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 31 毫秒
1.
《Inorganica chimica acta》1988,147(1):103-107
The electron spin resonance (ESR) spectrum of the reaction product of superoxide ion, O2, with chloro-5,10,15,20-tetraphenylporphyrinatochromium(III) [Cr(III)(TPP)Cl] shows strong hyperfine interactions with the metal nucleus and the metal ligand, indicating the formation of a superoxide adduct, Cr(IV)(TPP)(Cl)(O2). The formation of this superoxide adduct was also confirmed by UV-Vis spectroscopy. The reactive character of this superoxide adduct was investigated by ESR spectrometry. It was found that Cr(IV)(TPP)(Cl)(O2) can oxidize t-butylamine and triphenylphosphine to give the corresponding radical species, respectively, but not pyridine, 4-cyanopyridine or imidazole. These results indicate that the reactive character of Cr(IV)(TPP)- (Cl)(O2) resembles that of the free superoxide ion.  相似文献   

2.
Camptothecin (CPT) is an anticancer drug that inhibits topoisomerase I (Topo I) by forming a ternary DNA-CPT-Topo I complex. However, it has also been shown that UVA-irradiated CPT in the absence of Topo I produces significant DNA damage to cancer cells. In this work, we explored and identified free radicals generated in these processes. From the low-temperature EPR spectrum of Cu(II)-CPT complex, a proximity between Cu(II) ion and 20-hydroxy group of lactone E ring of CPT is proposed. Upon irradiation (λ = 365 nm) of the Cu(II)-CPT complex in de-oxygenated dimethylsulfoxide (DMSO), the EPR signal of Cu(II) measured in situ at room temperature shows formal first-order exponential decay with a formal half-life of 11 min. By the use of a specific Cu(I) chelating agent, neocuproine, it was shown that, during this process, Cu(II) is reduced to Cu(I). The loss in EPR signal intensity of the Cu(II)-CPT complex upon irradiation is accompanied by the appearance of a new EPR signal at g ≈ 2.0022. Application of the spin trap nitrosodurene (ND) revealed that the main radical product formed upon continuous irradiation of CPT in DMSO solutions is the hydroxyl radical (trapped in DMSO as the CH3 adduct) and superoxide radical. Application of 2,2,6,6-tetramethyl-4-piperidinol has revealed that irradiation of CPT in aerated DMSO solution also leads to formation of singlet oxygen (1O2). Our spectroscopic experiments indicate that CPT is a promising photosensitizer and that radicals and singlet oxygen generated upon illumination play a central role in DNA cleavage and in the induction of apoptosis in cancer cells.  相似文献   

3.
Abstract

We previously reported that irradiation of titanium dioxide (TiO2) in ethanol generates both singlet oxygen (1O2) and superoxide anion (O2·-) as measured by EPR spectroscopy. The present study describes the production of reactive oxygen species upon irradiation of TiO2 in aqueous suspension as determined by EPR spectroscopy using 2,2,6,6-tetramethyl-4-piperidone (4-oxo-TMP) and 5,5- dimethyl-pyrroline-N-oxide (DMPO). Photoproduction of 1O2 by suspended TiO2, detected as 2,2,6,6-tetramethyl-4-piperidone-N-oxyl (4-oxo-TEMPO), was measured in water and deuterium oxide (D2O) in the presence or absence of sodium azide (NaN3) and under air or argon atmospheres. Production of a DMPO-OH adduct was examined in 4-oxo-TMP containing medium in the presence or absence of dimethyl sulfoxide (DMSO). The signal for the DMPO spin adduct of superoxide anion was not observed in aqueous conditions. Kinetic analysis revealed that 1O2 was produced at the surface of irradiated TiO2 in aqueous suspension as was observed in ethanol. Kinetic analysis revealed that the formation of DMPO-OH adduct reflects oxidation of DMPO by 1O2 rather than the trapping of the hydroxyl radical produced by the reaction of photo-exited TiO2 and water. The production of large amounts of 1O2 by TiO2 in aqueous suspension as compared to those in ethanol and possible formation of hydroxyl radical in aqueous suspension but not in alcohol, suggest that irradiation of TiO2 in aqueous environments is biologically more important than that in non-aqueous media.  相似文献   

