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 46 毫秒
1.
The oxidant effect on the codeposition of two metals in a porous electrode (PE) was studied experimentally and by numerical calculations. The presence of an oxidant in a solution always leads to degradation of the dynamics and final characteristics of the target process, namely, to a decrease in the current efficiency, the final mass, and the average rate of metal deposition and to a less uniform distribution of metals in the PE. The effect depends on the PE operating conditions and involves less uniform potential distribution in PE and anodic dissolution of metals. The former component of the effect dominates when the concentrations of the metal ions and oxidant are comparable. A considerable deceleration of silver deposition in the presence of an oxidant and anodic dissolution of copper at the first stage of the electrolysis of the second and subsequent portions of solution were confirmed experimentally using as an example Ag and Cu extraction from an alkaline thiosulfate solution in the presence of a Fe3+ trilonate complex used as an oxidant.  相似文献   

2.
Using an extended dynamic model of liquid flow-through porous electrode (PE), the effect of kinetics of deposition of individual components and conditions of potentiostatic electrolysis on the dynamics and final parameters (the cathodic deposit weight, the ratio between the amounts of components, and the spatial distribution of components) of codeposition of two metals M1 and M2 is studied. An equipotential PE operating in the direct-flow mode in the absence of anodic dissolution of electronegative component M2 is considered. The effects of concentration, exchange currents, a difference between the equilibrium potentials M1 and M2, a prescribed voltage on PE, and solution flow velocity and direction are analyzed. It is shown that, for this version of codeposition of metals, the rates of M1 and M2 deposition averaged over the PE width are constant in time. However, this does not mean that their local deposition rates are constant. The general tendency is that the metal deposition rate on the rear part of PE decreases with the time, whereas the deposition rate on the frontal zone of PE, which is closer to the anode, increases. As a result, the final profiles for M1 and M2, which are calculated for equal deposition times taking into account and ignoring the redistribution of current during the deposition, differ essentially.  相似文献   

3.
The effect of the solution volume and deposition potential on the dynamics and final parameters of the deposition process was studied using a mathematical model of codeposition of the two metals within a one-dimensional flow porous electrode (PE). The initial PE characteristics, kinetic parameters of the overall polarization curve, and volume solution flow rate were fixed. It was found that a decrease in the solution volume at a high flow rate apart from the expected consequences (a decrease in the metal deposition rate in time and an increase in the time of the porous matrix filling by the deposit) results in a significant improvement of the uniformity of distribution of electropositive component M1 and a dramatic extension of the deposition region of electronegative component M2. The positive effect of a decrease in the solution volume is enhanced at an increase in the cathodic potential and the closing in of partial currents of the deposited metals. Eventually, these conditions provide simultaneously a high rate of metal recovery and maximum PE filling by the cathodic deposit.  相似文献   

4.
The characteristics of the copper deposition process from electrochemical cells using a solution of Cu(CF3SO3)2 in low molecular weight poly(ethylene glycol), PEG, have been investigated by cyclic voltammetry, potential steps and impedance spectroscopy. The results suggest that under non-equilibrium at the cathodic side the process mainly involves the reduction of Cu(II) ions to copper metal while at the anodic side a chemical reaction competes with the electrochemical dissolution.  相似文献   

5.
The earlier developed dynamic model of a flow-through electrode is used for studying how the variations in initial conductivity of a porous matrix κs,ini and a metal deposit affect the rate of metal deposition from an oxidant-containing solution for the direct-flow operation mode of the porous electrode. It is found that in contrast to an oxidant-free solution in which the decrease of κs,ini improves the uniformity of deposit distribution inside the porous cathode and increases the deposit final mass m f, the opposite situation is observed in the presence of an oxidant, namely, a decrease in κs,ini, under otherwise similar conditions reduces the deposit mass and leads to its specific spatial distribution. The final metal deposit is divided into two separate fragments (rear and front) with a region of low conductivity of the initial porous matrix in between. Dynamics of the current and metal redistribution within the porous electrode, the reasons for the formation and stabilization of the rear fragment of coating, the correlation between the metal deposition rate and changes in the anodic zone position and intensity are discussed. It is shown that with the appearance of a specific profile of deposit distribution, the dependence of m f on the metal conductivity develops a limit that differs considerably from the deposit final mass for an equipotential porous electrode.  相似文献   

