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 56 毫秒
1.
The chemical diffusion coefficient of oxygen vacancies and oxygen ion conductivity in lanthanum cobaltite LaCoO3 were determined by the polarization method as functions of oxygen partial pressure \(p_{O_2 } \) (atm) and temperature T(K) over the ranges ?4 ≤ log \(p_{O_2 } \) ≤ 0 and 1173 K ≤ T ≤ 1323 K. The mobilities (cm2/(V s)) of oxygen vacancies calculated over the temperature range studied satisfy the inequalities 1.8 × 10?5\(v_{v_0 } \) ≤ 3.4 × 10?5. The transfer numbers of oxygen vacancies were calculated. These numbers change depending on oxygen partial pressure over the range 5 × 10?7t 0 ≤ 1 × 10?5. The activation energy of self-diffusion of oxygen vacancies was found to be E a= 104 ± 10 kJ/mol (1.1 ± 0.1 eV).  相似文献   

2.
The direct electrochemistry of redox enzymes (or proteins) has received more and more attention[1—9]. These studies developed an electrochemical basis for the investigation of enzyme structure, mechanisms of redox transformations of enzyme molecules and metabolic processes involving redox transformations. From these studies, one can also find potential appli-cations of enzymes in biotechnology. For example, if an enzyme immobilized on electrode surface is ca-pable of the direct electron tra…  相似文献   

3.
Basic equations of classical polarography are transposed and tested withmanganese and europium at tracer scale. Using the previously presented automatic system, standard dispersion for experimental points is about 3% and residual activity before the wave 3±2% compared with diffusion activity Ad. Results with manganese in 10?3M ≤[LiCl]≤10?1M and 2≤pH≤4 solutions, give an experimental yield about 100%, E1/2=?1.47 V/SCE and α?0.85. The last values agree with results achieved at a weighable scale. For potentials <?2.00 V/SCE a screening effect with supporting electrolyte ions is observed. The screening effect influence on the diffusion coefficient is taken into consideration and results for europium (R?100%, E1/2??1.95 V/ECS, α=1 at pH 2.7) are in good agreement with the literature. Radiopolarograms for Mn and Eu and mixtures of both are similar. Thus the range of polarographic analysis is enlarged.  相似文献   

4.
A study regarding the electrodeposition of nickel oxide particles on the activated multi-walled carbon nanotubes from 2 M NaOH solution containing Ni(NO3)2 and EDTA was carried out. The electrodeposition process was carried out using an optimized double-pulse sequence of potentials: E 1 = ?0.2 V vs. SCE (t 1 = 0.3 s) and E 2 = 0.7 V vs. SCE (t 2 = 0.03 s). Spectroscopic XPS investigations and SEM analysis were used in order to characterize the surface and morphology of the studied modified electrode. Cyclic voltammetry and chronoamperometry were used in order to evaluate the electrochemical/amperometric performance of the GC/MWCNT-Ni electrode toward the oxidation of some aliphatic alcohols in strong alkaline medium.  相似文献   

5.
Differences of potentials of anodic and cathodic peaks (ΔE p) are determined in cyclic voltammograms of dihydroxybenzene/p-benzoquinone redox systems at an electrode made of a graphite-epoxy composite in a wide pH range. The data obtained (ΔE p = 29 ± 1 mV) are close to the thermodynamic values for two-electron reversible reactions. This indicates that the electrode mechanically renewed by cutting a 0.5-μm surface layer directly in a test solution exhibits a high activity in such electrochemical reactions. The potentials of anodic and cathodic peaks are proportional to the pH of the supporting electrolyte solution in the range from 1.0 to 9.0. A change of 58 ± 1 mV in E p per unit pH for all isomers shows that the first stage of the oxidation of each dihydroxybenzene isomer involves one electron and is accompanied by the detachment of one hydrogen ion, that is, an intermediate oxidation product, semiquinone, is formed. Despite the closeness of the potentials of hydroquinone and pyrocatechol peaks (ΔE = 100 mV), a scheme is proposed for the selective voltammetric determination of dihydroxybenzene isomers in a 0.1 M HCl solution in hydroquinone-pyrocatechol, pyrocatechol-resorcinol, and hydroquinone-resorcinol binary mixtures. The concentrations of hydroquinone and pyrocatechol are found from cathodic peaks and that of resorcinol, from the anodic peak. The results are well reproducible and contain no systematic error.  相似文献   