4.
The complex formation of Co(II) with N-donor ligands in dimethylsulfoxide (DMSO) is investigated by means of calorimetric and spectroscopic methods. The ligands considered in this work are tripodal polyamines and polypyridines: 2,2′,2′′-triaminotriethylamine (TREN), tris(2-(methylamino)ethyl)amine (Me3TREN), tris(2-(dimethylamino)ethyl)amine (Me6TREN), tris[(2-pyridyl)methyl]amine (TPA) and 6,6′-bis-[bis-(2-pyridylmethyl)aminomethyl]-2,2′-bipyridine (BTPA).These ligands are characterized by a systematic modification of the donor groups in order to study how their structure is related to the stability of the complexes formed and to their ability to bind oxygen. A comparison with thermodynamic data for similar Cd(II) systems as well as with data referred to linear tetra-amines in DMSO is also made. The solvent effect on the nature and stability of the species formed is discussed. DFT calculations are carried out to explain the trend in thermodynamic parameters for Me6TREN. Only Co(TREN)2+ is able to bind oxygen and two successive species (μ-superoxo and μ-peroxo) are formed. The kinetics of oxygen uptake by Co(TREN)2+ is found to be less solvent-dependent than other Co(II)-polyamine complexes when the formation of the mononuclear μ-superoxo complex is considered.  相似文献   

5.
Strong evidence suggests that the stretching vibration of the bound oxygen can be perturbed by an accidentally degenerate porphyrin ring mode, resulting in two split frequencies. In the Co(II)(TpivPP) (pyridine) 18O2 complex, we demonstrate that the ν(18O—18O) mode, after being shifted from its ν(16O—16O) value at 1,156 cm-1, undergoes a resonance interaction with the 1,080 cm-1 porphyrin mode, giving rise to two lines at 1,067 and 1,089 cm-1. In the O2 complex of Co(II) mesoporphyrin IX-substituted sperm whale myoglobin, we observed a dramatic intensity increase at 1,132 cm-1 upon 16O218O2 substitution, which is due to the reappearance of the 1,132-cm-1 porphyrin mode after the removal of resonance conditions. A decrease in O2 binding affinity, caused by the proximal base tension, corresponds to an increase in the Co—O2 stretching frequency. The ν(Co—O2) at 527 cm-1 for the low affinity Co(II)(TpivPP)(1,2-Me2Im) O2 complex is 11 cm-1 higher than the 516-cm-1 value for the high affinity complex (with N-MeIm replacing 1,2-Me2Im). However, in the corresponding iron complexes the reverse behavior is observed, i.e., the ν(Fe—O2) decreases for the (1,2-Me2Im) complex. There is a 24-cm-1 difference in the Co—O2 stretching frequencies between Co(II)(TpivPP)(N-MeIm)O2 (at 516 cm-1) and oxy meso CoMb (at 540 cm-1), suggesting a protein induced distortion of the Co—O—O linkage. However, the values for ν(Fe—O2) are nearly identical between Fe(II)(TpivPP)(N-MeIm)O2 (at 571 cm-1) and oxy Mb (at 573 cm-1), indicating that O2 binds to myoglobin in the same manner as in the sterically unhindered “picket fence” complex. Evidence is presented that suggests the presence of two dioxygen stretching frequencies due to two different conformers in each of the N-MeIm and 1,2-Me2Im complex of oxy Co(II)(TpivPP).  相似文献   

