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.
Pulse-generation of O2- by a flash was used to determine the reactivity of O2-, O2- was produced within 10 ms by a flash of light through the excitation of FMN in the presence of N,N,N',N'-tetramethylethylenediamine and oxygen. Kinetic analysis of cytochrome c reduction by O2- generated by flash yielded the reaction rate constant between cytochrome c and O2- and the spontaneous disproportionation rate constant of O2-. We applied it for superoxide dismutase assay using a linear relation between superoxide dismutase concentration and the apparent rate constant of cytochrome c reduction by O2-. The catalytic rate constant and activation energy at pH 7.3 of bovine liver Cu,Zn-superoxide dismutase were found to be 1.75 x 10(9) M-1 . s-1 at 25 degrees C and 26.9 kJ . M-1, respectively. The kinetics of O2- decay can be also monitored at 240 nm in this flash-photometric system and gave the spontaneous disproportionation rate constant of O2- and the catalytic rate constant of superoxide dismutase.  相似文献   

2.
In order to determine the reactivity on hemocyanin from Androctonus australis, the reaction of superoxide anion has been investigated using pulse radiolysis. The kinetics of O2- decays have been studied in aqueous buffered media at various basic pH (8, 8.5 and 9), first in the absence and then in the presence of hemocyanin (in oxygenated solutions containing formate anion 0.16 mol.l-1). We have shown that, in the presence of hemocyanin, O2- decay is a first-order process whose apparent rate constant is proportional to protein concentration (10(-7) to 10(-6) mol.l-1) and pH independent between 8 to 9. A second-order rate constant of 3.5 +/- 0.1.10(7) mol-1.l.s-1, has been deduced for the catalytic rate constant of hemocyanin with O2-. Meanwhile, this activity is smaller than that described for free copper, eukaryotic Cu-Zn-SOD or some copper chelates. We have verified that apohemocyanin--the copper deprived protein--does not exhibit such an activity vs. SOD (superoxide dismutase).  相似文献   

3.
The reaction between native myeloperoxidase and hydrogen peroxide, yielding Compound II, was investigated using the stopped-flow technique. The pH dependence of the apparent second-order rate constant showed the existence of a protonatable group on the enzyme with a pKa of 4.9. This group is ascribed to the distal histidine imidazole, which must be deprotonated to enable the reaction of Compound I with hydrogen peroxidase to take place. The rate constant for the formation of Compound II by hydrogen peroxide was 3.5.10(4) M-1.s-1. During the reaction of myeloperoxidase with H2O2, rapid reduction of added cytochrome c was observed. This reduction was inhibitable by superoxide dismutase, and this demonstrates that superoxide anion radicals are generated. When potassium ferrocyanide was used as an electron donor to generate Compound II from Compound I, the pH dependence of the apparent second-order rate constant indicated involvement of a group with a pKa of 4.5. However, with ferrocyanide as an electron donor, protonation of the group was necessary to enable the reaction to take place. The rate constant for the generation of Compound II by ferrocyanide was 1.6.10(7) M-1.s-1. We also investigated the reaction of Compound II with hydrogen peroxide, yielding Compound III. Formation of Compound III (k = 50 M-1.s-1) proceeded via two different pathways, one of which was inhibitable by tetranitromethane. We further investigated the stability of Compound II and Compound III as a function of pH, ionic strength and enzyme concentration. The half-life values of both Compound II and Compound III were independent of the enzyme concentration and ionic strength. The half-life value of Compound III was pH-dependent, showing a decreasing stability with increasing pH, whereas the stability of Compound II was independent of pH over the range 3-11.  相似文献   

