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.
A homogeneous Mn-dependent peroxidase (MnP) was purified from the extracellular culture fluid of the lignin-degrading white rot fungus Phlebia radiata by anion exchange chromatography. The enzyme had a molecular weight of 49,000 and pI 3.8. It was a glycoprotein, containing carbohydrate moieties accounting for 10% of the molecular weight. Mn-peroxidase was capable of oxidizing phenolic compounds in the presence of H2O2, whereas the effect on nonphenolic lignin model compounds was insignificant. MnP contained protoporphyrin IX as a prosthetic group. During enzymatic reactions H2O2 converted the native MnP to compound II. Mn2+ was essential in completing the catalytic cycle by returning the enzyme to its native state. The oxidation of ultimate substrates was dependent on superoxide radicals, O2- and probably on Mn3+ generated during the catalytic cycle. MnP exhibited high activity of NADH oxidation without exogenously added H2O2. It was shown to produce H2O2 at the expense of NADH.  相似文献   

2.
A novel preparation method for surfactant-MnP-Mn(II) ternary complex utilizing water-in-oil emulsions has been developed. The surfactant-MnP complex was spectroscopically characterized, strongly suggesting that the heme environment of the surfactant-MnP complex in benzene is identical to that of native MnP in the aqueous buffer. o-Phenylenediamine oxidation catalyzed by the surfactant-MnP-Mn(II) ternary complex was performed in benzene. The ternary complex efficiently catalyzed the oxidation, and the complex was catalytically stable. Kinetic experiments revealed that the reaction mechanism was as follows: MnP is oxidized by H(2)O(2) and the oxidized intermediate catalyzes the oxidation of Mn(II) to Mn(III) and the latter, after complexed with malonate, readily oxidizes o-PDA inside the complex. Thus, the organic substrate o-PDA, but not Mn(III), shuttled between the surfactant-MnP-Mn(II) ternary complex and organic solvent.  相似文献   

3.
Mn-peroxidase (MnP), a biotechnologically important enzyme was purified for the first time from a plant source Musa paradisiaca (banana) stem, which is an agro-waste easily available after harvest of banana fruits. MnP was earlier purified only from the fungal sources. The enzyme was purified from stem juice by ultrafiltration and anion-exchange column chromatography on diethylamino ethylcellulose with 8-fold purification and purification yield of 65%. The enzyme gave a single protein band in SDS-PAGE corresponding to molecular mass 43 kDa. The Native-PAGE of the enzyme also gave a single protein band, confirming the purity of the enzyme. The UV/VIS spectrum of the purified enzyme differed from the other heme peroxidases, as the Soret band was shifted towards lower wavelength and the enzyme had an intense absorption band around 250 nm. The K(m) values using MnSO4 and H2O2 as the substrates of the purified enzyme were 21.0 and 9.5 microM, respectively. The calculated k(cat) value of the purified enzyme using Mn(II) as the substrate in 50 mM lactate buffer (pH 4.5) at 25 degrees C was 6.7s(-1), giving a k(cat)/K(m) value of 0.32 microM(-1)s(-1). The k(cat) value for the MnP-catalyzed reaction was found to be dependent of the Mn(III) chelator molecules malonate, lactate and oxalate, indicating that the enzyme oxidized chelated Mn(II) to Mn(III). The pH and temperature optima of the enzyme were 4.5 and 25 degrees C, respectively. The enzyme in combination with H2O2 liberated bromine and iodine in presence of KBr and KI respectively. All these enzymatic characteristics were similar to those of fungal MnP. The enzyme has the potential as a green brominating and iodinating agent in combination with KBr/KI and H2O2.  相似文献   

4.
NADH oxidation by manganese peroxidase (MnP) was done in a reaction mixture including either alpha-hydroxy acid or acetate. The oxidation in the former reaction mixture was inhibited by a catalase and was accelerated by exogenous H2O2, while the oxidation in the latter reaction mixture was inhibited by a superoxide dismutase and was not accelerated by the exogenous H2O2. These results indicated that there are significant differences between the two reaction systems, particularly, in the active oxygen species involved in the reactions. Additionally, the experiment of MnP reduction with Mn(II) suggests that MnP has a separate catalytic activity other than an oxidation of Mn(II) to Mn(III) in the reaction mixture including acetate.  相似文献   

