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 chemical properties of the nitro radical of four types of nitroheterocyclic compounds, nitrofurans, 2-nitroimidazoles, 4(5)-nitroimidazoles, 5-nitroimidazoles, having radiosensitizing and cytotoxic properties, have been studied by pulse radiolysis. The acid-base equilibria involving the nitro radical, the imidazole ring and some residues on the heterocycle have been determined. The pH-dependence of the rate of the disproportionation reaction of the nitro radical have been extensively studied. While the nitro radical derived from nitrofurans, 4- and 5-nitroimidazoles had a second-order decay, those of the 2-nitroimidazoles were found to decay through simultaneous first-order and second-order processes. Intrinsic second-order rate constants of the decay of the radical species in its various acidic and basic forms, could be determined. The intrinsic rate constants that determine the overall decay rates in the physiologically important 6 to 7.5 pH-range could be related to the one-electron redox potential E7(1). The implication of such chemical properties to enzyme-catalyzed reduction processes and to the mechanisms of radiosensitization and cytotoxicity of nitroheterocyclic compounds are briefly discussed. Pharmacological properties such as in vitro radiosensitization efficiency or metabolic reduction rates could be related to two of the nitro radical intrinsic disproportionation rates.  相似文献   

2.
The reduction of flavin in hepatic NADH-cytochrome b5 reductase by the hydrated electron (eaq-) was investigated by pulse radiolysis. The eaq- reduced the flavin of NADH-cytochrome b5 reductase to form the red semiquinone between pH 5 and 9. The spectrum of the red semiquinone differs from that of enzyme reduced by dithionite in the presence of NAD+. After the first phase of the reduction, conversion of the red to blue semiquinone was observed at acidic pH. Resulting products are the blue (neutral) or red (anionic) semiquinone or a mixture of the two forms. The pK value for this flavin radical was approximately 6.3. Subsequently, the semiquinone form reacted by dismutation to form the oxidized and the fully reduced forms of the enzyme with a rate constant of 1 x 10(3) M-1 s-1 at pH 7.1. In the presence of NAD+, eaq- reacted with NAD+ to yield NAD(.). Subsequently, NAD. transferred an electron to NAD+-bound oxidized enzyme to form the blue and red semiquinone or mixture of the two forms of the enzyme, where pK value of this flavin radical was approximately 6.3. The blue semiquinone obtained at acidic pH was found to convert to the red semiquinone with a first order rate constant of 90 s-1, where the rates were not affected by pH or the concentration of NAD+. The final product is NAD+-bound red semiquinone of the enzyme.  相似文献   

3.
Reactivities of several proanthocyanidins (monomers of condensed tannins) and gallate esters (representing hydrolyzable tannins) with hydroxyl radicals, azide radicals, and superoxide anions were investigated using pulse radiolysis combined with kinetic spectroscopy. We determined the scavenging rate constants and the decay kinetics of the aroxyl radicals both at the wavelength of the semiquinone absorption (275 nm) and the absorption band of the gallate ester ketyl radical (400-420 nm). For most compounds second-order decay kinetics were observed, which reflect disproportionation of the semiquinones. In the case of the oligomeric hydrolysable tannins, pentagalloyl glucose and tannic acid, the decay kinetics were more complex involving sequential first-order and second-order reactions, which could only be resolved by kinetic modeling. A correlation of the reaction rates with hydroxyl radicals (k*OH) with the number of adjacent aromatic hydroxyl groups (i.e., representing catechol and/or pyrogallol structures) was obtained for both condensed and hydrolyzable tannins. Similar correlation for the reactions with azide radicals and superoxide anions are less obvious, but exist as well. We consider proanthocyanidins superior radical scavenging agents as compared with the monomeric flavonols and flavones and propose that these substances rather than the flavonoids proper represent the antioxidative principle in red wine and green tea.  相似文献   

