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 156 毫秒
1.
Alkene monooxygenase (AMO) from Rhodococcus rhodochrous (formerly Nocardia corallina) B-276 is a three-component enzyme system encoded by the four-gene operon amoABCD. AMO catalyses the stereoselective epoxygenation of aliphatic alkenes, yielding primarily R enantiomers. The presumed site of alkene oxygenation is a dinuclear iron centre similar to that in the soluble methane monooxygenases of methanotrophic bacteria, to which AMO exhibits a significant degree of amino acid sequence identity. The AMO complex was not expressed in Escherichia coli, at least partly because that host did not produce all of the AMO polypeptides. Expression of AMO was achieved in Streptomyces lividans by cloning the AMO genes into the thiostrepton-inducible expression plasmid pIJ6021. No background of AMO activity was detected in S. lividans cells without amoABCD and expression of AMO activity, at a level comparable to that from wild-type R. rhodochrous B-276, coincided with appearance of the AMO subunits. Recombinant AMO activity in cell-free extracts of S. lividans was stimulated by the addition of NADH and produced R-epoxypropane with comparable enantiomeric excess to AMO purified from the original organism. Although the whole AMO complex could not be expressed in E. coli, the functional coupling protein (AmoB) and reductase (AmoD) were expressed individually in E. coli as fusions with glutathione S-transferase. The expression systems described here now allow structure/function studies on AMO to be carried out by site-directed mutagenesis.  相似文献   

2.
Alkene monooxygenase (AMO) from Rhodococcus rhodochrous B-276 is a 3-component enzyme system encoded by the 4-gene operon amoABCD, which catalyzes the stereoselective epoxidation of aliphatic alkenes yielding primarily the R enantiomer. With propene as the substrate, wild-type AMO yields R-epoxypropane with an enantiomeric excess (e.e.) of 83%. The presumed site of alkene oxidation is a dinuclear iron center situated within the large subunit of the epoxygenase component, AmoC. Substantial problems with the expression of recombinant AMO were previously overcome. In this study, the authors have further developed this expression system to allow amoC to be subjected to mutagenesis by means of error-prone PCR, with the aim of developing a system that could be used to manipulate the enantioselectivity of the enzyme. The mutants were screened for altered stereoselectivity in the propene/epoxypropane reaction by a whole-cell assay, solvent extraction, and chiral gas chromatography analysis protocol that is suitable for scale up to several thousand mutants and that is estimated to detect differences in e.e. of as little as 5%.  相似文献   

3.
Soluble methane monooxygenase (sMMO) is a three-component enzyme that catalyses dioxygen- and NAD(P)H-dependent oxygenation of methane and numerous other substrates. Oxygenation occurs at the binuclear iron active centre in the hydroxylase component (MMOH), to which electrons are passed from NAD(P)H via the reductase component (MMOR), along a pathway that is facilitated and controlled by the third component, protein B (MMOB). We previously demonstrated that electrons could be passed to MMOH from a hexapeptide-modified gold electrode and thus cyclic voltammetry could be used to measure the redox potentials of the MMOH active site. Here we have shown that the reduction current is enhanced by the presence of catalase or if the reaction is performed in a flow-cell, probably because oxygen is reduced to hydrogen peroxide, by MMOH at the electrode surface and the hydrogen peroxide then inactivates the enzyme unless removed by catalase or a continuous flow of solution. Hydrogen peroxide production appears to be inhibited by MMOB, suggesting that MMOB is controlling the flow of electrons to MMOH as it does in the presence of MMOR and NAD(P)H. Most importantly, in the presence of MMOB and catalase, the electrode-associated MMOH oxygenates acetonitrile to cyanoaldehyde and methane to methanol. Thus the electochemically driven sMMO showed the same catalytic activity and regulation by MMOB as the natural NAD(P)H-driven reaction and may have the potential for development into an economic, NAD(P)H-independent oxygenation catalyst. The significance of the production of hydrogen peroxide, which is not usually observed with the NAD(P)H-driven system, is also discussed.  相似文献   

