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 hydroquinone (H2Q) by a μ-oxo-bridged diiron(III,III) complex, Fe2(μ-O)(phen)4(H2O)2]4+ (1) has been investigated in aqueous media at 25.0 °C in presence of an excess of 1,10-phenanthroline (phen). The overall redox rate increases with increase in [H+]. The title complex (1) and its conjugate bases, [Fe2(μ-O)(phen)4(OH)2]3+(2) and [Fe2(μ-O)(phen)4(OH)2]2+ (3), participate in the reaction with H2Q as the only kinetically reactive reducing species. Rate constants (in dm3 mol−1 s−1) for the parallel reactions (1) + H2Q → Products, (2) + H2Q → Products and that for (3) + H2Q → Products are, respectively, 500 ± 40, 100 ± 6 and 30 ± 2. Substantial rate retardation in D2O media in comparison to that in H2O media suggests that electron transfer is coupled with proton movements in the rate-determining step.  相似文献   

2.
The kinetics of the oxidation of promazine and chlorpromazine by hexaimidazolcobalt(III) were studied in the presence of a large excess of cobalt(III) and H+ ions using u.v.–vis. spectroscopy ([CoIII] = (1–6) × 10−3 m, [ptz] = (2.5–10) × 10−5 m, [H+] = 0.05–0.8 m, I = 1.0 m (H+, Na+, Cl), T = 333–353 K, l = 1 cm). In each case, the reversible reaction leads to formation of cobalt(II) species and a stable cationic radical. A linear dependence of the pseudo-first-order rate constant (kobs) on [CoIII] with a non-zero intercept was established for both phenothiazine derivatives. A marked difference in the observed reaction rate for promazine and chlorpromazine is associated with the difference in its ability to undergo oxidation and is consistent with a trend in the redox potential changes for these reductants. The activation parameters for reactions studied were determined. Mechanistic consequences of all the results are discussed.  相似文献   

3.
The title MnIV complex, [Mn(LH2)3]4+ (LH2 = biguanide = H2NC(NH)NHC(NH)NH2), an authentic two‐electron oxidant, quantitatively oxidizes hydrazine (H2NNH2) to dinitrogen in the pH interval 2.00–3.50. The net four‐electron oxidation of hydrazine is provided by two MnIV as established by stoichiometric studies. The overall reaction is composed of two parallel paths:  相似文献   

4.
The oxidation of acyclovir by diperiodatocuprate(III) in aqueous alkaline media, at a constant ionic strength of 0.01 mol?dm?3, was studied spectrophotometrically at 25?°C. The reaction between acyclovir and DPC in alkaline media exhibits 1:4 stoichiometry (acyclovir:diperiodatocuprate(III)). The main oxidation products were identified by a spot test, along with infrared and liquid chromatography mass spectral studies. The oxidation reaction is first order with regard to the diperiodatocuprate(III) concentration, but has less than unit order in the acyclovir concentration and negative fractional orders in the periodate and alkali concentrations. Intervention of free radicals was observed in the reaction. The oxidation reaction in alkaline media was shown to proceed via a diperiodatocuprate(III)?Cacyclovir complex, which decomposes slowly in a rate determining step followed by subsequent fast steps to give the products. A suitable mechanism is proposed for these observations. The reaction constants involved in the different steps of the mechanism were calculated. The activation parameters with respect to the slow step of the mechanism, along with the thermodynamic quantities, were determined and discussed.  相似文献   

5.
The synthesis of α‐substituted carbonyl compounds is of great importance due to their ubiquity in both natural and man‐made biologically active compounds. The field of hypervalent iodine chemistry has been a great contributor to access these molecules. For example, the α‐oxidation of carbonyl compounds has been one of the most investigated iodine(III)‐mediated stereoselective transformations. Yet, it is also the transformation that has met the most challenge in terms of achieving high stereoselectivities. The different mechanistic pathways of the iodine(III)‐mediated α‐tosyloxylation of ketones have been investigated. The calculations suggest an unprecedented iodine(III)‐promoted enolization process. Indications that iodonium intermediates could serve as proficient Lewis acids are reported. This concept could have broad impact and foster new developments in the field of hypervalent iodine chemistry.  相似文献   

6.
μ‐Oxodiiron(III) species are air‐stable and unreactive products of autoxidation processes of monomeric heme and non‐heme iron(II) complexes. Now, the organometallic [(LNHC)FeIII‐(μ‐O)‐FeIII(LNHC)]4+ complex 1 (LNHC is a macrocyclic tetracarbene) is shown to be reactive in C?H activation without addition of further oxidants. Studying the oxidation of dihydroanthracene, it was found that 1 thermally disproportionates in MeCN solution into its oxoiron(IV) ( 2 ) and iron(II) components; the former is the active species in the observed oxidation processes. Possible cleavage scenarios for 1 are shown by scrambling experiments and structural characterization of an unprecedented adduct of 1 and oxoiron(IV) complex 2 . Kinetic analysis gave an equilibrium constant for the disproportionation of 1 , which is very small (Keq=7.5±2.5×10?8 m ). Increasing Keq might by a useful strategy for circumventing the formation of dead‐end μ‐oxodiiron(III) products during Fe‐based homogeneous oxidation catalysis.  相似文献   