4.
Pulse radiolysis and steady-state X-radiolysis have been used to investigate the radiation inactivation of aldolase from rabbit muscle. Both eaq-and OH readily react with aldolase, and contribute to inactivation. The radical anions (CNS)2-and (Br)2-react with aldolase at neutral pH. The progressive addition of alkali results in an increase in the second-order rate constants, with an apparent pK approximately 10 +/- 0-3, and with the formation of an unstable intermediate, lambdamax approximately 400 nm resembling a phenoxyl radical. Steady-state radiolysis in the presence of (CNS)2-and (Br)2- at alkaline pH results in increased aldolase inactivation, with a pK of enzyme inactivation similar to that observed for reaction of the radical anions. We propose that a reaction of the radical anoins with tyrosine residues accounts for the resultant inactivation.  相似文献   

5.
This study examined the photo-induced generation of reactive oxygen species (ROS) by the carcinogenic iron(III)-NTA complex. Iron(III)-NTA complex (1:1) has three conformations (type (a) in acidic conditions of pH 1-6, type (n) in neutral conditions of pH 3-9, and type (b) in basic conditions of pH 7-10) with two pK(a) values (pK(a1) approximately 4, pK(a2) approximately 8). The iron(III)-NTA complex was reduced to iron(II) under cool-white fluorescent light without the presence of any reducing agent, and the reduction rates of the three conformations of iron(III)-NTA were in the order type (a)>type (n)>type (b) as reported previously (Akai K. et al., Free Radic. Res. 38, 951-962, 2004). ROS generation was investigated by electron paramagnetic resonance (EPR) spectroscopy with a spin-trapping technique. Apparent EPR signals attributed to PBN/*(13)CH(3) and PBN/*OCH(3) spin adducts were observed after incubation of the iron(III)-NTA complex was mixed with alpha-phenyl-tert-butylnitrone (PBN) and (13)C-DMSO in an aerobic condition. The addition of catalase effectively attenuated the PBN adducts, but superoxide dismutase enhanced them. Taken together, these results indicate that the iron(III)-NTA complex is spontaneously reduced to the iron(II)-NTA complex by light under acidic to neutral pH, and in turn transfers an electron to molecular oxygen to form ROS.  相似文献   

6.
Reduced "leuco" dyes such as dichlorodihydrofluorescein (DCFH(2)) are widely used as profluorescent probes for oxidative stress, although they require a catalyst to be oxidized by hydrogen peroxide and react indiscriminately with oxidizing radicals and the fluorescent product (DCF) is a potential photosensitizer of superoxide generation. In this study, key properties of the radical intermediate in oxidation ("semiquinone," DCFH(.-)/DCF(.)(-)) were measured, to help understand the reactions that can occur in biological systems. The intermediate was generated by oxidizing DCFH(2) or reducing DCF by radiolytically generated radicals and monitoring the reactions using kinetic spectrophotometry. The semiquinone showed pH-sensitive absorption spectral changes, decay kinetics (both in the absence and in the presence of oxygen), and reduction potential, all corresponding to prototropic dissociations with pK(a)'s of approximately 7.1 and 9.0. DCFH(2) has pK(a)'s in a similar region (8-9) and hence pH variations are potentially important in the use of this probe. The rate constant for reaction of the semiquinone with oxygen at pH 7.4 is 5.3 x 10(8) M(-1) s(-1): this reaction, rather than disproportionation of DCFH(.-)/DCF(.)(-), generates DCF in biological systems, concomitantly forming superoxide and hence H(2)O(2) to cycle the catalyst. The midpoint reduction potential of the couple DCF,H(+)/DCFH() is approximately -0.75 V vs. NHE at pH 7.4; DCF is unlikely to be reduced rapidly by common flavoprotein reductases.  相似文献   

