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 156 毫秒
1.
The hydrothermal reactions of (Ph4P)[VO2Cl2] and H2C2O4 at 150 and 125°C yield (Ph4P)2[V2O2(H2O)2(C2O4)3]·4H2O (1) and (Ph4P)[VOCl(C2O4)] (2), respectively. The structure of the molecular anion of 1 consists of a binuclear unit of oxovanadium(IV) octahedra bridged by a bisbidentate oxalate group. The VO6 coordination geometry at each vanadium site is defined by a terminal oxo group, an aquo ligand, and four oxygen donors — two from the bisbidentate bridging oxalate and two from the terminal bidentate oxalate. The structure of 2 consists of discrete Ph4P+ cations occupying regions between [VOCl(C2O4)] spiral chains. The structure of the one-dimensional anionic chain exhibits V(IV) octahedra bridged by bisbidentate oxalate groups. Crystal data: 1·4H2O, monoclinic P21/n, A = 12.694(3), B = 12.531(3), C = 17.17(3) Å, β = 106.32(2)°, V = 2621.3(13) Å3, Z = 2, Dcalc = 1.501 g cm−3, structure solution and refinement converged at a conventional residual of 0.0518; 2, tetragonal P43, A = 12.145(2), C = 15.991(3) Å, V = 2358.7(12) Å3, Z = 4, R = 0.0452.  相似文献   

2.
[NBun4]2[W(C3Se5)3] (C3Se52− = 1,3-diselenole-2-selone-4,5- diselenolate(2−)) was prepared by the reaction of Na2[C3Se5] with WCl6 in ethanol, followed by addition of [NBun4]Br. The cyclic voltammogram in dichloromethane exhibits two oxidation peaks at −0.04 and +0.03 V (versus SCE). The complex reacted with [Fe(C5Me5)2][BF4], iodine or [TTF]3[BF4]2 (TTF·+ = the tetrathiafulvalenium radical cation) in acetonitrile to afford the oxidized complexes [Fe(C5Me5)2]0.5[W(C3Se5)3], [NBun4]0.1[W(C3Se5)3] and [TTF]0.5[W(C3Se5)3], respectively. Current-controlled electrochemical oxidation of the complex in acetonitrile gave [NBun4]0.6[W(C3Se5)3]. The oxidized complexes exhibit electrical conductivities of 4.7×10 −5−1.5×10−3 S cm−1 at room temperature measured for compacted pellets. Electronic absorption, IR and ESR spectra of these complexes are discussed.  相似文献   

3.
Cp#2Yb (Cp#=C5H4(CH2)2NMe2) has been obtained by reaction of YbI2(THF)2 with 2 equiv. of C5H4(CH2CH2NMe2)K in THF. The X-ray structure analysis shows a bent structure with intramolecular coordination of both nitrogen atoms to ytterbium. The reaction of C60-fullerene with Cp#2Yb leads to the formation of the fullerenide derivative [Cp#2Yb]2C60, which shows an ESR signal in the solid state and in THF solution at room temperature (solid: ΔH = 50 G, G = 1.9992; solution: ΔH = 10 G, G = 2.0001) and a magnetic moment of 3.6 BM. The lutetium fullerenides CpLu(C60)(DME) (3) and Cp*Lu(C60)(DME)(C6H5CH3) (4), (Cp = η5−C5H5, Cp* = η5−C5Me5), were obtained by reaction of C60 with CpLu(C10H8) (DME) and Cp*Lu(C10H8) (DME) in toluene. Both complexes are paramagnetic (μeff = 1.4 and 0.9 BM) and exhibit temperature-dependent ESR signals (293 K: g = 1.992 and 2.0002 respectively).  相似文献   

4.
Roger N.F. Thorneley 《BBA》1974,333(3):487-496
1. Single reduced methyl viologen (MV.+) acts as an electron donor in a number of enzyme systems. The large changes in extinction coefficient upon oxidation (λmax 600 nm; MV.+, = 1.3 · 104 M−1 · cm−1; oxidised form of methyl viologen (MV2+), = 0.0) make it ideally suited to kinetic studies of electron transfer reactions using stopped-flow and standard spectrophotometric techniques.

