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 15 毫秒
1.
The environmentally friendly high‐energy density salt (TRTR)(PA) (TRTR = 3, 4′‐bis‐1, 2,4‐1H‐triazole, PA = 2, 4,6‐trinitrophenol, picric acid) was synthesized and characterized. The X‐ray single crystal diffraction results illustrate that the structure of title salt belongs to the monoclinic system, space group P21/c. Many parallel relationships exist in the molecule, as well as a strong intramolecular π–π stacking interaction. The DSC result shows only one exothermal decomposition step at 229.1 °C. The TG‐DTG curve demonstrates a 75.9 % mass loss from 180 °C to 300 °C at a rate of 3.01 % · K–1. Experimental data show that the combustion heat approximately equals to TNT (–15.22 MJ · kg–1) and the enthalpy of formation is +332.2 kJ · mol–1. Non–isothermal kinetic and thermodynamic parameters were obtained by two methods (Kissinger and Ozawa). Detonation pressure and velocity were calculated to be 23.4 GPa and 7.32 km · s–1, respectively. Additionally, the sensitivities towards impact and friction were assessed with relevant standard methods.  相似文献   

2.
Determination of Temperature Dependent Partial Pressures in Closed Systems – a New Method. The Heat of Formation for PtI2(s) A new method to determine temperature dependent partial pressures of gaseous species in equilibria with condensed phases in closed systems (silica ampoules) at temperatures up to 1000 °C and pressures pi 0.01 < pi < 10 bar is presented. It is based on the determination of the change of mass in the gasphase caused by solid-gas transition at higher temperatures of substances which are deposited at one end of the ampoule. The results of the measurements give informations about reaction mechanisms, enthalpies and entropies. The reliability of the method is demonstrated at the example of the system Pt/I2. The heat of formation and the entropy of PtI2(s) (δBH°(PtI2(s), 298) = –51.4 kJ · mol–1, S°(PtI2(s), 298) = 119.3 J · K–1 · mol–1) are computed from experimental results. The heat of thermal decomposition of PtI2(s) was reconsidered by Knudsen Mass Spectrometry.  相似文献   

3.
The mineral sabugalite (HAl)0.5[(UO2)2(PO4)]2⋅8H2O, has been studied using a combination of energy dispersive X-ray analysis, X-ray diffraction, dynamic and controlled rate thermal analysis techniques. X-ray diffraction shows that the starting material in the thermal decomposition is sabugalite and the product of the thermal treatment is a mixture of aluminium and uranyl phosphates. Four mass loss steps are observed for the dehydration of sabugalite at 48°C (temperature range 39 to 59°C), 84°C (temperature range 59 to 109°C), 127°C (temperature range 109 to 165°C) and around 270°C (temperature range 175 to 525°C) with mass losses of 2.8, 6.5, 2.3 and 4.4%, respectively, making a total mass loss of water of 16.0%. In the CRTA experiment mass loss stages were found at 60, 97, 140 and 270°C which correspond to four dehydration steps involving the loss of 2, 6, 6 and 2 moles of water. These mass losses result in the formation of four phases namely meta(I)sabugalite, meta(II)sabugalite, meta(III)sabugalite and finally uranyl phosphate and alumina phosphates. The use of a combination of dynamic and controlled rate thermal analysis techniques enabled a definitive study of the thermal decomposition of sabugalite. While the temperature ranges and the mass losses vary due to the different experimental conditions, the results of the CRTA analysis should be considered as standard data due to the quasi-equilibrium nature of the thermal decomposition process. The online version of the original article can be found at  相似文献   

4.
Dental composites can be improved by heat treatment, as a possible way to increase mechanical properties due to additional cure (post-cure). Direct dental composites are essentially similar to the indirect ones, supposing they have the same indication. Therefore, to establish a heat treatment protocol for direct composites, using as indirect (photoactivated by continuous and pulse-delay techniques), a characterization (TG/DTG and DSC) is necessary to determine parameters, such as mass loss by thermal decomposition, heat of reaction and glass transition temperature (T g). By the results of this study, a heat treatment could be carried out above 160 °C (above T g, and even higher than the endset exothermic event) and under 180 °C (temperature of significant initial mass loss).  相似文献   