7.
Previous studies on the chlorination reaction catalyzed by horseradish peroxidase using chlorite as the source of chlorine detected the formation of a chlorinating intermediate that was termed Compound X (Shahangian, S., and Hager, L.P. (1982) J. Biol. Chem. 257, 11529-11533). These studies indicated that at pH 10.7, the optical absorption spectrum of Compound X was similar to the spectrum of horseradish peroxidase Compound II. Compound X was shown to be quite stable at alkaline pH values. This study was undertaken to examine the relationship between the oxidation state of the iron protoporphyrin IX heme prosthetic group in Compound X and the chemistry of the halogenating intermediate. The experimental results show that the optical absorption properties and the oxidation state of the heme prosthetic group in horseradish peroxidase are not directly related to the presence of the activated chlorine atom in the intermediate. The oxyferryl porphyrin heme group in alkaline Compound X can be reduced to a ferric heme species that still retains the activated chlorine atom. Furthermore, the reaction of chlorite with horseradish peroxidase at acidic pH leads to the secondary formation of a green intermediate that has the spectral properties of horseradish peroxidase Compound I (Theorell, H. (1941) Enzymologia 10, 250-252). The green intermediate also retains the activated chlorine atom. By analogy to peroxidase Compound I chemistry, the heme prosthetic group in the green chlorinating intermediate must be an oxyferryl porphyrin pi-cation radical species (Roberts, J. E., Hoffman, B. M., Rutter, R. J., and Hager, L. P. (1981) J. Am. Chem. Soc. 103, 7654-7656). To be consistent with traditional peroxidase nomenclature, the red alkaline form of Compound X has been renamed Compound XII, and the green acidic form has been named Compound XI. The transfer of chlorine from the chlorinating intermediate to an acceptor molecule follows an electrophilic (rather than a free radical) path. A mechanism for the reaction is proposed in which the activated chlorine atom is bonded to a heteroatom on an active-site amino acid side chain. Transient state kinetic studies show that the initial intermediate, Compound XII, is formed in a very fast reaction. The second-order rate constant for the formation of Compound XII is approximately 1.1 x 10(7) M-1 s-1. The rate of formation of Compound XII is strongly pH-dependent. At pH 9, the second-order rate constant for the formation of Compound XII drops to 1.5 M-1 s-1. At acidic pH values, Compound XII undergoes a spontaneous first-order decay to yield Compound XI.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
Geijer P  Deák Z  Styring S 《Biochemistry》2000,39(23):6763-6772
We have studied the pH effect on the S(0) and S(2) multiline electron paramagnetic resonance (EPR) signals from the water-oxidizing complex of photosystem II. Around pH 6, the maximum signal intensities were detected. On both the acidic and alkaline sides of pH 6, the intensities of the EPR signals decreased. Two pKs were determined for the S(0) multiline signal; pK(1) = 4.2 +/- 0.2 and pK(2) = 8.0 +/- 0.1, and for the S(2) multiline signal the pKs were pK(1) = 4.5 +/- 0.1 and pK(2) = 7.6 +/- 0.1. The intensity of the S(0)-state EPR signal was partly restored when the pH was changed from acidic or alkaline pH back to pH approximately 6. In the S(2) state we observed partial recovery of the multiline signal when going from alkaline pH back to pH approximately 6, whereas no significant recovery of the S(2) multiline signal was observed when the pH was changed from acidic pH back to pH approximately 6. Several possible explanations for the intensity changes as a function of pH are discussed. Some are ruled out, such as disintegration of the Mn cluster or decay of the S states and formal Cl(-) and Ca(2+) depletion. The altered EPR signal intensities probably reflect the protonation/deprotonation of ligands to the Mn cluster or the oxo bridges between the Mn ions. Also, the possibility of decreased multiline signal intensities at alkaline pH as an effect of changed redox potential of Y(Z) is put forward.  相似文献   

9.
Semin BK  Seibert M 《Biochemistry》2004,43(21):6772-6782
Flash-probe fluorescence spectroscopy was used to compare the pH dependence of charge recombination between Y(Z)(*) and Q(a)(-) in Mn-depleted, photosystem II membranes [PSII(-Mn)] and in membranes with the high-affinity (HA(Z)) Mn-binding site blocked by iron [PSII(-Mn,+Fe); Semin, B. K., Ghirardi, M. L., and Seibert, M. (2002) Biochemistry 41, 5854-5864]. The apparent half-time for fluorescence decay (t(1/2)) in PSII(-Mn) increased from 9 ms at pH 4.4 to 75 ms at pH 9.0 [with an apparent pK (pK(app)) of 7.1]. The actual fluorescence decay kinetics can be fit to one exponential component at pH <6.0 (t(1/2) = 9.5 ms), but it requires an additional component at pH >6.0 (t(1/2) = 385 ms). Similar measurements with PSII(-Mn,+Fe) membranes show that iron binding has little effect on the maximum and minimum t(1/2) values measured at alkaline and acidic pHs but that it does shift the pK(app) from 7.1 to 6.1 toward the more acidic pK(app) value typical of intact membranes. Light-induced Fe(II) blocking of the PSII(-Mn) membrane is accompanied by a decrease in buffer Fe(II) concentration. This decrease was not the result of Fe(II) binding, but rather of its oxidation at two sites, the HA(Z) site and the low-affinity site. M?ssbauer spectroscopy at 80 K on PSII(-Mn,+Fe) samples, prepared under conditions providing the maximal blocking effect but minimizing the amount of nonspecifically bound iron cations, supports this conclusion since this method detected only Fe(III) cations bound to the membranes. Correlation of the kinetics of Fe(II) oxidation with the blocking parameters showed that blocking occurs after four to five Fe(II) cations were oxidized at the HA(Z) site. In summary, the blocking of the HA(Z) Mn-binding site by iron in PSII(-Mn) membranes not only prevents the access of exogenous donors to Y(Z) but also partially restores the properties of the hydrogen bond net found in intact PS(II), which in turn controls the rate of electron transport to Y(Z).  相似文献   

