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.
During alkaline water electrolysis, additional energy losses occur owing the presence of bubbles in the solution, particularly close to both the gas-evolving electrodes.For both hydrogen and oxygen-evolving disc electrodes (diameters from 0.2 to 2.0 mm) in KOH solutions, the reduced increase in ohmic resistance, ΔR*, has been determined by the alternating current—impedance method.It has been found that, for hydrogen-evolving electrodes, log ΔR* = 11 + b log i, where the exponent b at 0.1 A cm?2 < i < 5 A cm?2 does not depend on the diameter, position and material of the electrode, pressure and temperature but does significantly depend on KOH concentration. The factor a1, however, being dependent on the position, height and material of electrode, temperature and KOH concentration. ΔR* cannot be expressed for the oxygen-evolving electrode by a general equation, due to the coalescence behaviour of oxygen bubbles.Moreover, it has been established that the Bruggemann equation is useful to determine the ohmic resistance of a solution layer containing gas bubbles of different size at which each bubbles adheres to the electrode surface.  相似文献   

2.
The convective diffusion model, prevailing in radial thin layer flow cells (thickness 2h ? 10?2 cm) with central solution inlet at ultra-low flow rates J ? 10?2 cm3 s?1, has been investigated by numerical finite-difference techniques for the transport-controlled conversion of metal ions at a coulometric asymmetrical or symmetrical metal detector, mounted on one or on both sides of the thin layer, respectively. At symmetrical detectors, the concentration profiles and the conversion efficiency have been evaluated by assuming negligible radial diffusion and are in good agreement with previous eigenvalue solutions. At unsymmetrical detectors, the mass transfer study has been carried out in consideration of radial diffusion. The diffusional contribution to the radial flux becomes increasingly apparent in the concentration profiles and the conversion efficiency, as the internal detector boundary r0 is shifted downstream at constant h and J, or as J and h are lowered proportionally at constant r0. This effect can be diminished by choosing appropriate correlations between r0, h and J, which are established together with the limiting conditions for coulometric detector operation.  相似文献   

3.
《Electrochimica acta》1988,33(6):769-779
The effect of current density used for total gas evolution, ig, or for gas in bubbles, ib, and the effect of solution—flow velocity, vs,0, and temperature, T, on the detached bubbles parameters, viz average Sauter bubble radius, Rs,d,av, average bubble radius, Rd,av, average of the square of the bubble radii, (R2d)av, average of the third power of the bubble radii, (R3d)av, detached bubble frequency, ω, and efficiency of gas bubble evolution, ηb, for a chlorine-, a hydrogen- and an oxygen-evolving electrode at forced convection have been determined experimentally. The electrodes consist of a thin wire electrode with a small height. It has been found that:
  • 1.The average Sauter bubble radius is proportional to in4b and decreases linearly with increasing vs,0 at constant ig. The constant n4 depends on the temperature, the nature of the gas evolved, the electrode material and the electrolyte and is independent of vs,0. Similar correlations were obtained for Rd,av, (R2d)av and (R3d)av.
  • 2.The frequency of detached bubbles is proportional to in5b and increases linearly with increasing vs,0 at constant ig. The constant n5 depends on the temperature, the nature of the gas evolved, the electrode material and the electrolyte and is independent of vs,0.
  • 3.The efficiency of gas bubble evolution ηb increases with increasing ig and approaches a limiting value.
  相似文献   

4.
Cellulose triacetate (CTA) membranes were fabricated via a modified nonsolvent induced phase separation (NIPS) method. Different solvent-nonsolvent compositions in first coagulation bath (FCB) were introduced to optimize CTA membrane structures. The effects of FCB compositions, immersion time and mass ratio of solvent (N-methyl-2-pyrrolidone, NMP) and nonsolvent (water, ethanol, ethylene glycol and glycerol) on membrane morphology and performance were systematically investigated. Prospective membranes with a dense bottom layer and a scaffold-like top layer were obtained under room temperature, owing to the low relative energy difference (RED) between nonsolvent and polymer as well as the high viscosity of FCBs. A high water flux Jv (12.6 L m?2 h?1) and a low reserve salt flux Js (1.32 g m?2 h?1) were obtained in the optimized membrane, with a structural parameter S of 119 μm. Compared with membranes prepared via conventional NIPS method and commercial CTA forward osmosis (FO) membranes, a remarkable improvement of Js/Jv value and S value was achieved, indicating membranes with single dense-layer structure might suffer less from internal concentration polymerization (ICP) which is the main obstacle for the development of FO process. This study might help us pave the way to design superior CTA membrane structures for forward osmosis application.  相似文献   