6.
A nickel(II) complex, [Ni(taetacn)](ClO4)2 ? H2O, where taetacn = 1,4,7-tris(2-aminoethyl)-1,4,7-triazacyclononane was synthesized. The crystal structure was determined by the single-crystal X-ray diffraction method at 293 K. The complex crystallizes in the orthorhombic space group Pna21 with a = 16.004(2) Å, b = 10.186(1) Å, c = 13.937(2) Å, V = 2271.9(5) Å3, Dx = 1.56 g cm?3, Dm = 1.59 g cm?3 (floatation method), and Z = 4. The R1 [I > 2σ(I)] and wR2 (all data) values are 0.0636 and 0.1672, respectively, for all 4845 independent reflections. The compound is composed of octahedral nickel(II) cation with three 2-aminoethyl pendant groups of taetacn, tetrahedral ClO 4 ? anion, and a water molecule of crystallization. Electronic spectra are consistent with the octahedral geometry. Temperature dependence of the magnetic susceptibility (4.5–300 K) can be interpreted considering the zero-field splitting of the nickel(II) ion (g = 2.14, D = 3.72 cm?1, and = 300 × 10?6 cm3 mol?1). Cyclic voltammetry in DMF showed quasi-reversible and irreversible oxidation waves (Epa = 0.54 V, Epc = 0.45 V; Epa = 1.16 V, Epc = 0.71 V vs. Ag/Ag+).  相似文献   

7.
The effects of axial ligands on electron‐transfer and proton‐coupled electron‐transfer reactions of mononuclear nonheme oxoiron(IV) complexes were investigated by using [FeIV(O)(tmc)(X)]n+ ( 1 ‐X) with various axial ligands, in which tmc is 1,4,8,11‐tetramethyl‐1,4,8,11‐tetraazacyclotetradecane and X is CH3CN ( 1 ‐NCCH3), CF3COO? ( 1 ‐OOCCF3), or N3? ( 1 ‐N3), and ferrocene derivatives as electron donors. As the binding strength of the axial ligands increases, the one‐electron reduction potentials of 1 ‐X (Ered, V vs. saturated calomel electrode (SCE)) are more negatively shifted by the binding of the more electron‐donating axial ligands in the order of 1 ‐NCCH3 (0.39) > 1 ‐OOCCF3 (0.13) > 1 ‐N3 (?0.05 V). Rate constants of electron transfer from ferrocene derivatives to 1 ‐X were analyzed in light of the Marcus theory of electron transfer to determine reorganization energies (λ) of electron transfer. The λ values decrease in the order of 1 ‐NCCH3 (2.37) > 1 ‐OOCCF3 (2.12) > 1 ‐N3 (1.97 eV). Thus, the electron‐transfer reduction becomes less favorable thermodynamically but more favorable kinetically with increasing donor ability of the axial ligands. The net effect of the axial ligands is the deceleration of the electron‐transfer rate in the order of 1 ‐NCCH3 > 1 ‐OOCCF3 > 1 ‐N3. In sharp contrast to this, the rates of the proton‐coupled electron‐transfer reactions of 1 ‐X are markedly accelerated in the presence of an acid in the opposite order: 1 ‐NCCH3 < 1 ‐OOCCF3 < 1 ‐N3. Such contrasting effects of the axial ligands on the electron‐transfer and proton‐coupled electron‐transfer reactions of nonheme oxoiron(IV) complexes are discussed in light of the counterintuitive reactivity patterns observed in the oxo transfer and hydrogen‐atom abstraction reactions by nonheme oxoiron(IV) complexes (Sastri et al. Proc. Natl. Acad. Sci. U.S.A. 2007 , 104, 19 181–19 186).  相似文献   