6.
The influence of the ligand/metal ion concentration ratio on the shape, peak current and peak potential of curves obtained by anodic stripping voltammetry (a.s.v.) at the hanging mercury drop electrode is described, particularly with respect to the use of a.s.v. for speciation of metal ions at very low concentrations as is often found in natural waters. The lead(II)/triethylenetetramine system is used as a model of a fully labile reversible system. It is shown that the total metal ion concentration at the electrode surface (CoM) during the stripping step may be much larger (30–300 times in typical conditions) than that in the bulk solution (CM), the exact value depending on the deposition time td. Consequently, changes in the peak characteristics are observed when the ligand/metal concentration ratio in the bulk of the solution, CL/CM, is less than 1000. Semi-empirical equations, experimentally tested, are given, which enable CoM/CM to be estimated for a specified solution and a.s.v. conditions, which correct for the “surface concentration effect” when a.s.v. is used to measure complexation, and which describe the influence of the parameters such as stirring efficiency, radius of the mercury drop and CL/CM. The implications of the results are discussed for determinations of total metal ion in complex media, of speciation based on peak-potential shifts or stripping voltammetric curves, and of complexation capacity.  相似文献   

7.
The influence of hydrogen sulfide (10–100 mg/1) on the Armco iron anodic dissolution in an aerated 0.17 M Na2SO4 solution is investigated. During a potentiostatic anodic polarization, the hydrogen sulfide introduction makes the current increase stepwise. The magnitude of the increase depends on the duration of preliminary anodic polarization, electrode potential, and hydrogen sulfide concentration. The anodic metal dissolution activation by hydrogen sulfide is explained by chemical conversion of the oxide-hydroxide passivating film into iron sulfide that is generated at the metal surface in the form of a porous film and does not hinder the electrode dissolution. Dedicated to the ninetieth anniversary of Ya.M. Kolotyrkin’s birth.  相似文献   

8.
The deposition-dissolution mechanism of lithium on stainless steel and calcium electrodes in 1 M LiAlCl4 -thionyl-chloride solution is studied by pulse galvanostatic and ac techniques. The metal -solution interfacial capacitance of the stainless steel electrode is about 30 μF cm?2 which is higher by an order of magnitude than the capacitance of lithium-coated stainless steel and either pure or lithium-coated calcium. The lower capacitance is attributed to the existence of a solid electrolyte interphase (SEI) on the coated stainless steel or the calcium electrode.Significantly different is observed upon deposition of lithium on stainless steel or calcium. Deposition on stainless steel takes place only after prior formation of a SEI on the electrode (by passage of about 20 mC cm?2), while deposition on calcium starts immediately after the electrode capacitance has been charged (by about 5 μC cm?2). Furthermore, deposition of about 3% of a monolayer of lithium on calcium is enough to stabilize its potential at 0.0 V vs. LiRE.On the lithium-coated stainless steel electrode, a linear relationship between the current and over-potential is observed for up to 700 mV. This indicates a Tafel slope > V. During lithium deposition on stainless steel, the SEI resistivity is about 1.5 × 107 Ω-cm and its thickness is about 10 nm.Under open circuit potential, the deposited lithium corrodes at an apparent rate of 100 μA cm?2. Rapid fluctuations of the electrode potential during the corrosion or dissolution process are accounted for by a break and repair mechanism of metallic contact between lithium deposited within the SEI and the current collector.  相似文献   

9.
The regularities of electrochemical deposition and dissolution of gold in the mixed sulfite-thiocarbamide electrolytes in the absence and in the presence of sodium sulfide additive are studied by using the voltammetric measurements on a renewable electrode and quarts microgravimetry. It is shown that, in the cathodic metal deposition, an addition of sodium sulfide promotes the depolarization effect, which is caused by the presence of thiocarbamide in the solution. Under the anodic polarization of gold in the mixed sulfite-thiocarbamide solution with pH < 10, the gold dissolution rate is insignificant. An addition of 10?5 M Na2S to this solution dramatically accelerates the process. At pH > 10, the gold dissolution in the sulfite-thiocarbamide electrolyte is observed even in the solution free of Na2S additive. It is evidenced that this is associated with spontaneous accumulation of sulfide-containing species in the solution, probably, as a result of thiocarbamide hydrolysis; the rate of hydrolysis steeply increases with increasing pH value.  相似文献   