5.
The concentration of O2 in 1 M H2SO4 solution in the vicinity of the O2-evolving smooth Pt anode was measured as a function of anodic cdia using the galvanostatic potential—transient method.The solution near the O2-evolving anode was supersaturated with O2. When the anodic current was interrupted, supersaturated concentration C* decreased at a rate proportional to C*C0, where C0 is the solubility of O2 in the electrolyte at 1 atm. The rate constant of the decay of the supersaturation under the open circuit condition was measured to be 0.069 sec?1 at 25°C.At ia < 40 mA/cm2 there was a linear relation between log(C*C0) and log ia. At ia > 200 mA/cm2, however, the supersaturation exhibited a limiting value of 9.0 × 10?2 mol/l.  相似文献   

6.
Radioactive mercury, 203Hg, was deposited both on smooth Pt and amalgamated Pt electrodes in 1 N KOH, in the region of hydrogen bubble evolution. The relation found between the cd for deposition of mercury, iR, and for evolution of hydrogen, iG, is in accord with the equation derived previously[1] for systems with a surface incompletely covered with hydrogen bubbles. The heat transfer coefficient on a platinum electrode located on the wall of a rectangular channel was measured under defined hydrodynamic conditions during hydrogen evolution in 0·5 N KOH. The measured dependence of the heat flow and cd of hydrogen evolution is in accord with the relationship obtained earlier[1] for systems with fully covered surface by the evolved gas bubbles.  相似文献   

7.
Electrochemical oxidation and reduction of H2O2 on Ag were studied in alkaline solution of 10?3?0.3 M H2O2 and 2 × 10?3 ?1.0 M KOH under N2 bubbling. Steady i-φ curves obtained by a cyclic potential sweep method in a potential range where no electrode oxidation takes place, lead to the following results: (1) icd (A cm?2) (cathodic limiting current density) = 1.0 × [H2O2]1.0T (M), (2) i1d (A cm?2 (anodic limiting one) = icd ([KOH] ? [H2O2]T) or 1.0 × [KOH] < [H2O2]T), (3) φm (V) (mixed potential) = 0.126-0.060 log [KOH]1.0 and (4) (?φ/?i)φ=φm (Ωcm2) (reaction resistance at φ = φm) = 0.057 × [H2O2]?1.0T (M?1), where [H2O2]T designates a total H2O2 concentration and the others have their usual meanings.The above results are explained by the following mechanism; HO?2 formed by the reversible chemical reaction, H2O2 + OH ? HO?2 + H2O, is oxidised in anodic reaction by two steps: HO?2
HO2 (a) + e? and HO2(a) + OH? → O2 + H2O + e?, whereas in cathodic reaction, H2O2 is reduced by H2O2 + e?
OH(a) + OH?, OH(a) + e? → OH?. Here,
designates a rate determining step,Catalytic decomposition of H2O2 on the electrode is also discussed.  相似文献   

