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 15 毫秒
1.
The oxidation of methionine by freshly prepared colloidal manganese dioxide in aqueous as well as micellar media was studied spectrophotometrically at 35°C. The reaction between methionine and MnO2 in both media exhibits 1:1 stoichiometry (methionine:MnO2). The oxidation reaction is first order with regard to the MnO2 concentration, but is fractional-order in the methionine concentration and HClO4 concentrations. A catalytic effect of nonionic surfactant TX-100 on the rate of oxidation was observed and reaction rate was found to be proportional to {k′ + k″ [TX-100]}, where k′ and k″ are the rate constants in absence and presence of surfactant, respectively. The use of surfactant micelles is highlighted as, in favorable cases; the micelles help the redox reactions by bringing the reactants in a close proximity through hydrogen bonding. The oxidation reaction in aqueous and micellar media is shown to proceed via methionine–MnO2 and methionine–MnO2–TX-100 complexes, respectively, which decomposes slowly in a rate determining step to give methionine sulfoxide as the product. A suitable mechanism is proposed for these observations.  相似文献   

2.
Kinetics of oxidation of DL-malic acid by water soluble colloidal MnO2 (prepared from potassium permanganate and sodium thiosulfate solutions) have been studied spectrophotometrically in the absence and presence of nonionic Triton X-100 surfactant. The reaction is autocatalytic and manganese(II) (reduction product of the colloidal MnO2) may be the autocatalyst. The order of the reaction is first in colloidal [MnO2] as well as in [malic acid] both in the absence and presence of the surfactant. The reaction has acid-dependent and acid-independent paths and, in the former case, the order is fractional in [H+]. The effect of externally added manganese(II) is complex. The results show that the rate constant increases as the manganese(II) concentration is increased. It is not possible to predict the exact dependence of the rate constants on manganese(II) concentration, which has a series of reactions with other reactants. In the presence of TX-100, the observed effect on k is catalytic up to a certain [TX-100]; thereafter, an inhibitory effect follows. The catalytic effect is explained in terms of the mathematical model proposed by Tuncay et al. (in Colloids Surf A Physicochem Eng Aspects 149:279 3). Activation parameters associated with the observed rate constants (kobs/k) have also been evaluated and discussed.  相似文献   

3.
Electrochemical behavior of poly-3,4-ethylenedioxythiophene composites with manganese dioxide (PEDOT/MnO2) has been investigated by cyclic voltammetry and electrochemical quartz crystal microbalance at various component ratios and in different electrolyte solutions. The electrochemical formation of PEDOT film on the electrode surface and PEDOT/MnO2 composite film during the electrochemical deposition of manganese dioxide into the polymer matrix was gravimetrically monitored. The mass of manganese dioxide deposited into PEDOT at different time of electrodeposition and apparent molar mass values of species involved into mass transfer during redox cycling of PEDOT/MnO2 composites were evaluated. It was found that during the redox cycling of PEDOT/MnO2 composite films with various MnO2 content, the oppositely directed fluxes of counterions (anions and cations) occur, resulting in a change of the slope of linear parts of the Δf–E plots with changing the mass fraction of MnO2 in the composite film.Rectangular shape of cyclic voltammograms of PEDOT/MnO2 composites with different loadings of manganese dioxide was observed, which is characteristic of the pseudocapacitive behavior of the composite material. Specific capacity values of PEDOT/MnO2 composites obtained from cyclic voltammograms were about 169 F g?1. The specific capacity, related to the contribution of manganese dioxide component, was about 240 F g?1.  相似文献   

4.
A method for determining the rate constant of disproportionation of 2,5-dichlorosemiquinone radicals (k 6) from unsteady-state kinetic data for the initiated chain reaction of N,N′-diphenyl-1,4-benzoquinone diimine with 2,5-dichlorohydroquinone have been developed, and two variants of this method are presented. The method is based on the study of the unsteady-state disappearance kinetics of one of the initial reactants (quinone diimine) in its initiated chain reaction with hydroquinone. The unsteadiness of the reaction is due to the presence of semiquinone radicals or initiator radicals accumulated before the start of the reaction at a concentration exceeding the steady-state concentration of semiquinone radicals in the chain reaction. The variants of the method differ in the order of mixing the reactants and initiator, on which the nature and concentration of the radicals accumulated in the system before the reaction depend. In the first variant, a quinone diimine + initiator solution is initially prepared and initiator radicals are accumulated. Hydroquinone is added to this solution (start of the reaction). In the second variant, a hydroquinone + initiator solution is initially prepared and semiquinone radicals from hydroquinone are accumulated. Quinone imine is then added to the solution (start of the reaction). The disproportionation rate constant of semiquinone radicals (k 6) is derived from the dependence of the decrease in the quinone imine concentration in a certain short time (∼20 s) after the start of the reaction on the initiation rate. The rate constant k 6 in benzene is (7.3 ± 3.7) × 106 l mol−1 s−1 according to the first variant of the method and (5.0 ± 2.2) × 106 l mol−1 s−1 according to the second one.  相似文献   