4.
The soluble methane monooxygenase (sMMO) complex from Methylococcus capsulatus (Bath) catalyses oxygen- and NAD(P)H-dependent oxygenation of methane, propene, and other substrates. Whole-complex sMMO oxygenase activity requires all three sMMO components: the hydroxylase, the reductase, and protein B. Also, in the presence of hydrogen peroxide, the hydroxylase alone catalyzes substrate oxygenation via the peroxide shunt reaction. We investigated the effect of amine cross-linking on hydroxylase activity to probe the role of a gross conformational change that occurs in the hydroxylase upon binding of the other protein components. The cross-linker inhibited hydroxylase activity in the whole complex, but this effect was due to covalent modification of primary amine groups rather than cross-linking. Covalent modification of arginine side-chains on the hydroxylase had a similar effect, but, most remarkably, neither form of modification affected the activity of the hydroxylase via the peroxide shunt reaction. It was shown that covalent modification of positively charged groups on the hydroxylase, which occurred at multiple sites, interfered with its physical and functional interactions with protein B and with the passage of electrons from the reductase. These results indicate that protein B and the reductase of the sMMO complex interact via positively charged groups on the surface of the hydroxylase to induce a conformational change that is necessary for delivery of electrons into the active site of the hydroxylase. Modification of positively charged groups on protein B had no effect on its function, consistent with the hypothesis that positively charged groups on the hydroxylase interact with negative charges on protein B. Thus, we have discovered a means of specifically inactivating the interactions between the sMMO complex while preserving the catalytic activity of the hydroxylase active site which provides a new method of studying intercomponent interactions within sMMO.  相似文献   

5.
Phenylacetylene was investigated as a differential inhibitor of ammonia monooxygenase (AMO), soluble methane monooxygenase (sMMO) and membrane-associated or particulate methane monooxygenase (pMMO) in vivo. At phenylacetylene concentrations > 1 microM, whole-cell AMO activity in Nitrosomonas europaea was completely inhibited. Phenylacetylene concentrations above 100 microM inhibited more than 90% of sMMO activity in Methylococcus capsulatus Bath and Methylosinus trichosporium OB3b. In contrast, activity of pMMO in M. trichosporium OB3b, M. capsulatus Bath, Methylomicrobium album BG8, Methylobacter marinus A45 and Methylomonas strain MN was still measurable at phenylacetylene concentrations up to 1,000 microM. AMO of Nitrosococcus oceanus has more sequence similarity to pMMO than to AMO of N. europaea. Correspondingly, AMO in N. oceanus was also measurable in the presence of 1,000 microM phenylacetylene. Measurement of oxygen uptake indicated that phenylacetylene acted as a specific and mechanistic-based inhibitor of whole-cell sMMO activity; inactivation of sMMO was irreversible, time dependent, first order and required catalytic turnover. Corresponding measurement of oxygen uptake in whole cells of methanotrophs expressing pMMO showed that pMMO activity was inhibited by phenylacetylene, but only if methane was already being oxidized, and then only at much higher concentrations of phenylacetylene and at lower rates compared with sMMO. As phenylacetylene has a high solubility and low volatility, it may prove to be useful for monitoring methanotrophic and nitrifying activity as well as identifying the form of MMO predominantly expressed in situ.  相似文献   

