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 352 毫秒
1.
The regularities of silver anodic dissolution are studied by using the voltammetry (at the potential scan rates from 5 to 1000 mV/s) on the electrode, which was renewed immediately in the solution by cutting-off a thin surface metal layer, and quartz microgravimetry, for various concentrations of sodium thiosulfate (0.05–0.2 M). It is shown that, in the potential range from 0 to 0.4 V (normal hydrogen electrode), the polarization curves reflect the silver dissolution, whereas the contribution of oxidation of S2O32− ions is insignificant. At low potential scan rates, the process kinetics is of mixed nature. The kinetics and mechanism of anodic process are studied by using the measurements at high potential scan rates (100–200 mV/s) and the calculations of equilibrium composition of near-electrode layer. It is found that the exchange current in the electrolytes studied is 5 × 10−5 A/cm2, the transfer coefficient α is approximately 0.5, and both parameters are virtually independent of the concentration of S2O32− ions. The reaction order of silver dissolution with respect to the ligand $ \left. {\frac{{\partial logi}} {{\partial logc}}} \right|_E $ \left. {\frac{{\partial logi}} {{\partial logc}}} \right|_E is close to unity and is independent of potential. With regard for the literature data on the adsorption of thiosulfate ions on silver, this result is interpreted as the evidence for the involvement of one S2O32− ion from bulk solution, along with adsorbed ligands, in the elementary act of metal dissolution.  相似文献   

2.
A method for estimating the critical temperatures (T b) of thermal explosion for energetic materials is derived from Semenov’s thermal explosion theory and the non-isothermal kinetic equation dα/dt=A 0 T B f(α)e−E/RT using reasonable hypotheses. The final formula of calculating the value of T b is $ \left( {\frac{B} {{T_b }} + \frac{E} {{RT_b^2 }}} \right) $ \left( {\frac{B} {{T_b }} + \frac{E} {{RT_b^2 }}} \right) (T bT e0=1. The data needed for the method, E and T e0, can be obtained from analyses of the non-isothermal DSC curves. When B=0.5 the critical temperature (T b) of thermal explosion of azido-acetic-acid-2-(2-azido-acetoxy)-ethylester (EGBAA) is determined as 475.65 K.  相似文献   

3.
Density of the water-ethylene glycol binary mixtures was measured in the entire range of compositions in the temperature range 278–333.15 K (6 values) at atmospheric pressure using a vibration densimeter. Mixtures with low concentrations of ethylene glycol were studied at 15 temperatures in the range of 274–333.15 K. Excess molar volumes V m E , the partial molar volumes of water -V 1 and ethylene glycol, -V 2, the coefficients of thermal volume expansion α of the mixture, the partial molar volume coefficients of thermal expansion of water $ \bar V_1 $ \bar V_1 and ethylene $ \bar V_2 $ \bar V_2 were calculated. Excess molar volumes were described using the Redlich-Kister equation. The density ρ of the mixture was found to increase with the increasing ethylene glycol concentration at all temperatures, but at low content of ethylene glycol the dependence ρ = f(T) of the mixture at ∼276.5 K passed through a maximum. The coefficient α increases sharply in the composition range 0 < x < 0.2, in the range 0.5 < x <1 remains almost unchanged, and at T > 277 K is positive for all compositions. The dependences $ \bar \alpha _1 $ \bar \alpha _1 = f (x) and $ \bar \alpha _2 $ \bar \alpha _2 = f (x) are complex in whole temperature range and are characterized by the presence of an extremum. V m E values are negative at all temperatures, and upon increase in the temperature the deviation from ideality decreases (x is the mole fraction of ethylene glycol).  相似文献   

4.
Potentiometric method was used to measure the redox potentials of Tm3+/Tm2+ in a eutectic melt of sodium, potassium, and cesium chlorides relative to a chlorine reference electrode in the temperature range 823–973 K. The main thermodynamic characteristics of the redox reaction TmCl2(solution) + 1/2Cl2(g) ⇆ TmCl3(solution) were calculate from the conditional standard potentials $ E_{{{Tm^{3 + } } \mathord{\left/ {\vphantom {{Tm^{3 + } } {Tm^{2 + } }}} \right. \kern-\nulldelimiterspace} {Tm^{2 + } }}}^* $ E_{{{Tm^{3 + } } \mathord{\left/ {\vphantom {{Tm^{3 + } } {Tm^{2 + } }}} \right. \kern-\nulldelimiterspace} {Tm^{2 + } }}}^* .  相似文献   

