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 62 毫秒
1.
LiCoxMn1−xPO4/C nanocomposites (0 ≤ x ≤ 1.0) were prepared by a combination of spray pyrolysis at 300 °C and wet ball-milling followed by heat treatment at 500 °C for 4 h in 3% H2 + N2 atmosphere. X-ray diffraction analysis indicated that all samples had the single phase olivine structures indexed by orthorhombic Pmna. The lattice parameters linearly decreased with increasing cobalt content, which confirmed the existence of solid solutions. It was clearly seen from the scanning electron microscopy observation that the LiCoxMn1−xPO4/C samples were agglomerates with approximately 100 nm primary particles. The LiCoxMn1−xPO4/C nanocomposites were used as cathode materials for lithium batteries, and electrochemical performance was comparatively investigated with cyclic voltammetry and galvanostatic charge–discharge test using the Li?1 M LiPF6 in EC:DMC = 1:1?LiCoxMn1−xPO4/C cells at room temperature. The cells at 0.05 C charge–discharge rate delivered first discharge capacities of 165 mAh g−1 (96% of theoretical capacity) at x = 0, 136 mAh g−1 at x = 0.2, 132 mAh g−1 at x = 0.5, 125 mAh g−1 at x = 0.8 and 132 mAh g−1 (79% of theoretical capacity) at x = 1.0, respectively. While the first discharge capacity increased with the cobalt content at high charge–discharge rates more than 0.5 C due to higher electronic conductivity of LiCoPO4 in comparison with LiMnPO4, the cycleability of cell became worse with increasing the amount of cobalt. The existence of Mn2+ seemed to enhance the cycleability of LiCoxMn1−xPO4/C nanocomposite cathode.  相似文献   

2.
SmYb1−xMgxZr2O7−x/2 (0 ≤ x ≤ 0.15) ceramics are pressureless-sintered at 1973 K for 10 h in air. The structure and electrical conductivity of SmYb1−xMgxZr2O7−x/2 ceramics are investigated by the X-ray diffraction, scanning electron microscopy and impedance spectroscopy measurements. SmYb1−xMgxZr2O7−x/2 ceramics exhibit a defect fluorite-type structure. The measured electrical conductivities of SmYb1−xMgxZr2O7−x/2 ceramics obey the Arrhenius relation, and electrical conductivity of each composition increases with increasing temperature from 673 to 1173 K. At identical temperature levels, the electrical conductivity of SmYb1−xMgxZr2O7−x/2 ceramics gradually increases with increasing magnesia content. SmYb1−xMgxZr2O7−x/2 ceramics are oxide-ion conductors in the oxygen partial pressure range of 1.0 × 10−4 to 1.0 atm at all test temperature levels. The electrical conductivity obtained in SmYb1−xMgxZr2O7−x/2 ceramics reaches the highest value of 2.72 × 10−3 S cm−1 at 1173 K for the SmYb0.85Mg0.15Zr2O6.925 ceramic.  相似文献   

3.
Proton-conducting solid oxide fuel cells, incorporating BaZr0.1Ce0.7Y0.2O3−δ (BZCY) electrolyte, NiO–BZCY anode, and Sm0.5Sr0.5CoO3−δ–Ce0.8Sm0.2O2−δ (SSC–SDC) cathode, were successfully fabricated by a combined co-pressing and printing technique after a one-step co-firing process at 1100, 1150, or 1200 °C. Scanning electron microscope (SEM) results revealed that the co-firing temperature significantly affected not only the density of the electrolyte membrane but the grain size and porosity of the electrodes. Influences of the co-firing temperature on the electrochemical performances of the single cells were also studied in detail. Using wet hydrogen (2% H2O) as the fuel and static air as the oxidant, the cell co-fired at 1150 °C showed the highest maximum power density (PDmax) of 552 and 370 mW cm−2 at 700 and 650 °C, respectively, while the one co-fired at 1100 °C showed the highest PDmax of 276 and 170 mWcm−2 at 600 and 550 °C, respectively. The Arrhenius equation was proposed to analyze the dependence of the PDmax on the operating temperature, and revealed that PDmax of the cell co-fired at a lower temperature was less dependent on operating temperature. The influences of the co-firing temperature on the resistances of the single cells, which were estimated from the electrochemical impedance spectroscopy measured under open circuit conditions, were also investigated.  相似文献   

