首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 187 毫秒
1.
The surface composition of anodically oxidized SIMFUEL (doped uranium dioxide) has been determined as a function of applied potential over the range −500 to +500 mV (versus SCE). Cathodically cleaned UO2 specimens were anodically oxidized for 1 h and subsequently analyzed by XPS. Using published binding energies, the U (4f7/2) and O (1s) peaks were resolved into contributions from UIV, UV, UVI, OII, OH and H2O. It was shown that over the potential range −500 to approximately +50 mV a thin surface layer of UIV/UV oxide (UO2+x) formed. At more positive potentials, a UVI hydrated oxide (UO3·yH2O) was deposited on the electrode surface. At very positive potentials (≥400 mV) the rapid anodic formation and hydrolysis of UO22+ led to local acidification in pores in the deposited UO3·yH2O layer and its subsequent re-dissolution.  相似文献   

2.
The reaction of H2O2 on SIMFUEL electrodes has been studied electrochemically and under open circuit conditions in 0.1 mol l−1 NaCl (pH 9.8). The composition of the oxidized UO2 surface was determined by X-ray photoelectron spectroscopy (XPS). Peroxide reduction was found to be catalyzed by the formation of a mixed UIV/UV (UO2+x) surface layer, but to be blocked by the formation of UVI (UO22+) species on the electrode surface. The formation of this UVI layer blocks both H2O2 reduction and oxidation, thereby inhibiting the potentially rapid H2O2 decomposition process to H2O and O2. Decomposition is found to proceed at a rate controlled by desorption or reduction of the adsorbed O2 species. Reduction of O2 is coupled to the slow oxidative dissolution of UO2 and formation of a corrosion product deposit of UO3·yH2O.  相似文献   

3.
The electrodeposition of MoxRe1−xOy films (0.6 ≤ x ≤ 1) on indium-tin oxide (ITO) coated glass substrates from acidic peroxo-polymolybdo-perrhenate solutions is described. Trends in film growth were established as a function of potential from +0.4 V to −0.7 V vs Ag/AgCl by analyzing the composition and stoichiometry of the deposit using inductively coupled plasma mass spectrometry (ICPMS) and X-ray photoelectron spectroscopy (XPS). These experiments show that the concentration of rhenium increases linearly with the deposition potential and that the deposits are mixed-valent containing up to five different metal oxidation states (i.e., MoIV, MoV, MoVI, Re0, ReIV). Electroanalytical techniques were used to explore the deposition mechanism, including chronocoulometry, cyclic voltammetry, spectroelectrochemistry, and electrochemical quartz crystal nanogravimetry (EQCN). At potentials positive to −0.26 V, perrhenate (ReVIIO4) behaves as a redox mediator to accelerate the deposition of a mixed-valent molybdenum oxide, but at more negative potentials mixed molybdenum-rhenium oxides are produced.  相似文献   

4.
LiNi1−yCoyO2 (y=0.1, 0.3 and 0.5) were synthesized by solid state reaction method at 800 °C and 850 °C from LiOH·H2O, NiO and Co3O4 as starting materials. The electrochemical properties of the synthesized LiNi1−yCoyO2 were investigated. As the content of Co decreases, particle size decreases rapidly and particle size distribution gets more homogeneous. When the particle size is compared at the same composition, the particles synthesized at 850 °C are larger than those synthesized at 800 °C. LiNi0.7Co0.3O2 synthesized at 850 °C has the largest intercalated and deintercalated Li quantity Δx among LiNi1−yCoyO2 (y=0.1, 0.3 and 0.5). LiNi0.7Co0.3O2 synthesized at 850 °C has the largest first discharge capacity (178 mAh/g), followed by LiNi0.7Co0.3O2 (162 mAh/g) synthesized at 800 °C. LiNi0.7Co0.3O2 synthesized at 800 °C has discharge capacities of 162 and 125 mAh/g at n=1 and n=5, respectively.  相似文献   

