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 78 毫秒
1.
The kinetics of microperoxidase-11 (MP-11) in the oxidation reaction of guaiacol (AH) by hydrogen peroxide was studied, taking into account the inactivation of enzyme during reaction by its suicide substrate, H2O2. Concentrations of substrates were so selected that: 1) the reaction was first-order in relation to benign substrate, AH and 2) high ratio of suicide substrate to the benign substrate, [H2O2]>>[AH]. Validation and reliability of the obtained kinetic equations were evaluated in various nonlinear and linear forms. Fitting of experimental data into the obtained integrated equation showed a close match between the kinetic model and the experimental results. Indeed, a similar mechanism to horseradish peroxidase was found for the suicide-peroxide inactivation of MP-11. Kinetic parameters of inactivation including the intact activity of MP-11, αi, and the apparent inactivation rate constant, ki, were obtained as 0.282 ± 0.006 min? 1 and 0.497 ± 0.013 min? 1 at [H2O2] = 1.0 mM, 27°C, phosphate buffer 5.0 mM, pH = 7.0. Results showed that inactivation of microperoxidase as a peroxidase model enzyme can occur even at low concentrations of hydrogen peroxide (0.4 mM).  相似文献   

2.
The kinetics of the reduction by aniline and a series of substituted anilines of a peroxidatically active intermediate, formed by oxidation of deuteroferriheme with hydrogen peroxide, have been studied by stopped-flow spectrophotometry. The reaction with aniline was first order with respect to [intermediate] and showed first-order saturation kinetics with respect to [aniline]. The second-order rate constant was 2.0 ± 0.2 × 105 M?1 sec?1 at 25°C (independent of pH in the range 6.60–9.68) compared with the value of 2.4 × 105 M?1 sec?1 for the reaction of aniline with horseradish peroxidase Compound I. The effect of aniline substituents upon reactivity towards the heme intermediate closely paralled those reported for reaction with the enzymic intermediate. Anilines bearing electron-donating substituents reacted more rapidly and those bearing electron-withdrawing substituents more slowly than the unsubstituted amine. The rate constants for the heme intermediate reactions (kdfh)found to be related to those for the enzymic reactions (khrp) by the equation:log kDFH= 0.65log kHRP+ 1.96 with a correlation coefficient of 0. 98.  相似文献   

3.
Hamster liver glutathione peroxidase was purified to homogeneity in three chromatographic steps and with 30% yield. The purified enzyme had a specific activity of approximately 500 μmol cumene hydroperoxide reduced/min/mg of protein at 37 °C, pH 7.6, and 0.25 mm GSH. The enzyme was shown to be a tetramer of indistinguishable subunits, the molecular weight of which was approximately 23,000 as estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. A single isoelectric point of 5.0 was attributed to the active enzyme. Amino acid analysis determined that selenocysteine, identified as its carboxymethyl derivative, was the only form of selenium. One residue of cysteine was found to be present in each glutathione peroxidase subunit. The presence of tryptophan was colorimetrically determined. Pseudo-first-order kinetics of inactivation of the enzyme by iodoacetate was observed at neutral pH with GSH as the only reducing agent. An optimal pH of 8.0 at 37 °C and an activation energy of 3 kcal/mol at pH 7.6 were found. A ter-uni-ping-pong mechanism was shown by the use of an integrated-rate equation. At pH 7.6, the apparent second-order rate constants for reaction of glutathione peroxidase with hydroperoxides were as follows: k1 (t-butyl hydroperoxide), 7.06 × 105 mm min?1; k1 (cumene hydroperoxide), 1.04 × 106 mm?1 min?1; k1 (p-menthane hydroperoxide), 1.2 × 106 mm?1 min?1; k1 (diisopropylbenzene hydroperoxide), 1.7 × 106 mm?1 min?1; k1 (linoleic acid hydroperoxide), 2.36 × 106 mm?1 min?1; k1 (ethyl hydroperoxide), 2.5 × 106 mm?1 min?1; and k1 (hydrogen peroxide), 2.98 × 106 mm?1 min?1. It is concluded that for bulky hydroperoxides, the more hydrophobic the substrate, the faster its reduction by glutathione peroxidase.  相似文献   