10.
The voltammetric responses of copper and silver had been extensively studied and compared in a variety of non-aqueous solvents such as acetonitrile (AN), propylene carbonate (PC) and sulfolane containing two different supporting electrolytes namely triethylaminetrishydrogen fluoride (TEA.3HF) and tetrabutylammonium tetrafluoroborate (TBABF4). The dissolution rate and surface transformation on the electrode surfaces as a result of anodic polarization was investigated using atomic absorption spectroscopy (AAS) and scanning electron microscopy (SEM), respectively. In solvent-free TEA.3HF medium, the copper electrode shows high charge recovery ratio (Q c/Q a), and the difference between the initial anodic and cathodic potentials, obtained at a current density of 2 mA cm−2, is around 0.11 V, suggesting that in this medium, Cu can certainly serve as reference electrode. On the other hand, on Ag electrode, substantial dissolution was observed leading to very high anodic (Q a) and cathodic (Q c) charges, and the surface morphology after the cyclic polarization results in roughened surface with large pores. The effects of incorporating AN and water as additives in TEA.3HF on the solubility and stability of these metal fluoride films are also reported. The dissolution pattern and film formation behavior of these two metals in the different solvents containing fluoride and fluoroborate ionic species have several qualitative similarities, as noted from cyclic voltammetry responses and SEM morphology. Anodic dissolution and precipitation process for both Cu and Ag depends significantly on the nature of supporting electrolytes as well as solvents. In AN containing 0.1 M TEA.3HF, the dissolution of Cu and Ag electrodes was very high. Fluoride salts of Cu show lesser solubility than Ag in those solvents, while fluoroborate salts exhibit the reverse trend. The AAS data suggest that for a particular salt, which may be either fluoride or fluoroborate of Cu and Ag, the relative solubility decreases in the order AN > PC > sulfolane.  相似文献   

11.
Electrode materials for supercapacitors are at present commonly evaluated and selected by their mass specific capacitance (CM, F g−1). However, using only this parameter may be a misleading practice because the electrode capacitance also depends on kinetics, and may not increase simply by increasing material mass. It is therefore important to complement CM by the practically accessible electrode specific capacitance (CE, F cm−2) in material selection. Poly[3,4-ethylene-dioxythiophene] (PEDOT) has a mass specific capacitance lower than other common conducting polymers, e.g. polyaniline. However, as demonstrated in this communication, this polymer can be potentiostatically grown to very thick films (up to 0.5 mm) that were porous at both micro- and nanometer scales. Measured by both cyclic voltammetry and electrochemical impedance spectrometry, these thick PEDOT films exhibited electrode specific capacitance (CE, F cm−2) increasing linearly with the film deposition charge, approaching 5 F cm−2, which is currently the highest amongst all reported materials.  相似文献   

12.
A new method for the microdetermination of tungsten in organic compounds such as phosphotungstate and silicotungstate of basic organic compounds is described. The sample is burnt in a modified oxygen flask, the products are absorbed in a sodium hydroxide solution, and complete dissolution of WO3formed is achieved by boiling the solution for 6–7 min. After removal of carbon dioxide and appropriate pH adjustment, W(VI) is titrated potentiometrically with 10 mMPb(NO3)2by using a lead(II) ion-selective electrode as indicator electrode and a double-junction reference electrode. Accurate and precise results are obtained. The absolute error does not exceed 0.14%; the recoveries of tungsten are in the range from 99.74 to 100.00%. The mean relative standard deviation is 0.10%.  相似文献   

