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 93 毫秒
1.
In this work, samples of Y0.07Sr0.93Ti1-xFexO3-δ with 20, 40, 60 and 80 mol% of iron amount were prepared by a low-temperature polymer precursor method. The SEM-EDS analysis proved that analyzed Y0.07Sr0.93Ti1-xFexO3-δ samples were composites of two Ti- and Fe-rich perovskite samples. This kind of composite consists of two phases in which one has a good ionic and the other electronic conductivity, which makes such a composite a potential mixed ionic and electronic conductors (MIECs) material. The total electrical conductivities of analyzed samples were measured in air atmosphere (cathode conditions in Solid Oxide Fuel Cell). The values changed from ∼10−3 to 10−1 S cm−1 and depended on the ratio between two observed perovskite phases. The 0.12 S cm−1 conductivity value at 800 °C for sample with the highest amount of Fe-rich perovskite in the structure makes this composite material a candidate for air electrode in electrochemical devices.  相似文献   

2.
Pt/Eu2O3-CeO2 materials with different Eu concentrations were prepared and applied to toluene destruction, and the remarkable promotion impact of EuOx on Pt/CeO2 can be observed. The characterization results reveal that the presence of EuOx significantly enhances the redox property, lattice O concentration, and Ce3+ ratio of the Pt/CeO2 material, which facilitates the dispersion and activity of Pt active sites and thus accelerates the decomposition process of toluene. Among all catalysts, a sample with an Eu content of 2.5 at.% (Pt/EC-2.5) possesses the best catalytic activity with 0.09 vol% of toluene completely destructed at 200 °C under a relatively high GHSV of 50000 h?1. The possible reaction pathway and mechanism of toluene combustion over Pt/Eu2O3-CeO2 samples are presented according to in-situ DRIFTS, which confirms that the toluene oxidation process obeys the Mars-van Krevelen mechanism with aldehydes and ketones as primary organic intermediates.  相似文献   

3.
The structures of the low-and high-temperature modifications of lithium orthotantalate, Li3TaO4, have been determined by neutron and X-ray diffraction methods. The low-temperature, or β, phase has symmetry C2c and lattice parameters a1 = 8.500(3), b1 = 8.500(3), c1 = 9.344(3)Å, and β = 117.05(2)°. The high-temperature, or α, phase has symmetry P2 and lattice parameters ah = 6.018(1), bh = 5.995(1), ch = 12.865(2)Å, and βh = 103.53(2)°. Both structures are ordered. The β-phase has a rock salt-type structure with a 3 : 1 ordering of the Li+ and Ta5+ ions. Its structure can be generated from the low-temperature modification by means of a complex pattern of shifts of the Ta5+ ions.  相似文献   

4.
The heat of immersion in water was measured at 25°C for three iron(III) oxides using a twin-type microcalorimeter. One of the samples was commercial α-Fe2O3 (sample C) and the other two (samples M and F) were prepared by calcining magnetite and iron(III) hydroxide in air at various temperatures, Tp, from 300 to 700°C. The samples were evacuated at outgassing temperature, To, between room temperature and 500°C at a pressure of 1 × 10?2?2.7 × 10?2N m?2 for 6 h. The heat of immersion, hi(J m?2), of samples C and M increased with an increase in To and showed the maximum hi at To =400°C, while sample F did not show the maximum up to To =500°C. The systematic correlation was not observed between hi and Tp of sample F. The heat of reproduction of the surface hydroxyl group on sample F was approximately estimated as 6.6 × 104 J mole?1 H2O.  相似文献   

5.
采用乙二胺辅助的水热法制备了纳米片聚结的Co3O4微球. 利用多种分析技术表征了其物化性质,并评价了其对甲苯燃烧的催化活性. 结果表明,由添加1.0 ml乙二胺经140 ℃水热处理12 h后制得的Co3O4样品呈纳米片聚结的微球状表面形貌. Co3O4微球样品的比表面积约为66 m2 g-1. 与体相Co3O4样品相比,Co3O4微球样品具有较高的氧吸附物种浓度和较好的低温还原性. 当空速为20000 ml g-1 h-1时,在Co3O4微球样品上甲苯转化率达到50%和90%时的反应温度分别为230和254 ℃. 这与该样品具有较大的比表面积、较高的氧吸附物种浓度和较好的低温还原性相关.  相似文献   

