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 125 毫秒
1.
The α-tocopheroxyl radical was generated voltammetrically by one-electron oxidation of the α-tocopherol anion (r1/2=−0.73 V versus Ag|Ag+) that was prepared by reacting α-tocopherol with Et4NOH in acetonitrile (with Bu4NPF6 as the supporting electrolyte). Cyclic voltammograms recorded at variable scan rates (0.05–10 V s−1), temperatures (−20 to 20°C) and concentrations (0.5–10 mM) were modelled using digital simulation techniques to determine the rate of bimolecular self-reaction of α-tocopheroxyl radicals. The k values were calculated to be 3×103 l mol−1 s−1 at 20°C, 2×103 l mol−1 s−1 at 0°C and 1.2×103 l mol−1 s−1 at −20°C. In situ electrochemical-EPR experiments performed at a channel electrode confirmed the existence of the α-tocopheroxyl radical.  相似文献   

2.
Four short- and long-alkyl-multiamine ligands L1–L4 have been synthesized and characterized. The catalytic efficiency of complex CuL1 and functional metallomicelles CuL2–CuL4 were comparatively investigated for the hydrolysis of bis(p-nitrophenyl) phosphate (BNPP) in buffered solution at 30 °C. The ternary kinetic model for metallomicellar catalysis was suggested to analyze the experimental data. The kinetic and thermodynamic parameters kN, KT and pKa were obtained. The results indicated that the complexes with 1:1 ratio of ligands L2–L4 to copper(II) ion were the kinetic active catalysts, and the deprotonized Cu(II) complex formed by activated water molecule was the real active species for BNPP catalytic hydrolysis. The real rate constant of the reaction catalyzed by CuL1–CuL4 was 4.00 × 10−6, 7.44 × 10−5, 1.42 × 10−4 and 4.10 × 10−4 s−1, respectively. The effects of ligand and microenvironment on the hydrolytic reaction of BNPP have been discussed in detail.  相似文献   

3.
This report describes the use of a piezoelectric quartz crystal (PQC) sensor to investigate the nonspecific adsorption of fibrinogen (FN) and sodium dodecyl sulfate (SDS) onto a self-assembled monolayer (SAM) of alkanethiols on gold. The change in adsorption mass was monitored in situ by the PQC sensor. A kinetics model was proposed to describe the adsorption of the FN and SDS on the hydrophobic SAM surface. The adsorption kinetics parameters were determined from the responses of the PQC. The adsorption and desorption rate constants of the FN on the SAM surface were estimated to be (6.18 ± 0.53) × 103 M−1 s−1 and (6.74 ± 0.72) × 10−3 s−1, respectively. The rate constants for the adsorption and desorption of SDS on the SAM are (24.3 ± 1.4 M−1 s−1) and (1.52 ± 0.11) × 10−2 s−1, respectively. The adsorption of SDS on the SAM was reversible. The fractional coverage of the FN on the SAM surface was estimated from kinetics analyses to be 42–86% for the FN concentration range 25–400 μg/ml. Over 80% of the FN is irreversibly adsorbed on the SAM surface with respect to dilution of the bulk phase. The fraction of FN reversibly adsorbed increases with the bulk concentration of FN.  相似文献   

4.
Degradation of polyoxyethylene chain of non-ionic surfactant (TritonX-100) by chromium(VI) has been studied spectrophotometrically under different experimental conditions. The reaction rate bears a first-order dependence on the [Cr(VI)] under pseudo-first-order conditions, [TritonX-100]  [Cr(VI)] in presence of 1.16 mol dm−3 perchloric acid. The observed rate constant (kobs) was 3.3 × 10−4 to 3.5 × 10−4 s−1 and the half-life (t1/2) was 33–35 min for chromium(VI). The effects of total [TritonX-100] and [H+] on the reaction rate were determined. Reducing nature of non-ionic TritonX-100 surfactant is found to be due to the presence of –OH group in the polyoxyethylene chain. It was observed that monomeric and non-ionic micelles of TritonX-100 were oxidized by chromium(VI). When [TritonX-100] was less than its critical micelle concentration (cmc) the kobs values increased from 0.76 × 10−4 to 1.5 × 10−4 s−1. As the [TritonX-100] was greater than the cmc, the kobs values increases from 2.1 × 10−4 to 8.2 × 10−4 s−1 in presence of constant [HClO4] (1.16 mol dm−3) at 40 °C. A comparison was made of the oxidative degradation rates of TritonX-100 with different metal ion oxidants. The order of the effectiveness of different oxidants was as follows: permanganate > diperiodatoargentate(III) > chromium(VI) > cerium(IV).  相似文献   

