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 890 毫秒
1.
Manganese peroxidase and lignin peroxidase are ligninolytic heme-containing enzymes secreted by the white-rot fungus Phanerochaete chrysosporium. Despite structural similarity, these peroxidases oxidize different substrates. Veratryl alcohol is a typical substrate for lignin peroxidase, while manganese peroxidase oxidizes chelated Mn2+. By a single mutation, S168W, we have added veratryl alcohol oxidase activity to recombinant manganese peroxidase expressed in Escherichia coli. The kcat for veratryl alcohol oxidation was 11 s-1, Km for veratryl alcohol approximately 0.49 mM, and Km for hydrogen peroxide approximately 25 microM at pH 2.3. The Km for veratryl alcohol was higher and Km for hydrogen peroxide was lower for this manganese peroxidase mutant compared to two recombinant lignin peroxidase isoenzymes. The mutant retained full manganese peroxidase activity and the kcat was approximately 2.6 x 10(2) s-1 at pH 4.3. Consistent with relative activities with respect to these substrates, Mn2+ strongly inhibited veratryl alcohol oxidation. The single productive mutation in manganese peroxidase suggested that this surface tryptophan residue (W171) in lignin peroxidase is involved in catalysis.  相似文献   

2.
Stopped-flow techniques were used to investigate the kinetics of the formation of manganese peroxidase compound I (MnPI) and of the reactions of MnPI and manganese peroxidase compound II (MnPII) with p-cresol and MnII. All of the rate data were obtained from single turnover experiments under pseudo-first order conditions. In the presence of H2O2 the formation of MnPI is independent of pH over the range 3.12-8.29 with a second-order rate constant of (2.0 +/- 0.1) x 10(6) M-1 s-1. The activation energy for MnPI formation is 20 kJ mol-1. MnPI formation also occurs with organic peroxides such as peracetic acid, m-chloroperoxybenzoic acid, and p-nitroperoxybenzoic acid with second-order rate constants of 9.7 x 10(5), 9.5 x 10(4), and 5.9 x 10(4) M-1 s-1, respectively. The reactions of MnPI and MnPII with p-cresol strictly obeyed second-order kinetics. The second-order rate constant for the reaction of MnPII with p-cresol is extremely low, (9.5 +/- 0.5) M-1 s-1. Kinetic analysis of the reaction of MnII with MnPI and MnPII showed a binding interaction with the oxidized enzymes which led to saturation kinetics. The first-order dissociation rate constants for the reaction of MnII with MnPI and MnPII are (0.7 +/- 0.1) and (0.14 +/- 0.01) s-1, respectively, when the reaction is conducted in lactate buffer. Rate constants are considerably lower when the reactions are conducted in succinate buffer. Single turnover experiments confirmed that MnII serves as an obligatory substrate for MnPII and that both oxidized forms of the enzyme form productive complexes with MnII. Finally, these results suggest the alpha-hydroxy acids such as lactate facilitate the dissociation of MnIII from the enzyme.  相似文献   

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

4.
When neutrophils ingest bacteria, they discharge superoxide and myeloperoxidase into phagosomes. Both are essential for killing of the phagocytosed micro-organisms. It is generally accepted that superoxide is a precursor of hydrogen peroxide which myeloperoxidase uses to oxidize chloride to hypochlorous acid. Previously, we demonstrated that superoxide modulates the chlorination activity of myeloperoxidase by reacting with its ferric and compound II redox states. In this investigation we used pulse radiolysis to determine kinetic parameters of superoxide reacting with redox forms of myeloperoxidase and used these data in a steady-state kinetic analysis. We provide evidence that superoxide reacts with compound I and compound III. Our estimates of the rate constants for the reaction of superoxide with compound I, compound II, and compound III are 5 x 10(6) M-1 s-1, 5.5 +/- 0.4 x 10(6) M-1 s-1, and 1.3 +/- 0.2 x 10(5) M-1 s-1, respectively. These reactions define new activities for myeloperoxidase. It will act as a superoxide dismutase when superoxide reacts consecutively with ferric myeloperoxidase and compound III. It will also act as a superoxidase by using hydrogen peroxide to oxidize superoxide via compound I and compound II. The favorable kinetics of these reactions indicate that, within the confines of a phagosome, superoxide will react with myeloperoxidase and affect the reactions it will catalyze. These interactions of superoxide and myeloperoxidase will have a major influence on the way neutrophils use oxygen to kill bacteria. Consequently, superoxide should be viewed as a cosubstrate that myeloperoxidase uses to elicit bacterial killing.  相似文献   