6.
H2O2 production by skeletal muscle mitochondria oxidizing palmitoylcarnitine was examined under two conditions: the absence of respiratory chain inhibitors and the presence of myxothiazol to inhibit complex III. Without inhibitors, respiration and H2O2 production were low unless carnitine or malate was added to limit acetyl-CoA accumulation. With palmitoylcarnitine alone, H2O2 production was dominated by complex II (44% from site IIF in the forward reaction); the remainder was mostly from complex I (34%, superoxide from site IF). With added carnitine, H2O2 production was about equally shared between complexes I, II, and III. With added malate, it was 75% from complex III (superoxide from site IIIQo) and 25% from site IF. Thus complex II (site IIF in the forward reaction) is a major source of H2O2 production during oxidation of palmitoylcarnitine ± carnitine. Under the second condition (myxothiazol present to keep ubiquinone reduced), the rates of H2O2 production were highest in the presence of palmitoylcarnitine ± carnitine and were dominated by complex II (site IIF in the reverse reaction). About half the rest was from site IF, but a significant portion, ∼40 pmol H2O2·min−1·mg protein−1, was not from complex I, II, or III and was attributed to the proteins of β-oxidation (electron-transferring flavoprotein (ETF) and ETF-ubiquinone oxidoreductase). The maximum rate from the ETF system was ∼200 pmol H2O2·min−1·mg protein−1 under conditions of compromised antioxidant defense and reduced ubiquinone pool. Thus complex II and the ETF system both contribute to H2O2 productionduring fatty acid oxidation under appropriate conditions.  相似文献   

7.
Dehydrogenases that use ubiquinone as an electron acceptor, including complex I of the respiratory chain, complex II, and glycerol-3-phosphate dehydrogenase, are known to be direct generators of superoxide and/or H2O2. Dihydroorotate dehydrogenase oxidizes dihydroorotate to orotate and reduces ubiquinone to ubiquinol during pyrimidine metabolism, but it is unclear whether it produces superoxide and/or H2O2 directly or does so only indirectly from other sites in the electron transport chain. Using mitochondria isolated from rat skeletal muscle we establish that dihydroorotate oxidation leads to superoxide/H2O2 production at a fairly high rate of about 300 pmol H2O2·min−1·mg protein−1 when oxidation of ubiquinol is prevented and complex II is uninhibited. This H2O2 production is abolished by brequinar or leflunomide, known inhibitors of dihydroorotate dehydrogenase. Eighty percent of this rate is indirect, originating from site IIF of complex II, because it can be prevented by malonate or atpenin A5, inhibitors of complex II. In the presence of inhibitors of all known sites of superoxide/H2O2 production (rotenone to inhibit sites in complex I (site IQ and, indirectly, site IF), myxothiazol to inhibit site IIIQo in complex III, and malonate plus atpenin A5 to inhibit site IIF in complex II), dihydroorotate dehydrogenase generates superoxide/H2O2, at a small but significant rate (23 pmol H2O2·min−1·mg protein−1), from the ubiquinone-binding site. We conclude that dihydroorotate dehydrogenase can generate superoxide and/or H2O2 directly at low rates and is also capable of indirect production at higher rates from other sites through its ability to reduce the ubiquinone pool.  相似文献   

8.
Kinetic evaluation of the oxidation of oxymyoglobin (MbO2) to metmyoglobin (Mb+) by bis(dimethylglyoximato)cobalt nitrosyl [Co(NO)(DMGH)2] has established that the mechanism of this transformation involves initial dissociation of nitric oxide from Co(NO)(DMGH)2, followed by direct oxidation of MbO2 by nitric oxide. Nitrate formation accompanies the production of Mb+ and is proposed to arise from isomerization of the initially formed peroxynitrite ion. By comparative kinetic determinations with nitrosyl transfer from the cobalt nitrosyl reagent to deoxyhemoglobin, the rate constant for oxidation of MbO2 by nitric oxide is calculated to be 31 X 106 M?1sec?1 at 10.0°C in phosphate-buffered media at pH 7.0. Bis(dimethylglyoximato)cobalt(II), the cobalt complex formed by nitric oxide dissociation from Co(NO)(DMGH)2, is an effective trap for dioxygen liberated from MbO2. The resulting μ-peroxo- or μ-superoxo-dicobaloxime(III) oxidizes deoxymyoglobin to metmyoglobin at a rate that is competitive with oxidation induced by Co(NO)(DMGH)2.  相似文献   

