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 reactions of acetylacetonato cobalt (III) ion in sodium hydroxide solutions have been studied spectrophotometrically over a range of temperatures and hydroxide ion concentrations. The activation enthalpy, ΔH was 70.6 kJ mol?1 and the activation entropy, ΔS was ? 119 JK?1mol?1, with a rate law of kobs = k2 [OH?]2. A mechanism involving initial de-chelation of the acetylacetone ligand is suggested. The rate of exchange of methyl hydrogen of the acetylacetone ligand was studied, using proton nuclear magnetic resonance. The rate law was kobs = k [OH?]. Initial de-chelation is also suggested as a mechanism for this process. The 13C nuclear magnetic resonance spectrum of the complex is reported.  相似文献   

2.
The rates of formation of penta-ammineglycinecobalt(III) ion from aquopenta-amminecobalt(III) ion and glycine in acidic media have been studied spectrophotometrically at different glycine concentration and different pH in the range of 50–70°C. The ΔH≠ and ΔSz≠ values are 27.6 kcal mole?1 and +5.2 e. u. respectively, and increase in ionic strength causes only a slight acceleration of the rate. The results are consistent with a mechanism involving outer-sphere association between the aquopenta-amminecobalt(III) complex and glycine, followed by its transformation into the product by an essentially dissociative process in which rupture of the Co(III)? OH2 bond is primarily important in the transition state (SN1IP mechanism).  相似文献   

3.
The kinetics of the interactions between three sulfur‐containing ligands, thioglycolic acid, 2‐thiouracil, glutathione, and the title complex, have been studied spectrophotometrically in aqueous medium as a function of the concentrations of the ligands, temperature, and pH at constant ionic strength. The reactions follow a two‐step process in which the first step is ligand‐dependent and the second step is ligand‐independent chelation. Rate constants (k1 ~10?3 s?1 and k2 ~10?5 s?1) and activation parameters (for thioglycolic acid: ΔH1 = 22.4 ± 3.0 kJ mol?1, ΔS1 = ?220 ± 11 J K?1 mol?1, ΔH2 = 38.5 ± 1.3 kJ mol?1, ΔS2 = ?204 ± 4 J K?1 mol?1; for 2‐thiouracil: ΔH1 = 42.2 ± 2.0 kJ mol?1, ΔS1 = ?169 ± 6 J K?1 mol?1, ΔH2 = 66.1 ± 0.5 kJ mol?1, ΔS2 = ?124 ± 2 J K?1 mol?1; for glutathione: ΔH1 = 47.2 ± 1.7 kJ mol?1, ΔS1 = ?155 ± 5 J K?1mol?1, ΔH2 = 73.5 ± 1.1 kJ mol?1, ΔS2 = ?105 ± 3 J K?1 mol?1) were calculated. Based on the kinetic and activation parameters, an associative interchange mechanism is proposed for the interaction processes. The products of the reactions have been characterized from IR and ESI mass spectroscopic analysis. A rate law involving the outer sphere association complex formation has been established as   相似文献   

4.
The rate of formation of Cr(III)-EDTA complexes from Cr(C2O4) and EDTA at pH ca. 8–9.5 has been determined spectrophotometrically. From the observed dependence of the rate on pH and EDTA concentration a probable mechanism has been suggested for the overall change involving two concurrent paths; corresponding δH and δS values have also been determined. An increase in ionic strength has been found to increase the overall rate as well as that of each individual path as is expected in a reaction between ions of the same charge type.  相似文献   

5.
The kinetics of oxidation of benzhydrol and its p-substituted derivatives (YBH, where Y=H, Cl, Br, NO2, CH3, and OCH3) by sodium N-chloro-p-toluenesulfonamide or chloramine-T (CAT), catalyzed by ruthenium(III) chloride, in the presence of hydrochloric acid in 30% (v/v) MeOH medium has been studied at 35°C. The reaction rate shows a first-order dependence on [CAT]O and a fractional-order each on [ YBH]O, [Ru(III)], and [H+]. The reaction also has a negative fractional-order (−0.35) behavior in the reduction product of CAT, p-toluenesulfonamide (PTS). The increase in MeOH content of the solvent medium retards the rate. The variation of ionic strength of the medium has negligible effect on the rate. Rate studies in D2O medium show that the solvent isotope effect, k′H2O/k′D2O, is equal to 0.60. Proton inventory studies have been made in H2O(SINGLEBOND)D2O mixtures. The rates correlate satisfactorily with Hammett σ relationship. The LFE relationship plot is biphasic and the reaction constant ρ=−2.3 for electron donating groups and ρ=−0.32 for electron withdrawing groups at 35°C. Activation parameters ΔH, ΔS, and ΔG have been calculated. The parameters, ΔH and ΔS, are linearly related with an isokinetic temperature β=334 K indicating enthalpy as a controlling factor. A mechanism consistent with the observed kinetics has been proposed. © 1997 John Wiley & Sons, Inc.  相似文献   

