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 468 毫秒
1.
The catalytic oxidation of iron(II) with oxygen occurs along with an autocatalytic reaction between palladium(II) tetraaqua complex and iron(II) aqua complex in an oxygen atmosphere. The reaction is catalyzed by a compound of palladium in an intermediate oxidation state, presumably by a small palladium cluster formed in the course of the reduction of palladium(II) tetraaqua complex with iron(II) aqua complex.  相似文献   

2.
Insertion of molecular oxygen into a palladium(II) hydride bond to form an (eta1-hydroperoxo)palladium(II) complex is reported. The hydroperoxo palladium(II) product has been crystallographically characterized. A second-order rate law (first-order in palladium and first-order in oxygen) is observed for the reaction and a large kinetic isotope effect implicates Pd-H bond cleavage in the rate-determining step. The results of studies with radical inhibitors and light suggest that the reaction does not proceed by a radical chain mechanism.  相似文献   

3.
The kinetics of hydrogen peroxide oxidation of Fe(II) to Fe(III) complexed with tetraazamacrocyclic ligand was studied, and a decrease in the reaction rate was observed in the presence of nitrogeneous bases, capable of forming hexacoordinated complexes with tetraazamacrocyclic compound of iron(II). The rate of reaction is proprotional to the concentration of the iron complex and hydrogen peroxide and inversely proportional to the concentration of the nitrogeneous base. A mechanism for the course of the reaction has been proposed, and the rate constants of the oxidation of the pentacoordinated iron(II) complexes have been calculated. It was shown that the addition of the fifth donor particle (in particular imidazole) activates the iron(II) atom with respect to the oxidation reaction. It was found that a tetraazamacrocyclic complex of iron(II) is capable of displaying a peroxidase type activity.Translated from Teoreticheskaya Eksperimental'naya Khimiya, Vol. 22, No. 3, pp. 309–316, May–June, 1986.  相似文献   

4.
A micellar solution of cetylpyridinium chloride (CPC) can separate into two phases due to a temperature change or to the addition of salts. Platinum(II), (IV) and palladium(II) reacted with chloride ions to form stable anionic complexes of PtCl4(2-), PtCl6(2-) and PdCl4(2-), respectively, and were adsorbed onto the CPC gel phase. The CPC phase plays the role of an ion-exchange adsorbent for the anionic complexes. By such a procedure, the precious metals of platinum and palladium could be separated from base metals such as copper, zinc and iron. The kinetic separation was performed by a ligand exchange reaction of the palladium(II) chloro-complex with EDTA at 60 degrees C. The anionic palladium(II)-EDTA complex could not bind the opposite charged CP+ and was desorbed from the CPC phase. In the aqueous phase, the recovery of palladium(II) by the double-desorption was 101.1 +/- 1.2%. The platinum(II) and (IV) chloro-complexes were stable for at least 30 min and remained in the CPC phase.  相似文献   

5.
The mechanism of styrene carbonylation in the presence of a palladium(II) complex has been investigated by quantum chemical methods. The migratory insertion of styrene into the palladium-hydrogen bond of hydride π-complexes is considered in detail. This process determines the regioselectivity of the reaction.  相似文献   

6.
The extraction properties of (RS)-1-(4-chlorophenyl)-4,4-dimethyl-3-(1H-1,2,4-triazol-1-ylmethyl)-pentan-3-ol (with chloroform as a diluent) with respect to palladium(II) were studied. Palladium(II) was found to be efficiently extracted by the reagent from 0.1–6 M HCl solutions by the coordination mechanism. The rate of palladium(II) recovery depends on the hydrochloric acid and chloride ion concentrations in the aqueous phase. Conditions for the selective separation of palladium(II) and copper(II) from nickel(II), cobalt(II), and iron(III) were determined.  相似文献   

7.
The kinetics of substitution of chloride ions for ethylendiamine in a bis-ethylenediamine complex of palladium(II) is studied in solutions with pH 0.3–2.0 and the ionic strength of 0.11–1.0 M by measuring the transients of limiting diffusion currents of the electroreduction of palladium(II) ethylenediamine complexes on a rotating disk electrode. The first reaction orders in hydrogen and chloride ions are found at the ionic strength of 1 M (NaClO4). The activation energy of the homogeneous reaction under study is determined from the temperature dependence of its rate constant. The mechanism of substitution of chloride ions for ethylenediamine is discussed.  相似文献   

8.
We found a reaction of palladium(II) with picramine-epsilon to give a colored complex with an absorption maximum at 556 nm and a molar absorption coefficient of (2.01 ± 0.02) × 104. Conditions of spectrophotometric reaction were selected for determining palladium(II) in nitric acid solutions or solutions containing aqua regia without the conversion of initial complexes into other forms. A rapid procedure was developed for the direct spectrophotometric determination of palladium(II) in concentrations down to 0.5 mg/L in various process solutions (including those containing nitric acid and aqua regia) within 15 min.  相似文献   

