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 46 毫秒
1.
Effective radium content and radon exhalation rates in soil samples have been measured by ??Sealed Can Technique?? using LR-115 type II plastic track detectors. The soil samples were collected from Farrukhabad city of Utter Pradesh, India. The values of effective radium content were found to vary from 5.39 to 34.56?Bq?kg?1 with an average value of 16.58?Bq?kg?1 and a standard deviation of 7.16. The mass and surface exhalation rate has been found to vary from 0.41?×?10?6 to 2.64?×?10?6?Bq?kg?1?d?1 and 1.41?×?10?6 to 9.10?×?10?6?Bq?m?2?d?1, respectively. All the values of radium content in soil samples of study area were found to be quite lower than the permissible value of 370?Bq?kg?1 recommended by Organization for Economic Cooperation and Development.  相似文献   

2.
Cadmium thiourea reinickate undergoes two-stage thermal decomposition on heating. The DTG peak temperatures are 291 and 469°C and the corresponding DTA temperatures are 255 and 490°C. The kinetic parameters for the first stage decomposition are E* ≈ 120kJ mole?1; Z ≈ 1.2 × 108 cm3 mole?1 sec?1 and ΔS* ≈ ?95 J mole?1 K?1. For the second stage, E* ≈ 133 kJ mole?1; Z ≈ 6.1 × 105 cm?1 mole?1 sec?1 and ΔS* ≈ ?142 J mole?1 K?1.  相似文献   

3.
Poly(α,α-dimethyl-β-propiolactone) (PPL), known as poly(pivalolactone), has been studied by differential scanning calorimetry (DSC) and small-angle X-ray diffraction (SAXR). DSC measurements indicate the presence of two melting endotherms. Peak 1 and Peak 2, the latter at lower temperatures. Peak 1 is relatively unaffected by the crystallization temperature and its relative intensity decreases with heating rate. Peak 2 is greatly influenced by the crystallization temperature of the sample and its relative intensity increases with heating rate. Peak 2 is associated with the true melting of the PPL samples and Peak 1 with a recrystallization process during the heating cycle. SAXR long periods increase with crystallization and annealing temperatures. Similar increases in density, in melting temperature, in lamella thickness, and in degree of crystallinity have been observed. These results lead to a thermodynamic melting temperature of 268 ± 3 for PPL, and to interfacial free energies of, respectively. 13 × 10?7 J cm?2 and (43 ± 4) × 10?7 J cm?2 for the lateral surface and the fold surface of the PPL crystal.  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.
Novel films consisting of multi-walled carbon nanotubes (MWCNTs) were fabricated by means of chemical vapor deposition with decomposition of either acetonitrile (ACN) or benzene (BZ) using ferrocene as catalyst. The electrochemical responses of MWCNT-based films towards the ferrocyanide/ferricyanide, [Fe(CN)6]3?/4? redox couple were probed by means of cyclic voltammetry and electrochemical impedance spectroscopy at 25.0?±?0.5?°C. Both MWCNT-based films exhibit Nernstian response towards [Fe(CN)6]3?/4? with some slight kinetic differences. Namely, heterogeneous electron transfer rate constants lying in ranges of 2.69?×?10?2?C1.7?×?10?3 and 9.0?×?10?3?C2.6?×?10?3?cm·s?1 were obtained at v?=?0.05?V·s?1 for MWCNTACN and MWCNTBZ, respectively. The detection limit of MWCNTACN, estimated to be about 4.70?×?10?7?mol·L?1 at v?=?0.05?V·s?1, tends to become slightly poorer with the increase of the scan rate, namely at v?=?0.10?V·s?1 the detection limit of 1.70?×?10?6?mol·L?1 was determined. Slightly poorer response ability was exhibited by MWCNTBZ; specifically the detection limits of 1.57?×?10?6 and 4.35?×?10?6?mol·L?1 were determined at v?=?0.05 and v?=?0.10?V·s?1, respectively. The sensitivities of MWCNTACN and MWCNTBZ towards [Fe(CN)6]3?/4? were determined as 1.60?×?10?7 and 1.51?×?10?7?A·L·mol?1·cm?2, respectively. The excellent electrochemical performance of MWCNTACN is attributed to the presence of incorporated nitrogen in the nanotube??s structure.  相似文献   

