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 93 毫秒
1.
Samuels NM  Klinman JP 《Biochemistry》2005,44(43):14308-14317
Copper amine oxidase (CAO) is a dual-functioning enzyme that catalyzes the biosynthesis of a self-derived coenzyme and subsequent oxidative deamination of primary amines. The organic cofactor, 2,4,5-trihydroxyphenylalanine quinone (TPQ), is generated from the post-translational modification of an active site tyrosine (Y405) in a reaction shown to be dependent on both molecular oxygen and a mononuclear copper center. Previous investigations of Cu(II)-dependent cofactor formation in the Hansenula polymorpha amine oxidase (HPAO) provided evidence for the coordination of the precursor tyrosine in forming a ligand-to-metal charge transfer complex as a means of activating the tyrosyl ring for direct attack by triplet-state dioxygen. To further delineate the role of the metal in facilitating this complex series of reactions, apo-HPAO was reconstituted with alternate metals of varying reduction potentials and Lewis acidities [Ni(II), Co(II), Mn(II), Fe(II), and Fe(III)] and the consequence of each substitution on TPQ biogenesis examined. Ni(II) was found to support the transformation of the precursor tyrosine to the quinone cofactor to yield a mature enzyme competent for methylamine oxidation. Detailed kinetic analysis of the mechanism of TPQ biogenesis for the Ni(II)-substituted enzyme has led to the proposal of a direct electron transfer from the metal-coordinated tyrosinate to dioxygen as the dominant rate-limiting step.  相似文献   

2.
DuBois JL  Klinman JP 《Biochemistry》2005,44(34):11381-11388
The copper amine oxidases (CAOs) catalyze the O(2)-dependent, two-electron oxidation of amines to aldehydes at an active site that contains Cu(II) and topaquinone (TPQ) cofactor. TPQ arises from the autocatalytic, post-translational oxidation of a tyrosine side chain in the active site. Monooxygenation within the ring of tyrosine at a single Cu(II) site is unique in biology and occurs as an early step in the formation of TPQ. The mechanism of this reaction has been further examined in the CAO from Hansenula polymorpha (HPAO). When a Clark electrode fitted to a custom-made, gastight apparatus over a range of initial concentrations of O(2) was used, rates of O(2) consumption at levels greater than air are seen to be reduced relative to earlier results, yielding K(D)(apparent) = 216 microM for O(2). This is consistent with a mechanism in which O(2) binds reversibly to the active site, triggering a conformational change that promotes ligation of tyrosinate to Cu(II). The activated Cu(II)-tyrosinate species has been proposed to react with O(2) in a rate-limiting step, although it was also possible that breakdown of a putative peroxy-intermediate controlled TPQ formation. To test the latter hypothesis, Cu(II)-free HPAO was prepared with 3,5-ring-[(2)H(2)]-tyrosine incorporated throughout the primary sequence. The absence of an isotope effect on the rate of TPQ formation eliminates cleavage of this C-H bond in a proposed Cu(II)-aryl-peroxide intermediate as a rate limiting step. The role of methionine 634, previously found to moderate O(2) binding during the catalytic cycle, is shown here to serve a similar function in TPQ formation. As with catalysis, the rate of TPQ formation correlates with the volume of the hydrophobic side chain at position 634, implicating similar binding sites for O(2) during catalysis and cofactor biogenesis.  相似文献   

