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.
The effect of bicarbonate on the rates of the H2O2 oxidation of cysteine, gluthathione, and N-acetylcysteine to the corresponding disulfides was investigated. The relative oxidation rates at pH 8 for the different thiols are inversely related to the pKa values of the thiol groups, and the reactive nucleophiles are identified as the thiolate anions or their kinetic equivalents. The second-order rate constants at 25 °C for the reaction of the thiolate anions with hydrogen peroxide are 17 ± 2 M−1 s−1 for all three substrates. In the presence of bicarbonate (>25 mM), the observed rate of thiolate oxidation is increased by a factor of two or more, and the catalysis is proposed to be associated with the formation of peroxymonocarbonate from the equilibrium reaction of hydrogen peroxide with bicarbonate (via CO2). The calculated second-order rate constants for the direct reaction of the three thiolate anions with peroxymonocarbonate fall within the range of 900-2000 M−1 s−1. Further oxidation of disulfides by peroxymonocarbonate results in the formation of thiosulfonate and sulfonate products. These results strongly suggest that peroxymonocarbonate should be considered as a reactive oxygen species in aerobic metabolism with relevance in thiol oxidations.  相似文献   

2.
Human serum heme–albumin (HSA-heme-Fe) displays globin-like properties. Here, kinetics of O2-mediated oxidation of ferrous nitrosylated HSA-heme-Fe (HSA-heme-Fe(II)-NO) is reported. Values of the first-order rate constants for O2-mediated oxidation of HSA-heme-Fe(II)-NO (i.e., for ferric HSA-heme-Fe formation) and for NO dissociation from HSA-heme-Fe(II)-NO (i.e., for NO replacement by CO) are k = 9.8 × 10−5 and 8.3 × 10−4 s−1, and h = 1.3 × 10−4 and 8.5 × 10−4 s−1, in the absence and presence of rifampicin, respectively, at pH = 7.0 and T = 20.0 °C. The coincidence of values of k and h indicates that NO dissociation represents the rate limiting step of O2-mediated oxidation of HSA-heme-Fe(II)-NO. Mixing HSA-heme-Fe(II)-NO with O2 does not lead to the formation of the transient adduct(s), but leads to the final ferric HSA-heme-Fe derivative. These results reflect the fast O2-mediated oxidation of ferrous HSA-heme-Fe and highlight the role of drugs in modulating allosterically the heme-Fe-atom reactivity.  相似文献   

3.
Catalase-peroxidases (KatGs) are unique bifunctional heme peroxidases that exhibit peroxidase and substantial catalase activities. Nevertheless, the reaction pathway of hydrogen peroxide dismutation, including the electronic structure of the redox intermediate that actually oxidizes H2O2, is not clearly defined. Several mutant proteins with diminished overall catalase but wild-type-like peroxidase activity have been described in the last years. However, understanding of decrease in overall catalatic activity needs discrimination between reduction and oxidation reactions of hydrogen peroxide. Here, by using sequential-mixing stopped-flow spectroscopy, we have investigated the kinetics of the transition of KatG compound I (produced by peroxoacetic acid) to its ferric state by trapping the latter as cyanide complex. Apparent bimolecular rate constants (pH 6.5, 20 °C) for wild-type KatG and the variants Trp122Phe (lacks KatG-typical distal adduct), Asp152Ser (controls substrate access to the heme cavity) and Glu253Gln (channel entrance) are reported to be 1.2 × 104 M− 1 s− 1, 30 M− 1 s− 1, 3.4 × 103 M− 1 s− 1, and 8.6 × 103 M− 1 s− 1, respectively. These findings are discussed with respect to steady-state kinetic data and proposed reaction mechanism(s) for KatG. Assets and drawbacks of the presented method are discussed.  相似文献   

