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.
J.C Scaiano  L.C Stewart 《Polymer》1982,23(6):913-915
The photochemistry of poly(phenyl vinyl ketone) in solution has been examined over a temperature range using nanosecond laser flash photolysis techniques. Experiments were performed at low temperatures to eliminate the interference from biradical signals and allow for the separation in time of the behaviour of quencher-free and quencher-containing polymers. For the former the kinetics of triplet decay in toluene-d8 lead to log (As?1) = 9.90 ± 0.50 and Ea = (3.56 ± 0.30) kcal mol?1 where A is the preexponential factor and Ea the activation energy. The extrapolated triplet lifetime at 20°C is 57 ns.  相似文献   

2.
The operation of the SDERF-cell in the study of the electron transfer kinetics of the Fe(CN)4?6/Fe(CN)3?6-system in 1 M KCl and 1 M KNO3-solutions at a stationary Pt-disk electrode is reported. The experimental current—overpotential curves are recorded by linear sweep voltammetry and analysed by two different methods using the theoretical relationship derived for a stationary disk electrode placed in a free rotating fluid. Both methods give the same value for the experimental rate constant k*. The effects of the temperature (0° to 40°C) and of the ratio of the rotor radius (rr) to the electrode radius (re)(rr/re = 0.50 to 0.81) have been studied. The activation energy for the redox process in 1 M KCl and 1 M KNO3 are: Ea = 3.4 ± 0.6 kcal/mol and Ea = 3.7 ± 0.7 kcal/mol respectively, while the k*-values at 25°C are: k* = (5.67 ± 0.41) × 10?3 cm.s?1 and k* = (4.53 ± 0.29) × 10?3 cm.s?1 respectively. The difference from the standard rate constant k0 ? 0.100 cm.s?1 is explained by the effect of the cell-geometry characterized by the G-factor, so that k° = Gk*, where G ? 19 for our cell.  相似文献   

3.
Epoxy resin of 9,9′-bis-(3,5-dibromo-4-hydroxyphenyl) anthrone-10 (EANBr, EEW 490) was synthesized and was characterized by IR and 1HNMR . EANBr and EPK3251 cured resin (EANBrC) were characterized by DSC and TGA at 10°Cmin?1 under nitrogen atmosphere. Broad DSC endothermic transitions of EANBr (265.3 °C) and EANBrC (291.4 °C) are due to some physical change and further confirmed by no weight loss in their TG thermograms. EANBr and EANBrC are thermally stable up to 340 °C and 310 °C, respectively. EANBr has followed single step degradation kinetics, while EANBrC has followed two step degradation kinetics. EANBr followed apparently zero order kinetics, while EANBrC followed apparently second order (1.80) and first order (0.89) degradation kinetics, respectively. Ea and A values of EANBrC (299.7 kJmol?1 and 6.32?×?1020 s?1) were found higher than that of EANBr (201 kJmol?1 and 2.45?×?1013 s?1) due to more rigid nature of EANBrC. The ΔS* value of the first step degradation of EANBrC (146.3 JK?1 mol?1) was found much more than that of EANBr (4.6 JK?1 mol?1). Jute – EANBr composite (J-EANBr) was prepared by compression molding technique at 120 °C for 5 h and under 20 Bar pressure. The observed tensile strength, flexural strength, electric strength and volume resistivity of J-EANBr are 24.7 MPa, 19.0 MPa, 1.8 kVmm?1 and 3.5?×?1012 ohm cm, respectively. Water absorption in J-EANBr was carried out at 30 ± 2 °C against distilled water, 10% NaCl, 10% HCl, 10% HNO3, 10% H2SO4, 10% NaOH, and 10% KOH and also in boiling water. The equilibrium time and equilibrium water content for J-EANBr in different environments are 384–432 h; 12.7–15.2%, respectively. The observed equilibrium water content and diffusivity trends of J-EANBr are KOH>H2SO4>HCl>NaOH>H2O>NaCl and H2O>NaCl>NaOH>H2SO4>HCl>KOH, respectively. Good thermo-mechanical, electrical properties and excellent hydrolytic stability of J-EANBr may be useful for high temperature applications in diverse fields.  相似文献   