8.
The kinetics of the layer formation process and the corrosion process have been determined from the relation between the stationary (ic,0) and non-stationary (ic) corrosion rates and the rate of layer formation (iι) at passive iron in acid solutions. The rates of both processes depend on the potential difference ε2,3 at the passive-oxide/electrolyte-solution interface. The fact that ic and ic,0 are independent of the electrode potential is explained by a nearly constant composition of the passive oxide in contact with the solution (Fermi level within the band gap).Linear relations, such as log iι+ = a + (αι+c+)·log ic, are obtained, which are explained by charge-transfer hindrance. The apparent charge-transfer coefficients are αι+ = 1·43 and αι? = 0·57 for the layer formation and removal reaction, and αc+ = 0·84 for the corrosion reaction. From the pH dependence (pH = 0.35–2·90) of iι and ic,0 the reaction orders of the hydrogen ions are found to be approximately μnι+ = ?1 and νc = 0. The corrosion cd depends on the sulphate concentration, with a reaction order νs = ß = 0·16. From this, the kinetics of the oxide formation H2O·aq ? OH? · ox + H+·aq (preceding equilibrium), OH?·ox?O2?·ox + H+·aq (rate-determining) and the kinetics of the corrosion process Fe3+·ox + SO42?·aq ? FeSO4+·ad (adsorption equilibrium, Temkin conditions), FeSO4+·ad → FeSO4+·aq (rate-determining), with following dissociation of the complex can be deduced. The apparent charge-transfer coefficients are interpreted by αι+ = 1 + αι, αι? = 1 ? αι and αc+ = α + 2ßγs (gaι, α, true charge-transfer coefficients).  相似文献   

9.
Steady state anodic polarization curves were taken for Armco iron and in some experiments for high purity Puratronic iron in KOH solutions in the concentration range 5 x 10?2-5M. After the 30 min cathodic pretreatment, well reproducible anodic Tafel plots are obtained. The experimentally obtained diagnostic criteria ba = 67–70 mV dec?1, nOH? = 1.1 and nHFeO2?= ?0.45 are interpreted by the anodic reaction mechanism in which FeOHadsand Fe(OH)2,ads appear as the intermediates adsorbed under Temkin conditions, the primary stable product of the electrode reaction being HFeO2?, and the final product Fe (OH)2, formed by precipitation.  相似文献   

10.
The effects of inorganic electrolytes (NaCl, MgCl2, CaCl2) in aqueous solutions on oxygen transfer in a bubble column were studied. Electrolyte concentrations (c) below and above the critical concentrations for bubble coalescence (ctc), and six superficial gas velocities (vsg), were evaluated. The volumetric mass transfer (kLa) and the mass transfer (kL) coefficients were experimentally determined. It was found that the concentration of electrolytes reduced the kL, but the interfacial area (a) increased enough to result in a net increase of kLa. Using as independent variable a normalizing variable (cr = c/ctc), and maintaining fixed vsg, similar values of kLa were observed regardless the kind of electrolyte; the same happened for kL. This suggests that cr quantifies the structural effects that these solutes exert on mass transfer. Also, once cr = 1 was reached, no significant variations were found in kLa and kL for constant vsg. It is concluded that the gradual inhibition of bubble coalescence (cr < 1) governs the significant changes in hydrodynamics and mass transfer via the reduction of bubble size and the consequent increment of a and gas holdup (?g). Finally, regarding the effects of vsg on mass transfer, transition behaviors between those expected for isolated bubbles and bubble swarms were observed.  相似文献   

11.
Electrochemical reactions were assigned to the voltammetric waves obtained in alkaline KMnO4 and K2MnO4 solutions. Plots of ip (AC°ν12)?1 and iapicp?1 νs log ν indicate that at slow potential scan rates in MnO?4 solutions ranging from 0·07 to 0·19 F in NaOH the first step in MnO?4 reduction is a l-electron reversible charge transfer with no coupled chemical reaction; at fast scan rates (? V s?1 the process becomes quasi-reversible. The standard rate constant and the activation energy for the MnO?4/MnO2?4 charge transfer step were estimated. Plots of ip (AC°ν12)?1 and iapicp?1 νs log ν indicate that in 4·O F KOH and at potential scan rates between 0·005 and 2·V s?1 the reduction of MNO2?4 to MnO3?4 is a reversible charge transfer with no coupled chemical reaction. The diffusion coefficients of MnO?4 and MnO2?4 in alkaline solutions are reported.  相似文献   