4.
The kinetics of the formation of the purple complex [FeIII(EDTA)O2]3−, between FeIII-EDTA and hydrogen peroxide was studied as a function of pH (8.22-11.44) and temperature (10-40 °C) in aqueous solutions using a stopped-flow method. The reaction was first-order with respect to both reactants. The observed second-order rate constants decrease with an increase in pH and appear to be related to deprotonation of FeIII-EDTA ([Fe(EDTA)H2O] ⇔ Fe(EDTA)OH]2− + H+). The rate law for the formation of the complex was found to be d[FeIIIEDTAO2]3−/dt=[(k4[H+]/([H+] + K1)][FeIII-EDTA][H2O2], where k4=8.15±0.05×104 M−1 s−1 and pK1=7.3. The steps involved in the formation of [Fe(EDTA)O2]3− are briefly discussed.  相似文献   

5.
Quenching of Trp phosphorescence in proteins by diffusion of solutes of various molecular sizes unveils the frequency-amplitude of structural fluctuations. To cover the sizes gap between O2 and acrylamide, we examined the potential of acrylonitrile to probe conformational flexibility of proteins. The distance dependence of the through-space acrylonitrile quenching rate was determined in a glass at 77 K, with the indole analog 2-(3-indoyl) ethyl phenyl ketone. Intensity and decay kinetics data were fitted to a rate, k(r) = k0 exp[−(rr0)/re], with an attenuation length re = 0.03 nm and a contact rate k0 = 3.6 × 1010 s−1. At ambient temperature, the bimolecular quenching rate constant (kq) was determined for a series of proteins, appositely selected to test the importance of factors such as the degree of Trp burial and structural rigidity. Relative to kq = 1.9 × 109 M−1s−1 for free Trp in water, in proteins kq ranged from 6.5 × 106 M−1s−1 for superficial sites to 1.3 × 102 M−1s−1 for deep cores. The short-range nature of the interaction and the direct correlation between kq and structural flexibility attest that in the microsecond-second timescale of phosphorescence acrylonitrile readily penetrates even compact protein cores and exhibits significant sensitivity to variations in dynamical structure of the globular fold.  相似文献   

6.
C.L. Greenstock  R.W. Miller 《BBA》1975,396(1):11-16
The rate of reaction between superoxide anion (O¯.2) and 1,2-dihydroxybenzene-3,5-disulfonic acid (tiron) was measured with pulse radiolysis-generated O¯.2. A kinetic spectrophotometric method utilizing competition betweenp-benzoquinoneand tiron for O¯.2 was employed. In this system, the known rate of reduction ofp-benzoquinonewas compared with the rate of oxidation of tiron to the semiquinone. From the concentration dependence of the rate of tiron oxidation, the absolute second order rate constant for the reaction was determined to be 5 · 108 M?·s?1. Ascorbat reduced O¯.2 to hydrogen peroxide with a rate constant of 108 M?1 · s?1 as determined by the same method. The tiron semiquinone may be used as an indicator free radical for the formation of superoxide anion in biological systems because of the rapid rate of oxidation of the catechol by O¯.2 compared to the rate of O¯.2 formation in most enzymatic systems.Tiron oxidation was used to follow the formation of superoxide anion in swollen chloroplasts. The chloroplasts photochemically reduced molecular oxygen which was further reduced to hydrogen peroxide by tiron. Tiron oxidation specifically required O¯.2 since O2 was consumed in the reaction and tiron did not reduce the P700 cation radical or other components of Photosystem I under anaerobic conditions.  相似文献   

7.
Human myeloperoxidase (MPO) uses hydrogen peroxide generated by the oxidative burst of neutrophils to produce an array of antimicrobial oxidants. During this process MPO is irreversibly inactivated. This study focused on the unknown role of hydrogen peroxide in this process. When treated with low concentrations of H2O2 in the absence of reducing substrates, there was a rapid loss of up to 35% of its peroxidase activity. Inactivation is proposed to occur via oxidation reactions of Compound I with the prosthetic group or amino acid residues. At higher concentrations hydrogen peroxide acts as a suicide substrate with a rate constant of inactivation of 3.9 × 10−3 s−1. Treatment of MPO with high H2O2 concentrations resulted in complete inactivation, Compound III formation, destruction of the heme groups, release of their iron, and detachment of the small polypeptide chain of MPO. Ten of the protein’s methionine residues were oxidized and the thermal stability of the protein decreased. Inactivation by high concentrations of H2O2 is proposed to occur via the generation of reactive oxidants when H2O2 reacts with Compound III. These mechanisms of inactivation may occur inside neutrophil phagosomes when reducing substrates for MPO become limiting and could be exploited when designing pharmacological inhibitors.  相似文献   

