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 500 毫秒
1.
A novel and ideal dense catalytic membrane reactor for the reaction of partial oxidation of methane to syngas (POM) was constructed from the stable mixed conducting perovskite material of BaCo0.4Fe0.4Zr0.2O3– and the catalyst of LiLaNiO/-Al2O3. The POM reaction was performed successfully. Not only was a short induction period of 2 h obtained, but also a high catalytic performance of 96–98% CH4 conversion, 98–99% CO selectivity and an oxygen permeation flux of 5.4–5.8 mlcm–2min–1 (1.9–2.0 molm–2S–1Pa–1) at 850°C were achieved. Moreover, the reaction has been steadily carried out for more than 2200 h, and no interaction between the membrane material and the catalyst took place.  相似文献   

2.
Co/MgO catalysts with high Co-loading (>28 wt%) are able to initiate the reaction of methane with oxygen at temperatures around 500 °C. High conversions of methane ( 70%) and very high selectivities for hydrogen and carbon monoxide ( 90%) are obtained at very high reactant gas space velocities (105–106 h–1). The temperature of the catalyst at the conditions of partial oxidation of methane to form syngas was found to be extremely high (1200–1300 °C); it is about 600–850 °C higher than that previously reported by others. At these temperatures, high temperature homogeneous reactions may prevail. It is suggested that combustion of methane to carbon dioxide occurs on the catalyst with major heat release and that methane and water, respectively methane and carbon dioxide are reformed thermally in an endothermic reaction leading to syngas.  相似文献   

3.
The partial oxidation of methane was studied over -Al2O3-supported catalysts for Rh loadings between 0.01 and 5.0 wt%. It was found that the activity and selectivities for loadings between 0.5 and 5.0 wt% are almost the same. As an example, detailed information is presented for the 1.0 wt% Rh/-Al2O3, which provides at 750°C (furnace temperature) an activity of 82% and selectivities of 96% to CO and 98% to H2, at a gas hourly space velocity (GHSV) of 720000 ml g–1 h–1. Its activity remained stable during our experiment which lasted 120 h. Possible explanations for this high stability are proposed based on TPR and XRD experiments. Pulse reactions with small pulses of CH4 and CH4/O2 (2/1) were performed over the reduced and unreduced Rh catalysts to probe the mechanistic aspects of the reaction. The partial oxidation of methane to syngas was found to be initiated by metallic rhodium sites, since the CO selectivity increased with increasing number of such sites.  相似文献   

4.
Platinum–Cobalt-loaded NaY zeolite (Pt–Co/NaY) membranes were synthesized for continuous, single-step nonoxidative conversion of methane to higher hydrocarbons (C2+) and hydrogen. During isothermal operation at 300 °C, CH4 flowed on the feed side of the membrane whilst H2 flowed through the sweep side of the membrane. The C2+ products formed continuously on the H2 sweep side. The results indicate that the Pt–Co/NaY catalytic membrane can overcome the two-step limitation for nonoxidative CH4 conversion.  相似文献   

5.
Low-temperature SOFCs using biomass-produced gases as fuels   总被引:1,自引:0,他引:1  
The electromotive force (e.m.f) is calculated for solid oxide fuel cells (SOFCs) based on doped ceria electrolytes using biomass-produced gases (BPG, 14.7% CO, 14.2% CO2, 15.3% H2, 4.2% CH4, and 51% N2) as fuels and air as oxidant. It reveals that the BPG derived e.m.f. is very close to hydrogen when doped ceria is used as the electrolyte. A 35-m-thick samaria-doped ceria based single cell was tested between 450 and 650°C using BPG as fuel. Maximum power density of about 700 mW cm–2 was achieved at 650 °C. The open-circuit voltage at 450 °C was 0.96 V, close to the calculated value. However, the cell power density using BPG as fuel was relatively lower than that using humidified hydrogen (3% H2O), and close to that using humidified methane (3% H2O). Impedance measurements indicate that the relatively lower power output may be attributed to the high anode--electrolyte interfacial polarization resistance when BPG is used as fuel.  相似文献   