5.
The coordination of nitric oxide (NO) to cobalt(II) phthalocyanine (CoPc) in dimethyl sulphoxide (DMSO) has been studied. CoPc coordinates with NO in a 1:1 ratio, forming a CoPc(NO) species. The IR band observed at 1680 cm−1 is assigned to the coordinated NO. In the presence of excess NO, pseudo first order kinetics were followed. The observed rate constant, kf, was determined to be 15.0±0.3 dm−3 mol−1 s−1 and the equilibrium constant was K=5.4±0.4×104dm3 mol−1. Solution or adsorbed CoPc catalyses the reduction of NO. The products of reduction include NH3 and NH2OH.  相似文献   

6.
Raman spectroscopy was used as a method in the kinetic investigation of the thermal decomposition of solid (NH4)2CrO4. Time-dependent measurements of the intensity of the totally symmetric stretching Cr---O mode of (NH4)2CrO4 have been made between 343 and 363 K. A short initial acceleratory period is observed at lower temperatures and the decomposition reaction decelerates after the maximum decomposition rate has been reached at all temperatures. These results can be interpreted in terms of the Avrami-Erofe'ev law 1 − (χr)1/2 = kt, where χr is the fraction of reactant at time t. At 358 K, k is equal to 1.76 ± 0.01 × 10−3 sec−1 for microcrystals and for powdered samples. Activation energies of 97 ± 10 and 49 ± 0.9 kJ mole−1 have been calculated for microcrystalline and powdered samples, respectively.  相似文献   

7.
The diffusion of strontium and zirconium in single crystal BaTiO3 was investigated in air at temperatures between 1000 °C and 1250 °C. Thin films of SrTiO3, deposited by spin coating a precursor solution and thin films of zirconium, deposited onto the sample surfaces by sputtering, were used as diffusion sources. The diffusion profiles were measured by SIMS depth profiling on a time-of-flight secondary ion mass spectrometer (ToF-SIMS). The diffusion coefficients of strontium and zirconium were given by DSr = 3.6 × 102.0±4.4 exp[−(543 ± 117) kJ mol−1/(RT)] cm2 s−1 and DZr = 1.1 × 101.0±2.1 exp[−(489 ± 56) kJ mol−1/(RT)] cm2 s−1. The results are discussed in terms of different diffusion mechanisms in the perovskite structure of BaTiO3.  相似文献   

8.
The kinetics of the reaction of the CH3CHBr, CHBr2 or CDBr2 radicals, R, with HBr have been investigated in a temperature-controlled tubular reactor coupled to a photoionization mass spectrometer. The CH3CHBr (or CHBr2 or CDBr2) radical was produced homogeneously in the reactor by a pulsed 248 nm exciplex laser photolysis of CH3CHBr2 (or CHBr3 or CDBr3). The decay of R was monitored as a function of HBr concentration under pseudo-first-order conditions to determine the rate constants as a function of temperature. The reactions were studied separately from 253 to 344 K (CH3CHBr + HBr) and from 288 to 477 K (CHBr2 + HBr) and in these temperature ranges the rate constants determined were fitted to an Arrhenius expression (error limits stated are 1σ + Student’s t values, units in cm3 molecule−1 s−1, no error limits for the third reaction): k(CH3CHBr + HBr) = (1.7 ± 1.2) × 10−13 exp[+ (5.1 ± 1.9) kJ mol−1/RT], k(CHBr2 + HBr) = (2.5 ± 1.2) × 10−13 exp[−(4.04 ± 1.14) kJ mol−1/RT] and k(CDBr2 + HBr) = 1.6 × 10−13 exp(−2.1 kJ mol−1/RT). The energy barriers of the reverse reactions were taken from the literature. The enthalpy of formation values of the CH3CHBr and CHBr2 radicals and an experimental entropy value at 298 K for the CH3CHBr radical were obtained using a second-law method. The result for the entropy value for the CH3CHBr radical is 305 ± 9 J K−1 mol−1. The results for the enthalpy of formation values at 298 K are (in kJ mol−1): 133.4 ± 3.4 (CH3CHBr) and 199.1 ± 2.7 (CHBr2), and for α-C–H bond dissociation energies of analogous compounds are (in kJ mol−1): 415.0 ± 2.7 (CH3CH2Br) and 412.6 ± 2.7 (CH2Br2), respectively.  相似文献   