9.
The reactions between superoxide free radical anion (.O2) with the halocarbons CCl4, CHCl3, BrCH2CH2Br(EDB), decachloro-biphenyl (DCBP), and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in dimethyl sulphoxide (DMSO) results in the emission of chemiluminescence (CL). The chemiluminescence reactions are characterized as having biphasic second order kinetics, CL wavelengths between 350 nm and 650 nm, and exhibiting perturbation by chemicals reactive with singlet oxygen. These data suggest that singlet oxygen species are the excited state responsible for the light emissions. Polarographic studies confirm .O2 consumption and halide release in the reactions, while gas liquid chromatography and NBT reduction demonstrate the decomposition of the halocarbons into products. A chemiluminescent reaction mechanism is proposed involving reductive dehalogenation of the halocarbons and the generation of singlet oxygen. The significance of singlet oxygen generation is discussed with respect to a general mechanism for explaining the rapid initiation of lipid peroxidative membrane damage in halocarbon toxigenicity in animal and plant tissues.  相似文献   

10.
Copper amine oxidases (CAOs) are a large family of proteins that use molecular oxygen to oxidize amines to aldehydes with the concomitant production of hydrogen peroxide and ammonia. CAOs utilize two cofactors for this reaction: topaquinone (TPQ) and a Cu(II) ion. Two mechanisms for oxygen reduction have been proposed for these enzymes. In one mechanism (involving inner-sphere electron transfer to O2), Cu(II) is reduced by TPQ, forming Cu(I), to which O2 binds, forming a copper–superoxide complex. In an alternative mechanism (involving outer-sphere electron transfer to O2), O2 is directly reduced by TPQ, without reduction of Cu(II). Substitution of Cu(II) with Co(II) has been used to distinguish between the two mechanisms in several CAOs. Because it is unlikely that Co(II) could be reduced to Co(I) in this environment, an inner-sphere mechanism, as described above, is prevented. We adapted metal replacement methods used for other CAOs to the amine oxidase from pea seedlings (PSAO). Cobalt-substituted PSAO (CoPSAO) displayed nominal catalytic activity: k cat is 4.7% of the native k cat, and K M (O2) for CoPSAO is substantially (22-fold) higher. The greatly reduced turnover number for CoPSAO suggests that PSAO uses the inner-sphere mechanism, as has been predicted from 18O isotope effect studies (Mukherjee et al. in J Am Chem Soc 130:9459–9473, 2008), and is catalytically compromised when constrained to operate via outer-sphere electron transfer to O2. This study, together with previous work, provides strong evidence that CAOs use both proposed mechanisms, but each homolog may prefer one mechanism over the other.  相似文献   

11.
Although different routes for the S-nitrosation of cysteinyl residues have been proposed, the main in vivo pathway is unknown. We recently demonstrated that direct (as opposed to autoxidation-mediated) aerobic nitrosation of glutathione is surprisingly efficient, especially in the presence of Mg2+. In the present study we investigated this reaction in greater detail. From the rates of NO decay and the yields of nitrosoglutathione (GSNO) we estimated values for the apparent rate constants of 8.9±0.4 and 0.55±0.06 M−1 s−1 in the presence and absence of Mg2+. The maximum yield of GSNO was close to 100% in the presence of Mg2+ but only about half as high in its absence. From this observation we conclude that, in the absence of Mg2+, nitrosation starts by formation of a complex between NO and O2, which then reacts with the thiol. Omission of superoxide dismutase (SOD) reduced by half the GSNO yield in the absence of Mg2+, demonstrating O2 formation. The reaction in the presence of Mg2+ seems to involve formation of a Mg2+•glutathione (GSH) complex. SOD did not affect Mg2+-stimulated nitrosation, suggesting that no O2 is formed in that reaction. Replacing GSH with other thiols revealed that reaction rates increased with the pKa of the thiol, suggesting that the nucleophilicity of the thiol is crucial for the reaction, but that the thiol need not be deprotonated. We propose that in cells Mg2+-stimulated NO/O2-induced nitrosothiol formation may be a physiologically relevant reaction.  相似文献   