12.
Repetitive potentio- and galvanokinetic traces of Ni rods yielded anodic peaks (or arrests) corresponding to probable formations of α and β Ni(OH)2 and NiO(OH) prior to O2 evolution. Cathodic half cycles consistently showed only two reduction processes that are attributed to NiO(OH)→ Ni(OH)2åNi. Potentiostatic polarization resulted in multi-peaks i-t traces, in [NaOH] ? 1 moll?1; the main peak exhibiting the general form characteristic of instantaneous nucleation processes. The maximum current intensity, ip, of the potentiokinetic traces as well as the parameter imtm of the i-t curves showed negligible dependence on NaOH in the range 0.01–1 moll?1. In the range of 1–200 mVs?1 scan rate, ipv. Employing an aqueous media of mixtures of different concentrations of NaCl plus a constant [NaOH] resulted in: (a) significant increases of both ip and imtm and (b) nearly linear i-vsol12 relation. Achievement of electrode passivity was retarted by Cl? but eventually attained. As [Cl?]/[OH?] exceeded ~ 3 passivity showed signs of breaking down. The passivation current showed continous rise with increasing [Cl?]. A mechanism of Cl? attack involving surface adsorption, increasing solubility of an intermediate Ni hydroxide species that nucleates into passive film and peptization of the deposited oxide, by Cl?, is discussed.  相似文献   

13.
On the basis of potential theory, the equations of motion and the associated pressure field are derived for an idealized growing spherical gas bubble rising in an inviscid liquid under the influence of gravity from a horizontal plate-orifice and from a free standing nozzle. The dimensionless pressure (pN?p0)/πga at the gas source N (where p0 is the undisturbed ambient pressure, a the instantaneous bubble radius and π the liquid density) is found to rise to a maximum when the dimensionless bubble position h/a = 1·55 for bubbles formed at a plate orifice and (125/48)13 ? 1·38 for bubbles formed at a free standing nozzle. These positions of maximum pressure correspond well to experimentally observed positions at which the gas supply to bubbles grown at constant flow rate in water is cut off by the collapse of the neck linking bubble and gas source. The volumes of bubbles at this instant are predicted theoretically and compare well with experimentally determined values.  相似文献   

14.
The pyroelectric materials have immense applications in the uncooled infrared thermal detectors. However, owing to increasing environmental concerns due to Pb element, it is required to explore novel, high-performance, environmental-friendly pyroelectric materials. This is the first study to report about the pyroelectric properties of lead-free NaNbO3 ceramics, which displayed a high pyroelectric coefficient of 1.85 × 10?8 C cm-2 K?1 and figures of merit as Fi = 0.67 × 10?10 m V?1, Fv = 3.33 × 10?2 m2 C?1, and Fd = 5.32 × 10-5 Pa?1/2 at room temperature. Also, highly temperature-stable pyroelectric characteristics were also observed in NaNbO3 ceramics due to the high depolarization temperature of 280 ℃. The high pyroelectric properties and temperature stability were a result of the electric field induced irreversible phase transition from antiferroelectric to ferroelectric. Hence, we can conclude that lead-free NaNbO3 ceramics are a novel and promising candidate for pyroelectric detectors in a wide temperature range, especially for large area detectors and pyroelectric point detector.  相似文献   

15.
Potentiostatic and galvanostatic pulse measurements were carried out to investigate the anodic oxygen evolution at platinum electrodes in 1N H2SO4 in dependence on the oxide layer thickness d and the electrode potential ε. The thickness d (1·5–10 Å) was obtained from cathodic charging curves. Further, the temperature dependence (0°–81°C) was evaluated from Bowden's measurements. Summarizing, the current io2 follows the relation, log i = A - (E0a - αFη)/2·3 RT - d/do. The experimental activation energy Eo = Eoa = αFη decreases linearly with increasing overvoltage η. The linear decrease of log i with increasing d, which is given by the term d/do, is correlated to the probability of the quantum mechanical tunnel transition of the electron from adsorbed ions, OH?ad or O2?ad respectively, through the oxide layer to the metal. Similar effects of the oxide layer thickness on the current density were observed in the case of the oxygen evolution at iridium, the CO-oxidation on platinum, and the reduction of Cl? and Ce4+ at platinum. In these cases a rate determining electron transfer through the oxide layer is also assumed.  相似文献   