3.
Copper, a mediator of redox chemistries in biology, is often found in enzymes that bind and reduce dioxygen. Among these, the copper amine oxidases catalyze the oxidative deamination of primary amines utilizing a type(II) copper center and 2,4,5-trihydroxyphenylalanine quinone (TPQ), a covalent cofactor derived from the post-translational modification of an active site tyrosine. Previous studies established the dependence of TPQ biogenesis on Cu(II); however, the dependence of cofactor formation on the biologically relevant Cu(I) ion has remained untested. In this study, we demonstrate that the apoform of the Hansenula polymorpha amine oxidase readily binds Cu(I) under anaerobic conditions and produces the quinone cofactor at a rate of 0.28 h(-1) upon subsequent aeration to yield a mature enzyme with kinetic properties identical to the protein product of the Cu(II)-dependent reaction. Because of the change in magnetic properties associated with the oxidation of copper, electron paramagnetic resonance spectroscopy was employed to investigate the nature of the rate-limiting step of Cu(I)-dependent cofactor biogenesis. Upon aeration of the unprocessed enzyme prebound with Cu(I), an axial Cu(II) electron paramagnetic resonance signal was found to appear at a rate equivalent to that for the cofactor. These data provide strong evidence for a rate-limiting release of superoxide from a Cu(II)(O(2)(.)) complex as a prerequisite for the activation of the precursor tyrosine and its transformation for TPQ. As copper is trafficked to intracellular protein targets in the reduced, Cu(I) state, these studies offer possible clues as to the physiological significance of the acquisition of Cu(I) by nascent H. polymorpha amine oxidase.  相似文献   

4.
Goto Y  Klinman JP 《Biochemistry》2002,41(46):13637-13643
Copper amine oxidases (CAOs) contain 2,4,5-trihydroxyphenylalanyl quinone (TPQ) and a copper ion in their active sites, catalyzing amine oxidation to aldehyde and ammonia concomitant with the reduction of molecular oxygen to hydrogen peroxide. Kinetic studies on the CAO from bovine serum (BSAO) [Su and Klinman (1999) Biochemistry 37, 12513-12525] and the recent reports on the cobalt substituted form of the enzyme from Hansenula polymorpha (HPAO) [Mills and Klinman (2000) J. Am. Chem. Soc. 122, 9897-9904, and Mills et al. (2002) Biochemistry, 41, 10577-10584] support pre-binding of molecular oxygen prior to a rate-limiting electron transfer from the reduced form of TPQ (p-aminohydroquinone form) to dioxygen. Although there is significant sequence homology between BSAO and HPAO, k(cat)/K(m)(O2) for BSAO under the optimal condition is one order of magnitude lower than that for HPAO. From a comparison of amino acid sequences for BSAO and HPAO, together with the X-ray crystal structure of HPAO, a plausible dioxygen pre-binding site has been identified that involves Y407, L425, and M634 in HPAO; the latter two residues are altered in BSAO to A490 and T695. To determine which of these residues plays a greater role in dioxygen chemistry, k(cat)/K(m)(O2) was determined in HPAO for the M634 --> T and L425 --> A mutants. The L425 --> A mutation does not alter k(cat)/K(m)(O2) to a large extent, whereas the M634 --> T decreased k(cat)/K(m)(O2) by one order of a magnitude, creating a catalyst that is similar to BSAO. A series of mutants at M634 (to F, L, and Q) were, therefore, prepared in HPAO and characterized with regard to k(cat)/K(m)(O2) as a function of pH. Structure reactivity correlations show a linear relationship of rate with side chain volume, rather than hydrophobicity, indicating that dioxygen reactivity increases with the bulk of the residue at position 634. This site also shows specificity for O2, in relation to the co-gas N2, since substitution of the inert gas N2 by either Ar or He has no effect on measured rates. In particular, He gas is expected to have little affinity for protein at 1 atmospheric pressure, implying little or no binding by N2 as well.  相似文献   