5.
Tetraamminezinc(II) dipermanganate ([Zn(NH3)4](MnO4)2; 1 ) was prepared, and its structure was elucidated with XRD‐Rietveld‐refinement and vibrational‐spectroscopy methods. Compound 1 has a cubic lattice consisting of a 3D H‐bound network built from blocks formed by four MnO anions and four [Zn(NH3)4]2+ cations. The other four MnO anions are located in a crystallographically different environment, namely in the cavities formed by the attachment of the building blocks. A low‐temperature quasi‐intramolecular redox reaction producing NH4NO3 and amorphous ZnMn2O4 could be established occurring even at 100°. Due to H‐bonds between the [Zn(NH3)4]2+ cation and the MnO anion, a redox reaction took place between the NH3 and the anion; thus, thermal deammoniation of compound 1 cannot be used to prepare [Zn(NH3)2](MnO4)2 (contrary to the behavior of the analogous perrhenate (ReO ) complex). In solution‐phase deammoniation, a temperature‐dependent hydrolysis process leading to the formation of Zn(OH)2 and NH4MnO4 was observed. Refluxing 1 in toluene offering the heat convecting medium, followed by the removal of NH4NO3 by washing with H2O, proved to be an easy and convenient technique for the synthesis of the amorphous ZnMn2O4.  相似文献   

6.
    
Summary By using -manganese dioxide (MnO2) as an indicator electrode, the electrode reaction of ethylenediaminetetraacetic acid (EDTA) and other complexanes was investigated. A characteristic cathodic peak of manganese dioxide which increases in the presence of complexanes was observed. It was clarified that the manganese dioxide reduction on the electrode is accelarated by the complex formation with complexanes. The MnO2 electrode seems to be useful for determining a chelating agent such as the EDTA-type complexanes.
Untersuchung der Elektrodenreaktion von Komplexanen an der -Mangandioxid-Elektrode in saurer Lösung im Hinblick auf ihre Bestimmung
Zusammenfassung Die Elektrodenreaktion von EDTA, IDA, NTA und DTPA wurde an der -MnO2-Indicatorelektrode untersucht. Ein charakteristischer kathodischer Peak wurde beobachtet, der auf der Reduktion des MnO2 beruht und in Gegenwart von Komplexanen zunimmt. Die Reduktion wird durch die Komplexbildung beschleunigt. Die Elektrode scheint zur Bestimmung von Chelatbildnern vom EDTA-Typ geeignet zu sein.
  相似文献   

7.
The voltammetric exploration of ruthenocene RuCp2 in aqueous solution via abrasive modification on a basal plane pyrolytic graphite electrode is reported. It is found that ruthenocene undergoes a one‐electron electrochemically irreversible oxidation to ruthenocenium, which then rapidly dimerizes to form [RuCp2] . Then, depending on the anion in solution, either disproportionates to [RuCp2]2+ and subsequently decomposes to ruthenium oxide, or is anion stabilized. SEM images and cyclic voltammetry of the electrode surface indicates that in the presence of smaller anions the dimer is not stabilized, whereas for larger anions, such as sodium toluene sulfonate, the dimer is stabilized and decomposition is inhibited.  相似文献   

8.
Summary Oxidation of Mn aq 2+ by HSO 5 in acetate buffer to manganese(IV) is autocatalytic, and obeys a rate expression of the general form -d[MnII]/dt = k0[MnII] + k1[MnII][MnOx]. The first-order (k0) and heterogenetic (k1) rate constants show first-order dependences on [HSO 5 ] and on 1/[H+]. The reaction is catalyzed by the addition of the chelating ligand glycine; k1 shows a first-order dependence on [glycine] at a fixed pH. This catalysis is ascribed to complexation, whereby the redox potential for Mn(gly) n (2–n)+ is lower than that for Mn aq 2+ , facilitating oxidation. The stoichiometry of the reaction is Mn2+: HSO 5 = 11, and the manganese(IV) oxide formed is of battery-active grade. Purity of the recovered product is not affected by the presence of high concentrations of natural sugars in the initial solution.  相似文献   