4.
The reduction kinetics of NADPH:cytochrome P-450 reductase have been investigated by the laser flash photolysis technique, using the semiquinone of 5-deazariboflavin (5-dRfH.) as the reductant. Transients observed at 470 nm at neutral pH indicated that the oxidized reductase was reduced via second-order kinetics with a rate constant of 6.8 X 10(7) M-1 s-1. The second-order rate constant corresponding to the formation of the protein-bound semiquinone (measured at 585 nm) was essentially the same as that obtained at 470 nm (7.1 X 10(7) M-1 s-1). Subsequent to this rapid formation of protein-bound semiquinone, a partial exponential decay was observed at 585 nm. The rate of this decay remained invariant with protein concentration between pH 5.0 and 7.0, and a first-order rate constant of 70 s-1 was obtained for this process. This is assigned to intramolecular electron transfer from FADH. to FMN. Prior reduction of the enzyme to the one-electron level led to a decrease in both the second-order rate constant for reduction (2 X 10(7) M-1 s-1) and the first-order intraflavin electron transfer rate constant (15 s-1). The protein-bound FAD moiety of FMN-depleted reductase was reduced by 5-dRfH. with a second-order rate constant that was identical with that observed with the native enzyme (6.9 X 10(7) M-1 s-1). However, with this species no significant decay of the FAD semiquinone was observed at 585 nm following its rapid formation, consistent with the above assignment of this kinetic process.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Aqueous solutions of osmium tetroxide (OsO4) have been injected into arthritic knees for the past 45 years to chemically destroy diseased tissue, in a procedure termed "chemical synovectomy." Arthritis is an inflammatory disease. The primary inflammatory chemical species are the superoxide anion radical (O2.-) and nitric oxide (.NO), which combine to form the peroxynitrite anion (ONOO-). Here we show that OsO4 does not react with ONOO- but very efficiently catalyzes the dismutation of O2.- to O2 and H2O2. Using the pulse-radiolysis technique, the catalytic rate constant has been determined to be (1.43+/-0.04) x 10(9) M-1 s-1, independent of the pH in the 5.1-8.7 range. This value is about half that for the natural Cu,Zn-superoxide dismutase (Cu,Zn-SOD). Per unit mass, OsO4 is about 60 times more active than Cu,Zn-SOD. The catalytically active couple is OsVIII/OsVII, OsVIII oxidizing O2.- to O2 with a bimolecular rate constant of k=(2.6+/-0.1)x10(9) M-1 s-1 and OsVII reducing it to H2O2 with a bimolecular rate constant of (1.0+/-0.1)x10(9) M-1 s-1. Although lower valent osmium species are intrinsically poor catalysts, they are activated through oxidation by O2.- to the catalytic OsVIII/OsVII redox couple. The OsVIII/OsVII catalyst is stable to biochemicals other than proteins and peptides comprising histidine, cysteine, and dithiols.  相似文献   

6.
Peroxynitrite-mediated tyrosine nitration catalyzed by superoxide dismutase.   总被引:69,自引:0,他引:69  
Peroxynitrite (ONOO-), the reaction product of superoxide (O2-) and nitric oxide (NO), may be a major cytotoxic agent produced during inflammation, sepsis, and ischemia/reperfusion. Bovine Cu,Zn superoxide dismutase reacted with peroxynitrite to form a stable yellow protein-bound adduct identified as nitrotyrosine. The uv-visible spectrum of the peroxynitrite-modified superoxide dismutase was highly pH dependent, exhibiting a peak at 438 nm at alkaline pH that shifts to 356 nm at acidic pH. An equivalent uv-visible spectrum was obtained by Cu,Zn superoxide dismutase treated with tetranitromethane. The Raman spectrum of authentic nitrotyrosine was contained in the spectrum of peroxynitrite-modified Cu,Zn superoxide dismutase. The reaction was specific for peroxynitrite because no significant amounts of nitrotyrosine were formed with nitric oxide (NO), nitrogen dioxide (NO2), nitrite (NO2-), or nitrate (NO3-). Removal of the copper from the Cu,Zn superoxide dismutase prevented formation of nitrotyrosine by peroxynitrite. The mechanism appears to involve peroxynitrite initially reacting with the active site copper to form an intermediate with the reactivity of nitronium ion (NO2+), which then nitrates tyrosine on a second molecule of superoxide dismutase. In the absence of exogenous phenolics, the rate of nitration of tyrosine followed second-order kinetics with respect to Cu,Zn superoxide dismutase concentration, proceeding at a rate of 1.0 +/- 0.1 M-1.s-1. Peroxynitrite-mediated nitration of tyrosine was also observed with the Mn and Fe superoxide dismutases as well as other copper-containing proteins.  相似文献   