8.
The direct immobilization of glucose oxidase (GOD) on TiO2/SiO2 nanocomposite and its application as glucose biosensor were investigated. The room-temperature phosphorescence of TiO2/SiO2 nanocomposite can be quenched by hydrogen peroxide (H2O2). The detection of glucose may be accomplished by monitoring the formation of hydrogen peroxide which generated in the oxidation process of glucose with the catalysis of GOD. To our surprise, by using a 96-hole polyporous plate accessory of fluorescence spectrophotometer, the biosensor exhibits excellent linear response to glucose concentrations ranging from 1.0 × 10−9 to 1.0 × 10−2 M with a detection limit of 1.2 × 10−10 M. The TiO2/SiO2 nanocomposite can be used as both supporting material and signal transducer. The phosphorescence intensity and color of the biosensor change obviously and even could be observed with naked eyes by continuous addition of glucose. Based on the room-temperature phosphorescence of TiO2/SiO2 nanocomposite, a new method of solid substrate-room-temperature phosphorimetry (SS-RTP) for glucose determination is proposed. A glucose biosensor was fabricated with wide determination concentration range, low detection limit, high sensitivity, and fast response time. And the biosensor has been successfully applied to the determination of glucose in human blood serum. The coacervation of GOD enzyme and its interaction with TiO2/SiO2 nanocomposite enlarge the surface area and enhance the chemical stability of GOD. The nice biocompatibility, large surface area, good chemical stability and nontoxicity of the TiO2/SiO2 nanocomposite have made this material suitable for functioning as biosensor.  相似文献   

9.
Human peroxiredoxin 5 (PRDX5) catalyzes different peroxides reduction by enzymatic substitution mechanisms. Enzyme oxidation caused an increase in Trp84 fluorescence, allowing performing pre-steady state kinetic measurements. The technique was validated by comparing with data available from the literature or obtained herein by alternative approaches. PRDX5 reacted with organic hydroperoxides with rate constants in the 106-107 M−1 s−1 range, similar to peroxynitrite-mediated PRDX5 oxidation, whereas its reaction with hydrogen peroxide was slower (105 M−1 s−1). The method allowed determining the kinetics of intramolecular disulfide formation as well as thioredoxin 2-mediated reduction. The reactivities of PRDXs with peroxides were surprisingly high considering thiol pKa, indicating that other protein determinants are involved in PRDXs specialization. The order of reactivities between PRDX5 towards oxidizing substrates differ from other PRDXs studied, pointing to a selective action of PRDXs with respect to peroxide detoxification, helping to rationalize the multiple enzyme isoforms present even in the same cellular compartment.  相似文献   

10.
The reactions of a dioxotetraamine Cu(II) complex [Cu(H−2L)] (L is 6-(9-fluorenyl)-1,4,8,11-tetraazaandencane-5,7-dione)with O2 − were investigated by electrochemistry, UV-Vis spectrophotometry and pulse radiolysis, respectively. In DMSO solution, [CuII(H−2L)] was oxidized into [CuIII(H−2L)]+ by O2 −, a consecutive reaction was observed with [CuIII(H−2L)(O2 2−)] − as intermediates (k1=1.71×103 M−1 s−1, k2=1.2×10−2 s−1). The mechanism of O2 − dismutation catalyzed by the complex involved alternate oxidation and reduction of Cu(II) by O2 − and the kcat is 6.07 × 107 M−1 s−1 (pH 7.4).  相似文献   