6.
Pyrite is considered to be the major carrier of mercury in coal. Here, the chemical characteristics of two natural pyrite samples of different weathering degrees were characterized by time-of-flight secondary ion mass spectrometry (TOF-SIMS). Thermal stability of Hg was also analyzed via temperature programmed desorption experiment (TPD). Characteristic ions such as S, Fe+, FeS, and FeS2 were detected on the surface of fresh pyrite. The release temperature of Hg ranged between 180°C and 300°C, and the characteristic peak of black HgS was recorded. In addition, abundant Fe2O3, FeSO, SO4, and HSO4 were detected on the surface of weathered pyrite, and the release temperature of Hg therein was mainly distributed at 260°C to 380°C and 520°C to 600°C, corresponding to the characteristic peaks of red HgS and HgSO4, respectively. The results show that pyrite is acidified during weathering and that Hg forms in pyrite are transformed from the original state (HgS) to HgSO4.  相似文献   

7.
Na3GdCl6: Single Crystals of the Low Temperature Form by Metallothermic Reduction of GdCl3 with Na Single crystals of Na3GdCl6-I (low-temperature form, transition to form II at 205°C) are obtained by reaction of GdCl3 with Na (tantalum tube, 700°C, 9 d). The crystal structure [a = 700.72(8), c = 1879.1(3) pm, c/a = 2,682, Vm = 160.40(3) cm3 mol?1, trigonal, R3 (No. 148), Z = 3] may be derived from the LiSbF6 type: Na2 ? Li Gd ? Sb, Cl: ¨AB¨ or h6, with two Na 1 statistically distributed over four “octahedral” in terstices.  相似文献   

8.
The influence of Bi2O3 particles size at the sub-micron scale on the thermal excitation threshold and combustion performance of nano-thermite systems was investigated. Three formulas were designed and prepared, Al(100 nm)/Bi2O3(170 nm), Al(100 nm)/Bi2O3(370 nm) and Al(100 nm)/Bi2O3(740 nm). The samples were characterized and tested by SEM, XRD, and DSC techniques. Electrical ignition and combustion experiments were performed. The results showed that with the increase of the particle size of Bi2O3, in the case of slow linear heating, the exothermic heat decreased (1051.2 J g−1, 527.3 J g−1 and 243.6 J g−1) and the thermal excitation threshold temperature increased (564.52 °C, 658.1 °C and 810.9 °C). Simultaneously, the state of the thermite reaction correspondingly changed to solid-solid, liquid-solid and liquid-liquid thermite reaction. In the case of rapid heating , the increase in particle size increased the excitation current (0.561A, 0.710A and 0.837A). During the combustion process, the thermite system with the smallest Bi2O3 particle size showed the largest combustion rate, and that with the largest particle size had the longest combustion duration.  相似文献   

9.
Two new cobalt complexes were successfully synthesized from the reaction of binaphthyl Schiff base 2 with Co(OAc)2 in the presence of sodium methoxide at 80 °C for 24 h and Co(acac)3 in toluene under reflux. Their unique crystal structures are unambiguously disclosed by X‐ray analysis. Complex 3 is triclinic, space group P1 , unit cell dimensions a = 10.742(2) Å, b = 11.153(2) Å, c = 12.715 Å, α = 79.865(3) °, β = 76.053 °, γ = 72.532(4) °, volume 1401.3(5) Å3, Z = 2. Complex 4 is triclinic, space group P1 , unit cell dimensions a = 10.801(2) Å, b = 12.554(3) Å, c = 15.219(3) Å, α = 105.672(4) °, β = 103.048 °, γ = 104.594(4) °, volume 1824.8(7) Å3, Z = 2, calculated density 1.428 Mg m−3. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