4.
Ni0.4Co x Cd0.6?x Fe2O4 ferrites (x = 0.0, 0.2, 0.4, and 0.6) were prepared by autocombustion synthesis and characterized by XRD, SEM, and other physical methods. XRD reveals the formation of spinel structure without any impurity phase. With increasing x, the lattice parameter was found to decrease from 8.60 to 8.37 Å. The FTIR spectra show the presence of tetrahedral site (at 571.12 cm?1) and octahedral site (at 407.07 cm?1). SEM images indicated the formation of agglomerated grains with a size of about 4.3 μm. The resistivity of the ferrites was found to decrease from 24.53 to 10.02 × 108 Ω cm while the drift mobility to increase from 2.30 to 4.41 × 108 cm2 V?1 s?1 with increasing x.  相似文献   

5.
Biomass-based polyol obtained by chemical liquefaction technology is a potential substitute for polyether or polyester polyol in preparation of degradable polymers. To obtain the favorable biomass-based polyol products, one important emphasis is to reveal the liquefaction kinetics. The liquefaction kinetics of different corn stalk (CS) fractions, i.e. whole CS, ear husk and leaf blade, were investigated in this work. The liquefactions were catalyzed with sulfuric acid at 120–180 °C for 15–90 min. The results indicated that the apparent reaction rate constant (k), apparent activation energy (E), ΔG′, and ΔH′ of liquefaction reactions differed remarkably with different CS fractions. The highest k of 1.8 × 10?4 s?1 was obtained from ear husk liquefaction at 120 °C, which was twofold and 2.7-fold higher than those of whole CS and leaf blade, respectively. However, k is not correlated with the stalk heterogeneity at temperature over 120 °C. The calculated E ear husk, E whole CS and E leaf blade were 65.88, 81.64 and 85.23 kJ mol?1, respectively. ΔG′ and ΔH′ values of ear husk liquefaction reactions were lower than those of the other two fractions. This work was the first comparison of kinetics with different biomass fractions, casting light on the effect of heterogeneity on liquefaction, and suggesting that CS fractions should be given themselves optimum applications in future.  相似文献   

6.
Potentiostatic and galvanostatic pulse measurements were carried out to investigate the anodic oxygen evolution at platinum electrodes in 1N H2SO4 in dependence on the oxide layer thickness d and the electrode potential ε. The thickness d (1·5–10 Å) was obtained from cathodic charging curves. Further, the temperature dependence (0°–81°C) was evaluated from Bowden's measurements. Summarizing, the current io2 follows the relation, log i = A - (E0a - αFη)/2·3 RT - d/do. The experimental activation energy Eo = Eoa = αFη decreases linearly with increasing overvoltage η. The linear decrease of log i with increasing d, which is given by the term d/do, is correlated to the probability of the quantum mechanical tunnel transition of the electron from adsorbed ions, OH?ad or O2?ad respectively, through the oxide layer to the metal. Similar effects of the oxide layer thickness on the current density were observed in the case of the oxygen evolution at iridium, the CO-oxidation on platinum, and the reduction of Cl? and Ce4+ at platinum. In these cases a rate determining electron transfer through the oxide layer is also assumed.  相似文献   

7.
In this study, the ceramic powders of Ce1?xGdxO2?x/2 and Ce1?xNdxO2?x/2 (x=0.05, 0.10, 0.15, 0.20 and 0.25) were synthesized by ultrasound assisted co-precipitation method. The ionic conductivity was studied as a function of dopant concentration over the temperature range of 300–800 °C in air, using the impedance spectroscopy. The maximum ionic conductivity, σ800 °C=4.01×10?2 Scm?1 with the activation energy, Ea=0.828 kJmol?1 and σ800 °C=3.80×10?2 Scm?1 with the activation energy, Ea=0.838 kJmol?1 were obtained for Ce0.90Gd0.10O1.95 and Ce0.85Nd0.15O1.925 electrolytes, respectively. The average grain size was found to be in the range of 0.3–0.6 μm for gadolinium doped ceria and 0.2–0.4 μm for neodymium doped ceria. The uniformly fine crystallite sizes (average 12–13 nm) of the ultrasound assisted prepared powders enabled sintering of the samples into highly dense (over 95%) ceramic pellets at 1200 °C (5 °C min?1) for 6 h.  相似文献   

