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.
Electrolytically obtained Ni–Co–P alloys exhibit amorphous structure. On heating these alloys to 773 K crystalline nickel and cobalt phosphide phases, and also nickel crystallites, are formed. After cathode-anode polarization of the amorphous and crystalline Ni–Co–P alloys in 5m KOH for 18 h, oxygen-cobalt compounds together with crystalline or amorphous nickel phosphides appear on the electrode surface. The evolution of oxygen from 5m KOH was studied on Ni–Co–P alloys prepared in this way. The Tafel parameters were determined and it was ascertained that the values of directional coefficients of the Tafel lines are comparable with those obtained for spinel NiCo2O4 oxides, while calculated values of the exchange currents for the oxygen evolution reaction are dependent on the chemical composition of the alloy. The rate of electrolytic oxygen evolution is virtually identical for amorphous and crystalline Ni–Co–P alloys of the same chemical composition. The highest rate of oxygen evolution was found for the alloy containing 15–20% Ni, 66–73% Co and 12–14% P.  相似文献   

2.
Potentiodynamic and potentiostatic polarization, and the rotating disk electrode technique were used to study the reduction characteristics of iodate (IO3 ) ion on copper (Cu). Depending on the relative concentrations of IO3 and H+ two pH regimes were observed. The cathodic current in the first regime (pH > ,3) was controlled by H+ diffusion from the solution to the metal surface. In the second regime (pH > 3 and up to 102 m IO3 concentration) the cathodic current was found to be under mixed control, involving reaction control via the electrochemical reduction of IO3 and transport control via the diffusion of I2(aq). It was concluded that IO3 was an effective oxidant for Cu chemical mechanical polishing (CMP) with strongly acidic (pH < 3) slurries but it was not convenient reagent as an oxidant for Cu CMP with weakly acidic (pH > 3) slurries.  相似文献   

3.
When CF4 is bubbled through molten fluorides containing dissolved oxides within the range 900–1020° C, a chemical reaction takes place which decreases the oxide concentration and yields CO2 and F ions. The possibility of the reaction between the CF4 evolved at the anode and dissolved alumina, occurring during the anode effect in aluminium reduction cells, is discussed. This reaction provides a simple and convenient method for removing oxides and hydroxides from molten fluorides.  相似文献   

4.
The electrocatalytic properties of thin Co3O4 and NiCo2O4 films prepared on CdO-nonconductive glass by the method of chemical spray pyrolysis were investigated for oxygen evolution in varying KOH concentrations. The Tafel slopes were close to (2.3RT/F) V per decade for both oxides. The reaction order with respect to [OH] was found to be approximately 1.3. It was observed that Co3O4 is more active than NiCo2O4 towards oxygen evolution. A mechanism for oxygen evolution involving the electrochemical adsorption of OH as a fast step and the electrochemical desorption of OH forming an intermediate H2O2 as the rate-determining step has been suggested.  相似文献   

5.
The gas–solid reaction between methane and the lattice oxygen of Ni, Co, and Fe-oxides loaded on various support materials produced a synthesis gas (hydrogen and carbon monoxide) at 600–800 °C. Metal oxides were reduced to metals or lower valence oxides, and they were re-oxidized to oxides by introducing air after the reaction. Thus, production of hydrogen or synthesis gas free from nitrogen can be achieved alternatively without using pure oxygen. As a metal oxide, Fe2O3 and Rh2O3-loaded on Y2O3 exhibited the highest H2 selectivity of 60.1% with a moderate CH4 conversion of 54% and a high lattice oxygen utilization of 84% at 800 °C.  相似文献   

6.
The gaseous species desorbed from porous silicon (PS) were investigated using the method of temperature programmed desorption (TPD) and fourier transform infrared spectroscopy (FTIR). Silicon wafers (25–50 cm, p, FZ) were anodised in 40% HF and HF/C2H5OH electrolytes. The PS samples were linearly heated at 1.5 K s1 using a custom built heating unit in a oil-free pump backed vacuum chamber at a base pressure of <108 torr. A quadrupole mass spectrometer, which was used as the detector, was fitted in line of sight of the sample at a distance of about 6 mm. It was observed that silane was liberated during the heating of porous silicon samples produced from both electrolytes. The peak temperature at which this occurred was at 570 ± 10 K. This temperature coincides with the temperature of silicon-silicon bond breakage in Si–SiH3 groups on the pore walls, as shown by the FTIR results. It is proposed that silane formation involves the reaction of the Si-silyl group with moisture: Si–SiH3 + H2O Si–OH + SiH4.  相似文献   