9.
One kinetic model for the oxidation of iodide ion by peroxydisulfate ion in aqueous solution is proposed. The reaction is regarded as \documentclass{article}\pagestyle{empty}\begin{document} {\rm S}_2 {\rm O}_8^{2 -} + {\rm I}^ - {\rm IS}_2 {\rm O}_8^{3 -} \end{document}, followed by the reaction \documentclass{article}\pagestyle{empty}\begin{document} {\rm IS}_2 {\rm O}_8^{3 -} + {\rm I}l_2 + 2{\rm SO}_4^{2 -} \end{document}. If the initial rates V are obtained from the formation of the iodine molecules, the reaction rate constant k1 and the ratio k2/k-1 can be estimated by plotting the values of [S2O82?][I?]/V against that of 1/[I?]. The extrapolated value for k1 is 2.20×10?2 L/mol-sec and k2/k-1 is calculated to be 4.25×102 mol/L at 27°C in a solution with an ionic strength of 0.420.  相似文献   

10.
The kinetics of the anionic polymerization of octamethylcyclotetrasiloxane (D4) initiated by α-methylstyrene living polymer in tetrahydrofuran was studied. The following kinetic scheme was postulated: Initiation: Propagation: where S- and M represent the initiator and D4, respectively. At a living end concentration of 0.0377 mole/l. and a monomer concentration of 1.5 mole/l. in tetrahydrofuran at 25°C. the following kinetic data were obtained: k1 = 2.3 × 10?4 l./mole-sec., k2 < 2.3 × 10?5 sec.?1, k3 = 2.75 × 10?2l./mole-sec. k4 ≈ 1.17 × 10?2 sec.?1, K1 > 10 l./mole and K2 ≈ 2.35 l./mole. The rate constants k1 and k3 were found to be dependent on the concentration of anions. This is attributed to the dissociation of ion pairs to free ions at lower concentration. Under the experimental conditions studied the majority of the anions were present in the form of ion pairs. The reactivity of the free ions is about 100 times greater than that of ion pairs. There is no temperature effect on K2, indicating zero ΔH and positive ΔS in the propagation reaction.  相似文献   

11.
The photooxidation of chloral was studied by infrared spectroscopy under steady-state conditions with irradiation of a blackblue fluorescent lamp (300 nm < λ < 400 nm, λmax = 360 nm) at 296 ± 2 K. The products were hydrogen chloride, carbon monoxide, carbon dioxide, and phosgen. The kinetic results reveal that the reaction proceeds via chain reaction of the Cl atom: The results lead to the conclusion that mechanism (B) is confirmed to be more likely than mechanism (A), which was favored at one time by Heicklen for the mechanism of the oxidation of trichloromethyl radicals by oxygen molecules: The ratio of the initial rates of CO and CO2 formation gave k7/k6 = 4.23M?1, and the lower limit of reaction (5) was found to be 3.7 × 108M?1 sec?1.  相似文献   

12.
Using dimethyl peroxide as a thermal source of methoxy radicals overthe temperature range of 110–160°C, and the combination of methoxy radicals and nitrogen dioxide as a reference reaction: a value was determined of the rate constant for the reaction of methoxy radicals with oxygen: is independent of nitrogen dioxide or oxygen concentration and added inert gas (carbon tetrafluoride). No heterogeneous effects were detected. The value of k4 is given by the expression In terms of atmospheric chemistry, this corresponds to a value of 105.6 M?1·sec?1 at 298 K. Extrapolation to temperatures where the combustion of organic compounds has been studied (813 K) produces a value of 107.7 M?1·sec?1 for k4. Under these conditions, reaction (4) competes with hydrogen abstraction or disproportionation reactions of the methoxy radical and its decomposition (3): In particular k3 is in the falloff region under these conditions. It is concluded that reaction (4) takes place as the result of a bimolecular collision process rather than via the formation of a cyclic complex.  相似文献   

