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 578 毫秒
1.
EPR spin trapping experiments on bacterial oxalate decarboxylase from Bacillus subtilis under turn-over conditions are described. The use of doubly (13)C-labeled oxalate leads to a characteristic splitting of the observed radical adducts using the spin trap N-tert-butyl-α-phenylnitrone linking them directly to the substrate. The radical was identified as the carbon dioxide radical anion which is a key intermediate in the hypothetical reaction mechanism of both decarboxylase and oxidase activities. X-ray crystallography had identified a flexible loop, SENS161-4, which acts as a lid to the putative active site. Site directed mutagenesis of the hinge amino acids, S161 and T165 was explored and showed increased radical trapping yields compared to the wild type. In particular, T165V shows approximately ten times higher radical yields while at the same time its decarboxylase activity was reduced by about a factor of ten. This mutant lacks a critical H-bond between T165 and R92 resulting in compromised control over its radical chemistry allowing the radical intermediate to leak into the surrounding solution.  相似文献   

2.
Oxalate decarboxylases and oxalate oxidases are members of the cupin superfamily of proteins that have many common features: a manganese ion with a common ligand set, the substrate oxalate, and dioxygen (as either a unique cofactor or a substrate). We have hypothesized that these enzymes share common catalytic steps that diverge when a carboxylate radical intermediate becomes protonated. The Bacillus subtilis decarboxylase has two manganese binding sites, and we proposed that Glu162 on a flexible lid is the site 1 general acid. We now demonstrate that a decarboxylase can be converted into an oxidase by mutating amino acids of the lid that include Glu162 with specificity switches of 282,000 (SEN161-3DAS), 275,000 (SENS161-4DSSN), and 225,000 (SENS161-4DASN). The structure of the SENS161-4DSSN mutant showed that site 2 was not affected. The requirement for substitutions other than of Glu162 was, at least in part, due to the need to decrease the Km for dioxygen for the oxidase reaction. Reversion of decarboxylase activity could be achieved by reintroducing Glu162 to the SENS161-4DASN mutant to give a relative specificity switch of 25,600. This provides compelling evidence for the crucial role of Glu162 in the decarboxylase reaction consistent with it being the general acid, for the role of the lid in controlling the Km for dioxygen, and for site 1 being the sole catalytically active site. We also report the trapping of carboxylate radicals produced during turnover of the mutant with the highest oxidase activity. Such radicals were also observed with the wild-type decarboxylase.  相似文献   

3.
Oxalate decarboxylase converts oxalate to formate and carbon dioxide and uses dioxygen as a cofactor despite the reaction involving no net redox change. We have successfully used Fourier transform infrared spectroscopy to monitor in real time both substrate consumption and product formation for the first time. The assignment of the peaks was confirmed using [(13)C]oxalate as the substrate. The K(m) for oxalate determined using this assay was 3.8-fold lower than that estimated from a stopped assay. The infrared assay was also capable of distinguishing between oxalate decarboxylase and oxalate oxidase activity by the lack of formate being produced by the latter. In D(2)O, the product with oxalate decarboxylase was C-deuterio formate rather than formate, showing that the source of the hydron was solvent as expected. Large solvent deuterium kinetic isotope effects were observed on V(max) (7.1 +/- 0.3), K(m) for oxalate (3.9 +/- 0.9), and k(cat)/K(m) (1.8 +/- 0.4) indicative of a proton transfer event during a rate-limiting step. Semiempirical quantum mechanical calculations on the stability of formate-derived species gave an indication of the stability and nature of a likely enzyme-bound formyl radical catalytic intermediate. The capability of the enzyme to bind formate under conditions in which the enzyme is known to be active was determined by electron paramagnetic resonance. However, no enzyme-catalyzed exchange of the C-hydron of formate was observed using the infrared assay, suggesting that a formyl radical intermediate is not accessible in the reverse reaction. This restricts the formation of potentially harmful radical intermediates to the forward reaction.  相似文献   

4.
The reactions catalyzed by orotidine monophosphate decarboxylase, oxalate decarboxylase, organomercurial lyase and phosphopantothenoylcysteine decarboxylase involve putative high-energy carbanion intermediates that cannot be stabilized by delocalization. Mechanistic and structural studies on each of these enzymes are described that suggest different strategies for carbanion stabilization. Both orotidine monophosphate decarboxylase and organomercurial lyase are likely to avoid carbanion formation by protonating the fragmenting bond, oxalate decarboxylase stabilizes an acyl carbanion using an adjacent radical and phosphopantothenoylcysteine decarboxylase stabilizes its carbanion by delocalization into a transient thioaldehyde.  相似文献   