4.
Electrocatalysts of the general formula IrxRu1−xO2 were prepared using Adams’ fusion method. The crystallite characterization was examined via XRD, and the electrochemical properties were examined via cyclic voltammetry (CV) in, linear sweep voltammetry (LSV) and chronopotentiometry measurements in 0.5 M H2SO4. The electrocatalysts were applied to a membrane electrode assembly (MEA) and studied in situ in an electrolysis cell through electrochemical impedance spectroscopy (EIS) and stationary current density–potential relations were investigated. The IrxRu1−xO2 (x = 0.2, 0.4, 0.6) compounds were found to be more active than pure IrO2 and more stable than pure RuO2. The most active electrocatalyst obtained had a composition of Ir0.2Ru0.8O2. With an Ir0.2Ru0.8O2 anode, a 28.4% Pt/C cathode and the total noble metal loading of 1.7 mg cm−2, the potential of water electrolysis was 1.622 V at 1 A cm−2 and 80 °C.  相似文献   

5.
A new series of rare earth solid solutions Sc2−xYxW3O12 was successfully synthesized by the conventional solid-state method. Effects of doping ion yttrium on the crystal structure, morphology and thermal expansion property of as-prepared Sc2−xYxW3O12 ceramics were investigated by X-ray diffraction (XRD), thermogravimetric analysis (TG), field emission scanning electron microscope (FE-SEM) and thermal mechanical analyzer (TMA). Results indicate that the obtained Sc2−xYxW3O12 samples with Y doping of 0≤x≤0.5 are in the form of orthorhombic Sc2W3O12-structure and show negative thermal expansion (NTE) from room temperature to 600 °C; while as-synthesized materials with Y doping of 1.5≤x≤2 take hygroscopic Y2W3O12·nH2O-structure at room temperature and exhibit NTE only after losing water molecules. It is suggested that the obvious difference in crystal structure leads to different thermal expansion behaviors in Sc2−xYxW3O12. Thus it is proposed that thermal expansion properties of Sc2−xYxW3O12 can be adjusted by the employment of Y dopant; the obtained Sc1.5Y0.5W3O12 ceramic shows almost zero thermal expansion and its average linear thermal expansion coefficient is −0.00683×10−6 °C−1 in the 25–250 °C range.  相似文献   

6.
NO decomposition in solid electrolyte cells was investigated in the presence of excess O2. The results show that NO is decomposed via an electrocatalytic mechanism rather than electrolysis in the range of 1–4 V of applied voltage. The NO is catalytically decomposed to N2 on the cathode surface and O2– produced in situ is transferred through the yttria-stabilized zirconia (YSZ) to the anode by direct current (d.c.) and then is evolved in the form of O2, which helps to maintain the active state of the cathode. In a Pd/YSZ/Pd cell, the palladium metal surface is the active site for NO decomposition, while in the RuO2/Pd/YSZ/Pd cell, the partially reduced RuO x (0 < x < 2) is the main active site for NO decomposition. At 600 °C, the rate-determining step for the overall transportation of O2– from cathode to anode in the RuO2/Pd/YSZ/Pd cell is the transportation of O2– at the cathode Pd/YSZ interface. The transportation rate of O2– at the cathode M/YSZ interface decreases in the order of Ag > Au > Pd > Pt. Substitution of the Pd cathode by Ag leads to an increase in current density by a factor of 3.5. A higher NO decomposition parameter (=13.4) is also achieved at a lower temperature of 500 °C.  相似文献   