9.
The interaction of palladium(II) and platinum(II) with tin(II) chloride in hydrochloric acid solutions was studied by flow-injection (FI) spectrophotometry. It was found using kinetic measurements in the stopped flow mode that the composition of detected products and the rate of their formation depend on the concentrations of tin(II) and chloride ions in the reaction zone and on the acidity of the solution. Optimal FI conditions were found, and the selectivity of interaction of palladium(II) with tin(II) chloride was estimated for the detection of the signal at 407 nm (yellow form) and 646 nm (green form). It was demonstrated that the reaction of the formation of yellow platinum(IV) complexes is slower than that for palladium(II), especially at rather low concentrations of hydrochloric acid in the reaction flow. Based on the detection of green complexes of palladium(II) with tin(II) chloride, a flow injection method was proposed for the selective spectrophotometric determination of palladium(II) in the presence of other platinum-group metals. The height of the recorded peak is directly proportional to the concentration of palladium(II) in the injected solution in the range of 0.028–0.300 mM. The method was used for the analysis of pharmaceuticals and industrial catalysts.  相似文献   

10.
The redox reaction between cobalt(II) and gold(III) chloride in the presence of 1.10-phenanthroline or 2,2'-bipyridine was studied, and a titration of the cobalt(II) complex with a gold(III) chloride solution was developed. A 4-fold amount of 1,10-phenanthroline or 2,2'-bipyridine was necessary for rapid quantitative reaction; the permissible pH range was 1.5–5. The oxidation of the cobalt(II) complex proceeds rapidly at 40–50°C, and a direct potentiometric titration was possible. The following maximum errors were obtained: 3.3% for 0.2–1.0 mg Co, 2.0% for 1–5 mg Co, and 0.70% for 10–40 mg Co. The following ions did not interfere: Ni(II), Zn(II), Pb(II), Cd(II), Mn(II), Fe(II), Cr(III), Al(III), Th(IV), Se(IV), Ti(IV), U(VI), Mo(VI), SO2-4 and PO3-4. Even small quantities of silver(I), copper(II), palladium(II), mercury(II)and iron(III) interfered. The method was applied to the determination of high cobalt contents in high-temperature nickel-base alloys.  相似文献   

11.
The electrochemical behavior of palladium(II) complexes of Tropeolin 0 (Tr0) was studied by linear-potential-sweep voltammetry in acetate–ammonia buffer solutions. The optimum conditions for determining palladium(II) with Tr0 were found. The composition of the complex was found to be Pd : Tr0 = 1 : 2. A procedure was proposed for determining palladium(II) with a detection limit of 2.54 × 10–7M. The procedure was used for determining palladium(II) in capacitors.  相似文献   

12.
Interaction of adenine (A) with dichloro-[1-alkyl-2-(α-naphthylazo)imidazole] palladium(II) [Pd(α-NaiR)Cl2], 1 and dichloro-[1-alkyl-2-(β-naphthylazo)imidazole] palladium(II) [Pd(β-NaiR)Cl2], 2 {where R=Me (a), Et (b) or Bz (c)} in MeCN-water (50% v/v) medium to yield [{1-alkyl-2-(α-naphthylazo)imidazole}(adenine)]palladium(II) perchlorates (3a, 3b, 3c) and [{1-alkyl-2-(β-naphthylazo)imidazole}(adenine)]palladium(II) perchlorates (4a, 4b, 4c) was studied. The products were characterized by physico-chemical and spectroscopic methods. The reaction kinetics were second order overall, being first order in both the Pd(II) complex and adenine. The effect of adding chloride was consistent with rate-limiting dissociation of chloride from the complex. Thermodynamic parameters were determined from temperature variation experiments. The second-order rate constant k 2 corroborates with the experimental ΔH° values, while the negative values of ΔS° indicate that the reaction proceeds through an associative inner sphere mechanism.  相似文献   

13.
Odashima T  Yamaguchi M  Ishii H 《Talanta》1995,42(9):1229-1237
The kinetics of complexation reactions of five water-soluble heterocyclic hydrazones with nickel(II) and palladium(II) ions have been investigated by stopped-flow spectrophotometry. Rates of complexations with nickel(II) and palladium(II) in the absence of chloride ion were found to be proportional to the first order of the ligand and metal ion concentrations and to the inverse first order of the hydrogen ion concentration except for the complexation of alpha-(2-benzimidazolyl)-alpha-(5-nitro-2-pyridyl)hydrazono-3-toluenesulfonic acid with palladium(II). Rates of complexation with palladium(II) in the presence of chloride ion were best described by a two-term expression, both terms being first order in the palladium ion and ligand concentrations and inverse first order in the hydrogen ion concentration. The first term has zero dependence of the chloride ion concentration, whereas the second is first order with respect to the chloride ion concentration. The rate constant for each complexation reaction was determined. The complexation of the hydrazones with nickel(II) was estimated to go according to an Eigen mechanism and that with palladium (II) according to the associative mechanism.  相似文献   