7.
NiAl2O4 spinel was formed by solid state reaction. Its electrical conductivity was measured in the temperature range of 680–940 °C and under various oxygen-rich environments, as well as under reducing conditions. From the temperature dependence of the conductivity, the activation energies for conduction increase for decreasing oxygen partial pressures. From the partial oxygen pressure dependence, the defect structure of the material was analysed. The conductivity change with respect to P O2 can be attributed to singly and doubly ionized nickel vacancies. The chemical diffusivity of the oxide was determined by conductivity relaxation upon abrupt change in P O2 in the surrounding atmosphere. The oxygen chemical diffusion coefficient is of the order of magnitude of 10–4 cm2 s–1.  相似文献   

8.
Three Na-based thermochemical cycles for capturing CO2 from air are considered: (1) a NaOH/NaHCO3/Na2CO3/Na2O cycle with 4 reaction steps, (2) a NaOH/NaHCO3/Na2CO3 cycle with 3 reactions steps, and (3) a Na2CO3/NaHCO3 cycle with 2 reaction steps. Depending on the choice of CO2 sorbent – NaOH or Na2CO3 – the cycles are closed by either NaHCO3 or Na2CO3 decomposition, followed by hydrolysis of Na2CO3 or Na2O, respectively. The temperature requirements, energy inputs, and expected products of the reaction steps were determined by thermodynamic equilibrium and energy balance computations. The total thermal energy requirement for Cycles 1, 2, and 3 are 481, 213, and 390 kJ/mol of CO2 captured, respectively, when heat exchangers are employed to recover the sensible heat of hot streams. Isothermal and dynamic thermogravimetric runs were carried out on the pertinent carbonation, decomposition, and hydrolysis reactions. The extent of the NaOH carbonation with 500 ppm CO2 in air at 25 °C – applied in Cycles 1 and 2 – reached 9% after 4 h, while that for the Na2CO3 carbonation with water-saturated air – applied in Cycle 3 – was 3.5% after 2 h. Thermal decomposition of NaHCO3 – applied in all three cycles – reached completion after 3 min in the 90–200 °C range, while that of Na2CO3 – applied in Cycle 1 – reached completion after 15 min in the 1000–1400 °C range. The significantly slow reaction rates for the carbonation steps and, consequently, the relatively large mass flow rates required, introduce process complications in the scale-up of the reactor technology and impede the application of Na-based sorbents for capturing CO2 from air.  相似文献   

9.
Mo–V–Nb–P–O-based catalysts with a tetragonal tungsten bronze-type (TTB) structure have been prepared hydrothemally from a H3PMo12O40 Keggin-type heteropolyacid. These catalysts have been tested in the oxidation of C3–C4 olefins (propene, isobutene and 1-butene). Although the catalytic performance depends on the nature of the olefin fed the TTB-type catalysts prepared in the presence of elements of the V and VI groups such as Te, Sb and Bi have shown a high selectivity to partial oxidation products, especially that with Te. However, in the absence of these elements the TTB-catalysts present a high catalytic activity to deep oxidation. The selectivity to partial oxidation products decreases in the order: MoVNbPTe- > MoVNbPSb- > MoVNbPBi- > MoVNbP-TTB catalysts. The reaction products obtained in the oxidation of each olefin will be discussed according to their corresponding reaction mechanism and the characteristics of catalysts.  相似文献   

10.
Hydrogenated nanocrystalline silicon (nc-Si:H) thin films were deposited on Si wafers at room temperature by plasma-enhanced chemical vapor deposition (PECVD): a mixture of SiH4 and H2 was introduced into the evacuated reaction chamber. The films were postdeposition annealed at temperatures of 400–1100°C. The silicon nanocrystallites (nc-Si) in the films range from ∼4.0 to ∼8.0 nm in size, depending on the hydrogen flow rates as well as the annealing conditions. The relative fractions of the Si-H3, Si-H2, and Si-H bonds in the nc-Si:H films varied sensitively with the heat-treatment conditions. Local radial distribution function (RDF) analysis of the films was performed by using selected area electron diffraction methods. A model for the nanostructure of the nc-Si:H films was suggested to take into account the variation in the size and chemical bonds of the nanocrystallites as well as the amorphous matrix.Original English Text Copyright © 2005 by Fizika i Khimiya Stekla, Shim, Cho.This article was submitted by the authors in English.  相似文献   

