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 517 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 anodic dissolution of UIVO2 has been studied at 60 °C in 0.1 mol dm−3 KCl using a range of electrochemical methods and X-ray photoelectron spectroscopy (XPS). The results were compared to previous results obtained at 22 °C. This comparison shows that the threshold for the onset of anodic dissolution (−400 mV versus SCE) is not noticeably changed by this increase in temperature. However, both the oxidation of the surface (to UIV/VO2+x) and the rate of anodic dissolution (as UVIO22+) leading to the formation of a UVIO3·yH2O deposit are accelerated at the higher temperature. The XPS analysis shows that the conversion of UV-UVI occurs at lower potentials at 60 °C. Consequently, once the surface becomes blocked by the presence of a UVIO3·yH2O deposit, rapid dissolution coupled to uranyl ion hydrolysis causes the development of locally acidified sites within the fuel surface at lower potentials at the higher temperature.  相似文献   

3.
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   

4.
The intercalation and deintercalation mechanisms of lithium into V2O5 thin films prepared by thermal oxidation of vanadium metal have been studied by X-ray photoelectron spectroscopy (XPS) using a direct anaerobic and anhydrous transfer from the glove box (O2 and H2O < 1ppm), where the samples were electrochemically treated, to the XPS analysis chamber. Vanadium in the as-prepared oxide films is mostly (from 93 to 96% depending on samples) in a pentavalent state (V5+) with a stoichiometric O/V concentration ratio fitting that of V2O5. Four to seven percent of VO2 is also observed. After the 1st and the 2nd intercalation steps at E = 3.3 and 2.8 V versus Li/Li+, respectively, the V2p core level spectra evidence a partial reduction to V4+ states with a remaining concentration of 73 and 56% of V5+, in agreement with the intercalation of about 1/2 mol of Li per V2O5 mol at each intercalation step. Intercalated lithium was observed at a binding energy of 56.1 eV for Li1s. Changes of the electronic structure of the V2O5 thin film after intercalation are evidenced by the observation, at a binding energy of 1.3 eV, of occupied V3d states (V4+) originally empty in the pristine film (V5+). The V2p and Li1s core level spectra show that the process of Li intercalation is partially irreversible. In the first cycle, 34 and 14% of the vanadium ions remain in the V4+ state after deintercalation at E = 3.4 and 3.8 V versus Li/Li+, respectively, indicating a partially irreversible process already after the 1st deintercalation. The analyses of C1s and O1s XP spectra show the formation of a solid-electrolyte interface (SEI). The analyzed surface layer includes lithium carbonate and Li-alkoxides.  相似文献   

5.
The electrochemical reduction of hydrogen peroxide has been studied on uranium dioxide electrodes. The reduction kinetics are found to be influenced by dissolved carbonate/bicarbonate ions. The formation of hydrated UVI species on the electrode surface is avoided in carbonate solutions, allowing H2O2 reduction to proceed at less cathodic potentials than in carbonate-free solutions. At more cathodic potentials, the adsorption of carbonate ions on the active reduction sites inhibits the H2O2 reduction reaction. Over a narrow potential region, the reduction of peroxide is catalyzed by coadsoption of H2O2 and HCO3/CO32−. The pH dependence of the H2O2 reduction reaction appears to be stronger in carbonate solutions than in solutions that do not contain carbonate. This can be attributed to the displacement of inhibiting CO32−/HCO3 adsorbed ions by OH.  相似文献   

