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 15 毫秒
1.
The first detailed inhibition study of recombinant vanadium chloroperoxidase (rVCPO) using hydroxylamine, hydrazine and azide has been carried out. Hydroxylamine inhibits rVCPO both competitively and uncompetitively. The competitive inhibition constant K(ic) and the uncompetitive inhibition constant K(iu) see are 40 and 80 microM, respectively. The kinetic data suggest that rVCPO may form a hydroxylamido complex, hydroxylamine also seems to react with the peroxovanadate complex during turnover. The kinetic data show that the type of inhibition for hydrazine and azide is uncompetitive with the uncompetitive inhibition constant K(iu) of 350 microM and 50 nM, respectively, showing that in particular azide is a very potent inhibitor of this enzyme. Substitution of vanadate in the active site by phosphate also leads to inactivation of vanadium chloroperoxidase. However, the presence of H(2)O(2) clearly prevents the inactivation of the enzyme by phosphate. This shows that pervanadate is bound much more strongly to the enzyme than vanadate.  相似文献   

2.
Mutation studies were performed on active-site residues of vanadium chloroperoxidase from the fungus Curvularia inaequalis, an enzyme which exhibits both haloperoxidase and phosphatase activity and is related to glucose-6-phosphatase. The effects of mutation to alanine on haloperoxidase activity were studied for the proposed catalytic residue His-404 and for residue Asp-292, which is located close to the vanadate cofactor. The mutants were strongly impaired in their ability to oxidize chloride but still oxidized bromide, although they inactivate during turnover. The effects on the optical absorption spectrum of vanadium chloroperoxidase indicate that mutant H404A has a reduced affinity for the cofactor, whereas this affinity is unchanged in mutant D292A. The effect on the phosphatase activity of the apoenzyme was investigated for six mutants of putative catalytic residues. Effects of mutation of His-496, Arg-490, Arg-360, Lys-353, and His-404 to alanine are in line with their proposed roles in nucleophilic attack, transition-state stabilization, and leaving-group protonation. Asp-292 is excluded as the group that protonates the leaving group. A model based on the mutagenesis studies is presented and may serve as a template for glucose-6-phosphatase and other related phosphatases. Hydrolysis of a phospho-histidine intermediate is the rate-determining step in the phosphatase activity of apochloroperoxidase, as shown by burst kinetics.  相似文献   

3.
Directed evolution was performed on vanadium chloroperoxidase from the fungus Curvularia inaequalis to increase its brominating activity at a mildly alkaline pH for industrial and synthetic applications and to further understand its mechanism. After successful expression of the enzyme in Escherichia coli, two rounds of screening and selection, saturation mutagenesis of a "hot spot," and rational recombination, a triple mutant (P395D/L241V/T343A) was obtained that showed a 100-fold increase in activity at pH 8 (k(cat) = 100 s(-1)). The increased K(m) values for Br(-) (3.1 mm) and H(2)O(2) (16 microm) are smaller than those found for vanadium bromoperoxidases that are reasonably active at this pH. In addition the brominating activity at pH 5 was increased by a factor of 6 (k(cat) = 575 s(-1)), and the chlorinating activity at pH 5 was increased by a factor of 2 (k(cat) = 36 s(-1)), yielding the "best" vanadium haloperoxidase known thus far. The mutations are in the first and second coordination sphere of the vanadate cofactor, and the catalytic effects suggest that fine tuning of residues Lys-353 and Phe-397, along with addition of negative charge or removal of positive charge near one of the vanadate oxygens, is very important. Lys-353 and Phe-397 were previously assigned to be essential in peroxide activation and halide binding. Analysis of the catalytic parameters of the mutant vanadium bromoperoxidase from the seaweed Ascophyllum nodosum also adds fuel to the discussion regarding factors governing the halide specificity of vanadium haloperoxidases. This study presents the first example of directed evolution of a vanadium enzyme.  相似文献   