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

6.
Both cyclooxygenase and peroxidase reactions of prostaglandin H synthase were studied in the presence and absence of diethyldithiocarbamate and glycerol at 4 degrees C in phosphate buffer (pH 8.0). Diethyldithiocarbamate reacts with the high oxidation state intermediates of prostaglandin H synthase; it protects the enzyme from bleaching and loss of activity by its ability to act as a reducing agent. For the reaction of diethyldithiocarbamate with compound I, the second-order rate constant k2,app, was found to fall within the range of 5.8 x 10(6) +/- 0.4 x 10(6) M-1.s-1 less than k2,app less than 1.8 x 10(7) +/- 0.1 x 10(7) M-1.s-1. The reaction of diethyldithiocarbamate with compound II showed saturation behavior suggesting enzyme-substrate complex formation, with kcat = 22 +/- 3 s-1, Km = 67 +/- 10 microM, and the second-order rate constant k3,app = 2.0 x 10(5) +/- 0.2 x 10(5) M-1.s-1. In the presence of both diethyldithiocarbamate and 30% glycerol, the parameters for compound II are kcat = 8.8 +/- 0.5 s-1, Km = 49 +/- 7 microM, and k3,app = 1.03 x 10(5) +/- 0.07 x 10(5) M-1.s-1. The spontaneous decay rate constants of compounds I and II (in the absence of diethyldithiocarbamate) are 83 +/- 5 and 0.52 +/- 0.05 s-1, respectively, in the absence of glycerol; in the presence of 30% glycerol they are 78 +/- 5 and 0.33 +/- 0.02 s-1, respectively. Neither cyclooxygenase activity nor the rate constant for compound I formation using 5-phenyl-4-pentenyl-1-hydroperoxide is altered by the presence of diethyldithiocarbamate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Catalase-peroxidases (KatGs) are unique peroxidases exhibiting a high catalase activity and a peroxidase activity with a wide range of artificial electron donors. Exchange of tyrosine 249 in Synechocystis KatG, a distal side residue found in all as yet sequenced KatGs, had dramatic consequences on the bifunctional activity and the spectral features of the redox intermediate compound II. The Y249F variant lost catalase activity but retained a peroxidase activity (substrates o-dianisidine, pyrogallol, guaiacol, tyrosine, and ascorbate) similar to the wild-type protein. In contrast to wild-type KatG and similar to monofunctional peroxidases, the formation of the redox intermediate compound I could be followed spectroscopically even by addition of equimolar hydrogen peroxide to ferric Y249F. The corresponding bimolecular rate constant was determined to be (1.1 +/- 0.1) x 107 m-1 s-1 (pH 7 and 15 degrees C), which is typical for most peroxidases. Additionally, for the first time a clear transition of compound I to an oxoferryl-like compound II with peaks at 418, 530, and 558 nm was monitored when one-electron donors were added to compound I. Rate constants of reaction of compound I and compound II with tyrosine ((5.0 +/- 0.3) x 104 m-1 s-1 and (1.7 +/- 0.4) x 102 m-1 s-1) and ascorbate ((1.3 +/- 0.2) x 104 m-1 s-1 and (8.8 +/- 0.1) x 101 m-1 s-1 at pH 7 and 15 degrees C) were determined by using the sequential stopped-flow technique. The relevance of these findings is discussed with respect to the bifunctional activity of KatGs and the recently published first crystal structure.  相似文献   

8.
NADH chemistry ancillary to the oscillatory peroxidase-oxidase (PO) reaction has been reexamined. Previously, (NAD)2 has been thought of as a terminal, inert product of the PO reaction. We now show that (NAD)2 is a central reactant in this system. Although we found traces of the dimer after several hours of the PO reaction, no accumulation of the dimer occurred, regardless of the reaction time or the number of oscillations. (NAD)2 can convert horseradish peroxidase (HRP) compound I (CpI) to compound II (CpII) with apparent rate constant (2.7 +/- 0.2) x 105 M-1.s-1 and CpII to HRP at 1 x 105 M-1.s-1. Moreover, a reduction of HRP compound III (CpIII) to CpI by (NAD)2 occurs with a rate constant faster than 5 x 106 M-1.s-1. The (NAD)2 reduction of CpIII provides an alternative to the reduction by NAD radical suggested by Yokota and Yamazaki. HRP catalyzes oxidation of alpha-NADH, not only the beta anomer as previously assumed. Rate constants of alpha- and beta-NADH reactions with CpI are (7.4 +/- 0.4) x 105 M-1.s-1, and (1.7 +/- 0.2) x 105 M-1.s-1, and with CpII are estimated as 5 x 104 M-1.s-1, and 4 x 104 M-1.s-1. Apparent rate constants of reduction of methylene blue (MB) to leuco-methylene blue (MBH) are 3.8 x 104 M-1.s-1 for NADH and 6.4 x 104 M-1.s-1 for NAD dimer, (NAD)2, while reoxidation of MBH proceeds at (2.1 +/- 0.2) x 103 M-1.s-1 All the rates were measured in 0.1 M acetate buffer, pH 5.1.  相似文献   