13.
The interaction between the Co sulfide coating formed on a glassy carbon electrode and Cu(I)-ammonia complexes solution was investigated by cyclic voltammetry in 0.1 M KClO4, 0.1 M NaOH and 0.05 M H2SO4 solutions. It was determined that, after treating the cobalt sulfide coating formed by two deposition cycles with Cu(I)-ammonia complexes (0.4 M, pH 8.8–9.0, τ=180 s, T=25±1°C), an exchange occurs between the coating components and Cu(I). Copper(I) substitutes 75% of the Co(III) compounds present in the coating (~1.81×10–7 mol cm–2) because of Cu2O (1.36×10–7 mol cm–2) formation. The rest of the Co(II) and Co(III) sulfide compounds are also replaced by copper with formation of Cu2– x S with a stoichiometric coefficient close to 2 (~1.9). After modifying the cobalt sulfide coatings with Cu(I) ions, the total amount of metal (Co+Cu) increases, owing to the sorption of Cu(I) compounds. In addition, the number of deposition cycles decreases from 3 to 1.5 [1 cycle involves cobalt sulfide layer formation and 0.5 cycle is attributed to modifying by Cu(I) ions]. The coatings modified in the above-mentioned manner may be successfully used for plastic electrochemical metallization as Cu2– x S coatings formed by three deposition cycles. Electronic Publication  相似文献   

14.
The dissolution of Cu in solutions of HNO3 of different concentrations has been studied by the thermometric method. Starting from the initial temperature, Ti, the temperature—time curves exhibit an induction period followed by a rapid rise in temperature to a maximum value, Tm, attained t min after the start of the reaction. Tm increases and t decreases with increase of the acid concentration, M. ΔT (i.e.Tm ? Ti) and the reaction number (R.N. = (Tm ? Ti)/t) vary with M according to: ΔT = k(M ? M0) and R.N. = A1Mn, where k, M0, A1 and n are constants.The effect of varying concentrations of HCl, H2SO4 and H3PO4 on the R.N. of Cu in 3.5 M HNO3 was examined. Small amounts of these acids lower the R.N. (inhibition) due to the displacement of an active species on the surface of the metal by the anion of the acid. Larger additions of the acids accelerate dissolution. The concentration at which the added acid changes from corrosion-inhibitor to accelerator varies as HCl < H2SO4 < H3PO4. This sequence is considered to parallel the strength of adsorption of the respective anions. The results of experiments with salt additions confirm this view; all salts act only as dissolution-retardants. Calculations pertaining to the effect of the various ions on the R.N. support the conclusion that the dissolution of Cu in HNO3 is autocatalytic in nature, and depends on the [H+]/[NO3?] ratio.Cu does not dissolve in air-free, cold HCl. Attack takes place, however, in the presence of KNO3. Under these conditions attack is of the pitting- rather than the general type. The temperature rises suddenly after an incubation period, which decreases in length with increase of the amount of the added salt.Proof of the involvement of HNO2 in the autocatalytic cycle of Cu dissolution in HNO3 is obtained from the results of urea additions to the solution.  相似文献   

15.
The generation of porous metal oxides by removal of template molecules from inorganic polymers formed by sol-gel type hydrolysis and condensation of metal alkoxides is described. The template molecules include organic polymers, copper (II) ions in hybrid copper oxide/silica sols and copper (II) bis(hexafluorocetylacetonate) (hfac). Neutron scattering experiments on the system in which polyacrylic acid (Mw=2,000 Daltons) is used as an organic template to generate microporous tin oxide show that removal of the template generates skeletal voids. In a second series of experiments, mixed copper/silicon oxide xerogels were prepared by hydrolysis of mixtures of Si(OEt)4 and Cu(OCH2CH(CH3)N(CH3)H)2 in the ratios of Si:Cu=2:1, 4:1, 9:1. Selective removal (etching) of the copper component generates porous silica. Neutron scattering data and BET surface area measurements are consistent with the creation of pores with molecular dimensions (micropores, 10 Å or less). In the third strategy, Si(OEt)4 is hydrolyzed in the presence of Cu(hfac)2, a volatile, inert inorganic template, in a 4 to 1 molar ratio. Removal of the template from the xerogel at 100°C in vacuo affords microporous silica.  相似文献   

16.
The synthesis of linear symmetric ethynyl- and acetylide-amidinates of the coinage metals is presented. Starting with the desilylation of the complexes [{Me3SiC≡CC(NDipp)2}2M2] (Dipp=2,6-diisopropylphenyl) (M=Cu, Au) it is demonstrated that this compound class is suitable to serve as a versatile metalloligand. Deprotonation with n-butyllithium and subsequent salt metathesis reactions yield symmetric tetranuclear gold(I) acetylide complexes of the form [{(PPh3)AuC≡CC(NDipp)2}2M2] (M=Cu, Au). The corresponding Ag complex [{(PPh3)AuC≡CC(NDipp)2}2Ag2] was obtained by a different route via metal rearrangement. All compounds show bright blue or blue-green microsecond long phosphorescence in the solid state, hence their photophysical properties were thoroughly investigated in a temperature range of 20–295 K. Emission quantum yields of up to 41 % at room temperature were determined. Furthermore, similar emissions with quantum yields of 15 % were observed for the two most brightly luminescent complexes in thf solution.  相似文献   