9.
Recombination rate coefficients of protonated and deuterated ions KrH+, KrD+, XeH+ and XeD+ were measured using Flowing Afterglow with Langmuir Probe (FALP). Helium at 1600 Pa and at temperature 250 K was used as a buffer gas in the experiments. Kr, Xe, H2 and D2 were introduced to a flow tube to form the desired ions. Because of small differences in proton affinities of Kr, D2 and H2 mixtures of ions, KrD+/D3+ and KrH+/H3+ are formed in the afterglow plasma, influencing the plasma decay. To obtain a recombination rate coefficient for a particular ion, the dependencies on partial pressures of gases used in the ion formation were measured. The obtained rate coefficients, αKrD+(250 K) = (0.9 ± 0.3) × 10−8 cm3 s−1 and αXeD+(250 K) = (8 ± 2) × 10−8 cm3 s−1 are compared with αKrH+(250 K) = (2.0 ± 0.6) × 10−8 cm3 s−1 and αXeH+(250 K) = (8 ± 2) × 10−8 cm3 s−1.  相似文献   

10.
Quercetin can effectively accumulate at multi-walled carbon nanotubes-paraffin oil paste electrodes (CNTPE) and cause a sensitive anodic peak at around 0.32 V (vs. SCE) in a 0.10 M phosphate buffer solution (pH = 4.0). Under optimized conditions, the anodic peak current is linear to quercetin concentration in the ranges of 2.0 × 10− 9−1.0 × 10− 7 M and 1.0 × 10− 7−2.0 × 10− 5 M, and the regression equations are ip (μA) = 0.0017 + 0.928c (μM, r = 0.999) and ip (μA) = 0.183 + 0.0731c (μM, r = 0.995), respectively. This paste electrode can be regenerated by repetitively cycling in a blank solution for about 2 min. A 1.0 × 10− 6 M quercetin solution is measured for 10 times using the same electrode regenerated after every determination, and the relative standard deviation of the peak current is 1.7%. The method has been applied to the determination of quercetin in hydrolysate product of rutin and the recovery is 99.2–102.6%. In comparison with graphite paste electrode, carbon nanotubes-nujol paste electrode and carbon nanotubes casting film modified glassy carbon electrode, the CNTPE gives higher ratio of signal to background current and better defined voltammetric peak.  相似文献   

11.
A simple, sensitive, selective and rapid kinetic catalytic method has been developed for the determination of Hg(II) ions at micro-level. This method is based on the catalytic effect of Hg(II) ion on the rate of substitution of cyanide in hexacyanoruthenate(II) with nitroso-R-salt (NRS) in aqueous medium and provides good accuracy and precision. The concentration of Hg(II) catalyst varied from 4.0 to 10.0 × 10−6 M and the progress of reaction was followed spectrophotometrically at 525 nm (λmax of purple-red complex [Ru(CN)5NRS]3−,  = 3.1 × 103 M−1 s−1) under the optimized reaction conditions; 8.75 × 10−5 M [Ru(CN)64−], 3.50 × 10−4 M [nitroso-R-salt], pH 7.00 ± 0.02, ionic strength, I = 0.1 M (KCl), temp 45.0 ± 0.1 °C. The linear calibration curves, i.e. calibration equations between the absorbance at fixed times (t = 15, 20 and 25 min) versus concentration of Hg(II) ions were established under the optimized experimental conditions. The detection limit was found to be 1.0 × 10−7 M of Hg(II). The effect of various foreign ions on the proposed method has also been studied and discussed. The method has been applied to the determination of mercury(II) in aqueous solutions.  相似文献   

12.
The hydrogen abstraction reaction of 1,1,1,2-tetrafluoroethane (HFC-134a) by chlorine radical is investigated by theoretical calculations. Equilibrium geometries and harmonic vibrational frequencies of the reactants, transition state, and products are calculated using high-level ab initio methods. Rate constants of forward and backward reactions for the temperatures from 200 to 1000 K are calculated using classical transition state theory with Eckart tunneling correction, fitted in the expressions kf (T) = 1.19 × 10−23T3.93exp (−1110/T), and kb (T) = 8.86 × 10−24T3.32exp (−959/T) cm3 molecule−1 s−1 for forward and backward reactions, respectively, and are in reasonable agreement with the available experimental values.  相似文献   