6.
A radiotracer study was carried out in a trickle bed reactor (TBR) independently filled with two different types of packing i.e., hydrophobic and hydrophilic. The study was aimed at to estimate liquid holdup and investigate the dispersion characteristics of liquid phase with both types of packing at different operating conditions. Water and H2 gas were used as aqueous and gas phase, respectively. The liquid and gas flow rates used ranged from 0.83?×?10?7?C16.67?×?10?7?m3/s and 0?C3.33?×?10?4?m3 (std)/s, respectively. Residence time distribution (RTD) of liquid phase was measured using 82Br as radiotracer and about 10?MBq activity was used in each run. Mean residence time (MRT) and holdup of liquid phase were estimated from the measured RTD data. An axial dispersion with exchange model was used to simulate the measured RTD curves and model parameters (Peclet number and MRT) were obtained. At higher liquid flow rates, the TBR behaves as a plug flow reactor, whereas at lower liquid flow rates, the flow was found to be highly dispersed. The results of investigation indicated that the dispersion of liquid phase is higher in case of hydrophobic packing, whereas holdup is higher in case of hydrophilic packing.  相似文献   

7.
The decomposition mechanism of 5-Aza-2??-deoxycytidine has been studied by the use of computational techniques. Optimized structures for all of the stationary points in the gas phase were investigated at B3LYP/6-31+G(d,p) level of theory. Single-point energies were determined employing the ab initio MP2 method in conjunction with the 6-311++G(d,p) basis set. Five possible pathways, paths 1?C5, were evaluated. In each pathway, the direct (A-paths 1?C5) and water-assisted (B-paths 1?C5) processes were considered. Meanwhile, the local microhydration model with the direct participation of three water molecules around the reaction centers was adopted to mimic the system for the water-assisted decomposition mechanisms above, where one water molecule is the nucleophilic reactant and the other two are the auxiliary molecules located on each side of the nucleophilic water. The results in the gas phase exhibit that the energy barriers of the water-assisted pathways based on the local microhydration model decrease dramatically by about 15?C20?kcal/mol as compared with those of the direct pathways because of the contribution of the auxiliary water molecules. In addition, bulk solvent effects of water were determined by means of the self-consistent reaction field based on the conductor-like polarized continuum model and Monte Carlo simulation with free energy perturbation (MC-FEP) technique, respectively. Our computational results indicate that B-path 3 in the decomposition reaction of 5-azadC is the most favorable, where the calculated rate constant (1.68?×?10?3?min?1) using the MC-FEP method is within the range of the experimentally determined values [(5.89?±?0.54)?×?10?3?min?1 by UV and (1.46?±?0.08)?×?10?3?min?1 by NMR].  相似文献   

8.
Thermal and thermochemical investigations of natural hydroxyl-bearing copper sulfate Cu3SO4(OH)4??antlerite have been carried out. The stages of its thermal decomposition have been studied employing the Fourier-transform IR spectroscopy. The enthalpy of formation of antlerite from the elements ??f H m o (298.15?K)?=?(?1750?±?10)?kJ?mol?1 has been determined by the method of oxide melt solution calorimetry. Using value of S m o (298.15?K), equal to (263.46?±?0.47)?J?K?1?mol?1, obtained earlier by the method of adiabatic calorimetry, the Gibbs energy value of ??f G m o (298.15?K)?=?(?1467?±?10)?kJ?mol?1 has been calculated.  相似文献   

9.
The ultraviolet (λ = 2537 A.) photolysis of a degassed mixed phenyl and methyl polysiloxane liquid is examined in terms of gas and crosslinking yields. Results are compared to the published values obtained by ionizing irradiation of this type of molecule. It is shown that ultraviolet radiation is less efficient by two orders of magnitude in producing decomposition (i.e., gaseous products) than is ionizing radiation. The comparisons for crosslinking efficiencies are less certain, but the yields seem to have much more similar values in this case based on a spectroscopic estimation of crosslinking (i.e., analysis for substituted phenylcyclohexadiene formation). The gas quantum yields were ?H2 = 2.6 × 10?5, ?CH4 = 0.63 × 10?5, ?C2H6 ≈ 0.12 × 10?5, and ?C2H2 ≈ 0.06 × 10?5.  相似文献   