12.
The biological effects of ultraviolet radiation (UV), such as DNA damage, mutagenesis, cellular aging, and carcinogenesis, are in part mediated by reactive oxygen species (ROS). The major intracellular ROS intermediate is hydrogen peroxide, which is synthesized from superoxide anion (O2) and further metabolized into the highly reactive hydroxyl radical. In this study, we examined the involvement of mitochondria in the UV‐induced H2O2 accumulation in a keratinocyte cell line HaCaT. Respiratory chain blockers (cyanide‐p‐trifluoromethoxy‐phenylhydrazone and oligomycin) and the complex II inhibitor (theonyltrifluoroacetone) prevented H2O2 accumulation after UV. Antimycin A that inhibits electron flow from mitochondrial complex III to complex IV increased the UV‐induced H2O2 synthesis. The same effect was seen after incubation with rotenone, which blocks electron flow from NADH‐reductase (complex I) to ubiquinone. UV irradiation did not affect mitochondrial transmembrane potential (ΔΨm). These data indicate that UV‐induced ROS are produced at complex III via complex II (succinate‐Q‐reductase). J. Cell. Biochem. 80:216–222, 2000. © 2000 Wiley‐Liss, Inc.  相似文献   

13.
Three new compounds are reported with the tetradentate ligand (N,N′-bis(2-Pyridylmethyl)-1,3-propanediamine) (abbreviated as pypn), two mononuclear compounds i.e. [Co(pypn)(C2O4)](ClO4) (1), [Mn(pypn)(C2O4)](ClO4) (2) and one dinuclear compound [Ni2(pypn)2(C2O4)](ClO4)2(C2H6O)1/4(H2O) (3). In the Co(III) and Mn(II) complexes the oxalate behaves as bidentate ligand, chelating the metal in the O,O′ mode, whereas in the Ni(II) compound the oxalate behaves as tetradentate ligand binding each Ni(II) ion by two oxygen atoms and bridging the two metallic centers.The synthesis, X-ray crystal structure of all three compounds and their spectroscopic properties are presented in detail. The geometry around the Co3+, Mn3+, Ni2+ ions is essentially octahedrally based, while the stabilization of the crystal lattice in all cases is maintained by interesting hydrogen bond systems.  相似文献   

14.
Freshly-added iron only slightly affected the growth of iron-sufficient cells of the green alga Scenedesmus incrassatulus Bohl, strain R-83, but induced accumulation of malondialdehyde (MDA) in cells and excretion of MDA in the medium. These effects were stronger in response to Fe2+ as compared to Fe3+, but Fe3+ induced the release of more iron-binding chelators from these cells than Fe2+. Fe3+ added either in dark or in light induced release of equal concentrations of iron-complexing agents, part of which formed strong chelates with iron in the medium. Exogenously added hydrogen peroxide inhibited iron-induced release of chelators but the effect was removed by addition of the hydroxyl radical scavenger dimethylsulfoxide (DMSO). Malondialdehyde also inhibited the release of chelators. Release of chelators was induced in the absence of iron salts by photoexcited chlorophyll (Chl). The Chl-induced release was efficiently inhibited by singlet oxygen scavengers such as dimethylfuran, -carotene, sodium azide and vitamin B6, and stimulated in D2O or DMSO. Exogenously added catalase inhibited the release more than added superoxide dismutase. The Fe3-induced release of chelators was also inhibited by scavengers of singlet oxygen, but was not affected by sodium azide and by ethanol. Hence both H2O2 and singlet oxygen were involved in induction of chelator release in the absence of iron in light. The induction of chelator release by iron in dark involved H2O2, but not singlet oxygen.  相似文献   