10.
A model compound for the chromophore within the purple nonfluorescent GFP-like chromoprotein asFP595 was synthesized. The postulated structure of the chromophore, 2-acetyl-4-(p-hydroxybenzylidene)-1-methyl-5-imidazolone, was taken from the high-resolution crystal structure analysis of intact asFP595 [Quillin, M. L., Anstrom, D., Shu, X., O'Leary, S., Kallio, K., Lukyanov, K. A., and Remington, S. J. (2005) Kindling Fluorescent Protein from Anemonia sulcata: Dark-State Structure at 1.38 A Resolution, Biochemistry 44, 5774-5787]. Erlenmeyer lactonization and oxidation of the methylene group attached to the heteroaromatic moiety with selenium dioxide were used at the key stages of the synthesis. The spectral properties of the model chromophore in solution and their dependence on the pH and polarity of the solvent were investigated. In water, the chromophore was found to exist in two forms, neutral and anionic, with a pK(a) of 7.1. In a dimethylformamide solution, the spectral properties of the anionic form closely match those of the native protein, demonstrating that under these conditions, the compound is an excellent model for the chromophore within native asFP595.  相似文献   

11.
D-Glyceraldehyde 3-phosphate forms adducts with thiols. These adducts, which are presumed to be hemithioacetals, equilibrate rapidly with the unhydrated form of the aldehyde, which is the subtrate for D-glyceraldehyde 3-phosphate dehydrogenase. The adduct provides a substrate buffer system whereby a constant low free aldehyde concentration can be maintained during the oxidation of aldehyde by the enzyme and NAD+. With this system, the kinetics of the association of the aldehyde with the enzyme were examined. The rate profile for this reaction is a single exponential process, showing that all four active sites of the enzyme have equivalent and independent reactivity towards the aldehyde, with an apparent second-order rate constant of 5 X 10(7)M-1-S-1 at pH8.0 and 21 degrees C. The second-order rate constant becomes 8 X 10(7)M-1-S-1 when account is taken of the forward and reverse catalytic rate constants of the dehydrogenase. The pH-dependence of the observed rate constant is consistent with a requirement for the unprotonated form of a group of pK 6.1, which is the pK observed for second ionization of glyceraldehyde 3-phosphate. The rate of phosphorolysis of the acyl-enzyme intermediate during the steady-state oxidative phosphorylation of the aldehyde was studied, and is proportional to the total Pi concentration up to at least 1 mM-Pi at pH 7.5. The pH-dependence of the rate of NADH generation under these conditions can be explained by the rate law d[NADA]/dt = k[acy] holoenzyme][PO4(3-)-A1, where thioester bond, although kinetically indistinguishable rate equations for the reaction are possible. The rates of the phosphorolysis reaction and of the aldehyde-association reaction decrease with increasing ionic strength, suggesting that the active site of the enzyme has cationic groups which are involved in the reaction of the enzyme with anionic substrates.  相似文献   