5.
Uranium–neodymium mixed oxides (U1−yNdy)Ox (y=0.2–0.85) were prepared by citrate gel-combustion and characterized by XRD. Single phase fluorite structure was observed up to y=0.80. For solid solutions with y>0.80 additional lines pertaining to hexagonal neodymium oxide were observed. Lattice thermal expansion of these samples was investigated by using high temperature X-ray diffraction (HTXRD). The coefficients of thermal expansion for (U1−yNdy)Ox for y=0.2, 0.4, 0.6, and 0.8 in the temperature range 298–1973 K were found to be 16.46, 16.64, 16.79, and 16.89×10−6 K−1, respectively. Heat capacity and enthalpy increment measurements were carried out by using DSC and drop calorimetry in the temperature range 298–800 K and 800–1800 K respectively. The Cp,m values at 298 K for (U1−yLay)Ox (y=0.2, 0.4, 0.6, and 0.8) are 63.4, 64.3, 61.8, and 58.9 J K−1 mol−1 respectively.  相似文献   

6.
Preparation of anodes for oxygen evolution in seawater electrolysis was carried out. Manganese-molybdenum double oxides, Mn1−xMoxO2+x, prepared by anodic deposition from MnSO4-Na2MoO4 solutions showed the 100% oxygen evolution efficiency at a current density of 1000 A m−2 in 0.5 M NaCl at 30 °C and pH 12, but an increase in solution temperature resulted in dissolution of the oxides as molybdate and permanganate ions. In order to increase the stability of the electrodes at higher temperatures the addition of iron to the manganese-molybdenum oxides was performed by anodic deposition in MnSO4-Na2MoO4-FeNH4(SO4)2 solutions. The electrodes thus prepared showed the 100% oxygen evolution efficiency at 1000 A m−2 in 0.5 M NaCl at 30-90 °C, when proper amounts of molybdenum and iron were contained. The iron addition also enhanced the oxygen evolution efficiency. The electrodes were not composed of oxide mixtures but triple oxides, Mn1−xyMoxFeyO2+x−0.5y, consisting of Mn4+, Mo6+ and Fe3+. The formation of the triple oxides seemed responsible for enhancement of both oxygen evolution efficiency and stability.  相似文献   

7.
The corrosion of nuclear fuel under waste disposal conditions is likely to be influenced by the bicarbonate/carbonate content of the groundwater since it increases the solubility of the UVI corrosion product, [UO2]2+. As one of the half reactions involved in the corrosion process, the anodic dissolution of SIMFUEL (UO2) has been studied in bicarbonate/carbonate solutions (pH 9.8) using voltammetric and potentiostatic techniques and electrochemical impedance spectroscopy. The reaction proceeds by two consecutive one electron transfer reactions (UIV → UV → UVI). At low potentials (≤250 mV (vs. SCE) the rate of the first electron transfer reaction is rate determining irrespective of the total carbonate concentration. At potentials >250 mV (vs. SCE) the formation of a UVIO2CO3 surface layer begins to inhibit the dissolution rate and the current becomes independent of potential indicating rate control by the chemical dissolution of this layer.  相似文献   

8.
9.
In this study, single crystal V3O7·H2O nanobelts were successfully synthesized using a simple hydrothermal route, in which templates or catalysts were absent. The synthesized V3O7·H2O nanobelts are highly crystalline and have lengths up to several tens of micrometers. The width and thickness of the nanobelts are found to be about 30-50 and 30 nm, respectively. A lithium battery using V3O7·H2O nanobelts as the positive electrode exhibits a high initial discharge capacity of 409 mAh g−1, corresponding to the formation of LixV3O7·H2O (x = 4.32). Such a high degree of electrochemical performance is attributed to the intrinsic properties of the single-crystalline V3O7·H2O nanobelts.  相似文献   

10.
A systematic investigation was conducted of the surface properties and the HER at electrodes of nominal composition Ti/RhxTi(1−x)Oy prepared by thermal decomposition (Tcal: 500 °C; tcal: 2 h; O2 flux: 5 dm3 min−1) from salt precursor solutions dissolved in 6.0 mol dm−3 HNO3. Films were characterized ex situ by SEM, EDX, XPS and XRD and in situ by open circuit potential measurements and CV. The electrochemical behaviour was investigated by CV as function of the anodic, Eλ,a, and cathodic, Eλ,c, switching potentials showing the Rh surface oxidation states strongly depend on these experimental variables. Surface Rh-sites are reduced to metallic rhodium in the cathodic potential region while higher oxidation states (I-III) are formed at more positive potentials (E ≥ 0.5 V/RHE). Hydrogen adsorption and desorption peaks as well as a short double layer charging region are observed at intermediate potential values. The HER was investigated by Tafel coefficients and reaction order with respect to H+ as function of nominal Rh-content.  相似文献   