5.
Q Su  J P Klinman 《Biochemistry》1999,38(26):8572-8581
Glucose oxidase catalyzes the oxidation of glucose by molecular dioxygen, forming gluconolactone and hydrogen peroxide. A series of probes have been applied to investigate the activation of dioxygen in the oxidative half-reaction, including pH dependence, viscosity effects, 18O isotope effects, and solvent isotope effects on the kinetic parameter Vmax/Km(O2). The pH profile of Vmax/Km(O2) exhibits a pKa of 7.9 +/- 0.1, with the protonated enzyme form more reactive by 2 orders of magnitude. The effect of viscosogen on Vmax/Km(O2) reveals the surprising fact that the faster reaction at low pH (1.6 x 10(6) M-1 s-1) is actually less diffusion-controlled than the slow reaction at high pH (1.4 x 10(4) M-1 s-1); dioxygen reduction is almost fully diffusion-controlled at pH 9.8, while the extent of diffusion control decreases to 88% at pH 9.0 and 32% at pH 5.0, suggesting a transition of the first irreversible step from dioxygen binding at high pH to a later step at low pH. The puzzle is resolved by 18O isotope effects. 18(Vmax/Km) has been determined to be 1.028 +/- 0.002 at pH 5.0 and 1.027 +/- 0.001 at pH 9.0, indicating that a significant O-O bond order decrease accompanies the steps from dioxygen binding up to the first irreversible step at either pH. The results at high pH lead to an unequivocal mechanism; the rate-limiting step in Vmax/Km(O2) for the deprotonated enzyme is the first electron transfer from the reduced flavin to dioxygen, and this step accompanies binding of molecular dioxygen to the active site. In combination with the published structural data, a model is presented in which a protonated active site histidine at low pH accelerates the second-order rate constant for one electron transfer to dioxygen through electrostatic stabilization of the superoxide anion intermediate. Consistent with the proposed mechanisms for both high and low pH, solvent isotope effects indicate that proton transfer steps occur after the rate-limiting step(s). Kinetic simulations show that the model that is presented, although apparently in conflict with previous models for glucose oxidase, is in good agreement with previously published kinetic data for glucose oxidase. A role for electrostatic stabilization of the superoxide anion intermediate, as a general catalytic strategy in dioxygen-utilizing enzymes, is discussed.  相似文献   

6.
The catalytic reaction of copper/topa quinone (TPQ) containing amine oxidase consists of the initial, well-characterized, reductive half-reaction and the following, less studied, oxidative half-reaction. We have analyzed the oxidative half-reaction catalyzed by phenylethylamine oxidase from Arthrobacter globiformis (AGAO) by rapid-scan stopped-flow measurements. Upon addition of dioxygen to the substrate-reduced AGAO at pH 8.2, the absorption bands derived from the semiquinone (TPQ(sq)) and aminoresorcinol forms of the TPQ cofactor disappeared within the dead time (<1 ms) of the measurements, indicating that the reaction of the substrate-reduced enzyme with dioxygen is very rapid. Concomitantly, an early intermediate exhibiting an absorption band at about 410 nm was formed, which then decayed with a rate constant of 390 +/- 50 s(-1). This intermediate was detected more prominently in the reaction in D2O buffer (pD 8.1) and was assigned to a Cu(II)-peroxy species. The assignment was based on the observation that addition of H2O2 to the substrate-reduced AGAO under anaerobic conditions led to the formation of a new band at about 415 nm, accompanied by partial quenching of absorption bands derived from TPQ(sq). Other intermediates exhibiting absorption bands at about 310 and 340 nm were also observed in the oxidative half-reaction. Kinetics of the disappearance of these latter bands did not correspond with that of the Cu(II)-peroxy band at 410 nm but did well with that of the increase of the 480 nm absorption band due to the reoxidized TPQ. Rapid increase of the absorption in the 320-370 nm region was also observed for the reaction of the substrate-reduced, Ni-substituted enzyme with dioxygen. On the basis of these results, a possible mechanism is proposed for the oxidative half-reaction of the bacterial copper amine oxidase.  相似文献   

7.
Schwartz B  Dove JE  Klinman JP 《Biochemistry》2000,39(13):3699-3707
A detailed kinetic analysis of oxygen consumption during TPQ biogenesis has been carried out on a yeast copper amine oxidase. O(2) is consumed in a single, exponential phase, the rate of which responds linearly to dissolved oxygen concentration. This behavior is observed up to conditions of maximally obtainable oxygen concentrations. In contrast, no viscosity effect is observed on rate, implicating a high K(m) for O(2). Binding of oxygen appears to occur faster than its consumption and to result in displacement of the precursor tyrosine onto copper to form a charge-transfer species, described in the the preceding paper of this issue [Dove, J. E., Schwartz, B., Williams, N. K., and Klinman, J. P. (2000) Biochemistry 39, 3690-3698). Reaction between this intermediate and O(2) is proposed to occur in a rate-limiting step, and to proceed more rapidly when the tyrosine is deprotonated. This rate-limiting step in cofactor biogenesis does not display a solvent isotope effect and is, thus, uncoupled from proton transfer. Comparisons are drawn between the proposed biogenesis mechanism and that for the oxidation of reduced cofactor during catalytic turnover in the mature enzyme.  相似文献   