10.
Polymerization of styrene initiated by triflic acid in CH2Cl2 solution was reexamined, using a new stopped-flow device working in high purity conditions over a wide temperature range. Monomer and styryl cation were followed simultaneously through their respective absorbances at 290 and 340 nm. Initiation is very rapid, and cations concentration reaches a plateau the duration of which is depending on temperature. In our conditions (I0 = 0.5 − 9.10−3M, M0/I0 = 1 to 20), cations concentration is so low at room temperature that it is almost unmeasurable. At −65°C, it is 100 times higher, remains constant for several seconds and complete termination takes place within a minute or more. Such a profile of cation evolution agrees with an equilibrium situation between initiation and a much more temperature-dependent backward deprotonation. Apparent initial rate of initiation is first order with respect to monomer, but the order with respect to initiator was found very high and variable with temperature (from 4.5 at −65°C to 3 at −20°C). This supports the presence, even if they are in low concentration, of acid high agregates, the reactivity of which increases with size. A first order monomer consumption is observed during the plateau, which leads to kp values ranging from 103 at −65°C to 9.104 M−1.s−1 at −10°C (Ep# = 43 kJ.mol−1). The disappearance of cations, which follows the plateau, slows down and becomes unimolecular when monomer consumption is complete, and kt values range from 6.10−2s−1 at −65°C to 1.2s−1 at −23°C (Et# = 33 kJ.mol−1).  相似文献   

11.
Lead halide perovskite has triggered a lot of research due to its superior optical properties. However, halide perovskite materials have poor environmental stabilities and are easily affected by external factors such as water and heat, resulting in structural decomposition and performance failure. Contrary to this commonplace concept, it is found that CsPbBr3 (CPB) can convert to CsPb2Br5 (CP2B5) partially when meeting a small amount of water, and the CsPbBr3@CsPb2Br5 (CPB@CP2B5) composite is synthesized by an in situ method accordingly. The CPB@CP2B5 composite shows an enhanced catalytic performance compared with pure CPB, as well as a dramatically synergistic effect of photo and thermal for catalytic CO2 hydrogenation. The CO production rate of CPB@CP2B5 is determined as 69 μmol g−1 h−1 under light irradiation at 200 °C, which is 156.8 and 43.4 times higher than that under pure photo (0.44 μmol g−1 h −1) and pure thermal (1.59 μmol g−1 h −1) condition, respectively. Meanwhile, the CPB@CP2B5 sample is also stable, which shows no significant decline in the catalytic activity during 8 cycles of repeated experiments. The probable mechanism is explored by utilizing a series of in situ characterizations.  相似文献   

12.
A facile preparation of polyimide–silica gel hybrids by the simultaneous in-situ formation of polyimides during the hydrolysis–condensation of tetramethoxysilane (TMOS) is reported here. The hydrolysis and condensation of TMOS was carried out in a solution of DMAc containing 5% LiCl, CaCl2 or ZnCl2 and the seven-membered cyclic polyimide intermediate. The seven-membered cyclic intermediates, precursors of polyimides, were derived from the low-temperature polycondensation of dianhydrides [benzophenonetetracarboxylic dianhydride (BTDA), pyromellitic dianhydride (PMDA), and 4,4-bis(hexafluoroisopropylidene)phthalic dianhydride (6FDA)] and di-isocyanates [isophorone di-isocyanate (IPDI), toluene di-isocyanate (TDI), hexamethylene di-isocyanate (HDI) and 4,4′-diphenylmethane di-isocyanate (MDI)]. These intermediates could readily be converted to the corresponding polyimides. Films were cast from the resulting mixtures and the solvent was gradually evaporated at 130 °C to result in the formation of clear, transparent, pale yellow or amber-colored hybrid films in which the salts were dispersed at the molecular level. Pyrolysis of polyimide–silica gel hybrids at 600 °C gave mesoporous silica. Silica gel obtained from hybrids HPI-8 (containing no salt) and HPI-11 (containing ZnCl2) had a pore radius (BJH method) of 2.9 nm, while that from hybrid HPI-9 (containing LiCl) had a pore radius of 11.4 nm. The surface areas (BET method) obtained were 203 m2 g−1, 19 m2 g−1 and 285 m2 g−1, while the pore volumes were 0.373 cm3 g−1, 0.158 cm3 g−1 and 0.387 cm3 g−1, respectively, for samples obtained from hybrids HPI-8, HPI-9 and HPI-11. © 1997 by John Wiley & Sons, Ltd.  相似文献   