5.
Purified manganese peroxidase (MnP) from Phanerochaete chrysosporium oxidizes nonphenolic beta-1 diarylpropane lignin model compounds in the presence of Tween 80, and in three- to fourfold lower yield in its absence. In the presence of Tween 80, 1-(3',4'-diethoxyphenyl)-1-hydroxy-2-(4'-methoxyphenyl)propane (I) was oxidized to 3,4-diethoxybenzaldehyde (II), 4-methoxyacetophenone (III) and 1-(3',4'-diethoxyphenyl)-1-oxo-2-(4'-methoxyphenyl)propane (IV), while only 3,4-diethoxybenzaldehyde (II) and 4-methoxyacetophenone (III) were detected when the reaction was conducted in the absence of Tween 80. In contrast to the oxidation of this substrate by lignin peroxidase (LiP), oxidation of substrates by MnP did not proceed under anaerobic conditions. When the dimer (I) was deuterated at the alpha position and subsequently oxidized by MnP in the presence of Tween 80, yields of 3,4-diethoxybenzaldehyde, 4-methoxyacetophenone remained constant, while the yield of the alpha-keto dimeric product (IV) decreased by approximately sixfold, suggesting the involvement of a hydrogen abstraction mechanism. MnP also oxidized the alpha-keto dimeric product (IV) to yield 3,4-diethoxybenzoic acid (V) and 4-methoxyacetophenone (III), in the presence and, in lower yield, in the absence of Tween 80. When the reaction was performed in the presence of 18O2, both products, 3,4-diethoxybenzoic acid and 4-methoxyacetophenone, contained one atom of 18O. Finally, MnP oxidized the substrate 1-(3',5'-dimethoxyphenyl)-1-hydroxy-2-(4'-methoxyphenyl)propane (IX) to yield 3,5-dimethoxybenzaldehyde (XI), 4-methoxyacetophenone (III) and 1-(3',5'-dimethoxyphenyl)-1-oxo-2-(4'-methoxyphenyl)propane (X). In sharp contrast, LiP was not able to oxidize IX. Based on these results, we propose a mechanism for the MnP-catalyzed oxidation of these dimers, involving hydrogen abstraction at a benzylic carbon, rather than electron abstraction from an aromatic ring.  相似文献   

6.
A Mn(II)-dependent peroxidase found in the extracellular medium of ligninolytic cultures of the white rot fungus, Phanerochaete chrysosporium, was purified by DEAE-Sepharose ion-exchange chromatography, Blue Agarose chromatography, and gel filtration on Sephadex G-100. Sodium dodecyl sulfate-gel electrophoresis indicated that the homogeneous protein has an Mr of 46,000. The absorption spectrum of the enzyme indicates the presence of a heme prosthetic group. The pyridine hemochrome absorption spectrum indicates that the enzyme contained one molecule of heme as iron protoporphyrin IX. The absorption maximum of the native enzyme (406 nm) shifted to 433 nm in the reduced enzyme and to 423 nm in the reduced-CO complex. Both CN- and N-3 readily bind to the native enzyme, indicating an available coordination site and that the heme iron is high spin. The absorption spectrum of the H2O2 enzyme complex, maximum at 420 nm, is similar to that of horseradish peroxidase compound II. P. chrysosporium peroxidase activity is dependent on Mn(II), with maximal activity attained above 100 microM. The enzyme is also stimulated to varying degrees by alpha-hydroxy acids (e.g., malic, lactic) and protein (e.g., gelatin, albumin). The peroxidase is capable of oxidizing NADH and a wide variety of dyes, including Poly B-411 and Poly R-481. Several of the substrates (indigo trisulfonate, NADH, Poly B-411, variamine blue RT salt, and Poly R-481) are oxidized by this Mn(II)-dependent peroxidase at considerably faster rates than those catalyzed by horseradish peroxidase. The enzyme rapidly oxidizes Mn(II) to Mn(III); the latter was detected by the characteristic absorption spectrum of its pyrophosphate complex. Inhibition of the oxidation of the substrate diammonium 2,2-azino-bis(3-ethyl-6-benzothiazolinesulfonate) (ABTS) by Na-pyrophosphate suggests that Mn(III) plays a role in the enzyme mechanism.  相似文献   