6.
Glutathione reductase has been found to catalyze an NAD(P)H-dependent electron transfer to 2,4,6-trinitrobenzenesulfonate (TNBS). In the presence of oxygen TNBS is not consumed in the reaction, but is rapidly reoxidized with concomitant production of hydrogen peroxide. Cytochrome c can replace oxygen as the final electron acceptor, indicating that a one-electron transfer takes place. The rate is slightly higher in the absence than in the presence of oxygen, ruling out superoxide anion as an obligatory intermediate in cytochrome c reduction. In the absence of oxygen (or cytochrome c), TNBS limits the reaction and accepts a total of four electrons. The TNBS-dependent NADPH (or NADH) oxidation is markedly stimulated by NADP+, and to a smaller extent also by NAD+. The TNBS-dependent reactions are inhibited by excess of NADPH but not by NADH. The kinetics of these reactions are consistent with a branching reaction mechanism in which a pathway including a ternary complex between the two-electron reduced enzyme and NADP+ has the highest turnover. NADPH-dependent reductions of ferricyanide or 2,6-dichloroindophenol catalyzed by glutathione reductase are also markedly influenced by NADP+. Evidently NADP+ facilitates a shift of the catalyzed reaction from the normal two-electron reduction of glutathione disulfide to a more unspecific one-electron reduction of other acceptors. Spectral as well as kinetic data suggest that the rate of radical formation limits the reactions with the artificial electron acceptors and that NADP+ promotes this rate-limiting step.  相似文献   

7.
Naphthalene 1,2-dioxygenase (NDOS) catalyzes the NAD(P)H and O(2)-dependent oxidation of naphthalene to (+)-cis-(1R,2S)-dihydroxy-1,2-dihydronaphthalene. NDOS consists of three protein components: a flavo-[2Fe-2S] reductase (NDR), a ferredoxin electron transfer protein (NDF), and an (alphabeta)(3) oxygenase (NDO) containing a mononuclear iron site and a Rieske-type [2Fe-2S] cluster in each alpha-subunit. The active site is built across a subunit-subunit boundary, and each subunit contributes one type of metal center. Our previous studies have shown that NDO with both metal centers reduced is capable of an O(2)-coupled single turnover to yield the correct cis-diol product in the absence of the NDR and NDF components (Wolfe, M. D., Parales, J. V., Gibson, D. T., and Lipscomb, J. D. (2001) J. Biol. Chem. 276, 1945-1953). It is shown here that addition of H(2)O(2) to NDO allows reaction with naphthalene to rapidly yield the correct product in a "peroxide shunt" reaction that does not require a reduced Rieske cluster. The mononuclear Fe(2+) center is oxidized during turnover, while the Rieske cluster remains in the oxidized state. Peroxide shunt turnover in the presence of (18)O-labeled H(2)O(2), H(2)O, or O(2) shows that both oxygen atoms in the product derive primarily from H(2)O(2). The peroxide shunt halts after one turnover despite the presence of excess H(2)O(2) and naphthalene, but this is not the result of enzyme inactivation. Rather, it appears that the product cannot be released when the mononuclear iron is in the Fe(3+) state, blocking a second turnover. This work supports the hypotheses that the cis-dihydroxylation activity of NDOS requires only the NDO component, that a peroxo intermediate is formed during normal catalysis, and that product release requires an additional reducing equivalent beyond those necessary for the first turnover.  相似文献   

8.
Concentrated urine formation in the kidney is accompanied by conditions that favor the accumulation of reactive oxygen species (ROS). Under hyperosmotic conditions, medulla cells accumulate glycine betaine, which is an osmolyte synthesized by betaine aldehyde dehydrogenase (BADH, EC 1.2.1.8). All BADHs identified to date have a highly reactive cysteine residue at the active site, and this cysteine is susceptible to oxidation by hydrogen peroxide. Porcine kidney BADH incubated with H(2)O(2) (0-500 μM) lost 25% of its activity. However, pkBADH inactivation by hydrogen peroxide was limited, even after 120 min of incubation. The presence of coenzyme NAD(+) (10-50 μM) increased the extent of inactivation (60%) at 120 min of reaction, but the ligands betaine aldehyde (50 and 500 μM) and glycine betaine (100 mM) did not change the rate or extent of inactivation as compared to the reaction without ligand. 2-Mercaptoethanol and dithiothreitol, but not reduced glutathione, were able to restore enzyme activity. Mass spectrometry analysis of hydrogen peroxide inactivated BADH revealed oxidation of M278, M243, M241 and H335 in the absence and oxidation of M94, M327 and M278 in the presence of NAD(+). Molecular modeling of BADH revealed that the oxidized methionine and histidine residues are near the NAD(+) binding site. In the presence of the coenzyme, these oxidized residues are proximal to the betaine aldehyde binding site. None of the oxidized amino acid residues participates directly in catalysis. We suggest that pkBADH inactivation by hydrogen peroxide occurs via disulfide bond formation between vicinal catalytic cysteines (C288 and C289).  相似文献   