6.
The selective oxidation of methane with molecular oxygen over MoOx/La–Co–O and MoOx/ZrO2 catalysts to methanol/formaldehyde has been investigated in a specially designed high-pressure continuous-flow reactor. The properties of the catalysts, such as crystal phase, structure, reducibility, ion oxidation state, surface composition and the specific surface area have been characterized with the use of XRD, LRS, TPR, XPS and BET methods. MoOx/La–Co–O catalysts showed high selectivity to methanol formation while MoOx/ZrO2 revealed the property for the formation of formaldehyde in the selective oxidation of methane. 7 wt MoOx/La–Co–O catalyst gave 6.7 methanol yield (ca. 60 methanol selectivity) at 420°C and 4.2 MPa. On the other hand, the maximal yield of formaldehyde ca. 4 (47.8 formaldehyde selectivity) was obtained over 12wt MoOx/ZrO2 catalyst at 400 °C and 5.0MPa. 7MoOx/La–Co–O catalyst showed higher modified H2-consumption than 12MoOx/ZrO2 catalyst. The reducibility and the O/O2– ratio of the catalysts may play important roles on the catalytic performance. The proper reducibility and the O/O2– ratio enhanced the production of methanol in selective oxidation of methane. [MoO4]2– species in MoOx/ZrO2 catalysts enable selective oxidation of methane to formaldehyde.  相似文献   

7.
La2O3 promoted CaO [La/Ca (mol/mol) = 0.05] catalyst shows very high activity and selectivity (methane conversion: 25%, C2-selectivity: 66% and C2-space-time-yield: 864 mmol ·g–1 (cat.)·h–1) with no catalyst deactivation in oxidative coupling of methane to C2-hydrocarbons at 800 ° C.  相似文献   

8.
A reaction kinetic study has been performed for the reduction of nitrobenzene on a Cu electrode in 1m H2SO4 in a 5050 (Vol%) mixture of water and 1-propanol at 27°C. The study was carried out on a rotating disc electrode for which the current-potential data were supplemented with product-concentration measurements. The resulting rate expressions represent a reaction mechanism for the reduction of nitrobenzene to aniline and p-aminophenol through the common intermediate phenylhydroxylamine, and incorporate the dependence on reactant concentration and potential for the three predominant reaction pathways. The three major reaction steps were studied independently by performing experiments in which phenylhydroxylamine only was used as the reactant to complement those experiments in which nitrobenzene was used. The kinetic expressions found from measuring the rates of the individual reactions were consistent with the results of experiments in which all the reactions were carried out simultaneously. The expressions obtained are suitable for use in reactor design, modelling and control, and of equal importance, the methodology outlined to extract kinetic parameters from the current and concentration data serves as a model for application to other reaction systems.Nomenclature A electrode area (cm2) - D diffusion coefficient (cm2 s–1) - E electrode potential (V) - F Faraday's constant, 96485 (C mol–1) - i H current density due to the hydrogen evolution reaction (A cm–2) - I current (A) - I k kinetic current (A) - I L limiting current (A) - k 1 rate constant for the reduction of nitrobenzene to phenylhydroxylamine (cm s–1) - k 2 rate constant for the reduction of phenylhydroxylamine to aniline (cm s–1) - k 3 rate constant for the rearrangement of phenylhydroxylamine to p-aminophenol (s–1) - n number of electrons per equivalent - T temperature (K) - X fractional conversion of phenylhydroxylamine to p-aminophenol Greek i diffusion layer thickness of speciesi (cm) - conductivity (cm–1 ohm–1) - viscosity (g cm–1 s–1) - kinematic viscosity (cm2 s–1) - density (g cm–3) - rotation speed of electrode (s–1)  相似文献   