12.
3-Nitrotyrosine has been reported as an important biomarker of oxidative stress that may play a role in a variety of diseases. In this work, transient UV-visible absorption spectra and kinetics observed during the reaction of the hydrated electron, e(aq)(-), with 3-nitrotyrosine and derivatives thereof were investigated. The absorption spectra show characteristics of aromatic nitro anion radicals. The absorptivity of radical anion product at 300 nm is estimated to be (1.0 ± 0.2) × 10(4) M(-1) cm(-1) at pH 7.3. The rate constants determined for the reaction of e(aq)(-) with 3-nitrotyrosine, N-acetyl-3-nitrotyrosine ethyl ester and glycylnitrotyrosylglycine at neutral pH (3.0 ± 0.3) × 10(10) M(-1) s(-1), (2.9 ± 0.2) × 10(10) M(-1) s(-1) and (1.9 ± 0.2) × 10(10) M(-1) s(-1), respectively, approach the diffusion-control limit and are almost two orders of magnitude higher than those for the reactions with tyrosine and tyrosine-containing peptides. The magnitude of the rate constants supports reaction of e(aq)(-) at the nitro group, and the product absorbance at 300 nm is consistent with formation of the nitro anion radical. The pH dependence of the second-order rate constant for e(aq)(-) decay (720 nm) in the presence of 3-nitrotyrosine shows a decrease with increasing pH, consistent with unfavorable electrostatic interactions. The pH dependence of the second-order rate constant for formation of radical anion (300 nm) product suggests that deprotonation of the amino group slows the rate, which indicates that deamination to form the 1-carboxy-2-(4-hydroxy-3-nitrophenyl)ethyl radical occurs. We conclude that the presence of the nitro group activates tyrosine and derivatives toward reaction with e(aq)(-) and can affect the redox chemistry of biomolecules exposed to oxidative stress.  相似文献   

13.
Flash photolysis techniques have been utilized to investigate the reactivity patterns of flavin radical species. Rate constants for disproportionation were found to be la the following order: lumiflavin>FMN>FAD and neutral radicals>anionic radicals. Neutral flavin radicals react with oxygen at a rate which is at least 104 times slower than the anionic species. No evidence for an intermediate complex or adduct is obtained in this reaction. The pK values for the ionization of the neutral flavin radicals are in the order FAD>FMN>riboflavin=limiflavin. The rates of reaction of ferricyanide with flavin radicals are essentially independent of pH, whereas benzoquinone reacts slightly more slowly (5 times) with the neutral flavin radical than with the anionic form. Cytochromec reacts at least ten times more slowly with flavin radicals than does ferricyanide.  相似文献   

14.
The bifunctional hypoxia-specific cytotoxin RB90745, has a nitroimidazole moiety attached to an imidazo[1,2,-a]quinoxaline mono-N-oxide with a spacer/linking group. The reduction chemistry of the drug was studied by pulse radiolysis using the one electron reductant CO2˙-. As N-oxides and nitro compounds react with CO2˙- at diffusion controlled rates, initial reaction produced a mixture of the nitro radical (λmax 410 nm) and the N-oxide radical (λmax 550 nm) in a few microseconds. Subsequently an intramolecular electron transfer (IET) was observed (k = 1.0 ± 0.25 × 103 s-1 at pH 5-9), from the N-oxide to the more electron-affinic nitro group. This was confirmed by the first order decay rate of the radical at 550 nm and formation at 410 nm, which was independent of both the concentration of the parent compound and the radicals. The rates of electron transfer and the decay kinetics of the nitro anion radicals were pH dependent and three different pKaS could be estimated for the one electron reduced species: 5.6 (nitroimidazole group) and 4.3, and 7.6 (N-oxide function). The radicals react with oxygen with rate constants of 3.1 × 107 and 2.8 × 106 dm3 mol-1 s-1 observed at 575 nm and 410 nm respectively. Steady state radiolysis studies indicated four electron stoichiometry for the reduction of the compound.  相似文献   