11.
The oxidation of an anticancer drug 5-fluorouracil (5-FU) by diperiodatoargentate(III) (DPA) was carried out both in the absence and presence of osmium(VIII) catalyst in alkaline medium at 27 °C and a constant ionic strength of 0.20 mol dm−3 spectrophotometrically attached with HI-TECH SFA-12 stopped flow accessory. The oxidation products in both the cases were identified as fluoroketene and Ag(I). The stoichiometry is same in both cases, i.e., [5-FU]:[DPA] = 1:1. The reaction was of first order in both catalysed and uncatalysed cases, with respect to [DPA] and was less than unit order in [5-FU] and negative fraction in [alkali]. The order in Os(VIII) was unity. In both cases [Ag(H3IO6)2] itself is the active species of DPA. The uncatalysed reaction in alkaline medium has been shown to proceed via a DPA-5-fluorouracil complex, which decomposes in a rate determining step to give the products. In catalysed reaction, it has been shown to proceed via a Os(VIII)-5-fluorouracil complex, which further reacts with one molecule of DPA in a rate determining step to give the products. The reaction constants involved in the different steps of the mechanisms were calculated for both the reactions. The catalytic constant (kCat.const.) was also calculated for catalysed reaction at different temperatures. The activation parameters with respect to slow step of the mechanisms were computed and discussed for both the cases. The thermodynamic quantities were also determined for both reactions.  相似文献   

12.
Exposure of winter rye leaves grown at 20°C and an irradiance of either 50 or 250 μmol m−2 s−1 to high light stress (1600 μmol m−2 s−1, 4 h) at 5°C resulted in photoinhibition of PSI measured in vivo as a 34% and 31% decrease in ΔA820/A820 (P700+). The same effect was registered in plants grown at 5°C and 50 μmol m−2 s−1. This was accompanied by a parallel degradation of the PsaA/PsaB heterodimer, increase of the intersystem e pool size as well as inhibition of PSII photochemistry measured as Fv/Fm. Surprisingly, plants acclimated to high light (800 μmol m−2 s−1) or to 5°C and moderate light (250 μmol m−2 s−1) were fully resistant to photoinhibition of PSI and did not exhibit any measurable changes at the level of PSI heterodimer abundance and intersystem e pool size, although PSII photochemistry was reduced to 66% and 64% respectively. Thus, we show for the first time that PSI, unlike PSII, becomes completely resistant to photoinhibition when plants are acclimated to either 20°C/800 μmol m−2 s−1 or 5°C/250 μmol m−2 s−1 as a response to growth at elevated excitation pressure. The role of temperature/light dependent acclimation in the induction of selective tolerance to PSI photoinactivation is discussed.  相似文献   

13.
The reaction of with H2O2 in 1.0 M HClO4/LiClO4 was found to be first-order in both reactants and the [H+] dependence of the second-order rate constant is given by k2obs = b/[H+], b at 25 °C is 26.4 ± 0.5 s−1. The rate law shows a simple inverse dependence on [H+] that is consistent with a rapidly maintained equilibrium between and its hydrolyzed form Co(H2O)5(OH)2+, followed by the rate controlling step, i.e. oxidation of H2O2 by Co(H2O)5(OH)2+.  相似文献   

14.
15.
Kinetics of the reaction of octacarbonyl dicobalt with ethyl diazoacetate leading to [μ2-{ethoxycarbonyl(methylene)}-μ2-(carbonyl)-bis(tricarbonyl-cobalt)] (Co-Co) (1), dinitrogen, and carbon monoxide were investigated at 10 °C in heptane solution. The initial rate of the reaction was measured by following both the gas evolution and the decrease of the octacarbonyl dicobalt concentration. The rate is first order with respect to octacarbonyl dicobalt and a complex order with respect to ethyl diazoacetate and carbon monoxide depending on the ratio of their concentrations. This is in accord with the formation of a heptacarbonyl dicobalt reactive intermediate (k1 (10 °C) = (1.22 ± 0.06) × 10−3 s−1) for which carbon monoxide and ethyl diazoacetate compete (k−1/k2 (10 °C) = 1.34 ± 0.07).  相似文献   

