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.
Oxalato(amine)(trien)cobalt(III) and salicylato(amine)(trien)cobalt(III) perchlorates have been synthesized and tentatively assigned a cis α configuration. The dissociation constants of the complexes have been determined. The base hydrolysis of the complexes have been investigated at 30, 35, 38.6°C and I = 1.0 mol dm−3. The rate laws for the oxalato and salicylato complexes are −dln[Complex]T/dt = k2[OH] and −dln[Complex]T = k1 + k2[OH] respectively. For the oxalato complex, k2(30°C) = (2.95 ± 0.05) × 10−2 dm3 mol−1 s−1, ΔH1 = (109 ± 4) KJ mol−1, ΔS1 = (85 ± 13) JK−1 mol−1 and for the salicylato complex, k1(30°C) = (1.58 ± 0.28) × 10−4 s−1, ΔH1 = (166 ± 7) KJ mol−1, ΔS1 = (229 ± 21) JK−1 mol−1 and k2(30°C) = (3.56 ± 0.10) × 10−3 dm3 mol−1 s−1, ΔH1 = (101 ± 6) KJ mol−1, ΔS1 = (42 ± 20) JK−1 mol−1. The base hydrolysis reactions of the complexes were followed in presence of imidazole and ethanolamine, in the presence of added anions and also in D2O medium. The results are discussed in terms of δ SN1CB mechanism involving rate limiting CoO bond fission for both k1 and k2 paths. However, the possibility of CO bond cleavage in the base hydrolysis of the oxalato complex is not ruled out.  相似文献   

2.
3.
Summary The kinetics of reversible complexation of Ni(OH2) inf6 sup2+ with oxygen-bonded glycinatocobalt(III) substrates N4-Co(glyH)gly2+ [N4 = (en)2 or trien; glyH = H3N+CH2-COO] have been investigated by the stopped-flow technique in the 20–35° C range, at pH = 6.08–6.82 and I = 0.3 mol dm–3. The formation of N4Co(glyH)glyNi4+ occurred via the reaction of Ni(OH2) inf6 sup2+ with the deprotonated form of the cobalt(III) substrates, N4Co-(glyH)gly2+. The rate and activation parameters for the formation and dissociation of the binuclear species are reported. The formation rate constants k f (at 25° C), activation enthalpy and entropy H , S for N4Co-(glyH)glyNi4+ are 320±49, 341 ± 52dm3mol–1 s–1, 78 ± 7, 79 ± 5 kJmol–1 and 64 ± 24, 69 ± 18 JK–1 mol–1 for the ethylenediamine and triethylenetetraminecobalt(III) substrates, respectively. This result indicates that the rate and activation parameters are virtually independent of the nature of N4 moities, which strongly suggests that the formation of mono-bonded species occurs via entry of one of the pendant NH2 groups into the coordination sphere of nickel(II) via a rate-limiting Ni-OH2 bond dissociation mechanism (Id). The binuclear species exist in dynamic equilibrium between the monodentate and chelated forms, with the chelate form predominating. The low values of spontaneous dissociation rate constant for the binuclear species (k r- 0.095–1 at 25° C) in comparison with the high values of dissociation rate constants of monodentate nickel(II) complexes reported in the literature also support the chelate nature of the binuclear species.  相似文献   

4.
A study has been made on the oxidation of bis(2,2,6, 2-terpyridine)-iron(II), Fe(tpy) 2 2+ by manganese (IV) using stopped-flow spectrophotometry in H2SO4–H3PO4 mixtures. The reaction is first order in each the substrate and the oxidant. The rate of the reaction increases with hydrogen ion concentration. A plausible mechanism is proposed considering the protonated forms of manganese(IV) as reactive oxidizing species. The reaction obeys the rate law
  相似文献   

5.
Summary The reaction of ()-(tetren)CoOH2+ with S2O 3 2- in the 7.25–8.28 pH range at 20–40 °C yielded S- (yellow) and O- (purple) bonded thiosulfato(tetren)cobalt(III) complexes, the former in larger quantities. The rate determining step is preceded by diffusion-controlled ion-pair [(tetren)CoOH2+,S2O 3 2- ] formation. Replacement of coordinated OH- by S2O 3 2- is interpreted in terms of an internal conjugate base mechanism: (tetren)CoOH2+ (tetren-H)CoOH 2 2+ , the reactive amido conjugate base being generated by intramolecular proton transfer from the coordinated NH site.In acid medium the S-bonded (tetren)Co(S2O3)+ is highly stable to redox decomposition, in contrast to its pentaammine analogue. The complex however, undergoes base hydrolysis yielding the corresponding hydroxo complex. The rate and activation parameters for the base hydrolysis have been reported. Photolysis of O- and S-bonded isomers of [(tetren)CoS2O3]+ in acidic medium at 254 and 313 nm, respectively, yielded aquation products accompanied by some decomposition of S2O 3 2- .  相似文献   