5.
Electrospray ionization coupled with low energy collision induced dissociation (CID) in an ion trap mass spectrometer was used to examine the fragmentation patterns of the [M + Na]+ of eight pairs of heptapeptides containing α- or β-Asp residues in second and sixth amino acid positions, respectively. Selective cleavages at the peptide backbone C-terminal to two Asp residues were observed, which generated a series of C-terminal y5 ions and N-terminal b6 ions. Two typical ions: [ \texty5 + \textNa-\textH ] + {\left[ {{{\text{y}}_{{5}}} + {\text{Na}}-{\text{H}}} \right]^{ + }} and [ \textb6 + \textNa + \textOH ] + {\left[ {{{\text{b}}_{{6}}} + {\text{Na}} + {\text{OH}}} \right]^{ + }} , produced by α-Asp containing peptides were noted to be much more abundant than those of the peptides with β-Asp, which could be used for distinction of the isomers in Asp2 and Asp6, respectively. In addition, a series of internal ions generated by simultaneous cleavages at Asp residues were detected. Competitive reactions of carboxylic groups occurred between Asp6 side chain and C-terminus. Formation mechanisms of most product ions are proposed. The results obtained in this work are significant since low energy CID has been demonstrated to be effective for the distinction of Asp isomers.  相似文献   

6.
Cathodic oxygen reduction on the XC-72R carbon black modified by the products of pyrolysis of cobalt 5,10,15,20-tetrakis(4-methoxyphenyl)porphyrin (CoTMPP) (XC-72M) was studied in acidic and neutral electrolytes. Formation of new active centers on XC-72M is confirmed by voltammetric curves (specific charge density grows as compared to the XC-72R carbon black by 2–2.5 times) using the methods of rotating disk electrode (a shift in half-wave potential E 1/2 by 600 mV) and rotating ring-disk electrode (the fraction of the direct reaction increases to 70%). Herewith, the $\frac{{\partial E_{1/2} }} {{\partial pH}}$\frac{{\partial E_{1/2} }} {{\partial pH}} value in the range of pH 0.3–8.5 is −60 mV. It is shown that proton necessarily participates in the slow stage of the first electron transfer for the further occurrence of the direct reaction to water. At a transition from acidic solutions to neutral ones, the polarization curves converge for XC-72M and XC-72R, which is due to a decrease in the concentration of proton in the solution and variation of the mechanism of the oxygen reduction slow stage.  相似文献   

7.
The content of oxygen in Ca0.6 − y Sr0.4La y MnO3 − δ, where y = 0 and 0.05, was determined by coulometric titration over the temperature range 650–950°C at oxygen partial pressure in the gas phase varied from 10−4 to 1 atm. The results were used to calculate the partial molar enthalpy, Δ$ \bar H $ \bar H O(δ), and entropy, Δ$ \bar S $ \bar S O(δ), of oxygen in manganites. Changes in the Δ$ \bar H $ \bar H O(δ) and Δ$ \bar S $ \bar S O(δ) dependences caused by the introduction of lanthanum are evidence of the formation of local clusters of the double perovskite type in the Ca0.6Sr0.4MnO3 − δ matrix.  相似文献   

8.
The mechanism of the Co(II) catalytic electroreduction of water insoluble CoR2 salt in the presence of cysteine was developed. CoR2 = cobalt(II) cyclohexylbutyrate is the component of a carbon paste electrode. Electrode surface consecutive reactions are: (a) fast (equilibrium) reaction of the complex formation, (b) rate-determining reversible reaction of the promoting process of CoR(Ac+) complex formation, (c) rate-determining irreversible reaction of the electroactive complex formation with ligand-induced adsorption, and (d) fast irreversible reaction of the electroreduction. Reactions (a,b) connected with CoR2 dissolution and reactions (c,d) connected with CoR2 electroreduction are catalyzed by . Regeneration of (reactions “b,d”) and accumulation of atomic Co(0) (reaction “d”) take place. Experimental data [Sugawara et al., Bioelectrochem Bioenergetics 26:469, 1991]: i a vs E (i a is anodic peak, E is cathodic accumulation potential), i a vs , and i a vs pH have been quantitatively explained.  相似文献   