4.
Vanadium compounds are shown to have a mitogenic effect on fibroblast cells. The effects of vanadate, vanadyl and pervanadate on the proliferation and morphological changes of Swiss 3T3 cells in culture are compared. Vanadium derivatives induced cell proliferation in a biphasic manner, with a toxic-like effect at doses over 50mM, after 24h of incubation. Vanadyl and vanadate were equally potent at 2.5–10mM. At 50mM vanadate inhibited cell proliferation, whereas slight inhibition was observed at 100mM of vanadyl. At 10mM pervanadate was as potent as vanadate and vanadyl in stimulating fibroblast proliferation, but no effect was observed at lower concentrations. A pronounced cytotoxic-like effect was induced by pervanadate at 50mM. All of these effects were accompanied by morphological changes: transformation of fibroblast shape from polygonal to fusiform; retraction with cytoplasm condensation; and loss of lamellar processes. The magnitude of these transformations correlates with the potency of vanadium derivatives to induce a cytotoxic-like effect: pervanadate>vanadate>vanadyl. These data suggest that the oxidation state and coordination geometry of vanadium determine the degree of the cytotoxicity.  相似文献   

5.
It has been reported that vanadate-stimulated oxidation of NAD(P)H by microsomal systems can proceed anaerobically, in contrast to the general notion that the oxidation proceeds exclusively by an O(2-)-dependent free radical chain mechanism. The current study indicates that microsomal systems are endowed with a vanadate-reductase property, involving a NAD(P)H-dependent electron transport cytochrome P450 system. Our ESR measurements demonstrated the formation of a vanadium(IV) species in a mixture containing vanadate, rat liver microsomes, and NAD(P)H. This vanadium(IV) species was identified as the vanadyl ion (VO2+) by comparison with the ESR spectrum of VOSO4. The initial rate of vanadium(IV) formation depends linearly on the concentration of microsomes. The Michaelis-Menten constants were found to be: km = 1.25 mM and Vmax = 0.066 mumol (min)-1 (mg microsomes)-1, respectively. Pretreatment of the microsomes with carbon monoxide or K3Fe(CN)6 reduced vanadium(IV) generation, suggesting that the NAD(P)H-dependent electron transport cytochrome P450 system plays a significant role in the microsomal reduction of vanadate. Measurements under argon or in the presence of superoxide dismutase caused only minor (less than 10%) reductions in vanadium(IV) generation. The VO2+ species was also detected in NAD(P)H oxidation by fructose plus vanadate, a reaction known to proceed via an O(2-)-mediated chain mechanism. However, the amount of vanadium(IV) generated by this reaction was an order of magnitude smaller than that by the microsomal system and was inhibitable by superoxide dismutase, affirming the conclusion that the microsomal/NAD(P)H system is endowed with the (O(2-)-independent) vanadium(V) reductase property.  相似文献   

6.
Binding of vanadate to human serum transferrin   总被引:1,自引:0,他引:1  
Human serum transferrin specifically and reversibly binds 2 equiv of vanadate at the two metal-binding sites of the protein. The vanadium(V)-transferrin complex can be formed either by the addition of vanadate to apotransferrin or by the air oxidation of the vanadyl(IV)-transferrin complex. The formation of the vanadium complex can be blocked by loading the apotransferrin with iron(III), and bound vanadium can be displaced from the protein by the subsequent addition of either gallium(III) or iron(III). The binding constant for the second equiv of vanadate is 10(6.5) in 0.1 M hepes, pH 7.4 at 25 degrees C. The binding constant for the first equiv of vanadate is probably very similar, although no quantitative value could be determined. Although transferrin reacts with the vanadate anion, studies on the transferrin model compound ethylenebis(o-hydroxyphenylglycine) indicate that at pH 9.5, the vanadium is binding at the metal-binding site as a dioxovanadium(V) cation coordinated to two phenolic residues at each binding site. This bound cation appears to be protonated over the pH range 9.5-6.5, as shown by changes in the difference uv spectrum of the transferrin complex, to produce an oxohydroxo species. Further decreases in the pH lead to dissociation of the vanadium-transferrin complex.  相似文献   

