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 33 毫秒
1.
The reaction between Au(I), generated by reaction of thallium(I) with Au(III), and peroxydisulphate was studied in 5 mol dm?3 hydrochloric acid. The reaction proceeds with the formation of an ion‐pair between peroxydisulphate and chloride ion as the Michealis–Menten plot was linear with intercept. The ion‐pair thus formed oxidizes AuCl2? in a slow two‐electron transfer step without any formation of free radicals. The ion‐pair formation constant and the rate constant for the slow step were determined as 113 ± 20 dm?3 mol?1 and 5.0 ± 1.0 × 10?2 dm3 mol?1 s?1, respectively. The reaction was retarded by hydrogen ion, and formation of unreactive protonated form of the reductant, HAuCl2, causes the rate inhibition. From the hydrogen ion dependence of the reaction rate, the protonation constant was calculated to be as 0.6 ± 0.1 dm3 mol?1. The activation parameters were determined and the values support the proposed mechanism. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 589–594, 2002  相似文献   

2.
The enthalpy of hydrolysis of solid tungsten azide pentafluoride in alkaline aqueous solution (1.O mol dm?3 KOH; 298.2K) is ?578 kJ mol?1. Hence its enthalpy of formation is ?1170 kJ mol?1.  相似文献   

3.
The kinetics of the binding of the neurotoxin acrylamide to the cysteine residue of glutathione has been studied. At 37 °C and pH 7.3 the second order rate constant has been determined to be 0.72 ± 0.06 mol?1 dm3 min?1 by thermospray mass spectrometry. The critical energy at pH 11.5 measured over the temperature range 10–37°C by fast atom bombardment mass spectrometry was measured as 24.6 kJ mol?1.  相似文献   

4.
Tautomerization of 2-benzylidene-4-methyl-3-oxo-pentanoic acid phenylamide has been studied by NMR and GC-MS. The two tautomers were separated on an HP-5 column, which enabled the kinetic and the thermodynamic behavior of on-column interconversion to be investigated. The enol-to-imide tautomerization was found to occur primarily in the stationary phase. By treating the column as a reactor, the interconversion was investigated as a function of retention time and oven temperature. This enabled determination of the rate constant (0.0605 s?1) by monitoring the increase of the less gas stable tautomer at a constant temperature of 260 °C and determination of the activation energy of the reaction for the net tautomerization (52.0 kJ mol?1), because it was found that the reaction obeyed pseudo first-order kinetics. The enthalpy and the entropy changes (?H=1.68 kJ mol?1, ?S=3.54 J K?1 mol?1) for the enol-to-imide reaction in the stationary phase were also obtained.  相似文献   

5.
Abstract— The equilibrium constants, Kc, for complexation between methyl viologen dication (MV2+) and Rose Bengal, or Eosin Y, decrease with increasing ionic strength. At zero ionic strength Kc is 6500 (± 500) mol?1 dm3 for Rose Bengal and 3200 (± 200) mol?1 dm3 for Eosin Y, and these values decrease to 1500 (± 100) and 680 (± 40) mol?1 dm3, respectively, at an ionic strength of 0.1 mol dm?3. Kc is independent of pH between 4.5 and 10. ΔH is -25 (± 1) kJ mol?1 for complexation with either dye, whereas ΔS is -15 (± 3) J K?1 mol?1 for Rose Bengal, and - 23 (± 3) J K?1 mol?1 for Eosin Y. The complexation constant for Rose Bengal and the neutral viologen, 4,4'-bipyridinium-N, N'-di(propylsulphonate), (4,4'-BPS), is 420 (± 35) mol?1 dm3, and independent of ionic strength. No complexation could be observed for either Rose Bengal or Eosin with another neutral viologen, 2,2'-bipyridinium-N,N'-di(propylsulphonate), (2,2'-BPS). MV2+ quenches the triplet state of Rose Bengal with a rate constant of 7 × 109 mol?1 dm3 s?1, and this rate constant decreases slightly as ionic strength increases. The cage escape yield following quenching, Φcc is very low (Φcc= 0.02 (± 0.005), and independent of ionic strength. 4,4'-BPS quenches the triplet state of Rose Bengal with a rate constant of 2.2 (± 0.1) × 109 mol?1 dm3 s?1, and gives a cage escape yield of 0.033 (± 0.006). 2,2'-BPS quenches the Rose Bengal triplet with a rate constant of 6 (± 1) × 108 mol?1 dm3 s?1 and gives a cage escape yield of 0.07 (± 0.01). Conductivity measurements indicate that MV2+(Cl?)2 is completely dissociated at concentrations below 2 × 10?2 mol dm?3.  相似文献   