9.
The electrochemical properties of boron-doped diamond (BDD) polycrystalline films grown on tungsten wire substrates using ethanol as a precursor are described. The results obtained show that the use of ethanol improves the electrochemistry properties of “as-grown” BDD, as it minimizes the graphitic phase upon the surface of BDD, during the growth process. The BDD electrodes were characterized by Raman spectroscopy, scanning electronic microscopy, cyclic voltammetry (CV), and electrochemical impedance spectroscopy (EIS). The boron-doping levels of the films were estimated to be ∼1020 B/cm3. The electrochemical behavior was evaluated using the and redox couples and dopamine. Apparent heterogeneous electro-transfer rate constants were determined for these redox systems using the CV and EIS techniques. values in the range of 0.01–0.1 cm s−1 were observed for the and redox couples, while in the special case of dopamine, a lower value of 10−5 cm s−1 was found. The obtained results showed that the use of CH3CH2OH (ethanol) as a carbon source constitutes a promising alternative for manufacturing BDD electrodes for electroanalytical applications.  相似文献   

10.
The data on the permittivities of crystalline 1,4-dithiane and 1,4-dithiane-1,4-dioxide were used to calculate the molar deformation polarizations of the C-S (P ∞, C-S = 3.84 cm3/mol) and S$ \underline \ldots $ \underline \ldots O ($ P_{\infty ,S\underline \ldots O} $ P_{\infty ,S\underline \ldots O} = 4.34 cm3/mol) bonds within the framework of the additive scheme suggested by Levin. These data were used to calculate the deformation permittivity of dimethylsulfoxide (DMSO) at 298.15 K, ɛ∞, DMSO = 2.36, and the dipole correlation factor of pure DMSO, g dip = 1.055.  相似文献   

11.
The heat capacity and density of solutions of sodium and potassium perchlorates in N-methylpyrrolidone (MP) at 298.15 K were studied by calorimetry and densimetry. The standard partial molar heat capacities $ \bar C_{p2}^ \circ $ \bar C_{p2}^ \circ and volumes $ \bar V_2^ \circ $ \bar V_2^ \circ of NaClO4 and KClO4 in MP were calculated. The standard heat capacities $ \bar C_{pi}^ \circ $ \bar C_{pi}^ \circ and volumes $ \bar V_i^ \circ $ \bar V_i^ \circ of the perchlorate ion in an MP solution at 298.15 K were determined. The results are discussed with allowance for the specifics of solvation in the solutions of the salts under study. The coordination number of the ClO4 ion in an MP solution at 298.15 K was calculated.  相似文献   

12.
Colloidal indigo is reduced to an aqueous solution of leuco-indigo in a mediated two-electron process converting the water-insoluble dye into the water-soluble leuco form. The colloidal dye does not interact directly with the electrode surface, and to employ an electrochemical process for this reduction, the redox mediator 1,8-dihydroxyanthraquinone (1,8-DHAQ) is used to transfer electrons from the electrode to the dye. The mediated reduction process is investigated at a (500-kHz ultrasound-assisted) rotating disc electrode, and the quantitative analysis of voltammetric data is attempted employing the Digisim numerical simulation software package. At the most effective temperature, 353 K, the diffusion coefficient for 1,8-DHAQ is (0.84±0.08)×10−9 m2 s−1, and it is shown that an apparently kinetically controlled reaction between the reduced form of the mediator and the colloidal indigo occurs within the diffusion layer at the electrode surface. The apparent bimolecular rate constant k app=3 mol m−3 s−1 for the rate law \fracd[ \textleuco - \textindigo ] dt = k\textapp ×[ \textmediator ] ×[ \textindigo ]\frac{{d{\left[ {{\text{leuco}} - {\text{indigo}}} \right]}}} {{dt}} = k_{{{\text{app}}}} \times {\left[ {{\text{mediator}}} \right]} \times {\left[ {{\text{indigo}}} \right]} is determined and attributed to a mediator diffusion controlled dissolution of the colloid particles. The average particle size and the number of molecules per particles are estimated from the apparent bimolecular rate constant and confirmed by scanning electron microscopy.  相似文献   

