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 kinetics of oxidation of CoIIHEDTA {HEDTA = N-(2-hydroxyethyl)ethylenediamine-N,N,N-triacetic acid} by vanadate ion have been studied in aqueous acid in the pH range 0.75–5.4 at 43–57 °C. The reaction exhibits second-order kinetics; first-order in each of the reactants. The reaction rate is a maximum at pH = 2.1. A mechanism is proposed in which the species [CoIIHEDTA(H2O)] and VO2 + react to form an intermediate which decompose slowly to give pentadentate CoIIIHEDTA(H2O) and VIV as final products. The rate law was derived and the activation parameters calculated: H* = 26.96 kJ mol–1 and S* = –311.08 JK–1 mol–1.  相似文献   

2.
The effect of ion pair formation on the kinetics of the decay of the photoisomers and triplet states of cationic benzimidazolocyanine dyes is studied by flash photolysis. An increase in the rate constant of the reversecis-trans isomerization of the photoisomers is observed when ion pairs are formed (in nonpolar solvents). In the case of benzimidazolocyanine dyes with the I anion, ion-pair formation causes an increase in the rate constant of decay of the triplet state. Acceleration of S1 S0 internal conversion is discovered for the dyes with I–1 Translated fromIvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 507–512, March, 1993.  相似文献   

3.
The kinetics of oxidation of tartaric acid by Ce(IV) in the absence and presence of acrylamide has been investigated spectrophotometrically in aqueous H2SO4–HClO4 media at a constant ionic strength 2.0M and 25°C. Oxidation of tartaric acid in both cases was first order with respect to Ce(IV). Kinetic data showed that the reaction involves the formation of an unstable complex and an intermediate free radical. The activation parameters were calculated to be E a =91.3±0.4 kJ-mol–1, S=20.2±1.0 J-mol–1-K–1, H=88.8±0.4 kJ-mol–1. A polymerization mechanism is discussed.  相似文献   

4.
The kinetics of anation of hexaaquachromium(III) by thiocyanates follows the rate law: –d[complex]/dt=k[NCS] (20–40°C, [NCS]=0.1–0.6M, I=2.0M, pH=1.0). The specific salt effect has been studied for five media: NaCl, NaBr, NaClO4, KCl and CsCl. The series of chloride (Na+, K+ and Cs+) salts show a negligible effect on the anation rate. On the contrary, the series of sodium salts (Cl, Br and ClO 4 ) reveal a marked difference in the reaction rate. The anation rate decreases sharply with the ionic strength increase (I=0.2–2.0M, NaCl). The results were interpreted within the frame of fast equilibria of ion-pair formation followed by an interchange mechanism step. The difference of reaction rate is a result of competition between anions (thiocyanates and supporting electrolyte anions) to the complex cation at an ion-pair formation process.  相似文献   

5.
Raman spectroscopic measurements were performed at ambient temperature onaqueous silica-bearing solutions (0.005 < m Si < 0.02; 0 < pH < 14). The spectraare consistent with the formation of monomeric Si(OH)o 4, SiO(OH) 3 andSiO2(OH)2– 2 species at acid to neutral, basic, and strongly basic pH, respectively.Raman spectra of aqueous Al-bearing solutions at basic pH confirm thepredominance of the Al(OH) 4 species in a wide concentration range (0.01 < m Al < 0.1).Raman spectra of basic solutions (12.4 < pH < 14.3), containing both Al andSi, exhibit a strong decrease in intensities of SiO(OH) 3, SiO2(OH)2– 2, andAl(OH) 4 bands in comparison with Al-free Si-bearing and Si-free Al-bearingsolutions of the same metal concentration and pH, suggesting the formation ofsoluble Al—Si complexes. The amounts of complexed Al and Si derived fromthe measurements of the Al and Si band intensities in strongly basic solutions(pH 14) are consistent with the formation, between Al(OH) 4 andSiO2(OH)2– 2, of the single Al—Si dimer SiAlO3(OH)3– 4 according to the reactionSiO2(OH)2– 2 + Al(OH) 4 SiAlO3(OH)3– 4 + H2OAt lower pH ( 12.5) the changes in band intensities are consistent with theformation of several, likely more polymerized, Al—Si complexes.  相似文献   