7.
Manganese peroxidase (MnP) is one of two extracellular peroxidases believed to be involved in lignin biodegradation by the white-rot basidiomycete Phanerochaete chrysosporium. The enzyme oxidizes Mn(II) to Mn(III), which accumulates in the presence of Mn(III) stabilizing ligands. The Mn(III) complex in turn can oxidize a variety of organic substrates. The stoichiometry of Mn(III) complex formed per hydrogen peroxide consumed approaches 2:1 as enzyme concentration increases at a fixed concentration of peroxide or as peroxide concentration decreases at a fixed enzyme concentration. Reduced stoichiometry below 2:1 is shown to be due to Mn(III) complex decomposition by hydrogen peroxide. Reaction of Mn(III) with peroxide is catalyzed by Cu(II), which explains an apparent inhibition of MnP by Cu(II). The net decomposition of hydrogen peroxide to form molecular oxygen also appears to be the only observable reaction in buffers that do not serve as Mn(III) stabilizing ligands. The nonproductive decomposition of both Mn(III) and peroxide is an important finding with implications for proposed in vitro uses of the enzyme and for its role in lignin degradation. Steady-state kinetics of Mn(III) tartrate and Mn(III) malate formation by the enzyme are also described in this paper, with results largely corroborating earlier findings by others. Based on a comparison of pH effects on the kinetics of enzymatic Mn(III) tartrate and Mn(III) malate formation, it appears that pH effects are not due to ionizations of the Mn(III) complexing ligand.  相似文献   

8.
The hybrid Mn-peroxidase of the fungus Panus tigrinus 8/18 oxidized NADH in the absence of hydrogen peroxide, this being accompanied by the consumption of oxygen. The reaction of NADH oxidation started after a period of induction and completely depended on the presence of Mn(II). The reaction was inhibited in the presence of catalase and super-oxide dismutase. Oxidation of NADH by the enzyme or by manganese(III)acetate was accompanied by the production of hydrogen peroxide and superoxide radicals. In the presence of NADH, the enzyme was transformed into a catalytically inactive oxidized form (compound III), and the latter was inactivated with bleaching of the heme. The substrate of the hybrid Mn-peroxidase (Mn(II)) reduced compound III to yield the native form of the enzyme and prevented its inactivation. It is assumed that the hybrid Mn-peroxidase used the formed hydrogen peroxide in the usual peroxidase reaction to produce Mn(III), which was involved in the formation of hydrogen peroxide and thus accelerated the peroxidase reaction. The reaction of NADH oxidation is a peroxidase reaction and the consumption of oxygen is due to its interaction with the products of NADH oxidation. The role of Mn(II) in the oxidation of NADH consisted in the production of hydrogen peroxide and the protection of the enzyme from inactivation.__________Translated from Biokhimiya, Vol. 70, No. 4, 2005, pp. 568–574.Original Russian Text Copyright © 2005 by Lisov, Leontievsky, Golovleva.  相似文献   