6.
《Electroanalysis》2004,16(20):1690-1696
The electrode mechanism of Mo(VI) reduction was studied under catalytic adsorptive stripping mode by means of square‐wave voltammetry (SWV). Mo(VI) creates a stable surface active complex with mandelic acid. The electrode reaction of Mo(VI)‐mandelic acid system undergoes as one‐electron reduction, exhibiting properties of a surface electrode process. In the presence of chlorate, bromate, and hydrogen peroxide, the electrode reaction is transposed into a catalytic mechanism. The experimental results are compared with the recent theory for surface catalytic reaction, enabling qualitative characterization of the electrode mechanism in the presence of different catalytic agents. Utilizing both the method of “split SW peaks” and “quasireversible maximum” the standard redox rate constant of Mo(VI)‐mandelic acid system was estimates as ks=150±5 s?1. By fitting the experimental and theoretical results, the following catalytic rate constants have been estimated: (8.0±0.5)×104 mol?1 dm3 s?1, (1.0±0.1)×105 mol?1 dm3 s?1, and (3.2±0.1)×106 mol?1 dm3 s?1, for hydrogen peroxide, chlorate, and bromate, respectively.  相似文献   

7.
The effect of water vapor pressure on the molecular weight of molten poly (ethylene terephthalate) has been followed by measurement of the changes which occur in the intrinsic viscosity and the end-group concentrations upon hydrolysis. It was found that phosphoric acid is highly effective as a stabilizer; the hydrolysis rate constant (Kh) decreasing from 7.6 × 10?7 mol?1s?1 for nonstabilized to 2.5 × 10?7 mol?1s?1 for the stabilized polyester at 290°C and water vapor pressure of 20 mmHg.  相似文献   

8.
The kinetics of formation of the 1?:?1 complex of chromium(III) with 1,3-propanediamine-N,N′-diacetate-N,N′-di-3-propionate (1,3-pddadp) were followed spectrophotometrically at λ max?=?557?nm. The reaction was first-order in chromium(III). Increasing the 1,3-pddadp concentration from 2.2?×?10?2 to 0.11?mol?dm?3 accelerated the reaction rate. Increasing the hydrogen ion concentration from 1.995?×?10?5 to 6.31?×?10?4 mol?dm?3 retarded the reaction rate. The reaction rate was also retarded by increasing ionic strength and dielectric constant of the reaction medium. A mechanism was suggested to account for the results obtained which involves ion-pair formation between the various reactants. Values of 22?kJ?mol?1 and ?115?J?K?1 mol?1 were obtained for the energy and the entropy of activation, respectively, which indicate an associative mechanism. The logarithm of the formation constant of the 1?:?1 complex formed was 11.3.  相似文献   

9.
For a temperature range of −11.8–92.6°C, the propagation rate constant kp of styrene has been determined with the use of pulsed-laser polymerization (PLP). The temperature dependency of the obtained kp data was evaluated using the Arrhenius equation. The NLLS error-in-variables method (EVM) is recommended for this fit. The resulting activation energy is 32.6 kJ mol−1 and the pre-exponential factor is 107.66 dm3 mol−1 s−1. A joint confidence interval for these parameters is given. © 1996 John Wiley & Sons, Inc.  相似文献   