7.
Both vanadate and hydrogen peroxide (H2O2) are known to have insulin-mimetic effects. We previously reported that the mixture of vanadate plus H2O2 results in the generation of a peroxide(s) of vanadate, which strongly enhances IGF-II binding to rat adipocytes (Kadota et al., 1987b). We now report that pervanadate mimics insulin in isolated rat adipocytes to (1) stimulate lipogenesis, (2) inhibit epinephrine-stimulated lipolysis, and (3) stimulate protein synthesis. The efficacy of pervanadate is comparable to that of insulin. However, it is 10(2)-10(3) times more potent than vanadate alone. Exposure of intact rat adipocytes to pervanadate was found to activate the WGA-purified insulin receptor tyrosine kinase assayed with the exogenous substrate poly(Glu80/Tyr20) in a dose-dependent manner to a maximum of 1464% of control at 10(-3) M compared with a maximum insulin effect of 1046% at 10(-6) M. In contrast, in vitro assayed autophosphorylation of the WGA-purified extract was increased 3-fold after exposure of intact cells to insulin but not significantly increased after pervanadate. Furthermore, high concentrations of pervanadate (10(-5) M) inhibited subsequent in vitro added insulin-stimulated autophosphorylation. In vitro addition of pervanadate to WGA-purified receptors could not stimulate autophosphorylation or exogenous tyrosine kinase activity and did not inhibit insulin-stimulated autophosphorylation. Labeling of intact adipocytes with [32P]orthophosphate followed by exposure to 10(-4) M pervanadate increased insulin receptor beta-subunit phosphorylation (7.9 +/- 3.0)-fold, while 10(-7) M insulin and 10(-4) vanadate increased labeling (5.3 +/- 1.8)- and (1.1 +/- 0.2)-fold, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The stability of 11 vanadium compounds is tested under physiological conditions and in administration fluids. Several compounds including those currently used as insulin-mimetic agents in animal and human studies are stable upon dissolution in distilled water but lack such stability in distilled water at pH7. Complex lability may result in decomposition at neutral pH and thus may compromise the effectiveness of these compounds as therapeutic agents; Even well characterized vanadium compounds are surprisingly labile. Sufficiently stable complexes such as the VEDTA complex will only slowly reduce, however, none of the vanadium compounds currently used as insulin-mimetic agents show the high stability of the VEDTA complex. Both the bis(maltolato)oxovanadium(IV) and peroxovanadium complexes extend the insulin-mimetic action of vanadate in reducing cellular environments probably by increased lifetimes under physiological conditions and/or by decomposing to other insulin mimetic compounds. For example, treatment with two equivalents of glutathione or other thiols the (dipicolinato)peroxovanadate(V) forms 9dipicolinato)oxovanadate(V) and vanadate, which are both insulin-mimetic vanadium(V) compounds and can continue to act. The reactivity of vanadate under physiological conditions effects a multitude of biological responses. Other vanadium complexes may mimic insulin but not induce similar responses if the vanadate formation is blocked or reduced. We conclude that three properties, stability, lability and redox chemistry are critical to prolong the half-life of the insulin-mimetic form of vanadium compounds under physiological conditions and should all be considered in development of vanadium-based oral insulin-mimetic agents.Abbreviations ADP adenosine 5-diphosphate - ATP adenosine 5-triphosphate - ADP-V adenosine 5-diphosphate-vanadate - bpV bis(peroxo)oxovanadium(V) - (bpV)2 bis(peroxo)oxovanadium(V) dimer - bpVpic bis(peroxo)picolinatooxovanadate(V) - 13C carbon-13 - EDTA ethylenediaminetetraacetic acid - EPR electron paramagnetic resonance - EXSY exchange spectroscopy - 1H proton - HSG glutathione - NAD -nicotinamide adenine dinucleotide - NADP -nicotinamide adenine dinucleotide phosphate - NADV -nicotinamide adenine dinucleotide vanadate - NMR nuclear magnetic resonance (also referred to as magnetic resonance imaging) - pVdipic (dipicolinato)peroxovanadate(V) - Vcit (citrato)dioxovanadate(V) - VEDTA (ethylenediaminetetraacetato)dioxovanadate(V) - Vmalto bis(maltolato)-oxovanadium(IV) - Voxal bis(oxalato)dioxovanadate(V) - 51V vanadium-51 - V1 vanadate monomer - V2 vanadate dimer - V4 vanadate tetramer - V5 vanadate pentamer - UV-vis spectroscopy ultraviolet-visible spectroscopy  相似文献   