7.
1-Methyl-4-phenyl-2,3-dihydropyridinium (MPDP+), a metabolic product of the nigrostriatal toxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), has been shown to generate superoxide radicals during its autoxidation process. The generation of superoxide radicals was detected as a 5,5-dimethyl-1-pyrroline-N-oxide (DMPO).O2- spin adduct by spin trapping in combination with EPR techniques. The rate of formation of spin adduct was dependent not only on the concentrations of MPDP+ and oxygen but also on the pH of the system. Superoxide dismutase inhibited the spin adduct formation in a dose-dependent manner. The ability of DMPO to trap superoxide radicals, generated during the autoxidation of MPDP+, and of superoxide dismutase to effectively compete with this reaction for the available O2-, has been used as a convenient competition reaction to quantitatively determine various kinetic parameters. Thus, using this technique the rate constant for scavenging of superoxide radical by superoxide dismutase was found to be 7.56 x 10(9) M-1 s-1. The maximum rate of superoxide generation at a fixed spin trap concentration using different amounts of MPDP+ was found to be 4.48 x 10(-10) M s-1. The rate constant (K1) for MPDP+ making superoxide radical was found to be 3.97 x 10(-6) s-1. The secondary order rate constant (KDMPO) for DMPO-trapping superoxide radicals was found to be 10.2 M-1 s-1. The lifetime of superoxide radical at pH 10.0 was calculated to be 1.25 s. These values are in close agreement to the published values obtained using different experimental techniques. These results indicate that superoxide radicals are produced during spontaneous oxidation of MPDP+ and that EPR spin trapping can be used to determine the rate constants and lifetime of free radicals generated in aqueous solutions. It appears likely that the nigrostriatal toxicity of MPTP/MPDP+ leading to Parkinson's disease may largely be due to the reactivity of these radicals.  相似文献   

8.
1. At neutral pH ferricytochrome c is reduced by the superoxide anion radical (O2-), without loss of enzymatic activity, by a second order process in which no intermediates are observed. The yield of ferrocytochrome c (82-104%), as related to the amount of O2- produced, is slightly dependent on the concentration of sodium formate in the matrix solution. 2. The reaction (k1 equals (1.1+/-0.1) - 10(6) M-1 - s-1 at pH 7.2, I equals 4 mM and 21 degrees C) can be inhibited by superoxide dismutase and trace amounts of copper ions. The inhibition by copper ions is removed by EDTA without interference in the O2- reduction reaction. 3. The second-order rate constant for the reaction of O2- with ferricytochrome c depends on the pH of the matrix solution, decreasing rapidly at pH greater than 8. The dependence of the rate constant on the pH can be explained by assuming that only the neutral form of ferricytochrome c reacts with O2- and that the alkaline form of the hemoprotein is unreactive. From studies at pH 8.9, the rate for the transition from the alkaline to the neutral form of ferricytochrome c can be estimated to be 0.3 s-1 (at 21 degrees C and I equals 4 mM). 4. The second-order rate constant for the reaction of O2- with ferricytochrome c is also dependent on the ionic strength of the medium. From a plot of log k1 versus I1/2-(I + alphaI1/2)-1 we determined the effective charge on the ferricytochrome c molecule as +6.3 and the rate constant at I equals 0 as (3.1+/-0.1) - 10(6) M-1 - s-1 (pH 7.1, 21 degrees C). 5. The possibility that singlet oxygen is formed as a product of the reaction of O2- with ferricytochrome c can be ruled out on thermodynamic grounds.  相似文献   

9.
The formation of Compound I from Aspergillus niger catalase and methyl hydroperoxide (CH3OOH) has been investigated kinetically by means of rapid-scanning stopped-flow techniques. The spectral changes during the reaction showed distinct isobestic points. The second-order rate constant and the activation energy for the formation of Compound I were 6.4 x 10(3) M-1s-1 and 10.4 kcal.mol-1, respectively. After formation of Compound I, the absorbance at the Soret peak returned slowly to the level of ferric enzyme with a first-order rate constant of 1.7 x 10(-3) s-1. Spectrophotometric titration of the enzyme with CH3OOH indicates that 4 mol of peroxide react with 1 mol of enzyme to form 1 mol of Compound I. The amount of Compound I formed was proportional to the specific activity of the catalase. The irreversible inhibition of catalase by 3-amino-1,2,4-triazole (AT) was observed in the presence of CH3OOH or H2O2. The second-order rate constant of the catalase-AT formation in CH3OOH was 3.0 M-1 min-1 at 37 degrees C and pH 6.8 and the pKa value was estimated to be 6.10 from the pH profile of the rate constant of the AT-inhibition. These results indicate that A. niger catalase forms Compound I with the same properties as other catalases and peroxidases, but the velocity of the Compound I formation is lower than that of the others.  相似文献   