2. A convenient electrochemical preparation of large amounts of MV.+ has been developed.

3. A commercial stopped-flow apparatus was modified in order to obtain a high degree of anaerobicity.

4. The reaction of MV.+ with O2 produced H2O2 (k > 5 · 106 M−1 · s−1, pH 7.5, 25 °C). H2O2 subsequently reacted with excess MV.+ (k = 2.3 · 103 M−1 · s−1, pH 7.5, 25 °C) to produce water. The kinetics of this reaction were complex and have only been interpreted over a limited range of concentrations.

5. The results support the theory that the herbicidal action of methyl viologen (Paraquat, Gramoxone) is due to H2O2 (or radicals derived from H2O2) induced damage of plant cell membrane.  相似文献   


5.
Iron(II) salts in aqueous solution, or iron(III) salts in the presence of an O√2 generating system, can activate dioxygen to produce hydroxyl radicals. These are detected indirectly by their ability to degrade deoxyribose with the formation of thiobarbituric acid-reactive (TBA) products. Iron salts also catalyse the peroxidation of phospholipids resulting in the formation of TBA-reactive products. Hydroxyl radicals were responsible for the degradation of deoxyribose but not for the observed peroxidation of phospholipid. The function of O√2 in both deoxyribose degradation and phospholipid peroxidation seems to be that of reducing iron(III) into iron(II).  相似文献   

6.
In the present study, using the technique of EPR spin trapping with DMPO a spin trap, we demonstrated formation of thiyl radicals from thiol-containing angiotensin converting enzyme (ACE) inhibitor captopril (CAP) and from its stereoisomer epicaptopril (EPICAP), a non-ACE inhibitor, in the process of .OH radical scavenging. Splitting constants of DMPO/thiyl radical adducts were identical for both thiols and were aN = 15.3 G, and aH = 16.2 G. Bimolecular rate constants for the reaction of CAP and EPICAP with .OH radicals were close to a diffusion-controlled rate (≈ 2 × 1010 M−1s−1). Our data also show that both CAP and EPICAP reduce Fe(III) ions and that their respective thiyl radicals are formed in this reaction. In the presence of Fe(III), H2O2, and CAP, or EPICAP, .OH radicals were produced by a thiol-driven Fenton mechanism. Copper(II) ions were also reduced by these thiols, but no thiyl radicals could be detected in these reactions, and no .OH or other Fenton oxidants were observed in the presence of H2O2. Our data show direct evidence that thiol groups of CAP and EPICAP are involved in scavenging of .OH radicals. The direct .OH radical scavenging, together with the reductive “repair” of other sites of .OH radical attack, may contribute to the known protective effect of CAP against ischemia/reperfusion-induced arrhythmias. The formation of reactive thiyl radicals in the reactions of the studied compounds with .OH radicals and with Fe(III) ions may play a role in some of the known adverse effects of CAP.  相似文献   

7.
Many copper and iron complexes can be reduced by O-2 as well as by H2O2. According to the rates of reduction and the concentration of O-2 and H2O2, the metal complexes may serve either as catalyst of O-2 dismutation or as catalysts of the reaction between O-2 and H2O2 to form OH' radical (Haber-Weiss reaction). Various factors which influence whether metal complexes protect the biological systems from superoxide toxicity or enhance it are discussed.  相似文献   

8.
Three-dimensionally (3D) ordered macroporous active carbon has been fabricated and used as electrode substrate for the direct electrochemistry of horse heart cytochrome c (Cyt c). The Cyt c immobilized on the surface of the ordered macroporous active carbon shows a pair of well-defined and nearly reversible redox waves at the formal potential of −0.033 V in pH 6.8 phosphate buffer solution. The interaction between Cyt c and the 3D macroporous active carbon makes the formal potential shift negatively compared to that of Cyt c in solution. Spectrophotometric and electrochemical methods have been used to investigate the interaction between Cyt c and the porous active carbon. The immobilized Cyt c maintains its biological activity, and shows a surface controlled electrode process with the electron-transfer rate constant (ks) of 17.6 s−1 and the charge-transfer coefficient (a) of 0.52, and displays the features of a peroxidase in the electrocatalytic reduction of hydrogen peroxide (H2O2). A potential application of the Cyt c-immobilized porous carbon electrode as a biosensor to monitor H2O2 has been investigated. The steady-state current response increases linearly with H2O2 concentration from 2.0 × 10−5 to 2.4 × 10−4 mol l−1. The detection limit (3σ) for determination of H2O2 has been found to be 1.46 × 10−5 mol l−1.  相似文献   

9.
The kinetics and equilibria of complex formation by Ga(III) with NCS in aqueous solution have been measured over a range of acidities and temperatures, the contributing paths to the reaction resolved, and their rate constants and activation parameters determined. The hydrolysis equilibria required to carry out this resolution of kinetic behaviour have also been measured.