6.
The coordination circumstances and redox reactions of UO2 2+ in the aqueous solution concentrated by calcium chloride, such as CaCl2·6H2O (6.9 M CaCl2), were studied by Raman spectroscopy and electrochemical methods. The frequency of the O=U=O symmetrical stretching vibration suggested that the complex formation of UO2 2+ with Cl leads to the weakening of U=O bond. In the electrochemical measurements, two-step cathodic currents were observed at −0.090 and −0.4 V (vs. Ag|AgCl) corresponding to the reduction of UO2 2+ to UO2 + and that of UO2 + to UO2, respectively. It was found that UO2 + formed at first cathodic current was disproportionated to form UO2 2+ and UO2. The UO2 was identified by X-ray diffraction analysis. Electrolytic deposition of UO2 was observed in 6.9–4.7 M CaCl2 and in 14 M LiCl. When small amount of proton, i.e., 0.005 M was coexisted in 6.9 M CaCl2, UO2 2+ was reduced to form U4+ instead of UO2.  相似文献   

7.
The distribution of lithium in V2O5/V lower oxide duplex thin films prepared by thermal oxidation of V metal was analysed by XPS and ToF-SIMS after intercalation at 2.8 V versus Li/Li+ and de-intercalation at 3.8 V following cycling between 3.8 and 2.8 V in 1 M LiClO4-PC. XPS analysis of the intercalated thin film evidenced a partial reduction (43 at.% V4+) of the V2O5 surface, the modification of its electronic structure and the presence of Li, consistent with the formation of the δ-LixV2O5 (0.9 ≤ x ≤ 1) phase. The Li in-depth distribution measured by ToF-SIMS shows a maximum in the outer layer of V2O5, but Li is also found at the oxide film/metal substrate interface indicating its diffusion across the inner layer of V lower oxides. The analyses performed after de-intercalation on the samples cycled 12, 120 and 300 times reveal the effect of aging on the trapping of lithium. A significant reduction (17-22 at.% V4+) of the V2O5 surface was measured after 300 cycles. The Li in-depth distribution shows a maximum at the interface between the outer layer of V2O5 and the inner layer of lower oxides. Aging favours the accumulation of lithium at this interface with a resulting enlarged distribution enriching the sub-surface of the outer layer of V2O5 and the inner layer of lower oxides after 300 cycles. Lithium is also found, but in smaller quantities, at the oxide film/metal substrate interface. Measurements performed in the non-electrochemically treated surface areas of the de-intercalated samples revealed the same type of modifications, evidencing the diffusion of lithium along the interfaces where it is trapped.  相似文献   

8.
The oxidation and reduction of carbonate, GR(CO3), and sulphate, GR(SO4), green rusts (GR) have been studied through electrochemical techniques, electrochemical quartz crystal microbalance (EQCM), FTIR, XRD and SEM. The used samples were made of thin films electrodeposited on gold substrate. The results from the present work, from our previous studies and from literature were compiled in order to establish a general scheme for the formation and transformation pathways involving carbonate or sulphate green rusts. Depending on experimental conditions, two routes of redox transformations occur. The first one corresponds to reaction via solution and leads to the formation of ferric products such as goethite or lepidocrocite (oxidation) or to the release of FeII ions into the solution (reduction) with soluble FeII-FeIII complexes acting as intermediate species. The second way is solid-state reaction that involve conversion of lattice Fe2+ into Fe3+ and deprotonation of OH groups in octahedra sheets (solid-state oxidation) or conversion of lattice Fe3+ into Fe2+ and protonation of OH groups (solid-state reduction). The solid-state oxidation implies the complete transformation of GR(CO3) or GR(SO4) to ferric oxyhydroxycarbonate exGRc-Fe(III) or ferric oxyhydroxysulphate exGRs-Fe(III), for which the following formulas can be proposed, FeIII6(OH)(12−2y)(O)(2+y)(H2O)(y)(CO3) or FeIII6(OH)(12−2z)(O)(2+z)(H2O)(6+z)(SO4) with 0 ≤ y or z ≤ 2. The solid-state reduction gives ferrous hydroxycarbonate exGRc-Fe(II) or ferrous hydroxysulphate exGRs-Fe(II), which may have the following chemical formulas, [FeII6(OH)10(H2O)2]·[CO3, 2H2O] or [FeII6(OH)10(H2O)2]·[SO4, 8H2O].  相似文献   