9.
The reduction of prostaglandin H synthase compound II to native enzyme by phenol and by hydroquinone, in the presence of diethyldithiocarbamate as a stabilizing agent, was studied by rapid scan spectrometry and transient state kinetics at 4.0 +/- 0.5 degrees C in 0.1 M phosphate buffer, pH 8.0. The plot of pseudo-first-order rate constants for the conversion of prostaglandin H synthase compound II to native enzyme versus phenol concentration was linear with a non-zero intercept. The second-order rate constant was determined from the slope to be (5.3 +/- 0.3) x 10(5) M-1 s-1. For the reduction by hydroquinone, the second-order rate constant was determined from pointwise measurements of the pseudo-first-order rate constant to be (2.1 +/- 0.4) x 10(6) M-1 s-1. Rapid scan spectrum results also showed the reduction of compound I to compound II by both phenol and hydroquinone. Thus reduction of both compound I and compound II is one electron process. Our results suggest that the tyrosyl radical, detected in the presence of oxidizing agents, is formed by intramolecular electron transfer from the tyrosyl residue to the porphyrin pi-cation radical, and this reaction tends to disappear in the presence of sufficient reducing substrate. These in vitro results support speculation that there is a role of the peroxidase component of prostaglandin H synthase in benzene-induced toxicity. In the present work, the effect of indomethacin on the reduction of prostaglandin H synthase compound II by diethyldithiocarbamate, phenol, and hydroquinone was also investigated. Results revealed, for the first time, that indomethacin is an inhibitor of the peroxidase activity of prostaglandin H synthase, although not as effectively as in its well-known inhibition of cyclooxygenase activity.  相似文献   

10.
Myeloperoxidase is the most abundant protein in neutrophils and catalyzes the production of hypochlorous acid. This potent oxidant plays a central role in microbial killing and inflammatory tissue damage. 4-Aminobenzoic acid hydrazide (ABAH) is a mechanism-based inhibitor of myeloperoxidase that is oxidized to radical intermediates that cause enzyme inactivation. We have investigated the mechanism by which benzoic acid hydrazides (BAH) are oxidized by myeloperoxidase, and we have determined the features that enable them to inactivate the enzyme. BAHs readily reduced compound I of myeloperoxidase. The rate constants for these reactions ranged from 1 to 3 x 10(6) M-1 s-1 (15 degrees C, pH 7.0) and were relatively insensitive to the substituents on the aromatic ring. Rate constants for reduction of compound II varied between 6.5 x 10(5) M-1 s-1 for ABAH and 1.3 x 10(3) M-1 s-1 for 4-nitrobenzoic acid hydrazide (15 degrees C, pH 7.0). Reduction of both compound I and compound II by BAHs adhered to the Hammett rule, and there were significant correlations with Brown-Okamoto substituent constants. This indicates that the rates of these reactions were simply determined by the ease of oxidation of the substrates and that the incipient free radical carried a positive charge. ABAH was oxidized by myeloperoxidase without added hydrogen peroxide because it underwent auto-oxidation. Although BAHs generally reacted rapidly with compound II, they should be poor peroxidase substrates because the free radicals formed during peroxidation converted myeloperoxidase to compound III. We found that the reduction of ferric myeloperoxidase by BAH radicals was strongly influenced by Hansch's hydrophobicity constants. BAHs containing more hydrophilic substituents were more effective at converting the enzyme to compound III. This implies that BAH radicals must hydrogen bond to residues in the distal heme pocket before they can reduce the ferric enzyme. Inactivation of myeloperoxidase by BAHs was related to how readily they were oxidized, but there was no correlation with their rate constants for reduction of compounds I or II. We propose that BAHs destroy the heme prosthetic groups of the enzyme by reducing a ferrous myeloperoxidase-hydrogen peroxide complex.  相似文献   