5.
Experimental evidence is presented for a unimolecular gas-phase Wagner-Meerwein shift in neopentyl chloride pyrolysis. In the decomposition of α,α-neopentyl chloride-d2 at 445°C, maximally inhibited by cyclohexene, the initial products were isotopically pure 2-methyl-1-butene-d2 and 2-methyl-2-butene-d1. Rearrangement, accompanied by loss of either α- or γ-hydrogen in the formation of hydrogen chloride, is consistent with an incipient ion-pair type of transition state. The cyclohexene maximally inhibited pyrolysis of neopentyl chloride was also examined over the temperature range 424–478°C and Arrhenius parameters of E, 258.7 kJ/mole and logA/sec?1, 13.78, were determined.  相似文献   

6.
Evaporation of Plasticizer from NEPE Type Propellant   总被引:1,自引:0,他引:1  
Using the method of dynamic thermogravimetry and differential scanning calorimetry in the heating rate range 0.46–10.0 deg–1 min–1, evaporation of the plasticizer from propellant samples of the NEPE type was investigated. The experiments were carried out in an open system in a flow of pure argon at atmospheric pressure. Nitroglycerin is the main mass fraction of the plasticizer. The activation energy E of the gross evaporation–diffusion process is determined by various methods. Heat of evaporation of the plasticizer ΔHv is estimated. It is shown that in the early stage of evaporation the values of E and ΔHv practically coincide. At a temperature of 298.15 K ΔHv = 89 ± 4 kJ mol–1, which is in satisfactory agreement with the literature data for heat of evaporation of pure nitroglycerin. With any way of preventing free removal of the plasticizer from the surface of the samples on the DSC thermograms successive exothermic peaks of the thermal decomposition of the plasticizer and the octogen are observed, which are not realized in the open system for the indicated heating rates at T < 190°C.  相似文献   

7.
Simultaneous thermogravimetry (TG) and differential thermal analysis (DTA) techniques were used for the characterization the thermal degradation of loratadine, ethyl-4-(8-chloro-5,6-dihydro-11H-benzo[5,6]cyclohepta[1,2-b]pyridin-11-ylidine)-1-piperidinecarboxylate. TG analysis revealed that the thermal decomposition occurs in one step in the 200–400°C range in nitrogen atmosphere. DTA and DSC curves showed that loratadine melts before the decomposition and the decomposition products are volatile in nitrogen. In air the decomposition follows very similar profile up to 300°C, but two exothermic events are observed in the 170–680°C temperature range. Flynn–Wall–Ozawa method was used for the solid-state kinetic analysis of loratadine thermal decomposition. The calculated activation energy (E a) was 91±1 kJ mol–1 for α between 0.02 and 0.2, where the mass loss is mainly due to the decomposition than to the evaporation of the decomposition products.  相似文献   

8.
In this work, a kinetic study on the thermal degradation of carbon fibre reinforced epoxy is presented. The degradation is investigated by means of dynamic thermogravimetric analysis (TG) in air and inert atmosphere at heating rates from 0.5 to 20°C min−1 . Curves obtained by TG in air are quite different from those obtained in nitrogen. A three-step loss is observed during dynamic TG in air while mass loss proceeded as a two step process in nitrogen at fast heating rate. To elucidate this difference, a kinetic analysis is carried on. A kinetic model described by the Kissinger method or by the Ozawa method gives the kinetic parameters of the composite decomposition. Apparent activation energy calculated by Kissinger method in oxidative atmosphere for each step is between 40–50 kJ mol−1 upper than E a calculated in inert atmosphere. The thermo-oxidative degradation illustrated by Ozawa method shows a stable apparent activation energy (E a ≈130 kJ mol−1 ) even though the thermal degradation in nitrogen flow presents a maximum E a for 15% mass loss (E a ≈60 kJ mol−1 ). This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

9.
The elimination kinetics of the title compounds have been examined over the temperature range of 270–320°C and pressure range of 19–117 torr. The reactions, carried out in seasoned vessels, with the free-radical suppressor toluene always present, are homogeneous, unimolecular, and follow a first-order rate law. The products of 2-hydroxy-2-methylbutyric acid are 2-butanone, CO, and H2O; while of 2-ethyl-2-hydroxybutyric acid are 3-pentanone, CO, and H2O. The rate coefficient is expressed by the following Arrhenius equation: for 2-hydroxy-2-methylbutyric acid, log k1(s?1 = (12.87 ± 0.19) ? (171.2 ± 2.1) kJ mol?1 (2.303 RT)?1; and for 2-ethyl 2-hydroxybutyric acid, log k1s?1) = (12.13 ± 0.34) ? (159.4 ± 3.7) kJ mol?1 (2.303 RT)?1. Augmentation of alkyl bulkiness at the 2-position of the 2-hydroxycarboxylic acids showed an increase in the rate of dehydration. The electron release of alkyl groups, rather than steric acceleration, appears to enhance the pyrolysis decomposition of these substrates. These reactions are believed to proceed through a semi-polar five-membered cyclic transition type of mechanism. © 1995 John Wiley & Sons, Inc.  相似文献   