6.
The kinetics of the reduction of N2 to N2H4 and NH3 by TiIII-MoIII hydroxide was studied at pH I I and 303-333 K, and the activation energies for these reactions and also for the reaction N2H4 2 NH3 were determined (29, 70, and 25 kJ mol respectively). It was concluded that -90 % of ammonia was formed by the direct reduction of N2 without intermediate formation of hydrazine. A mechanism of this reaction is suggested, which includes the proton insertion into the N-N bond favored by an enhanced electron density at the nitrogen atoms, according to the data of the quantum-mechanical calculation.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1402–1405, June, 1996.  相似文献   

7.
The mechanism and kinetics of the reaction between an aqueous solution of CO2 and coordinate ions, cis-[Cr(C2O4)(L–L)(OH2)2]+, where L–L denotes histamine (hm) or pyridoxamine (pm), were investigated using the stopped-flow technique. The studies were carried out at 5 to 25 °C over the pH range 6.04–8.15 at a fixed ionic strength solution (1 M NaClO4). The results enabled determination of the number of steps of the reactions studied. Based on the kinetic equations, rate constants were determined for each step. Finally, thermodynamic activation parameter values H, were calculated for the reaction studied from temperature relationships.  相似文献   

8.
The kinetics of the mercury(II) catalysed ligand exchange of the hexacyanoferrate(II) complex with the N-methylpyrazinium ion (Mpz+) in a potassium hydrogen phthalate buffer medium has been investigated at 25.0 ± 0.1 °C, pH = 5.0 ± 0.02 and ionic strength, I = 0.1 M (KNO3). The reaction was followed spectrophotometrically in the aqueous medium by measuring the increase in absorbance of the intense blue complex [Fe(CN)5Mpz]2– at its max 655 nm. The effect of pH, and the concentrations of [Fe(CN)6 4–] and Mpz+ on the reaction rate have been studied and analysed. The varying catalytic activity of mercury(II) as a function of concentration has also been explained. The kinetic data suggest that substitution follows an interchange dissociative (I d) mechanism and occurs via formation of a solvent-bound intermediate. The effects of the dielectric constant of the medium on the reaction rates have been used to visualize the formation of a polar activated complex and an interchange dissociative mechanism for the reaction. A mechanism has been proposed in order to interpret the kinetic data. Kinetic evidence is reported for the displacement of CN by Mpz+ in [Fe(CN)6 4–]. Activation parameters for the catalysed and uncatalysed reaction have been evaluated, and lend further support to the proposed mechanism.  相似文献   

9.
The interaction of thymidine, a nucleoside, with hydroxopentaaquarhodium(III), [Rh(H2O)5(OH)]2+ ion in aqueous medium is reported and the possible mode of binding is discussed. The kinetics of interaction between thymidine and [Rh(H2O)5OH]2+ has been studied spectrophotometrically as a function of [Rh(H2O)5OH2+], [thymidine], pH and temperature. The reaction has been monitored at 298 nm, the max of the substituted complex, and where the spectral difference between the reactant and product is a maximum. The reaction rate increases with [thymidine] and reaches a limiting value at a higher ligand concentration. From the experimental findings an associative interchange mechanism for the substitution process is suggested. The activation parameters (H=47.8 ± 5.7 kJ mol–1, S=–173 ± 17 J K–1 mol–1) supports our proposition. The negative G0 (–13.8 kJ mol–1) for the first equilibrium step also supports the spontaneous formation of the outer sphere association complex.  相似文献   