7.
This paper reports on the composition and flow rate of outlet gas and current density during the reforming of CH4 with CO2 using three different electrochemical cells: cell A, with Ni−GDC (Gd-doped ceria: Ce0.8Gd0.2O1.9) cathode/porous GDC electrolyte/Cu−GDC anode, cell B, with Cu−GDC cathode/ porous GDC electrolyte/Cu−GDC anode and cell C, with Ru−GDC cathode/ porous GDC electrolyte/ Cu−GDC anode. In the cathode, CO2 reacts with supplied electrons to form CO fuel and O2− ions (CO2+2e→CO+O2−). Too low affinity of Cu cathode to CO2 in cell B reduced the reactivity of the CO2 with electrons. The CO fuel, O2− ions and CH4 gas were transported to the anode through the porous GDC mixed conductor of O2− ions and electrons. In the anode, CH4 reacts with O2− ions to produce CO and H2 fuels (CH4+O2−→2 H2+CO+2e). The reforming efficiency at 700−800 °C was lowest in cell B and highest in cell A. The Cu anode in cells A and C worked well to oxidize CH4 with O2− ions (2Cu+O2−→Cu2O+2e, Cu2O+CH4→2Cu+CO+2H2). However, a blockage of the outlet gas occurred in all the cells at 700−800 °C. The gas flow is inhibited due to a reduction in pore size in the cermet cathode, as well as sintering and grain growth of Cu metal in the anode during the reforming.  相似文献   

8.
The electrochemical properties of Sr1−xCexMnO3 (SCM, 0.1≤x≤0.4)–Gd0.2Ce0.8O2−x (GDC) composite cathodes were determined by impedance spectroscopy. The study focused on the doping effect of Ce in the composite cathodes. Single-phase perovskite was obtained for 0.1≤x≤0.3 in SCM. No reaction occurred between the Sr0.7Ce0.3MnO3 electrode and the GDC electrolyte at an operating temperature of 800 °C for 100 h. In the single phase perovskite region, lattice expansion occurred due to the reduction of Mn4+ to Mn3+ at B-sites, and this was attributed to an increase in Ce content. Ce doping enhanced the electrode performance of SCM–GDC composite cathodes, and best electrode performance was achieved for the Sr0.7Ce0.3MnO3–GDC composite cathode (0.93 Ω cm2 and 0.47 Ω cm2 at 750 °C and 800 °C, respectively). The improvement in electrode performance was attributed to increases in charge carriers induced by a shift of some Mn from +4 to +3 and to the formation of surface oxygen vacancies caused by Mn4+ to Mn3+ conversion at high temperatures.  相似文献   

9.
In this work, a series of Fe3−xTixO4 (0 ≤ x ≤ 0.78) was synthesized using a new soft chemical method. The synthetic Fe3−xTixO4 were characterized using X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), Mössbauer spectroscopy, thermogravimetric and differential scanning calorimetry (TG–DSC) analyses. The results showed that they were spinel structures and Ti was introduced into their structures.Then, decolorization of methylene blue (MB) by Fe3−xTixO4 in the presence of H2O2 at neutral pH values was studied using UV–vis spectra, dissolved organic carbon (DOC) and element C analyses. Furthermore, the degradation products remained in reaction solution after the decolorization were identified using ionic chromatography (IC), 13C nuclear magnetic resonance spectra (NMR), liquid chromatography and mass spectrometry (LC–MS). Although small amounts of MB were mineralized, the aromatic rings in MB were destroyed completely after the decolorization. Decolorization of MB by Fe3−xTixO4 in the presence of H2O2 was promoted remarkably with the increase of Ti content in Fe3−xTixO4 due to the enhancement of both adsorption and degradation of MB on Fe3−xTixO4.  相似文献   

10.
The NO storage properties of MnO x /support materials (5–50 wt% MnO x loading) was experimentally investigated in the presence of O2 and H2O between 50 and 700 °C applying a non-isothermal temperature-programmed method. In dependence on MnO x loading and NO supply, the materials show an intermediate decrease of NO storage capacity between 200 and 300 °C. This effect is caused by decomposition of surface nitrites with release of NO into the gas phase as proved by in situ DRIFT measurement. The interpretation is corroborated by modelling of the underlying adsorption/desorption reaction steps, considering the different thermal stability of nitrite/nitrate surface species.  相似文献   

