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.
New polymeric solid electrolyte films, consisting of crosslinked poly(N-vinylpyrrolidone) (PVPD) as matrix, and surfactant, sodium deoxycholate (NaDC), lithium deoxycholate (LiDC), sodium laulylsulfate (R12OSO3Na), or sodium palmitate (R15COONa) as electrolyte salt, are prepared; their basic structure and conductivity dependence on temperature are reported. The structure of the electrolytes is amorphous. Their conductivity is 3.1 × 10?5 S cm?1 (containing NaDC), 8.42 × 10?6 S cm?1 (LiDC), 2.18 × 10?4 S cm?1 (R12OSO3Na), and 7.27 × 10?5 S cm?1 (R15COONa) at 20°C. Their temperature dependence of the conductivity is similar to that of liquid electrolyte rather than that of usual polymeric solid electrolyte, i.e., the WLF-type dependence. The values of activation energy of conductivity (Ea) were PVPD, 25.5 kJ mol?1; PVPD/NaDC, 21.4 kJ mol?1; PVPD/LiDC, 25.3 kJ mol?1; PVPD/R12OSO3Na, 17.2 kJ mol?1; PVPD/R15COONa, 18.7 kJ mol?1. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
The polymerization of phenylacetylene to polyphenylacetylene was accomplished with the combined catalysts triethyl aluminum and titanium tetraethoxide. The progress of the reaction was monitored by gas chromatography. The parameters included temperature (?80, 25, 140°C), solvent (benzene, chlorobenzene, toluene, cyclohexane, and nitrobenzene), mole ratio of catalysts (Al/Ti; 1.5, 3.0, 4.5, 6.0, and 9.0), aging times of catalysts (2, 10, and 40 min), and order of addition of reagents. Derivatives of polyphenylacetylene were obtained by the acylation of polyphenylacetylene with p-nitrobenzoyl chloride, the sulfonation of polyphenylacetylene with benzenesulfonyl chloride, and the formation of polyphenylacetylene complexes with complexing agents such as bromine, iodine, iodine chloride, boron trifluoride, and ferric chloride. A new phenylacetylene-acetylene product mixture was produced by the polymerization of phenylacetylene and acetylene at 25 and ?80°C. The electrical conductivity of polyphenylacetylene and its derivatives is in the range of 10?10?10?3 Ω?1 cm?1.  相似文献   

3.
Iodine and nylon-6 form adducts containing 70–90 wt-% of iodine on heating at 115–145°C. The adducts have electrical conductivities ranging from 10?7 to 10?3 S cm?1 at 25°C, and the electrical conductivity increases with increasing content of iodine of the adduct. The activation energies of the electrical conduction for the adducts prepared at 115°C and containing 69.2, 81.8, 87.1, and 90.0 wt-% of iodine are 94.6, 67.0, 52.9, and 46.1 kJ/mol, respectively. Polyamides other than nylon-6 also form similar semiconducting adducts with iodine. IR and NMR spectroscopic analyses of the iodine—nylon-6 adducts indicate profound changes in the structure of nylon-6 on adduct formation and suggest the formation of a ?CN+H species. The iodine-nylon-6 adducts prepared at 115°C and containing more than 82 wt-% of iodine serve as good active materials of positive electrodes in lithium-iodine solid electrolyte galvanic cells (outer diameter = 11.6 mm; outer thickness = 2.0 mm). The current efficiencies of the galvanic cells at 500 kΩ load are about 50% based on the iodine added. Discharge at 100 kΩ load gives lower current efficiencies. The galvanic cell has an internal resistance of about 5 kΩ at 25°C before discharge, and the internal resistance increases to about 100 kΩ at about 40% discharge. The dependence of the internal resistance during discharge have been determined.  相似文献   

4.
Electropolymerization of cyanogen in acetonitrile containing an electrolyte yields a poly(cyanogen). Its structure involves an open-structured dimer derived from the heterocyclic anion C7N7? mixed with a sequence of nitrile-substituted, conjugated carbon–nitrogen bonds. Although this polymer is an insulating solid, pyrolysis in vacuo to 700°C leads to highly conducting carbon–nitrogen pyropolymers via crosslinking and nitrogen elimination. The 700° pyropolymer has a carbon–nitrogen ratio of 5:1, a room temperature conductivity of 1 Ω?1 cm?1, and an activation energy for conduction of ~0.03 eV.  相似文献   

5.
X-Ray diffraction, density, and electrical conductivity measurements were performed on the perovskite-like mixed oxide La0.84Sr0.16MnO3. A rhombohedral crystalline structure with lattice parameters a = 3.893 Å and α = 90°29′16″ was assigned to the powder prepared by standard ceramic technique. Its theoretical density is therefore 6.576 g/cm3, while the experimental density was determined as 6.48 g/cm3. The conductivity measured at 1000°C is 133 Ω?1 cm?1. The temperature dependence of the conductivity indicates that the charge carriers are small polarons. The activation energy of the mobility is 9.6 kJ/mole.  相似文献   