10.
Abstract Laser flash photolysis and pulse radiolysis have led to the characterisation of several shortlived intermediates formed after irradiation of retinoic acid and retinyl acetate in hexane or methanol. For retinoic acid, the triplet state, wavelength maximum 440 nm, extinction coefficient 7.3 × 104 dm3 mol?1 cm?1, decay constant 6.2 × 105 s?1, is formed with a quantum yield of 0.012 for 347 nm excitation. The radical cation, absorption maximum 590 nm, extinction coefficient ~7 × 104 dm3mol?1 cm?1, is formed in a biphotonic process. The radical anion, absorption maximum 510nm in hexane, 480 nm in methanol where its extinction coefficient is 1.2 × 105 dm3mol?1 cm?1, appears to decay partially in methanol into another longer-lived neutral radical, wavelength maximum 420 nm, by loss of OH?. For retinyl acetate, the triplet state, absorption maximum 395 nm, extinction coefficient 7.9 × 104dm3mol?1 cm?1, decay constant 1.2 × 106s?1 is formed with a quantum yield of 0.025 for 347 nm excitation. Monophotonic photoelimination of OCOCH3? in methanol produces the retinylic carbenium ion, wavelength maximum 590 nm, whose decay is enhanced by ammonia, k ~ 2 × 106 dm3 mol?1 s?1 and retarded by water. The radical cation also has a wavelength maximum at 590 nm, its extinction coefficient being ~ 1.0 × 105 dm3mol1 cm?1. The long-lived transient absorption with maximum at 385 nm, extinction coefficient 1.0 × 105 dm3mol?1 cm?1, obtained from the reaction of the solvated electron with retinyl acetate in methanol may be due to either the radical anion itself or more likely the radical resulting from elimination of OCOCH3? from this anion. These results suggest that skin photosensitivity caused by retinyl acetate might be greater than that due to retinoic acid.  相似文献   

11.
The volatilisation of ferrocene (Fc), dissolved in the ionic liquid N‐butyl‐N‐methylpyrrolidinium bis(trifluoromethylsulfonyl)imide, [C4mpyrr][NTf2], to the gas phase has been indirectly monitored by cyclic voltammetry and chronoamperometry. Simulation of the observed trends in concentration with time using a simple model allowed quantification of the process. Volatilisation of dissolved Fc under flowing wet and dry dinitrogen gas (N2) was found to be kinetically limited with a rate constant in the region of 2×10?7 cm s?1. The activation energy of diffusion for Fc was found to be 28.2±0.7 kJ mol?1, while the activation energy of volatilisation of Fc from [C4mpyrr][NTf2] to dry N2 was found to be 85±2 kJ mol?1.  相似文献   

12.
The kinetics of interaction between di-μ-hydroxobis(1,10-phenanthroline)dipalladium(II) perchlorate and thioglycolic acid and with glutathione has been studied spectrophotometrically in aqueous medium as a function of the complex concentration as well as the ligand concentrations, pH, and temperature at constant ionic strength. The observed pseudo-first-order rate constants k obs (s?1) obeyed the equation k obs = k 1[Nu] (Nu = nucleophile). At pH = 6.5, the interaction with thioglycolic acid shows two distinct consecutive steps and both steps are dependent on the concentration of thioglycolic acid. The rate constants for the process are: k 1 ≈ 10?5 s?1 and k 2 ≈ 10?3 dm3 · mol?1 · s?1. The association equilibrium constant (K E) for the outer sphere complex formation has been evaluated together with the rate constants for the two subsequent steps. The other bio-active ligand, glutathione, showed a single step reaction depending on [ligand] with a second-order anation rate constant: the 102 (k 2) values are (61.72, 79.20, 109.24 and 154.33) dm3 · mol?1 · s?1 at 20, 25, 30 and 35 °C, respectively. On the basis of the kinetic observations and evaluated activation parameters, plausible associative mechanisms are proposed for both interaction processes.  相似文献   

13.
The spectrocoulometric technique reported earlier is applied to verify the mechanism and to evaluate the contributions kBi of the individual bases to the total rate constant k of the hydrolysis of the tris (1,10-phenanthroline) iron(III) complex, Fe (phen)3+3. Both normal and “open-circuit” spectrocoulometric experiments are used. Partial rate constants for four bases in the acetate-buffered solutions are kH2O=(3.4±1.2) × 10?4s?1 (kH2O includes the H2O concentration), kOH=(1.20±0.06)×107 mol?1dm3s?1, kphen=(1.4±0.2) mol?1dm3s?1, kAc=(3.8±0.3)×10?2 mol?1dm3s?1, at 25°C and ionic strength 0.5 mol dm?3. The Fe(phen)3+3 hydrolysis, with (phen)2 (H2O) Fe-O-Fe (H2O) (phen)4+2 formation, is first order with respect to Fe (phen)3+3 and the bases present in the solution. The rate-determining step in the hydrolysis is the entry of a base to the coordinating sphere of the complex, as in the hydrolysis of the analogous 2,2′-bipyridyl complex.  相似文献   