10.
The pyrolysis of hydrated bis(pyrazinecarboxylate)copper(II) under an argon atmosphere proceeds via the loss of the water molecules at 84–95°C, ΔH=40.4 kJ (mol H2O)?1 followed by the thermal decomposition of the complex at 284–325°C, ΔH=97.0 kJ·mol?1, yielding 0.72 mole of pyrazine, 0.28 mole of bipyrazine, and 2 mole of CO2 per mole of complex.  相似文献   

11.
Low temperature heat capacities of N-(p-methylphenyl)-N'-(2-pyridyl)urea were determined by adiabatic calorimetry method in the temperature range from 80 to 370 K. It was found that there was not any heat anomaly in this temperature region. Based on the experimental data, some thermodynamic function results were obtained. Thermal stability and decomposition characteristics analysis of N-(p-methylphenyl)-N'-(2-pyridyl)urea were carried out by DSC and TG. The results indicated that N-(p-methylphenyl)-N'-(2-pyridyl)urea started to melt at ca. 426 K (153°C) and the melting peak located at 447.01 K (173.86°C). The melting enthalpy was 204.445 kJ mol-1 (899.6 J g-1). The decomposition peak of N-(p-methylphenyl)-N'-(2-pyridyl)urea was found at 499.26 K (226.11°C) from DSC curve. This result was similar with that from TG and DTG experiment, in which the mass loss peak was determined as 500.4 K (227.2°C). This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

12.
The application of hypergolic ionic liquids as propellant fuels is a newly emerging area in the fields of chemistry and propulsion science. Herein, a new class of [imidazolyl?amine?BH2]+‐cation‐based ionic liquids, which included fuel‐rich anions, such as dicyanamide (N(CN)2?) and cyanoborohydride (BH3CN?) anions, were synthesized and characterized. As expected, all of the ionic liquids exhibited spontaneous combustion upon contact with the oxidizer 100 % HNO3. The densities of these ionic liquids varied from 0.99–1.12 g cm?3, and the heats of formation, predicted based on Gaussian 09 calculations, were between ?707.7 and 241.8 kJ mol?1. Among them, the salt of compound 5 , that is, (1‐allyl‐1H‐imidazole‐3‐yl)?(trimethylamine)?dihydroboronium dicyanamide, exhibited the lowest viscosity (168 MPa s), good thermal properties (TgTd>130 °C), and the shortest ignition‐delay time (18 ms) with 100 % HNO3. These ionic fuels, as “green” replacements for toxic hydrazine‐derivatives, may have potential applications as bipropellant formulations.  相似文献   

13.
Thermal behavior of KClO4/Mg pyrotechnic mixtures heated in air was investigated by thermal analysis. Effects of oxygen balance and heating rates on the TG?CDSC curves of mixtures were examined. Results showed that DSC curves of the mixtures had two exothermic processes when heated from room temperature to 700?°C, and TG curve exhibited a slight mass gain followed by a two-stage mass fall and then a significant mass increase. The exothermic peak at lower temperature and higher temperature corresponded to the ignition process and afterburning process, respectively. Under the heating rate of 10?°C?min?1, the peak temperatures for ignition and afterburning process of stoichiometric KClO4/Mg (58.8/41.2) was 543 and 615?°C, respectively. When Mg content increased to 50%, the peak ignition temperature decreased to 530?°C, but the second exothermic peak changed little. Reaction kinetics of the two exothermic processes for the stoichiometric mixture was calculated using Kissinger method. Apparent activation energies for ignition and afterburning process were 153.6 and 289.5?kJ?mol?1, respectively. A five-step reaction pathway was proposed for the ignition process in air, and activation energies for each step were also calculated. These results should provide reference for formula design and safety storage of KClO4/Mg-containing pyrotechnics.  相似文献   