Unlike the other reported complexation reactions of Ga(III) in aqueous solution, the separate reaction pathways can be assigned with no ambiguity. At 25 °C and ionic strength 0.5 M, the observed forward rate constant for the complex formation is described by {k1 + k2K1h/[H+] + k3K1hK2h/[H+]2} M−1 s−1. For these conditions, the first and second successive hydrolysis constants of Ga(H2O)63+ are given by pK1h = 3.69 ± 0.01 and pK2h = 3.74 ± 0.04. The rate constants corresponding to the reactions of the species Ga(H2O)63+, Ga(H2O)5(OH)2+ and Ga(H2O)4(OH)2+ with NCS are k1 = 57 ± 4 M−1 −1, k2 = (1.08 ± 0.01) × 105 M−1 s−1 and k3 = 3 × 106 M−1 s−1 respectively. The complexation equilibrium quotient [GaNCS2+]/([Ga3+][NCS]) has been independently determined by spectrophotometric titration to be 20.8 ± 0.3 M−1 at 25 °C and ionic strength 0.5 M.

These kinetic results lead to an interpretation of the data, and a reinterpretation of other data for aquo-Ga(III) complex formation kinetics from the literature which support the assignment of a dissociative interchange mechanism for these reactions rather than the associative activation mode sometimes proposed.  相似文献   


10.
The aqueous chemistry of vanadium with physiologically relevant ligands constitutes a subject of burgeoning research, extending from bacterial metalloenzymic functions to human-health physiology. Vanadium, in the form of VCl3 and V2O5, reacted expediently with citric acid, in a 1:2 molar ratio in water at pH4, and, in the presence of various cations, afforded crystalline materials bearing the general formula (Cat)2[V2O4(C6H6O7)2nH2O (A) (Cat+=Na+, NH4 +, n=2; Me4N+, K+, n=4). Exploration of the reactivity of A toward H2O2 yielded the peroxo-containing complexes (Cat)2[V2O2(O2)2(C6H6O7)2]·2H2O (B) (Cat+=K+, NH4 +). Both classes of compounds were characterized analytically and spectroscopically. The X-ray structures of complexes A and B emphasize the exceptional stability of the dimeric rhombic unit V2O2, which is retained upon H2O2 reaction, and the preserved mode of coordination of the citrate ligand as a doubly deprotonated moiety. In these complexes, typical six and eight coordination numbers were observed for the Na+ and K+ counter-ions, respectively. The variety of synthetic approaches leading to A, along with the stepwise and direct assembly and isolation of peroxo-compounds (B), denotes the significance of reaction pathways and intermediates in vanadium(III–V)–citrate synthetic chemistry. Hence, a systematic investigation of reactivity modes in aqueous vanadium–citrate systems emerges as a crucial tool for the establishment of chemical interconnectivity among low MW complex species, potentially participating in the intricate biodistribution of that metal ion in biological fluids.  相似文献   