14.
The hydrolysis kinetics of CICN have been reinvestigated from pH 0.0–10.5 and from 18–40°C. In the pH range from 1–5, the hydrolysis rate is invariant and the activation parameters (ΔH? = 84 kJ mol?1 and ΔS? = ?84 J mol?1 K?1) are consistent with water attack. In basic solution the rate is first order each in CICN and OH? concentrations with parameters ΔH? and ΔS? equal to 82 kJ mol?1 and + 54 J mol?1 K?1, respectively. The rate constants with 20 other donors have been measured. Nitrogen nucleophiles are more reactive than oxygen donors, and an alpha-effect is seen. The constants follow a pattern indicative of attack at carbon. Cyanate in its acid form reacts with nucleophiles. Further points on the cyanate rate–pH profile have been obtained. A chromate-catalyzed hydrolysis can contribute between pH 5–10. Some studies were made of the reaction of cyanate with hydrogen peroxide. Free energy correlations are presented.  相似文献   

15.
Herein, we report that the thermodynamic barrier for solution-phase electron transfer (ET) between Cu(II) and Fe(II) in aqueous acidic media can be overcome through the addition of 2,9-dimethyl-1,10-phenanthroline (Neocuproine [NC]) to the reaction mixture. A detailed discussion of the kinetic and mechanistic aspects of this coordination-inspired ET is presented. We attribute the observed change in the thermodynamic feasibility to the change in the reduction potential of Cu(II)–Cu(I) couple on its ligation with NC. The reaction was found to be slow, following first-order kinetics with respect to each Cu(II) and Fe(II). In the presence of excess NC, the reaction was observed to proceed with a pseudo-second-order rate constant of 3.37?±?0.05?dm3?mol?1?s?1 at 298?K, with an activation barrier of ca. 26.22?kJ?mol?1. The slow reaction is attributed to the significant reorganization energy associated with large-scale changes in the coordination sphere of the oxidant. A two-step mechanism that explains the experimental observations is proposed for the investigated reaction.  相似文献   

16.
The kinetics of the formation of the titanium‐peroxide [TiO2+2] complex from the reaction of Ti(IV)OSO4 with hydrogen peroxide and the hydrolysis of hydroxymethyl hydroperoxide (HMHP) were examined to determine whether Ti(IV)OSO4 could be used to distinguish between hydrogen peroxide and HMHP in mixed solutions. Stopped‐flow analysis coupled to UV‐vis spectroscopy was used to examine the reaction kinetics at various temperatures. The molar absorptivity (ε) of the [TiO2+2] complex was found to be 679.5 ± 20.8 L mol?1 cm?1 at 405 nm. The reaction between hydrogen peroxide and Ti(IV)OSO4 was first order with respect to both Ti(IV)OSO4 and H2O2 with a rate constant of 5.70 ± 0.18 × 104 M?1 s?1 at 25°C, and an activation energy, Ea = 40.5 ± 1.9 kJ mol?1. The rate constant for the hydrolysis of HMHP was 4.3 × 10?3 s?1 at pH 8.5. Since the rate of complex formation between Ti(IV)OSO4 and hydrogen peroxide is much faster than the rate of hydrolysis of HMHP, the Ti(IV)OSO4 reaction coupled to time‐dependent UV‐vis spectroscopic measurements can be used to distinguish between hydrogen peroxide and HMHP in solution. © 2007 Wiley Periodicals, Inc. Int J Chem Kinet 39: 457–461, 2007  相似文献   