10.
The decomposition of polybromostyryl carbanions (PBS?), obtained by anionic polymerization of 4-bromostyrene in tetrahydrofuran (THF), was investigated in the dark in a temperature range of ?6–?21°C. It was accompanied by the evolution of bromine anions and by the formation of polymeric allylic carbanions (λmax = 575 nm; εmax = 6800 eq?1·L·cm?1). The reaction mechanism was elucidated. The rate constant of the unimolecular rate-determining step of the process was 1.3 × 10?5 s?1 and 9.7 × 10?5 s?1 at ?21 and ?6°C, respectively. Its apparent energy of activation Eapp = 18.38 Kcal/mol. The polybromostyrenes with allylic carbanions at their ends may decompose further. Their “dark” decomposition yielded 1,3-butadiene-1,3-diphenyl-macromers. The mechanisms of decomposition of the PBS? carbanions and the dark decomposition of the polybromostyryl allylic carbanions are analogous. The rate constant of the latter process was 2.5 × 10?6 s?1 at ?6°C. The anionic polymerization of prepared macromers can be initiated in THF at ?78°C by α-methylstyryl carbanions, which do not react, however, with PBS? carbanions. “Comblike” polymacromers were prepared in which each branch had a molecular weight of about 50,000. The overall molecular weight of the polymacromer was estimated to be about 1 × 106. It has been assumed that the 2–1 mode of addition to the diene group of the macromer is predominant during its polymerization. The 3–4 mode of addition followed by proton shift represents the termination step. The 4–3 mode of addition was ruled out on the basis of spectroscopic evidence.  相似文献   

11.
A multivariate calibration procedure based on singular value decomposition (SVD) and the Ho-Kashyap algorithm is used for the tensammetric determination of the cationic detergents Hyamine 1622, benzalkonium chloride (BACl), N-cetyl-N,N,N-trimethylammonium bromide (CTABr) and mixtures of CTABr and BACl. The sensitivity and accuracy depend strongly on the nature of the detergent. Acceptable accuracy is obtained with a two-step calculation procedure in which calibration constants for the total concentration range of interest are used to guide the choice of a more specific set of calibration constants which are valid for a much smaller concentration span. For Hyamine 1622, concentrations in the range 5 × 10?6?2 × 10?4 M could be determined with an accuracy of ± 10?6 M. For CTABr, these numbers were 3 × 10?6?2 × 10?4 M and ± 5 × 10?7 M; for BACl, they were 2 × 10?3?9 × 10?2 g l?1 and ± 1 × 10?3 g l?1. In the mixtures of CTABr and BACl, the accuracies were ± 3 × 10?6 M and × 1 × 10?3 g l?1, respectively.  相似文献   

12.
Hongzhe Tian 《Chromatographia》2011,73(5-6):457-462
A method for determination of chlormequat (CCC) residue in fruits by liquid chromatography?Ctandem mass spectrometry (LC?CMS?CMS) was developed. Residue of CCC was extracted from samples with methanol?Cwater (v/v, 1:1) containing 1.0% acetic acid, cleaned up by strong cationic exchange (SCX) cartridge, and then determined by LC?CMS?CMS. The method showed good linearity over the concentration range 0.002?C5.0 mg kg?1 with correlation coefficient above 0.997. The limit of detection (LOD) and limit of quantitation (LOQ) for CCC were 5 × 10?4 mg kg?1 (S/N = 3) and 0.002 mg kg?1 (S/N = 10), respectively. Recoveries for CCC at three spiked levels (0.025, 0.050, and 0.20 mg kg?1) were in the range 80?C102%. Estimation of measurement uncertainty was calculated for CCC at the level of 0.025 mg kg?1 in fruits. The results demonstrated that the uncertainty of recovery was the main contribution to the combined standard uncertainty. The relative combined standard uncertainties associated with the method ranged from 11 to 13%, depending on the sample matrices.  相似文献   