9.
In the presence of chloroperoxidase, indole was oxidized by H2O2 to give oxindole as the major product. Under most conditions oxindole was the only product formed, and under optimal conditions the conversion was quantitative. This reaction displayed maximal activity at pH 4.6, although appreciable activity was observed throughout the entire pH range investigated, namely pH 2.5-6.0. Enzyme saturation by indole could not be demonstrated, up to the limit of indole solubility in the buffer. The oxidation kinetics were first-order with respect to indole up to 8 mM, which was the highest concentration of indole that could be investigated. On the other hand, 2-methylindole was not affected by H2O2 and chloroperoxidase, but was a strong inhibitor of indole oxidation. The isomer 1-methylindole was a poor substrate for chloroperoxidase oxidation, and a weak inhibitor of indole oxidation. These results suggest the possibility that chloroperoxidase oxidation of the carbon atom adjacent to the nitrogen atom in part results from hydrogen-bonding of the substrate N-H group to the enzyme active site.  相似文献   

10.
Vanadium compounds are known to stimulate the oxidation of NAD(P)H, but the mechanism remains unclear. This reaction was studied spectrophotometrically and by electron spin resonance spectroscopy (ESR) using vanadium in the reduced state (+4, vanadyl) and the oxidized state (+5, vanadate). In 25 mM sodium phosphate buffer at pH 7.4, vanadyl was slightly more effective in stimulating NADH oxidation than was vanadate. Addition of a superoxide generating system, xanthine/xanthine oxidase, resulted in a marked increase in NADH oxidation by vanadyl, and to a lesser extent, by vanadate. Decreasing the pH with superoxide present increased NADH oxidation for both vanadate and vanadyl. Addition of hydrogen peroxide to the reaction mixture did not change the NADH oxidation by vanadate, regardless of concentration or pH. With vanadyl however, addition of hydrogen peroxide greatly enhanced NADH oxidation which further increased with lower pH. Use of the spin trap DMPO in reaction mixtures containing vanadyl and hydrogen peroxide or a superoxide generating system resulted in the detection by ESR of hydroxyl. In each case, the hydroxyl radical signal intensity increased with vanadium concentration. Catalase was able to inhibit the formation of the DMPO--OH adduct formed by vanadate plus superoxide. These results show that the ability of vanadium to act in a Fenton-type reaction is an important process in the vanadium-stimulated oxidation of NADH.  相似文献   

11.
The chloroperoxidase-catalyzed reactions of NAD(P)H with H2O2 in the presence of Cl- or Br- have been characterized. With 1 mol H2O2 per mol of NADH, one atom of 36Cl was incorporated into the 264-nm-absorbing intermediate product. This species was oxidized enzymatically by a second mole of H2O2 to a species distinct from NAD+, which retained one Cl atom. Spectroscopically identical species were also produced by reaction of NADH with one and two molar ratios of HOCl, respectively. These data indicate that, with respect to halogenation activities, chloroperoxidase functions similarly to myeloperoxidase, i.e., produces HOCl as the first product of Cl- oxidation by H2O2. Moreover, rapid chlorination of NAD(P)H followed by oxidation may be an important and highly lethal microbicidal effect of HOCl produced by myeloperoxidase in activated neutrophils.  相似文献   

12.
A radical species of monochlorodimedone has been characterized by its high reactivity with molecular O2. Horseradish peroxidase greatly accelerated O2 uptake by acidic solutions of this substrate; the enzymatic reaction required exogenous H2O2 only with freshly prepared substrate solutions, and the total substrate oxidized was equal to the sum of H2O2 added and O2 consumed. However, with excess Br- and horseradish peroxidase, or high Br- or Cl- and chloroperoxidase, a 1:1 stoichiometry between H2O2 and substrate was observed. In the absence of halide, the stoichiometry of the chloroperoxidase-catalyzed oxidation of monochlorodimedone changed to two molecules of the organic donor per H2O2. Moreover, in the absence of halide, at substrate:H2O2 ratios greater than 2.0, chloroperoxidase catalyzed significant O2 uptake; this enzyme-dependent autoxidation of monochlorodimedone also occurred in the presence of Cl- or Br-, when H2O2 was limiting. These data, and recent evidence from this laboratory for free hypohalous acid as the first product of chloroperoxidase-catalyzed halide oxidation [B. W. Griffin (1983) Biochem. Biophys. Res. Commun. 116, 873-879], strongly support a mixed enzymatic/nonenzymatic radical chain process as the mechanism for halogenation of monochlorodimedone by chloroperoxidase. Both horseradish peroxidase and chloroperoxidase can catalyze either bromination or oxidation of this substrate, depending on the experimental conditions. Implications of these results for the mechanism of HOCl formation catalyzed by chloroperoxidase are considered.  相似文献   