11.
In order to produce thin films of crystalline V2O5, vanadium metal was thermally oxidised at 500 °C under oxygen pressures between 250 and 1000 mbar for 1-5 min. The oxide films were characterised by X-ray photoelectron spectroscopy (XPS), atomic force microscopy (AFM), X-ray diffraction (XRD) and Rutherford backscattering spectrometry (RBS). The lithium intercalation performance of the oxide films was investigated by cyclic voltammetry (CV), chronopotentiometry and electrochemical impedance spectroscopy (EIS). It was shown that the composition, the crystallinity and the related lithium intercalation properties of the thin oxide films were critically dependent on the oxidation conditions. The formation of crystalline V2O5 films was stimulated by higher oxygen pressure and longer oxidation time. Exposure for 5 min at 750 mbar O2 at 500 °C resulted in a surface oxide film composed of V2O5, and consisting of crystallites up to 200 nm in lateral size. The thickness of the layer was about 100 nm. This V2O5 oxide film was found to have good cycling performance in a potential window between 3.8 and 2.8 V, with a stable capacity of 117 ± 10 mAh/g at an applied current density of 3.4 μA/cm2. The diffusion coefficients corresponding to the two plateaus at 3.4 and 3.2 V were determined from the impedance measurements to (5.2 and 3.0) × 10−13 cm2 s−1, respectively. Beneath the V2O5 layer, lower oxides (mainly VO2) were found close to the metal. At lower oxygen pressure and shorter exposure times, the oxide films were less crystalline and the amount of V4+ increased in the surface oxide film, as revealed by XPS. At intermediate oxygen pressures and exposure times a mixture of crystalline V2O5 and V6O13 was found in the oxide film.  相似文献   

12.
The influence of iron doping level in Ba0.5Sr0.5Co1−yFeyO3−δ (y = 0.0-1.0) (BSCF) oxides on their phase structure, oxygen nonstoichiometry, electrical conductivity, performance as symmetrical cell electrode and oxygen permeating membranes was systematically investigated. A cubic perovskite structure was observed for all the compositions with the presence of iron. The increase of iron doping level resulted in the decrease of the lattice constant, room-temperature oxygen nonstoichiometry, total electrical conductivity, and the increase of area specific resistance (ASR) as cathode with samaria doped ceria electrolyte. However, promising cathode performance with an ASR as low as 0.613 Ω cm2 was still obtained at 600 °C for Ba0.5Sr0.5FeO3−δ (BSF). The ceramic membranes composing of BSCF with various iron doping level are all oxygen semi-permeable at elevated temperatures. The increase of iron doping level resulted in the decrease of oxygen permeation flux from JO2 = 2.28 μmol cm−2 s−1 (STP) for Ba0.5Sr0.5Co0.8Fe0.2O3−δ (BSCF5582) to ∼0.45 μmol cm−2 s−1 (STP) at 900 °C for BSF (y = 1.0) with the same membrane thickness of 1.1 mm, alongside with the change of the rate-determination step from the oxygen surface exchange to the slow oxygen bulk diffusion. The formation of composite oxide with a proper electronic conducting phase and the thin film technology are important for their prospective application as cathode in IT-SOFCs and oxygen permeating membrane, respectively.  相似文献   

13.
LiNiO2 and LiNi1−yMyO2 (M = Zn and Ti, y = 0.005, 0.01, 0.025, 0.05, and 0.1) were synthesized with a solid-state reaction method by calcination at 750 °C for 30 h under oxygen stream after preheating at 450 °C for 5 h in air. LiNi0.995Zn0.005O2 among the Zn-substituted samples and LiNi0.995Ti0.005O2 among the Ti-substituted samples showed the best electrochemical properties. For similar values of y, LiNi1−yTiyO2 had in general better electrochemical properties than LiNi1−yZnyO2. Electrochemical properties seem to be closely related to R-factor but less related to I0 0 3/I1 0 4 value. In the FT-IR absorption spectra of LiNiO2 and LiNi1−yMyO2 (M = Zn and Ti, y = 0.005, 0.01, 0.025, 0.05 and 0.1), Li2CO3 was detected even if it is not observed from XRD pattern, with the samples LiNi1−yZnyO2 (y = 0.05 and 0.1) showing Li2ZnO2 additionally. The smaller cation mixing of the Ti-substituted samples is considered to lead to their better electrochemical properties than the Zn-substituted samples.  相似文献   