13.
For getting an insight into the mechanism of atmospheric autoxidation of sulfur(IV), the kinetics of this autoxidation reaction catalyzed by CoO, Co2O3 and Ni2O3 in buffered alkaline medium has been studied, and found to be defined by Eqs. I and II for catalysis by cobalt oxides and Ni2O3, respectively.
(I)
(II)
The values of empirical rate parameters were: A{0.22(CoO), 0.8 L mol−1s−1 (Co2O3)}, K 1{2.5 × 102 (Ni2O3)}, K 2{2.5 × 102(CoO), 0.6 × 102 (Co2O3)} and k 1{5.0 × 10−2(Ni2O3), 1.0 × 10−6(CoO), 1.7 × 10−5 s−1(Co2O3)} at pH 8.20 (CoO and Co2O3) and pH 7.05 (Ni2O3) and 30 °C. This is perhaps the first study in which the detailed kinetics in the presence of ethanol, a well known free radical scavenger for oxysulfur radicals, has been carried out, and the rate laws for catalysis by cobalt oxides and Ni2O3 in the presence of ethanol were Eqs. III and IV, respectively.
(III)
(IV)
For comparison, the effect of ethanol on these catalytic reactions was studied in acidic medium also. In addition, alkaline medium, the values of the inhibition factor C were 1.9 × 104 and 4.0 × 10L mol−1 s for CoO and Co2O3, respectively; for Ni2O3, C was only 3.0 × 102 only. On the other hand, in acidic medium, the values of this factor were all low: 20 (CoO), 0.7 (Co2O3) and 1.4 (Ni2O3). Based on these results, a radical mechanism for CoO and Co2O3 catalysis in alkaline medium, and a nonradical mechanism for Ni2O3 in both alkaline and acidic media and for cobalt oxides in acidic media are proposed.  相似文献   

14.
Kinetic order-disorder phase transitions (space group space group $ I\bar 4 $ I\bar 4 ↔ space group I41/a) have been considered for nonactivated and activated scheelite compounds (Na0.5Gd0.5)WO4 (NGW), (Na0.5Gd0.5)MoO4 (NGM), (Na0.5La0.5)WO4(NLW), and (Na0.5La0.5)MoO4 (NLM) synthesized by the Czochralski method and structural and growth sources of crystal dissymmetrization have been suggested. For NGW, it was shown that an increase in the difference between the content of Gd and Na in two positions of the structure with space group $ I\bar 4 $ I\bar 4 and their ratio leads to an increase in the deviation from centrosymmetry. On the basis of available literature data and our results, it was demonstrated that the degree of order depends on the initial composition of the reaction mixture, crystal growth and cooling rates, activator concentrations, and postgrowth treatment conditions. The inconsistency between X-ray diffraction data and asynchronous second harmonic generation studies was explained by the possibility of formation of centrosymmetric superstructures and/or local ordering of atoms.  相似文献   

15.
The partial mixing enthalpies of the components (Δm $ \bar H $ \bar H i ) of the Ni-Ga melts were measured using the high-temperature isoperibolic calorimetry at 1770 ± 5 K in wide concentration interval. The limiting partial mixing enthalpy of gallium in a liquid nickel (Δm $ \bar H $ \bar H Ga) is −95.5 ± 19.8 kJ mol−1, and similar function of nickel in liquid gallium (Δm $ \bar H $ \bar H Ni) is −74.5 ± 16.4 kJ mol−1. The integral mixing enthalpy of liquid alloys of this system was calculated from partial enthalpies for the whole concentration area (Δm H min = −32.1 ± 2.7 kJ mol−1 at x Ni = 0.5). The Δm H value of liquid nickel-gallium alloys independence of the temperature is confirmed. Enthalpies of formation of liquid (Δm H) phases of Ni-Ga system were compared with ones for solid (Δf H) phases of this system. An analysis of d-metals influence on formation energy of Ga-d-Me melts was made using the values of Δf H for intermediate phases of these systems. The article was translated by the authors.  相似文献   

16.
The stoichiometries, kinetics and mechanism of the reduction of tetraoxoiodate(VII) ion, IO4 to the corresponding trioxoiodate(V) ion, IO3 by n-(2-hydroxylethyl)ethylenediaminetriacetatocobaltate(II) ion, [CoHEDTAOH2] have been studied in aqueous media at 28 °C, I = 0.50 mol dm−3 (NaClO4) and [H+] = 7.0 × 10−3 mol dm−3. The reaction is first order in [Oxidant] and [Reductant], and the rate is inversely dependent on H+ concentration in the range 5.00 × 10−3 ≤ H+≤ 20.00 × 10−3 mol dm−3 studied. A plot of acid rate constant versus [H+]−1 was linear with intercept. The rate law for the reaction is:
- \frac[ \textCoHEDTAOH2 - ]\textdt = ( a + b[ \textH + ] - 1 )[ \textCoHEDTAOH2 - ][ \textIO4 - ] - {\frac{{\left[ {{\text{CoHEDTAOH}}_{2}^{ - } } \right]}}{{{\text{d}}t}}} = \left( {a + b\left[ {{\text{H}}^{ + } } \right]^{ - 1} } \right)\left[ {{\text{CoHEDTAOH}}_{2}^{ - } } \right]\left[ {{\text{IO}}_{4}^{ - } } \right]  相似文献   