9.
A two-step hydrothermal process was developed to synthesize hydrous 30RuO2-70SnO2 composites with much better capacitive performances than those fabricated through the normal hydrothermal process, co-annealing method, or modified sol-gel procedure. A very high specific capacitance of RuO2 (CS,Ru), ca. 1150 F g−1, was obtained when this composite was synthesized via this two-step hydrothermal process with annealing in air at 150 °C for 2 h. The voltammetric currents of this annealed composite were found to be quasi-linearly proportional to the scan rate of CV (up to 500 mV s−1), demonstrating its excellent power property. From Raman, UV-vis spectroscopic and TEM analyses, the reduction in mean particulate size is clearly found for this two-step oxide composite, attributable to the co-precipitation of (RuδSn1−δ)O2·xH2O onto partially dissolved SnO2·xH2O and the formation of (RuδSn1−δ)O2·xH2O crystallites in the second step. This effect significantly promotes the utilization of RuO2 (i.e., very high CS,Ru). The excellent capacitive performances, very similar to that of RuO2·xH2O, suggest the deposition of RuO2-enriched (RuδSn1−δ)O2·xH2O onto SnO2·xH2O seeds as well as the individual formation of (RuδSn1−δ)O2·xH2O crystallites in the second hydrothermal step.  相似文献   

10.
Sr-hexaferrites prepared by co-precipitation method and calcined at 700-1000 °С have been characterized by thermogravimetric and differential thermal analysis (TG-DTA), Fourier transformed infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), X-ray diffraction (XRD), hydrogen temperature-programmed reduction (H2-TPR), and Ar adsorption techniques. It has been shown that hexaferrite phase formed after calcination at 700 °С is amorphous and its crystallization occurs at 800 °С. Specific surface area (SBET) of the samples calcined at 700 °С is 30-60 m2/g. Reduction in hydrogen proceeds in several steps, Fe(III) in the hexaferrite structure being practically reduced to Fe0. Amount of hydrogen necessary for the reduction of the samples decrease in the order: SrMn2Fe10O19 > SrFe12O19 > SrMn6Fe6O19 > SrMn2Al10O19. Surface composition of the ferrites differs from bulk. According to XPS data, the surface is enriched with strontium. Sr segregation is most probably explained by the formation of surface carbonates and hydroxocarbonates. The main components on the surface are in oxidized states: Mn3+ and Fe3+. Maximum activity in the methane oxidation is achieved for the SrMnxFe12−xO19 (0 ? x ? 2) catalysts. These samples are characterized by highest amount of the hexaferrite phase, which promotes change of oxidation state Mn(Fe)3+ ↔ Mn(Fe)2+.  相似文献   

11.
A series of ZnxMg1 − xGa2O4:Co2+ spinels (x = 0, 0.25, 0.5, 0.75, and 1.0) was successfully produced through low-temperature burning method by using Mg(NO3)2·4H2O, Zn(NO3)2·6H2O, Ga(NO3)3·6H2O, CO(NH2)2, NH4NO3, and Co(NO3)2·6H2O as raw materials. The product was characterized by X-ray diffraction, transmission electron microscopy, and photoluminescence spectroscopy. The product was not merely a simple mixture of MgGa2O4 and ZnGa2O4; rather, it formed a solid solution. The lattice constant of ZnxMg1 − xGa2O4:Co2+ (0 ≤ x ≤ 1.0) crystals has a good linear relationship with the doping density, x. The synthesized products have high crystallinities with neat arrays. Based on an analysis of the form and position of the emission spectrum, the strong emission peak around the visible region (670 nm) can be attributed to the energy level transition [4T1(4P) → 4A2(4F)] of Co2+ in the tetrahedron. The weak emission peak in the near-infrared region can be attributed to the energy level transition [4T1(4P) → 4T2(4F)] of Co2+ in the tetrahedron.  相似文献   