10.
Trace amounts of adventitious transition metals in buffer solutions can serve as catalysts for many oxidative processes. To fully understand what role these metals may play it is necessary that buffer solutions be 'catalytic metal free'. We demonstrate here that ascorbate can be used in a quick and easy test to determine if near-neutral buffer solutions are indeed 'catalytic metal free'. In buffers which have been rendered free of catalytic metals we have found that ascorbate is quite stable, even at pH 7. The first-order rate constant for the loss of ascorbate in an air-saturated catalytic metal free solution is less than 6 X 10(-7) s-1 at pH 7.0. This upper limit appears to be set by the inability to completely eliminate catalytic metal contamination of solutions and glassware. We conclude that in the absence of catalytic metals, ascorbate is stable at pH 7.  相似文献   

11.
Superoxide dismutase and Fe3+EDTA catalyzed the nitration by peroxynitrite (ONOO-) of a wide range of phenolics including tyrosine in proteins. Nitration was not mediated by a free radical mechanism because hydroxyl radical scavengers did not reduce either superoxide dismutase or Fe3+EDTA-catalyzed nitration and nitrogen dioxide was not a significant product from either catalyst. Rather, metal ions appear to catalyze the heterolytic cleavage of peroxynitrite to form a nitronium-like species (NO2+). The calculated energy for separating peroxynitrous acid into hydroxide ion and nitronium ion is 13 kcal.mol-1 at pH 7.0. Fe3+EDTA catalyzed nitration with an activation energy of 12 kcal.mol-1 at a rate of 5700 M-1.s-1 at 37 degrees C and pH 7.5. The reaction rate of peroxynitrite with bovine Cu,Zn superoxide dismutase was 10(5) M-1.s-1 at low superoxide dismutase concentrations, but the rate of nitration became independent of superoxide dismutase concentration above 10 microM with only 9% of added peroxynitrite yielding nitrophenol. We propose that peroxynitrite anion is more stable in the cis conformation, whereas only a higher energy species in the trans conformation can fit in the active site of Cu,Zn superoxide dismutase. At high superoxide dismutase concentrations, phenolic nitration may be limited by the rate of isomerization from the cis to trans conformations of peroxynitrite as well as by competing pathways for peroxynitrite decomposition. In contrast, Fe3+EDTA appears to react directly with the cis anion, resulting in greater nitration yields.  相似文献   

12.
A relatively pure and stable compound III of bovine spleen myeloperoxidase was prepared from native enzyme using the aerobic oxidation of dihydroxyfumarate to generate O2-(.). Spectral scans show well defined peaks at 450 and 625 nm and an isosbestic point between compound III and native enzyme at 440 nm. Compound III decayed to native enzyme without any detectable intermediate. The rate of decay was faster at alkaline pH values and also in the presence of superoxide dismutase. Ascorbic acid reduces compound III to native enzyme with a second order rate constant of (4.0 +/- 0.1) x 10(2) M-1 s-1. The ascorbic acid reduction of compound III has potential physiological relevance since it could help maintain the catalytic cycle of myeloperoxidase to generate the bactericidal agent hypochlorous acid.  相似文献   

