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 93 毫秒
1.
The kinetics of the process [Pt(SNS)(R-py)]2+ + Cl → [Pt(SNS)Cl]+ + R-py {SNS = 2,6-bis(methylsulfanylmethyl)pyridine; R-py = meta- or para-substituted pyridines covering a wide range of basicity} were studied in methanol at 25 °C. The reactions obey the usual two-term rate law observed in the substitution reactions of square-planar d8 complexes. The plots of log k2 {k2 = second-order rate constants} against the pKa of the heterocycles conjugate acids highlighted a different sensitivity of the two groups of N-donors to changes in basicity, thepara-substituted pyridines (4R-py) showing a weaker dependence on pKa than the meta-substituted (3R-py). The results have been explained on the basis of a π-acidity difference between 3R-py and 4R-py which influences the reaction ground state.  相似文献   

2.
The kinetics of oxidative addition of CH3I to [Rh(FcCOCHCOCF3)(CO)(PPh3)], where Fc = ferrocenyl and (FcCOCHCOCF3) = fctfa = ferrocenoylacetonato, have been studied utilizing UV/Vis, IR, 1H and 31P NMR techniques. Three definite sets of reactions involving isomers of at least two distinctly different classes of RhIII-alkyl and two different classes of RhIII-acyl species were observed. Rate constants for this reaction in CHCl3 at 25 °C, applicable to the reaction sequence below, were determined as k1 = 0.00611(1) dm3 mol−1 s−1, k−1 = 0.0005(1) s−1, k3 = 0.00017(2) s−1 and k4 = 0.0000044(1) s−1 while k−3 ? k3 and k−4 ? k4 but both ≠0. The indeterminable equilibrium K2 was fast enough to be maintained during RhI depletion in the first set of reactions and during the RhIIIalkyl2 formation in the second set of reactions. From a 1H and 31P NMR study in CDCl3, Kc1 was found to be 0.68, Kc2 = 2.57, Kc3 = 1.00, Kc4 = 4.56 and Kc5 = 1.65.  相似文献   

3.
The rate constants for the reactions of OH radicals with CF3OCHFCF3, and CF3CHFCF3 have been measured over the temperature range 250-430 K. Kinetic measurements have been carried out using the flash photolysis, and laser photolysis methods combined, respectively, with the laser induced fluorescence technique. The influence of impurities in the samples has been investigated by using gas chromatography. No sizable effect of impurities was found on the measured rate constants of these fluorinated compounds, if the purified samples were used in the measurements. The following Arrhenius expressions were determined: k(CF3OCHFCF3) = (4.39 ± 1.38) × 10−13 exp[−(1780 ± 100)/T] cm3 molecule−1 s−1, and k(CF3CHFCF3) = (6.19 ± 2.07) × 10−13 exp[−(1830 ± 100)/T] cm3 molecule−1 s−1.  相似文献   

4.
Chemometric analysis of ultraviolet-visible (UV-vis) spectra for pH values 1.0, 3.3, 5.3, and 6.9 was used to investigate the kinetics and the structural transformations of anthocyanins in extracts of calyces of hibiscus flowers of the Hibiscus acetosella Welw. ex Finicius for the first time. Six different species were detected: the quinoidal base (A), the flavylium cation (AH+), the pseudobase or carbinol pseudobase (B), cis-chalcone (CC), trans-chalcone (Ct), and ionized cis-chalcone (CC). Four equilibrium constant values were calculated using relative concentrations, hydration, pKh = 2.60 ± 0.01, tautomeric, KT = 0.14 ± 0.01, acid-base, pKa = 4.24 ± 0.04, and ionization of the cis-chalcone, pKCC=8.74±1.5×10−2. The calculated protonation rate of the tautomers is KH+=0.08±7.6×10−3. These constants are in excellent agreement with those measured previously in salt form. From a kinetic viewpoint, the situation encountered is interesting since the reported investigation is limited to visible light absorption in acid medium. These models have not been reported in the literature.  相似文献   