5.
Oxalate decarboxylase is a manganese-dependent enzyme that catalyzes the conversion of oxalate to formate and carbon dioxide. We have determined the structure of oxalate decarboxylase from Bacillus subtilis at 1.75 A resolution in the presence of formate. The structure reveals a hexamer with 32-point symmetry in which each monomer belongs to the cupin family of proteins. Oxalate decarboxylase is further classified as a bicupin because it contains two cupin folds, possibly resulting from gene duplication. Each oxalate decarboxylase cupin domain contains one manganese binding site. Each of the oxalate decarboxylase domains is structurally similar to oxalate oxidase, which catalyzes the manganese-dependent oxidative decarboxylation of oxalate to carbon dioxide and hydrogen peroxide. Amino acid side chains in the two metal binding sites of oxalate decarboxylase and the metal binding site of oxalate oxidase are very similar. Four manganese binding residues (three histidines and one glutamate) are conserved as well as a number of hydrophobic residues. The most notable difference is the presence of Glu333 in the metal binding site of the second cupin domain of oxalate decarboxylase. We postulate that this domain is responsible for the decarboxylase activity and that Glu333 serves as a proton donor in the production of formate. Mutation of Glu333 to alanine reduces the catalytic activity by a factor of 25. The function of the other domain in oxalate decarboxylase is not yet known.  相似文献   

6.
Oxalate decarboxylase (EC 4.1.1.2) catalyzes the conversion of oxalate to formate and carbon dioxide and utilizes dioxygen as a cofactor. By contrast, the evolutionarily related oxalate oxidase (EC 1.2.3.4) converts oxalate and dioxygen to carbon dioxide and hydrogen peroxide. Divergent free radical catalytic mechanisms have been proposed for these enzymes that involve the requirement of an active site proton donor in the decarboxylase but not the oxidase reaction. The oxidase possesses only one domain and manganese binding site per subunit, while the decarboxylase has two domains and two manganese sites per subunit. A structure of the decarboxylase together with a limited mutagenesis study has recently been interpreted as evidence that the C-terminal domain manganese binding site (site 2) is the catalytic site and that Glu-333 is the crucial proton donor (Anand, R., Dorrestein, P. C., Kinsland, C., Begley, T. P., and Ealick, S. E. (2002) Biochemistry 41, 7659-7669). The N-terminal binding site (site 1) of this structure is solvent-exposed (open) and lacks a suitable proton donor for the decarboxylase reaction. We report a new structure of the decarboxylase that shows a loop containing a 3(10) helix near site 1 in an alternative conformation. This loop adopts a "closed" conformation forming a lid covering the entrance to site 1. This conformational change brings Glu-162 close to the manganese ion, making it a new candidate for the crucial proton donor. Site-directed mutagenesis of equivalent residues in each domain provides evidence that Glu-162 performs this vital role and that the N-terminal domain is either the sole or the dominant catalytically active domain.  相似文献   

7.
S-Adenosylmethionine decarboxylase (AdoMetDC) is a pyruvoyl-dependent enzyme that catalyzes the formation of the aminopropyl group donor in the biosynthesis of the polyamines spermidine and spermine. The enzyme is synthesized as a protein precursor and is activated by an autocatalytic serinolysis reaction that creates the pyruvoyl group. The autoprocessing reaction proceeds via an N --> O acyl rearrangement, generating first an oxyoxazolidine anion intermediate followed by an ester intermediate. A similar strategy is utilized in self-catalyzed protein splicing reactions and in autoproteolytic activation of protein precursors. Mutation of Ser68 to alanine in human AdoMetDC prevents processing by removing the serine side chain necessary for nucleophilic attack at the adjacent carbonyl carbon atom. We have determined the X-ray structure of the S68A mutant and have constructed models of the proenzyme and the oxyoxazolidine intermediate. Formation of the oxyoxazolidine intermediate is promoted by a hydrogen bond from Cys82 and stabilized by a hydrogen bond from Ser229. These observations are consistent with mutagenesis studies, which show that the C82S and C82A mutants process slowly and that the S229A mutant does not process at all. Donation of a proton by His243 to the nitrogen atom of the oxyoxazolidine ring converts the oxyoxazolidine anion to the ester intermediate. The absence of a base to activate the hydroxyl group of Ser68 suggests that strain may play a role in the cleavage reaction. Comparison of AdoMetDC with other self-processing proteins shows no common structural features. Comparison to histidine decarboxylase and aspartate decarboxylase shows that these pyruvoyl-dependent enzymes evolved different catalytic strategies for forming the same cofactor.  相似文献   