6.
The kinetics of formation of [Cr(NH3)5N3]2+ in the reaction of [Cr(NH3)5(OH2)3+] in NaN3–HN3 buffer media (pH ca. 3.4–4.3) has been studied spectrophotometrically under different conditions for elucidation of reaction mechanism. The reaction occurs in two concurrent paths involving reactions of the aquo complex and its conjugate base (the hydroxo complex) with azide ion (gross rate constants k1 and k2 respectively). Results, including values of the activation parameters (ΔH≠ and ΔS≠) are in agreement with SN1 IP mechanism for both the paths. The observed relativ nucleophilicityo of N3? compared to that reported in literature for Cl? and NCS? is N > NCS? > Cl?.  相似文献   

7.
The influence of placing thioether linkages trans to a site of nitrito substitution and spontaneous nitrito-tonitro isomerization is reported for the [CoQS(H2O)]3+ cation where QS is 1,11-diamino-3,6,9-trithiaundecane. Preparation and characterization is described for the aqua and nitrito complexes. Rate data for the substitution process is presented at 17.7, 25.0 and 35.0°C. It is consistent with the mechanism first proposed by Basolo and Pearson in which N2O3 is the nitrosation agent. [CoQS(H2O)]3+ is three hundred times more reactive than [Co(NH3)5H2O]3+ under identical conditions. Isomerization is dramatically slower than the conversion of [CoQS(H2O)3+ to [CoQS(ONO)]2+. The isomerization process was studied at 5 wavelengths, 3 temperatures and various conditions of acid and nitrite ion at an ionic strength of 0.11–0.60 M. Studies at 25°C give kisom = 1.21 ± 0.12 × 10?4 sec ?1. Similar determinations at 17.7 and 35.0°C give kisom = 3.84 ± 0.65 × 10?5 sec?1 and 3.59 ± 0.13 × 10?4 sec?1 respectively. The thermodynamic activation parameters ΔH, ΔG, and ΔS obtained from an Eyring plot gives ΔH = 111.3 kJ/mol, ΔS = + 53 J/molK and ΔG = 95.4 kJ/mol. These results are discussed in the context of present knowledge and experience with other cobalt(III) ligand systems.  相似文献   

8.
The rate constants for the reaction of 2,6‐bis(trifluoromethanesulfonyl)‐4‐nitroanisole with some substituted anilines have been measured by spectrophotometric methods in methanol at various temperatures. The data are consistent with the SNAr mechanism. The effect of substituents on the rate of reaction has been examined. Good linear relationships were obtained from the plots of log k1 against Hammett σpara constants values at all temperature with negative ρ values (?1.68 to ?1.11). Activation parameters ΔH varied from 41.6 to 54.3 kJ mol?1 and ΔS from ?142.7 to ?114.6 J mol?1 K?1. The δΔH and δΔS reaction constants were determined from the dependence of ΔH and ΔS activation parameters on the σ substituent constants, by analogy with the Hammett equation. A plot of ΔH versus ΔS for the reaction gave good straight line with 177°C isokinetic temperature. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 203–210, 2010  相似文献   

9.
The kinetics of the oxidation of ketorolac by hexacyanoferrate(III) (HCF) in aqueous alkaline medium at a constant ionic strength of 0.75 mol·dm?3 was studied spectrophotometrically at 300 K. A plausible mechanism was proposed and the rate law was derived. The mechanism of oxidation of ketorolac (KET) in alkaline medium has been shown to proceed via a KET-HCF complex, which decomposes in a slow step followed by other fast steps to give the products. The main oxidative product was identified as (2,3-dihydro-1-hydroxy-1H-pyrrolizin-5-yl-)(phenyl)methanone and is characterized by its LC–ESI–MS spectrum. Thermodynamic parameters of various equilibria of the mechanism were calculated and activation parameters ΔH , ΔS , ΔG and log10 A were found to be 29.9 kJ·mol?1, ?220 J·K?1·mol?1, 96 kJ·mol?1 and 2.70 respectively.  相似文献   