5.
The oxidation of a series of substituted pyridines by dimethyldioxirane (1) produced the expected N-oxides in quantitative yields. The second order rate constants (k2) for the oxidation of a series of substituted pyridines (2a-g) by dimethyldioxirane were determined in dried acetone at 23 °C. An excellent correlation with Hammett sigma values was found (ρ = −2.91, r = 0.995). Kinetic studies for the oxidation of 4-trifluoromethylpyridine by 1 were carried out in the following dried solvent systems: acetone (k2 = 0.017 M−1 s−1), carbon tetrachloride/acetone (7:3; k2 = 0.014 M−1 s−1), acetonitrile/acetone (7:3; k2 = 0.047 M−1 s−1), and methanol/acetone (7:3; k2 = 0.68 M−1 s−1). Kinetic studies of the oxidation of pyridine by 1 versus mole fraction of water in acetone [k2 = 0.78 M−1 s−1 (χ = 0) to k2 = 11.1 M−1 s−1 (χ = 0.52)] were carried out. The results showed the reaction to be very sensitive to protic, polar solvents.  相似文献   

6.
A pyrimethanil-imprinted polymer (P1) was prepared by iniferter-mediated photografting a mixture of methacrylic acid and ethylene dimethacrylate onto homemade near-monodispersed chloromethylated polydivinylbenzene beads. The chromatographic behaviour of a column packed with these imprinted beads was compared with another column packed with irregular particles obtained by grinding a bulk pyrimethanil-imprinted polymer (P2). The comparison was made using the kinetic model of non-linear chromatography, studying the elution of the template and of two related substances, cyprodinil and mepanipyrim. Extension of the region of linearity, capacity factors for the template and the related substances, column selectivity, binding site heterogeneity, apparent affinity constant (K) and lumped kinetic association (ka) and dissociation rate constant (kd) were studied during a large interval of solute concentration, ranging between 1 and 2000 μg/ml. From the experimental results obtained, in the linearity region of solute concentration column selectivity and binding site heterogeneity remained essentially the same for the two columns, while column capacity (at 20 μg/ml, P1 = 23.1, P2 = 11.5), K (at 20 μg/ml, P1 = 8.3 × 106 M−1, P2 = 2.5 × 106 M−1) and ka (at 20 μg/ml, P1 = 3.5 μM−1 s−1, P2 = 0.47 μM−1 s−1) significantly increased and kd (at 20 μg/ml, P1 = 0.42 s−1, P2 = 0.67 s−1) decreased for the column packed with the imprinted beads. These results are consistent with an influence of the polymerisation method on the morphology of the resulting polymer and not on the molecular recognition properties due to the molecular imprinting process.  相似文献   

7.
The fluorophilicity of a series of hydrocarbon and fluorocarbon-functionalized nicotinic acid esters (nicotinates) is measured from their partitioning behavior (log Kp) in the biphasic solvent system of perfluoro(methylcyclohexane) (PFMC) and toluene. The chain length of the hydrocarbon or fluorocarbon alkyl group of the ester ranges from one to twelve carbon atoms. Knowledge of the fluorophilicity of these solutes is relevant to the design of these prodrugs for fluorocarbon-based drug delivery. The experimental log Kp values range from −1.72 to −3.40 for the hydrocarbon nicotinates and −1.64 to 0.13 for the fluorinated nicotinates, where only the prodrug with the longest fluorinated chain (2,2,3,3,4,4,5,5,6,6,7,7,8,8,8-pentadecafluorooctyl nicotinic acid ester) partitions preferentially into the fluorinated phase (log Kp = 0.13). Predictions of the partition coefficients using solubility parameters calculated from group contribution techniques or molecular dynamics simulation are in reasonable agreement for the perhydrocarbon nicotinates and short chained perfluorinated nicotinates (≈0.3-39% deviation). Significant deviations from experimental partition coefficients (greater than 100%) are observed for the longest chain perfluoroalkyl nicotinates.  相似文献   