8.
杨桃提取物体外清除氧自由基作用   总被引:1,自引:0,他引:1  
从杨桃果中提取得到三种提取物为匀浆提取物、蛋白提取物和多糖提取物。采用化学发光法测定这三种提取物清除氧自由基的活性,实验结果:匀浆提取物清除羟自由基(·OH)和H2O2的活性大小相近,而清除超氧阴离子自由基(O2–·)的活性较小,其IC50约为前两者的4倍。蛋白提取物清除O2–·和·OH的活性大小相近,而清除H2O2的活性明显小于前两者,IC50约为前两者的9倍。多糖提取物清除.OH的活性明显大于清除O2–·和H2O2的活性,其IC50约为O2–·的1/22,约为H2O2的1/65。结果表明,杨桃果具有清除O2–·、·OH和H2O2的作用,不同提取物对这些活性氧自由基的清除能力有所不同。  相似文献   

9.
Anaerobic reduction of hydrogen peroxide in a xanthine/xanthine oxidase system by adriamycin semiquinone in the presence of chelators and radical scavengers was investigated by direct electron paramagnetic resonance and spin trapping techniques. Under these conditions, adriamycin semiquinone appears to react with hydrogen peroxide forming the hydroxyl radical in the presence of chelators such as ethylenediaminetetraacetic acid and diethylenetriaminepentaacetic acid. In the absence of chelators, a related, but unknown oxidant is formed. In the presence of desferrioxamine, adriamycin semiquinone does not disappear in the presence of hydrogen peroxide at a detectable rate. The presence of adventitious iron is therefore implicated during adriamycin semiquinone-catalyzed reduction of hydrogen peroxide. Formation of alpha-hydroxyethyl radical and carbon dioxide radical anion from ethanol and formate, respectively, was detected by spin trapping. Both the hydroxyl radical and the related oxidant react with these scavengers, forming the corresponding radical. In the presence of scavengers from which reducing radicals are formed, the rate of consumption of hydrogen peroxide in this system is increased. This result can be explained by a radical-driven Fenton reaction.  相似文献   

10.
Rates of autoxidation of NAD(P)H initiated by hydroperoxyl radical, the acid form of superoxide anion radical which was generated by xanthine/xanthine oxidase, followed a typical autoxidation kinetic equation. Second-order rate constants for the reactions of NADPH and NADH with hydroperoxyl radical were found to be 9.82 +/- 0.13 x 10(4) M-1s-1 and 9.26 +/- 0.58 x 10(4) M-1s-1 at 25 degrees C, respectively. Rates of the reactions between NAD(P)H and superoxide to give degraded products other than NAD(P)+ were also investigated.  相似文献   

11.
Oxalate decarboxylase (EC 4.1.1.2) catalyses the conversion of oxalate into carbon dioxide and formate. It requires manganese and, uniquely, dioxygen for catalysis. It forms a homohexamer and each subunit contains two similar, but distinct, manganese sites termed sites 1 and 2. There is kinetic evidence that only site 1 is catalytically active and that site 2 is purely structural. However, the kinetics of enzymes with mutations in site 2 are often ambiguous and all mutant kinetics have been interpreted without structural information. Nine new site-directed mutants have been generated and four mutant crystal structures have now been solved. Most mutants targeted (i) the flexibility (T165P), (ii) favoured conformation (S161A, S164A, D297A or H299A) or (iii) presence (Delta162-163 or Delta162-164) of a lid associated with site 1. The kinetics of these mutants were consistent with only site 1 being catalytically active. This was particularly striking with D297A and H299A because they disrupted hydrogen bonds between the lid and a neighbouring subunit only when in the open conformation and were distant from site 2. These observations also provided the first evidence that the flexibility and stability of lid conformations are important in catalysis. The deletion of the lid to mimic the plant oxalate oxidase led to a loss of decarboxylase activity, but only a slight elevation in the oxalate oxidase side reaction, implying other changes are required to afford a reaction specificity switch. The four mutant crystal structures (R92A, E162A, Delta162-163 and S161A) strongly support the hypothesis that site 2 is purely structural.  相似文献   

12.
Oxidation of veratryl alcohol by lignin peroxidase (LiP) was potently inhibited by oxalic acid. The inhibition analysis with Lineweaver-Burk plots clearly showed that the type of inhibition is non-competitive. The enzymatic oxidation of veratryl alcohol in the presence of 14C-oxalic acid yielded radioactive carbon dioxide. The results indicate that the apparent inhibition of LiP is caused by reduction of the veratryl alcohol cation radical intermediate back to the substrate level by oxalate, which is concomitantly oxidized to carbon dioxide.  相似文献   