10.
Summary The reactions of monochromatically photolyzed aqueous [W(CN)8]3– solutions in the pH range 1–13 were found to consist of a two-step sequence involving a primary inter-molecular redox process and a consecutive thermal chain process, both leading to the formation of octacyanotungstate(IV) as the main reaction product. The initial quantum yields, 0, are practically constant for the entire pH range and equal to 0.81±0.05 mole/Einstein. The final quantum yields, f, for both neutral and alkaline solution, were found to be greater than 1, having also an approximately constant value of 3.42±0.25. The empirical rate law for the post-irradiation reaction under excess of [W(CN)8]3– and at constant pH is given by: rate = kobs[R]2[W(CN) 8 4– ]–2 where R is a one-electron reductant. Kinetic data suggest a mechanism of the Haber-Weiss type for the thermal reduction of [W(CN)8]3– by H2O2.  相似文献   

11.
Guo  Yanhe  Ge  Qingchun  Lin  Hai  Lin  Huakuan  Zhu  Shourong 《Transition Metal Chemistry》2003,28(6):668-675
The ligands 1,10-N,N-bis(2-hydroxymethylbenzoyl)-1,4,7,10-tetraazadecane (L1) and 1,11-N,N-bis(2-hydroxymethylbenzoyl)-1,4,8,11-tetraazaundecane (L2) have been synthesized. The stability constants of NiII complexes of ligands L1 and L2 have been studied at 25 °C using pH titrations. The kinetics of general acid (HCl, 0.04–2.34 mol dm–3) or buffer (DEPP or DESPEN, 0.05 mol dm–3, pH 4.83–5.72)-catalyzed dissociation of these NiII complexes have been investigated at 25 °C using a stopped-flow spectrophotometer. The ionic strength of solution was controlled at I = 2.34 mol dm–3 (KCl + HCl) and I = 0.1 mol dm–3 (KNO3, buffer), respectively. The kinetic dissociation of NiII complexes catalyzed by HCl obeys the equilibrium k obs = k 1d + k 2H[H+], whereas in buffer solution the observed rate constant k obs = k d + k 1H[H+]. At pH < 1.5, both the proton-assisted and direct protonation pathways contribute to the rates, whereas solvation is the dominant pathway at pH > 6. In the 4.8–5.7 pH range, the complexes dissociate mainly through a proton-assisted pathway.  相似文献   

12.
The effects of anti-symmetric electrolytes (CaCl2, Ca(SCN)2, MgCl2, and/or Mg(SCN)2) and pH on the phase behavior, the -potential, the hydrodynamic diameter and the surface charge density of an oil-inwater type (O/W-type) microemulsion formed in solutions of an amphoteric surfactant (N ,N -dimethyl-N -lauroyllysine, DMLL)/n-octane/1-pentanol/brine have been examined. The formation of the microemulsion in the presence of CaCl2 and/or Ca(SCN)2 is of Winsor-type with an increase in the concentration of 1-pentanol. Particularly, microemulsion is not formed by the addition of Ca(SCN)2 in a pH region less than 2.6. The -potential and the surface charge density of the microemulsion in the presence of CaCl2 decrease with an increase in pH and show slightly positive values in the isoelectric region (pH 5-7), while, in the presence of Ca(SCN)2, the -potential and the surface charge density show negative values in the same region at which the net charge of DMLL molecules becomes almost zero. The hydrodynamic diameters in the presence of CaCl2 show a maximum value around pH 2.5, whereas, in the presence of Ca(SCN)2, the minimum value is around pH 5.5. Similar tendencies are recognized in results for the -potential, the hydrodynamic diameter and the surface charge density of the O/W-type microemulsion in the presence of MgCl2 and Mg(SCN)2. A new formula to estimate the binding constants (K) of Ca2+, Mg2+, Cl, and SCN to the hydrophilic groups in DMLL molecules and the adsorption density of DMLL molecules on the oil/water interface (N) in the presence of antisymmetric electrolytes has been derived.K for Ca2+, Mg2+, Cl, and SCN was found to beK Ca=0.12M–1,K Mg=0.14 M–1,K Cl=0.0084±0.0016 M–1, respectively.N for DMLL molecules in the presence of CaCl2, Ca(SCN)2, MgCl2 and/or Mg(SCN)2 was found to be 0.50 nm–2, 0.38 nm–2, 0.44 nm–2, and 0.47 nm–2, respectively; and the surfactant (DMLL) numbers per O/W-type microemulsion droplet change from a few hundreds to a few thousands with changing pH. The larger the hydrodynamic diameter of the O/W-type microemulsion, the greater the number of DMLL molecules adsorbed on the O/W-type microemulsion surfaces.  相似文献   