11.
We have investigated the kinetics for the peroxidase-type reaction of mangano microperoxidase 8 (Mn(III)-MP8) by the time-resolved and single-wavelength stopped-flow technique. The formation of intermediate and its subsequent reaction with substrates were studied separately. Oxidation of Mn(III)-MP8 by H2O2 at pH 10.7 yields an intermediate (1) with a rate constant of 2.9 x10(4) M-1 s-1. The formation of 1 exhibits no deuterium solvent isotope effect, favoring the homolytic cleavage of the Mn(III)-MP8 bound hydroperoxide. The rate for the formation of 1 increases sharply as the pH increases and no other intermediate was detected in the entire pH range. Addition of substrate to 1 leads to the regeneration of Mn(III)-MP8. Monitoring the conversion of 1 to Mn(III)-MP8 allows the determination of the substrate reactivity. The substrate reactivity varies by more than two orders of magnitude ranging from 1.04 x 10(6) M-1 s-1 for ascorbic acid to 4.61 x 10(3) M-1s-1 for aniline. It is linearly correlated with the reduction potential for most of the substrates studied, with the easier oxidized species showing greater reactivity. The substrate reactivity drops rapidly as the pH increases. The substrate reactivity at pH 10.7 for the Mn(III)-MP8 system is smaller than that of the corresponding Fe(III)-MP8 system by 2- to 25-fold, depending on the substrate used.  相似文献   

12.
The reaction kinetics of the peroxidase activity of prostaglandin H synthase have been examined with 15-hydroperoxyeicosatetraenoic acid and hydrogen peroxide as substrates and tetramethylphenylenediamine as cosubstrate. The apparent Km and Vmax values for both hydroperoxides were found to increase linearly with the cosubstrate concentration. The overall reaction kinetics could be interpreted in terms of an initial reaction of the synthase with hydroperoxide to form an intermediate equivalent to horseradish peroxidase Compound I, followed by reduction of this intermediate by cosubstrate to regenerate the resting enzyme. The rate constants estimated for the generation of synthase Compound I were 7.1 X 10(7) M-1 s-1 with the lipid hydroperoxide and 9.1 X 10(4) M-1 s-1 with hydrogen peroxide. The rate constants estimated for the rate-determining step in the regeneration of resting enzyme by cosubstrate were 9.2 X 10(6) M-1 s-1 in the case of the reaction with lipid hydroperoxide and 3.5 X 10(6) M-1 s-1 in the case of reaction with hydrogen peroxide. The intrinsic affinities of the synthase peroxidase for substrate (Ks) were estimated to be on the order of 10(-8) M for lipid hydroperoxide and 10(-5) M for hydrogen peroxide. These affinities are quite similar to the reported affinities of the synthase for these hydroperoxides as activators of the cyclooxygenase. The peroxidase activity was found to be progressively inactivated during the peroxidase reaction. The rate of inactivation of the peroxidase was increased by increases in hydroperoxide level, and decreased by increases in peroxidase cosubstrate. The inactivation of the peroxidase appeared to occur by a hydroperoxide-dependent process, originating from synthase Compound I or Compound II.  相似文献   

13.
Stopped-flow techniques were utilized to investigate the kinetics of the reaction of lignin peroxidase compounds I and II (LiPI and LiPII) with veratryl alcohol (VA). All rate data were collected from single turnover experiments under pseudo first-order conditions. The reaction of LiPI with VA strictly obeys second-order kinetics over the pH range 2.72-5.25 as demonstrated by linear plots of the pseudo first-order rate constants versus concentrations of VA. The second-order rate constants are strongly dependent on pH and range from 2.62 x 10(6) M-1 s-1 (pH 2.72) to 1.45 x 10(4) M-1 s-1 (pH 5.25). The reaction of LiPII and VA exhibits saturation behavior when the observed pseudo first-order rate constants are plotted against VA concentrations. The saturation phenomenon is quantitatively explained by the formation of a 1:1 LiPII-substrate complex. Results of kinetic and rapid scan spectral analyses exclude the formation of a LiPII-VA cation radical complex. The first-order dissociation rate constant and the equilibrium dissociation constant for the LiPII reaction are also pH dependent. Binding of VA to LiPII is controlled by a heme-linked ionizable group of pKa approximately 4.2. The pH profiles of the second-order rate constants for the LiPI reaction and of the first-order dissociation constants for the LiPII reaction both demonstrate two pKa values at approximately 3.0 and approximately 4.2. Protonated oxidized enzyme intermediates are most active, suggesting that only electron transfer, not proton uptake from the reducing substrate, occurs at the enzyme active site. These results are consistent with the one-electron oxidation of VA to an aryl cation radical by LiPI and LiPII.  相似文献   