6.
《Polyhedron》2001,20(15-16):1885-1890
The macrocycle L, prepared by template condensation of bis-6,6′′-(α-methylhidrazino)-4′-phenyl-2,2′:6′′,2′-terpyridine with glyoxal, forms a stable crystalline complex of cobalt(II) [Co(L)(H2O)2][PF6]2 which has been used as a starting material to prepare, for electrochemical studies, a series of seven coordinate cobalt(II) complexes [Co(L)X2][PF6]2 (X=pyridine, 4-cianopyridine, 4-aminopyridine, 4-dimethylaminopyridine, pirazine, imidazole, 1-methylimidazole, 2-methylimidazole, and trimethylphosphite). Cyclic voltammetry of the aquo complex in DMSO show one reversible reduction wave at −1.35 V versus Ag  AgBF4 reference electrode and controlled potential electrolysis in the presence of trimethylphosphite affords a diamagnetic species which has been assigned as a mononuclear d8 Co(I) species. The crystal and molecular structure of [Co(L)(imidazole)2][PF6]2·Me2CO shows the metal to be in a pentagonal-bipyramidal N7 environment.  相似文献   

7.
Precipitation behaviors of Np(IV), (V) and (VI) in sulfuric acid and ammonium sulfate solutions with hexammine cobalt(III) ion have been studied. The optimum conditions of precipitating Np as crystalline compounds are 0.4 M H2SO4 and 0.06 M [Co(NH3)6]Cl3 for Np(IV), larger than 0.5 M (NH4)2SO4 and 0.03 M [Co(NH3)6]Cl3 for Np(V), and the higher the concentrations the better for Np(VI). Np(IV) is completely removed from ammonium sulfate solution.Composition of the Np(IV) compound from the sulfuric acid solution is [Co(NH3)6]2[Np(SO4)5nH2O. And composition of the Np(V) compounds is possibly [Co(NH3)6][NpO2(SO4)2]·α[Co(NH3)6]2(SO4)3·nH2O (α = 2.20−0.32). Solubilities of the Np(IV), (V) and (VI) compounds in water are found to be 4.9 mgNp/l, 520 mgNp/l, and 250 mgNp/l, respectively.  相似文献   

8.
Two new complex anions, [Cr(N3)(S-pdtra)]– and [Cr(N3)(edtrp)]–, were obtained in solution by N3–/HN3 anation of the aqua analogues (S-pdtra = S-propane-1,2-diamine-N,N,N-triacetate, edtrp = ethylenediamine-N,N,N-tripropionate). Aquation of these species in acidic media leads to the same geometrical isomers as those used for the synthesis. The aquation rate is strongly dependent upon [H] and is substantially higher in D2O than in H2O. Protonation of the coordinated azide was not observed spectrophotometrically. The rate law and activation parameters have been determined and discussed.  相似文献   

9.

Abstract  

The surfactant complex ion cis-[Co(tmd)2(C12H25NH2)2]3+ (tmd = 1,3-propanediamine, C12H25NH2 = dodecylamine) has been synthesized and characterized by elemental analysis and spectral data. In addition we have determined the critical micelle concentration of the surfactant–cobalt(III) complex and studied the kinetics and mechanism of the complex with ferrocyanide anion. The reaction is found to be second order, and the second-order rate constant increases with increasing initial concentration of the surfactant–cobalt(III) complex due to the presence of self-micelles formed by the complex itself. The thermodynamic parameters were determined. The results have been analyzed.  相似文献   

10.
The kinetics of ruthenium(III) catalyzed oxidation of chloramphenicol (CHP) by diperiodatocuprate(III) (DPC) in aqueous alkaline medium at a constant ionic strength of 0.1 mol l−1 was studied spectrophotometrically. The reaction between DPC and CHP in alkaline medium exhibits 1: 2 stoichiometry (CHP: DPC). The main oxidation products were identified by spot test, IR, NMR, and GC-MS spectral studies. The reaction is first order with respect to ruthenium(III) and DPC concentrations. The order with respect to chloramphenicol concentration varies from first order to zero order as the chloramphenicol concentration increases. As the alkali concentration increases the reaction rate increases with fractional order dependence on alkali concentration. Increase in periodate concentration decreases the rate. A mechanism adequately describing the observed regularities is proposed. The reaction constants involved in the different steps of the mechanism were calculated. The activation parameters with respect to limiting step of the mechanism are computed and discussed. Thermodynamic quantities are determined.  相似文献   