10.
The base hydrolysis of (αβS) (salicylato) (tetraethylenepentamine)cobalt(III) has been investigated in MeOH + water and DMSO + water media (0–70% (v/v) cosolvents) at 20.0 ? t°C ? 35.0 and I = 0.10 mol dm?3 (ClO4?). The phenoxide species [(tetren)CoO2CC6H4O]+ undergoes both OH?-independent and OH?-catalyzed hydrolysis via SN1ICB and SN1CB mechanism, respectively. The OH?-independent hydrolysis of the phenoxide species is catalyzed by both DMSO + water and MeOH + water media, the former exerting a much stronger rate accelerating effect than the latter. The OH?-catalyzed reaction is strongly accelerated by DMSO + water medium but insensitive to the composition of MeOH + water medium up to 40% (v/v) MeOH beyond which it was not detectable under the experimental conditions. Data analysis has been attempted on the basis of the solvent stabilizing and destabilizing effects on the initial state and transition state of the concerned reactions. The nonlinear variation of the activation parameters, ΔH and ΔS, with solvent compositions presumably indicates that the solvent structural effects mediate the energetics of solvation of the initial state and transition state of the concerned reactions. The linearity in ΔH vs. ΔS plot accomodating all data for k1 and k2 paths in DMSO + water and MeOH + water further suggests that the solvent effects on these parameters are mutually compensatory.  相似文献   

11.
Cobalt Chelates for Hydrogenation Catalysts. II. Hydride Formation with [Co(dmgH)2] and [Co(dpnH)]+ In the presence of benzil as scavanger for the hydridocomplexes [Co(dpnH)]+ and [Co(dmgH)2] the hydride formation in water/n-propanol (50% v/v) becomes the rate determining step, and the ligand hydrogenation is completely suppressed in the case of [Co(dpnH)]+, but only partially in the case of [Co(dmgH)2]. The rate of hydride formation in both cases is 2nd order with respect to the complex, and the activation parameters ([Co(dmgH)2]: ΔH = 48.4 ± 1.0 kJ · mol–1, ΔS = ?57.4 ± 3.4J · mol?1 · K?1, [Co(dpnH)]+: ΔH = 52.7 = 0.4 kJ · mol?1, ΔS = ?59.8 ± 1.2J · mol?1 · K?1) indicate a H2-activation by homolytic splitting for both complexes. Some sources of error and possible causes for the missing activity of [Co(tim)]2+ are discussed.  相似文献   

12.
The metal ion (M2+) catalysed dissociation of cis-diaquobisoxalatochromate into the tetraaquomonooxalato complex in aqueous perchloric acid medium which follows the rate law — d(complex)/dt = {kH[H+] + kM[M2+]}[complex] has been studied. Based on kM values the order of catalysing effect of the different metal ions studied is Cu2+ > Ni2+ > Co2+ > Mn2+, which is also the order of stabilities (KMOx) of the monooxalato complexes of these metal ions; in fact the plot of log kM vs. log KMOx is linear. This together with the relative values of ΔH and ΔS for the H+ catalysed and M2+ catalysed paths is in agreement with a mechanism involving chelation of the catalysing cation through the free carbonyl oxygens of the oxalate ligand bound to Cr(III), followed by the dissociation of the Cr(III)? O bonds with simultaneous entry of two water molecules into the coordination sphere of Cr(III).  相似文献   

13.
L-脯氨酸独有的亚胺基使其在生物医药领域具有许多独特的功能,并广泛用作不对称有机化合物合成的有效催化剂。本文在碱性介质中研究了二(氢过碘酸)合银(III)配离子氧化 L-脯氨酸的反应。经质谱鉴定,脯氨酸氧化后的产物为脯氨酸脱羧生成的 γ-氨基丁酸盐;氧化反应对脯氨酸及Ag(III) 均为一级;二级速率常数 k′ 随 [IO4-] 浓度增加而减小,而与 [OHˉ] 的浓度几乎无关;推测反应机理应包括 [Ag(HIO6)2]5-与 [Ag(HIO6)(H2O)(OH)]2-之间的前期平衡,两种Ag(III)配离子均作为反应的活性组分,在速控步被完全去质子化的脯氨酸平行地还原,两速控步对应的活化参数为: k1 (25 oC)=1.87±0.04(mol·L-1)-1s-1,∆ H1=45±4 kJ · mol-1, ∆ S1=-90±13 J· K-1·mol-1 and k2 (25 oC) =3.2±0.5(mol·L-1)-1s-1, ∆ H2=34±2 kJ · mol-1, ∆ S2=-122 ±10 J· K-1·mol-1。本文第一次发现 [Ag(HIO6)2]5-配离子也具有氧化反应活性。  相似文献   