15.
At higher concentrations, deferoxamine (DFO) reacts with hydroxyl radicals to produce a stable nitroxide free radical. Formation and decay of this nitroxide radical was investigated and compared with a novel modified pentastarch conjugate of DFO (MPS-DFO). Photolytic generation of hydroxyl radicals from H2O2 in the presence of free DFO produced a nitroxide radical with coupling constants of aN = 8.0 G and aH = 6.5 G. Under the same experimental conditions, equimolar concentrations of MPS-DFO produced an ESR signal of reduced intensity while iron-saturated MPS-DFO produced no signal. Incubation of free DFO with pentastarch (i.e., without conjugation) greatly decreased the intensity of the nitroxide radical signal. Using a spin-trapping technique with 5,5-dimethyl-1-pyrroline N-oxide (DMPO), the pentastarch vehicle was shown to inhibit the DMPO-OH adduct formation. The decay of the DFO nitroxide radical decayed with a second-order rate constant while that of MPS-DFO decayed with a first-order rate constant. Thus, a novel derivative of DFO may provide some additional benefit in limiting DFO nitroxide radical formation and might explain the reported reduced in vivo toxicity of MPS-DFO relative to free DFO.  相似文献   

16.
Combined optical and conductimetric measurements in aqueous solution indicate that at high pH (greater than or equal to 10).OH radicals react with the phenoxide form of 2,4-dihydroxybenzoic acid to form transiently phenoxyl radicals and a small amount of hydroxyeyclohexadienyl (HCHD) radicals by 150 ns. The respective yields of 88 and 12% of the total.OH radical yield were deduced from conductance and optical changes as well as from studies using a low potential reductant. The HCHD radical possesses a pKa of 8.0 +/- 0.1 and the constructed spectrum of the deprotonated forms of HCHD has a lambda max at 420 nm with a minimum extinction coefficient of approximately 7250 M-1 cm-1. The red shift in lambda max and increase in extinction coefficient compared to the revised spectral properties of the protonated form of the HCHD radical (lambda max at 390 nm with extinction coefficient of approximately 4500 M-1 cm-1), together with the pKa of the HCHD radical, provide an explanation for the pH-dependent spectral changes of the so-called highly absorbing intermediate II species, observed in the functioning of the enzyme p-hydroxybenzoate hydroxylase. These results add further to the evidence in support of the proposal that intermediate II is composed of species which absorb similarly to the flavin 4(a)-hydroxide and a form of the substrate/product such as the HCHD radical (Anderson, R. F., Patel, K. B., and Stratford, M. R. L. (1987) J. Biol. Chem. 262, 17475-17479).  相似文献   

17.
The reactions of the cysteine, glutathione and penicillamine thiyl radicals with oxygen and their parent thiols in frozen aqueous solutions have been elucidated through electron spin resonance spectroscopy. The major sulfur radicals observed are: (1) thiyl radicals, RS.; (2) disulfide radical anions. RSSR-.; (3) perthiyl radicals, RSS. and upon introduction of oxygen; (4) sulfinyl radicals, RSO., where R represents the remainder of the cysteine, glutathione or penicillamine moiety. The radical product observed depends on the pH, concentration of thiol, and presence or absence of molecular oxygen. We find that the sulfinyl radical is a ubiquitous intermediate in the free radical chemistry of these important biological compounds, and also show that peroxyl radical attack on thiols may lead to sulfinyl radicals. We elaborate the observed reaction sequences that lead to sulfinyl radicals, and, using 17O isotopic substitution studies, demonstrate that the oxygen atom in sulfinyl radicals originates from dissolved molecular oxygen. In addition, the glutathione thiyl radical is found to abstract hydrogen from the alpha-carbon position on the cysteine residue of glutathione to form a carbon-centered radical.  相似文献   

18.
The oxidation-reduction potential values for the two electron transfers to glucose oxidase were obtained at pH 5.3, where the neutral radical is the stable form, and at pH 9.3, where the anion radical is the stable form. The midpoint potentials at 25 degrees were: pH 5.3 EFl1ox + e- H+ equilibrium EFlH. Em1 = -0.063 +/- 0.011 V EFlH. + e- + H+ equilibrium EFlredH2 Em2 = -0.065 +/- 0.007 V pH 9.3 EFlox + e- EFi- Em1 = -0.200 +/- 0.010 V EFi- + e- + H+ equilibrium EFlredH- Em2 = -0.240 +/- 0.005 V All potentials were measured versus the standard hydrogen electrode (SHE). The potentials indicated that glucose oxidase radicals are stabilized by kinetic factors and not by thermodynamic energy barriers. The pK for the glucose oxidase radical was 7.28 from dead time stopped flow measurements and the extinction coefficient of the neutral semiquinone was 4140 M-1 cm-1 at 570 nm. Both radical forms reacted with oxygen in a second order fashion. The rate at 25 degrees for the neutral semiquinone was 1.4 X 10(4) M-1 s-1; that for the anion radical was 3.5 X 10(4) M-1 s-1. The rate of oxidation of the neutral radical changed by a factor of 9 for a temperature difference of 22 degrees. For the anion radical, the oxidation rate changed by a factor of 6 for a 22 degrees change in temperature. We studied the oxygen reactivity of the 2-electron reduced form of the enzyme over a wide wavelength range and failed to detect either oxygenated flavin derivatives or semiquinoid forms as intermediates. The rate of reoxidation of fully reduced glucose oxidase at pH 9.3 was dependent on ionic strength.  相似文献   