13.
Adsorption (at a low temperature) of nitrogen on the protonic zeolite H-Y results in hydrogen bonding of the adsorbed N2 molecules with the zeolite Si(OH)Al Brønsted-acid groups. This hydrogen-bonding interaction leads to activation, in the infrared, of the fundamental N–N stretching mode, which appears at 2334 cm−1. From infrared spectra taken over a temperature range, the standard enthalpy of formation of the OH···N2 complex was found to be ΔH0 = −15.7(±1) kJ mol−1. Similarly, variable-temperature infrared spectroscopy was used to determine the standard enthalpy change involved in formation of H-bonded CO complexes for CO adsorbed on the zeolites H-ZSM-5 and H-FER; the corresponding values of ΔH0 were found to be −29.4(±1) and −28.4(±1) kJ mol−1, respectively. The whole set of results was analysed in the context of other relevant data available in the literature.  相似文献   

14.
We compared the binding affinity of 6-propyl-2-thiouracil (PTU) with native and destabilized human serum albumin (HSA) as a model to assess the binding ability of albumin in patients suffering from chronic liver or renal diseases. Urea (U) and guanidine hydrochloride (Gu·HCl) at a concentration of 3.0 M were used as denaturation agents.Increasing the concentration of PTU from 0.8 × 10−5 to 1.20 × 10−4 M in the systems with HSA causes a decrease in fluorescence intensity of the protein excited with both 280 and 295 nm wavelengths. The results indicate that urea and Gu·HCl bind to the carbonyl group and then to the NH-group. To determine binding constants we used the Scatchard plots. The presence of two classes of HSA–PTU binding sites was observed. The binding constants (Kb) are equal to 1.99 × 104 M−1 and 1.50 × 104 M−1 at λex = 280 nm, 5.20 × 104 M−1 and 1.65 × 104 M−1 at λex = 295 nm. At λex = 280 nm the number of drug molecules per protein molecule is aI = 1.45 and aII = 1.32 for I and II binding sites, respectively. At λex = 295 nm they are aI = 0.63 and aII = 1.54 for the I and II binding sites.The estimation of the binding ability of changed albumin in the uremic and diabetic patients suffering from chronic liver or renal diseases is very important for safety and effective therapy.  相似文献   

15.
The hollow fiber composite membrane involving Zr0.84Y0.16O1.92 (YSZ) as an oxygen ionic conductor and La0.8Sr0.2MnO3−δ (LSM) as an electronic conductor was explored for oxygen separation application. The hollow fiber precursor was prepared by the phase-inversion process, and transformed to a gas-tight ceramic by sintering at 1350 °C. The as-prepared fiber exhibited a thermal expansion coefficient of 11.1 × 10−6 K−1 and a three-point bending strength of 152 ± 12 MPa. An oxygen permeation flux of 2.1 × 10−7 mol cm−2 s−1 was obtained under air/He gradient at 950 °C for a hollow fiber of length 57.00 mm and wall thickness 0.16 mm. The oxygen permeation flux remained unchanged when the sweeping gas was changed from helium to high concentration of CO2. Considering the satisfactory trade-off between the permeability and stability, the YSZ–LSM hollow fiber is promising for oxygen production applications.  相似文献   

16.
The voltammetric responses observed for carbohydrates and polyalcohols at 0.60 V in 0.10 M NaOH are significantly larger at preanodized CuMn (95:5) electrodes as compared to preanodized pure Cu electrodes. Apparent values for the number of electrons transferred (napp) and the corresponding values of heterogeneous rate constants (kapp) are estimated for selected reactants from the slopes and intercepts, respectively, of Koutecký–Levich plots of background-corrected voltammetric currents obtained at CuMn and Cu rotated disk electrodes (RDEs). Values of napp (and kapp) for sorbitol and glucose are 11.8 (9.2×10−3 cm s−1) and 11.7 (8.0×10−3 cm s−1), respectively, at a CuMn RDE. These are compared to the values 10.4 (1.8×10−3 cm s−1) and 9.6 (2.0×10−3 cm s−1) for sorbitol and glucose, respectively, at a Cu RDE. The larger sensitivities observed at the CuMn RDE in comparison to the Cu RDE are concluded to be the beneficial result of larger kapp values at the alloy electrode. Furthermore, the larger kapp values are speculated to result from enhanced preadsorption of the reactant species at Mn(IV) sites in the preanodized CuMn surface. In flow-injection measurements, the peak signals obtained for successive injections of glucose using a CuMn electrode (0.60 V vs. SCE) were quite stable with a standard deviation of 1.5%. However, large day-to-day variations (±15%) observed in the average peak signals are attributed to the temperature sensitivities of the kapp value and the diffusion coefficient for glucose.  相似文献   