17.
The ionic conductivity of Na,Zr and Na,Sn silicates of the lovozerite family (Na8 − x H x ZrSi6O18 structural type, space group R $ \bar 3 $ \bar 3 m) was studied in the temperature range of 293–800 K using the impedance spectroscopy method (5−5 × 105 Hz). The compositions of the studied compounds were obtained using the method of hydrothermal synthesis in the MO2-SiO2-NaOH-H2O and MO2-SiO2-CaO-NaOH-H2O (M = Zr, Sn) systems at 573–823 K. The samples for electrophysical studies were prepared according to the ceramic technology. It was found that isovalent cation substitutions of Sn4+ → Zr4+ in Na8M4+Si6O18 and Na6CaM4+Si6O18 and H+ → Na+ in Na8 − x H x ZrSi6O18 result in an increase in the ionic conductivity by 2–3 orders of magnitude, without affecting the ionic transport activation energy (0.6–0.7 eV). The best electrolytic characteristics are typical for the Na5H3ZrSi6O18 compound, for which the ionic conductivity value is 5 × 10−4 S/cm at 573 K.  相似文献   

18.
The sodium hydrogen oxovanadate [Na2(H2O)8] 2H2[V10O28] · 4H2O was synthesized and studied by TGA, X-ray diffraction, and NMR and IR spectroscopy. The crystals are triclinic, space group P $ \bar 1 $ \bar 1 , a = 8.545(7) ?, b = 10.827(2) ?, c = 11.627(2) ?, α = 105.48(3)°, β = 99.38(3)°, γ = 101.29(3)°, V = 989.9(3) ?3, ρ(calcd) = 2,381 g/cm3, Z= 1.  相似文献   

19.
Zusammenfassung In der vorliegenden Arbeit wird das Sedimentations-gleichgewicht des Systems Na6[H2W12O40]-NaClO4-H2O bei 25 °C untersucht. Die scheinbare Ladungszahl des Natriummetawolframats wird als Funktion der NaClO4-MolalitätmNaClO4 ermittelt. Aus dieser Abhängigkeit kann für die summarische Dissoziation des NatriummetawolframatsNa2[H2W12O4o] 6Na+ + [H2W12O40]6– die DissoziationskonstanteK in Gegenwart von NaClO4 als Fremdelektrolyt zuK=0,08 mol6·kg–6 bestimmt werden.Der Kreuzdifferentialquotient ( PXz , ist der Trace-Aktivitätskoeffizient des Natriummetawolframats) wird ebenfalls als Funktion der NaClO4-Molalität ermittelt.Die Konzentrationsabhängigkeit des Trace-Aktivitätskoeffizienten zeigt einen Verlauf, wie er unseres Wissens nach noch nicht von anderen Autoren beobachtet wurde.
Summary An ultracentrifugal study of the sedimentation equilibrium of the system Na2[H2W12O40]-NaClO4-H2Oat 25 °C is presented. The dependency of the apparent charge of the sodium metatungstate on NaClO4-molality is investigated. From these data the dissociation constant for the dissociation of the sodium metatungstate in the presence of NaClO4 according to the equationNa[H2W12O40] 6Na+ + [H2W12O40]6– is calculated toK=0.08 mol6·kg–6. The cross differential quotient ( PXz is the trace activity coefficient of the sodium metatungstate) shows a behavior in its dependence on NaClO4-molality which to our knowledge has not yet been reported by other authors.


Mit 5 Abbildungen und 1 Tabelle  相似文献   

20.
The potentiometric response of 100-nm spherical K1.3Mn8O16 particles versus K+ ions has been studied in aqueous media using a polymeric technology. The stoichiometry of this material evolves in potassium nitrate solution towards K1.08Mn8O16. A stable and reversible response has been obtained with a sensitivity of 47 mV dec?1 in the range from 8?×?10?5 to 1 mol·L?1, and a rather good selectivity towards Li+, Na+, Mg2+ and Ca2+ $\left( {{\text{log K}}_{{{{\text{K}}^{\text{ + }} } \mathord{\left/ {\vphantom {{{\text{K}}^{\text{ + }} } {{\text{X}}^{n + } }}} \right. \kern-0em} {{\text{X}}^{n + } }}} \approx - {\text{3}}} \right)$ . We assume that this potentiometric response is the result of the ability of K1.08Mn8O16 to specifically adsorb K+ ions.  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