9.
Conclusions The introduction of hafnium dioxide into high-purity active magnesia accelerates its sintering and with 0.1–0.3 mol % HfO2 a ceramic with an apparent density of 99–100% of the theoretical (3.56–3.63 g/cm3) can be obtained after firing at 1300–1400°C. The optimum quantity of HfO2 additive is close to 0.25 mol %.The calcination temperature of the mixture of magnesium hydroxide and additive obtained by precipitation from a solution of MgCl2 and Hf(SO4)2 with ammonia, and the fabrication pressure do not greatly affect the final density of the ceramics.Sintering of spectrally pure MgO containing 0.25 mol. % HfO2 begins at 950°C, then the apparent density grows rapidly with rise in firing temperature, approaching 3.40 g/cm2 at 1100°C. Selective recrystallization at these low temperatures is slow and sintering is not accompanied by a substantial grain growth. At 1300°C and higher firing temperatures densification of the ceramics approaches the limit in several minutes.The mechanism of the transfer of substance during the sintering of these specimens is volume self-diffusion from the grain boundaries to the surfaces of the bridges formed between them. The energy of activation of this process [3.9 eV (6.2 × 10–16 J) in the 1000–1300°C range] and the coefficient of self-diffusion for MgO calculated in accordance with this (6 × 10–14 cm2/sec at 1100°C) correspond to existing data on the diffusion of magnesium into MgO.With the incorporation of 0.25 mol. % HfO2 in less pure magnesia obtained from chemically pure MgCl2, sintering is little different from the sintering of spectrally pure MgO, but the limiting apparent density of the ceramics in this case is somewhat lower-of the order of 99% of the theoretical.  相似文献   

10.
The characteristics of the effluents from the preparatory pickling step of zinc plating are presented and the various methods of oxidizing ferrous to ferric chloride are briefly considered. An electrochemical oxidation method is proposed to recover these effluents by using an electrochemical cell with three-dimensional electrodes and an anion selective membrane. A near exhausted hydrochloric acid solution was used as catholyte. The experimental data obtained from the proposed cell show a faradic yield of 100% and easy control of the parasitic reactions. The three-dimensional anode was modelled and it is shown that at high values of current only the felt entrance region works efficiently.Nomenclature A membrane surface (cm2) - a specific felt surface (cm–1) - C concentration difference (mol dm-–3) - D average diffusion coefficient through the membrane (cm2 s–1) - i n felt wall flux of species (mol cm–2 s–1) - j total current density (A cm–2) - j 0 exchange current density (A cm–2) - j 1 current density in matrix (A cm–2) - j 2 current density in felt solution (A cm–2) - j n transfer current density (A cm–2) - L thickness of felt electrode (cm) - L m thickness of membrane (cm) - M transport of ferrous and ferric ions through the membrane (mol) - N superficial flux of ion reactant (mol cm–2 s–1) - u superficial fluid velocity (cm s–1) - x distance through felt electrode (cm) - R universal gas constant (8.3143 J mol–1 K–1) - T absolute temperature (K) - t time (s) Greek letters a, c anodic and cathodic transfer coefficient - local overpotential ( = 12) (V) - conductivity of solution (mS cm–1) - µ solution viscosity (Pa s) - solution density (g cm–3) - conductivity of solid matrix (mS cm–1) - 1 electrostatic potential in matrix phase (V) - 2 electrostatic potential in solution (V)  相似文献   

11.
It is shown that the underpotential deposition of zinc on AISI 4340 steel and Inconel 718 alloys inhibits the hydrogen evolution reaction and the degree of hydrogen ingress. In the presence of monolayer coverage of zinc on the substrate surfaces, the hydrogen evolution current densities are reduced 46% and 68% compared with the values obtained on bare AISI 4340 steel and Inconel 718 alloy, respectively. As a consequence, the underpotential deposition of zinc on AISI 4340 steel and Inconel 718 alloy membrane reduces the steady state hydrogen permeation current density by 51% and 40%, respectively.List of symbols C S surface hydrogen concentration (mol cm–3) - D hydrogen diffusion coefficient (cm2 S–1) - E potential (V) - E pdep predeposition potential (V) - F Faraday constant (96 500 C mol–1) - i current density (A cm–2) - i a HER current density in the absence of predeposition of zinc (A cm–2) - i 0 exchange current density (A cm–2) - i p HER current density in the presence of predeposition of zinc (A cm–2) - j t transition hydrogen permeation current density (A cm–2) - j o initial hydrogen permeation current density (A cm–2) - j steady state hydrogen permeation current density (A cm–2) - k thickness dependent absorption-adsorption constant (mol cm–3) - L membrane thickness (cm) - Q max maximum charge required for one complete layer of atoms on a surface (C cm–2) - t time (s) Greek symbols c cathodic transfer coefficient, dimensionless - H hydrogen surface coverage, dimensionless - Zn zinc surface coverage, dimensionless - work function (eV) - = t D/L 2 (dimensionless time)  相似文献   