13.
The filamentous cyanobacterium Anabaena PCC 7120 (now renamed Nostoc PCC 7120) possesses two genes for superoxide dismutase (SOD). One is an iron-containing (FeSOD) whereas the other is a manganese-containing superoxide dismutase (MnSOD). Localization experiments and analysis of the sequence showed that the FeSOD is cytosolic, whereas the MnSOD is a membrane-bound homodimeric protein containing one transmembrane helix, a spacer region, and a soluble catalytic domain. It is localized in both cytoplasmic and thylakoid membranes at the same extent with the catalytic domains positioned either in the periplasm or the thylakoid lumen. A phylogenetic analysis revealed that generally the highly homologous MnSODs of filamentous cyanobacteria are unique in being membrane-bound. Two recombinant variants of Anabaena MnSOD lacking either the hydrophobic region (MnSOD(Delta 28)) or the hydrophobic and the linker region (MnSOD(Delta 60)) are shown to exhibit the characteristic manganese peak at 480 nm, an almost 100% occupancy of manganese per subunit, a specific activity using the ferricytochrome assay of (660 +/- 90) unit mg-1 protein and a dissociation constant for the inhibitor azide of (0.84 +/- 0.05) mm. Using stopped-flow spectroscopy it is shown that the decay of superoxide in the presence of various (MnSOD(Delta 28)) or (MnSOD(Delta 60)) concentrations is first-order in enzyme concentration allowing the calculation of catalytic rate constants which increase with decreasing pH: 8 x 10(6) m-1 s-1 (pH 10) and 6 x 10(7) m-1 s-1 (pH 7). The physiological relevance of these findings is discussed with respect to the bioenergetic peculiarities of cyanobacteria.  相似文献   

14.
The mechanism of lignin peroxidase (LiP) was examined using bovine pancreatic ribonuclease A (RNase) as a polymeric lignin model substrate. SDS/PAGE analysis demonstrates that an RNase dimer is the major product of the LiP-catalyzed oxidation of this protein. Fluorescence spectroscopy and amino acid analyses indicate that RNase dimer formation is due to the LiP-catalyzed oxidation of Tyr residues to Tyr radicals, followed by intermolecular radical coupling. The LiP-catalyzed polymerization of RNase in strictly dependent on the presence of veratryl alcohol (VA). In the presence of 100 microM H2O2, relatively low concentrations of RNase and VA, together but not individually, can protect LiP from H2O2 inactivation. The presence of RNase strongly inhibits VA oxidation to veratraldehyde by LiP; whereas the presence of VA does not inhibit RNase oxidation by LiP. Stopped-flow and rapid-scan spectroscopy demonstrate that the reduction of LiP compound I (LiPI) to the native enzyme by RNase occurs via two single-electron steps. At pH 3.0, the reduction of LiPI by RNase obeys second-order kinetics with a rate constant of 4.7 x 10(4) M-1.s-1, compared to the second-order VA oxidation rate constant of 3.7 x 10(5) M-1.s-1. The reduction of LiP compound II (LiPII) by RNase also follows second-order kinetics with a rate constant of 1.1 x 10(4) M-1.s-1, compared to the first-order rate constant for LiPII reduction by VA. When the reductions of LiPI and LiPIi are conducted in the presence of both VA and RNase, the rate constants are essentially identical to those obtained with VA alone. These results suggest that VA is oxidized by LiP to its cation radical which, while still in its binding site, oxidizes RNase.  相似文献   

15.
Stopped-flow techniques were used to investigate the kinetics of the formation of manganese peroxidase compound I (MnPI) and of the reactions of MnPI and manganese peroxidase compound II (MnPII) with p-cresol and MnII. All of the rate data were obtained from single turnover experiments under pseudo-first order conditions. In the presence of H2O2 the formation of MnPI is independent of pH over the range 3.12-8.29 with a second-order rate constant of (2.0 +/- 0.1) x 10(6) M-1 s-1. The activation energy for MnPI formation is 20 kJ mol-1. MnPI formation also occurs with organic peroxides such as peracetic acid, m-chloroperoxybenzoic acid, and p-nitroperoxybenzoic acid with second-order rate constants of 9.7 x 10(5), 9.5 x 10(4), and 5.9 x 10(4) M-1 s-1, respectively. The reactions of MnPI and MnPII with p-cresol strictly obeyed second-order kinetics. The second-order rate constant for the reaction of MnPII with p-cresol is extremely low, (9.5 +/- 0.5) M-1 s-1. Kinetic analysis of the reaction of MnII with MnPI and MnPII showed a binding interaction with the oxidized enzymes which led to saturation kinetics. The first-order dissociation rate constants for the reaction of MnII with MnPI and MnPII are (0.7 +/- 0.1) and (0.14 +/- 0.01) s-1, respectively, when the reaction is conducted in lactate buffer. Rate constants are considerably lower when the reactions are conducted in succinate buffer. Single turnover experiments confirmed that MnII serves as an obligatory substrate for MnPII and that both oxidized forms of the enzyme form productive complexes with MnII. Finally, these results suggest the alpha-hydroxy acids such as lactate facilitate the dissociation of MnIII from the enzyme.  相似文献   