9.
The peroxidation of C18 unsaturated fatty acids by fungal manganese peroxidase (MnP)/Mn(II) and by chelated Mn(III) was studied with application of three different methods: by monitoring oxygen consumption, by measuring conjugated dienes and by thiobarbituric acid-reactive substances (TBARS) formation. All tested polyunsaturated fatty acids (PUFAs) were oxidized by MnP in the presence of Mn(II) ions but the rate of their oxidation was not directly related to degree of their unsaturation. As it has been shown by monitoring oxygen consumption and conjugated dienes formation the linoleic acid was the most easily oxidizable fatty acid for MnP/Mn(II) and chelated Mn(III). However, when the lipid peroxidation (LPO) activity was monitored by TBARS formation the linolenic acid gave the highest results. High accumulation of TBARS was also recorded during peroxidation of linoleic acid initiated by MnP/Mn(II). Action of Mn(III)-tartrate on the PUFAs mimics action of MnP in the presence of Mn(II) indicating that Mn(III) ions are involved in LPO initiation. Although in our experiments Mn(III) tartrate gave faster than MnP/Mn(II) initial oxidation of the unsaturated fatty acids with consumption of O2 and formation of conjugated dienes the process was not productive and did not support further development of LPO. The higher effectiveness of MnP/Mn(II)-initiated LPO system depends on the turnover of manganese provided by MnP. It is proposed that the oxygen consumption assay is the best express method for evaluation of MnP- and Mn(III)-initiated peroxidation of C18 unsaturated fatty acids.  相似文献   

10.
Manganese peroxidase (MnP), which normally oxidizes Mn2+ to Mn3+, is rapidly and completely inactivated in an H2O2-dependent reaction by 2 equivalents of sodium azide. The inactivation is paralleled by formation of the azidyl radical and high yield conversion of the prosthetic heme into a meso-azido adduct. The meso-azido enzyme is oxidized by H2O2 to a Compound II-like species with the Soret band red-shifted 2 nm relative to that of native Compound II. The time-dependent decrease in this Compound II-like spectrum (t1/2 = 2.3 h) indicates that the delta-meso azido heme is more rapidly degraded by H2O2 than the prosthetic heme of control enzyme (t1/2 = 4.8 h). MnP is also inactivated by phenyl-, methyl-, and ethylhydrazine. The phenylhydrazine reaction is too rapid for kinetic analysis, but KI = 402 microM and kinact = 0.22/min for the slower inactivation by methylhydrazine. Reaction with phenylhydrazine at pH 4.5 does not yield iron-phenyl, N-phenyl, or meso-phenyl heme adducts. Ethylhydrazine inactivates the enzyme both at pH 4.5 and 7.0, but only detectably produces delta-meso-ethyl-heme at pH 7.0. Reconstitution of apo-MnP with hemin or delta-meso-ethylheme yields enzyme with, respectively, 50 and 5% of the native activity. The delta-meso-alkyl group thus suppresses most of the catalytic activity of the enzyme even though a Compound II-like species is still formed with H2O2. Finally, Co2+ inhibits the enzyme competitively with respect to Mn2+ but does not inhibit its inactivation by azide or the alkylhydrazines. The results argue that substrates interact with the heme edge in the vicinity of the delta-meso-carbon. They also suggest that Mn2+ and Co2+ bind to a common site close to the delta-meso-carbon without blocking the approach of small molecules to the heme edge. An active site model is proposed that accommodates these results.  相似文献   

11.
Site-directed mutations R177A and R177K in the gene encoding manganese peroxidase isozyme 1 (mnp1) from Phanerochaete chrysosporium were generated. The mutant enzymes were expressed in P. chrysosporium during primary metabolic growth under the control of the glyceraldehyde-3-phosphate dehydrogenase gene promoter, purified to homogeneity, and characterized by spectroscopic and kinetic methods. The UV-vis spectra of the ferric and oxidized states and resonance Raman spectra of the ferric state were similar to those of the wild-type enzyme, indicating that the heme environment was not significantly affected by the mutations at Arg177. Apparent K(m) values for Mn(II) were approximately 20-fold greater for the R177A and R177K MnPs than for wild-type MnP. However, the apparent K(m) values for the substrates, H(2)O(2) and ferrocyanide, and the k(cat) values for Mn(II) and ferrocyanide oxidation were similar to those of the wild-type enzyme. The second-order rate constants for compound I (MnPI) reduction of the mutant MnPs by Mn(II) were approximately 10-fold lower than for wild-type MnP. In addition, the K(D) values calculated from the first-order plots of MnP compound II (MnPII) reduction by Mn(II) for the mutant enzymes were approximately 22-fold greater than for wild-type MnP. In contrast, the first-order rate constants for MnPII reduction by Mn(II) were similar for the mutant and wild-type MnPs. Furthermore, second-order rate constants for the wild-type and mutant enzymes for MnPI formation, for MnPI reduction by bromide, and for MnPI and MnPII reduction by ferrocyanide were not significantly changed. These results indicate that both the R177A and R177K mutations specifically affect the binding of Mn, whereas the rate of electron transfer from Mn(II) to the oxidized heme apparently is not affected.  相似文献   