8.
Two radicals have been detected previously by electron paramagnetic resonance (EPR) and electron nuclear double resonance (ENDOR) spectroscopies in bovine cytochrome oxidase after reaction with hydrogen peroxide, but no correlation could be made with predicted levels of optically detectable intermediates (P(M), F and F(z.rad;)) that are formed. This work has been extended by optical quantitation of intermediates in the EPR/ENDOR sample tubes, and by comparison with an analysis of intermediates formed by reaction with carbon monoxide in the presence of oxygen. The narrow radical, attributed previously to a porphyrin cation, is detectable at low levels even in untreated oxidase and increases with hydrogen peroxide treatments generally. It is presumed to arise from a side-reaction unrelated to the catalytic intermediates. The broad radical, attributed previously to a tryptophan radical, is observed only in samples with a significant level of F(z.rad;) but when F(z.rad;) is generated with hydrogen peroxide, is always accompanied by the narrow radical. When P(M) is produced at high pH with CO/O(2), no EPR-detectable radicals are formed. Conversion of the CO/O(2)-generated P(M) into F(z.rad;) when pH is lowered is accompanied by the appearance of a broad radical whose ENDOR spectrum corresponds to a tryptophan cation. Quantitation of its EPR intensity indicates that it is around 3% of the level of F(z.rad;) determined optically. It is concluded that low pH causes a change of protonation pattern in P(M) which induces partial electron redistribution and tryptophan cation radical formation in F(z.rad;). These protonation changes may mimic a key step of the proton translocation process.  相似文献   

9.
In the reaction cycle of cytochrome c oxidase from Rhodobacter sphaeroides, one of the steps that are coupled to proton pumping, the oxo-ferryl-to-oxidized transition (F --> O), displays a large kinetic deuterium isotope effect of about 7. In this study we have investigated in detail the dependence of the kinetics of this reaction step ?k(FO)(chi) on the fraction (chi) D(2)O in the enzyme solution (proton-inventory technique). According to a simplified version of the Gross-Butler equation, from the shape of the graph describing k(FO)(chi)/k(FO)(0), conclusions can be drawn concerning the number of protonatable sites involved in the rate-limiting proton-transfer reaction step. Even though the proton-transfer reaction during the F --> O transition takes place over a distance of at least 30 A and involves a large number of protonatable sites, the proton-inventory analysis displayed a linear dependence, which indicates that the entire deuterium isotope effect of 7 is associated with a single protonatable site. On the basis of experiments with site-directed mutants of cytochrome c oxidase, this localized proton-transfer rate control is proposed to be associated with glutamate (I-286) in the D-pathway. Consequently, the results indicate that proton transfer from the glutamate controls the rate of all events during the F --> O reaction step. The proton-inventory analysis of the overall enzyme turnover reveals a nonlinear plot characteristic of at least two protonatable sites involved in the rate-limiting step in the transition state, which indicates that this step does not involve proton transfer through the same pathway (or through the same mechanism) as during the F --> O transition.  相似文献   