8.
The kinetics of aluminium deposition from NaClAlCl3 and NaClKClAlCl3 melts (cAlCl3 < 0.4 mol%) was studied by linear sweep voltammetry and potential step amperometry. The reduction of AlCl3 on tungsten and aluminium electrodes was found to be diffusion controlled. The diffusion coefficients of AlCl3 were: 3.5 × 10?5 cm2 s?1 at 820°C in NaClAlCl3, 2.7 × 10?5cm2s?1 at 825°C, and 2.1 × 10?5cm2s?1 at 705°C in KClNaClAlCl3. The rate constant for AlCl3 reduction at these conditions was found to be in the order of 0.2 cm s?1, in good agreement with extrapolated literature data.  相似文献   

9.
《Ceramics International》2022,48(2):1956-1962
A series of (In1-xAlx)2O3 (0.1 ≤ x ≤ 0.6) films with tunable bandgap were grown on MgO (100) substrates by MOCVD. The influences of chemical compositions and growth temperatures on the film properties were studied systematically. XRD analyses indicated that the film quality degraded from crystalline to amorphous structure as Al concentration (x) increased. The (In1-xAlx)2O3 films prepared at 700 °C exhibited better film crystallinity than those of the ones grown at 600 °C. The films prepared at 700 °C with x = 0.1–0.3 showed an epitaxial In2O3 <111> orientation with the corresponding growth relationship of In2O3 (111)∥MgO (100). The film with x = 0.2 exhibited the best crystallinity and the largest grain size of 25.9 nm. The Hall mobilities and resistivities of the films were influenced evidently by Al concentrations. The Hall mobility showed a monotonous decrease from 12 to 1.1 cm2V?1s?1 as x increased from 0.1 to 0.6. The lowest resistivity of 9.2 × 10?3 Ω cm was acquired for the film with x = 0.2. The average transmittances in the visible region for all the films were beyond 83%. The bandgap of the (In1-xAlx)2O3 films can be regulated in the range of 3.85–4.88 eV by changing Al concentrations from 0.1 to 0.6.  相似文献   

10.
Belle Ayr subbituminous coal was dried with gases including nitrogen, air, and nitrogen-air mixtures (simulated flue gases) to study the effect of drying on the coal characteristics in preparation for subsequent liquefaction experiments. Drying was carried out in micro-, laboratory- and bench-scale units at temperatures from ambient to 200 °C. The net moisture-free oxygen content of the coal increased with time and temperature to 3 wt%. Volatile oxygen-containing species, other than carbon oxides, that may have been released during drying were not investigated as the objective was to characterize oxidation kinetics and changes in coal properties. Two distinct kinetic regimes of oxygen consumption were observed during drying; an initial high-rate period of EA?42–55 kJ mol?1 followed by one of low rate, EA?13 kJ mol?1. A Powhatan No. 5 (Pittsburgh seam) bituminous coal, which initially had much lower oxygen content than the Belle Ayr coal (7.9 versus 23.3 wt%), gave analogous drying and oxidation results; however, the maximum net moisture-free oxygen uptake was ≈8 wt%.  相似文献   

11.
Mary E Galvin  Gary E Wnek 《Polymer》1982,23(6):795-797
Composites of low density polyethylene, LDPE, and polyacetylene, (CH)x, were prepared by polymerization of acetylene in LDPE films impregnated with a Ti(OBu)4Et3Al Ziegler-Natta catalyst. LDPE films were immersed in a toluene solution of this catalyst at 70°C to accomplish impregnation. Polymerization of acetylene in the LDPE films was carried out at 100°–110°C. The resulting composite films remain flexible and tough upon prolonged air exposure and those containing <ca. 5 wt.% (CH)x soften upon heating to 120°–160°C. The films can be rendered conductive upon exposure to a 2 wt%. solution of l2 in pentane. Ultimate conductivities of ca. 10Ω?1cm?1 can be obtained. The conductivities of the doped composites decay more slowly as compared with l2-doped (CH)x films. The synthetic approach, with some simple modifications, can be used in the construction of many electrically conductive, all-polymer composites having a combination of desirable physical, mechanical and electrical properties.  相似文献   