12.
Long service life IrO2/Ta2O5 electrodes for electroflotation   总被引:1,自引:0,他引:1  
Ti/IrO2-Ta2O5 electrodes prepared by thermal decomposition of the respective chlorides were successfully employed as oxygen evolving electrodes for electroflotation of waste water contaminated with dispersed peptides and oils. Service lives and rates of dissolution of the Ti/IrO2-Ta2O5 electrodes were measured by means of accelerated life tests, e.g. electrolysis in 0.5M H2SO4 at 25°C and j = 2 A cm–2. The steady-state rate of dissolution of the IrO2 active layer was reached after 600–700 h (0.095 g Ir h–1 cm–2) which is 200–300 times lower than the initial dissolution rate. The steady-state rate of dissolution of iridium was found to be proportional to the applied current density (in the range 0.5–3 A cm–2 ). The oxygen overpotential increased slightly during electrolysis (59–82 mV for j = 0.1 A cm–2 ) and the increase was higher for the lower content of iridium in an active surface layer. The service life of Ti/IrO2 (65 mol%)-Ta2O5 (35 mol%) in industrial conditions of electrochemical devices was estimated to be greater than five years.List of symbols a constant in Tafel equation (mV) - b slope in Tafel equation (mV dec–1) - E potential (V) - f mole fraction of iridium in the active layer - j current density (A cm–2) - l number of layers - m Ir content of iridium in the active layer (mg cm–2) - r dissolution rate of the IrO2 active layer (g Ir h–1 cm–2) - T c calcination temperature (°C) - O 2 oxygen overpotential (mV) - O 2 difference in oxygen overpotential (mV) - A service life in accelerated service life tests (h) - S service life in accelerated service life tests related to 0.1 mg Ir cm–2 (h) - p polarization time in accelerated service life tests (h)  相似文献   

13.
In situ Raman spectroscopy at temperatures up to 500°C is used for the first time to identify vanadium species on the surface of a vanadium oxide based supported molten salt catalyst during SO2 oxidation. Vanadia/silica catalysts impregnated with Cs2SO4 were exposed to various SO2/O2/SO3 atmospheres and in situ Raman spectra were obtained and compared to Raman spectra of unsupported model V2O5–Cs2SO4 and V2O5–Cs2S2O7 molten salts. The data indicate that (1) the VV complex VVO2(SO4)2 3– (with characteristic bands at 1034 cm–1 due to (V=O) and 940 cm–1 due to sulfate) and Cs2SO4 dominate the catalyst surface after calcination; (2) upon admission of SO3/O2 the excess sulfate is converted to pyrosulfate and the VV dimer (VVO)2O(SO4)4 4– (with characteristic bands at 1046 cm–1 due to (V=O), 830 cm–1 due to bridging S–O along S–O–V and 770 cm–1 due to V–O–V) is formed and (3) admission of SO2 causes reduction of VV to VIV (with the (V=O) shifting to 1024 cm–1) and to VIV precipitation below 420°C.  相似文献   

14.
Equilibrium shifts of methane steam reforming in membrane reactors consisting of either tetramethoxysilane‐derived amorphous hydrogen‐selective silica membrane and rhodium catalysts, or hexamethyldisiloxane‐derived membrane and nickel catalysts is experimentally demonstrated. The hexamethyldisiloxane‐derived silica membrane showed stable permeance as high as 8 × 10?8 mol m?2 s?1 Pa?1 of H2 after exposure to 76 kPa of vapor pressure at 773 K for 60 h, which was a much better performance than that from the tetramethoxysilane‐derived silica membrane. Furthermore, the better silica membrane also maintained selectivity of H2/N2 as high as 103 under the above hydrothermal conditions. The degree of the equilibrium shifts under various feedrate and pressure conditions coincided with the order of H2 permeance. In addition, the equilibrium shift of methane steam reforming was stable for 30 h with an S/C ratio of 2.5 at 773 K using a membrane reactor integrated with hexamethyldisiloxane‐derived membrane and nickel catalyst. © 2010 American Institute of Chemical Engineers AIChE J, 2011  相似文献   