13.
A quantitative cytochemical method for ornithine decarboxylase activity   总被引:1,自引:0,他引:1  
Although decarboxylases, particularly ornithine decarboxylase, are of considerable importance in cell metabolism, it has been impossible to demonstrate their activity histochemically, as this depends on trapping carbon dioxide at neutral pH values. A new reagent, lead hydroxyisobutyrate, has been shown capable of such trapping. It has been applied to the demonstration of ornithine decarboxylase activity in mouse kidney. Optimal concentrations of substrate, co-factor and trapping agent, as well as the pH optimum, have been determined for cryostat sections stabilized with a collagen polypeptide. The activity was inhibited by the specific ornithine decarboxylase inhibitor alpha-difluoromethyl ornithine.  相似文献   

14.
A general radiochemical method for estimating the activity of amino acid decarboxylases is reported. This method utilizes ion-exchange cartridges to separate unreacted radiolabeled amino acid substrates from product amines, which can then readily be quantitated by liquid scintillation counting. The assay is simple, rapid, and more sensitive than standard 14CO2 trapping procedures if uniformly labeled amino acid substrates are utilized. Acidic, basic, and aromatic amino acid decarboxylases can be assayed with the appropriate choice of cation or anion exchangers. The utility of the method is demonstrated for aspartate-alpha-decarboxylase, tyrosine decarboxylase, and lysine decarboxylase where kinetic parameters are comparable to values obtained by standard radiochemical 14CO2 trapping assays.  相似文献   

15.
Lipoic acid and its reduced form, dihydrolipoic acid, are thought to be strong antioxidants. There are also reports of dihydrolipoic acid acting as a pro-oxidant under certain circumstances. This article reports the direct observation by ESR spectrometry of the disulfide radical anion and the spin trapping of the primary thiyl radical formed from the oxidation of dihydrolipoic acid through thiol pumping with phenol and horseradish peroxidase. The disulfide radical anion reacts rapidly with oxygen to form the reactive radical superoxide, which is also trapped. The radical species formed show a potential for pro-oxidant activity of this compound. Although antioxidants, in general, have been shown to have pro-oxidant potential, the pro-oxidant chemistry of dihydrolipoic acid has not been well characterized.  相似文献   

16.
J L Popp  B Kalyanaraman  T K Kirk 《Biochemistry》1990,29(46):10475-10480
Veratryl alcohol (3,4-dimethoxybenzyl alcohol) appears to have multiple roles in lignin degradation by Phanerochaete chrysosporium. It is synthesized de novo by the fungus. It apparently induces expression of lignin peroxidase (LiP), and it protects LiP from inactivation by H2O2. In addition, veratryl alcohol has been shown to potentiate LiP oxidation of compounds that are not good LiP substrates. We have now observed the formation of Mn3+ in reaction mixtures containing LiP, Mn2+, veratryl alcohol, malonate buffer, H2O2, and O2. No Mn3+ was formed if veratryl alcohol or H2O2 was omitted. Mn3+ formation also showed an absolute requirement for oxygen, and oxygen consumption was observed in the reactions. This suggests involvement of active oxygen species. In experiments using oxalate (a metabolite of P. chrysosporium) instead of malonate, similar results were obtained. However, in this case, we detected (by ESR spin-trapping) the production of carbon dioxide anion radical (CO2.-) and perhydroxyl radical (.OOH) in reaction mixtures containing LiP, oxalate, veratryl alcohol, H2O2, and O2. Our data indicate the formation of oxalate radical, which decays to CO2 and CO2.-. The latter reacts with O2 to form O2.-, which then oxidizes Mn2+ to Mn3+. No radicals were detected in the absence of veratryl alcohol. These results indicate that LiP can indirectly oxidize Mn2+ and that veratryl alcohol is probably a radical mediator in this system.  相似文献   

17.
The one-electron oxidation of (bi)sulfite is catalyzed by peroxidases to yield the sulfur trioxide radical anion (SO3-), a predominantly sulfur-centered radical as shown by studies with 33S-labeled (bi)sulfite. This radical reacts with molecular oxygen to form a peroxyl radical. The subsequent reaction of this peroxyl radical with (bi)sulfite has been proposed to form the sulfate anion radical, which is nearly as strong an oxidant as the hydroxyl radical. We used the spin trapping electron spin resonance technique to provide for the first time direct evidence for sulfate anion radical formation during (bi)sulfite peroxidation. The sulfate anion radical is known to react with many compounds more commonly thought of as hydroxyl radical scavengers such as formate and ethanol. Free radicals derived from these scavengers are trapped in systems where (bi)sulfite peroxidation has been inhibited by these scavengers.  相似文献   