12.
The Mn-peroxidase from the fungus Panus tigrinus 8/18 is a hybrid enzyme. It catalyzes both Mn2+-dependent and Mn2+-independent oxidation of organic substrates. The spectral properties of intermediates and the pathway of the catalytic cycle are typical of hybrid Mn-peroxidases. The enzyme catalyzes the "oxidase" reaction (NADH oxidation) without peroxide and with the presence of Mn2+, which takes part in hydrogen peroxide production via Mn3+ and preserves the enzyme from inactivation. With the presence of organic mediators, the hybrid Mn-peroxidase oxidizes nonphenolic compounds: aromatic alcohols and a nonphenolic lignin model compound. The degree of conversion of 2,4,6-trichlorophenol is higher with the presence of l-hydroxybenzotriazole.  相似文献   

13.
Catalytic mechanisms and regulation of lignin peroxidase.   总被引:3,自引:0,他引:3  
Lignin peroxidase (LiP) is a fungal haemoprotein similar to the lignin-synthesizing plant peroxidases, but it has a higher oxidation potential and oxidizes dimethoxylated aromatic compounds to radical cations. It catalyses the degradation of lignin models but in vitro the outcome is net lignin polymerization. LiP oxidizes veratryl alcohol to radical cations which are proposed to act by charge transfer to mediate in the oxidation of lignin. Phenolic compounds are, however, preferentially oxidized, but transiently inactivate the enzyme. Analysis of the catalytic cycle of LiP shows that in the presence of veratryl alcohol the steady-state turnover intermediate is Compound II. We propose that veratryl alcohol is oxidized by the enzyme intermediate Compound I to a radical cation which now participates in charge-transfer reactions with either veratryl alcohol or another reductant, when present. Reduction of Compound II to native state may involve a radical product of veratryl alcohol or radical product of charge transfer. Phenoxy radicals, by contrast, cannot engage in charge-transfer reactions and reaction of Compound II with H2O2 ensues to form the peroxidatically inactive intermediate, Compound III. Regulation of LiP activity by phenolic compounds suggests feedback control, since many of the products of lignin degradation are phenolic. Such control would lower the concentration of phenolics relative to oxygen and favour degradative ring-opening reactions.  相似文献   

14.
The production of manganese peroxidase (MnP) by Irpex lacteus, purified to electrophoretic homogeneity by acetone precipitation, HiPrep Q and HiPrep Sephacryl S-200 chromatography, was shown to correlate with the decolorization of textile industry wastewater. The MnP was purified 11.0-fold, with an overall yield of 24.3%. The molecular mass of the native enzyme, as determined by gel filtration chromatography, was about 53 kDa. The enzyme was shown to have a molecular mass of 53.2 and 38.3 kDa on SDS-PAGE and MALDI-TOF mass spectrometry, respectively, and an isoelectric point of about 3.7. The enzyme was optimally active at pH 6.0 and between 30 and 40 degrees C. The enzyme efficiently catalyzed the decolorization of various artificial dyes and oxidized Mn (II) to Mn (III) in the presence of H(2)O(2). The absorption spectrum of the enzyme exhibited maxima at 407, 500, and 640 nm. The amino acid sequence of the three tryptic peptides was analyzed by ESI Q-TOF MS/MS spectrometry, and showed low similarity to those of the extracellular peroxidases of other white-rot basidiomycetes.  相似文献   