4.
The effect of sulfhydryl oxidase on the rate of disulfide bond formation and polypeptide chain folding in reductively denatured chymotrypsinogen A has been investigated using an immobilized zymogen preparation and a cylindrical quartz flow-through fluorescence cell. Enzymatic oxidation of the 10 sulfhydryl groups in reduced chymotrypsinogen followed first order kinetics at pH 7.0 with an apparent first order rate constant governing sulfhydryl group disappearance of 4.2 × 10?2 min?1. This provides a t12 of 16.3 min for the sulfhydryl oxidase-catalyzed oxidation, whereas 165 min are required for nonenzymatic aerobic oxidation of one-half the sulfhydryl groups. Refolding of the reductively denatured polypeptide chains, monitored by changes in protein fluorescence, did not follow first order kinetics characteristic of a simple two-state mechanism, nor did the return of trypsin activatability. It appears that at least one intermediate must exist in such refolding, in both the uncatalyzed and sulfhydryl oxidase-catalyzed processes. Estimation of the rate constants governing refolding, assuming a single intermediate between the denatured and native states, provided values of 3 × 10?2 min?1 and 7 × 10?3 min?1 for uncatalyzed autoxidation and 4 × 10?2 min?1 and 1.1 × 10?2 min?1 for the sulfhydryl oxidase-catalyzed transition. Thus, enzymic catalysis of disulfide bond formation can lead to apparent catalysis of protein refolding as monitored both by fluorescence and by acquisition of biological function.  相似文献   

5.
The initial reactions involved in anaerobic aniline degradation by the sulfate-reducing Desulfobacterium anilini were studied. Experiments for substrate induction indicated the presence of a common pathway for aniline and 4-aminobenzoate, different from that for degradation of 2-aminobenzoate, 2-hydroxybenzoate, 4-hydroxybenzoate, or phenol. Degradation of aniline by dense cell suspensions depended on CO2 whereas 4-aminobenzoate degradation did not. If acetyl-CoA oxidation was inhibited by cyanide, benzoate accumulated during degradation of aniline or 4-aminobenzoate, indicating an initial carboxylation of aniline to 4-aminobenzoate, and further degradation via benzoate of both substrates. Extracts of alinine or 4-aminobenzoategrown cells activated 4-aminobenzoate to 4-aminobenzoyl-CoA in the presence of CoA, ATP and Mg2+. 4-Aminobenzoyl-CoA-synthetase showed a K m for 4-aminobenzoate lower than 10 M and an activity of 15.8 nmol · min-1 · mg-1. 4-Aminobenzoyl-CoA was reductively deaminated to benzoyl-CoA by cell extracts in the presence of low-potential electron donors such as titanium citrate or cobalt sepulchrate (2.1 nmol · min-1 · mg-1). Lower activities for the reductive deamination were measured with NADH or NADPH. Reductive deamination was also indicated by benzoate accumulation during 4-aminobenzoate degradation in cell suspensions under sulfate limitation. The results provide evidence that aniline is degraded via carboxylation to 4-aminobenzoate, which is activated to 4-aminobenzoyl-CoA and further metabolized by reductive deamination to benzoyl-CoA.  相似文献   

6.
Kinetics of microperoxidase-11 (MP-11) as a heme–peptide enzyme model in oxidation reaction of guaiacol (AH) by hydrogen peroxide was studied in the presence of amino acids, taking into account the inactivation of MP-11 during reaction by its suicide substrate, H2O2. Reliability of the kinetic equation was evaluated by non-linear mathematical fitting. Fitting of experimental data into a new integrated kinetic relation showed a close match between the kinetic model and the experimental data. Indeed, it was found that the mechanism of suicide-peroxide inactivation of MP-11 in the presence of amino acids is different from MP-11 and/or horseradish peroxidase. In this mechanism, amino acids compete with hydrogen peroxide for the sixth co-ordination position of iron atom in the heme group through a competitive inhibition mechanism.The proposed model can successfully determine the kinetic parameters including inactivation by hydrogen peroxide as well as the inhibitory rate constants by the amino acid inhibitor.Kinetic parameters of inactivation including the initial activity of MP-11, α0, the apparent inactivation rate constant, ki and the apparent inhibition rate constant for cysteine, kI were obtained 0.282 ± 0.006 min?1, 0.497 ± 0.013 min?1 and 1.374 ± 0.007 min?1 at [H2O2] = 1.0 mM, 27 °C, phosphate buffer 5.0 mM, pH 7.0. Results showed that inactivation and inhibition of microperoxidase as a peroxidase model enzyme occurred simultaneously even at low concentrations of hydrogen peroxide (0.4 mM). This kinetic analysis based on the suicide-substrate inactivation of microperoxidase-11, provides a tool and model for studying peroxidase models in the presence of reversible inhibitors. The introduced inhibition procedure can be used in designing activity tunable and specific protected enzyme models in the hidden and reversibly inhibited forms, which do not undergo inactivation.  相似文献   