18.
Xanthine oxidase is generally recognized as a key enzyme in purine catabolism, but its structural complexity, low substrate specificity, and specialized tissue distribution suggest other functions that remain to be fully identified. The potential of xanthine oxidase to generate superoxide radical anion, hydrogen peroxide, and peroxynitrite has been extensively explored in pathophysiological contexts. Here we demonstrate that xanthine oxidase turnover at physiological pH produces a strong one-electron oxidant, the carbonate radical anion. The radical was shown to be produced from acetaldehyde oxidation by xanthine oxidase in the presence of catalase and bicarbonate on the basis of several lines of evidence such as oxidation of both dihydrorhodamine 123 and 5,5-dimethyl-1-pyrroline-N-oxide and chemiluminescence and isotope labeling/mass spectrometry studies. In the case of xanthine oxidase acting upon xanthine and hypoxanthine as substrates, carbonate radical anion production was also evidenced by the oxidation of 5,5-dimethyl-1-pyrroline-N-oxide and of dihydrorhodamine 123 in the presence of uricase. The results indicated that Fenton chemistry occurring in the bulk solution is not necessary for carbonate radical anion production. Under the conditions employed, the radical was likely to be produced at the enzyme active site by reduction of a peroxymonocarbonate intermediate whose formation and reduction is facilitated by the many xanthine oxidase redox centers. In addition to indicating that the carbonate radical anion may be an important mediator of the pathophysiological effects of xanthine oxidase, the results emphasize the potential of the bicarbonate-carbon dioxide pair as a source of biological oxidants.  相似文献   

19.
1. The metabolism of oxalate by the pink-pigmented organisms, Pseudomonas AM1, Pseudomonas AM2, Protaminobacter ruber and Pseudomonas extorquens has been compared with that of the non-pigmented Pseudomonas oxalaticus. 2. During growth on oxalate, all the organisms contain oxalyl-CoA decarboxylase, formate dehydrogenase and oxalyl-CoA reductase. This is consistent with oxidation of oxalate to carbon dioxide taking place via oxalyl-CoA, formyl-CoA and formate as intermediates, and also reduction of oxalate to glyoxylate taking place via oxalyl-CoA. 3. The pink-pigmented organisms, when grown on oxalate, contain l-serine–glyoxylate aminotransferase and hydroxypyruvate reductase but do not contain glyoxylate carboligase. The converse of this obtains in oxalate-grown Ps. oxalaticus. This indicates that, in contrast with Ps. oxalaticus, synthesis of C3 compounds from oxalate by the pink-pigmented organisms occurs by a variant of the `serine pathway' used by Pseudomonas AM1 during growth on C1 compounds. 4. Evidence in favour of this scheme is provided by the finding that a mutant of Pseudomonas AM1 that lacks hydroxypyruvate reductase is not able to grow on oxalate.  相似文献   

20.
The o-, m-, and p-nitrobenzyl chlorides are reduced aerobically and anaerobically by NADPH and rat hepatic microsomes. Under aerobic conditions, these nitro anion radicals reduce oxygen to superoxide as demonstrated by oxygen consumption and spin trapping of superoxide with 5,5-dimethyl-1-pyrroline N-oxide. At low oxygen concentration, the p- and o-nitro anion radicals undergo intramolecular electron transfer and decompose to carbon-centered nitrobenzyl radicals, which can be spin-trapped with t-nitrosobutane. The p-nitrobenzyl (o-nitrobenzyl) radical adduct was characterized by a nitrogen hyperfine splitting of 16.5 G (17.1 G) and two equivalent beta-hydrogen hyperfine splittings of 10.6 G (14.4 G). The spin trap 5,5-dimethyl-1-pyrroline N-oxide also yields adducts characteristic of carbon-centered free radicals. This unimolecular decomposition is much faster than the disproportionation decay, which is characteristic of most nitro anion radicals, and the primary o- and p-nitrobenzyl chloride anion radicals never achieve detectable concentrations. The nitrobenzyl radical trapping is not inhibited by metyrapone or CO. In contrast, the m-nitrobenzyl anion radical does achieve a detectable steady-state concentration, which is increased 20% by either metyrapone or a CO atmosphere.  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