14.
Current work at Lawrence Livermore National Laboratory (LLNL) includes both understanding properties of old explosives and measuring properties of new ones. The necessity to know and understand the properties of energetic materials is driven by the need to improve performance and enhance stability to various stimuli, such as thermal, friction and impact insult. This paper will concentrate on the physical properties of RX-55-AE-5, which is formulated from heterocyclic explosive, 2,6-diamino-3,5-dinitropyrazine-1-oxide, LLM-105, and 2.5% Viton A. Differential scanning calorimetry (DSC) was used to measure a specific heat capacity, C p, of≈0.950 J g−1 °C−1, and a thermal conductivity, κ, of≈0.475 W m−1 °C−1. The LLNL kinetics modeling code Kinetics05 and the Advanced Kinetics and Technology Solutions (AKTS) code thermokinetics were both used to calculate Arrhenius kinetics for decomposition of LLM-105. Both obtained an activation energy barrier E≈180 kJ mol−1 for mass loss in an open pan. Thermal mechanical analysis, TMA, was used to measure the coefficient of thermal expansion (CTE). The CTE for this formulation was calculated to be ≈61 μm m−1 °C−1. Impact, spark, friction are also reported.  相似文献   

15.
The kinetics of methacryloyl fluoride (MAF) homopolymerization was investigated in methyl ethyl ketone (MEK) with azobis(isobutyronitrile) as initiator. The rate of polymerization (Rp) followed the expression Rp = k[AIBN]0.55[MAF]1.18. The overall activation energy was calculated as 74.4 kJ/mol. The relative reactivity ratios of MAF(M2) copolymerization with styrene (r1 = 0.083, r2 = 0.14), and methyl methacrylate (r1 = 0.48, r2 = 0.81) in methyl ethyl ketone were obtained. Application of the Qe scheme (in styrene copolymerization) led to Q = 2.22 and e = 1.31. The glass transition temperature (Tg) of poly(MAF) was 90°C by thermomechanical analysis. Thermogravimetry of poly(MAF) showed a 10% weight loss of 228°C in air.  相似文献   

16.
The rate coefficients for the gas-phase pyrolyses of a series of structurally related secondary acetates have been measured in a static system over the temperature range of 289.1–359.5°C and the pressure range 50.0–203.0 torr. The temperature dependence of the rate coefficients is given by the following Arrhenius equations: for 3-hexyl acetate, log k1 (s?) = (12.12 ± 0.33) ? (176.1 ± 3.9)kJ/mol/2.203RT; for 5-methyl-3-hexyl acetate, log k1 (s?) = (13.17 ± 0.20) ? (186.2 ± 2.3)kJ/mol/2.303RT; and for 5,5-dimethyl-3-hexyl acetate, log k1 (s?) = (12.70 ± 0.19) ? (177.4 ± 2.2)kJ/mol/2.303RT. The direction of elimination of these esters has shown from the invariability of olefin distributions at different temperatures and percentages of decomposition that steric hindrance is a determining factor in the eclipsed cis conformation. Moreover, a more detailed analysis indicates that the greater the alkyl–alkyl interaction, the less favored the elimination tends to be. Otherwise, an increase of alkyl–hydrogen interaction caused steric acceleration to be the determining factor.  相似文献   

17.
The surface segregation of In and S from a dilute Cu(In,S) ternary alloy were measured using Auger electron spectroscopy coupled with a linear programmed heater. The alloy was linearly heated and cooled at constant rates. Segregation data of a linear heat run showed surface segregation of In that reached a maximum surface coverage of 25% followed by S, which reached a coverage of 30%. It was found that after In had reached a maximum surface coverage, it started to desegregate as soon as the S enriched the surface until In was completely replaced by S. The segregation parameters, namely, the pre‐exponential factor (D0), activation energy (Q), segregation energy (ΔG?) and interaction energy (Ω) were extracted from the measured segregation data for both In and S segregation in Cu by simulating the measured segregation data with a theoretical segregation model (modified Darken model). The segregation parameters obtained for In segregation in Cu are D0 = 1.8 ± 0.5 × 10?5 m2 s?1, Q = 184.3 ± 1.0 kJ.mol?1, ΔG? = ?61.4 ± 1.4 kJ.mol‐1, ΩCu?In = 3.0 ± 0.4 kJ.mol?1; for S segregation in Cu the parameters are D0 = 8.9 ± 0.5 × 10?3 m2 s?1, Q = 212.8 ± 3.0 kJ.mol?1, ΔG? = ?120.0 ± 3.5 kJ.mol?1, ΩCu?S = 23.0 ± 2.0 kJ mol?1 and the In and S interaction parameter is ΩIn?S = ?4.0 ± 0.5 kJ.mol?1. The initial parameters used for the Darken calculations were extracted from fits performed with the Fick's and Guttmann model. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