10.
Welford RW  Lam A  Mirica LM  Klinman JP 《Biochemistry》2007,46(38):10817-10827
The mechanism of the first electron transfer from reduced cofactor to O2 in the catalytic cycle of copper amine oxidases (CAOs) remains controversial. Two possibilities have been proposed. In the first mechanism, the reduced aminoquinol form of the TPQ cofactor transfers an electron to the copper, giving radical semiquinone and Cu(I), the latter of which reduces O2 (pathway 1). The second mechanism invokes direct transfer of the first electron from the reduced aminoquinol form of the TPQ cofactor to O2 (pathway 2). The debate over these mechanisms has arisen, in part, due to variable experimental observations with copper amine oxidases from plant versus other eukaryotic sources. One important difference is the position of the aminoquinol/Cu(II) to semiquinone/Cu(I) equilibrium on anaerobic reduction with amine substrate, which varies from almost 0% to 40% semiquinone/Cu(I). In this study we have shown how protein structure controls this equilibrium by making a single-point mutation at a second-sphere ligand to the copper, D630N in Hansenula polymorpha amine oxidase, which greatly increases the concentration of the cofactor semiquinone/Cu(I) following anaerobic reduction by substrate. The catalytic properties of this mutant, including 18O kinetic isotope effects, point to a conservation of pathway 2, despite the elevated production of the cofactor semiqunone/Cu(I). Changes in kcat/Km[O2] are attributed to an impact of D630N on an increased affinity of O2 for its hydrophobic pocket. The data in this study indicate that changes in cofactor semiquinone/Cu(I) levels are not sufficient to alter the mechanism of O2 reduction and illuminate how subtle features are able to control the reduction potential of active site metals in proteins.  相似文献   

11.
Glyoxal oxidase of P. chrysosporium is a radical copper oxidase that catalyzes oxidation of aldehydes to carboxylic acids coupled to dioxygen reduction to H(2)O(2). In addition to known substrates, glycerol is also found to be a substrate for glyoxal oxidase. During enzyme turnover, glyoxal oxidase undergoes a reversible inactivation, probably caused by loss of the active site free radical, resulting in short-lasting enzyme activities and undetectable substrate conversions. Enzyme activity could be extended by including two additional enzymes, horseradish peroxidase and catalase, in addition to a redox chemical activator, such as Mn(III) (or Mn(II)+H(2)O(2)) or hexachloroiridate. Using this three-enzyme system glycerol was converted in glyceric acid in a two-step reaction, with glyceraldehyde as intermediate. A possible operation mechanism is proposed in which the three enzymes would work coordinately allowing to maintain a sustained glyoxal oxidase activity. In the course of its catalytic cycle, glyoxal oxidase alternates between two functional and interconvertible reduced and oxidized forms resulting from a two-electron transfer process. However, glyoxal oxidase can also undergo an one-electron reduction to a catalytically inactive form lacking the active site free radical. Horseradish peroxidase could use glyoxal oxidase-generated H(2)O(2) to oxidize Mn(II) to Mn(III) which, in turn, would reoxidize and reactivate the inactive form of glyoxal oxidase. Catalase would remove the excess of H(2)O(2) generated during the reaction. In spite of the improvement achieved using the three-enzyme system, glyoxal oxidase inactivation still occurred, which resulted in low substrate conversions. Possible causes of inactivation, including end-product inhibition, are discussed.  相似文献   

12.
Copper amine oxidases are homodimeric enzymes that catalyze two reactions: first, a self-processing reaction to generate the 2,4,5-trihydroxyphenylalanine (TPQ) cofactor from an active site tyrosine by a single turnover mechanism; second, the oxidative deamination of primary amine substrates with the production of aldehyde, hydrogen peroxide, and ammonia catalyzed by the mature enzyme. The importance of active site residues in both of these processes has been investigated by structural studies and site-directed mutagenesis in enzymes from various organisms. One conserved residue is a tyrosine, Tyr369 in the Escherichia coli enzyme, whose hydroxyl is hydrogen bonded to the O4 of TPQ. To explore the importance of this site, we have studied a mutant enzyme in which Tyr369 has been mutated to a phenylalanine. We have determined the X-ray crystal structure of this variant enzyme to 2.1 A resolution, which reveals that TPQ adopts a predominant nonproductive conformation in the resting enzyme. Reaction of the enzyme with the irreversible inhibitor 2-hydrazinopyridine (2-HP) reveals differences in the reactivity of Y369F compared with wild type with more efficient formation of an adduct (lambda(max) = 525 nm) perhaps reflecting increased mobility of the TPQ adduct within the active site of Y369F. Titration with 2-HP also reveals that both wild type and Y369F contain one TPQ per monomer, indicating that Tyr369 is not essential for TPQ formation, although we have not measured the rate of TPQ biogenesis. The UV-vis spectrum of the Y369F protein shows a broader peak and red-shifted lambda(max) at 496 nm compared with wild type (480 nm), consistent with an altered electronic structure of TPQ. Steady-state kinetic measurements reveal that Y369F has decreased catalytic activity particularly below pH 6.5 while the K(M) for substrate beta-phenethylamine increases significantly, apparently due to an elevated pK(a) (5.75-6.5) for the catalytic base, Asp383, that should be deprotonated for efficient binding of protonated substrate. At pH 7.0, the K(M) for wild type and Y369F are similar at 1.2 and 1.5 microM, respectively, while k(cat) is decreased from 15 s(-1) in wild type to 0.38 s(-1), resulting in a 50-fold decrease in k(cat)/K(M) for Y369F. Transient kinetics experiments indicate that while the initial stages of enzyme reduction are slower in the variant, these do not represent the rate-limiting step. Previous structural and solution studies have implicated Tyr369 as a component of a proton shuttle from TPQ to dioxygen. The moderate changes in kinetic parameters observed for the Y369F variant indicate that if this is the case, then the absence of the Tyr369 hydroxyl can be compensated for efficiently within the active site.  相似文献   