13.
A novel and simple biosensor based on poly(indoleacetic acid) film-modified electrode (PIAA/CPE) was fabricated by electrochemical polymerization of indoleacetic acid on a carbon paste electrode (CPE) through cyclic voltammetry. The resulting electrode was characterized by scanning electron microscopy, and the electrochemical behaviors of dopamine (DA) and epinephrine (EP) at the electrode were studied. It was illustrated that PIAA/CPE had excellent electrochemical catalytic activities toward DA and EP. The anodic peak currents (I pa) were dramatically enhanced by about seven-fold for DA and ten times for EP at PIAA/CPE. Thus, the determinations of DA and EP were carried out using PIAA/CPE successfully. The linear responses were obtained in the range of 3.0?×?10?7~7.0?×?10?4 and 1.0?×?10?6 ~8.0?×?10?4 mol L?1 with the detection limits (3σ) of 1?×?10?7 and 4?×?10?7 mol L?1 corresponding with DA and EP, respectively. Moreover, the cathodic peaks of DA and EP were well-separated with a potential difference about 325 mV in pH 5.3 phosphate-buffered saline, so simultaneous determination of DA and EP was carried out in this paper. Additionally, the interference studies showed that the PIAA/CPE exhibited excellent selectivity in the presence of ascorbic acid (AA). With good selectivity and sensitivity, the present method has been successfully applied to the determination of DA and EP in pharmaceutical samples.  相似文献   

14.
Thermal decomposition behavior and non‐isothermal decomposition reaction kinetics of nitrate ester plasticized polyether NEPE propellant containing ammonium dinitramide (ADN), which is one of the most important high energetic materials, were investigated by DSC, TG and DTG at 0.1 MPa. The results show that there are four exothermic peaks on DTG curves and four mass loss stages on TG curves at a heating rate of 2.5 K·min?1 under 0.1 MPa, and nitric ester evaporates and decomposes in the first stage, ADN decomposes in the second stage, nitrocellulose and cyclotrimethylenetrinitramine (RDX) decompose in the third stage, and ammonium perchlorate decomposes in the fourth stage. It was also found that the thermal decomposition processes of the NEPE propellant with ADN mainly have two mass loss stages with an increase in the heating rate, that is the result of the decomposition heats of the first two processes overlap each other and the mass content of ammonium perchlorate is very little which is not displayed in the fourth stage at the heating rate of 5, 10, and 20 K·min?1 probably. It was to be found that the exothermal peak temperatures increased with an increase in the heating rate. The reaction mechanism was random nucleation and then growth, and the process can be classified as chemical reaction. The kinetic equations of the main exothermal decomposition reaction can be expressed as: dα/dt=1012.77(3/2)(1?α)[?ln(1?α)]1/3 e?1.723×104/T. The critical temperatures of the thermal explosion (Tbe and Tbp) obtained from the onset temperature (Te) and the peak temperature (Tp) on the condition of β→0 are 461.41 and 458.02 K, respectively. Activation entropy (ΔS), activation enthalpy (ΔH), and Gibbs free energy (ΔG) of the decomposition reaction are ?7.02 J·mol?1·K?1, 126.19 kJ·mol?1, and 129.31 kJ·mol?1, respectively.  相似文献   

15.
The kinetics of decomposition of organocobalt chelates in the pH range of 2.2–7.0 has been studied. It has been shown that the rate constant of decomposition of the octyl chelate complex at 20°C changes from ~3 × 10?3 to ~6 × 10?6 s?1 in the above pH range. The rate constants of decomposition of complexes with ethyl, octyl, and cetyl ligands, as estimated at 20°C and pH 8.3, are 1.69 × 10?4, 1.39 × 10?4, and 2.42 × 10?5 s?1, respectively. As evidenced by emission spectrometry measurements, ~100% of organocobalt chelates with ethyl and isopropyl ligands occur in the aqueous phase, while organocobalt chelates with octyl and cetyl ligands are partitioned between monomer and aqueous phases. The rates of initiation of the emulsion polymerization of styrene have been measured by the inhibited polymerization procedure. It has been demonstrated that among three tested compounds (diphenyl picryl hydrazyl, hydroquinone, and benzoquinone), benzoquinone has been found to be a suitable inhibitor for the polymerization under study. The rates of initiation of styrene polymerization at 30°C for organocobalts with ethyl, octyl, and cyclohexyl ligands are 1.0 × 10?7, 1.04 × 10?7, and 3.7 × 10?6 mol/(l s), respectively. The rate constant of decomposition of the organocobalt complex with the octyl ligand at 30°C is 2.28 × 10?5 s?1, and the efficiency of initiation with this complex is 0.95.  相似文献   