9.
Renal hyperosmotic conditions may produce reactive oxygen species, which could have a deleterious effect on the enzymes involved in osmoregulation. Hydrogen peroxide was used to provoke oxidative stress in the environment of betaine aldehyde dehydrogenase in vitro. Enzyme activity was reduced as hydrogen peroxide concentration was increased. Over 50% of the enzyme activity was lost at 100 μM hydrogen peroxide at two temperatures tested. At pH 8.0, under physiological ionic strength conditions, peroxide inhibited the enzyme. Initial velocity assays of betaine aldehyde dehydrogenase in the presence of hydrogen peroxide (0-200 μM) showed noncompetitive inhibition with respect to NAD(+) or to betaine aldehyde at saturating concentrations of the other substrate at pH 7.0 or 8.0. Inhibition data showed that apparent V(max) decreased 40% and 26% under betaine aldehyde and NAD(+) saturating concentrations at pH 8.0, while at pH 7.0 V(max) decreased 40% and 29% at betaine aldehyde and NAD(+) saturating concentrations. There was little change in apparent Km(NAD) at either pH, while Km(BA) increased at pH 7.0. K(i) values at pH 8 and 7 were calculated. Our results suggest that porcine kidney betaine aldehyde dehydrogenase could be inhibited by hydrogen peroxide in vivo, thus compromising the synthesis of glycine betaine.  相似文献   

10.
The inactivation of glutamine synthetase (GS; EC 6.3.1.2) by metal-catalyzed oxidation (MCO) systems was studied in several Prochlorococcus strains, including the axenic PCC 9511. GS was inactivated in the presence of various oxidative systems, either enzymatic (as NAD(P)H+NAD(P)H-oxidase+Fe(3+)+O(2)) or non-enzymatic (as ascorbate+Fe(3+)+O(2)). This process required the presence of oxygen and a metal cation, and is prevented under anaerobic conditions. Catalase and peroxidase, but not superoxide dismutase, effectively protected the enzyme against inactivation, suggesting that hydrogen peroxide mediates this mechanism, although it is not directly responsible for the reaction. Addition of azide (an inhibitor of both catalase and peroxidase) to the MCO systems enhanced the inactivation. Different thiols induced the inactivation of the enzyme, even in the absence of added metals. However, this inactivation could not be reverted by addition of strong oxidants, as hydrogen peroxide or oxidized glutathione. After studying the effect of addition of the physiological substrates and products of GS on the inactivation mechanism, we could detect a protective effect in the case of inorganic phosphate and glutamine. Immunochemical determinations showed that the concentration of GS protein significantly decreased by effect of the MCO systems, indicating that inactivation precedes the degradation of the enzyme.  相似文献   

11.
Cytochrome c is degraded by a large excess of hydrogen peroxide, leading to opening of the heme porphyrin ring and loss of the Soret absorption bands. The kinetic parameters of this reaction have been determined, and it is shown that a small concentration of oxygen is liberated at the same rate as degradation. Low-level chemiluminescence and release of a hydroxylating species also accompany heme destruction. It is proposed that heme iron activates hydrogen peroxide to a more powerful oxidant, perhaps the hydroxyl radical, which remains bound to the heme iron and initiates attack on the porphyrin ring. Chemiluminescence appears to result from a side reaction involving singlet oxygen attack on the alpha-methene bridge, yielding a dioxetane. The in vivo degradation of cytochrome c by excess hydrogen peroxide may interfere with respiration, accelerate aging, and enhance the metabolism of carcinogens.  相似文献   