6.
Transport properties of ionic salt CsH5(PO4)2 are studied by the impedance method. The salt’s bulk conductivity ranges from 10?8 to 10?4 S cm?1 in the temperature interval 90 to 145°C. The apparent activation energy is high (1.6–2.0 eV). The conductivity is slightly anisotropic: it is maximum in the [001] direction and minimum in the [100] direction (~5.6 and 1 times × 10?6 S cm?1, respectively, at 130°C). The conductivity of polycrystalline samples is higher by 1–2 orders of magnitude, and the activation energy drops to 1.05 eV due to the formation of a pseudoliquid layer with a high proton mobility at the intercrystallite boundary. The salt’s thermodynamic properties are examined by differential scanning calorimetry and thermogravimetry. No phase transitions are discovered in the salt up to the melting point (151.6°C), with the melting enthalpy equal to ~34 kJ mol?1. The crystallization occurs at lower temperatures (107°C) and the crystallization enthalpy (?18 kJ mol?1) is lower than the melting enthalpy. The melting is accompanied by slow decomposition of the salt. Factors affecting the proton transport in the salt are analyzed.  相似文献   

7.
In order to appreciate the excellent catalytic effect of iodine on the alcoholyses of alkoxysilanes more precisely, the rates of the reaction, Et3SiOBun + BusOH ? Et3SiOBus + BunOH, were determined at various iodine concentrations. Both forward and reverse reactions are first order with respect to butoxysilane and to butanol, and pseudo first-order rate constants were measured at 40°, 30°, and 20°C on reaction mixtures containing both butanols in excess by means of gas-liquid chromatography. The observed rate constants as a function of iodine concentration gave linear relationships, and from these data the catalytic coefficients of iodine were evaluated: The enthalpies and the entropies of activation were estimated to be 53.2 kJ mol?1, ?103 J K?1 mol?1 (forward, 30°C) and 51.8 kJ mol?1, minus;100 J K?1 mol?1 (reverse, 30°C).  相似文献   

8.
Solid solutions based on cesium monogallate CsGaO2 are synthesized in the Ga2O3-TiO2-Cs2O system. Their crystalline structure and also temperature and concentration conductivity dependences are studied. The cesium cation character of conductivity is confirmed. The most conducting samples contain an excess of cesium oxide and have the structure of high-temperature γ-modification of KAlO2. Their specific conductivity is (5.0–6.7) × 10?3 S cm?1 at 400 °C, (2.5–5.0) × 10?2 S cm?1 at 700°C at the activation energy of 33–35 kJ/mol?1.  相似文献   

9.
The preparation and characterization of the cocrystalline solid–organic sodium ion electrolyte NaClO4(DMF)3 (DMF=dimethylformamide) is described. The crystal structure of NaClO4(DMF)3 reveals parallel channels of Na+ and ClO4? ions. Pressed pellets of microcrystalline NaClO4(DMF)3 exhibit a conductivity of 3×10?4 S cm?1 at room temperature with a low activation barrier to conduction of 25 kJ mol?1. SEM revealed thin liquid interfacial contacts between crystalline grains, which promote conductivity. The material melts gradually between 55–65 °C, but does not decompose, and upon cooling, it resolidifies as solid NaClO4(DMF)3, permitting melt casting of the electrolyte into thin films and the fabrication of cells in the liquid state and ensuring penetration of the electrolyte between the electrode active particles.  相似文献   

10.
In order to understand the mobility of uranium it is very important to know about its sorption kinetics and the thermodynamics behind the sorption process on soil. In the present study the sorption kinetics of uranium was studied in soil and the influence parameters to the sorption process, such as initial uranium concentration, pH, contact time and temperature were investigated. Distribution coefficient of uranium on soil was measured by laboratory batch method. Experimental isotherms evaluated from the distribution coefficients were fit to Langmuir, Freundlich and Dubinin?CRadushkevich (D?CR) models. The sorption energy for uranium from the D?CR adsorption isotherm was calculated to be 7.07?kJ?mol?1.The values of ??H and ??S were calculated to be 37.33?kJ?mol?1 and 162?J?K?1?mol?1, respectively. ??G at 30?°C was estimated to be ?11.76?kJ?mol?1. From sorption kinetics of uranium the reaction rate was calculated to be 1.6?×?10?3?min?1.  相似文献   