18.
Reduced graphene oxide/metal ferrite (rGO/MFe2O4, M = Cu, Co, Ni) nanohybrids are successfully prepared through a simple, one-step hydrothermal method. The rGO/MFe2O4 hybrids are characterized by XRD, TEM. The rGO/MFe2O4 nanohybrids demonstrate amazing catalytic activity on thermal decomposition of ammonium perchlorate (AP) based propellants. DSC results indicate that the high-temperature decomposition (HTD) temperature of propellants added with rGO/MFe2O4 nanohybrids (3 wt%), could decrease from 325.9 °C to 259.9 °C, 268.8 °C, 271.9 °C, 306.9 °C, respectively. The HTD activation energy on a conversion degree (α) range from 0.05 to 0.95 of propellant samples were investigated by two model-free methods Flynne–Walle–Ozawa (FWO) and Kissinger–Akahira–Sunose (KAS). The results showed that both methods had similar values of Ea, and they match well with each other. A strong dependence of Ea on α revealed a complex decomposition process. The model-fitting analysis suggested the HTD process of propellant samples with or without catalysts both followed Mampel (First order) reaction model.  相似文献   

19.
This article studies the thermokinetics and safety parameters of cumene hydroperoxide (CHP) manufactured in the first oxidation tower. Vent sizing package 2 (VSP2), an adiabatic calorimeter, was employed to determine reaction kinetics, the exothermic onset temperature (T 0), reaction order (n), ignition runaway temperature (T C, I), etc. The n value and activation energy (E a) of 15?mass% CHP were calculated to be 0.5 and 120.2?kJ?mol?1, respectively. The heat generation rate (Q g) of 15?mass% CHP compared with hS (cooling rate)?=?6.7?J?min?1?K?1 of heat balance, the T S,E and the critical extinction temperature (T C, E) under 110?°C of ambient temperature (T a) were calculated 111 and 207?°C, respectively. The Q g of 15?mass% CHP compared with hS?=?0.3?J?min?1?K?1 of heat balance was applied to determine the T C, I that was evaluated to be 116?°C. This article describes the best operating conditions when handling CHP, starting from the first oxidation tower.  相似文献   

20.
Decomposition has been studied in the chemistry of perfluoropolyethers (PFPE), thus far no molecular structure information is reported. TG-MS is a tool to follow the off gassing of decomposition for clues. We selected two PFPEs that have different properties: Krytox® XHT-1000 and Fomblin Z60 heating to normal decomposition and catalytic decomposition in the presence of alumina powder. Comparing the decomposition fragment intensities, the molecular structure of the branched Krytox® XHT-1000 oil is more stable than the blocky Fomblin Z60. We see aluminum-containing fluorine fragments in the rapid decomposition of oils in contact with alumina powder. It has been suggested the formation of Al(O6?n F n ), where n = 1, 2, and 3, in which the fluorine atoms are selectively associated with aluminum atom. The major decomposition products are small and large fragments of fluorocarbons and perfluoroalkoxy. In the absence of alumina powder, Krytox XHT-1000 shows only a loss of 13 mass/% after several hours at 330 °C, whereas in the presence of 1 mass/% alumina powder the oil has rapidly decomposed to 67 mass/% of its original mass within 15 min. Fomblin Z60, a product might not be designed for high temperature, exposing to the same conditions at 330 °C for several hours and shows a loss of 98 mass/% alone, but in the presence of 1 mass/% alumina powder shows a loss of 98 mass/% in 3.6 min. When 3 mass/% of two new developmental additives were added to the both oils, the catalytic decomposition in the presence of 1 mass/% alumina powder was significantly reduced in Krytox® XHT-1000, showing only a loss of 23 mass/% in 4 h, but nearly all weight for Z60 in 60 min. In the oil grades that contain the new additives, we see the fragments of Al–O–S, and F–Al–O–S. The sulfur-containing compound has been reported ionically bonded to oxide in a tripod configuration of alumina surface, which shields the formation of Al–F.  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