14.
Direct reactions of peroxidases with Trolox C (a vitamin E analogue) and vitamin E were observed in 50% (v/v) methanol. The kinetic results revealed that the reaction of horseradish peroxidase intermediate Compound II with Trolox C and vitamin E was the rate-determining step, and the rate constants were estimated to be 1.7 x 10(3) and 5.1 x 10(2) M-1.s-1, respectively. Peroxidases catalyzed the one-electron oxidation of Trolox C and vitamin E, and the vitamin E phenoxyl radicals resulting from the peroxidase reactions were detected by continuous-flow ESR spectroscopy.  相似文献   

15.
Stopped-flow rapid scan techniques were used to obtain a spectrum of nearly homogeneous lignin peroxidase compound I (LiPI) under pseudo-first order conditions at the unusually low pH optimum (3.0) for the enzyme. The LiPI spectrum had a Soret band at 407 nm with approximately 60% reduced intensity and a visible maximum at 650 nm. Under steady-state conditions a Soret spectrum for lignin peroxidase compound II (LiPII) was also obtained. The Soret maximum of LiPII at 420 nm was only approximately 15% reduced in intensity compared to native LiP. Transient state kinetic results confirmed the pH independence of LiPI formation over the pH range 3.06-7.39. The rate constant was (6.5 +/- 0.2) x 10(5) M-1 S-1. Addition of excess veratryl alcohol to LiPI resulted in its reduction to LiPII with subsequent reduction of LiPII to the native enzyme. Reactions of LiPI and LiPII with veratryl alcohol exhibited marked pH dependencies. For the LiPI reaction the rate constants ranged from 2.5 x 10(6) M-1 S-1 at pH 3.06 to 4.1 x 10(3) M-1 S-1 at pH 7.39; for the LiPII reaction, 1.6 x 10(5) M-1 S-1 (pH 3.06) to 2.3 x 10(3) M-1 S-1 (pH 5.16). These single turnover experiments demonstrate directly that the pH dependence of these reactions dictates the overall pH dependence of this novel enzyme. These results are consistent with the one-electron oxidation of veratryl alcohol to an aryl cation radical by LiPI and by LiPII.  相似文献   

16.
Elementary reactions have been studied quantitatively in the complex overall process catalyzed by horseradish peroxidase whereby isobutyraldehyde and molecular oxygen react to form triplet state acetone and formic acid. The rate constant for the reaction of the enol form of isobutyraldehyde with compound I of peroxidase is (8 +/- 1) X 10(6) M-1 s-1 and with compound II (1.3 +/- 0.3) X 10(6) M-1 s-1. Neither the enolate anion nor the keto form is reactive. The reactivity of enols with peroxidase parallels that of unionized phenols and a common mechanism is proposed. The overall catalyzed reaction of isobutyraldehyde and oxygen consists of an initial burst followed by a steady state phase. The burst is caused by the following sequence: 1) an initial high yield of compound I is formed from reaction of native enzyme with the autoxidation product of isobutyraldehyde, a peracid and 2) compound I rapidly depletes the equilibrium pool of enol which is present. After this burst a steady state phase is observed in which the rate-limiting step is the conversion of the keto to the enol form of the aldehyde catalyzed by phosphate buffer. The rate constant for the keto form reacting with phosphate is (8.7 +/- 0.6) X 10(-5) M-1 s-1. All constants were measured in dilute aqueous ethanol at 35 degrees C, pH 7.4, and ionic strength 0.67 M. Both the initial burst of light and the steady state emission from triplet acetone can be observed with the naked eye. Since the magnitude of the burst is a measure of the equilibrium amount of enol, the keto-enol equilibrium constant is readily calculated and hence also the rate constant for conversion of enol to keto. The keto-enol equilibrium constant is unaffected by phosphate which therefore acts as a true catalyst.  相似文献   