11.
[RuII(Me2edda)(H2O)2] (1), Me2edda2− = N,N′-dimethylethylenediaminediacetate, exhibits a sterically-controlled molecular recognition in forming η2 and η4 olefin complexes. 1 exists with an N2O2 in-plane set of chelate donors and axial H2O ligands. The two CH3 functionalities of Me2edda2− are poised above and below the N2O2 plane of the glycinato rings. Studies herein of the 2,2′-bipyridine complex, [RuII(Me2edda)(bpy)], with bidentate bpy chelation as established via 1H NMR and electrochemical methods show 1 to be ligated in the S,S configuration with the glycinato rings in-plane as a cis-O form. 1 is sterically discriminating in forming η2 complexes with smaller olefins (ethylene, 2-propene, cis-2-butene, methyl vinyl ketone and 3-cyclohexene-1-methanol), but rejects larger decorated ring structures and branched olefins (1,2-dimethyluracil, cyclohexene-1-one 2-methyl-2-propene). η2 complexes of 1 have characteristic RuII/III DPP waves near 0.55 V which vary slightly with olefin structure. Potentially bidendate dienes (1,3-butadiene, 1,3-cyclohexadiene and 2,5-norbornadiene (nbd) form η4 complexes as shown by RuII/III waves between 0.94 and 1.30 V, indicate of a highly stabilized RuII center by π-backboning. An η2η4 ‘equilibrium’ with apparent K = 22 at 25 °C is observed for nbd coordinated to 1. (The η2 and η4 distribution may be a kinetic one and not a thermodynamic one). To allow formation of the cis η4 complexes, 1 must undergo a shift of one or both glycinato donors from the N2O2 plane into the axial site away from the dimethyl functionalities. η4 chelation by 1,3-butadiene has been confirmed by 1H NMR spectral assignments of two [RuII(Me2edda)] isomers, one in the axial rans-O glycinato configuration, e.g. 1,3-butadiene is bidentate in the original N2O2 plane and a second unsymmetrical glycinato arrangement with in-plane and axial glycinato as well as in-plane and axial η4-1,3-butadiene coordination. [RuII(hedta)(H2O)] (2), hedta3− = N-hydrpxyethylenediaminetriacetate, is less discriminating for olefin structures, forming η2 complexes with all eleven olefins and dienes mentioned for studies with 1. However, 2 does not undergo displacement of a carboxylate donor by the second olefin unit of a diene [RuII(hedta)(diene)] complexes possess a pendant non-coordinated olefin and on η2-bound olefin in the complex, indicated by a normal RuII(pac)(olefin)RuII/III wave near 0.55 V.  相似文献   

12.
The binuclear cyanoferrate, tetraphenylphosphonium pentacyanoiron(III)-μ-cyano-amminetetracyanoiron(III), [(C6H5)4P]4[Fe2(CN)10NH3]4−, was synthesized by air oxidation of aqueous solutions of Na3[Fe(CN)5NH3] · 3H2O. Single crystal X-ray diffraction studies show the compound to contain the binuclear, cyano-bridged anion, [(NC)5Fe---NC---Fe(CN)4NH3]4−. This compound is structurally identical to the one prepared by A. Ludi et al., [Inorg. Chim. Acta, 34, 113 (1979)], with the exception that [Fe(CN)6]3− is not required for the synthesis of this compound. The Fe(III) atoms are antiferromagnetically coupled through the CN bridge, as shown by a maximum in the magnetic susceptibility at 50 K. The electronic and IR spectra of the complex in the solid state and in solution are discussed.  相似文献   

13.
The reaction of meso-tetrakis (4-dimethoxyphenyl) porphinatomanganese(II), MnTPOMeP, with TCNE (TCNE = tetracyanoethylene) leads to the formation of [MnTPOMeP]+ [TCNE] and [MnTPOMeP]+[OC(CN)C(CN)2]. The single-crystal X-ray structures of the latter as well as [Cu(bipy)2Cl]+ [OC(CN)C(CN)2] were determined. The former has a disordered [OC(CN)C(CN)2] bridging via C and O between a pair of MnIII sites, whereas the latter has an isolated [OC(CN)C(CN)2] unbound to CuII. The IR characterization for μ2-C,O bound [OC(CN)C(CN)2] is at 2219m and 2196s (νCN) cm−1 and at 1558s (νCO) cm−1 while for unbound [OC(CN)C(CN)2] it is at 2210m, 2203m, 2181m (νCN) cm−1 and at 1583s (νCO) cm−1.  相似文献   