8.
The hydrogen peroxide-oxidation of o-phenylenediamine (OPD) catalyzed by horseradish peroxidase (HRP) at 37 °C in 50 mM phosphate buffer (pH 7.0) was studied by calorimetry. The apparent molar reaction enthalpy with respect to OPD and hydrogen peroxide were −447 ± 8 kJ mol−1 and −298 ± 9 kJ mol−1, respectively. Oxidation of OPD by H2O2 catalyzed by HRP (1.25 nM) at pH 7.0 and 37 °C follows a ping-pong mechanism. The maximum rate Vmax (0.91 ± 0.05 μM s−1), Michaelis constant for OPD Km,S (51 ± 3 μM), Michaelis constant for hydrogen peroxide Km,H2O2 (136 ± 8 μM), the catalytic constant kcat (364 ± 18 s−1) and the second-order rate constants k+1 = (2.7 ± 0.3) × 106 M−1 s−1 and k+5 = (7.1 ± 0.8) × 106 M−1 s−1 were obtained by the initial rate method.  相似文献   

9.
It is shown by the B3LYP/6-311+G(2d,p)//B3LYP/6-31G(d) calculations that the hexacyano derivative of aza-acepentalene is an extremely powerful superacid both in the gas phase and in DMSO as evidenced by the ΔHacid = 255.1  kcal mol−1 and pKa (DMSO) = −26.5. Its synthesis is strongly recommended, in particular, since the related conjugate base hexachloro aza-acepentalenide anion was prepared recently.  相似文献   

10.
β-Cyclodextrin formed the most robust complexes with o-carboranols 1b and 1c in aqueous solution, and the association constants estimated from NMR titration studies indicated Ka >1 × 106 M−1 and Ka = 6 × 105 M−1, respectively.  相似文献   

11.
12.
Hydroboration reactions of 1-octene and 1-hexyne with H2BBr·SMe2 in CH2Cl2 were studied as a function of concentration and temperature, using 11B NMR spectroscopy. The reactions exhibited saturation kinetics. The rate of dissociation of dimethyl sulfide from boron at 25 °C was found to be (7.36 ± 0.59 and 7.32 ± 0.90) × 10−3 s−1 for 1-octene and 1-hexyne, respectively. The second order rate constants, k2, for hydroboration worked out to be 7.00 ± 0.81 M s−1 and 7.03 ± 0.70 M s−1, while the overall composite second order rate constants, k K, were (3.30 ± 0.43 and 3.10 ± 0.37) × 10−2 M s−1, respectively at 25 °C. The entropy and enthalpy values were found to be large and positive for k1, whilst for k2 these were large and negative, with small values for enthalpies. This is indicative of a limiting dissociative (D) for the dissociation of Me2S and associative mechanism (A) for the hydroboration process. The overall activation parameters, ΔH and ΔS, were found to be 98 ± 2 kJ mol−1 and +56 ± 7 J K−1 mol−1 for 1-octene whilst, in the case of 1-hexyne these were found out to be 117 ± 7 kJ mol−1 and +119 ± 24 J K−1 mol−1, respectively. When comparing the kinetic data between H2BBr·SMe2 and HBBr2·SMe2, the results showed that the rate of dissociation of Me2S from H2BBr·SMe2 is on average 34 times faster than it is in the case of HBBr2·SMe2. Similarly, the rate of hydroboration with H2BBr·SMe2 was found to be on average 11 times faster than it is with HBBr2·SMe2. It is also clear that by replacing a hydrogen substituent with a bromine atom in the case of H2BBr·SMe2 the mechanism for the overall process changes from limiting dissociative (D) to interchange associative (Ia).  相似文献   

13.
The diffusion-controlled rate constants, kd, of various quenching reactions, [Ru(L)3]2+∗ (L = bpy, phen and 4,7-(CH3)2phen) + [Fe(CN)6]3−, were measured through fluorescence measurements. From them, the effective values of viscosity coefficients for several methanol + water mixtures were calculated. These coefficients were checked through calculations of the rate constants of the reaction [IrCl6]2− + [Ru(bpy)3]2+∗, which were also obtained by fluorescence quenching measurements. The agreement between the two sets of data (experimental and predicted) is excellent. Besides, the trends of association, kd, and dissociation, kd, rate constants for 2+/3−, 2+/2− and 2+/2+ reactions in methanol-water mixtures are discussed. The use of effective diffusion coefficients for estimating kd and kd allowed us to obtain the intrinsic electron transfer rate constant, ket, for the activation-diffusion-controlled process between [Ru(bpy)3]2+∗ and [Co(NH3)5Cl]2+ complexes from the observed (quenching) rate constant. The influence of methanol-water mixtures on ket was rationalized by using the Marcus electron-transfer treatment.  相似文献   