11.
The kinetics of oxidation of the neutralized -hydroxy acids: lactic, -hydroxyisobutyric, mandelic, benzilic and atrolactic acids by tris(pyridine-2-carboxylato)manganese(III) have been studied. The reactions were carried out in a Na(pic)-picH [Na(pic) = sodium salt of pyridine-2-carboxylic acid and picH = pyridine-2-carboxylic acid] buffer medium in the 4.89–6.10pH range. The oxidation rate was found to be independent of pH, and rate follows the order: benzilate > mandelate >atrolactate>lactate > -hydroxy isobutyrate. The oxidation products are MeCHO, Me2CO, PhCHO, Ph2CO and PhCOMe for the respective reactions. A mechanism is proposed involving intermediate formation of hepta-coordinated MnIII complexes in a fast step. The complexes then decompose to give free radicals and MnII in the rate determining step. The free radicals subsequently react with another molecule of the MnIII species to give the respective carbonyl compounds in a fast step.  相似文献   

12.
13.
The positive, liquid secondary ion (LSI) mass spectra of six cobalt(III) and three chromium(III) (β-diketonates ligand = L?) were examined in a 3-nitrobenzyl alcohol matrix. The complexes of both metals yield clean, matrix-free mass spectra, but there are important differences between them. The cobalt compounds show prominent peaks assignable to the molecular ion, CoL 3 + , of the monomeric chelates, together with abundant dimeric ions, such as Co2L 4 + and Co2L 3 + ; in contrast, chromium complexes show protonated monomers, CrL3H+, in addition to ionized monomers, CrL 3 + , and only minor formation of dimeric ions. The collisionally-activated dissociation (CAD) mass spectrum of Co2L 4 + shows fragmentation to CoL 2 + and Co2L 3 + . That of Co2L 3 + shows fragmentation only to dimeric ions, including Co2L 2 + and, for thienyl or phenyl substituted ligands, to Co2L2Ar+ or Co2LAr+ (Ar = thienyl or phenyl). Neither Co2L 4 + nor Co2L 3 + dissociates to the CoL 3 + ion. The LSI mass spectrum of a mixture of two different cobalt chelates shows dimeric ions containing both types of ligand, which can be explained by ion-molecule reactions in the selvedge region. The differing behaviors of the cobalt and chromium complexes is attributed to the relatively greater stability of the +2 oxidation state for cobalt than for chromium.  相似文献   

14.
The substitution of bis(2,4,6-tripyridyl 1,3,5-triazine)iron(II), \textFe(TPTZ) 2 2 + {\text{Fe(TPTZ)}}_{ 2}^{{ 2 { + }}} by 2,2′,6,2″-terpyridine (terpy) occurs on a time scale of about 6 m. The kinetics of this reaction was followed by stopped-flow spectrophotometry in the pH range of 3.6–5.6 in acetate buffer. The data indicate that the reaction occurs in two consecutive steps: kinetic data for both steps were acquired simultaneously and analyzed independently. The first step is assigned to the reaction between \textFe(TPTZ) 2 2 + {\text{Fe(TPTZ)}}_{ 2}^{{ 2 { + }}} and terpy to give Fe(TPTZ)(terpy)2+, followed by its reaction with another terpy molecule to give the final product, \textFe(terpy) 2 2 + {\text{Fe(terpy)}}_{ 2}^{{ 2 { + }}} . The rate of the reaction increases with increases in [terpy] and pH. The kinetic and activation parameters determined for both steps suggest that they involve both associative and dissociative paths. The ternary complex Fe(TPTZ)(terpy)2+ has been prepared, and the kinetics of its reaction with terpy suggest that this reaction is identical with the second step of the \textFe(TPTZ) 2 2 + {\text{Fe(TPTZ)}}_{ 2}^{{ 2 { + }}} -terpy system, supporting the proposed mechanism.  相似文献   

15.
Summary The kinetics of the oxidation of ascorbic acid by diaquatetrakis (2,2-bipyridine)--oxo diruthenium(III) in aqueous HClO4 were investigated. The dependence of the second order rate constantk 2 on [H+] is given by k 2=a+b[H+], indicating that both the undissociated form and the monoanion of ascorbic acid are reactive. Marcus theory was used to estimate the redox potential for the RuIII-O-RuIII/RuIII-O-RuII couple and a feasible mechanism has been proposed to explain the results.  相似文献   