17.
The C‐phenyl‐Ntert‐butylnitrone/azobisisobutyronitrile pair is able to impart control to the radical polymerization of n‐butyl acrylate as long as a two‐step process is implemented, that is, the prereaction of the nitrone and the initiator in toluene at 85 °C for 4 h followed by the addition and polymerization of n‐butyl acrylate at 110 °C. The structure of the in situ formed nitroxide has been established from kinetic and electron spin resonance data. The key parameters (the dissociation rate constant, combination rate constant, and equilibrium constant) that govern the process have been evaluated. The equilibrium constant between the dormant and active species is close to 1.6 × 10?12 mol L?1 at 110 °C. The dissociation rate constant and the activation energy for the C? ON bond homolysis are 1.9 × 10?3 s?1 and 122 ± 15 kJ mol?1, respectively. The rate constant of recombination between the propagating radical and the nitroxide is as high as 1.2 × 109 L mol?1 s?1. Finally, well‐defined poly(n‐butyl acrylate)‐b‐polystyrene block copolymers have been successfully prepared. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 6299–6311, 2006  相似文献   

18.
The kinetics and mechanism of the reaction of complexation of iron(III) with 2,4-octanedione and 2,4-nonanedione have been investigated spectrophotometrically in aqueous solution at 10°C and ionic strength 0.5 mol dm?3 NaClO4. The equilibrium constants of the mono-complexes have been determined. The mechanism proposed to account for the kinetic data involves a double reversible pathway where both Fe3+ and Fe(OH)2+ react with the enol tautomer of the ligand. 2,4-Octanedione reacts with Fe3+ and Fe(OH)2+ with rate constants of 0.65 dm3 mol?1 s?1, and 14.07 dm3 mol?1 s?1, respectively. For 2,4-nonanedione complexation the rate constants determined are 0.49 dm3 mol?1 s?1, and 11.39 dm3 mol?1 s?1, respectively. Some discussions are made on the basis of Eigen-Wilkins theory considering the effect of solvent exchange on the complex formation. © 1993 John Wiley & Sons, Inc.  相似文献   

19.
The surface segregation of In and S from a dilute Cu(In,S) ternary alloy were measured using Auger electron spectroscopy coupled with a linear programmed heater. The alloy was linearly heated and cooled at constant rates. Segregation data of a linear heat run showed surface segregation of In that reached a maximum surface coverage of 25% followed by S, which reached a coverage of 30%. It was found that after In had reached a maximum surface coverage, it started to desegregate as soon as the S enriched the surface until In was completely replaced by S. The segregation parameters, namely, the pre‐exponential factor (D0), activation energy (Q), segregation energy (ΔG?) and interaction energy (Ω) were extracted from the measured segregation data for both In and S segregation in Cu by simulating the measured segregation data with a theoretical segregation model (modified Darken model). The segregation parameters obtained for In segregation in Cu are D0 = 1.8 ± 0.5 × 10?5 m2 s?1, Q = 184.3 ± 1.0 kJ.mol?1, ΔG? = ?61.4 ± 1.4 kJ.mol‐1, ΩCu?In = 3.0 ± 0.4 kJ.mol?1; for S segregation in Cu the parameters are D0 = 8.9 ± 0.5 × 10?3 m2 s?1, Q = 212.8 ± 3.0 kJ.mol?1, ΔG? = ?120.0 ± 3.5 kJ.mol?1, ΩCu?S = 23.0 ± 2.0 kJ mol?1 and the In and S interaction parameter is ΩIn?S = ?4.0 ± 0.5 kJ.mol?1. The initial parameters used for the Darken calculations were extracted from fits performed with the Fick's and Guttmann model. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

20.
The kinetics of oxidation of amino acids viz. glycine, alanine, and threonine with bismuth(V) in HClO4–HF medium have been studied. The kinetics of the oxidation of all these amino acids exhibit similar rate laws. The second-order rate constants were calculated to be 2.04 × 10?2 dm3 mol?1 and 2.72 × 10?2 dm3 mol?1 s?1 for glycine and alanine, respectively, at 35°C and 5.9 × 10?2 dm3 mol?1 s?1 for threonine at 25°C. All the possible reactive species of both bismuth(V) and amino acids have been discussed and a most probable kinetic model in each reaction has been envisaged. © 1994 John Wiley & Sons, Inc.  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