13.
The kinetics of dimethyl sulfoxide (DMSO) oxidation by peroxomonophosphoric acid (PMPA) in aqueous medium at 308 K and I = 0.4 mol/dm3 follow the rate expressions In the pH range from 0 to 2, where k1 and k2 are 5.092 × 10?1 dm3/mol sec and ? 0, respectively; in the pH range from 4 to 7, where k2 = 8.127 × 10?3 and k3 = 2.90 × 10?3 dm3/mol sec; and in the pH range from 10 to 13.6, where k4 ? 0, and k5 = 3.08 × 10?2 dm3/mol sec. The reaction is interpreted in terms of mechanisms involving an electrophilic and a nucleophilic attack of the peroxomonophosphoric acid species, respectively, in acid and alkaline regions, on the sulfur atom of the sulfoxide molecule giving rise to S-type transition states followed by oxygen-oxygen bond fission to form the products.  相似文献   

14.
The reaction between chromium(VI) and L-ascorbic acid has been studied by spectrophotometry in the presence of aqueous citrate buffers in the pH range 5.69–7.21. The reaction is slowed down by an increase of the ionic strength. At constant ionic strength, manganese(II) ion does not exert any appreciable inhibition effect on the reaction rate. The rate law found is where Kp is the equilibrium constant for protonation of chromate ion and kr is the rate constant for the redox reaction between the active forms of the oxidant (hydrogenchromate ion) and the reductant (L-hydrogenascorbate ion). The activation parameters associated with rate constant kr are Ea = 20.4 ± 0.9 kJ mol?1, ΔH = 17.9 ± 0.9 kJ mol?1, and ΔS=?152 ± 3 J K?1 mol?1. The reaction thermodynamic magnitudes associated with equilibrium constant Kp are ΔH0 = 16.5 ± 1.1 kJ mol?1 and ΔS0 = 167 ± 4 J K?1 mol?1. A mechanism in accordance with the experimental data is proposed for the reaction. © 1993 John Wiley & Sons, Inc.  相似文献   

15.
The pseudo–first‐order reaction rate constants (k0, s?1) for the reaction of carbon dioxide in aqueous solutions of sodium taurate (NaTau) and sodium prolinate (NaPr) were measured using a stopped‐flow technique at a temperature range of 298–313 K. The solutions concentration varied from 5 to 50 mol m?3 and from 4 to 12 mol m?3 for NaTau and NaPr, respectively. Comparing the k0 values, aqueous NaPr was found to react very fast with CO2 as compared with the industrial standard aqueous monoethanolamine (MEA) and aqueous sodium taurate (NaTau) was found to react slower than aqueous MEA at the concentration range considered in this work. For the studied amino acid salts, the order of the reactions was found to be unity with respect to the amino acid salt concentration. Proposed reaction mechanisms such as termolecular and zwitterion reaction mechanisms for the reaction of CO2 with aqueous solutions were used for calculating the second‐order reaction rate constants (k2, m3 mol?1 s?1). The formation of zwitterion during the reaction with CO2 was found to be the rate‐determining step, and the deprotonation of zwitterion was instantaneous compared to the reverse reaction of zwitterion to form an amino acid salt. The contribution of water was established to be significant for the deprotonation of zwitterion. Comparing the pseudo–first‐order reaction rate constants (k0, s?1) of various amino acid salts with CO2, NaPr was found to be the faster reacting amino acid salt. The activation energy for NaTau was found to be 48.1 kJ mol?1 and that of the NaPr was found to be 12 kJ mol?1. The Arrhenius expressions for the reaction between CO2 and the studied amino acid salts are   相似文献   

16.
For the first time, the electrochemical oxygen reduction reaction (ORR), was investigated using cyclic voltammetry (CV) on the electrodeposited manganese oxide (MnO x )-modified glassy carbon (MnO x -GC) electrode in the room temperature ionic liquids (RTILs) of EMIBF4, i.e. 1-ethyl-3-methylimidazolium tetrafluoroborate (EMIBF4). The results demonstrated that, after being modified by MnO x on a GC electrode, the reduction peak current of oxygen was increased to some extent, while the oxidation peak current, corresponding to the oxidation of superoxide anion, i.e., O2 was attenuated in some degree, suggesting that MnO x could catalyze ORR in RTILs of EMIBF4, which is consistent with the results obtained in aqueous solution. To accelerate the electron transfer rate, multi-walled carbon nanotubes (MWCNTs) was modified the GC electrode, and then MnO x was electrodeposited onto the MWCNTs-modified GC electrode to give rise to a MnO x /MWCNTs-modified GC electrode, consequently, the improved standard rate constant, ks, originated from the modified MWCNTs, along with the modification of electrodeposited MnO x , showed us a satisfactory electrocatalysis for ORR in RTILs of EMIBF4. Published in Russian in Elektrokhimiya, 2009, Vol. 45, No. 3, pp. 340–345. The article is published in the original.  相似文献   