12.
Radish plasmalemma-enriched fractions show an NAD(P)H-ferricyanide or NAD(P)H-cytochrome c oxidoreductase activity which is not influenced by pH in the 4.5-7.5 range. In addition, at pH 4.5-5.0, NAD(P)H elicits an oxygen consumption (NAD(P)H oxidation) inhibited by catalase or superoxide dismutase (SOD), added either before or after NAD(P)H addition. Ferrous ions stimulate NAD(P)H oxidation, which is again inhibited by SOD and catalase. Hydrogen peroxide does not stimulate NADH oxidation, while it does stimulate Fe2+-induced NADH oxidation. NADH oxidation is unaffected by salicylhydroxamic acid and Mn2+, is stimulated by ferulic acid, and inhibited by KCN, EDTA and ascorbic acid. Moreover, NADH induces the conversion of epinephrine to adrenochrome, indicating that anion superoxide is formed during its oxidation. These results provide evidence that radish plasma membranes contain an NAD(P)H-ferricyanide or cytochrome c oxidoreductase and an NAD(P)H oxidase, active only at pH 4.5-5.0, able to induce the formation of anion superoxide, that is then converted to hydrogen peroxide. Ferrous ions, sparking a Fenton reaction, would stimulate NAD(P)H oxidation.  相似文献   

13.
Oxygenated and peroxy states of the cytochrome d complex of Escherichia coli have been proposed as intermediates in the reaction mechanism of this ubiquinol oxidase. In this report, several stable states of the purified enzyme were examined spectroscopically at room temperature. As purified, the cytochrome d complex exists in an oxygenated state characterized by an absorbance band at 650 nm. Removal of oxygen results in loss of absorbance at this wavelength, which is restored upon the return of oxygen. The presence of one oxygen molecule in the oxygenated state was quantified by measuring oxygen released when excess hydrogen peroxide was added to the oxygenated state by passage of argon generates a "partially reduced" state with an absorbance peak at 628 nm, apparently due to reduced cytochrome d. Addition of equimolar hydrogen peroxide to the fully oxidized state produces the peroxy state. This peroxy state is also formed upon addition of excess hydrogen peroxide to the oxygenated state via a stable intermediate termed "peroxy intermediate." It is likely that 1) the oxygenated state consists of one molecule of oxygen bound to reduced heme d, and 2) there are at least two stable states that have bound peroxide at room temperature, the peroxy state and a newly discovered peroxy intermediate.  相似文献   

14.
Localization of fatty acid beta-oxidation system in microbodies of Candida tropicalis cells growing on n-alkanes was studied. Microbodies isolated from the yeast cells showed palmitate-dependent activities of NAD reduction, acetyl-CoA formation and oxygen consumption. When sodium azide, an inhibitor of catalase, was added to the system, palmitate-dependent formation of hydrogen peroxide was observed. Stoichiometric study revealed that two moles of NAD were reduced per one mole of oxygen consumed in the absence of sodium azide and the presence of the inhibitor doubled the oxygen consumption by microbodies without an appreciable change in NAD reduction. These results indicate that the yeast microbodies contain beta-oxidation system of fatty acid, and that catalase located in the organelles participates in the degradation of hydrogen peroxide to be formed at the step of dehydrogenation of acyl-CoA.  相似文献   