13.
The thermal dehydration of La[Co(CN)6]⋅5H2O proceeded through at least three stages from the temperature range of30~230°C, and an abrupt mass loss occurred around 350°C and the perovskite type oxide,LaCoO3 was obtained at 1000°C. After dehydration, the color of the anhydride changed from white to pale blue around 230°C and furthermore, the color changed to blue around 290°C. These color changes were discussed on the basis of the changes of coordination structures around Co ions. In La[Co(CN)6]⋅5H2O, Co3+ ions lie at the center of the Oh crystal field consisted of six CN ions. However, in the pale blue specimen, Co3+ ions were situated in the center of D4h crystal field which was distorted the Oh one by lengthening of the trans CN ions along z-axis. In the blue specimen, Co3+ ions were reduced to Co2+ ions which were situated in the Td crystal field formed by four CNions as [Co(CN)4]2–. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

14.
Propane dehydrogenation (PDH) has great potential to meet the increasing global demand for propylene, but the widely used Pt-based catalysts usually suffer from short-term stability and unsatisfactory propylene selectivity. Herein, we develop a ligand-protected direct hydrogen reduction method for encapsulating subnanometer bimetallic Pt–Zn clusters inside silicalite-1 (S-1) zeolite. The introduction of Zn species significantly improved the stability of the Pt clusters and gave a superhigh propylene selectivity of 99.3 % with a weight hourly space velocity (WHSV) of 3.6–54 h−1 and specific activity of propylene formation of 65.5 mol gPt−1 h−1 (WHSV=108 h−1) at 550 °C. Moreover, no obvious deactivation was observed over PtZn4@S-1-H catalyst even after 13000 min on stream (WHSV=3.6 h−1), affording an extremely low deactivation constant of 0.001 h−1, which is 200 times lower than that of the PtZn4/Al2O3 counterpart under the same conditions. We also show that the introduction of Cs+ ions into the zeolite can improve the regeneration stability of catalysts, and the catalytic activity kept unchanged after four continuous cycles.  相似文献   

15.
A kinetic study of the anionic polymerization of dodecamethylcyclohexasiloxane (D6) has been carried out in toluene at 25°C with the cryptate Li+ + [211] as the counterion. The rate constants of propagation of D6 and of the formation of cyclic oligomers have been determined. The results are in good agreement with those reported previously for the anionic polymerization of D3 and D4 under the same conditions. The rate constants of propagation and of formation of D6 have been found to be equal to 1.1 l-mol−1.h−1 and 0.05 h−1, respectively. In the case of the anionic polymerization of cyclosiloxanes with Li+ + [211] as the counterion, the order of reactivities of Dx as a function of × is as follows:   相似文献   

16.
A series of NiMoW/P-Al2O3 catalysts with different Mo/W ratios (sample containing Mo only, Mo/W = 2: 1, Mo/W = 1: 1, Mo/W = 1: 2, and sample containing W only; P2O5 content of the support 2.0 wt %) were synthesized. The precursors of the active phase were the heteropoly acids H3PMo12O40?nH2O and H3PW12O40?nH2O, and also nickel citrate. The sulfide phase in the samples was studied by high-resolution transmission electron microscopy and X-ray photoelectron spectroscopy; the catalytic activity of the samples in dibenzothiophene hydrodesulfurization and naphthalene hydrogenation was determined. For the dibenzothiophene hydrogenolysis in the presence of quinoline and naphthalene (content in the model mixture, wt %: dibenzothiophene 0.3, naphthalene 1.5, and quinoline 0.5), kHDS for different samples is in the range 17.6–42.5 h–1 at 275°C and 24.6–45.9 h–1 at 300°C. For the naphthalene hydrogenation, kHYD varies from 0.79 to 1.89 h–1 at 275°C and from 0.91 to 3.78 h–1 at 300°C. The sample based on molybdenum showed the highest activity in hydrogenation and hydrodesulfurization.  相似文献   