14.
This paper discusses the kinetic simulation of TiCl4--coinitiated living carbocationic isobutylene (IB) polymerizations governed by dormant-active equilibria, using a mechanistic model. Two kinetic models were constructed from the same underlying mechanism: one using a commercial simulation software package (Predici®), and the other using the method of moments. Parameter estimation from experimental batch reactor data with Predici yielded a rate constant of propagation kp = 4.64 × 108 ± 2.75 × 108 L/mol s, with no constraints imposed. This agrees with kp data measured with diffusion clock and competition methods, but disagrees with kinetically obtained kp values. Estimation of rate constants with Predici® and the GREG parameter estimation software packages revealed that it was difficult to estimate the complete set of kinetic parameters, due to correlated effects of the parameters on model predictions. Estimability analysis confirmed that some of the strongly correlating parameters could not be estimated simultaneously using the available experimental data. Using kp = 6 × 108 ± 2.75  × 108 L/mol s measured by Mayr, and using starting estimates of other rate constants defined by experimentally observed correlations, yielded the set of rate constants required for the simulations. Both kinetic models yielded good agreement with experimental data, with the exception of Mw values that slightly diverged from the theoretically predicted ‘MwMn = constant’ relationship. This may indicate the occurrence of a minor side reaction. However, the kp/k−1 = 17.5 L/mol average run length calculated from measured and simulated MWD data agrees well with earlier literature values.  相似文献   

15.
In this article, we present a systematic study on IgG and Fab fragment of anti-IgG molecules using fluorescence auto- and cross-correlation spectroscopy to investigate their diffusion characteristics, binding kinetics, and the effect of small organic molecule, urea on their binding. Through our analysis, we found that the diffusion coefficient for IgG and Fab fragment of anti-IgG molecules were 37 ± 2 μm2 s−1 and 56 ± 2 μm2 s−1, respectively. From the binding kinetics study, the respective forward (ka) and backward (kd) reaction rates were (5.25 ± 0.25) × 106 M−1 s−1 and 0.08 ± 0.005 s−1, respectively and the corresponding dissociation binding constant (KD) was 15 ± 2 nM. We also found that urea inhibits the binding of these molecules at 4 M concentration due to denaturation.  相似文献   

16.
The inhibitory effects of five hydroxyanthraquinones (HAQs) from root and rhizoma of Rheum officinale Baill, a traditional Chinese medicinal (TCM) herb, on Staphylococcus aureus growth were investigated by calorimetry. The power-time curves of S.aureus with and without HAQ were acquired and the extent and duration of inhibitory effects on the metabolism evaluated by growth rate constants (k1, k2), half inhibitory ratio (IC50), maximum heat output (Pmax) and peak time (tp). The value of k1 and k2 of S. aureus in the presence of the five HAQs decreased with the increasing concentrations of HAQs. Moreover, Pmax was reduced and the value of tp increased with increasing concentrations of the five drugs. The inhibitory activity varied for different drugs. IC50 of the five HAQs was 4 μg ml−1 for emodin, 3.5 μg ml−1 for rhein, 10 μg ml−1 for aloe-emodin, 1000 μg ml−1 for chrysophanol, 1600 μg ml−1 for physcion. The sequence of antimicrobial activity of the five HAQs: rhein > emodin > aloe-emodin > chrysophanol > physicion.  相似文献   