7.
A bis(μ‐oxido)dinickel(III) complex was synthesized and characterized by single crystal X‐ray diffraction, resonance Raman, and ESI‐mass measurements. Magnetic susceptibility measurements by SQUID and EPR spectroscopy reveal that the complex has a triplet ground state, which is unprecedented for high‐valent metal (M) complexes with [M2(μ‐O)2] diamond core. DFT studies indicate ferromagnetic coupling of the nickel(III) centers. The complex exhibits hydrogen abstraction reactivity and oxygenation reactivity toward external substrates.  相似文献   

8.
碱性介质中, 在离子强度不变的条件下, 用广度法研究了三价银配合物氧化1,4-丁二胺的动力学及机理. 反应对三价银配合物和都是一级反应, 二级反应速率常数(k’) 随碱浓度的增大而增大,随糕碘酸根离子浓度的增大而减小. 据此提出了适合此反应的反应机理, 并计算得到了反应的热力学参数.  相似文献   

9.
A comparative kinetic study of the reactions of two mixed valence manganese(III,IV) complexes of macrocyclic ligands, [L1MnIV(O)2MnIIIL1], 1 (L1 = 1,4,8,11‐tetraazacyclotetradecane) and [L2MnIV(O)2MnIIIL2], 2 (L2 = 1,4,7,10‐tetraazacyclododecane) with thiosulfate has been carried out by spectrophotometry in aqueous buffer at 30°C. Reaction between complex 1 and thiosulfate follows a first‐order rate saturation kinetics. The pH dependency and kinetic evidences suggest the participation of two complex species of MnIII(μ‐O)2MnIV under the experimental conditions. Detailed kinetic study shows that reduction of 2 proceeds through an autocatalytic path where the intermediate (MnIII)2 species has been assumed to catalyze the reaction. The difference in the reaction mechanisms is ascribed to the difference in stability of the intermediate complex species, the evidence for which comes from the electrochemical behavior of the complexes and time dependent EPR spectroscopic measurements during the reduction of 2 . © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 36: 119–128, 2004  相似文献   

10.
The reaction of [RuIII(edta)(SCN)]2? (edta4? = ethylenediaminetetraacetate; SCN? = thiocyanate ion) with the peroxomonosulfate ion (HSO5?) has been studied by using stopped‐flow and rapid scan spectrophotometry as a function of [RuIII(edta)], [HSO5?], and temperature (15–30ºC) at constant pH 6.2 (phosphate buffer). Spectral analyses and kinetic data are suggestive of a pathway in which HSO5? effects the oxidation of the coordinated SCN? by its direct attack at the S‐atom (of SCN?) coordinated to the RuIII(edta). The high negative value of entropy of activation (ΔS = ?90 ± 6 J mol?1 deg?1) is consistent with the values reported for the oxygen atom transfer process involving heterolytic cleavage of the O‐O bond in HSO5?. Formation of SO42?, SO32?, and OCN? was identified as oxidation products in ESI‐MS experiments. A detailed mechanism in agreement with the spectral and kinetic data is presented.  相似文献   

11.
Two new complexes of the Ln2(oda)3·nH2O (oda =–O2CCH2OCH2CO2–) series are reported, i.e. {[Pr2(C4H4O5)3(H2O)3]·5H2O}n and {[Nd2(C4H4O5)3(H2O)6]·C4H6O5·‐2H2O}n. The former is isostructural with the reported La analogue, while the latter is a new structural variety within the series. Each compound exhibits two independent nine‐coordinated Ln centres showing a variety of coordination geometries.  相似文献   

12.
The two isomorphous lanthanide coordination polymers, {[Ln2(C6H4NO2)2(C8H4O4)(OH)2(H2O)]·H2O}n (Ln = Er and Tm), contain two crystallographically independent Ln ions which are both eight‐coordinated by O atoms, but with quite different coordination environments. In both crystal structures, adjacent Ln atoms are bridged by μ3‐OH groups and carboxylate groups of isonicotinate and benzene‐1,2‐dicarboxylate ligands, forming infinite chains in which the Er...Er and Tm...Tm distances are in the ranges 3.622 (3)–3.894 (4) and 3.599 (7)–3.873 (1) Å, respectively. Adjacent chains are further connected through hydrogen bonds and π–π interactions into a three‐dimensional supramolecular framework.  相似文献   