12.
The degradation process of commercial grade Lexan® was investigated by thermogravimetric technique under isothermal experimental conditions at four different operating temperatures: 375 °C, 387.5 °C, 400 °C and 425 °C. The kinetic triplet (E a , A, f(α)) was determined using conventional and Weibull kinetic analysis. The applied kinetic procedure shows that the investigated degradation process can be described by two-parameter autocatalytic ?esták–Berggren (SB) reaction model. It was established that the degradation process of Lexan® can be described by the following kinetic triplet: E a? =?158.3 kJ mol?1, A?=?8.80?×?109 min?1 and f(α)?=?α 0.33 (1???α)1.62. It was established that the operating temperature has an influence on the values of SB reaction orders (m and n) (0.27?m?n??1, represent the composite value from a complex degradation reaction and can not compare with the dissociation energy of the weak bonds in bisphenol-A polycarbonate. Also, it was concluded that the Weibull shape parameter (β) shows that the considered process occurs under the same reaction mechanism, independently on operating temperature (T), i.e. the change of rate-limiting step does not occur (β?ddf) of apparent activation energies for considered degradation process. On the other hand, it was shown that the experimentally evaluated density distribution function of apparent activation energies represents the intermediate case between the calculated density distribution functions at 375 °C and 425 °C.  相似文献   

13.
We investigated annealing effects of La1?xSrxMnO3 (x = 0–0.6) on electrical resistivity and the temperature coefficient of resistivity (TCR). The annealed samples’ resistivity was lower than those of non-annealed samples. For example, annealing changed the resistivity of x = 0.3 at 25 °C from 4.50 × 10?5 to 3.71 × 10?5 Ω m. Remarkable difference in TCR was observed after annealing, for x = 0.3, 0.45, and 0.5. For x = 0.3, the TCR after annealing was 4000 ppm/°C, which was 1250 ppm/°C greater than that before annealing. We investigated (1) crystal phase, (2) Mn average valence, (3) Mott insulator–metal transition temperature, and (4) microstructure. The microstructure was remarkably varied for annealed x = 0.3 and 0.5. The average grain size of the x = 0.3 increased from 1.60 up to 2.38 μm. Results show that annealing affects resistivity and TCR because of grain growth during annealing.  相似文献   

14.
Electrochemical reactions were assigned to the voltammetric waves obtained in alkaline KMnO4 and K2MnO4 solutions. Plots of ip (AC°ν12)?1 and iapicp?1 νs log ν indicate that at slow potential scan rates in MnO?4 solutions ranging from 0·07 to 0·19 F in NaOH the first step in MnO?4 reduction is a l-electron reversible charge transfer with no coupled chemical reaction; at fast scan rates (? V s?1 the process becomes quasi-reversible. The standard rate constant and the activation energy for the MnO?4/MnO2?4 charge transfer step were estimated. Plots of ip (AC°ν12)?1 and iapicp?1 νs log ν indicate that in 4·O F KOH and at potential scan rates between 0·005 and 2·V s?1 the reduction of MNO2?4 to MnO3?4 is a reversible charge transfer with no coupled chemical reaction. The diffusion coefficients of MnO?4 and MnO2?4 in alkaline solutions are reported.  相似文献   