16.
In order to investigate the mechanism of mercuration reaction of substituted ben-zylideneanilines, kinetic measurements of these reactions at different temperatures (40-60℃) inmethanol-l,4-dioxane (1/1, V/V) were carried out and Hammett ρ value for C-phenyl substituentsof-0.61 for the N-(substituted benzylidene)-4-toluidine series was obtained. Thermodynamicparameters E_a, △S~≠ were obtained for the reaction of different. N-(substituted benzylidene)-4-toluidines. It was found that this ortho-mercuration was brought about by an internal cyclometal-lation process involving the imino-moiety.  相似文献   

17.
Two new chromium(III)–nicotinate complexes, cis-[Cr(C2O4)2(O-nic)(H2O)] and cis-[Cr(C2O4)2(N-nic)(H2O)], were obtained and characterized in solution (where O-nic=O-bonded and N-nic=N-bonded nicotinic acid). The kinetics of nicotinate ligand liberation were studied spectrophotometrically in the 0.1–1.0 m HClO4 range, at I=1.0 m. The rate equations were determined and a mechanism is proposed. The rate of Cr–O bond breaking is [H+] dependent: kobs=kHQH[H+], where kH is the acid-catalyzed rate constant and QH is the protonation constant of the nonbonded oxygen atom in the O-coordinated ligand. The Cr–N bond breaking proceeds via two paths: spontaneous and acid-catalyzed; kobs=k0 + kHQH[H+], where k0 and kH are the spontaneous and acid catalyzed rate constants and QH is the protonation constant of the carboxylic group in the N-bonded nicotinic acid. The results demonstrate by comparison that Cr–N bond breaking is a much slower process than Cr–O bond fission.  相似文献   

18.
The solubilities of tris(2,2,6,6-tetramethyl-3,5-heptanedionate) cobalt(III) (Co(thd)3) and chromium(III) (Cr(thd)3) in supercritical carbon dioxide (scCO2) were measured at temperatures ranging from 313 to 343 K. The measurements were carried out using a circulation-type apparatus with a UV–vis spectrometer. The solubilities of both Co(thd)3 and Cr(thd)3 increased as both the density of scCO2 and the temperature increased, which has the same tendency as cobalt(III) acetylacetonate (Co(acac)3) and chromium(III) acetylacetonate (Cr(acac)3) had in our previous work. The solubilities of Cr(thd)3 were higher than that of Co(thd)3, and the solubilities of Co(thd)3 and Cr(thd)3 were about 50- and 70-fold higher than those of Co(acac)3 and Cr(acac)3, respectively. The measured solubilities of the metal complexes were correlated with the equation based on Chrastil's equation. The parameters were determined by correlating the experimental data for each metal complex, and the correlated results well reproduced the experimental data, especially Co(thd)3. Moreover, the charge density distributions on the molecular surface of CO2 and the metal complexes used in the measurement were estimated by the quantum chemical calculation and the COSMO-RS to clear the effect of the molecular structure of the metal complexes on the affinity for CO2.  相似文献   

19.
In this paper, multicolored micelles were prepared by coordination of lanthanide(III) (europium(III) (Eu(III)) and terbium(III) (Tb(III))) ions with block copolymer in different molar ratios of n Eu(III)/n Tb(III). The micelles formed by polymer–Eu(III)/Tb(III) could emit higher quantum yield luminescence than the mixture of polymer–Eu(III) micelles and polymer–Tb(III) micelles. The micelles containing Eu(III) and Tb(III) could emit a yellow-green color, and the intensity varied with the molar ratios of n Eu(III)/n Tb(III). In the constant concentrations of Eu(III) and 1,10-phenanthroline (Phen), the intensity of 5D07F2 increased with the addition of Tb(III), and the intensity of 5D47F5 decreased with the increasing of Eu(III) in the constant concentrations of Tb(III) and Phen. All the multicolored micelles could be spin-coated as intensity-tunable films.  相似文献   

20.
New approaches to the synthesis of poly(2,2′,3,3′-indole) were developed based on the photochemical dehydropolycondensation of indole in the presence of iodine and the photochemical polycondensation of 1-(1H-indol-3-yl)-2-iodo-1-ethanone in the absence of catalyst and solvent. A suggested mechanism for the formation of the oligomeric chain in these reactions includes the intermediate formation of 3,3′-diindole with subsequent polycondensation via elimination of hydrogen atoms at position 2 of the dimer pyrrole fragment.  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