15.
《Inorganica chimica acta》1988,142(2):223-228
The Co(EDTA) complex in aqueous solution gives rise to a specific interaction with I ions as evidenced by a new, relatively intense band formed at 290–300 nm. This specific interaction is attributed to the formation of an ion-pair between Co(EDTA) and I, even though they are like charged ions. Irradiation of this ion-pair in air- equilibrated solutions with 313 nm light, causes the reduction of the Co(EDTA) to Co(EDTA)2− and the oxidation of the I ion to I3. The results obtained are interpreted on the basis of a mechanism in which Co(EDTA)2− and I· are the primary photoproducts. The I· radical is then scavenged by I to yield I2, which subsequently disproportionates to I3 and I and reoxidizes Co(EDTA)2− to Co(EDTA). At the beginning of the photoreaction, the I2 decay is equally distributed on the two reaction pathways. It was possible to determine a value of 0.2 ± 0.05 for the photoreaction quantum yield and an efficiency of the primary photochemical step almost unitary. A schematic representation of the energetics of the overall reaction is reported.  相似文献   

16.
Reaction betwenn molecular oxygen and polystyrene covalently bonded Co(II) protoporphyrin IX complex, which was prepaired by the incorporation of a cobaltous ion into the metal-free porphyrin polymer, was studied in the presence of N-ethylimidazole by measuring visible absorption and electron spin resonance spectra. It was found that the complex forms a monomeric oxygen adduct reversibly at low temperature dependent on oxygen pressure. In the presence of molecular oxygen, a new electron spin resonance signal due to the oxygen complex at giso=2.02 shows no superhyperfine splitting structure in fluid toluene solution even at ?80 °C, but it was observed in frozen toulene glass solution at ?120°C, The oxygen adducts of the complexes between C0(II) protoporphyrin IX dimethyl ester and N-ethylimidazole and copoly(styrene-N-vinylimidazole) showed eight resolved superhyperfine splitting at ?40 and ?60°C, respectively. The polymer covalently bonded Co(II) complex with N-ethylimidazole was oxidized at room temperature under oxygen atmosphere. It was suggested that a Co(II) porphyrin–oxygen adduct with an axial ligand may be oxidised monomolecularly at high temperature.  相似文献   

17.
When photosystem II (PSII) is exposed to excess light, singlet oxygen (1O2) formed by the interaction of molecular oxygen with triplet chlorophyll. Triplet chlorophyll is formed by the charge recombination of triplet radical pair 3[P680•+Pheo•−] in the acceptor-side photoinhibition of PSII. Here, we provide evidence on the formation of 1O2 in the donor side photoinhibition of PSII. Light-induced 1O2 production in Tris-treated PSII membranes was studied by electron paramagnetic resonance (EPR) spin-trapping spectroscopy, as monitored by TEMPONE EPR signal. Light-induced formation of carbon-centered radicals (R) was observed by POBN-R adduct EPR signal. Increased oxidation of organic molecules at high pH enhanced the formation of TEMPONE and POBN-R adduct EPR signals in Tris-treated PSII membranes. Interestingly, the scavenging of R by propyl gallate significantly suppressed 1O2. Based on our results, it is concluded that 1O2 formation correlates with R formation on the donor side of PSII due to oxidation of organic molecules (lipids and proteins) by long-lived P680•+/TyrZ. It is proposed here that the Russell mechanism for the recombination of two peroxyl radicals formed by the interaction of R with molecular oxygen is a plausible mechanism for 1O2 formation in the donor side photoinhibition of PSII.  相似文献   

18.
A determination method for Co(II), Fe(II) and Cr(III) ions by luminol‐H2O2 system using chelating reagents is presented. A metal ion‐chelating ligand complex with a Co(II) ion and a chelating reagent like ethylenediaminetetraacetic acid (EDTA) produced highly enhanced chemiluminescence (CL) intensity as well as longer lifetime in the luminol‐H2O2 system compared to metals that exist as free ions. Whereas free Cu(II) and Pb(II) ions had a strong catalytic effect on the luminol‐H2O2 system, significantly, the complexes of Cu(II) and Pb(II) with chelating reagents lost their catalytic activity due to the chelating reagents acting as masking agents. Based on the observed phenomenon, it was possible to determine Co(II), Fe(II) and Cr(III) ions with enhanced sensitivity and selectivity using the chelating reagents of the luminol‐H2O2 system. The effects of ligand, H2O2 concentration, pH, buffer solution and concentrations of chelating reagents on CL intensity of the luminol‐H2O2 system were investigated and optimized for the determination of Co(II), Fe(II) and Cr(III) ions. Under optimized conditions, the calibration curve of metal ions was linear over the range of 2.0 × 10‐8 to 2.0 × 10‐5 M for Co(II), 1.0 × 10‐7 to 2.0 × 10‐5 M for Fe (II) and 2.0 × 10‐7 to 1.0 × 10‐4 M for Cr(III). Limits of detection (3σ/s) were 1.2 × 10‐8, 4.0 × 10‐8 and 1.2 × 10‐7 M for Co(II), Fe(II) and Cr(III), respectively. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