13.
Approaches to the rational design of vanadium-based semi-synthetic enzymes and biomimetic models as catalysts for enantioselective oxidations are reviewed. Incorporation of vanadate ion into the active site of phytase (E.C. 3.1.3.8), which in vivo mediates the hydrolysis of phosphate esters, afforded a semi-synthetic peroxidase. It catalyzed the enantioselective oxidation of prochiral sulfides with H2O2 affording the S-sulfoxide, e.g. in 66% ee at quantitative conversion of thioanisole. Under the reaction conditions the semi-synthetic vanadium peroxidase was stable for more than 3 days with only a slight decrease in turnover frequency. Amongst the transition-metal oxoanions that are known to be potent inhibitors of phosphatases, only vanadate resulted in a semi-synthetic peroxidase when incorporated into phytase. In a biomimetic approach, vanadium complexes of chiral Schiff base complexes were encapsulated in the super cages of a hydrophobic zeolite Y. Unfortunately, these ship-in-a-bottle complexes afforded only racemic sulfoxide in the catalytic oxidation of thioanisole with H2O2.  相似文献   

14.
Vanadium haloperoxidases have been reported to mediate the oxidation of halides to hypohalous acid and the sulfoxidation of organic sulfides to the corresponding sulfoxides in the presence of hydrogen peroxide. However, traditional heme peroxidase substrates were reported not to be oxidized by vanadium haloperoxidases. Surprisingly, we have now found that the recombinant vanadium chloroperoxidase from the fungus Curvularia inaequalis catalyzes the oxidation of 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid) (ABTS), a classical chromogenic heme peroxidase substrate. The enzyme mediates the oxidation of ABTS in the presence of hydrogen peroxide with a turnover frequency of 11 s(-1) at its pH optimum of 4.0. The Km of the recombinant enzyme for ABTS was observed to be approximately 35 microM at this pH value. In addition, the bleaching of an industrial sulfonated azo dye, Chicago Sky Blue 6B, catalyzed by the recombinant vanadium chloroperoxidase in the presence of hydrogen peroxide is reported.  相似文献   

15.
The direct effect of different vanadium compounds on acid phosphatase (ACP) activity was investigated. Vanadate and vanadyl but not pervanadate inhibited the wheat germ ACP activity. These vanadium derivatives did not alter the fibroblast Swiss 3T3 soluble fraction ACP activity. Using inhibitors of tyrosine phosphatases (PTPases), the wheat germ ACP was partially characterized as a PTPase. This study suggests that the inhibitory ability of different vanadium derivatives to modulate ACP activity seems to depend on the geometry around the vanadium atom more than on the oxidation state. Our results indicate a correlation between the PTPase activity and the sensitivity to vanadate and vanadyl cation.  相似文献   