11.
The products of reactions occurring in a CH4–O2–NO x mixture have been investigated by in situ FTIR spectroscopy. The results show that low temperatures favor the formation of HCHO and CH3OH, while high temperatures favor that of C2H4. Possible reaction mechanisms based on the in situ observations are briefly discussed.  相似文献   

12.
Results of a study of phase formation in the Al2O3 – ZrO2 – SiC system sintered under vacuum and in a reducing medium at 1350 – 1550°C are reported. Conditions for preparation of Al2O3 and ZrO2 powders from hydroxides are specified and the effect of specific surface, temperature, and holding time on the reaction between oxide and carbide components is considered. Results of microstructural, x-ray phase, thermogravimetric, and chemical analyses of precursor materials and end products are discussed. Optimum composition (20% SiC, 15% ZrO2, 65% Al2O3 ) and dispersity of the mixture for obtaining a composite with a strength of about 200 MPa by a method other than high-temperature toughening are determined.  相似文献   

13.
Metal pyrophosphates (M–P2O7, where M is V, Zr, Cr, Mg, Mn, Ni or Ce) have been found to catalyze the oxidative dehydrogenation of propane to propene. The reaction was conducted at 1 atm, 450–550°C and feed flowrate of 75 cm3/min (20 cm3/min propane, 5 cm3/min oxygen and the balance is helium). All catalysts showed increase in degrees of conversion and decrease in olefins selectivity with increase in reaction temperature. At 550°C, MnP2O7 exhibited the highest activity (40.7% conversion) and total olefins (C3H6 and C2H4) yield (29.3%). The other catalysts, indicated by their respective metals, may be ranked (based on olefins yield) as V (16.9%) < Cr (17.5%) < Ce (25.1%) < Zr (26.2%) < Ni (26.8%) < Mg (27.9%). The reactivity of the lattice oxygen was estimated from energy of formation of the corresponding metal oxides. Correlation between the selectivity to propene and the standard energy of formation was attempted. Although there was no clear correlation, the result suggested that the lattice oxygen play a key role in the selectivity-determining step.  相似文献   

14.
Supported gold, palladium and gold–palladium catalysts have been used to oxidatively dehydrogenate cyclohexane and cyclohexenes to their aromatic counterpart. The supported metal nanoparticles decreased the activation temperature of the dehydrogenation reaction. We found that the order of reactivity was Pd ≥ Au–Pd > Au supported on TiO2. Attempts were made to lower the reaction temperature whilst retaining high selectivity. The space-time yield of benzene from cyclohexane at 473 K was determined to be 53.7 mol/kgcat/h rising to 87.3 mol/kgcat/h at 673 K for the Pd catalyst. Increasing the temperature in this case improved conversion at a detriment to the benzene selectivity. Oxidative dehydrogenation of cyclohexene over AuPd/TiO2 or Pd/TiO2 catalysts was found to be very effective (conversion >99% at 423 K). These results indicate that the first step in the reaction sequence of cyclohexane to cyclohexene is the slowest step. These initial results suggest that in a fixed-bed reactor the oxidative dehydrogenation in the presence of oxygen, palladium and gold–palladium catalysts are readily able to surpass current literature examples and with further modification should yield even higher performance.  相似文献   

15.
Several palladium on alumina and ceria/alumina catalysts were prepared and oxidized in air between 400 and 1000°C. The metal dispersion was determined by hydrogen titration of adsorbed oxygen. Dispersions above 50% were maintained on 0.2% Pd/Al2O3 up to 900°C. Adding 5.0% ceria, or increasing the metal loading to 2.5%, greatly reduces the thermal stability of the palladium, such that the dispersion falls rapidly at 600°C. The rates of methane oxidation (moles of CO2/g Pd h) at 250°C and 5% excess oxygen are nearly equal on 0.22–2.50% Pd/3.5–5.2% CeO2/Al2O3, dispersion 14–42%, and 0.20–0.46% Pd/Al2O3, dispersion 59–86%, but are 10 to 20 times lower than the rate on 2.3% Pd/Al2O3, dispersion 11%. The lower rate of methane oxidation on ceria-promoted and highly dispersed palladium on alumina might be due to the conversion of the palladium into less active palladium oxide during reaction.  相似文献   