16.
The reaction of hydrogen peroxide with the copper-zinc bovine-liver superoxide dismutase at low molar ratios (0.2-20.0) of H2O2/active site between pH 7.3-10.0 leads to the loss of native enzyme as a distinct form monitored by electrophoresis. The pH dependence of the loss of native enzyme between 7.3 and 9.0 indicates the involvement of a conjugate base on the enzyme of pKa of 8.7 +/- 0.1. The rate of loss of the native enzyme is first order with respect to the concentration of both enzyme and hydrogen peroxide between pH 7.3 and 9.0 with no evidence for binding of peroxide. A second-order rate constant of 3.0 +/- 1.0 M-1 s-1 is obtained from these data. At pH 10.0 the reaction is first order with respect to enzyme concentration but saturable in H2O2. All data are consistent with the interpretation that H2O2 reacts with the enzyme at the lower pH where the reaction is dependent upon the conjugate base of a functional group on the enzyme. At the higher pH, the data are consistent with the reaction of HO2- and H2O2 with the dismutase. The dissociation constant for HO2- calculated from the kinetic data at pH 10.0 is between 25-50 microM and the rate constant for the breakdown of the HO2- dismutase complex is 1.10 + 0.05 x 10(-2) s-1. The change in the electrophoretic pattern at all pH values is accompanied by the loss of the ability of the enzyme to bind copper. Weakly bound or free copper can be detected using bathocuproine disulfonate. Furthermore copper-defficient forms of the enzyme can be detected by staining gels of the peroxide-treated dismutase with diethyldithiocarbamate.  相似文献   

17.
To help settle controversy as to whether the chelating agent diethylenetriaminepentaacetate (DTPA) supports or prevents hydroxyl radical production by superoxide/hydrogen peroxide systems, we have reinvestigated the question by spectroscopic, kinetic, and thermodynamic analyses. Potassium superoxide in DMSO was found to reduce Fe(III)DTPA. The rate constant for autoxidation of Fe(II)DTPA was found (by electron paramagnetic resonance spectroscopy) to be 3.10 M-1 s-1, which leads to a predicted rate constant for reduction of Fe(III)DTPA by superoxide of 5.9 x 10(3) M-1 s-1 in aqueous solution. This reduction is a necessary requirement for catalytic production of hydroxyl radicals via the Fenton reaction and is confirmed by spin-trapping experiments using DMPO. In the presence of Fe(III)DTPA, the xanthine/xanthine oxidase system generates hydroxyl radicals. The reaction is inhibited by both superoxide dismutase and catalase (indicating that both superoxide and hydrogen peroxide are required for generation of HO.). The generation of hydroxyl radicals (rather than oxidation side-products of DMPO and DMPO adducts) is attested to by the trapping of alpha-hydroxethyl radicals in the presence of 9% ethanol. Generation of HO. upon reaction of H2O2 with Fe(II)DTPA (the Fenton reaction) can be inhibited by catalase, but not superoxide dismutase. The data strongly indicate that iron-DTPA can catalyze the Haber-Weiss reaction.  相似文献   

18.
The mechanism of the enzymic reaction of an iron-containing superoxide dismutase purified from the marine bacterium Photobacterium leiognathi was studied by using pulse radiolysis. Measurements of activity were done with two different preparations of enzyme containing either 1.6 or 1.15 g-atom of iron/mol. In both cases, identical values of the second-order rate constant for reaction between superoxide dismutase and the superoxide ion in the pH range 6.2-9.0 (k=5.5 X 10(8) M-1-S-1 at pH 8.0) were found. As with the bovine erythrocuprein, there was no evidence for substrate saturation. The effects of reducing agents (H2O2, sodium ascorbate or CO2 radicals) on the visible and the electron-paramagnetic-resonance spectra of the superoxide dismutase containing 1.6 g-atom of ferric iron/mol indicate that this enzyme contains two different types of iron. Turnover experiments demonstrate that only that fraction of the ferric iron that is reduced by H2O2 is involved in the catalysis, being alternately oxidized and reduced by O2; both the oxidation and the reduction steps have a rate constant equal to that measured under turnover conditions. These results are interpreted by assuming that the superoxide dismutase isolated from the organism contains 1 g-atom of catalytic iron/mol and a variable amount of non-catalytic iron. This interpretation is discused in relation to the stoicheiometry reported for iron-containing superoxide dismutases prepared from several other organisms.  相似文献   