14.
This work demonstrates that anodic deposition of vanadium oxide (denoted as VOx·nH2O) can be considered as the chemical co-precipitation of V5+ and V4+ oxy-/hydroxyl species and the accumulation of V5+ species at the vicinity of electrode surface is the key factor for the successful anodic deposition of VOx·nH2O at a potential much more negative than the equilibrium potential of the oxygen evolution reaction (OER). The results of in situ UV-vis spectra show that the V4+/V5+ ratio near the electrode surface can be controlled by varying the applied potential, leading to different, three-dimensional (3D) nanostructures of VOx·nH2O. The accumulation of V5+ species due to V4+ oxidation at potentials ≥0.4 V (vs. Ag/AgCl) has been found to be very similar to the phenomenon by adding H2O2 in the deposition solution. The X-ray photoelectron spectroscopic (XPS) results show that all VOx·nH2O deposits can be considered as aggregates consisting of mixed V5+ and V4+ oxy-/hydroxyl species with the mean oxidation state significantly increasing with the applied electrode potential.  相似文献   

15.
Amorphous Ru1−yCryO2/TiO2 nanotube composites were synthesized by loading different amount of Ru1−yCryO2 on TiO2 nanotubes via a reduction reaction of K2Cr2O7 with RuCl3·nH2O at pH 8, followed by drying in air at 150 °C. Cyclic voltammetry and galvanostatic charge/discharge tests were applied to investigate the performance of the Ru1−yCryO2/TiO2 nanotube composite electrodes. For comparison, the performance of amorphous Ru1−yCryO2 was also studied. The results demonstrated that the three dimensional nanotube network of TiO2 offered a solid support structure for active materials Ru1−yCryO2, allowed the active material to be readily available for electrochemical reactions, and increased the utilization of active materials. A maximum specific capacitance 1272.5 F/g was obtained with the proper amount of Ru1−yCryO2 loaded on the TiO2 nanotubes.  相似文献   

16.
A freeze-drying precursor method was used to obtain submicrometric powders of ceria-based materials such as Ce1−xGdxO2−δ (x=0, 0.01, 0.05, 0.10 and 0.20), 80%CeO2–20%ZrO2, 80%CeO2–20%Al2O3 and (1−y)Ce0.99Gd0.01O2−δ– (y)Al2O3 (y=0.01, 0.02, 0.05, 0.10 and 0.30) at temperatures as low as 400 °C. The phase formation and evolution with the temperature was studied by X-ray diffraction (XRD). Also, temperature programmed reduction (TPR) was performed to investigate the reducibility of the ceramic powders. It was observed that after reduction of the ceria-based materials the fluorite structure of the samples was retained. The TPR profiles showed two peaks which are associated to the surface and bulk ceria reduction processes. Likewise, after the TPR measurements the resulting powders have showed high phase stability and reproducibility. XPS results confirmed the reduction of Ce4+ to (Ce3++Ce4+) ratio with alumina doping.  相似文献   