14.
Two compounds, [Eu(H2O)7][Al(OH)6Mo6O18] · 4H2O (1) and {(C2H5NO2)2[Eu(H2O)5]}[Al(OH)6Mo6O18] · 10H2O (2), have been synthesized by conventional solution method and determined by single-crystal X-ray diffraction. Compound 1 shows a 1D chain structure built up of alternating Anderson-type polyanions [Al(OH)6Mo6O18]3− and hydrated rare-earth ions Eu3+. Compound 2 displays a 3D supramolecular network structure containing 1D sandglass-like channels along c axis, which were occupied by repetitive array of (H2O)8 clusters. Extensive hydrogen bonds play an important role in the formation of the 3D structures of 1 and 2. Luminescence measurements reveal that 1 and 2 exhibit intense red and orange fluorescent emission at room temperature, respectively. Origin of the distinct emission can be assigned to the different site symmetries of Eu3+ centers in the two compounds. These results are consistent with the crystal structures of the two compounds.  相似文献   

15.
Culture (NAD(P)H) fluorescence dynamics have been used to provide information on culture behaviour when Xanthomonas campestris was grown in a bioreactor. Culture fluorescence decreased by 1150 units in response to an increase in extracellular pH from 3.1 to 7.6. A mathematical model incorporating the effect of pH on the bulk NADH depletion reaction simulated the experimental data. The rates of bulk NADH formation and depletion reactions were 1 s−1 and 719 (M h−1)−1 s−1, respectively. Subsequent to the initial NADH decrease, the culture fluorescence increased to within 200 units of its original value, with a concomitant decrease in oxygen uptake rate (OUR) from 7.3 to 3 mM h−1. A mathematical model incorporating the hypothesis that the culture manipulated its OUR to increase its NADH level, simulated the experimental data. In addition, it was inferred from culture fluorescence that the intracellular oxygen availability becomes insufficient at or below 10% extracellular dissolved oxygen value. Studies on H2O2 addition to X. campestris, to optimize the liquid-phase oxygen supply, showed no change in metabolic state, as indicated by NADH fluorescence, until 1.4 mmol H2O2 (g cell)−1 and a significant decrease above that. Investigations on the reasons for decreases in NADH fluorescence suggested a DNA-damaging Fenton reaction as the probable reason for the observed NADH decrease on addition of H2O2.  相似文献   

16.
An electron-rich iron(III) porphyrin complex (meso-tetramesitylporphinato)iron(III) chloride [Fe(TMP)Cl], was found to catalyze the epoxidation of olefins by aqueous 30% H2O2 when the reaction was carried out in the presence of 5-chloro-1-methylimidazole (5-Cl-1-MeIm) in aprotic solvent. Epoxides were the predominant products with trace amounts of allylic oxidation products, indicating that Fenton-type oxidation reactions were not involved in the olefin epoxidation reactions. cis-Stilbene was stereospecifically oxidized to cis-stilbene oxide without giving isomerized trans-stilbene oxide product, demonstrating that neither hydroperoxy radical (HOO·) nor oxoiron(IV) porphyrin [(TMP)FeIV=O] was responsible for the olefin epoxidations. We also found that the reactivities of other iron(III) porphyrin complexes such as (meso-tetrakis(2,6-dichlorophenyl)porphinato)iron(III) chloride [Fe(TDCPP)Cl], (meso-tetrakis(2,6-difluorophenyl)porphinato)iron(III) chloride [Fe(TDFPP)Cl], and (meso-tetrakis(pentafluorophenyl)porphinato)iron(III) chloride [Fe(TPFPP)Cl] were significantly affected by the presence of the imidazole in the epoxidation of olefins by H2O2. These iron porphyrin complexes did not yield cyclohexene oxide in the epoxidation of cyclohexene by H2O2 in the absence of 5-Cl-1-MeIm in aprotic solvent; however, addition of 5-Cl-1-MeIm to the reaction solutions gave high yields of cyclohexene oxide with the formation of trace amounts of allylic oxidation products. We proposed, on the basis of the results of mechanistic studies, that the role of the imidazole is to decelerate the O–O bond cleavage of an iron(III) hydroperoxide porphyrin (or H2O2–iron(III) porphyrin adduct) and that the intermediate transfers its oxygen to olefins prior to the O–O bond cleavage.  相似文献   