13.
Van Eps N  Szundi I  Einarsdóttir O 《Biochemistry》2000,39(47):14576-14582
We describe a new method for studying rapid biological reactions involving dioxygen. This approach is based on the photolysis of a synthetic caged dioxygen carrier, which produces dioxygen on a fast time scale. The method was used to investigate the reduction of dioxygen to water by cytochrome c oxidase at room temperature following photolysis of a (mu-peroxo)(mu-hydroxo)bis[bis(bipyridyl)c obalt(III)] complex. The fact that dioxygen is generated in situ on a nanosecond or faster time scale avoids potential complications related to the fate of photodissociated CO in a conventional CO flow-flash experiment. The cobalt complex is stable at room temperature under anaerobic conditions and releases dioxygen upon irradiation at 355 nm with a quantum yield of 0.04. The complex does not react with reduced cytochrome oxidase or its reducing agents within the mixing time of the experiment, and its photoproducts do not interfere with the kinetics of the dioxygen reduction. The oxidation of the reduced cytochrome oxidase was monitored between 500 and 750 nm using a gated optical spectrometric multichannel analyzer following photodissociation of the cobalt complex. The data were analyzed using singular value decomposition and global exponential fitting, and two apparent lifetimes (380 +/- 50 micros and 1.7 +/- 0.2 ms) were resolved and compared to results from a conventional CO flow-flash experiment. The results show that approximately 90 microM dioxygen can be generated upon a single laser pulse and that this approach can be used to study other fast biological reactions involving O(2).  相似文献   

14.
X Sun  X Shi  N S Dalal 《FEBS letters》1992,303(2-3):213-216
In the presence of hydrogen peroxide (H2O2), xanthine oxidase has been found to catalyze sulfur trioxide anion radical (SO3.-) formation from sulfite anion (SO3(2-)). The SO3.- radical was identified by ESR (electron spin resonance) spin trapping, utilizing 5,5-dimethyl-l-pyrroline-l-oxide (DMPO) as the spin trap. Inactivated xanthine oxidase does not catalyze SO3.- radical formation, implying a specific role for this enzyme. The initial rate of SO3.- radical formation increases linearly with xanthine oxidase concentration. Together, these observations indicate that the SO3.- generation occurs enzymatically. These results suggest a new property of xanthine oxidase and perhaps also a significant step in the mechanism of sulfite toxicity in cellular systems.  相似文献   