7.
Submitochondrial particles (SMP) were produced from Jerusalem artichoke (Helianthus tuberosus L.) mitochondria by sonication and differential centrifugation. The SMP were about 50% inside-out as measured by the access of reduced cytochrome c to cytochrome c oxidase. Uncoupled NADH oxidation (1 mM NADH) by the SMP was 120 nmol O2 min?1mg?1, which was reduced to 98 nmol O2 min?1 (mg mitochondrial protein)?1 in the presence of EGTA. In contrast, the oxidation of NADH by intact mitochondria was completely inhibited by EGTA (from 182 to 14 nmol O2 min?1mg?1). The EGTA-resistant NADH oxidation by the SMP is ascribed to the NADH dehydrogenase(s) on the inside of the inner membrane and exposed to the medium in the inside-out SMP. In the presence of EGTA it could be shown that two NADH dehydrogenase activities were present in the SMP. One had an apparent Km of 7 μM for NADH, a Vmax of 80 nmol NADH min?1mg?1, and was rotenone-sensitive. This dehydrogenase is equivalent to the mammalian Complex I NADH dehydrogenase. The other dehydrogenase, which was rotenone-resistant, had a Km of 80 μM and a Vmax of 131 nmol NADH min?1mg?1; it is probably responsible for the rotenone-resistant oxidation of organic acids often observed in plant mitochondria. The redox poise of the pyridine nucleotides had only a small effect on the relative rates of the two internal dehydrogenases. Electron flow through these dehydrogenases appears, therefore, to be regulated mainly by the concentration of NADH in the matrix of the mitochondria.  相似文献   

8.
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   

9.
Laccase-2 is a highly conserved multicopper oxidase that functions in insect cuticle pigmentation and tanning. In many species, alternative splicing gives rise to two laccase-2 isoforms. A comparison of laccase-2 sequences from three orders of insects revealed eleven positions at which there are conserved differences between the A and B isoforms. Homology modeling suggested that these eleven residues are not part of the substrate binding pocket. To determine whether the isoforms have different kinetic properties, we compared the activity of laccase-2 isoforms from Tribolium castaneum and Anopheles gambiae. We partially purified the four laccases as recombinant enzymes and analyzed their ability to oxidize a range of laccase substrates. The predicted endogenous substrates tested were dopamine, N-acetyldopamine (NADA), N-β-alanyldopamine (NBAD) and dopa, which were detected in T. castaneum previously and in A. gambiae as part of this study. Two additional diphenols (catechol and hydroquinone) and one non-phenolic substrate (2,2′-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid)) were also tested. We observed no major differences in substrate specificity between the A and B isoforms. Dopamine, NADA and NBAD were oxidized with catalytic efficiencies ranging from 51 to 550 min?1 mM?1. These results support the hypothesis that dopamine, NADA and NBAD are endogenous substrates for both isoforms of laccase-2. Catalytic efficiencies associated with dopa oxidation were low, ranging from 8 to 30 min?1 mM?1; in comparison, insect tyrosinase oxidized dopa with a catalytic efficiency of 201 min?1 mM?1. We found that dopa had the highest redox potential of the four endogenous substrates, and this property of dopa may explain its poor oxidation by laccase-2. We conclude that laccase-2 splice isoforms are likely to oxidize the same substrates in vivo, and additional experiments will be required to discover any isoform-specific functions.  相似文献   