17.
Complex dynamical behavior has been observed in the oxidation of hydroxylamine by bromate in acidic sulfate medium. The reaction shows clock type kinetics in closed conditions and an aperiodic oscillations if gaseous products are removed from the system with a constant flow-rate. The reduction kinetics of bromate ions with excess hydroxylamine has been studied in the presence of allyl alcohol. The observed pseudo-first-order rate constant kobs has been found to follow the expression where [hydroxylamine] is total initial hydroxylamine concentration, K1 = 0.5 M?1, K2 = 106 M?1, and k = 2.57 × 103 M?1 s?1 at 298.15 K and I = 2.0 M. The rate constant for the bromine oxidation of hydroxylamine in sulfuric aqueous solution has been determined. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
Given the species A1 and A2, the competition among the three different elementary processes (1) (2) (3) is frequently found in thermal and photochemical reaction systems. In the present paper, an analytical resolution of the system (1)–(3), performed under plausible contour conditions, namely, finite initial molar concentrations for both reactants, [A2]0 and [A1]0, and nonzero reaction rate coefficients k1, k2, and k3, leads to the equation [A1] = ((δ[A2]γ ? [A2])/β) ? α, where α = k1/2k3, γ = β + 1 = 2k3/k2, and δ = ([A2]0 + β[A1]0 + β α))/[A2]0γ. The comparison with a numerical integration employing the fourth‐order Runge–Kutta algorithm for the well‐known case of the oxidation of organic compounds by ferrate ion is performed. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 562–566, 2010  相似文献   

19.
Water-soluble colloidal manganese dioxide has been used to oxidize l-tyrosine in aqueous-acidic medium. The kinetics of the reaction was studied in the absence and presence of non-ionic surfactant (TX-100) using a spectrophotometric technique. As the reaction was fast under pseudo-first-order conditions ([l-tyrosine]  [MnO2]), the rate constants as a function of [l-tyrosine], [MnO2], [HClO4] and temperature were obtained under second-order conditions. The rate of the reaction increased and decreased with the increase in [l-tyrosine] and [MnO2], respectively. Perchloric acid, sodium pyrophosphate and sodium fluoride showed catalytic effect. The effect of externally added manganese(II) sulphate is complex. It is not possible to predict the exact dependence of the rate constants on manganese(II) concentration, which has a series of reactions with other reactants. The reaction is inhibited by the non-ionic surfactant TX-100. Activation parameters have been evaluated using Arrhenius and Eyring equations. Based on observed kinetic results, a probable mechanism for the reaction has been proposed which corresponds to fast adsorption of the reductant and hydrogen ion on the surface of colloidal MnO2 followed by one-step two-electron transfer rate limiting process.  相似文献   

20.
The mechanism by which an excess of iron(II) ion reacts with aqueous chlorine dioxide to produce iron(III) ion and chloride ion has been determined. The reaction proceeds via the formation of chlorite ion, which in turn reacts with additional iron(II) to produce the observed products. The first step of the process, the reduction of chlorine dioxide to chlorite ion, is fast compared to the subsequent reduction of chlorite by iron(II). The overall stoichiometry is The rate is independent of pH over the range from 3.5 to 7.5, but the reaction is assisted by the presence of acetate ion. Thus the rate law is given by At an ionic strength of 2.0 M and at 25°C, ku = (3.9 ± 0.1) × 103 L mol?1 s?1 and kc = (6 ± 1) × 104 L mol?1 s?1. The formation constant for the acetatoiron(II) complex, Kf, at an ionic strength of 2.0 M and 25°C was found to be (4.8 ± 0.8) × 10?2 L mol?1. The activation parameters for the reaction were determined and compared to those for iron(II) ion reacting directly with chlorite ion. At 0.1 M ionic strength, the activation parameters for the two reactions were found to be identical within experimental error. The values of ΔH? and ΔS? are 64 ± 3 kJ mol?1 and + 40 ± 10 J K?1 mol?1 respectively. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 554–565, 2004  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