15.
Stability constants for cadmium(II) complexes with tetraethylenepentamineheptaacetic acid (TPHA, H7L) were determined by the pH titration method. In an aqueous solution (μ = 0.1), three complex species, CdH2L, CdHL and CdL are confirmed. The structure of uninuclear complexes are discussed. The formation constants of the complexes stated above have been calculated as follows (at 25 ± 0.1°C): log KCdL = 15.35, log KCdHL = 13.33 and log KCdH2L = 7.89. The polarographic behaviour of the cadmium(II) in the presence of TPHA was studied over the pH range 3–5. Mechanisms of the electrode processes were elucidated and electrochemical kinetic parameters were evaluated from dependence of the half-wave potentials on the hydrogen ion and TPHA concentration. In the presence of an excess of TPHA, the wave B is assigned to the irreversible reductions of the complex, CdH2L3? (pH range 3–4) or CdHL4? (pH range 4–5). The electrode reaction can be written:
and
Where ke (the rate constant) = 2.3 × 10?2 cm s?1 and ke = 1.59 × 10?4 cm s?1. The other polarographic methods were also used in the elucidation of the electrode process.  相似文献   

16.
Aluminum gallium oxide (AGO) films were prepared on conventional c-plane sapphire by pulsed laser deposition (PLD). In the current PLD-AGO studies, target composition or growth temperature is usually the main deposition variable, and the other growth conditions are fixed. This would make it difficult to fully understand the theory and characterization of AGO films. In this study, several growth parameters such as target composition, gas atmosphere, laser repetition frequency, growth pressure, and substrate temperature (Ts) were all modulated to realize and optimize the AGO growth. When the (AlxGa1-x)2O3 target with the Al content larger than 20?at% was used, a serious target poisoning phenomenon occurred, leading to the extremely unstable growth rate. In comparison to the AGO film grown with argon atmosphere, the higher transparency was reached in the film prepared with oxygen atmosphere due to the relative abundance of oxygen. Because of the homogeneous oxygen reduction, the AGO film with the higher crystal quality was obtained at a higher laser repetition frequency. With an increment of growth pressure, the Al content of AGO film was increased. The growth of AGO film at the higher Ts would cause the higher bandgap value, smoother surface, and growth rate degradation. Additionally, the crystal quality of AGO film can be also improved both by increasing the growth pressure and Ts. The better characterization can be reached in the AGO film grown using the (Al0.05Ga0.95)2O3 target with oxygen atmosphere at the working pressure of 2?×?10?1 Torr, the laser repetition frequency of 10?Hz, and the Ts of 800?°C. When the metal-semiconductor-metal photodetector fabricated with this AGO active layer, the best performance including the photocurrent of 7.56?×?10?8 A, dark current of 1.59?×?10–12 A, and photo/dark current ratio of 4.76?×?104 (@5?V and 240?nm) were achieved.  相似文献   

17.
In this study, La0.4Sr0.6CoO3‐δ (LSC) oxide was synthesized via an EDTA‐citrate complexing process and its application as a mixed‐conducting ceramic membrane for oxygen separation was systematically investigated. The phase structure of the powder and microstructure of the membrane were characterized by XRD and SEM, respectively. The optimum condition for membrane sintering was developed based on SEM and four‐probe DC electrical conductivity characterizations. The oxygen permeation fluxes at various temperatures and oxygen partial pressure gradients were measured by gas chromatography method. Fundamental equations of oxygen permeation and transport resistance through mixed conducting membrane were developed. The oxygen bulk diffusion coefficient (Dv) and surface exchange coefficient (Kex) for LSC membrane were derived by model regression. The importance of surface exchange kinetics at each side of the membrane on oxygen permeation flux under different oxygen partial pressure gradients and temperatures were quantitatively distinguished from the oxygen bulk diffusion. The maximum oxygen flux achieved based on 1.6‐mm‐thick La0.4Sr0.6CoO3‐δ membrane was ~4.0 × 10?7 mol cm?2 s?1at 950°C. However, calculation results show theoretical oxygen fluxes as high as 2.98 × 10?5 mol cm?2 s?1 through a 5‐μm‐thick LSC membrane with ideal surface modification when operating at 950°C for air separation. © 2009 American Institute of Chemical Engineers AIChE J, 2009  相似文献   