15.
Soluble methane mono-oxygenase (sMMO) of Methylococcus capsulatus (Bath) catalyses the O2-dependent and NAD(P)H-dependent oxygenation of methane and numerous other substrates. During purification, the sMMO enzyme complex, which comprises three components and has a molecular mass in excess of 300 kDa, becomes inactivated because of cleavage of just 12 amino acids from the N-terminus of protein B, which is the smallest component of sMMO and the only one without prosthetic groups. Here we have shown that cleavage of protein B, to form the inactive truncated protein B', continued to occur when intact protein B was repeatedly separated from protein B' and all detectable contaminants, giving compelling evidence that the protein was cleaved autocatalytically. The rate of autocatalytic cleavage decreased when the residues flanking the cleavage site were mutated, but the position of cleavage was unaltered. Analysis of a series of incremental truncates showed that residue(s) essential for the activity of sMMO, and important in determining the stability of protein B, lay in the region Ser4-Tyr7. Protein B was shown to possess intrinsic nucleophilic activity, which we propose initiates the cleavage reaction via a novel mechanism. Proteins B and B' were detected in approximately equal amounts in the cell, showing that truncation of protein B is biologically relevant. Increasing the growth-medium copper concentration, which inactivates sMMO, did not alter the extent of in vivo cleavage, therefore the conditions under which cleavage of protein B may fulfil its proposed role as a regulator of sMMO remain to be identified.  相似文献   

16.
A purely chemical system for NAD(P)H oxidation to biologically active NAD(P)+ has been developed and characterized. Suitable amounts of EDTA, manganous ions and mercaptoethanol, combined at physiological pH, induce nucleotide oxidation through a chain length also involving molecular oxygen, which eventually undergoes quantitative reduction to hydrogen peroxide. Mn2+ is specifically required for activity, while both EDTA and mercaptoethanol can be replaced by analogs. Optimal molar ratios of chelator/metal ion (2:1) yield an active coordination compound which catalyzes thiol autoxidation to thiyl radical. The latter is further oxidized to disulfide by molecular oxygen whose one-electron reduction generates superoxide radical. Superoxide dismutase (SOD) inhibits both thiol oxidation and oxygen consumption as well as oxidation of NAD(P)H if present in the mixture. A tentative scheme for the chain length occurring in the system is proposed according to stoichiometry of reactions involved. Two steps appear of special importance in nucleotide oxidation: (a) the supposed transient formation of NAD(P). from the reaction between NAD(P)H and thiyl radicals; (b) the oxidation of the reduced complex by superoxide to keep thiol oxidation cycling.  相似文献   

17.
Redox reactions were carried out in aerobiosis and anaerobiosis between NAD(P) dimers or NAD(P)H and pyrroloquinoline quinone (PQQ) in different buffers. The buffer system and pH significantly affected the oxidation rates of nucleotides and the ESR signal intensity of the PQQ(*) radical formed in anaerobiosis by comproportion between the quinone and quinol forms. The relative reactivity of the four nucleotides toward PQQ was affected by pH and buffer nature. PQQ, which behaves as an electron shuttle from nucleotides to oxygen, was first converted to PQQH(2) and then rapidly reoxidized by oxygen, with formation of hydrogen peroxide. Both NAD(P) dimers and NAD(P)H consumed 1 mol of oxygen per mole of reacted molecule of pyridine nucleotide, yielding 1 or 2 mol of NAD(P)(+) from NAD(P)H or from NAD(P) dimers, respectively. Chelating agents such as EDTA and phytate strongly decreased the reaction rate and the PQQ(*) radical signal intensity. Kinetics carried out in the presence of metal ions showed instead an increased reaction rate in the order Ca(2+) > Mg(2+) > Na(+) > K(+). Spectrofluorimetric measurements of PQQ with increasing concentrations of Ca(2+) showed a fluorescence quenching and shift of the maximum emission toward lower wavelengths, while other metal ions showed minor effects, if any. Therefore, it is demonstrated that Ca(2+) binds to PQQ, probably forming a complex which is more reactive with both one-electron (NAD(P) dimers) or two-electron donors (NAD(P)H) in nonenzymic reactions. It is important to recall that Ca(2+) was already found to play active role in PQQ-containing enzymes.  相似文献   