11.
Lanthanum sulfophenyl phosphate (LaSPP) was synthesized by m-sulfophenyl phosphonic acid and lanthanum nitrate. UV-Vis spectrophotometry and Fourier-transform infrared spectroscopy indicate that the desired product was obtained and its elementary composition and typical layered structure were determined by energy dispersive X-ray spectroscopy and scanning electron microscopy. Transmission electron microscopy (TEM) proved its typical layered structure and X-ray diffraction spectroscopy indicated its good crystallinity and the interlayer distance of about 15.67 Å, which matches the value obtained by TEM (2.0 nm). Thermogravimetry and differential thermal analysis revealed good thermal stability of LaSPP. Proton conductivity of LaSPP was measured at different temperatures and relative humidities (RH), reaching values of 0.123 S cm?1 at 150°C and 100 % RH. Proton transfer activation energy was 22.52 kJ mol?1. At 160°C and 50 % RH, the conductivity was 0.096 S cm?1. In the drying oven, the conductivity retained the value of 1.118 × 10?2 S cm?1. The results show that LaSPP is a highly effective inorganic-organic conductor.  相似文献   

12.
The rates of an ene reaction between 4-phenyl-1,2,4-triazoline-3,5-dione and hex-1-ene were studied in a temperature range of 15–40 °C and in a pressure range of 1–2013 bar. The enthalpy of reaction in 1,2-dichloroethane (?158.2±1.0 kJ mol?1), the enthalpy (51.3±0.5 kJ mol?1), entropy (122±2 J mol?1 K?1), and volume of activation (?31.0±1.0 cm3 mol?1), and the volume of this reaction (?26.6±0.3 cm3 mol?1) were determined. The high exothermic effect of the reaction suggests its irreversibility.  相似文献   

13.
Pyrolytic characteristics and kinetics of pistachio shell were studied using a thermogravimetric analyzer in 50?C800?°C temperature range under nitrogen atmosphere at 2, 10, and 15?°C?min?1 heating rates. Pyrolysis process was accomplished at four distinct stages which can mainly be attributed to removal of water, decomposition of hemicellulose, decomposition of cellulose, and decomposition of lignin, respectively. The activation energies, pre-exponential factors, and reaction orders of active pyrolysis stages were calculated by Arrhenius, Coats?CRedfern, and Horowitz?CMetzger model-fitting methods, while activation energies were additionaly determined by Flynn?CWall?COzawa model-free method. Average activation energies of the second and third stages calculated from model-fitting methods were in the range of 121?C187 and 320?C353?kJ?mol?1, respectively. The FWO method yielded a compatible result (153?kJ?mol?1) for the second stage but a lower result (187?kJ?mol?1) for the third stage. The existence of kinetic compensation effect was evident.  相似文献   

14.
At room temperature and below, the proton NMR spectrum of N-(trideuteriomethyl)-2-cyanoaziridine consists of two superimposed ABC patterns assignable to two N-invertomers; a single time-averaged ABC pattern is observed at 158.9°C. The static parameters extracted from the spectra in the temperature range from –40.3 to 23.2°C and from the high-temperature spectrum permit the calculation of the thermodynamic quantities ΔH0 = ?475±20 cal mol?1 (?1.987 ± 0.084 kJ mol?1) and ΔS0 = 0.43±0.08 cal mol?1 K?1 (1.80±0.33 J mol?1 K?1) for the cis ? trans equilibrium. Bandshape analysis of the spectra broadened by non-mutual three-spin exchange in the temperature range from 39.4–137.8°C yields the activation parameters ΔHtc = 17.52±0.18 kcal mol?1 (73.30±0.75 kJ mol?1), ΔStc = ?2.08±0.50 cal mol?1 K?1 (?8.70±2.09 J mol?1 K?1) and ΔGtc (300 K) = 18.14±0.03 kcal mol?1 (75.90±0.13 kJ mol?1) for the transcis isomerization. An attempt is made to rationalize the observed entropy data in terms of the principles of statistical thermodynamics.  相似文献   

15.
Thermal analysis was used to study the influence of CaCl2 and urea as possible chemical additives inhibiting coal oxidation process at temperatures 100?C300?°C. Weight increase due to oxygen chemisorption and corresponding amount of evolved heat were evaluated as main indicative parameters. TA experiments with different heating rates enabled determination of effective activation energy E a as a dependence of conversion. In the studied range of temperatures, the interaction of oxygen with (untreated) coal was confirmed rather as a complex process giving effective activation energies changing continuously from 70?kJ?mol?1 (at about 100?°C) to ca. 180?kJ?mol?1 at temperatures about 250?°C. The similar trend in E a was found when chemical agents were added to the coal. However, while the presence of CaCl2 leads to higher values of the effective activation energies during the whole temperature range, urea causes increase in E a only at temperatures below 200?°C. Exceeding the temperature 200?°C, the presence of urea in the coal induces decrease in activation energy of the oxidation process indicating rather catalysing than inhibiting action on coal oxidation. Thus, CaCl2 can only be recommended as a ??real?? inhibitor affecting interaction of coal with oxygen at temperatures up to 300?°C.  相似文献   