10.
The synthesis of glutamate from 2-oxoglutarate generated by the citric acid cycle and ammonium acetate has been studied in brain mitochondria of synaptic or non synaptic origin. Non synaptic brain mitochondria synthesise glutamate at twice the rate (1.3 nmol. min?1. mg protein?1) of synaptic mitochondria (0.65 nmol. min?1. mg protein?1) when pyruvate is the precursor for 2-oxoglutarate, but at a similar rate (0.9 and 0.7 nmol. min?1, mg protein?1) when 3 hydroxybutyrate is the precursor. Glutamate synthesis from ammonium acetate and extramitochondrially addcd 2-oxoglutarate (5 mM) by both synaptic and nonsynaptic mitochondria was 5-fold higher (5-6nmol. min?1. mg protein?1) than glutamate synthesis from endogenously produced 2-oxoglutarate. In the uncoupled state (or un-coupler + oligomycin) the rate was reduced by half. (2.5-3 nmol. min?1. mg protein?1) as compared to mitochondria synthesising glutamate in states 3 or 4 (± oligomycin). The changes in brain mitochondrial nicotinamide nucleotide redox state have been monitored by fluorimetric, spectrophotometric and enzymatic techniques during glutamate synthesis and compared with liver mitochondria under similar conditions. On the instigation of glutamate synthesis by NH+4 addition a significant NAD(P)H oxidation occurs with liver mitochondria but no detectable change occurs with brain mitochondria. Leucine (2 mM) causes a doubling of glutamate synthesis by both synaptic and non synaptic brain mitochondria with no detectable change in the NAD(P)H redox state. The results are discussed with respect to the control of glutamate synthesis by mitochondrial redox potential and the possible intramitochondrial compartmentation of this process.  相似文献   

11.
Oxygenase‐containing cyanobacteria constitute promising whole‐cell biocatalysts for oxyfunctionalization reactions. Photosynthetic water oxidation thereby delivers the required cosubstrates, that is activated reduction equivalents and O2, sustainably. A recombinant Synechocystis sp. PCC 6803 strain showing unprecedentedly high photosynthesis‐driven oxyfunctionalization activities is developed, and its technical applicability is evaluated. The cells functionally synthesize a heterologous cytochrome P450 monooxygenase enabling cyclohexane hydroxylation. The biocatalyst‐specific reaction rate is found to be light‐dependent, reaching 26.3 ± 0.6 U gCDW?1 (U = μmol min?1 and cell dry weight [CDW]) at a light intensity of 150 µmolphotons m?2 s?1. In situ substrate supply via a two‐liquid phase system increases the initial specific activity to 39.2 ± 0.7 U gCDW?1 and stabilizes the biotransformation by preventing cell toxification. This results in a tenfold increased specific product yield of 4.5 gcyclohexanol gCDW?1 as compared to the single aqueous phase system. Subsequently, the biotransformation is scaled from a shake flask to a 3 L stirred‐tank photobioreactor setup. In situ O2 generation via photosynthetic water oxidation allows a nonaerated process operation, thus circumventing substrate evaporation as the most critical factor limiting the process performance and stability. This study for the first time exemplifies the technical applicability of cyanobacteria for aeration‐independent light‐driven oxyfunctionalization reactions involving highly toxic and volatile substrates.  相似文献   

12.
《Biomarkers》2013,18(4):267-272
Abstract

Sulphonamide hypersensitivity reactions are believed to be mediated through reactive intermediates derived from oxidation of the paraamino group to form sulphonamide hydroxylamines. Sulphamethoxazole hydroxylamine (SMX-HA) can be acetylated by N-acetyltransferase (NAT) enzymes to form an acetoxy metabolite (acetoxySMX). In the current studies, acetoxySMX was found to be not toxic over the concentration range of 0 to 500 μM towards a human lymphoblastoid cell line (RPMI 1788) or a human hepatoma cell line (HepG2). Further, transient expression of NAT1 in COS-1 cells or stable transfection of NAT1 andNAT2 in HepG2 cells did not alter the toxicity of SMX-HA in vitro. The activity of NAT1 in isolated mononuclear leucocytes (a reflection of systemic NAT1 activity) determined with paraaminobenzoic acid as a substrate was not different between controls (n = 11) or patients with a known hypersensitivity reaction (n = 5) (4.1 ±1.2 nmol min?1mg?1 vs 5.7 ± 1.4 nmol min?1 mg?1). Thus, acetoxy SMX is unlikely to play a significant role in mediating SMX hypersensitivity reactions anda constitutive deficiency in NAT1 activity is not a common finding in patients susceptible to SMX hypersensitivity reactions.  相似文献   