18.
The inactivation of glutamine synthetase (GS; EC 6.3.1.2) by metal-catalyzed oxidation (MCO) systems was studied in several Prochlorococcus strains, including the axenic PCC 9511. GS was inactivated in the presence of various oxidative systems, either enzymatic (as NAD(P)H+NAD(P)H-oxidase+Fe3++O2) or non-enzymatic (as ascorbate+Fe3++O2). This process required the presence of oxygen and a metal cation, and is prevented under anaerobic conditions. Catalase and peroxidase, but not superoxide dismutase, effectively protected the enzyme against inactivation, suggesting that hydrogen peroxide mediates this mechanism, although it is not directly responsible for the reaction. Addition of azide (an inhibitor of both catalase and peroxidase) to the MCO systems enhanced the inactivation. Different thiols induced the inactivation of the enzyme, even in the absence of added metals. However, this inactivation could not be reverted by addition of strong oxidants, as hydrogen peroxide or oxidized glutathione. After studying the effect of addition of the physiological substrates and products of GS on the inactivation mechanism, we could detect a protective effect in the case of inorganic phosphate and glutamine. Immunochemical determinations showed that the concentration of GS protein significantly decreased by effect of the MCO systems, indicating that inactivation precedes the degradation of the enzyme.  相似文献   

19.
Nagababu E  Rifkind JM 《Biochemistry》2000,39(40):12503-12511
The reaction of Fe(II) hemoglobin (Hb) but not Fe(III) hemoglobin (metHb) with hydrogen peroxide results in degradation of the heme moiety. The observation that heme degradation was inhibited by compounds, which react with ferrylHb such as sodium sulfide, and peroxidase substrates (ABTS and o-dianisidine), demonstrates that ferrylHb formation is required for heme degradation. A reaction involving hydrogen peroxide and ferrylHb was demonstrated by the finding that heme degradation was inihibited by the addition of catalase which removed hydrogen peroxide even after the maximal level of ferrylHb was reached. The reaction of hydrogen peroxide with ferrylHb to produce heme degradation products was shown by electron paramagnetic resonance to involve the one-electron oxidation of hydrogen peroxide to the oxygen free radical, superoxide. The inhibition by sodium sulfide of both superoxide production and the formation of fluorescent heme degradation products links superoxide production with heme degradation. The inability to produce heme degradation products by the reaction of metHb with hydrogen peroxide was explained by the fact that hydrogen peroxide reacting with oxoferrylHb undergoes a two-electron oxidation, producing oxygen instead of superoxide. This reaction does not produce heme degradation, but is responsible for the catalytic removal of hydrogen peroxide. The rapid consumption of hydrogen peroxide as a result of the metHb formed as an intermediate during the reaction of reduced hemoglobin with hydrogen peroxide was shown to limit the extent of heme degradation.  相似文献   

20.
A compact automated analyser which could analyse constituents in biological fluids with a small sample volume and in a short time has been developed. The instrument was composed of a flow injection analysis system equipped with chemiluminometric detection and an immobilized enzyme column reactor used in combination. Chemiluminescence has high sensitivity, and its reaction proceeds very quickly. Furthermore, an immobilized enzyme column reactor can produce a sufficient amount of hydrogen peroxide from compounds in serum in a short time. When enzymes are used as reagents for the analysis of substances in blood or blood serum, the final signals emitted by different enzyme reactions are usually not only hydrogen peroxide but also ammonia, NAD(P)H and so on. However, the practical chemiluminescence method for ammonia and NAD(P)H has not been established. We have discovered a new practical method for ammonia and NAD(P)H using an enzyme column reactor consisting of both immobilized L -glutamate dehydrogenase and L -glutamate oxidase. The determinations of glucose and uric acid in serum by chemiluminometry after production of hydrogen peroxide by the respective oxidases are presented. A newly chemiluminometric determination of ammonia, NAD(P)H and its applications to other enzymatic analyses that give ammonia and NAD(P)H as a final signal are also described.  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