15.
The oxidation states of the Mn atoms in three derivatives of Mn catalase have been characterized using a combination of X-ray absorption near-edge structure (XANES) and EPR spectroscopies. The as-isolated enzyme has an average oxidation state of Mn(III) and contains a Mn(III) form, together with a reduced Mn(II) form and a variable amount (10-25%) of a Mn(III)/Mn(IV) mixed-valence derivative. Treatment with NH2OH rapidly reduces the majority of the enzyme to a Mn(II) derivative with no loss of activity. Inactivation by treatment with NH2OH + H2O2 converts all of the enzyme to a mixed-valence Mn(III)/Mn(IV) form. The inactive, mixed-valence derivative can be completely reactivated by long-term (greater than 1 h) anaerobic incubation with NH2OH, giving a reduced Mn(II)/Mn(II) derivative. These data suggest a catalytic model in which the enzyme cycles between a reduced Mn(II)/Mn(II) state and an oxidized Mn(III)/Mn(III) state.  相似文献   

16.
Unlike general peroxidases, Pleurotus ostreatus MnP2 was reported to have a unique property of direct oxidization of high-molecular-weight compounds, such as Poly R-478 and RNase A. To elucidate the mechanism for oxidation of polymeric substrates by MnP2, a series of mutant enzymes were produced by using a homologous gene expression system, and their reactivities were characterized. A mutant enzyme with an Ala substituting for an exposing Trp (W170A) drastically lost oxidation activity for veratryl alcohol (VA), Poly R-478, and RNase A, whereas the kinetic properties for Mn(2+) and H(2)O(2) were substantially unchanged. These results demonstrated that, in addition to VA, the high-molecular-weight substrates are directly oxidized by MnP2 at W170. Moreover, in the mutants Q266F and V166/168L, amino acid substitution(s) around W170 resulted in a decreased activity only for the high-molecular-weight substrates. These results, along with the three-dimensional modeling of the mutants, suggested that the mutations caused a steric hindrance to access of the polymeric substrates to W170. Another mutant, R263N, contained a newly generated N glycosylation site and showed a higher molecular mass in sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis. Interestingly, the R263N mutant exhibited an increased reactivity with VA and high-molecular-weight substrates. The existence of an additional carbohydrate modification and the catalytic properties in this mutant are discussed. This is the first study of a direct mechanism for oxidation of high-molecular-weight substrates by a fungal peroxidase using a homologous gene expression system.  相似文献   

17.
Incubation of stimulated neutrophils with sulfhydryl (RSH) compounds or ascorbic acid (ascorbate) results in rapid superoxide (O2-)-dependent oxidation of these reducing agents. Oxidation of RSH compounds to disulfides (RSSR) is faster than the rate of O2- production by the neutrophil NADPH-oxidase, whereas about one ascorbate is oxidized per O2-. Ascorbate is oxidized to dehydroascorbate, which is also oxidized but at a slower rate. Oxidation is accompanied by a large increase in oxygen (O2) uptake that is blocked by superoxide dismutase. Lactoferrin does not inhibit, indicating that ferric (Fe3+) ions are not required, and Fe3+-lactoferrin does not catalyze RSH or ascorbate oxidation. Two mechanisms contribute to oxidation: 1) O2- oxidizes ascorbate or reduced glutathione and is reduced to hydrogen peroxide (H2O2), which also oxidizes the reductants. O2- reacts directly with ascorbate, but reduced glutathione oxidation is mediated by the reaction of O2- with manganese (Mn2+). The H2O2-dependent portion of oxidation is mediated by myeloperoxidase-catalyzed oxidation of chloride to hypochlorous acid (HOCl) and oxidation of the reductants by HOCl. 2) O2- initiates Mn2+-dependent auto-oxidation reactions in which RSH compounds are oxidized and O2 is reduced. Part of this oxidation is due to the RSH-oxidase activity of myeloperoxidase. This activity is blocked by superoxide dismutase but does not require O2- production by the NADPH-oxidase, indicating that myeloperoxidase produces O2- when incubated with RSH compounds. It is proposed that an important role for O2- in the cytotoxic activities of phagocytic leukocytes is to participate in oxidation of reducing agents in phagolysosomes and the extracellular medium. Elimination of these protective agents allows H2O2 and products of peroxidase/H2O2/halide systems to exert cytotoxic effects.  相似文献   