12.
Electrode-potential-dependent activation energies for electron transfer have been calculated using a local reaction center model and constrained variation theory for the oxygen reduction reaction on platinum in base. Results for four one-electron transfer steps are presented. For the first, O2(ads) is predicted to be reduced to adsorbed superoxide, O2(ads), which dissociates with a low activation barrier to O(ads) + O(ads). Then a proton transfer form H2O(ads) to O(ads) takes place, forming OH(ads) + OH(aq). The second electron transfer reacts O(ads) with H2O(aq) to form a second OH(ads) + OH(aq). The third and fourth electron transfers react the two OH(ads) with two H2O(aq) to form two H2O(ads) + two OH(aq). All three different surface reduction reactions are predicted to have reversible potentials in the −0.24 V(SHE) to −0.29 V(SHE) range for 0.1 M base and activation energies for the superoxide formation step are close to the experimentally observed range in 0.1 M base for the overall four-electron to water over the three low index (1 1 0) (1 0 0) and (1 1 1) surfaces: 0.38-0.49 eV at 0.35 eV respectively at 0.88 V(RHE). Predicted reversible potentials for forming O2(ads) are compared with estimates from the experimental literature. The difference between the acid mechanism, where the peroxyl radical, OOH(ads) is the first reduction intermediate, and the base mechanism, where superoxide, O2(ads) is the first reduction intermediate, is discussed.  相似文献   

13.
A KNbO3 nanoneedles (KNs) based hydrogen peroxide (H2O2) biosensor was first proposed. Perovskite-type KNs can directly catalyze H2O2. The mechanism can be explained by Molecular Orbital Principles, with the formation of σ-bonding between the eg orbital of surface niobium ions and surface adsorbed oxygen-related intermediate species. Direct electron transfer between the Horseradish peroxidase (HRP) and electrode surface was achieved. Co-catalyst system of both HRP and KNbO3 was introduced to the oxidation of H2O2, thus the as-prepared biosensor exhibited high sensitivity (750 μA mM−1 cm−2) and ultrafast response (1–2 s) to H2O2. Therefore, KNs provide a promising material for enzymes assembly and sensing application.  相似文献   

14.
Anodic oxidation of Pd in basic solutions (0.1 M KOHaq and 0.1 M NaOHaq) has been examined via cyclic voltammetry (CV) and an electrochemical quartz crystal microbalance (EQCM). Admittance tests show that Pd(II) layer behaves as a rigid one. The anodic vertex potential influences mass response during formation of the Pd(II) layer. For low anodic vertex potentials, obtained absolute mass per mole values suggest Pd(OH)2 or PdO·H2O to be oxidation products. At this stage of the oxidation process, contribution from adsorbed H2O/OH in Pd(II) layer formation could explain the lower-than-expected mass gain, although the extent of H2O/OH adsorption is unclear. The mass gain decreases with further increase in the anodic vertex potential, eventually reaching the value of ca. 8 g mol−1 at about 700 mV vs. SCE. Comparing the influence of vertex potential in CV experiments on the mass and reduction potential of the Pd(II) species points to the formation of PdO at higher oxidation potentials. At this stage of the process, a fraction of the PdO species is generated during transformation of previously formed Pd(OH)2/PdO·H2O. A shift of the main Pd(II) reduction potential peak depends on both the anodic vertex potential and on the composition of the Pd(II) film. The order of the Pd(II) reduction process is the opposite of that observed for the oxidation process. The Pd(IV) species formed at E ≥ 500 mV vs. SCE and those reduced between 50 and 350 mV are hydrated or contain hydroxyl groups.  相似文献   