16.
Nonlinear optical absorption at a wavelength of 1080 nm for nanocomposite thin films made from polymers and single-walled carbon nanotubes (SWNTs) was studied by the longitudinal scanning (z-scan) technique. Two SWNT types differing in the synthesis procedure (HipCO and arc evaporation techniques) were used for the preparation of nanocomposites based on the polymers carboxymethylcellulose (CMC) and poly(vinyl alcohol) (PVAL). The nonlinear absorption coefficients were measured to be ?5.0 × 10?7 and ?3.9 × 10?7 cm/W for the CMC-SWNT/HipCO and CMC-SWNT/Arc composite films and ?1.7 × 10?7 and ?0.9 × 10?7 cm/W for the PVAL-SWNT/HipCO and PVAL-SWNT/Arc films, respectively. It was found that the film nanocomposites based on carboxymethylcellulose had a higher absolute value of the nonlinear absorption coefficient than the films in which PVAL was used as the polymer matrix.  相似文献   

17.
Lewis‐acid catalyzed degradation of poly(isobutylene‐co‐isoprene) (butyl rubber) in the presence of an alkoxybenzene compound was studied as a new route toward low molecular weight multifunctional polyisobutylenes. Simultaneous cleavage and functionalization of butyl rubber was conducted at ?70 °C and ?40 °C under TiCl4 or AlCl3 catalysis in 60/40 hexane/methylene chloride cosolvents in the presence of (3‐bromopropoxy)benzene (BPB) for various times up to 24 h. The butyl rubber (EXXON? Butyl 365) possessed M n = 1.91 × 105 g/mol, PDI = 1.66 (GPC/MALLS), and 2.30 mol % isoprene units (nearly exclusively trans ?1,4). At ?70 °C with TiCl4, molecular weight was reduced to various values within the range 7 to 11 × 103 g/mol depending on conditions; lower BPB concentration produced lower molecular weight. However, the ratio of isobutylene repeat units to BPB units (IB/Q ) remained constant at about 43:1, which is approximately the same as the ratio of isobutylene to isoprene repeat units (IB/IP) in the starting butyl rubber (42.5:1). At ?40 °C with TiCl4, molecular weight was reduced to about 5 × 103 g/mol, and IB/Q was reduced below IB/IP, indicating nearly a difunctional telechelic structure. AlCl3 was a more active catalyst and produced results similar to TiCl4 at ?40 °C, even when used at seven times lower concentration. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55 , 1991–1997  相似文献   

18.
Reaction with hydroxyl radicals (OH) is the major pathway for removal of cyclic volatile methyl siloxanes (cVMS) from air. We present new measurements of second‐order rate constants for reactions of the cVMS octamethylcyclotetrasiloxane (D4), decamethylcyclopentasiloxane (D5), and dodecamethylcyclohexasiloxane (D6) with OH determined at temperatures between 313 and 353 K. Our measurements were made using the method of relative rates with cyclohexane as a reference substance and were conducted in a 140‐mL gas‐phase reaction chamber with online mass spectrometry analysis. When extrapolated to 298 K, our measured reaction rate constants of D4 and D5 with the OH radical are 1.9 × 10?12 (95% confidence interval (CI): (1.7–2.2) × 10?12) and 2.6 × 10?12 (CI: (2.3–2.9) × 10?12) cm3 molecule?1 s?1, respectively, which are 1.9× and 1.7× faster than previous measurements. Our measured rate constant for D6 is 2.8 × 10?12 (CI: (2.5–3.2) × 10?12) cm3 molecule?1 s?1 and to our knowledge there are no comparable laboratory measurements in the literature. Reaction rates for D5 were 33% higher than for D4 (CI: 30–37%), whereas the rates for D6 were only 8% higher than for D5 (CI: 5–10%). The activation energies of the reactions of D4, D5, and D6 with OH were not statistically different and had a value of 4300 ± 2800 J/mol.  相似文献   