15.
CoO-rare earth oxide catalysts (particularly CoO-Yb2O3) show high activity and selectivity in the oxidative conversion of methane to CO and H2 with very high productivity at low temperatures ( 700 °C as low as 300 °C).  相似文献   

16.
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   

17.
Conductivities of aqueous ZnSO4–H2SO4 solutions are reported for a wide range of ZnSO4 and H2SO4 concentrations (ZnSO4 concentrations of 01.2 M and H2SO4 concentrations of 02 M) at 25°C, 40°C and 60°C. The results indicate that the solution conductivity at a given ZnSO4 concentration is controlled by the H2SO4 (H+) concentration. The variation of the specific conductivity with ZnSO4 concentration is complex, and depends on the H2SO4 concentration. At H2SO4 concentrations lower than about 0.25 M, the addition of ZnSO4 increases the solution conductivity, likely because the added Zn2+ and SO 4 2– ions increase the total number of conducting ions. However, at H2SO4 concentrations higher than about 0.25 M, the solution conductivity decreases upon the addition of ZnSO4. This behaviour is attributed to decreases in the amount of free water (through solvation effects) upon the addition of ZnSO4, which in turn lowers the Grotthus-type conduction of the H+ ions. At H2SO4 concentrations of about 0.25 M, the addition of ZnSO4 does not appreciably affect the solution conductivity, possibly because the effects of increasing concentrations of Zn2+ and SO 4 2– ions are balanced by decreases in Grotthus conduction.Nomenclature a ion size parameter (m) - a * Bjerrum distance of closest approach (m) - C stoichiometric concentration (mol m–3 or mol L–1) - I ionic strength (mol L–1) - k constant in Kohlrausch's law - M molar concentration (mol L–1) - T absolute temperature (K) - z i electrochemical valence of speciesi (equiv. mol–1) - z (z |z |)1/2=2 for ZnSO4 - z + valence of cation in salt (=+2 for Zn2+) - z valence of anion in salt (=–2 for SO 4 2– ) Greek letters fraction of ZnSO4 dissociated - specific conductivity (–1 m–1) - expt measured specific conductivity (–1 m–1) - equivalent conductivity (–1 m2 equiv.–1) - equivalent conductivity at infinite dilution (–1 m2 equiv.–1) - 0 equivalent conductivity calculated using Equation 2 (–1 m2 equiv.–1) - cale measured equivalent conductivity (–1 m2 equiv.–1) - expt equivalent conductivity of ioni at infinite dilution (–1 m2 equiv.–1) - reciprocal of radius of ionic cloud (m–1) - viscosity of solvent (Pa s) - dielectric constant - ± mean molar activity coefficient - density (g cm–3)  相似文献   

18.
The rate of ion-exchange between an aqueous solution of platinum tetramine and a Nafion® 117 membrane in H+ form is studied. Experimental data are collected using extended X-ray absorption fine structure (EXAFS) spectroscopy in dispersive mode. Results are obtained for various platinum tetramine concentrations in the solution and different hydrodynamic regimes at the membrane-solution interface. A shift from a layer diffusion controlled rate (L) to a membrane diffusion controlled rate (M) is observed when the salt concentration and the stirring of the solution are increased. Time dependent fractional concentration in platinum tetramine inside the membrane are computed for the two limiting cases of diffusion (L and M). Good agreement is found between experimental and simulated data. The role of the rate of ion-exchange on the electrochemical performances of electrode-membrane-electrode composites for water electrolysis applications is discussed.List of symbols A surface of the membrane contacting the solution (1.0 cm2) - A Pt geometrical area of the platinum RDE (cm2) - C i concentration of species i in the bulk solution (mol cm–3) - C i /* concentration of species i in the membrane (mol cm–3) - C sulfonate concentration in the membrane (mol cm–3) - C 0 concentration at the membrane-solution interface (mol cm–3) - D i diffusion coefficient of species i in the solution (cm2 s–1) - D i /* diffusion coefficient of species i in the membrane (cm2 s–1) - F Faraday (96 500 C mol–1) - I L limiting current of diffusion (A) - K equilibrium constant = at equilibrium = 48 (298 K) - l diffusion layer thickness (cm) - L diameter of the cell (1.2cm) - n number of electron exchanged (2) during the electrooxidation of [Pt(NH3)4]2+ - N i fractional concentration of species i in the membrane = z i C i */C - Re Reynolds number defined as l/ - Sc Schmidt number defined as /D - t 1/2 time at which 50% of the ion-exchange is achieved - velocity of the solution (cm s–1) - V volume of the membrane (0.01 cm3) - z i charge beared by species i - kinematic viscosity of the solution (0.0114 cm2 s–1) assumed to be equal to that of pure water at 298 K - membrane thickness (cm) - rotation speed of the RDE (rad s–1)  相似文献   