17.
The low-dimensional halide perovskites have attracted increasing attention due to their improved moisture stability, reduced defects, and suppressed ions migration in many optoelectronic devices such as solar cells, light-emitting diodes, X-ray detectors, and so on. However, they are still limited by their large band gap and short charge carriers’ diffusion length. Here, we demonstrate that the introduction of metal ions into organic interlayers of two-dimensional (2D) perovskite by cross-linking the copper paddle-wheel cluster-based lead bromide ([Cu(O2C−(CH2)3−NH3)2]PbBr4) perovskite single crystals with coordination bonds can not only significantly reduce the perovskite band gap to 0.96 eV to boost the X-ray induced charge carriers, but can also selectively improve the charge carriers’ transport along the out-of-plane direction and blocking the ions motion paths. The [Cu(O2C−(CH2)3−NH3)2]PbBr4 single-crystal device can reach a record charges/ions collection ratio of 1.69×1018±4.7 % μGyair−1 s, and exhibit a large sensitivity of 1.14×105±7% μC Gyair−1 cm−2 with the lowest detectable dose rate of 56 nGyair s−1 under 120 keV X-rays irradiation. In addition, [Cu(O2C−(CH2)3−NH3)2]PbBr4 single-crystal detector exposed to the air without any encapsulation shows excellent X-ray imaging capability with long-term operational stability without any attenuation of 120 days.  相似文献   

18.
Three-dimensionally macroporous perovskite-type oxides EuFeO3 (EFO-3DOM, EFO-sucrose-1, EFO-sucrose-2, and EFO-sucrose-3, respectively) have been prepared using the polymethyl methacrylate-templating method in the absence or presence of sucrose. Physicochemical properties of the materials were characterized by means of a number of analytical techniques, and their catalytic activities were evaluated for the total oxidation of toluene. It is shown that all of the EFO samples were of single-phase orthorhombic crystal structure with a 3DOM architecture. The sucrose addition during the preparation process had a great effect on the surface area and porous structure of the final product. A clear correlation of surface area, surface oxygen species concentration, and low-temperature reducibility with the catalytic performance was observed. The EFO-sucrose-1 catalyst performed the best, giving the T50% and T90% of 312 and 347 °C at space velocity = 20,000 mL/(g h), respectively. The apparent activation energies of the 3DOM-structured EFO samples were in the range of 82–97 kJ/mol. It is concluded that the higher surface area and oxygen adspecies concentration and better low-temperature reducibility account for the good catalytic activity of EFO-sucrose-1.  相似文献   

19.
Cobalt-free perovskite oxide La0.5Sr0.5Fe0.8Cu0.2O3  δ (LSFC) was applied as both anode and cathode for symmetrical solid oxide fuel cells (SSOFCs). The LSFC shows a reversible transition between a cubic perovskite phase in air and a mixture of SrFeLaO4, a K2NiF4-type layered perovskite oxide, metallic Cu and LaFeO3 in reducing atmosphere at elevated temperature. The average thermal expansion coefficient of LSFC in air is 17.7 × 10 6 K 1 at 25 °C to 900 °C. By adopting LSFC as initial electrodes to fabricate electrolyte supported SSOFCs, the cells generate maximum power output of 1054, 795 and 577 mW cm 2 with humidified H2 fuel (~ 3% H2O) and 895, 721 and 482 mW cm 2 with humidified syngas fuel (H2:CO = 1:1) at 900, 850 and 800 °C, respectively. Moreover, the cell with humidified H2 fuel demonstrates a reasonable stability at 800 °C under 0.7 V for 100 h.  相似文献   

20.
The molecular dimensions of polydipropylsiloxamer were studied by intrinsic viscosity measurements in toluene and in 2-pentanone. The relationships between the molecualr weight and the intrinsic viscosity were found to be: [η]25°C., toluene = 4.35 × 10?4 M0.58; [η]θ(10°C.), toluene = 1.09 × 10?3 M0.5; [η]θ(76°C.), 2-pentanone = 8.71 × 10?4 M0.5. This held reasonably well for molecular weights from 25,000 to 3000,000. The root-mean-square end-to-end length ratio, (r02 /M)1/2 as calculated from the constant K, exceeds the free rotation value by approximately 100%. The disparity is greater than that found with polydimethylsiloxamer, indicating a lower degree of flexibility for the polydipropylsiloxamer. This is largely due to the short range steric interaction between near neighboring units of the chain. Gel permeation chromatography was also employed to demonstrate the lower degree of flexibility for polydipropylsiloxamer as compared with polydimethylsiloxamer.  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