16.
The standard Gibbs free energy, enthalpy, and entropy of complex formation of five solid molecular complexes of iodine have been determined by comparing the e.m.f.'s of galvanic cells having either solid iodine or the iodine complex as cathode. All of the complexes were found to have a negative enthalpy of formation, which was in the range ?5 to ?14 kJ mol?1, except for one very weak complex. The relative stability of the complexes was largely determined by the standard entropy of formation which varied from +18 J K?1 mol?1, for the most stable of the complexes studied, to ?21 J K?1 mol?1.  相似文献   

17.
The sequential segregation of Sn and Sb to the surface of a Cu(111) single crystal was measured in the temperature range 400–1100 K by Auger electron spectroscopy. It was found that Sn with the higher diffusion coefficient first segregates to the surface and then is replaced by the slower‐segregating Sb. The results were fitted by a ternary segregation model yielding segregation energies (ΔGSn = 76.3 kJ mol?1, ΔGSb = 95.9 kJ mol?1), interaction parameters (ΩSnCu = 3.8 kJ mol?1, ΩSbCu = 16.2 kJ mol?1, ΩSnSb = ?5.3 kJ mol?1) and diffusion coefficients (D0(Sn) = 1.8 × 10?5 m2 s?1, ESn = 173 kJ mol?1, D0(Sb) = 6.0 × 10?5 m2 s?1, ESb = 205 kJ mol?1) for both species. The validity of the interaction coefficients and segregation energies was verified using the Guttman equations for equilibrium segregation in ternary systems. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

18.
The combustion enthalpy of diosgenin was determined by oxygen-bomb calorimetry. The standard mole combustion enthalpy and the standard mole formation enthalpy have been calculated to be ?16098.68 and ?528.52 kJ mol?1, respectively. Fusion enthalpy and melting temperature for diosgenin were also measured to be ?34.43 kJ mol?1 and 212.33°C, respectively, according to differential scanning calorimetry (DSC) data. These studies can provide useful thermodynamic data for this compound.  相似文献   

19.
Ionic liquid monomer couples were prepared by the neutralization of 1‐vinylimidazole with vinylsulfonic acid or 3‐sulfopropyl acrylate. These ionic liquid monomer couples were viscous liquid at room temperature and showed low glass transition temperature (Tg) at ?83 °C and ?73 °C, respectively. These monomer couples were copolymerized to prepare ion conductive polymer matrix. Thus prepared ionic liquid copolymers had no carrier ions, and they showed very low ionic conductivity of below 10?9 S cm?1. Equimolar amount of lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) to imidazolium salt unit was then added to generate carrier ions in the ionic liquid copolymers. Poly(vinylimidazolium‐co‐vinylsulfonate) containing equimolar LiTFSI showed the ionic conductivity of 4 × 10?8 S cm?1 at 30 °C. Advanced copolymer, poly(vinylimidazolium‐co‐3‐sulfopropyl acrylate) which has flexible spacer between the anionic charge and polymer main chain, showed the ionic conductivity of about 10?6 S cm?1 at 30 °C, which is 100 times higher than that of copolymer without spacer. Even an excess amount of LiTFSI was added, the ionic conductivity of the copolymer kept this conductivity. This tendency is completely different from the typical polyether systems. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

20.
Tannase has been extensively applied to synthesize gallic acid esters. Bioimprinting technique can evidently enhance transesterification-catalyzing performance of tannase. In order to promote the practical utilization of the modified tannase, a few enzymatic characteristics of the enzyme and its kinetic and thermodynamics properties in synthesis of propyl gallate by transesterification in anhydrous medium have been studied. The investigations of pH and temperature found that the imprinted tannase holds an optimum activity at pH?5.0 and 40?°C. On the other hand, the bioimprinting technique has a profound enhancing effect on the adapted tannase in substrate affinity and thermostability. The kinetic and thermodynamic analyses showed that the modified tannase has a longer half-time of 1,710?h at 40?°C; the kinetic constants, the activation energy of reversible thermal inactivation, and the activation energy of irreversible thermal inactivation, respectively, are 0.054?mM, 17.35?kJ?mol?1, and 85.54?kJ?mol?1 with tannic acid as a substrate at 40?°C; the free energy of Gibbs (??G) and enthalpy (??H) were found to be 97.1 and 82.9?kJ?mol-1 separately under the same conditions.  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