19.
Dehydroaromatization of methane to benzene occurs over a 2 wt% Mo/ZSM-5 catalyst at 700C under non-oxidizing conditions. Following an initial induction period, during which CH4 reactant reduces the original Mo6+ ions in the zeolite to Mo2C and deposition of coke occurs, a benzene selectivity of 70% at a CH4 conversion of 8–10% could be sustained for more than 16 h. X-ray photoelectron spectroscopy and X-ray powder diffraction measurements indicate that the reduced Mo is highly dispersed in the channels of the zeolite. Initial activation of CH4 reactant occurs on Mo2C sites, leading to the formation of C2H4 as the primary product. The latter then undergoes subsequent oligomerization reactions on acidic sites of the zeolite to form aromatic products.  相似文献   

20.
Experimental measurements on free convection mass transfer in open cavities are described. The electrochemical deposition of copper at the inner surface of a cathodically polarized copper cylinder, open at one end and immersed in acidified copper sulphate was used to make the measurements. The effects on the rate of mass transfer of the concentration of the copper sulphate, the viscosity of the solution, the angle of orientation, and the dimensions of the cylinder were investigated. The data are presented as an empirical relation between the Sherwood number, the Rayleigh number, the Schmidt number, the angle of orientation and the ratio of the diameter to the depth of the cylinder. Comparison of the results with the available heat transfer data was not entirely satisfactory for a number of reasons that are discussed in the paper.Nomenclature C b bulk concentration of Cu++ (mol cm–3) - C b bulk concentration of H2SO4 (mol cm–3) - C o concentration of Cu++ at cathode (mol cm–3) - C o concentration of H2SO4 at cathode (mol cm–3) - D cavity diameter (cm) - D diffusivity of CuSO4 (cm2 s–1) - D diffusivity of H2SO4 (cm2 s–1) - Gr Grashof number [dimensionless] (=Ra/Sc) - g acceleration due to gravity (=981 cm s–2) - H cavity depth (cm) - h coefficient of heat transfer (Wm –2 K–1) - i L limiting current density (mA cm–2) - K mass transfer coefficient (cm s–1) - K 1,K 2 parameters in Equation 1 depending on the angle of orientation () of the cavity (see Table 3 for values) [dimensionless] - k thermal conductivity (W m–1 K–1) - L * characteristic dimension of the system (=D for cylindrical cavity) (cm) - m exponent on the Rayleigh number in Equation 1 (see Table 3 for values) [dimensionless] - Nu Nusselt number (=hL * k–1) [dimensionless] - n exponent on the Schmidt number in Equation 1 (see Table 3 for values) [dimensionless] - Pr Prandtl number (=v/k) [dimensionless] - Ra Rayleigh number (defined in Equation 2) [dimensionless] - Sc Schmidt number (=v/D) [dimensionless] - Sh Sherwood number (=KD/D) [dimensionless] - t H+ transference number for H+ [dimensionless] - t Cu++ transference number for Cu++ [dimensionless] - specific densification coefficient for CuSO4 [(1/)/C] (cm3 mol–1) - specific densification coefficient for H2SO4 [(1/)/C] (cm3 mol–1) - k thermal diffusivity (cm2 s–1) - dynamic viscosity of the electrolyte (g cm–1 s–1) - kinematic viscosity of the electrolyte (= /)(cm2 s–1) - density of the electrolyte (g cm–3) - angle of orientation of the cavity measured between the axis of the cavity and gravitational vector (see Fig. 1) [degrees] - parameter of Hasegawaet al. [4] (=(2H/D))5/4 Pr– 1/2) [dimensionless]  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