16.
17.
The anaerobic oxidation of cysteine, Cys, by Mn(III) in acetic acid solutions has been followed by use of a stopped-flow spectrophotometric method at a temperature of 20 °C. The formation and disappearance of the [Mn(OAc)2Cys] complex was monitored at 350 nm. The rate depends strongly on the acetic acid concentration (and hence also on pH) and led to the conclusion that more than one cysteine-containing species was involved. These mono-cysteinyl complexes are formed by the loss of two protons from the cysteine - one from the - SH and the other from either the -NH3+ or, more likely, the -COOH which is partially protonated at the low pH values involved (0.5-2.5). The rate-determining reprotonation of the bound -COO (or -NH2) is then accompanied by internal electron transfer yielding Mn(II) and the cysteinyl radical, Cys•, which then dimerises to form (inactive) cystine. At high acetic acid concentrations (60-90% AcOH) the tris-acetato species, [Mn(OAc)3], predominates together with some of the bis-complex, [Mn(OAc)2]+, and the active species is [Mn(OAc)2Cys] which decomposes with a rate constant of k2=16.8±0.9 M−1 s−1. At low acetic acid concentrations (20-30% AcOH) the mono-acetato species predominates and the reactive species is [Mn(OH)Cys] for which the rate of decomposition=k2=(1.32±0.11)×104 M−1 s−1. The relative values of the rate constants obtained are discussed, as is the bonding of cysteine to manganese(III).  相似文献   

18.
The luminescent complex [Pt(terpy)OH]BF4 undergoes photoinduced electron transfer reactions with phenyl amine electron donors and nitrophenyl electron acceptors. Stern-Volmer analysis of the quenching of metal-to-ligand charge transfer phosphorescence (3MLCT) was used to calculate bimolecular rate constants for electron transfer. Rate constants vary from 108 to >1010 M−1 s−1, depending on the thermodynamic driving force of the electron transfer reaction, with rate constants indicating that [Pt(terpy)OH]BF4* is a powerful photo-oxidant. Aromatic triplet energy acceptors can also quench the 3MLCT emission.  相似文献   

19.
When glycolate was metabolized in peroxisomes isolated from leaves of spinach beet (Beta vulgaris L., var. vulgaris) formate was produced. Although the reaction mixture contained glutamate to facilitate conversion of glycolate to glycine, the rate at which H2O2 became “available” during the oxidation of [1-14C]glycolate was sufficient to account for the breakdown of the intermediate [1-14C]glyoxylate to formate (C1 unit) and 14CO2. Under aerobic conditions formate production closely paralleled 14CO2 release from [1-14C]glycolate which was optimal between pH 8.0 and pH 9.0 and was increased 3-fold when the temperature was raised from 25 to 35 C, or when the rate of H2O2 production was increased artificially by addition of an active preparation of fungal glucose oxidase.  相似文献   

20.
The oxidation of Mn2+-pyrophosphate to Mn3+ by superoxide (O2?) was quantitative as evidenced from the formation of Mn3+-pyrophosphate and hydrogen peroxide and from the inhibition by superoxide dismutase. Using the competitive relation between Mn2+-pyrophosphate and superoxide dismutase for the O2?, the rate constant of Mn2+ oxidation was estimated to be about 6 × 106m?1 s?1. The oxidation of Mn2+-pyrophosphate by illuminated chloroplasts was also indicated to be stoichiometrically induced by O2?. In the presence of saturating amounts of the Mn2+, a double enhancement of hydrogen peroxide production and triple uptake of oxygen were found, as expected from the oxidation of Mn2+-pyrophosphate by O2?. Anaerobiosis or superoxide dismutase annuled these increments. We propose that the O2? generated as the sole initial step of the Mehler reaction oxidized Mn2+-pyrophosphate, and we discuss the role of free manganese in chloroplasts.  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