14.
The kinetics of reversible complexation of oxalatopentaammine cobalt(III) with Ni2+ has been investigated in MeOH + water media (0–50 (v/v) % MeOH) at 15.0–35.0°C and I = 0.10 mol dm?3. Analysis of rate data indicates that the monobonded complex [(NH3)5 · CoOCOCO2Ni]3+ in which Ni2+ is bound to the end carboxylate group is the possible reaction intermediate. The formation and dissociation rates of such a species are rate limiting for the overall formation and dissociation of the binuclear species, in which Ni2+ is chelated by the oxalate moiety. The rate and activation parameters for formation and dissociation of the binuclear species are moderately solvent sensitive and solvent structural effects are discernible in the nonlinear variations of ΔH and ΔS with solvent composition. The log kr vs. Grunwald Winstein parameter (Y) plot for the dissociation of the binuclear species is markedly nonlinear.  相似文献   

15.
Hindered internal rotation about the C‐N single bonds joining the thiuram disulfide was studied by 1H NMR complete line‐shaped analysis in different dimethyl sulfoxide‐chloroform (DMSO‐CDCl3) mixtures. From the temperature dependence of methyls proton spectra, activation parameters (Ea, ΔH, ΔS, and ΔG) were obtained. The Arrhenius plots showed a distinct isokinetic temperature at about 35 °C at which the exchange rate is more or less independent of the solvent composition. The resulting ΔH against TΔS plot showed a firmly good linear correlation, indicating the existence of an enthalpy‐entropy composition in an exchange process.  相似文献   

16.
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   

17.
Cobalt Chelates as Hydrogenation Catalysts. III. Hydride Formation of the [Co(dpnH)]+ Catalyst in the Presence of Pyridine as Axial Base The rate of the hydrogen uptake for [Co(dpnH)]+ is 2nd order with respect to the complex concentration and depends on the amount of added pyridine, in accord with the assumption of a more active mono- and an inactive bis-pyridine adduct. The rate law, formulated on this basis agrees very well with the measured data. The ΔH - and ΔRH°-values, calculated from the temperature dependence of the rate constants and equilibrium constants are in agreement with the suggested model, whereas the ΔS and ΔRS° values do not correspond completely with these expectations.  相似文献   

18.
The kinetics of pyridine exchange on trans-[MO2(py)4]+ have been followed by 1H-NMR in CD3NO2 for M = Re, Tc: k298S?1 = (5.5 ± 0.1) × 10?6, 0.04 ± 0.02; ΔH/kJmol?1 = 111 ± 3, 101 ± 9; ΔS/JK?1mol?1 = +28 ± 10, +68 ± 35. For the Rev complex, pyridine and oxygen exchanges have been measured simultaneously by 1H- and 17O-NMR in deuterated water: k298/s?1 = (8.6 ± 0.2) × 10?6 (py), (14.5 ± 0.3) × 10?6 (oxygen); ΔH/kJmol?1 = 111 ± 1, 91 ± 1; ΔS /JK?1mol?1 = +32 ± 3, ?32 ± 4. For both complexes, the rate law for pyridine exchange is first-order in complex and zero-order in pyridine; together with the activation parameter values, and the fact that the rate does not depend significantly on the nature of the solvent, this strongly implies the operation of a dissociative mechanism. The ratio of pyridine exchange rates for the Tc and Re complexes at room temperature is ca. 8000. The consequences of these observations for radiopharmaceutical synthesis are discussed.  相似文献   

19.
Solvolyses of tosylates IIB and IIC in a variety of solvents were found to give excellent linear log k -YBnOTs plots (R = 0.997 and SD = 0.022) and negative ΔS for IIB , but slight deviations (R = 0.987 and SD = 0.047) from lineation and positive ΔS for several cases for IIC . Limiting SN1 mechanisms with steric hindrance to resonance and solvent intervention at the cationic transition state could thus be confirmed. The inconsistency between theoretical and observed trend of ΔH is attributed to the neglect of solvent effect in MO calculations. Deductions linking theoretical calculation and solvolytic experiments should be drawn with caution.  相似文献   

20.
The kinetics of oxygen exchange between water (H2O, D2O) and 18O-labelled bromate ion has been investigated over the range of 1.7 ≤ pH ≤ 14.3 and 20 ≤ °C ≤ 95. At 60° and ionic strength I ? 1.0M (NaNO3), the experimental results were consistent with the rate laws (R in moll?1 s?1): From the temperature dependence of the rate constants the activation parameters ΔH, ΔS and ΔC were derived. In the acid-catalysed region the form of the rate law and the direction of the solvent isotope effect were the same as previously found, but the numerical values of ΔH and k2H/k2D differ considerably. For the spontaneous and the OH?-catalysed exchange reactions bimolecular displacement mechanisms are proposed.  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