17.
By means of earlier developed dynamic model of porous electrode, numerical analysis is given of the effect of the division into smaller parts of optimal volume V opt of metal-containing solution circulating through the porous electrode, which ensures both achieving of the metal preset recovery and the prescribed final porosity in the critical cross-section of the porous cathode, most strongly filled with the deposit. It is shown that the result depends significantly on the presence (or absence) of oxidant ions in the solution. In the absence of oxidant, the division of V opt is desirable; it entails some improving of the distribution uniformity and increase in the deposit eventual mass. On the contrary, in the presence of oxidant such a division brings about marked negative consequences: (1) the metal deposit stronger localization at the porous electrode front edge and (2) decrease in the deposit’s final mass. It was shown that these phenomena result from the formation of anodic zone in the porous electrode rear side in the initial stage of the processing of the second and subsequent portions of solution. This is promoted by low polarization in this part of the porous electrode and the presence of a soluble metal deposited here during the final stage of electrolysis of the solution preceding portion.  相似文献   

18.
Adding a microscopic quantity of sodium sulfide (~10?5 M) into acid solutions of thiourea leads to a dramatic acceleration of anodic dissolution of gold. The acceleration effect is greater at larger thiourea concentrations (c) and longer times of the electrode contact with solution (Δt) before the beginning of measurements. The effect diminishes after a polarization curve passes through a maximum at E ? 0.5 V. Regularities of the gold dissolution in a solution containing 0.1 M thiourea and 0.5 M H2SO4 at given values of c and Δt are studied with use made of the technique of renewing the electrode surface by cutting off a thin surface layer of metal. The discovered regularities are given an explanation which is based on the assumption that the dissolution process is catalyzed by sulfide ions adsorbed on the electrode surface.  相似文献   

19.
The mechanism of alkaline earth metal tungstate formation during soft solution processing was studied by cyclic voltammetry, electrochemical impedance spectroscopy and by direct in situ observation of the surface changes using atomic force microscopy. The electrochemical oxidation of W to WO3 was followed by dissolution of WO3 and, with some delay, by precipitation of tungstates at the metal surface. The same Tafel slopes observed in Li+, Ba2+, Sr2+ and Ca2+ containing solutions indicate that the course of the oxidation process is independent of the cation present in solution. The observed differences in the current-voltage curves outside the Tafel region are accounted for by the different film-forming tendencies of the various alkaline earth metal cations. The growth of tungstate layers at the W substrate decreases the electrochemically active area and limits the production of WO4 2– at later stages of deposition. At low potentials (E<0.2 V) the oxidation of W is the rate-controlling step. At higher potentials, however, the dissolution process slows down due to a relative decrease of the pH in the electrode vicinity and dissolution becomes the rate-limiting step. Electronic Publication  相似文献   

20.
Lability concepts are elaborated for metal complexes with soft (3D) and hard (2D) aqueous nanoparticles. In the presence of a non-equilibrium sensor, e.g. a voltammetric electrode, the notion of lability for nanoparticulate metal complexes, M-NP, reflects the ability of the M-NP to maintain equilibrium with the reduced concentration of the electroactive free M2+ in its diffusion layer. Since the metal ion binding sites are confined to the NP body, the conventional reaction layer in the form of a layer adjacent to the electrode surface is immaterial. Instead an intraparticulate reaction zone may develop at the particle/medium interface. Thus the chemodynamic features of M-NP complexes should be fundamentally different from those of molecular systems in which the reaction layer is a property of the homogeneous solution (μ?=?(D M/k a )1/2). For molecular complexes, the characteristic timescale of the electrochemical technique is crucial in the lability towards the electrode surface. In contrast, for nanoparticulate complexes it is the dynamics of the exchange of the electroactive metal ion with the surrounding medium that governs the effective lability towards the electrode surface.  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