–992 to ∼1121 °C (primary mullite) side by side with a crystalline cubic phase (Si-Al spinel) detected in the ∼983–1030 °C temperature interval; (c) growth of mullite crystals from ∼1136 °C, (d) high (or β) cristobalite crystallization at T > ∼1200 °C and (e) secondary mullite crystallization at T > ∼1300 °C. The calculated activation energy for the kaolinite dehydration was 115 kJ/mol; for the mullite nucleation was 278 kJ/mol and for the growth of mullite process was 87 kJ/mol; finally for cristobalite nucleation the calculated apparent activation energy was 481 kJ/mol.  相似文献   

18.
Polycrystalline ZrO2-8 mol%Y2O3 was investigated by combining several experimental techniques on identical materials sintered out of the same high purity powder. The mechanical loss spectrum (damping and elastic modulus) was measured in a large frequency and temperature range (10−2Hz–1.5kHz; −150 to 1400°C). Damping due to point defect relaxation at low temperature and to viscoelastic relaxation at high temperature was revealed. The creep resistance was investigated with four-point bending tests (stress and temperature ranges: 20–75 MPa, 1100–1290°C), indicating Nabarro-Herring creep as the main rate-controlling mechanism. Both viscoelastic deformation and creep seem to be controlled by cation diffusion. Measurements of the 96Zr tracer diffusivity by secondary ion mass spectrometry at 1125–1460°C yielded an activation enthalpy of 460 kJ/mol. Close values were obtained for creep (440 kJ/mol) and viscoelastic relaxation (530 kJ/mol). Finally, the ionic DC-conductivity of these electrolytes was measured with high accuracy in the range 300–1250°C.  相似文献   

19.
Permeabilities and diffusivities of methyl chloride and benzene vapors at low activities in FEP Teflon membranes were measured in a continuous-flow permeation cell, at temperatures ranging from 47°C to 150°C. In all cases investigated, the permeabilities and diffusivities were independent of the penetrant partial pressure, and the permeation process was well described by a Henry's law sorption–Fickian diffusion model. The activation energies for permeation and diffusion and the sorption enthalpy were respectively 34.8 kJ/mol, 50.1 kJ/mol, and ?15.3 kJ/mol for methyl chloride and 49.5 kJ/mol, 69.1 kJ/mol, and ?19.6 kJ/mol for benzene. The diffusional activation energies for these two substances and other low molecular weight hydrocarbons correlate reasonably well with the Lennard–Jones collision diameters of the penetrant molecules. The solubilities correlate approximately with the penetrant boiling points, but the highly polar nature of methyl chloride and the aromaticity of benzene lead to deviations between the solubilities of these substances and those of nonpolar aliphatics with the same tendency to condense from the vapor phase.  相似文献   

20.
The hydrolysis of polyethylene terephthalate (PET) pellets in nitric acid was investigated to determine the kinetic parameters. Experiments were conducted with cylindrical shaped pellets in 13M nitric acid at 80, 90, and 100°C respectively. Also, an experiment was conducted with a waste PET bottle sheet in 9.5M nitric acid at 100°C. The kinetics of the reaction was explained by the shrinking core model with surface chemical reaction as rate controlling step and accounting for surface area reduction due to the deposition of the product terephathalic acid (TPA) on the reaction surface. The activation energy for the reaction was found to be 135 kJ/mol. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 87:1781–1783, 2003  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