8.
The electrochemical reactions of ruthenium(II) bis(triethylenediamine)tetra-tret-butyl-phthalocyaninate in dimethylformamide are studied. Two reversible redox reactions on the platinum amalgam electrode are revealed at the potentials of ?0.73 and ?1.16 V (Ag/AgCl). Similarly to several other phthalocyanines, these redox reactions correspond to the successive transfer of two electrons to phthalocyanine ring. A new phenomenon, which has not been reported in the literature for phthalocyanines, namely, the cathodic polymerization, is discovered. Thus formed polymer is redox-active, and only one cathodic reaction at the potentials from ?0.78 to ?0.84 V (a shift in the cathodic direction takes place as the film thickness increases) is observed in the polymer. In addition, the polymer exhibits also considerable electron conductivity that enables one to perform various electrochemical reactions in a wide potential range on the electrode modified with the polymer.  相似文献   

9.
Steady-state polarization curves are compared in solutions of 0.5 M H2SO4 + O2 (saturated), 0.5 M H2SO4 + (0.005–0.1) M CH3OH, and 0.5 M H2SO4 + (0.005–0.1) M CH3OH + O2 (saturated) on a Pt/Pt electrode. A considerable difference is found between the currents in mixed solutions and those expected based on the principle of additivity of currents in CH3OH and O2 individual solutions. The surface coverages with the CH3OH and O2 adsorption products are determined in the potential range of 0.2–0.9 V (RHE). Open-circuit potentials are measured in mixed solutions. The obtained results suggest that the direct heterogeneous interaction between methanol and oxygen occurs alongside with faradaic reactions. This is assumed to lead to a decrease in methanol electrooxidation currents at E ≥ 0.8 V and their increase at E ≤ 0.65 V.  相似文献   

10.
Data on the thermogravimetry, spectroscopy, and electrical charge transfer as functions of T, aH2O, and aO2 for niobates and tantalates of alkali-earth metals with structure disordering of the oxygen sublattice, which can show high-temperature proton conduction, are summarized. It is shown that in the solid solution series with decreasing x (that is, with the increasing of the oxygen vacancies concentration) the proton conductivity increase, which is caused by the increasing of both the concentration of proton defects formed in the structure (in compliance with the formula Sr6 ? 2x M 2 + 2x +5 O10(OH)2?6x and their mobility. The proton transfer dominates for the compositions with x < 0.15 at temperatures below 550°C. In the solid solutions (Ba1?y Ca y )6Nb2O11 (0.23 ≤ y ≤ 0.47) characterized by equal concentration of oxygen vacancies, with the increasing of barium content (correspondingly, with the increasing of the lattice parameter) the oxygen-ion conductivity (at aH2O = 3 × 10?5) grows monotonically, which is caused by the decreasing of the oxygen atom migration energy and increasing of their mobility. In this series, the proton conductivity (at aH2O = 2 × 10?2) increased. It was shown, by using IR-spectroscopy and the 1H NMR method, that the protons exist in the complex oxide structure mainly as energy-wise nonequivalent OH? groups: isolated, closely set, and paired, whose quantitative ratios are determined by the coordination preference of the B-sublattice elements.  相似文献   

11.
The distribution of NO molecules desorbed from a Pt(111) surface due to valence electron excitation over rotational energy levels N(J) is analyzed using a simple impulse-induced model. A linear dependence is found between lnN(J) and (Er)1/2, where Er is the rotational energy of the desorbed molecules. The lifetime of the excited state and the critical time of residence in the excited state estimated using this dependence are found to be close to one another (~10?15 s). The frequency and amplitude of the tilting vibrations of the adsorbed molecules in the excited state are estimated.  相似文献   

12.
This work describes the development of a simple, fast and low-cost method for determining prazosin (PRA) in pharmaceutical samples by flow injection analysis with multiple-pulse amperometric (FIA-MPA) detection using a boron-doped diamond film electrode. Electrochemical detection of PRA was optimized in phosphate buffer pH 4.0 by cyclic voltammetry, in which PRA presented two oxidation processes around at 0.97 and 1.40 V versus Ag/AgCl (3.0 mol L?1 KCl). In these conditions, PRA also showed one reduction process at ?0.75 V that is dependent on the oxidation processes. Thus, the determination of PRA by FIA-MPA detection consisted on the application of a two-potential waveform, E 1 (generator potential)?=?1.6 V/400 ms and E 2 (collector potential)?=??1.0 V/30 ms, with sample loop of 150 μL and flow rate of 3.0 mL min?1. The method showed good repeatability (RSD?<?3.0 %) and high analytical frequency (70 injections per h). The working linear range was obtained from 2 to 200 μmol L?1 with a limit of detection of 0.5 μmol L?1. The recovery tests in all samples were approximately 100 %, and the results were compared with chromatographic methods.  相似文献   