17.
The Ca3−xB2O6:xDy3+ (0.0 ≤ x ≤ 0.105) and Ca2.95−yDy0.05B2O6:yLi+ (0 ≤ y ≤ 0.34) phosphors were synthesized at 1100 °C in air by solid-state reaction route. The as-synthesized phosphors were characterized by X-ray powder diffraction (XRD), scanning electron microscope (SEM), photoluminescence excitation (PLE) and photoluminescence (PL) spectra. The PLE spectra show the excitation peaks from 300 to 400 nm is due to the 4f-4f transitions of Dy3+. This mercury-free excitation is useful for solid state lighting and light-emitting diodes (LEDs). The emission of Dy3+ ions upon 350 nm excitation is observed at 480 nm (blue) due to the 4F9/2 → 6H15/2 transitions, 575 nm (yellow) due to 4F9/2 → 6H13/2 transitions and a weak 660 nm (red) due to 4F9/2 → 6H11/2 emissions, respectively. The optimal PL intensity of the Ca3−xB2O6:xDy3+ phosphors is found to be x = 0.05. Moreover, the PL results from Ca2.95−yDy0.05B2O6:yLi+ phosphors show that Dy3+ emissions can be enhanced with the increasing codopant Li+ content till y = 0.22. By simulation of white light, the CIE of the investigated phosphors can be tuned by varying the content of Li+ ions, and the optimal CIE value (0.300, 0.298) is realized when the content of Li+ ions is y = 0.22. All the results imply that the Ca2.95−yDy0.05B2O6:yLi+ phosphors could be potentially used as white LEDs.  相似文献   

18.
Ba1−xPrxCo1−yFeyO3−δ (BPCF) perovskite oxides have been synthesized and investigated as cathode materials for low temperature solid oxide fuel cells (LT-SOFCs). Compared with those of Ba0.5Sr0.5Co0.8Fe0.2O3−δ (BSCF) and Sm0.5Sr0.5CoO3 (SSCo) cathode materials, BPCF has a lower polarization resistance at decreased temperatures. In particular, Ba0.5Pr0.5Co0.8Fe0.2O3−δ showed the lowest polarization loss among the different compositions as a cathode material for LT-SOFCs. The area specific resistance (ASR) of Ba0.5Pr0.5Co0.8Fe0.2O3−δ as a cathode material is 0.70 and 0.185 Ω cm2 at 500 °C and 550 °C, respectively. The maximum power density of the cell BPCF/SDC/Ni-SDC with humidified hydrogen as fuel and air as oxidant reaches 860 mW cm−2 at 650 °C.  相似文献   

19.
Perovskite solid solutions of (La0.6Sr0.4)(X1−yMgy)O3−δ (X = Ti, Zr, Al) were prepared by a coprecipitation method using corresponding aqueous solutions and ammonium carbonate solution. The freeze-dried powders were sintered in air at 1000-1500 °C for 1-36 h. Single phase solid solutions were produced in the compositions of (La0.6Sr0.4)(Zr0.6Mg0.4)O3−δ and (La0.6Sr0.4)(Al0.9Mg0.1)O3−δ where (3 − δ) < 3. For the compositions of X = Ti and Zr for y = 0.1 where (3 − δ) > 3, two phases including perovskite solid solution were produced at 1400-1500 °C. The stability of perovskite solid solution was closely related to the fraction of lattice oxygen atom (3 − δ). A relatively high conductivity was measured for (La0.6Sr0.4)(Al0.9Mg0.1)O3−δ (σ = 4.15 × 10−4 S/cm at 600 °C, activation energy 113.4 kJ/mol). The influence of fraction of oxide ion vacancy on the activation energy was small for δ = 0.1-0.3 of perovskite solid solution.  相似文献   

20.
We report on the use of the polyoxometalate acids of the series [PMo(12 − n)VnO40](3 + n)− (n = 0-3) as electrocatalysts in both the anode and the cathode of polymer-electrolyte membrane (PEM) fuel cells. The heteropolyacids were incorporated as catalysts in a commercial gas diffusion electrode based on Vulcan XC-72 carbon which strongly adsorbed a low loading of the catalyst, ca. 0.1 mg/cm2. The moderate activity observed was independent of the number of vanadium atoms in the polyoxometalate. In the anode the electrochemistry is dominated by the V3+/4+ couple. With a platinum reference wire in contact with the anode, polarization curves are obtained withVOC of 650 mV and current densities of 10 mA cm−2 at 100 mV at 80 °C. These catalysts showed an order of magnitude more activity on the cathode after moderate heat treatment than on the anode,VOC = 750 mV, current densities of 140 mA cm−2 at 100 mV. The temperature dependence of the catalysts was also investigated and showed increasing current densities could be achieved on the anode up to 139 °C and the cathode to 100 °C showing the potential for these materials to work at elevated temperatures.  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-23

京公网安备 11010802026262号