15.
 The structure of a new biological redox cofactor – topaquinone (TPQ), the quinone of 2,4,5-trihydroxyphenylalanine – was elucidated in 1990. TPQ is the cofactor in most copper-containing amine oxidases. It is produced by post-translational modification of a strictly conserved active-site tyrosine residue. Recent work has established that TPQ biogenesis proceeds via a novel self-processing pathway requiring only the protein, copper, and molecular oxygen. The oxidation of tyrosine to TPQ by dioxygen is a six-electron process, which has intriguing mechanistic implications because copper is a one-electron redox agent, and dioxygen can function as either a two-electron or four-electron oxidant. This review adopts an historical perspective in discussing the structure and reactivity of TPQ in amine oxidases, and then assesses what is currently understood about the mechanism of the oxidation of tyrosine to produce TPQ. Aspects of the structures and chemistry of related cofactors, such as the Tyr-Cys radical in galactose oxidase and the lysine tyrosylquinone of lysyl oxidase, are also discussed. Received: 23 May 1998 / Accepted: 19 October 1998  相似文献   

16.
Formation of oxygen radicals during reduction of H(2)O(2) or diperoxovanadate with vanadyl sulfate or ferrous sulfate was indicated by the 1:2:2:1 electron spin resonance (ESR) signals of the DMPO adduct typical of standard ()OH radical. Signals derived from diperoxovanadate remained unchanged in the presence of ethanol in contrast to those from H(2)O(2). This gave the clue that they represent a different radical, possibly (*)OV(O(2))(2+), formed on breaking a peroxo-bridge of diperoxovanadate complex. The above reaction mixtures evolved dioxygen or, when NADH was present, oxidized it rapidly which was accompanied by consumption of dioxygen. Operation of a cycle of peroxovanadates including this new radical is suggested to explain these redox activities both with vanadyl and ferrous sulfates. It can be triggered by ferrous ions released from cellular stores in the presence of catalytic amounts of peroxovanadates.  相似文献   

17.
Selenocystamine (RSe-SeR) was shown to catalyze the oxygen-mediated oxidation of excess GSH to glutathione disulfide, at neutral pH and ambient PO2. This glutathione oxidase activity required the heterolytic reduction of the diselenide bond, which produced two equivalents of the selenolate derivative selenocysteamine (RSe-), via the transient formation of a selenenylsulfide intermediate (RSe-SG). Formation of RSe- was the only reaction observed in anaerobic conditions. At ambient PO2, the kinetics and stoichiometry of GSSG production as well as that of GSH and oxygen consumptions demonstrated that RSe- performed a three-step reduction of oxygen to water. The first step was a one-electron transfer from RSe- to dioxygen, yielding superoxide and a putative selenyl radical RSe., which decayed very rapidly to RSe-SeR. In the second step, RSe- reduced superoxide to hydrogen peroxide through a much faster one-electron transfer, also associated with the decay of RSe. to RSe-SeR. The third step was a two-electron transfer from RSe- to hydrogen peroxide, again much faster than oxygen reduction, which resulted in the production of RSe-SG, presumably via a selenenic acid intermediate (RSeOH) which was trapped by excess GSH. This third step was studied on exogenous hydroperoxide in anaerobic conditions, and it could be eliminated from the glutathione oxidase cycle in the presence of excess catalase. The role of RSe- as a one- and two-electron reductant was confirmed by competitive carboxymethylation with iodoacetate. RSe- was able to rapidly reduce ferric cytochrome c to its ferrous derivative. The overall rate of catalytic glutathione oxidation was GSH concentration dependent and oxygen concentration independent. Excess glutathione reductase and NADPH increased the catalytic oxidation of GSH, probably by switching the rate-limiting step from selenylsulfide to diselenide cleavage. When GSH was substituted for dithiothreitol, it was shown to reduce RSe-SeR to RSe- in a fast and quantitative reaction, and selenocystamine behaved as a dithiothreitol oxidase, whose catalytic cycle was dependent on oxygen concentration. The oxidase cycle of glutathione was inhibited by mercaptosuccinate, while that of dithiothreitol was not affected. When mercaptosuccinate was substituted for GSH, a stable selenenylsulfide was formed. These observations suggest that electrostatic interactions affect the reductive cleavage of diselenide and selenenylsulfide linkages. This study illustrates the ease of one-electron transfers from RSe- to a variety of reducible substrates. Such free radical mechanisms may explain much of the cytotoxicity of alkylselenols, and they demonstrate that selenocystamine is a poor catalytic model of the enzyme glutathione peroxidase.  相似文献   