13.
The field dependences of photocurrent, the two-beam coupling gain coefficient, and the grating formation time constant in polymer composites made from polyvinylcarbazole (PVK) and single-wall carbon nanotubes (SWNTs) were measured under the conditions of one-photon SWNT excitation with continuous laser radiation at a wavelength of 1550 nm. Carbon nanotubes are responsible for optical electronic absorption up to ~2000 nm in this composite. The dependence of the quantum efficiency of generation of mobile charge carriers on the electric field E 0 as determined from the photocurrent coincides with the curves calculated via the Onsager equation expanded to the (E 0)4 term, at a quantum yield of thermalized electron-hole pairs of η0 = 0.07 and a charge separation distance in the pair of r 0 = 9.8 Å. An analysis of the photorefractive characteristics showed that the admixture of fullerene C60 in an amount of 3 wt % to the PVK composite with 0.26 wt % SWNT leads to a twofold increase in the beam-coupling gain coefficient. In the PVK-matrix composite containing 0.26 wt % SWNT and 3 wt % C60, the beam-coupling gain coefficient Γ of a 1550-nm laser beam and the net gain Γ-α are 32 and ~27 cm?1, respectively, at a constant field of E 0 = 140 V/μm.  相似文献   

14.
The electrochemical study of electron transport between Au electrodes and the redox molecule Os[(bpy)2(PyCH2 NH2CO-]ClO4 tethered to molecular linkers of different length (1.3 to 2.9 nm) to Au surfaces has shown an exponential decay of the rate constant k ET 0 with a slope β = 0.53 consistent with through bond tunneling to the redox center. Electrochemical gating of single osmium molecules in an asymmetric tunneling nano-gap between a Au(111) substrate electrode modified with the redox molecules and a Pt-Ir tip of a scanning tunneling microscope was achieved by independent control of the reference electrode potential in the electrolyte, E ref ? E s, and the tip-substrate bias potential, E bias. Enhanced tunneling current at the osmium complex redox potential was observed as compared to the off resonance set point tunneling current with a linear dependence of the overpotential at maximum current vs. the E bias. This corresponds to a sequential two-step electron transfer with partial vibration relaxation from the substrate Au(111) to the redox molecule in the nano-gap and from this redox state to the Pt-Ir tip according to the model of Kuznetsov and Ulstrup (J Phys Chem A 104: 11531, 2000). Comparison of short and long linkers of the osmium complex has shown the same two-step ET (electron transfer) behavior due to the long time scale in the complete reduction-oxidation cycle in the electrochemical tunneling spectroscopy (EC-STS) experiment as compared to the time constants for electron transfer for all linker distances, k ET 0.  相似文献   

15.
Poly(N-vinylcarbazole) layers containing tetra-5-crown-5-gallium phthalocyaninate (R4Pc)Ga(OH) are shown to possess photoelectric and photorefractive sensitivity at a wavelength of 1064 nm. This effect is associated with the formation of supramolecular ensembles of (R4Pc)Ga(OH) molecules with electronic optical absorption in the near-IR range and nonlinear optical properties. For the composite containing 5 wt % (R4Pc)Ga(OH) supramolecular ensembles, the dependence of the quantum efficiency of mobile-charge photogeneration on electric field E 0 is well fit by the Onsager equation expanded to E 0 3 at a quantum yield of electron-hole pairs of φ0 = 0.9 s with an initial separation radius of r 0 = 9.8 Å susceptibility χ(3) equal to 1.85 × 10?10 esu is measured via the well-known method of electric-field-induced second-harmonic generation. Two-beam-coupling gain coefficient Γ is found to be 80 cm?1 at E 0 = 120 V/μm.  相似文献   