13.
The dose-dependent effect of intravenously infused synthetic somatostatin-14 on basal and postprandial insulin and gastrin release was assessed in anesthetized rats.Infusion of 1 ng · kg?1 · min?1 elicited a significant reduction of basal and postprandial insulin levels compared to the saline control group. At 15 ng · kg?1 · min?1 basal insulin was not affected but postprandial insulin levels were still significantly reduced. At 30 ng · kg?1 · min?1 neither basal nor stimulated insulin levels were affected. At the highest concentration of 120 ng · kg?1 · min?1 basal and postprandial insulin levels were suppressed similar to the lowest infusion rate of 1 ng · kg?1 · min?1. Basal gastrin levels were significantly reduced only at the highest rate of 120 ng · kg?1 · min?1. A significant reduction of postprandial gastrin levels was observed at 15 ng · kg?1 · min?1 and all higher infusion rates employed. Measurements of plasma somatostatin-like immunoreactivity (SLI) demonstrated that plasma SLI levels during the lowest infusion rate of 1 ng · kg?1 · min?1 were not different from the controls. No significant rise of plasma SLI levels was observed in response to the test meal. The higher infusion rates elicited a dose-dependent increase in plasma SLI levels. These data demonstrate that in rats somatostatin exerts a biological effect on insulin release at very low doses while certain greater infusion rates have no suppressive effect. Gastrin secretion is inhibited in a more linear pattern.  相似文献   

14.
Hydrogen peroxide amplifies the chemiluminescence in the oxidation of luminol by sodium hypochlorite. A linear relationship between concentration of hydrogen peroxide and light intensity was found in the concentration range 5 × 10?8?7.5 × 10?6 mol/l. At 7.5 × 10?6 mol/l H2O2 the chemiluminescence is amplified 550—fold. The chemiluminescence spectra of these reactions have a wavelength maximum at 431 nm independent of the concentration of hydrogen peroxide. The results indicate that hydrogen peroxide is a necessary component in the chemiluminescent oxidation of the luminol by sodium hypochlorite.  相似文献   

15.
The inhibition of neuraminidase from Clostridium chauvoei (jakari strain) with partially purified methanolic extracts of some plants used in Ethnopharmacological practice was evaluated. Extracts of two medicinal plants, Tamarindus indicus and Combretum fragrans at 100–1000 μg/ml, both significantly reduced the activity of the enzyme in a dose-dependent fashion (P < 0.001).

The estimated IC50 values for Tamarindus indicus and Combretum fragrans were 100 and 150 μ/ml respectively. Initial velocity studies conducted, using fetuin as substrate revealed a non-competitive inhibition with the Vmax significantly altered from 500 μmole min?1 mg?1 to 240μmole min?1 mg?1 and 340 μmole min?1 mg?1 in the presence of Tamarindus indicus and Combretum fragrans respectively. The KM remained unchanged at 0.42 mM. The computed Index of physiological efficiency was reduced from 1.19 min?1 to 0.57 min?1 and 0.75 min?1 with Tamarindus indicus and Combretum fragrans as inhibitors respectively.  相似文献   

16.
N-Nitrosodimethylamine (NDMA) is an emerging contaminant of concern. N-nitrodimethylamine (DMNA) is a structural analog to NDMA. NDMA and DMNA have been found in drinking water, groundwater, and other media and are of concern due their toxicity. The authors evaluated biotransformation of NDMA and DMNA by cultures enriched from contaminated groundwater growing on benzene, butane, methane, propane, or toluene. Maximum specific growth rates of enriched cultures on butane (μmax = 1.1 h?1) and propane (μmax = 0.65 h?1) were 1 to 2 orders of magnitude higher than those presented in the literature. Growth rates of mixed cultures grown on benzene (μmax = 1.3 h?1), methane (μmax = 0.09 h?1), and toluene (μmax = 0.99 h?1) in these studies were similar to those presented in the literature. NDMA biotransformation rates for methane oxidizers (υmax = 1.4 ng min?1 mg?1) and toluene oxidizers (υmax = 2.3 ng min?1 mg?1) were comparable to those presented in the literature, whereas the biotransformation rate for propane oxidizers (υmax = 0.37 ng min?1 mg?1) was lower. NDMA biotransformation rates for benzene oxidizers (υmax = 1.02 ng min?1 mg?1) and butane oxidizers (υmax = 1.2 ng min?1 mg?1) were comparable to those reported for other primary substrates. These studies showed that DMNA biotransformation rates for benzene (υmax = 0.79 ng min?1 mg?1), butane (υmax = 1.0 ng min?1 mg?1), methane (υmax = 2.1 ng min?1 mg?1), propane (υmax = 1.46 ng min?1 mg?1), and toluene (υmax = 0.52 ng min?1 mg?1) oxidizers were all comparable. These studies highlight potential bioremediation methods for NDMA and DMNA in contaminated groundwater.  相似文献   