19.
Antioxidant properties of melatonin: a pulse radiolysis study   总被引:5,自引:0,他引:5  
Various one-electron oxidants such as OH*, tert-BuO*, CCl3OO*, Br2*- and N3*, generated pulse radiolytically in aqueous solutions at pH 7, were scavenged by melatonin to form two main absorption bands with lambda(max) = 335 nm and 500 nm. The assignment of the spectra and determination of extinction coefficients of the transients have been reported. Rate constants for the formation of these species ranged from 0.6-12.5x10(9) dm3 mol(-1) s(-1). These transients decayed by second order, as observed in the case of Br2*- and N3* radical reactions. Both the NO2* and NO* radicals react with the substrate with k = 0.37x10(7) and 3x10(7) dm3 mol(-1) s(-1), respectively. At pH approximately 2.5, the protonated form of the transient is formed due to the reaction of Br2*- radical with melatonin, pKa ( MelH* <=> Mel* + H+) = 4.7+/-0.1. Reduction potential of the couple (Mel*/MelH), determined both by cyclic voltammetric and pulse-radiolytic techniques, gave a value E(1)7 = 0.95+/-0.02 V vs. NHE. Repair of guanosine radical and regeneration of melatonin radicals by ascorbate and urate ions at pH 7 have been reported. Reactions of the reducing radicals e(aq)- and H* atoms with melatonin have been shown to occur at near diffusion rates.  相似文献   

20.
Catalysis of the H2O2-dependent oxidation of 3,4-dimethoxybenzyl (veratryl) alcohol by the hemoprotein ligninase isolated from wood-decaying fungus, Phanerochaete chrysosporium Burds, is characterized. The reaction yields veratraldehyde and exhibits a stoichiometry of one H2O2 consumed per aldehyde formed. Ping-pong steady-state kinetics are observed for H2O2 (KM = 29 microM) and veratryl alcohol (KM = 72 microM) at pH 3.5. The magnitude of the turnover number varies from 2 to 3 s-1 at this pH, depending on the preparation of the enzyme. Each preparation of enzyme consists of a mixture of active and inactive enzyme. Extensive steady-state kinetic studies of several different preparations of enzyme, suggest a mechanism in which H2O2 reacts with enzyme to form an intermediate that subsequently reacts with the alcohol to return the enzyme to the resting state. The pH dependence of the overall reaction indicates that an ionization occurs having an apparent pK alpha approximately 3.1. The activity is, thus, nearly zero at pH 5 and increases to a maximum near pH approximately 2. However, the enzyme is unstable at this low pH. Transient-state kinetic studies reveal that, upon reaction of ligninase with H2O2, spectral changes occur in the Soret region, which, by analogy to previous studies of horseradish peroxidase, are consistent with formation of Compounds I and II. The active form of the enzyme appears to react rapidly with H2O2; we observed a positive correlation between the turnover number of the enzyme preparation and the extent of a rapid reaction between H2O2 and ligninase to form Compound I. Free radical cations derived from veratryl alcohol do not appear to be released from the enzyme during catalysis; however, other substrates are known to be converted to cation radicals (Kersten, P., Tien, M., Kalyanaraman, B., and Kirk, T.K. (1985) J. Biol. Chem. 260, 2609-2612). Our results are generally consistent with a classical peroxidase mechanism for the action of ligninase on lignin-like substrates.  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