11.
Powders of spinel Li4Ti5−xVxO12 (0 ≤ x ≤ 0.3) were successfully synthesized by solid-state method. The structure and properties of Li4Ti5−xVxO12 (0 ≤ x ≤ 0.3) were examined by X-ray diffraction (XRD), Raman spectroscopy (RS), scanning electronic microscope (SEM), galvanostatic charge–discharge test and cyclic voltammetry (CV). XRD shows that the V5+ can partially replace Ti4+ and Li+ in the spinel and the doping V5+ ion does almost not affect the lattice parameter of Li4Ti5O12. Raman spectra indicate that the Raman bands corresponding to the Li–O and Ti–O vibrations have a blue shift due to the doping vanadium ions, respectively. SEM exhibits that Li4Ti5−xVxO12 (0.05 ≤ x ≤ 0.25) samples have a relative uniform morphology with narrow size distribution. Charge–discharge test reveals that Li4Ti4.95V0.05O12 has the highest initial discharge capacity and cycling performance among all samples cycled between 1.0 and 2.0 V; Li4Ti4.9V0.1O12 has the highest initial discharge capacity and cycling performance among all samples cycled between 0.0 and 2.0 V or between 0.5 and 2.0 V. This excellent cycling capability is mainly due to the doping vanadium. CV reveals that electrolyte starts to decompose irreversibly below 1.0 V, and SEI film of Li4Ti5O12 was formed at 0.7 V in the first discharge process; the Li4Ti4.9V0.1O12 sample has a good reversibility and its structure is very advantageous for the transportation of lithium-ions.  相似文献   

12.
J. Jiang 《Electrochimica acta》2005,50(24):4778-4783
Samples of the layered cathode materials, Li[NixLi(1/3−2x/3)Mn(2/3−x/3)]O2 (x = 1/12, 1/4, 5/12, and 1/2), were synthesized at 900 °C. Electrodes of these samples were charged in Li-ion coin cells to remove lithium. The charged electrode materials were rinsed to remove the electrolyte salt and then added, along with EC/DEC solvent or 1 M LiPF6 EC/DEC, to stainless steel accelerating rate calorimetry (ARC) sample holders that were then welded closed. The reactivity of the samples with electrolyte was probed at two states of charge. First, for samples charged to near 4.45 V and second, for samples charged to 4.8 V, corresponding to removal of all mobile lithium from the samples and also concomitant release of oxygen in a plateau near 4.5 V. Li[NixLi(1/3−2x/3)Mn(2/3−x/3)]O2 samples with x = 1/4, 5/12 and 1/2 charged to 4.45 V do not react appreciably till 190 °C in EC/DEC. Li[NixLi(1/3−2x/3)Mn(2/3−x/3)]O2 samples charged to 4.8 V versus Li, across the oxygen release plateau, start to significantly react with EC/DEC at about 130 °C. However, their high reactivity is similar to that of Li0.5CoO2 (4.2 V) with 1 μm particle size. Therefore, Li[NixLi(1/3−2x/3)Mn(2/3−x/3)]O2 samples showing specific capacity of up to 225 mAh/g may be acceptable for replacing LiCoO2 (145 mAh/g to 4.2 V) from a safety point of view, if their particle size is increased.  相似文献   

13.
The deactivation of a Pt/Ba/Al2O3 NO x -trap model catalyst submitted to SO2 treatment and/or thermal ageing at 800 °C was studied by H2 temperature programmed reduction (TPR), X-ray diffraction (XRD) and NO x storage capacity measurements.The X-ray diffractogram of the fresh sample exhibits peaks characteristic for barium carbonate. Thermal ageing leads to the decomposition of barium carbonate and to the formation of BaAl2O4. The TPR profile of the sulphated sample shows the presence of (i) surface aluminium sulphates, (ii) surface barium sulphates, (iii) bulk barium sulphates. The exposure to SO2 after ageing leads to a small decrease of the surface barium-based sulphates, expected mainly as aluminate barium sulphates. This evolution can be attributed to a sintering of the storage material. TPR experiments also show that thermal treatment at 800 °C after the exposure to SO2 involves the decomposition of aluminium surface sulphates to give mainly bulk barium sulphates, also pointed out by XRD. Thus, the thermal treatment at 800 °C leads to a stabilization of the sulphates.These results are in accordance with the NO x storage capacity measurements. On non-sulphated catalysts, the treatment at 800 °C induces to a decrease of the NO x storage capacity, showing that barium aluminate presents a lower NO x storage capacity than barium carbonate. Sulphation strongly decreases the NO x storage capacity of catalysts, whatever the initial thermal treatment, showing that barium sulphates inhibit the NO2 adsorption. Moreover, the platinum activity for the NO to NO2 oxidation is lowered by thermal treatments.  相似文献   