17.
Abstract: A previous study of the metabolism of 6-[18F]-fluoro-l -3,4-dihydroxyphenylalanine (FDOPA) in rats pretreated with carbidopa contained information amenable to kinetic analysis. Using these data, tracer transfer coefficients and metabolic rate constants were estimated. After intravenous injection, FDOPA in circulation was O-methylated (kD0 = 0.055 min?1), and the metabolite (O-methyl-FDOPA) escaped from plasma with a rate constant (kM?1) of 0.01 min?1. The initial clearance of FDOPA to striatum (KD1) was 0.07 ml g?1 min?1, and the equilibrium distribution volume (VDe) was 0.67 ml g?1. The initial clearance of O-methyl-FDOPA to striatum (KM1) was 0.08 ml g?1 min?1, and the equilibrium distribution volume (VMe) was 0.75 ml g?1. The rate constant of FDOPA decarboxylation (kD3) was 0.17 min?1 in striatum. The elimination of 6-[18F]fluorodopamine (FDA) from striatum suggested an apparent rate constant for monoamine oxidase activity (k7) of 0.055 min?1. 6-[18F]Fluorohomovanillic acid (FHVA) was formed from 6-[18F]fluoro-l -3,4-dihydroxyphenylacetic acid with a rate constant (k11) of 0.083 min?1, and FHVA was eliminated from striatum (k9) with a rate constant of 0.12 min?1. The steady-state concentration ratios of FDA and its metabolites were shown to be functions of these rate constants.  相似文献   

18.
Starch suspensions have been treated with dilute, aqueous bromine at 30° in the pH range 6–8; no adsorption of oxidant occurred. The oxidation kinetics were first-order in bromine and in accordance with the rate law d [bromine]/dtk [starch] [bromine], except for a minor, initial rapid-phase in the oxidation of cereal starches, which is attributed to an enhanced reactivity of the granule surface. The apparent first-order rate-constants were 2.0–2.8 x 10?3 min?1, except for retrograded amylose oxidised at pH 8 when the value was 5.6 x 10?3 min?1. The i.r. spectra of the products indicated the presence of carboxylate and aldehyde groups. The functional group contents were determined quantitatively. Oxidation of the amylose at pH 6–7 introduced carbonyl groups, whereas at pH 8 carbonyl and carboxylate were found in equal amounts. For waxy-maize starch oxidised at pH 6–8, the carbonyl content was twice that of carboxylate. Acid hydrolysis of the product obtained by oxidation of amylose proceeded at pH 8 according to first-order kinetics. Chromatographic analysis of the anionic components of the hydrolysate indicated the presence of D-glucurono-6,3-lactone, D-gluconic acid, and an unidentified acidic ketose.  相似文献   

19.
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   

20.
The oxygen-transporting protein, hemocyanin (Hc), of the garden snail Helix aspersa maxima (HaH) was isolated and kinetically characterized. Kinetic parameters of the reaction of catalytic oxidation of catechol to quinone, catalyzed by native HaH were determined: the V max value amounted to 22 nmol min?1 mg?1, k cat to 1.1 min?1. Data were compared to those reported for other molluscan Hcs and phenoloxidases (POs). The o-diphenoloxidase activity of the native HaH is about five times higher than the activity determined for the Hcs of the terrestrial snail Helix pomatia and of the marine snail Rapana thomasiana (k cat values of 0.22 and 0.25 min?1, respectively). The K m values obtained for molluscan Hcs from different species are comparable to those for true POs, but the low catalytic efficiency of Hcs is probably related to inaccessibility of the active sites to potential substrates. Upon treatment of HaH with subtilisin DY, the enzyme activity against substrate catechol was considerably increased. The relatively high proteolytically induced o-diPO activity of HaH allowed using it for preparation of a biosensor for detection of catechol.  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