16.
Conclusions It was established that a high-density yttrium ceramic (relative density 95–98%) can be produced by adding 20–30% TiB2 and sintering in vacuo or hot molding at 1500°C. It is assumed that the high density of the material is the result of the formation of reaction products in the interaction of the components of the TiB2-Y2O3 system.Translated from Ogneupory, No. 3, pp. 53–55, March, 1977.  相似文献   

17.
We have applied photoelectron spectroscopy to investigate the surface composition after different surface treatments involving Br2–H2O mixtures in order to study wet chemical etching. Emersion experiments from Br2–H2O solution are compared with model experiments, in which Br2–H2O adsorbate and coadsorbate mixtures react with clean GaAs(110) surfaces. Our results indicate that Ga- and As-bromides formed initially are hydrolyzed to form the respective oxides. Without addition of Br2, only slight oxidation of the surface takes place. There is an enrichment of Ga due to loss of As both in adsorption as well as in emersion experiments. Since in emersion experiments only a final situation is analyzed, the relative influence of surface reactivity and subsequent solvation effects cannot be distinguished easily, while model experiments give clear information on reaction products formed intermediately. However, model experiments differ in environment and temperature from the real solid–liquid interface. The presented results demonstrate that a combination of emersion and model experiments provide valuable insight into the mechanism of wet chemical etching on a microscopic level.  相似文献   

18.
The processes of phase formation in the Nd2O3–Ho2O3–SrO–Al2O3 system are investigated in the temperature range 1100–1530°C. It is revealed that the structural–chemical mechanism of formation of (Nd x Ho1 – x )2SrAl2O7 solid solutions depends on the temperature and composition. An increase in the holmium content and in the temperature leads to a crossover from the mechanism including the formation of LnAlO3 and LnSrAlO4 as intermediate products to the mechanism in which the interaction of Ln 2O3 and SrAl2O4 is a limiting stage.  相似文献   

19.
Oxygen reduction on stainless steel   总被引:2,自引:0,他引:2  
Oxygen reduction was studied on AISI 304 stainless steel in 0.51 m NaCl solution at pH values ranging from 4 to 10. A rotating disc electrode was employed. It was found that oxygen reduction is under mixed activation-diffusion control. The reaction order with respect to oxygen was found to be one. The values of the Tafel slope depend on the potential scan direction and pH of the solution, and range from – 115 to – 180 mV dec–1. The apparent number of electrons exchanged was calculated to be four, indicating the absence of H2O2 formation.Nomenclature B =0.62 nFcD 2/31/6 - c bulk concentration of dissolved oxygen (mol dm–3) - D molecular diffusion coefficient of oxygen (cm2 s–1) - E electrode potential (V) - EH standard electrode potential (V) - E H 0 Faraday constant (96 500 As mol–1) - I current (A) - j current density (A cm–2) - j k kinetic current density (A cm–2) - j L limiting current density (A cm–2) - m reaction order with respect to dissolved oxygen molecule - M molar mass (g mol–1) - n number of transferred electrons per molecule oxygen - density (g cm–3) - kinematic viscosity (cm2 s–1) - angular velocity (s–1)  相似文献   

20.
Solid oxide fuel cells are studied under direct methane feeding with 10–70% CH4. When either La0.58Sr0.4Co0.2Fe0.8O3−δ (LSCF)–Ce0.9Gd0.1O1.95 (GDC) or Ni-added LSCF–GDC composite is used as the anode, the oscillations of the electrical current and the formation rates of CO and CO2 occur. The oscillation of the electrical current can be explained by a mechanism of periodic oxidation–reduction of the bulk lattice of the anode, with the determining factor being the build-up of the concentration of the oxygen vacancies to a certain extent. As the methane concentration increases, the current density increases and becomes larger with Ni addition. Higher methane concentration leads to higher possibility to induce the oscillation, to start it earlier, and to result in a larger amplitude. Ni addition inhibites the occurrence of the oscillation of the electrical current but promotes that of the CO2 formation rate.  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