16.
The effect of γ-radiation on electrical conductivity of Cd x Hg1 ? x Te (0.25 ≤ x ≤ 0.95) single crystals in weak and strong electric fields has been investigated. It has been shown that at relatively low fields in which charge carriers are still warm, the dependence (Δσ/σ0) ~ E 2 is observed because of the dominance of scattering on acoustic lattice vibrations with the increasing electric field strength; with a further increase in the field strength, the carriers become hot and the dependence of σ upon E becomes linear. The effect of irradiation by Γ-rays on the dependence of σ upon E in these samples is explained by a significant concentration of the intrinsic impurity centers in the crystals and their redistribution with the increasing radiation dose.  相似文献   

17.
The potentials of six approximation functions described in the literature and used to simulate bremsstrahlung X-radiation in electron probe microanalysis are analyzed within a uniform approach. An algorithm based on the use of wave spectra is proposed for determining the parameters of the functions. It is found that the calculated values of some parameters of approximation functions considerably differ from those presented by the authors of the functions. It is found that the quality of approximating bremsstrahlung X-radiation is strongly affected by the method of calculating the average atomic number of the sample and the corrections for matrix effects. In the wavelength range 0.1 ≤ λ ≤ 1.2 nm for atomic numbers 10 ≤ Z ≤ 83 and accelerating voltages 15 ≤ E 0 ≤ 25 kV, the best function ensures background estimation with a relative standard deviation of 3–6%, which is comparable with the relative standard deviation of the direct background measurement in routine electron probe microanalysis. The use of the calculation method of the background account at λ > 1.2 nm is complicated because of the significant uncertainty of the absorption factor in the long-wavelength region for the majority of elements.  相似文献   

18.
This paper reported a simple method for sulfanilamide determination by redox process electroanalysis of oxidation products (SFDox) formed in situ on glassy carbon electrode. The CV experiments showed a reversible process after applied E acc = + 1.06 V and t acc = 1 s, in 0.1 mol L?1 BRBS (pH = 2.0) at 50 mV s?1. Different voltammetric scan rates (from 10 to 450 mV s?1) suggested that the redox peaks of SFDox on the glassy carbon electrode (GCE) is an adsorption-controlled process. Square-wave voltammetry (SWV) method optimized conditions showed a linear response to SFD from 3.00 to 250.0 μmol L?1 (R = 0.998) with a limit of detection of 0.638 μmol L?1 and limit of quantification of 2.0 μmol L?1. The developed the SWV method was successfully used in the determination of SFD pharmaceutical formulation and human serum. The SFD quantification results in pharmaceutical obtained by SWV-GCE were comparable to those found by official analytical protocols.  相似文献   

19.
A procedure is developed for preparing conducting films by their casting from polymer solutions containing polyaniline in the form of a protonated emeraldin base and polyimides in two cosolvents, N-methylpyrrolidone or m-cresol. Self-supporting films cast from composites based on polyimides and camphorsulfonic acid-protonated polyaniline combine a conductivity of 10?1?10?2S/cm with good mechanical properties: elastic modulus E = 2.0?2.4 GPa, breaking strength σb = 55?60 MPa, and elongation at break ?b = 8?10%. It has been shown that, when m-cresol and N-methylpyrrolidone are used as cosolvents, the maximum film conductivity is achieved at polyaniline amounts in the composites of 20 and 3%, respectively. In the latter case, films with good strength parameters are formed.  相似文献   

20.
2-Mercapto-5-methyl-1,3,4-thiadiazol (MMTD) and 2,5-dimercapto-1,3,4-thiadiazol (DMTD) were studied by differential pulse cathodic stripping voltammetry (DPCSV). The influence of buffer, pH, accumulation potential (Eacc), and accumulation time (tacc) was investigated. It was stated that the concentration of the buffer affects the height of DPCSV peaks. The best analytical signals were recorded in acetate buffer at pH 4.3 and a buffer concentration of 0.01 mol/L for MMTD and 0.02 mol/L for DMTD, Eacc = 0.2 V, and tacc = 120 s for MMTD and 180 s for DMTD. A linear dependence was found from 1 to 8 × 10?8 mol/L for MMTD and from 1 × 10?8 to 1 × 10?7 mol/L for DMTD. The influence of cations [Cu(II), Co(II)] was also considered.  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