18.
Red tide phytoplankton Chattonella marina is known to produce reactive oxygen species (ROS), such as superoxide anion (O(2)(-)), hydrogen peroxide (H(2)O(2)) and hydroxyl radical (&z.rad;OH), under normal physiological conditions. Although several lines of evidence suggest that ROS are involved in the mortality of fish exposed to C. marina, the mechanism of ROS generation in C. marina remains to be clarified. In this study, we found that the cell-free supernatant prepared from C. marina cells showed NAD(P)H-dependent O(2)(-) generation, and this response was inhibited by diphenyleneiodonium, an inhibitor of mammalian NADPH oxidase. When the cell-free supernatant of C. marina was analyzed by immunoblotting using antibody raised against the human neutrophil cytochrome b558 large subunit (gp91phox), a main band of approximately 110 kDa was detected. The cell surface localization of the epitope recognized with this antibody was also demonstrated in C. marina by indirect immunofluorescence. Furthermore, Southern blot analysis performed on genomic DNA of C. marina with a probe covering the C-terminal region of gp91phox suggested the presence of a single-copy gene coding for gp91phox homologous protein in C. marina. These results provide evidence for the involvement of an enzymatic system analogous to the neutrophil NADPH oxidase as a source of O(2)(-) production in C. marina.  相似文献   

19.
Urate oxidase (EC 1.7.3.3 or UOX) catalyzes the conversion of uric acid using gaseous molecular oxygen to 5-hydroxyisourate and hydrogen peroxide in absence of any cofactor or transition metal. The catalytic mechanism was investigated using X-ray diffraction, electron spin resonance spectroscopy (ESR), and quantum mechanics calculations. The X-ray structure of the anaerobic enzyme-substrate complex gives credit to substrate activation before the dioxygen fixation in the peroxo hole, where incoming and outgoing reagents (dioxygen, water, and hydrogen peroxide molecules) are handled. ESR spectroscopy establishes the initial monoelectron activation of the substrate without the participation of dioxygen. In addition, both X-ray structure and quantum mechanic calculations promote a conserved base oxidative system as the main structural features in UOX that protonates/deprotonates and activate the substrate into the doublet state now able to satisfy the Wigner's spin selection rule for reaction with molecular oxygen in its triplet ground state.  相似文献   

20.
Xanthine oxidase has been hypothesized to be an important source of biological free radical generation. The enzyme generates the superoxide radical, .O2- and has been widely applied as a .O2- generating system; however, the enzyme may also generate other forms of reduced oxygen. We have applied electron paramagnetic resonance (EPR) spectroscopy using the spin trap 5,5'-dimethyl-1-pyrroline-N-oxide (DMPO) to characterize the different radical species generated by xanthine oxidase along with the mechanisms of their generation. Upon reaction of xanthine with xanthine oxidase equilibrated with air, both DMPO-OOH and DMPO-OH radicals are observed. In the presence of ethanol or dimethyl sulfoxide, alpha-hydroxyethyl or methyl radicals are generated, respectively, indicating that significant DMPO-OH generation occurred directly from OH rather than simply from the breakdown of DMPO-OOH. Superoxide dismutase totally scavenged the DMPO-OOH signal but not the DMPO-OH signal suggesting that .O2- was not required for .OH generation. Catalase markedly decreased the DMPO-OH signal, while superoxide dismutase + catalase totally scavenged all radical generation. Thus, xanthine oxidase generates .OH via the reduction of O2 to H2O2, which in turn is reduced to .OH. In anaerobic preparations, the enzyme reduces H2O2 to .OH as evidenced by the appearance of a pure DMPO-OH signal. The presence of the flavin in the enzyme is required for both .O2- and .OH generation confirming that the flavin is the site of O2 reduction. The ratio of .O2- and .OH generation was affected by the relative concentrations of dissolved O2 and H2O2. Thus, xanthine oxidase can generate the highly reactive .OH radical as well as the less reactive .O2- radical. The direct production of .OH by xanthine oxidase in cells and tissues containing this enzyme could explain the presence of oxidative cellular damage which is not prevented by superoxide dismutase.  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