19.
The photophysics of the complex forming reaction of Ca2+ and Fura-2 are investigated using steady-state and time-resolved fluorescence measurements. The fluorescence decay traces were analyzed with global compartmental analysis yielding the following values for the rate constants at room temperature in aqueous solution with BAPTA as Ca2+ buffer: k01 = 1.2 x 10(9)s-1, k21 = 1.0 x 10(11) M-1 s-1, k02 = 5.5 x 10(8) s-1, k12 = 2.2 x 10(7) s-1, and with EGTA as Ca2+ buffer: k01 = 1.4 x 10(9) s-1, k21 = 5.0 x 10(10) M-1 s-1, k02 = 5.5 x 10(8) s-1, k12 = 3.2 x 10(7) s-1. k01 and k02 denote the respective deactivation rate constants of the Ca2+ free and bound forms of Fura-2 in the excited state. k21 represents the second-order rate constant of binding of Ca2+ and Fura-2 in the excited state, whereas k12 is the first-order rate constant of dissociation of the excited Ca2+:Fura-2 complex. The ionic strength of the solution was shown not to influence the recovered values of the rate constants. From the estimated values of k12 and k21, the dissociation constant K*d in the excited state was calculated. It was found that in EGTA Ca2+ buffer pK*d (3.2) is smaller than pKd (6.9) and that there is negligible interference of the excited-state reaction with the determination of Kd and [Ca2+] from fluorimetric titration curves. Hence, Fura-2 can be safely used as an Ca2+ indicator. From the obtained fluorescence decay parameters and the steady-state excitation spectra, the species-associated excitation spectra of the Ca2+ free and bound forms of Fura-2 were calculated at intermediate Ca2+ concentrations.  相似文献   

20.
D Y Cai  M Tien 《Biochemistry》1990,29(8):2085-2091
The oxycomplexes (compound III, oxyperoxidase) of two lignin peroxidase isozymes, H1 (pI = 4.7) and H8 (pI = 3.5), were characterized in the present study. After generation of the ferroperoxidase by photochemical reduction with deazoflavin in the presence of EDTA, the oxycomplex is formed by mixing ferroperoxidase with O2. The oxycomplex of isozyme H8 is very stable, with an autoxidation rate at 25 degrees C too slow to measure at pH 3.5 or 7.0. In contrast, the oxycomplex of isozyme H1 has a half-life of 52 min at pH 4.5 and 29 min at pH 7.5 at 25 degrees C. The decay of isozyme H1 oxycomplex follows a single exponential. The half-lives of lignin peroxidase oxycomplexes are much longer than those observed with other peroxidases. The binding of O2 to ferroperoxidase to form the oxycomplex was studied by stopped-flow methods. At 20 degrees C, the second-order rate constants for O2 binding are 2.3 X 10(5) and 8.9 X 10(5) M-1 s-1 for isozyme H1 and 6.2 X 10(4) and 3.5 X 10(5) M-1 s-1 for isozyme H8 at pH 3.6 and pH 6.8, respectively. The dissociation rate constants for the oxycomplex of isozyme H1 (3.8 Z 10(-3) s-1) and isozyme H8 (1.0 X 10(-3) s-1) were measured at pH 3.6 by CO trapping. Thus, the equilibrium constants (K, calculated from kon/koff) for both isozymes H1 (7.0 X 10(7) M-1) and H8 (6.2 X 10(7) M-1) are higher than that of myoglobin (1.9 Z 10(6) M-1).(ABSTRACT TRUNCATED AT 250 WORDS)  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