14.
Li3xLa2/3−xTiO3 (LLTO) powder with different lithium contents (nominal 3x = 0.03–0.75) was synthesized via a simple sol–gel route and then calcination of gel-derived precursor at 900 °C which was much below the calcination temperature required for synthesizing the LLTO powder via solid state reaction route. The LLTO powder of sub-micron sized particles, derived from such sol–gel method, showed almost no aggregation. Starting from the sol–gel-derived powder, the LLTO ceramics with different lithium contents were prepared at different sintering temperatures of 1250 and 1350 °C. It demonstrated that our sol–gel route is quite simple and convenient compared to the previous sol–gel method and requires lower temperature for the LLTO. Our results also illustrated that lithium content significantly affects the structure and ionic conductivity of the LLTO ceramics. The dependence of the ionic conductivity on the lithium content, lattice structure, microstructure and sintering temperature was investigated systematically.  相似文献   

15.
Pt on ceria (CeOx) particles supported on carbon black (CB) were synthesized using the combined process of hot precipitation and impregnation methods. During 30 cycles of cyclic voltammetry pre-treatment in the potential ranging from −0.2 to 1.3 V (V vs. Ag/AgCl), it was observed that a small amount of CeOx, which consisted of the interface region between Pt and CeOx, remained on Pt particles. Other free CeOx particles were dissolved into H2SO4 aqueous solution. To develop the Pt-CeOx/CB catalyst, the surface chemical states, the net chemical composition, morphology and electrochemical behavior in H2SO4 aqueous solution were characterized. Our microanalysis and electrochemical analysis indicate that the active CeO2 with high specific surface area provides the continuous amorphous cerium oxide (Ce3+, Ce4+) layer with pores on the surface of Pt particles. It is concluded that the amorphous cerium oxide layer on Pt inhibits the oxidation of Pt surface and contributes to enhancement of the activity on Pt cathode. The single cell performance was also improved using the Pt-CeOx/CB cathode. Based on all data, it is expected that the design based on characterization of the interface between Pt and small amount of amorphous cerium oxide layer could help in preparation of more active Pt catalyst.  相似文献   

16.
YBaCo3ZnO7 + Gd0.2Ce0.8O1.9 (GDC) composites with various GDC contents (0-70 wt.%) have been investigated as cathode materials for intermediate temperature solid oxide fuel cells (SOFC). The effect of GDC incorporation on the microstructure, electrochemical properties, and thermal expansion behavior of the YBaCo3ZnO7 + GDC composites has been studied. The composite cathodes consist of smaller particles with larger surface area compared to the pure YBaCo3ZnO7 cathode, which is beneficial for providing extended triple-phase boundary (TPB) where the oxygen reduction reaction (ORR) occurs. Among the various compositions investigated, the YBaCo3ZnO7 + GDC (50:50 wt.%) composite is found to be optimum with the lowest polarization resistance (0.28 Ω cm2 at 600 °C) compared to that of pure YBaCo3ZnO7 (0.62 Ω cm2 at 600 °C). Anode-supported single cell SOFC fabricated with the YBaCo3ZnO7 + GDC (50:50 wt.%) composite cathode also exhibits excellent performance with a maximum power density of 743 mW/cm2 at 750 °C. Additionally, the YBaCo3ZnO7 + GDC (50:50 wt.%) composite shows a low thermal expansion coefficient (TEC) of 10.7 × 10−6 °C−1, which provides good compatibility with those of standard SOFC electrolytes.  相似文献   