15.
Ni/Al2O3 with the doping of CeO2 was found to have useful activity to reform ethane and propane with steam under Solid Oxide Fuel Cells (SOFCs) conditions, 700-900 °C. CeO2-doped Ni/Al2O3 with 14% ceria doping content showed the best reforming activity among those with the ceria content between 0 and 20%. The amount of carbon formation decreased with increasing Ce content. However, Ni was easily oxidized when more than 16% of ceria was doped. Compared to conventional Ni/Al2O3, 14%CeO2-doped Ni/Al2O3 provides significantly higher reforming reactivity and resistance toward carbon deposition. These enhancements are mainly due to the influence of the redox properties of doped ceria. Regarding the temperature programmed reduction experiments (TPR-1), the redox properties and the oxygen storage capacity (OSC) for the catalysts increased with increasing Ce doping content. In addition, it was also proven in the present work that the redox of these catalysts are reversible, according to the temperature programmed oxidation (TPO) and the second time temperature programmed reduction (TPR-2) results.During the reforming process, in addition to the reactions on Ni surface, the gas-solid reactions between the gaseous components presented in the system (C2H6, C3H8, C2H4, CH4, CO2, CO, H2O, and H2) and the lattice oxygen (Ox) on ceria surface also take place. The reactions of adsorbed surface hydrocarbons with the lattice oxygen (Ox) on ceria surface (CnHm+OxnCO+m/2(H2)+Oxn) can prevent the formation of carbon species on Ni surface from hydrocarbons decomposition reaction (CnHmnC+m/2H2). Moreover, the formation of carbon via Boudard reaction (2CO⇔CO2+C) is also reduced by the gas-solid reaction of carbon monoxide (produced from steam reforming) with the lattice oxygen (CO+Ox⇔CO2+Ox−1).  相似文献   

16.
The roles of adsorbed hydroxyl radicals, OH, at a high temperature and adsorbed hydrogen atoms, H, in an acidic solution were investigated in the electrochemical reactions on Pt electrode by using potentiodynamic polarisation experiment, cyclic voltammetry and constant-potential electrolysis combined with UV/VIS analysis. From the analysis of the polarisation curves obtained from Pt electrode in a 0.185 M H3BO3 solution at 473 K, it was found that the reducing capability of dissolved hydrogen is significantly enhanced due to the increases of the mass transfer and the electron transfer rates. Especially, it is suggested that the stable Pt-OHad plays a significant role in the passivation reaction in the potential range from 0.60 to 0.75 VSHE. From the analyses of the experimental results for the electrochemical reduction of UO22+ ions on Pt surface in a 1.0 M HClO4 solution, it is recognised that the reduction reaction of UO22+ to U4+ ions is strongly dependent on the hydrogen atoms adsorbed on Pt electrode (indirect reduction of UO22+) as well as on the electrons transferred from Pt electrode (direct reduction of UO22+). In addition, the reduction mechanism of UO22+ ions involved in Pt-Had is also proposed.  相似文献   

17.
Nano-crystalline strontium hexaferrite (SrFe12O19) powder was synthesized using the classical co-precipitation and microemulsion methods. The precursors were obtained by precipitating Sr2+ and Fe2+ ions using tetramethylammonium hydroxide and calcinating at different temperatures ranging from 400 °C to 1000 °C in air. The influence of the Sr2+/Fe3+ mol ratio and the calcination temperature on the product formation and magnetic properties were studied. The formation of nanosized particles of SrFe12O19 with a relatively high saturation magnetization Ms = 64 Am2/kg, remanent magnetization of Mr = 39 Am2/kg and a coercitivity of Hc = 5.5 kOe was achieved at a Sr2+/Fe3+ mol ratio of 1:8 calcined at 900 °C. The formation of the SrFe12O19 was inspected using XRD analysis, thermogravimetric analysis (TGA), differential thermal analysis (DTA), TEM, and magnetic measurements.  相似文献   