14.
Reduction of palladium(II) glycinate complexes in strongly acid 0.5 M NaClO4 solutions (pH 0.6 and 1.0) with variable palladium(II) complex and free glycine concentration was studied by the taking of cyclic voltammograms at palladium rotating disc electrode. It is shown that it was a chelate monoglycinate palladium(II) complex that was present in all studied solutions and underwent the reduction. The diffusion coefficient of the chelate monoglycinate palladium(II) complex D = (6.5 ± 0.5) × 10−6 cm2/s was determined from the limiting diffusion current of the complex reduction. The monoglycinate palladium(II) complex reduction occurred in the double-layer segment of the palladium charging curve; it was not complicated by hydrogen adsorption at electrodes. The palladium(II) complex reduction half-wave potential was determined (E 1/2 = ∼0.300 to 0.330 V (SCE)). It is shown that the decreasing of the number of ligands coordinated by palladium via nitrogen atom facilitates the complex reduction process. In particular, the reduction potentials of palladium(II) complexes with different ligand number at palladium electrode shifted markedly toward negative potentials in the series: Pdgly+ < Pd(gly)2 < Pd(gly)42−.  相似文献   

15.
The kinetics of the reaction between iron(II) and vanadium(V) have been investigated in the pH range 2.6–4.2 where decavanadates and VO2+ coexist in equilibrium. Under these conditions, the observed kinetic pattern is radically different from the one reported for the reaction in strong acid medium. In the pH range employed, the reaction rate is not appreciably altered by variation in the stoichiometric vanadium(V) concentration due to the operation of the equilibrium between the reactive species, VO2+, and the unreactive species, decavanadates. The reaction, however, obeys first‐order kinetics with respect to Fe(II). In the presence of salicylic acid, which imparts considerable reactivity to iron(II) by reducing the reduction potential of iron(III)/iron(II) couple by forming a stronger complex with iron(III) than iron(II), the kinetic results provide evidence for the participation of decavanadates in the electron transfer. The mechanism under both conditions is discussed. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 535–541, 2000  相似文献   

16.
A catalytic-spectrophotometric method for the determination of traces of palladium(II) is proposed. The reaction is based on the catalytic action of palladium(II) on the reduction of azure I (λmax = 647 nm) by sodium hypophosphite. The various variables affecting the sensitivity were studied, and a study of interfering ions was also carried out. The reaction gave a detection limit of 4.3 ng/mL palladium(II) and good reproducibility with a relative standard deviation of 1.53–1.98% in the palladium(II) concentration range 40–200 ng/mL. The method yielded another linear range (5–40 ng/mL) when using slightly different conditions. In this case, the detection limit was 0.78 ng/mL palladium(II), and the relative standard deviation for ten replicate analyses of 20 ng/mL palladium(II) was 2.05%. The method was applied to the determination of palladium in a sample of activated charcoal. The text was submitted by the authors in English.  相似文献   

17.
Summary Gel chromatographic behaviour of tris(1, 10-phenanthroline)iron(II), tris(2,2′-bipyridine)iron(II) and tris(glycinato)cobalt(III) on Sephadex G-10 or G-25 was investigated by TLC with 0.001–1.0M NaCl as the eluent. The zone shapes and RM values of tris(1,10-phenanthroline)iron(II) and tris(2,2′-bipyridine)iron(II) were appreciably dependent on the sample and eluent concentration, while the neutral complex, tris(glycinato) cobalt(III), exhibited the round zones with constant RM values. The order of RM values was found to be tris(glycinato)cobalt(III<tris(2,2∔pyridine)iron(II)<tris-(1,10-phenanthroline)iron(II) in all systems studied, although the reverse trend was expected when assuming the chromatographic behaviour of solute compounds to be controlled by the “sieving effect”. The comparison of the behaviour on Sephadex G gels with that on CM-cellulose revealed that the predominant mechanism involved is not the sieving effect, but ion-exchange and/or hydrophobic interaction.  相似文献   

18.
Anomalous results for the reaction between palladium(II) and fluorescent complexes of 8-quinolinol-5-sulfonic acid were found to be due to a slow reaction. Heating the solution permits the determination of palladium(II) in combination with copper(II) or nickel(II), the two cations being titrated stepwise The approximate value for the overall formation constant of the PdL22- complex was determined.  相似文献   

19.
Summary Oxidation of the diaqua(nitrilotriacetato)cobaltate(II) complex, [CoIInta(H2O)2]-, by NBS has been studied in aqueous medium. The kinetics of the reaction in the presence of an iron(II) catalyst obey the rate law: Catalysis by iron(II) is believed to be due to the oxidation of iron(II) to iron(III), which acts as the oxidizing agent. The thermodynamic activation parameters were calculated and we propose that electron transfer proceeds through an inner-sphere mechanism. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

20.
In this work, a mechanism of the reaction between palladium(II) and maleic acid has been theoretically investigated in detail. Upon spectroscopic studies, a structure for this complex had been proposed in which the hydrogen maleate was coordinated to palladium through a C=C double bond and one carboxylic group. Experimentally, a three‐pathways mechanism had been proposed for this process, which consists of a slow ring closure (kSRC) and two bimolecular processes (k3k4). The kSRC pathway is the direct and dominant route. The k3 and k4 pathways commence with the attack of a maleic acid to hydrogen maleate in the palladium complex. Using the quantum mechanical approach, different pathways were investigated. Considering the solvent effects, the activation energies of three pathways and other contingent pathways were calculated and compared with each other and experimental results.  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