18.
The effect of radiation dose (10–30 kGy) on the thermal decomposition of poly(ethylene terephthalate) was studied using thermogravimetric analysis (TGA), differential scanning calorimetry (DSC), and X‐ray diffraction (XRD) analysis. The TGA and DSC were carried out in a flowing nitrogen atmosphere at heating rates of 5 and 30°C/min for TGA and 10°C/min for DSC. The degradation process was composed of three overlapping stages. The second stage, at which a rapid degradation occurs, was studied in detail. The process was found to follow a second‐order kinetics and was independent of radiation dose or heating rate. The reaction rate constant (k) was found to depend on the heating rate and iradiation dose. The apparent activation energy (Q) and the logarithm of the preexponential rate constant (log A) were found to decrease linearly with the increase in dose at rates of 3.32 kJ mol?1 kGy?1 and 0.177 s?1 kGy?1 with intercepts of 249 kJ/mol and 12.26 s?1 for Q and log A of unirradiated fabric, respectively. A direct relationship was found between the percentage decrease in Q and log A and the percentage decrease in the temperature corresponding to 50% conversion (T50%) for samples irradiated at different doses. It was found that a decrease in T50% by 1% resulted in a decrease in Q and log A by 1.855 and 2.1%, respectively. Changes in Q and log A resulting from radiation, mechanical and thermal treatments, or their combinations can be predicted from the shift in T50%. The history of the fibers substantially affected the thermal properties. DSC and XRD studies revealed changes in the fabric crystallinity. DSC measurements indicated a linear increase in heat of fusion with dose increase at a rate of 0.855 kJ kg?1 kGy?1. XRD analysis confirmed structural changes, rearrangement by plane rotations, and formation of compact crystalline lattice with patterns characterizing irradiated samples. An attempt to explain the dependency of the apparent activation energy on dose was given. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 92: 3710–3720, 2004  相似文献   

19.
In this work, the relationship between the structural mechanisms and macroscopic electrical properties of the Nb-modified 0.96(Bi0.5Na0.84K0.16TiO3)–0.04SrTiO3 (BNKT–ST) system were elucidated by using temperature dependent and in situ synchrotron X-ray diffraction (XRD) techniques. For the composition x?=?0.0175, a large-signal piezoelectric coefficient (Smax/Emax?=?d33*) of 735 pm?V?1 at 6?kV mm?1 was observed at room temperature. Interestingly, at a higher temperature of 110?°C, the sample still showed a large d33* of 570 pm V?1. Furthermore, the temperature-invariant electrostrictive coefficient for this sample was found to be 0.0285?m4?C?2 over the temperature range of 25–170?°C. Moreover, the energy density for x?=?0.030 sample was ~1.0?J?cm?3 with an energy storage efficiency of ?70% in the temperature range of 25–135?°C. These results suggest that the synthesized Nb-modified BNKT–ST system is promising for the design of ceramic actuators as well as capacitor applications.  相似文献   

20.
《Ceramics International》2022,48(1):415-426
The oxygen transport membrane (OTM) has huge application prospects in gas separation and carbon neutralization based on oxygen enriched combustion. In this paper, the family 60 wt.%Ce0.9Pr0.1O2-δ-40 wt.%Pr0.6Sr0.4Fe1-xInxO3-δ (CPO-PSF1-xIxO, x = 0.01, 0.025, 0.05, 0.075, 0.1) cobalt-free dual-phase MIEC OTMs doped with indium have been successfully prepared by Pechini method. The phase structure, surface morphology, element distribution, oxygen permeability, and long-term operation stability of these OTMs are systematically explored. Among these OTMs, the champion oxygen permeable flux of CPO-PSF0.99I0.01O reaches 1.07 mL min?1·cm?2 and 0.80 mL min?1·cm?2 at 1000 °C under air/He gradient and air/CO2 gradient. Meanwhile, CPO-PSF0.99I0.01O maintains the value of 0.80 mL min?1·cm?2 steadily at 1000 °C for 100 h when pure CO2 as the sweep gas. The surface element distribution and phase structure of the OTMs after long-term oxygen permeability reaction are investigated by XRD, SEM combining with EDS, where the spent membranes retain the same structure and component as the fresh membranes, demonstrating that the In-doped OTMs have an excellent CO2 tolerance. Suitable indium substitution for iron of these OTMs not only improves the oxygen permeability, but also maintains the long-term reaction stability of the material.  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