19.
The complexation of thiabendazole (TBZ) with the cucurbit[6]uril (Q[6]), cucurbit[7]uril (Q[7]) and symmetric tetramethyl-cucurbit[6]uril (TMeQ[6]) in aqueous solution has been investigated using UV–vis and fluorespectrometry. The experimental results show 1:1 host–guest inclusion complexes at pH 6.5 for all three macrocyclic hosts, and the corresponding formation constants by UV and fluorescence methods are (5.37?±?1.05)?×?104?L?mol?1 and (1.47?±?0.41)?×?104?L?mol?1 for the Q[6]-TBZ system (7.76?±?0.51)?×?104?L?mol?1 and (9.36?±?0.22)?×?104?L?mol?1 for the Q[7]-TBZ system (1.28?±?0.78)?×?104?L?mol?1 and (2.69?±?0.55)?×?104?L?mol?1 for the TMeQ[6]-TBZ system, respectively. Based on the enhancement of the fluorescence intensity of TBZ with the addition of Q[n]s in neutral media, a fluorespectrometry method for the determination of TBZ in aqueous solution in the presence of Q[n] was established. In the range of 6.0?×?10?8?mol?L?1–8.0?×?10?6?mol?L?1 a linear relationship was obtained between fluorescence intensity and TBZ concentration. The detection limit was found to be between 5.51 and 8.85?×?10?9?mol?L?1. The interference of coexisting ions was found to be slight. The proposed method has been successfully applied to the determination of TBZ in different aqueous solutions with satisfactory recoveries of 92–103%. The method seems to be suitable for environmental water analysis.  相似文献   

20.
Thermal decomposition of neat TBP, acid-solvates (TBP·1.1HNO3, TBP·2.4HNO3) (prepared by equilibrating neat TBP with 8 and 15.6?M nitric acid) with and without the presence of additives such as uranyl nitrate, sodium nitrate and sodium nitrite, mixtures of neat TBP and nitric acid of different acidities, 1.1?M TBP solutions in diluents such as n-dodecane (n-DD), n-octane and isooctane has been studied using an adiabatic calorimeter. Enthalpy change and the activation energy for the decomposition reaction derived from the calorimetric data wherever possible are reported in this article. Neat TBP was found to be stable up to 255?°C, whereas the acid-solvates TBP·1.1HNO3 and TBP·2.4HNO3 decomposed at 120 and 111?°C, respectively, with a decomposition enthalpy of ?495.8?±?10.9 and ?1115.5?±?8.2?kJ?mol?1 of TBP. Activation energy and pre exponential factor derived from the calorimetric data for the decomposition of these acid-solvates were found be 108.8?±?3.7, 103.5?±?1.4?kJ?mol?1 of TBP and 6.1?×?1010 and 5.6?×?109?S?1, respectively. The thermochemical parameters such as, the onset temperature, enthalpy of decomposition, activation energy and the pre-exponential factor were found to strongly depend on acid-solvate stoichiometry. Heat capacity (C p ), of neat TBP and the acid-solvates (TBP·1.1HNO3 and TBP·2.4HNO3) were measured at constant pressure using heat flux type differential scanning calorimeter (DSC) in the temperature range 32?C67?°C. The values obtained at 32?°C for neat TBP, acid-solvates TBP·1.1HNO3 and TBP·2.4HNO3 are 1.8, 1.76 and 1.63?J?g?1?K?1, respectively. C p of neat TBP, 1.82?J?g?1?K?1, was also measured at 27?°C using ??hot disk?? method and was found to agree well with the values obtained by DSC method.  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