17.
LiNi1−xCoxO2 (x = 0, 0.1, 0.2) cathode materials were successfully synthesized by a rheological phase reaction method with calcination time of 0.5 h at 800 °C. All obtained powders are pure phase with α-NaFeO2 structure (R-3m space group). The samples deliver an initial discharge capacity of 182, 199 and 189 mAh g−1 (25 mA g−1, 4.35-3.0 V), respectively. The reaction mechanism was also discussed, which consists of a series of defect reactions. As a result of these defect reactions, the reaction of forming LiNi1−xCoxO2 takes place in high speed.  相似文献   

18.
Cobalt ferrite nanoparticles (CoxFe3−xO4) and chitosan (CS) film were used to immobilize/adsorb hemoglobin (Hb) to create a protein electrode to study the direct electron transfer between the redox centers of the proteins and the electrode. X-ray diffraction (XRD) and transmission electron microscopy (TEM) revealed that the CoxFe3−xO4 particles were nanoscale in size and formed an ordered layered structure. The native structure of the immobilized Hb was preserved as indicated by Fourier-transform infrared (FTIR) and UV–visible (UV–vis) spectroscopy. The Hb-CoxFe3−xO4–CS modified electrode showed a pair of well-defined and quasi-reversible cyclic voltammetric peaks at −0.373 V (vs. SCE) and exhibited appreciable electrocatalytic activity for the reduction of H2O2. The catalysis currents increased linearly with H2O2 concentration in a wide range of 5.0 × 10−8 to 1.0 × 10−3 mol L−1 with a detection limit of 1.0 × 10−8 mol L−1 (S/N = 3) and had long-term stability. Finally, the proposed method was applied to investigate the coexistence of hydrogen peroxide with the interfering substances. Experimental results showed that the ascorbic acid, glucose, l-cysteine, uric acid, and dopamine at corresponding concentrations did not influence the detection of H2O2.  相似文献   

19.
Electrochemical and thermal properties of Co3(PO4)2- and AlPO4-coated LiNi0.8Co0.2O2 cathode materials were compared. AlPO4-coated LiNi0.8Co0.2O2 cathodes exhibited an original specific capacity of 170.8 mAh g−1 and had a capacity retention (89.1% of its initial capacity) between 4.35 and 3.0 V after 60 cycles at 150 mA g−1. Co3(PO4)2-coated LiNi0.8Co0.2O2 cathodes exhibited an original specific capacity of 177.6 mAh g−1 and excellent capacity retention (91.8% of its initial capacity), which was attributed to a lithium-reactive Co3(PO4)2 coating. The Co3(PO4)2 coating material could react with LiOH and Li2CO3 impurities during annealing to form an olivine LixCoPO4 phase on the bulk surface, which minimized any side reactions with electrolytes and the dissolution of Ni4+ ions compared to the AlPO4-coated cathode. Differential scanning calorimetry results showed Co3(PO4)2-coated LiNi0.8Co0.2O2 cathode material had a much improved onset temperature of the oxygen evolution of about 218 °C, and a much lower amount of exothermic-heat release compared to the AlPO4-coated sample.  相似文献   

20.
The NO x storage performance at low temperature (100–200 °C) has been studied for model NO x storage catalysts. The catalysts were prepared by sequentially depositing support, metal oxide and platinum on ceramic monoliths. The support material consisted of acidic aluminium silicate, alumina or basic aluminium magnesium oxide, and the added metal oxide was either ceria or barium oxide. The NO x conversion was evaluated under net-oxidising conditions with transients between lean and rich gas composition and the NO x storage performance was studied by isothermal adsorption of NO2 followed by temperature programmed desorption of adsorbed species. The maximum in NO x storage capacity was observed at 100 °C for all samples studied. The Pt/BaO/Al2O3 catalyst stored about twice the amount of NO x compared with the Pt/Al2O3 and Pt/CeO2/Al2O3 samples. The storage capacity increased with increasing basicity of the support material, i.e. Pt/Al2O3·SiO2 < Pt/Al2O3 < Pt/Al2O3 · MgO. Water did not significantly affect the NO x storage performance for Pt/Al2O3 or Pt/BaO/Al2O3.  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