16.
Electrochemical response of regio-random and regio-controlled poly(3-hexylthiophene), P3HexTh, was investigated by cyclic voltammetry. P3HexTh underwent electrochemical oxidation at about 0.4 V vs. Ag+/Ag in a THF solution, and the peak anode electric current, ipa, was proportional to the sweeping rate v; ipa=const×v1/2. These data indicated that diffusion of the P3HexTh molecule in the solution was important to determine ipa. Application of a Matsuda's equation with assumptions gave a diffusion coefficient, D, of about 1×10−7 cm2 s−1 at molecular weight of about 5000, and the D value steeply decreased with increase in the molecular weight.  相似文献   

17.
This study investigates optimization of various competitive adsorption parameters for removal of Cd(II), Ni(II) and Pb(II) from aqueous solutions by commercial activated carbon (AC) using the Taguchi method. Adsorption parameters such as initial metal concentration of each metal ion (C0,i ), initial pH (pH0), adsorbent dosage (m) and contact time (t) in batch technique were studied to observe their effects on the total adsorption capacity of metals onto activated carbon (q tot ). The adsorbent dosage has been found to be the most significant parameter. Interactions between C0,Cd ×C0,Ni , C0,Cd ×C0,Pb and C0,Ni ×C0,Pb have been considered for simultaneous metal ions adsorption. The optimum condition for adsorption of metal ions were obtained with C0,i =100 mg L?1, pH0=7, m=2 g L?1 and t=80 min. Finally, experimental results showed that a multi-staged adsorptive treatment would be necessary to reach the minimal discharge standards of metal ions in the effluent.  相似文献   

18.
The temperature dependence of various kinetic parameters for the copper electrodeposition reaction and for the ferrous to ferric ion oxidation has been experimentally determined. Potentiodynamic sweeps were carried out in a reactive electrodialysis cell in the 30- range in order to obtain expressions for each parameter as a function of temperature. In the conditions studied, for both reactions, the exchange current density depends on temperature as ln[i0(A/m2)]=-a/T+b and the limiting current density, as |iL(A/m2)|=cT-d. For the copper electrodeposition reaction, the charge transfer coefficient is constant, whereas for the ferrous to ferric ion oxidation it depends on temperature as α=fT2+gT+h (a-h are constants). These expressions can be used in a temperature-dependent model of a membrane-based copper electrowinning cell.  相似文献   

19.
The cocurrent upward mode was employed to absorb pure oxygen into water in bubble columns packed with Koch (Sulzer) motionless mixers. The liquid-side volumetric mass transfer coefficient, KLa, in the packed bubble column was found to be always larger than that in the unpacked bubble column. In the range of liquid velocities from 6.7 cm/sec to 39.9 cm/sec, the value of KLa in the packed bubble column increased with the increasing liquid velocity while that in the unpacked bubble column was almost independent of the liquid velocity. The equation of the formKLa= mνlβ? was successfully adopted to correlate the KLa data.  相似文献   

20.
C.K Chai  N.G McCrum 《Polymer》1982,23(4):589-597
Volume relaxation of isotactic polypropylene was studied at the temperature region of the α-relaxation by the technique of thermal stimulation (or temperature-jump). It is shown that the interpretation of volume relaxation experiments with viscoelastic models fails unless the limiting values of the bulk compliance are assumed to be temperature dependent. Much simplification in the theory of volume relaxation follows if the rigorous formalism of viscoelasticity is adopted, particularly if the treatment by one model is intended for the interpretation of controlled experiments at a single temperature, for both volume and mechanical relaxation. With this model an experiment is designed, based on the Boltzmann superposition principle. The activation enthalpy for volume relaxation, ΔHv, can then be determined if the Boltzmann principle is applied at an appropriate reduced time, tav(T), where av(T) is the Arrhenius shift factor. For polypropylene at 40°C ΔHv = 35 kcal mol1, equal to ΔHJ the activation enthalpy for creep (at 40°C).  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