19.
Although quinones represent a class of organic compounds that may exert toxic effects both in vitro and in vivo, the molecular mechanisms involved in quinone species toxicity are still largely unknown, especially in the presence of transition metals, which may both induce the transformation of the various quinone species and result in generation of harmful reactive oxygen species. In this study, the oxidation of 1,4-naphthohydroquinone (NH2Q) in the absence and presence of nanomolar concentrations of Cu(II) in 10 mM NaCl solution over a pH range of 6.5–7.5 has been investigated, with detailed kinetic models developed to describe the predominant mechanisms operative in these systems. In the absence of copper, the apparent oxidation rate of NH2Q increased with increasing pH and initial NH2Q concentration, with concomitant oxygen consumption and peroxide generation. The doubly dissociated species, NQ2−, has been shown to be the reactive species with regard to the one-electron oxidation by O2 and comproportionation with the quinone species, both generating the semiquinone radical (NSQ). The oxidation of NSQ by O2 is shown to be the most important pathway for superoxide (O2) generation with a high intrinsic rate constant of 1.0×108 M−1 s−1. Both NSQ and O2 served as chain-propagating species in the autoxidation of NH2Q. Cu(II) is capable of catalyzing the oxidation of NH2Q in the presence of O2 with the oxidation also accelerated by increasing the pH. Both the uncharged (NH2Q0) and the mono-anionic (NHQ) species were found to be the kinetically active forms, reducing Cu(II) with an intrinsic rate constant of 4.0×104 and 1.2×107 M−1 s−1, respectively. The presence of O2 facilitated the catalytic role of Cu(II) by rapidly regenerating Cu(II) via continuous oxidation of Cu(I) and also by efficient removal of NSQ resulting in the generation of O2. The half-cell reduction potentials of various redox couples at neutral pH indicated good agreement between thermodynamic and kinetic considerations for various key reactions involved, further validating the proposed mechanisms involved in both the autoxidation and the copper-catalyzed oxidation of NH2Q in circumneutral pH solutions.  相似文献   

20.
One-pot reaction of cobalt(II) nitrate hexahydrate Co(NO3)2 · 6H2O with H2salpn (N,N′-bis(salicylidene)-1,3-diaminopropane) in presence of a large excess of sodium azide (NaN3) gives the new Co(III) compound {Na[CoIII(μ-salpn)(μ1,1-N3)2]}n (1), which was characterized by single crystal X-ray diffraction analysis. The crystal structure shows polymeric 1D complex generated by the hexadentate Schiff base salpn2− and two crystallographically different azide ligands. The two nitrogen atoms of the salpn ligand are bonded to the cobalt(III) ion while each phenoxo oxygen atom is bonded to the same Co(III) ion and to two equivalent sodium ions. Each azide ligand acts with the end-on bridging coordination mode between Co(III) and Na(I) ions. The Co(III) ion adopts a distorted octahedral geometry arising from two oxygen and two nitrogen atoms of the salpn ligand and from two nitrogen atoms of the two crystallographically different azide ligands in trans positions. Such [Co(salpn)(N3)2] entities are connected each other by sodium ions through four oxygen atoms of two equivalent Schiff base ligands and two nitrogen atom of the two different azide ligands to generate the 1D structure of 1.  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