17.
Two major peroxidases are secreted by the fungus Pleurotus eryngii in lignocellulose cultures. One is similar to Phanerochaete chrysosporium manganese-dependent peroxidase. The second protein (PS1), although catalyzing the oxidation of Mn2+ to Mn3+ by H2O2, differs from the above enzymes by its manganese-independent activity enabling it to oxidize substituted phenols and synthetic dyes, as well as the lignin peroxidase (LiP) substrate veratryl alcohol. This is by a mechanism similar to that reported for LiP, as evidenced by p-dimethoxybenzene oxidation yielding benzoquinone. The apparent kinetic constants showed high activity on Mn2+, but methoxyhydroquinone was the natural substrate with the highest enzyme affinity (this and other phenolic substrates are not efficiently oxidized by the P. chrysosporium peroxidases). A three-dimensional model was built using crystal models from four fungal peroxidase as templates. The model suggests high structural affinity of this versatile peroxidase with LiP but shows a putative Mn2+ binding site near the internal heme propionate, involving Glu36, Glu40, and Asp181. A specific substrate interaction site for Mn2+ is supported by kinetic data showing noncompetitive inhibition with other peroxidase substrates. Moreover, residues reported as involved in LiP interaction with veratryl alcohol and other aromatic substrates are present in peroxidase PS1 such as His82 at the heme-channel opening, which is remarkably similar to that of P. chrysosporium LiP, and Trp170 at the protein surface. These residues could be involved in two different hypothetical long range electron transfer pathways from substrate (His82-Ala83-Asn84-His47-heme and Trp170-Leu171-heme) similar to those postulated for LiP.  相似文献   

18.
Compared to other monomeric heme proteins and the heme peroxidases, the Glycera dibranchiata monomer hemoglobin components III and IV exhibit very slow cyanide binding kinetics. This is agreement with the previously reported behavior of component II. Similar to component II, components III and IV have been studied under pseudo-first-order conditions at pH 6.0, 7.0, 8.0, and 9.0 by using a 100-250-fold excess of potassium cyanide at each pH. At 20 degrees C with micromolar protein concentrations, kobs for component III varies between 7.08 x 10(-5) s-1 at pH 6.0 and 100-fold cyanide excess and 1.06 x 10(-2) s-1 at pH 9.0 and 250-fold cyanide excess. For component IV, the values are 2.03 x 10(-4) s-1 for 100-fold cyanide excess at pH 6.0 and 4.13 x 10(-2) s-1 for 250-fold cyanide excess at pH 9.0. In comparison to other heme proteins, our analysis shows that the bimolecular rate constant (klapp) is small. For example, at pH 7.0, it is 3.02 x 10(-1) M-1 s-1 for component III and 1.82 M-1 s-1 for component IV, compared to 400 M-1 s-1 for sperm whale metmyoglobin, 692 M-1 s-1 for soybean metleghemoglobin a, 111 M-1 s-1 for guinea pig methemoglobin, and 1.1 x 10(5) M-1 s-1 for cytochrome c peroxidase. Our results also show that the dissociation rates (k-lapp) are extremely slow and no larger than 10(-6) s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Rate constants for the reaction between horseradish peroxidase compound I and p-cresol have been determined at several values of pH between 2.98 and 10.81. These rate constants were used to construct a log (rate) versus pH profile from which it is readily seen that the most reactive form of the enzyme is its most basic form within this pH range so that base catalysis is occurring. At the maximum rate a second order rate constant of (5.1 +/- 0.3) x 10(-7) M-1 s-1 at 25 degrees is obtained. The activation energy of the reaction at the maximum rate was determined from an Arrhenius plot to be 5.0 +/- 0.5 kcal/mol. Evidence for an exception to the generally accepted enzymatic cycle of horseradish peroxidase is presented. One-half molar equivalent of p-cresol can convert compound I quantitatively to compound II at high pH, whereas usually this step requires 1 molar equivalent of reductant. The stoichiometry of this reaction is pH-dependent.  相似文献   

20.
The cDNA clone lambda ML-1 encoding one of the extracellular lignin peroxidases from the white rot fungus, Phanerochaete chrysosporium, was heterologously expressed in an active form using a recombinant baculovirus system. The glycosylated extracellular form of the recombinant protein contained the ferriprotoporphyrin IX moiety and was capable of oxidizing both iodide and the model lignin compound, veratryl alcohol. In comparative peroxidase assays using guaiacol and Mn(II), the recombinant lignin peroxidase did not appear to be Mn(II) dependent. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis demonstrated that the heterologously expressed peroxidase had an apparent molecular weight similar to that of the native fungal isozyme H8. The elution profile of the active recombinant enzyme derived by ion-exchange chromatography and immunoblot analysis using an anti-H8 monoclonal antibody provided further evidence that the lambda ML-1 DNA encodes the lignin peroxidase H8.  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