13.
Apparent equilibrium constants and calorimetric enthalpies of reaction have been measured for the reaction L-tryptophan(aq) + H2O(l) = indole(aq) + pyruvate(aq) + ammonia(aq) which is catalyzed by L-tryptophanase. High-pressure liquid-chromatography and microcalorimetery were used to perform these measurements. The equilibrium measurements were performed as a function of pH, temperature, and ionic strength. The results have been interpreted with a chemical equilibrium model to obtain thermodynamic quantities for the reference reaction: L-tryptophan(aq) + H2O(l) = indole(aq) + pyruvate(aq) + NH 4 + (aq). At T=25°C and Im=O the results for this reaction are: Ko=(1.05±0.13)×10–4, G°=(22.71±0.33) kJ-mol–1, H°=(62.0±2.3) kJ-mol–1, and S°=(132±8) J-K–1-mol–1. These results have been used together with thermodynamic results from the literature to calculate standard Gibbs energies of formation, standard enthalpies of formation, standard molar entropies, standard molar heat capacities, and standard transformed formation properties for the substances participating in this reaction.Presented at the Symposium, 76th CSC Congress, Sherbrooke, Quebec, May 30–June 3, 1993, honoring Professor Donald Patterson on the occasion of his 65th birthday.  相似文献   

14.
Summary The kinetics of formation and dissociation of the binuclear complex of CoII with histidinato(pentaammine)CoIII have been studied at 10.0°Ct°C25°C and I = 0.3 mol dm–3 (ClO inf4 sup– ). The formation of the binuclear complex, [(NH3)5CoIIILCoII]4+ (L = histidinate), in the 5.7–6.8 pH range involves the reaction of Co(OH2) inf6 sup2+ with the deprotonated, (NH3)5CoL2+, and monoprotonated, (NH3)5CoLH3+, forms of the complex. The rate and activation parameters for the formation are consistent with an I d mechanism. The binuclear species undergoes dissociation to yield the parent CoIII substrate and Co(OH2) inf6 sup2+ via spontaneous and acid-catalysed paths. Comparison of spontaneous dissociation rate of the binuclear complex with other related systems indicated the chelate nature of the binuclear species.  相似文献   

15.
The transient decay kinetics of electrons generated in thin cadmium sulfide films by short laser pulses was studied by the microwave photoconductivity method (9 and 36 GHz) at 295 K. The films were prepared by the pulverization method from thiocarbamide coordination compounds. At the high light intensity I 0 > 1014 photon cm–2 per pulse, the decay kinetics of photoelectrons corresponded to a reaction of the second order. Analysis of the kinetic data made it possible to determine the rate constant of recombination of free electrons and holes: k 3 2(±1)·10–13 cm3 s–1.  相似文献   

16.
Summary The kinetics and mechanism of the system [FeHIDA-(OH)2]+5CN[Fe(CN)5OH+HIDA2–+OH (HIDA=N-(2-hydroxyethyl) (iminodiacetate) at pH=9.5±0.02, I=0.1 M and at 25±0.1°C have been studied spectrophotometrically at 395 nm ( max of [Fe(CN)5OH]3–]. The reaction has three distinguishable stages; the first is formation of [Fe(CN)5OH]3–, the second is conversion of [Fe(CN)5OH]3– into [Fe(CN)6]3–, and last is the reduction of [Fe(CN)6]3– to [Fe(CN)6]4– by the HIDA2– released in the first stage. The first stage shows variable-order dependence on cyanide concentration, unity at high cyanide concentration and zero at low cyanide concentration. The second stage exhibits first-order dependence on the concentration of [Fe(CN)5OH]3– as well as on cyanide. The reverse reaction between [Fe(CN)5OH]3– and HIDA2– is first-order in each of these species and inverse first-order in cyanide. On the basis of forward and reverse rate studies, a five-step mechanism has been proposed for the first stage. The first step involves a slow loss of one OH, by a cyanide-independent path.  相似文献   