16.
Jameson GN  Jin W  Krebs C  Perreira AS  Tavares P  Liu X  Theil EC  Huynh BH 《Biochemistry》2002,41(45):13435-13443
The catalytic step that initiates formation of the ferric oxy-hydroxide mineral core in the central cavity of H-type ferritin involves rapid oxidation of ferrous ion by molecular oxygen (ferroxidase reaction) at a binuclear site (ferroxidase site) found in each of the 24 subunits. Previous investigators have shown that the first detectable reaction intermediate of the ferroxidase reaction is a diferric-peroxo intermediate, F(peroxo), formed within 25 ms, which then leads to the release of H(2)O(2) and formation of ferric mineral precursors. The stoichiometric relationship between F(peroxo), H(2)O(2), and ferric mineral precursors, crucial to defining the reaction pathway and mechanism, has now been determined. To this end, a horseradish peroxidase-catalyzed spectrophotometric method was used as an assay for H(2)O(2). By rapidly mixing apo M ferritin from frog, Fe(2+), and O(2) and allowing the reaction to proceed for 70 ms when F(peroxo) has reached its maximum accumulation, followed by spraying the reaction mixture into the H(2)O(2) assay solution, we were able to quantitatively determine the amount of H(2)O(2) produced during the decay of F(peroxo). The correlation between the amount of H(2)O(2) released with the amount of F(peroxo) accumulated at 70 ms determined by M?ssbauer spectroscopy showed that F(peroxo) decays into H(2)O(2) with a stoichiometry of 1 F(peroxo):H(2)O(2). When the decay of F(peroxo) was monitored by rapid freeze-quench M?ssbauer spectroscopy, multiple diferric mu-oxo/mu-hydroxo complexes and small polynuclear ferric clusters were found to form at rate constants identical to the decay rate of F(peroxo). This observed parallel formation of multiple products (H(2)O(2), diferric complexes, and small polynuclear clusters) from the decay of a single precursor (F(peroxo)) provides useful mechanistic insights into ferritin mineralization and demonstrates a flexible ferroxidase site.  相似文献   

17.
We used a proteomic approach to identify proteins that associate with keratins 8 or 18 (K8/K18) in a pervanadate-dependent manner. Pervanadate triggers Ran-K8/K18 binding and a gel-migration-shift of Ran from 25 to 27 kDa, which does not occur upon exposure to H2O2 or vanadate or if pervanadate is excluded during cell solubilization. Generation of 27-kDa Ran is not related to hyperphosphorylation, is heat-insensitive, but occurs upon conversion of Ran cysteines to cysteic acid. The pervanadate-mediated Ran cysteine --> cysteic acid oxidation and its related gel migration shift affects other proteins including actin. Mutation of the three Ran cysteines (Cys-85, -112, and -120) showed that Ran Cys-112 oxidation generates 27-kDa Ran and accounts for its keratin binding. Proteasome inhibition accentuates Ran-keratin binding after cell exposure to pervanadate. Therefore, cell-free exposure to pervanadate causes cysteine to cysteic acid oxidation of Ran and several other proteins and Ran-K8/K18 association. In cells, stabilization of oxidized Ran by proteasome inhibition promotes Ran-keratin interaction. Keratin sequestration of oxidized Ran may provide a back-up protective mechanism in some cases of oxidative injury.  相似文献   

18.
The chloroperoxidase catalyzed oxidation of methyl phenyl sulfide to (R)-methyl phenyl sulfoxide was investigated, both in batch and membrane reactors, using as oxidant H2O2, or O2 in the presence of either dihydroxyfumaric acid or ascorbic acid. The effects of pH and nature and concentration of the oxidants on the selectivity, stability, and productivity of the enzyme were evaluated. The highest selectivity was displayed by ascorbic acid/O2, even though the activity of chloroperoxidase with this system was lower than that obtained with the others. When the reaction was carried out in a membrane reactor, it was possible to reuse the enzyme for several conversion cycles. The results obtained with ascorbic acid/O2 and dihydroxyfumaric acid/O2 as oxidants do not seem to be compatible with either a mechanism involving hydroxyl radicals as the active species or with the hypothesis that oxidation occurs through the initial formation of H2O2. Copyright 1999 John Wiley & Sons, Inc.  相似文献   

19.
The Caldariomyces fumago chloroperoxidase was successfully expressed in Aspergillus niger. The recombinant enzyme was produced in the culture medium as an active protein and could be purified by a three-step purification procedure. The catalytic behavior of recombinant chloroperoxidase (rCPO) was studied and compared with that of native CPO. The specific chlorination activity (47 units/nmol) of rCPO and its pH optimum (pH 2.75) were very similar to those of native CPO. rCPO catalyzes the oxidation of various substrates in comparable yields and selectivities to native CPO. Indole was oxidized to 2-oxindole with 99% selectivity and thioanisole to the corresponding R-sulfoxide (enantiomeric excess >98%). Incorporation of (18)O from labeled H(2)18O(2) into the oxidized products was 100% in both cases.  相似文献   

20.
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