17.
D.F. Zhou  Y.J. Xia  J.X. Zhu  J. Meng   《Solid State Sciences》2009,11(9):1587-1591
Ce6−xDyxMoO15−δ (0.0 ≤ x ≤ 1.8) were synthesized by modified sol–gel method. Structural and electrical properties were investigated by means of X-ray diffraction (XRD), Raman, X-ray photoelectron spectroscopy (XPS) and electrochemical impedance spectroscopy (EIS). The XRD patterns showed that the materials were single phase with a cubic fluorite structure. Impedance spectroscopy measurement in the temperature range between 350 °C and 800 °C indicated a sharp increase in conductivity for the system containing small amount of Dy2O3. The Ce5.6Dy0.4MoO15−δ detected to be the best conducting phase with the highest conductivity (σt = 8.93 × 10−3 S cm−1) is higher than that of Ce5.6Sm0.4MoO15−δ (σt = 2.93 × 10−3 S cm−1) at 800 °C, and the corresponding activation energy of Ce5.6Dy0.4MoO15−δ (0.994 eV) is lower than that of Ce5.6Sm0.4MoO15−δ (1.002 eV).  相似文献   

18.
Pulse radiolysis transient UV–visible absorption spectroscopy was used to study the UV–visible absorption spectrum (225–575 nm) of the phenyl radical, C6H5(), and kinetics of its reaction with NO. Phenyl radicals have a strong broad featureless absorption in the region of 225–340 nm. In the presence of NO phenyl radicals are converted into nitrosobenzene. The phenyl radical spectrum was measured relative to that of nitrosobenzene. Based upon σ(C6H5NO)270 nm=3.82×10−17 cm2 molecule−1 we derive an absorption cross-section for phenyl radicals at 250 nm, σ(C6H5())250 nm=(2.75±0.58)×10−17 cm2 molecule−1. At 295 K in 200–1000 mbar of Ar diluent k(C6H5()+NO)=(2.09±0.15)×10−11 cm3 molecule−1 s−1.  相似文献   

19.
The kinetics of the CCl2 + Br2 and CCl2 + NO2 reactions have been studied at temperatures between 266 and 365 K using laser photolysis/photoionization mass spectrometry. Dichloromethylene biradicals were produced by the pulsed laser photolysis of CCl4. The bimolecular rate coefficients of the CCl2 + Br2 reaction can be described by the Arrhenius expression k1 = (7.05 ± 1.75) × 10−12 exp[(3.52 ± 0.63) kJ mol−1/RT] cm3 molecule−1 s−1. CCl2Br was observed as a primary product of this reaction. Interestingly, the bimolecular rate coefficients of the CCl2 + NO2 reaction were observed to depend weakly on the bath gas density and to possess a negative temperature dependence.  相似文献   

20.
The mediated oxidation of N-acetyl cysteine (NAC) and glutathione (GL) at the palladized aluminum electrode modified by Prussian blue film (PB/Pd–Al) is described. The catalytic activity of PB/Pd–Al was explored in terms of FeIII[FeIII(CN)6]/FeIII[FeII(CN)6]1− system by taking advantage of the metallic palladium layer inserted between PB film and Al, as an electron-transfer bridge. The best mediated oxidation of NAC and GL on the PB/Pd–Al electrode was achieved in 0.5 M KNO3 + 0.2 M potassium acetate of pH 2. The mechanism and kinetics of the catalytic oxidation reactions of the both compounds were monitored by cyclic voltammetry and chronoamperometry. The charge transfer-rate limiting step as well as overall oxidation reaction of NAC or GL is found to be a one-electron abstraction. The values of transfer coefficients α, catalytic rate constant k and diffusion coefficient D are 0.5, 3.2 × 102 M−1 s−1 and 2.45 × 10−5 cm2 s−1 for NAC and 0.5, 2.1 × 102 M−1 s−1 and 3.7 × 10−5 cm2 s−1 for GL, respectively. The modifying layers on the Pd–Al substrate have reproducible behavior and a high level of stability in the electrolyte solutions. The modified electrode is exploited for hydrodynamic amperometry of NAC and GL. The amperometric calibration graph is linear in concentration ranges 2 × 10−6–40 × 10−6 for NAC and 5 × 10−7–18 × 10−6 M for GL and the detection limits are 5.4 × 10−7 and 4.6 × 10−7 M, respectively.  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