13.
The biological effect, aquation, and kinetics of oxidation of bis(2‐aminobenzothiazole)dichlorocobalt(II) complex by periodate in aqueous acidic solutions were studied. The complex exhibited a broad resistance toward the studied pathogens. The average value of the aquation constant was calculated spectrophotometrically as 2.55 × 10?5 mol2 dm?6. Kinetics of the oxidation reaction showed first‐order dependence on each reactant concentration and increased by increasing pH over the 3.80–4.80 range, 35–50°C, and decreased by increasing the ionic strength over the 0.1–0.5 mol dm?3 range. The polymerization of acrylonitrile was taken as an evidence for an inner‐sphere mechanism through the formation of free radical intermediates of Co(III) complexes, which were slowly converted to the final Co(III)products.  相似文献   

14.
In acetate buffer media (pH 4.5–5.4) thiosulfate ion (S2O32?) reduces the bridged superoxo complex, [(NH3)4CoIII(μ‐NH2,μ‐O2)CoIII(NH3)4]4+ ( 1 ) to its corresponding μ‐peroxo product, [(NH3)4CoIII(μ‐NH2,μ‐O2)CoIII(NH3)4]3+ ( 2 ) and along a parallel reaction path, simultaneously S2O32? reacts with 1 to produce the substituted μ‐thiosulfato‐μ‐superoxo complex, [(NH3)4CoIII(μ‐S2O3,μ‐O2)CoIII(NH3)4]3+ ( 3 ). The formation of μ‐thiosulfato‐μ‐superoxo complex ( 3 ) appears as a precipitate which on being subjected to FTIR shows absorption peaks that support the presence of Co(III)‐bound S‐coordinated S2O32? group. In reaction media, 3 readily dissolves to further react with S2O32? to produce μ‐thiosulfato‐μ‐peroxo product, [(NH3)4CoIII(μ‐S2O3,μ‐O2)CoIII(NH3)4]2+ ( 4 ). The observed rate (k0) increases with an increase in [TThio] ([TThio] is the analytical concentration of S2O32?) and temperature (T), but it decreases with an increase in [H+] and the ionic strength (I). Analysis of the log At versus time data (A is the absorbance of 1 at time t) reveals that overall the reaction follows a biphasic consecutive reaction path with rate constants k1 and k2 and the change of absorbance is equal to {a1 exp(–k1t) + a2 exp(–k2t)}, where k1 > k2.  相似文献   

15.
Three new μ‐oxamido‐bridged heterodinuclear copper (II)‐chromium (III) complexes formulated [Cu(Me2oxpn)Cr‐(L)2](NO3)3, where Me2oxpn denotes N,N'‐bis(3‐amino‐2, 2‐dimethylpropyl)oxamido dianion and L represents 5‐methyl‐1,10‐phenanthroline (Mephen), 4,7‐diphenyl‐1,10‐phenanthroline (Ph2phen) or 2,2′‐bipyridine (bpy), have been synthesized and characterized by elemental analyses, IR and electronic spectral studies, magnetic moments of room‐temperature and molar conductivity measurements. It is proposed that these complexes have oxamido‐bridged structures consisting of planar copper (II) and octahedral chromium (III) ions. The variable temperature magnetic susceptibilities (4.2–300 K) of complexes [Cu(Me2oxpn)Cr(Ph2phen)2](NO3)3 (1) and [Cu(Me2oxpn)Cr(Mephen)2] (NO3)3 (2) were further measured and studied, demonstrating the ferromagnetic interaction between the adjacent chromium (III) and copper (II) ions through the oxamido‐bridge in both complexes 1 and 2. Based on the spin Hamiltonian, ? = ‐ 2J?1 · ?2, the exchange integrals J were evaluated as + 21.5 an?1 for 1 and + 22.8 cm?1 for 2.  相似文献   

16.
The solid‐state structure of the title compound, [Na2Mn2(C32H56N2OSi2)2O2] or [1,8‐C10H6(NSiiPr3)2Mn(μ3‐O)Na(THF)]2, which lies across a crystallographic twofold axis, exhibits a central [Mn2O2Na2]4+ core, with two oxide groups, each triply bridging between the two MnIII ions and an Na+ ion. Additional coordination is provided to each MnIII centre by a 1,8‐C10H6(NSiiPr3)2 [1,8‐bis(triisopropylsilylamido)naphthalene] ligand and to the Na+ centres by a tetrahydrofuran molecule. The presence of an additional Na...H—C agostic interaction potentially contributes to the distortion around the bridging oxide group.  相似文献   

17.
18.
19.
The molecule of the title compound, [Mn4Al(CH3)2(C3H7O2)4I5(C4H8O)], contains one AlIII and four MnII ions. Two Mn atoms are five‐coordinate in the form of a trigonal bipyramid or a square pyramid. The two other Mn atoms are six‐coordinate with an octahedral geometry. The fourcoordinate Al atom is linked to the manganese core by μ‐Oalkoxo bridges, forming an almost planar five‐membered ring.  相似文献   

20.
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