18.
A series of cerium oxide based materials for hydrogen production from water was studied by temperature-programmed reduction, X-ray diffraction and X-ray photoelectron spectroscopy (XPS) in addition to thermal reactions with water vapour. The addition of uranium ions into CeO2, making mixed oxides Ce x U1?x O2, resulted in noticeable modification of the reduction properties of CeO2; with the main observations being the decrease in reduction temperature and the increase of hydrogen consumption when compared to CeO2 alone. XPS U4f of the as prepared Ce0.5U0.5O2 showed the presence of large amounts of U6+ cations at 380.9 eV in addition to the U4+ cations at 379.9 eV; the ratio U4+ to U6+ cations was found equal to 0.35. XPS Ce3d showed, on the contrary, considerable amount of Ce3+ cations with an estimated ratio of Ce3+ to Ce4+ = ca. 0.5. Ar-ions sputtering results in decreasing the U6+ contribution and a dramatic increase of the Ce3+ contribution. The decrease of U6+ cations was, however, not mirrored by the increase in Ce3+ cations. After five minutes of Ar ions sputtering (1 kV, 10 mA) the surface and near surface Ce3d line shapes looked closer to those of Ce2O3 with prominent Ce3d5/2 and Ce3d3/2 lines at 885.6 and 904.0 eV attributed to v′ and u′, respectively. The Ce x U1?x O2 series was tested for hydrogen production from water (where x = 0, 0.25, 0.5, 0.75 and 1). All uranium containing oxides had higher activity than CeO2 or UO2 alone. Ce0.75U0.25O2 was found to have the highest activity in the studied series; about one order of magnitude higher than that of CeO2 alone at the same temperature. The reason for the enhanced activity is linked to the ease by which oxygen ions are removed from the oxide materials.  相似文献   

19.
Free acids of the iron substituted heteropoly acids (HPA), H7[(P2W17O61)FeIII(H2O)] (HFe1) and H18[(P2W15O56)2FeIII2(H2O)2] (HFe2) were prepared from the salts K7[(P2W17O61)FeIII(H2O)] (KFe1) and Na12[(P2W15O56)2FeIII4(H2O)2] (NaFe4), respectively. The iron-substituted HPA were adsorbed on to XC-72 carbon based GDLs to form HPA doped GDEs after water washing with HPA loadings of ca. 1 μmol. The HPA was detected throughout the GDL by EDX. Solution electrochemistry of the free acids are reported for the first time in sulfate buffer, pH 1-3. The hydrogen oxidation reaction was catalyzed by KFe1 at 0.33 V, with an exchange current density of 38 mA/cm2. Moderate activity for the oxygen reduction reaction was observed for the iron substituted HPA, which was dramatically improved by selectively removing oxygen atoms from the HPA by cycling the fuel cell cathode under N2 followed by reoxidation to give a restructured oxide catalyst. The nanostructured oxide achieved an OCV of 0.7 V with a Tafel slope of 115 mV/decade. Cycling the same catalysts in oxygen resulted in an improved catalyst/ionomer/carbon configuration with a slightly higher Tafel slope, 128 mV/decade but a respectable current density of 100 mA/cm2 at 0.2 V.  相似文献   

20.
To understand the concentration overpotential in the polymer electrolyte fuel cell (PEFC), we have performed an in situ analysis of the oxygen partial pressure (p[O2]CL/PEM) at the interface between the cathode catalyst layer (CL) and the polymer electrolyte membrane (PEM). Diffusion-limited oxygen reduction current was measured, with Pt probes inserted into the PEM, during cell operation by supplying H2 to the anode and O2 + N2 to the cathode at 80 °C. It was found that the p[O2]CL/PEM decreased by ca. 20% when the current density was stepped from 0 to 2.0 A cm−2 at p[O2]gas = 54 kPa and 100% RH at the cathode inlet, irrespective of the oxygen utilization UO2 (from 10% to 50%). Such a change in p[O2]CL/PEM might result in a concentration overpotential of ca. 10 mV, based on the Tafel slope of 120 mV decade−1 in the high current density region. It was also found that ohmic losses in the ionomer phase of the CL increased with decreasing humidity, from 100% to 80% RH, and became a dominant factor in the increased total overpotential, while the corresponding concentration overpotential was unchanged. The present results provide new insight into the transport of oxygen and water at the CL/PEM interface, especially at the high current densities required for the electric vehicle application.  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