17.
The title reaction has been studied spectrophotometrically in aqueous medium as a function of [substrate complex], [ligand], pH and temperature at constant ionic strength. At the physiological pH (7.4) the interaction with azide shows two distinct consecutive steps, i.e., it shows a non-linear dependence on the concentration of N3 ; both processes are [ligand]-dependent. The rate constant for the processes are: k 110–3 s–1 and k 210–5 s–1. The activation parameters calculated from Eyring plots are: H 1 = 14.8 ± 1 kJ mol–1, S 1 = –240 ± 3 J K–1 mol–1, H 2 = 44.0 ± 1.5 kJ mol–1 and S 2 = –190 ± 4 J K–1 mol–1. Based on the kinetic and activation parameters an associative interchange mechanism is proposed for the interaction process. From the temperature dependence of the outersphere association equilibrium constant, the thermodynamic parameters calculated are: H 1 0 = 4.4 ± 0.9 kJ mol–1, S 1 0 = 64 ± 3 J K–1 mol–1 and H 2 0 = 14.2 ± 2.9 kJ mol–1, S 2 0 = 90 ± 9 J K–1 mol–1, which gives a negative G 0 value at all temperatures studied, supporting the spontaneous formation of an outersphere association complex.  相似文献   

18.
Summary The title reaction has been followed spectrophotometrically at 325 nm (max of [Mn(CN)6]3–) under pseudo-first order conditions with cyanide in a large excess at pH=10.0, I=0.1M (NaClO4) and 25°C. The reaction follows first-order kinetics in [MnEDTA(OH)]2– and exhibits variable-order dependence in [CN] one at high cyanide concentration, and two at low cyanide concentration. The product of above reaction has been identified as [Mn(CN)6]3–.The kinetics of the reverse reaction,i.e., the reaction of [Mn(CN)6]3- with EDTA4– have also been followed spectrophotometrically. This reactions is first-order with respect to both [Mn(CN) 6 3– ] and [EDTA4–] and exhibits an inverse first-order dependence on [CN]. A six-step mechanism has been proposed in which the penultimate step is rate-determining. The activation parameters have been obtained and support the postulated mechanism.  相似文献   

19.
The kinetics of base hydrolysis of the trans-[Cr(NH3)2(NCS)4] anion follows the rate law: -d[complex]/dt = k 0 + k 1[OH] (50–70 °C, [OH] = 0.1–1.9 M and = 2.0 M). The specific salt effect has been investigated for eight aqueous media: NaCl, NaBr, NaI, NaClO4, KCl, KBr, CsCl and CsBr. The alkali-independent path (k 0) does not show any specific effect of inert electrolyte ions, the activation parameters: H = 113.5 ± 0.4 kJ mol–1 and S = 24.1 ± 1.3 J mol–1 K–1 are interpreted in the frame of a dissociative interchange mechanism (I d). For the alkali-dependent path (k 1) the specific salt effect is observed for cations of the inert electrolyte, showing an important role for ion-pair formation between the cations and reagent complex anion in the activation process. A linear correlation between lnk 1 and lnK 0 (K 0 – ion-pair formation constant) has been found for the cations studied. The dissociative, via conjugate base, mechanism (D CB) has been proposed for the alkali-dependent path.  相似文献   

20.
Summary The kinetics of anation of the hydroxopentaaqua-rhodium(III) ion by pyridine-2-aldoxime have been studied spectrophotometrically as a function of pH, substrate concentration, entering ligand, temperature and solvent polarity. The reaction rate increases with an increase in pH, temperature and ligand concentration, and reaches a limiting value at high ligand concentration. The activation parameters (H # = 87.5 kJ mol–1 and S = –52.3 JK–1 mol–1) have been calculated from an Eyring plot and are compared with aqua exchanges in other systems. All the experimental facts lead to an associative interchange (I a) mechanism with the prior formation of an outer-sphere association complex. Solvent polarity effects lead to the same conclusion.  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