17.
The kinetics of the anation reactions of [M(RNH2)5H2O]3+ (M = Rh, R = H, Me, Et, Pr; M = Cr, R = H, Me, Pr) with several ligands (H3PO4/H2PO4, H3PO3/H2PO3, CF3COO, Br, Cl, SCN) have been studied at different temperatures and acidities at I = 1.0 M (LiClO4. Results obtained for the anation rate constants and thermal activation parameters are compared with the previously published data for R = H, in order to establish the effects of the amine substituents in the reaction mechanism proposed for the substitution reactions of these complexes. The results obtained are interpreted on the basis of a mechanism where the bond formation process is more important in the substitution on M = Cr complexes than in that of the M = Rh complexes, as already pointed out for the published ΔΛ values for the water exchange on these systems. A simple Langford-Gray classification becomes inadequate to describe these situations where the increase of the steric demand of the amine substituents shifta the Ia-Id classification to the Id side, although no dramatic changes in the reaction mechanism are found. It is concluded that a More O'Ferall ‘continuous’ type of approach to the mechanism classification of the substitution reactions is much more useful in this case.  相似文献   

18.
Oxygen radical scavengers have been shown to prevent the development of ischemic preconditioning, suggesting that reactive oxygen species (ROS) might be involved in this phenomenon. In the present study, we have investigated whether direct exposure to ROS produced by photoactivated Rose Bengal (RB) could mimic the protective effects of ischemic preconditioning.

Methods In vitro generation of ROS from photoactivated RB in a physiological buffer was first characterised by ESR spectroscopy in the presence of 2,2,6,6-tetramethyl-1-piperidone (oxoTEMP) or 5,5-dimethyl-1-pyrroline-N-oxide (DMPO). In a second part of the study, isolated rat hearts were exposed for 2.5 min to photoactivated RB. After 5 min washout, hearts underwent 30 min no-flow normothermic ischemia followed by 30 min of reperfusion.

Results and Conclusions The production of singlet oxygen (1O2) by photoactivated RB in the perfusion medium was evidenced by the ESR detection of the nitroxyl radical oxoTEMPO. Histidine completely inhibited oxoTEMPO formation. In addition, the use of DMPO has indicated that (i) superoxide anions (O·-2) are produced directly and (ii) hydroxyl radicals (HO·) are formed indirectly from the successive O·-2 dismutation and the Fenton reaction. In the perfusion experiments, myocardial post-ischemic recovery was dramatically impaired in hearts previously exposed to the ROS produced by RB photoactivation (1O2, O·-2, H2O2 and HO·) as well as when 1O2 was removed by histidine (50 mM) addition. However, functional recovery was significantly improved when hearts were exposed to photoactivated RB in presence of superoxide dismutase (105 IU/L) and catalase (106 IU/L).

Further studies are now required to determine whether the cardioprotective effects of Rose Bengal in presence of O·-2 and H2O2 scavengers are due to singlet oxygen or to other species produced by Rose Bengal degradation.  相似文献   

19.
[MnL](ClO4)2 (L = N,N′,N″-tris(2-hydroxypropyl)-1,4,7-triazacyclononane) has been tested for catalyzing sulfide oxidation. In the presence of this complex, ethyl phenyl sulfide, butyl sulfide and phenyl sulfide are completely oxidized to the corresponding sulfoxides and sulfones with H2O2 as the oxidant. 2-Chloroethyl phenyl sulfide oxidation yield 2-chloroethyl phenyl sulfone and phenyl vinyl sulfone. In ethyl phenyl sulfide oxidation, effects of complex and H2O2 concentration and temperature on the reaction rate have been discussed. Through controlling reaction conditions, ethyl phenyl sulfoxide and ethyl phenyl sulfone may be produced selectively. The UV–Vis and electron paramagnetic resonance (EPR) studies on catalyst solution indicate that metal centre of the complex is transformed from Mn(II) to Mn(IV) after the addition of H2O2. At 25 °C, rate constant for ethyl phenyl sulfide oxidation is 4.38 × 10−3 min−1.  相似文献   

20.
The interaction of horse ferricytochrome c with the reagents [Fe(EDTA)(H2O)] and [Cr(CN)6]3− were studied at pH 7 and 25°C by 1H-NMR spectroscopy. Two binding regions near to the heme crevice of cytochrome c were identified. Both regions bound both reagents but they exhibited different selectivities.

The relevance of this finding to the electron-transfer function of cytochrome c is discussed.  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