18.
A purified and electrophoretically homogeneous blue laccase from the litter-decaying basidiomycete Stropharia rugosoannulata with a molecular mass of approximately 66 kDa oxidized Mn(2+) to Mn(3+), as assessed in the presence of the Mn chelators oxalate, malonate, and pyrophosphate. At rate-saturating concentrations (100 mM) of these chelators and at pH 5.0, Mn(3+) complexes were produced at 0.15, 0.05, and 0.10 micromol/min/mg of protein, respectively. Concomitantly, application of oxalate and malonate, but not pyrophosphate, led to H(2)O(2) formation and tetranitromethane (TNM) reduction indicative for the presence of superoxide anion radical. Employing oxalate, H(2)O(2) production, and TNM reduction significantly exceeded those found for malonate. Evidence is provided that, in the presence of oxalate or malonate, laccase reactions involve enzyme-catalyzed Mn(2+) oxidation and abiotic decomposition of these organic chelators by the resulting Mn(3+), which leads to formation of superoxide and its subsequent reduction to H(2)O(2). A partially purified manganese peroxidase (MnP) from the same organism did not produce Mn(3+) complexes in assays containing 1 mM Mn(2+) and 100 mM oxalate or malonate, but omitting an additional H(2)O(2) source. However, addition of laccase initiated MnP reactions. The results are in support of a physiological role of laccase-catalyzed Mn(2+) oxidation in providing H(2)O(2) for extracellular oxidation reactions and demonstrate a novel type of laccase-MnP cooperation relevant to biodegradation of lignin and xenobiotics.  相似文献   

19.
Lignin peroxidase (LiP) and manganese peroxidase (MnP) are structurally similar heme-containing enzymes secreted by white-rot fungi. Unlike MnP, which is only specific for Mn(2+), LiP has broad substrate specificity, but it is not known if this versatility is due to multiple substrate-binding sites. We report here that a S168W variant of MnP from Phanerochaete chrysosporium not only retained full Mn(2+) oxidase activity, but also, unlike native or recombinant MnP, oxidized a multitude of LiP substrates, including small molecule and polymeric substrates. The kinetics of oxidation of most nonpolymeric substrates by the MnP variant and LiP were similar. The stoichiometries for veratryl alcohol oxidation by these two enzymes were identical. Some readily oxidizable substrates, such as guaiacol and ferrocyanide, were oxidized by MnP S168W and LiP both specifically and nonspecifically while recombinant MnP oxidized these substrates only nonspecifically. The functional similarities between this MnP variant and LiP provide evidence for the broad substrate specificity of a single oxidation site near the surface tryptophan.  相似文献   

20.
When l-thyroxine activates the oxidation of NADH by peroxidase+H(2)O(2), little removal of phenolic-ring iodine atoms becomes apparent until most of the NADH has been oxidized, after which it increases markedly. This extensive deiodination is accompanied by loss of the ability of thyroxine to catalyse the oxidation of NADH by peroxidase+H(2)O(2). The slight deiodination observed before the appearance of extensive deiodination is somewhat higher when the effect of thyroxine on NADH oxidation is greater, and lower when thyroxine has exerted a slighter effect. ICN (but not I(2) or thyronine) catalyses NADH oxidation, in both the presence and the absence of peroxidase+H(2)O(2): thyroxine+peroxidase+H(2)O(2) are thus comparable with ICN alone in their effects on NADH oxidation. The obvious conclusion from the above observation, namely that the active moiety is the halogen liberated from thyroxine (or ICN) is, however, not directly supported by some of the results obtained by measuring the degree of deiodination of thyroxine in the system. In an attempt to reconcile some apparently contradictory conclusions, it is suggested that, when thyroxine activates oxidation of NADH by peroxidase+H(2)O(2), the diphenyl ether structure is undergoing cyclic deiodination and iodination. This would be accompanied by the maintenance in the reaction medium of an oxidized form of iodine, similar to that liberated by ICN, which would be the actual active moiety, until the NADH concentration becomes so low that the diphenyl ether structure is ruptured oxidatively. An alternative explanation is that thyroxine is oxidized to a form that either oxidizes NADH or loses iodine in competing reactions.  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