17.
Heat capacity and enthalpy increments of ternary bismuth tantalum oxides Bi4Ta2O11, Bi7Ta3O18 and Bi3TaO7 were measured by the relaxation time method (2-280 K), DSC (265-353 K) and drop calorimetry (622-1322 K). Temperature dependencies of the molar heat capacity in the form Cpm=445.8+0.005451T−7.489×106/T2 J K−1 mol−1, Cpm=699.0+0.05276T−9.956×106/T2 J K−1 mol−1 and Cpm=251.6+0.06705T−3.237×106/T2 J K−1 mol−1 for Bi3TaO7, Bi4Ta2O11 and for Bi7Ta3O18, respectively, were derived by the least-squares method from the experimental data. The molar entropies at 298.15 K, S°m(298.15 K)=449.6±2.3 J K−1 mol−1 for Bi4Ta2O11, S°m(298.15 K)=743.0±3.8 J K−1 mol−1 for Bi7Ta3O18 and S°m(298.15 K)=304.3±1.6 J K−1 mol−1 for Bi3TaO7, were evaluated from the low-temperature heat capacity measurements.  相似文献   

18.
The determination of pKa value for the unstable chromium(VI) peroxide, CrO(O2)2(H2O) in aqueous solution is presented. The pKa value is found to be (1.55 ± 0.03). The kinetic decomposition of chromium(VI) peroxide is dependent on the concentration of hydrogen peroxide in the pH range between 2.5 and 4.0. We have proposed the possible explanation for the formation of triperoxo chromium complex of hydrogen peroxide which is dependent on decomposition. Activation of coordinate peroxide in chromium(VI) peroxide observed in the kinetic studies is by reduction of thiolato-cobalt(III) complex. The rate constant (M−1 s−1, 15 °C) for the oxygen atom transfer reaction from CrO(O2)2(OH) to (en)2Co(SCH2CH2NH2)2+ is found to be (25.0 ± 1.3).  相似文献   

19.
Iron and ruthenium classical and non-classical hydrides of the type [MH(N–N)P3]+ and [M(η2-H2)(N–N)P3]2+ {M = Fe, Ru; N–N = 2,2′-bipyridine (bpy), 1,10-phenanthroline (phen); P = phosphites} were reported in 2004 together with an evaluation of the pseudo-aqueous pKa values of the η2-H2 complexes. The non-classical hydrides, even if doubly charged, showed a relatively low acidity, their pKa values ranging between −5.4 and −4.3. Moreover, ruthenium(II) derivatives showed to be more acidic than the corresponding iron(II) complexes. Information about the structural and electronic proprieties of complexes of this type, which allowed to better understand the role of both the metal centres and the ancillary ligands in the acidity of the co-ordinated hydrogen molecule, was obtained on the basis of DFT B3LYP calculations.  相似文献   

20.
Carina M.M. Machado 《Talanta》2007,71(3):1352-1363
This work describes the application of polarography, a technique scarcely used for modelling and optimisation of stability constants, in the study of copper complexes with [(2-hydroxy-1,1-bis(hydroxymethyl)ethyl)amino]-1-propanesulfonic acid (TAPS). Direct current polarography (DCP), using low total copper ion and large total ligand to total copper concentration, enabled the full characterization of Cu-(TAPS)x-(OH)y system, whose complexation occurs in the pH range of copper hydrolysis and Cu(OH)2 precipitation. Cu-(TAPS)x-(OH)y system was studied by DCP and glass electrode potentiometry (GEP) in aqueous solution at fixed total ligand to total metal concentrations ratios and varied pH values (25.0 °C; I = 0.1 M, KNO3). The predicted model, as well as the overall stability constants values, are (as log β): CuL+ = 4.2, CuL2 = 7.8, CuL2(OH) = 13.9 and CuL2(OH)22− = 18.94. GEP only allowed confirming the stability constants for CuL+ and CuL2 and was used to determine the pKa of TAPS, 8.342.Finally, a briefly comparative analysis between TAPS and other structural related buffers was done. Evaluation based on log βCuL versus pKa revealed that TES, TRIS, TAPS and AMPSO coordinated via amino and hydroxymethylgroups forming a five-membered chelate ring. For BIS-TRIS and TAPSO, and possibly DIPSO, one or more five-membered chelate rings involving additional hydroxyl groups are also likely formed.  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